Processing math: 32%
Research article

Existence and asymptotic behavior of normalized solutions for the mass supercritical fractional Kirchhoff equations with general nonlinearities

  • Received: 25 October 2024 Revised: 31 December 2024 Accepted: 03 January 2025 Published: 10 January 2025
  • MSC : 35A15, 35J20

  • In this paper, we studied a fractional Kirchhoff equation with mass supercritical general nonlinearities. Under some suitable conditions, we obtained the existence of ground state normalized solutions for this equation. Moreover, we presented the asymptotic behavior of normalized solutions to the above equation as c0+ and c+.

    Citation: Min Shu, Haibo Chen, Jie Yang. Existence and asymptotic behavior of normalized solutions for the mass supercritical fractional Kirchhoff equations with general nonlinearities[J]. AIMS Mathematics, 2025, 10(1): 499-533. doi: 10.3934/math.2025023

    Related Papers:

    [1] Zhongxiang Wang . Existence and asymptotic behavior of normalized solutions for the modified Kirchhoff equations in R3. AIMS Mathematics, 2022, 7(5): 8774-8801. doi: 10.3934/math.2022490
    [2] Huanhuan Wang, Kexin Ouyang, Huiqin Lu . Normalized ground states for fractional Kirchhoff equations with critical or supercritical nonlinearity. AIMS Mathematics, 2022, 7(6): 10790-10806. doi: 10.3934/math.2022603
    [3] Zhidan Shu, Jianjun Zhang . Normalized multi-bump solutions of nonlinear Kirchhoff equations. AIMS Mathematics, 2024, 9(6): 16790-16809. doi: 10.3934/math.2024814
    [4] Shengbin Yu, Jianqing Chen . Asymptotic behavior of the unique solution for a fractional Kirchhoff problem with singularity. AIMS Mathematics, 2021, 6(7): 7187-7198. doi: 10.3934/math.2021421
    [5] Yuan Shan, Baoqing Liu . Existence and multiplicity of solutions for generalized asymptotically linear Schrödinger-Kirchhoff equations. AIMS Mathematics, 2021, 6(6): 6160-6170. doi: 10.3934/math.2021361
    [6] Jie Yang, Haibo Chen . Normalized solutions for Kirchhoff-Carrier type equation. AIMS Mathematics, 2023, 8(9): 21622-21635. doi: 10.3934/math.20231102
    [7] Adel M. Al-Mahdi . The coupling system of Kirchhoff and Euler-Bernoulli plates with logarithmic source terms: Strong damping versus weak damping of variable-exponent type. AIMS Mathematics, 2023, 8(11): 27439-27459. doi: 10.3934/math.20231404
    [8] Zhiqiang Li, Yanzhe Fan . On asymptotics of solutions for superdiffusion and subdiffusion equations with the Riemann-Liouville fractional derivative. AIMS Mathematics, 2023, 8(8): 19210-19239. doi: 10.3934/math.2023980
    [9] Ruonan Liu, Tomás Caraballo . Random dynamics for a stochastic nonlocal reaction-diffusion equation with an energy functional. AIMS Mathematics, 2024, 9(4): 8020-8042. doi: 10.3934/math.2024390
    [10] Aníbal Coronel, Fernando Huancas . New results for the non-oscillatory asymptotic behavior of high order differential equations of Poincaré type. AIMS Mathematics, 2022, 7(4): 6420-6444. doi: 10.3934/math.2022358
  • In this paper, we studied a fractional Kirchhoff equation with mass supercritical general nonlinearities. Under some suitable conditions, we obtained the existence of ground state normalized solutions for this equation. Moreover, we presented the asymptotic behavior of normalized solutions to the above equation as c0+ and c+.



    The purpose of this paper is to study the existence and blow-up behavior of positive solutions for the following fractional Kirchhoff equation:

    {(a+bRN|(Δ)s2u|2dx)(Δ)su+V(x)u+ωu=f(u),  xRN,uHs(RN), (A)

    having prescribed mass

    RN|u|2dx=c, (1.1)

    where a,b,c are positive constants, 1N<4s, s(0,1), VC(RN,R), fC(R,R), ωR is a Lagrange multiplier, and (Δ)s denotes the fractional Laplacian operator defined as

    (Δ)su=F1(|ξ|2sF(u)),ξRN,

    where F denotes the Fourier transform on RN. It is well-known that it can also be computed by

    (Δ)sv(x)=CN,sP.V.RNv(x)v(y)|xy|N+2sdy,

    if v is smooth enough, where CN,s is the normalization constant, and P.V. denotes a Cauchy principle value, see [13,15,18].

    When s=1,f(u)=|u|p2u, V(x)=0, (A) reduces to the following case, i.e.,

    {(a+bRN|u|2dx)Δu+ωu=|u|p2u,  xRN,uH1(RN). (1.2)

    The sharp existence and the concentration behavior of the normalized solution of (1.2) in the mass subcritical, supercritical and critical cases were established in [7,14,22]. In fact, the author obtained the solutions by looking for critical points of the following functional:

    I(u)=a2RN|u|2dx+b4(RN|u|2dx)21pRN|u|pdx

    constrained on the L1 sphere in H1(RN),

    Sc={uH1(RN):uL2(RN)=c>0}.

    In [22], Ye showed that the constrained minimization problem

    Ic2:=infScI(u) (1.3)

    admits a minimizer if, and only if, c>cp with p(0,2+4N] or ccp with p(2+4N,2+8N), and there is no minimizer of (1.3) if p[2+8N,+). In [23], Ye studied (1.3) with p=2+8N and obtained that there exists a mountain pass critical point of I(u) on Sc if c>c. In the mass subcritical case, a complete classification with respect to the exponent p for its L2 normalized critical points can be deduced from some simple energy estimates in [24]. To be precise, they gained existence and the uniqueness of the mountain pass type minimum for (1.3) with p(2+8N,2) or p=2+8N and c>c. Moreover, if f(u) satisfies some suitable conditions, the authors of [8] studied the blow-up behavior of minimizers (1.3).

    To the best of our knowledge, if b0, then Eq (A ) with a prescribed mass has been studied in [16,17,21]. Frank-Lenzmann-Silvestre [4] established the uniqueness result of the positive ground state solution of the equation

    (Δ)su+u=|u|4sNu,

    which is an important foundation for blow-up analysis. In [2], Du et al. explored the existence, nonexistence and mass concentration of L2-normalized solutions for nonlinear fractional Schrödinger equations with nonnegative potentials

    (Δ)su+V(x)u=μu+βf(u),

    under the following assumptions:

    (f1): fC(R,R),|f(t)|c1(|t|+|t|p1) for some c1>0 and 2<p<2+4sN.

    (f2): fC(R,R),|f(t)|c1(|t|+|t|p1) for some c1>0 and 2+4sN<p<2s=2NN2s.

    (f3): there exist ν>2+4sN and r0>0 such that

    0<νF(t)tf(t),forall|t|r0.

    For the case of power function f(t)=tp2t with 2<p<2s, Du et al. conducted a complete classification of the existence and nonexistence of minimizers, except that p=2+4s/N. Very recently, Bao-Lv-Ou [1] investigated the following fractional Schrödinger equation with prescribed mass:

    (Δ)su=μu+a(x)|u|p2u,

    where s(0,1),2+4sN<p<2s. The existence of the bounded state normalized solution under various conditions on a(x) was demonstrated in [1]. For more recent works about the fractional Schrödinger or Kirchhoff equation, see [10,12,20] and the references therein.

    It is worth pointing out that p=2+4s/N is the mass critical exponent related to (A). However, Eq (A) involving the general potential and nonlinearities has not yet been resolved. An interesting question now is whether the same existence or nonexistence results occurs for the nonhomogeneous nonlinearities and mass supercritical case of (A). On the other hand, there have been no previous articles studying the asymptotic behavior of solutions to (A). Therefore, our goal is to fill the gaps in these areas. More precisely, in the first part of the paper, we prove the existence of solutions for (A) with V(x)0.

    Before describing more details, let's introduce the following fractional Gagliardo-Nirenberg-Sobolev inequality in [2] and Hardy inequality in [5].

    Lemma 1.1. [2] For uHs(RN) and q(0,2s2), the fractional Gagliardo-Nirenberg-Sobolev inequality

    RN|u|q+2dxCopt(RN|(Δ)s2u|2dx)Nq4s(RN|u|2dx)q+22Nq4s (1.4)

    is attained at a function ϕq(x) with the following properties:

    (i) ϕq(x) is radial, positive, and strictly decreasing in |x|.

    (ii) ϕq(x) belongs to H2s+1(RN)C(RN) and satisfies

    C11+|x|2s+Nϕq(x)C21+|x|2s+N,xRN,

    where Ci(i=1,2) are positive constants.

    (iii) ϕq(x) is the unique solution of the fractional Schrödinger equation

    {Nq4s(Δ)su+[1+q4(2Ns)]uuq+1=0,uHs(RN),q(0,2s2).

    (iv) Copt=q+22ϕqq2.

    Lemma 1.2. [5] Let s(0,1) and N>2s. Then, for all uDs,2(RN),

    RN|u(x)|2|x|2sdxHN,su2Ds,2(RN), (1.5)

    where

    HN,s=2πN2Γ2(N+2s4)Γ2(N2s4)Γ2(N+2s2)|Γ(s)|.

    Lemma 1.3. [13] Let s(0,1) and N>2s. Then, there exists a constant S>0 such that for any uDs,2(RN),

    u2L2s(RN)S1u2Ds,2(RN).

    In this paper, we will require f(x) to satisfy the following conditions:

    (H1) fC(R,R) and odd.

    (H2) There exist some λ,γR+×R+ with

    {2+8sN<λγ<2s:=2NN2s,ifN2s,2+8sN<λγ<2s:=+,ifN=2s,

    such that

    0<λF(t)f(t)tγF(t),fort0,whereF(t)=t0f(τ)dτ.

    (H3) The function ˜F(t):=12f(t)tF(t) is of class C1 and

    ˜F(t)tλ˜F(t),tR,

    where λ is given in (H2).

    We assume that V(x) is a radial function and satisfies the following assumptions:

    (V1) V(x)C1(RN)Lp(RN),p(N2,),lim|x|V(x)=0,infxRNV(x)=0.

    (V2) There exists κ1[0,s) such that one of the following two conditions holds

    (ⅰ)V(x)x2aκ1HN,s|x|2s, for any xRN{0}, where HN,s is given in Lemma 1.2;

    (ⅱ)max{V(x)x,0}LN2s(RN)2aκ1S, where S is given in Lemma 1.3.

    (V3) There exists κ2(0,N(λ2)4s4) such that

    V(x)xaκ22HN,s|x|2s,

    for any xRN{0}.

    Let us introduce the space of radial functions in Hs(RN) defined by

    Hsrad(RN)={uHs(RN):u(x)=u(|x|)}.

    It is standard to see that critical points of the energy functional

    J(u):=a2RN|(Δ)s2u|2dx+b4(RN|(Δ)s2u|2dx)2+12RNV(x)u2dxRNF(u)dx

    restricted to the (mass) constraint

    Sc:={uHs(RN):RN|u|2dx=c}

    are normalized solutions of (A).

    From (H1) and (H2), there exist C1,C2>0 such that for each tR,

    C1min{|t|λ,|t|γ}F(t)C2max{|t|λ,|t|γ}C2(|t|λ+|t|γ), (1.6)
    (λ21)F(t)˜F(t)(γ21)F(t), (1.7)

    where C2=F(1). It follows from (1.4) that there exists C3>0, for any uHs(RN),

    uλLλ(RN)C3uλDs,2(RN)uλλL2(RN),uγLγ(RN)C3uγDs,2(RN)uγγL2(RN), (1.8)

    where λ=(λ2)N2s(2,λ),γ=(γ2)N2s(2,γ). Let

    L:=saκ1aN(γ21)C2C3, (1.9)
    ϱ:=min{(L2/2cλλ+cγγ)1λ2,1}, (1.10)
    K:=2N(N2s)γ(γ2)c[(122s(λ2)N2κ2(λ2)N)aϱ+(142s(λ2)N)bϱ2]. (1.11)

    (V4) supxRN(V(x)+12sV(x)x)<K.

    Our first result is as follows.

    Theorem 1. Let s(0,1) and N>2s. Assume that (V1)(V4) and (H1)(H3) hold. Then, Eq (A) admits at least a radial solution.

    The second purpose of this article is to establish the existence results of ground state solutions for (A) wih V(x)=0.

    Theorem 2. Let 1N<4s, s(0,1), V(x)=0, and suppose that f satisfies (H1)(H3). Then, for all c>0 fixed, Eq (A) admits a ground state normalized solution (ωc,uc) with ωc>0 and ucHsrad(RN).

    Remark 1. Theorems 1 and 2 extend and complement the previous results on the fractional Kirchhoff equation with a mass super critical general nonlinearities. Particularly, unlike [8], in which V(x)=0, we apply a new deformation argument for the constrained functional on Sc.

    Remark 2. We now point out some difficulties faced in Theorems 1 and 2.

    (i) When V(x)>0 and the nonlinearity f is general mass supercritical, it prevents us from obtaining the compactness. We will apply a new deformation argument for the constraint functional with a new type of Palais-Smale condition denoted by (PSP)m.

    (ii) The simplest case of the function f satisfying the assumptions (H1)(H3) is f(t)=|t|p2t with 2+8s/N<p<2s. Naturally, the class of general nonlinearities satisfying these assumptions is much more difficult than this homogeneous case.

    (iii) Due to the appearance of the Kirchhoff nonlocal term

    RN|(Δ)s2u|2dx(Δ)su,

    Eq (A) is no longer point by point identity. Compared with [8], it is worth noting that (A) is a double nonlocal equation, and the decay estimates of test function near infinity are different from those in the case of the classical local problem; we thus borrow ideas of [3] for nonlocal operators to establish the decay estimates. This phenomenon has caused some mathematical difficulties, making research on such problems particularly interesting.

    Our other aim is to study the behavior of the normalized solution uc given in Theorem 2 as c0 and c+. In this direction, we need to assume that

    (H4) limt0+f(t)tλ1=μ1>0.

    (H5) limt+f(t)tγ1=μ2>0.

    The following results demonstrate the asymptotic behavior of uc in the sense of C2,αloc(RN) as well as Hs(RN).

    Theorem 3. Let 1N<4s, s(12,1), V(x)=0, and suppose that f satisfies (H1)(H5). For c>0, let (ωc,uc) be given by Theorem 2, then

    vc(x):=ω12λcuc(xω12sc)Q(x)inC2,αloc(RN),asc+,

    where Q is the unique radial positive solution of

    {a(Δ)sQ+Q=μ1Qλ1,inRN,lim|x|+Q(x)=0.

    Theorem 4. Let 1N<4s, s(12,1), V(x)=0, and suppose that f satisfies (H1)(H5). For c>0, let vc and Q be given by Theorem 3, then

    vc(x)Q(x)inHs(RN),asc+.

    Theorem 5. Let 1N<4s, s(12,1), V(x)=0, and suppose that f satisfies (H1)(H5). For c>0, let (ωc,uc) be given by Theorem 2, then

    ˉvc(x):=ω12λcuc(uc1sDs,2(RN)ω12scx)U(x)inC2,αloc(RN),asc0+,

    where U is the unique radial positive solution of

    {b(Δ)sU+U=μ2Uγ1,inRN,lim|x|+U(x)=0.

    Theorem 6. Let 1N<4s, s(12,1), V(x)=0, and suppose that f satisfies (H1)(H5). For c>0, let ˜vc and U be given by Theorem 5, then

    ˉvc(x)U(x)inHs(RN),asc0+.

    Throughout the paper, we use the following notations:

    Lq(RN) denotes the Lebesgue space with the norm

    uLq(RN)=(RN|u|qdx)1/q.

    ● For any xRN and R>0, BR(x):={yRN:|yx|<R}.

    C indicates positive numbers that may be different in different lines.

    The rest of this paper is organized as follows. Section 2 is dedicated to some preliminary notations and lemmas. In Section 3, we obtain the radial solutions for Eq (A) with V(x)0 and Theorem 1 will be proved there. In Section 4, we derive the existence of ground state normalized solution for problem (A) and give the proof of Theorems 2. In Section 5, we deal with asymptotic property of minimizers to problem (A) by proving Theorems 3–6.

    In this section, we provide some lemmas that will be frequently used in the rest of this article.

