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

Discretized Radial Projections in \Rd\R^{d}

Kevin Ren
(June 2022)
Abstract

We generalize a Furstenberg-type result of Orponen-Shmerkin to higher dimensions, leading to an ε\varepsilon-improvement in Kaufman’s projection theorem for hyperplanes and an unconditional discretized radial projection theorem in the spirit of Orponen-Shmerkin-Wang. Our proof relies on a new incidence estimate for δ\delta-tubes and a quasi-product set of δ\delta-balls in \Rd\R^{d}.

1 Introduction

Let XX be a set in \Rn\R^{n}, and define the radial projection πx(y):=yx|yx|Sn1\pi_{x}(y):=\frac{y-x}{|y-x|}\in S^{n-1}. We wish to study the size of radial projections πx(Y)\pi_{x}(Y) of YY, where xx is taken in some set XX. Recently, Orponen, Shmerkin, and Wang [19] proved a strong radial projection theorem in two dimensions, but they prove a conditional result in higher dimensions. In this paper, we shall remove the condition dimH(X)k1k+η(k)\dim_{H}(X)\geq k-\frac{1}{k}+\eta(k) in higher dimensions, which answers Conjecture 1.5 of [26] and improves Theorem 1.9 of [19]. We also improve upon the previously known result d1dmin(dimH(X),dimH(Y))+η(d,dimH(X),dimH(Y))\frac{d-1}{d}\min(\dim_{H}(X),\dim_{H}(Y))+\eta(d,\dim_{H}(X),\dim_{H}(Y)) of [25, Theorem 6.15].

Theorem 1.1.

Let X,Y\RdX,Y\subset\R^{d} be Borel sets with dimH(X),dimH(Y)k\dim_{H}(X),\dim_{H}(Y)\leq k. If XX is not contained in a kk-plane, then

supxXdimH(πx(Y{x}))min(dimH(X),dimH(Y)).\sup_{x\in X}\dim_{H}(\pi_{x}(Y\setminus\{x\}))\geq\min(\dim_{H}(X),\dim_{H}(Y)).

In fact, we can prove the following slicing result, which improves Proposition 6.8 of [19] and makes progress towards answering Conjecture 1.10 of [19].

Corollary 1.2.

Let s(d2,d]s\in(d-2,d], then there exists ε(s,d)>0\varepsilon(s,d)>0 such that the following holds. Let μ,ν\mu,\nu be Borel probability measures on \Rd\R^{d} with disjoint supports that satisfy s(μ),s(ν)<\mathcal{E}_{s}(\mu),\mathcal{E}_{s}(\nu)<\infty and dimH(spt(ν))<s+ε(s,d)\dim_{H}(\mathrm{spt}(\nu))<s+\varepsilon(s,d). Further, assume that μ,ν\mu,\nu don’t simultaneously give full measure to any affine (d1)(d-1)-plane H\RdH\subset\R^{d}. Then there exist restrictions of μ,ν\mu,\nu to subsets of positive measure (which we keep denoting μ,ν\mu,\nu) such that the following holds. For almost every affine 2-plane W\RdW\subset\R^{d} (with respect to the natural measure on the affine Grassmanian), if the sliced measures μW\mu_{W}, νW\nu_{W} on WW is non-trivial, then they don’t simultaneously give full measure to any line. In other words,

(γd,2×μ){(V,x):μV,x()νV,x()=|μV,x||νV,x|>0 for some 𝔸(V+x,1)}=0,(\gamma_{d,2}\times\mu)\{(V,x):\mu_{V,x}(\ell)\nu_{V,x}(\ell)=|\mu_{V,x}||\nu_{V,x}|>0\text{ for some }\ell\in\mathbb{A}(V+x,1)\}=0,

where we parametrize affine 2-planes as V+xV+x, for x\Rdx\in\R^{d} and VV in the Grassmannian Gr(d,2)\mathrm{Gr}(d,2) with the rotationally invariant Haar measure γd,2\gamma_{d,2}.

We also deduce an ε\varepsilon-improvement in Kaufman’s projection theorem for hyperplanes. The proof is a standard higher-dimensional generalization of the details in [18, Section 3.2] and we will omit it. For σSn1\sigma\in S^{n-1}, let πσ\pi_{\sigma} be projection in the direction orthogonal to σ\sigma.

Theorem 1.3.

For every k<s<tdk<s<t\leq d, there exists ε(s,t)\varepsilon(s,t) such that the following holds. Let EE be an analytic set in \Rd\R^{d} with dimH(E)=t\dim_{H}(E)=t. Then

dimH{σSd1:dimH(πσ(E))s}sε.\dim_{H}\{\sigma\in S^{d-1}:\dim_{H}(\pi_{\sigma}(E))\leq s\}\leq s-\varepsilon.
Remark 1.4.

Kaufman’s theorem is sharp when s=ks=k and t(k,k+1]t\in(k,k+1] because EE can be contained within a (k+1)(k+1)-plane.

We also derive a higher-dimensional version of Beck’s theorem (unlike in the discrete setting, the higher-dimensional version cannot proved by projection onto a generic 2D plane). The proof again follows similarly to the 2D version presented in [19, Corollary 1.4].

Corollary 1.5.

Let X\RdX\subset\R^{d} be a Borel set such that dimH(XH)=dimHX\dim_{H}(X\setminus H)=\dim_{H}X for all kk-planes HH. Then, the line set (X)\mathcal{L}(X) spanned by pairs of distinct points in XX satisfies

dimH((X))min{2dimHX,2k}.\dim_{H}(\mathcal{L}(X))\geq\min\{2\dim_{H}X,2k\}.

1.1 Connections and related work

Radial projections have also been used to study the Falconer distance set problem, which asks for lower bounds on the Hausdorff dimension of the distance set Δ(X):={|xy|:x,yX}\Delta(X):=\{|x-y|:x,y\in X\} given the value of dimH(X)\dim_{H}(X) for some X\RdX\in\R^{d}. In two dimensions, Wolff [33] used Fourier analysis to show that if dimH(X)43\dim_{H}(X)\geq\frac{4}{3}, then Δ(X)\Delta(X) has positive Lebesgue measure. Using Orponen’s radial projection theorem [17], Guth-Iosevich-Ou-Wang [7] used a good-bad tube decomposition and decoupling to improve the threshold to dimH(X)54\dim_{H}(X)\geq\frac{5}{4}. See also works of Keleti-Shmerkin [14] [14], Shmerkin [28], Liu [15], and Stull [29] which provide better lower bounds for dimH(Δ(X))\dim_{H}(\Delta(X)) given that dimH(X)(1,54)\dim_{H}(X)\in(1,\frac{5}{4}). In higher dimensions, the works of Du-Iosevich-Ou-Wang-Zhang [3] and Wang-Zheng [32] used a good-bad tube decomposition using Orponen’s radial projection theorem and decoupling techniques [17] to provide state-of-the-art results when the dimension dd is even; when dd is odd, a more classical approach purely based on decoupling gave the best estimates [6], [9]. More recently, Shmerkin and Wang [27] prove a radial projection theorem in the spirit of this paper to provide an improved lower bound when dimH(X)=d2\dim_{H}(X)=\frac{d}{2}, d=2,3d=2,3; using their framework combined with updated results of [19], one can show for example that dimH(Δ(X))58\dim_{H}(\Delta(X))\geq\frac{5}{8} when X\R3X\subset\R^{3} satisfies dimH(X)=32\dim_{H}(X)=\frac{3}{2}. In fact, all of these works prove lower bounds on the size of the pinned distance set, Δx(X):={|xy|:yX}\Delta_{x}(X):=\{|x-y|:y\in X\}. In the forthcoming companion papers [4], [5], we use Theorem 1.1 to improve the lower bounds for the Falconer distance set problem in all dimensions dd.

Very recently, radial projections in dimension 22 have been used to prove the ABC sum-product conjecture and Furstenberg set conjecture, and yield progress on the discretized sum-product problem [21], [22]. It is natural to wonder whether the exciting progress in 2 dimensions will generalize to higher dimensions. The starting point of the breakthrough work of [21] (which was also used in [22]) is a sharp radial projection theorem in 2 dimensions, [19, Theorem 1.1]. We hope to use our higher dimensional radial projection theorem to prove analogous results to [21], [22] in all dimensions.

1.2 Discretized results

We deduce Theorem 1.1 from δ\delta-discretized versions. The following notation will be used throughout this paper.

Definition 1.6.

Let P\RdP\subset\R^{d} be a bounded nonempty set, d1d\geq 1. Let δ>0\delta>0 be a dyadic number, and let 0sd0\leq s\leq d and C>0C>0. We say that PP is a (δ,s,C,k)(\delta,s,C,k)-set if for every (r,k)(r,k)-plate HH with r[δ,1]r\in[\delta,1], we have

|PH|δC|P|δrs.|P\cap H|_{\delta}\leq C\cdot|P|_{\delta}\cdot r^{s}.

If kk is not specified, we default to k=0k=0 (which becomes a standard definition from [18] because (r,0)(r,0)-plates are rr-balls).

Definition 1.7.

Let 𝒯\Rd\mathcal{T}\subset\R^{d} be a bounded nonempty set of dyadic δ\delta-tubes, d2d\geq 2. Let δ>0\delta>0 be a dyadic number, and let 0sd0\leq s\leq d, 0kd20\leq k\leq d-2, and C>0C>0. We say that 𝒯\mathcal{T} is a (δ,s,C,k)(\delta,s,C,k)-set of tubes if for every (r,k+1)(r,k+1)-plate HH and δr1\delta\leq r\leq 1, we have

|𝒯H|C|𝒯|rs.|\mathcal{T}\cap H|\leq C\cdot|\mathcal{T}|\cdot r^{s}. (1.1)

If kk is not specified, we default to k=0k=0. We also say 𝒯\mathcal{T} is a (δ,s,C,k)(\delta,s,C,k)-set of tubes from scale r1r_{1} to r2r_{2} if the non-concentration condition (1.1) holds for r2rr1r_{2}\leq r\leq r_{1}.

A (δ,s,C,k)(\delta,s,C,k)-set of balls cannot be concentrated in a small neighborhood of a kk-plane, while a (δ,s,C,k)(\delta,s,C,k)-set of tubes cannot be concentrated in a small neighborhood of a (k+1)(k+1)-plane.

The main ingredient in the proof of Theorem 1.1 is an ε\varepsilon-improvement to the (dual) Furstenberg set problem that generalizes Theorem 1.3 in [18] to higher dimensions.

Theorem 1.8.

For any 0k<d10\leq k<d-1, 0s<k+10\leq s<k+1, s<tds<t\leq d, κ>0\kappa>0, there exists ε(s,t,κ,k,d)>0\varepsilon(s,t,\kappa,k,d)>0 such that the following holds for all small enough δ2\delta\in 2^{-\mathbb{N}}, depending only on s,t,κ,k,ds,t,\kappa,k,d. Let 𝒫𝒟δ\mathcal{P}\subset\mathcal{D}_{\delta} be a (δ,t,δε)(\delta,t,\delta^{-\varepsilon})-set with 𝒫[0,1)d\cup\mathcal{P}\subset[0,1)^{d}, and let 𝒯𝒯δ\mathcal{T}\subset\mathcal{T}^{\delta} be a family of δ\delta-tubes. Assume that for every p𝒫p\in\mathcal{P}, there exists a (δ,s,δε,0)(\delta,s,\delta^{-\varepsilon},0) and (δ,κ,δε,k)(\delta,\kappa,\delta^{-\varepsilon},k)-set 𝒯(p)𝒯\mathcal{T}(p)\subset\mathcal{T} such that TpT\cap p\neq\emptyset for all T𝒯(p)T\in\mathcal{T}(p). Then |𝒯|δ2sε|\mathcal{T}|\geq\delta^{-2s-\varepsilon}.

Remark 1.9.

The condition of 𝒯(p)\mathcal{T}(p) being a (δ,κ,δε,k+1)(\delta,\kappa,\delta^{-\varepsilon},k+1)-set is to prevent the counterexample in (say) \R3\R^{3} when s=1,t(1,2]s=1,t\in(1,2], and 𝒯\mathcal{T} is a maximal set of δ2\delta^{-2} many essentially distinct tubes in [0,1]2[0,1]^{2}. This condition is automatically taken care of when s>ks>k: any (δ,s,δε,1)(\delta,s,\delta^{-\varepsilon},1)-set is a (δ,κ,δε,k+1)(\delta,\kappa,\delta^{-\varepsilon},k+1)-set with κ=sk\kappa=s-k.

Remark 1.10.

We can make this decay around kk-plane assumption assuming that (1) PP is a (δ,κ,δε,k+1)(\delta,\kappa,\delta^{-\varepsilon},k+1)-set and (2) for pPp\in P, |𝒯(p)P|δε|P||\mathcal{T}(p)\cap P|\geq\delta^{\varepsilon}|P|. This will be useful for radial projection estimates, since we can guarantee (1) by Theorem B.1 of [25] and (2) because we can get rid of δε|P|\lesssim\delta^{\varepsilon}|P| many pairs (p,q)(p,q) for a fixed pp.

In fact, we can prove the following refined version of Theorem 1.8.

Theorem 1.11.

For any 0k<d10\leq k<d-1, 0s<k+10\leq s<k+1, max(s,k)<td\max(s,k)<t\leq d, κ>0\kappa>0, r01r_{0}\leq 1, there exists ε(s,t,κ,k,d)>0\varepsilon(s,t,\kappa,k,d)>0 such that the following holds for all small enough δ/r02(0,δ0)\delta/r_{0}\in 2^{-\mathbb{N}}\cap(0,\delta_{0}), with δ0\delta_{0} depending only on s,t,κ,k,ds,t,\kappa,k,d. Let HH be a (r0,k+1)(r_{0},k+1)-plate, 𝒫𝒟δH\mathcal{P}\subset\mathcal{D}_{\delta}\cap H be a (δ,t,(δ/r0)ε)(\delta,t,(\delta/r_{0})^{-\varepsilon})-set with 𝒫[0,1)d\cup\mathcal{P}\subset[0,1)^{d}, and let 𝒯𝒯δH\mathcal{T}\subset\mathcal{T}^{\delta}\cap H be a family of δ\delta-tubes. Assume that for every p𝒫p\in\mathcal{P}, there exists a set 𝒯(p)𝒯\mathcal{T}(p)\subset\mathcal{T} such that:

  • TpT\cap p\neq\emptyset for all T𝒯(p)T\in\mathcal{T}(p);

  • 𝒯(p)\mathcal{T}(p) is a (δ,s,(δ/r0)εr0ks,0)(\delta,s,(\delta/r_{0})^{-\varepsilon}r_{0}^{k-s},0)-set down from scale rr;

  • 𝒯(p)\mathcal{T}(p) is a (δ,κ,(δ/r0)εr0κ,k)(\delta,\kappa,(\delta/r_{0})^{-\varepsilon}r_{0}^{-\kappa},k)-set.

Then |𝒯|(δr0)εδ2sr02(sk)|\mathcal{T}|\geq(\frac{\delta}{r_{0}})^{-\varepsilon}\delta^{-2s}r_{0}^{2(s-k)}.

Remark 1.12.

(a) Given fixed k,κk,\kappa, the value of ε\varepsilon can be chosen uniformly in a compact subset of {(s,t):0s<k+1,max(s,k)<td}\{(s,t):0\leq s<k+1,\max(s,k)<t\leq d\}. Indeed, if ε>0\varepsilon>0 works for (s,t)(s,t), then ε2\frac{\varepsilon}{2} works in the ε2\frac{\varepsilon}{2}-neighborhood of (s,t)(s,t).

(b) Conjecture: can we replace the condition of being in HH by 𝒯(p)\mathcal{T}(p) being a (δ,k,(δ/r0)ε,0)(\delta,k,(\delta/r_{0})^{-\varepsilon},0)-set from scales 11 to r0r_{0}?

Using Theorem 1.11, a bootstrap argument based on [19] gives the following.

Theorem 1.13.

Let k{1,2,,d1}k\in\{1,2,\cdots,d-1\}, k1<σ<skk-1<\sigma<s\leq k, and ε>0\varepsilon>0. There exist N,K0N,K_{0} depending on σ,s,k\sigma,s,k, and η(ε)>0\eta(\varepsilon)>0 (with η(1)=1\eta(1)=1) such that the following holds. Fix r01r_{0}\leq 1, and KK0K\geq K_{0}. Let μ,ν\mu,\nu be 1\sim 1-separated ss-dimensional measures with constant Cμ,CνC_{\mu},C_{\nu} supported on E1,E2E_{1},E_{2}, which lie in B(0,1)B(0,1). Assume that |μ|,|ν|1|\mu|,|\nu|\leq 1. Let AA be the pairs of (x,y)E1×E2(x,y)\in E_{1}\times E_{2} that lie in some K1K^{-1}-concentrated (r0,k)(r_{0},k)-plate. Then there exists a set BE1×E2B\subset E_{1}\times E_{2} with μ×ν(B)Kη\mu\times\nu(B)\lesssim K^{-\eta} such that for every xE1x\in E_{1} and rr-tube TT through xx, we have

ν(T(A|xB|x))rσr0σ(k1)+NεKN.\nu(T\setminus(A|_{x}\cup B|_{x}))\lesssim\frac{r^{\sigma}}{r_{0}^{\sigma-(k-1)+N\varepsilon}}K^{N}.

The implicit constant may depend on Cμ,Cν,σ,s,kC_{\mu},C_{\nu},\sigma,s,k.

Remark 1.14.

(a) It is not assumed that μ,ν\mu,\nu are a probability measures, just that μ(B(0,1)),ν(B(0,1))1\mu(B(0,1)),\nu(B(0,1))\leq 1.

(b) If α>d1\alpha>d-1, then the numerology of Theorem 1.13 doesn’t apply. Instead, Orponen’s radial projection theorem [17] in dimension dd applies. The result (stated in [7, Lemma 3.6] for d=2d=2, but can be generalized to all dimensions dd) is that for γ=ε/C\gamma=\varepsilon/C, there exists a set BE1×E2B\subset E_{1}\times E_{2} with μ1×μ2(B)rγ\mu_{1}\times\mu_{2}(B)\leq r^{\gamma} such that for every xE1x\in E_{1} and δ\delta-tube TT through xx, we have

μ2(TB|x)rd1ε.\mu_{2}(T\setminus B|_{x})\lesssim r^{d-1-\varepsilon}.

Note that the set AA of “concentrated pairs” is not needed here.

(c) If rr0r\sim r_{0}, we can obtain a slightly better result by projecting to a generic kk-dimensional subspace and following the argument in [3, Section 3.2]. The result is that for γ=ε/C\gamma=\varepsilon/C, there exists a set BE1×E2B\subset E_{1}\times E_{2} with μ1×μ2(B)δγ\mu_{1}\times\mu_{2}(B)\leq\delta^{\gamma} such that for every xE1x\in E_{1} and rr-tube TT through xx, we have

μ2(TB|x)rk1ε.\mu_{2}(T\setminus B|_{x})\lesssim r^{k-1-\varepsilon}.

The set AA is again not needed in this case. The main novelty of Theorem 1.13 comes when r<r0r<r_{0}.

1.3 Proof ideas

The main proof ideas for Theorem 1.8 are as follows:

  1. 1.

    Perform a standard multiscale decomposition argument due to [18] to reduce the original problem to two building blocks: the case when 𝒫\mathcal{P} is a (δ,s)(\delta,s)-set and when 𝒫\mathcal{P} is a tt-regular set. The first case doesn’t happen all the time and has no loss by an elementary incidence argument, so we focus on gaining an ε\varepsilon-improvement in the second case. A tt-regular set 𝒫\mathcal{P} has the special property that |𝒫Q||\mathcal{P}\cap Q| is still a (Δ,t)(\Delta,t)-set for Q𝒟Δ(𝒫)Q\in\mathcal{D}_{\Delta}(\mathcal{P}), Δ=δ1/2\Delta=\delta^{1/2}.

  2. 2.

    If 𝒫\mathcal{P} is tt-regular with Δ=δ1/2\Delta=\delta^{1/2}, we may find a Δ\Delta-tube 𝐓\mathbf{T} such that upon dilation of 𝐓\mathbf{T} to [0,1]d[0,1]^{d}, we obtain a new Furstenberg problem with the ball set having a quasi-product structure. See Appendix A of [18].

  3. 3.

    Finally, we will use discretized sum-product type arguments to conclude an ε\varepsilon-improvement to the dual Furstenberg problem assuming 𝒫=X×Y\Rd1×\R\mathcal{P}=X\times Y\subset\R^{d-1}\times\R has a quasi-product structure. In very rough terms, we shall lift YY to have dimension close to 11, and apply multi-linear Kakeya. This idea of lifting the dimension was found in He’s work on a higher-rank discretized sum-product theorem [11] in a slightly different context.

To prove Theorem 1.11, we use a similar multiscale decomposition argument as in (1) to reduce to two building blocks: a smaller version of the setting of Theorem 1.11 and a smaller version of Theorem 1.8. The smaller version of Theorem 1.11 has no loss by an elementary incidence argument, and the smaller version of Theorem 1.8 admits a gain.

For Theorem 1.13, we first prove the case when μ,ν\mu,\nu are supported in a r0Kr_{0}K plate (where KK is a small power of r01r_{0}^{-1}). This uses a similar argument as in [19, Lemma 2.8]. The general case follows from applying this special case many times.

1.4 Structure of the paper

In Section 2, we introduce some key concepts that will be used throughout the paper. In Sections 3 through 5, we prove Theorem 1.8 first for quasi-product sets following ideas of [10], and then for regular sets and finally for general sets following [18]. In Section 6, we prove Theorem 1.11 from Theorem 1.8. In Section 7, we generalize a radial projection theorem of Shmerkin [25, Theorem 6.3] that enables us to assume our sets have power decay around kk-planes. In Section 8, we prove Theorem 8.1 following ideas from [19]. Finally, in Section 9, we prove Theorem 1.1 and 1.2 from the discretized results.

Acknowledgments. The author is supported by a NSF GRFP fellowship. The author would like to thank Xiumin Du, Tuomas Orponen, Yumeng Ou, Pablo Shmerkin, Hong Wang, and Ruixiang Zhang for helpful discussions. We thank Paige Bright and Yuqiu Fu for suggesting to include a higher-dimensional version of Beck’s theorem in this paper.

2 Preliminaries

This section will summarize the argument of [18], and in lieu of proofs (with the exception of Proposition 2.9), we either refer the reader to [18] or defer the proof to a later section.

2.1 Definitions

We use ABA\lesssim B to denote ACBA\leq CB for some constant CC. We use ANBA\lesssim_{N}B to indicate the constant CC can depend on NN. We will also use ABA\lessapprox B in future proofs; its exact meaning will always be clarified when used.

For a finite set AA, let |A||A| denote the cardinality of AA. If AA is infinite, let |A||A| denote the Lebesgue measure of AA.

For a set AA, let Ac=\RdAA^{c}=\R^{d}\setminus A.

For a tube TT, let (T)\ell(T) denote the central line segment of TT.

For a set EE, let E(δ)E^{(\delta)} be the δ\delta-neighborhood of EE.

For AX×YA\subset X\times Y and xXx\in X, define the slice A|x={yY:(x,y)A}A|_{x}=\{y\in Y:(x,y)\in A\} and A|y={xX:(x,y)A}A|^{y}=\{x\in X:(x,y)\in A\}.

For a measure μ\mu and a set GG, define the restricted measure μ|G\mu|_{G} by μ|G(A)=μ(GA)\mu|_{G}(A)=\mu(G\cap A). The renormalized restricted measure is μG=1μ(G)μ|G\mu_{G}=\frac{1}{\mu(G)}\mu|_{G}.

For vectors v1,,vi\Rdv_{1},\cdots,v_{i}\in\R^{d}, 1id1\leq i\leq d, the quantity |v1vi||v_{1}\wedge\cdots\wedge v_{i}| is the non-negative volume of the parallelepiped spanned by v1v_{1} through viv_{i}.

B(x,r)B(x,r) is the ball in \Rd\R^{d} of radius rr centered at xx. We also use the notation BrB_{r} for an arbitrary rr-ball in \Rd\R^{d}.

For sets A,BA,B and PA×BP\subset A\times B, let A+𝑃B:={a+b:(a,b)P}A\overset{P}{+}B:=\{a+b:(a,b)\in P\}.

Definition 2.1.

We say μ\mu supported in \Rd\R^{d} is an α\alpha-dimensional measure with constant CμC_{\mu} if μ(Br)Cμrα\mu(B_{r})\leq C_{\mu}r^{\alpha} for all r1r\leq 1 and balls BrB_{r} of radius rr.

2.2 Plates

We work in \Rd\R^{d}. An (r,k)(r,k)-plate is the rr-neighborhood of a kk-dimensional hyperplane in \Rd\R^{d}. We construct a set r,k\mathcal{E}_{r,k} of (r,k)(r,k)-plates with the following properties:

  • Each (r2,k)(\frac{r}{2},k)-plate intersecting B(0,1)B(0,1) lies in at least one plate of r,k\mathcal{E}_{r,k};

  • For srs\geq r, every (s,k)(s,k)-plate contains (sr)(k+1)(dk)\lesssim\left(\frac{s}{r}\right)^{(k+1)(d-k)} many (r,k)(r,k)-plates of r,k\mathcal{E}_{r,k}.

For example, when k=1k=1 and d=2d=2, we can simply pick r1\sim r^{-1} many rr-tubes in each of an rr-net of directions. This generalizes to higher kk and dd via a standard rr-net argument, but we haven’t seen it in the literature, so we provide a precise construction.

An rr-net of a metric space is a subset SS such that B(x,r)B(y,r)=B(x,r)\cap B(y,r)=\emptyset for xy,x,ySx\neq y,x,y\in S. The affine Grassmanian manifold 𝔸(k,d)\mathbb{A}(k,d) is the set of all kk-planes in \Rd\R^{d}. By counting degrees of freedom, we see that dim𝔸(k,d)=(k+1)(dk)\dim\mathbb{A}(k,d)=(k+1)(d-k). Any such plane is uniquely V=V0+aV=V_{0}+a for some kk-dimensional subspace V0V_{0} and aV0a\in V_{0}^{\perp}. For V=V0+aV=V_{0}+a and W=W0+bW=W_{0}+b, define their distance d𝔸d_{\mathbb{A}} to be (following Section 3.16 of [16]):

d𝔸(V,W)=πV0πW0op+|ab|,d_{\mathbb{A}}(V,W)=\|\pi_{V_{0}}-\pi_{W_{0}}\|_{op}+|a-b|,

where πV0:\RdV0\pi_{V_{0}}:\R^{d}\to V_{0} and πW0:\RdW0\pi_{W_{0}}:\R^{d}\to W_{0} are orthogonal projections, and op\|\cdot\|_{op} is the usual operator norm for linear maps. Let 𝔸0(k,d)\mathbb{A}_{0}(k,d) be the submanifold of kk-planes V0+aV_{0}+a with aB(0,10)a\in B(0,10). Since the manifold (𝔸0(k,d),d𝔸)(\mathbb{A}_{0}(k,d),d_{\mathbb{A}}) is compact and smooth, it can be covered by finitely many charts that are 1\sim 1-bilipschitz to a subset of \R(k+1)(dk)\R^{(k+1)(d-k)}.

From a maximal crcr-net 𝒩\mathcal{N} of the set of affine planes of 𝔸0(k,d)\mathbb{A}_{0}(k,d) with c>0c>0 a sufficiently small constant, we can construct a set r,k\mathcal{E}_{r,k} of (r,k)(r,k)-plates whose central planes are the elements of 𝒩\mathcal{N}. We now check the two properties for r,k\mathcal{E}_{r,k}.

To prove the first property, let HH be a (r2,k)(\frac{r}{2},k)-plate intersecting B(0,1)B(0,1). Then the central plane P=PHP=P_{H} must lie at distance 2cr\leq 2cr from some element QQ of 𝒩\mathcal{N} (otherwise, we can add it to the net). Let P=P0+aP=P_{0}+a and Q=Q0+bQ=Q_{0}+b. Hence, πP0πQ0op2cr\|\pi_{P_{0}}-\pi_{Q_{0}}\|_{op}\leq 2cr and |ab|2cr|a-b|\leq 2cr, so for xPB(0,10)x\in P\cap B(0,10) (so xaP0x-a\in P_{0}),

|πQ0(xa)(xa)|2cr|xa|2cr(|x|+|a|)40cr.|\pi_{Q_{0}}(x-a)-(x-a)|\leq 2cr|x-a|\leq 2cr(|x|+|a|)\leq 40cr.

Now, note that πQ0(xa)+bQ\pi_{Q_{0}}(x-a)+b\in Q. It is close to xx if c<1100c<\frac{1}{100}:

|πQ0(xa)+bx|40cr+|ab|50cr<r2.|\pi_{Q_{0}}(x-a)+b-x|\leq 40cr+|a-b|\leq 50cr<\frac{r}{2}.

We have proved PB(0,10)Q(r/2)P\cap B(0,10)\subset Q^{(r/2)} and thus P(r/2)B(0,10)Q(r)P^{(r/2)}\cap B(0,10)\subset Q^{(r)}. Hence, HH is contained in the (r,k)(r,k)-plate with central plane QQ.

To prove the second property, we note that the set of kk-planes in 𝔸(k,d)\mathbb{A}(k,d) whose intersection with B(0,10)B(0,10) is contained in a given (s,k)(s,k)-plate is contained in an O(s)O(s)-ball BB of 𝔸(k,d)\mathbb{A}(k,d). First suppose BB is contained within some coordinate chart; we would like to prove that |𝒩B|(sr)(k+1)(dk)|\mathcal{N}\cap B|\lesssim\left(\frac{s}{r}\right)^{(k+1)(d-k)}. To show this, note that {B(x,r):x𝒩B}\{B(x,r):x\in\mathcal{N}\cap B\} is a packing of B(r)B^{(r)} with finitely overlapping rr-balls. Now map the chart to \R(k+1)(dk)\R^{(k+1)(d-k)}. Since the map only distorts distances by a constant factor, we can pack |𝒩B||\mathcal{N}\cap B| many finitely overlapping c1rc_{1}r-balls into a ball of radius O(s)O(s). Thus by a volume argument, we have |𝒩B|(sr)(k+1)(dk)|\mathcal{N}\cap B|\lesssim\left(\frac{s}{r}\right)^{(k+1)(d-k)}. Since there are finitely many charts, we can apply the argument to BB intersecting each chart, which proves the second property.

We specialize our discussion to tubes. For each scale δ\delta, let 𝒯δ\mathcal{T}^{\delta} be a cover of [0,1]d[0,1]^{d} with δ\delta-tubes such that every δ2\frac{\delta}{2}-tube (and in particular every rr-tube with r<δ2r<\frac{\delta}{2}) is contained in at least 11 and at most CdC_{d} many tubes of 𝒯δ\mathcal{T}^{\delta}. Slightly abusing notation (á la [18]), we will also use 𝒯,𝒯δ,𝒯Δ\mathcal{T},\mathcal{T}_{\delta},\mathcal{T}_{\Delta} to represent sets of tubes, where the subscript δ\delta helpfully indicates a set of δ\delta-tubes.

In Theorem 1.13, we pay attention to certain plates with disproportionately much mass.

Definition 2.2.

We say that a (r,k)(r,k)-plate HH is cc-concentrated on μ\mu if μ(H)c\mu(H)\geq c.

Other notation is following [18]. Unlike [18], we work with ordinary rather than dyadic tubes. The advantage of dyadic tubes is that every 2n2^{-n}-tube is in a unique 2m2^{-m}-tube if n>mn>m; thus, dyadic tubes will avoid the CdC_{d} loss incurred by the finitely overlapping cover 𝒯δ\mathcal{T}^{\delta}. However, dyadic tubes have the disadvantage that they don’t behave well under rotations or dilations, and it would be more cumbersome to define (δ,s,C,k)(\delta,s,C,k)-sets of dyadic tubes (whereas the definition for ordinary tubes is more geometric). Thus, in principle it is possible to work with dyadic tubes and save on the CdC_{d} loss, but it doesn’t affect our numerology in the end (since our losses will depend badly on dd anyway), so we chose to work with ordinary tubes throughout.

Definition 2.3.

[18] Let P\RdP\subset\R^{d} be a bounded nonempty set, d1d\geq 1. Let δ>0\delta>0 be a dyadic number, and let 0sd0\leq s\leq d and C>0C>0. We say that PP is a (δ,s,C)(\delta,s,C)-set if

|PQ|δC|P|δrs,Q𝒟r(\Rd),r[δ,1].|P\cap Q|_{\delta}\leq C\cdot|P|_{\delta}\cdot r^{s},\qquad Q\in\mathcal{D}_{r}(\R^{d}),r\in[\delta,1].
Definition 2.4.

Let 𝒯\Rd\mathcal{T}\subset\R^{d} be a bounded nonempty set of dyadic δ\delta-tubes, d2d\geq 2. Let δ>0\delta>0 be a dyadic number, and let 0sd0\leq s\leq d, 0kd20\leq k\leq d-2, and C>0C>0. We say that 𝒯\mathcal{T} is a (δ,s,C,k)(\delta,s,C,k)-set of tubes if for every (r,k+1)(r,k+1)-plate HH and δr1\delta\leq r\leq 1, we have

|𝒯H|C|𝒯|rs.|\mathcal{T}\cap H|\leq C\cdot|\mathcal{T}|\cdot r^{s}.

If kk is not specified, we default to k=0k=0.

The following is a simpler interpretation of (δ,s,C,k)(\delta,s,C,k)-set if the tubes all pass through the same point.

Definition 2.5.

Let σ(t)Sd1\sigma(t)\in S^{d-1} be the slope of the central axis of tt.

Lemma 2.6.

Let 𝒯\mathcal{T} be a set of δ\delta-tubes intersecting pp. Then if 𝒯\mathcal{T} is a (δ,s,C,k)(\delta,s,C,k)-set, then σ(𝒯)\sigma(\mathcal{T}) is a (δ,s,O(C),k)(\delta,s,O(C),k)-set. Conversely, if σ(𝒯)\sigma(\mathcal{T}) is a (δ,s,C,k)(\delta,s,C,k)-set, then 𝒯\mathcal{T} is a (δ,s,O(C),k)(\delta,s,O(C),k)-set.

Proof.

Let πp:\RdSd1\pi_{p}:\R^{d}\to S^{d-1} denote spherical projection through pp. Then πp(tB(p,1/2))\pi_{p}(t\setminus B(p,1/2)) is well-defined and equals σ(t)\sigma(t), up to an additive loss of CδC\delta. Fix a (r,k)(r,k)-plate HSd1H\in S^{d-1}. Then the set of tubes with slope in HH and passing through pp must lie in a (r+Cδ,k+1)(r+C\delta,k+1)-plate p(Cδ)+πp1(H)p^{(C\delta)}+\pi_{p}^{-1}(H). Conversely, for any (r,k+1)(r,k+1)-plate WW containing pp, the set of possible slopes of tubes through pp contained in WW is contained in a (r+Cδ,k)(r+C\delta,k)-plate (πp(Wp))Cδ(\pi_{p}(W-p))^{C\delta}. ∎

We will need the following lemma from [18].

Lemma 2.7 ([18], Lemma 2.7).

Let P[2,2]dP\subset[-2,2]^{d} be a (δ,s,C)(\delta,s,C)-set. Then PP contains a δ\delta-separated (δ,s,Od(C))(\delta,s,O_{d}(C))-subset PP^{\prime} with |P|δs|P^{\prime}|\leq\delta^{-s}.

First, since (δ,κ,δε,k)(\delta,\kappa,\delta^{-\varepsilon},k)-sets are (δ,κ,δε,k)(\delta,\kappa,\delta^{-\varepsilon},k^{\prime})-sets for any k<kk^{\prime}<k, we can assume that ks<k+1k\leq s<k+1. Next, since (δ,t,δε)(\delta,t,\delta^{-\varepsilon})-sets are (δ,t,δε)(\delta,t^{\prime},\delta^{-\varepsilon})-sets for t<tt^{\prime}<t, we may assume tk+1t\leq k+1. In particular, we get ts1t-s\leq 1, a useful assumption.

We record a useful geometric fact about (r,k)(r,k)-plates.

Lemma 2.8.

Fix Csep1C_{\mathrm{sep}}\geq 1, then there exists r0r_{0} depending on CsepC_{\mathrm{sep}} such that the following is true for r<r0r<r_{0}. If (x,y)(x,y) lie in an (r,k)(r,k)-plate HH and |xy|=Csep1|x-y|=C_{\mathrm{sep}}^{-1}, then any rr-tube TT through x,yx,y will lie in H(CCsepr)H^{(CC_{\mathrm{sep}}r)}, which is a (CCsepr,k)(CC_{\mathrm{sep}}r,k)-plate.

Proof.

For CC sufficiently large: If TT does not lie in H(CCsepr)H^{(CC_{\mathrm{sep}}r)}, then HTH\cap T will be contained in a (2Csep)1(2C_{\mathrm{sep}})^{-1}-tube segment of TT. ∎

2.3 An Elementary Estimate

We prove a classical estimate which can be viewed as Theorem 1.8 with ε=0\varepsilon=0. We won’t need the fact that 𝒯(p)\mathcal{T}(p) is a (δ,κ,δε,k)(\delta,\kappa,\delta^{-\varepsilon},k)-set. The d=2d=2 case is proven as Proposition 2.13 and Corollary 2.14 of [18]. For higher dimensions, the proof is similar and we sketch the details. Let AδBA\lessapprox_{\delta}B denote the inequality

AClog(1δ)CB.A\leq C\cdot\log(\frac{1}{\delta})^{C}B.
Proposition 2.9.

Let 0std10\leq s\leq t\leq d-1, and let CP,CT1C_{P},C_{T}\geq 1. Let 𝒫𝒟δ\mathcal{P}\subset\mathcal{D}_{\delta} be a (δ,t,CP)(\delta,t,C_{P})-set. Assume that for every p𝒫p\in\mathcal{P} there exists a (δ,s,CT)(\delta,s,C_{T})-family 𝒯(p)𝒯δ\mathcal{T}(p)\subset\mathcal{T}^{\delta} of dyadic δ\delta-tubes with the property that TpT\cap p\neq\emptyset for all T𝒯(p)T\in\mathcal{T}(p), and |𝒯(p)|=M|\mathcal{T}(p)|=M for some M1M\geq 1.

Let 𝒯𝒯δ\mathcal{T}\subset\mathcal{T}^{\delta} be arbitrary, and define I(𝒫,𝒯)={(p,T)𝒫×𝒯:T𝒯(p)}I(\mathcal{P},\mathcal{T})=\{(p,T)\in\mathcal{P}\times\mathcal{T}:T\in\mathcal{T}(p)\}. Then

|I(𝒫,𝒯)|δCPCT(Mδs)θ/2|𝒯|1/2|𝒫|,|I(\mathcal{P},\mathcal{T})|\lessapprox_{\delta}\sqrt{C_{P}C_{T}}\cdot(M\delta^{s})^{\theta/2}\cdot|\mathcal{T}|^{1/2}|\mathcal{P}|,

where θ=θ(s,t)=d1td1s[0,1]\theta=\theta(s,t)=\frac{d-1-t}{d-1-s}\in[0,1]. (If s=t=d1s=t=d-1, then θ(s,t)=0\theta(s,t)=0.)

The following corollary of Proposition 2.9 is the form we will use.

Corollary 2.10.

Let 0std10\leq s\leq t\leq d-1, and let CP,CT1C_{P},C_{T}\geq 1. Let 𝒫𝒟δ\mathcal{P}\subset\mathcal{D}_{\delta} be a (δ,t,CP)(\delta,t,C_{P})-set. Assume that for every p𝒫p\in\mathcal{P} there exists a (δ,s,CT)(\delta,s,C_{T})-family 𝒯(p)𝒯δ\mathcal{T}(p)\subset\mathcal{T}^{\delta} of dyadic δ\delta-tubes with the property that TpT\cap p\neq\emptyset for all T𝒯(p)T\in\mathcal{T}(p), and |𝒯(p)|=M|\mathcal{T}(p)|=M for some M1M\geq 1. If 𝒯=p𝒫𝒯(p)\mathcal{T}=\cup_{p\in\mathcal{P}}\mathcal{T}(p), then

|𝒯|δ(CPCT)1Mδs(Mδs)tsd1s.|\mathcal{T}|\gtrapprox_{\delta}(C_{P}C_{T})^{-1}\cdot M\delta^{-s}\cdot(M\delta^{s})^{\frac{t-s}{d-1-s}}.

(If s=t=d1s=t=d-1, then tsd1s=0\frac{t-s}{d-1-s}=0.)

Remark 2.11.

To use Corollary 2.10, we need td1t\leq d-1. Fortunately, this is a harmless assumption because s<d1s<d-1, and changing tt to min(t,d1)\min(t,d-1) makes the hypothesis of Theorem 1.8 weaker.

Proof.

