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

A fractional Laplacian problem with critical nonlinearity

  • Received: 22 March 2021 Accepted: 18 May 2021 Published: 01 June 2021
  • MSC : 35B09, 35J20

  • In this paper, using the mountain pass theorem we obtain a positive solution to the fractional Laplacian problem

    {(Δ)su=g(x)(uk)q1++un+2sn2sin Ω,u>0in Ω,u=0on Ω,

    where ΩRn is a bounded smooth domain, 0<s<1,2q<2n/(n2s) and k(0,) is an arbitrary number. The function g:ΩR is a nonnegative continuous function satisfying some integrability condition.

    Citation: Xiuhong Long, Jixiu Wang. A fractional Laplacian problem with critical nonlinearity[J]. AIMS Mathematics, 2021, 6(8): 8415-8425. doi: 10.3934/math.2021488

    Related Papers:

    [1] Mengyu Wang, Xinmin Qu, Huiqin Lu . Ground state sign-changing solutions for fractional Laplacian equations with critical nonlinearity. AIMS Mathematics, 2021, 6(5): 5028-5039. doi: 10.3934/math.2021297
    [2] Fugeng Zeng, Peng Shi, Min Jiang . Global existence and finite time blow-up for a class of fractional p-Laplacian Kirchhoff type equations with logarithmic nonlinearity. AIMS Mathematics, 2021, 6(3): 2559-2578. doi: 10.3934/math.2021155
    [3] Zhilin Li, Guoping Chen, Weiwei Long, Xinyuan Pan . Variational approach to p-Laplacian fractional differential equations with instantaneous and non-instantaneous impulses. AIMS Mathematics, 2022, 7(9): 16986-17000. doi: 10.3934/math.2022933
    [4] Kun Cheng, Wentao Huang, Li Wang . Least energy sign-changing solution for a fractional p-Laplacian problem with exponential critical growth. AIMS Mathematics, 2022, 7(12): 20797-20822. doi: 10.3934/math.20221140
    [5] Maria Alessandra Ragusa, Abdolrahman Razani, Farzaneh Safari . Existence of positive radial solutions for a problem involving the weighted Heisenberg p()-Laplacian operator. AIMS Mathematics, 2023, 8(1): 404-422. doi: 10.3934/math.2023019
    [6] Lixiong Wang, Ting Liu . Existence and regularity results for critical (p,2)-Laplacian equation. AIMS Mathematics, 2024, 9(11): 30186-30213. doi: 10.3934/math.20241458
    [7] Wangjin Yao . Variational approach to non-instantaneous impulsive differential equations with p-Laplacian operator. AIMS Mathematics, 2022, 7(9): 17269-17285. doi: 10.3934/math.2022951
    [8] Zhiqiang Li . The finite time blow-up for Caputo-Hadamard fractional diffusion equation involving nonlinear memory. AIMS Mathematics, 2022, 7(7): 12913-12934. doi: 10.3934/math.2022715
    [9] Adnan Khaliq, Mujeeb ur Rehman . Fixed point theorem combined with variational methods for a class of nonlinear impulsive fractional problems with derivative dependence. AIMS Mathematics, 2021, 6(2): 1943-1953. doi: 10.3934/math.2021118
    [10] Yan Ning, Daowei Lu, Anmin Mao . Existence and subharmonicity of solutions for nonsmooth p-Laplacian systems. AIMS Mathematics, 2021, 6(10): 10947-10963. doi: 10.3934/math.2021636
  • In this paper, using the mountain pass theorem we obtain a positive solution to the fractional Laplacian problem

    \begin{equation*} \label{eq: main problem} \begin{cases} (-\Delta)^su = g(x)(u-k)_{+}^{q-1}+u^{\frac{n+2s}{n-2s}}& \rm{in}\ \,\Omega,\\ u>0& \rm{in}\ \,\Omega,\\ u = 0& \rm{on}\ \,\partial\Omega, \end{cases} \end{equation*}

    where ΩRn is a bounded smooth domain, 0<s<1,2q<2n/(n2s) and k(0,) is an arbitrary number. The function g:ΩR is a nonnegative continuous function satisfying some integrability condition.



    In recent years, motivated by problems that appear in anomalous diffusions in plasmas, flames propagation and chemical reactions in liquids etc, fractional Laplacian problems have received extensive attention in the literature. Also, due to its rich multi-origins, fractional Laplacian operators turn out to have many different definitions. When studying fractional problems in the whole space Rn, the fractional Laplacian (Δ)s is usually defined via the Fourier transformation (Δ)sf(x)=(|2πiξ|2sˆf(ξ)), see e.g. Frank and Lenzmann [12] and Di Nezza, Palatucci and Valdinoci [11]. Equivalently, this operator can be written as the difference quotient

    (Δ)sf(x)=cn,sP.V.Rnf(x)f(y)|xy|n+2sdy (1.1)

    for sufficiently regular f, where cn,s is a normalization constant, for a proof see e.g. [11]. An advantage of the difference quotient type definition is that it can be directly extended to define the so-called fractional p-Laplacian operator

    (Δ)spf(x)=cn,sP.V.Rnf(x)f(y)|xy|n+spdy,

    and the so called regional fractional Laplacian operator

    (Δ)sΩf(x)=cn,sP.V.Ωf(x)f(y)|xy|n+2sdy,

    in case the problem under consideration is restricted to a bounded domain ΩRn, see e.g. Chen [10] and the references therein. It is beyond the size of this paper to present all the other definitions for fractional Laplacian operators in the literature. We refer the interested readers to the references mentioned above.

