Discretized Radial Projections in
Abstract
We generalize a Furstenberg-type result of Orponen-Shmerkin to higher dimensions, leading to an -improvement in Kaufman’s projection theorem for hyperplanes and an unconditional discretized radial projection theorem in the spirit of Orponen-Shmerkin-Wang. Our proof relies on a new incidence estimate for -tubes and a quasi-product set of -balls in .
1 Introduction
Let be a set in , and define the radial projection . We wish to study the size of radial projections of , where is taken in some set . Recently, Orponen, Shmerkin, and Wang [19] proved a strong radial projection theorem in two dimensions, but they prove a conditional result in higher dimensions. In this paper, we shall remove the condition in higher dimensions, which answers Conjecture 1.5 of [26] and improves Theorem 1.9 of [19]. We also improve upon the previously known result of [25, Theorem 6.15].
Theorem 1.1.
Let be Borel sets with . If is not contained in a -plane, then
In fact, we can prove the following slicing result, which improves Proposition 6.8 of [19] and makes progress towards answering Conjecture 1.10 of [19].
Corollary 1.2.
Let , then there exists such that the following holds. Let be Borel probability measures on with disjoint supports that satisfy and . Further, assume that don’t simultaneously give full measure to any affine -plane . Then there exist restrictions of to subsets of positive measure (which we keep denoting ) such that the following holds. For almost every affine 2-plane (with respect to the natural measure on the affine Grassmanian), if the sliced measures , on is non-trivial, then they don’t simultaneously give full measure to any line. In other words,
where we parametrize affine 2-planes as , for and in the Grassmannian with the rotationally invariant Haar measure .
We also deduce an -improvement in Kaufman’s projection theorem for hyperplanes. The proof is a standard higher-dimensional generalization of the details in [18, Section 3.2] and we will omit it. For , let be projection in the direction orthogonal to .
Theorem 1.3.
For every , there exists such that the following holds. Let be an analytic set in with . Then
Remark 1.4.
Kaufman’s theorem is sharp when and because can be contained within a -plane.
We also derive a higher-dimensional version of Beck’s theorem (unlike in the discrete setting, the higher-dimensional version cannot proved by projection onto a generic 2D plane). The proof again follows similarly to the 2D version presented in [19, Corollary 1.4].
Corollary 1.5.
Let be a Borel set such that for all -planes . Then, the line set spanned by pairs of distinct points in satisfies
1.1 Connections and related work
Radial projections have also been used to study the Falconer distance set problem, which asks for lower bounds on the Hausdorff dimension of the distance set given the value of for some . In two dimensions, Wolff [33] used Fourier analysis to show that if , then has positive Lebesgue measure. Using Orponen’s radial projection theorem [17], Guth-Iosevich-Ou-Wang [7] used a good-bad tube decomposition and decoupling to improve the threshold to . See also works of Keleti-Shmerkin [14] [14], Shmerkin [28], Liu [15], and Stull [29] which provide better lower bounds for given that . In higher dimensions, the works of Du-Iosevich-Ou-Wang-Zhang [3] and Wang-Zheng [32] used a good-bad tube decomposition using Orponen’s radial projection theorem and decoupling techniques [17] to provide state-of-the-art results when the dimension is even; when is odd, a more classical approach purely based on decoupling gave the best estimates [6], [9]. More recently, Shmerkin and Wang [27] prove a radial projection theorem in the spirit of this paper to provide an improved lower bound when , ; using their framework combined with updated results of [19], one can show for example that when satisfies . In fact, all of these works prove lower bounds on the size of the pinned distance set, . In the forthcoming companion papers [4], [5], we use Theorem 1.1 to improve the lower bounds for the Falconer distance set problem in all dimensions .
Very recently, radial projections in dimension have been used to prove the ABC sum-product conjecture and Furstenberg set conjecture, and yield progress on the discretized sum-product problem [21], [22]. It is natural to wonder whether the exciting progress in 2 dimensions will generalize to higher dimensions. The starting point of the breakthrough work of [21] (which was also used in [22]) is a sharp radial projection theorem in 2 dimensions, [19, Theorem 1.1]. We hope to use our higher dimensional radial projection theorem to prove analogous results to [21], [22] in all dimensions.
1.2 Discretized results
We deduce Theorem 1.1 from -discretized versions. The following notation will be used throughout this paper.
