Skip to main content
Log in

Riesz Transforms Characterizations of Hardy Spaces \(H^1\) for the Rational Dunkl Setting and Multidimensional Bessel Operators

  • Published:
The Journal of Geometric Analysis Aims and scope Submit manuscript

Abstract

We characterize the Hardy space \(H^1\) in the rational Dunkl setting associated with the reflection group \(\mathbb {Z}_2^n\) by means of special Riesz transforms. As a corollary we obtain Riesz transforms characterization of \(H^1\) for product of Bessel operators in \((0,\infty )^n\).

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Institutional subscriptions

Similar content being viewed by others

References

  1. Amri, B., Sifi, M.: Riesz transform for Dunkl transform. Ann. Math. Blaise Pascal 19, 247–262 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  2. Anker, J-Ph, Ben Salem, N., Dziubański, J., Hamda, N.: The Hardy space \(H^1\) in the rational Dunkl setting. Constr. Approx. 42, 93–128 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  3. Burkholder, D.L., Gundy, R.F., Silverstein, M.L.: A maximal function characterisation of the class \(H^p\). Trans. Am. Math. Soc. 157, 137–153 (1971)

    MathSciNet  MATH  Google Scholar 

  4. Coifman, R.R.: A real variable characterization of \(H^p\). Studia Math. 51, 269–274 (1974)

    MathSciNet  MATH  Google Scholar 

  5. Coifman, R.R., Weiss, G.L.: Extensions of Hardy spaces and their use in analysis. Bull. Am. Math. Soc. 83, 569–615 (1977)

    Article  MathSciNet  MATH  Google Scholar 

  6. Deleaval, L.: Fefferman-Stein inequalities for the \({\mathbb{Z}}^d_2\) Dunkl maximal operator. J. Math. Anal. Appl. 360, 711–726 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  7. de Jeu, M.F.E.: The Dunkl transform. Invent. Math. 113, 147–162 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  8. Dunkl, C.F.: Reflection groups and orthogonal polynomials on the sphere. Math. Z. 197, 33–60 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  9. Dunkl, C.F.: Differential-difference operators associated to reflection groups. Trans. Amer. Math. 311, 167–183 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  10. Dunkl, C.F.: Hankel transforms associated to finite reflection groups. In: Proceedings of the Special Session on Hypergeometric Functions on Domains of Positivity, Jack polynomials and applications. Proceedings, Tampa 1991, Contemp. Math., vol. 138, pp. 123–138 (1989)

  11. Dunkl, C.F.: Integral kernels with reflection group invariance. Canad. J. Math. 43, 1213–1227 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  12. Dziubański, J., Preisner, M., Wróbel, B.: Multivariate Hörmander-type multiplier theorem for the Hankel transform. J. Fourier Anal. Appl. 19, 417–437 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  13. Evans, L.C.: Partial Differential Equations. Graduate Studies in Mathematics, vol. 19. AMS, Providence, RI (1998)

    Google Scholar 

  14. Fefferman, C., Stein, E.M.: \(H^p\) spaces of several variables. Acta Math. 129, 137–193 (1972)

    Article  MathSciNet  MATH  Google Scholar 

  15. Macías, R.A., Segovia, C.: A decomposition into atoms of distributions on spaces of homogeneous type. Adv. Math. 33, 271–309 (1979)

    Article  MathSciNet  MATH  Google Scholar 

  16. Muckenhoupt, B., Stein, E.: Classical expansions and their relation to conjugate harmonic functions. Trans. Am. Math. Soc. 118, 17–92 (1965)

    Article  MathSciNet  MATH  Google Scholar 

  17. NIST Digital Library of Mathematical Functions. http://dlmf.nist.gov

  18. Opdam, E.M.: Lecture notes on Dunkl operators for real and complex reflection groups. MSJ Memoirs, vol. 8. Mathematical Society of Japan, Tokyo (2000)

  19. Rösler, M.: Generalized Hermite polynomials and the heat equation for Dunkl operators. Commun. Math. Phys. 192, 519–542 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  20. Rösler, M.: Dunkl operators: theory and applications. In: Orthogonal polynomials and special functions (Leuven, 2002). Lecture Notes in Mathematics, vol. 1817, pp. 93–135. Springer, Berlin (2003)

  21. Rösler, M., Voit, M.: Dunkl theory, convolution algebras, and related Markov processes. In: Harmonic and stochastic analysis of Dunkl processes, Collection Travaux en cours, vol. 71, pp. 1–112. Hermann, Paris (2008)

  22. Stein, E.M., Weiss, G.L.: On the theory of harmonic functions of several variables I (the theory of \(H^p\) spaces). Acta Math. 103, 25–62 (1960)

    Article  MathSciNet  MATH  Google Scholar 

  23. Stein, E.M.: Singular Integrals and Differentiability of Functions. Princeton University Press, Princeton, NJ (1971)

    Google Scholar 

  24. Stein, E.M.: Harmonic Analysis (Real-Variable Methods, Orthogonality, and Oscillatory Integrals). Princeton Mathematics Series, vol. 43. Princeton University Press, Princeton, NJ (1993)

    Google Scholar 

  25. Thangavelu, S., Xu, Y.: Convolution operator and maximal function for the Dunkl transform. J. Anal. Math. 97, 25–55 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  26. Uchiyama, A.: A maximal function characterization of \(H^p\) on the space of homogeneous type. Trans. Am. Math. Soc. 262(2), 579–592 (1980)

    MathSciNet  MATH  Google Scholar 

  27. Watson, G.N.: A Treatise on the Theory of Bessel Functions. Cambridge University Press, Cambridge, MA (1995)

    MATH  Google Scholar 

Download references

Acknowledgments

The author wishes to thank Jean-Philippe Anker, Paweł Głowacki and Rysiek Szwarc for their remarks. The author is greatly indebted to Bartosz Trojan for his suggestions which shortened the original proof of Theorem 1.3. Finally, the author wants to thank the referee for her/his careful reading of the manuscript and helpful comments which improved the presentation of the paper. Research supported by the Polish National Science Center (Narodowe Centrum Nauki, grant DEC-2012/05/B/ST1/00672) and by the University of Orléans.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jacek Dziubański.

