This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

An extension problem and Hardy’s inequality for the fractional Laplace-Beltrami operator on Riemannian symmetric spaces of noncompact type

Mithun Bhowmik and Sanjoy Pusti (Mithun Bhowmik) Department of Mathematics, Indian Institute of Science, Bangalore-560012, India [email protected], [email protected] (Sanjoy Pusti) Department of Mathematics, IIT Bombay, Powai, Mumbai-400076, India [email protected]
Abstract.

In this paper we study an extension problem for the Laplace-Beltrami operator on Riemannian symmetric spaces of noncompact type and use the solution to prove Hardy-type inequalities for fractional powers of the Laplace-Beltrami operator. Next, we study the mapping properties of the extension operator. In the last part we prove Poincaré-Sobolev inequalities on these spaces.

Key words and phrases:
Hardy’s inequality, fractional Laplacian, extension problem, Riemannian symmetric spaces
2010 Mathematics Subject Classification:
Primary 43A85; Secondary 26A33, 22E30

1. Introduction

In recent years there has been intensive research on various kinds of inequalities for fractional order operators because of their applications to many areas of analysis (see for instance [8, 19, 39] and the references therein). The classical definitions of the fractional operator in terms of the Fourier analysis involve functional analysis and singular integrals. They are nonlocal objects. This fact does not allow to apply local PDE techniques to treat nonlinear problems for the fractional operators. To overcome this difficulty, in the Euclidean case, Caffarelli and Silvestre [11] studied the extension problem associated to the Laplacian and realised the fractional power as the map taking Dirichlet data to the Neumann data. On a certain class of noncompact manifolds, this definition of the fractional Laplacian through an extension problem has been studied by Banika et al. [6].

In the first part of this article we will concern with the Hardy-type inequalities for the fractional operators. Let Δn=j=1n2xj2\Delta_{\mathbb{R}^{n}}=\sum_{j=1}^{n}\frac{\partial^{2}}{\partial x_{j}^{2}} denote the Euclidean Laplacian on n\mathbb{R}^{n}. For 0<s<n/20<s<n/2 and fCc(n)f\in C_{c}^{\infty}(\mathbb{R}^{n}), the Hardy’s inequality for fractional powers of the Laplacian states the following

(1.1) n|f(x)|2|x|2s𝑑x4sΓ(n2s4)2Γ(n+2s4)2(Δn)sf,f.\int_{\mathbb{R}^{n}}\frac{|f(x)|^{2}}{|x|^{2s}}\,dx\leq 4^{-s}\frac{\Gamma\left(\frac{n-2s}{4}\right)^{2}}{\Gamma\left(\frac{n+2s}{4}\right)^{2}}\langle(-\Delta_{\mathbb{R}^{n}})^{s}f,f\rangle.

This is a generalization of the original Hardy’s inequality proved for the gradient n\nabla_{\mathbb{R}^{n}} of ff: for n3n\geq 3,

(1.2) (n2)24n|f(x)|2|x|2𝑑xn|nf(x)|2𝑑x, for fCc(n).\frac{(n-2)^{2}}{4}\int_{\mathbb{R}^{n}}\frac{|f(x)|^{2}}{|x|^{2}}dx\leq\int_{\mathbb{R}^{n}}|\nabla_{\mathbb{R}^{n}}f(x)|^{2}~{}dx,\>\>\textit{ for }f\in C_{c}(\mathbb{R}^{n}).

The constant appearing in the equation (1.1) is sharp [7, 27, 41]. It is also known that the equality is not obtained in the class of functions for which both sides of the inequality (1.1) are finite. Using a ground state representation, Frank, Lieb, and Seiringer gave a different proof of the inequality (1.1) when 0<s<min{1,n/2}0<s<\min\{1,n/2\} which improved the previous results [19]. There is another version of Hardy’s inequality where the homogeneous weight function |x|2s|x|^{-2s} is replaced by non-homogeneous one:

(1.3) n|f(x)|2(δ2+|x|2)2s𝑑x4sΓ(n2s4)Γ(n+2s4)δ2s(Δn)sf,f,δ>0.\int_{\mathbb{R}^{n}}\frac{|f(x)|^{2}}{(\delta^{2}+|x|^{2})^{2s}}~{}dx\leq 4^{-s}\frac{\Gamma\left(\frac{n-2s}{4}\right)}{\Gamma\left(\frac{n+2s}{4}\right)}~{}\delta^{-2s}~{}\langle(-\Delta_{\mathbb{R}^{n}})^{s}f,f\rangle,\>\>\delta>0.

Here also the constant is sharp and equality is achieved for the functions (δ2+|x|2)(n2s)/2(\delta^{2}+|x|^{2})^{-(n-2s)/2} and their translates [10].

Generalization of the classical Hardy’s inequality (1.2) to Riemannian manifolds was intensively pursued after the seminal work of Carron [12], see for instance [9, 18, 28, 29, 30, 42]. In [12], the following weighted Hardy’s inequality was obtained on a complete noncompact Riemannian manifold MM:

Mηα|gϕ|2𝑑vg(C+α12)2Mηαϕ2η2𝑑vg,\int_{M}\eta^{\alpha}|\nabla_{g}\phi|^{2}~{}dv_{g}\geq\left(\frac{C+\alpha-1}{2}\right)^{2}\int_{M}\eta^{\alpha}\frac{\phi^{2}}{\eta^{2}}~{}dv_{g},

where ϕCc(Mη1{0}),α,C>1,C+α1>0\phi\in C_{c}^{\infty}(M-\eta^{-1}\{0\}),~{}\alpha\in\mathbb{R},~{}C>1,~{}C+\alpha-1>0 and the weight function η\eta satisfies |Mη|=1|\nabla_{M}\eta|=1 and |ΔMη|C/η|\Delta_{M}\eta|\geq C/\eta in the sense of distribution. Here g,dvg\nabla_{g},dv_{g} denote respectively the Riemannian gradient and Riemannian measure on MM. In the case of Cartan-Hadamard manifold MM of dimension NN (namely, a manifold which is complete, simply-connected, and has everywhere non-positive sectional curvature), the geodesic distance function d(x,x0)d(x,x_{0}), where x0Mx_{0}\in M, satisfies all the assumptions of the weight η\eta and the above inequality holds with the best constant (N2)2/4(N-2)^{2}/4, see [30]. Analogues of Hardy-type inequalities for fractional powers of the sublaplacian are also known, for instance, the work by P. Ciatti, M. Cowling and F. Ricci for stratified Lie groups [14]. There the authors have not paid attention to the sharpness of the constants. Recently, in [37], Roncal and Thangavelu have proved analogues of Hardy-type inequalities with sharp constants for fractional powers of the sublaplacian on the Heisenberg group. For recent results on the Hardy-type inequalities for the fractional operators we refer [10, 36, 38].

Our first aim in this article is to prove analogues of Hardy’s inequalities (1.1) and (1.3) for fractional powers of the Laplace-Beltrami operator Δ\Delta on Riemannian symmetric space XX of noncompact type. We have the following analogue of Hardy’s inequality in the non-homogeneous case.

Theorem 1.1.

Let 0<σ<10<\sigma<1 and y>0y>0. Then there exists a constant Cσ>0C_{\sigma}>0 such that for FHσ(X)F\in H^{\sigma}(X)

(Δ)σF,FCσy2σ({x:|x|2+y2<1}|F(x)|2(y2+|x|2)2σ𝑑x+{x:|x|2+y21}|F(x)|2(y2+|x|2)σ𝑑x).\left\langle(-\Delta)^{\sigma}F,F\right\rangle\geq C_{\sigma}\,y^{2\sigma}\left(\int_{\left\{x:|x|^{2}+y^{2}<1\right\}}\frac{|F(x)|^{2}}{(y^{2}+|x|^{2})^{2\sigma}}\,dx+\int_{\left\{x:|x|^{2}+y^{2}\geq 1\right\}}\frac{|F(x)|^{2}}{(y^{2}+|x|^{2})^{\sigma}}\,dx\right).
Remark 1.2.

In contrast with the inequality (1.3) for the Euclidean space, we get an improvement in the theorem above. This comes as a consequence of the geometry of the symmetric space. In the following theorem also we get similar improvement.

For the homogeneous weight function, we prove the following analogue of Hardy’s inequality on XX.

Theorem 1.3.

Let 0<σ<10<\sigma<1. Then there exists a constant Cσ>0C^{\prime}_{\sigma}>0 such that for FCc(X)F\in C_{c}^{\infty}(X)

(Δ)σF,FCσ({x:|x|<1}|F(x)|2|x|2σ𝑑x+{x:|x|1}|F(x)|2|x|σ𝑑x).\left\langle(-\Delta)^{\sigma}F,F\right\rangle\geq C^{\prime}_{\sigma}\left(\int_{\left\{x:|x|<1\right\}}\frac{|F(x)|^{2}}{|x|^{2\sigma}}\,dx+\int_{\left\{x:|x|\geq 1\right\}}\frac{|F(x)|^{2}}{|x|^{\sigma}}\,dx\right).

Given σ(0,1)\sigma\in(0,1), the fractional Laplacian (Δn)σ(-\Delta_{\mathbb{R}^{n}})^{\sigma} on n\mathbb{R}^{n} is defined as a pseudo-differential operator by

((Δn)σf)(ξ)=|ξ|2σf(ξ),ξn,\mathcal{F}\left((-\Delta_{\mathbb{R}^{n}})^{\sigma}f\right)(\xi)=|\xi|^{2\sigma}\mathcal{F}f(\xi),\>\>\xi\in\mathbb{R}^{n},

where f\mathcal{F}f is the Fourier transform of ff given by

f(ξ)=(2π)n/2nf(x)eixξ𝑑x,ξn.\mathcal{F}f(\xi)=(2\pi)^{-n/2}\int_{\mathbb{R}^{n}}f(x)~{}e^{-ix\cdot\xi}~{}dx,\>\>\xi\in\mathbb{R}^{n}.

It can also be written as the singular integral

(Δn)σf(x)=cn,σP.V.nf(x)f(y)|xy|n+2σ𝑑y,(-\Delta_{\mathbb{R}^{n}})^{\sigma}f(x)=c_{n,\sigma}P.V.\int_{\mathbb{R}^{n}}\frac{f(x)-f(y)}{|x-y|^{n+2\sigma}}~{}dy,

where cn,σc_{n,\sigma} is a positive constant. Caffarelli and Silvestre have developed in [11] an equivalent definition of the fractional Laplacian (Δn)σ,σ(0,1)(-\Delta_{\mathbb{R}^{n}})^{\sigma},\sigma\in(0,1), using an extension problem to the upper half-space +n+1\mathbb{R}^{n+1}_{+}. For a function f:nf:\mathbb{R}^{n}\rightarrow\mathbb{R}, consider the solution u:n×[0,+)u:\mathbb{R}^{n}\times[0,+\infty)\rightarrow\mathbb{R} of the following differential equation

(1.4) Δnu+(12σ)yuy+2uy2=0,y>0;\displaystyle\Delta_{\mathbb{R}^{n}}u+\frac{(1-2\sigma)}{y}\frac{\partial u}{\partial y}+\frac{\partial^{2}u}{\partial y^{2}}=0,\>\>y>0;
u(x,0)=f(x),xn.\displaystyle u(x,0)=f(x),\>\>\>\>x\in\mathbb{R}^{n}.

Then the fractional Laplacian of ff can be computed as

(Δn)σf=22σ1Γ(σ)Γ(1σ)limy0+y12σuy.(-\Delta_{\mathbb{R}^{n}})^{\sigma}f=-2^{2\sigma-1}\frac{\Gamma(\sigma)}{\Gamma(1-\sigma)}\lim_{y\rightarrow 0^{+}}y^{1-2\sigma}\frac{\partial u}{\partial y}.

The Poisson kernel for the fractional Laplacian (Δn)σ(-\Delta_{\mathbb{R}^{n}})^{\sigma} in n\mathbb{R}^{n} is

Kσ(x,y)=cn,σy2σ(|x|2+y2)σ+n2,K_{\sigma}(x,y)=c_{n,\sigma}\frac{y^{2\sigma}}{(|x|^{2}+y^{2})^{\sigma+\frac{n}{2}}},

and then u(x,y)=fnKσu(x,y)=f\ast_{\mathbb{R}^{n}}K_{\sigma}. Therefore

(Δn)σf=22σ1Γ(σ)Γ(1σ)limy0+y12σy(fnKσ)(x).(-\Delta_{\mathbb{R}^{n}})^{\sigma}f=-2^{2\sigma-1}\frac{\Gamma(\sigma)}{\Gamma(1-\sigma)}\lim_{y\rightarrow 0^{+}}y^{1-2\sigma}\frac{\partial}{\partial y}(f\ast_{\mathbb{R}^{n}}K_{\sigma})(x).

Later, Stinga and Torrea [39] showed that one can define the fractional Laplacian on a domain Ωn\Omega\subset\mathbb{R}^{n} through the extension (1.4) using the heat-diffusion semigroup generated by the Laplacian ΔΩ\Delta_{\Omega} provided that the heat kernel associated with ΔΩ\Delta_{\Omega} exists and it satisfies some decay properties. Since the heat kernel on general noncompact manifolds has been extensively studied depending on the underlying geometry, Banica et al. in [6] take this approach to define the fractional Laplace-Beltrami operator on some noncompact manifolds which in particular, include the Riemannian symmetric spaces of noncompact type. Let 𝐝{\bf d} be a Riemannian metric on a Riemannian symmetric space XX and Δ\Delta be the corresponding Laplace-Beltrami operator on XX. Also, let 𝐠{\bf g} be the product metric on X×+X\times\mathbb{R}^{+} given by 𝐠=𝐝+dy2{\bf g}={\bf d}+dy^{2}. For σ>0\sigma>0, let Hσ(X)H^{\sigma}(X) denote the Sobolev space on XX (defined in Section 2). In [6, Theorem 1.1], the following result is proved for the Riemannian symmetric space XX of noncompact type of arbitrary rank.

Theorem 1.4.

((Banica; Gonźalez; Sáez)) Let σ(0,1)\sigma\in(0,1). Then for any given fHσ(X)f\in H^{\sigma}(X), there exists a unique solution of the extension problem

(1.5) Δu+(12σ)yuy+2uy2=0,y>0;\displaystyle\Delta u+\frac{(1-2\sigma)}{y}\frac{\partial u}{\partial y}+\frac{\partial^{2}u}{\partial y^{2}}=0,\>\>y>0;
u(x,0)=f(x),xX.\displaystyle u(x,0)=f(x),\>\>\>\>x\in X.

Moreover, the fractional Laplace-Beltrami operator on XX can be recovered through

(1.6) (Δ)σf(x)=22σ1Γ(σ)Γ(1σ)limy0+y12σuy(x,y).(-\Delta)^{\sigma}f(x)=-2^{2\sigma-1}\frac{\Gamma(\sigma)}{\Gamma(1-\sigma)}\lim_{y\rightarrow 0^{+}}y^{1-2\sigma}\frac{\partial u}{\partial y}(x,y).

The following theorem gives an alternative expression of a solution of the extension problem (1.5), which will be useful for us. The proof is similar to [39, Theorem 1.1]. See also [6, Theorem 3.1] . For the sake of completeness we give a proof in section 33.

Theorem 1.5.

Let fDom(Δ)σf\in Dom(-\Delta)^{\sigma}. A solution of (1.5) is given by

(1.7) u(x,y)=1Γ(σ)0etΔ(Δ)σf(x)ey2/4tdtt1σ,u(x,y)=\frac{1}{\Gamma(\sigma)}\int_{0}^{\infty}e^{t\Delta}(-\Delta)^{\sigma}f(x)e^{-y^{2}/4t}~{}\frac{dt}{t^{1-\sigma}},

and uu is related to (Δ)σf(-\Delta)^{\sigma}f by the equation (1.6). Moreover, the following Poisson formula for uu holds:

(1.8) u(x,y)=Xf(ζ)Pyσ(ζ1x)𝑑ζ=(fPyσ)(x),u(x,y)=\int_{X}f(\zeta)P_{y}^{\sigma}(\zeta^{-1}x)~{}d\zeta=(f\ast P_{y}^{\sigma})(x),

where

(1.9) Pyσ(x)=y2σ4σΓ(σ)0ht(x)ey2/4tdtt1+σ.P_{y}^{\sigma}(x)=\frac{y^{2\sigma}}{4^{\sigma}\Gamma(\sigma)}\int_{0}^{\infty}h_{t}(x)~{}e^{-y^{2}/4t}~{}\frac{dt}{t^{1+\sigma}}.

All these identities in theorem above are to be understood in the L2L^{2} sense. The mapping properties of the Poisson operator PσP_{\sigma} on n\mathbb{R}^{n} which maps boundary value ff to the solution uu of the extension problem (1.4) were studied by Möllers et al. [34]. In the same paper, the authors have also obtained a similar result for Heisenberg groups. On the Euclidean spaces, they proved the following

Theorem 1.6 (Möllers; Ø\Orsted; Zhang).

Let 0<σ<n20<\sigma<\frac{n}{2}. Then

  1. (1)

    Pσ:Hσ(n)Hσ+1/2(n×+)P_{\sigma}:H^{\sigma}(\mathbb{R}^{n})\rightarrow H^{\sigma+1/2}(\mathbb{R}^{n}\times\mathbb{R}^{+}) is isometric up to a constant.

  2. (2)

    PσP_{\sigma} extends to a bounded operator from Lp(n)L^{p}(\mathbb{R}^{n}) to Lq(n×+)L^{q}(\mathbb{R}^{n}\times\mathbb{R}_{+}), for 1<p1<p\leq\infty and q=n+1npq=\frac{n+1}{n}p (Figure 1, (a)).

In [13], Chen proved that for particular values p=2nn2σp=\frac{2n}{n-2\sigma} and q=2n+2n2σq=\frac{2n+2}{n-2\sigma}, there exists a sharp constant CC such that

PσfLq(n)CfLp(n), for fLp(n),\|P_{\sigma}f\|_{L^{q}(\mathbb{R}^{n})}\leq C\|f\|_{L^{p}(\mathbb{R}^{n})},\>\>\textit{ for }f\in L^{p}(\mathbb{R}^{n}),

and the optimizer of this inequality are translations, dilations and multiples of the function

f(x)=(1+|x|2)n2+σ.f(x)=\left(1+|x|^{2}\right)^{-\frac{n}{2}+\sigma}.

Our second main aim in this article is to study the mapping properties of the “Poisson operator” TσT_{\sigma} given by

(1.10) Tσf(x,y)=fPyσ,xX,y>0,T_{\sigma}f(x,y)=f\ast P_{y}^{\sigma},\>\>x\in X,\>y>0,

which maps ff to the solution uu of the extension problem (1.5) related to the Laplace-Beltrami operator on Riemannian symmetric spaces of noncompact type. The following analogue of Theorem 1.6 is our main result in this direction.

Theorem 1.7.

Let dimX=n\dim X=n and 0<σ<10<\sigma<1. Then

  1. (1)

    Tσ:Hσ(X)Hσ+1/2(X×+)T_{\sigma}:H^{\sigma}(X)\rightarrow H^{\sigma+1/2}(X\times\mathbb{R}_{+}) is isometric up to a constant.

  2. (2)

    TσT_{\sigma} extends to a bounded operator from Lp(X)L^{p}(X) to Lq(X×+)L^{q}(X\times\mathbb{R}_{+}), for 1<p<1<p<\infty and p<qn+1npp<q\leq\frac{n+1}{n}p; and from L1(X)L^{1}(X) to Lq(X)L^{q}(X), for 1<q<n+1n1<q<\frac{n+1}{n} (Figure 1, (b)).

Refer to caption
Figure 1. (a) Euclidean (b) Symmetric spaces
Remark 1.8.

In contrast with Theorem 1.6 on Euclidean space, the exponents p,qp,q in Theorem 1.7 on XX can vary over a much larger region (see in the figure 1 above). This striking phenomenon comes as a consequence of the Kunze-Stein phenomenon. The Kunze-Stein phenomenon, proved by Cowling [16] on connected semi-simple Lie groups GG with finite center, says that the convolution inequality

L2(G)Lp(G)L2(G),L^{2}(G)\ast L^{p}(G)\subset L^{2}(G),

holds for p[1,2)p\in[1,2). We note that above inequalities on Euclidean space are only valid for p=1p=1. We use the following generalize version [17, Theorem 2.2, (ii)]: let kLq(X)k\in L^{q}(X), for 1<q21<q\leq 2 and let 1p<q1\leq p<q. Then the map ffkf\mapsto f\ast k is bounded from Lp(X)L^{p}(X) to Lq(X)L^{q}(X).

An explicit expression of the heat kernel is known for certain symmetric spaces. Using this in section 5, we write the precise expression of the kernel PyσP_{y}^{\sigma} in the case of complex and rank one symmetric spaces.

The final topic we shall deal with here is analogues of the Poincaré-Sobolev inequalities for the fractional Laplace-Beltrami operator on XX. In [33], Mancini and Sandeep proved the following optimal Poincaré-Sobolev inequalities for the Laplace-Beltrami operator Δn\Delta_{\mathbb{H}^{n}} on the real hyperbolic space n\mathbb{H}^{n} of dimension n3n\geq 3.

Theorem 1.9.

((Mancini; Sandeep)) Let n3n\geq 3. Then for 2<p2nn22<p\leq\frac{2n}{n-2}, there exists S=Sn,p>0S=S_{n,p}>0 such that for all uCc(n)u\in C_{c}^{\infty}(\mathbb{H}^{n}),

(Δn(n1)2/4))1/2uL2(n)2SuLp(n)2.\|\left(-\Delta_{\mathbb{H}^{n}}-(n-1)^{2}/4)\right)^{1/2}u\|_{L^{2}(\mathbb{H}^{n})}^{2}\geq S\|u\|_{L^{p}(\mathbb{H}^{n})}^{2}.

In case of real hyperbolic space 3\mathbb{H}^{3} of dimension three, Benguria, Frank and Loss [8] proved that the best constant S3S_{3} in the theorem above is the same as the best sharp Sobolev constant for the first order Sobolev inequality on 3\mathbb{H}^{3}. Recently, using Green kernel estimates Li, Lu, Yang [31, Theorem 6.2] proved the following Poincaré-Sobolev inequalities for the fractional Laplace-Beltrami operator Δn\Delta_{\mathbb{H}^{n}} on n\mathbb{H}^{n}.

Theorem 1.10.

((Li; Lu; Yang)) Let n3n\geq 3 and 1σ<31\leq\sigma<3. Then there exists a constant C=Cn,σ,p>0C=C_{n,\sigma,p}>0 such that

(Δn(n1)2/4)σ4uL2(n)2CuL2nnσ(n)2, for uHσ2(n).\|\left(-\Delta_{\mathbb{H}^{n}}-(n-1)^{2}/4\right)^{\frac{\sigma}{4}}u\|_{L^{2}(\mathbb{H}^{n})}^{2}\geq C\|u\|^{2}_{L^{\frac{2n}{n-\sigma}}(\mathbb{H}^{n})},\>\>\textit{ for }u\in H^{\frac{\sigma}{2}}(\mathbb{H}^{n}).

