Skip to main content
Log in

\(\ell ^p\left( \mathbb {Z}^d\right) \)-estimates for discrete operators of Radon type: variational estimates

  • Published:
Inventiones mathematicae Aims and scope

Abstract

We prove \(\ell ^p\left( {\mathbb {Z}}^d\right) \) bounds for \(p\in (1, \infty )\), of r-variations \(r\in (2, \infty )\), for discrete averaging operators and truncated singular integrals of Radon type. We shall present a new powerful method which allows us to deal with these operators in a unified way and obtain the range of parameters of p and r which coincide with the ranges of their continuous counterparts.

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.

Similar content being viewed by others

References

  1. Bourgain, J.: Pointwise ergodic theorems for arithmetic sets. With an appendix by the author, Harry Furstenberg, Yitzhak Katznelson and Donald S. Ornstein. Publ. Math. Paris 69(1), 5–45 (1989)

    Article  MATH  Google Scholar 

  2. Carbery, A., Christ, M., Wright, J.: Multidimensional van der Corput and sublevel set estimates. J. Am. Math. Soc. 12(4), 981–1015 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  3. Christ, M.: A \({T}(b)\) theorem with remarks on analytic capacity and the Cauchy integral. Colloq. Math. 2(60–61), 601–628 (1990)

    MathSciNet  MATH  Google Scholar 

  4. Cotlar, M.: A unified theory of Hilbert transforms and ergodic theorems. Rev. Mat. Cuyana 1(2), 105–167 (1955)

    MathSciNet  Google Scholar 

  5. deLeeuw, K.: On \(L^p\) multipliers. Ann. Math. 81, 364–379 (1965)

    Article  MathSciNet  Google Scholar 

  6. Duoandikoetxea, J., Rubio de Francia, J.L.: Maximal and singular integral operators via Fourier transform estimates. Invent. Math. 84(3), 541–561 (1986)

    Article  MathSciNet  MATH  Google Scholar 

  7. Hughes, K., Krause, B., Trojan, B.: The maximal function and conditional square function control the variation: an elementary proof. Proc. Amer. Math. Soc. 144(8), 3583–3588 (2016)

  8. Ionescu, A.D., Wainger, S.: \(L^p\) boundedness of discrete singular Radon transforms. J. Am. Math. Soc. 19(2), 357–383 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  9. Jones, R.L., Kaufman, R., Rosenblatt, J.M., Wierdl, M.: Oscillation in ergodic theory. Ergod. Theory Dyn. Syst. 18(4), 889–935 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  10. Jones, R.L., Seeger, A., Wright, J.: Strong variational and jump inequalities in harmonic analysis. Trans. Am. Math. Soc. 6711–6742 (2008)

  11. Krause, B.: Polynomial ergodic averages converge rapidly: variations on a theorem of Bourgain. arXiv:1402.1803 (2014)

  12. Krause, B.: Some optimizations for (maximal) multipliers in \({L}^{p}\). arXiv:1402.1804 (2014)

  13. Lepingle, D.: La variation d’ordre \(p\) des semi-martingales. Z. Wahrscheinlichkeitstheorie Verw. Gebiete 36(4), 295–316 (1976)

    Article  MathSciNet  MATH  Google Scholar 

  14. Magyar, A., Stein, E.M., Wainger, S.: Discrete analogues in harmonic analysis: spherical averages. Ann. Math. 189–208 (2002)

  15. Mirek, M., Stein, E.M., Trojan, B.: \({\ell }^{p}\left({\mathbb{Z}}^d\right)\)-estimates for discrete operators of Radon type: Maximal functions and vector-valued estimates. arXiv:1512.07518 (2015)

  16. Mirek, M., Trojan, B.: Discrete maximal functions in higher dimensions and applications to ergodic theory. Am. J. Math. 138(6), 1495–1532 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  17. Nazarov, F., Oberlin, R., Thiele, Ch.: A Calderón Zygmund decomposition for multiple frequencies and an application to an extension of a lemma of Bourgain. Math. Res. Lett. 17(3), 529–545 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  18. Pisier, G., Xu, Q.: The strong \(p\)-variation of martingales and orthogonal series. Probab. Theory Relat. Fields 77(4), 497–514 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  19. Stein, E.M.: Harmonic Analysis: Real-Variable Methods, Orthogonality, and Oscillatory Integrals. Princeton University Press, Princeton (1993)

    MATH  Google Scholar 

  20. Stein, E.M., Wainger, S.: Discrete analogues in harmonic analysis, I: \(\ell ^2\) estimates for singular Radon transforms. Am. J. Math. 121(6), 1291–1336 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  21. Stein, E.M., Wainger, S.: Oscillatory integrals related to Carleson’s theorem. Math. Res. Lett. 8, 789–800 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  22. Zorin-Kranich, P.: Variation estimates for averages along primes and polynomials. J. Funct. Anal. 268(1), 210–238 (2015)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Mariusz Mirek.

Additional information

Mariusz Mirek was partially supported by NCN Grant DEC-2015/19/B/ST1/01149. Elias M. Stein was partially supported by NSF Grant DMS-1265524.

Appendix: Variational estimates for the continuous analogues

Appendix: Variational estimates for the continuous analogues

This section is intended to provide r-variational estimates for averaging and truncated singular operators of Radon type in the continuous settings. These kinds of questions were extensively discussed in [10], see also the references given there. Here we propose a different approach. Firstly, we will discuss long variation estimates. We give a new proof of Lépingle’s inequality which will be very much in spirit of good-\(\lambda \) inequalities. Secondly, we present a new approach to short variation estimates which is based on vector-valued bounds from [15]. This observation, as far as we know, has not been used in this context before. To fix notation let \(\mathcal {P}=\left( \mathcal {P}_1,\ldots , \mathcal {P}_{d_0}\right) : \mathbb {R}^k \rightarrow \mathbb {R}^{d_0}\) be a polynomial mapping whose components \(\mathcal {P}_j\) are real valued polynomials on \(\mathbb {R}^k\) such that \(\mathcal {P}_j(0) = 0\). One of the main objects of our interest will be

$$\begin{aligned} {\mathcal {M}}_t^{\mathcal {P}}f(x)=\frac{1}{|G_t|}\int _{G_t}f\left( x-\mathcal P(y)\right) {\, \mathrm{d}}y \end{aligned}$$

for \(x\in \mathbb {R}^{d_0}\) where G is an open bounded convex subset of \(\mathbb {R}^k\), containing the origin and

$$\begin{aligned} G_t = \left\{ x \in \mathbb {R}^k : t^{-1} x \in G \right\} . \end{aligned}$$

For any \(r\in [1, \infty )\) the r-variational seminorm \(V_r\) of complex-valued functions \((a_t(x): t>0)\) is defined by

$$\begin{aligned} V_r\left( a_t(x): t>0\right) =\sup _{0< t_0<\ldots <t_J} \left( \sum _{j=0}^J|a_{t_{j+1}}(x)-a_{t_j}(x)|^r\right) ^{1/r} \end{aligned}$$

where the supremum is taken over all finite increasing sequences. In order to avoid some problems with measureability of \(V_r\left( a_t(x): t>0\right) \) we assume that \((0, \infty )\ni t\mapsto a_t(x)\) is always a continuous function for every \(x\in \mathbb {R}^{d_0}\). The main result of this section is the following theorem.

Theorem 9.1

For every \(1< p < \infty \) and \(r\in (2, \infty )\) there is \(C_{p, r} > 0\) such that for all \(f \in L^p\left( \mathbb {R}^{d_0}\right) \)

$$\begin{aligned} \left||V_r\left( {\mathcal {M}}_t^\mathcal {P}f: t>0\right) \right||_{L^p}\le C_{p, r}\Vert f\Vert _{L^p}. \end{aligned}$$

Moreover, the constant \(C_{p, r}\le C_p\frac{r}{r-2}\) for some \(C_p>0\) which is independent of the coefficients of the polynomial mapping \(\mathcal {P}\).

