Skip to content
BY 4.0 license Open Access Published by De Gruyter January 22, 2019

Reverse Stein–Weiss Inequalities on the Upper Half Space and the Existence of Their Extremals

  • Lu Chen , Guozhen Lu and Chunxia Tao EMAIL logo

Abstract

The purpose of this paper is four-fold. First, we employ the reverse weighted Hardy inequality in the form of high dimensions to establish the following reverse Stein–Weiss inequality on the upper half space:

+ n + n | x | α | x - y | λ f ( x ) g ( y ) | y | β d y d x C n , α , β , p , q f L q ( + n ) g L p ( + n )

for any nonnegative functions fLq(+n), gLp(+n), and p,q(0,1), β<1-np or α<-nq, λ>0 satisfying

n - 1 n 1 p + 1 q - α + β + λ - 1 n = 2 .

Second, we show that the best constant of the above inequality can be attained. Third, for a weighted system analogous to the Euler–Lagrange equations of the reverse Stein–Weiss inequality, we obtain the necessary conditions of existence for any positive solutions using the Pohozaev identity. Finally, in view of the stereographic projection, we give a spherical form of the Stein–Weiss inequality and reverse Stein–Weiss inequality on the upper half space +n.

1 Introduction

The classical Stein–Weiss inequality, established by Stein and Weiss [41] in the 1950s, states that

(1.1) n n | x | - α | x - y | - λ f ( x ) g ( y ) | y | - β d x d y C n , α , β , p , q f L q ( n ) g L p ( n ) ,

where p, q, α, β and λ satisfy the following conditions:

1 q + 1 p + α + β + λ n = 2 , 1 q + 1 p 1 ,
α + β 0 , α < n q , β < n p , 0 < λ < n .

The smallest Cn,α,β,p,q for (1.1) to hold is often referred as the best constant of the Stein–Weiss inequality. Throughout this paper, we always let p, q denote the conjugate of p, q, respectively. We note that an alternative proof of establishing the Stein–Weiss inequalities has been recently found in [25] by using conditions on weights to guarantee the weighted boundedness of fractional integrals given in [37]. Lieb [29] used the method based on the symmetrization argument and the Riesz rearrangement to establish the existence of extremals for inequality (1.1) in the case p<q and α,β0. Furthermore, in the case of p=q, the extremals cannot be expected to exist (see Lieb [29] and also Herbst [28] for the case λ=n-1, p=q=2, α=0, β=1). In the case of p=q, Beckner [5, 4] obtained the sharp constant of the Stein–Weiss inequalities (1.1) by establishing a Stein–Weiss lemma with a general kernel. The precise estimate of the sharp constant of the Stein–Weiss inequalities for the case of pq was also established in [4]. For more results about proving precise estimates for Stein–Weiss functionals in conjunction with the study of Pitt’s type inequalities and their multilinear versions, we refer the reader to the works of Beckner [2, 3, 1, 6, 7]. We note that the existence of extremal functions for the Stein–Weiss inequalities in the case p<q, under the assumption α+β0, has been established by Chen, Lu and Tao [13], which extends Lieb’s result under the stronger assumption that α0 and β0, using the concentration-compactness principle of Lions [31, 32].

In the special case of α=β=0, the Stein–Weiss inequality (1.1) becomes the Hardy–Littlewood–Sobolev (HLS) inequality, which can be stated as follows:

(1.2) n n | x - y | - λ f ( x ) g ( y ) d x d y C n , p , q f L q ( n ) g L p ( n ) ,

with 1<q, p<, 0<λ<n and 1q+1p+λn=2. The above inequality was established by Hardy and Littlewood [27] for n=1 and Sobolev [38] for n2 (see also [39, 40]). Lieb and Loss [30] applied the layer cake representation formula to give an explicit upper bound for the sharp constant Cn,p,q. More precisely, he showed that the best constant Cn,p,q satisfies the following estimate:

C n , p , q n n - λ ( π λ 2 Γ ( 1 + n 2 ) ) λ n 1 q p ( ( λ q n ( q - 1 ) ) λ n + ( λ p n ( p - 1 ) ) λ n ) .

In the diagonal case q=p=2n2n-λ, Lieb [29] classified the positive solutions of the Euler–Lagrange equations related to the Hardy–Littlewood–Sobolev inequality and obtained the best constant

C n , p , q = π λ 2 Γ ( n 2 - λ 2 ) Γ ( n - λ 2 ) ( Γ ( n ) Γ ( n 2 ) ) 1 - λ n .

Furthermore, Lieb also established the existence of extremals of (1.2) for general index p and q through the rearrangement inequality. Recently, Frank and Lieb [21, 22, 23], and Carlen, Carrillo and Loss [8] developed a rearrangement-free argument to obtain the sharp Hardy–Littlewood–Sobolev inequality and extremal functions (1.2). In particular, in [23], Frank and Lieb established a remarkable sharp Hardy–Littlewood–Sobolev inequality on the Heisenberg group which extends the celebrated work of Jerison and Lee on the sharp Sobolev inequality on the Heisenberg group. The Stein–Weiss inequality on the Heisenberg group was also established in [25] by using a condition on weighted inequalities for integral operators in [37]. Han [24] proved the existence of extremals of the Hardy–Littlewood–Sobolev inequality on the Heisenberg group for all the case of q and p through the concentration-compactness principle of Lions [31, 32]. More recently, the existence of the extremal functions of the Stein–Weiss inequalities proved in [25] have also been established by Chen, Lu and Tao [13]. The radial symmetry of positive solutions for the Hardy–Sobolev equation and system were studied in [10, 16, 15, 34]. For more results about the Stein–Weiss inequality, see [3, 5] and the references therein.

We mention that we recently have established the following Hardy–Littlewood–Sobolev and Stein–Weiss inequalities with the fractional Poisson kernel in [12, 14]:

(1.3) + n + n | ξ | - α f ( ξ ) P ( x , ξ , γ ) g ( x ) | x | - β d ξ d x C n , α , β , p , q g L q ( + n ) f L p ( + n ) ,

where

P ( x , ξ , γ ) = x n ( | x - ξ | 2 + x n 2 ) n + 2 - γ 2 , 2 γ < n ,

f L p ( + n ) , gLq(+n), and p,q(1,) satisfy

n - 1 n 1 p + 1 q + α + β + 2 - γ n = 1 .

We also proved that there exist extremals for the Stein–Weiss inequality (1.3), and these extremals must be radially decreasing about the origin. We further provided the regularity and asymptotic estimates of positive solutions to the integral systems which are the Euler–Lagrange equations of the extremals to the Stein–Weiss inequality (1.3) with the fractional Poisson kernel. These results are inspired by the work of Hang, Wang and Yan [26], where the Hardy–Littlewood–Sobolev type inequality was first established when γ=2 and α=β=0. The proof of the Stein–Weiss inequality (1.3) with the fractional Poisson kernel relies on our work on the Hardy–Littlewood–Sobolev inequality with the fractional Poisson kernel [14].

We now return to the discussion on the reverse Stein–Weiss inequalities. Chen, Liu, Lu and Tao [11] established the following reverse Stein–Weiss inequality which states that for n1, p,q(0,1), λ>0, 0α<-nq and 0β<-np satisfying

1 p + 1 q - α + β + λ n = 2 ,

there exists a constant Cn,α,β,p,q>0 such that for any nonnegative functions fLq(n), gLp(n), there holds

(1.4) n n | x | α | x - y | λ f ( x ) g ( y ) | y | β d x d y C n , α , β , p , q f L q ( n ) g L p ( n ) .

Chen, Liu, Lu and Tao [11] used Fourier analysis techniques and the fractional integral operators acting on radial functions to establish the above inequality (1.4). They also obtained the extremal functions for all the indices by using symmetry and a rearrangement argument. The work of [11] was motivated by the sharp reverse Hardy–Littlewood–Sobolev inequality by Beckner [7, 9], and by Dou and Zhu [18]. The result of Beckner in [7] was used by Carneiro [9] to establish the sharp inequality for the Strichartz norm. In the case q=p=2n2n+λ, Beckner [7], and Dou and Zhu [18] established the sharp constants and extremal functions for the reverse Hardy–Littlewood–Sobolev inequality. Subsequently, Ngô and Nguyen [35] verified that extremal functions still exist for general p and q through adapting similar arguments to Lieb in [29]. We note that the range of the exponents in the reverse Stein–Weiss inequality (1.4) are quite different from those in the Stein–Weiss inequality (1.1). They are mainly reflected in the difference that the power of the kernel |x-y| is negative in (1.1) and positive in (1.4), and p,q>1 in (1.1) and p,q<1 in (1.4).