    In this paper, we are concerned with the fractional Laplacian operator (Δ)s on a bounded domain defined via the spectral of the Laplacian operator Δ. This type of fractional Laplacian operator was first defined by Cabré and Tan [7] for s=1/2 and then extended by Brändle, Colorado, de Pablo and Sánchez [4] to the whole range s(0,1). This type of definition can be viewed as a discrete version of Fourier transformation for functions on a bounded domain. More precisely, let s(0,1) and ΩRn be a bounded smooth domain. Let {λk,φk}k1 be the corresponding eigenvalues and eigenfunctions of Δ on Ω with Dirichlet boundary value φk|Ω=0 and normalization φkL2(Ω)=1. Then, for u=k1ckφk satisfying k1c2kλ2sk<, we define

    (Δ)su=k1λskckφk. (1.2)

    Therefore, a natural function space of (Δ)s in the sense of functional analysis is given by

    Hs0(Ω)={u=ckφkL2(Ω):u2Hs0(Ω)c2kλsk<}

    such that

    (Δ)su,vΩ(Δ)s/2u(Δ)s/2vdx=k1ckdkλsk

    is well defined for every u=ckφk,v=dkφkHs0(Ω). Observe that a remarkable feature of the operator (Δ)s is that it is nonlocal, in the sense that for any point xΩ, the value of (Δ)su at point x can be obtained only if one knows the distribution of u in the whole domain, so that it is possible derive all of the coefficients ckΩuφkdx. For a systematical study on this type of fractional Laplacian operators, we refer to e.g. Cabré and Tan [7] for the case s=1/2 and Brändle et al. [4] for the general case s(0,1).

    Now our problem under consideration can be stated as below. Fix s(0,1) and consider the following fractional Laplacian problem

    {(Δ)su=g(x)(uk)q1++u2s1inΩ,u>0inΩ,u=0onΩ, (1.3)

    where 2s=2n/(n2s), 2q<2s and k(0,) is an arbitrary positive number. A positive function uHs0(Ω) is called a (weak) solution of problem (1.3) if for every vHs0(Ω), there holds

    Ω(Δ)s/2u(Δ)s/2vdx=Ω(g(x)(uk)q1++u2s1)vdx.

    Precise assumptions on the coefficient function g will be given soon. Our aim is to prove that under appropriate conditions on the function g and the parameters q,n, there does exist at least one positive solution to the above problem.

    Problem (1.3) is not totally new. Problems of type (1.3) have been extensively studied in the literature with different leading operators. In the local case (i.e., s=1) with k=0, problem of type (1.3) dates back to the famous work Brézis and Nirenberg [6], where positive solutions to the critical growth problem was obtained:

    {Δu=λu+u21inΩ,u>0inΩ,u=0onΩ.

    In this case, the function g is the constant function λ. Since then, there are numerous extensions and important variants which have been studied. In [1], Ambrosetti, Brezis and Cerami considered the problem combined with concave and convex nonlinearities

    {Δu=λuq+upinΩ,u>0inΩ,u=0onΩ,

    where 0<q<1<p and λ>0. They proved among many other results that when λ is sufficiently small, this problem has a positive solution, see [1, Theorem 2.1] for more results. Quite recently, this problem was extended by Barrios, Colorado, Servadei, and Soria [3] to the nonlocal setting

    {(Δ)su=λuq+upinΩ,u>0inΩ,u=0onRnΩ,

    where 0<s<1, 0<q<p=n+2sn2s and λ>0, and (Δ)su is defined via the difference quotient (1.1). Note that in this problem the boundary condition is in fact assumed on the complement of Ω. They obtained similar results [3, Theorem 1.1, 1.2] as that of [1, Theorem 2.1]. In [15] Servadei and Valdinoci considered the even more general integrodifferential problem of Brézis and Nirenberg type

    {LKu=λu+f(x,u)+|u|4sn2suinΩ,u=0onRnΩ,

    where s(0,1) and LK is a nonlocal operator with kernel function K such that (Δ)s plays a model case. Under various natural assumptions they proved several existence results to this problem, see [15, Theorem 1–3]. In particular, as an application, they obtained a positive solution in the model case LK=(Δ)s and f0, see [15, Theorem 4]. Similar to [3], but with the leading operator (Δ)s defined in the spectral way (1.2), Barrios, Colorado, de Pablo and Sánchez [2] considered the problem

    {(Δ)su=λuq+upinΩ,u>0inΩ,u=0onΩ,

    with 0<q<p=n+2sn2s. Then also obtained existence and nonexistence results under various assumptions on λ, see [2, Theorem 1.1, 1.2, 1.3] for details. For the case p<n+2sn2s, see Br¨andle, Colorado, de Pablo and Sánchez [4].

    In the case gconstant, Gazzola [13] studied this type of problem with p-Laplacian operator Δp as the leading term,

    {div(|u|p2u)=g(x)(uk)q1++up1inΩ,u>0inΩ,u=0onΩ

    where 1<p<n, p=np/(np) and positive solutions were obtained among other results, see [13, Theorem 2]. Inspired by the work of Gazzola [13] and Cabré and Tan [7], Wang studied problem (1.3) with s=1/2 and obtained a positive solution to her problem. We mention that Li and Xiang [14] extended Wang's work [17] to the setting of regional fractional Laplacion problems.