Suppose that \(K\in {\mathcal {C}}^1\left( \mathbb {R}^k{\setminus }\{0\}\right) \) is a Calderón–Zygmund kernel satisfying the differential inequality

$$\begin{aligned} {\left|y \right|}^k {|{K(y)} |} + {\left|y \right|}^{k+1} {\left|\nabla K(y) \right|} \le 1 \end{aligned}$$
(9.1)

for all \(y \in \mathbb {R}^k{\setminus }\{0\}\) and the cancellation condition

$$\begin{aligned} \int _{G_t {\setminus } G_s}K(y){\, \mathrm{d}}y=0 \end{aligned}$$

for every \(t> s > 0\). We consider a truncated singular Radon transform defined by

$$\begin{aligned} {\mathcal {T}}_t^{\mathcal {P}}f(x)=\int _{G_t^c}f\left( x-{\mathcal {P}}(y)\right) K(y){\, \mathrm{d}}y \end{aligned}$$

for \(x\in \mathbb {R}^{d_0}\) and \(t>0\). The second main result is the following theorem.

Theorem 9.2

For every \(p \in (1, \infty )\) and \(r\in (2, \infty )\) there is \(C_{p, r} > 0\) such that for all \(f \in L^p\left( \mathbb {R}^{d_0}\right) \)

$$\begin{aligned} \left||V_r\left( {\mathcal {T}}_t^\mathcal {P}f: t>0\right) \right||_{L^p}\le C_{p, r}\Vert f\Vert _{L^p}. \end{aligned}$$
(9.2)

Moreover, the constant \(C_{p, r}\le C_p\frac{r}{r-2}\) for some \(C_p>0\) which is independent of the coefficients of the polynomial mapping \(\mathcal {P}\).

We immediately see that (9.2) remains true for the operator

$$\begin{aligned} \tilde{{\mathcal {T}}}_t^{\mathcal {P}}f(x)={\mathrm{p.v.}}\int _{G_t}f\left( x-{\mathcal {P}}(y)\right) K(y) {\, \mathrm{d}}y. \end{aligned}$$

Let

$$\begin{aligned} \deg \mathcal {P}= \max \left\{ \deg \mathcal {P}_j : 1 \le j \le d_0\right\} . \end{aligned}$$

It is convenient to work with the set

$$\begin{aligned} \Gamma = \left\{ \gamma \in \mathbb {Z}^k {\setminus }\{0\} : 0 < |\gamma | \le \deg \mathcal {P}\right\} \end{aligned}$$

with the lexicographic order. Then each \(\mathcal {P}_j\) can be expressed as

$$\begin{aligned} \mathcal {P}_j(x) = \sum _{\gamma \in \Gamma } c_j^\gamma x^\gamma \end{aligned}$$

for some \(c_j^\gamma \in \mathbb {R}\). Let us denote by d the cardinality of the set \(\Gamma \). We identify \(\mathbb {R}^d\) with the space of all vectors whose coordinates are labeled by multi-indices \(\gamma \in \Gamma \). Let A be the diagonal \(d \times d\) matrix such that

$$\begin{aligned} (A v)_\gamma = {|{\gamma } |} v_\gamma . \end{aligned}$$

For \(t > 0\) we set

$$\begin{aligned} t^{A}=\exp (A\log t) \end{aligned}$$

i.e. \(t^A x=(t^{|\gamma |}x_{\gamma }: \gamma \in \Gamma )\) for any \(x\in \mathbb {R}^d\). Next, we introduce the canonical polynomial mapping

$$\begin{aligned} \mathcal {Q}= \left( {\mathcal {Q}_\gamma }: {\gamma \in \Gamma }\right) : \mathbb {R}^k \rightarrow \mathbb {R}^d \end{aligned}$$

where \(\mathcal {Q}_\gamma (x) = x^\gamma \) and \(x^\gamma =x_1^{\gamma _1}\cdot \ldots \cdot x_k^{\gamma _k}\). The coefficients \(\left( {c_j^\gamma }: {\gamma \in \Gamma , j \in \{1, \ldots , d_0\}}\right) \) define a linear transformation \(L: \mathbb {R}^d \rightarrow \mathbb {R}^{d_0}\) such that \(L\mathcal {Q}= \mathcal {P}\). Indeed, it is enough to set

$$\begin{aligned} (L v)_j = \sum _{\gamma \in \Gamma } c_j^\gamma v_\gamma \end{aligned}$$

for each \(j \in \{1, \ldots , d_0\}\) and \(v \in \mathbb {R}^d\). Now, proceeding as in Lemma 2.3 we can reduce the matters (see also [5] or [19, p. 515]) to the canonical polynomial mapping. To simplify the notation we will write \({\mathcal {M}}_t={\mathcal {M}}_t^{\mathcal {Q}}\) and \({\mathcal {T}}_t={\mathcal {T}}_t^{\mathcal {Q}}\).

1.1 Long variations

In this subsection we give a new proof of Lépingle’s inequality. Since we will appeal to the results from [10] we are going to follow their notation and therefore we will work with a more general setup than it is necessary for our further purposes.

We consider a slightly more general dilation structure

$$\begin{aligned} t^A=\exp (A\log t) \end{aligned}$$

for any \(t>0\), where A is a \(d\times d\) matrix whose eigenvalues have positive real parts. We say that any regular quasi-norm \(\rho :\mathbb {R}^d\rightarrow [0, \infty )\) is homogeneous with respect to the dilations \((t^A: t>0)\) if \(\rho (t^Ax)=t\rho (x)\) for any \(x\in \mathbb {R}^d\) and \(t>0\). Here \(\mathbb {R}^d\) endowed with a quasi-norm \(\rho \) and the Lebesgue measure will be considered as a space of homogeneous type with the quasi-metric induced by \(\rho \).

In this setting let us recall Christ’s construction of dyadic cubes [3].

Lemma 9.1

([3]) There exists a collection of open sets \(\{Q_{\alpha }^k: k\in \mathbb {Z}\text { and } \alpha \in I_k\}\) and constants \(D>1\), \(\delta , \eta >0\) and \(0<C_1, C_2\) such that

  1. (i)

    \(\left| \mathbb {R}^d{\setminus }\bigcup _{\alpha \in I_k}Q_{\alpha }^k\right| =0\) for all \(k\in \mathbb {Z}\);

  2. (ii)

    if \(l\le k\) then either \(Q_{\beta }^l\subseteq Q_{\alpha }^k\) or \(Q_{\beta }^l\cap Q_{\alpha }^k = \emptyset \);

  3. (iii)

    for each \((l, \beta )\) and \(l\le k\), there exists a unique \(\alpha \) such that \(Q_{\beta }^l\subseteq Q_{\alpha }^k\);

  4. (iv)

    each \(Q_{\alpha }^k\) contains some ball \(B(z_{\alpha }^k, \delta D^k)\) and \({\text {diam}}(Q_{\alpha }^k)\le C_1 D^k\);

  5. (v)

    for each \((\alpha , k)\) and \(t>0\) we have \(|\{x\in Q_{\alpha }^k: {\mathrm{dist}}(x, \mathbb {R}^d{\setminus } Q_{\alpha }^k)\le tD^k\}|\le C_2 t^{\eta }|Q_{\alpha }^k|\).

Now two comments are in order. Firstly, each cube \(Q_{\alpha }^k\) contains a ball and is contained in some ball, each with radius \(\simeq D^k\). Secondly, the quasi-metric is translation invariant thus we see that for each \((\alpha , k)\) the measure of \(Q_{\alpha }^k\) is \(\simeq D^{{\mathrm{tr}}(A)k}\). In particular, there is \(R > 0\) such that for all \(k \in \mathbb {Z}\), \(\alpha \in I_k\) and \(\beta \in I_{k+1}\)

$$\begin{aligned} |Q_\alpha ^{k+1}| \le R |Q_\beta ^k|. \end{aligned}$$
(9.3)

The collection \(\{Q_{\alpha }^k: k\in \mathbb {Z}\text { and } \alpha \in I_k\}\) will be called the collection of dyadic cubes in \(\mathbb {R}^d\) adapted to the dilation group \((t^A: t>0)\). In view of Lemma 9.1, it gives rise to an atomic filtration. Namely, for each \(k\in \mathbb {Z}\) let \(\mathcal {F}_k=\sigma (\{Q_{\alpha }^l: \alpha \in I_l \text { and } l \ge -k\})\) be the \(\sigma \)-algebra generated by the cubes at level at least \(-k\). Then

$$\begin{aligned} \mathcal {F}_k \subset \mathcal {F}_{k+1}. \end{aligned}$$

For a localy integrable function f we set

$$\begin{aligned} \mathbb {E}_k f(x) = \mathbb {E}[f | \mathcal {F}_k] (x) = \frac{1}{|Q_{\alpha }^k|} \int _{Q_{\alpha }^k}f(y){\, \mathrm{d}}y \end{aligned}$$

provided \(Q_{\alpha }^k\) is the unique dyadic cube containing \(x\in \mathbb {R}^d\). Thanks to Lemma 9.1 (i), it is true for almost all x. We define a martingale difference by

$$\begin{aligned} \mathbb {D}_k f = \mathbb {E}_k f - \mathbb {E}_{k-1} f \end{aligned}$$

Finally, the maximal function and the square function are given by

$$\begin{aligned} Mf=\sup _{k\in \mathbb {Z}}|{\mathbb {E}}_kf| \quad \text { and } \quad Sf=\left( \sum _{k\in \mathbb {Z}}|{\mathbb {D}}_k f|^2\right) ^{1/2}, \end{aligned}$$

respectively.

