Home Existence of Groundstates for a Class of Nonlinear Choquard Equations in the Plane
Article Open Access

Existence of Groundstates for a Class of Nonlinear Choquard Equations in the Plane

  • Luca Battaglia and Jean Van Schaftingen ORCID logo
Published/Copyright: January 11, 2017

Abstract

We prove the existence of a nontrivial groundstate solution for the class of nonlinear Choquard equations

- Δ u + u = ( I α * F ( u ) ) F ( u ) in  2 ,

where Iα is the Riesz potential of order α on the plane 2 under general nontriviality, growth and subcriticality on the nonlinearity FC1(,).

MSC 2010: 35J91; 35J20

1 Introduction

We are interested in the existence of nontrivial solutions to the class of nonlinear Choquard equations of the form

(P) - Δ u + u = ( I α * F ( u ) ) F ( u ) in  N ,

where N={1,2,}, Δ is the standard Laplacian operator on the Euclidean space N, Iα:N is the Riesz potential of order α(0,N) defined for each xN{0} by

I α ( x ) = Γ ( N - α 2 ) Γ ( α 2 ) π N 2 2 α | x | N - α ,

and a nonlinearity is described by the function FC1(,). Solutions of equation (P) are at least formally critical points of the energy functional defined for a function u:N by

(1.1) ( u ) = 1 2 N ( | u | 2 + | u | 2 ) - 1 2 N ( I α * F ( u ) ) F ( u ) .

In the particular case where F(s)=s2/2 for each s, solutions to the Choquard equation (P) are standing waves solutions of the Hartree equation. In particular, when N=3 and α=2, problem (P) has arisen in various fields of physics: quantum mechanics [20], one-component plasma [11] and self-gravitating matter [15]. In these cases, many existence results have been obtained in literature, with both variational [11, 13, 14] and ordinary differential equations techniques [21, 15, 6] (see also the review [18]). Such methods extend also to the case of homogeneous nonlinearities [16].

When the nonlinearity F is not any more homogeneous, it has been shown that the Choquard equation (P) has a nontrivial solution if the nonlinearity F satisfies the following hypotheses [17]:

  1. There exists s0 such that F(s0)0.

  2. There exists C>0 such that

    | F ( s ) | C ( | s | α N + | s | α + 2 N - 2 )

    for every s>0.

  3. There holds

    lim s 0 F ( s ) / | s | 1 + α N = 0 = lim s 0 F ( s ) / | s | N + α N - 2 .

The solution u is a groundstate, in the sense that u minimizes the value of the functional among all nontrivial solutions. Assumptions (F0), (F1) and (F2) are rather mild and reasonable and are “almost necessary” in the sense of Berestycki and Lions [4]: the nontriviality of the nonlinearity condition (F0) is clearly necessary to have a nontrivial solution; assumption (F1) secures a proper variational formulation of problem (P) by ensuring that the energy functional is well-defined on the natural Sobolev space H1(N) through the Hardy–Littlewood–Sobolev and Sobolev inequalities; condition (F2) is a sort of subcriticality condition with respect to the limiting-case embeddings. The analysis by a Pohožaev identity shows that assumptions (F1) and (F2) are necessary in the homogeneous case F(s)=sp/p (see [16]).

The results in [17] can thus be seen as a counterpart for Choquard-type equations of the result of Berestycki and Lions [4] which give similar “almost necessary” conditions for the existence of a groundstate to the equation

(1.2) - Δ u + u = G ( u ) in  N .

The latter equation can be at least formally obtained by (P) by passing to the limit as α0 and setting G=F2/2.

Whereas the above-mentioned almost necessary conditions for existence of the Choquard equation (P) and for the scalar field equation (1.2) have been obtained in higher dimensions N3, the latter result has been extended to the two-dimensional case [3] under the following assumptions:

  1. There exists s0 such that G(s0)>|s0|22.

  2. For every θ>0 there exists C=Cθ>0 such that |G(s)|Cθmin{1,s2}eθ|s|2 for every s>0.

  3. lim s 0 G ( s ) / | s | 2 < 1 2 .

This raises naturally the question whether there is a similar existence result for the Choquard equation (P) in the planar case.

In the present work, we provide a general existence result for groundstate solutions of problem (P) in the planar case N=2, which is a two-dimensional counterpart of [17] and a counterpart for the Choquard equation of [3]. The counterparts of (F0), (F1), (F2) we need are the following:

  1. There exists s0 such that F(s0)0.

  2. For every θ>0 there exists C=Cθ>0 such that |F(s)|Cθmin{1,|s|α2}eθ|s|2 for every s>0.

  3. lim s 0 F ( s ) / | s | 1 + α 2 = 0 .

Our main result reads as follows.

Theorem 1.1.

If N=2 and FC1(R,R) satisfies conditions (F0), (F1) and (F2), then problem (P) has a groundstate solution uH1(R2){0}, namely the function u solves (P) and

( u ) = c := inf { ( v ) v H 1 ( 2 ) { 0 } is a solution of (P) } .

Let us discuss the assumptions of Theorem 1.1. As above, assumption (F0) is necessary for the existence of a nontrivial solution. As before, condition (F1) ensures the needed well-definedness of the energy functional on the whole space H1(2). It has a different shape, because in dimension N=2 the critical nonlinearity for Sobolev embeddings is not anymore a power but rather an exponential-type nonlinearity. More precisely, the integral of min{1,u2}eθ|u|2 on 2 is uniformly controlled on H01(B1) if and only if θB1|u|24π (see [19, 1]); this is why the parameter θ>0 appears in condition (F1). It will appear that condition (F1) is strong enough at infinity. Indeed, by integrating the function F, it is possible to observe that

(1.3) lim | s | | F ( s ) | + | F ( s ) | | s | e θ | s | 2 = 0

for every θ>0. A subcriticality condition still needs to be imposed around 0; that is the goal of the subcriticality condition (F2).

