Home Least energy sign-changing solutions for Schrödinger-Poisson systems with potential well
Article Open Access

Least energy sign-changing solutions for Schrödinger-Poisson systems with potential well

  • Xiao-Ping Chen and Chun-Lei Tang EMAIL logo
Published/Copyright: August 25, 2022

Abstract

In this article, we investigate the existence of least energy sign-changing solutions for the following Schrödinger-Poisson system

Δ u + V ( x ) u + K ( x ) ϕ u = f ( u ) , x R 3 , Δ ϕ = K ( x ) u 2 , x R 3 ,

where the functions V ( x ) , K ( x ) have finite limits as x satisfying some mild assumptions. By combining variational methods with the global compactness lemma, we obtain a least energy sign-changing solution with exactly two nodal domains, and its energy is strictly larger than twice that of least energy solutions.

MSC 2010: 35A15; 35B38; 35J50

1 Introduction and main results

In this article, we are concerned with the following Schrödinger-Poisson system:

Δ u + V ( x ) u + K ( x ) ϕ u = f ( u ) , x R 3 , Δ ϕ = K ( x ) u 2 , x R 3 , ( SP )

where the functions V ( x ) , K ( x ) , and the nonlinearity f satisfy the following assumptions:

( V 0 ) V C ( R 3 , R ) , 0 < V lim x V ( x ) < + , and there exists α 0 > 0 such that

(1.1) α 0 inf u H 1 ( R 3 ) { 0 } R 3 ( u 2 + V ( x ) u 2 ) d x R 3 u 2 d x > 0 ;

( K 0 ) K L 2 ( R 3 ) , there exist γ > 0 and C K > 0 such that

0 K ( x ) C K e γ x , for all x R 3 , K ( x ) 0 ;

  1. f C ( R , R ) and lim t 0 f ( t ) t = 0 ;

  2. lim t f ( t ) t 4 t = 0 ;

  3. there exists ζ > 0 such that F ( ± ζ ) > 0 , where F ( ζ ) = 0 ζ f ( s ) d s ;

  4. f ( t ) t 3 is a non-decreasing function on t ( , 0 ) ( 0 , ) .

System ( SP ) or the more general one

(1.2) Δ u + V ( x ) u + K ( x ) ϕ u = f ( x , u ) , x R 3 , Δ ϕ = K ( x ) u 2 , x R 3 ,

has been studied extensively, where V , K C ( R 3 , R ) and f C 1 ( R 3 × R , R ) . The Schrödinger-Poisson system, of the form similar to system (1.2), also known as the Schrödinger-Maxwell system, was first proposed by Benci and Fortunato [3] to describe the interaction of a charged particle with the electrostatic field in quantum mechanics. For more detailed mathematical and physical background of the Schrödinger-Poisson system, we refer the readers to [4] and references therein.

When K ( x ) 0 , system (1.2) is reduced to the following well-known Schrödinger equation:

(1.3) Δ u + V ( x ) u = f ( x , u ) , x R 3 ,

which has received increasingly more attention under various assumptions on the potential and the nonlinearities. For more results about problem (1.3), we also refer the readers to [12,17,20] and references therein.

In recent years, there has been increasing attention to system (1.2), especially on the existence of positive solutions, least energy (or ground state) solutions, multiple solutions, sign-changing solutions, fractional problems, and so on, see [2,3,8, 9,10,11,13,22,23, 24,25,26, 27,29,30] and references therein. Since R 3 ϕ u u 2 d x is homogeneous of degree 4, the greatest part of the literature focuses on system (1.2) assume that f ( x , u ) either satisfies the 3-superlinear condition:

( SF ) lim t F ( x , t ) t 4 = + , where F ( x , t ) = 0 t f ( x , s ) d s ,

or satisfies the following well-known Ambrosetti-Rabinowitz (AR) condition:

( AR ) there exists μ > 4 such that 0 < μ F ( x , t ) f ( x , t ) t , for all t R 3 \ { 0 } .

When inf x R 3 V ( x ) > 0 , the authors always assumed that V C ( R 3 , R ) is radial or satisfies suitable assumptions to ensure the compactness of the embedding

H u H 1 ( R 3 ) : R 3 V ( x ) u 2 d x < + L q ( R 3 )

for all q [ 2 , 6 ) . Recently, via the variational methods and the Brouwer degree, Wang and Zhou [27] proved that system (1.2) when f ( x , u ) replaced by u p 1 u ( 3 < p < 5 ) has a least energy sign-changing solution with V C ( R 3 , R + ) satisfying lim x V ( x ) = . Afterward, combining variational methods with the quantitative deformation lemma, Shuai and Wang [22] proved that system (1.2) has a least energy sign-changing solution when f ( x , u ) f ( u ) satisfies ( f 1 ), ( f 2 ), (SF), and the following monotonicity condition:

( NE ) f ( t ) t 3 is an increasing function on t ( , 0 ) ( 0 , ) ,

and the potential V C ( R 3 , R + ) satisfies some assumptions that guarantee the compactness of the Sobolev embedding. More results about the existence of sign-changing solutions with the compactness of the Sobolev embedding, we refer the readers to [13] and references therein.

When V C ( R 3 , R ) satisfies lim x V ( x ) = V sup x R 3 V ( x ) inf x R 3 V ( x ) > 0 , based on variational methods in association with the quantitative deformation lemma, Alves et al. [2] proved that system ( SP ) possesses a least energy sign-changing solution, when the nonlinearity f C 1 ( R , R ) satisfies ( f 1 ), ( f 2 ), (NE), and (SF), and the potential V : R 3 R is locally Hölder continuous and satisfies the following assumption:

( V 1 ) there exist R 0 > 0 and ρ : ( R 0 , ) ( 0 , ) a non-increasing function such that

lim r ρ ( r ) e ϱ r = , for any ϱ > 0 , V ( x ) V ρ ( x ) , for any x R 0 .

We especially refer to [9], where Chen and Tang studied the following Schrödinger-Poisson system:

(1.4) Δ u + V ( x ) u + ϕ u = a ( x ) f ( u ) , x R 3 , Δ ϕ = u 2 , x R 3 ,

when V C ( R 3 , R ) satisfies ( V 0 ) , ( V 1 ) and a C ( R 3 , R ) satisfies the following condition:

( a 0 ) inf x R 3 a ( x ) = a lim x a ( x ) > 0 ,

they proved that system (1.4) possesses a least energy sign-changing solution with exactly two nodal domains, where f satisfies ( f 1 ), ( f 2 ), (NE), and the following super-quadratic condition:

( CF ) lim t F ( t ) t 3 = + , where F ( t ) = 0 t f ( s ) d s .

Under the same assumptions on f , when V satisfies ( V 0 ) and the following assumption:

( V 2 ) there exist μ > 0 and C V > 0 such that

V ( x ) V C V e μ x , for all x R 3 .

Chen and Tang [9] also obtained the existence of least energy sign-changing solutions for system (1.4). The key point in [2,9] is to overcome the lack of compactness of the Sobolev embedding, and authors chose a class of least energy sign-changing critical points { u R } of the corresponding functional in B R ( 0 ) as a minimizing sequence for the nodal level, then proved that u R converges toward a sign-changing critical point of as R , provided that m < c + c . Here we must point out that the methods to estimate energy used in [2,9] heavily rely on that ϕ w , ϕ v have the same decay as w , v (where w , v denote the positive least energy solutions of system (1.4) and the limit problem, respectively). To the best of our knowledge, even if w has compact support, ϕ w decays at infinity as 1 x , the exponential decay of ϕ w , ϕ v cannot hold. More results about the existence of sign-changing solutions of system ( SP ) or similar systems, we refer the readers to [10,11,26,30] for the critical case, [29] for 3-linear case, [13] for the quadratic, and sub-quadratic case, and so on.

Motivated by the aforementioned works, we will further study the existence of least energy sign-changing solutions for system ( SP ) with lack of compactness and improve the results received in [2,9] by using some much weaker conditions ( f 3 ) and ( f 4 ). As we know, there are no works concerning the existence of least energy sign-changing solutions for system ( SP ) with ( f 3 ) until now.

Now, we are ready to state our main results.

Theorem 1.1

Assume that ( V 0 ) , ( V 1 ) , ( K 0 ) , and ( f 1 )–( f 4 ) hold. If γ < 2 V , then system ( SP ) possesses a least energy sign-changing solution which has exactly two nodal domains, and its energy is strictly larger than twice that of least energy solutions.

Theorem 1.2

Assume that ( V 0 ) , ( V 2 ) , ( K 0 ) , and ( f 1 )–( f 4 ) hold. If μ < min { γ , V 2 } , and γ < 2 V , then all the conclusions of Theorem 1.1remain true.

Remark 1.3

Compared with [1,2,9], we prove the existence of least energy sign-changing solutions for system ( SP ) without compactness by using a much weaker growth condition ( f 3 ). Note that ( f 4 ) is weaker than the usual monotonicity condition (NE).

Remark 1.4

Conditions ( V 1 ) , ( V 2 ) , and ( K 0 ) are key to estimate energy (see Lemmas 3.2 and 3.3). Observe that by ( K 0 ) , for any x R 3 ,

0 K ( x ) C K , lim x K ( x ) = 0 .

There are many functions satisfying ( V 0 ) and ( V 1 ) , such as V ( x ) = V e x θ with V > 1 and 0 < θ < 1 , and a simple example of V ( x ) satisfying ( V 0 ) and ( V 2 ) is given by V ( x ) = V e μ x with V > 1 and μ > 0 . Let f ( t ) = t 3 1 + t for t R , then for some constant C > 0 ,

F ( t ) = 1 3 t 3 + 1 2 t 2 t + ln 1 + t + C , if t 0 , 1 3 t 3 + 1 2 t 2 + t + ln 1 t + C , if t < 0 .

It is evident to prove that f satisfies conditions ( f 1 ) ( f 4 ) , but not satisfy ( SF ) or ( CF ) .

The proof is based on variational methods in association with the global compactness lemma. The main difficulties lie in three aspects. The first difficulty is to prove that the sign-changing Nehari manifold is non-empty. In this article, we assume that the nonlinearity f satisfies ( f 3 ) ((AR) or (SF) may be not hold), which means that there would be a competition between the non-local term and the nonlinearity. In view of this, we construct a non-empty set S (defined by (2.16)) and apply some new techniques to deal with this difficulty, see Lemma 2.1.

The second difficulty is to construct a sign-changing Palais-Smale (PS) sequence of at the nodal level m . Here we borrow an idea from [6] to overcome it, see Lemma 3.4.

The third difficulty is to obtain the compactness of bounded (PS) sequence of energy functional , we use the concentration-compactness principle of Lions (see [28, Lemma 1.21]) to apply the global compactness lemma for the functional , see Lemma 3.5, which contains a relationship among least energy nodal level m of , least energy level c of , and least energy level c of the limiting functional , see Lemmas 3.2 and 3.3. Such a relationship can help us to obtain compactness of bounded (PS) sequence.