    We claim that the condition (V2) yields that for any uHs(RN),

    RNV(x)xu2dx2aκ1u2Ds,2(RN). (2.1)

    Indeed, if (i) of (V2) holds, from Lemma 1.2, we deduce that

    RNV(x)xu2dx2aκ1HN,sRNu2|x|2dx2aκ1u2Ds,2(RN).

    If (ii) of (V2) holds, by the Sobolev embedding inequality, we observe that

    RNV(x)xu2dx(RN|max{V(x)x,0}|N2sdx)2sN(RN|u|2NN2sdx)N2sNS1max{V(x)x,0}LN2s(RN)u2Ds,2(RN)2aκ1u2Ds,2(RN).

    Define

    P(u):=sau2Ds,2(RN)+sbu4Ds,2(RN)NRN˜F(u)dx12RN(V(x)x)u2dx, (2.2)

    and

    Mc:={uSc:P(u)<0}.

    Then

    Mc:={uSc:P(u)=0}.

    Set

    mc:=infMcJ(u).

    Lemma 2.1. Let s(0,1) and N>2s. Assume that (H1)(H3) and (V1)(V4) hold, then there exists ˉm>0 such that mcˉm>0. Moreover,

    ˉm(122s(λ2)N2κ2(λ2)N)aϱ+(142s(λ2)N)bϱ2, (2.3)

    where ϱ is given in (1.10).

    Proof. For any uMc, applying (1.7), (1.8), (2.1), and (2.2), we obtain that

    0=sau2Ds,2(RN)+sbu4Ds,2(RN)NRN˜F(u)dx12RN(V(x)x)u2dxsau2Ds,2(RN)N(γ21)RNF(u)dxaκ1u2Ds,2(RN)(sκ1)au2Ds,2(RN)N(γ21)C2(uλLλ(RN)+uγLγ(RN))(sκ1)au2Ds,2(RN)N(γ21)C2C3(uλDs,2(RN)cλλ2+uγDs,2(RN)cγγ2). (2.4)

    Set

    M1=N(γ21)C2C3cλλ2,M2=N(γ21)C2C3cγγ2, (2.5)
    g(t)=(sκ1)aM1tλ2M2tγ2.

    Since g(t) is decreasing on [0,+), there exists a unique t0>0 such that

    (sκ1)aM1tλ20M2tγ20=0, (2.6)

    and t0 is dependent on κ1,λ,γ,C2,C3,c. It follows from (2.4) that g(uDs,2(RN))0. Therefore, for any uMc, we observe that uDs,2(RN)t0 and

    J(u)=a2RN|(Δ)s2u|2dx+b4(RN|(Δ)s2u|2dx)2+12RNV(x)u2dxRNF(u)dxa2RN|(Δ)s2u|2dx+b4(RN|(Δ)s2u|2dx)22λ2RN˜F(u)dx=(122s(λ2)N)au2Ds,2(RN)+(142s(λ2)N)bu4Ds,2(RN)+1λ2RN(V(x)x)u2dx(122s(λ2)N2κ2λ2)au2Ds,2(RN)+(142s(λ2)N)bu4Ds,2(RN),

    due to (1.7), (2.2), (V3), and Lemma 1.2. Let

    ˉm:=(122s(λ2)N2κ2λ2)at20+(142s(λ2)N)bt40,

    then, mcˉm>0. From (1.9), (2.5), and (2.6), we conclude that

    L=cλλ2tλ20+cγγ2tγ20(cλλ+cγγ)12(t2(λ2)0+t2(γ2)0)12,

    which yields that

    t2(λ2)0+t2(γ2)0L2cλλ+cγγ.

    If L<cγγ2+cλλ2, thus g(1)<0, and we conclude that 0<t0<1 and

    t2(λ2)0>12(t2(λ2)0+t2(γ2)0)L2/2cλλ+cγγ,

    that is, t0>(L2/2cλλ+cγγ)12(λ2). If Lcγγ2+cλλ2, thus g(1)0, and t01. Therefore, t20ϱ and (2.3) holds.

    As in [9], we define

    mΓ=infγΓmaxt[0,1]J(γ(t)),

    where

    Γ={γC([0,1],Sc)):γ(0)Mc,γ(1)Sc¯Mc,J(γ(0))<12mc,J(γ(1))<12mc}.

    We will show that mΓ is well-defined.

    Lemma 3.1. Let s(0,1) and N>2s. Assume that (H1)(H3) and (V1)(V4) hold, then Γ.

    Proof. For any uSc, define

    uτ(x):=τN2u(τx),τ>0.

    It's not difficult to see uτ(x)Sc. From (1.6), for any τ1,

    RNF(uτ)dxC1min{uλLλ(RN),uγLγ(RN)}τλ22N, (3.1)

    and for any 0<τ<1,

    C1min{uλLλ(RN),uγLγ(RN)}τγ22NRNF(uτ)dxC2(uλLλ(RN)τλ22N+uγLγ(RN)τγ22N). (3.2)

    According to (3.1) and

    0RNV(xτ)u2dxVL(RN)c,

    we observe that

    J(uτ)=aτ2s2u2Ds,2(RN)+bτ4s4u4Ds,2(RN)+12RNV(xτ)u2dxRNF(uτ)dxaτ2s2u2Ds,2(RN)+bτ4s4u4Ds,2(RN)+12VL(RN)cC1min{uλLλ(RN),uγLγ(RN)}τλ22N,

    which yields that J(uτ), as τ+. On the other side, from (3.2), we conclude that

    J(uτ)=aτ2s2u2Ds,2(RN)+bτ4s4u4Ds,2(RN)+12RNV(xτ)u2dxRNF(uτ)dxaτ2s2u2Ds,2(RN)+bτ4s4u4Ds,2(RN)C2(uλLλ(RN)τλ22N+uγLγ(RN)τγ22N),

    and

    J(uτ)aτ2s2u2Ds,2(RN)+bτ4s4u4Ds,2(RN)+12RNV(xτ)u2dxC1min{uλLλ(RN),uγLγ(RN)}τγ22N,

    which leads to J(uτ)0+, as τ0+, recalling lim|x|V(x)=0. From (V2), (V3), and (1.5), we see that

    2κ2τ2su2Ds,2(RN)RN(V(x)x)u2τ(x)dx2κ1τ2su2Ds,2(RN).

    Combining with (1.7), (2.2), and (3.1), we deduce that

    P(uτ)=asτ2su2Ds,2(RN)+bsτ4su4Ds,2(RN)NRN˜F(uτ)dx12RN(V(x)x)u2τdx(s+κ2)aτ2su2Ds,2(RN)+bsτ4su4Ds,2(RN)N(λ21)C1min{uλLλ(RN),uγLγ(RN)}τλ22N,

    and

    P(uτ)(sκ1)aτ2su2Ds,2(RN)+bsτ4su4Ds,2(RN)N(γ21)C2(uλLλ(RN)τλ22N+uγLγ(RN)τγ22N),

    which yields that P(uτ) as τ+, and P(uτ)0+ as τ0+. Therefore, for any given uSc, we can take τ0>0 large enough, and τ1(0,1) small enough such that

    J(uτ1)<mc2,P(uτ1)>0,J(uτ0)<0,P(uτ0)<0.

    Thus, taking γ0(t):=uτ0(1t)+τ1t, we see that γ0(t)Γ.

    Lemma 3.2. Let s(0,1) and N>2s. Assume that (H1)(H3) and (V1)(V4) hold, then mΓmcˉm.

    Proof. For any γΓ, there exists tγ[0,1] such that γ(tγ)Mc. Therefore,

    maxt[0,1]J(γ(t))J(γ(tγ))infMcJ(u)=mc.

    Together with Lemma 2.1, the conclusion holds.

    Now in view of Lemma 3.1, we can apply a new deformation argument for the constraint functional on Sc with a new type of Palais-Smale condition denoted by (PSP)m. The functional J satisfies the (PSP)m condition on Sc, if, and only if, any (PSP)m sequence {un}Sc satisfying

    J(un)m,J(un)TunSc0,P(un)0, (3.3)

    has a strongly convergent subsequence.

    Lemma 3.3. Let s(0,1) and N>2s. Assume that (H1)(H3) and (V1)(V4) hold. If {un}Sc is a (PSP)m sequence satisfying (3.3) with mˉm, then

    (i) {un} is bounded in Hs(RN).

    (ii) there exists ω>0 such that the sequence of Lagrange multipliers ωn satisfying

    (a+bRN|(Δ)s2un|2dx)(Δ)sun+V(x)un+ωnun=f(un)+on(1)

    converges to ω in the sense of subsequence.

    Proof. (ⅰ) For every mR, let {un}Sc be a (PSP)m sequence satisfying (3.3). From (3.3), (V1), (V3) and (1.7), we conclude that

    m+on(1)=J(un)2N(λ2)P(un)=(122sN(λ2))aun2Ds,2(RN)+(142sN(λ2))bun4Ds,2(RN)+12RNV(x)u2ndx+2λ2RN˜F(un)dxRNF(un)dx+1N(λ2)RN(V(x)x)u2ndx(122sN(λ2)2κ2N(λ2))aun2Ds,2(RN)+(142sN(λ2))bun4Ds,2(RN),

    which implies that {un} is bounded in Hs(RN), recalling that κ2(0,N(λ2)4s4) and λ>2+8sN. Then, up to a subsequence, unu weakly in Hs(RN). From (3.3), there exists {ωn}R such that

    (a+bRN|(Δ)s2un|2dx)(Δ)sun+V(x)un+ωnun=f(un)+on(1)inHs(RN).

    (ⅱ) By (3.3), we see that

    \begin{equation*} \begin{aligned} &m+o_{n}(1) = J(u_{n}) -\frac{1}{2s} \mathcal{P}(u_{n})\\ = & -\frac{b}{4} \|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} +\frac{1}{2} \int_{\mathbb{R}^{N}} V(x)u_{n}^{2} \mathrm{d}x +\frac{N}{2s} \int_{ \mathbb{R}^{N}} \widetilde{F}(u_{n}) \mathrm{d}x-\int_{ \mathbb{R}^{N}} F(u_{n}) \mathrm{d}x +\frac{1}{4s} \int_{ \mathbb{R}^{N}} (\nabla V(x)\cdot x)u_{n}^{2} \mathrm{d}x. \end{aligned} \end{equation*}

    Combining with (1.7), we observe that

    \begin{equation*} 2\left[ \frac{N}{2s} \left(\frac{\gamma}{2}-1\right) -1 \right] \int_{ \mathbb{R}^{N}} F(u_{n}) \mathrm{d}x +\int_{ \mathbb{R}^{N}} \left[ V +\frac{1}{2s} (\nabla V(x)\cdot x) \right] u_{n}^{2} \mathrm{d}x \geqslant 2m+o_{n}(1). \end{equation*}

    Set

    \begin{equation*} \begin{aligned} &y_{n}: = \int_{ \mathbb{R}^{N}} F(u_{n}) \mathrm{d}x, \; \; z_{n}: = \int_{ \mathbb{R}^{N}} \left[ V +\frac{1}{2s} (\nabla V(x)\cdot x) \right] u_{n}^{2} \mathrm{d}x, \\ &\bar{\lambda}: = 2\left[ \frac{N}{2s} \left(\frac{\lambda}{2}-1\right) -1 \right], \; \; \bar{\gamma}: = 2\left[ \frac{N}{2s} \left(\frac{\gamma}{2}-1\right) -1 \right], \; \; \bar{\alpha}: = \frac{2N-(N-2s)\gamma}{2s}. \end{aligned} \end{equation*}

    Obviously, y_{n}\geqslant 0 , \bar{\gamma} , \bar{\alpha} > 0 , and

    \begin{equation} \bar{\gamma}y_{n}+z_{n} \geqslant 2m+o_{n}(1). \end{equation} (3.4)

    Applying (V_{4}) , (1.11), and Lemma 2.1, there exists \delta > 0 such that

    \begin{equation*} \sup\limits_{x\in\mathbb{R}^{N}} \left( V +\frac{1}{2s} (\nabla V(x)\cdot x) \right) < \frac{2s\bar{\alpha}\bar{m}} {(\bar{\gamma}+\bar{\alpha})c}-\delta. \end{equation*}

    Noting that m\geqslant \bar{m} , we obtain

    \begin{equation} V +\frac{1}{2s} (\nabla V(x)\cdot x) \leqslant \sup\limits_{x\in\mathbb{R}^{N}} \left( V +\frac{1}{2s} (\nabla V(x)\cdot x) \right) < \frac{2s\bar{\alpha}m} {(\bar{\gamma}+\bar{\alpha})c}-\delta. \end{equation} (3.5)

    Let

    \begin{equation*} z_{c, m}: = \frac{2s\bar{\alpha}m} {(\bar{\gamma}+\bar{\alpha})} -\frac{\delta c}{2} < 2m, \end{equation*}

    and from (3.5), we deduce that

    \begin{equation} z_{n}\leqslant \frac{2s\bar{\alpha}m} {(\bar{\gamma}+\bar{\alpha})} -\delta c < z_{c, m}. \end{equation} (3.6)

    Using (3.4)–(3.6), we deduce that

    \begin{equation} z_{n} < \bar{\alpha}y_{n}-\hat{m}+o(1), \end{equation} (3.7)

    where \hat{m} = \bar{\alpha}\bar{y}-\bar{z} , and (\bar{y}, \bar{z}) satisfies

    \begin{equation*} \begin{cases} \bar{\gamma}\bar{y}+\bar{z} = 2m, \\ \bar{z} = z_{c, m}. \end{cases} \end{equation*}

    Therefore,

    \begin{equation} \hat{m} = \frac{2\bar{\alpha}}{\bar{\gamma}} m-\frac{\bar{\alpha}+\bar{\gamma}}{\bar{\gamma}}z_{c, m} = \frac{\bar{\alpha}+\bar{\gamma}}{2\bar{\gamma}} \delta c > 0. \end{equation} (3.8)

    Gathering (1.7), (2.2), (3.7), and (3.8), we conclude that

    \begin{equation} \begin{aligned} \omega_{n}c = &\omega_{c}\|u_{n}\|_{L^{2}(\mathbb{R}^{N})}^{2}\\ = &-a \|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} -b \|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} - \int_{\mathbb{R}^{N}} V(x)u_{n}^{2} \mathrm{d}x +\int_{ \mathbb{R}^{N}} f(u_{n})u_{n} \mathrm{d}x+o_{n}(1)\\ = &-\frac{N}{s} \int_{\mathbb{R}^{N}} \widetilde{F}(u_{n}) \mathrm{d}x -\frac{1}{2s} \int_{\mathbb{R}^{N}} (\nabla V(x)\cdot x) u_{n}^{2} \mathrm{d}x- \int_{\mathbb{R}^{N}} V(x)u_{n}^{2} \mathrm{d}x +\int_{ \mathbb{R}^{N}} f(u_{n})u_{n} \mathrm{d}x+o_{n}(1)\\ = & \int_{\mathbb{R}^{N}} \left( -\frac{N-2s}{2s}f(u_{n})u_{n} +\frac{N}{s}F(u_{n}) \right) \mathrm{d}x -\int_{\mathbb{R}^{N}} \left( V+\frac{1}{2s} \nabla V(x)\cdot x \right) u_{n}^{2} \mathrm{d}x +o_{n}(1)\\ \geqslant& \bar{\alpha}y_{n}-z_{n}+o_{n}(1)\\ > &\hat{m}+o_{n}(1). \end{aligned} \end{equation} (3.9)

    Meanwhile, from (3.9), (H_{2}) , (V_{1}), (V_{3}) , and Lemmas 1.1 and 1.2, we infer that

    \begin{equation} \begin{aligned} \omega_{n}c = &\int_{\mathbb{R}^{N}} \left( -\frac{N-2s}{2s}f(u_{n})u_{n} +\frac{N}{s}F(u_{n}) \right) \mathrm{d}x -\int_{\mathbb{R}^{N}} \left( V+\frac{1}{2s} \nabla V(x)\cdot x \right) u_{n}^{2} \mathrm{d}x +o_{n}(1)\\ \leqslant& \left( -\frac{(N-2s)\lambda}{2s} +\frac{N}{s} \right) \int_{\mathbb{R}^{N}} F(u_{n}) \mathrm{d}x +\frac{a\kappa_{2}}{4s} \|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +o_{n}(1)\\ \leqslant& C+o_{n}(1). \end{aligned} \end{equation} (3.10)

    (3.9) and (3.10) imply that \omega_{n}\to \omega > 0 .