Appendix

Appendix

It is well known that

$$\begin{aligned} \int _{{\mathbb {R}}} h^{\{j\}}_{ t}(x,y)\, d\mu _j(y)=1, \quad d\mu _j(y)= |y|^{2k_j}\, dy. \end{aligned}$$
(7.2)

It was proved in [2] that \(h^{\{j\}}_t (x,y)\) has the following global behavior :

$$\begin{aligned} h^{\{j\}}_{ t}(x,y)\,\asymp \, {\left\{ \begin{array}{ll} t^{-k_j-\frac{1}{2}}\, e^{-\frac{x^2+ y^2}{4 t}} &{}\quad \text {if }| x y|\le t,\\ t^{-\frac{1}{2}}\,(x y)^{-k_j}\, e^{-\frac{(x-y)^2}{4 t}} &{}\quad \text {if }x y\ge t,\\ t^{\frac{1}{2}}\,(- x y)^{-k_j-1}\, e^{-\frac{(x+y)^2}{4 t}} &{}\quad \text {if }-x y\ge t.\\ \end{array}\right. } \end{aligned}$$
(7.3)

From (7.3) we easily conclude that

$$\begin{aligned}&0<{\mathbf {h}}_{t}({\mathbf {x}},{\mathbf {y}})\le \frac{C}{{\varvec{\mu }}(B({\mathbf {x}},\sqrt{t}))} \text {for all } \ {\mathbf {x}}, {\mathbf {y}} \in {\mathbb {R}}^n \text {and } t>0; \end{aligned}$$
(7.4)
$$\begin{aligned}&{\mathbf {h}}_{t}({\mathbf {x}},{\mathbf {y}})\le \frac{C}{{\varvec{\mu }}(B({\mathbf {x}},\sqrt{t}))} e^{-c|{\mathbf {x}}|^2\slash t} \text { for } |{\mathbf {x}}|>2n|{\mathbf {y}}|. \end{aligned}$$
(7.5)

We shall need the following inequalities for volumes of the Euclidean balls (see [2])

$$\begin{aligned} \big (\tfrac{R}{r}\big )^{n}\lesssim \tfrac{\varvec{\mu }(\mathbf {B}(\mathbf {x}, R))}{\varvec{\mu }(\mathbf {B}(\mathbf {x}, r))} \lesssim \big (\tfrac{R}{r}\big )^{\mathbf {N}}, \qquad \forall \;\mathbf {x}\in {\mathbb {R}}^n,\,\forall \;R\ge r>0. \end{aligned}$$
(7.6)

The subordination formula (2.1) combined with (7.1) and (7.2) implies

$$\begin{aligned} \int _{{\mathbb {R}}^n} P_t({\mathbf {x}},{\mathbf {y}})\, d{\varvec{\mu }} ({\mathbf {x}})=\int _{{\mathbb {R}}^n} P_t({\mathbf {x}},{\mathbf {y}})\, d{\varvec{\mu }} ({\mathbf {y}})=1. \end{aligned}$$
(7.7)

Lemma 7.1

There is a constant \(C>0\) such that

$$\begin{aligned} 0< P_t({\mathbf {x}},{\mathbf {y}})\le \frac{C}{{\varvec{\mu }} (B({\mathbf {x}}, t))}. \end{aligned}$$
(7.8)

Moreover, for every \(0<\delta <\frac{1}{\mathbf N}\) there is a constant \(C_\delta \) such that

$$\begin{aligned} P_t({\mathbf {x}},{\mathbf {y}})\le \frac{C_\delta }{{\varvec{\mu }} (B({\mathbf {x}}, t))}\left( 1+\frac{ {\varvec{\mu }} (B({\mathbf {x}}, |{\mathbf {x}}|))}{{\varvec{\mu }} (B({\mathbf {x}},t))}\right) ^{-1-\delta } \text { for } |{\mathbf {x}}|>2n|{\mathbf {y}}|. \end{aligned}$$
(7.9)

Proof

To see (7.8) we use (2.1) together with (7.4) and (7.6) and obtain

$$\begin{aligned}\begin{aligned} P_t({\mathbf {x}},{\mathbf {y}})&\lesssim \int _0^{\frac{1}{4}} \frac{1}{{\varvec{\mu }} (B({\mathbf {x}},t))} \frac{{\varvec{\mu }} (B({\mathbf {x}},t))}{{\varvec{\mu }} (B({\mathbf {x}},\frac{t}{2\sqrt{u}}))}\frac{du}{\sqrt{u}}+ \int _{\frac{1}{4}}^\infty \frac{e^{-u}}{{\varvec{\mu }} (B({\mathbf {x}},t))} \frac{{\varvec{\mu }} (B({\mathbf {x}},t))}{{\varvec{\mu }} (B({\mathbf {x}},\frac{t}{2\sqrt{u}}))}\frac{du}{\sqrt{u}}\\&\lesssim \frac{1}{\mu (B({\mathbf {x}},t))} \left( \int _0^{\frac{1}{4}} u^{n\slash 2}\frac{d}{\sqrt{u}}+\int _{\frac{1}{4}}^\infty e^{-u} u^{\mathbf N\slash 2}\frac{du}{\sqrt{u}}\right) \lesssim \frac{1}{\mu (B({\mathbf {x}},t))}. \end{aligned}\end{aligned}$$

The proof of the lower bound of \(P_t({\mathbf {x}},{\mathbf {y}})\) is obvious.

