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

A note on local smoothing estimates for fractional Schrödinger equations

Shengwen Gan Shengwen Gan
Deparment of Mathematics, Massachusetts Institute of Technology, USA
[email protected]
Changkeun Oh Changkeun Oh
Department of Mathematics, University of Wisconsin-Madison, USA
[email protected]
 and  Shukun Wu Shukun Wu
Department of Mathematics
California Institute of Technology, USA
[email protected]
Abstract.

We improve local smoothing estimates for fractional Schrödinger equations for α(0,1)(1,)\alpha\in(0,1)\cup(1,\infty).

1. Introduction

Let n3n\geq 3 and α(0,1)(1,)\alpha\in(0,1)\cup(1,\infty). Consider the fractional Schrödinger equation

(1.1) {iut+(Δ)α/2u=0u(x,0)=g(x).\begin{cases}i\frac{\partial u}{\partial t}+(-\Delta)^{\alpha/2}u=0\\ u(x,0)=g(x).\end{cases}

The case α=2\alpha=2 corresponds to the Schrödinger equation. For g𝒮(n1)g\in\mathcal{S}(\mathbb{R}^{n-1}), the solution of (1.1) can be written as

(1.2) eit(Δ)α/2g(x):=n1g^(ξ)e(xξ+t|ξ|α)𝑑ξ.e^{it(-\Delta)^{\alpha/2}}g(x):=\int_{\mathbb{R}^{n-1}}\widehat{g}(\xi)e(x\cdot\xi+t|\xi|^{\alpha})\,d\xi.

Here, we use the notation e(t):=e2πite(t):=e^{2\pi it}. Denote the standard Bessel potential space by Wβ,pW^{\beta,p}. For simplicity, we define βc=βc(n)\beta_{c}=\beta_{c}(n) satisfying βcα=(n1)(121p)1p\frac{\beta_{c}}{\alpha}=(n-1)(\frac{1}{2}-\frac{1}{p})-\frac{1}{p}.

Conjecture 1.1 (Local smoothing conjecture).

Fix a number α(0,1)(1,)\alpha\in(0,1)\cup(1,\infty). For p>2nn1p>\frac{2n}{n-1}, it holds that

(1.3) eit(Δ)α/2gLx,tp(n1×[0,1])gWβ,p(n1)\Big{\|}e^{it(-\Delta)^{\alpha/2}}g\Big{\|}_{L^{p}_{x,t}(\mathbb{R}^{n-1}\times[0,1])}\lesssim\|g\|_{W^{\beta,p}(\mathbb{R}^{n-1})}

for every β>βc\beta>\beta_{c}.

Let pnp_{n} be the exponent for which the Fourier restriction problem for paraboloid in n\mathbb{R}^{n} is verified in [HZ20]. Note that when n4n\geq 4 this is the best exponent for the restriction problem. Our main theorem is as follows.

Theorem 1.2.

We consider α\alpha in the range below

(1.4) {α(0,1)(1,) for n=3α(1,) for n>3.\begin{cases}\alpha\in(0,1)\cup(1,\infty)\text{ for }n=3\\ \alpha\in(1,\infty)\text{ for }n>3.\end{cases}

For ppnp\geq p_{n}, it holds that

(1.5) eit(Δ)α/2gLx,tp(n1×[0,1])gWβ,p(n1)\Big{\|}e^{it(-\Delta)^{\alpha/2}}g\big{\|}_{L^{p}_{x,t}(\mathbb{R}^{n-1}\times[0,1])}\lesssim\|g\|_{W^{\beta,p}(\mathbb{R}^{n-1})}

for every β>βc\beta>\beta_{c}.

Our strategy is to adapt the method in the Fourier restriction problem to the local smoothing. We remark that the case α=2\alpha=2 is already known. Actually, in the work of Rogers [Rog08], it is proved that the local smoothing estimate for Schrödinger equation follows from the restriction estimate for paraboloid. Thus, when α=2\alpha=2, the recent restriction results in [HZ20] and [Wan18] already imply certain local smoothing estimates. However, for other α\alpha, an analogous implication has not been discovered and our result is new. For the case α>1\alpha>1, we adapt the ideas of [GOW21a] and improve the results in [GRY20] (Theorem 1.6), [GMZ20], and in [GOW21a] (Corollary 1.5). The case 0<α<10<\alpha<1 is slightly different from the case α>1\alpha>1 as the manifold associated to the operator has negative Gaussian curvature. We combine the ideas of [GO20] and [GOW21a], and improve the results in [RS10].

One of our main tools is the polynomial partitioning, which was introduced by Guth to the study of oscillatory integral operators in [Gut16] and [Gut18]. Another main tool is the polynomial Wolff axiom, which was first formulated by Guth and Zahl in [GZ18], and later proved by Katz and Rogers in [KR18]. Furthermore, a refined version called the nested polynomial Wolff axiom was verified independently in [HRZ19] and [Zah21]. What’s more, there is a stronger result which is a slice version of polynomial Wolff (see (3.69) and (5.17)). This result was implicitly proved in [Zah21], and we will use it in our proof.

In the appendix, we explore some connections between local smoothing estimates for the fractional Schrödinger equations and restriction type estimates. Our argument is based on the ideas in [GOW+21b], and it generalizes the result in [Rog08].

This article is not intended to be self-contained and refers to [HZ20] and [GO20].

Organization of the paper. In Section 2, we reduce the local smoothing estimate to a localized estimate (2.8), and review the wave packet decomposition. In Section 3, we prove Theorem 1.2 for the model case (α,n)=(2,3)(\alpha,n)=(2,3), and later we will use similar arguments to prove it for other cases. In Section 4 and 5, we prove Theorem 1.2 for the case α<1\alpha<1 and α>1\alpha>1, respectively. In the appendix, we study the relationship between the local smoothing and Fourier restriction estimates.

Notation.

  • Denote Brn(x)B^{n}_{r}(x) the ball of radius rr centered at xx in n\mathbb{R}^{n}. We sometimes write Br(x)B_{r}(x) or BrB_{r} for Brn(x)B_{r}^{n}(x) provided that there will be no confusion. We also sometimes use BRnB_{R}^{n} and BnB^{n} for BRn(0)B_{R}^{n}(0) and B1n(0)B_{1}^{n}(0), respectively.

  • We write A(R)RapDec(R)BA(R)\leq\mathrm{RapDec}(R)B to mean that for any power β\beta, there is a constant CNC_{N} such that

    A(R)CNRNBfor all R1.A(R)\leq C_{N}R^{-N}B\;\;\text{for all $R\geq 1$}.

Acknowledgement. The authors would like to thank Andreas Seeger for helpful discussions on the case α<1\alpha<1. The authors would also like to thank Shaoming Guo, Hong Wang and Ruixiang Zhang for valuable discussions on the appendix of this manuscript. C.O. was partially supported by the NSF grant DMS-1800274.

2. Preliminaries

In this section, we reduce the local smoothing estimate to the localized version of it. We also review the wave packet decomposition.

2.1. Some reductions

In this subsection, we do some reductions to make Theorem 1.2 more like a Fourier restriction problem.

We apply some change of variables and assume that the integral of tt runs over [1/2,1][1/2,1]. We first consider functions gg whose Fourier support is on the annulus |ξ|1|\xi|\sim 1. For such gg, we will prove that

(2.1) eit(Δ)α/2gLx,tp(n1×[R/2,R])R(n1)(121p)+ϵgLp(n1).\big{\|}e^{it(-\Delta)^{\alpha/2}}g\big{\|}_{L^{p}_{x,t}(\mathbb{R}^{n-1}\times[R/2,R])}\lesssim R^{(n-1)(\frac{1}{2}-\frac{1}{p})+\epsilon}\|g\|_{L^{p}(\mathbb{R}^{n-1})}.

Let us assume this inequality for a moment and prove Theorem 1.2. By the Littlewood-Paley decomposition, it suffices to show (1.5) for functions g1g_{1} whose Fourier support is contained in the annulus {ξ:|ξ|(R)1/α}\{\xi:|\xi|\sim(R)^{1/\alpha}\}. By a simple scaling argument, we have

(2.2) eiRt(Δ)α/2g(x)=R(n1)/α(eit(Δ)α/2g1)(R1/αx)e^{iRt(-\Delta)^{\alpha/2}}g(x)=R^{-(n-1)/\alpha}(e^{it(-\Delta)^{\alpha/2}}g_{1})(R^{-1/\alpha}x)

where

(2.3) g1^(ξ)=:g^(R1/αξ).\widehat{g_{1}}(\xi)=:\widehat{g}(R^{-1/\alpha}\xi).

Notice that the function gg has Fourier support in the unit ball. By applying (2.1) and a scaling argument, we obtain

(2.4) eit(Δ)α/2g1Lp(n1×[1/2,1])RϵR(n1)(121p)g1Wβ,p(n1)\|e^{it(-\Delta)^{\alpha/2}}g_{1}\|_{L^{p}(\mathbb{R}^{n-1}\times[1/2,1])}\lesssim R^{\epsilon}R^{(n-1)(\frac{1}{2}-\frac{1}{p})}\|g_{1}\|_{W^{\beta,p}(\mathbb{R}^{n-1})}

for every function g1g_{1} whose Fourier support is in a ball of radius R1/αR^{1/\alpha}.

We have showed that (1.5) follows from (2.1). We can make a further reduction. By a standard localization argument (for example, see Lemma 8 of [Rog08] or Corollary 2.2 of [GMZ20]), it suffices to prove

(2.5) eit(Δ)α/2gLx,tp(BRn1×[R/2,R])R(n1)(121p)+ϵgLp(n1).\big{\|}e^{it(-\Delta)^{\alpha/2}}g\big{\|}_{L^{p}_{x,t}(B_{R}^{n-1}\times[R/2,R])}\lesssim R^{(n-1)(\frac{1}{2}-\frac{1}{p})+\epsilon}\|g\|_{L^{p}(\mathbb{R}^{n-1})}.

for every function gg whose Fourier support is on |ξ|1|\xi|\sim 1.

We will reformulate the inequality (2.5) using a different language in order to be consistent with the paper for the Fourier restriction problem. Let SS be a hypersurface in n\mathbb{R}^{n}, represented by the graph of a function hh. Define the extension operator corresponding to the hypersurface SS by

(2.6) Ef(x)=[1,1]n1f(ξ)e(ξx+h(ξ)xn)𝑑ξ,Ef(x)=\int_{[-1,1]^{n-1}}f(\xi)e\big{(}\xi\cdot x^{\prime}+h(\xi)x_{n}\big{)}\,d\xi,

where x=(x,xn)n1×x=(x^{\prime},x_{n})\in\mathbb{R}^{n-1}\times\mathbb{R}. Notice that the hypersurface {(ξ,|ξ|α)}\{(\xi,|\xi|^{\alpha})\} has all positive principal curvatures for the case α>1\alpha>1, and it has principal curvatures with different signs for the case α<1\alpha<1. Therefore, the inequality (2.5) follows from the following theorem.

Theorem 2.1.

Suppose that h:n1h:\mathbb{R}^{n-1}\rightarrow\mathbb{R} satisfies

(2.7) {ξξh has nonzero eigenvalues, for n=3ξξh has only positive eigenvalues, for n>3.\begin{cases}\nabla_{\xi\xi}h\text{ has nonzero eigenvalues, for }n=3\\ \nabla_{\xi\xi}h\text{ has only positive eigenvalues, for }n>3.\end{cases}

For ppnp\geq p_{n}, it holds that

(2.8) EfLp(BRn)Cp,ϵRϵR(n1)(121p)f^Lp(n1)\|Ef\|_{L^{p}(B_{R}^{n})}\leq C_{p,\epsilon}R^{\epsilon}R^{(n-1)(\frac{1}{2}-\frac{1}{p})}\|\widehat{f}\|_{L^{p}(\mathbb{R}^{n-1})}

for every number R1R\geq 1 and function ff supported in [1,1]n1[-1,1]^{n-1}.

2.2. Wave packet decomposition

In this subsection, we review the wave packet decomposition.

Let 1rR1\leq r\leq R. We consider a hypersurface {ξ,h(ξ)}\{\xi,h(\xi)\} and a function ff on Bn1B^{n-1}. For any nn-dimensional ball BrB_{r}, let us do wave packet decomposition of ff associated to this ball.

Take a collection Θr\Theta_{r} of finitely overlapping balls θ\theta of radius r1/2r^{-1/2}, and take ψθ\psi_{\theta} a smooth partition of unity adapted to this cover, and write f=θΘrψθff=\sum_{\theta\in\Theta_{r}}\psi_{\theta}f. We next cover n1\mathbb{R}^{n-1} by finitely overlapping balls of radius r(1+δ)/2\sim r^{(1+\delta)/2} centered at vr(1+δ)/2n1v\in r^{(1+\delta)/2}\mathbb{Z}^{n-1}. For each vv, let ηv\eta_{v} be a smooth partition of unity adapted to this cover. We now have the decomposition

(2.9) f=(θ,v)Θr×r(1+δ)/2n1ηv(ψθf).f=\sum_{(\theta,v)\in\Theta_{r}\times r^{(1+\delta)/2}\mathbb{Z}^{n-1}}\eta_{v}^{\vee}*(\psi_{\theta}f).

We want to make the summand compactly supported, so multiply a function ψ~θ\tilde{\psi}_{\theta}, which is supported on 2θ2\theta and equal to one on r1/2r^{-1/2}-neighborhood of the support of ψθ\psi_{\theta}. Since ηv\eta_{v}^{\vee} decays rapidly outside of a ball of radius r(1+δ)/2\sim r^{-(1+\delta)/2} centered at the origin, we have

(2.10) f=(θ,v)Θr×r(1+δ)/2n1ψ~θ(ηv(ψθf))+RapDec(r)f2.f=\sum_{(\theta,v)\in\Theta_{r}\times r^{(1+\delta)/2}\mathbb{Z}^{n-1}}\tilde{\psi}_{\theta}(\eta_{v}^{\vee}*(\psi_{\theta}f))+\mathrm{RapDec}(r)\|f\|_{2}.

Define fθ,vf_{\theta,v} as the summand on the right hand side of (2.12). Denote by wθw_{\theta} the center of the ball θ\theta. Let (θ,v)(\theta,v) be given, and define the tube Tθ,vT_{\theta,v} by

(2.11) Tθ,v:={(x,xn)Br:|x+xnωh(ωθ)+v|r1/2+δ}.T_{\theta,v}:=\big{\{}(x^{\prime},x_{n})\in B_{r}:|x^{\prime}+x_{n}\partial_{\omega}h(\omega_{\theta})+v|\leq r^{1/2+\delta}\big{\}}.

Define 𝕋[Br]\mathbb{T}[B_{r}] to be the collection of the tubes Tθ,vT_{\theta,v}. If the set Tθ,vT_{\theta,v} is empty, then we do not include this set in 𝕋[Br]\mathbb{T}[B_{r}]. We sometimes use the notation

(2.12) fTθ,v:=fθ,v=ψ~θ(ηv(ψθf)).f_{T_{\theta,v}}:=f_{\theta,v}=\tilde{\psi}_{\theta}(\eta_{v}^{\vee}*(\psi_{\theta}f)).

This finishes the wave packet decomposition on the ball BrB_{r}.

We next consider the ball Br(x0)B_{r}(x_{0}) of radius rr centered at the point x0x_{0}. Define Tθ,v(x0):=Tθ,v+x0T_{\theta,v}(x_{0}):=T_{\theta,v}+x_{0} and the collection of the tubes

(2.13) 𝕋[Br(x0)]:={Tθ,v(x0):(θ,v)Θr×r(1+δ)/2n1}.\mathbb{T}[B_{r}(x_{0})]:=\{T_{\theta,v}(x_{0}):(\theta,v)\in\Theta_{r}\times r^{(1+\delta)/2}\mathbb{Z}^{n-1}\}.

Define the phase function ϕ(x,ω):=xω+xnh(ω)\phi(x,\omega):=x^{\prime}\cdot\omega+x_{n}h(\omega), and the wave packet

(2.14) fTθ,v(x0)(ω):=e(ϕ(x0,ω))(f()e(ϕ(x0,)))θ,v(ω).f_{T_{\theta,v}(x_{0})}(\omega):=e(-\phi(x_{0},\omega))\Big{(}f(\cdot)e(\phi(x_{0},\cdot))\Big{)}_{\theta,v}(\omega).

Notice that we have the decomposition

(2.15) f=(θ,v)fTθ,v(x0)+RapDec(r)f2.f=\sum_{(\theta,v)}f_{T_{\theta,v}(x_{0})}+\mathrm{RapDec}(r)\|f\|_{2}.

A key property of the wave packet is as follows. We refer to Proposition 2.6 of [Gut16] for the details.

Proposition 2.2.

For every xBr(x0)Tθ,v(x0)x\in B_{r}(x_{0})\setminus T_{\theta,v}(x_{0}), we have

(2.16) EfTθ,v(𝐱0)(x)=RapDec(r)f2.Ef_{T_{\theta,v}({\bf x}_{0})}(x)=\mathrm{RapDec}(r)\|f\|_{2}.

Therefore, on the ball Br(x0)B_{r}(x_{0}), we have the wave packet decomposition

(2.17) Ef(x)=T𝕋[Br(x0)]EfT(x)+RapDec(r)f2,Ef(x)=\sum_{T\in\mathbb{T}[B_{r}(x_{0})]}Ef_{T}(x)+\mathrm{RapDec}(r)\|f\|_{2},

and the function EfTEf_{T} decays rapidly outside of the tube TT.

Next we discuss how to compare wave packets at different scales. Suppose that we have two scales 1ρr1\leq\rho\leq r. Consider balls Bρ(x1)Br(x0)B_{\rho}(x_{1})\subset B_{r}(x_{0}). Given 𝕎𝕋[Bρ(x1)]\mathbb{W}\subset\mathbb{T}[B_{\rho}(x_{1})], define 𝕎\uparrow\!\mathbb{W} to be a collection of the tubes in 𝕋[Br(x0)]\mathbb{T}[B_{r}(x_{0})] satisfying that for every element Tθ,v𝕎T_{\theta,v}\in\uparrow\!\mathbb{W} there exists Tθ,v𝕎T_{\theta^{\prime},v^{\prime}}\in\mathbb{W} such that

(2.18) dist(θ,θ)ρ1/2 and dist(Tθ,v(x0)Bρ(x1),Tθ,v(x1))r1/2+δ.\mathrm{dist}(\theta,\theta^{\prime})\lesssim\rho^{-1/2}\text{ \; and \; }\mathrm{dist}(T_{\theta,v}(x_{0})\cap B_{\rho}(x_{1}),T_{\theta^{\prime},v^{\prime}}(x_{1}))\lesssim r^{1/2+\delta}.