    In what follows, we consider that V(x) is radial, and then we can choose H^{s}_{rad}(\mathbb{R}^{N}) as the workspace. We take \mathcal{S}_{c}^{rad} : = \mathcal{S}_{c}\cap H^{s}_{rad}(\mathbb{R}^{N}).

    Lemma 3.4. Assume that (V_{1}) (V_{4}) and (H_{1}) (H_{3}) hold. Then, J(u) satisfies the (PSP)_{m} condition on \mathcal{S}_{c}^{rad} for m\geqslant \bar{m} .

    Proof. Let \{u_{n}\}\subset \mathcal{S}^{rad}_{c} be a (PSP)_{m} sequence satisfying (3.3) with m\geqslant \bar{m} . Due to Lemma 3.3, there exist u\in H^{s}_{rad}(\mathbb{R}^{N}) and \omega > 0 such that, up to a subsequence, u_{n}\rightharpoonup u weakly in H^{s}_{rad}(\mathbb{R}^{N}) and

    \begin{equation} \left(a+b\int_{\mathbb{R}^{N}} |(-\Delta)^{\frac{s}{2}}u|^{2} \mathrm{d}x \right)(-\Delta)^{s}u +V(x)u +\omega u = f(u). \end{equation} (3.11)

    Then, using the compact embedding H_{rad}^{s}(\mathbb{R}^{N}) \hookrightarrow\hookrightarrow L^{q}(\mathbb{R}^{N}) for all q\in (2, 2_{s}^{*}) , we deduce that u_{n}\to u strongly in L^{q}(\mathbb{R}^{N}) and

    \begin{equation} \int_{\mathbb{R}^{N}} f(u_{n})u_{n} \mathrm{d}x \to \int_{\mathbb{R}^{N}} f(u)u \mathrm{d}x. \end{equation} (3.12)

    Due to (V_{1}) , for p\in (\frac{N}{2s}, +\infty) , \|u_{n}-u\|_{L^{2p{'}}(\mathbb{R}^{N})} \to 0 as n\to\infty , where p^{'} = \frac{p}{p-1} . Hence, by u_{n}\rightharpoonup u weakly in H^{s}_{rad}(\mathbb{R}^{N}) , we observe that

    \begin{equation} \int_{\mathbb{R}^{N}} V(x)u_{n}^{2} \mathrm{d}x -\int_{\mathbb{R}^{N}} V(x)u^{2} \mathrm{d}x = \int_{\mathbb{R}^{N}} V(x)(u-u_{n})^{2} \mathrm{d}x+o_{n}(1) \leqslant \|V\|_{L^{p}(\mathbb{R}^{N})} \|u_{n}-u\|_{L^{2p{'}}(\mathbb{R}^{N})} = o_{n}(1). \end{equation} (3.13)

    Therefroe, by (3.11)–(3.13), we see that

    \begin{equation*} a (\|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} -\|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{2}) +b (\|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} -\|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} )+\omega (\|u_{n}\|_{L^{2}(\mathbb{R}^{N})}^{2} -\|u\|_{L^{2}(\mathbb{R}^{N})}^{2}) = o_{n}(1), \end{equation*}

    which implies that u_{n}\to u strongly in H^{s}_{rad}(\mathbb{R}^{N}) .

    To prove the existence of nontrivial solutions, we use the following deformation result given by [9].

    Define

    \begin{equation*} \mathcal{K}_{m} : = \{u\in \mathcal{S}_{c}: J(u) = m, dJ|_{\mathcal{S}_{c}} = 0, \mathcal{P}(u) = 0\}. \end{equation*}

    Lemma 3.5. ([9, Proposition 4.5]) Assume that J satisfies the (PSP)_{m} condition on \mathcal{S}_{c} . For any neighborhood D of \mathcal{K}_{m} (if \mathcal{K}_{m} = \emptyset , D = \emptyset ) and any \tilde{\varepsilon} > 0 , there exists \varepsilon\in (0, \tilde{\varepsilon}) and \eta\in C([0, 1]\times \mathcal{S}_{c}, \mathcal{S}_{c}) such that

    (i) \eta(0, u) = u , for u\in \mathcal{S}_{c} ;

    (ii) \eta(t, u) = u , for t\in [0, 1] if u\in [J\leqslant m-\tilde{\varepsilon}]_{\mathcal{S}_{c}} ;

    (iii) t\longmapsto J(\eta(t, u)) is nonincreasing for u\in \mathcal{S}_{c} ;

    (iv) \eta(1, [J\leqslant m-\varepsilon]_{\mathcal{S}_{c}}\setminus D) \subset [J\leqslant m-\varepsilon]_{\mathcal{S}_{c}}, \eta(1, [J\leqslant m+\varepsilon]_{\mathcal{S}_{c}}) \subset [J\leqslant m-\varepsilon]_{\mathcal{S}_{c}} \cup D .

    Proof of Theorem 1. Let

    \begin{equation*} \mathcal{K}_{m_{\Gamma}} : = \{u\in \mathcal{S}_{c}: J(u) = m_{\Gamma}, dJ|_{\mathcal{S}^{rad}_{c}} = 0, \mathcal{P}(u) = 0\} = \emptyset, \end{equation*}

    then U = \emptyset . It follows from Lemmas 3.2 and 3.4 that J satisfies the (PSP)_{m_{\Gamma}} condition. Due to m_{\Gamma}\geqslant m_{c} > 0 , taking \tilde{\varepsilon} = m_{\Gamma}-\frac{m_{c}}{2} , and using Lemma 3.5, we obtain that there exist \varepsilon\in (0, \tilde{\varepsilon}), \eta\in C([0, 1]\times\mathcal{S}_{c}^{rad}, \mathcal{S}_{c}^{rad}) satisfying

    \begin{equation} \eta(t, u) = u, \; \; \forall t\in [0, 1], \; \; u\in [J\leqslant m_{\Gamma}-\tilde{\varepsilon}]_{\mathcal{S}^{rad}_{c}}, \end{equation} (3.14)
    \begin{equation} \eta(1, [J\leqslant m_{\Gamma}+\varepsilon]_{\mathcal{S}^{rad}_{c}} \setminus U )\subset [J\leqslant m_{\Gamma}-\varepsilon]_{\mathcal{S}^{rad}_{c}}, \end{equation} (3.15)

    According to the definition of m_{\Gamma} and Lemma 3.1, there exists \gamma\in \Gamma such that

    \begin{equation*} \max\limits_{t\in [0, 1]} J(\gamma(t)) < m_{\Gamma}+\varepsilon. \end{equation*}

    Let us define \tilde{\gamma}: = \eta(1, \gamma(t)) , and claim that \tilde{\gamma}\in \Gamma . Indeed, J(\gamma(0)) < \frac{1}{2}m_{c} = m_{\Gamma}-\tilde{\varepsilon} , which implies that \gamma(0)\in [J\leqslant m_{\Gamma}-\tilde{\varepsilon}]_{\mathcal{S}^{rad}_{c}} . By (3.14), we observe that \tilde{\gamma}(0) = \eta(1, \gamma(0)) = \gamma(0) and \tilde{\gamma}(1) = \gamma(1) analogously. By (3.15), we conclude that

    \begin{equation*} J(\eta(1, \gamma(t))) \leqslant m_{\Gamma}-\varepsilon. \end{equation*}

    Therefore, J(\tilde{\gamma}(t))\leqslant m_{\Gamma}-\varepsilon for any t\in [0, 1] . We have

    \begin{equation*} m_{\Gamma} = \inf\limits_{\gamma\in \Gamma}\max\limits_{t\in [0, 1]}J(\gamma(t)) \leqslant \max\limits_{t\in [0, 1]}J(\tilde{\gamma}(t)) \leqslant m_{\Gamma}-\varepsilon, \end{equation*}

    which yields a contradiction. Hence, \mathcal{K}_{m_{\Gamma}}\neq \emptyset . Thie implies that (\mathcal{A}) admits a radial solution.

    In this section, we consider the existence of ground state solutions for (\mathcal{A}) with V(x) = 0 . From now on, in this article, we always assume that (H_ {1}) (H_ {3}) hold and will not further mention it.

    Lemma 4.1. Let 1\leqslant N<4s , s\in (0, 1) , and suppose that f satisfies (H_{1}) (H_{3}) . If u\in H^{s}(\mathbb{R}^{N}) is a nontrivial solution of Eq (\mathcal{A}), then u\in\mathcal{M} , where

    \begin{equation*} \mathcal{M}: = \{u\in H^{s}(\mathbb{R}^{N}) : \mathcal{P}(u) = 0\}, \end{equation*}

    and

    \begin{equation*} \mathcal{P}(u) : = sa\|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +sb\|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} -N \int_{\mathbb{R}^{N}} \widetilde{F}(u) \mathrm{d}x. \end{equation*}

    Proof. Let u be a solution to Eq (\mathcal{A}), and we derive that

    \begin{equation} a\|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +b\|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} +\omega \|u\|_{L^{2}(\mathbb{R}^{N})}^{2} = \int_{\mathbb{R}^{N}}f(u)u \mathrm{d}x. \end{equation} (4.1)

    Meanwhile, u satisfies the following Pohožaev identity:

    \begin{equation} (N-2s) \left( a\|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +b\|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} \right) +N\omega\|u\|_{L^{2}(\mathbb{R}^{N})}^{2} = 2N \int_{\mathbb{R}^{N}}F(u) \mathrm{d}x. \end{equation} (4.2)

    Combining (4.1) and (4.2), we conclude that

    \begin{equation} sa\|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +sb\|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} = N \int_{\mathbb{R}^{N}} \widetilde{F}(u) \mathrm{d}x. \end{equation} (4.3)

    Define

    \begin{equation*} J_{u}(\tau): = J(u_{\tau}), \; \; \tau > 0, \end{equation*}

    where u_{\tau}(x): = \tau^{\frac{N}{2}}u(\tau x), \tau > 0 . We observe that

    \begin{equation} \mathcal{P}(u_{\tau}) = \tau(J_{u})^{'}(\tau), \; \; \tau > 0. \end{equation} (4.4)

    Particularly, it holds that \mathcal{P}(u) = (J_{u})^{'}(1) . By exploiting it, we can deduce the following result.

    Lemma 4.2. Let u\in \mathcal{S}_{c} . Then, \tau > 0 is a critical point of J_{u}(\tau) if, and only if, u_{\tau}\in \mathcal{M}_{c} , where

    \begin{equation*} \mathcal{M}_{c}: = \mathcal{M}\cap \mathcal{S}_{c}. \end{equation*}

    Lemma 4.3. For each critical point of J|_{\mathcal{M}_{c}} , if (J_{u})^{''}(1)\neq 0 , then there exists a \omega\in \mathbb{R} such that

    \begin{equation*} J'(u)+\omega u = 0. \end{equation*}

    Proof. Let u be a critical point of J(u) constrained on \mathcal{M}_{c} , then there exist \omega, \nu\in \mathbb{R} such that

    \begin{equation} J'(u)+\omega u+\nu \mathcal{P}'(u) = 0. \end{equation} (4.5)

    Therefore, we need to show that \nu = 0 . Let

    \begin{equation*} \Phi(u): = J(u) +\frac{\omega}{2} \|u\|_{L^{2}(\mathbb{R}^{N})}^{2} +\nu \mathcal{P}(u), \end{equation*}

    and it is the corresponding energy functional of (4.5). Thanks to (4.5), one sees that u satisfies the corresponding Pohozaev identity

    \begin{equation} (\Phi_{u})'(1) : = \frac{d}{d\tau}\Phi(u_{\tau})|_{\tau = 1} = 0. \end{equation} (4.6)

    By (4.4), we observe that

    \begin{equation*} \Phi_{u}(\tau) = \Phi(u_{\tau}) = J(u_{\tau}) +\frac{\omega}{2} \|u\|_{L^{2}(\mathbb{R}^{N})}^{2} +\nu \mathcal{P}(u_{\tau}) = J(u_{\tau}) +\frac{\omega}{2} \|u\|_{L^{2}(\mathbb{R}^{N})}^{2} +\nu \tau(J_{u})^{'}(\tau), \end{equation*}

    which yields

    \begin{equation*} (\Phi_{u})^{'}(\tau) = (1+\nu)(J_{u})^{'}(\tau) +\nu \tau (J_{u})^{''}(\tau). \end{equation*}

    Together with (4.6), one gets

    \begin{equation*} \begin{aligned} 0 = (\Phi_{u})'(1) = (1+\nu)(J_{u})^{'}(1) +\nu (J_{u})^{''}(1) = (1+\nu) \mathcal{P}(u) +\nu (J_{u})^{''}(1) = \nu (J_{u})^{''}(1). \end{aligned} \end{equation*}

    According to (J_{u})^{''}(1)\neq 0 , we derive that \nu = 0 .

    Lemma 4.4. Let 1\leqslant N < 4s , s\in (0, 1) , and suppose that f satisfies (H_{1}) (H_{3}) . Then if \omega\leqslant 0 , Eq (\mathcal{A}) has no nontrivial solution.

    Proof. Assume by contradiction that u\in H^{s}(\mathbb{R}^{N}) is a nontrivial solution of Eq (\mathcal{A}) with \omega\leqslant0 . It follows from (4.1) and (4.2), \omega\leqslant0 , and (H_{2}) that

    \begin{equation*} 0 \geqslant s\omega \|u\|_{L^{2}(\mathbb{R}^{N})}^{2} = \int_{\mathbb{\mathbb{R}^{N}}} \left[ NF(u)-\frac{N-2s}{2}f(u)u \right] \mathrm{d}x \geqslant 0. \end{equation*}

    Therefore, we derive that \omega = 0 and

    \begin{equation*} \int_{\mathbb{\mathbb{R}^{N}}} F(u) \mathrm{d}x = \int_{\mathbb{\mathbb{R}^{N}}} f(u)u \mathrm{d}x = \int_{\mathbb{\mathbb{R}^{N}}} \widetilde{F}(u) \mathrm{d}x = 0. \end{equation*}

    For \omega = 0 , by Lemma 4.1, we deduce that u\in \mathcal{M} , that is, (4.3) holds. This yields

    \begin{equation*} \|u\|_{D^{s, 2}(\mathbb{R}^{N})} = 0, \end{equation*}

    which contradicts u\not\equiv 0 in H^{s}(\mathbb{R}^{N}) .

    Lemma 4.5. Let 1\leqslant N < 4s , s\in (0, 1) , and suppose that f satisfies (H_{1}) (H_{3}) . Then, for any c > 0 fixed, there exists a \sigma_{c} > 0 such that

    \begin{equation*} \inf\{\tau > 0: \exists u\in \mathcal{S}_{c} \; with\; \|u\|_{D^{s, 2}(\mathbb{R}^{N})} = 1 \; such\; that\; u_{\tau}\in \mathcal{M}_{c}\} \geqslant \sigma_{c}, \end{equation*}

    i.e.,

    \begin{equation*} \inf\{\|u\|_{D^{s, 2}(\mathbb{R}^{N})}: u\in \mathcal{M}_{c}\} \geqslant \sigma_{c}. \end{equation*}

    Proof. Since u_{\tau}\in \mathcal{M}_{c} , we obtain that \mathcal{P}(u_{\tau}) = 0 . According to (4.3), we derive that

    \begin{equation*} sa \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +sb\tau^{2s} \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} = N\tau^{-N-2s}\int_{\mathbb{ \mathbb{R}}^{N}} \widetilde{F}(\tau^{\frac{N}{2}}u(x)) \mathrm{d}x. \end{equation*}

    Together with (H_{2}) , for \|u\|_{D^{s, 2}(\mathbb{R}^{N})} = 1 , we get that

    \begin{equation} sa < sa+sb\tau^{2s} \leqslant N\left( \frac{1}{2} -\frac{1}{\gamma} \right) \tau^{-N-2s} \int_{\mathbb{R}^{N}} f(\tau^{\frac{N}{2}}u(x)) \tau^{\frac{N}{2}}u(x) \mathrm{d}x. \end{equation} (4.7)

    From (H_{2}) , there exists some C > 0 such that

    \begin{equation*} f(t)t\leqslant C(|t|^{\lambda} +|t|^{\gamma}), \; \; \forall t\in \mathbb{R}. \end{equation*}

    Note that for u\in \mathcal{M}_{c} with \|u\|_{D^{s, 2}(\mathbb{R}^{N})} = 1 , by Lemma 1.1, there exists C > 0 such that

    \begin{equation} \|u\|^{\lambda}_{L^{\lambda}(\mathbb{R}^{N})}\leq C, \; \; \|u\|^{\gamma}_{L^{\gamma}(\mathbb{R}^{N})}\leq C. \end{equation} (4.8)

    Combining (4.7) and (4.8), we deduce that

    \begin{equation*} sa < CN\left( \frac{1}{2} -\frac{1}{\gamma} \right) \left( \tau^{\frac{N\lambda}{2}-N-2s} +\tau^{\frac{N\gamma}{2}-N-2s} \right), \end{equation*}

    which implies that \sigma_{c} > 0 .