Definition 1.6.
Let be a bounded nonempty set, . Let be a dyadic number, and let and . We say that is a -set if for every -plate with , we have
If is not specified, we default to (which becomes a standard definition from [18] because -plates are -balls).
Definition 1.7.
Let be a bounded nonempty set of dyadic -tubes, . Let be a dyadic number, and let , , and . We say that is a -set of tubes if for every -plate and , we have
(1.1) |
If is not specified, we default to . We also say is a -set of tubes from scale to if the non-concentration condition (1.1) holds for .
A -set of balls cannot be concentrated in a small neighborhood of a -plane, while a -set of tubes cannot be concentrated in a small neighborhood of a -plane.
The main ingredient in the proof of Theorem 1.1 is an -improvement to the (dual) Furstenberg set problem that generalizes Theorem 1.3 in [18] to higher dimensions.
Theorem 1.8.
For any , , , , there exists such that the following holds for all small enough , depending only on . Let be a -set with , and let be a family of -tubes. Assume that for every , there exists a and -set such that for all . Then .
Remark 1.9.
The condition of being a -set is to prevent the counterexample in (say) when , and is a maximal set of many essentially distinct tubes in . This condition is automatically taken care of when : any -set is a -set with .
Remark 1.10.
We can make this decay around -plane assumption assuming that (1) is a -set and (2) for , . This will be useful for radial projection estimates, since we can guarantee (1) by Theorem B.1 of [25] and (2) because we can get rid of many pairs for a fixed .
In fact, we can prove the following refined version of Theorem 1.8.
Theorem 1.11.
For any , , , , , there exists such that the following holds for all small enough , with depending only on . Let be a -plate, be a -set with , and let be a family of -tubes. Assume that for every , there exists a set such that:
-
•
for all ;
-
•
is a -set down from scale ;
-
•
is a -set.
Then .
Remark 1.12.
(a) Given fixed , the value of can be chosen uniformly in a compact subset of . Indeed, if works for , then works in the -neighborhood of .
(b) Conjecture: can we replace the condition of being in by being a -set from scales to ?
Theorem 1.13.
Let , , and . There exist depending on , and (with ) such that the following holds. Fix , and . Let be -separated -dimensional measures with constant supported on , which lie in . Assume that . Let be the pairs of that lie in some -concentrated -plate. Then there exists a set with such that for every and -tube through , we have
The implicit constant may depend on .
Remark 1.14.
(a) It is not assumed that are a probability measures, just that .
(b) If , then the numerology of Theorem 1.13 doesn’t apply. Instead, Orponen’s radial projection theorem [17] in dimension applies. The result (stated in [7, Lemma 3.6] for , but can be generalized to all dimensions ) is that for , there exists a set with such that for every and -tube through , we have
Note that the set of “concentrated pairs” is not needed here.
(c) If , we can obtain a slightly better result by projecting to a generic -dimensional subspace and following the argument in [3, Section 3.2]. The result is that for , there exists a set with such that for every and -tube through , we have
The set is again not needed in this case. The main novelty of Theorem 1.13 comes when .
1.3 Proof ideas
The main proof ideas for Theorem 1.8 are as follows:
-
1.
Perform a standard multiscale decomposition argument due to [18] to reduce the original problem to two building blocks: the case when is a -set and when is a -regular set. The first case doesn’t happen all the time and has no loss by an elementary incidence argument, so we focus on gaining an -improvement in the second case. A -regular set has the special property that is still a -set for , .
-
2.
If is -regular with , we may find a -tube such that upon dilation of to , we obtain a new Furstenberg problem with the ball set having a quasi-product structure. See Appendix A of [18].
-
3.
Finally, we will use discretized sum-product type arguments to conclude an -improvement to the dual Furstenberg problem assuming has a quasi-product structure. In very rough terms, we shall lift to have dimension close to , and apply multi-linear Kakeya. This idea of lifting the dimension was found in He’s work on a higher-rank discretized sum-product theorem [11] in a slightly different context.
To prove Theorem 1.11, we use a similar multiscale decomposition argument as in (1) to reduce to two building blocks: a smaller version of the setting of Theorem 1.11 and a smaller version of Theorem 1.8. The smaller version of Theorem 1.11 has no loss by an elementary incidence argument, and the smaller version of Theorem 1.8 admits a gain.