We begin with an application of Cauchy-Schwarz.

|I(𝒫,𝒯)|\displaystyle|I(\mathcal{P},\mathcal{T})| =T𝒯|{p𝒫:T𝒯(p)}|\displaystyle=\sum_{T\in\mathcal{T}}|\{p\in\mathcal{P}:T\in\mathcal{T}(p)\}|
|𝒯|1/2|{(T,P,P):T𝒯(p)𝒯(p)}|1/2.\displaystyle\leq|\mathcal{T}|^{1/2}\left|\{(T,P,P^{\prime}):T\in\mathcal{T}(p)\cap\mathcal{T}(p^{\prime})\}\right|^{1/2}.

Note that we have the following bounds:

|𝒯(p)𝒯(p)|min{CTM(δd(p,p)+δ)s,(1d(p,p)+δ)d1},|\mathcal{T}(p)\cap\mathcal{T}(p^{\prime})|\lesssim\min\left\{C_{T}\cdot M\cdot\left(\tfrac{\delta}{d(p,p^{\prime})+\delta}\right)^{s},\left(\tfrac{1}{d(p,p^{\prime})+\delta}\right)^{d-1}\right\}, (2.2)

where d(p,p)d(p,p^{\prime}) stands for the distance of the midpoints of pp and pp^{\prime}. To prove (2.2), observe that if T𝒯(p)𝒯(p)T\in\mathcal{T}(p)\cap\mathcal{T}(p^{\prime}), then TT lies in a δd(p,p)+δ\frac{\delta}{d(p,p^{\prime})+\delta}-tube with central line being the line between pp and pp^{\prime}. Thus, the first bound in (2.2) follows from 𝒯(p)\mathcal{T}(p) being a (δ,s,CT)(\delta,s,C_{T})-set with |𝒯(p)|=M|\mathcal{T}(p)|=M, and the second bound is the maximum number of essentially distinct δ\delta-tubes that can fit inside a δd(p,p)+δ\frac{\delta}{d(p,p^{\prime})+\delta}-tube.

Write θ:=θ(s,t):=(d1)t(d1)s[0,1]\theta:=\theta(s,t):=\frac{(d-1)-t}{(d-1)-s}\in[0,1]. (If s=t=d1s=t=d-1, we set θ:=0\theta:=0.) The parameter θ\theta is chosen so that t=sθ+(d1)(1θ)t=s\theta+(d-1)(1-\theta). Then (2.2) and the inequality min{a,b}aθb1θ\min\{a,b\}\leq a^{\theta}b^{1-\theta} imply that

|𝒯(p)𝒯(p)|(CTMδs)θd(p,p)t.|\mathcal{T}(p)\cap\mathcal{T}(p^{\prime})|\lesssim(C_{T}M\delta^{s})^{\theta}\cdot d(p,p^{\prime})^{-t}.

Since 𝒫\mathcal{P} is a (δ,t,CP)(\delta,t,C_{P})-set, for fixed p𝒫p\in\mathcal{P} we have

p(d(p,p)+δ)t2δ2j22tj|{p𝒫:d(p,p)2j}|δCP|𝒫|.\sum_{p^{\prime}}(d(p,p^{\prime})+\delta)^{-t}\lesssim\sum_{\sqrt{2}\cdot\delta\leq 2^{-j}\leq\sqrt{2}}2^{tj}|\{p^{\prime}\in\mathcal{P}:d(p,p^{\prime})\leq 2^{-j}\}|\lessapprox_{\delta}C_{P}\cdot|\mathcal{P}|.

We deduce that

p,p|𝒯(p)𝒯(p)|(CTMδs)θp,p(d(p,p)+δ)tδCP(CTMδs)θ|𝒫|2,\sum_{p,p^{\prime}}|\mathcal{T}(p)\cap\mathcal{T}(p^{\prime})|\lesssim(C_{T}M\delta^{s})^{\theta}\sum_{p,p^{\prime}}(d(p,p^{\prime})+\delta)^{-t}\lessapprox_{\delta}C_{P}(C_{T}M\delta^{s})^{\theta}\cdot|\mathcal{P}|^{2},

so

|I(𝒫,𝒯)|δCP1/2(CTMδs)θ/2|𝒯|1/2|𝒫|CPCT(Mδs)θ/2|𝒯|1/2|𝒫|.|I(\mathcal{P},\mathcal{T})|\lessapprox_{\delta}C_{P}^{1/2}(C_{T}M\delta^{s})^{\theta/2}\cdot|\mathcal{T}|^{1/2}|\mathcal{P}|\leq\sqrt{C_{P}C_{T}}\cdot(M\delta^{s})^{\theta/2}\cdot|\mathcal{T}|^{1/2}|\mathcal{P}|.

This proves Proposition 2.9, and Corollary 2.10 follows by observing |I(𝒫,𝒯)|M|𝒫||I(\mathcal{P},\mathcal{T})|\geq M|\mathcal{P}|. ∎

2.4 Multiscale analysis

Following Section 4 of [18], we would like to change scale from δ\delta to Δ>δ\Delta>\delta, while preserving the properties of 𝒯(p)\mathcal{T}(p). We say AδBA\lessapprox_{\delta}B if there exists an absolute constant C1C\geq 1 such that AC[log(1/δ)]CA\leq C\cdot[\log(1/\delta)]^{C}. We start by naming the objects in Theorem 1.8.

Definition 2.12.

Fix δ2,s[0,d1],C>0,M\delta\in 2^{-\mathbb{N}},s\in[0,d-1],C>0,M\in\mathbb{N}. We say that a pair (𝒫0,𝒯0)𝒟δ×𝒯δ(\mathcal{P}_{0},\mathcal{T}_{0})\subset\mathcal{D}_{\delta}\times\mathcal{T}^{\delta} is a (δ,s,C1,κ,C2,M)(\delta,s,C_{1},\kappa,C_{2},M)-nice configuration if for every p𝒫0p\in\mathcal{P}_{0}, there exists a (δ,s,C1,0)(\delta,s,C_{1},0) and (δ,κ,C2,k)(\delta,\kappa,C_{2},k)-set 𝒯(p)𝒯0\mathcal{T}(p)\subset\mathcal{T}_{0} with |𝒯(p)|=M|\mathcal{T}(p)|=M and such that TpT\cap p\neq\emptyset for all T𝒯(p)T\in\mathcal{T}(p).

Using the method of induction on scales, we would like to relate nice configurations at scale δ\delta to nice configurations at scales Δ,δΔ\Delta,\frac{\delta}{\Delta}, where δ<Δ1\delta<\Delta\leq 1. The following proposition, which combines Propositions 4.1 and 5.2 of [18], gives a way of doing so with only polylog losses. Our proof relies on the same ideas as [18], with some technical simplifications. We defer the proof to Section 6, where we prove a slightly more general version.

Proposition 2.13.

Fix dyadic numbers 0<δ<Δ10<\delta<\Delta\leq 1. Let (𝒫0,𝒯0)(\mathcal{P}_{0},\mathcal{T}_{0}) be a (δ,s,C1,κ,C2,M)(\delta,s,C_{1},\kappa,C_{2},M)-nice configuration. Then there exist sets 𝒫𝒫0\mathcal{P}\subset\mathcal{P}_{0}, 𝒯(p)𝒯0(p),p𝒫\mathcal{T}(p)\subset\mathcal{T}_{0}(p),p\in\mathcal{P}, and 𝒯Δ𝒯Δ\mathcal{T}_{\Delta}\subset\mathcal{T}^{\Delta} such that denoting 𝒯=p𝒫𝒯(p)\mathcal{T}=\cup_{p\in\mathcal{P}}\mathcal{T}(p) the following hold:

  1. (i)

    |𝒟Δ(𝒫)|δ|𝒟Δ(𝒫0)||\mathcal{D}_{\Delta}(\mathcal{P})|\approx_{\delta}|\mathcal{D}_{\Delta}(\mathcal{P}_{0})| and |𝒫Q|δ|𝒫0Q||\mathcal{P}\cap Q|\approx_{\delta}|\mathcal{P}_{0}\cap Q| for all Q𝒟Δ(𝒫)Q\in\mathcal{D}_{\Delta}(\mathcal{P}).

  2. (ii)

    There exists 𝐍\mathbf{N} such that |𝒯𝐓|𝐍|\mathcal{T}\cap\mathbf{T}|\sim\mathbf{N} for all 𝐓𝒯Δ\mathbf{T}\in\mathcal{T}_{\Delta}.

  3. (iii)

    (𝒟Δ(𝒫),𝒯Δ)(\mathcal{D}_{\Delta}(\mathcal{P}),\mathcal{T}_{\Delta}) is (Δ,s,CΔ1,κ,CΔ2,MΔ)(\Delta,s,C^{1}_{\Delta},\kappa,C^{2}_{\Delta},M_{\Delta})-nice for some CΔ1δC1C^{1}_{\Delta}\approx_{\delta}C_{1}, CΔ2δC2C^{2}_{\Delta}\approx_{\delta}C_{2}, and MΔ1M_{\Delta}\geq 1.

  4. (iv)

    For each Q𝒟Δ(𝒫)Q\in\mathcal{D}_{\Delta}(\mathcal{P}), let 𝒯Δ(Q)\mathcal{T}_{\Delta}(Q) be the tubes in 𝒯Δ\mathcal{T}_{\Delta} through QQ. Then for all 𝐓𝒯Δ(Q)\mathbf{T}\in\mathcal{T}_{\Delta}(Q), we have

    |{(p,T)(𝒫Q)×𝒯:T𝒯(p) and T𝐓}|δM|𝒫Q||𝒯Δ(Q)|.|\{(p,T)\in(\mathcal{P}\cap Q)\times\mathcal{T}:T\in\mathcal{T}(p)\text{ and }T\subset\mathbf{T}\}|\gtrapprox_{\delta}\frac{M\cdot|\mathcal{P}\cap Q|}{|\mathcal{T}_{\Delta}(Q)|}.
  5. (v)

    For each Q𝒟Δ(𝒫)Q\in\mathcal{D}_{\Delta}(\mathcal{P}), there exist CQ1δC1C^{1}_{Q}\approx_{\delta}C_{1}, CQ2δC2C^{2}_{Q}\approx_{\delta}C_{2}, MQ1M_{Q}\geq 1, a subset 𝒫Q𝒫Q\mathcal{P}_{Q}\subset\mathcal{P}\cap Q with |𝒫Q|Δ|𝒫Q||\mathcal{P}_{Q}|\gtrapprox_{\Delta}|\mathcal{P}\cap Q|, and a family of tubes 𝒯Q𝒯δ/Δ\mathcal{T}_{Q}\subset\mathcal{T}^{\delta/\Delta} such that (SQ(𝒫Q),𝒯Q)(S_{Q}(\mathcal{P}_{Q}),\mathcal{T}_{Q}) is (δ/Δ,s,CQ1,κ,CQ2,MQ)(\delta/\Delta,s,C^{1}_{Q},\kappa,C^{2}_{Q},M_{Q})-nice.

Furthermore, the families 𝒯Q\mathcal{T}_{Q} can be chosen so that

|𝒯0|Mδ|𝒯Δ|MΔ(maxQ𝒟Δ(𝒫)|𝒯Q|MQ).\frac{|\mathcal{T}_{0}|}{M}\gtrapprox_{\delta}\frac{|\mathcal{T}_{\Delta}|}{M_{\Delta}}\cdot\left(\max_{Q\in\mathcal{D}_{\Delta}(\mathcal{P})}\frac{|\mathcal{T}_{Q}|}{M_{Q}}\right). (2.3)

Iterate this proposition to get (for details, see [27, Corollary 4.1])

Corollary 2.14.

Fix N2N\geq 2 and a sequence {Δj}j=0n2\{\Delta_{j}\}_{j=0}^{n}\subset 2^{-\mathbb{N}} with

0<δ=ΔN<ΔN1<<Δ1<Δ0=1.0<\delta=\Delta_{N}<\Delta_{N-1}<\cdots<\Delta_{1}<\Delta_{0}=1.

Let (𝒫0,𝒯0)𝒟δ×𝒯δ(\mathcal{P}_{0},\mathcal{T}_{0})\subset\mathcal{D}_{\delta}\times\mathcal{T}^{\delta} be a (δ,s,C1,κ,C2,M)(\delta,s,C_{1},\kappa,C_{2},M)-nice configuration. Then there exists a set 𝒫𝒫0\mathcal{P}\subset\mathcal{P}_{0} such that:

  1. 1.

    |𝒟Δj(𝒫)|δ|𝒟Δj(𝒫0)||\mathcal{D}_{\Delta_{j}}(\mathcal{P})|\approx_{\delta}|\mathcal{D}_{\Delta_{j}}(\mathcal{P}_{0})| and |𝒫p|δ|𝒫0p||\mathcal{P}\cap\textbf{p}|\approx_{\delta}|\mathcal{P}_{0}\cap\textbf{p}|, 1jN1\leq j\leq N, p𝒟Δj(𝒫)\textbf{p}\in\mathcal{D}_{\Delta_{j}}(\mathcal{P}).

  2. 2.

    For every 0jN10\leq j\leq N-1 and p𝒟Δj\textbf{p}\in\mathcal{D}_{\Delta_{j}}, there exist numbers Cp1δC1C_{\textbf{p}}^{1}\approx_{\delta}C^{1}, Cp2δC2C_{\textbf{p}}^{2}\approx_{\delta}C^{2}, and Mp1M_{\textbf{p}}\geq 1, and a family of tubes 𝒯p𝒯Δj+1/Δj\mathcal{T}_{\textbf{p}}\subset\mathcal{T}^{\Delta_{j+1}/\Delta_{j}} with the property that (Sp(𝒫p),𝒯p)(S_{\textbf{p}}(\mathcal{P}\cap\textbf{p}),\mathcal{T}_{\textbf{p}}) is a (Δj+1/Δj,s,Cp1,κ,Cp2,Mp)(\Delta_{j+1}/\Delta_{j},s,C_{\textbf{p}}^{1},\kappa,C_{\textbf{p}}^{2},M_{\textbf{p}})-nice configuration.

Furthermore, the families 𝒯p\mathcal{T}_{\textbf{p}} can be chosen such that if pj𝒟Δj(𝒫)\textbf{p}_{j}\in\mathcal{D}_{\Delta_{j}}(\mathcal{P}) for 0jN10\leq j\leq N-1, then

|𝒯0|Mδj=0N1|𝒯pj|Mpj.\frac{|\mathcal{T}_{0}|}{M}\gtrapprox_{\delta}\prod_{j=0}^{N-1}\frac{|\mathcal{T}_{\textbf{p}_{j}}|}{M_{\textbf{p}_{j}}}.

Here, δ\gtrapprox_{\delta} means Nlog(1/δ)C\gtrsim_{N}\log(1/\delta)^{C}, and likewise for δ,δ\lessapprox_{\delta},\approx_{\delta}.

2.5 Uniform sets and branching numbers

The following exposition borrows heavily from [21, Section 2.3].

Definition 2.15.

Let n1n\geq 1 and

δ=Δn<Δn1<<Δ1Δ0=1\delta=\Delta_{n}<\Delta_{n-1}<\cdots<\Delta_{1}\leq\Delta_{0}=1

be a sequence of dyadic scales. We say that a set P[0,1)dP\subset[0,1)^{d} is {Δj}j=1n\{\Delta_{j}\}_{j=1}^{n}-uniform if there is a sequence {Nj}j=1n\{N_{j}\}_{j=1}^{n} such that Nj2N_{j}\in 2^{\mathbb{N}} and |PQ|Δj=Nj|P\cap Q|_{\Delta_{j}}=N_{j} for all j{1,2,,n}j\in\{1,2,\cdots,n\} and Q𝒟Δj1(P)Q\in\mathcal{D}_{\Delta_{j-1}}(P).

Remark 2.16.

By uniformity, we have |P|Δm=|PQ|Δm|P|Δ|P|_{\Delta_{m}}=|P\cap Q|_{\Delta_{m}}|P|_{\Delta_{\ell}} for 0<mn0\leq\ell<m\leq n and Q𝒟Δ(P)Q\in\mathcal{D}_{\Delta_{\ell}}(P).

As a result, we can always refine a set PP to be uniform:

Lemma 2.17.

Let P[0,1)dP\subset[0,1)^{d}, m,Tm,T\in\mathbb{N}, and δ=2mT\delta=2^{-mT}. Let Δj:=2jT\Delta_{j}:=2^{-jT} for 0jm0\leq j\leq m, so in particular δ=Δm\delta=\Delta_{m}. Then there is a {Δj}j=1m\{\Delta_{j}\}_{j=1}^{m}-uniform set PPP^{\prime}\subset P such that

|P|δ(2T)m|P|δ.|P^{\prime}|_{\delta}\geq(2T)^{-m}|P|_{\delta}.

In particular, if ε>0\varepsilon>0 and T1log(2T)εT^{-1}\log(2T)\leq\varepsilon, then |P|δε|P||P^{\prime}|\geq\delta^{\varepsilon}|P|.

Uniform sets can be encoded by a branching function.

Definition 2.18.

Let TT\in\mathbb{N}, and let 𝒫[0,1)d\mathcal{P}\subset[0,1)^{d} be a {Δj}j=1n\{\Delta_{j}\}_{j=1}^{n}-uniform set, with Δj:=2jT\Delta_{j}:=2^{-jT}, and with associated sequence {Nj}j=1n{1,,2dT}n\{N_{j}\}_{j=1}^{n}\subset\{1,\dots,2^{dT}\}^{n}. We define the branching function f:[0,n][0,dn]f:[0,n]\rightarrow[0,dn] by setting f(0)=0f(0)=0, and

f(j):=log|𝒫|2jTT=1Ti=1jlogNi,i{1,n},f(j):=\frac{\log|\mathcal{P}|_{2^{-jT}}}{T}=\frac{1}{T}\sum_{i=1}^{j}\log N_{i},\quad i\in\{1,\dots n\},

and then interpolating linearly between integers.

Definition 2.19.

Let sf(a,b)=f(b)f(a)bas_{f}(a,b)=\frac{f(b)-f(a)}{b-a} denote the slope of a line segment between (a,b)(a,b) and (f(a),f(b))(f(a),f(b)). We say that a function f:[0,n]f:[0,n]\rightarrow\mathbb{R} is ε\varepsilon-superlinear on [a,b][0,n][a,b]\subset[0,n], or that (f,a,b)(f,a,b) is ε\varepsilon-superlinear, if

f(x)f(a)+sf(a,b)(xa)ε(ba),x[a,b].f(x)\geq f(a)+s_{f}(a,b)(x-a)-\varepsilon(b-a),x\in[a,b].

We say that (f,a,b)(f,a,b) is ε\varepsilon-linear if

|f(x)f(a)sf(a,b)(xa)|ε(ba),x[a,b].|f(x)-f(a)-s_{f}(a,b)(x-a)|\leq\varepsilon(b-a),x\in[a,b].

The following lemma converts between branching functions and the uniform structure of PP. It is [18, Lemma 8.3] (or an immediate consequence of the definitions)

Lemma 2.20.

Let PP be a (Δi)i=1m(\Delta^{i})_{i=1}^{m}-uniform set in [0,1)d[0,1)^{d} with associated branching function ff, and let δ=Δm\delta=\Delta^{m}.

  1. (i)

    If ff is ε\varepsilon-superlinear on [0,m][0,m], then PP is a (δ,sf(0,m),OΔ(1)δε)(\delta,s_{f}(0,m),O_{\Delta}(1)\delta^{-\varepsilon})-set.

  2. (ii)

    If ff is ε\varepsilon-linear on [0,m][0,m], then PP is a (δ,sf(0,m),OΔ(1)δε,OΔ(1)δε)(\delta,s_{f}(0,m),O_{\Delta}(1)\delta^{-\varepsilon},O_{\Delta}(1)\delta^{-\varepsilon})-regular set between scales δ\delta and 11.

The crucial branching lemma is [18, Lemma 8.5] applied to the function 2df\frac{2}{d}\cdot f:

Lemma 2.21.

Fix s(0,1)s\in(0,1) and t(s,d]t\in(s,d]. For every ε>0\varepsilon>0 there is τ=τ(ε,s,t)>0\tau=\tau(\varepsilon,s,t)>0 such that the following holds: for every piecewise affine dd-Lipschitz function f:[0,m]\Rf:[0,m]\to\R such that

f(x)txεm for all x[0,m],f(x)\geq tx-\varepsilon m\text{ for all }x\in[0,m],

there exists a family of non-overlapping intervals {[cj,dj]}j=1n\{[c_{j},d_{j}]\}_{j=1}^{n} contained in [0,m][0,m] such that:

  1. 1.

    For each jj, at least one of the following alternatives holds:

    1. (a)

      (f,cj,dj)(f,c_{j},d_{j}) is ε\varepsilon-linear with sf(cj,dj)ss_{f}(c_{j},d_{j})\geq s;

    2. (b)

      (f,cj,dj)(f,c_{j},d_{j}) is ε\varepsilon-superlinear with sf(cj,dj)=ss_{f}(c_{j},d_{j})=s.

  2. 2.

    djcjτmd_{j}-c_{j}\geq\tau m for all jj;

  3. 3.

    |[0,m]j[cj,dj]|s,tεm|[0,m]\setminus\cup_{j}[c_{j},d_{j}]|\lesssim_{s,t}\varepsilon m.

2.6 Combinatorial and probabilistic preliminaries

In this section, we collect a few of the results from additive combinatorics and probability that will be used in the following sections.

First, we make the following observation (Lemma 19 of [10]) about intersections of high-probability events. (That lemma was stated for Lebesgue measure but the same proof works for general measures ν\nu.)

Lemma 2.22.

Let A\RdA\subset\R^{d} equipped with a measure ν\nu and Θ\Theta be an index set equipped with a probability measure μ\mu. Suppose there is K1K\geq 1 and for each θΘ\theta\in\Theta, a Borel subset AθA_{\theta} with ν(Aθ)ν(A)/K\nu(A_{\theta})\geq\nu(A)/K. Then

μq({(θ1,θ2,,θq):ν(Aθ1Aθ2Aθq)ν(A)2Kq})12Kq.\mu^{\otimes q}(\{(\theta_{1},\theta_{2},\cdots,\theta_{q}):\nu(A_{\theta_{1}}\cap A_{\theta_{2}}\cap\cdots\cap A_{\theta_{q}})\geq\frac{\nu(A)}{2K^{q}}\})\geq\frac{1}{2K^{q}}.

Next, we state Rusza’s triangular inequality [10, Lemma 21] (see also [23]):

Lemma 2.23.

For any sets A,B,C\RdA,B,C\subset\R^{d}, we have

|B|δ|AC|δd|AB|δ|BC|δ.|B|_{\delta}|A-C|_{\delta}\lesssim_{d}|A-B|_{\delta}|B-C|_{\delta}.

We also would like the Plünnecke-Rusza inequality, in the form stated by [10, Lemma 22]:

Lemma 2.24.

Let A,BA,B be bounded subsets of \Rd\R^{d}. For all K1K\geq 1, δ>0\delta>0, if |A+B|δK|B|δ|A+B|_{\delta}\leq K|B|_{\delta}, then for all k,1k,\ell\geq 1, we have

|kAA|δdKk+|B|δ.|kA-\ell A|_{\delta}\lesssim_{d}K^{k+\ell}|B|_{\delta}.

Here, kA=A++Ak timeskA=\underbrace{A+\cdots+A}_{k\text{ times}}.

In a similar spirit, the set of ww such that X+wXX+wX is small compared to |X||X| forms a ring. The following is a restatement of [11, Lemma 30(i,ii)] for \R\R. Note that End(\R)\R\mathrm{End}(\R)\simeq\R with identity 11.

Lemma 2.25.

Define Sδ(X;K)={w[K,K]:|X+wX|δK|X|δ}S_{\delta}(X;K)=\{w\in[-K,K]:|X+wX|_{\delta}\leq K|X|_{\delta}\}.

  1. (i)

    If aSδ(X;δε)a\in S_{\delta}(X;\delta^{-\varepsilon}) and b\Rb\in\R such that |ab|δ1ε|a-b|\leq\delta^{1-\varepsilon}, then bSδ(X;δO(ε))b\in S_{\delta}(X;\delta^{-O(\varepsilon)}).

  2. (ii)

    If 1,a,bSδ(X;δε)1,a,b\in S_{\delta}(X;\delta^{-\varepsilon}), then a+b,ab,aba+b,a-b,ab all belong to Sδ(X;δO(ε))S_{\delta}(X;\delta^{-O(\varepsilon)}).

The following theorem (a special case of Theorem 5 of [11]) is a quantitative statement that 12\frac{1}{2}-dimensional subrings of \R\R don’t exist. In fact, by repeated sum-product operations, we can get all of \R\R.

Theorem 2.26.

We work in \R1\R^{1}. Given κ,ε0>0\kappa,\varepsilon_{0}>0, there exist ε>0\varepsilon>0 and an integer s1s\geq 1 such that for δ<δ0(κ,ε0)\delta<\delta_{0}(\kappa,\varepsilon_{0}), the following holds. For every (κ,δε)(\kappa,\delta^{-\varepsilon})-set AB(0,δε)A\subset B(0,\delta^{-\varepsilon}), we have

B(0,δε0)As+B(0,δ),B(0,\delta^{\varepsilon_{0}})\subset\langle A\rangle_{s}+B(0,\delta),

where A1:=A(A)\langle A\rangle_{1}:=A\cup(-A) and for any integer s1s\geq 1, define As+1:=As(As+A1)(AsA1)\langle A\rangle_{s+1}:=\langle A\rangle_{s}\cup(\langle A\rangle_{s}+\langle A\rangle_{1})\cup(\langle A\rangle_{s}\cdot\langle A\rangle_{1}).

Finally, we shall need a discretized variant of the Balog-Szemerédi-Gowers theorem. Our version is closest to [20, Theorem 4.38], which is taken from [1, p. 196], which in turn refers to Exercise 6.4.10 in [31]. But the exercise is only sketched in [31], so for completeness, we provide a proof in Appendix A.

Theorem 2.27.

Let K1K\geq 1 and δ>0\delta>0 be parameters. Let A,BA,B be bounded subsets of \Rd\R^{d}, and let PA×BP\subset A\times B satisfy

|P|δK1|A|δ|B|δ and |{a+b:(a,b)P}|δK(|A|δ|B|δ)1/2.|P|_{\delta}\geq K^{-1}|A|_{\delta}|B|_{\delta}\text{ and }|\{a+b:(a,b)\in P\}|_{\delta}\leq K(|A|_{\delta}|B|_{\delta})^{1/2}.

Then one can find subsets AA,BBA^{\prime}\subset A,B^{\prime}\subset B satisfying

  • |A|δdK2|A|δ,|B|δdK2|B|δ|A^{\prime}|_{\delta}\gtrsim_{d}K^{-2}|A|_{\delta},|B^{\prime}|_{\delta}\gtrsim_{d}K^{-2}|B|_{\delta},

  • |A+B|δdK8(|A|δ|B|δ)1/2|A^{\prime}+B^{\prime}|_{\delta}\lesssim_{d}K^{8}(|A|_{\delta}|B|_{\delta})^{1/2},

  • |P(A×B)|d|A|δ|B|δK2|P\cap(A^{\prime}\times B^{\prime})|\gtrsim_{d}\frac{|A|_{\delta}|B|_{\delta}}{K^{2}}.

(Implicit constants depend on dd but not on δ,K\delta,K.)

We also need the following version of multi-linear Kakeya.

Theorem 2.28 (Theorem 1 in [2]).

Let 2kd2\leq k\leq d and 𝒯1,𝒯2,,𝒯k\mathcal{T}_{1},\mathcal{T}_{2},\cdots,\mathcal{T}_{k} be families of 11-tubes in \Rd\R^{d}. Then

\Rd(T1𝒯1Tk𝒯k|e(T1)e(Tk)|χT1Tk(x))1/(k1)𝑑xk,d(i=1k|𝒯i|)1/(k1).\int_{\R^{d}}\left(\sum_{T_{1}\in\mathcal{T}_{1}}\cdots\sum_{T_{k}\in\mathcal{T}_{k}}|e(T_{1})\wedge\cdots\wedge e(T_{k})|\chi_{T_{1}\cap\cdots\cap T_{k}}(x)\right)^{1/(k-1)}\,dx\lesssim_{k,d}\left(\prod_{i=1}^{k}|\mathcal{T}_{i}|\right)^{1/(k-1)}.

Here, e(Ti)e(T_{i}) is the unit vector in the direction of tube TiT_{i}.

2.7 Energy

Definition 2.29.

The (s,k)(s,k)-Riesz energy of a finite Borel measure μ\mu on \Rd\R^{d} is

Is,kδ(μ)=(|(x0x1)(x0xk)|+δ)s𝑑μ(x0)𝑑μ(xk).I_{s,k}^{\delta}(\mu)=\int(|(x_{0}-x_{1})\wedge\cdots\wedge(x_{0}-x_{k})|+\delta)^{-s}\,d\mu(x_{0})\cdots d\mu(x_{k}).

If k=1k=1 and δ=0\delta=0, we recover the usual ss-dimensional Riesz energy.

Lemma 2.30.
  1. (a)

    Fix 0<s<t0<s<t and a measure μ\mu with total mass CC. If μ(Hr)Crt\mu(H_{r})\leq Cr^{t} for every (r,k1)(r,k-1)-plate HrH_{r} and r>0r>0, then Is,k0(μ)s,tCkI_{s,k}^{0}(\mu)\lesssim_{s,t}C^{k}.

  2. (b)

    Fix 0<δ<120<\delta<\frac{1}{2}. If Isi,kiδ(μ)CI_{s_{i},k_{i}}^{\delta}(\mu)\leq C for 1im1\leq i\leq m, then spt(μ)\mathrm{spt}(\mu) contains a set which is simultaneously a (δ,siki,O(1)(Cm)1/kilogδ1,ki1)(\delta,\frac{s_{i}}{k_{i}},O(1)\cdot(Cm)^{1/k_{i}}\log\delta^{-1},k_{i}-1)-set for each ii.

Remark 2.31.

If k1=m=1k_{1}=m=1 in part (b), then we can drop the log factor (c.f. proof of Lemma A.6 in [18]). We don’t know if we can drop the log factor for k>1k>1 or m>1m>1.

Proof.

(a) Let ρi\rho_{i} be the distance between xix_{i} and the plane spanned by x0,,xi1x_{0},\cdots,x_{i-1}; notice that |(x0x1)(x0xk)|=i=1kρi|(x_{0}-x_{1})\wedge\cdots\wedge(x_{0}-x_{k})|=\prod_{i=1}^{k}\rho_{i}. Thus, we can rewrite Is,k(μ)I_{s,k}(\mu) as an iterated integral

𝑑μ(x0)ρ1s𝑑μ(x1)ρ2s𝑑μ(x2)ρks𝑑μ(xk).\int d\mu(x_{0})\int\rho_{1}^{-s}d\mu(x_{1})\int\rho_{2}^{-s}\,d\mu(x_{2})\cdots\int\rho_{k}^{-s}\,d\mu(x_{k}).

We will be done if we show for all 1ik1\leq i\leq k and choices of x0,,xi1x_{0},\cdots,x_{i-1}, that ρis𝑑μ(xi)C\int\rho_{i}^{-s}\,d\mu(x_{i})\lesssim C. Let HH be the span of x0x_{0} through xi1x_{i-1}, and observe that by definition, {xi:ρir}H(r)\{x_{i}:\rho_{i}\geq r\}\subset H^{(r)}, which is contained in a (r,k1)(r,k-1)-plate. Thus,

ρis𝑑μ(xi)C+ρ=2n,n1Cρtss,tC.\int\rho_{i}^{-s}\,d\mu(x_{i})\lesssim C+\sum_{\rho=2^{-n},n\geq 1}C\rho^{t-s}\lesssim_{s,t}C.

(b) Let Pi={x0spt(μ):(|(x0x1)(x0xki)|+δ)si𝑑μ(x1)𝑑μ(xki)<2mC}P_{i}=\{x_{0}\in\mathrm{spt}(\mu):\int(|(x_{0}-x_{1})\wedge\cdots\wedge(x_{0}-x_{k_{i}})|+\delta)^{-s_{i}}\,d\mu(x_{1})\cdots d\mu(x_{k_{i}})<2mC\}. By Markov’s inequality, μ(Pi)>112m\mu(P_{i})>1-\frac{1}{2m}, so by the union bound, P=i=1mPiP=\cap_{i=1}^{m}P_{i} satisfies μ(P)>12\mu(P)>\frac{1}{2}.

We claim that μ(PHr)Crsi/ki\mu(P\cap H_{r})\leq Cr^{s_{i}/k_{i}} for all (r,ki1)(r,k_{i}-1)-plates HrH_{r} and δr1\delta\leq r\leq 1, 1im1\leq i\leq m. Indeed, if PHr=P\cap H_{r}=\emptyset, then we are done. Otherwise, pick x0PHrx_{0}\in P\cap H_{r} and observe that if x1,x2,,xkiHrx_{1},x_{2},\cdots,x_{k_{i}}\in H_{r}, then |(x0x1)(x0xki)|+δr|(x_{0}-x_{1})\wedge\cdots\wedge(x_{0}-x_{k_{i}})|+\delta\lesssim r. Thus, we get μ(Hr)kirs2C\mu(H_{r})^{k_{i}}\cdot r^{-s}\leq 2C, so μ(PHr)(2Crs)1/ki\mu(P\cap H_{r})\leq(2Cr^{s})^{1/k_{i}}.

Finally, let Pc𝒟δ(P)P^{\prime}_{c}\subset\mathcal{D}_{\delta}(P) be those dyadic δ\delta-cubes pp such that μ(p)c\mu(p)\sim c. We know c=2n[δd,1]μ(Pc)14\sum_{c=2^{-n}\in[\delta^{d},1]}\mu(P^{\prime}_{c})\geq\frac{1}{4}, so by dyadic pigeonholing, some μ(Pc)(logδ1)1\mu(P^{\prime}_{c})\gtrsim(\log\delta^{-1})^{-1}. Then PcP^{\prime}_{c} will be a (δ,siki,O(1)(Cm)1/kilogδ1,ki1)(\delta,\frac{s_{i}}{k_{i}},O(1)\cdot(Cm)^{1/k_{i}}\log\delta^{-1},k_{i}-1)-set for all 1im1\leq i\leq m. ∎

3 Improved incidence estimates for quasi-product sets

The main novelty of this paper is the following Proposition, which is a higher-dimensional refinement of [20, Proposition 4.36] (see also [18, Proposition A.7]). It can be viewed as a variant of Theorem 1.8 for quasi-product sets.

Proposition 3.1.

Given 0k<d10\leq k<d-1, 0s<k+10\leq s<k+1, τ,κ>0\tau,\kappa>0, there exist η(s,k,κ,τ,d)>0\eta(s,k,\kappa,\tau,d)>0 and δ0(s,k,κ,τ,d)>0\delta_{0}(s,k,\kappa,\tau,d)>0 such that the following holds for all δ(0,δ0]\delta\in(0,\delta_{0}].

Let 𝐘(δ)[0,1)\mathbf{Y}\subset(\delta\cdot\mathbb{Z})\cap[0,1) be a (δ,τ,δη)(\delta,\tau,\delta^{-\eta})-set, and for each 𝐲𝐘\mathbf{y}\in\mathbf{Y}, assume that 𝐗𝐲(δ)d1[0,1)d1\mathbf{X}_{\mathbf{y}}\subset(\delta\cdot\mathbb{Z})^{d-1}\cap[0,1)^{d-1} is a (δ,κ,δη,k)(\delta,\kappa,\delta^{-\eta},k)-set with cardinality δs+η\geq\delta^{-s+\eta}. Let

𝐙=𝐲𝐘𝐗𝐲×{𝐲}.\mathbf{Z}=\bigcup_{\mathbf{y}\in\mathbf{Y}}\mathbf{X}_{\mathbf{y}}\times\{\mathbf{y}\}.

For every 𝐳𝐙\mathbf{z}\in\mathbf{Z}, assume that 𝒯(𝐳)\mathcal{T}(\mathbf{z}) is a set of δ\delta-tubes each making an angle 1100\geq\frac{1}{100} with the plane y=0y=0 with |𝒯(𝐳)|δs+η|\mathcal{T}(\mathbf{z})|\geq\delta^{-s+\eta} such that 𝐳T\mathbf{z}\in T for all T𝒯(𝐳)T\in\mathcal{T}(\mathbf{z}). Then |𝒯|δ2sη|\mathcal{T}|\geq\delta^{-2s-\eta}, where 𝒯=𝐳𝐙𝒯(𝐳)\mathcal{T}=\cup_{\mathbf{z}\in\mathbf{Z}}\mathcal{T}(\mathbf{z}).

Remark 3.2.

In contrast to Theorem 1.8 and [20, Proposition 4.36], we (perhaps surprisingly) don’t need any non-concentration assumptions on the tube sets 𝒯(𝐳)\mathcal{T}(\mathbf{z}) (even when d=2d=2). Instead, it suffices to have weak non-concentration assumptions on 𝐘\mathbf{Y} and 𝐗𝐲\mathbf{X}_{\mathbf{y}} for each 𝐲𝐘\mathbf{y}\in\mathbf{Y}. The non-concentration assumption on 𝐗𝐲\mathbf{X}_{\mathbf{y}} is necessary: otherwise, we can take s=ks=k, and let 𝐙\mathbf{Z} to be the δ\delta-balls contained in some (δ,k+1)(\delta,k+1)-plate HH, and 𝒯\mathcal{T} to be the δ\delta-tubes contained in HH.

3.1 An improved slicing estimate

We will eventually deduce Proposition 3.1 from the following slicing estimate.

Theorem 3.3.

For 0kd20\leq k\leq d-2, 0s<k+10\leq s<k+1, and 0<κ10<\kappa\leq 1, there exists ε>0\varepsilon>0 such that the following holds for sufficiently small δ<δ0(s,k,d,ε)\delta<\delta_{0}(s,k,d,\varepsilon). Let 𝒯\mathcal{T} be a (δ,κ,δε,k)(\delta,\kappa,\delta^{-\varepsilon},k)-set of δ\delta-tubes each making angle 1100\geq\frac{1}{100} with the plane y=0y=0 with |𝒯|δ2s+ε|\mathcal{T}|\geq\delta^{-2s+\varepsilon}. Let μ\mu be a probability measure on \R\R such that for all δr1\delta\leq r\leq 1, we have μ(Br)δεrκ\mu(B_{r})\leq\delta^{-\varepsilon}r^{\kappa}. Then there is a set 𝒟\R\mathcal{D}\subset\R with μ(𝒟)1δε\mu(\mathcal{D})\geq 1-\delta^{\varepsilon} such that the slice of 𝒯\cup\mathcal{T}^{\prime} at z=z0z=z_{0} has δ\delta-covering number δsε\geq\delta^{-s-\varepsilon}, for every subset 𝒯𝒯\mathcal{T}^{\prime}\subset\mathcal{T} with |𝒯|δε|𝒯||\mathcal{T}^{\prime}|\geq\delta^{\varepsilon}|\mathcal{T}| and x𝒟x\in\mathcal{D}.

Remark 3.4.

One should compare Theorem 3.3 to [10, Theorem 1]. Indeed, if k=0k=0 and d=2d=2, Theorem 3.3 is a direct corollary of [10, Theorem 1]. We can see this by using ball-tube duality, which turns 𝒯\mathcal{T} into a subset of \R2\R^{2}. Under this duality, the slice of 𝒯\cup\mathcal{T}^{\prime} at z=z0z=z_{0} becomes the orthogonal projection πz~0\pi_{\tilde{z}_{0}} to a line in the dual space, for some z~0S1\tilde{z}_{0}\in S^{1}. The map z0z~0z_{0}\to\tilde{z}_{0} induces a pushforward measure μ~\tilde{\mu} of μ\mu which still satisfies the non-concentration condition μ~(Br)δεrκ\tilde{\mu}(B_{r})\lesssim\delta^{-\varepsilon}r^{\kappa}, so we can apply [10, Theorem 1]. (For more details, see the proof of Proposition A.7 in [18].)

In higher dimensions, we can still use duality to turn 𝒯\mathcal{T} into a subset of 𝔸(d,1)\R2(d1)\mathbb{A}(d,1)\sim\R^{2(d-1)}, and then slices of 𝒯\cup\mathcal{T}^{\prime} become orthogonal projections to (d1)(d-1)-planes. Unfortunately, [10, Theorem 1] does not apply because the pushforward measure μ~\tilde{\mu} is still supported on a line in Sd1S^{d-1}. This approach is bound to fail because [10, Theorem 1] does not use the strong assumption that 𝒯\mathcal{T} is non-concentrated around (k+1)(k+1)-planes. Using this assumption is the key novelty of this proof.

Nonetheless, Theorem 3.3 will borrow many ideas from the proof of [10, Theorem 1] and He’s previous work [11]. Roughly, the strategy is as follows.

  • As in [10], reduce to the following slightly weaker statement: given 𝒯\mathcal{T} and μ\mu, we can find a subset 𝒯𝒯\mathcal{T}^{\prime}\subset\mathcal{T} such that the conclusion of Theorem 3.3 holds for 𝒯\mathcal{T}^{\prime} in place of 𝒯\mathcal{T}. This relies on a formal exhaustion argument.