For every 𝕎𝕋[Bρ(x1)]\mathbb{W}\subset\mathbb{T}[B_{\rho}(x_{1})], we define

(2.19) g|𝕎:=T𝕎gT,g|𝕎:=T𝕎gT.g|_{\mathbb{W}}:=\sum_{T\in\mathbb{W}}g_{T},\;\;\;\;\;g|_{\uparrow\mathbb{W}}:=\sum_{T\in\uparrow\mathbb{W}}g_{T}.

The lemma below relates small wave packets to larger wave packets.

Lemma 2.3 (Lemma 5.2 of [HZ20]).

Given gL1(Bn1)g\in L^{1}(B^{n-1}) and 𝕎𝕋[Bρ]{\mathbb{W}}\subset\mathbb{T}[B_{\rho}],

(2.20) g|𝕎=(g|𝕎)|𝕎+RapDec(r)g2.g|_{{\mathbb{W}}}=\big{(}g|_{\uparrow{\mathbb{W}}}\big{)}\big{|}_{\mathbb{W}}+\mathrm{RapDec}(r)\|g\|_{2}.

3. The discussion of a model case: α=2\alpha=2 in 3\mathbb{R}^{3}

In this section, we prove Theorem 2.1 for the case h(ξ)=|ξ|2h(\xi)=|\xi|^{2} and n=3n=3. The reason that we discuss this model case is to avoid irrelevant technical difficulties and shed light on our method. After we finish the proof of this model case, we will see in later sections that our method can be easily adapted to the proofs for other cases in Theorem 2.1. Actually, we classify them into two cases: 0<α<10<\alpha<1 in 3\mathbb{R}^{3}; α>1\alpha>1 in any dimensions. Both cases have been studied in the setting of Fourier restriction problem. In later sections, we will derive the local smoothing estimates by combining the techniques in Fourier restriction problem and the method for the model case discussed in this section.

In this section, we assume our Fourier extension operator is

Ef(x):=[1,1]2f(ξ)e(ξx+|ξ|2x3)𝑑ξ,Ef(x):=\int_{[-1,1]^{2}}f(\xi)e(\xi\cdot x^{\prime}+|\xi|^{2}x_{3})d\xi,

where x=(x,x3)3x=(x^{\prime},x_{3})\in\mathbb{R}^{3}. Our goal is to prove the following estimate.

Theorem 3.1.

For p13/4p\geq 13/4, we have

(3.1) EfLp(BR)Cp,ϵRϵR(12p)f^Lp(2)\|Ef\|_{L^{p}(B_{R})}\leq C_{p,\epsilon}R^{\epsilon}R^{(1-\frac{2}{p})}\|\widehat{f}\|_{L^{p}(\mathbb{R}^{2})}

for every number R1R\geq 1 and function ff supported in [1,1]2[-1,1]^{2}.

3.1. Broad function

We follow [Gut16] to define the broad function.

Let [1,1]2=τ[-1,1]^{2}=\cup\tau be a finitely overlapping covering, where τ\tau are K1K^{-1}-squares. We will later set K=eϵ10K=e^{\epsilon^{-10}}. We consider a decomposition f=τfτf=\sum_{\tau}f_{\tau}, where suppfττ\textup{supp}f_{\tau}\subset\tau.

For α(0,1)\alpha\in(0,1), we say that xx is α\alpha-broad for EfEf if:

maxτ|Efτ(x)|α|Ef(x)|.\max_{\tau}|Ef_{\tau}(x)|\leq\alpha|Ef(x)|.

We define BrαEf(x){\rm{Br}}_{\alpha}Ef(x) to be Ef(x)Ef(x) if xx is α\alpha-broad for EfEf and zero otherwise.

We will prove the following estimate for broad functions.

Theorem 3.2.

For any ϵ>0\epsilon>0, there exists K=K(ϵ)K=K(\epsilon) such that for every R1R\geq 1, we have

(3.2) BrKϵEfL13/4(BR)ϵRϵR2(12413)fL28/13f^L5/13.\begin{split}\|\mathrm{Br}_{K^{-\epsilon}}Ef\|_{L^{13/4}(B_{R})}\lesssim_{\epsilon}R^{\epsilon}R^{2(\frac{1}{2}-\frac{4}{13})}\|f\|_{L^{2}}^{8/13}\|\widehat{f}\|_{L^{\infty}}^{5/13}.\end{split}

Moreover, limϵ0K(ϵ)=+\lim_{\epsilon\rightarrow 0}K(\epsilon)=+\infty.

In the rest of this subsection, we quickly see how to prove (3.1) assuming Theorem 3.2. By interpolation, it suffices to prove

(3.3) EfL13/4(BR)CϵRϵR2(12413)fL28/13f^L5/13.\begin{split}\|Ef\|_{L^{13/4}(B_{R})}\leq C_{\epsilon}R^{\epsilon}R^{2(\frac{1}{2}-\frac{4}{13})}\|f\|_{L^{2}}^{8/13}\|\widehat{f}\|_{L^{\infty}}^{5/13}.\end{split}

We induct on RR. Assume (3.3) is true for radius R/2\leq R/2. We first show the following rescaling lemma.

Lemma 3.3.

Assuming (3.3) is true for radius R/2\leq R/2, then for any K1K^{-1}-square τ[1,1]2\tau\subset[-1,1]^{2}, we have

(3.4) EfτL13/4(BR)CϵK2+6413RϵR2(12413)fτL28/13f^L5/13.\begin{split}\|Ef_{\tau}\|_{L^{13/4}(B_{R})}\lesssim C_{\epsilon}K^{-2+6\cdot\frac{4}{13}}R^{\epsilon}R^{2(\frac{1}{2}-\frac{4}{13})}\|f_{\tau}\|_{L^{2}}^{8/13}\|\widehat{f}\|_{L^{\infty}}^{5/13}.\end{split}
Proof.

Since this parabolic rescaling argument is well-known, we only do the calculation for τ\tau centered at the origin.

Let g(ξ)=fτ(ξ/K)g(\xi)=f_{\tau}(\xi/K). We have

Eg(x)\displaystyle Eg(x) =[K1,K1]2fτ(ξ/K)e(ξx+|ξ|2x3)𝑑ξ\displaystyle=\int_{[-K^{-1},K^{-1}]^{2}}f_{\tau}(\xi/K)e(\xi\cdot x^{\prime}+|\xi|^{2}x_{3})d\xi
=K2[1,1]2f(ξ)e(ξKx+|ξ|2K2x3)𝑑ξ\displaystyle=K^{2}\int_{[-1,1]^{2}}f(\xi)e(\xi\cdot Kx^{\prime}+|\xi|^{2}K^{2}x_{3})d\xi
=K2Efτ(Kx,K2x3).\displaystyle=K^{2}Ef_{\tau}(Kx^{\prime},K^{2}x_{3}).

By change of variable, we have

EfτL13/4(BR)K14/13EgL13/4([R/K,R/K]2×[R/K2,R/K2]).\|Ef_{\tau}\|_{L^{13/4}(B_{R})}\leq K^{-14/13}\|Eg\|_{L^{13/4}([-R/K,R/K]^{2}\times[-R/K^{2},R/K^{2}])}.

Also, we have fτL2=K1gL2\|f_{\tau}\|_{L^{2}}=K^{-1}\|g\|_{L^{2}}, f^τL=K2g^L\|\widehat{f}_{\tau}\|_{L^{\infty}}=K^{-2}\|\widehat{g}\|_{L^{\infty}}. Since fτ=χτff_{\tau}=\chi_{\tau}f such that χ^τ\widehat{\chi}_{\tau} is L1L^{1}-bounded, we have g^L=K2f^τLK2f^L\|\widehat{g}\|_{L^{\infty}}=K^{2}\|\widehat{f}_{\tau}\|_{L^{\infty}}\lesssim K^{2}\|\widehat{f}\|_{L^{\infty}}.

We divide [R/K,R/K]2×[R/K2,R/K2][-R/K,R/K]^{2}\times[-R/K^{2},R/K^{2}] into K\sim K balls of radius R/K2R/K^{2}. Since gg is supported in [1,1]2[-1,1]^{2}. For each such ball BR/K2B_{R/K^{2}}, by the hypothesis we have

EgL13/4(BR/K2)Cϵ(R/K2)ϵ(R/K2)2(12413)gL28/13g^L5/13.\|Eg\|_{L^{13/4}(B_{R/K^{2}})}\leq C_{\epsilon}(R/K^{2})^{\epsilon}(R/K^{2})^{2(\frac{1}{2}-\frac{4}{13})}\|g\|_{L^{2}}^{8/13}\|\widehat{g}\|_{L^{\infty}}^{5/13}.

Summing over BR/K2B_{R/K^{2}}, we get

EgL13/4([R/K,R/K]2×[R/K2,R/K2])CϵK413(R/K2)ϵ(R/K2)2(12413)gL28/13g^L5/13.\|Eg\|_{L^{13/4}([-R/K,R/K]^{2}\times[-R/K^{2},R/K^{2}])}\leq C_{\epsilon}K^{\frac{4}{13}}(R/K^{2})^{\epsilon}(R/K^{2})^{2(\frac{1}{2}-\frac{4}{13})}\|g\|_{L^{2}}^{8/13}\|\widehat{g}\|_{L^{\infty}}^{5/13}.

Plugging in fτf_{\tau} and noting that f^τLf^L\|\widehat{f}_{\tau}\|_{L^{\infty}}\lesssim\|\widehat{f}\|_{L^{\infty}}, we obtain

EfτL13/4(BR)CϵK2+6413RϵR2(12413)fτL28/13f^L5/13,\|Ef_{\tau}\|_{L^{13/4}(B_{R})}\lesssim C_{\epsilon}K^{-2+6\cdot\frac{4}{13}}R^{\epsilon}R^{2(\frac{1}{2}-\frac{4}{13})}\|f_{\tau}\|_{L^{2}}^{8/13}\|\widehat{f}\|_{L^{\infty}}^{5/13},

which finishes the proof. ∎

We continue the proof of (3.3). Recall K=eϵ10K=e^{\epsilon^{-10}} and we set α=Kϵ\alpha=K^{-\epsilon}. By the definition of the broad function, we have

(3.5) |Ef(x)||BrαEf(x)|+α1supτ|Efτ(x)|.|Ef(x)|\leq|{\rm{Br}}_{\alpha}Ef(x)|+\alpha^{-1}\sup_{\tau}|Ef_{\tau}(x)|.

As a result, we obtain

(3.6) BR|Ef(x)|13/4BR|BrαEf(x)|13/4+α1τBR|Efτ(x)|13/4.\int_{B_{R}}|Ef(x)|^{13/4}\leq\int_{B_{R}}|{\rm{Br}}_{\alpha}Ef(x)|^{13/4}+\alpha^{-1}\sum_{\tau}\int_{B_{R}}|Ef_{\tau}(x)|^{13/4}.

Using Theorem 3.2, we can bound the first term by

110(CϵRϵR2(12413))13/4fL22f^L1.25,\frac{1}{10}\big{(}C_{\epsilon}R^{\epsilon}R^{2(\frac{1}{2}-\frac{4}{13})}\big{)}^{13/4}\|f\|_{L^{2}}^{2}\|\widehat{f}\|_{L^{\infty}}^{1.25},

if CϵC_{\epsilon} is large enough. For the second term, we apply Lemma 3.3 to obtain

(3.7) α1τBR|Efτ(x)|13/4Kϵτ(CϵK2+6413RϵR2(12413))13/4fτL22f^L1.25110(CϵRϵR2(12413))13/4fL22f^L1.25.\begin{split}\alpha^{-1}\sum_{\tau}\int_{B_{R}}|Ef_{\tau}(x)|^{13/4}&\lesssim K^{\epsilon}\sum_{\tau}\big{(}C_{\epsilon}K^{-2+6\cdot\frac{4}{13}}R^{\epsilon}R^{2(\frac{1}{2}-\frac{4}{13})}\big{)}^{13/4}\|f_{\tau}\|_{L^{2}}^{2}\|\widehat{f}\|_{L^{\infty}}^{1.25}\\ &\leq\frac{1}{10}\big{(}C_{\epsilon}R^{\epsilon}R^{2(\frac{1}{2}-\frac{4}{13})}\big{)}^{13/4}\|f\|_{L^{2}}^{2}\|\widehat{f}\|_{L^{\infty}}^{1.25}.\end{split}

Here we use τfτ22f22\sum_{\tau}\|f_{\tau}\|_{2}^{2}\lesssim\|f\|_{2}^{2} and we assume KK is large enough so that the negative power of KK cancels the implicit constant.

Combining things together, we proved (3.3).

In the remaining of the section, we prove Theorem 3.2. We choose K(ϵ)=eϵ10K(\epsilon)=e^{\epsilon^{-10}}. Without loss of generality, we can assume the wave packets of ff are contained in B10RB_{10R} since those outside B10RB_{10R} contribute little to the left hand side of (3.2). The proof includes several steps. First, we will do a one-step polynomial partitioning, where we will deal with three cases: cellular case, transverse case and tangent case. Next, we iterate the polynomial partitioning until we encounter the tangent case or the radius becomes small. Finally, we will combine the estimates from polynomial partitioning and the sliced polynomial Wolff estimate to derive our result.

3.2. One-step polynomial partitioning

First, we recall the polynomial partitioning lemma. Here we use the modified version which gives additional information on the radius of cells. For more details, we refer to [Wan18] page 10.

Lemma 3.4.

Let FF be a non-negative L1L^{1} function on n\mathbb{R}^{n}. Then for any D+D\in\mathbb{Z}^{+}, there is a non-zero polynomial PP of degree at most DD so that nZ(P)\mathbb{R}^{n}\setminus Z(P) is a disjoint union of Dn\sim D^{n} open sets OiO_{i}, and the integrals OiF\int_{O_{i}}F agree up to a factor of 2. Moreover, the polynomial PP is a product of non-singular polynomials, and each set OiO_{i} is contained in a ball of radius R/DR/D.

Take the parameters D=Rϵ6D=R^{\epsilon^{6}} and δ=ϵ2\delta=\epsilon^{2}.

Let us apply Lemma 3.4 to the function F=|1BRBrαEf|13/4F=|1_{B_{R}}\mathrm{Br}_{\alpha}Ef|^{13/4}. Then there exists a non-zero polynomial PP of degree at most DD so that 3Z(P)\mathbb{R}^{3}\setminus Z(P) is a disjoint union of D3\sim D^{3} cells OiO_{i} and so that for each ii,

(3.8) OiBR(BrαEf)13/4D3BR(BrαEf)13/4.\int_{O_{i}\cap B_{R}}(\mathrm{Br}_{\alpha}{Ef})^{13/4}\sim D^{-3}\int_{B_{R}}(\mathrm{Br}_{\alpha}{Ef})^{13/4}.

Moreover, we can assume that PP is a product of non-singular polynomials.

Define the wall and the cells by

(3.9) W:=NR1/2+δ(Z(P)),Oi(1):=(OiBR)W.W:=N_{R^{1/2+\delta}}(Z(P)),\;\;\;O_{i}^{(1)}:=(O_{i}\cap B_{R})\setminus W.

Here, the superscript “(1)(1)” on Oi(1)O_{i}^{(1)} indicates that we are doing the first polynomial partition. We have the following estimate

(3.10) BrαEfL13/4(BR)13/4Oi(1)BrαEfL13/4(Oi(1))13/4+BrαEfL13/4(W)13/4.\|\mathrm{Br}_{\alpha}Ef\|_{L^{13/4}(B_{R})}^{13/4}\lesssim\sum_{O_{i}^{(1)}}\|\mathrm{Br}_{\alpha}Ef\|_{L^{13/4}(O_{i}^{(1)})}^{13/4}+\|\mathrm{Br}_{\alpha}Ef\|_{L^{13/4}(W)}^{13/4}.

We say that we are in the cellular case if the first term dominates the second term. Otherwise, we say that we are in the wall case.

Let us deal with the integral of the broad function on WW. We cover WW by balls BkB_{k} of radius R1δR^{1-\delta}. For each BkB_{k}, we will define tangent tubes and transverse tubes associated to BkB_{k}.

Definition 3.5 (Tangent tubes).

Define 𝕋k,\mathbb{T}_{k,-} to be the set of all tubes T𝕋[BR]T\in\mathbb{T}[B_{R}] obeying the following conditions:

\bullet TBkWT\cap B_{k}\cap W\neq\emptyset.

\bullet If zz is any non-singular point of Z(P)Z(P) lying in 2Bk10T2B_{k}\cap 10T, then

(3.11) Angle(v(T),Tz(Z(P)))R1/2+2δ.\mathrm{Angle}(v(T),{T}_{z}(Z(P)))\leq R^{-1/2+2\delta}.

Recall that v(T)v(T) is the unit vector in the direction of the tube TT.

Definition 3.6 (Transverse tubes).

Define 𝕋k,+\mathbb{T}_{k,+} to be the set of all tubes T𝕋[BR]T\in\mathbb{T}[B_{R}] obeying the following conditions:

\bullet TBkWT\cap B_{k}\cap W\neq\emptyset.

\bullet There exists a non-singular point zz of Z(P)Z(P) lying in 2Bk10T2B_{k}\cap 10T, so that

(3.12) Angle(v(T),Tz(Z(P)))>R1/2+2δ.\mathrm{Angle}(v(T),{T}_{z}(Z(P)))>R^{-1/2+2\delta}.

Recall that [1,1]2[-1,1]^{2} is divided into K2\sim K^{2} squares τ\tau of diameter K1K^{-1}. We apply the wave packet decomposition (2.15) to fτf_{\tau}, so that

(3.13) fτ=Tfτ,T+RapDec(R)f2.f_{\tau}=\sum_{T}f_{\tau,T}+\mathrm{RapDec(R)}\|f\|_{2}.