    Lemma 4.6. Let 1\leqslant N < 4s , s\in (0, 1) , and suppose that f satisfies (H_{1}) (H_{3}) . Then, (J_{u})^{''}(1) < 0 , for each u\in\mathcal{M}_{c} , and \mathcal{M}_{c} is a natural constraint of J|_{\mathcal{S}_{c}} .

    Proof. From (4.4), by a direct caculation, we derive that

    \begin{equation*} \begin{aligned} (J_{u})^{''}(\tau) = &as(2s-1) \tau^{2s-2} \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +bs(4s-1) \tau^{4s-2} \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{4}\\ &+N(N+1) \tau^{-N-2} \int_{\mathbb{R}^{N}} \widetilde{F} (\tau^{\frac{N}{2}}u(x)) \mathrm{d}x -\frac{N^{2}}{2} \tau^{-\frac{N}{2}-2} \int_{\mathbb{R}^{N}} \widetilde{F}^{'} (\tau^{\frac{N}{2}}u(x)) u(x) \mathrm{d}x. \end{aligned} \end{equation*}

    Then,

    \begin{equation*} (J_{u})^{''}(1) = as(2s-1) \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +bs(4s-1) \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{4}+N(N+1) \int_{\mathbb{R}^{N}} \widetilde{F} (u) \mathrm{d}x -\frac{N^{2}}{2} \int_{\mathbb{R}^{N}} \widetilde{F}^{'} (u)u \mathrm{d}x. \end{equation*}

    Together with (H_{3}) and \mathcal{P}(u) = 0 , we conclude that

    \begin{equation*} \begin{aligned} (J_{u})^{''}(1) \leqslant &as(2s-1) \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +bs(4s-1) \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{4}+N(N+1) \int_{\mathbb{R}^{N}} \widetilde{F} (u) \mathrm{d}x -\frac{N^{2}\lambda}{2} \int_{\mathbb{R}^{N}} \widetilde{F} (u) \mathrm{d}x\\ = & as(2s-1) \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +bs(4s-1) \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{4}+\left( N+1-\frac{N\lambda}{2} \right) \left( sa\|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +sb\|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} \right)\\ = & \left( N+2s-\frac{N\lambda}{2} \right) sa\|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +\left( N+4s-\frac{N\lambda}{2} \right) sb\|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{4}. \end{aligned} \end{equation*}

    According to \lambda > 2+\frac{8s}{N} and Lemma 4.5, one can see that

    \begin{equation*} (J_{u})^{''}(1) \leqslant \left( N+2s-\frac{N\lambda}{2} \right) sa\sigma_{c}^{2} +\left( N+4s-\frac{N\lambda}{2} \right) sb\sigma_{c}^{4} < 0. \end{equation*}

    Recalling Lemma 4.3, we conclude that \mathcal{M}_{c} is a natrual contraint of J|_{\mathcal{S}_{c}} .

    Corollary 4.1. Let 1\leqslant N < 4s , s\in (0, 1) , and suppose that f satisfies (H_{1}) (H_{3}) . Then, for each u\in H^{s}(\mathbb{R}^{N}) \setminus \{0\} , there exists a unique \tau_{u} > 0 such that u_{\tau_{u}}\in \mathcal{M} . Moreover,

    \begin{equation*} J(u_{\tau_{u}}) = \max\limits_{\tau > 0} J(u_{\tau}). \end{equation*}

    Proof. Set c: = \|u\|^{2}_{L^{2}(\mathbb{R}^{N})} . From (H_{1}) and (H_{2}) , we observe that for each \tau\geqslant 0 and t\in \mathbb{R} ,

    \begin{equation} \begin{cases} \tau^{\gamma} F(t) \leqslant F(\tau t) \leqslant \tau^{\lambda} F(t), \; \; &{\rm if}\; \; \tau\leqslant 1, \\ \tau^{\lambda} F(t) \leqslant F(\tau t) \leqslant \tau^{\gamma} F(t), \; \; &{\rm if}\; \; \tau\geqslant 1. \end{cases} \end{equation} (4.9)

    Hence, we obtain

    \begin{equation} \frac{\lambda-2}{\gamma-2} \min\{\tau^{\lambda}, \tau^{\gamma}\} \widetilde{F}(t) \leqslant \widetilde{F}(\tau t) \leqslant \frac{\gamma-2}{\lambda-2} \max\{\tau^{\lambda}, \tau^{\gamma}\} \widetilde{F}(t), \end{equation} (4.10)

    which implies that

    \begin{equation*} \tau^{-N} \int_{\mathbb{R}^{N}} \widetilde{F}(\tau^{\frac{N}{2}}u) \mathrm{d}x = o(\tau^{4s}) \; \; {\rm as}\; \; \tau\to 0^{+}, \end{equation*}

    noting that \lambda > 2+\frac{8s}{N} . Recalling that

    \begin{equation*} \mathcal{P}(u_{\tau}) = sa\tau^{2s} \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +sb\tau^{4s} \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} -N\tau^{-N}\int_{\mathbb{ \mathbb{R}}^{N}} \widetilde{F}(\tau^{\frac{N}{2}}u(x)) \mathrm{d}x, \end{equation*}

    we can deduce that \mathcal{P}(u_{\tau}) > 0 for \tau > 0 small enough. Meanwhile, according to (4.4), one can see that (J_{u})^{'}(\tau) > 0 for \tau > 0 small enough. Then, there exists a \tau_{0} > 0 such that J_{u}(\tau) is increasing in \tau\in (0, \tau_{0}) .

    On the other hand, from \int_{\mathbb{R}^{N}} F(u) \mathrm{d}x > 0 , (4.9), and \lambda > 2+\frac{8s}{N} , we derive that

    \begin{equation*} \tau^{-N-4s} \int_{\mathbb{R}^{N}} F(\tau^{\frac{N}{2}}u) \mathrm{d}x \geqslant \tau^{\frac{\lambda N}{2}-N-4s} \int_{\mathbb{R}^{N}} F(u) \mathrm{d}x \to+\infty, \; \; {\rm as}\; \; \tau\to+\infty, \end{equation*}

    which yields that

    \begin{equation*} \lim\limits_{\tau\to+\infty} J_{u}(\tau) = \lim\limits_{\tau\to+\infty} \tau^{4s} \left( \frac{a\tau^{-2s}}{2} \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +\frac{b}{4} \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} - \tau^{-N-4s} \int_{\mathbb{R}^{N}} F(\tau^{\frac{N}{2}}u) \mathrm{d}x \right) = -\infty. \end{equation*}

    Hence, there exist some \tau_{1} > \tau_{0} such that

    \begin{equation*} J_{u}(\tau_{1}) = \max\limits_{\tau > 0} J(u_{\tau}), \end{equation*}

    and (J_{u})^{'}(\tau_{1}) = 0 . Then, by Lemma 4.2, we conclude that u_{\tau_{1}}\in \mathcal{M} . Next, we will prove that \tau_{1} is unique. Assume by contradiction that there exists \tau_{2} > 0 such that u_{\tau_{2}}\in \mathcal{M} . From Lemma 4.6, we know that \tau_{1} and \tau_{2} are strict local maximum points of J_{u}(\tau) . Without loss of generality, we suppose that \tau_{1} < \tau_{2} . Thus, there exist some \tau_{3}\in (\tau_{1}, \tau_{2}) such that

    \begin{equation*} J_{u}(\tau_{3}) = \min\limits_{\tau\in[\tau_{1}, \tau_{2}]} J(u_{\tau}), \end{equation*}

    which indicates that \tau_{3} is a local minimum point of J_{u}(\tau) . Then, (J_{u})^{'}(\tau_{1}) = 0 and u_{\tau_{3}}\in \mathcal{M} with (J_{u_{\tau_{3}}})^{''}(1) = (J_{u})^{''}(\tau_{3}) \geqslant 0 , which contradicts Lemma 4.6.

    Corollary 4.2. Let 1\leqslant N < 4s , s\in (0, 1) , and suppose that f satisfies (H_{1}) (H_{3}) . For each u\in H^{s}(\mathbb{R}^{N})\setminus \{0\} , let \tau_{u} be given in Corollary 4.1. Then,

    \begin{equation*} \begin{cases} &\tau_{u} = 1 \Leftrightarrow (J_{u})'(1) = 0 \Leftrightarrow \mathcal{P}(u) = 0, \\ &\tau_{u} > 1 \Leftrightarrow (J_{u})'(1) > 0 \Leftrightarrow \mathcal{P}(u) > 0, \\ & \tau_{u} < 1 \Leftrightarrow (J_{u})'(1) < 0 \Leftrightarrow \mathcal{P}(u) < 0. \end{cases} \end{equation*}

    Proof. By Corollary 4.1, we obtain that

    \begin{equation*} J_{u}(\tau_{u}) = \max\limits_{\tau > 0} J_{u}(\tau). \end{equation*}

    Moreover,

    \begin{equation*} (J_{u})^{'}(\tau) > 0, \; \; {\rm for}\; 0 < \tau < \tau_{u}, \; \; {\rm and}\; \; (J_{u})^{'}(\tau) < 0, \; \; {\rm for}\; \tau > \tau_{u}. \end{equation*}

    The conclusion follows from (4.4).

    Lemma 4.7. Let 1\leqslant N < 4s , s\in (0, 1) , and suppose that f satisfies (H_{1}) (H_{3}) . Then J|_{\mathcal{M}_{c}} is coercive.

    Proof. For u\in \mathcal{M}_{c} , from (H_{2}) , we observe that

    \begin{equation} sa\|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +sb\|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} = N \int_{\mathbb{R}^{N}} \widetilde{F}(u) \mathrm{d}x \geqslant \frac{N(\lambda-2)}{2} \int_{\mathbb{R}^{N}} F(u) \mathrm{d}x. \end{equation} (4.11)

    Therefore,

    \begin{equation} \begin{aligned} J(u) = &\frac{a}{2} \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +\frac{b}{4} \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} -\int_{\mathbb{R}^{N}} F(u) \mathrm{d}x\\ \geqslant& \frac{a}{2} \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +\frac{b}{4} \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} -\frac{2s}{N(\lambda-2)} \left( a \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +b \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} \right)\\ = & \frac{N(\lambda-2)-4s}{2N(\lambda-2)} a \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{2}+ \frac{N(\lambda-2)-8s}{4N(\lambda-2)} b \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{4}. \end{aligned} \end{equation} (4.12)

    Thanks to \lambda > 2+\frac{8s}{N} , we complete the proof.

    For given c > 0 , let us define

    \begin{equation*} m_{c}: = \inf\limits_{u\in \mathcal{M}_{c}} J(u) = \inf\limits_{u\in\mathcal{S}_{c}} \max\limits_{\tau > 0} J(u_{\tau}). \end{equation*}

    Since u is a solution to (\mathcal{A}) satisfying (1.1), u must belong to \mathcal{M}_{c} . If u attains m_{c} , we can assert that u is the least energy solution, i.e., ground state solution. It follows from (4.12) and Lemma 4.5 that m_{c}>0 for each c>0 .

    For each u\in H^{s}(\mathbb{R}^{N}) , let u^{*} be the symmetric radial decreasing rearrangement of u . By (H_{1}) , without loss of generality, we assume that u is nonnegative. Then, we can obtain that

    \begin{equation*} \begin{aligned} \int_{\mathbb{R}^{N}} F(u) \mathrm{d}x = & \int_{\mathbb{R}^{N}} \left( \int_{0}^{u(x)} f(t) \mathrm{d}t \right) \mathrm{d}x = \int_{0}^{\infty} f(t)|\{x: u(x) > t\}| \mathrm{d}t\\ = & \int_{0}^{\infty} f(t)|\{x: u^{*}(x) > t\}| \mathrm{d}t = \int_{\mathbb{R}^{N}} F(u^{*}) \mathrm{d}x. \end{aligned} \end{equation*}

    From [11, Lemma 2.3], one can see that

    \begin{equation*} \iint_{\mathbb{R}^{2N}} \frac{|u^{*}(x)-u^{*}(y)|^{2}}{|x-y|^{N+2s}} \mathrm{d}x \mathrm{d}y \leqslant \iint_{\mathbb{R}^{2N}} \frac{|u(x)-u(y)|^{2}}{|x-y|^{N+2s}} \mathrm{d}x \mathrm{d}y. \end{equation*}

    Thus, we derive that

    \begin{equation} J(u^{*}) \leqslant J(u). \end{equation} (4.13)

    Set

    \begin{equation*} \mathcal{S}_{c}^{rad} : = \mathcal{S}_{c} \cap H_{rad}^{s}(\mathbb{R}^{N}), \; \; \mathcal{M}^{rad} : = \mathcal{M} \cap H_{rad}^{s}(\mathbb{R}^{N}), \; \; \mathcal{P}_{c}^{rad} : = \mathcal{P}_{c} \cap H_{rad}^{s}(\mathbb{R}^{N}). \end{equation*}

    Define

    \begin{equation*} m_{c}^{rad} : = \inf\limits_{u\in\mathcal{S}^{rad}_{c}} \max\limits_{\tau > 0} J(u_{\tau}), \end{equation*}

    and we can obtain that

    \begin{equation*} m_{c}^{rad} = \inf\limits_{u\in\mathcal{M}^{rad}_{c}} J(u). \end{equation*}

    Moreover, we find the following.

    Lemma 4.8. Let 1\leqslant N < 4s , s\in (0, 1) , and suppose that f satisfies (H_{1}) (H_{3}) . Then,

    \begin{equation*} m_{c}^{rad} = m_{c}. \end{equation*}

    Proof. Since \mathcal{S}_{c}^{rad} \subset \mathcal{S}_{c}, it is clear that m_{c}^{rad} \geqslant m_{c}. On the other side, for each t > 0 ,

    \begin{equation*} \begin{aligned} &|\{x: u^{*}_{\tau}(x) > t\}| = |\{x: \tau^{\frac{N}{2}}u^{*}(\tau x) > t\}| = |\{y: \tau^{\frac{N}{2}}u^{*}(y) > t\}|\tau^{-N}\\ = &\tau^{-N}|\{y: u^{*}(y) > \tau^{-\frac{N}{2}}t\}| = \tau^{-N}|\{y: u(y) > \tau^{-\frac{N}{2}}t\}| = \tau^{-N}|\{y: \tau^{\frac{N}{2}}u(y) > t\}|\\ = &|\{x: \tau^{\frac{N}{2}}u(\tau x) > t\}| = |\{x: u_{\tau}(x) > t\}| = |\{x: (u_{\tau}(x))^{*} > t\}|. \end{aligned} \end{equation*}

    Hence, it holds true that

    \begin{equation*} u_{\tau}^{*} = (u_{\tau})^{*}, \; \; \forall \tau\in \mathbb{R}^{+}. \end{equation*}

    As a consequence of (4.13), for each u\in\mathcal{M}_{c} , we obtain that

    \begin{equation*} J(u_{\tau}^{*}) = J((u_{\tau})^{*}) \leqslant J(u_{\tau}) \leqslant \max\limits_{t > 0}J(u_{t}) = J(u), \; \; \forall \tau\in \mathbb{R}^{+}, \end{equation*}

    which yields that m_{c}^{rad} \leqslant m_{c} , by the arbitrary of u\in \mathcal{M}_{c} , and this ends the proof.

