Home Mathematics On a critical Choquard-Kirchhoff p-sub-Laplacian equation in ℍn
Article Open Access

On a critical Choquard-Kirchhoff p-sub-Laplacian equation in ℍn

  • Sihua Liang EMAIL logo , Patrizia Pucci , Yueqiang Song and Xueqi Sun
Published/Copyright: August 29, 2024

Abstract

This article is devoted to the study of a critical Choquard-Kirchhoff p -sub-Laplacian equation on the entire Heisenberg group H n , where the Kirchhoff function K can be zero at zero, i.e., the equation can be degenerate, and involving a nonlinearity, which is critical in the sense of the Hardy-Littlewood-Sobolev inequality. We first establish the concentration-compactness principle for the p -sub-Laplacian Choquard equation on the Heisenberg group, and we then prove existence results.

MSC 2010: 35J20; 35R03; 46E35

Dedicated to Professor Ermanno Lanconelli on the occasion of his 80th birthday, with high feelings of admiration for his notable contributions in Mathematics and great affection.

1 Introduction and main results

Recently, critical equations on the Heisenberg group H n have been widely studied because of their important applications in quantum mechanics, number theory, partial differential equations, and other fields. It is worth noting that the Heisenberg group is topologically Euclidean, but analytically non-Euclidean. This is why, some basic properties, such as dilatations, need to be developed again [33]. Folland and Stein [8] were the first mathematicians who worked on the subelliptic analysis of the Heisenberg group. As in [7], they consistently were devoted to creating a generalization of the analysis for more general stratified groups. In the celebrated book [9], Rothschild and Stein extended these results for general vector fields satisfying the Hormander’s conditions. The book also laid the foundation of anisotropic analysis.

In this article, we consider the critical Choquard-Kirchhoff problem involving p -sub-Laplacian

(ℰ) K ( u p , V p + D u H p p ) ( Δ H , p u + V ( ξ ) u p 2 u ) = H n u ( η ) p λ η 1 ξ H λ d η u p λ 2 u + μ f ( ξ , u ) , 0 < λ < Q , 1 < p < Q , p λ = p ( 2 Q λ ) 2 ( Q p ) , 2 p λ > p * = p Q Q p ,

where Δ H , p div H ( D H u H p 2 D H u ) is the p -sub-Laplacian on the Heisenberg group H n , and the number p λ is the critical exponent in the sense of the Hardy-Littlewood-Sobolev inequality. The Kirchhoff function K and the potential function V satisfy the suitable natural assumptions, which are listed in the following.

The reason of studying equation ( ) is based on two main reasons. On the one hand, equation ( ) has several important applications. Indeed, it is well-known that the Heisenberg group appears in various areas of physics and science, such as quantum theory (uncertainty principle, commutation relations). On the other hand, in the development of the theory of partial differential equations in the Heisenberg group and, more generally, in sub-Riemannian manifolds, it is important to explore new challenging problems as well as to figure out, in this emerging variety of results, whether or not the standard methods developed in the Euclidean spaces can be adapted to this new context.

In [34], existence of solutions for critical p -sub-Laplacian equations in the entire Heisenberg group is proved via the concentration-compactness principle. In [35], the results of [34] are completed and existence of entire solutions of critical ( p , q ) equations with Hardy terms on the Heisenberg group is proved. Sun et al. [44] considered existence of infinitely nontrivial solutions for critical Kirchhoff p -sub-Laplacian equations on the Heisenberg group via the concentration-compactness principle and the mountain pass theorem. Liu et al. [24] dealt with existence and multiplicity of nontrivial solutions for a new nonlocal Schrödinger-Poisson system with the superlinear and suberlinear terms on the first Heisenberg group H 1 . For other results, we refer to [14,17,25,31,32,42].

Turning back to the critical Choquard-Kirchhoff problem on the Heisenberg group, we can see that the results are relatively scarce. Via the mountain pass theorem, the linking theorem and other methods, Goel and Sreenadh [12] dealt with the Brézis-Nirenberg-type result of the Dirichlet problem with Choquard-type nonlinearity. With the application of the limit index theory and the concentration-compactness principle for the Choquard equation, Sun et al. [45] considered the following noncooperative Choquard-Kirchhoff system:

K ( D H u 2 2 ) Δ H u = Ω u ( η ) 2 λ η 1 ξ H λ d η u 2 λ 2 u + F u ( ξ , u , v ) , in Ω , K ( D H v 2 2 ) Δ H v = Ω v ( η ) 2 λ η 1 ξ H λ d η v 2 λ 2 v + F v ( ξ , u , v ) , in Ω , u = 0 , v = 0 , in H n \ Ω ,

where Ω is a smooth bounded domain in H n , and 2 λ = ( 2 Q λ ) ( Q 2 ) is the Hardy-Littlewood-Sobolev critical exponent on the Heisenberg group, when p = 2 .

The inspiration of this article comes from some recent results. In the Euclidean setting, what is well known is that the Choquard equation

(1.1) Δ u + V ( x ) u = ( K μ u 2 ) u + λ f ( x , u ) , in R n ,

was first studied in the pioneering work of Pekar [28] for the modeling of quantum mechanics. It was subsequently considered to be an approximation of the Hartree-Fock theory [4]. In 1998, Penrose [29] identified it as a self-gravitational collapse model of a quantum mechanical wave function. The mathematicians who first investigated the existence and symmetry of the solution to equation (1.1) were Lieb [20] and Lions [21]. Since then, equations of type (1.1) have been widely treated (see, e.g., [1,27,50]). In particular, the study of critical Choquard-Kirchhoff-type problems has been widely studied. In [39], the following Schrödinger-Choquard-Kirchhoff-type fractional p -Laplacian equation

( a + b u s , p p ( θ 1 ) ) [ ( Δ u ) p s u + V ( x ) u p 2 u ] = λ f ( x , u ) + R n u p μ , s x y μ d y u p μ , s 2 u , in R n ,

where ( Δ ) p s is the fraction p -Laplacian, and p μ , s = ( p n p μ 2 ) ( n p s ) is the critical exponent in the sense of the Hardy-Littlewood-Sobolev inequality. With the help of the mountain pass theorem and the Ekeland variational principle, existence of nonnegative solutions to the aforementioned equation is obtained in both the sublinear and superlinear cases. Later, with the help of the concentration-compactness principle in the fractional setting and the Kajikiya new version of the symmetric mountain pass lemma, Wang et al. [47] proved existence of infinitely many solutions for the fractional p -Laplacian equations. Furthermore, there are many contributions in the interesting degenerate Kirchhoff-type problems. Using the Kajikiya new version of the symmetric mountain pass lemma, Liang et al. [19] dealt with existence and multiplicity of solutions of a class of degenerate Kirchhoff-type Schrödinger-Choquard equations involving the fractional p -Laplacian. Then, Song and Shi [41] showed existence of infinitely many solutions for a class of degenerate p -fractional Kirchhoff equations, involving critical Hardy-Sobolev nonlinearities. For recent results on degenerate Kirchhoff stationary problems, we refer to [6,38,49] and references therein.

For critical Choquard equations in unbounded domains, there is no doubt that the concentration-compactness principle for the Choquard equation is the most powerful tool to overcome the difficulties caused by the lack of compactness. Consequently, the study of the critical Choquard equation has a profound connection with the concentration phenomenon, which takes place when one looks for sequences of approximate solutions. Ivanov and Vassilev [15] extended the concentration-compactness principle á la Lions from the Euclidean setting to the general Carnot groups, and so to the Heisenberg group, which is the special interesting subcase of Carnot groups. In [33], the concentration-compactness principle has been given for systems with critical nonlinearities and Hardy terms in the Heisenberg group. Moreover, the article [36] dealt with the concentration-compactness principle involving Hardy terms on the classical and fractional spaces in the Heisenberg group. Sun et al. [43] considered the concentration-compactness principle for the Choquard equation on the Heisenberg group. This article presents a generalization of the results of [43]. To our best knowledge, there do not seem to be concentration-compactness principle theorems for the critical Choquard equation involving p -Laplacian on the Heisenberg group. Moreover, it also seems that there are no applications of these results on the Heisenberg group setting.

Inspired by the aforementioned studies, the main purpose in this article is to prove existence results for equation ( ). We first establish a version of the concentration-compactness principle for the Choquard equation involving p -Laplacian on the Heisenberg group. Then, we use this result to show that the ( P S ) c condition at special levels c is true. Finally, using the mountain pass theorem, we obtain existence and multiplicity of solutions for equation ( ) in both the nondegenerate and degenerate cases on the entire Heisenberg group.

Let us now list the key natural assumptions we require in the main results. The potential function V and the Kirchhoff function K : R 0 + R 0 + are supposed to satisfy

  1. V : H n R + is a continuous non-decreasing function, and there exists V 0 > 0 such that V V 0 > 0 in H n .

  2. There exists k 0 > 0 such that inf t 0 K ( t ) = k 0 .

  3. The Kirchhoff function K is a continuous in R 0 + , and there exists θ [ 1 , p p ) such that

    K ( t ) t θ K ( t ) , for all t R 0 + ,

    where K ( t ) = 0 t K ( s ) d s and p * = p Q ( Q p ) .

  4. The Kirchhoff function K is also non-decreasing in R 0 + , and there exists k 1 > 0 such that

    K ( t ) k 1 t θ 1 ,

    for all t R + and K ( 0 ) = 0 , where θ [ 1 , p p ) is the number given in ( K 2 ) .

A typical example of K is given by K ( t ) = a + b t θ 1 for t R 0 + , where a R 0 + , b R 0 + and a + b > 0 . Equation ( ) is called non-degenerate when a > 0 , b 0 , while equation ( ) is called degenerate when a = 0 , b > 0 . Similarly, in the general case, equation ( ) is said to be non-degenerate when inf t 0 K ( t ) > 0 , i.e., when ( K 1 ) holds, while it is called degenerate when K ( 0 ) = 0 and K ( t ) > 0 for t > 0 .

The function f satisfies the following hypotheses:

  1. f : H n × R R is a Carathéodory function such that f is odd with respect to the second variable.

  2. There exists an exponent q , with p θ < q < r < p , such that

    f ( ξ , t ) a ( ξ ) t q 1 , a.e. ξ H n and all t R ,

    where 0 a L ϑ ( H n ) L ( H n ) and ϑ = r ( r q ) .

  3. There exists σ ( p θ , p ) such that

    0 < σ F ( ξ , t ) f ( ξ , t ) t , for all t R + ,

    where F ( ξ , t ) = 0 t f ( ξ , s ) d s .

The natural solution space for equation ( ) is denoted by S V 1 , p ( H n ) and is properly defined in Section 2 together with the necessary preliminaries. Let us first state the existence results for equation ( ) in the non-degenerate case.

Theorem 1.1

(Existence in the non-degenerate case) Let ( V ) , ( K 1 ) ( K 2 ) , and ( f 1 ) ( f 3 ) hold. Then, there exists μ 1 > 0 such that equation ( ) processes a nontrivial solution in S V 1 , p ( H n ) for any μ μ 1 .

Theorem 1.2

(Multiplicity in the non-degenerate case) Under the assumptions of Theorem 1.1, there exists a constant μ 2 > 0 such that equation ( ) has at least k pairs of nontrivial solutions in S V 1 , p ( H n ) for all μ μ 2 .

Finally, the next existence results for equation ( ) are also obtained in the degenerate case.

Theorem 1.3

(Existence in the degenerate case) Let ( V ) , ( K 2 ) ( K 3 ) , and ( f 1 ) ( f 3 ) be satisfied. Then, there exists μ 3 > 0 such that equation ( ) has a nontrivial solution in S V 1 , p ( H n ) for any μ μ 3 .

Theorem 1.4

(Multiplicity in the degenerate case) Under the assumptions of Theorem 1.3, there exists a constant μ 4 > 0 such that equation ( ) has at least k pairs of nontrivial solutions in S V 1 , p ( H n ) for all μ μ 4 .

The table of contents of this article is as follows. In Section 2, we present some preliminary results related to the Heisenberg group. Section 3 is devoted to the proof of the concentration-compactness principle for the p -Laplacian Choquard equation on the Heisenberg group, which is useful to prove the Palais-Smale condition at some special energy levels. Section 4 is dedicated to the proofs of the existence and multiplicity Theorems 1.1 and 1.2 for equation ( ) in the non-degenerate case. Finally, Section 5 deals with the proof of Theorems 1.3 and 1.4, i.e., to the proofs of existence and multiplicity of solutions for equation ( ) in the degenerate case.

2 Preliminaries

Section 2 starts with a brief review of the basic properties of the Heisenberg group. For a complete treatment on the Heisenberg group, we refer to [11,15,18,26] and references therein.

The Heisenberg group H n is a Lie group of topological dimension 2 n + 1 , i.e., the Lie group that has R 2 n + 1 as a background manifold, endowed with the following group law:

τ : H n H n , τ ξ ( ξ ) = ξ ξ ,

with

ξ ξ = x + x , y + y , t + t + 2 i = 1 n ( x y y x ) , for all ξ , ξ H n .

The inverse is given by ξ 1 = ξ , so that ( ξ ξ ) 1 = ( ξ ) 1 ξ 1 .

The Lie algebra of left invariant vector fields is generated by the following vector fields:

X j = x j + 2 y j t , Y j = y j 2 x j t , T = t , j = 1 , 2 , , n .

It is straightforward to obtain that for all j , k = 1 , 2 , , n ,

[ X j , X k ] = [ Y j , Y k ] = X j , t = Y j , t = 0 and [ X j , Y k ] = 4 δ j k t .

A vector field in the span of [ X j , Y j ] j = 1 n is called horizontal.

The horizontal gradient of a C 1 function u : H n R is defined by

D H u = j = 1 n [ ( X j u ) X j + ( Y j u ) Y j ] .

Hence, D H u span { X j , Y j } j = 1 n .

The natural inner product in the span of { X j , Y j } j = 1 n is given by

( X , Y ) H = j = 1 n ( x j y j + x ˜ j y ˜ j ) ,

where X = { x j X j + x ˜ j Y j } j = 1 n and Y = { y j X j + y ˜ j Y j } j = 1 n . The inner product ( , ) H produces the Hilbertian norm

D H u H = ( D H u , D H u ) H

for the horizontal vector field D H u .

For any horizontal vector field function X = X ( ξ ) , X = { x j X j + x ˜ j Y j } j = 1 n of class C 1 ( H n , R 2 n ) , the horizontal divergence of X is given by

div H X = j = 1 n [ X j ( x j ) + Y j ( x ˜ j ) ] .

In a similar way, if u C 2 ( H n ) , then the Kohn-Spencer Laplacian, in other words, equivalently the horizontal Laplacian in H n , or the sub-Laplacian, of u is

Δ H u = j = 1 n ( X j 2 + Y j 2 ) u = j = 1 n 2 x j 2 + 2 y j 2 + 4 y j 2 x j t 4 x j 2 y j t u + 4 z 2 2 u t 2 .

Applying the famous Theorem 1.1 due to Hörmander in [13], we see that the operator Δ H is hypoelliptic. In particular, Δ H u = div H D H u for each u C 2 ( H n ) .

A well-known generalization of the Kohn-Spencer Laplacian is the horizontal p-Laplacian, 1 < p < , on the Heisenberg group, defined as

Δ H , p φ = div H ( D H φ H p 2 D H φ ) ,

for all φ C c ( H n ) .

For R > 0 , a natural group of dilation on H n is defined by δ R ( ξ ) = ( R x , R y , R 2 t ) for all ξ = ( x , y , t ) H n . Hence,

δ R ( ξ 0 ξ ) = δ R ( ξ 0 ) δ R ( ξ ) , for all ξ 0 , ξ H n .

It is easily seen that the Jacobian determinant of dilatations δ s : H n H n is constant in H n and equal to R 2 n + 2 . This is why the natural number Q = 2 n + 2 is called the homogeneous dimension of H n .

The anisotropic dilation structure on Heisenberg group leads to the Korányi norm, which is defined by

r ( ξ ) = r ( z , t ) = ( z 4 + t 2 ) 1 4 , for all ξ = ( z , t ) H n , z = ( x , y ) .

Hence, the homogeneous degree of the Korányi norm with respect to the dilations δ R , R > 0 , is equal to 1. Indeed, for all ξ = ( z , t ) H n ,

r ( δ R ( ξ ) ) = r ( R z , R 2 t ) = ( R z 4 + R 4 t 2 ) 1 4 = R r ( ξ ) .

The Korányi distance is defined as follows:

d K o ( ξ , ξ ) = r ( ξ 1 ξ ) , for all ( ξ , ξ ) H n × H n .

The Korányi open ball with radius R centered at ξ 0 is the set B R ( ξ 0 ) = { ξ H n : d K o ( ξ 0 , ξ ) < R } . For the sake of simplicity, B R ( O ) is denoted by B R , when ξ 0 = O = ( 0 , 0 ) is the natural origin of H n . Moreover, the complement of B R ( ξ 0 ) , i.e., H n \ B R ( ξ 0 ) , is denoted by B R ( ξ 0 ) c and the complement of B R ( O ) by B R c .

The Lebesgue measure on R 2 n + 1 is invariant under the left translations of the Heisenberg group. The Haar measures d ξ on Lie groups are unique up to constant multipliers, which is the reason why the Haar measure on H n corresponds to the ( 2 n + 1 ) -Lebesgue measure. In what follows, E denotes the measure of a measurable subset E of H n . Moreover, the Haar measure on H n is Q -homogeneous with respect to dilations δ R . Therefore, as shown in [18], the topological dimension 2 n + 1 of H n is strictly less than its Hausdorff dimension Q = 2 n + 2 . Indeed, for any measurable subset E of H n and any R > 0 ,

δ R ( E ) = R Q E and d ( δ R ξ ) = R Q d ξ .

In particular, if E = B R , then B R = R Q B 1 .

Now, we present some properties on the classical Sobolev spaces on the Heisenberg group H n , recalling that the underlying measure in use here is the Haar measure on H n , which agrees with the Lebesgue measure on R 2 n + 1 . For 1 p < , the linear space

L p ( H n ) = u : H n R measurable : H n u p d ξ <

endowed with the L p -norm p , i.e.,

u p = H n u p d ξ 1 p ,

is a separable Banach space and also reflexive, if 1 < p < .

For 1 p < the classical horizontal Sobolev space H W 1 , p ( H n ) on the Heisenberg group consists of all functions u L p ( H n ) such that the horizontal gradient D H u exists in the sense of distributions and D H u H L p ( H n ) , endowed with the natural norm

u H W 1 , p ( H n ) = H n u p d ξ + H n D H u H p d ξ 1 p = ( u p p + D H u p p ) 1 p , D H u p = D H u H p .

The horizontal Sobolev space H W 1 , p ( H n ) is a separable Banach space, while is reflexive, provided that 1 < p < . Moreover, C c ( H n ) is dense in H W 1 , p ( H n ) for every p , with 1 p < .

Let us assume from now on that 1 < p < and that condition ( V ) holds. The natural solution space for equation ( ) is the space

S V 1 , p ( H n ) = u H W 1 , p ( H n ) : H n V ( ξ ) u ( ξ ) p d ξ <

with the norm

(2.1) u = u S V 1 , p ( H n ) = ( u p , V p + D H u p p ) 1 p , u p , V = H n V ( ξ ) u p d ξ 1 p .

Extending the proof of Lemma 10 in [37] from the Euclidean context to the Heisenberg group, we see that S V 1 , p ( H n ) is a separable reflexive Banach space. Moreover, when 1 < p < Q , the embeddings

S V 1 , p ( H n ) H W 1 , p ( H n ) L ( H n )

are continuous for all , with p p , where

p = p Q Q p

is the critical Sobolev exponent associated with p in the Heisenberg context.

Furthermore, if 1 p < Q , then the embedding of S V 1 , p ( H n ) L ( B R ) is compact for all , with 1 < p and all R > 0 thanks to assumption ( V ) [5,30].

The celebrated Folland-Stein space S 1 , p ( H n ) , 1 p < Q , is defined as the completion of C c ( H n ) with respect to the norm

D H u p = H n D H u H p d ξ 1 p .

The Folland-Stein inequality says that if 1 < p < Q , then there exists a constant C = C ( p , Q ) > 0 such that

(2.2) φ p * C D H φ p , for all φ C c ( H n ) .

In [46], when 1 < p < Q , Vassilev proved the existence of extremal functions of (2.2) by the concentration-compactness method of Lions. However, the best constant

(2.3) S = inf u S 1 , p ( H n ) u 0 D H u p p u p p

of (2.2) is not explicitly determined yet.

Proposition 2.1

(Hardy-Littlewood-Sobolev inequality [8]) If , s > 1 and 0 < λ < Q , with

1 + λ Q + 1 s = 2 ,

then there exists a sharp constant C ( , s , λ , Q ) such that

(2.4) H n × H n u ( ξ ) v ( η ) η 1 ξ H λ d η d ξ C ( , s , λ , Q ) u v s ,

for all u L ( H n ) and v L s ( H n ) .

If = s = 2 Q ( 2 Q λ ) , then

C ( λ , Q ) = C ( , s , λ , Q ) = π N + 1 2 N 1 N ! λ Q N ! Γ ( ( Q λ ) 2 ) Γ 2 ( ( 2 Q λ ) 2 ) ,

where Γ is the Gamma function. In this case, equality holds in (2.4) if and only if either u 0 or v 0 , or u Constant v , where

v ( ξ ) = c U ( δ R ( ξ 0 1 ξ ) ) , ξ H n ,

for some c C , R > 0 , ξ 0 H n and where U is defined by

U ( ξ ) = U ( x , y , t ) = ( t 2 + ( 1 + x 2 + y 2 ) 2 ) ( 2 Q λ ) 4 ,

for all ξ = ( x , y , t ) H n .

Proposition 2.1 guarantees that the integral

(2.5) H n × H n u ( ξ ) β u ( η ) β η 1 ξ H λ d η d ξ

is well defined if u β L ( H n ) for some > 1 such that

2 + λ Q = 2 .

The Folland-Stein inequality (2.2) shows that (2.5) is well defined for u S 1 , p ( H n ) only if

p β Q p Q p ,

i.e., β has to satisfy

(2.6) p λ β p λ , where p λ = p ( 2 Q λ ) 2 Q and p λ = p ( 2 Q λ ) 2 ( Q p ) .

Hence, it is quite natural to call p λ the lower critical exponent and p λ the upper critical exponent on the Heisenberg group. Therefore, Proposition 2.1 shows that

H n × H n u ( ξ ) p λ u ( η ) p λ η 1 ξ H λ d η d ξ C ( λ , Q ) u p 2 p λ ,

where

(2.7) p λ 2 Q 2 Q λ = p .

In our context, we define

(2.8) S H G , p = inf u S 1 , p ( H n ) u 0 D H u p p u F L , p p , where u F L , p = H n × H n u p λ u p λ η 1 ξ H λ d η d ξ 1 2 p λ ,

where p λ is given in (2.6). Proposition 2.1 implies also that

S H G , p inf u S 1 , p ( H n ) u 0 D H u p p C ( λ , Q ) p 2 p λ u p p = C ( λ , Q ) p 2 p λ S ,

where S is the best constant introduced in (2.3).

Lemma 2.2

(Proposition 4.1 of [10]) Let 0 < λ < Q . There exists a function h : H n R , which is even, i.e., h ( ξ 1 ) = h ( ξ ) , ξ H n , positively homogeneous of degree α = ( Q + λ ) 2 , i.e., h ( t ξ ) = t α h ( ξ ) , ξ H n , and t > 0 , and of class L 2 Q Q + λ ( H n ) such that

1 η 1 ξ H λ = H n h ( η 1 σ ) h ( ξ 1 σ ) d σ ,

for all ξ , η H n

Lemma 2.3

Let 0 < λ < Q and X F L , p = { u : H n R : u F L , p < } . Then, F L , p defined by (2.8) is a norm on X F L , p . Moreover, X F L , p = ( X F L , p , F L , p ) is a Banach space.

Proof

Thanks to Lemma 2.2, we obtain for all u L p ( H n ) ,

(2.9) u F L , p 2 p λ = H n × H n u ( ξ ) p λ u ( η ) p λ η 1 ξ H λ d η d ξ = H n ( ξ λ u p λ ) u p λ d ξ = H n ( h u p λ ) 2 d ξ .

Consequently, the non-negativity, the definiteness, i.e., u F L , p = 0 if and only if the u = 0 , the absolute homogeneity of F L , p are clearly satisfied. It remains to show the triangle inequality. Indeed, using Minkowski’s inequality, we obtain

( h u + v p λ ) 2 = H n h ( η 1 ξ ) u ( η ) + v ( η ) p λ d η 1 p λ 2 p λ = H n h ( η 1 ξ ) u ( η ) + v ( η ) p λ d η 1 p λ 2 p λ H n h ( η 1 ξ ) u ( η ) p λ d η 1 p λ + H n h ( η 1 ξ ) v ( η ) p λ d η 1 p λ 2 p λ = H n h ( η 1 ξ ) u ( η ) p λ d η 2 2 p λ + H n h ( η 1 ξ ) v ( η ) p λ d η 2 2 p λ 2 p λ .

Applying the Minkowski inequality again, we obtain

H n ( h u + v p λ ) 2 d ξ 1 2 p λ H n H n h ( η 1 ξ ) u ( η ) p λ d η 2 2 p λ + H n h ( η 1 ξ ) v ( η ) p λ d η 2 2 p λ 2 p λ d ξ 1 2 p λ H n H n h ( η 1 ξ ) u ( η ) p λ d η 2 d ξ 1 2 p λ + H n H n h ( η 1 ξ ) v ( η ) p λ d η 2 d ξ 1 2 p λ = H n ( h u p λ ) 2 d ξ 1 2 p λ + H n ( h v p λ ) 2 d ξ 1 2 p λ ,

which implies

u + v F L u F L + v F L .

Hence, X F L , p is a real vector space and F L , p is a norm on it.

Let us finally verify that X F L , p = ( X F L , p , F L , p ) is a Banach space. Let ( u k ) k be a Cauchy sequence in X F L , p . In order to complete this proof, it suffices to verify that a subsequence converges in X F L , p .

We extract a subsequence ( v k ) k , v k = u n k , such that

v k + 1 v k p 1 2 k , for all k 1 .

In fact, choosing n 1 such that u m u k p 1 2 for all m , k n 1 , then choose n 2 n 1 such that u m u k p 2 2 , for all m , k n 2 etc. We claim that ( v k ) k converges in X F L , p . Indeed,

v k + 1 v k p 2 k , for all k 1 .

Let

g k = k = 1 n v k + 1 v k p

such that

g k p ε .

Therefore, by the monotone convergence theorem, we have g k g a.e. H n and in L p ( H n ) , with g L p ( H n ) .

On the other hand, using Hardy-Littlewood-Sobolev inequality and Minkowski’s inequality, we obtain

H n ( h v m v k p λ ) 2 d ξ = H n H n v m ( η ) v k ( η ) p λ v m ( ξ ) v k ( ξ ) p λ η 1 ξ H λ d η d ξ C ( λ , Q ) v m v k p 2 p λ = C ( λ , Q ) v m v m 1 + + v k + 1 v k p 2 p λ = C ( λ , Q ) ( v m v m 1 ) + + ( v k + 1 v k ) p 2 p λ = C ( λ , Q ) ( H n ( v m v m 1 ) + + ( v k + 1 v k ) p d ξ ) 2 p λ p C ( λ , Q ) H n ( v m v m 1 ) p d ξ 2 p λ p + + H n v k + 1 v k p d ξ 2 p λ p < ε .

It follows that ( v k ) k is a Cauchy sequence a.e. in H n and in L p ( H n ) . Hence, ( v k ) k converges to some measurable function u , which is a.e. finite in H n . Moreover, for all k 2 ,

H n ( h v k u p λ ) 2 d ξ < ε ,

a.e. in H n . This shows that u L p ( H n ) . Therefore, v k u F L , p 0 . Moreover, the entire Cauchy sequence ( u k ) k converges to u in X F L , p . In conclusion, X F L , p = ( X F L , p , F L , p ) is a Banach space, as required.□

Lemma 2.4

Let ( u k ) k be a bounded sequence in X F L , p . If u k u in L p ( H n ) and u k u in a.e. H n as k , then

u k F L , p 2 p λ u k u F L , p 2 p λ u F L , p 2 p λ ,

as k .

Proof

Using the Clarkson and Mil’man theorems, u k u p λ 0 in L 2 Q 2 Q λ ( H n ) as k , by (2.7). A similar argument as in the famous proof of the Brézis-Lieb lemma shows that

(2.10) u k p λ u k u p λ u p λ in L 2 Q 2 Q λ ( H n ) ,

as k . The Hardy-Littlewood-Sobolev inequality implies that

(2.11) ξ λ ( u k p λ u k u p λ ) ξ λ u p λ , in L 2 Q λ ( H n ) ,

as k . Thanks to (2.9), we obtain

(2.12) u k F L , p 2 p λ u k u F L , p 2 p λ = H n ξ λ ( u k p λ ) u k p λ d ξ H n ξ λ ( u k u p λ ) u k u p λ d ξ = H n [ ξ λ ( u k p λ u k u p λ ) ] ( u k p λ u k u p λ ) d ξ + 2 H n [ ξ λ ( u k p λ u k u p λ ) ] u k u p λ d ξ .

Substituting (2.10) and (2.11) into (2.12), we obtain the desired result as k .□

3 Palais-Smale condition

From now on, we assume that the structural conditions ( V ) , ( K 2 ) ( K 3 ) , and ( f 1 ) ( f 3 ) are satisfied. Hence, equation ( ) is variational and the underlying functional is

(3.1) μ ( u ) = 1 p K ( u p ) 1 2 p λ u F L , p 2 p λ μ H n F ( ξ , u ) d ξ ,

for all u S V 1 , p ( H n ) , where is defined in (2.1).

Moreover, μ is of class C 1 ( S V 1 , p ( H n ) ) and its Fréchet derivative is given for all u S V 1 , p ( H n ) by

μ ( u ) , v = K ( u p ) H n { u ( ξ ) p 2 u ( ξ ) v ( ξ ) + D H u ( ξ ) H p 2 ( D H u ( ξ ) , D H v ( ξ ) ) } d ξ H n × H n u ( η ) p λ u ( ξ ) p λ 2 u ( ξ ) v ( ξ ) η 1 ξ H λ d η d ξ + μ H n f ( ξ , u ) v d ξ ,

pointwise for any v S V 1 , p ( H n ) . Therefore, the critical points of functional μ are exactly the (weak) solutions of equation ( ).

Now, we state the concentration-compactness principle of Choquard equation involving p -Laplacian on the Heisenberg group, which is essential tool to prove that the Palais-Smale condition holds at special levels.

Theorem 3.1

Let ( u k ) k be a bounded sequence in S 1 , p ( H n ) such that u k u in S 1 , p ( H n ) and u k u a.e. in H n . Furthermore, let ν , ω and ζ be three finite nonnegative Radon measures on H n . Assume that as k ,

H n u k ( η ) p λ η 1 ξ H λ d η u k ( ξ ) p λ d ξ ν , D H u k H p d ξ ω , u k p d ξ ζ , in ( H n ) .

Introduce the real numbers ν , ω , and ζ defined by

ν = lim R lim ¯ n ξ R H n u k ( η ) p λ η 1 ξ H λ d η u k ( ξ ) p λ d ξ , ω = lim R lim ¯ n ξ R D H u k H p d ξ , ζ = lim R lim ¯ n ξ R u k p d ξ .

Then, there exist a at most countable sequence of points { z j } j J H n and families of positive numbers { ν j } j J , { ζ j } j J , and { ω j } j J such that

(3.2) ν = H n u ( η ) p λ η 1 ξ H λ d η u ( ξ ) p λ d ξ + j J ν j δ z j , j J ν j 1 p λ < ,

(3.3) ω D H u H p d ξ + j J ω j δ z j ,

(3.4) ζ u p d ξ + j J ζ j δ z j ,

and

(3.5) S H G , p ν j p 2 p λ ω j , ν j Q 2 Q λ C ( λ , Q ) Q 2 Q λ ζ j ,

where δ ξ is the Dirac-mass of mass 1 concentrated at ξ H n .

Finally, the energy at infinity verifies

(3.6) lim ¯ k u k F L , p 2 p λ = ν + H n d ν ,

(3.7) lim ¯ k D H u k p p = H n d ω + ω ,

(3.8) lim ¯ k D H u k p p = H n d ζ + ζ ,

and

(3.9) C ( λ , Q ) 2 Q 2 Q λ ν 2 Q 2 Q λ ζ H n d ζ + ζ , S 2 C ( λ , Q ) p p λ ν p p λ ω H n d ω + ω .

Proof

Let ( u k ) k be the sequence as in the statement of the theorem. Hence, the sequence k v k = u k u converges weakly to 0 in S 1 , p ( H n ) and a.e. in H n . Lemma 2.4 implies that

D H v k H p d ξ ϖ , where ϖ = ω D H u H p ,

H n v k ( η ) p λ η 1 ξ H λ d η v k ( ξ ) p λ d ξ κ , with κ = ν H n u ( η ) p λ η 1 ξ H λ d η u ( ξ ) p λ d ξ ,

v k p d ξ ς , provided that ς = ζ u p .

In order to verify the possible concentration at finite points, we first show that

(3.10) H n ( ξ λ φ v k ( ξ ) p λ ) φ v k ( ξ ) p λ d ξ H n ( ξ λ v k ( ξ ) p λ ) φ ( ξ ) p λ φ v k ( ξ ) p λ d ξ 0 ,

for any arbitrarily fixed φ C c ( H n ) . To see this, first put

(3.11) ϒ k ( ξ ) = ( ξ λ φ v k ( ξ ) p λ ) φ v k ( ξ ) p λ ( ξ λ v k ( ξ ) p λ ) φ ( ξ ) p λ φ v k ( ξ ) p λ .

Since φ C c ( H n ) , for every δ > 0 , there exists M > 0 such that

(3.12) ξ M ϒ k ( ξ ) d ξ < δ , for all k .

Since the Riesz potential defines a linear operator and v k ( ξ ) 0 a.e. in H n as k , we obtain that

H n v k ( η ) p λ η 1 ξ H λ d η 0 , a.e. in H n .

Therefore, also ϒ k ( ξ ) 0 a.e. in H n . By (3.11), for all ξ H n ,

ϒ k ( ξ ) = H n ( φ ( η ) p λ φ ( ξ ) p λ ) v k ( η ) p λ η 1 ξ H λ d η φ v k ( ξ ) p λ = H n P ( ξ , η ) v k ( η ) p λ d η φ v k ( ξ ) p λ ,

where H n × H n ( ξ , η ) P ( ξ , η ) = φ ( η ) p λ φ ( ξ ) p λ η 1 ξ H λ .

Moreover, for almost all ξ , there exists R > 0 large enough such that

H n P ( ξ , η ) v k ( η ) p λ d η = η R P ( ξ , η ) v k ( η ) p λ d η φ ( ξ ) p λ η R v k ( η ) p λ η 1 ξ H λ d η .

For each ξ , the function P ( ξ , ) L r ( B R ) , 1 r < Q λ if λ > 1 , 1 r if 0 < λ 1 , as proved in [23]. By the Young inequality, there exists s > 2 Q λ such that

B M B R P ( ξ , η ) v k ( η ) p λ d η s d ξ 1 s C φ P ( ξ , η ) r v k p λ 2 Q 2 Q λ C φ ,

where M is given in (3.12). It can be easily seen that for R > 0 large enough

B M φ ( ξ ) p λ η R v k ( η ) p λ η 1 ξ H λ d η s d ξ 1 s C ,

and hence, we have

B M H n P ( ξ , η ) v k ( η ) p λ d η s d ξ 1 s C φ .

Therefore, we can obtain for θ > 0 small enough

B M ϒ k ( ξ ) 1 + θ d ξ B M H n P ( ξ , η ) v k ( η ) p λ d η s d ξ 1 s B M φ v k p d ξ 2 Q λ 2 Q C φ .

Combining this with the fact and ϒ k 0 a.e. in H n , we obtain that

B M ϒ k ( ξ ) d ξ 0 as k .

Hence, this and (3.12) imply that

H n ϒ k ( ξ ) d ξ 0 , as k .

The Hardy-Littlewood-Sobolev inequality (2.4) yields for all φ C c ( H n ) that

H n H n φ v k ( η ) p λ η 1 ξ H λ d η φ v k ( ξ ) p λ d ξ C ( λ , Q ) φ v k p 2 p λ .

Using (3.10), we obtain as k that

H n φ ( ξ ) 2 p λ H n v k ( η ) p λ η 1 ξ H λ d η v k ( ξ ) p λ d ξ C ( λ , Q ) φ v k p 2 p λ + o ( 1 ) .

Letting k , we obtain

(3.13) H n φ ( ξ ) 2 p λ d κ C ( λ , Q ) H n φ p d ς 2 Q λ Q .

Applying Lemma 1.2 of [22], we see that (3.4) holds true.

Taking φ = χ { ζ j } , j J , in (3.13), we have

ν j Q 2 Q λ C ( λ , Q ) Q 2 Q λ ζ j , for all j J .

The definition of S H G , p gives

H n H n φ v k ( ξ ) p λ η 1 ξ H λ d η φ v k ( ξ ) p λ d ξ p 2 p λ S H G , p H n D H ( φ v k ) p d ξ .

The elementary -norm inequality in R 2 N yields

(3.14) D H ( φ v k ) φ D H v k D H φ v k .

Now, v k 0 in L ( B R ) for every , with 1 < p , and every R > 0 , thanks to assumption ( V ) . Thus, the right-hand side of (3.14) goes to 0 as k . Together with (3.10), we obtain as k ,

H n H n φ ( ξ ) 2 p λ v k ( η ) p λ η 1 ξ H λ d η v k ( ξ ) p λ d ξ p 2 p λ S H G , p H n φ p D H v k p d ξ + o ( 1 ) .

Passing to the limit as k , we have

(3.15) H n φ ( ξ ) 2 p λ d κ p 2 p λ S H G , p H n φ p d ϖ .

Once again applying Lemma 1.2 in [22], we obtain that (3.3) holds.

Let φ χ { ζ j } , j J , in (3.15). Thus,

S H G , p ν j p 2 p λ ω j , j J .

Hence, (3.2) and (3.5) hold.

Next, we prove the possible loss of mass at infinity. Using the similar technique as in [36], we can prove (3.7) and (3.8). Let ϕ C c ( H n ) be such that 0 ϕ 1 , ϕ = 0 in B 1 and ϕ = 1 in B 2 c . Taking R > 1 and put ϕ R ( ξ ) = ϕ ( δ 1 R ( ξ ) ) , ξ H n . For every R > 1 , we have

lim ¯ k H n × H n v k ( η ) p λ v k ( ξ ) p λ η 1 ξ H λ d η d ξ = lim ¯ k H n × H n u k ( η ) p λ u k ( ξ ) p λ η 1 ξ H λ d η d ξ H n × H n u ( η ) p λ u ( ξ ) p λ η 1 ξ H λ d η d ξ = lim ¯ k H n × H n u k ( η ) p λ u k ( ξ ) p λ ϕ R ( ξ ) η 1 ξ H λ d η d ξ + H n × H n u k ( η ) p λ u k ( ξ ) p λ ( 1 ϕ R ( ξ ) ) η 1 ξ H λ d η d ξ H n × H n u ( η ) p λ u ( ξ ) p λ η 1 ξ H λ d η d ξ = lim ¯ k H n × H n u k ( η ) p λ u k ( ξ ) p λ ϕ R ( ξ ) η 1 ξ H λ d η d ξ + H n ( 1 ϕ R ) d ν + H n × H n u ( η ) p λ u ( ξ ) p λ ( 1 ϕ R ( ξ ) ) η 1 ξ H λ d η d ξ H n × H n u ( η ) p λ u ( ξ ) p λ η 1 ξ H λ d η d ξ .

As R the Lebesgue theorem gives (3.6), as required. Moreover, by the Hardy-Littlewood-Sobolev inequality, we have

ν = lim R lim ¯ k H n H n u k ( η ) p λ η 1 ξ H λ d η ϕ R u k ( ξ ) p λ d ξ C ( λ , Q ) lim R lim ¯ k H n u k p d ξ H n ϕ R u k p d ξ 2 Q λ 2 Q = C ( λ , Q ) ζ H n d ζ + ζ 2 Q λ 2 Q ,

which implies

C ( λ , Q ) 2 Q 2 Q λ ν 2 Q 2 Q λ ζ H n d ζ + ζ ,

i.e., (3.9) 1 holds true. Similarly, according to the definition of S H G , p and ν , we have

ν = lim R lim ¯ k H n H n u k ( η ) p λ η 1 ξ H λ d η ϕ R u k ( ξ ) p λ d ξ C ( λ , Q ) lim R lim ¯ k H n u k p d ξ H n u k ϕ R p d ξ 2 Q λ 2 Q S 2 p λ p C ( λ , Q ) lim R lim ¯ k H n D H u k H p d ξ H n D H ( ϕ R u k ) H p d ξ p λ p = S 2 p λ p C ( λ , Q ) ω H n d ω + ω p λ p ,

which implies that

S 2 C ( λ , Q ) p p λ ν p p λ ω H n d ω + ω ,

namely, (3.9) 2 is verified. This ends the proof of the theorem.□

Throughout this article, S V 1 , p ( H n ) denotes the dual of S V 1 , p ( H n ) = ( S V 1 , p ( H n ) , ) , where is defined in (2.1). Furthermore, we say that ( u k ) k S V 1 , p ( H n ) is a Palais-Smale sequence of μ at level c μ , briefly ( P S ) c μ , if

(3.16) μ ( u k ) c μ , and μ ( u k ) 0 , in S V 1 , p ( H n )

as k .

Lemma 3.2

Assume that K satisfies ( K 1 ) ( K 2 ) , f satisfies ( f 1 ) ( f 3 ) , and ( V ) holds. Let ( u k ) k S V 1 , p ( H n ) be a ( P S ) c μ sequence of the functional μ at some level c μ . If

(3.17) 0 < c μ < 1 σ 1 2 p λ ( k 0 S C ( λ , Q ) p 2 p λ ) 2 p λ 2 p λ p ,

where S H G , p is defined in (2.8) and k 0 in the non-degeneracy positive number in ( K 1 ) , then there exists a subsequence of ( u k ) k strongly convergent in S V 1 , p ( H n ) .

Proof

Let ( u k ) k be a ( P S ) c μ sequence for μ , with c μ satisfying (3.17). We proceed diving the proof into two steps.

Step 1. We claim that ( u k ) k is bounded in S V 1 , p ( H n ) .

By ( K 1 ) , ( K 2 ) , and ( f 3 ) , as k ,

(3.18) c μ + 1 + o ( 1 ) u k = μ ( u k ) 1 σ μ ( u k ) , u k = 1 p K ( u k p ) 1 σ K ( u k p ) u k p + 1 σ 1 2 p λ u k F L , p 2 p λ + μ H n 1 σ f ( ξ , u k ) u k F ( ξ , u k ) d ξ 1 θ p σ k 0 p u k p + 1 σ 1 2 p λ u k F L , p 2 p λ + μ H n 1 σ f ( ξ , u k ) u k F ( ξ , u k ) d ξ 1 θ p σ k 0 p u k p .

Since σ > p θ by ( f 3 ) , inequality (3.18) shows that ( u k ) k is bounded in S V 1 , p ( H n ) , as claimed.

Step 2. We claim that ( u k ) k admits a strongly convergent subsequence in S V 1 , p ( H n ) .

Theorem 3.1 and Step 1 show that, up to a subsequence, there exists some u S V 1 , p ( H n ) such that

(3.19) u k u , a.e. in ( H n ) ,

u k u , in S V 1 , p ( H n ) ,

D H u k H p d ξ ω D H u H p d ξ + j J ω j δ ξ j ,

H n u k ( η ) p λ η 1 ξ H λ d η u k p λ d ξ ν = H n u ( η ) p λ η 1 ξ H λ d η u p λ d ξ + j J ν j δ z j ,

(3.20) S H G , p ν j p 2 p λ ω j , for all j J .

Moveover, we obtain

(3.21) limsup k H n H n u ( ξ ) p λ u ( η ) p λ η 1 ξ H λ d η d ξ = ν ( H n ) + ν , limsup k H n D H u k H p d ξ = ω ( H n ) + ω , S C p 2 p λ ω p 2 p λ ω .

Next, we are going to prove the following two claims.

Claim 1. J = .

Assume by contradiction that J and take j J . Fix a cut-off function φ C c ( H n ) such that 0 φ 1 , φ ( O ) = 1 , and supp φ = B 1 ¯ . Let ε > 0 and put φ ε ( ξ ) = φ ( δ 1 ε ( ξ ) ) , ξ H n . Then, ( u k φ ε ) k is bounded in S V 1 , p ( H n ) . Obviously, μ ( u k ) , u k φ ε 0 , i.e., as k ,

(3.22) K ( u k p ) H n D H u k p φ ε d ξ + H n V ( ξ ) u k p φ ε d ξ = μ H n f ( ξ , u k ) u k φ ε d ξ + H n × H n u ( η ) p λ u ( ξ ) p λ φ ε η 1 ξ H λ d η d ξ + o k ( 1 ) .

On the one hand, thanks to ( f 2 ) and the Hölder inequality, we have

(3.23) H n f ( ξ , u k ) u k φ ε d ξ H n a ( ξ ) u k q φ ε d ξ 2 a ϑ ( B ε ( z j ) ) u k r ( B ε ( z j ) ) q C a ϑ ( B ε ( z j ) ) ,

where ϑ and r are given in assumption ( f 2 ) . Hence, it follows from (3.23) that

(3.24) lim ε 0 + lim k H n f ( ξ , u k ) u k φ ε d ξ = 0 .

On the other hand, let δ > 0 be arbitrary and fixed. Invoking Theorem 3.1, we deduce

(3.25) lim k H n D H u k H p φ ε d ξ = H n φ ε d ω H n D H u H p φ ε d ξ + ω j .

We also obtain that

(3.26) lim ε 0 + lim k H n V ( ξ ) u k p φ ε d ξ = 0 ,

(3.27) lim k H n × H n u ( η ) p λ u ( ξ ) p λ φ ε η 1 ξ H λ d η d ξ = H n φ ε d ν + ν j .

Substituting (3.24)–(3.27) into (3.22) and using ( K 1 ) , as ε 0 + and k in (3.22), we obtain that

k 0 ω j ν j .

Using this and (3.20), we obtain that

(3.28) ν j ( k 0 S H G , p ) 2 p λ 2 p λ p ( k 0 S C ( λ , Q ) p 2 p λ ) 2 p λ 2 p λ p .

According to μ ( u k ) c μ and μ ( u k ) 0 as k , it follows from (3.28) that

(3.29) c μ = lim k μ ( u k ) = lim k μ ( u k ) 1 σ μ ( u k ) , u k lim k 1 σ 1 2 p λ H n × H n u k ( ξ ) p λ u k ( η ) p λ η 1 ξ H λ d η d ξ 1 σ 1 2 p λ ν j 1 σ 1 2 p λ ( k 0 S C ( λ , Q ) p 2 p λ ) 2 p λ 2 p λ p > c μ ,

thanks to assumption (3.17) and the fact that θ p < σ < p < 2 p λ by ( f 3 ) and (2.6). This is impossible. Therefore, J = .

Claim 2. ν = 0 .

To obtain the possible concentration of mass at infinity, we suppose on the contrary that ν > 0 . Define a cut-off function ϕ C c ( H n ) such that 0 ϕ 1 , ϕ = 0 in B 1 , and ϕ = 1 in B 2 c . Take R > 0 and put ϕ R ( ξ ) = ϕ ( δ 1 R ( ξ ) ) , ξ H n . Again, the fact that ( u k ) k is bounded in S V 1 , p ( H n ) yields that ( u k ϕ R ) k is bounded in S V 1 , p ( H n ) . Consequently, μ ( u k ) , u k ϕ R 0 as k , which means

(3.30) K ( u k p ) H n D H u k H p ϕ R d ξ + H n V ( ξ ) u k p ϕ R d ξ = μ H n f ( ξ , u k ) u k ϕ R d ξ + H n × H n u k ( η ) p λ u k ( ξ ) p λ ϕ R η 1 ξ H λ d η d ξ + o k ( 1 ) .

Analogous to the argument of (3.24), we obtain that

(3.31) lim R lim k H n f ( ξ , u k ) u k ϕ R d ξ = 0 .

Proceeding as in (3.25) and (3.26), we obtain

(3.32) lim k H n D H u k H p ϕ R d ξ = H n ϕ R d μ H n D H u H p ϕ R d ξ + ω

and

(3.33) lim R lim k H n V ( ξ ) u k p ϕ R d ξ = 0 .

Also, we deduce that

(3.34) lim k H n × H n u k ( η ) p λ u k ( ξ ) p λ ϕ R η 1 ξ H λ d η d ξ = H n ϕ R d ν = H n × H n u ( η ) p λ u ( ξ ) p λ ϕ R η 1 ξ H λ d η d ξ + ν .

Letting R , k in (3.30), we obtain from (3.31)–(3.34) and ( K 1 ) that

(3.35) k 0 ω ν .

The combination of (3.21) with (3.35) gives

ν ( k 0 S C ( λ , Q ) p 2 p λ ) 2 p λ 2 p λ p .

Hence,

c μ = lim k μ ( u k ) = lim k μ ( u k ) 1 σ μ ( u k ) , u k 1 σ 1 2 p λ lim k H n × H n u k ( η ) p λ u k ( ξ ) p λ ϕ R η 1 ξ H λ d η d ξ 1 σ 1 2 p λ ν 1 σ 1 2 p λ ( k 0 S C ( λ , Q ) p 2 p λ ) 2 p λ 2 p λ p > c μ ,

which is impossible. Thus, ν = 0 .

Together with the fact that J = and ν = 0 , we have

lim k H n × H n u k ( η ) p λ u k ( ξ ) p λ η 1 ξ H λ d η d ξ = H n × H n u ( η ) p λ u ( ξ ) p λ η 1 ξ H λ d η d ξ .

The Brézis-Lieb lemma for the classical Lebesgue spaces yields from the last equality and (3.19) that

H n × H n u k ( η ) u ( η ) p λ u k ( ξ ) u ( ξ ) p λ η 1 ξ H λ d η d ξ 0 ,

i.e., u k u p λ 0 in L 2 Q 2 Q λ ( H n ) . Therefore, we have

H n × H n u k ( η ) p λ u k ( ξ ) p λ 2 u k ( u k u ) η 1 ξ H λ d η d ξ 0 .

Furthermore, we obtain

H n f ( ξ , u k ) ( u k u ) d ξ 0 .

For the sake of simplicity, let us introduce for all v S V 1 , p ( H n ) the linear functional L ( v ) on S V 1 , p ( H n ) , defined for all ϖ S V 1 , p ( H n ) by

(3.36) L ( v ) , ϖ = H n D H u H p 2 D H u D H ϖ d ξ + H n V ( ξ ) v p 2 v ϖ d ξ .

The Hölder inequality implies at once that L ( v ) is continuous and satisfies

L ( v ) , ϖ v ϖ , for all ϖ S V 1 , p ( H n ) .

Consequently, the weak convergence of ( u k ) k in S V 1 , p ( H n ) and the boundedness of K ( u k p ) K ( u p ) k in R show that

(3.37) lim k [ K ( u k p ) K ( u p ) ] L ( u ) , u k u = 0 .

Obviously, μ ( u k ) , u k u 0 . Hence, (3.37) implies as k that

o k ( 1 ) = μ ( u k ) μ ( u ) , u k u = K ( u k p ) L ( u k ) , u k u K ( u p ) L ( u ) , u k u + ( K ( u k p ) K ( u p ) ) L ( u ) , u k u μ H n ( f ( ξ , u k ) u k f ( ξ , u ) u ) ( u k u ) d ξ H n ( ξ λ ( u k p λ 2 u k u p λ 2 u ) ) ( u k u ) d ξ = K ( u k p ) [ L ( u k ) , u k u L ( u ) , u k u ] + o k ( 1 ) .

This yields that

lim k K ( u k p ) [ L ( u k ) , u k u L ( u ) , u k u ] = 0 .

Since

K ( u k p ) [ L ( u k ) , u k u L ( u ) , u k u ] 0 ,

for all k , by ( K 1 ) , in particular, we obtain

lim k [ L ( u k ) , u k u L ( u ) , u k u ] = 0 .

Hence, this and (3.36) yield that

lim k u k u p = lim k ( L ( u k ) , u k u L ( u ) , u k u ) = 0 .

Hence, ( u k ) k strongly converges to u in S V 1 , p ( H n ) . This completes the proof.□

4 Non-degenerate case

This section is devoted to the proofs of Theorems 1.1 and 1.2, so that from here on, we assume that ( V ) , ( K 1 ) ( K 2 ) and ( f 1 ) ( f 3 ) are satisfied.

4.1 Proof of Theorem 1.1

A key tool for the proof of Theorem 1.1 is the mountain pass lemma, and we recall it in the version given, e.g., in [2,3].

Theorem 4.1

Let E be a real Banach space and C 1 ( E ) , with ( 0 ) = 0 . Suppose that

  1. There exists ρ , α > 0 such that ( u ) α for all u E , with u E = ρ ;

  2. There exists e E satisfying e E > ρ such that ( e ) < 0 .

Then,

c = inf γ Γ max 0 t 1 ( γ ( t ) ) α , where Γ = { γ C ( [ 0 , 1 ] , E ) : γ ( 0 ) = 1 , γ ( 1 ) = e } ,

and admits a ( P S ) c sequence ( u k ) k E .

Next, we show that the functional μ in (3.1) satisfies the geometric properties ( a ) and ( b ) of Theorem 4.1 when E = S V 1 , p ( H n ) .

Lemma 4.2

The functional μ satisfies the geometric properties (a)–(b) of Theorem 4.1.

Proof

Fix μ > 0 and u S V 1 , p ( H n ) , with 0 < u 1 . Assumptions ( K 1 ) , ( V ) , and (2.8) yield

μ ( u ) = 1 p K ( u p ) μ H n F ( ξ , u ) d ξ 1 2 p λ u F L , p 2 p λ k 0 θ p u p 2 q μ c u q D H u p 2 p λ 2 p λ S H G , p 2 p λ p k 0 θ p u p 2 q μ c u q u 2 p λ 2 p λ S H G , p 2 p λ p ,

where c is a positive constant. Then, we can select ρ , α > 0 such that μ ( u ) α for u = ρ , since 1 < p < θ p < q by ( K 2 ) and ( f 2 ) , while p < p * < 2 p λ by (2.6). Thus, ( a ) of Theorem 4.1 holds true.

Next, to prove ( b ) in Theorem 4.1, let ω C 0 ( H n ) with ω 0 and ω 0 . Assumption ( K 2 ) gives that

(4.1) K ( t ) K ( 1 ) t θ , for all t 1 .

Using ( f 1 ) ( f 3 ) , for all t > 1 so large that t ω p 1 , we obtain

μ ( t ω ) 1 p K ( t ω p ) μ H n F ( ξ , t ω ) d ξ 1 2 p λ t ω F L , p 2 p λ 1 p K ( 1 ) ω θ p t θ p 1 2 p λ ω F L , p 2 p λ t 2 p λ .

Since θ p < p < 2 p λ by ( K 2 ) and (2.6), we deduce that μ ( t 0 ω ) < 0 and t 0 ω > ρ for t 0 large enough. Hence, the function e = t 0 ω satisfies property ( b ) in Theorem 4.1. This completes the proof.□

Proof of Theorem 1.1

We claim that

(4.2) 0 < c μ = inf γ Γ max 0 t 1 μ ( γ ( t ) ) < 1 σ 1 2 p λ k 0 S C ( λ , Q ) p 2 p λ 2 p λ 2 p λ p ,

for all μ large enough. Clearly, if (4.2) holds, then Lemmas 3.2 and 4.2 and Theorem 4.1 show the existence of nontrivial critical points of μ at level c μ , provided that μ > 0 is sufficiently large. Consequently, the validity of (4.2) would end the proof.

To prove (4.2), take v 0 S V 1 , p ( H n ) such that

v 0 = 1 and lim t μ ( t v 0 ) = .

Then, sup t 0 μ ( t v 0 ) = μ ( t μ v 0 ) for some t μ > 0 . Therefore, t μ satisfies

(4.3) K ( t μ v 0 p ) t λ v 0 p = μ H n f ( ξ , t μ v 0 ) t μ v 0 d ξ + t μ v 0 F L , p 2 p λ .

Now, we claim that { t μ } μ > 0 is bounded. Indeed, let t μ 1 and μ > 0 . Then, ( K 2 ) , (4.1), and (4.3) show that

θ K ( 1 ) t μ 2 θ p θ K ( 1 ) ( t μ v 0 p ) θ K ( t μ v 0 p ) t μ v 0 p = μ H n f ( ξ , t μ v 0 ) t μ v 0 d ξ + t μ v 0 F L , p 2 p λ v 0 F L , p 2 p λ t μ 2 p λ .

Hence, { t μ } μ > 0 is bounded, since θ p < p < 2 p λ by ( K 2 ) and (2.6).

Now, we claim that t μ 0 as μ . Otherwise, we can suppose that there exists t 0 > 0 and a sequence ( μ k ) k , with μ k as k , such that t μ k t 0 as k . The Lebesgue-dominated convergence theorem implies as k that

H n f ( ξ , t μ k v 0 ) t μ k v 0 d ξ H n f ( ξ , t 0 v 0 ) t 0 v 0 d ξ R +

by ( f 3 ) . This implies that

μ k H n f ( ξ , t μ k v 0 ) t μ k v 0 d ξ , as k .

Therefore, from (4.3), we deduce that this is impossible.

Hence, t μ 0 as μ . Furthermore, from (4.3), we have

lim μ μ H n f ( ξ , t μ v 0 ) t μ v 0 d ξ = 0 and lim μ t μ v 0 F L , p 2 p λ = 0 .

Combining this with the fact that t μ 0 as μ and the definition of μ , we have

lim μ ( sup t 0 μ ( t v 0 ) ) = lim μ μ ( t μ v 0 ) = 0 .

Therefore, there exists μ 1 > 0 such that for any μ μ 1 ,

sup t 0 μ ( t v 0 ) < ( k 0 S C ( λ , Q ) p 2 p λ ) 2 p λ 2 p λ p .

Choosing e = t 1 v 0 , with t 1 so large that μ ( e ) < 0 , we obtain that

0 < c μ max t [ 0 , 1 ] μ ( γ ( t ) ) , taking γ ( t ) = t e .

Consequently,

0 < c μ sup t 0 μ ( t v 0 ) < k 0 S C ( λ , Q ) p 2 p λ 2 p λ 2 p λ p ,

for μ μ 1 . This completes the proof.□

4.2 Proof of Theorem 1.2

The Krasnoselskii genus theory [40] is a powerful tool to prove Theorem 1.2. Let X be a Banach space and denote by Λ the class of all closed subsets A X \ { 0 } that are symmetric with respect to the origin, i.e., u A implies u A .

Theorem 4.3

(Theorem 9.12 of [40]) Let X be an infinite dimensional Banach space and C 1 ( X ) be an even functional, with ( 0 ) = 0 . Assuming that X = Y Z , where Y is finite dimensional, and that satisfies

  1. There exist constants ρ , α > 0 such that ( u ) α for all u B ρ Z ;

  2. There exists Θ > 0 such that satisfies the ( P S ) c condition for all c , with c ( 0 , Θ ) ;

  3. For any finite dimensional subspace X ˜ X , there is R = R ( X ˜ ) > 0 such that ( u ) 0 on X ˜ \ B R .

Let Y = span { v 1 , , v k } , and for N k , inductively lt v N + 1 E N = span { v 1 , , v N } . Put R N = R ( E N ) and Ω N = B R N E N and define

G N = { ψ C ( Ω N , X ) : ψ B R N E N = id a n d ψ i s o d d }

and

Γ j = { ψ ( Ω N \ V ¯ ) : ψ G N , N j , V Λ , γ ( V ) N j } ,

where γ ( V ) is the Krasnoselskii genus of V. For j N , set

c j = inf E Γ j max u E ( u ) .

Thus, 0 c j c j + 1 and c j < Θ for j > k , then c j is a critical value of . Moreover, if c j = c j + 1 = = c j + m = c < Θ for j > k , then γ ( K c ) m + 1 , where

K c = { u X : ( u ) = c a n d ( u ) = 0 } .

Proof of Theorem 1.2

We are going to apply Theorem 4.3 to μ C 1 ( S V 1 , p ( H n ) ) , where X is the reflexive infinite dimensional real Banach space S V 1 , p ( H n ) . It follows from (3.1) that μ ( 0 ) = 0 and by ( f 2 ) that the functional μ is even. Let us divide the argument into the next four steps.

Step 1. A similar proof to verify ( a ) and ( b ) of Theorem 4.1 shows that μ satisfies ( i ) and ( i i i ) of Theorem 4.3.

Step 2. We claim that there exists a sequence ( Λ k ) k R + , with Λ k Λ k + 1 , such that

c k μ = inf E Γ k max u E μ ( u ) Λ k .

For this, applying an argument given in [48], the definition of c k μ implies that

c k μ = inf E Γ k max u E μ ( u ) inf E Γ k max u E K ( u p ) 1 2 p λ u F L , p 2 p λ .

Set

Λ k = inf E Γ k max u E K ( u p ) 1 2 p λ u F L , p 2 p λ .

Therefore, the definition of Γ k yields that Λ k < and Λ k Λ k + 1 .

Step 3. We claim that equation ( ) has at least k pairs of weak solutions.

The argument used to prove (4.2) shows that there exists μ 2 > 0 such that

c k μ Λ k < k 0 S C ( λ , Q ) p 2 p λ 2 p λ 2 p λ p ,

for all μ μ 2 .

Therefore, in any case, we have

0 < c 1 μ c 2 μ c k μ Λ k < k 0 S C ( λ , Q ) p 2 p λ 2 p λ 2 p λ p .

With the aid of the discussion of [40], we obtain that the levels c 1 μ c 2 μ c k μ are the critical values of μ .

If c j μ = c j + 1 μ for some j = 1 , 2 , , k 1 , then by Theorem 4.2 and Remark 2.12 in [2], the set K c j μ processes infinitely many distinct critical points. Hence, equation ( ) admits infinitely many (weak) solutions. In conclusion, equation ( ) has at least k pairs of solutions, as required.□

5 Degenerate case

This section is devoted to the study of equation ( ) in the degenerate case. From here on, we assume as usually conditions ( V ) and ( f 1 ) ( f 3 ) , but on K , we require only properties ( K 2 ) and ( K 3 ) , without further mentioning. We start by presenting a result, essential for the existence proofs of equation ( ).

Lemma 5.1

Let ( u k ) k S V 1 , p ( H n ) be a Palais-Smale sequence of functional μ at level c μ , that is, satisfying (3.16). If

(5.1) 0 < c μ < 1 σ 1 2 p λ k 1 S θ C ( λ , Q ) θ p 2 p λ 2 p λ 2 p λ θ p ,

then there exists a subsequence of ( u k ) k strongly convergent in S V 1 , p ( H n ) .

Proof

Let ( u k ) k S V 1 , p ( H n ) satisfy (3.16) at level c μ , verifying (5.1). Since equation ( ) is degenerate, we have to share the discussion into two cases.

Case 1: inf k 1 u k = 0 .

Here, either 0 is an accumulation point for the real sequence ( u k ) k and so there is a subsequence of ( u k ) k strongly convergent to u = 0 , or 0 is an isolated point of ( u k ) k and so there exists a subsequence, still denoted by ( u k ) k , such that inf k 1 u k > 0 . The first case is impossible, since the trivial solution cannot be a critical point at level c μ , being 0 = μ ( 0 ) = c μ > 0 . Consequently, only the latter case can take place, so that there is a subsequence, still denoted by ( u k ) k such that inf k 1 u k > 0 .

Case 2: d = inf k 1 u k > 0 .

We first prove sequence ( u k ) k is bounded in S V 1 , p ( H n ) . By (3.16), ( K 2 ) , ( K 3 ) , and ( f 3 ) , we have as k ,

(5.2) c μ + 1 + o ( 1 ) + u k = μ ( u k ) 1 σ μ ( u k ) , u k = 1 p K ( u k p ) 1 σ K ( u k p ) u k p μ H n F ( ξ , u k ) 1 σ f ( ξ , u k ) u k d ξ + 1 σ 1 2 p λ u k F L , p 2 p λ 1 θ p 1 σ K ( u k p ) u k p 1 θ p 1 σ k 1 u k θ p .

Hence, ( u k ) k is bounded in S V 1 , p ( H n ) , since θ p > 1 .

Now, as in the proof of Lemma 3.2, up to a subsequence, still denoted by ( u k ) k there exists some u S V 1 , p ( H n ) such that (3.19)–(3.21) hold.

Suppose, by contradiction, that J . We construct a smooth cut-off function φ ε , ε > 0 , a in the proof of Lemma 3.2. Hence, also ( u k φ ε ) k is bounded in S V 1 , p ( H n ) and so μ ( u k ) , u k φ ε 0 as k . Therefore, as k ,

(5.3) K ( u k p ) H n D H u k p φ ε d ξ + H n V ( ξ ) u k p φ ε d ξ = μ H n f ( ξ , u k ) u k φ ε d ξ + H n × H n u k ( η ) p λ u k ( ξ ) p λ φ ε η 1 ξ H λ d η d ξ + o k ( 1 ) .

Again, by ( f 2 ) and the Hölder inequality, we obtain again (3.24). By ( K 3 ) and the fact that d = inf k 1 u k > 0 , the left-hand side of (5.3) satisfies

(5.4) K ( u k p ) H n D H u k p φ ε d ξ + H n V ( ξ ) u k p φ ε d ξ K H n D H u k p φ ε d ξ + H n V ( ξ ) u k p φ ε d ξ H n D H u k p φ ε d ξ + H n V ( ξ ) u k p φ ε d ξ k 1 H n D H u k p φ ε d ξ + H n V ( ξ ) u k p φ ε d ξ θ .

To estimate each term at the right-hand side of (5.4), we first observe that (3.25) and (3.26) continue to hold. Consequently, substituting (3.25) and (3.26) into (5.4), we have

lim ε 0 + lim k K ( u k p ) H n D H u k p φ ε d ξ + H n V ( ξ ) u k p φ ε d ξ k 1 ω j θ .

Therefore, by (3.24) and (5.3), it follows that

ν j k 1 ω j θ .

Inserting this into (5.3), we have that either ν j = 0 or

(5.5) ν j ( k 1 S H G , p θ ) 2 p λ 2 p λ θ p .

We assume on the contrary that (5.5) holds. As discussed in the proof of Lemma 3.2, assumptions ( K 2 ) and ( f 3 ) show that

c μ 1 σ 1 2 p λ H n × H n u k ( η ) p λ u k ( ξ ) p λ φ ε η 1 ξ H λ d η d ξ .

Moreover,

c μ 1 σ 1 2 p λ ( k 1 S H G , p θ ) 2 p λ 2 p λ θ p 1 σ 1 2 p λ k 1 S θ C ( λ , Q ) θ p 2 p λ 2 p λ 2 p λ θ p .

This contradicts assumption (5.1). Therefore, ν j = 0 for any j J .

In what follows, we shall verify that ν = 0 . Choose a smooth cut-off function ϕ R as in the proof of Lemma 3.2. Since μ ( u k ) , u k ϕ R 0 as k . Thus, as k ,

(5.6) K ( u k p ) H n D H u k p ϕ R d ξ + H n V ( ξ ) u k p ϕ R d ξ = μ H n f ( ξ , u k ) u k ϕ R d ξ + H n × H n u k ( η ) p λ u k ( ξ ) p λ ϕ R η 1 ξ H λ d η d ξ + o k ( 1 ) .

As in the proof of Lemma 3.2, we obtain again (3.31). Applying in (5.6), a similar technique as in the proof of (5.4) and passing to the limit as R and k , we have

lim R lim k K ( u k p ) H n D H u k p ϕ R d ξ + H n V ( ξ ) u k p ϕ R d ξ k 1 ω θ .

Arguing as in the proof of Lemma 3.2, we obtain that

ν k 1 ω θ .

From this fact and (3.21), we have that either ν = 0 or

(5.7) ν 1 σ 1 2 p λ k 1 S θ C ( λ , Q ) θ p 2 p λ 2 p λ 2 p λ θ p .

Suppose, by contradiction, that (5.7) holds. Analogous to the proof in (3.29), it follows that

c μ 1 σ 1 2 p λ k 1 S θ C ( λ , Q ) θ p 2 p λ 2 p λ 2 p λ θ p .

This cannot occur by (5.1). Hence, ν = 0 .

In conclusion, J = and ν = 0 , so that

lim k H n × H n u k ( ξ ) p λ u k ( η ) p λ η 1 ξ H λ d η d ξ = H n × H n u ( ξ ) p λ u ( η ) p λ η 1 ξ H λ d η d ξ ,

as shown in the proof of Lemma 3.2. From the Brézis-Lieb lemma, the last equality, and (3.19), we deduce that

H n × H n u k ( ξ ) u ( ξ ) p λ u k ( η ) u ( η ) p λ η 1 ξ H λ d η d ξ 0 ,

as k . Therefore, we obtain that

H n × H n u k ( η ) p λ u k ( ξ ) p λ 2 u k ( u k u ) η 1 ξ H λ d η d ξ 0

and

H n f ( ξ , u k ) ( u k u ) d ξ 0 .

Let L be the operator given in (3.36). Therefore, by (3.37) and μ ( u k ) , u k u 0 , we obtain

lim k K ( u k p ) [ L ( u k ) , u k u L ( u ) , u k u ] = 0 .

Hence, by ( K 3 ) and the fact that d = inf k 1 u k > 0 , we obtain

lim k [ L ( u k ) , u k u L ( u ) , u k u ] = 0 .

Consequently,

lim k u k u p = lim k ( L ( u k ) , u k u L ( u ) , u k u ) = 0 .

Thus, u k u in S V 1 , p ( H n ) . This concludes the proof.□

Lemma 5.2

The functional μ satisfies the assumptions ( a ) and ( b ) of Theorem 4.1.

Proof

Fix μ > 0 . By ( K 2 ) ( K 3 ) , ( f 2 ) , and (2.8), for any u S V 1 , p ( H n ) , we obtain

μ ( u ) = 1 p K ( u p ) μ H n F ( ξ , u ) d ξ 1 2 p λ u F L , p 2 p λ 1 θ p K ( u p ) u p 2 q μ a r u p λ q 1 2 p λ u F L , p 2 p λ 1 θ p k 1 u θ p 2 q c μ u q D H u p 2 p λ 2 p λ S H G , p 2 p λ p 1 θ p k 1 u θ p 2 q c μ u q u 2 p λ 2 p λ S H G , p 2 p λ p ,

where c > 0 is a constant. Therefore, there exist ρ , α > 0 such that μ ( u ) α for u = ρ , since θ p < q and θ p < p < 2 p λ by ( f 2 ) , ( K 2 ) , and (2.6), respectively. Hence, ( a ) of Theorem 4.1 holds. Proceeding as in the proof of Lemma 4.2, we obtain that also ( b ) of Theorem 4.1 is verified.□

Proof of Theorem 1.3

Following the proof of Theorem 1.1, with the changes required in the degenerate case as shown in the proof of Lemma 5.1, we deduce that

c μ = inf γ Γ max t [ 0 , 1 ] μ ( γ ( t ) ) < 1 σ 1 2 p λ k 1 S θ C ( λ , Q ) θ p 2 p λ 2 p λ 2 p λ θ p .

The rest of the proof is as that of Theorem 1.1.□

Proof of Theorem 1.4

The proof is the same as that of Theorem 1.2.□

  1. Funding information: S. Liang was supported by the Science and Technology Development Plan Project of Jilin Province, China (No. YDZJ202201ZYTS582), the Young outstanding talents project of Scientific Innovation and Entrepreneurship in Jilin (No. 20240601048RC), the National Natural Science Foundation of China (No. 12371455), Natural Science Foundation of Changchun Normal University (No. CSJJ2023004GZR), the Research Foundation of Department of Education of Jilin Province (No. JJKH20230902KJ), and the Innovation and Entrepreneurship Talent Funding Project of Jilin Province (No. 2023QN21). Y. Song was supported by the Science and Technology Development Plan Project of Jilin Province, China (Grant No. 222614JC010793935), the National Natural Science Foundation of China (Grant No.12001061), and the Research Foundation of Department of Education of Jilin Province (Grant No. JJKH20220822KJ). P. Pucci is a member of the Gruppo Nazionale per l’Analisi Matematica, la Probabilità e le loro Applicazioni (GNAMPA) of the Istituto Nazionale di Alta Matematica (INdAM), and this research is under the auspices of INdAM.

  2. Author contributions: All authors contributed equally to the writing of this article. All authors read and approved the final manuscript.

  3. Conflict of interest: The authors declare that they have no competing interests.

References

[1] C. O. Alves, G. M. Figueiredo, and M. Yang, Multiple semiclassical solutions for a nonlinear Choquard equation with magnetic field, Asymptot. Anal. 96 (2016), no. 2, 135–159. 10.3233/ASY-151337Search in Google Scholar

[2] A. Ambrosetti and P. Rabinowitz, Dual variational methods in critical point theory and applications, J. Funct. Anal. 14 (1973), no. 4, 349–381. 10.1016/0022-1236(73)90051-7Search in Google Scholar

[3] J. Aubin and I. Ekeland, Applied Nonlinear Analysis, Wiley, New York, 1984. Search in Google Scholar

[4] A. Bongers, Existenzaussagen für die Choquard-Gleichung: Ein nichtlineares Eigenwertproblem der Plasma-Physik, Z. Angew. Math. Mech. 60 (1980), no. 7, T240–T242. Search in Google Scholar

[5] S. Bordoni and P. Pucci, Schrödinger-Hardy systems involving two Laplacian operators in the Heisenberg group, Bull. Sci. Math. 146 (2018), 50–88. 10.1016/j.bulsci.2018.03.001Search in Google Scholar

[6] M. Caponi and P. Pucci, Existence theorems for entire solutions of stationary Kirchhoff fractional p-Laplacian equations, Ann. Mat. Pura Appl. 195 (2016), no. 6, 2099–2129. 10.1007/s10231-016-0555-xSearch in Google Scholar

[7] G. Folland, Subelliptic estimates and function spaces on nilpotent Lie groups, Ark. Mat. 13 (1975), no. 2, 161–207. 10.1007/BF02386204Search in Google Scholar

[8] G. Folland and E. M. Stein, Estimates for the ∂¯b complex and analysis on the Heisenberg group, Commun. Pure Appl. Math. 27 (1974), 429–522. 10.1002/cpa.3160270403Search in Google Scholar

[9] G. Folland and E. Stein, Hardy Spaces on Homogeneous Groups, Princeton University Press, Princeton, NJ, 1982. 10.1515/9780691222455Search in Google Scholar

[10] R. Frank and E. Lieb, Sharp constants in several inequalities on the Heisenberg group, Ann. Math. 176 (2012), no. 1, 349–381. 10.4007/annals.2012.176.1.6Search in Google Scholar

[11] N. Garofalo and E. Lanconelli, Frequency functions on the Heisenberg group, the uncertainty principle and unique continuation, Ann. Inst. Fourier. 40 (1990), no. 2, 313–356. 10.5802/aif.1215Search in Google Scholar

[12] D. Goel and K. Sreenadh, Existence and nonexistence results for Kohn Laplacian with Hardy-Littlewood-Sobolev critical exponents, J. Math. Anal. Appl. 486 (2020), no. 2, Paper no. 123915, 29 pp. 10.1016/j.jmaa.2020.123915Search in Google Scholar

[13] L. Hörmander, Hypoelliptic second order differential equations, Acta Math. 119 (1967), 147–171. 10.1007/BF02392081Search in Google Scholar

[14] Y. Hu, Reversed Hardy-Littlewood-Sobolev inequalities with weights on the Heisenberg group, Adv. Nonlinear Anal. 12 (2023), no. 1, Paper no. 20230116, 25 pp. 10.1515/anona-2023-0116Search in Google Scholar

[15] S. Ivanov and D. Vassilev, Extremals for the Sobolev Inequality and the Quaternionic Contact Yamabe Problem, World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2011. 10.1142/9789814295710Search in Google Scholar

[16] D. Jerison and J. M. Lee, Intrinsic CR normal coordinates and the CR Yamabe problem, J. Differential Geom. 29 (1989), no. 2, 303–343. 10.4310/jdg/1214442877Search in Google Scholar

[17] J. Lan, X. He, and Y. Meng, Normalized solutions for a critical fractional Choquard equation with a nonlocal perturbation, Adv. Nonlinear Anal. 12 (2023), no. 1, Paper No. 20230112, 40pp. 10.1515/anona-2023-0112Search in Google Scholar

[18] G. P. Leonardi and S. Masnou, On the isoperimetric problem in the Heisenberg group Hn, Ann. Mat. Pura Appl. 184 (2005), no. 4, 533–553. 10.1007/s10231-004-0127-3Search in Google Scholar

[19] S. Liang and V. Rădulescu, Existence of infinitely many solutions for degenerate Kirchhoff-type Schrödinger-Choquard equations, Electron. J. Differ. Equations 2017 (2017), no. 230, 1–17.Search in Google Scholar

[20] E. Lieb, Existence and uniqueness of the minimizing solution of Choquardas nonlinear equation, Studies Appl. Math. 57 (1976/77), no. 2, 93–105. 10.1002/sapm197757293Search in Google Scholar

[21] 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

[22] P. L. Lions, The concentration-compactness principle in the calculus of variations, The limit case, Part 1, Rev. Mat. Iberoamericana 1 (1985), no. 1, 145–201. 10.4171/rmi/6Search in Google Scholar

[23] P. L. Lions, The concentration-compactness principle in the calculus of variations, The limit case, Part 2, Rev. Mat. Iberoamericana 1 (1985), no. 2, 45–121. 10.4171/rmi/12Search in Google Scholar

[24] Z. Liu, L. Tao, D. Zhang, S. Liang, and Y. Song, Critical nonlocal Schrödinger-Poisson system on the Heisenberg group, Adv. Nonlinear Anal. 11 (2022), no. 1, 482–502. 10.1515/anona-2021-0203Search in Google Scholar

[25] Z. Liu and D. Zhang, A new Kirchhoff-Schrödinger-Poisson type system on the Heisenberg group, Differ. Integral Equ. 34 (2021), no. 11–12, 621–639. 10.57262/die034-1112-621Search in Google Scholar

[26] A. Loiudice, Improved Sobolev inequalities on the Heisenberg group, Nonlinear Anal. 62 (2005), no. 5, 953–962. 10.1016/j.na.2005.04.015Search in Google Scholar

[27] 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

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

[29] R. Penrose, Quantum computation, entanglement and state reduction, R. Soc. Lond. Philos. Trans. Ser. A Math. Phys. Eng. Sci. 356 (1998), no. 1743, 1927–1939. 10.1098/rsta.1998.0256Search in Google Scholar

[30] P. Pucci, Critical Schrödinger-Hardy systems in the Heisenberg group, Discrete Contin. Dyn. Syst. Ser. S 12 (2019), no. 2, 375–400. 10.3934/dcdss.2019025Search in Google Scholar

[31] P. Pucci, Existence and multiplicity results for quasilinear elliptic equations in the Heisenberg group, Opuscula Math. 39 (2019), no. 2, 247–257. 10.7494/OpMath.2019.39.2.247Search in Google Scholar

[32] P. Pucci and L. Temperini, (p,Q) systems with critical singular exponential nonlinearities in the Heisenberg group, Open Math. 18 (2020), no. 1, 1423–1439. 10.1515/math-2020-0108Search in Google Scholar

[33] P. Pucci and L. Temperini, Concentration-compactness results for systems in the Heisenberg group, Opuscula Math. 40 (2020), no. 1, 151–163. 10.7494/OpMath.2020.40.1.151Search in Google Scholar

[34] P. Pucci and L. Temperini, Entire solutions for some critical equations in the Heisenberg group, Opuscula Math. 42 (2022), no. 2, 279–303. 10.7494/OpMath.2022.42.2.279Search in Google Scholar

[35] P. Pucci and L. Temperini, Critical equations with Hardy terms in the Heisenberg group, Rend. Circ. Mat. Palermo, II. Ser. 71 (2022), no. 3, 1049–1077. 10.1007/s12215-022-00741-ySearch in Google Scholar

[36] P. Pucci and L. Temperini, On the concentration-compactness principle for Folland-Stein spaces and for fractional horizontal Sobolev spaces, Math. Eng. 5 (2023), no. 1, Paper No. 007, 21 pp. 10.3934/mine.2023007Search in Google Scholar

[37] P. Pucci, M. Xiang, and B. Zhang, Multiple solutions for nonhomogeneous Schrödinger-Kirchhoff type equations involving the fractional p-Laplacian in RN, Calc. Var. Partial Differential Equations 54 (2015), no. 3, 2785–2806. 10.1007/s00526-015-0883-5Search in Google Scholar

[38] P. Pucci, M. Xiang, and B. Zhang, Existence and multiplicity of entire solutions for fractional p-Kirchhoff equations, Adv. Nonlinear Anal. 5 (2016), no. 1, 27–55. 10.1515/anona-2015-0102Search in Google Scholar

[39] P. Pucci, M. Xiang, and B. Zhang, Existence results for Schrödinger-Choquard-Kirchhoff equations involving the fractional p-Laplacian, Adv. Calc. Var. 12 (2017), no. 3, 253–275. 10.1515/acv-2016-0049Search in Google Scholar

[40] P. Rabinowitz, Minimax methods in critical point theory with applications to differential equations, in: CBME Regional Conference Series in Mathematics, vol. 65, American Mathematical Society, Providence, RI, 1986. 10.1090/cbms/065Search in Google Scholar

[41] Y. Song and S. Shi, On a degenerate p-fractional Kirchhoff equations involving critical Sobolev-Hardy nonlinearities, Mediterr. J. Math. 15 (2018), no. 1, Paper No. 17, 18 pp. 10.1007/s00009-017-1062-zSearch in Google Scholar

[42] X. Sun, S. Bai, and Y. Song, On the noncooperative Schrödinger-Kirchhoff system involving the critical nonlinearities on the Heisenberg group, Bound Value Probl. 2022 (2022), Paper No. 75, 19 pp. 10.1186/s13661-022-01657-3Search in Google Scholar

[43] X. Sun, Y. Song, and S. Liang, On the critical Choquard-Kirchhoff problem on the Heisenberg group, Adv. Nonlinear Anal. 12 (2023), no. 1, 210–236. 10.1515/anona-2022-0270Search in Google Scholar

[44] X. Sun, Y. Song, S. Liang, and B. Zhang, Critical Kirchhoff equations involving the p-sub-Laplacians operators on the Heisenberg group, Bull. Math. Sci. 13 (2023), no. 2, Paper No. 2250006, 26 pp. 10.1142/S1664360722500060Search in Google Scholar

[45] X. Sun, B. Yang, and Y. Song, Multiplicity of solutions for the noncooperative Choquard-Kirchhoff system involving Hardy-Littlewood-Sobolev critical exponent on the Heisenberg group, Rend. Circ. Mat. Palermo (2) 72 (2023), no. 7, 3439–3457. 10.1007/s12215-022-00833-9Search in Google Scholar

[46] D. Vassilev, Existence of solutions and regularity near the characteristic boundary for sub-Laplacian equations on Carnot groups, Pacific J. Math. 227 (2006), no. 2, 361–397. 10.2140/pjm.2006.227.361Search in Google Scholar

[47] L. Wang, T. Han, and J. Wang, Infinitely many solutions for Schrödinger-Choquard-Kirchhoff equations involving the fractional p-Laplacian, Acta. Math. Sin. - English Ser. 37 (2021), no. 2, 315–332. 10.1007/s10114-021-0125-zSearch in Google Scholar

[48] W. Wei and X. Wu, A multiplicity result for quasilinear elliptic equations involving critical Sobolev exponents, Nonlinear Anal. 18 (1992), no. 6, 559–567. 10.1016/0362-546X(92)90210-6Search in Google Scholar

[49] M. Xiang, B. Zhang, and V. D. Rădulescu, Superlinear Schrödinger-Kirchhoff type problems involving the fractional p-Laplacian and critical exponent, Adv. Nonlinear Anal. 9 (2020), no. 1, 690–709. 10.1515/anona-2020-0021Search in Google Scholar

[50] H. Ye, The existence of least energy nodal solutions for some class of Kirchhoff equations and Choquard equations in RN, J. Math. Anal. Appl. 431 (2015), no. 2, 935–954. 10.1016/j.jmaa.2015.06.012Search in Google Scholar

Received: 2024-02-01
Revised: 2024-06-06
Accepted: 2024-07-26
Published Online: 2024-08-29

© 2024 the author(s), published by De Gruyter

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

Downloaded on 5.12.2025 from https://www.degruyterbrill.com/document/doi/10.1515/agms-2024-0006/html
Scroll to top button