This article is organized as follows. In Section 2, we present the variational framework and prove some preliminary lemmas. In Section 3, we are devoted to proving Theorems 1.1 and 1.2.

2 Preliminaries

In this section, we present some notations, work space stuff, and some preliminary lemmas which are important for proving our main results. We begin this section by giving some notations.

  • H 1 ( R 3 ) is the usual Sobolev space endowed with the norm

    u H 1 = R 3 ( u 2 + u 2 ) d x 1 2 .

  • D 1 , 2 ( R 3 ) is the completion of C 0 ( R 3 ) with respect to the norm

    u D 1 , 2 = R 3 u 2 d x 1 2 .

  • L q ( R 3 ) is the usual Lebesgue space equipped with the norm

    u q = R 3 u q d x 1 q , for all q [ 1 , + ) .

  • ” and “ ” denote the weak and strong convergence, respectively.

  • o ( 1 ) denotes a quantity which goes to zero as n .

  • B r ( x ) { y R 3 : x y < r } .

  • C , C i ( i = 1 , 2 , ) denote positive constants which may change from line to line.

  • u + ( x ) max { u ( x ) , 0 } and u ( x ) min { u ( x ) , 0 } .

  • S denotes the best Sobolev constant for the embedding of D 1 , 2 ( R 3 ) in L 6 ( R 3 ) , that is,

    (2.1) S inf u D 1 , 2 ( R 3 ) { 0 } R 3 u 2 d x R 3 u 6 d x 1 3 .

2.1 Work space stuff

Throughout this article, we define

H u H 1 ( R 3 ) : R 3 V ( x ) u 2 d x < +

equipped with the inner product

u , v R 3 ( u v + V ( x ) u v ) d x

and the corresponding norm

u R 3 ( u 2 + V ( x ) u 2 ) d x 1 2 .

Since is equivalent to the usual norm of H 1 ( R 3 ) , the embedding H L q ( R 3 ) is continuous for any q [ 2 , 6 ] , thus there exists a constant S q > 0 (only depend on q ) such that

(2.2) u q S q u .

For u H 1 ( R 3 ) , the Lax-Milgram theorem implies that there is a unique ϕ u D 1 , 2 ( R 3 ) such that

(2.3) R 3 ϕ v d x = R 3 K ( x ) u 2 v d x ,

for v D 1 , 2 ( R 3 ) , that is, ϕ u is a weak solution of Δ ϕ = K ( x ) u 2 and the representation formula

(2.4) ϕ u ( x ) = 1 4 π R 3 K ( y ) u 2 ( y ) x y d y

holds. Hölder’s inequality and (2.1) imply that

(2.5) R 3 K ( x ) u 2 v d x K 2 u 2 3 v 6 = K 2 u 6 2 v 6 S 1 2 K 2 u 6 2 v D 1 , 2 .

Moreover, by (2.2), (2.3), and (2.5), there hold

(2.6) ϕ u D 1 , 2 S 1 2 K 2 u 6 2 S 1 2 S 6 2 K 2 u 2

and

(2.7) L ϕ u ( u ) R 3 K ( x ) ϕ u u 2 d x = 1 4 π R 3 R 3 K ( x ) K ( y ) u 2 ( x ) u 2 ( y ) x y d x d y S 1 S 6 4 K 2 2 u 4 .

It can derive from (2.7) and the Fubini theorem that

L ϕ u + ( u ) = R 3 K ( x ) ϕ u + ( u ) 2 d x = R 3 K ( x ) ϕ u ( u + ) 2 d x = L ϕ u ( u + ) .

Substituting (2.4) into system ( SP ), we rewrite system ( SP ) into the following equation:

Δ u + V ( x ) u + K ( x ) ϕ u u = f ( u ) , x R 3 .

Now, we define the energy functional : H R by

( u ) = 1 2 R 3 ( u 2 + V ( x ) u 2 ) d x + 1 4 R 3 K ( x ) ϕ u u 2 d x R 3 F ( u ) d x .

Obviously, is well defined in H and of class C 1 . Moreover, for any v H ,

( u ) , v = R 3 ( u v + V ( x ) u v ) d x + R 3 K ( x ) ϕ u u v d x R 3 f ( u ) v d x .

Recall that u is a weak solution of system ( SP ) if u H such that ( u ) , v = 0 for any v H . Furthermore, if u H is a weak solution of system ( SP ) with u ± 0 , then u is a sign-changing solution of system ( SP ).

Our main purpose in this article is to prove the existence of least energy sign-changing solutions of system ( SP ), in view of this, we will consider the following minimization problem:

m inf u ( u ) ,

where

{ u H : u ± 0 , ( u ) , u + = ( u ) , u = 0 } .

Obviously, the set contains all of the sign-changing solutions of system ( SP ).

Another purpose of this article is to show the relationship between the energy of least energy sign-changing solutions of system ( SP ) and that of least energy solutions. We can seek the least energy solution of system ( SP ) in the following Nehari manifold:

N { u H { 0 } : ( u ) , u = 0 } ,

and define the following minimization problem

c inf u N ( u ) .

Using a global compactness lemma which is related to the functional and its limit functional , we can deduce that ( PS ) m condition holds. Before applying the global compactness lemma, it is crucial to consider the “limiting problem” associated with system ( SP ), which is defined by

Δ u + V u = f ( u ) , x R 3 , ( SP )

with the corresponding energy functional

(2.8) ( u ) = 1 2 u 2 R 3 F ( u ) d x , where u 2 R 3 ( u 2 + V u 2 ) d x .

By a standard argument (see [5, Theorem A.VI]), we can show that C 1 ( H , R ) with

( u ) , v = R 3 ( u v + V u v ) d x R 3 f ( u ) v d x , for any v H .

Analogously, define the Nehari manifold and the least energy corresponding to ( SP ) as follows:

N { u H { 0 } : ( u ) , u = 0 } , c inf u N ( u ) .

2.2 Preliminary lemmas

In this subsection, we present some technical lemmas. We first give some properties on the nonlinearity f and its primitive F ( t ) = 0 t f ( s ) d s . Combining ( f 1 ) with ( f 2 ) , for any ε > 0 and p ( 2 , 6 ) , there exists C ε > 0 such that, for any t R ,

(2.9) f ( t ) ε t + C ε t 5 ,

(2.10) f ( t ) ε ( t + t 5 ) + C ε t p 1 .

Condition ( f 4 ) implies that, for any t R ,

(2.11) ( t ) 1 4 f ( t ) t F ( t ) 0 ,

and ( t ) is non-decreasing when t > 0 and non-increasing when t < 0 , which implies that

(2.12) f ( t ) t 2 3 f ( t ) t 0 .

Using ( f 4 ) again, one can easily check that for any t 1 and τ R ,

(2.13) f ( t τ ) t τ t 4 f ( τ ) τ ,

which means that for any t 1 and τ R ,

(2.14) L ϕ t τ ( t τ ) R 3 f ( t τ ) t τ d x = t 4 L ϕ τ ( τ ) R 3 f ( t τ ) t τ d x t 4 L ϕ τ ( τ ) R 3 f ( τ ) τ d x .

Different from the 3-superlinear case, in order to show and N , we have to overcome the competing effect of the non-local term. Now, we define sets S 0 and S as follows:

(2.15) S 0 u H : L ϕ u ( u ) R 3 f ( u ) u d x < 0 ,

(2.16) S u H : L ϕ u ( u ± ) R 3 f ( u ± ) u ± d x < 0 .

Lemma 2.1

Assume that ( V 0 ) , ( K 0 ) , and ( f 1 )–( f 4 ) hold. Then S and S 0 . Moreover, S and N S 0 .

Proof

We first claim that there exists z H such that

(2.17) R 3 F ( z ± ) d x > 0 .

Borrowing the method in [5, Proof of Theorem 2], for R > 0 , we define

w ( R , x ) ζ , for x R , ζ ( R + 1 r ) , for r = x [ R , R + 1 ] , 0 , for x R + 1 .

Then, w ( R , x ) H . It is not difficult to check that

R 3 F ( ± w ( R , x ) ) d x F ( ± ζ ) meas ( B R ( 0 ) ) meas ( B R + 1 ( 0 ) B R ( 0 ) ) ( max γ [ 0 , ζ ] F ( ± γ ) ) ,

where meas ( ) denotes the Lebesgue measure. Hence, there exist constants C 1 , C 2 > 0 such that

R 3 F ( ± w ( R , x ) ) d x C 1 R 3 C 2 R 2 .

So there exists R ˜ > 1 sufficiently large such that R 3 F ( ± w ( R ˜ , x ) ) d x > 0 . Let z ± ± w ( R ˜ , x ) , then (2.17) holds.

For any fixed u H , let u t ( x ) u ( t x ) for t 1 , then (2.11) and (2.13) imply that

(2.18) L ϕ t u t ( t u t ± ) R 3 f ( t u t ± ) t u t ± d x C K 2 R 3 R 3 ( t u t ( y ) ) 2 ( t u t ± ( x ) ) 2 4 π x y d x d y R 3 f ( t u t ± ) t u t ± d x = C K 2 t 1 R 3 R 3 u 2 ( y ) ( u ± ( x ) ) 2 4 π x y d x d y t 3 R 3 f ( t u ± ) t u ± d x C K 2 t 1 R 3 R 3 u 2 ( y ) ( u ± ( x ) ) 2 4 π x y d x d y t R 3 f ( u ± ) u ± d x C K 2 t 1 R 3 R 3 u 2 ( y ) ( u ± ( x ) ) 2 4 π x y d x d y 4 t R 3 F ( u ± ) d x .

Note that there exists z H such that R 3 F ( z ± ) d x > 0 , then (2.17) and (2.18) imply that

L ϕ t z t ( t z t ± ) R 3 f ( t z t ± ) t z t ± d x C K 2 t 1 R 3 R 3 z 2 ( y ) ( z ± ( x ) ) 2 4 π x y d x d y 4 t R 3 F ( z ± ) d x ,

as t + . Thus, taking v = t z t with t sufficiently large such that v S , which implies that S . Obviously, S S 0 , then S 0 . Moreover, by the definition of and N , it is easy to deduce that S and N S 0 . Thus, we complete the proof of Lemma 2.1.□

With the help of Lemma 2.1, we will prove that and N .

Lemma 2.2

Assume that ( V 0 ) , ( K 0 ) , and ( f 1 )–( f 4 ) hold. If u S , then there exists a unique pair ( s u , t u ) of positive numbers such that s u u + + t u u . If u S 0 , then there exists a unique t ( u ) > 0 such that t ( u ) u N .