Assumptions (F0), (F1) and (F2) are still almost necessary: in the case F(s)=spp, they are satisfied if and only if p>1+α2, and for p1+α2 the Choquard equation (P) has no nontrivial solutions (see [16]).

In order to prove Theorem 1.1 the constraint minimization technique used in [4, 3] for the local problem (1.2) does not seem to work, as it introduces a Lagrange multiplier that cannot be absorbed through a suitable dilation because of the presence of three different scalings in the equation and because of the nonhomogeneity of the nonlinearity.

Following [17], we use a mountain pass construction. We start by constructing a Palais–Smale sequence for the mountain pass level

(1.4) b := inf γ Γ sup t [ 0 , 1 ] ( γ ( t ) ) ,

where

Γ := { γ C ( [ 0 , 1 ] , H 1 ( 2 ) ) γ ( 0 ) = 0  and  ( γ ( 1 ) ) < 0 } .

To avoid relying on an Ambrosetti–Rabinowitz superlinearity condition, we use a scaling trick due to Jeanjean [9], which allows us to construct a Pohožaev–Palais–Smale sequence (Proposition 3.1), namely a Palais–Smale sequence which, in addition, satisfies asymptotically the Pohožaev identity

(1.5) 𝒫 ( u ) := 2 | u | 2 - ( 1 + α 2 ) 2 ( I α * F ( u ) ) F ( u ) = 0 .

Such a condition will imply quite directly the boundedness of the sequence in the space H1(2) and it will be crucial to get the convergence, hence the existence of a solution (Proposition 4.1).

We are left with showing that the solution u is actually a groundstate. To prove this, we first show that the solution u itself satisfies the Pohožaev identity (Proposition 5.2). This will follow by simple calculations once a suitable regularity result is established (Proposition 5.1); this regularity turns out to be easier to prove from assumption (F1) than in the higher-dimensional case [17] where a suitable nonlocal Brezis–Kato regularity had to be proved. The last ingredient that we need is an optimal path γvΓ associated to any solution v of (P). The construction of such paths (Proposition 5.3) is inspired by [10, 17] but it is more delicate in our two-dimensional case than in the higher dimensions N3, because dilations tv(/t)H1(N) are not anymore continuous at t=0 when N=2.

The content of the paper is the following: In Section 2 we provide some technical preliminaries. In Section 3 we construct the Pohožaev–Palais–Smale sequence. In Section 4 we show that the sequence converges to a solution of (P). In Section 5 we prove that u is actually a groundstate. In the last section we also state some qualitative result concerning the solutions, which can be proved directly following [17].

2 Preliminaries

In this section, we present some preliminary results which we will need throughout the rest of this paper. We start by reformulating in a more convenient form the Moser–Trudinger inequality of Adachi and Tanaka [1]. This quantitative estimate will play a crucial role throughout the paper.

Proposition 2.1 (Moser–Trudinger inequality).

For any β(0,4π) there exists C=Cβ>0 such that for every uH1(R2) satisfying

2 | u | 2 1

one has

2 min { 1 , | u | 2 } e β | u | 2 C β 2 | u | 2 .

Proof.

The result follows from the fact [1, Theorem 0.1] that under the conditions of the theorem one has

2 ( e β | u | 2 - 1 ) C 2 | u | 2 .

Together with the elementary inequalities valid for every s0, this yields

( 1 - 1 e ) max { 1 , s } e s e s - 1 max { 1 , s } e s ,

as desired. ∎

We will also use the Hardy–Littlewood–Sobolev inequality to deal with the nonlocal term (see for example [12, Theorem 4.3]).

Proposition 2.2 (Hardy–Littlewood–Sobolev inequality).

For any p[1,2α) and fLp(R2) there exists a constant C=Cα,p such that

I α f L 2 p 2 - α p ( 2 ) C f L p ( 2 ) .

Combining the last two results with the assumption on F and (1.3), we deduce that the energy functional is well-defined on H1(2).

Proposition 2.3.

If F satisfies (F1), then the energy functional I defined by (1.1) is well-defined and continuously differentiable.

Proof.

We first consider the superposition map defined for each uH1(2) and x2 by (u)(x)=F(u(x)). We claim that is well-defined and continuous as a map from H1(2) to L4/α(2). Indeed by assumption (F1), for every θ>0 and s we have

| F ( s ) | 4 α C θ 4 α min { 1 , s 2 } e 4 θ α | s | 2 .

If uH1(2), we take θ>0 such that 2|u|2<απ2θ. We observe that

| F ( u ) | 4 α C θ 4 α min { 1 , | u | 2 } e 4 θ α | u | 2

on 2, where the right-hand side is integrable in view of the Moser–Trudinger inequality (Proposition 2.1). Therefore, the map :H1(2)L4/α(2) is well-defined.