    Proof of Theorem 2. Thanks to Lemma 4.8, let \{u_{n}\}\subset \mathcal{M}_{c}^{rad} be such that J(u_{n})\to m_{c} > 0 . From Lemma 4.7, we obtain that \{u_{n}\} is bounded in H^{s}(\mathbb{R}^{N}) . We may suppose that up to a subsequence, u_{n}\rightharpoonup u in H^{s}(\mathbb{R}^{N}) . Then, for N = 2, 3 , using the compact embedding H_{rad}^{s}(\mathbb{R}^{N}) \hookrightarrow\hookrightarrow L^{q}(\mathbb{R}^{N}) for all q\in (2, 2_{s}^{*}) , we deduce that

    \begin{equation} \int_{\mathbb{R}^{N}} F(u_{n}) \mathrm{d}x \to \int_{\mathbb{R}^{N}} F(u) \mathrm{d}x. \end{equation} (4.14)

    For N = 1 , we may suppose that u_{n} = u_{n}^{*}, \forall n\in \mathbb{N} . Then, (4.14) also holds true.

    Now, we claim that u\neq 0 . Suppose by contradiction that u = 0 . Then, \int_{\mathbb{R}^{N}} \widetilde{F}(u_{n}) \mathrm{d}x = o_{n}(1) , and taking into account of \{u_{n}\}\subset \mathcal{M}_{c}^{rad} , we obtain that

    \begin{equation*} sa\|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +sb\|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} = o_{n}(1), \end{equation*}

    which contradicts Lemma 4.5.

    Since \{u_{n}\} is bounded in H^{s}(\mathbb{R}^{N}) , it is obvious that

    \begin{equation*} \omega_{n}: = -\frac{1}{c} \langle J'(u_{n}), u_{n} \rangle \end{equation*}

    is a bounded sequence. Particularly, applying (4.3), the definition of \widetilde{F} , and (H_{2}) , we derive that

    \begin{equation} \begin{aligned} \omega_{n}c = &\omega_{n} \|u_{n}\|_{L^{2}(\mathbb{R}^{N})}^{2} = -\langle J'(u_{n}), u_{n} \rangle \\ = &\int_{\mathbb{R}^{N}} f(u_{n})u_{n} \mathrm{d}x -a\|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} -b\|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{4}\\ = & \int_{\mathbb{R}^{N}} f(u_{n})u_{n} \mathrm{d}x -\frac{N}{s} \int_{\mathbb{R}^{N}} \widetilde{F}(u_{n}) \mathrm{d}x\\ = & \frac{N}{s} \int_{\mathbb{R}^{N}} F(u_{n}) \mathrm{d}x -\frac{N-2s}{2s} \int_{\mathbb{R}^{N}} f(u_{n})u_{n} \mathrm{d}x\\ \geqslant& \left[ \frac{N}{s}-\frac{(N-2s)\gamma}{2s} \right] \int_{\mathbb{R}^{N}} F(u_{n}) \mathrm{d}x. \end{aligned} \end{equation} (4.15)

    On the other hand, from (H_{2}) , it follows that

    \begin{equation} sa\|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +sb\|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} = N\int_{\mathbb{R}^{N}} \widetilde{F}(u_{n}) \mathrm{d}x \leqslant \frac{(\gamma-2)N}{2} \int_{\mathbb{R}^{N}} F(u_{n}) \mathrm{d}x. \end{equation} (4.16)

    Due to \gamma > 2+\frac{8s}{N} , by Lemma 4.5, there exist some \delta_{c} > 0 such that for any n\in \mathbb{N} ,

    \begin{equation*} \int_{\mathbb{R}^{N}} F(u_{n}) \mathrm{d}x \geqslant \delta_{c}. \end{equation*}

    Therefore, together with (4.15), there exist some \eta_{c} > 0 such that for any n\in\mathbb{N} ,

    \begin{equation*} \omega_{n} \geqslant \eta_{c}. \end{equation*}

    Hence, we can assume that \omega_{n}\to \omega_{c} > 0 . Up to a subsequence, if necessary, we can assume that \|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} \to A\geqslant 0 . It is easily seen that u_{c}\in H^{s}_{rad}(\mathbb{R}^{N}) is a solution of

    \begin{equation} \left(a+bA\right)(-\Delta)^{s}u +\omega_{c} u = f(u), \ \ x\in \mathbb{R}^{N}. \end{equation} (4.17)

    Then, u_{c}\in \mathcal{M}^{rad} and

    \begin{equation*} \begin{aligned} &sa\|u_{c}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +sbA\|u_{c}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} = N\int_{\mathbb{R}^{N}} \widetilde{F}(u_{c}) \mathrm{d}x\\ = & N\lim\limits_{n\to\infty} \int_{\mathbb{R}^{N}} \widetilde{F}(u_{n}) \mathrm{d}x = \lim\limits_{n\to\infty} \left( sa\|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +sb\|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} \right)\\ = & saA+sbA^{2}. \end{aligned} \end{equation*}

    Hence, s(a+bA)(A-\|u_{c}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2}) = 0 . By s > 0, a, b > 0 , we get \|u_{c}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} = A , which yields that u_{n}\to u_{c} in D^{s, 2}_{0}(\mathbb{R}^{N}) . So, (4.17) gives that u_{c} is a solution of

    \begin{equation*} \left(a+b\int_{\mathbb{R}^{N}} |(-\Delta)^{\frac{s}{2}}u|^{2} \mathrm{d}x\right)(-\Delta)^{s}u +\omega_{c} u = f(u), \; \; \ x\in \mathbb{R}^{N}. \end{equation*}

    As a consequence, we get that

    \begin{equation*} \begin{aligned} &a\|u_{c}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +b\|u_{c}\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} +\omega_{c} \|u_{c}\|_{L^{2}(\mathbb{R}^{N})}^{2} = \int_{\mathbb{R}^{N}} f(u_{c})u_{c} \mathrm{d}x = \int_{\mathbb{R}^{N}} f(u_{n})u_{n} \mathrm{d}x +o_{n}(1)\\ = &a\|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +b\|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} +\omega_{n} \|u_{n}\|_{L^{2}(\mathbb{R}^{N})}^{2}+o_{n}(1), \end{aligned} \end{equation*}

    which yields that

    \begin{equation*} \omega_{c}(c-\|u_{c}\|_{L^{2}(\mathbb{R}^{N})}^{2}) = 0. \end{equation*}

    So, u_{c}\in \mathcal{S}_{c} , and then u_{c}\in \mathcal{M}_{c} . Therefore,

    \begin{equation*} m_{c} \leqslant J(u_{c}) = \lim\limits_{n\to\infty} J(u_{n}) = m_{c}, \end{equation*}

    that is, (\omega_{c}, u_{c}) is a ground state normalized solution to Eq (\mathcal{A}). Recalling Lemma 4.4, we complete the proof.

    The aim of this section is to consider the continuity and the limit behavior of m_{c} and \omega_{c} as c\to 0^{+} as well as c\to+\infty .

    For c>0 , let (\omega_{c}, u_{c}) be the solution to (\mathcal{A}), which is given by Theorem 2. We remark that \omega_{c}>0 and u_{c}\in \mathcal{S}_{c}^{rad} satisfy

    \begin{equation*} \left(a+b\int_{\mathbb{R}^{N}} |(-\Delta)^{\frac{s}{2}}u_{c}|^{2} \mathrm{d}x\right)(-\Delta)^{s}u_{c} +\omega_{c} u_{c} = f(u_{c}), \; \; \ x\in \mathbb{R}^{N}, \end{equation*}

    and

    \begin{equation} J(u_{c}) = \frac{a}{2} \|u_{c}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +\frac{b}{4} \|u_{c}\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} -\int_{\mathbb{R}^{N}} F(u_{c}) \mathrm{d}x = m_{c}. \end{equation} (5.1)

    Lemma 5.1. Let 1\leqslant N < 4s , s\in (0, 1) , and suppose that f satisfies (H_{1}) (H_{3}) . Then, m_{c} is continuous with respect to c\in (0, +\infty) .

    Proof. Thanks to Theorem 2, we note that m_{c} is attained by a symmetric decreasing function u_{c}\in H^{s}_{rad}(\mathbb{R}^{N}) . Let c > 0 be fixed, and for each \{c_{n}\}\subset \mathbb{R}^{+} with c_{n}\to c as n\to+\infty , for the sake of simplicity, we denote (\omega_{c_{n}}, u_{c_{n}}) by (\omega_{n}, u_{n}) . Without loss of generality, we may suppose that for any n\in \mathbb{N} , \frac{c}{2} < c_{n} < 2c . Taking into account of (4.11), (4.16), and (5.1), we deduce that

    \begin{equation} \begin{aligned} &\left[ \frac{1}{2}- \frac{2s}{(\gamma-2)N} \right] a\|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +\left[ \frac{1}{4}- \frac{2s}{(\gamma-2)N} \right] b\|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{4}\\ \leqslant& m_{c_{n}} = J(u_{n}) \leqslant \left[ \frac{1}{2}- \frac{2s}{(\lambda-2)N} \right] a\|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +\left[ \frac{1}{4}- \frac{2s}{(\lambda-2)N} \right] b\|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{4}. \end{aligned} \end{equation} (5.2)

    Let u_{c/2} attain m_{c/2} . For \theta\in (1, 4), \theta u_{c/2}\in \mathcal{S}_{\theta^{2}c/2} . From Corollary 4.1, there exists a unique \tau_{\theta} > 0 such that (\theta u_{c/2})_{\tau_{\theta}}\in \mathcal{M}_{\theta^{2}c/2} and

    \begin{equation*} sa\tau_{\theta}^{2s} \|\theta u_{c/2}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +sb\tau_{\theta}^{4s} \|\theta u_{c/2}\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} = N \int_{\mathbb{ \mathbb{R}}^{N}} \widetilde{F}((\theta u_{c/2})_{\tau_{\theta}}) \mathrm{d}x. \end{equation*}

    From (4.10), we conclude that

    \begin{equation*} \begin{aligned} &sa\tau_{\theta}^{-2s} \|\theta u_{c/2}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +sb \|\theta u_{c/2}\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} = N\tau_{\theta}^{-4s} \int_{\mathbb{ \mathbb{R}}^{N}} \widetilde{F}((\theta u_{c/2})_{\tau_{\theta}}) \mathrm{d}x\\ = & N\tau_{\theta}^{-4s} \int_{\mathbb{ \mathbb{R}}^{N}} \widetilde{F} \left(\tau_{\theta}^{\frac{N}{2}}\theta u_{c/2}(\tau_{\theta}x) \right) \mathrm{d}x\\ \geqslant& C\tau_{\theta}^{-4s} \min\left\{ \left(\tau_{\theta}^{\frac{N}{2}}\theta \right)^{\lambda}, \left(\tau_{\theta}^{\frac{N}{2}}\theta \right)^{\gamma} \right\} \int_{\mathbb{ \mathbb{R}}^{N}} \widetilde{F}( u_{c/2}(\tau_{\theta}x)) \mathrm{d}x\\ = & C\tau_{\theta}^{-4s-N} \min\left\{ \left(\tau_{\theta}^{\frac{N}{2}}\theta \right)^{\lambda}, \left(\tau_{\theta}^{\frac{N}{2}}\theta \right)^{\gamma} \right\} \int_{\mathbb{ \mathbb{R}}^{N}} \widetilde{F}( u_{c/2}(y)) \mathrm{d}y, \end{aligned} \end{equation*}

    which leads to that \tau_{\theta} is bounded due to \frac{N\lambda}{2}-4s-N \geqslant \frac{N\gamma}{2}-4s-N > 0 . It easily follows from m_{\theta^{2}c/2} \leqslant J((\theta u_{c/2})_{\tau_{\theta}}) that \{m_{\theta^{2}c/2}: \theta\in (1, 4)\} is bounded. Together with (5.2), we conclude that \{u_{n}\} is bounded in H^{s}(\mathbb{R}^{N}) .

    Similar as the proof of Theorem 2, we can suppose that \omega_{n}\to \omega_{c} > 0 and u_{n}\rightharpoonup u_{c} in H^{s}(\mathbb{R}^{N}) , where u_{c}\in \mathcal{S}_{c} is a solution of

    \begin{equation*} \left(a+b\int_{\mathbb{R}^{N}} |(-\Delta)^{\frac{s}{2}}u|^{2} \mathrm{d}x\right)(-\Delta)^{s}u +\omega_{c} u = f(u), \; \; \ x\in \mathbb{R}^{N}. \end{equation*}

    Therefore,

    \begin{equation*} \lim\limits_{n\to\infty} m_{c_{n}} = \lim\limits_{n\to\infty} J(u_{n}) = J(u_{c}) \geqslant m_{c}. \end{equation*}

    If J(u_{c})\neq m_{c} , there exist some \delta > 0 such that for n large enough,

    \begin{equation} m_{c_{n}} \geqslant m_{c}+\delta. \end{equation} (5.3)

    Recall that \sqrt{\rho}u_{c}\in\mathcal{S}_{\rho c} , and let \tau_{\rho} > 0 be the unique number such that (\sqrt{\rho}u_{c})_{\tau_{\rho}} = \sqrt{\rho}(u_{c})_{\tau_{\rho}} \in \mathcal{M}_{\rho c} , that is,

    \begin{equation*} sa\rho\tau_{\rho}^{2s} \| u_{c}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +sb\rho^{2}\tau_{\rho}^{4s} \| u_{c}\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} -N \int_{\mathbb{ \mathbb{R}}^{N}} \widetilde{F}(\sqrt{\rho}(u_{c})_{\tau_{\rho}}) \mathrm{d}x = 0. \end{equation*}

    By Lemma 4.5, we can derive that \tau_{\rho} is bounded away from 0 as \rho approaches to 1. Thus, the uniqueness indicates that \tau_{\rho}\to 1 as \rho\to 1 . Therefore, for \rho close to 1 enough, we conclude that

    \begin{equation*} m_{\rho c} \leqslant J(\sqrt{\rho}(u_{c})_{\tau_{\rho}}) \to J(u_{c}) = m_{c}. \end{equation*}

    Hence, there exists N_{0}\in \mathbb{N} such that for n\geqslant N_{0} ,

    \begin{equation*} m_{c_{n}}\leqslant m_{c}+\frac{\delta}{2}, \end{equation*}

    which contradicts (5.3). Consequently, we obtain that J(u_{c}) = m_{c} and

    \begin{equation*} \lim\limits_{n\to\infty} m_{c_{n}} = m_{c}. \end{equation*}

    Lemma 5.2. Let 1\leqslant N < 4s , s\in (0, 1) , and suppose that f satisfies (H_{1}) (H_{3}) . Then,

    \begin{equation*} \lim\limits_{c\to 0^{+}} m_{c} = +\infty \; \; {\rm and}\; \; \lim\limits_{c\to +\infty} m_{c} = 0. \end{equation*}

    Proof. Using (H_{2}) and Lemma 1.1, there exists C > 0 such that

    \begin{equation*} \begin{aligned} &sa\|u_{c}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +sb\|u_{c}\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} = N\int_{\mathbb{R}^{N}} \widetilde{F}(u_{c}) \mathrm{d}x \leqslant \frac{(\gamma-2)N}{2} \int_{\mathbb{R}^{N}} F(u_{c}) \mathrm{d}x\\ \leqslant& C(\|u\|_{L^{\lambda}(\mathbb{R}^{N})}^{\lambda} +\|u\|_{L^{\gamma}(\mathbb{R}^{N})}^{\gamma}) \leqslant C\left( \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{\frac{N(\lambda-2)}{2s}} c^{\frac{\lambda}{2}-\frac{N(\lambda-2)}{4s}} +\|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{\frac{N(\gamma-2)}{2s}} c^{\frac{\gamma}{2}-\frac{N(\gamma-2)}{4s}} \right). \end{aligned} \end{equation*}