Recently, Dou [17] established the following double weighted Hardy–Littlewood–Sobolev inequality on the upper half space +n, which can be seen as an extension of the Stein–Weiss inequality (1.1):

(1.5) + n + n | x | - α | x - y | - λ f ( x ) g ( y ) | y | - β d y d x C n , α , β , p , q f L q ( + n ) g L p ( + n ) ,

with p, q, α, β and λ satisfying the following conditions:

1 q + n - 1 n p + α + β + λ + 1 n = 2 , 1 q + 1 p 1 ,
α + β 0 , α < n q , β < n - 1 p , 0 < λ < n - 1 .

In the case α, β>0, he applied the rearrangement inequality to obtain the existence of extremals for inequality (1.5) when p<q. In the special case of α=β=0, the Stein–Weiss inequality (1.5) on the upper half space +n becomes the following Hardy–Littlewood–Sobolev inequality on the upper half space:

+ n + n | x - y | - λ f ( x ) g ( y ) d y d x C n , α , β , p , q f L q ( + n ) g L p ( + n ) ,

where 1<q, p<, 0<λ<n-1 and 1q+n-1n1p+λ+1n=2. In the case of q=2n2n-λ and p=2n-22n-2-λ, Dou and Zhu [19] applied the moving spheres in integral form to classify the positive solutions of the Euler–Lagrange equations of the extremals for the Hardy–Littlewood–Sobolev inequality on +n and obtained the sharp constant. Ngô and Nguyen [36] studied the sharp reverse Hardy–Littlewood–Sobolev inequality on +n. More recently, the reverse Hardy–Littlewood–Sobolev and Stein–Weiss inequalities on the Heisenberg group have been established in [33].

Motivated by the aforementioned works [4, 11, 17, 19, 18, 35, 36], we will show in this paper that there exist a reverse version of the Stein–Weiss inequality on the upper half space +n. Furthermore, we prove that such an inequality has an extremal function for general indices. To answer these questions, we consider the following reverse Stein–Weiss inequality on the upper half space +n. Our first main result is the following.

Theorem 1.1.

For n>1, 0<p,q<1, β<1-np or α<-nq, λ>0 satisfying

n - 1 n p + 1 q - α + β + λ - 1 n = 2 ,

there exists some constant Cn,α,β,p,q>0 such that for any nonnegative functions fLq(R+n), gLp(R+n), there holds

(1.6) + n + n | x | α | x - y | λ f ( x ) g ( y ) | y | β d y d x C n , α , β , p , q f L q ( + n ) g L p ( + n ) .

Next, we will study whether the sharp constant Cn,α,β,p,q for the reverse Stein–Weiss inequality on the upper half space can be attained. Namely, we will consider whether there exist extremal functions for inequality (1.6). To this end, we can consider the following minimizing problem:

(1.7) C n , α , β , p , q := inf { V λ ( g ) L q ( + n ) : g 0 , g L p ( + n ) = 1 } ,

where the double weighted operator Vλ(g)(x) is given by

V λ ( g ) ( x ) = + n | x | α | x - y | λ g ( y ) | y | β d y .

It is easy to verify that the extremals for (1.6) are those solving the minimizing problem (1.7). Then we can prove that the constant Cn,α,β,p,q could actually be achieved.

Theorem 1.2.

For n>1, p,q(0,1), λ>0, 0α<-n-1q and 0β<1-np satisfying

n - 1 n p + 1 q - α + β + λ - 1 n = 2 ,

there exists some nonnegative function gLp(R+n) satisfying gLp(R+n)=1 and Vλ(g)Lq(R+n)=Cn,α,β,p,q.

Once we have established the existence of extremals of inequality (1.6), we can then consider the corresponding Euler–Lagrange system. First of all, one can minimize the functional

J ( f , g ) = + n + n | x | α | x - y | λ f ( x ) g ( y ) | y | β d y d x

under the constraint fLq(+n)=gLp(+n)=1. By the Euler–Lagrange multiplier theorem, we can derive the following integral system:

(1.8) { J ( f , g ) f ( x ) q - 1 = + n | x | α | x - y | λ g ( y ) | y | β d y , x + n , J ( f , g ) g ( y ) p - 1 = + n | y | β | x - y | λ f ( x ) | x | α d x , y + n .

Let u=c1fq-1, v=c2gp-1, 1q-1=-p1 and 1p-1=-p2, and pick two suitable constants c1 and c2. Then system (1.8) is simplified as

(1.9) { u ( x ) = + n | x | α | x - y | λ v - p 2 ( y ) | y | β d y , x + n , v ( y ) = + n | y | β | x - y | λ u - p 1 ( x ) | x | α d x , y + n ,

where 1p1-1+1p2-1n-1n=α+β+λn.

Next, we give some asymptotic estimates of positive solutions to system (1.9) around the origin and near infinity.

Theorem 1.3.

Suppose that (u,v) is a pair of positive Lebesgue measurable solutions of (1.9). Then u(x) and v(y) must satisfy the following asymptotic estimate around the origin and near infinity:

lim | x | u ( x ) | x | λ + α = + n v - p 2 ( y ) | y | β 𝑑 y , lim | y | v ( y ) | y | λ + β = + n u - p 1 ( x ) | x | α 𝑑 x ,
lim | x | 0 u ( x ) | x | α = + n v - p 2 ( y ) | y | β + λ 𝑑 y , lim | y | 0 v ( y ) | y | β = + n u - p 1 ( x ) | x | α + λ 𝑑 y .

Finally, we are interested in studying the following weighted integral system:

(1.10) { u ( x ) = n | x - y | λ | y | ν 2 v - p 2 ( y ) d y , x + n , v ( y ) = n | x - y | λ | x | ν 1 u - p 1 ( x ) d x , y + n .

With the help of the Pohozaev identity, we establish some necessary conditions for the existence of positive solutions to the weighted integral system (1.10).

Theorem 1.4.

For λ, ν1, ν2, p1, p2>0, suppose that there exists a pair of positive solutions (u,v)C1(R+n)×C1(R+n) satisfying (1.10). Then the following balance condition must hold:

n + ν 1 p 1 - 1 + n - 1 + ν 2 p 2 - 1 = λ .

An immediate corollary is that we can obtain the following Liouville theorem to the above integral system (1.10).

Corollary 1.5.

For λ, ν1, ν2, p1, p2>0, suppose that

n + ν 1 p 1 - 1 + n - 1 + ν 2 p 2 - 1 λ .

Then there does not exist a pair of positive solutions (u,v)C1(R+n)×C1(R+n) satisfying (1.10).

In the special case ν1=ν2=0, system (1.10) becomes

(1.11) { u ( x ) = + n | x - y | λ v - p 2 ( y ) d y , x + n , v ( y ) = + n | x - y | λ u - p 1 ( x ) d x , y + n .

Clearly, system (1.11) corresponds to the Euler–Lagrange system of the extremal functions of the reverse Hardy–Littlewood–Sobolev inequality on the upper half space +n:

+ n + n | x - y | λ f ( x ) g ( y ) d y d x C n , p , q f L q ( + n ) g L p ( + n )

for any nonnegative functions fLq(n), gLp(n) and p,q(0,1), λ>0 such that n-1n1p+1q-λ-1n=2, with 1q-1=-p1 and 1p-1=-p2. Therefore, we can deduce the following sufficient and necessary condition from Theorem 1.2 and Theorem 1.4.

Theorem 1.6.

For λ, p1, p2>0, the sufficient and necessary condition for the existence of a pair of positive solutions (u,v)C1(R+n)×C1(R+n) to system (1.11) is

n p 1 - 1 + n - 1 p 2 - 1 = λ .

As an application of Theorem 1.1, we also establish a spherical form of the reverse Stein–Weiss inequality (1.6) on the upper half space in the case of q=2n2n+λ+α+β and p=2n-22n-2+λ+α+β.

Theorem 1.7.

For n1, λ>0, β<-n-1p, q=2n2n+λ+α+β, p=2n-22n-2+λ+α+β, there exists a constant Cn,α,β,p,q>0 such that for any nonnegative functions fLq(R+n), gLp(R+n), there holds

𝕊 + n 𝕊 0 n | ξ - S ( 0 ) | α | ξ - η | λ F 1 ( ξ ) G 1 ( η ) | η - S ( 0 ) | β d η d ξ C n , α , β , p , q F L q ( 𝕊 + n ) G L p ( 𝕊 0 n ) ,

where

F ( ξ ) = ( 2 1 + | x | 2 ) - 2 n + λ + α + β 2 f ( x ) , F 1 ( ξ ) = ( 2 1 + | x | 2 ) - 2 n + λ + α + β 2 ( 1 1 + | x | 2 ) β 2 f ( x ) ,
G ( η ) = ( 2 1 + | y | 2 ) - 2 n - 2 + λ + α + β 2 g ( y ) , G 1 ( η ) = ( 2 1 + | y | 2 ) - 2 n - 2 + λ + α + β 2 ( 1 1 + | y | 2 ) α 2 g ( y ) ,

| ξ - η | is the chordal distance from ξ to η in Rn+1 and S is the stereographic projection RnSn{(0,0,,-1)}, S+n={(ξ1,,ξn+1Sn):ξn+1>0} and S0n={(ξ1,,ξn+1Sn):ξn+1=0}.