1.4 Structure of the paper
In Section 2, we introduce some key concepts that will be used throughout the paper. In Sections 3 through 5, we prove Theorem 1.8 first for quasi-product sets following ideas of [10], and then for regular sets and finally for general sets following [18]. In Section 6, we prove Theorem 1.11 from Theorem 1.8. In Section 7, we generalize a radial projection theorem of Shmerkin [25, Theorem 6.3] that enables us to assume our sets have power decay around -planes. In Section 8, we prove Theorem 8.1 following ideas from [19]. Finally, in Section 9, we prove Theorem 1.1 and 1.2 from the discretized results.
Acknowledgments. The author is supported by a NSF GRFP fellowship. The author would like to thank Xiumin Du, Tuomas Orponen, Yumeng Ou, Pablo Shmerkin, Hong Wang, and Ruixiang Zhang for helpful discussions. We thank Paige Bright and Yuqiu Fu for suggesting to include a higher-dimensional version of Beck’s theorem in this paper.
2 Preliminaries
This section will summarize the argument of [18], and in lieu of proofs (with the exception of Proposition 2.9), we either refer the reader to [18] or defer the proof to a later section.
2.1 Definitions
We use to denote for some constant . We use to indicate the constant can depend on . We will also use in future proofs; its exact meaning will always be clarified when used.
For a finite set , let denote the cardinality of . If is infinite, let denote the Lebesgue measure of .
For a set , let .
For a tube , let denote the central line segment of .
For a set , let be the -neighborhood of .
For and , define the slice and .
For a measure and a set , define the restricted measure by . The renormalized restricted measure is .
For vectors , , the quantity is the non-negative volume of the parallelepiped spanned by through .
is the ball in of radius centered at . We also use the notation for an arbitrary -ball in .
For sets and , let .
Definition 2.1.
We say supported in is an -dimensional measure with constant if for all and balls of radius .
2.2 Plates
We work in . An -plate is the -neighborhood of a -dimensional hyperplane in . We construct a set of -plates with the following properties:
-
•
Each -plate intersecting lies in at least one plate of ;
-
•
For , every -plate contains many -plates of .
For example, when and , we can simply pick many -tubes in each of an -net of directions. This generalizes to higher and via a standard -net argument, but we haven’t seen it in the literature, so we provide a precise construction.
An -net of a metric space is a subset such that for . The affine Grassmanian manifold is the set of all -planes in . By counting degrees of freedom, we see that . Any such plane is uniquely for some -dimensional subspace and . For and , define their distance to be (following Section 3.16 of [16]):
where and are orthogonal projections, and is the usual operator norm for linear maps. Let be the submanifold of -planes with . Since the manifold is compact and smooth, it can be covered by finitely many charts that are -bilipschitz to a subset of .
From a maximal -net of the set of affine planes of with a sufficiently small constant, we can construct a set of -plates whose central planes are the elements of . We now check the two properties for .
To prove the first property, let be a -plate intersecting . Then the central plane must lie at distance from some element of (otherwise, we can add it to the net). Let and . Hence, and , so for (so ),
Now, note that . It is close to if :
We have proved and thus . Hence, is contained in the -plate with central plane .
To prove the second property, we note that the set of -planes in whose intersection with is contained in a given -plate is contained in an -ball of . First suppose is contained within some coordinate chart; we would like to prove that . To show this, note that is a packing of with finitely overlapping -balls. Now map the chart to . Since the map only distorts distances by a constant factor, we can pack many finitely overlapping -balls into a ball of radius . Thus by a volume argument, we have . Since there are finitely many charts, we can apply the argument to intersecting each chart, which proves the second property.
We specialize our discussion to tubes. For each scale , let be a cover of with -tubes such that every -tube (and in particular every -tube with ) is contained in at least and at most many tubes of . Slightly abusing notation (á la [18]), we will also use to represent sets of tubes, where the subscript helpfully indicates a set of -tubes.
In Theorem 1.13, we pay attention to certain plates with disproportionately much mass.
Definition 2.2.
We say that a -plate is -concentrated on if .
