Home Solutions for the Problems Involving Fractional Laplacian and Indefinite Potentials
Article Open Access

Solutions for the Problems Involving Fractional Laplacian and Indefinite Potentials

  • Zhongwei Tang EMAIL logo and Lushun Wang
Published/Copyright: January 12, 2017

Abstract

In this paper, we consider a class of Schrödinger equations involving fractional Laplacian and indefinite potentials. By modifying the definition of the Nehari–Pankov manifold, we prove the existence and asymptotic behavior of least energy solutions. As the fractional Laplacian is nonlocal, when the bottom of the potentials contains more than one isolated components, the least energy solutions may localize near all the isolated components simultaneously. This phenomenon is different from the Laplacian.

MSC 2010: 35J60; 35J20

1 Introduction and Main Results

We consider the following Schrödinger equations involving fractional Laplacian and indefinite potentials:

(${\mathcal{P}_{\lambda}}$) ( - Δ ) s u + ( λ a ( x ) - δ ) u = | u | p - 2 u , x N ,

where 0<s<1, N>2s, 2<p2s*:=2NN-2s, a(x) is a nonnegative continuous function and the zero set a-1(0):={xN:a(x)=0} has a nonempty, smooth and connected interior part inta-1(0).

The fractional Laplacian has been studied extensively in recent years. In [9], Caffarelli and Silvestre showed that the fractional Laplacian (-Δ)s on d can be expressed as the Dirichlet-to-Neumann operator for a suitable local problem on the upper half space +d+1. For one-dimensional case, Frank and Lenzmann [16] proved the uniqueness and nondegeneracy of ground states for a class of semilinear fractional Laplacian equations with subcritical growth. The authors also obtained similar results in [17] for the higher-dimensional case. Via a Lyapunov–Schmidt reduction method, Dávila, del Pino and Wei [13] proved the existence and concentration of standing waves for a class of semilinear fractional Laplacian equations. For more details of fractional Laplacian, please see [3, 7, 8, 10, 11, 19, 28, 30] and the references therein.

Many papers deal with Schrödinger equations involving Laplacian and potential wells. For λ large enough, the operator -Δ+λa(x)-δ is positive definite if δ<λ1, where λ1 is the principle eigenvalue for -Δ on functions defined in inta-1(0). If δ>λ1, the operator -Δ+λa(x)-δ is indefinite. Take δ=-1, the positive definite case, Bartsch and Wang [6] considered (${\mathcal{P}_{\lambda}}$) with a more general nonlinearity f(x,u) satisfies some conditions. The authors proved that there exists Λ1>0 such that (${\mathcal{P}_{\lambda}}$) has a positive and a negative solution for λ>Λ1. Furthermore, if f(x,u) is odd in u, the authors also proved that for any k there exists Λk>0 such that (${\mathcal{P}_{\lambda}}$) has at least k pairs ±u1,±u2,,±uk of nontrivial solutions for λΛk. For more details about the positive definite case, please see [1, 2, 4, 21] and the references therein.

For the indefinite case, Ding and Wei [15] considered the following problem:

(1.1) { - Δ u ( x ) + λ V ( x ) u ( x ) = λ | u ( x ) | p - 2 u ( x ) + λ g ( x , u ) , x N , u ( x ) 0 as  | x | ,

where V(x) can be negative in some domains in N, g(x,u) is a perturbation term. Both for subcritical growth and critical growth, using variational methods, the authors proved that there exists Λ>0 such that for any λ>Λ, problem (1.1) admits at least one nontrivial solution. In [29], Szulkin and Weth gave a new minimax characterization of the corresponding critical value and hence reduced the indefinite problem to a definite one. The authors also presented a precise description of the Nehari–Pankov manifold which is useful even for other problems. Using this type of manifold, Bartsch and the first author [5] proved the existence of multi-bump solutions for semilinear equations with indefinite potentials. For more details about indefinite case, please see [14, 18] and the references therein.

In present paper, we are interested in the existence and concentration phenomenon of least energy solutions for fractional Laplacian with indefinite potentials. Motivated by Servadei and Valdinoci [24, 26, 27, 25], we consider our problem directly in the corresponding fractional Sobolev space without taking s-harmonic extension. We will prove that (${\mathcal{P}_{\lambda}}$) has a least energy solution which localizes near the potential well and changes sign for λ large. Our assumptions about a(x) and δ are as follows.

  1. a ( x ) C ( N , ) satisfies a(x)0 and Ω:=int{xN:a(x)=0} is nonempty, the boundary of Ω is smooth and Ω¯={xN:a(x)=0}.

  2. a 0 = 1 2 lim inf | x | a ( x ) > 0 .

  3. The operator (-Δ)s-δ defined in inta-1(0) is indefinite and nondegenerate. Namely, λk<δ<λk+1 for some k1. Here {λk} is the class of all eigenvalues for (-Δ)s on functions defined in inta-1(0).

Remark 1.1.

Assume (A1) and (A2) hold; we see that Ω is bounded in N. In fact, there exists R>0 such that

(1.2) Ω B R ( 0 ) , a ( x ) > a 0 for all  | x | > R ,

where BR(0) is the ball centered at 0 with radii R. Of course, we can replace (A2) by the following one.

  1. There exists a constant M>0 such that the Lebesgue measure of the set {xN:a(x)<M} is finite, i.e. |{xN:a(x)<M}|<+.

Let Hs(N) be the standard fractional Sobolev space defined by

H s ( N ) := { u : N N | u ( x ) - u ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y < + , N u 2 𝑑 x < + }

with the inner product

u , v := N N ( u ( x ) - u ( y ) ) ( v ( x ) - v ( y ) ) | x - y | N + 2 s 𝑑 x 𝑑 y + N u v 𝑑 x ,

and the induced norm is

u := ( N N | u ( x ) - u ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N u 2 𝑑 x ) 1 2 .

Let

E 0 := { u H s ( N ) : u = 0  a.e. in  N Ω }

be a subspace of Hs(N) endowed with a new norm

u 0 = ( N N | u ( x ) - u ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y ) 1 2 , u E 0 .

As described in Section 2.2, 0 and are equivalent and (E0,0) is a Hilbert space.

Denote

V λ = λ a ( x ) - δ , V λ + = max { 0 , V λ } , V λ - = max { 0 , - V λ } .

Let

E λ := { u H s ( N ) : N a ( x ) u 2 𝑑 x < }

be the Hilbert space endowed with the norm

u λ := ( N N | u ( x ) - u ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ + u 2 𝑑 x ) 1 2 for all  u E λ .

The energy functional corresponding to (${\mathcal{P}_{\lambda}}$) is

I λ ( u ) := 1 2 N N | u ( x ) - u ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y - 1 2 N V λ u 2 𝑑 x - 1 p N | u | p 𝑑 x for all  u E λ .

with Fréchet derivative

I λ ( u ) v := N N ( u ( x ) - u ( y ) ) ( v ( x ) - v ( y ) ) | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ u v 𝑑 x - N | u | p - 2 u v 𝑑 x for all  v E λ .

We say uλEλ is a solution to (${\mathcal{P}_{\lambda}}$) if uλEλ is a critical point of Iλ, i.e. Iλ(uλ)=0.

Let us denote Lλ:=(-Δ)s+λa(x)-δ and L0:=(-Δ)s-δ. Take {μk(L0)} be the class of all eigenvalues of L0 in E0(each eigenvalue is repeated according to its multiplicity). As proved in Section 2.3, if μ1(L0)>0, then (-Δ)s+λa(x)-δ is positive definite and the ground state of (${\mathcal{P}_{\lambda}}$) is positive for λ large enough. If μ1(L0)<0 and μk(L0)0 for any k, then (-Δ)s+λa(x)-δ is indefinite and the least energy solutions of (${\mathcal{P}_{\lambda}}$) are sign changing for λ large enough.

Let ek be an eigenfunction corresponding to μk(L0) for k=1,2,3,. We may assume that {ek}k1 is an orthogonal base of E0 and L2(Ω) (see Lemma 2.3 in Section 2.3). By (A3), E0 can be split as an orthogonal sum E0-E0+ according to the positive and negative eigenvalues of L0, i.e.

E 0 = E 0 - E 0 + ,

where

E 0 - = span { e 1 , e 2 , , e k } , E 0 + = { u E 0 : Ω u v 𝑑 x = 0  for all  v E 0 - } .

As proved in Section 2.3, we see that infσe(Lλ)λa0-δ, where σess(Lλ) is the essential spectrum of Lλ. Moreover, there exist a finite number of k such that μk(Lλ)<0, i.e. Lλ thus has finite Morse index. For large λ fixed, Lλ has finite eigenvalues below σess(Lλ). By the study of asymptotic behavior of eigenvalues for Lλ as λ+, we see that 0 is not an eigenvalue of Lλ for λ large. Therefore, Eλ can be split as an orthogonal sum Eλ=Eλ-Eλ+ according to the negative and positive eigenvalues of Lλ.

Note that for λ large enough, the space Eλ- are close to E0- (please see Section 2.3). In order to study the asymptotic behavior of least energy solutions for (${\mathcal{P}_{\lambda}}$), we need to modify the Nehari–Pankov manifold. To do that, let Pλ-:EλE0- and P0-:E0E0- be two orthogonal projections. Define the modified Nehari–Pankov manifold of Iλ by

𝒩 λ := { u E λ { 0 } : P λ - I λ ( u ) = 0 , I λ ( u ) u = 0 } E λ E 0 - ,

with the corresponding level

c λ := inf u 𝒩 λ I λ ( u ) .

As λ+, the limit problem of (${\mathcal{P}_{\lambda}}$) is

(1.3) { ( - Δ ) s u - δ u = | u | p - 2 u in  Ω , u = 0 in  N Ω

with the energy functional

I 0 ( u ) = 1 2 N N | u ( x ) - u ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y - 1 2 Ω δ u 2 𝑑 x - 1 p Ω | u | p 𝑑 x for all  u E 0 .

The Nehari–Pankov manifold of I0(u) is defined by

𝒩 0 = { u E 0 { 0 } : P 0 - I 0 ( u ) = 0 , I 0 ( u ) u = 0 }

with the corresponding level

c 0 := inf 𝒩 0 I 0 ( u ) ,

where I0(u) is the Fréchet derivative of I0(u).

Remark 1.2.

The Nehari–Pankov manifold was firstly introduced by Pankov [22]. This type of manifold coincides with the Nehari manifold if all eigenvalues are positive, i.e. the negative eigenfunction space is {0}. Moreover, the minimizer for c0 is a least energy solution of problem (1.3). By the definition of the modified Nehari–Pankov manifold 𝒩λ, it is easy to see that Pλ-u=P0-u for any uE0. As proved in Lemma 2.9, we see that the minimizer for cλ is indeed a least energy solution of (${\mathcal{P}_{\lambda}}$), the solution with smallest energy among all nontrivial solutions.

Our main results are as follows.

Theorem 1.3 (Existence).

Assume (A1), (A2), (A3) and one of the following conditions hold:

  1. N > 2 s , p is subcritical, i.e. 2<p2s*,

  2. N 4 s , p is critical, i.e. p=2s*.

Then there exists Λ>0 such that for any λΛ, (Pλ) admits a least energy solution uλEλ which achieves cλ.

Theorem 1.4 (Asymptotic Behavior).

Under the assumptions of Theorem 1.3, let uλ be a least energy solution to (Pλ) for λΛ. Then up to a subsequence, for any sequence λn+, the sequence {uλn} converges to a least energy solution of the limit problem (1.3) in Hs(RN).

The following is an extension of Theorem 1.4.

Theorem 1.5.

Under the assumptions of Theorem 1.3, let uλ be a nontrivial solution to (Pλ) for λΛ. Suppose that one of the following conditions holds:

  1. lim sup λ + I λ ( u λ ) < + for 2 < p < 2 s * ,

  2. lim sup λ + I λ ( u λ ) < s N S N 2 for p = 2 s * , where

    S := inf u H s ( N ) , u L 2 s * ( N ) = 1 N | u ( x ) - u ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y .

