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

New improvement to Falconer distance set problem in higher dimensions

Xiumin Du, Yumeng Ou, Kevin Ren and Ruixiang Zhang
Abstract.

We show that if a compact set EdE\subset\mathbb{R}^{d} has Hausdorff dimension larger than d2+1418d+4\frac{d}{2}+\frac{1}{4}-\frac{1}{8d+4}, where d3d\geq 3, then there is a point xEx\in E such that the pinned distance set Δx(E)\Delta_{x}(E) has positive Lebesgue measure. This improves upon bounds of Du–Zhang and Du–Iosevich–Ou–Wang–Zhang in all dimensions d3d\geq 3. We also prove lower bounds for Hausdorff dimension of pinned distance sets when dimH(E)(d21438d+4,d2+1418d+4)\dim_{H}(E)\in(\frac{d}{2}-\frac{1}{4}-\frac{3}{8d+4},\frac{d}{2}+\frac{1}{4}-\frac{1}{8d+4}), which improves upon bounds of Harris and Wang–Zheng in dimensions d3d\geq 3.

1. Introduction

A classical question in geometric measure theory, introduced by Falconer in the early 80s ([9]) is, how large does the Hausdorff dimension of a compact subset of d{\mathbb{R}}^{d}, d2d\geq 2 need to be to ensure that the Lebesgue measure of the set of its pairwise Euclidean distances is positive.

Let EdE\subset\mathbb{R}^{d} be a compact set, its distance set Δ(E)\Delta(E) is defined by

Δ(E):={|xy|:x,yE}.\Delta(E):=\{|x-y|:x,y\in E\}\,.
Conjecture.

[Falconer] Let d2d\geq 2 and EdE\subset\mathbb{R}^{d} be a compact set. Then

dimH(E)>d2|Δ(E)|>0.{\dim_{H}}(E)>\frac{d}{2}\Rightarrow|\Delta(E)|>0.

Here |||\cdot| denotes the Lebesgue measure and dimH(){\dim_{H}}(\cdot) is the Hausdorff dimension.

The main result in this paper improves the best-known dimensional threshold towards the Falconer conjecture in dimensions d3d\geq 3.

Theorem 1.1.

Let d3d\geq 3 and EdE\subset\mathbb{R}^{d} be a compact set. Then

dimH(E)>d2+1418d+4|Δ(E)|>0.{\dim_{H}}(E)>\frac{d}{2}+\frac{1}{4}-\frac{1}{8d+4}\Rightarrow|\Delta(E)|>0.

Falconer’s conjecture remains open in all dimensions as of today. It has attracted a great amount of attention in the past decades. To name a few landmarks: in 1985, Falconer [9] showed that |Δ(E)|>0|\Delta(E)|>0 if dimH(E)>d2+12{\dim_{H}}(E)>\frac{d}{2}+\frac{1}{2}. Bourgain [1] was the first to lower the threshold d2+12\frac{d}{2}+\frac{1}{2} in dimensions d=2,d=3d=2,d=3 and to use the theory of Fourier restriction in the Falconer problem. The thresholds were further improved by Wolff [29] to 43\frac{4}{3} in the case d=2d=2, and by Erdoğan [8] to d2+13\frac{d}{2}+\frac{1}{3} when d3d\geq 3. These records were only very recently rewritten:

{54,d=2,(Guth–Iosevich–Ou–Wang [10])95,d=3,(Du–Guth–Ou–Wang–Wilson–Zhang [4])d2+14+18d4,d3,(Du–Zhang [7])d2+14,d4 even,(Du–Iosevich–Ou–Wang–Zhang [5]).\begin{cases}\frac{5}{4},&d=2,\quad\qquad\text{(Guth--Iosevich--Ou--Wang \cite[cite]{[\@@bibref{}{guth2020falconer}{}{}]})}\\ \frac{9}{5},&d=3,\quad\qquad\text{(Du--Guth--Ou--Wang--Wilson--Zhang \cite[cite]{[\@@bibref{}{DGOWWZ}{}{}]})}\\ \frac{d}{2}+\frac{1}{4}+\frac{1}{8d-4},&d\geq 3,\quad\qquad\text{(Du--Zhang \cite[cite]{[\@@bibref{}{du2019sharp}{}{}]})}\\ \frac{d}{2}+\frac{1}{4},&d\geq 4\text{ even},\quad\text{(Du--Iosevich--Ou--Wang--Zhang \cite[cite]{[\@@bibref{}{du2021improved}{}{}]})}.\end{cases}

Our main result in this paper further improves the thresholds in all dimensions d3d\geq 3. Now, the gap between the best-known threshold and the conjectured one is 1418d+4\frac{1}{4}-\frac{1}{8d+4} when d3d\geq 3. This is the first time that the gap gets smaller than 14\frac{1}{4} and the first time that the gap in higher dimensions is smaller than that in dimension d=2d=2. Similar to [10, 5], we in fact prove a slightly stronger version of the main theorem regarding the pinned distance set.

Theorem 1.2.

Let d3d\geq 3 and EdE\subset\mathbb{R}^{d} be a compact set. Suppose that dimH(E)>d2+1418d+4\dim_{H}(E)>\frac{d}{2}+\frac{1}{4}-\frac{1}{8d+4}. Then there is a point xEx\in E such that the pinned distance set Δx(E)\Delta_{x}(E) has positive Lebesgue measure, where

Δx(E):={|xy|:yE}.\Delta_{x}(E):=\{|x-y|:\,y\in E\}.

We also get a Hausdorff dimension version if dimH(E)\dim_{H}(E) is lower than the threshold. Let f(α)=α2d+1d+1(d1)f(\alpha)=\alpha\cdot\frac{2d+1}{d+1}-(d-1). Then the following theorem gives a non-trivial result for dimH(E)(d212d+1,d2+d2d+1]\dim_{H}(E)\in(\frac{d^{2}-1}{2d+1},\frac{d^{2}+d}{2d+1}].

Theorem 1.3.

Let d2d\geq 2, 0<αd10<\alpha\leq d-1, and EdE\subset\mathbb{R}^{d} be a compact set. Suppose that dimH(E)=α\dim_{H}(E)=\alpha. Then for any ε>0\varepsilon>0,

dimH({xE:dimH(Δx(E))min(f(α),1)ε})<α.\dim_{H}(\{x\in E:\dim_{H}(\Delta_{x}(E))\leq\min(f(\alpha),1)-\varepsilon\})<\alpha.

In particular, supxEdimH(Δx(E))min(f(α),1)\sup_{x\in E}\dim_{H}(\Delta_{x}(E))\geq\min(f(\alpha),1), and if α(E)>0\mathcal{H}^{\alpha}(E)>0, then dimH(Δx(E))min(f(α),1)\dim_{H}(\Delta_{x}(E))\geq\min(f(\alpha),1) for α\mathcal{H}^{\alpha}-almost all xEx\in E.

In dimension d=2d=2, a lot of recent progress has been made. Liu [15] showed that if dimH(E)=α>1\dim_{H}(E)=\alpha>1, then supxEdimH(Δx(E))min(4α323,1)\sup_{x\in E}\dim_{H}(\Delta_{x}(E))\geq\min(\frac{4\alpha}{3}-\frac{2}{3},1). For small values of α>1\alpha>1, Shmerkin [24] improved this bound by using entropy, and later, Stull [27] (see also [fiedler2023dimension]) made a further improvement, he showed using algorithmic complexity theory that dimH(Δx(E))α4+12\dim_{H}(\Delta_{x}(E))\geq\frac{\alpha}{4}+\frac{1}{2} for many xx. Theorem 1.3 deals with the case dimH(E)1\dim_{H}(E)\leq 1.

When d3,dimH(E)=α>d2d\geq 3,\dim_{H}(E)=\alpha>\frac{d}{2}, the previous best-known results are as follows: Harris [12] proved that supxEdimH(Δx(E))min(2d1dα(d1),1)\sup_{x\in E}\dim_{H}(\Delta_{x}(E))\geq\min(\frac{2d-1}{d}\alpha-(d-1),1); for even dimensions d4d\geq 4, Wang–Zheng [28] improved this result to supxEdimH(Δx(E))min(2dd+1α2d2d22(d+1),1)\sup_{x\in E}\dim_{H}(\Delta_{x}(E))\geq\min(\frac{2d}{d+1}\alpha-\frac{2d^{2}-d-2}{2(d+1)},1). Theorem 1.3 further improves upon these two results.

However, getting a lower bound for dimH(Δx(E))\dim_{H}(\Delta_{x}(E)) given dimH(E)=d2\dim_{H}(E)=\frac{d}{2} is much more challenging. Falconer in [9] proved that dimH(Δ(E))12\dim_{H}(\Delta(E))\geq\frac{1}{2}, and a pinned version supxEdimH(Δx(E))12\sup_{x\in E}\dim_{H}(\Delta_{x}(E))\geq\frac{1}{2} was proved in [17]. The reason 12\frac{1}{2} is a natural barrier is that if there existed a 12\frac{1}{2}-dimensional ring R[1,2]R\subset[1,2], then the distance set of RdR^{d} is contained in R\sqrt{R}, so it has Hausdorff dimension 12\frac{1}{2}. Even though the works of Bourgain [2] and Katz–Tao [13] showed a quantitative discretized sum-product theorem that rules out the existence of RR, it is still a challenging problem to obtain explicit bounds for the discretized sum-product problem, which asks for the largest exponent γ>0\gamma>0 such that min(|A+A|δ,|AA|δ)|A|1+γ\min(|A+A|_{\delta},|A\cdot A|_{\delta})\gtrsim|A|^{1+\gamma} for any Katz–Tao (δ,12)(\delta,\frac{1}{2})-set A1A\subset\mathbb{R}^{1} (see [13] for the definition of such sets). See [11], [20], [16], and [23] for the best known bounds for the discretized sum-product problem.

For the Falconer distance set problem, the only explicit improvements known over 12\frac{1}{2} for dimH(Δ(E))\dim_{H}(\Delta(E)) when dimH(E)=d2\dim_{H}(E)=\frac{d}{2} were derived in [25] and [26] for d=2,d=3d=2,d=3, and the latter paper also proved box dimension results for d4d\geq 4. In personal communication, Shmerkin–Wang extended Stull’s bound in d=2d=2 to the case dimH(E)=1\dim_{H}(E)=1. Furthermore, by plugging in the sharp radial projection estimates of [19] into the proofs of Theorems 1.2 and 1.3 of [26], we get improved bounds over those stated in these theorems for d3d\geq 3. In summary, the previously known results are:

  • if d=2d=2, dimH(E)=1\dim_{H}(E)=1, then supxEdimH(Δx(E))34\sup_{x\in E}\dim_{H}(\Delta_{x}(E))\geq\frac{3}{4};

  • if d=3d=3, dimH(E)=32\dim_{H}(E)=\frac{3}{2}, then supxEdimH(Δx(E))58\sup_{x\in E}\dim_{H}(\Delta_{x}(E))\geq\frac{5}{8};

  • if d4d\geq 4, dimH(E)=d2\dim_{H}(E)=\frac{d}{2}, then supxEdimB(Δx(E))d+22(d+1)\sup_{x\in E}\dim_{B}(\Delta_{x}(E))\geq\frac{d+2}{2(d+1)}.

A key obstruction to Hausdorff dimension estimates in [26] when d4d\geq 4 is that d2\frac{d}{2}-dimensional measures don’t necessarily have decay around small neighborhoods of (d1)(d-1)-planes when d4d\geq 4, so one cannot apply Bourgain’s discretized projection theorem to such measures.

Note that in Theorem 1.3, f(d2)=d+22(d+1)>12f(\frac{d}{2})=\frac{d+2}{2(d+1)}>\frac{1}{2}. This matches the best-known bound 58\frac{5}{8} when d=3d=3 using an entirely different approach. And when d4d\geq 4, this is the first time one obtains an explicit improved bound over 12\frac{1}{2} for the Hausdorff dimension of pinned distance sets of a d2\frac{d}{2}-dimensional set, (and this matches the box dimension bound as in the above).

Finally, let us remark what is known when α<d2\alpha<\frac{d}{2}. Falconer [9] proved that dimH(Δ(E))min(αd12,0)\dim_{H}(\Delta(E))\geq\min(\alpha-\frac{d-1}{2},0); this was improved in dimensions d=2d=2 and 33 to supxEdimH(Δx(E))α+1d+1\sup_{x\in E}\dim_{H}(\Delta_{x}(E))\geq\frac{\alpha+1}{d+1} for α(d12,d2)\alpha\in(\frac{d-1}{2},\frac{d}{2}) by [19], [26]. In dimension d4d\geq 4 however, their approach only recovers box dimension estimates. Our bound supxEdimH(Δx(E))f(α)\sup_{x\in E}\dim_{H}(\Delta_{x}(E))\geq f(\alpha) from Theorem 1.3 is weaker than α+1d+1\frac{\alpha+1}{d+1} for all α<d2\alpha<\frac{d}{2}, but it works for Hausdorff dimension for all d3d\geq 3.

Old and new ideas

We will extend and refine the good tube/bad tube and decoupling method pioneered by [10] for dimension d=2d=2 and continued in [5] for even dimensions dd. To state our contributions, we find it helpful to first recall the good tube/bad tube decomposition in [10], [5]. Fix α>d2\alpha>\frac{d}{2} and a set EE with dimH(E)>α\dim_{H}(E)>\alpha; thus, we can find two α\alpha-dimensional Frostman measures μ1,μ2\mu_{1},\mu_{2} supported on separated subsets of EE. Then we fix a single scale rr, and consider the set of rr-tubes (i.e. rr-neighborhoods of lines) that intersect EE. We say that an rr-tube TT is good if μ2(T)rd2ε\mu_{2}(T)\lesssim r^{\frac{d}{2}-\varepsilon}, and bad otherwise. The paper [10] developed a decoupling framework that can handle good tubes very well, so it remains to show that there are very few bad tubes. When d=2d=2, this was solved by appealing to the n=2n=2 case of Orponen’s radial projection theorem, which roughly states that given a Frostman measure μ\mu in n\mathbb{R}^{n} with dimension >n1>n-1, most tubes satisfy μ2(T)rn1\mu_{2}(T)\lesssim r^{n-1}. When d>2d>2 is even, we can no longer directly apply the radial projection theorem because α\alpha is usually d1\leq d-1; to overcome this [5] came up with a partial fix by first projecting μ2\mu_{2} onto a generic (d2+1)(\frac{d}{2}+1)-dimensional subspace of d\mathbb{R}^{d} (assuming dd is even), so the projected measure μ2\mu_{2*} still remains α\alpha-dimensional, and then applying Orponen’s radial projection theorem to μ2\mu_{2*} in d2+1\mathbb{R}^{\frac{d}{2}+1}. Despite the seemingly wasteful application of the initial orthogonal projection, the threshold rd2εr^{\frac{d}{2}-\varepsilon} is sharp up to ε\varepsilon: the example of a (d2+1)(\frac{d}{2}+1)-hyperplane shows that rd2r^{\frac{d}{2}} is the best threshold one can get in general (and we get an even worse threshold rd12r^{\frac{d-1}{2}} for odd dimensions via an (d+12)(\frac{d+1}{2})-hyperplane example; see e.g. Section 5.2 of [5]).

To improve upon [5], we must find a way to deal with this hyperplane example. As it so happens, we are elated if this situation occurs: if our set EE is contained in a lower-dimensional hyperplane H=nH=\mathbb{R}^{n}, then we can simply work in n\mathbb{R}^{n} where existing distance set results are already better than what we need. Thus, it would be nice to establish a dichotomy: either we can improve upon the barrier rd2r^{\lfloor\frac{d}{2}\rfloor}, or it turns out that much of EE is contained in a lower-dimensional hyperplane. Fortunately, this wish is indeed true and is one of our main innovations of the paper. More precisely, we establish a framework, for the first time in the context of the distance problems, that analyzes how much a fractal measure concentrates near a low dimensional plane hence allows us to reduce the problem to the aforementioned dichotomy. We show that either (1) the measure is dictated by wave packets (i.e. microlocalized pieces of the measure) that display good transversality. In this case, desirable multilinear estimates or projection results can apply to deduce an improved threshold rαϵr^{\alpha-\epsilon} (a significant improvement over the barrier rd2r^{\lfloor\frac{d}{2}\rfloor} and would lead to gains in the decoupling step), or (2) the problem can be reduced to lower dimensions hence existing distance set results can apply.

On the technical level, this is achieved by a careful analysis of the interaction between the plates of intermediate dimensions (i.e. neighborhood of planes) and wave packets. Our starting point is a new radial projection theorem of the third author [22], which we precisely state as Theorem 3.1. Roughly, it states that most rr-tubes either have μ2\mu_{2}-mass rαε\lesssim r^{\alpha-\varepsilon} or lie in some heavy (rκ,α)(r^{\kappa},\lceil\alpha\rceil)-plate, which is the rκr^{\kappa}-neighborhood of a α\lceil\alpha\rceil-dimensional hyperplane with (μ1+μ2)(\mu_{1}+\mu_{2})-mass rη\gtrsim r^{\eta} (here, ηκε\eta\ll\kappa\ll\varepsilon). The main obstacle that we need to deal with then becomes the heavy (rκ,α)(r^{\kappa},\lceil\alpha\rceil)-plates, and a key novelty of this paper is an upgrade of the good/bad framework that analyzes the interaction between these plates and the wave packets.

Note that this strategy in fact indicates a key difference between the distance problem in two dimensions and that in higher dimensions: in dimension three and higher, it is possible to study the distance set using information from the distance problem in lower dimensions. This is apparently not possible in the plane, in which case the measure is assumed to have Hausdorff dimension greater than one hence cannot concentrate on a line. As far as we are aware of, it seems that such a special feature of the distance problem in the higher dimensional setting has not been explored in the literature before.

Here are some more details, where we fix d=3d=3 for concreteness. Following the setup in [5], we let μ1,μ2\mu_{1},\mu_{2} be α\alpha-dimensional Frostman measures supported on subsets E1,E2EE_{1},E_{2}\subset E with dist(E1,E2)1\text{dist}(E_{1},E_{2})\gtrsim 1. Let R0R_{0} be a large number and define the scales Rj=2jR0R_{j}=2^{j}R_{0}. Fix a scale jj; we shall work with Rj1/2+βR_{j}^{-1/2+\beta}-tubes. Let G(x)G(x) to be the set of yE1y\in E_{1} such that x,yx,y don’t both lie in some heavy (Rjκ,2)(R_{j}^{-\kappa},2)-plate. Define a good Rj1/2+βR_{j}^{-1/2+\beta}-tube to be one with mass Rjα/2+ε\lesssim R_{j}^{-\alpha/2+\varepsilon} and doesn’t lie in any heavy plate, and μ1,g\mu_{1,g} to be a part of μ1\mu_{1} from all good tubes. With these definitions, we can try to follow the framework of [5] and show that μ1|G(x)\mu_{1}|_{G(x)} and μ1,g\mu_{1,g} are close in L1L^{1}-norm. There are two subtleties with this approach. First, the error in μ1|G(x)μ1,gL1\|\mu_{1}|_{G(x)}-\mu_{1,g}\|_{L^{1}} may contain contributions from tubes TT at the border of G(x)G(x), i.e. TG(x)T\subset G(x) but 2TG(x)2T\not\subset G(x). To overcome this issue, we introduce a probabilistic wiggle. If we work with heavy (aRjκ,2)(aR_{j}^{-\kappa},2)-plates for a uniformly random a[1,2]a\in[1,2], then the borderline error will be small on average, so there exists a choice for aa to make the borderline error small.