Other notation is following [18]. Unlike [18], we work with ordinary rather than dyadic tubes. The advantage of dyadic tubes is that every -tube is in a unique -tube if ; thus, dyadic tubes will avoid the loss incurred by the finitely overlapping cover . However, dyadic tubes have the disadvantage that they don’t behave well under rotations or dilations, and it would be more cumbersome to define -sets of dyadic tubes (whereas the definition for ordinary tubes is more geometric). Thus, in principle it is possible to work with dyadic tubes and save on the loss, but it doesn’t affect our numerology in the end (since our losses will depend badly on anyway), so we chose to work with ordinary tubes throughout.
Definition 2.3.
[18] Let be a bounded nonempty set, . Let be a dyadic number, and let and . We say that is a -set if
Definition 2.4.
Let be a bounded nonempty set of dyadic -tubes, . Let be a dyadic number, and let , , and . We say that is a -set of tubes if for every -plate and , we have
If is not specified, we default to .
The following is a simpler interpretation of -set if the tubes all pass through the same point.
Definition 2.5.
Let be the slope of the central axis of .
Lemma 2.6.
Let be a set of -tubes intersecting . Then if is a -set, then is a -set. Conversely, if is a -set, then is a -set.
Proof.
Let denote spherical projection through . Then is well-defined and equals , up to an additive loss of . Fix a -plate . Then the set of tubes with slope in and passing through must lie in a -plate . Conversely, for any -plate containing , the set of possible slopes of tubes through contained in is contained in a -plate . ∎
We will need the following lemma from [18].
Lemma 2.7 ([18], Lemma 2.7).
Let be a -set. Then contains a -separated -subset with .
First, since -sets are -sets for any , we can assume that . Next, since -sets are -sets for , we may assume . In particular, we get , a useful assumption.
We record a useful geometric fact about -plates.
Lemma 2.8.
Fix , then there exists depending on such that the following is true for . If lie in an -plate and , then any -tube through will lie in , which is a -plate.
Proof.
For sufficiently large: If does not lie in , then will be contained in a -tube segment of . ∎
2.3 An Elementary Estimate
We prove a classical estimate which can be viewed as Theorem 1.8 with . We won’t need the fact that is a -set. The case is proven as Proposition 2.13 and Corollary 2.14 of [18]. For higher dimensions, the proof is similar and we sketch the details. Let denote the inequality
Proposition 2.9.
Let , and let . Let be a -set. Assume that for every there exists a -family of dyadic -tubes with the property that for all , and for some .
Let be arbitrary, and define . Then
where . (If , then .)
The following corollary of Proposition 2.9 is the form we will use.
Corollary 2.10.
Let , and let . Let be a -set. Assume that for every there exists a -family of dyadic -tubes with the property that for all , and for some . If , then
(If , then .)
Remark 2.11.
Proof.
We begin with an application of Cauchy-Schwarz.
Note that we have the following bounds:
(2.2) |
where stands for the distance of the midpoints of and . To prove (2.2), observe that if , then lies in a -tube with central line being the line between and . Thus, the first bound in (2.2) follows from being a -set with , and the second bound is the maximum number of essentially distinct -tubes that can fit inside a -tube.
2.4 Multiscale analysis
Following Section 4 of [18], we would like to change scale from to , while preserving the properties of . We say if there exists an absolute constant such that . We start by naming the objects in Theorem 1.8.
Definition 2.12.
Fix . We say that a pair is a -nice configuration if for every , there exists a and -set with and such that for all .
Using the method of induction on scales, we would like to relate nice configurations at scale to nice configurations at scales , where . The following proposition, which combines Propositions 4.1 and 5.2 of [18], gives a way of doing so with only polylog losses. Our proof relies on the same ideas as [18], with some technical simplifications. We defer the proof to Section 6, where we prove a slightly more general version.
Proposition 2.13.
Fix dyadic numbers . Let be a -nice configuration. Then there exist sets , , and such that denoting the following hold:
-
(i)
and for all .
-
(ii)
There exists such that for all .
-
(iii)
is -nice for some , , and .
-
(iv)
For each , let be the tubes in through . Then for all , we have
-
(v)
For each , there exist , , , a subset with , and a family of tubes such that is -nice.
Furthermore, the families can be chosen so that
(2.3) |
Iterate this proposition to get (for details, see [27, Corollary 4.1])
Corollary 2.14.
Fix and a sequence with
Let be a -nice configuration. Then there exists a set such that:
-
1.
and , , .
-
2.
For every and , there exist numbers , , and , and a family of tubes with the property that is a -nice configuration.
Furthermore, the families can be chosen such that if for , then
Here, means , and likewise for .
