Home Existence of a Positive Solution to a Nonlinear Scalar Field Equation with Zero Mass at Infinity
Article Open Access

Existence of a Positive Solution to a Nonlinear Scalar Field Equation with Zero Mass at Infinity

  • Mónica Clapp ORCID logo EMAIL logo and Liliane A. Maia
Published/Copyright: January 10, 2018

Abstract

We establish the existence of a positive solution to the problem

- Δ u + V ( x ) u = f ( u ) , u D 1 , 2 ( N ) ,

for N3, when the nonlinearity f is subcritical at infinity and supercritical near the origin, and the potential V vanishes at infinity. Our result includes situations in which the problem does not have a ground state. Then, under a suitable decay assumption on the potential, we show that the problem has a positive bound state.

MSC 2010: 35Q55; 35B09; 35J20

1 Introduction

This paper is concerned with the existence of a positive solution to the problem

(1.1) - Δ u + V ( x ) u = f ( u ) , u D 1 , 2 ( N ) ,

for N3, where the nonlinearity f is subcritical at infinity and supercritical near the origin, and the potential V vanishes at infinity. Our precise assumptions on V and f are stated below.

In their groundbreaking paper [10], Berestycki and Lions considered the case where Vλ is constant and f has superlinear growth. They showed that if f is subcritical, then problem (1.1) has a solution for λ>0 and it does not have a solution for λ<0. They also studied the limiting case V0, which they called the zero mass case. They showed that if f is subcritical at infinity and supercritical near the origin, then the problem

(1.2) - Δ u = f ( u ) , u D 1 , 2 ( N ) ,

has a ground state solution ω, which is positive, radially symmetric and decreasing in the radial direction.

The motivation for studying this type of equations came from some problems in particle physics related to the nonabelian gauge theory, which underlies strong interaction, called quantum chromodynamics (QCD). Their solutions give rise to some special solutions of the pure Yang–Mills equations via ’t Hooft’s ansatz, see [17].

For a radial potential V(|x|), Badiale and Rolando [4] established the existence of a positive radial solution to problem (1.1). On the other hand, under suitable hypotheses, but without assuming any symmetries on V, Benci, Grisanti and Micheletti showed in [7] that problem (1.1) has a positive least energy solution if V(x)0 for all xN and V(x)<0 on a set of positive measure. They also showed that, if V(x)0 for all xN and V(x)>0 on a set of positive measure, then this problem does not have a ground state solution, i.e., the corresponding variational functional does not attain its (least energy) mountain pass value. Other related results may be found in [2, 8, 9, 16].

The result that we present in this paper includes the existence of a positive bound state for positive or sign changing potentials which decay to 0 at infinity with a suitable velocity. More precisely, we assume that V and f have the following properties:

  1. V L N / 2 ( N ) L r ( N ) for some r>N/2 and N|V-|N/2<SN/2, where V-:=min{0,V} and S is the best constant for the embedding D1,2(N)L2(N), with 2:=2NN-2.

  2. There exist constants A0>0 and κ>max{2,N-2} such that

    V ( x ) A 0 ( 1 + | x | ) - κ for all  x N .

  1. f 𝒞 1 [ 0 , ) and there exist constants A1>0 and 2<p<2<q such that, for m=-1,0,1,

    | f ( m ) ( s ) | { A 1 | s | p - ( m + 1 ) if  | s | 1 , A 1 | s | q - ( m + 1 ) if  | s | 1 ,

    where f(-1):=F, f(0):=f, f(1):=f and F(s):=0sf(t)dt.

  2. There exists a constant θ>2 such that 0θF(s)f(s)s<f(s)s2 for all s>0.

  3. The function g(s):=sf(s)f(s) is a decreasing function of s>0 and

    lim s g ( s ) < 2 - 1 < lim s 0 g ( s ) .

Our main result is the following one.

Theorem 1.1.

Assume that (V1)(V2) and (f1)(f3) hold true. Then problem (1.1) has a positive solution.

It is easy to see that the model nonlinearity

f ( s ) := s q - 1 1 + s q - p

satisfies assumptions (f1)(f3).

We point out that assumptions (V1), (f1) and (f2) are quite natural and have been also considered in previous works, in particular, in [7]. Assumption (f3) guarantees that the limit problem (1.2) has a unique positive solution. This fact, together with some fine estimates, which involve assumption (V2), allows us to show the existence of a positive bound state for problem (1.1) when the ground state is not attained.

The positive mass case, in which the potential V tends to a positive constant at infinity, has been widely investigated. A brief account may be found in [13], where a result similar to Theorem 1.1 was obtained for subcritical nonlinearities. On the other hand, except for the case of the critical pure power nonlinearity, only few results are known for the zero mass case.

There are several delicate issues in dealing with the zero mass case. Already the variational formulation requires some care, because the energy space D1,2(N) is only embedded in L2(N). The growth assumptions (f1) on the nonlinearity, however, provide the basic interpolation and boundedness conditions that allow to establish the differentiability of the variational functional and to study its compactness properties. Benci and Fortunato, in [6], expressed these conditions in the framework of Orlicz spaces, which was also used and further developed in [2, 4, 7, 8, 9, 3, 16]. The crucial facts for our purposes are stated in Proposition 3.1 below.

Another sensitive issue is the lack of compactness. In the positive mass case, a fundamental tool for dealing with it is Lions’ vanishing lemma, whose proof relies deeply on the fact that the sequences involved are bounded in H1(N). Once again, assumption (f1) allowed us to obtain a suitable version of this result for sequences which are only bounded in D1,2(N) (see Lemma 3.5). This new version of Lions’ vanishing lemma plays a crucial role in the proof of the splitting lemma (Lemma 3.9), which describes the lack of compactness of the variational functional. When the ground state is not attained, due to the uniqueness of the positive solution to the limit problem (1.2), the splitting lemma provides an open interval of values at which the energy functional satisfies the Palais–Smale condition.

We give a topological argument to establish the existence of a critical value in this interval. This argument requires, on the one hand, some fine estimates which are based on the precise asymptotic decay for the solutions of the limit problem (1.2), obtained recently by Vétois in [18], and on a suitable deformation lemma for 𝒞1-manifolds that was proved by Bonnet in [11].

This paper is organized as follows. In Section 2 we collect the information that we need about the solutions of the limit problem. Section 3 is devoted to the study of the variational problem and, especially, to the compactness properties of the variational functional. In Section 4 we derive the estimates that we need, and we prove our main result.

2 The Limit Problem

We define f(u):=-f(-u) for u<0. Then f𝒞1() and it is an odd function. Note that, if u is a positive solution of problem (1.1) for this new function, it is also a solution of (1.1) for the original function f. Hereafter, f will denote this extension.

We consider the Hilbert space D1,2(N):={uL2(N):uL2(N,N)} with its standard scalar product and norm

u , v := N u v ,   u := ( N | u | 2 ) 1 / 2 .

In this section we collect the information that we need on the positive solutions to the limit problem (1.2).

Since f𝒞1() and f satisfies (f1), a classical result of Berestycki and Lions establishes the existence of a ground state solution ω𝒞2(N) to problem (1.2), which is positive, radially symmetric and decreasing in the radial direction, see [10, Theorem 4].

