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

Sharp estimates of noncommutative Bochner-Riesz means on two-dimensional quantum tori

Xudong Lai Xudong Lai: Institute for Advanced Study in Mathematics, Harbin Institute of Technology, Harbin, 150001, People’s Republic of China xudonglai@hit.edu.cn
Abstract.

In this paper, we establish the full LpL_{p} boundedness of noncommutative Bochner-Riesz means on two-dimensional quantum tori, which completely resolves an open problem raised in [7] in the sense of the LpL_{p} convergence for two dimensions. The main ingredients are sharp estimates of noncommutative Kakeya maximal functions and geometric estimates in the plane. We make the most of noncommutative theories of maximal/square functions, together with microlocal decompositions in both proofs of sharper estimates of Kakeya maximal functions and Bochner-Riesz means.

Key words and phrases:
Noncommutative LpL_{p} space, Bochner-Riesz means, Quantum tori, Kakeya maximal operator, Fourier series, Square function, Rotational algebras
2010 Mathematics Subject Classification:
Primary 46L52, 46L51, Secondary 46B15, 42B25
This work was supported by National Natural Science Foundation of China (No. 11801118, No. 12071098) and China Postdoctoral Science Foundation (No. 2017M621253, No. 2018T110279).

1. Introduction and main results

Inspired by operator algebras, harmonic analysis, noncommutative geometry and quantum probability, noncommutative harmonic analysis has rapidly developed recently (see e.g. [4, 7, 15, 16, 22, 27, 28, 35, 36, 38, 40, 60, 61]). The purpose of this paper is to study the noncommutative Bochner-Riesz means. We start by introducing the classical Bochner-Riesz means in Euclidean spaces.

It is well-known that the boundedness and convergence of Bochner-Riesz means are among the most important problems in harmonic analysis. The study of Bochner-Riesz means can also be regarded as making precise the sense in which the Fourier inversion formula holds. Recall that the Bochner-Riesz means on the usual torus 𝕋d\mathbb{T}^{d} are defined by

(1.1) BRλ(f)(x)=md(1|m|2R2)+λf^(m)e2πim,x{\text{\rm B}}_{R}^{\lambda}(f)(x)=\sum_{m\in{\mathbb{Z}}^{d}}(1-\frac{|m|^{2}}{R^{2}})^{\lambda}_{+}\widehat{f}(m)e^{2\pi i\langle m,x\rangle}

where λ0\lambda\geq 0, R>0R>0, (x)+=max{x,0}(x)_{+}=\max\{x,0\} and f^(m)=𝕋df(x)e2πim,x𝑑x\widehat{f}(m)=\int_{\mathbb{T}^{d}}f(x)e^{-2\pi i\langle m,x\rangle}dx. The central topic of Bochner-Riesz means is to seek the optimal range of λ\lambda such that BRλ(f){\text{\rm B}}_{R}^{\lambda}(f) converges to ff in some sense. In particular, the problem of the LpL_{p} convergence turns out to show (1||2R2)+λ:d(1-\frac{|\cdot|^{2}}{R^{2}})_{+}^{\lambda}:{\mathbb{Z}}^{d}\rightarrow{\mathbb{R}} is a uniform LpL_{p} Fourier multiplier in R>0R>0, which can be formulated as the so called Bochner-Riesz conjecture as follows (see e.g. [47]).

Conjecture.

Suppose λ>0\lambda>0 and 2dd+1+2λ<p<2dd12λ\frac{2d}{d+1+2\lambda}<p<\frac{2d}{d-1-2\lambda}. Then we have

supR>0BRλ(f)Lp(𝕋d)fLp(𝕋d).\sup_{R>0}\|{\text{\rm B}}_{R}^{\lambda}(f)\|_{L_{p}(\mathbb{T}^{d})}\lesssim\|f\|_{L_{p}(\mathbb{T}^{d})}.

One can also define the Bochner-Riesz means on d{\mathbb{R}}^{d} by

(1.2) BRλ(f)(x)=d(1|ξ|2R2)+λf^(ξ)e2πiξ,x𝑑ξ,B_{R}^{\lambda}(f)(x)=\int_{{\mathbb{R}}^{d}}\big{(}1-\tfrac{|\xi|^{2}}{R^{2}}\big{)}^{\lambda}_{+}\widehat{f}(\xi)e^{2\pi i\langle\xi,x\rangle}d\xi,

where f^\widehat{f} is the Fourier transform of ff on d{\mathbb{R}}^{d}. By the standard transference technique (see e.g. [17]), the uniform LpL_{p} boundedness of BRλ{\text{\rm B}}_{R}^{\lambda} in R>0R>0 on 𝕋d\mathbb{T}^{d} is equivalent to that of BRλB_{R}^{\lambda} on d{\mathbb{R}}^{d}. Because of this equivalence, in modern literature, researchers prefer to study the Bochner-Riesz means on d{\mathbb{R}}^{d}.

The study of Bochner-Riesz means originated from S. Bochner [1]. The necessity of the conditions of λ\lambda and pp in the Bochner-Riesz conjecture was given by C. Herz [20]. In dimension two, this conjecture has been completely resolved by L. Carleson and P. Sjölin [6], independently later by C. Fefferman [14], L. Hörmander [24] and A. Córdoba [10]. When dimension d3d\geq 3, the Bochner-Riesz conjecture is still open. We refer to some substantial progress in [3, 19, 31, 52, 53] and the references therein.

Concerning the pointwise convergence of BRλ(f)B_{R}^{\lambda}(f), it is natural to investigate the maximal Bochner-Riesz means defined by Bλ(f)(x)=supR>0|BRλ(f)(x)|B_{*}^{\lambda}(f)(x)=\sup_{R>0}|B_{R}^{\lambda}(f)(x)|. It is conjectured that BλB_{*}^{\lambda} is bounded on Lp(d)L_{p}({\mathbb{R}}^{d}) for λ>0\lambda>0 and 2d1d+2λ<p<2dd12λ\frac{2d-1}{d+2\lambda}<p<\frac{2d}{d-1-2\lambda} (see [51]), where the range of p2p\leq 2 is different from that of the Bochner-Riesz conjecture. It is clear that the study of BλB_{*}^{\lambda} is hander than that of BλB^{\lambda}. Up to now, this maximal Bochner-Riesz conjecture is even open for two dimensions. For some important progress, we refer the reader to [5, 33, 51] in the two-dimensional case and [31, 32, 46] for higher dimensions.

It should be pointed out that the study of Bochner-Riesz means is also quite related to several conjectures in harmonic analysis: Fourier restriction conjecture, local smoothing conjecture, maximal Kakeya function conjecture and Kakeya set conjecture (see e.g. [50]). To investigate the Bochner-Riesz conjecture, except some fundamental theories of maximal operators, Calderón-Zygmund operators, oscillatory integral operators, etc, researchers in harmonic analysis have invented many new and deep tools: bilinear or multilinear Fourier restriction (see [3, 31, 53]), incident geometry (see [2, 58]), decoupling and polynomial partitioning (see [19, 33]) in the last two decades. These new methods not only greatly improve the ranges of λ\lambda and pp in the study of Bochner-Riesz means, but also open new promising research directions in harmonic analysis.

On the other hand, many useful theories in harmonic analysis, such as Littlewood-Paley-Stein square functions, Hardy-Littlewood maximal operators, duality of H1H_{1}-BMO, Calderón-Zygmund operators and multiplier operators, have been successfully transferred to the noncommutative setting (see e.g. [4, 21, 23, 36, 37, 40, 62]). Motivated by the development of this noncommutative harmonic analysis, the noncommutative Bochner-Riesz means on quantum tori have been investigated partially by Z. Chen, Q. Xu and Z. Yin [7] with limited indexes of λ\lambda and pp. Due to the lack of commutativity, the study of noncommutative Bochner-Riesz means seems to be more challenging. For example, Z. Chen, Q. Xu and Z. Yin [7] established the boundedness of the maximal Bochner-Riesz means on the LpL_{p} space over quantum tori for λ>(d1)|121p|\lambda>(d-1)|\frac{1}{2}-\frac{1}{p}|, which is an analogue of a classical result by E. M. Stein [46], but with a much more technical proof. Compared with the fruitful theories of commutative Bochner-Riesz means, the noncommutative Bochner-Riesz means deserve to be investigated further. This leads to a natural question that whether we can transfer the modern and powerful tools mentioned before into the noncommutative setting and apply them to studying noncommutative Bochner-Riesz means. Since the study of two-dimensional Bochner-Riesz means is relatively simple (note the Bochner-Riesz conjecture is resolved in this case), in this paper we focus on two dimensions and our main purpose is to obtain the full boundedness of noncommutative Bochner-Riesz means on quantum tori by developing a new tool—the noncommutative Kakeya maximal function.

Quantum tori are also known as noncommutative tori or rotational algebras (see [44]). One can regard quantum tori as analogues of usual tori. Quantum tori are basic examples in operator algebras (see [11]) and are interesting objects in noncommutative geometry which have been extensively studied (see e.g. [8, 44, 54]). The research of analysis on quantum tori was started in [45, 56, 57] and the first systematic work of harmonic analysis on quantum tori was given later in [7]. For recent work related to quantum tori in the direction of noncommutative analysis, we refer to see [27, 29, 34, 43, 60, 61] and the references therein.

To illustrate our main results, we should give the definition of quantum torus. Suppose that d2d\geq 2, θ=(θk,j)1k,jd\theta=(\theta_{k,j})_{1\leq k,j\leq d} is a real skew symmetric d×dd\times d matrix. The dd-dimensional noncommutative torus 𝒜θ{\mathcal{A}}_{\theta} is a universal CC^{*}-algebra generated by dd unitary operators U1,,UdU_{1},\cdots,U_{d} satisfying the following commutation relation:

UkUj=e2πiθk,jUjUk,1k,jd.U_{k}U_{j}=e^{2\pi i\theta_{k,j}}U_{j}U_{k},\quad 1\leq k,j\leq d.

A well-known fact is that 𝒜θ{\mathcal{A}}_{\theta} admits a faithful tracial state τ\tau. Define 𝕋θd\mathbb{T}_{\theta}^{d} the weak *-closure of 𝒜θ{\mathcal{A}}_{\theta} in the GNS representation of τ\tau. We call 𝕋θd\mathbb{T}_{\theta}^{d} the dd-dimensional quantum torus. The state τ\tau also extends to a normal faithful state on 𝕋θd\mathbb{T}_{\theta}^{d}, which will be denoted again by τ\tau. Notice that when θ=0\theta=0, 𝒜θ=C(𝕋d){\mathcal{A}}_{\theta}=C(\mathbb{T}^{d}) and 𝕋θd=L(𝕋d)\mathbb{T}_{\theta}^{d}=L_{\infty}(\mathbb{T}^{d}). Thus quantum torus 𝕋θd\mathbb{T}_{\theta}^{d} is a deformation of classical torus 𝕋d\mathbb{T}^{d}. Let Lp(𝕋θd)L_{p}(\mathbb{T}_{\theta}^{d}) be the noncommutative space associated to pairs (𝕋θd,τ)(\mathbb{T}_{\theta}^{d},\tau) with the LpL_{p} norm given by xLp(𝕋θd)=(τ(|x|p))1/p\|x\|_{L_{p}(\mathbb{T}_{\theta}^{d})}=(\tau(|x|^{p}))^{1/p}.

In the following, we consider the Bochner-Riesz means on quantum tori which are defined by

(1.3) BRλ(f)=md(1|mR|2)+λf^(m)Um,fLp(𝕋θd),{\text{\rm B}}^{\lambda}_{R}(f)=\sum_{m\in{\mathbb{Z}}^{d}}\big{(}1-|\tfrac{m}{R}|^{2}\big{)}^{\lambda}_{+}\hat{f}(m)U^{m},\quad f\in L_{p}(\mathbb{T}^{d}_{\theta}),

where U=(U1,,Ud)U=(U_{1},\cdots,U_{d}), Um=U1m1UdmdU^{m}=U_{1}^{m_{1}}\cdots U_{d}^{m_{d}} and f^(m)=τ((Um)f)\widehat{f}(m)=\tau((U^{m})^{*}f). A fundamental problem raised in [7, Page 762] is that in which sense the Bochner-Riesz means converge back to ff. In this paper we consider this problem in two-dimensional case and state our main results as follows.

Theorem 1.1.

Suppose 0<λ<0<\lambda<\infty and 43+2λ<p<412λ\frac{4}{3+2\lambda}<p<\frac{4}{1-2\lambda}. Let BRλ{\text{\rm B}}_{R}^{\lambda} be the Bochner-Riesz means defined in (1.3) for d=2d=2. Then we have

supR>0BRλ(f)Lp(𝕋θ2)fLp(𝕋θ2).\sup_{R>0}\|{\text{\rm B}}_{R}^{\lambda}(f)\|_{L_{p}(\mathbb{T}_{\theta}^{2})}\lesssim\|f\|_{L_{p}(\mathbb{T}^{2}_{\theta})}.

Consequently for fLp(𝕋θ2)f\in L_{p}(\mathbb{T}_{\theta}^{2}), BRλ(f){\text{\rm B}}_{R}^{\lambda}(f) converges to ff in Lp(𝕋θ2)L_{p}(\mathbb{T}_{\theta}^{2}) as RR\rightarrow\infty.

This theorem is in fact a noncommutative version of the two-dimensional Bochner-Riesz conjecture. Thus we completely resolve an open problem raised in [7] in the sense of the LpL_{p} convergence for two dimensions. In the following, we briefly introduce the strategy used in the proof of Theorem 1.1.

Notice that BRλ(f){\text{\rm B}}_{R}^{\lambda}(f) is fully noncommutative and its analysis seems to be rather difficult. Nevertheless there is a clever trick that transfers the problem of multiplier operator on quantum tori to the operator-valued setting on usual tori (see [7]). Hence using this method, the LpL_{p} boundedness of Bochner-Riesz means on quantum tori can be reduced to that of the operator-valued Bochner-Riesz means on usual tori 𝕋d\mathbb{T}^{d} (see Theorem 5.1). Next we use the noncommutative transference of multiplier (see Theorem 5.4), we can transfer the study of the operator-valued Bochner-Riesz means on 𝕋d\mathbb{T}^{d} to that on d{\mathbb{R}}^{d}, in which case we can do analysis based on some known noncommutative theories of harmonic analysis. However to establish the full boundedness of two-dimensional operator-valued Bochner-Riesz means on 2{\mathbb{R}}^{2}, the previous noncommutative theories may not be sufficient and some new tools in harmonic analysis related to the geometry of Euclidean spaces should be brought in.

Our main new tool is the noncommutative Kakeya maximal function. Define the Kakeya average operator by

𝒦Rf(x)=|R|1Rf(xy)𝑑y,\mathcal{K}_{R}f(x)=|R|^{-1}\int_{R}f(x-y)dy,

where RR is a rectangle centered at the origin with arbitrary orientation and eccentricity NN (see Section 3 for its definition). As aforementioned, the study of the Kakeya maximal function supR|𝒦Rf(x)|\sup_{R}|\mathcal{K}_{R}f(x)| is another important problem in harmonic analysis related to Bochner-Riesz means (see e.g. [59]). Notice that the study of noncommutative Kakeya maximal functions is more difficult since it can not be defined directly. It is easy to see that 𝒦R\mathcal{K}_{R} could be dominated by the Hardy-Littlewood average operator with bound NN. This implies that the noncommutative Kakeya maximal operator is L2L_{2} bounded with norm N12N^{\frac{1}{2}}. In this paper, we shall establish its sharp L2L_{2} norm—logN\log N (see Theorem 3.1), which is crucial to our study of noncommutative Bochner-Riesz means. To the best knowledge of the author, it is the first time that a sharp estimate of the noncommutative Kakeya maximal function is obtained in noncommutative analysis. The proof here is quite technical and our strategy is the microlocal decomposition, together with theories of the Fourier transform and noncommutative square/maximal functions.

Below we sketch out the proof of the LpL_{p} boundedness of the operator-valued Bochner-Riesz means on 2{\mathbb{R}}^{2} (i.e. Theorem 4.1). To get the full boundedness of Bochner-Riesz means, by the duality and the noncommutative analytic interpolation theorem (see the appendix), it suffices to show the result for the case p=4p=4. We first make a dyadic decomposition: Bλ=kTkB^{\lambda}=\sum_{k}T_{k} and matters are reduced to proving the L4L_{4} norm of TkT_{k} has enough decay in kk. We next make a microlocal decomposition: Tk=lTk,lT_{k}=\sum_{l}T_{k,l} where the support of each Tk,lT_{k,l} lies in a small piece (denoted by Γk,l\Gamma_{k,l}) of annulus with the major direction e2πil2k2e^{2\pi il2^{-\frac{k}{2}}}. Notice that the L4L_{4} norm of Tk(f)T_{k}(f) has an expression

lTk,l(f)L4(𝒩)=llTk,l(f)Tk,l(f)L2(𝒩)12.\Big{\|}\sum_{l}T_{k,l}(f)\Big{\|}_{L_{4}({\mathcal{N}})}=\Big{\|}\sum_{l}\sum_{l^{\prime}}T_{k,l}(f)^{*}T_{k,l^{\prime}}(f)\Big{\|}_{L_{2}({\mathcal{N}})}^{\frac{1}{2}}.

If the major directions of these pieces are closed to each other (i.e. |ll|C|l-l^{\prime}|\leq C), then the above term is bounded by a column square function norm (l|Tk,l(f)|2)12L4(𝒩)\big{\|}\big{(}\sum_{l}|T_{k,l}(f)|^{2}\big{)}^{\frac{1}{2}}\big{\|}_{L_{4}({\mathcal{N}})}. If |ll|C|l-l^{\prime}|\geq C, then we need a very important geometric observation: {Γk,lΓk,l}l,l\{\Gamma_{k,l}-\Gamma_{k,l^{\prime}}\}_{l,l^{\prime}} is finite overlapped. With this geometric estimate, the L4L_{4} norm of TkfT_{k}f is bounded by a row square function norm (l|Tk,l(f)|2)12L4(𝒩)\big{\|}\big{(}\sum_{l}|T_{k,l}(f)^{*}|^{2}\big{)}^{\frac{1}{2}}\big{\|}_{L_{4}({\mathcal{N}})}. Consequently we use both column and row square function norms to control the L4L_{4} norm of Tk(f)T_{k}(f), which is consistent with the theory of noncommutative square functions. To estimate this column/row square functions, we should do more analysis for the kernel of Tk,lT_{k,l}. Roughly speaking, a key fact in our proof is that Tk,lT_{k,l} can be bounded by a sum of Kakeya average operators 𝒦R\mathcal{K}_{R}s where orientations of RRs are just in a fixed direction—the major direction of Γk,l\Gamma_{k,l}. By the dual theory between Lp(;)L_{p^{\prime}}({\mathcal{M}};\ell_{\infty}) and Lp(;1)L_{p}({\mathcal{M}};\ell_{1}), our estimates for noncommutative square functions can be reduced to that of Kakeya maximal functions, which will be systematically studied in Section 3.

The methods above heavily rely on the geometry of the plane. For the higher dimensional case, to get some nontrivial boundedness of Bochner-Riesz means, some more new tools in harmonic analysis should be transferred to the noncommutative setting. We hope to work this problem in the future.

This paper is organized as follows. First we give some preliminaries of noncommutative LpL_{p} spaces, noncommutative maximal/square functions and related lemmas in Section 2. In Section 3, we investigate the noncommutative Kakeya maximal function and establish its sharper estimate there. In Section 4, we obtain the full LpL_{p} boundedness of the operator-valued Bochner-Riesz means on 2{\mathbb{R}}^{2}. The proof is based on sharper estimates of noncommutative Kakeya maximal functions and a square function inequality studied in the previous sections. Section 5 is devoted to the study of Bochner-Riesz means on quantum tori. In this section, we first establish the full estimates of the operator-valued Bochner-Riesz means on usual tori 𝕋2\mathbb{T}^{2} and then transfer this result to that on two-dimensional quantum tori (i.e. Theorem 1.1). Finally for the reader’s convenience, we give a proof of the noncommutative analytic interpolation theorem in the appendix which may be known to experts.

Notation. Throughout this paper, the letter CC stands for a positive finite constant which is independent of the essential variables, not necessarily the same one in each occurrence. ABA\lesssim B means ACBA\leq CB for some constant CC. By the notation CεC_{\varepsilon} we mean that the constant depends on the parameter ε\varepsilon. ABA\approx B means that ABA\lesssim B and BAB\lesssim A. For any measurable set AdA\subset{\mathbb{R}}^{d}, we denote the Lebesgue measure by |A||A|. +{\mathbb{Z}}_{+} denotes the set of all nonnegative integers and +d=+××+{\mathbb{Z}}_{+}^{d}={{\mathbb{Z}}_{+}\times\cdots\times{\mathbb{Z}}_{+}} with dd-tuples product. For α+d\alpha\in{\mathbb{Z}}_{+}^{d} and xdx\in{\mathbb{R}}^{d}, xα=x1α1xdαdx^{\alpha}=x_{1}^{\alpha_{1}}\cdots x_{d}^{\alpha_{d}}. Set +=(0,){\mathbb{R}}_{+}=(0,\infty). s+\forall s\in{\mathbb{R}}_{+}, s\lfloor s\rfloor denotes the integer part of ss. We use LHS to represent left hand side of an expression. Given a function ff on 𝕋d\mathbb{T}^{d}, the Fourier transform of ff is defined by f^(k)=𝕋df(x)e2πixk𝑑x\widehat{f}(k)=\int_{\mathbb{T}^{d}}f(x)e^{-2\pi ix\cdot k}dx. For a function ff on d{\mathbb{R}}^{d}, define f\mathcal{F}f (or f^\hat{f}) and 1f\mathcal{F}^{-1}f (or fˇ\check{f}) the Fourier transform and the inversion Fourier transform of ff by

f(ξ)=de2πix,ξf(x)𝑑x,1f(ξ)=de2πix,ξf(x)𝑑x.\mathcal{F}f(\xi)=\int_{{\mathbb{R}}^{d}}e^{-2\pi i\langle x,\xi\rangle}f(x)dx,\ \ \ \ \mathcal{F}^{-1}f(\xi)=\int_{{\mathbb{R}}^{d}}e^{2\pi i\langle x,\xi\rangle}{f(x)dx}.

2. Preliminaries and some lemmas

In this section, we introduce some basic knowledge of noncommutative harmonic analysis including noncommutative LpL_{p} spaces, maximal functions, square functions and many operator-valued inequalities which are useful in this paper.

2.1. Noncommutative LpL_{p}-spaces

Let {\mathcal{M}} be a semifinite von Neumann algebra equipped with a normal semifinite faithful (n.s.f. in short) trace τ\tau. Denote by +{\mathcal{M}}_{+} the positive part of {\mathcal{M}} and let 𝒮+\mathcal{S_{+}} be the set of all x+x\in{\mathcal{M}}_{+} whose support projections have finite trace. Let 𝒮\mathcal{S} be the linear span of 𝒮+\mathcal{S_{+}}, then 𝒮\mathcal{S} is a weak * dense \ast-subalgebra of {\mathcal{M}}. Consider 0<p<0<p<\infty. For any x𝒮x\in\mathcal{S}, |x|p𝒮|x|^{p}\in\mathcal{S} and we set

xLp()=(τ(|x|p))1/p,x𝒮,\|x\|_{L_{p}({\mathcal{M}})}=\big{(}\tau(|x|^{p})\big{)}^{1/p},\ \ x\in\mathcal{S},

where |x|=(xx)12|x|=(x^{\ast}x)^{\frac{1}{2}} is the modulus of xx. Define the noncommutative LpL_{p} space associated with (,τ)({\mathcal{M}},\tau) by the completion of (𝒮,Lp())(\mathcal{S},\|\cdot\|_{L_{p}({\mathcal{M}})}) and set it as Lp()L_{p}({\mathcal{M}}). For convenience, if p=p=\infty, we define L()=L_{\infty}({\mathcal{M}})={\mathcal{M}} equipped with the operator norm \|\cdot\|_{{\mathcal{M}}}. Let Lp+()L_{p}^{+}({\mathcal{M}}) denote the positive part of Lp()L_{p}({\mathcal{M}}). A lot of basic properties of classical LpL_{p} spaces, such as Minkowski’s inequality, Hölder’s inequality, dual property, real and complex interpolation, have been transferred to this noncommutative setting. In particular, the following monotone properties are frequently used in this paper: for a,ba,b\in{\mathcal{M}} and α+\alpha\in{\mathbb{R}}_{+},

(2.1) aLp()bLp(), if 0ab;\|a\|_{L_{p}({\mathcal{M}})}\leq\|b\|_{L_{p}({\mathcal{M}})},\text{ if }0\leq a\leq b;
(2.2) aαbα, if 0ab and 0<α<1.a^{\alpha}\leq b^{\alpha},\text{ if $0\leq a\leq b$ and $0<\alpha<1$.}

For more about noncommutative LpL_{p} spaces, we refer to the very detailed introduction in the survey article [42] or the book [62].

In this paper, we are interested in the noncommutative LpL_{p} space on the tensor von Neumann algebra 𝒩=L(d)¯{\mathcal{N}}=L_{\infty}({\mathbb{R}}^{d})\overline{\otimes}{\mathcal{M}}. Set the tensor trace φ=d𝑑xτ\varphi=\int_{{\mathbb{R}}^{d}}dx\otimes\tau. Define the noncommutative space Lp(𝒩)L_{p}({\mathcal{N}}) associated with pairs (𝒩,φ)({\mathcal{N}},\varphi). Notice that Lp(𝒩)L_{p}({\mathcal{N}}) is isometric to Lp(d;Lp())L_{p}({\mathbb{R}}^{d};L_{p}({\mathcal{M}})) the Bochner LpL_{p} space on d{\mathbb{R}}^{d} with values in Lp()L_{p}({\mathcal{M}}).

