Skip to content
BY 4.0 license Open Access Published by De Gruyter November 20, 2020

Reversed Stein–Weiss Inequalities with Poisson-Type Kernel and Qualitative Analysis of Extremal Functions

  • Chunxia Tao EMAIL logo

Abstract

Through conformal map, isoperimetric inequalities are equivalent to the Hardy–Littlewood–Sobolev (HLS) inequalities involved with the Poisson-type kernel on the upper half space. From the analytical point of view, we want to know whether there exists a reverse analogue for the Poisson-type kernel. In this work, we give an affirmative answer to this question. We first establish the reverse Stein–Weiss inequality with the Poisson-type kernel, finding that the range of index 𝑝,q appearing in the reverse inequality lies in the interval (0,1), which is perfectly consistent with the feature of the index for the classical reverse HLS and Stein–Weiss inequalities. Then we give the existence and asymptotic behaviors of the extremal functions of this inequality. Furthermore, for the reverse HLS inequalities involving the Poisson-type kernel, we establish the regularity for the positive solutions to the corresponding Euler–Lagrange system and give the sufficient and necessary conditions of the existence of their solutions. Finally, in the conformal invariant index, we classify the extremal functions of the latter reverse inequality and compute the sharp constant. Our methods are based on the reversed version of the Hardy inequality in high dimension, Riesz rearrangement inequality and moving spheres.

MSC 2010: 42B37; 35B40; 45G15

1 Introduction

There is a well-known inequality proved by Carleman [5] which states that the inequality

(1.1)B1e2fdx14π(S1efdθ)2

holds for any smooth harmonic function 𝑓 in two-dimensional unit disk B12¯, and the equality holds if and only if f(x)=C or -2log|x-x0|+C for some x0R2B12¯ and CR. From the Riemann mapping theorem, we know that any simply connected compact two-dimensional manifold with boundary (M2,g) is isometric to (B12¯,e2wdx2). Obviously, if the Gauss curvature of (M2,g) is nonpositive, then 𝑤 is a subharmonic function. It follows from the maximum principle that inequality (1.1) also holds for subharmonic functions. Hence, (1.1) can be regarded as the isoperimetric inequality for the two-dimensional manifold (M2,g) with the nonpositive Gauss curvature. Naturally, a higher-dimensional form of (1.1) for the harmonic function 𝑓,

(1.2)B1f2nn-2dxn-nn-1ωn-1n-1(B1f2(n-1)n-2ds)nn-1,

could be used to represent the corresponding isoperimetric inequality for the manifold (Mn,g) equipped with the nonpositive scalar curvature, which is isometric to (B1n¯,w4n-2dx2) (𝑤 is subharmonic because of nonpositivity of scalar curvature). Inequality (1.2) was proved by Hang, Wang and Yan in [27] through the method of the subcritical approximation. At the same time, they provided an alternative proof in [26] by proving the inequality

(1.3)PfL2nn-2(R+n)n-n-22(n-1)ωn-n-22n(n-1)fL2(n-1)n-2(R+n),

where (Pf)(x,xn) ((x,xn)R+n)) is the convolution of f(x) with the Poisson kernel

(1.4)P(x,xn)=2nωnxn(|x|2+xn2)n2,

and we can see that (1.2) is equivalent to (1.3) by the Möbius transformation between R+n and B1.

It should be noted that the Poisson kernel P(x,xn) appearing in inequality (1.3) is a fundamental solution of the Laplacian operator on half space R+n, that is to say, P(x,xn) satisfies the equation

{-ΔP(x,xn)=0,(x,xn)R+n,P(x,0)=δ0(x),(x,0)R+n.

From Caffarelli and Silvestre’s works in [4], we know that the outer normal derivatives of the Poisson integral (Pf)(x) along the boundary R+n could be used to define (-Δ)12f in Rn-1 as long as 𝑓 is sufficiently smooth. In order to define general fractional Laplacian operator (-Δ)α2 for α(0,2), Caffarelli and Silvestre introduced an extension operator 𝐿 on the half space R+n by giving the operator L=Δ+1-αxn. It is not difficult to verify that the fundamental solution for the equation

{LP(x,xn))=0,x=(x,xn)R+n,P(x,0)=δ0(x),(x,0)R+n,

is

(1.5)Pb(x,xn)=C1xn1-b(|x|2+xn2)n-b2,

where b=1-α. Then the fractional Laplacian (-Δ)α2f in Rn-1 could be defined as

(-Δ)α2f(x)=limxn0-xn1-αPb(f)xn=αlimxn0Pb(f)(x,xn)-Pb(f)(x,0)xnα,

where

Pb(f)(x,xn)=Rn-1Pb(x-y,xn)f(y)dy.

We may naturally ask whether there exists an analogue of inequality (1.3) for Pb(f). Chen [12] gave a positive answer and proved the result

(1.6)PbfL2nn-2+b(R+n)Cn,bfL2(n-1)n-2+b(R+n)

for b<1. Dou, Guo and Zhu [15] discussed the above conformal invariant index inequality with the kernel

Pθ,γ(x.xn)=xn(|x|2+xn2)γ2.

For this kernel, Chen, Lu and Tao [10] established the general LpLq boundedness

(1.7)R+nR+nf(x)xn(|x-y|2+xn2)γ2g(y)dydxCn,γ,p,qfLq(R+n)gLp(R+n),

where p,q>1, and 2<γn. By using the result in [10], Chen, Liu, Lu and Tao [8] extended it to the doubly weighted form. Recently, the above conformal invariant index inequalities have also been extended to a more general kernel in [23], which takes the form

(1.8)Pθ,γ(x.xn)=xnθ(|x|2+xn2)γ2

with θ>0, γ>0.

Inequalities (1.2), (1.6) and (1.7) could be also seen as a new type of HLS inequality [28, 33]. The classical HLS inequality is originally the convolution in Rn of f(x) with the Riesz kernel 1|x|γ. Many geometric inequalities, such as the Gross logarithmic Sobolev inequality [24] and the Moser–Onofri–Beckner inequality [1] can be implied by the HLS inequality. In the 1950s, Stein and Weiss [34] established the doubly weighted HLS inequality, namely, the Stein–Weiss inequality

(1.9)RnRn|x|-α|x-y|-γf(x)g(y)|y|-βdxdyCn,α,β,p,qfLq(Rn)gLp(Rn),

where

p,q>1,α+β0,α<nq,β<np,0<γ<n.

Lieb [31] utilized the symmetric rearrangement argument to prove the existence of the extremals of the HLS inequalities. He also classified the extremal functions in the conformal index p=q=2n2n-γ. Furthermore, Chen, Li and Ou [13] and Li [30] classified all the positive solutions of the Euler–Lagrange system related with the HLS inequality in the conformal case through the method of moving plane and the method of moving sphere respectively. Carlen and Loss [6] and Frank and Lieb [20, 21] also gave the best constant and an accurate expression of extremal functions of the HLS inequality in the conformal index without using a rearrangement argument. For more results about the HLS inequalities on the Heisenberg group and the upper half space, one can see [9, 14, 17, 22, 19, 25].