In order to prove (7.9) it suffices to consider \(t\le |{\mathbf {x}}|\slash 2\). By (7.6), for every \(\delta >0\) and \(c>0\), we have

$$\begin{aligned} \left( 1+\frac{{\varvec{\mu }} (B({\mathbf {x}}, |{\mathbf {x}}|))}{{\varvec{\mu }} (B({\mathbf {x}},\sqrt{s}))} \right) ^{1+\delta }\le C_\delta \left( 1+\frac{|{\mathbf {x}}|}{\sqrt{s}}\right) ^{(1+\delta ) \mathbf N} \le C_{\delta , c} e^{c|{\mathbf {x}}|^2\slash s}, \text {for } \ s>0. \end{aligned}$$
(7.10)

Utilizing (7.5) together with (7.10) and proceeding similarly to the proof of (7.8) we have

$$\begin{aligned}&P_t({\mathbf {x}},{\mathbf {y}}) \lesssim \int _0^\infty \frac{e^{-u}}{{\varvec{\mu }} (B({\mathbf {x}}, \frac{t}{2\sqrt{u}}))} \left( 1+\frac{{\varvec{\mu }} (B({\mathbf {x}}, |{\mathbf {x}}|))}{{\varvec{\mu }} (B({\mathbf {x}}, \frac{t}{2\sqrt{u}}))}\right) ^{-1-\delta }\frac{du}{\sqrt{u}}\\&\quad = \int _0^{\infty } \frac{e^{-u}}{{\varvec{\mu }} (B({\mathbf {x}},t))} \frac{{\varvec{\mu }} (B({\mathbf {x}},t))}{{\varvec{\mu }} (B({\mathbf {x}},\frac{t}{2\sqrt{u}}))}\\&\qquad \times \left( 1+\frac{{\varvec{\mu }} (B({\mathbf {x}}, |{\mathbf {x}}|))}{{\varvec{\mu }} (B({\mathbf {x}}, t))} \frac{{\varvec{\mu }} (B({\mathbf {x}}, t))}{{\varvec{\mu }} (B({\mathbf {x}}, \frac{t}{2\sqrt{u}}))}\right) ^{-1-\delta } \frac{du}{\sqrt{u}}\\&\quad \lesssim \int _0^{\frac{1}{4}} \frac{1}{{\varvec{\mu }} (B({\mathbf {x}},t))} \frac{{\varvec{\mu }} (B({\mathbf {x}},t))}{{\varvec{\mu }} (B({\mathbf {x}},\frac{t}{2\sqrt{u}}))} \left( 1+\frac{{\varvec{\mu }} (B({\mathbf {x}}, |{\mathbf {x}}|))}{{\varvec{\mu }} (B({\mathbf {x}}, t))}\right) ^{-1-\delta }\\&\quad \quad \times \left( \frac{{\varvec{\mu }} (B({\mathbf {x}}, t))}{{\varvec{\mu }} (B({\mathbf {x}}, \frac{t}{2\sqrt{u}}))}\right) ^{-1-\delta } \frac{du}{\sqrt{u}}\\&\qquad + \int _{\frac{1}{4}}^{\infty } \frac{e^{-u}}{{\varvec{\mu }} (B({\mathbf {x}},t))} \frac{{\varvec{\mu }} (B({\mathbf {x}},t))}{{\varvec{\mu }} (B({\mathbf {x}},\frac{t}{2\sqrt{u}}))} \left( 1+\frac{{\varvec{\mu }} (B({\mathbf {x}}, |{\mathbf {x}}|))}{{\varvec{\mu }} (B({\mathbf {x}}, t))}\right) ^{-1-\delta } \frac{du}{\sqrt{u}}.\\ \end{aligned}$$

Fix \(0<\delta <\mathbf {N}^{-1}\). Applying (7.6) we obtain

$$\begin{aligned} \begin{aligned} P_t({\mathbf {x}},{\mathbf {y}})&\lesssim \int _0^{\frac{1}{4}} \frac{1}{{\varvec{\mu }} (B({\mathbf {x}},t))} \left( 1+\frac{{\varvec{\mu }} (B({\mathbf {x}}, |{\mathbf {x}}|))}{{\varvec{\mu }} (B({\mathbf {x}}, t))}\right) ^{-1-\delta } u^{-\delta \mathbf N\slash 2} \frac{du}{\sqrt{u}}\\&\qquad + \int _{\frac{1}{4}}^{\infty } \frac{e^{-u}u^{\mathbf N\slash 2}}{{\varvec{\mu }} (B({\mathbf {x}},t))} \left( 1+\frac{{\varvec{\mu }} (B({\mathbf {x}}, |{\mathbf {x}}|))}{{\varvec{\mu }} (B({\mathbf {x}}, t))}\right) ^{-1-\delta } \frac{du}{\sqrt{u}},\\ \end{aligned} \end{aligned}$$

which proves (7.9). \(\square \)

Proof of Theorem 2.1

The proof, which is in its spirit similar to that of the heat kernel characterization of \(H^1_\mathrm{{atom}}\) (see [2]), is based on the following result due to Uchiyama [26].