Observe that assumption (f1) implies that |f(s)|A1|s|2-1 and |f(s)|A1|s|2-2. Note also that assumption (f2) yields f(s)>0 if s>0. Therefore, a recent result of Vétois implies that every positive solution u to (1.2) satisfies the decay estimates

(2.1) A 2 ( 1 + | x | ) - ( N - 2 ) | u ( x ) | A 3 ( 1 + | x | ) - ( N - 2 ) , | u ( x ) | A 3 ( 1 + | x | ) - ( N - 1 )

for some positive constants A2 and A3 and, moreover, u is radially symmetric and strictly radially decreasing about some point x0N, see [18, Theorem 1.1 and Corollary 1.2].

Concerning uniqueness, Erbe and Tang showed that if f also satisfies (f3), then problem (1.2) has a unique fast decaying radial solution, up to translations, where fast decaying means that u is positive and there exists a constant c(0,) such that lim|x||x|N-2u(|x|)=c, see [14, Theorem 2] and the remark in the paragraph following it.

We summarize these results in the following statement.

Proposition 2.1.

Under assumptions (f1)(f3), the limit problem (1.2) has a unique positive solution ω, up to translations. Moreover, ωC2(RN), it is radially symmetric and strictly decreasing in the radial direction, and it satisfies the decay estimates (2.1).

3 The Variational Setting

For u,vD1,2(N), we set

(3.1) u , v V := N u v + V ( x ) u v , u V 2 := N ( | u | 2 + V ( x ) u 2 ) .

By assumption (V1), these expressions are well defined and, using the Sobolev inequality, we conclude that V is a norm in D1,2(N) which is equivalent to the standard one.

Let 2<p<2<q. The following proposition, combined with assumption (f1), provides the interpolation and boundedness properties that are needed to obtain a good variational problem.

Proposition 3.1.

Let α,β>0 and hC0(R). Assume αβpq and βq, and that there exists M>0 such that

| h ( s ) | M min { | s | α , | s | β } for every  s .

Then, for every t[qβ,pα], the map D1,2(RN)Lt(RN) given by uh(u) is well defined, continuous and bounded.

Proof.

The decomposition u=u1Ωu+u1NΩu, where Ωu:={xN:|u(x)|>1}, gives a continuous embedding of L2(N), and hence of D1,2(N), into the Orlicz space

L p ( N ) + L q ( N ) := { u : u = u 1 + u 2 , with  u 1 L p ( N ) , u 2 L q ( N ) } ,

whose norm is defined by

| u | p , q := inf { | u 1 | p + | u 2 | q : u = u 1 + u 2 , u 1 L p ( N ) , u 2 L q ( N ) } .

Therefore, our claim is a special case of [3, Proposition 3.5]. ∎

Let F(u):=0uf(s)ds. Assumption (f1) implies that |F(s)|A1|s|2 and |f(s)|A1|s|2-1. Therefore, the functionals Φ,Ψ:D1,2(N), given by

Φ ( u ) := N F ( u ) , Ψ ( u ) := N f ( u ) u ,

are well defined. Using Proposition 3.1, it is easy to show that Φ is of class 𝒞2 and Ψ is of class 𝒞1, see [9, Lemma 2.6] or [3, Proposition 3.8]. Hence, the functional IV:D1,2(N), given by

I V ( u ) := 1 2 N ( | u | 2 + V ( x ) u 2 ) - N F ( u ) ,

is of class 𝒞2, with derivative

I V ( u ) v = N ( u v + V ( x ) u v ) - N f ( u ) v , u , v D 1 , 2 ( N ) ,

and the functional JV:D1,2(N), defined by

J V ( u ) := I V ( u ) u = N ( | u | 2 + V ( x ) u 2 ) - N f ( u ) u ,

is of class 𝒞1.

The solutions to problem (1.1) are the critical points of the functional IV. The nontrivial ones lie on the set

𝒩 V := { u D 1 , 2 ( N ) : u 0 , J V ( u ) = 0 } .

We define

c V := inf u 𝒩 V I V ( u )

and we write I0, J0, 𝒩0 and c0 for the previous expressions with V=0.

The proofs of the next two lemmas use well-known arguments. We include them for the sake of completeness. Hereafter, C will denote a positive constant, not necessarily the same one.

Lemma 3.2.

  1. There exists ϱ > 0 such that u V ϱ for every u 𝒩 V .

  2. 𝒩 V is a closed 𝒞 1 -submanifold of D 1 , 2 ( N ) and a natural constraint for the functional I V .

  3. c V > 0 .

  4. If u 𝒩 V , then the function t I V ( t u ) is strictly increasing in [ 0 , 1 ) and strictly decreasing in ( 1 , ) . In particular,

    I V ( u ) = max t > 0 I V ( t u ) .

Proof.

(a) Assumption (f1) implies that |f(s)s|A1|s|2. So, using Sobolev’s inequality, we get

J V ( u ) u V 2 - C N | u | 2 u V 2 - C u V 2 for all  u D 1 , 2 ( N ) .

As 2>2, there exists ϱ>0 such that JV(u)>0 if 0<uVϱ. This proves (a).

(b) It follows from (a) that 𝒩V is closed in D1,2(N). Moreover, assumption (f2) yields

J V ( u ) u = 2 u V 2 - N f ( u ) u 2 - N f ( u ) u = N [ f ( u ) - f ( u ) u ] u < 0

for every u𝒩V. This implies that 0 is a regular value of the restriction of JV to D1,2(N){0}, which is of class 𝒞1. Hence, 𝒩V is a 𝒞1-submanifold of D1,2(N) and a natural constraint for IV.

(c) Let u𝒩V. From hypothesis (f2) and statement (a), we obtain that

I V ( u ) = I V ( u ) - 1 θ I V ( u ) u = ( 1 2 - 1 θ ) u V 2 + N ( 1 θ f ( u ) u - F ( u ) ) ( 1 2 - 1 θ ) u V 2 ( 1 2 - 1 θ ) ϱ 2 .

Hence, cV>0.

(d) Let u𝒩V. Then

d d t I V ( t u ) = 1 t J V ( t u ) = t u V 2 - N f ( t u ) u = t N ( f ( u ) - f ( t u ) t ) u
= t [ u > 0 ( f ( u ) u - f ( t u ) t u ) u 2 + u < 0 ( f ( u ) u - f ( t u ) t u ) u 2 ] .

Property (f2) implies that f(s)s is strictly increasing for s>0 and strictly decreasing for s<0. Therefore, ddtIV(tu)>0 if t(0,1) and ddtIV(tu)<0 if t(1,). This proves (d). ∎

Lemma 3.3.

If u is a solution of (1.1), with IV(u)[cV,2cV), then u does not change sign.

Proof.

If u is a solution of (1.1), then 0=IV(u)u±=JV(u±), where u+:=max{u,0} and u-:=min{u,0}. Thus, if u+0 and u-0, then u±𝒩V and

I V ( u ) = I V ( u + ) + I V ( u - ) 2 c V ,

contradicting our assumption. ∎

Lemma 3.4.

The limit problem (1.2) does not have a solution u with I0(u)(c0,2c0).