    Therefore,

    \begin{equation*} sa < sa+ sb\|u_{c}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} \leqslant C\left( \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{\frac{N(\lambda-2)-4s}{2s}} c^{\frac{\lambda}{2}-\frac{N(\lambda-2)}{4s}} +\|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{\frac{N(\gamma-2)-4s}{2s}} c^{\frac{\gamma}{2}-\frac{N(\gamma-2)}{4s}} \right), \end{equation*}

    which yields that

    \begin{equation*} \|u\|_{D^{s, 2}(\mathbb{R}^{N})} \to +\infty, \end{equation*}

    as c\to 0^{+} , since \frac{N(\lambda-2)-4s}{2s}, \frac{\lambda}{2}-\frac{N(\lambda-2)}{4s}, \frac{N(\gamma-2)-4s}{2s}, \frac{\gamma}{2}-\frac{N(\gamma-2)}{4s} > 0. Together with (4.12), we observe that

    \begin{equation*} \lim\limits_{c\to 0^{+}}m_{c} = +\infty. \end{equation*}

    For the case c\to +\infty , we choose a positive v\in H^{s}(\mathbb{R}^{N}) with \|v\|_{L^{2}(\mathbb{R}^{N})} = 1 . For c > 0 , it is easy to check that u = \sqrt{c}v\in \mathcal{S}_{c} . From Corollary 4.1, there exists a unique \tau_{c} > 0 such that u_{\tau_{c}}\in \mathcal{M}_{c} . We observe that c\tau_{c}^{2s}\to 0 as c\to+\infty . If not, there exists a sequence \{c_{n}\} with c_{n}\to+\infty as n\to+\infty and c_{n}\tau_{c_{n}}^{2s}\geqslant \delta > 0 , for all n\in \mathbb{N} . Applying (4.9), we obtain that

    \begin{equation*} \begin{aligned} &c_{n}^{-2} \tau_{c_{n}}^{-4s} \int_{ \mathbb{R}^{N}} \widetilde{F} (\sqrt{c_{n}}v_{\tau_{c_{n}}}) \mathrm{d}x = c_{n}^{-2} \tau_{c_{n}}^{-4s-N} \int_{ \mathbb{R}^{N}} \widetilde{F} \left(\sqrt{c_{n}}\tau_{c_{n}}^{\frac{N}{2}}v(y) \right) \mathrm{d}y\\ \geqslant& c_{n}^{-2} \tau_{c_{n}}^{-4s-N} \frac{\lambda-2}{\gamma-2} \min\left\{\left(\sqrt{c_{n}}\tau_{c_{n}}^{\frac{N}{2}}\right)^{\lambda}, \left(\sqrt{c_{n}}\tau_{c_{n}}^{\frac{N}{2}}\right)^{\gamma}\right\} \int_{ \mathbb{R}^{N}} \widetilde{F} (v(y)) \mathrm{d}y. \end{aligned} \end{equation*}

    According to

    \begin{equation*} c_{n}^{-2} \tau_{c_{n}}^{-4s-N} \left(\sqrt{c_{n}}\tau_{c_{n}}^{\frac{N}{2}}\right)^{\lambda} = \left( c_{n}\tau_{c_{n}}^{2s} \right)^{-2-\frac{N}{2s}+\frac{N\lambda}{4s}} c_{n}^{\frac{2N-(N-2s)\lambda}{4s}} \geqslant \delta^{-2-\frac{N}{2s}+\frac{N\lambda}{4s}} c_{n}^{\frac{2N-(N-2s)\lambda}{4s}} \to+\infty, \; \; {\rm as}\; n\to+\infty, \end{equation*}

    and

    \begin{equation*} c_{n}^{-2} \tau_{c_{n}}^{-4s-N} \left(\sqrt{c_{n}}\tau_{c_{n}}^{\frac{N}{2}}\right)^{\gamma} \to+\infty \; \; {\rm as}\; n\to+\infty, \end{equation*}

    we conclude that

    \begin{equation*} c_{n}^{-2} \tau_{c_{n}}^{-4s} \int_{ \mathbb{R}^{N}} \widetilde{F} (\sqrt{c_{n}}v_{\tau_{c_{n}}}) \mathrm{d}x \to +\infty, \; \; {\rm as}\; n\to+\infty. \end{equation*}

    Therefore,

    \begin{equation*} \begin{aligned} 0 = &\mathcal{P}(u_{\tau_{c_{n}}}) = sa\tau_{c_{n}}^{2s} \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +sb\tau_{c_{n}}^{4s} \|u\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} -N\int_{\mathbb{ \mathbb{R}}^{N}} \widetilde{F}(u_{\tau_{c_{n}}}) \mathrm{d}x\\ = & sac_{n}\tau_{c_{n}}^{2s} \|v\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +sbc_{n}^{2}\tau_{c_{n}}^{4s} \|v\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} -N\int_{\mathbb{ \mathbb{R}}^{N}} \widetilde{F}(\sqrt{c_{n}}v_{\tau_{c_{n}}}) \mathrm{d}x\\ = & c_{n}^{2} \tau_{c_{n}}^{4s} \left( sac_{n}^{-1}\tau_{c_{n}}^{-2s} \|v\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +sb \|v\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} -Nc_{n}^{-2} \tau_{c_{n}}^{-4s} \int_{\mathbb{ \mathbb{R}}^{N}} \widetilde{F}(\sqrt{c_{n}}v_{\tau_{c_{n}}}) \mathrm{d}x \right)\\ < &0, \end{aligned} \end{equation*}

    for n large enough, which is a contradiction. Hence, our claim c\tau_{c}^{2s}\to 0 as c\to+\infty holds true. Consequently, combining with w: = u_{\tau_{c}} = \sqrt{c}v_{\tau_{c}} \in \mathcal{M}_{c} , we deduce that

    \begin{equation*} \begin{aligned} m_{c} \leqslant J(w) = &\frac{a}{2} \|w\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +\frac{b}{4} \|w\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} -\int_{ \mathbb{R}^{N}} F(w) \mathrm{d}x\\ \leqslant& \frac{a}{2} \|w\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +\frac{b}{4} \|w\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} -\frac{2}{\gamma-2} \int_{ \mathbb{R}^{N}} \widetilde{F}(w) \mathrm{d}x\\ = & \frac{a}{2} \|w\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +\frac{b}{4} \|w\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} -\frac{2}{(\gamma-2)N} \left( sa\|w\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +sb\|w\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} \right)\\ = & \left( \frac{1}{2} -\frac{2s}{(\gamma-2)N} \right) a\|w\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +\left( \frac{1}{4} -\frac{2s}{(\gamma-2)N} \right) b\|w\|_{D^{s, 2}(\mathbb{R}^{N})}^{4}\\ = & \left( \frac{1}{2} -\frac{2s}{(\gamma-2)N} \right) ac\tau_{c}^{2s}\|v\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +\left( \frac{1}{4} -\frac{2s}{(\gamma-2)N} \right) b(c\tau_{c}^{2s})^{2} \|v\|_{D^{s, 2}(\mathbb{R}^{N})}^{4}\\ \to& 0, \; \; {\rm as}\; c\to +\infty. \end{aligned} \end{equation*}

    On the other hand, it follows from (4.12) and Lemma 4.5 that m_{c} > 0 for each c > 0 . This completes the proof.

    Lemma 5.3. Let 1\leqslant N < 4s , s\in (0, 1) , and suppose that f satisfies (H_{1}) (H_{3}) . Then,

    \begin{equation*} \lim\limits_{c\to 0^{+}} \omega_{c}c = +\infty \; \; {\rm and}\; \; \lim\limits_{c\to +\infty} \omega_{c}c = 0. \end{equation*}

    Proof. Combining (5.2) and Lemma 5.2, we derive that

    \begin{equation} \lim\limits_{c\to0^{+}} \|u_{c}\|_{D^{s, 2}(\mathbb{R}^{N})} = +\infty, \; \; {\rm and}\; \; \lim\limits_{c\to+\infty} \|u_{c}\|_{D^{s, 2}(\mathbb{R}^{N})} = 0. \end{equation} (5.4)

    Moreover, from \mathcal{P}(u_{c}) = 0 , we conclude that

    \begin{equation*} \int_{ \mathbb{R}^{N}} \widetilde{F}(u_{c}) \mathrm{d}x = \frac{1}{N} \left( as\|u_{c}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +bs\|u_{c}\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} \right). \end{equation*}

    Together with (5.4), we see that

    \begin{equation} \lim\limits_{c\to0^{+}} \int_{ \mathbb{R}^{N}} \widetilde{F}(u_{c}) \mathrm{d}x = +\infty, \; \; {\rm and}\; \; \lim\limits_{c\to+\infty} \int_{ \mathbb{R}^{N}} \widetilde{F}(u_{c}) \mathrm{d}x = 0. \end{equation} (5.5)

    Similar as (4.15), we derive that

    \begin{equation*} \begin{aligned} &\omega_{c}c = \omega_{c} \|u_{c}\|_{L^{2}(\mathbb{R}^{N})}^{2} = & \frac{N}{s} \int_{\mathbb{R}^{N}} F(u_{c}) \mathrm{d}x -\frac{N-2s}{2s} \int_{\mathbb{R}^{N}} f(u_{c})u_{c} \mathrm{d}x. \end{aligned} \end{equation*}

    By (H_{2}) , there exist C_{1}, C_{2} > 0 such that

    \begin{equation*} C_{1} \int_{ \mathbb{R}^{N}} \widetilde{F}(u_{c}) \mathrm{d}x \leqslant \omega_{c}c \leqslant C_{2} \int_{ \mathbb{R}^{N}} \widetilde{F}(u_{c}) \mathrm{d}x, \end{equation*}

    which clearly means

    \begin{equation*} \lim\limits_{c\to 0^{+}} \omega_{c}c = +\infty \; \; {\rm and}\; \; \lim\limits_{c\to +\infty} \omega_{c}c = 0, \end{equation*}

    recalling (5.5).

    Remark 3. For the two quantities \zeta(c), \varphi(c) , if there exist C_{1}, C_{2} > 0 independent of c such that

    \begin{equation*} C_{1}\zeta(c) \leqslant \varphi(c) \leqslant C_{2}\zeta(c), \end{equation*}

    we say \zeta(c) and \varphi(c) are comparable. Thus, by the proofs above in Section 5, it is easy to see that any two elements in the set

    \begin{equation*} \left\{ m_{c}, \omega_{c}c, \|u_{c}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} +\|u_{c}\|_{D^{s, 2}(\mathbb{R}^{N})}^{4} \int_{\mathbb{R}^{N}} F(u_{c}) \mathrm{d}x, \int_{\mathbb{R}^{N}} \widetilde{F}(u_{c}) \mathrm{d}x, \int_{\mathbb{R}^{N}} f(u_{c})u_{c} \mathrm{d}x \right\}, \end{equation*}

    are comparable.

    Let c_{n}\to+\infty . By Lemma 5.3, we know that \omega_{c_{n}}\to 0 . For convenience, we denote (\omega_{c_{n}}, u_{c_{n}}) as (\omega_{n}, c_{n}) . Set

    \begin{equation*} e_{n}: = \|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2}. \end{equation*}

    From (5.4), one sees that e_{n}\to 0 as n\to +\infty .

    Lemma 5.4. Let 1\leqslant N < 4s , s\in (\frac{1}{2}, 1) , and suppose that f satisfies (H_{1}) (H_{5}) . Let (\omega_{c}, u_{c}) be the solution given by Theorem 2. Then,

    \begin{equation*} \limsup\limits_{c\to+\infty} u_{c}(0) < +\infty. \end{equation*}

    Proof. Let us argue by contradiction. Assume that there exists a sequence \{c_{n}\}\to +\infty such that

    \begin{equation*} m_{n}: = u_{n}(0) = \max\limits_{x\in \mathbb{R}^{N}} u_{n}(x)\to+\infty. \end{equation*}

    Take x = y/m_{n}^{\frac{\gamma-2}{2s}} and define

    \begin{equation*} Q_{n}(y) = \frac{u_{n}\left( y/m_{n}^{\frac{\gamma-2}{2s}} \right)}{m_{n}}, \; \; y\in\mathbb{R}^{N}. \end{equation*}

    Thus, \max_{y\in \mathbb{R}^{N}} Q_{n}(y) = 1 and

    \begin{equation} (-\Delta)^{s}Q_{n}(y) = \frac{1}{a+be_{n}} \left[ \frac{f(m_{n}Q_{n}(y))}{m_{n}^{\gamma-1}} -\frac{\omega_{n}}{m_{n}^{\gamma-2}}Q_{n}(y) \right]. \end{equation} (5.6)

    It follows from (H_{1}) and (H_{2}) that there exists a C > 0 such that

    \begin{equation*} f(t)\leqslant C(t^{\gamma-1}+t^{\lambda-1}). \end{equation*}

    Therefore, we note that

    \begin{equation*} \frac{1}{a+be_{n}} \left[ \frac{f(m_{n}Q_{n}(y))}{m_{n}^{\gamma-1}} -\frac{\omega_{n}}{m_{n}^{\gamma-2}}Q_{n}(y) \right]\in L^{\infty}(\mathbb{R}^{N}). \end{equation*}

    Then, applying a similar argument to the proof of [19, Proposition 4.4], and passing to a subsequence if necessary, Q_{n}\to Q in C_{loc}^{2, \alpha}(\mathbb{R}^{N}) , for some \alpha\in (0, 1) . It is easy to see that Q satisfies, in weak sense,

    \begin{equation*} (-\Delta)^{s}Q = \frac{\mu_{2}}{a}Q^{\gamma-1}, \; \; {\rm in}\; \; \mathbb{R}^{N}. \end{equation*}

    According to [6, Theorem 1.5], we derive that Q = 0 , which contradicts Q(0) = 1 .

    Define

    \begin{equation*} \tilde{u}_{n}(x) : = \frac{1}{u_{n}(0)} u_{n} \left( \frac{x}{\omega_{n}^{\frac{1}{2s}}} \right). \end{equation*}

    By direct calculation, we see that \tilde{u}_{n}(0) = \|\tilde{u}_{n}\|_{L^{\infty}(\mathbb{R}^{N})} = 1 and

    \begin{equation} (a+be_{n}) (-\Delta)^{s} \tilde{u}_{n}(x) +\tilde{u}_{n}(x) = \frac{1}{\omega_{n}u_{n}(0)} f(u_{n}(0)\tilde{u}_{n}(x)). \end{equation} (5.7)

    Lemma 5.5. Let 1\leqslant N < 4s , s\in (\frac{1}{2}, 1) , and suppose that f satisfies (H_{1}) (H_{5}) . Let (\omega_{c}, u_{c}) be the solution given by Theorem 2. Then,

    \begin{equation*} \liminf\limits_{c\to+\infty} \frac{[u_{c}(0)]^{\lambda-2}}{ \omega_{c}} > 0. \end{equation*}

    Proof. We assume by contradiction that there exists a sequence c_{n}\to+\infty such that

    \begin{equation*} [u_{n}(0)]^{\lambda-2} = o_{n}(\omega_{n}). \end{equation*}

    Letting x = 0 in (5.7), one sees that

    \begin{equation*} \begin{aligned} 1 = &\tilde{u}_{n}(0) \leqslant (a+be_{n}) (-\Delta)^{s} \tilde{u}_{n}(0) +\tilde{u}_{n}(0) = \frac{1}{\omega_{n}u_{n}(0)} f(u_{n}(0)\tilde{u}(0))\\ \leqslant& \frac{C}{\omega_{n}u_{n}(0)} \left( [u_{n}(0)]^{\lambda-1} +[u_{n}(0)]^{\gamma-1} \right) \leqslant \frac{C[u_{n}(0)]^{\lambda-2}}{\omega_{n}} = o_{n}(1), \end{aligned} \end{equation*}

    which is a contradiction.

    Lemma 5.6. Let 1\leqslant N < 4s , s\in (\frac{1}{2}, 1) , and suppose that f satisfies (H_{1}) (H_{5}) . Let (\omega_{c}, u_{c}) be the solution given by Theorem 2. Then, u_{c}(0) = \|u_{c}\|_{L^{\infty}(\mathbb{R}^{N})} \to 0 , as c\to+\infty .

    Proof. Recalling that \omega_{c}\to 0^{+} and \|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}\to 0 as c\to+\infty , assume by contradiction that \liminf_{n\to+\infty}u_{n}(0) > 0 . According to f(u_{n})-\omega_{n}u_{n}\in L^{\infty}(\mathbb{R}^{N}) , applying a similar argument to the proof of [19, Proposition 4.4], and passing to a subsequence if necessary, we assume that u_{n}\to u_{c} in C_{loc}^{2, \alpha}(\mathbb{R}^{N}) , for some \alpha\in (0, 1) with u_{c}(0) = \max_{x\in \mathbb{R}^{N}}u_{c}(x) > 0 , and u_{c} is a nonnegative bounded radial solution of

    \begin{equation*} (-\Delta)^{s}u = \frac{1}{a}f(u) \geqslant 0 \; \; {\rm in}\; \; \mathbb{R}^{N}. \end{equation*}

    Then, by [6, Theorem 1.1], we derive that u_{c}\equiv0 , which contradicts u_{c}(0) > 0 .