Theorem 7.2

Assume that a set X is equipped with

  • a quasi-distance \(\widetilde{d}\) i.e., a distance except that the triangular inequality is replaced by the weaker condition

    $$\begin{aligned} \widetilde{d}(x,y)\le A\,\{\widetilde{d}(x,z)+\widetilde{d}(z,y)\}, \qquad \forall \;x,y,z\in X; \end{aligned}$$
  • a measure \(\mu \) whose values on quasi-balls satisfy

    $$\begin{aligned} \frac{r}{A}\le \mu (\widetilde{B}(x,r))\le r, \qquad \forall \;x\in X,\,\forall \;r>0\,; \end{aligned}$$
  • a continuous kernel \(K_r(x,y)\ge 0\) such that, for every \(r>0\) and \(x,y,y^{\prime }\in X\),

    • \(K_r(x,x)\ge \frac{1}{Ar}\),

    • \(K_r(x,y)\le r^{-1}\bigl ( 1+\frac{\widetilde{d}(x, y)}{r} \bigr )^{-1-\delta }\) ,

    • \(\bigl | K_r(x,y)-K_r(x,y^{\prime })\bigr |\le r^{-1}\bigl (1+\frac{\widetilde{d}(x, y)}{r}\bigr )^{-1-2\delta } \bigl (\frac{\widetilde{d}(y, y^{\prime })}{r}\bigr )^{\delta }\) when \(\widetilde{d}(y,y^{\prime }) \le \frac{r+\widetilde{d}(x, y)}{4 A}\) .

Here \(A\ge 1\) and \(\delta >0\). Then the following definitions of the Hardy space \(H^1(X)\) and their corresponding norms are equivalent :

  • Maximal definition : \(H^1_{\mathrm{max}, K_r}(X)\) consists of all functions \(f\in L^1(X,d\mu )\) such that

    $$\begin{aligned} K_*f(x)=\sup \nolimits _{\,r>0}\, \Bigl |{\displaystyle \int _X}K_r(x,y)\,f(y)\,d\mu (y)\Bigr | \end{aligned}$$

    belongs to \(L^1(X,d\mu )\) and the norm \(\Vert f\Vert _{H^1_{\mathrm{max}, K_r}(X)} = \Vert K_*f\Vert _{L^1(X,d\mu )}\).

  • Atomic definition : An atom for \(H^1_\mathrm{atom}(X,\widetilde{d})\) is a measurable function \(a:X\rightarrow {\mathbb {C}}\)  such that: a is supported in a quasi-ball \(\widetilde{B}\), \(\Vert a\Vert _{L^\infty }\lesssim \mu (\widetilde{B})^{-1}\) and \(\displaystyle \int _Xa \,d\mu =0\) (see [5, 15, 26]). Then \(H^1_\mathrm{atom}(X, \widetilde{d})\) consists of all functions \(f\in L^1(X,d\mu )\) which can be written as \(f=\sum _{\ell }\lambda _{\ell } a_{\ell }\), where the \(a_{\ell }\)’s are atoms and \(\sum _{\ell }|\lambda _{\ell }|<+\infty \), and the norm \(\Vert f\Vert _{H^1_\mathrm{atom}(X,\widetilde{d})}= \inf \sum _{\ell }|\lambda _{\ell }|\), where the infimum is taken over all such representations.

For \(X={\mathbb {R}}^n\), equipped with the Euclidean distance \(d(\mathbf {x},\mathbf {y})=|\mathbf {x}-\mathbf {y}|\) and the measure \({\varvec{\mu }}\) (see (1.5)), set

$$\begin{aligned} \widetilde{d}(\mathbf {x},\mathbf {y})=\inf \varvec{\mu }(B), \qquad \forall \;\mathbf {x},\mathbf {y}\in {\mathbb {R}}^n, \end{aligned}$$

where the infimum is taken over all closed balls B containing \(\mathbf {x}\) and \(\mathbf {y}\). Let \(t=t(\mathbf {x},r)\) be defined by \(\varvec{\mu }(B(\mathbf {x},\sqrt{t}))=r\). Then

$$\begin{aligned} {\varvec{\mu }}(\widetilde{B}({\mathbf {x}},r))\sim r \end{aligned}$$

and there exists a constant \(c>0\) such that

$$\begin{aligned} B(\mathbf {x},\sqrt{ t\,}) \subset \widetilde{B}(\mathbf {x},r) \subset B(\mathbf {x},c\sqrt{ t\,}), \end{aligned}$$
(7.11)

where \(\widetilde{B}(\mathbf {x},r)=\{ {\mathbf {y}}\in {\mathbb {R}}^n: \widetilde{d}({\mathbf {x}},{\mathbf {y}})<r\}\) (see, e.g., [2]).

Let us remark that thanks to (7.11) and (7.6) the atomic spaces \(H^1_\mathrm{atom}(X,\widetilde{d})\) and \(H^1_\mathrm{atom}\) (defined in Section 1) do coincide and \(\Vert f\Vert _{H^1_\mathrm{atom}(X,\widetilde{d})} \sim \Vert f\Vert _{H^1_\mathrm{atom}}\).