Proof.

If u is a solution of (1.2) such that I0(u)[c0,2c0), then, by Lemma 3.3, u does not change sign. So, by Proposition 2.1, we have that u=±ω, up to a translation. Hence, I0(u)=c0. ∎

The following version of Lions’ vanishing lemma plays a crucial role in the proof of Lemma 3.6 and of the splitting lemma (Lemma 3.9). Its proof was inspired by that of [1, Lemma 2]. We write

B R ( y ) := { x N : | x - y | < R } .

Lemma 3.5.

If (uk) is bounded in D1,2(RN) and there exists R>0 such that

lim k ( sup y N B R ( y ) | u k | 2 ) = 0 ,

then limkRNf(uk)uk=0.

Proof.

Fix ε(0,1) and set η:=22>1. For each k, consider the function

w k := { | u k | if  | u k | ε , ε - ( η - 1 ) | u k | η if  | u k | ε .

Observe that

| w k | 2 = η 2 ε - 2 ( η - 1 ) | u k | 2 ( η - 1 ) | u k | 2 η 2 | u k | 2 if  | u k | ε

and

| w k | 2 ε - ( 2 - 2 ) | u k | 2 | u k | 2 if  | u k | ε ,
| w k | 2 | u k | 2 - 2 | u k | 2 ε - ( 2 - 2 ) | u k | 2 if  | u k | ε .

Using these inequalities, we obtain that

| w k | 2 | u k | 2 , | w k | 2 ε - ( 2 - 2 ) | u k | 2 , | w k | 2 η 2 | u k | 2 .

Therefore, as (uk) is bounded in D1,2(N), we have that

w k H 1 ( N ) 2 := N | w k | 2 + | w k | 2 N η 2 | u k | 2 + N ε - ( 2 - 2 ) | u k | 2 C ,

i.e., (wk) is bounded in H1(N). Moreover,

lim k ( sup y N B R ( y ) | w k | 2 ) = 0 .

From Lions’ vanishing lemma (namely, [19, Lemma 1.21]), it follows that

w k 0 in  L s ( N ) for each  2 < s < 2 .

Now, using (f1), we obtain

| N f ( u k ) u k | A 1 ( | u k | 1 | u k | p + | u k | 1 | u k | q )
A 1 ( | u k | ε | u k | p + ε | u k | 1 | u k | q + | u k | ε | u k | q )
2 A 1 | u k | ε | u k | p + A 1 | u k | ε | u k | q
2 A 1 | u k | ε | w k | p + A 1 | u k | ε | u k | q - 2 | u k | 2
2 A 1 N | w k | p + A 1 ε q - 2 N | u k | 2 .

As (uk) is bounded in D1,2(N) and wk0 in Lp(N), we conclude that

| N f ( u k ) u k | C ε q - 2 .

Since ε(0,1) was arbitrarily chosen, the statement is proved. ∎

We write IV(u) and JV(u) for the gradients of IV and JV at u with respect to the scalar product (3.1).

Lemma 3.6.

Let (uk) be a sequence in D1,2(RN) such that IV(uk)d>0 and JV(uk)0. Then there exist a1>a0>0 such that, after passing to a subsequence,

a 0 u k V a 1 , a 0 J V ( u k ) V a 1 , | J V ( u k ) u k | a 0 .

Proof.

From assumption (f2), we get that

d + o ( 1 ) = I V ( u k ) I V ( u k ) + N F ( u k ) = 1 2 u k V 2 ,

and

d + o ( 1 ) = I V ( u k ) - 1 θ J V ( u k ) = ( 1 2 - 1 θ ) u k V 2 + N [ 1 θ f ( u k ) u k - F ( u k ) ] ( 1 2 - 1 θ ) u k V 2 .

Hence, (uk) is bounded and bounded away from 0 in D1,2(N).

By assumption (f1), for any vD1,2(N), we have that

| N [ f ( u k ) u k + f ( u k ) ] v | C N | u k | 2 - 1 | v | C | u k | 2 2 - 1 | v | 2 C u k 2 - 1 v C v V .

Therefore,

| J V ( u k ) , v V | = | 2 u k , v V - N [ f ( u k ) u k + f ( u k ) ] v | C v V .

This implies that (JV(uk)) is bounded in D1,2(N). Hence, after passing to a subsequence, we have that |JV(uk)uk|a0. Next, we show that a>0.

As JV(uk)0, we have that

0 < a 0 2 u k V 2 = N f ( u k ) u k + o ( 1 ) .

So, by Lemma 3.5, there exist δ>0 and a sequence (yk) in N such that

(3.2) B 1 ( y k ) | u k | 2 = sup y N B 1 ( y ) | u k | 2 > δ .

Set u~k:=uk(+yk). Replacing (u~k) by a subsequence, we have that u~ku weakly in D1,2(N) and u~ku in Lloc2(N). Inequality (3.2) implies that u0. Hence, there exists a subset Λ of positive measure such that u(x)0 for every xΛ. Assumption (f2) implies that f(s)s2-f(s)s>0 if s0. So, using Fatou’s lemma, we conclude that

a = lim k | J V ( u k ) u k | = lim k | 2 u k V 2 - N ( f ( u k ) u k 2 + f ( u k ) u k ) |
= lim k N ( f ( u k ) u k 2 - f ( u k ) u k ) = lim k N ( f ( u ~ k ) u ~ k 2 - f ( u ~ k ) u ~ k )
lim inf k Λ ( f ( u ~ k ) u ~ k 2 - f ( u ~ k ) u ~ k ) Λ ( f ( u ) u 2 - f ( u ) u ) > 0 .

This proves that a>0 and hence that (JV(uk)uk) is bounded away from 0 in . It follows that

J V ( u k ) V | J V ( u k ) u k | u k V a ~ > 0 .

This completes the proof. ∎

For σ, we set σ:=JV-1(σ) if σ0, and 0:=𝒩V. If |σ| is small enough and uσ, we write σIV(u) for the orthogonal projection of IV(u) onto the tangent space σ at u.

Recall that a sequence (uk) in D1,2(N) is said to be a (PS)d-sequence forIV if IV(uk)d and IV(uk)0.

Lemma 3.7.

Let σkR and ukMσk be such that σk0, IV(uk)d>0 and MσkIV(uk)0. Then (uk) is a (PS)d-sequence for IV.

Proof.

Let tk be such that

(3.3) I V ( u k ) = σ k I V ( u k ) + t k J V ( u k ) .

Taking the scalar product with uk, we get that

σ k = J V ( u k ) = I V ( u k ) u k = σ k I V ( u k ) , u k V + t k J V ( u k ) u k .

By Lemma 3.6, we have that (uk) is bounded and (JV(uk)uk) is bounded away from 0. So, as σk0 and σkIV(uk)0, we conclude that tk0. Moreover, as (JV(uk)) is bounded in D1,2(N), from equation (3.3) we get that IV(uk)0, as claimed. ∎

Lemma 3.8.