For related results and their sharpness, we refer the reader to [32, 40]. Our aim in the final section is to prove an analogue of the Poincaré-Sobolev inequality for the fractional Laplace-Beltrami operator Δ\Delta on XX which generalizes the above mentioned theorems. The idea of the proof is to use the estimate of the Bassel-Green-Riesz kernel due to Anker-Ji [4]. Since we are working on general Riemannian symmetric spaces of noncompact type, it is difficult to get the explicit values of the constants involve and we do not make attempt to get the optimal constant. Here is our final result. We refer the reader to the next section for the unexplained notation used in the theorem below.

Theorem 1.11.

Let dimX=n3\dim X=n\geq 3 and 0<σ<min{l+2|Σ0+|,n}0<\sigma<\min\{l+2|\Sigma_{0}^{+}|,n\}. Then for 2<p2nnσ2<p\leq\frac{2n}{n-\sigma} there exists S=Sn,σ,p>0S=S_{n,\sigma,p}>0 such that for all uHσ2(X)u\in H^{\frac{\sigma}{2}}(X),

(Δ|ρ|2)σ/4uL2(X)2SuLp(X)2.\|(-\Delta-|\rho|^{2})^{\sigma/4}u\|_{L^{2}(X)}^{2}\geq S\|u\|_{L^{p}(X)}^{2}.

2. Preliminaries

In this section, we describe the necessary preliminaries regarding semisimple Lie groups and harmonic analysis on Riemannian symmetric spaces. These are standard and can be found, for example, in [20, 24, 25, 26]. To make the article self-contained, we shall gather only those results which will be used throughout this paper.

2.1. Notations

Let GG be a connected, noncompact, real semisimple Lie group with finite centre and 𝔤\mathfrak{g} its Lie algebra. We fix a Cartan involution θ\theta of 𝔤\mathfrak{g} and write 𝔤=𝔨𝔭\mathfrak{g}=\mathfrak{k}\oplus\mathfrak{p} where 𝔨\mathfrak{k} and 𝔭\mathfrak{p} are +1+1 and 1-1 eigenspaces of θ\theta respectively. Then 𝔨\mathfrak{k} is a maximal compact subalgebra of 𝔤\mathfrak{g} and 𝔭\mathfrak{p} is a linear subspace of 𝔤\mathfrak{g}. The Cartan involution θ\theta induces an automorphism Θ\Theta of the group GG and K={gGΘ(g)=g}K=\{g\in G\mid\Theta(g)=g\} is a maximal compact subgroup of GG. Let BB denote the Cartan Killing form of 𝔤\mathfrak{g}. It is known that B𝔭×𝔭B\mid_{\mathfrak{p}\times\mathfrak{p}} is positive definite and hence induces an inner product and a norm B\|\cdot\|_{B} on 𝔭\mathfrak{p}. The homogeneous space X=G/KX=G/K is a smooth manifold. The tangent space of XX at the point o=eKo=eK can be naturally identified to 𝔭\mathfrak{p} and the restriction of BB on 𝔭\mathfrak{p} then induces a GG-invariant Riemannian metric 𝐝{\bf d} on XX. For xXx\in X and r>0r>0, we denote 𝐁(x,r){\bf B}(x,r) to be the ball of radius rr centered at xx in this metric.

Let 𝔞\mathfrak{a} be a maximal subalgebra in 𝔭\mathfrak{p}; then 𝔞\mathfrak{a} is abelian. We assume that dim𝔞=l\dim\mathfrak{a}=l, called the real rank of GG. We can identify 𝔞\mathfrak{a} endowed with the inner product induced from 𝔭\mathfrak{p} with d\mathbb{R}^{d} and let 𝔞\mathfrak{a}^{*} be the real dual of 𝔞\mathfrak{a}. The set of restricted roots of the pair (𝔤,𝔞)(\mathfrak{g},\mathfrak{a}) is denoted by Σ\Sigma. It consists of all α𝔞\alpha\in\mathfrak{a}^{*} such that

𝔤α={X𝔤|[Y,X]=α(Y)X, for all Y𝔞}\mathfrak{g}_{\alpha}=\left\{X\in\mathfrak{g}~{}|~{}[Y,X]=\alpha(Y)X,\>\>\textmd{ for all }Y\in\mathfrak{a}\right\}

is nonzero with mα=dim(𝔤α)m_{\alpha}=\dim(\mathfrak{g}_{\alpha}). We choose a system of positive roots Σ+\Sigma^{+} and with respect to Σ+\Sigma^{+}, the positive Weyl chamber 𝔞+={X𝔞|α(X)>0, for all αΣ+}\mathfrak{a}_{+}=\left\{X\in\mathfrak{a}~{}|~{}\alpha(X)>0,\>\>\textmd{ for all }\alpha\in\Sigma^{+}\right\}. We also let Σ0+\Sigma_{0}^{+} be the set of positive indivisible roots. We denote by

𝔫=αΣ+𝔤α.\mathfrak{n}=\oplus_{\alpha\in\Sigma^{+}}~{}\mathfrak{g}_{\alpha}.

Then 𝔫\mathfrak{n} is a nilpotent subalgebra of 𝔤\mathfrak{g} and we obtain the Iwasawa decomposition 𝔤=𝔨𝔞𝔫\mathfrak{g}=\mathfrak{k}\oplus\mathfrak{a}\oplus\mathfrak{n}. If N=exp𝔫N=\exp\mathfrak{n} and A=exp𝔞A=\exp\mathfrak{a} then NN is a Nilpotent Lie group and AA normalizes NN. For the group GG, we now have the Iwasawa decomposition G=KANG=KAN, that is, every gGg\in G can be uniquely written as

g=κ(g)expH(g)η(g),κ(g)K,H(g)𝔞,η(g)N,g=\kappa(g)\exp H(g)\eta(g),\>\>\>\>\kappa(g)\in K,H(g)\in\mathfrak{a},\eta(g)\in N,

and the map

(k,a,n)kan(k,a,n)\mapsto kan

is a global diffeomorphism of K×A×NK\times A\times N onto GG. Let nn be the dimension of XX then

n=l+αΣ+mα.n=l+\sum_{\alpha\in\Sigma^{+}}m_{\alpha}.

We always assume that n2n\geq 2. Let ρ\rho denote the half sum of all positive roots counted with their multiplicities:

ρ=12αΣ+mαα.\rho=\frac{1}{2}\sum_{\alpha\in\Sigma^{+}}m_{\alpha}~{}\alpha.

It is known that the L2L^{2}-spectrum of the Laplace-Beltrami operator Δ\Delta on XX is the half-line (,|ρ|2](-\infty,-|\rho|^{2}]. Let MM^{\prime} and MM be the normalizer and centralizer of 𝔞\mathfrak{a} in KK respectively. Then MM is a normal subgroup of MM^{\prime} and normalizes NN. The quotient group W=M/MW=M^{\prime}/M is a finite group, called the Weyl group of the pair (𝔤,𝔨)(\mathfrak{g},\mathfrak{k}). The Weyl group WW acts on 𝔞\mathfrak{a} by the adjoint action. It is known that WW acts as a group of orthogonal transformations (preserving the Cartan-Killing form) on 𝔞\mathfrak{a}. Each wWw\in W permutes the Weyl chambers and the action of WW on the Weyl chambers is simply transitive. Let A+=exp𝔞+A_{+}=\exp{\mathfrak{a_{+}}}. Since exp:𝔞A\exp:\mathfrak{a}\to A is an isomorphism we can identify AA with d\mathbb{R}^{d}. If A+¯\overline{A_{+}} denotes the closure of A+A_{+} in GG, then one has the polar decomposition G=KAKG=KAK, that is, each gGg\in G can be written as

g=k1(expY)k2,k1,k2K,Y𝔞.g=k_{1}(\exp Y)k_{2},\>\>k_{1},k_{2}\in K,Y\in\mathfrak{a}.

In the above decomposition, the AA component of 𝔤\mathfrak{g} is uniquely determined modulo WW. In particular, it is well defined in A+¯\overline{A_{+}}. The map (k1,a,k2)k1ak2(k_{1},a,k_{2})\mapsto k_{1}ak_{2} of K×A×KK\times A\times K into GG induces a diffeomorphism of K/M×A+×KK/M\times A_{+}\times K onto an open dense subset of GG. We extend the inner product on 𝔞\mathfrak{a} induced by BB to 𝔞\mathfrak{a}^{*} by duality, that is, we set

λ,μ=B(Yλ,Yμ),λ,μ𝔞,Yλ,Yμ𝔞,\langle\lambda,\mu\rangle=B(Y_{\lambda},Y_{\mu}),\>\>\>\>\lambda,\mu\in\mathfrak{a}^{*},~{}Y_{\lambda},Y_{\mu}\in\mathfrak{a},

where YλY_{\lambda} is the unique element in 𝔞\mathfrak{a} such that

λ(Y)=B(Yλ,Y), for all Y𝔞.\lambda(Y)=B(Y_{\lambda},Y),\>\>\>\>\textmd{ for all }Y\in\mathfrak{a}.

This inner product induces a norm, again denoted by |||\cdot|, on 𝔞\mathfrak{a}^{*},

|λ|=λ,λ12,λ𝔞.|\lambda|=\langle\lambda,\lambda\rangle^{\frac{1}{2}},\>\>\>\>\lambda\in\mathfrak{a}^{*}.

The elements of the Weyl group WW acts on 𝔞\mathfrak{a}^{*} by the formula

sYλ=Ysλ,sW,λ𝔞.sY_{\lambda}=Y_{s\lambda},\>\>\>\>\>\>s\in W,\>\lambda\in\mathfrak{a}^{*}.

Let 𝔞\mathfrak{a}_{\mathbb{C}}^{*} denote the complexification of 𝔞\mathfrak{a}^{*}, that is, the set of all complex-valued real linear functionals on 𝔞\mathfrak{a}. The usual extension of BB to 𝔞\mathfrak{a}_{\mathbb{C}}^{*}, using conjugate linearity is also denoted by BB. Through the identification of AA with d\mathbb{R}^{d}, we use the Lebesgue measure on d\mathbb{R}^{d} as the Haar measure dada on AA. As usual on the compact group KK, we fix the normalized Haar measure dkdk and dndn denotes a Haar measure on NN. The following integral formulae describe the Haar measure of GG corresponding to the Iwasawa and polar decomposition respectively. For any fCc(G)f\in C_{c}(G),

Gf(g)𝑑g\displaystyle\int_{G}{f(g)dg} =\displaystyle= K𝔞Nf(kexpYn)e2ρ(Y)𝑑n𝑑Y𝑑k\displaystyle\int_{K}\int_{\mathfrak{a}}\int_{N}f(k\exp Yn)~{}e^{2\rho(Y)}~{}dn~{}dY~{}dk
=\displaystyle= KA+¯Kf(k1ak2)J(a)𝑑k1𝑑a𝑑k2,\displaystyle\int_{K}{\int_{\overline{A_{+}}}{\int_{K}{f(k_{1}ak_{2})~{}J(a)~{}dk_{1}~{}da~{}dk_{2}}}},

where dYdY is the Lebesgue measure on d\mathbb{R}^{d} and for H𝔞+¯H\in\overline{\mathfrak{a}_{+}}

(2.1) J(expH)=cαΣ+(sinhα(H))mα{αΣ+(α(H)1+α(H))mα}e2ρ(H),J(\exp H)=c\prod_{\alpha\in\Sigma^{+}}\left(\sinh\alpha(H)\right)^{m_{\alpha}}\asymp\left\{\prod_{\alpha\in\Sigma^{+}}\left(\frac{\alpha(H)}{1+\alpha(H)}\right)^{m_{\alpha}}\right\}e^{2\rho(H)},

where cc (in the equality above) is a normalizing constant. If ff is a function on X=G/KX=G/K then ff can be thought of as a function on GG which is right invariant under the action of KK. It follows that on XX we have a GG invariant measure dxdx such that

(2.2) Xf(x)𝑑x=K/M𝔞+f(kexpY)J(expY)𝑑Y𝑑kM,\int_{X}f(x)~{}dx=\int_{K/M}\int_{\mathfrak{a}_{+}}f(k\exp Y)~{}J(\exp Y)~{}dY~{}dk_{M},

where dkMdk_{M} is the KK-invariant measure on K/MK/M.

2.2. Fourier analysis on XX

For a sufficiently nice function ff on XX, its Fourier transform f~\widetilde{f} is a function defined on 𝔞×K\mathfrak{a}_{\mathbb{C}}^{*}\times K given by

f~(λ,k)=Gf(g)e(iλρ)H(g1k)𝑑g,λ𝔞,kK,\widetilde{f}(\lambda,k)=\int_{G}f(g)e^{(i\lambda-\rho)H(g^{-1}k)}dg,\>\>\>\>\>\>\lambda\in\mathfrak{a}_{\mathbb{C}}^{*},\>\>k\in K,

whenever the integral exists [25, P. 199]. As MM normalizes NN the function kf~(λ,k)k\mapsto\widetilde{f}(\lambda,k) is right MM-invariant. It is known that if fL1(X)f\in L^{1}(X) then f~(λ,k)\widetilde{f}(\lambda,k) is a continuous function of λ𝔞\lambda\in\mathfrak{a}^{*}, for almost every kKk\in K (in fact, holomorphic in λ\lambda on a domain containing 𝔞\mathfrak{a}^{\ast}). If in addition, f~L1(𝔞×K,|𝐜(λ)|2dλdk)\widetilde{f}\in L^{1}(\mathfrak{a}^{*}\times K,|{\bf c}(\lambda)|^{-2}~{}d\lambda~{}dk) then the following Fourier inversion holds,

f(gK)=|W|1𝔞×Kf~(λ,k)e(iλ+ρ)H(g1k)|𝐜(λ)|2𝑑λ𝑑k,f(gK)=|W|^{-1}\int_{\mathfrak{a}^{*}\times K}\widetilde{f}(\lambda,k)~{}e^{-(i\lambda+\rho)H(g^{-1}k)}~{}|{\bf c}(\lambda)|^{-2}d\lambda~{}dk,

for almost every gKXgK\in X [25, Chapter III, Theorem 1.8, Theorem 1.9]. Here 𝐜(λ){\bf c}(\lambda) denotes Harish Chandra’s 𝐜{\bf c}-function. Moreover, ff~f\mapsto\widetilde{f} extends to an isometry of L2(X)L^{2}(X) onto L2(𝔞+×K,|𝐜(λ)|2dλdk)L^{2}(\mathfrak{a}^{*}_{+}\times K,|{\bf c}(\lambda)|^{-2}~{}d\lambda~{}dk) [25, Chapter III, Theorem 1.5]:

X|f(x)|2𝑑x=|W|1𝔞×K|f~(λ,k)|2|𝐜(λ)|2𝑑λ𝑑k.\int_{X}|f(x)|^{2}~{}dx=|W|^{-1}\int_{\mathfrak{a}^{\ast}\times K}|\widetilde{f}(\lambda,k)|^{2}~{}|{\bf c}(\lambda)|^{-2}~{}d\lambda~{}dk.

It is known [26, Ch. IV, prop 7.2] that there exists a positive number CC and dd\in\mathbb{N} such that for all λ𝔞+\lambda\in\mathfrak{a}_{+}^{*}

|𝐜(λ)|2\displaystyle|{\bf c}(\lambda)|^{-2} \displaystyle\leq C(1+|λ|)nl, for |λ|1;\displaystyle C(1+|\lambda|)^{n-l},\>\>\textit{ for }|\lambda|\geq 1;
\displaystyle\leq C(1+|λ|)d, for |λ|<1.\displaystyle C(1+|\lambda|)^{d},\>\>\textit{ for }|\lambda|<1.

We now specialize in the case of KK-biinvariant function ff on GG. Using the polar decomposition of GG we may view a KK-biinvariant integrable function ff on GG as a function on A+A_{+}, or by using the inverse exponential map we may also view ff as a function on 𝔞\mathfrak{a} solely determined by its values on 𝔞+\mathfrak{a}_{+}. Henceforth, we shall denote the set of KK-biinvariant functions in L1(G)L^{1}(G) by L1(K\G/K)L^{1}(K\backslash G/K). If fL1(K\G/K)f\in L^{1}(K\backslash G/K) then the Fourier transform f~\widetilde{f} reduces to the spherical Fourier transform f^(λ)\widehat{f}(\lambda) which is given by the integral

(2.4) f~(λ,k)=f^(λ):=Gf(g)ϕλ(g)𝑑g,\widetilde{f}(\lambda,k)=\widehat{f}(\lambda):=\int_{G}f(g)\phi_{-\lambda}(g)~{}dg,

for all kKk\in K where

(2.5) ϕλ(g)=Ke(iλ+ρ)(H(g1k))𝑑k,λ𝔞,\phi_{\lambda}(g)=\int_{K}e^{-(i\lambda+\rho)\big{(}H(g^{-1}k)\big{)}}~{}dk,\>\>\>\>\>\>\lambda\in\mathfrak{a}_{\mathbb{C}}^{*},

is Harish Chandra’s elementary spherical function. We now list down some well-known properties of the elementary spherical functions which are important for us ([4, Prop. 2.2.12], [20, Prop. 3.1.4]; [25, Lemma 1.18, P. 221]).

Theorem 2.1.
  1. (1)

    ϕλ(g)\phi_{\lambda}(g) is KK-biinvariant in gGg\in G and WW-invariant in λ𝔞\lambda\in\mathfrak{a}_{\mathbb{C}}^{*}.

  2. (2)

    ϕλ(g)\phi_{\lambda}(g) is CC^{\infty} in gGg\in G and holomorphic in λ𝔞\lambda\in\mathfrak{a}_{\mathbb{C}}^{*}.

  3. (3)

    The elementary spherical function ϕ0\phi_{0} satisfies the following global estimate:

    (2.6) ϕ0(expH){αΣ0+(1+α(H))}eρ(H), for all H𝔞+¯.\phi_{0}(\exp H)\asymp\left\{\prod_{\alpha\in\Sigma_{0}^{+}}\left(1+\alpha(H)\right)\right\}e^{-\rho(H)},\>\>\text{ for all }H\in\overline{\mathfrak{a}^{+}}.
  4. (4)

    For all λ𝔞+¯\lambda\in\overline{\mathfrak{a}_{+}^{*}} we have

    (2.7) |ϕλ(g)|ϕ0(g)1.|\phi_{\lambda}(g)|\leq\phi_{0}(g)\leq 1.

2.3. Function spaces on XX

For 1p<1\leq p<\infty we define

Lp(X×)={u|uLp(X×)p:=X×|u(x,y)|p𝑑x𝑑y<},L^{p}(X\times\mathbb{R})=\left\{u~{}|~{}\|u\|_{L^{p}(X\times\mathbb{R})}^{p}:=\int_{X\times\mathbb{R}}|u(x,y)|^{p}~{}dx~{}dy<\infty\right\},

and Lp(X×+)L^{p}(X\times\mathbb{R}_{+}) to be the subspace of Lp(X×)L^{p}(X\times\mathbb{R}) consisting of all functions u(x,y)u(x,y) which are even in the yy-variable. We also define L(X×+)L^{\infty}(X\times\mathbb{R}^{+}) analogously. For σ>0\sigma>0, the Sobolev space of order σ\sigma on XX is defined by

Hσ(X)={fL2(X)|fHσ(X)2:=𝔞×K|f~(λ,k)|2(|λ|2+|ρ|2)σ|𝐜(λ)|2𝑑λ𝑑k<}.H^{\sigma}(X)=\big{\{}f\in L^{2}(X)~{}|~{}\|f\|_{H^{\sigma}(X)}^{2}:=\int_{\mathfrak{a}^{\ast}\times K}|\tilde{f}(\lambda,k)|^{2}~{}(|\lambda|^{2}+|\rho|^{2})^{\sigma}~{}|{\bf c}(\lambda)|^{-2}~{}d\lambda~{}dk<\infty\big{\}}.

Similarly, for σ>0\sigma>0 we define Hσ(X×)H^{\sigma}(X\times\mathbb{R}) as the space of all functions uL2(X×)u\in L^{2}(X\times\mathbb{R}) such that

uHσ(X×)2:=𝔞×K|(u~(λ,k,)(ξ))|2(|λ|2+|ρ|2+ξ2)σ|𝐜(λ)|2𝑑λ𝑑k𝑑ξ<,\|u\|_{H^{\sigma}(X\times\mathbb{R})}^{2}:=\int_{\mathbb{R}}\int_{\mathfrak{a}^{\ast}\times K}|\mathcal{F}\left(\tilde{u}(\lambda,k,\cdot)(\xi)\right)|^{2}~{}\left(|\lambda|^{2}+|\rho|^{2}+\xi^{2}\right)^{\sigma}~{}|{\bf c}(\lambda)|^{-2}~{}d\lambda~{}dk~{}d\xi<\infty,

where u~(λ,k,)(ξ)\mathcal{F}\tilde{u}(\lambda,k,\cdot)(\xi) denotes the Euclidean Fourier transform of the function yu~(λ,k,y)y\mapsto\tilde{u}(\lambda,k,y) at the point ξ\xi\in\mathbb{R}, for almost every (λ,k)𝔞×K(\lambda,k)\in\mathfrak{a}^{\ast}\times K. Let Hσ(X×+)H^{\sigma}(X\times\mathbb{R}_{+}) be the subspace of Hσ(X×)H^{\sigma}(X\times\mathbb{R}) consisting of all elements u(x,y)u(x,y) which are even in the yy-variable.

2.4. Heat kernel on XX

For the details of the heat kernel hth_{t} on X=G/KX=G/K we refer [3, 4]. It is a family {ht:t>0}\{h_{t}:~{}t>0\} of smooth functions with the following properties:

  1. (a)

    htLp(K\G/K),p[1,]h_{t}\in L^{p}(K\backslash G/K),~{}~{}p\in[1,\infty],   for each t>0t>0.

  2. (b)

    For each t>0t>0, hth_{t} is positive with

    (2.8) Ght(g)𝑑g=1.\int_{G}h_{t}(g)~{}dg=1.
  3. (c)

    ht+s=hths,t,s>0.h_{t+s}=h_{t}*h_{s},~{}t,s>0.

  4. (d)

    For each fLp(G/K),p[1,)f\in L^{p}(G/K),~{}p\in[1,\infty) the function u(x,t)=fht(x)u(x,t)=f*h_{t}(x), for xXx\in X solves the heat equation

    Δxu(x,t)\displaystyle\Delta_{x}u(x,t) =\displaystyle= tu(x,t)\displaystyle\frac{\partial}{\partial t}u(x,t)
    u(,t)\displaystyle u(\cdot,t) \displaystyle\rightarrow f in Lp(X), as t0.\displaystyle f\textmd{ in }L^{p}(X),\textmd{ as }t\rightarrow 0.
  5. (e)

    The spherical Fourier transform of hth_{t} is given by

    (2.9) ht^(λ)=et(|λ|2+|ρ|2),λ𝔞.\widehat{h_{t}}(\lambda)=e^{-t(|\lambda|^{2}+|\rho|^{2})},\>\>\>\>\lambda\in\mathfrak{a}^{*}.