    Inspired by Wang's work [17], in this paper we study problem (1.3) with s belonging to the full range (0,1). Throughout we assume that 2q<2s and g satisfies

    (G1) gC(Ω) is a nonnegative nontrivial function, and

    (G2) gL2n/(2n(n2s)q)(Ω). In addition, if q=2, then gLn/2s(Ω)<S, where S is defined as in (1.6) in the below.

    Our main result reads as follows.

    Theorem 1.1. Assume that s(0,1) and g satisfies (G1) (G2). Then problem (1.3) admits a positive solution for every k(0,), provided either q2andn>2s(1+2/q), or q=2 and n=4s.

    It is clear that the nonlocality of (Δ)s makes it difficult to deal with problem (1.3) directly. Our strategy is to use a localization method. In the global case Ω=Rn, a localization principle for fractional Laplacian problem was first systematically developed by Caffarelli and Silvestre [9]. In our local case, a localization principle was also developed by Cabré and Tan [7] for the case s=1/2 and by Brändle et al. [4] for the general case s(0,1). To state the localization principle, we need to introduce some notations. Denote

    C=Ω×(0,)andLC=Ω×(0,)

    with coordinates (x,y)C for xΩ and y>0. According to Brändle et al. [4], for every uHs0(Ω), there exists a unique 2s-harmonic extension ˉu which is equal to u on Ω×{y=0} in the sense of trace with ˉu=0 on the lateral boundary LC, and satisfies the equation

    div(y12sˉu)=0inC.

    Following Brändle et al. [4], introduce the function space

    X2s0(C)=¯C0(Ω×[0,))X2s0(C)withzX2s0(C)=(κsCy12s|z|2dxdy)1/2,

    where κs=212sΓ(1s)/Γ(s) is a normalization constant such that the 2s-harmonic extension ˉu of u satisfies ˉuX2s0(C)=uHs0(Ω). Then, problem (1.3) is equivalent to

    {div(y12sˉu)=0,ˉu>0inC,ˉu=0onLC,ˉuν2s=g(x)(ˉuk)q1++ˉu2s1onΩ×{y=0}, (1.4)

    where

    ˉuν2s(x,0)=lim

    That is, u\in H_{0}^{s}({\Omega}) is a solution to problem (1.3) if and only if \bar{u}\in X_{0}^{2s}({\mathcal C}) is a solution to problem (1.4) which means that

    {\kappa}_{s}\int_{{\mathcal C}}y^{1-2s}{\nabla}\bar{u}\cdot{\nabla}\Phi {{\rm{d}}} x {{\rm{d}}} y = \int_{{\Omega}}\left(g(x)(\overline{u}-k)_{+}^{q-1}+\overline{u}^{2_{s}^{\ast}-1}\right)\Phi {{\rm{d}}} x

    holds for all \Phi\in X_{0}^{2s}({\mathcal C}) . This explains the localization method.

    So, to prove Theorem 1.1, we turn to the equivalent problem (1.4). Note that problem (1.4) is variational with the energy functional {\mathcal J}:X_{0}^{2s}({\mathcal C})\to{\mathbb R} being given by

    {\mathcal J}(\bar{u}) = \frac{1}{2}\|\bar{u}\|_{X_{0}^{2s}({\mathcal C})}^{2}-\int_{{\Omega}}\left(\frac{1}{q}g(x)(\bar{u}-k)_{+}^{q} +\frac{1}{2_{s}^{*}}\bar{u}_+^{2_{s}^{\ast}}\right) {{\rm{d}}} x

    for \bar{u}\in X_{0}^{2s}({\mathcal C}) . To find a critical point of {\mathcal J} , we will use the mountain pass theorem (see e.g. Struwe [16, Chapter 6]).

    Before ending the introduction, we record the following inequality (see [4, Theorem 2.1]) for later use: there exists C = C(n, s) > 0 such that, for every w\in X^{2s}({\mathbb R}_{+}^{n+1}) , the closure of C_{0}^{{\infty}}(\overline{{\mathbb R}_{+}^{n+1}}) under the seminorm \|u\|_{X^{2s}({\mathbb R}_{+}^{n+1})}^{2} = \int_{{\mathbb R}_{+}^{n+1}}{\kappa}_{s}y^{1-2s}|{\nabla} w(x, y)|^{2} {{\rm{d}}} x {{\rm{d}}} y , there holds

    \begin{equation} C\left(\int_{{\mathbb R}^{n}}|w(x,0)|^{2_{s}^{\ast}} {{\rm{d}}} x\right)^{2/2_{s}^{\ast}}\le\int_{{\mathbb R}_{+}^{n+1}}{\kappa}_{s}y^{1-2s}|{\nabla} w(x,y)|^{2} {{\rm{d}}} x {{\rm{d}}} y. \end{equation} (1.5)