Proof

First, we prove the existence of ( s u , t u ) . For s , t 0 , let ψ 1 ( s , t ) ( s u + + t u ) , s u + and ψ 2 ( s , t ) ( s u + + t u ) , t u , that is,

(2.19) ψ 1 ( s , t ) = s 2 u + 2 + s 4 L ϕ u + ( u + ) R 3 f ( s u + ) s u + d x + s 2 t 2 L ϕ u ( u + ) ,

(2.20) ψ 2 ( s , t ) = t 2 u 2 + t 4 L ϕ u ( u ) R 3 f ( t u ) t u d x + s 2 t 2 L ϕ u + ( u ) .

From (2.14) and (2.19), for any s 1 and t 0 , it follows that

(2.21) ψ 1 ( s , t ) s 2 u + 2 + s 4 L ϕ u + ( u + ) R 3 f ( u + ) u + d x + s 2 t 2 L ϕ u ( u + ) = s 2 u + 2 + s 4 L ϕ u ( u + ) R 3 f ( u + ) u + d x + s 2 ( t 2 s 2 ) L ϕ u ( u + ) .

For any fixed t 0 , using ( f 1 ) , ( f 2 ) , and (2.21), it is easy to verify that ψ 1 ( 0 , t ) = 0 , ψ 1 ( s , t ) > 0 for s > 0 sufficiently small and ψ 1 ( s , t ) < 0 for s > 0 sufficiently large due to u S . Thus, there exist 0 < r < R such that

(2.22) ψ 1 ( r , r ) > 0 , ψ 2 ( r , r ) > 0 ; ψ 1 ( R , R ) < 0 , ψ 2 ( R , R ) < 0 .

By (2.19) and (2.20), we obtain that the functions ψ 1 ( s , ) and ψ 2 ( , t ) are increasing in R + for any fixed s > 0 and t > 0 , respectively. Then, we can conclude from (2.22) that

ψ 1 ( r , t ) > 0 , ψ 1 ( R , t ) < 0 , for any t [ r , R ] ; ψ 2 ( s , r ) > 0 , ψ 2 ( s , R ) < 0 , for any s [ r , R ] .

Consequently, ( ψ 1 , ψ 2 ) ( 0 , 0 ) on the boundary of ( r , R ) × ( r , R ) . Applying Miranda’s theorem [19], there exists ( s u , t u ) ( r , R ) × ( r , R ) such that ψ 1 ( s u , t u ) = 0 and ψ 2 ( s u , t u ) = 0 . That is, s u u + + t u u , which implies that .

Second, we prove the uniqueness of the pair ( s u , t u ) . The proof will be divided into two cases.

Case 1: u .

If u , which implies that

(2.23) u ± 2 + L ϕ u ( u ± ) = R 3 f ( u ± ) u ± d x .

Now, we prove ( s u , t u ) = ( 1 , 1 ) is the unique pair of positive numbers such that s u u + + t u u . We can assume ( s 0 , t 0 ) is another pair of positive numbers such that s 0 u + + t 0 u , without loss of generality, we may suppose that 0 < s 0 t 0 . Hence, we obtain

(2.24) s 0 2 u + 2 + s 0 4 L ϕ u ( u + ) s 0 2 u + 2 + s 0 4 L ϕ u + ( u + ) + s 0 2 t 0 2 L ϕ u ( u + ) = R 3 f ( s 0 u + ) s 0 u + d x ,

(2.25) t 0 2 u 2 + t 0 4 L ϕ u ( u ) t 0 2 u 2 + t 0 4 L ϕ u ( u ) + s 0 2 t 0 2 L ϕ u + ( u ) = R 3 f ( t 0 u ) t 0 u d x .

Combining (2.23) with (2.24), we deduce that

(2.26) 1 1 s 0 2 u + 2 R 3 f ( u + ) ( u + ) 3 f ( s 0 u + ) ( s 0 u + ) 3 u + 4 d x .

If s 0 < 1 , the left side of (2.26) is negative, which is absurd because the right side is non-negative by ( f 4 ) . So, we obtain 1 s 0 t 0 . Meanwhile, by (2.23) and (2.25), one has

(2.27) 1 1 t 0 2 u 2 R 3 f ( u ) ( u ) 3 f ( t 0 u ) ( t 0 u ) 3 u 4 d x .

If t 0 > 1 , the left side of (2.27) is positive, which is a contradiction because the right side is non-positive by ( f 4 ) . So, we obtain 0 < s 0 t 0 1 . Consequently, s 0 = t 0 = 1 .

Case 2: u S .

If u S , we know there exists ( s u , t u ) of positive numbers such that s u u + + t u u . We can assume there exists another pair ( s , t ) of positive numbers such that s u + + t u . Let ν s u u + + t u u and ν s u u + + t u u , then

ν = s u s u s u u + + t u t u t u u = s s u ν + + t t u ν .

Thanks to ν , we have s u s u = t u t u = 1 , which yields to s u = s u , t u = t . Thus, we assert that ( s u , t u ) is the unique pair of positive numbers such that s u u + + t u u .

Similarly, we obtain that for all u S 0 , there exists a unique t ( u ) > 0 such that t ( u ) u N .□

Lemma 2.3

Assume that ( V 0 ) , ( K 0 ) , and ( f 1 )–( f 4 ) hold. Let u H with u ± 0 , then

  1. if ( u ) , u ± 0 , then 0 < s u , t u 1 , where ( s u , t u ) is obtained by Lemma 2.2;

  2. if { u n } , lim n ( u n ) = m , then m > 0 and C 1 u n ± C 2 for some C 1 , C 2 > 0 .

Proof

(1) Since s u u + + t u u , without loss of generality, suppose that 0 < t u s u , then

(2.28) s u 2 u + 2 + s u 4 L ϕ u ( u + ) R 3 f ( s u u + ) s u u + d x .

The assumption ( u ) , u + 0 yields that

(2.29) u + 2 + L ϕ u ( u + ) R 3 f ( u + ) u + d x .

Combining (2.28) with (2.29), we obtain

(2.30) 1 1 s u 2 u + 2 R 3 f ( u + ) ( u + ) 3 f ( s u u + ) ( s u u + ) 3 u + 4 d x .

If s u > 1 , the left side of (2.30) is positive, which is absurd because the right side is non-positive by ( f 4 ) . Thus, s u 1 . Hence, 0 < t u s u 1 , and finishes the proof of (1) of Lemma 2.3.

( 2 ) On one hand, by using (2.2), (2.9), and the fact that { u n } , we can arrive at

u n ± 2 R 3 f ( u n ± ) u n ± d x ε R 3 u n ± 2 d x + C ε R 3 u n ± 6 d x ε S 2 2 u n ± 2 + C ε S 6 6 u n ± 6 .

Choosing ε > 0 such that ( 1 ε S 2 2 ) > 0 , which implies that u n ± C 1 for some C 1 > 0 .

On the other hand, for any { u n } N , we obtain ( u n ) , u n = 0 . By (2.11), one has

m + o ( 1 ) = ( u n ) = ( u n ) 1 4 ( u n ) , u n = 1 4 u n 2 + R 3 ( u n ) d x 1 4 u n 2 1 4 u n ± 2 > 0 ,

which means that m > 0 and there is C 2 > 0 such that u n ± C 2 . Thus, C 1 u n ± C 2 .□

3 Proof of main results

In this section, we shall prove Theorems 1.1 and 1.2. This section is divided into three parts. The first part is to verify m < c + c and c < c . The second part is to construct a sign-changing ( PS ) m sequence for the functional . The third part is to prove that the ( PS ) m condition holds via a global compactness lemma.

3.1 Estimate energy on c and m

In this subsection, we are devoted to verifying m < c + c and c < c . Let w , v H be a positive least energy solution of system ( SP ) and problem ( SP ), respectively, that is,

( w ) = c , ( w ) = 0 ; ( v ) = c , ( v ) = 0 .

The exponential decay of w ( x ) , v ( x ) at infinity will be shown in the following lemma.

Lemma 3.1

For any 0 < θ < V , there exists C = C ( θ ) > 0 such that

(3.1) 0 < v ( x ) + w ( x ) C e θ x .

Proof

By applying a similar argument to [16, Theorem 1.11], it can deduce that v L ( R 3 ) and v ( x ) 0 as x . Then, for any 0 < θ < V , there exists R = R ( θ ) > 0 such that

V v f ( v ) V v ε v C ε v 5 θ 2 v , for all x R .

Thus, we have Δ v + θ 2 v 0 for all x R and there exists C = C ( θ ) > 0 such that v ( x ) C for x = R . Let v ¯ ( x ) = C e θ ( x R ) , a direct calculation derives that Δ v ¯ + θ 2 v ¯ 0 for x 0 . The maximum principle implies that v ( x ) C e θ ( x R ) for x R . Thus, v ( x ) C e θ x .

We can proceed as above (or see [15, Theorem 3.1]) to conclude that for any 0 < θ < V , there exists C = C ( θ ) > 0 such that w ( x ) C e θ x . Thus, the proof is completed.□

Setting v n ( x ) v ( x + n e 1 ) for each n N , where e 1 = ( 1 , 0 , 0 ) is the fixed unit vector in R 3 .

Lemma 3.2

Under assumptions ( V 0 ) , ( K 0 ) , and ( f 1 )–( f 4 ), suppose that ( V 1 ) or ( V 2 ) holds. Then, 0 < c < c .

Proof

By ( v ) , v = 0 and ( f 4 ) , there exist constants α > 0 sufficiently small and β sufficiently large such that

(3.2) ( α v ) , α v > 0 , ( β v ) , β v < 0 .

Observe that

R 3 ( V ( x ) V ) v n 2 d x + R 3 K ( x ) ϕ v n v n 2 d x = o ( 1 ) ,

then (3.2) implies that

(3.3) ( α v n ) , α v n ( α v ) , α v > 0 , ( β v n ) , β v n ( β v ) , β v < 0 ,

as n . In view of (3.3), we deduce that there exist n N sufficiently large and t ¯ ( α , β ) such that ( t ¯ v n ) , t ¯ v n = 0 , which implies that t ¯ v n N and ( t ¯ v n ) c .

By a direct calculation and the definition of c , we can obtain that

(3.4) c ( t ¯ v n ) = ( t ¯ v n ) + 1 2 R 3 ( V ( x ) V ) ( t ¯ v n ) 2 d x + 1 4 L ϕ t ¯ v n ( t ¯ v n ) = ( t ¯ v n ) + T 1 + T 2 .

Claim: For n sufficiently large, we have

(3.5) T 1 + T 2 < 0 .

If the above claim comes true, combining with (3.4), there holds

(3.6) c < ( t ¯ v n ) = ( t ¯ v ) ( v ) = c ,

which completes the proof of Lemma 3.2.

It suffices to verify inequality (3.5). Since γ < 2 V , we can choose 0 < θ < V such that γ < 2 θ . Then, it derives from (3.1), ( K 0 ), Hölder’s inequality, and x n e 1 n x that

(3.7) T 2 C ϕ v n 6 R 3 e γ x n e 1 v 2 6 5 d x 5 6 C e γ n R 3 e ( γ 2 θ ) x 6 5 d x 5 6 C 1 e γ n .