We need the following both side estimates of the heat kernel [4, Theorem 3.7].

Theorem 2.2.

Let κ\kappa be an arbitrary positive number. Then there exists positive constants C1,C2C_{1},C_{2} (depending on κ\kappa) such that

C1ht(expH)tn2(1+t)nl2|Σ0+|{αΣ0+(1+α(H)}e|ρ|2tρ(H)|H|24tC2,C_{1}\leq\frac{h_{t}(\exp H)}{t^{-\frac{n}{2}}(1+t)^{\frac{n-l}{2}-|\Sigma_{0}^{+}|}\left\{\prod_{\alpha\in\Sigma_{0}^{+}}(1+\alpha(H)\right\}e^{-|\rho|^{2}t-\rho(H)-\frac{|H|^{2}}{4t}}}\leq C_{2},

for all t>0t>0, and H𝔞+¯H\in\overline{\mathfrak{a}^{+}}, with |H|κ(1+t)|H|\leq\kappa(1+t).

For H𝔞+¯H\in\overline{\mathfrak{a}^{+}} with tHt\ll H, we will use the following global upper bound [3, Theorem 3.1]

(2.10) |ht(expH)|td1(1+|H|)d2e|ρ|2tρ(H)|H|2/(4t),|h_{t}(\exp H)|\leq t^{-d_{1}}(1+|H|)^{d_{2}}e^{-|\rho|^{2}t-\rho(H)-|H|^{2}/(4t)},

where d1d_{1} and d2d_{2} are positive constants depending on the position of H𝔞+¯H\in\overline{\mathfrak{a}^{+}} with respect to the walls and on the relative size of t>0t>0 and 1+|H|1+|H|.

3. Extension problem and kernel estimates

Since we are dealing with fractional operators, it is natural to relate the fractional Laplace-Beltrami operator acting on ff to the solution uu in (1.5). We proceed by proving Theorem 1.5 which will provide us an expression for the Poisson kernel of the extension operator. This will crucially be used throughout this paper.

Proof of Theorem 1.5.

Using the heat-diffusion semigroup generated by the Laplace-Beltrami operator, the first part of the theorem follows exactly as in [39, Theorem 1.1]. We will prove the second part. Let fDom(Δ)σf\in Dom(-\Delta)^{\sigma}, and uu be the solution of the extension problem (1.5) given by equation (1.7). It now follows that

u(,y),g=1Γ(σ)|ρ|20etλλσey2/4tdtt1σ𝑑Ef,g(λ),\left\langle u(\cdot,y),g\right\rangle=\frac{1}{\Gamma(\sigma)}\int_{|\rho|^{2}}^{\infty}\int_{0}^{\infty}e^{-t\lambda}\lambda^{\sigma}e^{-y^{2}/4t}\frac{dt}{t^{1-\sigma}}\,dE_{f,g}(\lambda),

for all gL2(X)g\in L^{2}(X), where dEf,g(λ)dE_{f,g}(\lambda) is the regular Borel complex measure of bounded variation concentrated on the spectrum [|ρ|2,)[|\rho|^{2},\infty) of Δ-\Delta with d|Ef,g|(|ρ|2,)fL2(X)gL2(X)d|E_{f,g}|(|\rho|^{2},\infty)\leq\|f\|_{L^{2}(X)}\|g\|_{L^{2}(X)}. By the Fubini’s theorem, putting rλ=y2/4tr\lambda=y^{2}/4t the above equation yields

u(,y),g\displaystyle\left\langle u(\cdot,y),g\right\rangle =\displaystyle= y2σ4σΓ(σ)0|ρ|2ey2/4rerλ𝑑Ef,g(λ)drr1+σ\displaystyle\frac{y^{2\sigma}}{4^{\sigma}\Gamma(\sigma)}\int_{0}^{\infty}\int_{|\rho|^{2}}^{\infty}e^{-y^{2}/4r}~{}e^{-r\lambda}\,dE_{f,g}(\lambda)\frac{dr}{r^{1+\sigma}}
=\displaystyle= y2σ4σΓ(σ)0erΔf,gL2(X)ey2/4rdrr1+σ\displaystyle\frac{y^{2\sigma}}{4^{\sigma}\Gamma(\sigma)}\int_{0}^{\infty}\left\langle e^{r\Delta}f,g\right\rangle_{L^{2}(X)}e^{-y^{2}/4r}\frac{dr}{r^{1+\sigma}}
=\displaystyle= y2σ4σΓ(σ)0erΔfey2/4rdrr1+σ,g.\displaystyle\left\langle\frac{y^{2\sigma}}{4^{\sigma}\Gamma(\sigma)}\int_{0}^{\infty}e^{r\Delta}f~{}e^{-y^{2}/4r}\frac{dr}{r^{1+\sigma}},g\right\rangle.

This proves that

u(x,y)=y2σ4σΓ(σ)0erΔf(x)ey2/4rdrr1+σ.u(x,y)=\frac{y^{2\sigma}}{4^{\sigma}\Gamma(\sigma)}\int_{0}^{\infty}e^{r\Delta}f(x)e^{-y^{2}/4r}\frac{dr}{r^{1+\sigma}}.

Now, using etΔf=fhte^{t\Delta}f=f\ast h_{t} and Fubini’s theorem, we get from the above equation that

u(x,y)=y2σ4σΓ(σ)X0f(xz1)ht(z)ey2/4tdtt1+σ𝑑z=fPyσ(x),u(x,y)=\frac{y^{2\sigma}}{4^{\sigma}\Gamma(\sigma)}\int_{X}\int_{0}^{\infty}f(xz^{-1})~{}h_{t}(z)e^{-y^{2}/4t}\,\frac{dt}{t^{1+\sigma}}~{}dz=f\ast P_{y}^{\sigma}(x),

where the kernel PyσP_{y}^{\sigma} is given by the equation (1.9). ∎

As in [39, Theorem 2.1], we have the following consequences of the theorem above.

Corollary 3.1.

Let u(x,y)=(fPyσ)(x)u(x,y)=(f\ast P_{y}^{\sigma})(x), for xX,y>0x\in X,y>0 be the solution of the extension problem (1.5) given in Theorem 1.5. Then

  1. (a)

    supy0|u(x,y)|supt0|fht|\sup_{y\geq 0}|u(x,y)|\leq\sup_{t\geq 0}|f\ast h_{t}| in XX.

  2. (b)

    u(,y)Lp(X)fLp(X)\|u(\cdot,y)\|_{L^{p}(X)}\leq\|f\|_{L^{p}(X)},   for all y0y\geq 0 and p[1,]p\in[1,\infty].

  3. (c)

    limy0+u(,y)=f\lim_{y\rightarrow 0^{+}}u(\cdot,y)=f in Lp(X)L^{p}(X), for p[1,)p\in[1,\infty).

Proof.

Part (a) follows from the expression of the Poisson kernel given in equation (1.9). For (b), we observe that etΔe^{t\Delta} has the contraction property in Lp(X)L^{p}(X). Hence, by equation (1.9) and Minkowski’s integral inequality it follows that

u(,y)Lp(X)y2σ4σΓ(σ)0fhtLp(X)ey2/4tdtt1+σfLp(X).\displaystyle\|u(\cdot,y)\|_{L^{p}(X)}\leq\frac{y^{2\sigma}}{4^{\sigma}\Gamma(\sigma)}\int_{0}^{\infty}\|f\ast h_{t}\|_{L^{p}(X)}~{}e^{-y^{2}/4t}~{}\frac{dt}{t^{1+\sigma}}\leq\|f\|_{L^{p}(X)}.

Similarly, for part (c) we observe that

u(,y)fLp(X)y2σ4σΓ(σ)0fhtfLp(X)ey2/4tdtt1+σ.\|u(\cdot,y)-f\|_{L^{p}(X)}\leq\frac{y^{2\sigma}}{4^{\sigma}\Gamma(\sigma)}\int_{0}^{\infty}\|f\ast h_{t}-f\|_{L^{p}(X)}~{}e^{-y^{2}/4t}~{}\frac{dt}{t^{1+\sigma}}.

Since fhtfLp(X)2fLp(X)\|f\ast h_{t}-f\|_{L^{p}(X)}\leq 2\|f\|_{L^{p}(X)}, using dominated convergence theorem the result follows from the fact that limt0+fht=f\lim_{t\rightarrow 0^{+}}f\ast h_{t}=f in Lp(X)L^{p}(X), for p[1,)p\in[1,\infty). ∎

For 0<σ<10<\sigma<1 and y>0y>0, let us define the function PyσP_{y}^{-\sigma} given by the equation (1.9), that is

Pyσ(x)=y2σ4σΓ(σ)0ht(x)ey2/4tdtt1σ, for xX.P_{y}^{-\sigma}(x)=\frac{y^{-2\sigma}}{4^{-\sigma}\Gamma(-\sigma)}\int_{0}^{\infty}h_{t}(x)~{}e^{-y^{2}/4t}~{}\frac{dt}{t^{1-\sigma}},\>\>\textit{ for }x\in X.

By the estimate of the heat kernel (Theorem 2.2), it follows that PyσP_{y}^{-\sigma} is well defined. For 0<σ<10<\sigma<1, we observe that Γ(σ):=Γ(1σ)σ<0\Gamma(-\sigma):=\frac{\Gamma(1-\sigma)}{-\sigma}<0 and hence Pyσ0P_{y}^{-\sigma}\leq 0. Since the heat kernel hth_{t} is KK-biinvariant so is the function PyσP_{y}^{-\sigma}. By (2.4) the spherical Fourier transform is given by

(3.1) Pyσ^(λ)=XPyσ(x)ϕλ(x)𝑑x=y2σ4σΓ(σ)0ht^(λ)ey2/4tdtt1σ, for λ𝔞.\widehat{P_{y}^{-\sigma}}(\lambda)=\int_{X}P_{y}^{-\sigma}(x)~{}\phi_{-\lambda}(x)~{}dx=\frac{y^{-2\sigma}}{4^{-\sigma}\Gamma(-\sigma)}\int_{0}^{\infty}\widehat{h_{t}}(\lambda)~{}e^{-y^{2}/4t}~{}\frac{dt}{t^{1-\sigma}},\>\>\textit{ for }\lambda\in\mathfrak{a}^{\ast}.

Interchange of the integration is possible by the Fubini’s theorem. Indeed, by (2.7) and (2.9)

0Xht(x)|ϕλ(x)|𝑑xey2/4tdtt1σ\displaystyle\int_{0}^{\infty}\int_{X}h_{t}(x)~{}|\phi_{-\lambda}(x)|~{}dx~{}e^{-y^{2}/4t}~{}\frac{dt}{t^{1-\sigma}} \displaystyle\leq 0Xht(x)ϕ0(x)𝑑xey2/4tdtt1σ\displaystyle\int_{0}^{\infty}\int_{X}h_{t}(x)~{}\phi_{0}(x)~{}dx~{}e^{-y^{2}/4t}~{}\frac{dt}{t^{1-\sigma}}
=\displaystyle= 0et|ρ|2ey2/4tdtt1σ<.\displaystyle\int_{0}^{\infty}e^{-t|\rho|^{2}}~{}e^{-y^{2}/4t}~{}\frac{dt}{t^{1-\sigma}}<\infty.

Moreover, PyσP_{y}^{-\sigma} is contained in the Sobolev space Hσ(X)H^{\sigma}(X). Indeed, by using (3.1), (2.9) and Minkowski’s integral inequality we get that

PyσHσ(X)=(𝔞|Pyσ^(λ)|2(|λ|2+|ρ|2)σ|𝐜(λ)|2𝑑λ)12\displaystyle\|P_{y}^{-\sigma}\|_{H^{\sigma}(X)}=\left(\int_{\mathfrak{a}^{\ast}}|\widehat{P_{y}^{-\sigma}}(\lambda)|^{2}~{}\left(|\lambda|^{2}+|\rho|^{2}\right)^{\sigma}~{}|{\bf c}(\lambda)|^{-2}~{}d\lambda\right)^{\frac{1}{2}}
\displaystyle\leq y2σ4σ|Γ(σ)|0(𝔞|h^t(λ)|2(|λ|2+|ρ|2)σ|𝐜(λ)|2𝑑λ)12ey2/4tdtt1σ\displaystyle\frac{y^{-2\sigma}}{4^{-\sigma}|\Gamma(-\sigma)|}\int_{0}^{\infty}\left(\int_{\mathfrak{a}^{\ast}}|\widehat{h}_{t}(\lambda)|^{2}~{}(|\lambda|^{2}+|\rho|^{2})^{\sigma}~{}|{\bf c}(\lambda)|^{-2}~{}d\lambda\right)^{\frac{1}{2}}~{}e^{-y^{2}/4t}~{}\frac{dt}{t^{1-\sigma}}
\displaystyle\leq y2σ4σ|Γ(σ)|0(𝔞et(|λ|2+|ρ|2)(|λ|2+|ρ|2)σ|𝐜(λ)|2𝑑λ)12e|ρ|22tey2/4tdtt1σ\displaystyle\frac{y^{-2\sigma}}{4^{-\sigma}|\Gamma(-\sigma)|}\int_{0}^{\infty}\left(\int_{\mathfrak{a}^{\ast}}~{}e^{-t(|\lambda|^{2}+|\rho|^{2})}~{}(|\lambda|^{2}+|\rho|^{2})^{\sigma}~{}|{\bf c}(\lambda)|^{-2}~{}d\lambda\right)^{\frac{1}{2}}~{}e^{-\frac{|\rho|^{2}}{2}t}~{}e^{-y^{2}/4t}~{}\frac{dt}{t^{1-\sigma}}
=\displaystyle= I1+I2,\displaystyle I_{1}+I_{2},

where

(3.2) I1=y2σ4σ|Γ(σ)|01(𝔞et(|λ|2+|ρ|2)(|λ|2+|ρ|2)σ|𝐜(λ)|2𝑑λ)12e|ρ|22tey2/4tdtt1σ,I_{1}=\frac{y^{-2\sigma}}{4^{-\sigma}|\Gamma(-\sigma)|}\int_{0}^{1}\left(\int_{\mathfrak{a}^{\ast}}~{}e^{-t(|\lambda|^{2}+|\rho|^{2})}~{}(|\lambda|^{2}+|\rho|^{2})^{\sigma}~{}|{\bf c}(\lambda)|^{-2}~{}d\lambda\right)^{\frac{1}{2}}~{}e^{-\frac{|\rho|^{2}}{2}t}~{}e^{-y^{2}/4t}~{}\frac{dt}{t^{1-\sigma}},

and I2I_{2} is defiend as above with the integration in the tt-variable over the interval [1,)[1,\infty). It is enough to show that both I1I_{1} and I2I_{2} are finite. We consider I1I_{1} first. Using the property (2.2) of |𝐜(λ)|2|{\bf c}(\lambda)|^{-2}, we estimate the inner integral in the equation above as follows

{λ𝔞:|λ|<1}et|λ|2(|λ|2+|ρ|2)σ+d𝑑λ+{λ𝔞:|λ|1}et|λ|2(|λ|2+|ρ|2)σ+nl𝑑λ\displaystyle\int_{\{\lambda\in\mathfrak{a}^{\ast}:|\lambda|<1\}}e^{-t|\lambda|^{2}}~{}(|\lambda|^{2}+|\rho|^{2})^{\sigma+d}~{}d\lambda+\int_{\{\lambda\in\mathfrak{a}^{\ast}:|\lambda|\geq 1\}}e^{-t|\lambda|^{2}}~{}(|\lambda|^{2}+|\rho|^{2})^{\sigma+n-l}~{}d\lambda
\displaystyle\leq C1+C21etr2r2(σ+nl)rl1𝑑r\displaystyle C_{1}+C_{2}\int_{1}^{\infty}e^{-tr^{2}}~{}r^{2(\sigma+n-l)}~{}r^{l-1}~{}dr
\displaystyle\leq C1+C2t(σ+nl/2).\displaystyle C_{1}+C_{2}~{}t^{-(\sigma+n-l/2)}.

It now follows from (3.2) that

I1C01t12(σ+nl/2)e|ρ|22tey2/4tdtt1σ<.\displaystyle I_{1}\leq C\int_{0}^{1}t^{-\frac{1}{2}(\sigma+n-l/2)}~{}e^{-\frac{|\rho|^{2}}{2}t}e^{-y^{2}/4t}~{}\frac{dt}{t^{1-\sigma}}<\infty.

On the other hand

I2\displaystyle I_{2} \displaystyle\leq C1(𝔞e1(|λ|2+|ρ|2)(|λ|2+|ρ|2)σ|𝐜(λ)|2𝑑λ)12e|ρ|22tey2/4tdtt1σ\displaystyle C\int_{1}^{\infty}\left(\int_{\mathfrak{a}^{\ast}}~{}e^{-1(|\lambda|^{2}+|\rho|^{2})}~{}(|\lambda|^{2}+|\rho|^{2})^{\sigma}~{}|{\bf c}(\lambda)|^{-2}~{}d\lambda\right)^{\frac{1}{2}}~{}e^{-\frac{|\rho|^{2}}{2}t}~{}e^{-y^{2}/4t}~{}\frac{dt}{t^{1-\sigma}}
\displaystyle\leq Ch1/2Hσ(X).\displaystyle C\|h_{1/2}\|_{H^{\sigma}(X)}.

This completes the proof that PyσHσ(X)P_{y}^{-\sigma}\in H^{\sigma}(X).

The proofs of Hardy’s inequalities is crucially depend on the following lemma.

Lemma 3.2.

For 0<σ<10<\sigma<1 and y>0y>0 we have, (Δ)σPyσ(x)=4σΓ(σ)y2σΓ(σ)Pyσ(x)(-\Delta)^{\sigma}P_{y}^{-\sigma}(x)=\frac{4^{\sigma}\Gamma(\sigma)}{y^{2\sigma}\Gamma(-\sigma)}~{}P_{y}^{\sigma}(x).

Proof.

Let fHσ(X)f\in H^{\sigma}(X) and u(x,y)=fPyσ(x)u(x,y)=f\ast P_{y}^{\sigma}(x) be the solution of the extension problem (1.5). For any gL2(X)g\in L^{2}(X) we have by equation (1.7) that

u(,y),g\displaystyle\left\langle u(\cdot,y),g\right\rangle =\displaystyle= 1Γ(σ)0etΔ(Δ)σf,gey2/4tdtt1σ\displaystyle\frac{1}{\Gamma(\sigma)}\int_{0}^{\infty}\left\langle e^{t\Delta}~{}(-\Delta)^{\sigma}f,g\right\rangle~{}e^{-y^{2}/4t}~{}\frac{dt}{t^{1-\sigma}}
=\displaystyle= 1Γ(σ)0|ρ|2etλλσ𝑑Ef,g(λ)ey2/4tdtt1σ\displaystyle\frac{1}{\Gamma(\sigma)}\int_{0}^{\infty}\int_{|\rho|^{2}}^{\infty}e^{-t\lambda}~{}\lambda^{\sigma}dE_{f,g}(\lambda)~{}e^{-y^{2}/4t}~{}\frac{dt}{t^{1-\sigma}}
=\displaystyle= 1Γ(σ)|ρ|2(λσ0etλtσ1ey2/4t𝑑t)𝑑Ef,g(λ).\displaystyle\frac{1}{\Gamma(\sigma)}\int_{|\rho|^{2}}^{\infty}\left(\lambda^{\sigma}\int_{0}^{\infty}e^{-t\lambda}~{}t^{\sigma-1}~{}e^{-y^{2}/4t}~{}dt\right)~{}dE_{f,g}(\lambda).

By using change of variable ty2/(4λr)t\rightarrow y^{2}/(4\lambda r) we get the following formula ([10], p. 2582, equation(2.5))

λσ0etλtσ1ey2/4t𝑑t=y2σ4σ0etλtσ1ey2/4t𝑑t.\lambda^{\sigma}\int_{0}^{\infty}e^{-t\lambda}~{}t^{\sigma-1}~{}e^{-y^{2}/4t}~{}dt=\frac{y^{2\sigma}}{4^{\sigma}}\int_{0}^{\infty}e^{-t\lambda}~{}t^{-\sigma-1}~{}e^{-y^{2}/4t}~{}dt.

Using this in the equation above it follows that

(3.4) u(,y),g\displaystyle\left\langle u(\cdot,y),g\right\rangle =\displaystyle= y2σ4σΓ(σ)0tσ1ey2/4t(|ρ|2etλ𝑑Ef,g(λ))𝑑t\displaystyle\frac{y^{2\sigma}}{4^{\sigma}\Gamma(\sigma)}\int_{0}^{\infty}t^{-\sigma-1}~{}e^{-y^{2}/4t}~{}\left(\int_{|\rho|^{2}}^{\infty}e^{-t\lambda}~{}dE_{f,g}(\lambda)\right)~{}dt
=\displaystyle= y2σ4σΓ(σ)0etΔf,gtσ1ey2/4t𝑑t.\displaystyle\frac{y^{2\sigma}}{4^{\sigma}\Gamma(\sigma)}\int_{0}^{\infty}\left\langle e^{t\Delta}f,g\right\rangle t^{-\sigma-1}~{}e^{-y^{2}/4t}~{}dt.

Therefore, from equations (3) and (3.4) we have

0etΔ(Δ)σf(x)ey2/4ttσ1𝑑t=y2σ4σ0etΔf(x)ey2/4ttσ1𝑑t.\int_{0}^{\infty}e^{t\Delta}~{}(-\Delta)^{\sigma}f(x)~{}e^{-y^{2}/4t}~{}t^{\sigma-1}dt=\frac{y^{2\sigma}}{4^{\sigma}}\int_{0}^{\infty}e^{t\Delta}f(x)~{}e^{-y^{2}/4t}~{}t^{-\sigma-1}~{}dt.

If we take the function ff to be the heat kernel ht1,t1>0h_{t_{1}},t_{1}>0, then the equation above reduces to

0(Δ)σ(ht+t1)(x)ey2/4ttσ1𝑑t=y2σ4σ0ht+t1(x)ey2/4ttσ1𝑑t.\int_{0}^{\infty}(-\Delta)^{\sigma}(h_{t+t_{1}})(x)~{}e^{-y^{2}/4t}~{}t^{\sigma-1}dt=\frac{y^{2\sigma}}{4^{\sigma}}\int_{0}^{\infty}h_{t+t_{1}}(x)~{}e^{-y^{2}/4t}~{}t^{-\sigma-1}~{}dt.