    Let

    \begin{equation} S = \inf\left\{ \frac{\int_{{\mathcal {\mathbb R}_{+}^{n+1}}}{\kappa}_{s}y^{1-2s}|{\nabla} w(x,y)|^{2} {{\rm{d}}} x {{\rm{d}}} y}{\left(\int_{{\mathbb R}^{n}}|w(x,0)|^{2_{s}^{\ast}} {{\rm{d}}} x\right)^{2/2_{s}^{\ast}}}:w\in X^{2s}({\mathbb R}_{+}^{n+1}),w\not\equiv0\right\} \end{equation} (1.6)

    be the best constant for inequality (1.5). It is known [4, Theorem 2.1] that S is attained by the functions

    \begin{eqnarray} U_{\epsilon,x_{0}}(x,y) = (P_{y}^{1-2s}\ast u_{{\epsilon},x_{0}})(x), & & (x,y)\in{\mathbb R}^{n}\times[0,{\infty}) \end{eqnarray} (1.7)

    for all {\epsilon} > 0 and x_{0}\in{\mathbb R}^{n} , where P_{y}^{1-2s}(x) = k_{1-2s}y^{-n}(1+|x|/y)^{-(n+2s)/2} is the so-called s -Poisson kernel and

    u_{{\epsilon},x_{0}}(x) = c_{n}\left(\frac {{\epsilon}}{|x-x_{0}|^{2}+{\epsilon}^{2}}\right)^{(n-2s)/2}.

    The constant c_{n} > 0 is chosen such that

    \int_{{\mathbb R}_{+}^{n+1}}{\kappa}_{s}y^{1-2s}|{\nabla} U_{{\epsilon},x_{0}}|^{2} {{\rm{d}}} x {{\rm{d}}} y = \int_{{\mathbb R}^{n}}|U_{{\epsilon},x_{0}}(x,0)|^{2_{s}^{\ast}} {{\rm{d}}} x = S^{n/2s}.

    Our notations are standard. We use C to denote positive constants that are different from line to line. For simplicity, we will use \|\cdot\| to denote the norm \|\cdot\|_{X_{0}^{2s}({\mathcal C})} throughout.

    This section is devoted to the proof of Theorem 1.1. Our method is to use the well known mountain pass theorem, see e.g. Struwe [16, Theorem 6.1]. The first lemma points out that

    Lemma 2.1. \mathcal{J} satisfies the mountain pass geometry.

    The proof is standard, and we omit the details. We shall also need

    Lemma 2.2. Under the assumptions (G_{1}) and (G_{2}) , there hold

    g(u_{m}-k)_{+}^{q-1}\rightarrow g(u-k)_{+}^{q-1}\qquad in \,L^{\frac{2n}{n+2s}}(\Omega)

    and

    g(u_{m}-k)_{+}^{q}\rightarrow g(u-k)_{+}^{q}\qquad in \,L^{2s}(\Omega)

    for every (u_{m})\subset X_{0}^{2s}({\mathcal C}) that converges weakly to u\in X_{0}^{2s}({\mathcal C}) .

    Lemma 2.2 is a consequence of Vitali's convergence theorem. We omit the details, see also Wang [17, Lemma 2.1] and Gazzola [13, Lemma 1].

    We say that (u_{m})\in X_{0}^{2s}({\mathcal C}) is a (PS)_c sequence, if \mathcal{J}(u_{m})\rightarrow c\in {\mathbb R} and \mathcal{J}'(u_{m})\rightarrow 0 as m\rightarrow \infty . Say that the functional \mathcal{J} satisfies (PS)_c compactness condition if (u_{m})\in X_{0}^{2s}({\mathcal C}) is a (PS)_c sequence, then it contains a convergent subsequence in X_{0}^{2s}({\mathcal C}) . Next lemma says that this happens when c is below a critical value, as that was observed by Brezis and Nirenberg [6].

    Lemma 2.3. \mathcal{J} satisfies the (PS)_{c} condition provided c < \frac {s}{n} S^{\frac{n}{2s}} .

    Proof. Let (u_{m})\subset X_{0}^{2s}({\mathcal C}) be a (PS)_c sequence. First we claim that (u_{m}) is a bounded sequence in X_{0}^{2s}({\mathcal C}) . This will follow from a simple combination of the (PS)_c assumption. Let \beta\geq2 . Then

    \mathcal{J}(u_{m}) -\frac{1}{\beta}\langle\mathcal{J}'(u_{m}),u_{m}\rangle = c+o(1)+o(1)\|u_{m}\|.

    Taking \beta = q when q > 2 yields

    \bigg(\frac{1}{2}-\frac{1}{q}\bigg)\|u_{m}\|^{2} +\frac{k}{q}\int_{\Omega}g(x)(u_{m}-k)_{+}^{q-1}dx +\bigg(\frac{1}{q}-\frac{1}{2_{s}^{*}}\bigg)\int_{\Omega}(u_m)_+^{2_{s}^{*}}dx = c+o(1)+o(1)\|u_{m}\|.

    Then the assumption g\geq0 implies that

    \begin{equation} \bigg(\frac{1}{2}-\frac{1}{q}\bigg)\|u_{m}\|^{2}\le c+o(1)+o(1)\|u_{m}\|. \end{equation} (2.1)

    When q = 2 , take 2_{s}^{*} > \beta > 2 . Then using the assumption (G_{2}) and Hölder's inequality gives

    \begin{equation} \bigg(\frac{1}{2}-\frac{1}{\beta}\bigg) \bigg(1-\frac{\|g\|_{L^{\frac{n}{2s}}}}{S}\bigg)\|u_{m}\|^{2}\le c+o(1)+o(1)\|u_{m}\|. \end{equation} (2.2)

    Now the boundedness of (u_{m})\subset X_{0}^{2s}({\mathcal C}) follows from (2.1) and (2.2).

