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

Small cap square function estimates

Shengwen Gan Shengwen Gan
Deparment of Mathematics
University of Wisconsin-Madison
Madison, WI-53706, USA
sgan7@wisc.edu
Abstract.

We introduce and prove small cap square function estimates for the unit parabola and the truncated light cone. More precisely, we study inequalities of the form

fpCα,p(R)(γΓα(R1)|fγ|2)1/2p,\|f\|_{p}\leq C_{\alpha,p}(R)\Big{\|}(\sum_{\gamma\in\Gamma_{\alpha}(R^{-1})}|f_{\gamma}|^{2})^{1/2}\Big{\|}_{p},

where Γα(R1)\Gamma_{\alpha}(R^{-1}) is the set of small caps of width RαR^{-\alpha}. We find sharp upper and lower bounds of the constant Cα,p(R)C_{\alpha,p}(R).

2020 Mathematics Subject Classification:
42B10

1. Introduction

In this paper, we study the square function estimates. We begin with the most general setting. Let Ωn\Omega\subset\mathbb{R}^{n} be a set in the frequency space, and suppose we are given a partition of Ω\Omega into subsets Σ={σ}\Sigma=\{\sigma\}:

Ω=σΣσ.\Omega=\bigsqcup_{\sigma\in\Sigma}\sigma.

We will only consider the case when σ\sigma are morally rectangles. For any function ff, we define fσ=(ψσf^)f_{\sigma}=(\psi_{\sigma}\widehat{f})^{\vee}, where ψσ\psi_{\sigma} is a smooth bump function adapted to σ\sigma. We will also assume suppf^Ω\textup{supp}\widehat{f}\subset\Omega in the following discussions. The inequality we are interested in is of the following form:

Square function estimate:

fpCp,Σ(σΣ|fσ|2)1/2p.\|f\|_{p}\leq C_{p,\Sigma}\|(\sum_{\sigma\in\Sigma}|f_{\sigma}|^{2})^{1/2}\|_{p}.

The goal is to find the best constant Cp,ΣC_{p,\Sigma} that works for all test functions ff.

This type of estimate is of huge interest in harmonic analysis. We briefly review some well-known results.

When Ω\Omega is the R1R^{-1}-neighborhood of the unit parabola 𝒫={(ξ,ξ2)2:|ξ|1}\mathcal{P}=\{(\xi,\xi^{2})\in\mathbb{R}^{2}:|\xi|\leq 1\} and Σ={σ}\Sigma=\{\sigma\} is the set of R1/2×R1\sim R^{-1/2}\times R^{-1}-caps that form a partition of Ω\Omega, then an argument of Córdoba-Fefferman (see also [1, Proposition 3.3]) gives

f4(σΣ|fσ|2)1/24.\|f\|_{4}\lesssim\|(\sum_{\sigma\in\Sigma}|f_{\sigma}|^{2})^{1/2}\|_{4}.

(Throughout this article, we suppress the \sim symbol for simplicity when the precise scale is unimportant.)

When Ω\Omega is the R1R^{-1}-neighborhood of the unit cone 𝒞={(ξ1,ξ2,ξ3)3:ξ3=ξ12+ξ22,1/2ξ31}\mathcal{C}=\{(\xi_{1},\xi_{2},\xi_{3})\in\mathbb{R}^{3}:\xi_{3}=\sqrt{\xi_{1}^{2}+\xi_{2}^{2}},1/2\leq\xi_{3}\leq 1\} and Σ={σ}\Sigma=\{\sigma\} are 1×R1/2×R11\times R^{-1/2}\times R^{-1}-caps that form a partition of Ω\Omega, then the sharp L4L^{4} square function estimate was proved by Guth-Wang-Zhang [6]:

f4(σΣ|fσ|2)1/24.\|f\|_{4}\lessapprox\|(\sum_{\sigma\in\Sigma}|f_{\sigma}|^{2})^{1/2}\|_{4}.

Here, ABA\lessapprox B means AϵRϵBA\lesssim_{\epsilon}R^{\epsilon}B for any ϵ>0\epsilon>0.

When Ω\Omega is certain neighborhood of a moment curve, it was studied by Gressman, Guo, Pierce, Roos and Yung [4]. The sharp L7L^{7} estimate was obtained by Maldague [7]. There are some other related results (see [8], [3]).

In the discussion above, we see that the size of caps in the partition of parabola is R1/2×R1R^{-1/2}\times R^{-1}; the size of caps in the partition of cone is 1×R1/2×R11\times R^{-1/2}\times R^{-1}. We usually call them the canonical partition. Besides the canonical partition of parabola and cone, Demeter, Guth and Wang [2] introduced the “small cap decoupling” which is the decoupling inequality for a finer partition than the canonical partition. Similarly, we can also ask the question about the small cap square function estimate.

The goal of this paper is to prove the sharp square function estimates for the small caps of parabola and cone. We will first define the small caps. Then we will introduce and study examples which give sharp lower bounds of the constants. Finally, we will prove the sharp bounds of the constants.

1.1. Small caps

1.1.1. Small caps for parabola

Let 𝒫:={(ξ,ξ2):ξ,|ξ|1}\mathcal{P}:=\{(\xi,\xi^{2}):\xi\in\mathbb{R},|\xi|\leq 1\} be the unit parabola, and NR1(𝒫)N_{R^{-1}}(\mathcal{P}) be its R1R^{-1}-neighborhood. For 1/2α11/2\leq\alpha\leq 1, let Γα(R1)\Gamma_{\alpha}(R^{-1}) be the partition of NR1(𝒫)N_{R^{-1}}(\mathcal{P}) into rectangular boxes of dimensions Rα×R1R^{-\alpha}\times R^{-1}. More precisely, each γΓα(R1)\gamma\in\Gamma_{\alpha}(R^{-1}) is of form

γ=(I×)NR1(𝒫),\gamma=(I\times\mathbb{R})\cap N_{R^{-1}}(\mathcal{P}),

where I[1,1]I\subset[-1,1] is an interval of length RαR^{-\alpha}. Note that we have #Γα(R1)Rα\#\Gamma_{\alpha}(R^{-1})\sim R^{\alpha}. Our square function estimate is

Theorem 1.

For suppf^NR1(𝒫)\textup{supp}\widehat{f}\subset N_{R^{-1}}(\mathcal{P}), we have

(1) fLp(2)Cα,p(R)(γΓα(R1)|fγ|2)1/2Lp(2),\|f\|_{L^{p}(\mathbb{R}^{2})}\lessapprox C_{\alpha,p}(R)\Big{\|}(\sum\limits_{\gamma\in\Gamma_{\alpha}(R^{-1})}|f_{\gamma}|^{2})^{1/2}\Big{\|}_{L^{p}(\mathbb{R}^{2})},

where

(2) Cα,p(R)={Rα(122p)p4α+2,R(α12)(121p) 2p4α+2.C_{\alpha,p}(R)=\begin{cases}R^{\alpha(\frac{1}{2}-\frac{2}{p})}\ \ \ \ \ p\geq 4\alpha+2,\\ R^{(\alpha-\frac{1}{2})(\frac{1}{2}-\frac{1}{p})}\ \ \ \ \ 2\leq p\leq 4\alpha+2.\end{cases}
Remark.

We remark that p4α+2p\geq 4\alpha+2 is equivalent to α(122p)(α12)(121p)\alpha(\frac{1}{2}-\frac{2}{p})\geq(\alpha-\frac{1}{2})(\frac{1}{2}-\frac{1}{p}). Therefore, (2) is equivalent to (up to constant) Cα,p(R)Rα(122p)+R(α12)(121p).C_{\alpha,p}(R)\sim R^{\alpha(\frac{1}{2}-\frac{2}{p})}+R^{(\alpha-\frac{1}{2})(\frac{1}{2}-\frac{1}{p})}.

1.1.2. Small caps for cone

Denote the truncated cone in 3\mathbb{R}^{3} by

𝒞:={(ξ1,ξ2,ξ3)3:ξ3=ξ12+ξ22,1/2ξ31}.\mathcal{C}:=\{(\xi_{1},\xi_{2},\xi_{3})\in\mathbb{R}^{3}:\xi_{3}=\sqrt{\xi_{1}^{2}+\xi_{2}^{2}},1/2\leq\xi_{3}\leq 1\}.

For 1/2β11/2\leq\beta\leq 1, let Γβ(R1)\Gamma_{\beta}(R^{-1}) be the partition of NR1(𝒞)N_{R^{-1}}(\mathcal{C}) into caps of dimensions 1×Rβ×R11\times R^{-\beta}\times R^{-1}. More precisely, we first choose a partition of 𝕊1\mathbb{S}^{1} into RβR^{-\beta}-arcs: 𝕊1=σ\mathbb{S}^{1}=\sqcup\sigma. For each arc σ\sigma, consider the R1R^{-1}-neighborhood of

{(ξ1,ξ2,ξ3)𝒞:(ξ1,ξ2)ξ12+ξ22σ},\{(\xi_{1},\xi_{2},\xi_{3})\in\mathcal{C}:\frac{(\xi_{1},\xi_{2})}{\sqrt{\xi_{1}^{2}+\xi_{2}^{2}}}\in\sigma\},

which is a cap of dimensions 1×Rβ×R11\times R^{-\beta}\times R^{-1}. Γβ(R1)\Gamma_{\beta}(R^{-1}) is the set of caps constructed in this way (see Figure 1). Note that #Γβ(R1)Rβ\#\Gamma_{\beta}(R^{-1})\sim R^{\beta}. Our square function estimate is

Figure 1. Small caps of the cone
Theorem 2.

For suppf^NR1(𝒞)\textup{supp}\widehat{f}\subset N_{R^{-1}}(\mathcal{C}), we have

(3) fLp(3)Cβ,p(R)(γΓβ(R1)|fγ|2)1/2Lp(3),\|f\|_{L^{p}(\mathbb{R}^{3})}\lessapprox C_{\beta,p}(R)\Big{\|}(\sum\limits_{\gamma\in\Gamma_{\beta}(R^{-1})}|f_{\gamma}|^{2})^{1/2}\Big{\|}_{L^{p}(\mathbb{R}^{3})},

where

(4) Cβ,p(R)={Rβ2p8,Rβ2+142p 4p8R(β12)(12p) 2p4.C_{\beta,p}(R)=\begin{cases}R^{\frac{\beta}{2}}\ \ \ \ \ p\geq 8,\\ R^{\frac{\beta}{2}+\frac{1}{4}-\frac{2}{p}}\ \ \ \ \ 4\leq p\leq 8\\ R^{(\beta-\frac{1}{2})(1-\frac{2}{p})}\ \ \ \ \ 2\leq p\leq 4.\end{cases}
Remark.

We remark that there is no interpolation argument in the proof of square function estimate. It is because that we cannot rewrite our square function estimate in the form of

TgXCgY,\|Tg\|_{X}\lesssim C\|g\|_{Y},

where X,YX,Y are some normed vector spaces and TT is a linear operator. Another way to see the interpolation argument is prohibited is by looking at the numerology in (4). We draw the graph of (1p,logRCβ,p(R))(\frac{1}{p},\log_{R}C_{\beta,p}(R)), where we ignore the CϵRϵC_{\epsilon}R^{\epsilon} factor in Cβ,p(R)C_{\beta,p}(R) (See Figure 2). We see the critical exponent p=8p=8 corresponds to a concave point (18,β2)(\frac{1}{8},\frac{\beta}{2}) in the graph. But if the interpolation argument works, then the graph should be convex which is a contradiction. Not being allowed to do interpolation will be the main difficulty in the proof. This means that we need to prove the estimate for all pp, but not only the critical pp. Let us consider the case β=1/2\beta=1/2. One critical exponent p=4p=4 was proved by Guth-Wang-Zhang [6]. The result for another critical exponent p=8p=8 and hence for p(4,8)p\in(4,8) is not included in [6]. We also remark that

Cβ,p(R)min{Rβ2,Rβ2+142p+R(β12)(121p)}.C_{\beta,p}(R)\sim\min\{R^{\frac{\beta}{2}},R^{\frac{\beta}{2}+\frac{1}{4}-\frac{2}{p}}+R^{(\beta-\frac{1}{2})(\frac{1}{2}-\frac{1}{p})}\}.
1p\frac{1}{p}14\frac{1}{4}12\frac{1}{2}logR(Cβ,p(R))\log_{R}\big{(}C_{\beta,p}(R)\big{)}
Figure 2.

1.2. Elementary tools

We briefly introduce the notion of dual rectangle and local orthogonality.

Definition 1.

Let RR be a rectangle of dimensions a×b×ca\times b\times c. Then the dual rectangle of RR, denoted by RR^{*}, is the rectangle centered at the origin of dimensions a1×b1×c1a^{-1}\times b^{-1}\times c^{-1}. Here RR^{*} is made from RR by letting the length of each edge of RR become the reciprocal.

From our definition, we see that if R2R_{2} is a translated copy of R1R_{1}, then R1=R2R_{1}^{*}=R_{2}^{*}. The motivation for defining dual rectangle is the following result.

Lemma 1.

For any rectangle RR, there exists a smooth function ωR\omega_{R} which satisfies 110𝟏R(x)ωR(x)10𝟏R(x)\frac{1}{10}\cdot{\bf 1}_{R}(x)\leq\omega_{R}(x)\leq 10\cdot{\bf 1}_{R}(x) for xRx\in R, and ωR\omega_{R} decays rapidly outside RR. Also, suppω^RR\textup{supp}\widehat{\omega}_{R}\subset R^{*}.

This lemma is very standard, so we omit the proof. The next result is the local orthogonality property.

Lemma 2.