The second issue is that we do not have the luxury of working at a single scale, so we need to introduce some ideas from [25, Appendix B]. Specifically, we will construct a decreasing sequence 3G0(x)G1(x)\mathbb{R}^{3}\supset G_{0}(x)\supset G_{1}(x)\supset\cdots such that 3G0(x)\mathbb{R}^{3}\setminus G_{0}(x) is contained in a (R0κ/2,2)(R_{0}^{-\kappa/2},2)-plate, μ(G0(x)G(x))R0κ/2\mu(G_{0}(x)\setminus G_{\infty}(x))\leq R_{0}^{-\kappa/2}, and Gj(x)G_{j}(x) is disjoint from all heavy (aRjκ,2)(aR_{j}^{-\kappa},2)-plates containing xx (where a[1,2]a\in[1,2] is the probabilistic wiggle). Since all the GjG_{j}’s are close in measure, we can combine multiscale information to show that μ1|G0(x)\mu_{1}|_{G_{0}(x)} and μ1,g\mu_{1,g} are close in L1L^{1}-norm.

The final step is to show G0(x)G_{0}(x) is large. This is equivalent to showing that any (R0κ/2,2)(R_{0}^{-\kappa/2},2)-plate has small (12\leq\frac{1}{2}) measure for R0R_{0} large enough. If this is not true, then by compactness we can find a hyperplane with nonzero measure, and then we reduce to Falconer in 2D. In this case, we have strong bounds for distance sets (e.g. see Wolff [29]) because dimH(E)>1.5\dim_{H}(E)>1.5 is large.

Remark 1.4.

In dimensions d4d\geq 4, a slightly simpler and more intuitive choice of good measures exists. We refer the interested reader to [6] for more details, where such a direction is pursued and same results in Theorems 1.1 and 1.2 (except for the d=3d=3 case) are obtained as a consequence of new weighted refined decoupling estimates. In dimension d=3d=3 though, the strategy still works but doesn’t seem to yield as good a result as the current paper.

We point out that, in some sense, the dimensional threshold we obtained already fully exploited the power of the methods of this paper and the companion paper [6]. The radial projection theorem of [22] is sharp in the following sense: let E1,E2E_{1},E_{2} each be unions of rαr^{-\alpha} many rr-balls satisfying an α\alpha-dimensional spacing condition such that dist(E1,E2)12\mathrm{dist}(E_{1},E_{2})\geq\frac{1}{2}, and let μ1,μ2\mu_{1},\mu_{2} be probability measures supported on E1,E2E_{1},E_{2} respectively such that each rr-ball in EiE_{i} has μi\mu_{i}-measure rαr^{\alpha}, i=1,2i=1,2. Then, we know that many rr-tubes can intersect at least one rr-ball in E1E_{1} and one rr-ball in E2E_{2}. Thus, many rr-tubes can have both μ1\mu_{1} and μ2\mu_{2}-mass at least rαr^{\alpha}, so rαr^{\alpha} is the best possible threshold for the good tubes in this paper. To make further progress, we suggest to look at improving the decoupling framework.

On the other hand, the threshold rαr^{\alpha} is already stronger than all the previously best known thresholds for good tubes in all dimensions d2d\geq 2 (see [10, 5]). For comparison, in two dimensions, it was shown that one can reduce to good rr-tubes whose measure is at most r\sim r, which is weaker than rαr^{\alpha}. The possibility of obtaining such a better threshold for the good tubes underscores another key difference between the distance problem in higher dimensions and that in the plane. Indeed, one can easily see that for an evenly distributed probablity measure in the plane, the threshold rr would be the best possible.

Outline of the paper

In Section 2, we outline two main estimates and prove Theorem 1.2 using them. In Section 3, we list several results that we will use to bound the bad part, including the new radial projection estimate by the third author [22] and two geometric lemmas governing the physical location of small plates with large mass. In Section 4, we construct the good measure and prove the first main estimate - Proposition 2.1. In Section 5, we prove the second main estimate - Proposition 2.2 using refined decoupling. In Section 6, we prove Theorem 1.3 using the two main estimates and a framework of Liu [15]. In Section 7, we give some remarks about the extension of Theorem 1.2 to more general norms and its connection with the Erdős distance problem.

Notations

Throughout the article, we write ABA\lesssim B if ACBA\leq CB for some absolute constant CC; ABA\sim B if ABA\lesssim B and BAB\lesssim A; AεBA\lesssim_{\varepsilon}B if ACεBA\leq C_{\varepsilon}B; ABA\lessapprox B if ACεRεBA\leq C_{\varepsilon}R^{\varepsilon}B for any ε>0\varepsilon>0, R>1R>1.

For a large parameter RR, RapDec(R){\rm RapDec}(R) denotes those quantities that are bounded by a huge (absolute) negative power of RR, i.e. RapDec(R)CNRN{\rm RapDec}(R)\leq C_{N}R^{-N} for arbitrarily large N>0N>0. Such quantities are negligible in our argument.

For xdx\in\mathbb{R}^{d} and t>0t>0, B(x,t)B(x,t) is the ball in d\mathbb{R}^{d} of radius tt centered at xx. A δ\delta-tube is the intersection of an infinite cylinder of radius δ\delta and B(0,10)B(0,10). This is not standard (usually, a tube is defined to be a finite δ\delta-cylinder with length 11), but this definition won’t cause problems and is slightly more convenient for us.

For a set EdE\subset\mathbb{R}^{d}, let Ec=dEE^{c}=\mathbb{R}^{d}\setminus E, and E(δ)E^{(\delta)} the δ\delta-neighborhood of EE. For subsets E1,E2dE_{1},E_{2}\subset\mathbb{R}^{d}, dist(E1,E2)\mathrm{dist}(E_{1},E_{2}) is their Euclidean distance.

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\}. Similar definition for A|yA|_{y}, when yYy\in Y.

For a measure μ\mu and a measurable set GG, define the restricted measure μ|G\mu|_{G} by μ|G(A)=μ(GA)\mu|_{G}(A)=\mu(G\cap A) for all measurable AdA\subset\mathbb{R}^{d}.

We say a measure μ\mu supported in d\mathbb{R}^{d} is an α\alpha-dimensional measure with constant CμC_{\mu} if it is a probability measure satisfying that

μ(B(x,t))Cμtα,xd,t>0.\mu(B(x,t))\leq C_{\mu}t^{\alpha},\qquad\forall x\in\mathbb{R}^{d},\,\forall t>0.

An (r,k)(r,k)-plate HH in d\mathbb{R}^{d} is the rr-neighborhood of a kk-dimensional affine plane in the ball Bd(0,10)B^{d}(0,10). More precisely,

H:={zB(0,10):dist(z,PH)<r},H:=\{z\in B(0,10):\mathrm{dist}(z,P_{H})<r\},

where PHP_{H} is a kk-dimensional affine plane, which is called the central plane of HH. We can also write H=PH(r)B(0,10)H=P_{H}^{(r)}\cap B(0,10). A CC-scaling of HH is

CH={zB(0,10):dist(z,PH)<Cr}.CH=\{z\in B(0,10):\mathrm{dist}(z,P_{H})<Cr\}.

Denote the surface of HH by

Surf(H):={zB(0,10):dist(z,PH)=r}.\mathrm{Surf}(H):=\{z\in B(0,10):\mathrm{dist}(z,P_{H})=r\}\,.

Since a δ\delta-tube is a (δ,1)(\delta,1)-plate, the same conventions apply to tubes.

To prevent the reader from feeling too attached to B(0,10)B(0,10), we will see in the proofs that B(0,10)B(0,10) can be replaced by any smooth bounded convex body that contains B(0,10)B(0,10).

We say that an (r,k)(r,k)-plate HH is γ\gamma-concentrated on μ\mu if μ(H)γ\mu(H)\geq\gamma.

Given a collection \mathcal{H} of plates in d\mathbb{R}^{d} and a point xdx\in\mathbb{R}^{d}, define (x):={H:xaH}\mathcal{H}(x):=\{H\in\mathcal{H}:x\in aH\}, where aa is a “master parameter” that will be defined later.

We will work with a collection r,k\mathcal{E}_{r,k} of essentially distinct (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 k,d(sr)(k+1)(dk)\lesssim_{k,d}\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, which can be found in [22, Section 2.2].

Acknowledgements.

XD is supported by NSF DMS-2107729 (transferred from DMS-1856475), NSF DMS-2237349 and Sloan Research Fellowship. YO is supported by NSF DMS-2142221 and NSF DMS-2055008. KR is supported by a NSF GRFP fellowship. RZ is supported by NSF DMS-2207281 (transferred from DMS-1856541), NSF DMS-2143989 and the Sloan Research Fellowship.

2. Main estimates

In this section, we outline two main estimates, from which Theorems 1.2 and 1.3 follow.

Let EdE\subset\mathbb{R}^{d} be a compact set with positive α\alpha-dimensional Hausdorff measure. Without loss of generality, assume that EE is contained in the unit ball, and there are subsets E1E_{1} and E2E_{2} of EE, each with positive α\alpha-dimensional Hausdorff measure, and dist(E1,E2)1\mathrm{dist}(E_{1},E_{2})\gtrsim 1. Then there exist α\alpha-dimensional probability measures μ1\mu_{1} and μ2\mu_{2} supported on E1E_{1} and E2E_{2}, respectively.

To relate the measures to the distance set, we consider their pushforward measures under the distance map. For a fixed point xE2x\in E_{2}, let dx:E1d^{x}:E_{1}\to\mathbb{R} be the pinned distance map given by dx(y):=|xy|d^{x}(y):=|x-y|. Then, the pushforward measure dx(μ1)d^{x}_{\ast}(\mu_{1}), defined as

ψ(t)dx(μ1)(t)=E1ψ(|xy|)𝑑μ1(y),\int_{\mathbb{R}}\psi(t)\,d^{x}_{\ast}(\mu_{1})(t)=\int_{E_{1}}\psi(|x-y|)\,d\mu_{1}(y),

is a natural measure that is supported on Δx(E1)\Delta_{x}(E_{1}).

In the following, we will construct another complex-valued measure μ1,gx\mu_{1,g}^{x} that is the good part of μ1\mu_{1} with respect to μ2\mu_{2} depending on xx, and study its pushforward under the map dxd^{x}. The main estimates are the following.

Proposition 2.1.

Let d2d\geq 2, k{1,2,,d1}k\in\{1,2,\cdots,d-1\}, k1<αkk-1<\alpha\leq k, and ε>0\varepsilon>0. Then there exists a small β(ε)>0\beta(\varepsilon)>0 such that the following holds for sufficiently large R0(β,ε)R_{0}(\beta,\varepsilon). Assume μ1,gx\mu_{1,g}^{x} has been constructed following the procedure in Section 4.2 below. Then there is a subset E2E2E_{2}^{\prime}\subset E_{2} with μ2(E2)1R0β\mu_{2}(E_{2}^{\prime})\geq 1-R_{0}^{-\beta} and for each xE2x\in E_{2}^{\prime}, there exists a set G(x)Bd(0,10)G(x)\subset B^{d}(0,10) where Bd(0,10)G(x)B^{d}(0,10)\setminus G(x) is contained within some (R0β,k)(R_{0}^{-\beta},k)-plate, such that the following estimate holds:

(2.1) dx(μ1|G(x))dx(μ1,gx)L1R0β.\|d_{*}^{x}(\mu_{1}|_{G(x)})-d_{*}^{x}(\mu_{1,g}^{x})\|_{L^{1}}\leq R_{0}^{-\beta}.
Proposition 2.2.

Let d2d\geq 2, 0<α<d0<\alpha<d, and ε>0\varepsilon>0. Then for sufficiently small β(ε)(0,ε)\beta(\varepsilon)\in(0,\varepsilon) in the construction of μ1,gx\mu_{1,g}^{x} in Section 4.2 below,

E2dx(μ1,gx)L22𝑑μ2(x)d,α,εd|ξ|αdd+1+ε|μ^1(ξ)|2𝑑ξ+R0d.\int_{E_{2}}\|d^{x}_{*}(\mu_{1,g}^{x})\|_{L^{2}}^{2}d\mu_{2}(x)\lesssim_{d,\alpha,\varepsilon}\int_{\mathbb{R}^{d}}|\xi|^{-\frac{\alpha d}{d+1}+\varepsilon}|\widehat{\mu}_{1}(\xi)|^{2}\,d\xi+R_{0}^{d}.
Remark 2.3.

Broadly speaking, G(x)G(x) removes the contributions from small plates that contain large mass. This makes it possible to efficiently apply the new radial projection theorem, Theorem 3.1.

Proof of Theorem 1.2 using Propositions 2.1 and 2.2.

For d3d\geq 3 and d2<α<d+12\frac{d}{2}<\alpha<\frac{d+1}{2}, let k=d2+1k=\lfloor\frac{d}{2}\rfloor+1 so that k1<αkk-1<\alpha\leq k. If μ1\mu_{1} gives nonzero mass to some kk-dimensional affine plane, then we are done by applying [9] to that kk-plane since α>d2k+12\alpha>\frac{d}{2}\geq\frac{k+1}{2}. Thus, assume μ1\mu_{1} gives zero mass to every kk-dimensional affine plane. By a compactness argument, there exists r0>0r_{0}>0 such that μ1(H)<11000\mu_{1}(H)<\frac{1}{1000} for any (r0,k)(r_{0},k)-plate HH. Now the two propositions tell us that there is a point xE2x\in E_{2} such that

dx(μ1|G(x))dx(μ1,gx)L111000,\displaystyle\|d_{*}^{x}(\mu_{1}|_{G(x)})-d_{*}^{x}(\mu_{1,g}^{x})\|_{L^{1}}\leq\frac{1}{1000},
dx(μ1,gx)L22Iλ(μ1)+R0d<,\displaystyle\|d_{*}^{x}(\mu_{1,g}^{x})\|^{2}_{L^{2}}\lesssim I_{\lambda}(\mu_{1})+R_{0}^{d}<\infty,

by choosing R0R_{0} sufficiently large. Here Iλ(μ1)I_{\lambda}(\mu_{1}) is the λ\lambda-dimensional energy of μ1\mu_{1} and λ=dαdd+1+ε\lambda=d-\frac{\alpha d}{d+1}+\varepsilon, by a Fourier representation for IλI_{\lambda}:

Iλ(μ)=|xy|λ𝑑μ(x)𝑑μ(y)=Cd,λd|ξ|λd|μ^(ξ)|2𝑑ξ.I_{\lambda}(\mu)=\int\int|x-y|^{-\lambda}d\mu(x)d\mu(y)=C_{d,\lambda}\int_{\mathbb{R}^{d}}|\xi|^{\lambda-d}|\widehat{\mu}(\xi)|^{2}\,d\xi.

One has Iλ(μ1)<I_{\lambda}(\mu_{1})<\infty if λ<α\lambda<\alpha, which is equivalent to α>d(d+1)2d+1=d2+1418d+4\alpha>\frac{d(d+1)}{2d+1}=\frac{d}{2}+\frac{1}{4}-\frac{1}{8d+4}. Now Bd(0,10)G(x)B^{d}(0,10)\setminus G(x) is contained in some (R0β,k)(R_{0}^{-\beta},k)-plate HH. If R0R_{0} is chosen sufficiently large such that R0β>r01R_{0}^{\beta}>r_{0}^{-1}, then

μ1(G(x))1μ1(H)>111000.\mu_{1}(G(x))\geq 1-\mu_{1}(H)>1-\frac{1}{1000}.

Since dx(μ1|G(x))d_{*}^{x}(\mu_{1}|_{G(x)}) is a positive measure, its L1L^{1} norm is μ1(G(x))>111000\mu_{1}(G(x))>1-\frac{1}{1000}. Thus,

Δx(E)|dx(μ1,gx)|=|dx(μ1,gx)|Δx(E)c|dx(μ1,gx)|121000|dx(μ1|G(x))dx(μ1,gx)|131000.\int_{\Delta_{x}(E)}|d_{*}^{x}(\mu_{1,g}^{x})|=\int|d_{*}^{x}(\mu_{1,g}^{x})|-\int_{\Delta_{x}(E)^{c}}|d_{*}^{x}(\mu_{1,g}^{x})|\\ \geq 1-\frac{2}{1000}-\int|d_{*}^{x}(\mu_{1}|_{G(x)})-d_{*}^{x}(\mu_{1,g}^{x})|\geq 1-\frac{3}{1000}.

On the other hand,

Δx(E)|dx(μ1,gx)||Δx(E)|1/2(|dx(μ1,gx)|2)1/2.\int_{\Delta_{x}(E)}|d_{*}^{x}(\mu_{1,g}^{x})|\leq|\Delta_{x}(E)|^{1/2}\left(\int|d_{*}^{x}(\mu_{1,g}^{x})|^{2}\right)^{1/2}.

Therefore, |Δx(E)|>0|\Delta_{x}(E)|>0. ∎

The proof of Theorem 1.3 is deferred until Section 6.

3. Radial projections and heavy plates

In this section, we list several results that we will use in Sections 4.4 and 4.6 to bound the bad part of μ1\mu_{1}.

We’ll use the following new radial projection estimate, which follows from [22, Theorem 1.13].

Theorem 3.1.

Let d2d\geq 2, k{1,2,,d1}k\in\{1,2,\cdots,d-1\}, k1<αkk-1<\alpha\leq k, and fix η,ε>0\eta,\varepsilon>0, and two α\alpha-dimensional measures μ1,μ2\mu_{1},\mu_{2} with constants Cμ1,Cμ2C_{\mu_{1}},C_{\mu_{2}} supported on E1,E2B(0,1)E_{1},E_{2}\subset B(0,1) respectively. There exists γ>0\gamma>0 depending on η,ε,α,k\eta,\varepsilon,\alpha,k such that the following holds. Fix δ<r<1\delta<r<1. Let AA be the set of pairs (x,y)E1×E2(x,y)\in E_{1}\times E_{2} satisfying that xx and yy lie in some δη\delta^{\eta}-concentrated (r,k)(r,k)-plate on μ1+μ2\mu_{1}+\mu_{2}. Then 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 δ\delta-tube TT through xx, we have

μ2(T(A|xB|x))δαrα(k1)δε.\mu_{2}(T\setminus(A|_{x}\cup B|_{x}))\lesssim\frac{\delta^{\alpha}}{r^{\alpha-(k-1)}}\delta^{-\varepsilon}.

The implicit constant may depend on η,ε,α,k,Cμ1,Cμ2\eta,\varepsilon,\alpha,k,C_{\mu_{1}},C_{\mu_{2}}.

Remark 3.2.

(a) The roles of μ1\mu_{1} and μ2\mu_{2} in Theorem 3.1 are interchangeable, so the conclusion also holds for μ1\mu_{1} instead of μ2\mu_{2}.

(b) If α>d1\alpha>d-1, then the numerology of Theorem 3.1 doesn’t apply. Instead, Orponen’s radial projection theorem [18] in dimension dd applies. The result (stated in [10, 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)δγ\mu_{1}\times\mu_{2}(B)\leq\delta^{\gamma} such that for every xE1x\in E_{1} and δ\delta-tube TT through xx, we have

μ2(TB|x)δd1ε.\mu_{2}(T\setminus B|_{x})\lesssim\delta^{d-1-\varepsilon}.

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

(c) If rδr\sim\delta, we can obtain a slightly better result by projecting to a kk-dimensional subspace and following the argument in [5, 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 δ\delta-tube TT through xx, we have

μ2(TB|x)δk1ε.\mu_{2}(T\setminus B|_{x})\lesssim\delta^{k-1-\varepsilon}.

The set AA is again not needed in this case. The main novelty of Theorem 3.1 comes when r>δr>\delta.

We will also need the following two lemmas from [22] (Lemmas 7.5 and 7.8) governing the physical location of small plates with large mass.