It was proved in [2] that the kernel \({\mathbf {h}}_t\) can be written in the form

$$\begin{aligned} {\mathbf {h}}_{ t}(\mathbf {x},{\mathbf {y}}) ={\mathbf {H}}_{ t}(\mathbf {x},{\mathbf {y}}) +{\mathbf {S}}_{t}(\mathbf {x},{\mathbf {y}}), \end{aligned}$$

where \({\mathbf {H}}_{ t}(\mathbf {x},{\mathbf {y}})\) and \({\mathbf { S}}_{t}(\mathbf {x},{\mathbf {y}})\) are nonnegative continuous functions such that there are \(C_1,\, C_2,\, C_4,\, \delta >0\) such that

$$\begin{aligned}&{\mathbf {H}}_t(\mathbf {x},\mathbf {x})\ge \tfrac{C_1}{{\varvec{\mu }} (B({\mathbf {x}},\sqrt{t}))}; \end{aligned}$$
(7.12)
$$\begin{aligned}&{\mathbf {H}}_t(\mathbf {x},\mathbf {y})\le \tfrac{C_2}{ {\varvec{\mu }} (B({\mathbf {x}},\sqrt{t}))} \,\bigl (1+\tfrac{\widetilde{d}(\mathbf {x},\mathbf {y})}{ {\varvec{\mu }} (B({\mathbf {x}},\sqrt{t}))} \bigr )^{-1-\delta }; \end{aligned}$$
(7.13)
$$\begin{aligned}&\bigl | {\mathbf {H}}_t (\mathbf {x},\mathbf {y})-{\mathbf {H}}_t(\mathbf {x},\mathbf {y}^{\prime })\bigr | \le \tfrac{C_4}{{\varvec{\mu }} ( B ({\mathbf {x}},\sqrt{t}))} \bigl ( 1+\tfrac{\widetilde{d}(\mathbf {x},\mathbf {y})}{{\varvec{\mu }} (B({\mathbf {x}},\sqrt{t}))} \bigr )^{-1-2\delta } \bigl (\tfrac{\widetilde{d}(\mathbf {y},\mathbf {y}^{\prime })}{{\varvec{\mu }} ( B({\mathbf {x}},\sqrt{t}))} \bigr )^{\frac{1}{\mathbf {N}}}\nonumber \\ \end{aligned}$$
(7.14)

for \(\widetilde{d}(\mathbf {y},\mathbf {y}^{\prime })\le C_3\max \,\{ {\varvec{\mu }} (B({\mathbf {x}},\sqrt{t})),\widetilde{d}(\mathbf {x},\mathbf {y})\}\), (the kernel \({\mathbf {S}}_t\) is denoted in [2] by \(\mathbf {P}_t\)). Moreover, the maximal function

$$\begin{aligned} {\mathbf {S}}_*f(x)=\sup _{t>0} \Big |\int {\mathbf {S}}_t({\mathbf {x}},{\mathbf {y}})f({\mathbf {y}})d{\varvec{\mu }}({\mathbf {y}}) \Big | \end{aligned}$$

is a bounded operator on \(L^1({\mathbb {R}}^n,\, d{\varvec{\mu }} )\).

Using subordination formula (2.1) we write

$$\begin{aligned} P_t({\mathbf {x}},{\mathbf {y}}) = U_t({\mathbf {x}},{\mathbf {y}}) + W_t({\mathbf {x}},{\mathbf {y}}), \end{aligned}$$
(7.15)

where

$$\begin{aligned} U_t({\mathbf {x}},{\mathbf {y}})=c_1\int _0^\infty e^{-u}{\mathbf {H}}_{t^2\slash 4u}({\mathbf {x}},{\mathbf {y}})\frac{du}{\sqrt{u}}, W_t({\mathbf {x}},{\mathbf {y}}) = c_1\int _0^\infty e^{-u}{\mathbf {S}}_{t^2\slash 4u}({\mathbf {x}},{\mathbf {y}})\frac{du}{\sqrt{u}}. \end{aligned}$$

Clearly, the maximal operator

$$\begin{aligned} W_*f({\mathbf {x}})=\sup _{t>0} \Big |\int W_t({\mathbf {x}},{\mathbf {y}})f({\mathbf {y}}) d{\varvec{\mu }}({\mathbf {y}})\Big | \end{aligned}$$

is bounded on \(L^1({\mathbb {R}}^n,\, d{\varvec{\mu }} )\), that is,

$$\begin{aligned} \Vert W_*f\Vert _{L^1({\mathbb {R}}^n,\, d{\varvec{\mu }} )} \le C \Vert f\Vert _{L^1({\mathbb {R}}^n,\, d{\varvec{\mu }} )}. \end{aligned}$$
(7.16)

\(\square \)

Our task is to prove the following lemma.

Lemma 7.3

There are constants \(C_1,\, C_2,\, C_4,\, \delta ' >0\) such that

$$\begin{aligned}&U_t(\mathbf {x},\mathbf {x})\ge \tfrac{C_1}{{\varvec{\mu }} (B ({\mathbf {x}},t))}; \end{aligned}$$
(7.17)
$$\begin{aligned}&U_t(\mathbf {x},\mathbf {y})\le \tfrac{C_2}{ {\varvec{\mu }} (B({\mathbf {x}},t))} \,\bigl (1+\tfrac{\widetilde{d}(\mathbf {x},\mathbf {y})}{ {\varvec{\mu }} (B({\mathbf {x}},t))} \bigr )^{-1-\delta '}; \end{aligned}$$
(7.18)
$$\begin{aligned}&\bigl | U_t (\mathbf {x},\mathbf {y})-U_t(\mathbf {x},\mathbf {y}^{\prime })\bigr | \le \tfrac{C_4}{{\varvec{\mu }} (B({\mathbf {x}}, t))} \bigl ( 1+\tfrac{\widetilde{d}(\mathbf {x},\mathbf {y})}{{\varvec{\mu }} (B({\mathbf {x}},t))} \bigr )^{-1-2\delta '} \bigl (\tfrac{\widetilde{d}(\mathbf {y},\mathbf {y}^{\prime })}{{\varvec{\mu }} ( B({\mathbf {x}},t))} \bigr )^{\delta '} \end{aligned}$$
(7.19)

for \(\widetilde{d}(\mathbf {y},\mathbf {y}^{\prime })\le C_3\max \,\{ {\varvec{\mu }} (B({\mathbf {x}},t )),\widetilde{d}(\mathbf {x},\mathbf {y})\}\) .