2.2. Noncommutative maximal functions

It is difficult to define a noncommutative maximal function straightforwardly since two general elements in a von Neumann algebra may not be comparable. This obstacle can be overcome by defining the maximal norm directly. We adopt the definition of the noncommutative maximal norm introduced by G. Pisier [41] and M. Junge [26].

Definition 2.1 (Lp(;)L_{p}({\mathcal{M}};\ell_{\infty})).

We define Lp(;)L_{p}({\mathcal{M}};\ell_{\infty}) the space of all sequences x={xn}nx=\{x_{n}\}_{n\in{\mathbb{Z}}} in Lp()L_{p}({\mathcal{M}}) which admits a factorization of the following form: there exist a,bL2p()a,b\in L_{2p}({\mathcal{M}}) and a bounded sequence y={yn}ny=\{y_{n}\}_{n\in{\mathbb{Z}}} in L()L_{\infty}({\mathcal{M}}) such that xn=aynbx_{n}=ay_{n}b, n\forall\;n\in{\mathbb{Z}}. The norm of xx in Lp(;)L_{p}({\mathcal{M}};\ell_{\infty}) is given by

xLp(;)=inf{aL2p()supnynL()bL2p()},\|x\|_{L_{p}({\mathcal{M}};\ell_{\infty})}=\inf\big{\{}\|a\|_{L_{2p}({\mathcal{M}})}\,\sup_{n\in{\mathbb{Z}}}\|y_{n}\|_{L_{\infty}({\mathcal{M}})}\,\|b\|_{L_{2p}({\mathcal{M}})}\big{\}},

where the infimum is taken over all factorizations of xx as above.

If x={xn}nx=\{x_{n}\}_{n\in{\mathbb{Z}}} is a sequence of positive elements, then xLp(;)x\in L_{p}({\mathcal{M}};\ell_{\infty}) if and only if there exists a positive element aLp()a\in L_{p}({\mathcal{M}}) such that 0<xna0<x_{n}\leq a, and

(2.3) xLp(;)=inf{aLp(): 0<xna,n}.\|x\|_{L_{p}({\mathcal{M}};\ell_{\infty})}=\inf\{\|a\|_{L_{p}({\mathcal{M}})}:\ 0<x_{n}\leq a,\forall n\in{\mathbb{Z}}\}.

Similarly if x={xn}nx=\{x_{n}\}_{n\in{\mathbb{Z}}} is a sequence of self-adjoint elements, then xLp(;)x\in L_{p}({\mathcal{M}};\ell_{\infty}) if and only if there exists a positive element aLp()a\in L_{p}({\mathcal{M}}) such that axna-a\leq x_{n}\leq a, and

(2.4) xLp(;)=inf{aLp():axna,n}.\|x\|_{L_{p}({\mathcal{M}};\ell_{\infty})}=\inf\{\|a\|_{L_{p}({\mathcal{M}})}:\ -a\leq x_{n}\leq a,\forall n\in{\mathbb{Z}}\}.

More generally, if Λ\Lambda is an index set, we define Lp(;(Λ))L_{p}({\mathcal{M}};\ell_{\infty}(\Lambda)) as the space of all x={xλ}λΛx=\{x_{\lambda}\}_{\lambda\in\Lambda} in Lp()L_{p}({\mathcal{M}}) that can be factorized as

xλ=ayλbwitha,bL2p(),yλL(),supλyλL()<.x_{\lambda}=ay_{\lambda}b\quad\mbox{with}\quad a,b\in L_{2p}({\mathcal{M}}),\;y_{\lambda}\in L_{\infty}({\mathcal{M}}),\;\sup_{\lambda}\|y_{\lambda}\|_{L_{\infty}({\mathcal{M}})}<\infty.

Then the norm of Lp(;(Λ))L_{p}({\mathcal{M}};\ell_{\infty}(\Lambda)) is defined by

xLp(;(Λ))=infxλ=ayλb{aL2p()supλΛyλL()bL2p()}.\|x\|_{L_{p}({\mathcal{M}};\ell_{\infty}(\Lambda))}=\inf_{x_{\lambda}=ay_{\lambda}b}\big{\{}\|a\|_{L_{2p}({\mathcal{M}})}\,\sup_{{\lambda}\in\Lambda}\|y_{\lambda}\|_{L_{\infty}({\mathcal{M}})}\,\|b\|_{L_{2p}({\mathcal{M}})}\big{\}}.

It was shown in [30] that xLp(;(Λ))x\in L_{p}({\mathcal{M}};\ell_{\infty}(\Lambda)) if and only if sup{xLp(;(J)):JΛ,J is finite}<\sup\big{\{}\|x\|_{L_{p}({\mathcal{M}};\ell_{\infty}(J))}\;:\;J\subset\Lambda,\;J\textrm{ is finite}\big{\}}<\infty and moreover in this case, the norm xLp(;(Λ))\|x\|_{L_{p}({\mathcal{M}};\ell_{\infty}(\Lambda))} is equal to the above supremum.

If x={xλ}λΛx=\{x_{\lambda}\}_{\lambda\in\Lambda} is positive (resp. self-adjoint), xLp(;(Λ))\|x\|_{L_{p}({\mathcal{M}};\ell(\Lambda))} has the similar property of (2.3) (resp. (2.4)).

We will often use supλΛxλLp()\|\sup\limits_{\lambda\in\Lambda}x_{\lambda}\|_{L_{p}({\mathcal{M}})} to represent xLp(;(Λ))\|x\|_{L_{p}({\mathcal{M}};\ell(\Lambda))}. However we point out that supλΛxλLp()\|\sup\limits_{\lambda\in\Lambda}x_{\lambda}\|_{L_{p}({\mathcal{M}})} is just a notation since supλΛxλ\sup\limits_{\lambda\in\Lambda}x_{\lambda} makes no sense in the noncommutative setting.

To study the dual property of the above spaces Lp(;)L_{p}({\mathcal{M}};\ell_{\infty}), we need to introduce another space.

Definition 2.2 (Lp(;1)L_{p}({\mathcal{M}};\ell_{1})).

Define Lp(;1)L_{p}({\mathcal{M}};\ell_{1}) as the space of all sequences {yn}\{y_{n}\} in Lp()L_{p}({\mathcal{M}}) which could be factorized as

yn=kuk,nvk,n,n,y_{n}=\sum_{k}u^{*}_{k,n}v_{k,n},\quad\forall n\in{\mathbb{Z}},

for two families {uk,n}k,n\{u_{k,n}\}_{k,n\in{\mathbb{Z}}} and {vk,n}k,n\{v_{k,n}\}_{k,n\in{\mathbb{Z}}} in L2p()L_{2p}({\mathcal{M}}) such that k,nuk,nuk,nLp()\sum_{k,n\in{\mathbb{Z}}}u^{*}_{k,n}u_{k,n}\in L_{p}({\mathcal{M}}) and k,nvk,nvk,nLp()\sum_{k,n\in{\mathbb{Z}}}v^{*}_{k,n}v_{k,n}\in L_{p}({\mathcal{M}}). Lp(;1)L_{p}({\mathcal{M}};\ell_{1}) is equipped with the norm

{yn}Lp(;1)=infk,nuk,nuk,nLp()12k,nvk,nvk,nLp()12,\|\{y_{n}\}\|_{L_{p}({\mathcal{M}};\ell_{1})}=\inf\Big{\|}\sum_{k,n\in{\mathbb{Z}}}u^{*}_{k,n}u_{k,n}\Big{\|}_{L_{p}({\mathcal{M}})}^{\frac{1}{2}}\Big{\|}\sum_{k,n\in{\mathbb{Z}}}v^{*}_{k,n}v_{k,n}\Big{\|}_{L_{p}({\mathcal{M}})}^{\frac{1}{2}},

where the infimun is taken over all decompositions of {yn}\{y_{n}\} as above.

It is not difficult to see that if yn0y_{n}\geq 0 for all nn\in{\mathbb{Z}}, {yn}Lp(;1)\{y_{n}\}\in L_{p}({\mathcal{M}};\ell_{1}) if and only if nynLp()\sum_{n\in{\mathbb{Z}}}y_{n}\in L_{p}({\mathcal{M}}) (see e.g. [62]). In such a case, we have the following equality

{yn}Lp(;1)=nynLp().\|\{y_{n}\}\|_{L_{p}({\mathcal{M}};\ell_{1})}=\Big{\|}\sum_{n\in{\mathbb{Z}}}y_{n}\Big{\|}_{L_{p}({\mathcal{M}})}.

We introduce the following basic duality theorem of Lp(;1)L_{p}({\mathcal{M}};\ell_{1}), which has been established by M. Junge and Q. Xu in [30].

Lemma 2.3.

(i). Suppose 1p<1\leq p<\infty. Let pp^{\prime} be the conjugate index: 1p+1p=1\frac{1}{p}+\frac{1}{p^{\prime}}=1. Then the dual space of Lp(;1)L_{p}({\mathcal{M}};\ell_{1}) is Lp(;)L_{p^{\prime}}({\mathcal{M}};\ell_{\infty}). The element x={xn}Lp(;)x=\{x_{n}\}\in L_{p^{\prime}}({\mathcal{M}};\ell_{\infty}) acts on Lp(;1)L_{p}({\mathcal{M}};\ell_{1}) as follows

x,y=nτ(xnyn),y={yn}Lp(;1).\langle x,y\rangle=\sum_{n\in{\mathbb{Z}}}\tau(x_{n}y_{n}),\quad\forall y=\{y_{n}\}\in L_{p}({\mathcal{M}};\ell_{1}).

(ii). Suppose 1p1\leq p\leq\infty. For any xLp(;)x\in L_{p}({\mathcal{M}};\ell_{\infty}), we have

{xn}Lp(;)=sup{nτ(xnyn):y={yn}Lp(;1)andyLp(;1)1}.\|\{x_{n}\}\|_{L_{p}({\mathcal{M}};\ell_{\infty})}=\sup\Big{\{}\sum_{n}\tau(x_{n}y_{n}):y=\{y_{n}\}\in L_{p^{\prime}}({\mathcal{M}};\ell_{1})\ \text{and}\ \|y\|_{L_{p^{\prime}}({\mathcal{M}};\ell_{1})}\leq 1\Big{\}}.

Moreover if xx is positive, then

{xn}Lp(;)=sup{nτ(xnyn):ynLp+()andnynLp()1}.\|\{x_{n}\}\|_{L_{p}({\mathcal{M}};\ell_{\infty})}=\sup\Big{\{}\sum_{n}\tau(x_{n}y_{n}):y_{n}\in L^{+}_{p^{\prime}}({\mathcal{M}})\ \text{and}\ \|\sum_{n}y_{n}\|_{L_{p^{\prime}}({\mathcal{M}})}\leq 1\Big{\}}.

2.3. Noncommutative square functions

To define the noncommutative square function, we should first introduce the so-called column and row function spaces. Let f={fj}f=\{f_{j}\} be a finite sequence in Lp()L_{p}({\mathcal{M}}) where 1p1\leq p\leq\infty. Define

{fj}Lp(;2r)=(|fj|2)12Lp(),{fj}Lp(;2c)=(|fj|2)12Lp().\|\{f_{j}\}\|_{L_{p}({\mathcal{M}};\ell_{2}^{r})}=\big{\|}(\sum|f^{\ast}_{j}|^{2})^{\frac{1}{2}}\big{\|}_{L_{p}({\mathcal{M}})},\ \|\{f_{j}\}\|_{L_{p}({\mathcal{M}};\ell_{2}^{c})}=\big{\|}(\sum|f_{j}|^{2})^{\frac{1}{2}}\big{\|}_{L_{p}({\mathcal{M}})}.
Definition 2.4 (Lp(;2rc)L_{p}({\mathcal{M}};\ell_{2}^{rc})).

We define the spaces Lp(;2rc)L_{p}({\mathcal{M}};\ell_{2}^{rc}) as follows:

  1. (i).

    If p2p\geq 2, Lp(;2rc)=Lp(;2c)Lp(;2r)L_{p}({\mathcal{M}};\ell_{2}^{rc})=L_{p}({\mathcal{M}};\ell_{2}^{c})\cap L_{p}({\mathcal{M}};\ell_{2}^{r}) equipped with the intersection norm:

    {fj}Lp(;2rc)=max{{fj}Lp(;2c),{fj}Lp(;2r)}.\|\{f_{j}\}\|_{L_{p}({\mathcal{M}};\ell_{2}^{rc})}=\max\{\|\{f_{j}\}\|_{L_{p}({\mathcal{M}};\ell_{2}^{c})},\|\{f_{j}\}\|_{L_{p}({\mathcal{M}};\ell_{2}^{r})}\}.
  2. (ii).

    If p<2p<2, Lp(;2rc)=Lp(;2c)+Lp(;2r)L_{p}({\mathcal{M}};\ell_{2}^{rc})=L_{p}({\mathcal{M}};\ell_{2}^{c})+L_{p}({\mathcal{M}};\ell_{2}^{r}) equipped with the sum norm:

    {fj}Lp(;2rc)=inf{{gj}Lp(;2c)+{hj}Lp(;2r)},\|\{f_{j}\}\|_{L_{p}({\mathcal{M}};\ell_{2}^{rc})}=\inf\{\|\{g_{j}\}\|_{L_{p}({\mathcal{M}};\ell_{2}^{c})}+\|\{h_{j}\}\|_{L_{p}({\mathcal{M}};\ell_{2}^{r})}\},

    where the infimun is taken over all decompositions fj=gj+hjf_{j}=g_{j}+h_{j} with gjg_{j} and hjh_{j} in Lp()L_{p}({\mathcal{M}}).

It is easy to see that L2(;2r)=L2(;2c)=L2(;2rc)L_{2}({\mathcal{M}};\ell_{2}^{r})=L_{2}({\mathcal{M}};\ell_{2}^{c})=L_{2}({\mathcal{M}};\ell_{2}^{rc}). Next we introduce some inequalities for Lp(;2rc)L_{p}({\mathcal{M}};\ell_{2}^{rc}). The first one is Hölder type inequality whose proof can be found in [62].

Lemma 2.5.

Let 0<p,q,r0<p,q,r\leq\infty be such that 1/r=1/p+1/q1/r=1/p+1/q. Then for any fLp(;2c)f\in L_{p}({\mathcal{M}};\ell^{c}_{2}) and gLq(;2c)g\in L_{q}({\mathcal{M}};\ell^{c}_{2}),

ifigiLr()(i|fi|2)12Lp()(i|gi|2)12Lq().\Big{\|}\sum_{i}f_{i}^{*}g_{i}\Big{\|}_{L_{r}({\mathcal{M}})}\leq\Big{\|}\Big{(}\sum_{i}|f_{i}|^{2}\Big{)}^{\frac{1}{2}}\Big{\|}_{L_{p}({\mathcal{M}})}\Big{\|}\Big{(}\sum_{i}|g_{i}|^{2}\Big{)}^{\frac{1}{2}}\Big{\|}_{L_{q}({\mathcal{M}})}.

The second one is the noncommutative Khintchine inequality for the Rademacher sequence as follows.

Lemma 2.6 (see [62]).

Let 1p<1\leq p<\infty and {xn}\{x_{n}\} be a finite sequence in Lp()L_{p}({\mathcal{M}}). Then

nxnεnLp(Ω;Lp()){xn}Lp(;2rc),\big{\|}\sum_{n}x_{n}\varepsilon_{n}\big{\|}_{L_{p}(\Omega;L_{p}({\mathcal{M}}))}\approx\|\{x_{n}\}\|_{L_{p}({\mathcal{M}};\ell_{2}^{rc})},

where {εn}\{\varepsilon_{n}\} is a Rademacher sequence on a probability space (Ω,P)(\Omega,P).

We also require some convexity inequalities for the operator-valued function in this paper. The following one is Cauchy-Schwarz type inequality which can found in [36, Page 9]).

Lemma 2.7.

Let (Σ,μ)(\Sigma,\mu) be a measure space. Suppose that f:Σf:\ \Sigma\rightarrow{\mathcal{M}} is a weak-* integrable function and g:Σg:\Sigma\rightarrow\mathbb{C} is an integrable function. Then

(2.5) |Σf(x)g(x)𝑑μ(x)|2Σ|f(x)|2𝑑μ(x)Σ|g(x)|2𝑑μ(x),\Big{|}\int_{\Sigma}f(x)g(x)d\mu(x)\Big{|}^{2}\leq\int_{\Sigma}|f(x)|^{2}d\mu(x)\int_{\Sigma}|g(x)|^{2}d\mu(x),

where \leq is understood as the partial order in the positive cone of {\mathcal{M}}.

Finally we introduce the following vector-valued Plancherel theorem which will be mostly used in square function estimates: Let 𝒩=L(d)¯{\mathcal{N}}=L_{\infty}({\mathbb{R}}^{d})\overline{\otimes}{\mathcal{M}}, then we have

(2.6) fL2(𝒩)=fL2(𝒩),\|\mathcal{F}f\|_{L_{2}({\mathcal{N}})}=\|f\|_{L_{2}({\mathcal{N}})},

which is a consequence of the fact L2()L_{2}({\mathcal{M}}) is a Hilbert space. Such vector-valued Plancherel theorem is sufficient in most part of our proof though sometimes we need a more general operator-valued Parseval’s relation: for f,gL2(L(d)¯)f,g\in L_{2}(L_{\infty}({\mathbb{R}}^{d})\overline{\otimes}{\mathcal{M}}), we have

(2.7) dg(x)f(x)𝑑x=d(g)(ξ)(f)(ξ)𝑑ξ.\int_{{\mathbb{R}}^{d}}g^{*}(x)f(x)dx=\int_{{\mathbb{R}}^{d}}(\mathcal{F}g)^{*}(\xi)\mathcal{F}(f)(\xi)d\xi.

3. Noncommutative Kakeya maximal functions

In this section, we study the boundedness of noncommutative Kakeya maximal functions. Before that we give several definitions and lemmas. We only consider d=2d=2. Set 𝒩=L(2)¯{\mathcal{N}}=L_{\infty}({\mathbb{R}}^{2})\overline{\otimes}{\mathcal{M}} and φ=2𝑑xτ\varphi=\int_{{\mathbb{R}}^{2}}dx\otimes\tau. Denote Lp(𝒩)L_{p}({\mathcal{N}}) the noncommutative LpL_{p} space associated with pairs (𝒩,φ)({\mathcal{N}},\varphi).

The main preliminaries of noncommutative maximal functions have been given in Subsection 2.2. We introduce the noncommutative Kakeya maximal function as follows. Define the eccentricity of a rectangle by the ratio of the length of its long side to that of its short side. Let NN be a positive integer. Define the set N\mathcal{R}_{N} as rectangles in the plane of arbitrary orientation whose center is the origin and eccentricity is NN. For fLp(𝒩)f\in L_{p}({\mathcal{N}}), we define the Kakeya average operator as follows

(3.1) 𝒦Rf(x)=|R|1Rf(xy)𝑑y,\mathcal{K}_{R}f(x)=|R|^{-1}\int_{R}f(x-y)dy,

where RR is a rectangle belonging to N\mathcal{R}_{N}. We are mainly interested in the maximal L2L_{2} norm of noncommutative Kakeya average operator, since it is a crucial estimate in the study of noncommutative Bochner-Riesz means.

Let us first give a trivial bound of its L2L_{2} norm. Recall that the Hardy-Littlewood average operator is defined by

MQf(x)=1|Q|Qf(xy)𝑑y,M_{Q}f(x)=\frac{1}{|Q|}\int_{Q}f(x-y)dy,

where QQ is a cube in 2{\mathbb{R}}^{2} with center zero and arbitrary orientation. For any rectangle RNR\in\mathcal{R}_{N}, there exists a cube QQ such that RQR\subset Q and l(Q)l(Q) equals to the length of long side of RR. Consider ff as a positive function in 𝒩{\mathcal{N}}. Then 𝒦Rf(x)NMQf(x)\mathcal{K}_{R}f(x)\leq NM_{Q}f(x). Using the noncommutative Hardy-Littlewood maximal operator is of weak type (1,1)(1,1) (see [36]), we get the Kakeya maximal operator is of weak type (1,1)(1,1) with bound NN. Applying the noncommutative Marcinkiewicz interpolation theorem in [30], together with the fact that the maximal operator of 𝒦R\mathcal{K}_{R} is of (,)(\infty,\infty), we get that the maximal operator of 𝒦R\mathcal{K}_{R} is of strong (2,2)(2,2) type with bound N12N^{\frac{1}{2}}. However this bound is quite rough and not sufficient for our later application. The following improved bound is our main result in this section.

Theorem 3.1.

Let 𝒦R\mathcal{K}_{R} be the Kakeya average operator defined in (3.1). Then for fL2(𝒩)f\in L_{2}({\mathcal{N}}), we have

supRN𝒦RfL2(𝒩)(logN)fL2(𝒩).\|\sup_{R\in\mathcal{R}_{N}}\mathcal{K}_{R}f\|_{L_{2}({\mathcal{N}})}\lesssim(\log N)\|f\|_{L_{2}({\mathcal{N}})}.

Combining the noncommutative Marcinkiewicz interpolation theorem (see [30]), together with a trivial weak type (1,1)(1,1) bound and a strong (,)(\infty,\infty) bound of 𝒦R\mathcal{K}_{R}, we immediately get the following corollary.

Corollary 3.2.

Let 𝒦R\mathcal{K}_{R} be defined in (3.1). Then for any 1<p<1<p<\infty, we have

supRN𝒦RLp(𝒩)Lp(𝒩){N2p1(logN)2p,if   1<p<2;(logN)2p,if   2p<.\|\sup_{R\in\mathcal{R}_{N}}\mathcal{K}_{R}\|_{L_{p}({\mathcal{N}})\rightarrow L_{p}({\mathcal{N}})}\lesssim\begin{cases}N^{\frac{2}{p}-1}(\log N)^{\frac{2}{p^{\prime}}},\,&\hbox{if}\,\,\ 1<p<2;\\ (\log N)^{\frac{2}{p}},\,&\hbox{if}\,\,\ 2\leq p<\infty.\end{cases}

It should be pointed out that the bound logN\log N in Theorem 3.1 is sharp even in the commutative case, see [18, Proposition 5.3.4]. A. Córdoba [9] first obtained the boundedness of the Kakeya maximal function on L2(2)L_{2}({\mathbb{R}}^{2}) with norm (logN)2(\log N)^{2}. The sharp bound logN\log N was later established by J. O. Strömberg [49] where he used several estimates of distribution functions and some geometric constructions. S. Wainger [55] also obtained the sharp bound logN\log N without a proof but he mentioned that the idea given by A. Nagel, E. M. Stein, S. Wainger [39] can be modified to his setting. It is well-known that the distribution function is difficult to deal with in the noncommutative setting (for example the weak (1,1)(1,1) boundedness problem is a challenge problem). Hence the method from J. O. Strömberg [49] may be difficult to be applied in the noncommutative setting. The strategies used in our proof below are the Fourier transform, square and maximal function theories and the microlocal decomposition, which are mainly motivated by [39] and [55].

Before giving the proof of Theorem 3.1, we introduce the noncommutative directional Hardy-Littlewood average operator defined by

(3.2) Mhe(f)(x)=12hhhf(xey)𝑑y,M^{e}_{h}(f)(x)=\frac{1}{2h}\int_{-h}^{h}f(x-ey)dy,

where ee is a unit vector in 2{\mathbb{R}}^{2}. By using the standard method of rotation, the definition of maximal norm in (2.3) and the fact one dimensional noncommutative Hardy-Littlewood maximal operator is of strong type (p,p)(p,p) for 1<p1<p\leq\infty (see [36]), the author and his collaborators recently established the following result in [21, Lemma 6.3].

Lemma 3.3.

Let ee be a unit vector. Define MheM_{h}^{e} in (3.2). Let 1<p1<p\leq\infty. Then we have

suph>0Mhe(f)Lp(𝒩)fLp(𝒩).\|\sup_{h>0}M_{h}^{e}(f)\|_{L_{p}({\mathcal{N}})}\lesssim\|f\|_{L_{p}({\mathcal{N}})}.

Now we are in a position to prove Theorem 3.1.

Proof of Theorem 3.1.

Let us start with several reductions. Without loss of generality, we suppose that ff is positive since the general case just follows by decomposing ff as linear combination of four positive functions.

To prove our estimate, it suffices to consider the case N=2mN=2^{m}. In fact, assume that we show this theorem for N=2mN=2^{m}, we can prove the general case for arbitrary NN as follows. For any positive integer NN, there exists a positive integer mm such that 2m1<N2m2^{m-1}<N\leq 2^{m}, i.e. m1<logNmm-1<\log N\leq m. Then for any RNR\in\mathcal{R}_{N}, by enlarging the long side of RR such that its eccentricity increases to 2m2^{m}, we get a new rectangle R~\tilde{R}. Then 𝒦Rf(x)2mN𝒦R~f(x).\mathcal{K}_{R}f(x)\leq\frac{2^{m}}{N}\mathcal{K}_{\tilde{R}}f(x). Therefore we get

supRN𝒦RfL2(𝒩)2mNsupR2m𝒦RfL2(𝒩)mfL2(𝒩)(logN)fL2(𝒩).\|\sup_{R\in\mathcal{R}_{N}}\mathcal{K}_{R}f\|_{L_{2}({\mathcal{N}})}\leq\frac{2^{m}}{N}\|\sup_{R\in\mathcal{R}_{2^{m}}}\mathcal{K}_{R}f\|_{L_{2}({\mathcal{N}})}\lesssim m\|f\|_{L_{2}({\mathcal{N}})}\approx(\log N)\|f\|_{L_{2}({\mathcal{N}})}.