If uku weakly in D1,2(RN), then the following statements hold true:

  1. u k V 2 = u k - u 2 + u V 2 + o ( 1 ) ,

  2. N | f ( u k ) - f ( u ) | | φ | = o ( 1 ) for every φ 𝒞 c ( N ) ,

  3. N F ( u k ) = N F ( u k - u ) + N F ( u ) + o ( 1 ) ,

  4. f ( u k ) - f ( u k - u ) f ( u ) in ( D 1 , 2 ( N ) ) .

Proof.

(a) Set vk:=uk-u. Assumption (V1) states that VLN/2(N)Lr(N), with r>N/2. As η:=2rr-1<2, we have that vk0 in Llocη(N). Given ε>0, we fix R>0 such that NBR(0)|V|N/2εN/2. Then

N | V | v k 2 = B R ( 0 ) | V | v k 2 + N B R ( 0 ) | V | v k 2 | V | r | v k | L η ( B R ( 0 ) ) 2 + | V | L N / 2 ( N B R ( 0 ) ) v k 2 C ε

for k large enough. It follows that

u k V 2 = u k - u V 2 + u V 2 + o ( 1 ) = u k - u 2 + u V 2 + o ( 1 ) .

(b) Let s,t. By the mean value theorem, there exists ζ(0,1) such that

| f ( s + t ) - f ( s ) | = | f ( s + ζ t ) | | t | A 1 min { | s + ζ t | p - 2 , | s + ζ t | q - 2 } | t |
A 1 min { ( | s | + | t | ) p - 2 , ( | s | + | t | ) q - 2 } | t |
(3.4) = h ( | s | + | t | ) | t | ,

where h(s):=A1min{|s|p-2,|s|q-2}. Applying Proposition 3.1 to this function, we get that {h(|u|+|uk-u|)} is bounded in Lp/(p-2)(N). So, as uku in Llocp(N), we conclude that

N | f ( u k ) - f ( u ) | | φ | N h ( | u | + | u k - u | ) | u k - u | | φ |
| h ( | u | + | u k - u | ) | p p - 2 | φ | p ( supp ( φ ) | u k - u | p ) 1 / p = o ( 1 ) .

(c) Arguing as in (b), we have that

| F ( s + t ) - F ( s ) | H ( | s | + | t | ) | t | for all  s , t ,

where H(s):=A1min{|s|p-1,|s|q-1}. Let ε>0 and set vk:=uk-u. Then, noting that |F(s)|A1|s|2 and using Proposition 3.1, we may choose R>1 such that

| x | > R | F ( u k ) - F ( v k ) - F ( u ) | | x | > R | F ( u k ) - F ( v k ) | + | x | > R | F ( u ) |
| x | > R H ( | v k | + | u | ) | u | + A 1 | x | > R | u | 2
| H ( | v k | + | u | ) | 2 / ( 2 - 1 ) ( | x | > R | u | 2 ) 1 / 2 + A 1 | x | > R | u | 2
C ( | x | > R | u | 2 ) 1 / 2 + A 1 | x | > R | u | 2 < ε .

On the other hand, as uku in Llocp(N), we have that

| x | R | F ( u k ) - F ( v k ) - F ( u ) | | x | R | F ( u k ) - F ( u ) | + 1 | x | R | F ( v k ) | + | x | 1 | F ( v k ) |
| x | R H ( | u k | + | u | ) | v k | + A 1 1 | x | R | v k | p + A 1 | x | 1 | v k | q
| H ( | u k | + | u | ) | p / ( p - 1 ) ( | x | R | v k | p ) 1 / p + C | x | R | v k | p
C ( | x | R | v k | p ) 1 / p + C | x | R | v k | p < ε

if k is large enough. This proves (c).

(d) Let φ𝒞c(N) and R>0. Set h(s):=A1min{|s|p-2,|s|q-2}. From (3.4) and Proposition 3.1, we get that

| x | > R | f ( u k ) - f ( u k - u ) | | φ | | x | > R h ( | u k | + | u | ) | u | | φ |
| h ( | u k | + | u | ) | 2 / ( 2 - 2 ) ( | x | > R | u | 2 ) 1 / 2 | φ | 2
C ( | x | > R | u | 2 ) 1 / 2 φ .

Moreover, as |f(u)|A1|u|2-1, we have that

| x | > R | f ( u ) | | φ | C ( | x | > R | u | 2 ) ( 2 - 1 ) / 2 φ .

Thus, given ε>0, we may choose R>0 large enough so that

| x | > R | f ( u k ) - f ( u k - u ) - f ( u ) | | φ | ε φ .

Next, we fix δ(0,1) such that η:=2δ(p,2) and ν:=ηη-1-δ[qq-2,pp-2]. As uku strongly in Llocη(N), from (3.4) and Proposition 3.1, we get that

| x | R | f ( u k ) - f ( u ) | | φ | | h ( | u | + | u k - u | ) | ν ( | x | R | u k - u | η ) 1 / η | φ | 2 ε φ

for k large enough and, similarly,

| x | R | f ( u k - u ) | | φ | | h ( | u k - u | ) | ν ( | x | R | u k - u | η ) 1 / η | φ | 2 ε φ .

Therefore,

| N ( f ( u k ) - f ( u k - u ) - f ( u ) ) φ | ε φ for  k  large enough.

This proves the claim. ∎

The following lemma is stated in [7] but its proof contains a gap. This can be fixed with the help of Lemma 3.5. We give the details.

Lemma 3.9 (Splitting lemma).

Let (uk) be a bounded (PS)d-sequence for IV. Then, after replacing (uk) by a subsequence, there exists a solution u of problem (1.1), a number mN{0}, m nontrivial solutions w1,,wm to the limit problem (1.2), and m sequences of points (yj,k)RN, 1jm, satisfying

  1. | y j , k | and | y j , k - y i , k | if i j ,

  2. u k - i = 1 m w i ( - y i , k ) u in D 1 , 2 ( N ) ,

  3. d = I V ( u ) + i = 1 m I 0 ( w i ) .

Proof.

Passing to a subsequence, we have that uku weakly in D1,2(N). It follows from Lemma 3.8 that

o ( 1 ) = I V ( u k ) φ = u k , φ V - N f ( u k ) φ = u , φ V - N f ( u ) φ + o ( 1 ) = I V ( u ) φ + o ( 1 )

for every φ𝒞c(N). Hence, u solves (1.1). Set u1,k:=uk-u. Then u1,k0 weakly in D1,2(N). Moreover, Lemma 3.8 implies that

(3.5) { I V ( u k ) = I 0 ( u 1 , k ) + I V ( u ) + o ( 1 ) , o ( 1 ) = I V ( u k ) = I 0 ( u 1 , k ) + o ( 1 ) in  ( D 1 , 2 ( N ) ) .

If u1,k0 strongly in D1,2(N), the proof is completed. So assume it does not. Then, as IV(u1,k)u1,k0, after passing to a subsequence, we have that

0 < C u 1 , k V 2 = N f ( u 1 , k ) u 1 , k + o ( 1 ) .

So, by Lemma 3.5, there exist δ>0 and a sequence (y1,k) in N such that

(3.6) B 1 ( y 1 , k ) | u 1 , k | 2 = sup y N B 1 ( y ) | u 1 , k | 2 > δ .