If V satisfies ( V 1 ) , for R 0 > 0 and ρ : a non-increasing function, we deduce that

(3.8) T 1 1 2 x n e 1 1 ( V ( x ) V ) ( t ¯ v n ) 2 d x ρ ( n + 1 ) 2 x n e 1 1 ( t ¯ v n ) 2 d x ,

for any n R 0 + 1 . (3.7), (3.8), and the boundedness of t ¯ imply that, for n sufficiently large,

(3.9) T 1 + T 2 e γ n C 1 ρ ( n + 1 ) 2 e γ n t ¯ 2 B 1 ( 0 ) v 2 d x < 0 .

If V satisfies ( V 2 ) , we can derive from x n e 1 n + x that

(3.10) T 1 C V R 3 e μ x n e 1 v 2 d x C V e μ n R 3 e μ x v 2 d x C 2 e μ n .

Since μ < γ , it derives from (3.7), (3.10), and the boundedness of t ¯ that

(3.11) T 1 + T 2 C 1 e γ n C 2 e μ n = e μ n ( C 1 e ( μ γ ) n C 2 ) < 0 ,

for n sufficiently large. Thus, when ( V 1 ) or ( V 2 ) holds, it follows from (3.9) or (3.11) that (3.5) is satisfied for n sufficiently large. This completes the proof of Lemma 3.2.□

Lemma 3.3

Under assumptions ( V 0 ) , ( K 0 ) , and ( f 1 )–( f 4 ), suppose that ( V 1 ) or ( V 2 ) holds. Then, 0 < m < c + c .

Proof

Let T [ α , β ] × [ α , β ] ( α , β are defined by Lemma 3.2). We first prove that there exists ( ξ 0 , κ 0 ) T such that ξ 0 w κ 0 v n for some large n N . Using w N and ( f 4 ) , we obtain

(3.12) ( α w ) , α w > 0 , ( β w ) , β w < 0 .

For any ξ , κ > 0 , let us define

Ψ ( ξ , κ ) ( ( ξ w κ v n ) , ( ξ w κ v n ) + , ( ξ w κ v n ) , ( ξ w κ v n ) ) .

Since v ( x ) 0 as x , it follows from (3.3) and (3.12) that there exists n 0 > 0 such that

(3.13) ( α w κ v n ) , ( α w κ v n ) + > 0 , ( β w κ v n ) , ( β w κ v n ) + < 0 ;

(3.14) ( ξ w α v n ) , ( ξ w α v n ) > 0 , ( ξ w β v n ) , ( ξ w β v n ) < 0 ,

for any n n 0 and ξ , κ [ α , β ] . Note that the function Ψ is continuous in T and by (3.13) and (3.14), we can apply Miranda’s theorem [19] to conclude that there exists ( ξ 0 , κ 0 ) T such that Ψ ( ξ 0 , κ 0 ) = ( 0 , 0 ) , which means that ξ 0 w κ 0 v n for any n n 0 . Consequently,

m ( ξ 0 w κ 0 v n ) sup ( s , t ) T ( s w t v n ) .

In view of this, it suffices to show that

sup ( s , t ) T ( s w t v n ) < c + c ,

for n sufficiently large. Indeed, by a direct calculation, we can obtain that, for any ( s , t ) T ,

(3.15) ( s w t v n ) = 1 2 s w t v n 2 + 1 4 L ϕ ( s w t v n ) ( s w t v n ) R 3 F ( s w t v n ) d x = ( s w ) + ( t v n ) + Π 1 + Π 2 + Π 3 + Π 4 ,

where

Π 1 = s t R 3 ( w v n + V ( x ) w v n ) d x , Π 2 = 1 4 L ϕ ( s w t v n ) ( s w t v n ) 1 4 L ϕ s w ( s w ) ,

Π 3 = R 3 F ( s w ) + F ( t v n ) F ( s w t v n ) d x , Π 4 = 1 2 R 3 ( V ( x ) V ) ( t v n ) 2 d x .

Claim: For n sufficiently large, we have

(3.16) Π 1 + Π 2 + Π 3 + Π 4 < 0 .

If the above claim comes true, combining with (3.15), there holds

sup ( s , t ) T ( s w t v n ) < ( s w ) + ( t v n ) ( w ) + ( v ) = c + c ,

which completes the proof of Lemma 3.3.

It suffices to prove inequality (3.16). Let us define Θ B n ( q + 1 ) ( 0 ) . For any 1 q 5 , by Hölder’s inequality and (3.1), we can obtain that

Θ w q v n d x R 3 w q + 1 d x q q + 1 Θ v n q + 1 d x 1 q + 1 C Θ e θ ( q + 1 ) x + n e 1 d x 1 q + 1 .

This, together with x + n e 1 n e 1 x = n x , we obtain that

(3.17) Θ w q v n d x C e θ n Θ e θ ( q + 1 ) x d x 1 q + 1 = C e θ n 0 n q + 1 e θ ( q + 1 ) r r 2 d r 1 q + 1 .

We now recall that, for any fixed t > 0 and N N , we have

e t r r N 1 d r = e t r P ( r ) ,

where

P ( r ) r N 1 t ( N 1 ) t 2 r N 2 + ( N 1 ) ( N 2 ) t 3 r N 3 + + ( 1 ) N + 1 ( N 1 ) ! t N .

In view of this, by taking t θ ( q + 1 ) and N 3 , then there holds

0 n q + 1 e θ ( q + 1 ) r r 2 d r = e θ n P n q + 1 = C 3 e θ n + C 4 ,

where C 3 , C 4 > 0 depend only on θ . The above estimate and (3.17) imply that

(3.18) Θ w q v n d x C e θ n e θ n q + 1 + e θ n C e θ n q q + 1 .

Analogously, by (3.1) and Hölder’s inequality, we have

(3.19) R 3 Θ w q v n d x R 3 Θ w q + 1 d x q q + 1 R 3 v ( x + n e 1 ) q + 1 d x 1 q + 1 C R 3 Θ e θ ( q + 1 ) x d x q q + 1 = C n q + 1 e θ ( q + 1 ) r r 2 d r q q + 1 C e θ n q q + 1 .

Then, (3.18) and (3.19) imply that

(3.20) R 3 w q v n d x C e θ n q q + 1 .

Meanwhile, we can proceed as above to obtain that

(3.21) R 3 w v n q d x C e θ n q q + 1 .

Since w is a positive least energy solution of system ( SP ), one has

(3.22) Π 1 = s t R 3 ( w v n + V ( x ) w v n ) d x = s t R 3 K ( x ) ϕ w w v n d x R 3 f ( w ) v n d x .

Then, (3.1), ( K 0 ) , and Hölder’s inequality, together with x + n e 1 n x imply that

(3.23) R 3 K ( x ) ϕ w w v n d x ϕ w 6 R 3 K ( x ) w v n 6 5 d x 5 6 C e θ n R 3 e γ x 6 5 d x 5 6 C e θ n .

According to (2.9), (3.20), (3.22), (3.23), and the boundedness of s , t , we can deduce that

(3.24) Π 1 s t R 3 K ( x ) ϕ w w v n d x + ε s t R 3 w v n d x + C ε s t R 3 w 5 v n d x C 5 e θ n 2 .

Indeed, by using (2.7), a simple calculation shows that

L ϕ ( s w t v n ) ( s w t v n ) = 1 4 π R 3 R 3 K ( x ) K ( y ) ( s w ( x ) t v n ( x ) ) 2 ( s w ( y ) t v n ( y ) ) 2 x y d x d y 1 4 π R 3 R 3 K ( x ) K ( y ) ( s 2 w 2 ( x ) + t 2 v n 2 ( x ) ) ( s 2 w 2 ( y ) + t 2 v n 2 ( y ) ) x y d x d y = 1 4 π R 3 R 3 K ( x ) K ( y ) s 4 w 2 ( x ) w 2 ( y ) x y d x d y + 1 4 π R 3 R 3 K ( x ) K ( y ) s 2 t 2 ( w 2 ( x ) v n 2 ( y ) + v n 2 ( x ) w 2 ( y ) ) x y d x d y + 1 4 π R 3 R 3 K ( x ) K ( y ) t 4 v n 2 ( x ) v n 2 ( y ) x y d x d y = L ϕ s w ( s w ) + 2 L ϕ s w ( t v n ) + L ϕ t v n ( t v n ) .

Then, it follows from the above inequality that

(3.25) Π 2 1 2 L ϕ s w ( t v n ) + 1 4 L ϕ t v n ( t v n ) = s 2 t 2 2 L ϕ w ( v n ) + t 4 4 L ϕ v n ( v n ) .

Since γ < 2 V , choosing 0 < θ < V satisfying γ < 2 θ . Thus, it derives from (3.1), ( K 0 ) , Hölder’s inequality, and x n e 1 n x that

(3.26) L ϕ w ( v n ) ϕ w 6 R 3 e γ x n e 1 v 2 6 5 d x 5 6 C e γ n R 3 e ( γ 2 θ ) x 6 5 d x 5 6 C e γ n ,

(3.27) L ϕ v n ( v n ) ϕ v n 6 R 3 e γ x n e 1 v 2 6 5 d x 5 6 C e γ n R 3 e ( γ 2 θ ) x 6 5 d x 5 6 C e γ n .

Thus, we can deduce from the boundedness of s , t , and (3.25)–(3.27) that

(3.28) Π 2 C 6 e γ n .

In view of (2.9), (3.20), (3.21), and [10, Lemma 3.2], we conclude that

(3.29) Π 3 4 R 3 ( f ( s w ) t v n + f ( s w ) t v n + f ( t v n ) s w + f ( t v n ) s w ) d x C R 3 ( s w t v n + s w 5 t v n + s w t v n 5 ) d x C 7 e θ n 2 .

If V satisfies ( V 1 ) , for R 0 > 0 and ρ : a non-increasing function given by ( V 1 ) , we deduce that

(3.30) Π 4 1 2 x n e 1 1 ( V ( x ) V ) ( t v n ) 2 d x ρ ( n + 1 ) 2 x n e 1 1 ( t v n ) 2 d x ,

for any n R 0 + 1 . Taking λ > 0 such that λ min { γ , θ 2 } , it follows from (3.24), (3.28)–(3.30) that, for n sufficiently large,

(3.31) Π 1 + Π 2 + Π 3 + Π 4 C 5 e θ n 2 + C 6 e γ n + C 7 e θ n 2 ρ ( n + 1 ) 2 x n e 1 1 ( t v n ) 2 d x e λ n C 6 e ( λ γ ) n + C 8 e λ θ 2 n ρ ( n + 1 ) 2 e λ n t 2 B 1 ( 0 ) v 2 d x < 0 .