    Next we prove that (u_{m}) has a convergent subsequence in X_{0}^{2s}({\mathcal C}) . Since we have proved the boundedness of (u_{m}) in X_{0}^{2s}({\mathcal C}) , we may assume without loss of generality that, up to a subsequence,

    \begin{eqnarray*} & & u_{m} \rightharpoonup u\;{\rm{in }}\;X_{0}^{2s}({\mathcal C}),\\ & & u_{m}(\cdot,0)\to u(\cdot,0)\;{\rm{in }}\;L^{2}({\Omega}),\\ & & u_{m}(\cdot,0)\to u(\cdot,0)\;{\rm{ a.e.\; in }}\;{\Omega}, \end{eqnarray*}

    for some u\in X_{0}^{2s}({\mathcal C}). Put v_{m} = u_{m}-u . It suffices to prove that \|v_{m}\|\rightarrow0 .

    It is standard to find that \mathcal{J}'(u) = 0. To examine the sequence closer, note that

    \|u_{m}\|^{2} = \|v_{m}+u\|^{2} = \|v_{m}\|^{2}+\|u\|^{2}+o(1).

    Using Lemma 2.2 we find that

    \int_{\Omega}g(u_{m}-k)_{+}^{q}\rightarrow \int_{\Omega}g(u-k)_{+}^{q},

    By Lemma 2 of Brézis and Lieb [5],

    \int_{\Omega}(u_{m})_+^{2_{s}^{\ast}} = \int_{\Omega}(v_{m})_+^{2_{s}^{*}} +\int_{\Omega}u_{+}^{2_{s}^{*}}+o(1).

    As a result, we obtain the first decomposition

    \begin{aligned} \mathcal{J}(u_{m})& = \frac 12 \|u_m\|^2-\int_{\Omega}\bigg(\frac{1}{q}g(x)(u_{m}-k)_{+}^{q}+\frac{1}{2_{s}^{*}}(u_{m})_{+}^{2_{s}^{*}}\bigg)dx\\ & = \frac{1}{2}(\|v_{m}\|^{2}+\|u\|^{2})-\int_{\Omega}\frac{1}{q}g(x)(u-k)_{+}^{q}-\frac{1}{2_{s}^{*}}\int_{\Omega}\bigg((v_{m})_{+}^{2_{s}^{*}}+u_{+}^{2_{s}^{*}}\bigg)+o(1)\\ & = \mathcal{J}(u)+\frac{1}{2}\|v_{m}\|^{2}-\frac{1}{2_{s}^{*}}\int_{\Omega}(v_{m})_{+}^{2_{s}^{*}}+o(1). \end{aligned}

    Similarly, we also have the second decomposition

    \begin{aligned} o(1) = \langle\mathcal{J}'(u_{m}),u_{m}\rangle & = \|u_m\|^2-\int_{\Omega}\bigg(g(x)(u_{m}-k)_{+}^{q-1} +(u_{m})_{+}^{2_{s}^{*}-1}\bigg)\cdot u_{m}dx \\ & = \|v_{m}\|^{2}+\|u\|^{2}-\int_{\Omega}g(x)(u-k)_{+}^{q-1}\cdot u-\int_{\Omega}(v_{m})_{+}^{2_{s}^{*}}-\int_{\Omega}u_{+}^{2_{s}^{*}}+o(1)\\ & = \langle\mathcal{J}'(u),u\rangle+\|v_{m}\|^{2} -\int_{\Omega}(v_{m})_{+}^{2_{s}^{\ast}}+o(1)\\ & = \|v_{m}\|^{2}-\int_{\Omega}(v_{m})_{+}^{2_{s}^{*}}+o(1), \end{aligned}

    where we used used the assumption \mathcal{J}'(u_{m})\to 0 in the first equality and the fact \mathcal{J}'(u) = 0 in the last line.

    To continue, note that \|v_{m}\|\leq C(\|u_{m}\|+\|u\|) is a bounded sequence. So we may assume \|v_{m}\|^{2}\rightarrow b\geq0 . Then the above second decomposition gives

    \int_{\Omega}(v_{m})_{+}^{2_{s}^{*}} = \|v_{m}\|^{2}+o(1)\rightarrow b.

    Recall the trace inequality (1.5). This implies Sb^{\frac{2}{2_{s}^{*}}}\leq b . As a consequence, we infer that either b = 0 or b\geq S^{\frac{n}{2s}} holds.

    We have to exclude the case b\geq S^{\frac{n}{2s}} . To this end, note that \mathcal{J}'(u) = 0 implies \mathcal{J}(u)\geq0 . Thus the first decomposition leads to

    \begin{equation*} \label{eq:result 03} \mathcal{J}(u_{m})\geq\frac{1}{2}\|v_{m}\|^{2}-\frac{1}{2_{s}^{*}} \int_{\Omega}(v_{m})_{+}^{2_{s}^{*}}+o(1). \end{equation*}

    If b\geq S^{\frac{n}{2s}} , then taking limit in this inequality gives c\geq \frac{s}{n}S^{\frac{n}{2s}} , which contradicts with our assumption c < \frac{s}{n}S^{\frac{n}{2s}} . Hence b = 0 . That is v_{m}\rightarrow 0 in X_{0}^{2s}({\mathcal C}) . The proof is finished.