2.5 Uniform sets and branching numbers
The following exposition borrows heavily from [21, Section 2.3].
Definition 2.15.
Let and
be a sequence of dyadic scales. We say that a set is -uniform if there is a sequence such that and for all and .
Remark 2.16.
By uniformity, we have for and .
As a result, we can always refine a set to be uniform:
Lemma 2.17.
Let , , and . Let for , so in particular . Then there is a -uniform set such that
In particular, if and , then .
Uniform sets can be encoded by a branching function.
Definition 2.18.
Let , and let be a -uniform set, with , and with associated sequence . We define the branching function by setting , and
and then interpolating linearly between integers.
Definition 2.19.
Let denote the slope of a line segment between and . We say that a function is -superlinear on , or that is -superlinear, if
We say that is -linear if
The following lemma converts between branching functions and the uniform structure of . It is [18, Lemma 8.3] (or an immediate consequence of the definitions)
Lemma 2.20.
Let be a -uniform set in with associated branching function , and let .
-
(i)
If is -superlinear on , then is a -set.
-
(ii)
If is -linear on , then is a -regular set between scales and .
The crucial branching lemma is [18, Lemma 8.5] applied to the function :
Lemma 2.21.
Fix and . For every there is such that the following holds: for every piecewise affine -Lipschitz function such that
there exists a family of non-overlapping intervals contained in such that:
-
1.
For each , at least one of the following alternatives holds:
-
(a)
is -linear with ;
-
(b)
is -superlinear with .
-
(a)
-
2.
for all ;
-
3.
.
2.6 Combinatorial and probabilistic preliminaries
In this section, we collect a few of the results from additive combinatorics and probability that will be used in the following sections.
First, we make the following observation (Lemma 19 of [10]) about intersections of high-probability events. (That lemma was stated for Lebesgue measure but the same proof works for general measures .)
Lemma 2.22.
Let equipped with a measure and be an index set equipped with a probability measure . Suppose there is and for each , a Borel subset with . Then
Lemma 2.23.
For any sets , we have
We also would like the Plünnecke-Rusza inequality, in the form stated by [10, Lemma 22]:
Lemma 2.24.
Let be bounded subsets of . For all , , if , then for all , we have
Here, .
In a similar spirit, the set of such that is small compared to forms a ring. The following is a restatement of [11, Lemma 30(i,ii)] for . Note that with identity .
Lemma 2.25.
Define .
-
(i)
If and such that , then .
-
(ii)
If , then all belong to .
The following theorem (a special case of Theorem 5 of [11]) is a quantitative statement that -dimensional subrings of don’t exist. In fact, by repeated sum-product operations, we can get all of .
Theorem 2.26.
We work in . Given , there exist and an integer such that for , the following holds. For every -set , we have
where and for any integer , define .
Finally, we shall need a discretized variant of the Balog-Szemerédi-Gowers theorem. Our version is closest to [20, Theorem 4.38], which is taken from [1, p. 196], which in turn refers to Exercise 6.4.10 in [31]. But the exercise is only sketched in [31], so for completeness, we provide a proof in Appendix A.
Theorem 2.27.
Let and be parameters. Let be bounded subsets of , and let satisfy
Then one can find subsets satisfying
-
•
,
-
•
,
-
•
.
(Implicit constants depend on but not on .)
We also need the following version of multi-linear Kakeya.
Theorem 2.28 (Theorem 1 in [2]).
Let and be families of -tubes in . Then
Here, is the unit vector in the direction of tube .
2.7 Energy
Definition 2.29.
The -Riesz energy of a finite Borel measure on is
If and , we recover the usual -dimensional Riesz energy.
Lemma 2.30.
-
(a)
Fix and a measure with total mass . If for every -plate and , then .
-
(b)
Fix . If for , then contains a set which is simultaneously a -set for each .
Remark 2.31.
If in part (b), then we can drop the log factor (c.f. proof of Lemma A.6 in [18]). We don’t know if we can drop the log factor for or .
Proof.
(a) Let be the distance between and the plane spanned by ; notice that . Thus, we can rewrite as an iterated integral
We will be done if we show for all and choices of , that . Let be the span of through , and observe that by definition, , which is contained in a -plate. Thus,
(b) Let . By Markov’s inequality, , so by the union bound, satisfies .