Then up to a subsequence, for any sequence λn+, the sequence {uλn} converges to a solution of the limit problem (1.3) in Hs(RN).

Recall that (-Δ)s is a nonlocal operator, please see the definition of (-Δ)s in Section 2.1. Compared with local operators, such as the Laplacian -Δ, nonlocal operators seem much more difficult to deal with. There are some different phenomena between the local and nonlocal operators. For instance:

Remark 1.6.

Assume Ω:=int{xN:a(x)=0}=Ω1Ω2,Ω1Ω2. Here Ω1 and Ω2 are two smooth domains in N. We see that the least energy solutions to (𝒫λ) may localize near Ω1 and Ω2 simultaneously for λ large enough (see Lemma 3.6). This phenomena cannot happen for local operators such as the Laplacian.

In the following sections, without especially stated, we always assume λk<δ<λk+1 for some k1, i.e. μk(L0)<0 and μk+1(L0)>0. Our paper is organized as follows. In Section 2, we give some preliminary results. More precisely, in Section 2.1, we introduce some definitions of fractional Laplacian. In Section 2.2, we introduce some fractional Sobolev spaces. The spectrum of Lλ and L0 is studied in Section 2.3 and some properties of the Modified Nehari–Pankov manifold are proved in Section 2.4. In Section 3, we prove the existence of least energy solutions for the limit problem (1.3). Section 4 is dedicated to the proof of our main results in the subcritical case and Section 5 focus on the proof of the main results in the critical case.

2 Preliminary Results

In this section, for the convenience of the readers, we present some definitions and notations. Firstly, we introduce some definitions of the fractional Laplacian (-Δ)s. Secondly, we consider some fractional Sobolev spaces, such as Eλ and E0. Then we discuss the spectrum of Lλ and L0. At last, we give some properties of the modified Nehari–Pankov manifold.

2.1 Introduction to Fractional Laplacian Operators

Let (N):={ϕC0(N):supxN(1+|x|2)k2|Dαϕ(x)|<+,k,|α|=0,1,2,} be the so-called Schwartz space with the seminorms

p n ( ϕ ) = sup x N | α | n | D α ϕ ( x ) | , n = 1 , 2 , .

We denote the topological dual of (N) by (N). As usual, for any ϕ, we denote by

𝔉 ϕ ( ξ ) = 1 ( 2 π ) N 2 N e - i ξ x ϕ ( x ) 𝑑 x

the Fourier transform of ϕ. For any ϕ, 0<s<1, we define the fractional Laplacian operator (-Δ)s as

(2.1) ( - Δ ) s ϕ ( x ) = C N , s P.V. N ϕ ( x ) - ϕ ( y ) | x - y | N + 2 s 𝑑 y = C N , s lim ϵ 0 N B ϵ ( x ) ϕ ( x ) - ϕ ( y ) | x - y | N + 2 s 𝑑 y .

Here P.V. is a commonly used abbreviation for “in the principle value sense” and CN,s is a dimensional constant that depends on N and s, precisely given by

C N , s = ( N 1 - cos ( ξ 1 ) | ξ | N + 2 s 𝑑 ξ ) - 1 .

According to (2.1), it is easy to see that (-Δ)s is a nonlocal operator. Due to the singularity of the kernel, the right-hand side of (2.1) is not well defined in general. In the case 0<s<12 the integral (2.1) is not really singular near x, see [20]. In order to get rid of P.V. in (2.1), one can redefine the fractional Laplacian operator (-Δ)s as

(2.2) ( - Δ ) s ϕ ( x ) = - C N , s 2 N u ( x + y ) + u ( x - y ) - 2 u ( x ) | y | N + 2 s 𝑑 y for all  ϕ , x N .

Now we take into account an alternative definition of the space Hs(N) via the Fourier transform. Precisely, we define

H ^ s := { u L 2 ( N ) : N ( 1 + | ξ | 2 s ) | 𝔉 ϕ ( ξ ) | 2 𝑑 ξ < + } .

The fractional Laplacian (-Δ)s can be viewed as a pseudo-differential operator of symbol |ξ|2s, namely for 0<s<1,

(2.3) ( - Δ ) s ϕ ( x ) = 𝔉 - 1 ( | ξ | 2 s ( 𝔉 u ) ) for all  u , ξ N .

As proved in [20], Hs(N) and H^s are coincide and the above definitions of (-Δ)s are also coincide.

Recently Caffarelli and Silvestre [7] considered the following boundary value problem in the half space +N+1:

{ ( t 1 - 2 s ϕ ~ ) = 0 in  + N + 1 , ϕ ~ ( x , 0 ) = ϕ ( x ) in  N .

Here ϕ~(x,t) is the so-called s-harmonic extension of ϕ, explicitly given as a convolution integral with the s-Poisson kernel ps(x,y),

ϕ ~ ( x , t ) = N p s ( x - z , t ) ϕ ( z ) 𝑑 z ,

where

p s ( x , t ) = d N , s t 4 s - 1 ( | x | 2 + | t | 2 ) N - 1 + 4 s 2

and dN,s achieves Nps(x,t)𝑑x=1. Then under suitable regularity, (-Δ)s is the Dirichlet-to-Neumann map for this problem, namely

(2.4) ( - Δ ) s ϕ ( x ) = - lim t 0 + t 1 - 2 s t ϕ ~ ( x , t ) .

Thus by this new transformation, the nonlocal operator (-Δ)s may be reduced to a local, possibly singular or degenerate operator on functions sitting in the higher-dimensional half space +N+1=N×(0,+).

The definitions (2.1), (2.2), (2.3) and (2.4) of the fractional Laplacian (-Δ)s are all equivalent for instance in Schwartz’s space of rapidly decreasing smooth functions. For more details about the fractional Laplacian, one refers to [9, 20] and the references therein.

In the present paper, we use (2.1) as the definition of (-Δ)s and for the simplicity and without loss of generality, we replace the constant CN,s by 1 throughout this paper.

2.2 Some Fractional Sobolev Spaces

As we know, Hs(N) can be continuously embedded into Lp(N) for 2p2s*, and compactly embedded into Llocp(N) for 2p<2s*. In particular, the following fractional Sobolev inequality holds:

(2.5) S = inf { N N | u ( x ) - u ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y : u H s ( N ) { 0 } , u L 2 s * ( N ) = 1 } > 0 .

In [12], Cotsiolis and Tavoularis showed that S is achieved by

v = V ( x ) V L 2 s * ( N ) ,

where

V ( x ) = κ ( μ 2 + | x - x 0 | 2 ) - N - 2 s 2

for κ{0}, μ and x0N fixed.

Let ΩN be a bounded smooth domain in N, and let Hs(Ω) be the standard fractional Sobolev space. The embedding Hs(Ω)Lp(Ω) is continuous for 1p2s*, and compact for 1p<2s*. For more details about the fractional Sobolev spaces, one refers to [20] and the references therein.

Recall that E0={uHs(N):u=0 a.e. in NΩ}. In [12], Cotsiolis and Tavoularis gave a sort of Poincaré–Sobolev inequality as follows:

Ω | u | 2 𝑑 x C N N | u ( x ) - u ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y for all  u E 0 ,

where C>0 depends on Ω and N. According to the definitions of the norms 0 and , we see that the two norms are equivalent in E0. Moreover, (E0,0) is also a Hilbert space (see [12, Lemma 7]). The embedding E0Lp(Ω) is continuous for 1p2s*, and compact for 1p<2s* (see [12, Lemma 8]).

We finish this subsection by giving the following uniformly imbedding lemma.

Lemma 2.1.

There exists a constant λ0>0 such that the embedding EλHs(RN) is continuous, uniformly in λ for λ>λ0.

Proof.

By (1.2) in Remark 1.1, ΩBR(0) for large R>0 and a(x)>a0 for each xNBR(0). Let ξ:N be a smooth cutoff function satisfying

(2.6) 0 ξ 1 , ξ = 1 in  B R ( 0 ) , ξ = 0 in  N B R + 1 ( 0 ) .

Take λ0=a0+δa0; then by (2.6), for λλ0, we have

(2.7) N ( 1 - ξ 2 ) u 2 𝑑 x = N B R ( 0 ) ( 1 - ξ 2 ) u 2 𝑑 x
(2.8) 1 a 0 N B R ( 0 ) ( λ a ( x ) - δ ) u 2 𝑑 x
= 1 a 0 N B R ( 0 ) V λ + u 2 𝑑 x
(2.9) 1 a 0 N V λ + u 2 𝑑 x .

By (2.6), Hölder’s inequality and the definition of S, we find that

(2.10) N ξ 2 u 2 𝑑 x = B R + 1 ( 0 ) ξ 2 u 2 𝑑 x
(2.11) | B R + 1 ( 0 ) | 1 - 2 2 s * ( N | u | 2 s * 𝑑 x ) 2 2 s *
(2.12) | B R + 1 ( 0 ) | 1 - 2 2 s * S N N | u ( x ) - u ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y .

According to (2.9), (2.12) and the definition of the norm , we obtain that

u 2 = N N | u ( x ) - u ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N u 2 𝑑 x
( | B R + 1 ( 0 ) | 1 - 2 2 s * + 1 ) N N | u ( x ) - u ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + 1 a 0 N V λ + u 2 𝑑 x
C ( N N | u ( x ) - u ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ + u 2 𝑑 x ) C u λ 2 .

Thus uCuλ, where C>0 is independent of λ. ∎

Remark 2.2.

For λ>λ0, since the embedding constant for EλHs(N) does not dependent on λ, Eλ keeps all properties of Hs(N). Precisely, for λ>λ0, Eλ can be continuously embedded into Lp(N) for 2p2s* and compactly into Llocp(N) for 2p<2s*. Moreover, all embedding constants do not depend on λ.

2.3 Eigenvalue Problems with Fractional Laplacian Operators

Firstly, we consider the spectrum of the operator L0 in E0. Recall that Ω=inta-1(0) is a smooth bounded domain in N. For 0<s<1, we say λ is an eigenvalue of (-Δ)s in E0 if there exists uE0 such that

(2.13) N N ( u ( x ) - u ( y ) ) ( v ( x ) - v ( y ) ) | x - y | N + 2 s 𝑑 x 𝑑 y = λ Ω u ( x ) v ( x ) 𝑑 x

for any vE0, i.e. (u,λ) is a weak solution to the equation

(2.14) { ( - Δ ) s u ( x ) = λ u ( x ) in  Ω , u ( x ) = 0 in  N Ω ,

A function uE0 satisfying (2.13) is called an eigenfunction corresponding to λ. For convenience, we also say λ is an eigenvalue of (2.14).

The following lemma described the eigenvalues and eigenfunctions to (2.14).

Lemma 2.3 ([26, Proposition 9]).

Let 0<s<1, N>2s, and let Ω be an open bounded set of RN.

  1. Problem ( 2.14 ) admits an eigenvalue λ 1 which is positive and that can be characterized as

    (2.15) λ 1 = min { N N | u ( x ) - u ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y : u E 0 , u L 2 ( Ω ) = 1 } .

  2. There exists a nonnegative function e 1 E 0 , which is an eigenfunction corresponding to λ 1 , attaining the minimum in ( 2.15 ), that is e 1 L 2 ( Ω ) = 1 and

    λ 1 = N N | e 1 ( x ) - e 1 ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y .

  3. λ 1 is simple, that is if u E 0 is a solution to the equation

    N N ( u ( x ) - u ( y ) ) ( ϕ ( x ) - ϕ ( y ) ) | x - y | N + 2 s 𝑑 x 𝑑 y = λ 1 Ω u ( x ) ϕ ( x ) 𝑑 x , u E 0 , for all  ϕ ( x ) E 0 ,

    then u = k e 1 , with k .

  4. The set of the eigenvalues of problem ( 2.14 ) consists of a sequence { λ k } k 1 with

    0 < λ 1 < λ 2 λ k λ k + 1

    and

    λ k + as  k + .