Lemma 3.3.

Let k1<skk-1<s\leq k and 0<r10<r\leq 1. There is N=N(s,k,d)N=N(s,k,d) such that the following holds: let ν\nu be an ss-dimensional measure with constant Cν1C_{\nu}\geq 1, and let r,k\mathcal{E}_{r,k} be the collection of essentially distinct (r,k)(r,k)-plates from the Notations part of Section 1. 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}. (The implicit constant only depends on k,dk,d and is independent of a,ra,r.)

Lemma 3.4.

Let 0<r<r010<r<r_{0}\lesssim 1 and s>k1s>k-1. Let \mathcal{H} be a collection of (r,k)(r,k)-plates, and let μ\mu be a compactly supported ss-dimensional measure with constant CμC_{\mu}. Then for all xspt(μ)x\in\mathrm{spt}\,(\mu) except a set of μ\mu-measure Cμ(rr0)s(k1)||2\lesssim C_{\mu}\left(\frac{r}{r_{0}}\right)^{s-(k-1)}|\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.

4. Construction of good measure and Proposition 2.1

In this section, we will construct the good measure μ1,gx\mu_{1,g}^{x} and prove Proposition 2.1. We will henceforth treat α,k,d,ε\alpha,k,d,\varepsilon in the hypothesis of Proposition 2.1 as fixed constants. To assist in the proof, we will eventually be choosing the following parameters: ε0(ε)\varepsilon_{0}(\varepsilon), κ(ε0)\kappa(\varepsilon_{0}), η(κ,ε0)\eta(\kappa,\varepsilon_{0}), β(κ,η,ε0)\beta(\kappa,\eta,\varepsilon_{0}). In terms of size, they satisfy

0<βηκε0ε.0<\beta\ll\eta\ll\kappa\ll\varepsilon_{0}\ll\varepsilon.

Here, ABA\ll B means “AA is much smaller than BB.” Unwrapping the dependences, we see that β\beta ultimately only depends on ε\varepsilon, which is what we want.

4.1. Smooth Partitions of Unity

We follow the first part of [5, Section 3.1]. Let R0R_{0} be a large power of 22 that will be determined later, and let Rj=2jR0R_{j}=2^{j}R_{0}. Construct a partition of unity

1=j0ψj,1=\sum_{j\geq 0}\psi_{j},

where ψ0\psi_{0} is supported in the ball |ω|2R0|\omega|\leq 2R_{0} and each ψj\psi_{j} for j1j\geq 1 is supported on the annulus Rj1|ω|Rj+1R_{j-1}\leq|\omega|\leq R_{j+1}. Importantly, we may choose ψj\psi_{j} such that ψˇjL1C\|\check{\psi}_{j}\|_{L^{1}}\leq C for some absolute constant CC and all j1j\geq 1. For example, choose ψj\psi_{j} to be Littlewood-Paley functions χ(x/Rj)χ(x/Rj1)\chi(x/R_{j})-\chi(x/R_{j-1}), where χ\chi is a smooth bump function that is 11 on B(0,1)B(0,1) and 0 outside B(0,2)B(0,2).

In d\mathbb{R}^{d}, cover the annulus Rj1|ω|Rj+1R_{j-1}\leq|\omega|\leq R_{j+1} by rectangular blocks τ\tau of dimensions approximately Rj1/2××Rj1/2×RjR_{j}^{1/2}\times\cdots\times R_{j}^{1/2}\times R_{j}, with the long direction of each block τ\tau being the radial direction. Choose a smooth “partition of unity” with respect to this cover such that

ψj=τψj,τ(ω).\psi_{j}=\sum_{\tau}\psi_{j,\tau}(\omega).

The functions ψj,τ\psi_{j,\tau} satisfy the following properties:

  • ψj,τ\psi_{j,\tau} is supported on τ\tau and ψj,τL1\|\psi_{j,\tau}\|_{L^{\infty}}\leq 1;

  • ψˇj,τ\check{\psi}_{j,\tau} is essentially supported on a Rj1/2××Rj1/2×Rj1R_{j}^{-1/2}\times\cdots\times R_{j}^{-1/2}\times R_{j}^{-1} box KK centered at 0, in the sense that |ψˇj,τ(x)|RapDec(Rj)|\check{\psi}_{j,\tau}(x)|\leq\mathrm{RapDec}(R_{j}) if dist(x,K)Rj1/2+β\mathrm{dist}(x,K)\gtrsim R_{j}^{-1/2+\beta} (and the implicit constant in the decay CNRjNC_{N}R_{j}^{-N} is universal only depending on NN);

  • ψˇj,τL11\|\check{\psi}_{j,\tau}\|_{L^{1}}\lesssim 1 (the implicit constant is universal).

For each (j,τ)(j,\tau), cover the unit ball in d\mathbb{R}^{d} with tubes TT of dimensions approximately Rj1/2+β××Rj1/2+β×20R_{j}^{-1/2+\beta}\times\cdots\times R_{j}^{-1/2+\beta}\times 20 with the long axis parallel to the long axis of τ\tau. The covering has uniformly bounded overlap, each TT intersects at most C(d)C(d) other tubes. We denote the collection of all these tubes as 𝕋j,τ\mathbb{T}_{j,\tau}. Let ηT\eta_{T} be a smooth partition of unity subordinate to this covering, so that for each choice of jj and τ\tau, T𝕋j,τηT\sum_{T\in\mathbb{T}_{j,\tau}}\eta_{T} is equal to 1 on the ball of radius 10 and each ηT\eta_{T} is smooth.

For each T𝕋j,τT\in\mathbb{T}_{j,\tau}, define an operator

MTf:=ηT(ψj,τf^),M_{T}f:=\eta_{T}(\psi_{j,\tau}\hat{f})^{\vee},

which, morally speaking, maps ff to the part of it that has Fourier support in τ\tau and physical support in TT. Define also M0f:=(ψ0f^)M_{0}f:=(\psi_{0}\hat{f})^{\vee}. We denote 𝕋j=τ𝕋j,τ\mathbb{T}_{j}=\cup_{\tau}\mathbb{T}_{j,\tau} and 𝕋=j1𝕋j\mathbb{T}=\cup_{j\geq 1}\mathbb{T}_{j}. Hence, for any L1L^{1} function ff supported on the unit ball, one has the decomposition

f=M0f+T𝕋MTf+RapDec(R0)fL1.f=M_{0}f+\sum_{T\in\mathbb{T}}M_{T}f+\text{RapDec}(R_{0})\|f\|_{L^{1}}.

See [10, Lemma 3.4] for a justification of the above decomposition. (Even though [10, Lemma 3.4] is stated in two dimensions, the argument obviously extends to higher dimensions.)

4.2. Heavy Plates and Good Tubes

In this subsection, we define good tubes. Actually, we will use three categories: good, acceptable, and non-acceptable. The idea is that Theorem 3.1 tells us that Rj1/2+βR_{j}^{-1/2+\beta}-tubes fall into one of three categories:

  • For RjηR_{j}^{-\eta}-concentrated (Rjκ,k)(R_{j}^{-\kappa},k)-plates, tubes in them can have large μ2\mu_{2}-mass. Call a tube inside one of these plates non-acceptable.

  • Many of the acceptable tubes TT are good, i.e. μ2(4T)Rjα/2+ε0\mu_{2}(4T)\lesssim R_{j}^{-\alpha/2+\varepsilon_{0}}.

  • By Theorem 3.1, there are not many tubes that are neither non-acceptable nor good.

The idea is that to form our good measure μ1,gx\mu_{1,g}^{x}, we keep contributions only from good tubes. By the third bullet, we are allowed to remove tubes that are neither non-acceptable nor good. To remove the non-acceptable tubes, we will instead remove the heavy plates. Next, we formalize this idea.

Let kk be the integer such that k1<αkk-1<\alpha\leq k. Let r,k\mathcal{E}_{r,k} be the cover of B(0,1)B(0,1) with (r,k)(r,k)-plates as described in the Introduction; every (r/2,k)(r/2,k)-plate is contained within some element of r,k\mathcal{E}_{r,k}. Let j\mathcal{H}_{j} be the set of (Rjκ,k)(R_{j}^{-\kappa},k)-plates in Rjκ,k\mathcal{E}_{R_{j}^{-\kappa},k} that are RjηR_{j}^{-\eta}-concentrated on μ1+μ2\mu_{1}+\mu_{2}; then Lemma 3.3 tells us |j|RjNη|\mathcal{H}_{j}|\lesssim R_{j}^{N\eta}. Let j=i=1ji\mathcal{H}_{\leq j}=\cup_{i=1}^{j}\mathcal{H}_{i} and =i=1i\mathcal{H}=\cup_{i=1}^{\infty}\mathcal{H}_{i}. Note that |j|RjNη|\mathcal{H}_{\leq j}|\lesssim R_{j}^{N\eta}.

Let Csep1C_{\text{sep}}\geq 1 be a constant such that dist(E1,E2)Csep11\mathrm{dist}(E_{1},E_{2})\geq C_{\text{sep}}^{-1}\gtrsim 1. We will eventually choose a “master parameter” a[99Csep,100Csep]a\in[99C_{\text{sep}},100C_{\text{sep}}]. For HH\in\mathcal{H}, we will use aHaH as proxies for HH when defining acceptable tubes and the good measure.

We briefly attempt to motivate the construction. The role of CsepC_{\text{sep}} is to make sure that tubes intersecting HE1H\cap E_{1} and HE2H\cap E_{2} actually lie inside aHaH. The role of aa is to introduce a probabilistic wiggle to make a key technical condition hold (see the control of Badj2\text{Bad}_{j}^{2} in Lemma 4.8).

Now, we fix a choice of aa and define the following. We say an Rj1/2+βR_{j}^{-1/2+\beta}-tube T𝕋jT\in\mathbb{T}_{j} is non-acceptable if there exists some Hmax(j,j)H\in\mathcal{H}_{\leq\max(j,j_{*})} such that 2T2T is contained in aHaH, where j=log2R0j_{*}=\log_{2}R_{0} (the motivation for introducing jj_{*} will be explained in the proof of Lemma 4.2). Otherwise, we say it is acceptable. Define a good Rj1/2+βR_{j}^{-1/2+\beta}-tube TT to be an acceptable tube with μ2(4T)Rjα/2+ε0\mu_{2}(4T)\leq R_{j}^{-\alpha/2+\varepsilon_{0}}. And define the good part of μ1\mu_{1} with respect to μ2\mu_{2} and xE2x\in E_{2} by

μ1,gx:=M0(μ1|G0(x))+T𝕋,T goodMTμ1.\mu_{1,g}^{x}:=M_{0}(\mu_{1}|_{G_{0}(x)})+\sum_{T\in\mathbb{T},\,T\text{ good}}M_{T}\mu_{1}.

The only dependence on xx comes in the M0(μ1|G0(x))M_{0}(\mu_{1}|_{G_{0}(x)}), which is crucial when we try to prove Proposition 2.2 later. We will define G0(x)G_{0}(x) in the next subsection, roughly speaking, G0(x)G_{0}(x) is obtained by removing heavy plates through xx at several scales.

As constructed, we may not get a good bound of the form dx(μ1,gx)dx(μ1)R0ε\|d^{x}_{*}(\mu_{1,g}^{x})-d^{x}_{*}(\mu_{1})\|\leq R_{0}^{-\varepsilon}. This is because μ1,gx\mu_{1,g}^{x} doesn’t include contributions from non-acceptable tubes while μ1\mu_{1} does. Instead, we need to work with a measure μ1x\mu_{1}^{x} depending on xx that removes the contributions of the non-acceptable tubes through xx. In fact, since non-acceptable tubes are contained in heavy plates, we should define μ1x\mu_{1}^{x} by removing these heavy plates “at different scales RjR_{j}” and make sure that summing over different scales still leads to good behavior (see Lemma 4.3). We make things rigorous in the next subsection.

4.3. Construction of G(x)G(x) and μ1x\mu_{1}^{x}

Recall that dist(E1,E2)Csep11\mathrm{dist}(E_{1},E_{2})\geq C_{\text{sep}}^{-1}\gtrsim 1. Thus, for xE2x\in E_{2}, we have E1B(x,Csep1)cE_{1}\subset B(x,C_{\text{sep}}^{-1})^{c}, a fact that underlies the rest of the paper.

Recall that a[99Csep,100Csep]a\in[99C_{\text{sep}},100C_{\text{sep}}] is the “master parameter” to be chosen later. For xE2x\in E_{2} and Hj(x)H\in\mathcal{H}_{j}(x), let F(x,aH)F(x,aH) be given by

F(x,aH):={yB(x,Csep1)cB(0,10):l(x,y)Surf(aH)=},F(x,aH):=\{y\in B(x,C_{\text{sep}}^{-1})^{c}\cap B(0,10):l(x,y)\cap\mathrm{Surf}(aH)=\emptyset\}\,,

where l(x,y)l(x,y) is the line through xx and yy. It is true that F(x,aH)aHF(x,aH)\subset aH; this will be proved in Lemma 4.1. Define j:=log2R0j_{*}:=\log_{2}R_{0} such that Rj=R02R_{j_{*}}=R_{0}^{2}. For j0j\geq 0, let

(4.1) Gj(x)=[B(x,Csep1)cB(0,10)]Hmax(j,j)(x)F(x,aH).G_{j}(x)=\left[B(x,C_{\text{sep}}^{-1})^{c}\cap B(0,10)\right]\setminus\cup_{H\in\mathcal{H}_{\leq\max(j,j_{*})}(x)}F(x,aH)\,.

Finally, we define G(x)=G0(x)B(x,Csep1)G(x)=G_{0}(x)\cup B(x,C_{\text{sep}}^{-1}) and

(4.2) μ1x:=j0μ1|Gj(x)ψˇj.\mu_{1}^{x}:=\sum_{j\geq 0}\mu_{1}|_{G_{j}(x)}*\check{\psi}_{j}.

It will be proved that G(x)G(x) satisfies the condition of Proposition 2.1 in the next subsection, and this is the only reason why we include B(x,Csep1)B(x,C_{\text{sep}}^{-1}) in G(x)G(x).

We now list some good properties of these definitions that will be critical later. First, the construction of Gj(x)G_{j}(x) ensures that Gj(x)Gj+1(x)G_{j}(x)\supset G_{j+1}(x) and

Gj(x)Gj+1(x)Hj+1(x)F(x,aH),G_{j}(x)\setminus G_{j+1}(x)\subset\cup_{H\in\mathcal{H}_{j+1}(x)}F(x,aH)\,,

with Gj(x)=Gj+1(x)G_{j}(x)=G_{j+1}(x) for j<jj<j_{*}. Next, Gj(x)G_{j}(x) keeps the contributions from the acceptable tubes through xx while discarding the non-acceptable tubes through xx.

Lemma 4.1.

Let T𝕋jT\in\mathbb{T}_{j} and x2TE2x\in 2T\cap E_{2}.

(a) If 2TB(x,Csep1)cGj(x)2T\cap B(x,C_{\text{sep}}^{-1})^{c}\subset G_{j}(x), then TT is acceptable;

(b) If 2TB(x,Csep1)cF(x,aH)2T\cap B(x,C_{\text{sep}}^{-1})^{c}\subset F(x,aH) for some Hmax(j,j)(x)H\in\mathcal{H}_{\leq\max(j,j_{*})}(x), then TT is non-acceptable.

Proof.

Let T𝕋jT\in\mathbb{T}_{j} and x2Tx\in 2T.

(a) If 2TB(x,Csep1)cGj(x)2T\cap B(x,C_{\text{sep}}^{-1})^{c}\subset G_{j}(x), we show TT is acceptable, i.e. 2TaH2T\not\subset aH for any Hmax(j,j)H\in\mathcal{H}_{\leq\max(j,j_{*})}. Fix Hmax(j,j)H\in\mathcal{H}_{\leq\max(j,j_{*})}. If xaHx\notin aH, then 2TaH2T\not\subset aH. If xaHx\in aH, then Hmax(j,j)(x)H\in\mathcal{H}_{\leq\max(j,j_{*})}(x). Choose y2TB(x,Csep1)cy\in 2T\cap B(x,C_{\text{sep}}^{-1})^{c} such that the line l(x,y)l(x,y) is parallel to the central line of TT. By assumption, we know that yGj(x)y\in G_{j}(x), so l(x,y)l(x,y) intersects Surf(aH)\mathrm{Surf}(aH). In particular, 2T2T intersects Surf(aH)\mathrm{Surf}(aH), and thus 2TaH2T\not\subset aH.

(b) If 2TB(x,Csep1)cF(x,aH)2T\cap B(x,C_{\text{sep}}^{-1})^{c}\subset F(x,aH) for some H(x)H\in\mathcal{H}_{\ell}(x) with max(j,j)\ell\leq\max(j,j_{*}), we show TT is non-acceptable. More precisely, we can show that 2T2T is contained in aHaH.

As promised before, we will prove F(x,aH)aHF(x,aH)\subset aH. Take yF(x,aH)y\in F(x,aH), and suppose yaHy\notin aH. Then d(y,PH)aRκd(y,P_{H})\geq aR_{\ell}^{-\kappa} and d(x,PH)<aRκd(x,P_{H})<aR_{\ell}^{-\kappa}. By continuity, we can find zz on the line segment between xx and yy such that d(z,PH)=aRκd(z,P_{H})=aR_{\ell}^{-\kappa}. By convexity of B(0,10)B(0,10), we have zB(0,10)z\in B(0,10). Thus, we have zSurf(aH)z\in\mathrm{Surf}(aH), which contradicts yF(x,aH)y\in F(x,aH), and so in fact yaHy\in aH. This proves F(x,aH)aHF(x,aH)\subset aH.

Now observe the following geometric fact: 2T2T is a tube through xB(0,1)x\in B(0,1) whose ends lie on B(0,10)B(0,10), and so 2TB(x,Csep1)c2T\cap B(x,C_{\text{sep}}^{-1})^{c} contains both ends of 2T2T. Also, by our initial assumption, we get

2TB(x,Csep1)cF(x,aH)aH.2T\cap B(x,C_{\text{sep}}^{-1})^{c}\subset F(x,aH)\subset aH.

Therefore, since aHaH is convex, we get 2TaH2T\subset aH. ∎

To estimate dx(μ1|G(x))dx(μ1,gx)L1\|d_{*}^{x}(\mu_{1}|_{G(x)})-d_{*}^{x}(\mu_{1,g}^{x})\|_{L^{1}}, it suffices to estimate μ1|G(x)μ1xL1\|\mu_{1}|_{G(x)}-\mu_{1}^{x}\|_{L^{1}} and dx(μ1x)dx(μ1,gx)L1\|d_{*}^{x}(\mu_{1}^{x})-d_{*}^{x}(\mu_{1,g}^{x})\|_{L^{1}}. This will be the content of the next two subsections.

4.4. Pruning E2E_{2} and estimating μ1|G(x)μ1xL1\|\mu_{1}|_{G(x)}-\mu_{1}^{x}\|_{L^{1}}

If all the Gj(x)G_{j}(x)’s were the same, then μ1|G(x)=μ1|G0(x)=μ1x\mu_{1}|_{G(x)}=\mu_{1}|_{G_{0}(x)}=\mu_{1}^{x}. This isn’t quite true, but in general, we will still get a good bound on μ1|G(x)μ1xL1\|\mu_{1}|_{G(x)}-\mu_{1}^{x}\|_{L^{1}} if μ1(Gj(x)Gj+1(x))\mu_{1}(G_{j}(x)\setminus G_{j+1}(x)) is small for all jj. This weaker assumption is in fact true for “most” xE2x\in E_{2}. To see this, we use a variant of Proposition 7.3 in [22] (see also Proposition B.1 of [25]).