If now the sequence (un)n converges to u in H1(2), then we can assume without loss of generality that

ν := sup n 2 | u | 2 < α π 2 θ

and that (un)n converges to u almost everywhere. Then, for some constant C0, we have

C ( min { 1 , | u | 2 } e 4 θ α | u | 2 + min { 1 , | u n | 2 } e 4 θ α | u n | 2 ) - | F ( u ) - F ( u n ) | 4 α 0

for each n almost everywhere in 2. By Fatou’s lemma we get

lim inf n 2 C ( min { 1 , | u | 2 } e 4 θ α | u | 2 + min { 1 , | u n | 2 } e 4 θ α | u n | 2 ) - | F ( u ) - F ( u n ) | 4 α 2 C 2 min { 1 , | u | 2 } e 4 θ α | u | 2 ,

and therefore

lim sup n 2 | F ( u ) - F ( u n ) | 4 α C lim sup n 2 min { 1 , | u n | 2 } e 4 θ α | u n | 2 - min { 1 , | u | 2 } e 4 θ α | u | 2 .

If we consider the set Anλ={x2|un(x)|λ}, we have by Lebesgue’s dominated convergence theorem

lim sup n 2 A n λ min { 1 , | u n | 2 } e 4 θ α | u n | 2
lim sup n 2 A n λ ( min { 1 , | u n | 2 } e 4 θ α | u n | 2 - min { 1 , | u | 2 } e 4 θ α min ( | u | 2 , λ 2 ) ) + 2 min { 1 , | u | 2 } e 4 θ α | u | 2
2 min { 1 , | u | 2 } e 4 θ α | u | 2

for every λ>0. On the other hand, we have by the Cauchy–Schwarz inequality, the Chebyshev inequality and the Moser–Trudinger inequality (Proposition 2.1)

A n λ min { 1 , | u n | 2 } e 4 θ α | u n | 2 | A n λ | 1 2 ( A n λ min { 1 , | u n | 2 } e 8 θ α | u n | 2 ) 1 2 C λ 2 | u n | 2 .

This allows us to conclude that the map :H1(2)L4/α(2) is continuous.

We now consider the map :H1(2)L4/(2+α)(2) defined for each uH1(2) by (u)=Fu. We observe that for every s we have

F ( s ) = 0 1 F ( τ s ) s d τ ,

and thus for almost every x2 we have

F ( u ( x ) ) = 0 1 F ( τ u ( x ) ) u ( x ) d τ .

Thus it follows from the first part of the proof that is well-defined from H1(2) to L4/(2+α)(2).

For the differentiability we consider a sequence (un)n converging strongly to u in H1(2). We observe that

( u n ) - ( u ) - ( u ) ( u n - u ) = 0 1 ( ( ( 1 - τ ) u + τ ( u n ) ) - ( u ) ) ( u n - u ) d τ

for each n, and thus by Hölder’s inequality

( u n ) - ( u ) - ( u ) ( u n - u ) L 4 / ( 2 + α ) 0 1 ( ( 1 - τ ) u + τ ( u n ) ) - ( u ) L 4 / α u n - u L 2 .

By the convergence of the sequence (un)n and the continuity of the functional , it follows that

( u n ) - ( u ) - ( u n ) ( u n - u ) L 4 / ( 2 + α ) = o ( u n - u L 2 ) as  n ,

that is, represents the Fréchet differential of the functional . Since is continuous, it follows that is of class C1.

Finally, we consider the quadratic form 𝒬 defined for fL4/(2+α) by

𝒬 ( f ) = 2 ( I α f ) f .

By the Hardy–Littlewood–Sobolev inequality (Proposition 2.2), the quadratic form 𝒬 is bounded on bounded sets of the space L4/(2+α)(2). This implies that 𝒬 is continuously differentiable and thus the functional

u H 1 ( 2 ) 𝒬 ( ( u ) , ( u ) ) = 2 ( I α F ( u ) ) F ( u )

is continuously differentiable. By the smoothness of the norm on a Hilbert space, we conclude that the functional is continuously differentiable. ∎

Finally, we will use the following improvement of Proposition 2.2 when one has some more Lp integrability.

Proposition 2.4.

For any p[1,2α), q(2α,+) and fLp(R2)Lq(R2) there exists C=Cα,p,q such that

I α f L ( 2 ) C ( f L p ( 2 ) + f L q ( 2 ) ) .

Proof.

The result is classical. We give its short proof for the convenience of the reader. By choosing p, q in those ranges we have (2-α)qq-1<2<(2-α)pp-1. Therefore, through splitting the integral and by the Hölder inequality, we get

| I α * f ( x ) | C 2 | f ( x - y ) | | y | 2 - α 𝑑 y
C ( B 1 d y | y | ( 2 - α ) q q - 1 ) 1 - 1 q f L q ( B 1 ( x ) ) + C ( B 1 d y | y | ( 2 - α ) p p - 1 ) 1 - 1 p f L p ( 2 B 1 ( x ) )
C ( f L p ( 2 ) + f L q ( 2 ) )

for every x2. ∎

3 Construction of a Pohožaev–Palais–Smale Sequence

In this section, we show the existence of a Pohožaev–Palais–Smale sequence at the level b defined by (1.4). In other words, we construct a sequence of almost critical points which asymptotically satisfies equation (P) and the Pohožaev identity (1.5).

Proposition 3.1.

Suppose the function FC1(R,R) satisfies assumptions (F0) and (F1). Then there exists a sequence (un)nN in H1(R2) such that the following hold:

  1. ( u n ) b as n .

  2. ( u n ) 0 strongly in H 1 ( 2 ) as n .

  3. 𝒫 ( u n ) 0 as n .

To prove Proposition 3.1, we first need to show that the energy functional has the mountain pass geometry, namely that the mountain pass level b is well-defined and nontrivial.

Lemma 3.2.

The critical level b defined by (1.4) satisfies b(0,+).