    Moreover, for any k 1 , the eigenvalues can be characterized as

    (2.16) λ k + 1 = min { 2 N | u ( x ) - u ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y : u 𝒫 k + 1 , u L 2 ( Ω ) = 1 } ,

    where

    𝒫 k + 1 = { u E 0 : 2 N ( u ( x ) - u ( y ) ) ( e j ( x ) - e j ( y ) ) | x - y | N + 2 s = 0 , j = 1 , , k } .

  5. For any k 1 , there exists a function e k + 1 , which is an eigenfunction corresponding to λ k + 1 , attaining the minimum in ( 2.16 ), that is e k + 1 L 2 ( Ω ) = 1 and

    λ k + 1 = N N | e k + 1 ( x ) - e k + 1 ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y .

  6. The sequence { e k } k 1 of eigenfunctions corresponding to λ k is an orthonormal basis of L 2 ( Ω ) and an orthogonal basis of E 0 .

  7. Each eigenvalue λ k has finite multiplicity; more precisely, if λ k is such that

    λ k - 1 < λ k = λ k + 1 = = λ k + h < λ k + h + 1

    for some h 0 , then the set of all eigenfunctions corresponding to λ k agrees with

    span { e k , e k + 1 , , e k + h } .

Next, we study the spectrum of Lλ in Eλ. It is easy to see that Lλ is self-adjoint in L2(N) and bounded from below by -δ. Denote

U λ ( ψ 1 , ψ 2 , , ψ m ) = inf { L λ ( ψ ) ψ : ψ E λ , ψ L 2 ( N ) = 1 , ( ψ , ψ j ) = 0 , j = 1 , 2 , , m } ,

where

L λ ( ψ ) ψ := N N | ψ ( x ) - ψ ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ ψ 2 𝑑 x , ψ L 2 ( N ) 2 := N ψ 2 𝑑 x , ( ψ , ψ j ) := N ψ ψ j 𝑑 x = 0 .

For k fixed, we define spectral values of Lλ by the k-th Rayleigh quotient

μ k ( L λ ) := sup { U λ ( ψ 1 , ψ 2 , , ψ k - 1 ) : ψ 1 , ψ 2 , , ψ k - 1 E λ } .

Obviously, μk(Lλ) is nondecreasing respect to k and λ. By [23, Theorem XIII.1 and Theorem XIII.2], we know that either μk(Lλ) is an eigenvalue of Lλ or μk(Lλ)=μk+1(Lλ)==infσess(Lλ). Denote μk(L0):=λk-δ. By Lemma 2.3, it is easy to see that {μk(L0)}k1 is the class of all eigenvalues of L0:=(-Δ)s-δ in E0. Moreover, μk(L0) can be characterized as

μ k ( L 0 ) = max S k - 1 min ψ S , ψ L 2 ( Ω ) = 1 { L 0 ( ψ ) ψ } ,

where

L 0 ( ψ ) ψ := N N | ψ ( x ) - ψ ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y - δ Ω ψ 2 𝑑 x ,

and k-1 denotes the collection of (k-1)-dimensional subspaces of E0. Since Lλ(ψ)ψ=L0(ψ)ψ for each ψE0, we have μk(Lλ)μk(L0). Thus in order to prove the existence of eigenvalues of Lλ in Eλ, we just need to prove that infσess(Lλ)+ as λ+.

Lemma 2.4.

Under the assumptions (A1)(A3), the essential spectrum σess(Lλ) is contained in [λa0-δ,+) for λ>λ0. Furthermore, infσess(Lλ)+ as λ+.

Proof.

The proof of this lemma is similar to [4, Proposition 2.3]. For the convenience of the reader, we give the details.

We set Wλ=Vλ-λa0+δ=λ(a(x)-a0) and write Wλ1=max{Wλ,0} and Wλ2=min{Wλ,0}. Obviously, for λ>λ0,

(2.17) σ ( ( - Δ ) s + W λ 1 + λ a 0 - δ ) [ λ a 0 - δ , + )

for Wλ10. Let Hλ:=(-Δ)s+Wλ1+λa0-δ; then Lλ=Hλ+Wλ2.

We claim that Wλ2 is a relative form compact perturbation of Lλ. Since Wλ2 is bounded, the form domain of Hλ is the same as the form domain Eλ of Lλ. Thus we have to show that EλEλ*, uWλ2u is compact. Select a bounded sequence {un}n1 in Eλ,λ; then according to Lemma 2.1, {un}n1 is also a bounded sequence in Hs(N). Thus there exists a function uHs(N) such that, up to a subsequence,

(2.18) { u n u weakly in  H s ( N ) , u n u strongly in  L loc 2 ( N ) , u n u  a.e. in N

as n+. Notice that a(x)>a0 for each xNBR (see (1.2) in Remark 1.1), then the support set of Wλ2 is contained in BR. Thus by Hölder’s inequality, and Lemma 2.1, for λ>λ0, we have

| N W λ 2 ( u n - u ) v 𝑑 x | = | B R W λ 2 ( u n - u ) v 𝑑 x | δ B R | ( u n - u ) v | 𝑑 x
δ ( B R ( u n - u ) 2 𝑑 x ) 1 2 v C ( B R ( u n - u ) 2 𝑑 x ) 1 2 v λ .

Hence by (2.18), we have

W λ 2 u n - W λ 2 u E λ * C ( B R ( u n - u ) 2 𝑑 x ) 1 2 0

as n+. Thus Wλ2 is a relative form compact perturbation of Lλ.

According to the Classical Weyl Theorem (see [23, Example 3, p. 117]), σess(Lλ)=σess(Hλ). Thus by (2.17), for λ>λ0, σess(Lλ)[λa0-δ,+). Moreover, infσess(Lλ)+ as λ+. ∎

At last, we deal with the asymptotic behavior of the eigenvalues of Lλ in Eλ. Let {μkλ}k1 be the class of all distinct eigenvalues of Lλ:=(-Δ)s+λa(x)-δ in Eλ, and let {μk}k1 be the class of all distinct eigenvalues of L0:=(-Δ)s-δ in E0. Without loss of generality, we can order these eigenvalues as follows:

μ 1 λ < μ 2 λ < μ 3 λ < < μ k λ λ < inf σ ess ( L λ ) ,

and

μ 1 < μ 2 < μ 3 < < μ k < 0 < μ k + 1 < .

Moreover, μkλλ+ as λ+, and μk+ as k+. Let Vjλ be the eigenfunction space corresponding to μjλ and let Vj be the eigenfunction space corresponding to μj. We say Vkλ converges to Vk, i.e. VkλVk as k+, if for any sequence λi and normalized eigenfunctions ψiVkλi, there exists a normalized eigenfunction ψVk such that ψiψ strongly in Hs(N) along a subsequence. We have the following lemma.

Lemma 2.5.

We have μkλμk and VkλVk as λ.

Proof.

We prove this lemma by induction.

Step 1. We firstly prove the case for k=1, i.e. up to a subsequence, we will prove that

λ n + , μ 1 λ n μ 1 , V 1 λ n V 1 as  n + .

According to the definition of μ1λn, it is easy to see that μ1λnμ1. Now we assume that ψnEλnis such that

(2.19) N ψ n 2 𝑑 x = 1 , N N | ψ n ( x ) - ψ n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N ( λ n a ( x ) - δ ) ψ n 2 𝑑 x = μ 1 λ n .

By (2.19), we have

ψ n λ n 2 = N N | ψ n ( x ) - ψ n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ n ψ N 2 𝑑 x + N V λ n - ψ n 2 𝑑 x = μ 1 λ n + N V λ n - ψ n 2 𝑑 x μ 1 + δ .

Then by Lemma 2.1, {ψn} is bounded in Hs(N). Up to a subsequence, there is ψHs(N) such that

(2.20) { ψ n ψ weakly in  H s ( N ) , ψ n ψ strongly in  L loc 2 ( N ) , ψ n ψ a.e. in  N

as λn+.

Claim 1.

We have ψ(x)=0 a.e. in RNΩ.

In fact, let Cm:={xN:a(x)>1m}. For fixed positive integer m, by (2.19), we have

C m ψ 2 𝑑 x m λ n N λ n a ( x ) ψ n 2 𝑑 x
m λ n ( N N | ψ n ( x ) - ψ n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N λ n a ( x ) ψ n 2 𝑑 x )
m λ n ( μ 1 + δ ) 0 ,

as λn+. Thus u(x)=0 a.e. in Cm for m=1,2,3,. Since m=1Cm=NΩ, it follows that u(x)=0 a.e. in NΩ.

Claim 2.

We have Ωψ2𝑑x=1.

As in Section 1 (see (1.2) in Remark 1.1), ΩBR(0) and a(x)>a0 for all |x|>R. Then by (2.19), we have

N B R ( 0 ) ψ n 2 𝑑 x 1 a 0 λ n N B R ( 0 ) λ n a ( x ) ψ n 2 𝑑 x
1 a 0 λ n ( N N | ψ n ( x ) - ψ n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N λ n a ( x ) ψ n 2 𝑑 x )
= 1 a 0 λ n N N | ψ n ( x ) - ψ n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + 1 a 0 λ n N λ n V λ n ψ n 2 𝑑 x + δ a 0 λ n N | ψ n | 2 𝑑 x
1 a 0 λ n ( μ 1 + δ ) 0

as λn+. Thus

(2.21) lim n + N B R ( 0 ) ψ n 2 𝑑 x = 0 .

Combine (2.19), (2.20), (2.21) and Claim 1 to have

Ω ψ 2 𝑑 x = B R ( 0 ) ψ 2 𝑑 x = lim n + B R ( 0 ) ψ n 2 𝑑 x = lim n + N ψ n 2 𝑑 x - lim n + N B R ( 0 ) ψ n 2 𝑑 x = 1 .

According to Claim 1 and Claim 2, we find that

(2.22) N ψ 2 𝑑 x = Ω ψ 2 𝑑 x = 1 .

Let ψn1=ψn-ψ. By (2.20) and Brézis–Lieb’s lemma,

(2.23) N ψ n 2 𝑑 x = N ψ 2 𝑑 x + N ( ψ n 1 ) 2 𝑑 x + o ( 1 )

and

(2.24) N N | ψ n ( x ) - ψ n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y = N N | ψ ( x ) - ψ ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N N | ψ n 1 ( x ) - ψ n 1 ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + o ( 1 )

as n+. Thus by (2.19), (2.22) and (2.23), we have ψnψ strongly in L2(N) as n+. By Claim 1, we obtain that ψE0. Thus by (2.15), (2.19), (2.22) and (2.24), we have

μ 1 := inf u E 0 , u L 2 ( Ω ) = 1 ( N N | u ( x ) - u ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y - Ω δ u 2 𝑑 x )
N N | ψ ( x ) - ψ ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y - Ω δ ψ 2 𝑑 x
lim n ( N N | ψ n ( x ) - ψ n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y - N δ ψ n 2 𝑑 x )
lim n μ 1 λ n μ 1 ,

which implies that μ1λnμ1 and ψnψ in Hs(N) as n.

Step 2. Suppose k2 and the results hold up to k-1. We need to prove the same results hold for the k-th eigenvalues.

Since Lλψ=L0ψ for all ψE0, by the k-th Rayleigh quotient descriptions of μkλ and μk, we have

lim sup λ + μ k λ μ k .

Just like in the case k=1, we can select λn+ and the normalized eigenfunctions ψnVkλn which is the eigenfunction space corresponding to μkλn satisfying

N N | ψ n ( x ) - ψ n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N λ n ( a ( x ) - δ ) ψ n 2 𝑑 x = μ k λ n

and

N ψ n 2 𝑑 x = 1 , ψ n V j λ n , j = 1 , 2 , 3 , , k - 1 .