Let RR be a rectangle and {fi}\{f_{i}\} is a set of functions. If {suppf^i+R}\{\textup{supp}\widehat{f}_{i}+R^{*}\} are finitely overlapping, then

(5) R|fi|2|fi|2|ωR|2.\int_{R}|\sum f_{i}|^{2}\lesssim\int\sum|f_{i}|^{2}|\omega_{R}|^{2}.
Proof.
R|fi|2|fiωR|2=|fiωR^|2.\displaystyle\int_{R}|\sum f_{i}|^{2}\lesssim\int|\sum f_{i}\omega_{R}|^{2}=\int|\sum\widehat{f_{i}\omega_{R}}|^{2}.

Note that fiωR^=fi^ωR^\widehat{f_{i}\omega_{R}}=\widehat{f_{i}}*\widehat{\omega_{R}} is supported in suppf^i+R\textup{supp}\widehat{f}_{i}+R^{*}. By the finitely overlapping property, we see the above is bounded by

|fiωR^|2=|fiωR|2.\lesssim\int\sum|\widehat{f_{i}\omega_{R}}|^{2}=\int\sum|f_{i}\omega_{R}|^{2}.

Remark.

Note that ωR\omega_{R} is essentially 𝟏R{\bf 1}_{R} by ignoring the rapidly decaying tail. It turns out that the tail is always harmless. Therefore, to get rid of some irrelevant technicalities, we will just ignore the rapidly decaying tail, and write (5) as

R|fi|2R|fi|2.\int_{R}|\sum f_{i}|^{2}\lesssim\int_{R}\sum|f_{i}|^{2}.

There is another notion called comparable. Given two rectangles R1,R2R_{1},R_{2}, we say R1R_{1} is essentially contained in R2R_{2}, if there exists a universal constant CC (say C=100C=100) such that

R1CR2.R_{1}\subset CR_{2}.

We say R1R_{1} and R2R_{2} are comparable if R1R_{1} is essentially contained in R2R_{2} and vice versa, i.e.,

1CR1R2CR1.\frac{1}{C}R_{1}\subset R_{2}\subset CR_{1}.

Throughout this paper, we will just ignore the unimportant constant CC, and just write R1R2R_{1}\subset R_{2} to denote that R1R_{1} is essentially contained in R2R_{2}.


Acknowledgement. The author would like to thank Prof. Larry Guth and Dominique Maldague for helpful discussions. The author also want to thank Changkeun Oh for the discussion of Theorem 1. The author also want to thank referees for carefully reading the manuscript and many helpful suggestions.

2. Small cap square function estimate for parabola

We prove Theorem 1 in this section. We begin with the sharp examples.

2.1. Sharp examples

There are two types of examples: concentrated example and flat example.

Case 1: p4α+2p\geq 4\alpha+2

We introduce the concentrated example. Choose ff such that f^(ξ)=ψNR1(𝒫)(ξ)\widehat{f}(\xi)=\psi_{N_{R^{-1}}(\mathcal{P})}(\xi), where ψNR1(𝒫)\psi_{N_{R^{-1}}(\mathcal{P})} is a smooth bump function supported in NR1(𝒫)N_{R^{-1}}(\mathcal{P}). We see that f(0)=f^(ξ)𝑑ξR1f(0)=\int\widehat{f}(\xi)d\xi\sim R^{-1}. Since f^\widehat{f} is supported in the unit ball centered at the origin, ff is locally constant in B(0,1)B(0,1). Therefore,

fpfLp(B(0,1))R1.\|f\|_{p}\geq\|f\|_{L^{p}(B(0,1))}\gtrsim R^{-1}.

We consider the right hand side of (1). By definition, for each γΓα(R1)\gamma\in\Gamma_{\alpha}(R^{-1}), f^γ\widehat{f}_{\gamma} is roughly a bump function supported in 2γ2\gamma. Let γ\gamma^{*} be the dual rectangle of γ\gamma which has dimensions Rα×RR^{\alpha}\times R and is centered at the origin. By an application of integration by parts and by ignoring the tails, we assume

fγ=1|γ|𝟏γ.f_{\gamma}=\frac{1}{|\gamma^{*}|}{\bf 1}_{\gamma^{*}}.

Here, “\approx” means up to a CϵRϵC_{\epsilon}R^{\epsilon} factor for any ϵ>0\epsilon>0. We will use the same notation throughout the paper.

We see that

(γΓα(R1)|fγ|2)1/2Lp(2)pR(1+α)pB(0,R)(γ𝟏γ)p/2.\Big{\|}(\sum\limits_{\gamma\in\Gamma_{\alpha}(R^{-1})}|f_{\gamma}|^{2})^{1/2}\Big{\|}_{L^{p}(\mathbb{R}^{2})}^{p}\sim R^{-(1+\alpha)p}\int_{B(0,R)}(\sum_{\gamma}{\bf 1}_{\gamma^{*}})^{p/2}.

We evaluate the integral above. There are two extreme regions: B(0,Rα)B(0,R^{\alpha}) where all the {γ}\{\gamma^{*}\} overlap; B(0,R)B(0,R/2)B(0,R)\setminus B(0,R/2) where {γ}\{\gamma^{*}\} is O(R2α1)O(R^{2\alpha-1})-overlapping. For the intermediate region B(0,r)B(0,r/2)B(0,r)\setminus B(0,r/2) (RαrRR^{\alpha}\leq r\leq R), we see that {γ}\{\gamma^{*}\} is O(r1R2α)O(r^{-1}R^{2\alpha})-overlapping. We may find a dyadic radius rr such that

(γ𝟏γ)p/2\displaystyle\int(\sum_{\gamma}{\bf 1}_{\gamma^{*}})^{p/2} B(0,r)B(0,r/2)(γ𝟏γ)p/2(r1R2α)p/2|B(0,r)|r2p2Rαp.\displaystyle\approx\int_{B(0,r)\setminus B(0,r/2)}(\sum_{\gamma}{\bf 1}_{\gamma^{*}})^{p/2}\lesssim(r^{-1}R^{2\alpha})^{p/2}|B(0,r)|\sim r^{2-\frac{p}{2}}R^{\alpha p}.

Since p4α+24p\geq 4\alpha+2\geq 4, the expression above is maximized when r=Rαr=R^{\alpha}. Plugging in, we obtain

(γ𝟏γ)p/2Rα(2+p2).\int(\sum_{\gamma}{\bf 1}_{\gamma^{*}})^{p/2}\lessapprox R^{\alpha(2+\frac{p}{2})}.

Plugging into (1), we have

R1Cα,p(R)R(1+α)Rα(2p+12),R^{-1}\lessapprox C_{\alpha,p}(R)R^{-(1+\alpha)}R^{\alpha(\frac{2}{p}+\frac{1}{2})},

which gives

Cα,p(R)Rα(122p).C_{\alpha,p}(R)\gtrapprox R^{\alpha(\frac{1}{2}-\frac{2}{p})}.

Case 2: 2p4α+22\leq p\leq 4\alpha+2

We introduce the flat example. Let θNR1(𝒫)\theta\subset N_{R^{-1}}(\mathcal{P}) be a R1/2×R1R^{-1/2}\times R^{-1}-cap. Choose ff such that f^(ξ)=ψθ(ξ)\widehat{f}(\xi)=\psi_{\theta}(\xi), where ψθ\psi_{\theta} is a smooth bump function supported in NR1(𝒫)N_{R^{-1}}(\mathcal{P}). Let θ\theta^{*} be the dual rectangle of θ\theta which has dimensions R1/2×RR^{1/2}\times R and is centered at the origin. By the locally constant property, ff is an L1L^{1} normalized function essentially supported in θ\theta^{*}. By ignoring the tails, we assume

f=1|θ|𝟏θ.f=\frac{1}{|\theta^{*}|}{\bf 1}_{\theta^{*}}.

We see that

fpR32R32p.\|f\|_{p}\sim R^{-\frac{3}{2}}R^{\frac{3}{2p}}.

We consider the right hand side of (1). By the same reasoning as in Case 1, for each γΓα(R1)\gamma\in\Gamma_{\alpha}(R^{-1}) with γθ\gamma\subset\theta, we know that f^γ\widehat{f}_{\gamma} is roughly a bump function supported in 2γ2\gamma. Therefore, we can assume

fγ=1|γ|𝟏γ.f_{\gamma}=\frac{1}{|\gamma^{*}|}{\bf 1}_{\gamma^{*}}.

We also note that γ1\gamma_{1}^{*} and γ2\gamma_{2}^{*} are comparable when γ1,γ2θ\gamma_{1},\gamma_{2}\subset\theta. We have

(γΓα(R1)|fγ|2)1/2Lp(2)\displaystyle\Big{\|}(\sum\limits_{\gamma\in\Gamma_{\alpha}(R^{-1})}|f_{\gamma}|^{2})^{1/2}\Big{\|}_{L^{p}(\mathbb{R}^{2})} R(1+α)((γθ𝟏γ)p/2)1/pR(1+α)#{γθ}1/2|γ|1/p\displaystyle\sim R^{-(1+\alpha)}\bigg{(}\int(\sum_{\gamma\subset\theta}{\bf 1}_{\gamma^{*}})^{p/2}\bigg{)}^{1/p}\sim R^{-(1+\alpha)}\#\{\gamma\subset\theta\}^{1/2}|\gamma^{*}|^{1/p}
R(1+α)R12(α12)R1+αp.\displaystyle\sim R^{-(1+\alpha)}R^{\frac{1}{2}(\alpha-\frac{1}{2})}R^{\frac{1+\alpha}{p}}.

Plugging into (1), we have

R32R32pCα,p(R)R(1+α)R12(α12)R1+αp,R^{-\frac{3}{2}}R^{\frac{3}{2p}}\lessapprox C_{\alpha,p}(R)R^{-(1+\alpha)}R^{\frac{1}{2}(\alpha-\frac{1}{2})}R^{\frac{1+\alpha}{p}},

which gives

Cα,p(R)R(α12)(121p).C_{\alpha,p}(R)\gtrapprox R^{(\alpha-\frac{1}{2})(\frac{1}{2}-\frac{1}{p})}.

2.2. Proof of Theorem 1

By the standard localization argument, it suffices to prove

fLp(BR)ϵ(Rα(122p)+R(α12)(121p))(γΓα(R1)|fγ|2)1/2p.\|f\|_{L^{p}(B_{R})}\lessapprox_{\epsilon}(R^{\alpha(\frac{1}{2}-\frac{2}{p})}+R^{(\alpha-\frac{1}{2})(\frac{1}{2}-\frac{1}{p})})\Big{\|}(\sum\limits_{\gamma\in\Gamma_{\alpha}(R^{-1})}|f_{\gamma}|^{2})^{1/2}\Big{\|}_{p}.

We introduce some notations. Throughout the proof, we use γ\gamma to denote caps of dimensions Rα×R1R^{-\alpha}\times R^{-1}. For R1/2Δ1R^{-1/2}\leq\Delta\leq 1, we will consider caps τ\tau of length Δ\Delta and thickness R1R^{-1}. We write |τ|=Δ|\tau|=\Delta to indicate the length of τ\tau. We will also partition the region BRB_{R} into rectangles of dimensions Rα×RR^{\alpha}\times R. For simplicity, we denote these rectangles by BRα×R.B_{R^{\alpha}\times R}. The longest direction of BRα×RB_{R^{\alpha}\times R} will be specified in the proof.

Let KlogRK\sim\log R and let mm\in\mathbb{N} be such that Km=R1/2K^{m}=R^{1/2}. By doing the broad-narrow reduction as in [2, Section 5.1] , we have

(6) fLp(BR)p\displaystyle\|f\|_{L^{p}(B_{R})}^{p} Cm|θ|=R1/2fθLp(BR)p\displaystyle\lesssim C^{m}\sum_{|\theta|=R^{-1/2}}\|f_{\theta}\|_{L^{p}(B_{R})}^{p}
(7) +CmKCR1/2Δ1Δ dyadic|τ|Δmaxτ1,τ2τ|τ1|=|τ2|=K1Δdist(τ1,τ2)(10K)1Δ(fτ1fτ2)1/2Lp(BR)p.\displaystyle+C^{m}K^{C}\sum_{\begin{subarray}{c}R^{-1/2}\leq\Delta\leq 1\\ \Delta\text{~{}dyadic}\end{subarray}}\sum_{|\tau|\sim\Delta}\max_{\begin{subarray}{c}\tau_{1},\tau_{2}\subset\tau\\ |\tau_{1}|=|\tau_{2}|=K^{-1}\Delta\\ \textup{dist}(\tau_{1},\tau_{2})\geq(10K)^{-1}\Delta\end{subarray}}\|(f_{\tau_{1}}f_{\tau_{2}})^{1/2}\|_{L^{p}(B_{R})}^{p}.

Note that CmKCϵRϵC^{m}K^{C}\lesssim_{\epsilon}R^{\epsilon}, for each ϵ>0\epsilon>0.

We first estimate the right hand side of (6).

Lemma 3.

Let θ\theta be a cap of length R1/2R^{-1/2}. Then,

fθLp(BR)R(α12)(121p)(γθ|fγ|2)1/2p.\|f_{\theta}\|_{L^{p}(B_{R})}\lesssim R^{(\alpha-\frac{1}{2})(\frac{1}{2}-\frac{1}{p})}\|(\sum_{\gamma\subset\theta}|f_{\gamma}|^{2})^{1/2}\|_{p}.
Proof.

We partition BRB_{R} into BRα×RB_{R^{\alpha}\times R}, where each BRα×RB_{R^{\alpha}\times R} is a translation of γ\gamma^{*} for γθ\gamma\subset\theta (note that for all γθ\gamma\subset\theta, γ\gamma^{*}’s are comparable). It suffices to prove for any BRα×RB_{R^{\alpha}\times R},

(8) fθLp(BRα×R)R(α12)(121p)(γθ|fγ|2)1/2Lp(ωBRα×R).\|f_{\theta}\|_{L^{p}(B_{R^{\alpha}\times R})}\lesssim R^{(\alpha-\frac{1}{2})(\frac{1}{2}-\frac{1}{p})}\|(\sum_{\gamma\subset\theta}|f_{\gamma}|^{2})^{1/2}\|_{L^{p}(\omega_{B_{R^{\alpha}\times R}})}.