Proof

Take \(0<\delta <\mathbf {N}^{-1}\). By (7.12) and the subordination formula we have

$$\begin{aligned} \begin{aligned} U_t({\mathbf {x}},{\mathbf {x}})&\gtrsim \int _1^\infty \frac{e^{-u}}{{\varvec{\mu }} (B({\mathbf {x}},\frac{t}{2\sqrt{u}}))} \frac{du}{\sqrt{u}} \gtrsim \int _1^\infty \frac{e^{-u}}{{\varvec{\mu }}(B({\mathbf {x}}, t))} \frac{{\varvec{\mu }}(B({\mathbf {x}}, t))}{{\varvec{\mu }} (B({\mathbf {x}},\frac{t}{2\sqrt{u}}))} \frac{du}{\sqrt{u}}\\&\gtrsim \frac{1}{{\varvec{\mu }}(B({\mathbf {x}}, t))} \int _1^\infty \frac{e^{-u}}{\sqrt{u}}\, du \gtrsim \frac{1}{{\varvec{\mu }}(B({\mathbf {x}}, t))}, \end{aligned} \end{aligned}$$

which proves (7.17).

The proof of (7.18) is similar to that of (7.9). Indeed, by (7.13) we have

$$\begin{aligned} \begin{aligned} U_t({\mathbf {x}},{\mathbf {y}})&\le \int _0^\infty \frac{e^{-u}}{{\varvec{\mu }}(B({\mathbf {x}}, t))} \frac{{\varvec{\mu }}(B({\mathbf {x}}, t))}{{\varvec{\mu }}(B({\mathbf {x}}, \frac{t}{2\sqrt{u}}))}\\&\qquad \left( 1+ \frac{\tilde{d}({\mathbf {x}},{\mathbf {y}})}{{{\varvec{\mu }}(B({\mathbf {x}}, t))}}\frac{{\varvec{\mu }}(B({\mathbf {x}}, t))}{{\varvec{\mu }}(B({\mathbf {x}}, \frac{t}{2\sqrt{u}}))}\right) ^{-1-\delta }\frac{du}{\sqrt{u}}\\&\le \int _0^{\frac{1}{4}} \frac{e^{-u}}{{\varvec{\mu }}(B({\mathbf {x}}, t))} \frac{{\varvec{\mu }}(B({\mathbf {x}}, t))}{{\varvec{\mu }}(B({\mathbf {x}}, \frac{t}{2\sqrt{u}}))} \left( 1+ \frac{\tilde{d}({\mathbf {x}},{\mathbf {y}})}{{{\varvec{\mu }}(B({\mathbf {x}}, t))}}\right) ^{-1-\delta }\\&\qquad \left( \frac{{\varvec{\mu }}(B({\mathbf {x}}, t))}{{\varvec{\mu }}(B({\mathbf {x}}, \frac{t}{2\sqrt{u}}))} \right) ^{-1-\delta } \frac{du}{\sqrt{u}}\\&\qquad + \int _{\frac{1}{4}}^\infty \frac{e^{-u}}{{\varvec{\mu }}(B({\mathbf {x}}, t))} \frac{{\varvec{\mu }}(B({\mathbf {x}}, t))}{{\varvec{\mu }}(B({\mathbf {x}}, \frac{t}{2\sqrt{u}}))} \left( 1+ \frac{\tilde{d}({\mathbf {x}},{\mathbf {y}})}{{{\varvec{\mu }}(B({\mathbf {x}}, t))}}\right) ^{-1-\delta }\frac{du}{\sqrt{u}}.\\ \end{aligned} \end{aligned}$$

Now using (7.6) we obtain (7.18).

Now we turn to the proof of (7.19). First we show that there is a constant \(C_4>0\) such that

$$\begin{aligned} \bigl | U_t (\mathbf {x},\mathbf {y})-U_t(\mathbf {x},\mathbf {y}^{\prime })\bigr | \le \tfrac{C_4}{{\varvec{\mu }} (B({\mathbf {x}}, t))} \bigl (\tfrac{\widetilde{d}(\mathbf {y},\mathbf {y}^{\prime })}{{\varvec{\mu }} ( B({\mathbf {x}},t))} \bigr )^{\frac{1}{\mathbf {N}}} \text { for every } {\mathbf {x}},{\mathbf {y}}, {\mathbf {y}}'\in {\mathbb {R}}^n. \end{aligned}$$
(7.20)

Since \(U_t({\mathbf {x}},{\mathbf {y}})\le C{\varvec{\mu }} (B({\mathbf {x}},t))^{-1}\) (see (7.18)), it suffices to prove (7.20) for \(\tilde{d}({\mathbf {y}},{\mathbf {y}}')\le {\varvec{\mu }} (B({\mathbf {x}},t))\). Let \(u_0\ge 1\slash 4\) be such that \({\varvec{\mu }}(B({\mathbf {x}},\frac{t}{2\sqrt{u_0}}))=\tilde{d}({\mathbf {y}},{\mathbf {y}}')\). Then, using (7.14) and (7.6), we have