The variational estimates for \(\left( \mathbb {E}_k f : k \in \mathbb {Z}\right) \) follow from estimates on \(\lambda \)-jump function \(J_{\lambda }\), see [1, 10, 18]. Recall, that for a sequence of complex numbers \((a_j: j\in \mathbb {Z})\) the function \(J_{\lambda }(a_j: j\in \mathbb {Z})\) is equal to the supremum over all \(J\in \mathbb {N}\) for which there exists a sequence of integers \(t_1<t_2<\ldots <t_J\) so that

$$\begin{aligned} |a_{t_{j+1}}-a_{t_j}|>\lambda \end{aligned}$$

for every \(j=1, 2,\ldots , N-1\). We immediately see that \(J_{\lambda }(a_j: j\in \mathbb {Z})\le \lambda ^{-r}V_r(a_j: j\in \mathbb {Z})^r\).

Theorem 9.3

([1, 18]) For each \(p \ge 1\) there exists \(B_p > 0\) such that for all \(f \in L^p\left( \mathbb {R}^d\right) \) and \(\lambda > 0\), if \(p\in (1, \infty )\)

$$\begin{aligned} \left\| \lambda \left( J_\lambda ({\mathbb {E}}_kf: k\in \mathbb {Z})\right) ^{1/2} \right\| _{L^p} \le B_p \Vert f\Vert _{L^p}, \end{aligned}$$

and if \(p=1\) then for any \(t > 0\) we have

$$\begin{aligned} \left| \left\{ x\in \mathbb {R}^d: \lambda \left( J_\lambda ({\mathbb {E}}_kf(x): k\in \mathbb {Z})\right) ^{1/2} > t \right\} \right| \le B_1 t^{-1} \Vert f\Vert _{L^1}. \end{aligned}$$

The next theorem is a new ingredient in proving Lépingle’s inequality and is inspired by [7].

Theorem 9.4

For each \(q \ge 2\) there is \(C_q > 0\) such that for all \(r > 2\) and \(\lambda > 0\)

$$\begin{aligned}&C_q \cdot \left| \left\{ x\in \mathbb {R}^d: V_r({\mathbb {E}}_kf(x): k\in \mathbb {Z})> \lambda \text { and } Mf(x) < \lambda /2\right\} \right| \nonumber \\&\quad \le \left| \left\{ x\in \mathbb {R}^d: Sf(x) > \lambda \right\} \right| + \lambda ^{-q} (r-2)^{-q/2} \int _{\{S(f) \le \lambda \}} Sf(x)^q {\, {\mathrm{d}}}x. \end{aligned}$$
(9.4)

Proof

By homogeneity, it suffices to prove the result with \(\lambda = 1\). Let \(B = \{x\in \mathbb {R}^d: Sf(x) > 1\}\), \(B^* = \{x\in \mathbb {R}^d: M {\mathbbm {1}_{{B}}}(x) > 1/(2R) \}\) and \(G = (B^*)^c\). By the maximal inequality, we have

$$\begin{aligned}&|B^*|=|\{x\in \mathbb {R}^d: M {\mathbbm {1}_{{B}}}(x)> 1/(2R) \}|\\&\quad \lesssim \int _{\mathbb {R}^d} {\mathbbm {1}_{{B}}}(x)^2 {\, \mathrm{d}}x = |\{x\in \mathbb {R}^d: Sf(x) > 1\}|. \end{aligned}$$

Therefore, it is enough to show that

$$\begin{aligned}&\left| \left\{ x\in G: V_r({\mathbb {E}}_kf(x): k\in \mathbb {Z}) > 1 \text { and } Mf(x) < 1/2\right\} \right| \\&\quad \lesssim (r-2)^{-q/2} \int _{B^c} Sf(x)^q {\, \mathrm{d}}x. \end{aligned}$$

We can pointwise dominate the variation (see [1])

$$\begin{aligned} V_r({\mathbb {E}}_kf: k\in \mathbb {Z})^r \le \sum _{l \in \mathbb {Z}} 2^{rl} J_{2^l}({\mathbb {E}}_kf: k\in \mathbb {Z}). \end{aligned}$$

Since \(Mf < 1/2\), the above sum runs over \(l \le 0\), which leads to the containment

$$\begin{aligned}&\left\{ x\in G: V_r({\mathbb {E}}_kf(x): k\in \mathbb {Z})> 1 \text { and } Mf(x) < 1/2 \right\} \nonumber \\&\quad \subseteq \left\{ x\in G: \sum _{l \le 0} 2^{rl} J_{2^l}({\mathbb {E}}_kf(x): k\in \mathbb {Z}) > 1 \right\} . \end{aligned}$$
(9.5)

For each \(n \in \mathbb {Z}\) we define \(U_n =\left\{ x\in \mathbb {R}^d: {\mathbb {E}}_n{\mathbbm {1}_{{B}}}(x) \le 1/2\right\} \). Notice that, if \(x \in G\) then \(x \in U_n\) for all \(n \in \mathbb {Z}\). Let

$$\begin{aligned} g(x) = \sum _{n \in \mathbb {Z}} {\mathbb {D}}_nf(x) \cdot {\mathbbm {1}_{{U_{n-1}}}}(x). \end{aligned}$$

We observe that \({\mathbb {E}}_ng(x) = {\mathbb {E}}_nf(x)\) for all \(x \in G\) and \(n \in \mathbb {Z}\). Indeed, \({\mathbb {D}}_nf \cdot {\mathbbm {1}_{{U_{n-1}}}}\) is \(\mathcal {F}_n\)-measurable and

$$\begin{aligned} {\mathbb {E}}_m\left( {\mathbb {D}}_nf\cdot {\mathbbm {1}_{{U_{n-1}}}}\right) = 0 \end{aligned}$$

for every \(m \le n-1\). Thus, for \(x \in G\) we have

$$\begin{aligned} {\mathbb {E}}_mg(x) = \sum _{n \le m} {\mathbb {D}}_nf(x) \cdot {\mathbbm {1}_{{U_{n-1}}}}(x) ={\mathbb {E}}_m f(x). \end{aligned}$$

Hence, by (9.5) and Hölder’s inequality with \(a=\frac{q}{2}\) and \(a'=\frac{q}{q-2}\), we obtain

$$\begin{aligned}&\left\{ x\in G: V_r({\mathbb {E}}_kf(x): k\in \mathbb {Z})> 1 \text { and } Mf(x) < 1/2 \right\} \\&\qquad \subseteq \left\{ x\in \mathbb {R}^d: \sum _{l \le 0}2^{rl} J_{2^l}({\mathbb {E}}_kg(x): k\in \mathbb {Z})> 1 \right\} \\&\qquad \qquad \subseteq \left\{ x\in G: \left( \sum _{l\le 0}2^{\frac{1}{2}l(r-2)\frac{q}{q-2}}\right) ^{\frac{q-2}{q}}\right. \\&\qquad \qquad \quad \left. \left( \sum _{l\le 0}2^{\frac{1}{2}l(r-2)\frac{q}{2}} \left( 2^{l}J_{2^l}({\mathbb {E}}_kg(x): k\in \mathbb {Z})^{1/2}\right) ^q\right) ^{\frac{2}{q}}>1\right\} . \end{aligned}$$