Lemma 4.2.

For every κ>0\kappa>0, there exists η0(α,k,κ)>0\eta_{0}(\alpha,k,\kappa)>0 such that for all η(0,η0]\eta\in(0,\eta_{0}] and R0R_{0} sufficiently large in terms of α,k,κ,Csep,η\alpha,k,\kappa,C_{\text{sep}},\eta, the following holds. Then there exists E2′′E2E_{2}^{\prime\prime}\subset E_{2} with μ2(E2E2′′)R0η\mu_{2}(E_{2}\setminus E_{2}^{\prime\prime})\leq R_{0}^{-\eta} such that the following assertions hold for xE2′′x\in E_{2}^{\prime\prime}.

  1. (a)

    The 100Csep100C_{\text{sep}}-scalings of elements in j(x)\mathcal{H}_{j}(x) are contained within some (Rjκ/2/3,k)(R_{j}^{-\kappa/2}/3,k)-plate through xx;

  2. (b)

    The 100Csep100C_{\text{sep}}-scalings of elements in j(x)\mathcal{H}_{\leq j_{*}}(x) are contained within some (R0κ/2,k)(R_{0}^{-\kappa/2},k)-plate through xx;

  3. (c)

    μ1(Gj(x)Gj+1(x))Rjη/2\mu_{1}(G_{j}(x)\setminus G_{j+1}(x))\lesssim R_{j}^{-\eta/2}, for any j0j\geq 0.

Proof.

First pick j0j\geq 0. By Lemma 3.4 and 300Csep<R0κ/2300C_{\text{sep}}<R_{0}^{\kappa/2} for sufficiently large R0R_{0}, we can find sets FjF_{j} with μ2(Fj)Rjκ(αk+1)/2|j|2\mu_{2}(F_{j})\lesssim R_{j}^{-\kappa(\alpha-k+1)/2}|\mathcal{H}_{j}|^{2} such that for xE2Fjx\in E_{2}\setminus F_{j}, the 100Csep100C_{\text{sep}}-scalings of elements in j(x)\mathcal{H}_{j}(x) are contained within some (Rjκ/2/3,k)(R_{j}^{-\kappa/2}/3,k)-plate through xx. Thus for xE2Fjx\in E_{2}\setminus F_{j}, assertion (a) is true.

By the same Lemma, we can find a set F0F_{0} with

μ2(F0)R0κ(αk+1)/2|j|2\mu_{2}(F_{0})\lesssim R_{0}^{-\kappa(\alpha-k+1)/2}|\mathcal{H}_{\leq j_{*}}|^{2}

such that for xE2F0x\in E_{2}\setminus F_{0}, the 100Csep100C_{\text{sep}}-scalings of elements in j(x)\mathcal{H}_{\leq j_{*}}(x) are contained within some (R0κ/2,k)(R_{0}^{-\kappa/2},k)-plate through xx. (Some of the plates in j\mathcal{H}_{\leq j_{*}} are too small, but we simply thicken them to (R0κ,k)(R_{0}^{-\kappa},k)-plates.) Thus for xE2F0x\in E_{2}\setminus F_{0}, assertion (b) is true.

Let E2′′=E2(F0j=jFj)E_{2}^{\prime\prime}=E_{2}\setminus(F_{0}\cup\bigcup_{j=j_{*}}^{\infty}F_{j}); then (using Rj=R02R_{j_{*}}=R_{0}^{2} and Lemma 3.3):

μ2(E2E2′′)R0κ(αk+1)/2|j|2+jjRjκ(αk+1)/2|j|2R0κ(αk+1)/2+4Nη+jjRjκ(αk+1)/2+2Nηj0Rj2ηηR02η,\mu_{2}(E_{2}\setminus E_{2}^{\prime\prime})\lesssim R_{0}^{-\kappa(\alpha-k+1)/2}|\mathcal{H}_{\leq j_{*}}|^{2}+\sum_{j\geq j_{*}}R_{j}^{-\kappa(\alpha-k+1)/2}|\mathcal{H}_{j}|^{2}\\ \lesssim R_{0}^{-\kappa(\alpha-k+1)/2+4N\eta}+\sum_{j\geq j_{*}}R_{j}^{-\kappa(\alpha-k+1)/2+2N\eta}\lesssim\sum_{j\geq 0}R_{j}^{-2\eta}\lesssim_{\eta}R_{0}^{-2\eta},

if ηη0\eta\leq\eta_{0} is chosen sufficiently small in terms of κ\kappa. Hence, if R0R_{0} is chosen sufficiently large in terms of η\eta, then R0ηR_{0}^{\eta} dominates the implicit constant and we get μ2(E2E2′′)R0η\mu_{2}(E_{2}\setminus E_{2}^{\prime\prime})\leq R_{0}^{-\eta}. Thus, E2′′E2E_{2}^{\prime\prime}\subset E_{2} satisfies the desired bound and assertions (a) and (b) are proved.

Let xE2′′x\in E_{2}^{\prime\prime}. For assertion (c), we observe that if Rj<R02R_{j}<R_{0}^{2}, then Gj(x)=Gj+1(x)G_{j}(x)=G_{j+1}(x). (This is the reason why the parameter jj_{*} was introduced.) Thus, assume RjR02R_{j}\geq R_{0}^{2}. Then

Gj(x)Gj+1(x)Hj+1(x)F(x,aH),G_{j}(x)\setminus G_{j+1}(x)\subset\bigcup_{H\in\mathcal{H}_{j+1}(x)}F(x,aH)\,,

which is contained within some (Rj+1κ/2/3,k)(R_{j+1}^{-\kappa/2}/3,k)-plate VV through xx using assertion (a). Let mm be such that Rj+1[Rm2/2,2Rm2]R_{j+1}\in[R_{m}^{2}/2,2R_{m}^{2}]. Since Rj+1κ/2/3Rmκ/2R_{j+1}^{-\kappa/2}/3\leq R_{m}^{-\kappa}/2, we know that VV is contained in some (Rmκ,k)(R_{m}^{-\kappa},k)-plate HRmκ,kH\in\mathcal{E}_{R_{m}^{-\kappa},k}. Note that, since a99Csepa\geq 99C_{\text{sep}}, we have HB(x,Csep1)cF(x,aH)H\cap B(x,C_{\text{sep}}^{-1})^{c}\subset F(x,aH). Therefore, if HmH\in\mathcal{H}_{m}, then VGm(x)=V\cap G_{m}(x)=\emptyset. Otherwise, we have μ1(V)Rmη\mu_{1}(V)\leq R_{m}^{-\eta} by definition of m\mathcal{H}_{m}. In either case, we have

μ1(Gj(x)Gj+1(x))μ1(VGm(x))RmηRjη/2,\mu_{1}(G_{j}(x)\setminus G_{j+1}(x))\leq\mu_{1}(V\cap G_{m}(x))\lesssim R_{m}^{-\eta}\lesssim R_{j}^{-\eta/2}\,,

as desired. ∎

Lemma 4.2 has two ramifications. First, Lemma 4.2(b) tells us that B(0,10)G(x)B(0,10)\setminus G(x) is contained within some (R0κ/2,k)(R_{0}^{-\kappa/2},k)-plate. Second, Lemma 4.2(c) tells us that μ1(Gj(x)Gj+1(x))\mu_{1}(G_{j}(x)\setminus G_{j+1}(x)) is small for all xE2′′x\in E_{2}^{\prime\prime}, so we can estimate μ1xμ1|G(x)L1\|\mu_{1}^{x}-\mu_{1}|_{G(x)}\|_{L^{1}}. We record these two observations and provide detailed proofs in the following lemma.

Lemma 4.3.

Let E2′′E_{2}^{\prime\prime} be the subset given in Lemma 4.2 and xE2′′x\in E_{2}^{\prime\prime}. Then B(0,10)G(x)B(0,10)\setminus G(x) is contained within some (R0κ/2,k)(R_{0}^{-\kappa/2},k)-plate and

μ1|G(x)μ1xL1R0η/2.\|\mu_{1}|_{G(x)}-\mu_{1}^{x}\|_{L^{1}}\lesssim R_{0}^{-\eta/2}.
Proof.

For the first assertion, we use the definition of G(x),G0(x)G(x),G_{0}(x) in (4.1) and the fact F(x,aH)aHF(x,aH)\subset aH proved in Lemma 4.1 to write

B(0,10)G(x)=[B(x,Csep1)cB(0,10)]G0(x)Hj(x)aH.B(0,10)\setminus G(x)=\left[B(x,C_{\text{sep}}^{-1})^{c}\cap B(0,10)\right]\setminus G_{0}(x)\subset\bigcup_{H\in\mathcal{H}_{\leq j_{*}}(x)}aH.

By Lemma 4.2(b), the rightmost expression is contained within some (R0κ/2,k)(R_{0}^{-\kappa/2},k)-plate.

For the second assertion, let G(x)=j=0Gj(x)G_{\infty}(x)=\bigcap_{j=0}^{\infty}G_{j}(x). First, by Lemma 4.2(c), we establish that for all j0j\geq 0,

(4.3) μ1(Gj(x)G(x))=i=jμ1(Gi(x)Gi+1(x))i=jRiη/2Rjη/2.\mu_{1}(G_{j}(x)\setminus G_{\infty}(x))=\sum_{i=j}^{\infty}\mu_{1}(G_{i}(x)\setminus G_{i+1}(x))\lesssim\sum_{i=j}^{\infty}R_{i}^{-\eta/2}\lesssim R_{j}^{-\eta/2}.

Also, μ1\mu_{1} is supported on B(x,Csep1)cB(x,C_{\text{sep}}^{-1})^{c}, so μ1|G(x)\mu_{1}|_{G(x)} and μ1|G0(x)\mu_{1}|_{G_{0}(x)} are the same measure. Thus, by (4.3), we get μ1|G(x)μ1|G(x)L1R0η/2\|\mu_{1}|_{G(x)}-\mu_{1}|_{G_{\infty}(x)}\|_{L^{1}}\lesssim R_{0}^{-\eta/2}, so it suffices to show that μ1|G(x)μ1xL1R0η/2\|\mu_{1}|_{G_{\infty}(x)}-\mu_{1}^{x}\|_{L^{1}}\lesssim R_{0}^{-\eta/2}. Indeed, using j0ψj=1\sum_{j\geq 0}\psi_{j}=1 and the definition (4.2) of μ1x\mu_{1}^{x}, we have

μ1xμ1|G(x)=j0(μ1|G(x)μ1|G(x))ψˇj=j0μ1|Gj(x)G(x)ψˇj,\mu_{1}^{x}-\mu_{1}|_{G_{\infty}(x)}=\sum_{j\geq 0}(\mu_{1}|_{G(x)}-\mu_{1}|_{G_{\infty}(x)})*\check{\psi}_{j}=\sum_{j\geq 0}\mu_{1}|_{G_{j}(x)\setminus G_{\infty}(x)}*\check{\psi}_{j},

and so by ψˇjL11\|\check{\psi}_{j}\|_{L^{1}}\lesssim 1, Young’s convolution inequality, and (4.3), we have

μ1xμ1|G(x)L1j0μ1(Gj(x)G(x))j0Rjη/2R0η/2.\|\mu_{1}^{x}-\mu_{1}|_{G_{\infty}(x)}\|_{L^{1}}\lesssim\sum_{j\geq 0}\mu_{1}(G_{j}(x)\setminus G_{\infty}(x))\lesssim\sum_{j\geq 0}R_{j}^{-\eta/2}\lesssim R_{0}^{-\eta/2}.

Remark 4.4.

We can make the last step more efficient using Abel summation. Let Pˇk=j=1kψˇj\check{P}_{k}=\sum_{j=1}^{k}\check{\psi}_{j}. Note that if we chose ψj\psi_{j} to be Littlewood-Paley functions χ(x/Rj)χ(x/Rj1)\chi(x/R_{j})-\chi(x/R_{j-1}) as in Section 4.1, then PˇkL1C\|\check{P}_{k}\|_{L^{1}}\leq C for some universal constant CC. Now, Abel summation gives

μ1xμ1|G(x)=j=0μ1|Gj(x)Gj+1(x)Pˇj,\mu_{1}^{x}-\mu_{1}|_{G_{\infty}(x)}=\sum_{j=0}^{\infty}\mu_{1}|_{G_{j}(x)\setminus G_{j+1}(x)}*\check{P}_{j},

and thus we get a slightly sharper bound μ1xμ1|G(x)L1μ(G0(x)G(x))\|\mu_{1}^{x}-\mu_{1}|_{G_{\infty}(x)}\|_{L^{1}}\lesssim\mu(G_{0}(x)\setminus G_{\infty}(x)). This can be useful in some potential applications where μ(G0(x)G(x))\mu(G_{0}(x)\setminus G_{\infty}(x)) is controlled but not μ(Gj(x)G(x))\mu(G_{j}(x)\setminus G_{\infty}(x)) as jj\to\infty.

4.5. Tube Geometry and bound of dx(μ1x)dx(μ1,gx)L1\|d_{*}^{x}(\mu_{1}^{x})-d_{*}^{x}(\mu_{1,g}^{x})\|_{L^{1}}

The goal of this subsection is to establish a bound on dx(μ1x)dx(μ1,gx)L1\|d_{*}^{x}(\mu_{1}^{x})-d_{*}^{x}(\mu_{1,g}^{x})\|_{L^{1}} in terms of the geometry of the tubes (Lemma 4.7). Showing such a bound will allow us to prove Proposition 2.1, thanks to Lemma 4.3.

Recall that

μ1x=j0μ1|Gj(x)ψˇj.\mu_{1}^{x}=\sum_{j\geq 0}\mu_{1}|_{G_{j}(x)}*\check{\psi}_{j}.

We first give a good approximation to μ1x\mu_{1}^{x}. The following lemma can be proved the same way as Lemma 3.4 in [10] and we omit the details. It shows that problems about the mysterious μ1x\mu_{1}^{x} actually reduce to problems about restricted measures of μ1\mu_{1}. We remark that the next few lemmas will be proved for all xE2x\in E_{2}, even though we only need these results for xE2′′x\in E_{2}^{\prime\prime} to prove Proposition 2.1. This is a minor difference: restricting to xE2′′x\in E_{2}^{\prime\prime} doesn’t make the lemmas easier to prove.

Lemma 4.5.

Let xE2x\in E_{2}. Then

μ1xM0(μ1|G0(x))j=1T𝕋jMT(μ1|Gj(x))L1RapDec(R0).\|\mu_{1}^{x}-M_{0}(\mu_{1}|_{G_{0}(x)})-\sum_{j=1}^{\infty}\sum_{T\in\mathbb{T}_{j}}M_{T}(\mu_{1}|_{G_{j}(x)})\|_{L^{1}}\leq\mathrm{RapDec}(R_{0}).

We make the following observations relating the geometry of tubes to analytic estimates.

Lemma 4.6.

Let T𝕋jT\in\mathbb{T}_{j} be an Rj1/2+βR_{j}^{-1/2+\beta}-tube and xE2x\in E_{2}.

  1. (a)

    If 2T2T doesn’t pass through xx, then dx(MTμ1)L1RapDec(Rj)\|d_{*}^{x}(M_{T}\mu_{1})\|_{L^{1}}\leq\mathrm{RapDec}(R_{j}) and dx(MT(μ1|Gj(x)))L1RapDec(Rj)\|d_{*}^{x}(M_{T}(\mu_{1}|_{G_{j}(x)}))\|_{L^{1}}\leq\mathrm{RapDec}(R_{j}).

  2. (b)

    If 2TB(x,Csep1)cGj(x)2T\cap B(x,C_{\text{sep}}^{-1})^{c}\subset G_{j}(x), then MTμ1MT(μ1|Gj(x))L1RapDec(Rj)\|M_{T}\mu_{1}-M_{T}(\mu_{1}|_{G_{j}(x)})\|_{L^{1}}\leq\mathrm{RapDec}(R_{j}).

  3. (c)

    If 2TB(x,Csep1)cF(x,aH)2T\cap B(x,C_{\text{sep}}^{-1})^{c}\subset F(x,aH) for some Hmax(j,j)(x)H\in\mathcal{H}_{\leq\max(j,j_{*})}(x), then MT(μ1|Gj(x))L1RapDec(Rj)\|M_{T}(\mu_{1}|_{G_{j}(x)})\|_{L^{1}}\leq\mathrm{RapDec}(R_{j}).

  4. (d)

    MT(μ1|Gj(x))L1\|M_{T}(\mu_{1}|_{G_{j}(x)})\|_{L^{1}} and MTμ1L1\|M_{T}\mu_{1}\|_{L^{1}} are both μ1(2T)+RapDec(Rj)\lesssim\mu_{1}(2T)+\mathrm{RapDec}(R_{j}).

Proof.

(a) This is Lemma 3.1 of [10] applied to μ1\mu_{1} and μ1|Gj(x)\mu_{1}|_{G_{j}(x)}; note that μ1|Gj(x)L1μ1L11\|\mu_{1}|_{G_{j}(x)}\|_{L^{1}}\leq\|\mu_{1}\|_{L^{1}}\leq 1.

(b) By assumption, μ1μ1|Gj(x)\mu_{1}-\mu_{1}|_{G_{j}(x)} is supported outside 2TB(x,Csep1)c2TE12T\cap B(x,C_{\text{sep}}^{-1})^{c}\supset 2T\cap E_{1}, so we can apply Lemma 3.2 of [10].

(c) By assumption, μ1|Gj(x)\mu_{1}|_{G_{j}(x)} is supported outside 2TE12T\cap E_{1}, so we can apply Lemma 3.2 of [10].

(d) This is a direct consequence of Lemma 3.2 of [10]. ∎

Using Lemma 4.6, we are able to compare μ1x\mu_{1}^{x} with μ1,gx\mu_{1,g}^{x}. We first need some definition. For xE2x\in E_{2}, j0j\geq 0, define Badj(x)\mathrm{Bad}_{j}(x) to be the union of 2T2T, where T𝕋jT\in\mathbb{T}_{j} is an Rj1/2+βR_{j}^{-1/2+\beta}-tube such that 2T2T passes through xx and either (1) 2TB(x,Csep1)c2T\cap B(x,C_{\text{sep}}^{-1})^{c} is not contained in Gj(x)G_{j}(x) or any F(x,aH)F(x,aH) for Hmax(j,j)(x)H\in\mathcal{H}_{\leq\max(j,j_{*})}(x); or (2) 2TB(x,Csep1)cGj(x)2T\cap B(x,C_{\text{sep}}^{-1})^{c}\subset G_{j}(x) and μ2(4T)>Rjα/2+ε0\mu_{2}(4T)>R_{j}^{-\alpha/2+\varepsilon_{0}}. By Lemma 4.1, (2) is morally the union of the acceptable but not good tubes through xx, while (1) is the union of the tubes through xx that are “borderline” between acceptable and non-acceptable.

One may compare the following lemma with Lemma 3.1 of [5].

Lemma 4.7.

Let xE2x\in E_{2} and Badj(x)\text{Bad}_{j}(x) be defined as above, j0\forall j\geq 0. Then,

dx(μ1x)dx(μ1,gx)L1j1Rj100βdμ1(Badj(x))+RapDec(R0).\|d_{*}^{x}(\mu_{1}^{x})-d_{*}^{x}(\mu_{1,g}^{x})\|_{L^{1}}\lesssim\sum_{j\geq 1}R_{j}^{100\beta d}\mu_{1}(\mathrm{Bad}_{j}(x))+\mathrm{RapDec}(R_{0}).
Proof.