Set vk:=u1,k(+y1,k). Passing to a subsequence, we have that vkw1 weakly in D1,2(N) and vkw1 in Lloc2(N). Inequality (3.6) implies that w10 and, as u1,k0 weakly in D1,2(N), we conclude that |y1,k|. Next, we shall show that w1 is a solution to the limit problem (1.2). Let φ𝒞c(N) and set φk:=φ(-y1,k). Using Lemma 3.8 and performing a change of variable, we obtain that

I 0 ( w 1 ) φ + o ( 1 ) = I 0 ( v k ) φ = I 0 ( u 1 , k ) φ k = o ( 1 ) .

This proves that w1 solves problem (1.2). Moreover, Lemma 3.8 implies that

I 0 ( v k ) = I 0 ( v k - w 1 ) + I 0 ( w 1 ) + o ( 1 ) ,
o ( 1 ) = I V ( v k ) = I 0 ( v k - w 1 ) + o ( 1 ) in  ( D 1 , 2 ( N ) ) .

Set u2,k:=u1,k-w1(-y1,k)=uk-u-w1(-y1,k). Then u2,k0 weakly in D1,2(N) and, after a change of variable, from identities (3.5) and (3.7), we obtain that

(3.7) { I 0 ( u 2 , k ) = I 0 ( u 1 , k ) - I 0 ( w 1 ) + o ( 1 ) = I V ( u k ) - I V ( u ) - I 0 ( w 1 ) + o ( 1 ) , I 0 ( u 2 , k ) = I 0 ( u 1 , k ) + o ( 1 ) = I V ( u k ) + o ( 1 ) = o ( 1 ) in  ( D 1 , 2 ( N ) ) .

If u2,k0 strongly in D1,2(N), the proof is completed. If not, we repeat the argument. After a finite number of steps, we will arrive to a sequence (um+1,k) which converges strongly to 0 in D1,2(N). This completes the proof. ∎

Corollary 3.10 (Compactness).

If cV is not attained by IV on NV, then the following statements hold true:

  1. c V c 0 .

  2. If σ k and u k σ k are such that σ k 0 , IV(uk)d(c0,2c0) and σkIV(uk)0, then (uk) contains a convergent subsequence.

Proof.

(a) Let (uk) be a minimizing sequence for IV on 𝒩V. By Ekeland’s variational principle and Lemma 3.6, we may assume that (uk) is a bounded (PS)cV-sequence for IV. As cV is not attained, the splitting lemma implies that cVc0.

(b) By Lemmas 3.6 and 3.7, we have that (uk) is a bounded (PS)d-sequence for IV. Arguing by contradiction, assume that (uk) does not contain a convergent subsequence. Then, the splitting lemma yields a solution w of the limit problem (1.2) with d=I0(w), contradicting Lemma 3.4. This proves our claim. ∎

4 Existence of a Positive Solution

The proof of our main result requires some delicate estimates. The following lemma will help us obtain them.

Lemma 4.1.

  1. If y 0 , y N , y0y and α , and β are positive constants such that α+β>N, then there exists C1=C1(α,β,|y-y0|)>0 such that

    N d x ( 1 + | x - R y 0 | ) α ( 1 + | x - R y | ) β C 1 R - μ

    for all R 1 , where μ := min { α , β , α + β - N } .

  2. If y 0 , y N { 0 } , and κ and ϑ are positive constants such that κ+2ϑ>N, then there exists C2=C2(κ,ϑ,|y0|,|y|)>0 such that

    N d x ( 1 + | x | ) κ ( 1 + | x - R y 0 | ) ϑ ( 1 + | x - R y | ) ϑ C 2 R - τ

    for all R 1 , where τ := min { κ , 2 ϑ , κ + 2 ϑ - N } .

Proof.

(a) After a suitable translation, we may assume that y=-y0. Let 2ρ:=|y-y0|>0. In the following, C will denote different positive constants which depend on α, β and ρ. If |x-Ry0|ρR, then |x-Ry|ρR. Hence,

B ρ R ( R y 0 ) d x ( 1 + | x - R y 0 | ) α ( 1 + | x - R y | ) β = B ρ R ( R y 0 ) d x ( 1 + | x - R y 0 | ) α ( ρ R ) β
= C R - β B ρ R ( 0 ) d x ( 1 + | x | ) α C ( R - β + R N - ( α + β ) ) C R - μ .

Similarly,

B ρ R ( R y ) d x ( 1 + | x - R y 0 | ) α ( 1 + | x - R y | ) β C ( R - α + R N - ( α + β ) ) C R - μ .

Let

H + := { z N : | z - R y | | z - R y 0 | } and H - := { z N : | z - R y | | z - R y 0 | } .

Setting x=Rz, we obtain

H + B ρ R ( R y 0 ) d x ( 1 + | x - R y 0 | ) α ( 1 + | x - R y | ) β H + B ρ R ( R y 0 ) d x ( 1 + | x - R y 0 | ) α + β
H + B ρ ( y 0 ) R N d x ( R | z - y 0 | ) α + β = C R N - ( α + β ) C R - μ .

Similarly,

H - B ρ R ( R y ) d x ( 1 + | x - R y 0 | ) α ( 1 + | x - R y | ) β C R - μ .

Since N[BρR(Ry0)BρR(Ry)]=[H+BρR(Ry0)][H-BρR(Ry)], the previous estimates yield (a).

(b) From Hölder’s inequality and inequality (a), we obtain

N d x ( 1 + | x | ) κ ( 1 + | x - R y 0 | ) ϑ ( 1 + | x - R y | ) ϑ
( N d x ( 1 + | x | ) κ ( 1 + | x - R y 0 | ) 2 ϑ ) 1 / 2 ( N d x ( 1 + | x | ) κ ( 1 + | x - R y | ) 2 ϑ ) 1 / 2 C 2 R - τ ,

as claimed. ∎

Let ω be the positive radial ground state of the limit problem (1.2). Fix y0N, with |y0|=1, and let B2(y0):={xN:|x-y0|2}. For R1 and each yB2(y0), we define

ω 0 R := ω ( - R y 0 ) , ω y R := ω ( - R y ) ,

and we set

ε R := N f ( ω 0 R ) ω y R = N f ( ω ( x - R y 0 ) ) ω ( x - R y ) d x .

As before, C will denote a positive constant, not necessarily the same one.

Lemma 4.2.

There exists a constant C3>0 such that

ε R C 3 R - ( N - 2 )

for all yB2(y0) and all R1.

Proof.

By assumption (f1), we have that |f(s)|A1|s|2-1. On the other hand, from estimates (2.1) and Lemma 4.1 (a), we obtain

(4.1) N ( ω 0 R ) 2 - 1 ω y R C N d x ( 1 + | x - R y 0 | ) N + 2 ( 1 + | x - R y | ) N - 2 C R - ( N - 2 ) .

Therefore,

N f ( ω 0 R ) ω y R A 1 N ( ω 0 R ) 2 - 1 ω y R C R - ( N - 2 )

for all yB2(y0) and all R1, as claimed. ∎

Lemma 4.3.

There exists a constant C4>0 such that

N f ( s ω 0 R ) t ω y R C 4 R - ( N - 2 )

for all s,t12, yB2(y0) and R1.

Proof.