Here, ωBRα×R\omega_{B_{R^{\alpha}\times R}} is a weight which =1=1 on BRα×RB_{R^{\alpha}\times R} and decays rapidly outside BRα×RB_{R^{\alpha}\times R}. And gLp(ω)\|g\|_{L^{p}(\omega)} is defined to be (|g|pω)1/p(\int|g|^{p}\omega)^{1/p}. We remark that we use ωBRα×R\omega_{B_{R^{\alpha}\times R}} instead of 𝟏BRα×R{\bf 1}_{B_{R^{\alpha}\times R}} is to make the local orthogonality and locally constant property rigorous. As such technicality is well-known (see for example in [1]), we will just pretend ωBRα×R=𝟏BRα×R\omega_{B_{R^{\alpha}\times R}}={\bf 1}_{B_{R^{\alpha}\times R}} for convenience.

We further do the partition

BRα×R=BR1/2×R,B_{R^{\alpha}\times R}=\bigsqcup B_{R^{1/2}\times R},

where each BR1/2×RB_{R^{1/2}\times R} is a translation of θ\theta^{*}. Since fθf_{\theta} is locally constant on each BR1/2×RB_{R^{1/2}\times R}, we have

fθLp(BRα×R)\displaystyle\|f_{\theta}\|_{L^{p}(B_{R^{\alpha}\times R})} =(BR1/2×RfθLp(BR1/2×R)p)1/p\displaystyle=\bigg{(}\sum_{B_{R^{1/2}\times R}}\|f_{\theta}\|_{L^{p}(B_{R^{1/2}\times R})}^{p}\bigg{)}^{1/p}
R32(1p12)(BR1/2×RfθL2(BR1/2×R)p)1/p\displaystyle\lesssim R^{\frac{3}{2}(\frac{1}{p}-\frac{1}{2})}\bigg{(}\sum_{B_{R^{1/2}\times R}}\|f_{\theta}\|_{L^{2}(B_{R^{1/2}\times R})}^{p}\bigg{)}^{1/p}
R32(1p12)fθL2(BRα×R).\displaystyle\leq R^{\frac{3}{2}(\frac{1}{p}-\frac{1}{2})}\|f_{\theta}\|_{L^{2}(B_{R^{\alpha}\times R})}.

By local orthogonality, Hölder’s inequality and noting p2p\geq 2, we have

fθL2(BRα×R)(γθ|fγ|2)1/2L2(BRα×R)R(1+α)(121p)(γθ|fγ|2)1/2Lp(BRα×R).\|f_{\theta}\|_{L^{2}(B_{R^{\alpha}\times R})}\lesssim\|(\sum_{\gamma\subset\theta}|f_{\gamma}|^{2})^{1/2}\|_{L^{2}(B_{R^{\alpha}\times R})}\leq R^{(1+\alpha)(\frac{1}{2}-\frac{1}{p})}\|(\sum_{\gamma\subset\theta}|f_{\gamma}|^{2})^{1/2}\|_{L^{p}(B_{R^{\alpha}\times R})}.

Combining the inequalities, we finish the proof of (8). ∎

By Lemma 3, the right hand side of (6) is bounded by

RϵR(α12)(121p)(θ(γθ|fγ|2)1/2pp)1/pCα,p(R)(γ|fγ|2)1/2p.R^{\epsilon}R^{(\alpha-\frac{1}{2})(\frac{1}{2}-\frac{1}{p})}\bigg{(}\sum_{\theta}\|(\sum_{\gamma\subset\theta}|f_{\gamma}|^{2})^{1/2}\|_{p}^{p}\bigg{)}^{1/p}\leq C_{\alpha,p}(R)\|(\sum_{\gamma}|f_{\gamma}|^{2})^{1/2}\|_{p}.

Next, we estimate (7). For any summand in (7), we will show that

(9) (fτ1fτ2)1/2Lp(BR)Cα,p(R)(γτ|fγ|2)1/2p.\|(f_{\tau_{1}}f_{\tau_{2}})^{1/2}\|_{L^{p}(B_{R})}\lesssim C_{\alpha,p}(R)\|(\sum_{\gamma\subset\tau}|f_{\gamma}|^{2})^{1/2}\|_{p}.

This will imply (7)1p{}^{\frac{1}{p}} Cα,p(R)(γ|fγ|2)1/2p\lessapprox C_{\alpha,p}(R)\|(\sum\limits_{\gamma}|f_{\gamma}|^{2})^{1/2}\|_{p}, and then finishes the proof of Theorem 1. It remains to prove (9).

Fix a Δ[R1/2,1]\Delta\in[R^{-1/2},1] and a τ\tau with |τ|=Δ|\tau|=\Delta. We first consider γτγ\bigcap_{\gamma\subset\tau}\gamma^{*}. It is easy to see γτγ\bigcap_{\gamma\subset\tau}\gamma^{*} is an Rα×RαΔ1R^{\alpha}\times R^{\alpha}\Delta^{-1}-rectangle when ΔRα1\Delta\geq R^{\alpha-1}; γτγ\bigcap_{\gamma\subset\tau}\gamma^{*} is an Rα×RR^{\alpha}\times R-rectangle when ΔRα1\Delta\leq R^{\alpha-1}. We consider these two cases separately.

Case 1: ΔRα1\Delta\geq R^{\alpha-1}

We choose a partition BR=BRα×RαΔ1B_{R}=\bigsqcup B_{R^{\alpha}\times R^{\alpha}\Delta^{-1}}, where each BRα×RαΔ1B_{R^{\alpha}\times R^{\alpha}\Delta^{-1}} is a translation of γτγ\bigcap_{\gamma\subset\tau}\gamma^{*}. We just need to show

(10) (fτ1fτ2)1/2Lp(BRα×RαΔ1)Cα,p(R)(γτ|fγ|2)1/2Lp(BRα×RαΔ1).\|(f_{\tau_{1}}f_{\tau_{2}})^{1/2}\|_{L^{p}(B_{R^{\alpha}\times R^{\alpha}\Delta^{-1}})}\lesssim C_{\alpha,p}(R)\|(\sum_{\gamma\subset\tau}|f_{\gamma}|^{2})^{1/2}\|_{L^{p}(B_{R^{\alpha}\times R^{\alpha}\Delta^{-1}})}.

Since each |fγ||f_{\gamma}| is locally constant on BRα×RαΔ1B_{R^{\alpha}\times R^{\alpha}\Delta^{-1}} when γτ\gamma\subset\tau, we have

(γτ|fγ|2)1/2Lp(BRα×RαΔ1)(R2αΔ1)12+1p(γτ|fγ|2)1/2L2(BRα×RαΔ1).\|(\sum_{\gamma\subset\tau}|f_{\gamma}|^{2})^{1/2}\|_{L^{p}(B_{R^{\alpha}\times R^{\alpha}\Delta^{-1}})}\sim(R^{2\alpha}\Delta^{-1})^{-\frac{1}{2}+\frac{1}{p}}\|(\sum_{\gamma\subset\tau}|f_{\gamma}|^{2})^{1/2}\|_{L^{2}(B_{R^{\alpha}\times R^{\alpha}\Delta^{-1}})}.

Since {fγ}γτ\{f_{\gamma}\}_{\gamma\subset\tau} are locally orthogonal on BRα×RαΔ1B_{R^{\alpha}\times R^{\alpha}\Delta^{-1}}, we have

(γτ|fγ|2)1/2L2(BRα×RαΔ1)fτL2(BRα×RαΔ1).\|(\sum_{\gamma\subset\tau}|f_{\gamma}|^{2})^{1/2}\|_{L^{2}(B_{R^{\alpha}\times R^{\alpha}\Delta^{-1}})}\sim\|f_{\tau}\|_{L^{2}(B_{R^{\alpha}\times R^{\alpha}\Delta^{-1}})}.

Therefore, (10) is reduced to

(11) (fτ1fτ2)1/2Lp(BRα×RαΔ1)Cα,p(R)(R2αΔ1)12+1pfτL2(BRα×RαΔ1).\|(f_{\tau_{1}}f_{\tau_{2}})^{1/2}\|_{L^{p}(B_{R^{\alpha}\times R^{\alpha}\Delta^{-1}})}\lesssim C_{\alpha,p}(R)(R^{2\alpha}\Delta^{-1})^{-\frac{1}{2}+\frac{1}{p}}\|f_{\tau}\|_{L^{2}(B_{R^{\alpha}\times R^{\alpha}\Delta^{-1}})}.

Next, we apply the parabolic rescaling. Recall that τ\tau is a cap of length Δ\Delta. We dilate by factor Δ1\Delta^{-1} in the tangent direction of τ\tau and dilate by factor Δ2\Delta^{-2} in the normal direction of τ\tau. Under the rescaling, we see that: τ\tau becomes the R1Δ2R^{-1}\Delta^{-2}-neighborhood of 𝒫\mathcal{P}; τ1\tau_{1} and τ2\tau_{2} become K1K^{-1}-separated caps with length K1K^{-1} and thickness R1Δ2R^{-1}\Delta^{-2}; the rectangle BRα×RαΔ1B_{R^{\alpha}\times R^{\alpha}\Delta^{-1}} in the physical space becomes BRαΔB_{R^{\alpha}\Delta}. Let g,g1,g2g,g_{1},g_{2} be the rescaled version of fτ,fτ1,fτ2f_{\tau},f_{\tau_{1}},f_{\tau_{2}} respectively. The inequality (11) becomes

(12) (g1g2)1/2Lp(BRαΔ)Cα,p(R)(R2αΔ1)12+1pΔ3(12+1p)gL2(BRαΔ).\|(g_{1}g_{2})^{1/2}\|_{L^{p}(B_{R^{\alpha}\Delta})}\lesssim C_{\alpha,p}(R)(R^{2\alpha}\Delta^{-1})^{-\frac{1}{2}+\frac{1}{p}}\Delta^{3(-\frac{1}{2}+\frac{1}{p})}\|g\|_{L^{2}(B_{R^{\alpha}\Delta})}.

We recall the following bilinear restriction estimate (see for example in [9]).

Lemma 4.

Let r>1r>1, K>1K>1. Suppose g1,g2g_{1},g_{2} satisfy suppg^1,suppg^2Nr2(𝒫)\textup{supp}\widehat{g}_{1},\textup{supp}\widehat{g}_{2}\subset N_{r^{-2}}(\mathcal{P}) and dist(suppg^1,suppg^2)>K1\textup{dist}(\textup{supp}\widehat{g}_{1},\textup{supp}\widehat{g}_{2})>K^{-1}. Then for p2p\geq 2 and rrr^{\prime}\geq r we have

(13) (g1g2)1/2Lp(Br)KO(1)r2p1(g1L2(Br)g2L2(Br))1/2.\|(g_{1}g_{2})^{1/2}\|_{L^{p}(B_{r^{\prime}})}\lesssim K^{O(1)}r^{\frac{2}{p}-1}\big{(}\|g_{1}\|_{L^{2}(B_{r^{\prime}})}\|g_{2}\|_{L^{2}(B_{r^{\prime}})}\big{)}^{1/2}.
Proof.

We just need to prove for r=rr^{\prime}=r. When p=2p=2, this is trivial. When p=4p=4, this is the bilinear restriction estimate. When p=p=\infty, we note that

(g1g2)1/2L(Br)2\displaystyle\|(g_{1}g_{2})^{1/2}\|_{L^{\infty}(B_{r})}^{2} g1L(Br)g2L(Br)g^1L1g^2L1\displaystyle\leq\|g_{1}\|_{L^{\infty}(B_{r})}\|g_{2}\|_{L^{\infty}(B_{r})}\leq\|\widehat{g}_{1}\|_{L^{1}}\|\widehat{g}_{2}\|_{L^{1}}
r2g^1L2g^1L2=r2g1L2g2L2.\displaystyle\lesssim r^{-2}\|\widehat{g}_{1}\|_{L^{2}}\|\widehat{g}_{1}\|_{L^{2}}=r^{-2}\|g_{1}\|_{L^{2}}\|g_{2}\|_{L^{2}}.

The second-last inequality is by Hölder and the condition on the support of g^1,g^2\widehat{g}_{1},\widehat{g}_{2}. The last inequality is by Plancherel. For other pp, the proof is by using Hölder to interpolate between p=2,4,p=2,4,\infty. ∎

We return to (12). Noting that RαΔ(RΔ2)1/2R^{\alpha}\Delta\geq(R\Delta^{2})^{1/2}, we apply the lemma above to bound the left hand side of (12) by (RαΔ)2p1gL2(BRαΔ)(R^{\alpha}\Delta)^{\frac{2}{p}-1}\|g\|_{L^{2}(B_{R^{\alpha}\Delta})}. It suffices to prove

(14) (RαΔ)2p1Cα,p(R)(R2αΔ1)12+1pΔ3(12+1p).(R^{\alpha}\Delta)^{\frac{2}{p}-1}\lesssim C_{\alpha,p}(R)(R^{2\alpha}\Delta^{-1})^{-\frac{1}{2}+\frac{1}{p}}\Delta^{3(-\frac{1}{2}+\frac{1}{p})}.

When p4p\geq 4, we use Cα,p(R)Rα(122p)C_{\alpha,p}(R)\gtrsim R^{\alpha(\frac{1}{2}-\frac{2}{p})}. Then (14) boils down to

(15) (RαΔ)2p1Rα(122p)(R2αΔ1)12+1pΔ3(12+1p),(R^{\alpha}\Delta)^{\frac{2}{p}-1}\lesssim R^{\alpha(\frac{1}{2}-\frac{2}{p})}(R^{2\alpha}\Delta^{-1})^{-\frac{1}{2}+\frac{1}{p}}\Delta^{3(-\frac{1}{2}+\frac{1}{p})},

which is equivalent to

Rα(122p)1,R^{\alpha(\frac{1}{2}-\frac{2}{p})}\gtrsim 1,

which is true since R1R\geq 1.

When p4p\leq 4, we use Cα,p(R)R(α12)(121p)C_{\alpha,p}(R)\gtrsim R^{(\alpha-\frac{1}{2})(\frac{1}{2}-\frac{1}{p})}. Then (14) boils down to

(16) (RαΔ)2p1R(α12)(121p)(R2αΔ1)12+1pΔ3(12+1p),(R^{\alpha}\Delta)^{\frac{2}{p}-1}\lesssim R^{(\alpha-\frac{1}{2})(\frac{1}{2}-\frac{1}{p})}(R^{2\alpha}\Delta^{-1})^{-\frac{1}{2}+\frac{1}{p}}\Delta^{3(-\frac{1}{2}+\frac{1}{p})},

which is equivalent to

R(α12)(121p)1,R^{(\alpha-\frac{1}{2})(\frac{1}{2}-\frac{1}{p})}\gtrsim 1,

which is true since α1/2.\alpha\geq 1/2.

Case 2: ΔRα1\Delta\leq R^{\alpha-1}

We choose a partition BR=BRα×RB_{R}=\bigsqcup B_{R^{\alpha}\times R}, where each BRα×RB_{R^{\alpha}\times R} is a translation of γτγ\bigcap_{\gamma\subset\tau}\gamma^{*}. We just need to show

(17) (fτ1fτ2)1/2Lp(BRα×R)Cα,p(R)(γτ|fγ|2)1/2Lp(BRα×R).\|(f_{\tau_{1}}f_{\tau_{2}})^{1/2}\|_{L^{p}(B_{R^{\alpha}\times R})}\lesssim C_{\alpha,p}(R)\|(\sum_{\gamma\subset\tau}|f_{\gamma}|^{2})^{1/2}\|_{L^{p}(B_{R^{\alpha}\times R})}.

Since each |fγ||f_{\gamma}| is locally constant on BRα×RB_{R^{\alpha}\times R} when γτ\gamma\subset\tau, we have

(γτ|fγ|2)1/2Lp(BRα×R)(Rα+1)12+1p(γτ|fγ|2)1/2L2(BRα×R).\|(\sum_{\gamma\subset\tau}|f_{\gamma}|^{2})^{1/2}\|_{L^{p}(B_{R^{\alpha}\times R})}\sim(R^{\alpha+1})^{-\frac{1}{2}+\frac{1}{p}}\|(\sum_{\gamma\subset\tau}|f_{\gamma}|^{2})^{1/2}\|_{L^{2}(B_{R^{\alpha}\times R})}.

Since {fγ}γτ\{f_{\gamma}\}_{\gamma\subset\tau} are locally orthogonal on BRα×RB_{R^{\alpha}\times R}, we have

(γτ|fγ|2)1/2L2(BRα×R)fτL2(BRα×R).\|(\sum_{\gamma\subset\tau}|f_{\gamma}|^{2})^{1/2}\|_{L^{2}(B_{R^{\alpha}\times R})}\sim\|f_{\tau}\|_{L^{2}(B_{R^{\alpha}\times R})}.

Therefore, (17) is reduced to

(18) (fτ1fτ2)1/2Lp(BRα×R)Cα,p(R)(Rα+1)12+1pfτL2(BRα×R).\|(f_{\tau_{1}}f_{\tau_{2}})^{1/2}\|_{L^{p}(B_{R^{\alpha}\times R})}\lesssim C_{\alpha,p}(R)(R^{\alpha+1})^{-\frac{1}{2}+\frac{1}{p}}\|f_{\tau}\|_{L^{2}(B_{R^{\alpha}\times R})}.

Next, we do the same parabolic rescaling as above. The rectangle BRα×RB_{R^{\alpha}\times R} in the physical space becomes BRαΔ×RΔ2B_{R^{\alpha}\Delta\times R\Delta^{2}}. Let g,g1,g2g,g_{1},g_{2} be the rescaled version of fτ,fτ1,fτ2f_{\tau},f_{\tau_{1}},f_{\tau_{2}} respectively. The inequality (18) becomes

(19) (g1g2)1/2Lp(BRαΔ×RΔ2)Cα,p(R)(Rα+1)12+1pΔ3(12+1p)gL2(BRαΔ×RΔ2).\|(g_{1}g_{2})^{1/2}\|_{L^{p}(B_{R^{\alpha}\Delta\times R\Delta^{2}})}\lesssim C_{\alpha,p}(R)(R^{\alpha+1})^{-\frac{1}{2}+\frac{1}{p}}\Delta^{3(-\frac{1}{2}+\frac{1}{p})}\|g\|_{L^{2}(B_{R^{\alpha}\Delta\times R\Delta^{2}})}.

To apply Lemma 4, we do the partition BRαΔ×RΔ2=BRΔ2B_{R^{\alpha}\Delta\times R\Delta^{2}}=\bigsqcup B_{R\Delta^{2}}. So, (19) is reduced to

(20) (g1g2)1/2Lp(BRΔ2)Cα,p(R)(Rα+1)12+1pΔ3(12+1p)gL2(BRΔ2).\|(g_{1}g_{2})^{1/2}\|_{L^{p}(B_{R\Delta^{2}})}\lesssim C_{\alpha,p}(R)(R^{\alpha+1})^{-\frac{1}{2}+\frac{1}{p}}\Delta^{3(-\frac{1}{2}+\frac{1}{p})}\|g\|_{L^{2}(B_{R\Delta^{2}})}.

By Lemma 4,

(g1g2)1/2Lp(BRΔ2)(RΔ2)2p1gL2(BRΔ2).\|(g_{1}g_{2})^{1/2}\|_{L^{p}(B_{R\Delta^{2}})}\lesssim(R\Delta^{2})^{\frac{2}{p}-1}\|g\|_{L^{2}(B_{R\Delta^{2}})}.

It suffices to prove

(21) (RΔ2)2p1Cα,p(R)R(α+1)(12+1p)Δ3(12+1p).(R\Delta^{2})^{\frac{2}{p}-1}\lesssim C_{\alpha,p}(R)R^{(\alpha+1)(-\frac{1}{2}+\frac{1}{p})}\Delta^{3(-\frac{1}{2}+\frac{1}{p})}.

When p4α+2p\geq 4\alpha+2, we use Cα,p(R)Rα(122p)C_{\alpha,p}(R)\gtrsim R^{\alpha(\frac{1}{2}-\frac{2}{p})}. Then (21) boils down to

(22) (RΔ2)2p1Rα(122p)R(α+1)(12+1p)Δ3(12+1p),(R\Delta^{2})^{\frac{2}{p}-1}\lesssim R^{\alpha(\frac{1}{2}-\frac{2}{p})}R^{(\alpha+1)(-\frac{1}{2}+\frac{1}{p})}\Delta^{3(-\frac{1}{2}+\frac{1}{p})},

which is equivalent to

Δ1p12Rαp+121p.\Delta^{\frac{1}{p}-\frac{1}{2}}\lesssim R^{-\frac{\alpha}{p}+\frac{1}{2}-\frac{1}{p}}.

Using ΔR12\Delta\geq R^{-\frac{1}{2}}, we just need to prove

R12p+14Rαp+121p.R^{-\frac{1}{2p}+\frac{1}{4}}\lesssim R^{-\frac{\alpha}{p}+\frac{1}{2}-\frac{1}{p}}.

The last inequality is equivalent to 1412pαp0\frac{1}{4}-\frac{1}{2p}-\frac{\alpha}{p}\geq 0, which is further equivalent to p4α+2p\geq 4\alpha+2. We also remark that this is the place where the critical exponent p=4α+2p=4\alpha+2 appears.

When 2p4α+22\leq p\leq 4\alpha+2, we use Cα,p(R)R(α12)(121p)C_{\alpha,p}(R)\gtrsim R^{(\alpha-\frac{1}{2})(\frac{1}{2}-\frac{1}{p})}. Then (21) boils down to

(23) (RΔ2)2p1R(α12)(121p)R(α+1)(12+1p)Δ3(12+1p),(R\Delta^{2})^{\frac{2}{p}-1}\lesssim R^{(\alpha-\frac{1}{2})(\frac{1}{2}-\frac{1}{p})}R^{(\alpha+1)(-\frac{1}{2}+\frac{1}{p})}\Delta^{3(-\frac{1}{2}+\frac{1}{p})},

which is equivalent to

Δ1p12R1412p,\Delta^{\frac{1}{p}-\frac{1}{2}}\lesssim R^{\frac{1}{4}-\frac{1}{2p}},

which is true since Δ1R1/2.\Delta^{-1}\leq R^{1/2}.

The proof of Theorem 1 is finished.

3. Small cap square function estimate for cone

We prove Theorem 2 in this section. We begin with the sharp examples.

3.1. Sharp examples

Choose ff such that f^=ψNR1(𝒞)(ξ)\widehat{f}=\psi_{N_{R^{-1}}(\mathcal{C})}(\xi), where ψNR1(𝒞)(ξ)\psi_{N_{R^{-1}}(\mathcal{C})}(\xi) is a smooth bump function supported in NR1(𝒞)N_{R^{-1}}(\mathcal{C}). We are going to calculate the lower bound of fp\|f\|_{p}, which is the left hand side of (3). We see that f(0)=f^(ξ)𝑑ξR1f(0)=\int\widehat{f}(\xi)d\xi\sim R^{-1}. Since f^\widehat{f} is supported in the unit ball centered at the origin, ff is locally constant in B(0,1)B(0,1). Therefore,

(24) fpfLp(B(0,1))R1.\|f\|_{p}\gtrsim\|f\|_{L^{p}(B(0,1))}\gtrsim R^{-1}.

We also estimate the integral of ff in the region {|x|R}\{|x|\sim R\}. We first do a canonical partition of NR1(𝒞)N_{R^{-1}}(\mathcal{C}) into 1×R1/2×R11\times R^{-1/2}\times R^{-1}-planks, denoted by

NR1(𝒞)=θ.N_{R^{-1}}(\mathcal{C})=\bigsqcup\theta.

Then we can write f=θfθf=\sum_{\theta}f_{\theta}, such that each f^θ\widehat{f}_{\theta} is a smooth bump function on θ\theta. Let θ\theta^{*} be the dual rectangle of θ\theta, so θ\theta^{*} has size 1×R1/2×R1\times R^{1/2}\times R and is centered at the origin. By an application of integration by parts, we can assume

|fθ|=1|θ|𝟏θ=R3/2𝟏θ.|f_{\theta}|=\frac{1}{|\theta^{*}|}{\bf 1}_{\theta^{*}}=R^{-3/2}{\bf 1}_{\theta^{*}}.

Now the key observation is that {θ}\{\theta^{*}\} are disjoint in B(0,R)B(0,910R)B(0,R)\setminus B(0,\frac{9}{10}R), so we see that

fp=θfθp\displaystyle\|f\|_{p}=\|\sum_{\theta}f_{\theta}\|_{p} θfθLp(B(0,R)B(0,910R))\displaystyle\geq\|\sum_{\theta}f_{\theta}\|_{L^{p}(B(0,R)\setminus B(0,\frac{9}{10}R))}
R3/2θ𝟏θLp(B(0,R)B(0,910R))\displaystyle\sim R^{-3/2}\|\sum_{\theta}{\bf 1}_{\theta^{*}}\|_{L^{p}(B(0,R)\setminus B(0,\frac{9}{10}R))}
R3/2(θ|θ|)1/p=R32+2p.\displaystyle\sim R^{-3/2}(\sum_{\theta}|\theta^{*}|)^{1/p}=R^{-\frac{3}{2}+\frac{2}{p}}.

Combining with (24), we see

(25) fpmax{R1,R32+2p}.\|f\|_{p}\gtrsim\max\{R^{-1},R^{-\frac{3}{2}+\frac{2}{p}}\}.

And we see the threshold for these two lower bounds to be equal is at p=4p=4.

For this same ff, we will estimate the upper bound of the right hand side of (3). Recall that γ\gamma is a 1×Rβ×R11\times R^{-\beta}\times R^{-1}-cap contained in NR1(𝒞)N_{R^{-1}}(\mathcal{C}), and by definition f^γ=ψγf^\widehat{f}_{\gamma}=\psi_{\gamma}\widehat{f}. Therefore, f^γ\widehat{f}_{\gamma} is a smooth bump function adapted to γ\gamma. By an application of integration by parts, we can assume

|fγ|=1|γ|𝟏γ.|f_{\gamma}|=\frac{1}{|\gamma^{*}|}{\bf 1}_{\gamma^{*}}.

Here, the dual rectangle γ\gamma^{*} is centered at the origin with size 1×Rβ×R1\times R^{\beta}\times R. See Figure 3: the rectangle on the left hand side is γ\gamma; the rectangle on the right hand side is γ\gamma^{*}.

Therefore, we can write

(26) (γΓβ(R1)|fγ|2)1/2pR1β((γ𝟏γ)p/2)1/p.\|(\sum_{\gamma\in\Gamma_{\beta}(R^{-1})}|f_{\gamma}|^{2})^{1/2}\|_{p}\sim R^{-1-\beta}\big{(}\int(\sum_{\gamma}{\bf 1}_{\gamma^{*}})^{p/2}\big{)}^{1/p}.
Figure 3.

Note that each γ\gamma^{*} is supported in B(0,R)B(0,R), so we rewrite

(γ𝟏γ)p/2=|r|R𝑑r{x3=r}(γ𝟏γ)p/2𝑑x1𝑑x2.\int(\sum_{\gamma}{\bf 1}_{\gamma^{*}})^{p/2}=\int_{|r|\leq R}dr\int_{\{x_{3}=r\}}(\sum_{\gamma}{\bf 1}_{\gamma^{*}})^{p/2}dx_{1}dx_{2}.

We are going to calculate {x3=r}(γ𝟏γ)p/2\int_{\{x_{3}=r\}}(\sum_{\gamma}{\bf 1}_{\gamma^{*}})^{p/2}. Here is the result:

Proposition 1.

For p2p\geq 2, we have

(27) {x3=r}(γ𝟏γ)p/2{R2β+Rpβ20r10,r1p4Rβp2+r2p2Rβp2+R2β10rRβ,r1p4Rβp2+R2βRβrR.\int_{\{x_{3}=r\}}(\sum_{\gamma}{\bf 1}_{\gamma^{*}})^{p/2}\approx\begin{cases}R^{2\beta}+R^{\frac{p\beta}{2}}&0\leq r\leq 10,\\ r^{1-\frac{p}{4}}R^{\beta\frac{p}{2}}+r^{2-\frac{p}{2}}R^{\beta\frac{p}{2}}+R^{2\beta}&10\leq r\leq R^{\beta},\\ r^{1-\frac{p}{4}}R^{\beta\frac{p}{2}}+R^{2\beta}&R^{\beta}\leq r\leq R.\end{cases}
Figure 4.
Proof.

Fix the plane {x3=r}\{x_{3}=r\}. For each γ\gamma^{*}, we set

γr:=γ{x3=r}.\gamma^{*}_{r}:=\gamma^{*}\cap\{x_{3}=r\}.

γr\gamma^{*}_{r} is a rectangle of size 1×Rβ1\times R^{\beta} in the plane {x3=r}\{x_{3}=r\}. Denote the center of γr\gamma^{*}_{r} by C(γr)C(\gamma^{*}_{r}). We see that C(γr)C(\gamma^{*}_{r}) lies on the circle

Sr:={x3=r,x12+x22=r},S_{r}:=\{x_{3}=r,\sqrt{x_{1}^{2}+x_{2}^{2}}=r\},

and the long direction of γr\gamma^{*}_{r} is tangent to SrS_{r} (see Figure 4). We can rewrite the left hand side of (27) as

2(γ𝟏γr)p/2.\int_{\mathbb{R}^{2}}(\sum_{\gamma}{\bf 1}_{\gamma^{*}_{r}})^{p/2}.

We also notice two useful facts: (1) #{γr}Rβ\#\{\gamma^{*}_{r}\}\sim R^{\beta}; (2) {C(γr)}\{C(\gamma^{*}_{r})\} are roughly rRβrR^{-\beta}-separated on the circle SrS_{r}.

Case 1: 0r100\leq r\leq 10

In this case, we see that {γr}\{\gamma^{*}_{r}\} essentially form a bush centered at the origin. Evaluating the concentrated part and spread-out part, we have

2(γ𝟏γr)p/2B(0,1)(γ𝟏γr)p/2+B(0,Rβ)B(0,12Rβ)(γ𝟏γr)p/2Rpβ2+R2β.\int_{\mathbb{R}^{2}}(\sum_{\gamma}{\bf 1}_{\gamma^{*}_{r}})^{p/2}\approx\int_{B(0,1)}(\sum_{\gamma}{\bf 1}_{\gamma^{*}_{r}})^{p/2}+\int_{B(0,R^{\beta})\setminus B(0,\frac{1}{2}R^{\beta})}(\sum_{\gamma}{\bf 1}_{\gamma^{*}_{r}})^{p/2}\sim R^{\frac{p\beta}{2}}+R^{2\beta}.

Case 2: 10rRβ10\leq r\leq R^{\beta}

For any point PγrP\in\bigcup\gamma^{*}_{r}, we are going to estimate γ𝟏γr(P)\sum_{\gamma}{\bf 1}_{\gamma^{*}_{r}}(P). Define

d(P):=dist(P,Sr).d(P):=\textup{dist}(P,S_{r}).

We see that any PγrP\in\bigcup\gamma^{*}_{r} satisfies d(P)Rβd(P)\lesssim R^{\beta}, and if PγrP\in\bigcup\gamma^{*}_{r} lies inside SrS_{r} then d(P)=0d(P)=0. For simplicity, we write d=d(P)d=d(P). We consider several cases:

P
Figure 5.
  1. (1)

    d10d\leq 10. In this case, PP lies in the 1010-neighborhood of SrS_{r}. Therefore,

    γ𝟏γr(P)=γ𝟏γrN10(Sr)(P)\sum_{\gamma}{\bf 1}_{\gamma^{*}_{r}}(P)=\sum_{\gamma}{\bf 1}_{\gamma^{*}_{r}\cap N_{10}(S_{r})}(P)

    Noting that γrN10(Sr)\gamma^{*}_{r}\cap N_{10}(S_{r}) is essentially a 1×r1/21\times r^{1/2}-rectangle centered at C(γr)(Sr)C(\gamma^{*}_{r})(\in S_{r}) and noting that {C(γr)}\{C(\gamma^{*}_{r})\} are rRβrR^{-\beta} separated, we have

    γ𝟏γrN10(Sr)(P)r1/2rRβ=r1/2Rβ.\sum_{\gamma}{\bf 1}_{\gamma^{*}_{r}\cap N_{10}(S_{r})}(P)\sim\frac{r^{1/2}}{rR^{-\beta}}=r^{-1/2}R^{\beta}.
  2. (2)

    10dr10\leq d\leq r. We claim in this case

    γ𝟏γr(P)Rβ(rd)1/2.\sum_{\gamma}{\bf 1}_{\gamma^{*}_{r}}(P)\sim R^{\beta}(rd)^{-1/2}.

    See Figure 5. By translation and rotation, we may assume SrS_{r} is centered at (r,0)(-r,0) and PP lies on the x2x_{2}-axis. By Pythagorean theorem, the coordinate of PP is (0,d(d+2r))(0,\sqrt{d(d+2r)}). Since drd\leq r, we may ignore some constant factor and write the coordinate of PP as

    (28) P=(0,(dr)1/2).P=(0,(dr)^{1/2}).

    The next step is to find the number of γr\gamma^{*}_{r} that pass through PP. Suppose PγrP\in\gamma^{*}_{r}. Since the center of γr\gamma^{*}_{r} lies in SrS_{r}, we may denote its coordinate by C(γr)=(r+rcosθ,rsinθ)C(\gamma^{*}_{r})=(-r+r\cos\theta,r\sin\theta). Let \ell be the line passing through C(γr)C(\gamma^{*}_{r}) and tangent to SrS_{r} (which is also the core line of γr\gamma^{*}_{r}):

    :yrsinθ=cosθsinθ(x+rrcosθ).\ell:y-r\sin\theta=-\frac{\cos\theta}{\sin\theta}(x+r-r\cos\theta).

    Since (dr)1/2Rβ(dr)^{1/2}\leq R^{\beta}, we see that PγrP\in\gamma^{*}_{r} is equivalent to dist(,P)12\textup{dist}(\ell,P)\leq\frac{1}{2}. By some calculation,

    dist(,P)\displaystyle\textup{dist}(\ell,P) =|(dr)1/2rsinθ+cosθsinθr(1cosθ)|1+cos2θsin2θ\displaystyle=\frac{|(dr)^{1/2}-r\sin\theta+\frac{\cos\theta}{\sin\theta}r(1-\cos\theta)|}{\sqrt{1+\frac{\cos^{2}\theta}{\sin^{2}\theta}}}
    =|sinθ(dr)1/2r(1cosθ)|\displaystyle=|\sin\theta(dr)^{1/2}-r(1-\cos\theta)|
    =2|(dr)1/2sinθ2cosθ2rsin2θ2|.\displaystyle=2|(dr)^{1/2}\sin\frac{\theta}{2}\cos\frac{\theta}{2}-r\sin^{2}\frac{\theta}{2}|.

    We just need to find the number of θ\theta such that dist(,P)1/2\textup{dist}(\ell,P)\leq 1/2. By symmetry, we just compute the positive solutions θ\theta that are close to 0. In this case, the inequality becomes

    (dr)1/2sinθ2cosθ2rsin2θ21/4.(dr)^{1/2}\sin\frac{\theta}{2}\cos\frac{\theta}{2}-r\sin^{2}\frac{\theta}{2}\leq 1/4.

    The meaningful solutions will be

    sinθ2\displaystyle\sin\frac{\theta}{2} (dr)1/2cosθ2drcos2θ2r2r\displaystyle\leq\frac{(dr)^{1/2}\cos\frac{\theta}{2}-\sqrt{dr\cos^{2}\frac{\theta}{2}-r}}{2r}
    =121(dr)1/2cosθ2+drcos2θ2r\displaystyle=\frac{1}{2}\frac{1}{(dr)^{1/2}\cos\frac{\theta}{2}+\sqrt{dr\cos^{2}\frac{\theta}{2}-r}}
    (dr)1/2.\displaystyle\sim(dr)^{-1/2}.

    In the last step, we use cosθ21\cos\frac{\theta}{2}\sim 1. Therefore, 0θ(dr)1/20\leq\theta\lesssim(dr)^{-1/2}. Since {C(γr)}\{C(\gamma^{*}_{r})\} have angle separation Rβ\sim R^{-\beta}, we see the number of γr\gamma^{*}_{r} that contains PP is Rβ(dr)1/2\sim R^{\beta}(dr)^{-1/2}.

  3. (3)

    rdRβr\leq d\leq R^{\beta}. We claim in this case

    γ𝟏γr(P)Rβd1.\sum_{\gamma}{\bf 1}_{\gamma^{*}_{r}}(P)\sim R^{\beta}d^{-1}.

    The calculation is exactly the same as above, with the only modification that we replace (28) by P=(0,d)P=(0,d).

Combining the three scenarios (1), (2), (3), we can estimate

2(γ𝟏γr)p/2\displaystyle\int_{\mathbb{R}^{2}}(\sum_{\gamma}{\bf 1}_{\gamma^{*}_{r}})^{p/2} =(d(P)10+10d(P)r+rd(P)Rβ)(γ𝟏γr(P))p/2dP\displaystyle=\bigg{(}\int_{d(P)\leq 10}+\int_{10\leq d(P)\leq r}+\int_{r\leq d(P)\leq R^{\beta}}\bigg{)}(\sum_{\gamma}{\bf 1}_{\gamma^{*}_{r}}(P))^{p/2}dP
r(r1/2Rβ)p/2+d[10,r] dyadicdr(Rβ(rd)1/2)p/2+d[r,Rβ] dyadicd2(Rβd1)p/2\displaystyle\sim r(r^{-1/2}R^{\beta})^{p/2}+\sum_{d\in[10,r]\textup{~{}dyadic}}dr(R^{\beta}(rd)^{-1/2})^{p/2}+\sum_{d\in[r,R^{\beta}]\textup{~{}dyadic}}d^{2}(R^{\beta}d^{-1})^{p/2}
r1p4Rβp2+r2p2Rβp2+R2β.\displaystyle\approx r^{1-\frac{p}{4}}R^{\beta\frac{p}{2}}+r^{2-\frac{p}{2}}R^{\beta\frac{p}{2}}+R^{2\beta}.

In the last line, we use \approx is because when p=4p=4, the summation is over logR\sim\log R same numbers instead of a geometric series.

Case 3: RβrRR^{\beta}\leq r\leq R

This is almost the same as Case 2. Actually, it is even simpler, since we only have scenarios (1) and (2) (with the range in (2) replaced by 10dr1R2β10\leq d\leq r^{-1}R^{2\beta} and noting r1R2βrr^{-1}R^{2\beta}\leq r). The same argument will give

2(γ𝟏γr)p/2\displaystyle\int_{\mathbb{R}^{2}}(\sum_{\gamma}{\bf 1}_{\gamma^{*}_{r}})^{p/2} =(d(P)10+10d(P)r1R2β)(γ𝟏γr(P))p/2dP\displaystyle=\bigg{(}\int_{d(P)\leq 10}+\int_{10\leq d(P)\leq r^{-1}R^{2\beta}}\bigg{)}(\sum_{\gamma}{\bf 1}_{\gamma^{*}_{r}}(P))^{p/2}dP
r(r1/2Rβ)p/2+d[10,r1R2β] dyadicdr(Rβ(rd)1/2)p/2\displaystyle\sim r(r^{-1/2}R^{\beta})^{p/2}+\sum_{d\in[10,r^{-1}R^{2\beta}]\textup{~{}dyadic}}dr(R^{\beta}(rd)^{-1/2})^{p/2}
r1p4Rβp2+R2β.\displaystyle\approx r^{1-\frac{p}{4}}R^{\beta\frac{p}{2}}+R^{2\beta}.

With (27), we can finally estimate

(γ𝟏γ)p/2\displaystyle\int(\sum_{\gamma}{\bf 1}_{\gamma^{*}})^{p/2} =|r|R𝑑r{x3=r}(γ𝟏γ)p/2𝑑x1𝑑x2\displaystyle=\int_{|r|\leq R}dr\int_{\{x_{3}=r\}}(\sum_{\gamma}{\bf 1}_{\gamma^{*}})^{p/2}dx_{1}dx_{2}
(0|r|10+10|r|Rβ+Rβ|r|R)(γ𝟏γ)p/2dx1dx2dr\displaystyle\bigg{(}\int_{0\leq|r|\leq 10}+\int_{10\leq|r|\leq R^{\beta}}+\int_{R^{\beta}\leq|r|\leq R}\bigg{)}(\sum_{\gamma}{\bf 1}_{\gamma^{*}})^{p/2}dx_{1}dx_{2}dr
R2β+Rpβ2+r[10,Rβ] dyadicr(r1p4Rβp2+r2p2Rβp2+R2β)\displaystyle\lesssim R^{2\beta}+R^{\frac{p\beta}{2}}+\sum_{r\in[10,R^{\beta}]\textup{~{}dyadic}}r(r^{1-\frac{p}{4}}R^{\beta\frac{p}{2}}+r^{2-\frac{p}{2}}R^{\beta\frac{p}{2}}+R^{2\beta})
+r[Rβ,R] dyadicr(r1p4Rβp2+R2β)\displaystyle\ \ \ \ +\sum_{r\in[R^{\beta},R]\textup{~{}dyadic}}r(r^{1-\frac{p}{4}}R^{\beta\frac{p}{2}}+R^{2\beta})
Rpβ2+Rβ(2+p4)+R(2p4)+pβ2+R1+2β\displaystyle\lesssim R^{\frac{p\beta}{2}}+R^{\beta(2+\frac{p}{4})}+R^{(2-\frac{p}{4})+\frac{p\beta}{2}}+R^{1+2\beta}
Rpβ2+R(2p4)+pβ2+R1+2β.\displaystyle\sim R^{\frac{p\beta}{2}}+R^{(2-\frac{p}{4})+\frac{p\beta}{2}}+R^{1+2\beta}.

The last step is because of Rβ(2+p4)Rpβ2+R(2p4)+pβ2R^{\beta(2+\frac{p}{4})}\leq R^{\frac{p\beta}{2}}+R^{(2-\frac{p}{4})+\frac{p\beta}{2}}.

Combining (25), (26) and plugging into (3), we obtain

max{R1,R32+2p}Cβ,p(R)R1β(Rβ2+R2p14+β2+R1+2βp).\max\{R^{-1},R^{-\frac{3}{2}+\frac{2}{p}}\}\lessapprox C_{\beta,p}(R)R^{-1-\beta}(R^{\frac{\beta}{2}}+R^{\frac{2}{p}-\frac{1}{4}+\frac{\beta}{2}}+R^{\frac{1+2\beta}{p}}).

Considering of the three cases 2p42\leq p\leq 4, 4p84\leq p\leq 8 and p8p\geq 8 will give us that the right hand side of (4) is actually the lower bound of Cβ,p(R)C_{\beta,p}(R) (up to RϵR^{\epsilon} factor).

3.2. Proof of Theorem 2

The difficult part of the proof will be in the range 4p84\leq p\leq 8. Recall from Remark Remark that we need to prove for all pp but not only the endpoint pp, since there is no interpolation argument. The main tool we are going to use is called the amplitude dependent wave envelope estimate by Guth-Maldague [5]. Before giving the proof, we introduce some notations from [6], [5].

Recall 𝒞\mathcal{C} is the truncated cone in 3\mathbb{R}^{3}:

𝒞:={ξ3:ξ3=x12+x22,1/2ξ31}.\mathcal{C}:=\{\xi\in\mathbb{R}^{3}:\xi_{3}=\sqrt{x_{1}^{2}+x_{2}^{2}},1/2\leq\xi_{3}\leq 1\}.

We have the canonical partition of NR1(𝒞)N_{R^{-1}}(\mathcal{C}) into 1×R1/2×R11\times R^{-1/2}\times R^{-1}-planks Θ={θ}\Theta=\{\theta\}:

NR1(𝒞)=θ.N_{R^{-1}}(\mathcal{C})=\bigsqcup\theta.

More generally, for any dyadic s[R1/2,1]s\in[R^{-1/2},1], we can partition the s2s^{2}-neighborhood of 𝒞\mathcal{C} into 1×s×s21\times s\times s^{2}-planks Ss={τs}\textbf{S}_{s}=\{\tau_{s}\}:

Ns2(𝒞)=τs.N_{s^{2}}(\mathcal{C})=\bigsqcup\tau_{s}.

Note in particular SR1/2=Θ\textbf{S}_{R^{-1/2}}=\Theta. For each ss and a frequency plank τsSs\tau_{s}\in\textbf{S}_{s}, we define the box UτsU_{\tau_{s}} in the physical space to be a rectangle centered at the origin of dimensions Rs2×Rs×RRs^{2}\times Rs\times R whose edge of length Rs2Rs^{2} (respectively RsRs, RR) is parallel to the edge of τs\tau_{s} with length 11 (respectively ss, s2s^{2}). Note that for any θΘ\theta\in\Theta, UθU_{\theta} is just θ\theta^{*} (the dual rectangle of θ\theta). Also, UτsU_{\tau_{s}} is the convex hull of θτsUθ\cup_{\theta\subset\tau_{s}}U_{\theta}.

We make a useful observation, which will be used later. For any θτs\theta\subset\tau_{s}, we see that θ\theta^{*} is a 1×R1/2×R1\times R^{1/2}\times R-plank. Define Uθ,sU_{\theta,s} to be the Rs2×Rs×RRs^{2}\times Rs\times R-plank which is made by dilating the corresponding edges of θ\theta^{*}. Our observation is that UτsU_{\tau_{s}} and Uθ,sU_{\theta,s} are comparable:

(29) 1CUθ,sUτsCUθ,s.\frac{1}{C}U_{\theta,s}\subset U_{\tau_{s}}\subset CU_{\theta,s}.

This is not hard to see by noting that the second longest edge of θ\theta^{*} form an angle s\lesssim s with the Rs×RRs\times R-face of UτsU_{\tau_{s}}. We just omit the proof.

We cover 3\mathbb{R}^{3} by translated copies of UτsU_{\tau_{s}}. We will use UUτsU\parallel U_{\tau_{s}} to indicate UU is one of the translated copies. If UUτsU\parallel U_{\tau_{s}}, then we define SUfS_{U}f by

(30) SUf=(θτs|fθ|2)1/2𝟏U.S_{U}f=\big{(}\sum_{\theta\subset\tau_{s}}|f_{\theta}|^{2}\big{)}^{1/2}{\bf 1}_{U}.

We can think of SUfS_{U}f as the wave envelope of ff localized in UU in the physical space and localized in τs\tau_{s} in the frequency space. We have the following inequality of Guth, Wang and Zhang (see [6, Theorem 1.5] ):

Theorem 3 (Wave envelope estimate).

Suppose suppf^NR1(𝒞)\textup{supp}\widehat{f}\subset N_{R^{-1}}(\mathcal{C}). Then

(31) f44CϵRϵR1/2s1τsSsUUτs|U|1SUf24,\|f\|_{4}^{4}\leq C_{\epsilon}R^{\epsilon}\sum_{R^{-1/2}\leq s\leq 1}\sum_{\tau_{s}\in\textbf{S}_{s}}\sum_{U\parallel U_{\tau_{s}}}|U|^{-1}\|S_{U}f\|_{2}^{4},

for any ϵ>0\epsilon>0.

There is a refined version of the wave envelope estimate proved by Guth and Maldague (See [5, Theorem 2] ):

Theorem 4 (Amplitude dependent wave envelope estimate).

Suppose suppf^NR1(𝒞)\textup{supp}\widehat{f}\subset N_{R^{-1}}(\mathcal{C}). Then for any α>0\alpha>0,

(32) α4|{x3:|f(x)|>α}|CϵRϵR1/2s1τsSsU𝒢τs(α)|U|1SUf24,\alpha^{4}|\{x\in\mathbb{R}^{3}:|f(x)|>\alpha\}|\leq C_{\epsilon}R^{\epsilon}\sum_{R^{-1/2}\leq s\leq 1}\sum_{\tau_{s}\in\textbf{S}_{s}}\sum_{U\in\mathcal{G}_{\tau_{s}}(\alpha)}|U|^{-1}\|S_{U}f\|_{2}^{4},

for any ϵ>0\epsilon>0. Here, 𝒢τs(α)={UUτs:|U|1SUf22|logR|1α2(#Ss)2}\mathcal{G}_{\tau_{s}}(\alpha)=\{U\parallel U_{\tau_{s}}:|U|^{-1}\|S_{U}f\|_{2}^{2}\gtrsim|\log R|^{-1}\frac{\alpha^{2}}{(\#\textbf{S}_{s})^{2}}\}.

Remark.

In the original paper [5], their definition for 𝒢τs(α)\mathcal{G}_{\tau_{s}}(\alpha) is

𝒢τs(α)={UUτs:|U|1SUf22α2(#τs)2},\mathcal{G}_{\tau_{s}}(\alpha)=\{U\parallel U_{\tau_{s}}:|U|^{-1}\|S_{U}f\|_{2}^{2}\gtrapprox\frac{\alpha^{2}}{(\#\tau_{s})^{2}}\},

where #τs=#{τsSs:fτs0}\#\tau_{s}=\#\{\tau_{s}\in\textbf{S}_{s}:f_{\tau_{s}}\not\equiv 0\}. Noting that #τs#Ss\#\tau_{s}\leq\#\textbf{S}_{s}, we see our 𝒢τs(α)\mathcal{G}_{\tau_{s}}(\alpha) is a bigger set, and hence our (32) is weaker than the original version ([5] Theorem 2).

Proof of Theorem 2.

Case 1: p8p\geq 8

This is just by Cauchy-Schwarz inequality, since #Γβ(R1)Rβ.\#\Gamma_{\beta}(R^{-1})\sim R^{\beta}.

Case 2: 2p42\leq p\leq 4

We have (31). By dyadic pigeonholing on ss, we can find ss such that

(33) f44τSsUUτ|U|1SUf24.\|f\|_{4}^{4}\lessapprox\sum_{\tau\in\textbf{S}_{s}}\sum_{U\parallel U_{\tau}}|U|^{-1}\|S_{U}f\|_{2}^{4}.

We fix this ss. Denote U:={U:UUτ for some τSs}\textbf{U}:=\{U:U\parallel U_{\tau}\textup{~{}for~{}some~{}}\tau\in\textbf{S}_{s}\}. Then the inequality above can be written as

(34) f44UU|U|1SUf24.\|f\|_{4}^{4}\lessapprox\sum_{U\in\textbf{U}}|U|^{-1}\|S_{U}f\|_{2}^{4}.

We remind readers that each UUU\in\textbf{U} has size Rs2×Rs×RRs^{2}\times Rs\times R. We also have the following L2L^{2} estimate:

(35) f22UUSUf22.\|f\|_{2}^{2}\sim\sum_{U\in\textbf{U}}\|S_{U}f\|^{2}_{2}.

We provide a quick proof for (35). We have

f22=τSsfτ22τSsUUτfτL2(U)2.\|f\|_{2}^{2}=\sum_{\tau\in\textbf{S}_{s}}\|f_{\tau}\|_{2}^{2}\sim\sum_{\tau\in\textbf{S}_{s}}\sum_{U\parallel U_{\tau}}\|f_{\tau}\|_{L^{2}(U)}^{2}.

Noting that {fθ:θτ}\{f_{\theta}:\theta\subset\tau\} are locally orthogonal on any translation of UτU_{\tau} and recalling (30), we have

f22τSsUUτUθτ|fθ|2=UUSUf22.\|f\|_{2}^{2}\sim\sum_{\tau\in\textbf{S}_{s}}\sum_{U\parallel U_{\tau}}\int_{U}\sum_{\theta\subset\tau}|f_{\theta}|^{2}=\sum_{U\in\textbf{U}}\|S_{U}f\|_{2}^{2}.

Next, we will do dyadic pigeonholing on SUf22\|S_{U}f\|_{2}^{2}. (Actually, we only need to prove a local version of the inequality, so we just care about those UU that intersect BRB_{R}. There are in total RO(1)R^{O(1)} of them.) We can find a number W>0W>0 and set U={UU:SUf22W}\textbf{U}^{\prime}=\{U\in\textbf{U}:\|S_{U}f\|_{2}^{2}\sim W\}, so that

(36) f44\displaystyle\|f\|_{4}^{4} |U|1#UW2,\displaystyle\lessapprox|U|^{-1}\#\textbf{U}^{\prime}W^{2},
(37) f22\displaystyle\|f\|_{2}^{2} #UW.\displaystyle\approx\#\textbf{U}^{\prime}W.

Since every UUU\in\textbf{U} has the same measure R3s2R^{3}s^{2}, there is no ambiguity to write |U|1|U|^{-1} in (36).

Let α\alpha be such that 1p=α4+1α2\frac{1}{p}=\frac{\alpha}{4}+\frac{1-\alpha}{2}. Then α=4(121p)\alpha=4(\frac{1}{2}-\frac{1}{p}). Applying Hölder’s inequality gives

fppf4αpf2(1α)p|U|p(121p)#UWp2|U|p(121p)UUSUf2p.\|f\|_{p}^{p}\leq\|f\|_{4}^{\alpha p}\|f\|_{2}^{(1-\alpha)p}\lessapprox|U|^{-p(\frac{1}{2}-\frac{1}{p})}\#\textbf{U}^{\prime}W^{\frac{p}{2}}\leq|U|^{-p(\frac{1}{2}-\frac{1}{p})}\sum_{U\in\textbf{U}}\|S_{U}f\|_{2}^{p}.

Next we are going to exploit more orthogonality for SUfS_{U}f. Suppose UUτU\parallel U_{\tau}. By definition

SUf22=Uθτ|fθ|2=Uθτ|γθfγ|2.\|S_{U}f\|_{2}^{2}=\int_{U}\sum_{\theta\subset\tau}|f_{\theta}|^{2}=\int_{U}\sum_{\theta\subset\tau}|\sum_{\gamma\subset\theta}f_{\gamma}|^{2}.

We remind readers that {τ}\{\tau\} are 1×s×s21\times s\times s^{2}-caps; {θ}\{\theta\} are 1×R1/2×R11\times R^{-1/2}\times R^{-1}-caps; {γ}\{\gamma\} are 1×Rβ×R11\times R^{-\beta}\times R^{-1}-caps. Since UU is too small for {fγ:γθ}\{f_{\gamma}:\gamma\subset\theta\} to be orthogonal on UU, we need to find a larger rectangle. First, let us look at the rectangles {γ:γθ}\{\gamma:\gamma\subset\theta\}. We want to find a rectangle νθ\nu_{\theta} as big as possible, such that {γ+ν:γθ}\{\gamma+\nu:\gamma\subset\theta\} are finitely overlapping. Actually, we can choose νθ\nu_{\theta} to be of size R1/2β×Rβ×R1R^{1/2-\beta}\times R^{-\beta}\times R^{-1} (here the edge of νθ\nu_{\theta} with length R1/2βR^{1/2-\beta} (respectively RβR^{-\beta}, R1R^{-1}) are parallel to the edge of θ\theta with length 11 (respectively R1/2R^{-1/2}, R1R^{-1}). See Figure 6: the left hand side is θ\theta and {γ:γθ}\{\gamma:\gamma\subset\theta\}; the right hand side is our νθ\nu_{\theta}. It is not hard to see {γ+νθ:γθ}\{\gamma+\nu_{\theta}:\gamma\subset\theta\} are finitely overlapping. Let νθ\nu_{\theta}^{*} be the dual of νθ\nu_{\theta} in the physical space, then νθ\nu_{\theta}^{*} has size Rβ12×Rβ×RR^{\beta-\frac{1}{2}}\times R^{\beta}\times R and we have the local orthogonality (we just ignore the rapidly decaying tail for simplicity):

νθ|γθfγ|2νθγθ|fγ|2\int_{\nu_{\theta}^{*}}|\sum_{\gamma\subset\theta}f_{\gamma}|^{2}\sim\int_{\nu_{\theta}^{*}}\sum_{\gamma\subset\theta}|f_{\gamma}|^{2}
θ\thetaγ\gammaR1/2R^{-1/2}RβR^{-\beta}νθ\nu_{\theta}R1/2βR^{1/2-\beta}RβR^{-\beta}
Figure 6.

Define

(38) Vθ=Uτ+νθ,V_{\theta}=U_{\tau}+\nu_{\theta}^{*},

which is a rectangle of size

max{Rs2,Rβ12}×max{Rs,Rβ}×R.\max\{Rs^{2},R^{\beta-\frac{1}{2}}\}\times\max\{Rs,R^{\beta}\}\times R.

We tile 3\mathbb{R}^{3} with translated copies of VθV_{\theta}, and we write VVθV\parallel V_{\theta} if VV is one of the tiles. Noting that Rβ12Rs2RβRs\frac{R^{\beta-\frac{1}{2}}}{Rs^{2}}\leq\frac{R^{\beta}}{Rs}, we will discuss three scenarios: 1. Rβ12Rs2RβRs1\frac{R^{\beta-\frac{1}{2}}}{Rs^{2}}\leq\frac{R^{\beta}}{Rs}\leq 1; 2. Rβ12Rs21RβRs\frac{R^{\beta-\frac{1}{2}}}{Rs^{2}}\leq 1\leq\frac{R^{\beta}}{Rs}; 3. 1Rβ12Rs2RβRs1\leq\frac{R^{\beta-\frac{1}{2}}}{Rs^{2}}\leq\frac{R^{\beta}}{Rs}.

\bullet If Rβ12Rs2RβRs1\frac{R^{\beta-\frac{1}{2}}}{Rs^{2}}\leq\frac{R^{\beta}}{Rs}\leq 1, then VθV_{\theta} is essentially UτU_{\tau}. In this case, we already have the orthogonality of {fγ:γθ}\{f_{\gamma}:\gamma\subset\theta\} on U(Uτ)U(\parallel U_{\tau}). Therefore,

fpp\displaystyle\|f\|_{p}^{p} |U|p(121p)τSsUUτ(Uθτ|γθfγ|2)p2\displaystyle\lessapprox|U|^{-p(\frac{1}{2}-\frac{1}{p})}\sum_{\tau\in\textbf{S}_{s}}\sum_{U\parallel U_{\tau}}\bigg{(}\int_{U}\sum_{\theta\subset\tau}|\sum_{\gamma\subset\theta}f_{\gamma}|^{2}\bigg{)}^{\frac{p}{2}}
|U|p(121p)τSsUUτ(Uγτ|fγ|2)p2\displaystyle\sim|U|^{-p(\frac{1}{2}-\frac{1}{p})}\sum_{\tau\in\textbf{S}_{s}}\sum_{U\parallel U_{\tau}}\bigg{(}\int_{U}\sum_{\gamma\subset\tau}|f_{\gamma}|^{2}\bigg{)}^{\frac{p}{2}}
τSsUUτU(γτ|fγ|2)p2\displaystyle\leq\sum_{\tau\in\textbf{S}_{s}}\sum_{U\parallel U_{\tau}}\int_{U}(\sum_{\gamma\subset\tau}|f_{\gamma}|^{2})^{\frac{p}{2}}
=τSs3(γτ|fγ|2)p2\displaystyle=\sum_{\tau\in\textbf{S}_{s}}\int_{\mathbb{R}^{3}}(\sum_{\gamma\subset\tau}|f_{\gamma}|^{2})^{\frac{p}{2}}
3(γΓβ(R1)|fγ|2)p/2.\displaystyle\leq\int_{\mathbb{R}^{3}}(\sum_{\gamma\in\Gamma_{\beta}(R^{-1})}|f_{\gamma}|^{2})^{p/2}.

\bullet In the other two scenarios, we proceed as follows.

fpp\displaystyle\|f\|_{p}^{p} |U|p(121p)τSsUUτ(Uθτ|fθ|2)p/2\displaystyle\lessapprox|U|^{-p(\frac{1}{2}-\frac{1}{p})}\sum_{\tau\in\textbf{S}_{s}}\sum_{U\parallel U_{\tau}}\bigg{(}\int_{U}\sum_{\theta\subset\tau}|f_{\theta}|^{2}\bigg{)}^{p/2}
|U|p(121p)τSsUUτ#{θτ}p21θτ(U|fθ|2)p/2\displaystyle\leq|U|^{-p(\frac{1}{2}-\frac{1}{p})}\sum_{\tau\in\textbf{S}_{s}}\sum_{U\parallel U_{\tau}}\#\{\theta\subset\tau\}^{\frac{p}{2}-1}\sum_{\theta\subset\tau}\bigg{(}\int_{U}|f_{\theta}|^{2}\bigg{)}^{p/2}
|U|p(121p)#{θτ}p21τSsθτVVθ(V|fθ|2)p/2\displaystyle\leq|U|^{-p(\frac{1}{2}-\frac{1}{p})}\#\{\theta\subset\tau\}^{\frac{p}{2}-1}\sum_{\tau\in\textbf{S}_{s}}\sum_{\theta\subset\tau}\sum_{V\parallel V_{\theta}}\bigg{(}\int_{V}|f_{\theta}|^{2}\bigg{)}^{p/2}
(By orthogonality)\displaystyle(\textup{By~{}orthogonality}) |U|p(121p)#{θτ}p21τSsθτVVθ(Vγθ|fγ|2)p/2\displaystyle\sim|U|^{-p(\frac{1}{2}-\frac{1}{p})}\#\{\theta\subset\tau\}^{\frac{p}{2}-1}\sum_{\tau\in\textbf{S}_{s}}\sum_{\theta\subset\tau}\sum_{V\parallel V_{\theta}}\bigg{(}\int_{V}\sum_{\gamma\subset\theta}|f_{\gamma}|^{2}\bigg{)}^{p/2}
(Hölder)\displaystyle(\textup{H\"{o}lder}) |U|p(121p)#{θτ}p21τSsθτVVθ|V|p(121p)V(γθ|fγ|2)p/2\displaystyle\leq|U|^{-p(\frac{1}{2}-\frac{1}{p})}\#\{\theta\subset\tau\}^{\frac{p}{2}-1}\sum_{\tau\in\textbf{S}_{s}}\sum_{\theta\subset\tau}\sum_{V\parallel V_{\theta}}|V|^{p(\frac{1}{2}-\frac{1}{p})}\int_{V}(\sum_{\gamma\subset\theta}|f_{\gamma}|^{2})^{p/2}
(|V||U|)p(121p)#{θτ}p21(γΓβ(R1)|fγ|2)12pp\displaystyle\leq\bigg{(}\frac{|V|}{|U|}\bigg{)}^{p(\frac{1}{2}-\frac{1}{p})}\#\{\theta\subset\tau\}^{\frac{p}{2}-1}\|(\sum_{\gamma\in\Gamma_{\beta}(R^{-1})}|f_{\gamma}|^{2})^{\frac{1}{2}}\|_{p}^{p}
=(max{RβRs,1}max{Rβ12Rs2,1})p(121p)(sR12)p21(γΓβ(R1)|fγ|2)12pp\displaystyle=\bigg{(}\max\{\frac{R^{\beta}}{Rs},1\}\max\{\frac{R^{\beta-\frac{1}{2}}}{Rs^{2}},1\}\bigg{)}^{p(\frac{1}{2}-\frac{1}{p})}(sR^{\frac{1}{2}})^{\frac{p}{2}-1}\|(\sum_{\gamma\in\Gamma_{\beta}(R^{-1})}|f_{\gamma}|^{2})^{\frac{1}{2}}\|_{p}^{p}

We just need to check

(39) (max{RβRs,1}max{Rβ12Rs2,1})p(121p)(sR12)p21R(β12)(p2).\bigg{(}\max\{\frac{R^{\beta}}{Rs},1\}\max\{\frac{R^{\beta-\frac{1}{2}}}{Rs^{2}},1\}\bigg{)}^{p(\frac{1}{2}-\frac{1}{p})}(sR^{\frac{1}{2}})^{\frac{p}{2}-1}\lesssim R^{(\beta-\frac{1}{2})(p-2)}.

* If Rβ12Rs21RβRs\frac{R^{\beta-\frac{1}{2}}}{Rs^{2}}\leq 1\leq\frac{R^{\beta}}{Rs}, then the left hand side of (39) equals R(β12)(p21)R^{(\beta-\frac{1}{2})(\frac{p}{2}-1)}, which is \leq the right hand side of (39).

* If 1Rβ12Rs2RβRs1\leq\frac{R^{\beta-\frac{1}{2}}}{Rs^{2}}\leq\frac{R^{\beta}}{Rs}, then the left hand side of (39) equals

(R2β2s2)p21,(R^{2\beta-2}s^{-2})^{\frac{p}{2}-1},

which is less than the right hand side of (39) since s1R1/2s^{-1}\leq R^{1/2}.

Case 3: 4p84\leq p\leq 8

Note that

fppα dyadicαp|{x3:|f(x)|α}|.\|f\|_{p}^{p}\sim\sum_{\alpha\text{~{}dyadic}}\alpha^{p}|\{x\in\mathbb{R}^{3}:|f(x)|\sim\alpha\}|.

We can assume the range of α\alpha is R100fαfR^{-100}\|f\|_{\infty}\leq\alpha\leq\|f\|_{\infty}. Other α\alpha are considered as negligible.

By dyadic pigeonholing, we can find α>0\alpha>0 such that

fpp(logR)αp|{x3:|f(x)|α}|+negligible term.\|f\|_{p}^{p}\lesssim(\log R)\cdot\alpha^{p}|\{x\in\mathbb{R}^{3}:|f(x)|\sim\alpha\}|+\textup{negligible~{}term}.

We just need to fix this α\alpha, and prove an upper bound for αp|{x3:|f(x)|>α}|\alpha^{p}|\{x\in\mathbb{R}^{3}:|f(x)|>\alpha\}|. By (32), we have

α4|{x3:|f(x)|>α}|CϵRϵR1/2s1τsSsU𝒢τs(α)|U|1SUf24.\alpha^{4}|\{x\in\mathbb{R}^{3}:|f(x)|>\alpha\}|\leq C_{\epsilon}R^{\epsilon}\sum_{R^{-1/2}\leq s\leq 1}\sum_{\tau_{s}\in\textbf{S}_{s}}\sum_{U\in\mathcal{G}_{\tau_{s}}(\alpha)}|U|^{-1}\|S_{U}f\|_{2}^{4}.

By pigeonholing again, we can find ss such that

(40) α4|{x3:|f(x)|>α}|τSsU𝒢τ(α)|U|1SUf24.\alpha^{4}|\{x\in\mathbb{R}^{3}:|f(x)|>\alpha\}|\lessapprox\sum_{\tau\in\textbf{S}_{s}}\sum_{U\in\mathcal{G}_{\tau}(\alpha)}|U|^{-1}\|S_{U}f\|_{2}^{4}.