We apply Lemma 4.5 and the definition of μ1,gx\mu_{1,g}^{x}. Note that the M0(μ1|G0(x))M_{0}(\mu_{1}|_{G_{0}(x)}) terms cancel in the resulting expression for μ1xμ1,gx\mu_{1}^{x}-\mu_{1,g}^{x} (this is the reason why μ1,gx\mu_{1,g}^{x} needs to depend on xx). Let g(T)=1g(T)=1 if TT is good and 0 otherwise. Thus, it suffices to prove for each j1j\geq 1,

T𝕋j[dx(MT(μ1|Gj(x)))dx(MTμ1)g(T)]L1\displaystyle\|\sum_{T\in\mathbb{T}_{j}}[d^{x}_{*}(M_{T}(\mu_{1}|_{G_{j}(x)}))-d^{x}_{*}(M_{T}\mu_{1})g(T)]\|_{L^{1}}
(4.4) \displaystyle\lesssim Rj100βdμ1(Badj(x))+RapDec(Rj).\displaystyle R_{j}^{100\beta d}\mu_{1}(\mathrm{Bad}_{j}(x))+\mathrm{RapDec}(R_{j}).

Let 𝕋j,bad\mathbb{T}_{j,\mathrm{bad}} be the set of tubes T𝕋jT\in\mathbb{T}_{j} such that 2T2T passes through xx and either condition (1) or (2) in the definition of Badj(x)\text{Bad}_{j}(x) holds. We claim that if T𝕋j,badT\notin\mathbb{T}_{j,\mathrm{bad}}, then

(4.5) dx(MT(μ1|Gj(x)))dx(MTμ1)g(T)L1RapDec(Rj),\|d^{x}_{*}(M_{T}(\mu_{1}|_{G_{j}(x)}))-d^{x}_{*}(M_{T}\mu_{1})g(T)\|_{L^{1}}\leq\mathrm{RapDec}(R_{j}),

while if T𝕋j,badT\in\mathbb{T}_{j,\mathrm{bad}}, then

(4.6) dx(MT(μ1|Gj(x)))dx(MTμ1)g(T)L1μ1(2T)+RapDec(Rj).\|d^{x}_{*}(M_{T}(\mu_{1}|_{G_{j}(x)}))-d^{x}_{*}(M_{T}\mu_{1})g(T)\|_{L^{1}}\lesssim\mu_{1}(2T)+\mathrm{RapDec}(R_{j}).

We show how the claim implies (4.4). For any yE1y\in E_{1}, d(x,y)1d(x,y)\gtrsim 1 and so there are Rj100βd\lesssim R_{j}^{100\beta d} many Rj1/2+βR_{j}^{-1/2+\beta}-tubes in 𝕋j\mathbb{T}_{j} passing through both xx and yy. Thus,

(4.7) T𝕋j,badμ1(2T)Rj100βdμ1(Badj(x)).\sum_{T\in\mathbb{T}_{j,\mathrm{bad}}}\mu_{1}(2T)\lesssim R_{j}^{100\beta d}\mu_{1}(\mathrm{Bad}_{j}(x)).

Combining (4.5), (4.6), (4.7) proves (4.4).

Now we prove the claim. Suppose T𝕋j,badT\notin\mathbb{T}_{j,\mathrm{bad}}. By working through the definition of 𝕋j,bad\mathbb{T}_{j,\mathrm{bad}}, we have three possibilities of TT: either

  1. (i)

    x2Tx\notin 2T;

  2. (ii)

    x2Tx\in 2T, 2TB(x,Csep1)cGj(x)2T\cap B(x,C_{\text{sep}}^{-1})^{c}\subset G_{j}(x) and μ2(4T)Rjα/2+ε0\mu_{2}(4T)\leq R_{j}^{-\alpha/2+\varepsilon_{0}}. Then TT is acceptable by Lemma 4.1(a), so it is good since μ2(4T)Rjα/2+ε0\mu_{2}(4T)\leq R_{j}^{-\alpha/2+\varepsilon_{0}}.

  3. (iii)

    x2Tx\in 2T, 2TB(x,Csep1)cF(x,aH)2T\cap B(x,C_{\text{sep}}^{-1})^{c}\subset F(x,aH) for some Hmax(j,j)(x)H\in\mathcal{H}_{\leq\max(j,j_{*})}(x). Then TT is non-acceptable by Lemma 4.1(b).

In case (i), we get (4.5) by Lemma 4.6(a), regardless of g(T)=0g(T)=0 or 11. In case (ii), we use Lemma 4.6(b) and g(T)=1g(T)=1, and in case (iii), we use Lemma 4.6(c) and g(T)=0g(T)=0. Thus, if T𝕋j,badT\notin\mathbb{T}_{j,\mathrm{bad}}, then (4.5) holds.

Now suppose T𝕋j,badT\in\mathbb{T}_{j,\mathrm{bad}}. Then we get (4.6) by applying Lemma 4.6(d), regardless of whether g(T)=0g(T)=0 or 11. This proves the claim. ∎

The crucial estimate about Badj(x)\mathrm{Bad}_{j}(x) is the following lemma, which will be proved in Section 4.6.

Lemma 4.8.

For every ε0>0\varepsilon_{0}>0, there exist η0(ε0),κ0(ε0)>0\eta_{0}(\varepsilon_{0}),\kappa_{0}(\varepsilon_{0})>0 such that for any η(0,η0],κ(0,κ0]\eta\in(0,\eta_{0}],\kappa\in(0,\kappa_{0}] and sufficiently small β\beta depending on ε0,η,κ\varepsilon_{0},\eta,\kappa, the following holds. In the construction of μ1,gx\mu_{1,g}^{x} in Section 4.2, we can choose some a[99Csep,100Csep]a\in[99C_{\text{sep}},100C_{\text{sep}}] such that for any j1j\geq 1, if we define

Badj:={(x,y)E2×E1:yBadj(x)},\mathrm{Bad}_{j}:=\{(x,y)\in E_{2}\times E_{1}:y\in\mathrm{Bad}_{j}(x)\}\,,

then μ2×μ1(Badj)Rj200βd.\mu_{2}\times\mu_{1}(\mathrm{Bad}_{j})\lesssim R_{j}^{-200\beta d}.

Using this, we complete the proof of Proposition 2.1.

Proof of Proposition 2.1.

Our goal is to find a large subset E2E2E_{2}^{\prime}\subset E_{2} with μ2(E2)1R0β\mu_{2}(E_{2}^{\prime})\geq 1-R_{0}^{-\beta} such that for each xE2x\in E_{2}^{\prime}, we have that Bd(0,10)G(x)B^{d}(0,10)\setminus G(x) is contained within some (R0β,k)(R_{0}^{-\beta},k)-plate and

(4.8) dx(μ1|G(x))dx(μ1,gx)L1R0β.\|d_{*}^{x}(\mu_{1}|_{G(x)})-d_{*}^{x}(\mu_{1,g}^{x})\|_{L^{1}}\leq R_{0}^{-\beta}.

First, let us determine the auxiliary parameters ε0,κ,η,β\varepsilon_{0},\kappa,\eta,\beta. (Recall that ε,α,k,d\varepsilon,\alpha,k,d are fixed constants.) We defer the choice of ε0(ε)\varepsilon_{0}(\varepsilon) to the next section, see Lemma 5.1. Then, choose κ=κ0(ε0)\kappa=\kappa_{0}(\varepsilon_{0}) in Lemma 4.8. Next, choose η=η(κ,ε0)\eta=\eta(\kappa,\varepsilon_{0}) to be the smaller of the two η0\eta_{0}’s in Lemma 4.2 and Lemma 4.8. Finally, choose β\beta to the smaller of the β(ε0,η,κ)\beta(\varepsilon_{0},\eta,\kappa) in Lemma 4.8 and min(η3,κ2)\min(\frac{\eta}{3},\frac{\kappa}{2}).

Now, we shall construct E2E_{2}^{\prime} by taking the set E2′′E_{2}^{\prime\prime} from Lemma 4.2 and removing some “bad parts” given by Lemma 4.8. Fix a choice of aa in the construction of μ1,gx\mu_{1,g}^{x} in Section 4.2 such that the conclusion in Lemma 4.8 holds for any j1j\geq 1. By Lemma 4.8, for each j1j\geq 1, we can find a set FjE2F_{j}\subset E_{2} with μ2(Fj)Rj50βd\mu_{2}(F_{j})\leq R_{j}^{-50\beta d} such that μ1(Badj(x))Rj150βd\mu_{1}(\mathrm{Bad}_{j}(x))\lesssim R_{j}^{-150\beta d} for xE2Fjx\in E_{2}\setminus F_{j}. Finally, define E2:=E2′′j1FjE_{2}^{\prime}:=E_{2}^{\prime\prime}\setminus\bigcup_{j\geq 1}F_{j}. We now verify that E2E_{2}^{\prime} satisfies the desired conditions.

First, observe that μ2(E2)μ(E2)μ(E2E2′′)j1Rj50βd>1R0β\mu_{2}(E_{2}^{\prime})\geq\mu(E_{2})-\mu(E_{2}\setminus E_{2}^{\prime\prime})-\sum_{j\geq 1}R_{j}^{-50\beta d}>1-R_{0}^{-\beta} if R0R_{0} is sufficiently large.

Next, fix xE2x\in E_{2}^{\prime}. Since xE2′′x\in E_{2}^{\prime\prime} and β<κ2\beta<\frac{\kappa}{2}, we get from the first part of Lemma 4.3 that B(0,10)G(x)B(0,10)\setminus G(x) is contained in some (R0β,k)(R_{0}^{-\beta},k)-plate. Now by the second part of Lemma 4.3, since βη3\beta\leq\frac{\eta}{3}, we have that

(4.9) dx(μ1|G(x))dx(μ1x)L1<12R0β.\|d_{*}^{x}(\mu_{1}|_{G(x)})-d^{x}_{*}(\mu_{1}^{x})\|_{L^{1}}<\frac{1}{2}R_{0}^{-\beta}.

For each xE2x\in E_{2}^{\prime}, Lemma 4.7 tells us (for some constant CC depending only on our parameters):

(4.10) dx(μ1x)dx(μ1,gx)L1Cj1Rj50βd+RapDec(R0)<12R0β\|d_{*}^{x}(\mu_{1}^{x})-d_{*}^{x}(\mu_{1,g}^{x})\|_{L^{1}}\leq C\cdot\sum_{j\geq 1}R_{j}^{-50\beta d}+\mathrm{RapDec}(R_{0})<\frac{1}{2}R_{0}^{-\beta}

if R0R_{0} is sufficiently large. Combining (4.9) and (4.10) via triangle inequality proves the desired equation (4.8). ∎

4.6. Control of bad part

The goal of this subsection is to prove Lemma 4.8. To do so, we will use the new radial projection estimate, Theorem 3.1.

Proof of Lemma 4.8.

Let Badj={(x,y)E2×E1:yBadj(x)}\mathrm{Bad}_{j}=\{(x,y)\in E_{2}\times E_{1}:y\in\mathrm{Bad}_{j}(x)\}. By definition of Badj(x)\mathrm{Bad}_{j}(x) above Lemma 4.7, we have BadjBadj1Badj2\mathrm{Bad}_{j}\subset\mathrm{Bad}_{j}^{1}\cup\mathrm{Bad}_{j}^{2}. Here Badj1\mathrm{Bad}_{j}^{1} is the set of pairs (x,y)E2×E1(x,y)\in E_{2}\times E_{1} such that x,yx,y lie in 2T2T for some Rj1/2+βR_{j}^{-1/2+\beta}-tube T𝕋jT\in\mathbb{T}_{j} with 2TB(x,Csep1)cGj(x)2T\cap B(x,C_{\text{sep}}^{-1})^{c}\subset G_{j}(x) and μ2(4T)Rjα/2+ε0\mu_{2}(4T)\geq R_{j}^{-\alpha/2+\varepsilon_{0}}. And Badj2\mathrm{Bad}_{j}^{2} is the set of pairs (x,y)E2×E1(x,y)\in E_{2}\times E_{1} satisfying that x,yx,y lie in 2T2T for some Rj1/2+βR_{j}^{-1/2+\beta}-tube T𝕋jT\in\mathbb{T}_{j} such that 2TB(x,Csep1)c2T\cap B(x,C_{\text{sep}}^{-1})^{c} is not contained in Gj(x)G_{j}(x) or any F(x,aH)F(x,aH) for Hmax(j,j)(x)H\in\mathcal{H}_{\leq\max(j,j_{*})}(x).

Our bound of μ2×μ1(Badj1)\mu_{2}\times\mu_{1}(\mathrm{Bad}^{1}_{j}) will not depend on the choice of aa in the construction of μ1,gx\mu_{1,g}^{x} in Section 4.2, while μ2×μ1(Badj2)\mu_{2}\times\mu_{1}(\mathrm{Bad}^{2}_{j}) will. For a given HH\in\mathcal{H}_{\ell} with max(j,j)\ell\leq\max(j,j_{*}), let Badj2(H)\mathrm{Bad}^{2}_{j}(H) be the set of pairs (x,y)E2×E1(x,y)\in E_{2}\times E_{1} satisfying that Hmax(j,j)(x)H\in\mathcal{H}_{\leq\max(j,j_{*})}(x) and x,yx,y lie in 2T2T for some Rj1/2+βR_{j}^{-1/2+\beta}-tube T𝕋jT\in\mathbb{T}_{j} such that 2TB(x,Csep1)c2T\cap B(x,C_{\text{sep}}^{-1})^{c} is not contained in dF(x,aH)\mathbb{R}^{d}\setminus F(x,aH) or F(x,aH)F(x,aH). We’ll prove

μ2×μ1(Badj1)Rj200βd\mu_{2}\times\mu_{1}(\mathrm{Bad}^{1}_{j})\lesssim R_{j}^{-200\beta d}

and that there exists aa such that

(4.11) Hmax(j,j)μ2×μ1(Badj2(H))Rj1/8,\sum_{H\in\mathcal{H}_{\leq\max(j,j_{*})}}\mu_{2}\times\mu_{1}(\mathrm{Bad}^{2}_{j}(H))\lesssim R_{j}^{-1/8}\,,

for any j0j\geq 0. Since Badj2Hmax(j,j)Badj2(H)\mathrm{Bad}_{j}^{2}\subset\bigcup_{H\in\mathcal{H}_{\leq\max(j,j_{*})}}\mathrm{Bad}_{j}^{2}(H), these two bounds will prove Lemma 4.8.

Upper bound of μ2×μ1(Badj1)\mu_{2}\times\mu_{1}(\mathrm{Bad}^{1}_{j}). This is a consequence of Theorem 3.1. By applying Theorem 3.1 with parameters (δ,r,δη,ε)=(Rj1/2+β,Rjκ/(200Csep),Rjη,ε0/4)(\delta,r,\delta^{\eta},\varepsilon)=(R_{j}^{-1/2+\beta},R_{j}^{-\kappa}/(200C_{\text{sep}}),R_{j}^{-\eta},\varepsilon_{0}/4), we can find γ>0\gamma>0 such that the following is true. There exists a set BE2×E1B\subset E_{2}\times E_{1} with μ2×μ1(B)Rjγ\mu_{2}\times\mu_{1}(B)\leq R_{j}^{-\gamma} such that for each yE1y\in E_{1} and Rj1/2+βR_{j}^{-1/2+\beta}-tube TT with 2T2T containing yy, we have (assuming κ\kappa and β\beta are chosen small enough in terms of ε0\varepsilon_{0}):

(4.12) μ2(2T(A|yB|y))CsepRj(1/2+β)αRjκ(αk+1)Rj(1/2+β)(ε0/4)Rjα/2+ε0/2,\mu_{2}(2T\setminus(A|_{y}\cup B|_{y}))\lesssim_{C_{\text{sep}}}\frac{R_{j}^{(-1/2+\beta)\alpha}}{R_{j}^{-\kappa(\alpha-k+1)}}\cdot R_{j}^{(-1/2+\beta)(-\varepsilon_{0}/4)}\leq R_{j}^{-\alpha/2+\varepsilon_{0}/2}\,,

where AA is the set of pairs (x,y)E2×E1(x,y)\in E_{2}\times E_{1} satisfying that xx and yy lie in some RjηR_{j}^{-\eta}-concentrated (Rjκ/(200Csep),k)(R_{j}^{-\kappa}/(200C_{\text{sep}}),k)-plate on μ1+μ2\mu_{1}+\mu_{2}. Since decreasing the values of β,γ\beta,\gamma makes the previous statement weaker, we may assume 200βd=γ200\beta d=\gamma.

Now, observe that since d(y,E2)Csep11d(y,E_{2})\geq C_{\text{sep}}^{-1}\gtrsim 1 for yE1y\in E_{1} and μ2\mu_{2} is a probability measure, there are at most Rjα/2ε0+O(β)\lesssim R_{j}^{\alpha/2-\varepsilon_{0}+O(\beta)} many tubes T𝕋jT\in\mathbb{T}_{j} with 2T2T containing yy satisfying μ2(4T)Rjα/2+ε0\mu_{2}(4T)\geq R_{j}^{-\alpha/2+\varepsilon_{0}}. Moreover, we claim the following.

Claim 1.. Let yE1y\in E_{1}. Suppose there exist xE2x\in E_{2} and T𝕋jT\in\mathbb{T}_{j} such that x,yx,y lie in 2T2T with 2TB(x,Csep1)cGj(x)2T\cap B(x,C_{\text{sep}}^{-1})^{c}\subset G_{j}(x). Then 2TA|y=2T\cap A|_{y}=\emptyset.

Assuming Claim 1, by (4.12) we get

μ2(Badj1|yB|y)Rjα/2ε0+O(β)Rjα/2+ε0/2Rj200βd.\mu_{2}(\mathrm{Bad}^{1}_{j}|_{y}\setminus B|_{y})\lesssim R_{j}^{\alpha/2-\varepsilon_{0}+O(\beta)}\cdot R_{j}^{-\alpha/2+\varepsilon_{0}/2}\leq R_{j}^{-200\beta d}\,.

Thus, μ2×μ1(Badj1B)Rj200βd\mu_{2}\times\mu_{1}(\mathrm{Bad}^{1}_{j}\setminus B)\lesssim R_{j}^{-200\beta d}, and so μ2×μ1(Badj1)Rj200βd\mu_{2}\times\mu_{1}(\mathrm{Bad}^{1}_{j})\lesssim R_{j}^{-200\beta d}.

It remains to prove Claim 1. Let yE1y\in E_{1}, xE2x\in E_{2}, and T𝕋jT\in\mathbb{T}_{j} be such that x,yx,y lie in 2T2T with 2TB(x,Csep1)cGj(x)2T\cap B(x,C_{\text{sep}}^{-1})^{c}\subset G_{j}(x). Suppose 2TA|y2T\cap A|_{y}\neq\emptyset. Pick a point x2TA|yx^{\prime}\in 2T\cap A|_{y}; by definition, we have xE22Tx^{\prime}\in E_{2}\cap 2T and there exists a RjηR_{j}^{-\eta}-concentrated (Rjκ/(200Csep),k)(R_{j}^{-\kappa}/(200C_{\text{sep}}),k)-plate HH^{\prime} such that xx^{\prime} and yy lie in HH^{\prime}. We also know x,yx^{\prime},y both belong to 2T2T. Since d(x,y)Csep1d(x^{\prime},y)\geq C_{\text{sep}}^{-1}, we have that 2T2T is contained in 100CsepH100C_{\text{sep}}H^{\prime}. This in turn is a RjηR_{j}^{-\eta}-concentrated (Rjκ/2,k)(R_{j}^{-\kappa}/2,k)-plate, so it must be contained in some HjH\in\mathcal{H}_{j}. Hence, 2TH2T\subset H.