Note that if |x|<1, then, for every yB2(y0) and R1,

1 + | x - R ( y - y 0 ) | < 1 + | x | + R | y - y 0 | < 4 R .

Assumption (f3) implies that f(s)s is strictly increasing for s>0; hence, so is f. Performing a change of variable and using estimate (2.1), we obtain

N f ( s ω 0 R ) t ω y R t N f ( 1 2 ω 0 R ) ω y R 1 2 B 1 ( R y 0 ) f ( 1 2 ω 0 R ) ω y R
1 4 [ min x B 1 ( 0 ) f ( 1 2 ω ( x ) ) ] B 1 ( 0 ) ω ( x - R ( y - y 0 ) ) d x
C B 1 ( 0 ) ( 1 + | x - R ( y - y 0 ) | ) - ( N - 2 ) d x C R - ( N - 2 )

for all s,t12, yB2(y0) and R1, as claimed. ∎

Note that Lemmas 4.2 and 4.3 yield

(4.2) C 4 R - ( N - 2 ) ε R := N f ( ω 0 R ) ω y R C 3 R - ( N - 2 )

for all yB2(y0) and R1.

Lemma 4.4.

For each b>1, there exists a constant Cb>0 such that

| N ( s f ( ω 0 R ) - f ( s ω 0 R ) ) ω y R | C b | s - 1 | ε R

for all s[0,b], yB2(y0) and R1.

Proof.

Fix t and set g(s):=sf(t)-f(st). By the mean value theorem, there exists ζ between 1 and s such that

| s f ( t ) - f ( s t ) | = | g ( s ) - g ( 1 ) | = | f ( t ) - f ( ζ t ) t | | s - 1 |
( | f ( t ) | + | f ( ζ t ) t | ) | s - 1 |
A 1 ( | t | 2 - 1 + | ζ | 2 - 2 | t | 2 - 1 ) | s - 1 |
A 1 ( 1 + b 2 - 2 ) | t | 2 - 1 | s - 1 | for all  s [ 0 , b ] ,

where the second-to-last inequality follows from assumption (f1). So, from inequalities (4.1) and (4.2), we obtain that

N ( s f ( ω 0 R ) - f ( s ω 0 R ) ) ω y R C | s - 1 | N ( ω 0 R ) 2 - 1 ω y R C | s - 1 | R - ( N - 2 ) C | s - 1 | ε R

for all s[0,b], yB2(0) and R1, as claimed. ∎

Lemma 4.5.

There exists τ>N-2 such that

N V + ( ω 0 R + ω y R ) 2 C R - τ

for every yB2(y0) and R1.

Proof.

From assumption (V2), estimates (2.1) and Lemma 4.1 (b), we immediately obtain that

N V + ( ω 0 R + ω y R ) 2 = N V + ( ω 0 R ) 2 + 2 N V + ω 0 R ω y R + N V + ( ω y R ) 2 C R - τ ,

with τ:=min{κ,2(N-2),κ+N-4}>N-2. ∎

For each R1, yB2(y0) and λ[0,1], we define

z λ , y R := λ ω 0 R + ( 1 - λ ) ω y R .

Lemma 4.6.

For each R1, yB2(y0) and λ[0,1], there exists a unique Tλ,yR>0 such that

T λ , y R z λ , y R 𝒩 V .

Moreover, there exist R01 and T0>2 such that Tλ,yR[0,T0) for all RR0, yB2(y0) and λ[0,1], and Tλ,yR is a continuous function of the variables λ,y and R.

Proof.

The proof is the same as that of [13, Lemma 3.2], with obvious changes. ∎

Lemma 4.7.

For λ=12, we have that Tλ,yR2 as R uniformly in yB2(y0).

Proof.

Since ω solves (1.2), we have that

J 0 ( ω 0 R + ω y R ) = J 0 ( ω 0 R ) + J 0 ( ω y R ) + 2 N ω 0 R ω y R - N [ f ( ω 0 R + ω y R ) - f ( ω 0 R ) ] ω 0 R - N [ f ( ω 0 R + ω y R ) - f ( ω y R ) ] ω y R
= 2 N ω ω y - y 0 R - N [ f ( ω 0 R + ω y R ) - f ( ω 0 R ) ] ω 0 R - N [ f ( ω 0 R + ω y R ) - f ( ω y R ) ] ω y R .

Set h(s):=A1min{|s|p-2,|s|q-2}|s|2-2. Using inequality (3.4), we get that

N | f ( ω 0 R + ω y R ) - f ( ω 0 R ) | ω 0 R N h ( ω 0 R + ω y R ) ω y R ω 0 R
| ω 0 R + ω y R | 2 2 - 2 ( N ( ω y R ) 2 / 2 ( ω 0 R ) 2 / 2 ) 2 / 2
C ( N ( ω y - y 0 R ) 2 / 2 ω 2 / 2 ) 2 / 2

and, similarly,

N | f ( ω 0 R + ω y R ) - f ( ω y R ) | ω y R C ( N ω 2 / 2 ( ω y - y 0 R ) 2 / 2 ) 2 / 2 .

From the above inequalities, we conclude that J0(ω0R+ωyR)=oR(1), where oR(1)0 as R uniformly in yB2(y0). Hence, for λ=12, we get that

J V ( 2 z λ , y R ) = J V ( ω 0 R + ω y R ) = J 0 ( ω 0 R + ω y R ) + N V ( ω 0 R + ω y R ) 2 = o R ( 1 ) .

This yields the claim. ∎

The proof of the next result follows that of [15, Lemma 2.1].

Lemma 4.8.

For each a>0, there exists Ca0 such that

F ( s + t ) - F ( s ) - F ( t ) - f ( s ) t - f ( t ) s - C a ( s t ) 1 + ν / 2

for all s,t[0,a] and ν(0,q-2).

Proof.

The inequality is clearly satisfied if s=0 or t=0. Assumption (f3) implies that f is increasing for s>0. Therefore,

F ( s + t ) - F ( s ) = s s + t f ( ζ ) d ζ f ( s ) t

for all s,t>0. Moreover, by (f2), we have that f(s)=o(|s|1+ν) for any ν(0,q-2). Hence, there exists Ma>0 such that f(s)Mas1+ν for all s[0,a]. It follows that

F ( s + t ) - F ( s ) - F ( t ) - f ( s ) t - f ( t ) s - F ( t ) - f ( t ) s - M a 0 t ζ 1 + ν d ζ - M a s t 1 + ν = - M a ( t 2 + ν 2 + ν + s t 1 + ν )

for all s,t[0,a]. So, if ts, we get that

F ( s + t ) - F ( s ) - F ( t ) - f ( s ) t - f ( t ) s - 3 2 M a ( s t ) 1 + ν / 2 .

As this expression is symmetric in s and t, it holds true also when st, and the proof is complete. ∎

With the previous lemmas on hand, we now prove the following estimate.

Proposition 4.9.

There exists R11 and, for each R>R1, a number ηR>0 such that

I V ( T λ , y R z λ , y R ) 2 c 0 - η R

for all λ[0,1] and all yB2(y0).

Proof.

To simplify the notation, let us set

s := T λ , y R λ , t := T λ , y R ( 1 - λ ) , ω 0 := ω 0 R , ω y := ω y R .