Similar to that in the commutative case, by symmetry and rotation, we only need to control the average over those rectangles that have eccentricity NN, but whose major axes make angles θk\theta_{k} with the xx-axis such that tanθk=k/N,k=0,,N1\tan\theta_{k}=k/N,k=0,\cdots,N-1 (see e.g. Section 3.11, Chapter X in [47]). We abuse notation and still define N\mathcal{R}_{N} as those preceding rectangles.

After these reductions, below we smooth the average operator. Set ukN=(N,k),k=0,,N1u^{N}_{k}=(N,k),k=0,\cdots,N-1 and let e2=(0,1)2e_{2}=(0,1)\in{\mathbb{R}}^{2}. Choose a nonnegative, radially decreasing and smooth function ψ\psi such that ψ(x)=1\psi(x)=1 if |x|<12|x|<\frac{1}{2} and suppψ{|x|<2}{\text{\rm supp}}\psi\subset\{|x|<2\}. Define ψh(t)=h1ψ(t/h)\psi_{h}(t)=h^{-1}\psi(t/h). To prove our theorem, it is sufficient to consider the average operator

Ahk,N(f)(x)=f(xukNte2s)ψh(t)ψh(s)𝑑s𝑑t,h>0.A_{h}^{k,N}(f)(x)=\int_{{\mathbb{R}}}\int_{{\mathbb{R}}}f(x-u^{N}_{k}t-e_{2}s)\psi_{h}(t)\psi_{h}(s)dsdt,\ h>0.

Indeed, since f0f\geq 0, by some elementary geometric observation, for any RNR\in\mathcal{R}_{N} with the major direction eiθe^{i\theta} where tanθ=kN\tan\theta=\frac{k}{N} for some k=0,,N1k=0,\cdots,N-1, there exists hh such that

𝒦Rf(x)CAhk,N(f)(x)\mathcal{K}_{R}f(x)\leq CA_{h}^{k,N}(f)(x)

where the constant CC is independent of k,N,fk,N,f. On the other hand, for 0kN10\leq k\leq N-1, h+h\in{\mathbb{R}}_{+}, there exists RNR\in\mathcal{R}_{N} with the major direction eiθe^{i\theta} such that tanθ=kN\tan\theta=\frac{k}{N},

Ahk,N(f)(x)C𝒦Rf(x).A_{h}^{k,N}(f)(x)\leq C\mathcal{K}_{R}f(x).

Therefore by the definition of positive maximal norm in (2.3), we get

supRN𝒦RfL2(𝒩)sup0k<Nh>0Ahk,N(f)L2(𝒩).\Big{\|}\sup_{R\in\mathcal{R}_{N}}\mathcal{K}_{R}f\Big{\|}_{L_{2}({\mathcal{N}})}{\approx}\Big{\|}\sup_{0\leq k<N\atop h>0}A_{h}^{k,N}(f)\Big{\|}_{L_{2}({\mathcal{N}})}.

Notice that there are two averages for two different directions in the operator Ahk,NA_{h}^{k,N}. Our next goal is to reduce it to one average with the help of the directional Hardy-Littlewood maximal operator. By our choice of ψ\psi, it is easy to see that

Ahk,N(f)(x)M2he2[Mhk,N(f)](x)A_{h}^{k,N}(f)(x)\lesssim M^{e_{2}}_{2h}[M_{h}^{k,N}(f)](x)

where M2he2M_{2h}^{e_{2}} is the directional Hardy-Littlewood average operator defined in (3.2) and Mhk,N(f)M_{h}^{k,N}(f) is defined as

Mhk,N(f)(x)=f(xukNt)ψh(t)𝑑t.M_{h}^{k,N}(f)(x)=\int_{{\mathbb{R}}}f(x-u^{N}_{k}t)\psi_{h}(t)dt.

Therefore by the L2L_{2} boundedness of maximal operator of MheM^{e}_{h} in Lemma 3.3, to prove our theorem, it suffices to show

(3.3) sup0k<Nh>0Mhk,N(f)L2(𝒩)(logN)fL2(𝒩).\Big{\|}\sup_{0\leq k<N\atop h>0}M_{h}^{k,N}(f)\Big{\|}_{L_{2}({\mathcal{N}})}\lesssim(\log N)\|f\|_{L_{2}({\mathcal{N}})}.

Next we reduce the study of Mhk,N(f)(x)M_{h}^{k,N}(f)(x) to its lacunary case M2jk,N(f)(x)M_{2^{j}}^{k,N}(f)(x). For any h>0h>0, there exists jj\in{\mathbb{Z}} such 2j1h<2j2^{j-1}\leq h<2^{j}. Then we get

Mhk,N(f)(x)=1hf(xukNt)ψ(th)𝑑t12jf(xukNt)ψ(t2j)𝑑t=M2jk,N(f)(x),M_{h}^{k,N}(f)(x)=\frac{1}{h}\int_{{\mathbb{R}}}f(x-u^{N}_{k}t)\psi(\frac{t}{h})dt\lesssim\frac{1}{2^{j}}\int_{{\mathbb{R}}}f(x-u^{N}_{k}t)\psi(\frac{t}{2^{j}})dt=M_{2^{j}}^{k,N}(f)(x),

where the second inequality just follows from the radially decreasing property of ψ\psi. Hence we only need to consider the lacunary operator M2jk,N(f)(x)M_{2^{j}}^{k,N}(f)(x). At present time, we conclude that the proof of our main theorem is reduced to show

(3.4) sup0k<NjM2jk,N(f)L2(𝒩)(logN)fL2(𝒩).\Big{\|}\sup_{0\leq k<N\atop j\in{\mathbb{Z}}}M_{2^{j}}^{k,N}(f)\Big{\|}_{L_{2}({\mathcal{N}})}\lesssim(\log N)\|f\|_{L_{2}({\mathcal{N}})}.

Let N=2mN=2^{m}. In the following, we will establish a key inequality

(3.5) sup0k<2mjM2jk,2m(f)L2(𝒩)CfL2(𝒩)+sup0k<2m1jM2jk,2m1(f)L2(𝒩)\Big{\|}\sup_{0\leq k<2^{m}\atop j\in{\mathbb{Z}}}M_{2^{j}}^{k,2^{m}}(f)\Big{\|}_{L_{2}({\mathcal{N}})}\leq C\|f\|_{L_{2}({\mathcal{N}})}+\Big{\|}\sup_{0\leq k<2^{m-1}\atop j\in{\mathbb{Z}}}M_{2^{j}}^{k,2^{m-1}}(f)\Big{\|}_{L_{2}({\mathcal{N}})}

with the constant CC independent of mm. Notice that when m=1m=1, M2j0,1(f)M2j+1e1(f)M_{2^{j}}^{0,1}(f)\leq M^{e_{1}}_{2^{j+1}}(f), where e1=(1,0)e_{1}=(1,0) and MheM_{h}^{e} is the noncommutative directional Hardy-Littlewood average operator in (3.2). Using the L2L_{2} boundedness of maximal MheM_{h}^{e} in Lemma 3.3, we get

supjM2j0,1(f)L2(𝒩)fL2(𝒩).\Big{\|}\sup_{j\in{\mathbb{Z}}}M_{2^{j}}^{0,1}(f)\Big{\|}_{L_{2}({\mathcal{N}})}\lesssim\|f\|_{L_{2}({\mathcal{N}})}.

Then it is easy to see that the required estimate (3.4) just follows from (3.5) with an induction argument.

The rest of this section is devoted to the proof of (3.5). For convenience, set eml=(1,l2m)e_{m}^{l}=(1,\frac{l}{2^{m}}). Define Γl={ξ:|ξ|ξ|,eml|c2m}\Gamma_{l}=\{\xi:|\langle\frac{\xi}{|\xi|},e_{m}^{l}\rangle|\leq\frac{c}{2^{m}}\} where cc is a constant independent of mm such that Γ1,Γ2,,Γ2m11\Gamma_{1},\Gamma_{2},\cdots,\Gamma_{2^{m-1}-1} are disjoint from each other. Notice that Γl\Gamma_{l}s are equally distributed in {(x,y):|x||y|,x0,y0}\{(x,y):|x|\leq|y|,x\leq 0,y\geq 0\} or {(x,y):|x||y|,x0,y0}\{(x,y):|x|\leq|y|,x\geq 0,y\leq 0\} (see Figure 1 later). We split ff as flf_{l} and rlr_{l} which are defined by

fl^(ξ)=χΓ2l+1Γ2l(ξ)f^(ξ),rl^(ξ)=χΓ2l+1cΓ2lc(ξ)f^(ξ).\widehat{f_{l}}(\xi)=\chi_{\Gamma_{2l+1}\cup\Gamma_{2l}}(\xi)\widehat{f}(\xi),\quad\widehat{r_{l}}(\xi)=\chi_{\Gamma_{2l+1}^{c}\cap\Gamma^{c}_{2l}}(\xi)\widehat{f}(\xi).

Recall that φ=2𝑑xτ\varphi=\int_{{\mathbb{R}}^{2}}dx\otimes\tau. Since M2jk,N(f)(x)M_{2^{j}}^{k,N}(f)(x) is positive in L2(𝒩)L_{2}({\mathcal{N}}), by the duality (see (ii) in Lemma 2.3), there exists a positive sequence {h2jk}L2(𝒩;1)\{h_{2^{j}}^{k}\}\in L_{2}({\mathcal{N}};\ell_{1}) with norm 0k<2mjh2jkL2(𝒩)1\|\sum\limits_{0\leq k<2^{m}\atop j\in{\mathbb{Z}}}h_{2^{j}}^{k}\|_{L_{2}({\mathcal{N}})}\leq 1 such that

sup0k<2mjM2jk,2m(f)L2(𝒩)=φ(0k<2mjM2jk,2m(f)h2jk)I+II\begin{split}\Big{\|}\sup_{0\leq k<2^{m}\atop j\in{\mathbb{Z}}}M_{2^{j}}^{k,2^{m}}(f)\Big{\|}_{L_{2}({\mathcal{N}})}&=\varphi\Big{(}\sum_{0\leq k<2^{m}\atop j\in{\mathbb{Z}}}M_{2^{j}}^{k,2^{m}}(f)h_{2^{j}}^{k}\Big{)}\leq I+II\end{split}

where

I=|φ(0k<2mj[M2j+1k2,2m1(f)]h2jk)|,I=\Big{|}\varphi\Big{(}\sum_{0\leq k<2^{m}\atop j\in{\mathbb{Z}}}[M_{2^{j+1}}^{\lfloor\frac{k}{2}\rfloor,2^{m-1}}(f)]h_{2^{j}}^{k}\Big{)}\Big{|},
II=|φ(0k<2mj[M2jk,2m(f)M2j+1k2,2m1(f)]h2jk)|.II=\Big{|}\varphi\Big{(}\sum_{0\leq k<2^{m}\atop j\in{\mathbb{Z}}}[M_{2^{j}}^{k,2^{m}}(f)-M_{2^{j+1}}^{\lfloor\frac{k}{2}\rfloor,2^{m-1}}(f)]h_{2^{j}}^{k}\Big{)}\Big{|}.

We consider the first term II. Notice that for every 0k<2m10\leq k<2^{m-1}, we have

(3.6) M2j+12k2,2m1(f)=M2j+12k+12,2m1(f).M_{2^{j+1}}^{\lfloor\frac{2k}{2}\rfloor,2^{m-1}}(f)=M_{2^{j+1}}^{\lfloor\frac{2k+1}{2}\rfloor,2^{m-1}}(f).

Recall that M2j+1k,2m1(f)M_{2^{j+1}}^{k,2^{m-1}}(f) is positive in 𝒩{\mathcal{N}}. Then by the duality (see (ii) in Lemma 2.3), the definition of L2(𝒩;)L_{2}({\mathcal{N}};\ell_{\infty}) in (2.3) and the preceding equality (3.6), we get

Isup0k<2mjM2j+1k2,2m1(f)L2(𝒩)=sup0k<2m1jM2j+1k,2m1(f)L2(𝒩)=sup0k<2m1jM2jk,2m1(f)L2(𝒩),\begin{split}I&\leq\Big{\|}\sup_{0\leq k<2^{m}\atop j\in{\mathbb{Z}}}M_{2^{j+1}}^{\lfloor\frac{k}{2}\rfloor,2^{m-1}}(f)\Big{\|}_{L_{2}({\mathcal{N}})}=\Big{\|}\sup_{0\leq k<2^{m-1}\atop j\in{\mathbb{Z}}}M_{2^{j+1}}^{k,2^{m-1}}(f)\Big{\|}_{L_{2}({\mathcal{N}})}\\ &=\Big{\|}\sup_{0\leq k<2^{m-1}\atop j\in{\mathbb{Z}}}M_{2^{j}}^{k,2^{m-1}}(f)\Big{\|}_{L_{2}({\mathcal{N}})},\end{split}

which is exact the second term in right side of (3.5).

Now we turn to IIII. To finish the proof of (3.5), we only need to show that IIII is controlled by fL2(𝒩)\|f\|_{L_{2}({\mathcal{N}})}. By making a dilation, it is easy to check that for 0l<2m10\leq l<2^{m-1},

(3.7) M2j+1l,2m1(f)(x)=M2j2l,2m(f)(x).M_{2^{j+1}}^{l,2^{m-1}}(f)(x)=M_{2^{j}}^{2l,2^{m}}(f)(x).

Therefore we see that the terms related to even kk in the sum of IIII equal to zero. Applying (3.6) and (3.7) again, we rewrite the odd terms in IIII as follows

II=|φ(0l<2m1j[M2j2l+1,2m(f)M2j2l,2m(f)]h2j2l+1)||φ(0l<2m1j[M2j2l+1,2m(fl)h2j2l+1])|+|φ(0l<2m1j[M2j2l,2m(fl)h2j2l+1])|+|φ(0l<2m1j[M2j2l+1,2m(rl)M2j2l,2m(rl)]h2j2l+1)|=:II1+II2+II3.\begin{split}II&=\Big{|}\varphi\Big{(}\sum_{0\leq l<2^{m-1}\atop j\in{\mathbb{Z}}}[M_{2^{j}}^{2l+1,2^{m}}(f)-M_{2^{j}}^{2l,2^{m}}(f)]h_{2^{j}}^{2l+1}\Big{)}\Big{|}\\ &\leq\Big{|}\varphi\Big{(}\sum_{0\leq l<2^{m-1}\atop j\in{\mathbb{Z}}}[M_{2^{j}}^{2l+1,2^{m}}(f_{l})h_{2^{j}}^{2l+1}]\Big{)}\Big{|}+\Big{|}\varphi\Big{(}\sum_{0\leq l<2^{m-1}\atop j\in{\mathbb{Z}}}[M_{2^{j}}^{2l,2^{m}}(f_{l})h_{2^{j}}^{2l+1}]\Big{)}\Big{|}\\ &\quad+\Big{|}\varphi\Big{(}\sum_{0\leq l<2^{m-1}\atop j\in{\mathbb{Z}}}[M_{2^{j}}^{2l+1,2^{m}}(r_{l})-M_{2^{j}}^{2l,2^{m}}(r_{l})]h_{2^{j}}^{2l+1}\Big{)}\Big{|}=:II_{1}+II_{2}+II_{3}.\end{split}

Let us consider II1II_{1} firstly. Using the duality in (ii) of Lemma 2.3, we get

II1sup0l<2m1jM2j2l+1,2m(fl)L2(𝒩).II_{1}\leq\big{\|}\sup_{0\leq l<2^{m-1}\atop j\in{\mathbb{Z}}}M_{2^{j}}^{2l+1,2^{m}}(f_{l})\big{\|}_{L_{2}({\mathcal{N}})}.

Recall eml=(1,l2m)e_{m}^{l}=(1,\frac{l}{2^{m}}). Let e~m2l+1\tilde{e}_{m}^{2l+1} be the unit vector in the direction along em2l+1e_{m}^{2l+1}, i.e. e~m2l+1=em2l+1/|em2l+1|\tilde{e}_{m}^{2l+1}={e}_{m}^{2l+1}/|{e}_{m}^{2l+1}|. Then it is straightforward to verify that for a positive function gg,

M2j2l+1,2m(g)(x)M2j+me~m2l+1(g)(x),M_{2^{j}}^{2l+1,2^{m}}(g)(x)\lesssim M^{\tilde{e}_{m}^{2l+1}}_{2^{j+m}}(g)(x),

where the right side of the above inequality is the directional Hardy-Littlewood average operator defined in (3.2). Now using the L2L_{2} boundedness of maximal operator of MheM^{e}_{h} in Lemma 3.3, we get for any positive function gL2(𝒩)g\in L_{2}({\mathcal{N}}),

(3.8) supjM2j2l+1,2m(g)L2(𝒩)gL2(𝒩).\big{\|}\sup_{j\in{\mathbb{Z}}}M_{2^{j}}^{2l+1,2^{m}}(g)\big{\|}_{L_{2}({\mathcal{N}})}\lesssim\|g\|_{L_{2}({\mathcal{N}})}.

Consequently (3.8) holds for any gL2(𝒩)g\in L_{2}({\mathcal{N}}) by decomposing it as linear combination of four positive functions.

Observe that M2j2l+1,2m(fl)M_{2^{j}}^{2l+1,2^{m}}(f_{l}) may not be positive. So we can not apply the maximal norm in (2.3). Recall that for any a𝒩a\in{\mathcal{N}}, we can decompose it as linear combination of two self-adjoint elements:

a=Re(a)+iIm(a),whereRe(a)=12(a+a),Im(a)=12i(aa).a=\text{Re}(a)+i\text{Im}(a),\quad\text{where}\quad\text{Re}(a)=\frac{1}{2}{(a+a^{*})},\ \text{Im}(a)=\frac{1}{2i}(a-a^{*}).

Hence we can write M2j2l+1,2m(fl)=Re[M2j2l+1,2m(fl)]+iIm[M2j2l+1,2m(fl)]M_{2^{j}}^{2l+1,2^{m}}(f_{l})=\text{Re}[M_{2^{j}}^{2l+1,2^{m}}(f_{l})]+i\text{Im}[M_{2^{j}}^{2l+1,2^{m}}(f_{l})]. Then utilizing Minkowski’s inequality, we get

II1sup0l<2m1jRe[M2j2l+1,2m(fl)]L2(𝒩)+sup0l<2m1jIm[M2j2l+1,2m(fl)]L2(𝒩).II_{1}\leq\big{\|}\sup_{0\leq l<2^{m-1}\atop j\in{\mathbb{Z}}}\text{Re}[M_{2^{j}}^{2l+1,2^{m}}(f_{l})]\big{\|}_{L_{2}({\mathcal{N}})}+\big{\|}\sup_{0\leq l<2^{m-1}\atop j\in{\mathbb{Z}}}\text{Im}[M_{2^{j}}^{2l+1,2^{m}}(f_{l})]\big{\|}_{L_{2}({\mathcal{N}})}.

We first consider the real part. Notice that M2j2l+1,2m(fl)=M2j2l+1,2m(fl)M_{2^{j}}^{2l+1,2^{m}}(f_{l})^{*}=M_{2^{j}}^{2l+1,2^{m}}(f_{l}^{*}). Then by Minkowski’s inequality and (3.8), we get

supjRe[M2j2l+1,2m(fl)]L2(𝒩)12supjM2j2l+1,2m(fl)L2(𝒩)+12supjM2j2l+1,2m(fl)L2(𝒩)flL2(𝒩).\begin{split}\big{\|}\sup_{j\in{\mathbb{Z}}}\text{Re}[M_{2^{j}}^{2l+1,2^{m}}(f_{l})]\big{\|}_{L_{2}({\mathcal{N}})}\leq&\frac{1}{2}\big{\|}\sup_{j\in{\mathbb{Z}}}M_{2^{j}}^{2l+1,2^{m}}(f_{l})\big{\|}_{L_{2}({\mathcal{N}})}\\ &+\frac{1}{2}\big{\|}\sup_{j\in{\mathbb{Z}}}M_{2^{j}}^{2l+1,2^{m}}(f_{l}^{*})\big{\|}_{L_{2}({\mathcal{N}})}\lesssim\|f_{l}\|_{L_{2}({\mathcal{N}})}.\end{split}

Rewrite this estimate via the equivalent definition of Lp(;)L_{p}({\mathcal{M}};\ell_{\infty}) in (2.4), we obtain that there exists Fl>0F_{l}>0 such that for each jj\in{\mathbb{Z}},

FlRe[M2j2l+1,2m(fl)]FlandFlL2(𝒩)flL2(𝒩).-F_{l}\leq\text{Re}[M_{2^{j}}^{2l+1,2^{m}}(f_{l})]\leq F_{l}\quad\text{and}\quad\|F_{l}\|_{L_{2}({\mathcal{N}})}\lesssim\|f_{l}\|_{L_{2}({\mathcal{N}})}.

Then by setting F=(lFl2)12F=(\sum_{l}F_{l}^{2})^{\frac{1}{2}}, we see that for each 0l<2m1,j0\leq l<2^{m-1},j\in{\mathbb{Z}}, we have FRe[M2j2l+1,2m(fl)]F-F\leq\text{Re}[M_{2^{j}}^{2l+1,2^{m}}(f_{l})]\leq F because of Fl(lFl2)12F_{l}\leq(\sum_{l}F_{l}^{2})^{\frac{1}{2}} which is just by an elementary inequality (2.2). Moreover we get

FL2(𝒩)=(lFlL2(𝒩)2)12(lflL2(𝒩)2)12=(τ2lχΓ2l+1Γ2l(ξ)|f^(ξ)|2dξ)1/2fL2(𝒩)\begin{split}\|F\|_{L_{2}({\mathcal{N}})}&=(\sum_{l}\|F_{l}\|_{L_{2}({\mathcal{N}})}^{2})^{\frac{1}{2}}\lesssim(\sum_{l}\|f_{l}\|_{L_{2}({\mathcal{N}})}^{2})^{\frac{1}{2}}\\ &=\Big{(}\tau\int_{{\mathbb{R}}^{2}}\sum_{l}\chi_{\Gamma_{2l+1}\cup\Gamma_{2l}}(\xi)|\widehat{f}(\xi)|^{2}d\xi\Big{)}^{1/2}\lesssim\|f\|_{L_{2}({\mathcal{N}})}\end{split}

where in the third equality we use vector-valued Plancherel’s theorem (2.6), the last inequality just follows from Γ1,,Γ2m11\Gamma_{1},\cdots,\Gamma_{2^{m-1}-1} are disjoint from each other and vector-valued Plancherel’s theorem (2.6) again. Thus we prove that

sup0l<2m1jRe[M2j2l+1,2m(fl)]L2(𝒩)fL2(𝒩).\big{\|}\sup_{0\leq l<2^{m-1}\atop j\in{\mathbb{Z}}}\text{Re}[M_{2^{j}}^{2l+1,2^{m}}(f_{l})]\big{\|}_{L_{2}({\mathcal{N}})}\lesssim\|f\|_{L_{2}({\mathcal{N}})}.

By applying the similar argument to the imaginary part, we could also get

sup0l<2m1jIm[M2j2l+1,2m(fl)]L2(𝒩)fL2(𝒩).\big{\|}\sup_{0\leq l<2^{m-1}\atop j\in{\mathbb{Z}}}\text{Im}[M_{2^{j}}^{2l+1,2^{m}}(f_{l})]\big{\|}_{L_{2}({\mathcal{N}})}\lesssim\|f\|_{L_{2}({\mathcal{N}})}.

Combining these estimates of real and imaginary parts, we obtain the desired estimate of II1II_{1}.

For the term II2II_{2}, using the similar argument as we have done in the proof of II1II_{1}, we also get that II2II_{2} is bounded by fL2(𝒩)\|f\|_{L_{2}({\mathcal{N}})}.

At last we turn to the term II3II_{3}. We first introduce an inequality as follows:

(3.9) |φ(ab)|2φ(|a|b)φ(|a|b),a,b𝒩withb0.|\varphi(ab)|^{2}\leq\varphi(|a|b)\varphi(|a^{*}|b),\quad\forall\ a,b\in{\mathcal{N}}\ \text{with}\ b\geq 0.

This inequality could be verified by writing aa as the polar decomposition a=u|a|a=u|a| and using Cauchy-Schwarz’s inequality,

|φ(ab)|2=|φ(b1/2u|a|b1/2)|2φ(b1/2u|a|ub1/2)φ(b1/2|a|b1/2)=φ(b1/2|a|b1/2)φ(b1/2|a|b1/2)=φ(|a|b)φ(|a|b).\begin{split}|\varphi(ab)|^{2}&=|\varphi(b^{1/2}u|a|b^{1/2})|^{2}\leq\varphi(b^{1/2}u|a|u^{*}b^{1/2})\varphi(b^{1/2}|a|b^{1/2})\\ &=\varphi(b^{1/2}|a^{*}|b^{1/2})\varphi(b^{1/2}|a|b^{1/2})=\varphi(|a^{*}|b)\varphi(|a|b).\end{split}

For simplicity, we define B2jl,2m(g)=M2j2l+1,2m(g)M2j2l,2m(g)B_{2^{j}}^{l,2^{m}}(g)=M_{2^{j}}^{2l+1,2^{m}}(g)-M_{2^{j}}^{2l,2^{m}}(g). Then by the above inequality (3.9) and Cauchy-Schwarz’s inequality, we have

II30l<2m1jφ(|B2jl,2m(rl)|h2j2l+1)12φ(|B2jl,2m(rl)|h2j2l+1)12φ(0l<2m1j|B2jl,2m(rl)|h2j2l+1)12φ(0l<2m1j|B2jl,2m(rl)|h2j2l+1)12sup0l<2m1j|B2jl,2m(rl)|L2(𝒩)12sup0l<2m1j|B2jl,2m(rl)|L2(𝒩)12,\begin{split}II_{3}&\leq\sum_{0\leq l<2^{m-1}\atop j\in{\mathbb{Z}}}\varphi(|B_{2^{j}}^{l,2^{m}}(r_{l})|h_{2^{j}}^{2l+1})^{\frac{1}{2}}\varphi(|B_{2^{j}}^{l,2^{m}}(r_{l})^{*}|h_{2^{j}}^{2l+1})^{\frac{1}{2}}\\ &\leq\varphi\Big{(}\sum_{0\leq l<2^{m-1}\atop j\in{\mathbb{Z}}}|B_{2^{j}}^{l,2^{m}}(r_{l})|h_{2^{j}}^{2l+1}\Big{)}^{\frac{1}{2}}\cdot\varphi\Big{(}\sum_{0\leq l<2^{m-1}\atop j\in{\mathbb{Z}}}|B_{2^{j}}^{l,2^{m}}(r_{l})^{*}|h_{2^{j}}^{2l+1}\Big{)}^{\frac{1}{2}}\\ &\leq\Big{\|}\sup_{0\leq l<2^{m-1}\atop j\in{\mathbb{Z}}}|B_{2^{j}}^{l,2^{m}}(r_{l})|\Big{\|}_{L_{2}({\mathcal{N}})}^{\frac{1}{2}}\cdot\Big{\|}\sup_{0\leq l<2^{m-1}\atop j\in{\mathbb{Z}}}|B_{2^{j}}^{l,2^{m}}(r_{l})^{*}|\Big{\|}_{L_{2}({\mathcal{N}})}^{\frac{1}{2}},\end{split}

where in the last inequality we apply the dual property (ii) in Lemma 2.3.

We first consider the part |B2jl,2m(rl)||B_{2^{j}}^{l,2^{m}}(r_{l})|. Our goal is to show that

(3.10) sup0l<2m1j|B2jl,2m(rl)|L2(𝒩)fL2(𝒩).\Big{\|}\sup_{0\leq l<2^{m-1}\atop j\in{\mathbb{Z}}}|B_{2^{j}}^{l,2^{m}}(r_{l})|\Big{\|}_{L_{2}({\mathcal{N}})}\lesssim\|f\|_{L_{2}({\mathcal{N}})}.

The strategy here is to use a square function to control the maximal function, which has been appeared in the proof of II1II_{1}. In fact applying an equivalent norm of Lp(;)L_{p}({\mathcal{M}};\ell_{\infty}) in (2.3), monotone properties in (2.2) and (2.1), we have

sup0l<2m1j|B2jl,2m(rl)|L2(𝒩)(0l<2m1j|B2jl,2m(rl)|2)12L2(𝒩).\Big{\|}\sup_{0\leq l<2^{m-1}\atop j\in{\mathbb{Z}}}|B_{2^{j}}^{l,2^{m}}(r_{l})|\Big{\|}_{L_{2}({\mathcal{N}})}\leq\Big{\|}\Big{(}\sum_{0\leq l<2^{m-1}\atop j\in{\mathbb{Z}}}|B_{2^{j}}^{l,2^{m}}(r_{l})|^{2}\Big{)}^{\frac{1}{2}}\Big{\|}_{L_{2}({\mathcal{N}})}.

It is straightforward to show that [M2j2l,2m(g)](ξ)=ψ^(2j+mem2l,ξ)g^(ξ)\mathcal{F}[M_{2^{j}}^{2l,2^{m}}(g)](\xi)=\widehat{\psi}(2^{j+m}\langle e^{2l}_{m},\xi\rangle)\widehat{g}(\xi). Then utilizing vector-valued Plancherel’s theorem (2.6) and the definitions of B2jl,2mB_{2^{j}}^{l,2^{m}}, rlr_{l}, we get that

(0l<2m1j|B2jl,2m(rl)|2)12L2(𝒩)2=τ2j0l<2m1|ψ^(2j+mem2l+1,ξ)ψ^(2j+mem2l,ξ)|2χΓ2l+1cΓ2lc(ξ)|f^(ξ)|2dξ.\begin{split}&\Big{\|}\Big{(}\sum_{0\leq l<2^{m-1}\atop j\in{\mathbb{Z}}}|B_{2^{j}}^{l,2^{m}}(r_{l})|^{2}\Big{)}^{\frac{1}{2}}\Big{\|}_{L_{2}({\mathcal{N}})}^{2}\\ &=\tau\int_{{\mathbb{R}}^{2}}\sum_{j\in{\mathbb{Z}}}\sum_{0\leq l<2^{m-1}}\Big{|}\widehat{\psi}\Big{(}2^{j+m}\langle e_{m}^{2l+1},\xi\rangle\Big{)}-\widehat{\psi}\Big{(}2^{j+m}\langle e_{m}^{2l},\xi\rangle\Big{)}\Big{|}^{2}\chi_{\Gamma_{2l+1}^{c}\cap\Gamma_{2l}^{c}}(\xi)|\widehat{f}(\xi)|^{2}d\xi.\end{split}

To prove (3.10), again by vector-valued Plancherel’s theorem (2.6), we only need to show the multiplier above is bounded, i.e.

(3.11) j0l<2m1|ψ^(2j+mem2l+1,ξ)ψ^(2j+mem2l,ξ)|2χΓ2l+1cΓ2lc(ξ)C\sum_{j\in{\mathbb{Z}}}\sum_{0\leq l<2^{m-1}}\Big{|}\widehat{\psi}\Big{(}2^{j+m}\langle e_{m}^{2l+1},\xi\rangle\Big{)}-\widehat{\psi}\Big{(}2^{j+m}\langle e_{m}^{2l},\xi\rangle\Big{)}\Big{|}^{2}\chi_{\Gamma_{2l+1}^{c}\cap\Gamma_{2l}^{c}}(\xi)\leq C

holds uniformly for ξ0\xi\neq 0. Fix ξ0\xi\neq 0. It suffices to consider the sum of ll such that ξΓ2l+1cΓ2lc\xi\in\Gamma_{2l+1}^{c}\cap\Gamma_{2l}^{c} which means that

|em2l+1,ξ|>c2m,|em2l,ξ|>c2m|\langle e_{m}^{2l+1},\xi^{\prime}\rangle|>c2^{-m},\quad|\langle e_{m}^{2l},\xi^{\prime}\rangle|>c2^{-m}

where ξ=ξ/|ξ|\xi^{\prime}=\xi/|\xi|. Such lower estimates may be not enough to prove (3.11). In the following, we obtain some better lower estimates via some geometric observations of the plane. Denote L={l: 0l<2m1,ξΓ2l+1cΓ2lc}L=\{l:\ 0\leq l<2^{m-1},\xi\in\Gamma_{2l+1}^{c}\cap\Gamma_{2l}^{c}\}. Then by using the mean value formula and Cauchy-Schwarz’s inequality, we see that

LHS(3.11)jlL[012j|ξ||ψ^(2j+mem2l+s,ξ)|𝑑s]201jlL[2j|ξ||ψ^(2j+mem2l+s,ξ)|]2ds.\begin{split}\text{LHS}\eqref{e:21multi}&\leq\sum_{j\in{\mathbb{Z}}}\sum_{l\in L}\Big{[}\int_{0}^{1}2^{j}|\xi|\cdot|\nabla\widehat{\psi}(2^{j+m}\langle e_{m}^{2l+s},\xi\rangle)|ds\Big{]}^{2}\\ &\leq\int_{0}^{1}\sum_{j\in{\mathbb{Z}}}\sum_{l\in L}\big{[}2^{j}|\xi|\cdot|\nabla\widehat{\psi}(2^{j+m}\langle e_{m}^{2l+s},\xi\rangle)|\big{]}^{2}ds.\end{split}

Let ee^{\bot} denote the orthogonal unit vector of ee in 2{\mathbb{R}}^{2}. Then we observe that eml,e_{m}^{l,\bot}s, where l=0,,2m11l=0,\cdots,2^{m-1}-1, are equally distributed with distance 2m2^{-m} in the plane {(x,y):|x||y|,x0,y0}\{(x,y):|x|\leq|y|,x\leq 0,y\geq 0\} or {(x,y):|x||y|,x0,y0}\{(x,y):|x|\leq|y|,x\geq 0,y\leq 0\}(see Figure 1 below).

Refer to caption
Figure 1. {eml}l\{e_{m}^{l}\}_{l} is equally distributed. Γl\Gamma_{l}s are disjoint from each other. |em2l,ξ||\langle e_{m}^{2l},\xi^{\prime}\rangle| is the distance between the point ξ\xi^{\prime} and the line along the direction em2l,e^{2l,\bot}_{m}.

Notice that the geometric illustration of |e,ξ||\langle e,\xi^{\prime}\rangle| is the distance between the point ξ\xi^{\prime} and the line along the direction ee^{\bot}. Recall that ξ\xi^{\prime} is a fixed unit vector. By the equally distributed property of {eml,}l{0,,2m11}\{e_{m}^{l,\bot}\}_{l\in\{0,\cdots,2^{m-1}-1\}}, we can separate the set LL into at most 2m12^{m-1} subsets L1,,L2m1L_{1},\cdots,L_{2^{m-1}} such that each LiL_{i} has cardinality less than some absolute constant and for each ii,

|em2l+1,ξ|>ci2m,|em2l,ξ|>ci2m,lLi.|\langle e_{m}^{2l+1},\xi^{\prime}\rangle|>ci2^{-m},\quad|\langle e_{m}^{2l},\xi^{\prime}\rangle|>ci2^{-m},\quad\forall l\in L_{i}.

This fact can be easily seen from the geometry in Figure 1. Therefore by the above inequality, for any s[0,1]s\in[0,1], lLil\in L_{i}, we get |em2l+s,ξ|>ci2m|\langle e_{m}^{2l+s},\xi^{\prime}\rangle|>ci2^{-m}. Since ψ^\widehat{\psi} is a Schwartz function, so ψ^\nabla\widehat{\psi} is smooth and rapidly decays at infinity. We finally conclude that

LHS(3.11)j1i2m122j|ξ|2min{1,(2j|ξ|i)100}1i2m1i2j[i2j|ξ|]2[1+(2j|ξ|i)]100C\begin{split}\text{LHS}\eqref{e:21multi}&\leq\sum_{j\in{\mathbb{Z}}}\sum_{1\leq i\leq 2^{m-1}}2^{2j}|\xi|^{2}\min\{1,(2^{j}|\xi|i)^{-100}\}\\ &\lesssim\sum_{1\leq i\leq 2^{m-1}}i^{-2}\sum_{j\in{\mathbb{Z}}}[i2^{j}|\xi|]^{2}[1+(2^{j}|\xi|i)]^{-100}\leq C\end{split}

which proves (3.10). For the term |B2jl,2m(rl)||B_{2^{j}}^{l,2^{m}}(r_{l})^{*}|, we can also obtain

sup0l<2m1j|B2jl,2m(rl)|L2(𝒩)fL2(𝒩)\Big{\|}\sup_{0\leq l<2^{m-1}\atop j\in{\mathbb{Z}}}|B_{2^{j}}^{l,2^{m}}(r_{l})^{*}|\Big{\|}_{L_{2}({\mathcal{N}})}\lesssim\|f\|_{L_{2}({\mathcal{N}})}

by using the same argument if observing that

[B2jl,2m(rl)](ξ)=[ψ^(2j+mem2l+1,ξ)ψ^(2j+mem2l,ξ)]χΓ2l+1cΓ2lc(ξ)f^(ξ).{\mathcal{F}}[B_{2^{j}}^{l,2^{m}}(r_{l})^{*}](\xi)=\Big{[}\widehat{\psi}\Big{(}-2^{j+m}\langle e_{m}^{2l+1},\xi\rangle\Big{)}-\widehat{\psi}\Big{(}-2^{j+m}\langle e_{m}^{2l},\xi\rangle\Big{)}\Big{]}\chi_{\Gamma_{2l+1}^{c}\cap\Gamma_{2l}^{c}}(-\xi)\widehat{f}(\xi).

Combining these estimates of |B2jl,2m(rl)||B_{2^{j}}^{l,2^{m}}(r_{l})| and |B2jl,2m(rl)||B_{2^{j}}^{l,2^{m}}(r_{l})^{*}|, we get II3II_{3} is majorized by fL2(𝒩)\|f\|_{L_{2}({\mathcal{N}})} which ends the proof this theorem. ∎

4. Bochner-Riesz means on L(2)¯L_{\infty}({\mathbb{R}}^{2})\overline{\otimes}{\mathcal{M}}

In this section, we study the operator-valued Bochner-Riesz means on Lp(𝒩)L_{p}({\mathcal{N}}), where 𝒩=L(2)¯{\mathcal{N}}=L_{\infty}({\mathbb{R}}^{2})\overline{\otimes}{\mathcal{M}} throughout this section. The main tools are noncommutative Kakeya maximal functions which have been investigated in the previous section. Our main result can be stated as follows.

Theorem 4.1.

Let 0<λ120<\lambda\leq\frac{1}{2} and 43+2λ<p<412λ\frac{4}{3+2\lambda}<p<\frac{4}{1-2\lambda}. Then for the Bochner-Riesz means BRλB^{\lambda}_{R} given in (1.2), set Bλ=B1λB^{\lambda}=B_{1}^{\lambda}, we have

(4.1) Bλ(f)Lp(𝒩)fLp(𝒩),supR>0BRλ(f)Lp(𝒩)fLp(𝒩).\|B^{\lambda}(f)\|_{L_{p}({\mathcal{N}})}\lesssim\|f\|_{L_{p}({\mathcal{N}})},\quad\sup_{R>0}\|B_{R}^{\lambda}(f)\|_{L_{p}({\mathcal{N}})}\lesssim\|f\|_{L_{p}({\mathcal{N}})}.

Consequently for fLp(𝒩)f\in L_{p}({\mathcal{N}}), BRλ(f)B_{R}^{\lambda}(f) converges to ff in Lp(𝒩)L_{p}({\mathcal{N}}) as RR\rightarrow\infty.

Before giving the proof, let us start with some definitions and lemmas. We first define the following Fourier multiplier on Lp(𝒩)L_{p}({\mathcal{N}}) for convenience.

Definition 4.2.

We say m:2m:{\mathbb{R}}^{2}\rightarrow{\mathbb{C}} is an Lp(𝒩)L_{p}({\mathcal{N}}) Fourier multiplier if the operator TmT_{m} defined by

Tm(f)^(ξ)=m(ξ)f^(ξ),f𝐒(2)𝒮,\widehat{T_{m}(f)}(\xi)=m(\xi)\widehat{f}(\xi),\quad f\in\mathbf{S}({\mathbb{R}}^{2})\otimes\mathcal{S},

extends to a bounded operator on Lp(𝒩)L_{p}({\mathcal{N}}), where 𝐒(2)\mathbf{S}({\mathbb{R}}^{2}) is the class of Schwartz functions on 2{\mathbb{R}}^{2} and 𝒮\mathcal{S} is the linear span of all x+x\in{\mathcal{M}}_{+} whose support projections have finite trace defined in Subsection 2.1. Denote by Mp(𝒩)M_{p}({\mathcal{N}}) the space of all Lp(𝒩)L_{p}({\mathcal{N}}) multipliers and Mp(𝒩)\|\cdot\|_{M_{p}({\mathcal{N}})} the LpL_{p} multiplier norm.

Notice that for Reλ>12\text{Re}\lambda>\frac{1}{2}, the convolution kernel associated to (1||2)+λ(1-|\cdot|^{2})^{\lambda}_{+} is integrable over 2{\mathbb{R}}^{2} with the bound e6|Imλ|2e^{6|\text{Im}\lambda|^{2}} (see e.g. [18, Proposition 5.2.2]). Then we immediately get the following lemma.

Lemma 4.3.

Let λ\lambda\in{\mathbb{C}}. If Reλ>12\text{Re}\lambda>\frac{1}{2}, then for all 1p1\leq p\leq\infty, we have

(1||2)+λMp(𝒩)e6|Imλ|2.\|(1-|\cdot|^{2})_{+}^{\lambda}\|_{M_{p}({\mathcal{N}})}\lesssim e^{6|\text{Im}\lambda|^{2}}.

Next we introduce a noncommutative Littlewood-Paley-Rubio de Francia’s square function inequality for equal intervals whose proof can be found in [25]. This inequality is also a key step in our study of Bochner-Riesz means.

Lemma 4.4.

Set 𝒩=L()¯{\mathcal{N}}=L_{\infty}({\mathbb{R}})\overline{\otimes}{\mathcal{M}}. Let IjI_{j}s be intervals of equal length with disjoint interior, jj\in{\mathbb{Z}} and jIj=\bigcup_{j\in{\mathbb{Z}}}I_{j}={\mathbb{R}}. Define Pjf^(ξ)=χIj(ξ)f^(ξ)\widehat{P_{j}f}(\xi)=\chi_{I_{j}}(\xi)\widehat{f}(\xi). Then for all 2p<2\leq p<\infty, we have

{Pj(f)}Lp(𝒩;2rc)fLp(𝒩).\|\{P_{j}(f)\}\|_{L_{p}({\mathcal{N}};\ell_{2}^{rc})}\lesssim\|f\|_{L_{p}({\mathcal{N}})}.

Now we begin to prove Theorem 4.1.

Proof of Theorem 4.1.

We first point out the norm convergence in Lp(𝒩)L_{p}({\mathcal{N}}) is a direct consequence of (4.1). In fact for any f𝐒(2)𝒮f\in\mathbf{S}({\mathbb{R}}^{2})\otimes\mathcal{S}, i.e. f(x)=i=1nψi(x)aif(x)=\sum_{i=1}^{n}\psi_{i}(x)a_{i} with ψi𝐒(2)\psi_{i}\in\mathbf{S}({\mathbb{R}}^{2}) and ai𝒮a_{i}\in\mathcal{S}, BRλ(f)B_{R}^{\lambda}(f) converges to ff in Lp(𝒩)L_{p}({\mathcal{N}}) as RR\rightarrow\infty. Then by the density argument and the second inequality in (4.1), we could get for every fLp(𝒩)f\in L_{p}({\mathcal{N}}), BRλ(f)B_{R}^{\lambda}(f) converges to ff in Lp(𝒩)L_{p}({\mathcal{N}}). So in our proof below, it is sufficient to consider (4.1). Since the multiplier in Mp(𝒩)M_{p}({\mathcal{N}}) is invariant under the dilation:

m(R1)Mp(𝒩)=mMp(𝒩),\|m(R^{-1}\cdot)\|_{M_{p}({\mathcal{N}})}=\|m\|_{M_{p}({\mathcal{N}})},

we only need to consider the first inequality in (4.1). Throughout the proof, we suppose that λ\lambda is a complex number with Reλ>0\text{Re}\lambda>0. When p=2p=2, the estimate just follows from vector-valued Plancherel’s theorem (2.6). By the duality, it is enough to show (4.1) holds for 2<p<412Reλ2<p<\frac{4}{1-2\text{Re}\lambda}.

We first make a dyadic decomposition of the multiplier. To do that, we choose a smooth function ϕ\phi supported in [12,12][-\frac{1}{2},\frac{1}{2}] and a smooth function ψ\psi supported in [18,58][\frac{1}{8},\frac{5}{8}] such that

ϕ(t)+k=0ψ(2k(1t))=1,t[0,1).\phi(t)+\sum_{k=0}^{\infty}\psi(2^{k}(1-t))=1,\quad\forall t\in[0,1).

Using this equality, we can decompose the multiplier (1|ξ|2)+λ(1-|\xi|^{2})_{+}^{\lambda} as follows

(4.2) (1|ξ|2)+λ=[ϕ(|ξ|)+k=0ψ(2k(1|ξ|))](1|ξ|2)+λ=ϕ(|ξ|)(1|ξ|2)λ+k=02kλ(2k(1|ξ|))λψ(2k(1|ξ|))(1+|ξ|)λ=:m00(ξ)+k=02kλmk(ξ).\begin{split}(1-|\xi|^{2})_{+}^{\lambda}&=\Big{[}\phi(|\xi|)+\sum_{k=0}^{\infty}\psi(2^{k}(1-|\xi|))\Big{]}(1-|\xi|^{2})^{\lambda}_{+}\\ &=\phi(|\xi|)(1-|\xi|^{2})^{\lambda}+\sum_{k=0}^{\infty}2^{-k\lambda}(2^{k}(1-|\xi|))^{\lambda}\psi(2^{k}(1-|\xi|))(1+|\xi|)^{\lambda}\\ &=:m_{00}(\xi)+\sum_{k=0}^{\infty}2^{-k\lambda}m_{k}(\xi).\end{split}

Below we give some observations of Fourier multipliers m0,0m_{0,0} and mkm_{k}. It is easy to see that m00m_{00} is a smooth function with compact support, hence the multiplier m00m_{00} lies in Mp(𝒩)M_{p}({\mathcal{N}}) for all 1p1\leq p\leq\infty. Notice that each function mkm_{k} is also smooth, radial and supported in a small annulus:

182k1|ξ|582k.\frac{1}{8}2^{-k}\leq 1-|\xi|\leq\frac{5}{8}2^{-k}.

Therefore each mkm_{k} also lies in Mp(𝒩)M_{p}({\mathcal{N}}) for 1p1\leq p\leq\infty, but its bound may depend on kk. To sum over the series in (4.2), it is crucial to determine exactly the bound of multiplier mkm_{k} in kk. Our main goal in this proof is to show that for each kk,

(4.3) mkM4(𝒩)(1+|k|)1/2(1+|λ|)3.\|m_{k}\|_{M_{4}({\mathcal{N}})}\lesssim(1+|k|)^{1/2}(1+|\lambda|)^{3}.

Suppose we have (4.3) for the moment. Then summing over the series in (4.2) with these estimates in (4.3), we get

Bλ(f)L4(𝒩)1[m00]fL4(𝒩)+k=02kReλ1[mk]fL4(𝒩)fL4(𝒩)+(1+|λ|)3fL4(𝒩)(1+|λ|)3fL4(𝒩).\begin{split}\|B^{\lambda}(f)\|_{L_{4}({\mathcal{N}})}&\lesssim\|{\mathcal{F}}^{-1}[{m_{00}}]*f\|_{L_{4}({\mathcal{N}})}+\sum_{k=0}^{\infty}2^{-k\text{Re}\lambda}\|{\mathcal{F}}^{-1}[m_{k}]*f\|_{L_{4}({\mathcal{N}})}\\ &\lesssim\|f\|_{L_{4}({\mathcal{N}})}+(1+|\lambda|)^{3}\|f\|_{L_{4}({{\mathcal{N}}})}\lesssim(1+|\lambda|)^{3}\|f\|_{L_{4}({\mathcal{N}})}.\end{split}

Applying the noncommutative Riesz-Thorin interpolation theorem (see [62]), we get (4.1) for 2<p<42<p<4. Next we use the analytic interpolation theorem (see Theorem A.1 in the appendix) to show the remaining part. Let ε>0\varepsilon>0 be a small constant (for example less than 12\frac{1}{2}). Then

Bλ(f)L4(𝒩)\displaystyle\|B^{\lambda}(f)\|_{L_{4}({\mathcal{N}})} (1+|Imλ|)3fL4(𝒩),\displaystyle\lesssim(1+|\text{Im}\lambda|)^{3}\|f\|_{L_{4}({\mathcal{N}})}, Reλ\displaystyle\text{Re}\lambda =ε;\displaystyle=\varepsilon;
Bλ(f)L(𝒩)\displaystyle\|B^{\lambda}(f)\|_{L_{\infty}({\mathcal{N}})} e6|Imλ|2fL(𝒩),\displaystyle\lesssim e^{6|\text{Im}\lambda|^{2}}\|f\|_{L_{\infty}({\mathcal{N}})}, Reλ\displaystyle\text{Re}\lambda =12+ε,\displaystyle=\frac{1}{2}+\varepsilon,

where the second inequality just follows from Lemma 4.3. Define a new operator TzT_{z} by

Tz(f)=B12z+ε(f).T_{z}(f)=B^{\frac{1}{2}z+\varepsilon}(f).

To apply Theorem A.1, we should verify the hypothesis of Theorem A.1. It is easy to see that zTzz\rightarrow T_{z} is an analytic family of linear operators with admissible growth and

Tiy(f)L4(𝒩)C1(1+|y|)3fL4(𝒩),T1+iy(f)L(𝒩)C2e32|y|2fL(𝒩).\begin{split}\|T_{iy}(f)\|_{L_{4}({\mathcal{N}})}&\leq C_{1}(1+|y|)^{3}\|f\|_{L_{4}({\mathcal{N}})},\\ \|T_{1+iy}(f)\|_{L_{\infty}({\mathcal{N}})}&\leq C_{2}e^{\frac{3}{2}|y|^{2}}\|f\|_{L_{\infty}({\mathcal{N}})}.\end{split}

Let pp and θ\theta satisfy 1p=1θ4+θ\frac{1}{p}=\frac{1-\theta}{4}+\frac{\theta}{\infty} and 12θ+ε=Reλ\frac{1}{2}\theta+\varepsilon=\text{Re}\lambda. Set M0(y)=C1(1+|y|)3M_{0}(y)=C_{1}(1+|y|)^{3} and M1(y)=C2e32|y|2M_{1}(y)=C_{2}e^{\frac{3}{2}|y|^{2}}. Then applying Theorem A.1, we get

Tθ(f)Lp(𝒩)M(θ)fLp(𝒩),\|T_{\theta}(f)\|_{L_{p}({\mathcal{N}})}\leq M(\theta)\|f\|_{L_{p}({\mathcal{N}})},

where M(θ)M(\theta) is a finite constant defined in Theorem A.1. This immediately implies our required estimate (4.1) since BReλ(f)=Tθ(f)B^{\text{Re}\lambda}(f)=T_{\theta}(f) and p=412Reλ+2ε<412Reλp=\frac{4}{1-2\text{Re}\lambda+2\varepsilon}<\frac{4}{1-2\text{Re}\lambda}.

Now we turn to prove (4.3). Fix kk in the rest of proof. To estimate the L4L_{4} norm of the Fourier multiplier mkm_{k}, we need an additional decomposition (the called microlocal decomposition) of mkm_{k} that takes into the radial nature of mkm_{k}. We usually identify the plane 2{\mathbb{R}}^{2} as complex plane {\mathbb{C}} by (x,y)z=x+iy(x,y)\leftrightarrow z=x+iy. The reason to do this is that the expression z=re2πiθz=re^{2\pi i\theta} in {\mathbb{C}} can be easily understood in the geometric point (note that rr is length of |z||z| and θ\theta is an argument). Next we define sectorial arcs as follows:

Γk,l={re2πiθ:(r,θ)+×[0,1),|θl2k2|<2k2,1582kr1182k},\Gamma_{k,l}=\{re^{2\pi i\theta}:\ (r,\theta)\in{\mathbb{R}}_{+}\times[0,1),|\theta-l2^{-\frac{k}{2}}|<2^{-\frac{k}{2}},1-\frac{5}{8}2^{-k}\leq r\leq 1-\frac{1}{8}2^{-k}\},

for all l{0,1,2,,2k/2}l\in\{0,1,2,\cdots,\lfloor 2^{k/2}\rfloor\}. We choose a smooth function ω\omega such that ω(u)=1\omega(u)=1 for |u|<14|u|<\frac{1}{4}, ω(u)=0\omega(u)=0 for |u|>1|u|>1 and lω(xl)=1\sum_{l\in{\mathbb{Z}}}\omega(x-l)=1 holds for all xx\in{\mathbb{R}}. Define

mk,l(re2πiθ)=mk(re2πiθ)ω(2k/2θl).m_{k,l}(re^{2\pi i\theta})=m_{k}(re^{2\pi i\theta})\omega(2^{k/2}\theta-l).

It should be pointed out that mk(re2πiθ)m_{k}(re^{2\pi i\theta}) is a function about variable rr. Then by our construction of ω\omega, we see that

mk(ξ)=lmk,l(ξ)=l=0r(k)mk,l(ξ),m_{k}(\xi)=\sum_{l\in{\mathbb{Z}}}m_{k,l}(\xi)=\sum_{l=0}^{r(k)}m_{k,l}(\xi),

where r(k)=2k/21r(k)=\lfloor 2^{k/2}\rfloor-1 if kk is even and r(k)=2k/2r(k)=\lfloor 2^{k/2}\rfloor if kk is odd. By some elementary calculations, it is not difficult to get that

(4.4) |rαθβmk,l(re2πiθ)|(1+|λ|)α2kα2k2β,suppmk,l(re2πiθ)Γk,l.|\partial_{r}^{\alpha}\partial_{\theta}^{\beta}m_{k,l}(re^{2\pi i\theta})|\lesssim(1+|\lambda|)^{\alpha}2^{k\alpha}2^{\frac{k}{2}\beta},\quad{\text{\rm supp}}\ m_{k,l}(re^{2\pi i\theta})\subset\Gamma_{k,l}.

Next we split all {mk,l}l\{m_{k,l}\}_{l} into five subsets whose supports satisfy the following conditions:

  1.   (a).

    suppmk,lQa=:{(x,y)2:x>0,|y|<|x|};{\text{\rm supp}}m_{k,l}\subsetneq Q_{a}=:\{(x,y)\in{\mathbb{R}}^{2}:x>0,|y|<|x|\};

  2.   (b).

    suppmk,lQb=:{(x,y)2:x<0,|y|<|x|};{\text{\rm supp}}m_{k,l}\subsetneq Q_{b}=:\{(x,y)\in{\mathbb{R}}^{2}:x<0,|y|<|x|\};

  3.   (c).

    suppmk,lQc=:{(x,y)2:y>0,|y|>|x|};{\text{\rm supp}}m_{k,l}\subsetneq Q_{c}=:\{(x,y)\in{\mathbb{R}}^{2}:y>0,|y|>|x|\};

  4.   (d).

    suppmk,lQd=:{(x,y)2:y<0,|y|>|x|};{\text{\rm supp}}m_{k,l}\subsetneq Q_{d}=:\{(x,y)\in{\mathbb{R}}^{2}:y<0,|y|>|x|\};

  5.   (e).

    The support suppmk,l{\text{\rm supp}}m_{k,l} intersects Qe={(x,y)2:|x|=|y|}.Q_{e}=\{(x,y)\in{\mathbb{R}}^{2}:|x|=|y|\}.

We first observe that there are only at most eight mk,lm_{k,l}s in the case (e). In such a case, it is straightforward to get that

(4.5) mk,lMp(𝒩)(1+|λ|)3\|m_{k,l}\|_{M_{p}({\mathcal{N}})}\lesssim(1+|\lambda|)^{3}

as we will see below (which will be pointed out in our later proof). By symmetry, we only need to concentrate on the case (a). Denote the index set ={l:Γk,lQa}\mathfrak{I}=\{l:\ \Gamma_{k,l}\subsetneq Q_{a}\}. For each ll\in\mathfrak{I}, define a Fourier multiplier operator Tk,lT_{k,l} by

Tk,lg^(ξ)=mk,l(ξ)g^(ξ).\widehat{T_{k,l}g}(\xi)={m_{k,l}}(\xi)\widehat{g}(\xi).

Our purpose below is to show lmk,l\sum_{l\in\mathfrak{I}}m_{k,l} satisfies (4.3), i.e.

(4.6) lTk,l(f)L4(𝒩)(1+|k|)12(1+|λ|)3fL4(𝒩).\Big{\|}\sum_{l\in\mathfrak{I}}T_{k,l}(f)\Big{\|}_{L_{4}({\mathcal{N}})}\lesssim(1+|k|)^{\frac{1}{2}}(1+|\lambda|)^{3}\|f\|_{L_{4}({\mathcal{N}})}.

By decomposing ff as linear combination of four positive functions, we may suppose that ff is positive. In the following we should separate the sum of ll\in\mathfrak{I} into two parts,

(4.7) lTk,l(f)L4(𝒩)4=τ2|llTk,l(f)(x)Tk,l(f)(x)|2𝑑xτ2|ll,|ll|103Tk,l(f)(x)Tk,l(f)(x)|2𝑑x+τ2|ll,|ll|>103Tk,l(f)(x)Tk,l(f)(x)|2𝑑x=:I+II.\begin{split}\Big{\|}\sum_{l\in\mathfrak{I}}T_{k,l}(f)\Big{\|}_{L_{4}({\mathcal{N}})}^{4}&=\tau\int_{{\mathbb{R}}^{2}}|\sum_{l\in\mathfrak{I}}\sum_{l^{\prime}\in\mathfrak{I}}T_{k,l}(f)^{*}(x)T_{k,l^{\prime}}(f)(x)|^{2}dx\\ &\lesssim\tau\int_{{\mathbb{R}}^{2}}|\sum_{l\in\mathfrak{I}}\sum_{l^{\prime}\in\mathfrak{I},|l-l^{\prime}|\leq 10^{3}}T_{k,l}(f)^{*}(x)T_{k,l^{\prime}}(f)(x)|^{2}dx\\ &\quad+\tau\int_{{\mathbb{R}}^{2}}|\sum_{l\in\mathfrak{I}}\sum_{l^{\prime}\in\mathfrak{I},|l-l^{\prime}|>10^{3}}T_{k,l}(f)^{*}(x)T_{k,l^{\prime}}(f)(x)|^{2}dx\\ &=:I+II.\end{split}

For the term II, the sum of ll^{\prime} is taking over l103ll+103l-10^{3}\leq l^{\prime}\leq l+10^{3} which is finite. By Hölder’s inequality of the square function in Lemma 2.5, we get

Ii=103103τ2|l,l+iTk,l(f)(x)Tk,l+i(f)(x)|2𝑑xi=103103(l|Tk,l(f)|2)12L4(𝒩)2(l+i|Tk,l+i(f)|2)12L4(𝒩)2(l|Tk,l(f)|2)12L4(𝒩)4.\begin{split}I&\lesssim\sum_{i=-10^{3}}^{10^{3}}\tau\int_{{\mathbb{R}}^{2}}|\sum_{l\in\mathfrak{I},l+i\in\mathfrak{I}}T_{k,l}(f)^{*}(x)T_{k,l+i}(f)(x)|^{2}dx\\ &\leq\sum_{i=-10^{3}}^{10^{3}}\Big{\|}\Big{(}\sum_{l\in\mathfrak{I}}|T_{k,l}(f)|^{2}\Big{)}^{\frac{1}{2}}\Big{\|}^{2}_{L_{4}({\mathcal{N}})}\Big{\|}\Big{(}\sum_{l+i\in\mathfrak{I}}|T_{k,l+i}(f)|^{2}\Big{)}^{\frac{1}{2}}\Big{\|}^{2}_{L_{4}({\mathcal{N}})}\\ &\lesssim\Big{\|}\Big{(}\sum_{l\in\mathfrak{I}}|T_{k,l}(f)|^{2}\Big{)}^{\frac{1}{2}}\Big{\|}^{4}_{L_{4}({\mathcal{N}})}.\end{split}

For the term IIII, using vector-valued Plancherel’s theorem (2.6), we get

II=τ2|ll,|ll|>103Tk,l(f)^Tk,l(f)^(ξ)|2𝑑ξ.II=\tau\int_{{\mathbb{R}}^{2}}\big{|}\sum_{l\in\mathfrak{I}}\sum_{l^{\prime}\in\mathfrak{I},|l-l^{\prime}|>10^{3}}\widehat{T_{k,l}(f)^{*}}*\widehat{T_{k,l^{\prime}}(f)}(\xi)\big{|}^{2}d\xi.

Since ff is positive, it is easy to see Tk,l(f)=1[mk,l~]fT_{k,l}(f)^{*}={\mathcal{F}}^{-1}[{\widetilde{m_{k,l}}}]*f where we use the notation a~()=a()\tilde{a}(\cdot)=a(-\cdot). Then we have

suppTk,l(f)^Γ~k,l,suppTk,l(f)^Γk,l,{\text{\rm supp}}\widehat{T_{k,l}(f)^{*}}\subset\tilde{\Gamma}_{k,l},\quad{\text{\rm supp}}\widehat{T_{k,l^{\prime}}(f)}\subset\Gamma_{k,l^{\prime}},

where Γ~k,l={x:xΓk,l}\tilde{\Gamma}_{k,l}=\{x:-x\in\Gamma_{k,l}\}. Applying these support conditions, we rewrite the above integral of IIII as

τ2|ll,|ll|>103Tk,l(f)^Tk,l(f)^(ξ)χΓ~k,l+Γk,l(ξ)|2𝑑ξ.\tau\int_{{\mathbb{R}}^{2}}\big{|}\sum_{l\in\mathfrak{I}}\sum_{l^{\prime}\in\mathfrak{I},|l-l^{\prime}|>10^{3}}\widehat{T_{k,l}(f)^{*}}*\widehat{T_{k,l^{\prime}}(f)}(\xi)\chi_{\tilde{\Gamma}_{k,l}+{\Gamma}_{k,l^{\prime}}}(\xi)\big{|}^{2}d\xi.

Next using the convexity operator inequality (2.5), this term is bounded by

τ2(ll,|ll|>103|Tk,l(f)^Tk,l(f)^(ξ)|2)(ll,|ll|>103χΓ~k,l+Γk,l(ξ))𝑑ξ.\tau\int_{{\mathbb{R}}^{2}}\Big{(}\sum_{l\in\mathfrak{I}}\sum_{l^{\prime}\in\mathfrak{I},|l-l^{\prime}|>10^{3}}|\widehat{T_{k,l}(f)^{*}}*\widehat{T_{k,l^{\prime}}(f)}(\xi)|^{2}\Big{)}\Big{(}\sum_{l\in\mathfrak{I}}\sum_{l^{\prime}\in\mathfrak{I},|l-l^{\prime}|>10^{3}}\chi_{\tilde{\Gamma}_{k,l}+{\Gamma}_{k,l^{\prime}}}(\xi)\Big{)}d\xi.

Before proceeding our proof further, we need the following geometric estimate.

Lemma 4.5.

There exists a constant CC independent of kk and ξ\xi such that

ll,|ll|>103χΓ~k,l+Γk,l(ξ)C.\sum_{l\in\mathfrak{I}}\sum_{l^{\prime}\in\mathfrak{I},|l-l^{\prime}|>10^{3}}\chi_{\tilde{\Gamma}_{k,l}+{\Gamma}_{k,l^{\prime}}}(\xi)\leq C.
Remark 4.6.

The geometric estimate of this lemma is different from the classical commutative one stated in [10]. The main difference is that there is no involution * of Tk,lfT_{k,l}f in the commutative case because of |a|4=|aa|2|a|^{4}=|aa|^{2} if aa is a complex number. This yields a geometric estimate in the form llχΓk,l+Γk,l(ξ)\sum_{l\in\mathfrak{I}}\sum_{l^{\prime}\in\mathfrak{I}}\chi_{\Gamma_{k,l}+\Gamma_{k,l^{\prime}}}(\xi) which is simpler than that of Lemma 4.5. In fact, all Γk,l\Gamma_{k,l}s are contained in {(x,y):x>0,|x|<|y|}\{(x,y):x>0,|x|<|y|\} and the pieces Γk,l+Γk,l\Gamma_{k,l}+\Gamma_{k,l^{\prime}}s are well distributed. However in the noncommutative setting, for an element aa\in{\mathcal{M}}, we have |a|4=|aa|2|a|^{4}=|a^{*}a|^{2}. Our argument above leads to the geometric estimate llχΓ~k,l+Γk,l(ξ)\sum_{l\in\mathfrak{I}}\sum_{l^{\prime}\in\mathfrak{I}}\chi_{\tilde{\Gamma}_{k,l}+\Gamma_{k,l^{\prime}}}(\xi) . Notice Γ~k,l\tilde{\Gamma}_{k,l} stays in an oppositive direction of Γk,l\Gamma_{k,l}, i.e. Γ~k,l{(x,y):x<0,|x|<|y|}\tilde{\Gamma}_{k,l}\subset\{(x,y):x<0,|x|<|y|\}. These pieces Γ~k,l+Γk,l\tilde{\Gamma}_{k,l}+\Gamma_{k,l^{\prime}}s may accumulate near the origin if Γk,l\Gamma_{k,l} and Γk,l\Gamma_{k,l^{\prime}} are close enough, which can be easily seen in view of geometric observation. Hence if ξ\xi is close to the origin, llχΓ~k,l+Γk,l(ξ)\sum_{l\in\mathfrak{I}}\sum_{l^{\prime}\in\mathfrak{I}}\chi_{\tilde{\Gamma}_{k,l}+\Gamma_{k,l^{\prime}}}(\xi) may be infinite. This is the reason why we split the sum of \mathfrak{I} into two parts in (4.7). Luckily we can show {Γ~k,l+Γk,l}l,l\{\tilde{\Gamma}_{k,l}+\Gamma_{k,l^{\prime}}\}_{l,l^{\prime}} is finite overlapped if |ll|>103|l-l^{\prime}|>10^{3}.

The proof of Lemma 4.5 will be given later. Applying this geometric estimate, we get IIII is bounded by

llτ2|Tk,l(f)^Tk,l(f)^(ξ)|2dξ=llτ2|Tk,l(f)(x)Tk,l(f)(x)|2dx=llτ2Tk,l(f)(x)Tk,l(f)(x)Tk,l(f)(x)Tk,l(f)(x)𝑑x=τ2(l|Tk,l(f)(x)|2)2𝑑x=(l|Tk,l(f)|2)12L4(𝒩)4,\begin{split}\sum_{l\in\mathfrak{I}}\sum_{l^{\prime}\in\mathfrak{I}}\tau\int_{{\mathbb{R}}^{2}}|\widehat{T_{k,l}(f)^{*}}&*\widehat{T_{k,l^{\prime}}(f)}(\xi)|^{2}d\xi=\sum_{l\in\mathfrak{I}}\sum_{l^{\prime}\in\mathfrak{I}}\tau\int_{{\mathbb{R}}^{2}}|{T_{k,l}(f)^{*}(x)}{T_{k,l^{\prime}}(f)}(x)|^{2}dx\\ &=\sum_{l\in\mathfrak{I}}\sum_{l^{\prime}\in\mathfrak{I}}\tau\int_{{\mathbb{R}}^{2}}{T_{k,l^{\prime}}(f)^{*}(x)}{T_{k,l}(f)}(x){T_{k,l}(f)^{*}(x)}{T_{k,l^{\prime}}(f)}(x)dx\\ &=\tau\int_{{\mathbb{R}}^{2}}\Big{(}\sum_{l\in\mathfrak{I}}|{T_{k,l}(f)^{*}(x)}|^{2}\Big{)}^{2}dx=\Big{\|}\Big{(}\sum_{l\in\mathfrak{I}}|{T_{k,l}(f)^{*}}|^{2}\Big{)}^{\frac{1}{2}}\Big{\|}^{4}_{L_{4}({\mathcal{N}})},\end{split}

where in the first equality we use vector-value’s Plancherel theorem (2.6) and the third equality follows from the tracial property of τ\tau: τ(ab)=τ(ba)\tau(ab)=\tau(ba). Combining these estimates of II and IIII, together with (4.7), we get

(4.8) lTk,l(f)L4(𝒩)4(l|Tk,l(f)|2)12L4(𝒩)4+(l|Tk,l(f)|2)12L4(𝒩)4.\Big{\|}\sum_{l\in\mathfrak{I}}T_{k,l}(f)\Big{\|}_{L_{4}({\mathcal{N}})}^{4}\lesssim\Big{\|}\Big{(}\sum_{l\in\mathfrak{I}}|{T_{k,l}(f)}|^{2}\Big{)}^{\frac{1}{2}}\Big{\|}^{4}_{L_{4}({\mathcal{N}})}+\Big{\|}\Big{(}\sum_{l\in\mathfrak{I}}|{T_{k,l}(f)^{*}}|^{2}\Big{)}^{\frac{1}{2}}\Big{\|}^{4}_{L_{4}({\mathcal{N}})}.

This kind of estimate is consistent with the estimate of square functions in noncommutative analysis in the sense that we should use both row and column square functions to control the LpL_{p} norm for p2p\geq 2 (see Subsection 2.3).

Now we should study carefully the multiplier mk,lm_{k,l}. Consider mk,0m_{k,0} firstly. Note that mk,0m_{k,0} is supported in a rectangle parallel to axes with side length 2k12^{-k-1} and 2k2+12^{-\frac{k}{2}+1} (along the ξ1\xi_{1}-axis and ξ2\xi_{2}-axis, respectively). Roughly speaking in such a rectangle ξ1r,ξ2θ\partial_{\xi_{1}}\approx\partial_{r},\partial_{\xi_{2}}\approx\partial_{\theta}. In fact by some straightforward calculations, we get the smooth function mk,0m_{k,0} satisfies

(4.9) |ξ1αξ2βmk,0(ξ1,ξ2)|α,β(1+|λ|)α2kα2k2β|\partial_{\xi_{1}}^{\alpha}\partial_{\xi_{2}}^{\beta}m_{k,0}(\xi_{1},\xi_{2})|\lesssim_{\alpha,\beta}(1+|\lambda|)^{\alpha}2^{k\alpha}2^{\frac{k}{2}\beta}

for all positive integers α\alpha and β\beta which is an analogue of (4.4). We can rewrite (4.9) as follows

|ξ1αξ2β[mk,0(2kξ1,2k2ξ2)]|α,β(1+|λ|)α.|\partial_{\xi_{1}}^{\alpha}\partial_{\xi_{2}}^{\beta}[m_{k,0}(2^{-k}\xi_{1},2^{-\frac{k}{2}}\xi_{2})]|\lesssim_{\alpha,\beta}(1+|\lambda|)^{\alpha}.

With these estimates and using integration by parts, we obtain that

(4.10) 232k|1[mk,0](2kx1,2k2x2)|(1+|λ|)3(1+|x1|+|x2|)3.2^{\frac{3}{2}k}|{\mathcal{F}}^{-1}[m_{k,0}](2^{k}x_{1},2^{\frac{k}{2}}x_{2})|\lesssim(1+|\lambda|)^{3}(1+|x_{1}|+|x_{2}|)^{-3}.

Let vlv_{l} and vlv_{l}^{\bot} be the unit vectors corresponding the directions e2πil2k2e^{2\pi il2^{-\frac{k}{2}}} and ie2πil2k2ie^{2\pi il2^{-\frac{k}{2}}}, respectively. According our definition, mk,l(ξ)=mk,0(Aξ)m_{k,l}(\xi)=m_{k,0}(A\xi) where A=(vlvl)A=\Big{(}{v_{l}\atop v_{l}^{\bot}}\Big{)}. By a rotation, we see that

(4.11) 1[mk,l](x1,x2)=1[mk,0](Ax)=1[mk,0](xvl,xvl).{\mathcal{F}}^{-1}[m_{k,l}](x_{1},x_{2})={\mathcal{F}}^{-1}[m_{k,0}](Ax)={\mathcal{F}}^{-1}[m_{k,0}](x\cdot v_{l},x\cdot v_{l}^{\bot}).

Hence combining with (4.10), we have

|1[mk,l](x1,x2)|232k(1+|λ|)3(1+2k|xvl|+2k2|xvl|)3,|{\mathcal{F}}^{-1}[m_{k,l}](x_{1},x_{2})|\lesssim 2^{-\frac{3}{2}k}(1+|\lambda|)^{3}(1+2^{-k}|x\cdot v_{l}|+2^{-\frac{k}{2}}|x\cdot v_{l}^{\bot}|)^{-3},

which immediately implies that

(4.12) supk>0supl1[mk,l]L1(2)(1+|λ|)3.\sup_{k>0}\sup_{l\in\mathfrak{I}}\|{\mathcal{F}}^{-1}[m_{k,l}]\|_{L_{1}({\mathbb{R}}^{2})}\lesssim(1+|\lambda|)^{3}.

Here we point out that (4.5) follows from (4.12). Obviously all the above estimates hold for mk,lm_{k,l}^{*} since mk,l=mk,l¯m_{k,l}^{*}=\overline{m_{k,l}}.

We turn to give some geometric observations of the support of mk,lm_{k,l}. Let Jk,lJ_{k,l} be the ξ2\xi_{2} projection of the support of mk,lm_{k,l}. If the support of mk,lm_{k,l} lies in the upper half plane of QaQ_{a} (i.e. ξ2>0\xi_{2}>0), then we see

Jk,l=×[(1582k)sin(2π2k2(l1)),(1182k)sin(2π2k2(l+1))].J_{k,l}={\mathbb{R}}\times[(1-\tfrac{5}{8}2^{-k})\sin(2\pi 2^{-\frac{k}{2}}(l-1)),(1-\tfrac{1}{8}2^{-k})\sin(2\pi 2^{-\frac{k}{2}}(l+1))].

Similarly if the support of mk,lm_{k,l} lies in the lower half plane of QaQ_{a} (i.e. ξ20\xi_{2}\leq 0), then

Jk,l=×[(1182k)sin(2π2k2(l1)),(1582k)sin(2π2k2(l+1))].J_{k,l}={\mathbb{R}}\times[(1-\tfrac{1}{8}2^{-k})\sin(2\pi 2^{-\frac{k}{2}}(l-1)),(1-\tfrac{5}{8}2^{-k})\sin(2\pi 2^{-\frac{k}{2}}(l+1))].

Since kk is a fixed integer, the sets Jk,lJ_{k,l}s are almost disjoint for different ll\in\mathfrak{I} because Γk,l\Gamma_{k,l} is only joint with Γk,l1\Gamma_{k,l-1} and Γk,l+1\Gamma_{k,l+1}. Next our goal is to construct congruent strips containing Jk,lJ_{k,l}. If we set J~k,l\tilde{J}_{k,l} as a strip centered at ξ2=(1382k)sin(2π2k2l)\xi_{2}=(1-\frac{3}{8}2^{-k})\sin(2\pi 2^{-\frac{k}{2}}l) with width 202k220\cdot 2^{-\frac{k}{2}}, i.e.

J~k,l=×[(1382k)sin(2π2k2l)102k2,(1382k)sin(2π2k2l)+102k2].\tilde{J}_{k,l}={\mathbb{R}}\times[(1-\tfrac{3}{8}2^{-k})\sin(2\pi 2^{-\frac{k}{2}}l)-10\cdot 2^{-\frac{k}{2}},(1-\tfrac{3}{8}2^{-k})\sin(2\pi 2^{-\frac{k}{2}}l)+10\cdot 2^{-\frac{k}{2}}].

Then Jk,lJ~k,lJ_{k,l}\subset\tilde{J}_{k,l}. For any σ\sigma\in{\mathbb{Z}}, υ{0,1,,39}\upsilon\in\{0,1,\cdots,39\}, define the strips as follows

Sk,σ,υ=×[40σ2k2+υ2k2,40(σ+1)2k2+υ2k2].S_{k,\sigma,\upsilon}={\mathbb{R}}\times[40\sigma 2^{-\frac{k}{2}}+\upsilon 2^{-\frac{k}{2}},40(\sigma+1)2^{-\frac{k}{2}}+\upsilon 2^{-\frac{k}{2}}].

Note that Sk,σ,υS_{k,\sigma,\upsilon} has width 402k240\cdot 2^{-\frac{k}{2}} and each J~k,l\tilde{J}_{k,l} must be contained in one of Sk,σ,υS_{k,\sigma,\upsilon} for some σ\sigma\in{\mathbb{Z}} and υ{0,1,,39}\upsilon\in\{0,1,\cdots,39\} in view of some simple geometric observation of Sk,σ,υS_{k,\sigma,\upsilon}. We say J~k,lSk,σl,υl\tilde{J}_{k,l}\subset S_{k,\sigma_{l},\upsilon_{l}} and define Sk,σl,υl=Bk,lS_{k,\sigma_{l},\upsilon_{l}}=B_{k,l}. Let fk,l=1[χBk,lf^]=(1[χBk,l])ff_{k,l}={\mathcal{F}}^{-1}[\chi_{B_{k,l}}\widehat{f}]=({\mathcal{F}}^{-1}[\chi_{B_{k,l}}])*f. According the definition of Tk,lT_{k,l}, we see

Tk,l(f)=1[mk,l]fk,l=Tk,l(fk,l).T_{k,l}(f)={\mathcal{F}}^{-1}[m_{k,l}]*f_{k,l}=T_{k,l}(f_{k,l}).

Now we come back to the estimate in (4.8). We first consider the column square function. Using the convexity operator inequality (2.5) and the uniform boundedness of 1[mk,l]{\mathcal{F}}^{-1}[m_{k,l}] in (4.12), we get

|Tk,l(f)(x)|2=|21[mk,l](xy)fk,l(y)𝑑y|22|1[mk,l](xy)|𝑑y2|1[mk,l](xy)||fk,l(y)|2𝑑y(1+|λ|)3|1[mk,l]||fk,l|2(x).\begin{split}|T_{k,l}(f)(x)|^{2}&=\Big{|}\int_{{\mathbb{R}}^{2}}{\mathcal{F}}^{-1}[m_{k,l}](x-y)f_{k,l}(y)dy\Big{|}^{2}\\ &\leq\int_{{\mathbb{R}}^{2}}|{\mathcal{F}}^{-1}[m_{k,l}](x-y)|dy\cdot\int_{{\mathbb{R}}^{2}}|{\mathcal{F}}^{-1}[m_{k,l}](x-y)|\cdot|f_{k,l}(y)|^{2}dy\\ &\lesssim(1+|\lambda|)^{3}|{\mathcal{F}}^{-1}[m_{k,l}]|*|f_{k,l}|^{2}(x).\end{split}

Plugging the above inequality into the column square function in (4.8) and applying the monotone property (2.1), we get

(l|Tk,l(f)|2)124L4(𝒩)(1+|λ|)6τ2(l|1[mk,l]||fk,l|2(x))2dx=(1+|λ|)6supg[τ2l|1[mk,l]||fk,l|2(x)g(x)dx]2\begin{split}\Big{\|}\Big{(}\sum_{l\in\mathfrak{I}}|{T_{k,l}(f)}|^{2}&\Big{)}^{\frac{1}{2}}\Big{\|}^{4}_{L_{4}({\mathcal{N}})}\lesssim(1+|\lambda|)^{6}\tau\int_{{\mathbb{R}}^{2}}\Big{(}\sum_{l\in\mathfrak{I}}|{\mathcal{F}}^{-1}[m_{k,l}]|*|f_{k,l}|^{2}(x)\Big{)}^{2}dx\\ &\ \ \ \ =(1+|\lambda|)^{6}\sup_{g}\Big{[}\tau\int_{{\mathbb{R}}^{2}}\sum_{l\in\mathfrak{I}}|{\mathcal{F}}^{-1}[m_{k,l}]|*|f_{k,l}|^{2}(x)g(x)dx\Big{]}^{2}\end{split}

where in the last equality we use the duality [Lp(𝒩)]=Lp(𝒩)[L_{p}({\mathcal{N}})]^{*}=L_{p^{\prime}}({\mathcal{N}}) and the supremum is taken over all positive gg in L2(𝒩)L_{2}({\mathcal{N}}) with norm less than one. Continue the above estimate, we obtain

(1+|λ|)6supg(τl2|fk,l(y)|2|[mk,l]|g(y)𝑑y)2(1+|λ|)6supgsupl|[mk,l]|gL2(𝒩)2l|fk,l|2L2(𝒩)2=(1+|λ|)6supgsupl|[mk,l]|gL2(𝒩)2(l|fk,l|2)12L4(𝒩)4,\begin{split}&\quad(1+|\lambda|)^{6}\sup_{g}\Big{(}\tau\sum_{l\in\mathfrak{I}}\int_{{\mathbb{R}}^{2}}|f_{k,l}(y)|^{2}\cdot|{\mathcal{F}}[m_{k,l}]|*g(y)dy\Big{)}^{2}\\ &\leq(1+|\lambda|)^{6}\sup_{g}\|\sup_{l\in\mathfrak{I}}|{\mathcal{F}}[m_{k,l}]|*g\|_{L_{2}({\mathcal{N}})}^{2}\big{\|}\sum_{l\in\mathfrak{I}}|f_{k,l}|^{2}\big{\|}_{L_{2}({\mathcal{N}})}^{2}\\ &=(1+|\lambda|)^{6}\sup_{g}\|\sup_{l\in\mathfrak{I}}|{\mathcal{F}}[m_{k,l}]|*g\|_{L_{2}({\mathcal{N}})}^{2}\big{\|}(\sum_{l\in\mathfrak{I}}|f_{k,l}|^{2})^{\frac{1}{2}}\big{\|}_{L_{4}({\mathcal{N}})}^{4},\end{split}

where the first inequality follows from the duality in Lemma 2.3. If we can prove the following two estimates:

(4.13) supl|[mk,l]|gL2(𝒩)(1+|λ|)3(1+k)gL2(𝒩),\|\sup_{l\in\mathfrak{I}}|{\mathcal{F}}[m_{k,l}]|*g\|_{L_{2}({\mathcal{N}})}\lesssim(1+|\lambda|)^{3}(1+k)\|g\|_{L_{2}({\mathcal{N}})},
(4.14) (l|fk,l|2)12L4(𝒩)fL4(𝒩),\big{\|}(\sum_{l\in\mathfrak{I}}|f_{k,l}|^{2})^{\frac{1}{2}}\big{\|}_{L_{4}({\mathcal{N}})}\lesssim\|f\|_{L_{4}({\mathcal{N}})},

then we finally get that

(l|Tk,l(f)|2)12L4(𝒩)(1+|λ|3)(1+k)12fL4(𝒩),\Big{\|}\Big{(}\sum_{l\in\mathfrak{I}}|{T_{k,l}(f)}|^{2}\Big{)}^{\frac{1}{2}}\Big{\|}_{L_{4}({\mathcal{N}})}\lesssim(1+|\lambda|^{3})(1+k)^{\frac{1}{2}}\|f\|_{L_{4}({\mathcal{N}})},

which is the required estimate of (4.6). For the row square function, we can use the similar method to obtain the desired estimate

(l|Tk,l(f)|2)12L4(𝒩)(1+|λ|3)(1+k)12fL4(𝒩).\Big{\|}\Big{(}\sum_{l\in\mathfrak{I}}|{T_{k,l}(f)}^{*}|^{2}\Big{)}^{\frac{1}{2}}\Big{\|}_{L_{4}({\mathcal{N}})}\lesssim(1+|\lambda|^{3})(1+k)^{\frac{1}{2}}\|f\|_{L_{4}({\mathcal{N}})}.

In fact, Tk,l(f)=[mk,l]fk,lT_{k,l}(f)^{*}={\mathcal{F}}[m_{k,l}^{*}]*f_{k,l}^{*}. Then repeating these arguments for the column square function above, we see the proof is reduced to the following two inequalities:

(4.15) supl|[mk,l]|gL2(𝒩)(1+|λ|)3(1+k)gL2(𝒩),\|\sup_{l\in\mathfrak{I}}|{\mathcal{F}}[m_{k,l}^{*}]|*g\|_{L_{2}({\mathcal{N}})}\lesssim(1+|\lambda|)^{3}(1+k)\|g\|_{L_{2}({\mathcal{N}})},
(4.16) (l|fk,l|2)12L4(𝒩)fL4(𝒩).\big{\|}(\sum_{l\in\mathfrak{I}}|f_{k,l}^{*}|^{2})^{\frac{1}{2}}\big{\|}_{L_{4}({\mathcal{N}})}\lesssim\|f\|_{L_{4}({\mathcal{N}})}.

Below we give the proofs of (4.13), (4.15) and (4.14), (4.16). We first consider (4.13) and (4.15). Recall (4.10), [mk,0]{\mathcal{F}}[m_{k,0}] is integrable over 2{\mathbb{R}}^{2} and satisfies

232k|[mk,0](2kx1,2k2x2)|(1+|λ|)3(1+|x|)3(1+|λ|)3s=02s22sχ[2s,2s]×[2s,2s](x).2^{\frac{3}{2}k}|{\mathcal{F}}[m_{k,0}](2^{k}x_{1},2^{\frac{k}{2}}{x_{2}})|\lesssim\frac{(1+|\lambda|)^{3}}{(1+|x|)^{3}}\lesssim(1+|\lambda|)^{3}\sum_{s=0}^{\infty}\frac{2^{-s}}{2^{2s}}\chi_{[-2^{s},2^{s}]\times[-2^{s},2^{s}]}(x).

By making a change of variables, we immediately get

|[mk,0](x)|(1+|λ|)3s=02s1|Rs|χRs(x),|{\mathcal{F}}[m_{k,0}](x)|\lesssim(1+|\lambda|)^{3}\sum_{s=0}^{\infty}{2^{-s}}\frac{1}{|R_{s}|}\chi_{R_{s}}(x),

where Rs=[2s+k,2s+k]×[2s+k2,2s+k2]R_{s}=[-2^{s+k},2^{s+k}]\times[-2^{s+\frac{k}{2}},2^{s+\frac{k}{2}}]. Notice that a general function [mk,l]{\mathcal{F}}[m_{k,l}] is obtained from [mk,0]{\mathcal{F}}[m_{k,0}] by a rotation A=(vlvl)A=\Big{(}{v_{l}\atop v_{l}^{\bot}}\Big{)} (see (4.11)). Therefore we get

|[mk,l](x)|(1+|λ|)3s=02s1|Rs,l|χRs,l(x),|{\mathcal{F}}[m_{k,l}](x)|\lesssim(1+|\lambda|)^{3}\sum_{s=0}^{\infty}{2^{-s}}\frac{1}{|R_{s,l}|}\chi_{R_{s,l}}(x),

where Rs,lR_{s,l} is a rectangle with principal axes along the directions vlv_{l} and vlv_{l}^{\bot} with half side length 2s+k2^{s+k} and 2s+k22^{s+\frac{k}{2}}, respectively. Since gg is positive, we see that

|[mk,l]|g(x)(1+|λ|)3s=02s1|Rs,l|Rs,lg(xy)𝑑y,|{\mathcal{F}}[m_{k,l}]|*g(x)\lesssim(1+|\lambda|)^{3}\sum_{s=0}^{\infty}2^{-s}\frac{1}{|R_{s,l}|}\int_{R_{s,l}}g(x-y)dy,

where \lesssim should be understood as partial order in the positive cone of L2(𝒩)L_{2}({\mathcal{N}}). For convenience, we set MRs,lg(x)=|Rs,l|1Rs,lg(xy)𝑑yM_{R_{s,l}}g(x)={|R_{s,l}|^{-1}}\int_{R_{s,l}}g(x-y)dy. Notice that MRs,lg=M2k/2gM_{R_{s,l}}g=M_{\mathcal{R}_{2^{k/2}}}g the Kakeya average operator defined in (3.1). Now using the estimate of the Kakeya maximal function in Theorem 3.1, we get

supl|[mk,l]|gL2(𝒩)(1+|λ|)3s=02ssups,lMRs,lgL2(𝒩)(1+|λ|)3s=02ssupM2k/2gL2(𝒩)(1+|λ|)3(1+k)gL2(𝒩),\begin{split}\|\sup_{l\in\mathfrak{I}}|{\mathcal{F}}[m_{k,l}]|*g\|_{L_{2}({\mathcal{N}})}&\lesssim(1+|\lambda|)^{3}\sum_{s=0}^{\infty}2^{-s}\|\sup_{s\in\mathbb{N},l\in\mathfrak{I}}M_{R_{s,l}}g\|_{L_{2}({\mathcal{N}})}\\ &\lesssim(1+|\lambda|)^{3}\sum_{s=0}^{\infty}2^{-s}\|\sup M_{\mathcal{R}_{2^{k/2}}}g\|_{L_{2}({\mathcal{N}})}\\ &\lesssim(1+|\lambda|)^{3}(1+k)\|g\|_{L_{2}({\mathcal{N}})},\end{split}

which is the just required estimate of (4.13). The proof of (4.15) is similar, we omit the details here.

We turn to the proofs of (4.14) and (4.16). Recall the strips

Sk,σ,υ=×[40σ2k2+υ2k2,40(σ+1)2k2+υ2k2]S_{k,\sigma,\upsilon}={\mathbb{R}}\times[40\sigma 2^{-\frac{k}{2}}+\upsilon 2^{-\frac{k}{2}},40(\sigma+1)2^{-\frac{k}{2}}+\upsilon 2^{-\frac{k}{2}}]

which is defined for σ\sigma\in{\mathbb{Z}} and υ{0,1,,39}\upsilon\in\{0,1,\cdots,39\}. These strips have width 402k240\cdot 2^{-\frac{k}{2}} and each J~k,l\tilde{J}_{k,l} belongs to one of them, which we call Sk,σl,υl=Bk,lS_{k,\sigma_{l},\upsilon_{l}}=B_{k,l}.

The family {Bk,l}l\{B_{k,l}\}_{l\in\mathfrak{I}} may have duplicate sets, so we split it into 4040 subfamilies by placing Bk,lB_{k,l} into different subfamilies if the indices υl\upsilon_{l} and υl\upsilon_{l^{\prime}} are different. Thus we write the set \mathfrak{I} as =1240\mathfrak{I}=\mathfrak{I}^{1}\cup\mathfrak{I}^{2}\cup\cdots\cup\mathfrak{I}^{40} where elements in each i\mathfrak{I}^{i} are different.

We next observe that the multiplier operator

fk,l^=χBk,lf^\widehat{f_{k,l}}=\chi_{B_{k,l}}\widehat{f}

satisfies fk,l=(idx1Pl)ff_{k,l}=(id_{x_{1}}\otimes P_{l})f where idx1id_{x_{1}} is the identity operator in x1x_{1} variable and PlP_{l} is an operator with the multiplier χ{ξ2:(ξ1,ξ2)Bk,l}\chi_{\{\xi_{2}:(\xi_{1},\xi_{2})\in B_{k,l}\}}. Now using Lemma 4.4, we get for 2<p<2<p<\infty,

{fk,l}liLp(𝒩;2rc)={(idx1Pl)(f)}liLp(𝒩;2rc)=({Pl(f(x1,))}liLp(L()¯;2rc)p𝑑x1)1p(f(x1,)Lp(L()¯)p𝑑x1)1p=fLp(𝒩).\begin{split}\big{\|}\{f_{k,l}\}_{l\in\mathfrak{I}^{i}}\big{\|}_{L_{p}({\mathcal{N}};\ell_{2}^{rc})}&=\big{\|}\{(id_{x_{1}}\otimes P_{l})(f)\}_{l\in\mathfrak{I}^{i}}\big{\|}_{L_{p}({\mathcal{N}};\ell_{2}^{rc})}\\ &=\Big{(}\int_{{\mathbb{R}}}\big{\|}\{P_{l}(f(x_{1},\cdot))\}_{l\in\mathfrak{I}^{i}}\Big{\|}^{p}_{L_{p}(L_{\infty}({\mathbb{R}})\overline{\otimes}{\mathcal{M}};\ell_{2}^{rc})}dx_{1}\Big{)}^{\frac{1}{p}}\\ &\lesssim\Big{(}\int_{{\mathbb{R}}}\|f(x_{1},\cdot)\|^{p}_{L_{p}(L_{\infty}({\mathbb{R}})\overline{\otimes}{\mathcal{M}})}dx_{1}\Big{)}^{\frac{1}{p}}=\|f\|_{L_{p}({\mathcal{N}})}.\end{split}

Specially the case p=4p=4 is our required estimate for (4.14) and (4.16). Now we prove (4.14) and (4.16) for one i\mathfrak{I}^{i}. Since =1240\mathfrak{I}=\mathfrak{I}^{1}\cup\mathfrak{I}^{2}\cdots\cup\mathfrak{I}^{40}, the full forms of (4.14) and (4.16) just follow from Minkowski’s inequality. ∎

In the remain part of this section, we turn to Lemma 4.5. This lemma is stated in [13, Exercise 3.4] without a proof. For the sake of self-containment, we give a proof here.

Proof of Lemma 4.5.

We only need to show that for each integer kk, the following estimate

l|ll|>103χΓk,lΓk,l(ξ)C\sum_{l\in\mathfrak{I}}\sum_{|l-l^{\prime}|>10^{3}}\chi_{\Gamma_{k,l}-{\Gamma}_{k,l^{\prime}}}(\xi)\leq C

holds. If kk is finite, there exist only finite many pairs of sets Γk,lΓk,l\Gamma_{k,l}-\Gamma_{k,l^{\prime}} depending on kk and this lemma is trivial. Therefore we can assume that kk is a large integer. Set δ=2k\delta=2^{-k}. Then δ\delta is sufficient small. For simplicity, denote the set Γk,l\Gamma_{k,l} by Γl\Gamma_{l}. Elements of ΓlΓl\Gamma_{l}-\Gamma_{l^{\prime}} have the form

E:=re2πi(l+α)δ12re2πi(l+α)δ12,E:=re^{2\pi i(l+\alpha)\delta^{\frac{1}{2}}}-r^{\prime}e^{2\pi i(l^{\prime}+\alpha^{\prime})\delta^{\frac{1}{2}}},

where α,α[1,1]\alpha,\alpha^{\prime}\in[-1,1] and r,r[158δ,118δ]r,r^{\prime}\in[1-\frac{5}{8}\delta,1-\frac{1}{8}\delta]. We first give some geometric observations of ΓlΓl\Gamma_{l}-\Gamma_{l^{\prime}}. Set

(4.17) w(l,l):=e2πilδ12e2πilδ12=2sin(π(ll)δ12)ieπi(l+l)δ12,w(l,l^{\prime}):=e^{2\pi il\delta^{\frac{1}{2}}}-e^{2\pi il^{\prime}\delta^{\frac{1}{2}}}=2\sin(\pi(l-l^{\prime})\delta^{\frac{1}{2}})ie^{\pi i(l+l^{\prime})\delta^{\frac{1}{2}}},

then it is easy to see that the direction of w(l,l)w(l,l^{\prime}) is along ieπi(l+l)δ12ie^{\pi i(l+l^{\prime})\delta^{\frac{1}{2}}}. Next we rewrite EE as follows

w(l,l)+[re2πi(l+α)δ12re2πi(l+α)δ12rw(l,l)]+[(r1)w(l,l)+(rr)e2πi(l+α)δ12].w(l,l^{\prime})+\Big{[}re^{2\pi i(l+\alpha)\delta^{\frac{1}{2}}}-re^{2\pi i(l^{\prime}+\alpha^{\prime})\delta^{\frac{1}{2}}}-rw(l,l^{\prime})\Big{]}+\Big{[}(r-1)w(l,l^{\prime})+(r-r^{\prime})e^{2\pi i(l^{\prime}+\alpha^{\prime})\delta^{\frac{1}{2}}}\Big{]}.

By some elementary calculations, we further get the second term above equals to

rw(l,l)(e2πiαδ12+e2πiαδ1222)+r(e2πilδ12+e2πilδ12)(e2πiαδ12e2πiαδ122).rw(l,l^{\prime})(\frac{e^{2\pi i\alpha\delta^{\frac{1}{2}}}+e^{2\pi i\alpha^{\prime}\delta^{\frac{1}{2}}}-2}{2})+r(e^{2\pi il^{\prime}\delta^{\frac{1}{2}}}+e^{2\pi il\delta^{\frac{1}{2}}})(\frac{e^{2\pi i\alpha\delta^{\frac{1}{2}}}-e^{2\pi i\alpha^{\prime}\delta^{\frac{1}{2}}}}{2}).

Combining this equality, we see that EE equals

w(l,l)+ir(e2πilδ12+e2πilδ12)sin2παδ12sin2παδ122+r(e2πilδ12+e2πilδ12)cos2παδ12cos2παδ122+E(r,r)\begin{split}w(l,l^{\prime})&+ir(e^{2\pi il^{\prime}\delta^{\frac{1}{2}}}+e^{2\pi il\delta^{\frac{1}{2}}})\frac{\sin 2\pi\alpha\delta^{\frac{1}{2}}-\sin 2\pi\alpha^{\prime}\delta^{\frac{1}{2}}}{2}\\ &+r(e^{2\pi il^{\prime}\delta^{\frac{1}{2}}}+e^{2\pi il\delta^{\frac{1}{2}}})\frac{\cos 2\pi\alpha\delta^{\frac{1}{2}}-\cos 2\pi\alpha^{\prime}\delta^{\frac{1}{2}}}{2}+E(r,r^{\prime})\end{split}

where

E(r,r)=(r1)w(l,l)+(rr)e2πi(l+α)δ12+rw(l,l)cos2παδ12+cos2παδ1222+irw(l,l)sin2παδ12+sin2παδ122.\begin{split}E(r,r^{\prime})&=(r-1)w(l,l^{\prime})+(r-r^{\prime})e^{2\pi i(l^{\prime}+\alpha^{\prime})\delta^{\frac{1}{2}}}+rw(l,l^{\prime})\frac{\cos 2\pi\alpha\delta^{\frac{1}{2}}+\cos 2\pi\alpha^{\prime}\delta^{\frac{1}{2}}-2}{2}\\ &\quad+irw(l,l^{\prime})\frac{\sin 2\pi\alpha\delta^{\frac{1}{2}}+\sin 2\pi\alpha^{\prime}\delta^{\frac{1}{2}}}{2}.\end{split}

On the other hand, utilizing the following equality

e2πilδ12+e2πilδ12=2cos(π(ll)δ12)eπi(l+l)δ12,e^{2\pi il^{\prime}\delta^{\frac{1}{2}}}+e^{2\pi il\delta^{\frac{1}{2}}}=2\cos(\pi(l-l^{\prime})\delta^{\frac{1}{2}})e^{\pi i(l+l^{\prime})\delta^{\frac{1}{2}}},

we see that i(e2πilδ12+e2πilδ12)i(e^{2\pi il^{\prime}\delta^{\frac{1}{2}}}+e^{2\pi il\delta^{\frac{1}{2}}}) has the same direction as w(l,l)w(l,l^{\prime}). Thus, (e2πilδ12+e2πilδ12)(e^{2\pi il^{\prime}\delta^{\frac{1}{2}}}+e^{2\pi il\delta^{\frac{1}{2}}}) is perpendicular to w(l,l)w(l,l^{\prime}). Before proceeding further, we need the following inequalities which will be frequently used in later manipulations:

(4.18) (i):|sint||t|;(ii):|1cost||t|2/2;(iii):|sint|2|t|π,|t|π2.\begin{split}\rm{(i):\ |\sin t|\leq|t|;\quad(ii):\ |1-\cos t|\leq{|t|}^{2}/2;\quad(iii):\ |\sin t|\geq\frac{2|t|}{\pi},\forall|t|\leq\frac{\pi}{2}}.\end{split}

By the first and second inequalities in (4.18), it is straightforward to verify that the error term satisfies

|E(r,r)|54δ+δ2+4π2δ+4π2δ|ll|50δ+4π2δ|ll|.|E(r,r^{\prime})|\leq\frac{5}{4}\delta+\frac{\delta}{2}+4\pi^{2}\delta+4\pi^{2}\delta|l-l^{\prime}|\leq 50\delta+4\pi^{2}\delta|l-l^{\prime}|.

We therefore conclude that ΓlΓl\Gamma_{l}-\Gamma_{l^{\prime}} contains in a rectangle R(l,l)R(l,l^{\prime}) centered at w(l,l)w(l,l^{\prime}) with half-length

4πδ12+50δ+4π2δ|ll|<100δ124\pi\delta^{\frac{1}{2}}+50\delta+4\pi^{2}\delta|l-l^{\prime}|<100\delta^{\frac{1}{2}}

in the direction along w(l,l)w(l,l^{\prime}) since δ\delta is sufficient small and half-width

4π2δ+50δ+4π2δ|ll|<90|ll|δ<100δ124\pi^{2}\delta+50\delta+4\pi^{2}\delta|l-l^{\prime}|<90|l-l^{\prime}|\delta<100\delta^{\frac{1}{2}}

in the direction along iw(l,l)iw(l,l^{\prime}) since |ll|>103|l-l^{\prime}|>10^{3}. Moreover R(l,l)B(w(l,l),150δ12)R(l,l^{\prime})\subset B(w(l,l^{\prime}),150\delta^{\frac{1}{2}}) a disk centered at w(l,l)w(l,l^{\prime}) with radius 150δ12150\delta^{\frac{1}{2}}.

Our next goal is to show that for a fixed (l,l)(,)(l,l^{\prime})\in(\mathfrak{I},\mathfrak{I}) with |ll|>103|l-l^{\prime}|>10^{3}, there exist only finite pairs (m,m)(,)(m,m^{\prime})\in(\mathfrak{I},\mathfrak{I}) with |mm|>103|m-m^{\prime}|>10^{3} such that (ΓmΓm)(ΓlΓl)(\Gamma_{m}-\Gamma_{m^{\prime}})\cap(\Gamma_{l}-\Gamma_{l^{\prime}})\neq\emptyset. Once we prove this fact, the geometric estimate in Lemma 4.5 immediately follows from it. If |w(l,l)w(m,m)|>300δ12|w(l,l^{\prime})-w(m,m^{\prime})|>300\delta^{\frac{1}{2}}, then (ΓmΓm)(ΓlΓl)=(\Gamma_{m}-\Gamma_{m^{\prime}})\cap(\Gamma_{l}-\Gamma_{l^{\prime}})=\emptyset. Therefore if these two sets intersect, we get

(4.19) |w(l,l)w(m,m)|300δ12.|w(l,l^{\prime})-w(m,m^{\prime})|\leq 300\delta^{\frac{1}{2}}.

In the following we should give a lower bound for |w(l,l)w(m,m)||w(l,l^{\prime})-w(m,m^{\prime})|. Before that we need a fundamental inequality which could be found in [18, Exercise 5.2.2]:

(4.20) |r1eiθ1r2eiθ2|min(r1,r2)sin|θ1θ2|,|θ1|,|θ2|<π4.|r_{1}e^{i\theta_{1}}-r_{2}e^{i\theta_{2}}|\geq\min(r_{1},r_{2})\sin|\theta_{1}-\theta_{2}|,\quad\forall\ |\theta_{1}|,|\theta_{2}|<\frac{\pi}{4}.

By some elementary calculations and using the inequality (4.20) with (iii) of (4.18), we get a lower bound as follows

|w(l,l)w(m,m)|=|2cos(π(lm)δ12)eπi(l+m)δ122cos(π(lm)δ12)eπi(l+m)δ12|4πcos(π4)|π[((l+m)π(l+m))]|δ12\begin{split}|w(l,l^{\prime})-w(m,m^{\prime})|&=|2\cos(\pi(l-m^{\prime})\delta^{\frac{1}{2}})e^{\pi i(l+m^{\prime})\delta^{\frac{1}{2}}}-2\cos(\pi(l^{\prime}-m)\delta^{\frac{1}{2}})e^{\pi i(l^{\prime}+m)\delta^{\frac{1}{2}}}|\\ &\geq\frac{4}{\pi}\cos(\frac{\pi}{4})|\pi[((l+m^{\prime})-\pi(l^{\prime}+m))]|\delta^{\frac{1}{2}}\end{split}

where we use our hypothesis that all the supports of mk,lm_{k,l} are contained in QaQ_{a} (i.e. |2πlδ12|,|2πlδ12|,|2πmδ12|,|2πmδ12|<π4|2\pi l\delta^{\frac{1}{2}}|,|2\pi l^{\prime}\delta^{\frac{1}{2}}|,|2\pi m\delta^{\frac{1}{2}}|,|2\pi m^{\prime}\delta^{\frac{1}{2}}|<\frac{\pi}{4}). Thus combining the above estimate and (4.19), we obtain

(4.21) |(ll)(mm)|<200.|(l-l^{\prime})-(m-m^{\prime})|<200.

However (4.21) is not enough to show that the number of (m,m)(m,m^{\prime}) is finite. In the following, we give more exact information of lower and upper bound of |w(l,l)w(m,m)||w(l,l^{\prime})-w(m,m^{\prime})|.

According our condition, |ll|>103|l-l^{\prime}|>10^{3} and |mm|>103|m-m^{\prime}|>10^{3}. Without loss of generality, we suppose that l>l,m>ml>l^{\prime},m>m^{\prime}. Then by (4.21), we get

llmm.l-l^{\prime}\approx m-m^{\prime}.

This implies |w(l,l)||w(m,m)|(ll)δ12|w(l,l^{\prime})|\approx|w(m,m^{\prime})|\approx(l-l^{\prime})\delta^{\frac{1}{2}} in view of the identity (4.17) and (i), (iii) of (4.18).

Next we see that the rectangle R(l,l)R({l,l^{\prime}}) is centered at point w(l,l)w(l,l^{\prime}) with a distance to the origin |w(l,l)|=2sin(π(ll)δ12)|w(l,l^{\prime})|=2\sin(\pi(l-l^{\prime})\delta^{\frac{1}{2}}). Recall that R(l,l)R({l,l^{\prime}}) has half-length 100δ12100\delta^{\frac{1}{2}} in the direction w(l,l)w(l,l^{\prime}) and in the direction iw(l,l)iw(l,l^{\prime}) has half-width 90(ll)δ90(l-l^{\prime})\delta (see Figure 2 below). This implies that R(l,l)R({l,l^{\prime}}) is far from the origin (0,0)(0,0). In fact, by (iii) of (4.18),

dist(R(l,l),(0,0))2sin(π(ll)δ12)100δ123(ll)δ12.dist(R(l,l^{\prime}),(0,0))\geq 2\sin(\pi(l-l^{\prime})\delta^{\frac{1}{2}})-100\delta^{\frac{1}{2}}\geq 3{(l-l^{\prime})\delta^{\frac{1}{2}}}.

Similar properties also hold for R(m,m)R({m,m^{\prime}}). Since ΓlΓl\Gamma_{l}-\Gamma_{l^{\prime}} intersects ΓmΓm\Gamma_{m}-\Gamma_{m^{\prime}}, then R(l,l)R({l,l^{\prime}}) intersects R(m,m)R(m,m^{\prime}). Note that the angle θ\theta between w(l,l)w(l,l^{\prime}) and w(m,m)w(m,m^{\prime}) is π|l+lmm|δ12\pi|l+l^{\prime}-m-m^{\prime}|\delta^{\frac{1}{2}} (see Figure 2 below).

Refer to caption
Figure 2. R(l,l)R({l,l^{\prime}}) has half-length 100δ12100\delta^{\frac{1}{2}} in the direction along w(l,l)w(l,l^{\prime}) and in the direction along iw(l,l)iw(l,l^{\prime}) has half-width 90(ll)δ90(l-l^{\prime})\delta. R(l,l)R(l,l^{\prime}) intersects R(m,m)R(m,m^{\prime}) at the point AA.

Then the distance of w(l,l)w(l,l^{\prime}) and w(m,m)w(m,m^{\prime}) has an upper bound

|w(l,l)w(m,m)|90(ll)δ+90(mm)δ+100δ12θ.|w(l,l^{\prime})-w(m,m^{\prime})|\leq 90(l-l^{\prime})\delta+90(m-m^{\prime})\delta+100\delta^{\frac{1}{2}}\theta.

On the other hand, the distance of w(l,l)w(l,l^{\prime}) and w(m,m)w(m,m^{\prime}) has the following lower bound

|w(l,l)w(m,m)|2(sin(π(ll)δ12)+sin(π(mm)δ12))sinθ24((ll)+(mm))|l+lmm|δ,\begin{split}|w(l,l^{\prime})-w(m,m^{\prime})|&\geq 2\Big{(}\sin(\pi(l-l^{\prime})\delta^{\frac{1}{2}})+\sin(\pi(m-m^{\prime})\delta^{\frac{1}{2}})\Big{)}\sin\tfrac{\theta}{2}\\ &\geq 4((l-l^{\prime})+(m-m^{\prime}))|l+l^{\prime}-m-m^{\prime}|\delta,\end{split}

where we use (iii) of (4.18). The upper and lower estimates, together with our hypothesis ll>103l-l^{\prime}>10^{3}, mm>103m-m^{\prime}>10^{3}, finally yield that

(4.22) |l+lmm|200.|l+l^{\prime}-m-m^{\prime}|\leq 200.

Combining (4.22) and (4.21), we get that |lm|<200|l-m|<200, |lm|<200|l^{\prime}-m^{\prime}|<200. Hence we prove that for a fixed (l,l)(l,l^{\prime}), there exist only finite pairs (m,m)(m,m^{\prime}) such that (ΓlΓl)(ΓmΓm)(\Gamma_{l}-\Gamma_{l^{\prime}})\cap(\Gamma_{m}-\Gamma_{m^{\prime}})\neq\emptyset, which completes our proof. ∎

5. Bochner-Riesz means on quantum tori

In this section, our goal is to establish the full boundedness of Bochner-Riesz means on two-dimensional quantum tori, i.e. Theorem 1.1. We first establish the corresponding results on usual torus based on Theorem 4.1 in Section 4.

5.1. Bochner-Riesz means on usual tori

Recall that the dd-torus 𝕋d\mathbb{T}^{d} is defined as d/d{\mathbb{R}}^{d}/{\mathbb{Z}}^{d}. We often set 𝕋d\mathbb{T}^{d} as the cube [0,1]d[0,1]^{d} with opposite sides identified. The function on 𝕋d\mathbb{T}^{d} can be regarded as an 11-periodic function on d{\mathbb{R}}^{d} in every coordinate. Haar measure on 𝕋d\mathbb{T}^{d} is the restriction of Lebesgue measure to [0,1]d[0,1]^{d} which is still denoted by dxdx.

Our previous main theorem of Bochner-Riesz means in Section 4 is in the frame of the tensor von Neumann algebra L(d)¯L_{\infty}({\mathbb{R}}^{d})\overline{\otimes}{\mathcal{M}}. To extend our main result to the fully noncommutative quantum torus, we should first transfer them to the setting of the tensor von Neumann algebra L(𝕋d)¯L_{\infty}(\mathbb{T}^{d})\overline{\otimes}{\mathcal{M}}.

Given a function f:𝕋df:\mathbb{T}^{d}\rightarrow{\mathcal{M}}, define the Bochner-Riesz means BRλ(f)(x){\text{\rm B}}^{\lambda}_{R}(f)(x) by (1.1). By the transference method (see Theorem 5.4 later), we can formulate the noncommutative Bochner-Riesz conjecture on tori with ranges of indexes rr and pp the same as that in the commutative case in the introduction. Our main result in this subsection is stated as follows.

Theorem 5.1.

Suppose 0<λ120<\lambda\leq\frac{1}{2} and 43+2λ<p<412λ\frac{4}{3+2\lambda}<p<\frac{4}{1-2\lambda}. Let BRλ{\text{\rm B}}_{R}^{\lambda} be defined in (1.1) for d=2d=2. Then we have

supR>0BRλ(f)Lp(L(𝕋2)¯)fLp(L(𝕋2)¯).\sup_{R>0}\|{\text{\rm B}}_{R}^{\lambda}(f)\|_{L_{p}(L_{\infty}(\mathbb{T}^{2})\overline{\otimes}{\mathcal{M}})}\lesssim\|f\|_{L_{p}(L_{\infty}(\mathbb{T}^{2})\overline{\otimes}{\mathcal{M}})}.

Consequently for fLp(L(𝕋2)¯)f\in L_{p}(L_{\infty}(\mathbb{T}^{2})\overline{\otimes}{\mathcal{M}}), BRλ(f){\text{\rm B}}_{R}^{\lambda}(f) converges to ff in Lp(L(𝕋2)¯)L_{p}(L_{\infty}(\mathbb{T}^{2})\overline{\otimes}{\mathcal{M}}) as RR\rightarrow\infty.

To prove Theorem 5.1, we should introduce some definitions and lemmas.

Definition 5.2.

For xdx\in{\mathbb{R}}^{d}, a bounded function mm on d{\mathbb{R}}^{d} taking values in {\mathbb{C}} is called regulated at the point xx if

limε01εd|t|ε(m(xt)m(x))𝑑t=0.\lim_{\varepsilon\rightarrow 0}\frac{1}{\varepsilon^{d}}\int_{|t|\leq\varepsilon}(m(x-t)-m(x))dt=0.

The function mm is called regulated if it is regulated at every xdx\in{\mathbb{R}}^{d}.

The above definition of regulated function was first appeared in [12]. Next we introduce the Fourier multiplier on L(𝕋d)¯L_{\infty}(\mathbb{T}^{d})\overline{\otimes}{\mathcal{M}}.

Definition 5.3.

Given a bounded function m:dm:{\mathbb{Z}}^{d}\rightarrow{\mathbb{C}}, we say {m(z)}zd\{m(z)\}_{z\in{\mathbb{Z}}^{d}} is a Fourier multiplier on Lp(L(𝕋d)¯)L_{p}(L_{\infty}(\mathbb{T}^{d})\overline{\otimes}{\mathcal{M}}) if the operator SS defined by

S(f)(x)=zdm(z)f^(z)e2πixz{S(f)}(x)=\sum_{z\in{\mathbb{Z}}^{d}}m(z)\widehat{f}(z)e^{2\pi ix\cdot z}

where f:𝕋df:\mathbb{T}^{d}\rightarrow{\mathcal{M}} and f^(z)=𝕋de2πizxf(x)𝑑x\widehat{f}(z)=\int_{\mathbb{T}^{d}}e^{-2\pi izx}f(x)dx, extends to a bounded operator on Lp(L(𝕋d)¯)L_{p}(L_{\infty}(\mathbb{T}^{d})\overline{\otimes}{\mathcal{M}}). The space of all these multipliers is denoted by Mp(L(𝕋d)¯)M_{p}(L_{\infty}(\mathbb{T}^{d})\overline{\otimes}{\mathcal{M}}). For simplicity, we denote the norm of Lp(L(𝕋d)¯)L_{p}(L_{\infty}(\mathbb{T}^{d})\overline{\otimes}{\mathcal{M}}) multiplier by Mp(L(𝕋d)¯)\|\cdot\|_{M_{p}(L_{\infty}(\mathbb{T}^{d})\overline{\otimes}{\mathcal{M}})}.

Theorem 5.4.

Let mm be a regulated function which lies in Mp(L(d)¯)M_{p}(L_{\infty}({\mathbb{R}}^{d})\overline{\otimes}{\mathcal{M}}) for some 1p<1\leq p<\infty. Then the sequence {m(z)}zd\{m(z)\}_{z\in{\mathbb{Z}}^{d}} belongs to Mp(L(𝕋d)¯)M_{p}(L_{\infty}(\mathbb{T}^{d})\overline{\otimes}{\mathcal{M}}) and more precisely

{m(z)}zdMp(L(𝕋d)¯)mMp(L(d)¯).\|\{m(z)\}_{z\in{\mathbb{Z}}^{d}}\|_{M_{p}(L_{\infty}(\mathbb{T}^{d})\overline{\otimes}{\mathcal{M}})}\leq\|m\|_{M_{p}(L_{\infty}({\mathbb{R}}^{d})\overline{\otimes}{\mathcal{M}})}.

Moreover for all R>0R>0, the sequence {m(z/R)}zd\{m(z/R)\}_{z\in{\mathbb{Z}}^{d}} lies in Mp(L(𝕋d)¯)M_{p}(L_{\infty}(\mathbb{T}^{d})\overline{\otimes}{\mathcal{M}}) and

supR>0{m(z/R)}zdMp(L(𝕋d)¯)mMp(L(d)¯).\sup_{R>0}\|\{m(z/R)\}_{z\in{\mathbb{Z}}^{d}}\|_{M_{p}(L_{\infty}(\mathbb{T}^{d})\overline{\otimes}{\mathcal{M}})}\leq\|m\|_{M_{p}(L_{\infty}({\mathbb{R}}^{d})\overline{\otimes}{\mathcal{M}})}.
Proof.

The proof of this theorem is quite similar to that in the commutative case (see e.g. the proof of Theorem 4.3.7 in [17]), so we omit the details of the proof here. ∎

Now we can apply Theorem 4.1 and Theorem 5.4 to show Theorem 5.1. Indeed notice that the function m(z)=(1|z|2)+λm(z)=(1-|z|^{2})_{+}^{\lambda} is continuous on d{\mathbb{R}}^{d}. Hence it is regulated. Now utilizing Theorem 4.1 and Theorem 5.4, it is easy to obtain Theorem 5.1.

5.2. Bochner-Riesz means on quantum tori

In this subsection, we finally consider the Bochner-Riesz means on two-dimensional quantum tori and prove our main results Theorem 1.1. We begin by introducing some notation.

Suppose that d2d\geq 2, θ=(θk,j)1k,jd\theta=(\theta_{k,j})_{1\leq k,j\leq d} is a real skew symmetric d×dd\times d matrix. The dd-dimensional noncommutative torus 𝒜θ{\mathcal{A}}_{\theta} is a universal CC^{*}-algebra generated by dd unitary operators U1,,UdU_{1},\cdots,U_{d} satisfying the following commutation relation:

UkUj=e2πiθk,jUjUk,1k,jd.U_{k}U_{j}=e^{2\pi i\theta_{k,j}}U_{j}U_{k},\quad 1\leq k,j\leq d.

By the unitary property of UkU_{k}s, if we multiply UkU^{*}_{k} in both left and right sides of the above equality, we get

UjUk=e2πiθk,jUkUj,1k,jd.U_{j}U^{*}_{k}=e^{2\pi i\theta_{k,j}}U_{k}^{*}U_{j},\quad 1\leq k,j\leq d.

For a multi-index k=(k1,,kd)dk=(k_{1},\cdots,k_{d})\in{\mathbb{Z}}^{d} and U=(U1,,Ud)U=(U_{1},\cdots,U_{d}), we define Uk=U1k1UdkdU^{k}=U_{1}^{k_{1}}\cdots U_{d}^{k_{d}}. We call

(5.1) f=kdαkUk,withαk,f=\sum_{k\in{\mathbb{Z}}^{d}}\alpha_{k}U^{k},\quad\text{with}\ \ \alpha_{k}\in{\mathbb{C}},

is a polynomial in UU if the sum in (5.1) is finite, i.e. αk0\alpha_{k}\neq 0 holds only for finite multi-indexes. Let 𝒫θ\mathcal{P}_{\theta} be the involution algebra of all such polynomials. Then 𝒫θ\mathcal{P}_{\theta} is dense in 𝒜θ{\mathcal{A}}_{\theta}. A well-known fact is that 𝒜θ{\mathcal{A}}_{\theta} admits a faithful tracial state τ\tau such that τ(U1k1Udkd)=1\tau(U_{1}^{k_{1}}\cdots U_{d}^{k_{d}})=1 if and only if k=(k1,,kd)=0k=(k_{1},\cdots,k_{d})=\textbf{0} where 0=(0,,0)\textbf{0}=(0,\cdots,0) (see e.g. [44]). Hence for any polynomial ff with form (5.1), we can define

τ(f)=α0.\tau(f)=\alpha_{\textbf{0}}.

Define 𝕋θd\mathbb{T}_{\theta}^{d} the weak *-closure of 𝒜θ{\mathcal{A}}_{\theta} in the GNS representation of τ\tau. We call 𝕋θd\mathbb{T}_{\theta}^{d} the dd-dimensional quantum torus associated to θ\theta. The state τ\tau also extends to a normal faithful state on 𝕋θd\mathbb{T}_{\theta}^{d}, which will be denoted again by τ\tau. Notice that when θ=0\theta=0, 𝒜θ=C(𝕋d){\mathcal{A}}_{\theta}=C(\mathbb{T}^{d}) and 𝕋θd=L(𝕋d)\mathbb{T}_{\theta}^{d}=L_{\infty}(\mathbb{T}^{d}). Thus quantum torus 𝕋θd\mathbb{T}_{\theta}^{d} is a deformation of classical torus 𝕋d\mathbb{T}^{d}.

Let Lp(𝕋θd)L_{p}(\mathbb{T}_{\theta}^{d}) be the noncommutative space associated with (𝕋θd,τ)(\mathbb{T}_{\theta}^{d},\tau). Using τ\tau is a state and Hölder’s inequality, we see that Lq(𝕋θd)Lp(𝕋θd)L_{q}(\mathbb{T}_{\theta}^{d})\subset L_{p}(\mathbb{T}^{d}_{\theta}) for 0<p<q<0<p<q<\infty. For any fL1(𝕋θd)f\in L_{1}(\mathbb{T}_{\theta}^{d}), there exists a formal Fourier series

fmdf^(m)Umf\sim\sum_{m\in{\mathbb{Z}}^{d}}\widehat{f}(m)U^{m}

where f^(m)=τ((Um)f)\widehat{f}(m)=\tau((U^{m})^{*}f) is called the Fourier coefficient of ff. Analogous to the classical analysis, a fundamental problem here is that when the Fourier series mdf^(m)Um\sum_{m\in{\mathbb{Z}}^{d}}\widehat{f}(m)U^{m} converges to ff. In the following, we consider the most important Bochner-Riesz means on quantum tori which is defined by

(5.2) BRλ(f)=md(1|mR|2)+λf^(m)Um.{\text{\rm B}}^{\lambda}_{R}(f)=\sum_{m\in{\mathbb{Z}}^{d}}\big{(}1-|\tfrac{m}{R}|^{2}\big{)}^{\lambda}_{+}\hat{f}(m)U^{m}.

The Bochner-Riesz means on quantum tori was firstly studied by Z. Chen, Q. Xu and Z. Yin [7]. A fundamental problem raised in [7, Page 762] is that in which senses BRλ(f){\text{\rm B}}^{\lambda}_{R}(f) converge back to ff. We consider the LpL_{p} convergence here. In fact similar to that of the commutative case, one can formulate the following problem of quantum Bochner-Riesz means.

Conjecture.

Suppose λ>0\lambda>0 and 2dd+1+2λ<p<2dd12λ\frac{2d}{d+1+2\lambda}<p<\frac{2d}{d-1-2\lambda}. Consider the Bochner-Riesz means defined in (5.2), then we have

supR>0BRλ(f)Lp(𝕋θd)fLp(𝕋θd).\sup_{R>0}\|{\text{\rm B}}_{R}^{\lambda}(f)\|_{L_{p}(\mathbb{T}_{\theta}^{d})}\lesssim\|f\|_{L_{p}(\mathbb{T}_{\theta}^{d})}.

The main result here is to show that the preceding conjecture holds for two dimensions, i.e. Theorem 1.1. For the reader’s convenience, we restate this theorem as follows.

Theorem 5.5.

Suppose 0<λ120<\lambda\leq\frac{1}{2} and 43+2λ<p<412λ\frac{4}{3+2\lambda}<p<\frac{4}{1-2\lambda}. Let the Bochner-Riesz means BRλ{\text{\rm B}}_{R}^{\lambda} be defined in (5.2) for d=2d=2. Then we have

supR>0BRλ(f)Lp(𝕋θ2)fLp(𝕋θ2).\sup_{R>0}\|{\text{\rm B}}_{R}^{\lambda}(f)\|_{L_{p}(\mathbb{T}_{\theta}^{2})}\lesssim\|f\|_{L_{p}(\mathbb{T}^{2}_{\theta})}.

Consequently for fLp(𝕋θ2)f\in L_{p}(\mathbb{T}_{\theta}^{2}), BRλ(f){\text{\rm B}}_{R}^{\lambda}(f) converges to ff in Lp(𝕋θ2)L_{p}(\mathbb{T}_{\theta}^{2}) as RR\rightarrow\infty.

The proof is based on a transference technique which is a standard method. Consider the tensor von Neumann algebra 𝒩θ=L(𝕋d)¯𝕋θd{\mathcal{N}}_{\theta}=L_{\infty}(\mathbb{T}^{d})\overline{\otimes}\mathbb{T}_{\theta}^{d} equipped with the tensor trace v=𝕋d𝑑xτv=\int_{\mathbb{T}^{d}}dx\otimes\tau. Let Lp(𝒩θ)L_{p}({\mathcal{N}}_{\theta}) be the noncommutative LpL_{p} space associated with (𝒩θ,v)({\mathcal{N}}_{\theta},v). Observe that

Lp(𝒩θ)Lp(𝕋d;Lp(𝕋θd))L_{p}({\mathcal{N}}_{\theta})\cong L_{p}(\mathbb{T}^{d};L_{p}(\mathbb{T}^{d}_{\theta}))

where the space on the right hand side is the Bochner LpL_{p} space on 𝕋d\mathbb{T}^{d} with values in Lp(𝕋θd)L_{p}(\mathbb{T}_{\theta}^{d}).