We define

(3.14) fτ,k,+:=T𝕋k,+fτ,T,fk,+:=τfτ,k,+.f_{\tau,k,+}:=\sum_{T\in\mathbb{T}_{k,+}}f_{\tau,T},\;\;f_{k,+}:=\sum_{\tau}f_{\tau,k,+}.

For II being a collection of these τ\tau’s, we define

(3.15) fI,k,+:=τIfτ,k,+.f_{I,k,+}:=\sum_{\tau\in I}f_{\tau,k,+}.

Define fk,f_{k,-} similarly. Lastly, we define the bilinear function

(3.16) Bil(Efk,):=τ,τnon adjacent|Efτ,k,|1/2|Efτ,k,|1/2.\mathrm{Bil}(Ef_{k,-}):=\sum_{\tau,\tau^{\prime}\ \textup{non adjacent}}|Ef_{\tau,k,-}|^{1/2}|Ef_{\tau^{\prime},k,-}|^{1/2}.

We have the following lemma:

Lemma 3.7 (Lemma 3.8 of [Gut16]).

If xBkWx\in B_{k}\cap W and α=Kϵ\alpha=K^{-\epsilon}, then

(3.17) Brα|Ef(x)|2(IBr2αEfI,k,+(x)+K100Bil(Efk,)(x)+RapDec(R)f2).{\rm{Br}}_{\alpha}|Ef(x)|\leq 2\bigg{(}\sum_{I}{\rm{Br}}_{2\alpha}Ef_{I,k,+}(x)+K^{100}{\rm{Bil}}(Ef_{k,-})(x)+{\rm RapDec}(R)\|f\|_{2}\bigg{)}.

Here the summation I\sum_{I} is summing over the roughly 2K2\sim 2^{K^{2}} subsets of the set of caps τ\tau.

Now we plug (3.17) into (3.10) to obtain:

(3.18) BrαEfL13/4(BR)13/4Oi(1)BrαEfL13/4(Oi(1))13/4+kIBr2αEfI,k,+L13/4(BkW)13/4+K100Bil(Efj,)L13/4(W)13/4+RapDec(R)f2\begin{split}\|{\rm{Br}}_{\alpha}Ef\|_{L^{13/4}(B_{R})}^{13/4}&\lesssim\sum_{O_{i}^{(1)}}\|{\rm{Br}}_{\alpha}Ef\|^{13/4}_{L^{13/4}(O_{i}^{(1)})}+\sum_{k}\sum_{I}\|{\rm{Br}}_{2\alpha}Ef_{I,k,+}\|^{13/4}_{L^{13/4}(B_{k}\cap W)}\\ &+K^{100}\|{\rm{Bil}}(Ef_{j,-})\|_{L^{13/4}(W)}^{13/4}+{\rm RapDec}(R)\|f\|_{2}\end{split}

There are three cases: If the first term on the right hand side dominates, we say we are in the cellular case; If the second term on the right hand side dominates, we say we are in the transverse case; If the third term on the right hand side dominates, we say we are in the tangent case.

3.2.1. Cellular case

Suppose that we are in the cellular case. Denote by 𝒪(1)\mathcal{O}^{(1)} the collection of the cells Oi(1)O_{i}^{(1)}, and by 𝕋i\mathbb{T}_{i} the collection of the tubes in 𝕋[BR]\mathbb{T}[B_{R}] intersecting the cell Oi(1)O_{i}^{(1)}. Notice that the cardinality of 𝒪(1)\mathcal{O}^{(1)} is comparable to D3D^{3}. Define

(3.19) fi:=τT𝕋ifτ,T.f_{i}:=\sum_{\tau}\sum_{T\in\mathbb{T}_{i}}f_{\tau,T}.

By Lemma 3.7 of [Gut16], for every xOi(1)x\in O_{i}^{(1)}, we have

(3.20) BrαEf(x)2Br2αEfi(x)+RapDec(R)fL2.\mathrm{Br}_{\alpha}Ef(x)\leq 2\mathrm{Br}_{2\alpha}Ef_{i}(x)+{\rm RapDec}(R)\|f\|_{L^{2}}.

Since we are in the cellular case, we have

(3.21) BrαEfL13/4(BR)13/4Oi(1)BrαEfL13/4(Oi(1))13/4Oi(1)Br2αEfiL13/4(Oi(1))13/4.\|\mathrm{Br}_{\alpha}Ef\|_{L^{13/4}(B_{R})}^{13/4}\lesssim\sum_{O_{i}^{(1)}}\|\mathrm{Br}_{\alpha}Ef\|_{L^{13/4}(O_{i}^{(1)})}^{13/4}\lesssim\sum_{O_{i}^{(1)}}\|\mathrm{Br}_{2\alpha}Ef_{i}\|_{L^{13/4}(O_{i}^{(1)})}^{13/4}.

Notice that by Lemma 3.2 of [Gut16], each tube T𝕋T\in\mathbb{T} can intersect at most D+1D+1 many cells in 𝒪(1)\mathcal{O}^{(1)}, thus

(3.22) ifi22Df22.\sum_{i}\|f_{i}\|_{2}^{2}\lesssim D\|f\|_{2}^{2}.

Since the cardinality of the cells 𝒪(1)\mathcal{O}^{(1)} is comparable to D3D^{3}, by pigeonholing, we can take a sub-collection of the cells so that the cardinality of which is comparable to D3D^{3} and for each Oi(1)O_{i}^{(1)} in this sub-collection we have

(3.23) fi22D2f22.\|f_{i}\|_{2}^{2}\lesssim D^{-2}\|f\|_{2}^{2}.

By abusing the notation, we still denote this sub-collection by 𝒪(1)\mathcal{O}^{(1)}. By (3.21) and (3.8), we have

(3.24) BrαEfL13/4(BR)13/4Oi(1)𝒪(1)Br2αEfiL13/4(Oi(1))13/4.\|\mathrm{Br}_{\alpha}Ef\|_{L^{13/4}(B_{R})}^{13/4}\lesssim\sum_{O_{i}^{(1)}\in\mathcal{O}^{(1)}}\|\mathrm{Br}_{2\alpha}Ef_{i}\|_{L^{13/4}(O_{i}^{(1)})}^{13/4}.

To indicate the relation between functions and cells, we write fOi(1):=fif_{O_{i}^{(1)}}:=f_{i} for any Oi(1)𝒪(1)O_{i}^{(1)}\in\mathcal{O}^{(1)}. For convenience, later we will drop the subscript ii to simply write fO(1)f_{O^{(1)}} for any O(1)𝒪(1)O^{(1)}\in\mathcal{O}^{(1)}.

3.2.2. Transverse case

Suppose that we are in the transverse case. We have

BRW|BrαEf(x)|13/4kIBkW(Br2αEfI,k,+)13/4.\begin{split}\int_{B_{R}\cap W}|\mathrm{Br}_{\alpha}Ef(x)|^{13/4}\lesssim\sum_{k}\sum_{I}\int_{B_{k}\cap W}\Big{(}\mathrm{Br}_{2\alpha}Ef_{I,k,+}\Big{)}^{13/4}.\end{split}

Choose the II that maximizes the right hand side, so we have

(3.25) BRW|BrαEf(x)|13/42K2kBkW(Br2αEfI,k,+)13/4.\begin{split}\int_{B_{R}\cap W}|\mathrm{Br}_{\alpha}Ef(x)|^{13/4}\lesssim 2^{K^{2}}\sum_{k}\int_{B_{k}\cap W}\Big{(}\mathrm{Br}_{2\alpha}Ef_{I,k,+}\Big{)}^{13/4}.\end{split}

We set 𝒪(1):={BkW}\mathcal{O}^{(1)}:=\{B_{k}\cap W\} to be the collection of these cells BkWB_{k}\cap W. For O(1)=BkW𝒪(1)O^{(1)}=B_{k}\cap W\in\mathcal{O}^{(1)}, we write fO(1):=fI,k,+f_{O^{(1)}}:=f_{I,k,+}. Note that fO(1)f_{O^{(1)}} is the sum of a subset of wave packets of ff, so

(3.26) fO(1)22f22.\|f_{O^{(1)}}\|_{2}^{2}\lesssim\|f\|_{2}^{2}.

By Lemma 5.7 of [Gut18], each tube belongs to at most O(D3)O(D^{3}) different sets 𝕋k,+\mathbb{T}_{k,+}, so we have

(3.27) O(1)𝒪(1)fO(1)22D3f22.\sum_{O^{(1)}\in\mathcal{O}^{(1)}}\|f_{O^{(1)}}\|_{2}^{2}\lesssim D^{3}\|f\|_{2}^{2}.

3.2.3. Tangent case

Suppose we are in the tangent case, then

(3.28) BRW|BrαEf(x)|13/4KO(1)kBkWBil(Efk,)13/4.\begin{split}\int_{B_{R}\cap W}|\mathrm{Br}_{\alpha}Ef(x)|^{13/4}&\lesssim K^{O(1)}\sum_{k}\int_{B_{k}\cap W}\mathrm{Bil}(Ef_{k,-})^{13/4}.\end{split}

We need the following estimate for the bilinear term.

Lemma 3.8 ((43) of [Gut16]).
(3.29) Bil(Efk,)L13/4(BkW)RO(δ)R152(τfτ,k,22)1/2.\begin{split}\|\mathrm{Bil}(Ef_{k,-})\|_{L^{13/4}(B_{k}\cap W)}\lesssim R^{O(\delta)}R^{\frac{1}{52}}\big{(}\sum_{\tau}\|f_{\tau,k,-}\|_{2}^{2}\big{)}^{1/2}.\end{split}

Choose the kk and τ\tau that maximizes the right hand side of (3.28) and apply Lemma 3.8, so we obtain

(3.30) BRW|BrαEf(x)|13/4RO(δ)R1/16fτ,k,213/4.\begin{split}\int_{B_{R}\cap W}|\mathrm{Br}_{\alpha}Ef(x)|^{13/4}&\lesssim R^{O(\delta)}R^{1/16}\|f_{\tau,k,-}\|_{2}^{13/4}.\end{split}

We set 𝒪(1)\mathcal{O}^{(1)} to consist only a single cell BkWB_{k}\cap W, and for O(1)=BkWO^{(1)}=B_{k}\cap W set fO(1):=fτ,k,f_{O^{(1)}}:=f_{\tau,k,-}. We also have

(3.31) fO(1)22f22,\|f_{O^{(1)}}\|_{2}^{2}\lesssim\|f\|_{2}^{2},

since fO(1)f_{O^{(1)}} is the sum of a subset of wave packets of ff.

3.3. Iteration

In the last subsection, we work with the function BrαEf{\rm{Br}}_{\alpha}Ef in the cell BRB_{R} and result in one the three cases. We will iterate this process in this subsection. We would like to state this iteration as an algorithm.

Algorithm 1 (Iteration of polynomial partitioning).

Suppose we begin with a function ff, a cell BRB_{R} and a small number α=Kϵ\alpha=K^{-\epsilon}. By iteratively doing the polynomial partitioning, we have the following output:

\bullet There exists an integer ss (1sϵO(1)1\leq s\leq\epsilon^{-O(1)}) which denotes the total number of iteration steps. There is a function STATE which we use to record the state of each step:

(3.32) STATE:{1,2,,s}{cell, trans, tang}.\textup{STATE}:\{1,2,\cdots,s\}\rightarrow\{\textup{cell,~{}trans,~{}tang}\}.

We stop at step ss because of one of the two reasons: either we first encounter the tangent state STATE(ss)=tang, or the radius at step ss is small rsRϵ/10r_{s}\sim R^{\epsilon/10} (rsr_{s} is as below).

\bullet At each step uu, u{1,,s}u\in\{1,\cdots,s\}, we have:

1. A radius rur_{u} which is defined recursively by

(3.33) ru={ru1/D,STATE(u)=cellru11δ,STATE(u)=trans.r_{u}=\left\{\begin{array}[]{cc}r_{u-1}/D,&\textup{STATE}(u)=\textup{cell}\\ r_{u-1}^{1-\delta},&\textup{STATE}(u)=\textup{trans}.\end{array}\right.

We also set r0=Rr_{0}=R. Here are another two important parameters defined by

(3.34) sc:=#{1is:STATE(i)=cell},\displaystyle s_{c}:=\#\{1\leq i\leq s:~{}\textup{STATE}(i)=\textup{cell}\},
(3.35) st:=#{1is:STATE(i)=trans}.\displaystyle s_{t}:=\#\{1\leq i\leq s:~{}\textup{STATE}(i)=\textup{trans}\}.

Note that we have

(3.36) sclogR/logD=ϵ6,\displaystyle s_{c}\lesssim\log R/\log D=\epsilon^{-6},
(3.37) stδ2=ϵ4.\displaystyle s_{t}\lesssim\delta^{-2}=\epsilon^{-4}.

2. A set of cells 𝒪(u)={O(u)}\mathcal{O}^{(u)}=\{O^{(u)}\} such that each O(u)O^{(u)} is contained in a rur_{u}-ball BO(u)B_{O^{(u)}}. Each O(u)O^{(u)} has a unique parent O(u1)𝒪(u1)O^{(u-1)}\in\mathcal{O}^{(u-1)}. We denote this relation by

(3.38) O(u)<O(u1).O^{(u)}<O^{(u-1)}.

Moreover we have the nested property for these cells. That is, for any cell Os𝒪sO_{s}\in\mathcal{O}_{s}, there exist unique O(u)𝒪(u)O^{(u)}\in\mathcal{O}^{(u)} (u=1,,s)(u=1,\cdots,s) such that

(3.39) Os<Os1<<O1(<BR).O_{s}<O_{s-1}<\cdots<O_{1}(<B_{R}).

3. A set of functions {fO(u)}O(u)𝒪(u)\{f_{O^{(u)}}\}_{O^{(u)}\in\mathcal{O}^{(u)}}.

4. There are three possible cases for each step uu: cellular case, transverse case and tangent case. The outputs for each case are the following:

Cellular case: If STATE(u)=\textup{STATE}(u)= cell, we have the following estimates:

(3.40) fO(u)22D2fO(u1)22,for O(u)<O(u1).\|f_{O^{(u)}}\|_{2}^{2}\lesssim D^{-2}\|f_{O^{(u-1)}}\|_{2}^{2},\ \ \textup{for~{}}O^{(u)}<O^{(u-1)}.
(3.41) O(u)fO(u)22DO(u1)fO(u1)22.\sum_{O^{(u)}}\|f_{O^{(u)}}\|_{2}^{2}\lesssim D\sum_{O^{(u-1)}}\|f_{O^{(u-1)}}\|_{2}^{2}.
(3.42) O(u1)Br2u1αEfO(u1)L13/4(O(u1))13/4O(u)Br2uαEfO(u)L13/4(O(u))13/4.\sum_{O^{(u-1)}}\|{\rm{Br}}_{2^{u-1}\alpha}Ef_{O^{(u-1)}}\|_{L^{13/4}(O^{(u-1)})}^{13/4}\lesssim\sum_{O^{(u)}}\|{\rm{Br}}_{2^{u}\alpha}Ef_{O^{(u)}}\|_{L^{13/4}(O^{(u)})}^{13/4}.

Transverse case: If STATE(u)=\textup{STATE}(u)= trans, we have the following estimates:

(3.43) fO(u)22fO(u1)22,for O(u)<O(u1).\|f_{O^{(u)}}\|_{2}^{2}\lesssim\|f_{O^{(u-1)}}\|_{2}^{2},\ \ \textup{for~{}}O^{(u)}<O^{(u-1)}.
(3.44) O(u)fO(u)22D3O(u1)fO(u1)22.\sum_{O^{(u)}}\|f_{O^{(u)}}\|_{2}^{2}\lesssim D^{3}\sum_{O^{(u-1)}}\|f_{O^{(u-1)}}\|_{2}^{2}.
(3.45) O(u1)Br2u1αEfO(u1)L13/4(O(u1))13/42K2O(u)Br2uαEfO(u)L13/4(O(u))13/4.\sum_{O^{(u-1)}}\|{\rm{Br}}_{2^{u-1}\alpha}Ef_{O^{(u-1)}}\|_{L^{13/4}(O^{(u-1)})}^{13/4}\lesssim 2^{K^{2}}\sum_{O^{(u)}}\|{\rm{Br}}_{2^{u}\alpha}Ef_{O^{(u)}}\|_{L^{13/4}(O^{(u)})}^{13/4}.

Tangent case: If STATE(u)=\textup{STATE}(u)= tang, so u=su=s, then we have the following estimates:

(3.46) fO(u)22fO(u1)22,for O(u)<O(u1).\|f_{O^{(u)}}\|_{2}^{2}\lesssim\|f_{O^{(u-1)}}\|_{2}^{2},\ \ \textup{for~{}}O^{(u)}<O^{(u-1)}.
(3.47) O(u)fO(u)22O(u1)fO(u1)22.\sum_{O^{(u)}}\|f_{O^{(u)}}\|_{2}^{2}\lesssim\sum_{O^{(u-1)}}\|f_{O^{(u-1)}}\|_{2}^{2}.
(3.48) O(u1)Br2u1αEfO(u1)L13/4(O(u1))13/4KO(1)O(u)Br2uαEfO(u)L13/4(O(u))13/4.\sum_{O^{(u-1)}}\|{\rm{Br}}_{2^{u-1}\alpha}Ef_{O^{(u-1)}}\|_{L^{13/4}(O^{(u-1)})}^{13/4}\lesssim K^{O(1)}\sum_{O^{(u)}}\|{\rm{Br}}_{2^{u}\alpha}Ef_{O^{(u)}}\|_{L^{13/4}(O^{(u)})}^{13/4}.
(3.49) Br2uαEfO(u)L13/4(O(u))13/4rs11/16+δO(u)fO(u)213/4.\|{\rm{Br}}_{2^{u}\alpha}Ef_{O^{(u)}}\|_{L^{13/4}(O^{(u)})}^{13/4}\lesssim r_{s-1}^{1/16+\delta}\sum_{O^{(u)}}\|f_{O^{(u)}}\|_{2}^{13/4}.
Proof.

The proof follows from the same idea as in the last subsection. Note that in the last subsection we have dealt with u=1u=1. We briefly explain how to pass from step u1u-1 to step uu.

Suppose we already obtained cells 𝒪(u1)={O(u1)}\mathcal{O}^{(u-1)}=\{O^{(u-1)}\} and functions

{fO(u1)}O(u1)𝒪(u1),\{f_{O^{(u-1)}}\}_{O^{(u-1)}\in\mathcal{O}^{(u-1)}},

and note that each cell O(u1)O^{(u-1)} is contained in a ball of radius ru1r_{u-1}. Now, we fix a cell O(u1)𝒪(u1)O^{(u-1)}\in\mathcal{O}^{(u-1)}. We apply the same argument as in the last subsection, with RR replaced by ru1r_{u-1}, BRB_{R} replaced by O(u1)O^{(u-1)}, ff replaced by fO(u1)f_{O^{(u-1)}}. We just give a sketch here.

As in the Subsection 3.1, we still use K1K^{-1}-squares {τ}\{\tau\} to cover [1,1]2[-1,1]^{2} and define the broad function BrαEfO(u1){\rm{Br}}_{\alpha}Ef_{O^{(u-1)}}. Apply polynomial partitioning to the function |1O(u1)Br2(u1)EfO(u1)|13/4|1_{O^{(u-1)}}{\rm{Br}}_{2^{(u-1)}}Ef_{O^{(u-1)}}|^{13/4}, then we obtain D3\sim D^{3} cells 𝒪(u)(O(u1))={Oi(u)}\mathcal{O}^{(u)}(O^{(u-1)})=\{O_{i}^{(u)}\} and a wall denoted by WO(u1)W_{O^{(u-1)}} such that an inequality similar to (3.10) holds:

(3.50) Br2u1αEfO(u1)L13/4(O(u1))13/4Oi(u)Br2uαEfL13/4(Oi(u))13/4+Br2uαEfL13/4(WO(u1))13/4.\|{\rm{Br}}_{2^{u-1}\alpha}Ef_{O^{(u-1)}}\|_{L^{13/4}(O^{(u-1)})}^{13/4}\lesssim\sum_{O_{i}^{(u)}}\|\mathrm{Br}_{2^{u}\alpha}Ef\|_{L^{13/4}(O_{i}^{(u)})}^{13/4}+\|\mathrm{Br}_{2^{u}\alpha}Ef\|_{L^{13/4}(W_{O^{(u-1)}})}^{13/4}.

To deal with the integral over wall, we cover WO(u1)W_{O^{(u-1)}} by balls BkB_{k} of radius ru11δr_{u-1}^{1-\delta}. For each BkB_{k}, similar to Definition 3.5 and 3.6, we can define the tangent tubes 𝕋k,\mathbb{T}_{k,-} and transverse tubes 𝕋k,+\mathbb{T}_{k,+} associated to BkB_{k}, but the tubes here are of dimensions ru11/2×ru11/2×ru1r_{u-1}^{1/2}\times r_{u-1}^{1/2}\times r_{u-1}. We also remark that we do wave packet decomposition for fO(1),τf_{O^{(1)},\tau} at scale ru1r_{u-1}. Similarly we can define the bilinear function Bil(EfO(u1),k,){\rm{Bil}}(Ef_{O^{(u-1)},k,-}) as in (3.16).

We also have the following lemma which is similar to Lemma 3.7:

Lemma 3.9.

If xBkWO(u1)x\in B_{k}\cap W_{O^{(u-1)}} and α=Kϵ\alpha=K^{-\epsilon}, then

(3.51) Br2u1α|EfO(u1)(x)|IBr2uαEfO(u1),I,k,+(x)+K100Bil(EfO(u1),k,)(x)+RapDec(R)f2.\begin{split}{\rm{Br}}_{2^{u-1}}\alpha|Ef_{O^{(u-1)}}(x)|\lesssim\sum_{I}{\rm{Br}}_{2^{u}\alpha}Ef_{O^{(u-1)},I,k,+}(x)+K^{100}{\rm{Bil}}(Ef_{O^{(u-1)},k,-})(x)\\ +{\rm RapDec}(R)\|f\|_{2}.\end{split}

Now we plug (3.51) into (3.50) to obtain:

(3.52) Br2u1αEfO(u1)L13/4(O(u1))13/4Oi(u)Br2uαEfO(u1)L13/4(Oi(u))13/4+kIBr2uαEfO(u1),I,k,+L13/4(BkWO(u1))13/4+K100Bil(EfO(u1),j,)L13/4(WO(u1))13/4+RapDec(R)f2.\begin{split}\|{\rm{Br}}_{2^{u-1}\alpha}Ef_{O^{(u-1)}}\|_{L^{13/4}(O^{(u-1)})}^{13/4}&\lesssim\sum_{O_{i}^{(u)}}\|{\rm{Br}}_{2^{u}\alpha}Ef_{O^{(u-1)}}\|^{13/4}_{L^{13/4}(O_{i}^{(u)})}\\ &+\sum_{k}\sum_{I}\|{\rm{Br}}_{2^{u}\alpha}Ef_{O^{(u-1)},I,k,+}\|^{13/4}_{L^{13/4}(B_{k}\cap W_{O^{(u-1)}})}\\ &+K^{100}\|{\rm{Bil}}(Ef_{O^{(u-1)},j,-})\|_{L^{13/4}(W_{O^{(u-1)}})}^{13/4}\\ &+{\rm RapDec}(R)\|f\|_{2}.\end{split}

Summing over O(u1)𝒪(u1)O^{(u-1)}\in\mathcal{O}^{(u-1)}, we obtain

(3.53) O(u1)Br2u1αEfO(u1)L13/4(O(u1))13/4O(u1)Oi(u)𝒪u(O(u1))Br2uαEfO(u1)L13/4(Oi(u))13/4+O(u1)kIBr2uαEfO(u1),I,k,+L13/4(BkWO(u1))13/4+K100O(u1)Bil(EfO(u1),j,)L13/4(WO(u1))13/4+RapDec(R)f2\begin{split}\sum_{O^{(u-1)}}\|{\rm{Br}}_{2^{u-1}\alpha}Ef_{O^{(u-1)}}\|_{L^{13/4}(O^{(u-1)})}^{13/4}&\lesssim\sum_{O^{(u-1)}}\sum_{O_{i}^{(u)}\in\mathcal{O}^{u}(O^{(u-1)})}\|{\rm{Br}}_{2^{u}\alpha}Ef_{O^{(u-1)}}\|^{13/4}_{L^{13/4}(O_{i}^{(u)})}\\ &+\sum_{O^{(u-1)}}\sum_{k}\sum_{I}\|{\rm{Br}}_{2^{u}\alpha}Ef_{O^{(u-1)},I,k,+}\|^{13/4}_{L^{13/4}(B_{k}\cap W_{O^{(u-1)}})}\\ &+K^{100}\sum_{O^{(u-1)}}\|{\rm{Bil}}(Ef_{O^{(u-1)},j,-})\|_{L^{13/4}(W_{O^{(u-1)}})}^{13/4}\\ &+{\rm RapDec}(R)\|f\|_{2}\end{split}

There are three cases: If the first term on the right hand side dominates, we set STATE(u)(u)=cell; If the second term on the right hand side dominates, we set STATE(u)(u)=trans; If the third term on the right hand side dominates, we set STATE(u)(u)=tang.

When STATE(u)(u)=cell, we can find cells 𝒪(u)={O(u)}\mathcal{O}^{(u)}=\{O^{(u)}\} which is a subset of O(u1)𝒪(u1)𝒪(u)(O(u1))\bigcup_{O^{(u-1)}\in\mathcal{O}^{(u-1)}}\mathcal{O}^{(u)}(O^{(u-1)}), and functions {fO(u)}\{f_{O^{(u)}}\}, so that similar to (3.23), (3.22) and (3.24), we have (3.40), (3.41) and (3.42).

When STATE(u)(u)=trans, we remark that (3.43), (3.44) and (3.45) are analogues of (3.26), (3.27) and (3.25).

When STATE(u)(u)=tang, we remark that (3.46) is an analogue of (3.31). (3.47) holds because each O(u1)O^{(u-1)} has only one child: O(u)<O(u1)O^{(u)}<O^{(u-1)} in the tangent case. (3.48) is an analogue of (3.28). Also, (3.49) is an analogue of (3.30)

We also remark that the relations “<<” between cells in 𝒪(u)\mathcal{O}^{(u)} and cells in 𝒪(u1)\mathcal{O}^{(u-1)} are apparently defined according to the polynomial partitioning process: we say O(u)<O(u1)O^{(u)}<O^{(u-1)} if O(u)O^{(u)} is obtained by doing polynomial partitioning to O(u1)O^{(u-1)}.

We finally remark that the subscript in Br2uα{\rm{Br}}_{2^{u}\alpha} is always less than 11, since 2sα2O(ϵ6)Kϵ<12^{s}\alpha\lesssim 2^{O(\epsilon^{-6})}K^{-\epsilon}<1, so the broad functions are always well-defined here. ∎

3.4. Put things together

As stated in Algorithm 1, the iteration stops due to one of the two reasons: STATE(ss)=tang; rsRϵ/10r_{s}\sim R^{\epsilon/10}.

Let us consider the case that the iteration stops due to the smallness of the radius rsr_{s}. Iterating (3.42) and (3.45), we have

(3.54) BrαEfL13/4(BR)13/4O(s)Br2sαEfO(s)L13/4(O(s))13/4.\|\mathrm{Br}_{\alpha}Ef\|_{L^{13/4}(B_{R})}^{13/4}\lesssim\sum_{O^{(s)}}\|{\rm{Br}}_{2^{s}\alpha}Ef_{O^{(s)}}\|_{L^{13/4}(O^{(s)})}^{13/4}.

Also note that the diameter of O(s)O^{(s)} is rsRϵ/10\lesssim r_{s}\sim R^{\epsilon/10}, so

(3.55) BrαEfL13/4(BR)13/4O(s)EfO(s)L13/4(O(s))13/4Rϵ/3O(s)fO(s)213/4Rϵ/3O(s)fO(s)22(maxO(s)fO(s)25/4).\begin{split}\|\mathrm{Br}_{\alpha}Ef\|_{L^{13/4}(B_{R})}^{13/4}\lesssim\sum_{O^{(s)}}\|Ef_{O^{(s)}}\|_{L^{13/4}(O^{(s)})}^{13/4}\lesssim R^{\epsilon/3}\sum_{O^{(s)}}\|f_{O^{(s)}}\|_{2}^{13/4}\\ \lesssim R^{\epsilon/3}\sum_{O^{(s)}}\|f_{O^{(s)}}\|_{2}^{2}\,\big{(}\max_{O^{(s)}}\|f_{O^{(s)}}\|_{2}^{5/4}\big{)}.\end{split}

By iterating (3.41) and (3.44) and noting D3stRO(ϵ2)D^{3s_{t}}\lesssim R^{O(\epsilon^{2})}, we have

(3.56) O(s)fO(s)22RO(ϵ2)Dscf22.\sum_{O^{(s)}}\|f_{O^{(s)}}\|_{2}^{2}\lesssim R^{O(\epsilon^{2})}D^{s_{c}}\|f\|_{2}^{2}.

By iterating (3.40) and (3.43), we have the inequality fO(s)22D2scf22\|f_{O^{(s)}}\|_{2}^{2}\lesssim D^{-2s_{c}}\|f\|_{2}^{2} for each O(s)𝒪(s)O^{(s)}\in\mathcal{O}^{(s)}. Since we assumed the wave packet of ff is contained in B10RB_{10R}, we have f^\widehat{f} is essentially supported in [10R,10R]2[-10R,10R]^{2}. By Hölder’s inequality, we have f2=f^2Rf^\|f\|_{2}=\|\widehat{f}\|_{2}\lesssim R\|\widehat{f}\|_{\infty}, so

(3.57) fO(s)22D2scR2f^2.\|f_{O^{(s)}}\|_{2}^{2}\lesssim D^{-2s_{c}}R^{2}\|\widehat{f}\|_{\infty}^{2}.

Plugging (3.56) and (3.57) into (3.55), we obtain

(3.58) BrαEfL13/4(BR)13/4Rϵ/2Dsc/2R5/4f22f^5/4Rϵ/2R5/4f22f^5/4,\begin{split}\|\mathrm{Br}_{\alpha}Ef\|_{L^{13/4}(B_{R})}^{13/4}\lesssim R^{\epsilon/2}D^{-s_{c}/2}R^{5/4}\|f\|_{2}^{2}\|\widehat{f}\|_{\infty}^{5/4}\leq R^{\epsilon/2}R^{5/4}\|f\|_{2}^{2}\|\widehat{f}\|_{\infty}^{5/4},\end{split}

which is (3.2).

Next, we consider the case that STATE(s)(s)=tang. Similar to (3.55), we have

(3.59) BrαEfL13/4(BR)13/4RO(ϵ2)O(s)EfO(s)L13/4(O(s))13/4.\|\mathrm{Br}_{\alpha}Ef\|_{L^{13/4}(B_{R})}^{13/4}\lesssim R^{O(\epsilon^{2})}\sum_{O^{(s)}}\|Ef_{O^{(s)}}\|_{L^{13/4}(O^{(s)})}^{13/4}.

By (3.49), we have

(3.60) BrαEfL13/4(BR)13/4RO(ϵ)rs1/16O(s)fO(s)213/4.\begin{split}\|\mathrm{Br}_{\alpha}Ef\|_{L^{13/4}(B_{R})}^{13/4}&\lesssim R^{O(\epsilon)}r_{s}^{1/16}\sum_{O^{(s)}}\|f_{O^{(s)}}\|_{2}^{13/4}.\end{split}

We also have (3.56) and (3.57). But here in the tangent case, we need a new estimate of fO(s)2\|f_{O^{(s)}}\|_{2} other than (3.57). We will obtain it using polynomial Wolff. Fix a cell O(s)O^{(s)}. Let us first recall how we obtain fO(s)f_{O^{(s)}} from ff. We have the chain that indicates the relations between cells at different scales:

Os<Os1<<O1<BR(=O0).O_{s}<O_{s-1}<\cdots<O_{1}<B_{R}(=O_{0}).

For 0us0\leq u\leq s, let BruB_{r_{u}} be the ball of radius rur_{u} that contains OuO_{u}. fO(u)f_{O^{(u)}} is obtained as follows. We do wave packet decomposition for fO(u1)f_{O^{(u-1)}} in Bru1B_{r_{u-1}} and choose a subset of wave packets of fO(u1)f_{O^{(u-1)}} that are related to the cell O(u)O^{(u)}. Then we define fO(u)f_{O^{(u)}} to be the sum of these chosen wave packets. We may denote by 𝕋O(u)\mathbb{T}_{O^{(u)}} the set of these wave packets and write the relation as

(3.61) fO(u)=T𝕋O(u)(fO(u1))T.f_{O^{(u)}}=\sum_{T\in\mathbb{T}_{O^{(u)}}}(f_{O^{(u-1)}})_{T}.

To compare wave packets at different scale, we need the following definition.

Definition 3.10.

For two tubes T=Tθ,v𝕋[Br(x)]T=T_{\theta,v}\in\mathbb{T}[B_{r}(x)] and T=Tθ,v𝕋[Br(x)]T^{\prime}=T^{\prime}_{\theta^{\prime},v^{\prime}}\in\mathbb{T}[B_{r^{\prime}}(x^{\prime})] with rrr^{\prime}\leq r. We say T<TT^{\prime}<T if

(3.62) dist(θ,θ)r1/2 and dist(T,T)r1/2+δ.\mathrm{dist}(\theta,\theta^{\prime})\lesssim r^{\prime-1/2}\;\;\text{ and }\;\;\mathrm{dist}(T,T^{\prime})\lesssim r^{1/2+\delta}.

Now we define the tube set

(3.63) 𝕋O(u1),O(u):={T𝕋O(u1):T<T for some T𝕋O(u)},\mathbb{T}_{O^{(u-1)},O^{(u)}}:=\{T\in\mathbb{T}_{O^{(u-1)}}:T^{\prime}<T\textup{~{}for~{}some~{}}T^{\prime}\in\mathbb{T}_{O^{(u)}}\},

and the function

fO(u1),O(u):=T𝕋O(u1),O(u)(fO(u1))T.f_{O^{(u-1)},O^{(u)}}:=\sum_{T\in\mathbb{T}_{O^{(u-1)},O^{(u)}}}(f_{O^{(u-1)}})_{T}.

Since fO(u1)f_{O^{(u-1)}} is concentrated in wave packets from 𝕋O(u1)\mathbb{T}_{O^{(u-1)}}, we have

fO(u1)=T𝕋O(u1)(fO(u1))T+RapDec(R)f2.f_{O^{(u-1)}}=\sum_{T\in\mathbb{T}_{O^{(u-1)}}}(f_{O^{(u-1)}})_{T}+{\rm RapDec}(R)\|f\|_{2}.

Plugging into (3.61) and using Lemma 2.3, we have

(3.64) fO(u)=T𝕋O(u)(fO(u1),O(u))T+RapDec(R)f2.f_{O^{(u)}}=\sum_{T\in\mathbb{T}_{O^{(u)}}}(f_{O^{(u-1)},O^{(u)}})_{T}+{\rm RapDec}(R)\|f\|_{2}.

Motivated by the definition (3.63), we now define

(3.65) 𝕋O(s):={T𝕋BR:T<T for some T𝕋O(s)},\mathbb{T}_{O^{(s)}}^{\sharp}:=\{T\in\mathbb{T}_{B_{R}}:T^{\prime}<T\textup{~{}for~{}some~{}}T^{\prime}\in\mathbb{T}_{O^{(s)}}\},

and set

(3.66) fO(s):=T𝕋O(s)fT.f_{{O^{(s)}}}^{\sharp}:=\sum_{T\in\mathbb{T}_{O^{(s)}}^{\sharp}}f_{T}.

Then we have

(3.67) fO(s)=T𝕋O(u)(fO(s))T+RapDec(R)f2.f_{O^{(s)}}=\sum_{T\in\mathbb{T}_{O^{(u)}}}(f_{{O^{(s)}}}^{\sharp})_{T}+{\rm RapDec}(R)\|f\|_{2}.

Before proving an upper bound for fO(s)2\|f_{O^{(s)}}\|_{2}, we state the polynomial Wolff estimate in 3\mathbb{R}^{3}. For its proof, we refer to [Wu20] (or Lemma 11.1 in [GJW21]).

Lemma 3.11 (Three dimensional polynomial Wolff).

Fix R>r>Rϵ/10>1R>r>R^{\epsilon/10}>1. Set δ=ϵ2\delta=\epsilon^{2}. Let SS be a two-dimensional algebraic variety of complexity at most E=O(Rϵ6)E=O(R^{\epsilon^{6}}). Let BrB_{r} be a ball contained in B10RB_{10R}. We define

(3.68) S~:=B10RT a r×r1/2+δ tubeTN2r1/2+δ(SBr)FatRr(T).\widetilde{S}:=B_{10R}\cap\bigcup_{\begin{subarray}{c}T\textup{~{}a~{}}r\times r^{1/2+\delta}\textup{~{}tube}\\ T\subset N_{2r^{1/2+\delta}}(S\cap B_{r})\end{subarray}}\textup{Fat}_{\frac{R}{r}}(T).

Here, FatRrT=RrT\textup{Fat}_{\frac{R}{r}}T=\frac{R}{r}T is the dilation of TT. We also define

S~:=T a r×r1/2+δ tube(T,e3)<1/10TS~T.\widetilde{S}^{\prime}:=\bigcup_{\begin{subarray}{c}T\textup{~{}a~{}}r\times r^{1/2+\delta}\textup{~{}tube}\\ \angle(T,e_{3})<1/10\\ T\subset\widetilde{S}\end{subarray}}T.

Then for every a[2R,2R]a\in[-2R,2R], we have

(3.69) |S~(2×{a})|C(ϵ)RϵR2r1/2.|\widetilde{S}^{\prime}\cap(\mathbb{R}^{2}\times\{a\})|\leq C(\epsilon)R^{\epsilon}R^{2}r^{-1/2}.

We are ready to prove the following estimate

Lemma 3.12.

For any O(s)𝒪(s)O^{(s)}\in\mathcal{O}^{(s)}, we have

(3.70) fO(s)22Rϵ2R2rs1/2f^2\|f_{O^{(s)}}\|_{2}^{2}\lesssim R^{\epsilon^{2}}R^{2}r_{s}^{-1/2}\|\widehat{f}\|_{\infty}^{2}
Proof.

From (3.67), we have by L2L^{2}-orthogonality that fO(s)2fO(s)2.\|f_{O^{(s)}}\|_{2}\lesssim\|f_{O^{(s)}}^{\sharp}\|_{2}. Set

X:=(T𝕋O(s)10T)2×{0}.X:=\big{(}\bigcup_{T\in\mathbb{T}^{\sharp}_{O^{(s)}}}10T\big{)}\cap\mathbb{R}^{2}\times\{0\}.

We claim that

(3.71) fO(s)21Xf^2.\|f^{\sharp}_{O^{(s)}}\|_{2}\lesssim\|1_{X}\widehat{f}\|_{2}.

If it holds, then we use polynomial Wolff estimate that |X|Rϵ2R2rs1/2|X|\lesssim R^{\epsilon^{2}}R^{2}r_{s}^{-1/2}, and together with Hölder’s inequality, so that we obtain

(3.72) fO(s)221Xf^22|X|f^2Rϵ2R2rs1/2f^2.\|f_{O^{(s)}}\|_{2}^{2}\lesssim\|1_{X}\widehat{f}\|_{2}^{2}\lesssim|X|\|\widehat{f}\|_{\infty}^{2}\lesssim R^{\epsilon^{2}}R^{2}r_{s}^{-1/2}\|\widehat{f}\|_{\infty}^{2}.

Now, we prove (3.71). From the notation (2.12), we can rewrite (3.66) as

(3.73) fO(s)=Tθ,v𝕋O(s)ψ~θ(ηv(ψθf)).f_{{O^{(s)}}}^{\sharp}=\sum_{T_{\theta,v}\in\mathbb{T}_{O^{(s)}}^{\sharp}}\tilde{\psi}_{\theta}\Big{(}\eta_{v}^{\vee}*\big{(}\psi_{\theta}f\big{)}\Big{)}.

So,

(3.74) fO(s)^=Tθ,v𝕋O(s)ψ~θ^(ηv(ψ^θf^))=Tθ,v𝕋O(s)ψ~θ^(ηv(ψ^θ(f^1X)))+Tθ,v𝕋O(s)ψ~θ^(ηv(ψ^θ(f^1Xc))).\begin{split}\widehat{f_{{O^{(s)}}}^{\sharp}}&=\sum_{T_{\theta,v}\in\mathbb{T}_{O^{(s)}}^{\sharp}}\widehat{\tilde{\psi}_{\theta}}*\Big{(}\eta_{v}\big{(}\widehat{\psi}_{\theta}*\widehat{f}\big{)}\Big{)}\\ &=\sum_{T_{\theta,v}\in\mathbb{T}_{O^{(s)}}^{\sharp}}\widehat{\tilde{\psi}_{\theta}}*\Big{(}\eta_{v}\big{(}\widehat{\psi}_{\theta}*(\widehat{f}1_{X})\big{)}\Big{)}+\sum_{T_{\theta,v}\in\mathbb{T}_{O^{(s)}}^{\sharp}}\widehat{\tilde{\psi}_{\theta}}*\Big{(}\eta_{v}\big{(}\widehat{\psi}_{\theta}*(\widehat{f}1_{X^{c}})\big{)}\Big{)}.\end{split}

We show that the second term on the right hand side is negligible. Note that ηv\eta_{v} is supported in Br1+δ2(v)B_{r^{\frac{1+\delta}{2}}}(v), and ψ^θ\widehat{\psi}_{\theta} is essentially supported in Br1/2(0)B_{r^{1/2}}(0), so ηv(ψ^θ(f^1Xc))\eta_{v}\big{(}\widehat{\psi}_{\theta}*(\widehat{f}1_{X^{c}})\big{)} is essentially supported in Br1+δ2(v)(Br1/2(0)+Xc)B_{r^{\frac{1+\delta}{2}}}(v)\cap(B_{r^{1/2}}(0)+X^{c}). By the definition of XX, we see that if Tθ,v𝕋O(s)T_{\theta,v}\in\mathbb{T}_{O^{(s)}}^{\sharp}, then 10Br1+δ2(v)Xc=10B_{r^{\frac{1+\delta}{2}}}(v)\cap X^{c}=\emptyset. As a result, we proved that ηv(ψ^θ(f^1Xc))\eta_{v}\big{(}\widehat{\psi}_{\theta}*(\widehat{f}1_{X^{c}})\big{)} is negligible. Ignoring the negligible term, we have

(3.75) fO(s)22=fO(s)^22=Tθ,v𝕋O(s)ψ~θ^(ηv(ψ^θ(f^1X)))22f^1X22,\|f_{{O^{(s)}}}^{\sharp}\|_{2}^{2}=\|\widehat{f_{{O^{(s)}}}^{\sharp}}\|_{2}^{2}=\bigg{\|}\sum_{T_{\theta,v}\in\mathbb{T}_{O^{(s)}}^{\sharp}}\widehat{\tilde{\psi}_{\theta}}*\Big{(}\eta_{v}\big{(}\widehat{\psi}_{\theta}*(\widehat{f}1_{X})\big{)}\Big{)}\bigg{\|}_{2}^{2}\lesssim\|\widehat{f}1_{X}\|_{2}^{2},

which finishes the proof. ∎

Let us return to (3.60). We have

(3.76) BrαEfL13/4(BR)13/4RO(ϵ)rs1/16O(s)fO(s)22maxO(s)fO(s)25/4(by (3.56))RO(ϵ)rs1/16Dscf22maxO(s)fO(s)25/4\begin{split}\|{\rm{Br}}_{\alpha}Ef\|_{L^{13/4}(B_{R})}^{13/4}&\lesssim R^{O(\epsilon)}r_{s}^{1/16}\sum_{O^{(s)}}\|f_{O^{(s)}}\|_{2}^{2}\max_{O^{(s)}}\|f_{O^{(s)}}\|_{2}^{5/4}\\ (\textup{by~{}}\eqref{est1})&\lesssim R^{O(\epsilon)}r_{s}^{1/16}D^{s_{c}}\|f\|_{2}^{2}\max_{O^{(s)}}\|f_{O^{(s)}}\|_{2}^{5/4}\end{split}

To estimate maxO(s)fO(s)25/4\max_{O^{(s)}}\|f_{O^{(s)}}\|_{2}^{5/4}, we combine 1/21/2 power of (3.57) and 1/81/8 power of (3.70) to obtain

maxO(s)fO(s)25/4DscR5/4rs1/16f^2.\max_{O^{(s)}}\|f_{O^{(s)}}\|_{2}^{5/4}\lesssim D^{-s_{c}}R^{5/4}r_{s}^{-1/16}\|\widehat{f}\|_{\infty}^{2}.

Plugging into (3.76), we obtain (3.2).

4. A proof for the case α<1\alpha<1

In this section, we prove Theorem 2.1 for the case: ξξh\nabla_{\xi\xi}h has eigenvalues of different signs and n=3n=3. The strategy is very similar to that for α=2\alpha=2. The main difference is, for the case α<1\alpha<1, the surface {(ξ,h(ξ)):ξ[1,1]2}\{(\xi,h(\xi)):\xi\in[-1,1]^{2}\} has negative Gaussian curvature. When it comes to the restriction problem, the obstacles from this difference is well understood in [GO20]. We will simply combine the ideas therein with the ideas in Section 3. Since the main idea is already addressed in Section 3, we give only a sketch of the proof here.

We first make some reductions. As in the previous section, by discarding the contribution outside BRB_{R}, we may assume that the wave packets of ff is contained in B10RB_{10R}. Moreover, it suffices to prove (2.8) for polynomials hh of the form

(4.1) h(ξ1,ξ2):=ξ1ξ2+a2,0ξ12+a2,2ξ22+i=3dj=0iai,jξ1ijξ2jh(\xi_{1},\xi_{2}):=\xi_{1}\xi_{2}+a_{2,0}\xi_{1}^{2}+a_{2,2}\xi_{2}^{2}+\sum_{i=3}^{d}\sum_{j=0}^{i}a_{i,j}\xi_{1}^{i-j}\xi_{2}^{j}

satisfying the condition

(4.2) |a2,0|+|a2,2|+100di=3dj=0i|ai,j|ϵ0.|a_{2,0}|+|a_{2,2}|+100^{d}\sum_{i=3}^{d}\sum_{j=0}^{i}|a_{i,j}|\leq\epsilon_{0}.

Here, ϵ0\epsilon_{0} is a small number. Once we prove (2.8) for such polynomials, the general case follows by a simple application of Taylor’s theorem (see Section 2 of [GO20] for the details). Hence, we may assume that hh satisfies the condition (4.2) and aim to prove

(4.3) EfL13/4(BR)CϵRϵR(1813)f^L13/4.\|Ef\|_{L^{13/4}(B_{R})}\leq C_{\epsilon}R^{\epsilon}R^{(1-\frac{8}{13})}\|\widehat{f}\|_{L^{13/4}}.

4.1. Broad-narrow reduction

We do broad-narrow reduction in this subsection. Since the surface has negative curvature, the broad-narrow reduction is more involved than that for the surface with positive curvature as we did in Section 3. We will follow the idea from [GO20].

4.1.1. Bad lines

Let c1(KL1)c_{1}(K_{L}^{-1}) be a small number given by

(4.4) c1(KL1):=1010dϵ0KL1.c_{1}(K_{L}^{-1}):=10^{-10d}\epsilon_{0}K_{L}^{-1}.

Take a collection 𝔻\mathbb{D} of points on the unit circle S1S^{1} such that 𝔻\mathbb{D} is a maximal c1(KL1)c_{1}(K_{L}^{-1})-separated set. For every ac1(KL1)[10,10]a\in c_{1}(K_{L}^{-1})\mathbb{Z}\cap[-10,10] and v=(v1,v2)𝔻v=(v_{1},v_{2})\in\mathbb{D} with |v2||v1||v_{2}|\leq|v_{1}|, let l1,a,vl_{1,a,v} denote the line passing through the point (0,a)(0,a) with direction vector vv. Similarly, for every ac1(KL1)[10,10]a\in c_{1}(K_{L}^{-1})\mathbb{Z}\cap[-10,10] and v=(v1,v2)𝔻v=(v_{1},v_{2})\in\mathbb{D} with |v2|>|v1||v_{2}|>|v_{1}|, let l2,a,vl_{2,a,v} denote the line passing through (a,0)(a,0) with direction vector vv.

Let us now define bad lines. Suppose that a line l=l1,a,vl=l_{1,a,v} is given. We define an affine transformation MlM_{l} associated to the line ll in the following way. First, let Ml,1M_{l,1} be the action of translation that sends (0,a)(0,a) to the origin and let Ml,2M_{l,2} be the rotation mapping vv to the point (1,0)(1,0). Define

(4.5) Ml:=Ml,2Ml,1.M_{l}:=M_{l,2}\circ M_{l,1}.

We write the polynomial (hMl1)(ξ1,ξ2)(h\circ M_{l}^{-1})(\xi_{1},\xi_{2}) as

(4.6) (hMl1)(ξ1,ξ2)=i=0dj=0ici,jξ1ijξ2j.(h\circ M_{l}^{-1})(\xi_{1},\xi_{2})=\sum_{i=0}^{d}\sum_{j=0}^{i}c_{i,j}\xi_{1}^{i-j}\xi_{2}^{j}.

By the assumption (4.2), we see that |c2,1|1001|c_{2,1}|\geq 100^{-1}. The line l1,a,vl_{1,a,v} is called a bad line provided that

(4.7) maxi=2,,d(|ci,0|)105dϵ0KL1=105dc1(KL1).\max_{i=2,\ldots,d}(|c_{i,0}|)\leq 10^{-5d}\epsilon_{0}K_{L}^{-1}=10^{5d}c_{1}(K_{L}^{-1}).

We define a bad line for l2,a,vl_{2,a,v} in a similar way, with the role of ξ1\xi_{1} and ξ2\xi_{2} in (4.6) and (4.7) exchanged. Take ι{1,2}\iota\in\{1,2\}. Let Lι,a,vL_{\iota,a,v} denote the c1(KL1)c_{1}(K_{L}^{-1})-neighborhood of lι,a,vl_{\iota,a,v} and call it a bad strip. We denote the collection of all the bad strips by

(4.8) 𝕃:={Lι,a,v:lι,a,v is a bad line;ι=1,2}.\mathbb{L}:=\{L_{\iota,a,v}:l_{\iota,a,v}\text{ is a bad line};\iota=1,2\}.

4.1.2. Broad function

Consider dyadic numbers

(4.9) KL=Kd+1KdK1K0=K.K_{L}=K_{d+1}\ll K_{d}\ll\cdots\ll K_{1}\ll K_{0}=K.

Let M:=1020dϵ01M:=10^{20d}\epsilon_{0}^{-1}. Define 𝒫(K1,A)\mathcal{P}(K^{-1},A) to be a collection of all dyadic squares with side length K1K^{-1} in a set AA. We sometimes use 𝒫(K1)\mathcal{P}(K^{-1}) for 𝒫(K1,[1,1]2)\mathcal{P}(K^{-1},[-1,1]^{2}).

For every α=(α0,,αd+1)(0,1)d+2\alpha=(\alpha_{0},\ldots,\alpha_{d+1})\in(0,1)^{d+2}, we say that x3x\in\mathbb{R}^{3} is α\alpha-broad if

(4.10) maxL1,,LM𝕃|τ𝒫(K1,i=1MLi)Efτ(x)|αd+1|Ef(x)|\max_{L_{1},\ldots,L_{M}\in\mathbb{L}}\Big{|}\sum_{\begin{subarray}{c}\tau\in\mathcal{P}(K^{-1},\,\cap_{i=1}^{M}L_{i})\end{subarray}}Ef_{\tau}(x)\Big{|}\leq\alpha_{d+1}|Ef(x)|

and

(4.11) maxτj𝒫(Kj1)|Efτj(x)|+maxυj𝒫(Kj1)maxL1,,LM𝕃|τ𝒫(K1,υj(i=1MLi))Efτ(x)|αj|Ef(x)|\begin{split}&\max_{\tau_{j}\in\mathcal{P}(K_{j}^{-1})}|Ef_{\tau_{j}}(x)|\\ &+\max_{\upsilon_{j}\in\mathcal{P}(K_{j}^{-1})}\max_{L_{1},\ldots,L_{M}\in\mathbb{L}}\Big{|}\sum_{\begin{subarray}{c}\tau\in\mathcal{P}(K^{-1},\,\upsilon_{j}\cap(\cap_{i=1}^{M}L_{i}))\end{subarray}}Ef_{\tau}(x)\Big{|}\leq\alpha_{j}|Ef(x)|\end{split}

for every j=0,,d+1j=0,\ldots,d+1. For α(0,1)d+2\alpha\in(0,1)^{d+2} and rr\in\mathbb{R}, we use the notation rα=(rα0,,rαd+1)r\alpha=(r\alpha_{0},\ldots,r\alpha_{d+1}). We let BrαEf(x)\mathrm{Br}_{\alpha}Ef(x) denote the function which is |Ef(x)||Ef(x)| if xx is an α\alpha-broad point of EfEf and 0 otherwise.

We will prove the following estimate.

Theorem 4.1.

For every ϵ>0\epsilon>0 there exist dyadic numbers K,K1,,Kd+1K,K_{1},\ldots,K_{d+1} with

(4.12) Kd+1(ϵ)Kd(ϵ)K1(ϵ)K0(ϵ)=KK_{d+1}(\epsilon)\ll K_{d}(\epsilon)\ll\cdots\ll K_{1}(\epsilon)\ll K_{0}(\epsilon)=K

and αjKjϵ\alpha_{j}\sim K_{j}^{-\epsilon} such that for every R1R\geq 1, we have

(4.13) BrαEfL13/4(BR)Cϵ,dRϵR2(12413)fL28/13f^L5/13.\begin{split}\|\mathrm{Br}_{\alpha}Ef\|_{L^{13/4}(B_{R})}\leq C_{\epsilon,d}R^{\epsilon}R^{2(\frac{1}{2}-\frac{4}{13})}\|f\|_{L^{2}}^{8/13}\|\widehat{f}\|_{L^{\infty}}^{5/13}.\end{split}

Moreover, limϵ0Kd+1(ϵ)+\lim_{\epsilon\rightarrow 0}K_{d+1}(\epsilon)\rightarrow+\infty.

4.2. Proof of Theorem 2.1 for the negative curvature case

In this subsection, we prove (4.3) by assuming Theorem 4.1. By interpolation, it suffices to prove

(4.14) EfL13/4(BR)Cϵ,dRϵR2(12413)fL28/13f^L5/13.\|Ef\|_{L^{13/4}(B_{R})}\leq C_{\epsilon,d}R^{\epsilon}R^{2(\frac{1}{2}-\frac{4}{13})}\|{f}\|_{L^{2}}^{8/13}\|\widehat{f}\|_{L^{\infty}}^{5/13}.

We use induction on the radius RR. Assume that (4.14) is true for radius R/2\leq R/2. By isotropic and anisotropic rescaling, we can prove the following lemmas.

Lemma 4.2 (cf. Lemma 3.2 of [GO20]).

Suppose that 1KRR/21\lesssim K\leq R^{\prime}\leq R/2. Under the induction hypothesis, for every square τ𝒫(K1)\tau\in\mathcal{P}(K^{-1}), we have

(4.15) EfτL13/4(BR)K2+6413Cϵ,d(R)ϵ(R)2(12413)fτL28/13f^L5/13.\|Ef_{\tau}\|_{L^{13/4}(B_{R^{\prime}})}\leq K^{-2+6\cdot\frac{4}{13}}C_{\epsilon,d}(R^{\prime})^{\epsilon}(R^{\prime})^{2(\frac{1}{2}-\frac{4}{13})}\|f_{\tau}\|_{L^{2}}^{8/13}\|\widehat{f}\|_{L^{\infty}}^{5/13}.

The proof of the above lemma is essentially the same as that for Lemma 3.3. We leave out the details.

Lemma 4.3 (cf. Lemma 3.3 of [GO20]).

Suppose that 1Kd+1RR/21\lesssim K_{d+1}\leq R^{\prime}\leq R/2. Under the induction hypothesis, for every L𝕃L\in\mathbb{L}, we have

(4.16) EfLL13/4(BR)(Kd+1)113Cϵ,d(R)ϵ(R)2(12413)fLL28/13f^L5/13.\|{Ef_{L}}\|_{L^{13/4}(B_{R^{\prime}})}\leq(K_{d+1})^{-\frac{1}{13}}C_{\epsilon,d}({R^{\prime}})^{\epsilon}(R^{\prime})^{2(\frac{1}{2}-\frac{4}{13})}\|f_{L}\|_{L^{2}}^{8/13}\|\widehat{f}\|_{L^{\infty}}^{5/13}.
Sketch of the proof of Lemma 4.3.

The proof is essentially given in Lemma 3.3 of [GO20], so we only give a sketch here. For simplicity, assume that our strip LL is [0,1]×[0,Kd+11][0,1]\times[0,K_{d+1}^{-1}]. We apply the rescaling (ξ1,ξ2)(ξ1,Kd+1ξ2)(\xi_{1},\xi_{2})\mapsto(\xi_{1},K_{d+1}\xi_{2}). After the rescaling, EfLEf_{L} becomes

(4.17) Kd+11[1,1]2fL(ξ1,Kd+11ξ2)e((x1,Kd+11x2,Kd+11x3)(ξ1,ξ2,h~(ξ)))𝑑ξ1𝑑ξ2K_{d+1}^{-1}\int_{[-1,1]^{2}}f_{L}(\xi_{1},K_{d+1}^{-1}\xi_{2})e((x_{1},K_{d+1}^{-1}x_{2},K_{d+1}^{-1}x_{3})\cdot(\xi_{1},\xi_{2},\tilde{h}(\xi)))\,d\xi_{1}d\xi_{2}

where h~\tilde{h} is some function satisfying the conditions (4.1) and (4.2). Define the function g(ξ1,ξ2)=fL(ξ1,Kd+11ξ2)g(\xi_{1},\xi_{2})=f_{L}(\xi_{1},K_{d+1}^{-1}\xi_{2}) and denote (4.17) by E~g(x1,Kd+11x2,Kd+11x3)\tilde{E}g(x_{1},K_{d+1}^{-1}x_{2},K_{d+1}^{-1}x_{3}). We apply the change of variables on the physical variables xx, and obtain

(4.18) EfLL13/4([0,R]3)(Kd+1)513E~gL13/4([0,R]×[0,R/K]×[0,R/K]).\|Ef_{L}\|_{L^{13/4}([0,R^{\prime}]^{3})}\lesssim(K_{d+1})^{-\frac{5}{13}}\|\tilde{E}g\|_{L^{13/4}([0,R^{\prime}]\times[0,R^{\prime}/K]\times[0,R^{\prime}/K])}.

We decompose the rectangular box into smaller squares of sidelength R/KR^{\prime}/K, and apply the induction hypothesis on RR. Then (4.18) is bounded by

(4.19) (Kd+1)513(RKd+1)ϵ(RKd+1)513g2813g^513.(K_{d+1})^{-\frac{5}{13}}(\frac{R^{\prime}}{K_{d+1}})^{\epsilon}(\frac{R^{\prime}}{K_{d+1}})^{\frac{5}{13}}\|g\|_{2}^{\frac{8}{13}}\|\widehat{g}\|_{\infty}^{\frac{5}{13}}.

By changing back to the original variables, the above term becomes

(4.20) (Kd+1)113ϵ(R)ϵ(R)513f2813f^513.(K_{d+1})^{-\frac{1}{13}-\epsilon}({R^{\prime}})^{\epsilon}({R^{\prime}})^{\frac{5}{13}}\|f\|_{2}^{\frac{8}{13}}\|\widehat{f}\|_{\infty}^{\frac{5}{13}}.

This completes the proof. ∎

Let us continue the proof of (4.14). By the definition of the broad function, we obtain

(4.21) |Ef(x)||BrαEf(x)|+j=0d+1αj1(maxτj𝒫(Kj1)|Efτj(x)|+maxυj𝒫(Kj1)maxL1,,LM𝕃|τ𝒫(K1,υj(i=1MLi))Efτ(x)|)+(αd+1)1maxL1,,LM𝕃|τ𝒫(K1,i=1MLi)Efτ(x)|.\begin{split}|Ef(x)|&\leq|\mathrm{Br}_{\alpha}{Ef}(x)|+\sum_{j=0}^{d+1}\alpha_{j}^{-1}\Bigg{(}\max_{\tau_{j}\in\mathcal{P}(K_{j}^{-1})}|Ef_{\tau_{j}}(x)|\\ &+\max_{\upsilon_{j}\in\mathcal{P}(K_{j}^{-1})}\max_{L_{1},\ldots,L_{M}\in\mathbb{L}}\Big{|}\sum_{\tau\in\mathcal{P}(K^{-1},\,\upsilon_{j}\cap(\cap_{i=1}^{M}L_{i}))}Ef_{\tau}(x)\Big{|}\Bigg{)}\\ &+(\alpha_{d+1})^{-1}\max_{L_{1},\ldots,L_{M}\in\mathbb{L}}\Big{|}\sum_{\tau\in\mathcal{P}(K^{-1},\,\cap_{i=1}^{M}L_{i})}Ef_{\tau}(x)\Big{|}.\end{split}

We raise both sides to the 13/413/4-th power, integrate over BRB_{R}, replace the max by l13/4l^{13/4}-norms, and obtain

(4.22) BR|Ef|13/4CBR|BrαEf|13/4+Cj=0d+1αj13/4(τj𝒫(Kj1)BR|Efτj|13/4+υj𝒫(Kj1)L1,,LM𝕃BR|τ𝒫(K1,υj(i=1MLi))Efτ|13/4)+C(αd+1)13/4L1,,LM𝕃BR|τ𝒫(K1,i=1MLi)Efτ|13/4.\begin{split}\int_{B_{R}}|Ef|^{13/4}&\leq C\int_{B_{R}}|\mathrm{Br}_{\alpha}{Ef}|^{13/4}+C\sum_{j=0}^{d+1}\alpha_{j}^{-13/4}\Bigg{(}\sum_{\tau_{j}\in\mathcal{P}(K_{j}^{-1})}\int_{B_{R}}|Ef_{\tau_{j}}|^{13/4}\\ &+\sum_{\upsilon_{j}\in\mathcal{P}(K_{j}^{-1})}\sum_{L_{1},\ldots,L_{M}\in\mathbb{L}}\int_{B_{R}}\bigg{|}\sum_{\tau\in\mathcal{P}(K^{-1},\,\upsilon_{j}\cap(\cap_{i=1}^{M}L_{i}))}Ef_{\tau}\bigg{|}^{13/4}\Bigg{)}\\ &+C(\alpha_{d+1})^{-13/4}\sum_{L_{1},\ldots,L_{M}\in\mathbb{L}}\int_{B_{R}}\bigg{|}\sum_{\tau\in\mathcal{P}(K^{-1},\,\cap_{i=1}^{M}L_{i})}Ef_{\tau}\bigg{|}^{13/4}.\end{split}

We apply Theorem 4.1 to the first term, apply Hölder’s inequality, and obtain

(4.23) BR|BrαEf|13/4CCϵ,d13/4Rϵ/100R2(12413)134fL22f^L541013d/4Cϵ,d13/4R13ϵ/4R2(12413)134fL22f^54,\begin{split}\int_{B_{R}}|\mathrm{Br}_{\alpha}Ef|^{13/4}&\leq CC_{\epsilon,d}^{13/4}R^{\epsilon/100}R^{2(\frac{1}{2}-\frac{4}{13})\frac{13}{4}}\|f\|_{L^{2}}^{2}\|\widehat{f}\|_{L^{\infty}}^{\frac{5}{4}}\\ &\leq 10^{-13d/4}C_{\epsilon,d}^{13/4}R^{13\epsilon/4}R^{2(\frac{1}{2}-\frac{4}{13})\frac{13}{4}}\|f\|_{L^{2}}^{2}\|\widehat{f}\|_{\infty}^{\frac{5}{4}},\end{split}

provided that RR is large enough so that

(4.24) 1013d/4Rϵ/100R13ϵ/4.10^{13d/4}R^{\epsilon/100}\leq R^{13\epsilon/4}.

This takes care of the contribution from the first term. Let us bound the second term on the right hand side of (4.22). We first break our ball BRB_{R} into smaller balls BR/2B_{R/2}, and apply Lemma 4.2 with R=R/2R^{\prime}=R/2, and obtain

(4.25) τj𝒫(Kj1)BR|Efτj|13/4Kj100ϵCϵ,d13/4R13ϵ/4R2(12413)134τj𝒫(Kj1)fτjL22f^L5/4Kj100ϵCϵ,d13/4R13ϵ/4R2(12413)134fL22f^L5/4.\begin{split}\sum_{\tau_{j}\in\mathcal{P}(K_{j}^{-1})}\int_{B_{R}}|Ef_{\tau_{j}}|^{13/4}&\leq K_{j}^{-100\epsilon}C_{\epsilon,d}^{13/4}R^{13\epsilon/4}R^{2(\frac{1}{2}-\frac{4}{13})\frac{13}{4}}\sum_{\tau_{j}\in\mathcal{P}(K_{j}^{-1})}\|f_{\tau_{j}}\|_{L^{2}}^{2}\|\widehat{f}\|_{L^{\infty}}^{5/4}\\ &\leq K_{j}^{-100\epsilon}C_{\epsilon,d}^{13/4}R^{13\epsilon/4}R^{2(\frac{1}{2}-\frac{4}{13})\frac{13}{4}}\|f\|_{L^{2}}^{2}\|\widehat{f}\|_{L^{\infty}}^{5/4}.\end{split}

Recall that αjKϵ\alpha_{j}\sim K^{-\epsilon}. Hence, the second term is also harmless. The third term can be dealt with by following the same argument. We leave out the details. Let us move on to the fourth term. By Lemma 4.3, we obtain

(4.26) L1,,LM𝕃BR|τ𝒫(K1,i=1MLi)Efτ|13/4(Kd+1)100ϵCϵ,d13/4R13ϵ/4R2(12413)134(L1,,LM𝕃fi=1MLiL22)f^L5/4.\begin{split}&\sum_{L_{1},\ldots,L_{M}\in\mathbb{L}}\int_{B_{R}}\bigg{|}\sum_{\tau\in\mathcal{P}(K^{-1},\,\cap_{i=1}^{M}L_{i})}Ef_{\tau}\bigg{|}^{13/4}\\ &\leq(K_{d+1})^{-100\epsilon}C_{\epsilon,d}^{13/4}R^{13\epsilon/4}R^{2(\frac{1}{2}-\frac{4}{13})\frac{13}{4}}\Big{(}\sum_{L_{1},\ldots,L_{M}\in\mathbb{L}}\|f_{\cap_{i=1}^{M}L_{i}}\|_{L^{2}}^{2}\Big{)}\|\widehat{f}\|_{L^{\infty}}^{5/4}.\end{split}

By Lemma 3.1 of [GO20], we have

(4.27) L1,,LM𝕃fi=1MLiL22CMMf22.\sum_{L_{1},\ldots,L_{M}\in\mathbb{L}}\|f_{\cap_{i=1}^{M}L_{i}}\|_{L^{2}}^{2}\leq CM^{M}\|f\|_{2}^{2}.

Notice that the term MMM^{M} is harmless. Also, recall that αd+1Kd+1ϵ\alpha_{d+1}\sim K_{d+1}^{-\epsilon}. Combining all the information, we obtain the desired bound for the fourth term. This completes the proof of (4.14).

Now it suffices to prove Theorem 4.1, which we will discuss in the next section.

4.3. Proof of Theorem 4.1

In this subsection, we prove Theorem 4.1. Fix ϵ<1/100\epsilon<1/100. Take the parameters D=Rϵ6D=R^{\epsilon^{6}} and δ=ϵ2\delta=\epsilon^{2}. Let KK be a large number depending on ϵ\epsilon but independent of RR. We write f=τfτf=\sum_{\tau}f_{\tau}. Here, τ\tau is a ball of radius K1/2K^{-1/2}. Then we apply the wave packet decomposition to fτf_{\tau}: fτ=Tfτ,Tf_{\tau}=\sum_{T}f_{\tau,T}. We sometimes write fTf_{T} for fτ,Tf_{\tau,T}. We follow the proof in Section 3.2, and obtain an inequality similar to (3.10):

(4.28) BrαEfL13/4(BR)13/4Oi(1)BrαEfL13/4(Oi(1))13/4+BrαEfL13/4(W)13/4.\|\mathrm{Br}_{\alpha}Ef\|_{L^{13/4}(B_{R})}^{13/4}\lesssim\sum_{O_{i}^{(1)}}\|\mathrm{Br}_{\alpha}Ef\|_{L^{13/4}(O_{i}^{(1)})}^{13/4}+\|\mathrm{Br}_{\alpha}Ef\|_{L^{13/4}(W)}^{13/4}.

Define fif_{i} by

(4.29) fi:=τT𝕋ifτ,T.f_{i}:=\sum_{\tau}\sum_{T\in\mathbb{T}_{i}}f_{\tau,T}.

By Lemma 4.2 of [GO20], for every xOi(1)x\in O_{i}^{(1)}, we have

(4.30) BrαEf(x)Br2αEfi(x)+O(R900fL2).\mathrm{Br}_{\alpha}Ef(x)\leq\mathrm{Br}_{2\alpha}Ef_{i}(x)+O(R^{-900}\|f\|_{L^{2}}).

Cover the wall WW by balls BkB_{k} of radius R1δR^{1-\delta}. Define 𝕋k,\mathbb{T}_{k,-} and 𝕋k,+\mathbb{T}_{k,+} as in Definition 3.5 and 3.6. Define the functions fτ,k,+,fk,+f_{\tau,k,+},f_{k,+} and fI,k,+f_{I,k,+} as in (3.14) and (3.15). Lastly, we define the bilinear function. Compared with (3.16), the bilinear function is more involved here.

(4.31) Bil(Ef):=(τ,τ)𝒫(K1)×𝒫(K1):(τ,τ)is a good pair|Efτ|1/2|Efτ|1/2.\mathrm{Bil}(Ef):=\sum_{\begin{subarray}{c}(\tau,\tau^{\prime})\in\mathcal{P}(K^{-1})\times\mathcal{P}(K^{-1}):\\ (\tau,\tau^{\prime})\,\text{is a good pair}\end{subarray}}|Ef_{\tau}|^{1/2}|Ef_{\tau^{\prime}}|^{1/2}.

In the above formula, we use the notation “good pair” for which we didn’t give the definition in this paper. In fact, the definition is not important here because we will simply cite the following two lemmas as a black box. We refer to Section 3.2 of [GO20] for the definition of the “good pair” and more discussions.

Lemma 4.4 (Lemma 5.4 of [GO20]).

Suppose that α(0,1)d+2\alpha\in(0,1)^{d+2} satisfies

(4.32) KjϵαjKj+1100d,Kd+1ϵαd+110100dK_{j}^{-\epsilon}\leq\alpha_{j}\leq K_{j+1}^{-100d},\;\;\;\;K_{d+1}^{-\epsilon}\leq\alpha_{d+1}\leq 10^{-100d}

for every j=0,,dj=0,\ldots,d. Then for every xBkWx\in B_{k}\cap W

(4.33) BrαEf(x)2IBr200d2αEfI,k,+(x)+K0100Bil(Efk,)(x)+R50f2.\begin{split}\mathrm{Br}_{\alpha}Ef(x)&\leq 2\sum_{I}\mathrm{Br}_{200d^{2}\alpha}Ef_{I,k,+}(x)\\ &+K_{0}^{100}\mathrm{Bil}(Ef_{k,-})(x)+R^{-50}\|f\|_{2}.\end{split}

The summation I\sum_{I} runs over all possible collections of squares with sidelength K01K_{0}^{-1}. This summation does not play a significant role.

Lemma 4.5 ((5.127) of [GO20]).
(4.34) Bil(Efk,)L13/4(BkW)RCδR152(τ𝒫(K1)fτ,k,22)1/2.\begin{split}\|\mathrm{Bil}(Ef_{k,-})\|_{L^{13/4}(B_{k}\cap W)}\lesssim R^{C\delta}R^{\frac{1}{52}}\big{(}\sum_{\tau\in\mathcal{P}(K^{-1})}\|f_{\tau,k,-}\|_{2}^{2}\big{)}^{1/2}.\end{split}

The above lemma is proved in (5.127) of [GO20], which is a consequence of the L4L^{4}-estimate (Lemma 5.6 of [GO20]). We refer to the paper for the proof.

By Lemma 4.4, we have the counterpart of (3.18):

(4.35) BrαEfL13/4(BR)13/4Oi(1)BrαEfL13/4(Oi(1))13/4+k,IBr(200d2)αEfI,k,+L13/4(BkW)13/4+K100Bil(Efj,)L13/4(W)13/4+RapDec(R)f2.\begin{split}\|{\rm{Br}}_{\alpha}Ef\|_{L^{13/4}(B_{R})}^{13/4}&\lesssim\sum_{O_{i}^{(1)}}\!\|{\rm{Br}}_{\alpha}Ef\|^{13/4}_{L^{13/4}(O_{i}^{(1)})}+\sum_{k,I}\|{\rm{Br}}_{(200d^{2})\alpha}Ef_{I,k,+}\|^{13/4}_{L^{13/4}(B_{k}\cap W)}\\ &+K^{100}\|{\rm{Bil}}(Ef_{j,-})\|_{L^{13/4}(W)}^{13/4}+{\rm RapDec}(R)\|f\|_{2}.\end{split}

The only difference between this inequality and (3.18) is that BrαEfI,k,+\mathrm{Br}_{\alpha}Ef_{I,k,+} is replaced by Br(200d2)αEfI,k,+\mathrm{Br}_{(200d^{2})\alpha}Ef_{I,k,+}. This does not make any change for the rest of the proof. Note that the bilinear function also has the nice estimate (4.34) as a counterpart of (3.29).

So far, we have finished the part of the proof corresponding to Section 3.1 and Section 3.2. For each main estimate in Section 3.1 or Section 3.2, we can find in this section its counterpart which has been adapted to the negative curvature setting. What remains is to do iteration as we did in Section 3.3. Although we are in the negative curvature setting, we can follow the arguments in Section 3.3 line by line to finish the proof of Theorem 4.1. We do not reproduce it here.

5. A proof for the case α>1\alpha>1

In this section, we prove that when the surface {(ξ,h(ξ)),ξ[1,1]n1}\{(\xi,h(\xi)),\xi\in[-1,1]^{n-1}\} has positive principle curvatures we have

(5.1) EfLp(BR)Cp,ϵRϵR(n1)(121p)f^Lp\|Ef\|_{L^{p}(B_{R})}\leq C_{p,\epsilon}R^{\epsilon}R^{(n-1)(\frac{1}{2}-\frac{1}{p})}\|\widehat{f}\|_{L^{p}}

for every ppnp\geq p_{n}. Using Bourgain-Guth’s broad-narrow reduction in [BG11], (5.1) boils down to the following kk-broad estimate

(5.2) EfBLk,Ap(BR)Cp,ϵRϵR(n1)(121p)fL22/pf^L12/p,\|Ef\|_{\mathrm{BL}_{k,A}^{p}(B_{R})}\leq C_{p,\epsilon}R^{\epsilon}R^{(n-1)(\frac{1}{2}-\frac{1}{p})}\|{f}\|_{L^{2}}^{2/p}\|\widehat{f}\|_{L^{\infty}}^{1-2/p},

where the kk-broad norm is defined in [Gut18] Page 86. The reduction from (5.1) to (5.2) is quite standard, and we refer to [Gut18] Section 9 for details.

The main estimate is as follows.

Theorem 5.1.

Let 2kn12\leq k\leq n-1 and

(5.3) p>pn(k):=2+62(n1)+(k1)i=kn12i2i+1.p>p_{n}(k):=2+\frac{6}{2(n-1)+(k-1)\prod_{i=k}^{n-1}\frac{2i}{2i+1}}.

Then for every ϵ>0\epsilon>0 and R1R\geq 1, we have

(5.4) EfBLk,Ap(BR)Cp,ϵRϵR(n1)(121p)fL22/pf^L12/p.\|Ef\|_{\mathrm{BL}_{k,A}^{p}(B_{R})}\leq C_{p,\epsilon}R^{\epsilon}R^{(n-1)(\frac{1}{2}-\frac{1}{p})}\|f\|_{L^{2}}^{2/p}\|\widehat{f}\|_{L^{\infty}}^{1-2/p}.

Before we start the proof, let us recall some notations from [HZ20].

Definition 5.2.

A grain is defined to be a pair (S,Br)(S,B_{r}) where SnS\subset\mathbb{R}^{n} is a transverse complete intersection and BrnB_{r}\subset\mathbb{R}^{n} is a ball of radius r>0r>0. A multigrain is an (m+1)(m+1)-tuple of grains

(5.5) Sm=(𝒢0,,𝒢m),𝒢i=(Si,Bri)\vec{S}_{m}=(\mathcal{G}_{0},\ldots,\mathcal{G}_{m}),\;\;\;\;\mathcal{G}_{i}=(S_{i},B_{r_{i}})

satisfying

  1. (1)

    codimSi=i\mathrm{codim}S_{i}=i for 0im0\leq i\leq m,

  2. (2)

    SmSm1S0S_{m}\subset S_{m-1}\subset\ldots\subset S_{0},

  3. (3)

    BrmBrm1Br0B_{r_{m}}\subset B_{r_{m-1}}\subset\ldots\subset B_{r_{0}}.

Definition 5.3.

Let Sm=(𝒢0,,𝒢m)\vec{S}_{m}=(\mathcal{G}_{0},\ldots,\mathcal{G}_{m}) be a multigrain with

(5.6) 𝒢i=(Si,Bri(yi))for  0im.\mathcal{G}_{i}=(S_{i},B_{r_{i}}(y_{i}))\;\;\text{for}\;\;0\leq i\leq m.

Define 𝕋[Sm]\mathbb{T}[\vec{S}_{m}] to be a sub-collection of 𝕋[Br0(y0)]\mathbb{T}[B_{r_{0}}(y_{0})] for which element satisfies the following nested tube hypothesis: there exists Tθi,vi𝕋[Bri(yi)]T_{\theta_{i},v_{i}}\in\mathbb{T}[B_{r_{i}}(y_{i})] for 1im1\leq i\leq m such that

  1. (1)

    dist(θi,θj)rj1/2\mathrm{dist}(\theta_{i},\theta_{j})\lesssim r_{j}^{-1/2},

  2. (2)

    dist(Tθj,vj(yj),Tθi,vi(yi))ri1/2+δ\mathrm{dist}(T_{\theta_{j},v_{j}}(y_{j}),T_{\theta_{i},v_{i}}(y_{i}))\lesssim r_{i}^{1/2+\delta},

  3. (3)

    Tθj,vj(yj)Nrj1/2+δjSjT_{\theta_{j},v_{j}}(y_{j})\subset N_{r_{j}^{1/2+\delta_{j}}}S_{j}

hold for all 0ijm0\leq i\leq j\leq m.

The proof of Theorem 5.1 is based on an algorithm of polynomial partitioning. Let us state it below.

Algorithm 2 ([HZ20] page 9).

Inputs: Fix R1R\gg 1 and let ff be a smooth function satisfying

(5.7) EfBLk,Ap(BR)Cp,ϵRϵfL2.\|Ef\|_{\mathrm{BL}_{k,A}^{p}(B_{R})}\geq C_{p,\epsilon}R^{\epsilon}\|{f}\|_{L^{2}}.

Then we have the following Outputs:

\bullet 𝒪\mathcal{O} a finite collection of open subsets of n\mathbb{R}^{n} of diameter at most RϵR^{\epsilon}.

\bullet A codimension mm (0mnk0\leq m\leq n-k) and an integer parameter 1Am+1A1\leq A_{m+1}\leq A.

\bullet An (m+1)(m+1)-tuple of:

  1. (1)

    Scales r=(r0,,rm)\vec{r}=(r_{0},\ldots,r_{m}) satisfying R=r0>r1>>rm;R=r_{0}>r_{1}>\ldots>r_{m};

  2. (2)

    Large parameters D=(D1,,Dm+1)\vec{D}=(D_{1},\ldots,D_{m+1}).

\bullet For 0lm0\leq l\leq m a family 𝒮l\vec{\mathcal{S}}_{l} of level ll multigrains. Each Sl𝒮l\vec{S}_{l}\in\vec{\mathcal{S}}_{l} has multiscale rl=(r0,,rl)\vec{r}_{l}=(r_{0},\ldots,r_{l}).

\bullet For 0lm0\leq l\leq m an assigmment of a function fSlf_{\vec{S}_{l}} to each Sl𝒮l\vec{S}_{l}\in\vec{\mathcal{S}}_{l}.

The above data is chosen so that the following properties hold:

  1. (1)

    Define M(r,D):=(i=1mDi)mδ(i=1mri(βi1βi)/2Di(βi1βm)/2)M(\vec{r},\vec{D}):=\big{(}\prod_{i=1}^{m}D_{i}\big{)}^{m\delta}\big{(}\prod_{i=1}^{m}r_{i}^{(\beta_{i-1}-\beta_{i})/2}D_{i}^{(\beta_{i-1}-\beta_{m})/2}\big{)}. Then

    (5.8) EfBLk,Ap(BR)M(r,D)f21βm(O𝒪EfOBLk,Am+1pm(O)pm)βm/pm\|Ef\|_{\mathrm{BL}_{k,A}^{p}(B_{R})}\leq M(\vec{r},\vec{D})\|f\|_{2}^{1-\beta_{m}}\Big{(}\sum_{O\in\mathcal{O}}\|Ef_{O}\|_{\mathrm{BL}_{k,A_{m+1}}^{p_{m}}(O)}^{p_{m}}\Big{)}^{\beta_{m}/p_{m}}
  2. (2)
    (5.9) O𝒪EfO22(i=1m+1Di1+δ)Rϵf22\sum_{O\in\mathcal{O}}\|Ef_{O}\|_{2}^{2}\lesssim\big{(}\prod_{i=1}^{m+1}D_{i}^{1+\delta}\big{)}R^{\epsilon}\|f\|_{2}^{2}
  3. (3)

    For 1lm1\leq l\leq m

    (5.10) maxO𝒪fO22rll/2i=l+1m+1ri1/2Di(ni)+δRϵmaxSl𝒮lfSl22,\max_{O\in\mathcal{O}}\|f_{O}\|_{2}^{2}\lesssim r_{l}^{-l/2}\prod_{i=l+1}^{m+1}r_{i}^{-1/2}D_{i}^{-(n-i)+\delta}R^{\epsilon}\max_{\vec{S}_{l}\in\vec{\mathcal{S}}_{l}}\|f_{\vec{S}_{l}}\|_{2}^{2},

    where rm+1:=1r_{m+1}:=1.

  4. (4)

    For each multigrain Sl𝒮l\vec{S}_{l}\in\vec{\mathcal{S}}_{l}, we define

    (5.11) fSl:=T𝕋[Sl]fT,f^{\sharp}_{\vec{S}_{l}}:=\sum_{T\in\mathbb{T}[\vec{S}_{l}]}f_{T},

    where 𝕋[Sl]\mathbb{T}[\vec{S}_{l}] is defined in Definition 5.3. For 1lm1\leq l\leq m,

    (5.12) fSl22rll/2(i=1lri1/2Diδ)RϵfSl22.\|f_{\vec{S}_{l}}\|_{2}^{2}\lesssim r_{l}^{l/2}\big{(}\prod_{i=1}^{l}r_{i}^{-1/2}D_{i}^{\delta}\big{)}R^{\epsilon}\|f^{\sharp}_{\vec{S}_{l}}\|_{2}^{2}.

This finishes the statement of the algorithm. We are now ready to prove Theorem 5.1. Since O𝒪O\in\mathcal{O} has radius at most RϵR^{\epsilon}, we have

(5.13) EfOBLk,Am+1pm(O)RϵfO2.\|Ef_{O}\|_{\mathrm{BL}_{k,A_{m+1}}^{p_{m}}(O)}\lesssim R^{\epsilon}\|f_{O}\|_{2}.

Combining this inequality with (5.8) and (5.9), we have

(5.14) EfBLk,Ap(BR)i=1m+1ri(βi1βi)/2Diβi1/2(1/21/p)+O(δ)Rϵf22/pmaxO𝒪fO212/p.\|Ef\|_{\mathrm{BL}_{k,A}^{p}(B_{R})}\lesssim\prod_{i=1}^{m+1}r_{i}^{(\beta_{i-1}-\beta_{i})/2}D_{i}^{\beta_{i-1}/2-(1/2-1/p)+O(\delta)}R^{\epsilon}\|f\|_{2}^{2/p}\max_{O\in\mathcal{O}}\|f_{O}\|_{2}^{1-2/p}.

In order to estimate maxOfO2\max_{O}\|f_{O}\|_{2}, we will apply (5.10) and (5.12), and combine them with the following lemma. This lemma plays a role of the counterpart of Lemma 4.3 of [HZ20].

Lemma 5.4.

For mlnm\leq l\leq n,

(5.15) maxSl𝒮lfSl22(i=1lri1/2)RϵRn1f^2.\max_{\vec{S}_{l}\in\vec{\mathcal{S}}_{l}}\|f^{\sharp}_{\vec{S}_{l}}\|_{2}^{2}\lesssim\Big{(}\prod_{i=1}^{l}r_{i}^{-1/2}\Big{)}R^{\epsilon}R^{n-1}\|\widehat{f}\|_{\infty}^{2}.

The main ingredient of the proof of Lemma 5.4 is the nested polynomial Wolff.

Lemma 5.5 (Nested polynomial Wolff).

Fix r0r1rm>0r_{0}\geq r_{1}\geq\cdots\geq r_{m}>0 and ρ0ρ1ρm>0\rho_{0}\geq\rho_{1}\geq\cdots\geq\rho_{m}>0 so that 1ρmrmρm1rm1ρ0r01\geq\frac{\rho_{m}}{r_{m}}\geq\frac{\rho_{m-1}}{r_{m-1}}\geq\cdots\geq\frac{\rho_{0}}{r_{0}}. Let S0S1SmS_{0}\supset S_{1}\supset\cdots\supset S_{m} be semi-algebraic sets of complexity at most EE such that for each ii, SiS_{i} is a manifold with codimension ii contained in Bri(xi)B_{r_{i}}(x_{i}). We recursively define another sequence of sets S~i\widetilde{S}_{i} (0im)(0\leq i\leq m). Set S~m:=N2ρm(Sm)\widetilde{S}_{m}:=N_{2\rho_{m}}(S_{m}). For each i=1,,mi=1,\cdots,m, define

(5.16) S~i1:=N2ρi1(Si1)T a ρi×ri tubeTS~iFatri1ri(T).\widetilde{S}_{i-1}:=N_{2\rho_{i-1}}(S_{i-1})\cap\bigcup_{\begin{subarray}{c}T\textup{~{}a~{}}\rho_{i}\times r_{i}\textup{~{}tube}\\ T\subset\widetilde{S}_{i}\end{subarray}}\textup{Fat}_{\frac{r_{i-1}}{r_{i}}}(T).

We also define

S~0:=T a ρ0×r0 tube(T,en)<1/10TS~0T\widetilde{S}_{0}^{\prime}:=\bigcup_{\begin{subarray}{c}T\textup{~{}a~{}}\rho_{0}\times r_{0}\textup{~{}tube}\\ \angle(T,e_{n})<1/10\\ T\subset\widetilde{S}_{0}\end{subarray}}T

Then for every a[2r0,2r0]a\in[-2r_{0},2r_{0}], we have

(5.17) |S~0(n1×{a})|C(n,E,ϵ)r0ϵr0n1i=1mρiri.|\widetilde{S}_{0}^{\prime}\cap(\mathbb{R}^{n-1}\times\{a\})|\leq C(n,E,\epsilon)r_{0}^{\epsilon}r_{0}^{n-1}\prod_{i=1}^{m}\frac{\rho_{i}}{r_{i}}.
Remark 5.6.

This is a higher dimensional version of Lemma 3.11. One minor difference is that the constant C(n,E,ϵ)C(n,E,\epsilon) here behaves badly when the complexity EE is big, so in order to apply this lemma we have to set EE as an absolute constant that only depends on ϵ\epsilon.

Since the proof of Lemma 5.5 is implicitly contained in [Zah21], we just sketch the proof here. First, by (2.33) in [Zah21], we have the estimate

|S~0|C(n,E,ϵ)r0ϵr0ni=1mρiri.|\widetilde{S}_{0}|\leq C(n,E,\epsilon)r_{0}^{\epsilon}r_{0}^{n}\prod_{i=1}^{m}\frac{\rho_{i}}{r_{i}}.

Next, we follow the proof of Lemma 2.6 in [Zah21] where we set SS to be S~0\widetilde{S}_{0}^{\prime}. From (2.17) in [Zah21] where we set λ=r0\lambda=r_{0} and A=1A=1, we obtain (5.17). We are ready to prove Lemma 5.4.

Proof of Lemma 5.4.

The proof is the same as that for Lemma 3.12. Set

(5.18) X=(T𝕋[Sl]T)(n1×{0}).X=\Big{(}\bigcup_{T\in\mathbb{T}[\vec{S}_{l}]}T\Big{)}\cap(\mathbb{R}^{n-1}\times\{0\}).

We see that fSl^\widehat{f^{\sharp}_{\vec{S}_{l}}} is essentially supported in XX, so

(5.19) fSl^221XfSl^221Xf^22.\|\widehat{f^{\sharp}_{\vec{S}_{l}}}\|_{2}^{2}\lesssim\|1_{X}\widehat{f^{\sharp}_{\vec{S}_{l}}}\|_{2}^{2}\lesssim\|1_{X}\widehat{f}\|_{2}^{2}.

Next, we use Lemma 5.5 to estimate |X||X|. Set m=lm=l and ρi=ri1/2+δi\rho_{i}=r_{i}^{1/2+\delta_{i}} (0il0\leq i\leq l) in Lemma 5.5 so that we obtain a set S~0\widetilde{S}_{0}^{\prime} and the estimate (5.17). By the definition of 𝕋[Sl]\mathbb{T}[\vec{S}_{l}] in Definition 5.3, we see that

T𝕋[Sl]TS~0.\bigcup_{T\in\mathbb{T}[\vec{S}_{l}]}T\subset\widetilde{S}_{0}^{\prime}.

As a result,

|X||S~0(n1×{a})|C(n,E,ϵ)r0ϵr0n1i=1mρiri.|X|\leq|\widetilde{S}_{0}\cap(\mathbb{R}^{n-1}\times\{a\})|\leq C(n,E,\epsilon)r_{0}^{\epsilon}r_{0}^{n-1}\prod_{i=1}^{m}\frac{\rho_{i}}{r_{i}}.

Noting that r0=Rr_{0}=R, we apply Hölder’s inequality to (5.19) to finish the proof of Lemma 5.4. ∎

By (5.10), (5.12), and the above lemma, we have

(5.20) maxO𝒪fO22RϵRn1(i=1mri1/2Diδ)(i=1lri1/2)(i=l+1m+1Di(ni))f^2\max_{O\in\mathcal{O}}\|f_{O}\|_{2}^{2}\lesssim R^{\epsilon}R^{n-1}\Big{(}\prod_{i=1}^{m}r_{i}^{-1/2}D_{i}^{\delta}\Big{)}\Big{(}\prod_{i=1}^{l}r_{i}^{-1/2}\Big{)}\Big{(}\prod_{i=l+1}^{m+1}D_{i}^{-(n-i)}\Big{)}\|\widehat{f}\|_{\infty}^{2}

for all 0lm0\leq l\leq m. We take a geometric average over ll with weight 0γl10\leq\gamma_{l}\leq 1. Then

(5.21) maxO𝒪fO22RϵRn1i=1m+1ri(1+σi)/2Di(ni)(1σi)+O(δ)f^2\max_{O\in\mathcal{O}}\|f_{O}\|_{2}^{2}\lesssim R^{\epsilon}R^{n-1}\prod_{i=1}^{m+1}r_{i}^{-(1+\sigma_{i})/2}D_{i}^{-(n-i)(1-\sigma_{i})+O(\delta)}\|\widehat{f}\|_{\infty}^{2}

where σi:=j=imγj\sigma_{i}:=\sum_{j=i}^{m}\gamma_{j} for 0im0\leq i\leq m and σm+1=0\sigma_{m+1}=0. Applying this inequality to the right hand side of (5.14), we obtain

(5.22) EfBLk,Ap(BR)R(n1)(12p)i=1m+1riXiDiYi+O(δ)f22/pf^12/p.\|Ef\|_{\mathrm{BL}_{k,A}^{p}(B_{R})}\lesssim R^{(n-1)(1-\frac{2}{p})}\prod_{i=1}^{m+1}r_{i}^{X_{i}}D_{i}^{Y_{i}+O(\delta)}\|f\|_{2}^{2/p}\|\widehat{f}\|_{\infty}^{1-2/p}.

Here XiX_{i} and YiY_{i} are given on page 12 of [HZ20]. Following the same optimization process therein gives the desired bound (5.4). We do not reproduce it.

6. Appendix: Approach using pseudo-conformal transformation

In this section, we briefly explain how the pseudo-conformal transformation can be applied to local smoothing problems. Similar ideas date back to [Car92], [Rog08].

Define a “restriction type” operator by

(6.1) ERf(x,t):=n1e(R2t|xtyR|αα1)ψ(xtyR)f(y)𝑑yE_{R}f(x,t):=\int_{\mathbb{R}^{n-1}}e\big{(}\frac{R^{2}}{t}\big{|}\frac{x-ty}{R}\big{|}^{\frac{\alpha}{\alpha-1}}\big{)}\psi\big{(}\frac{x-ty}{R}\big{)}f(y)\,dy

for fL1([0,1]n1)f\in L^{1}([0,1]^{n-1}). Here, ψ\psi is a compactly supported smooth function.

Conjecture 6.1 (Restriction type conjecture).

For p>2nn1p>\frac{2n}{n-1}, ϵ>0\epsilon>0, and R1R\geq 1, it holds that

(6.2) ERfLp([0,R]n1×[R/2,R])Cp,ϵRϵfLp.\|E_{R}f\|_{L^{p}([0,R]^{n-1}\times[R/2,R])}\leq C_{p,\epsilon}R^{\epsilon}\|f\|_{L^{p}}.

Notice that when α=2\alpha=2, this conjecture follows from the restriction conjecture for a paraboloid by the Taylor’s theorem on the function ψ\psi. In [Rog08], it is proved that the restriction estimate for a paraboloid implies a local smoothing estimate for the Schrödinger equation. The goal of this section is to generalize his theorem to the general fractional Schrödinger equations.

Theorem 6.2.

Let α(0,1)(1,)\alpha\in(0,1)\cup(1,\infty). The extension type estimate (6.2) for pp implies the local smoothing estimate (1.3) for every β>βc\beta>\beta_{c} for the same pp.

The wave packets of ERfE_{R}f behave very similarly to that for a restriction operator for paraboloid. It looks very plausible to prove Theorem 1.2 for the case α>1\alpha>1 by using the above theorem and following the arguments in [GOW+21b]. On the other hand, for the case α<1\alpha<1, the operator ERfE_{R}f behaves differently from that for the case α>1\alpha>1—the manifold associated with the operator has a negative Gaussian curvature. In particular, the operator does not seem to have a transverse equidistribution property, which is a key ingredient in [GOW+21b]. Even though the restriction estimate for surfaces with negative Gaussian curvature is well understood in [GO20] and [BMV20], the arguments therein do not simply rely on the properties of wave packets. Thus, it is not straightforward to the authors whether their ideas can be applied to the operator ERfE_{R}f. We do not explore in this direction here.

Sketch of the proof of Theorem 6.2.

Recall that in Section 2 we showed that (1.3) follows by (2.5). So it remains to prove

(6.3) eit(Δ)α/2fLx,tp(BRn1×[R/2,R])R(n1)(121p)+ϵfLp(n1)\Big{\|}e^{it(-\Delta)^{\alpha/2}}f\big{\|}_{L^{p}_{x,t}(B_{R}^{n-1}\times[R/2,R])}\lesssim R^{(n-1)(\frac{1}{2}-\frac{1}{p})+\epsilon}\|f\|_{L^{p}(\mathbb{R}^{n-1})}

for ff whose Fourier support is contained in the annulus {ξn1:|ξ|1}\{\xi\in\mathbb{R}^{n-1}:|\xi|\sim 1\}. By plugging a smooth cutoff function ψ\psi with a suitable support (ψ(ξ)=1\psi(\xi)=1 on the annulus {ξn1:|ξ|1}\{\xi\in\mathbb{R}^{n-1}:|\xi|\sim 1\} and ψ\psi is supported on a slightly thicker annulus), we may write

(6.4) eit(Δ)α/2f(x)=n1ψ(ξ)f^(ξ)e(xξ+t|ξ|α)𝑑ξ.e^{it(-\Delta)^{\alpha/2}}f(x)=\int_{\mathbb{R}^{n-1}}\psi(\xi)\widehat{f}(\xi)e(x\cdot\xi+t|\xi|^{\alpha})\,d\xi.

In this way, we can write

(6.5) eit(Δ)α/2f(x)=Ktf(x),e^{it(-\Delta)^{\alpha/2}}f(x)=K_{t}*f(x),

where the kernel Kt(x)K_{t}(x) is defined as

(6.6) Kt(x)=nψ(ξ)e(xξ+t|ξ|α)𝑑ξ=eit(Δ)α/2ψ^(x).K_{t}(x)=\int_{\mathbb{R}^{n}}\psi(\xi)e(x\cdot\xi+t|\xi|^{\alpha})\,d\xi=e^{it(-\Delta)^{\alpha/2}}\widehat{\psi}(x).

Since ψ\psi is a smooth function, it follows that Kt(x)K_{t}(x) decays rapidly outside BRB_{R}. Hence one can assume that ff is supported in an RR ball in n1\mathbb{R}^{n-1} without loss of generality.

Let us rewrite the kernel as

(6.7) Kt(x)=n1ψ(ξ)e(t(ξxt+|ξ|α))𝑑ξ.K_{t}(x)=\int_{\mathbb{R}^{n-1}}\psi(\xi)e\big{(}t(\xi\cdot\frac{x}{t}+|\xi|^{\alpha})\big{)}\,d\xi.

Note that this kernel decays rapidly for x,tx,t on the region |x/t|1|x/t|\gtrsim 1. Hence, we may assume that |x/t|1|x/t|\lesssim 1. In particular, since tt is restricted to the range [R/2,R][R/2,R], one can assume |x|R|x|\lesssim R. For simplicity, let us introduce the notation x~:=x/t\tilde{x}:=x/t. Define the phase function ϕ(ξ):=ξx~+|ξ|α\phi(\xi):=\xi\cdot\tilde{x}+|\xi|^{\alpha}. Then

(6.8) ξϕ(ξ)=x~+α|ξ|α2ξ\nabla_{\xi}\phi(\xi)=\tilde{x}+\alpha|\xi|^{\alpha-2}\xi

and this function vanishes only at ξ=ξc:=(x~|x~|1)(α1|x~|)1/(α1)\xi=\xi_{c}:=(-\tilde{x}|\tilde{x}|^{-1})(\alpha^{-1}|\tilde{x}|)^{1/(\alpha-1)}. By the method of stationary phase (see Theorem 7.7.5 of [H0̈3]) and considering only the main term (the other terms can be handled similarly), one can pretend that

(6.9) Kt(x)t(n1)/2e(tϕ(ξc))ψ(ξc).K_{t}(x)\sim t^{-(n-1)/2}e(t\phi(\xi_{c}))\psi(\xi_{c}).

Simple calculation gives

(6.10) ϕ(ξc)=α1α11(1α)|t1x|αα1.\phi(\xi_{c})=\alpha^{-\frac{1}{\alpha-1}-1}(1-\alpha)|t^{-1}{x}|^{\frac{\alpha}{\alpha-1}}.

The factor α1α11(1α)\alpha^{-\frac{1}{\alpha-1}-1}(1-\alpha) can be treated as 1 after using some scaling on the variable tt. So, our operator eit(Δ)α/2f(x)e^{it(-\Delta)^{\alpha/2}}f(x) can be morally written as

(6.11) Ktf(x)t(n1)/2ne(t|xyt|αα1)ψ(xyt)f(y)𝑑y=:t(n1)/2Tf(x,t)\begin{split}K_{t}*f(x)&\sim t^{-(n-1)/2}\int_{\mathbb{R}^{n}}e\big{(}t\big{|}\frac{x-y}{t}\big{|}^{\frac{\alpha}{\alpha-1}}\big{)}\psi(\frac{x-y}{t})f(y)\,dy\\ &=:t^{-(n-1)/2}Tf(x,t)\end{split}

where |t|αR|t|\sim_{\alpha}R.

Now let us prove (6.3). Via (6.11), what we need to prove becomes

(6.12) TfLx,tp(BRn1×[R/2,R])Rn12R(n1)(121p)+ϵfLp(n1).\|Tf\|_{L^{p}_{x,t}(B_{R}^{n-1}\times[R/2,R])}\lesssim R^{\frac{n-1}{2}}R^{(n-1)(\frac{1}{2}-\frac{1}{p})+\epsilon}\|f\|_{L^{p}(\mathbb{R}^{n-1})}.

Define

(6.13) T~f(x,t):=ne(t1|xty|αα1)ψ(xty)f(y)𝑑y\tilde{T}f(x,t):=\int_{\mathbb{R}^{n}}e\big{(}t^{-1}|x-ty|^{\frac{\alpha}{\alpha-1}}\big{)}\psi(x-ty)f(y)\,dy

so that

(6.14) Tf(x,t)=T~f(xt,1t).Tf(x,t)=\tilde{T}f\big{(}\frac{x}{t},\frac{1}{t}\big{)}.

Employing the pseudo-conformal transformation (x/t,1/t)(x,t)(x/t,1/t)\mapsto({x},{t}), we have

(6.15) TfLp(BRn1×[R/2,R])Rn+1pT~fLp([0,1]n1×[1/(2R),1/R]).\|Tf\|_{L^{p}(B_{R}^{n-1}\times[R/2,R])}\sim R^{\frac{n+1}{p}}\|\tilde{T}f\|_{L^{p}([0,1]^{n-1}\times[1/(2R),1/R])}.

Note that

(6.16) T~f(R1x,R2t)=Rn1ERg(x,t),\tilde{T}f(R^{-1}x,R^{-2}t)=R^{n-1}E_{R}g(x,t),

where g(y):=f(Ry)g(y):=f(Ry) and ERgE_{R}g is the extension operator (6.1). We apply some rescaling on the physical variables: (x,t)(R1x,R2t)(x,t)\mapsto(R^{-1}x,R^{-2}t) and on the frequency variables: ξRξ\xi\mapsto R\xi. After this rescaling, we obtain

(6.17) T~fLp([0,1]n1×[1/(2R),1/R])Rn1n+1pERgLp(BRn1×[R/2,R]).\|\tilde{T}f\|_{L^{p}([0,1]^{n-1}\times[1/(2R),1/R])}\sim R^{n-1-\frac{n+1}{p}}\|E_{R}g\|_{L^{p}(B_{R}^{n-1}\times[R/2,R])}.

The hypothesis (6.2) gives

(6.18) ERgLp(BRn1×[R/2,R])RϵgLp=R(n1)/pfLp.\begin{split}\|E_{R}g\|_{L^{p}(B_{R}^{n-1}\times[R/2,R])}\lesssim R^{\epsilon}\|g\|_{L^{p}}=R^{-(n-1)/p}\|f\|_{L^{p}}.\end{split}

Combining all the inequalities (6.15), (6.17), and (6.18), we finally get

(6.19) TfLp(BRn1×[R/2,R])Rn1R(n1)/pRϵfLp.\|Tf\|_{L^{p}(B_{R}^{n-1}\times[R/2,R])}\lesssim R^{n-1}R^{-(n-1)/p}R^{\epsilon}\|f\|_{L^{p}}.

This completes the proof of (6.12), and thus, the proof of Theorem 6.2. ∎

References

  • [BG11] Jean Bourgain and Larry Guth. Bounds on oscillatory integral operators based on multilinear estimates. Geom. Funct. Anal., 21(6):1239–1295, 2011.
  • [BMV20] Stefan Buschenhenke, Detlef Müller, and Ana Vargas. Fourier restriction for smooth hyperbolic 2-surfaces. preprint, arXiv:2010.10449, 2020.
  • [Car92] Anthony Carbery. Restriction implies Bochner-Riesz for paraboloids. Math. Proc. Cambridge Philos. Soc., 111(3):525–529, 1992.
  • [GJW21] Shengwen Gan, Yifan Jing, and Shukun Wu. New bounds for Stein’s square function in 3\mathbb{R}^{3}. preprint, arXiv:2104.14731, 2021.
  • [GMZ20] Chuanwei Gao, Changxing Miao, and Jiqiang Zheng. Improved local smoothing estimate for the fractional schrödinger operator, 2020.
  • [GO20] Shaoming Guo and Changkeun Oh. A restriction estimate for surfaces with negative gaussian curvatures, 2020.
  • [GOW21a] Shengwen Gan, Changkeun Oh, and Shukun Wu. New bounds for stein’s square functions in higher dimensions. preprint, arXiv:2108.11567, 2021.
  • [GOW+21b] Shaoming Guo, Changkeun Oh, Hong Wang, Shukun Wu, and Ruixiang Zhang. The Bochner-Riesz problem: an old approach revisited. preprint, arXiv:2104.11188, 2021.
  • [GRY20] Shaoming Guo, Joris Roos, and Po-Lam Yung. Sharp variation-norm estimates for oscillatory integrals related to Carleson’s theorem. Anal. PDE, 13(5):1457–1500, 2020.
  • [Gut16] Larry Guth. A restriction estimate using polynomial partitioning. J. Amer. Math. Soc., 29(2):371–413, 2016.
  • [Gut18] Larry Guth. Restriction estimates using polynomial partitioning ii. Acta Math., 221(1):81–142, 09 2018.
  • [GZ18] Larry Guth and Joshua Zahl. Polynomial Wolff axioms and Kakeya-type estimates in 4\mathbb{R}^{4}. Proceedings of the London Mathematical Society, 117(1):192–220, 2018.
  • [H0̈3] Lars Hörmander. The analysis of linear partial differential operators. I. Classics in Mathematics. Springer-Verlag, Berlin, 2003. Distribution theory and Fourier analysis, Reprint of the second (1990) edition [Springer, Berlin; MR1065993 (91m:35001a)].
  • [HRZ19] Jonathan Hickman, Keith M Rogers, and Ruixiang Zhang. Improved bounds for the Kakeya maximal conjecture in higher dimensions. arXiv preprint arXiv:1908.05589, 2019.
  • [HZ20] Jonathan Hickman and Joshua Zahl. A note on fourier restriction and nested polynomial Wolff axioms. arXiv:2010.02251, 2020.
  • [KR18] Nets Hawk Katz and Keith M. Rogers. On the polynomial Wolff axioms. Geometric and Functional Analysis, 28(6):1706–1716, 2018.
  • [Rog08] Keith M. Rogers. A local smoothing estimate for the Schrödinger equation. Adv. Math., 219(6):2105–2122, 2008.
  • [RS10] Keith M. Rogers and Andreas Seeger. Endpoint maximal and smoothing estimates for Schrödinger equations. J. Reine Angew. Math., 640:47–66, 2010.
  • [Wan18] Hong Wang. A restriction estimate in 3\mathbb{R}^{3} using brooms. arXiv:1802.04312, 2018.
  • [Wu20] Shukun Wu. On the Bochner-Riesz operator in 3{\mathbb{R}^{3}}. preprint, arXiv:2008.13043, 2020.
  • [Zah21] Joshua Zahl. New Kakeya estimates using Gromov’s algebraic lemma. Advances in Mathematics, 380:107596, 2021.