By assumption, we know 2TB(x,Csep1)cF(x,aH)2T\cap B(x,C_{\text{sep}}^{-1})^{c}\not\subset F(x,aH) (where a=99Csepa=99C_{\text{sep}}), so there exists z2TB(x,Csep1)cz\in 2T\cap B(x,C_{\text{sep}}^{-1})^{c} such that (x,z)\ell(x,z) intersects Surf(aH)\mathrm{Surf}(aH) at some point ww. Let PP be the kk-plane through xx parallel to the central plane PHP_{H} of HH. Since x,w,zx,w,z are collinear, we have

d(w,P)d(z,P)=d(w,x)d(z,x).\frac{d(w,P)}{d(z,P)}=\frac{d(w,x)}{d(z,x)}.

However, we know that wSurf(aH)w\in\mathrm{Surf}(aH), so d(w,PH)=aRjκd(w,P_{H})=aR_{j}^{-\kappa}. Also d(P,PH)Rjκd(P,P_{H})\leq R_{j}^{-\kappa} since x2THx\in 2T\subset H. Thus by triangle inequality, we get d(w,P)(a1)Rjκd(w,P)\geq(a-1)R_{j}^{-\kappa}. While d(z,P)2Rjκd(z,P)\leq 2R_{j}^{-\kappa} (since z2THz\in 2T\subset H), so d(w,P)d(z,P)a12\frac{d(w,P)}{d(z,P)}\geq\frac{a-1}{2}. On the other hand, d(w,x)20d(w,x)\leq 20 and d(z,x)Csep1d(z,x)\geq C_{\text{sep}}^{-1}, so d(w,x)d(z,x)20Csep\frac{d(w,x)}{d(z,x)}\leq 20C_{\text{sep}}. Hence, we get a1220Csep\frac{a-1}{2}\leq 20C_{\text{sep}}, contradiction to a=99Csepa=99C_{\text{sep}} and Csep1C_{\text{sep}}\geq 1.

Upper bound of μ2×μ1(Badj2)\mu_{2}\times\mu_{1}(\mathrm{Bad}^{2}_{j}). We will prove there exists a choice for aa such that for all j,j,\ell satisfying max(j,j)\ell\leq\max(j,j_{*}), we have

(4.13) Fj,(a)=Hμ2×μ1(Badj2(H))Rj1/4.F_{j,\ell}(a)=\sum_{H\in\mathcal{H}_{\ell}}\mu_{2}\times\mu_{1}(\mathrm{Bad}_{j}^{2}(H))\leq R_{j}^{-1/4}.

Then given jj, summing (4.13) over max(j,j)\ell\leq\max(j,j_{*}) gives (4.11) (use max(j,j)logRj\max(j,j_{*})\lesssim\log R_{j}). To prove (4.13), we fix j,j,\ell and upper bound the measure of the set of aa’s for which (4.13) fails. We would like to apply Markov’s inequality, so we compute the expectation of Fj,(a)F_{j,\ell}(a) over aa. Let P(a)=1Csep𝟙[99Csep,100Csep]daP(a)=\frac{1}{C_{\text{sep}}}\mathbbm{1}_{[99C_{\text{sep}},100C_{\text{sep}}]}\,da be a probability measure of a[99Csep,100Csep]a\in[99C_{\text{sep}},100C_{\text{sep}}]. Then we have

IFj,(a)𝑑P(a)=\displaystyle\int_{I}F_{j,\ell}(a)\,dP(a)= E1E2IH1Badj2(H)(x,y)dP(a)dμ2(x)dμ1(y)\displaystyle\int_{E_{1}}\int_{E_{2}}\int_{I}\sum_{H\in\mathcal{H}_{\ell}}1_{\mathrm{Bad}^{2}_{j}(H)}(x,y)\,dP(a)d\mu_{2}(x)d\mu_{1}(y)\,
\displaystyle\leq supxE2,yE1IH1Badj2(H)(x,y)dP(a).\displaystyle\sup_{x\in E_{2},y\in E_{1}}\int_{I}\sum_{H\in\mathcal{H}_{\ell}}1_{\mathrm{Bad}^{2}_{j}(H)}(x,y)\,dP(a).

The following claim shows that it is unlikely that (x,y)Badj2(H)(x,y)\in\mathrm{Bad}_{j}^{2}(H) for a given HH.

Claim 2. Suppose (x,y)Badj2(H)(x,y)\in\mathrm{Bad}_{j}^{2}(H), where H(x)H\in\mathcal{H}_{\ell}(x) with max(j,j)\ell\leq\max(j,j_{*}). Let z1,z2z_{1},z_{2} be the intersections of the line through x,yx,y with B(0,10)B(0,10). Then for one of i=1,2i=1,2, we have |d(zi,PH)aRκ|Rj1/2+β|d(z_{i},P_{H})-aR_{\ell}^{-\kappa}|\lesssim R_{j}^{-1/2+\beta}, where PHP_{H} is the central plane of HH.

Proof. By definition of Badj2(H)\mathrm{Bad}_{j}^{2}(H), there exists T𝕋jT\in\mathbb{T}_{j} with 2T2T containing both x,yx,y such that 2TB(x,Csep1)c2T\cap B(x,C_{\text{sep}}^{-1})^{c} is not contained in dF(x,aH)\mathbb{R}^{d}\setminus F(x,aH) or F(x,aH)F(x,aH).

Fix a large constant C>0C>0. If d(zi,PH)aRκ>CRj1/2+βd(z_{i},P_{H})-aR_{\ell}^{-\kappa}>CR_{j}^{-1/2+\beta} for i=1i=1 or 22, then we claim 2TB(x,Csep1)cdF(x,aH)2T\cap B(x,C_{\text{sep}}^{-1})^{c}\subset\mathbb{R}^{d}\setminus F(x,aH). Note that for any y2TB(x,Csep1)cy^{\prime}\in 2T\cap B(x,C_{\text{sep}}^{-1})^{c}, one of the intersections zz^{\prime} of the line through x,yx,y^{\prime} with B(0,10)B(0,10) is contained in B(zi,CRj1/2+β)B(z_{i},CR_{j}^{-1/2+\beta}), and so by triangle inequality, we get d(z,PH)>aRκd(z^{\prime},P_{H})>aR_{\ell}^{-\kappa}. (This is why the B(x,Csep1)cB(x,C_{\text{sep}}^{-1})^{c} is important: it is not true that for all y2Ty^{\prime}\in 2T, one of the intersections zz^{\prime} of the line through x,yx,y^{\prime} with B(0,10)B(0,10) is contained in B(zi,CRj1/2+β)B(z_{i},CR_{j}^{-1/2+\beta}). Take yy^{\prime} such that (yx)(yx)(y-x)\perp(y^{\prime}-x), for instance.) This shows that 2TB(x,Csep1)cdF(x,aH)2T\cap B(x,C_{\text{sep}}^{-1})^{c}\subset\mathbb{R}^{d}\setminus F(x,aH).

If on the other hand d(zi,PH)<aRκCRj1/2+βd(z_{i},P_{H})<aR_{\ell}^{-\kappa}-CR_{j}^{-1/2+\beta} for both i=1,2i=1,2, then a similar argument shows that 2TB(x,Csep1)cF(x,aH)2T\cap B(x,C_{\text{sep}}^{-1})^{c}\subset F(x,aH). Thus, we have established the contrapositive of the claim. ∎

Using Claim 2, observe that, for a fixed pair (x,y)E2×E1(x,y)\in E_{2}\times E_{1} and a fixed H(x)H\in\mathcal{H}_{\ell}(x) with max(j,j)\ell\leq\max(j,j_{*}) (so RRj2R_{\ell}\leq R_{j}^{2}), we have (x,y)Badj2(H)(x,y)\in\mathrm{Bad}_{j}^{2}(H) for aa lying in two intervals each of length Rj1/2+βRκRj1/2+β+2κ\lesssim R_{j}^{-1/2+\beta}R_{\ell}^{\kappa}\lesssim R_{j}^{-1/2+\beta+2\kappa}. Recall that by Lemma 3.3 and RRj2R_{\ell}\leq R_{j}^{2} we have ||Rj2Nη|\mathcal{H}_{\ell}|\lesssim R_{j}^{2N\eta}. Therefore,

HI1Badj2(H)(x,y)𝑑P(a)\displaystyle\sum_{H\in\mathcal{H}_{\ell}}\int_{I}1_{\mathrm{Bad}^{2}_{j}(H)}(x,y)\,dP(a)\lesssim Rj2NηRj1/2+β+2κ|I|\displaystyle\frac{R_{j}^{2N\eta}R_{j}^{-1/2+\beta+2\kappa}}{|I|}
\displaystyle\sim Rj1/2+β+2Nη+2κ,\displaystyle R_{j}^{-1/2+\beta+2N\eta+2\kappa},

Thus, assuming β,η,κ\beta,\eta,\kappa are small enough and R0R_{0} is large enough, by Markov’s inequality we have |{a:Fj,(a)>Rj1/4}|Rj1/8|\{a:F_{j,\ell}(a)>R_{j}^{-1/4}\}|\leq R_{j}^{-1/8}. By the union bound, (4.13) fails for some j,j,\ell satisfying max(j,j)\ell\leq\max(j,j_{*}) only in a set of measure

j=0max(j,j)Rj1/8j=0j=0jR01/8+=0j=Rj1/8\displaystyle\sum_{j=0}^{\infty}\sum_{\ell\leq\max(j,j_{*})}R_{j}^{-1/8}\leq\sum_{j=0}^{j_{*}}\sum_{\ell=0}^{j_{*}}R_{0}^{-1/8}+\sum_{\ell=0}^{\infty}\sum_{j=\ell}^{\infty}R_{j}^{-1/8}
\displaystyle\lesssim R01/8(log2R0)2+0R1/8R01/8[(log2R0)2+1]<1,\displaystyle R_{0}^{-1/8}\cdot(\log_{2}R_{0})^{2}+\sum_{\ell\geq 0}R_{\ell}^{-1/8}\lesssim R_{0}^{-1/8}\cdot\left[(\log_{2}R_{0})^{2}+1\right]<1,

if R0R_{0} is chosen sufficiently large. ∎

5. Refined decoupling and Proposition 2.2

In this section, we prove Proposition 2.2, which will complete the proof of Theorem 1.2. This part of the argument proceeds very similarly as [5, Section 4] and [10, Section 5].

Let σr\sigma_{r} be the normalized surface measure on the sphere of radius rr. The main estimates in the proof of Proposition 2.2 are the following.

Lemma 5.1.

For any α>0\alpha>0, r>10R0r>10R_{0}, and ε0\varepsilon_{0} sufficiently small depending on α,ε\alpha,\varepsilon:

E2|μ1,gxσ^r(x)|2𝑑μ2(x)εrαdd+1+εr(d1)|μ^1|2ψr𝑑ξ+RapDec(r),\int_{E_{2}}|\mu_{1,g}^{x}*\widehat{\sigma}_{r}(x)|^{2}d\mu_{2}(x)\lesssim_{\varepsilon}r^{-\frac{\alpha d}{d+1}+\varepsilon}r^{-(d-1)}\int|\widehat{\mu}_{1}|^{2}\psi_{r}d\xi+{\rm RapDec}(r),

where ψr\psi_{r} is a weight function which is 1\sim 1 on the annulus r1|ξ|r+1r-1\leq|\xi|\leq r+1 and decays off of it. To be precise, we could take

ψr(ξ)=(1+|r|ξ||)100.\psi_{r}(\xi)=\left(1+|r-|\xi||\right)^{-100}.
Lemma 5.2.

For any α>0\alpha>0, r>0r>0, we have

E2|μ1,gxσ^r(x)|2𝑑μ2(x)(r+1)d1r(d1).\int_{E_{2}}|\mu_{1,g}^{x}*\widehat{\sigma}_{r}(x)|^{2}\,d\mu_{2}(x)\lesssim(r+1)^{d-1}r^{-(d-1)}\,.
Proof of Proposition 2.2, given Lemmas 5.1 and 5.2.

Note that

dx(μ1,gx)(t)=td1μ1,gxσt(x).d_{*}^{x}(\mu_{1,g}^{x})(t)=t^{d-1}\mu_{1,g}^{x}*\sigma_{t}(x).

Since μ1,gx\mu_{1,g}^{x} is essentially supported in the R01/2+βR_{0}^{-1/2+\beta} neighborhood of E1E_{1}, for xE2x\in E_{2}, we only need to consider t1t\sim 1. Hence, up to a loss of RapDec(R0)\text{RapDec}(R_{0}) which is negligible in our argument, we have

dx(μ1,gx)L22𝑑μ2(x)\displaystyle\int\|d_{*}^{x}(\mu_{1,g}^{x})\|^{2}_{L^{2}}\,d\mu_{2}(x) 0|μ1,gxσt(x)|2𝑑μ2(x)td1𝑑t\displaystyle\lesssim\int_{0}^{\infty}\int|\mu_{1,g}^{x}*\sigma_{t}(x)|^{2}\,d\mu_{2}(x)t^{d-1}dt
=0|μ1,gxσ^r(x)|2𝑑μ2(x)rd1𝑑r,\displaystyle=\int_{0}^{\infty}\int|\mu_{1,g}^{x}*\widehat{\sigma}_{r}(x)|^{2}\,d\mu_{2}(x)r^{d-1}dr,

where in the second step, we used a limiting process and an L2L^{2}-identity proved by Liu [14, Theorem 1.9]: for any Schwartz function ff on d,d2\mathbb{R}^{d},d\geq 2, and any xdx\in\mathbb{R}^{d},

0|fσt(x)|2td1𝑑t=0|fσ^r(x)|2rd1𝑑r.\int_{0}^{\infty}|f*\sigma_{t}(x)|^{2}\,t^{d-1}\,dt=\int_{0}^{\infty}|f*\widehat{\sigma}_{r}(x)|^{2}\,r^{d-1}\,dr.

For r10R0r\leq 10R_{0} we use Lemma 5.2, and for r>10R0r>10R_{0} we use Lemma 5.1. The small rr contribution to dx(μ1,gx)L22𝑑μ2(x)\int\|d_{*}^{x}(\mu_{1,g}^{x})\|^{2}_{L^{2}}d\mu_{2}(x) is

010R0(r+1)d1𝑑rR0d.\displaystyle\int_{0}^{10R_{0}}(r+1)^{d-1}\,dr\lesssim R_{0}^{d}.

The large rr contribution is (dropping the negligible RapDec(r)\mathrm{RapDec}(r) term)

10R0drαdd+1+εψr(ξ)|μ^1(ξ)|2𝑑ξ𝑑r\displaystyle\int_{10R_{0}}^{\infty}\int_{\mathbb{R}^{d}}r^{-\frac{\alpha d}{d+1}+\varepsilon}\psi_{r}(\xi)|\widehat{\mu}_{1}(\xi)|^{2}\,d\xi dr
d|ξ|αdd+1+ε|μ^1(ξ)|2𝑑ξ.\displaystyle\lesssim\int_{\mathbb{R}^{d}}|\xi|^{-\frac{\alpha d}{d+1}+\varepsilon}|\widehat{\mu}_{1}(\xi)|^{2}\,d\xi.

The proof of Proposition 2.2 is thus complete upon verification of Lemmas 5.1 and 5.2. ∎

Proof of Lemma 5.2.

We follow the proof of Proposition 5.3 in [10], case r<10R0r<10R_{0}. Since μ2\mu_{2} is a probability measure, it suffices to upper bound supx|μ1,gxσ^r(x)|\sup_{x}|\mu_{1,g}^{x}*\widehat{\sigma}_{r}(x)|. Fix xx and note that

|μ1,gxσ^r(x)|2μ1,gx^L1(dσr)2μ1,gx^L2(dσr)2.|\mu_{1,g}^{x}*\widehat{\sigma}_{r}(x)|^{2}\leq\|\widehat{\mu_{1,g}^{x}}\|^{2}_{L^{1}(d\sigma_{r})}\leq\|\widehat{\mu_{1,g}^{x}}\|^{2}_{L^{2}(d\sigma_{r})}\,.

Then by the approximately orthogonal argument in [10, Proof of Proposition 5.3], we have

μ1,gx^L2(dσr)2r(d1)(|μ1|G0(x)^|2+|μ1^|2)ψr𝑑ξ.\|\widehat{\mu_{1,g}^{x}}\|^{2}_{L^{2}(d\sigma_{r})}\lesssim r^{-(d-1)}\int(|\widehat{\mu_{1}|_{G_{0}(x)}}|^{2}+|\widehat{\mu_{1}}|^{2})\psi_{r}d\xi.

Finally, since μ^1Lμ1L1=1\|\widehat{\mu}_{1}\|_{L^{\infty}}\leq\|\mu_{1}\|_{L^{1}}=1, μ1|G0(x)^Lμ1|G0(x)L11\|\widehat{\mu_{1}|_{G_{0}(x)}}\|_{L^{\infty}}\leq\|\mu_{1}|_{G_{0}(x)}\|_{L^{1}}\leq 1, and ψr𝑑ξ(r+1)d1\int\psi_{r}\,d\xi\lesssim(r+1)^{d-1}, we get the desired result. ∎

5.1. Refined decoupling estimates

The key ingredient in the proof of Lemma 5.1 is the following refined decoupling theorem, which is derived by applying the l2l^{2} decoupling theorem of Bourgain and Demeter [3] at many different scales.

Here is the setup. Suppose that SdS\subset\mathbb{R}^{d} is a compact and strictly convex C2C^{2} hypersurface with Gaussian curvature 1\sim 1. For any ε>0\varepsilon>0, suppose there exists 0<βε0<\beta\ll\varepsilon satisfying the following. Suppose that the 1-neighborhood of RSRS is partitioned into R1/2××R1/2×1R^{1/2}\times...\times R^{1/2}\times 1 blocks θ\theta. For each θ\theta, let 𝕋θ\mathbb{T}_{\theta} be a set of finitely overlapping tubes of dimensions R1/2+β××R1/2+β×1R^{-1/2+\beta}\times\cdots\times R^{-1/2+\beta}\times 1 with long axis perpendicular to θ\theta, and let 𝕋=θ𝕋θ\mathbb{T}=\cup_{\theta}\mathbb{T}_{\theta}. Each T𝕋T\in\mathbb{T} belongs to 𝕋θ\mathbb{T}_{\theta} for a single θ\theta, and we let θ(T)\theta(T) denote this θ\theta. We say that ff is microlocalized to (T,θ(T))(T,\theta(T)) if ff is essentially supported in 2T2T and f^\widehat{f} is essentially supported in 2θ(T)2\theta(T).

Theorem 5.3.

[10, Corollary 4.3] Let pp be in the range 2p2(d+1)d12\leq p\leq\frac{2(d+1)}{d-1}. For any ε>0\varepsilon>0, suppose there exists 0<βε0<\beta\ll\varepsilon satisfying the following. Let 𝕎𝕋\mathbb{W}\subset\mathbb{T} and suppose that each T𝕎T\in\mathbb{W} lies in the unit ball. Let W=|𝕎|W=|\mathbb{W}|. Suppose that f=T𝕎fTf=\sum_{T\in\mathbb{W}}f_{T}, where fTf_{T} is microlocalized to (T,θ(T))(T,\theta(T)). Suppose that fTLp\|f_{T}\|_{L^{p}} is \sim constant for each T𝕎T\in\mathbb{W}. Let YY be a union of R1/2R^{-1/2}-cubes in the unit ball each of which intersects at most MM tubes T𝕎T\in\mathbb{W}. Then

fLp(Y)εRε(MW)121p(T𝕎fTLp2)1/2.\|f\|_{L^{p}(Y)}\lesssim_{\varepsilon}R^{\varepsilon}\left(\frac{M}{W}\right)^{\frac{1}{2}-\frac{1}{p}}\left(\sum_{T\in\mathbb{W}}\|f_{T}\|_{L^{p}}^{2}\right)^{1/2}.

The proof of Lemma 5.1 using Theorem 5.3 proceeds almost identically as in [5, Lemma 4.1], as the xx dependence of the good measure doesn’t exist in the regime r>10R0r>10R_{0}. We include the proof below for the sake of completeness.

5.2. Proof of Lemma 5.1

Assume r>10R0r>10R_{0}. By definition,

μ1,gxσ^r=j:RjrT𝕋j:T goodMTμ1σ^r+RapDec(r).\mu_{1,g}^{x}*\widehat{\sigma}_{r}=\sum_{j:R_{j}\sim r}\sum_{T\in\mathbb{T}_{j}:T\textrm{ good}}M_{T}\mu_{1}*\widehat{\sigma}_{r}+{\rm RapDec}(r).

The key point to notice is that μ1,gxσ^r\mu_{1,g}^{x}*\widehat{\sigma}_{r} is independent of xx.

The contribution of RapDec(r){\rm RapDec}(r) is already taken into account in the statement of Lemma 5.1. Hence without loss of generality we may ignore the tail RapDec(r){\rm RapDec}(r) in the argument below.

Let η1\eta_{1} be a bump function adapted to the unit ball and define

fT=η1(MTμ1σ^r).f_{T}=\eta_{1}\left(M_{T}\mu_{1}*\widehat{\sigma}_{r}\right).

One can easily verify that fTf_{T} is microlocalized to (T,θ(T))(T,\theta(T)).

Let p=2(d+1)d1p=\frac{2(d+1)}{d-1}. By dyadic pigeonholing, there exists λ>0\lambda>0 such that

|μ1,gxσ^r(x)|2𝑑μ2(x)logr|fλ(x)|2𝑑μ2(x),\int|\mu_{1,g}^{x}*\widehat{\sigma}_{r}(x)|^{2}\,d\mu_{2}(x)\lesssim\log r\int|f_{\lambda}(x)|^{2}d\mu_{2}(x),

where

fλ=T𝕎λfT,𝕎λ:=j:Rjr{T𝕋j:T good ,fTLpλ}.f_{\lambda}=\sum_{T\in\mathbb{W}_{\lambda}}f_{T},\quad\mathbb{W}_{\lambda}:=\bigcup_{j:R_{j}\sim r}\Big{\{}T\in\mathbb{T}_{j}:T\text{ good },\|f_{T}\|_{L^{p}}\sim\lambda\Big{\}}.

Next, we divide the unit ball into r1/2r^{-1/2}-cubes qq and sort them. Denote

𝒬M:={r1/2-cubes q:q intersects M tubes T𝕎λ}.\mathcal{Q}_{M}:=\{r^{-1/2}\textrm{-cubes }q:q\textrm{ intersects }\sim M\textrm{ tubes }T\in\mathbb{W}_{\lambda}\}.

Let YM:=q𝒬MqY_{M}:=\bigcup_{q\in\mathcal{Q}_{M}}q. Since there are only logr\sim\log r many choices of MM, there exists MM such that

|μ1,gxσ^r(x)|2𝑑μ2(x)(logr)2YM|fλ(x)|2𝑑μ2(x).\int|\mu_{1,g}^{x}*\widehat{\sigma}_{r}(x)|^{2}\,d\mu_{2}(x)\lesssim(\log r)^{2}\int_{Y_{M}}|f_{\lambda}(x)|^{2}d\mu_{2}(x)\,.

Since fλf_{\lambda} only involves good wave packets, by considering the quantity

q𝒬MT𝕎λ:Tqμ2(q),\sum_{q\in\mathcal{Q}_{M}}\sum_{T\in\mathbb{W}_{\lambda}:T\cap q\neq\emptyset}\mu_{2}(q),

we get

(5.1) Mμ2(𝒩r1/2(YM))|𝕎λ|rα2+ε0,M\mu_{2}(\mathcal{N}_{r^{-1/2}}(Y_{M}))\lesssim|\mathbb{W}_{\lambda}|r^{-\frac{\alpha}{2}+\varepsilon_{0}},

where 𝒩r1/2(YM)\mathcal{N}_{r^{-1/2}}(Y_{M}) is the r1/2r^{-1/2}-neighborhood of YMY_{M}.

The rest of the proof of Lemma 5.1 will follow from Theorem 5.3 and estimate (5.1).

By Hölder’s inequality and the observation that fλf_{\lambda} has Fourier support in the 11-neighborhood of the sphere of radius rr, one has

YM|fλ(x)|2𝑑μ2(x)(YM|fλ|p)2/p(YM|μ2η1/r|p/(p2))12/p,\int_{Y_{M}}|f_{\lambda}(x)|^{2}\,d\mu_{2}(x)\lesssim\left(\int_{Y_{M}}|f_{\lambda}|^{p}\right)^{2/p}\left(\int_{Y_{M}}|\mu_{2}*\eta_{1/r}|^{p/(p-2)}\right)^{1-2/p},

where η1/r\eta_{1/r} is a bump function with integral 11 that is essentially supported on the ball of radius 1/r1/r.

To bound the second factor, we note that η1/rrd\eta_{1/r}\sim r^{d} on the ball of radius 1/r1/r and rapidly decaying off it. Using the fact that μ2(B(x,t))tα,xd,t>0\mu_{2}(B(x,t))\lesssim t^{\alpha},\forall x\in\mathbb{R}^{d},\forall t>0, we have

μ2η1/rrdα.\|\mu_{2}*\eta_{1/r}\|_{\infty}\lesssim r^{d-\alpha}\,.

Therefore,

YM|μ2η1/r|p/(p2)μ2η1/r2/(p2)YM𝑑μ2η1/rr2(dα)/(p2)μ2(𝒩r1/2(YM)).\begin{split}\int_{Y_{M}}|\mu_{2}*\eta_{1/r}|^{p/(p-2)}\lesssim&\|\mu_{2}*\eta_{1/r}\|_{\infty}^{2/(p-2)}\int_{Y_{M}}d\mu_{2}*\eta_{1/r}\\ \lesssim&r^{2(d-\alpha)/(p-2)}\mu_{2}(\mathcal{N}_{r^{-1/2}}(Y_{M})).\end{split}

By Theorem 5.3, the first factor can be bounded as follows:

(YM|fλ|p)2/p(M𝕎λ)12/pT𝕎λfTLp2(rα2+ε0μ2(𝒩r1/2(YM)))12/pT𝕎λfTLp2,\begin{split}\left(\int_{Y_{M}}|f_{\lambda}|^{p}\right)^{2/p}\lessapprox&\left(\frac{M}{\mathbb{W}_{\lambda}}\right)^{1-2/p}\sum_{T\in\mathbb{W}_{\lambda}}\|f_{T}\|_{L^{p}}^{2}\\ \lesssim&\left(\frac{r^{-\frac{\alpha}{2}+\varepsilon_{0}}}{\mu_{2}(\mathcal{N}_{r^{-1/2}}(Y_{M}))}\right)^{1-2/p}\sum_{T\in\mathbb{W}_{\lambda}}\|f_{T}\|_{L^{p}}^{2},\end{split}

where the second step follows from (5.1).

Combining the two estimates together, one obtains

YM|fλ(x)|2𝑑μ2(x)r2dpα(12+1p)+O(ε0)T𝕎λfTLp2.\int_{Y_{M}}|f_{\lambda}(x)|^{2}\,d\mu_{2}(x)\lesssim r^{\frac{2d}{p}-\alpha(\frac{1}{2}+\frac{1}{p})+O(\varepsilon_{0})}\sum_{T\in\mathbb{W}_{\lambda}}\|f_{T}\|_{L^{p}}^{2}.

Observe that fTLp\|f_{T}\|_{L^{p}} has the following simple bound:

fTLpfTL|T|1/pσr(θ(T))1/2|T|1/pMTμ1^L2(dσr)=r(12p+14)(d1)+O(β)MTμ1^L2(dσr).\begin{split}\|f_{T}\|_{L^{p}}\lesssim&\|f_{T}\|_{L^{\infty}}|T|^{1/p}\lesssim\sigma_{r}(\theta(T))^{1/2}|T|^{1/p}\|\widehat{M_{T}\mu_{1}}\|_{L^{2}(d\sigma_{r})}\\ =&r^{-(\frac{1}{2p}+\frac{1}{4})(d-1)+O(\beta)}\|\widehat{M_{T}\mu_{1}}\|_{L^{2}(d\sigma_{r})}.\end{split}

Plugging this back into the above formula, one obtains

YM|fλ(x)|2𝑑μ2(x)r2dp(α+d1)(12+1p)+O(ε0)T𝕎λMTμ1^L2(dσr)2rαdd+1+εr(d1)|μ^1|2ψr𝑑ξ,\begin{split}\int_{Y_{M}}|f_{\lambda}(x)|^{2}\,d\mu_{2}(x)\lesssim&r^{\frac{2d}{p}-(\alpha+d-1)(\frac{1}{2}+\frac{1}{p})+O(\varepsilon_{0})}\sum_{T\in\mathbb{W}_{\lambda}}\|\widehat{M_{T}\mu_{1}}\|_{L^{2}(d\sigma_{r})}^{2}\\ \lesssim&r^{-\frac{\alpha d}{d+1}+\varepsilon}r^{-(d-1)}\int|\widehat{\mu}_{1}|^{2}\psi_{r}\,d\xi,\end{split}

where p=2(d+1)/(d1)p=2(d+1)/(d-1) and we have used orthogonality and chosen βε0ε\beta\ll\varepsilon_{0}\ll\varepsilon. The proof of Lemma 5.1 and hence Proposition 2.2 is complete.

6. Proof of Theorem 1.3

We will prove Theorem 1.3 following the approach in [15]. We shall use the following criteria to determine the Hausdorff dimension of pinned distance sets.

Lemma 6.1.

[15, Lemma 3.1] Given a compact set EdE\subset\mathbb{R}^{d}, xdx\in\mathbb{R}^{d} and a probability measure μE\mu_{E} on EE. Suppose there exist τ(0,1]\tau\in(0,1], K+K\in\mathbb{Z}_{+}, β>0\beta>0 such that

μE({y:|yx|Dk})<2kβ\mu_{E}(\{y:|y-x|\in D_{k}\})<2^{-k\beta}

for any

Dk=j=1MIj,D_{k}=\bigcup_{j=1}^{M}I_{j},

where k>Kk>K, M2kτM\leq 2^{k\tau} are arbitrary integers and each IjI_{j} is an arbitrary interval of length 2k\approx 2^{-k}. Then

dimH(Δx(E))τ.\dim_{H}(\Delta_{x}(E))\geq\tau.

The next proposition is a key step in the proof of Theorem 1.3, which can be viewed as a discretized variant of it.

Proposition 6.2.

Let d2,1kd1d\geq 2,1\leq k\leq d-1, k1<αkk-1<\alpha\leq k, and τ<min(f(α),1)\tau<\min(f(\alpha),1). There exists β>0\beta>0 depending on τ,α,k\tau,\alpha,k such that the following holds for sufficiently small δ<δ0(τ,α,k)\delta<\delta_{0}(\tau,\alpha,k). Let μ1,μ2\mu_{1},\mu_{2} be α\alpha-dimensional measures with 1\sim 1 separation and constant CαC_{\alpha} supported on E1,E2E_{1},E_{2} respectively. Then there exists a set FδE2F_{\delta}\subset E_{2} with μ2(Fδ)δβ2\mu_{2}(F_{\delta})\lesssim\delta^{\beta^{2}} such that for all xE2Fδx\in E_{2}\setminus F_{\delta}, there exists a set W(x)W(x) that is contained within some (2δβ2,k)(2\delta^{\beta^{2}},k)-plate such that

μ1({y:|xy|J}W(x))δβ2/2,\mu_{1}(\{y:|x-y|\in J\}\setminus W(x))\leq\delta^{\beta^{2}/2},

where JJ is any union of δτ\leq\delta^{-\tau} many intervals each of length δ\sim\delta.

Next we prove Proposition 6.2. The proof reproduces the argument in [15, Section 4] with some minor simplifications.

Proof.

First, instead of working with μ1\mu_{1}, we will use a mollified version that removes the high frequency contributions. Let ϕC0(d)\phi\in C_{0}^{\infty}(\mathbb{R}^{d}) be supported on B(0,1)B(0,1) and satisfy ϕ0\phi\geq 0, ϕ=1\int\phi=1, and ϕ1\phi\geq 1 on B(0,12)B(0,\frac{1}{2}). This ϕ\phi will be fixed for the rest of the proof (in particular, it does not depend on δ\delta, and subsequent implicit constants may depend on ϕ\phi). Let ϕδ()=δdϕ(δ1)\phi_{\delta}(\cdot)=\delta^{-d}\phi(\delta^{-1}\cdot) and μ1δ=μ1ϕδ\mu_{1}^{\delta}=\mu_{1}*\phi_{\delta}. The crucial point is that μ1δ\mu_{1}^{\delta} is supported in a δ\delta-neighborhood of the support of μ1\mu_{1} and in fact serves as a good approximation for μ1\mu_{1} down to scale δ\delta, but μ1δ\mu_{1}^{\delta} is rapidly decaying at frequencies much larger than δ1\delta^{-1}.

Fix a small ε>0\varepsilon>0. We apply Proposition 2.1 with R0=δβR_{0}=\delta^{-\beta} and the measure μ1δ\mu_{1}^{\delta}, which is still an α\alpha-dimensional measure with constant comparable to CαC_{\alpha} (independent of δ\delta). (We make δ\delta sufficiently small to ensure that R0R_{0} is sufficiently large.) Then there is a subset E2E2E_{2}^{\prime}\subset E_{2} so that μ2(E2)1δβ2\mu_{2}(E_{2}^{\prime})\geq 1-\delta^{\beta^{2}} and for each xE2x\in E_{2}^{\prime}, there exists a set G(x)dG(x)\subset\mathbb{R}^{d} where Bd(0,10)G(x)B^{d}(0,10)\setminus G(x) is contained within some (δβ2,k)(\delta^{\beta^{2}},k)-plate H(x)H(x) such that

dx(μ1δ|G(x))dx(μ1,gδ,x)L1δβ2.\|d_{*}^{x}(\mu^{\delta}_{1}|_{G(x)})-d_{*}^{x}(\mu_{1,g}^{\delta,x})\|_{L^{1}}\leq\delta^{\beta^{2}}.

We will define W(x)=H(x)(δ)W(x)=H(x)^{(\delta)}, which satisfies the condition for W(x)W(x). Let

𝒥δτ={j=1MIj:Mδτ, each Ij is an open interval of lengthδ}.\mathcal{J}^{\tau}_{\delta}=\left\{\bigcup_{j=1}^{M}I_{j}:M\leq\delta^{-\tau},\text{ each }I_{j}\text{ is an open interval of length}\sim\delta\right\}.

Let FF^{\prime} be the set of points xE2x\in E_{2}^{\prime} such that

supJ𝒥δτJ(δ)dx(μ1δ|G(x))(t)𝑑tδβ2/2.\sup_{J\in\mathcal{J}^{\tau}_{\delta}}\int_{J^{(\delta)}}d_{*}^{x}(\mu_{1}^{\delta}|_{G(x)})(t)\,dt\geq\delta^{\beta^{2}/2}.

Now, define Fδ:=F(E2E2)F_{\delta}:=F^{\prime}\cup(E_{2}\setminus E_{2}^{\prime}). Then for any xE2Fδ=E2Fx\in E_{2}\setminus F_{\delta}=E_{2}^{\prime}\setminus F^{\prime}, we can claim the following.

Claim. For all xE2Fx\in E_{2}^{\prime}\setminus F^{\prime} and J𝒥δτJ\in\mathcal{J}_{\delta}^{\tau}, we have

μ1({y:|xy|J}W(x))δβ2/2.\mu_{1}(\{y:|x-y|\in J\}\setminus W(x))\leq\delta^{\beta^{2}/2}.

Proof of Claim. Note that if yW(x)y\notin W(x), then B(y,δ)G(x)B(y,\delta)\subset G(x). For xE2Fx\in E_{2}^{\prime}\setminus F^{\prime} and J𝒥δτJ\in\mathcal{J}_{\delta}^{\tau}, we have

δβ2/2\displaystyle\delta^{\beta^{2}/2} |xz|J(δ)μ1δ|G(x)(z)dz\displaystyle\geq\int_{|x-z|\in J^{(\delta)}}\mu_{1}^{\delta}|_{G(x)}(z)\,dz
=δd|xz|J(δ),zG(x)ϕ(δ1(zy))𝑑μ1(y)𝑑z\displaystyle=\delta^{-d}\iint_{|x-z|\in J^{(\delta)},z\in G(x)}\phi(\delta^{-1}(z-y))d\mu_{1}(y)dz
δd|xy|J,|yz|δ,yW(x)ϕ(δ1(zy))𝑑μ1(y)𝑑z\displaystyle\geq\delta^{-d}\iint_{|x-y|\in J,|y-z|\leq\delta,y\notin W(x)}\phi(\delta^{-1}(z-y))d\mu_{1}(y)dz
|xy|J,yW(x)𝑑μ1(y)B(0,1)ϕ(u)𝑑u\displaystyle\geq\int_{|x-y|\in J,y\notin W(x)}d\mu_{1}(y)\int_{B(0,1)}\phi(u)\,du
=μ1({y:|xy|J}W(x)).\displaystyle=\mu_{1}(\{y:|x-y|\in J\}\setminus W(x)).\qed

Recall that μ2(E2E2)δβ2\mu_{2}(E_{2}\setminus E_{2}^{\prime})\leq\delta^{\beta^{2}}. So it remains to show μ2(F)δβ2\mu_{2}(F^{\prime})\lesssim\delta^{\beta^{2}} (assuming good choice for β,ε\beta,\varepsilon). For xFx\in F^{\prime}, we have

supJ𝒥δτJ(δ)\displaystyle\sup_{J\in\mathcal{J}^{\tau}_{\delta}}\int_{J^{(\delta)}} |dx(μ1,gδ,x)(t)|dt\displaystyle|d_{*}^{x}(\mu_{1,g}^{\delta,x})(t)|\,dt
supJ𝒥δτJ(δ)dx(μ1δ|G(x))(t)dtdx(μ1δ|G(x))dx(μ1,gδ,x)L1\displaystyle\geq\sup_{J\in\mathcal{J}^{\tau}_{\delta}}\int_{J^{(\delta)}}d_{*}^{x}(\mu_{1}^{\delta}|_{G(x)})(t)\,dt-\|d_{*}^{x}(\mu_{1}^{\delta}|_{G(x)})-d_{*}^{x}(\mu_{1,g}^{\delta,x})\|_{L^{1}}
δβ2/2δβ2δβ2.\displaystyle\geq\delta^{\beta^{2}/2}-\delta^{\beta^{2}}\geq\delta^{\beta^{2}}.

Then by Cauchy-Schwarz, we have for xFx\in F^{\prime},

(supJ𝒥δτ|Jδ|12)(|dx(μ1,gδ,x)(t)|2𝑑t)12supJ𝒥δτJ(δ)|dx(μ1,gδ,x)(t)|𝑑tδβ2.\left(\sup_{J\in\mathcal{J}_{\delta}^{\tau}}|J^{\delta}|^{\frac{1}{2}}\right)\left(\int|d_{*}^{x}(\mu_{1,g}^{\delta,x})(t)|^{2}dt\right)^{\frac{1}{2}}\geq\sup_{J\in\mathcal{J}^{\tau}_{\delta}}\int_{J^{(\delta)}}|d_{*}^{x}(\mu_{1,g}^{\delta,x})(t)|\,dt\geq\delta^{\beta^{2}}.