Similar to the proof in Step 1, we have

{ ψ n ψ weakly in  H s ( N ) , ψ n ψ strongly in  L loc 2 ( N ) , ψ n ψ a.e. in  N

for some ψE0 satisfying Ω|ψ|2𝑑x=1. Since ψnVjλn, j=1,2,,k-1, and VjλnVj as n+, it follows that ψVj, j=1,2,,k-1, and

μ k N N | ψ ( x ) - ψ ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y - Ω δ ψ 2 𝑑 x
lim n ( N N | ψ n ( x ) - ψ n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y - Ω δ ψ n 2 𝑑 x )
lim n μ k λ n μ k .

This induces that μkλnμk,VkλnVk as n+. ∎

Corollary 2.6.

For λ large, the operator (-Δ)s+λa+a0 on Eλ is nondegenerate and has finite Morse index dj:=dimEλ- uniformly in λ.

2.4 Modified Nehari–Pankov Manifold

Recall the modified Nehari–Pankov manifold

𝒩 λ := { u E λ { 0 } : P λ - I λ ( u ) = 0 , I λ ( u ) u = 0 } E λ E 0 -

and the Nehari–Pankov manifold

𝒩 0 := { u E 0 { 0 } : P 0 - I 0 ( u ) = 0 , I 0 ( u ) u = 0 } E 0 E 0 - .

The corresponding levels are

c λ := inf u 𝒩 λ I λ ( u ) and c 0 := inf u 𝒩 0 I 0 ( u ) .

For the Nehari–Pankov manifold 𝒩0, we have the following lemma due to Szulkin and Weth [29].

Lemma 2.7.

For any ωE0E0-, set

H ω := { v + t ω : v E 0 - , t > 0 } .

The following properties hold:

  1. 𝒩 0 = { w E 0 E 0 - : ( I 0 | H w ) = 0 } .

  2. For every w E 0 + { 0 } there exist t w > 0 and φ ( w ) E 0 - such that

    H w 𝒩 0 = { φ ( w ) + t w w } .

  3. For every w 𝒩 0 and every u H w { w } there holds I 0 ( u ) < I 0 ( w ) .

  4. c 0 = inf u 𝒩 0 I 0 ( u ) > 0 .

Proof.

The proof of this lemma is similar to the proofs of [29, Lemma 2.2–2.6]; we omit the details. ∎

For the modified Nehari–Pankov manifold, we have the following lemma.

Lemma 2.8.

For any ωEλE0-, set

H ^ ω := { v + t ω : v E 0 - , t > 0 } .

The following properties hold:

  1. 𝒩 λ = { w E λ E 0 - : ( I λ | H ^ w ) = 0 } .

  2. Let

    E ^ λ + := { ω E λ + : N ω e i 𝑑 x = 0 , i = 1 , 2 , , k } .

    Then for any w E ^ λ + { 0 } there exist t w > 0 and φ ( w ) E 0 - such that

    H ^ w 𝒩 λ = { φ ( w ) + t w w } .

  3. For any w 𝒩 λ and u H ^ w { w } there holds I λ ( u ) < I λ ( w ) .

  4. c λ = inf u 𝒩 λ I λ ( u ) τ > 0 for some small τ > 0 independent of large λ.

Proof.

The proofs of (i)–(iii) are obvious. We just need to prove (iv).

Firstly, we claim that there exists a μ0>0 such that for any λ>μ0 and uEλ+, we have

2 N ( u ( x ) - u ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ u 2 𝑑 x C ( 2 N ( u ( x ) - u ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ + u 2 𝑑 x )

for some C>0 which is independent of λ.

In fact, for any uEλ+, we have u=uλ++uλ-, uλ+Eλ+, uλ-Eλ-. Note that

2 N ( u λ + ( x ) - u λ + ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ + ( u λ + ) 2 𝑑 x = 2 N ( u λ + ( x ) - u λ + ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ ( u λ + ) 2 𝑑 x + N V λ - ( u λ + ) 2 𝑑 x
μ k + 1 ( L λ ) + δ μ k + 1 ( L λ ) 2 N ( u λ + ( x ) - u λ + ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ ( u λ + ) 2 𝑑 x .

Then a simple computation shows

2 N ( u ( x ) - u ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ u 2 𝑑 x = 2 N ( u λ + ( x ) - u λ + ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ ( u λ + ) 2 𝑑 x
+ 2 N ( u λ - ( x ) - u λ - ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ ( u λ - ) 2 𝑑 x
μ k + 1 ( L λ ) μ k + 1 ( L λ ) + δ ( 2 N ( u λ + ( x ) - u λ + ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ + ( u λ + ) 2 𝑑 x )
+ 2 N ( u λ - ( x ) - u λ - ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ ( u λ - ) 2 𝑑 x
μ k + 1 ( L λ ) μ k + 1 ( L λ ) + δ ( 2 N ( u ( x ) - u ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ + u 2 𝑑 x )
+ { δ μ k + 1 ( L λ ) + δ ( 2 N ( u λ - ( x ) - u λ - ( y ) ) 2 | x - y | N + 2 s d x d y + N V λ + ( u λ - ) 2 d x )
    - μ k + 1 ( L λ ) μ k + 1 ( L λ ) + δ N V λ - [ ( u λ - ) 2 + 2 u λ + u λ - ] d x } ,

since

u λ - = i = 1 k ( N u e λ , i 𝑑 x ) e λ , i = i = 1 k [ N u ( e λ , i - e i ) 𝑑 x ] e λ , i ,

where ei and eλ,i are the eigenfunction of L0 and Lλ, respectively. Note that eλ,iei in Hs(N) for each i as λ+. Then we can easily get

2 N ( u λ - ( x ) - u λ - ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ ( u λ - ) 2 𝑑 x + N V λ - [ ( u λ - ) 2 + 2 u λ + u λ - ] 𝑑 x = o ( 1 ) u λ 2 .

Since μk+1(Lλ)μk+1(L0) as λ+, there exists μ0>λ0 such that for any λ>μ0, we have

2 N ( u ( x ) - u ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ u 2 𝑑 x μ k + 1 ( L 0 ) μ k + 1 ( L 0 ) + δ ( 2 N ( u ( x ) - u ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ + u 2 𝑑 x ) .

Secondly, let

S α := { u E λ + : 2 N ( u ( x ) - u ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ u 2 𝑑 x = α 2 } .

For any uSα, by the Sobolev inequality, we have

N | u | p C [ 2 N ( u ( x ) - u ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ + u 2 𝑑 x ] p 2
C [ 2 N ( u ( x ) - u ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ u 2 𝑑 x ] p 2 C α p .

Therefore, for α>0 small enough, we have Iλ(u)12α2-Cαp14α2>0. This implies that infSαIλ(u)>0.

Finally, for any ω𝒩λ, ω+=ω-ω-H^ω, take t>0 small enough such that tω+HωSα, thus take τ-14α2. By (iii), we have

I λ ( ω ) > I λ ( t ω + ) inf S α I λ ( u ) τ > 0 ,

which implies that cλτ>0. ∎

At the end of this subsection, in the following lemma we will prove that the minimizer for cλ is indeed a least energy solution for (${\mathcal{P}_{\lambda}}$).

Lemma 2.9.

For λ large enough, uNλ is an achieved function for cλ, i.e. cλ=Iλ(u). Then u is a least energy solution of (${\mathcal{P}_{\lambda}}$).

Proof.

Note that each solution for (${\mathcal{P}_{\lambda}}$) belongs to Nλ; thus we just need to show that the achieved function u is indeed a solution.

We denote

G 0 ( u ) = I λ ( u ) u , G i ( u ) = I λ ( u ) e i , i = 1 , 2 , , k .

Then the modified Nehari–Pankov manifold

𝒩 λ = { u E λ { 0 } : G i ( u ) = 0 , i = 1 , 2 , , k } .

According to the Lagrange Multiplier Theorem, there exists (λ0,λ1,,λk)k such that

I λ ( u ) + λ 0 G 0 ( u ) + λ 1 G 1 ( u ) + + λ k G k ( u ) = 0 .

Multiplying u and ei on both sides of the above equation, respectively, for i=1,2,,k, we have the following system:

{ a 00 λ 0 + a 0 , 1 λ 1 + + a 0 k λ k = 0 , a 10 λ 0 + a 1 , 1 λ 1 + + a 1 k λ k = 0 , a k 0 λ 0 + a k , 1 λ 1 + + a k k λ k = 0 ,

where

a 00 = ( p - 2 ) N | u | p 𝑑 x , a i i = ( p - 1 ) N | u | p - 2 u e i 2 𝑑 x - μ i ( L 0 ) N e i 2 𝑑 x ,
a 0 , i = a i , 0 = ( p - 2 ) N | u | p - 2 u e i 𝑑 x , a i j = a j i = ( p - 1 ) N | u | p - 2 e i e j 𝑑 x

for i,j=1,2,,k. It is easy to show that the coefficient matrix of the above system is positive definite. Thus the solution of the above system is (λ0,λ1,,λk)=(0,0,,0) which implies that u is indeed a solution to (${\mathcal{P}_{\lambda}}$). ∎

3 Existence of Least Energy Solutions to (1.3)

In this section, we present an existence result for the least energy solutions to (1.3). Recall the energy functional of (1.3):

I 0 ( u ) = 1 2 N N | u ( x ) - u ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y - 1 2 Ω δ u 2 𝑑 x - 1 p Ω | u | p 𝑑 x for all  u E 0 .

And the Nehari–Pankov manifold related to I0 is

𝒩 0 = { u E 0 { 0 } : P 0 - I 0 ( u ) = 0 , I 0 ( u ) u = 0 } ,

with the corresponding least level

c 0 := inf 𝒩 0 I 0 ( u ) .

We state our main result in this section as follows.

Proposition 3.1.

Assume that Ω:=int{xRN:a(x)=0} is a smooth bounded domain in RN, c0:=infN0I0(u) and one of the following conditions hold:

  1. N > 2 s , p is subcritical, i.e. 2<p<2s*,

  2. N 4 s , p is critical, i.e. p=2s*.

Then c0 is achieved by a sign-changing function which is a least energy solution to (1.3).

To show that, we get a (PS)c0 sequence by Ekeland’s variational principle and study the boundedness and compactness of the (PS)c0 sequence. We have the following lemmas.

Lemma 3.2.

Let 2<p2s*, N>2s. Then there exists a (PS)c0 sequence {un}n1, i.e.

I 0 ( u n ) c 0 , I 0 ( u n ) 0 as  n + .

Proof.

Indeed, similar arguments can be fund in the proof of [22, Theorem 5.7]. This is a direct result of Ekeland’s variational principle. For the convenience of the reader, here we give a sketch of the argument.

By the definition of c0, we can easily get a minimizing sequence {vn}𝒩0 satisfying

I 0 ( v n ) c 0  as  n .

Take

ϵ n = { I 0 ( v n ) - c 0 , if  I 0 ( v n ) > c 0 , 1 n , if  I 0 ( v n ) = c 0 .

Then by Ekeland’s variational principle, there exists a sequence {un}E0 such that

I 0 ( u n ) I 0 ( v n ) , u n - v n 0 ϵ n 1 2 , ( I 0 ( u n ) ) 1 E 0 ϵ n 1 2 ,

where E0 is the dual space of E0 and (I0(vn))1 denotes the orthogonal projection of I0(un) onto the tangent space of 𝒩0 at un. One can show that there is a constant C>0 which is independent of n such that

I 0 ( u n ) E 0 C ( I 0 ( u n ) ) 1 E 0 .

Let n+; we have I0(un)c0, I0(un)0 as n+. ∎

Lemma 3.3.

Let 2<p2s*, N>2s, {un}n1 be a (PS)c0 sequence. Then {un}n1 is bounded in E0.

Proof.

According to the definition of (PS)c0 sequences, there exists a positive integer number M such that for each n>M, we have

(3.1) c 0 + 1 + u n 0 I 0 ( u n ) - 1 2 I 0 ( u n ) u n = ( 1 2 - 1 p ) Ω | u n | p 𝑑 x

and

(3.2) c 0 + 1 + u n 0 I 0 ( u n ) - 1 p I 0 ( u n ) u n = p - 2 2 p ( N N ( u ( x ) - u ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y - δ Ω u 2 𝑑 x ) .