We fix this ss. We also remind readers the definition of 𝒢τ(α)\mathcal{G}_{\tau}(\alpha):

𝒢τ(α):={UUτ:|U|1Uθτ|fθ|2(αs)2},\mathcal{G}_{\tau}(\alpha):=\{U\parallel U_{\tau}:|U|^{-1}\int_{U}\sum_{\theta\subset\tau}|f_{\theta}|^{2}\gtrapprox(\alpha s)^{2}\},

since #Sss1\#\textbf{S}_{s}\sim s^{-1}. Continuing the estimate in (40), we have

α4|{x3:|f(x)|>α}|\displaystyle\alpha^{4}|\{x\in\mathbb{R}^{3}:|f(x)|>\alpha\}| τSsU𝒢τ(α)|U|1(Uθτ|fθ|2)2\displaystyle\lessapprox\sum_{\tau\in\textbf{S}_{s}}\sum_{U\in\mathcal{G}_{\tau}(\alpha)}|U|^{-1}\bigg{(}\int_{U}\sum_{\theta\subset\tau}|f_{\theta}|^{2}\bigg{)}^{2}
τSsU𝒢τ(α)|U|1(Uθτ|fθ|2)p2(|U|(αs)2)2p2.\displaystyle\lessapprox\sum_{\tau\in\textbf{S}_{s}}\sum_{U\in\mathcal{G}_{\tau}(\alpha)}|U|^{-1}\bigg{(}\int_{U}\sum_{\theta\subset\tau}|f_{\theta}|^{2}\bigg{)}^{\frac{p}{2}}\bigg{(}|U|(\alpha s)^{2}\bigg{)}^{2-\frac{p}{2}}.

Moving the power of α\alpha to the left hand side, we obtain

(41) αp|{x3:|f(x)|>α}|τSsU𝒢τ(α)|U|1p2(Uθτ|fθ|2)p2s4p.\alpha^{p}|\{x\in\mathbb{R}^{3}:|f(x)|>\alpha\}|\lessapprox\sum_{\tau\in\textbf{S}_{s}}\sum_{U\in\mathcal{G}_{\tau}(\alpha)}|U|^{1-\frac{p}{2}}\bigg{(}\int_{U}\sum_{\theta\subset\tau}|f_{\theta}|^{2}\bigg{)}^{\frac{p}{2}}s^{4-p}.

Our final goal is to prove that the right hand side above is

(42) Rβp2+p42(γ|fγ|2)1/2pp.\lessapprox R^{\frac{\beta p}{2}+\frac{p}{4}-2}\|(\sum_{\gamma}|f_{\gamma}|^{2})^{1/2}\|_{p}^{p}.

To do that, we again need to exploit the orthogonality of {fγ:γθ}\{f_{\gamma}:\gamma\subset\theta\}. The argument is different from that in Case 2: 2p42\leq p\leq 4. In Case 2: 2p42\leq p\leq 4, we expand the integration domain UU to a bigger rectangle VV to get orthogonality, whereas here we are going to use Cauchy-Schwarz inequality.

We discuss the geometry of these caps. Fix a τSs\tau\in\textbf{S}_{s}. By definition, UτU_{\tau} is a Rs2×Rs×RRs^{2}\times Rs\times R-rectangle in the physical space. Then UτU_{\tau}^{*} is a R1s2×R1s1×R1R^{-1}s^{-2}\times R^{-1}s^{-1}\times R^{-1}-rectangle. We make the following observation: for each θτ\theta\subset\tau, we can show that UτU^{*}_{\tau} is comparable to another rectangle, which has the same size but with the edges parallel to the corresponding edges of θ\theta. We explain it with more details. Let Uθ,sU_{\theta,s} be the Rs2×Rs×RRs^{2}\times Rs\times R-rectangle which is made from the 1×R1/2×R1\times R^{1/2}\times R-rectangle θ\theta^{*} by dilating the corresponding edges. Then Uθ,sU_{\theta,s}^{*} is a R1s2×R1s1×R1R^{-1}s^{-2}\times R^{-1}s^{-1}\times R^{-1}-rectangle whose edges are parallel to the corresponding edges of the 1×R1/2×R11\times R^{-1/2}\times R^{-1}-rectangle θ\theta. We want to show UτU_{\tau}^{*} and Uθ,sU_{\theta,s}^{*} are comparable. This is equivalent to show UτU_{\tau} and Uθ,sU_{\theta,s} are comparable, which is an observation we made at (29). Therefore, for any θτ\theta\subset\tau, we can assume the edges of UτU_{\tau}^{*} are parallel to the corresponding edges of θ\theta.

Fix a UUτU\parallel U_{\tau}, then U=UτU^{*}=U_{\tau}^{*}. See Figure 7: on the left is θ\theta and {γ:γθ}\{\gamma:\gamma\subset\theta\}; on the middle is our UU^{*}. We will discuss two scenarios depending on whether RβR^{-\beta} (the width of γ\gamma) is bigger than R1s1R^{-1}s^{-1} (the width of UU^{*}).

θ\thetaγ\gammaR1/2R^{-1/2}RβR^{-\beta}R1s2R^{-1}s^{-2}R1s1R^{-1}s^{-1}UU^{\ast}π\pi
Figure 7.

\bullet If RβR1s1R^{-\beta}\geq R^{-1}s^{-1}, then we see that {γ+U:γθ}\{\gamma+U^{*}:\gamma\subset\theta\} are finitely overlapping. This means that {fγ:γθ}\{f_{\gamma}:\gamma\subset\theta\} are locally orthogonal on UU:

U|γθfγ|2Uγθ|fγ|2.\int_{U}|\sum_{\gamma\subset\theta}f_{\gamma}|^{2}\lesssim\int_{U}\sum_{\gamma\subset\theta}|f_{\gamma}|^{2}.

Therefore,

fpp\displaystyle\|f\|_{p}^{p} |U|1p2τSsUUs(Uθτ|γθfγ|2)p2s4p\displaystyle\lessapprox|U|^{1-\frac{p}{2}}\sum_{\tau\in\textbf{S}_{s}}\sum_{U\parallel U_{s}}\bigg{(}\int_{U}\sum_{\theta\subset\tau}|\sum_{\gamma\subset\theta}f_{\gamma}|^{2}\bigg{)}^{\frac{p}{2}}s^{4-p}
|U|1p2τSsUUs(Uγτ|fγ|2)p2s4p\displaystyle\lesssim|U|^{1-\frac{p}{2}}\sum_{\tau\in\textbf{S}_{s}}\sum_{U\parallel U_{s}}\bigg{(}\int_{U}\sum_{\gamma\subset\tau}|f_{\gamma}|^{2}\bigg{)}^{\frac{p}{2}}s^{4-p}
(Hölder)\displaystyle(\textup{H\"{o}lder}) s4pτSsUUsU(γτ|fγ|2)p2\displaystyle\leq s^{4-p}\sum_{\tau\in\textbf{S}_{s}}\sum_{U\parallel U_{s}}\int_{U}(\sum_{\gamma\subset\tau}|f_{\gamma}|^{2})^{\frac{p}{2}}
=s4pτSs3(γτ|fγ|2)p2\displaystyle=s^{4-p}\sum_{\tau\in\textbf{S}_{s}}\int_{\mathbb{R}^{3}}(\sum_{\gamma\subset\tau}|f_{\gamma}|^{2})^{\frac{p}{2}}
s4p3(γΓβ(R1)|fγ|2)p/2.\displaystyle\leq s^{4-p}\int_{\mathbb{R}^{3}}(\sum_{\gamma\in\Gamma_{\beta}(R^{-1})}|f_{\gamma}|^{2})^{p/2}.

We just need to check

s4pRβp2+p42.s^{4-p}\leq R^{\frac{\beta p}{2}+\frac{p}{4}-2}.

Plugging s1R1/2s^{-1}\leq R^{-1/2}, the inequality above is reduced to

Rp/4Rβp2,R^{p/4}\leq R^{\frac{\beta p}{2}},

which is true since β1/2\beta\geq 1/2.

\bullet If RβR1s1R^{-\beta}\leq R^{-1}s^{-1}, we will define a set of new planks which we call π\pi. See on the right hand side of Figure 7. We partition θ\theta into a set of 1×R1s1×R11\times R^{-1}s^{-1}\times R^{-1}-planks, which we denoted by {π:πθ}\{\pi:\pi\subset\theta\}. If the partition is well chosen (the size of caps can vary within a constant multiple), we can assume each γ\gamma fits into one π\pi, so we define

fπ:=γπfγ.f_{\pi}:=\sum_{\gamma\subset\pi}f_{\gamma}.

Now, our key observation is that {π+U:πθ}\{\pi+U^{*}:\pi\subset\theta\} are finitely overlapping. This is true by noting that: the width of UU^{*} and π\pi are both R1s1R^{-1}s^{-1}; the angle between the longest edge of π\pi and UU^{*} is less than R1/2R^{-1/2} and R1s2R1/2R1s1R^{-1}s^{-2}\cdot R^{-1/2}\leq R^{-1}s^{-1}. Therefore, we have that {fπ:πθ}\{f_{\pi}:\pi\subset\theta\} are locally orthogonal on UU, i.e.,

(43) U|πθfπ|2Uπθ|fπ|2.\int_{U}|\sum_{\pi\subset\theta}f_{\pi}|^{2}\lesssim\int_{U}\sum_{\pi\subset\theta}|f_{\pi}|^{2}.

Another step of Cauchy-Schwarz will give

(44) Uπθ|fπ|2=Uπθ|γπfγ|2#{γπ}Uγθ|fγ|2=RβR1s1Uγθ|fγ|2.\int_{U}\sum_{\pi\subset\theta}|f_{\pi}|^{2}=\int_{U}\sum_{\pi\subset\theta}|\sum_{\gamma\subset\pi}f_{\gamma}|^{2}\leq\#\{\gamma\subset\pi\}\int_{U}\sum_{\gamma\subset\theta}|f_{\gamma}|^{2}=R^{\beta}R^{-1}s^{-1}\int_{U}\sum_{\gamma\subset\theta}|f_{\gamma}|^{2}.

As a result, we obtain

U|fθ|2RβR1s1Uγθ|fγ|2.\int_{U}|f_{\theta}|^{2}\lesssim R^{\beta}R^{-1}s^{-1}\int_{U}\sum_{\gamma\subset\theta}|f_{\gamma}|^{2}.

Summing over θτ\theta\subset\tau, we obtain

Uθτ|fθ|2RβR1s1Uγτ|fγ|2.\int_{U}\sum_{\theta\subset\tau}|f_{\theta}|^{2}\lesssim R^{\beta}R^{-1}s^{-1}\int_{U}\sum_{\gamma\subset\tau}|f_{\gamma}|^{2}.

Therefore,

fpp\displaystyle\|f\|_{p}^{p} |U|1p2τSsUUτ(Uθτ|fθ|2)p/2s4p\displaystyle\lessapprox|U|^{1-\frac{p}{2}}\sum_{\tau\in\textbf{S}_{s}}\sum_{U\parallel U_{\tau}}\bigg{(}\int_{U}\sum_{\theta\subset\tau}|f_{\theta}|^{2}\bigg{)}^{p/2}s^{4-p}
|U|1p2(RβR1s1)p2τSsUUτ(Uγτ|fγ|2)p/2s4p\displaystyle\lesssim|U|^{1-\frac{p}{2}}(R^{\beta}R^{-1}s^{-1})^{\frac{p}{2}}\sum_{\tau\in\textbf{S}_{s}}\sum_{U\parallel U_{\tau}}\bigg{(}\int_{U}\sum_{\gamma\subset\tau}|f_{\gamma}|^{2}\bigg{)}^{p/2}s^{4-p}
(Hölder)\displaystyle(\textup{H\"{o}lder}) s4p(RβR1s1)p2τSsUUτU(γτ|fγ|2)p/2\displaystyle\leq s^{4-p}(R^{\beta}R^{-1}s^{-1})^{\frac{p}{2}}\sum_{\tau\in\textbf{S}_{s}}\sum_{U\parallel U_{\tau}}\int_{U}(\sum_{\gamma\subset\tau}|f_{\gamma}|^{2})^{p/2}
s4p(RβR1s1)p2(γΓβ(R1)|fγ|2)12pp.\displaystyle\leq s^{4-p}(R^{\beta}R^{-1}s^{-1})^{\frac{p}{2}}\|(\sum_{\gamma\in\Gamma_{\beta}(R^{-1})}|f_{\gamma}|^{2})^{\frac{1}{2}}\|_{p}^{p}.

We just need to check

s4p(RβR1s1)p2Rβp2+p42,s^{4-p}(R^{\beta}R^{-1}s^{-1})^{\frac{p}{2}}\leq R^{\frac{\beta p}{2}+\frac{p}{4}-2},

which is equivalent to

s43p2R3p42.s^{4-\frac{3p}{2}}\leq R^{\frac{3p}{4}-2}.

Plugging s1R1/2s^{-1}\leq R^{1/2} and noting that 43p2<04-\frac{3p}{2}<0, we prove the result.

References

  • [1] C. Demeter. Fourier restriction, decoupling and applications, volume 184. Cambridge University Press, 2020.
  • [2] C. Demeter, L. Guth, and H. Wang. Small cap decouplings. Geometric and Functional Analysis, 30(4):989–1062, 2020.
  • [3] P. Gressman, L. Pierce, J. Roos, and P.-L. Yung. A new type of superorthogonality. Proceedings of the American Mathematical Society, 152(02):665–675, 2024.
  • [4] P. T. Gressman, S. Guo, L. B. Pierce, J. Roos, and P.-L. Yung. Reversing a philosophy: from counting to square functions and decoupling. The Journal of Geometric Analysis, 31(7):7075–7095, 2021.
  • [5] L. Guth and D. Maldague. Amplitude dependent wave envelope estimates for the cone in 3\mathbb{R}^{3}. arXiv preprint arXiv:2206.01093, 2022.
  • [6] L. Guth, H. Wang, and R. Zhang. A sharp square function estimate for the cone in 3\mathbb{R}^{3}. Annals of Mathematics, 192(2):551–581, 2020.
  • [7] D. Maldague. A sharp square function estimate for the moment curve in 3\mathbb{R}^{3}. arXiv preprint arXiv:2210.17436, 2022.
  • [8] L. B. Pierce. On superorthogonality. The Journal of Geometric Analysis, 31:7096–7183, 2021.
  • [9] T. Tao. A sharp bilinear restriction estimate for paraboloids. Geometric and Functional Analysis, 13:1359–1384, 2003.