$$\begin{aligned}&\displaystyle \int _0^{u_0} e^{-u}|{\mathbf {H}}_{\frac{t^2}{4u}}({\mathbf {x}},{\mathbf {y}})-{\mathbf {H}}_{\frac{t^2}{ 4u}}({\mathbf {x}},{\mathbf {y}}')|\frac{du}{\sqrt{u}}\\&\quad \displaystyle \lesssim \int _0^{u_0} \frac{e^{-u}}{{\varvec{\mu }} (B(\mathbf x,\frac{t}{2\sqrt{u}}))} \left( \frac{\tilde{d}({\mathbf {y}},\mathbf y')}{{\varvec{\mu }}(B({\mathbf {x}},\frac{t}{2\sqrt{u}}))}\right) ^{\frac{1}{ \mathbf N} } \frac{du}{\sqrt{u}}\\&\quad \displaystyle = \int _0^{u_0} \frac{e^{-u}}{{\varvec{\mu }} (B({\mathbf {x}},t))} \left( \frac{\tilde{d}({\mathbf {y}},\mathbf y')}{{\varvec{\mu }}(B({\mathbf {x}},t))}\right) ^{\frac{1}{ \mathbf N} } \left( \frac{\mu (B({\mathbf {x}},t))}{\mu (B(\mathbf {x},\frac{t}{2\sqrt{u}}))}\right) ^{1+\frac{1}{\mathbf N}} \frac{du}{\sqrt{u}}\\&\quad \displaystyle \lesssim \frac{1}{{\varvec{\mu }} (B({\mathbf {x}},t))} \left( \frac{\tilde{d}({\mathbf {y}},\mathbf y')}{{\varvec{\mu }}(B({\mathbf {x}},t))}\right) ^{\frac{1}{ \mathbf N} }\\&\quad \quad \displaystyle \left( \int _0^{1\slash 4} e^{-u}u^{n(1+\frac{1}{\mathbf N})\slash 2}\frac{du}{\sqrt{u}}+\int _{1\slash 4}^{u_0}e^{-u}u^{\mathbf N(1+\mathbf N^{-1})\slash 2} \frac{du}{\sqrt{u}}\right) \\&\quad \displaystyle \lesssim \frac{1}{{\varvec{\mu }} (B({\mathbf {x}},t))} \left( \frac{\tilde{d}({\mathbf {y}},\mathbf y')}{{\varvec{\mu }}(B({\mathbf {x}},t))}\right) ^{\frac{1}{ \mathbf N} }. \end{aligned}$$

Similarly, by (7.13), we get

$$\begin{aligned}\begin{aligned} \int _{u_0}^\infty e^{-u}|{\mathbf {H}}_{\frac{t^2}{4u}}({\mathbf {x}},{\mathbf {y}})-{\mathbf {H}}_{\frac{t^2}{ 4u}}({\mathbf {x}},{\mathbf {y}}')|\frac{du}{\sqrt{u}}&\lesssim \int _{u_0}^\infty \frac{e^{-u}}{{\varvec{\mu }}(B({\mathbf {x}},t))}\left( \frac{\mu (B({\mathbf {x}},t))}{ \mu (B({\mathbf {x}},\frac{t}{2\sqrt{u}}))}\right) \frac{du}{\sqrt{u}} \\&\lesssim \frac{1}{{\varvec{\mu }}(B({\mathbf {x}},t))} \int _{u_0}^\infty e^{-u}u^{\mathbf N\slash 2}\frac{du}{\sqrt{u}}\\&\lesssim \frac{1}{{\varvec{\mu }}(B({\mathbf {x}},t))}u_0^{-\mathbf N\slash 2}. \end{aligned} \end{aligned}$$

Since

$$\begin{aligned} \frac{\tilde{d}({\mathbf {y}},{\mathbf {y}}')}{ {\varvec{\mu }} ( B({\mathbf {x}},t))}= \frac{{\varvec{\mu }} (B({\mathbf {x}},\frac{t}{2\sqrt{u_0}}))}{{\varvec{\mu }} (B({\mathbf {x}},t))}\gtrsim u_0^{-\mathbf N\slash 2}, \end{aligned}$$

(see (7.6)), we obtain (7.20).

We are now in a position to continue the proof of (7.19).

If \(\tilde{d}({\mathbf {x}},{\mathbf {y}})\le {\varvec{\mu }} (B({\mathbf {x}},t))\) then (7.19) follows from (7.20).

If \(\tilde{d}({\mathbf {x}},{\mathbf {y}})>{\varvec{\mu }} (B({\mathbf {x}},t))\) and \(\tilde{d}({\mathbf {y}},{\mathbf {y}}') <\tilde{d}({\mathbf {x}},{\mathbf {y}})\slash (2A)\), then \(\tilde{d}({\mathbf {x}},{\mathbf {y}})\le 2A\tilde{d} ({\mathbf {x}},{\mathbf {y}}')\). Hence, from (7.18) we conclude that

$$\begin{aligned} \bigl | U_t (\mathbf {x},\mathbf {y})-U_t(\mathbf {x},\mathbf {y}^{\prime })\bigr | \le \tfrac{C_2'}{{\varvec{\mu }} (B({\mathbf {x}}, t))} \bigl ( 1+\tfrac{\widetilde{d}(\mathbf {x},\mathbf {y})}{{\varvec{\mu }} (B({\mathbf {x}},t))} \bigr )^{-1-\delta } . \end{aligned}$$
(7.21)

Consequently, we deduce (7.19) (with perhaps small \(\delta '>0\)) from (7.20) and (7.21).