Then s,t[0,T0) if RR0, with R01 and T0>2, as in Lemma 4.6, and

I V ( T λ , y R z λ , y R ) = I V ( s ω 0 + t ω y )
= s 2 2 N | ω 0 | 2 + s 2 2 N V ω 0 2 + t 2 2 N | ω y | 2 + t 2 2 N V ω y 2
+ s t N ω 0 ω y + s t N V ω 0 ω y - N F ( s ω 0 + t ω y )
= I 1 + I 2 + I 3 + I 4 + I 5 ,

where

I 1 = s 2 2 N | ω 0 | 2 - N F ( s ω 0 ) + t 2 2 N | ω y | 2 - N F ( t ω y ) ,
I 2 = s 2 2 N V ω 0 2 + t 2 2 N V ω y 2 + s t N V ω 0 ω y ,
I 3 = s t N ω 0 ω y ,
I 4 = - N [ F ( s ω 0 + t ω y ) - F ( s ω 0 ) - F ( t ω y ) - f ( s ω 0 ) t ω y - f ( t ω y ) s ω 0 ] ,
I 5 = - N f ( s ω 0 ) t ω y - N f ( t ω y ) s ω 0 .

Next, we estimate each Ii, with i=1,,5. As ω0 and ωy are ground states of the limit problem (1.2), Lemma 3.2 (d) yields

I 1 = I 0 ( s ω 0 ) + I 0 ( t ω y ) I 0 ( ω 0 ) + I 0 ( ω y ) = 2 c 0 .

From Lemma 4.5 and estimates (4.2), we get that

I 2 C R - τ = o ( ε R ) .

Lemma 4.8, with ν(2N-2,q-2), and Lemma 4.1 (a), with α=β=(1+ν2)(N-2), imply, for some μ>N-2, that

I 4 C | s t | 1 + ν 2 N ( ω 0 ω y ) 1 + ν 2 C R - μ = o ( ε R ) .

We write the sum of the remaining terms as

I 3 + I 5 = t 2 N [ s f ( ω 0 ) - f ( s ω 0 ) ] ω y + s 2 N [ t f ( ω y ) - f ( t ω y ) ] ω 0 - 1 2 N f ( s ω 0 ) t ω y - 1 2 N f ( t ω y ) s ω 0 .

By Lemma 4.4, there exists a constant C>0 such that

t 2 N | s f ( ω 0 ) - f ( s ω 0 ) | ω y + s 2 N | t f ( ω y ) - f ( t ω y ) | ω 0 C ( | s - 1 | + | t - 1 | ) ε R

for all s,t[0,T0], yB2(y0) and RR0. Moreover, Lemma 4.3 yields a constant C0>0 such that

1 2 N f ( s ω 0 ) t ω y + 1 2 N f ( t ω y ) s ω 0 C 0 ε R

for all s,t12, yB2(y0) and RR0. By Lemma 4.7, if λ=12, then s,t1 as R. Therefore, there exist R1R0 and δ(0,12) such that

I 3 + I 5 - C 0 2 ε R

for all λ[12-δ,12+δ], yB2(y0) and RR1. Summing up, we have shown that

(4.3) I V ( s ω 0 + t ω y ) 2 c 0 - C 0 2 ε R + o ( ε R )

for all λ[12-δ,12+δ], yB2(y0), RR1.

On the other hand, by Lemma 3.2 (d), there exists ϑ(0,c0) such that

I 1 = I 0 ( s ω 0 ) + I 0 ( t ω y ) 2 c 0 - ϑ

for all λ[0,12-δ][12+δ,1], yB2(y0) and R sufficiently large. Since I2++I5O(εR), we conclude that

(4.4) I V ( s ω 0 + t ω y ) 2 c 0 - 2 ϑ + O ( ε R )

for all λ[0,12-δ][12+δ,1], yB2(y0) and R sufficiently large.

Inequalities (4.3) and (4.4) yield the statement of the proposition. ∎

Lemma 4.10.

For any δ>0, there exists R2>0 such that

I V ( T λ , y R z λ , y R ) < c 0 + δ

for λ=1, every yB2(y0) and RR2. In particular, cVc0.

Proof.

By Lemma 4.6, Tλ,yR is bounded uniformly in λ,y and R. So, from Lemmas 3.2 (d) and 4.5, we obtain that

I V ( T 1 , y R z 1 , y R ) = I 0 ( T 1 , y R ω y R ) + ( T 1 , y R ) 2 N V ( ω y R ) 2 c 0 + o R ( 1 ) ,

where oR(1)0 as R uniformly in yB2(y0), and the claim is proved. ∎

For c, set

I V c := { u D 1 , 2 ( N ) : I V ( u ) c } .

Lemma 4.11 (Deformation).

If cV is not attained by IV on NV, then cV=c0. If, moreover, IV does not have a critical value in (c0,2c0), then, for any given δ,η(0,c04), there exists a continuous function

π : 𝒩 V I V 2 c 0 - η 𝒩 V I V c 0 + δ

such that π(u)=u for all uNVIVc0+δ.

Proof.

If cV is not attained, Corollary 3.10 (a) and Lemma 4.10 imply that cV=c0.

Recall that 𝒩V is a 𝒞1-manifold. From Lemma 3.6 and Corollary 3.10 (b), we have that the following statement is true: If σk and ukσk are such that σk0, IV(uk)d(c0,2c0) and either σkIV(uk)0 or JV(uk)0, then (uk) contains a convergent subsequence. (In fact, Lemma 3.6 says that (JV(uk)) must be bounded away from 0). This allows us to apply [11, Theorem 2.5] to conclude that there exists ε^>0 such that for each ε(0,ε^), there exists a homeomorphism ϕ:𝒩V𝒩V such that

  1. ϕ ( u ) = u if IV(u)[d-ε,d+ε],

  2. I V ( ϕ ( u ) ) I V ( u ) for all u𝒩V,

  3. I V ( ϕ ( u ) ) d - ε for all u𝒩V, with IV(u)d+ε.

Our claim follows easily from this fact. ∎

Let 𝔟:L2(N){0}N be a barycenter map, i.e., a continuous map such that

𝔟 ( u ( - y ) ) = 𝔟 ( u ) + y and 𝔟 ( u Θ - 1 ) = Θ ( 𝔟 ( u ) )

for all uL2(N){0} and yN, and every linear isometry Θ of N, see [5, 12]. Note that 𝔟(u)=0 if u is radial.

Lemma 4.12.

If cV is not attained by IV on NV, then there exists δ>0 such that

𝔟 ( u ) 0 for all  u 𝒩 V I V c 0 + δ .

Proof.

The proof is the same as that of [13, Lemma 3.11]. ∎

Proof of Theorem 1.1.

If cV is attained by IV at some u𝒩V, then u is a nontrivial solution of problem (1.1). So assume that cV is not attained. Then, by Lemma 4.11, cV=c0. We will show that IV has a critical value in (c0,2c0).