By Hölder’s inequality, we have

(3.3) Ω u 2 𝑑 x | Ω | 1 - 2 p ( Ω | u | p 𝑑 x ) 2 p .

Combining (3.1)–(3.3), we can easily get that for each n>M,

u n 0 2 = N N ( u ( x ) - u ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y C ( c 0 + 1 + u n 0 ) ,

where the constant C does not depend on n. Thus {un}n1 is bounded in E0. ∎

Lemma 3.4.

Assume that one of the following conditions holds:

  1. N > 2 s , p is subcritical, i.e. 2<p<2s*,

  2. N 4 s , p is critical, i.e. p=2s*.

Then the (PS)c0 condition holds. Namely, let {un}n1 be a (PS)c0 sequence. Then there must exist uE0 such that, up to a subsequence, unu strongly in E0 as n+.

Proof.

We consider two cases.

Case 1: p is subcritical, i.e. 2<p<2s*. In fact, by Lemma 3.3, there exists a function uE0, up to a subsequence, still denoted by {un}n1, such that

(3.4) { u n u weakly in  E 0 , u n u strongly in  L 2 ( Ω )  and  L p ( Ω ) , u n u a.e. in  Ω

as n+. According to the definition of (PS)c0 sequence and (3.4), we have

(3.5) o ( 1 ) = ( I 0 ( u n ) - I 0 ( u ) ) ( u n - u )
(3.6) = N N ( ( u n - u ) ( x ) - ( u n - u ) ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y - δ Ω ( u n - u ) 2 𝑑 x - Ω ( | u n | p - 2 u n - | u | p - 2 u ) ( u n - u ) 𝑑 x .

By (3.4) and the Lebesgue Dominated Theorem, we can easily get that

| u n | p - 2 u n | u | p - 2 u  strongly in  L p p - 1 ( Ω ) .

By Hölder’s inequality and (3.4), we have

(3.7) | Ω ( | u n | p - 2 u n - | u | p - 2 u ) ( u n - u ) 𝑑 x | ( Ω | | u n | p - 2 u n - | u | p - 2 u | p p - 1 𝑑 x ) p - 1 p ( Ω | u n - u | p 𝑑 x ) 1 p 0

as n+. Combining (3.4), (3.6) and (3.7), we have as n+,

N N ( ( u n - u ) ( x ) - ( u n - u ) ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y 0 .

Thus limn+un-u0=0.

Case 2: p is critical, i.e. p=2s*. To begin with, by the next lemma, we see

(3.8) 0 < c 0 < s N S N 2 s .

As proved in Lemma 3.3, {un} in bounded in E0. Taking if necessary a subsequence, we may assume that

(3.9) { u n u in  E 0  and  L 2 s * ( Ω ) , u n u strongly in  L 2 ( Ω ) , u n u a.e. in  Ω .

By (3.9) and the definition of (PS)c0 sequences, for each vE0, we have

(3.10) 0 = lim n + I 0 ( u n ) v
(3.11) = lim n + N N ( u n ( x ) - u n ( y ) ) ( v ( x ) - v ( y ) ) | x - y | N + 2 s 𝑑 x 𝑑 y - δ lim n + Ω u n v 𝑑 x - lim n + Ω | u n | 2 s * - 2 u n v 𝑑 x
(3.12) = N N ( u ( x ) - u ( y ) ) ( v ( x ) - v ( y ) ) | x - y | N + 2 s 𝑑 x 𝑑 y - δ Ω u v 𝑑 x - Ω | u | 2 s * - 2 u v 𝑑 x ,

i.e. I0(u)=0. By (3.12), we have

(3.13) I 0 ( u ) = I 0 ( u ) - 1 2 I 0 ( u ) u = ( 1 2 - 1 2 s * ) Ω | u | 2 s * 𝑑 x 0 .

Let vn=un-u, due to Brézis–Lieb’s lemma and (3.9), we have as n+,

(3.14) u n 0 2 = v n 0 2 + u 0 2 + o ( 1 ) , u n L 2 s * ( Ω ) 2 s * = v n L 2 s * ( Ω ) 2 s * + u L 2 s * ( Ω ) 2 s * .

By (3.9) and (3.14), we have

(3.15) o ( 1 ) = I 0 ( u n ) , u n
(3.16) = N N ( u n ( x ) - u n ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y - δ Ω u n 2 𝑑 x - Ω | u n | 2 s * 𝑑 x
(3.17) = N N ( u ( x ) - u ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y - δ Ω u 2 𝑑 x - Ω | u | 2 s * 𝑑 x + N N ( v n ( x ) - v n ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y - Ω | v n | 2 s * 𝑑 x + o ( 1 )
(3.18) = N N ( v n ( x ) - v n ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y - Ω | v n | 2 s * 𝑑 x + o ( 1 )

and

(3.19) I 0 ( u n ) = 1 2 N N ( u n ( x ) - u n ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y - δ 2 Ω u n 2 𝑑 x - 1 2 s * Ω | u n | 2 s * 𝑑 x
= ( 1 2 N N ( u ( x ) - u ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y - δ 2 Ω u 2 𝑑 x - 1 2 s * Ω | u | 2 s * 𝑑 x )
(3.20) + 1 2 N N ( v n ( x ) - v n ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y - 1 2 s * Ω | v n | 2 s * 𝑑 x + o ( 1 )
(3.21) = I 0 ( u ) + 1 2 N N ( v n ( x ) - v n ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y - 1 2 s * Ω | v n | 2 s * 𝑑 x + o ( 1 ) .

We assume that b=limn+Ω|vn|2s*𝑑x.

(i) If b=0, we complete the proof by (3.18).

(ii) If b>0, by (3.18) and Sobolev inequality, we have

b = lim n + Ω | v n | 2 s * 𝑑 x = lim n + N N ( v n ( x ) - v n ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y S ( lim n + Ω | v n | 2 s * 𝑑 x ) 2 2 s * = S b 2 2 s * .

Thus bSN2s. On the other hand, by (3.13), (3.18) and (3.21), we have

c 0 = lim n + I 0 ( u n ) 1 2 lim n + N N ( v n ( x ) - v n ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y - 1 2 s * lim n + Ω | v n | 2 s * 𝑑 x = ( 1 2 - 1 2 s * ) b .

Thus b<SN2s due to the estimate of c0, see (3.8). This leads to a contradiction. ∎

We prove (3.8) by the following lemma.

Lemma 3.5.

Assume that N4s, p=2s* is critical and S is defined as (2.5) in Section 2.2. Then

0 < c 0 < s N S N 2 s .

Proof.

According to (iv) in Lemma 2.7, we know that c0>0. We just need to verify that c0<sNSN2s. Without loss of generality, we may assume that 0Ω and B4αΩ for some α>0. Let

u ^ = U ( x ) U ( x ) L 2 s * ( N ) ,

where

U ( x ) := κ ( μ 2 + | x | 2 ) - N - 2 s 2 .

As we have described in Section 2.2, u^ is a minimizer for S. Let

U ϵ ( x ) = ϵ 2 s - N 2 u * ( x ϵ ) ,

where u*(x)=u^(xS12s) and ϵ>0 is small. Put uϵ=η(x)Uϵ(x), where η(x) is a smooth cutoff function satisfying

η = 1 in  B α , η = 0 in  N B 2 α , 0 η 1 .

By a standard argument which is similar to [25, Proposition 7.2], we have for N4s and ϵ>0 small enough,

(3.22) N N ( u ϵ ( x ) - u ϵ ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y - δ Ω | u ϵ ( x ) | 2 𝑑 x ( Ω | u ϵ | 2 s * 𝑑 x ) 2 2 s * < S .

Let Huϵ:=span{E0-,uϵ}. According to [25, Proposition 7.3] and some calculation, we have

M ϵ := max { N N ( u ( x ) - u ( y ) ) 2 | x - y | N + 2 s 𝑑 x 𝑑 y - δ Ω | u ( x ) | 2 𝑑 x : u H u ϵ { 0 } , Ω | u ϵ | 2 s * 𝑑 x = 1 } < S

for ϵ>0 small and N4s. For each uHuϵ{0}, we have

I 0 ( u ) max t 0 I 0 ( t u ) = s N ( u 0 2 - δ u L 2 ( Ω ) 2 u L 2 s * ( Ω ) 2 ) .

Thus

max u H u ϵ I 0 ( u ) s N M ϵ N 2 < s N S N 2 s

for ϵ>0 small and N4s. Lemma 2.7 immediately implies that c0<sNSN2s for N4s. ∎

Now we come to give the proof of Proposition 3.1.

Proof of Proposition 3.1.

According to the above lemmas, there is a sequence {un}E0 such that I0(un)c0, I0(un)0, and unu strongly in E0. This implies that I0(u)=c0, I0(u)=0. Thus u is a least energy solution to (1.3).

For δλ1, u is sign-changing. In fact, let e1 be the principle eigenfunction corresponding to the principle eigenvalue λ1 of L0 defined in E0. We may assume that e1 is positive. Multiplying e1 on both sides of equation (1.3) and integrate over N, we have

N N ( u ( x ) - u ( y ) ) ( e 1 ( x ) - e 1 ( y ) ) | x - y | N + 2 s 𝑑 x 𝑑 y - δ Ω u e 1 𝑑 x = Ω | u | p - 2 u e 1 𝑑 x .

This implies that

(3.23) ( λ 1 - δ ) Ω u e 1 𝑑 x = Ω | u | p - 2 u e 1 𝑑 x .

If u keeps sign, we may assume that u0. Thus the left side of (3.23) is nonpositive, but the right side is positive. This leads to a contradiction. Thus u changes sign. This completes the proof of Proposition 3.1. ∎

Remark 3.6.

Assume Ω=Ω1Ω2 is the interior part of the zero set a-1(0), where Ω1 and Ω2 are two smooth, connected and bounded domains in N with Ω1Ω2. Suppose u is a least energy solution of (1.3) with u(x)=0 in Ω1 and u(x)0 in Ω2. Then for any xΩ1,

( - Δ ) s u ( x ) = Ω 2 - u ( y ) | x - y | N + 2 s 𝑑 y

might be nonzero. For example, for 0<δ<λ1, u is nontrivial and keeps sign in Ω2. On one hand,

( - Δ ) s u ( x ) = Ω 2 - u ( y ) | x - y | N + 2 s 𝑑 y

is positive or negative. But on the other hand, for xΩ1,

( - Δ ) s u ( x ) = δ u ( x ) + u ( x ) p - 1 = 0 .

Thus u(x)0 in both Ω1 and Ω2. However, for the Laplacian, the local case, u=0 in Ω if and only if Δu=0. This is the difference between the local and nonlocal problems.

Remark 3.7.

Recently, Raffaella Servadei and Enrico Valdinoci considered the following problem:

(3.24) { ( - Δ ) s u ( x ) - δ u ( x ) = f ( x , u ( x ) ) in  Ω , u ( x ) = 0 in  N Ω ,

where 0<s<1, N>2s, and ΩN is a bounded domain with Lipschitz boundary and the nonlinearity f(x,u) satisfies some growth conditions. When f(x,u) is subcritical, if δ=0, they proved in [24] the existence of ground states to (3.24) by the Mountain Pass Theorem. If δλ1, they obtained in [26] a positive solution to (3.24) by the Mountain Pass Theorem for δ<λ1 and a sign-changing solution by the Linking Theorem for δλ1. When f(x,u) is critical, the authors also obtained the similar results to (3.24). For more details, please see [27] and [25].

4 The Proof of Main Results in the Subcritical Case

In this section, we focus on the proof of the main results in the subcritical case and we divide it into several subsections. To be precise, Theorem 1.3 is proved in Section 4.1 and Theorems 1.41.5 are proved in Section 4.2. Throughout this section, we always assume 2<p<2s* and N>2s.

4.1 The Proof of Theorem 1.3 in the Subcritical Case

In this subsection, we prove the existence and asymptotic behavior of least energy solutions to (${\mathcal{P}_{\lambda}}$) in the subcritical case. Similar to the proof of the previous section, we consider the existence, boundedness and the compactness of a (PS)cλ sequence. We have the following lemmas.