Taking t10t_{1}\rightarrow 0, we get from dominated convergent theorem that

(3.5) 0(Δ)σht(x)ey2/4ttσ1𝑑t=y2σ4σ0ht(x)ey2/4ttσ1𝑑t.\int_{0}^{\infty}(-\Delta)^{\sigma}h_{t}(x)~{}e^{-y^{2}/4t}~{}t^{\sigma-1}dt=\frac{y^{2\sigma}}{4^{\sigma}}\int_{0}^{\infty}h_{t}(x)~{}e^{-y^{2}/4t}~{}t^{-\sigma-1}~{}dt.

Using (3.5) and (1.9) we get

(Δ)σPyσ(x)\displaystyle(-\Delta)^{\sigma}P_{y}^{-\sigma}(x) =\displaystyle= y2σ4σΓ(σ)0(Δ)σht(x)ey2/4ttσ1𝑑t\displaystyle\frac{y^{-2\sigma}}{4^{-\sigma}\Gamma(-\sigma)}\int_{0}^{\infty}(-\Delta)^{\sigma}h_{t}(x)~{}e^{-y^{2}/4t}~{}t^{\sigma-1}~{}dt
=\displaystyle= 1Γ(σ)0ht(x)ey2/4ttσ1𝑑t\displaystyle\frac{1}{\Gamma(-\sigma)}\int_{0}^{\infty}h_{t}(x)~{}e^{-y^{2}/4t}~{}t^{-\sigma-1}~{}dt
=\displaystyle= 4σΓ(σ)y2σΓ(σ)Pyσ(x).\displaystyle\frac{4^{\sigma}\Gamma(\sigma)}{y^{2\sigma}\Gamma(-\sigma)}~{}P_{y}^{\sigma}(x).

This completes the proof. ∎

We will now compute the asymptotic behaviour of the Poisson kernel PyσP_{y}^{\sigma} for arbitrary rank Riemannian symmetric spaces of noncompact type. We use this estimate crucially for the remaining part of this article.

Theorem 3.3.

For 1<σ<1,σ0-1<\sigma<1,\sigma\not=0 and y>0y>0 we have

Γ(σ)Pyσ(x)\displaystyle\Gamma(\sigma)\,P_{y}^{\sigma}(x) \displaystyle\asymp y2σ4σ|x|2+y2l/21/2σ|Σ0+|ϕ0(x)e|ρ||x|2+y2, for |x|2+y21,\displaystyle\frac{y^{2\sigma}}{4^{\sigma}}\sqrt{|x|^{2}+y^{2}}^{-l/2-1/2-\sigma-|\Sigma_{0}^{+}|}\phi_{0}(x)e^{-|\rho|\sqrt{|x|^{2}+y^{2}}},\textmd{ for }|x|^{2}+y^{2}\geq 1,
\displaystyle\asymp y2σ(|x|2+y2)n/2σ, for |x|2+y2<1.\displaystyle y^{2\sigma}~{}\left(|x|^{2}+y^{2}\right)^{-n/2-\sigma},\textmd{ for }~{}|x|^{2}+y^{2}<1.
Proof.

We first assume that |x|2+y2<1|x|^{2}+y^{2}<1. In this case, we will use the following local expansion of the heat kernel ht(x)h_{t}(x)

(3.6) ht(x)=e|x|2/4ttn/2v0(x)+ec|x|2/t𝒪(tn/2+1),h_{t}(x)=e^{-|x|^{2}/4t}t^{-n/2}v_{0}(x)+e^{-c|x|^{2}/t}\mathcal{O}\left(t^{-n/2+1}\right),

where v0(x)=(4π)n/2+𝒪(|x|2)v_{0}(x)=(4\pi)^{-n/2}+\mathcal{O}(|x|^{2}) and c<1/4c<1/4 ([3, (3.9), p. 278]). Using this we have

Γ(σ)Pyσ(x)\displaystyle\Gamma(\sigma)P_{y}^{\sigma}(x) =\displaystyle= y2σ4σ01(e|x|2/4ttn/2v0(x)+ec|x|2/t𝒪(tn/2+1))ey2/4tdtt1+σ\displaystyle\frac{y^{2\sigma}}{4^{\sigma}}\int_{0}^{1}\left(e^{-|x|^{2}/4t}t^{-n/2}v_{0}(x)+e^{-c|x|^{2}/t}\mathcal{O}\left(t^{-n/2+1}\right)\right)e^{-y^{2}/4t}~{}\frac{dt}{t^{1+\sigma}}
+y2σ4σ1ht(x)ey2/4tdtt1+σ\displaystyle+\frac{y^{2\sigma}}{4^{\sigma}}\int_{1}^{\infty}h_{t}(x)~{}e^{-y^{2}/4t}~{}\frac{dt}{t^{1+\sigma}}
=\displaystyle= y2σ4σv0(x)01e(|x|2+y2)/4ttn/21σ𝑑t\displaystyle\frac{y^{2\sigma}}{4^{\sigma}}v_{0}(x)\int_{0}^{1}e^{-(|x|^{2}+y^{2})/4t}~{}t^{-n/2-1-\sigma}~{}dt
+y2σ4σ01e(c|x|2+y2/4)/t𝒪(tn/2+1)t1σ𝑑t+y2σ4σ1ht(x)ey2/4tdtt1+σ.\displaystyle+\frac{y^{2\sigma}}{4^{\sigma}}\int_{0}^{1}e^{-(c|x|^{2}+y^{2}/4)/t}~{}\mathcal{O}(t^{-n/2+1})~{}t^{-1-\sigma}~{}dt+\frac{y^{2\sigma}}{4^{\sigma}}\int_{1}^{\infty}h_{t}(x)~{}e^{-y^{2}/4t}~{}\frac{dt}{t^{1+\sigma}}.

We write the right-hand side of the equation above as I1+I2+I3I_{1}+I_{2}+I_{3}, where I1,I2I_{1},I_{2} and I3I_{3} are the first, second and third term respectively. Then applying change of variable s=(|x|2+y2)/(4t)s=\left(|x|^{2}+y^{2}\right)/(4t), we have

I1=4n2y2σ(|x|2+y2)n/2σv0(x)(|x|2+y2)/4essn/2+σ1𝑑s.I_{1}=4^{\frac{n}{2}}y^{2\sigma}~{}\left(|x|^{2}+y^{2}\right)^{-n/2-\sigma}~{}v_{0}(x)\int_{(|x|^{2}+y^{2})/4}^{\infty}e^{-s}~{}s^{n/2+\sigma-1}~{}ds.

As |x|2+y2<1|x|^{2}+y^{2}<1,

1essn/2+σ1𝑑s(|x|2+y2)/4essn/2+σ1𝑑s0essn/2+σ1𝑑s.\int_{1}^{\infty}e^{-s}~{}s^{n/2+\sigma-1}~{}ds\leq\int_{(|x|^{2}+y^{2})/4}^{\infty}e^{-s}~{}s^{n/2+\sigma-1}~{}ds\leq\int_{0}^{\infty}e^{-s}~{}s^{n/2+\sigma-1}~{}ds.

This implies that for |x|2+y2<1|x|^{2}+y^{2}<1

I1y2σ(|x|2+y2)n/2σ,I_{1}\asymp y^{2\sigma}\left(|x|^{2}+y^{2}\right)^{-n/2-\sigma},

as v0(x)=(4π)n/2+𝒪(|x|2)v_{0}(x)=(4\pi)^{-n/2}+\mathcal{O}(|x|^{2}). For I2I_{2}, using c<1/4c<1/4 we have that

I2\displaystyle I_{2} \displaystyle\leq Cy2σ4σ01ec(|x|2+y2)/t𝒪(tn/2+1)t1σ𝑑t\displaystyle C\frac{y^{2\sigma}}{4^{\sigma}}\int_{0}^{1}e^{-c(|x|^{2}+y^{2})/t}~{}\mathcal{O}(t^{-n/2+1})~{}t^{-1-\sigma}~{}dt
\displaystyle\leq Cy2σ(|x|2+y2)n/2σ+1c(|x|2+y2)essn/2+σ2𝑑s\displaystyle C\,y^{2\sigma}~{}\left(|x|^{2}+y^{2}\right)^{-n/2-\sigma+1}~{}\int_{c(|x|^{2}+y^{2})}^{\infty}e^{-s}~{}s^{n/2+\sigma-2}~{}ds
\displaystyle\leq Cy2σ(|x|2+y2)n/2σ0essn/2+σ1𝑑s\displaystyle C\,y^{2\sigma}~{}\left(|x|^{2}+y^{2}\right)^{-n/2-\sigma}~{}\int_{0}^{\infty}e^{-s}~{}s^{n/2+\sigma-1}~{}ds
\displaystyle\leq Cy2σ(|x|2+y2)n/2σ.\displaystyle C\,y^{2\sigma}~{}\left(|x|^{2}+y^{2}\right)^{-n/2-\sigma}.

For the integral I3I_{3}, using Theorem 2.2 we get that for |x|2+y2<1|x|^{2}+y^{2}<1,

I3=y2σ4σ1ht(x)ey2/4tdtt1+σCy2σCy2σ(|x|2+y2)n/2σ.I_{3}=\frac{y^{2\sigma}}{4^{\sigma}}\int_{1}^{\infty}h_{t}(x)~{}e^{-y^{2}/4t}~{}\frac{dt}{t^{1+\sigma}}\leq Cy^{2\sigma}\leq Cy^{2\sigma}\left(|x|^{2}+y^{2}\right)^{-n/2-\sigma}.

This proves that for |x|2+y2<1|x|^{2}+y^{2}<1,

Γ(σ)Pyσ(x)y2σ(|x|2+y2)n/2σ.\Gamma(\sigma)P_{y}^{\sigma}(x)\asymp y^{2\sigma}~{}(|x|^{2}+y^{2})^{-n/2-\sigma}.

We will now assume that |x|2+y21|x|^{2}+y^{2}\geq 1. Let us fix a positive number κ>4\kappa>4. We proceed as in the proof of [4, Theorem 4.3.1].

Γ(σ)Pyσ(x)\displaystyle\Gamma(\sigma)P_{y}^{\sigma}(x) =\displaystyle= y2σ4σ0ht(x)ey2/4tdtt1+σ\displaystyle\frac{y^{2\sigma}}{4^{\sigma}}\int_{0}^{\infty}h_{t}(x)~{}e^{-y^{2}/4t}~{}\frac{dt}{t^{1+\sigma}}
=\displaystyle= y2σ4σ{I4+I5+I6},\displaystyle\frac{y^{2\sigma}}{4^{\sigma}}\{I_{4}+I_{5}+I_{6}\},

where the quantities I4,I5I_{4},I_{5} and I6I_{6} are defined by the integration of the above integrand ht(x)ey2/4tt1σh_{t}(x)~{}e^{-y^{2}/4t}~{}t^{-1-\sigma} over the intervals [0,κ1b),[κ1b,κb)\big{[}0,\kappa^{-1}b\big{)},\big{[}\kappa^{-1}b,\kappa b\big{)} and [κb,)\big{[}\kappa b,\infty\big{)} with b=|x|2+y2/(2|ρ|)b=\sqrt{|x|^{2}+y^{2}}/(2|\rho|) respectively. For the integral I5I_{5}, using Theorem 2.2 and the asymptotic of ϕ0\phi_{0} in Theorem 2.1 (3) we get the following:

I5\displaystyle I_{5} \displaystyle\asymp κ1|x|2+y22|ρ|κ|x|2+y22|ρ|tn/2(1+t)(nl)/2|Σ0+|{αΣ0+(1+α(x))}e|ρ|2tρ(logx)|x|2/4tey2/4tdtt1+σ\displaystyle\int_{\frac{\kappa^{-1}\sqrt{|x|^{2}+y^{2}}}{2|\rho|}}^{\frac{\kappa\sqrt{|x|^{2}+y^{2}}}{2|\rho|}}t^{-n/2}(1+t)^{(n-l)/2-|\Sigma_{0}^{+}|}\left\{\prod_{\alpha\in\Sigma_{0}^{+}}(1+\alpha(x))\right\}e^{-|\rho|^{2}t-\rho(\log x)-|x|^{2}/4t}e^{-y^{2}/4t}\frac{dt}{t^{1+\sigma}}
\displaystyle\asymp κ1|x|2+y22|ρ|κ|x|2+y22|ρ|tl/2|Σ0+|ϕ0(x)e|ρ|2te(|x|2+y2)/4tdtt1+σ\displaystyle\int_{\frac{\kappa^{-1}\sqrt{|x|^{2}+y^{2}}}{2|\rho|}}^{\frac{\kappa\sqrt{|x|^{2}+y^{2}}}{2|\rho|}}t^{-l/2-|\Sigma_{0}^{+}|}\phi_{0}(x)e^{-|\rho|^{2}t}e^{-(|x|^{2}+y^{2})/4t}\frac{dt}{t^{1+\sigma}}
=\displaystyle= κ1κ(s|x|2+y2/2|ρ|)l/2σ1|Σ0+|ϕ0(x)es|ρ||x|2+y2/2e|ρ||x|2+y2/(2s)(|x|2+y22|ρ|)𝑑s\displaystyle\int_{\kappa^{-1}}^{\kappa}(s\sqrt{|x|^{2}+y^{2}}/2|\rho|)^{-l/2-\sigma-1-|\Sigma_{0}^{+}|}\phi_{0}(x)e^{-s|\rho|\sqrt{|x|^{2}+y^{2}}/2}e^{-|\rho|\sqrt{|x|^{2}+y^{2}}/(2s)}\left(\frac{\sqrt{|x|^{2}+y^{2}}}{2|\rho|}\right)ds
\displaystyle\asymp (|x|2+y22|ρ|)l/2σ|Σ0+|ϕ0(x)κ1κe|x|2+y2|ρ|(s+1/s)/2𝑑s.\displaystyle\left(\frac{\sqrt{|x|^{2}+y^{2}}}{2|\rho|}\right)^{-l/2-\sigma-|\Sigma_{0}^{+}|}\phi_{0}(x)\int_{\kappa^{-1}}^{\kappa}e^{-\sqrt{|x|^{2}+y^{2}}|\rho|(s+1/s)/2}ds.

The last both side estimate follows because

κ(l/2+σ+1+|Σ0+|)s(l/2+σ+1+|Σ0+|)κ(l/2+σ+1+|Σ0+|).\kappa^{-(l/2+\sigma+1+|\Sigma_{0}^{+}|)}\leq s^{-(l/2+\sigma+1+|\Sigma_{0}^{+}|)}\leq\kappa^{(l/2+\sigma+1+|\Sigma_{0}^{+}|)}.

Now, using the fact that

κ1κe|ρ||x|2+y2(s+1/s)/2𝑑s|ρ|1/2(|x|2+y2)1/4e|ρ||x|2+y2,\int_{\kappa^{-1}}^{\kappa}e^{-|\rho|\sqrt{|x|^{2}+y^{2}}(s+1/s)/2}ds\asymp|\rho|^{-1/2}(|x|^{2}+y^{2})^{-1/4}e^{-|\rho|\sqrt{|x|^{2}+y^{2}}},

(this follows by the Laplace method [15, Ch 5]) we get from the above equation that

I5\displaystyle I_{5} \displaystyle\asymp (|x|2+y2)l/21/2σ|Σ0+|ϕ0(x)e|ρ||x|2+y2.\displaystyle\left(\sqrt{|x|^{2}+y^{2}}\right)^{-l/2-1/2-\sigma-|\Sigma_{0}^{+}|}\phi_{0}(x)e^{-|\rho|\sqrt{|x|^{2}+y^{2}}}.

For the third integral I6I_{6}, we will use the fact that κ>4\kappa>4. Using Theorem 2.2, we get

I6\displaystyle I_{6} \displaystyle\leq ϕ0(x)κ|x|2+y2/(2|ρ|)tl/2|Σ0+|1σe|ρ|2te(|x|2+y2)/4t𝑑t\displaystyle\phi_{0}(x)\int_{\kappa\sqrt{|x|^{2}+y^{2}}/(2|\rho|)}^{\infty}t^{-l/2-|\Sigma_{0}^{+}|-1-\sigma}e^{-|\rho|^{2}t}e^{-(|x|^{2}+y^{2})/4t}dt
\displaystyle\leq ϕ0(x)(κ|x|2+y2/(2|ρ|))l/2|Σ0+|1κ|x2|+y2/(2|ρ|)tσe|ρ|2te(|x|2+y2)/(4t)𝑑t\displaystyle\phi_{0}(x)\left(\kappa\sqrt{|x|^{2}+y^{2}}/(2|\rho|)\right)^{-l/2-|\Sigma_{0}^{+}|-1}\int_{\kappa\sqrt{|x^{2}|+y^{2}}/(2|\rho|)}^{\infty}t^{-\sigma}e^{-|\rho|^{2}t}e^{-(|x|^{2}+y^{2})/(4t)}dt
\displaystyle\leq Cϕ0(x)(|x|2+y2)l/2|Σ0+|1e|ρ|2k|x|2+y2/(4|ρ|)κ|x2|+y2/(2|ρ|)tσe|ρ|2t/2e(|x|2+y2)/(4t)𝑑t\displaystyle C\phi_{0}(x)\left(\sqrt{|x|^{2}+y^{2}}\right)^{-l/2-|\Sigma_{0}^{+}|-1}~{}e^{-|\rho|^{2}k\sqrt{|x|^{2}+y^{2}}/(4|\rho|)}\int_{\kappa\sqrt{|x^{2}|+y^{2}}/(2|\rho|)}^{\infty}t^{-\sigma}e^{-|\rho|^{2}t/2}e^{-(|x|^{2}+y^{2})/(4t)}dt
\displaystyle\leq C(|x|2+y2)l/2|Σ0+|1/2ϕ0(x)e(|ρ|+η)|x|2+y2,\displaystyle C\left(\sqrt{|x|^{2}+y^{2}}\right)^{-l/2-|\Sigma_{0}^{+}|-1/2}~{}\phi_{0}(x)e^{-(|\rho|+\eta)\sqrt{|x|^{2}+y^{2}}},

where η=|ρ|κ/4|ρ|>0\eta=|\rho|\kappa/4-|\rho|>0. For the first integral I4I_{4}, we use heat kernel Gaussian estimate (2.10) and the estimate of ϕ0\phi_{0} in Theorem 2.1 to obtain the following

I4\displaystyle I_{4} \displaystyle\leq 0κ1|x|2+y2/(2|ρ|)td1(1+|x|)d2e|ρ|2tρ(logx)e(|x|2+y2)/(4t)dtt1+σ\displaystyle\int_{0}^{\kappa^{-1}\sqrt{|x|^{2}+y^{2}}/(2|\rho|)}t^{-d_{1}}(1+|x|)^{d_{2}}e^{-|\rho|^{2}t-\rho(\log x)}e^{-(|x|^{2}+y^{2})/(4t)}\frac{dt}{t^{1+\sigma}}
\displaystyle\leq (1+|x|)d2|Σ0+|ϕ0(x)0κ1|x|2+y2/(2|ρ|)e(|x|2+y2)/(4t)t1σd1𝑑t\displaystyle(1+|x|)^{d_{2}-|\Sigma_{0}^{+}|}\phi_{0}(x)\int_{0}^{\kappa^{-1}\sqrt{|x|^{2}+y^{2}}/(2|\rho|)}e^{-(|x|^{2}+y^{2})/(4t)}t^{-1-\sigma-d_{1}}dt
=\displaystyle= (1+|x|)d2|Σ0+|ϕ0(x)0κ1|x|2+y2/(2|ρ|)e(|x|2+y2)/(8t)e(|x|2+y2)/(8t)t1σd1𝑑t\displaystyle(1+|x|)^{d_{2}-|\Sigma_{0}^{+}|}\phi_{0}(x)\int_{0}^{\kappa^{-1}\sqrt{|x|^{2}+y^{2}}/(2|\rho|)}e^{-(|x|^{2}+y^{2})/(8t)}e^{-(|x|^{2}+y^{2})/(8t)}t^{-1-\sigma-d_{1}}dt
\displaystyle\leq C(1+|x|)d2|Σ0+|ϕ0(x)e|ρ|κx2+y240κ1|x|2+y2/(2|ρ|)e(|x|2+y2)/(8t)t1σd1𝑑t\displaystyle C(1+|x|)^{d_{2}-|\Sigma_{0}^{+}|}\phi_{0}(x)\,e^{-|\rho|\kappa\frac{\sqrt{x^{2}+y^{2}}}{4}}~{}\int_{0}^{\kappa^{-1}\sqrt{|x|^{2}+y^{2}}/(2|\rho|)}e^{-(|x|^{2}+y^{2})/(8t)}t^{-1-\sigma-d_{1}}dt
\displaystyle\leq C(1+|x|)d2|Σ0+|ϕ0(x)e(|ρ|+ϵ)|x|2+y2(|x|2+y2)σd1,\displaystyle C(1+|x|)^{d_{2}-|\Sigma_{0}^{+}|}\phi_{0}(x)\,e^{-(|\rho|+\epsilon)\sqrt{|x|^{2}+y^{2}}}(|x|^{2}+y^{2})^{-\sigma-d_{1}},

for some ϵ>0\epsilon>0, as κ>4\kappa>4.

This completes the proof. ∎

To prove Hardy’s inequalities we use an integral representation for the operator (Δ)σ(-\Delta)^{\sigma}. The following function

(3.7) P0σ(x)=0ht(x)dtt1+σ,P_{0}^{\sigma}(x)=\int_{0}^{\infty}h_{t}(x)\frac{dt}{t^{1+\sigma}},

serves as the kernel of the integral representation. We state both sides estimate of P0σP_{0}^{\sigma}, whose proof is exactly the same as of Theorem 3.3.

Theorem 3.4.

For any α>n/2\alpha>-n/2 the following asymptotic estimates holds:

P0α(x)\displaystyle P_{0}^{\alpha}(x) \displaystyle\asymp |x|l/21/2α|Σ0+|ϕ0(x)e|ρ||x|, for |x|1,\displaystyle|x|^{-l/2-1/2-\alpha-|\Sigma_{0}^{+}|}\phi_{0}(x)e^{-|\rho||x|},\textmd{ for }|x|\geq 1,
\displaystyle\asymp |x|n2α, for |x|<1.\displaystyle|x|^{-n-2\alpha},\textmd{ for }~{}|x|<1.
Corollary 3.5.

Let χ\chi be the characteristic function of the unit ball in XX and α>0\alpha>0. Then the function (1χ)P0α(1-\chi)P_{0}^{\alpha} is in Lp(X)L^{p}(X) for 1p1\leq p\leq\infty.

Proof.

For 1<p1<p\leq\infty, the result follows trivially from the asymptotic formula in Theorem 3.4. We prove the case p=1p=1. We recall from (2.2) that

{xX:|x|>1}P0α(x)𝑑xC{H𝔞+¯:|H|>1}P0α(expH)e2ρ(H)𝑑H.\int_{\{x\in X:|x|>1\}}P_{0}^{\alpha}(x)~{}dx\leq C\int_{\{H\in\overline{\mathfrak{a}_{+}}:|H|>1\}}P_{0}^{\alpha}(\exp H)~{}e^{2\rho(H)}~{}dH.