For inequality (1.9), what will happen if p,q(0,1)? Beckner [2] and Dou and Zhu [16] found that the exponent 𝛾 of the kernel becomes negative and the inequalities turn into reversed inequalities. In fact, they established the reverse type of the HLS inequality and derived the existence of extremal functions in the diagonal case p=q=2n2n-γ with γ<0. Subsequently, Chen, Liu, Lu and Tao [7] established the reverse Stein–Weiss inequality

RnRn|x|α|x-y|γf(x)g(y)|y|βdxdyCn,α,β,p,qfLq(Rn)gLp(Rn),

where p,q(0,1), γ>0, 0α<-nq and 0β<-np. Motivated by the above works, thus a natural questions arises: for the Poisson-type kernel such as (1.4), (1.5) and (1.8), what will happen to the inequalities when 0<p,q<1? We find that the inequalities could become reversed if the exponents 𝜃 and 𝛾 appearing in the kernel (1.8) are negative. To be specific, we derive the following general results.

Theorem 1.1

For n>1, 0<p,q<1, γ,θ>0 and either β<1-np or α<-nq+θ satisfying

(1.10)n-1np+1q-α+β+γ-1-θn=2,

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

(1.11)R+nR+n|x|αf(x)P(x,γ,θ,y)g(y)|y|βdydxCn,γ,θ,α,β,p,qfLq(R+n)gLp(R+n),

where

P(x,γ,θ,y)=|x-y|γxnθ,1p+1p=1,1q+1q=1.
Remark 1.2

Inequality (1.11) is called the reverse Stein–Weiss inequality with the Poisson-type kernel. When θ=0, the following reverse Stein–Weiss inequality on the upper half space was studied by Chen, Lu and Tao in [11]:

R+nR+n|x|αf(x)|x-y|γg(y)|y|βdydxCn,γ,α,β,p,qfLq(R+n)gLp(R+n).

When θ=α=β=0, it becomes the reverse HLS inequality on the upper half space studied by Ngô [32].

If (f,g)Cc(R+n)×Cc(R+n), the left integral of (1.11) is finite, which implies the sharp constant Cn,γ,θ,α,β,p,q is finite. A natural question is whether the sharp constant Cn,γ,θ,α,β,p,q can be attained. To do this, we can consider the minimizing problem

(1.12)Cn,γ,θ,α,β,p,q:=inf{Vγ,θ(g)Lq(R+n):g0, 0gLp(R+n)=1},

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

Vγ,θ(g)(x)=R+n|x|αP(x,γ,θ,y)g(y)|y|βdy.

Then we can prove that the constant Cn,γ,θ,α,β,p,q could actually be achieved.

Theorem 1.3

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

n-1np+1q-α+β+γ-1-θn=2,

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

Obviously, the extremals of reverse inequality (1.11) satisfy the following Euler–Lagrange system up to a constant:

(1.13){J(f,g)f(x)q-1=R+n|x|αP(x,γ,θ,y)g(y)|y|βdy,xR+n,J(f,g)g(y)p-1=R+n|y|βP(x,γ,θ,y)f(x)|x|αdx,yR+n.

Let u=c1fq-1, v=c2gp-1, 1q-1=-p1 and -p2=1p-1, and pick two suitable constants c1 and c2. Then system (1.13) is simplified as

(1.14){u(x)=R+n|x|αP(x,γ,θ,y)v-p2(y)|y|βdy,xR+n,v(y)=R+n|y|βP(x,γ,θ,y)u-p1(x)|x|αdx,yR+n,

where 1p1-1+1p2-1n-1n=α+β+γ-θn.

We also give some asymptotic estimates of positive solutions to system (1.14) around the origin and near infinity.

Theorem 1.4

Suppose that (u,v) is a pair of positive Lebesgue measurable solutions of (1.14). Then

lim|x|u(x)xnθ|x|γ+α=R+nv-p2(y)|y|βdy,lim|y|v(y)|y|γ+β=R+nu-p1(x)|x|αxn-θdx,
lim|x|0u(x)xnθ|x|α=R+nv-p2(y)|y|β+γdy,lim|y|0v(y)|y|β=R+nu-p1(x)|x|α+γxn-θdx.

When α=β=0, (1.11) becomes the reverse HLS inequality with the Poisson-type kernel

(1.15)R+nR+nf(x)|x-y|γg(y)xnθdydxCn,γ,θ,p,qfLq(R+n)gLp(R+n),

where 0<p,q<1, γ>0, θ>0 satisfying

n-1np+1q-γ-1-θn=2.

Accordingly, system (1.14) becomes

(1.16){u(x)=R+nP(x,γ,θ,y)v-p2(y)dy,xR+n,v(y)=R+nP(x,γ,θ,y)u-p1(x)dx,yR+n.

Then we have the following regularity result.

Theorem 1.5

Assume that (u,v) is a pair of positive solutions of integral system (1.16). Then

(u,v)C1(R+n)×C1(R+n).

From Theorem 1.3, we can deduce that if

(1.17)np1-1+n-1p2-1=γ-θ,

then system (1.16) must have positive solutions. A natural question arises, whether identity (1.17) is a necessary condition for the existence of positive solutions of system (1.16). We give a positive answer in the following theorem.

Theorem 1.6

For 𝛾, p1, p2>0, there exists a pair of positive solutions (u,v) satisfying (1.16) if and only if

np1-1+n-1p2-1=γ-θ.

When q=2n2n+γ-2θ, p=2(n-1)2n+γ-2, using the method of moving spheres in integral forms (see [30]), we can classify the positive solutions for system (1.16). Then we can derive the accurate forms of extremals for inequality (1.15) in the conformal index.

Theorem 1.7

For γ>2θ, q=2n2n+γ-2θ, p=2(n-1)2n+γ-2, there exists some constant Cn,γ,θ>0 such that, for any nonnegative functions fLq(R+n), gLp(R+n), there holds

(1.18)R+nR+nf(x)P(x,γ,θ,y)g(y)dydxCn,γ,θfLq(R+n)gLp(R+n),

where

Cn,γ,θ=2θ(nwn)-2n+γ-22(n-1)(Bn(Bn|ξ-ζ|γ(1-|ξ-zo|2)θdζ)-2nγ-2θdξ)-γ-2θ2n,z0=(0,-1)Rn-1×R.

The equality holds if and only if

g(y)=c(1|y-y0|2+d2)2n+γ-22andf(x)=C(R+n|x-y|γxnθ(1|y-y0|2+d2)2n+γ-22dy)2n+γ-2θ2θ-γ

for some c,d,C>0 and y0R+n.

For x0=(0,-2), z0=(0,-1)Rn-1×R, consider the conformal transformation Tx0:ξBnξx0R+n given by

Tx0(ξ)=ξx0=4(ξ-x0)|ξ-x0|2+x0R+n.

Via easy computation, one can note that Tx0-1(x)=xx0=4(x-x0)|x-x0|2+x0Bn, which has the same expression with Tx0. Through this conformal transformation, we can easily show that inequality (1.18) is equivalent to the following sharp reversed integral inequality on the ball Bn with Cn,γ,θ=2θC~n,γ,θ.