Lemma 4.1 (Existence of (PS)cλ Sequences).

Let λΛ0 be fixed. Then there exists a (PS)cλ sequence {un}n1.

Proof.

The proof of this lemma is similar to Lemma 3.2; we omit the details. ∎

Lemma 4.2 (Boundedness of (PS)cλ Sequences).

Let λΛ0 be fixed and let {un}n1 be a (PS)cλ sequence. Then {un}n1 is bounded in Eλ.

Proof.

According to the definition of (PS)cλ sequence, for n large,

(4.1) ( 1 2 - 1 p ) N | u n | p 𝑑 x = I λ ( u n ) - 1 2 I λ ( u n ) u n c λ + 1 + 1 2 I λ ( u n ) λ u n λ c λ + 1 + u n λ .

On the other hand, as in Section 1 (see (1.2) in Remark 1.1),

Ω B R ( 0 ) , a ( x ) > a 0 for all  | x | > R .

Then for any λΛ0>λ0, we have

V λ ( x ) = λ a ( x ) - δ > a 0 > 0 .

Thus the support of Vλ- is contained in BR(0), i.e.

(4.2) supp V λ - ( x ) B R ( 0 ) .

Since Vλ-<δ, it follows that for λΛ0, by (4.2) and Hölder’s inequality,

(4.3) N V λ - u n 2 𝑑 x = B R ( 0 ) V λ - u n 2 𝑑 x δ B R ( 0 ) u n 2 𝑑 x C ( B R ( 0 ) | u n | p 𝑑 x ) 2 p C ( N | u n | p 𝑑 x ) 2 p ,

where C is a constant which does not depend on n. By the definition of (PS)cλ sequences, for n large we have

(4.4) ( 1 2 - 1 p ) ( N N | u n ( x ) - u n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ u n 2 𝑑 x )
(4.5) = p - 2 2 p ( N N | u n ( x ) - u n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ + u n 2 𝑑 x ) - p - 2 2 p N V λ - u n 2 𝑑 x
(4.6) = I λ ( u n ) - 1 p I λ ( u n ) u n c λ + 1 + u n λ .

Combining (4.1), (4.3) and (4.6), we have for n large,

u n λ 2 = N N | u n ( x ) - u n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ + u n 2 𝑑 x C ( c λ + 1 + u n λ ) ,

where C>0 does not depend on n. Thus {un}n1 is bounded in Eλ. ∎

Lemma 4.3 (Converges of (PS)cλ Sequences).

There exists a positive constant Λ1>Λ0 such that, for λΛ1, any (PS)cλ sequence {un} of Iλ with cλc0 converges strongly in Eλ along a subsequence to a least energy solution uλ of (${\mathcal{P}_{\lambda}}$) with Iλ(uλ)=cλ, where c0 is the least energy to problem (1.3) with subcritical exponent, i.e. 2<p<2s*.

Proof.

Let {un}n1Eλ be a (PS)cλ sequence of Iλ, where λ>Λ0. Then due to Lemma 4.2, the sequence {un}n1 is bounded in Eλ. Thus up to a subsequence, there exists a function uλEλ such that

(4.7) { u n u λ weakly in  E λ , u n u λ weakly in  L p ( N ) , u n u λ strongly in  L loc q ( N ) ,  2 q < 2 s * , u n u λ a.e. in  N .

We first claim that uλ is a solution to (𝒫λ). Indeed, for any vEλ, we have

I λ ( u n ) v = N N ( u n ( x ) - u n ( y ) ) ( v ( x ) - v ( y ) ) | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ u n v 𝑑 x - N | u n | p - 2 u n v 𝑑 x
(4.8) = u n , v λ - N V λ - u n v 𝑑 x - N | u n | p - 2 u n v 𝑑 x .

Notice that suppVλ-BR(0) for λ>Λ0 and Vλ-δ (see (1.2) in Remark 1.1). Then by Hölder’s inequality and the third equality in (4.7), we have

| N V λ - ( u n - u λ ) v 𝑑 x 𝑑 y | = | B R ( 0 ) V λ - ( u n - u λ ) v 𝑑 x |
(4.9) δ B R ( 0 ) | ( u n - u λ ) v | 𝑑 x δ ( B R ( 0 ) | u n - u λ | 2 𝑑 x ) 1 2 ( N v 2 𝑑 x ) 1 2 0

as n+. Hence by (4.7), (4.8) and (4.9), we have

I λ ( u n ) v I λ ( u λ ) v as  n + .

Since Iλ(un)0 strongly in Eλ as n+, thus Iλ(uλ)=0 in Eλ. This implies that uλ is a solution to (𝒫λ). Moreover,

(4.10) I λ ( u λ ) = I λ ( u λ ) - 1 2 I λ ( u λ ) u λ = ( 1 2 - 1 p ) N | u λ | p 𝑑 x 0 .

Let vn=un-uλ, by Brézis–Lieb’s lemma, we have

(4.11) u n λ 2 = u λ λ 2 + v n λ 2 + o ( 1 ) , u n L p ( N ) p = u λ L p ( N ) p + v n L p ( N ) p + o ( 1 ) .

For λ>Λ0, since {un}n1 is bounded in Eλ, it follows from Lemma 2.1 that {un}n1 is also bounded in Hs(N). Replace v by un+uλ in (4.9), it is easy to get that as n+,

(4.12) N V λ - v n 2 𝑑 x 0 , N V λ - u n 2 𝑑 x N V λ - u λ 2 𝑑 x .

Thus by (4.11) and (4.12), as n+ we have

(4.13) I λ ( u n ) = 1 2 N N | u n ( x ) - u n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + 1 2 N V λ u n 2 𝑑 x - 1 p N | u n | p 𝑑 x
(4.14) = 1 2 u n λ 2 - 1 2 N V λ - u n 2 𝑑 x - 1 p N | u n | p 𝑑 x
(4.15) = 1 2 u n λ 2 - 1 2 N V λ - u λ 2 𝑑 x - 1 p N | u n | p 𝑑 x + o ( 1 )
(4.16) = I λ ( u λ ) + I λ ( v n ) + o ( 1 )

and

(4.17) I λ ( u n ) u n = N N | u n ( x ) - u n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ u n 2 𝑑 x - N | u n | p 𝑑 x
(4.18) = u n λ 2 - N V λ - u n 2 𝑑 x - N | u n | p 𝑑 x
(4.19) = u n λ 2 - N V λ - u λ 2 𝑑 x - N | u n | p 𝑑 x + o ( 1 )
(4.20) = I λ ( v n ) v n + o ( 1 ) .

We may assume that

b = lim n + N | v n | p 𝑑 x 0 .

If b=0, then unuλ in Lp(N) as n+, hence

lim n + N | u n | p 𝑑 x = N | u λ | p 𝑑 x .

Since Iλ(uλ)=0, it follows from (4.11) and (4.12) that

c λ = lim n + I λ ( u n ) = lim n + ( I λ ( u n ) - 1 2 I λ ( u n ) u n ) = ( 1 2 - 1 p ) lim n + N | u n | p 𝑑 x
= ( 1 2 - 1 p ) N | u λ | p 𝑑 x = I λ ( u λ ) - 1 2 I λ ( u λ ) u λ = I λ ( u λ ) .

Thus uλ is a least energy solution to (${\mathcal{P}_{\lambda}}$). Furthermore, by (4.11) and (4.12), we have

c λ = lim n + I λ ( u n ) = lim n + ( I λ ( u n ) - 1 p I λ ( u n ) ) = ( 1 2 - 1 p ) lim n + ( u n λ 2 - N V λ - | u n | 2 𝑑 x )
( 1 2 - 1 p ) ( u λ λ 2 - N V λ - | u λ | 2 𝑑 x ) = I λ ( u λ ) = c λ ,

which implies that limn+unλ2=uλλ2. Hence unuλ strongly in Eλ by (4.11).

If b>0, we claim that there exists a positive constant Λ which does not depend on λ such that bΛ. In fact, since Iλ(un)0 in Eλ as n+ and {un}n1 is bounded in Eλ, we have

(4.21) I λ ( u n ) u n 0 as  n + .

Thus by (4.20), (4.21), (4.12) and Lemma 2.1, we have

b = lim n + N | v n | p 𝑑 x = lim n + v n λ 2 C lim inf n + v n 2 C ( lim inf n + N | v n | p 𝑑 x ) 2 p = C b 2 p ,

where the constant C does not depend on λ. Select Λ:=Cpp-2, thus bΛ.

At last, we claim that there exists a constant Λ1>Λ0 such that b<12Λ for any λ>Λ1. In fact, according to (4.10), (4.16), (4.20) and (4.21), we have

c 0 c λ lim n + I λ ( v n ) = lim n + ( I λ ( v n ) - 1 2 I λ ( v n ) v n )
(4.22) = ( 1 2 - 1 p ) lim n + N | v n | p 𝑑 x = ( 1 2 - 1 p ) b .

Due to the third equality in (4.7), combining (4.20) and (4.21) we have

b = lim n + N | v n | p 𝑑 x = lim n + v n λ 2 lim n + N B R ( 0 ) V λ | v n | 2 𝑑 x
(4.23) ( λ a 0 - δ ) lim n + N B R ( 0 ) | v n | 2 𝑑 x .

Thus by (4.22) and (4.23), we have

(4.24) lim n + N B R ( 0 ) | v n | 2 𝑑 x b λ a 0 - δ C λ a 0 - δ ,

where C=2pp-2c0. According to (4.10), (4.16), (4.20), (4.21) and (4.24), we have

c 0 c λ lim n + I λ ( v n ) = lim n + ( I λ ( v n ) - 1 p I λ ( v n ) v n )
(4.25) = p - 2 2 p lim n + ( v n λ 2 - N V λ - v n 2 𝑑 x ) = p - 2 2 p lim n + v n λ 2 .

Thus by the third equality in (4.7), Hölder’s inequality, the definition of S and (4.25), we have

b = lim n + N | v n | p 𝑑 x lim n + ( N | v n | 2 𝑑 x ) p θ 2 ( N | v n | 2 s * 𝑑 x ) p ( 1 - θ ) 2 *
lim n + ( N | v n | 2 𝑑 x ) p θ 2 ( S - 1 N N | v n ( x ) - v n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y ) p ( 1 - θ ) 2
C lim n + ( N | v n | 2 𝑑 x ) p θ 2 lim n + v n λ p ( 1 - θ )
C lim n + ( N B R ( 0 ) | v n | 2 𝑑 x ) p θ 2 C ( 1 λ a 0 - δ ) p θ 2 ,

where 1p=θ2+1-θ2s* and C is independent of λ. Thus there exists Λ1>Λ0 such that b<12Λ for any λΛ1 which leads to a contradiction. ∎

The following gives the boundedness of cλ from both sides.

Lemma 4.4 (Boundedness for cλ).

For N>2s and λΛ1, we have

0 < τ < c λ c 0 .

Proof.

Since 𝒩0𝒩λ, it follows from the definition of cλ and c0 that cλc0<+, where c0 is defined as in Section 1. By (iv) in Lemma 2.7, cλ>τ>0. Therefore τ<cλc0. ∎

Proof of Theorem 1.3 in the subcritical case.

This is a direct result of Lemma 4.1 to Lemma 4.4. ∎

4.2 The Proofs of Theorem 1.4 and Theorem 1.5 in the Subcritical Case

In this subsection, we mainly focus on the study of the asymptotic behavior of the least energy solution to (${\mathcal{P}_{\lambda}}$) in the subcritical case, namely the proof of Theorem 1.4. Since the proof of Theorem 1.5 can be done similarly with the proof of Theorem 1.4, we will not give the details here.

Proof of Theorem 1.4 in the subcritical case.