If V satisfies ( V 2 ) , we can derive from x n e 1 n + x that

(3.32) Π 4 C V R 3 e μ x n e 1 v 2 d x C V e μ n R 3 e μ x v 2 d x C 9 e μ n .

Since μ < V 2 , choosing 0 < θ < V such that 2 μ < θ , it derives from μ < γ , (3.24), (3.28), (3.29), and (3.32) that

(3.33) Π 1 + Π 2 + Π 3 + Π 4 C 5 e θ n 2 + C 6 e γ n + C 7 e θ n 2 C 9 e μ n = e μ n C 6 e ( μ γ ) n + C 10 e ( 2 μ θ ) n 2 C 9 < 0 ,

for n sufficiently large. Thus, when ( V 1 ) or ( V 2 ) holds, it follows from (3.31) or (3.33) that (3.16) is satisfied for n sufficiently large. This completes the proof of Lemma 3.3.□

3.2 Construction of the sign-changing ( PS ) m sequence

In this subsection, we are devoted to constructing a sign-changing ( PS ) m sequence, we borrow an idea from [6] to construct it. Let P be the cone of nonnegative functions in H , Q [ 0 , 1 ] × [ 0 , 1 ] and Σ be the set of continuous maps σ such that, for any s , t [ 0 , 1 ] ,

  1. σ ( s , 0 ) = 0 , σ ( 0 , t ) P and σ ( 1 , t ) P ;

  2. ( σ ) ( s , 1 ) 0 , R 3 f ( σ ( s , 1 ) ) ( σ ( s , 1 ) ) d x σ ( s , 1 ) 2 + L ϕ σ ( s , 1 ) ( σ ( s , 1 ) ) 2 .

For each u H with u ± 0 , let σ ( s , t ) = k t ( 1 s ) u + + k t s u , where k > 0 , s , t [ 0 , 1 ] . It is easy to check that σ ( s , t ) Σ for k > 0 sufficiently large, which means that Σ . Define

l ( u , v ) R 3 f ( u ) u d x u 2 + L ϕ u ( u ) + L ϕ v ( u ) , if u 0 ; 0 , if u = 0 .

And u if and only if l ( u + , u ) = l ( u , u + ) = 1 . Let us define

U u H : 1 2 < l ( u + , u ) < 3 2 , 1 2 < l ( u , u + ) < 3 2 .

Lemma 3.4

There exists a sequence { u n } U such that ( u n ) m and ( u n ) 0 .

Proof

We divide the proof into three claims.

Claim 1: inf σ Σ sup u σ ( Q ) ( u ) = inf u ( u ) = m .

In fact, for any u , there exists σ ( s , t ) = k t ( 1 s ) u + + k t s u Σ for k > 0 sufficiently large. Then, from Lemma 2.2, we can arrive at

( u ) = max s , t 0 ( s u + + t u ) sup u σ ( Q ) ( u ) inf σ Σ sup u σ ( Q ) ( u ) ,

and then

inf u ( u ) inf σ Σ sup u σ ( Q ) ( u ) .

To complete Claim 1, it remains to prove that for each σ Σ , there is u σ σ ( Q ) such that

sup u σ ( Q ) ( u ) ( u σ ) inf u ( u ) ,

which implies that

inf σ Σ sup u σ ( Q ) ( u ) inf u ( u ) .

Indeed, for each σ Σ and t [ 0 , 1 ] , it is easy to verify that σ ( 0 , t ) P and σ ( 1 , t ) P , thus

(3.34) l ( σ + ( 0 , t ) , σ ( 0 , t ) ) l ( σ ( 0 , t ) , σ + ( 0 , t ) ) = l ( σ + ( 0 , t ) , σ ( 0 , t ) ) 0 ,

(3.35) l ( σ + ( 1 , t ) , σ ( 1 , t ) ) l ( σ ( 1 , t ) , σ + ( 1 , t ) ) = l ( σ ( 1 , t ) , σ + ( 1 , t ) ) 0 .

Meanwhile, from the definition of Σ , for any σ Σ and s [ 0 , 1 ] , using the elementary inequality b a + d c b + d a + c for any a , b , c , d > 0 , we can deduce that

l ( σ + ( s , 1 ) , σ ( s , 1 ) ) + l ( σ ( s , 1 ) , σ + ( s , 1 ) ) R 3 f ( σ ( s , 1 ) ) ( σ ( s , 1 ) ) d x σ ( s , 1 ) 2 + L ϕ σ ( s , 1 ) ( σ ( s , 1 ) ) 2 .

Thus,

(3.36) l ( σ + ( s , 1 ) , σ ( s , 1 ) ) + l ( σ ( s , 1 ) , σ + ( s , 1 ) ) 2 0 ,

(3.37) l ( σ + ( s , 0 ) , σ ( s , 0 ) ) + l ( σ ( s , 0 ) , σ + ( s , 0 ) ) 2 = 2 < 0 .

Combining Miranda’s theorem [19] with (3.34)–(3.37), there exists ( s σ , t σ ) Q such that

0 = l ( σ + ( s σ , t σ ) , σ ( s σ , t σ ) ) l ( σ ( s σ , t σ ) , σ + ( s σ , t σ ) ) = l ( σ + ( s σ , t σ ) , σ ( s σ , t σ ) ) + l ( σ ( s σ , t σ ) , σ + ( s σ , t σ ) ) 2 ,

then

l ( σ + ( s σ , t σ ) , σ ( s σ , t σ ) ) = l ( σ ( s σ , t σ ) , σ + ( s σ , t σ ) ) = 1 .

That is, for any σ Σ , there exists u σ = σ ( s σ , t σ ) σ ( Q ) . Hence,

inf σ Σ sup u σ ( Q ) ( u ) = inf u ( u ) = m .

Claim 2: There exists a ( PS ) m sequence { u n } H for the functional .

Consider a minimizing sequence { w n } and σ n ( s , t ) = k t ( 1 s ) w n + + k t s w n Σ , then

lim n max w σ n ( Q ) ( w ) = lim n ( w n ) .

Using a variant form of the classical deformation lemma [21] due to Hofer [14], we claim that there exists { u n } H such that, as n ,

(3.38) ( u n ) m , ( u n ) 0 , dist ( u n , σ n ( Q ) ) 0 .

Suppose by contradiction, there is δ > 0 such that σ n ( Q ) V δ = for n sufficiently large, where

V δ = { u H : v H , s . t . v u δ , ( v ) δ , ( v ) m δ } .

By [14], there exists a map η C ( [ 0 , 1 ] × H , H ) satisfying for some ε ( 0 , m 2 ) and all t [ 0 , 1 ] ,

  1. η ( 0 , u ) = u , η ( t , u ) = η ( t , u ) ;

  2. η ( t , u ) = u , for any u m ε ( H m + ε ) ;

  3. η ( 1 , m + ε 2 V δ ) m ε 2 ;

  4. η ( 1 , ( m + ε 2 P ) V δ ) m ε 2 P , where d = { u H : ( u ) d } .

Since lim n max w σ n ( Q ) ( w ) = m , we can choose n sufficiently large such that

(3.39) σ n ( Q ) m + ε 2 , σ n ( Q ) V δ = .

Let us define σ ˜ n ( s , t ) η ( 1 , σ n ( s , t ) ) for all ( s , t ) Q . Then we declare that σ ˜ n Σ , it derives from (3.39) and property (iii) of η that σ ˜ n ( Q ) m ε 2 , which leads to a contradiction with

m = inf σ Σ sup w σ ( Q ) ( w ) max w σ ˜ n ( Q ) ( w ) m ε 2 .

Indeed, property (ii) of η and σ n Σ imply σ ˜ n ( s , 0 ) = η ( 1 , σ n ( s , 0 ) ) = η ( 1 , 0 ) = 0 . On one hand, it follows from σ n ( 0 , t ) P , (3.39), and property (iv) of η that σ ˜ n ( 0 , t ) P . On the other hand, thanks to σ n ( 1 , t ) P and (3.39), we see that σ n ( 1 , t ) ( m + ε 2 P ) V δ , which implies that σ ˜ n ( 1 , t ) = η ( 1 , σ n ( 1 , t ) ) P in view of properties (i) and (iv) of η . Then, σ ˜ n satisfies property ( I ) . Moreover, using the fact that ( σ n ) ( s , 1 ) 0 and property (ii) of η , we have σ ˜ n ( s , 1 ) = η ( 1 , σ n ( s , 1 ) ) = σ n ( s , 1 ) , then σ ˜ n satisfies property ( ii ) . Therefore, it can deduce from the continuity of η and σ n that σ ˜ n Σ .

Claim 3: The sequence { u n } obtained in Claim 2 satisfies { u n } U for n sufficiently large.

Since ( u n ) 0 , we have ( u n ) , u n ± = o ( 1 ) . Then it suffices to prove that u n ± 0 , which means that l ( u n + , u n ) 1 , l ( u n , u n + ) 1 , and thus { u n } U for n sufficiently large. By (3.38), there exists a sequence { ν n } such that

(3.40) ν n = s n w n + + t n w n σ n ( Q ) , ν n u n 0 .

Hence, in order to obtain u n ± 0 , we show that s n w n + 0 and t n w n 0 for n sufficiently large. By { w n } and Lemma 2.3(2), we have C 1 w n ± C 2 for some C 1 , C 2 > 0 . Now, we only need to show that s n 0 and t n 0 for n sufficiently large. Suppose by contradiction that s n 0 , by the continuity of and (3.40), we obtain

m = lim n ( ν n ) = lim n ( s n w n + + t n w n ) = lim n ( t n w n ) .

Then, by (2.9), Lemma 2.2, and C 1 w n ± C 2 , choosing ε > 0 such that ( 1 ε S 2 2 ) > 0 , thus

m = lim n ( w n ) = lim n max s , t > 0 ( s w n + + t w n ) lim n max s > 0 ( s w n + + t n w n ) = lim n max s > 0 s 2 2 w n + 2 + s 4 4 L ϕ w n + ( w n + ) R 3 F ( s w n + ) d x + s 2 t n 2 2 L ϕ w n ( w n + ) + ( t n w n ) lim n max s > 0 s 2 2 w n + 2 R 3 F ( s w n + ) d x + lim n ( t n w n ) lim n max s > 0 ( 1 ε S 2 2 ) s 2 2 w n + 2 C ε S 6 6 s 6 6 w n + 6 + m max s > 0 { C 3 s 2 C 4 s 6 } + m > m ,

which leads to a contradiction for some C 3 , C 4 > 0 . Then { u n } U for n sufficiently large.□

3.3 The compactness condition

In this subsection, we are ready to establish a global compactness lemma inspired by [25, Lemma 4.2] (see also [24, Lemma 4.3]), which is a key point to prove our main results, then prove the ( PS ) m condition holds.