  • Then, as in [10], reduce this slightly weaker to the following even weaker statement: there exists z0E:=sptμz_{0}\in E:=\mathrm{spt}\mu such that the slice of 𝒯\cup\mathcal{T} at z=z0z=z_{0} has δ\delta-covering number δsε\geq\delta^{-s-\varepsilon}. This relies on additive combinatorics (e.g. the Balog-Szemerédi-Gowers theorem) and some probability.

  • Assume this is false: that for all z0Ez_{0}\in E, the slice of 𝒯\cup\mathcal{T} at z=z0z=z_{0} has δ\delta-covering number δs\lessapprox\delta^{-s}. Using additive combinatorics as in [11], the same conclusion is true for all z0Ez_{0}\in E^{\prime}, which is the set of sums or differences of mm many terms, each of which is a product of mm elements of EE. (Here, mm will be a fixed large integer.)

  • Finally, if mm is sufficiently large in terms of κ,ε\kappa,\varepsilon, then EE^{\prime} contains a large interval [0,δε][0,\delta^{\varepsilon}] (c.f. [11, Theorem 5]). Essentially, we have a set of δ2s\gtrapprox\delta^{-2s} many tubes 𝒯\mathcal{T}, each containing δ1\gtrapprox\delta^{-1} many δ\delta-balls, such that the union of the δ\delta-balls has cardinality δ(s+1)\lessapprox\delta^{-(s+1)}. Without further restrictions, this Furstenberg-type problem doesn’t lead to a contradiction: take s=ks=k and 𝒯\mathcal{T} to be the set of δ\delta-tubes in a (δ,k+1)(\delta,k+1)-plate. Luckily, our set of tubes 𝒯\mathcal{T} is still a (δ,κ,δO(ε),k)(\delta,\kappa,\delta^{-O(\varepsilon)},k)-set, which rules out this counterexample. Indeed, we may finish using multi-linear Kakeya.

The reader be warned: we shall execute this strategy in reverse order. This is mainly because the main innovation of the paper is the fourth bullet point.

3.2 An improved Furstenberg estimate

The following estimate complements work of Zahl [34]: we prove an ε\varepsilon-improvement on the union of tubes under a mild (k+1)(k+1)-plane non-concentration for the set of tubes. As in Zahl [34], the key technique is multilinear Kakeya.

Theorem 3.5.

For any 0k<d10\leq k<d-1, 0s<k+10\leq s<k+1, 0<κ10<\kappa\leq 1, there exists ε>0\varepsilon>0 such that the following holds for sufficiently small δ>0\delta>0. Let 𝒯\mathcal{T} be a (δ,κ,δε,k)(\delta,\kappa,\delta^{-\varepsilon},k)-set of δ\delta-tubes with |𝒯|δ2s+ε|\mathcal{T}|\geq\delta^{-2s+\varepsilon}, and for each t𝒯t\in\mathcal{T}, let PtP_{t} be a set of δ\delta-balls intersecting tt such that |Pt|δ1+ε|P_{t}|\geq\delta^{-1+\varepsilon}. Then |Pt|δ(s+1)ε|\cup P_{t}|\gtrsim\delta^{-(s+1)-\varepsilon}.

Proof.

The proof below is lossy and can possibly be improved (say by induction on scale). Also, the ε\varepsilon can be determined explicitly in terms of the parameters but we choose not to do so here.

We use \gtrapprox notation to hide δCε\delta^{-C\varepsilon} terms, where CC can depend on the other parameters. Let P=PtP=\cup P_{t}, and suppose |P|δ(s+1)|P|\lessapprox\delta^{-(s+1)}. Let 𝒯(p)\mathcal{T}(p) be the set of tubes in 𝒯\mathcal{T} through pp. Use a bush argument to upper bound |𝒯(p)||\mathcal{T}(p)|:

δ(s+1)|P||tp(PtB(p,δ2ε))|δ2dεtp|PtB(p,δ2ε)|δ1+(2d+1)ε|𝒯(p)|.\delta^{-(s+1)}\gtrapprox|P|\geq|\cup_{t\ni p}(P_{t}\setminus B(p,\delta^{2\varepsilon}))|\geq\delta^{2d\varepsilon}\sum_{t\ni p}|P_{t}\setminus B(p,\delta^{2\varepsilon})|\geq\delta^{-1+(2d+1)\varepsilon}|\mathcal{T}(p)|.

Thus, |𝒯(p)|δs|\mathcal{T}(p)|\lessapprox\delta^{-s} for all pPp\in P. We get the following inequality chain

δ2s1δ1|𝒯|I(P,𝒯)δs|P|δ2s1.\delta^{-2s-1}\lessapprox\delta^{-1}|\mathcal{T}|\lessapprox I(P,\mathcal{T})\lessapprox\delta^{-s}|P|\lessapprox\delta^{-2s-1.}

This means I(P,𝒯)δ2s1I(P,\mathcal{T})\approx\delta^{-2s-1}, |P|δs1|P|\approx\delta^{-s-1}, and |𝒯|δ2s|\mathcal{T}|\approx\delta^{-2s}. Now perform a dyadic pigeonholing to extract a subset PPP^{\prime}\subset P such that |𝒯(p)|[M,2M]|\mathcal{T}(p)|\in[M,2M] for all pPp\in P^{\prime} and I(P,𝒯)δ2s1I(P^{\prime},\mathcal{T})\approx\delta^{-2s-1}. We know from before that MδsM\leq\delta^{-s}, and δ2s1I(P,𝒯)M|P|M|P|δ2s1\delta^{-2s-1}\lessapprox I(P^{\prime},\mathcal{T})\approx M|P^{\prime}|\lessapprox M|P|\lessapprox\delta^{-2s-1}, so MδsM\approx\delta^{-s} and |P|δs|P^{\prime}|\approx\delta^{-s}. (This type of dyadic pigeonholing will also be used later. We also remark that dyadic pigeonholing was not necessary to achieve this step; simply let PP^{\prime} be the set of pPp\in P satisfying |𝒯(p)|δs+Cη|\mathcal{T}(p)|\geq\delta^{-s+C\eta} for some large CC, and use the bound on I(P,𝒯)I(P,\mathcal{T}) to get a lower bound for |P||P^{\prime}|.)

Now, we claim that PP^{\prime} is a (δ,κ,δO(ε),k+1)(\delta,\kappa,\delta^{-O(\varepsilon)},k+1)-set. Fix δ<r<1\delta<r<1 and let HrH_{r} be a (r,k+1)(r,k+1)-plate. We first bound I(PHr,𝒯)I(P^{\prime}\cap H_{r},\mathcal{T}). Letting HrH_{r^{\prime}} be the (r,k+1)(r^{\prime},k+1)-plate that is a dilate of HrH_{r} with the same center, we have

I(PHr,𝒯)\displaystyle I(P^{\prime}\cap H_{r},\mathcal{T}) r=2[r,1]I(PHr,HrHr/2)\displaystyle\leq\sum_{r^{\prime}=2^{-\mathbb{N}}\cap[r,1]}I(P^{\prime}\cap H_{r},H_{r^{\prime}}\setminus H_{r^{\prime}/2})
rrrδ|𝒯Hr|\displaystyle\leq\sum_{r^{\prime}}\frac{r}{r^{\prime}\delta}\cdot|\mathcal{T}\cap H_{r^{\prime}}|
rrrδ|𝒯|δε(r)κ\displaystyle\leq\sum_{r^{\prime}}\frac{r}{r^{\prime}\delta}\cdot|\mathcal{T}|\delta^{-\varepsilon}(r^{\prime})^{\kappa}
δ1|𝒯|δεrκ.\displaystyle\lesssim\delta^{-1}|\mathcal{T}|\delta^{-\varepsilon}r^{\kappa}.

Thus, since I(PHr,𝒯)δs|PHr|I(P^{\prime}\cap H_{r},\mathcal{T})\gtrapprox\delta^{-s}|P^{\prime}\cap H_{r}|, we have |PHr|δs1|𝒯|rκδ(s+1)rκ|P|rκ|P^{\prime}\cap H_{r}|\lessapprox\delta^{s-1}|\mathcal{T}|r^{\kappa}\approx\delta^{-(s+1)}r^{\kappa}\approx|P^{\prime}|r^{\kappa}.

Finally, since I(P,𝒯)|P|δsδ2s1I(P^{\prime},\mathcal{T})\gtrapprox|P^{\prime}|\delta^{-s}\gtrapprox\delta^{-2s-1} and |𝒯|δ2s|\mathcal{T}|\lessapprox\delta^{-2s}, by dyadic pigeonholing there exists a subset 𝒯𝒯\mathcal{T}^{\prime}\subset\mathcal{T} with |𝒯||𝒯||\mathcal{T}^{\prime}|\approx|\mathcal{T}| such that each t𝒯t\in\mathcal{T}^{\prime} contains δ1\approx\delta^{-1} many δ\delta-balls in PP^{\prime}. Now since I(P,𝒯)δ1|𝒯|δ2s1I(P^{\prime},\mathcal{T}^{\prime})\gtrapprox\delta^{-1}|\mathcal{T}^{\prime}|\gtrapprox\delta^{-2s-1}, |P|δs1|P^{\prime}|\lessapprox\delta^{-s-1}, and |𝒯(p)|δs|\mathcal{T}(p)|\lessapprox\delta^{-s} for all pPp\in P^{\prime}, by dyadic pigeonholing we can find P~P\tilde{P}\subset P^{\prime} with |P~||P||\tilde{P}|\gtrapprox|P^{\prime}| such that each pP~p\in\tilde{P} lies in δs\approx\delta^{-s} many tubes in 𝒯\mathcal{T}^{\prime}.

Now, we are in good shape to apply multilinear Kakeya. For pP~p\in\tilde{P}, let 𝒯(p)\mathcal{T}(p) be the tubes in 𝒯\mathcal{T}^{\prime} through pp. By a bush argument, 𝒯(p)\cup\mathcal{T}(p) contains δ(s+1)\gtrapprox\delta^{-(s+1)} many δ\delta-balls in PP. Since PP^{\prime} is a (δ,κ,δO(ε),k+1)(\delta,\kappa,\delta^{-O(\varepsilon)},k+1)-set, there are δ(s+1)(k+3)\gtrapprox\delta^{-(s+1)(k+3)} many (k+3)(k+3)-tuples of points (p0,p1,,pk+2)(p_{0},p_{1},\cdots,p_{k+2}) such that p0p_{0} and pip_{i} lie on some tube ti𝒯it_{i}\in\mathcal{T}_{i} and |e(t1)e(tk+2)|1|e(t_{1})\wedge\cdots\wedge e(t_{k+2})|\gtrapprox 1 (where e(t)e(t) is the unit vector in the direction of tube tt). Thus, there is a choice of p1,,pk+2p_{1},\cdots,p_{k+2} such that there are δ(s+1)\gtrapprox\delta^{-(s+1)} many valid choices for p0p_{0}. But this leads to a contradiction by the following argument. Let 𝒯i\mathcal{T}_{i} be the tubes of 𝒯\mathcal{T} through pip_{i}, 1ik+21\leq i\leq k+2; then by a rescaled version of Multilinear Kakeya (Theorem 2.28), the number of valid choices for p0p_{0} is (i=1k+2|𝒯i|)1/(k+1)δs(k+2)/(k+1)\lessapprox\left(\prod_{i=1}^{k+2}|\mathcal{T}_{i}|\right)^{1/(k+1)}\lessapprox\delta^{-s(k+2)/(k+1)}, which (using s<k+1s<k+1) is much smaller than δ(s+1)\delta^{-(s+1)} provided that ε,δ\varepsilon,\delta are sufficiently small in terms of the parameters. This contradiction completes the proof. ∎

3.3 From Furstenberg to weak slicing

This subsection contains ideas from [11] and [10].

Theorem 3.6.

For 0kd20\leq k\leq d-2, 0s<k+10\leq s<k+1, and 0<κ10<\kappa\leq 1, there exists ε>0\varepsilon>0 such that the following holds for sufficiently small δ<δ0(s,k,d,ε)\delta<\delta_{0}(s,k,d,\varepsilon). Let 𝒯\mathcal{T} be a (δ,κ,δε,k)(\delta,\kappa,\delta^{-\varepsilon},k)-set of δ\delta-tubes each making angle 1100\geq\frac{1}{100} with the plane y=0y=0 with |𝒯|δ2s+ε|\mathcal{T}|\geq\delta^{-2s+\varepsilon}. Let μ\mu be a probability measure on \R\R such that for all δr1\delta\leq r\leq 1, we have μ(Br)δεrκ\mu(B_{r})\leq\delta^{-\varepsilon}r^{\kappa}. Then there exists z0sptμz_{0}\in\mathrm{spt}\mu such that the slice of 𝒯\cup\mathcal{T} at z=z0z=z_{0} has δ\delta-covering number δsε\geq\delta^{-s-\varepsilon}.

Proof.

We use \lessapprox to denote CδCε\leq C\delta^{-C\varepsilon}, where CC may depend on κ,s\kappa,s.

Let E:=sptμE:=\mathrm{spt}\mu; without loss of generality, assume EE is closed. Let z1=infEz_{1}=\inf E and z2=supEz_{2}=\sup E; then d(z1,z2)δ2ε/κd(z_{1},z_{2})\geq\delta^{2\varepsilon/\kappa} since μ(B(z1,δ2ε/κ))δεδ2εδε<1\mu(B(z_{1},\delta^{2\varepsilon/\kappa}))\leq\delta^{-\varepsilon}\cdot\delta^{2\varepsilon}\leq\delta^{\varepsilon}<1.

Let X=𝒯(z=z1)X=\cup\mathcal{T}(z=z_{1}) and Y=𝒯(z=z2)Y=\cup\mathcal{T}(z=z_{2}); we are given that |X|δ,|Y|δδs|X|_{\delta},|Y|_{\delta}\lessapprox\delta^{-s}. On the other hand, since each T𝒯T\in\mathcal{T} passes through O(1)O(1) many elements in XX and O(1)O(1) many elements in YY, we get that

δ2s|X|δ|Y|δ|𝒯|δ2s,\delta^{-2s}\gtrapprox|X|_{\delta}|Y|_{\delta}\gtrsim|\mathcal{T}|\gtrapprox\delta^{-2s},

so in fact, |X|,|Y|δs|X|,|Y|\approx\delta^{-s} and |𝒯|δ2s|\mathcal{T}|\approx\delta^{-2s}.

Let E:=[z1,z2](B(z1,δ2ε/κ)B(z2,δ2ε/κ))E^{\prime}:=[z_{1},z_{2}]\setminus(B(z_{1},\delta^{2\varepsilon/\kappa})\cup B(z_{2},\delta^{2\varepsilon/\kappa})); then μ(E)12δε12\mu(E^{\prime})\geq 1-2\delta^{\varepsilon}\geq\frac{1}{2} for δ\delta small enough.

Let f(z)=zz1z2zf(z)=\frac{z-z_{1}}{z_{2}-z}; note that on EE^{\prime}, we have that ff is 1\approx 1-bilipschitz, and f(z)1f(z)\approx 1 for all zEz\in E^{\prime}.

The problem condition literally states |(z2z)X+(zz1)Y|δδs|(z_{2}-z)X+(z-z_{1})Y|_{\delta}\lessapprox\delta^{-s} for zEz\in E^{\prime}; since (z2z)1(z_{2}-z)\approx 1, we can divide through by (z2z)(z_{2}-z) to get

|X+f(z)Y|δδs for zE.|X+f(z)Y|_{\delta}\lessapprox\delta^{-s}\text{ for }z\in E^{\prime}.

Now pick an arbitrary zEz^{\prime}\in E^{\prime}. In particular, we get |X+f(z)Y|δδs|X+f(z^{\prime})Y|_{\delta}\lessapprox\delta^{-s}, so by Lemma 2.23, we have for all zEz\in E^{\prime},

|Xf(z)f(z)X|δ|X+f(z)Y|δ|f(z)Y+f(z)f(z)X|δ|f(z)Y|δδs.|X-\frac{f(z)}{f(z^{\prime})}X|_{\delta}\leq\frac{|X+f(z)Y|_{\delta}|f(z)Y+\frac{f(z)}{f(z^{\prime})}X|_{\delta}}{|f(z)Y|_{\delta}}\lessapprox\delta^{-s}.

In addition, since |X+f(z)Y|δδs|X|δ|X+f(z^{\prime})Y|_{\delta}\lessapprox\delta^{-s}\lessapprox|X|_{\delta}, the Plünnecke-Rusza inequality (Lemma 2.24) gives |X+X|δ|Y|δδs|X+X|_{\delta}\lessapprox|Y|_{\delta}\lessapprox\delta^{-s}.

Define μ~=g(μ)\tilde{\mu}=g_{*}(\mu), the pushforward of g(z)=f(z)f(z)g(z)=\frac{f(z)}{f(z^{\prime})}; then gg (like ff) is 1\gtrapprox 1-bilipschitz on EE^{\prime}, so μ~\tilde{\mu} also satisfies a non-concentration condition μ~(Br)rκ\tilde{\mu}(B_{r})\lessapprox r^{\kappa} for δr1\delta\leq r\leq 1. Now pick ε0>0\varepsilon_{0}>0, and assume ε\varepsilon is chosen sufficiently small in terms of ε0\varepsilon_{0}. By the iterated sum-product Theorem 2.26, we can find an integer m1m\geq 1 such that for δ<δ0(κ,ε,ε0)\delta<\delta_{0}(\kappa,\varepsilon,\varepsilon_{0}),

B(0,δε0)Am+B(0,δ),B(0,\delta^{\varepsilon_{0}})\subset\langle A\rangle_{m}+B(0,\delta),

where A1:=A(A)\langle A\rangle_{1}:=A\cup(-A) and for any integer m1m\geq 1, define Am+1:=Am(Am+A1)(AmA1)\langle A\rangle_{m+1}:=\langle A\rangle_{m}\cup(\langle A\rangle_{m}+\langle A\rangle_{1})\cup(\langle A\rangle_{m}\cdot\langle A\rangle_{1}).

By applying the ring structure Lemma 2.25 many times, we see that B(0,δε0)Sδ(X;δOm(ε))B(0,\delta^{\varepsilon_{0}})\subset S_{\delta}(X;\delta^{-O_{m}(\varepsilon)}) and since 1Sδ(X;δO(ε))1\in S_{\delta}(X;\delta^{-O(\varepsilon)}), that B(1,δε0)Sδ(X;δOm(ε))B(1,\delta^{\varepsilon_{0}})\subset S_{\delta}(X;\delta^{-O_{m}(\varepsilon)}). By definition of SδS_{\delta} and Lemma 2.23, we get for wB(1,δε0)w\in B(1,\delta^{\varepsilon_{0}}),

|X+wf(z)Y|δ|XwX|δ|wX+wf(z)Y|δ|wX|δδs.|X+wf(z^{\prime})Y|_{\delta}\leq\frac{|X-wX|_{\delta}|wX+wf(z^{\prime})Y|_{\delta}}{|wX|_{\delta}}\lessapprox\delta^{-s}.

In other words, for all z0I:=f1([(1δε0)f(z),(1+δε0)f(z)])z_{0}\in I:=f^{-1}([(1-\delta^{\varepsilon_{0}})f(z^{\prime}),(1+\delta^{\varepsilon_{0}})f(z^{\prime})]), the slice 𝒯(z=z0)\cup\mathcal{T}(z=z_{0}) has δ\delta-covering number δs\lessapprox\delta^{-s}. Since ff is 1\approx 1-bilipschitz and f(z)1f(z^{\prime})\approx 1, we have |I|δε0|I|\approx\delta^{\varepsilon_{0}}.

Now, we seek a contradiction to Theorem 3.5. For every t𝒯t\in\mathcal{T}, we let PtP_{t} be the δ\delta-balls on tt with zz-coordinate in II. We observe the following:

  • Recall our assumption that 𝒯\mathcal{T} is a (δ,κ,δε,k)(\delta,\kappa,\delta^{-\varepsilon},k)-set of δ\delta-tubes with |𝒯|δ2s+ε|\mathcal{T}|\geq\delta^{-2s+\varepsilon}.

  • |Pt|δ1|I|δ1+ε0|P_{t}|\gtrsim\delta^{-1}|I|\gtrapprox\delta^{-1+\varepsilon_{0}}.

  • |Pt||zI(X+f(z)Y)|δ(s+1)|\cup P_{t}|\leq|\cup_{z\in I}(X_{*}+f(z)Y_{*})|\lessapprox\delta^{-(s+1)}.

Thus, if ε,ε0\varepsilon,\varepsilon_{0} are sufficiently small in terms of s,k,κs,k,\kappa, then we contradict Theorem 3.5.

3.4 An intermediate slicing result

This subsection contains ideas from [10].

Let (𝒯,ε)\mathcal{E}(\mathcal{T},\varepsilon) be the set of exceptional slices,

(𝒯,ε)={z0\R:𝒯𝒯,|𝒯|δε|𝒯|,|𝒯(z=z0)|<δsε}.\mathcal{E}(\mathcal{T},\varepsilon)=\{z_{0}\in\R:\exists\mathcal{T}^{\prime}\subset\mathcal{T},|\mathcal{T}^{\prime}|\geq\delta^{\varepsilon}|\mathcal{T}|,|\cup\mathcal{T}^{\prime}(z=z_{0})|<\delta^{-s-\varepsilon}\}.

Just like in [10, Proposition 25], we will prove a weaker version of Theorem 3.3; the stronger version follows from a formal exhaustion argument which we present in the next subsection.

Theorem 3.7.

With assumptions of Theorem 3.3, there exists 𝒯𝒯\mathcal{T}^{\prime}\subset\mathcal{T} such that μ((𝒯))δε\mu(\mathcal{E}(\mathcal{T}^{\prime}))\leq\delta^{\varepsilon}.

Proof.

We use \lessapprox to denote CδCε\leq C\delta^{-C\varepsilon}, where CC may depend on κ,s\kappa,s. Let π:\Rd\Rd1\pi:\R^{d}\to\R^{d-1} be the projection onto the plane orthogonal to the zz-axis. For a tube tt, let t(z)=π(t{z=z})t(z^{\prime})=\pi(t\cap\{z=z^{\prime}\}), and for a set of tubes 𝒯\mathcal{T}, let 𝒯(z)\mathcal{T}(z^{\prime}) denote the slice π(𝒯{z=z})\pi(\mathcal{T}\cap\{z=z^{\prime}\}).

We follow the argument in [10, Proof of Proposition 7]. Suppose Theorem 3.7 is false. We can find z1z_{1} and a subset 𝒯′′′𝒯\mathcal{T}^{\prime\prime\prime}\subset\mathcal{T} with |𝒯′′′|δε|𝒯||\mathcal{T}^{\prime\prime\prime}|\geq\delta^{\varepsilon}|\mathcal{T}| such that |𝒯′′′(z1)|<δsε|\cup\mathcal{T}^{\prime\prime\prime}(z_{1})|<\delta^{-s-\varepsilon}. For this 𝒯′′′\mathcal{T}^{\prime\prime\prime} we have μ((𝒯′′′))δε\mu(\mathcal{E}(\mathcal{T}^{\prime\prime\prime}))\geq\delta^{\varepsilon}, hence μ((𝒯′′′)B(z1,δ3ε/κ))δεδ2ε>0\mu(\mathcal{E}(\mathcal{T}^{\prime\prime\prime})\setminus B(z_{1},\delta^{3\varepsilon/\kappa}))\geq\delta^{\varepsilon}-\delta^{2\varepsilon}>0 by the non-concentration property of μ\mu. Thus, we can find z2z_{2} with |z1z2|1|z_{1}-z_{2}|\gtrapprox 1 and 𝒯′′𝒯′′′\mathcal{T}^{\prime\prime}\subset\mathcal{T}^{\prime\prime\prime} with |𝒯′′|δ2ε|𝒯||\mathcal{T}^{\prime\prime}|\geq\delta^{2\varepsilon}|\mathcal{T}| such that X:=|𝒯′′(z1)|<δsεX:=|\cup\mathcal{T}^{\prime\prime}(z_{1})|<\delta^{-s-\varepsilon} and Y:=|𝒯′′(z2)|<δsεY:=|\cup\mathcal{T}^{\prime\prime}(z_{2})|<\delta^{-s-\varepsilon}. Since every t𝒯′′t\in\mathcal{T}^{\prime\prime} passes through a point in XX and a point in YY, and since there are 1\lessapprox 1 many tubes through given points xXx\in X and yYy\in Y, we can find 𝒯𝒯′′\mathcal{T}^{\prime}\subset\mathcal{T}^{\prime\prime} with |𝒯||𝒯|δ2s|\mathcal{T}^{\prime}|\gtrapprox|\mathcal{T}|\gtrapprox\delta^{-2s} such that for every xX,yYx\in X,y\in Y, there is at most one tube in 𝒯\mathcal{T}^{\prime} through x,yx,y. In particular,

δ2s|𝒯||X|δ|Y|δδ2s,\delta^{-2s}\lessapprox|\mathcal{T}^{\prime}|\leq|X|_{\delta}|Y|_{\delta}\lessapprox\delta^{-2s},

and so |X|δ,|Y|δδs|X|_{\delta},|Y|_{\delta}\gtrapprox\delta^{-s}, |𝒯||𝒯|δ2s|\mathcal{T}|\lessapprox|\mathcal{T}^{\prime}|\lessapprox\delta^{-2s}.

For this 𝒯\mathcal{T}^{\prime} we have μ((𝒯))δε\mu(\mathcal{E}(\mathcal{T}^{\prime}))\geq\delta^{\varepsilon}, so defining 𝒟=(𝒯)(B(z1,δ3ε/κ)B(z2,δ3ε/κ))\mathcal{D}=\mathcal{E}(\mathcal{T}^{\prime})\setminus(B(z_{1},\delta^{3\varepsilon/\kappa})\cup B(z_{2},\delta^{3\varepsilon/\kappa})), we have μ(𝒟)δε2δ2ε>δ2ε\mu(\mathcal{D})\geq\delta^{\varepsilon}-2\delta^{2\varepsilon}>\delta^{2\varepsilon}.

Claim 1. For z=az1+(1a)z2𝒟z=az_{1}+(1-a)z_{2}\in\mathcal{D}, we have a,1a1a,1-a\gtrapprox 1. Furthermore, there exists XzXX_{z}\subset X, YzXY_{z}\subset X, and 𝒯z𝒯\mathcal{T}_{z}\subset\mathcal{T}^{\prime} with |Xz|δ,|Yz|δδs,|𝒯z|δ2s|X_{z}|_{\delta},|Y_{z}|_{\delta}\gtrapprox\delta^{-s},|\mathcal{T}_{z}|\gtrapprox\delta^{-2s} such that |Xz+1aaYz|δs|X_{z}+\frac{1-a}{a}Y_{z}|\lessapprox\delta^{-s} and for each t𝒯zt\in\mathcal{T}_{z}, we have t(z1)Xz(δ)t(z_{1})\in X_{z}^{(\delta)} and t(z2)Yz(δ)t(z_{2})\in Y_{z}^{(\delta)}.

Proof. The first claim is evident by definition of 𝒟\mathcal{D}. For the second claim, since z(𝒯)z\in\mathcal{E}(\mathcal{T}^{\prime}), there exists 𝒯z𝒯\mathcal{T}^{\prime}_{z}\subset\mathcal{T}^{\prime} such that |𝒯z|δ2s|\mathcal{T}^{\prime}_{z}|\gtrapprox\delta^{-2s} and |𝒯z(z)|δδs|\mathcal{T}^{\prime}_{z}(z)|_{\delta}\lessapprox\delta^{-s}. Now notice that for each xX,yYx\in X,y\in Y there is at most one tube t𝒯t\in\mathcal{T}^{\prime} passing through x,yx,y. Let PP be the set of (x,y)X×Y(x,y)\in X\times Y with exactly one tube tx,y𝒯zt_{x,y}\in\mathcal{T}^{\prime}_{z} passing through x,yx,y. So |P||𝒯z|δ2s|P|\geq|\mathcal{T}^{\prime}_{z}|\gtrapprox\delta^{-2s}. We also observe that |ax+(1a)ytx,y(z)|δ|ax+(1-a)y-t_{x,y}(z)|\leq\delta, and so |aX+𝑃(1a)Y|δ|𝒯z(z)|2δδ2s|aX\overset{P}{+}(1-a)Y|_{\delta}\leq|\mathcal{T}^{\prime}_{z}(z)|_{2\delta}\lessapprox\delta^{-2s}. Thus, by the Balog-Szemerédi-Gowers theorem 2.27, we can find XzXX_{z}\subset X, YzYY_{z}\subset Y, and 𝒯z𝒯z\mathcal{T}_{z}\subset\mathcal{T}^{\prime}_{z} such that |aXz|δ,|(1a)Yz|δδs,|𝒯z|δ2s|aX_{z}|_{\delta},|(1-a)Y_{z}|_{\delta}\gtrapprox\delta^{-s},|\mathcal{T}_{z}|\gtrapprox\delta^{-2s}, |aXz+(1a)Yz|δs|aX_{z}+(1-a)Y_{z}|\lessapprox\delta^{-s}, and for each t𝒯zt\in\mathcal{T}_{z}, we have t(z1)Xz(δ)t(z_{1})\in X_{z}^{(\delta)} and t(z2)Yz(δ)t(z_{2})\in Y_{z}^{(\delta)}. Then |Xz|δ,|Yz|δδs|X_{z}|_{\delta},|Y_{z}|_{\delta}\gtrapprox\delta^{-s} and |Xz+(1a)aYz|δs|X_{z}+\frac{(1-a)}{a}Y_{z}|\lessapprox\delta^{-s}, proving the Claim. ∎

Now, we apply Lemma 2.22 to the sets Xz(δ)×Yz(δ)X_{z}^{(\delta)}\times Y_{z}^{(\delta)}, the measure 1μ(𝒟)μ|𝒟\frac{1}{\mu(\mathcal{D})}\mu|_{\mathcal{D}}, and K=δCεK=\delta^{-C\varepsilon} for a sufficiently large CC. The result, after applying Fubini’s theorem, is that we can find zz_{*}, X:=XzXX_{*}:=X_{z_{*}}\subset X, Y:=YzYY_{*}:=Y_{z_{*}}\subset Y, and a subset 𝒟𝒟\mathcal{D}^{\prime}\subset\mathcal{D} with μ(𝒟)μ(𝒟)1\mu(\mathcal{D}^{\prime})\gtrapprox\mu(\mathcal{D})\gtrapprox 1 and z𝒟z_{*}\in\mathcal{D}^{\prime} such that for all z𝒟z\in\mathcal{D}^{\prime}, we have

|X(δ)Xz(δ)||Y(δ)Yz(δ)|δ2(n1)δ2s.|X_{*}^{(\delta)}\cap X_{z}^{(\delta)}||Y_{*}^{(\delta)}\cap Y_{z}^{(\delta)}|\gtrapprox\delta^{2(n-1)}\delta^{-2s}.

Since |X(δ)Xz(δ)||X|δδs|X_{*}^{(\delta)}\cap X_{z}^{(\delta)}|\lesssim|X|_{\delta}\lessapprox\delta^{-s} and |Y(δ)Yz(δ)||Y|δδs|Y_{*}^{(\delta)}\cap Y_{z}^{(\delta)}|\lesssim|Y|_{\delta}\lessapprox\delta^{-s}, we have in fact |X(δ)Xz(δ)|,|Y(δ)Yz(δ)|δs|X_{*}^{(\delta)}\cap X_{z}^{(\delta)}|,|Y_{*}^{(\delta)}\cap Y_{z}^{(\delta)}|\approx\delta^{-s}. In particular, |Xz(δ)|,|Yz(δ)|δs|X_{z}^{(\delta)}|,|Y_{z}^{(\delta)}|\approx\delta^{-s} for all z𝒟z\in\mathcal{D}^{\prime}.

The next leg of the proof is to show:

Claim 2. For all z𝒟z\in\mathcal{D}^{\prime}, if we write z=az1+(1a)z2z=az_{1}+(1-a)z_{2}, then |X+1aaY|δδs|X_{*}+\frac{1-a}{a}Y_{*}|_{\delta}\lessapprox\delta^{-s}.

Proof. Note that Claim 1 tells us |Xz+1aaYz|δδs|X_{z}+\frac{1-a}{a}Y_{z}|_{\delta}\lessapprox\delta^{-s}. Combining this with the Rusza triangle inequality (Lemma 2.23), X(δ)Xz(δ)Xz(δ)X_{*}^{(\delta)}\cap X_{z}^{(\delta)}\subset X_{z}^{(\delta)}, and |A(δ)|δd|A|δ|A^{(\delta)}|_{\delta}\sim_{d}|A|_{\delta} for any subset AA of the doubling metric space \Rd\R^{d}, we have

|XzX(δ)Xz(δ)|δ|Xz(δ)Xz(δ)|δ|XzXz|δ|Xz+1aaYz|δ2|1aaYz|δs.|X_{z}-X_{*}^{(\delta)}\cap X_{z}^{(\delta)}|_{\delta}\lesssim|X_{z}^{(\delta)}-X_{z}^{(\delta)}|_{\delta}\lesssim|X_{z}-X_{z}|_{\delta}\lesssim\frac{|X_{z}+\frac{1-a}{a}Y_{z}|_{\delta}^{2}}{|\frac{1-a}{a}Y_{z}|}\lessapprox\delta^{-s}.

The same argument shows (where z=az1+(1a)z2z_{*}=a_{*}z_{1}+(1-a_{*})z_{2}):

|XX(δ)Xz(δ)|δ|XX|δ|X+1aaY|δ2|1aaY|δs.|X_{*}-X_{*}^{(\delta)}\cap X_{z}^{(\delta)}|_{\delta}\lesssim|X_{*}-X_{*}|_{\delta}\lesssim\frac{|X_{*}+\frac{1-a_{*}}{a_{*}}Y_{*}|_{\delta}^{2}}{|\frac{1-a_{*}}{a_{*}}Y_{*}|}\lessapprox\delta^{-s}.

Thus, by Lemma 2.23 again, we have

|XXz|δ|XzX(δ)Xz(δ)|δ|XX(δ)Xz(δ)|δ|X(δXz(δ)|δδs.|X_{*}-X_{z}|_{\delta}\lesssim\frac{|X_{z}-X_{*}^{(\delta)}\cap X_{z}^{(\delta)}|_{\delta}|X_{*}-X_{*}^{(\delta)}\cap X_{z}^{(\delta)}|_{\delta}}{|X_{*}^{(\delta}\cap X_{z}^{(\delta)}|_{\delta}}\lessapprox\delta^{-s}.

Similarly, we have |YYz|δδs|Y_{*}-Y_{z}|_{\delta}\lessapprox\delta^{-s}. A final application of Lemma 2.23 gives

|X+1aaY|δ\displaystyle|X_{*}+\frac{1-a}{a}Y_{*}|_{\delta} |Xz+1aaY|δ|XzX|δ|Xz|δ\displaystyle\lesssim\frac{|X_{z}+\frac{1-a}{a}Y_{*}|_{\delta}|X_{z}-X_{*}|_{\delta}}{|X_{z}|_{\delta}}
|Xz+1aaYz|δ|XzX|δ|1aa(YzY)|δ|Xz|δ|1aaYz|δ\displaystyle\lesssim\frac{|X_{z}+\frac{1-a}{a}Y_{z}|_{\delta}|X_{z}-X_{*}|_{\delta}|\frac{1-a}{a}(Y_{z}-Y_{*})|_{\delta}}{|X_{z}|_{\delta}|\frac{1-a}{a}Y_{z}|_{\delta}}
δsδsδsδsδsδs.\displaystyle\lessapprox\frac{\delta^{-s}\delta^{-s}\delta^{-s}}{\delta^{-s}\delta^{-s}}\leq\delta^{-s}.

This proves Claim 2. ∎

Finally, we seek a contradiction by applying Theorem 3.6 to 𝒯z\mathcal{T}_{z_{*}} and μ|𝒟\mu|_{\mathcal{D}^{\prime}}. We satisfy the condition (if ε\varepsilon is sufficiently small) because Claim 1 and |𝒯|δ2s|\mathcal{T}|\lessapprox\delta^{-2s} tell us that 𝒯z\mathcal{T}_{z_{*}} is a (δ,κ,δO(ε),k)(\delta,\kappa,\delta^{-O(\varepsilon)},k)-set with |𝒯z|δ2s|\mathcal{T}_{z_{*}}|\gtrapprox\delta^{-2s}. But we violate the conclusion (if ε\varepsilon is sufficiently small) because Claim 2 tells us that |aX+(1a)Y|δδs|aX_{*}+(1-a)Y_{*}|_{\delta}\lessapprox\delta^{-s}. This contradiction finishes the proof of Theorem 3.7.

3.5 Formal exhaustion argument

Using Theorem 3.7, we prove the following proposition, which implies Theorem 3.3 with a different value for ε\varepsilon.

Proposition 3.8.

For 0k<d10\leq k<d-1, 0s<k+10\leq s<k+1, and 0<κ10<\kappa\leq 1, there exists ε>0\varepsilon>0 such that the following holds for sufficiently small δ<δ0(s,k,d,ε)\delta<\delta_{0}(s,k,d,\varepsilon). Let 𝒯\mathcal{T} be a (δ,κ,δε/2,k)(\delta,\kappa,\delta^{-\varepsilon/2},k)-set of δ\delta-tubes each making angle 1100\geq\frac{1}{100} with the plane y=0y=0 with |𝒯|δ2s+ε/2|\mathcal{T}|\geq\delta^{-2s+\varepsilon/2}. Let μ\mu be a probability measure on \R\R such that for all δr1\delta\leq r\leq 1, we have μ(Br)δεrκ\mu(B_{r})\leq\delta^{-\varepsilon}r^{\kappa}. Then μ((𝒯,ε3))δε/2\mu(\mathcal{E}(\mathcal{T},\frac{\varepsilon}{3}))\leq\delta^{\varepsilon/2}.

The idea is the following. A first application of Theorem 3.7 gives a subset 𝒯𝒯\mathcal{T}^{\prime}\subset\mathcal{T} with μ((𝒯,ϵ))δϵ\mu(\mathcal{E}(\mathcal{T}^{\prime},\epsilon))\leq\delta^{\epsilon}. Either 𝒯\mathcal{T}^{\prime} is large enough in which case we are done or we can cut 𝒯\mathcal{T}^{\prime} out of 𝒯\mathcal{T} and apply Theorem 1.8 again. This will give us another subset 𝒯\mathcal{T}^{\prime}. Then we iterate until the union of these sets 𝒯\mathcal{T}^{\prime} is large enough.

Proof.

Let N0N\geq 0 be an integer. Suppose we have already constructed pairwise disjoint sets 𝒯1,,𝒯N\mathcal{T}_{1},\cdots,\mathcal{T}_{N} such that μ((𝒯i,ϵ))δϵ\mu(\mathcal{E}(\mathcal{T}_{i},\epsilon))\leq\delta^{\epsilon} for every i=1,,Ni=1,\cdots,N. Either we have

|𝒯i=1N𝒯i|δϵ2|𝒯|,\left|\mathcal{T}\setminus\bigcup_{i=1}^{N}\mathcal{T}_{i}\right|\leq\delta^{\frac{\epsilon}{2}}|\mathcal{T}|, (3.4)

in which case we stop, or the set 𝒯i=1N𝒯i\mathcal{T}\setminus\bigcup_{i=1}^{N}\mathcal{T}_{i} satisfies the conditions of Theorem 3.3. In the latter case Theorem 3.3 gives us 𝒯N+1𝒯i=1N𝒯i\mathcal{T}_{N+1}\subset\mathcal{T}\setminus\bigcup_{i=1}^{N}\mathcal{T}_{i} with μ((AN+1,ϵ))δϵ\mu(\mathcal{E}(A_{N+1},\epsilon))\leq\delta^{\epsilon}. By construction, 𝒯N+1\mathcal{T}_{N+1} is disjoint with any of the 𝒯i\mathcal{T}_{i}, i=1,,Ni=1,\cdots,N.

When this procedure ends write 𝒯0=i=1N𝒯i\mathcal{T}_{0}=\bigcup_{i=1}^{N}\mathcal{T}_{i}. Then (3.4) says |𝒯𝒯0|δε2|𝒯||\mathcal{T}\setminus\mathcal{T}_{0}|\leq\delta^{\frac{\varepsilon}{2}}|\mathcal{T}|. Moreover, since the 𝒯i\mathcal{T}_{i}’s are disjoint, |𝒯0|=i=1N|𝒯i||\mathcal{T}_{0}|=\sum_{i=1}^{N}|\mathcal{T}_{i}|.

Set ai=|𝒯i||𝒯0|a_{i}=\frac{|\mathcal{T}_{i}|}{|\mathcal{T}_{0}|}. We claim that

(𝒯,ε3)IiI(𝒯i,ε),\mathcal{E}(\mathcal{T},\frac{\varepsilon}{3})\subset\bigcup_{I}\bigcap_{i\in I}\mathcal{E}(\mathcal{T}_{i},\varepsilon),

where the index set II runs over subsets of {1,2,,n}\{1,2,\cdots,n\} with iIaiδε2\sum_{i\in I}a_{i}\geq\delta^{\frac{\varepsilon}{2}}. Since μ((𝒯i)))δε\mu(\mathcal{E}(\mathcal{T}_{i})))\leq\delta^{\varepsilon} for all ii, the desired upper bound μ((𝒯,ε3))δε/2\mu(\mathcal{E}(\mathcal{T},\frac{\varepsilon}{3}))\leq\delta^{\varepsilon/2} then follows immediately from Markov’s inequality applied to the event ai𝟙𝒯i\sum a_{i}\mathbbm{1}_{\mathcal{T}_{i}} (or [10, Lemma 20]).

We will now show the claim. Let z0(A,ε3)z_{0}\in\mathcal{E}(A,\frac{\varepsilon}{3}), so there exists 𝒯𝒯\mathcal{T}^{\prime}\subset\mathcal{T} with |𝒯|δε3|𝒯||\mathcal{T}^{\prime}|\geq\delta^{\frac{\varepsilon}{3}}|\mathcal{T}| and |πz0(𝒯)|δδsε3|\pi_{z_{0}}(\mathcal{T}^{\prime})|_{\delta}\leq\delta^{-s-\frac{\varepsilon}{3}}. Consider the index set II defined as

I={1in|𝒯𝒯i|δε|𝒯i|}.I=\{1\leq i\leq n\mid|\mathcal{T}^{\prime}\cap\mathcal{T}_{i}|\geq\delta^{\varepsilon}|\mathcal{T}_{i}|\}.

We have

δε/2|𝒯|\displaystyle\delta^{\varepsilon/2}|\mathcal{T}| |𝒯||𝒯𝒯0|\displaystyle\leq|\mathcal{T}^{\prime}|-|\mathcal{T}\setminus\mathcal{T}_{0}|
i=1n|𝒯𝒯i|\displaystyle\leq\sum_{i=1}^{n}|\mathcal{T}^{\prime}\cap\mathcal{T}_{i}|
iI|𝒯i|+iIδε|𝒯i|\displaystyle\lesssim\sum_{i\in I}|\mathcal{T}_{i}|+\sum_{i\notin I}\delta^{\varepsilon}|\mathcal{T}_{i}|
iIai|𝒯|+δε|𝒯|\displaystyle\lesssim\sum_{i\in I}a_{i}|\mathcal{T}|+\delta^{\varepsilon}|\mathcal{T}|

Hence iIaiδε2\sum_{i\in I}a_{i}\geq\delta^{\frac{\varepsilon}{2}}. On the other hand, for all iIi\in I, since

|πz0(𝒯𝒯i))|δ|πz0(𝒯)|δδsε3,|\pi_{z_{0}}(\mathcal{T}^{\prime}\cap\mathcal{T}_{i}))|_{\delta}\leq|\pi_{z_{0}}(\mathcal{T}^{\prime})|_{\delta}\leq\delta^{-s-\frac{\varepsilon}{3}},

we have z0(𝒯i,ε)z_{0}\in\mathcal{E}(\mathcal{T}_{i},\varepsilon) for all iIi\in I. This finishes the proof of the claim. ∎

3.6 Proof of Proposition 3.1

This subsection is based on Section A.7 of [18].

We restate Proposition 3.1.

Proposition 3.9.

Given 0k<d10\leq k<d-1, 0s<k+10\leq s<k+1, τ,κ>0\tau,\kappa>0, there exist η(s,k,κ,τ,d)>0\eta(s,k,\kappa,\tau,d)>0 and δ0(s,k,κ,τ,d)>0\delta_{0}(s,k,\kappa,\tau,d)>0 such that the following holds for all δ(0,δ0]\delta\in(0,\delta_{0}].

Let 𝐘(δ)[0,1)\mathbf{Y}\subset(\delta\cdot\mathbb{Z})\cap[0,1) be a (δ,τ,δη)(\delta,\tau,\delta^{-\eta})-set, and for each 𝐲𝐘\mathbf{y}\in\mathbf{Y}, assume that 𝐗𝐲(δ)d1[0,1)d1\mathbf{X}_{\mathbf{y}}\subset(\delta\cdot\mathbb{Z})^{d-1}\cap[0,1)^{d-1} is a (δ,κ,δη,k)(\delta,\kappa,\delta^{-\eta},k)-set with cardinality δs+η\geq\delta^{-s+\eta}. Let

𝐙=𝐲𝐘𝐗𝐲×{𝐲}.\mathbf{Z}=\bigcup_{\mathbf{y}\in\mathbf{Y}}\mathbf{X}_{\mathbf{y}}\times\{\mathbf{y}\}.

For every 𝐳𝐙\mathbf{z}\in\mathbf{Z}, assume that 𝒯(𝐳)\mathcal{T}(\mathbf{z}) is a set of δ\delta-tubes each making an angle 1100\geq\frac{1}{100} with the plane y=0y=0 with |𝒯(𝐳)|δs+η|\mathcal{T}(\mathbf{z})|\geq\delta^{-s+\eta} such that 𝐳T\mathbf{z}\in T for all T𝒯(𝐳)T\in\mathcal{T}(\mathbf{z}). Then |𝒯|δ2sη|\mathcal{T}|\geq\delta^{-2s-\eta}, where 𝒯=𝐳𝐙𝒯(𝐳)\mathcal{T}=\cup_{\mathbf{z}\in\mathbf{Z}}\mathcal{T}(\mathbf{z}).

Proof.

Let ABA\lessapprox B denote ACδCηBA\leq C\delta^{-C\eta}B for some absolute constant C1C\geq 1. A (δ,u,m)(\delta,u,m)-set stands for a (δ,u,CδCη,m)(\delta,u,C\delta^{-C\eta},m)-set.

First, without loss of generality, assume |𝒯(𝐳)|=δs+η|\mathcal{T}(\mathbf{z})|=\delta^{-s+\eta} for each 𝐳𝐙\mathbf{z}\in\mathbf{Z}.

Suppose |𝒯|δ2sη|\mathcal{T}|\leq\delta^{-2s-\eta}. Let

𝒯(𝐲)=𝐱𝐗𝐲𝒯(𝐱,𝐲).\mathcal{T}(\mathbf{y})=\bigcup_{\mathbf{x}\in\mathbf{X}_{\mathbf{y}}}\mathcal{T}(\mathbf{x},\mathbf{y}).

Since each tube in 𝒯(𝐲)\mathcal{T}(\mathbf{y}) has angle 1100\geq\frac{1}{100} with the plane y=0y=0, it only intersects O(1)O(1) many δ\delta-balls (𝐱,𝐲)(\mathbf{x},\mathbf{y}) for a given 𝐲\mathbf{y}. Since |𝒯(𝐱,𝐲)|δs|\mathcal{T}(\mathbf{x},\mathbf{y})|\gtrapprox\delta^{-s} for each 𝐱𝐗𝐲\mathbf{x}\in\mathbf{X}_{\mathbf{y}}, we get |𝒯(𝐲)|δs|𝐗𝐲||\mathcal{T}(\mathbf{y})|\gtrapprox\delta^{-s}|\mathbf{X}_{\mathbf{y}}|. With the counter-assumption |𝒯|δ2s|\mathcal{T}|\lessapprox\delta^{-2s}, this forces |𝐗𝐲|δs|\mathbf{X}_{\mathbf{y}}|\lessapprox\delta^{-s} for each 𝐲𝐘\mathbf{y}\in\mathbf{Y}. On the other hand, |𝐗𝐲|δs|\mathbf{X}_{\mathbf{y}}|\gtrapprox\delta^{-s} and so |𝒯|δ2s|\mathcal{T}|\approx\delta^{-2s}.

Now, we check that 𝒯(𝐲)\mathcal{T}(\mathbf{y}) is a (δ,κ,δO(η),k)(\delta,\kappa,\delta^{-O(\eta)},k)-set. Pick a (r,k+1)(r,k+1)-plane HH. We claim that either 𝒯(𝐲)H=\mathcal{T}(\mathbf{y})\cap H=\emptyset or H(y=𝐲)H(y=\mathbf{y}) is contained in a (O(r),k)(O(r),k)-plate. Indeed, if H(y=𝐲)H(y=\mathbf{y}) is not contained within a (Cr,k)(Cr,k)-plate, then HH is contained within the O(C1)O(C^{-1})-neighborhood of the plane y=𝐲y=\mathbf{y}, which means that HH cannot contain any tubes of 𝒯(𝐲)\mathcal{T}(\mathbf{y}) if CC is large enough (since the tubes of 𝒯(𝐲)\mathcal{T}(\mathbf{y}) have angle 1100\geq\frac{1}{100} with that plane). Thus, we may assume H(y=𝐲)H(y=\mathbf{y}) is contained within a (Cr,k)(Cr,k)-plate, which means

|𝒯(𝐲)H|=|𝐱𝐗𝐲H𝒯(𝐱,𝐲)H||𝐗𝐲H|δs+η|𝐗𝐲|rκδs+ηrκ|𝒯(𝐲)|.|\mathcal{T}(\mathbf{y})\cap H|=|\bigcup_{\mathbf{x}\in\mathbf{X}_{\mathbf{y}}\cap H}\mathcal{T}(\mathbf{x},\mathbf{y})\cap H|\\ \leq|\mathbf{X}_{\mathbf{y}}\cap H|\cdot\delta^{-s+\eta}\lessapprox|\mathbf{X}_{\mathbf{y}}|r^{\kappa}\cdot\delta^{-s+\eta}\lessapprox r^{\kappa}|\mathcal{T}(\mathbf{y})|.

Since |𝒯(𝐲)||𝒯||\mathcal{T}(\mathbf{y})|\approx|\mathcal{T}| for each 𝐲𝐘\mathbf{y}\in\mathbf{Y}, there is a subset 𝒯¯𝒯\overline{\mathcal{T}}\subset\mathcal{T} such that |𝒯||𝒯¯||\mathcal{T}|\approx|\overline{\mathcal{T}}| and each T𝒯¯T\in\overline{\mathcal{T}} belongs to |𝐘|\approx|\mathbf{Y}| of the sets 𝒯(𝐲)\mathcal{T}(\mathbf{y}). We show 𝒯¯\overline{\mathcal{T}} is a (δ,κ,δO(η),k)(\delta,\kappa,\delta^{-O(\eta)},k)-set. Indeed, given a (r,k+1)(r,k+1)-plate HH, we have

|𝒯¯H|T𝒯¯H1|𝐘|𝐲𝐘𝟙𝒯(y)(T)1|𝐘|𝐲𝐘|𝒯¯(𝐲)H|1|𝐘|𝐲𝐘rκ|𝒯(y)|rκ|𝒯¯|.|\overline{\mathcal{T}}\cap H|\approx\sum_{T\in\overline{\mathcal{T}}\cap H}\frac{1}{|\mathbf{Y}|}\sum_{\mathbf{y}\in\mathbf{Y}}\mathbbm{1}_{\mathcal{T}(y)}(T)\\ \lessapprox\frac{1}{|\mathbf{Y}|}\sum_{\mathbf{y}\in\mathbf{Y}}|\overline{\mathcal{T}}(\mathbf{y})\cap H|\lessapprox\frac{1}{|\mathbf{Y}|}\sum_{\mathbf{y}\in\mathbf{Y}}r^{\kappa}|\mathcal{T}(y)|\leq r^{\kappa}|\overline{\mathcal{T}}|.

Finally, we refine 𝐘\mathbf{Y} further: since

𝐲𝐘|𝒯¯𝒯(y)|=T𝒯¯|{𝐲𝐘:T𝒯(𝐘)}||𝒯¯||𝐘|,\sum_{\mathbf{y}\in\mathbf{Y}}|\overline{\mathcal{T}}\cap\mathcal{T}(y)|=\sum_{T\in\overline{\mathcal{T}}}|\{\mathbf{y}\in\mathbf{Y}:T\in\mathcal{T}(\mathbf{Y})\}|\approx|\overline{\mathcal{T}}||\mathbf{Y}|,

we can find a subset 𝐘¯𝐘\overline{\mathbf{Y}}\subset\mathbf{Y} with the property that |𝒯¯(y)|:=|𝒯¯𝒯(y)||𝒯¯||\overline{\mathcal{T}}(y)|:=|\overline{\mathcal{T}}\cap\mathcal{T}(y)|\approx|\overline{\mathcal{T}}| for each y𝐘¯y\in\overline{\mathbf{Y}}. Also, 𝐘¯\overline{\mathbf{Y}} is still a (δ,τ,δO(η))(\delta,\tau,\delta^{-O(\eta)})-set.

Now for each 𝐲𝐘¯\mathbf{y}\in\overline{\mathbf{Y}}, the large subset 𝒯¯(y)𝒯¯\overline{\mathcal{T}}(y)\subset\overline{\mathcal{T}} has small covering number |𝐗𝐲|δs|\mathbf{X}_{\mathbf{y}}|\approx\delta^{-s}. On the other hand, |𝒯¯|δ2s|\overline{\mathcal{T}}|\approx\delta^{-2s}. This contradicts Theorem 3.3 if η\eta is chosen sufficiently small in terms of the ε\varepsilon of the theorem. ∎

4 Improved incidence estimates for regular sets

In this section, we prove a version of Theorem 1.8 for regular sets.

Definition 4.1.

Let δ22\delta\in 2^{-2\mathbb{N}} be a dyadic number. Let C,K>0C,K>0, and let 0sd0\leq s\leq d. A non-empty set 𝒫𝒟δ\mathcal{P}\subset\mathcal{D}_{\delta} is called (δ,s,C,K)(\delta,s,C,K)-regular if 𝒫\mathcal{P} is a (δ,s,C,0)(\delta,s,C,0)-set, and

|𝒫|δ1/2Kδs/2.|\mathcal{P}|_{\delta^{1/2}}\leq K\cdot\delta^{-s/2}.
Theorem 4.2.

For any 0s,k<d10\leq s,k<d-1, max(s,k)<td\max(s,k)<t\leq d, κ>0\kappa>0, there exists ε(s,t,κ,k,d)>0\varepsilon(s,t,\kappa,k,d)>0 such that the following holds for all small enough δ2\delta\in 2^{-\mathbb{N}}, depending only on s,t,κ,k,ds,t,\kappa,k,d. Let 𝒫𝒟δ\mathcal{P}\subset\mathcal{D}_{\delta} be a (δ,t,δε,δε)(\delta,t,\delta^{-\varepsilon},\delta^{-\varepsilon})-regular set. Assume that for every p𝒫p\in\mathcal{P}, there exists a (δ,s,δε,0)(\delta,s,\delta^{-\varepsilon},0) and (δ,κ,δε,k)(\delta,\kappa,\delta^{-\varepsilon},k)-set 𝒯(p)𝒯\mathcal{T}(p)\subset\mathcal{T} with |𝒯(p)|=M|\mathcal{T}(p)|=M such that TpT\cap p\neq\emptyset for all T𝒯(p)T\in\mathcal{T}(p). Then |𝒯|Mδsε|\mathcal{T}|\geq M\delta^{-s-\varepsilon}.

4.1 Initial reductions

This subsection is based on Sections 6 and A.1-A.3 of [18].

In this section, let ABA\lessapprox B denote ACδCεBA\leq C\delta^{-C\varepsilon}B for some constant C1C\geq 1 depending only on s,t,κ,k,ds,t,\kappa,k,d. Also, let 𝒫Q:={p𝒫:pQ}\mathcal{P}\cap Q:=\{p\in\mathcal{P}:p\subset Q\}.

The proof will be based on contradiction, so assume |𝒯|Mδsε|\mathcal{T}|\leq M\delta^{-s-\varepsilon}. Let’s rename 𝒫\mathcal{P} to 𝒫0\mathcal{P}_{0} and 𝒯\mathcal{T} to 𝒯0\mathcal{T}_{0}, reserving 𝒫,𝒯\mathcal{P},\mathcal{T} for the use of Proposition 2.13.

By Corollary 2.10, we have 1(Mδs)tsd1s1\gtrapprox(M\delta^{s})^{\frac{t-s}{d-1-s}}, so MδsM\lessapprox\delta^{-s} and |𝒯|δ2s|\mathcal{T}|\lessapprox\delta^{-2s}. But 𝒯(p)\mathcal{T}(p) is a (δ,s,δε)(\delta,s,\delta^{-\varepsilon})-set, so MδsM\approx\delta^{-s}. Finally, by Lemma 2.7, we may assume |𝒫0|δt|\mathcal{P}_{0}|\approx\delta^{-t} (passing to subsets will preserve the (δ,t,δε,δε)(\delta,t,\delta^{-\varepsilon},\delta^{-\varepsilon})-regularity of 𝒫0\mathcal{P}_{0}).

The next reduction will make the value |𝒫0Q||\mathcal{P}_{0}\cap Q| uniform for different Q𝒟Δ(𝒫0)Q\in\mathcal{D}_{\Delta}(\mathcal{P}_{0}). Let 𝒬0=𝒟Δ(𝒫0)\mathcal{Q}_{0}=\mathcal{D}_{\Delta}(\mathcal{P}_{0}). By (δ,t,δε,δε)(\delta,t,\delta^{-\varepsilon},\delta^{-\varepsilon})-regularity of 𝒫0\mathcal{P}_{0}, we have |𝒬0|Δt|\mathcal{Q}_{0}|\lessapprox\Delta^{-t}. On the other hand, since 𝒫0\mathcal{P}_{0} is a (δ,t)(\delta,t)-set, we have that for all Q𝒬0Q\in\mathcal{Q}_{0},

|𝒫0Q|Δt|\mathcal{P}_{0}\cap Q|\lessapprox\Delta^{-t} (4.5)

This means |𝒬0|Δt|\mathcal{Q}_{0}|\gtrapprox\Delta^{-t}. Hence, |𝒬0|Δt|\mathcal{Q}_{0}|\approx\Delta^{-t}. Now using (4.5) again and |𝒫0|Δ2t|\mathcal{P}_{0}|\approx\Delta^{-2t}, there exists 𝒬0𝒬0\mathcal{Q}_{0}^{\prime}\subset\mathcal{Q}_{0} with |𝒬0||𝒬0||\mathcal{Q}_{0}^{\prime}|\gtrapprox|\mathcal{Q}_{0}| such that for each Q𝒬0Q\in\mathcal{Q}_{0}^{\prime},

|𝒫0Q|Δt|\mathcal{P}_{0}\cap Q|\approx\Delta^{-t} (4.6)

Using (4.6), we quickly check that 𝒬0\mathcal{Q}_{0}^{\prime} is a (Δ,t)(\Delta,t)-set. Indeed, for r(Δ,1)r\in(\Delta,1) and Qr𝒟rQ_{r}\in\mathcal{D}_{r}, we have

|𝒬0Qr|(4.6)Δt|𝒫0Qr|Δt|𝒫0|rt|𝒬0|rt.|\mathcal{Q}_{0}^{\prime}\cap Q_{r}|\overset{\eqref{eqn:good squares}}{\approx}\Delta^{t}\cdot|\mathcal{P}_{0}\cap Q_{r}|\lessapprox\Delta^{t}\cdot|\mathcal{P}_{0}|\cdot r^{t}\approx|\mathcal{Q}_{0}^{\prime}|\cdot r^{t}. (4.7)

(The second inequality uses that 𝒫0\mathcal{P}_{0} is a (Δ,t)(\Delta,t)-set.) Let 𝒫0=Q𝒬0𝒫Q\mathcal{P}_{0}^{\prime}=\bigcup_{Q\in\mathcal{Q}_{0}^{\prime}}\mathcal{P}\cap Q; then |𝒫0||𝒫||\mathcal{P}_{0}^{\prime}|\approx|\mathcal{P}| and |𝒫0Q|=|𝒫0Q|Δt|\mathcal{P}_{0}^{\prime}\cap Q|=|\mathcal{P}_{0}\cap Q|\approx\Delta^{-t} for Q𝒫0Q\in\mathcal{P}_{0}^{\prime}. Apply Proposition 2.13 to find 𝒫𝒫0,𝒯(p)𝒯0(p)\mathcal{P}\subset\mathcal{P}_{0}^{\prime},\mathcal{T}(p)\subset\mathcal{T}_{0}(p), 𝒯Δ\mathcal{T}_{\Delta}, 𝒯\mathcal{T}, and 𝒯Q\mathcal{T}_{Q}. Let 𝒬=𝒟Δ(𝒫)\mathcal{Q}=\mathcal{D}_{\Delta}(\mathcal{P}).

Claim. MΔΔsM_{\Delta}\approx\Delta^{-s} and |𝒯Δ|δs|\mathcal{T}_{\Delta}|\lessapprox\delta^{-s}.

Proof. By Proposition 2.13iii, we know that (𝒟Δ(𝒫),𝒯Δ)(\mathcal{D}_{\Delta}(\mathcal{P}),\mathcal{T}_{\Delta}) is (Δ,s,CΔ1,κ,CΔ2,MΔ)(\Delta,s,C_{\Delta}^{1},\kappa,C_{\Delta}^{2},M_{\Delta})-nice, so MΔΔsM_{\Delta}\gtrapprox\Delta^{-s}. Also, by Corollary 2.10, we have that

|𝒯Δ|MΔδs/2(MΔδs/2)tsd1s.|\mathcal{T}_{\Delta}|\gtrapprox M_{\Delta}\delta^{-s/2}\cdot(M_{\Delta}\delta^{s/2})^{\frac{t-s}{d-1-s}}. (4.8)

Next, for any Q𝒬Q\in\mathcal{Q}, we know that (SQ(𝒫Q),𝒯Q)(S_{Q}(\mathcal{P}\cap Q),\mathcal{T}_{Q}) is (Δ,s,CQ1,κ,CQ2,MQ)(\Delta,s,C_{Q}^{1},\kappa,C_{Q}^{2},M_{Q})-nice. Recall that

SQ(𝒫Q)={SQ(p):p𝒫,pQ}𝒟Δ.S_{Q}(\mathcal{P}\cap Q)=\{S_{Q}(p):p\in\mathcal{P},p\subset Q\}\subset\mathcal{D}_{\Delta}.

We also know |𝒫Q||𝒫0Q|δt/2|\mathcal{P}\cap Q|\approx|\mathcal{P}_{0}^{\prime}\cap Q|\approx\delta^{-t/2} and 𝒫\mathcal{P} is a (δ,t)(\delta,t)-set, so by a similar check to (4.7), we get that SQ(𝒫Q)S_{Q}(\mathcal{P}\cap Q) is a (Δ,t)(\Delta,t)-set. Thus by Corollary 2.10, we have

|𝒯Q|MQδs/2|\mathcal{T}_{Q}|\gtrapprox M_{Q}\cdot\delta^{-s/2}

But by our counterassumption |𝒯0|δ2s|\mathcal{T}_{0}|\lessapprox\delta^{-2s}, we get from (2.3) in Proposition 2.13 and MδsM\gtrapprox\delta^{-s},

δ2s|𝒯Δ|MΔ|𝒯Q|MQM|𝒯Δ|MΔδ3s/2.\delta^{-2s}\gtrapprox\frac{|\mathcal{T}_{\Delta}|}{M_{\Delta}}\cdot\frac{|\mathcal{T}_{Q}|}{M_{Q}}\cdot M\gtrapprox\frac{|\mathcal{T}_{\Delta}|}{M_{\Delta}}\cdot\delta^{-3s/2}.

Thus, |𝒯Δ|MΔδs/2|\mathcal{T}_{\Delta}|\lessapprox M_{\Delta}\delta^{-s/2}. Substitute into (4.8) to get

δs/2|TΔ|MΔδs/2(MΔδs/2)tsd1s.\delta^{-s/2}\gtrapprox\frac{|T_{\Delta}|}{M_{\Delta}}\gtrapprox\delta^{-s/2}\cdot(M_{\Delta}\delta^{s/2})^{\frac{t-s}{d-1-s}}.

Thus, MΔδs/21M_{\Delta}\delta^{s/2}\lessapprox 1, so MΔΔsM_{\Delta}\lessapprox\Delta^{-s} and |𝒯Δ|δs|\mathcal{T}_{\Delta}|\lessapprox\delta^{-s}, proving the Claim. ∎

Thus, we get the higher-dimensional analogues of properties (H1-2), (G1-4) of [18] except we only know |𝒯|δ2s|\mathcal{T}|\lessapprox\delta^{-2s} and not |𝒯|δ2s|\mathcal{T}|\gtrapprox\delta^{-2s}. But this is not a limitation. We repeat and relabel these properties here:

  1. (G1)

    |𝒬|Δt|\mathcal{Q}|\approx\Delta^{-t} and |𝒫Q|Δt|\mathcal{P}\cap Q|\approx\Delta^{-t} for all Q𝒬Q\in\mathcal{Q}.

  2. (G2)

    Every tube 𝐓𝒯Δ\mathbf{T}\in\mathcal{T}_{\Delta} satisfies |𝒯𝐓|δs|\mathcal{T}\cap\mathbf{T}|\lessapprox\delta^{-s}.

  3. (G3)

    For every square Q𝒬Q\in\mathcal{Q}, there corresponds a (Δ,s,0)(\Delta,s,0)-set and (Δ,κ,k)(\Delta,\kappa,k)-set 𝒯Δ(Q)𝒯Δ\mathcal{T}_{\Delta}(Q)\subset\mathcal{T}_{\Delta} of cardinality MΔΔs\approx M_{\Delta}\approx\Delta^{-s} such that 𝐓Q\mathbf{T}\cap Q\neq\emptyset for all 𝐓𝒯Δ(Q)\mathbf{T}\in\mathcal{T}_{\Delta}(Q).

  4. (G4)

    |𝒯|δ2s|\mathcal{T}|\lessapprox\delta^{-2s} and |𝒯Δ|Δ2s|\mathcal{T}_{\Delta}|\approx\Delta^{-2s}.

  5. (G5)

    For 𝐓𝒯Δ(Q)\mathbf{T}\in\mathcal{T}_{\Delta}(Q), we have

    |{(p,T)(𝒫Q)×𝒯:T𝒯(p)𝐓}|Δst.|\{(p,T)\in(\mathcal{P}\cap Q)\times\mathcal{T}:T\in\mathcal{T}(p)\cap\mathbf{T}\}|\gtrapprox\Delta^{-s-t}.

Item 1 follows from Proposition 2.13i.

Item 3 follows from Proposition 2.13iii and Claim.

Item 4 follows from |𝒯0|δ2s|\mathcal{T}_{0}|\lessapprox\delta^{-2s} and Claim.

Item 5 follows from Proposition 2.13iv and the estimation M|𝒫Q|/|𝒯Δ(Q)|ΔstM\cdot|\mathcal{P}\cap Q|/|\mathcal{T}_{\Delta}(Q)|\approx\Delta^{-s-t}, which uses item 1, item 3, and the fact MδsM\approx\delta^{-s} we proved at the beginning of the argument.

Item 2 follows from Proposition 2.13ii, the fact that a given δ\delta-tube lies in 1\lesssim 1 many of the 𝐓\mathbf{T}’s in 𝒯Δ\mathcal{T}_{\Delta}, and item 4:

δ2s|𝒯|𝐓𝒯Δ|𝒯𝐓||𝒯Δ|𝐍Δ2s𝐍.\delta^{-2s}\gtrapprox|\mathcal{T}|\gtrsim\sum_{\mathbf{T}\in\mathcal{T}_{\Delta}}|\mathcal{T}\cap\mathbf{T}|\sim|\mathcal{T}_{\Delta}|\cdot\mathbf{N}\approx\Delta^{-2s}\cdot\mathbf{N}.

4.2 Transferring angular non-concentration to ball non-concentration

This subsection is based on Section A.4 of [18].

We first recall some notation. For a unit vector σ\Rd\sigma\in\R^{d}, define πσ(v):=v(vσ)σ\pi_{\sigma}(\vec{v}):=\vec{v}-(\vec{v}\cdot\sigma)\sigma to be the orthogonal projection to the orthogonal complement of σ\sigma. For a δ\delta-tube TT, let σ(T)Sd1\sigma(T)\in S^{d-1} denote the direction of TT.

In this subsection, we fix a Q𝒟Δ(𝒫)Q\in\mathcal{D}_{\Delta}(\mathcal{P}). Our goal is to show that for many 𝐓𝒯Δ(Q)\mathbf{T}\in\mathcal{T}_{\Delta}(Q), the Δ1\Delta^{-1}-rescaled version of πσ(𝐓)((𝒫Q))\pi_{\sigma(\mathbf{T})}(\cup(\mathcal{P}\cap Q)) contains a (Δ,s,0)(\Delta,s,0) and (Δ,κ,k)(\Delta,\kappa^{\prime},k)-set for some κ>0\kappa^{\prime}>0. This is the content of the next Proposition 4.3, which is a higher-dimensional extension of Lemma A.6 of [18]. The proposition encodes the following principle: If we have a set of orthogonal projections in Gr(d,d1)\mathrm{Gr}(d,d-1) (which we view as Sd1S^{d-1}) that don’t concentrate around kk-planes, and we have a tt-dimensional set XX with t>kt>k, then many projections of XX will not concentrate around kk-planes.

Proposition 4.3.

Let 0max(s,k)<td0\leq\max(s,k)<t\leq d, κ>0\kappa>0, and 𝐀,𝐁>0\mathbf{A},\mathbf{B}>0. Let 𝒫\mathcal{P} be a (Δ,t,Δ𝐀ε)(\Delta,t,\Delta^{-\mathbf{A}\varepsilon})-set in [0,1)d[0,1)^{d}, and let ΓSd1\Gamma\subset S^{d-1} be a (Δ,s,Δ𝐀ε,0)(\Delta,s,\Delta^{-\mathbf{A}\varepsilon},0)-set and (Δ,κ,Δ𝐀ε,k)(\Delta,\kappa,\Delta^{-\mathbf{A}\varepsilon},k)-set. There exists a subset ΣΓ\Sigma\subset\Gamma with |Σ|12|Γ||\Sigma|\geq\frac{1}{2}|\Gamma| such that the following holds for all σΣ\sigma\in\Sigma: if 𝒫𝒫\mathcal{P}^{\prime}\subset\mathcal{P} is an arbitrary subset of cardinality |𝒫|Δ𝐁ε|𝒫||\mathcal{P}^{\prime}|\geq\Delta^{\mathbf{B}\varepsilon}|\mathcal{P}|, then πσ(𝒫)\pi_{\sigma}(\mathcal{P}^{\prime}) contains a (Δ,1k+1min(tk2,κ),Δ𝐂(𝐀+𝐁)ε,k)(\Delta,\frac{1}{k+1}\min(\frac{t-k}{2},\kappa),\Delta^{-\mathbf{C}(\mathbf{A}+\mathbf{B})\varepsilon},k) and (Δ,s,Δ𝐂(𝐀+𝐁),0)(\Delta,s,\Delta^{-\mathbf{C}(\mathbf{A}+\mathbf{B})},0)-set, where 𝐂1\mathbf{C}\geq 1 is absolute depending on kk.

Proof.

We will use a variation of the energy argument due to Kaufman [13] in the form used to prove [18, Lemma A.6]. An alternate proof can follow [10, Lemma 27], but this approach would give weaker bounds.

Let μ\mu be the Δ\Delta-discretized probability measure corresponding to 𝒫\mathcal{P},

μ:=1|𝒫|q𝒫d|qΔd,\mu:=\frac{1}{|\mathcal{P}|}\sum_{q\in\mathcal{P}}\frac{\mathcal{L}^{d}|_{q}}{\Delta^{d}},

where d\mathcal{L}^{d} is dd-dimensional Lebesgue measure. Since 𝒫\mathcal{P} is a (Δ,t,Δ𝐀ε)(\Delta,t,\Delta^{-\mathbf{A}\varepsilon})-set, we have μ(B(x,r))rt\mu(B(x,r))\lessapprox r^{t} for all r>δr>\delta, and it’s also true for r<δr<\delta since μ\mu behaves like Lebesgue measure at small scales. We will choose a uniformly random σΓ\sigma\in\Gamma and consider what happens to the energy of μ\mu under projection by σ\sigma. By linearity of expectation and definition of energy,

Es,1:=𝔼σ[Is,1Δ(πσμ)]=𝔼σ[(|πσ(x0x1)|+δ)s]𝑑μ(x0)𝑑μ(x1).E_{s,1}:=\mathbb{E}_{\sigma}[I_{s,1}^{\Delta}(\pi_{\sigma}\mu)]=\int\mathbb{E}_{\sigma}[(|\pi_{\sigma}(x_{0}-x_{1})|+\delta)^{-s}]\,d\mu(x_{0})d\mu(x_{1}).

Since Γ\Gamma is a (Δ,s)(\Delta,s)-set, we have 𝔼σ[(|πσ(x0x1)|+Δ)s](logΔ1)Δ𝐀ε|x0x1|s\mathbb{E}_{\sigma}[(|\pi_{\sigma}(x_{0}-x_{1})|+\Delta)^{-s}]\lesssim(\log\Delta^{-1})\cdot\Delta^{-\mathbf{A}\varepsilon}|x_{0}-x_{1}|^{-s} (c.f. [13]), and so Es,1Is,10(μ)1E_{s,1}\lessapprox I_{s,1}^{0}(\mu)\lessapprox 1 by Lemma 2.30(a) and s<ts<t.

Analogously, we have (let β=min(κ,tk2)\beta=\min(\kappa,\frac{t-k}{2})):

Eβ,k+1:=𝔼σ[Iβ,k+1Δ(πσμ)]=𝔼σ[(|i=1k+1πσ(x0xi)|+Δ)β]𝑑μ(x0)𝑑μ(xk+1).E_{\beta,k+1}:=\mathbb{E}_{\sigma}[I_{\beta,k+1}^{\Delta}(\pi_{\sigma}\mu)]=\int\mathbb{E}_{\sigma}\left[\left(\left|\bigwedge_{i=1}^{k+1}\pi_{\sigma}(x_{0}-x_{i})\right|+\Delta\right)^{-\beta}\right]\,d\mu(x_{0})\cdots d\mu(x_{k+1}).

Observe that

|i=1k+1πσ(x0xi)|=|σi=1k+1πσ(x0xi)|=|σi=1k+1(x0xi)|=|i=1k+1(x0xi)|ρ,\left|\bigwedge_{i=1}^{k+1}\pi_{\sigma}(x_{0}-x_{i})\right|=\left|\sigma\wedge\bigwedge_{i=1}^{k+1}\pi_{\sigma}(x_{0}-x_{i})\right|=\left|\sigma\wedge\bigwedge_{i=1}^{k+1}(x_{0}-x_{i})\right|=\left|\bigwedge_{i=1}^{k+1}(x_{0}-x_{i})\right|\cdot\rho,

where ρ\rho is the distance from σ\sigma to the plane spanned by x1x0x_{1}-x_{0} through xk+1x0x_{k+1}-x_{0}. (The first equality follows since σ\sigma is orthogonal to each πσ(x0xi)\pi_{\sigma}(x_{0}-x_{i}). The second equality follows since \wedge is multilinear and σσ=0\sigma\wedge\sigma=0. The third equality follows by the geometric definition of wedge product as a volume of a parallelepiped.) Thus, since Γ\Gamma is a (Δ,κ,k)(\Delta,\kappa,k)-set and βκ\beta\leq\kappa, we have

𝔼σ[(|i=1k+1πσ(x0xi)|+Δ)β]\displaystyle\mathbb{E}_{\sigma}\left[\left(\left|\bigwedge_{i=1}^{k+1}\pi_{\sigma}(x_{0}-x_{i})\right|+\Delta\right)^{-\beta}\right] ρ=2n(Δ,1)Δ𝐀ερκβ|i=1k+1(x0xi)|β\displaystyle\lesssim\sum_{\rho=2^{-n}\in(\Delta,1)}\Delta^{-\mathbf{A}\varepsilon}\cdot\rho^{\kappa-\beta}\left|\bigwedge_{i=1}^{k+1}(x_{0}-x_{i})\right|^{-\beta}
(logΔ1)Δ𝐀ε|i=1k+1(x0xi)|β,\displaystyle\lesssim(\log\Delta^{-1})\cdot\Delta^{-\mathbf{A}\varepsilon}\left|\bigwedge_{i=1}^{k+1}(x_{0}-x_{i})\right|^{-\beta},

and so Eβ,k+1Iβ,k+10(μ)1E_{\beta,k+1}\lessapprox I_{\beta,k+1}^{0}(\mu)\lessapprox 1 by Lemma 2.30(a) and β<tk\beta<t-k.

Consequently, by Markov’s inequality we can find ΣΓ\Sigma\subset\Gamma with |Σ|12|Γ||\Sigma|\geq\frac{1}{2}|\Gamma| such that for each σΣ\sigma\in\Sigma, we have Is,1Δ(πσμ)Δ2C1𝐀εI_{s,1}^{\Delta}(\pi_{\sigma}\mu)\leq\Delta^{-2C_{1}\mathbf{A}\varepsilon} and Iβ,k+1Δ(πσμ)Δ2C1𝐀εI_{\beta,k+1}^{\Delta}(\pi_{\sigma}\mu)\leq\Delta^{-2C_{1}\mathbf{A}\varepsilon}. For any 𝒫𝒫\mathcal{P}^{\prime}\subset\mathcal{P} with |𝒫|Δ𝐁ε|𝒫||\mathcal{P}^{\prime}|\geq\Delta^{\mathbf{B}\varepsilon}|\mathcal{P}|, we have Is,1Δ(πσμ𝒫)Δ(2C1𝐀+2𝐁)εI_{s,1}^{\Delta}(\pi_{\sigma}\mu_{\mathcal{P}^{\prime}})\leq\Delta^{-(2C_{1}\mathbf{A}+2\mathbf{B})\varepsilon} and Iβ,k+1Δ(πσμ𝒫)Δ(2C1𝐀+(k+2)𝐁)εI_{\beta,k+1}^{\Delta}(\pi_{\sigma}\mu_{\mathcal{P}^{\prime}})\leq\Delta^{-(2C_{1}\mathbf{A}+(k+2)\mathbf{B})\varepsilon}, where μ𝒫=1μ(𝒫)μ|𝒫\mu_{\mathcal{P}^{\prime}}=\frac{1}{\mu(\mathcal{P}^{\prime})}\mu|_{\mathcal{P}^{\prime}} is the renormalized restriction of μ\mu to 𝒫\mathcal{P}^{\prime}. Then Lemma 2.30(b) gives the desired conclusion. ∎

4.3 Finding a special Δ\Delta-tube

This subsection is based on Section A.4 of [18].

Apply Proposition 4.3 to SQ(𝒫Q)S_{Q}(\mathcal{P}\cap Q), which is a (Δ,t)(\Delta,t)-set using 1 and the fact that 𝒫\mathcal{P} is a (Δ,t)(\Delta,t)-set. Define

𝒯Δπ(Q)={𝐓𝒯Δ(Q):σ(𝐓)Σ(Q)},Q𝒬,\mathcal{T}_{\Delta}^{\pi}(Q)=\{\mathbf{T}\in\mathcal{T}_{\Delta}(Q):\sigma(\mathbf{T})\in\Sigma(Q)\},\qquad Q\in\mathcal{Q},

where Σ(Q)\Sigma(Q) is the set of good directions of cardinality |Σ(Q)|14|σ(Q)||𝒯Δ(Q)||\Sigma(Q)|\geq\frac{1}{4}|\sigma(Q)|\sim|\mathcal{T}_{\Delta}(Q)| (since for a given direction, there are 1\sim 1 many Δ\Delta-tubes in that direction that intersect QQ). Then 𝒯Δπ(Q),Q𝒬\mathcal{T}_{\Delta}^{\pi}(Q),Q\in\mathcal{Q} remain (Δ,s)(\Delta,s)-sets of cardinality Δs\approx\Delta^{-s}, and so the properties 1-5 remain valid upon replacing 𝒯Δ(Q)\mathcal{T}_{\Delta}(Q) with 𝒯Δ(Q)\mathcal{T}_{\Delta}(Q). (We leave 𝒯Δ\mathcal{T}_{\Delta} unchanged, so only 3 and 5 are affected.) Thus, 𝒫Q\mathcal{P}\cap Q for Q𝒬Q\in\mathcal{Q} and their large subsets have nice projections in the sense of Proposition 4.3 in every direction orthogonal to the tubes 𝐓𝒯Δπ(Q)\mathbf{T}\in\mathcal{T}_{\Delta}^{\pi}(Q). We keep the symbol “π\pi” as a reminder of this fact.

The next goal is to find a tube 𝐓0\mathbf{T}_{0} with the following properties:

  1. (P1)

    The set {Q𝒬:𝐓0𝒯Δπ(Q)}\{Q\in\mathcal{Q}:\mathbf{T}_{0}\in\mathcal{T}_{\Delta}^{\pi}(Q)\} contains a (Δ,ts)(\Delta,t-s)-subset, which we denote 𝐓0(𝒬)\mathbf{T}_{0}(\mathcal{Q}).

  2. (P2)

    |𝒯𝐓0|Δ2s|\mathcal{T}\cap\mathbf{T}_{0}|\lessapprox\Delta^{-2s}.

  3. (P3)

    For each Q𝐓0(𝒬)Q\in\mathbf{T}_{0}(\mathcal{Q}), there exists a subset 𝒫Q𝒫Q\mathcal{P}_{Q}\subset\mathcal{P}\cap Q such that

    |𝒫Q|Δt and |𝒯(p)𝐓0|Δs for all p𝒫Q.|\mathcal{P}_{Q}|\approx\Delta^{-t}\text{ and }|\mathcal{T}(p)\cap\mathbf{T}_{0}|\approx\Delta^{-s}\text{ for all }p\in\mathcal{P}_{Q}.
  4. (P4)

    Let σ\sigma be the direction of 𝐓\mathbf{T}. Then πσ(SQ(𝒫Q))\pi_{\sigma}(S_{Q}(\mathcal{P}_{Q})) contains a (Δ,κ,k)(\Delta,\kappa^{\prime},k)-set with cardinality Δs\gtrapprox\Delta^{-s}, where κ:=1k+1min(tk2,κ)\kappa^{\prime}:=\frac{1}{k+1}\min(\frac{t-k}{2},\kappa).

To get 1- 3, we will mostly follow Section A.4 of [18]. (We have used the fact that 𝒯\mathcal{T} is a (Δ,κ,Δ𝐀ε,k)(\Delta,\kappa,\Delta^{-\mathbf{A}\varepsilon},k)-set, by converting it into ball concentration near (k+1)(k+1)-planes in Proposition 4.3; the rest of the argument will only use the fact that 𝒯\mathcal{T} is a (Δ,s,Δ𝐀ε,0)(\Delta,s,\Delta^{-\mathbf{A}\varepsilon},0)-set.) First, we refine the sets 𝒬\mathcal{Q} and 𝒯Δπ(Q)\mathcal{T}_{\Delta}^{\pi}(Q) further to ensure that the family {Q𝒬:𝐓𝒯Δπ(Q)}\{Q\in\mathcal{Q}:\mathbf{T}\in\mathcal{T}_{\Delta}^{\pi}(Q)\} will be (Δ,ts)(\Delta,t-s)-sets for 𝐓𝒯Δ\mathbf{T}\in\mathcal{T}_{\Delta}. Indeed, we have

𝐓𝒯ΔQ,Q𝒬QQ𝟙𝒯Δπ(Q)𝒯Δπ(Q)(𝐓)d(Q,Q)ts\displaystyle\sum_{\mathbf{T}\in\mathcal{T}_{\Delta}}\sum_{\begin{subarray}{c}Q,Q^{\prime}\in\mathcal{Q}\\ Q\neq Q^{\prime}\end{subarray}}\frac{\mathbbm{1}_{\mathcal{T}_{\Delta}^{\pi}(Q)\cap\mathcal{T}_{\Delta}^{\pi}(Q^{\prime})}(\mathbf{T})}{d(Q,Q^{\prime})^{t-s}} =Q,Q𝒬,QQ|𝒯Δπ(Q)𝒯Δπ(Q)|d(Q,Q)ts\displaystyle=\sum_{Q,Q^{\prime}\in\mathcal{Q},Q\neq Q^{\prime}}\frac{|\mathcal{T}_{\Delta}^{\pi}(Q)\cap\mathcal{T}_{\Delta}^{\pi}(Q^{\prime})|}{d(Q,Q^{\prime})^{t-s}}
Q,Q𝒬,QQ1d(Q,Q)tΔ2t.\displaystyle\lessapprox\sum_{Q,Q^{\prime}\in\mathcal{Q},Q\neq Q^{\prime}}\frac{1}{d(Q,Q^{\prime})^{t}}\lessapprox\Delta^{-2t}.

The first \lessapprox inequality uses the fact that 𝒯Δπ(Q)\mathcal{T}_{\Delta}^{\pi}(Q) is a (Δ,s)(\Delta,s)-set of tubes with |𝒯Δπ(Q)|Δs|\mathcal{T}_{\Delta}^{\pi}(Q)|\approx\Delta^{-s}, and the second \lessapprox inequality uses the fact that 𝒬\mathcal{Q} is a (Δ,t)(\Delta,t)-set with |𝒬|Δt|\mathcal{Q}|\approx\Delta^{-t}.

Thus, by Markov’s inequality, for a fixed absolute large constant C1C\geq 1, we have

Q,Q𝒬QQ𝟙𝒯Δπ(Q)𝒯Δπ(Q)(𝐓)d(Q,Q)tsΔCε+2(st)\sum_{\begin{subarray}{c}Q,Q^{\prime}\in\mathcal{Q}\\ Q\neq Q^{\prime}\end{subarray}}\frac{\mathbbm{1}_{\mathcal{T}_{\Delta}^{\pi}(Q)\cap\mathcal{T}_{\Delta}^{\pi}(Q^{\prime})}(\mathbf{T})}{d(Q,Q^{\prime})^{t-s}}\geq\Delta^{-C\varepsilon+2(s-t)} (4.9)

can only hold for ΔCε2s\lessapprox\Delta^{C\varepsilon-2s} many tubes 𝐓𝒯Δ\mathbf{T}\in\mathcal{T}_{\Delta}.

Claim 2. If C1C\geq 1 is sufficiently large, then there exists a subset 𝒬¯𝒬\overline{\mathcal{Q}}\subset\mathcal{Q} with |𝒬¯|12|𝒬||\overline{\mathcal{Q}}|\geq\frac{1}{2}|\mathcal{Q}| such that for all Q0𝒬¯Q_{0}\in\overline{\mathcal{Q}}, at most half of the tubes 𝐓𝒯Δπ(Q0)\mathbf{T}\in\mathcal{T}_{\Delta}^{\pi}(Q_{0}) satisfy (4.9).

Proof. Suppose this is not true: there exists a set 𝒬bad\mathcal{Q}_{\text{bad}} such that for Q0𝒬Q_{0}\in\mathcal{Q}, at least 12|𝒯Δπ(Q0)|\frac{1}{2}|\mathcal{T}_{\Delta}^{\pi}(Q_{0})| many tubes 𝐓𝒯Δπ(Q0)\mathbf{T}\in\mathcal{T}_{\Delta}^{\pi}(Q_{0}) satisfy (4.9). Then apply Corollary 2.10 to 𝒬bad\mathcal{Q}_{\text{bad}} and the bad parts of 𝒯Δπ(Q0)\mathcal{T}_{\Delta}^{\pi}(Q_{0}), which are still (Δ,s(\Delta,s)-sets. By Corollary 2.10, we have Δ2s\gtrapprox\Delta^{-2s} many Δ\Delta-tubes in 𝒯Δ\mathcal{T}_{\Delta} that satisfy (4.9). But we observed before that (4.9) only holds for ΔCε2s\lessapprox\Delta^{C\varepsilon-2s} many tubes 𝐓𝒯Δ\mathbf{T}\in\mathcal{T}_{\Delta}. By choosing CC large enough (and δ\delta small enough), we obtain a contradiction. ∎

In what follows, the CC in Claim 2 will be absorbed into the \lessapprox notation. Replace 𝒬\mathcal{Q} by 𝒬¯\overline{\mathcal{Q}} and 𝒯Δπ(Q)\mathcal{T}_{\Delta}^{\pi}(Q) by their good subsets without changing notation. All of the properties 1-5 remain valid, and

Q,Q𝒬QQ𝟙𝒯Δπ(Q)𝒯Δπ(Q)(𝐓)d(Q,Q)tsΔ2(st),𝐓𝒯Δπ(Q0),Q0𝒬.\sum_{\begin{subarray}{c}Q,Q^{\prime}\in\mathcal{Q}\\ Q\neq Q^{\prime}\end{subarray}}\frac{\mathbbm{1}_{\mathcal{T}_{\Delta}^{\pi}(Q)\cap\mathcal{T}_{\Delta}^{\pi}(Q^{\prime})}(\mathbf{T})}{d(Q,Q^{\prime})^{t-s}}\lessapprox\Delta^{2(s-t)},\qquad\mathbf{T}\in\mathcal{T}_{\Delta}^{\pi}(Q_{0}),Q_{0}\in\mathcal{Q}. (4.10)

Now, we will find 𝐓0𝒯Δ\mathbf{T}_{0}\in\mathcal{T}_{\Delta} satisfying

|𝐓0(𝒬)|:=|{Q𝒬:𝐓0𝒯Δπ(Q)}|Δst.|\mathbf{T}_{0}(\mathcal{Q})|:=|\{Q\in\mathcal{Q}:\mathbf{T}_{0}\in\mathcal{T}_{\Delta}^{\pi}(Q)\}|\gtrapprox\Delta^{s-t}. (4.11)

Indeed, the average tube works, because of the following: since |𝒯Δ|Δ2s,|𝒬|Δt|\mathcal{T}_{\Delta}|\approx\Delta^{-2s},|\mathcal{Q}|\approx\Delta^{-t}, and |𝒯Δπ(Q)|Δs|\mathcal{T}_{\Delta}^{\pi}(Q)|\approx\Delta^{-s} (by 4, 1, 3 respectively), we have

1|𝒯Δ|𝐓𝒯Δ|{Q𝒬:𝐓0𝒯Δπ(Q)}|=1|𝒯Δ|Q𝒬|𝒯Δπ(Q)||𝒬|ΔsΔ2sΔst.\frac{1}{|\mathcal{T}_{\Delta}|}\sum_{\mathbf{T}\in\mathcal{T}_{\Delta}}|\{Q\in\mathcal{Q}:\mathbf{T}_{0}\in\mathcal{T}_{\Delta}^{\pi}(Q)\}|=\frac{1}{|\mathcal{T}_{\Delta}|}\sum_{Q\in\mathcal{Q}}|\mathcal{T}_{\Delta}^{\pi}(Q)|\approx\frac{|\mathcal{Q}|\cdot\Delta^{-s}}{\Delta^{-2s}}\approx\Delta^{s-t}.

Now, we show that using (4.10) and (4.11), the family 𝐓0(𝒬){Q𝒬:Q𝐓0}\mathbf{T}_{0}(\mathcal{Q})\subset\{Q\in\mathcal{Q}:Q\cap\mathbf{T}_{0}\neq\emptyset\} contains a (Δ,ts)(\Delta,t-s)-set, which proves item 1. Indeed, rewrite (4.10) as

Q,Q𝐓0(𝒬)QQ1d(Q,Q)tsΔ2(st).\sum_{\begin{subarray}{c}Q,Q^{\prime}\in\mathbf{T}_{0}(\mathcal{Q})\\ Q\neq Q^{\prime}\end{subarray}}\frac{1}{d(Q,Q^{\prime})^{t-s}}\lessapprox\Delta^{2(s-t)}. (4.12)

Let

𝐓0(𝒬):={Q𝐓0(𝒬):Q𝐓0(𝒬){Q}d(Q,Q)stΔstCε}.\mathbf{T}_{0}^{\prime}(\mathcal{Q}):=\{Q\in\mathbf{T}_{0}(\mathcal{Q}):\sum_{Q^{\prime}\in\mathbf{T}_{0}(\mathcal{Q})\setminus\{Q\}}d(Q,Q^{\prime})^{s-t}\leq\Delta^{s-t-C\varepsilon}\}. (4.13)

By Markov’s inequality on (4.12), we have |𝐓0(𝒬)𝐓0(𝒬)|Δst+Cε|\mathbf{T}_{0}(\mathcal{Q})\setminus\mathbf{T}_{0}^{\prime}(\mathcal{Q})|\lessapprox\Delta^{s-t+C\varepsilon}. Hence, if CC is chosen large enough, we have by (4.11), |𝐓0(𝒬)|12|𝐓0(𝒬)|Δst|\mathbf{T}_{0}^{\prime}(\mathcal{Q})|\geq\frac{1}{2}|\mathbf{T}_{0}(\mathcal{Q})|\gtrapprox\Delta^{s-t}. By Markov’s inequality on (4.13), we have that for all Q𝐓0(𝒬)Q\in\mathbf{T}_{0}^{\prime}(\mathcal{Q}) and r(δ,1)r\in(\delta,1),

|{Q𝐓0(𝒬):d(Q,Q)r}|ΔstCεrts.|\{Q^{\prime}\in\mathbf{T}_{0}(\mathcal{Q}):d(Q,Q^{\prime})\leq r\}|\leq\Delta^{s-t-C\varepsilon}r^{t-s}.

Thus, 𝐓0(𝒬)\mathbf{T}_{0}^{\prime}(\mathcal{Q}) is a (Δ,ts)(\Delta,t-s)-set, which proves 1.

To get 2, we use 2.

|𝒯𝐓0|δs=Δ2s.|\mathcal{T}\cap\mathbf{T}_{0}|\lessapprox\delta^{-s}=\Delta^{-2s}.

By 5, we have

|{(p,T)(𝒫Q)×𝒯:T𝒯(p)𝐓0}|Δst.|\{(p,T)\in(\mathcal{P}\cap Q)\times\mathcal{T}:T\in\mathcal{T}(p)\cap\mathbf{T}_{0}\}|\gtrapprox\Delta^{-s-t}. (4.14)

Fix Q𝐓0(𝒬)Q\in\mathbf{T}_{0}(\mathcal{Q}). Since |𝒫Q|Δt|\mathcal{P}\cap Q|\approx\Delta^{-t} by 1 and |𝒯(p)𝐓0|Δs|\mathcal{T}(p)\cap\mathbf{T}_{0}|\lessapprox\Delta^{-s} since 𝒯(p)\mathcal{T}(p) is a (δ,s)(\delta,s)-set, we use (4.14) to find a subset 𝒫Q𝒫Q\mathcal{P}_{Q}\subset\mathcal{P}\cap Q with

|𝒫Q||𝒫Q|Δt and |𝒯(p)𝐓0|Δs for all p𝒫Q.|\mathcal{P}_{Q}|\approx|\mathcal{P}\cap Q|\approx\Delta^{-t}\text{ and }|\mathcal{T}(p)\cap\mathbf{T}_{0}|\approx\Delta^{-s}\text{ for all }p\in\mathcal{P}_{Q}.

This verifies 3. Finally, we get 4 by |𝒫Q|Δ𝐁ε|𝒫Q||\mathcal{P}_{Q}|\geq\Delta^{\mathbf{B}\varepsilon}|\mathcal{P}\cap Q| for some constant 𝐁1\mathbf{B}\geq 1 and Proposition 4.3.

4.4 Product-like structure

This subsection is based on Section A.6 of [18].

Our goal is to find a product-type structure and apply Proposition 3.1. Choose coordinates such that the yy-axis is in the direction of 𝐓0\mathbf{T}_{0}, and let π(𝐱,y):=𝐱\Rd1\pi(\mathbf{x},y):=\mathbf{x}\in\R^{d-1} denote the orthogonal projection to the orthogonal complement of the yy-axis. Define the function Δ1(𝐱,y)=(Δ1𝐱,y)\Delta^{-1}(\mathbf{x},y)=(\Delta^{-1}\mathbf{x},y). If T𝐓0T\in\mathbf{T}_{0}, then Δ1T\Delta^{-1}T is roughly a Δ\Delta-tube: it is contained in some CΔC\Delta-tube and contains a cΔc\Delta-tube for some universal constants c,C>0c,C>0. This technicality will not cause issues in what follows.

For each Q𝐓0(𝒬)Q\in\mathbf{T}_{0}(\mathcal{Q}), let 𝐲QΔ[0,1)\mathbf{y}_{Q}\in\Delta\cdot\mathbb{Z}\cap[0,1) be a point such that the plane y=𝐲Qy=\mathbf{y}_{Q} intersects QQ. By 1, we know that 𝐘={𝐲Q:Q𝐓0(𝒬)}\mathbf{Y}=\{\mathbf{y}_{Q}:Q\in\mathbf{T}_{0}(\mathcal{Q})\} is a (Δ,ts)(\Delta,t-s)-set. By 4, we know that for each 𝐲𝐘\mathbf{y}\in\mathbf{Y} that π(Δ1(𝒫Q))\pi(\Delta^{-1}(\mathcal{P}\cap Q)) contains a (Δ,κ,k)(\Delta,\kappa^{\prime},k)-set 𝐗𝐲\mathbf{X}_{\mathbf{y}}^{\prime} with cardinality Δs\gtrapprox\Delta^{-s}. Let 𝐗𝐲(Δ)d[0,1]d\mathbf{X}_{\mathbf{y}}\subset(\Delta\cdot\mathbb{Z})^{d}\cap[0,1]^{d} that is 𝐗𝐲\mathbf{X}_{\mathbf{y}}^{\prime} rounded to the nearest multiple of Δ\Delta.

Now, let L=(Δ)B(0,Δ(d+1))L=(\Delta\cdot\mathbb{Z})\cap B(0,\Delta(\sqrt{d}+1)) and 𝒯(𝐙)={σ(T)+x:T𝒯𝐓0,xL}\mathcal{T}(\mathbf{Z})=\{\sigma(T)+x:T\in\mathcal{T}\cap\mathbf{T}_{0},x\in L\}. Clearly, |𝒯(𝐙)|d|𝒯𝐓0|Δ2s|\mathcal{T}(\mathbf{Z})|\lesssim_{d}|\mathcal{T}\cap\mathbf{T}_{0}|\lessapprox\Delta^{-2s} by 2. On the other hand, we show that |𝒯(𝐳)|:=|{𝐓𝒯(𝐙):𝐳𝐓}|Δs|\mathcal{T}(\mathbf{z})|:=|\{\mathbf{T}\in\mathcal{T}(\mathbf{Z}):\mathbf{z}\in\mathbf{T}\}|\gtrapprox\Delta^{-s} for any (𝐱,𝐲)𝐙(\mathbf{x},\mathbf{y})\in\mathbf{Z}. This follows since 𝐳=(𝐱,𝐲Q)\mathbf{z}=(\mathbf{x},\mathbf{y}_{Q}) for some QQ and 𝐱𝐗𝐲\mathbf{x}\in\mathbf{X}_{\mathbf{y}}. Let p𝒫Qp\in\mathcal{P}_{Q} such that d(π(Δ1p),𝐱)Δd(\pi(\Delta^{-1}p),\mathbf{x})\leq\Delta. We know d((π(Δ1p),𝐲Q),Δ1p)Δd((\pi(\Delta^{-1}p),\mathbf{y}_{Q}),\Delta^{-1}p)\leq\Delta since QQ has diameter Δ\Delta, so by triangle inequality, we have d(Δ1p,𝐳)(d+1)Δd(\Delta^{-1}p,\mathbf{z})\leq(\sqrt{d}+1)\Delta. Thus, 𝒯(𝐳)\mathcal{T}(\mathbf{z}) contains {σ(T)+x:T𝒯(p)𝐓0}\{\sigma(T)+x:T\in\mathcal{T}(p)\cap\mathbf{T}_{0}\} for some suitable xLx\in L. By 3, we get the desired cardinality estimate |𝒯(𝐳)|Δs|\mathcal{T}(\mathbf{z})|\approx\Delta^{-s}.

Finally, we apply Proposition 3.1 to the sets 𝐙\mathbf{Z} and 𝒯(𝐙)\mathcal{T}(\mathbf{Z}) to obtain a contradiction if ε>0\varepsilon>0 is sufficiently small. This proves Theorem 4.2.

5 Improved incidence estimates for general sets

In this section, we will prove the following refinement of Theorem 1.8, following Sections 7-9 of [18].

Theorem 5.1.

For any 0k<d10\leq k<d-1, 0s<k+10\leq s<k+1, s<tds<t\leq d, κ>0\kappa>0, there exist ε(s,t,κ,k,d)>0\varepsilon(s,t,\kappa,k,d)>0 and η(s,t,κ,k,d)>0\eta(s,t,\kappa,k,d)>0 such that the following holds for all small enough δ2\delta\in 2^{-\mathbb{N}}, depending only on s,t,κ,k,ds,t,\kappa,k,d. Let 𝒫𝒟δ\mathcal{P}\subset\mathcal{D}_{\delta} be a (δ,t,δε)(\delta,t,\delta^{-\varepsilon})-set with 𝒫[0,1)d\cup\mathcal{P}\subset[0,1)^{d}, and let 𝒯𝒯δ\mathcal{T}\subset\mathcal{T}^{\delta} be a family of δ\delta-tubes. Assume that for every p𝒫p\in\mathcal{P}, there exists a (δ,s,δλ,0)(\delta,s,\delta^{-\lambda},0) and (δ,κ,δλ,k)(\delta,\kappa,\delta^{-\lambda},k)-set 𝒯(p)𝒯\mathcal{T}(p)\subset\mathcal{T} with |𝒯(p)|=M|\mathcal{T}(p)|=M such that TpT\cap p\neq\emptyset for all T𝒯(p)T\in\mathcal{T}(p). Then |𝒯|Mδsε|\mathcal{T}|\geq M\delta^{-s-\varepsilon}.

The original theorem follows from taking ε=η\varepsilon=\eta and pigeonholing, since M(δs+ε,δd)M\in(\delta^{-s+\varepsilon},\delta^{-d}).

Proof.

Before anything else, we state the dependencies of the parameters: ε0(s,t,κ,k,d)\varepsilon_{0}(s,t,\kappa,k,d), ε(ε0,s,t,κ,k,d),T(ε),τ(s,t,ε),η(ε0,τ)\varepsilon(\varepsilon_{0},s,t,\kappa,k,d),T(\varepsilon),\tau(s,t,\varepsilon),\eta(\varepsilon_{0},\tau).

First, choose T=T(ε)T=T(\varepsilon) such that 2logTTε\frac{2\log T}{T}\leq\varepsilon. By Lemma 2.17 we may find a subset 𝒫𝒫\mathcal{P}^{\prime}\subset\mathcal{P} with |𝒫|δε|𝒫||\mathcal{P}^{\prime}|\geq\delta^{\varepsilon}|\mathcal{P}| that is {2jT}j=1m\{2^{-jT}\}_{j=1}^{m}-uniform for 2mT=δ2^{-mT}=\delta with associated sequence {Nj}j=1m\{N_{j}\}_{j=1}^{m}. Thus, 𝒫\mathcal{P}^{\prime} is a (δ,t,δ2ε)(\delta,t,\delta^{-2\varepsilon})-set. Replacing 𝒫\mathcal{P} with 𝒫\mathcal{P}^{\prime} and ε\varepsilon with ε2\frac{\varepsilon}{2}, we may assume from the start that 𝒫\mathcal{P} is {2jT}j=1m\{2^{-jT}\}_{j=1}^{m}-uniform.

Let ff be the corresponding branching function. Since 𝒫\mathcal{P} is a (δ,t,δε)(\delta,t,\delta^{-\varepsilon})-set, we have f(x)txεmf(x)\geq tx-\varepsilon m for all x[0,m]x\in[0,m].

Let {[cj,dj]}j=1n\{[c_{j},d_{j}]\}_{j=1}^{n} be the intervals from Proposition 2.14 applied with parameters s,t,εs,t,\varepsilon, corresponding to a sequence 0<δ=Δn<Δn1<<Δ1<Δ0=10<\delta=\Delta_{n}<\Delta_{n-1}<\cdots<\Delta_{1}<\Delta_{0}=1. We can partition {0,1,,n1}=𝒮\{0,1,\cdots,n-1\}=\mathcal{S}\cup\mathcal{B}, “structured” and “bad” scales such that:

  • ΔjΔj+1δτ\frac{\Delta_{j}}{\Delta_{j+1}}\geq\delta^{-\tau} for all j𝒮j\in\mathcal{S}, and j(Δj/Δj+1)δε\prod_{j\in\mathcal{B}}(\Delta_{j}/\Delta_{j+1})\leq\delta^{-\varepsilon};

  • For each j𝒮j\in\mathcal{S} and 𝐩𝒟Δj(𝒫)\mathbf{p}\in\mathcal{D}_{\Delta_{j}}(\mathcal{P}), the set 𝒫j:=S𝐩(𝒫𝐩)\mathcal{P}_{j}:=S_{\mathbf{p}}(\mathcal{P}\cap\mathbf{p}) is either

    1. (i)

      an (tj,Δj+1/Δj,(Δj/Δj+1)ε,(Δj/Δj+1)ε)(t_{j},\Delta_{j+1}/\Delta_{j},(\Delta_{j}/\Delta_{j+1})^{\varepsilon},(\Delta_{j}/\Delta_{j+1})^{\varepsilon})-regular set, where tj(s,2)t_{j}\in(s,2);

    2. (ii)

      a (s,Δj+1/Δj,(Δj/Δj+1)ε)(s,\Delta_{j+1}/\Delta_{j},(\Delta_{j}/\Delta_{j+1})^{\varepsilon})-set.

  • jS(Δj/Δj+1)tj|𝒫|j(Δj+1/Δj)d|𝒫|δOs,t,d(ε)\prod_{j\in S}(\Delta_{j}/\Delta_{j+1})^{t_{j}}\geq|\mathcal{P}|\cdot\prod_{j\in\mathcal{B}}(\Delta_{j+1}/\Delta_{j})^{d}\geq|\mathcal{P}|\delta^{O_{s,t,d}(\varepsilon)}.

Apply Proposition 2.14 and ΔjΔj+1δτ\frac{\Delta_{j}}{\Delta_{j+1}}\geq\delta^{-\tau} to get a family of tubes 𝒯p𝒯Δj+1/Δj\mathcal{T}_{\textbf{p}}\subset\mathcal{T}^{\Delta_{j+1}/\Delta_{j}} with the property that (Sp(𝒫p),𝒯p)(S_{\textbf{p}}(\mathcal{P}\cap\textbf{p}),\mathcal{T}_{\textbf{p}}) is a (Δj+1/Δj,s,Cj1,κ,Cj2Mp)(\Delta_{j+1}/\Delta_{j},s,C_{j}^{1},\kappa,C_{j}^{2}M_{\textbf{p}})-nice configuration for some Cj1,Cj2δ(Δj+1/Δj)τ1ηC_{j}^{1},C_{j}^{2}\lessapprox_{\delta}(\Delta_{j+1}/\Delta_{j})^{-\tau^{-1}\eta} and

|𝒯0|Mδj=0N1|𝒯pj|Mpj.\frac{|\mathcal{T}_{0}|}{M}\gtrapprox_{\delta}\prod_{j=0}^{N-1}\frac{|\mathcal{T}_{\textbf{p}_{j}}|}{M_{\textbf{p}_{j}}}.

Let 𝒮1={jS:tjs+t2}\mathcal{S}_{1}=\{j\in S:t_{j}\geq\frac{s+t}{2}\} and 𝒮2=𝒮𝒮1\mathcal{S}_{2}=\mathcal{S}\setminus\mathcal{S}_{1}. Then

jS1(Δj/Δj+1)tj|𝒫|δOs,t,d(ε)jS2(Δj/Δj+1)s+t2δts2+Os,t,d(ε).\prod_{j\in S_{1}}(\Delta_{j}/\Delta_{j+1})^{t_{j}}\geq|\mathcal{P}|\delta^{O_{s,t,d}(\varepsilon)}\prod_{j\in S_{2}}(\Delta_{j}/\Delta_{j+1})^{-\frac{s+t}{2}}\geq\delta^{\frac{t-s}{2}+O_{s,t,d}(\varepsilon)}.

For j𝒮1j\in\mathcal{S}_{1} we apply Theorem 4.2 with parameters s,s+t2s,\frac{s+t}{2}, and for j𝒮2j\in\mathcal{S}_{2} we apply Corollary 2.10. If ε0(s,t,κ,k,d)\varepsilon_{0}(s,t,\kappa,k,d) is the η\eta from Theorem 4.2, then for τ1η<ε0\tau^{-1}\eta<\varepsilon_{0}, we get

|𝒯0|Mδj𝒮1(ΔjΔj+1)sε0j𝒮2(ΔjΔj+1)s+O(ε)δs(1ε)(ts2+Os,t,d(ε))ε0+O(ε)δsε\frac{|\mathcal{T}_{0}|}{M}\gtrapprox_{\delta}\prod_{j\in\mathcal{S}_{1}}\left(\frac{\Delta_{j}}{\Delta_{j+1}}\right)^{-s-\varepsilon_{0}}\cdot\prod_{j\in\mathcal{S}_{2}}\left(\frac{\Delta_{j}}{\Delta_{j+1}}\right)^{-s+O(\varepsilon)}\geq\delta^{-s(1-\varepsilon)-(\frac{t-s}{2}+O_{s,t,d}(\varepsilon))\varepsilon_{0}+O(\varepsilon)}\geq\delta^{-s-\varepsilon}

as long as ε\varepsilon is taken small enough in terms of ε0,s,t,d\varepsilon_{0},s,t,d. ∎

6 Sets contained in an (r0,k)(r_{0},k)-plate

We restate Theorem 1.11.

Theorem 6.1.

For any 0k<d10\leq k<d-1, 0s<k+10\leq s<k+1, max(s,k)<td\max(s,k)<t\leq d, κ>0\kappa>0, r01r_{0}\leq 1, there exists ε(s,t,κ,k,d)>0\varepsilon(s,t,\kappa,k,d)>0 such that the following holds for all small enough δ/r02(0,δ0)\delta/r_{0}\in 2^{-\mathbb{N}}\cap(0,\delta_{0}), with δ0\delta_{0} depending only on s,t,κ,k,ds,t,\kappa,k,d. Let HH be a (r0,k+1)(r_{0},k+1)-plate, 𝒫𝒟δH\mathcal{P}\subset\mathcal{D}_{\delta}\cap H be a (δ,t,(δ/r0)ε)(\delta,t,(\delta/r_{0})^{-\varepsilon})-set with 𝒫[0,1)d\cup\mathcal{P}\subset[0,1)^{d}, and let 𝒯𝒯δH\mathcal{T}\subset\mathcal{T}^{\delta}\cap H be a family of δ\delta-tubes. Assume that for every p𝒫p\in\mathcal{P}, there exists a set 𝒯(p)𝒯\mathcal{T}(p)\subset\mathcal{T} such that:

  • TpT\cap p\neq\emptyset for all T𝒯(p)T\in\mathcal{T}(p);

  • 𝒯(p)\mathcal{T}(p) is a (δ,s,(δ/r0)εr0ks,0)(\delta,s,(\delta/r_{0})^{-\varepsilon}r_{0}^{k-s},0)-set down from scale rr;

  • 𝒯(p)\mathcal{T}(p) is a (δ,κ,(δ/r0)εr0κ,k)(\delta,\kappa,(\delta/r_{0})^{-\varepsilon}r_{0}^{-\kappa},k)-set.

Then |𝒯|(δr0)εδ2sr02(sk)|\mathcal{T}|\geq(\frac{\delta}{r_{0}})^{-\varepsilon}\delta^{-2s}r_{0}^{2(s-k)}.

6.1 Multiscale analysis

We will use Theorem 1.8 to prove Theorem 1.11. Let SHS_{H} be the dilation sending HH to [0,1]d[0,1]^{d}. Then 𝒫,𝒯(p)\mathcal{P},\mathcal{T}(p), and 𝒯\mathcal{T} become deformed under SHS_{H}, but they satisfy the following statistics assumptions for r[δr0,1]r\in[\frac{\delta}{r_{0}},1]:

|𝒫SH1(Q)|(δr0)ε|𝒫|rt,\displaystyle|\mathcal{P}\cap S_{H}^{-1}(Q)|\leq\left(\frac{\delta}{r_{0}}\right)^{-\varepsilon}\cdot|\mathcal{P}|\cdot r^{t}, Q𝒟r(\Rd),\displaystyle\qquad Q\in\mathcal{D}_{r}(\R^{d}), (6.15)
|𝒯(p)SH1(𝐓)|(δr0)ε|𝒯(p)|rs,\displaystyle|\mathcal{T}(p)\cap S_{H}^{-1}(\mathbf{T})|\leq\left(\frac{\delta}{r_{0}}\right)^{-\varepsilon}\cdot|\mathcal{T}(p)|\cdot r^{s}, 𝐓r-tube,\displaystyle\qquad\mathbf{T}\quad r\text{-tube}, (6.16)
|𝒯(p)SH1(W)|(δr0)ε|𝒯(p)|rκ,\displaystyle|\mathcal{T}(p)\cap S_{H}^{-1}(W)|\leq\left(\frac{\delta}{r_{0}}\right)^{-\varepsilon}\cdot|\mathcal{T}(p)|\cdot r^{\kappa}, W(r,k+1)-plate.\displaystyle\qquad W\quad(r,k+1)\text{-plate}. (6.17)

To prove (6.15), observe that SH1(Q)S_{H}^{-1}(Q) is contained in an rr-ball, and then we use that 𝒫\mathcal{P} is a (δ,t,(δ/r0)ε,0)(\delta,t,(\delta/r_{0})^{-\varepsilon},0)-set.

To prove (6.16), observe that SH1(𝐓)S_{H}^{-1}(\mathbf{T}) is contained in a box with kk sides of length rr and dkd-k sides of length rr0rr_{0}. This box can be covered by r0k\sim r_{0}^{-k} many rr0rr_{0}-balls. Finally, use that 𝒯(p)\mathcal{T}(p) is a (δ,s,(δ/r0)εr0ks,0)(\delta,s,(\delta/r_{0})^{-\varepsilon}r_{0}^{k-s},0)-set.

To prove (6.17), observe that SH1(W)S_{H}^{-1}(W) is contained in a (rr0,k)(rr_{0},k)-plate.

Using these observations, we obtain the following refinement of Proposition 2.13. We use (δ,s,C1r0ks,κ,C2,M)(\delta,s,C_{1}r_{0}^{k-s},\kappa,C_{2},M)-nice configuration down from scale r0r_{0} to indicate that 𝒯(p)\mathcal{T}(p) is a (δ,s,C1r0ks,0)(\delta,s,C_{1}r_{0}^{k-s},0)-set down from scale r0r_{0}.

Proposition 6.2.

Fix dyadic numbers 0<δ=δr0<Δ10<\delta^{\prime}=\frac{\delta}{r_{0}}<\Delta\leq 1. Let (𝒫0,𝒯0)(\mathcal{P}_{0},\mathcal{T}_{0}) be a (δ,s,C1r0ks,κ,C2,M)(\delta,s,C_{1}r_{0}^{k-s},\kappa,C_{2},M)-nice configuration down from scale r0r_{0}, and assume 𝒫0H\mathcal{P}_{0}\subset H for some (r0,k)(r_{0},k)-plate HH. Then there exist refinements 𝒫𝒫0\mathcal{P}\subset\mathcal{P}_{0}, 𝒯(p)𝒯0(p),p𝒫\mathcal{T}(p)\subset\mathcal{T}_{0}(p),p\in\mathcal{P}, and 𝒯Δ(Q)𝒯Δ\mathcal{T}_{\Delta}(Q)\subset\mathcal{T}^{\Delta} such that denoting 𝒯Δ=Q𝒟Δ(SH(𝒫))𝒯Δ(Q)\mathcal{T}_{\Delta}=\cup_{Q\in\mathcal{D}_{\Delta}(S_{H}(\mathcal{P}))}\mathcal{T}_{\Delta}(Q) and 𝒯=p𝒫𝒯(p)\mathcal{T}=\cup_{p\in\mathcal{P}}\mathcal{T}(p) the following hold:

  1. (i)

    |𝒟Δ(SH(𝒫))|Δ|𝒟Δ(SH(𝒫0))||\mathcal{D}_{\Delta}(S_{H}(\mathcal{P}))|\approx_{\Delta}|\mathcal{D}_{\Delta}(S_{H}(\mathcal{P}_{0}))| and |SH(𝒫)Q|Δ|SH(𝒫0)Q||S_{H}(\mathcal{P})\cap Q|\approx_{\Delta}|S_{H}(\mathcal{P}_{0})\cap Q| for all Q𝒟Δ(𝒫)Q\in\mathcal{D}_{\Delta}(\mathcal{P}).

  2. (ii)

    We have |𝒯𝐓||𝒯0||𝒯Δ||\mathcal{T}\cap\mathbf{T}|\lessapprox\frac{|\mathcal{T}_{0}|}{|\mathcal{T}_{\Delta}|} for all 𝐓𝒯Δ\mathbf{T}\in\mathcal{T}_{\Delta}.

  3. (iii)

    (𝒟Δ(SH(𝒫)),𝒯Δ)(\mathcal{D}_{\Delta}(S_{H}(\mathcal{P})),\mathcal{T}_{\Delta}) is (Δ,s,CΔ1,κ,CΔ2,MΔ)(\Delta,s,C^{1}_{\Delta},\kappa,C^{2}_{\Delta},M_{\Delta})-nice for some CΔ1ΔC1C^{1}_{\Delta}\approx_{\Delta}C_{1}, CΔ2ΔC2C^{2}_{\Delta}\approx_{\Delta}C_{2}, and MΔ1M_{\Delta}\geq 1.

  4. (iv)

    For all 𝐓𝒯Δ(Q)\mathbf{T}\in\mathcal{T}_{\Delta}(Q), we have

    |{(p,T)𝒫×𝒯:T𝒯(p) and TSH1(𝐓)}|ΔM|SH(𝒫)Q||𝒯Δ(Q)|.|\{(p,T)\in\mathcal{P}\times\mathcal{T}:T\in\mathcal{T}(p)\text{ and }T\subset S_{H}^{-1}(\mathbf{T})\}|\gtrapprox_{\Delta}\frac{M\cdot|S_{H}(\mathcal{P})\cap Q|}{|\mathcal{T}_{\Delta}(Q)|}.
  5. (v)

    For each Q𝒟Δ(SH(𝒫2))Q\in\mathcal{D}_{\Delta}(S_{H}(\mathcal{P}_{2})), there exist CQ1ΔC1C^{1}_{Q}\approx_{\Delta}C_{1}, CQ2ΔC2C^{2}_{Q}\approx_{\Delta}C_{2}, MQ1M_{Q}\geq 1, a subset 𝒫Q𝒫Q\mathcal{P}_{Q}\subset\mathcal{P}\cap Q with |𝒫Q|Δ|𝒫Q||\mathcal{P}_{Q}|\gtrapprox_{\Delta}|\mathcal{P}\cap Q| and a family of tubes 𝒯Q𝒯δ/Δ\mathcal{T}_{Q}\subset\mathcal{T}^{\delta/\Delta} such that (SH1SQ(SH(𝒫2)Q),𝒯Q)(S_{H}^{-1}\circ S_{Q}(S_{H}(\mathcal{P}_{2})\cap Q),\mathcal{T}_{Q}) is (δ/Δ,s,CQ1r0ks,κ,CQ2,MQ)(\delta/\Delta,s,C^{1}_{Q}r_{0}^{k-s},\kappa,C^{2}_{Q},M_{Q})-nice down from scale r0r_{0}.

Furthermore, the families 𝒯Q\mathcal{T}_{Q} can be chosen so that

|𝒯0|MΔ|𝒯Δ(𝒯)|MΔ(maxQ𝒟Δ(𝒫2)|𝒯Q|MQ).\frac{|\mathcal{T}_{0}|}{M}\gtrapprox_{\Delta}\frac{|\mathcal{T}^{\Delta}(\mathcal{T})|}{M_{\Delta}}\cdot\left(\max_{Q\in\mathcal{D}_{\Delta}(\mathcal{P}_{2})}\frac{|\mathcal{T}_{Q}|}{M_{Q}}\right). (6.18)
Proof.

The proof will involve many dyadic pigeonholing steps.

Step 1: construct 𝒯Δ(Q)\mathcal{T}_{\Delta}(Q). For a given Q𝒟Δ(SH(𝒫0)):=𝒬0Q\in\mathcal{D}_{\Delta}(S_{H}(\mathcal{P}_{0})):=\mathcal{Q}_{0}, we claim that we can find a subset 𝒫Q𝒫0SH1(Q)\mathcal{P}_{Q}\subset\mathcal{P}_{0}\cap S_{H}^{-1}(Q) with |𝒫Q|Δ|𝒫0SH1(Q)||\mathcal{P}_{Q}|\approx_{\Delta}|\mathcal{P}_{0}\cap S_{H}^{-1}(Q)| and a family of dyadic Δ\Delta-tubes 𝒯¯Δ(Q)\overline{\mathcal{T}}_{\Delta}(Q) intersecting QQ such that the following holds:

  1. (T1)

    𝒯¯Δ(Q)\overline{\mathcal{T}}_{\Delta}(Q) is a (Δ,s,CΔ1,0)(\Delta,s,C_{\Delta}^{1},0)-set and (Δ,κ,CΔ2,k)(\Delta,\kappa,C_{\Delta}^{2},k)-set for some CΔ1,CΔ2ΔC1C_{\Delta}^{1},C_{\Delta}^{2}\approx_{\Delta}C_{1}.

  2. (T2)

    there exists a constant HQΔM|𝒫Q|/|𝒯¯Δ(Q)|H_{Q}\approx_{\Delta}M\cdot|\mathcal{P}_{Q}|/|\overline{\mathcal{T}}_{\Delta}(Q)| such that

    |{(p,T)𝒫Q×𝒯0:T𝒯0(p) and TSH1(𝐓)}|HQ,𝐓𝒯¯Δ(Q).|\{(p,T)\in\mathcal{P}_{Q}\times\mathcal{T}_{0}:T\in\mathcal{T}_{0}(p)\text{ and }T\subset S_{H}^{-1}(\mathbf{T})\}|\gtrsim H_{Q},\qquad\mathbf{T}\in\overline{\mathcal{T}}_{\Delta}(Q).

This claim generalizes [18, Proposition 4.1] and relies on the same dyadic pigeonholing steps; for brevity, we only state these steps and refer the reader to [18] for the detailed proof. (We essentially follow the same proof for 2, and we introduce a nice shortcut to derive 1 from 2.) Let 𝒯Δ(Q)𝒯Δ\mathcal{T}_{\Delta}(Q)\subset\mathcal{T}^{\Delta} be a minimal finitely overlapping cover of SH(𝒯Q):=p𝒫0QSH(𝒯0(p))S_{H}(\mathcal{T}_{Q}):=\cup_{p\in\mathcal{P}_{0}\cap Q}S_{H}(\mathcal{T}_{0}(p)) by Δ\Delta-tubes. For p𝒫0Qp\in\mathcal{P}_{0}\cap Q, define

𝒯Δ,j(p)={𝐓𝒯Δ(Q):2j1<|{T𝒯(p):TSH1(𝐓)}|2j}.\mathcal{T}_{\Delta,j}(p)=\{\mathbf{T}\in\mathcal{T}_{\Delta}(Q):2^{j-1}<|\{T\in\mathcal{T}(p):T\subset S_{H}^{-1}(\mathbf{T})\}|\leq 2^{j}\}.

Since |𝒯Δ(Q)|100Δ2(d1)|\mathcal{T}_{\Delta}(Q)|\lesssim 100\Delta^{-2(d-1)} and Mj2j|𝒯Δ,j(p)|M\lesssim\sum_{j}2^{j}\cdot|\mathcal{T}_{\Delta,j}(p)|, we in fact have

MMΔ2(d1)/2002jM2j|𝒯Δ,j(p)|M\lesssim\sum_{M\Delta^{2(d-1)}/200\leq 2^{j}\leq M}2^{j}\cdot|\mathcal{T}_{\Delta,j}(p)|

Thus, by dyadic pigeonholing, there exists j=j(p)j=j(p) such that 2j|𝒯Δ,j(p)|ΔM2^{j}\cdot|\mathcal{T}_{\Delta,j}(p)|\approx_{\Delta}M. Another dyadic pigeonholing allows us to find 𝒫Q𝒫0Q\mathcal{P}_{Q}\subset\mathcal{P}_{0}\cap Q such that j(p)j(p) is constant for p𝒫Qp\in\mathcal{P}_{Q}. This is the desired refinement 𝒫Q\mathcal{P}_{Q} of 𝒫0Q\mathcal{P}_{0}\cap Q. Finally, let

𝒯Δ,i(Q):={𝐓𝒯Δ(Q):2i1<|{p𝒫Q:𝐓𝒯Δ(p)}|2i}.\mathcal{T}_{\Delta,i}(Q):=\{\mathbf{T}\in\mathcal{T}_{\Delta}(Q):2^{i-1}<|\{p\in\mathcal{P}_{Q}:\mathbf{T}\in\mathcal{T}_{\Delta}(p)\}|\leq 2^{i}\}.

Then by a similar dyadic pigeonholing (for calculations, see [18, Proposition 4.1]), there is ii such that

1200|𝒫Q|Δd12i|𝒫Q| and 2i+j|𝒯Δ,i(Q)|ΔM|𝒫Q|.\frac{1}{200}|\mathcal{P}_{Q}|\Delta^{d-1}\leq 2^{i}\leq|\mathcal{P}_{Q}|\text{ and }2^{i+j}\cdot|\mathcal{T}_{\Delta,i}(Q)|\approx_{\Delta}M\cdot|\mathcal{P}_{Q}|. (6.19)

Finally, we define 𝒯¯Δ(Q):=𝒯Δ,i(Q)\overline{\mathcal{T}}_{\Delta}(Q):=\mathcal{T}_{\Delta,i}(Q), which is the desired refinement of 𝒯Δ(Q)\mathcal{T}_{\Delta}(Q).

We check 2 holds with HQ=2i+jH_{Q}=2^{i+j}, which satisfies HQΔM|𝒫Q|/|𝒯¯Δ|H_{Q}\approx_{\Delta}M\cdot|\mathcal{P}\cap Q|/|\overline{\mathcal{T}}_{\Delta}| by (6.19) and |𝒫Q|Δ|𝒫Q||\mathcal{P}_{Q}|\approx_{\Delta}|\mathcal{P}\cap Q|. With this choice of HQH_{Q}, fix 𝐓𝒯¯Δ\mathbf{T}\in\overline{\mathcal{T}}_{\Delta} and note that

|{(p,T)𝒫Q×𝒯0:T𝒯0(p),TSH1(𝐓)}|=p𝒫Q|{T𝒯(p):TSH1(𝐓)}|2j|{p𝒫Q:𝐓𝒯Δ(p)}|2i+j=H.|\{(p,T)\in\mathcal{P}_{Q}\times\mathcal{T}_{0}:T\in\mathcal{T}_{0}(p),T\subset S_{H}^{-1}(\mathbf{T})\}|=\sum_{p\in\mathcal{P}_{Q}}|\{T\in\mathcal{T}(p):T\subset S_{H}^{-1}(\mathbf{T})\}|\\ \geq 2^{j}|\{p\in\mathcal{P}_{Q}:\mathbf{T}\in\mathcal{T}_{\Delta}(p)\}|\geq 2^{i+j}=H.

To check 1, we first pick a rr-tube 𝐓r\mathbf{T}_{r} with rΔr\geq\Delta. Then by 2 and (6.16),

|{𝐓𝒯¯Δ:𝐓𝐓r}|1H|{(p,T)𝒫Q×𝒯0:T𝒯0(p),TSH1(𝐓r)}|1H|𝒫Q|C1MrsC1|𝒯¯Δ|rs.|\{\mathbf{T}\in\overline{\mathcal{T}}_{\Delta}:\mathbf{T}\subset\mathbf{T}_{r}\}|\lesssim\frac{1}{H}|\{(p,T)\in\mathcal{P}_{Q}\times\mathcal{T}_{0}:T\in\mathcal{T}_{0}(p),T\subset S_{H}^{-1}(\mathbf{T}_{r})\}|\\ \lesssim\frac{1}{H}|\mathcal{P}_{Q}|\cdot C_{1}Mr^{s}\lessapprox C_{1}|\overline{\mathcal{T}}_{\Delta}|r^{s}.

Thus, 𝒯¯Δ(Q)\overline{\mathcal{T}}_{\Delta}(Q) is a (Δ,s,CΔ1,0)(\Delta,s,C_{\Delta}^{1},0)-set with CΔ1ΔC1C_{\Delta}^{1}\approx_{\Delta}C_{1}. Doing the same calculation with an (r,k+1)(r,k+1)-plank instead of an rr-tube, we get that 𝒯¯Δ(Q)\overline{\mathcal{T}}_{\Delta}(Q) is a (Δ,κ,CΔ2,k)(\Delta,\kappa,C_{\Delta}^{2},k)-set with CΔ2ΔC1C_{\Delta}^{2}\approx_{\Delta}C_{1}. This proves 1 and thus the claim.

Step 2: uniformity of |𝒯0𝐓||\mathcal{T}_{0}\cap\mathbf{T}|. By the pigeonhole principle, we can find M¯Δ1\overline{M}_{\Delta}\geq 1 and a subset 𝒬𝒟Δ(𝒫)\mathcal{Q}\subset\mathcal{D}_{\Delta}(\mathcal{P}) with |𝒬|Δ|𝒬0||\mathcal{Q}|\approx_{\Delta}|\mathcal{Q}_{0}| such that |𝒯¯Δ(Q)|M¯Δ|\overline{\mathcal{T}}_{\Delta}(Q)|\sim\overline{M}_{\Delta} for all Q𝒬Q\in\mathcal{Q}. Write

𝒯¯Δ=Q𝒬𝒯¯Δ(Q).\overline{\mathcal{T}}_{\Delta}=\bigcup_{Q\in\mathcal{Q}}\overline{\mathcal{T}}_{\Delta}(Q).

Next, by another dyadic pigeonholing, we can find a subset 𝒯¯Δ𝒯¯Δ\overline{\mathcal{T}}_{\Delta}^{\prime}\subset\overline{\mathcal{T}}_{\Delta} such that I(𝒬,𝒯¯Δ)I(𝒬,𝒯¯Δ)I(\mathcal{Q},\overline{\mathcal{T}}_{\Delta}^{\prime})\gtrapprox I(\mathcal{Q},\overline{\mathcal{T}}_{\Delta}) and |𝒯0𝐓|NΔ|\mathcal{T}_{0}\cap\mathbf{T}|\sim N_{\Delta} for all 𝐓𝒯¯Δ\mathbf{T}\in\overline{\mathcal{T}}_{\Delta}^{\prime}. Also, |𝒯¯Δ(Q)|M¯Δ|\overline{\mathcal{T}}_{\Delta}(Q)|\lesssim\overline{M}_{\Delta} for all Q𝒬Q\in\mathcal{Q}. Thus, we can find 𝒬𝒬\mathcal{Q}^{\prime}\subset\mathcal{Q} with |𝒬|Δ|𝒬||\mathcal{Q}^{\prime}|\approx_{\Delta}|\mathcal{Q}|, and for each Q𝒬Q\in\mathcal{Q}^{\prime} a subset 𝒯Δ(Q)\mathcal{T}_{\Delta}(Q) of cardinality M¯Δ\approx\overline{M}_{\Delta}, such that 𝒯Δ(Q)𝒯¯Δ\mathcal{T}_{\Delta}(Q)\subset\overline{\mathcal{T}}_{\Delta}^{\prime}. In other words,

|𝒯0𝐓|NΔ for 𝐓𝒯Δ(Q).|\mathcal{T}_{0}\cap\mathbf{T}|\sim N_{\Delta}\text{ for }\mathbf{T}\in\mathcal{T}_{\Delta}(Q).

Thus, we obtain item ii.

|𝒯0||𝒯Δ|min𝐓𝒯Δ|𝒯0𝐓||𝒯Δ|NΔ.|\mathcal{T}_{0}|\geq|\mathcal{T}_{\Delta}|\cdot\min_{\mathbf{T}\in\mathcal{T}_{\Delta}}|\mathcal{T}_{0}\cap\mathbf{T}|\sim|\mathcal{T}_{\Delta}|\cdot N_{\Delta}. (6.20)

Reduce the families 𝒯Δ(Q)\mathcal{T}_{\Delta}(Q) such that their cardinality is MΔ:=min(|𝒯Δ(Q)|:Q𝒬})δM¯ΔM_{\Delta}:=\min(|\mathcal{T}_{\Delta}(Q)|:Q\in\mathcal{Q}\})\approx_{\delta}\overline{M}_{\Delta}. By 1, 𝒯Δ(Q)\mathcal{T}_{\Delta}(Q) remains a (Δ,s,CΔ1,0)(\Delta,s,C_{\Delta}^{1},0) and (Δ,κ,CΔ2,k)(\Delta,\kappa,C_{\Delta}^{2},k)-set with CΔ1,CΔ2δC1C_{\Delta}^{1},C_{\Delta}^{2}\approx_{\delta}C_{1}.

Finally, define

𝒫=Q𝒬𝒫Q,\mathcal{P}=\bigcup_{Q\in\mathcal{Q}}\mathcal{P}_{Q},

where 𝒬\mathcal{Q} is the latest refinement of 𝒬0\mathcal{Q}_{0}. Since |𝒫Q|Δ|𝒫0SH1(Q)||\mathcal{P}_{Q}|\approx_{\Delta}|\mathcal{P}_{0}\cap S_{H}^{-1}(Q)|, we get that item i holds.

For p𝒫Q=𝒫Qp\in\mathcal{P}_{Q}=\mathcal{P}\cap Q, Q𝒬Q\in\mathcal{Q}, define

𝒯(p)=𝐓𝒯Δ(Q)(𝒯0(p)𝐓),𝒯=p𝒫𝒯(p),𝒯Δ=Q𝒬𝒯Δ(Q).\mathcal{T}(p)=\bigcup_{\mathbf{T}\in\mathcal{T}_{\Delta}(Q)}(\mathcal{T}_{0}(p)\cap\mathbf{T}),\mathcal{T}=\bigcup_{p\in\mathcal{P}}\mathcal{T}(p),\mathcal{T}_{\Delta}=\bigcup_{Q\in\mathcal{Q}}\mathcal{T}_{\Delta}(Q).

Thus, (𝒟Δ(𝒫),𝒯Δ)=(𝒬,𝒯Δ)(\mathcal{D}_{\Delta}(\mathcal{P}),\mathcal{T}_{\Delta})=(\mathcal{Q},\mathcal{T}_{\Delta}) is a (Δ,s,CΔ1,κ,CΔ2,MΔ)(\Delta,s,C_{\Delta}^{1},\kappa,C_{\Delta}^{2},M_{\Delta})-nice configuration, establishing item iii. To summarize, in this step, we refined 𝒬\mathcal{Q} and 𝒯Δ(Q)\mathcal{T}_{\Delta}(Q) for Q𝒬Q\in\mathcal{Q}, so 2/iv still holds (with same HQH_{Q} and a weaker implied constant).

Step 3: uniformity of 𝒯(p)\mathcal{T}(p) and construct 𝒯Q\mathcal{T}_{Q}. This step will be devoted to verifying v and (6.18). We will not change 𝒫,𝒯\mathcal{P},\mathcal{T}, or 𝒯Δ\mathcal{T}_{\Delta}.

Fix Q𝒬Q\in\mathcal{Q}, and let 𝒫Q=𝒫SH1(Q)\mathcal{P}_{Q}=\mathcal{P}\cap S_{H}^{-1}(Q). Define

𝒯(Q)=p𝒫Q𝒯(p).\mathcal{T}(Q)=\bigcup_{p\in\mathcal{P}_{Q}}\mathcal{T}(p).

By dyadic pigeonholing and 2, we can find a Δ\approx_{\Delta}-comparable subset of 𝒫Q\mathcal{P}_{Q} (which we keep denoting 𝒫Q\mathcal{P}_{Q}) such that

|𝒯(p)|ΔM,p𝒫Q.|\mathcal{T}(p)|\approx_{\Delta}M,\qquad p\in\mathcal{P}_{Q}.

Next,

|𝒯(Q)|𝐓𝒯Δ(Q)|𝒯𝐓|ΔMΔNΔ.|\mathcal{T}(Q)|\leq\sum_{\mathbf{T}\in\mathcal{T}_{\Delta}(Q)}|\mathcal{T}\cap\mathbf{T}|\lessapprox_{\Delta}M_{\Delta}\cdot N_{\Delta}. (6.21)

For a given pQp\in Q, we consider the tube packet 𝕌(p):=𝒯(p)SH1(Q)\mathbb{U}(p):=\mathcal{T}(p)\cap S_{H}^{-1}(Q) (discarding duplicate tubelets). Each tubelet u𝕌(p)u\in\mathbb{U}(p) lies in at most Δ2(d1)\Delta^{-2(d-1)} many tubes of 𝒯(p)\mathcal{T}(p), so by dyadic pigeonholing, we can refine 𝒯(p)\mathcal{T}(p) by a logΔ1\log\Delta^{-1} factor to ensure that each tubelet u𝕌(p)u\in\mathbb{U}(p) lies in m(p)\sim m(p) many tubes of 𝒯(p)\mathcal{T}(p), and there are M(p)ΔMm(p)M(p)\approx_{\Delta}\frac{M}{m(p)} many distinct tubelets through pp. By refining 𝒫Q\mathcal{P}_{Q} by a (logΔ1)(\log\Delta^{-1})-factor, we may assume m(p)mQm(p)\approx m_{Q} for each p𝒫Qp\in\mathcal{P}_{Q}. Now, define

𝒫Q:=SH1SQSH(𝒫Q) and 𝒯Q:=p𝒫QSH1SQSH(𝕌(p)).\mathcal{P}^{Q}:=S_{H}^{-1}\circ S_{Q}\circ S_{H}(\mathcal{P}_{Q})\text{ and }\mathcal{T}_{Q}:=\bigcup_{p\in\mathcal{P}_{Q}}S_{H}^{-1}\circ S_{Q}\circ S_{H}(\mathbb{U}(p)).

Since tubelets are essentially distinct and each tubelet in any 𝕌(p)\mathbb{U}(p) corresponds to mQ\approx m_{Q} many tubes in 𝒯(Q)\mathcal{T}(Q), we obtain:

|𝒯(Q)|Δ|p𝒫Q𝕌(p)|mQ|𝒯Q|MMQ.|\mathcal{T}(Q)|\gtrapprox_{\Delta}\left|\bigcup_{p\in\mathcal{P}_{Q}}\mathbb{U}(p)\right|\cdot m_{Q}\gtrsim|\mathcal{T}_{Q}|\cdot\frac{M}{M_{Q}}. (6.22)

Then (6.18) will follow by combining (6.20), (6.21), and (6.22).

We finally check (𝒫Q,𝒯Q)(\mathcal{P}^{Q},\mathcal{T}_{Q}) is a (δ/Δ,s,CQ1r0ks,κ,CQ2,MQ)(\delta/\Delta,s,C_{Q}^{1}r_{0}^{k-s},\kappa,C_{Q}^{2},M_{Q})-nice configuration down from scale r0r_{0}. First, for any δ¯<r<r0\overline{\delta}<r<r_{0}, we have for any (r,0)(r,0)-plank HH in Sd1S^{d-1},

|σ(𝒯Q)H|Δ1mQ|σ(𝒯(p))H|Δ1mQCMrs=CMQrs.|\sigma(\mathcal{T}_{Q})\cap H|\sim_{\Delta}\frac{1}{m_{Q}}|\sigma(\mathcal{T}(p))\cap H|\lessapprox_{\Delta}\frac{1}{m_{Q}}\cdot C\cdot M\cdot r^{s}=C\cdot M_{Q}\cdot r^{s}.

Thus, σ(𝒯Q)\sigma(\mathcal{T}_{Q}) is a (δ¯,s,CQ1,0)(\overline{\delta},s,C_{Q}^{1},0)-set down from scale r0r_{0} with CQ1ΔC1C_{Q}^{1}\approx_{\Delta}C_{1}. Similarly, σ(𝒯Q)\sigma(\mathcal{T}_{Q}) is a (δ¯,κ,CQ2,k)(\overline{\delta},\kappa,C_{Q}^{2},k)-set with CQ2ΔC2C_{Q}^{2}\approx_{\Delta}C_{2}. This shows item v and thus the proof of the Proposition.

6.2 Good multiscale decomposition

The idea is to apply Proposition 6.2, then apply Theorem 1.8 to bound |𝒯Δ||\mathcal{T}_{\Delta}| and Corollary 6.5 to bound |𝒯Q||\mathcal{T}_{Q}|. Unfortunately, while we use pigeonholing to ensure that 𝒟Δ(SH(𝒫))\mathcal{D}_{\Delta}(S_{H}(\mathcal{P})) is a (Δ,t)(\Delta,t)-set, we don’t know that SH1SQ(SH(𝒫)Q)S_{H}^{-1}\circ S_{Q}(S_{H}(\mathcal{P})\cap Q) is a (δΔ,t)(\frac{\delta}{\Delta},t)-set. In fact, we won’t show this statement, but rather a slightly weaker statement that is good enough. For this, a good choice of Δ\Delta based on the branching structure of 𝒫\mathcal{P} is needed.

First, we explain the pigeonholing preliminaries.

Lemma 6.3.

Given PHrP\subset H_{r}, a (r0,k)(r_{0},k)-plane, there is a subset PPP^{\prime}\subset P with |P|δ(log(r0δ))1|P|δ|P^{\prime}|_{\delta}\geq(\log(\frac{r_{0}}{\delta}))^{-1}|P|_{\delta} such that |QSH(P)||Q\cap S_{H}(P)| is constant for all Q𝒟δ/r0(SH(P))Q\in\mathcal{D}_{\delta/r_{0}}(S_{H}(P)).

Proof.

Let f(N)={|PSH1(Q)|δ:Q𝒟δ/r0([0,1]d),|PSH1(Q)|δ[N,2N]}f(N)=\sum\{|P\cap S_{H}^{-1}(Q)|_{\delta}:Q\in\mathcal{D}_{\delta/r_{0}}([0,1]^{d}),|P\cap S_{H}^{-1}(Q)|_{\delta}\in[N,2N]\}. Then N dyadicf(N)=|P|δ\sum_{N\text{ dyadic}}f(N)=|P|_{\delta}. For each NN, either f(N)=0f(N)=0 or Nf(N)(r0/δ)dNN\leq f(N)\leq(r_{0}/\delta)^{d}\cdot N. Hence, if N0N_{0} is the largest NN for which f(N)>0f(N)>0, we get f(N0)N0>M<N0(δ/r0)d/100 dyadicf(M)f(N_{0})\geq N_{0}>\sum_{M<N_{0}(\delta/r_{0})^{d}/100\text{ dyadic}}f(M). Thus, we have

N0(δ/r0)d/100<M<N0 dyadicf(M)>12|P|δ.\sum_{N_{0}(\delta/r_{0})^{d}/100<M<N_{0}\text{ dyadic}}f(M)>\frac{1}{2}|P|_{\delta}.

Thus, by dyadic pigeonholing, there exists M(N0(δ/r0)d/100,N0)M\in(N_{0}(\delta/r_{0})^{d}/100,N_{0}) such that f(M)120d(log(r0δ))1|P|δf(M)\geq\frac{1}{20d}(\log(\frac{r_{0}}{\delta}))^{-1}|P|_{\delta}. ∎

The next step is to make SH1SQ(SH(𝒫)Q)S_{H}^{-1}\circ S_{Q}(S_{H}(\mathcal{P})\cap Q) satisfy a tt^{\prime}-dimensional spacing condition with tt^{\prime} just slightly less than tt, for all Q𝒟Δ(SH(𝒫))Q\in\mathcal{D}_{\Delta}(S_{H}(\mathcal{P})) at a certain scale Δ\Delta. To do so, we need the following lemma.

Lemma 6.4.

Fix C,ε>0C,\varepsilon>0, and let δr0=Δm\frac{\delta}{r_{0}}=\Delta^{m} and PHrP\subset H_{r} be a (δ,t,C,0)(\delta,t,C,0)-set in HrH_{r}, a (r0,k)(r_{0},k)-plane. Let L=log(r0δ)(log(1/Δ))mL=\log(\frac{r_{0}}{\delta})\cdot(\log(1/\Delta))^{m}. If t<tdε1εt^{\prime}<\frac{t-d\varepsilon}{1-\varepsilon}, then there exists mεkmm\varepsilon\leq k\leq m and a subset PPP^{\prime}\subset P with |P|L1|P||P^{\prime}|\geq L^{-1}|P| such that for any kjmk\leq j\leq m, Q𝒟Δk(P)Q\in\mathcal{D}_{\Delta^{k}}(P^{\prime}), and R𝒟Δj(P)QR\in\mathcal{D}_{\Delta^{j}}(P^{\prime})\cap Q, we have

|PSH1(R)||PSH1(Q)|Δ(jk)t,|P^{\prime}\cap S_{H}^{-1}(R)|\leq|P^{\prime}\cap S_{H}^{-1}(Q)|\cdot\Delta^{(j-k)t^{\prime}}, (6.23)

and for δrδr0\delta\leq r\leq\frac{\delta}{r_{0}} and a ball BrB_{r}, we have

|PBr|CL|PSH1(Q)|(rΔk)t.|P^{\prime}\cap B_{r}|\leq C\cdot L\cdot|P^{\prime}\cap S_{H}^{-1}(Q)|\left(\frac{r}{\Delta^{k}}\right)^{t^{\prime}}. (6.24)
Proof.

Throughout this proof we will not distinguish between mεm\varepsilon and mε\lceil m\varepsilon\rceil.

First, we will make SH(P)S_{H}(P) uniform at scales 1,Δ,Δ2,,Δm=δr01,\Delta,\Delta^{2},\cdots,\Delta^{m}=\frac{\delta}{r_{0}}. By Lemmas 6.3 and 2.17, we can find |P|L1|P||P^{\prime}|\geq L^{-1}|P| such that there is a sequence (Nj)j=1n(N_{j})_{j=1}^{n} with |SH(P)Q|Δk=Nk|S_{H}(P^{\prime})\cap Q|_{\Delta^{k}}=N_{k} for all 1kn1\leq k\leq n and Q𝒟Δk(P)Q\in\mathcal{D}_{\Delta^{k}}(P^{\prime}).

Let mεkmm\varepsilon\leq k\leq m be the largest index such that Nk|P|Δmεd+(kmε)tN_{k}\geq|P^{\prime}|\Delta^{m\varepsilon\cdot d+(k-m\varepsilon)t^{\prime}} for one (equivalently all) Q𝒟Δk(P)Q\in\mathcal{D}_{\Delta^{k}}(P^{\prime}). Certainly k=mεk=m\varepsilon is a valid index since |𝒟Δmε|=Δdmε|\mathcal{D}_{\Delta^{m\varepsilon}}|=\Delta^{-dm\varepsilon}.

Now, we will check the given conditions. By maximality of kk, we have for kjmk\leq j\leq m,

Nj|P|Δdmε+(jmε)tNkΔ(jk)t.N_{j}\leq|P^{\prime}|\Delta^{dm\varepsilon+(j-m\varepsilon)t^{\prime}}\leq N_{k}\Delta^{(j-k)t^{\prime}}.

Noticing that |PSH1(Q)|=|SH(P)Q||P^{\prime}\cap S_{H}^{-1}(Q)|=|S_{H}(P^{\prime})\cap Q| and likewise for R𝒟Δj(P)QR\in\mathcal{D}_{\Delta^{j}}(P^{\prime})\cap Q, this proves (6.23).

To check (6.24), we recall that LNkL|P|Δdmε+(kmε)t|P|Δdmε+(kmε)tLN_{k}\geq L\cdot|P^{\prime}|\Delta^{dm\varepsilon+(k-m\varepsilon)t^{\prime}}\geq|P|\Delta^{dm\varepsilon+(k-m\varepsilon)t^{\prime}}. Using rδr0=Δmr\leq\frac{\delta}{r_{0}}=\Delta^{m}, ttdε1εt^{\prime}\leq\frac{t-d\varepsilon}{1-\varepsilon}, and that PP is a (δ,t,C)(\delta,t,C)-set, we have

|PBr||PBr|C|P|rtCNkL(rΔk)t.|P^{\prime}\cap B_{r}|\leq|P\cap B_{r}|\leq C|P|r^{t}\leq C\cdot N_{k}L\left(\frac{r}{\Delta^{k}}\right)^{t^{\prime}}.

Finally, we will need the following variant of Corollary 2.10.

Corollary 6.5.

Let 0max(s,k)<td10\leq\max(s,k)<t\leq d-1, δr1\delta\leq r\leq 1, and let CP1,CT0C_{P}\geq 1,C_{T}\geq 0. Let 𝒫𝒟δ\mathcal{P}\subset\mathcal{D}_{\delta} be a set contained in an (r0,k+1)(r_{0},k+1)-plate HH satisfying the following conditions:

  • For all δr0r1\frac{\delta}{r_{0}}\leq r\leq 1 and balls BrB_{r}, we have

    |𝒫SH1(Br)|CP|𝒫|rt.|\mathcal{P}\cap S_{H}^{-1}(B_{r})|\leq C_{P}\cdot|\mathcal{P}|\cdot r^{t}. (6.25)
  • For all δrδr0\delta\leq r\leq\frac{\delta}{r_{0}} and balls BrB_{r}, we have

    |𝒫Br|CP|𝒫|rt.|\mathcal{P}\cap B_{r}|\leq C_{P}\cdot|\mathcal{P}|\cdot r^{t}. (6.26)

Assume that for every p𝒫p\in\mathcal{P} there exists a family 𝒯(p)𝒯δ\mathcal{T}(p)\subset\mathcal{T}^{\delta} of dyadic δ\delta-tubes satisfying the following conditions:

  • TpT\cap p\neq\emptyset for all T𝒯(p)T\in\mathcal{T}(p);

  • |𝒯(p)𝐓|CT|𝒯(p)|r0ksxs|\mathcal{T}(p)\cap\mathbf{T}|\leq C_{T}\cdot|\mathcal{T}(p)|\cdot r_{0}^{k-s}x^{s} for all xx-tubes 𝐓\mathbf{T} with δxr0\delta\leq x\leq r_{0}.

Further assume that |𝒯(p)|=M|\mathcal{T}(p)|=M for some M1M\geq 1. If 𝒯=p𝒫𝒯(p)\mathcal{T}=\cup_{p\in\mathcal{P}}\mathcal{T}(p), then

|𝒯|(CPCT)1Mr0skδs.|\mathcal{T}|\gtrsim(C_{P}C_{T})^{-1}\cdot Mr_{0}^{s-k}\delta^{-s}.
Proof.

Let

jP(𝒫,𝒯)={(q,t)𝒫×𝒯(p):t𝒯(q)}j_{P}(\mathcal{P},\mathcal{T})=\{(q,t)\in\mathcal{P}\times\mathcal{T}(p):t\in\mathcal{T}(q)\}

We have the following:

Lemma 6.6.

For all p𝒫p\in\mathcal{P}, we have jp(𝒫,𝒯)s,t,kCPCT|𝒫|Mr0ksδsj_{p}(\mathcal{P},\mathcal{T})\lesssim_{s,t,k}C_{P}C_{T}|\mathcal{P}|\cdot Mr_{0}^{k-s}\delta^{s}.

Proof.

We count jp(𝒫,𝒯)j_{p}(\mathcal{P},\mathcal{T}) by first choosing a dyadic δ<r<1\delta<r<1, then counting the number of q𝒫q\in\mathcal{P} with |pq|r|p-q|\sim r, then finally counting the number of t𝒯t\in\mathcal{T} that pass through p,qp,q.

If r>δr0r>\frac{\delta}{r_{0}}, we claim that if |xy|[r,2r]|x-y|\in[r,2r] and some tube through x,yx,y lies in HH, then |SH(x)SH(y)|100r|S_{H}(x)-S_{H}(y)|\leq 100r, so ySH1(B100r(SH(x)))y\in S_{H}^{-1}(B_{100r}(S_{H}(x))).

To prove this, we may assume r0150r_{0}\leq\frac{1}{50}, as otherwise we can use the simple fact |SH(x)SH(y)|r01|xy|100r|S_{H}(x)-S_{H}(y)|\leq r_{0}^{-1}|x-y|\leq 100r. Now choose a coordinate system such that the first k+1k+1 axes correspond to the long sides of HH, and the remaining axes correspond to the short sides of HH. Let xy=(a,b)\Rk+1×\Rdk1x-y=(\vec{a},\vec{b})\in\R^{k+1}\times\R^{d-k-1}. Then |a||xy|r|\vec{a}|\leq|x-y|\leq r. Furthermore, we have |b|50r0|a||\vec{b}|\leq 50r_{0}|\vec{a}|, otherwise any tube through x,yx,y would be roughly orthogonal to HH and intersect HH in a subtube with length 2r012r_{0}\leq 1, contradiction. Thus, we have |SH(x)SH(y)||a|+r01|b|100r|S_{H}(x)-S_{H}(y)|\leq|\vec{a}|+r_{0}^{-1}|\vec{b}|\leq 100r.

Using the claim and condition (6.25), we see that there are CP|𝒫|rt\lesssim C_{P}|\mathcal{P}|\cdot r^{t} many choices for qq. For each qq, the set of tubes t𝒯(p)t\in\mathcal{T}(p) passing through qq lies in a δr\frac{\delta}{r}-tube, so by the tube non-concentration condition (and noting that δr<r0\frac{\delta}{r}<r_{0}), we have CTMr0ks(δr)sC_{T}\cdot Mr_{0}^{k-s}\left(\frac{\delta}{r}\right)^{s} choices for tt.

Thus, the contribution to jp(𝒫,𝒯)j_{p}(\mathcal{P},\mathcal{T}) for a given dyadic r>δr0r>\frac{\delta}{r_{0}} is CPCT|𝒫|Mr0ksδsrstC_{P}C_{T}|\mathcal{P}|\cdot Mr_{0}^{k-s}\delta^{s}\cdot r^{s-t}, and summing over dyadic rr gives CPCT|𝒫|Mr0ksδsOst(1)C_{P}C_{T}|\mathcal{P}|\cdot Mr_{0}^{k-s}\delta^{s}\cdot O_{s-t}(1).

If r<δr0r<\frac{\delta}{r_{0}}, then by condition 6.26 we see that there are CP|𝒫|rt\lesssim C_{P}|\mathcal{P}|\cdot r^{t} many choices for qq. For each qq, the set of tubes t𝒯(p)t\in\mathcal{T}(p) passing through qq lies in a δr\frac{\delta}{r}-tube 𝐓δ/r\mathbf{T}_{\delta/r}. We note that δr>r0\frac{\delta}{r}>r_{0}, so the tube non-concentration doesn’t apply directly, but luckily we note that 𝐓δ/rH\mathbf{T}_{\delta/r}\cap H can be covered by (δrr0)k(\frac{\delta}{rr_{0}})^{k} many r0r_{0}-tubes. Thus, by using tube non-concentration at scale r0r_{0}, we have CTMr0k(δrr0)kC_{T}\cdot Mr_{0}^{k}\cdot(\frac{\delta}{rr_{0}})^{k} choices for tt.

Thus, the contribution to jp(𝒫,𝒯)j_{p}(\mathcal{P},\mathcal{T}) for a given dyadic r<δr0r<\frac{\delta}{r_{0}} is (after some manipulation)

CPCT|P|Mr0k(δr0)t(rr0δ)tk.C_{P}C_{T}|P|Mr_{0}^{k}\left(\frac{\delta}{r_{0}}\right)^{t}\cdot\left(\frac{rr_{0}}{\delta}\right)^{t-k}.

Since t>max(k,s)t>\max(k,s), the sum is CPCT|P|Mr0k(δr0)s\lesssim C_{P}C_{T}|P|Mr_{0}^{k}\left(\frac{\delta}{r_{0}}\right)^{s}.

Adding up both r>δr0r>\frac{\delta}{r_{0}} and r<δr0r<\frac{\delta}{r_{0}} contributions, we prove the Lemma. ∎

For t𝒯t\in\mathcal{T}, let 𝒫(t)={p𝒫:t𝒯(p)}\mathcal{P}(t)=\{p\in\mathcal{P}:t\in\mathcal{T}(p)\}. By Cauchy-Schwarz, we have

(M|𝒫|)2=(t𝒯|𝒫(t)|)2|𝒯|t𝒯|𝒫(t)|2=|𝒯|p𝒫jp(𝒫,𝒯).(M|\mathcal{P}|)^{2}=\left(\sum_{t\in\mathcal{T}}|\mathcal{P}(t)|\right)^{2}\leq|\mathcal{T}|\sum_{t\in\mathcal{T}}|\mathcal{P}(t)|^{2}=|\mathcal{T}|\sum_{p\in\mathcal{P}}j_{p}(\mathcal{P},\mathcal{T}).

By Lemma 6.6, we get

|𝒯|M2|𝒫|2CPCT|𝒫|2Mr0ksδs=(CPCT)1Mr0skδs.|\mathcal{T}|\geq\frac{M^{2}|\mathcal{P}|^{2}}{C_{P}C_{T}|\mathcal{P}|^{2}Mr_{0}^{k-s}\delta^{s}}=(C_{P}C_{T})^{-1}Mr_{0}^{s-k}\delta^{-s}.

Proof of Theorem 6.1.

A small reduction: we would like to assume |𝒯(p)|M|\mathcal{T}(p)|\sim M for all p𝒫p\in\mathcal{P}. To assume this, we first observe that |𝒯(p)|M0=(δ/r0)εr0ksδs|\mathcal{T}(p)|\geq M_{0}=(\delta/r_{0})^{-\varepsilon}r_{0}^{k-s}\delta^{-s} for all p𝒫p\in\mathcal{P}. On the other hand, if for at least half of the p𝒫p\in\mathcal{P} (call them 𝒫\mathcal{P}^{\prime}) we have |𝒯(p)|M0(δ/r0)1|\mathcal{T}(p)|\geq M_{0}(\delta/r_{0})^{-1}, then we are immediately done by Corollary 6.5 applied to 𝒫\mathcal{P}^{\prime} and 𝒯(p)\mathcal{T}(p). Thus, by reducing 𝒫\mathcal{P} if necessary, we may assume |𝒯(p)|(M0,M0(δ/r0)1)|\mathcal{T}(p)|\in(M_{0},M_{0}(\delta/r_{0})^{-1}). Then by reducing 𝒫\mathcal{P} further by a log(δ/r0)1\lesssim\log(\delta/r_{0})^{-1} factor, we may assume |𝒯(p)|(M,2M)|\mathcal{T}(p)|\in(M,2M) for some M(M0,M0(δ/r0)1)M\in(M_{0},M_{0}(\delta/r_{0})^{-1}). Finally, we may remove some tubes from each 𝒯(p)\mathcal{T}(p) to make |𝒯(p)|=M|\mathcal{T}(p)|=M. Then (𝒫0,𝒯0)(\mathcal{P}_{0},\mathcal{T}_{0}) is a (δ,s,C1r0ks,κ,C2,M)(\delta,s,C_{1}r_{0}^{k-s},\kappa,C_{2},M)-nice configuration.

Pick β(s,t,k)>0\beta(s,t,k)>0 such that tdβ1β>max(s,k)\frac{t-d\beta}{1-\beta}>\max(s,k), and let t=12(tdβ1β+max(s,k))t^{\prime}=\frac{1}{2}(\frac{t-d\beta}{1-\beta}+\max(s,k)). Pick Δ>0\Delta>0 such that log(1/Δ)<Δε\log(1/\Delta)<\Delta^{-\varepsilon}. Find Δ=Δk(δ/r0,(δ/r0)β)\Delta^{\prime}=\Delta^{k}\in(\delta/r_{0},(\delta/r_{0})^{\beta}) such that the conclusion of Lemma 6.4 holds. Now by Proposition 6.2, we have

|𝒯|M|𝒯Q|MQ|𝒯Δ(𝒯)|MΔ.\frac{|\mathcal{T}|}{M}\geq\frac{|\mathcal{T}_{Q}|}{M_{Q}}\cdot\frac{|\mathcal{T}^{\Delta^{\prime}}(\mathcal{T})|}{M_{\Delta^{\prime}}}.

If ε<βη2\varepsilon<\beta\eta^{2}, where η(s,t,κ,k,d)\eta(s,t,\kappa,k,d) is the parameter in Theorem 5.1, we have |𝒯Δ(𝒯)|MΔ(Δ)sε\frac{|\mathcal{T}^{\Delta^{\prime}}(\mathcal{T})|}{M_{\Delta^{\prime}}}\geq(\Delta^{\prime})^{-s-\sqrt{\varepsilon}}.

Pick QQ. Then SH1SQ(SH(𝒫)Q)S_{H}^{-1}\circ S_{Q}(S_{H}(\mathcal{P})\cap Q) satisfies the conditions of Corollary 6.5 with CP=(δr0)εLΔdC_{P}=\left(\frac{\delta}{r_{0}}\right)^{-\varepsilon}\cdot L\cdot\Delta^{-d}. Thus, we have |𝒯Q|MQ(δr0)εΔdL1r0sk(δΔ)s\frac{|\mathcal{T}_{Q}|}{M_{Q}}\geq\left(\frac{\delta}{r_{0}}\right)^{-\varepsilon}\Delta^{d}L^{-1}\cdot r_{0}^{s-k}\left(\frac{\delta}{\Delta^{\prime}}\right)^{-s}. Using these two bounds and M(δ/r0)εr0skδsM\geq(\delta/r_{0})^{\varepsilon}r_{0}^{s-k}\delta^{-s}, we get

|𝒯|(δr0)2εεβΔdL1r02(sk)δ2s.|\mathcal{T}|\geq\left(\frac{\delta}{r_{0}}\right)^{2\varepsilon-\sqrt{\varepsilon}\beta}\Delta^{-d}L^{-1}r_{0}^{2(s-k)}\cdot\delta^{-2s}.

It remains to choose ε<β2/100\varepsilon<\beta^{2}/100 and also for δr0\frac{\delta}{r_{0}} small enough, we have Δd<(δr0)ε\Delta^{-d}<\left(\frac{\delta}{r_{0}}\right)^{-\varepsilon} and L(δr0)εΔεm(δr0)2εL\leq\left(\frac{\delta}{r_{0}}\right)^{-\varepsilon}\Delta^{-\varepsilon m}\leq\left(\frac{\delta}{r_{0}}\right)^{-2\varepsilon}. Thus, |𝒯|(δr0)εr02(sk)δ2s|\mathcal{T}|\geq\left(\frac{\delta}{r_{0}}\right)^{-\varepsilon}r_{0}^{2(s-k)}\cdot\delta^{-2s} and we are done. ∎

7 Power decay around kk-planes

In this section, we will roughly deal with the following situation:

  • μ,ν\mu,\nu are ss-Frostman measures with k1<skk-1<s\leq k;

  • ν\nu gives mass ε\leq\varepsilon to any (r0,k)(r_{0},k)-plate.

In other words, ν\nu does not concentrate around (r0,k)(r_{0},k)-plates. We would like to understand the ν\nu-mass of (r,k)(r,k)-plates for rr much smaller than r0r_{0}. A result of Shmerkin [25, Proposition B.1] says that there exist r1(r0,s,k),κ(s,k)>0r_{1}(r_{0},s,k),\kappa(s,k)>0, a subset XsptμX\subset\mathrm{spt}\mu with μ(X)>1O(ε)\mu(X)>1-O(\varepsilon), and for each xXx\in X, a subset YxsptνY_{x}\subset\mathrm{spt}\nu with μ(Yx)>1O(ε)\mu(Y_{x})>1-O(\varepsilon) such that ν(HYx)rη\nu(H\cap Y_{x})\leq r^{\eta} for all rr1r\leq r_{1} and (r,k)(r,k)-plates HH through xx. Thus, we do obtain a power decay for sufficiently small rr. But what is the optimal starting point of the power decay? Can we hope for a power decay ν(HYx)K(rr0)η\nu(H\cap Y_{x})\lesssim K(\frac{r}{r_{0}})^{\eta} for all (r,k)(r,k)-plates through xx? The answer is yes, and indeed we shall prove it by making small but meaningful tweaks to Shmerkin’s argument. But before stating our result, we shall introduce some convenient notation. We define thin kk-plates, a generalization of thin tubes, as follows.

Definition 7.1.

Let K,t0K,t\geq 0, 1kd11\leq k\leq d-1, and c(0,1]c\in(0,1]. Let μ,ν(\Rd)\mu,\nu\in\mathbb{P}(\R^{d}) supported on X,YX,Y. Fix GX×YG\subset X\times Y. We say (μ,ν)(\mu,\nu) has (t,K,c)(t,K,c)-thin kk-plates on GG down from scale r0r_{0} if

ν(HG|x)Krt for all r(0,r0) and all (r,k)-plates H containing x.\nu(H\cap G|_{x})\leq K\cdot r^{t}\quad\text{ for all }r\in(0,r_{0})\text{ and all }(r,k)\text{-plates }H\text{ containing }x. (7.27)
Remark 7.2.

In this paper, we will choose G=(AB)cG=(A\cup B)^{c} where μ×ν(B)\mu\times\nu(B) is small. (The complement is taken with respect to \Rd×\Rd\R^{d}\times\R^{d}.) In this case, the equation (7.27) becomes

ν(H(A|xB|x))Krt for all r(0,r0) and all (r,k)-plates H containing x.\nu(H\setminus(A|_{x}\cup B|_{x}))\leq K\cdot r^{t}\quad\text{ for all }r\in(0,r_{0})\text{ and all }(r,k)\text{-plates }H\text{ containing }x.

Now, we can state the main proposition, which generalizes and extends Proposition B.1 of [25]. It may be of independent interest.

Proposition 7.3.

Let 1kd11\leq k\leq d-1 and k1<skk-1<s\leq k. There exist η(κ,k,d)>0\eta(\kappa,k,d)>0 and K0(κ,k,d)>0K_{0}(\kappa,k,d)>0 with the following property. Fix r01r_{0}\leq 1 and KK0K\geq K_{0}. Suppose that μ,ν\mu,\nu are positive measures with |μ|,|ν|1|\mu|,|\nu|\geq 1 and for any (r,k1)(r,k-1)-plate HH, we have

μ(H)Cμrκ,\displaystyle\mu(H)\leq C_{\mu}r^{\kappa},
ν(H)Cνrκ.\displaystyle\nu(H)\leq C_{\nu}r^{\kappa}.

Let AX×YA\subset X\times Y be the pairs of points that lie in some K1K^{-1}-concentrated (r0,k)(r_{0},k)-plate. Then there exists BB with μ×ν(B)K0K1\mu\times\nu(B)\leq K_{0}K^{-1} such that (μ,ν)(\mu,\nu) have (η,Kr0η)(\eta,Kr_{0}^{-\eta})-thin kk-plates on (AB)c(A\cup B)^{c}. (The complement is taken with respect to \Rd×\Rd\R^{d}\times\R^{d}.)

Remark 7.4.

(a) We can apply Proposition 7.3 in case μ,ν\mu,\nu are ss-dimensional with s>k1s>k-1.

(b) In Proposition B.1 of [25], the exponents for μ,ν\mu,\nu are allowed to differ. The proof of Proposition 7.3 is easily modified to include this detail.

To prove Proposition 7.3, we need the following two lemmas. Fix rr0r\leq r_{0}. The first says that there are few dense (r,k)(r,k)-plates, and the second says that for most xXx\in X, the dense (r,k)(r,k)-plates through xx lie in some (r0,k)(r_{0},k)-plate.

Lemma 7.5.

There is N=N(κ,k,d)N=N(\kappa,k,d) such that the following holds: let ν\nu be a measure with mass 1\leq 1 such that ν(W)Cνρκ\nu(W)\leq C_{\nu}\rho^{\kappa} for all (ρ,k1)(\rho,k-1)-plates WW, 1>ρ>r1>\rho>r. Let r,k\mathcal{E}_{r,k} be a set of (r,k)(r,k)-plates such that every (s,k)(s,k)-plate contains (sr)(k+1)(dk)\lesssim\left(\frac{s}{r}\right)^{(k+1)(d-k)} many rr-plates of r,k\mathcal{E}_{r,k} (as in Section 2.2). Let ={Hr,k:ν(H)a}\mathcal{H}=\{H\in\mathcal{E}_{r,k}:\nu(H)\geq a\}. Then ||(Cνa)N|\mathcal{H}|\lesssim(\frac{C_{\nu}}{a})^{N}.

Lemma 7.5 follows from the condition on r,k\mathcal{E}_{r,k} and the following generalization of [25, Lemma B.3]. In the case a=δηa=\delta^{\eta}, the resulting bound is stronger but the assumption is also stronger.

Lemma 7.6.

Suppose ν(W)Cνρκ\nu(W)\leq C_{\nu}\rho^{\kappa} for all (ρ,k1)(\rho,k-1)-plates WW, 1>ρ>r1>\rho>r. Then there exists a family of a1\lesssim a^{-1} many (r(Cν/a2)1/κ,k)(r(C_{\nu}/a^{2})^{1/\kappa},k)-plates {Tj}\{T_{j}\} such that every (r,k)(r,k)-plate HH with ν(H)a\nu(H)\geq a is contained in some plate TjT_{j}.

Proof.

Choose a maximal set of (r,k)(r,k)-plates {Yj}j=1m\{Y_{j}\}_{j=1}^{m} such that

  1. 1.

    ν(Yi)a\nu(Y_{i})\geq a,

  2. 2.

    ν(YiYj)a2/2\nu(Y_{i}\cap Y_{j})\leq a^{2}/2 for 1i<jm1\leq i<j\leq m.

We claim m2a1m\leq 2a^{-1}. Indeed, if S=i=1mν(Yi)S=\sum_{i=1}^{m}\nu(Y_{i}) and f=i=1m𝟙Yif=\sum_{i=1}^{m}\mathbbm{1}_{Y_{i}}, then, then

S2=(f𝑑ν)2f2𝑑ν=S+1i<jmν(YiYj)S+m2a2/2.S^{2}=\left(\int f\,d\nu\right)^{2}\leq\int f^{2}\,d\nu=S+\sum_{1\leq i<j\leq m}\nu(Y_{i}\cap Y_{j})\leq S+m^{2}a^{2}/2. (7.28)

Now, Sma>2S\geq ma>2, so S2S>S22S^{2}-S>\frac{S^{2}}{2}. Combining with (7.28) gives S2<m2a2S^{2}<m^{2}a^{2}, a contradiction.

Let {Tj}j=1m\{T_{j}\}_{j=1}^{m} be the (r(Cν/a2)1/κ,k)(r(C_{\nu}/a^{2})^{1/\kappa},k)-plates with same central kk-plane as YjY_{j}. We show the problem condition. Given an (r,k)(r,k)-plate HH with ν(H)a\nu(H)\geq a, by maximality there exists YjY_{j} such that ν(HYj)a2/2\nu(H\cap Y_{j})\geq a^{2}/2. Thus, if (H,Yj)\angle(H,Y_{j}) is the largest principal angle between the central planes of HH and YjY_{j}, then HYjH\cap Y_{j} is contained in a box of dimensions

1××1(k1) times×r/(H,Yj)×r××rdk times.\underbrace{1\times\cdots\times 1}_{(k-1)\text{ times}}\times r/\angle(H,Y_{j})\times\underbrace{r\times\cdots\times r}_{d-k\text{ times}}.

Thus, HYjH\cap Y_{j} is contained in a (r/(H,Yj),k1)(r/\angle(H,Y_{j}),k-1)-plate, so ν(HYj)Cν(r/(H,Yj))κ\nu(H\cap Y_{j})\leq C_{\nu}(r/\angle(H,Y_{j}))^{\kappa}. Thus, (H,Yj)r(Cν/a2)1/κ\angle(H,Y_{j})\lesssim r(C_{\nu}/a^{2})^{1/\kappa}, so HH is contained in TjT_{j}. ∎

Remark 7.7.

We would like to present an alternative proof of Lemma 7.5, which was the original one found by the author. It gives slightly worse bounds but we believe it is slightly more motivated.

If a>1a>1 then =\mathcal{H}=\emptyset, so assume a1a\leq 1. Let ξ=(a2Cν)1/κ1\xi=\left(\frac{a}{2C_{\nu}}\right)^{1/\kappa}\leq 1. By induction, for each 0ik0\leq i\leq k, there exist x0,,xix_{0},\cdots,x_{i} such that |x0x1xi|ξi|x_{0}\wedge x_{1}\wedge\cdots\wedge x_{i}|\geq\xi^{i} and that lie in at least ||(a2)i+1|\mathcal{H}|(\frac{a}{2})^{i+1} many elements of \mathcal{H}.

The base case i=1i=1 is trivial. For the inductive step, suppose x0,,xix_{0},\cdots,x_{i} are found. Let Ω\Omega be the r~\tilde{r}-neighborhood of the span of x1,,xix_{1},\cdots,x_{i}. Then since μ(Ω)Cνξκ<12a\mu(\Omega)\leq C_{\nu}\xi^{\kappa}<\frac{1}{2}a for every HH\in\mathcal{H}, we have μ(HΩ)12a\mu(H\setminus\Omega)\geq\frac{1}{2}a. Thus, there is xi+1\RdΩx_{i+1}\in\R^{d}\setminus\Omega such that x0,,xi+1x_{0},\cdots,x_{i+1} lie in at least ||(a2)(i+2)|\mathcal{H}|(\frac{a}{2})^{(i+2)} many elements of \mathcal{H}, and by construction, |x0x1xi+1|ξi+1|x_{0}\wedge x_{1}\wedge\cdots\wedge x_{i+1}|\geq\xi^{i+1}. This completes the inductive step and thus the proof of the claim.

Finally, the set of (r,k)(r,k)-plates through x1,,xkx_{1},\cdots,x_{k} must lie in a (rξk,k)(r\xi^{-k},k)-plate, so at most ξk(k+1)(dk)\xi^{-k(k+1)(d-k)} many (r,k)(r,k)-plates of r,k\mathcal{E}_{r,k} can lie in it. Thus, ||(a2)(k+1)ξk(k+1)(dk)(Cνa)N|\mathcal{H}|\leq(\frac{a}{2})^{-(k+1)}\xi^{-k(k+1)(d-k)}\lesssim(\frac{C_{\nu}}{a})^{N}.

The following lemma is in the same spirit as [25, Proposition B.2].

Lemma 7.8.

Let \mathcal{H} be a collection of (r,k)(r,k)-plates, and suppose μ(W)Cμρκ\mu(W)\leq C_{\mu}\rho^{\kappa} for all (ρ,k1)(\rho,k-1)-plates WW, 1>ρ>r1>\rho>r. Then for all xXx\in X except a set of μ\mu-measure Cμ(rr0)κ||2\leq C_{\mu}\left(\frac{r}{r_{0}}\right)^{\kappa}|\mathcal{H}|^{2}, there exists an (r0,k)(r_{0},k)-plate that contains every (r,k)(r,k)-plate in \mathcal{H} that passes through xx.

Proof.

The exceptional set is contained in the set of xXx\in X that lies in two plates of \mathcal{H} with “angle” 1r0\geq\frac{1}{r_{0}}. The intersection of two such plates is contained in a box with dimensions r××rdk times×rr0×1××1k1 times\underbrace{r\times\cdots\times r}_{d-k\text{ times}}\times\frac{r}{r_{0}}\times\underbrace{1\times\cdots\times 1}_{k-1\text{ times}}, which in turn is contained in a (rr0,k1)(\frac{r}{r_{0}},k-1)-plate (since r01r_{0}\leq 1). Thus, by assumption on μ\mu, this box has mass Cμ(rr0)s(k1)\lesssim C_{\mu}\left(\frac{r}{r_{0}}\right)^{s-(k-1)}. Finally, there are ||2|\mathcal{H}|^{2} pairs of plates in \mathcal{H}. ∎

Proof of Proposition 7.3.

Fix rr0r\leq r_{0}, and let η=κ4N\eta=\frac{\kappa}{4N}, where NN is the constant in Lemma 7.5. We may assume N2N\geq 2. By Lemmas 7.5 and 7.8, we can find a set ErE_{r} with μ(Er)K2(rr0)η\mu(E_{r})\leq K^{-2}\left(\frac{r}{r_{0}}\right)^{\eta} and, for each xErx\notin E_{r}, a set Pr(x)YP_{r}(x)\subset Y that is either empty or a (r01/2r1/2,k)(r_{0}^{1/2}r^{1/2},k)-plate through xx such that ν(W)K(rr0)η\nu(W)\leq K\left(\frac{r}{r_{0}}\right)^{\eta} for every WW intersecting YPr(x)Y\setminus P_{r}(x).

Now, let E=n0Er0K2nE=\cup_{n\geq 0}E_{r_{0}K^{-2^{n}}} and P(x)=n0Pr0K2n(x)P(x)=\cup_{n\geq 0}P_{r_{0}K^{-2^{n}}}(x). We claim that μ(E)K1\mu(E)\leq K^{-1} and if xEx\notin E, then ν(P(x)A|x)K1\nu(P(x)\setminus A|_{x})\lesssim K^{-1}. Then if rr0K1r\geq r_{0}K^{-1}, then ν(W)1K(rr0)η\nu(W)\leq 1\leq K\left(\frac{r}{r_{0}}\right)^{\eta} for η<1\eta<1; for any r0K2nr<r0K2n1r_{0}K^{-2^{n}}\leq r<r_{0}K^{-2^{n-1}}, we have for any (r,k)(r,k)-plate WW,

ν(WPr0K2n(x))K(r0K2n1r0)ηK(rr0)η/2\nu(W\setminus P_{r_{0}K^{-2^{n}}}(x))\leq K\left(\frac{r_{0}K^{-2^{n-1}}}{r_{0}}\right)^{\eta}\leq K\left(\frac{r}{r_{0}}\right)^{\eta/2}

Then (μ,ν)(\mu,\nu) have (η/2,K2r0η/2,1K1)(\eta/2,K^{2}r_{0}^{-\eta/2},1-K^{-1})-thin kk-plates relative to AA.

To prove the first claim, we observe that μ(E)K2n=0Kη2nK1\mu(E)\leq K^{-2}\sum_{n=0}^{\infty}K^{-\eta 2^{n}}\leq K^{-1} if K0K_{0} is sufficiently large in terms of η\eta.

Next, by definition of Pr0K2n1P_{r_{0}K^{-2^{n-1}}}, we have ν(Pr0K2n(x)Pr0K2n1(x))K(r0K2n1r0)η/2K12n2η\nu(P_{r_{0}K^{-2^{n}}}(x)\setminus P_{r_{0}K^{-2^{n-1}}}(x))\leq K\left(\frac{r_{0}K^{-2^{n-1}}}{r_{0}}\right)^{\eta/2}\leq K^{1-2^{n-2}\eta}. We also have the bound ν(Pr02n(x)A|x)K1\nu(P_{r_{0}2^{-n}}(x)\setminus A|_{x})\leq K^{-1} from the given condition (note that Pr02n(x)P_{r_{0}2^{-n}}(x) is a (r02(n1),k)(r_{0}2^{-(n-1)},k)-plate). Thus,

ν(P(x)A|x)\displaystyle\nu(P(x)\setminus A|_{x}) n=0logη1ν(Pr0K2n(x)A|x)+n=logη1ν(Pr0K2n(x)Pr0K2n1(x))\displaystyle\leq\sum_{n=0}^{\log\eta^{-1}}\nu(P_{r_{0}K^{-2^{n}}}(x)\setminus A|_{x})+\sum_{n=\log\eta^{-1}}^{\infty}\nu(P_{r_{0}K^{-2^{n}}}(x)\setminus P_{r_{0}K^{-2^{n-1}}}(x))
logη1K1+n=logη1K12n2η\displaystyle\leq\log\eta^{-1}\cdot K^{-1}+\sum_{n=\log\eta^{-1}}^{\infty}K^{1-2^{n-2}\eta}
K1,\displaystyle\lesssim K^{-1},

if K0K_{0} is chosen large enough. ∎

8 Radial projection estimates

In this section, we will first prove a key special case, and then the general case of Theorem 1.13.

8.1 Maximal plate concentration case

This subsection is based on ideas from [19].

Theorem 8.1.

Let k{1,2,,d1}k\in\{1,2,\cdots,d-1\}, k1<σ<skk-1<\sigma<s\leq k, and fix K1K\geq 1. There exists NN\in\mathbb{N} and K0K_{0} depending on σ,s,k\sigma,s,k such that the following holds. Fix r01r_{0}\leq 1 and K1,K2K0K_{1},K_{2}\geq K_{0}. Let μ,ν\mu,\nu be 1\sim 1-separated ss-dimensional measures with constant Cμ,CνC_{\mu},C_{\nu} supported on E1,E2E_{1},E_{2}, which lie in an (r0,k)(r_{0},k)-plate HrH_{r}. Assume that |μ|,|ν|1|\mu|,|\nu|\leq 1. Let AA be the pairs of (x,y)E1×E2(x,y)\in E_{1}\times E_{2} that lie in some K11K_{1}^{-1}-concentrated (r0K2,k)(\frac{r_{0}}{K_{2}},k)-plate. Then there exists a set BE1×E2B\subset E_{1}\times E_{2} with μ×ν(B)K11\mu\times\nu(B)\lesssim K_{1}^{-1} such that for every xE1x\in E_{1} and rr-tube TT through xx, we have

μ(T(A|xB|x))rσr0σ(k1)(K1K2)N.\mu(T\setminus(A|_{x}\cup B|_{x}))\lesssim\frac{r^{\sigma}}{r_{0}^{\sigma-(k-1)}}(K_{1}K_{2})^{N}.

The implicit constant may depend on s,ks,k.

Theorem 8.1 is the special case of Theorem 1.13 where (μ,ν)(\mu,\nu) are concentrated in a (r0K,k)(r_{0}K,k)-plate for some small Kr0K\ll r_{0} (we call this the maximal plate concentration case). For this, we closely follow the bootstrapping approach of [19]. There are three ingredients.

  • The next Proposition 8.2 will be the base case for the bootstrapping argument (σ=0\sigma=0).

  • Proposition 7.3 will ensure power decay for μ,ν\mu,\nu around kk-planes.

  • Theorem 1.11 will be used in the bootstrapping step to upgrade σ\sigma to σ+η\sigma+\eta.

Proposition 8.2.

Let 1kd11\leq k\leq d-1 and k1<skk-1<s\leq k, then there exists N=N(s,k)N=N(s,k) such that the following holds. Fix KK0K\geq K_{0}. Then for any ss-dimensional measures μ,ν\mu,\nu with constant 1\sim 1 contained in the r0r_{0}-neighborhood of a kk-plane and d(μ,ν)1d(\mu,\nu)\gtrsim 1, there exists BX×YB\subset X\times Y with μ×ν(B)K1\mu\times\nu(B)\leq K^{-1} such that (μ,ν)(\mu,\nu) has (0,KNr0k1)(0,K^{N}r_{0}^{k-1})-thin tubes on BcB^{c} down from scale r0r_{0}.

Proof.

Let μ~,ν~\tilde{\mu},\tilde{\nu} be the projected measures on the kk-plane. Then μ~,ν~\tilde{\mu},\tilde{\nu} satisfy ss-dimensional Frostman conditions for r0r1r_{0}\leq r\leq 1. Let

B={(x,y):x,yT for some r0-tube T with ν(T)KNr0k1.}B=\{(x,y):x,y\in T\text{ for some }r_{0}\text{-tube }T\text{ with }\nu(T)\geq K^{N}r_{0}^{k-1}.\}

The rest is a standard argument following [8, Proof of Lemma 3.6]. Define the radial projection Py(x)=xy|xy|P_{y}(x)=\frac{x-y}{|x-y|}. Orponen’s radial projection theorem [17, Equation (3.5)] can be written in the form (where p=p(s,k)>1p=p(s,k)>1):

Pxμ~Lpp𝑑μ~(x)1.\int\|P_{x}\tilde{\mu}\|_{L^{p}}^{p}\,d\tilde{\mu}(x)\lesssim 1. (8.29)

To effectively use (8.29), we will show that |Px(B|x)||P_{x}(B|_{x})| is small for xXx\in X. Indeed, let 𝒯x\mathcal{T}_{x} be a minimal set of finitely overlapping 2r02r_{0}-tubes through xx such that any r0r_{0}-tube through xx with ν(T)KNr0k1\nu(T)\geq K^{N}r_{0}^{k-1} lies in a 2r02r_{0}-tube in 𝒯x\mathcal{T}_{x}. Then each 2r02r_{0}-tube in 𝒯x\mathcal{T}_{x} has ν\nu-measure KNr0k1\geq K^{N}r_{0}^{k-1}. Since d(x,ν)1d(x,\nu)\gtrsim 1, we conclude that |𝒯x|KNr01k|\mathcal{T}_{x}|\lesssim K^{-N}r_{0}^{1-k}. Therefore, since the Lebesgue measure |Px(T)|r0k1|P_{x}(T)|\lesssim r_{0}^{k-1} for a 2r02r_{0}-tube TT through xx, we obtain |Px(B|x)|KN|P_{x}(B|_{x})|\lesssim K^{-N}. Finally, we can use Holder’s inequality and (8.29) to upper bound μ×ν(B)\mu\times\nu(B):

μ×ν(B)\displaystyle\mu\times\nu(B) =ν(B|x)𝑑μ(x)\displaystyle=\int\nu(B|_{x})d\mu(x)
=(Px(B|x)Px(ν))𝑑μ(x)\displaystyle=\int\left(\int_{P_{x}(B|_{x})}P_{x}(\nu)\right)d\mu(x)
supx|Px(B|x)|11/pPxνLpdμ(x)\displaystyle\leq\sup_{x}|P_{x}(B|_{x})|^{1-1/p}\int\|P_{x}\nu\|_{L^{p}}d\mu(x)
KN(11/p).\displaystyle\lesssim K^{-N(1-1/p)}.

Choose N=1+(11/p)1N=1+(1-1/p)^{-1} to finish (the implicit constant is dominated by KK0K\geq K_{0} if K0K_{0} is large enough). ∎

The bootstrapping step is as follows:

Proposition 8.3.

Let k{1,,d1}k\in\{1,\cdots,d-1\}, 0σk0\leq\sigma\leq k, max(σ,k1)<sk\max(\sigma,k-1)<s\leq k, κ>0\kappa>0. There exist η(σ,s,κ,k,d)\eta(\sigma,s,\kappa,k,d) and K0(η,k)>0K_{0}(\eta,k)>0 such that the following holds. Fix r01r_{0}\leq 1 and KK0K\geq K_{0}. Let μ,ν\mu,\nu be K1\sim K^{-1}-separated ss-dimensional measures with constant KK supported on X,YX,Y, which lie in an (r0,k)(r_{0},k)-plate HH. Let GX×YG\subset X\times Y. Suppose that (μ,ν)(\mu,\nu) and (ν,μ)(\nu,\mu) have (σ,Kr0(σ(k1)))(\sigma,Kr_{0}^{-(\sigma-(k-1))})-thin tubes and (κ,Kr0κ)(\kappa,Kr_{0}^{-\kappa})-thin kk-plates on GG down from scale r0r_{0}. Then there exists a set BX×YB\subset X\times Y with μ×ν(B)K1\mu\times\nu(B)\leq K^{-1} such that (μ,ν)(\mu,\nu) and (ν,μ)(\nu,\mu) have (σ+η,Kd+1r0(σ+η(k1)))(\sigma+\eta,K^{d+1}r_{0}^{-(\sigma+\eta-(k-1))})-thin tubes on GBG\setminus B down from scale r0r_{0}. Furthermore, η(σ,s,κ,k,d)\eta(\sigma,s,\kappa,k,d) is bounded away from zero on any compact subset of {(σ,s,κ,k):max(σ,k1)<skd1}\{(\sigma,s,\kappa,k):\max(\sigma,k-1)<s\leq k\leq d-1\}.

Remark 8.4.

The reader is advised to set r0=1r_{0}=1 in the following argument, in which case it is a straightforward modification of [19, Lemma 2.8], with one small technical exception in the proof of the concentrated case, where we improve upon the dyadic pigeonholing step. Also if r0=1r_{0}=1, then the simpler Theorem 1.8 can be used instead of Theorem 1.11 in the proof.

Proof.

We are given that for all r(0,r0]r\in(0,r_{0}],

ν(TG|x)Krσr0σ(k1) for all r-tubes T containing xX,\displaystyle\nu(T\cap G|_{x})\leq K\cdot\frac{r^{\sigma}}{r_{0}^{\sigma-(k-1)}}\text{ for all }r\text{-tubes }T\text{ containing }x\in X, (8.30)
ν(WG|x)Krσr0σ(k1) for all (r,k)-plates W containing xX,\displaystyle\nu(W\cap G|_{x})\leq K\cdot\frac{r^{\sigma}}{r_{0}^{\sigma-(k-1)}}\text{ for all }(r,k)\text{-plates }W\text{ containing }x\in X, (8.31)
μ(TG|y)Krσr0σ(k1) for all r-tubes T containing yY,\displaystyle\mu(T\cap G|^{y})\leq K\cdot\frac{r^{\sigma}}{r_{0}^{\sigma-(k-1)}}\text{ for all }r\text{-tubes }T\text{ containing }y\in Y, (8.32)
μ(WG|y)Krσr0σ(k1) for all (r,k)-plates W containing yY.\displaystyle\mu(W\cap G|^{y})\leq K\cdot\frac{r^{\sigma}}{r_{0}^{\sigma-(k-1)}}\text{ for all }(r,k)\text{-plates }W\text{ containing }y\in Y. (8.33)

For xXx\in X and rr0r\leq r_{0}, let 𝒯x,r′′\mathcal{T}^{\prime\prime}_{x,r} denote the rr-tubes through xx such that

ν(TG|x)Kd+1rσ+ηr0σ+η(k1).\nu(T\cap G|_{x})\geq K^{d+1}\cdot\frac{r^{\sigma+\eta}}{r_{0}^{\sigma+\eta-(k-1)}}. (8.34)

Now, let 𝒯x,r\mathcal{T}^{\prime}_{x,r} denote a covering of 𝒯x,r′′\mathcal{T}^{\prime\prime}_{x,r} by essentially distinct 2r2r-tubes. Then for xXx\in X, since d(x,Y)K1d(x,Y)\geq K^{-1}, we have that the tubes in 𝒯x,r\mathcal{T}^{\prime}_{x,r} have Kd1\lesssim K^{d-1}-overlap on ν\nu, so |𝒯x,r|r(σ+η)r0(σ+η(k1))|\mathcal{T}^{\prime}_{x,r}|\lesssim\frac{r^{-(\sigma+\eta)}}{r_{0}^{-(\sigma+\eta-(k-1))}}. For a dyadic r(0,r0]r\in(0,r_{0}], let Hr={(x,y)G:y𝒯x,r}H_{r}=\{(x,y)\in G:y\in\cup\mathcal{T}^{\prime}_{x,r}\}, where 𝒯x,r\cup\mathcal{T}^{\prime}_{x,r} denotes the union of the tubes in 𝒯x,r\mathcal{T}^{\prime}_{x,r}.

Claim. There are η(σ,s,κ,k,d)>0\eta(\sigma,s,\kappa,k,d)>0 and K0(η)>0K_{0}(\eta)>0 such that the following holds for KK0K\geq K_{0}. If rr0<K1/η\frac{r}{r_{0}}<K^{-1/\eta}, then μ×ν(Hr)2(rr0)η\mu\times\nu(H_{r})\leq 2\left(\frac{r}{r_{0}}\right)^{\eta}. Furthermore, η(σ,s,κ,k,d)\eta(\sigma,s,\kappa,k,d) is bounded away from zero on any compact subset of {(σ,s,κ,k,d):max(σ,k1)<skd1}\{(\sigma,s,\kappa,k,d):\max(\sigma,k-1)<s\leq k\leq d-1\}.

We will be done if we show the claim. Indeed, let B1=rr0 dyadic HrB_{1}=\cup_{r\leq r_{0}\text{ dyadic }}H_{r}; then for any dyadic rr0r\leq r_{0} and any rr-tube TT through some xXx\in X, we either have T𝒯x,rT\in\mathcal{T}^{\prime}_{x,r}, which means TG|xB1|x=T\cap G|_{x}\setminus B_{1}|_{x}=\emptyset, or the negation of (8.34) holds. In either case, we get

ν(TG|xB1|x)Kd+1rσ+ηr0σ+η(k1).\nu(T\cap G|_{x}\setminus B_{1}|_{x})\leq K^{d+1}\cdot\frac{r^{\sigma+\eta}}{r_{0}^{\sigma+\eta-(k-1)}}. (8.35)

We have (8.35) for dyadic rr0r\leq r_{0}, but it also holds for all rr0r\leq r_{0} at the cost of introducing a multiplicative factor of 2σ+η2k+12^{\sigma+\eta}\leq 2^{k+1} on the RHS of (8.35). Thus, (μ,ν)(\mu,\nu) have (σ+η,2k+1Kdr0(σ+η(k1)))(\sigma+\eta,2^{k+1}\cdot K^{d}r_{0}^{-(\sigma+\eta-(k-1))})-thin tubes on GB1G\setminus B_{1} down from scale r0r_{0}. Now we move to upper-bounding μ×ν(B1)\mu\times\nu(B_{1}). By (8.30) and (8.34), we have HrH_{r}\neq\emptyset for all r>r0Kd/ηr>r_{0}K^{-d/\eta}, and so if KK0K\geq K_{0} from Claim, then

μ×ν(B1)rr0Kd/η dyadic μ×ν(Hr)rr0Kd/η dyadic 2(rr0)ηCηKd.\mu\times\nu(B_{1})\leq\sum_{r\leq r_{0}K^{-d/\eta}\text{ dyadic }}\mu\times\nu(H_{r})\leq\sum_{r\leq r_{0}K^{-d/\eta}\text{ dyadic }}2\left(\frac{r}{r_{0}}\right)^{\eta}\leq C_{\eta}K^{-d}.

Let K0K_{0} be the maximum of the value of K0K_{0} from Claim, 2Cη2C_{\eta}, and 2k+12^{k+1}. Since d2d\geq 2, we get μ×ν(B1)12K1\mu\times\nu(B_{1})\leq\frac{1}{2}K^{-1} and (μ,ν)(\mu,\nu) have (σ+η,Kd+1r0(σ+η(k1)))(\sigma+\eta,K^{d+1}r_{0}^{-(\sigma+\eta-(k-1))})-thin tubes on GB1G\setminus B_{1} down from scale r0r_{0}. We can analogously find B2X×YB_{2}\subset X\times Y with μ×ν(B2)12K1\mu\times\nu(B_{2})\leq\frac{1}{2}K^{-1} such that (ν,μ)(\nu,\mu) have (σ+η,Kd+1r0(σ+η(k1)))(\sigma+\eta,K^{d+1}r_{0}^{-(\sigma+\eta-(k-1))})-thin tubes on GB2G\setminus B_{2} down from scale r0r_{0}, and so B=B1B2B=B_{1}\cup B_{2} would be a good choice. Now we turn to proving the Claim.

Proof of Claim. We will choose η=min{12(6+15(d1)smax(σ,k1))2,15ε2}\eta=\min\{\frac{1}{2}(6+\frac{15(d-1)}{s-\max(\sigma,k-1)})^{2},\frac{1}{5}\varepsilon^{2}\}, where ε\varepsilon is obtained from Theorem 1.11. From Remark 1.12 and the continuity of the function (s,σ,k)(smax(σ,k1))1(s,\sigma,k)\mapsto(s-\max(\sigma,k-1))^{-1}, we see that η(σ,s,κ,k,d)\eta(\sigma,s,\kappa,k,d) is bounded away from zero on any compact subset of {(σ,s,κ,k,d):max(σ,k1)<skd1}\{(\sigma,s,\kappa,k,d):\max(\sigma,k-1)<s\leq k\leq d-1\}.

Suppose that Claim is false. Let 𝐗={xX:ν(Hr)(rr0)η}\mathbf{X}=\{x\in X:\nu(H_{r})\geq\left(\frac{r}{r_{0}}\right)^{\eta}\}. Then μ(𝐗)(rr0)η\mu(\mathbf{X})\geq\left(\frac{r}{r_{0}}\right)^{\eta}.

Recall that for xXx\in X, the fiber Hr|xH_{r}|_{x} is covered by 𝒯x,r\mathcal{T}^{\prime}_{x,r}, which is a set of cardinality r(σ+η)r0(σ+η(k1))\lesssim\frac{r^{-(\sigma+\eta)}}{r_{0}^{-(\sigma+\eta-(k-1))}}. Let

𝒯x={T𝒯x,r:ν(THr|x)rσ+3ηr0σ+3η(k1)},Yx=(Hr|x)𝒯x.\mathcal{T}_{x}=\{T\in\mathcal{T}^{\prime}_{x,r}:\nu(T\cap H_{r}|_{x})\geq\frac{r^{\sigma+3\eta}}{r_{0}^{\sigma+3\eta-(k-1)}}\},\qquad Y_{x}=(H_{r}|_{x})\cap\bigcup\mathcal{T}_{x}.

Then ν(Yx)(rr0)η(rr0)2η(rr0)2η\nu(Y_{x})\geq\left(\frac{r}{r_{0}}\right)^{\eta}-\left(\frac{r}{r_{0}}\right)^{2\eta}\geq\left(\frac{r}{r_{0}}\right)^{2\eta} for all x𝐗x\in\mathbf{X}. Furthermore, for every T𝒯xT\in\mathcal{T}_{x}, we have

rσ+3ηr0σ+3η(k1)ν(TYx)rσηr0ση(k1).\frac{r^{\sigma+3\eta}}{r_{0}^{\sigma+3\eta-(k-1)}}\leq\nu(T\cap Y_{x})\leq\frac{r^{\sigma-\eta}}{r_{0}^{\sigma-\eta-(k-1)}}. (8.36)

The upper bound follows from YxHr|xG|xY_{x}\subset H_{r}|_{x}\subset G|_{x}, (8.30), and K(rr0)ηK\leq\left(\frac{r}{r_{0}}\right)^{-\eta}. In fact, we have in general,

ν(T(ρ)Yx)(rr0)ηρη,ρ[r,1],T𝒯x.\nu(T^{(\rho)}\cap Y_{x})\leq\left(\frac{r}{r_{0}}\right)^{-\eta}\rho^{\eta},\qquad\rho\in[r,1],T\in\mathcal{T}_{x}.

We also take the time to state the thin plates assumption:

ν(W(ρ)Yx)(rr0)κηρ[r,1],W is (ρ,k)-plate.\nu(W^{(\rho)}\cap Y_{x})\leq\left(\frac{r}{r_{0}}\right)^{\kappa-\eta}\qquad\rho\in[r,1],W\text{ is }(\rho,k)\text{-plate}.

Since 𝒯x\cup\mathcal{T}_{x} covers YxY_{x}, we get by the upper bound in (8.36), |𝒯x|rσ+ηr0σ+η+(k1)ν(Yx)rσ+3ηr0σ+3η+(k1)|\mathcal{T}_{x}|\gtrsim\frac{r^{-\sigma+\eta}}{r_{0}^{-\sigma+\eta+(k-1)}}\nu(Y_{x})\geq\frac{r^{-\sigma+3\eta}}{r_{0}^{-\sigma+3\eta+(k-1)}}. Hence, 𝒯x\mathcal{T}_{x} is a (r,σ,r0(σ(k1))(rr0)5η)(r,\sigma,r_{0}^{-(\sigma-(k-1))}\left(\frac{r}{r_{0}}\right)^{-5\eta})-set and (r,κ,r0κ(rr0)5η,k1)(r,\kappa,r_{0}^{-\kappa}\left(\frac{r}{r_{0}}\right)^{-5\eta},k-1)-set for each x𝐗x\in\mathbf{X}.

Let γ=15ηsmax(σ,k1)\gamma=\frac{15\eta}{s-\max(\sigma,k-1)}. Call a tube T𝒯xT\in\mathcal{T}_{x} concentrated if there is a ball BTB_{T} with radius (rr0)γ\left(\frac{r}{r_{0}}\right)^{\gamma} such that

ν(TBTYx)13ν(TYx).\nu(T\cap B_{T}\cap Y_{x})\geq\frac{1}{3}\cdot\nu(T\cap Y_{x}). (8.37)

Suppose that there is 𝐗𝐗\mathbf{X}^{\prime}\subset\mathbf{X} with μ(𝐗)μ(𝐗)/2\mu(\mathbf{X}^{\prime})\geq\mu(\mathbf{X})/2 such that for each x𝐗x\in\mathbf{X}^{\prime}, at least half the tubes of 𝒯x\mathcal{T}_{x} are non-concentrated. Since μ(𝐗)12μ(𝐗)/212(rr0)2η\mu(\mathbf{X}^{\prime})\geq\frac{1}{2}\mu(\mathbf{X})/2\geq\frac{1}{2}\left(\frac{r}{r_{0}}\right)^{2\eta} and μ\mu is Frostman with constant K(rr0)ηK\leq\left(\frac{r}{r_{0}}\right)^{-\eta}, we can find a (r,σ,(rr0)3η)(r,\sigma,\left(\frac{r}{r_{0}}\right)^{-3\eta})-set P𝐗P\subset\mathbf{X}^{\prime}. For each x𝐗x\in\mathbf{X}^{\prime}, the set of non-concentrated tubes 𝒯x𝒯x\mathcal{T}^{\prime}_{x}\subset\mathcal{T}_{x} is a (r,σ,2r0(σ(k1))(rr0)5η)(r,\sigma,2r_{0}^{-(\sigma-(k-1))}\left(\frac{r}{r_{0}}\right)^{-5\eta})-set and (r,κ,2r0κ(rr0)5η,k1)(r,\kappa,2r_{0}^{-\kappa}\left(\frac{r}{r_{0}}\right)^{-5\eta},k-1)-set. Let 𝒯=xP𝒯x\mathcal{T}=\cup_{x\in P}\mathcal{T}^{\prime}_{x}. By Lemma 2.8, since d(X,Y)K1d(X,Y)\geq K^{-1}, we have that 𝒯\mathcal{T} is contained in the O(K)r0O(K)\cdot r_{0}-neighborhood of HH. Now, we apply Theorem 1.11 with r¯0:=min(O(K)r0,1)\overline{r}_{0}:=\min(O(K)\cdot r_{0},1). Since K(rr0)ηK\leq\left(\frac{r}{r_{0}}\right)^{-\eta} and σk\sigma\leq k, we still have that for each x𝐗x\in\mathbf{X}^{\prime}, the set of non-concentrated tubes 𝒯x\mathcal{T}_{x}^{\prime} is a (r,σ,2r¯0(σ(k1))(rr¯0)7η)(r,\sigma,2\overline{r}_{0}^{-(\sigma-(k-1))}\left(\frac{r}{\overline{r}_{0}}\right)^{-7\eta})-set and (r,κ,2r¯0κ(rr¯0)7η,k1)(r,\kappa,2\overline{r}_{0}^{-\kappa}\left(\frac{r}{\overline{r}_{0}}\right)^{-7\eta},k-1)-set. At this point, let us remark that implicit constants are dominated by (rr¯0)ηKη\left(\frac{r}{\overline{r}_{0}}\right)^{-\eta}\geq K^{\eta} if KK0(η)K\geq K_{0}(\eta) is chosen large enough.

Then if ηε2/4\eta\leq\varepsilon^{2}/4, where ε\varepsilon is obtained from Theorem 1.11, then

|𝒯|r2σ2ηr¯02(σ(k1))2ηr2σηr02(σ(k1))η.|\mathcal{T}|\geq\frac{r^{-2\sigma-2\sqrt{\eta}}}{\overline{r}_{0}^{-2(\sigma-(k-1))-2\sqrt{\eta}}}\geq\frac{r^{-2\sigma-\sqrt{\eta}}}{r_{0}^{-2(\sigma-(k-1))-\sqrt{\eta}}}.

In other words, we get a gain of (rr0)η\left(\frac{r}{r_{0}}\right)^{-\sqrt{\eta}}, which means a two-ends argument gives an immediate contradiction. Specifically, by (8.36) and (8.37), we have for each non-concentrated T𝒯T\in\mathcal{T}, ν×ν({(x,y):x,yT,d(x,y)(rr0)γ})23ν(TYx)2r2σ+6ηr02σ+6η2(k1)\nu\times\nu(\{(x,y):x,y\in T,d(x,y)\geq\left(\frac{r}{r_{0}}\right)^{\gamma}\})\geq\frac{2}{3}\nu(T\cap Y_{x})^{2}\geq\frac{r^{2\sigma+6\eta}}{r_{0}^{2\sigma+6\eta-2(k-1)}}. Thus, by Fubini, there exists a pair (x,y)(x,y) with d(x,y)(rr0)γd(x,y)\geq\left(\frac{r}{r_{0}}\right)^{\gamma} such that x,yTx,y\in T for r2σ+6ηr02σ+6η2(k1)|𝒯|(rr0)η+6η\gtrsim\frac{r^{2\sigma+6\eta}}{r_{0}^{2\sigma+6\eta-2(k-1)}}|\mathcal{T}|\geq\left(\frac{r}{r_{0}}\right)^{-\sqrt{\eta}+6\eta} many tubes T𝒯T\in\mathcal{T}. However, since d(x,y)(rr0)γd(x,y)\geq\left(\frac{r}{r_{0}}\right)^{\gamma}, we have that x,yx,y can only lie in (rr0)(d1)γ\lesssim\left(\frac{r}{r_{0}}\right)^{-(d-1)\gamma} many essentially distinct 2r2r-tubes. Since η6η(d1)γ\sqrt{\eta}-6\eta\geq(d-1)\gamma, we get a contradiction.

Now we focus on the concentrated case: assume there is a subset 𝐗𝐗\mathbf{X}^{\prime}\subset\mathbf{X} with μ(𝐗)μ(𝐗)/2\mu(\mathbf{X}^{\prime})\geq\mu(\mathbf{X})/2 such that at least half of the tubes in 𝒯x\mathcal{T}_{x} are concentrated for all x𝐗x\in\mathbf{X}^{\prime}. This case is where we use the fact that ν\nu is a ss-dimensional measure. Let 𝒯x\mathcal{T}_{x}^{\prime} denote the concentrated tubes and {BT:T𝒯x}\{B_{T}:T\in\mathcal{T}_{x}^{\prime}\} denote the corresponding heavy (rr0)γ\left(\frac{r}{r_{0}}\right)^{\gamma}-balls. Because the family 𝒯x\mathcal{T}_{x} has KK-overlap on spt(ν)\mathrm{spt}(\nu), the set

H={(x,y):x𝐗,yTBTYx for some T𝒯x}H^{\prime}=\{(x,y):x\in\mathbf{X}^{\prime},y\in T\cap B_{T}\cap Y_{x}\text{ for some }T\in\mathcal{T}_{x}^{\prime}\}

has measure

(μ×ν)(H)K1μ(𝐗)infx𝐗|𝒯x|infx𝐗,T𝒯xν(TBTYx)(rr0)2ηrσ+3ηr0(σ3η(k1))rσ+3ηr0σ+3η(k1)=(rr0)8η.(\mu\times\nu)(H^{\prime})\gtrsim K^{-1}\cdot\mu(\mathbf{X}^{\prime})\cdot\inf_{x\in\mathbf{X}^{\prime}}|\mathcal{T}_{x}^{\prime}|\cdot\inf_{x\in\mathbf{X}^{\prime},T\in\mathcal{T}_{x}^{\prime}}\nu(T\cap B_{T}\cap Y_{x})\\ \gtrsim\left(\frac{r}{r_{0}}\right)^{2\eta}\cdot\frac{r^{-\sigma+3\eta}}{r_{0}^{-(\sigma-3\eta-(k-1))}}\cdot\frac{r^{\sigma+3\eta}}{r_{0}^{\sigma+3\eta-(k-1)}}=\left(\frac{r}{r_{0}}\right)^{8\eta}.

Notice that if (x,y)H(x,y)\in H^{\prime}, then there is a tube T(x,y)𝒯rT(x,y)\in\mathcal{T}^{r} containing x,yx,y such that

ν(B(y,2(r/r0)γ)T(x,y))rσ+3ηr0σ+3η(k1).\nu(B(y,2(r/r_{0})^{\gamma})\cap T(x,y))\gtrsim\frac{r^{\sigma+3\eta}}{r_{0}^{\sigma+3\eta-(k-1)}}.

Thus, ν\nu can’t be too concentrated near yy:

ν(B(y,r))Krs12ν(B(y,2(rr0)γ)T(x,y)),\nu(B(y,r))\leq K\cdot r^{s}\leq\frac{1}{2}\nu(B(y,2\left(\frac{r}{r_{0}}\right)^{\gamma})\cap T(x,y)),

assuming 4η<sσ4\eta<s-\sigma and k1<sk-1<s. (The relevant inequalities are K(rr0)ηK\leq\left(\frac{r}{r_{0}}\right)^{-\eta} and rsσ3ηr0sσ3ηr0k1σ3ηr^{s-\sigma-3\eta}\leq r_{0}^{s-\sigma-3\eta}\leq r_{0}^{k-1-\sigma-3\eta}.)

Therefore, for each (x,y)H(x,y)\in H^{\prime}, we can choose a dyadic number rξ(x,y)(r/r0)γr\leq\xi(x,y)\leq(r/r_{0})^{\gamma} such that

ν(A(y,ξ(x,y),2ξ(x,y))T(x,y))(rr0)σ+4η(ξ(x,y)(r/r0)γ)ηr0k,\nu(A(y,\xi(x,y),2\xi(x,y))\cap T(x,y))\geq\left(\frac{r}{r_{0}}\right)^{\sigma+4\eta}\left(\frac{\xi(x,y)}{(r/r_{0})^{\gamma}}\right)^{\eta}r_{0}^{k},

where the annulus A(y,ξ,2ξ):=B(y,2ξ)B(y,ξ)A(y,\xi,2\xi):=B(y,2\xi)\setminus B(y,\xi). (One remark: [19] used dyadic pigeonholing at this step, but we can’t do this because then we would introduce a logr01\log r_{0}^{-1} factor. Fortunately, we are allowed to introduce the decaying tail (ξ(x,y)(r/r0)γ)η\left(\frac{\xi(x,y)}{(r/r_{0})^{\gamma}}\right)^{\eta}, which is summable in ξ(x,y)\xi(x,y).)

Then, recalling that (μ×ν)(H)(rr0)7η(\mu\times\nu)(H^{\prime})\gtrsim\left(\frac{r}{r_{0}}\right)^{7\eta}, we can further find rξ(rr0)γr\leq\xi\leq\left(\frac{r}{r_{0}}\right)^{\gamma} such that

(μ×ν)(H′′)(rr0)8η(ξ(x,y)(r/r0)γ)η, where H′′={(x,y)H:ξ(x,y)=ξ}G.(\mu\times\nu)(H^{\prime\prime})\geq\left(\frac{r}{r_{0}}\right)^{8\eta}\left(\frac{\xi(x,y)}{(r/r_{0})^{\gamma}}\right)^{\eta},\text{ where }H^{\prime\prime}=\{(x,y)\in H^{\prime}:\xi(x,y)=\xi\}\subset G.

By Fubini, we can find yYy\in Y such that μ(H′′|y)(rr0)8η(ξ(x,y)(r/r0)γ)η\mu(H^{\prime\prime}|^{y})\geq\left(\frac{r}{r_{0}}\right)^{8\eta}\left(\frac{\xi(x,y)}{(r/r_{0})^{\gamma}}\right)^{\eta}. Then by construction, H′′|yH^{\prime\prime}|^{y} can be covered by a collection of tubes 𝒯y𝒯r\mathcal{T}_{y}\subset\mathcal{T}^{r} containing yy that satisfy

ν(A(y,ξ,2ξ)T)ν(A(y,ξ(x,y),2ξ(x,y))T(x,y))(rr0)σ+4η(ξ(x,y)(r/r0)γ)ηr0k.\nu(A(y,\xi,2\xi)\cap T)\geq\nu(A(y,\xi(x,y),2\xi(x,y))\cap T(x,y))\geq\left(\frac{r}{r_{0}}\right)^{\sigma+4\eta}\left(\frac{\xi(x,y)}{(r/r_{0})^{\gamma}}\right)^{\eta}r_{0}^{k}.

Finally, we claim that 𝒯y\mathcal{T}_{y} contains a subset 𝒯y\mathcal{T}_{y}^{\prime} whose directions are separated by (r/ξ)\geq(r/\xi), such that |𝒯y|μ(H′′|y)rη(ξr0r)σr0k|\mathcal{T}_{y}^{\prime}|\gtrsim\mu(H^{\prime\prime}|^{y})\cdot r^{\eta}\cdot\left(\frac{\xi r_{0}}{r}\right)^{\sigma}r_{0}^{-k} if ξ>rr0\xi>\frac{r}{r_{0}} and |𝒯y|μ(H′′|y)rη(ξr0r)σr0k|\mathcal{T}_{y}^{\prime}|\gtrsim\mu(H^{\prime\prime}|^{y})\cdot r^{\eta}\cdot\left(\frac{\xi r_{0}}{r}\right)^{\sigma}r_{0}^{-k} if ξ>rr0\xi>\frac{r}{r_{0}} if r<ξ<rr0r<\xi<\frac{r}{r_{0}}. Indeed, if ξ>rr0\xi>\frac{r}{r_{0}}, then any r/ξr/\xi-tube 𝐓\mathbf{T} containing yy has

μ(𝐓H′′|y)μ(𝐓A|y)K(rξr0)σr0k(rr0)η(rξr0)σr0k.\mu(\mathbf{T}\cap H^{\prime\prime}|^{y})\leq\mu(\mathbf{T}\cap A|^{y})\leq K\cdot\left(\frac{r}{\xi r_{0}}\right)^{\sigma}r_{0}^{k}\leq\left(\frac{r}{r_{0}}\right)^{\eta}\cdot\left(\frac{r}{\xi r_{0}}\right)^{\sigma}r_{0}^{k}.

If ξ<rr0\xi<\frac{r}{r_{0}}, then any r/ξr/\xi-tube 𝐓\mathbf{T} containing yy lies in the union of (rξr0)k(\frac{r}{\xi r_{0}})^{k} many r0r_{0}-tubes, and so

μ(𝐓H′′|y)μ(𝐓A|y)K(rξr0)kr0k(rr0)η(rξ)k.\mu(\mathbf{T}\cap H^{\prime\prime}|^{y})\leq\mu(\mathbf{T}\cap A|^{y})\leq K\cdot\left(\frac{r}{\xi r_{0}}\right)^{-k}r_{0}^{k}\leq\left(\frac{r}{r_{0}}\right)^{\eta}\cdot\left(\frac{r}{\xi}\right)^{k}.

Thus, if ξ>rr0\xi>\frac{r}{r_{0}}, then it takes μ(H′′|y)rη(ξr0r)σr0k\gtrsim\mu(H^{\prime\prime}|^{y})\cdot r^{\eta}\cdot\left(\frac{\xi r_{0}}{r}\right)^{\sigma}r_{0}^{-k} many (r/ξ)(r/\xi)-tubes to cover H′′|yH^{\prime\prime}|^{y}, and perhaps even more to cover 𝒯y\cup\mathcal{T}_{y}. We may now choose 𝒯y𝒯y\mathcal{T}_{y}^{\prime}\subset\mathcal{T}_{y} to be a maximal subset with (r/ξ)(r/\xi)-separated directions to prove the claim for ξ>rr0\xi>\frac{r}{r_{0}}. A similar argument holds for ξ<rr0\xi<\frac{r}{r_{0}}.

Finally, let’s first assume ξ>rr0\xi>\frac{r}{r_{0}}. Since 𝒯y\mathcal{T}^{y} has bounded overlap in \RdB(y,ξ)\R^{d}\setminus B(y,\xi), we obtain

(rr0)σ+13η(ξ(r/r0)γ)2η(ξr0r)σr0kinfT𝒯yν(A(y,ξ,2ξ)T)|𝒯y|ν(B(y,2ξ))C(2ξ)s.\left(\frac{r}{r_{0}}\right)^{\sigma+13\eta}\left(\frac{\xi}{(r/r_{0})^{\gamma}}\right)^{2\eta}\cdot\left(\frac{\xi r_{0}}{r}\right)^{\sigma}r_{0}^{-k}\\ \lesssim\inf_{T\in\mathcal{T}_{y}^{\prime}}\nu(A(y,\xi,2\xi)\cap T)\cdot|\mathcal{T}_{y}^{\prime}|\lesssim\nu(B(y,2\xi))\leq C\cdot(2\xi)^{s}.

We will obtain a contradiction if we show the opposite inequality holds, for γ=15ηsmax(k,σ)\gamma=\frac{15\eta}{s-\max(k,\sigma)}. Since 2η+σ<s2\eta+\sigma<s and ξ(rr0)γ\xi\leq(\frac{r}{r_{0}})^{\gamma}, it suffices to check ξ=(rr0)γ\xi=(\frac{r}{r_{0}})^{\gamma}.

If ξ<rr0\xi<\frac{r}{r_{0}}, then we obtain

(rr0)σ+13η(ξ(r/r0)γ)2η(ξr0r)kr0kinfT𝒯yν(A(y,ξ,2ξ)T)|𝒯y|ν(B(y,2ξ))C(2ξ)s.\left(\frac{r}{r_{0}}\right)^{\sigma+13\eta}\left(\frac{\xi}{(r/r_{0})^{\gamma}}\right)^{2\eta}\cdot\left(\frac{\xi r_{0}}{r}\right)^{k}r_{0}^{-k}\\ \lesssim\inf_{T\in\mathcal{T}_{y}^{\prime}}\nu(A(y,\xi,2\xi)\cap T)\cdot|\mathcal{T}_{y}^{\prime}|\lesssim\nu(B(y,2\xi))\leq C\cdot(2\xi)^{s}.

Again, since 2η+k<s2\eta+k<s, it suffices to check ξ=rr0\xi=\frac{r}{r_{0}}.

This proves the result. ∎

Proof of Theorem 8.1.

By Propositions 7.3 (with r0K2\frac{r_{0}}{K_{2}} for r0r_{0}) and 8.2, there exists a set B0X×YB_{0}\subset X\times Y with μ×ν(B0)K11\mu\times\nu(B_{0})\lesssim K_{1}^{-1} such that (μ,ν)(\mu,\nu) and (ν,μ)(\nu,\mu) have (0,K1Nr0k1)(0,K_{1}^{N}r_{0}^{k-1})-thin tubes on B0cB_{0}^{c} down from scale r0r_{0}, and (μ,ν)(\mu,\nu) and (ν,μ)(\nu,\mu) have (κ,K1(r0K2)κ)(\kappa,K_{1}\left(\frac{r_{0}}{K_{2}}\right)^{-\kappa})-thin kk-plates on (AB0)c(A\cup B_{0})^{c}. Then iterate Proposition 8.3 applied to a uniform η(σ,s,κ,k,d)\eta(\sigma,s,\kappa,k,d). So initially we have K=max(K1NK2κ,K0(η,k))K=\max(K_{1}^{N}K_{2}^{\kappa},K_{0}(\eta,k)), and after each iteration, KK becomes Kd+1K^{d+1}. After iterating η1\lesssim\eta^{-1} many times and letting B1X×YB_{1}\subset X\times Y be the union of the BB’s outputted from the Proposition (so μ×ν(B1)K1K11\mu\times\nu(B_{1})\lesssim K^{-1}\leq K_{1}^{-1}), we find that (μ,ν)(\mu,\nu) and (ν,μ)(\nu,\mu) have (σ,K(d+1)η1r0k1)(\sigma,K^{(d+1)\eta^{-1}}r_{0}^{k-1})-thin tubes on (AB0B1)c(A\cup B_{0}\cup B_{1})^{c}. Then we can take B:=B0B1B:=B_{0}\cup B_{1} to be our desired set. ∎

8.2 Proof of Theorem 1.13, general case

We will prove Theorem 1.13, which we restate here.

Theorem 8.5.

Let k{1,2,,d1}k\in\{1,2,\cdots,d-1\}, k1<σ<skk-1<\sigma<s\leq k, and ε>0\varepsilon>0. There exist N,K0N,K_{0} depending on σ,s,k\sigma,s,k, and η(ε)>0\eta(\varepsilon)>0 (with η(1)=1\eta(1)=1) such that the following holds. Fix r01r_{0}\leq 1, and KK0K\geq K_{0}. Let μ,ν\mu,\nu be 1\sim 1-separated ss-dimensional measures with constant Cμ,CνC_{\mu},C_{\nu} supported on E1,E2E_{1},E_{2}, which lie in B(0,1)B(0,1). Assume that |μ|,|ν|1|\mu|,|\nu|\leq 1. Let AA be the pairs of (x,y)E1×E2(x,y)\in E_{1}\times E_{2} that lie in some K1K^{-1}-concentrated (r0,k)(r_{0},k)-plate. Then there exists a set BE1×E2B\subset E_{1}\times E_{2} with μ×ν(B)Kη\mu\times\nu(B)\lesssim K^{-\eta} such that for every xE1x\in E_{1} and rr-tube TT through xx, we have

ν(T(A|xB|x))rσr0σ(k1)+NεKN.\nu(T\setminus(A|_{x}\cup B|_{x}))\lesssim\frac{r^{\sigma}}{r_{0}^{\sigma-(k-1)+N\varepsilon}}K^{N}.

The implicit constant may depend on Cμ,Cν,σ,s,kC_{\mu},C_{\nu},\sigma,s,k.

Remark 8.6.

Note that in Theorem 8.1, we demand the stronger conclusion μ×ν(B)K1\mu\times\nu(B)\lesssim K^{-1}.

The idea is to apply Theorem 8.1 at different scales. As a start, if ε=1\varepsilon=1, then we can directly apply Theorem 8.1 with K1=K2=KK_{1}=K_{2}=K (and thus we may take η(1)=1\eta(1)=1).

We may assume ε=1M\varepsilon=\frac{1}{M} for some MM. Let NN be the large constant in Lemma 7.5, and let ηn=(N+2)nM\eta_{n}=(N+2)^{n-M}. For 1nM1\leq n\leq M, let AnA_{n} be the pairs of (x,y)E1×E2(x,y)\in E_{1}\times E_{2} that lie in some KηnK^{-\eta_{n}}-concentrated (r0nε,k)(r_{0}^{n\varepsilon},k)-plate. We remark that AM=AA_{M}=A.

Lemma 8.7.

Fix n1n\geq 1. There exists a set BnAnB_{n}\subset A_{n} with μ×ν(Bn)Kηn\mu\times\nu(B_{n})\lesssim K^{-\eta_{n}} such that for every xE1x\in E_{1} and rr-tube through xx that intersects An|xA_{n}|_{x}, we have

ν(T(An+1|xB|x))rσr0nε(σ(k1))+NεKN.\nu(T\setminus(A_{n+1}|_{x}\cup B|_{x}))\lesssim\frac{r^{\sigma}}{r_{0}^{n\varepsilon(\sigma-(k-1))+N\varepsilon}}K^{N}.
Proof.

By Lemma 2.8, there exists an absolute constant CC such that every rr-tube through some (x,y)An(x,y)\in A_{n} lies in some KηnK^{-\eta_{n}}-concentrated (CKn,k)(CK^{-n},k)-plate. We can find a collection \mathcal{H} of essentially distinct KηnK^{-\eta_{n}}-concentrated (2CKn,k)(2CK^{-n},k)-plates such that each KηnK^{-\eta_{n}}-concentrated (CKn,k)(CK^{-n},k)-plate is contained within some element of \mathcal{H}. By Lemma 7.5, ||KNηn|\mathcal{H}|\lesssim K^{-N\eta_{n}}. By construction, every rr-tube through some (x,y)An(x,y)\in A_{n} is contained in some member of \mathcal{H}. Apply Theorem 8.1 to each HH\in\mathcal{H} with measures μ|H,ν|H\mu|_{H},\nu|_{H} and K1Kηn+1K_{1}\rightarrow K^{-\eta_{n+1}}, K22Cr0εK_{2}\rightarrow 2Cr_{0}^{-\varepsilon}, and r0r0nεr_{0}\rightarrow r_{0}^{n\varepsilon} to obtain a set BHB_{H} with μ×ν(BH)Kηn+1\mu\times\nu(B_{H})\lesssim K^{-\eta_{n+1}}. Let Bn=HBHB_{n}=\cup_{H\in\mathcal{H}}B_{H}, and then μ×ν(Bn)KNηnKηn+1<Kηn\mu\times\nu(B_{n})\leq K^{N\eta_{n}}\cdot K^{-\eta_{n+1}}<K^{-\eta_{n}} since (N+1)ηn<ηn+1(N+1)\eta_{n}<\eta_{n+1}. ∎

Proof of Theorem 8.5.

Let B=n=1MBnB=\cup_{n=1}^{M}B_{n}; then μ×ν(B)Kη0\mu\times\nu(B)\leq K^{-\eta_{0}}. Fix an rr-tube TT and xE1x\in E_{1}. Let nM1n\leq M-1 be the largest number such that TT passes through points in An|xA_{n}|_{x}. Then by Lemma 8.7, we have ν(T(An+1|xB|x))rσKn(σ(k1)KN\nu(T\setminus(A_{n+1}|_{x}\cup B|_{x}))\lesssim\frac{r^{\sigma}}{K^{-n(\sigma-(k-1)}}K^{N}. If m<M1m<M-1, then TAn+1|x=T\cap A_{n+1}|_{x}=\emptyset. In any case, we have ν(T(A|xB|x))rσr0σ(k1)+NεKN\nu(T\setminus(A|_{x}\cup B|_{x}))\lesssim\frac{r^{\sigma}}{r_{0}^{\sigma-(k-1)+N\varepsilon}}K^{N}, completing the proof of Theorem 8.5.

9 Corollaries of Radial Projection Estimates

We prove a variant of Corollary 1.1.

Proposition 9.1.

Fix s(k1,k]s\in(k-1,k] and η>0\eta>0. Let μ,ν𝒫(\Rd)\mu,\nu\in\mathcal{P}(\R^{d}) be measures with s(μ),s(ν)<\mathcal{E}_{s}(\mu),\mathcal{E}_{s}(\nu)<\infty and 1\sim 1-separated supports. Suppose that μ(H)=ν(H)=0\mu(H)=\nu(H)=0 for each kk-plane H𝔸(\Rd,k)H\in\mathbb{A}(\R^{d},k). Then for μ\mu-almost all xx, for all sets YY of positive ν\nu-measure,

dimH(πxY)sη.\dim_{H}(\pi_{x}Y)\geq s-\eta.
Proof.

The proof is standard and follows [27, Proof of Proposition 6.9]. By Lemma 2.30, by passing to subsets of nearly full measure and replacing ss by an arbitrary s<ss^{\prime}<s, we may assume that μ(Br),ν(Br)rs\mu(B_{r}),\nu(B_{r})\lesssim r^{s} for all r(0,1]r\in(0,1].

Fix ε>0\varepsilon>0. By a compactness argument, there exists r0>0r_{0}>0 such that μ(H),ν(H)<ε\mu(H),\nu(H)<\varepsilon for all (r0,k)(r_{0},k)-plates HH. In Theorem 8.1, we know that for ε>0\varepsilon>0 sufficiently small, the set A=A=\emptyset. Thus, there exists BX×YB\subset X\times Y with μ×ν(B)ε\mu\times\nu(B)\lesssim\varepsilon such that for every xXx\in X and rr-tube through xx, we have

ν(TB|x)η,ε,srsη.\nu(T\setminus B|_{x})\lesssim_{\eta,\varepsilon,s}r^{s-\eta}.

Thus, there is a set XX with μ(X)>1O(ε)\mu(X)>1-O(\varepsilon) such that if xXx\in X, then

dimH(πxY)sη for all Y with ν(Y)O(ε).\dim_{H}(\pi_{x}Y)\geq s-\eta\text{ for all }Y\text{ with }\nu(Y)\geq O(\varepsilon).

Taking ε0\varepsilon\to 0 completes the proof. ∎

Using this, we prove Corollary 1.2.

Corollary 9.2.

Let s(d2,d]s\in(d-2,d], then there exists ε(s,d)>0\varepsilon(s,d)>0 such that the following holds. Let μ,ν\mu,\nu be Borel probability measures on \Rd\R^{d} with disjoint supports that satisfy s(μ),s(ν)<\mathcal{E}_{s}(\mu),\mathcal{E}_{s}(\nu)<\infty and dimH(spt(ν))<s+ε(s,d)\dim_{H}(\mathrm{spt}(\nu))<s+\varepsilon(s,d). Further, assume that μ,ν\mu,\nu don’t simultaneously give full measure to any affine (d1)(d-1)-plane H\RdH\subset\R^{d}. Then there exist restrictions of μ,ν\mu,\nu to subsets of positive measure (which we keep denoting μ,ν\mu,\nu) such that the following holds. For almost every affine 2-plane W\RdW\subset\R^{d} (with respect to the natural measure on the affine Grassmanian), if the sliced measures μW\mu_{W}, νW\nu_{W} on WW is non-trivial, then they don’t simultaneously give full measure to any line. In other words,

(γd,2×μ){(V,x):μV,x()νV,x()=|μV,x||νV,x|>0 for some 𝔸(V+x,1)}=0(\gamma_{d,2}\times\mu)\{(V,x):\mu_{V,x}(\ell)\nu_{V,x}(\ell)=|\mu_{V,x}||\nu_{V,x}|>0\text{ for some }\ell\in\mathbb{A}(V+x,1)\}=0

where we parametrize affine 2-planes as V+xV+x, for x\Rdx\in\R^{d} and VV in the Grassmannian Gr(d,2)\mathrm{Gr}(d,2) with the rotationally invariant Haar measure γd,2\gamma_{d,2}.

Proof.

First, if μ(H)>0\mu(H)>0 for some affine (d1)(d-1)-plane HH, then ν(Hc)>0\nu(H^{c})>0 where HcH^{c} denotes the complement of HH in \Rd\R^{d}. By restricting μ\mu to HH and ν\nu to HcH^{c} (and calling the results μ,ν\mu,\nu), we see that the sliced measures μW\mu_{W} and νW\nu_{W} can’t give full mass to any line \ell for any affine (d1)(d-1)-plane WW, for the simple reason that μW()>0\mu_{W}(\ell)>0 forces H\ell\subset H, and νW()>0\nu_{W}(\ell)>0 forces Hc\ell\subset H^{c}. Likewise, we are done if ν(H)>0\nu(H)>0 for some affine (d1)(d-1)-plane H\RdH\subset\R^{d}. Thus, assume μ(H)=ν(H)=0\mu(H)=\nu(H)=0 for all affine (d1)(d-1)-planes HH.

With this assumption, the remainder of the proof is nearly identical to the proof of Proposition 6.8 in [27], except using Proposition 9.1 instead of [27, Proposition 6.9]. One can take ε(s,d)\varepsilon(s,d) to be arbitrarily close to s(d2)s-(d-2). ∎

Finally, we can deduce Theorem 1.1 from either Proposition 9.1 or Proposition 9.2, see [19, Section 4] for details. The only case not yet considered in this paper is when either μ,ν\mu,\nu gives positive mass to a kk-plane. But this special case was considered in [19, Section 4] (briefly, if XX gives positive mass to some kk-plane, then radial projections become orthogonal projections and then we apply Kaufman’s projection theorem; if YY gives positive mass to some kk-plane HH, then for xHx\notin H, we have dimH(πx(Y))=dimH(Y)\dim_{H}(\pi_{x}(Y))=\dim_{H}(Y).)

Appendix A Proof of Balog-Szemerédi-Gowers

By a standard covering argument (e.g. see Section 3 of [12]), Theorem 2.27 follows from the case δ=0\delta=0, which we prove below.

Theorem A.1 (refined Theorem 4.1 of [30]).

Let K1K\geq 1 be a parameter. Let A,BA,B be finite subsets of \Rn\R^{n}, and let PA×BP\subset A\times B satisfy |P|K1|A||B||P|\geq K^{-1}|A||B|. Suppose that |A+𝑃B|K(|A||B|)1/2|A\overset{P}{+}B|\leq K(|A||B|)^{1/2}, where A+𝑃B={a+b:(a,b)P}A\overset{P}{+}B=\{a+b:(a,b)\in P\}. Then one can find subsets AA,BBA^{\prime}\subset A,B^{\prime}\subset B with |A|116K2|A|,|B|116K2|B||A^{\prime}|\geq\frac{1}{16K^{2}}|A|,|B^{\prime}|\geq\frac{1}{16K^{2}}|B| such that |A+B|212K8(|A||B|)1/2|A^{\prime}+B^{\prime}|\leq 2^{12}K^{8}(|A||B|)^{1/2} and |P(A×B)||A||B|16K2|P\cap(A^{\prime}\times B^{\prime})|\geq\frac{|A||B|}{16K^{2}}.

Proof.

We follow the exposition in [24].

Claim. There exist subsets AA,BBA^{\prime}\subset A,B^{\prime}\subset B with |P(A×B)||A||B|16K2|P\cap(A^{\prime}\times B^{\prime})|\geq\frac{|A||B|}{16K^{2}}, such that for each aA,bBa\in A^{\prime},b\in B^{\prime}, there are |A||B|212K5\geq\frac{|A||B|}{2^{12}K^{5}} many pairs (a,b)A×B(a^{\prime},b^{\prime})\in A\times B such that (a,b)(a,b^{\prime}), (a,b)(a^{\prime},b^{\prime}), and (a,b)P(a^{\prime},b)\in P.

Assuming the claim, we will see how the theorem follows. First, we get |A||B||P(A×B)||A||B|16K2|A^{\prime}||B^{\prime}|\geq|P\cap(A^{\prime}\times B^{\prime})|\geq\frac{|A||B|}{16K^{2}}. Since |A||A||A^{\prime}|\leq|A| and |B||B||B^{\prime}|\leq|B|, we get |A||A|16K2|A|\geq\frac{|A|}{16K^{2}} and |B||B|16K2|B|\geq\frac{|B|}{16K^{2}}.

Next, for aA,bBa\in A^{\prime},b\in B^{\prime}, we have

a+b=(a+b)(a+b)+(a+b).a+b=(a+b^{\prime})-(a^{\prime}+b^{\prime})+(a^{\prime}+b).

Thus, there are |A||B|212K5\geq|A||B|2^{-12}K^{-5} many solutions to a+b=xy+za+b=x-y+z with x,y,zA+𝑃Bx,y,z\in A\overset{P}{+}B. Since |A+𝑃B|K(|A||B|)1/2|A\overset{P}{+}B|\leq K(|A||B|)^{1/2}, we get |A+B|K3(|A||B|)3/2|A||B|212K5=212K8|A|1/2|B|1/2|A^{\prime}+B^{\prime}|\lesssim\frac{K^{3}(|A||B|)^{3/2}}{|A||B|2^{-12}K^{-5}}=2^{12}K^{8}|A|^{1/2}|B|^{1/2}.

Now, we prove the claim. For convenience, we can prune PP to satisfy |P|=K1|A||B||P|=K^{-1}|A||B| (this is not necessary but will make the proof look nicer). Treat (AB,P)(A\cup B,P) as a bipartite graph with an edge between aAa\in A and bBb\in B if (a,b)P(a,b)\in P. Then we want to find A,BA^{\prime},B^{\prime} such that there are many paths of length 33 between any aA,bBa\in A^{\prime},b\in B^{\prime}.

The average degree of a vertex in AA is K1|B|K^{-1}|B|. Thus, if we delete the vertices in AA with degree 12K1|B|\leq\frac{1}{2}K^{-1}|B|, then at least 12K|A||B|\frac{1}{2K}|A||B| many edges remain. Let EE be the set of edges. For vABv\in A\cup B, let N(v)N(v) be the set of neighbors of vv.

Now pick a vertex bBb\in B. On average, it has |E||B|12K|A|\frac{|E|}{|B|}\geq\frac{1}{2K}|A| many neighbors.

Now, we say (a,a)A2(a,a^{\prime})\in A^{2} is bad if |N(a)N(a)|<1128K3|B||N(a)\cap N(a^{\prime})|<\frac{1}{128K^{3}}|B|. For vBv\in B, let Badv\mathrm{Bad}_{v} be the set of bad pairs in N(v)2N(v)^{2}. There are (|A|2)\binom{|A|}{2} many pairs in AA, so (expectation is taken over uniformly chosen vBv\in B)

𝔼[|Badv|]<(|A|2)1128K3<|A|2256K3.\mathbb{E}[|\mathrm{Bad}_{v}|]<\binom{|A|}{2}\cdot\frac{1}{128K^{3}}<\frac{|A|^{2}}{256K^{3}}.

If Abad,vA_{bad,v} is the set of vertices of AA that lie in at least |A|32K2\frac{|A|}{32K^{2}} many pairs of BvB_{v}, then

𝔼[|Abad,v|]2𝔼[|Bv|]|A|/(32K2)<|A|4K.\mathbb{E}[|A_{bad,v}|]\leq\frac{2\mathbb{E}[|B_{v}|]}{|A|/(32K^{2})}<\frac{|A|}{4K}.

Finally, let Av=N(v)Abad,vA_{v}=N(v)\setminus A_{bad,v}. Then by linearity of expectation,

𝔼[|Av|]=𝔼[|N(v)|]𝔼[|Abad,v|]>|A|2K|A|4K=|A|4K.\mathbb{E}[|A_{v}|]=\mathbb{E}[|N(v)|]-\mathbb{E}[|A_{bad,v}|]>\frac{|A|}{2K}-\frac{|A|}{4K}=\frac{|A|}{4K}.

Thus, there exists vBv\in B such that |Av|>|A|4K|A_{v}|>\frac{|A|}{4K}. Then, let A=AvA^{\prime}=A_{v} and

B={wB:|N(w)A||A|16K2.B^{\prime}=\{w\in B:|N(w)\cap A^{\prime}|\geq\frac{|A|}{16K^{2}}.

Let E(X,Y)E(X,Y) be the number of edges between XX and YY. We first check that E(A,B)|A||B|16K2E(A^{\prime},B^{\prime})\geq\frac{|A||B|}{16K^{2}}. Indeed, since every vertex of AA has degree |B|2K\geq\frac{|B|}{2K}, we have

|E(A,B)||A||B|2K|A||B|28K2.|E(A^{\prime},B)|\geq\frac{|A^{\prime}||B|}{2K}\geq\frac{|A||B|^{2}}{8K^{2}}.

On the other hand, every vertex in BBB\setminus B^{\prime} corresponds to fewer than |A|16K2\frac{|A|}{16K^{2}} many edges of AA^{\prime}, so |E(A,BB)||A||B|216K2|E(A^{\prime},B\setminus B^{\prime})|\leq\frac{|A||B|^{2}}{16K^{2}}. Hence, |E(A,B)||A||B|216K2|E(A^{\prime},B^{\prime})|\geq\frac{|A||B|^{2}}{16K^{2}}.

Finally, for any vAv\in A^{\prime}, wBw\in B^{\prime}, we know that ww has at least |A|16K2\frac{|A|}{16K^{2}} many neighbors in AA^{\prime}, and fewer than |A|32K2\frac{|A|}{32K^{2}} of those form a bad pair with ww. For the remaining |A|32K2\geq\frac{|A|}{32K^{2}} vertices vv^{\prime} that do not form a bad pair with ww, there are |B|128K3\geq\frac{|B|}{128K^{3}} many vertices wBw^{\prime}\in B that are common neighbors of v,vv,v^{\prime}. Thus, we get at least |A|32K|B|128K3=|A||B|212K5\frac{|A|}{32K}\cdot\frac{|B|}{128K^{3}}=\frac{|A||B|}{2^{12}K^{5}} many paths (v,w,v,w)(v,w^{\prime},v^{\prime},w) between vv and ww. ∎

References

  • [1] Jean Bourgain. The discretized sum-product and projection theorems. Journal d’Analyse Mathématique, 112(1):193–236, 2010.
  • [2] A. Carbery and S.I. Valdimarsson. The endpoint multilinear Kakeya theorem via the Borsuk–Ulam theorem. Journal of Functional Analysis, 264(7):1643–1663, 2013.
  • [3] X. Du, A. Iosevich, Y. Ou, H. Wang, and R. Zhang. An improved result for Falconer’s distance set problem in even dimensions. Mathematische Annalen, 380(3):1215–1231, 2021.
  • [4] X. Du, Y. Ou, K. Ren, and R. Zhang. New improvement to Falconer distance set problem in higher dimensions. arXiv preprint, 2023.
  • [5] X. Du, Y. Ou, K. Ren, and R. Zhang. Weighted refined decoupling estimates and application to Falconer distance set problem. arXiv preprint, 2023.
  • [6] X. Du and R. Zhang. Sharp l2l^{2} estimates of the Schrödinger maximal function in higher dimensions. Annals of Mathematics, 189(3):837–861, 2019.
  • [7] L. Guth, A. Iosevich, Y. Ou, and H. Wang. On Falconer’s distance set problem in the plane. Inventiones Mathematicae, 219(3):779–830, 2020.
  • [8] L. Guth, N. Solomon, and H. Wang. Incidence estimates for well spaced tubes. Geometric and Functional Analysis, 29(6):1844–1863, 2019.
  • [9] T.L.J. Harris. Low-dimensional pinned distance sets via spherical averages. The Journal of Geometric Analysis, 31(11):11410–11416, 2021.
  • [10] W. He. Orthogonal projections of discretized sets. Journal of Fractal Geometry, 7(3):271–317, 2020.
  • [11] Weikun He. Discretized sum-product estimates in matrix algebras. Journal d’Analyse Mathématique, 139(2):637–676, 2019.
  • [12] Nets Hawk Katz and Terence Tao. Some connections between falconer’s distance set conjecture and sets of furstenburg type. New York J. Math, 7:149–187, 2001.
  • [13] R. Kaufman. On Hausdorff dimension of projections. Mathematika, 15(2):153–155, 1968.
  • [14] Tamás Keleti and Pablo Shmerkin. New bounds on the dimensions of planar distance sets. Geometric and Functional Analysis, 29(6):1886–1948, 2019.
  • [15] B. Liu. Hausdorff dimension of pinned distance sets and the L2{L}^{2}-method. Proceedings of the American Mathematical Society, 148(1):333–341, 2020.
  • [16] P. Mattila. Geometry of sets and measures in Euclidean spaces: fractals and rectifiability. Number 44. Cambridge university press, 1999.
  • [17] T. Orponen. On the dimension and smoothness of radial projections. Analysis & PDE, 12(5):1273–1294, 2018.
  • [18] T. Orponen and P. Shmerkin. On the Hausdorff dimension of Furstenberg sets and orthogonal projections in the plane. arXiv preprint arXiv:2106.03338, 2021.
  • [19] T. Orponen, P. Shmerkin, and H. Wang. Kaufman and Falconer estimates for radial projections and a continuum version of Beck’s Theorem. arXiv preprint arXiv:2209.00348, 2022.
  • [20] Tuomas Orponen. An improved bound on the packing dimension of furstenberg sets in the plane. J. Eur. Math. Soc.(JEMS), 22(3):797–831, 2020.
  • [21] Tuomas Orponen and Pablo Shmerkin. Projections, Furstenberg sets, and the abcabc sum-product problem, 2023.
  • [22] Kevin Ren and Hong Wang. Furstenberg sets estimate in the plane, 2023.
  • [23] I.Z. Ruzsa. On the cardinality of A+A{A}+{A} and AA{A}-{A}. In Combinatorics (Proc. Fifth Hungarian Colloq., Keszthely, 1976), volume 2, pages 933–938, 1978.
  • [24] Adam Sheffer. The Balog-Szemerédi-Gowers theorem, 2016.
  • [25] P. Shmerkin. A non-linear version of Bourgain’s projection theorem. Journal of the European Mathematical Society, 2022.
  • [26] P. Shmerkin and H. Wang. On the distance sets spanned by sets of dimension d/2d/2 in d\mathbb{R}^{d}. arXiv preprint arXiv:2112.09044, 2021.
  • [27] P. Shmerkin and H. Wang. Dimensions of Furstenberg sets and an extension of Bourgain’s projection theorem. arXiv preprint arXiv:2211.13363, 2022.
  • [28] Pablo Shmerkin. Improved bounds for the dimensions of planar distance sets. J. Fractal Geom., 8(1):27–51, 2021.
  • [29] D.M. Stull. Pinned distance sets using effective dimension. arXiv preprint arXiv:2207.12501, 2022.
  • [30] B. Sudakov, E. Szemerédi, and V.H. Vu. On a question of Erdős and Moser. 2005.
  • [31] T. Tao and V.H. Vu. Additive combinatorics, volume 105. Cambridge University Press, 2006.
  • [32] Z. Wang and J. Zheng. An improvement of the pinned distance set problem in even dimensions. In Colloquium Mathematicum, volume 170, pages 171–191. Instytut Matematyczny Polskiej Akademii Nauk, 2022.
  • [33] T. Wolff. Decay of circular means of Fourier transforms of measures. Int. Math. Res. Not., 10:547–567, 1999.
  • [34] J. Zahl. Unions of lines in n\mathbb{R}^{n}. arXiv preprint arXiv:2208.02913, 2022.