Let Γ\Gamma be a small circular cone in 𝔞+¯\overline{\mathfrak{a}_{+}} around the ρ\rho-axis. By introducing polar coordinates in Γ\Gamma and using (2.6) we get

{HΓ:|H|>1}P0α(expH)e2ρ(H)𝑑H\displaystyle\int_{\{H\in\Gamma:|H|>1\}}P_{0}^{\alpha}(\exp H)~{}e^{2\rho(H)}~{}dH
\displaystyle\leq C{HΓ:|H|>1}|H|l/21/2αeρ(H)|ρ||H|𝑑H\displaystyle C\int_{\{H\in\Gamma:|H|>1\}}|H|^{-l/2-1/2-\alpha}~{}e^{\rho(H)-|\rho||H|}~{}dH
\displaystyle\leq C1rl/21/2αrl10ν(sinξ)l2er(1cosξ)𝑑ξ𝑑r.\displaystyle C\int_{1}^{\infty}r^{-l/2-1/2-\alpha}r^{l-1}~{}\int_{0}^{\nu}(\sin\xi)^{l-2}~{}e^{-r(1-\cos\xi)}~{}d\xi~{}dr.

Since sinξξ\sin\xi\sim\xi and 1cosξξ21-\cos\xi\sim\xi^{2}, the inner integral behaves like r1/2l/2r^{1/2-l/2}. Consequently, the integral above is finite. On the other hand, eρ(H)|ρ||H|e^{\rho(H)-|\rho||H|} decreases exponentially outside Γ\Gamma, and therefore

{H𝔞+¯\Γ:|H|>1}P0α(expH)e2ρ(H)𝑑H={H𝔞+¯\Γ:|H|>1}Hl/21/2αeρ(H)|ρ||H|𝑑H<.\int_{\{H\in\overline{\mathfrak{a}_{+}}\backslash\Gamma:|H|>1\}}P_{0}^{\alpha}(\exp H)~{}e^{2\rho(H)}~{}dH=\int_{\{H\in\overline{\mathfrak{a}_{+}}\backslash\Gamma:|H|>1\}}H^{-l/2-1/2-\alpha}~{}e^{\rho(H)-|\rho||H|}~{}dH<\infty.

This completes the proof. ∎

4. Fractional Hardy inequalities

This section aims to prove two versions of the Hardy’s inequalities for fractional powers of the Laplace-Beltrami operator on XX, namely Theorem 1.1 and Theorem 1.3 with homogeneous and non-homogeneous weight functions respectively. In order to prove these inequalities, we will follow similar ideas used by Frank et al. [19] in the case of Euclidean Laplacian. Therefore, we need to establish ground state representations for the operators (Δ)σ(-\Delta)^{\sigma}. We start with the following integral representations of (Δ)σ(-\Delta)^{\sigma} on XX. For the cases of real hyperbolic spaces, analogues integral representations were proved in [6, Theorem 2.5].

Lemma 4.1.

Let 0<σ<1/20<\sigma<1/2. Then for all fCc(X)f\in C_{c}^{\infty}(X) we have

(Δ)σf(x)=1|Γ(σ)|X(f(x)f(z))P0σ(z1x)𝑑z,(-\Delta)^{\sigma}f(x)=\frac{1}{|\Gamma(-\sigma)|}~{}\int_{X}\left(f(x)-f(z)\right)P_{0}^{\sigma}(z^{-1}x)~{}dz,

where P0σP_{0}^{\sigma} is defined in (3.7).

Proof.

Let fCc(X)f\in C_{c}^{\infty}(X). Using the numerical identity

λσ=1|Γ(σ)|0(1etλ)dtt1+σ,λ>0,\lambda^{\sigma}=\frac{1}{|\Gamma(-\sigma)|}\int_{0}^{\infty}\left(1-e^{-t\lambda}\right)~{}\frac{dt}{t^{1+\sigma}},\>\>\lambda>0,

and the spectral theorem we have

(Δ)σf(x)=1|Γ(σ)|0(f(x)etΔf(x))dtt1+σ.(-\Delta)^{\sigma}f(x)=\frac{1}{|\Gamma(-\sigma)|}\int_{0}^{\infty}\left(f(x)-e^{t\Delta}f(x)\right)~{}\frac{dt}{t^{1+\sigma}}.

By (2.8) it follows that

(4.1) f(x)etΔf(x)=f(x)fht(x)=X(f(x)f(xz1))ht(z)𝑑z.f(x)-e^{t\Delta}f(x)=f(x)-f\ast h_{t}(x)=\int_{X}(f(x)-f(xz^{-1}))~{}h_{t}(z)~{}dz.

Thus, we have the following representation

(Δ)σf(x)=1|Γ(σ)|0X(f(x)f(xz1))ht(z)𝑑zdtt1+σ.(-\Delta)^{\sigma}f(x)=\frac{1}{|\Gamma(-\sigma)|}\int_{0}^{\infty}\int_{X}\left(f(x)-f(xz^{-1})\right)h_{t}(z)~{}dz~{}\frac{dt}{t^{1+\sigma}}.

We now show that the right-hand side is absolutely integrable and hence, interchange of the order of integral is possible. Then the result follows by the change of variable zz1xz\mapsto z^{-1}x. To show absolute integrability let us define

I1=1|Γ(σ)|0{zX:|z|<1}|f(x)f(xz1)|ht(z)𝑑zdtt1+σ,\displaystyle I_{1}=\frac{1}{|\Gamma(-\sigma)|}\int_{0}^{\infty}\int_{\{z\in X:|z|<1\}}\left|f(x)-f(xz^{-1})\right|h_{t}(z)~{}dz~{}\frac{dt}{t^{1+\sigma}},
I2=1|Γ(σ)|0{zX:|z|1}|f(x)f(xz1)|ht(z)𝑑zdtt1+σ.\displaystyle I_{2}=\frac{1}{|\Gamma(-\sigma)|}\int_{0}^{\infty}\int_{\{z\in X:|z|\geq 1\}}\left|f(x)-f(xz^{-1})\right|h_{t}(z)~{}dz~{}\frac{dt}{t^{1+\sigma}}.

For the integral I2I_{2}, we use the fact that P0σL1(X)P_{0}^{\sigma}\in L^{1}(X) away from the origin (Corollary 3.5). Indeed, we have that

{zX:|z|1}0|f(x)f(xz1)|ht(z)dtt1+σ𝑑zfL(X){zX:|z|1}P0σ(z)𝑑z<.\int_{\{z\in X:|z|\geq 1\}}\int_{0}^{\infty}|f(x)-f(xz^{-1})|~{}h_{t}(z)~{}\frac{dt}{t^{1+\sigma}}~{}dz\leq\|f\|_{L^{\infty}(X)}\int_{\{z\in X:|z|\geq 1\}}P_{0}^{\sigma}(z)~{}dz<\infty.

Therefore, by Fubini’s theorem I2I_{2} is also finite. For I1I_{1} we first observe by the fundamental theorem of calculus (see the proof of equation (34) in [2]) that

(4.2) |f(x)f(x(expH))||H|01|f(xexp(sH))|𝑑s|H|fL(X),|f(x)-f\left(x(\exp H)\right)|\leq|H|\int_{0}^{1}|\nabla f\left(x\exp(sH)\right)|~{}ds\leq|H|~{}\|\nabla f\|_{L^{\infty}(X)},

for xX,H𝔞x\in X,H\in\mathfrak{a}. Using the above estimate and the fact that P0σ(x)|x|n2σP_{0}^{\sigma}(x)\asymp|x|^{-n-2\sigma} around the origin (Theorem 3.4) it follows that

|z|<10|f(x)f(xz1)|ht(z)dtt1+σ𝑑z\displaystyle\int_{|z|<1}\int_{0}^{\infty}|f(x)-f(xz^{-1})|~{}h_{t}(z)~{}\frac{dt}{t^{1+\sigma}}~{}dz
\displaystyle\leq CfL(X){H𝔞+¯:|H|<1}|H||H|n2σJ(expH)𝑑H\displaystyle C\|\nabla f\|_{L^{\infty}(X)}~{}\int_{\{H\in\overline{\mathfrak{a}_{+}}:|H|<1\}}|H|~{}|H|^{-n-2\sigma}~{}J(\exp H)~{}dH
\displaystyle\leq CfL(X)01r1n2σrn1𝑑r,\displaystyle C\|\nabla f\|_{L^{\infty}(X)}\int_{0}^{1}r^{1-n-2\sigma}~{}r^{n-1}~{}dr,

and the right-hand side is finite if 0<σ<1/20<\sigma<1/2. This completes the proof. ∎

Remark 4.2.

If rank(X)=1rank(X)=1, then for 1/2σ<11/2\leq\sigma<1 the integral formula in Lemma 4.1 exists in principal value sense. To see this, let 𝔞=span{H0}\mathfrak{a}=span\{H_{0}\} with |H0|=1|H_{0}|=1. Clearly, for σ>0\sigma>0, the integral I2I_{2} is absolutely convergent and we can interchange the order of the integral. On the other hand the formula (2.2) yields

I1=1|Γ(σ)|011(f(x)f(xexp(sH0)))ht(exp(sH0))J(exp(sH0))𝑑sdtt1+σ.I_{1}=\frac{1}{|\Gamma(-\sigma)|}\int_{0}^{\infty}\int_{-1}^{1}\big{(}f(x)-f\left(x\exp(-sH_{0})\right)\big{)}~{}h_{t}\left(\exp(sH_{0})\right)~{}J\left(\exp(sH_{0})\right)~{}ds\frac{dt}{t^{1+\sigma}}.

We now define F(s):=f(xexp(sH0))F(s):=f\left(x\exp(sH_{0})\right), for ss\in\mathbb{R}. Since fCc(X)f\in C_{c}^{\infty}(X), it follows that for each xXx\in X, the function FCc()F\in C_{c}^{\infty}(\mathbb{R}). By using the Taylor development of FF, we get that

I1=1|Γ(σ)|011(sF(x)+s2F′′(x)2!+𝒪(s3))ht(exp(sH0))J(exp(sH0))𝑑sdtt1+σ.I_{1}=\frac{1}{|\Gamma(-\sigma)|}\int_{0}^{\infty}\int_{-1}^{1}\left(sF^{\prime}(x)+\frac{s^{2}F^{\prime\prime}(x)}{2!}+\mathcal{O}(s^{3})\right)~{}h_{t}\left(\exp(sH_{0})\right)~{}J\left(\exp(sH_{0})\right)~{}ds\frac{dt}{t^{1+\sigma}}.

Since the heat kernel hth_{t} and the Jacobian JJ is even, the first order term vanishes. Hence, using the fact that P0σ(x)|x|n2σP_{0}^{\sigma}(x)\sim|x|^{-n-2\sigma}, around the origin (Theorem 3.4), it follows that

I1\displaystyle I_{1} \displaystyle\leq Cf001s2ht(exp(sH0))sn1𝑑sdtt1+σ=Cf01sn+1sn2σ𝑑s,\displaystyle C_{f}\int_{0}^{\infty}\int_{0}^{1}s^{2}~{}h_{t}\left(\exp(sH_{0})\right)~{}s^{n-1}~{}ds~{}\frac{dt}{t^{1+\sigma}}=C_{f}\int_{0}^{1}~{}s^{n+1}~{}s^{-n-2\sigma}~{}ds,

which is finite if 0<σ<10<\sigma<1. Hence, the required integral formula exists as a principal value sense. For the case of higher rank symmetric spaces, neither the heat kernel ht(exp())h_{t}\left(\exp(\cdot)\right) nor the Jacobian J(exp())J\left(\exp(\cdot)\right) is, in general, radial function on 𝔞\mathfrak{a}. They are only Weyl group invariant. This is the main difficulty that we could not prove the integral formula in the lemma above for 1/2σ<11/2\leq\sigma<1 in case of rank(X)>1rank(X)>1.

Lemma 4.3.

Let 0<σ<10<\sigma<1. Then, for all fHσ(X)f\in H^{\sigma}(X)

(Δ)σf,f=12|Γ(σ)|XX|f(z)f(x)|2P0σ(z1x)𝑑z𝑑x.\langle(-\Delta)^{\sigma}f,f\rangle=\frac{1}{2|\Gamma(-\sigma)|}\int_{X}\int_{X}\left|f(z)-f(x)\right|^{2}P_{0}^{\sigma}(z^{-1}x)~{}dz~{}dx.
Proof.

We first prove that for 0<σ<10<\sigma<1 and fCc(X)f\in C_{c}^{\infty}(X) the quantity

(4.3) 12|Γ(σ)|XX|f(z)f(x)|2P0σ(z1x)𝑑z𝑑x<.\frac{1}{2|\Gamma(-\sigma)|}\int_{X}\int_{X}\left|f(z)-f(x)\right|^{2}P_{0}^{\sigma}(z^{-1}x)~{}dz~{}dx<\infty.

To show this let us assume suppf𝐁(o,m)supp~{}f\subset{\bf B}(o,m) for some m>1m>1 and define

I1=12|Γ(σ)|𝐁(o,2m)X|f(z)f(x)|2P0σ(z1x)𝑑z𝑑x,\displaystyle I_{1}=\frac{1}{2|\Gamma(-\sigma)|}\int_{{\bf B}(o,2m)}\int_{X}\left|f(z)-f(x)\right|^{2}P_{0}^{\sigma}(z^{-1}x)~{}dz~{}dx,
I2=12|Γ(σ)|X\𝐁(o,2m)X|f(z)f(x)|2P0σ(z1x)𝑑z𝑑x.\displaystyle I_{2}=\frac{1}{2|\Gamma(-\sigma)|}\int_{X\backslash{\bf B}(o,2m)}\int_{X}\left|f(z)-f(x)\right|^{2}~{}P_{0}^{\sigma}(z^{-1}x)~{}dz~{}dx.

Since suppf𝐁(o,m)supp~{}f\subset{\bf B}(o,m) it follows that

I2\displaystyle I_{2} =\displaystyle= 12|Γ(σ)|X\𝐁(o,2m)𝐁(o,m)|f(z)|2P0σ(z1x)𝑑z𝑑x\displaystyle\frac{1}{2|\Gamma(-\sigma)|}\int_{X\backslash{\bf B}(o,2m)}\int_{{\bf B}(o,m)}\left|f(z)\right|^{2}~{}P_{0}^{\sigma}(z^{-1}x)~{}dz~{}dx
\displaystyle\leq fL(X)22|Γ(σ)|𝐁(o,m)X\𝐁(o,2m)P0σ(z1x)𝑑x𝑑z\displaystyle\frac{\|f\|_{L^{\infty}(X)}^{2}}{2|\Gamma(-\sigma)|}\int_{{\bf B}(o,m)}\int_{X\backslash{\bf B}(o,2m)}P_{0}^{\sigma}(z^{-1}x)~{}dx~{}dz
\displaystyle\leq fL(X)22|Γ(σ)||𝐁(o,m)|X\𝐁(o,m)P0σ(x)𝑑x<.\displaystyle\frac{\|f\|_{L^{\infty}(X)}^{2}}{2|\Gamma(-\sigma)|}|{\bf B}(o,m)|\int_{X\backslash{\bf B}(o,m)}P_{0}^{\sigma}(x)~{}dx<\infty.

The last term is finite because of the fact that P0σP_{0}^{\sigma} is integrable away from the origin (Corollary 3.5). To show that I1I_{1} is finite we write it as follows

I1\displaystyle I_{1} =\displaystyle= 𝐁(o,2m)𝐁(0,3m)|f(z)f(x)|2P0σ(z1x)𝑑z𝑑x\displaystyle\int_{{\bf B}(o,2m)}\int_{{\bf B}(0,3m)}\left|f(z)-f(x)\right|^{2}P_{0}^{\sigma}(z^{-1}x)~{}dz~{}dx
+𝐁(o,2m)X\𝐁(0,3m)|f(z)f(x)|2P0σ(z1x)𝑑z𝑑x.\displaystyle+\int_{{\bf B}(o,2m)}\int_{X\backslash{\bf B}(0,3m)}\left|f(z)-f(x)\right|^{2}P_{0}^{\sigma}(z^{-1}x)~{}dz~{}dx.

Using change of variable zxz1z\mapsto xz^{-1} in the first integral, the estimate (4.2) and the asymptotic estimates of P0σP_{0}^{\sigma} in Theorem 3.4 it follows that

I1\displaystyle I_{1} \displaystyle\leq 𝐁(o,2m)𝐁(o,5m)|f(xz1)f(x)|2P0σ(z)𝑑z𝑑x\displaystyle\int_{{\bf B}(o,2m)}\int_{{\bf B}(o,5m)}\left|f(xz^{-1})-f(x)\right|^{2}P_{0}^{\sigma}(z)~{}dz~{}dx
+𝐁(o,2m)X\𝐁(o,3m)|f(z)f(x)|2P0σ(z1x)𝑑z𝑑x\displaystyle+\int_{{\bf B}(o,2m)}\int_{X\backslash{\bf B}(o,3m)}\left|f(z)-f(x)\right|^{2}P_{0}^{\sigma}(z^{-1}x)~{}dz~{}dx
\displaystyle\leq CfL(X)2𝐁(o,2m)𝑑x{H𝔞+¯:|H|<5m}|H|2Hn2σJ(expH)𝑑H\displaystyle C\|\nabla f\|_{L^{\infty}(X)}^{2}\int_{{\bf B}(o,2m)}~{}dx~{}\int_{\{H\in\overline{\mathfrak{a}_{+}}:|H|<5m\}}|H|^{2}~{}H^{-n-2\sigma}~{}J(\exp H)~{}dH
+C𝐁(o,2m)fL(X)2X\𝐁(o,3m)P0σ(z1x)𝑑z𝑑x\displaystyle+C\int_{{\bf B}(o,2m)}\|f\|_{L^{\infty}(X)}^{2}~{}\int_{X\backslash{\bf B}(o,3m)}P_{0}^{\sigma}(z^{-1}x)~{}dz~{}dx
\displaystyle\leq CfL(X)205mr2n2σrn1𝑑r+CfL(X)2X\𝐁(o,m)P0σ(z)𝑑z.\displaystyle C\|\nabla f\|_{L^{\infty}(X)}^{2}\int_{0}^{5m}r^{2-n-2\sigma}~{}r^{n-1}~{}dr+C\|f\|_{L^{\infty}(X)}^{2}\int_{X\backslash{\bf B}(o,m)}P_{0}^{\sigma}(z)~{}dz.

The first term of the above quantity is finite provided σ<1\sigma<1 and the second one finite by Corollary 3.5. This completes the proof of the fact the quantity in (4.3) is finite.

Let 0<σ<1/20<\sigma<1/2 and fCc(X)f\in C_{c}^{\infty}(X). By the integral representation in Lemma 4.1 it follows that

(Δ)σf,f=1|Γ(σ)|XX(f(x)f(z))P0σ(z1x)f(x)¯𝑑z𝑑x.\left\langle(-\Delta)^{\sigma}f,f\right\rangle=\frac{1}{|\Gamma(-\sigma)|}\int_{X}\int_{X}\left(f(x)-f(z)\right)P_{0}^{\sigma}(z^{-1}x)~{}\overline{f(x)}~{}dz~{}dx.

As the kernel P0σP_{0}^{\sigma} is symmetric, that is P0σ(x)=P0σ(x1)P_{0}^{\sigma}(x)=P_{0}^{\sigma}(x^{-1}), the above quantity is also equals to

1|Γ(σ)|XX(f(z)f(x))P0σ(z1x)f(z)¯𝑑x𝑑z.\frac{1}{|\Gamma(-\sigma)|}\int_{X}\int_{X}\left(f(z)-f(x)\right)P_{0}^{\sigma}(z^{-1}x)~{}\overline{f(z)}~{}dx~{}dz.

By adding them up we get that

(Δ)σf,f=12|Γ(σ)|XX|f(z)f(x)|2P0σ(z1x)𝑑z𝑑x.\langle(-\Delta)^{\sigma}f,f\rangle=\frac{1}{2|\Gamma(-\sigma)|}\int_{X}\int_{X}\left|f(z)-f(x)\right|^{2}P_{0}^{\sigma}(z^{-1}x)~{}dz~{}dx.

The justification of the change of order of integration follows from (4.3). By the analytic continuation, we extend the range of σ\sigma to 0<σ<10<\sigma<1 provided fCc(X)f\in C_{c}^{\infty}(X). Indeed, the functions σΓ(σ)\sigma\mapsto-\Gamma(-\sigma) and σ(Δ)σf,f\sigma\mapsto\langle(-\Delta)^{\sigma}f,f\rangle are holomorphic on S={w:0<w<1}S=\{w\in\mathbb{C}:0<\Re w<1\}. Hence their product F(σ)=Γ(σ)(Δ)σf,fF(\sigma)=-\Gamma(-\sigma)~{}\langle(-\Delta)^{\sigma}f,f\rangle is also holomorphic on SS. On the other hand, since right-hand side of (4.3) is finite for 0<σ<10<\sigma<1, by the Morera’s theorem it follows that the function GG defined by

G(σ)=12XX|f(z)f(x)|2P0σ(z1x)𝑑z𝑑x.G(\sigma)=\frac{1}{2}\int_{X}\int_{X}\left|f(z)-f(x)\right|^{2}P_{0}^{\sigma}(z^{-1}x)~{}dz~{}dx.

is holomorphic on SS. Since F(σ)=G(σ)F(\sigma)=G(\sigma) for 0<σ<1/20<\sigma<1/2 we get that F(σ)=G(σ)F(\sigma)=G(\sigma) for all σS\sigma\in S, in particular, for 0<σ<10<\sigma<1.

By approximating any function fHσ(X)f\in H^{\sigma}(X) by a sequence of functions fkCc(X)f_{k}\in C_{c}^{\infty}(X), we complete the proof. This uses the fact that P0σ(x)|x|n2σP_{0}^{\sigma}(x)\asymp|x|^{-n-2\sigma} around the origin and the rest follows as in the proof of Lemma 5.1 in [37]. ∎

We now establish ground state representation for the operator (Δ)σ(-\Delta)^{\sigma} as a consequence of the integral representation proved in Lemma 4.3. As in the Euclidean case, we define the following error term. For 0<σ<10<\sigma<1 and y>0y>0 we let,

Hyσ[F]=(Δ)σF,F4σΓ(σ)y2σΓ(σ)X|F(x)|2(Pyσ(x)Pyσ(x))𝑑x.H_{y}^{\sigma}[F]=\langle(-\Delta)^{\sigma}F,F\rangle-\frac{4^{\sigma}\Gamma(\sigma)}{y^{2\sigma}\Gamma(-\sigma)}\int_{X}|F(x)|^{2}\left(\frac{P_{y}^{\sigma}(x)}{P_{y}^{-\sigma}(x)}\right)\,dx.
Theorem 4.4.