    Lemma 5.7. Let 1\leqslant N < 4s , s\in (\frac{1}{2}, 1) , and suppose that f satisfies (H_{1}) (H_{5}) . Let (\omega_{c}, u_{c}) be the solution given by Theorem 2. Then,

    \begin{equation*} \limsup\limits_{c\to+\infty} \frac{[u_{c}(0)]^{\lambda-2}}{\omega_{c}} < +\infty. \end{equation*}

    Proof. Assume by contradiction that there exists a sequence c_{n}\to+\infty such that

    \begin{equation*} \frac{[u_{n}(0)]^{\lambda-2}}{\omega_{n}} \to+\infty. \end{equation*}

    Thus,

    \begin{equation} \lim\limits_{n\to+\infty} \frac{\omega_{n}}{[u_{n}(0)]^{\lambda-2}} = 0. \end{equation} (5.8)

    Set

    \begin{equation*} \hat{u}_{n}(x) : = \frac{1}{u_{n}(0)} u_{n} \left( \frac{x}{[u_{n}(0)]^{\frac{\lambda-2}{2s}}} \right). \end{equation*}

    A direct calculation shows that \hat{u}_{n}(0) = \|\hat{u}_{n}\|_{L^{\infty}(\mathbb{R}^{N})} = 1 and

    \begin{equation} (a+be_{n}) (-\Delta)^{s} \hat{u}_{n} = \frac{f(u_{n}(0)\hat{u})}{[u_{n}(0)]^{\lambda-1}} -\frac{\omega_{n}}{[u_{n}(0)]^{\lambda-2}}\hat{u}_{n}. \end{equation} (5.9)

    From (H_{2}) , Lemma 5.4 and (5.8), we derive that the right side of (5.9) is of L^{\infty}(\mathbb{R}^{N}) . Therefore, applying a similar argument to the proof of [19, Proposition 4.4] and passing to a subsequence if necessary, we assume that \hat{u}_{n}\to \hat{u}_{c} in C_{loc}^{2, \alpha}(\mathbb{R}^{N}) , for some \alpha\in (0, 1) . Combining Lemma 5.6 and (H_{4}) , noting that e_{n}\to 0 , we deduce that \hat{u}_{c} is a nonnegative bounded radial solution of

    \begin{equation*} (-\Delta)^{s}\hat{u}_{c} = \frac{\mu_{1}}{a}\hat{u}_{c}^{\lambda-1}, \; \; {\rm in}\; \; \mathbb{R}^{N}. \end{equation*}

    Thanks to [6, Theorem 1.5], we derive that \hat{u}_{c} = 0 , which contradicts \hat{u}_{c}(0) = 1 .

    Lemma 5.8. Let 1\leqslant N < 4s , s\in (\frac{1}{2}, 1) , and suppose that f satisfies (H_{1}) (H_{5}) . Let (\omega_{c}, u_{c}) be the solution given by Theorem 2. Then, \tilde{u}_{n}\to 0 as |x|\to+\infty uniformly for large n\in \mathbb{N} .

    Proof. (5.7) can be rewritten as

    (-\Delta)^s\tilde{u}_{n} +\tilde{u}_{n} = g_n(x)\quad{\rm \in}\ \mathbb{R}^N,

    where g_n(x) = -\frac{\tilde{u}_{n}(x)}{(a+be_{n})} +\frac{f(u_{n}(0)\tilde{u}_{n}(x))}{\omega_{n}u_{n}(0)(a+be_{n})} . By Lemma 5.4, it is clear that g_n\in L^\infty(\mathbb{R}^N) . Moreover, it is uniformly bounded for sufficiently large n . Due to the fact \{u_n\} converges strongly in H^s(\mathbb{R}^N) , the interpolation inequality yields that there exists g\in L^r(\mathbb{R}^N) such that g_n\rightarrow g in L^r(\mathbb{R}^N) for r\in [2, +\infty) . Thus, by [3], we observe that

    \tilde{u}_{n}(x) = \int_{ \mathbb{R}^N} K(x-y)g_n(y) \mathrm{d}y,

    where K is a Bessel potential and it satisfies the following properties:

    (D_1) K is positive, radially symmetric, and smooth in \mathbb{R}^N\setminus\{0\} .

    (D_2) There exists C > 0 such that K(x)\leq \frac{C}{|x|^{N+2s}} for x\in \mathbb{R}^N\setminus\{0\} .

    (D_3) K\in L^{r}(\mathbb{R}^N) for r\in [1, \frac{1}{1-s}) . For any \sigma > 0 , we see that

    0\leqslant \tilde{u}_{n}(x) \leqslant \int_{ \mathbb{R}^N} K(x-y)|g_n(y)| \mathrm{d}y = \int_{\{|x-y|\geqslant \frac{1}{\sigma}\}} K(x-y)|g_n(y)| \mathrm{d}y +\int_{\{|x-y| < \frac{1}{\sigma}\}} K(x-y)|g_n(y)| \mathrm{d}y.

    It follows from (D_2) that

    \begin{equation} \int_{\{|x-y| \geqslant \frac{1}{\sigma}\}} K(x-y)|g_n(y)| \mathrm{d}y \leqslant C\|g_n\|_\infty \int_{\{|x-y| \geqslant \frac{1}{\sigma}\}} \frac{1}{|x-y|^{N+2s}} \mathrm{d}y \leqslant C\sigma^{2s}. \end{equation} (5.10)

    By Hölder's inequality and (D_3) , we obtain

    \begin{equation*} \begin{aligned} &\int_{\{|x-y| < \frac{1}{\sigma}\}} K(x-y)|g_n(y)| \mathrm{d}y\\ &\leqslant \int_{\{|x-y| < \frac{1}{\sigma}\}} K(x-y)|g_n(y)-g| \mathrm{d}y+ \int_{\{|x-y| < \frac{1}{\sigma}\}} K(x-y)|g(y)| \mathrm{d}y\\ &\leqslant \left(\int_{ \mathbb{R}^N} |K|^2 \mathrm{d}y \right)^{\frac{1}{2}} \left( \int_{ \mathbb{R}^N} |g_n-g|^2 \mathrm{d}y \right)^{\frac{1}{2}}+ \left( \int_{ \mathbb{R}^N} |K|^2 \mathrm{d}y \right)^{\frac{1}{2}} \left( \int_{\{|x-y| < \frac{1}{\sigma}\}} |g|^2 \mathrm{d}y \right)^{\frac{1}{2}}. \end{aligned} \end{equation*}

    This yields that there exists R_1 > 0 independent of \sigma > 0 such that

    \begin{equation} \int_{|x-y| < \frac{1}{\sigma}} K(x-y)|g_n(y)| \mathrm{d}y \leqslant \sigma\quad {\rm uniformly\ for\ large\ } n\ {\rm and\ } |x|\geq R_1. \end{equation} (5.11)

    Here, we use the fact 2 < \frac{1}{1-s} and \left(\int_{\{|x-y| < \frac{1}{\sigma}\}} |g|^2 \mathrm{d}y\right)^{\frac{1}{2}} \rightarrow 0 as |x|\rightarrow +\infty . Combining (5.10) and (5.11), we obtain

    0\leq \tilde{u}_{n}(x) \leqslant C(\sigma^{2s}+\sigma^s), \quad {\rm as}\ |x|\geq R_1\ {\rm uniformly\ for\ large\ n, }

    and, thus, \tilde{u}_{n}(x)\rightarrow 0 as |x|\rightarrow +\infty uniformly for large n .

    Proof of Theorem 3. For each sequence \{c_{n}\}\to+\infty , we define

    \begin{equation*} v_{n}(x): = \omega_{n}^{\frac{1}{2-\lambda}} u_{n} \left( \frac{x}{\omega_{n}^{\frac{1}{2s}}} \right) = \frac{u_{n}(0)}{\omega_{n}^{1/(\lambda-2)}} \tilde{u}_{n}. \end{equation*}

    According to Lemma 5.7, we observe that

    \begin{equation} \limsup\limits_{n\to+\infty} \sup\limits_{x\in\mathbb{R}^{N}} v_{n}(x) = \limsup\limits_{n\to+\infty} \omega_{n}^{\frac{1}{2-\lambda}} u_{n}(0) < +\infty. \end{equation} (5.12)

    Together with Lemmas 5.5, 5.7, and 5.8, we deserve that \lim_{|x|\to+\infty} v_{n}(x) = 0 uniformly for large n . Moreover, v_{n} is the solution of

    \begin{equation} (a+be_{n}) (-\Delta)^{s} v_{n}(x) +v_{n}(x) = \frac{f(\omega_{n}^{\frac{1}{\lambda-2}}v_{n})}{\omega_{n}^{\frac{\lambda-1}{\lambda-2}}}. \end{equation} (5.13)

    Hence, a similar argument to the proof of [19, Proposition 4.4] implies that v_{n}\to Q in C_{loc}^{2, \alpha}(\mathbb{R}^{N}) for some \alpha\in (0, 1) . Recalling that \omega_{n}\to 0^{+} and \|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}\to 0 as n\to+\infty , from (H_{4}) , we obtain that Q is nontrivial nonnegative solution of

    \begin{equation*} \begin{cases} a(-\Delta)^{s}Q+Q = \mu_{1}Q^{\lambda-1}, \; \; {in}\; \mathbb{R}^{N}, \\ \lim\nolimits_{|x|\to+\infty}Q(x) = 0. \end{cases} \end{equation*}

    From [4], we know that Q is a radial, positive, and strictly decreasing in x .

    Proof of Theorem 4. From (5.13), we have

    \begin{equation} (-\Delta)^{s} v_{n}(x) +\frac{1}{a+b\|u_{n}\| ^{2}_{D^{s, 2}(\mathbb{R}^{N})}} \left[ 1-\frac{f(\omega_{n}^{\frac{1}{\lambda-2}}v_{n})}{\omega_{n}^{\frac{\lambda-1}{\lambda-2}}v_{n}} \right] v_{n} = 0. \end{equation} (5.14)

    Since (5.12) and \omega_{n}\to 0^{+} , by (H_{4}) , we observe that

    \begin{equation*} f(\omega_{n}^{\frac{1}{\lambda-2}}v_{n}) = (\mu_{1}+o_{n}(1)) \omega_{n}^{\frac{\lambda-1}{\lambda-2}} [v_{n}(x)]^{\lambda-1}, \end{equation*}

    as n\to+\infty . Together with \|u_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}\to 0 as n\to+\infty , it follows from (5.14) that

    \begin{equation} (-\Delta)^{s} v_{n}(x) +\left(\frac{1}{a}+o_{n}(1)\right) \left[ 1-(\mu_{1}+o_{n}(1)) [v_{n}(x)]^{\lambda-2} \right] v_{n} = 0. \end{equation} (5.15)

    Noting that \lim_{|x|\to+\infty}v_{n}(x) = 0 uniformly in n\in\mathbb{N} , there exist R > 0 large enough and N_{0}\in \mathbb{N} such that

    \begin{equation*} \left(\frac{1}{a}+o_{n}(1)\right) \left[ 1-(\mu_{1}+o_{n}(1)) [v_{n}(x)]^{\lambda-2} \right] > \frac{1}{2a}, \end{equation*}

    for |x| > R and n\geqslant N_{0} . Therefore,

    \begin{equation*} (-\Delta)^{s} v_{n}(x) +\frac{1}{2a} v_{n}(x) \leqslant 0, \; \; {\rm for}\; \; |x| > R, \; n\geqslant N_{0}. \end{equation*}

    Arguing as in the proof of [19, Lemma 5.6], we see that

    \begin{equation*} v_{n}(x) \leqslant \frac{C}{1+|x|^{N+2s}}, \; \; {\rm for}\; \; |x| > R, \; n\geqslant N_{0}. \end{equation*}

    Thus, v_{n}\to Q in L^{2}(\mathbb{R}^{N}) . By direct calculation, we see that

    \begin{equation*} \|v_{n}\|_{L^{2}(\mathbb{R}^{N})}^{2} = \omega_{n}^{\frac{N(\lambda-2)-4s}{2s(\lambda-2)}}c_{n}, \; \; {\rm and}\; \; \|v_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} = \omega_{n}^{\frac{N(\lambda-2)-4s}{2s(\lambda-2)}}e_{n}/\omega_{n}. \end{equation*}

    Combining with Remark 3 and recalling that \omega_{n}\to 0 , we deduce that \|v_{n}\|_{L^{2}(\mathbb{R}^{N})}^{2} and \|v_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} are comparable. According to the fact v_{n}\to Q in L^{2}(\mathbb{R}^{N}) , there exist C_{3}, C_{4} > 0 such that

    \begin{equation*} C_{3} \leqslant \|v_{n}\|_{L^{2}(\mathbb{R}^{N})}^{2} \leqslant C_{4}, \end{equation*}

    for all n\in \mathbb{N} . Therefore, there exist C_{5}, C_{6} > 0 such that

    \begin{equation*} C_{5} \leqslant \|v_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} \leqslant C_{6}. \end{equation*}

    Hence, \{v_{n}\} is a bounded sequence in H^{s}(\mathbb{R}^{N}) . Up to a subsequence, v_{n}\rightharpoonup v in H^{s}(\mathbb{R}^{N}) . Moreover, from Lemma 1.1 and the fact v_{n}\to Q in L^{2}(\mathbb{R}^{N}) , one gets that v_{n}\to Q in L^{q}(\mathbb{R}^{N}), q\in [2, 2_{s}^{*}) . Applying (5.15), we observe that \|v_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} \to \|Q\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} , which yields v_{n}\to Q in H^{s}(\mathbb{R}^{N}) .

    Let c_{n}\to 0^{+} . From Lemma 5.3 and (5.4), we obtain that \omega_{n}\to +\infty and e_{n}\to +\infty as n\to+\infty .

    Lemma 5.9. Let 1\leqslant N < 4s , s\in (\frac{1}{2}, 1) , and suppose that f satisfies (H_{1}) (H_{5}) . Let (\omega_{c}, u_{c}) be the solution given by Theorem 2. Then,

    \begin{equation*} \liminf\limits_{c\to 0^{+}} u_{c}(0) = +\infty, \end{equation*}

    and

    \begin{equation*} \liminf\limits_{c\to0^{+}} \frac{[u_{c}(0)]^{\gamma-2}}{\omega_{c}} > 0. \end{equation*}

    Proof. For each sequence \{c_{n}\}\to 0^{+} , set

    \begin{equation*} \tilde{v}_{n}(x) : = \frac{1}{u_{n}(0)} u_{n} \left( \frac{e_{n}^{\frac{1}{2s}}}{\omega_{n}^{\frac{1}{2s}}} x\right). \end{equation*}

    A direct calculation shows that \tilde{v}_{n}(0) = \|\tilde{v}_{n}\|_{L^{\infty}(\mathbb{R}^{N})} = 1 and

    \begin{equation} \left(\frac{a}{e_{n}}+b\right) (-\Delta)^{s} \tilde{v}_{n} +\tilde{v}_{n} = \frac{1}{\omega_{n}u_{n}(0)} f(u_{n}(0)\tilde{v}_{n}). \end{equation} (5.16)

    Letting x = 0 in (5.16), and applying (H_{2}) , there exists C > 0 such that

    \begin{equation*} \begin{aligned} 1 \leqslant& \left(\frac{a}{e_{n}}+b\right) (-\Delta)^{s} \tilde{v}_{n}(0) +\tilde{v}_{n}(0) = \frac{1}{\omega_{n}u_{n}(0)} f(u_{n}(0)\tilde{v}_{n}(0))\\ \leqslant& \frac{C}{\omega_{n}u_{n}(0)} \left( u_{n}^{\lambda-1}(0) +u_{n}^{\gamma-1}(0) \right) = \frac{C}{\omega_{n}} \left( u_{n}^{\lambda-2}(0) +u_{n}^{\gamma-2}(0) \right), \end{aligned} \end{equation*}

    which yields that u_{n}(0)\to+\infty , since \omega_{n}\to+\infty, \gamma > \lambda > 2 . Moreover, by \lambda\leqslant \gamma , we deserve that

    \begin{equation*} \liminf\limits_{n\to+\infty} \frac{[u_{n}(0)]^{\gamma-2}}{\omega_{n}} \geqslant \frac{1}{2C} > 0. \end{equation*}

    By the arbitrary of c_{n} , we complete the proof.

    Lemma 5.10. Let 1\leqslant N < 4s , s\in (\frac{1}{2}, 1) , and suppose that f satisfies (H_{1}) (H_{5}) . Let (\omega_{c}, u_{c}) be the solution given by Theorem 2. Then,

    \begin{equation*} \limsup\limits_{c\to0^{+}} \frac{[u_{c}(0)]^{\gamma-2}}{\omega_{c}} < +\infty. \end{equation*}

    Proof. We assume by contradiction that there exists a sequence c_{n}\to 0^{+} such that

    \begin{equation*} \frac{[u_{n}(0)]^{\gamma-2}}{\omega_{n}} \to +\infty, \end{equation*}

    which implies that \omega_{n} = o([u_{n}(0)]^{\gamma-2}) . Set

    \begin{equation*} \hat{v}_{n}(x) : = \frac{1}{u_{n}(0)} u_{n} \left( \frac{e_{n}^{\frac{1}{2s}}}{u_{n}^{\frac{\gamma-2}{2s}}(0)} x\right). \end{equation*}

    Then, \hat{v}_{n}(0) = \|\hat{v}_{n}\|_{L^{\infty}(\mathbb{R}^{N})} = 1 and

    \begin{equation} \left(\frac{a}{e_{n}}+b\right) (-\Delta)^{s} \hat{v}_{n} +\frac{\omega_{n}}{u^{\gamma-2}_{n}(0)} \hat{v}_{n} = \frac{1}{u^{\gamma-1}_{n}(0)} f(u_{n}(0)\hat{v}_{n}). \end{equation} (5.17)

    Note that

    \begin{equation*} \frac{1}{u^{\gamma-1}_{n}(0)} f(u_{n}(0)\hat{v}_{n}) \leqslant \frac{C}{u^{\gamma-1}_{n}(0)} \left( u^{\gamma-1}_{n}(0) \hat{v}_{n}^{\gamma-1} +u^{\lambda-1}_{n}(0) \hat{v}_{n}^{\lambda-1} \right) = C\left( \hat{v}_{n}^{\gamma-1} +u^{\lambda-\gamma}_{n}(0) \hat{v}_{n}^{\lambda-1} \right). \end{equation*}

    Together with u_{n}(0)\to +\infty and \lambda\leqslant \gamma , we conclude that

    \begin{equation*} \frac{1}{u^{\gamma-1}_{n}(0)} f(u_{n}(0)\hat{v}_{n}) \leqslant C\left( \hat{v}_{n}^{\gamma-1} + \hat{v}_{n}^{\lambda-1} \right). \end{equation*}

    This indicates that \frac{1}{u^{\gamma-1}_{n}(0)} f(u_{n}(0)\hat{v}_{n}) is of L^{\infty}(\mathbb{R}^{N}) . Therefore, applying a similar argument to the proof of [19, Proposition 4.4], and passing to a subsequence if necessary, we assume that \hat{v}_{n}\to \hat{v} in C_{loc}^{2, \alpha}(\mathbb{R}^{N}) , for some \alpha\in (0, 1) . Combining Lemma 5.6 and (H_{4}) , noting that e_{n}\to +\infty , we deduce that \hat{v} is a nonnegative bounded radial solution of

    \begin{equation*} (-\Delta)^{s}\hat{v} = \frac{\mu_{2}}{a}\hat{v}^{\gamma-1}, \; \; {\rm in}\; \; \mathbb{R}^{N}. \end{equation*}

    Thanks to [6, Theorem 1.5], we derive that \hat{v} = 0 , which contradicts \hat{v}(0) = 1 .

    Proof of Theorem 5. For each sequence \{c_{n}\}\to 0^{+} , set

    \begin{equation*} \bar{v}_{n}(x) : = \omega_{n}^{\frac{1}{2-\lambda}} u_{n} \left( \frac{\|u_{n}\|^{\frac{1}{s}}_{D^{s, 2}(\mathbb{R}^{N})}}{\omega_{n}^{\frac{1}{2s}}} x \right). \end{equation*}

    Furthermore, \bar{v}_{n} is the solution of

    \begin{equation} (ae_{n}^{-1}+b) (-\Delta)^{s} \bar{v}_{n}(x) +\bar{v}_{n}(x) = \frac{f(\omega_{n}^{\frac{1}{\gamma-2}}\bar{v}_{n})} {\omega_{n}^{\frac{\gamma-1}{\gamma-2}}}. \end{equation} (5.18)

    From (H_{2}) , we obtain that

    \begin{equation*} \frac{f(\omega_{n}^{\frac{1}{\gamma-2}} \bar{v}_{n})} {\omega_{n}^{\frac{\gamma-1}{\gamma-2}}} \leqslant \frac{C}{\omega_{n}^{\frac{\gamma-1}{\gamma-2}}} \left( \omega^{\frac{\gamma-1}{\gamma-2}}_{n} \bar{v}_{n}^{\gamma-1} +\omega^{\frac{\lambda-1}{\gamma-2}}_{n} \bar{v}_{n}^{\lambda-1} \right) = C\left( \bar{v}_{n}^{\gamma-1} +\omega^{\frac{\lambda-\gamma}{\gamma-2}}_{n} \bar{v}_{n}^{\lambda-1} \right). \end{equation*}

    Combining \omega_{n}\to +\infty and \lambda\leqslant \gamma , we deduce that \frac{f(\omega_{n}^{\frac{1}{\gamma-2}} \bar{v}_{n})} {\omega_{n}^{\frac{\gamma-1}{\gamma-2}}}\in L^{\infty}(\mathbb{R}^{N}) . Hence, a similar discussion to the proof of Lemma 5.8 and Theorem 3 implies that \bar{v}_{n}\to U in C_{loc}^{2, \alpha}(\mathbb{R}^{N}) , for some \alpha\in (0, 1) . Recalling that \omega_{n}\to +\infty and e_{n}\to +\infty as n\to+\infty , from (H_{4}) , we obtain that U is nontrivial nonnegative solution of

    \begin{equation*} \begin{cases} b(-\Delta)^{s}U+U = \mu_{2}U^{\gamma-1}, \; \; {in}\; \mathbb{R}^{N}, \\ \lim\nolimits_{|x|\to+\infty} U(x) = 0. \end{cases} \end{equation*}

    From [4], we know that U is radial, positive, and strictly decreasing in x .

    Proof of Theorem 6. Letting c_{n}\to 0^{+} and recalling (5.18), we derive that

    \begin{equation} (-\Delta)^{s} \bar{v}_{n}(x) +\frac{1}{(ae_{n}^{-1}+b)} \left[ 1- \frac{f(\omega_{n}^{\frac{1}{\gamma-2}}\bar{v}_{n})} {\omega_{n}^{\frac{\gamma-1}{\gamma-2}} \bar{v}_{n}(x)} \right] \bar{v}_{n}(x) = 0. \end{equation} (5.19)

    By the proof of Theorem 5, we also obtain that \lim_{|x|\to+\infty}\bar{v}_{n}(x) = 0 uniformly in n\in\mathbb{N} . It follows from (H_{2}) and \omega_{n}\to +\infty as n\to +\infty that

    \begin{equation*} \begin{aligned} \frac{f(\omega_{n}^{\frac{1}{\gamma-2}}\bar{v}_{n})} {\omega_{n}^{\frac{\gamma-1}{\gamma-2}} \bar{v}_{n}(x)} \leqslant C\left( \bar{v}_{n}^{\gamma-2} + \bar{v}_{n}^{\lambda-2} \right) \to 0, \end{aligned} \end{equation*}

    as |x|\to +\infty uniformly in n\in\mathbb{N} . Hence, there exist R > 0 large enough and N_{1}\in \mathbb{N} such that

    \begin{equation*} \frac{1}{(ae_{n}^{-1}+b)} \left[ 1- \frac{f(\omega_{n}^{\frac{1}{\gamma-2}}\bar{v}_{n})} {\omega_{n}^{\frac{\gamma-1}{\gamma-2}} \bar{v}_{n}(x)} \right] > \frac{1}{2b}, \end{equation*}

    for |x| > R and n\geqslant N_{1} . Arguing similarly as in the proof of Theorem 4, we can show that \bar{v}_{n}\to U in L^{2}(\mathbb{R}^{N}) . By direct calculation, we see that

    \begin{equation*} \|\bar{v}_{n}\|_{L^{2}(\mathbb{R}^{N})}^{2} = \omega_{n}^{\frac{N}{2s}-\frac{2}{\gamma-2}} e_{n}^{-\frac{N}{2s}}c_{n}, \; \; {\rm and}\; \; \|\bar{v}_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} = \omega_{n}^{\frac{N}{2s}-\frac{2}{\gamma-2}} e_{n}^{-\frac{N}{2s}}\frac{e_{n}}{\omega_{n}}. \end{equation*}

    Combining with Remark 3 and recalling that e_{n}\to +\infty , we deduce that \|\bar{v}_{n}\|_{L^{2}(\mathbb{R}^{N})}^{2} and \|\bar{v}_{n}\|_{D^{s, 2}(\mathbb{R}^{N})}^{2} are comparable. Similar as the proof of Theorem 4, we observe that \bar{v}_{n}\to U in H^{s}(\mathbb{R}^{N}) .

    Min Shu, Haibo Chen and Jie Yang: Conceptualization, Methodology, Validation, Writing-original draft, Writing-review & editing. All authors contributed equally to this work.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    The authors thank the anonymous referees for their valuable comments and nice suggestions to improve the results.

    J. Yang is supported by Natural Science Foundation of Hunan Province of China (Nos. 2023JJ30482 and 2022JJ30463), Research Foundation of Education Bureau of Hunan Province (Nos. 23A0558 and 22A0540), and Aid Program for Science and Technology Innovative Research Team in Higher Educational Institutions of Hunan Province.

    On behalf of all authors, the corresponding author states that there is no conflict of interest.



    [1] X. Bao, Y. Lv, Z. Ou, Normalized bound state solutions of fractional Schrödinger equations with general potential, Complex Var. Elliptic, in press. https://doi.org/10.1080/17476933.2024.2338436
    [2] M. Du, L. X. Tian, J. Wang, F. B. Zhang, Existence of normalized solutions for nonlinear fractional Schrödinger equations with trapping potentials, Proc. Roy. Soc. Edinb. A, 149 (2019), 617–653. https://doi.org/10.1017/prm.2018.41 doi: 10.1017/prm.2018.41
    [3] P. Felmer, A. Quaas, J. Tan, Positive solutions of nonlinear Schrödinger equation with the fractional Laplacian, Proc. Roy. Soc. Edinb. A, 142 (2012), 1237–1262. https://doi.org/10.1017/S0308210511000746 doi: 10.1017/S0308210511000746
    [4] R. Frank, E. Lenzmann, L. Silvestre, Uniqueness of radial solutions for the fractional Laplacian, Commun. Pur. Appl. Math., 69 (2016), 1671–1726. https://doi.org/10.1002/cpa.21591 doi: 10.1002/cpa.21591
    [5] R. Frank, R. Seiringer, Non-linear ground state representations and sharp Hardy inequalities, J. Funct. Anal., 255 (2008), 3407–3430. https://doi.org/10.1016/j.jfa.2008.05.015 doi: 10.1016/j.jfa.2008.05.015
    [6] M. Fall, T. Weth, Nonexistence results for a class of fractional elliptic boundary value problems, J. Funct. Anal., 263 (2012), 2205–2227. http://dx.doi.org/10.1016/j.jfa.2012.06.018 doi: 10.1016/j.jfa.2012.06.018
    [7] H. L. Guo, Y. M. Zhang, H. S. Zhou, Blow-up solutions for a Kirchhoff type elliptic equation with trapping potential, Commun. Pur. Appl. Anal., 17 (2018), 1875–1897. https://doi.org/10.3934/cpaa.2018089 doi: 10.3934/cpaa.2018089
    [8] Q. He, Z. Lv, Y. Zhang, X. Zhong, Existence and blow up behavior of positive normalized solution to the Kirchhoff equation with general nonlinearities: mass super-critical case, J. Differ. Equations, 356 (2023), 375–406. https://doi.org/10.1016/j.jde.2023.01.039 doi: 10.1016/j.jde.2023.01.039
    [9] N. Ikoma, K. Tanaka, A note on deformation argument for L^{2} normalized solutions of nonlinear Schrödinger equations and systems, Adv. Differential Equations, 24 (2019), 609–646. https://doi.org/10.57262/ade/1571731543 doi: 10.57262/ade/1571731543
    [10] L. Z. Kong, H. B. Chen, Normalized ground states for fractional Kirchhoff equations with Sobolev critical exponent and mixed nonlinearities, J. Math. Phys., 64 (2023), 061501. https://doi.org/10.1063/5.0098126 doi: 10.1063/5.0098126
    [11] L. Li, K. Teng, J. Yang, H. Chen, Properties of minimizers for the fractional Kirchhoff energy functional, J. Math. Phys., 64 (2023), 081504. https://doi.org/ 10.1063/5.0157267 doi: 10.1063/5.0157267
    [12] L. T. Liu, H. B. Chen, J. Yang, Normalized solutions to the fractional Kirchhoff equations with a perturbation, Appl. Anal., 102 (2023), 1229–1249. https://doi.org/10.1080/00036811.2021.1979222 doi: 10.1080/00036811.2021.1979222
    [13] E. Nezza, G. Palatucci, E. Valdinoci, Hitchhiker's guide to the fractional Sobolev spaces, Bull. Sci. Math., 136 (2012), 521–573. https://doi.org/10.1016/j.bulsci.2011.12.004 doi: 10.1016/j.bulsci.2011.12.004
    [14] T. Phan, Blow-up profile of Bose-Einstein condensate with singular potentials, J. Math. Phys., 58 (2017), 072301. https://doi.org/10.1063/1.4995393 doi: 10.1063/1.4995393
    [15] S. Secchi, Ground state solutions for nonlinear fractional Schrödinger equations in \mathbb{R}^{N}, J. Math. Phys., 54 (2013), 031501. https://doi.org/10.1063/1.4793990 doi: 10.1063/1.4793990
    [16] Y. Su, Z. Liu, Semiclassical states to nonlinear Choquard equation with critical growth, Isr. J. Math., 255 (2023), 729–762. https://doi.org/10.1007/s11856-023-2485-9 doi: 10.1007/s11856-023-2485-9
    [17] Y. Su, Z. Liu, Semi-classical states for the Choquard equations with doubly critical exponents: existence, multiplicity and concentration, Asymptotic Anal., 132 (2023), 451–493. https://doi.org/10.3233/ASY-221799 doi: 10.3233/ASY-221799
    [18] L. Silvestre, Regularity of the obstacle problem for a fractional power of the Laplace operator, Commun. Pur. Appl. Math., 60 (2007), 67–112. https://doi.org/10.1002/cpa.20153 doi: 10.1002/cpa.20153
    [19] K. M. Teng, R. P. Agarwal, Existence and concentration of positive ground state solutions for nonlinear fractional Schrödinger-Poisson system with critical growth, Math. Method. Appl. Sci., 41 (2018), 8258–8293. https://doi.org/10.1002/mma.5289 doi: 10.1002/mma.5289
    [20] J. Yang, H. Chen, L. Liu, Limiting behaviors of constrained minimizers for the mass subcritical fractional NLS equations, Anal. Math. Phys., 14 (2024), 32. https://doi.org/10.1007/s13324-024-00899-x doi: 10.1007/s13324-024-00899-x
    [21] S. Yao, H. Chen, V. Rădulescu, J. Sun, Normalized solutions for lower critical Choquard equations with critical Sobolev perturbations, SIAM J. Math. Anal., 54 (2022), 3696–3723. https://doi.org/10.1137/21M1463136 doi: 10.1137/21M1463136
    [22] H. Y. Ye, The sharp existence of constrained minimizers for a class of nonlinear Kirchhoff equations, Math. Method. Appl. Sci., 38 (2015), 2663–2679. https://doi.org/10.1002/mma.3247 doi: 10.1002/mma.3247
    [23] H. Y. Ye, The existence of normalized solutions for L^{2}-critical constrained problems related to Kirchhoff equations, Z. Angew. Math. Phys., 66 (2015), 1483–1497. https://doi.org/10.1007/s00033-014-0474-x doi: 10.1007/s00033-014-0474-x
    [24] X. Y. Zeng, Y. M. Zhang, Existence and uniqueness of normalized solutions for the Kirchhoff equation, Appl. Math. Lett., 74 (2017), 52–59. https://doi.org/10.1016/j.aml.2017.05.012 doi: 10.1016/j.aml.2017.05.012
  • Reader Comments
  • © 2025 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(602) PDF downloads(90) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog