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

\stackMath

On restricted projections to planes in 3\mathbb{R}^{3}

Shengwen Gan Department of Mathematics
Massachusetts Institute of Technology
Cambridge, MA 02142-4307, USA
shengwen@mit.edu
Shaoming Guo Department of Mathematics
University of Wisconsin-Madison
Madison, WI-53706, USA
shaomingguo@math.wisc.edu
Larry Guth Department of Mathematics
Massachusetts Institute of Technology
Cambridge, MA 02142-4307, USA
lguth@math.mit.edu
Terence L. J. Harris Department of Mathematics
Cornell University
Ithaca
NY 14853-4201
USA
Department of Mathematics
University of Wisconsin-Madison
Madison, WI-53706, USA terry.harris@wisc.edu
Dominique Maldague Department of Mathematics
Massachusetts Institute of Technology
Cambridge, MA 02142-4307, USA
dmal@mit.edu
 and  Hong Wang Department of Mathematics
UCLA
Los Angeles, CA 90095, USA
hongwang@math.ucla.edu
Abstract.

Let γ:[0,1]𝕊2\gamma:[0,1]\rightarrow\mathbb{S}^{2} be a non-degenerate curve in 3\mathbb{R}^{3}, that is to say, det(γ(θ),γ(θ),γ′′(θ))0\det\big{(}\gamma(\theta),\gamma^{\prime}(\theta),\gamma^{\prime\prime}(\theta)\big{)}\neq 0. For each θ[0,1]\theta\in[0,1], let Vθ=γ(θ)V_{\theta}=\gamma(\theta)^{\perp} and let πθ:3Vθ\pi_{\theta}:\mathbb{R}^{3}\rightarrow V_{\theta} be the orthogonal projections. We prove that if A3A\subset\mathbb{R}^{3} is a Borel set, then for a.e. θ[0,1]\theta\in[0,1] we have dim(πθ(A))=min{2,dimA}\textup{dim}(\pi_{\theta}(A))=\min\{2,\textup{dim}A\}.

More generally, we prove an exceptional set estimate. For A3A\subset\mathbb{R}^{3} and 0s20\leq s\leq 2, define Es(A):={θ[0,1]:dim(πθ(A))<s}E_{s}(A):=\{\theta\in[0,1]:\textup{dim}(\pi_{\theta}(A))<s\}. We have dim(Es(A))max{1+sdim(A),0}\textup{dim}(E_{s}(A))\leq\max\{1+s-\textup{dim}(A),0\}. We also prove that if dim(A)>2\textup{dim}(A)>2, then for a.e. θ[0,1]\theta\in[0,1] we have 2(πθ(A))>0\mathcal{H}^{2}(\pi_{\theta}(A))>0.

Key words and phrases:
decoupling inequalities, superlevel set
2020 Mathematics Subject Classification:
42B15, 42B20

1. Introduction

Let 𝕊23\mathbb{S}^{2}\subset\mathbb{R}^{3} denote the unit sphere. Let γ:[0,1]𝕊2\gamma:[0,1]\to\mathbb{S}^{2} be a C2C^{2} curve. We say that γ\gamma is non-degenerate if

(1) det(γ(θ),γ(θ),γ′′(θ))0,\det(\gamma(\theta),\gamma^{\prime}(\theta),\gamma^{\prime\prime}(\theta))\neq 0,

for every θ[0,1]\theta\in[0,1]. A model example for the non-degenerate curve is γ:θ(cosθ2,sinθ2,12)\gamma_{\circ}:\theta\mapsto(\frac{\cos\theta}{\sqrt{2}},\frac{\sin\theta}{\sqrt{2}},\frac{1}{\sqrt{2}}) (θ[0,1])(\theta\in[0,1]).

For a given θ[0,1]\theta\in[0,1], let Vθ=γ(θ)V_{\theta}=\gamma(\theta)^{\perp} denote the orthogonal complement of γ(θ)\gamma(\theta) in 3\mathbb{R}^{3}, and let πθ:3γ(θ)\pi_{\theta}:\mathbb{R}^{3}\to\gamma(\theta)^{\perp} denote the orthogonal projection onto γ(θ)\gamma(\theta)^{\perp}. For α>0\alpha>0 and a Borel set E3E\subset\mathbb{R}^{3}, we will use α(E)\mathcal{H}^{\alpha}(E) to denote the α\alpha-dimensional Hausdorff measure of EE. Moreover, we use dimX\textup{dim}X to denote the Hausdorff dimension of a set XX.

Theorem 1.

Suppose A3A\subset\mathbb{R}^{3} is a Borel set of Hausdorff dimension α\alpha. For 0s<20\leq s<2, define the exceptional set

Es={θ[0,1]:dim(πθ(A))<s}.E_{s}=\{\theta\in[0,1]:\textup{dim}(\pi_{\theta}(A))<s\}.

Then we have

dim(Es)max{1+sα,0}.\textup{dim}(E_{s})\leq\max\{1+s-\alpha,0\}.

As an immediate corollary, we have:

Corollary 1.

Suppose A3A\subset\mathbb{R}^{3} is a Borel set of Hausdorff dimension α\alpha. Then we have

dim(πθ(A))=min{2,α}, for a.e. θ[0,1].\textup{dim}(\pi_{\theta}(A))=\min\{2,\alpha\},\textup{~{}for~{}a.e.~{}}\theta\in[0,1].
Theorem 2.

Suppose A3A\subset\mathbb{R}^{3} is a Borel set of Hausdorff dimension greater than 22. Then

(2) 2(πθ(A))>0,\mathcal{H}^{2}(\pi_{\theta}(A))>0,

for almost every θ[0,1]\theta\in[0,1].

1.1. Background of the problems

The projection theory dates back to Marstrand [18], who showed that if AA is a Borel set in 2\mathbb{R}^{2}, then the projection of AA onto almost every line through the origin has Hausdorff dimension min{1,dimA}\min\{1,\textup{dim}A\}. This was generalized to higher dimensions by Mattila [19], who showed that if AA is a Borel set in n\mathbb{R}^{n}, then the projection of AA onto almost every kk-plane through the origin has Hausdorff dimension min{k,dimA}\min\{k,\textup{dim}A\}. It is more general to consider the projection problem when the direction set is restricted to some submanifold of the Grassmannian. To have a better understanding of this restricted projection problem, the first step is to study the problem in 3\mathbb{R}^{3}. Fässler and Orponen made a conjecture about restricted projections to lines and planes (see Conjecture 1.6 in [5]), and there has been much related research (see for example [5], [2], [16], [15], [17], [21], [22], [23], [13], [14]). For more of an introduction to this problem, we refer to [13]. Recently, Käenmäki-Orponen-Venieri [17] and Pramanik-Yang-Zahl [25] proved one half of the conjecture: restricted projections to lines. In this paper, we resolve another half of the conjecture: restricted projections to planes.

1.2. An overview of the high-low method

The high-low method is a powerful tool developed recently in Fourier analysis. There are many applications of the high-low method, see for example [11], [3], [12], [10], [6], [7].

In this subsection, we briefly discuss how the high-low method can be used to study projection theory. As a warm-up, we study Marstrand’s projection theorem from another point of view, using the high-low method.

Theorem 3 (Marstrand’s projection theorem).

For each θ[0,π]\theta\in[0,\pi], define Lθ:={x2:arg(x)=θ}L_{\theta}:=\{x\in\mathbb{R}^{2}:\arg(x)=\theta\} and let pθ:2Lθp_{\theta}:\mathbb{R}^{2}\rightarrow L_{\theta} be the projection. Suppose A2A\subset\mathbb{R}^{2} is a Borel set, then we have dim(pθ(A))=min{1,dimA}\textup{dim}(p_{\theta}(A))=\min\{1,\textup{dim}A\} for a.e. θ[0,π]\theta\in[0,\pi].

We will frequently use the following definition.

Definition 1.

For a number δ>0\delta>0 and any set XX, we use |X|δ|X|_{\delta} to denote the maximal number of δ\delta-separated points in XX.

Marstrand’s projection theorem can be reduced to the following discretized version. We do not show how reduction works in this section, but we will give the full details in the later section when we prove our main theorems.

Proposition 1.

Fix 0<s<10<s<1. Fix a small scale δ>0\delta>0 and set Θ=δ[0,π]\Theta=\delta\mathbb{Z}\cap[0,\pi]. For each θΘ\theta\in\Theta, define Lθ:={x2:arg(x)=θ}L_{\theta}:=\{x\in\mathbb{R}^{2}:\arg(x)=\theta\} and let pθ:2Lθp_{\theta}:\mathbb{R}^{2}\rightarrow L_{\theta} be the projection. Suppose AB2(0,1)A\subset B^{2}(0,1) is a union of disjoint δ\delta-balls with measure |A|=δ2a|A|=\delta^{2-a}, or equivalently |A|δδa|A|_{\delta}\sim\delta^{-a}. We also assume there is a subset ΘΘ\Theta^{\prime}\subset\Theta with #Θδ1\#\Theta^{\prime}\gtrsim\delta^{-1}, such that for any θΘ\theta\in\Theta^{\prime}, pθ(A)p_{\theta}(A) (which is a union of line segments of length δ\delta in LθL_{\theta}) satisfies the ss-dimensional condition: For each rδr\geq\delta and line segment IrLθI_{r}\subset L_{\theta} of length rr, we have

(3) |pθ(A)Ir|δ(r/δ)s.|p_{\theta}(A)\cap I_{r}|_{\delta}\lesssim(r/\delta)^{s}.

Then,

(4) δasδs.\delta^{-a}\lesssim_{s}\delta^{-s}.
Proof.

For each θΘ\theta\in\Theta^{\prime}, let 𝕋θ\mathbb{T}_{\theta} be a set of δ×1\delta\times 1 tubes that cover pθ1(pθ(A))B2(0,1)p_{\theta}^{-1}(p_{\theta}(A))\cap B^{2}(0,1) and hence cover AA. We can also assume that 𝕋θ\mathbb{T}_{\theta} satisfies a similar ss-dimensional condition that is inherited from pθ(A)p_{\theta}(A). For each θ\theta, let SθS_{\theta} be a δ1×1\delta^{-1}\times 1-tube centered at the origin whose longest direction forms an angle θ\theta with xx-axis. We see that SθS_{\theta} is dual to the tubes in 𝕋θ\mathbb{T}_{\theta}. Now, for each Tθ𝕋θ,T_{\theta}\in\mathbb{T}_{\theta}, we choose a bump function ψθ\psi_{\theta} satisfying the following properties: ψTθ1\psi_{T_{\theta}}\geq 1 on TθT_{\theta}, ψTθ\psi_{T_{\theta}} decays rapidly outside TθT_{\theta}, and suppψ^TθSθ\mathrm{supp}\widehat{\psi}_{T_{\theta}}\subset S_{\theta}.

Define functions

fθ:=Tθ𝕋θψTθandf=θΘfθ.f_{\theta}:=\sum_{T_{\theta}\in\mathbb{T}_{\theta}}\psi_{T_{\theta}}\ \ \ \ \textup{and}\ \ \ \ f=\sum_{\theta\in\Theta^{\prime}}f_{\theta}.

By definition, f(x)#Θf(x)\gtrsim\#\Theta^{\prime} for any xAx\in A, so we simply have

|A|(#Θ)2A|f|2.|A|(\#\Theta^{\prime})^{2}\lesssim\int_{A}|f|^{2}.

We can do better by performing a high-low decomposition for ff.

Let KK be a large number that will be determined later. Let ηlow(ξ)\eta_{low}(\xi) be a smooth bump function adapted to B2(0,K1δ1)B^{2}(0,K^{-1}\delta^{-1}) and let ηhigh=1ηlow\eta_{high}=1-\eta_{low}. We have the following high-low decomposition for ff:

f=flow+fhigh,f=f_{low}+f_{high},

where f^low=ηlowf^\widehat{f}_{low}=\eta_{low}\widehat{f} and f^high=ηhighf^\widehat{f}_{high}=\eta_{high}\widehat{f}. We will show that the high part dominates on AA. Actually, for xAx\in A, we have

(5) δ1f(x)|fhigh(x)|+|flow(x)|.\delta^{-1}\lesssim f(x)\leq|f_{high}(x)|+|f_{low}(x)|.

By definition, flow(x)=ηlowf(x)=ηlow(θΘTθ𝕋θψTθ)(x)f_{low}(x)=\eta_{low}^{\vee}*f(x)=\eta_{low}^{\vee}*\big{(}\sum_{\theta\in\Theta^{\prime}}\sum_{T_{\theta}\in\mathbb{T}_{\theta}}\psi_{T_{\theta}}\big{)}(x). Note that |ηlow||\eta_{low}^{\vee}| is morally an L1L^{1}-normalized bump function at B2(0,Kδ)B^{2}(0,K\delta). By the ss-dimensional condition of 𝕋θ\mathbb{T}_{\theta}, we have

ηlow(Tθ𝕋θψTθ)(x)K1#{Tθ𝕋θ:TθB2(x,CKδ)}Ks1.\eta_{low}^{\vee}*\big{(}\sum_{T_{\theta}\in\mathbb{T}_{\theta}}\psi_{T_{\theta}}\big{)}(x)\lesssim K^{-1}\#\{T_{\theta}\in\mathbb{T}_{\theta}:T_{\theta}\cap B^{2}(x,CK\delta)\neq\emptyset\}\lesssim K^{s-1}.

In the last step, we use the condition (3) in Proposition 1. As a result, we have

|flow(x)|#ΘKs1δ1Ks1.|f_{low}(x)|\lesssim\#\Theta^{\prime}K^{s-1}\lesssim\delta^{-1}K^{s-1}.

Since s<1s<1, by choosing KK large enough (depending on ss) and plugging into (5), we see that

|fhigh(x)|C1δ1|flow(x)|δ1.|f_{high}(x)|\geq C^{-1}\delta^{-1}-|f_{low}(x)|\gtrsim\delta^{-1}.

We obtain that

(6) |A|δ2A|fhigh|2.|A|\delta^{-2}\lesssim\int_{A}|f_{high}|^{2}.

Next, we will use a strong separation property for the high part. Note that f^high=θΘηhighf^θ\widehat{f}_{high}=\sum_{\theta\in\Theta^{\prime}}\eta_{high}\widehat{f}_{\theta}, and {supp(ηhighf^θ)}θ\{\mathrm{supp}(\eta_{high}\widehat{f}_{\theta})\}_{\theta} is at most O(K)O(K)-overlapped. We have

|fhigh|2=|f^high|2K2θΘ|ηhighf^θ|2K2θΘ|fθ|2K2δs.\int|f_{high}|^{2}=\int|\widehat{f}_{high}|^{2}\lesssim K^{2}\sum_{\theta\in\Theta^{\prime}}\int|\eta_{high}\widehat{f}_{\theta}|^{2}\lesssim K^{2}\sum_{\theta\in\Theta^{\prime}}\int|f_{\theta}|^{2}\lesssim K^{2}\delta^{-s}.

In the last inequality, we used the ss-dimensional condition of pθ(A)p_{\theta}(A).

As a result, we have

|A|δ2sδs,|A|\delta^{-2}\lesssim_{s}\delta^{-s},

which implies δasδs\delta^{-a}\lesssim_{s}\delta^{-s}. ∎

Notation.

  1. (1)

    For two positive real numbers R1R_{1} and R2R_{2}, we say that R1R2R_{1}\lesssim R_{2} if there exists a large constant CC, depending on relevant parameters, such that R1CR2R_{1}\leq CR_{2}; we say that R1R2R_{1}\ll R_{2} if R1R2/CR_{1}\leq R_{2}/C.

  2. (2)

    We use dim(E)\textup{dim}(E) for the Hausdorff dimension of EE.

  3. (3)

    For a given Borel measure μ\mu supported on 3\mathbb{R}^{3} and the projection πθ\pi_{\theta}, the pushforward measure πθ#μ\pi_{\theta\#}\mu, supported on γ(θ)\gamma(\theta)^{\perp}, is defined by

    (πθ#μ)(E):=μ((πθ)1(E)),(\pi_{\theta\#}\mu)(E):=\mu((\pi_{\theta})^{-1}(E)),

    for every Borel Eγ(θ)E\subset\gamma(\theta)^{\perp}.

  4. (4)

    Let μ\mu be a compactly supported Borel measure on 3\mathbb{R}^{3}. Take α>0\alpha>0. Define

    cα(μ):=supx3,r>0μ(B(x,r))rα,c_{\alpha}(\mu):=\sup_{x\in\mathbb{R}^{3},r>0}\frac{\mu(B(x,r))}{r^{\alpha}},

    where B(x,r)B(x,r) is the ball of radius rr centered at x3x\in\mathbb{R}^{3}.

  5. (5)

    For r1r\geq 1 and a rectangular box T3T\subset\mathbb{R}^{3}, we use rTrT to mean the dilation of TT by rr with respect to the center of TT, unless otherwise stated. For r>0r>0 and E3E\subset\mathbb{R}^{3}, we use rEr\cdot E to mean {rx:xE}\{rx:x\in E\}.

  6. (6)

    We often use m(E)m(E) the Lebesgue measure of the set E3E\subset\mathbb{R}^{3}; that is, m(E)=3(E)m(E)=\mathcal{H}^{3}(E).

  7. (7)

    By a measure we always assume that it is non-negative, unless stated otherwise.


Acknowledgement. S. Guo is partly supported by NSF-1800274 and NSF-2044828. L. Guth is supported by a Simons Investigator Award. H. Wang is supported by NSF Grant DMS-2055544. D. Maldague is supported by the NSF under Award No. 2103249.

2. Proof of Theorem 1

In this section we prove Theorem 1. First, we make some remarks on γ\gamma. We can cut γ\gamma into several small pieces and work on each of them. From now on, we assume that γ:[0,a]𝕊2\gamma:[0,a]\to\mathbb{S}^{2} is C2C^{2} and non-degenerate, and satisfies

(7) γ(0)=(0,0,1),γ(0)=(1,0,0),|γ(θ)|=1,θ[0,a].\gamma(0)=(0,0,1),\ \ \gamma^{\prime}(0)=(1,0,0),\ |\gamma^{\prime}(\theta)|=1,\forall\theta\in[0,a].

Here a>0a>0 is sufficiently small depending on γ\gamma. Since the parameter aa does not play any role here, we may pretend a=1a=1. We need some notation.

Definition 2 ((δ,s)(\delta,s)-set).

Let PnP\subset\mathbb{R}^{n} be a bounded set. Let δ>0\delta>0 be a dyadic number, and let 0sd0\leq s\leq d. We say that PP is a (δ,s)(\delta,s)-set if

|PBr|δ(r/δ)s,|P\cap B_{r}|_{\delta}\lesssim(r/\delta)^{s},

for any BrB_{r} being a ball of radius rr with δr1\delta\leq r\leq 1.

Let t\mathcal{H}^{t}_{\infty} denote the tt-dimensional Hausdorff content which is defined as

t(B):=inf{ir(Bi)t:BiBi}.\mathcal{H}^{t}_{\infty}(B):=\inf\{\sum_{i}r(B_{i})^{t}:B\subset\cup_{i}B_{i}\}.

Here, each BiB_{i} in the covering is a cube and r(Bi)r(B_{i}) is the length of the cube. We recall the following result (see [5] Lemma 3.13).

Lemma 1.

Let δ,s>0\delta,s>0, and BnB\subset\mathbb{R}^{n} with s(B):=κ>0\mathcal{H}_{\infty}^{s}(B):=\kappa>0. Then there exists a δ\delta-separated (δ,s)(\delta,s)-set PBP\subset B with cardinality #Pκδs\#P\gtrsim\kappa\delta^{-s}.

Our main effort will be devoted to the proof of the following theorem.

Theorem 4.

Fix 0<s<20<s<2. For each ε>0\varepsilon>0, there exists Cs,εC_{s,\varepsilon} so that the following holds. Let δ>0\delta>0. Let HB3(0,1)H\subset B^{3}(0,1) be a union of disjoint δ\delta-balls and we use #H\#H to denote the number of δ\delta-balls in HH. Let Θ\Theta be a δ\delta-separated subset of [0,1][0,1] such that Θ\Theta is a (δ,t)(\delta,t)-set and #Θ(logδ1)2δt\#\Theta\gtrsim(\log\delta^{-1})^{-2}\delta^{-t} for some t>0t>0. Assume for each θΘ\theta\in\Theta, we have a collection of δ×δ×1\delta\times\delta\times 1-tubes 𝕋θ\mathbb{T}_{\theta} pointing in direction γ(θ)\gamma(\theta). Each 𝕋θ\mathbb{T}_{\theta} satisfies the ss-dimensional condition:

  1. (1)

    #𝕋θδs\#\mathbb{T}_{\theta}\lesssim\delta^{-s},

  2. (2)

    #{T𝕋θ:TBr}(rδ)s\#\{T\in\mathbb{T}_{\theta}:T\cap B_{r}\}\lesssim(\frac{r}{\delta})^{s}, for any BrB_{r} being a ball of radius rr (δr1)(\delta\leq r\leq 1).

We also assume that each δ\delta-ball contained in HH intersects (logδ1)2#Θ\gtrsim(\log\delta^{-1})^{-2}\#\Theta many tubes from θΘ𝕋θ\cup_{\theta\in\Theta}\mathbb{T}_{\theta}. Then

#Θ#HCs,εδ1sε.\#\Theta\#H\leq C_{s,\varepsilon}\delta^{-1-s-\varepsilon}.

2.1. δ\delta-discretization of the projection problem

In this subsection we show how Theorem 4 implies Theorem 1. Before starting the proof, we state a very useful lemma. We use the following notation. For any δ=2k\delta=2^{-k} (k+k\in\mathbb{N}^{+}), let 𝒟δ\mathcal{D}_{\delta} denote the lattice of δ\delta-squares in [0,1]2[0,1]^{2}. For technical reasons, we remove the top edge and the right edge of each δ\delta-square so that they are disjoint.

Lemma 2.

Suppose X[0,1]2X\subset[0,1]^{2} with dimX<s\textup{dim}X<s. Then for any ε>0\varepsilon>0, there exist dyadic squares 𝒞2k𝒟2k\mathcal{C}_{2^{-k}}\subset\mathcal{D}_{2^{-k}} (k>0)(k>0) so that

  1. (1)

    Xk>0D𝒞2kD,X\subset\bigcup_{k>0}\bigcup_{D\in\mathcal{C}_{2^{-k}}}D,

  2. (2)

    k>0D𝒞2kr(D)sε\sum_{k>0}\sum_{D\in\mathcal{C}_{2^{-k}}}r(D)^{s}\leq\varepsilon,

  3. (3)

    𝒞2k\mathcal{C}_{2^{-k}} satisfies the ss-dimensional condition: For l<kl<k and any D𝒟2lD\in\mathcal{D}_{2^{-l}}, we have #{D𝒞2k:DD}2(kl)s\#\{D^{\prime}\in\mathcal{C}_{2^{-k}}:D^{\prime}\subset D\}\leq 2^{(k-l)s}.

Proof of the lemma.

Consider all the covering 𝒞\mathcal{C} of XX by dyadic lattice squares that satisfy condition (1), (2) in Lemma 2, i.e., 𝒞k>0𝒟2k\mathcal{C}\subset\bigcup_{k>0}\mathcal{D}_{2^{-k}}, XD𝒞DX\subset\bigcup_{D\in\mathcal{C}}D and D𝒞r(D)sε\sum_{D\in\mathcal{C}}r(D)^{s}\leq\varepsilon. We also assume all the dyadic squares in 𝒞\mathcal{C} are disjoint. We will define an order “<<” between any two of such coverings 𝒞,𝒞\mathcal{C},\mathcal{C}^{\prime}. First, we define the kk-th covering number of 𝒞\mathcal{C} by

ck(𝒞):=#(𝒞𝒟2k),c_{k}(\mathcal{C}):=\#(\mathcal{C}\cap\mathcal{D}_{2^{-k}}),

which is the number of 2k2^{-k}-squares in the covering 𝒞\mathcal{C}.

We say 𝒞<𝒞\mathcal{C}<\mathcal{C}^{\prime}, if they satisfy: (1) There is a maximal k00k_{0}\geq 0 such that 𝒞𝒟2k=𝒞𝒟2k\mathcal{C}\cap\mathcal{D}_{2^{-k}}=\mathcal{C}^{\prime}\cap\mathcal{D}_{2^{-k}} (k<k0k<k_{0}), and 𝒞𝒟2k0𝒞𝒟2k0\mathcal{C}\cap\mathcal{D}_{2^{-k_{0}}}\subset\mathcal{C}^{\prime}\cap\mathcal{D}_{2^{-k_{0}}}; (2) For any xXx\in X, the square in 𝒞\mathcal{C}^{\prime} that covers xx contains the square in 𝒞\mathcal{C} that covers xx. It is not hard to check the transitivity: If 𝒞<𝒞\mathcal{C}<\mathcal{C}^{\prime} and 𝒞<𝒞′′\mathcal{C}^{\prime}<\mathcal{C}^{\prime\prime}, then 𝒞<𝒞′′\mathcal{C}<\mathcal{C}^{\prime\prime}.

Suppose 𝒞\mathcal{C} is a covering that is maximal with respect to the order <<. Then we can show that 𝒞\mathcal{C} satisfies condition (3)(3) in Lemma 2. Suppose by contradiction, there exist l<kl<k and D𝒟2lD\in\mathcal{D}_{2^{-l}} so that

(8) #{D𝒞𝒟2k:DD}>2(kl)s.\#\{D^{\prime}\in\mathcal{C}\cap\mathcal{D}_{2^{-k}}:D^{\prime}\subset D\}>2^{(k-l)s}.

We define another covering 𝒞\mathcal{C}^{\prime} by adding DD to 𝒞\mathcal{C} and deleting {D𝒞{D}:DD}\{D^{\prime}\in\mathcal{C}\setminus\{D\}:D^{\prime}\subset D\} from 𝒞\mathcal{C}. It is easy to check that 𝒞\mathcal{C}^{\prime} is still a covering of XX. By (8), we can also check D𝒞r(D)s<D𝒞r(D)sε\sum_{D\in\mathcal{C}^{\prime}}r(D)^{s}<\sum_{D\in\mathcal{C}}r(D)^{s}\leq\varepsilon, so 𝒞\mathcal{C}^{\prime} satisfies (2)(2) in Lemma 2. However, 𝒞<𝒞\mathcal{C}<\mathcal{C}^{\prime} which contradicts the maximality of 𝒞\mathcal{C}.

Now, it suffices to find a maximal element among all the coverings that satisfy condition (1),(2)(1),(2) in Lemma 2. First of all, such covering exists by the definition of Hausdorff dimension and dimX<s\textup{dim}X<s. By Zorn’s lemma, it suffices to find an upper bound for any ascending chain.

Let {𝒞j}jJ\{\mathcal{C}_{j}\}_{j\in J} be an infinite chain of coverings of XX. Define

𝒞=jJiJ𝒞i𝒞j𝒞i.\mathcal{C}=\bigcap_{j\in J}\bigcup_{\begin{subarray}{c}i\in J\\ \mathcal{C}_{i}\geq\mathcal{C}_{j}\end{subarray}}\mathcal{C}_{i}.

We show that 𝒞\mathcal{C} is an upper bound of the chain. First, we show that 𝒞\mathcal{C} covers XX. For xXx\in X and jj, let Dx(j)D^{(j)}_{x} be the largest dyadic square in iJ𝒞i𝒞j𝒞i\bigcup_{\begin{subarray}{c}i\in J\\ \mathcal{C}_{i}\geq\mathcal{C}_{j}\end{subarray}}\mathcal{C}_{i} containing xx. By the definition of the partial order and the fact that chains are totally ordered, Dx(j)=DxD^{(j)}_{x}=D_{x} is independent of jj, and thus Dx𝒞D_{x}\in\mathcal{C}. This shows that 𝒞\mathcal{C} is a covering of XX. It also shows that the squares in 𝒞\mathcal{C} are disjoint. Let KK\in\mathbb{N}. Choose jJj\in J such that 𝒞i𝒟2k=𝒞j𝒟2k\mathcal{C}_{i}\cap\mathcal{D}_{2^{-k}}=\mathcal{C}_{j}\cap\mathcal{D}_{2^{-k}} for all 0kK0\leq k\leq K and all 𝒞i𝒞j\mathcal{C}_{i}\geq\mathcal{C}_{j}. Then

k=0KD𝒞𝒟2kr(D)sk=0KD𝒞j𝒟2kr(D)sε.\sum_{k=0}^{K}\sum_{D\in\mathcal{C}\cap\mathcal{D}_{2^{-k}}}r(D)^{s}\leq\sum_{k=0}^{K}\sum_{D\in\mathcal{C}_{j}\cap\mathcal{D}_{2^{-k}}}r(D)^{s}\leq\varepsilon.

Letting KK\to\infty gives

D𝒞r(D)sε.\sum_{D\in\mathcal{C}}r(D)^{s}\leq\varepsilon.

So, 𝒞\mathcal{C} satisfies condition (2)(2). By definition, it is easy to check 𝒞i𝒞\mathcal{C}_{i}\leq\mathcal{C} for every 𝒞i\mathcal{C}_{i} in the initial chain. This proves that 𝒞\mathcal{C} is an upper bound of the chain. ∎

Remark 1.

Besides [0,1]2[0,1]^{2}, this lemma holds for other compact metric spaces, for example [0,1]n[0,1]^{n} or 𝕊2\mathbb{S}^{2}. The proof is exactly the same.

Proof that Theorem 4 implies Theorem 1.

Suppose A3A\subset\mathbb{R}^{3} is a Borel set of Hausdorff dimension α\alpha. We may assume AB3(0,1)A\subset B^{3}(0,1). Recall the definition of the exceptional set

Es:={θ[0,1]:dimπθ(A)<s}.E_{s}:=\{\theta\in[0,1]:\textup{dim}\pi_{\theta}(A)<s\}.

Recall the definition of the tt-dimensional Hausdorff content is given by

t(B):=inf{ir(Bi)t:BiBi}.\mathcal{H}^{t}_{\infty}(B):=\inf\{\sum_{i}r(B_{i})^{t}:B\subset\cup_{i}B_{i}\}.

A property for the Hausdorff content is that

dim(B)=sup{t:t(B)>0}.\textup{dim}(B)=\sup\{t:\mathcal{H}^{t}_{\infty}(B)>0\}.

We choose a<dim(A),t<dim(Es)a<\textup{dim}(A),t<\textup{dim}(E_{s}). Then t(Es)>0\mathcal{H}_{\infty}^{t}(E_{s})>0, and by Frostman’s lemma there exists a probability measure νA\nu_{A} supported on AA satisfying νA(Br)ra\nu_{A}(B_{r})\lesssim r^{a} for any BrB_{r} being a ball of radius rr. We may assume t>0t>0, otherwise dim(Es)=0\textup{dim}(E_{s})=0. We only need to prove

a1+st,a\leq 1+s-t,

since then we can send adim(A)a\rightarrow\textup{dim}(A) and tdim(Es)t\rightarrow\textup{dim}(E_{s}). As aa and tt are fixed, we may assume t(Es)1\mathcal{H}_{\infty}^{t}(E_{s})\sim 1 is a constant.

Fix a θEs\theta\in E_{s}. By definition, we have dimπθ(A)<s\textup{dim}\pi_{\theta}(A)<s. We also fix a small number ϵ\epsilon_{\circ} which we will later send to 0. By Lemma 2, we can find a covering of πθ(A)\pi_{\theta}(A) by disks 𝔻θ={D}\mathbb{D}_{\theta}=\{D\}, each of which has radius 2j2^{-j} for some integer j>|log2ϵ|j>|\log_{2}\epsilon_{\circ}|. We define 𝔻θ,j:={D𝔻θ:r(D)=2j}\mathbb{D}_{\theta,j}:=\{D\in\mathbb{D}_{\theta}:r(D)=2^{-j}\}. Lemma 2 yields the following properties:

(9) D𝔻θr(D)s<1;\sum_{D\in\mathbb{D}_{\theta}}r(D)^{s}<1;

For each jj and rr-ball BrVθB_{r}\subset V_{\theta} with 2jr12^{-j}\leq r\leq 1, we have

(10) #{D𝔻θ,j:DBr}(r2j)s.\#\{D\in\mathbb{D}_{\theta,j}:D\subset B_{r}\}\lesssim(\frac{r}{2^{-j}})^{s}.

For each θEs\theta\in E_{s}, we can find such a 𝔻θ\mathbb{D}_{\theta}. We also define the tube sets 𝕋θ,j:={πθ1(D):D𝔻θ,j}B3(0,1)\mathbb{T}_{\theta,j}:=\{\pi^{-1}_{\theta}(D):D\in\mathbb{D}_{\theta,j}\}\cap B^{3}(0,1), 𝕋θ=j𝕋θ,j\mathbb{T}_{\theta}=\bigcup_{j}\mathbb{T}_{\theta,j}. Each tube in 𝕋θ,j\mathbb{T}_{\theta,j} has dimensions 2j×2j×12^{-j}\times 2^{-j}\times 1 and direction γ(θ)\gamma(\theta). One easily sees that AT𝕋θTA\subset\bigcup_{T\in\mathbb{T}_{\theta}}T. By pigeonholing, there exists j(θ)j(\theta) such that

(11) νA(A(T𝕋θ,j(θ)T))110j(θ)2νA(A)=110j(θ)2.\nu_{A}(A\cap(\cup_{T\in\mathbb{T}_{\theta,j(\theta)}}T))\geq\frac{1}{10j(\theta)^{2}}\nu_{A}(A)=\frac{1}{10j(\theta)^{2}}.

For each j>|log2ϵ|j>|\log_{2}\epsilon_{\circ}|, define Es,j:={θEs:j(θ)=j}E_{s,j}:=\{\theta\in E_{s}:j(\theta)=j\}. Then we obtain a partition of EsE_{s}:

Es=jEs,j.E_{s}=\bigsqcup_{j}E_{s,j}.

By pigeonholing again, there exists jj such that

(12) t(Es,j)110j2t(Es)110j2.\mathcal{H}_{\infty}^{t}(E_{s,j})\geq\frac{1}{10j^{2}}\mathcal{H}_{\infty}^{t}(E_{s})\sim\frac{1}{10j^{2}}.

In the rest of the poof, we fix this jj. We also set δ=2j(<ϵ)\delta=2^{-j}(<\epsilon_{\circ}). By Lemma 1, there exists a δ\delta-separated (δ,t)(\delta,t)-set ΘEs,j\Theta\subset E_{s,j} with cardinality #Θ(logδ1)2δt\#\Theta\gtrsim(\log\delta^{-1})^{-2}\delta^{-t}.

Next, we consider the set S:={(x,θ)A×Θ:xT𝕋θ,jT}S:=\{(x,\theta)\in A\times\Theta:x\in\cup_{T\in\mathbb{T}_{\theta,j}}T\}. We also use μ\mu to denote the counting measure on Θ\Theta. Define the section of SS:

Sx={θ:(x,θ)S},Sθ:={x:(x,θ)S}.S_{x}=\{\theta:(x,\theta)\in S\},\ \ \ S_{\theta}:=\{x:(x,\theta)\in S\}.

By (11) and Fubini, we have

(13) (νA×μ)(S)110j2μ(Θ).(\nu_{A}\times\mu)(S)\geq\frac{1}{10j^{2}}\mu(\Theta).

This implies

(14) (νA×μ)({(x,θ)S:μ(Sx)120j2μ(Θ)})120j2μ(Θ).(\nu_{A}\times\mu)\bigg{(}\Big{\{}(x,\theta)\in S:\mu(S_{x})\geq\frac{1}{20j^{2}}\mu(\Theta)\Big{\}}\bigg{)}\geq\frac{1}{20j^{2}}\mu(\Theta).

since

(15) (νA×μ)({(x,θ)S:μ(Sx)120j2μ(Θ)})120j2a(A)μ(Θ).(\nu_{A}\times\mu)\bigg{(}\Big{\{}(x,\theta)\in S:\mu(S_{x})\leq\frac{1}{20j^{2}}\mu(\Theta)\Big{\}}\bigg{)}\leq\frac{1}{20j^{2}}\mathcal{H}_{\infty}^{a}(A)\mu(\Theta).

By (14), we have

(16) νA({xA:μ(Sx)120j2μ(Θ)})120j2.\nu_{A}\bigg{(}\Big{\{}x\in A:\mu(S_{x})\geq\frac{1}{20j^{2}}\mu(\Theta)\Big{\}}\bigg{)}\geq\frac{1}{20j^{2}}.

We are ready to apply Theorem 4. Recall δ=2j\delta=2^{-j} and #Θ(logδ1)2δt\#\Theta\gtrsim(\log\delta^{-1})^{-2}\delta^{-t}. By (16), we can find a δ\delta-separated subset of {xA:#Sx120j2#Θ}\{x\in A:\#S_{x}\geq\frac{1}{20j^{2}}\#\Theta\} with cardinality (logδ1)2δa\gtrsim(\log\delta^{-1})^{-2}\delta^{-a}. We denote the δ\delta-neighborhood of this set by HH, which is a union of δ\delta-balls. For each δ\delta-ball BδB_{\delta} contained in HH, we see that there are (logδ1)2#Θ\gtrsim(\log\delta^{-1})^{-2}\#\Theta many tubes from θΘ𝕋θ,j\cup_{\theta\in\Theta}\mathbb{T}_{\theta,j} that intersect BδB_{\delta}. We can now apply Theorem 4 to obtain

(logδ1)4δat#Θ#HCs,εδ1sε.(\log\delta^{-1})^{-4}\delta^{-a-t}\lesssim\#\Theta\#H\leq C_{s,\varepsilon}\delta^{-1-s-\varepsilon}.

Letting ϵ0\epsilon_{\circ}\rightarrow 0 (and hence δ0\delta\rightarrow 0) and then ε0\varepsilon\rightarrow 0, we obtain a+t1+sa+t\leq 1+s.

2.2. Proof of Theorem 4

The proof of Theorem 4 is base on the L6L^{6} decoupling inequality for cone which is well-understood. For convenience, we will prove the following version of Theorem 4 after rescaling xδ1xx\mapsto\delta^{-1}x.

Theorem 5.

Fix 0<s<20<s<2. For each ε>0\varepsilon>0, there exists Cs,εC_{s,\varepsilon} so that the following holds. Let δ>0\delta>0. Let HB3(0,δ1)H\subset B^{3}(0,\delta^{-1}) be a union of δa\delta^{-a} many disjoint unit balls so that HH has measure |H|δa|H|\sim\delta^{-a}. Let Θ\Theta be a δ\delta-separated subset of [0,1][0,1] so that Θ\Theta is a (δ,t)(\delta,t)-set and #Θ(logδ1)2δt\#\Theta\gtrsim(\log\delta^{-1})^{-2}\delta^{-t}. Assume for each θΘ\theta\in\Theta, we have a collection of 1×1×δ11\times 1\times\delta^{-1}-tubes 𝕋θ\mathbb{T}_{\theta} pointing in direction γ(θ)\gamma(\theta). 𝕋θ\mathbb{T}_{\theta} satisfies the ss-dimensional condition:

  1. (1)

    #𝕋θδs\#\mathbb{T}_{\theta}\lesssim\delta^{-s},

  2. (2)

    #{T𝕋θ:TBr}rs\#\{T\in\mathbb{T}_{\theta}:T\cap B_{r}\}\lesssim r^{s}, for any BrB_{r} being a ball of radius rr (1rδ1)(1\leq r\leq\delta^{-1}).

We also assume that each unit ball contained in HH intersects |logδ1|2#Θ\gtrsim|\log\delta^{-1}|^{-2}\#\Theta many tubes from θ𝕋θ\cup_{\theta}\mathbb{T}_{\theta}. Then

δtaCs,εδ1sε.\delta^{-t-a}\leq C_{s,\varepsilon}\delta^{-1-s-\varepsilon}.

We first discuss the geometry of γ\gamma. Let γ\gamma be the non-degenerate curve as discussed in the beginning of this section. We have |γ(θ)|=1|\gamma^{\prime}(\theta)|=1. For convenience, we define

(17) 𝐞1(θ):=γ(θ),𝐞2(θ):=γ(θ),𝐞3(θ):=γ(θ)×γ(θ).{\mathbf{e}}_{1}(\theta):=\gamma(\theta),\ \ {\mathbf{e}}_{2}(\theta):=\gamma^{\prime}(\theta),\ \ {\mathbf{e}}_{3}(\theta):=\gamma(\theta)\times\gamma^{\prime}(\theta).

We see that {𝐞1(θ),𝐞2(θ),𝐞3(θ)}\{{\mathbf{e}}_{1}(\theta),{\mathbf{e}}_{2}(\theta),{\mathbf{e}}_{3}(\theta)\} form a Frenet coordinate along γ\gamma. Define the corresponding conical surface Γ:={r𝐞3(θ):1/2r1,θ[0,1]}\Gamma:=\{r{\mathbf{e}}_{3}(\theta):1/2\leq r\leq 1,\theta\in[0,1]\}.

We first show that Γ\Gamma satisfies the same non-degenerate condition as the standard cone. Note that we have the following formulae for the Frenet coordinate:

(18) 𝐞1(θ)=𝐞2(θ),\displaystyle{\mathbf{e}}_{1}^{\prime}(\theta)={\mathbf{e}}_{2}(\theta),
(19) 𝐞2(θ)=𝐞1(θ)+κ(θ)𝐞3(θ),\displaystyle{\mathbf{e}}_{2}^{\prime}(\theta)=-{\mathbf{e}}_{1}(\theta)+\kappa(\theta){\mathbf{e}}_{3}(\theta),
(20) 𝐞3(θ)=κ(θ)𝐞2(θ),\displaystyle{\mathbf{e}}_{3}^{\prime}(\theta)=-\kappa(\theta){\mathbf{e}}_{2}(\theta),

where κ(θ)=𝐞2(θ),𝐞3(θ)>0\kappa(\theta)=\langle{\mathbf{e}}_{2}^{\prime}(\theta),{\mathbf{e}}_{3}(\theta)\rangle>0.

First, we show that Γ\Gamma is a C2C^{2} surface. We will do this by finding a reparametrization s=s(θ)s=s(\theta) so that 𝐞3(θ(s)){\mathbf{e}}_{3}(\theta(s)) is a C2C^{2} function of ss. Choose

s(θ)=0θκ(t)𝑑t,s(\theta)=\int_{0}^{\theta}\kappa(t)dt,

and then dθds=κ(θ)1\frac{d\theta}{ds}=\kappa(\theta)^{-1}. We have

dds𝐞3=dθdsddθ𝐞3=𝐞2.\frac{d}{ds}{\mathbf{e}}_{3}=\frac{d\theta}{ds}\cdot\frac{d}{d\theta}{\mathbf{e}}_{3}=-{\mathbf{e}}_{2}.

Since θ=θ(s)\theta=\theta(s) is C1C^{1}, we have dds𝐞3=𝐞2(θ(s))\frac{d}{ds}{\mathbf{e}}_{3}=-{\mathbf{e}}_{2}(\theta(s)) is C1C^{1} with respect to ss, and therefore 𝐞3(θ(s)){\mathbf{e}}_{3}(\theta(s)) is C2C^{2} with respect to ss. Moreover, det(𝐞3(θ(s)),dds𝐞3(θ(s)),d2ds2𝐞3(θ(s)))=θ(s)det(𝐞3(θ(s)),𝐞2(θ(s)),𝐞1(θ(s)))\det\left(\mathbf{e}_{3}(\theta(s)),\frac{d}{ds}\mathbf{e}_{3}(\theta(s)),\frac{d^{2}}{ds^{2}}\mathbf{e}_{3}(\theta(s))\right)=\theta^{\prime}(s)\det(\mathbf{e}_{3}(\theta(s)),-\mathbf{e}_{2}(\theta(s)),\mathbf{e}_{1}(\theta(s))) by the above, which is nonvanishing since θ(s)\theta^{\prime}(s) is nonvanishing.


For any large scale RR, there is a standard partition of NR1ΓN_{R^{-1}}\Gamma into planks σ\sigma of dimensions R1×R1/2×1R^{-1}\times R^{-1/2}\times 1:

NR1Γ=σ.N_{R^{-1}}\Gamma=\bigcup\sigma.

For any Schwartz function ff, we define fσ:=(1σf^)f_{\sigma}:=(1_{\sigma}\widehat{f})^{\vee} as usual. We have the following L6L^{6}-decoupling inequality for these planks.

Theorem 6 (Bourgain-Demeter [1]).

For any Schwartz ff with f^NR1Γ\widehat{f}\subset N_{R^{-1}}\Gamma, we have

(21) f6εRε(σ:R1×R1/2×1fσ62)1/2.\|f\|_{6}\lesssim_{\varepsilon}R^{\varepsilon}\big{(}\sum_{\sigma:R^{-1}\times R^{-1/2}\times 1}\|f_{\sigma}\|_{6}^{2}\big{)}^{1/2}.
Remark 2.

We will actually apply Theorem 6 to a slightly different cone

(22) ΓK1={r𝐞3(θ):K1r1,θ[0,1]},\Gamma_{K^{-1}}=\{r{\mathbf{e}}_{3}(\theta):K^{-1}\leq r\leq 1,\theta\in[0,1]\},

for some K(logδ1)O(1)K\sim(\log\delta^{-1})^{O(1)}. Compared with Γ\Gamma, we see that ΓK1\Gamma_{K^{-1}} is at distance K1K^{-1} from the origin, but we still have a similar decoupling inequality. Instead of (21), we have

(23) f6εKO(1)Rε(σ:R1×R1/2×1fσ62)1/2.\|f\|_{6}\lesssim_{\varepsilon}K^{O(1)}R^{\varepsilon}\big{(}\sum_{\sigma:R^{-1}\times R^{-1/2}\times 1}\|f_{\sigma}\|_{6}^{2}\big{)}^{1/2}.

The idea is to partition ΓK1\Gamma_{K^{-1}} into O(K)\sim O(K) many parts, each of which is roughly a cone for which we can apply Theorem 6. By triangle inequality, this results in an additional factor KO(1)K^{O(1)}. It turns out that this factor is not harmful, since we will set K(logR)O(1)K\sim(\log R)^{O(1)} which can be absorbed into RεR^{\varepsilon}.

We are ready to prove Theorem 5.

Proof of Theorem 5.

Recall that 𝕋θ\mathbb{T}_{\theta} is a collection of 1×1×δ11\times 1\times\delta^{-1}-tubes pointing to direction γ(θ)=𝐞1(θ)\gamma(\theta)={\mathbf{e}}_{1}(\theta). We consider the dual of each TθT_{\theta} in the frequency space. For each θ\theta, we define PθP_{\theta} to be a slab centered at the origin that has dimensions 1×1×δ1\times 1\times\delta, and its shortest direction is parallel to 𝐞1(θ){\mathbf{e}}_{1}(\theta). We see that PθP_{\theta} is the dual rectangle of each Tθ𝕋θT_{\theta}\in\mathbb{T}_{\theta}. Now, for each Tθ𝕋θT_{\theta}\in\mathbb{T}_{\theta}, we choose a bump function ψTθ\psi_{T_{\theta}} satisfying the following properties: ψTθ1\psi_{T_{\theta}}\geq 1 on TθT_{\theta}, ψTθ\psi_{T_{\theta}} decays rapidly outside TθT_{\theta}, and suppψ^TθPθ\mathrm{supp}\widehat{\psi}_{T_{\theta}}\subset P_{\theta}.

Define functions

fθ=Tθ𝕋θψTθandf=θΘfθ.f_{\theta}=\sum_{T_{\theta}\in\mathbb{T}_{\theta}}\psi_{T_{\theta}}\qquad\text{and}\qquad f=\sum_{\theta\in\Theta}f_{\theta}.

From our definitions, we see that for any xHx\in H, we have f(x)#{Tθ𝕋θ:xT}(logδ1)2#Θf(x)\gtrsim\#\{T\in\cup_{\theta}\mathbb{T}_{\theta}:x\in T\}\gtrsim(\log\delta^{-1})^{-2}\#\Theta. Therefore, we obtain

(24) |H|(logδ1)2p(#Θ)p2pH|f|p,|H|(\log\delta^{-1})^{-2p}(\#\Theta)^{p}\leq 2^{p}\int_{H}|f|^{p},

for any pp. For our purpose, we just choose p=6p=6, so we have

(25) |H|(logδ1)12(#Θ)626H|f|6,|H|(\log\delta^{-1})^{-12}(\#\Theta)^{6}\leq 2^{6}\int_{H}|f|^{6},

Our goal is to find an upper bound for the right hand side of (25). We will decompose PθP_{\theta} into pieces and estimate the contribution of f^θ\widehat{f}_{\theta} from each piece.

Let us discuss the decomposition for PθP_{\theta}. Recall that PθP_{\theta} is a 1×1×δ1\times 1\times\delta-slab centered at the origin with normal direction 𝐞1(θ){\mathbf{e}}_{1}(\theta). Recall (17), we can write Pθ={i=13ξi𝐞i(θ):|ξ1|δ,|ξ2|1,|ξ3|1}P_{\theta}=\{\sum_{i=1}^{3}\xi_{i}{\mathbf{e}}_{i}(\theta):|\xi_{1}|\leq\delta,|\xi_{2}|\leq 1,|\xi_{3}|\leq 1\}.

Pθ,highP_{\theta,high}Pθ,lowP_{\theta,low}Pθ,δ1/2P_{\theta,\delta^{1/2}}Pθ,λP_{\theta,\lambda}𝒆2(θ)\boldsymbol{e}_{2}(\theta)𝒆1(θ)\boldsymbol{e}_{1}(\theta)𝒆3(θ)\boldsymbol{e}_{3}(\theta)
Figure 1. High-low decomposition for PθP_{\theta}
Definition 3.

See Figure 1. Let KK be a large number which we will choose later. (Actually, we will choose K(logδ1)O(1)K\sim(\log\delta^{-1})^{O(1)}) Define the high part of PθP_{\theta} as

Pθ,high:={i=13ξi𝐞i(θ):|ξ1|δ,K1|ξ2|1,|ξ3|1}.P_{\theta,high}:=\{\sum_{i=1}^{3}\xi_{i}{\mathbf{e}}_{i}(\theta):|\xi_{1}|\leq\delta,K^{-1}\leq|\xi_{2}|\leq 1,|\xi_{3}|\leq 1\}.

Define the low part of PθP_{\theta} as

Pθ,low:={i=13ξi𝐞i(θ):|ξ1|δ,|ξ2|K1,|ξ3|K1}.P_{\theta,low}:=\{\sum_{i=1}^{3}\xi_{i}{\mathbf{e}}_{i}(\theta):|\xi_{1}|\leq\delta,|\xi_{2}|\leq K^{-1},|\xi_{3}|\leq K^{-1}\}.

For dyadic numbers λ(δ1/2,K1]\lambda\in(\delta^{1/2},K^{-1}], define

Pθ,λ:={i=13ξi𝐞i(θ):|ξ1|δ,12λ|ξ2|λ,K1|ξ3|1}.P_{\theta,\lambda}:=\{\sum_{i=1}^{3}\xi_{i}{\mathbf{e}}_{i}(\theta):|\xi_{1}|\leq\delta,\frac{1}{2}\lambda\leq|\xi_{2}|\leq\lambda,K^{-1}\leq|\xi_{3}|\leq 1\}.

In particular, we define

Pθ,δ1/2:={i=13ξi𝐞i(θ):|ξ1|δ,|ξ2|δ1/2,K1|ξ3|1}.P_{\theta,\delta^{1/2}}:=\{\sum_{i=1}^{3}\xi_{i}{\mathbf{e}}_{i}(\theta):|\xi_{1}|\leq\delta,|\xi_{2}|\leq\delta^{1/2},K^{-1}\leq|\xi_{3}|\leq 1\}.
Remark 3.

We obtain a partition of PθP_{\theta} as

Pθ=Pθ,highPθ,lowλPθ,λ.P_{\theta}=P_{\theta,high}\bigsqcup P_{\theta,low}\bigsqcup_{\lambda}P_{\theta,\lambda}.

We see that Pθ,λP_{\theta,\lambda} consists of four planks of dimensions δ×λ×1\sim\delta\times\lambda\times 1 whose longest side is along direction 𝐞3(θ){\mathbf{e}}_{3}(\theta). Here λ\lambda plays a role of angular parameter in the sense that Pθ,λP_{\theta,\lambda} are roughly those points in PθPθ,lowP_{\theta}\setminus P_{\theta,low} so that the lines connecting them with the origin form an angle λ\sim\lambda with 𝐞3(θ){\mathbf{e}}_{3}(\theta).

We choose a smooth partition of unity adapted to this covering which we denote by ηθ,high,ηθ,low,ηθ,λ\eta_{\theta,high},\eta_{\theta,low},\eta_{\theta,\lambda}, so that

(26) ηθ,high+ηθ,low+δ1/2λK1ηθ,λ=1\eta_{\theta,high}+\eta_{\theta,low}+\sum_{\delta^{1/2}\leq\lambda\leq K^{-1}}\eta_{\theta,\lambda}=1

on PθP_{\theta}. Since suppf^θPθ\mathrm{supp}\widehat{f}_{\theta}\subset P_{\theta}, we also obtain a decomposition of fθf_{\theta}

(27) fθ=fθ,high+fθ,low+δ1/2λK1fθ,λ,f_{\theta}=f_{\theta,high}+f_{\theta,low}+\sum_{\delta^{1/2}\leq\lambda\leq K^{-1}}f_{\theta,\lambda},

where f^θ,high=ηθ,highf^θ,f^θ,low=ηθ,lowf^θ,f^θ,λ=ηθ,λf^θ.\widehat{f}_{\theta,high}=\eta_{\theta,high}\widehat{f}_{\theta},\widehat{f}_{\theta,low}=\eta_{\theta,low}\widehat{f}_{\theta},\widehat{f}_{\theta,\lambda}=\eta_{\theta,\lambda}\widehat{f}_{\theta}. Similarly, we have a decomposition of ff

(28) f=fhigh+flow+δ1/2λK1fλ,f=f_{high}+f_{low}+\sum_{\delta^{1/2}\leq\lambda\leq K^{-1}}f_{\lambda},

where fhigh=θfθ,high,flow=θfθ,low,fλ=θfθ,λ.f_{high}=\sum_{\theta}f_{\theta,high},f_{low}=\sum_{\theta}f_{\theta,low},f_{\lambda}=\sum_{\theta}f_{\theta,\lambda}.

Recalling (25) and using triangle inequality, we have

(29) |H|(logδ1)12(#Θ)6H|flow|6+H|fhigh|6+(logδ1)O(1)λH|fλ|6.|H|(\log\delta^{-1})^{-12}(\#\Theta)^{6}\lesssim\int_{H}|f_{low}|^{6}+\int_{H}|f_{high}|^{6}+(\log\delta^{-1})^{O(1)}\sum_{\lambda}\int_{H}|f_{\lambda}|^{6}.

We will discuss three cases depending on which term on the right hand side of (29) dominates.

Case 1: Low case

If the first term on the right hand side of (29) dominates, we say we are in the low case. Actually, we will see that we are never in the low case by showing

(30) H|flow|6C1|H|(logδ1)12(#Θ)6,\int_{H}|f_{low}|^{6}\leq C^{-1}|H|(\log\delta^{-1})^{-12}(\#\Theta)^{6},

for some large constant CC. This means the low term on the right hand side of (29) will not dominate. By properly choosing KK, we can show a pointwise bound for flowf_{low}:

(31) |flow(x)|C1(logδ1)2#Θ.|f_{low}(x)|\leq C^{-1}(\log\delta^{-1})^{-2}\#\Theta.

This will immediately imply (30). Let us focus on (31).

Recall that flow=θfθ,low=θfθηθ,lowf_{low}=\sum_{\theta}f_{\theta,low}=\sum_{\theta}f_{\theta}*\eta_{\theta,low}^{\vee}. Since ηθ,low\eta_{\theta,low} is a bump function at Pθ,lowP_{\theta,low}, we see that ηθ,low\eta^{\vee}_{\theta,low} is a bump function essentially supported in the dual of Pθ,lowP_{\theta,low}. Denote the dual of Pθ,lowP_{\theta,low} by Tθ,KT_{\theta,K} which is a K×K×δ1K\times K\times\delta^{-1}-tube parallel to 𝐞1(θ){\mathbf{e}}_{1}(\theta). One has

|ηθ,low|1|Tθ,K|ψTθ,K.|\eta_{\theta,low}^{\vee}|\lesssim\frac{1}{|T_{\theta,K}|}\psi_{T_{\theta,K}}.

Here ψTθ,K\psi_{T_{\theta,K}} is a bump function =1=1 on Tθ,KT_{\theta,K} and decays rapidly outside Tθ,KT_{\theta,K}.

By definition, fθ=TθψTθf_{\theta}=\sum_{T_{\theta}}\psi_{T_{\theta}} is the sum of bump function of tubes. We have

(32) |flow|θTθ𝕋θψTθ1|Tθ,K|ψTθ,K.\displaystyle|f_{low}|\lesssim\sum_{\theta}\sum_{T_{\theta}\in\mathbb{T}_{\theta}}\psi_{T_{\theta}}*\frac{1}{|T_{\theta,K}|}\psi_{T_{\theta,K}}.

If we ignore the rapidly decaying tails, we have

(33) |flow(x)|θ1K2#{Tθ𝕋θ:TθB100K(x)}.|f_{low}(x)|\lesssim\sum_{\theta}\frac{1}{K^{2}}\#\{T_{\theta}\in\mathbb{T}_{\theta}:T_{\theta}\cap B_{100K}(x)\not=\emptyset\}.

Recalling the condition (2) in Theorem 5, we have

#{Tθ𝕋θ:TθB100K(x)}(100K)s.\#\{T_{\theta}\in\mathbb{T}_{\theta}:T_{\theta}\cap B_{100K}(x)\not=\emptyset\}\lesssim(100K)^{s}.

This implies

(34) |flow(x)|Ks2#Θ.|f_{low}(x)|\lesssim K^{s-2}\#\Theta.

Since s<2s<2, by choosing K(logδ1)22sK\sim(\log\delta^{-1})^{\frac{2}{2-s}}, we obtain (31).

In the rest of the proof, we may pretend KK is a large constant, since any (logδ1)O(1)(\log\delta^{-1})^{O(1)}-loss is allowable (see Remark 2).

Case 2: High case

If the second term on the right hand side of (29) dominates, we say we are in the high case.

Since for any x3x\in\mathbb{R}^{3} there is at most 11 tube in 𝕋θ\mathbb{T}_{\theta} pass through xx, we have |fθ(x)|1|f_{\theta}(x)|\lesssim 1 (here we use 1\lesssim 1 instead of 1\leq 1 to take care of the rapidly decaying tail). Recalling the definition fhigh=θηθ,highfθf_{high}=\sum_{\theta}\eta^{\vee}_{\theta,high}*f_{\theta} and noting that each ηθ,high\eta^{\vee}_{\theta,high} is L1L^{1} bounded, we have

|fhigh(x)|θ|fθ(x)|#Θ.|f_{high}(x)|\lesssim\sum_{\theta}|f_{\theta}(x)|\leq\#\Theta.

We see that

(35) H|fhigh|6(#Θ)4|fhigh|2=(#Θ)4|θfθ,high|2.\int_{H}|f_{high}|^{6}\lesssim(\#\Theta)^{4}\int|f_{high}|^{2}=(\#\Theta)^{4}\int|\sum_{\theta}f_{\theta,high}|^{2}.

Next we will show that {suppf^θ,high}θΘ\{\mathrm{supp}\widehat{f}_{\theta,high}\}_{\theta\in\Theta} are finitely overlapping, i.e., {Pθ,high}θΘ\{P_{\theta,high}\}_{\theta\in\Theta} are finitely overlapping. (Actually they are O(K)O(K)-overlapping. But since the KO(1)K^{O(1)}-loss are acceptable, we may just pretend K1K\lesssim 1. See also Remark 2.) If this is true, then we have

(36) H|fhigh|6(#Θ)4θ|fθ,high|2.\int_{H}|f_{high}|^{6}\lesssim(\#\Theta)^{4}\int\sum_{\theta}|f_{\theta,high}|^{2}.

Since θ|fθ,high|2=θ|ηθ,highfθ|2θ|fθ|2θ(#𝕋θ)δ1.\int\sum_{\theta}|f_{\theta,high}|^{2}=\int\sum_{\theta}|\eta^{\vee}_{\theta,high}*f_{\theta}|^{2}\leq\int\sum_{\theta}|f_{\theta}|^{2}\sim\sum_{\theta}(\#\mathbb{T}_{\theta})\delta^{-1}. We obtain

(37) |H|(logδ1)12(#Θ)6|fhigh|6(#Θ)4θ(#𝕋θ)δ1(#Θ)4δst1,|H|(\log\delta^{-1})^{-12}(\#\Theta)^{6}\lesssim\int|f_{high}|^{6}\lesssim(\#\Theta)^{4}\sum_{\theta}(\#\mathbb{T}_{\theta})\delta^{-1}\lesssim(\#\Theta)^{4}\delta^{-s-t-1},

which implies

(38) δatδs1ε.\delta^{-a-t}\lesssim\delta^{-s-1-\varepsilon}.

Now we prove that {Pθ,high}θΘ\{P_{\theta,high}\}_{\theta\in\Theta} are finitely overlapping. First, recall that

Pθ,high={i=13ξi𝐞i(θ):|ξ1|δ,K1|ξ2|1,|ξ3|1}.P_{\theta,high}=\{\sum_{i=1}^{3}\xi_{i}{\mathbf{e}}_{i}(\theta):|\xi_{1}|\leq\delta,K^{-1}\leq|\xi_{2}|\leq 1,|\xi_{3}|\leq 1\}.

We see that Pθ,highP_{\theta,high} is contained in the δ\delta-neighborhood of the plane

Πθ,high={ξ2𝐞2(θ)+ξ3𝐞3(θ):K1|ξ2|1,|ξ3|1}.\Pi_{\theta,high}=\{\xi_{2}{\mathbf{e}}_{2}(\theta)+\xi_{3}{\mathbf{e}}_{3}(\theta):K^{-1}\leq|\xi_{2}|\leq 1,|\xi_{3}|\leq 1\}.

To show the finitely overlapping property, we just need to show: For any θ[0,1]\theta\in[0,1] and θ=θ+Δ[0,1]\theta^{\prime}=\theta+\Delta\in[0,1] with CδΔC1C\delta\leq\Delta\leq C^{-1} (for some bounded CC to be determined later), if ξΠθ,high\xi\in\Pi_{\theta,high}, then

dist(ξ,Πθ+Δ,high)>10δ.\textup{dist}(\xi,\Pi_{\theta+\Delta,high})>10\delta.

Write ξ=a𝐞2(θ)+b𝐞3(θ)=aγ(θ)+bγ(θ)×γ(θ)\xi=a{\mathbf{e}}_{2}(\theta)+b{\mathbf{e}}_{3}(\theta)=a\gamma^{\prime}(\theta)+b\gamma(\theta)\times\gamma^{\prime}(\theta), where |a|[K1,1]|a|\in[K^{-1},1] and |b|1|b|\leq 1. Since the normal direction of Πθ+Δ,high\Pi_{\theta+\Delta,high} is γ(θ+Δ)\gamma(\theta+\Delta), it suffices to prove

(39) |γ(θ+Δ)(aγ(θ)+bγ(θ)×γ(θ))|10δ.|\gamma(\theta+\Delta)\cdot\big{(}a\gamma^{\prime}(\theta)+b\gamma(\theta)\times\gamma^{\prime}(\theta)\big{)}|\geq 10\delta.

By Taylor’s expansion, we have γ(θ+Δ)=γ(θ)+Δγ(θ)+O(Δ2)\gamma(\theta+\Delta)=\gamma(\theta)+\Delta\gamma^{\prime}(\theta)+O(\Delta^{2}). We see the left hand side of (39) is |aΔ|γ(θ)|2+O(Δ2)||(aO(Δ))Δ|10δ|a\Delta|\gamma^{\prime}(\theta)|^{2}+O(\Delta^{2})|\geq|\big{(}a-O(\Delta)\big{)}\Delta|\geq 10\delta, if CC is large enough (depending on KK).

Case 3: λ\lambda-middle case (δ1/2λK1)(\delta^{1/2}\leq\lambda\leq K^{-1}) If the term (logδ1)O(1)λH|fλ|6(\log\delta^{-1})^{O(1)}\sum_{\lambda}\int_{H}|f_{\lambda}|^{6} on the right hand side of (29) dominates, we say we are in the λ\lambda-middle case. We remark that when λ\lambda is close to K1K^{-1}, H|fλ|6\int_{H}|f_{\lambda}|^{6} can be estimated in a similar way as in the High case. We will be interested in the cone

ΓK1={r𝐞3(θ):K1r1,θ[0,1]}.\Gamma_{K^{-1}}=\{r{\mathbf{e}}_{3}(\theta):K^{-1}\leq r\leq 1,\theta\in[0,1]\}.

Recall Remark 2 that we still have the decoupling inequality for this cone.

We first discuss the case that λ=δ1/2\lambda=\delta^{1/2}.

Case 3.1: λ=δ1/2\lambda=\delta^{1/2}

When λ=δ1/2\lambda=\delta^{1/2}, we have fδ1/2=θfθ,δ1/2f_{\delta^{1/2}}=\sum_{\theta}f_{\theta,\delta^{1/2}}, where each f^θ,δ1/2\widehat{f}_{\theta,\delta^{1/2}} is supported in Pθ,δ1/2P_{\theta,\delta^{1/2}}. Note that Pθ,δ1/2P_{\theta,\delta^{1/2}} consists of two pieces: One is

Pθ,δ1/2+:={i=13ξi𝐞i(θ):|ξ1|δ,|ξ2|δ1/2,K1ξ31}.P_{\theta,\delta^{1/2}}^{+}:=\{\sum_{i=1}^{3}\xi_{i}{\mathbf{e}}_{i}(\theta):|\xi_{1}|\leq\delta,|\xi_{2}|\leq\delta^{1/2},K^{-1}\leq\xi_{3}\leq 1\}.

the other is

Pθ,δ1/2:={i=13ξi𝐞i(θ):|ξ1|δ,|ξ2|δ1/2,1ξ3K1}.P_{\theta,\delta^{1/2}}^{-}:=\{\sum_{i=1}^{3}\xi_{i}{\mathbf{e}}_{i}(\theta):|\xi_{1}|\leq\delta,|\xi_{2}|\leq\delta^{1/2},-1\leq\xi_{3}\leq-K^{-1}\}.

We note that Pθ,δ1/2+P_{\theta,\delta^{1/2}}^{+} lies in the δ\delta-neighborhood of ΓK1\Gamma_{K^{-1}}. Symmetrically, Pθ,δ1/2P_{\theta,\delta^{1/2}}^{-} lies in the δ\delta-neighborhood of ΓK1\Gamma_{K^{-1}}^{-}, which is the reflection of ΓK1\Gamma_{K^{-1}} with respect to the origin. We can write fθ,δ1/2=fθ,δ1/2++fθ,δ1/2f_{\theta,\delta^{1/2}}=f_{\theta,\delta^{1/2}}^{+}+f_{\theta,\delta^{1/2}}^{-}, so that suppf^θ,δ1/2+Pθ,δ1/2+\mathrm{supp}\widehat{f}_{\theta,\delta^{1/2}}^{+}\subset P_{\theta,\delta^{1/2}}^{+} and suppf^θ,δ1/2Pθ,δ1/2\mathrm{supp}\widehat{f}_{\theta,\delta^{1/2}}^{-}\subset P_{\theta,\delta^{1/2}}^{-}. We also write fδ1/2+=θfθ,δ1/2+f_{\delta^{1/2}}^{+}=\sum_{\theta}f_{\theta,\delta^{1/2}}^{+} and fδ1/2=θfθ,δ1/2f_{\delta^{1/2}}^{-}=\sum_{\theta}f_{\theta,\delta^{1/2}}^{-}, and then fδ1/2=fδ1/2++fδ1/2f_{\delta^{1/2}}=f_{\delta^{1/2}}^{+}+f_{\delta^{1/2}}^{-}. We have

(40) |fδ1/2|6|fδ1/2+|6+|fδ1/2|6.\int|f_{\delta^{1/2}}|^{6}\lesssim\int|f_{\delta^{1/2}}^{+}|^{6}+\int|f_{\delta^{1/2}}^{-}|^{6}.

By symmetry, we only estimate |fδ1/2+|6\int|f_{\delta^{1/2}}^{+}|^{6}.

Note that Pθ,δ1/2+P_{\theta,\delta^{1/2}}^{+} and Pθ,δ1/2+P_{\theta^{\prime},\delta^{1/2}}^{+} are essentially the same when |θθ|δ1/2|\theta-\theta^{\prime}|\lesssim\delta^{1/2}; Pθ,δ1/2+P_{\theta,\delta^{1/2}}^{+} and Pθ,δ1/2+P_{\theta^{\prime},\delta^{1/2}}^{+} are essentially distinct when |θθ|δ1/2|\theta-\theta^{\prime}|\gtrsim\delta^{1/2}. We can choose a partition of Nδ(ΓK1)N_{\delta}(\Gamma_{K^{-1}}) by finitely overlapping planks of dimensions 10δ×10δ1/2×110\delta\times 10\delta^{1/2}\times 1, denoted by {R}\{R\}. We attach each Pθ,δ1/2+P_{\theta,\delta^{1/2}}^{+} to one of the RR and denote by θR\theta\prec R, if Pθ,δ1/2+RP_{\theta,\delta^{1/2}}^{+}\subset R. We see that for each RR, there are δ1/2\lesssim\delta^{-1/2} many Pθ,δ1/2P_{\theta,\delta^{1/2}} attached to it. We define fR:=θRfθ,δ1/2+f_{R}:=\sum_{\theta\prec R}f_{\theta,\delta^{1/2}}^{+}. The Fourier support of fRf_{R} is contained in RR, by Theorem 6, we have

|fδ1/2+|6=|RfR|6εδε(#{R:fR0})2RfR66.\int|f_{\delta^{1/2}}^{+}|^{6}=\int|\sum_{R}f_{R}|^{6}\lesssim_{\varepsilon}\delta^{-\varepsilon}(\#\{R:f_{R}\neq 0\})^{2}\sum_{R}\|f_{R}\|^{6}_{6}.

By pigeonholing, we may pass to a subset of {R}\{R\} so that #{θ:θR}\#\{\theta:\theta\prec R\} are all comparable. For simplicity, we write #{θ:θR}\#\{\theta:\theta\prec R\} as #{θR}\#\{\theta\prec R\}, and write the number of {R}\{R\} after pigeonholing as #{R}\#\{R\}. For each RR, we have by triangle inequality:

fR66#{θR}5θRfθ,δ1/2+66#{θR}5θRfθ66.\|f_{R}\|_{6}^{6}\lesssim\#\{\theta\prec R\}^{5}\sum_{\theta\prec R}\|f_{\theta,\delta^{1/2}}^{+}\|_{6}^{6}\lesssim\#\{\theta\prec R\}^{5}\sum_{\theta\prec R}\|f_{\theta}\|_{6}^{6}.

We obtain that

|fδ1/2+|6εδε#{R}2#{θR}5θfθ66.\int|f_{\delta^{1/2}}^{+}|^{6}\lesssim_{\varepsilon}\delta^{-\varepsilon}\#\{R\}^{2}\#\{\theta\prec R\}^{5}\sum_{\theta}\|f_{\theta}\|_{6}^{6}.

Note that #{R}#{θR}#Θ\#\{R\}\#\{\theta\prec R\}\lesssim\#\Theta, #{θR}δt/2\#\{\theta\prec R\}\lesssim\delta^{-t/2} (by the (δ,t)(\delta,t)-spacing of Θ\Theta), and θfθ66#Θδ1s\sum_{\theta}\|f_{\theta}\|_{6}^{6}\lesssim\#\Theta\delta^{-1-s}. We obtain

|fδ1/2+|6εδε(#Θ)3δ3t/21s.\int|f_{\delta^{1/2}}^{+}|^{6}\lesssim_{\varepsilon}\delta^{-\varepsilon}(\#\Theta)^{3}\delta^{-3t/2-1-s}.

Similarly, we have

|fδ1/2|6εδε(#Θ)3δ3t/21s.\int|f_{\delta^{1/2}}^{-}|^{6}\lesssim_{\varepsilon}\delta^{-\varepsilon}(\#\Theta)^{3}\delta^{-3t/2-1-s}.

As a result, we obtain

(41) |H|(#Θ)6εδ2ε(#Θ)3δ3t/21s.|H|(\#\Theta)^{6}\lesssim_{\varepsilon}\delta^{-2\varepsilon}(\#\Theta)^{3}\delta^{-3t/2-1-s}.

Combined with #Θ(logδ2)2δt\#\Theta\gtrsim(\log\delta^{-2})^{-2}\delta^{-t}, we have δa3t/2δε1s\delta^{-a-3t/2}\lesssim\delta^{-\varepsilon-1-s} which is even better than what we aimed.

Case 3.2: λ(δ1/2,K1]\lambda\in(\delta^{1/2},K^{-1}]

For λ\lambda being a dyadic scale in (δ1/2,K1](\delta^{1/2},K^{-1}], we see that f^θ,λ\widehat{f}_{\theta,\lambda} is supported in Pθ,λP_{\theta,\lambda} which consists of four separated planks (Pθ,δ1/2P_{\theta,\delta^{1/2}} only consists of two planks). As in the proof of case λ=δ1/2\lambda=\delta^{1/2}, we will write fθ,λf_{\theta,\lambda} as the sum of four functions each of which has Fourier support in one of the planks of Pθ,λP_{\theta,\lambda}. We will estimate for one of the planks. We define

(42) Qθ:={i=13ξi𝐞i(θ):|ξ1|δ,12λξ2λ,K1ξ31}.Q_{\theta}:=\{\sum_{i=1}^{3}\xi_{i}{\mathbf{e}}_{i}(\theta):|\xi_{1}|\leq\delta,\frac{1}{2}\lambda\leq\xi_{2}\leq\lambda,K^{-1}\leq\xi_{3}\leq 1\}.

Roughly speaking, QθQ_{\theta} is the top-right plank of Pθ,λP_{\theta,\lambda} and the distance between QθQ_{\theta} and the line 𝐞3(θ)\mathbb{R}{\mathbf{e}}_{3}(\theta) is 12λ\geq\frac{1}{2}\lambda. For simplicity, we may assume fθ,λf_{\theta,\lambda} has Fourier support in QθQ_{\theta}.

We discuss some geometric properties for the planks {Qθ}θΘ\{Q_{\theta}\}_{\theta\in\Theta}. First of all, there is a canonical finitely overlapping covering of Nλ2(ΓK1)N_{\lambda^{2}}(\Gamma_{K^{-1}}) by planks of dimensions λ2×λ×1\lambda^{2}\times\lambda\times 1. More precisely, we choose Σ=λ[0,1]\Sigma=\lambda\mathbb{Z}\cap[0,1] to be a set of λ\lambda-lattice points. For each σΣ\sigma\in\Sigma, define

Rσ:={i=13ξi𝐞i(σ):|ξ1|C1λ2,|ξ2|C1λ,C11K1ξ3C1},R_{\sigma}:=\{\sum_{i=1}^{3}\xi_{i}{\mathbf{e}}_{i}(\sigma):|\xi_{1}|\leq C_{1}\lambda^{2},|\xi_{2}|\leq C_{1}\lambda,C_{1}^{-1}K^{-1}\leq\xi_{3}\leq C_{1}\},

where C1C_{1} is a large constant. We see that {Rσ}\{R_{\sigma}\} form a finitely overlapping covering of Nλ2(ΓK1)N_{\lambda^{2}}(\Gamma_{K^{-1}}). We have the following three properties:

Lemma 3.

For Qθ,RσQ_{\theta},R_{\sigma} defined above, we have

  1. (1)

    If |θσ|λ|\theta-\sigma|\lesssim\lambda, then QθQ_{\theta} is contained in RσR_{\sigma}.

  2. (2)

    If |θθ|λ1δ|\theta-\theta^{\prime}|\lesssim\lambda^{-1}\delta, then QθQ_{\theta} and QθQ_{\theta^{\prime}} are essentially the same.

  3. (3)

    If |θθ|λ1δ|\theta-\theta^{\prime}|\gtrsim\lambda^{-1}\delta, then QθQ_{\theta} and QθQ_{\theta^{\prime}} are disjoint.

Before proving the lemma, we see how it can be used to finish the proof of Theorem 5. Motivated by Property (2) and (3), we define 𝒯=(λ1δ)[0,1]\mathcal{T}=(\lambda^{-1}\delta)\mathbb{Z}\cap[0,1], and for each τ𝒯\tau\in\mathcal{T} define

Sτ:={i=13ξi𝐞i(τ):|ξ1|δ,|ξ2|C2λ,C21K1ξ3C2},S_{\tau}:=\{\sum_{i=1}^{3}\xi_{i}{\mathbf{e}}_{i}(\tau):|\xi_{1}|\leq\delta,|\xi_{2}|\leq C_{2}\lambda,C_{2}^{-1}K^{-1}\leq\xi_{3}\leq C_{2}\},

where C2C_{2} is a large constant but much smaller than C1C_{1}. Note that SτS_{\tau} has the same dimensions as QθQ_{\theta} up to a C2C_{2}-dilation.

Now we have three subsets of [0,1][0,1]:

Θ=δ[0,1],𝒯=(λ1δ)[0,1],Σ=λ[0,1].\Theta=\delta\mathbb{Z}\cap[0,1],\ \ \mathcal{T}=(\lambda^{-1}\delta)\mathbb{Z}\cap[0,1],\ \ \Sigma=\lambda\mathbb{Z}\cap[0,1].

We will define a relationship between their elements. For any θΘ\theta\in\Theta, we attach it to a τ𝒯\tau\in\mathcal{T} such that |θτ|λ1δ|\theta-\tau|\leq\lambda^{-1}\delta, which we denote by θτ\theta\prec\tau. For any τ𝒯\tau\in\mathcal{T}, we attach it to a σΣ\sigma\in\Sigma such that |τσ|λ|\tau-\sigma|\leq\lambda, which we denote by τσ\tau\prec\sigma. We also write θσ\theta\prec\sigma if there is a τ\tau such that θτ\theta\prec\tau and τσ\tau\prec\sigma.

By property (1), if θσ\theta\prec\sigma, then QθRσQ_{\theta}\subset R_{\sigma}. By property (2), for a given τ𝒯\tau\in\mathcal{T}, all the planks in {Qθ:θτ}\{Q_{\theta}:\theta\prec\tau\} are essentially the same and contained in SτS_{\tau}. By property (3), if Qθ,QθQ_{\theta},Q_{\theta^{\prime}} lie in different SτS_{\tau}, then Qθ,QθQ_{\theta},Q_{\theta^{\prime}} are disjoint.

Before estimating |fλ|6\int|f_{\lambda}|^{6}, we may apply a pigeonhole argument to pass to subsets of Θ,𝒯,Σ\Theta,\mathcal{T},\Sigma (still denoted by Θ,𝒯,Σ\Theta,\mathcal{T},\Sigma), so that #{θΘ:θτ}\#\{\theta\in\Theta:\theta\prec\tau\} are comparable for τ𝒯\tau\in\mathcal{T}, and #{τ𝒯:τσ}\#\{\tau\in\mathcal{T}:\tau\prec\sigma\} are comparable for σΣ\sigma\in\Sigma. For convinience, we write #{θΘ:θτ}\#\{\theta\in\Theta:\theta\prec\tau\} as #{θτ}\#\{\theta\prec\tau\}, and write #{τ𝒯:τσ}\#\{\tau\in\mathcal{T}:\tau\prec\sigma\} as #{τσ}\#\{\tau\prec\sigma\}.

We define

fτ:=θτfθ,λ,f_{\tau}:=\sum_{\theta\prec\tau}f_{\theta,\lambda},
fσ:=θσfθ,λ=τσfτ.f_{\sigma}:=\sum_{\theta\prec\sigma}f_{\theta,\lambda}=\sum_{\tau\prec\sigma}f_{\tau}.

By decoupling for Nλ2(ΓK1)=σRσN_{\lambda^{2}}(\Gamma_{K^{-1}})=\bigsqcup_{\sigma}R_{\sigma}, we have

(43) |fλ|6=|σfσ|6δε#{σ}2σ|fσ|6.\int|f_{\lambda}|^{6}=\int|\sum_{\sigma}f_{\sigma}|^{6}\lesssim\delta^{-\varepsilon}\#\{\sigma\}^{2}\sum_{\sigma}\int|f_{\sigma}|^{6}.

By the trivial decoupling for Rσ=τσSτR_{\sigma}=\bigsqcup_{\tau\prec\sigma}S_{\tau}, we have

(44) |fσ|6=|τσfτ|6#{τσ}4τσ|fτ|6.\int|f_{\sigma}|^{6}=\int|\sum_{\tau\prec\sigma}f_{\tau}|^{6}\lesssim\#\{\tau\prec\sigma\}^{4}\sum_{\tau\prec\sigma}\int|f_{\tau}|^{6}.

By Hölder’s inequality, we have

(45) |fτ|6=|θτfθ,λ|6#{θτ}5θτ|fθ,λ|6.\int|f_{\tau}|^{6}=\int|\sum_{\theta\prec\tau}f_{\theta,\lambda}|^{6}\lesssim\#\{\theta\prec\tau\}^{5}\sum_{\theta\prec\tau}\int|f_{\theta,\lambda}|^{6}.

Combining the three inequalities, we obtain

(46) |fλ|6δε#{σ}2#{τσ}4#{θτ}5θ|fθ,λ|6.\int|f_{\lambda}|^{6}\lesssim\delta^{-\varepsilon}\#\{\sigma\}^{2}\#\{\tau\prec\sigma\}^{4}\#\{\theta\prec\tau\}^{5}\sum_{\theta}\int|f_{\theta,\lambda}|^{6}.

Note that #{σ}#{τσ}#{θτ}#Θ\#\{\sigma\}\#\{\tau\prec\sigma\}\#\{\theta\prec\tau\}\lesssim\#\Theta, #{τσ}#{θτ}(λδ1)t\#\{\tau\prec\sigma\}\#\{\theta\prec\tau\}\lesssim(\lambda\delta^{-1})^{t} (by the (δ,t)(\delta,t)-spacing of Θ\Theta), and #{θτ}λt\#\{\theta\prec\tau\}\lesssim\lambda^{-t} (by the (δ,t)(\delta,t)-spacing of Θ\Theta). We also note that θ|fθ,λ|6(#𝕋)δ1#Θδs1\sum_{\theta}\int|f_{\theta,\lambda}|^{6}\lesssim(\#\mathbb{T})\delta^{-1}\lesssim\#\Theta\delta^{-s-1}. We obtain

(47) |fλ|6δε(#Θ)2(λδ1)2tλt(#𝕋)δ1δε(#Θ)3λtδs12t.\int|f_{\lambda}|^{6}\lesssim\delta^{-\varepsilon}(\#\Theta)^{2}(\lambda\delta^{-1})^{2t}\lambda^{-t}(\#\mathbb{T})\delta^{-1}\lesssim\delta^{-\varepsilon}(\#\Theta)^{3}\lambda^{t}\delta^{-s-1-2t}.

Plugging into (29) and noting δt#Θ(logδ1)2δt\delta^{-t}\gtrsim\#\Theta\gtrsim(\log\delta^{-1})^{-2}\delta^{-t}, we obtain

δatδ2ελtδs1,\delta^{-a-t}\lesssim\delta^{-2\varepsilon}\lambda^{t}\delta^{-s-1},

which is better than we aimed because of the factor λt\lambda^{t}.

𝒆2(θ)\boldsymbol{e}_{2}(\theta)𝒆1(θ)\boldsymbol{e}_{1}(\theta)𝒆3(θ)\boldsymbol{e}_{3}(\theta)PθP_{\theta}SτS_{\tau}RσR_{\sigma}
Figure 2. Relation between planks

It remains to prove Lemma 3. Before proving the lemma, we give some intuition on why the lemma should be true. See Figure (2). We first cover Nλ2ΓK1N_{\lambda^{2}}\Gamma_{K^{-1}} by gray planks RσR_{\sigma} of dimensions λ2×λ×1\sim\lambda^{2}\times\lambda\times 1. Fix a RσR_{\sigma}, we draw all the black slabs PθP_{\theta} of dimensions δ×1×1\delta\times 1\times 1 whose corresponding QθQ_{\theta} is contained in RσR_{\sigma}. Morally speaking, PθRσQθP_{\theta}\cap R_{\sigma}\approx Q_{\theta} which is a δ×λ×1\delta\times\lambda\times 1-plank. One tricky thing is that different PθP_{\theta} may have essentially same QθQ_{\theta}, which is the reason to introduce SτS_{\tau} (the thick-black planks in the Figure of dimensions δ×λ×1\delta\times\lambda\times 1). Suppose we have a partition Rσ=τσSτR_{\sigma}=\bigcup_{\tau\prec\sigma}S_{\tau}. We see that each PθRσP_{\theta}\cap R_{\sigma} is contained in one of the SτS_{\tau}. If so, then we define θτ.\theta\prec\tau. We can talk about the intuition on the numerology of these planks.

  1. (1)

    #{Rσ}=λ1\#\{R_{\sigma}\}=\lambda^{-1},

  2. (2)

    #{Qθ:QθRσ}#{θσ}=δ1λ\#\{Q_{\theta}:Q_{\theta}\subset R_{\sigma}\}\sim\#\{\theta\prec\sigma\}=\delta^{-1}\lambda,

  3. (3)

    #{Sτ:SτRσ}|Rσ||Sτ|=δ1λ2\#\{S_{\tau}:S_{\tau}\subset R_{\sigma}\}\sim\frac{|R_{\sigma}|}{|S_{\tau}|}=\delta^{-1}\lambda^{2},

  4. (4)

    #{Qθ:QθSτ}#{Qθ:QθRσ}#{Sτ:SτRσ}=λ1\#\{Q_{\theta}:Q_{\theta}\subset S_{\tau}\}\sim\frac{\#\{Q_{\theta}:Q_{\theta}\subset R_{\sigma}\}}{\#\{S_{\tau}:S_{\tau}\subset R_{\sigma}\}}=\lambda^{-1}.

By (2), we see Property (1) in Lemma 3 should be true. By (4), we see Property (2) and (3) in Lemma 3 should be true.

Proof of Lemma 3.

Recall 𝐞1(θ)=γ(θ),𝐞2(θ)=γ(θ),𝐞3(θ)=γ(θ)×γ(θ){\mathbf{e}}_{1}(\theta)=\gamma(\theta),{\mathbf{e}}_{2}(\theta)=\gamma^{\prime}(\theta),{\mathbf{e}}_{3}(\theta)=\gamma(\theta)\times\gamma^{\prime}(\theta). Defining

κ(θ)=𝐞2(θ),𝐞3(θ)(1),\kappa(\theta)=\langle{\mathbf{e}}_{2}^{\prime}(\theta),{\mathbf{e}}_{3}(\theta)\rangle(\gtrsim 1),

we have

(48) 𝐞1(θ)=𝐞2(θ),\displaystyle{\mathbf{e}}_{1}^{\prime}(\theta)={\mathbf{e}}_{2}(\theta),
(49) 𝐞2(θ)=𝐞1(θ)+κ(θ)𝐞3(θ),\displaystyle{\mathbf{e}}_{2}^{\prime}(\theta)=-{\mathbf{e}}_{1}(\theta)+\kappa(\theta){\mathbf{e}}_{3}(\theta),
(50) 𝐞3(θ)=κ(θ)𝐞2(θ).\displaystyle{\mathbf{e}}_{3}^{\prime}(\theta)=-\kappa(\theta){\mathbf{e}}_{2}(\theta).

To prove Property (1), write θ=σ+Δ\theta=\sigma+\Delta with |Δ|λ|\Delta|\lesssim\lambda. Any ξQθ\xi\in Q_{\theta} can be written as ξ=a𝐞1(θ)+b𝐞2(θ)+c𝐞3(θ)\xi=a{\mathbf{e}}_{1}(\theta)+b{\mathbf{e}}_{2}(\theta)+c{\mathbf{e}}_{3}(\theta) with |a|δ,|b|λ,|c|1|a|\leq\delta,|b|\leq\lambda,|c|\leq 1. By Taylor’s expansion, we have

(51) ξ=\displaystyle\xi= a(𝐞1(σ)+Δ𝐞2(σ))+b(𝐞2(σ)Δ𝐞1(σ)+Δκ(σ)𝐞3(σ))\displaystyle a({\mathbf{e}}_{1}(\sigma)+\Delta{\mathbf{e}}_{2}(\sigma))+b({\mathbf{e}}_{2}(\sigma)-\Delta{\mathbf{e}}_{1}(\sigma)+\Delta\kappa(\sigma){\mathbf{e}}_{3}(\sigma))
+c(𝐞3(σ)Δκ(σ)𝐞2(σ))+O(Δ2)\displaystyle+c({\mathbf{e}}_{3}(\sigma)-\Delta\kappa(\sigma){\mathbf{e}}_{2}(\sigma))+O(\Delta^{2})
=\displaystyle= (abΔ)𝐞1(σ)+(aΔ+bcΔκ(σ))𝐞2(σ)+(bΔκ(σ)+c)𝐞3(σ)+O(Δ2).\displaystyle(a-b\Delta){\mathbf{e}}_{1}(\sigma)+(a\Delta+b-c\Delta\kappa(\sigma)){\mathbf{e}}_{2}(\sigma)+(b\Delta\kappa(\sigma)+c){\mathbf{e}}_{3}(\sigma)+O(\Delta^{2}).

One can easily check ξRσ\xi\in R_{\sigma}. The Property (2) can also be proved by using (51). For the Property (3), we have proved a special case λ1\lambda\sim 1 in the High case, but here we need to do more work. We may assume λ<<1\lambda<<1. Consider the plane

Πθ={ξ2𝐞2(θ)+ξ3𝐞3(θ):12λξ2λ,K1ξ31}.\Pi_{\theta}=\{\xi_{2}{\mathbf{e}}_{2}(\theta)+\xi_{3}{\mathbf{e}}_{3}(\theta):\frac{1}{2}\lambda\leq\xi_{2}\leq\lambda,K^{-1}\leq\xi_{3}\leq 1\}.

We see QθQ_{\theta} is the δ\delta-neighborhood of Πθ\Pi_{\theta}. We just need to show: For any θ[0,1]\theta\in[0,1] and θ=θ+Δ[0,1]\theta^{\prime}=\theta+\Delta\in[0,1] with Cλ1δΔC1C\lambda^{-1}\delta\leq\Delta\leq C^{-1} (for some bounded CC to be determined later), if ξΠθ\xi\in\Pi_{\theta}, then

dist(ξ,Πθ+Δ)>10δ.\textup{dist}(\xi,\Pi_{\theta+\Delta})>10\delta.

Write ξ=a𝐞2(θ)+b𝐞3(θ)\xi=a{\mathbf{e}}_{2}(\theta)+b{\mathbf{e}}_{3}(\theta), where a[12λ,λ]a\in[\frac{1}{2}\lambda,\lambda] and b[K1,1]b\in[K^{-1},1].

We consider two scenarios: (1) Cλ1δΔCλC\lambda^{-1}\delta\leq\Delta\leq C\lambda, (2) CλΔC1C\lambda\leq\Delta\leq C^{-1}. If we are in the first scenario, since the normal direction of Πθ+Δ\Pi_{\theta+\Delta} is 𝐞1(θ+Δ){\mathbf{e}}_{1}(\theta+\Delta), it suffices to prove

(52) |𝐞1(θ+Δ)(a𝐞2(θ)+b𝐞3(θ))|10δ.|{\mathbf{e}}_{1}(\theta+\Delta)\cdot\big{(}a{\mathbf{e}}_{2}(\theta)+b{\mathbf{e}}_{3}(\theta)\big{)}|\geq 10\delta.

By Taylor’s expansion, we have 𝐞1(θ+Δ)=𝐞1(θ)+Δ𝐞2(θ)+Δ22(𝐞1(θ)+κ(θ)𝐞3(θ))+o(Δ2){\mathbf{e}}_{1}(\theta+\Delta)={\mathbf{e}}_{1}(\theta)+\Delta{\mathbf{e}}_{2}(\theta)+\frac{\Delta^{2}}{2}(-{\mathbf{e}}_{1}(\theta)+\kappa(\theta){\mathbf{e}}_{3}(\theta))+o(\Delta^{2}). We see the left hand side of (52) is |aΔ+Δ22κ(θ)b+o(Δ2)|aΔo(Δ2)=(ao(Δ))Δ10CaΔ10δ|a\Delta+\frac{\Delta^{2}}{2}\kappa(\theta)b+o(\Delta^{2})|\geq a\Delta-o(\Delta^{2})=(a-o(\Delta))\Delta\geq\frac{10}{C}a\Delta\gtrsim 10\delta, when ΔCλ\Delta\leq C\lambda and λ<<1\lambda<<1. If we are in the second scenario, we show that

(53) 𝐞2(θ+Δ)(a𝐞2(θ)+b𝐞3(θ))10λ.{\mathbf{e}}_{2}(\theta+\Delta)\cdot(a{\mathbf{e}}_{2}(\theta)+b{\mathbf{e}}_{3}(\theta))\geq 10\lambda.

By Taylor’s expansion, we have 𝐞2(θ+Δ)=𝐞2(θ)+Δ(𝐞1(θ)+κ(θ)𝐞3(θ))+O(Δ2){\mathbf{e}}_{2}(\theta+\Delta)={\mathbf{e}}_{2}(\theta)+\Delta(-{\mathbf{e}}_{1}(\theta)+\kappa(\theta){\mathbf{e}}_{3}(\theta))+O(\Delta^{2}). We see the left hand side of (53) is |a+Δκ(θ)b+O(Δ2)|Δκ(θ)baO(Δ2)10λ|a+\Delta\kappa(\theta)b+O(\Delta^{2})|\geq\Delta\kappa(\theta)b-a-O(\Delta^{2})\geq 10\lambda if the constant CC is big enough. ∎

3. Proof of Theorem 2

For a small positive number δ\delta and E[0,1]E\subset[0,1], we use Λδ(E)\Lambda_{\delta}(E) to denote a maximal δ\delta-separated subset of EE. By definition, #Λδ(E)|E|δ\#\Lambda_{\delta}(E)\sim|E|_{\delta}. If E=[0,1]E=[0,1], then we abbreviate Λδ(E)\Lambda_{\delta}(E) as Λδ\Lambda_{\delta} and just choose it to be the δ\delta-lattice points in [0,1][0,1]. A rectangular box of dimensions δ×δ×1\delta\times\delta\times 1 will be referred to as a δ\delta-tube. For each θΛδ\theta\in\Lambda_{\delta}, there is a set of finitely overlapping collection of δ\delta-tubes that cover 3\mathbb{R}^{3} whose long sides are parallel to γ(θ)\gamma(\theta).

In order to prove Theorem 2, we need the following result about incidence estimate.

3.1. An incidence estimate

Theorem 7.

Let Λδ\Lambda_{\delta} be a δ\delta-net of [0,1][0,1] for some δ>0\delta>0. Given a small constant ε>0\varepsilon>0 and 0<α20<\alpha\leq 2, let μ\mu be a finite nonzero Borel measure supported in the unit ball in 3\mathbb{R}^{3} with cα(μ):=supx3,r>0μ(B(x,r))rα<c_{\alpha}(\mu):=\sup_{x\in\mathbb{R}^{3},r>0}\frac{\mu(B(x,r))}{r^{\alpha}}<\infty. Suppose that 𝕎\mathbb{W} is a set of δ\delta-tubes, with directions in {γ(θ):θΛδ}\{\gamma(\theta):\theta\in\Lambda_{\delta}\}, such that each 𝕎θ\mathbb{W}_{\theta} is disjoint, where we use 𝕎θ\mathbb{W}_{\theta} to denote the subset of tubes in 𝕎\mathbb{W} that points to direction γ(θ)\gamma(\theta). Suppose also that

(54) T𝕎χT(x)δε1,xsupp(μ).\sum_{T\in\mathbb{W}}\chi_{T}(x)\gtrsim\delta^{\varepsilon-1},\qquad\forall\,x\in\mathrm{supp}(\mu).

Then

(55) |𝕎|Cε,αμ(3)cα(μ)1δ(1+αO(ε)),|\mathbb{W}|\geq C_{\varepsilon,\alpha}\cdot\mu(\mathbb{R}^{3})c_{\alpha}(\mu)^{-1}\delta^{-(1+\alpha-O(\sqrt{\varepsilon}))},

where the constant Cε,αC_{\varepsilon,\alpha} is allowed to depend on α\alpha and ε\varepsilon but not on δ\delta, and O(ε)O(\sqrt{\varepsilon}) can be taken to be 1010ε10^{10}\sqrt{\varepsilon}. We also remark that Cε,αCε′′,αC_{\varepsilon^{\prime},\alpha}\leq C_{\varepsilon^{\prime\prime},\alpha} for εε′′\varepsilon^{\prime}\leq\varepsilon^{\prime\prime}.

Proof of Theorem 7.

The main argument of the proof is similar to that of Theorem 4, except for the “low case”.

For a given δ\delta-tube Tθ𝕎θT_{\theta}\in\mathbb{W}_{\theta}, we denote the dual slab of TθT_{\theta} by PθP_{\theta} which is of dimensions δ1×δ1×1\delta^{-1}\times\delta^{-1}\times 1 and centered at the origin. Recalling (17), we can write

Pθ={i=13ξi𝐞i(θ):|ξ1|1,|ξ2|δ1,|ξ3|δ1}.P_{\theta}=\left\{\sum_{i=1}^{3}\xi_{i}\mathbf{e}_{i}(\theta):|\xi_{1}|\leq 1,|\xi_{2}|\leq\delta^{-1},|\xi_{3}|\leq\delta^{-1}\right\}.
Remark 4.

The slab PθP_{\theta} here is just the δ1\delta^{-1}-dilation of that in the proof of Theorem 4. So are the tubes TθT_{\theta} and Pθ,high,Pθ,low,Pλ,θP_{\theta,high},P_{\theta,low},P_{\lambda,\theta} that we will define right now.

Let ϕT:3\phi_{T}:\mathbb{R}^{3}\to\mathbb{R} be a non-negative function with ϕT^\widehat{\phi_{T}} supported on PθP_{\theta} and ϕT(x)1\phi_{T}(x)\gtrsim 1 for every xTx\in T. Set

(56) fθ(x):=T𝕎θϕT(x),f(x):=T𝕎ϕT(x).f_{\theta}(x):=\sum_{T\in\mathbb{W}_{\theta}}\phi_{T}(x),\ \ \ f(x):=\sum_{T\in\mathbb{W}}\phi_{T}(x).

The assumption (54) implies that

(57) f(x)δε1,f(x)\gtrsim\delta^{\varepsilon-1},

for every xsupp(μ)x\in\mathrm{supp}(\mu).

Next we will do the frequency decomposition for fθf_{\theta}. Similar to Definition 3, we make the following definitions.

Definition 4.

(See Figure 1.) Let K=δεK=\delta^{-\sqrt{\varepsilon}}. Define the high part of PθP_{\theta} as

Pθ,high:={i=13ξi𝐞i(θ):|ξ1|1,K1δ1|ξ2|δ1,|ξ3|δ1}.P_{\theta,high}:=\left\{\sum_{i=1}^{3}\xi_{i}{\mathbf{e}}_{i}(\theta):|\xi_{1}|\leq 1,K^{-1}\delta^{-1}\leq|\xi_{2}|\leq\delta^{-1},|\xi_{3}|\leq\delta^{-1}\right\}.

Define the low part of PθP_{\theta} as

Pθ,low:={i=13ξi𝐞i(θ):|ξ1|1,|ξ2|K1δ1,|ξ3|K1δ1}.P_{\theta,low}:=\left\{\sum_{i=1}^{3}\xi_{i}{\mathbf{e}}_{i}(\theta):|\xi_{1}|\leq 1,|\xi_{2}|\leq K^{-1}\delta^{-1},|\xi_{3}|\leq K^{-1}\delta^{-1}\right\}.

For dyadic numbers λ(δ1/2,K1]\lambda\in(\delta^{1/2},K^{-1}], define

Pθ,λ:={i=13ξi𝐞i(θ):|ξ1|1,12λδ1|ξ2|λδ1,K1δ1|ξ3|δ1}.P_{\theta,\lambda}:=\left\{\sum_{i=1}^{3}\xi_{i}{\mathbf{e}}_{i}(\theta):|\xi_{1}|\leq 1,\frac{1}{2}\lambda\delta^{-1}\leq|\xi_{2}|\leq\lambda\delta^{-1},K^{-1}\delta^{-1}\leq|\xi_{3}|\leq\delta^{-1}\right\}.

In particular, we define

Pθ,δ1/2:={i=13ξi𝐞i(θ):|ξ1|1,|ξ2|δ1/2,K1δ1|ξ3|δ1}.P_{\theta,\delta^{1/2}}:=\left\{\sum_{i=1}^{3}\xi_{i}{\mathbf{e}}_{i}(\theta):|\xi_{1}|\leq 1,|\xi_{2}|\leq\delta^{-1/2},K^{-1}\delta^{-1}\leq|\xi_{3}|\leq\delta^{-1}\right\}.

Similarly to (27) and (28), we have

(58) fθ=fθ,high+fθ,low+δ1/2λK1fθ,λ,f_{\theta}=f_{\theta,high}+f_{\theta,low}+\sum_{\delta^{1/2}\leq\lambda\leq K^{-1}}f_{\theta,\lambda},

and

(59) f=fhigh+flow+δ1/2λK1fλ.f=f_{high}+f_{low}+\sum_{\delta^{1/2}\leq\lambda\leq K^{-1}}f_{\lambda}.

Since f(x)δε1f(x)\gtrsim\delta^{\varepsilon-1}, there are two cases:

Case 1: High case We can find a Borel set FF satisfying

μ(F)μ(3),\mu(F)\gtrsim\mu(\mathbb{R}^{3}),

and

(60) δε1|fhigh(x)+λfλ(x)|,xF.\delta^{\varepsilon-1}\lesssim\left\lvert f_{high}(x)+\sum_{\lambda}f_{\lambda}(x)\right\rvert,\quad\forall\,x\in F.

Case 2: Low case We can find a Borel set FF satisfying

μ(F)μ(3),\mu(F)\gtrsim\mu(\mathbb{R}^{3}),

and

(61) δε1|flow(x)|,xF.\delta^{\varepsilon-1}\lesssim\left\lvert f_{low}(x)\right\rvert,\quad\forall\,x\in F.

Assume first that we are in the high case. We raise both sides of (60) to the sixth power, integrate with respect to dμd\mu, and obtain

(62) μ(3)δ6(1ε)(logδ1)O(1)|fhigh|6+λ|fλ|6dμ.\mu(\mathbb{R}^{3})\delta^{-6(1-\varepsilon)}\lesssim(\log\delta^{-1})^{O(1)}\int|f_{high}|^{6}+\sum_{\lambda}|f_{\lambda}|^{6}d\mu.

Since the functions on the right hand side are locally constant on δ\delta-balls, together with the upper density condition on μ\mu, we obtain

(63) μ(3)cα(μ)1δ6(1ε)+3α(logδ1)O(1)|fhigh(x)|6+λ|fλ(x)|6dx.\mu(\mathbb{R}^{3})c_{\alpha}(\mu)^{-1}\delta^{-6(1-\varepsilon)+3-\alpha}\lesssim(\log\delta^{-1})^{O(1)}\int|f_{high}(x)|^{6}+\sum_{\lambda}|f_{\lambda}(x)|^{6}\,dx.

We can just use (37) and (47) with t=1t=1 and #Θδ1\#\Theta\sim\delta^{-1}, noting there is a scaling difference, so that the right hand side above is bounded by

δO(ε)δ2|𝕎|.\delta^{-O(\sqrt{\varepsilon})}\delta^{-2}|\mathbb{W}|.

It follows that

|𝕎|μ(3)cα(μ)1δ(1+αO(ε)).|\mathbb{W}|\gtrsim\mu(\mathbb{R}^{3})c_{\alpha}(\mu)^{-1}\delta^{-(1+\alpha-O(\sqrt{\varepsilon}))}.

This finishes the proof if we are in the high case (60).

Now we assume that we are in the low case (61). For each T𝕎T\in\mathbb{W}, the support of ϕT\savestack\tmpbox\stretchto\scaleto\scalerel[width("ηlow")] 0.5ex\stackon[1pt]ηlow\tmpbox\phi_{T}*\savestack{\tmpbox}{\stretchto{\scaleto{\scalerel*[width("\eta_{low}")]{\kern-0.6pt\bigwedge\kern-0.6pt}{\rule[-505.89pt]{4.30554pt}{505.89pt}}}{}}{0.5ex}}\stackon[1pt]{\eta_{low}}{\scalebox{-1.0}{\tmpbox}} is essentially a thickened tube of TT with dimensions 1×Kδ×Kδ1\times K\delta\times K\delta where K=δεK=\delta^{-\sqrt{\varepsilon}}. We use T~\widetilde{T} to denote this thickened tube. Let 𝕎~\widetilde{\mathbb{W}} be the collection of these thickened tubes obtained from 𝕎\mathbb{W}, and we only keep those essentially distinct tubes (each tube intersects 1\lesssim 1 other tubes of those whose angle is within Kδ\lesssim K\delta of its own, and every T𝕎T\in\mathbb{W} is contained in some T~\widetilde{T} from 𝕎~\widetilde{\mathbb{W}}).

Write 𝕎~\widetilde{\mathbb{W}} as a disjoint union

(64) 𝕎~=𝕎~heavy𝕎~light\widetilde{\mathbb{W}}=\widetilde{\mathbb{W}}_{\mathrm{heavy}}\bigcup\widetilde{\mathbb{W}}_{\mathrm{light}}

where 𝕎~light\widetilde{\mathbb{W}}_{\mathrm{light}} is the collection of thickened tubes T~\widetilde{T} that contain C1K3ε\leq C^{-1}K^{3-\sqrt{\varepsilon}} tubes from 𝕎\mathbb{W}. Here CC is a large universal constant which is much larger than the implicit (universal) constant in (61). Note that

(65)

As a consequence, we see that (61) can be upgraded to

(66) δε1|T~𝕎~heavyT𝕎:TT~ϕT\savestack\tmpbox\stretchto\scaleto\scalerel[width("ηlow")] 0.5ex\stackon[1pt]ηlow\tmpbox(x)|,xF.\delta^{\varepsilon-1}\lesssim\Big{|}\sum_{\widetilde{T}\in\widetilde{\mathbb{W}}_{\mathrm{heavy}}}\sum_{\begin{subarray}{c}T^{\prime}\in\mathbb{W}:T^{\prime}\subset\widetilde{T}\end{subarray}}\phi_{T^{\prime}}*\savestack{\tmpbox}{\stretchto{\scaleto{\scalerel*[width("\eta_{low}")]{\kern-0.6pt\bigwedge\kern-0.6pt}{\rule[-505.89pt]{4.30554pt}{505.89pt}}}{}}{0.5ex}}\stackon[1pt]{\eta_{low}}{\scalebox{-1.0}{\tmpbox}}(x)\Big{|},\ \forall x\in F.

Next, by (66) and the fact that if xT~x\in\widetilde{T} then #{T𝕎:xTT~}K\#\{T\in\mathbb{W}:x\in T\subset\widetilde{T}\}\lesssim K, we conclude that

(67) T~𝕎~heavyχT~(x)K1δε1,xF.\sum_{\widetilde{T}\in\widetilde{\mathbb{W}}_{\mathrm{heavy}}}\chi_{\widetilde{T}}(x)\gtrsim K^{-1}\delta^{\varepsilon-1},\ \forall x\in F.

Write

(68) δ1+εK1=(δK)1+ε~,ε~:=ε1ε.\delta^{-1+\varepsilon}K^{-1}=(\delta K)^{-1+\widetilde{\varepsilon}},\ \widetilde{\varepsilon}:=\frac{\varepsilon}{1-\sqrt{\varepsilon}}.

The tubes in 𝕎~heavy\widetilde{\mathbb{W}}_{\mathrm{heavy}} satisfy the induction hypothesis at the scale KδK\delta with the new parameter ε~\widetilde{\varepsilon}. Hence

|𝕎||𝕎~heavy|K3ε\displaystyle|\mathbb{W}|\geq|\widetilde{\mathbb{W}}_{\mathrm{heavy}}|K^{3-\sqrt{\varepsilon}} μ(3)cα(μ)1Cε~,α(δ1K1)(α+11010ε~)K3ε\displaystyle\geq\mu(\mathbb{R}^{3})c_{\alpha}(\mu)^{-1}C_{\widetilde{\varepsilon},\alpha}\left(\delta^{-1}K^{-1}\right)^{(\alpha+1-10^{10}\sqrt{\widetilde{\varepsilon}})}K^{3-\sqrt{\varepsilon}}

Elementary computation shows that

(69) ε~ε=ε~εε~+εε2(1ε)3ε4.\sqrt{\widetilde{\varepsilon}}-\sqrt{\varepsilon}=\frac{\widetilde{\varepsilon}-\varepsilon}{\sqrt{\widetilde{\varepsilon}}+\sqrt{\varepsilon}}\leq\frac{\varepsilon}{2(1-\sqrt{\varepsilon})}\leq\frac{3\varepsilon}{4}.

Therefore,

(70) |𝕎|Cε~,αμ(3)cα(μ)1δ(α+11010ε)δ10103ε4K3(1+α1010ε~)εCε~,αμ(3)cα(μ)1δ(α+11010ε)K2αδ10103ε4δ1010ε+ε.\begin{split}|\mathbb{W}|&\geq C_{\widetilde{\varepsilon},\alpha}\cdot\mu(\mathbb{R}^{3})c_{\alpha}(\mu)^{-1}\delta^{-(\alpha+1-10^{10}\sqrt{\varepsilon})}\delta^{10^{10}\frac{3\varepsilon}{4}}K^{3-(1+\alpha-10^{10}\sqrt{\widetilde{\varepsilon}})-\sqrt{\varepsilon}}\\ &\geq C_{\widetilde{\varepsilon},\alpha}\cdot\mu(\mathbb{R}^{3})c_{\alpha}(\mu)^{-1}\delta^{-(\alpha+1-10^{10}\sqrt{\varepsilon})}K^{2-\alpha}\delta^{10^{10}\frac{3\varepsilon}{4}}\delta^{-10^{10}\varepsilon+\varepsilon}.\end{split}

Since α2\alpha\leq 2, the induction closes. ∎

We have the following corollary of Theorem 7.

Corollary 2.

Let α(0,3)\alpha\in(0,3) and α1(0,min{2,α})\alpha_{1}\in(0,\min\{2,\alpha\}). Let μ\mu be a finite non-zero Borel measure supported on the unit ball in 3\mathbb{R}^{3} with cα(μ)1c_{\alpha}(\mu)\leq 1. Let δ>0\delta>0 be a small number. Let Λδ\Lambda_{\delta} be a δ\delta-net of [0,1][0,1]. For each θΛδ\theta\in\Lambda_{\delta}, let 𝔻θ\mathbb{D}_{\theta} be a disjoint collection of at most μ(3)δα1\mu(\mathbb{R}^{3})\delta^{-\alpha_{1}} balls of radius δ\delta in πθ(3)\pi_{\theta}(\mathbb{R}^{3}). Then there exists ε>0\varepsilon>0, depending only on α\alpha and α1\alpha_{1}, such that

(71) δθΛδ(πθ#μ)(D𝔻θD)C(α,α1)μ(3)δε,\delta\sum_{\theta\in\Lambda_{\delta}}(\pi_{\theta\#}\mu)\Big{(}\bigcup_{D\in\mathbb{D}_{\theta}}D\Big{)}\leq C(\alpha,\alpha_{1})\mu(\mathbb{R}^{3})\delta^{\varepsilon},

where the constant C(α,α1)C(\alpha,\alpha_{1}) depends on α,α1\alpha,\alpha_{1}, but not on δ\delta.

Proof of Corollary 2.

We argue by contradiction. Suppose that for every ε>0\varepsilon>0, there exist δ>0\delta>0 and 𝔻θ\mathbb{D}_{\theta}, a disjoint collection of at most μ(3)δα1\mu(\mathbb{R}^{3})\delta^{-\alpha_{1}} balls of radius δ\delta in πθ(3)\pi_{\theta}(\mathbb{R}^{3}) for each θΛδ\theta\in\Lambda_{\delta}, such that (71) fails. Note that for each Dθ𝔻θD_{\theta}\in\mathbb{D}_{\theta}, πθ1(Dθ)B3(0,1)\pi_{\theta}^{-1}(D_{\theta})\cap B^{3}(0,1) is a δ\delta-tube. We denote these tubes by

(72) 𝕎θ:={πθ1(Dθ)B3(0,1):Dθ𝔻θ},\mathbb{W}_{\theta}:=\{\pi_{\theta}^{-1}(D_{\theta})\cap B^{3}(0,1):D_{\theta}\in\mathbb{D}_{\theta}\},

and write 𝕎:=θ𝕎θ\mathbb{W}:=\cup_{\theta}\mathbb{W}_{\theta}. Define

(73) F:={xsupp(μ):δθΛδT𝕎θχT(x)C(α,α1)δε/2}F:=\Big{\{}x\in\mathrm{supp}(\mu):\delta\sum_{\theta\in\Lambda_{\delta}}\sum_{T\in\mathbb{W}_{\theta}}\chi_{T}(x)\geq C(\alpha,\alpha_{1})\delta^{\varepsilon}/2\Big{\}}

Note that by our contradiction assumption, we have

(74) C(α,α1)μ(3)δεδθΛδ(πθ#μ)(D𝔻θD)=δ3θT𝕎θχT(x)dμ(x).\begin{split}C(\alpha,\alpha_{1})\mu(\mathbb{R}^{3})\delta^{\varepsilon}&\leq\delta\sum_{\theta\in\Lambda_{\delta}}(\pi_{\theta\#}\mu)\Big{(}\bigcup_{D\in\mathbb{D}_{\theta}}D\Big{)}\\ &=\delta\int_{\mathbb{R}^{3}}\sum_{\theta}\sum_{T\in\mathbb{W}_{\theta}}\chi_{T}(x)d\mu(x).\end{split}

This further implies

(75) μ(F)C(α,α1)μ(3)δε.\mu(F)\gtrsim C(\alpha,\alpha_{1})\mu(\mathbb{R}^{3})\delta^{\varepsilon}.

Note that by definition, for every xFx\in F, it holds that

(76) T𝕎χT(x)δ1+ε.\sum_{T\in\mathbb{W}}\chi_{T}(x)\gtrsim\delta^{-1+\varepsilon}.

We apply Theorem 7 to the measure μ\mu restricted to FF and obtain

(77) |𝕎|δεδ1min{2,α}+O(ε)μ(3).|\mathbb{W}|\gtrsim\delta^{\varepsilon}\delta^{-1-\min\{2,\alpha\}+O(\sqrt{\varepsilon})}\mu(\mathbb{R}^{3}).

By pigeonholing, this further implies that there exists θ\theta such that

(78) |𝕎θ|δεδmin{2,α}+O(ε)μ(3).|\mathbb{W}_{\theta}|\gtrsim\delta^{\varepsilon}\delta^{-\min\{2,\alpha\}+O(\sqrt{\varepsilon})}\mu(\mathbb{R}^{3}).

This is a contradiction to the assumption that |𝔻θ|δα1μ(3)|\mathbb{D}_{\theta}|\lesssim\delta^{-\alpha_{1}}\mu(\mathbb{R}^{3}) when ε\varepsilon is chosen to be small enough. ∎

The corollary below is a special case of Corollary 2. It is recorded below for a later application.

Corollary 3.

Let α[2,3]\alpha\in[2,3] and α[0,2)\alpha^{*}\in\left[0,2\right). Let μ\mu be a finite non-zero Borel measure supported on the unit ball in 3\mathbb{R}^{3} with cα(μ)1c_{\alpha}(\mu)\leq 1. Let δ>0\delta>0 be a small number. Let Λδ\Lambda_{\sqrt{\delta}} be a δ\sqrt{\delta}-net of [0,1][0,1]. For each θΛδ\theta\in\Lambda_{\sqrt{\delta}}, let 𝔻θ\mathbb{D}_{\theta} be a disjoint collection of at most δα+12μ(3)\delta^{-\frac{\alpha^{*}+1}{2}}\mu(\mathbb{R}^{3}) rectangles of dimension δ×δ\delta\times\sqrt{\delta} in πθ(3)\pi_{\theta}(\mathbb{R}^{3}) whose long sides point in the γ(θ)\gamma^{\prime}(\theta) direction. Then there exists ε>0\varepsilon>0, depending only on α\alpha and α\alpha^{*}, such that

(79) δθΛδ(πθ#μ)(D𝔻θD)C(α,α)δεμ(3),\sqrt{\delta}\sum_{\theta\in\Lambda_{\sqrt{\delta}}}(\pi_{\theta\#}\mu)\Big{(}\bigcup_{D\in\mathbb{D}_{\theta}}D\Big{)}\leq C(\alpha,\alpha^{*})\delta^{\varepsilon}\mu(\mathbb{R}^{3}),

where the constant C(α,α)C(\alpha,\alpha^{*}) depends on γ,α,α\gamma,\alpha,\alpha^{*}, but not on δ\delta.

Proof of Corollary 3.

For each θΛδ\theta\in\Lambda_{\sqrt{\delta}} and recalling (17), define

(80) Uθ:={i=13xi𝐞i(θ):|x1|1,|x2|δ,|x3|δ},U_{\theta}:=\left\{\sum_{i=1}^{3}x_{i}{\mathbf{e}}_{i}(\theta):|x_{1}|\leq 1,|x_{2}|\leq\sqrt{\delta},|x_{3}|\leq\delta\right\},

which is a δ×δ×1\delta\times\sqrt{\delta}\times 1-plank. By a simple geometric observation, we have that

(81) πθ(Wθ)Cπθ(Wθ),\pi_{\theta^{\prime}}(W_{\theta})\subset C\pi_{\theta}(W_{\theta}),

for every θ,θ[0,1]\theta,\theta^{\prime}\in[0,1] with |θθ|δ|\theta^{\prime}-\theta|\leq\sqrt{\delta}, where CC is a constant depending only on γ.\gamma. For each θΛδ\theta\in\Lambda_{\sqrt{\delta}} we have a set of δ×δ×1\delta\times\sqrt{\delta}\times 1-planks

𝕌θ={πθ1(Dθ)B3(0,1)}Dθ𝔻θ,\mathbb{U}_{\theta}=\{\pi_{\theta}^{-1}(D_{\theta})\cap B^{3}(0,1)\}_{D_{\theta}\in\mathbb{D}_{\theta}},

which are essentially the translates of UθU_{\theta}. For each θΛδ\theta^{\prime}\in\Lambda_{\delta}, we choose a θΛδ\theta\in\Lambda_{\sqrt{\delta}} with |δδ|δ|\delta-\delta^{\prime}|\leq\sqrt{\delta}. We partition each plank in 𝕌θ\mathbb{U}_{\theta} into δ×δ×1\delta\times\delta\times 1-tubes with direction γ(θ)\gamma(\theta^{\prime}) and denote all these tubes by 𝕎θ\mathbb{W}_{\theta^{\prime}}. We also define 𝔻θ=πθ(𝕎θ)\mathbb{D}_{\theta^{\prime}}^{\prime}=\pi_{\theta}(\mathbb{W}_{\theta^{\prime}}) which is a collection of at most μ(3)δα21\mu(\mathbb{R}^{3})\delta^{-\frac{\alpha*}{2}-1} balls of radius δ\delta in πθ(3)\pi_{\theta}(\mathbb{R}^{3}).

Now we can apply Corollary 2 to the sets {𝔻θ}θΛδ\{\mathbb{D}_{\theta^{\prime}}^{\prime}\}_{\theta^{\prime}\in\Lambda_{\delta}} with α1=α2+1\alpha_{1}=\frac{\alpha^{*}}{2}+1. We obtain

δθΛδ(πθ#μ)(D𝔻θD)δθΛδ(πθ#μ)(D𝔻θD)C(α,α)μ(3)δε.\sqrt{\delta}\sum_{\theta\in\Lambda_{\sqrt{\delta}}}(\pi_{\theta\#}\mu)\Big{(}\bigcup_{D\in\mathbb{D}_{\theta}}D\Big{)}\sim\delta\sum_{\theta\in\Lambda_{\delta}}(\pi_{\theta\#}\mu)\Big{(}\bigcup_{D\in\mathbb{D}_{\theta}^{\prime}}D\Big{)}\leq C(\alpha,\alpha^{*})\mu(\mathbb{R}^{3})\delta^{\varepsilon}.

By Frostman’s Lemma (see for instance [20, page 112]), Theorem 2 is an immediate consequence of the following theorem.

Theorem 8.

Let γ:[0,1]𝕊2\gamma:[0,1]\to\mathbb{S}^{2} be C2C^{2} and non-degenerate. If μ\mu is a compactly supported Borel measure on 3\mathbb{R}^{3} such that cα(μ)<c_{\alpha}(\mu)<\infty for some α>2\alpha>2, then πθ#μ\pi_{\theta\#}\mu is absolutely continuous with respect to 2\mathcal{H}^{2}, for a.e. θ[0,1]\theta\in[0,1].

To prove Theorem 8, we will cut γ\gamma into finitely many pieces with a number depending only on γ\gamma, and work on one piece. From now on, we assume that γ:[0,a]𝕊2\gamma:[0,a]\to\mathbb{S}^{2} is C2C^{2} and non-degenerate, and satisfy

(82) γ(0)=(0,0,1),γ(0)=(1,0,0),|γ(θ)|=1,θ[0,a].\gamma(0)=(0,0,1),\ \ \gamma^{\prime}(0)=(1,0,0),\ |\gamma^{\prime}(\theta)|=1,\forall\theta\in[0,a].

Here a>0a>0 is sufficiently small depending on γ\gamma.

3.2. Decomposition of the frequency space

In this subsection, we discuss the decomposition of the frequency space ξ3\mathbb{R}^{3}_{\xi}. Recall the cone that we are considering:

Γ={r(γ×γ)(θ):r,θ[0,a]}.\Gamma=\{r(\gamma\times\gamma^{\prime})(\theta):r\in\mathbb{R},\theta\in[0,a]\}.

For any ξ=r(γ×γ)(θ)Γ\xi=r(\gamma\times\gamma^{\prime})(\theta)\in\Gamma, there are three directions that we would like to specify: the normal direction γ(θ)\gamma(\theta); the tangent direction γ(θ)\gamma^{\prime}(\theta); and the flat (or radial) direction γ×γ(θ)\gamma\times\gamma^{\prime}(\theta). We want to decompose ξ3\mathbb{R}^{3}_{\xi} into regions according to the distance from the origin, the distance from the cone Γ\Gamma, and the angular parameter. We give the precise definition below.

For a given integer kk, define

(83) Θk:=(2k/2)[0,a].\Theta_{k}:=\big{(}2^{-k/2}\mathbb{N}\big{)}\cap[0,a].

Fix kjk\leq j. If k<jk<j and {+,}\Box\in\{+,-\}, then we define the plank in the forward/backward light cone

(86) τ(θ,j,k,)={λ1(γ×γ)(θ)+λ2γ(θ)+λ3γ(θ):|λ1|2j,|λ2|Cγ12k/2+j,|λ3|2k+j,sgnλ1==sgnλ3}.\tau(\theta,j,k,\Box)=\Big{\{}\lambda_{1}\left(\gamma\times\gamma^{\prime}\right)(\theta)+\lambda_{2}\gamma^{\prime}(\theta)+\lambda_{3}\gamma(\theta)\\ :|\lambda_{1}|\sim 2^{j},|\lambda_{2}|\leq C_{\gamma}^{-1}2^{-k/2+j},|\lambda_{3}|\sim 2^{-k+j},\mathrm{sgn}\lambda_{1}=\Box=-\mathrm{sgn}\lambda_{3}\Big{\}}.

Here Cγ>0C_{\gamma}>0 is some large constant that depends only on γ\gamma. It is chosen such that the distance from τ(θ,j,k)\tau(\theta,j,k) to cone {(γ×γ)(θ):θ[0,a]}\{(\gamma\times\gamma^{\prime})(\theta):\theta\in[0,a]\} is comparable to 2jk2^{j-k}. Let us digest a little bit about the plank τ(θ,j,k)\tau(\theta,j,k): λ12j\lambda_{1}\sim 2^{j} is the distance from the origin; |λ3|2k+j|\lambda_{3}|\sim 2^{-k+j} is the distance from the cone Γ\Gamma; and |λ2|2k/2+j|\lambda_{2}|\lesssim 2^{-k/2+j} is the angular parameter of the plank.

If k=jk=j, then for {+,}\Box\in\{+,-\} we define

(89) τ(θ,j,j,)={|λ1|(γ×γ)(θ)+λ2γ(θ)+λ3γ(θ):λ12j,|λ2|2j/2,|λ3|1,sgnλ1=}.\tau(\theta,j,j,\Box)=\\ \Big{\{}|\lambda_{1}|\left(\gamma\times\gamma^{\prime}\right)(\theta)+\lambda_{2}\gamma^{\prime}(\theta)+\lambda_{3}\gamma(\theta):\lambda_{1}\sim 2^{j},\quad|\lambda_{2}|\lesssim 2^{j/2},\quad|\lambda_{3}|\lesssim 1,\mathrm{sgn}\lambda_{1}=\Box\Big{\}}.

Let

(90) Λj,k:={τ(θ,j,k,):θΘk,{,+}},\Lambda_{j,k}:=\{\tau(\theta,j,k,\Box):\theta\in\Theta_{k},\Box\in\{-,+\}\},

and

(91) Λj:=kjΛj,k,Λ:=jΛj.\Lambda_{j}:=\bigcup_{k\leq j}\Lambda_{j,k},\ \ \Lambda:=\bigcup_{j\in\mathbb{N}}\Lambda_{j}.

Roughly speaking, for k<jk<j, Λj,k\Lambda_{j,k} forms a canonical covering of the part of {ξ3:|ξ|2j;dist(ξ,Γ)2k+j}\{\xi\in\mathbb{R}^{3}:|\xi|\sim 2^{j};\textup{dist}(\xi,\Gamma)\sim 2^{-k+j}\} outside the cone Γ\Gamma by 2j×2k/2+j×2k+j2^{j}\times 2^{-k/2+j}\times 2^{-k+j}-planks. Each Λj\Lambda_{j} forms a covering of the 1\sim 1-neighbourhood of {ξΓ:|ξ|2j}\{\xi\in\Gamma:|\xi|\sim 2^{j}\}.

For each τΛ\tau\in\Lambda, we define 𝕋τ\mathbb{T}_{\tau} to be a set of planks of dual dimensions to τ\tau (but scaled by 2kδ2^{k\delta} in each direction where δ>0\delta>0 and τΛj,k\tau\in\Lambda_{j,k}) and forming a finitely overlapping covering of 3\mathbb{R}^{3}. We will refer to 𝕋τ\mathbb{T}_{\tau} as the wave packets determined by the plank τ\tau. Now, we discuss the wave packet decomposition. For each τΛ\tau\in\Lambda, we can choose a smooth bump function ψτ\psi_{\tau} supported in 2τ2\tau and choose a smooth bump function ψ0\psi_{0} supported in the unit ball, so that we have the partition of the unity

ψ0(ξ)+τΛψτ(ξ)=1,\psi_{0}(\xi)+\sum_{\tau\in\Lambda}\psi_{\tau}(\xi)=1,

on the union of the τ\tau’s. For each T𝕋τT\in\mathbb{T}_{\tau}, we can choose a smooth function ηT\eta_{T} which is essentially supported in TT (with rapidly decaying tail outside of TT), such that suppη^Tτ\mathrm{supp}\widehat{\eta}_{T}\subset\tau and

T𝕋τηT(x)=1.\sum_{T\in\mathbb{T}_{\tau}}\eta_{T}(x)=1.

For any T𝕋τT\in\mathbb{T}_{\tau}, we define the wave packet

MTμ:=ηT(μ\savestack\tmpbox\stretchto\scaleto\scalerel[width("ψτ")] 0.5ex\stackon[1pt]ψτ\tmpbox).M_{T}\mu:=\eta_{T}\big{(}\mu*\savestack{\tmpbox}{\stretchto{\scaleto{\scalerel*[width("\psi_{\tau}")]{\kern-0.6pt\bigwedge\kern-0.6pt}{\rule[-505.89pt]{4.30554pt}{505.89pt}}}{}}{0.5ex}}\stackon[1pt]{\psi_{\tau}}{\scalebox{-1.0}{\tmpbox}}\big{)}.
Lemma 4.

For τΛj,k\tau\in\Lambda_{j,k} and T𝕋τT\in\mathbb{T}_{\tau}, we have

MTμL1(3)23kδμ(2T)+CN2kNμ(3),\left\lVert M_{T}\mu\right\rVert_{L^{1}(\mathbb{R}^{3})}\lesssim 2^{3k\delta}\mu(2T)+C_{N}2^{-kN}\mu\left(\mathbb{R}^{3}\right),

for every N1N\geq 1.

Proof of Lemma 4.

Note that |\savestack\tmpbox\stretchto\scaleto\scalerel[width("ψτ")] 0.5ex\stackon[1pt]ψτ\tmpbox(x)|ϕτ(x)|\savestack{\tmpbox}{\stretchto{\scaleto{\scalerel*[width("\psi_{\tau}")]{\kern-0.6pt\bigwedge\kern-0.6pt}{\rule[-505.89pt]{4.30554pt}{505.89pt}}}{}}{0.5ex}}\stackon[1pt]{\psi_{\tau}}{\scalebox{-1.0}{\tmpbox}}(x)|\lesssim\phi_{\tau^{*}}(x), where ϕτ(x)\phi_{\tau^{*}}(x) is an L1L^{1} normalized function essentially supported in τ\tau^{*} (the dual plank of τ\tau). So, we have |ηT||\savestack\tmpbox\stretchto\scaleto\scalerel[width("ψτ")] 0.5ex\stackon[1pt]ψτ\tmpbox||ηT||\eta_{T}|*|\savestack{\tmpbox}{\stretchto{\scaleto{\scalerel*[width("\psi_{\tau}")]{\kern-0.6pt\bigwedge\kern-0.6pt}{\rule[-505.89pt]{4.30554pt}{505.89pt}}}{}}{0.5ex}}\stackon[1pt]{\psi_{\tau}}{\scalebox{-1.0}{\tmpbox}}|\lesssim|\eta_{T}|. Therefore,

|MTμ||μ|(|ηT||\savestack\tmpbox\stretchto\scaleto\scalerel[width("ψτ")] 0.5ex\stackon[1pt]ψτ\tmpbox|)|μ||ηT|23kδμ(2T)+CN2kNμ(3).\int|M_{T}\mu|\lesssim\int|\mu|\big{(}|\eta_{T}|*|\savestack{\tmpbox}{\stretchto{\scaleto{\scalerel*[width("\psi_{\tau}")]{\kern-0.6pt\bigwedge\kern-0.6pt}{\rule[-505.89pt]{4.30554pt}{505.89pt}}}{}}{0.5ex}}\stackon[1pt]{\psi_{\tau}}{\scalebox{-1.0}{\tmpbox}}|\big{)}\lesssim\int|\mu||\eta_{T}|\lesssim 2^{3k\delta}\mu(2T)+C_{N}2^{-kN}\mu(\mathbb{R}^{3}).

Lemma 5.

For τΛj,k\tau\in\Lambda_{j,k} and θ[0,a]\theta\in[0,a] with

(92) |θθτ|2k(1/2δ),|\theta-\theta_{\tau}|\geq 2^{-k(1/2-\delta)},

it holds that

(93) πθ#MTfL1(2)δ,N2kN|τ|fL1(3).\left\lVert\pi_{\theta\#}M_{T}f\right\rVert_{L^{1}(\mathcal{H}^{2})}\lesssim_{\delta,N}2^{-kN}|\tau|\|f\|_{L^{1}(\mathbb{R}^{3})}.

for every N1N\geq 1, T𝕋τT\in\mathbb{T}_{\tau} and fL1(3)f\in L^{1}(\mathbb{R}^{3}).

Proof of Lemma 5.

We start by writing

(94) MTf(x)=ηT(x)3f^(ξ)ψτ(ξ)eix,ξ𝑑ξ.M_{T}f(x)=\eta_{T}(x)\int_{\mathbb{R}^{3}}\widehat{f}(\xi)\psi_{\tau}(\xi)e^{i\langle x,\xi\rangle}d\xi.

By identifying the complex measure πθ#MTf\pi_{\theta\#}M_{T}f with its Radon-Nikodym derivative with respect to 2\mathcal{H}^{2}, we obtain

(95) πθ#MTf(x)=3f^(ξ)ψτ(ξ)[ηT(x+tγ(θ))eitγ(θ),ξ𝑑t]eix,ξ𝑑ξ\begin{split}\pi_{\theta\#}M_{T}f(x)=\int_{\mathbb{R}^{3}}\widehat{f}(\xi)\psi_{\tau}(\xi)\Big{[}\int_{\mathbb{R}}\eta_{T}(x+t\gamma(\theta))e^{it\langle\gamma(\theta),\xi\rangle}dt\Big{]}e^{i\langle x,\xi\rangle}d\xi\end{split}

for every xγ(θ)x\in\gamma(\theta)^{\perp}. It therefore suffices to show that

|ηT(x+tγ(θ))eitξ,γ(θ)𝑑t|N2kN,ξτ,x3.\left\lvert\int_{\mathbb{R}}\eta_{T}(x+t\gamma(\theta))e^{it\langle\xi,\gamma(\theta)\rangle}\,dt\right\rvert\lesssim_{N}2^{-kN},\quad\forall\,\xi\in\tau,\quad\forall\,x\in\mathbb{R}^{3}.

Integration by parts will finish the proof. For more details, we refer to Lemma 2.6 in [13]. ∎

3.3. Proof of Theorem 8: good part and bad part

The main idea is to divide the wave packets into two parts, called the good part and the bad part. We will prove an L1L^{1} estimate for the bad part and an L2L^{2} estimate for the good part.

Let α>2\alpha>2 be as in Theorem 8. Let ϵ>0\epsilon>0 be a small number (note ϵ\epsilon is different from ε\varepsilon) and let α0>0\alpha_{0}>0 be determined later (we will later let α02\alpha_{0}\nearrow 2). For jkj\geq k and τΛj,k\tau\in\Lambda_{j,k}, define

𝕋τ,b:={T𝕋τ:μ(4T)C2k(α0+1)/2α(jk)+103jϵ},𝕋τ,g=𝕋τ𝕋τ,b.\mathbb{T}_{\tau,b}:=\left\{T\in\mathbb{T}_{\tau}:\mu(4T)\geq C2^{-k(\alpha_{0}+1)/2-\alpha(j-k)+10^{3}j\epsilon}\right\},\qquad\mathbb{T}_{\tau,g}=\mathbb{T}_{\tau}\setminus\mathbb{T}_{\tau,b}.

Define

(96) μb=jk[jϵ,j]τΛj,kT𝕋τ,bMTμ,μg=μμb.\mu_{b}=\sum_{j\in\mathbb{N}}\sum_{\begin{subarray}{c}k\in[j\epsilon,j]\end{subarray}}\sum_{\tau\in\Lambda_{j,k}}\sum_{T\in\mathbb{T}_{\tau,b}}M_{T}\mu,\ \ \ \mu_{g}=\mu-\mu_{b}.

We remark that the wave packets of μb\mu_{b} are those that have heavy μ\mu-mass and not too far away from the cone Γ\Gamma. We have

μ=μg+μb.\mu=\mu_{g}+\mu_{b}.

We remark that μg=μg,α,ϵ,α0\mu_{g}=\mu_{g,\alpha,\epsilon,\alpha_{0}} and μb=μb,α,ϵ,α0\mu_{b}=\mu_{b,\alpha,\epsilon,\alpha_{0}} depends on parameters α,ϵ,α0\alpha,\epsilon,\alpha_{0}, but for simplicity we just omit them.

Theorem 8 follows from Lemma 6 and Lemma 7 below.

Lemma 6.

Let α>2\alpha>2 and ϵ1\epsilon\ll 1. Fix α0<2\alpha_{0}<2. For all Borel measures μ\mu supported on the unit ball in 3\mathbb{R}^{3} with cα(μ)1c_{\alpha}(\mu)\leq 1, it holds that

|πθ#μb|𝑑2𝑑θ1,\int\int\left\lvert\pi_{\theta\#}\mu_{b}\right\rvert\,d\mathcal{H}^{2}d\theta\lesssim 1,

where μb\mu_{b} is defined by (96), and the implicit constant depends on α,α0,ϵ\alpha,\alpha_{0},\epsilon and μ\mu.

Lemma 7.

Let α>2\alpha>2. Then for α0<2\alpha_{0}<2 sufficiently close to 22, and ϵ>0\epsilon>0 small enough depending on α\alpha and α0\alpha_{0}, and δϵ\delta\ll\epsilon,

(97) |πθ#μg|2𝑑2𝑑θ1,\int\int\left\lvert\pi_{\theta\#}\mu_{g}\right\rvert^{2}\,d\mathcal{H}^{2}\,d\theta\lesssim 1,

where the implicit constant depends on α,α0,ϵ\alpha,\alpha_{0},\epsilon and μ\mu.

Proof of Lemma 6.

By definition, we first write

(98) |πθ#μb|𝑑2𝑑θjk[jϵ,j]τΛj,kT𝕋τ,b|πθ#MTμ|𝑑2𝑑θ.\int\int\left\lvert\pi_{\theta\#}\mu_{b}\right\rvert d\mathcal{H}^{2}d\theta\leq\int\sum_{j\in\mathbb{N}}\sum_{\begin{subarray}{c}k\in[j\epsilon,j]\end{subarray}}\sum_{\tau\in\Lambda_{j,k}}\sum_{T\in\mathbb{T}_{\tau,b}}\int\left\lvert\pi_{\theta\#}M_{T}\mu\right\rvert d\mathcal{H}^{2}d\theta.

By the triangle inequality, this is

(99) jk[jϵ,j]τΛj,k:|θτθ|<2k(1/2+δ)T𝕋τ,b|πθ#MTμ|d2dθ\displaystyle\leq\sum_{j}\sum_{\begin{subarray}{c}k\in[j\epsilon,j]\end{subarray}}\int\int\sum_{\begin{subarray}{c}\tau\in\Lambda_{j,k}:\\ \left\lvert\theta_{\tau}-\theta\right\rvert<2^{k(-1/2+\delta)}\end{subarray}}\sum_{T\in\mathbb{T}_{\tau,b}}\left\lvert\pi_{\theta\#}M_{T}\mu\right\rvert d\mathcal{H}^{2}d\theta
(100) +jk[jϵ,j]τΛj,k:|θτθ|2k(1/2+δ)T𝕋τ,b|πθ#MTμ|d2dθ.\displaystyle\qquad+\sum_{j}\sum_{\begin{subarray}{c}k\in[j\epsilon,j]\end{subarray}}\int\int\sum_{\begin{subarray}{c}\tau\in\Lambda_{j,k}:\\ \left\lvert\theta_{\tau}-\theta\right\rvert\geq 2^{k(-1/2+\delta)}\end{subarray}}\sum_{T\in\mathbb{T}_{\tau,b}}\left\lvert\pi_{\theta\#}M_{T}\mu\right\rvert d\mathcal{H}^{2}d\theta.

By Lemma 5, the contribution from (100) is

δ,ϵ,Njk[ϵj,j]23j32kkNμ(3)δ,ϵjk[ϵj,j]2jμ(3)μ(3),\lesssim_{\delta,\epsilon,N}\sum_{j}\sum_{k\in[\epsilon j,j]}2^{3j-\frac{3}{2}k-kN}\mu\left(\mathbb{R}^{3}\right)\lesssim_{\delta,\epsilon}\sum_{j}\sum_{k\in[\epsilon j,j]}2^{-j}\mu\left(\mathbb{R}^{3}\right)\lesssim\mu(\mathbb{R}^{3}),

By choosing N>100ε1N>100\varepsilon^{-1}.

To estimate (99), we discretize the integration in θ\theta and bound it by

(101) jk[jϵ,j]θΘk2k/2τΛj,k:|θτθ|<2k(1/2+δ)T𝕋τ,b|πθ#MTμ|d2\sum_{j}\sum_{\begin{subarray}{c}k\in[j\epsilon,j]\end{subarray}}\sum_{\theta\in\Theta_{k}}2^{-k/2}\int\sum_{\begin{subarray}{c}\tau\in\Lambda_{j,k}:\\ \left\lvert\theta_{\tau}-\theta\right\rvert<2^{k(-1/2+\delta)}\end{subarray}}\sum_{T\in\mathbb{T}_{\tau,b}}\left\lvert\pi_{\theta\#}M_{T}\mu\right\rvert d\mathcal{H}^{2}

By Lemma 4 the contribution from (99) is

μ(3)+jk[jϵ,j]θ2k/2τΛj,k:|θτθ|<2k(1/2+δ)T𝕋τ,b23jδμ(2T)\displaystyle\lesssim\mu\left(\mathbb{R}^{3}\right)+\sum_{j}\sum_{\begin{subarray}{c}k\in[j\epsilon,j]\end{subarray}}\sum_{\theta}2^{-k/2}\sum_{\begin{subarray}{c}\tau\in\Lambda_{j,k}:\\ \left\lvert\theta_{\tau}-\theta\right\rvert<2^{k(-1/2+\delta)}\end{subarray}}\sum_{T\in\mathbb{T}_{\tau,b}}2^{3j\delta}\mu(2T)
(102) μ(3)+jk[jϵ,j]θ2k/22100jδμ(Bj,k(θ)),\displaystyle\lesssim\mu\left(\mathbb{R}^{3}\right)+\sum_{j}\sum_{\begin{subarray}{c}k\in[j\epsilon,j]\end{subarray}}\sum_{\theta}2^{-k/2}2^{100j\delta}\mu(B_{j,k}(\theta)),

where

Bj,k(θ)=τΛj,k:|θτθ|<2k(1/2+δ)T𝕋τ,b2T.B_{j,k}(\theta)=\bigcup_{\begin{subarray}{c}\tau\in\Lambda_{j,k}:\\ |\theta_{\tau}-\theta|<2^{k(-1/2+\delta)}\end{subarray}}\bigcup_{T\in\mathbb{T}_{\tau,b}}2T.

For fixed jj and kk, let {Bl}l\{B_{l}\}_{l} be a finitely overlapping cover of the unit ball in 3\mathbb{R}^{3} by balls of radius 2(jk)2^{-(j-k)}. For each θ\theta and ll let

Bj,k,l(θ)=τΛj,k:|θτθ|<2k(1/2+δ)T𝕋τ,b:2TBl2T.B_{j,k,l}(\theta)=\bigcup_{\begin{subarray}{c}\tau\in\Lambda_{j,k}:\\ |\theta_{\tau}-\theta|<2^{k(-1/2+\delta)}\end{subarray}}\bigcup_{\begin{subarray}{c}T\in\mathbb{T}_{\tau,b}:\\ 2T\cap B_{l}\neq\emptyset\end{subarray}}2T.

Let μj,k\mu_{j,k} be the pushforward of μ\mu under x2jk2kδxx\mapsto 2^{j-k-2k\delta}x. Denote

(103) Bj,k,l(θ):=2jk2kδBj,k,l(θ),B~l={2jk2kδbl+y:|y|1,y3},B^{\prime}_{j,k,l}(\theta):=2^{j-k-2k\delta}\cdot B_{j,k,l}(\theta),\ \ \widetilde{B}_{l}=\left\{2^{j-k-2k\delta}b_{l}+y:|y|\leq 1,y\in\mathbb{R}^{3}\right\},

with blb_{l} the centre of BlB_{l}, and define

(104) μ~j,k,l=2α(jk2kδ)μj,kχB~l.\widetilde{\mu}_{j,k,l}=2^{\alpha(j-k-2k\delta)}\cdot\mu_{j,k}\chi_{\widetilde{B}_{l}}.

Then

(105) θμ(Bj,k(θ))θlμ(Bj,k,l(θ))l2α(jk2kδ)θμ~j,k,l(Bj,k,l(θ)).\sum_{\theta}\mu(B_{j,k}(\theta))\leq\sum_{\theta}\sum_{l}\mu(B_{j,k,l}(\theta))\leq\sum_{l}2^{-\alpha(j-k-2k\delta)}\sum_{\theta}\widetilde{\mu}_{j,k,l}(B^{\prime}_{j,k,l}(\theta)).

Note that for each θ\theta, the set Bj,k,l(θ)B^{\prime}_{j,k,l}(\theta) is contained in a union of planks of dimensions 1×2k/2×2k1\times 2^{-k/2}\times 2^{-k}; the number of planks is

(106) 2kα0+12+Cδkμ~j,k,l(3),\lesssim 2^{k\frac{\alpha_{0}+1}{2}+C\delta k}\widetilde{\mu}_{j,k,l}(\mathbb{R}^{3}),

for some large constant CC, and each plank overlaps 210kδ\lesssim 2^{10k\delta} of the others. Moreover cα(μ~j,k,l)1c_{\alpha}\left(\widetilde{\mu}_{j,k,l}\right)\leq 1 and μ~j,k,l\widetilde{\mu}_{j,k,l} is supported in a ball of radius 1. Therefore, by applying the triangle inequality and Corollary 3, we can find δ\delta^{\prime}, depending only on α\alpha and α0\alpha_{0}, such that

2k/2θμ~j,k,l(Bj,k,l(θ))2Cδk+Cδkμ~j,k,l(3)2Cδk+Cδk2α(jk2kδ)μj,k(B~l),2^{-k/2}\sum_{\theta}\widetilde{\mu}_{j,k,l}(B^{\prime}_{j,k,l}(\theta))\lesssim 2^{-C^{\prime}\delta^{\prime}k+C^{\prime}\delta k}\widetilde{\mu}_{j,k,l}(\mathbb{R}^{3})\\ \lesssim 2^{-C^{\prime}\delta^{\prime}k+C^{\prime}\delta k}2^{\alpha(j-k-2k\delta)}\mu_{j,k}(\widetilde{B}_{l}),

for some large constant CC^{\prime}, whose precise value is not important. Putting this into (105) yields

(109) (105)2k/22Cδk+Cδkμ(3).\eqref{tildemu}\lesssim 2^{k/2}2^{-C^{\prime}\delta^{\prime}k+C^{\prime}\delta k}\mu(\mathbb{R}^{3}).

Substituting this into (102) and then (98) gives

(110) (98)jk[ϵj,j]2100δj2Cδk+Cδkμ(3)\eqref{badbound}\lesssim\sum_{j}\sum_{\begin{subarray}{c}k\in[\epsilon j,j]\end{subarray}}2^{100\delta j}2^{-C^{\prime}\delta^{\prime}k+C^{\prime}\delta k}\mu(\mathbb{R}^{3})

Recall that δ\delta^{\prime} depends only on α\alpha and α0\alpha_{0}. We just need to pick δ\delta to be sufficiently small, and will finish the proof. ∎

Proof of Lemma 7.

Take ϵmin{α2,2α0}\epsilon\ll\min\{\alpha-2,2-\alpha_{0}\}. Given xπθ(3)x\in\pi_{\theta}(\mathbb{R}^{3}), note that

(111) πθ#μg(x)=μg(x+tγ(θ))𝑑t.\pi_{\theta\#}\mu_{g}(x)=\int\mu_{g}(x+t\gamma(\theta))dt.

Fix the coordinate (𝐞1,𝐞2,𝐞3)=(γ(θ),γ(θ)×γ(θ),γ(θ))({\mathbf{e}}_{1},{\mathbf{e}}_{2},{\mathbf{e}}_{3})=(\gamma^{\prime}(\theta),\gamma(\theta)\times\gamma^{\prime}(\theta),\gamma(\theta)). Any xπθ(3)x\in\pi_{\theta}(\mathbb{R}^{3}) can be written in this coordinate as x=(x1,x2,0)x=(x_{1},x_{2},0). We can also rewrite (111) as

πθ#μg(x1,x2)=μg(x1,x2,t)𝑑t.\pi_{\theta\#}\mu_{g}(x_{1},x_{2})=\int\mu_{g}(x_{1},x_{2},t)dt.

Doing the Fourier transfom in the (𝐞1,𝐞2)({\mathbf{e}}_{1},{\mathbf{e}}_{2})-plane, we have

(πθ#μg)(η1,η2)=μg(x1,x2,t)ei(x1η1+x2η2)𝑑t=μg^(η1,η2,0)=μg^(η1γ(θ)+η2(γ×γ)(θ)).(\pi_{\theta\#}\mu_{g})^{\wedge}(\eta_{1},\eta_{2})=\int\mu_{g}(x_{1},x_{2},t)e^{-i(x_{1}\eta_{1}+x_{2}\eta_{2})}\,dt\\ =\widehat{\mu_{g}}(\eta_{1},\eta_{2},0)=\widehat{\mu_{g}}(\eta_{1}\gamma^{\prime}(\theta)+\eta_{2}(\gamma\times\gamma^{\prime})(\theta)).

By Plancherel’s theorem,

(114) |πθ#μg|2𝑑2𝑑θ=2|μg^(η1γ(θ)+η2(γ×γ)(θ))|2𝑑η𝑑θ.\int\int\left\lvert\pi_{\theta\#}\mu_{g}\right\rvert^{2}\,d\mathcal{H}^{2}\,d\theta=\int\int_{\mathbb{R}^{2}}\left\lvert\widehat{\mu_{g}}\left(\eta_{1}\gamma^{\prime}(\theta)+\eta_{2}\left(\gamma\times\gamma^{\prime}\right)(\theta)\right)\right\rvert^{2}\,d\eta\,d\theta.

Roughly speaking,

μg=μ0+μg,1+μg,2,\mu_{g}=\mu_{0}+\mu_{g,1}+\mu_{g,2},

where μ0\mu_{0} is roughly μ(3)𝟏B3(0,1)\mu(\mathbb{R}^{3})\boldsymbol{1}_{B^{3}(0,1)} with rapidly decaying tail outside B3(0,1)B^{3}(0,1), μg,1\mu_{g,1} is the sum of good wave packets which have controlled mass, and μg,2\mu_{g,2} is the sum of wave packets which are far away from the cone Γ\Gamma. A formula for μg,1\mu_{g,1} is

μg,1=jk[jϵ,j]τΛj,kT𝕋τ,gMTμ.\mu_{g,1}=\sum_{j\in\mathbb{N}}\sum_{\begin{subarray}{c}k\in[j\epsilon,j]\end{subarray}}\sum_{\tau\in\Lambda_{j,k}}\sum_{T\in\mathbb{T}_{\tau,g}}M_{T}\mu.

The above used that 𝕋τ,b\mathbb{T}_{\tau,b} is empty when τΛj,k\tau\in\Lambda_{j,k}, kjϵk\leq j\epsilon and jj is sufficiently large, which follows from the Frostman condition on μ\mu.

Claim 1.

Let τΛj,k\tau\in\Lambda_{j,k} with k<jk<j. If there exist θ[0,a]\theta\in[0,a] and (η1,η2)(\eta_{1},\eta_{2}) satisfying

(115) η1γ(θ)+η2(γ×γ)(θ)supp(ψτ),\eta_{1}\gamma^{\prime}(\theta)+\eta_{2}\left(\gamma\times\gamma^{\prime}\right)(\theta)\in\mathrm{supp}(\psi_{\tau}),

then it holds that |η2|2j|\eta_{2}|\sim 2^{j} and |η1|2jk/2|\eta_{1}|\sim 2^{j-k/2}.

Proof of Claim 1.

Recalling (86), we may assume

τ=τ(θ,j,k)={λ1γ(θ)+λ2(γ×γ)(θ)+λ3γ(θ):|λ1|2k/2+j,λ22j,|λ3|2k+j}.\tau=\tau(\theta^{\prime},j,k)\\ =\Big{\{}\lambda_{1}\gamma^{\prime}(\theta^{\prime})+\lambda_{2}\left(\gamma\times\gamma^{\prime}\right)(\theta^{\prime})+\lambda_{3}\gamma(\theta^{\prime}):|\lambda_{1}|\lesssim 2^{-k/2+j},\lambda_{2}\sim 2^{j},|\lambda_{3}|\sim 2^{-k+j}\Big{\}}.

If η1γ(θ)+η2(γ×γ)(θ)τ(θ,j,k),\eta_{1}\gamma^{\prime}(\theta)+\eta_{2}\left(\gamma\times\gamma^{\prime}\right)(\theta)\in\tau(\theta^{\prime},j,k), we discuss some geometric observations. Noting that |λ3|2k+j|\lambda_{3}|\sim 2^{-k+j} in the definition of τ(θ,j,k)\tau(\theta^{\prime},j,k), we see that η1γ(θ)+η2(γ×γ)(θ)τ(θ,j,k)\eta_{1}\gamma^{\prime}(\theta)+\eta_{2}\left(\gamma\times\gamma^{\prime}\right)(\theta)\notin\tau(\theta^{\prime},j,k) if θ=θ\theta^{\prime}=\theta; we also note that if |θθ|2k/2|\theta^{\prime}-\theta|\gg 2^{-k/2} are too far apart, then η1γ(θ)+η2(γ×γ)(θ)τ(θ,j,k)\eta_{1}\gamma^{\prime}(\theta)+\eta_{2}\left(\gamma\times\gamma^{\prime}\right)(\theta)\notin\tau(\theta^{\prime},j,k). Therefore we must have |θθ|2k/2|\theta^{\prime}-\theta|\sim 2^{-k/2}. In this case, in order for η1γ(θ)+η2(γ×γ)(θ)τ(θ,j,k)\eta_{1}\gamma^{\prime}(\theta)+\eta_{2}\left(\gamma\times\gamma^{\prime}\right)(\theta)\in\tau(\theta^{\prime},j,k), we must have |η2|2j|\eta_{2}|\sim 2^{j} and |η1|2jk/2|\eta_{1}|\sim 2^{j-k/2}, which finishes the proof. ∎

By Claim 1,we see that (114) is bounded by

(118) 1+{|η1||η2|1ϵ}|μ^(η1γ(θ)+η2(γ×γ)(θ))|2𝑑η𝑑θ\displaystyle 1+\int\int_{\{|\eta_{1}|\geq|\eta_{2}|^{1-\epsilon}\}}\left\lvert\widehat{\mu}\left(\eta_{1}\gamma^{\prime}(\theta)+\eta_{2}\left(\gamma\times\gamma^{\prime}\right)(\theta)\right)\right\rvert^{2}\,d\eta\,d\theta
(119) +{|η1|<|η2|1ϵ}|μg,1^(η1γ(θ)+η2(γ×γ)(θ))|2𝑑η𝑑θ.\displaystyle\quad+\int\int_{\{|\eta_{1}|<|\eta_{2}|^{1-\epsilon}\}}\left\lvert\widehat{\mu_{g,1}}\left(\eta_{1}\gamma^{\prime}(\theta)+\eta_{2}\left(\gamma\times\gamma^{\prime}\right)(\theta)\right)\right\rvert^{2}\,d\eta\,d\theta.

For the first term, the change of variables

(120) ξ=ξ(η,θ)=η1γ(θ)+η2(γ×γ)(θ)\xi=\xi(\eta,\theta)=\eta_{1}\gamma^{\prime}(\theta)+\eta_{2}\left(\gamma\times\gamma^{\prime}\right)(\theta)

has Jacobian

(121) |(ξ1,ξ2,ξ3)(η1,η2,θ)(η1,η2,θ)|\displaystyle\left\lvert\frac{\partial(\xi_{1},\xi_{2},\xi_{3})}{\partial(\eta_{1},\eta_{2},\theta)}(\eta_{1},\eta_{2},\theta)\right\rvert =|η1||det((γ×γ)(θ),γ(θ),γ′′(θ))|\displaystyle=\left\lvert\eta_{1}\right\rvert\left\lvert\det\left(\left(\gamma\times\gamma^{\prime}\right)(\theta),\gamma^{\prime}(\theta),\gamma^{\prime\prime}(\theta)\right)\right\rvert
=|η1||γ(θ),γ′′(θ)|=|η1|,\displaystyle=|\eta_{1}|\left\lvert\left\langle\gamma(\theta),\gamma^{\prime\prime}(\theta)\right\rangle\right\rvert=|\eta_{1}|,

where in the last step we used

(122) γ(t),γ(t)0γ(t),γ′′(t)1.\langle\gamma(t),\gamma^{\prime}(t)\rangle\equiv 0\implies\langle\gamma(t),\gamma^{\prime\prime}(t)\rangle\equiv-1.

Applying this change of variables to (118) gives

(118)1+|ξ|1|ξ|ϵ1|μ^(ξ)|2𝑑ξ1+Iαϵ(μ)1.\displaystyle\eqref{energymu}\lesssim 1+\int_{|\xi|\geq 1}\left\lvert\xi\right\rvert^{\epsilon-1}\left\lvert\widehat{\mu}(\xi)\right\rvert^{2}\,d\xi\lesssim 1+I_{\alpha-\epsilon}(\mu)\lesssim 1.

Here Iαϵ(μ)=|ξ|αε3|μ^(ξ)|2𝑑ξI_{\alpha-\epsilon}(\mu)=\int|\xi|^{\alpha-\varepsilon-3}|\widehat{\mu}(\xi)|^{2}\,d\xi is the (αϵ)(\alpha-\epsilon)-energy of μ\mu and we used the fact that α>2\alpha>2 and ϵ\epsilon is sufficiently small. The last step is because cα(μ)<c_{\alpha}(\mu)<\infty.

It remains to bound the contribution from μg,1\mu_{g,1}, in (119). By frequency disjointness,

(123) 2|μg,1^(η1γ(θ)+η2(γ×γ)(θ))|2𝑑η𝑑θjk[ϵj,j]τΛj,k2|T𝕋τ,gMTμ^(η1γ(θ)+η2(γ×γ)(θ))|2𝑑η𝑑θ.\begin{split}&\int\int_{\mathbb{R}^{2}}\left\lvert\widehat{\mu_{g,1}}\left(\eta_{1}\gamma^{\prime}(\theta)+\eta_{2}\left(\gamma\times\gamma^{\prime}\right)(\theta)\right)\right\rvert^{2}\,d\eta\,d\theta\\ &\lesssim\sum_{j}\sum_{k\in[\epsilon j,j]}\sum_{\tau\in\Lambda_{j,k}}\int\int_{\mathbb{R}^{2}}\Big{|}\sum_{T\in\mathbb{T}_{\tau,g}}\widehat{M_{T}\mu}\left(\eta_{1}\gamma^{\prime}(\theta)+\eta_{2}\left(\gamma\times\gamma^{\prime}\right)(\theta)\right)\Big{|}^{2}d\eta\,d\theta.\end{split}

Consider the case k<jk<j and k=jk=j separately. In the former case, we apply the change of variables as in (120) and obtain

(128) τΛj,k2|T𝕋τ,gMTμ^(η1γ(θ)+η2(γ×γ)(θ))|2𝑑η𝑑θτΛj,k2j+k/23|T𝕋τ,gMTμ^(ξ)|2𝑑ξτΛj,kT𝕋τ,g2j+k/22O(δ)k3|MTμ(x)|2𝑑x.\sum_{\tau\in\Lambda_{j,k}}\int\int_{\mathbb{R}^{2}}\Big{|}\sum_{T\in\mathbb{T}_{\tau,g}}\widehat{M_{T}\mu}\left(\eta_{1}\gamma^{\prime}(\theta)+\eta_{2}\left(\gamma\times\gamma^{\prime}\right)(\theta)\right)\Big{|}^{2}d\eta\,d\theta\\ \\ \lesssim\sum_{\tau\in\Lambda_{j,k}}2^{-j+k/2}\int_{\mathbb{R}^{3}}\Big{|}\sum_{T\in\mathbb{T}_{\tau,g}}\widehat{M_{T}\mu}(\xi)\Big{|}^{2}d\xi\\ \lesssim\sum_{\tau\in\Lambda_{j,k}}\sum_{T\in\mathbb{T}_{\tau,g}}2^{-j+k/2}2^{O(\delta)k}\int_{\mathbb{R}^{3}}\Big{|}M_{T}\mu(x)\Big{|}^{2}dx.

When k=jk=j, we show that (128) holds as well. To see this, we first observe that for each fixed TT, in order for

(129) MTμ^(η1γ(θ)+η2(γ×γ)(θ))\widehat{M_{T}\mu}\left(\eta_{1}\gamma^{\prime}(\theta)+\eta_{2}\left(\gamma\times\gamma^{\prime}\right)(\theta)\right)

not to vanish, θ\theta has to take values on an interval of length 2j/22^{-j/2}; next, we apply the two dimensional Plancherel’s theorem in the η1\eta_{1} and η2\eta_{2} variables for every fixed θ\theta, and (128) follows from the uncertainty principle.

We continue to estimate (128) and do not distinguish k<jk<j and k=jk=j anymore. We have

(130) (128)\displaystyle\eqref{220623e5_33} τΛj,kT𝕋τ,g2j+k/22O(δ)k3|MTμ(x)|2𝑑x\displaystyle\lesssim\sum_{\tau\in\Lambda_{j,k}}\sum_{T\in\mathbb{T}_{\tau,g}}2^{-j+k/2}2^{O(\delta)k}\int_{\mathbb{R}^{3}}\Big{|}M_{T}\mu(x)\Big{|}^{2}dx
(131) =2j+k/22O(δ)kτΛj,kT𝕋τ,gfTdμ\displaystyle=2^{-j+k/2}2^{O(\delta)k}\int\sum_{\tau\in\Lambda_{j,k}}\sum_{T\in\mathbb{T}_{\tau,g}}f_{T}d\mu

where

(132) fT:=(ηTMTμ)\savestack\tmpbox\stretchto\scaleto\scalerel[width("ψτ")] 0.5ex\stackon[1pt]ψτ\tmpbox,f_{T}:=(\eta_{T}M_{T}\mu)*\savestack{\tmpbox}{\stretchto{\scaleto{\scalerel*[width("\psi_{\tau}")]{\kern-0.6pt\bigwedge\kern-0.6pt}{\rule[-505.89pt]{4.30554pt}{505.89pt}}}{}}{0.5ex}}\stackon[1pt]{\psi_{\tau}}{\scalebox{-1.0}{\tmpbox}},

and from (130) to (131) we applied Fubini and expanded the square. We cut the unit ball into small balls BιB_{\iota} of radius 2j+k2^{-j+k} and let νι\nu_{\iota} be the restriction of μ\mu to 210δkBι2^{10\delta k}B_{\iota}. By Cauchy-Schwarz,

(133) (131)2j+k/22O(δ)kιμ(210kδBι)1/2(|τΛj,kT𝕋τ,gfT|2𝑑νι)1/2\eqref{220623e5_36}\lesssim 2^{-j+k/2}2^{O(\delta)k}\sum_{\iota}\mu(2^{10k\delta}B_{\iota})^{1/2}\Big{(}\int\Big{|}\sum_{\tau\in\Lambda_{j,k}}\sum_{T\in\mathbb{T}_{\tau,g}}f_{T}\Big{|}^{2}d\nu_{\iota}\Big{)}^{1/2}

Let ζj\zeta_{j} be a non-negative bump function such that ζj^(ξ)=1\widehat{\zeta_{j}}(\xi)=1 for |ξ|2j+10|\xi|\leq 2^{j+10}. By the Fourier support information of fTf_{T}, we have

(134) |τΛj,kT𝕋τ,gfT|2𝑑νι=|τΛj,kT𝕋τ,gfT|2d(νιζj)\int\Big{|}\sum_{\tau\in\Lambda_{j,k}}\sum_{T\in\mathbb{T}_{\tau,g}}f_{T}\Big{|}^{2}d\nu_{\iota}=\int\Big{|}\sum_{\tau\in\Lambda_{j,k}}\sum_{T\in\mathbb{T}_{\tau,g}}f_{T}\Big{|}^{2}d(\nu_{\iota}*\zeta_{j})

By pigeonholing, we can find a subset

(135) 𝕎ιτΛj,k{T𝕋τ,g:TBι}\mathbb{W}_{\iota}\subset\bigcup_{\tau\in\Lambda_{j,k}}\{T\in\mathbb{T}_{\tau,g}:T\cap B_{\iota}\neq\emptyset\}

such that fT2\big{\|}f_{T}\big{\|}_{2} is constant up to a factor of 22 as TT varies over 𝕎ι\mathbb{W}_{\iota}, and

(136) (134)2O(δ)j|T𝕎ιfT|2d(νιζj)\displaystyle\eqref{220623e5_39}\lesssim 2^{O(\delta)j}\int\Big{|}\sum_{T\in\mathbb{W}_{\iota}}f_{T}\Big{|}^{2}d(\nu_{\iota}*\zeta_{j})

By pigeonholing again and by Hölder’s inequality, there is a disjoint union YY of balls QQ of radius 2j2^{-j}, such that

(137) |T𝕎ιfT|2d(νιζj)2O(δ)jT𝕎ιfTLp(Y)2(Y(νιζj)pp2)12p,\int\Big{|}\sum_{T\in\mathbb{W}_{\iota}}f_{T}\Big{|}^{2}d(\nu_{\iota}\ast\zeta_{j})\lesssim 2^{O(\delta)j}\big{\|}\sum_{T\in\mathbb{W}_{\iota}}f_{T}\big{\|}_{L^{p}(Y)}^{2}\left(\int_{Y}\left(\nu_{\iota}\ast\zeta_{j}\right)^{\frac{p}{p-2}}\right)^{1-\frac{2}{p}},

and such that each QYQ\subseteq Y intersects M\sim M planks 3T3T as TT varies over 𝕎ι\mathbb{W}_{\iota}, for some dyadic number MM. By rescaling and then applying the refined decoupling inequality in Theorem 9 from the Appendix A, the first term in (137) satisfies

(138) T𝕎ιfTLp(Y)2(3j3k2)(121p)+O(ϵ)j(M|𝕎ι|)121p(T𝕎ιfT22)1/2.\displaystyle\Big{\|}\sum_{T\in\mathbb{W}_{\iota}}f_{T}\Big{\|}_{L^{p}(Y)}\lesssim 2^{(3j-\frac{3k}{2})(\frac{1}{2}-\frac{1}{p})+O(\epsilon)j}\left(\frac{M}{\left\lvert\mathbb{W}_{\iota}\right\rvert}\right)^{\frac{1}{2}-\frac{1}{p}}\left(\sum_{T\in\mathbb{W}_{\iota}}\left\lVert f_{T}\right\rVert_{2}^{2}\right)^{1/2}.

For the second term in (137), the assumption that cα(μ)c_{\alpha}(\mu) is finite implies that

(139) νιζj2j(3α).\|\nu_{\iota}\ast\zeta_{j}\|_{\infty}\lesssim 2^{j(3-\alpha)}.

Hence by Hölder’s inequality and the definition of 𝕋τ,g\mathbb{T}_{\tau,g}, we have

(140) Y(νιζj)pp2\displaystyle\int_{Y}\left(\nu_{\iota}\ast\zeta_{j}\right)^{\frac{p}{p-2}} 22j(3α)p2M1T𝕎ι(νιζj)(3T)\displaystyle\lesssim 2^{\frac{2j(3-\alpha)}{p-2}}M^{-1}\sum_{T\in\mathbb{W}_{\iota}}\left(\nu_{\iota}\ast\zeta_{j}\right)(3T)
(141) 22j(3α)p2M1|𝕎ι|2k(α0+1)/2α(jk).\displaystyle\lesssim 2^{\frac{2j(3-\alpha)}{p-2}}M^{-1}\left\lvert\mathbb{W}_{\iota}\right\rvert 2^{-k(\alpha_{0}+1)/2-\alpha(j-k)}.

Combining (138) and (141), we obtain

(142) (137)2O(ϵ)j2j(3α)+k(121p)(4+2αα0)T𝕎ιfT22.\eqref{holder}\lesssim 2^{O(\epsilon)j}2^{j(3-\alpha)+k\left(\frac{1}{2}-\frac{1}{p}\right)\left(-4+2\alpha-\alpha_{0}\right)}\sum_{T\in\mathbb{W}_{\iota}}\left\lVert f_{T}\right\rVert_{2}^{2}.

Note that

(143) fT2MTμ2\displaystyle\|f_{T}\|_{2}\lesssim\|M_{T}\mu\|_{2}

for every TT. Substituting into (142) and then into (130) yields

(144) τΛj,kT𝕋τ,g3|MTμ(x)|2𝑑x\displaystyle\sum_{\tau\in\Lambda_{j,k}}\sum_{T\in\mathbb{T}_{\tau,g}}\int_{\mathbb{R}^{3}}\Big{|}M_{T}\mu(x)\Big{|}^{2}dx
(145) ιμ(210kδBι)1/22O(ϵ)j212j(3α)+12k(121p)(4+2αα0)(T𝕎ιfT22)1/2.\displaystyle\lesssim\sum_{\iota}\mu(2^{10k\delta}B_{\iota})^{1/2}2^{O(\epsilon)j}2^{\frac{1}{2}j(3-\alpha)+\frac{1}{2}k\left(\frac{1}{2}-\frac{1}{p}\right)\left(-4+2\alpha-\alpha_{0}\right)}\Big{(}\sum_{T\in\mathbb{W}_{\iota}}\left\lVert f_{T}\right\rVert_{2}^{2}\Big{)}^{1/2}.

By Cauchy-Schwarz in the sum over ι\iota, we obtain

(146) τΛj,kT𝕋τ,g3|MTμ(x)|2𝑑x2O(ϵ)j2j(3α)+k(121p)(4+2αα0).\displaystyle\sum_{\tau\in\Lambda_{j,k}}\sum_{T\in\mathbb{T}_{\tau,g}}\int_{\mathbb{R}^{3}}\Big{|}M_{T}\mu(x)\Big{|}^{2}dx\lesssim 2^{O(\epsilon)j}2^{j(3-\alpha)+k\left(\frac{1}{2}-\frac{1}{p}\right)\left(-4+2\alpha-\alpha_{0}\right)}.

By substituting back into (130), we obtain

(147) (128)jϵjkj2j+k/22O(ϵ)j2j(3α)+k(121p)(4+2αα0).\eqref{220623e5_33}\lesssim\sum_{j}\sum_{\epsilon j\leq k\leq j}2^{-j+k/2}2^{O(\epsilon)j}2^{j(3-\alpha)+k\left(\frac{1}{2}-\frac{1}{p}\right)\left(-4+2\alpha-\alpha_{0}\right)}.

In the end, we pick p=4p=4 and finish the proof. ∎

Appendix A Refined decoupling inequality

The refined decoupling inequality stated here is a natural analogue of the refined decoupling inequality for the paraboloid from [9]. The shortest length of a plank dual to an R1/2R^{-1/2}-cap in the cone is 1\approx 1, rather than R1/2\approx R^{1/2} in the case of the paraboloid, so the setup uses unit cubes instead of R1/2R^{1/2}-cubes. The argument is similar to the paraboloid case, using induction and with Lorentz rescaling in place of parabolic rescaling, but the use of unit cubes requires the induction to be carried out over a finer sequence of scales (similarly to the induction setup in [4]). The refined decoupling inequalities in [13, 14] used tubes rather than planks, and the use of planks here is a significant reason for the improved result on the projection problem (at least in Theorem 2). Much of the proof is similar to [13, 14], with only the differences outlined above.

For each R1R\geq 1 let ΞR={jR1/2:j}[0,a].\Xi_{R}=\left\{jR^{-1/2}:j\in\mathbb{Z}\right\}\cap[0,a]. For each θΞR\theta\in\Xi_{R}, let

(148) τR(θ)={x1γ(θ)+x2γ(θ)+x3(γ×γ)(θ):1x12,|x2|R1/2,|x3|R1}.\tau_{R}(\theta)=\bigg{\{}x_{1}\gamma(\theta)+x_{2}\gamma^{\prime}(\theta)+x_{3}(\gamma\times\gamma^{\prime})(\theta):1\leq x_{1}\leq 2,\,|x_{2}|\leq R^{-1/2},\,|x_{3}|\leq R^{-1}\bigg{\}}.

If it is clear from the context which RR is used, then we often abbreviate τR(θ)\tau_{R}(\theta) to τ(θ)\tau(\theta). Let 𝒫R1={τ(θ):θΞR}.\mathcal{P}_{R^{-1}}=\left\{\tau(\theta):\theta\in\Xi_{R}\right\}. For τ=τ(θ)𝒫R1\tau=\tau(\theta)\in\mathcal{P}_{R^{-1}}, denote θτ:=θ\theta_{\tau}:=\theta. Let

(149) Tτ,0={x1(γ×γ)(θτ)+x2γ(θτ)+x3γ(θτ):|x1|R,|x2|R1/2,|x3|1}.T^{\circ}_{\tau,0}=\bigg{\{}x_{1}\left(\gamma\times\gamma^{\prime}\right)(\theta_{\tau})+x_{2}\gamma^{\prime}(\theta_{\tau})+x_{3}\gamma(\theta_{\tau}):|x_{1}|\leq R,\,|x_{2}|\leq R^{1/2},\,|x_{3}|\leq 1\bigg{\}}.

Moreover, we will use 𝕋τ\mathbb{T}^{\circ}_{\tau} to denote the collection of translates of Tτ,0T^{\circ}_{\tau,0} that cover B(0,R)B(0,R). For a fixed small constant δ>0\delta>0, denote Tτ,0:=RδTτ,0T_{\tau,0}:=R^{\delta}T^{\circ}_{\tau,0}, and 𝕋τ:={RδT:T𝕋τ}\mathbb{T}_{\tau}:=\{R^{\delta}T:T\in\mathbb{T}^{\circ}_{\tau}\}. For T𝕋τT\in\mathbb{T}_{\tau}, set τ(T)=τ\tau(T)=\tau.

Definition 5.

Fix T𝕋τT\in\mathbb{T}_{\tau}. We say that a function fT:3f_{T}:\mathbb{R}^{3}\to\mathbb{C} is a TT-function if fT^\widehat{f_{T}} is supported on τ(T)\tau(T) and

(150) fTL(B(0,R)T)δR10000fT2.\left\lVert f_{T}\right\rVert_{L^{\infty}\left(B(0,R)\setminus T\right)}\lesssim_{\delta}R^{-10000}\left\lVert f_{T}\right\rVert_{2}.
Theorem 9.

Let γ:[a,b]𝕊2\gamma:[a,b]\to\mathbb{S}^{2} be a C2C^{2} curve with det(γ,γ,γ′′)\det(\gamma,\gamma^{\prime},\gamma^{\prime\prime}) nonvanishing. Let B1B\geq 1 be such that

(151) |det(γ,γ,γ′′)|B1,\left\lvert\det\left(\gamma,\gamma^{\prime},\gamma^{\prime\prime}\right)\right\rvert\geq B^{-1},

and

(152) γC2[a,b]B.\left\lVert\gamma\right\rVert_{C^{2}[a,b]}\leq B.

Let R1R\geq 1 and suppose that

f=T𝕎fT,f=\sum_{T\in\mathbb{W}}f_{T},

where each fTf_{T} is a TT-function and

𝕎τ𝒫R1𝕋τ.\mathbb{W}\subseteq\bigcup_{\tau\in\mathcal{P}_{R^{-1}}}\mathbb{T}_{\tau}.

Assume that for all T,T𝕎T,T^{\prime}\in\mathbb{W},

(153) fT2fT2.\big{\|}f_{T}\big{\|}_{2}\sim\big{\|}f_{T^{\prime}}\big{\|}_{2}.

Let YY be a disjoint union of unit balls in B(0,R)B(0,R), each of which intersects at most MM sets 2T2T with T𝕎T\in\mathbb{W}. Then for 2p62\leq p\leq 6,

fLp(Y)ϵ,δRϵ(M|𝕎|)121p(T𝕎fTp2)1/2.\left\lVert f\right\rVert_{L^{p}(Y)}\lesssim_{\epsilon,\delta}R^{\epsilon}\left(\frac{M}{\left\lvert\mathbb{W}\right\rvert}\right)^{\frac{1}{2}-\frac{1}{p}}\left(\sum_{T\in\mathbb{W}}\left\lVert f_{T}\right\rVert_{p}^{2}\right)^{1/2}.
Proof.

Assume that [a,b]=[1,1][a,b]=[-1,1]. Fix ϵ(0,1/2)\epsilon\in(0,1/2), δ0=ϵ100\delta_{0}=\epsilon^{100}, δ(0,δ0)\delta\in(0,\delta_{0}),

Rmin{B103/ϵ,2105/ϵ},R\geq\min\left\{B^{10^{3}/\epsilon},2^{10^{5}/\epsilon}\right\},

and assume inductively that a (superficially) stronger version of the theorem holds with K2K^{2}-cubes instead of unit cubes, where K=Rδ2K=R^{\delta^{2}}, for all scales smaller than R~:=R/K2\widetilde{R}:=R/K^{2}, for all curves γ\gamma satisfying (151) and (152), and for all B1B\geq 1.

For each τ𝒫R1/2(Γ(γ))\tau\in\mathcal{P}_{R^{-1/2}}(\Gamma(\gamma)), let κ=κ(τ)𝒫K1(Γ(γ))\kappa=\kappa(\tau)\in\mathcal{P}_{K^{-1}}(\Gamma(\gamma)) be the element of 𝒫K1(Γ(γ))\mathcal{P}_{K^{-1}}(\Gamma(\gamma)) which minimises |θτθκ||\theta_{\tau}-\theta_{\kappa}|. For each κ\kappa, let

κ,0={x1(γ×γ)(θκ)|(γ×γ)(θκ)|+x2γ(θκ)|γ(θκ)|+x3γ(θκ):|x1|R1+δ,|x2|R1+δ/K,|x3|R1+δ/K2},\Box_{\kappa,0}=\Big{\{}x_{1}\frac{\left(\gamma\times\gamma^{\prime}\right)(\theta_{\kappa})}{\left\lvert\left(\gamma\times\gamma^{\prime}\right)(\theta_{\kappa})\right\rvert}+x_{2}\frac{\gamma^{\prime}(\theta_{\kappa})}{\left\lvert\gamma^{\prime}(\theta_{\kappa})\right\rvert}+x_{3}\gamma(\theta_{\kappa}):\\ |x_{1}|\leq R^{1+\delta},\,|x_{2}|\leq R^{1+\delta}/K,\,|x_{3}|\leq R^{1+\delta}/K^{2}\Big{\}},

and

κ={=aγ(θκ)+bγ(θκ)|γ(θκ)|+κ,0:a((1/10)R1+δK2),b((1/10)R1+δK1)}.\mathbb{P}_{\kappa}=\Big{\{}\Box=a\gamma(\theta_{\kappa})+b\frac{\gamma^{\prime}(\theta_{\kappa})}{\left\lvert\gamma^{\prime}(\theta_{\kappa})\right\rvert}+\Box_{\kappa,0}:\\ a\in\left((1/10)R^{1+\delta}K^{-2}\right)\mathbb{Z},\quad b\in\left((1/10)R^{1+\delta}K^{-1}\right)\mathbb{Z}\Big{\}}.

Let =κ𝒫K1(Γ(γ))κ\mathbb{P}=\bigcup_{\kappa\in\mathcal{P}_{K^{-1}}(\Gamma(\gamma))}\mathbb{P}_{\kappa}. Given any τ\tau and corresponding κ=κ(τ)\kappa=\kappa(\tau),

(158) |(γ×γ)(θτ),γ(θκ)|B7K1,\left\lvert\left\langle(\gamma\times\gamma^{\prime})(\theta_{\tau}),\gamma^{\prime}(\theta_{\kappa})\right\rangle\right\rvert\leq B^{-7}K^{-1},

and

(159) |(γ×γ)(θτ),γ(θκ)|B7K2.\left\lvert\left\langle(\gamma\times\gamma^{\prime})(\theta_{\tau}),\gamma(\theta_{\kappa})\right\rangle\right\rvert\leq B^{-7}K^{-2}.

It follows that for each T𝕋τT\in\mathbb{T}_{\tau}, there are 1\sim 1 sets κ(τ)\Box\in\mathbb{P}_{\kappa(\tau)} with T10T\cap 10\Box\neq\emptyset, and moreover T100T\subseteq 100\Box whenever T10T\cap 10\Box\neq\emptyset. For each such TT let =(T)κ\Box=\Box(T)\in\mathbb{P}_{\kappa} be some choice such that T10T\cap 10\Box\neq\emptyset, and let 𝕎\mathbb{W}_{\Box} be the set of TT’s associated to \Box.

For each κ\kappa and κ\Box\in\mathbb{P}_{\kappa}, let {Q}Q\left\{Q_{\Box}\right\}_{Q_{\Box}} be a finitely overlapping cover of 100100\Box by translates of the ellipsoid

{x1γ(θκ)+x2γ(θκ)|γ(θκ)|+x3(γ×γ)(θκ)|(γ×γ)(θκ)|:(|x1|2+(|x2|K1)2+(|x3|K2)2)1/2K~2}.\Bigg{\{}x_{1}\gamma(\theta_{\kappa})+x_{2}\frac{\gamma^{\prime}(\theta_{\kappa})}{\left\lvert\gamma^{\prime}(\theta_{\kappa})\right\rvert}+x_{3}\frac{\left(\gamma\times\gamma^{\prime}\right)(\theta_{\kappa})}{\left\lvert\left(\gamma\times\gamma^{\prime}\right)(\theta_{\kappa})\right\rvert}:\\ \left(|x_{1}|^{2}+\left(|x_{2}|K^{-1}\right)^{2}+\left(|x_{3}|K^{-2}\right)^{2}\right)^{1/2}\leq\widetilde{K}^{2}\Bigg{\}}.

Using Poisson summation, let {ηQ}Q𝒬\{\eta_{Q_{\Box}}\}_{Q_{\Box}\in\mathcal{Q}_{\Box}} be a smooth partition of unity such that on 10310^{3}\Box,

Q𝒬ηQ=1,\sum_{Q_{\Box}\in\mathcal{Q}_{\Box}}\eta_{Q_{\Box}}=1,

and such that each ηQ\eta_{Q_{\Box}} satisfies

ηQ1,ηQL(3Q)R10000,\|\eta_{Q_{\Box}}\|_{\infty}\lesssim 1,\quad\|\eta_{Q_{\Box}}\|_{L^{\infty}(\mathbb{R}^{3}\setminus Q_{\Box})}\lesssim R^{-10000},

and

|ηQ(x)|dist(x,Q)10000x3,|\eta_{Q_{\Box}}(x)|\lesssim\textup{dist}(x,Q_{\Box})^{-10000}\quad\forall\,x\in\mathbb{R}^{3},

with ηQ^\widehat{\eta_{Q_{\Box}}} supported in

{ξ1γ(θκ)+ξ2γ(θκ)|γ(θκ)|+ξ3(γ×γ)(θκ)|(γ×γ)(θκ)|:|ξ1|K~,|ξ2|K~K,|ξ3|K~K2}.\bigg{\{}\xi_{1}\gamma(\theta_{\kappa})+\xi_{2}\frac{\gamma^{\prime}(\theta_{\kappa})}{\left\lvert\gamma^{\prime}(\theta_{\kappa})\right\rvert}+\xi_{3}\frac{\left(\gamma\times\gamma^{\prime}\right)(\theta_{\kappa})}{\left\lvert\left(\gamma\times\gamma^{\prime}\right)(\theta_{\kappa})\right\rvert}\\ :|\xi_{1}|\leq\widetilde{K},\quad|\xi_{2}|\leq\widetilde{K}K,\quad|\xi_{3}|\leq\widetilde{K}K^{2}\bigg{\}}.

By dyadic pigeonholing,

fLp(Y)logRT𝕎ηYfTLp(Y)+R1000(T𝕎fTp2)1/2,\left\lVert f\right\rVert_{L^{p}(Y)}\lesssim\log R\left\lVert\sum_{\Box}\sum_{T\in\mathbb{W}_{\Box}}\eta_{Y_{\Box}}f_{T}\right\rVert_{L^{p}(Y)}+R^{-1000}\left(\sum_{T\in\mathbb{W}}\left\lVert f_{T}\right\rVert_{p}^{2}\right)^{1/2},

where, for each \Box, YY_{\Box} is a union over a subset of the sets QQ_{\Box}, and ηY\eta_{Y_{\Box}} is the corresponding sum over ηQ\eta_{Q_{\Box}}, such that each QYQ_{\Box}\subseteq Y_{\Box} intersects a number #[M(),2M())\#\in[M^{\prime}(\Box),2M^{\prime}(\Box)) different sets (1.5)T(1.5)T with T𝕎T\in\mathbb{W}_{\Box}, up to a factor of 2. By pigeonholing again,

T𝕎ηYfTLp(Y)(logR)2𝔹T𝕎ηYfTLp(Y),\left\lVert\sum_{\Box}\sum_{T\in\mathbb{W}_{\Box}}\eta_{Y_{\Box}}f_{T}\right\rVert_{L^{p}(Y)}\lesssim(\log R)^{2}\left\lVert\sum_{\Box\in\mathbb{B}}\sum_{T\in\mathbb{W}_{\Box}}\eta_{Y_{\Box}}f_{T}\right\rVert_{L^{p}(Y)},

where |𝕎|\left\lvert\mathbb{W}_{\Box}\right\rvert and M=M()M^{\prime}=M^{\prime}(\Box) are constant over 𝔹\Box\in\mathbb{B}, up to a factor of 2. By one final pigeonholing step,

𝔹T𝕎ηYfTLp(Y)logR𝔹T𝕎ηYfTLp(Y),\left\lVert\sum_{\Box\in\mathbb{B}}\sum_{T\in\mathbb{W}_{\Box}}\eta_{Y_{\Box}}f_{T}\right\rVert_{L^{p}(Y)}\lesssim\log R\left\lVert\sum_{\Box\in\mathbb{B}}\sum_{T\in\mathbb{W}_{\Box}}\eta_{Y_{\Box}}f_{T}\right\rVert_{L^{p}(Y^{\prime})},

where YY^{\prime} is a union over K2K^{2}-balls QYQ\subseteq Y such that each ball 2Q2Q intersects a number #[M′′,2M′′)\#\in[M^{\prime\prime},2M^{\prime\prime}) of the sets YY_{\Box} in a set of strictly positive Lebesgue measure, as \Box varies over 𝔹\mathbb{B}. Fix QYQ\subseteq Y^{\prime}. By the decoupling theorem for generalised C2C^{2} cones, followed by Hölder’s inequality,

𝔹T𝕎ηYfTLp(Q)CϵB100Kϵ/100(M′′)121p(𝔹T𝕎ηYfTLp(2Q)p)1/p+R900(T𝕎fTp2)1/2.\left\lVert\sum_{\Box\in\mathbb{B}}\sum_{T\in\mathbb{W}_{\Box}}\eta_{Y_{\Box}}f_{T}\right\rVert_{L^{p}(Q)}\\ \leq C_{\epsilon}B^{100}K^{\epsilon/100}\left(M^{\prime\prime}\right)^{\frac{1}{2}-\frac{1}{p}}\left(\sum_{\Box\in\mathbb{B}}\left\lVert\sum_{T\in\mathbb{W}_{\Box}}\eta_{Y_{\Box}}f_{T}\right\rVert_{L^{p}(2Q)}^{p}\right)^{1/p}\\ +R^{-900}\left(\sum_{T\in\mathbb{W}}\left\lVert f_{T}\right\rVert_{p}^{2}\right)^{1/2}.

Summing over QQ gives

fLp(Y)Cϵ(logR)100B100Kϵ/100(M′′)121p×(𝔹T𝕎fTLp(Y)p)1/p+R800(T𝕎fTp2)1/2.\|f\|_{L^{p}(Y)}\lesssim C_{\epsilon}\left(\log R\right)^{100}B^{100}K^{\epsilon/100}\left(M^{\prime\prime}\right)^{\frac{1}{2}-\frac{1}{p}}\\ \times\left(\sum_{\Box\in\mathbb{B}}\left\lVert\sum_{T\in\mathbb{W}_{\Box}}f_{T}\right\rVert_{L^{p}(Y_{\Box})}^{p}\right)^{1/p}+R^{-800}\left(\sum_{T\in\mathbb{W}}\left\lVert f_{T}\right\rVert_{p}^{2}\right)^{1/2}.

This will be bounded using the inductive assumption, following a Lorentz rescaling.

For each θ[1,1]\theta\in[-1,1], define the Lorentz rescaling map L=LθL=L_{\theta} at θ\theta by

L[x1γ(θ)+x2γ(θ)|γ(θ)|+x3(γ×γ)(θ)|(γ×γ)(θ)|]=x1γ(θ)+Kx2γ(θ)|γ(θ)|+K2x3(γ×γ)(θ)|(γ×γ)(θ)|.L\left[x_{1}\gamma(\theta)+x_{2}\frac{\gamma^{\prime}(\theta)}{\left\lvert\gamma^{\prime}(\theta)\right\rvert}+x_{3}\frac{\left(\gamma\times\gamma^{\prime}\right)(\theta)}{\left\lvert\left(\gamma\times\gamma^{\prime}\right)(\theta)\right\rvert}\right]\\ =x_{1}\gamma(\theta)+Kx_{2}\frac{\gamma^{\prime}(\theta)}{\left\lvert\gamma^{\prime}(\theta)\right\rvert}+K^{2}x_{3}\frac{\left(\gamma\times\gamma^{\prime}\right)(\theta)}{\left\lvert\left(\gamma\times\gamma^{\prime}\right)(\theta)\right\rvert}.

Let

γ~(ϕ)=L(γ(ϕ))|L(γ(ϕ))|,ϕ[1,1].\widetilde{\gamma}(\phi)=\frac{L(\gamma(\phi))}{\left\lvert L(\gamma(\phi))\right\rvert},\qquad\phi\in[-1,1].

Then for any ϕ[1,1]\phi\in[-1,1],

γ~(ϕ)=πγ~(ϕ)(L(γ(ϕ)))|L(γ(ϕ))|,\widetilde{\gamma}^{\prime}(\phi)=\frac{\pi_{\widetilde{\gamma}(\phi)^{\perp}}\left(L(\gamma^{\prime}(\phi))\right)}{\left\lvert L(\gamma(\phi))\right\rvert},

and

γ~′′(ϕ)=πγ~(ϕ)(L(γ′′(ϕ)))|L(γ(ϕ))|L(γ(ϕ)),L(γ(ϕ))πγ~(ϕ)(L(γ(ϕ)))|L(γ(ϕ))|3.\widetilde{\gamma}^{\prime\prime}(\phi)=\frac{\pi_{\widetilde{\gamma}(\phi)^{\perp}}\left(L(\gamma^{\prime\prime}(\phi))\right)}{\left\lvert L(\gamma(\phi))\right\rvert}-\frac{\left\langle L(\gamma(\phi)),L(\gamma^{\prime}(\phi))\right\rangle\pi_{\widetilde{\gamma}(\phi)^{\perp}}\left(L(\gamma^{\prime}(\phi))\right)}{\left\lvert L(\gamma(\phi))\right\rvert^{3}}.

Hence

det(γ~,γ~,γ~′′)\displaystyle\det(\widetilde{\gamma},\widetilde{\gamma}^{\prime},\widetilde{\gamma}^{\prime\prime}) =1|Lγ|3det(Lγ,Lγ,Lγ′′)\displaystyle=\frac{1}{\left\lvert L\circ\gamma\right\rvert^{3}}\det\left(L\circ\gamma,L\circ\gamma^{\prime},L\circ\gamma^{\prime\prime}\right)
=K3|Lγ|3det(γ,γ,γ′′).\displaystyle=\frac{K^{3}}{\left\lvert L\circ\gamma\right\rvert^{3}}\det\left(\gamma,\gamma^{\prime},\gamma^{\prime\prime}\right).

Let ε=(105B10)1\varepsilon=(10^{5}B^{10})^{-1}, and for fixed θ[1+ε,1ε]\theta\in[-1+\varepsilon,1-\varepsilon], let

(171) γ^(ϕ)=γ~(θ+K1ϕ),ϕ[ε,ε].\widehat{\gamma}(\phi)=\widetilde{\gamma}(\theta+K^{-1}\phi),\quad\phi\in[-\varepsilon,\varepsilon].

The assumption that γC2[1,1]B\left\lVert\gamma\right\rVert_{C^{2}[-1,1]}\leq B yields

1|L(γ(ϕ))|1+10Bε,ϕ[θεK1,θ+εK1].1\leq|L(\gamma(\phi))|\leq 1+10B\varepsilon,\quad\forall\,\phi\in[\theta-\varepsilon K^{-1},\theta+\varepsilon K^{-1}].

Similarly,

|L(γ(ϕ))L(γ(θ))|10εBK,ϕ[θεK1,θ+εK1].\left\lvert L(\gamma^{\prime}(\phi))-L(\gamma^{\prime}(\theta))\right\rvert\leq 10\varepsilon BK,\quad\forall\,\phi\in[\theta-\varepsilon K^{-1},\theta+\varepsilon K^{-1}].

It follows that

|det(γ^,γ^,γ^′′)|(2B)1\left\lvert\det\left(\widehat{\gamma},\widehat{\gamma}^{\prime},\widehat{\gamma}^{\prime\prime}\right)\right\rvert\geq(2B)^{-1}

on [ε,ε][-\varepsilon,\varepsilon], and that

γ^C2[ε,ε]2B.\left\lVert\widehat{\gamma}\right\rVert_{C^{2}[-\varepsilon,\varepsilon]}\leq 2B.

For each 𝔹\Box\in\mathbb{B}, given T𝕎T\in\mathbb{W}_{\Box}, let gT=fTLg_{T}=f_{T}\circ L, where L=Lθκ()L=L_{\theta_{\kappa(\Box)}}. Then

(172) T𝕎fTLp(Y)K3pT𝕎gTLp(L1Y).\left\lVert\sum_{T\in\mathbb{W}_{\Box}}f_{T}\right\rVert_{L^{p}(Y_{\Box})}\leq K^{\frac{3}{p}}\left\lVert\sum_{T\in\mathbb{W}_{\Box}}g_{T}\right\rVert_{L^{p}(L^{-1}Y_{\Box})}.

The inequalities (158) and (159) imply that for each T𝕎T\in\mathbb{W}_{\Box}, the set L1(T)L^{-1}(T) is a equivalent (up to a factor 1.01) to a plank of length R~1+δ\widetilde{R}^{1+\delta} in its longest direction parallel to L1(γ×γ)(θτ(T))L^{-1}(\gamma\times\gamma^{\prime})(\theta_{\tau(T)}), of length R~1/2+δ\widetilde{R}^{1/2+\delta} in its medium direction, and of length R~δ\widetilde{R}^{\delta} in its shortest direction. The ellipsoids QQ_{\Box} are rescaled to K~2\widetilde{K}^{2}-balls L1(Q)L^{-1}(Q_{\Box}). Moreover, it will be shown that

(175) L(τ){x1γ~(θτ)+x2γ~(θτ)|γ~(θτ)|+x3(γ~×γ~)(θτ)|(γ~×γ~)(θτ)|:1x12.01,|x2|(1.01)R~1/2,|x3|R~1}.L(\tau)\subseteq\bigg{\{}x_{1}\widetilde{\gamma}(\theta_{\tau})+x_{2}\frac{\widetilde{\gamma}^{\prime}(\theta_{\tau})}{\left\lvert\widetilde{\gamma}^{\prime}(\theta_{\tau})\right\rvert}+x_{3}\frac{\left(\widetilde{\gamma}\times\widetilde{\gamma}^{\prime}\right)(\theta_{\tau})}{\left\lvert\left(\widetilde{\gamma}\times\widetilde{\gamma}^{\prime}\right)(\theta_{\tau})\right\rvert}\\ :1\leq x_{1}\leq 2.01,\,|x_{2}|\leq(1.01)\widetilde{R}^{-1/2},\,|x_{3}|\leq\widetilde{R}^{-1}\bigg{\}}.

To prove this, let

x=x1γ(θτ)+x2γ(θτ)|γ(θτ)|+x3(γ×γ)(θτ)|(γ×γ)(θτ)|τ,x=x_{1}\gamma(\theta_{\tau})+x_{2}\frac{\gamma^{\prime}(\theta_{\tau})}{\left\lvert\gamma^{\prime}(\theta_{\tau})\right\rvert}+x_{3}\frac{(\gamma\times\gamma^{\prime})(\theta_{\tau})}{\left\lvert(\gamma\times\gamma^{\prime})(\theta_{\tau})\right\rvert}\in\tau,

where

x1[1,2],|x2|R1/2,|x3|R1.x_{1}\in[1,2],\quad|x_{2}|\leq R^{-1/2},\quad|x_{3}|\leq R^{-1}.

The vector (γ~×γ~)(θτ)\left(\widetilde{\gamma}\times\widetilde{\gamma}^{\prime}\right)(\theta_{\tau}) is parallel to L1((γ×γ)(θτ))L^{-1}((\gamma\times\gamma^{\prime})(\theta_{\tau})), since L1((γ×γ)(θτ))L^{-1}((\gamma\times\gamma^{\prime})(\theta_{\tau})) is orthogonal to γ~(θτ)\widetilde{\gamma}(\theta_{\tau}) and γ~(θτ)\widetilde{\gamma}^{\prime}(\theta_{\tau}). The inequality

|L1((γ×γ)(θτ))|K2|(γ×γ)(θτ)|\left\lvert L^{-1}((\gamma\times\gamma^{\prime})(\theta_{\tau}))\right\rvert\geq K^{-2}\left\lvert(\gamma\times\gamma^{\prime})(\theta_{\tau})\right\rvert

gives

(176) |Lx,L1((γ×γ)(θτ))|L1((γ×γ)(θτ))||R~1.\left\lvert\left\langle Lx,\frac{L^{-1}((\gamma\times\gamma^{\prime})(\theta_{\tau}))}{\left\lvert L^{-1}((\gamma\times\gamma^{\prime})(\theta_{\tau}))\right\rvert}\right\rangle\right\rvert\leq\widetilde{R}^{-1}.

Moreover,

|Lx,πL(γ(θτ))(L(γ(θτ)))|πL(γ(θτ))(L(γ(θτ)))||\displaystyle\left\lvert\left\langle Lx,\frac{\pi_{L(\gamma(\theta_{\tau}))^{\perp}}\left(L\left(\gamma^{\prime}(\theta_{\tau})\right)\right)}{\left\lvert\pi_{L(\gamma(\theta_{\tau}))^{\perp}}\left(L\left(\gamma^{\prime}(\theta_{\tau})\right)\right)\right\rvert}\right\rangle\right\rvert
=|x2L(γ(θτ)|γ(θτ)|)+x3L((γ×γ)(θτ)|(γ×γ)(θτ)|),πL(γ(θτ))(L(γ(θτ)))|πL(γ(θτ))(L(γ(θτ)))||\displaystyle\quad=\left\lvert\left\langle x_{2}L\left(\frac{\gamma^{\prime}(\theta_{\tau})}{\left\lvert\gamma^{\prime}(\theta_{\tau})\right\rvert}\right)+x_{3}L\left(\frac{(\gamma\times\gamma^{\prime})(\theta_{\tau})}{\left\lvert(\gamma\times\gamma^{\prime})(\theta_{\tau})\right\rvert}\right),\frac{\pi_{L(\gamma(\theta_{\tau}))^{\perp}}\left(L\left(\gamma^{\prime}(\theta_{\tau})\right)\right)}{\left\lvert\pi_{L(\gamma(\theta_{\tau}))^{\perp}}\left(L\left(\gamma^{\prime}(\theta_{\tau})\right)\right)\right\rvert}\right\rangle\right\rvert
(177) (1.01)R~1/2.\displaystyle\quad\leq(1.01)\widetilde{R}^{-1/2}.

For the direction L(γ(θτ))L(\gamma(\theta_{\tau})),

(178) Lx,L(γ(θτ))|L(γ(θτ))|=x1|L(γ(θτ))|+O(K2R1/2).\left\langle Lx,\frac{L(\gamma(\theta_{\tau}))}{|L(\gamma(\theta_{\tau}))|}\right\rangle=x_{1}|L(\gamma(\theta_{\tau}))|+O(K^{2}R^{-1/2}).

Combining (176), (177) and (178) gives (175).

Inductively applying the theorem at scale R~\widetilde{R} gives

(172)\displaystyle\eqref{inductthis} Cϵ,δB1010/ϵRϵK2ϵ(M|𝕎|)121p(T𝕎fTp2)1/2\displaystyle\lesssim C_{\epsilon,\delta}B^{10^{10}/\epsilon}R^{\epsilon}K^{-2\epsilon}\left(\frac{M^{\prime}}{\left\lvert\mathbb{W}_{\Box}\right\rvert}\right)^{\frac{1}{2}-\frac{1}{p}}\left(\sum_{T\in\mathbb{W}_{\Box}}\left\lVert f_{T}\right\rVert_{p}^{2}\right)^{1/2}

for each 𝔹\Box\in\mathbb{B}. Hence

fLp(Y)Cϵ,δB1010/ϵKϵ(MM′′|𝕎|)121p(𝔹(T𝕎fTp2)p/2)1/p.\|f\|_{L^{p}(Y)}\\ \leq C_{\epsilon,\delta}B^{10^{10}/\epsilon}K^{-\epsilon}\left(\frac{M^{\prime}M^{\prime\prime}}{\left\lvert\mathbb{W}_{\Box}\right\rvert}\right)^{\frac{1}{2}-\frac{1}{p}}\left(\sum_{\Box\in\mathbb{B}}\left(\sum_{T\in\mathbb{W}_{\Box}}\left\lVert f_{T}\right\rVert_{p}^{2}\right)^{p/2}\right)^{1/p}.

By the dyadically constant property of fTp\|f_{T}\|_{p}, this is

Cϵ,δB1010/ϵKϵ(MM′′|𝕎|)121p(|𝔹||𝕎||𝕎|)1p(T𝕎fT22)1/2.\lesssim C_{\epsilon,\delta}B^{10^{10}/\epsilon}K^{-\epsilon}\left(\frac{M^{\prime}M^{\prime\prime}}{\left\lvert\mathbb{W}\right\rvert}\right)^{\frac{1}{2}-\frac{1}{p}}\left(\frac{\left\lvert\mathbb{B}\right\rvert\left\lvert\mathbb{W}_{\Box}\right\rvert}{\left\lvert\mathbb{W}\right\rvert}\right)^{\frac{1}{p}}\left(\sum_{T\in\mathbb{W}}\left\lVert f_{T}\right\rVert_{2}^{2}\right)^{1/2}.

The second bracketed term is 1\lesssim 1, since

|𝕎|=T𝕎1𝔹T𝕎:=(T)1𝔹T𝕎:=(T)1|𝔹||𝕎|.\left\lvert\mathbb{W}\right\rvert=\sum_{T\in\mathbb{W}}1\geq\sum_{\Box\in\mathbb{B}}\sum_{\begin{subarray}{c}T\in\mathbb{W}:\\ \Box=\Box(T)\end{subarray}}1\sim\sum_{\Box\in\mathbb{B}}\sum_{\begin{subarray}{c}T\in\mathbb{W}:\\ \Box=\Box(T)\end{subarray}}1\geq\left\lvert\mathbb{B}\right\rvert\left\lvert\mathbb{W}_{\Box}\right\rvert.

It remains to show that MM′′MM^{\prime}M^{\prime\prime}\lesssim M. Let QYQ\subseteq Y^{\prime} be any R1/2R^{1/2}-ball. By definition of MM,

M\displaystyle M T𝕎:2TQ𝔹:=(T)1\displaystyle\gtrsim\sum_{\begin{subarray}{c}T\in\mathbb{W}:\\ 2T\cap Q\neq\emptyset\end{subarray}}\sum_{\begin{subarray}{c}\Box\in\mathbb{B}:\\ \Box=\Box(T)\end{subarray}}1
=𝔹T𝕎:=(T)2TQ1\displaystyle=\sum_{\Box\in\mathbb{B}}\sum_{\begin{subarray}{c}T\in\mathbb{W}:\\ \Box=\Box(T)\\ 2T\cap Q\neq\emptyset\end{subarray}}1
𝔹T𝕎:2TQ1.\displaystyle\geq\sum_{\Box\in\mathbb{B}}\sum_{\begin{subarray}{c}T\in\mathbb{W}_{\Box}:\\ 2T\cap Q\neq\emptyset\end{subarray}}1.

By definition of MM^{\prime} and M′′M^{\prime\prime},

MM′′\displaystyle M^{\prime}M^{\prime\prime} 𝔹:m(Y2Q)>0M\displaystyle\sim\sum_{\begin{subarray}{c}\Box\in\mathbb{B}:\\ m(Y_{\Box}\cap 2Q)>0\end{subarray}}M^{\prime}
𝔹:m(Y2Q)>0QYMm(Q2Q)m(Y2Q)\displaystyle\leq\sum_{\begin{subarray}{c}\Box\in\mathbb{B}:\\ m(Y_{\Box}\cap 2Q)>0\end{subarray}}\sum_{Q_{\Box}\subseteq Y_{\Box}}M^{\prime}\frac{m(Q_{\Box}\cap 2Q)}{m(Y_{\Box}\cap 2Q)}
𝔹:m(Y2Q)>0QYT𝕎:Q(1.5)Tm(Q2Q)m(Y2Q)\displaystyle\sim\sum_{\begin{subarray}{c}\Box\in\mathbb{B}:\\ m(Y_{\Box}\cap 2Q)>0\end{subarray}}\sum_{Q_{\Box}\subseteq Y_{\Box}}\sum_{\begin{subarray}{c}T\in\mathbb{W}_{\Box}:\\ Q_{\Box}\cap(1.5)T\neq\emptyset\end{subarray}}\frac{m(Q_{\Box}\cap 2Q)}{m(Y_{\Box}\cap 2Q)}
=𝔹:m(Y2Q)>0T𝕎QY:Q(1.5)Tm(Q2Q)m(Y2Q)\displaystyle=\sum_{\begin{subarray}{c}\Box\in\mathbb{B}:\\ m(Y_{\Box}\cap 2Q)>0\end{subarray}}\sum_{T\in\mathbb{W}_{\Box}}\sum_{\begin{subarray}{c}Q_{\Box}\subseteq Y_{\Box}:\\ Q_{\Box}\cap(1.5)T\neq\emptyset\end{subarray}}\frac{m(Q_{\Box}\cap 2Q)}{m(Y_{\Box}\cap 2Q)}
(179) 𝔹:m(Y2Q)>0T𝕎:2TQQYm(Q2Q)m(Y2Q)\displaystyle\leq\sum_{\begin{subarray}{c}\Box\in\mathbb{B}:\\ m(Y_{\Box}\cap 2Q)>0\end{subarray}}\sum_{\begin{subarray}{c}T\in\mathbb{W}_{\Box}:\\ 2T\cap Q\neq\emptyset\end{subarray}}\sum_{Q_{\Box}\subseteq Y_{\Box}}\frac{m(Q_{\Box}\cap 2Q)}{m(Y_{\Box}\cap 2Q)}
𝔹:T𝕎:2TQ1\displaystyle\lesssim\sum_{\Box\in\mathbb{B}:}\sum_{\begin{subarray}{c}T\in\mathbb{W}_{\Box}:\\ 2T\cap Q\neq\emptyset\end{subarray}}1
M.\displaystyle\lesssim M.

The inequality (179) above follows from the observation that if Q2QQ_{\Box}\cap 2Q\neq\emptyset, and if T𝕎T\in\mathbb{W}_{\Box} is such that Q(1.5)TQ_{\Box}\cap(1.5)T\neq\emptyset, then 2TQ2T\cap Q\neq\emptyset. ∎

Appendix B Decoupling for C2C^{2} cones

We have been using the decoupling inequality for C2C^{2} cones in 3\mathbb{R}^{3} in a few places above, but it may have not been written down in the literature. In the appendix, we state it and sketch its proof. We start with the decoupling for C2C^{2} curves on 2\mathbb{R}^{2}.

Theorem 10.

Let γ:[1,1]\gamma:[-1,1]\to\mathbb{R} with γ(0)=γ(0)=0\gamma(0)=\gamma^{\prime}(0)=0 be C2C^{2} and satisfy γ′′(t)0\gamma^{\prime\prime}(t)\neq 0 for every t[1,1]t\in[-1,1]. Then

(180) E[1,1]fL6(2)γ,ϵδϵ(I[0,1],|I|=δEIfL6(2)2)1/2,\Big{\|}E_{[-1,1]}f\Big{\|}_{L^{6}(\mathbb{R}^{2})}\lesssim_{\gamma,\epsilon}\delta^{-\epsilon}\Big{(}\sum_{I\subset[0,1],|I|=\delta}\big{\|}E_{I}f\big{\|}_{L^{6}(\mathbb{R}^{2})}^{2}\Big{)}^{1/2},

for every ϵ>0\epsilon>0 and δ(0,1)\delta\in(0,1). Here

(181) EIf(x,y):=If(t)ei(xt+yγ(t))𝑑t,E_{I}f(x,y):=\int_{I}f(t)e^{i(xt+y\gamma(t))}dt,

for an interval I[1,1]I\subset[-1,1].

One can follow the same argument as in [8] to prove Theorem 10. We leave out the proof. Via the bootstrapping argument as in Bourgain and Demeter [1] and Pramanik and Seeger [24], one can prove the following decoupling for C2C^{2} cones.

Theorem 11.

Let γ:[1,1]\gamma:[-1,1]\to\mathbb{R} with γ(0)=γ(0)=0\gamma(0)=\gamma^{\prime}(0)=0 be C2C^{2} and satisfy γ′′(t)0\gamma^{\prime\prime}(t)\neq 0 for every t[1,1]t\in[-1,1]. Then

(182) [1,1]fL6(3)γ,ϵδϵ(I[0,1],|I|=δIfL6(3)2)1/2,\Big{\|}\mathcal{E}_{[-1,1]}f\Big{\|}_{L^{6}(\mathbb{R}^{3})}\lesssim_{\gamma,\epsilon}\delta^{-\epsilon}\Big{(}\sum_{I\subset[0,1],|I|=\delta}\big{\|}\mathcal{E}_{I}f\big{\|}_{L^{6}(\mathbb{R}^{3})}^{2}\Big{)}^{1/2},

for every ϵ>0\epsilon>0 and δ(0,1)\delta\in(0,1). Here

(183) If(x,y,z):=[1,2]×If(s,t)ei(sx+sty+sγ(t)z)𝑑s𝑑t,\mathcal{E}_{I}f(x,y,z):=\int_{[1,2]\times I}f(s,t)e^{i(sx+sty+s\gamma(t)z)}dsdt,

for an interval I[1,1]I\subset[-1,1].

We give a sketch of the proof of Theorem 11. By the triangle inequality, we can assume that ff is supported on [1,1+δϵ]×[0,δϵ][1,1+\delta^{\epsilon}]\times[0,\delta^{\epsilon}]. The key observation in [1] is that the cone

(184) {s(1,t,γ(t)):1s1+δϵ,0tδμ}\{s(1,t,\gamma(t)):1\leq s\leq 1+\delta^{\epsilon},0\leq t\leq\delta^{\mu}\}

is in the δ2μ+ϵ\delta^{2\mu+\epsilon}-neighborhood of the cylinder

(185) {(s,t,γ(t)):1s1+δϵ,0tδμ},\{(s,t,\gamma(t)):1\leq s\leq 1+\delta^{\epsilon},0\leq t\leq\delta^{\mu}\},

for every μϵ\mu\geq\epsilon. To see this, let us take one point s(1,t,γ(t))s(1,t,\gamma(t)) from (184), and we will show that its distance to (s,st,γ(st))(s,st,\gamma(st)), which lies in (185), is δ2μ+ϵ\lesssim\delta^{2\mu+\epsilon}. This amounts to proving

(186) |γ(st)sγ(t)|δ2μ+ϵ.|\gamma(st)-s\gamma(t)|\lesssim\delta^{2\mu+\epsilon}.

Note that

(187) |γ(st)sγ(t)||1s||γ(t)|+|γ(st)γ(t)|.|\gamma(st)-s\gamma(t)|\lesssim|1-s||\gamma(t)|+|\gamma(st)-\gamma(t)|.

The desired bound (186) follows from Taylor’s expansion and mean value theorems.

After proving (186), one can then apply Theorem 10 iteratively, in the same way as in [1], and finish the proof of Theorem 11. We leave out the iteration step.

References

  • [1] J. Bourgain and C. Demeter. The proof of the 2\ell^{2} decoupling conjecture. Ann. of Math., pages 351–389, 2015.
  • [2] C. Chen. Restricted families of projections and random subspaces. Real Anal. Exchange, 43(2):347–358, 2018.
  • [3] C. Demeter, L. Guth, and H. Wang. Small cap decouplings. Geom. Funct. Anal., 30(4):989–1062, 2020.
  • [4] X. Du and R. Zhang. Sharp L2L^{2} estimates of the Schrödinger maximal function in higher dimensions. Ann. of Math. (2), 189(3):837–861, 2019.
  • [5] K. Fässler and T. Orponen. On restricted families of projections in 3\mathbb{R}^{3}. Proc. Lond. Math. Soc., 109(2):353–381, 2014.
  • [6] Y. Fu, L. Guth, and D. Maldague. Sharp superlevel set estimates for small cap decouplings of the parabola. arXiv preprint arXiv:2107.13139, 2021.
  • [7] S. Gan and S. Wu. Square function estimates for conical regions. arXiv preprint arXiv:2203.12155, 2022.
  • [8] S. Guo, Z. K. Li, P.-L. Yung, and P. Zorin-Kranich. A short proof of 2\ell^{2} decoupling for the moment curve. Amer. J. Math., 143(6):1983–1998, 2021.
  • [9] L. Guth, A. Iosevich, Y. Ou, and H. Wang. On Falconer’s distance set problem in the plane. Invent. Math., 219(3):779–830, 2020.
  • [10] L. Guth, D. Maldague, and H. Wang. Improved decoupling for the parabola. arXiv preprint arXiv:2009.07953, 2020.
  • [11] L. Guth, N. Solomon, and H. Wang. Incidence estimates for well spaced tubes. Geom. Funct. Anal., 29(6):1844–1863, 2019.
  • [12] L. Guth, H. Wang, and R. Zhang. A sharp square function estimate for the cone in 3\mathbb{R}^{3}. Ann. of Math., 192(2):551–581, 2020.
  • [13] T. L. J. Harris. Improved bounds for restricted projection families via weighted fourier restriction. arXiv preprint arXiv:1911.00615, 2019.
  • [14] T. L. J. Harris. Restricted families of projections onto planes: the general case of nonvanishing geodesic curvature. arXiv preprint arXiv:2107.14701, 2021.
  • [15] E. Järvenpää, M. Järvenpää, and T. Keleti. Hausdorff dimension and non-degenerate families of projections. J. Geom. Anal., 24(4):2020–2034, 2014.
  • [16] E. Järvenpää, M. Järvenpää, F. Ledrappier, and M. Leikas. One-dimensional families of projections. Nonlinearity, 21(3):453, 2008.
  • [17] A. Käenmäki, T. Orponen, and L. Venieri. A Marstrand-type restricted projection theorem in 3\mathbb{R}^{3}. arXiv preprint arXiv:1708.04859, 2017.
  • [18] J. M. Marstrand. Some fundamental geometrical properties of plane sets of fractional dimensions. Proc. Lond. Math. Soc., 3(1):257–302, 1954.
  • [19] P. Mattila. Hausdorff dimension, orthogonal projections and intersections with planes. Ann. Acad. Sci. Fenn. Ser. AI Math, 1(2):227–244, 1975.
  • [20] P. Mattila. Geometry of sets and measures in Euclidean spaces, volume 44 of Cambridge Stud. Adv. Math. Cambridge University Press, Cambridge, 1995. Fractals and rectifiability.
  • [21] D. Oberlin and R. Oberlin. Application of a fourier restriction theorem to certain families of projections in 3\mathbb{R}^{3}. J. Geom. Anal., 25(3):1476–1491, 2015.
  • [22] T. Orponen. Hausdorff dimension estimates for restricted families of projections in 3\mathbb{R}^{3}. Adv. Math., 275:147–183, 2015.
  • [23] T. Orponen and L. Venieri. Improved bounds for restricted families of projections to planes in 3\mathbb{R}^{3}. Int. Math. Res. Not. IMRN, 2020(19):5797–5813, 2020.
  • [24] M. Pramanik and A. Seeger. LpL^{p} regularity of averages over curves and bounds for associated maximal operators. Amer. J. Math., 129(1):61–103, 2005.
  • [25] M. Pramanik, T. Yang, and J. Zahl. A Furstenberg-type problem for circles, and a Kaufman-type restricted projection theorem in 3\mathbb{R}^{3}. arXiv preprint arXiv:2207.02259, 2022.