Since 𝒩0𝒩λ, it follows from the definition of cλ and c0 that cλc0. Taking any sequence λn(>Λ1)+ as n+, up to a subsequence we have that cλnkc0. Let un be a least energy solution to (𝒫λn); then Iλn(un)=cn, Iλn(un)=0. We firstly claim that {un} is bounded in Hs(N). Indeed, since

(4.26) c λ n = I λ n ( u n ) - 1 2 I λ n ( u n ) u n = ( 1 2 - 1 p ) N | u n | p 𝑑 x

and

(4.27) c λ n = I λ n ( u n ) - 1 p I λ n ( u n ) u n = p - 2 2 p ( N N | u n ( x ) - u n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ u n 2 𝑑 x )

and by Hölder’s inequality and (1.2), we have

(4.28) N V λ n - u n 2 𝑑 x = B R ( 0 ) V λ n - u n 2 𝑑 x δ B R ( 0 ) u n 2 𝑑 x δ | B R ( 0 ) | 1 - 2 p ( N | u n | p 𝑑 x ) 2 p .

Then by (4.26), (4.27) and (4.28), we have

u n λ n 2 = N N | u n ( x ) - u n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ u n 2 𝑑 x + N V λ - u n 2 𝑑 x C ( c λ n + c λ n 2 p ) C ( c 0 + c 0 2 p ) ,

where the constant C>0 does not depend on n. Thus by Lemma 2.1, {un} is bounded in Hs(N). Thus up to a subsequence, we have

(4.29) { u n u weakly in  H s ( N ) , u n u strongly in  L loc q ( N )  for  2 q < 2 s * , u n u a.e. in  N .

Now we claim that u|Ωc=0, where Ωc:={x:xNΩ}.

Indeed, let Cm:={xN:a(x)>1m}, m=1,2,. Since {un}n1 is bounded in Hs(N), we have for any fixed m,

c 0 c λ n = I λ n ( u n ) = I λ n ( u n ) - 1 p I λ n ( u n ) u n
= ( 1 2 - 1 p ) ( N N | u n ( x ) - u n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ n u n 2 𝑑 x )
( 1 2 - 1 p ) ( λ n N a ( x ) u n 2 𝑑 x - δ N u n 2 𝑑 x )
( 1 2 - 1 p ) ( λ n m C m u n 2 𝑑 x - C ) .

Let n+. We have Cmun2𝑑x0. By (4.29) and Fatou’s lemma,

0 C m u 2 𝑑 x lim inf n + C m u n 2 𝑑 x = 0 .

Then u=0 a.e. in Cm. Since NΩ=m=1+Cm, it follows that u=0 a.e. in NΩ. Now we come to show that uE0 is a solution to (1.3). Indeed, according to (4.29) , we have for each vE0,

I λ n ( u n ) v = N N ( u n ( x ) - u n ( y ) ) ( v ( x ) - v ( y ) ) | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ n u n v 𝑑 x - N | u n | p - 2 u n v 𝑑 x
= N N ( u n ( x ) - u n ( y ) ) ( v ( x ) - v ( y ) ) | x - y | N + 2 s 𝑑 x 𝑑 y - δ Ω u n v 𝑑 x - Ω | u n | p - 2 u n v 𝑑 x
N N ( u ( x ) - u ( y ) ) ( v ( x ) - v ( y ) ) | x - y | N + 2 s 𝑑 x 𝑑 y - δ Ω u v 𝑑 x - Ω | u | p - 2 u v 𝑑 x = I 0 ( u ) v

as n+. Since Iλn(un)=0, it follows that I0(u)=0. Hence, u is a solution to (1.2). Moreover,

(4.30) I 0 ( u ) = I 0 ( u ) - 1 2 I 0 ( u ) u = ( 1 2 - 1 2 s * ) N | u | 2 s * 𝑑 x 0 .

At this moment, we want to show that, up to subsequence, unu strongly in Lp(N) as n+. We do it by a contradiction argument. Let vn=un-u and we assume that limn+N|vn|p𝑑x=b>0. By (4.29) and Brézis–Lieb’s lemma,

u n 2 = u 2 + v n 2 + o ( 1 ) , N | u n | p 𝑑 x = N | u | p 𝑑 x + N | v n | p 𝑑 x + o ( 1 ) as  n + .

Then we can easily get

(4.31) 0 = I λ n ( u n ) u n = I 0 ( u ) u + I λ n ( v n ) v n + o ( 1 ) = I λ n ( v n ) v n + o ( 1 )

and

(4.32) I λ n ( u n ) = I 0 ( u ) + I λ n ( v n ) + o ( 1 ) .

According to (4.30), (4.31)and (4.32), we have

(4.33) c 0 c λ n = I λ n ( u n ) I λ n ( v n ) + o ( 1 ) = I λ n ( v n ) - 1 2 I λ n ( v n ) v n + o ( 1 )
(4.34) = ( 1 2 - 1 p ) N | v n | p 𝑑 x + o ( 1 ) = ( 1 2 - 1 p ) b + o ( 1 ) .

By (4.29), (4.31) and (1.2), we have

b = N | v n | p 𝑑 x + o ( 1 ) = v n λ n 2 + o ( 1 ) N B R ( 0 ) V λ n | v n | 2 𝑑 x + o ( 1 ) ( λ n a 0 - δ ) N B R ( 0 ) | v n | 2 𝑑 x + o ( 1 ) .

Let n+. By (4.34), we have

lim n + N B R ( 0 ) | v n | 2 𝑑 x = 0 .

Then by (4.29), we have

(4.35) lim n + N | v n | 2 𝑑 x = 0 .

Combining Hölder’s inequality, Sobolev imbedding inequality (4.29) and (4.35), we have

N | v n | p 𝑑 x ( N | v n | 2 𝑑 x ) p θ 2 ( N | v n | 2 s * 𝑑 x ) p ( 1 - θ ) 2 * C ( N | v n | 2 𝑑 x ) p θ 2 v n p ( 1 - θ ) 2 0

as n+. Thus b=0 which contradict to our assumption b>0.

At last we show that u is indeed a least energy solution to (1.3) and I0(u)=c0. Indeed,

lim n + I λ n ( v n ) = lim n + ( I λ n v n - 1 2 I λ n ( v n ) v n ) = ( 1 2 - 1 p ) lim n + N | v n | p 𝑑 x = 0 .

Thus by (4.32), we have I0(u)=kτ>0. Therefore, c0k=I0(u)c0, i.e. I0(u)=c0 and u is a least energy solution to (1.3). This completes the proof of Theorem 1.4. ∎

Proof of Theorem 1.5 in the subcritical case.

Let c^0:=lim supλ+Iλ(uλ), c^λ:=Iλ(uλ), where uλ is a nontrivial solution to (${\mathcal{P}_{\lambda}}$). Similar to the proof of Theorem 1.4, we complete the proof. ∎

5 The Proof of Main Results in the Critical Case

In this section, we deal with the proof of the main theorems in the critical case. In Section 5.1 we prove Theorem 1.3, and Theorems 1.41.5 are proved in Section 5.2. Throughout this section, we assume p=2s* and N4s without especially stated.

5.1 The Proof of Theorem 1.3 in the Critical Case

In this subsection, by showing series lemmas, we complete the proof of Theorem 1.3 in the critical case. These lemmas deal with the existence, boundedness, compactness of (PS)cλ sequences in the critical case and we list them as follows.

Lemma 5.1 (Existence of (PS)cλ Sequences).

Let λΛ0 fixed. Then there exists a (PS)cλ sequence {un}n1, i.e.

I λ ( u n ) c λ , I λ ( u n ) 0 as  n + .

Lemma 5.2 (Boundedness of (PS)cλ Sequences).

Let λΛ0 be fixed and let {un}n1 be a (PS)cλ sequence. Then {un}n1 is bounded in Eλ.

Remark 5.3.

The proofs of Lemma 5.1 and Lemma 5.2 are similar to the proofs of Lemma 4.1 and Lemma 4.2, respectively, just replacing p by 2s*; we omit them.

The following lemma is about the compactness of (PS)cλ sequences in the critical case.

Lemma 5.4 (Converges of (PS)cλ Sequences).

For λΛ0, {un}n1 is a (PS)cλ sequence of Iλ with cλ<sNSN2s, where S is the best embedding constant for Hs(RN)L2s*(RN). Then up to a subsequence, there exists uλEλ such that unuλ in Eλ. Furthermore, uλ is a least energy solution of (Pλ) which satisfies Iλ(uλ)=cλ.

Proof.

By Lemma 5.2, we know that {un}n1 is bounded in Eλ. Thus there is a function uλEλ such that, up to a subsequence,

(5.1) { u n u λ weakly in  E λ , u n u λ weakly in  L 2 * ( N ) , u n u λ strongly in  L loc 2 ( N ) , u n u λ a.e. in  N .

Similar to the proof of Lemma 4.3, by (5.1), one can easily check that Iλ(uλ)=0 and Iλ(uλ)0. Then uλ is a solution to (${\mathcal{P}_{\lambda}}$). Let vn=un-uλ. Then by the first and second equality in (5.1) and Brézis–Lieb’s lemma, we have as n+,

(5.2) u n λ 2 = u λ λ 2 + v n λ 2 + o ( 1 ) ,
(5.3) N | u n | 2 s * 𝑑 x = N | u λ | 2 s * 𝑑 x + N | v n | 2 s * 𝑑 x + o ( 1 ) .

Up to a subsequence, we may assume that b=limn+N|vn|2s*𝑑x. By (1.2), the third and fourth equality in (5.1), we can easily get

lim n + N V λ - u n 2 𝑑 x = N V λ - u λ 2 𝑑 x , lim n + N V λ - ( u n - u λ ) 2 𝑑 x = 0 .

By (5.2) and (5.3), one has

(5.4) o ( 1 ) = I λ ( u n ) u n = I λ ( u λ ) u λ + I λ ( v n ) v n + o ( 1 ) = I λ ( v n ) v n + o ( 1 ) ,
(5.5) I λ ( u n ) = I λ ( u λ ) + I λ ( v n ) + o ( 1 )

as n+. Thus on one hand, by (5.4) and (5.5), we have

c λ lim n + I λ ( v n ) = lim n + ( I λ ( v n ) - 1 2 I λ ( v n ) v n ) = ( 1 2 - 1 p ) lim n + N | v n | 2 s * 𝑑 x = s N b .

Taking into account of cλ<sNSN2s, we have b<SN2s. On the other hand, by (5.1) and (5.4), we have

b = lim n + N | v n | 2 s * 𝑑 x
(5.6) = lim n + ( N N | v n ( x ) - v n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ v n 2 𝑑 x )
(5.7) = lim n + N N | v n ( x ) - v n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + lim n + N B R ( 0 ) V λ v n 2 𝑑 x + lim n + B R ( 0 ) V λ v n 2 𝑑 x
(5.8) = lim n + N N | v n ( x ) - v n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + lim n + N B R ( 0 ) V λ v n 2 𝑑 x + 0
(5.9) lim n + N N | v n ( x ) - v n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y S lim n + ( N | v n | 2 s * 𝑑 x ) 2 2 s * .

If b>0, then bSN2s, which contradicts b<SN2s. Hence b=0. Replace b by 0 in (5.9), we can easily get limn+vnλ=0. Thus unuλ strongly in Eλ. Furthermore, by (5.3), we have

I λ ( u λ ) = I λ ( u ) - 1 2 I λ ( u λ ) u λ = ( 1 2 - 1 2 s * ) N | u λ | 2 s * 𝑑 x = ( 1 2 - 1 2 s * ) lim n + N | u n | 2 s * 𝑑 x = lim n + I λ ( u n ) = c λ .

Thus uλ is a least energy solution to (1.3). ∎

The boundedness of cλ from two sides is given by the following lemma.

Lemma 5.5 (Boundedness for cλ).

For N4s and λΛ0, we have

0 < τ < c λ < s N S N 2 s .

Proof.

By the definition of cλ we know that cλc0, where c0 is defined as in Section 1. Thus by Lemma 5.5, we know that c0<sNSN2s. By (iv) in Lemma 2.8, we have τ<cλ<sNSN2s. ∎

Proof of Theorem 1.3 in the critical case.

Select Λ2:=Λ0. This is a direct result of Lemma 5.1, Lemma 5.2, Lemma 5.4 and Lemma 5.5. ∎

5.2 The Proofs of Theorem 1.4 and Theorem 1.5 in the Critical Case

In this subsection, we first prove Theorem 1.4 in the critical case. Since the proof of Theorem 1.5 in the critical case is also similar to Theorem 1.4 in the critical case, we just give a sketch.

Proof of Theorem 1.4 in the critical case.

Since 𝒩0𝒩λ, it follows from the definition of cλ and c0 that we can take any sequence λn(>Λ2)+ as n+ so that we have up to a subsequence cλnkc0. Let un be a least energy solution to (𝒫λn). Then Iλn(un)=cn, Iλn(un)=0. Similar to the subcritical case, we can prove that {un} is bounded in Hs(N). Thus, there exists a function u such that

(5.10) { u n u weakly in  H s ( N ) , u n u weakly in  L 2 s * ( N ) , u n u strongly in  L loc 2 ( N ) , u n u a.e. in  N .

Indeed, using similar arguments as in the subcritical case, we have u|Ωc=0, where Ωc=:{x:xNΩ}. On the other hand, uE0 is a solution to equation (1.3) with I0(u)0. Now we come to show that, up to a subsequence, unu in L2s*(N) as n+.

We prove this by a contradiction argument. Assume limn+N|vn|2s*𝑑x=b>0, where vn=un-u. By Brézis–Lieb’s lemma,

(5.11) u n 2 = u 2 + v n 2 + o ( 1 ) ,
(5.12) N | u n | 2 s * 𝑑 x = N | u | 2 s * 𝑑 x + N | v n | 2 s * 𝑑 x + o ( 1 )

as n+. Then by the third equality in (5.10), (5.11) and (5.12), we can easily get

(5.13) 0 = I λ n ( u n ) u n = I 0 ( u ) u + I λ n ( v n ) v n + o ( 1 ) = I λ n ( v n ) v n + o ( 1 )

and

(5.14) I λ n ( u n ) = I 0 ( u ) + I λ n ( v n ) + o ( 1 ) .

On the one hand, by (5.13) and (5.14), we have

k = lim n + c λ n = lim n + I λ n ( u n ) lim n + I λ n ( v n )
= lim n + ( I λ n ( v n ) - 1 2 I λ n ( v n ) v n ) = ( 1 2 - 1 2 s * ) N | v n | 2 s * 𝑑 x = s N b .

Thus by the third equality in (5.10), (5.13) and the definition of S, we have

(5.15) b = lim n + N | v n | 2 s * 𝑑 x = lim n + ( N N | v n ( x ) - v n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ n v n 2 𝑑 x )
(5.16) = lim n + ( N N | v n ( x ) - v n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ n + v n 2 𝑑 x - N V λ n - v n 2 𝑑 x )
(5.17) = lim n + ( N N | v n ( x ) - v n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y + N V λ n + v n 2 𝑑 x ) - 0
(5.18) lim n + N N | v n ( x ) - v n ( y ) | 2 | x - y | N + 2 s 𝑑 x 𝑑 y
(5.19) S ( lim n + N | v n | 2 s * 𝑑 x ) 2 2 s * = S b 2 2 s * .

Then bSN2s. Therefore, c0ksNbsNSN2s>c0 which leads to a contradiction and thus we have that unu in Hs(N) as n+. Furthermore, uE0 is a solution of (1.3) and

I 0 ( u ) = I 0 ( u ) - 1 2 I 0 ( u ) u = ( 1 2 - 1 2 s * ) N | u | 2 s * 𝑑 x = lim n + I λ n ( u n ) = lim n + c λ n = k .

Since k>0, it follows that I0(u)c0. Thus c0k=I0(u)c0, i.e. I0(u)=c0. This implies that u is a least energy solution to (1.3). ∎

Proof of Theorem 1.5 in the critical case.

Let c^0:=lim supλ+Iλ(uλ) and c^λ:=Iλ(uλ), where uλ is a nontrivial solution to (𝒫λ). The left arguments are the same as the proof of Theorem 1.4 in the critical case and we omit it. ∎


Communicated by Zhi-Qiang Wang


Award Identifier / Grant number: 11571040

Funding statement: The first author was supported by National Science Foundation of China (11571040).

Acknowledgements

This paper was partially finished while the first author visiting the Mathematical Department of College of Staten Island and also Graduate Center at CUNY, He wishes to express his gratitude to Professor Marcello Lucia for the enlightening discussion and the hospitality during his stay.

References

[1] C. O. Alves, D. C. De Morais Filho and M. A. S. Souto, Multiplicity of positive solutions for a class of problems with critical growth in N, Proc. Edinb. Math. Soc. (2) 52 (2009), 1–21. 10.1017/S0013091507000028Search in Google Scholar

[2] C. O. Alves, A. B. Nóbrega and M. Yang, Multi-bump solutions for Choquard equation with deepening potential well, Calc. Var. Partial Differential Equations 55 (2016), Article ID 48. 10.1007/s00526-016-0984-9Search in Google Scholar

[3] B. Barrios, E. Colorado, A. de Pablo and U. Sánchez, On some critical problems for the fractional Laplacian operator, J. Differential Equations 252 (2012), 6133–6162. 10.1016/j.jde.2012.02.023Search in Google Scholar

[4] T. Bartsch, A. Pankov and Z. Wang, Nonlinear Schrödinger equations with steep pontential well, Commun. Contemp. Math. 3 (2001), 549–569. 10.1142/S0219199701000494Search in Google Scholar

[5] T. Bartsch and Z. Tang, Multibump solutions of nonlinear Schrödinger equations with steep potential well and indefinite potential, Discrete Contin. Dyn. Syst. 33 (2013), 7–26. 10.3934/dcds.2013.33.7Search in Google Scholar

[6] T. Bartsch and Z. Wang, Existence and multiplicity results for some superlinear elliptic problems on N, Comm. Partial Differential Equations 20 (1995), 1725–1741. 10.1080/03605309508821149Search in Google Scholar

[7] X. Cabré and Y. Sire, Nonlinear equations for fractional Laplacians I: Regularity, maximum principles, and Hamiltonian estimates, Ann. Inst. H. Poincaré Anal. Non Linéaire 31 (2014), 23–53. 10.1016/j.anihpc.2013.02.001Search in Google Scholar

[8] L. Caffarelli, S. Salsa and L. Silvestre, Regularity estimates for the solution and the free boundary of the obstacle problem for the fractional Laplacian, Invent. Math. 171 (2008), no. 2, 425–461. 10.1007/s00222-007-0086-6Search in Google Scholar

[9] L. Caffarelli and L. Silvestre, An extension problem related to the fractional Laplacian, Comm. Partial Differential Equations 32 (2007), 1245–1260. 10.1080/03605300600987306Search in Google Scholar

[10] M. Cheng, Bound state for the fractional Schrödinger equation with unbounded potential, J. Math. Phys. 53 (2012), Article ID 043507. 10.1063/1.3701574Search in Google Scholar

[11] W. Choi, S. Kim and K. Lee, Asymptotic behavior of solutions for nonlinear elliptic problems with the fractional Laplacian, J. Funct. Anal. 266 (2014), 6531–6598. 10.1016/j.jfa.2014.02.029Search in Google Scholar

[12] A. Cotsiolis and N. Tavoularis, Best constants for Sobolev inequalities for higher order fractional derivatives, J. Math. Anal. Appl. 295 (2004), 225–236. 10.1016/j.jmaa.2004.03.034Search in Google Scholar

[13] J. Dávila, M. Del Pino and J. Wei, Concentrating standing waves for the fractional nonlinear Schrödinger equation, J. Differential Equations 256 (2014), 858–892. 10.1016/j.jde.2013.10.006Search in Google Scholar

[14] Y. Ding, Variational Methods for Strongly Indefinite Problems, Interdiscip. Math. Sci. 7, World Scientific, Hackensack, 2007. 10.1142/6565Search in Google Scholar

[15] Y. Ding and J. Wei, Semi-classical states for nonlinear Schrödinger equations with sign-changing potentials, J. Funct. Anal. 251 (2007), 546–572. 10.1016/j.jfa.2007.07.005Search in Google Scholar

[16] R. L. Frank and E. Lenzmann, Uniqueness and nondegeneracy of ground states for (-Δ)sQ+Q-Qα+1=0 in , Acta Math. 210 (2013), 261–318. 10.1007/s11511-013-0095-9Search in Google Scholar

[17] R. L. Frank and E. Lenzmann, Uniqueness of radial solutions for the fractional Laplacian, Comm. Pure Appl. Math. (2015), 10.1002/cpa.21591. 10.1002/cpa.21591Search in Google Scholar

[18] Y. Guo and Z. Tang, Sign changing bump solutions for Schrödinger equations involving critical growth and indefinite potential wells, J. Differential Equations 259 (2015), 6038–6071. 10.1016/j.jde.2015.07.015Search in Google Scholar

[19] T. Jin, Y. Li and J. Xiong, On a fractional Nirenberg problem. I: Blow up analysis and compactness of solutions, J. Eur. Math. Soc. (JEMS) 16 (2014), 1111–1171. 10.4171/JEMS/456Search in Google Scholar

[20] E. D. Nezza, G. Palatucci and E. Valdinoci, Hitchhiker’s guide to the fractional Sobolev space, Bull. Sci. Math. 136 (2012), 521–573. 10.1016/j.bulsci.2011.12.004Search in Google Scholar

[21] M. Niu and Z. Tang, Least energy solutions for nonlinear Schödinger equation involving half Laplacian and potential wells, Commun. Pure Appl. Anal. 15 (2016), 1215–1231. 10.3934/cpaa.2016.15.1215Search in Google Scholar

[22] A. Pankov, Periodic nonlinear Schrödinger equations with application to photonic crystals, Milan J. Math. 257 (2005), 563–574. 10.1007/s00032-005-0047-8Search in Google Scholar

[23] M. Reed and B. Simon, Methods of Modern Mathematical Physics. Vol. IV, Academic Press, New York, 1978. Search in Google Scholar

[24] R. Servadei and E. Valdinoci, Mountain pass solutions for nonlocal elliptic operators, J. Math. Anal. Appl. 389 (2012), 887–898. 10.1016/j.jmaa.2011.12.032Search in Google Scholar

[25] R. Servadei and E. Valdinoci, The Yamabe equation in a nonlocal setting, Adv. Nonlinear Anal. 2 (2013), 235–270. 10.1515/anona-2013-0008Search in Google Scholar

[26] R. Servadei and E. Valdinoci, Variational methods for nonlocal operators of elliptic type, Discrete Contin. Dyn. Syst. 33 (2013), 2105–2137. 10.3934/dcds.2013.33.2105Search in Google Scholar

[27] R. Servadei and E. Valdinoci, The Brezis–Nerenberg result for the fractional Laplacian, Trans. Amer. Math. Soc. 367 (2015), 67–102. 10.1090/S0002-9947-2014-05884-4Search in Google Scholar

[28] L. Silvestre, Regularity of the obstacle problem for a fractional power of the Laplace operator, Comm. Pure Appl. Math. 60 (2007), 67–112. 10.1002/cpa.20153Search in Google Scholar

[29] A. Szulkin and T. Weth, Ground state solutions for some indefinite variational problems, J. Funct. Anal. 257 (2009), 3802–3822. 10.1016/j.jfa.2009.09.013Search in Google Scholar

[30] S. Yan, J. Yang and X. Yu, Equations involving fractional Laplacian operator: Compactness and application, J. Funct. Anal. 269 (2015), 47–79. 10.1016/j.jfa.2015.04.012Search in Google Scholar

Received: 2016-02-21
Revised: 2016-12-02
Accepted: 2016-12-13
Published Online: 2017-01-12
Published in Print: 2017-07-01

© 2017 Walter de Gruyter GmbH, Berlin/Boston

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

Downloaded on 29.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2016-6015/html
Scroll to top button