Lemma 4.12 allows us to choose δ(0,c04) such that

𝔟 ( u ) 0 for all  u 𝒩 V I V c 0 + δ

and, by Proposition 4.9 and Lemma 4.10, we may choose R1 and η(0,c04) such that

I V ( T λ , y R z λ , y R ) { 2 c 0 - η for all  λ [ 0 , 1 ]  and all  y B 2 ( y 0 ) , c 0 + δ for  λ = 1  and all  y B 2 ( y 0 ) .

Define ι:B2(y0)𝒩VIV2c0-η by

ι ( ( 1 - λ ) y 0 + λ y ) := T λ , y R z λ , y R , with  λ [ 0 , 1 ] , y B 2 ( y 0 ) .

Arguing by contradiction, assume that IV does not have a critical value in (c0,2c0). Then, by Lemma 4.11, there exists a continuous function

π : 𝒩 V I V 2 c 0 - η 𝒩 V I V c 0 + δ

such that π(u)=u for all u𝒩VIVc0+δ. The function ψ:B2(y0)B2(y0), given by

ψ ( x ) := 2 ( 𝔟 π ι ) ( x ) | ( 𝔟 π ι ) ( x ) | ,

is well defined and continuous, and ψ(y)=y for every yB2(y0). This is a contradiction. Therefore, IV must have a critical point u𝒩V, with IV(u)(c0,2c0).

By Lemma 3.3, u does not change sign and, since f is odd, -u is also a solution of (1.1). This proves that problem (1.1) has a positive solution. ∎


Communicated by Vieri Benci


Award Identifier / Grant number: 237661

Award Identifier / Grant number: IN104315

Award Identifier / Grant number: 308173/2014-7

Award Identifier / Grant number: 193.000.939/2015

Award Identifier / Grant number: 0193.001300/2016

Funding statement: The first author was supported by CONACYT grant 237661 (Mexico) and UNAM-DGAPA-PAPIIT grants IN104315 (Mexico). The second author was supported by FAPDF grants 193.000.939/2015 and 0193.001300/2016, CNPq/PQ 308173/2014-7 (Brazil) and PROEX/CAPES (Brazil).

References

[1] A. Azzollini, V. Benci, T. D’Aprile and D. Fortunato, Existence of static solutions of the semilinear Maxwell equations, Ric. Mat. 55 (2006), no. 2, 283–297. 10.1007/s11587-006-0016-8Search in Google Scholar

[2] A. Azzollini and A. Pomponio, Compactness results and applications to some “zero mass” elliptic problems, Nonlinear Anal. 69 (2008), no. 10, 3559–3576. 10.1016/j.na.2007.09.041Search in Google Scholar

[3] M. Badiale, L. Pisani and S. Rolando, Sum of weighted Lebesgue spaces and nonlinear elliptic equations, NoDEA Nonlinear Differential Equations Appl. 18 (2011), no. 4, 369–405. 10.1007/s00030-011-0100-ySearch in Google Scholar

[4] M. Badiale and S. Rolando, Elliptic problems with singular potential and double-power nonlinearity, Mediterr. J. Math. 2 (2005), no. 4, 417–436. 10.1007/s00009-005-0055-5Search in Google Scholar

[5] T. Bartsch and T. Weth, Three nodal solutions of singularly perturbed elliptic equations on domains without topology, Ann. Inst. H. Poincaré Anal. Non Linéaire 22 (2005), no. 3, 259–281. 10.1016/j.anihpc.2004.07.005Search in Google Scholar

[6] V. Benci and D. Fortunato, Towards a unified field theory for classical electrodynamics, Arch. Ration. Mech. Anal. 173 (2004), no. 3, 379–414. 10.1007/s00205-004-0324-7Search in Google Scholar

[7] V. Benci, C. R. Grisanti and A. M. Micheletti, Existence and non existence of the ground state solution for the nonlinear Schroedinger equations with V()=0, Topol. Methods Nonlinear Anal. 26 (2005), no. 2, 203–219. 10.12775/TMNA.2005.031Search in Google Scholar

[8] V. Benci, C. R. Grisanti and A. M. Micheletti, Existence of solutions for the nonlinear Schrödinger equation with V()=0, Contributions to Nonlinear Analysis, Progr. Nonlinear Differential Equations Appl. 66, Birkhäuser, Basel (2006), 53–65. 10.1007/3-7643-7401-2_4Search in Google Scholar

[9] V. Benci and A. M. Micheletti, Solutions in exterior domains of null mass nonlinear field equations, Adv. Nonlinear Stud. 6 (2006), no. 2, 171–198. 10.1515/ans-2006-0204Search in Google Scholar

[10] H. Berestycki and P.-L. Lions, Nonlinear scalar field equations. I. Existence of a ground state, Arch. Ration. Mech. Anal. 82 (1983), no. 4, 313–345. 10.1007/BF00250555Search in Google Scholar

[11] A. Bonnet, A deformation lemma on a 𝒞1 manifold, Manuscripta Math. 81 (1993), no. 3–4, 339–359. 10.1007/BF02567863Search in Google Scholar

[12] G. Cerami and D. Passaseo, The effect of concentrating potentials in some singularly perturbed problems, Calc. Var. Partial Differential Equations 17 (2003), no. 3, 257–281. 10.1007/s00526-002-0169-6Search in Google Scholar

[13] M. Clapp and L. A. Maia, A positive bound state for an asymptotically linear or superlinear Schrödinger equation, J. Differential Equations 260 (2016), no. 4, 3173–3192. 10.1016/j.jde.2015.09.059Search in Google Scholar

[14] L. Erbe and M. Tang, Structure of positive radial solutions of semilinear elliptic equations, J. Differential Equations 133 (1997), no. 2, 179–202. 10.1006/jdeq.1996.3194Search in Google Scholar

[15] G. Evéquoz and T. Weth, Entire solutions to nonlinear scalar field equations with indefinite linear part, Adv. Nonlinear Stud. 12 (2012), no. 2, 281–314. 10.1515/ans-2012-0206Search in Google Scholar

[16] M. Ghimenti and A. M. Micheletti, Existence of minimal nodal solutions for the nonlinear Schrödinger equations with V()=0, Adv. Differential Equations 11 (2006), no. 12, 1375–1396. 10.57262/ade/1355867589Search in Google Scholar

[17] B. Gidas, Euclidean Yang–Mills and related equations, Bifurcation Phenomena in Mathematical Physics and Related Topics (Cargèse 1979), NATO Adv. Study Inst. Ser. C Math. Phys. Sci. 54, Reidel, Boston (1980), 243–267. 10.1007/978-94-009-9004-3_15Search in Google Scholar

[18] J. Vétois, A priori estimates and application to the symmetry of solutions for critical p-Laplace equations, J. Differential Equations 260 (2016), no. 1, 149–161. 10.1016/j.jde.2015.08.041Search in Google Scholar

[19] M. Willem, Minimax Theorems, Progr. Nonlinear Differential Equations Appl. 24, Birkhäuser, Boston, 1996. 10.1007/978-1-4612-4146-1Search in Google Scholar

Received: 2017-12-01
Accepted: 2017-12-05
Published Online: 2018-01-10
Published in Print: 2018-11-01

© 2018 Walter de Gruyter GmbH, Berlin/Boston

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

Downloaded on 28.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2017-6044/html
Scroll to top button