Proof.

We start by showing the finiteness of b, which will be done as in [17, Proposition 2.1]. By the definition of the set b, it is sufficient to show that Γ, which in turn is equivalent to find u0H1(2) such that (u0)<0. By assumption (F0), we can take s0 such that F(s0)0 and we find

2 ( I α * F ( s 0 𝟏 B 1 ) ) F ( s 0 𝟏 B 1 ) = F ( s 0 ) 2 B 1 B 1 I α ( x - y ) 𝑑 x 𝑑 y > 0 .

Therefore, by the density of smooth functions in Lq(2) there will be v0H1(2) with

2 ( I α * F ( v 0 ) ) F ( v 0 ) > 0 .

We consider now, for t>0, the function vt:2 defined for x2 by vt(x):=v0(xt). This function verifies

( v t ) = 1 2 2 | v 0 | 2 + t 2 2 2 | v 0 | 2 - t 2 + α 2 2 ( I α * F ( v 0 ) ) F ( v 0 ) .

Therefore, for some t00, the function u0:=vt0 satisfies (u0)<0.

Let us now show that b>0. By the definition of b, this is equivalent to showing that there exists ε>0 such that for every path γΓ there exists tγ[0,1] with (γ(tγ))ε>0. We first assume that uH1(2) and 2(|u|2+|u|2)δ1. In particular, since 2|u|21, Proposition 2.1 applies to u with β=2π. Therefore, by Propositions 2.2 and 2.1 and by (1.3), we have

2 ( I α * F ( u ) ) F ( u ) C ( 2 | F ( u ) | 4 2 + α ) 1 + α 2
C ( 2 min { 1 , | u | 2 } e 2 π | u | 2 ) 1 + α 2
C ( 2 | u | 2 ) 1 + α 2 ,

which is smaller than 142(|u|2+|u|2) if δ is small enough. It follows then that if 2(|u|2+|u|2)δ, we have

( u ) 1 4 2 ( | u | 2 + | u | 2 ) .

We now take an arbitrary path γΓ. Since (γ(1))<0<142(|γ(tγ)|2+|γ(tγ)|2), we have

2 ( | γ ( 1 ) | 2 + | γ ( 1 ) | 2 ) > δ > 0 = 2 ( | γ ( 0 ) | 2 + | γ ( 0 ) | 2 ) .

Therefore, there exists tγ(0,1) such that 2(|γ(tγ)|2+|γ(tγ)|2)=δ, and hence (γ(tγ))δ4. The lemma follows by taking ε:=δ4. ∎

Proof of Proposition 3.1.

We follow [9, Chapter 2], [8, Chapter 4] and [17, Proposition 2.1]. We consider the map Φ given by

Φ : × H 1 ( 2 ) H 1 ( 2 ) , ( σ , v ) Φ ( σ , v ) ( x ) := v ( e - σ x )

and the functional ~=Φ, i.e.,

~ ( σ , v ) = ( Φ ( σ , u ) ) = 1 2 2 | v | 2 + e 2 σ 2 2 | v | 2 - e ( 2 + α ) σ 2 2 ( I α * F ( v ) ) F ( v ) ,

which is well-defined and Fréchet-differentiable on the Hilbert space ×H1(2).

We define now the class of paths

Γ ~ := { γ ~ C ( [ 0 , 1 ] , × H 1 ( 2 ) ) γ ~ ( 0 ) = ( 0 , 0 )  and  ~ ( γ ~ ( 1 ) ) < 0 } .

Since we have Γ={Φγ~γ~Γ~}, the mountain pass levels of and ~ coincide, namely

b = inf γ ~ Γ ~ sup t [ 0 , 1 ] ~ ( γ ~ ( t ) ) .

Since, by Lemma 3.2, the mountain pass level b is not trivial, we can thus apply the minimax principle (see [24, Theorem 2.9]) and we find a sequence ((σn,vn))n in ×H1(2) such that

~ ( σ n , v n ) n b and ~ ( σ n , v n ) n 0  strongly in  ( × H 1 ( 2 ) ) .

Writing explicitly the derivative of ~ as

~ ( σ n , v n ) [ h , w ] = ( Φ ( σ n , v n ) ) [ Φ ( σ n , w ) ] + 𝒫 ( Φ ( σ n , v n ) ) h ,

we see that the conclusion follows by taking un=Φ(σn,vn). ∎

4 Convergence of the Pohožaev–Palais–Smale Sequence

In this section, we will construct a nontrivial solution of (P) from the sequence given by Proposition 3.1.

Proposition 4.1.

Suppose that the function FC1(R,R) satisfies (F1) and (F2) and the sequence (un)nN in H1(R2) satisfies the following conditions:

  1. ( u n ) is uniformly bounded.

  2. ( u n ) 0 strongly in ( H 1 ( 2 ) ) as n .

  3. 𝒫 ( u n ) 0 as n .

Then, up to subsequences, one of the following occurs:

  1. either u n 0 strongly in H 1 ( 2 ) as n ,

  2. or there exist u H 1 ( 2 ) { 0 } solving ( (P) ) and a sequence ( x n ) n in 2 such that u n ( - x n ) u weakly in H 1 ( 2 ) as n .

We follow the strategy of [17, Proposition 2.2]. Since the gradient does not appear in the Pohožaev identity (1.5), it will be more delicate to show that the nonlocal term does not vanish.

Proof of Proposition 4.1.

We assume that the first alternative does not hold, namely

lim inf n 2 ( | u n | 2 + | u n | 2 ) > 0 .