For every xdx\in{\mathbb{Z}}^{d}, we define πx\pi_{x} as

πx(Uk)=e2πixkUk=e2πix1k1e2πixdkdUk1Ukd\pi_{x}(U^{k})=e^{2\pi ix\cdot k}U^{k}=e^{2\pi ix_{1}k_{1}}\cdots e^{2\pi ix_{d}k_{d}}U^{k_{1}}\cdots U^{k_{d}}

which is an isomorphism of 𝕋θd\mathbb{T}_{\theta}^{d}. It is easy to see that τ(πx(f))=τ(f)\tau(\pi_{x}(f))=\tau(f) for any x𝕋dx\in\mathbb{T}^{d}. Hence πx\pi_{x} is trace preserving. Therefore it extends to an isometry on Lp(𝕋θd)L_{p}(\mathbb{T}^{d}_{\theta}) for 1p<1\leq p<\infty, i.e.

(5.3) πx(f)Lp(𝕋θd)=fLp(𝕋θd).\|\pi_{x}(f)\|_{L_{p}(\mathbb{T}_{\theta}^{d})}=\|f\|_{L_{p}(\mathbb{T}^{d}_{\theta})}.

The following transference method has been showed by Z. Chen, Q. Xu and Z. Yin in [7, Proposition 2.1 & Corollary 2.2].

Lemma 5.6.

For any fLp(𝕋θd)f\in L_{p}(\mathbb{T}^{d}_{\theta}), the function f~:xπx(f)\tilde{f}:x\rightarrow\pi_{x}(f) is continuous from 𝕋d\mathbb{T}^{d} to Lp(𝕋θd)L_{p}(\mathbb{T}^{d}_{\theta}) (with respect to the weak *-topology for p=p=\infty). Moreover f~Lp(𝒩θ)\tilde{f}\in L_{p}({\mathcal{N}}_{\theta}) and f~Lp(𝒩θ)=fLp(𝕋θd)\|\tilde{f}\|_{L_{p}({\mathcal{N}}_{\theta})}=\|f\|_{L_{p}(\mathbb{T}^{d}_{\theta})} for 1p1\leq p\leq\infty. Thus, ff~f\rightarrow\tilde{f} is an isometric embedding from Lp(𝕋θd)L_{p}(\mathbb{T}_{\theta}^{d}) to Lp(𝒩θ)L_{p}({\mathcal{N}}_{\theta}).

Proof of Theorem 5.5.

The proof just follows from Theorem 5.1 and Lemma 5.6. In fact, by the density argument, it suffices to consider ff as a polynomial k2f^(k)Uk\sum_{k\in{\mathbb{Z}}^{2}}\hat{f}(k)U^{k}. Define f~\tilde{f} in Lemma 5.6. Then f~Lp(𝒩θ)\tilde{f}\in L_{p}({\mathcal{N}}_{\theta}) and f~Lp(𝒩θ)=fLp(𝕋θ2)\|\tilde{f}\|_{L_{p}({\mathcal{N}}_{\theta})}=\|f\|_{L_{p}(\mathbb{T}^{2}_{\theta})} by Lemma 5.6. Using Theorem 5.1, we get

(5.4) supR>0BRλ(f~)Lp(𝕋2;Lp(𝕋θ2))f~Lp(𝕋2;Lp(𝕋θ2)).\sup_{R>0}\|{\text{\rm B}}_{R}^{\lambda}(\tilde{f})\|_{L_{p}(\mathbb{T}^{2};L_{p}(\mathbb{T}_{\theta}^{2}))}\lesssim\|\tilde{f}\|_{L_{p}(\mathbb{T}^{2};L_{p}(\mathbb{T}_{\theta}^{2}))}.

On the other hand, it is easy to see f~^(k)=f^(k)Uk\widehat{\tilde{f}}(k)=\widehat{f}(k)U^{k}. According the definition of Bochner-Riesz means,

BRλ(f~)(x)=k2(1|kR|2)+λf~^(k)e2πikx=k2(1|kR|2)+λf^(k)e2πikxUk=πx[BRλ(f)].\begin{split}{\text{\rm B}}^{\lambda}_{R}(\tilde{f})(x)=\sum_{k\in{\mathbb{Z}}^{2}}\big{(}1-\big{|}\tfrac{k}{R}\big{|}^{2}\big{)}_{+}^{\lambda}\widehat{\tilde{f}}(k)e^{2\pi ik\cdot x}=\sum_{k\in{\mathbb{Z}}^{2}}\big{(}1-\big{|}\tfrac{k}{R}\big{|}^{2}\big{)}_{+}^{\lambda}\widehat{f}(k)e^{2\pi ik\cdot x}U^{k}=\pi_{x}[{\text{\rm B}}^{\lambda}_{R}(f)].\end{split}

This, together with (5.3) and (5.4), implies the desired estimate in Theorem 5.5. The convergence follows from the standard limiting argument. ∎

Appendix A Interpolation of analytic families of operators on noncommutative LpL_{p} spaces

In this appendix, we state precisely an analytic interpolation theorem which may be known to experts. Let 𝒮\mathcal{S} be the linear span of all x+x\in{\mathcal{M}}_{+} whose support projections have finite trace. Suppose that TzT_{z} is a linear operator mapping 𝒮\mathcal{S} to itself for every zz in the closed strip S¯={z:0Rez1}\bar{S}=\{z\in{\mathbb{C}}:0\leq\text{Re}z\leq 1\}. We say the family {Tz}z\{T_{z}\}_{z} is analytic if the function

zτ(gTz(f))z\rightarrow\tau(gT_{z}(f))

is analytic in the open strip S={z:0<Rez<1}S=\{z\in{\mathbb{C}}:0<\text{Re}z<1\} and continuous on S¯\bar{S} for any functions ff and gg in 𝒮\mathcal{S}. Moreover we say the analytic family {Tz}z\{T_{z}\}_{z} is of admissible growth if there exists a constant 0<a<π0<a<\pi such that

ea|Imz|log|τ(gTz(f))|<e^{-a|\text{Im}z|}\log|\tau(gT_{z}(f))|<\infty

for all zS¯z\in\bar{S}. Now we can state the following analytic interpolation theorem.

Theorem A.1.

Suppose that TzT_{z} is an analytic family of linear operators of admissible growth. Let p0,p1,q0,q1(0,)p_{0},p_{1},q_{0},q_{1}\in(0,\infty) and assume that M0,M1M_{0},M_{1} are positive functions on {\mathbb{R}} such that

(A.1) supyeb|y|logMj(y)<\sup_{y\in{\mathbb{R}}}e^{-b|y|}\log M_{j}(y)<\infty

for j=0,1j=0,1 and some b(0,π)b\in(0,\pi). Let p,θp,\theta satisfy 1p=1θp0+θp1\frac{1}{p}=\frac{1-\theta}{p_{0}}+\frac{\theta}{p_{1}} and 1q=1θq0+θq1\frac{1}{q}=\frac{1-\theta}{q_{0}}+\frac{\theta}{q_{1}}. Suppose that

(A.2) Tiy(f)Lq0()M0(y)fLp0(),T1+iy(f)Lq1()M1(y)fLp1()\begin{split}\|T_{iy}(f)\|_{L_{q_{0}}({\mathcal{M}})}&\leq M_{0}(y)\|f\|_{L_{p_{0}}({\mathcal{M}})},\quad\|T_{1+iy}(f)\|_{L_{q_{1}}({\mathcal{M}})}\leq M_{1}(y)\|f\|_{L_{p_{1}}({\mathcal{M}})}\end{split}

hold for all f𝒮f\in\mathcal{S}. Then for any θ(0,1)\theta\in(0,1), we have

Tθ(f)Lq()M(θ)fLp(),\|T_{\theta}(f)\|_{L_{q}({\mathcal{M}})}\leq M(\theta)\|f\|_{L_{p}({\mathcal{M}})},

where for 0<t<10<t<1,

M(t)=exp{sin(πt)2[logM0(y)cosh(πy)cos(πt)+logM1(y)cosh(πy)+cos(πt)]𝑑y}.M(t)=\exp\Big{\{}\frac{\sin(\pi t)}{2}\int_{-\infty}^{\infty}\Big{[}\frac{\log M_{0}(y)}{\cosh(\pi y)-\cos(\pi t)}+\frac{\log M_{1}(y)}{\cosh(\pi y)+\cos(\pi t)}\Big{]}dy\Big{\}}.

To prove this theorem, we need an extension of the three lines theorem which could be found in [48, Page 206, Lemma 4.2].

Lemma A.2.

Suppose that FF is analytic on the open strip SS and continuous on its closure such that

supzS¯ea|Imz|log|F(z)|<\sup_{z\in\bar{S}}e^{-a|\text{Im}z|}\log|F(z)|<\infty

for some a(0,π)a\in(0,\pi). Then for any 0<x<10<x<1, we have

|F(x)|exp{sin(πx)2[log|F(iy)|cosh(πy)cos(πx)+log|F(1+iy)|cosh(πy)+cos(πx)]𝑑y}.|F(x)|\leq\exp\Big{\{}\frac{\sin(\pi x)}{2}\int_{-\infty}^{\infty}\Big{[}\frac{\log|F(iy)|}{\cosh(\pi y)-\cos(\pi x)}+\frac{\log|F(1+iy)|}{\cosh(\pi y)+\cos(\pi x)}\Big{]}dy\Big{\}}.
Proof of Theorem A.1.

The proof is quite similar to that in the commutative case. Let f,g𝒮f,g\in\mathcal{S} with polar decompositions f=u|f|f=u|f| and g=v|g|g=v|g|. Without loss of generality, we may suppose that fLp()=1=gLq()\|f\|_{L_{p}({\mathcal{M}})}=1=\|g\|_{L_{q^{\prime}}({\mathcal{M}})}. By the duality, to prove our theorem, it suffices to show

|τ(gTθ(f))|M(θ).|\tau(gT_{\theta}(f))|\leq M({\theta}).

For zS¯z\in\bar{S}, define f(z)=u|f|p(1z)p0+pzp1f(z)=u|f|^{\frac{p(1-z)}{p_{0}}+\frac{pz}{p_{1}}} and g(z)=v|g|q(1z)q0+qzq1g(z)=v|g|^{\frac{q^{\prime}(1-z)}{q^{\prime}_{0}}+\frac{q^{\prime}z}{q^{\prime}_{1}}} where the continuous functional calculus is defined by complex powers of positive operators. By the density argument, we could suppose that |f||f| and |g||g| are linear combinations of mutually orthogonal projections of finite trace, i.e.

|f|=j=1nαjej,|g|=k=1mβke~k|f|=\sum_{j=1}^{n}\alpha_{j}e_{j},\quad|g|=\sum_{k=1}^{m}\beta_{k}\tilde{e}_{k}

where αj\alpha_{j}s, βk\beta_{k}s are real and eje_{j}s, e~k\tilde{e}_{k}s are mutually orthogonal basis. Then

f(z)=j=1nαjp(1z)p0+pzp1uej.f(z)=\sum_{j=1}^{n}\alpha_{j}^{\frac{p(1-z)}{p_{0}}+\frac{pz}{p_{1}}}ue_{j}.

Therefore the function zf(z)z\rightarrow f(z) is an analytic function on {\mathbb{C}} taking values in {\mathcal{M}}. Similar properties hold for the function zg(z)z\rightarrow g(z). Define

F(z)=τ(g(z)Tz(f(z))).F(z)=\tau(g(z)T_{z}(f(z))).

Then we have

F(z)=j=1nk=1mαjp(1z)p0+pzp1βkq(1z)q0+qzq1τ(ve~kTz(uej)).F(z)=\sum_{j=1}^{n}\sum_{k=1}^{m}\alpha_{j}^{\frac{p(1-z)}{p_{0}}+\frac{pz}{p_{1}}}\beta_{k}^{\frac{q^{\prime}(1-z)}{q^{\prime}_{0}}+\frac{q^{\prime}z}{q^{\prime}_{1}}}\tau(v\tilde{e}_{k}T_{z}(ue_{j})).

By our assumption, τ(ve~kTz(uej))\tau(v\tilde{e}_{k}T_{z}(ue_{j})) is analytic. Hence F(z)F(z) is an analytic function satisfying the hypothesis of Lemma A.2. Recall a property of polar decomposition: |f|=|f|uu=uu|f||f|=|f|u^{*}u=u^{*}u|f|, then by the continuous functional calculus of |f||f|, we obtain

(A.3) ω(|f|)=ω(|f|)uu=uuw(|f|),\omega(|f|)=\omega(|f|)u^{*}u=u^{*}uw(|f|),

where ω\omega is a continuous function on +{\mathbb{R}}_{+}. Since

f(iy)=u|f|iyp(1p11p0)+pp0,f(iy)=u|f|^{iyp(\frac{1}{p_{1}}-\frac{1}{p_{0}})+\frac{p}{p_{0}}},

then by (A.3), we get

|f(iy)|2=f(iy)f(iy)=|f|iyp(1p11p0)+pp0uu|f|iyp(1p11p0)+pp0=|f|2pp0.\begin{split}|f(iy)|^{2}=f^{*}(iy)f(iy)=|f|^{-iyp(\frac{1}{p_{1}}-\frac{1}{p_{0}})+\frac{p}{p_{0}}}u^{*}u|f|^{iyp(\frac{1}{p_{1}}-\frac{1}{p_{0}})+\frac{p}{p_{0}}}=|f|^{\frac{2p}{p_{0}}}.\end{split}

Therefore we get f(iy)Lp0()=1\|f(iy)\|_{L_{p_{0}}({\mathcal{M}})}=1. Similarly f(1+iy)Lp1()=1=g(iy)Lq0()=g(1+iy)Lq1()\|f(1+iy)\|_{L_{p_{1}}({\mathcal{M}})}=1=\|g(iy)\|_{L_{q^{\prime}_{0}}({\mathcal{M}})}=\|g(1+iy)\|_{L_{q^{\prime}_{1}}({\mathcal{M}})}. Hölder’s inequality and our assumption show that for all yy\in{\mathbb{R}},

|F(iy)|Tiy(f(iy))Lq0()g(iy)Lq0()M0(y)f(iy)Lp0()g(iy)Lq0()=M0(y).\begin{split}|F(iy)|&\leq\|T_{iy}(f(iy))\|_{L_{q_{0}}({\mathcal{M}})}\|g(iy)\|_{L_{q^{\prime}_{0}}({\mathcal{M}})}\\ &\leq M_{0}(y)\|f(iy)\|_{L_{p_{0}}({\mathcal{M}})}\|g(iy)\|_{L_{q^{\prime}_{0}}({\mathcal{M}})}=M_{0}(y).\end{split}

Similarly for all yy\in{\mathbb{R}}, |F(1+iy)|M1(y)|F(1+iy)|\leq M_{1}(y). Now applying Lemma A.2 with the preceding two estimates and notice that cosh(πy)=12(eπy+eπy)1cos(πx)\cosh(\pi y)=\frac{1}{2}(e^{\pi y}+e^{-\pi y})\geq 1\geq\cos(\pi x), we get

|τ(gTθ(f))|=|F(θ)|M(θ),|\tau(gT_{\theta}(f))|=|F(\theta)|\leq M(\theta),

which implies the required estimate. ∎

Acknowledgement

The author would like to thank Guixiang Hong for some helpful suggestions when preparing this paper and the referees for their very careful reading and valuable suggestions.


References

  • [1] S. Bochner, Summation of multiple Fourier series by spherical means, Trans. Amer. Math. Soc. 40 (1936), no. 2, 175–207.
  • [2] J. Bourgain, On the dimension of Kakeya sets and related maximal inequalities, Geom. Funct. Anal. 9 (1999), no. 2, 256–282.
  • [3] J. Bourgain and L. Guth, Bounds on oscillatory integral operators based on multilinear estimates, Geom. Funct. Anal. 21 (2011), no. 6, 1239–1295.
  • [4] L. Cadilhac, Weak boundedness of Calderón-Zygmund operators on noncommutative L1L_{1}-spaces, J. Funct. Anal. 274 (2018), no. 3, 769–796.
  • [5] A. Carbery, The boundedness of the maximal Bochner-Riesz operator on L4(𝐑2)L^{4}({\bf R}^{2}), Duke Math. J. 50 (1983), no. 2, 409–416.
  • [6] L. Carleson and P. Sjölin, Oscillatory integrals and a multiplier problem for the disc, Studia Math. 44 (1972), 287–299.
  • [7] Z. Chen, Q. Xu, and Z. Yin, Harmonic analysis on quantum tori, Comm. Math. Phys. 322 (2013), no. 3, 755–805.
  • [8] A. Connes, Noncommutative geometry, Academic Press, Inc., San Diego, CA, 1994.
  • [9] A. Córdoba, The Kakeya maximal function and the spherical summation multipliers, Amer. J. Math. 99 (1977), no. 1, 1–22.
  • [10] A. Córdoba, A note on Bochner-Riesz operators, Duke Math. J. 46 (1979), no. 3, 505–511.
  • [11] K. R. Davidson, CC^{*}-algebras by example, Fields Institute Monographs, vol. 6, American Mathematical Society, Providence, RI, 1996.
  • [12] K. de Leeuw, On LpL_{p} multipliers, Ann. of Math. (2) 81 (1965), 364–379.
  • [13] C. Demeter, Fourier restriction, decoupling, and applications, Cambridge Studies in Advanced Mathematics, vol. 184, Cambridge University Press, Cambridge, 2020.
  • [14] C. Fefferman, A note on spherical summation multipliers, Israel J. Math. 15 (1973), 44–52.
  • [15] A. González-Pérez, M. Junge, and J. Parcet, Singular integrals in quantum euclidean spaces, Mem. Amer. Math. Soc. To appear. arXiv: 1705,01081.
  • [16] by same author, Smooth Fourier multipliers in group algebras via Sobolev dimension, Ann. Sci. Éc. Norm. Supér. (4) 50 (2017), no. 4, 879–925.
  • [17] L. Grafakos, Classical Fourier analysis, Third ed., Graduate Texts in Mathematics, vol. 249, Springer, New York, 2014.
  • [18] by same author, Modern Fourier analysis, Third ed., Graduate Texts in Mathematics, vol. 250, Springer, New York, 2014.
  • [19] L. Guth, J. Hickman, and M. Iliopoulou, Sharp estimates for oscillatory integral operators via polynomial partitioning, Acta Math. 223 (2019), no. 2, 251–376.
  • [20] C. S. Herz, On the mean inversion of Fourier and Hankel transforms, Proc. Nat. Acad. Sci. U.S.A. 40 (1954), 996–999.
  • [21] G. Hong, X. Lai, and B. Xu, Maximal singular integral operators acting on noncommutative LpL_{p}-spaces, arXiv:2009.03827.
  • [22] G. Hong, B. Liao, and S. Wang, Noncommutative maximal ergodic inequalities associated with doubling conditions, Duke Math. J. 170 (2021), no. 2, 205–246.
  • [23] G. Hong, S. Wang, and X. Wang, Pointwise convergence of noncommutative fourier series, arXiv:1908.00240.
  • [24] L. Hörmander, Oscillatory integrals and multipliers on FLpFL^{p}, Ark. Mat. 11 (1973), 1–11.
  • [25] T. P. Hytönen, J. L. Torrea, and D. V. Yakubovich, The Littlewood-Paley-Rubio de Francia property of a Banach space for the case of equal intervals, Proc. Roy. Soc. Edinburgh Sect. A 139 (2009), no. 4, 819–832.
  • [26] M. Junge, Doob’s inequality for non-commutative martingales, J. Reine Angew. Math. 549 (2002), 149–190.
  • [27] M. Junge, T. Mei, and J. Parcet, Smooth Fourier multipliers on group von Neumann algebras, Geom. Funct. Anal. 24 (2014), no. 6, 1913–1980.
  • [28] by same author, Noncommutative Riesz transforms—dimension free bounds and Fourier multipliers, J. Eur. Math. Soc. (JEMS) 20 (2018), no. 3, 529–595.
  • [29] M. Junge, S. Rezvani, and Q. Zeng, Harmonic analysis approach to Gromov-Hausdorff convergence for noncommutative tori, Comm. Math. Phys. 358 (2018), no. 3, 919–994.
  • [30] M. Junge and Q. Xu, Noncommutative maximal ergodic theorems, J. Amer. Math. Soc. 20 (2007), no. 2, 385–439.
  • [31] S. Lee, Improved bounds for Bochner-Riesz and maximal Bochner-Riesz operators, Duke Math. J. 122 (2004), no. 1, 205–232.
  • [32] by same author, Square function estimates for the Bochner-Riesz means, Anal. PDE 11 (2018), no. 6, 1535–1586.
  • [33] X. Li and S. Wu, New estimates of the maximal Bochner-Riesz operator in the plane, Math. Ann. 378 (2020), no. 3-4, 873–890.
  • [34] E. McDonald, F. Sukochev, and X. Xiong, Quantum differentiability on quantum tori, Comm. Math. Phys. 371 (2019), no. 3, 1231–1260.
  • [35] by same author, Quantum differentiability on noncommutative Euclidean spaces, Comm. Math. Phys. 379 (2020), no. 2, 491–542.
  • [36] T. Mei, Operator valued Hardy spaces, Mem. Amer. Math. Soc. 188 (2007), no. 881, vi+64.
  • [37] T. Mei and M. de la Salle, Complete boundedness of heat semigroups on the von Neumann algebra of hyperbolic groups, Trans. Amer. Math. Soc. 369 (2017), no. 8, 5601–5622.
  • [38] T. Mei and J. Parcet, Pseudo-localization of singular integrals and noncommutative Littlewood-Paley inequalities, Int. Math. Res. Not. IMRN (2009), no. 8, 1433–1487.
  • [39] A. Nagel, E. M. Stein, and S. Wainger, Differentiation in lacunary directions, Proc. Nat. Acad. Sci. U.S.A. 75 (1978), no. 3, 1060–1062.
  • [40] J. Parcet, Pseudo-localization of singular integrals and noncommutative Calderón-Zygmund theory, J. Funct. Anal. 256 (2009), no. 2, 509–593.
  • [41] G. Pisier, Non-commutative vector valued LpL_{p}-spaces and completely pp-summing maps, Astérisque (1998), no. 247, vi+131.
  • [42] G. Pisier and Q. Xu, Non-commutative LpL^{p}-spaces, Handbook of the geometry of Banach spaces, Vol. 2, North-Holland, Amsterdam, 2003, pp. 1459–1517.
  • [43] É. Ricard, LpL_{p}-multipliers on quantum tori, J. Funct. Anal. 270 (2016), no. 12, 4604–4613.
  • [44] M. A. Rieffel, Noncommutative tori—a case study of noncommutative differentiable manifolds, Contemp. Math., vol. 105, Amer. Math. Soc., Providence, RI, 1990, pp. 191–211.
  • [45] M. Spera, Sobolev theory for noncommutative tori, Rend. Sem. Mat. Univ. Padova 86 (1991), 143–156 (1992).
  • [46] E. M. Stein, Localization and summability of multiple Fourier series, Acta Math. 100 (1958), 93–147.
  • [47] by same author, Harmonic analysis: real-variable methods, orthogonality, and oscillatory integrals, Princeton Mathematical Series, vol. 43, Princeton University Press, Princeton, NJ, 1993.
  • [48] E. M. Stein and G. Weiss, Introduction to Fourier analysis on Euclidean spaces, Princeton University Press, Princeton, N.J., 1971, Princeton Mathematical Series, No. 32.
  • [49] J. O. Strömberg, Maximal functions associated to rectangles with uniformly distributed directions, Ann. of Math. (2) 107 (1978), no. 2, 399–402.
  • [50] T. Tao, The Bochner-Riesz conjecture implies the restriction conjecture, Duke Math. J. 96 (1999), no. 2, 363–375.
  • [51] by same author, On the maximal Bochner-Riesz conjecture in the plane for p<2p<2, Trans. Amer. Math. Soc. 354 (2002), no. 5, 1947–1959.
  • [52] T. Tao and A. Vargas, A bilinear approach to cone multipliers. I. Restriction estimates, Geom. Funct. Anal. 10 (2000), no. 1, 185–215.
  • [53] T. Tao, A. Vargas, and L. Vega, A bilinear approach to the restriction and Kakeya conjectures, J. Amer. Math. Soc. 11 (1998), no. 4, 967–1000.
  • [54] J. C. Várilly, An introduction to noncommutative geometry, EMS Series of Lectures in Mathematics, European Mathematical Society (EMS), Zürich, 2006.
  • [55] S. Wainger, Applications of Fourier transforms to averages over lower-dimensional sets, Proc. Sympos. Pure Math., XXXV, Part, Amer. Math. Soc., Providence, R.I., 1979, pp. 85–94.
  • [56] N. Weaver, Lipschitz algebras and derivations of von Neumann algebras, J. Funct. Anal. 139 (1996), no. 2, 261–300.
  • [57] by same author, Mathematical quantization, Studies in Advanced Mathematics, Chapman & Hall/CRC, Boca Raton, FL, 2001.
  • [58] T. H. Wolff, An improved bound for Kakeya type maximal functions, Rev. Mat. Iberoam. 11 (1995), no. 3, 651–674.
  • [59] by same author, Recent work connected with the Kakeya problem, Prospects in mathematics (Princeton, NJ, 1996), Amer. Math. Soc., Providence, RI, 1999, pp. 129–162.
  • [60] R. Xia, X. Xiong, and Q. Xu, Characterizations of operator-valued Hardy spaces and applications to harmonic analysis on quantum tori, Adv. Math. 291 (2016), 183–227.
  • [61] X. Xiong, Q. Xu, and Z. Yin, Sobolev, Besov and Triebel-Lizorkin spaces on quantum tori, Mem. Amer. Math. Soc. 252 (2018), no. 1203, vi+118.
  • [62] Q. Xu, Noncommutative Lp{L}_{p}-spaces and Martingale inequalities, Book manuscript (2007).