For J𝒥δτJ\in\mathcal{J}_{\delta}^{\tau}, JJ and J(δ)J^{(\delta)} can both be covered by δτ\lesssim\delta^{-\tau} many intervals each of length δ\sim\delta, so supJ𝒥δτ|Jδ|1/2δ(1τ)/2\sup_{J\in\mathcal{J}_{\delta}^{\tau}}|J^{\delta}|^{1/2}\lesssim\delta^{(1-\tau)/2}. Thus, for xFx\in F^{\prime},

dx(μ1,gδ,x)L22δ2β2(1τ)q.\|d_{*}^{x}(\mu_{1,g}^{\delta,x})\|^{2}_{L^{2}}\geq\delta^{2\beta^{2}-(1-\tau)}q.

Integrate over FF^{\prime} and apply Proposition 2.2 to get

δ2β2(1τ)μ2(F)\displaystyle\delta^{2\beta^{2}-(1-\tau)}\mu_{2}(F^{\prime}) dx(μ1,gδ,x)L22𝑑μ2(x)\displaystyle\leq\int\|d_{*}^{x}(\mu_{1,g}^{\delta,x})\|^{2}_{L^{2}}d\mu_{2}(x)
|μ1δ^(ξ)|2|ξ|αdd+1+ε𝑑ξ+δdβ+RapDec(δ)\displaystyle\leq\int|\widehat{\mu_{1}^{\delta}}(\xi)|^{2}|\xi|^{-\frac{\alpha d}{d+1}+\varepsilon}\,d\xi+\delta^{-d\beta}+\mathrm{RapDec}(\delta)
|μ1^(ξ)|2|ϕ^(δξ)|2|ξ|αdd+1+ε𝑑ξ+δdβ+RapDec(δ)\displaystyle\leq\int|\widehat{\mu_{1}}(\xi)|^{2}|\widehat{\phi}(\delta\xi)|^{2}|\xi|^{-\frac{\alpha d}{d+1}+\varepsilon}\,d\xi+\delta^{-d\beta}+\mathrm{RapDec}(\delta)
|ξ|δ1β|μ1^(ξ)|2|ξ|αdd+1+ε𝑑ξ+δdβ+RapDec(δ).\displaystyle\leq\int_{|\xi|\leq\delta^{-1-\beta}}|\widehat{\mu_{1}}(\xi)|^{2}|\xi|^{-\frac{\alpha d}{d+1}+\varepsilon}\,d\xi+\delta^{-d\beta}+\mathrm{RapDec}(\delta)\,.

Since τ<1\tau<1, the condition |ξ|δ1β|\xi|\leq\delta^{-1-\beta}, which implies |ξ|1βδ(1+β)(1β)δ1|\xi|^{1-\beta}\leq\delta^{-(1+\beta)(1-\beta)}\leq\delta^{-1}, gives us

δ3β2+(1τ)|ξ|(3β2(1τ))(1β)=|ξ|(1τ)+O(β).\delta^{-3\beta^{2}+(1-\tau)}\leq|\xi|^{(3\beta^{2}-(1-\tau))(1-\beta)}=|\xi|^{-(1-\tau)+O(\beta)}.

Thus,

μ2(F)\displaystyle\mu_{2}(F^{\prime})
\displaystyle\leq δ2β2+(1τ)|ξ|δ1β|μ1^(ξ)|2|ξ|αdd+1+ε𝑑ξ+δdβ2β2+(1τ)+RapDec(δ)\displaystyle\delta^{-2\beta^{2}+(1-\tau)}\int_{|\xi|\leq\delta^{-1-\beta}}|\widehat{\mu_{1}}(\xi)|^{2}|\xi|^{-\frac{\alpha d}{d+1}+\varepsilon}\,d\xi+\delta^{-d\beta-2\beta^{2}+(1-\tau)}+\mathrm{RapDec}(\delta)
\displaystyle\leq δβ2|μ1^(ξ)|2|ξ|αdd+1+ε(1τ)+O(β)𝑑ξ+δdβ2β2+(1τ)+RapDec(δ).\displaystyle\delta^{\beta^{2}}\int|\widehat{\mu_{1}}(\xi)|^{2}|\xi|^{-\frac{\alpha d}{d+1}+\varepsilon-(1-\tau)+O(\beta)}\,d\xi+\delta^{-d\beta-2\beta^{2}+(1-\tau)}+\mathrm{RapDec}(\delta)\,.

Finally, since τ<f(α)=α2d+1d+1(d1)\tau<f(\alpha)=\alpha\cdot\frac{2d+1}{d+1}-(d-1) and τ<1\tau<1, we may choose small ε\varepsilon and β\beta such that

dβ2β2+(1τ)>β2,\displaystyle-d\beta-2\beta^{2}+(1-\tau)>\beta^{2},
αdd+1+ε(1τ)+O(β)<d+α.\displaystyle-\frac{\alpha d}{d+1}+\varepsilon-(1-\tau)+O(\beta)<-d+\alpha.

This guarantees the energy integral

|μ1^(ξ)|2|ξ|αdd+1+ε(1τ)+O(β)𝑑ξ\int|\widehat{\mu_{1}}(\xi)|^{2}|\xi|^{-\frac{\alpha d}{d+1}+\varepsilon-(1-\tau)+O(\beta)}\,d\xi

to be finite (in fact, Cμ\lesssim C_{\mu}) and so μ2(F)δβ2\mu_{2}(F^{\prime})\lesssim\delta^{\beta^{2}}. ∎

We will also need a result of Shmerkin to prove Theorem 1.3.

Theorem 6.3.

[25, Theorem 6.3 and Theorem B.1] Fix 1kd11\leq k\leq d-1, c>0c>0. Given κ1,κ2>0\kappa_{1},\kappa_{2}>0, there is γ>0\gamma>0 (depending continuously on κ1,κ2\kappa_{1},\kappa_{2}) such that the following holds.

Let μ,ν\mu,\nu be probability measures on Bd(0,1)B^{d}(0,1) satisfying decay conditions

μ(V)\displaystyle\mu(V) Cμrκ1,\displaystyle\leq C_{\mu}r^{\kappa_{1}},
ν(V)\displaystyle\nu(V) Cνrκ2,\displaystyle\leq C_{\nu}r^{\kappa_{2}},

for any (r,k1)(r,k-1)-plate VV, and 0<r10<r\leq 1. Suppose ν\nu gives zero mass to every kk-dimensional affine plane. Then for all xx in a set EE of μ\mu-measure 1c\geq 1-c there is a set K(x)K(x) with ν(K(x))1c\nu(K(x))\geq 1-c such that

(6.2) ν(WK(x))rγ,\nu(W\cap K(x))\leq r^{\gamma},

for any r(0,r0]r\in(0,r_{0}] and any (r,k)(r,k)-plate WW passing through xx, where r0>0r_{0}>0 depends only on d,μ,Cν,κ2,cd,\mu,C_{\nu},\kappa_{2},c.

Finally, the set {(x,y):xE,yK(x)}\{(x,y):x\in E,y\in K(x)\} is compact.

We now prove the following theorem that implies the second claim of Theorem 1.3 (that supxEdimH(Δx(E))min(f(α),1)\sup_{x\in E}\dim_{H}(\Delta_{x}(E))\geq\min(f(\alpha),1)). (Specifically, apply Theorem 6.4 with αε\alpha-\varepsilon for any ε>0\varepsilon>0.) The first and third claims then follow by applying the second claim of Theorem 1.3 to the set {xE:dimH(Δx(E))min(f(α),1)ε}\{x\in E:\dim_{H}(\Delta_{x}(E))\leq\min(f(\alpha),1)-\varepsilon\} and taking a sequence of εn0\varepsilon_{n}\to 0.

Theorem 6.4.

Let 0<αd10<\alpha\leq d-1. Suppose E1,E2Bd(0,1)E_{1},E_{2}\subset B^{d}(0,1) are separated by 1\sim 1, and each of them has positive α\alpha-dimensional Hausdorff measure. Then there exists xE1E2x\in E_{1}\cup E_{2} with dimH(Δx(E1))min(f(α),1)\dim_{H}(\Delta_{x}(E_{1}))\geq\min(f(\alpha),1).

Proof.

Let 1kd11\leq k\leq d-1, k1<αkk-1<\alpha\leq k. Let μ1,μ2\mu_{1},\mu_{2} be α\alpha-dimensional measures supported on E1,E2E_{1},E_{2} respectively. Suppose μ1\mu_{1} gives nonzero mass to some kk-dimensional affine plane HH. We have three possible cases:

  • If k3k\geq 3, then α>k1k+12\alpha>k-1\geq\frac{k+1}{2}, so by [21], there exists xE1x\in E_{1} such that |Δx(E1)|>0|\Delta_{x}(E_{1})|>0.

  • If k=2k=2, then by [15, Theorem 1.1] we have dimH(Δx(E1))min(f(α),1)\dim_{H}(\Delta_{x}(E_{1}))\geq\min(f(\alpha),1) for some xE1x\in E_{1}.

  • If k=1k=1, then for all xE1Hx\in E_{1}\cap H, we have dimH(Δx(E1))dimH(E1H)α>f(α)=min(f(α),1)\dim_{H}(\Delta_{x}(E_{1}))\geq\dim_{H}(E_{1}\cap H)\geq\alpha>f(\alpha)=\min(f(\alpha),1).

Now assume μ1\mu_{1} gives zero mass to every kk-dimensional affine plane. Also, note that μ1\mu_{1} and μ2\mu_{2} satisfy decay conditions

μi(V)Cμirα(k1),i=1,2,\mu_{i}(V)\lesssim C_{\mu_{i}}r^{\alpha-(k-1)},\quad i=1,2,

for any (r,k1)(r,k-1)-plate VV, and 0<r10<r\leq 1. Then we can use Theorem 6.3 to find r0,γ>0r_{0},\gamma>0 and a set E2E2E_{2}^{\prime}\subset E_{2} with μ2(E2)12\mu_{2}(E_{2}^{\prime})\geq\frac{1}{2} such that for any xE2x\in E_{2}^{\prime}, there exists K(x)E1K(x)\subset E_{1} with μ1(K(x))12\mu_{1}(K(x))\geq\frac{1}{2} such that for any (r,k)(r,k)-plate HH containing xx with rr0r\leq r_{0},

μ1(K(x)H)rγ.\mu_{1}(K(x)\cap H)\leq r^{\gamma}.

Additionally, the set {(x,y):xE2,yK(x)}\{(x,y):x\in E_{2}^{\prime},y\in K(x)\} is compact, which in particular means that K(x)K(x) is compact for all xE2x\in E_{2}^{\prime}.

Fix 0<τ<min(f(α),1)0<\tau<\min(f(\alpha),1). We apply Proposition 6.2 at all sufficiently small dyadic scales δ\delta. By the Borel-Cantelli lemma, a.e. xE2x\in E_{2}^{\prime} lie in finitely many of the FδF_{\delta}. For such xx, we have for all sufficiently small δ>0\delta>0 and any JJ which is a union of δτ\leq\delta^{-\tau} many intervals each of length δ\sim\delta,

μ1({y:|xy|\displaystyle\mu_{1}(\{y:|x-y| J}K(x))\displaystyle\in J\}\cap K(x))
μ1({y:|xy|J}W(x))+μ1(K(x)W(x))\displaystyle\leq\mu_{1}(\{y:|x-y|\in J\}\setminus W(x))+\mu_{1}(K(x)\cap W(x))
δβ2/2+δγβ2.\displaystyle\lesssim\delta^{\beta^{2}/2}+\delta^{\gamma\beta^{2}}.

Hence, by Lemma 6.1 applied to the restricted measure μ1|K(x)\mu_{1}|_{K(x)}, we see that dimH(Δx(E1))τ\dim_{H}(\Delta_{x}(E_{1}))\geq\tau for a.e. xE2x\in E_{2}^{\prime}. Taking a sequence of τmin(f(α),1)\tau\to\min(f(\alpha),1), we see that dimH(Δx(E1))min(f(α),1)\dim_{H}(\Delta_{x}(E_{1}))\geq\min(f(\alpha),1) for a.e. xE2x\in E_{2}^{\prime}. ∎

7. Other norms and connections with Erdős distance problem

Theorem 1.2 also extends to more general norms. For a symmetric convex body KK in d\mathbb{R}^{d}, let K\|\cdot\|_{K} be the norm with unit ball KK.

Theorem 7.1.

Let d3d\geq 3. Let KK be a symmetric convex body in d\mathbb{R}^{d} whose boundary is CC^{\infty} smooth and has strictly positive curvature. Let EdE\subset\mathbb{R}^{d} be a compact set. Suppose that dimH(E)>d2+1418d+4\dim_{H}(E)>\frac{d}{2}+\frac{1}{4}-\frac{1}{8d+4}. Then there is a point xEx\in E such that the pinned distance set ΔK,x(E)\Delta_{K,x}(E) has positive Lebesgue measure, where

ΔK,x(E):={xyK:yE}.\Delta_{K,x}(E):=\{\|x-y\|_{K}:\,y\in E\}.

The argument for this generalization is similar to that in [10]. Indeed, the definition of bad tubes and heavy plates depends only on the geometry of d\mathbb{R}^{d} and not the specific norm involved (note that the rr-neighborhood of a set AA is still defined via Euclidean metric, not the new norm’s metric). The change of norm only affects the conversion from geometry to analysis, as manifested in Lemma 4.6(a) and our use of Liu’s L2L^{2}-identity [14, Theorem 1.9] in the proof of Proposition 2.2. These considerations were already done in [10] (see also [5]); we omit the details.

As discussed in [10, 5], one can go from Falconer-type results to Erdős-type results.

Definition 7.2.

Let PP be a set of NN points contained in [0,1]d{[0,1]}^{d}. Define the measure

(7.3) dμPs(x)=N1NdspPχB(N1s(xp))dx,d\mu^{s}_{P}(x)=N^{-1}\cdot N^{\frac{d}{s}}\cdot\sum_{p\in P}\chi_{B}(N^{\frac{1}{s}}(x-p))\,dx,

where χB\chi_{B} is the indicator function of the ball of radius 11 centered at the origin. We say that PP is ss-adaptable if there exists CC independent of NN such that

(7.4) Is(μP)=|xy|s𝑑μPs(x)𝑑μPs(y)C.I_{s}(\mu_{P})=\int\int{|x-y|}^{-s}\,d\mu^{s}_{P}(x)\,d\mu^{s}_{P}(y)\leq C.

It is not difficult to check that if the points in set PP are separated by distance cN1/scN^{-1/s}, then (7.4) is equivalent to the condition

(7.5) 1N2pp|pp|sC,\frac{1}{N^{2}}\sum_{p\not=p^{\prime}}{|p-p^{\prime}|}^{-s}\leq C,

where the exact value of CC may be different from line to line. In dimension dd, it is also easy to check that if the distance between any two points of PP is N1/d\gtrsim N^{-1/d}, then (7.5) holds for any s[0,d)s\in[0,d), and hence PP is ss-adaptable.

Using the same argument as in [5], from Theorem 7.1 we get the following Erdős-type result.

Proposition 7.3.

Let d3d\geq 3. Let KK be a symmetric convex body in d\mathbb{R}^{d} whose boundary is CC^{\infty} smooth and has strictly positive curvature. Let PP be a set of NN points contained in [0,1]d[0,1]^{d}.

(a). If the distance between any two points of PP is N1/d\gtrsim N^{-1/d}, then there exists xPx\in P such that

|ΔK,x(P)|N1d2+1418d+4=N2d+1d(d+1).\left|\Delta_{K,x}(P)\right|\gtrapprox N^{\frac{1}{\frac{d}{2}+\frac{1}{4}-\frac{1}{8d+4}}}=N^{\frac{2d+1}{d(d+1)}}.

(b). More generally, if PP is sns_{n}-adaptable for a decreasing sequence (sn)n=1(s_{n})_{n=1}^{\infty} converging to d2+1418d+4\frac{d}{2}+\frac{1}{4}-\frac{1}{8d+4}, then there exists xPx\in P such that

|ΔK,x(P)|N2d+1d(d+1).\left|\Delta_{K,x}(P)\right|\gtrapprox N^{\frac{2d+1}{d(d+1)}}.

References

  • [1] J. Bourgain. Hausdorff dimension and distance sets. Israel J. Math., 87(1-3):193–201, 1994.
  • [2] J. Bourgain. On the Erdős-Volkmann and Katz-Tao ring conjectures. Geometric and Functional Analysis, 13(2):334–365, 2003.
  • [3] J. Bourgain and C. Demeter. The proof of the l2l^{2} decoupling conjecture. Ann. of Math. (2), 182(1):351–389, 2015.
  • [4] X. Du, L. Guth, Y. Ou, H. Wang, B. Wilson, and R. Zhang. Weighted restriction estimates and application to Falconer distance set problem. Amer. J. Math., 143(1):175–211, 2021.
  • [5] 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.
  • [6] X. Du, Y. Ou, K. Ren, and R. Zhang. Weighted refined decoupling estimates and application to Falconer distance set problem. arXiv preprint, 2023.
  • [7] 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.
  • [8] M.B. Erdoğan. A bilinear Fourier extension theorem and applications to the distance set problem. Int. Math. Res. Not., 23:1411–1425, 2005.
  • [9] K.J. Falconer. On the Hausdorff dimensions of distance sets. Mathematika, 32(2):206–212, 1985.
  • [10] 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.
  • [11] Larry Guth, Nets Hawk Katz, and Joshua Zahl. On the discretized sum-product problem. International Mathematics Research Notices, 2021(13):9769–9785, 2021.
  • [12] T.L.J. Harris. Low-dimensional pinned distance sets via spherical averages. The Journal of Geometric Analysis, 31(11):11410–11416, 2021.
  • [13] 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.
  • [14] B. Liu. An l2l^{2}-identity and pinned distance problem. Geom. Funct. Anal., 29(1):283–294, 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] András Máthé and William Lewis O’Regan. Discretised sum-product theorems by shannon-type inequalities. arXiv preprint arXiv:2306.02943, 2023.
  • [17] D. Oberlin and R. Oberlin. Spherical means and pinned distance sets. Commun. Korean Math. Soc., 30(1):23–34, 2015.
  • [18] T. Orponen. On the dimension and smoothness of radial projections. Analysis & PDE, 12(5):1273–1294, 2018.
  • [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 and Pablo Shmerkin. Projections, Furstenberg sets, and the abcabc sum-product problem, 2023.
  • [21] Y. Peres and W. Schlag. Smoothness of projections, Bernoulli convolutions, and the dimension of exceptions. Duke Math. J., 104(1):193–251, 2000.
  • [22] K. Ren. Discretized radial projections in d\mathbb{R}^{d}. arXiv preprint, 2023.
  • [23] Kevin Ren and Hong Wang. Furstenberg sets estimate in the plane, 2023.
  • [24] P. Shmerkin. Improved bounds for the dimensions of planar distance sets. Journal of Fractal Geometry, 8(1):27–51, 2020.
  • [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] D.M. Stull. Pinned distance sets using effective dimension. arXiv preprint arXiv:2207.12501, 2022.
  • [28] 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.
  • [29] T. Wolff. Decay of circular means of Fourier transforms of measures. Int. Math. Res. Not., 10:547–567, 1999.

Xiumin Du, Northwestern University, [email protected]

Yumeng Ou, University of Pennsylvania, [email protected]

Kevin Ren, Princeton University, [email protected]

Ruixiang Zhang, UC Berkeley, [email protected]