    To proceed, we may assume that 0\in {\Omega} and g(0) > 0 with no loss of generality. By the assumption (G{1}) we can assume B_{r}(0)\subset \{x\in\Omega:g(x) > g(0)\} for some r > 0 sufficiently small. For simplicity, write u_{{\epsilon}} = u_{{\epsilon}, 0}, U_{{\epsilon}} = U_{{\epsilon}, 0} , where u_{{\epsilon}, x_{0}}, U_{{\epsilon}, x_{0}} are defined as in (1.7). Let 0 < \rho < r and take a cut-off function \eta\in C^{\infty}(\overline{\mathcal{C}}) such that 0\leq\eta\leq1 and \eta(x, y)\equiv1 for |x| < \frac{\rho}{2} and y\geq0 , and \eta(x, y)\equiv0 for |x|\geq\rho . These auxiliary functions and parameters are used to construct a special path in X_{0}^{2s}({\mathcal C}) starting from the origin such that the following lemma holds.

    Lemma 2.4. For {\epsilon} > 0 sufficiently small there holds

    \max\limits_{t\ge 0} \mathcal{J}(t\eta U_{{\epsilon}}) < \frac sn S^{\frac{n}{2s}}.

    Proof. Write v_{{\epsilon}} = \eta U_{{\epsilon}} and \gamma(t) = tv_{{\epsilon}} for t\geq0 . Suffices to show that \max_{t > 0}\mathcal{J}(\gamma(t)) < \frac{s}{n}S^{\frac{n}{2s}} for {\epsilon} sufficiently small. By a direct computation we have

    \begin{aligned} \mathcal{J}(\gamma(t))& = \frac{1}{2}\|tv_{{\epsilon}}\|^{2} -\int_{\Omega}\frac{1}{q}g(x)(tv_{{\epsilon}}-k)_{+}^{q} -\frac{1}{2_{s}^{*}}\int_{\Omega}(t v_{{\epsilon}})^{{2_{s}^{*}}}\\ & = \frac{t^{2}}{2}\|v_{{\epsilon}}\|^{2}-\int_{\Omega}\frac{1}{q}g(x)(tv_{{\epsilon}}-k)_{+}^{q} -\frac{t^{2_{s}^{*}}}{2_{s}^{*}}\int_{\Omega}v_{{\epsilon}}^{{2_{s}^{*}}}. \end{aligned}

    Since the second term on the right hand side is nonnegative, there holds

    \mathcal{J}(\gamma(t)) \le \frac{t^{2}}{2}\|v_{{\epsilon}}\|^{2} -\frac{t^{2_{s}^{*}}}{2_{s}^{*}}\int_{\Omega}v_{{\epsilon}}^{{2_{s}^{*}}}\to -{\infty} \qquad{\rm{ as }} \; t \to \infty.

    Also note that v_{\epsilon} is a bounded function. So for t sufficiently small, we have (tv_{\epsilon}-k)_+\equiv 0 . Hence for t sufficiently small,

    \mathcal{J}(\gamma(t)) = \frac{t^{2}}{2}\|v_{{\epsilon}}\|^{2} -\frac{t^{2_{s}^{*}}}{2_{s}^{*}}\int_{\Omega}v_{{\epsilon}}^{{2_{s}^{*}}} > 0 \qquad{\rm{ for }}\; t \to 0.

    So \mathcal{J}(\gamma(t)) achieves a positive maximum on (0, {\infty}) .

    Let t_{{\epsilon}} be such that \mathcal{J}(\gamma(t_{{\epsilon}})) = \max_{t > 0}\mathcal{J}(\gamma(t)) . We claim that for {\epsilon} sufficiently small, there exist constants C_{1}, C_{2} independent of {\epsilon} such that

    \begin{equation} 0 < C_{1}\leq t_{{\epsilon}}\leq C_{2}. \end{equation} (2.3)

    Notice that

    \mathcal{J}(\gamma(t))\leq\frac{t^{2}}{2}\|v_{{\epsilon}}\|^{2} -\frac{t^{2_{s}^{*}}}{2_{s}^{\ast}}\int_{\Omega}v_{{\epsilon}}^{{2_{s}^{\ast}}}\le 0\qquad {\rm{for }}\; t\ge \overline{t}_{{\epsilon}},

    where \overline{t}_{{\epsilon}} = ({2_{s}^{*}\|v_{{\epsilon}}\|^{2}}/ {2\int_{\Omega}v_{{\epsilon}}^{{2_{s}^{*}}}})^{(n-2s)/{4s}} . We infer that \mathcal{J}(\gamma(t))\leq0 holds for all t\geq \overline{t}_{{\epsilon}} , which implies

    t_{{\epsilon}}\leq \overline{t}_{{\epsilon}}\leq C_{2}\qquad {\rm{as }}\; {\epsilon} \to 0

    for some C_2 > 0 independent of {\epsilon} . On the other hand, by Lemma 2.1, t_{{\epsilon}} cannot converge to zero. Therefore there exists C_{1} > 0 such that

    t_{{\epsilon}}\geq C_{1}

    for {\epsilon} sufficiently small. The claim is proved.