Define

$$\begin{aligned} A_{q, r}=\left( \sum _{l\le 0}2^{\frac{1}{2}l(r-2)\frac{q}{q-2}}\right) ^{\frac{q-2}{q}} ={\mathcal {O}}\left( (r-2)^{-\frac{q-2}{q}}\right) . \end{aligned}$$

Now Theorem 9.3 immediately leads to the majorization

$$\begin{aligned}&\left| \left\{ x\in G: V_r({\mathbb {E}}_kf(x): k\in \mathbb {Z})> 1 \text { and } Mf(x) < 1/2 \right\} \right| \\&\quad \le \left| \left\{ x\in G: \sum _{l\le 0}2^{\frac{1}{2}l(r-2)\frac{q}{2}}\left( 2^{l}J_{2^l}({\mathbb {E}}_kg(x): k\in \mathbb {Z})^{1/2}\right) ^q>A_{q, r}^{-\frac{q}{2}}\right\} \right| \\&\quad \lesssim A_{q, r}^{\frac{q}{2}} \sum _{l\le 0}2^{\frac{1}{2}l(r-2)\frac{q}{2}} \int _{\mathbb {R}^d} |g(x)|^q {\, \mathrm{d}}x\\&\quad \lesssim (r-2)^{1-q/2-1} \int _{\mathbb {R}^d} {|{g(x)} |}^q {\, \mathrm{d}}x\lesssim (r-2)^{-q/2} \int _{\mathbb {R}^d} {|{g(x)} |}^q {\, \mathrm{d}}x. \end{aligned}$$

Next, the square function S is bounded from below on \(L^q\left( \mathbb {R}^d\right) \), therefore

$$\begin{aligned} \int _{\mathbb {R}^d} {|{g(x)} |}^q {\, {\mathrm{d}}}x \lesssim \int _{\mathbb {R}^d} Sg(x)^q {\, {\mathrm{d}}}x = \int _{\mathbb {R}^d} \left( \sum _{k \in \mathbb {Z}} {|{{\mathbb {D}}_k f(x)} |}^2 \cdot {\mathbbm {1}_{{U_{k-1}}}}(x) \right) ^{q/2} {\, {\mathrm{d}}}x. \end{aligned}$$

Since for \(x \in U_{k-1}\), we have \({\mathbb {E}}_{k-1}{\mathbbm {1}_{{B}}}(x) \le 1/(2R)\), by the doubling property (9.3) we get \(\mathbb {E}_k{\mathbbm {1}_{{B}}}(x) \le 1/2\). Hence,

$$\begin{aligned} {\mathbbm {1}_{{U_{k-1}}}}(x) \le 2 \cdot \mathbb {E}_k{\mathbbm {1}_{{B^c}}}(x), \end{aligned}$$

and

$$\begin{aligned} \int _{\mathbb {R}^d} {|{g(x)} |}^q {\, {\mathrm{d}}}x \lesssim \int _{\mathbb {R}^d} \left( \sum _{k \in \mathbb {Z}} {|{{\mathbb {D}}_kf(x)} |}^2 \cdot {\mathbb {E}}_k{\mathbbm {1}_{{B^c}}}(x) \right) ^{q/2} {\, {\mathrm{d}}}x. \end{aligned}$$

We observe that for \(q = 2\) we have

$$\begin{aligned} \int _{\mathbb {R}^d}\sum _{k \in \mathbb {Z}} {|{{\mathbb {D}}_kf(x)} |}^2 \cdot {\mathbb {E}}_k{\mathbbm {1}_{{B^c}}}(x) {\, {\mathrm{d}}}x = \int _{B^c} Sf(x)^2 {\, {\mathrm{d}}}x. \end{aligned}$$

For \(q > 2\) let \(\tilde{q} = q/2 > 1\) and \(\tilde{q}'\) be its dual exponent. Then for \(h \in L^{\tilde{q}'}\left( \mathbb {R}^d\right) \)

$$\begin{aligned}&\int _{\mathbb {R}^d}\left( \sum _{k \in \mathbb {Z}} {|{{\mathbb {D}}_k f(x)} |}^2 \cdot {\mathbb {E}}_k {\mathbbm {1}_{{B^c}}}(x)\right) h(x) {\, {\mathrm{d}}}x\\&\quad = \sum _{k \in \mathbb {Z}}\int _{\mathbb {R}^d} {|{{\mathbb {D}}_k f(x)} |}^2 \cdot {\mathbb {E}}_k {\mathbbm {1}_{{B^c}}}(x)h(x) {\, {\mathrm{d}}}x \\&\quad = \sum _{k \in \mathbb {Z}} \int _{B^c} {|{{\mathbb {D}}_k f(x)} |}^2 \cdot {\mathbb {E}}_k h(x) {\, {\mathrm{d}}}x. \end{aligned}$$

Therefore, by Hölder’s inequality, we get

$$\begin{aligned} \int _{\mathbb {R}^d}\left( \sum _{k \in \mathbb {Z}} {|{{\mathbb {D}}_kf(x)} |}^2 \cdot {\mathbb {E}}_k{\mathbbm {1}_{{B^c}}}(x)\right) h(x) {\, {\mathrm{d}}}x \le \left( \int _{B^c} Sf(x)^q {\, {\mathrm{d}}} x \right) ^{2/q} ||M h ||_{\tilde{q}'}. \end{aligned}$$

Taking the supremum over all \(h \in L^{\tilde{q}'}\left( \mathbb {R}^d\right) \) we conclude

$$\begin{aligned} \int _{\mathbb {R}^d} \left( \sum _{k \in \mathbb {Z}} {|{{\mathbb {D}}_kf(x)} |}^2 \cdot {\mathbb {E}}_k{\mathbbm {1}_{{B^c}}}(x) \right) ^{q/2} {\, {\mathrm{d}}}x\lesssim \int _{B^c} Sf(x)^q {\, {\mathrm{d}}}x, \end{aligned}$$

which finishes the proof. \(\square \)

Now, using Theorem 9.1 we prove Lépingle’s inequality for the sequence \(({\mathbb {E}}_kf: k\in \mathbb {Z})\).

Theorem 9.5

([13]) For each \(p \in (1, \infty )\) there exists \(C_p > 0\) such that for all \(r\in (2, \infty )\) and \(f \in L^p\left( \mathbb {R}^d\right) \) we have

$$\begin{aligned} \left\| V_r({\mathbb {E}}_kf: k\in \mathbb {Z}) \right\| _{L^p} \le C_p\frac{r}{r-2} \Vert f\Vert _{L^p}, \end{aligned}$$

Moreover, \(V_r({\mathbb {E}}_kf: k\in \mathbb {Z})\) is also weak type (1, 1).

Proof

Given \(p > 1\) we take \(q=2p>2\). By (9.4), for \(f\in L^p\left( \mathbb {R}^d\right) \cap L^q\left( \mathbb {R}^d\right) \) we have

$$\begin{aligned}&C_q \cdot \left| \left\{ x \in \mathbb {R}^d : V_r(\mathbb {E}_k f : k \in \mathbb {Z})> \lambda \right\} \right| \le \left| \left\{ x \in \mathbb {R}^d : S f(x)> \lambda \right\} \right| \\&\quad + \left| \left\{ x \in \mathbb {R}^d : M f(x) > \lambda /2 \right\} \right| \\&\quad + \lambda ^{-q} (r-2)^{-q/2} \int _{\{Sf \le \lambda \}} S f(x)^q {\, {\mathrm{d}}} x. \end{aligned}$$

Therefore, we get

$$\begin{aligned} \left\| V_r(\mathbb {E}_kf: k\in \mathbb {Z})\right\| _{L^p}^p= & {} p\int _{0}^{\infty }\lambda ^{p-1}\left| \left\{ x\in \mathbb {R}^d: V_r(\mathbb {E}_kf(x): k\in \mathbb {Z})>\lambda \right\} \right| {\, {\mathrm{d}}}\lambda \\\lesssim & {} {\left||{Sf} \right||}_{L^p}^p + {\left||{Mf} \right||}_{L^p}^p\\&+ (r-2)^{-p}\int _{0}^{\infty }\lambda ^{p-q-1}\int _{\{S(f) \le \lambda \}} Sf(x)^q {\mathrm{d}}x{\, \mathrm{d}}\lambda \\\lesssim & {} \Vert f\Vert _{L^p}^p+(r-2)^{-p} \int Sf(x)^q\int _{Sf(x)}^{\infty }\lambda ^{p-q-1} {\mathrm{d}}\lambda {\, \mathrm{d}}x\\\lesssim & {} \Vert f\Vert _{L^p}^p+(r-2)^{-p}\Vert Sf\Vert _{L^p}^p\lesssim \left( 1+(r-2)^{-p}\right) \Vert f\Vert _{L^p}^p. \end{aligned}$$

For \(p=1\) it suffices to apply the Calderón–Zygmund decomposition and the desired claim follows. \(\square \)

Long variational bounds for \(\mathcal {M}_t\) follows the same line as in [10]. For every \(f\in L^p\left( \mathbb {R}^d\right) \) with \(p\in (1, \infty )\) we obtain

$$\begin{aligned}&\left||V_r\left( {\mathcal {M}}_{2^n} f: n\in \mathbb {Z}\right) \right||_{L^p}\le \left||V_r\left( {\mathbb {E}}_nf: n\in \mathbb {Z}\right) \right||_{L^p}\nonumber \\&\quad +\left\| \left( \sum _{n\in \mathbb {N}} \left| {\mathcal {M}}_{2^n} f- {\mathbb {E}}_nf\right| ^2\right) ^{1/2}\right\| _{L^p}. \end{aligned}$$
(9.6)

The first term in (9.6) is bounded by Theorem 9.5, whereas the square function can be estimated as in [10, Proof of Theorem 1.1]. In the next theorem we consider long variational estimates for \({\mathcal {T}}_{t}\).

Theorem 9.6

For every \(p \in (1, \infty )\) there is \(C_p > 0\) such that for all \(r\in (2, \infty )\) and \(f \in L^p\left( \mathbb {R}^{d}\right) \)

$$\begin{aligned} \left||V_r\left( {\mathcal {T}}_{2^n} f: n\in \mathbb {Z}\right) \right||_{L^p}\le C_p \frac{r}{r-2} {\left||{f} \right||}_{L^p}. \end{aligned}$$

Proof

Let \(\Phi \in {\mathcal {C}}^{\infty }\left( \mathbb {R}^d\right) \) with compact support and integral one. Then we have the following decomposition (see [6])

$$\begin{aligned} \sum _{j\ge n}{\mathcal {T}}_{2^j}f&=\Phi _{2^n}*\left( {\mathcal {T}}f-\sum _{j<n}\mu _{2^j}*f\right) +(\delta _0-\Phi _{2^n})*\sum _{j\ge n}\mu _{2^j}*f\nonumber \\&=\Phi _{2^n}*{\mathcal {T}}f-\left( \Phi _{2^n}*\sum _{j< n}\mu _{2^j}\right) *f+\sum _{j\ge 0}(\delta _0-\Phi _{2^n})*\mu _{2^{j+n}}*f, \end{aligned}$$
(9.7)

where

$$\begin{aligned} \mu _{2^j}*f(x)=\int _{G_{2^{j+1}}{\setminus } G_{2^j}}f\left( x-\mathcal {Q}(y)\right) K(y){\, \mathrm{d}}y. \end{aligned}$$

The variational estimates for the first term in (9.7) follows by [10, Theorem 1.1, Lemma 2.1] and the boundedness for the operator \({\mathcal {T}}\), see [6, 19]. For the second term we apply the Littlewood–Paley theory since for each \(n \in \mathbb {N}\)

$$\begin{aligned} \Phi _{2^n} * \sum _{j < n} \mu _{2^j} \end{aligned}$$

is a Schwartz function with integral zero. Thus, by (2.3),

$$\begin{aligned} V_r\left( \Phi _{2^n} * \sum _{j< n} \mu _{2^j}*f : n \in \mathbb {Z}\right) \lesssim \left( \sum _{n \in \mathbb {Z}} \left|\Phi _{2^n} * \sum _{j < n} \mu _{2^j} * f \right|^2\right) ^{1/2}. \end{aligned}$$

For the last term we have

$$\begin{aligned}&\left\| V_r\left( \sum _{j\ge 0}(\delta _0-\Phi _{2^n})*\mu _{2^{j+n}}*f: n\in \mathbb {Z}\right) \right\| _{L^p}\\&\quad \le \sum _{j\ge 0} \left\| \left( \sum _{n\in \mathbb {Z}} \left| (\delta _0-\Phi _{2^n})*\mu _{2^{j+n}}*f\right| ^2\right) ^{1/2}\right\| _{L^p} \end{aligned}$$

and due to the Littlewood–Paley theory one can show that there is \(C_p>0\) and \(\delta _p>0\) such that

$$\begin{aligned} \left\| \left( \sum _{n\in \mathbb {Z}}\left| (\delta _0-\Phi _{2^n})*\mu _{2^{j+n}} *f\right| ^2\right) ^{1/2}\right\| _{L^p} \le C_p2^{-\delta _p j}\Vert f\Vert _{L^p}. \end{aligned}$$

This completes the proof of Theorem 9.6. \(\square \)

1.2 Short variations for averaging operators

To deal with short variations we need a counterpart of Lemma 2.2.

Lemma 9.2

Let \(u < v\) be real numbers and \(a:[u, v]\rightarrow \mathbb {C}\) be a differentiable function. For any \(h \in \mathbb {N}\) and the sequence \(\left( {s_j}: {0 \le j \le h}\right) \) with \(s_j=u+h^{-1}(v-u)j\) we have for every \(r\in [1, \infty )\)

$$\begin{aligned} V_r\left( a(t): t\in [u, v)\right) \lesssim \left( \sum _{j=0}^h|a(s_j)|^r\right) ^{1/r} +\left( \sum _{j=0}^{h-1}\left( \int _{s_j}^{s_{j+1}}|a'(t)|{\, \mathrm{d}}t\right) ^r\right) ^{1/r}. \end{aligned}$$
(9.8)

Moreover, if \(p\ge r\) then

$$\begin{aligned}&\left( \sum _{j=0}^h|a(s_j)|^r\right) ^{1/r} +\left( \sum _{j=0}^{h-1}\left( \int _{s_j}^{s_{j+1}}|a'(t)|{\, \mathrm{d}}t\right) ^r\right) ^{1/r}\\&\quad \lesssim h^{1/r-1/p}\left( \sum _{j=0}^h|a(s_j)|^p\right) ^{1/p} +h^{1/r-1}(v-u)^{1-1/p}\left( \int _{u}^{v}|a'(t)|^p{\, \mathrm{d}}t\right) ^{1/p}. \end{aligned}$$

Proof

Fix \(h \in \mathbb {N}\) and consider the sequence \(\left( {s_j}: {0 \le j \le h}\right) \) such that \(s_j=u+h^{-1}(v-u)j\). Then

$$\begin{aligned} V_r\left( a(t): t\in [u, v)\right)&\lesssim \left( \sum _{j=0}^h|a(s_j)|^r\right) ^{1/r} +\left( \sum _{j=0}^{h-1}V_1\left( a(t): t\in [s_j, s_{j+1})\right) ^r\right) ^{1/r}\\&\lesssim \left( \sum _{j=0}^h|a(s_j)|^r\right) ^{1/r} +\left( \sum _{j=0}^{h-1}\left( \int _{s_j}^{s_{j+1}}|a'(t)|{\, \mathrm{d}}t\right) ^r\right) ^{1/r}. \end{aligned}$$

If \(p \ge r\), by Hölder’s inequality, we get

$$\begin{aligned}&\left( \sum _{j=0}^h|a(s_j)|^r\right) ^{1/r} +\left( \sum _{j=0}^{h-1}\left( \int _{s_j}^{s_{j+1}}|a'(t)|{\, \mathrm{d}}t\right) ^r\right) ^{1/r}\\&\quad \le h^{1/r - 1/p} \left( \sum _{j=0}^h|a(s_j)|^p\right) ^{1/p} \\&\quad +h^{1/r-1/p}\left( \sum _{j=0}^{h-1}\left( \int _{s_j}^{s_{j+1}}|a'(t)|{\, \mathrm{d}}t\right) ^p\right) ^{1/p}. \end{aligned}$$

For the second term we again use Hölder’s inequality to obtain

$$\begin{aligned}&h^{1/r-1/p}\left( \sum _{j=0}^{h-1}\left( \int _{s_j}^{s_{j+1}}|a'(t)|{\, \mathrm{d}}t\right) ^p\right) ^{1/p} \lesssim h^{1/r-1/p}\left( \sum _{j=0}^{h-1}(s_{j+1}\right. \\&\quad \left. -s_j)^{p(1-1/p)}\left( \int _{s_j}^{s_{j+1}}|a'(t)|^p{\, \mathrm{d}}t\right) \right) ^{1/p}\\&\quad \lesssim h^{1/r-1}(v-u)^{1-1/p}\left( \int _{u}^{v}|a'(t)|^p{\, \mathrm{d}}t\right) ^{1/p} \end{aligned}$$

where in the last estimate we have used \(s_{j+1}-s_j=(v-u)/h\). This completes the proof of the lemma. \(\square \)

Now the task is to prove that for every \(p\in (1, \infty )\) there are \(C_p>0\) such that for every \(f\in L^p\left( \mathbb {R}^d\right) \) we have

$$\begin{aligned} \left||\left( \sum _{n\in \mathbb {Z}} V_2\left( \left( {\mathcal {M}}_t-{\mathcal {M}}_{2^n}\right) f: t\in \left[ 2^n, 2^{n+1}\right) \right) ^2\right) ^{1/2} \right||_{L^p}\le C_p\Vert f\Vert _{L^p}. \end{aligned}$$
(9.9)

We may assume that f is a Schwartz function. Let \(S_j\) be a Littlewood–Paley projection \(\mathcal {F}(S_j g)(\xi )=\phi _j(\xi )\mathcal {F}g (\xi )\) associated with \(\left( \phi _j: j\in \mathbb {Z}\right) \) a smooth partition of unity of \(\mathbb {R}^d{\setminus }\{0\}\) such that for each \(j \in \mathbb {Z}\) we have \(0\le \phi _j\le 1\) and

$$\begin{aligned} {\text {supp}}\, \phi _j \subseteq \left\{ \xi \in \mathbb {R}^d: 2^{-1}< {|{2^{jA}\xi } |}_{\infty } < 2 \right\} \end{aligned}$$

and for \(\xi \in \mathbb {R}^d{\setminus }\{0\}\)

$$\begin{aligned} \sum _{j\in \mathbb {Z}}\phi _j(\xi )=1. \end{aligned}$$

We are going to prove that for every \(p\in (1, \infty )\) there are \(C_p>0\) and \(\delta _p>0\) such that for every \(j\in \mathbb {Z}\) we have

$$\begin{aligned} \left||\left( \sum _{n\in \mathbb {Z}} V_2\left( \left( {\mathcal {M}}_t-{\mathcal {M}}_{2^n}\right) S_{j+n}f: t\in \left[ 2^n, 2^{n+1}\right) \right) ^2\right) ^{1/2} \right||_{L^p}\le C_p2^{-\delta _p|j|}\Vert f\Vert _{L^p}. \end{aligned}$$
(9.10)

Applying (9.8) with \(h = 2^{\varepsilon {|{j} |}}\), \(u=2^n\) and \(v=2^{n+1}\) we obtain that

$$\begin{aligned}&\left||\left( \sum _{n\in \mathbb {Z}} V_2\left( \left( {\mathcal {M}}_t-{\mathcal {M}}_{2^n}\right) S_{j+n}f: t\in \left[ 2^n, 2^{n+1}\right) \right) ^2\right) ^{1/2} \right||_{L^p} \\&\quad \lesssim \left||\left( \sum _{n\in \mathbb {Z}}\sum _{l=0}^h \left| \left( {\mathcal {M}}_{s_l}-{\mathcal {M}}_{2^n}\right) S_{j+n}f\right| ^2\right) ^{1/2} \right||_{L^p} \\&\quad + \left||\left( \sum _{n\in \mathbb {Z}}\sum _{l=0}^{h-1}\left( \int _{s_l}^{s_{l+1}} \left| \frac{{\mathrm{d}}}{{\, \mathrm{d}}t}{\mathcal {M}}_{t} S_{j+n}f\right| {\, \mathrm{d}}t\right) ^2\right) ^{1/2} \right||_{L^p}=I_p^1+I_p^2. \end{aligned}$$

1.2.1 The estimates for \(I_p^1\)

First, using vector-valued estimates from [15, Theorem A.1] together with the Littlewood–Paley theory we get

$$\begin{aligned} \begin{aligned} I_p^1&\lesssim 2^{\varepsilon |j|/2} \left||\left( \sum _{n\in \mathbb {Z}} \sup _{t>0}\left| {\mathcal {M}}_{t}S_{j+n}f\right| ^2\right) ^{1/2} \right||_{L^p} \\&\lesssim 2^{\varepsilon |j|/2} \left\| \left( \sum _{n\in \mathbb {N}}|S_{j+n}f|^2\right) ^{1/2}\right\| _{L^p} \lesssim 2^{\varepsilon |j|/2}\Vert f\Vert _{L^p}. \end{aligned} \end{aligned}$$
(9.11)

Next, we are going to refine the estimate (9.11) for \(p=2\). Let \(m_t\) be the multiplier associated with the operator \({\mathcal {M}}_t\). By van der Corput’s lemma [21], for each \(t\in [2^n, 2^{n+1})\) we have

$$\begin{aligned} |m_t(\xi )-m_{2^n}(\xi )|\lesssim \min \left\{ 1, |2^{nA}\xi |_\infty , |2^{nA}\xi |_\infty ^{-1/d}\right\} . \end{aligned}$$

Therefore, by Plancherel’s theorem

$$\begin{aligned} I_2^1= & {} \left( \sum _{l=0}^h\int _{\mathbb {R}^d}\sum _{n\in \mathbb {Z}}\left| (m_{s_l}(\xi )-m_{2^n} (\xi ))\phi _{j+n}(\xi ){\mathcal {F}}f(\xi )\right| ^2{\, {\mathrm{d}}}\xi \right) ^{1/2}\nonumber \\\lesssim & {} 2^{-|j|/d+\varepsilon |j|/2}\left( \int _{\mathbb {R}^d}\sum _{n\in \mathbb {Z}}\left| \phi _{j+n}(\xi ){\mathcal {F}}f(\xi )\right| ^2{\mathrm{d}}\xi \right) ^{1/2} \lesssim 2^{-{|{j} |}/d + \varepsilon {|{j} |} /2}\Vert f\Vert _{L^2}.\nonumber \\ \end{aligned}$$
(9.12)

Interpolating (9.11) with (9.12) and choosing appropriate \(\varepsilon < 2/d\) we get

$$\begin{aligned} I_p^1\lesssim 2^{-\delta _p|j|}\Vert f\Vert _{L^p}. \end{aligned}$$

1.2.2 The estimates for \(I_p^2\)

Since G is an open bounded convex set containing the origin, with the aid of the spherical coordinates we may write

$$\begin{aligned} \mathcal {M}_t g(x) = \frac{1}{t^k {|{G} |}} \int _{S^{k-1}} \int _0^{r(\omega ) t} g\left( x - \mathcal {Q}(r \omega ) \right) r^{k-1} {\, {\mathrm{d}}} r {\, {\mathrm{d}}}\sigma (\omega ) \end{aligned}$$

where \(S^{k-1}\) is a unit sphere in \(\mathbb {R}^k\) and \(\sigma \) is the surface measure on \(S^{k-1}\). We observe that if g is a Schwartz function

$$\begin{aligned} \frac{{\mathrm{d}}}{{\mathrm{d}} t} \mathcal {M}_t g(x)&= -k \frac{1}{t^{k+1} {|{G} |}} \int _{S^{k-1}} \int _0^{r(\omega ) t} g\left( x - \mathcal {Q}(r \omega )\right) r^{k-1} {\, \mathrm{d}}r {\, \mathrm{d}}\sigma (\omega ) \nonumber \\&\quad + \frac{1}{t^k {|{G} |}} \int _{S^{k-1}} g\left( x - \mathcal {Q}(r(\omega ) t \omega )\right) r(\omega )^k t^{k-1} {\, \mathrm{d}}\sigma (\omega ). \end{aligned}$$
(9.13)

Change of the order of integration and differentiation is permited since g is bounded. Hence, if \(t \in [s_l, s_{l+1})\) and \(s_l, s_{l+1} \simeq 2^n\), by Tonnelli’s theorem, we get

$$\begin{aligned} \sum _{l=0}^{h-1}\int _{s_l}^{s_{l+1}} \left| \frac{{\mathrm{d}}}{{\mathrm{d}} t} \mathcal {M}_t g(x) \right| {\, \mathrm{d}} t&\lesssim \mathcal {M}_{2^{n+1}} {|{g} |}(x) + \frac{1}{2^{nk} |G|} \int _{2^n}^{2^{n+1}} \int _{S^{k-1}} \left| g\left( x \right. \right. \\&\quad \left. \left. - \mathcal {Q}(r(\omega ) t \omega )\right) \right| r(\omega )^k t^{k-1} {\, \mathrm{d}}\sigma (\omega ) {\, \mathrm{d}}t \\&\lesssim \mathcal {M}_{2^{n+1}} {|{g} |} (x). \end{aligned}$$

Therefore, we obtain

$$\begin{aligned} I_p^2\lesssim & {} \left||\left( \sum _{n\in \mathbb {Z}}\left( \mathcal {M}_{2^{n+1}} {|{S_{j+n}f} |} \right) ^2\right) ^{1/2} \right||_{L^p}\nonumber \\\lesssim & {} \left||\left( \sum _{n \in \mathbb {Z}}\sup _{t > 0} \left( \mathcal {M}_t {|{S_{j + n} f} |} \right) ^2 \right) ^{1/2} \right||_{L^p} \lesssim {\left||{f} \right||}_{L^p}, \end{aligned}$$
(9.14)

where the last inequality follows by the same line of reasoning as (9.11).

Next, we refine the estimates of \(I_p^2\) for \(p = 2\). Let \(\tilde{m}_{t}\) be the multiplier associated with the operator \(\frac{{\mathrm{d}}}{{\mathrm{d}}t}{\mathcal {M}}_{t}\). We have

$$\begin{aligned} \tilde{m}_t(\xi ) = - \frac{k}{t^{k+1} {|{G} |}} \int _{G_t} e^{2\pi i {\xi \cdot \mathcal {Q}(x)}} {\, \mathrm{d}}x + \frac{1}{t^k {|{G} |}} \int _{S^{k-1}} e^{2\pi i {\xi \cdot \mathcal {Q}(r(\omega ) t \omega )}} r(\omega )^k t^{k-1} {\, \mathrm{d}}\sigma (\omega ). \end{aligned}$$
(9.15)

Indeed, by (9.13), we have

$$\begin{aligned} \mathcal {F}\left( \frac{{\mathrm{d}}}{{\mathrm{d}} t} \mathcal {M}_t g\right) (\xi )&= -\frac{k}{t^{k+1} {|{G} |}} \int _{G_t} e^{2\pi i {\xi \cdot \mathcal {Q}(x)}} {\, \mathrm{d}} x \mathcal {F}g(\xi ) \\&\quad + \frac{1}{t^k {|{G} |}} \int _{\mathbb {R}^d} e^{2\pi i {\xi \cdot x}} \int _{S^{k-1}} g\left( x \right. \\&\quad \left. - \mathcal {Q}(r(\omega ) t \omega )\right) r(\omega )^k t^{k-1} {\, \mathrm{d}}\sigma (\omega ) {\, \mathrm{d}}x. \end{aligned}$$

Again, for the second term we need to justify the change of integrations. Let

$$\begin{aligned} R = \sup \left\{ {\left|\mathcal {Q}(y) \right|} : y \in G_t \right\} . \end{aligned}$$

Then for all \({\left|x \right|} \ge 2 R\) and \(|y|\le R\) we have

$$\begin{aligned} {\left|x - y \right|} \ge \frac{{\left|x \right|}}{2}, \end{aligned}$$

thus

$$\begin{aligned} \left| g\left( x - \mathcal {Q}(y)\right) \right| \lesssim \left( 1 + {\left|x \right|}\right) ^{-2d}. \end{aligned}$$
(9.16)

By Fubini’s theorem we get the claim. Moreover, we see that for \(t\simeq 2^n\) we obtain \(|\tilde{m}_{t}(\xi )|\lesssim 2^{-n}\).

Now, using (9.15), by the Cauchy–Schwarz inequality and Plancherel’s theorem we get

$$\begin{aligned} I_2^2\le & {} \left||\left( \sum _{n\in \mathbb {Z}}\frac{2^n}{h}\int _{2^n}^{2^{n+1}} \left| \frac{{\mathrm{d}}}{{\mathrm{d}}t}{\mathcal {M}}_{t} S_{j+n}f\right| ^2{\, \mathrm{d}}t \right) ^{1/2} \right||_{L^2} \nonumber \\= & {} \left( \sum _{n\in \mathbb {Z}}\frac{2^n}{h}\int _{2^n}^{2^{n+1}}\int _{\mathbb {R}^d} \left| \tilde{m}_{t}(\xi ) \phi _{j+n}(\xi ){\mathcal {F}}f(\xi )\right| ^2{\, \mathrm{d}}\xi {\, \mathrm{d}} t\right) ^{1/2}\nonumber \\\lesssim & {} 2^{-\varepsilon |j|/2} \left( \sum _{n\in \mathbb {Z}}\int _{\mathbb {R}^d} \left| \phi _{j+n}(\xi ){\mathcal {F}}f(\xi )\right| ^2{\, \mathrm{d}}\xi \right) ^{1/2} \lesssim 2^{-\varepsilon |j|/2}\Vert f\Vert _{L^2}\qquad \qquad \end{aligned}$$
(9.17)

for \(0<\varepsilon <1/d\). Thus interpolation of (9.14) with (9.17) gives

$$\begin{aligned} I_p^2\lesssim 2^{-\delta _p|j|}\Vert f\Vert _{L^p} \end{aligned}$$

and the proof of (9.10) is completed.

1.3 Short variations for truncated singular integral operators

We are going to show that for any \(p\in (1, \infty )\) there are \(C_p>0\) and \(\delta _0 > 0\) such that for every \(j \in \mathbb {Z}\) and \(f\in L^p\left( \mathbb {R}^d\right) \) we have

$$\begin{aligned} \left||\left( \sum _{n\in \mathbb {Z}} V_2\left( \left( {\mathcal {T}}_t-{\mathcal {T}}_{2^n}\right) S_{j +n} f: t\in \left[ 2^n, 2^{n+1}\right) \right) ^2 \right) ^{1/2} \right||_{L^p} \le C_p 2^{-\delta _p {|{j} |}} \Vert f\Vert _{L^p}. \end{aligned}$$
(9.18)

We may assume that f is a Schwartz function. The proof of (9.18) follows the same line as the one for the averaging operator. By Lemma 9.2, for \(h = 2^{\varepsilon {|{j} |}}\), \(u=2^n\) and \(v=2^{n+1}\) we obtain

$$\begin{aligned}&\left||\left( \sum _{n\in \mathbb {Z}} V_2\left( \left( {\mathcal {T}}_t-{\mathcal {T}}_{2^n}\right) S_{j+n}f: t\in \left[ 2^n, 2^{n+1}\right) \right) ^2\right) ^{1/2} \right||_{L^p}\\&\quad \lesssim \left||\left( \sum _{n\in \mathbb {Z}}\sum _{l=0}^h \left| \left( {\mathcal {T}}_{s_l}-{\mathcal {T}}_{2^n}\right) S_{j+n}f\right| ^2\right) ^{1/2} \right||_{L^p} \\&\quad + \left||\left( \sum _{n\in \mathbb {Z}}\sum _{l=0}^{h-1}\left( \int _{s_l}^{s_{l+1}} \left| \frac{{\mathrm{d}}}{{\, \mathrm{d}}t}\left( {\mathcal {T}}_{t}-{\mathcal {T}}_{2^n}\right) S_{j+n}f\right| {\, \mathrm{d}}t\right) ^2\right) ^{1/2} \right||_{L^p} =I_p^1+I_p^2. \end{aligned}$$

1.3.1 The estimates for \(I_p^1\)

By the vector-valued estimates from [15, Theorem A.1] and the Littlewood–Paley theory we get

$$\begin{aligned} \begin{aligned} I_p^1&\lesssim 2^{\varepsilon |j|/2} \left||\left( \sum _{n\in \mathbb {Z}} \sup _{t>0}\left( {\mathcal {M}}_{t}|S_{j+n}f|\right) ^2\right) ^{1/2} \right||_{L^p} \\&\lesssim 2^{\varepsilon |j|/2} \left\| \left( \sum _{n\in \mathbb {N}}|S_{j+n}f|^2\right) ^{1/2}\right\| _{L^p} \lesssim 2^{\varepsilon |j|/2}\Vert f\Vert _{L^p}. \end{aligned} \end{aligned}$$
(9.19)