Remark 1.8.

Under the same assumptions of Theorem 1.6, if we furthermore assume that α, β satisfy 0<α<-n-1q and 0<β<-n-1p, then we can deduce from Theorem 1.2 that the constant Cn,α,β,p,q can be attained.

In view of the Stein–Weiss inequality (1.5) and stereographic projection, we can also obtain the following Stein–Weiss inequality on the upper sphere 𝕊+n.

Theorem 1.9.

For n1, 0<λ<n-1, α<nq, β<np, α+β0, q=2n2n-λ-α-β, p=2n-22n-2-λ-α-β, there exists a constant Cn,α,β,p,q>0 such that for any functions fLq(R+n), gLp(R+n), there holds

𝕊 + n 𝕊 0 n | ξ - S ( 0 ) | - α | ξ - η | - λ H 1 ( ξ ) T 1 ( η ) | η - S ( 0 ) | - β d η d ξ C n , α , β , p , q H L q ( 𝕊 + n ) T L p ( 𝕊 0 n ) ,

where

H ( ξ ) = ( 2 1 + | x | 2 ) - 2 n - λ - α - β 2 f ( x ) , H 1 ( ξ ) = ( 2 1 + | x | 2 ) - 2 n - λ - α - β 2 ( 1 1 + | x | 2 ) - β 2 f ( x ) ,
T ( η ) = ( 2 1 + | y | 2 ) - 2 n - 2 - λ - α - β 2 g ( y ) , T 1 ( η ) = ( 2 1 + | y | 2 ) - 2 n - 2 - λ - α - β 2 ( 1 1 + | y | 2 ) - α 2 g ( y ) .

This paper is organized as follows. In Section 2, we employ the reverse weighted Hardy inequality to establish the reverse Stein–Weiss inequality on the upper half space and prove Theorem 1.1. In Section 3, we obtain the existence results for extremals of the inequality (1.6) with the help of the rearrangement inequality and prove Theorem 1.2. In Section 4, we investigate some asymptotic estimates of positive solutions to the corresponding Euler–Lagrange system around the origin and near infinity and establish Theorem 1.3. In Section 5, we use the Pohozaev identity in integral form to obtain necessary conditions for the existence of any positive solutions of system (1.10) and prove Theorem 1.4. The reverse Stein–Weiss inequality and the Stein–Weiss inequality on the upper sphere 𝕊+n are considered in Sections 6 and 7 through the stereographic projection. Theorems 1.7 and 1.9 are proved in these two sections. We mention that the reverse Stein–Weiss inequality on the sphere was already established in [11].

2 The Proof of Theorem 1.1

Throughout this section, we shall establish the reverse Stein–Weiss inequality on the upper half space. We need the following reverse weighted Hardy inequality on the upper half space. Our proof of the reverse weighted Hardy inequality on the upper half space is similar to that of the weighted Hardy inequality on whole space proved by Drábek, Heing and Kufner [20]. However, it needs a more careful analysis on the upper half space. In order to state the reverse Stein–Weiss inequality on the upper half space, we give the following notations. Define

B R ( x ) = { y n : | y - x | < R , x n } ,
B R n - 1 ( x ) = { y + n : | y - x | < R , x + n } ,
B R + ( x ) = { y = ( y 1 , y 2 , , y n ) B R ( x ) : y n > 0 , x + n } .

For x=0, we write BR=BR(0), BRn-1=BRn-1(0), BR+=BR+(0).

Lemma 2.1.

Let W(x) and U(y) be nonnegative locally integrable functions defined on R+n and R+n, respectively. For 0<p<1, q<0 and g0 on R+n, there holds

(2.1) ( + n W ( x ) ( B | x | n - 1 g ( y ) 𝑑 y ) q 𝑑 x ) 1 q C 0 ( p , q ) ( + n g p ( y ) U ( y ) 𝑑 y ) 1 p

if

(2.2) 0 < A 0 = inf R > 0 { ( | x | R W ( x ) 𝑑 x ) 1 q ( | y | R U 1 - p ( y ) 𝑑 y ) 1 p } < + ,

and

( + n W ( x ) ( + n B 2 | x | n - 1 g ( y ) 𝑑 y ) q 𝑑 x ) 1 q C 1 ( p , q ) ( + n g p ( v ) U ( v ) 𝑑 v ) 1 p

if

0 < A 1 = inf R > 0 { ( | x | R W ( x ) 𝑑 x ) 1 q ( | y | R U 1 - p ( y ) 𝑑 y ) 1 p } < + .

Proof.

Given any R>0, and x+n, y+n, through polar coordinates x=tξ and y=rη, with ξB1+, ηSn-2, the left-hand side of (2.1) and the infimum in (2.2) can be written as

+ n W ( x ) ( B | x | n - 1 g ( y ) 𝑑 y ) q 𝑑 x = 0 B 1 + t n - 1 W ( t ξ ) ( 0 t S n - 2 r n - 2 g ( r η ) 𝑑 η 𝑑 r ) q 𝑑 ξ 𝑑 t

and

(2.3) ( R t n - 1 B 1 + W ( t ξ ) 𝑑 ξ 𝑑 t ) 1 q ( 0 R r n - 2 S n - 2 U 1 - p ( r η ) 𝑑 η 𝑑 r ) 1 p A 0 ,

respectively. Define

h ( t ) = ( S n - 2 0 t r n - 2 U 1 - p ( r η ) 𝑑 r 𝑑 η ) 1 p p , H n ( t ) = S n - 2 0 t r n - 2 ( U 1 p ( r η ) h ( r ) ) - p 𝑑 r 𝑑 η ,
F n ( r ) = S n - 2 r n - 2 ( g ( r η ) U 1 p ( r η ) h ( r ) ) p 𝑑 η , W n ( t ) = B 1 + t n - 1 W ( t ξ ) 𝑑 ξ .

This, together with the reverse Hölder inequality, yields that

0 t S n - 2 r n - 2 g ( r η ) 𝑑 η 𝑑 r = S n - 2 0 t r n - 2 p g ( r η ) U 1 p ( r η ) h ( r ) r n - 2 p ( U 1 p ( r η ) h ( r ) ) - 1 𝑑 r 𝑑 η
( S n - 2 0 t r n - 2 ( g ( r η ) U 1 p ( r η ) h ( r ) ) p 𝑑 r 𝑑 η ) 1 p ( S n - 2 0 t r n - 2 ( U 1 p ( r η ) h ( r ) ) - p 𝑑 r 𝑑 η ) 1 p
(2.4) = ( 0 t F n ( r ) 𝑑 r ) 1 p × ( H n ( t ) ) 1 p .

With the help of (2.4) and the reverse Minkowski inequality, we derive that

+ n W ( x ) ( B | x | n - 1 g ( y ) 𝑑 y ) q 𝑑 x 0 W n ( t ) ( 0 t F n ( r ) 𝑑 r ) q p ( H n ( t ) ) q p 𝑑 t
= 0 ( 0 ( W n ( t ) ( H n ( t ) ) q p ) p q F n ( r ) χ ( r < t ) d r ) q p d t
(2.5) ( 0 F n ( r ) ( r W n ( t ) ( H n ( t ) ) q p 𝑑 t ) p q 𝑑 r ) q p ,

where χ denotes the characteristic function. In view of (2.3) and Un(r)=Sn-2rn-2U1-p(rη)𝑑η, it follows that

H n ( t ) = S n - 2 0 t r n - 2 ( U 1 p ( r η ) h ( r ) ) - p 𝑑 r 𝑑 η
= S n - 2 0 t r n - 2 U - p p ( r η ) ( S n - 2 0 r ρ n - 2 U 1 - p ( ρ ζ ) 𝑑 ρ 𝑑 ζ ) - 1 p 𝑑 η 𝑑 r
= 0 t U n ( r ) ( 0 r U n ( ρ ) 𝑑 ρ ) - 1 p 𝑑 r = p 0 t d d r ( 0 r U n ( ρ ) 𝑑 ρ ) 1 p 𝑑 r
= p ( 0 t U n ( r ) 𝑑 r ) 1 p = p ( 0 t S n - 2 r n - 2 U 1 - p ( r η ) 𝑑 η 𝑑 r ) 1 p
(2.6) p A 0 ( t r n - 1 B 1 + W ( r ξ ) 𝑑 ξ 𝑑 r ) - 1 q = p A 0 ( t W n ( r ) 𝑑 r ) - 1 q .

Similarly, we also obtain

r W n ( t ) ( t W n ( τ ) 𝑑 τ ) - 1 p 𝑑 t = p ( r W n ( t ) 𝑑 t ) 1 p = p ( r B 1 + t n - 1 W ( t ξ ) 𝑑 ξ 𝑑 t ) 1 p
(2.7) p A 0 q p ( 0 r t n - 2 S n - 2 U 1 - p ( t η ) 𝑑 η 𝑑 t ) - q p p = p A 0 q p h - q ( r ) .

Combining (2.5), (2.6) and (2.7), we conclude that

+ n W ( x ) ( B | x | n - 1 g ( y ) 𝑑 y ) q 𝑑 x ( 0 F n ( r ) ( r W n ( t ) ( H n ( t ) ) q p 𝑑 t ) p q 𝑑 r ) q p
p ( p ) q p A 0 q ( 0 F n ( r ) h - p ( r ) 𝑑 r ) q p
= p ( p ) q p A 0 q ( 0 S n - 2 r n - 2 ( g ( r η ) U 1 p ( r η ) h ( r ) ) p h - p ( r ) 𝑑 η 𝑑 r ) q p
= p ( p ) q p A 0 q ( 0 S n - 2 r n - 2 g p ( r η ) U ( r η ) 𝑑 η 𝑑 r ) q p
= p ( p ) q p A 0 q ( + n g p ( y ) U ( y ) 𝑑 y ) q p .

The proof of Lemma 2.1 is completed. ∎

Now we are in position to give the proof of Theorem 1.1.

Proof of Theorem 1.1.

It is easy to verify that inequality (1.6) is equivalent to the following inequality:

E λ ( g ) | x | α L q ( + n ) C n , α , β , p , q g | y | - β L p ( + n ) ,

where

E λ ( g ) ( x ) = + n | x - y | λ g ( y ) d y

and

1 q = n - 1 n 1 p - α + β + λ - 1 n - 1 .

Then it follows that

E λ ( g ) | x | α L q ( + n ) ( + n ( + n | x | α | x - y | λ g ( y ) d y ) q d x ) 1 q
( + n ( B | x | / 2 n - 1 | x | α | x - y | λ g ( y ) d y ) q d x ) 1 q + ( + n ( + n B 2 | x | n - 1 | x | α | x - y | λ g ( y ) d y ) q d x ) 1 q
(2.8) = I 1 + I 2 .

Therefore, it suffices to show that

I 1 C n , α , β , p , q g | y | - β L p ( + n ) or I 2 C n , α , β , p , q g | y | - β L p ( + n ) .

We only prove that if β<-n-1p, then I1Cn,α,β,p,qg|y|-βLp(+n).

In fact, we can write

I 1 ( 1 2 ) λ ( + n | x | ( α + λ ) q ( B | x | / 2 n - 1 g ( y ) d y ) q d x ) 1 q .

Taking W(x)=|2x|(α+λ)q and U(y)=|y|-βp in (2.1) and combining it with (2.8), we conclude that

E λ ( g ) | x | α L q ( + n ) C n , α , β , p , q g | y | - β L p ( + n ) ,

as long as W(x)=|2x|(α+λ)q and U(y)=|y|-βp satisfy the assumption of Lemma 2.1. From

1 q = n - 1 n 1 p - α + β + λ - 1 n - 1 and β < 1 - n p ,

we deduce that nq>-α-λ. Then it follows that for any R>0,

(2.9) | x | R W ( x ) d x = | x | R | 2 x | ( α + λ ) q d x = 2 ( α + λ ) q B 1 + d ξ R t ( α + λ ) q + n - 1 d t = C 1 ( n , λ , α , q ) R n + ( α + λ ) q

and

(2.10) | y | R U 1 - p ( y ) 𝑑 y = | y | R ( | y | - β p ) 1 - p 𝑑 y = S n - 2 𝑑 η 0 R r β p + n - 2 𝑑 r = C 2 ( n , λ , β , p ) R n - 1 + β p .

This, together with (2.9) and (2.10), leads to

( | x | R W ( x ) 𝑑 x ) 1 q ( | y | R U 1 - p ( y ) 𝑑 y ) 1 p > C ( n , α , β , λ , p ) R n - 1 p + n q + α + β + λ = C ( n , α , β , λ , p ) ,

where C(n,α,β,λ,p)=min{C1(n,λ,α,q), C2(n,λ,β,p)} and n-1p+nq+α+β+λ=0. The proof of Theorem 1.1 is completed. ∎

3 The Proof of Theorem 1.2

In this section, we shall establish the existence result of extremals of inequality (1.6) using the rearrangement inequality. Let g be a nonnegative function defined on +n and denote its symmetric decreasing rearrangement by g (see [11] for specific rearrangement definition). We will need the following lemma in the proof of Theorem 1.2. (see, e.g., [29, 36])

Lemma 3.1.

  1. If g ( y ) is radially symmetric, then V λ ( g ) ( x , x n ) is radially symmetric with respect to x , where x = ( x , x n ) + n .

  2. For any nonnegative function g L p ( + n ) , there holds

    V λ ( g ) L q ( + n ) V λ ( g * ) | L q ( + n ) .

In order to obtain the desired minimizer of (1.7), we also need the following lemma. The proof of Lemma 3.2 is similar to that of [19, Lemma 3.1] and [36, Lemma 3]. However, with the presence of the weights, our proof is more involved and slightly more complicated.

Lemma 3.2.

Under the same hypothesis of Theorem 1.2, if we furthermore assume that gLp(R+n) is nonnegative and radially symmetric with g(r)εr-n-1p, then there exists a constant C, independent of ε, such that for any 0<t<p, there holds

V λ ( g ) L q ( + n ) C ε 1 - p t g L p ( + n ) p t .

Proof.

Define G: and h:+n by

G ( v ) = e v ( n - 1 ) p g ( e v ) and h ( x , x n ) = V λ ( g ) ( x , x n ) ,

respectively. One can easily verify that

(3.1) w n - 2 1 p G L p ( ) = g L p ( + n ) , G L ( ) ε ,

and Vλ(g)(x,xn) is radially symmetric in x. Define H:×+ by

H ( u , x n ) = e u n q h ( e u , e u x n ) .

Then

(3.2) w n - 2 1 q H L q ( + 2 ) = h L q ( + n ) .

Let l be some vector sitting on +n with length l. A straightforward calculation leads to

H ( u , x n ) = e u n q + n | e 2 u + e 2 u x n 2 | α 2 ( | e u - y | 2 + e 2 u x n 2 ) λ 2 g ( y ) | y | β d y
= e u ( n q + α ) + n | 1 + x n 2 | α 2 ( e 2 u - 2 e u y + y 2 + e 2 u x n 2 ) λ 2 g ( y ) | y | β d y
= e u ( n q + α ) - + S n - 2 | 1 + x n 2 | α 2 ( e 2 u - 2 e u e v + e 2 v + e 2 u x n 2 ) λ 2 g ( e v ) e v ( n - 1 + β ) d θ d v
= e u ( n q + α + λ 2 ) - + S n - 2 | 1 + x n 2 | α 2 ( e u - v ( 1 + x n 2 ) - 2 1 θ + e v - u ) λ 2 g ( e v ) e v ( n - 1 + λ 2 + β ) d θ d v .

In view of the index equality nq+n-1+α+β+λ=n-1p, an explicit form of H can be obtained as follows:

H ( u , x n ) = - + L n , α , β ( u - v , x n ) G ( v ) 𝑑 v ,

where Ln,α,β(u,xn) is given by

L n , α , β ( u , x n ) = { e u ( n q + α + λ 2 ) | 1 + x n 2 | α 2 ( ( e u ( 1 + x n 2 ) - 2 + e - u ) λ 2 + ( e u ( 1 + x n 2 ) + 2 + e - u ) λ 2 ) , n = 2 , e u ( n q + α + λ 2 ) S n - 2 | 1 + x n 2 | α 2 ( e u ( 1 + x n 2 ) - 2 1 θ + e - u ) λ 2 d θ , n 3 .

It is easy to check that

L n , α , β ( u , x n ) e u ( n q + α + λ 2 ) | 1 + x n 2 | α 2 ( e u ( 1 + x n 2 ) + e - u ) λ 2

as u2+xn2+. According to the assumption of Theorem 1.2, for any s<0, we can calculate

(3.3) ( 0 L q ( u , x n ) d x n ) s q d u < + .

For any 0<t<p, we can choose s1 in such a way that 1t+1s1=1+1q. It is clear that s1<0, and with the help of the reverse Young inequality, we obtain

| H ( u , x n ) | q d u ( | L ( u , x n ) | s 1 d u ) q s 1 ( | G | t d u ) q t .

Taking the integral with respect to the variable xn over [0,+) for both sides, by the Minkowski inequality, (3.3) and (3.1), we have

+ 2 | H ( u , x n ) | q d u d x n ( | G | t d u ) q t 0 ( L s 1 ( u , x n ) d u ) q s 1 d x n
( | G | t d u ) q t ( ( 0 L q ( u , x n ) d x n ) s 1 q d u ) q s 1
G L ( ) q ( 1 - p t ) ( | G | p d u ) q t
ε q ( 1 - p t ) ( + n | g | p d y ) q t .

Through (3.2), we derive that

V λ ( g ) L q ( + n ) C ε 1 - p t g L p ( + n ) p t .

The proof of Lemma 3.2 is completed. ∎

Now we turn our attention to the existence of the extremals of the reverse Stein–Weiss inequality on the upper half space +n. Our proof can be divided into two steps. Step 1. Choose a proper minimizing sequence {gj}j satisfying gj(1)c0 for any j, and find a potential minimizer g for (1.7).

Let {gj}j be a minimizing sequence for problem (1.7). According to Lemma 3.1, without loss of generality, we can assume that {gj}j is nonnegative radially symmetric and monotonously decreasing with gjLp(+n)=1.

For any R>0, we obtain

v n - 1 g j p ( R ) R n - 1 ω n - 2 0 R g j p ( r ) r n - 2 𝑑 r ω n - 2 0 + g j p ( r ) r n - 2 𝑑 r = + n g j p ( x ) 𝑑 x = 1 ,

where vn-1 and wn-2 denote the volume of the unit ball and the surface area of the unit sphere in n-1, respectively. The above estimate implies that there exists some constant C independent of j such that for any R>0,

(3.4) 0 g j ( R ) C R - n - 1 p .

Set

a j = sup r > 0 r n - 1 p g j ( r ) ,

in view of gjLp(+n)=1 , Vλ(gj)Lq(+n)Cn,α,β,p,q< and Lemma 3.2, we derive that aj2c0 for some c0>0. Select λj>0 satisfying λj(n-1)/pgj(λj)>c0, and define

g ~ j ( y ) = λ j n - 1 p g j ( λ j y ) .

It is not hard to verify that g~jLp=gjLp=1 and Vλg~jLq=VλgjLq. Replacing the sequence {gj(y)}j with the new sequence {g~j(y)}j, still denoted by {gj(y)}j, we obtain that gj(1)c0 for any j.

Following Lieb’s argument based on the Helly theorem, we can choose a subsequence such that gjg a.e. in +n. It is clearly that g is nonnegative radially symmetric and decreasing. Step 2. Verify that g is a minimizer of (1.7).

According to the definition of (Vλgj)(x), for all x+n and j*, we have

( V λ g j ) ( x ) c 0 | x | α | y | 1 | x - y | λ | y | β d y F ( x ) ,

where

F ( x ) { | x | α if  | x | 1 , | x | λ + α if  | x | > 1 .

For each x+n, set

k ( x ) = lim inf j ( V λ g j ( x ) ) .

One can employ Fatou’s lemma to obtain

+ n ( F q ( x ) - k q ( x ) ) 𝑑 x = + n lim inf j [ F q ( x ) - ( V λ g j ) q ( x ) ] d x
lim inf j + n [ F q ( x ) - ( V λ g j ) q ( x ) ] 𝑑 x
(3.5) = + n F q ( x ) 𝑑 x - ( C n , α , β , p , q ) q .

Since 1q=n-1n1p-α+β+λ-1n-1, β<1-np, α<-nq, we can verify that

+ n F q ( x ) 𝑑 x < .

Combining this with (3.5), we derive that

(3.6) + n F q ( x ) 𝑑 x + n k q ( x ) 𝑑 x ( C n , α , β , p , q ) q .

This estimate implies that the set {x+n:0<k(x)<+} has positive measure. Hence, we can take a point x1+n and extract a subsequence of Vλgj, still denoted by Vλgj, such that

lim j V λ g j ( x 1 ) = a 1 ( 0 , + ) .

By the reverse Hölder inequality, there holds

+ n | x 1 | α | x 1 - y | λ g j ( y ) | y | β d y C α , λ | y | | x 1 | 2 g j ( y ) | y | β d y C α , λ ( | y | | x 1 | 2 g j τ ( y ) d y ) 1 τ ( | y | | x 1 | 2 | y | τ β d y ) 1 τ ,

where 0<τ<1 and 1τ+1τ=1. Since 0<β<1-np, we can choose τ, satisfying τ>p and τβ>1-n, such that the integral |y||x1|2|y|τβdy is finite. Therefore, it follows that

(3.7) | y | | x 1 | 2 g j τ ( y ) 𝑑 y 1 .

This, together with (3.4), yields that for any sufficiently large R, there holds

| y | R g j τ ( y ) 𝑑 y = | y | | x 1 | 2 g j τ ( y ) 𝑑 y + | x 1 | 2 | y | R g j τ ( y ) 𝑑 y C ( R ) .

On the other hand, for sufficiently large R, we also have

+ n | x 1 | α | x 1 - y | λ g j ( y ) | y | β d y R λ 3 4 R | y | R g j ( y ) | y | β d y
R λ g j ( R ) 3 4 R | y | R | y | β d y
R λ + n - 1 + β g j ( R ) .

Thus, we obtain gj(R)R1-n-λ-β. Note that 0<α<-nq, and so we conclude that

(3.8) lim R lim j | y | R g j p ( y ) 𝑑 y = 0 .

Combining (3.7) and (3.8), we derive that

lim j + n g j p ( y ) 𝑑 y = lim R lim j | y | R g j p ( y ) 𝑑 y + lim R lim j | y | R g j p ( y ) 𝑑 y
= lim R lim j | y | R g j p ( y ) 𝑑 y
= lim R | y | R g p ( y ) 𝑑 y
= + n g p ( y ) 𝑑 y ,

which implies that gLp(R+n)=1. It remains to show that VλgLq(+n)=Cn,α,β,p,q. By Fatou’s lemma, we have

(3.9) ( + n k q ( x ) ( x ) 𝑑 x ) 1 q = ( + n lim inf j ( V λ g j ) q ( x ) d x ) 1 q ( + n ( V λ g ) q ( x ) 𝑑 x ) 1 q .

Gathering (3.6) and (3.9), we can conclude that

C n , α , β , p , q ( + n k q ( x ) ( x ) 𝑑 x ) 1 q ( + n ( V λ g ) q ( x ) 𝑑 x ) 1 q C n , α , β , p , q ,

which implies that g is actually a minimizer of (1.7). This completes the proof of Theorem 1.2.

4 The Proof of Theorem 1.3

Throughout this section, we investigate some asymptotic estimates of positive solutions of system (1.9) around the origin and near infinity. We need the following lemma.

Lemma 4.1.

Given α, β, p1, p2, λ>0, assume that (u,v) is a pair of positive Lebesgue measurable solutions of (1.9). Then

+ n ( 1 + | y | λ ) v - p 2 ( y ) | y | β 𝑑 y < , + n ( 1 + | x | λ ) u - p 1 ( x ) | x | α 𝑑 x < .

Furthermore, for some constants C1,C21, xR+n, yR+n,

1 C 1 ( 1 + | x | λ ) u ( x ) | x | α C 1 ( 1 + | x | λ ) , 1 C 2 ( 1 + | y | λ ) v ( y ) | y | β C 2 ( 1 + | y | λ ) .

Proof.

We shall only show

(4.1) 1 C 1 ( 1 + | x | λ ) u ( x ) | x | α C 1 ( 1 + | x | λ )

and

(4.2) + n ( 1 + | y | λ ) v - p 2 ( y ) | y | β 𝑑 y < .

Since (u,v) is a pair of positive Lebesgue measurable solutions of (1.9), we obtain

(4.3) meas { x + n : u ( x ) < + } > 0 , meas { y + n : v ( y ) < + } > 0 .

Then, there exist R>1 and some measurable set E such that

E { y + n : v ( y ) < R } B R n - 1 ,

satisfying |E|>1R.

For |x|>2R>2, one can write

u ( x ) | x | α = + n | x - y | λ v - p 2 ( y ) | y | β d y
C E ( 1 + | x | λ ) v - p 2 ( y ) | y | β 𝑑 y
C ( 1 + | x | λ ) R - p 2 E | y | β d y
C R ( 1 + | x | λ ) .

For 0<|x|2R, one can write

u ( x ) | x | α ( 1 + | x | λ ) 1 1 + ( 2 R ) λ R - p 2 E | x - y | λ | y | β d y
1 1 + ( 2 R ) λ R - p 2 E { | x - y | | y | } | x - y | λ + β d y + 1 1 + ( 2 R ) λ R - p 2 E { | x - y | | y | } | y | λ + β d y
c 1 + ( 2 R ) λ R - p 2 ( | E { | x - y | | y | } | 1 + λ + β n - 1 + | E { | x - y | | y | } | 1 + λ + β n - 1 )
c 0 1 + ( 2 R ) λ R - p 2 - 1 - λ + β n - 1 .

Then, for any x+n,

u ( x ) | x | α 1 C 1 ( 1 + | x | λ ) .

Thus, we obtain the left-hand side of the inequality in (4.1).

A similar argument shows that for any y+n{0},

(4.4) v ( y ) | y | β 1 C 2 ( 1 + | y | λ ) .

Next, we are ready to show that

+ n ( 1 + | y | λ ) v - p 2 ( y ) | y | β d y < .

By (4.3), there exists some x¯+n such that

u ( x ¯ ) = + n | x ¯ | α | x - y | λ v - p 2 ( y ) | y | β d y < + .

Then

+ n ( 1 + | y | λ ) v - p 2 ( y ) | y | β d y C x ¯ | y | < 1 2 | x ¯ | | x ¯ - y | λ v - p 2 ( y ) | y | β d y + C x ¯ | y | > 2 | x ¯ | | x ¯ - y | λ v - p 2 ( y ) | y | β d y
(4.5) + 1 2 | x ¯ | | y | 2 | x ¯ | ( 1 + | y | λ ) v - p 2 ( y ) | y | β 𝑑 y .

This, together with (4.4), implies that (4.2).

For x+n, with the help of (4.5), we can verify that

u ( x ) | x | α ( 1 + | x | λ ) = + n | x - y | λ ( 1 + | x | λ ) v - p 2 ( y ) | y | β 𝑑 y + n ( 1 + | y | λ ) v - p 2 ( y ) | y | β 𝑑 y < + .

The proof of Lemma 4.1 is completed. ∎

Now, we are in position to start our proof of Theorem 1.3. The proof is carried out in two parts. Part I. We give asymptotic estimates of u and v around the origin.

For x+n satisfying 0<|x|<1 and y+n satisfying 0<|y|<1, with the help of Lemma 4.1, we have

+ n | x - y | λ v - p 2 ( y ) | y | β d y C λ + n ( 1 + | y | λ ) v - p 2 ( y ) | y | β d y <

and

+ n | x - y | λ u - p 1 ( x ) | x | α d x C λ + n ( 1 + | x | λ ) u - p 1 ( x ) | x | α d x < .

The dominated convergence theorem applied to u and v shows that

lim | x | 0 u ( x ) | x | α = lim | x | 0 + n | x - y | λ v - p 2 ( y ) | y | β d y = + n v - p 2 ( y ) | y | λ + β d y

and

lim | y | 0 v ( y ) | y | β = lim | y | 0 + n | x - y | λ u - p 1 ( x ) | x | α d x = + n u - p 1 ( x ) | x | λ + α d x ,

which accomplishes the proof of Part I. Part II. We give asymptotic estimates of u and v near infinity.

For x+n satisfying |x|>1 and y+n satisfying |y|>1, thanks to Lemma 4.1, we obtain

| x | - λ + n | x - y | λ | y | β v - p 2 ( y ) d y C λ + n ( 1 + | y | λ ) v - p 2 ( y ) | y | β d y <

and

| y | - λ + n | x - y | λ u - p 1 ( x ) | x | α d x C λ + n ( 1 + | x | λ ) u - p 1 ( x ) | x | α d x < .

Then it follows from the dominated convergence theorem that

lim | x | u ( x ) | x | λ + α = + n v - p 2 ( y ) | y | β 𝑑 y

and

lim | y | v ( y ) | y | λ + β = + n u - p 1 ( x ) | x | α 𝑑 x .

This accomplishes the proof of Part II.

5 The Proof of Theorem 1.4

In this section, we obtain some necessary conditions for the existence of positive solutions to weighted integral system (1.10). For λ, ν1, ν2, p1, p2>0, assume that (u,v) is a pair of positive solutions of the following integral system:

{ u ( x ) = + n | x - y | λ | y | ν 2 v - p 2 ( y ) d y , x + n , v ( y ) = + n | x - y | λ | x | ν 1 u - p 1 ( x ) d x , y + n .

Integration by parts shows that

B R + | x | ν 1 u - p 1 ( x ) ( x u ( x ) ) d x = 1 1 - p 1 B R + | x | ν 1 x ( u 1 - p 1 ( x ) ) d x
= 1 1 - p 1 B R + u 1 - p 1 ( x ) R 1 + ν 1 d σ - n + ν 1 1 - p 1 B R + | x | ν 1 u 1 - p 1 ( x ) d x

and

B R n - 1 | y | ν 2 v - p 2 ( y ) ( y v ( y ) ) d y = 1 1 - p 2 B R n - 1 v 1 - p 2 ( y ) R 1 + ν 2 d σ - n - 1 + ν 2 1 - p 2 B R n - 1 | y | ν 2 v 1 - p 2 ( y ) d y .

Thanks to Lemma 4.1, we obtain

+ n | x | ν 1 u 1 - p 1 ( x ) d x < , + n | y | ν 2 v 1 - p 2 ( y ) d y < .

Therefore, there exists R=Rj+ such that

R 1 + ν 1 B R + u 1 - p 1 ( x ) 𝑑 σ 0 , R 1 + ν 2 B R n - 1 v 1 - p 2 ( y ) 𝑑 σ 0 .

Then we derive that

+ n | x | ν 1 u - p 1 ( x ) ( x u ( x ) ) d x + + n | y | ν 2 v - p 2 ( y ) ( y v ( y ) ) d y
(5.1) = - n + ν 1 1 - p 1 + n | x | ν 1 u 1 - p 1 ( x ) d x - n - 1 + ν 2 1 - p 2 + n | y | ν 2 v 1 - p 2 ( y ) d y .

In view of the integral system (1.10), we also have

u ( x ) x = d u ( ρ x ) d p | ρ = 0 = + n x ( x - y ) | x - y | λ - 2 | y | ν 2 v - p 2 ( y ) 𝑑 y

and

v ( y ) y = d u ( ρ y ) d p | ρ = 0 = + n y ( y - x ) | x - y | λ - 2 | x | ν 1 u - p 1 ( x ) 𝑑 x .

Then it follows that

+ n | x | ν 1 u - p 1 ( x ) ( x u ( x ) ) d x = λ + n + n x ( x - y ) | x - y | λ - 2 | x | ν 1 | y | ν 2 u - p 1 ( x ) v - p 2 ( y ) d y d x

and

+ n | y | ν 2 v - p 2 ( y ) ( y v ( y ) ) d y = λ + n + n y ( y - x ) | x - y | λ - 2 | y | ν 2 | x | ν 1 v - p 2 ( y ) u - p 1 ( x ) d x d y
= - λ + n + n y ( x - y ) | x - y | λ - 2 | y | ν 2 | x | ν 1 v - p 2 ( y ) u - p 1 ( x ) 𝑑 y 𝑑 x .

Therefore, we conclude that

+ n | x | ν 1 u - p 1 ( x ) ( x u ( x ) ) d x + + n | y | ν 2 v - p 2 ( y ) ( y v ( y ) ) d y = λ + n + n | x - y | λ | x | ν 1 | y | ν 2 v - p 2 ( y ) u - p 1 ( x ) d y d x
= λ + n | x | ν 1 u 1 - p 1 ( x ) d x = λ + n | y | ν 2 v 1 - p 2 ( y ) d y .

This, together with (5.1), implies that n+ν1p1-1+n-1+ν2p2-1=λ. The proof of Theorem 1.4 is completed.

6 The Proof of Theorem 1.7

In this section, we employ the stereographic projection to give a spherical form of the reverse Stein–Weiss inequality on the upper half space +n in the special case q=2n2n+λ+α+β and p=2n-22n-2+λ+α+β.

Consider the stereographic projection S:xnξ𝕊n{(0,0,,-1)} given by

ξ i = 2 x i 1 + | x | 2 for  i = 1 , 2 , , n , ξ n + 1 = 1 - | x | 2 1 + | x | 2 .

It is easy to check that S transforms +n into 𝕊+n={(ξ1,,ξn+1𝕊n):ξn+1>0} and transforms +n into 𝕊0n={(ξ1,,ξn+1𝕊n):ξn+1=0}. For x+n, yn, ξ𝕊+n and η𝕊0n, one can refer to [30] to obtain

| S ( x ) - S ( y ) | = ( 4 | x - y | 2 ( 1 + | x | 2 ) ( 1 + | x | 2 ) ) 1 2 , d ξ = ( 2 1 + | x | 2 ) n d x , d η = ( 2 1 + | y | 2 ) n - 1 d y .

Given λ, α, β>0, q=2n2n+λ+α+β, p=2n-22n-2+λ+α+β, ξ𝕊n, η𝕊0n, fLq(+n), gLp(+n), define

F ( ξ ) = ( 2 1 + | x | 2 ) - 2 n + λ + α + β 2 f ( x ) , F 1 ( ξ ) = ( 2 1 + | x | 2 ) - 2 n + λ + α + β 2 ( 1 1 + | x | 2 ) β 2 f ( x ) ,
G ( η ) = ( 2 1 + | y | 2 ) - 2 n - 2 + λ + α + β 2 g ( y ) , G 1 ( η ) = ( 2 1 + | y | 2 ) - 2 n - 2 + λ + α + β 2 ( 1 1 + | y | 2 ) α 2 g ( y ) ,

where x=S-1(ξ), y=S-1(η). A straightforward calculation shows

𝕊 + n | F ( ξ ) | q d ξ = + n | f ( x ) | q ( 2 1 + | x | 2 ) - q 2 n + λ + α + β 2 ( 2 1 + | x | 2 ) n d x = + n | f ( x ) | q d x

and

𝕊 0 n | G ( η ) | p d η = + n | g ( y ) | p ( 2 1 + | y | 2 ) - p 2 n - 2 + λ + α + β 2 ( 2 1 + | y | 2 ) n - 1 d y = + n | g ( y ) | p d y .

Recall the reversed Stein–Weiss inequality on the half space +n:

+ n + n | x | α | x - y | λ f ( x ) g ( y ) | y | β d x d y C n , α , β , p , q f L q ( + n ) g L p ( + n ) ,

where p, q(0,1), α, β and λ>0 satisfy n-1n1p+1q-α+β+λ-1n=2.

Then it follows from stereographic projection that

𝕊 + n 𝕊 0 n | ξ - S ( 0 ) | α | ξ - η | λ F 1 ( ξ ) G 1 ( η ) | η - S ( 0 ) | β d η d ξ
= + n + n ( 4 | x | 2 1 + | x | 2 ) α 2 ( 4 | x - y | 2 ( 1 + | x | 2 ) ( 1 + | y | 2 ) ) λ 2 ( 2 1 + | x | 2 ) - λ + α + β 2 f ( x )
    × ( 1 1 + | x | 2 ) β 2 ( 2 1 + | y | 2 ) - λ + α + β 2 g ( y ) ( 1 1 + | y | 2 ) α 2 ( 4 | y | 2 1 + | y | 2 ) β 2 d y d x
= + n + n | x | α | x - y | λ f ( x ) g ( y ) | y | β d y d x
C n , α , β , p , q ( + n | f ( x ) | q d x ) 1 q ( + n | g ( y ) | q d y ) 1 p
= C n , α , β , p , q ( 𝕊 + n | F ( ξ ) | q d ξ ) 1 q ( 𝕊 0 n | G ( η ) | p d η ) 1 p ,

which accomplishes the proof of Theorem 1.7.

7 The Proof of Theorem 1.9

In this section, we also give a spherical form of the Stein–Weiss inequality on the upper half space +n by choosing q=2n2n-λ-α-β, p=2n-22n-2-λ-α-β.

Recall the classical Stein–Weiss inequality on the upper half space +n which can be stated as follows:

+ n + n | x | - α | x - y | - λ f ( x ) g ( y ) | y | - β d y d x C n , α , β , p , q f L q ( + n ) g L p ( + n ) ,

where p, q, α, β and λ satisfy the following conditions:

1 q + n - 1 n 1 p + α + β + λ + 1 n = 2 , 1 q + 1 p 1 ,
α + β 0 , α < n q , β < n - 1 p , 0 < λ < n - 1 .

For q=2n2n-λ-α-β, p=2n-22n-2-λ-α-β, ξ𝕊+n, η𝕊0n, fLq(+n), gLp(+n), set

H ( ξ ) = ( 2 1 + | x | 2 ) - 2 n - λ - α - β 2 f ( x ) , H 1 ( ξ ) = ( 2 1 + | x | 2 ) - 2 n - λ - α - β 2 ( 1 1 + | x | 2 ) - β 2 f ( x ) ,
T ( η ) = ( 2 1 + | y | 2 ) - 2 n - 2 - λ - α - β 2 g ( y ) , T 1 ( η ) = ( 2 1 + | y | 2 ) - 2 n - 2 - λ - α - β 2 ( 1 1 + | y | 2 ) - α 2 g ( y ) ,

where x=S-1(ξ),y=S-1(η). An easy computation yields that

𝕊 + n | H ( ξ ) | q d ξ = + n | f ( x ) | q ( 2 1 + | x | 2 ) - q 2 n - λ - α - β 2 ( 2 1 + | x | 2 ) n d x = + n | f ( x ) | q d x

and

𝕊 0 n | T ( η ) | p d η = + n | g ( y ) | p ( 2 1 + | y | 2 ) - p 2 n - 2 - λ - α - β 2 ( 2 1 + | y | 2 ) n - 1 d y = + n | g ( y ) | p d y .

Combining this with the Stein–Weiss inequality (1.5) and the stereographic projection, we conclude that

𝕊 + n 𝕊 0 n | ξ - S ( 0 ) | - α | ξ - η | - λ H 1 ( ξ ) T 1 ( η ) | η - S ( 0 ) | - β d η d ξ
= + n + n ( 4 | x | 2 1 + | x | 2 ) - α 2 ( 4 | x - y | 2 ( 1 + | x | 2 ) ( 1 + | y | 2 ) ) - λ 2 ( 2 1 + | x | 2 ) λ + α + β 2 f ( x )
    × ( 1 1 + | x | 2 ) - β 2 ( 2 1 + | y | 2 ) λ + α + β 2 g ( y ) ( 1 1 + | y | 2 ) - α 2 ( 4 | y | 2 1 + | y | 2 ) - β 2 d y d x
= + n + n | x | - α | x - y | - λ f ( x ) g ( y ) | y | - β d y d x
C n , α , β , p , q ( + n | f ( x ) | q d x ) 1 q ( + n | g ( y ) | p d y ) 1 p
= C n , α , β , p , q ( 𝕊 + n | H ( ξ ) | q d ξ ) 1 q ( 𝕊 0 n | T ( η ) | p d η ) 1 p ,

which accomplishes the proof of Theorem 1.9.


Communicated by Luis Caffarelli


Award Identifier / Grant number: 11371056

Funding statement: The first and third authors were partly supported by grant from the NNSF of China (No.11371056), the second author was partly supported by a US NSF grant and a grant from the Simons foundation.