    To further compute the maximum \mathcal{J}(\gamma(t_{{\epsilon}})) , according to [2], there hold

    \|v_{{\epsilon}}\|^{2} = S^{\frac{n}{2s}}+O({\epsilon}^{n-2s}) \quad \;{\rm{and}}\; \quad \int_{\Omega}v_{{\epsilon}}^{{2_{s}^{*}}} = S^{\frac{n}{2s}}+O({\epsilon}^{n})

    for {\epsilon} sufficiently small. Thus

    \begin{aligned} \frac{t_{{\epsilon}}^{2}}{2}\|v_{{\epsilon}}\|^{2} -\frac{t_{{\epsilon}}^{2_{s}^{*}}}{2_{s}^{*}}\int_{\Omega}v_{{\epsilon}}^{{2_{s}^{*}}}& = \frac{t_{{\epsilon}}^{2}}{2}(S^{\frac{n}{2s}}+O({\epsilon}^{n-2s})) -\frac{t_{{\epsilon}}^{2_{s}^{*}}}{2_{s}^{*}}(S^{\frac{n}{2s}}+O({\epsilon}^{n}))\\ & = \frac{s}{n}S^{\frac{n}{2s}}+\left(\frac{t_{{\epsilon}}^{2}-1}{2} -\frac{t_{{\epsilon}}^{2_{s}^{*}}-1}{2_{s}^{*}}\right)S^{\frac{n}{2s}}+O({\epsilon}^{n-2s})\\ &\leq \frac{s}{n}S^{\frac{n}{2s}}+O({\epsilon}^{n-2s}), \end{aligned}

    where we used in the last line the fact that \frac{t^{2}-1}{2}-\frac{t^{2_{s}^{*}}-1}{2_{s}^{*}}\le 0 for all t\ge 0 . Therefore we derive the estimate

    \begin{equation} \mathcal{J}(\gamma t_{{\epsilon}})\leq\frac{s}{n}S^{\frac{n}{2s}} -\int_{\Omega}\frac{1}{q}g(x)(t_{{\epsilon}}v_{{\epsilon}}-k)_{+}^{q}+O({\epsilon}^{n-2s}). \end{equation} (2.4)

    We need to estimate \int_{{\Omega}}g(t_{{\epsilon}}v_{{\epsilon}}-k)_{+}^{q} . For {\epsilon} sufficiently small, we always have

    \int_{{\Omega}}g(t_{{\epsilon}}v_{{\epsilon}}-k)_{+}^{q}\ge g(0)\int_{B_{\rho/2}(0)}(t_{{\epsilon}}U_{{\epsilon}}-k)_{+}^{q}.

    In the case q\ge 2 and n > 2s(1+2/q) , there holds

    \int_{B_{\rho/2}(0)}(t_{{\epsilon}}U_{{\epsilon}}-k)_{+}^{q} \ge C\int_{B_{{\epsilon}}(0)}{\epsilon}^{-\frac{n-2s}{2}q} = C{\epsilon}^{n-\frac{n-2s}{2}q}

    for some C > 0 independent of {\epsilon} , where we have used the estimate (2.3) for t_{{\epsilon}} . In the case q = 2, n = 4s , note that on the set \{{\epsilon} < |x| < {\epsilon}^{3/4}\} , there holds

    U_{{\epsilon}}(x)\ge\frac{{\epsilon}^{s}}{({2|x|^{2}})^{s}}\ge\frac{{\epsilon}^{-\frac{1}{2}s}}{2^{s}} > 2k

    for {\epsilon} sufficiently small. Hence,

    g(0)\int_{B_{\rho/2}(0)}(t_{{\epsilon}}u_{{\epsilon}}-k)_{+}^{2}\ge C\int_{\{{\epsilon} < |x| < {\epsilon}^{3/4}\}}\frac{{\epsilon}^{2s}}{|x|^{4s}} = C{\epsilon}^{2s}|\ln{\epsilon}|

    for some C > 0 independent of {\epsilon} . Thus, there exists C > 0 such that

    \begin{equation} \int_{{\Omega}}g(t_{{\epsilon}}v_{{\epsilon}}-k)_{+}^{q}\ge\begin{cases}C{\epsilon}^{n-\frac{n-2s}{2}q}, & \;{\rm{if }}\;q\ge 2\;{\rm{and}}\;n > 2s(1+2/q)\\ C{\epsilon}^{2s}\ln{\epsilon}& \;{\rm{if }}\;q = 2,n = 4s. \end{cases} \end{equation} (2.5)

    Combining (2.4) and (2.5) yields

    \mathcal{J}(\gamma t_{{\epsilon}})\leq \begin{cases}\frac{s}{n}S^{\frac{n}{2s}}- C{\epsilon}^{n-\frac{n-2s}{2}q}\left(1-{\epsilon}^{\frac{n-2s}{2}q-2s}\right) & \;{\rm{if }}\;q\ge 2\;{\rm{and}}\;n > 2s(1+2/q),\\ \frac{s}{n}S^{\frac{n}{2s}}- C{\epsilon}^{2s}(|\ln{\epsilon}|-1)& \;{\rm{if }}\;q = 2,n = 4s. \end{cases}

    In the case q\ge 2\;{\rm{and}}\;n > 2s(1+2/q) , it holds \frac{n-2s}{2}q-2s > 0 . Therefore, in both cases we can deduce that

    \mathcal{J}(\gamma t_{{\epsilon}}) < \frac{s}{n}S^{\frac{n}{2s}},

    provided {\epsilon} is sufficiently small. The proof is complete.

    Now we are in the position to prove Theorem 1.1.