Corollary 1.8

For γ>2θ, q=2n2n+γ-2θ, p=2(n-1)2n+γ-2, the following sharp inequality holds for any nonnegative functions fLq(Bn), gLp(Bn):

(1.19)BnBnf(ξ)|ξ-ζ|γg(ζ)(1-|ξ-zo|2)θdζdξC~n,γ,θfLq(Bn)gLp(Bn),

where

C~n,γ,θ=(nwn)-2n+γ-22(n-1)(Bn(Bn|ξ-ζ|γ(1-|ξ-zo|2)θdζ)-2nγ-2θdξ)-γ-2θ2n,z0=(0,-1)Rn-1×R.
Remark 1.9

In the proof of Theorem 1.7, we classify the extremal functions of inequality (1.18). This together with the conformal transformation implies that g1 could be the extremal function of inequality (1.19). Then the sharp constant C~n,γ,θ and Cn,γ,θ can be easily computed.

2 The Proof of Theorem 1.1

Throughout this section, we shall establish the reverse Stein–Weiss inequality with the Poisson-type kernel. We need the following lemma (see [11]) which is the reverse version of the classical high-dimensional Hardy inequality (see [18, 14]).

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)(R+nW(x)(B|x|n-1g(y)dy)qdx)1qC0(p,q)(R+ngp(y)U(y)dy)1p

if

(2.2)0<A0=infR>0{(|x|RW(x)dx)1q(|y|RU1-p(y)dy)1p}<+,

and

(R+nW(x)(R+nB2|x|n-1g(y)dy)qdx)1qC1(p,q)(R+ngp(v)U(v)dv)1p

if

0<A1=infR>0{(|x|RW(x)dx)1q(|y|RU1-p(y)dy)1p}<+.

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

Proof

Let

Eγ,θ(g)(x)=R+n|x-y|γxnθg(y)dy.

By a dual argument, we can verify that inequality (1.11) is equivalent to

Eγ,θ(g)|x|αLq(R+n)Cn,α,β,p,qg|y|-βLp(R+n).

If β<1-np, we take W(x)=|x|(α+γ-θ)q and U(y)=|y|-βp in (2.1). Then, for any R>0, we have

|x|RW(x)dx=C1Rn+(α+γ-θ)qand|y|RU1-p(y)dy=C2Rn-1+βp.

Hence,

(|x|RW(x)dx)1q(|y|RU1-p(y)dy)1p>C(n,α,β,γ,θ,p)Rn-1p+nq+α+β+γ-θ=C(n,α,β,γ,θ,p),

where C(n,α,β,γ,θ,p)=min{C1, C2} and the last equality follows from condition (1.10) of Theorem 1.1. Hence, W(x) and U(y) satisfy condition (2.2). From (2.1), we have

Eγ,θ(g)|x|αLq(R+n)=(R+n(R+n|x|α|x-y|γxnθg(y)dy)qdx)1q(R+n(B|x|/2n-1|x|α|x-y|γxnθg(y)dy)qdx)1q(12)γ(R+n|x|(α+γ-θ)q(B|x|/2n-1g(y)dy)qdx)1qC(R+ngp(y)|y|-βpdy)1p,

where B|x|/2n-1={yR+n:|y|<R}. Hence, inequality (1.11) holds. If α<-nq+θ, we take W(x)=|x|(α-θ)q and U(y)=|y|-(γ+β)p. Through the same discussion as above, we can also prove that the inequality holds for the case α<-nq+θ. Thus, we complete the proof of Theorem 1.1. ∎

3 The Proof of Theorem 1.3

In this section, we shall use the method in spirit of Lieb’s works in [31] to establish the existence of extremals of reverse inequality (1.11) involved with the Poisson-type. Let ARn be a Borel set of finite Lebesgue measure. The symmetric rearrangement A* of the set 𝐴 is the open ball centered at the origin whose volume is that of 𝐴. The decreasing rearrangement of 𝑓 is the function f*:A*R defined by

f*(x)=inf{α:|{y:|f(y)|α}|vn|x|n},

where vn is the volume of the unit ball in Rn.

Through [3, 7, 11, 32], we derive the following lemma.

Lemma 3.1

The following statements hold.

  1. If g(y) is radially symmetric, then we have that Vγ,θ(g)(x,xn) is radially symmetric with respect to x, where x=(x,xn)R+n.

  2. For any nonnegative function gLp(R+n), we have Vγ,θ(g)Lq(R+n)Vγ,θ(g*)|Lq(R+n).

We also need the following lemma.

Lemma 3.2

Under the same hypothesis as Theorem 1.3, if we furthermore assume that gLp(R+n) is nonnegative, radially symmetric with g(r)εr-n-1p, then, for any 0<t<p, there holds

Vγ,θ(g)Lq(R+n)Cε1-ptgLp(R+n)pt.

Proof

Define G:RR by G(v)=ev(n-1)pg(ev) and H:R×R+R by H(u,xn)=eunqVγ,θ(g)(eu,euxn). We can easily verify that

(3.1)wn-21pGLp(R)=gLp(R+n),GL(R)ε,
(3.2)wn-21qHLq(R+2)=Vγ,θ(g)Lq(R+n).
Through calculation, we can get

H(u,xn)=eunqR+n|e2u+e2uxn2|α2(|eu-y|2+e2uxn2)γ2(euxn)θg(y)|y|βdy=eu(nq+α)R+n|1+xn2|α2(e2u-2euy+|y|2+e2uxn2)γ2(euxn)θg(y)|y|βdy=eu(nq+α)-+Sn-2|1+xn2|α2(e2u-2euev+e2v+e2uxn2)γ2(euxn)θg(ev)ev(n-1+β)dξdv=eu(nq+α+γ2)-+Sn-2|1+xn2|α2(eu-v(1+xn2)-21ξ+ev-u)γ2(euxn)θg(ev)ev(n-1+γ2+β)dξdv,

where eu denotes the vector sitting on R+n with length eu, so do ev and 1. It follows from

nq+n-1+α+β+γ-θ=n-1p

that 𝐻 can be written as

H(u,xn)=-+Ln,α,β(u-v,xn)G(v)dv,

where Ln,α,β(u,xn) is given by

Ln,α,β(u,xn)={eu(nq+α+γ2-θ)|1+xn2|α2xn-θ((eu(1+xn2)-2+e-u)γ2+(eu(1+xn2)+2+e-u)γ2),n=2,eu(nq+α+γ2-θ)Sn-2|1+xn2|α2xn-θ(eu(1+xn2)-21ξ+e-u)γ2dξ,n3.

We can check that

Ln,α,β(u,xn)eu(nq+α+γ2-θ)|1+xn2|α2xn-θ(eu(1+xn2)+e-u)γ2

as u2+xn2+. From the assumption of Theorem 1.3, we derive that, for any s<0,

(3.3)R(0Lq(u,xn)dxn)sqdu<+.

For any 0<t<p, we can choose s1<0 such that 1t+1s1=1+1q. Through the reverse Young inequality, we obtain

R|H(u,xn)|qdu(R|L(u,xn)|s1du)qs1(R|G|tdu)qt.

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

R+2|H(u,xn)|qdudxn(R|G|tdu)qt0(RLs1(u,xn)du)qs1dxn(R|G|tdu)qt(R(0Lq(u,xn)dxn)s1qdu)qs1GL(R)q(1-pt)(R|G|pdu)qtεq(1-pt)(R+n|g|pdy)qt.

Through (3.2), we derive that

Vγ,θ(g)Lq(R+n)Cε1-ptgLp(R+n)pt.

Then the proof of Lemma 3.2 is completed. ∎

Now we turn to prove Theorem 1.3.

Proof

According to Lemma 3.1, we can deduce that if 𝑔 is a minimizer for problem (1.12), 𝑔 must be radially decreasing. Likewise, if {gj}j is a minimizing sequence satisfying

limjVγ,θ(gj)Lq(R+n)=Cn,γ,θ,α,β,p.q

with gjLp(R+n)=1, without loss of generality, we can assume that {gj}j is nonnegative, radially symmetric and monotonously decreasing.

For any R>0, we obtain

wn-2n-1gjp(R)Rn-1ωn-20Rgjp(r)rn-2drωn-20+gjp(r)rn-2dr=R+ngjp(x)dx=1,

which implies that there exists some constant 𝐶 independent of 𝑗 such that, for any R>0,

(3.4)0gj(R)CR-n-1p.

Set aj=supr>0rn-1pgj(r). According to Lemma 3.2, we can deduce that aj2c0 for some c0>0. Select γj>0 satisfying γjn-1pgj(γj)>c0, and let g~j(y)=γjn-1pgj(γjy). Then g~jLp=gjLp=1, Vγ,θg~jLq=Vγ,θgjLq. 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.

Lieb’s argument based on the Helly theorem allow us to choose a subsequence such that gjg a.e. in R+n. It is obvious that 𝑔 is nonnegative and radially decreasing.

We are left to prove that

Vγ,θ(g)Lq(R+n)=Cn,γ,θ,α,β,p.qandgLp(R+n)=1.

According to the definition of Vγ,θ(gj)(x), for all xR+n and jN*, we have

Vγ,θ(gj)(x)c0|x|α|y|1|x-y|γxnθ|y|βdyF(x),

where

F(x){|x|αxn-θif|x|1,|x|γ+αxn-θif|x|>1.

For each xR+n, set k(x)=lim infj(Vγ,θgj(x)). Fatou’s lemma gives that

(3.5)R+nFq(x)-kq(x)dx=R+nlim infj[Fq(x)-(Vγ,θgj)q(x)]lim infjR+nFq(x)-(Vγ,θgj)q(x)dx=R+nFq(x)dx-(Cn,γ,θ,α,β,p,q)q.

Since 1q=n-1n1p-α+β+γ-1-θn-1, β<1-np, α<-n-1q+θ, we have R+nFq(x)dx<. Combining this with (3.5), we derive that

(3.6)R+nFq(x)dxR+nkq(x)dx(Cn,α,β,p,q)q,

which implies that the set {xR+n:0<k(x)<+} has positive measure. Hence, we can take a point x1R+n and extract a subsequence of Vγ,θgj, still denoted by Vγ,θgj, satisfying limjVγ,θgj(x1)=a1(0,+). We can employ the reverse Hölder inequality to obtain

R+n|x1|αP(x1,γ,θ,y)gj(y)|y|βdyCα,γ|y||x1|2gj(y)|y|βdyCα,γ(|y||x1|2gjτ(y)dy)1τ(|y||x1|2|y|τβdy)1τ,

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,

(3.7)|y||x1|2gjτ(y)dy1.

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

|y|Rgjτ(y)dy=|y||x1|2gjτ(y)dy+|x1|2|y|Rgjτ(y)dyC(R).

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

R+nP(x,γ,θ,y)gj(y)|y|βdyRγ34R|y|Rgj(y)|y|βdyRγgj(R)34R|y|R|y|βdyRγ+n-1+βgj(R).

Thus, we obtain gj(R)R1-n-γ-β. Because of 0<α<-n-1q+θ, we have

(3.8)limRlimj|y|Rgjp(y)dy=0.

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

limjR+ngjp(y)dy=limRlimj|y|Rgjp(y)dy+limRlimj|y|Rgjp(y)dy=limRlimj|y|Rgjp(y)dy=limR|y|Rgp(y)dy=R+ngp(y)dy,

which implies that gLp(R+n)=1. By Fatou’s lemma, we have

(3.9)(R+nkq(x)(x)dx)1q=(R+nlim infj(Vγ,θgj)q(x)dx)1q(R+n(Vγ,θg)q(x)dx)1q.

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

Cn,γ,θ,α,β,p,q(R+nkq(x)(x)dx)1q(R+n(Vγ,θg)q(x)dx)1qCn,γ,θ,α,β,p,q,

which implies that Vγ,θgLq(R+n)=Cn,γ,θ,α,β,p,q. Thus, we complete the proof of Theorem 1.3. ∎

4 The Proof of Theorem 1.4

Lemma 4.1

Let α,β,p1,p2,γ>0, and let (u,v) be a pair of positive Lebesgue measurable solutions of (1.14). Then the following holds.

  1. There exist some constants C1,C21 such that, for xR+n, yR+n{0}, we have

    1C1(1+|x|γ)u(x)xnθ|x|αC1(1+|x|γ),
    (4.1)1C2(1+|y|γ)v(y)|y|βC2(1+|y|γ).

  2. We have

    (4.2)R+n(1+|y|γ)v-p2(y)|y|βdy<,
    (4.3)R+n(1+|x|γ)xn-θu-p1(x)|x|αdx<.

Proof

Since (u,v) is a pair of positive Lebesgue measurable solutions of (1.14), it follows that

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

Then there exist R>1 and some measurable set 𝐸 such that E{yR+n:v(y)<R}BRn-1 satisfying |E|>1R. Write E=E1E2, where E1=E{y:|x-y||y|}, E2=E{y:|x-y||y|}. When 0<|x|2R, we have

u(x)xnθ|x|α(1+|x|γ)11+(2R)γR-p2E|x-y|γ|y|βdy11+(2R)γR-p2E1|x-y|γ+βdy+11+(2R)γR-p2E2|y|γ+βdy11+(2R)γR-p2E1*|y|γ+βdy+11+(2R)γR-p2E2*|y|γ+βdyC(n,γ,β)1+(2R)γR-p2|E1|1+γ+βn-1+C(n,γ,β)1+(2R)γR-p2|E2|1+γ+βn-1c01+(2R)γR-p2-1-γ+βn-1.

When |x|>2R>2, we have |x-y||x|-|y||x|214(1+|x|) for yE. Then

u(x)xnθ|x|α=R+n|x-y|γv-p2(y)|y|βdyCE(1+|x|γ)v-p2(y)|y|βdyC(1+|x|γ)R-p2E|y|βdyCR(1+|x|γ).

Thus, for any xR+n,

u(x)xnθ|x|α1C1(1+|x|γ).

A similar argument shows that, for any yR+n{0}, we get

(4.5)v(y)|y|β1C2(1+|y|γ).

By (4.4), there exists some x¯R+n such that

u(x¯)x¯nθ=R+n|x¯|α|x¯-y|γv-p2(y)|y|βdy<+.

Then

(4.6)R+n(1+|y|γ)v-p2(y)|y|βdyCx¯|y|<12|x¯||x¯-y|γv-p2(y)|y|βdy+Cx¯|y|>2|x¯||x¯-y|γv-p2(y)|y|βdy+12|x¯||y|2|x¯|(1+|y|γ)v-p2(y)|y|βdy.

This together with (4.5) implies (4.2).

For xR+n, with the help of (4.6), we can verify that

u(x)xnθ|x|α(1+|x|γ)=R+n|x-y|γ(1+|x|γ)v-p2(y)|y|βdyR+n(1+|y|γ)v-p2(y)|y|βdy<+.

Likewise, we can get (4.1) and (4.3). Then Lemma 4.1 is completed. ∎

Now we continue the proof of Theorem 1.4.

Proof

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

R+n|x-y|γv-p2(y)|y|βdyCγR+n(1+|y|γ)v-p2(y)|y|βdy<,
R+n|x-y|γxn-θu-p1(x)|x|αdxCγR+n(1+|x|γ)xn-θu-p1(x)|x|αdx<.
The dominated convergence theorem applied to 𝑢 and 𝑣 shows that
lim|x|0u(x)xnθ|x|α=lim|x|0R+n|x-y|γv-p2(y)|y|βdy=R+nv-p2(y)|y|γ+βdy,
lim|y|0v(y)|y|β=lim|y|0R+n|x-y|γxn-1u-p1(x)|x|αdx=R+nu-p1(x)|x|γ+αxn-θdx,
which give asymptotic estimates of 𝑢, 𝑣 around the origin.

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

|x|-γR+n|x-y|γ|y|βv-p2(y)dyCγR+n(1+|y|γ)v-p2(y)|y|βdy<,
|y|-γR+n|x-y|γxn-θu-p1(x)|x|αdxCγR+n(1+|x|γ)xn-θu-p1(x)|x|αdx<.
Then it follows from the dominated convergence theorem that

lim|x|u(x)xnθ|x|γ+α=R+nv-p2(y)|y|βdyandlim|y|v(y)|y|γ+β=R+nu-p1(x)xn-θ|x|αdx.

This accomplishes the proof of part II. ∎

5 The Proof of Theorem 1.5

In this section, we will give the regular estimate for positive solutions of equations (1.16). We shall prove that if (u,v) is a pair of positive solutions of system (1.16), then (u,v)C1(R+n)×C1(R+n).

We just show u(x)C1(R+n). For x0R+n, we take δ=12dist{x0,R+n}.

Step 1. We first prove that u(x) is continuous at x0. It is easy to obtain that, for xBδ(x0), there exists a constant C(x0,γ,δ,θ) such that P(x,γ,θ,y)v-p2(y)(1+|y|γ)v-p2(y). From Lemma 4.1, we know that

R+n(1+|y|γ)v-p2(y)dy<,
u(x)=R+nP(x,γ,θ,y)v-p2(y)dyC(x0,γ,δ,θ)R+n(1+|y|γ)v-p2(y)dy.
By the Lebesgue dominated convergence theorem, we have

limxx0u(x)=R+nlimxx0P(x,γ,θ,y)v-p2(y)dy=R+nP(x0,γ,θ,y)v-p2(y)dy.

Thus, we get that u(x) is continuous at x0.

Step 2. We prove that, when γ>1, u(x)C1(R+n). For xBδ(x0), we have

|P(x,γ,θ,y)xi|=|γ|x-y|γ-2(xi-yi)xnθ|γ|x-y|γ-1δθC(x0,γ,δ,θ)(1+|y|γ-1),i=1,,n-1,
|P(x,γ,θ,y)xn|=|γ|x-y|γ-2xnxnθ-θ|x-y|γxnθ+1|C(x0,γ,δ,θ)(1+|y|γ-1)+C(x0,γ,δ,θ)(1+|y|γ).
From Lemma 4.1, we have

R+n(1+|y|γ-1)v-p2(y)dy=|y|>1(1+|y|γ-1)v-p2(y)dy+|y|<1(1+|y|γ-1)v-p2(y)dy|y|>1(1+|y|γ)v-p2(y)dy+2|y|<1v-p2(y)dy<.

Likewise, By the Lebesgue dominated convergence theorem, for i=1,,n, u(x)xi|x=x0 exists and u(x)xi|x=x0 is continuous at x0.

Step 3. We prove that, when 0<γ<1, for i=1,,n-1, u(x)xi|x=x0 exists and u(x)xi is continuous at x0. According to the definition of derivative, for i-1,,n-1, we have

u(x)xi|x=x0=limh0u(x0+hei)-u(x0)h=limh0R+n|x0+hei-y|γ-|x0-y|γ(x0)nθhv-p2(y)dy,

where ei denotes the unit vector in the direction of xi. For 0<γ<1, A,B>0, we have the inequality

(5.1)|Aγ-Bγ||B|γ-1|A-B|,

which implies

||x0+hei-y|γ-|x0-y|γ(x0)nθh|v-p2(y)|x0-y|γ-1||x0+hei-y|-|x0-y||(x0)nθ|h|v-p2(y)|x0-y|γ-1|hei|(x0)nθ|h|v-p2(y)=|x0-y|γ-1(x0)nθv-p2(y).

From Lemma 4.1, we know that

R+n|x0-y|γ-1(x0)nθv-p2(y)dy=|x0-y|<1|x0-y|γ-1(x0)nθv-p2(y)dy+|x0-y|>1|x0-y|γ-1(x0)nθv-p2(y)dy|x0-y|<1|x0-y|γ-1(x0)nθdy+|x0-y|>1v-p2(y)(x0)nθdy<.

By the Lebesgue dominated convergence theorem, we can get

u(x)xi|x=x0=R+nlimh0|x0+hei-y|γ-|x0-y|γ(x0)nθhv-p2(y)dy=R+nγ|x0-y|γ-2((x0)i-yi)(x0)nθv-p2(y)dy.

So u(x) is differentiable for xi at x0, i=1,,n-1.

Set

uiϵ(x)=R+nγ|x-y|γ-2(xi-yi)xnθv-p2(y)η(x-yϵ)dy,

where

η(t)={0if|t|1,01if 1<|t|<2,1if|t|2.

We claim that uiϵ(x) is continuous. In fact, for fixed x0,

limxx0uiϵ(x)=limxx0R+nγ|x-y|γ-2(xi-yi)xnθv-p2(y)η(x-yϵ)dy=limxx0|x-y|>ϵγ|x-y|γ-2(xi-yi)xnθv-p2(y)η(x-yϵ)dy.

For xBδ(x0), we have

||x-y|γ-2|xi-yi|xnθv-p2(y)η(x-yϵ)|ϵγ-1δθ(1+|y|)v-p2(y).

From Lemma 4.1, we have

R+n(1+|y|)v-p2(y)<.

From the Lebesgue dominated convergence theorem, we deduce that uiϵ(x) is continuous.

On the other hand, we can prove that uiϵ(x) converges uniformly to u(x)xi on Bδ(x0) as ϵ0. In fact, for all xBδ(x0), we can write

|uxi(x)-uiϵ(x)|=|x-y|2ϵγ|x-y|γ-2(xi-yi)xnθv-p2(y)(1-η(x-yϵ))dy|x-y|2ϵγ|x-y|γ-2|xi-yi|xnθv-p2(y)dy|x-y|2ϵγ|x-y|γ-1δθv-p2(y)dy|x-y|2ϵγ|x-y|γ-1δθC(1+|y|γ)-p2dy|x-y|2ϵγ|x-y|γ-1δθCdy=Cγ(2ϵ)γ+n-2δθ.

So uiϵ(x) converges uniformly to u(x)xi on Bδ(x0) as ϵ0. From the fact that uiϵ(x) is continuous, we can deduce that u(x)xi is continuous in Bδ(x0). Thus, u(x)xi is continuous at x0 for i=1,,n-1.

Step 4. We prove that, when 0<γ1, u(x)xn|x=x0 exists and u(x)xn is continuous at x0. From the definition of derivative, we get

u(x)xn=limh0u(x+hen)-u(x)h=limh0R+n(|x+hen-y|γ(xn+h)θ-|x-y|γxnθ)v-p2(y)dyh=limh0R+n|x+hen-y|γ-|x-y|γ(xn+h)θhv-p2(y)dy-limh0R+n(xn+h)θ|x-y|γ-|x-y|γxnθ(xn+h)θxnθhv-p2(y)dy=h1(x)-h2(x).

We first discuss h1(x). For fixed x0 and |h|<12(x0)n, through (5.1), we have

| | x 0 + h e n - y | γ - | x 0 - y | γ ( ( x 0 ) n + h ) θ h | v - p 2 ( y ) | x 0 - y | γ - 1 | | x 0 + h e n - y | - | x 0 - y | | | ( x 0 ) n + h | θ | h | v - p 2 ( y ) 2 θ | x 0 - y | γ - 1 ( x 0 ) n θ v - p 2 ( y )

Using the similar arguments of step 3, we can get

h1(x0)=R+nlimh0|x0+hen-y|γ-|x0-y|γ((x0)n+h)θhv-p2(y)dy=γR+n|x0-y|γ-2(x0)n(x0)nθv-p2(y)dy

and that h1(x) is continuous at x0.

Now we discuss h2(x). For xBδ(x0) and some t(0,1), we have

|(xn+h)θ|x-y|γ-|x-y|γxnθ(xn+h)θxnθh|v-p2(y)=|θ(xn+th)θ-1|x-y|γ(xn+h)θxnθh|v-p2(y)C(x0,γ,δ,θ)(1+|y|γ)v-p2(y).

By Lemma 4.1 and the Lebesgue dominated convergence theorem, we derive

h2(x0)=R+nlimh0|x0+hen-y|γ-|x0-y|γ((x0)n+h)θhv-p2(y)dy=γR+n|x-y|γ-2(x0)n(x0)nθv-p2(y)dy

and that h2(x) is continuous at x0. Thus, we can get that, when 0<γ1, u(x) is differentiable for xn at x0 and u(x)xn is continuous at x0.

6 The Proof of Theorem 1.6

In this section, we shall give the proof of Theorem 1.6. We only need to prove that if (u,v) is a pair of positive solutions of the integral system

{u(x)=R+nP(x,γ,θ,y)v-p2(y)dy,xR+n,v(y)=R+nP(x,γ,θ,y)u-p1(x)dx,yR+n,

then the balanced condition np1-1+n-1p2-1=γ-θ must hold.

In view of Theorem 1.5, we know that (u,v)C1(R+n)×C1(R+n). Then integration by parts shows that

BR+u-p1(x)(xu(x))dx=11-p1BR+x(u1-p1(x))dx=11-p1BR+u1-p1(x)Rdσ-n1-p1BR+u1-p1(x)dx,
BRn-1v-p2(y)(yv(y))dy=11-p2BRn-1v1-p2(y)Rdσ-n-11-p2BRn-1v1-p2(y)dy.
Thanks to Lemma 4.1, we obtain

R+nu1-p1(x)dx<,R+nv1-p2(y)dy<.

Therefore, there exists Rj+ such that

RjBRj+u1-p1(x)dσ0,RjBRjn-1v1-p2(y)dσ0.

Then we derive

(6.1)R+nu-p1(x)(xu(x))dx+R+nv-p2(y)(yv(y))dy=-n1-p1R+nu1-p1(x)dx-n-11-p2R+nv1-p2(y)dy.

In view of integral system (1.14),

v(y)y=dv(ρy)dp|ρ=0=γR+nP(x,γ,θ,y)|y-x|-2(y-x)yu-p1(x)dx,
u(x)x=du(ρx)dρ|ρ=0=γR+nP(x,γ,θ,y)|x-y|-2(x-y)xv-p2(y)dy-θR+nP(x,γ,θ,y)v-p2(y)dy.
Then it follows that

R+nu-p1(x)(xu(x))dx+R+nv-p2(y)(yv(y))dy=(γ-θ)R+nR+nP(x,γ,θ,y)u-p1(x)v-p2(y)dydx=(γ-θ)R+nu1-p1(x)dx=(γ-θ)R+nv1-p2(y)dy.

This together with (6.1) implies that np1-1+n-1p2-1=γ-θ. Then we accomplish the proof of Theorem 1.6.

7 The Proof of Theorem 1.7

In this section, we will give the accurate forms of extremals for inequality (1.11) in the conformal index q=2n2n+γ-2θ, p=2(n-1)2n+γ-2. Obviously, the extremals (f,g) of inequality (1.15) in the conformal index under the suitable transformation u=c1fq-1, v=c2gp-1 satisfies the conformal integral equation

(7.1){u(x)=R+nP(x,γ,θ,y)v-2n+γ-2γ(y)dy,xR+n,v(y)=R+nP(x,γ,θ,y)u-2n+γ-2θγ-2θ(x)dx,yR+n.

Therefore, in order to give the accurate forms of extremals (f,g), it suffices to prove the following proposition.

Proposition 7.1

Assume that (u,v) is a pair of positive solutions of the conformal integral equation (7.1). Then v(y) must take the form as

v(y)=(d2+|y-y0|2c)γ2,yR+n,

for some y0R+n, c>0 and d>0.

Proof

We divide the proof into five steps. We first give some notations. Let 𝑢 and 𝑣 be positive functions defined on R+n×R+n. For y0R+n and λ>0, we define

uλ,y0(x)=(|x-y0|λ)γ-2θu(xλ,y0),vλ,y0(y)=(|y-y0|λ)γv(yλ,y0),

where (x,y)R+n×R+n and

xλ,y0=y0+λ2(x-y0)|x-y0|2,yλ,y0=y0+λ2(y-y0)|y-y0|2.

For R>0, define

BR(y0)={xRn:|x-y0|<R},
BRn-1(y0)={yR+n:|y-y0|<R},
BR+(y0)={x=(x1,x2,,xn)BR(y0):xn>0}.
For y0=0, we write BR=BR(0), BRn-1=BRn-1(0), BR+=BR+(0).

Step 1. Prove that, for any y0R+n, there exists λ0(y0)>0 such that, for any λ(0,λ0), there holds

uλ,y0(x)u(x),vλ,y0(y)v(y)

for xR+nBλ+(y0), yR+nBλn-1(y0).

From Theorem 1.5, we know that (u,v)C1(R+n)×C1(R+n). Then it follows that there exists s0>0 such that

y(|y-y0|-γ2v(y))(y-y0)<0for allyBs0n-1(y0),

which implies that

vλ,y0(y)>v(y)for allyBs0n-1(y0)Bλn-1(y0).

On the other hand, a direction application of Lemma 4.1 gives that, for small λ0(0,s0) and 0<λ<λ0, there holds

(7.2)vλ,y0(y)=(|y-y0|λ)γv(λ2(y-y0)|y-y0|2)(|y-y0|λ0)γinfBs0n-1(y0)vv(y)for allyRnBs0n-1(y0).

This together with (7.2) yields vλ,y0(y)v(y) for yR+nBλn-1(y0). Similarly, we can also derive that uλ,y0(x)u(x) for xR+nBλ+(y0).

Step 2. For each y0R+n, we could define

λ¯(y0)=sup{μ:uλ,y0(x)u(x),vλ,y0(y)v(y)for all 0<λ<μ,xR+nBλ+(y0),yR+nBλn-1(y0)}.

It follows from step 1 that λ¯(y0) is well defined. Now we prove that, for any y0R+n, λ¯(y0)<+.

In fact, it follows from the definition of λ¯(y0) that, for any 0<λ<λ¯(y0), there holds

(7.3)uλ,y0(x)u(x),vλ,y0(y)v(y)for allxR+nBλ+(y0),yR+nBλn-1(y0),

which together with (7.3) and Theorem 1.4 implies that

0<R+nv-2n+γ-2γ(y)dy=lim|x||x|-γu(x)xnθlim|x||x|-γ(|x-y0|λ)γu(xλ,y0)(xλ,y0)nθ=λ-γlimx(y0,0)u(x)xnθ=λ-γR+nv-2n+γ-2γ(y)|y-y0|γdy.

This gives λ¯(y0)<.

Step 3. From the definition of λ¯, we see that there are the following four cases:

  1. uλ¯,y0(x)=u(x),vλ¯,y0(y)=v(y),

  2. uλ¯,y0(x)>u(x),vλ¯,y0(y)=v(y),

  3. uλ¯,y0(x)=u(x),vλ¯,y0(y)>v(y),

  4. uλ¯,y0(x)>u(x),vλ¯,y0(y)>v(y).

Now we prove that case (ii) and case (iii) are impossible.

Let (u,v) be a pair of positive Lebesgue measurable solutions for integral system (7.1). Through the identity

|x|x|2-y|y|2|=|x-y||x||y|,

we have

|xλ,y0-yλ,y0|=λ2|x-y||x-y0||y-y0|and|x-yλ,y0|=|(xλ,y0)λ,y0-yλ,y0|=|x-y0||xλ,y0-y||y-y0|.

Then, for xR+n, yR+n, straight calculations give that

(7.4)uλ,y0(x)-u(x)=R+nBλn-1(y0)K(y0,λ,θ,x,y)(v-2n+γ-2γ(y)-vλ,y0-2n+γ-2γ(y))dy,
(7.5)vλ,y0(y)-v(y)=R+nBλ+(y0)K(y0,λ,θ,y,x)(u-2n+γ-2θγ-2θ(x)-uλ,y0-2n+γ-2θγ-2θ(x))dx,
where
K(y0,λ,θ,x,y)=(|x-y0|λ)γ|xλ,y0-y|γxnθ-|x-y|γxnθ,
K(y0,λ,θ,y,x)=(|y-y0|λ)γ|yλ,y0-x|γxnθ-|x-y|γxnθ.
Furthermore, we also show that K(y0,λ,θ,x,y),K(y0,λ,θ,y,x)>0 when |x-y0|,|y-y0|>λ. In fact, we may assume y0=0 and get

K(0,λ,θ,x,y)=(|x|λ)γ|xλ,0-y|γ1xnθ-|x-y|γ1xnθ=(λ2-2xy+|x|2|y|2λ2)γ21xnθ-(|x|2-2xy+|y|2)γ21xnθ=((|x|2-2xy+|y|2+(|x|2-λ2)(|y|2λ2-1))γ2-(|x|2-2xy+|y|2)γ2)1xnθ>0

if |x|,|y|>λ. Similarly, we can also derive that K(0,λ,θ,x,y)>0 for |x|,|y|>λ. Then (ii) and (iii) being impossible is a direct result of (7.4) and (7.5).

Step 4. Now we start to prove that case (iv) is also impossible. For convenience, we may assume y0=0 and use notations λ¯=λ¯(0), uλ=uλ,0 and xλ=xλ,0, yλ=yλ,0. We carry out the argument by contradiction. Suppose that case (iv) is valid. We will show that there exists an ε>0 such that

uλ(x)u(x),vλ(y)v(y)for allλ¯<λ<λ¯+εandxR+nBλn-1,yR+nBλ+.

By (7.4) and (7.5), with y0=0, λ=λ¯ and the Fatou lemma, one can calculate

lim|x|inf|x|-γ(uλ¯(x)-u(x))=lim|x|inf|x|-γR+nBλ¯n-1K(0,λ¯,θ,x,y)(v-2n+γ-2γ-vλ¯-2n+γ-2γ)dy,
lim|y|inf|y|-γ(vλ¯(y)-v(y))=lim|y|inf|y|-γR+nBλ¯+K(0,λ¯,θ,y,x)(u-2n+γ-2θγ-2θ-uλ¯-2n+γ-2θγ-2θ)dx.
Then it follows that
lim|x|inf|x|-γxnθ(uλ¯(x)-u(x))R+nBλ¯n-1((|y|λ¯)γ-1)(v-2n+γ-2γ-vλ¯-2n+γ-2γ)dy>0,
lim|y|inf|y|-γ(vλ¯(y)-v(y))R+nBλ¯+((|x|λ¯)γ-1)1xnθ(u-2n+γ-2θγ-2θ-uλ-2n+γ-2θγ-2θ)dx>0.
Then, for any xR+nBλ¯+1+ and yR+nBλ¯+1n-1, there exists ε1(0,1) such that

uλ¯(x)-u(x)ε1|x|γxn-θ,vλ¯(y)-v(y)ε1|y|γ.

We can use continuity of 𝑢 and 𝑣 with respect to variable 𝜆 to obtain

uλ(x)-u(x)ε1|x|γxnθ+(uλ(x)-uλ¯(x))ε12|x|γxnθfor allxR+nBλ¯+1+,λ¯λλ¯+ε2,
(7.6)vλ(y)-v(y)ε1|y|γ+(vλ(y)-vλ¯(y))ε12|y|γfor allyR+nBλ¯+1+,λ¯λλ¯+ε2,
for sufficiently small ε2. Thus, it suffices to verify that, for all λ¯<λ<λ¯+ϵ,
uλ(x)u(x)forxBλ¯+1+Bλ+,
vλ(y)v(y)foryBλ¯+1n-1Bλn-1.

We only show that uλ(x)u(x) for all λ¯<λ<λ¯+ϵ. By step 3, we can write

(7.7)uλ(x)-u(x)=R+nBλn-1K(0,λ,θ,x,y)(v-2n+γ-2γ(y)-vλ,0-2n+γ-2γ(y))dyBλ¯+1n-1Bλn-1K(0,λ,θ,x,y)(v-2n+γ-2γ(y)-vλ,0-2n+γ-2γ(y))dy+Bλ¯+3n-1Bλ¯+2n-1K(0,λ,θ,x,y)(v-2n+γ-2γ(y)-vλ,0-2n+γ-2γ(y))dy.

One can employ (7.6) to derive that

(7.8)vλ(y)-v(y)ε2|y|γε1(λ¯+2)γ

for yBλ¯+3n-1Bλ¯+2n-1. Obviously, there exists δ1>0 dependent of λ¯ such that

v-2n+γ-2γ(y)-vλ,0-2n+γ-2γ(y)>v-2n+γ-2γ(y)-(v+ε1(λ¯+2)γ)-2n+γ-2γδ1.

For any xBλ¯+1+Bλ+ and yBλ¯+3n-1Bλ¯+2n-1, there exist δ~2, δ2>0 such that

(7.9)K(0,λ,θ,x,y)=xn-θ((|x|λ)γ|λ2x|x|2-y|γ-|y-x|γ)=xn-θ((|x|λ)2|λ2x|x|2-y|2)γ2-xn-θ(|y-x|2)γ2=δ~2xn-θ((|x|λ)2|λ2x|x|2-x|2-|y-x|2)δ2xn-θ(|x|-λ).

For any λ¯λλ¯+ε and yBλ¯+1n-1Bλn-1, direct calculations yield

vλ(y)-vλ¯(y)=(|y|λ)γv(λ2y|y|2)-(|y|λ¯)γv(λ¯2y|y|2)=|y|γv(λ2y|y|2)λγλ¯γ(λ¯γ-λγ)+(|y|λ¯)γv(y)(λ2y|y|2-λ¯2y|y|2),

which gives

(7.10)|vλ¯-2n+γ-2γ(y)-vλ-2n+γ-2γ(y)|C1ε,

where y(λ2y|y|2,λ¯2y|y|2).

For xBλ¯+1+Bλ+, we obtain

(7.11)Bλ¯+1n-1Bλn-1K(0,λ,θ,x,y)dy=Bλ¯+1n-1Bλn-1xn-θ(|x|λ)γ|λ2x|x|2-y|γ-xn-θ|x-y|γdyCxn-θ(|x|-λ)+Bλ¯+1n-1Bλn-1xn-θ|λ2x|x|2-y|γ-xn-θ|x-y|γdyC2xn-θ(|x|-λ).

By (7.7), (7.8), (7.9), (7.10) and (7.11), for sufficiently small ε>0, we conclude that

uλ(x)-u(x)xn-θ(δ1δ2Bλ¯+3n-1Bλ¯+2n-1dy-C1C2ε)(|x|-λ)0,

which leads to a contradiction with the definition of λ¯. This proves that case (iv) is also impossible.

Step 5. From the above analysis, we know that only (i) can happen. By a calculus lemma [29, Lemma 11.1], we know that any C1 positive solution 𝑣 must take the form

v(y)=(d2+|y-y0|2c)γ2,yR+n,

for some y0R+n, c>0 and d>0. Subsequently, the extremal function 𝑔 of inequality (1.18) must be of the form

g(y)=c(1|y-y0|2+d2)2n+γ-22

for some c>0, d>0 and y0R+n. This accomplishes the proof of Theorem 1.7. ∎

  1. Communicated by: Guozhen Lu

References

[1] W. Beckner, Sharp Sobolev inequalities on the sphere and the Moser–Trudinger inequality, Ann. of Math. (2) 138 (1993), no. 1, 213–242. 10.2307/2946638Search in Google Scholar

[2] 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

[3] H. J. Brascamp and E. H. Lieb, Best constants in Young’s inequality, its converse, and its generalization to more than three functions, Adv. Math. 20 (1976), no. 2, 151–173. 10.1007/978-3-642-55925-9_35Search in Google Scholar

[4] L. Caffarelli and L. Silvestre, An extension problem related to the fractional Laplacian, Comm. Partial Differential Equations 32 (2007), no. 7–9, 1245–1260. 10.1080/03605300600987306Search in Google Scholar

[5] T. Carleman, Zur Theorie der Minimalflächen, Math. Z. 9 (1921), no. 1–2, 154–160. 10.1007/BF01378342Search in Google Scholar

[6] E. A. Carlen and M. Loss, Extremals of functionals with competing symmetries, J. Funct. Anal. 88 (1990), no. 2, 437–456. 10.1016/0022-1236(90)90114-ZSearch in Google Scholar

[7] 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

[8] L. Chen, Z. Liu, G. Lu and C. Tao, Stein–Weiss inequalities with the fractional Poisson kernel, Rev. Mat. Iberoam. 36 (2020), no. 5, 1289–1308. 10.4171/rmi/1167Search in Google Scholar

[9] L. Chen, G. Lu and C. Tao, Existence of extremal functions for the Stein–Weiss inequalities on the Heisenberg group, J. Funct. Anal. 277 (2019), no. 4, 1112–1138. 10.1016/j.jfa.2019.01.002Search in Google Scholar

[10] L. Chen, G. Lu and C. Tao, Hardy–Littlewood–Sobolev inequalities with the fractional Poisson kernel and their applications in PDEs, Acta Math. Sin. (Engl. Ser.) 35 (2019), no. 6, 853–875. 10.1007/s10114-019-8417-2Search in Google Scholar

[11] L. Chen, G. Lu and C. Tao, Reverse Stein–Weiss inequalities on the upper half space and the existence of their extremals, Adv. Nonlinear Stud. 19 (2019), no. 3, 475–494. 10.1515/ans-2018-2038Search in Google Scholar

[12] S. Chen, A new family of sharp conformally invariant integral inequalities, Int. Math. Res. Not. IMRN 2014 (2014), no. 5, 1205–1220. 10.1093/imrn/rns248Search in Google Scholar

[13] W. Chen, C. Li and B. Ou, Classification of solutions for an integral equation, Comm. Pure Appl. Math. 59 (2006), no. 3, 330–343. 10.1002/cpa.20116Search in Google Scholar

[14] 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

[15] J. Dou, Q. Guo and M. Zhu, Subcritical approach to sharp Hardy–Littlewood–Sobolev type inequalities on the upper half space, Adv. Math. 312 (2017), 1–45. 10.1016/j.aim.2017.03.007Search in Google Scholar

[16] 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

[17] 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

[18] 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

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

[20] 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

[21] 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

[22] 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

[23] M. Gluck, Subcritical approach to conformally invariant extension operators on the upper half space, J. Funct. Anal. 278 (2020), no. 1, Article ID 108082. 10.1016/j.jfa.2018.08.012Search in Google Scholar

[24] L. Gross, Logarithmic Sobolev inequalities, Amer. J. Math. 97 (1975), no. 4, 1061–1083. 10.2307/2373688Search 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] F. Hang, X. Wang and X. Yan, An integral equation in conformal geometry, Ann. Inst. H. Poincaré Anal. Non Linéaire 26 (2009), no. 1, 1–21. 10.1016/j.anihpc.2007.03.006Search in Google Scholar

[28] 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

[29] Y. Li and L. Zhang, Liouville-type theorems and Harnack-type inequalities for semilinear elliptic equations, J. Anal. Math. 90 (2003), 27–87. 10.1007/BF02786551Search in Google Scholar

[30] Y. Y. Li, Remark on some conformally invariant integral equations: the method of moving spheres, J. Eur. Math. Soc. (JEMS) 6 (2004), no. 2, 153–180. 10.4171/JEMS/6Search in Google Scholar

[31] 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

[32] 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

[33] 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

[34] 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: 2020-08-23
Accepted: 2020-10-13
Published Online: 2020-11-20
Published in Print: 2021-02-01

© 2020 Walter de Gruyter GmbH, Berlin/Boston

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

Downloaded on 13.5.2024 from https://www.degruyter.com/document/doi/10.1515/ans-2020-2112/html
Scroll to top button