Let 0<σ<10<\sigma<1 and y>0y>0. If FCc(X)F\in C_{c}^{\infty}(X) and G(x)=F(x)(Pyσ(x))1G(x)=F(x)~{}\left(P^{-\sigma}_{y}(x)\right)^{-1} then

Hyσ[F]=12|Γ(σ)|XX|G(x)G(z)|2Pyσ(x)Pyσ(z)P0σ(z1x)𝑑x𝑑z.H_{y}^{\sigma}[F]=\frac{1}{2|\Gamma(-\sigma)|}\int_{X}\int_{X}|G(x)-G(z)|^{2}~{}P_{y}^{-\sigma}(x)P_{y}^{-\sigma}(z)~{}P_{0}^{-\sigma}(z^{-1}x)~{}dx~{}dz.
Proof.

Let f,gHσ(X)f,g\in H^{\sigma}(X). From Lemma 4.3 we get that

(4.4) (Δ)σf,g=12|Γ(σ)|XX(f(z)f(x))(g(z)g(x))¯P0σ(z1x)𝑑z𝑑x.\left\langle(-\Delta)^{\sigma}f,g\right\rangle=\frac{1}{2|\Gamma(-\sigma)|}\int_{X}\int_{X}\left(f(z)-f(x)\right)~{}\overline{\left(g(z)-g(x)\right)}~{}P_{0}^{\sigma}(z^{-1}x)~{}dz~{}dx.

Let us assume g=Pyσg=P_{y}^{-\sigma}, and f(x)=|F(x)|2g(x)1f(x)=|F(x)|^{2}~{}g(x)^{-1}. Then the right-hand side of (4.4) reduces to

12|Γ(σ)|XX(|F(z)|2g(z)|F(x)|2g(x))(g(z)g(x))¯P0σ(z1x)𝑑z𝑑x\displaystyle\frac{1}{2|\Gamma(-\sigma)|}\int_{X}\int_{X}\left(\frac{|F(z)|^{2}}{g(z)}-\frac{|F(x)|^{2}}{g(x)}\right)~{}\overline{\left(g(z)-g(x)\right)}~{}P_{0}^{\sigma}(z^{-1}x)~{}dz~{}dx
=\displaystyle= 12|Γ(σ)|XX(|F(x)F(z)|2|F(x)g(x)F(z)g(z)|2g(x)g(z))P0σ(z1x)𝑑z𝑑x.\displaystyle\frac{1}{2|\Gamma(-\sigma)|}\int_{X}\int_{X}\left(|F(x)-F(z)|^{2}-\left|\frac{F(x)}{g(x)}-\frac{F(z)}{g(z)}\right|^{2}~{}g(x)g(z)\right)~{}P_{0}^{\sigma}(z^{-1}x)~{}dz~{}dx.

Also, using Lemma 3.2 the left-hand side of (4.4) reduces to

(Δ)σf,g\displaystyle\left\langle(-\Delta)^{\sigma}f,g\right\rangle =\displaystyle= (Δ)σ(|F(x)|2/g(x)),g(x)\displaystyle\left\langle(-\Delta)^{\sigma}(|F(x)|^{2}/g(x)),g(x)\right\rangle
=\displaystyle= (|F(x)|2/g(x)),(Δ)σPyσ\displaystyle\left\langle\left(|F(x)|^{2}/g(x)\right),(-\Delta)^{\sigma}P_{y}^{-\sigma}\right\rangle
=\displaystyle= 4σΓ(σ)y2σΓ(σ)(|F(x)|2/g(x)),Pyσ\displaystyle\frac{4^{\sigma}\Gamma(\sigma)}{y^{2\sigma}\Gamma(-\sigma)}\left\langle(|F(x)|^{2}/g(x)),P_{y}^{\sigma}\right\rangle
=\displaystyle= 4σΓ(σ)y2σΓ(σ)X|F(x)|2Pyσ(x)Pyσ(x)𝑑x.\displaystyle\frac{4^{\sigma}\Gamma(\sigma)}{y^{2\sigma}\Gamma(-\sigma)}\int_{X}|F(x)|^{2}\frac{P_{y}^{\sigma}(x)}{P_{y}^{-\sigma}(x)}~{}dx.

Therefore, equating the left-hand and right-hand sides of the equation (4.4) we have

4σΓ(σ)y2σΓ(σ)X|F(x)|2Pyσ(x)Pyσ(x)𝑑x=12|Γ(σ)|XX|F(x)F(z)|2P0σ(z1x)𝑑x𝑑z\displaystyle\frac{4^{\sigma}\Gamma(\sigma)}{y^{2\sigma}\Gamma(-\sigma)}\int_{X}|F(x)|^{2}\frac{P_{y}^{\sigma}(x)}{P_{y}^{-\sigma}(x)}~{}dx=\frac{1}{2|\Gamma(-\sigma)|}\int_{X}\int_{X}\left|F(x)-F(z)\right|^{2}P_{0}^{\sigma}(z^{-1}x)~{}dx~{}dz
\displaystyle- 12|Γ(σ)|XX|F(x)g(x)F(z)g(z)|2g(x)g(z)P0σ(z1x)𝑑z𝑑x.\displaystyle\frac{1}{2|\Gamma(-\sigma)|}\int_{X}\int_{X}\left|\frac{F(x)}{g(x)}-\frac{F(z)}{g(z)}\right|^{2}~{}g(x)g(z)~{}P_{0}^{\sigma}(z^{-1}x)~{}dz~{}dx.

By Lemma 4.3 the first term in the right-hand side of the above equation is equals to (Δ)σF,F\langle(-\Delta)^{\sigma}F,F\rangle. Hence, it follows that

(Δ)σF,F4σΓ(σ)y2σΓ(σ)X|F(x)|2Pyσ(x)Pyσ(x)𝑑x\displaystyle\langle(-\Delta)^{\sigma}F,F\rangle-\frac{4^{\sigma}\Gamma(\sigma)}{y^{2\sigma}\Gamma(-\sigma)}\int_{X}|F(x)|^{2}\frac{P_{y}^{\sigma}(x)}{P_{y}^{-\sigma}(x)}~{}dx
=\displaystyle= 12|Γ(σ)|XX|G(x)G(z)|2Pyσ(x)Pyσ(x)P0σ(z1x)𝑑x𝑑z,\displaystyle\frac{1}{2|\Gamma(-\sigma)|}\int_{X}\int_{X}|G(x)-G(z)|^{2}~{}P_{y}^{-\sigma}(x)P_{y}^{-\sigma}(x)~{}P_{0}^{-\sigma}(z^{-1}x)~{}dx~{}dz,

where G(x)=F(x)Pyσ(x)1G(x)=F(x)P_{y}^{-\sigma}(x)^{-1}. This completes the proof. ∎

We have already observed that for 0<σ<10<\sigma<1, Γ(σ)<0\Gamma(-\sigma)<0 and hence Pyσ0P_{y}^{-\sigma}\leq 0. Therefore, as a corollary of Theorem 4.4 we get the following result.

Corollary 4.5.

For a fixed y>0y>0 and 0<σ<10<\sigma<1 we have

(Δ)σF,F4σy2σX|F(x)|2(Γ(σ)Γ(σ)Pyσ(x)Pyσ(x))𝑑x, for FHσ(X).\left\langle(-\Delta)^{\sigma}F,F\right\rangle\geq\frac{4^{\sigma}}{y^{2\sigma}}\int_{X}|F(x)|^{2}\left(\frac{\Gamma(\sigma)}{\Gamma(-\sigma)}\frac{P_{y}^{\sigma}(x)}{P_{y}^{-\sigma}(x)}\right)\,dx,\>\>\textit{ for }F\in H^{\sigma}(X).
Remark 4.6.

By Lemma 3.2 it follows that the equality in the expression above is achieved for the function F=PyσF=P_{y}^{-\sigma}. Therefore, the constant 4σΓ(σ)/y2σ|Γ(σ)|4^{\sigma}\Gamma(\sigma)/y^{2\sigma}|\Gamma(-\sigma)| appeared in the corollary above is sharp.

Now, using the estimate of PyσP_{y}^{\sigma} (Theorem 3.3) in Corollary 4.5 we get Theorem 1.1.

Proof of Theorem 1.1.

From Theorem 3.3 we have

Γ(σ)Γ(σ)Pyσ(x)Pyσ(x){y4σ(|x|2+y2)σ if |x|2+y21y4σ(|x|2+y2)2σ if |x|2+y2<1.\frac{\Gamma(\sigma)}{\Gamma(-\sigma)}\frac{P_{y}^{\sigma}(x)}{P_{y}^{-\sigma}(x)}\asymp\left\{\begin{array}[]{lll}\frac{y^{4\sigma}}{(|x|^{2}+y^{2})^{\sigma}}&\text{ if }|x|^{2}+y^{2}\geq 1\\ \frac{y^{4\sigma}}{(|x|^{2}+y^{2})^{2\sigma}}&\text{ if }|x|^{2}+y^{2}<1.\end{array}\right.

Therefore, from Corollary 4.5 we have

(Δ)σF,FCσy2σ({x:|x|2+y2<1}|F(x)|2(y2+|x|2)2σ𝑑x+{x:|x|2+y21}|F(x)|2(y2+|x|2)σ𝑑x).\left\langle(-\Delta)^{\sigma}F,F\right\rangle\geq C_{\sigma}\,y^{2\sigma}\left(\int_{\left\{x:|x|^{2}+y^{2}<1\right\}}\frac{|F(x)|^{2}}{(y^{2}+|x|^{2})^{2\sigma}}\,dx+\int_{\left\{x:|x|^{2}+y^{2}\geq 1\right\}}\frac{|F(x)|^{2}}{(y^{2}+|x|^{2})^{\sigma}}\,dx\right).

We now prove Hardy’s inequality corresponding to the homogeneous weight function (Theorem 1.3). To prove this theorem we need the following expression of the error term.

Theorem 4.7.

Let 0<σ<10<\sigma<1 and α>(2σ+n)/4\alpha>(2\sigma+n)/4. Then for FCc(X)F\in C_{c}^{\infty}(X) and G(x)=F(x)(P0α(x))1G(x)=F(x)~{}\left(P_{0}^{-\alpha}(x)\right)^{-1} we have

(4.6) (Δ)σF,FΓ(α)Γ(ασ)X|F(x)|2(P0σα(x)P0α(x))𝑑x\displaystyle\left\langle(-\Delta)^{\sigma}F,F\right\rangle-\frac{\Gamma(\alpha)}{\Gamma(\alpha-\sigma)}\int_{X}|F(x)|^{2}\left(\frac{P_{0}^{\sigma-\alpha}(x)}{P_{0}^{-\alpha}(x)}\right)\,dx
=\displaystyle= 12|Γ(σ)|XX|G(x)G(z)|2P0α(x)P0α(z)P0σ(z1x)𝑑x𝑑z,\displaystyle\frac{1}{2|\Gamma(-\sigma)|}\int_{X}\int_{X}|G(x)-G(z)|^{2}~{}P_{0}^{-\alpha}(x)P_{0}^{-\alpha}(z)~{}P_{0}^{\sigma}(z^{-1}x)~{}dx~{}dz,

where the function P0αP_{0}^{-\alpha} is defined by (3.7).

Proof.

Since α>n/4\alpha>n/4, we observe from Theorem 3.4 that P0αL2(X)P_{0}^{-\alpha}\in L^{2}(X). As before by Fubini theorem the spherical Fourier transform of P0αP_{0}^{-\alpha} is given by

P0α^(λ)=0et(|λ|2+|ρ|2)dtt1α=Γ(α)(|λ|2+|ρ|2)α,λ𝔞.\widehat{P_{0}^{-\alpha}}(\lambda)=\int_{0}^{\infty}e^{-t(|\lambda|^{2}+|\rho|^{2})}\frac{dt}{t^{1-\alpha}}=\Gamma(\alpha)~{}\left(|\lambda|^{2}+|\rho|^{2}\right)^{-\alpha},\>\>\lambda\in\mathfrak{a}^{\ast}.

Since α>(2σ+n)/4\alpha>(2\sigma+n)/4, it follows that P0αHσ(X)P_{0}^{-\alpha}\in H^{\sigma}(X). Indeed, using (2.2) we get that

𝔞|P0α^(λ)|2(|λ|2+|ρ|2)σ|𝐜(λ)|2𝑑λC+C{𝔞:|λ|1}(|λ|2+|ρ|2)2α+σ(1+|λ|)nl𝑑λ,\displaystyle\int_{\mathfrak{a}^{\ast}}|\widehat{P_{0}^{-\alpha}}(\lambda)|^{2}~{}(|\lambda|^{2}+|\rho|^{2})^{\sigma}~{}|{\bf c}(\lambda)|^{-2}~{}d\lambda\leq C+C^{\prime}\int_{\{\mathfrak{a}^{\ast}:|\lambda|\geq 1\}}(|\lambda|^{2}+|\rho|^{2})^{-2\alpha+\sigma}~{}(1+|\lambda|)^{n-l}~{}d\lambda,

which is finite. We recall from (4.4) that for f,gHσ(X)f,g\in H^{\sigma}(X)

(4.7) (Δ)σf,g=12|Γ(σ)|XX(f(z)f(x))(g(z)g(x))¯P0σ(z1x)𝑑z𝑑x.\left\langle(-\Delta)^{\sigma}f,g\right\rangle=\frac{1}{2|\Gamma(-\sigma)|}\int_{X}\int_{X}\left(f(z)-f(x)\right)~{}\overline{\left(g(z)-g(x)\right)}~{}P_{0}^{\sigma}(z^{-1}x)~{}dz~{}dx.

If we put g(x)=P0α(x)g(x)=P_{0}^{-\alpha}(x) and f(x)=|F(x)|2(P0α(x))1f(x)=|F(x)|^{2}(P_{0}^{-\alpha}(x))^{-1} in the equation above, then the left-hand side reduces to

(Δ)σf,g\displaystyle\left\langle(-\Delta)^{\sigma}f,g\right\rangle =\displaystyle= 𝔞(|λ|2+|ρ|2)σf^(λ)g^(λ)|𝐜(λ)|2𝑑λ\displaystyle\int_{\mathfrak{a}^{\ast}}\left(|\lambda|^{2}+|\rho|^{2}\right)^{\sigma}\widehat{f}(\lambda)\,\widehat{g}(\lambda)\,|{\bf c}(\lambda)|^{-2}\,d\lambda
=\displaystyle= Γ(α)𝔞(|λ|2+|ρ|2)σαf^(λ)|𝐜(λ)|2𝑑λ\displaystyle\Gamma(\alpha)\int_{\mathfrak{a}^{\ast}}\left(|\lambda|^{2}+|\rho|^{2}\right)^{\sigma-\alpha}\widehat{f}(\lambda)\,|{\bf c}(\lambda)|^{-2}\,d\lambda
=\displaystyle= Γ(α)Γ(ασ)𝔞P0σα^(λ)f^(λ)|𝐜(λ)|2𝑑λ\displaystyle\frac{\Gamma(\alpha)}{\Gamma(\alpha-\sigma)}\int_{\mathfrak{a}^{\ast}}\widehat{P_{0}^{\sigma-\alpha}}(\lambda)~{}\widehat{f}(\lambda)\,|{\bf c}(\lambda)|^{-2}\,d\lambda
=\displaystyle= Γ(α)Γ(ασ)X|F(x)|2P0σα(x)P0α(x)𝑑x.\displaystyle\frac{\Gamma(\alpha)}{\Gamma(\alpha-\sigma)}\int_{X}|F(x)|^{2}\frac{P_{0}^{\sigma-\alpha}(x)}{P_{0}^{-\alpha}(x)}\,dx.

The right-hand side of the equation (4.7) becomes (see (4))

(4.8) 12|Γ(σ)|XX(|F(x)F(z)|2|F(x)g(x)F(z)g(z)|2g(x)g(z))P0σ(z1x)𝑑z𝑑x.\frac{1}{2|\Gamma(-\sigma)|}\int_{X}\int_{X}\left(|F(x)-F(z)|^{2}-\left|\frac{F(x)}{g(x)}-\frac{F(z)}{g(z)}\right|^{2}~{}g(x)g(z)\right)~{}P_{0}^{\sigma}(z^{-1}x)~{}dz~{}dx.

Hence, equating both sides of the equation (4.7) we have

Γ(α)Γ(ασ)X|F(x)|2P0σα(x)P0α(x)𝑑x=12|Γ(σ)|XX|F(x)F(z)|2P0σ(z1x)𝑑z𝑑x\displaystyle\frac{\Gamma(\alpha)}{\Gamma(\alpha-\sigma)}\int_{X}|F(x)|^{2}\frac{P_{0}^{\sigma-\alpha}(x)}{P_{0}^{-\alpha}(x)}\,dx=\frac{1}{2|\Gamma(-\sigma)|}\int_{X}\int_{X}|F(x)-F(z)|^{2}~{}P_{0}^{\sigma}(z^{-1}x)~{}dz~{}dx
12|Γ(σ)|XX|F(x)g(x)F(z)g(z)|2g(x)g(z)P0σ(z1x)𝑑z𝑑x.\displaystyle-\frac{1}{2|\Gamma(-\sigma)|}\int_{X}\int_{X}\left|\frac{F(x)}{g(x)}-\frac{F(z)}{g(z)}\right|^{2}~{}g(x)g(z)~{}P_{0}^{\sigma}(z^{-1}x)~{}dz~{}dx.

By Lemma 4.3 the first term in the right-hand side of the above equation is equals to (Δ)σF,F\langle(-\Delta)^{\sigma}F,F\rangle and hence the required identity follows. ∎

Proof of Theorem 1.3.

Since σ<1\sigma<1 and n2n\geq 2, we can choose a positive α\alpha such that 2σ+n/4<α<n/22\sigma+n/4<\alpha<n/2. From Theorem 3.4 above it follows that

P0σα(x)P0α(x)\displaystyle\frac{P_{0}^{\sigma-\alpha}(x)}{P_{0}^{-\alpha}(x)} \displaystyle\asymp |x|2σ, for |x|<1;\displaystyle|x|^{-2\sigma},\>\>\textit{ for }|x|<1;
\displaystyle\asymp |x|σ, for |x|1.\displaystyle|x|^{-\sigma},\>\>\textit{ for }|x|\geq 1.

Therefore, it follows from Theorem 4.7 that

(Δ)σF,FCσ(|x|<1|F(x)|2|x|2σ𝑑x+|x|1|F(x)|2|x|σ𝑑x).\left\langle(-\Delta)^{\sigma}F,F\right\rangle\geq C_{\sigma}\left(\int_{|x|<1}\frac{|F(x)|^{2}}{|x|^{2\sigma}}\,dx+\int_{|x|\geq 1}\frac{|F(x)|^{2}}{|x|^{\sigma}}\,dx\right).

5. Mapping properties of Poisson Operator

In this section we prove Theorem 1.7. We start with the following lemma.

Lemma 5.1.

For 0<σ<10<\sigma<1 and 1<q<n+1n1<q<\frac{n+1}{n}, the function (x,y)Pyσ(x)Lq(X×+)(x,y)\mapsto P_{y}^{\sigma}(x)\in L^{q}(X\times\mathbb{R}_{+}).

Proof.

We first observe from (2.1) that for H𝔞H\in\mathfrak{a} with |H|<1|H|<1, the Jacobian J(expH)J(\exp H) corresponding to the polar decomposition is of order |H|nl|H|^{n-l}. From Theorem 3.3 it follows that

|x|2+y2<1|Pyσ(x)|q𝑑x𝑑y\displaystyle\int_{|x|^{2}+y^{2}<1}|P_{y}^{\sigma}(x)|^{q}~{}dx~{}dy \displaystyle\leq C|x|2+y2<1y2σq(|x|2+y2)nq/2σq𝑑x𝑑y\displaystyle C\int_{|x|^{2}+y^{2}<1}y^{2\sigma q}(|x|^{2}+y^{2})^{-nq/2-\sigma q}~{}dx~{}dy
\displaystyle\leq Cy=01{H𝔞+¯:|H|<1}y2σq(|H|2+y2)nq/2σq|H|nl𝑑H𝑑y\displaystyle C\int_{y=0}^{1}\int_{\{H\in\overline{\mathfrak{a^{+}}}:|H|<1\}}y^{2\sigma q}(|H|^{2}+y^{2})^{-nq/2-\sigma q}\,|H|^{n-l}~{}dH~{}dy
=\displaystyle= 0101y2σq(r2+y2)nq/2σqrnlrl1𝑑r𝑑y\displaystyle\int_{0}^{1}\int_{0}^{1}y^{2\sigma q}(r^{2}+y^{2})^{-nq/2-\sigma q}~{}r^{n-l}~{}r^{l-1}~{}dr~{}dy
\displaystyle\leq 01(0(1+s2)nq/2σqsn1𝑑s)ynnq𝑑y.\displaystyle\int_{0}^{1}\left(\int_{0}^{\infty}(1+s^{2})^{-nq/2-\sigma q}s^{n-1}~{}ds\right)~{}y^{n-nq}~{}dy.

We now use the following fact from [22, 3.251, (2); p.324]

(5.1) 0xμ1(1+x2)ν1𝑑x=12B(μ/2,(1νμ/2)), if μ>0, and (ν+μ/2)<1.\int_{0}^{\infty}x^{\mu-1}(1+x^{2})^{\nu-1}~{}dx=\frac{1}{2}B\left(\mu/2,(1-\nu-\mu/2)\right),\textmd{ if }\Re\mu>0,\textmd{ and }\Re(\nu+\mu/2)<1.

In our case, μ=n\mu=n and ν=nq/2σq+1\nu=-nq/2-\sigma q+1. Hence, ν+μ/2<1\nu+\mu/2<1 if and only if q>n/(n+2σ)q>n/(n+2\sigma). Therefore, if q>n/(n+2σ)q>n/(n+2\sigma) the above integral reduces to

12B(n/2,(nq/2+σqn/2))01ynnq𝑑y.\displaystyle\frac{1}{2}B\left(n/2,(nq/2+\sigma q-n/2)\right)\int_{0}^{1}y^{n-nq}~{}dy.

This is finite only if q<(1+n)/nq<(1+n)/n. Hence, for n/(n+2σ)<q<1+1nn/(n+2\sigma)<q<1+\frac{1}{n},

|x|2+y21|Pyσ(x)|q𝑑x𝑑y<.\int_{|x|^{2}+y^{2}\leq 1}|P_{y}^{\sigma}(x)|^{q}~{}dx~{}dy<\infty.