    Proof of Theorem 1.1. First choose e = t_0 \eta U_{\epsilon} , where {\epsilon} and \eta are chosen as in Lemma 2.4 and t_0 is sufficiently large such that {\mathcal J}(t_0 \eta U_{\epsilon}) < 0 , and then let

    \Gamma = \left\{\gamma\in C\left([0,1], X_{0}^{2s}({\mathcal C})\right):\gamma(0) = 0 \;{\rm{and}}\; {\gamma}(1) = e\right\}

    and

    \begin{equation*} \label{eq: critical value} c_{0} = \inf\limits_{\gamma\in\Gamma}\max\limits_{t\geq0}\mathcal{J}(\gamma(t)). \end{equation*}

    By Lemma 2.4, we have c_0 < \frac sn S^{n/2s} . Since \mathcal{J} satisfies the geometry of mountain pass, there exists a sequence (u_{m})\subset X_{0}^{2s}({\mathcal C}) satisfying {\mathcal J}(u_{m})\to c_{0} and \mathcal{J}^{\prime}(u_{m})\to0 as m\to{\infty} . Therefore, Lemma 2.3 implies that problem (1.4) admits a nonnegative nontrivial solution. Finally, a maximum principle of Cabré and Sire [8, Remark 4.2] implies that the solution is positive in {\Omega} . The proof is complete.

    This work is supported by the Project of Hubei University of Arts and Science (No. XK2021023). The first author would like to thank Miao Chen for many useful help and suggestions. She would also like to thank the second author for pointing out the problems. Both authors would like to thank the anonymous referees for valuable comments that greatly improve this manuscript.

    All authors declare no conflict of interest in this paper.



    [1] A. Ambrosetti, H. Brezis, G. Cerami, Combined effects of concave and convex nonlinearities in some elliptic problems, J. Funct. Anal., 122 (1994), 519–543. doi: 10.1006/jfan.1994.1078
    [2] B. Barrios, E. Colorado, A. de Pablo, U. Sánchez, On some critical problems for the Laplacian operator, J. Differ. Equations, 252 (2012), 6133–6162. doi: 10.1016/j.jde.2012.02.023
    [3] B. Barrios, E. Colorado, R. Servadei, F. Soria, A critical fractional equation with concave-convex power nonlinearities, Ann. Inst. H. Poincaré Anal. Non Linéaire, 32 (2015), 875–900. doi: 10.1016/j.anihpc.2014.04.003
    [4] C. Brändle, E. Colorado, A. de Pablo, U. Sánchez, A concave-convex elliptic problem involving the fractional Laplacian, Proc. Roy. Soc. Edinburgh Sect. A, 143 (2013), 39–71. doi: 10.1017/S0308210511000175
    [5] H. Brézis, E. Lieb, A relation between pointwise convergence of functions and convergence of functionals, Proc. Amer. Math. Soc., 88 (1983), 486–490. doi: 10.1090/S0002-9939-1983-0699419-3
    [6] H. Brézis, L. Nirenberg, Positive solutions of nonlinear elliptic equations involving critical Sobolev exponents, Comm. Pure Appl. Math., 36 (1983), 437–477. doi: 10.1002/cpa.3160360405
    [7] X. Cabré, J. Tan, Positive solutions of nonlinear problems involving the square root of the Laplacian, Adv. Math., 224 (2010), 2052–2093. doi: 10.1016/j.aim.2010.01.025
    [8] X. Cabré, J. Sire, Nonlinear equations for fractional Laplacians, Ⅰ: Regularity, maximum principles, and Hamiltonian estimates, Ann. Inst. H. Poincaré Anal. Non Linéaire, 31 (2014), 23–53. doi: 10.1016/j.anihpc.2013.02.001
    [9] L. A. Caffarelli, L. Silvestre, An extension problem related to the fractional Laplacian, Commun. Partial Differ. Equations, 32 (2007), 1245–1260. doi: 10.1080/03605300600987306
    [10] H. Chen, The Dirichlet elliptic problem involving regional fractional Laplacian, J. Math. Phys., 59 (2018), 071504. doi: 10.1063/1.5046685
    [11] E. Di Nezza, G. Palatucci, E. Valdinoci, Hitchhiker's guide to the fractional Sobolev spaces, Bull. Sci. Math., 136 (2012), 521–573. doi: 10.1016/j.bulsci.2011.12.004
    [12] R. L. Frank, E. Lenzmann, Uniqueness of non-linear ground states for fractional Laplacians in {\mathbb R}, Acta Math., 210 (2013), 261–318. doi: 10.1007/s11511-013-0095-9
    [13] F. Gazzola, Critical growth quasilinear elliptic problems with shifting subcritical perturbation, Differ. Integr. Equations, 14 (2001), 513–528.
    [14] Q. Li, C. L. Xiang, The fractional problem with shifting subcritical perturbation, in press.
    [15] R. Servadei, E. Valdinoci, The Brezis-Nirenberg result for the fractional Laplacian, Trans. Am. Math. Soc., 367 (2015), 67–102.
    [16] M. Struwe, Variational methods. Applications to nonlinear partial differential equations and Hamiltonian systems, Ergebnisse der Mathematik und ihrer Grenzgebiete, 34, Berlin: Springer-Verlag, Second edition, 1996.
    [17] J. Wang, On a nonlocal problem with critical Sobolev growth, Appl. Math. Lett., 99 (2020), 105959. doi: 10.1016/j.aml.2019.06.030
  • Reader Comments
  • © 2021 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

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

Metrics

Article views(2998) PDF downloads(107) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog