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

\stackMath

An exceptional set estimate for restricted projections to lines in 3\mathbb{R}^{3}

Shengwen Gan Department of Mathematics
Massachusetts Institute of Technology
Cambridge, MA 02142-4307, USA
shengwen@mit.edu
Larry Guth Department of Mathematics
Massachusetts Institute of Technology
Cambridge, MA 02142-4307, USA
lguth@math.mit.edu
 and  Dominique Maldague Department of Mathematics
Massachusetts Institute of Technology
Cambridge, MA 02142-4307, USA
dmal@mit.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 lθ={tγ(θ):t}l_{\theta}=\{t\gamma(\theta):t\in\mathbb{R}\} and ρθ:3lθ\rho_{\theta}:\mathbb{R}^{3}\rightarrow l_{\theta} be the orthogonal projections. We prove an exceptional set estimate. For any Borel set A3A\subset\mathbb{R}^{3} and 0s10\leq s\leq 1, define Es(A):={θ[0,1]:dim(ρθ(A))<s}E_{s}(A):=\{\theta\in[0,1]:\dim(\rho_{\theta}(A))<s\}. We have dim(Es(A))1+sdim(A)2\dim(E_{s}(A))\leq 1+\frac{s-\dim(A)}{2}.

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

1. Introduction

If γ:[0,1]𝕊2\gamma:[0,1]\rightarrow\mathbb{S}^{2} is a curve that satisfies the non-degenerate condition

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

then we call γ\gamma a non-degenerate curve. 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]).

In this paper, we study the the projections in 3\mathbb{R}^{3} whose directions are determined by γ\gamma. For each θ[0,1]\theta\in[0,1], let Vθ3V_{\theta}\subset\mathbb{R}^{3} be the 22-dimensional subspace that is orthogonal to γ(θ)\gamma(\theta) and let lθ3l_{\theta}\subset\mathbb{R}^{3} be the 11-dimensional subspace spanned by γ(θ)\gamma(\theta). We also define πθ:3Vθ\pi_{\theta}:\mathbb{R}^{3}\rightarrow V_{\theta} to be the orthogonal projection onto VθV_{\theta}, and define ρθ:3lθ\rho_{\theta}:\mathbb{R}^{3}\rightarrow l_{\theta} to be the orthogonal projection onto lθl_{\theta}. We use dimX\dim X to denote the Hausdorff dimension of set XX. Let us state our main results.

Theorem 1.

Suppose A3A\subset\mathbb{R}^{3} is a Borel set of Hausdorff dimension α\alpha. For 0s10\leq s\leq 1, define the exceptional set

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

Then we have

(2) dim(Es)max{0,1+sα2}.\dim(E_{s})\leq\max\{0,1+\frac{s-\alpha}{2}\}.

As a corollary, we have

Corollary 1.

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

dim(ρθ(A))=min{1,α}, for a.e. θ[0,1].\dim(\rho_{\theta}(A))=\min\{1,\alpha\},\textup{~{}for~{}a.e.~{}}\theta\in[0,1].
Remark 1.

The proof of Theorem 1 relies on the small cap decoupling for the general cone. We also remark that, for the set of directions determined by the model curve γ\gamma_{\circ}, Käenmäki, Orponen and Venieri can prove the exceptional set estimate with upper bound dim(Es)α+s2α\dim(E_{s})\leq\frac{\alpha+s}{2\alpha} when α1\alpha\leq 1 (see [6] Theorem 1.3). The novelty of our paper is that we prove a Falconer-type exceptional set estimate for general non-degenerate curve, hence Corollary 1.

Remark 2.

Pramanik, Yang and Zahl [10] have also recently proved Corollary 1 with an exceptional set estimate of the form dim(Es)s\dim(E_{s})\leq s, compared to (2). Their proof is based on some incidence estimates for curves in the spirit of Wolff’s circular maximal function estimate. The estimates in [10] hold for curves that are only C2C^{2}, which requires a very different proof from earlier work of Wolff and others on these problems.

Remark 3.

It is also an interesting question to ask for the estimate of the set

E={θ[0,1]:1(ρθ(A))=0}E=\{\theta\in[0,1]:\mathcal{H}^{1}(\rho_{\theta}(A))=0\}

which consists of directions to which the projection of AA has zero measure. We notice that recently Harris [5] proved that

(3) dim({θ[0,1]:1(ρθ(A))=0})4dimA3.\dim(\{\theta\in[0,1]:\mathcal{H}^{1}(\rho_{\theta}(A))=0\})\leq\frac{4-\dim A}{3}.

Intuitively, one may think of EE as E1E_{1} (E1E_{1} is defined in (1)). The main result of this paper (2) yields dim(E1)3dimA2\dim(E_{1})\leq\frac{3-\dim A}{2} which is better than the bound 4dimA3\frac{4-\dim A}{3}. This shows that (3) cannot imply (2).

Now we briefly discuss the history of projection theory. Projection theory dates back to Marstrand[8], 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,\dim A\}. This was generalized to higher dimensions by Mattila[9], 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,\dim A\}. More recently, Fässler and Orponen [2] started to consider the projection problems when the direction set is restricted to some submanifold of Grassmannian. Such problems are known as the restricted projection problem. Fässler and Orponen made conjectures about restricted projections to lines and planes in 3\mathbb{R}^{3} (see Conjecture 1.6 in [2]). In this paper, we give an answer to the conjecture about the projections to lines.

2. Projection to one dimensional family of lines

In this section, we prove Theorem 1. Theorem 1 will be a result of an incidence estimate that we are going to state later. Recall that γ:[0,1]𝕊2\gamma:[0,1]\rightarrow\mathbb{S}^{2} a non-degenerate curve.

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.

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}\}.

We recall the following result (see [2] 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,s)(\delta,s)-set PBP\subset B with cardinality #Pκδs\#P\gtrsim\kappa\delta^{-s}.

Next, we state a useful lemma. It is proved in [3], but we still provide the full details here. The lemma roughly says that given a set XX of Hausdorff dimension less than ss, then we can find a covering of XX by squares of dyadic lengths which satisfy a certain ss-dimensional condition. Let us use 𝒟2k\mathcal{D}_{2^{-k}} to denote the lattice squares of length 2k2^{-k} in [0,1]2[0,1]^{2}.

Lemma 2.

Suppose X[0,1]2X\subset[0,1]^{2} with dimX<s\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 Lemma 2.

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 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

(4) #{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 {D𝒞𝒟2k:DD}\{D^{\prime}\in\mathcal{C}\cap\mathcal{D}_{2^{-k}}: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 (4), 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\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 a maximal element of the chain. First, we show that 𝒞\mathcal{C} covers XX. For xXx\in X, let D(j)D^{(j)} 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, D(j)D^{(j)} is independent of jj, and thus D(j)𝒞D^{(j)}\in\mathcal{C}. This shows that 𝒞\mathcal{C} is a covering of XX. 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 a maximal element. ∎

Remark 4.

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.

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

Theorem 2.

Fix 0<s<10<s<1. 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×1×1\delta\times 1\times 1-slabs 𝕊θ\mathbb{S}_{\theta} with normal direction γ(θ)\gamma(\theta). 𝕊θ\mathbb{S}_{\theta} satisfies the ss-dimensional condition:

  1. (1)

    #𝕊θδs\#\mathbb{S}_{\theta}\lesssim\delta^{-s},

  2. (2)

    #{S𝕊θ:SBr}(rδ)s\#\{S\in\mathbb{S}_{\theta}:S\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 slabs from θ𝕊θ\cup_{\theta}\mathbb{S}_{\theta}. Then

(#Θ)4#HCs,εδ2ts2ε.(\#\Theta)^{4}\#H\leq C_{s,\varepsilon}\delta^{-2t-s-2-\varepsilon}.

2.1. δ\delta-discretization of the projection problem

We show Theorem 2 implies Theorem 1 in this subsection.

Proof of Theorem 1 assuming Theorem 2.

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]:dimρθ(A)<s}.E_{s}=\{\theta\in[0,1]:\dim\rho_{\theta}(A)<s\}.

If dim(Es)=0\dim(E_{s})=0, then there is nothing to prove. Therefore, we assume dim(Es)>0\dim(E_{s})>0. 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 dimension is that

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

We choose a<dim(A),t<dim(Es)a<\dim(A),t<\dim(E_{s}). Then t(Es)>0\mathcal{H}^{t}_{\infty}(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 only need to prove

a2+s2t,a\leq 2+s-2t,

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

Fix a θEs\theta\in E_{s}. By definition we have dimρθ(A)<s\dim\rho_{\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 ρθ(A)\rho_{\theta}(A) by intervals 𝕀θ={I}\mathbb{I}_{\theta}=\{I\}, each of which has length 2j2^{-j} for some integer j>|log2ϵ|j>|\log_{2}\epsilon_{\circ}|. We define 𝕀θ,j:={I𝕀θ:r(I)=2j}\mathbb{I}_{\theta,j}:=\{I\in\mathbb{I}_{\theta}:r(I)=2^{-j}\} (Here r(I)r(I) denotes the length of II). Lemma 2 yields the following properties:

(5) I𝕀θr(I)s1;\sum_{I\in\mathbb{I}_{\theta}}r(I)^{s}\leq 1;

For each jj and rr-interval IrlθI_{r}\subset l_{\theta}, we have

(6) #{I𝕀θ,j:IIr}(r2j)s.\#\{I\in\mathbb{I}_{\theta,j}:I\subset I_{r}\}\lesssim(\frac{r}{2^{-j}})^{s}.

For each θEs\theta\in E_{s}, we can find such a 𝕀θ\mathbb{I}_{\theta}. We also define the slab sets 𝕊θ,j:={ρθ1(I):I𝕀θ,j}B3(0,1)\mathbb{S}_{\theta,j}:=\{\rho^{-1}_{\theta}(I):I\in\mathbb{I}_{\theta,j}\}\cap B^{3}(0,1), 𝕊θ:=j𝕊θ,j\mathbb{S}_{\theta}:=\bigcup_{j}\mathbb{S}_{\theta,j}. Each slab in 𝕊θ,j\mathbb{S}_{\theta,j} has dimensions 2j×1×12^{-j}\times 1\times 1 and normal direction γ(θ)\gamma(\theta). One easily sees that AS𝕊θSA\subset\bigcup_{S\in\mathbb{S}_{\theta}}S. By pigeonholing, there exists j(θ)j(\theta) such that

(7) νA(A(S𝕊θ,j(θ)S))110j(θ)2νA(A)=110j(θ)2.\nu_{A}\big{(}A\cap(\cup_{S\in\mathbb{S}_{\theta,j(\theta)}}S)\big{)}\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

(8) 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\delta=2^{-j}. By Lemma 1, there exists a (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 U:={(x,θ)A×Θ:xS𝕊θ,jS}U:=\{(x,\theta)\in A\times\Theta:x\in\cup_{S\in\mathbb{S}_{\theta,j}}S\}. We also use μ\mu to denote the counting measure on Θ\Theta (note that Θ\Theta is a finite set). Define the section of UU:

Ux={θ:(x,θ)U},Uθ:={x:(x,θ)U}.U_{x}=\{\theta:(x,\theta)\in U\},\ \ \ U_{\theta}:=\{x:(x,\theta)\in U\}.

By (7) and Fubini, we have

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

This implies

(10) (νA×μ)({(x,θ)U:μ(Ux)120j2μ(Θ)})120j2μ(Θ),(\nu_{A}\times\mu)\bigg{(}\Big{\{}(x,\theta)\in U:\mu(U_{x})\geq\frac{1}{20j^{2}}\mu(\Theta)\Big{\}}\bigg{)}\geq\frac{1}{20j^{2}}\mu(\Theta),

since

(11) (νA×μ)({(x,θ)U:μ(Ux)120j2μ(Θ)})120j2μ(Θ).(\nu_{A}\times\mu)\bigg{(}\Big{\{}(x,\theta)\in U:\mu(U_{x})\leq\frac{1}{20j^{2}}\mu(\Theta)\Big{\}}\bigg{)}\leq\frac{1}{20j^{2}}\mu(\Theta).

By (10), we have

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

We are ready to apply Theorem 2. Recall δ=2j\delta=2^{-j} and #Θ(logδ1)2δt\#\Theta\gtrsim(\log\delta^{-1})^{-2}\delta^{-t}. By (12) and noting that νA(Bδ)δa\nu_{A}(B_{\delta})\lesssim\delta^{a}, we can find a δ\delta-separated subset of {xA:#Ux120j2#Θ}\{x\in A:\#U_{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 slabs from θΘ𝕊θ,j\cup_{\theta\in\Theta}\mathbb{S}_{\theta,j} that intersect BδB_{\delta}. We can now apply Theorem 2 to obtain

(logδ1)8δa4t(#Θ)4#HCs,εδ2ts2ε.(\log\delta^{-1})^{-8}\delta^{-a-4t}\lesssim(\#\Theta)^{4}\#H\leq C_{s,\varepsilon}\delta^{-2t-s-2-\varepsilon}.

Letting ϵ0\epsilon_{\circ}\rightarrow 0 (and hence δ0\delta\rightarrow 0) and then ε0\varepsilon\rightarrow 0, we obtain a2+s2ta\leq 2+s-2t. ∎

2.2. Proof of Theorem 2

For convenience, we will prove the following version of Theorem 2 after rescaling xδ1xx\mapsto\delta^{-1}x.

Theorem 3.

Fix 0<s<10<s<1. 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×δ1×δ11\times\delta^{-1}\times\delta^{-1}-slabs 𝕊θ\mathbb{S}_{\theta} with normal direction γ(θ)\gamma(\theta). 𝕊θ\mathbb{S}_{\theta} satisfies the ss-dimensional condition:

  1. (1)

    #𝕊θδs\#\mathbb{S}_{\theta}\lesssim\delta^{-s},

  2. (2)

    #{S𝕊θ:SBr}rs\#\{S\in\mathbb{S}_{\theta}:S\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 slabs from θ𝕋θ\cup_{\theta}\mathbb{T}_{\theta}. Then

(#Θ)4#HCs,εδ2ts2ε.(\#\Theta)^{4}\#H\leq C_{s,\varepsilon}\delta^{-2t-s-2-\varepsilon}.

We define the cone

(13) Γ:={rγ(θ):1/2r1,θ[0,1]}.\Gamma:=\{r\gamma(\theta):1/2\leq r\leq 1,\theta\in[0,1]\}.

For any large scale RR, there is a standard partition of NR1ΓN_{R^{-1}}\Gamma into planks σR1/2\sigma_{R^{-1/2}} of dimensions R1×R1/2×1R^{-1}\times R^{-1/2}\times 1:

NR1Γ=σR1/2.N_{R^{-1}}\Gamma=\bigcup\sigma_{R^{-1/2}}.

Here, the subscript of σR1/2\sigma_{R^{-1/2}} denotes its angular size. For any function ff and plank σ=ψR1/2\sigma=\psi_{R^{-1/2}}, we define fσ:=(1σf^)f_{\sigma}:=(1_{\sigma}\widehat{f})^{\vee} as usual. The main tool we need is the following fractal small cap decoupling for the cone Γ\Gamma.

Theorem 4 (fractal small cap decoupling).

Suppose Nδ(Γ)=γN_{\delta}(\Gamma)=\bigcup\gamma, where each γ\gamma is a δ×δ×1\delta\times\delta\times 1-cap. Given a function gg, we say gg is tt-spacing if suppg^γΓgγ\mathrm{supp}\widehat{g}\subset\cup_{\gamma\in\Gamma_{g}}\gamma, where Γg\Gamma_{g} is a set of δ×δ×1\delta\times\delta\times 1-caps from the partition of Nδ(Γ)N_{\delta}(\Gamma) and satisfies:

(14) #{γΓg:γσr}(r/δ)t, for any r2×r×1plank σrNr2Γ(δr1).\#\{\gamma\in\Gamma_{g}:\gamma\subset\sigma_{r}\}\lesssim(r/\delta)^{t},\textup{~{}for~{}any~{}}r^{2}\times r\times 1-\textup{plank~{}}\sigma_{r}\subset N_{r^{2}}\Gamma\ \ (\delta\leq r\leq 1).

If gg is tt-spacing, then we have

(15) Bδ1|g|4εδεtγ|gγ|4.\int_{B_{\delta^{-1}}}|g|^{4}\lesssim_{\varepsilon}\delta^{-\varepsilon-t}\sum_{\gamma}\int|g_{\gamma}|^{4}.

Small cap decoupling for the cone was studied by the second and third authors in [7], where they proved amplitude-dependent versions of the wave envelope estimates (Theorem 1.3) of [4]. Wave envelope estimates are a more refined version of square function estimates, and sharp small cap decoupling is a straightforward corollary. For certain choices of conical small caps, the critical LpcL^{p_{c}} exponent is pc=4p_{c}=4 (as is the case in our Theorem 4). When pc=4p_{c}=4, the sharp small cap decoupling inequalities follow already from the wave envelope estimates of [4]. A version of this was first observed in Theorem 3.6 of [1] and was later thoroughly explained in §10 of [7]. To prove Theorem 4 above, we repeat the derivation of small cap decoupling from the wave envelope estimates of [4] but incorporate the extra ingredient of tt-spacing.

Remark 5.

We will actually apply Theorem 4 to a slightly different cone

(16) ΓK1={rγ(θ):K1r1,θ[0,1]}.\Gamma_{K^{-1}}=\{r\gamma(\theta):K^{-1}\leq r\leq 1,\theta\in[0,1]\}.

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 fractal small cap decoupling for ΓK1\Gamma_{K^{-1}}. Instead of (15), we have

(17) Bδ1|g|4εKO(1)δεtγ|gγ|4.\int_{B_{\delta^{-1}}}|g|^{4}\lesssim_{\varepsilon}K^{O(1)}\delta^{-\varepsilon-t}\sum_{\gamma}\int|g_{\gamma}|^{4}.

The idea is to partition ΓK1\Gamma_{K^{-1}} into K\sim K many parts, each of which is roughly a cone that we can apply Theorem 4 to. By triangle inequality, it gives (17) with an additional factor KO(1)K^{O(1)}. It turns out that this factor is not harmful, since we will set K(logδ1)O(1)K\sim(\log\delta^{-1})^{O(1)} which can be absorbed into δε\delta^{-\varepsilon}.

We postpone the proof of Theorem 4 to the next subsection, and first show how it implies Theorem 3.

Proof of Theorem 3 assuming Theorem 4.

We consider the dual of each Sθ𝕊θS_{\theta}\in\mathbb{S}_{\theta} in the frequency space. For each θΘ\theta\in\Theta, we define τθ\tau_{\theta} to be a tube centered at the origin that has dimensions δ×δ×1\delta\times\delta\times 1, and its direction is γ(θ)\gamma(\theta). We see that τθ\tau_{\theta} is the dual of each Sθ𝕊θS_{\theta}\in\mathbb{S}_{\theta}. Now, for each Sθ𝕊θS_{\theta}\in\mathbb{S}_{\theta}, we choose a bump function ψSθ\psi_{S_{\theta}} satisfying the following properties: ψSθ1\psi_{S_{\theta}}\geq 1 on SθS_{\theta}, ψSθ\psi_{S_{\theta}} decays rapidly outside SθS_{\theta}, and suppψ^Sθτθ\mathrm{supp}\widehat{\psi}_{S_{\theta}}\subset\tau_{\theta}.

Define functions

fθ=Sθ𝕊θψSθandf=θΘfθ.f_{\theta}=\sum_{S_{\theta}\in\mathbb{S}_{\theta}}\psi_{S_{\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)(logδ1)2#Θf(x)\gtrsim(\log\delta^{-1})^{-2}\#\Theta. Therefore, we obtain

(18) |H|(#Θ)4H|f|4,|H|(\#\Theta)^{4}\lessapprox\int_{H}|f|^{4},

where “\lessapprox” means “(logδ1)O(1)\lesssim(\log\delta^{-1})^{O(1)}”.

Next, we will do a high-low decomposition for each τθ\tau_{\theta}.

Definition 3.

Let KK be a large number which we will choose later. Define the high part of τθ\tau_{\theta} as

τθ,high:={ξτθ:K1|ξγ(θ)|1}.\tau_{\theta,high}:=\{\xi\in\tau_{\theta}:K^{-1}\leq|\xi\cdot\gamma(\theta)|\leq 1\}.

Define the low part of τθ\tau_{\theta} as

τθ,low:=τθτθ,high={ξτθ:|ξγ(θ)|K1}.\tau_{\theta,low}:=\tau_{\theta}\setminus\tau_{\theta,high}=\{\xi\in\tau_{\theta}:|\xi\cdot\gamma(\theta)|\leq K^{-1}\}.

We choose a smooth partition of unity adapted to the covering τθ=τθ,highτθ,low\tau_{\theta}=\tau_{\theta,high}\bigcup\tau_{\theta,low} which we denote by ηθ,high,ηθ,low\eta_{\theta,high},\eta_{\theta,low}, so that

ηθ,high+ηθ,low=1\eta_{\theta,high}+\eta_{\theta,low}=1

on τθ\tau_{\theta}. The key observation is that {suppη^θ,high}θ\{\mathrm{supp}\widehat{\eta}_{\theta,high}\}_{\theta} are at most O(K)O(K)-overlapping and form a canonical covering of Nδ(ΓK1)N_{\delta}(\Gamma_{K^{-1}}). (See the definition of ΓK1\Gamma_{K^{-1}} in (16)).

Since suppf^θτθ\mathrm{supp}\widehat{f}_{\theta}\subset\tau_{\theta}, we also obtain a decomposition of fθf_{\theta}

(19) fθ=fθ,high+fθ,low,f_{\theta}=f_{\theta,high}+f_{\theta,low},

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

(20) f=fhigh+flow,f=f_{high}+f_{low},

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

Recall that for xHx\in H, we have

(logδ1)2#Θf(x)|fhigh(x)|+|flow(x)|.(\log\delta^{-1})^{-2}\#\Theta\lesssim f(x)\leq|f_{high}(x)|+|f_{low}(x)|.

We will show that by properly choosing KK, we have

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

Recall that flow=θfθηθ,lowf_{low}=\sum_{\theta}f_{\theta}*\eta^{\vee}_{\theta,low}. Since ηθ,low\eta_{\theta,low} is a bump function at τθ,low\tau_{\theta,low}, we see that ηθ,low\eta_{\theta,low}^{\vee} is an L1L^{1}-normalized bump function essentially supported in the dual of τθ,low\tau_{\theta,low}. Denote the dual of τθ,low\tau_{\theta,low} by Sθ,KS_{\theta,K} which is a δ1×δ1×K\delta^{-1}\times\delta^{-1}\times K-slab whose normal direction is γ(θ)\gamma(\theta). One actually has

|ηθ,low|1|Sθ,K|ψSθ,K.|\eta^{\vee}_{\theta,low}|\lesssim\frac{1}{|S_{\theta,K}|}\psi_{S_{\theta,K}}.

Here, ψSθ,K\psi_{S_{\theta,K}} is bump function =1=1 on Sθ,KS_{\theta,K} and decays rapidly outside Sθ,KS_{\theta,K}. Ignoring the rapidly decaying tails, we have

|flow(x)|θ1K#{Sθ𝕊θ:SθB100K(x)}.|f_{low}(x)|\lesssim\sum_{\theta}\frac{1}{K}\#\{S_{\theta}\in\mathbb{S}_{\theta}:S_{\theta}\cap B_{100K}(x)\neq\emptyset\}.

Recalling the condition (2) in Theorem 3, we have

#{Sθ𝕊θ:SθB100K(x)}(100K)s.\#\{S_{\theta}\in\mathbb{S}_{\theta}:S_{\theta}\cap B_{100K}(x)\neq\emptyset\}\lesssim(100K)^{s}.

This implies

|flow(x)|Ks1#Θ.|f_{low}(x)|\lesssim K^{s-1}\#\Theta.

Since s<1s<1, by choosing K(logδ1)21sK\sim(\log\delta^{-1})^{\frac{2}{1-s}}, we obtain (21). This shows that for sHs\in H, we have

(logδ1)2#Θ|f(x)||fhigh|.(\log\delta^{-1})^{-2}\#\Theta\lesssim|f(x)|\lesssim|f_{high}|.

We define g=fhighg=f_{high}. By remark (5), we actually see that {τθ,high}\{\tau_{\theta,high}\} form a KK-overlapping covering of Nδ(ΓK)N_{\delta}(\Gamma_{K}), and we have the decoupling inequality (17). By (18), we have

|H|δ4tH|f|4Bδ1|fhigh|4.|H|\delta^{-4t}\lesssim\int_{H}|f|^{4}\lesssim\int_{B_{\delta^{-1}}}|f_{high}|^{4}.

By (17), it is further bounded by

εδtεθ|fθ,high|4δtεθ|Sθ𝕊θψSθ|4.\lesssim_{\varepsilon}\delta^{-t-\varepsilon}\sum_{\theta}\int|f_{\theta,high}|^{4}\lesssim\delta^{-t-\varepsilon}\sum_{\theta}\int|\sum_{S_{\theta}\in\mathbb{S}_{\theta}}\psi_{S_{\theta}}|^{4}.

Since the slabs in 𝕊θ\mathbb{S}_{\theta} are essentially disjoint, the above expression is bounded by

δtεθSθ𝕊θ|ψSθ|4δtεθSθ𝕊θ|Sθ|δtεδst1.\lesssim\delta^{-t-\varepsilon}\sum_{\theta}\int\sum_{S_{\theta}\in\mathbb{S}_{\theta}}|\psi_{S_{\theta}}|^{4}\sim\delta^{-t-\varepsilon}\sum_{\theta}\sum_{S_{\theta}\in\mathbb{S}_{\theta}}|S_{\theta}|\sim\delta^{-t-\varepsilon}\delta^{-s-t-1}.

This implies (#Θ)4#Hεδ2ts1ε(\#\Theta)^{4}\#H\lesssim_{\varepsilon}\delta^{2t-s-1-\varepsilon}.

2.3. Proof of Theorem 4

The proof of Theorem 4 is based on an inequality of Guth, Wang and Zhang. Let us first introduce some notation from their paper [4]. Let Γ\Gamma_{\circ} denote the standard cone in 3\mathbb{R}^{3}:

Γ:={(rcosθ,rsinθ,r):1/2r1,θ[0,2π]}.\Gamma_{\circ}:=\{(r\cos\theta,r\sin\theta,r):1/2\leq r\leq 1,\theta\in[0,2\pi]\}.

We can partition the δ\delta-neighborhood of Γ\Gamma_{\circ} into δ×δ1/2×1\delta\times\delta^{1/2}\times 1-planks Σ={σ}\Sigma=\{\sigma\}:

Nδ(Γ)=σ.N_{\delta}(\Gamma_{\circ})=\bigsqcup\sigma.

More generally, for any dyadic ss in the range δ1/2s1\delta^{1/2}\leq s\leq 1, we can partition the s2s^{2}-neighborhood of Γ\Gamma_{\circ} into s2×s×1s^{2}\times s\times 1-planks 𝒯s={τs}\mathcal{T}_{s}=\{\tau_{s}\}:

Ns2(Γ)=τs.N_{s^{2}}(\Gamma_{\circ})=\bigsqcup\tau_{s}.

For each ss and frequency plank τs𝒯s\tau_{s}\in\mathcal{T}_{s}, we define the box UτsU_{\tau_{s}} in the physical space to be a rectangle centered at the origin of dimensions δ1×δ1s×δ1s2\delta^{-1}\times\delta^{-1}s\times\delta^{-1}s^{2} whose edge of length δ1\delta^{-1} (respectively δ1s\delta^{-1}s, δ1s2\delta^{-1}s^{2}) is parallel to the edge of τs\tau_{s} with length s2s^{2} (respectively ss, 11). Note that for any σΣ\sigma\in\Sigma, UσU_{\sigma} is just the dual rectangle of σ\sigma. Also, UτsU_{\tau_{s}} is the convex hull of στsUσ\cup_{\sigma\subset\tau_{s}}U_{\sigma}.

If UU is a translated copy of UτsU_{\tau_{s}}, then we define SUfS_{U}f by

(22) SUf=(στs|fσ|2)1/2𝟏U.S_{U}f=\big{(}\sum_{\sigma\subset\tau_{s}}|f_{\sigma}|^{2}\big{)}^{1/2}\boldsymbol{1}_{U}.

We can think of SUfS_{U}f as the wave envelope of ff localized in UU in the physical space and localized in τs\tau_{s} in the frequency space. We have the following inequality of Guth, Wang and Zhang (see [4] Theorem 1.5):

Theorem 5 (Wave envelope estimate).

Suppose suppf^Nδ(Γ)\mathrm{supp}\widehat{f}\subset N_{\delta}(\Gamma_{\circ}). Then

(23) f44Cεδεδ1/2s1τs𝒯sUUτs|U|1SUf24.\|f\|_{4}^{4}\leq C_{\varepsilon}\delta^{-\varepsilon}\sum_{\delta^{1/2}\leq s\leq 1}\sum_{\tau_{s}\in\mathcal{T}_{s}}\sum_{U\parallel U_{\tau_{s}}}|U|^{-1}\|S_{U}f\|_{2}^{4}.

Although the theorem above is stated for the standard cone Γ\Gamma_{\circ}, it is also true for general cone Γ\Gamma (see (13)). The appendix of [7] shows how to adapt the inductive proof of Guth-Wang-Zhang for Γ\Gamma_{\circ} to the case of a general cone Γ\Gamma.

As we did for Γ\Gamma_{\circ}, we can also define the δ×δ1/2×1\delta\times\delta^{1/2}\times 1-planks Σ={σ}\Sigma=\{\sigma\} and s2×s×1s^{2}\times s\times 1-planks 𝒯s={τs}\mathcal{T}_{s}=\{\tau_{s}\}, which form a partition of certain neighborhood of Γ\Gamma. We can similarly define the wave envelope SUfS_{U}f for suppf^Nδ(Γ)\mathrm{supp}\widehat{f}\subset N_{\delta}(\Gamma). We have the following estimate for general cone.

Theorem 6 (Wave envelope estimate for general cone).

Suppose suppf^Nδ(Γ)\mathrm{supp}\widehat{f}\subset N_{\delta}(\Gamma). Then

(24) f44Cεδεδ1/2s1τs𝒯sUUτs|U|1SUf24.\|f\|_{4}^{4}\leq C_{\varepsilon}\delta^{-\varepsilon}\sum_{\delta^{1/2}\leq s\leq 1}\sum_{\tau_{s}\in\mathcal{T}_{s}}\sum_{U\parallel U_{\tau_{s}}}|U|^{-1}\|S_{U}f\|_{2}^{4}.

We are ready to prove Theorem 4.

Proof of Theorem 4.

By pigeonholing, we can assume all the wave packet of gγg_{\gamma} have amplitude 1\sim 1, so we have

(25) |gγ|4|gγ|2.\int|g_{\gamma}|^{4}\sim\int|g_{\gamma}|^{2}.

Apply Theorem 6 to gg, we have

g44Cεδεδ1/2s1τs𝒯sUUτs|U|1SUg24.\|g\|_{4}^{4}\leq C_{\varepsilon}\delta^{-\varepsilon}\sum_{\delta^{1/2}\leq s\leq 1}\sum_{\tau_{s}\in\mathcal{T}_{s}}\sum_{U\parallel U_{\tau_{s}}}|U|^{-1}\|S_{U}g\|_{2}^{4}.

For fixed s,τs,UUτss,\tau_{s},U\parallel U_{\tau_{s}}, let us analyze the quantity SUg22\|S_{U}g\|_{2}^{2} on the right hand side. By definition,

SUg22=Uστs|gσ|2.\|S_{U}g\|_{2}^{2}=\int_{U}\sum_{\sigma\subset\tau_{s}}|g_{\sigma}|^{2}.

Note that UU has dimensions δ1×δ1s×δ1s2\delta^{-1}\times\delta^{-1}s\times\delta^{-1}s^{2}, so its dual UU^{*} has dimensions δ×δs1×δs2\delta\times\delta s^{-1}\times\delta s^{-2}. We will apply local orthogonality to each fσf_{\sigma} on UU. Let {β}\{\beta\} be a set of (δs1)2×δs1×1(\delta s^{-1})^{2}\times\delta s^{-1}\times 1-planks that form a finitely overlapping covering of Nδs1(Γ)N_{\delta s^{-1}}(\Gamma). We see that UU^{*} and each β\beta have the same angular size δs1\delta s^{-1}. For reader’s convenience, we recall that we have defined three families of planks: {γ:γΓg}\{\gamma:\gamma\in\Gamma_{g}\} of dimensions δ×δ×1\delta\times\delta\times 1; {β}\{\beta\} of dimensions (δs1)2×δs1×1(\delta s^{-1})^{2}\times\delta s^{-1}\times 1; {σ}\{\sigma\} of dimensions δ×δ1/2×1\delta\times\delta^{1/2}\times 1.

Since δ1/2s1\delta^{1/2}\leq s\leq 1, we have the nested property for these planks: each γ(Γg)\gamma(\in\Gamma_{g}) is contained in 100100-dilation of some β\beta and each β\beta is contained in 100100-dilation of some σ\sigma. We simply denote this relationship by γβ,βσ\gamma\subset\beta,\beta\subset\sigma. We can write

gσ=βσgβ,gβ=γβgγ.g_{\sigma}=\sum_{\beta\subset\sigma}g_{\beta},\ \ \ g_{\beta}=\sum_{\gamma\subset\beta}g_{\gamma}.

Choose a smooth bump function ψU\psi_{U} at UU satisfying: |ψU|𝟏U|\psi_{U}|\gtrsim\boldsymbol{1}_{U}, ψU\psi_{U} decays rapidly outside UU, and ψ^U\widehat{\psi}_{U} is supported in UU^{*}. We have

U|gσ|2|ψUβσgβ|2.\int_{U}|g_{\sigma}|^{2}\lesssim\int|\psi_{U}\sum_{\beta\subset\sigma}g_{\beta}|^{2}.

Since (ψUgβ)U+β(\psi_{U}g_{\beta})^{\wedge}\subset U^{*}+\beta and by a geometric observation that {U+β}βσ\{U^{*}+\beta\}_{\beta\subset\sigma} are finitely overlapping, we have

U|gσ|2βσ|ψUgβ|2=βσ|ψUγβgγ|2βσ#{γβ}γβ|ψUgγ|2.\int_{U}|g_{\sigma}|^{2}\lesssim\int\sum_{\beta\subset\sigma}|\psi_{U}g_{\beta}|^{2}=\int\sum_{\beta\subset\sigma}|\psi_{U}\sum_{\gamma\subset\beta}g_{\gamma}|^{2}\lesssim\int\sum_{\beta\subset\sigma}\#\{\gamma\subset\beta\}\sum_{\gamma\subset\beta}|\psi_{U}g_{\gamma}|^{2}.

Summing over στs\sigma\subset\tau_{s}, we get

(26) SUg22=Uστs|gσ|2\displaystyle\|S_{U}g\|_{2}^{2}=\int_{U}\sum_{\sigma\subset\tau_{s}}|g_{\sigma}|^{2} στsβσ#{γβ}γβ|ψUgγ|2\displaystyle\lesssim\int\sum_{\sigma\subset\tau_{s}}\sum_{\beta\subset\sigma}\#\{\gamma\subset\beta\}\sum_{\gamma\subset\beta}|\psi_{U}g_{\gamma}|^{2}
(27) (στsβσ#{γβ})(supβγβ|gγ|2)|ψU|2\displaystyle\lesssim\big{(}\sum_{\sigma\subset\tau_{s}}\sum_{\beta\subset\sigma}\#\{\gamma\subset\beta\}\big{)}(\sup_{\beta}\sum_{\gamma\subset\beta}|g_{\gamma}|^{2})\int|\psi_{U}|^{2}
(28) (Since gγ1)\displaystyle(\textup{Since~{}}\|g_{\gamma}\|_{\infty}\leq 1)\ \ \ #{γτs}#{γβ}|U|\displaystyle\lesssim\#\{\gamma\subset\tau_{s}\}\#\{\gamma\subset\beta\}|U|
(29) (By the t-spacing condition)\displaystyle(\textup{By~{}the~{}}t\textup{-spacing~{}condition})\ \ \ (s/δ)t(δs1/δ)t|U|\displaystyle\lesssim(s/\delta)^{t}(\delta s^{-1}/\delta)^{t}|U|
(30) =δt|U|.\displaystyle=\delta^{-t}|U|.

Therefore, we have

(31) U|U|1SUg24δtUSUg22=δtστs|gσ|2δtγτs|gγ|2.\sum_{U}|U|^{-1}\|S_{U}g\|_{2}^{4}\lesssim\delta^{-t}\sum_{U}\|S_{U}g\|_{2}^{2}=\delta^{-t}\int\sum_{\sigma\subset\tau_{s}}|g_{\sigma}|^{2}\lesssim\delta^{-t}\int\sum_{\gamma\subset\tau_{s}}|g_{\gamma}|^{2}.

The last inequality is by the L2L^{2} orthogonality.

Noting (25), we have

(32) g44Cεδεtδ1/2s1γ|gγ|2Cεδεtδ1/2s1γ|gγ|4Cεδ2εtγ|gγ|4.\|g\|_{4}^{4}\leq C_{\varepsilon}\delta^{-\varepsilon-t}\sum_{\delta^{1/2}\leq s\leq 1}\sum_{\gamma}\int|g_{\gamma}|^{2}\sim C_{\varepsilon}\delta^{-\varepsilon-t}\sum_{\delta^{1/2}\leq s\leq 1}\sum_{\gamma}\int|g_{\gamma}|^{4}\lesssim C_{\varepsilon}\delta^{-2\varepsilon-t}\sum_{\gamma}\int|g_{\gamma}|^{4}.

References

  • [1] C. Demeter, L. Guth, and H. Wang. Small cap decouplings. Geometric and Functional Analysis, 30(4):989–1062, 2020.
  • [2] K. Fässler and T. Orponen. On restricted families of projections in 3\mathbb{R}^{3}. Proceedings of the London Mathematical Society, 109(2):353–381, 2014.
  • [3] S. Gan, S. Guo, L. Guth, T. L. Harris, D. Maldague, and H. Wang. On restricted projections to planes in 3\mathbb{R}^{3}. arXiv preprint arXiv:2207.13844, 2022.
  • [4] L. Guth, H. Wang, and R. Zhang. A sharp square function estimate for the cone in 3\mathbb{R}^{3}. Annals of Mathematics, 192(2):551–581, 2020.
  • [5] T. L. Harris. Length of sets under restricted families of projections onto lines. arXiv preprint arXiv:2208.06896, 2022.
  • [6] A. Käenmäki, T. Orponen, and L. Venieri. A Marstrand-type restricted projection theorem in 3\mathbb{R}^{3}. arXiv preprint arXiv:1708.04859, 2017.
  • [7] D. Maldague and L. Guth. Amplitude dependent wave envelope estimates for the cone in 3\mathbb{R}^{3}, 2022.
  • [8] J. M. Marstrand. Some fundamental geometrical properties of plane sets of fractional dimensions. Proceedings of the London Mathematical Society, 3(1):257–302, 1954.
  • [9] P. Mattila. Hausdorff dimension, orthogonal projections and intersections with planes. Ann. Acad. Sci. Fenn. Ser. AI Math, 1(2):227–244, 1975.
  • [10] 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.