On the other hand for q>1q>1, using the estimate of Jacobian in (2.1) and the asymptotic behaviour of ϕ0\phi_{0} given in (2.6), it follows from Theorem 3.3 that

|x|2+y21|Pyσ(x)|q𝑑x𝑑y\displaystyle\int_{|x|^{2}+y^{2}\geq 1}|P_{y}^{\sigma}(x)|^{q}~{}dx~{}dy
|x2|+y21y2σq(4σΓ(σ))q(|x|2+y2)(l/2+|Σ0+|+σ+ 1/2)qe|ρ|q|x|2+y2|ϕ0(x)|q𝑑x𝑑y\displaystyle\leq\int_{|x^{2}|+y^{2}\geq 1}\frac{y^{2\sigma q}}{\left(4^{\sigma}\Gamma(\sigma)\right)^{q}}\left(\sqrt{|x|^{2}+y^{2}}\right)^{-(l/2+\,|\Sigma_{0}^{+}|+\,\sigma+\,1/2)q}~{}e^{-|\rho|q\sqrt{|x|^{2}+y^{2}}}~{}|\phi_{0}(x)|^{q}~{}dx~{}dy
C|x|2+y21y2σqe|ρ|(q+1)2|x|2+y2e|ρ|(q1)2|x|2+y2|ϕ0(x)|q𝑑x𝑑y\displaystyle\leq C\int_{|x|^{2}+y^{2}\geq 1}y^{2\sigma q}~{}e^{-\frac{|\rho|(q+1)}{2}\sqrt{|x|^{2}+y^{2}}}~{}e^{-\frac{|\rho|(q-1)}{2}\sqrt{|x|^{2}+y^{2}}}~{}|\phi_{0}(x)|^{q}~{}dx~{}dy
C{(H,y)𝔞+¯×(0,):|H|2+y21}y2σqe|ρ|(q1)|y|2e|ρ|(q+1)|H|2|H||Σ0+|qeqρ(H)e2ρ(H)𝑑H𝑑y\displaystyle\leq C\int_{\left\{(H,y)\in\overline{\mathfrak{a}^{+}}\times(0,\infty):|H|^{2}+y^{2}\geq 1\right\}}y^{2\sigma q}~{}e^{-\frac{|\rho|(q-1)|y|}{2}}~{}e^{-\frac{|\rho|(q+1)|H|}{2}}~{}|H|^{|\Sigma_{0}^{+}|q}\,e^{-q\,\rho(H)}~{}e^{2\rho(H)}~{}dH~{}dy
(0y2σqe|ρ|(q1)|y|/2𝑑y)(𝔞+¯|H||Σ0+|qe32(q1)ρ(H)𝑑H)<.\displaystyle\leq\left(\int_{0}^{\infty}y^{2\sigma q}~{}e^{-|\rho|(q-1)|y|/2}~{}dy\right)\left(\int_{\overline{\mathfrak{a^{+}}}}|H|^{|\Sigma_{0}^{+}|q}~{}e^{-\frac{3}{2}(q-1)\rho(H)}~{}dH\right)<\infty.

This completes the proof.

We are now in a position to prove Theorem 1.7. We follow similar ideas which are used to the proof of [34, Theorem B].

Proof of Theorem 1.7.

We first prove (1)(1). Let uu be the solution of (1.5) with boundary value fHσ(X)f\in H^{\sigma}(X), and let

𝒰(λ,k,η)=(u~(λ,k))(η), for λ𝔞,kK,η+\mathcal{U}(\lambda,k,\eta)=\mathcal{F}\left(\widetilde{u}(\lambda,k)\right)(\eta),\>\textit{ for }\lambda\in\mathfrak{a}^{\ast},k\in K,\eta\in{\mathbb{R}}_{+}

be the composition of the Helgason and the Euclidean Fourier transform on X×X\times\mathbb{R}. Multiplying y2y^{2} on both sides of the equation (1.5) and taking the composition of Helgason and Euclidean Fourier transform on X×X\times\mathbb{R} it follows that

2η2((|λ|2+|ρ|2+η2)𝒰(λ,k,η))(12σ)η(η𝒰(λ,k,η))=0\frac{\partial^{2}}{\partial\eta^{2}}\left((|\lambda|^{2}+|\rho|^{2}+\eta^{2})~{}\mathcal{U}(\lambda,k,\eta)\right)-(1-2\sigma)\frac{\partial}{\partial\eta}\left(\eta~{}\mathcal{U}(\lambda,k,\eta)\right)=0

which is equivalent to

(5.2) {(|λ|2+|ρ|2+η2)22η+(3+2σ)ηη+(1+2σ)}𝒰(λ,k,η)=0.\left\{(|\lambda|^{2}+|\rho|^{2}+\eta^{2})\frac{\partial^{2}}{\partial^{2}\eta}+(3+2\sigma)\eta\frac{\partial}{\partial\eta}+(1+2\sigma)\right\}\mathcal{U}(\lambda,k,\eta)=0.

Let t=η|λ|2+|ρ|2t=\frac{\eta}{\sqrt{|\lambda|^{2}+|\rho|^{2}}} and we define

v(λ,k,t)=𝒰(λ,k,η).v(\lambda,k,t)=\mathcal{U}(\lambda,k,\eta).

Then equation (5.2) reduces to

𝒟σ,tv(λ,k,t):={(1+t2)d2dt2+(2σ+3)tddt+(2σ+1)}v(λ,k,t)=0.\mathcal{D}_{\sigma,t}v(\lambda,k,t):=\left\{(1+t^{2})\frac{d^{2}}{dt^{2}}+(2\sigma+3)t\frac{d}{dt}+(2\sigma+1)\right\}v(\lambda,k,t)=0.

Since f(x)=u(x,0)f(x)=u(x,0) for xXx\in X, by the Euclidean Fourier inversion formula we have

f~(λ,k)=u(,0)~(λ,k)=12π𝒰(λ,k,η)𝑑η=|λ|2+|ρ|22πv(λ,k,t)𝑑t.\widetilde{f}(\lambda,k)=u(\cdot,0)^{\widetilde{}}(\lambda,k)=\frac{1}{\sqrt{2\pi}}\int_{\mathbb{R}}\mathcal{U}(\lambda,k,\eta)~{}d\eta=\frac{\sqrt{|\lambda|^{2}+|\rho|^{2}}}{\sqrt{2\pi}}\int_{\mathbb{R}}v(\lambda,k,t)~{}dt.

Therefore, the function vv satisfies

𝒟σ,tv(λ,k,t)=0, and v(λ,k,t)𝑑t=2π|λ|2+|ρ|2f~(λ,k),\mathcal{D}_{\sigma,t}v(\lambda,k,t)=0,\text{ and }\int_{\mathbb{R}}v(\lambda,k,t)\,dt=\frac{\sqrt{2\pi}}{\sqrt{|\lambda|^{2}+|\rho|^{2}}}\widetilde{f}(\lambda,k),

for almost every (λ,k)𝔞×K(\lambda,k)\in\mathfrak{a}^{\ast}\times K. Hence, the function vv is given by

(5.3) v(λ,k,t)=2π|λ|2+|ρ|2f~(λ,k)ψ(t),v(\lambda,k,t)=\frac{\sqrt{2\pi}}{\sqrt{|\lambda|^{2}+|\rho|^{2}}}\widetilde{f}(\lambda,k)~{}\psi(t),

where ψ\psi satisfies

(5.4) 𝒟σ,tψ=0,andψ(t)𝑑t=1.\mathcal{D}_{\sigma,t}\psi=0,\>\>\textmd{and}\>\>\int_{\mathbb{R}}\psi(t)~{}dt=1.

The equation Dσ,tψ=0D_{\sigma,t}\psi=0 has a fundamental system of solutions spanned by

ψ1(t)=2F1(12,σ+12;12;t2)=(1+t2)σ1/2,\displaystyle\psi_{1}(t)=\,_{2}F_{1}\left(\frac{1}{2},\sigma+\frac{1}{2};\frac{1}{2};-t^{2}\right)=(1+t^{2})^{-\sigma-1/2},
ψ2(t)=t2F1(1,σ+1;32;t2).\displaystyle\psi_{2}(t)=t\,_{2}F_{1}\left(1,\sigma+1;\frac{3}{2};-t^{2}\right).

Using (5.3) it is now easy to check that

(5.5) 𝔞×K×|𝒰(λ,k,η)|2(|λ|2+|ρ|2+η2)σ+12|𝐜(λ)|2𝑑λ𝑑k𝑑η\displaystyle\int_{\mathfrak{a}^{\ast}\times K\times\mathbb{R}}|\mathcal{U}(\lambda,k,\eta)|^{2}(|\lambda|^{2}+|\rho|^{2}+\eta^{2})^{\sigma+\frac{1}{2}}~{}|{\bf c}(\lambda)|^{-2}~{}d\lambda~{}dk~{}d\eta
=\displaystyle= 𝔞×K×|v(λ,k,t)|2(|λ|2+|ρ|2)σ+1(1+t2)σ+12|𝐜(λ)|2𝑑λ𝑑k𝑑t\displaystyle\int_{\mathfrak{a}^{\ast}\times K\times\mathbb{R}}|v(\lambda,k,t)|^{2}~{}(|\lambda|^{2}+|\rho|^{2})^{\sigma+1}~{}\left(1+t^{2}\right)^{\sigma+\frac{1}{2}}~{}|{\bf c}(\lambda)|^{-2}~{}d\lambda~{}dk~{}dt
=\displaystyle= 2π𝔞×K|f^(λ,k)|2(|λ|2+|ρ2|)σ|𝐜(λ)|2𝑑λ𝑑k|ψ(t)|2(1+t2)σ+12𝑑t.\displaystyle 2\pi\int_{\mathfrak{a}^{\ast}\times K}|\widehat{f}(\lambda,k)|^{2}~{}(|\lambda|^{2}+|\rho^{2}|)^{\sigma}~{}|{\bf c}(\lambda)|^{-2}~{}d\lambda~{}dk\int_{\mathbb{R}}|\psi(t)|^{2}~{}(1+t^{2})^{\sigma+\frac{1}{2}}~{}dt.

Since fHσf\in H^{\sigma}, it follows that uHσ+12u\in H^{\sigma+\frac{1}{2}} if and only if ψL2(,(1+t2)σ+12dt)\psi\in L^{2}(\mathbb{R},(1+t^{2})^{\sigma+\frac{1}{2}}\,dt). It is easy to check from the asymptotic properties of hypergeometric function that ψ2L2(,(1+t2)σ+12dt)\psi_{2}\notin L^{2}(\mathbb{R},(1+t^{2})^{\sigma+\frac{1}{2}}\,dt) (see [1, Theorem 2.3.2]). Hence, we choose ψ(t)\psi(t) to be a constant multiple of ψ1(t)=(1+t2)σ12\psi_{1}(t)=(1+t^{2})^{-\sigma-\frac{1}{2}}. From (5.1) we get that ψ1L1()=πΓ(σ)/Γ(σ+12)\|\psi_{1}\|_{L^{1}(\mathbb{R})}=\sqrt{\pi}\,\Gamma(\sigma)/\Gamma(\sigma+\frac{1}{2}). Hence, using (5.4) it follows that

ψ(t)=Γ(σ+1/2)πΓ(σ)ψ1(t).\psi(t)=\frac{\Gamma(\sigma+1/2)}{\sqrt{\pi}\Gamma(\sigma)}\psi_{1}(t).

We now observe that

|ψ(t)|2(1+t2)σ+12𝑑z=Γ(σ+1/2)πΓ(σ),\int_{\mathbb{R}}|\psi(t)|^{2}~{}(1+t^{2})^{\sigma+\frac{1}{2}}~{}dz=\frac{\Gamma(\sigma+1/2)}{\sqrt{\pi}\Gamma(\sigma)},

and hence from (5.5),

uHσ+12(X×+)2=2πΓ(σ+12)Γ(σ)fHσ(X).\displaystyle\|u\|_{H^{\sigma+\frac{1}{2}}(X\times\mathbb{R}_{+})}^{2}=\frac{2\sqrt{\pi}\Gamma(\sigma+\frac{1}{2})}{\Gamma(\sigma)}\|f\|_{H^{\sigma}(X)}.

This completes the proof of part (1). We now prove part (2). We first observe that

TσfLq(X×+)q=0fPyσLq(X)q𝑑y.\|T_{\sigma}f\|_{L^{q}(X\times\mathbb{R}_{+})}^{q}=\int_{0}^{\infty}\|f\ast P_{y}^{\sigma}\|_{L^{q}(X)}^{q}~{}dy.

Also, from Theorem 3.3 it follows that for each y>0y>0 the function PyσLq(X)P_{y}^{\sigma}\in L^{q}(X), for all q>1q>1. Therefore, by Kunze-Stein phenomenon (Remark 1.8), for 1p<q21\leq p<q\leq 2

fPyσLq(X)CfLp(X)PyσLq(X).\|f\ast P_{y}^{\sigma}\|_{L^{q}(X)}\leq C\|f\|_{L^{p}(X)}\|P_{y}^{\sigma}\|_{L^{q}(X)}.

Therefore, by Lemma 5.1 it follows that

(5.6) Tσ:Lp(X)Lq(X×+),T_{\sigma}:L^{p}(X)\rightarrow L^{q}(X\times\mathbb{R}_{+}),

is a bounded map, for 1p<q<(n+1)/n1\leq p<q<(n+1)/n. We also observe that

(5.7) Tσ:L(X)L(X×+),T_{\sigma}:L^{\infty}(X)\rightarrow L^{\infty}(X\times\mathbb{R}_{+}),

is a bounded map, as the integral XPyσ(x)𝑑x=1\int_{X}P_{y}^{\sigma}(x)\,dx=1 for all y>0y>0. By Riesz Thorin interpolation theorem it now follows from (5.6) and (5.7) that

(5.8) Tσ:Lp(X)Lq(X×+),T_{\sigma}:L^{p}(X)\rightarrow L^{q}(X\times\mathbb{R}_{+}),

is bounded for 1p<1\leq p<\infty and p<q<(n+1n)pp<q<(\frac{n+1}{n})p. We now prove that

TσfLq(X×+)CfLp(X),\|T_{\sigma}f\|_{L^{q}(X\times\mathbb{R}_{+})}\leq C\|f\|_{L^{p}(X)},

for p>1p>1 and q=(n+1n)pq=(\frac{n+1}{n})p. By (5.7) and Marcinkiewicz interpolation theorem it is enough to show that

Tσ:L1(X)L(1+n)/n,(X×+).T_{\sigma}:L^{1}(X)\rightarrow L^{(1+n)/n,\infty}(X\times\mathbb{R}_{+}).

Using Theorem 3.3 and the boundedness of the function ϕ0\phi_{0} we get that

|Tσf(x,y)|X|f(z)|Pyσ(z1x)|f(z)|𝑑zCynfL1(X)\displaystyle|T_{\sigma}f(x,y)|\leq\int_{X}|f(z)|P_{y}^{\sigma}(z^{-1}x)~{}|f(z)|~{}dz\leq Cy^{-n}~{}\|f\|_{L^{1}(X)}
+Cy2σ|z1x|2+y21(|z1x|2+y2)(l/21/2σ|Σ0+|)e|ρ|(|z1x|2+y2)||f(z)|𝑑z\displaystyle+Cy^{2\sigma}\int_{|z^{-1}x|^{2}+y^{2}\geq 1}\sqrt{(|z^{-1}x|^{2}+y^{2})}^{(-l/2-1/2-\sigma-|\Sigma_{0}^{+}|)}~{}e^{-|\rho|\sqrt{(|z^{-1}x|^{2}+y^{2})}|}|f(z)|~{}dz
CynfL1(X)+CynfL1(X)supy+(y2σ+ne|ρ|y)\displaystyle\leq Cy^{-n}~{}\|f\|_{L^{1}(X)}+Cy^{-n}~{}\|f\|_{L^{1}(X)}~{}\sup{y\in\mathbb{R}_{+}}\left(y^{2\sigma+n}e^{-|\rho|y}\right)
CynfL1(X).\displaystyle\leq Cy^{-n}~{}\|f\|_{L^{1}(X)}.

Hence, |Taf(x,y)|>λ|T_{a}f(x,y)|>\lambda implies that y(CfL1(X)λ)1n=by\leq\left(\frac{C\|f\|_{L^{1}(X)}}{\lambda}\right)^{\frac{1}{n}}=b (say). Then Chebyshev’s inequality yields

m({(x,y)X×+:|Taf(x,y)|>λ})\displaystyle m\left(\{(x,y)\in X\times\mathbb{R}_{+}:|T_{a}f(x,y)|>\lambda\}\right)
=\displaystyle= m({(x,y)X×+:y<b,|Taf(x,y)|>λ})\displaystyle m\left(\{(x,y)\in X\times\mathbb{R}_{+}:y<b,|T_{a}f(x,y)|>\lambda\}\right)
\displaystyle\leq 1λ{(x,y)X×+:y<b}|Taf(x,y)|𝑑x𝑑y\displaystyle\frac{1}{\lambda}\int_{\{(x,y)\in X\times\mathbb{R}_{+}:y<b\}}|T_{a}f(x,y)|~{}dx~{}dy
\displaystyle\leq CaλX|f(z)|{(x,y)X×+:y<b}Pyσ(z1x)𝑑x𝑑y𝑑z\displaystyle\frac{C_{a}}{\lambda}\int_{X}|f(z)|\int_{\{(x,y)\in X\times\mathbb{R}_{+}:y<b\}}P_{y}^{\sigma}(z^{-1}x)~{}dx~{}dy~{}dz
\displaystyle\leq CσλfL1(X)b=Cσ(fL1(X)λ)1+1n.\displaystyle\frac{C_{\sigma}}{\lambda}\|f\|_{L^{1}(X)}\,b=C_{\sigma}\left(\frac{\|f\|_{L^{1}(X)}}{\lambda}\right)^{1+\frac{1}{n}}.

The last inequality follows because of the fact that XPyσ(x)𝑑x=1\int_{X}P_{y}^{\sigma}(x)\,dx=1 for all y>0y>0. This completes the proof. ∎

6. Expression of the kernel PyσP_{y}^{\sigma}

In the case of n\mathbb{R}^{n} and of the Heisenberg groups the function PyσP_{y}^{\sigma} is the classical Poisson kernel. In the case of symmetric spaces, we only have the integral expression as in Theorem 1.5 and the both-sides estimates (Theorem 3.3) for PyσP_{y}^{\sigma}. In this section we write the precise expression of PyσP_{y}^{\sigma} for complex and rank one symmetric spaces using the expression of the heat kernel.

6.1. GG is complex

In this case we have the following formula for the heat kernel [5]

ht(expH)=(4πt)n/2e|ρ|2t(αΣ+α(H)sinhα(H))eH2/4t,t>0,H𝔞.h_{t}(\exp H)=(4\pi t)^{-n/2}~{}e^{-|\rho|^{2}t}~{}\left(\prod_{\alpha\in\Sigma^{+}}\frac{\alpha(H)}{\sinh\alpha(H)}\right)~{}e^{-H^{2}/4t},\>\>t>0,~{}H\in\mathfrak{a}.

It now follows from the definition (1.9) of PyσP_{y}^{\sigma}

Pyσ(expH)=y2σ4σΓ(σ)(4πt)n/2(αΣ+α(H)sinhα(H))0tn/2σ1e|ρ|2te(|H|2+y2)/4tdtt1+σ\displaystyle P_{y}^{\sigma}(\exp H)=\frac{y^{2\sigma}}{4^{\sigma}\Gamma(\sigma)}(4\pi t)^{-n/2}~{}\left(\prod_{\alpha\in\Sigma^{+}}\frac{\alpha(H)}{\sinh\alpha(H)}\right)~{}\int_{0}^{\infty}t^{-n/2-\sigma-1}~{}e^{-|\rho|^{2}t}~{}e^{-\left(|H|^{2}+y^{2}\right)/4t}~{}\frac{dt}{t^{1+\sigma}}
=\displaystyle= y2σΓ(σ)21n/2σπn/2(αΣ+α(H)sinhα(H))(|H|2+y2|ρ|)(n+2σ)/2Kn/2σ(|H|2+y2|ρ|).\displaystyle\frac{y^{2\sigma}}{\Gamma(\sigma)}~{}2^{1-n/2-\sigma}~{}\pi^{-n/2}~{}\left(\prod_{\alpha\in\Sigma^{+}}\frac{\alpha(H)}{\sinh\alpha(H)}\right)~{}\left(\frac{\sqrt{|H|^{2}+y^{2}}}{|\rho|}\right)^{-(n+2\sigma)/2}~{}K_{-n/2-\sigma}(\sqrt{|H|^{2}+y^{2}}|\rho|).

Here the last equality follows from the formula [22, 3.471(9), p. 368], and Kn/2σK_{-n/2-\sigma} is the modified Bessel function (defined in [22, 8.407 (1), p. 911]).

6.2. XX is of rank one

Let F=,,HF=\mathbb{R},\mathbb{C},H, or OO be the real numbers, the complex numbers, the quaternions or the Cayley octonions respectively. The rank one symmetric spaces can be realized as the hyperbolic space n(F)\mathbb{H}^{n}(F). Here the subscript nn denotes the dimension over the base field FF. Using the expression of the heat kernel [5, 21] we have the following results.

  1. (1)

    X=n()X=\mathbb{H}^{n}(\mathbb{R}), and n3n\geq 3 odd. Using the formula [22, 3.471(9), p. 368] we get

    Pyσ(x)\displaystyle P_{y}^{\sigma}(x) =\displaystyle= c0t1/2eρ2tey2/4t(1sinhxx)(n1)/2e|x|2/4tdtt1+σ\displaystyle c\int_{0}^{\infty}t^{-1/2}~{}e^{-\rho^{2}t}~{}e^{-y^{2}/4t}\left(-\frac{1}{\sinh x}~{}\frac{\partial}{\partial x}\right)^{(n-1)/2}~{}e^{-|x|^{2}/4t}~{}\frac{dt}{t^{1+\sigma}}
    =\displaystyle= c(1sinhxx)(n1)/20t3/2σeρ2te(|x|2+y2)/4t𝑑t\displaystyle c\left(-\frac{1}{\sinh x}~{}\frac{\partial}{\partial x}\right)^{(n-1)/2}\int_{0}^{\infty}t^{-3/2-\sigma}~{}e^{-\rho^{2}t}~{}e^{-(|x|^{2}+y^{2})/4t}~{}dt
    =\displaystyle= c(1sinhxx)(n1)/2(|x|2+y2ρ)σ1/2Kσ1/2(ρ|x|2+y2).\displaystyle c\left(-\frac{1}{\sinh x}~{}\frac{\partial}{\partial x}\right)^{(n-1)/2}\left(\frac{\sqrt{|x|^{2}+y^{2}}}{\rho}\right)^{-\sigma-1/2}~{}K_{-\sigma-1/2}(\rho\sqrt{|x|^{2}+y^{2}}).
  2. (2)

    X=n()X=\mathbb{H}^{n}(\mathbb{R}), and n2n\geq 2 even. Using the formula [22, 3.471(9), p. 368] we get

    Pyσ(x)\displaystyle P_{y}^{\sigma}(x)
    =\displaystyle= c0t1/2eρ2tey2/4txsinhzcosh2zcosh2x(1sinhzz)n/2e|z|2/4t𝑑zdtt1+σ\displaystyle c\int_{0}^{\infty}t^{-1/2}~{}e^{-\rho^{2}t}~{}e^{-y^{2}/4t}~{}\int_{x}^{\infty}\frac{\sinh z}{\sqrt{\cosh^{2}z-\cosh^{2}x}}\left(-\frac{1}{\sinh z}~{}\frac{\partial}{\partial z}\right)^{n/2}~{}e^{-|z|^{2}/4t}~{}dz~{}\frac{dt}{t^{1+\sigma}}
    =\displaystyle= cxsinhzcosh2zcosh2x(1sinhzz)n/20t3/2σeρ2te(|z|2+y2)/4t𝑑t𝑑z\displaystyle c\int_{x}^{\infty}\frac{\sinh z}{\sqrt{\cosh^{2}z-\cosh^{2}x}}\left(-\frac{1}{\sinh z}~{}\frac{\partial}{\partial z}\right)^{n/2}~{}\int_{0}^{\infty}t^{-3/2-\sigma}~{}e^{-\rho^{2}t}~{}e^{-\left(|z|^{2}+y^{2}\right)/4t}~{}dt~{}dz
    =\displaystyle= cxsinhzcosh2zcosh2x(1sinhzz)n/2(|z|2+y2ρ)σ1/2Kσ1/2(ρ|z|2+y2)𝑑z.\displaystyle c\int_{x}^{\infty}\frac{\sinh z}{\sqrt{\cosh^{2}z-\cosh^{2}x}}\left(-\frac{1}{\sinh z}~{}\frac{\partial}{\partial z}\right)^{n/2}\left(\frac{\sqrt{|z|^{2}+y^{2}}}{\rho}\right)^{-\sigma-1/2}~{}K_{-\sigma-1/2}(\rho\sqrt{|z|^{2}+y^{2}})~{}dz.
  3. (3)

    X=n(F)X=\mathbb{H}^{n}(F) where F=,HF=\mathbb{C},H or OO. Then there exist constants c1,c2,,cn/2c_{1},c_{2},\cdots,c_{n/2} such that

    Pyσ(x)\displaystyle P_{y}^{\sigma}(x) =\displaystyle= 0t1/2eρ2tj=1n/2cjxsinhzcosh2zsinh2x(coshz)j+1d(12πsinhzz)j+mα/2\displaystyle\int_{0}^{\infty}t^{-1/2}~{}e^{-\rho^{2}t}\sum_{j=1}^{n/2}c_{j}\int_{x}^{\infty}\frac{\sinh z}{\sqrt{\cosh^{2}z-\sinh^{2}x}}~{}(\cosh z)^{j+1-d}~{}\left(-\frac{1}{2\pi\sinh z}~{}\frac{\partial}{\partial z}\right)^{j+m_{\alpha}/2}
    e|z|2/4tdzdtt1+σ\displaystyle e^{-|z|^{2}/4t}~{}dz~{}\frac{dt}{t^{1+\sigma}}
    =\displaystyle= cσj=1d/2cjxsinhzcosh2zsinh2x(coshz)j+1dρ1+2σ(12πsinhzz)j+mα/2\displaystyle c_{\sigma}\sum_{j=1}^{d/2}c_{j}\int_{x}^{\infty}\frac{\sinh z}{\sqrt{\cosh^{2}z-\sinh^{2}x}}~{}(\cosh z)^{j+1-d}~{}\rho^{1+2\sigma}~{}\left(-\frac{1}{2\pi\sinh z}~{}\frac{\partial}{\partial z}\right)^{j+m_{\alpha}/2}
    2(2|z|2+y2ρ)σ+1/2Kσ1/2(ρ|z|2+y2)dz,\displaystyle 2\left(\frac{2}{\sqrt{|z|^{2}+y^{2}}\rho}\right)^{\sigma+1/2}~{}K_{-\sigma-1/2}(\rho\sqrt{|z|^{2}+y^{2}})~{}dz,

    where the constant cσc_{\sigma} depends only on σ\sigma.

7. Poincaré-Sobolev inequality

In this section we prove Theorem 1.11. For the convenience of the reader we restate the theorem here.

Theorem 7.1.

Let dimX=n3\dim X=n\geq 3 and 0<σ<min{l+2|Σ0+|,n}0<\sigma<\min\{l+2|\Sigma_{0}^{+}|,n\}. Then for 2<p2nnσ2<p\leq\frac{2n}{n-\sigma} there exists S=Sn,σ,p>0S=S_{n,\sigma,p}>0 such that for all fHσ2(X)f\in H^{\frac{\sigma}{2}}(X)

(7.1) (Δ|ρ|2)σ/4fL2(X)2SfLp(X)2.\|(-\Delta-|\rho|^{2})^{\sigma/4}f\|_{L^{2}(X)}^{2}\geq S\|f\|_{L^{p}(X)}^{2}.
Proof.

We first observe that it is enough to prove the result for fCc(X)f\in C_{c}^{\infty}(X). It also suffices to show that

(7.2) Xf(x)(Δ|ρ|2)σ/2f(x)𝑑xCfLp(X)2.\int_{X}f(x)~{}(-\Delta-|\rho|^{2})^{-\sigma/2}f(x)~{}dx\leq C\|f\|_{L^{p^{\prime}}(X)}^{2}.

Indeed, if (7.2) holds, then by Hölder’s inequality

|f,g|\displaystyle\left|\left\langle f,g\right\rangle\right| =\displaystyle= |(Δ|ρ|2)σ/4f,(Δ|ρ|2)σ/4g|\displaystyle\left|\left\langle\left(-\Delta-|\rho|^{2}\right)^{\sigma/4}f,\left(-\Delta-|\rho|^{2}\right)^{-\sigma/4}g\right\rangle\right|
\displaystyle\leq (Δ|ρ|2)σ/4fL2(X)(Δ|ρ|2)σ/4gL2(X)\displaystyle\left\|\left(-\Delta-|\rho|^{2}\right)^{\sigma/4}f\right\|_{L^{2}(X)}\,\,\left\|\left(-\Delta-|\rho|^{2}\right)^{-\sigma/4}g\right\|_{L^{2}(X)}
=\displaystyle= (Δ|ρ|2)σ/2f,f1/2(Δ|ρ|2)σ/2g,g1/2\displaystyle\left\langle\left(-\Delta-|\rho|^{2}\right)^{\sigma/2}f,f\right\rangle^{1/2}\left\langle\left(-\Delta-|\rho|^{2}\right)^{-\sigma/2}g,g\right\rangle^{1/2}
\displaystyle\leq C12(Δ|ρ|2)σ/2f,f12gLp(X),\displaystyle C^{\frac{1}{2}}\left\langle\left(-\Delta-|\rho|^{2}\right)^{\sigma/2}f,f\right\rangle^{\frac{1}{2}}\|g\|_{L^{p^{\prime}}(X)},

and hence

fLp(X)Cn12(Δ|ρ|2)σ/2f,f12.\|f\|_{L^{p}(X)}\leq C_{n}^{\frac{1}{2}}\left\langle\left(-\Delta-|\rho|^{2}\right)^{\sigma/2}f,f\right\rangle^{\frac{1}{2}}.

We now prove (7.2). Let kσk_{\sigma} be the Schwartz kernel for the operator (Δ|ρ|2)σ/2(-\Delta-|\rho|^{2})^{-\sigma/2}. We have the following well-known estimates due to Anker and Ji [4, Theorem 4.2.2], for 0<σ<l+2|Σ0+|0<\sigma<l+2|\Sigma_{0}^{+}|

kσ(x)\displaystyle k_{\sigma}(x) |x|σl2|Σ0+|ϕ0(x),|x|1,\displaystyle\asymp|x|^{\sigma-l-2|\Sigma_{0}^{+}|}~{}\phi_{0}(x),\>\>|x|\geq 1,
|x|σn,|x|<1.\displaystyle\asymp|x|^{\sigma-n},\>\>|x|<1.

To prove (7.2), it is enough to show that

fkσLp(X)CfLp(X).\|f\ast k_{\sigma}\|_{L^{p}(X)}\leq C\|f\|_{L^{p^{\prime}}(X)}.

Let χ\chi be the characteristic function of the unit ball 𝐁(o,1){\bf B}(o,1) and kσ0(x)=χ(x)kσ(x)k_{\sigma}^{0}(x)=\chi(x)~{}k_{\sigma}(x) and kσ=kk0k_{\sigma}^{\infty}=k-k_{0}. Now, by Young’s inequality we have that

fkσ0Lp(X)CfLp(X)kσ0Lp/2(X),\|f\ast k_{\sigma}^{0}\|_{L^{p}(X)}\leq C\|f\|_{L^{p^{\prime}}(X)}~{}\|k_{\sigma}^{0}\|_{L^{p/2}(X)},

and

kσ0Lp/2(X)p201|t|(σn)p/2|t||Σ+|tl1𝑑t.\displaystyle\|k_{\sigma}^{0}\|_{L^{p/2}(X)}^{\frac{p}{2}}\asymp\int_{0}^{1}|t|^{(\sigma-n)p/2}~{}|t|^{|\Sigma^{+}|}~{}t^{l-1}~{}dt.

The right-hand side is finite if p<2nnσp<\frac{2n}{n-\sigma}. Using the fact that for r<1r<1, the volume of the ball B(o,r)B(o,r) in XX is of order rnr^{n}, it is easy to check that kσ0Lnnσ,(X)k_{\sigma}^{0}\in L^{\frac{n}{n-\sigma},\infty}(X). By Young’s inequality for weak type spaces [23, Theorem 1.4.24. page 63] it follows that

fkσ0L2nnσ(X)CfL2nn+σ(X).\|f\ast k_{\sigma}^{0}\|_{L^{\frac{2n}{n-\sigma}}(X)}\leq C\|f\|_{L^{\frac{2n}{n+\sigma}}(X)}.

Therefore, we have for all p2nnσp\leq\frac{2n}{n-\sigma},

(7.4) fkσ0Lp(X)CfLp(X).\|f\ast k_{\sigma}^{0}\|_{L^{p}(X)}\leq C\|f\|_{L^{p^{\prime}}(X)}.

Next, we shall show that for p>2p>2,

fkσLp(X)CpfLp(X).\|f\ast k_{\sigma}^{\infty}\|_{L^{p}(X)}\leq C_{p}\|f\|_{L^{p^{\prime}}(X)}.

To prove this we shall use complex interpolation theorem and the idea of [35, Theorem 4.1]. For z12\Re z\geq-\frac{1}{2}, we define an analytic family of linear operators TzT_{z} from (X,dx)(X,dx) to itself as follows:

Tzf=f(kσ)1+z.T_{z}f=f\ast(k_{\sigma}^{\infty})^{1+z}.

For z=12+iyz=-\frac{1}{2}+iy, we have

TzfL(X)\displaystyle\|T_{z}f\|_{L^{\infty}(X)} =\displaystyle= f(kσ)12+iyL(X)\displaystyle\|f\ast(k_{\sigma}^{\infty})^{\frac{1}{2}+iy}\|_{L^{\infty}(X)}
\displaystyle\leq Csup{xX:|x|1}φ0(x)12|x|(σl)/2|Σ0+|fL1(X)\displaystyle C\,\sup_{\{x\in X:|x|\geq 1\}}\varphi_{0}(x)^{\frac{1}{2}}|x|^{(\sigma-l)/2-|\Sigma_{0}^{+}|}\,\|f\|_{L^{1}(X)}
\displaystyle\leq CfL1(X).\displaystyle C\|f\|_{L^{1}(X)}.

For z=ϵ+iy,ϵ>0z=\epsilon+iy,\epsilon>0, we have

TzfL2(X)2\displaystyle\|T_{z}f\|_{L^{2}(X)}^{2} =\displaystyle= K|f~(λ,k)|2|(kσ)1+ϵ+iy^(λ)|2|𝐜(λ)|2𝑑λ𝑑k\displaystyle\int_{\mathbb{R}}\int_{K}\left|\widetilde{f}(\lambda,k)\right|^{2}\left|\widehat{(k_{\sigma}^{\infty})^{1+\epsilon+iy}}(\lambda)\right|^{2}|{\bf c}(\lambda)|^{-2}\,d\lambda~{}dk
\displaystyle\leq sup|(kσ)1+ϵ+iy^(λ)|2fL2(X)2.\displaystyle\sup\left|\widehat{(k_{\sigma}^{\infty})^{1+\epsilon+iy}}(\lambda)\right|^{2}\|f\|_{L^{2}(X)}^{2}.

Now, by Theorem 2.1 it follows that for λ𝔞\lambda\in\mathfrak{a}^{\ast} and ϵ>0\epsilon>0

|(kσ)1+ϵ+iy^(λ)|\displaystyle|\widehat{(k_{\sigma}^{\infty})^{1+\epsilon+iy}}(\lambda)| \displaystyle\leq |{xX:|x|1}|x|(σl2|Σ0+|)(1+ϵ+iy)(ϕ0(x))1+ϵ+iyϕλ(x)𝑑x|\displaystyle\left|\int_{\{x\in X:|x|\geq 1\}}|x|^{(\sigma-l-2|\Sigma_{0}^{+}|)(1+\epsilon+iy)}~{}(\phi_{0}(x))^{1+\epsilon+iy}~{}\phi_{-\lambda}(x)~{}dx\right|
\displaystyle\leq {xX:|x|1}ϕ0(x)2+ϵ𝑑x<,\displaystyle\int_{\{x\in X:|x|\geq 1\}}\phi_{0}(x)^{2+\epsilon}~{}dx<\infty,

and hence TzfL(X)fL2(X)\|T_{z}f\|_{L^{\infty}(X)}\leq\|f\|_{L^{2}(X)}. Hence, by analytic interpolation for p>2p>2,

(7.5) fkσLp(X)=T0fLp(X)CfLp(X).\|f\ast k_{\sigma}^{\infty}\|_{L^{p}(X)}=\|T_{0}f\|_{L^{p}(X)}\leq C\|f\|_{L^{p^{\prime}}(X)}.

Therefore, from (7.4) and from (7.5), it follows that for all 2<p2nnσ2<p\leq\frac{2n}{n-\sigma},

fkσLp(X)CfLp(X).\|f\ast k_{\sigma}\|_{L^{p}(X)}\leq C\|f\|_{L^{p^{\prime}}(X)}.

This completes the proof.

As a corollary of the theorem above we have the following

Corollary 7.2.

Let 2<p2nn22<p\leq\frac{2n}{n-2} and dimX=n3\dim X=n\geq 3. Then there exists Sn,p>0S_{n,p}>0 such that for all uH1(X)u\in H^{1}(X),

uL2(X)2|ρ|2uL2(X)2Sn,puLp(X)2.\|\nabla u\|_{L^{2}(X)}^{2}-|\rho|^{2}\|u\|_{L^{2}(X)}^{2}\geq S_{n,p}\|u\|_{L^{p}(X)}^{2}.

Acknowledgement: The first author is supported by the Post Doctoral fellowship from IIT Bombay. The second author is supported partially by SERB, MATRICS, MTR/2017/000235. The authors are thankful to Swagato K Ray for numerous useful discussions and detailed comments. The authors are also grateful to Sundaram Thangavelu for valuable suggestions for the improvement of the paper.

References

  • [1] Andrews, George E.; Askey, Richard; Roy, Ranjan; Special functions. Encyclopedia of Mathematics and its Applications, 71. Cambridge University Press, Cambridge, 1999
  • [2] Anker, J.-P.; LpL_{p} Fourier multipliers on Riemannian symmetric spaces of the noncompact type, Ann. of Math. (2) 132 (1990), no. 3, 597-628.
  • [3] Anker, J.-P.; Sharp estimates for some functions of the Laplacian on noncompact symmetric spaces. Duke Math. J. 65 (1992), no. 2, 257–297.
  • [4] Anker, J.-P.; Ji, L. Heat kernel and Green function estimates on noncompact symmetric spaces, Geom. Funct. Anal. 9 (1999), no. 6, 1035-1091.
  • [5] Anker, J. P.; Ostellari, P. The heat kernel on noncompact symmetric spaces, In Lie groups and symmetric spaces, Amer. Math. Soc. Transl. Ser. 2, 210, Amer. Math. Soc., Providence, RI, 2003, 27-46.
  • [6] Banica, V.; Gonźalez, María del Mar; Sáez, Mariel. Some constructions for the fractional Laplacian on noncompact manifolds, Rev. Mat. Iberoam. 31 (2015), no. 2, 681-712.
  • [7] Beckner, W.; Pitt’s inequality and the uncertainty principle, Proc. Amer. Math. Soc. 123 (1995), 1897-1905.
  • [8] Benguria, R. D.; Frank, R. L.; Loss, M. The sharp constant in the Hardy-Sobolev-Mazýa inequality in the three dimensional upper half-space. Math. Res. Lett. 15 (2008), no. 4, 613–622.
  • [9] Berchio, E.; Ganguly, D.; Grillo, G.; Sharp Poincaré-Hardy and Poincaré-Rellich inequalities on the hyperbolic space, J. Funct. Anal. 272 (2017), 1661-1703.
  • [10] Boggarapu, Pradeep; Roncal, Luz; Thangavelu, Sundaram. On extension problem, trace Hardy and Hardy’s inequalities for some fractional Laplacians, Commun. Pure Appl. Anal. 18 (2019), no. 5, 2575-2605.
  • [11] Caffarelli, Luis; Silvestre, Luis An extension problem related to the fractional Laplacian. Comm. Partial Differential Equations 32 (2007), no. 7-9, 1245–1260.
  • [12] Carron, G.; Inégalitès de Hardy sur les variétés riemanniennes non-compactes, J. Math. Pures Appl. (9) 76 (1997), no. 10, 883-891.
  • [13] Chen, S. A new family of sharp conformally invariant integral inequalities. Int.Math. Res. Not. 2014:1205–1220.
  • [14] Ciatti, P.; Cowling, M.G.; Ricci, F.; Hardy and uncertainty inequalities on stratified Lie groups, Adv. Math. 277 (2015) 365-387.
  • [15] Copson, E.T.; Asymptotic Expansions, Cambridge Univ. Press, 1965.
  • [16] Cowling, M.G.; The Kunze-Stein Phenomenon, Ann. of Math., Vol. 107, No. 2 (1978), pp. 209-234.
  • [17] Cowling, M.; Giulini, S.; Meda, S.; LpLqL^{p}-L^{q} estimates for functions of the Laplace-Beltrami operator on noncompact symmetric spaces I, Duke Math. J. 72 (1993), no. 1, 109-150. MR1242882
  • [18] D’Ambrosio, L.; Dipierro, S.; Hardy inequalities on Riemannian manifolds and applications, Ann. Inst. H. Poinc. Anal. Non Lin. 31 (2014), 449-475.
  • [19] Frank, R.L.; Lieb, E. H.; Seiringer, R.; Hardy-Lieb-Thirring inequalities for fractional Schrödinger operators, J. Amer. Math. Soc. 21 (2008) 925-950.
  • [20] Gangolli, R.; Varadarajan V. S., Harmonic Analysis of Spherical Functions on Real Reductive Groups, Springer-Verlag, Berlin, 1988.
  • [21] Giulini, S.; Mauceri, G.; Almost everywhere convergence of Riesz means on certain noncompact symmetric spaces, Ann. Mat. Pura Appl. (4) 159 (1991), 357-369.
  • [22] Gradshteyn, I.S.; Ryzhik, I.M.; Table of Integrals, Series, and Products, 7th ed. (Elsevier/Academic Press, Amsterdam, 2007).
  • [23] Grafakos, L.; Classical Fourier Analysis, Grad. Texts in Math., Vol. 249, Springer-Verlag, New York, 2008.
  • [24] Helgason, S., Differential geometry, Lie groups, and symmetric spaces, Graduate Studies in Mathematics, 34, American Mathematical Society, Providence, RI, 2001.
  • [25] Helgason, S.,Geometric Analysis on Symmetric Spaces, Mathematical Surveys and Monographs 39. American Mathematical Society, Providence, RI, 2008.
  • [26] Helgason, S., Groups and geometric analysis, Integral geometry, invariant differential operators, and spherical functions, Mathematical Surveys and Monographs, 83. American Mathematical Society, Providence, RI, 2000.
  • [27] Herbst, I.W.; Spectral theory of the operator (p2+m2)1/2Ze2/r(p^{2}+m^{2})^{1/2}-Ze^{2}/r, Comm. Math. Phys. 53 (1977), 285-294.
  • [28] Kombe, I.; Ozaydin, M.; Improved Hardy and Rellich inequalities on Riemannian manifolds, Trans. Amer. Math. Soc. 361 (2009) 6191-6203.
  • [29] Kombe, I.; Ozaydin, M.; Hardy-Poincaré, Rellich and uncertainty principle inequalities on Riemannian manifolds, Trans. Amer. Math. Soc. 365 (2013), 5035-5050.
  • [30] Kristaly, A.; Sharp uncertainty principles on Riemannian manifolds: the influence of curvature, Journal de Mathématiques Pures et Appliquées, 119 (2018), 326-346.
  • [31] Li, Jungang; Lu, Gouzhen; Yang Q.; Sharp Adams and Hardy-Adams inequalities of any fractional order on hyperbolic spaces of all dimensions, Transactions of the American Mathematical Society 373 (2020), 3483-3513.
  • [32] Lu, G., Yang, Q.; Paneitz operators on hyperbolic spaces and high order Hardy-Sobolev-Maz’ya inequalities on half spaces, American Journal of Mathematics 141 (2019), 1777-1816.
  • [33] Mancini, Gianni; Sandeep, Kunnath On a semilinear elliptic equation in n\mathbb{H}^{n}. Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 7 (2008), no. 4, 635–671.
  • [34] Möllers, Jan; Ø\Orsted, Bent; Zhang, Genkai. On boundary value problems for some conformally invariant differential operators, Comm. Partial Differential Equations 41 (2016), no. 4, 609-643.
  • [35] Ray, Swagato K.; Sarkar, Rudra P. Fourier and Radon transform on harmonic NA groups. Trans. Amer. Math. Soc. 361 (2009), no. 8, 4269-4297.
  • [36] Roncal, Luz; Thangavelu, Sundaram. An Extension Problem and Trace Hardy Inequality for the Sub-Laplacian on HH-type Groups, International Mathematics Research Notices, https://doi.org/10.1093/imrn/rny137.
  • [37] Roncal, Luz; Thangavelu, Sundaram. Hardy’s inequality for fractional powers of the sublaplacian on the Heisenberg group, Adv. Math. 302 (2016), 106-158.
  • [38] Ruzhansky, M.; Suragan D.; Hardy and Rellich inequalities, identities, and sharp remainders on homogeneous groups, Advances in Mathematics 317 (2017), 799-822.
  • [39] Stinga, Pablo Raúl; Torrea, José Luis. Extension problem and Harnack’s inequality for some fractional operators, Comm. Partial Differential Equations 35 (2010), no. 11, 2092-2122.
  • [40] Varopoulos, N. Th. Sobolev inequalities on Lie groups and symmetric spaces. J. Funct. Anal. 86 (1989), no. 1, 19-40.
  • [41] Yafaev, D.; Sharp constants in the Hardy-Rellich inequalities, J. Funct. Anal. 168 (1999), 121-144.
  • [42] Yang, Q.; Su, D.; Kong, Y.; Hardy inequalities on Riemannian manifolds with negative curvature, Commun. Contemp. Math. 16 (2014), 1350043.