For \(p = 2\) we get better estimate. Let \(m_{2^n, t}\) be the multiplier associated with the operator \({\mathcal {T}}_t - {\mathcal {T}}_{2^n}\) for \(t \in [2^n, 2^{n+1})\). By van der Corput’s lemma [21] (or more precisely the method of proof of van der Corput lemma from [21]) we have

$$\begin{aligned} |m_{t, 2^n}(\xi )| \lesssim \min \left\{ 1, |2^{nA}\xi |_\infty , |2^{nA}\xi |_\infty ^{-1/d}\right\} . \end{aligned}$$

Therefore, by Plancherel’s theorem

$$\begin{aligned} I_2^1= & {} \left( \sum _{l=0}^h\int _{\mathbb {R}^d}\sum _{n\in \mathbb {Z}}\left| (m_{s_l, 2^n}(\xi )\phi _{j+n}(\xi ){\mathcal {F}}f(\xi )\right| ^2{\, \mathrm{d}}\xi \right) ^{1/2}\nonumber \\\lesssim & {} 2^{-|j|/d+\varepsilon |j|/2} \left( \int _{\mathbb {R}^d}\sum _{n\in \mathbb {Z}}\left| \phi _{j+n}(\xi ) {\mathcal {F}}f(\xi )\right| ^2{\mathrm{d}}\xi \right) ^{1/2} \lesssim 2^{-{|{j} |}/d + \varepsilon {|{j} |} /2}\Vert f\Vert _{L^2}.\nonumber \\ \end{aligned}$$
(9.20)

Interpolating (9.19) with (9.20) and choosing appropriate \(\varepsilon < 2/d\) we get

$$\begin{aligned} I_p^1\lesssim 2^{-\delta _p|j|}\Vert f\Vert _{L^p}. \end{aligned}$$

1.3.2 The estimates for \(I_p^2\)

Since G is an open bounded convex set containing the origin, with the aid of the spherical coordinates we may write

$$\begin{aligned} \left( {\mathcal {T}}_t-{\mathcal {T}}_{2^n}\right) g(x) =- \int _{S^{k-1}} \int _{r(\omega ) 2^n}^{r(\omega ) t} g\left( x - \mathcal {Q}(r \omega ) \right) K(r \omega ) r^{k-1} {\, \mathrm{d}} r {\, \mathrm{d}}\sigma (\omega ) \end{aligned}$$

where \(S^{k-1}\) is a unit sphere in \(\mathbb {R}^k\) and \(\sigma \) is the surface measure on \(S^{k-1}\). We observe that if g is a Schwartz function

$$\begin{aligned} \frac{{\mathrm{d}}}{{\mathrm{d}} t} \left( {\mathcal {T}}_t-{\mathcal {T}}_{2^n} \right) g(x)&=- t^{k-1} \int _{S^{k-1}} g\left( x - \mathcal {Q}(r(\omega ) t \omega )\right) K(r(\omega ) t \omega ) r(\omega )^k {\, \mathrm{d}}\sigma (\omega ). \end{aligned}$$
(9.21)

Change of the order of integration and differentiation is permited since g is bounded and the kernel K satifies (9.1). Hence, if \(t \in [s_l, s_{l+1})\) and \(s_l, s_{l+1} \simeq 2^n\), by Tonnelli’s theorem, we get

$$\begin{aligned} \sum _{l=0}^{h-1}\int _{s_l}^{s_{l+1}} \left| \frac{{\mathrm{d}}}{{\mathrm{d}} t} \left( {\mathcal {T}}_t-{\mathcal {T}}_{2^n}\right) g(x) \right| {\, \mathrm{d}} t \lesssim \mathcal {M}_{2^{n+1}} {|{g} |}(x). \end{aligned}$$

Therefore, we obtain

$$\begin{aligned} I_p^2\lesssim & {} \left||\left( \sum _{n\in \mathbb {Z}}\left( \mathcal {M}_{2^{n+1}} {|{S_{j+n}f} |} \right) ^2\right) ^{1/2} \right||_{L^p}\nonumber \\\lesssim & {} \left||\left( \sum _{n \in \mathbb {Z}} \sup _{t > 0} \left( \mathcal {M}_t {|{S_{j + n} f} |} \right) ^2 \right) ^{1/2} \right||_{L^p} \lesssim {\left||{f} \right||}_{L^p}. \end{aligned}$$
(9.22)

Now we refine the estimate of \(I_p^2\) for \(p = 2\). Let \(\tilde{m}_{t, 2^n}\) be the multiplier associated with the operator \(\frac{{\mathrm{d}}}{{\mathrm{d}}t} \left( {\mathcal {T}}_t-{\mathcal {T}}_{2^n}\right) \). We have

$$\begin{aligned} \tilde{m}_{t, 2^n}(\xi ) = - t^{k-1} \int _{S^{k-1}} e^{2\pi i {\xi \cdot \mathcal {Q}(r(\omega ) t \omega )}} r(\omega )^k K(r(\omega ) t \omega ) {\, \mathrm{d}}\sigma (\omega ). \end{aligned}$$
(9.23)

Indeed, by (9.21), we have

$$\begin{aligned} \mathcal {F}\left( \frac{{\mathrm{d}}}{{\mathrm{d}} t} \left( {\mathcal {T}}_t-{\mathcal {T}}_{2^n}\right) g\right) (\xi )&=- t^{k-1} \int _{\mathbb {R}^d} e^{2\pi i {\xi \cdot x}} \int _{S^{k-1}} g\left( x \right. \\&\quad \left. - \mathcal {Q}(r(\omega ) t \omega )\right) r(\omega )^k K(r(\omega ) t \omega ) {\, \mathrm{d}}\sigma (\omega ) {\, \mathrm{d}}x \end{aligned}$$

and thanks to estimates (9.1) and (9.16) we may change the order of integrations. Note that if \(t\simeq 2^n\) we obtain \(|\tilde{m}_{t, 2^n}(\xi )|\lesssim 2^{-n}\).

Now, using (9.23), by the Cauchy–Schwarz inequality and Plancherel’s theorem we get

$$\begin{aligned} I_2^2\le & {} \left||\left( \sum _{n\in \mathbb {Z}}\frac{2^n}{h}\int _{2^n}^{2^{n+1}} \left| \frac{{\mathrm{d}}}{{\mathrm{d}}t}\left( \mathcal {T}_{t} - \mathcal {T}_{2^n}\right) S_{j+n}f\right| ^2{\, \mathrm{d}}t \right) ^{1/2} \right||_{L^2} \nonumber \\= & {} \left( \sum _{n\in \mathbb {Z}}\frac{2^n}{h}\int _{2^n}^{2^{n+1}}\int _{\mathbb {R}^d} \left| \tilde{m}_{t,2^n}(\xi ) \phi _{j+n}(\xi ){\mathcal {F}}f(\xi )\right| ^2{\, \mathrm{d}}\xi {\, \mathrm{d}} t \right) ^{1/2}\nonumber \\\lesssim & {} 2^{-\varepsilon |j|/2} \left( \sum _{n\in \mathbb {Z}}\int _{\mathbb {R}^d} \left| \phi _{j+n}(\xi ){\mathcal {F}}f(\xi )\right| ^2{\, \mathrm{d}}\xi \right) ^{1/2} \lesssim 2^{-\varepsilon |j|/2}\Vert f\Vert _{L^2}\nonumber \\ \end{aligned}$$
(9.24)

for \(0<\varepsilon <1/d\). Thus interpolation of (9.22) with (9.24) gives

$$\begin{aligned} I_p^2\lesssim 2^{-\delta _p|j|}\Vert f\Vert _{L^p} \end{aligned}$$

and the proof of (9.18) is completed.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mirek, M., Stein, E.M. & Trojan, B. \(\ell ^p\left( \mathbb {Z}^d\right) \)-estimates for discrete operators of Radon type: variational estimates . Invent. math. 209, 665–748 (2017). https://doi.org/10.1007/s00222-017-0718-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00222-017-0718-4

Navigation