References

[1] W. Beckner, Geometric inequalities in Fourier anaylsis Essays on Fourier Analysis in Honor of Elias M. Stein (Princeton 1991), Princeton Math. Ser. 42, Princeton University, Princeton (1995), 36–68. 10.1515/9781400852949.36Search in Google Scholar

[2] W. Beckner, Pitt’s inequality and the uncertainty principle, Proc. Amer. Math. Soc. 123 (1995), no. 6, 1897–1905. 10.1090/S0002-9939-1995-1254832-9Search in Google Scholar

[3] W. Beckner, Sharp inequalities and geometric manifolds, J. Fourier Anal. Appl. 3 (1997), 825–836. 10.1007/BF02656488Search in Google Scholar

[4] W. Beckner, Pitt’s inequality with sharp convolution estimates, Proc. Amer. Math. Soc. 136 (2008), no. 5, 1871–1885. 10.1090/S0002-9939-07-09216-7Search in Google Scholar

[5] W. Beckner, Weighted inequalities and Stein–Weiss potentials, Forum Math. 20 (2008), no. 4, 587–606. 10.1515/FORUM.2008.030Search in Google Scholar

[6] W. Beckner, Multilinear embedding estimates for the fractional Laplacian, Math. Res. Lett. 19 (2012), no. 1, 175–189. 10.4310/MRL.2012.v19.n1.a14Search in Google Scholar

[7] W. Beckner, Functionals for multilinear fractional embedding, Acta Math. Sin. (Engl. Ser.) 31 (2015), no. 1, 1–28. 10.1007/s10114-015-4321-6Search in Google Scholar

[8] E. A. Carlen, J. A. Carrillo and M. Loss, Hardy–Littlewood–Sobolev inequalities via fast diffusion flows, Proc. Natl. Acad. Sci. USA 107 (2010), no. 46, 19696–19701. 10.1073/pnas.1008323107Search in Google Scholar PubMed PubMed Central

[9] E. Carneiro, A sharp inequality for the Strichartz norm, Int. Math. Res. Not. IMRN 2009 (2009), no. 16, 3127–3145. 10.1093/imrn/rnp045Search in Google Scholar

[10] L. Chen, Z. Liu and G. Lu, Symmetry and regularity of solutions to the weighted Hardy–Sobolev type system, Adv. Nonlinear Stud. 16 (2016), no. 1, 1–13. 10.1515/ans-2015-5005Search in Google Scholar

[11] L. Chen, Z. Liu, G. Lu and C. Tao, Reverse Stein–Weiss inequalities and existence of their extremal functions, Trans. Amer. Math. Soc. 370 (2018), no. 12, 8429–8450. 10.1090/tran/7273Search in Google Scholar

[12] L. Chen, Z. Liu, G. Lu and C. Tao, Stein–Weiss inequalities with the fractional Poisson kernel, preprint (2018), https://arxiv.org/abs/1807.04906; to appear in Rev. Mat. Iberoam. 10.4171/rmi/1167Search in Google Scholar

[13] L. Chen, G. Lu and C. Tao, Existence of extremal functions for the Stein–Weiss inequalities on the Heisenberg group, preprint (2018), https://arxiv.org/abs/1807.04699. 10.1016/j.jfa.2019.01.002Search in Google Scholar

[14] L. Chen, G. Lu and C. Tao, Hardy–Littlewood–Sobolev inequality with fractional Poisson kernel and its appliaction in PDEs, Acta Math. Sin. (Engl. Ser.), to appear. Search in Google Scholar

[15] W. Chen, C. Jin, C. Li and J. Lim, Weighted Hardy–Littlewood–Sobolev inequalities and systems of integral equations, Discrete Contin. Dyn. Syst. 2005 (2005), 164–172. Search in Google Scholar

[16] W. Chen and C. Li, The best constant in a weighted Hardy–Littlewood–Sobolev inequality, Proc. Amer. Math. Soc. 136 (2008), no. 3, 955–962. 10.1090/S0002-9939-07-09232-5Search in Google Scholar

[17] J. Dou, Weighted Hardy–Littlewood–Sobolev inequalities on the upper half space, Commun. Contemp. Math. 18 (2016), no. 5, Article ID 1550067. 10.1142/S0219199715500674Search in Google Scholar

[18] J. Dou and M. Zhu, Reversed Hardy–Littewood–Sobolev inequality, Int. Math. Res. Not. IMRN 2015 (2015), no. 19, 9696–9726. 10.1093/imrn/rnu241Search in Google Scholar

[19] J. Dou and M. Zhu, Sharp Hardy–Littlewood–Sobolev inequality on the upper half space, Int. Math. Res. Not. IMRN 2015 (2015), no. 3, 651–687. 10.1093/imrn/rnt213Search in Google Scholar

[20] P. Drábek, H. P. Heinig and A. Kufner, Higher-dimensional Hardy inequality, General Inequalities. 7 (Oberwolfach 1995), Internat. Ser. Numer. Math. 123, Birkhäuser, Basel (1997), 3–16. 10.1007/978-3-0348-8942-1_1Search in Google Scholar

[21] R. L. Frank and E. H. Lieb, Inversion positivity and the sharp Hardy–Littlewood–Sobolev inequality, Calc. Var. Partial Differential Equations 39 (2010), no. 1–2, 85–99. 10.1007/s00526-009-0302-xSearch in Google Scholar

[22] R. L. Frank and E. H. Lieb, A new, rearrangement-free proof of the sharp Hardy–Littlewood–Sobolev inequality, Spectral Theory, Function Spaces and Inequalities, Oper. Theory Adv. Appl. 219, Birkhäuser, Basel (2012), 55–67. 10.1007/978-3-0348-0263-5_4Search in Google Scholar

[23] R. L. Frank and E. H. Lieb, Sharp constants in several inequalities on the Heisenberg group, Ann. of Math. (2) 176 (2012), no. 1, 349–381. 10.4007/annals.2012.176.1.6Search in Google Scholar

[24] X. Han, Existence of maximizers for Hardy–Littlewood–Sobolev inequalities on the Heisenberg group, Indiana Univ. Math. J. 62 (2013), no. 3,737–751. 10.1512/iumj.2013.62.4976Search in Google Scholar

[25] X. Han, G. Lu and J. Zhu, Hardy–Littlewood–Sobolev and Stein–Weiss inequalities and integral systems on the Heisenberg group, Nonlinear Anal. 75 (2012), no. 11, 4296–4314. 10.1016/j.na.2012.03.017Search in Google Scholar

[26] F. Hang, X. Wang and X. Yan, Sharp integral inequalities for harmonic functions, Comm. Pure Appl. Math. 61 (2008), no. 1, 54–95. 10.1002/cpa.20193Search in Google Scholar

[27] G. H. Hardy and J. E. Littlewood, Some properties of fractional integrals. I, Math. Z. 27 (1928), no. 1, 565–606. 10.1007/BF01171116Search in Google Scholar

[28] I. W. Herbst, Spectral theory of the operator (p2+m2)1/2-Ze2/r, Comm. Math. Phys. 53 (1977), no. 3, 285–294. 10.1007/BF01609852Search in Google Scholar

[29] E. H. Lieb, Sharp constants in the Hardy–Littlewood–Sobolev and related inequalities, Ann. of Math. (2) 118 (1983), no. 2, 349–374. 10.1007/978-3-642-55925-9_43Search in Google Scholar

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

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

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

[33] G. Lu and C. Tao, Reverse Hardy–Littlewood–Sobolev and Stein–Weiss inequalities on the Heisenberg group, preprint (2018). Search in Google Scholar

[34] G. Lu and J. Zhu, Symmetry and regularity of extremals of an integral equation related to the Hardy–Sobolev inequality, Calc. Var. Partial Differential Equations 42 (2011), no. 3–4, 563–577. 10.1007/s00526-011-0398-7Search in Google Scholar

[35] Q. A. Ngô and V. H. Nguyen, Sharp reversed Hardy–Littlewood–Sobolev inequality: The case of whole space n, preprint (2016), https://arxiv.org/abs/1508.02041v2. 10.1093/imrn/rnw108Search in Google Scholar

[36] Q. A. Ngô and V. H. Nguyen, Sharp reversed Hardy-Littlewood-Sobolev inequality on the half space 𝐑+n, Int. Math. Res. Not. IMRN 2017 (2017), no. 20, 6187–6230. 10.1093/imrn/rnw108Search in Google Scholar

[37] E. Sawyer and R. L. Wheeden, Weighted inequalities for fractional integrals on Euclidean and homogeneous spaces, Amer. J. Math. 114 (1992), no. 4, 813–874. 10.2307/2374799Search in Google Scholar

[38] S. L. Sobolev, On a theorem in functional analysis (in Russian), Mat. Sb. 4 (1938), 471–497. 10.1090/trans2/034/02Search in Google Scholar

[39] E. M. Stein, Singular Integrals and Differentiability Properties of Functions, Princeton Math. Ser. 30, Princeton University, Princeton, 1970. 10.1515/9781400883882Search in Google Scholar

[40] E. M. Stein, Harmonic Analysis: Real-variable Methods, Orthogonality, and Oscillatory Integrals, Princeton Math. Ser. 43, Princeton University, Princeton, 1993. 10.1515/9781400883929Search in Google Scholar

[41] E. M. Stein and G. Weiss, Fractional integrals on n-dimensional Euclidean space, J. Math. Mech. 7 (1958), 503–514. 10.1512/iumj.1958.7.57030Search in Google Scholar

Received: 2018-08-11
Revised: 2018-12-04
Accepted: 2018-12-05
Published Online: 2019-01-22
Published in Print: 2019-08-01

© 2019 Walter de Gruyter GmbH, Berlin/Boston

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

Downloaded on 9.5.2024 from https://www.degruyter.com/document/doi/10.1515/ans-2018-2038/html
Scroll to top button