Writing

1 2 2 | u n | 2 + α 2 ( 2 + α ) 2 | u n | 2 = ( u n ) - 𝒫 ( u n ) 2 + α

for each n, we deduce that the sequence (un)n is bounded in the space H1(2). Since (un)0 in H1(2) as n, we have (un)[un]0 as n, therefore

2 ( I α * F ( u n ) ) F ( u n ) u n = 2 ( | u n | 2 + | u n | 2 ) - ( u n ) [ u n ] 1 C .

Taking C0supn2(|un|2+|un|2), we can apply Proposition 2.1 to 1C0un with β=2π and we obtain

2 min { 1 , u n 2 } e 2 π C 0 | u n | 2 C 2 π 2 | u n | 2 C 0 C 2 π

for each n. Moreover, we also have

2 | u n | 2 = ( 1 + α 2 ) 2 ( I α * F ( u n ) ) F ( u n ) + 𝒫 ( u n )
= ( 1 + α 2 ) 2 ( I α * F ( u n ) ) F ( u n ) + o ( 1 )

as n. Therefore, from Proposition 2.2 and by (1.3) we get

1 C 2 ( I α * F ( u n ) ) F ( u n ) u n C ( 2 | F ( u n ) | 4 2 + α 2 ( | F ( u n ) | | u n | ) 4 2 + α ) 2 + α 4
C ( 2 min { 1 , | u n | 2 } e 2 π C 0 | u n | 2 ) 1 + α 2 C ′′ ( 2 | u n | 2 ) 1 + α 2
= C ′′ ( ( 1 + α 2 ) 2 ( I α * F ( u n ) ) F ( u n ) + o ( 1 ) ) 1 + α 2 ,

namely 2(Iα*F(un))F(un) is bounded from above from zero when n.

We now want to prove that un does not vanish. We will use the following inequality [13, Lemma I.1] (see also [24, Lemma 1.21], [16, Lemma 2.3] and [22, (2.4)]):

2 | u n | p C 2 ( | u n | 2 + | u n | 2 ) ( sup x 2 B 1 ( x ) | u n | p ) 1 - 2 p .

We will show that the term on the right-hand side is bounded from below by a positive constant for every p>2. By assumption (F2) and (1.3), for every ε>0 there exists Cε,θ>0 such that

| F ( s ) | 4 2 + α ε min { 1 , | s | 2 } e θ | s | 2 + C ε , θ | s | p .

Therefore,

( sup x 2 B 1 ( x ) | u n | p ) 1 - 2 p 1 C 2 | u n | p 2 ( | u n | 2 + | u n | 2 )
1 C C 0 C ε ( 2 | F ( u n ) | 4 2 + α - ε 2 min { 1 , | u n | 2 } e 2 π C 0 | u n | 2 )
1 C ε ( ( 2 ( I α * F ( u n ) ) F ( u n ) ) 2 2 + α - ε C 2 | u n | 2 )
1 C ε ( 1 C - ε C C 0 ) .

From the arbitrariness of the quantity ε, we get B1(xn)|un|p1C for some xn2, for n large enough.

We can now consider the translated sequence (un(-xn))n. Since problem (P) is invariant by translation, this sequence will satisfy the hypotheses of the present proposition, hence we will still denote it as (un)n and we will assume that xn=0 for all n. Since lim infnB1|un|p>0, we can assume that this sequence (un)n converges weakly to uH1(2){0}. We just have to show that u solves (P).

The sequence (un)n is bounded in H1(2). Thus, the sequence (F(un))n is bounded in Lp(2) for every p42+α. Moreover, up to subsequences, unu almost everywhere as n, so by the continuity of the function F we also have F(un)F(u) almost everywhere as n. This implies that F(un)F(u) weakly in Lp(2) for every such p as n. Since 2α>42+α, by Propositions 2.2 and 2.4 we get

I α * F ( u n ) I α * F ( u ) weakly in  L 4 / ( 2 - α ) ( 2 ) L ( 2 )  as  n .

By condition (F1) and Proposition 2.1, the sequence (F(un))n is bounded in Lp(2) for every p[2α,+), and by continuity F(un)F(u) almost everywhere as n. Therefore, F(un)F(u) strongly in Llocq(2) for every q[1,+) as n, hence

( I α * F ( u n ) ) F ( u n ) n ( I α * F ( u ) ) F ( u ) in  L loc r ( 2 )  for all  r [ 1 , + ) .

Therefore, for every φC01(2) we have

2 ( u φ + u φ ) = lim n 2 ( u n φ + u n φ )
= lim n 2 ( I α * F ( u n ) ) F ( u n ) φ
= 2 ( I α * F ( u ) ) F ( u ) φ ,

namely u solves the Choquard equation (P). ∎

Corollary 4.2.

If F satisfies conditions (F0), (F1) and (F2), then problem (P) has a nontrivial solution uH1(R2).

Proof.

By Proposition 3.1, the functional admits a Pohožaev–Palais–Smale sequence (un)n at the level b. We apply Proposition 4.1 to (un)n. If the first alternative occurred, then we would have (un)(0)=0 as n, in contradiction with Lemma 3.2. Therefore, the second alternative must occur, and in particular we get a solution uH1(2){0} of (P). ∎

5 From Solutions to Groundstates

We start by providing a local regularity result for solutions of (P). This result can be obtained quite directly because our growth assumption (F1) gives a good control on Iα*F(u), which, in turn, permits to apply a standard bootstrap method. The equivalent result in higher dimension N3 is more delicate to prove (see [17, Theorem 2]) because of the relative weakness of assumption (F1).

Proposition 5.1.

If FC1(R,R) satisfies condition (F1) and if the function uH1(R2) solves problem (P), then uWloc2,p(R2) for every p1.

Proof.

By (1.3) and Lemma 2.1 we deduce that if vH1(2), then F(v)Lp(2) for every p42+α. Since 2α>42+α, by the inequality from Proposition 2.4 we get Iα*F(v)L(2).

Therefore, any solution u of (P) verifies

| - Δ u + u | C | F ( u ) |

with F(u)Llocp(2) for every p1 because of (F1). By standard (interior) regularity theory on bounded domains (see for example [7, Chapter 9]) we deduce that uWloc2,p(2). ∎

The extra regularity just proved allows us to prove that solutions of (P) satisfy the Pohožaev identity (1.5). The proof of the Pohožaev identity is classical and it is based on testing (P) against a suitable cut-off of xu(x), therefore it will be skipped. Details can be found in [17, Theorem 3].

Proposition 5.2 (Pohožaev Identity).

If FC1(R,R) satisfies (F1) and uH1(R2)Wloc2,2(R2) solves (P), then

𝒫 ( u ) = 2 | u | 2 - ( 1 + α 2 ) 2 ( I α * F ( u ) ) F ( u ) = 0 .

The Pohožaev identity allows us to show that the mountain pass solution is actually a groundstate. We will argue like [10, Lemma 2.1] and [17, Proposition 2.1], associating to any solution v a path γvΓ passing through v. The main difficulty here is that the integral of |u|2 is invariant by dilation, therefore we are not allow to join v with 0 by just taking dilations tv(t). To overcome this difficulty, we will combine properly dilatations and multiplication by constants [10].

Proposition 5.3.

If FC1(R,R) satisfies (F1) and if vH1(R2){0} solves (P), then there exists a path γvC([0,1],H1(R2)) such that the following hold:

  1. γ v ( 0 ) = 0 ;

  2. γ v ( 1 2 ) = v ;

  3. ( γ v ( t ) ) < ( v ) for every t [ 0 , 1 ] { 1 2 } ;

  4. ( γ v ( 1 ) ) < 0 .

Proof.

We consider the path γ~:[0,+)H1(2) given for each τ[0,) by

( γ ~ ( τ ) ) ( x ) := { τ τ 0 v ( x τ 0 ) if  τ τ 0 , v ( x τ ) if  τ τ 0 ,

with τ01 to be chosen later. The function γ~ is clearly continuous on the interval [0,+) and in particular at its boundary 0. For ττ0 Proposition 5.2 gives

( γ ~ ( τ ) ) = 1 2 2 | v | 2 + τ 2 2 2 | v | 2 - τ 2 + α 2 2 ( I α * F ( v ) ) F ( v )
= 1 2 2 | v | 2 + ( τ 2 2 - τ 2 + α 2 + α ) 2 | v | 2 ,

which attains its strict maximum in τ=1 and is negative for ττ1, for some τ11. For ττ0 we use (1.3) with θ=(1+α2)π and then apply Proposition 2.1 to the function γ~(τ)/(2|γ~(τ)|2)1/2 to obtain

(5.1) 2 | F ( γ ~ ( τ ) ) | 4 2 + α C 2 min { 1 , | γ ~ ( τ ) | 2 } e 2 π | γ ~ ( τ ) | 2 C 2 | γ ~ ( τ ) | 2 2 | γ ~ ( τ ) | 2 = C τ 0 2 2 | v | 2 .

Therefore, because of the Pohožaev identity (Proposition 5.2) and the Hardy–Littlewood–Sobolev inequality (Proposition 2.2), we have

( γ ~ ( τ ) ) = τ 2 2 τ 0 2 2 | v | 2 + τ 2 2 2 | v | 2 - 2 ( I α * F ( γ ~ ( τ ) ) ) F ( γ ~ ( τ ) )
1 2 2 | v | 2 + τ 2 2 2 | v | 2 + C ( 2 | F ( γ ~ ( τ ) ) | 4 2 + α ) 1 + α 2 .

Therefore, in view of (5.1) and the Pohožaev identity again, we deduce that

( γ ~ ( τ ) ) 1 2 2 | v | 2 + τ 0 2 2 2 | v | 2 + C τ 0 2 + α ( 2 | v | 2 ) 1 + α 2
= ( v ) + ( τ 0 2 2 - α 2 ( 2 + α ) ) 2 | v | 2 + C τ 0 2 + α ( 2 | v | 2 ) 1 + α 2 ,

which is strictly less than (v) if τ0=τ0(v) is chosen small enough. Therefore, the function γ~ verifies the following properties:

  1. γ ~ ( 0 ) = 0 ;

  2. γ ~ ( 1 ) = v ;

  3. ( γ ~ ( τ ) ) < ( v ) for every t[0,τ1]{1};

  4. ( γ ~ ( τ 1 ) ) < 0 .

To get the required γv it suffices to take a suitable change of variable γv(t):=γ~(T(τ)) for some function TC([0,1],) satisfying T(0)=0, T(1)=12 and T(τ1)=1. ∎

We are now in a position to prove the main theorem of this work.

Proof of Theorem 1.1.

Let (un)n be the Pohožaev–Palais–Smale sequence given by Proposition 3.1. Then, by Proposition 4.1, it converges weakly to a solution uH1(2){0} of (P). By the definition of groundstates, (u)c and, by Proposition 5.2, we have 𝒫(u)=0 (Proposition 5.2 is applicable in view of Proposition 5.1). Arguing as in [17, Theorem 1], we get successively

( u ) = 1 2 2 | u | 2 + α 2 ( 2 + α ) 2 | u | 2
lim inf n ( 1 2 2 | u n | 2 + α 2 ( 2 + α ) 2 | u n | 2 )
= lim inf n ( ( u n ) - 𝒫 ( u n ) 2 + α ) = b .

If vH1(2){0} is another solution of the Choquard equation (P), we apply Proposition 5.3 to v and obtain

( v ) = sup t [ 0 , 1 ] ( γ v ( t ) ) inf γ Γ sup t [ 0 , 1 ] ( γ ( t ) ) = b .

By the definition of groundstates and the fact that the solution v is arbitrary, one has bc. Putting everything together, we get

c ( u ) b c ,

hence (u)=b=c. ∎

We point out, as a corollary of the proof of Theorem 1.1, that the convergence in Proposition 4.1 turns out to be actually a strong convergence in H1(2) and that this gives as a byproduct a compactness property of the set of groundstates of (P).

Corollary 5.4.

Let (un)nN be a Pohožaev–Palais–Smale sequence satisfying the assumptions of Proposition 4.1 and in addition

lim n ( u n ) = c .

Then there exist uH1(R2){0} solving (P) and a sequence (xn)nN in R2 such that, up to subsequences, un(-xn)nu strongly in H1(R2). Moreover, the set of groundstates

𝒮 c := { u H 1 ( 2 ) u solves (P) and  ( u ) = c }

is compact, up to translations, in H1(R2).

Proof.

We apply Proposition 4.1; the first alternative is excluded by our assumption and the continuity of the functional at 0. Therefore we get, up to translations, unu as n in H1(2) and the function uH1(2){0} solves (P). As in the proof of Theorem 1.1, we get

lim inf n ( 1 2 2 | u n | 2 + α 2 ( 2 + α ) 2 | u n | 2 ) c = 1 2 2 | u | 2 + α 2 ( 2 + α ) 2 | u | 2 ,

from which it follows that unu strongly in H1(2) as n.

To show the compactness of the set of groundstates 𝒮c, we consider an arbitrary sequence (un)n in 𝒮c. Because of Proposition 5.2, it verifies 𝒫(un)=0 for every n, so it satisfies the hypotheses of Proposition 4.1 and of the first part of the present corollary. Therefore, up to subsequences and translations, it will converge to some u which solves (P) and, by the continuity of the functional in H1(2), we get u𝒮c. ∎

We conclude this paper by the following result on additional qualitative properties of the solution u.

Proposition 5.5.

If F is even and nondecreasing on (0,) and u is a groundstate solution of (P), then u has constant sign and is radially symmetric with respect to some point aRN.

Proof.

The proof is the same as [17, Propositions 5.2 and 5.3]. We briefly sketch the argument for the convenience of the reader.

To prove the constant-sign property, consider the path γu defined in Proposition 5.3. Since F is an even function, (|v|)=(v) for every vH1(2), hence

( | γ u ( t ) | ) < ( | γ u ( 1 2 ) | ) = b

for every t[0,1]{12}. From this one easily deduces that the function |u| is a groundstate solution of (P). Since F0, we can apply the strong maximum principle and get |u|>0, namely u has constant sign. Without loss of generality we assume now that u0.

For the symmetry, we follow the strategy of Bartsch, Weth and Willem [2] and its adaptation to the Choquard equation [17, 16]. For any closed half space H2 we consider the reflection σH with respect to H and define, for every uH1(2), the polarization (see for example [5])

u H ( x ) := { max { u ( x ) , u ( σ H ( x ) ) } if  x H , min { u ( x ) , u ( σ H ( x ) ) } if  x H .

We first observe that (see [5, Lemma 5.3])

2 | u H | 2 + | u H | 2 = 2 | u | 2 + | u | 2 .

Moreover, since F is nondecreasing on (0,+), we have (Fu)H=F(uH) and thus, in view of the rearrangement inequality for the Riesz potential,

2 ( I α F ( u H ) ) F ( u H ) = 2 ( I α F ( u ) H ) F ( u ) H 2 ( I α F ( u ) ) F ( u )

with equality if and only if either (Fu)H=Fu or (Fu)H=FuσH (see [16, Lemma 5.3]). Thus, it follows that (uH)(u), with equality holding if and only if either F(uH)=F(u) or F(uH)=F(uσH) on 2. From this and the definition of the level b, it follows that uH is a groundstate solution of (P), hence either F(uH)=F(u) or F(uH)=F(uσH) on 2. In the former case, we easily get f(uH)=f(u), hence uH=u; in the latter, we similarly get uH=uσH. Since the hyperplane H is arbitrary, in either case we conclude that the function u is radially symmetric with respect to some point a2 (see [16, Lemma 5.4] and [23, Proposition 3.15]). ∎


Communicated by Zhi-Qiang Wang


Award Identifier / Grant number: T.1110.14

Funding statement: This work was supported by the Projet de Recherche (Fonds de la Recherche Scientifique–FNRS) T.1110.14 “Existence and asymptotic behavior of solutions to systems of semilinear elliptic partial differential equations”.

References

[1] S. Adachi and K. Tanaka, Trudinger type inequalities in N and their best exponents, Proc. Amer. Math. Soc. 128 (2000), no. 7, 2051–2057. 10.1090/S0002-9939-99-05180-1Search in Google Scholar

[2] T. Bartsch, T. Weth and M. Willem, Partial symmetry of least energy nodal solutions to some variational problems, J. Anal. Math. 96 (2005), 1–18. 10.1007/BF02787822Search in Google Scholar

[3] H. Berestycki, T. Gallouët and O. Kavian, Équations de champs scalaires euclidiens non linéaires dans le plan, C. R. Acad. Sci. Paris Sér. I Math. 297 (1983), no. 5, 307–310. Search in Google Scholar

[4] H. Berestycki and P.-L. Lions, Nonlinear scalar field equations. II: Existence of infinitely many solutions, Arch. Ration. Mech. Anal. 82 (1983), no. 4, 347–375. 10.1007/BF00250556Search in Google Scholar

[5] F. Brock and A. Y. Solynin, An approach to symmetrization via polarization, Trans. Amer. Math. Soc. 352 (2000), no. 4, 1759–1796. 10.1090/S0002-9947-99-02558-1Search in Google Scholar

[6] P. Choquard, J. Stubbe and M. Vuffray, Stationary solutions of the Schrödinger–Newton model. An ODE approach, Differential Integral Equations 21 (2008), no. 7–8, 665–679. 10.57262/die/1356038617Search in Google Scholar

[7] D. Gilbarg and N. S. Trudinger, Elliptic Partial Differential Equations of Second Order, 2nd ed., Grundlehren Math. Wiss. 224, Springer, Berlin, 1983. Search in Google Scholar

[8] J. Hirata, N. Ikoma and K. Tanaka, Nonlinear scalar field equations in N: Mountain pass and symmetric mountain pass approaches, Topol. Methods Nonlinear Anal. 35 (2010), no. 2, 253–276. Search in Google Scholar

[9] L. Jeanjean, Existence of solutions with prescribed norm for semilinear elliptic equations, Nonlinear Anal. 28 (1997), no. 10, 1633–1659. 10.1016/S0362-546X(96)00021-1Search in Google Scholar

[10] L. Jeanjean and K. Tanaka, A remark on least energy solutions in 𝐑N, Proc. Amer. Math. Soc. 131 (2003), no. 8, 2399–2408. 10.1090/S0002-9939-02-06821-1Search in Google Scholar

[11] E. H. Lieb, Existence and uniqueness of the minimizing solution of Choquard’s nonlinear equation, Stud. Appl. Math. 57 (1976/77), no. 2, 93–105. 10.1007/978-3-642-55925-9_37Search in Google Scholar

[12] E. H. Lieb and M. Loss, Analysis, 2nd ed., Grad. Stud. Math. 14, American Mathematical Society, Providence, 2001. Search in Google Scholar

[13] P.-L. Lions, The Choquard equation and related questions, Nonlinear Anal. 4 (1980), no. 6, 1063–1072. 10.1016/0362-546X(80)90016-4Search in Google Scholar

[14] G. P. Menzala, On the nonexistence of solutions for an elliptic problem inunbounded domains, Funkcial. Ekvac. 26 (1983), no. 3, 231–235. Search in Google Scholar

[15] I. M. Moroz, R. Penrose and P. Tod, Spherically-symmetric solutions of the Schrödinger–Newton equations, Classical Quantum Gravity 15 (1998), no. 9, 2733–2742. 10.1088/0264-9381/15/9/019Search in Google Scholar

[16] V. Moroz and J. Van Schaftingen, Groundstates of nonlinear Choquard equations: Existence, qualitative properties and decay asymptotics, J. Funct. Anal. 265 (2013), no. 2, 153–184. 10.1016/j.jfa.2013.04.007Search in Google Scholar

[17] V. Moroz and J. Van Schaftingen, Existence of groundstates for a class of nonlinear Choquardequations, Trans. Amer. Math. Soc. 367 (2015), no. 9, 6557–6579. 10.1090/S0002-9947-2014-06289-2Search in Google Scholar

[18] V. Moroz and J. Van Schaftingen, A guide to the Choquard equation, J. Fixed Point Theory Appl. (2016), 10.1007/s11784-016-0373-1. 10.1007/s11784-016-0373-1Search in Google Scholar

[19] J. Moser, A sharp form of an inequality by N. Trudinger, Indiana Univ. Math. J. 20 (1970/71), 1077–1092. 10.1512/iumj.1971.20.20101Search in Google Scholar

[20] S. I. Pekar, Untersuchungen über die Elektronentheorie der Kristalle, Akademie-Verlag, Berlin, 1954. 10.1515/9783112649305Search in Google Scholar

[21] P. Tod and I. M. Moroz, An analytical approach to the Schrödinger–Newton equations, Nonlinearity 12 (1999), no. 2, 201–216. 10.1088/0951-7715/12/2/002Search in Google Scholar

[22] J. Van Schaftingen, Interpolation inequalities between Sobolev and Morrey–Campanato spaces: A common gateway to concentration-compactness and Gagliardo–Nirenberg interpolation inequalities, Port. Math. 71 (2014), no. 3–4, 159–175. 10.4171/PM/1947Search in Google Scholar

[23] J. Van Schaftingen and M. Willem, Symmetry of solutions of semilinear elliptic problems, J. Eur. Math. Soc. (JEMS) 10 (2008), no. 2, 439–456. 10.4171/JEMS/117Search in Google Scholar

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

Received: 2016-04-12
Revised: 2016-12-07
Accepted: 2016-12-08
Published Online: 2017-01-11
Published in Print: 2017-07-01

© 2017 Walter de Gruyter GmbH, Berlin/Boston

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

Downloaded on 24.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ans-2016-0038/html
Scroll to top button