It remains to consider the case when \(\tilde{d}({\mathbf {x}},{\mathbf {y}}) > {\varvec{\mu }} (B({\mathbf {x}},t))\) and \(\tilde{d}({\mathbf {y}},{\mathbf {y}}') \ge \tilde{d}({\mathbf {x}},{\mathbf {y}})\slash (2A)\). Recall that \(\tilde{d}({\mathbf {y}},{\mathbf {y}}')\le {\varvec{\mu }} (B({\mathbf {x}},t))\). Thus \(\tilde{d}({\mathbf {x}},{\mathbf {y}})\sim {\varvec{\mu }} (B({\mathbf {x}},t))\). So, finally, using (7.20) we have

$$\begin{aligned}&\bigl | U_t (\mathbf {x},\mathbf {y})-U_t(\mathbf {x},\mathbf {y}^{\prime })\bigr | \le \tfrac{C_4}{{\varvec{\mu }} (B({\mathbf {x}}, t))} \bigl (\tfrac{\widetilde{d}(\mathbf {y},\mathbf {y}^{\prime })}{{\varvec{\mu }} ( B({\mathbf {x}},t))} \bigr )^{\frac{1}{\mathbf {N}}}\\&\quad \le \tfrac{C_4}{{\varvec{\mu }} (B({\mathbf {x}}, t))} \bigl (\tfrac{\widetilde{d}(\mathbf {y},\mathbf {y}^{\prime })}{{\varvec{\mu }} ( B({\mathbf {x}},t))} \bigr )^{\frac{1}{\mathbf {N}}}\bigl ( 1+\tfrac{\widetilde{d}(\mathbf {x},\mathbf {y})}{{\varvec{\mu }} (B({\mathbf {x}},t))} \bigr )^{-1-\delta }. \end{aligned}$$

This completes the proof of Lemma 7.3. \(\square \)

Set \(K_r({\mathbf {x}},{\mathbf {y}})=U_t({\mathbf {x}},{\mathbf {y}})\), where \(r={\varvec{\mu }} (B({\mathbf {x}},t))\). Now Theorem 2.1 follows from (7.15), boundedness of the maximal function \(W_*\) on \(L^1({\mathbb {R}}^n, {\varvec{\mu }})\), and the Uchiyama theorem (see Theorem 7.2) combined with Lemma 7.3.

Now we turn to prove (2.4). Recall that \(P_t({\mathbf {x}},{\mathbf {y}})>0\). So, by (7.7), the operator \(P_*\) is bounded on \(L^\infty ({\mathbb {R}}^n, d{\varvec{\mu }})\). Thanks to (7.18) and (7.16), it is of weak-type (1,1). Finally, from the Marcinkiewicz interpolation theorem we conclude that \(P_*\) is bounded on \(L^p({\mathbb {R}}^n,\, d{\varvec{\mu }})\) for \(1<p<\infty \). \(\square \)

Proof of Proposition 2.2

Fix \(\varepsilon >0\). There is \(R>0\) such that \(|g({\mathbf {x}})|<\varepsilon \) for \(|{\mathbf {x}}|>R\). Write

$$\begin{aligned} g=g\chi _{B(0,R)}+g\chi _{B(0,R)^c}=: g_1+ g_2. \end{aligned}$$

From (7.7) we get \(|P_tg_2({\mathbf {x}})|<\varepsilon \) for every \(t>0\) and \({\mathbf {x}}\in {\mathbb {R}}^n\). Now using (7.8) we obtain

$$\begin{aligned} | P_tg_1({\mathbf {x}})| \le \frac{C}{{\varvec{\mu }} (B({\mathbf {x}}, t))} \Vert g_1\Vert _{L^1({\mathbb {R}}^n, \, d{\varvec{\mu }})}\rightarrow 0 \text { as }\ t\rightarrow \infty . \end{aligned}$$

On the other hand, if t remains in a bounded interval and \(|{\mathbf {x}}|>2nR\), applying (7.9) we have

$$\begin{aligned} |P_tg_1({\mathbf {x}})| \le \frac{C}{{\varvec{\mu }}(B({\mathbf {x}}, t))} \Big (1+\frac{{\varvec{\mu }} (B({\mathbf {x}}, |{\mathbf {x}}|))}{{\varvec{\mu }}(B({\mathbf {x}}, t))}\Big )^{-1-\delta } \Vert g_1\Vert _{L^1({\mathbb {R}}^n, \, d{\varvec{\mu }})}\rightarrow 0 \text { as } |{\mathbf {x}}|\rightarrow \infty . \end{aligned}$$

The proof of the first part of Proposition 2.2 is complete.

In order to prove the second part of the proposition we fix \(\varepsilon >0\). We claim that

$$\begin{aligned} \lim _{|{\mathbf {x}}|\rightarrow \infty } P_\varepsilon f({\mathbf {x}}) =0. \end{aligned}$$

To prove the claim let \(\varepsilon ' >0\). Take \(R>0\) large enough such that \(\int _{|{\mathbf {y}}|>R} |f({\mathbf {y}})|\, d{\varvec{\mu }} ({\mathbf {y}})\le \varepsilon ' {\varvec{\mu }}(B(0, \varepsilon ))\). Write \(f=f\chi _{B(0,R)}+f\chi _{B(0,R)^c}=:f_1+f_2\). Then, by (7.7) and (7.8) we have \(|P_\varepsilon f_2|\le \varepsilon '\). On the other hand from the first part of the proposition we conclude that \(\lim _{|{\mathbf {x}}|\rightarrow \infty } P_\varepsilon f_1({\mathbf {x}})=0\), which gives the claim. Now (2.5) follows from the first part of Proposition 2.2, since \(P_{t+\varepsilon } f=P_t(P_\varepsilon f)\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Dziubański, J. Riesz Transforms Characterizations of Hardy Spaces \(H^1\) for the Rational Dunkl Setting and Multidimensional Bessel Operators. J Geom Anal 26, 2639–2663 (2016). https://doi.org/10.1007/s12220-015-9642-2

Download citation

  • Received:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s12220-015-9642-2

Keywords

Mathematics Subject Classification

Navigation