Lemma 3.5

[Global compactness lemma] Assume that ( V 0 ) , ( K 0 ) , and ( f 1 ) ( f 4 ) hold. Let { u n } U be a bounded ( PS ) m sequence for the functional , then there exists u 0 H such that u n u 0 with ( u 0 ) = 0 , and integer l N { 0 } , sequence { y n k } R 3 , non-trivial points w k H for k = 1 , 2 , , l such that

  1. y n k + and y n k y n k + for k k ;

  2. w k 0 and ( w k ) = 0 for k = 1 , 2 , , l ;

  3. u n u 0 k = 1 l w k ( y n k ) 0 ;

  4. ( u n ) ( u 0 ) + k = 1 l ( w k ) .

Moreover, if l = 0 , then u n u 0 in H without w k and { y n k } .

Proof

Let us prove that ( u 0 ) = 0 . Since { u n } is bounded in H , we may assume that, up to a subsequence, u n u 0 in H , u n u 0 in L loc q ( R 3 ) for q [ 1 , 6 ) , and u n ( x ) u 0 ( x ) a . e . in R 3 . In order to prove that ( u 0 ) = 0 . It is sufficient to check that ( u 0 ) , φ = 0 for all φ C 0 ( R 3 ) . Observe that

(3.41) ( u n ) , φ ( u 0 ) , φ = u n u 0 , φ + R 3 K ( x ) ( ϕ u n u n ϕ u 0 u 0 ) φ d x R 3 ( f ( u n ) f ( u 0 ) ) φ d x .

Recalling that u n u 0 in H , then u n u 0 , φ 0 .

By using (2.1), (2.6), [24, Lemma 2.4], and Hölder’s inequality, we can deduce that

R 3 K ( x ) ( ϕ u n u n ϕ u 0 u 0 ) φ d x = R 3 K ( x ) ϕ u n u 0 ( u n u 0 ) φ d x + o ( 1 ) C K φ R 3 ϕ u n u 0 6 d x 1 6 supp φ u n u 0 6 5 d x 5 6 + o ( 1 ) C K S 1 2 φ ϕ u n u 0 D 1 , 2 supp φ u n u 0 6 5 d x 5 6 + o ( 1 ) C K S 1 S 6 2 φ K 2 u n u 0 2 supp φ u n u 0 6 5 d x 5 6 + o ( 1 ) C supp φ u n u 0 6 5 d x 5 6 0 .

In view of this, it remains to prove that

(3.42) R 3 ( f ( u n ) f ( u 0 ) ) φ d x 0 .

By (2.9) and Young’s inequality, we conclude that

( f ( u n ) f ( u 0 ) ) φ u n φ + u 0 φ + C 1 u n 5 φ + C 2 u 0 5 φ u n u 0 φ + 2 u 0 φ + C 1 u n u 0 5 φ + C 3 u 0 5 φ ε u n u 0 2 + C ε φ 2 + 2 u 0 φ + ε u n u 0 6 + C 1 C ε φ 6 + C 3 u 0 5 φ .

Let us define G ε , n ( x ) max { ( f ( u n ) f ( u 0 ) ) φ ε u n u 0 2 ε u n u 0 6 , 0 } , then

0 G ε , n ( x ) C ε φ 2 + 2 u 0 φ + C 1 C ε φ 6 + C 3 u 0 5 φ L 1 ( R 3 ) ,

and G ε , n ( x ) 0 a.e. in R 3 . By the Lebesgue dominated convergence theorem, one has

R 3 G ε , n ( x ) d x 0 , as n .

Therefore,

limsup n R 3 ( f ( u n ) f ( u 0 ) ) φ d x limsup n R 3 G ε , n ( x ) d x + ε limsup n R 3 u n u 0 2 d x + ε limsup n R 3 u n u 0 6 d x C 4 ε .

It follows from the arbitrariness of ε that (3.42) is satisfied. Thus, recalling that ( u n ) 0 , we deduce from (3.41) that ( u 0 ) = 0 . Now, we prove that ( u 0 ) 0 . Indeed, by (2.11), we have

(3.43) ( u 0 ) = ( u 0 ) 1 4 ( u 0 ) , u 0 = 1 4 u 0 2 + R 3 ( u 0 ) d x 1 4 u 0 2 0 .

The remaining proof will be divided into three steps.

Step 1. Set z n 1 u n u 0 , then we have z n 1 0 in H . Let us define

δ limsup n + sup y R 3 B 1 ( y ) z n 1 2 d x .

Case 1 (Vanishing): δ = 0 . That is,

sup y R 3 B 1 ( y ) z n 1 2 d x 0 .

Using the concentration-compactness principle of Lions [28, Lemma 1.21], we obtain that z n 1 0 in L q ( R 3 ) for any q [ 2 , 6 ) . Since ( u 0 ) = 0 , we have

z n 1 2 = ( u n ) , z n 1 ( u 0 ) , z n 1 + R 3 K ( x ) ( ϕ u 0 u 0 ϕ u n u n ) z n 1 d x + R 3 ( f ( u n ) f ( u 0 ) ) z n 1 d x = R 3 K ( x ) ( ϕ u 0 u 0 ϕ u n u n ) z n 1 d x + R 3 ( f ( u n ) f ( u 0 ) ) z n 1 d x + o ( 1 ) .

By using (2.1), (2.6), [24, Lemma 2.4] and Hölder’s inequality, we can deduce that

R 3 K ( x ) ( ϕ u n u n ϕ u 0 u 0 ) z n 1 d x = R 3 K ( x ) ϕ z n 1 ( z n 1 ) 2 d x + o ( 1 ) C K R 3 ϕ z n 1 6 d x 1 6 R 3 z n 1 12 5 d x 5 6 + o ( 1 ) C K S 1 2 ϕ z n 1 D 1 , 2 R 3 z n 1 12 5 d x 5 6 + o ( 1 ) C R 3 z n 1 12 5 d x 5 6 0 .

Using (2.10) and Hölder’s inequality, we deduce that

R 3 ( f ( u n ) f ( u 0 ) ) z n 1 d x R 3 [ ε ( u n + u 0 + u n 5 + u 0 5 ) + C ε ( u n p 1 + u 0 p 1 ) ] z n 1 d x ε C 5 + C ε ( u n p p 1 + u 0 p p 1 ) z n 1 p ε C 5 + C 6 C ε z n 1 p ,

where C 5 , C 6 are independent of ε and n . By the arbitrariness of ε and z n 1 p 0 , we obtain

R 3 ( f ( u n ) f ( u 0 ) ) z n 1 d x 0 , as n .

Therefore, z n 1 0 in H . Thus, z n 1 0 in H , and Lemma 3.5 holds with l = 0 .

Case 2 (Non-vanishing): δ > 0 . Assume that there exists a sequence { y n 1 } R 3 such that

B 1 ( y n 1 ) z n 1 2 d x > δ 2 > 0 .

Then, along a subsequence if necessary, we have, for some w 1 H ,

y n 1 + , z n 1 ( + y n 1 ) w 1 , ( w 1 ) = 0 .

Let us define z n 1 ˜ ( ) z n 1 ( + y n 1 ) . Then, z n 1 ˜ is bounded in H and we may assume that z n 1 ˜ w 1 in H , z n 1 ˜ w 1 in L loc q ( R 3 ) for q [ 2 , 6 ) , and z n 1 ˜ ( x ) w 1 ( x ) a.e. in R 3 . Since

B 1 ( 0 ) z n 1 ˜ 2 d x > δ 2 ,

then

B 1 ( 0 ) w 1 2 d x > δ 2 ,

and then w 1 0 . But it can derive from z n 1 0 in H that { y n 1 } must be unbounded. Passing to a subsequence, we assume that y n 1 + .

Now, it remains to prove that ( w 1 ) = 0 . Indeed, similar to the proof of (3.41), we see that

( z n 1 ˜ ) , φ ( w 1 ) , φ 0 , for any fixed φ C 0 ( R 3 ) .

Therefore, it suffices to prove that ( z n 1 ˜ ) , φ 0 for any fixed φ C 0 ( R 3 ) . We conclude that

(3.44) ( z n 1 ) , φ ( y n 1 ) = R 3 z n 1 φ ( y n 1 ) d x + R 3 V ( x ) z n 1 φ ( y n 1 ) d x + R 3 K ( x ) ϕ z n 1 z n 1 φ ( y n 1 ) d x R 3 f ( z n 1 ) φ ( y n 1 ) d x = R 3 z n 1 ˜ φ d x + R 3 V ( x + y n 1 ) z n 1 ˜ φ d x + R 3 K ( x + y n 1 ) ϕ z n 1 ˜ z n 1 ˜ φ d x R 3 f ( z n 1 ˜ ) φ d x .

Since z n 1 0 in H , similar to the proof of (3.41), we obtain that

( z n 1 ) , φ ( y n 1 ) ( 0 ) , φ ( y n 1 ) 0 ,

which implies that

( z n 1 ) , φ ( y n 1 ) 0 .

Then, it deduces from (3.44) that

(3.45) R 3 z n 1 ˜ φ d x + R 3 V ( x + y n 1 ) z n 1 ˜ φ d x + R 3 K ( x + y n 1 ) ϕ z n 1 ˜ z n 1 ˜ φ d x R 3 f ( z n 1 ˜ ) φ d x 0 .

Since y n 1 + as n and for any φ C 0 ( R 3 ) , by ( V 0 ) and ( K 0 ) , one can prove that

(3.46) R 3 ( V ( x + y n 1 ) V ) z n 1 ˜ φ d x 0 , as n ,

and

(3.47) R 3 K ( x + y n 1 ) ϕ z n 1 ˜ z n 1 ˜ φ d x 0 , as n .

Thus, we use (3.45) minus ( z n 1 ˜ ) , φ , by (3.46) and (3.47), we deduce that

( z n 1 ˜ ) , φ 0 .

Therefore, ( w 1 ) = 0 .

Step 2. In the following, we are devoted to showing that

(3.48) ( z n 1 ) = m ( u 0 ) + o ( 1 ) , ( z n 1 ) = ( u n ) ( u 0 ) + o ( 1 ) .

Indeed, it follows from the Brézis-Lieb lemma [7] that

(3.49) z n 1 2 = u n 2 u 0 2 + o ( 1 ) .

We can deduce from Lemma 2.4 in [24] that

(3.50) L ϕ z n 1 ( z n 1 ) = L ϕ u n ( u n ) L ϕ u 0 ( u 0 ) + o ( 1 ) .

Lemma 3.2 in [18] implies that

(3.51) R 3 F ( z n 1 ) d x = R 3 F ( u n ) d x R 3 F ( u 0 ) d x + o ( 1 ) ,

(3.52) R 3 f ( z n 1 ) z n 1 d x = R 3 f ( u n ) u n d x R 3 f ( u 0 ) u 0 d x + o ( 1 ) .

Then, using (3.49)–(3.52), it is easy to verify that

( z n 1 ) = m ( u 0 ) + o ( 1 ) .

Through direct calculations, we deduce that

(3.53) ( z n 1 ) = ( u n ) u n , u 0 + 1 2 u 0 2 1 2 R 3 ( V ( x ) V ) ( u n u 0 ) 2 d x 1 4 L ϕ u n ( u n ) + R 3 ( F ( u n ) F ( z n 1 ) ) d x = ( u n ) ( u 0 ) u n u 0 , u 0 1 2 R 3 ( V ( x ) V ) ( u n u 0 ) 2 d x 1 4 ( L ϕ u n ( u n ) L ϕ u 0 ( u 0 ) ) + R 3 F ( u n ) F ( z n 1 ) F ( u 0 ) d x .

In view of ( V 1 ) and the locally compactness of Sobolev embedding, we see that

(3.54) R 3 ( V ( x ) V ) ( u n u 0 ) 2 d x = o ( 1 ) .

By ( K 0 ) , we have lim x K ( x ) = 0 , then for any ε > 0 , there exists R ( ε ) > 0 such that

x R ( ε ) K ( x ) ϕ z n 1 ( z n 1 ) 2 d x ε .

We also have

lim n x R ( ε ) K ( x ) ϕ z n 1 ( z n 1 ) 2 d x C lim n x R ( ε ) ϕ z n 1 6 d x 1 6 x R ( ε ) z n 1 12 5 d x 5 6 = 0 .

Combining the aforementioned two inequalities with (3.50), we can deduce that

(3.55) L ϕ u n ( u n ) L ϕ u 0 ( u 0 ) = o ( 1 ) .

Substituting (3.51), (3.54), and (3.55) into (3.53), there holds

( z n 1 ) = ( u n ) ( u 0 ) + o ( 1 ) .

Step 3. Set z n 2 ( ) z n 1 ( ) w 1 ( y n 1 ) , then z n 2 0 in H . We can similarly check that

(3.56) ( z n 2 ) = ( u n ) ( u 0 ) ( w 1 ) + o ( 1 ) ,

(3.57) ( z n 2 ) = ( z n 1 ) ( w 1 ) + o ( 1 ) ,

(3.58) ( z n 2 ) , φ = ( u n ) , φ ( u 0 ) , φ ( w 1 ) , φ + o ( 1 ) = o ( 1 ) .

Hence, using (3.48), we see that

( u n ) = ( u 0 ) + ( z n 1 ) + o ( 1 ) = ( u 0 ) + ( w 1 ) + ( z n 2 ) + o ( 1 ) .

Then, by ( w 1 ) = 0 , similar to (3.43), we obtain ( w 1 ) 0 , together with (3.43) imply that

( z n 2 ) = ( u n ) ( u 0 ) ( w 1 ) + o ( 1 ) m .

Similar to the argument in Step 1, let us define

δ 1 limsup n + sup y R 3 B 1 ( y ) z n 2 2 d x .

If vanishing case occurs, then z n 2 0 , i.e., u n u 0 w 1 ( y n 1 ) 0 , and thus Lemma 3.5 holds with k = 1 . If non-vanishing case occurs, then there exists a sequence { y n 2 } R 3 and w 2 H such that z n 2 ˜ ( x ) = z n 2 ( x + y n 2 ) w 2 in H . By using (3.58), we can similarly check that ( w 2 ) = 0 . Furthermore, z n 2 0 in H implies that y n 2 + and y n 1 y n 2 + . By iterating this procedure, we obtain sequences of points { y n k } R 3 such that y n k + and y n k y n k + for k k and z n k = z n k 1 w k 1 ( y n k 1 ) with k 2 such that

z n k 0 in H , ( w k ) = 0

and

(3.59) u n 2 u 0 2 j = 1 k 1 w j ( y n j ) 2 = u n u 0 j = 1 k 1 w j ( y n j ) 2 + o ( 1 ) ,

(3.60) ( u n ) ( u 0 ) j = 1 k 1 ( w j ) ( z n k ) = o ( 1 ) .

Since { u n } is bounded, (3.59) and (3.60) imply that the iteration stops at some finite l + 1 N . Therefore, z n l + 1 0 in H , by (3.59) and (3.60), it is easy to verify that (3) and (4) hold.□

Based on Lemma 3.5, we can prove that the ( PS ) m condition for the functional holds.

Lemma 3.6

Under assumptions ( V 0 ) , ( K 0 ) , and ( f 1 ) ( f 4 ) , suppose that ( V 1 ) or ( V 2 ) holds. Then any sequence { u n } U satisfying

( u n ) m ( 0 , c + c ) , ( u n ) 0

contains a convergent subsequence in H.

Proof

Since { u n } is bounded in H , passing to a subsequence, we may assume that there exists u 0 H such that u n u 0 in H . In view of Lemma 3.5, there exists l N { 0 } and { y n k } R 3 with y n k + for k = 1 , 2 , , l , and u 0 H , w k H such that

( u 0 ) = 0 , u n u 0 k = 1 l w k ( y n k ) 0 and ( u n ) ( u 0 ) + k = 1 l ( w k ) ,

where w k are non-trivial critical points of for k = 1 , 2 , , l . Since c < c (see Lemma 3.2), it follows from (4) of Lemma 3.5 that l 1 , that is, l = 0 or l = 1 .

Suppose that u 0 = 0 . In this case, since m > 0 , we have that l = 1 , and therefore

(3.61) u n w 1 ( y n 1 ) 0 .

From { u n } U and the definition of U , we obtain 1 2 < l ( u n ± , u n ) < 3 2 , then (2.2) and (2.9) imply that

1 2 u n ± < R 3 f ( u n ± ) u n ± d x ε R 3 u n ± 2 d x + C ε R 3 u n ± 6 d x ε S 2 2 u n ± 2 + C ε S 6 6 u n ± 6 ,

choosing ε > 0 such that 1 2 ε S 2 2 > 0 , which implies that there exists ρ > 0 such that

(3.62) u n ± ρ > 0 .

Since y n 1 + and { u n } U , (3.61) and (3.62) imply that ( w 1 ) ± N . Hence,

c + c > m = ( w 1 ) = ( ( w 1 ) + ) + ( ( w 1 ) ) 2 c ,

contradicting to c < c . So, u 0 0 and we can use ( u 0 ) = 0 to obtain that u 0 N and ( u 0 ) c .

If l = 1 , we obtain

m = lim n ( u n ) = ( u 0 ) + k = 1 l ( w k ) = ( u 0 ) + ( w 1 ) c + c ,

which leads to a contradiction with m < c + c . Thus, l = 0 , that is, u n u 0 in H .□

Proof of Theorems 1.1 and 1.2

In view of Lemma 3.4, there exists a sequence { u n } U such that ( u n ) m and ( u n ) 0 as n . Lemmas 3.5 and 3.6 imply that { u n } admits a convergent subsequence in H . Thus, under assumptions of Theorem 1.1 or Theorem 1.2, there exists u 0 H such that u n u 0 in H , then ( u 0 ) = m and ( u 0 ) = 0 . Furthermore, from { u n } U and the definition of U , we see that 1 2 < l ( u n ± , u n ) < 3 2 , similar to the proof of (3.62), we can obtain that

u 0 ± = lim n u n ± ρ > 0 ,

that is, u 0 ± 0 . Therefore, u 0 is a least energy sign-changing solution of system ( SP ).

Next, we prove m > 2 c . By Lemma 2.2, there exist s ¯ , t ¯ > 0 such that s ¯ u 0 + , t ¯ u 0 N , then

m = ( u 0 ) ( s ¯ u 0 + + t ¯ u 0 ) = ( s ¯ u 0 + ) + ( t ¯ u 0 ) + s ¯ 2 t ¯ 2 2 L ϕ u 0 ( u 0 + ) > ( s ¯ u 0 + ) + ( t ¯ u 0 ) 2 c .

The proof is completed by showing that u 0 has exactly two nodal domains. By contradiction, we assume that u 0 has at least three nodal domains such that

u 0 = u 1 + u 2 + u 3

with u i 0 , u 1 0 , u 2 0 , and supp ( u i ) supp ( u j ) = for i j ( i , j = 1 , 2 , 3 ) and

( u 0 ) , u i = 0 , for i = 1 , 2 , 3 .

Setting v u 1 + u 2 , we deduce that v + = u 1 and v = u 2 , i.e., v ± 0 . In view of Lemma 2.2, there exists a unique pair ( s v , t v ) of positive numbers such that s v v + + t v v , then

( s v v + + t v v ) m .

Moreover, using the fact that ( u 0 ) = 0 , it follows that ( v ) , v ± < 0 , then by Lemma 2.3(1), we have that ( s v , t v ) ( 0 , 1 ] × ( 0 , 1 ] . On the other hand, by (2.11), we conclude that

(3.63) 0 = 1 4 ( u 0 ) , u 3 = 1 4 u 3 2 + 1 4 L ϕ u 1 ( u 3 ) + 1 4 L ϕ u 2 ( u 3 ) + 1 4 L ϕ u 3 ( u 3 ) 1 4 R 3 f ( u 3 ) u 3 d x 1 4 u 3 2 + 1 4 L ϕ u 1 ( u 3 ) + 1 4 L ϕ u 2 ( u 3 ) + 1 4 L ϕ u 3 ( u 3 ) R 3 F ( u 3 ) d x < ( u 3 ) + 1 4 L ϕ u 1 ( u 3 ) + 1 4 L ϕ u 2 ( u 3 ) .

Then, by (3.63), we can deduce that

m ( s v u 1 + t v u 2 ) = ( s v u 1 + t v u 2 ) 1 4 ( s v u 1 + t v u 2 ) , s v u 1 + t v u 2 = s v 2 4 u 1 2 + t v 2 4 u 2 2 + R 3 ( s v u 1 ) d x + R 3 ( t v u 2 ) d x 1 4 u 1 2 + 1 4 u 2 2 + R 3 ( u 1 ) d x + R 3 ( u 2 ) d x = ( u 1 ) + ( u 2 ) + 1 2 L ϕ u 1 ( u 2 ) + 1 4 L ϕ u 1 ( u 3 ) + 1 4 L ϕ u 2 ( u 3 ) < ( u 1 ) + ( u 2 ) + ( u 3 ) + 1 2 L ϕ u 1 ( u 2 ) + 1 2 L ϕ u 1 ( u 3 ) + 1 2 L ϕ u 2 ( u 3 ) = ( u 0 ) = m ,

which is a contradiction. That is, u 3 = 0 , and u 0 has exactly two nodal domains.□

Acknowledgments

The authors would like to thank editors and the anonymous referees for a careful reading of the results and for making suggestions to improve the original manuscript.

  1. Funding information: This study was supported by National Natural Science Foundation of China (No. 11971393).

  2. Conflict of interest: The authors state no conflict of interest.

  3. Data availability statement: Data sharing is not applicable to this article as no new data were created or analyzed in this study.

References

[1] C. O. Alves and M. A. S. Souto, Existence of least energy nodal solution for a Schrödinger-Poisson system in bounded domains, Z. Angew. Math. Phys. 65 (2014), no. 6, 1153–1166.10.1007/s00033-013-0376-3Search in Google Scholar

[2] C. O. Alves, M. A. S. Souto and S. H. M. Soares, A sign-changing solution for the Schrödinger-Poisson equation, Rocky Mountain J. Math. 47 (2017), no. 1, 1–25.10.1216/RMJ-2017-47-1-1Search in Google Scholar

[3] V. Benci and D. Fortunato, An eigenvalue problem for the Schrödinger-Maxwell equations, Topol. Methods Nonlinear Anal. 11 (1998), no. 2, 283–293.10.12775/TMNA.1998.019Search in Google Scholar

[4] V. Benci and D. Fortunato, Solitary waves of the nonlinear Klein-Gordon equation coupled with the Maxwell equations, Rev. Math. Phys. 14 (2002), no. 4, 409–420.10.1142/S0129055X02001168Search in Google Scholar

[5] H. Berestycki and P.-L. Lions, Nonlinear scalar field equations. I. Existence of a ground state, Arch. Rational Mech. Anal. 82 (1983), no. 4, 313–345.10.1007/BF00250555Search in Google Scholar

[6] G. Cerami, S. Solimini and M. Struwe, Some existence results for superlinear elliptic boundary value problems involving critical exponents, J. Funct. Anal. 69 (1986), no. 3, 289–306.10.1016/0022-1236(86)90094-7Search in Google Scholar

[7] H. Brézis and E. Lieb, A relation between pointwise convergence of functions and convergence of functionals, Proc. Amer. Math. Soc. 88 (1983), no. 3, 486–490.10.1007/978-3-642-55925-9_42Search in Google Scholar

[8] J. Q. Chen, Multiple positive solutions of a class of non autonomous Schrödinger-Poisson systems, Nonlinear Anal. Real World Appl. 21 (2015), 13–26.10.1016/j.nonrwa.2014.06.002Search in Google Scholar

[9] S. T. Chen and X. H. Tang, Ground state sign-changing solutions for asymptotically cubic or super-cubic Schrödinger-Poisson systems without compact condition, Comput. Math. Appl. 74 (2017), no. 3, 446–458.10.1016/j.camwa.2017.04.031Search in Google Scholar

[10] X.-P. Chen and C.-L. Tang, Least energy sign-changing solutions for Schrödinger-Poisson system with critical growth, Commun. Pure Appl. Anal. 20 (2021), no. 6, 2291–2312.10.3934/cpaa.2021077Search in Google Scholar

[11] X.-P. Chen and C.-L. Tang, Positive and sign-changing solutions for critical Schrödinger-Poisson systems with sign-changing potential, Qual. Theory. Dyn. Syst. 21 (2022), no. 3, Paper No. 89.10.1007/s12346-022-00628-4Search in Google Scholar

[12] M. F. Furtado, L. A. Maia and E. S. Medeiros, Positive and nodal solutions for a nonlinear Schrödinger equation with indefinite potential, Adv. Nonlinear Stud. 8 (2008), no. 2, 353–373.10.1515/ans-2008-0207Search in Google Scholar

[13] L. H. Gu, H. Jin, and J. J. Zhang, Sign-changing solutions for nonlinear Schrödinger-Poisson systems with subquadratic or quadratic growth at infinity, Nonlinear Anal. 198 (2020), 111897, 16 pp.10.1016/j.na.2020.111897Search in Google Scholar

[14] H. Hofer, Variational and topological methods in partially ordered Hilbert spaces, Math. Ann. 261 (1982), no. 4, 493–514.10.1007/BF01457453Search in Google Scholar

[15] G. B. Li and S. S. Yan, Eigenvalue problems for quasilinear elliptic equations on RN, Comm. Partial Differ. Equ. 14 (1989), no. 8–9, 1291–1314.10.1080/03605308908820654Search in Google Scholar

[16] G. B. Li, Some properties of weak solutions of nonlinear scalar field equations, Ann. Acad. Sci. Fenn. Math. 15 (1990), no. 1, 27–36.10.5186/aasfm.1990.1521Search in Google Scholar

[17] J. Liu, J.-F. Liao and C.-L. Tang, Ground state solution for a class of Schrödinger equations involving general critical growth term, Nonlinearity 30 (2017), no. 3, 899–911.10.1088/1361-6544/aa5659Search in Google Scholar

[18] Z. S. Liu and S. J. Guo, Existence of positive solutions for Kirchhoff-type problems, Nonlinear Anal. 120 (2015), 1–13.10.1016/j.na.2014.12.008Search in Google Scholar

[19] C. Miranda, Unaosservazione su un teorema di Brouwer, Boll. Un. Mat. Ital. 3 (1940), no. 25–7. (Italian)Search in Google Scholar

[20] J. J. Nie and Q. Q. Li, Multiplicity of sign-changing solutions for a supercritical nonlinear Schrödinger equation, Appl. Math. Lett. 109 (2020), 106569, 7pp.10.1016/j.aml.2020.106569Search in Google Scholar

[21] P. H. Rabinowitz, Variational methods for nonlinear eigenvalue problems, Eigenvalues of non-linear problems (Centro Internaz. Mat. Estivo (C.I.M.E.), III Ciclo, Varenna, 1974), Edizioni Cremonese, Rome, 1974, pp. 139–195.10.1007/978-3-642-10940-9_4Search in Google Scholar

[22] W. Shuai and Q. F. Wang, Existence and asymptotic behavior of sign-changing solutions for the nonlinear Schrödinger-Poisson system in R3, Z. Angew. Math. Phys. 66 (2015), no. 6, 3267–3282.10.1007/s00033-015-0571-5Search in Google Scholar

[23] J. J. Sun and S. W. Ma, Ground state solutions for some Schrödinger-Poisson systems with periodic potentials, J. Differ. Equ. 260 (2016), no. 3, 2119–2149.10.1016/j.jde.2015.09.057Search in Google Scholar

[24] K. M. Teng, Existence of ground state solutions for the nonlinear fractional Schrödinger-Poisson system with critical Sobolev exponent, J. Differ. Equ. 261 (2016), no. 6, 3061–3106.10.1016/j.jde.2016.05.022Search in Google Scholar

[25] K. M. Teng, Ground state solutions for the non-linear fractional Schrödinger-Poisson system, Appl. Anal. 98 (2019), no. 11, 1959–1996.10.1080/00036811.2018.1441998Search in Google Scholar

[26] D.-B. Wang, H.-B. Zhang, and W. Guan, Existence of least-energy sign-changing solutions for Schrödinger-Poisson system with critical growth, J. Math. Anal. Appl. 479 (2019), no. 2, 2284–2301.10.1016/j.jmaa.2019.07.052Search in Google Scholar

[27] Z. P. Wang and H.-S. Zhou, Sign-changing solutions for the nonlinear Schrödinger-Poisson system in R3, Calc. Var. Partial Differ. Equ. 52 (2015), no. 3–4, 927–943.10.1007/s00526-014-0738-5Search in Google Scholar

[28] M. Willem, Minimax Theorems, Birkhäuser, Boston, 1996.10.1007/978-1-4612-4146-1Search in Google Scholar

[29] X.-J. Zhong and C.-L. Tang, Ground state sign-changing solutions for a Schrödinger-Poisson system with a 3-linear growth nonlinearity, J. Math. Anal. Appl. 455 (2017), no. 2, 1956–1974.10.1016/j.jmaa.2017.04.010Search in Google Scholar

[30] X.-J. Zhong and C.-L. Tang, Ground state sign-changing solutions for a Schrödinger-Poisson system with a critical nonlinearity in R3, Nonlinear Anal. Real World Appl. 39 (2018), 166–184.10.1016/j.nonrwa.2017.06.014Search in Google Scholar

Received: 2021-12-05
Revised: 2022-07-08
Accepted: 2022-07-30
Published Online: 2022-08-25

© 2022 Xiao-Ping Chen and Chun-Lei Tang, published by De Gruyter

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

Articles in the same Issue

  1. Research Articles
  2. Editorial
  3. A sharp global estimate and an overdetermined problem for Monge-Ampère type equations
  4. Non-degeneracy of bubble solutions for higher order prescribed curvature problem
  5. On fractional logarithmic Schrödinger equations
  6. Large solutions of a class of degenerate equations associated with infinity Laplacian
  7. Chemotaxis-Stokes interaction with very weak diffusion enhancement: Blow-up exclusion via detection of absorption-induced entropy structures involving multiplicative couplings
  8. Asymptotic mean-value formulas for solutions of general second-order elliptic equations
  9. Weighted critical exponents of Sobolev-type embeddings for radial functions
  10. Existence and asymptotic behavior of solitary waves for a weakly coupled Schrödinger system
  11. On the Lq-reflector problem in ℝn with non-Euclidean norm
  12. Existence of normalized solutions for the coupled elliptic system with quadratic nonlinearity
  13. Normalized solutions for a class of scalar field equations involving mixed fractional Laplacians
  14. Multiplicity and concentration of semi-classical solutions to nonlinear Dirac-Klein-Gordon systems
  15. Multiple solutions to multi-critical Schrödinger equations
  16. Existence of solutions to contact mean-field games of first order
  17. The regularity of weak solutions for certain n-dimensional strongly coupled parabolic systems
  18. Uniform stabilization for a strongly coupled semilinear/linear system
  19. Existence of nontrivial solutions for critical Kirchhoff-Poisson systems in the Heisenberg group
  20. Existence of ground state solutions for critical fractional Choquard equations involving periodic magnetic field
  21. Least energy sign-changing solutions for Schrödinger-Poisson systems with potential well
  22. Lp Hardy's identities and inequalities for Dunkl operators
  23. Global well-posedness analysis for the nonlinear extensible beam equations in a class of modified Woinowsky-Krieger models
  24. Gradient estimate of the solutions to Hessian equations with oblique boundary value
  25. Sobolev-Gaffney type inequalities for differential forms on sub-Riemannian contact manifolds with bounded geometry
  26. A Liouville theorem for the Hénon-Lane-Emden system in four and five dimensions
  27. Regularity of degenerate k-Hessian equations on closed Hermitian manifolds
  28. Principal eigenvalue problem for infinity Laplacian in metric spaces
  29. Concentrations for nonlinear Schrödinger equations with magnetic potentials and constant electric potentials
  30. A general method to study the convergence of nonlinear operators in Orlicz spaces
  31. Existence of ground state solutions for critical quasilinear Schrödinger equations with steep potential well
  32. Global existence of the two-dimensional axisymmetric Euler equations for the Chaplygin gas with large angular velocities
  33. Existence of two solutions for singular Φ-Laplacian problems
  34. Existence and multiplicity results for first-order Stieltjes differential equations
  35. Concentration-compactness principle associated with Adams' inequality in Lorentz-Sobolev space
Downloaded on 11.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2022-0021/html
Scroll to top button