We claim that for all -plates and , . Indeed, if , then we are done. Otherwise, pick and observe that if , then . Thus, we get , so .
Finally, let be those dyadic -cubes such that . We know , so by dyadic pigeonholing, some . Then will be a -set for all . ∎
3 Improved incidence estimates for quasi-product sets
The main novelty of this paper is the following Proposition, which is a higher-dimensional refinement of [20, Proposition 4.36] (see also [18, Proposition A.7]). It can be viewed as a variant of Theorem 1.8 for quasi-product sets.
Proposition 3.1.
Given , , , there exist and such that the following holds for all .
Let be a -set, and for each , assume that is a -set with cardinality . Let
For every , assume that is a set of -tubes each making an angle with the plane with such that for all . Then , where .
Remark 3.2.
In contrast to Theorem 1.8 and [20, Proposition 4.36], we (perhaps surprisingly) don’t need any non-concentration assumptions on the tube sets (even when ). Instead, it suffices to have weak non-concentration assumptions on and for each . The non-concentration assumption on is necessary: otherwise, we can take , and let to be the -balls contained in some -plate , and to be the -tubes contained in .
3.1 An improved slicing estimate
We will eventually deduce Proposition 3.1 from the following slicing estimate.
Theorem 3.3.
For , , and , there exists such that the following holds for sufficiently small . Let be a -set of -tubes each making angle with the plane with . Let be a probability measure on such that for all , we have . Then there is a set with such that the slice of at has -covering number , for every subset with and .
Remark 3.4.
One should compare Theorem 3.3 to [10, Theorem 1]. Indeed, if and , Theorem 3.3 is a direct corollary of [10, Theorem 1]. We can see this by using ball-tube duality, which turns into a subset of . Under this duality, the slice of at becomes the orthogonal projection to a line in the dual space, for some . The map induces a pushforward measure of which still satisfies the non-concentration condition , so we can apply [10, Theorem 1]. (For more details, see the proof of Proposition A.7 in [18].)
In higher dimensions, we can still use duality to turn into a subset of , and then slices of become orthogonal projections to -planes. Unfortunately, [10, Theorem 1] does not apply because the pushforward measure is still supported on a line in . This approach is bound to fail because [10, Theorem 1] does not use the strong assumption that is non-concentrated around -planes. Using this assumption is the key novelty of this proof.
Nonetheless, Theorem 3.3 will borrow many ideas from the proof of [10, Theorem 1] and He’s previous work [11]. Roughly, the strategy is as follows.
- •
-
•
Then, as in [10], reduce this slightly weaker to the following even weaker statement: there exists such that the slice of at has -covering number . This relies on additive combinatorics (e.g. the Balog-Szemerédi-Gowers theorem) and some probability.
-
•
Assume this is false: that for all , the slice of at has -covering number . Using additive combinatorics as in [11], the same conclusion is true for all , which is the set of sums or differences of many terms, each of which is a product of elements of . (Here, will be a fixed large integer.)
-
•
Finally, if is sufficiently large in terms of , then contains a large interval (c.f. [11, Theorem 5]). Essentially, we have a set of many tubes , each containing many -balls, such that the union of the -balls has cardinality . Without further restrictions, this Furstenberg-type problem doesn’t lead to a contradiction: take and to be the set of -tubes in a -plate. Luckily, our set of tubes is still a -set, which rules out this counterexample. Indeed, we may finish using multi-linear Kakeya.
The reader be warned: we shall execute this strategy in reverse order. This is mainly because the main innovation of the paper is the fourth bullet point.
3.2 An improved Furstenberg estimate
The following estimate complements work of Zahl [34]: we prove an -improvement on the union of tubes under a mild -plane non-concentration for the set of tubes. As in Zahl [34], the key technique is multilinear Kakeya.
Theorem 3.5.
For any , , , there exists such that the following holds for sufficiently small . Let be a -set of -tubes with , and for each , let be a set of -balls intersecting such that . Then .
Proof.
The proof below is lossy and can possibly be improved (say by induction on scale). Also, the can be determined explicitly in terms of the parameters but we choose not to do so here.
We use notation to hide terms, where can depend on the other parameters. Let , and suppose . Let be the set of tubes in through . Use a bush argument to upper bound :
Thus, for all . We get the following inequality chain
This means , , and . Now perform a dyadic pigeonholing to extract a subset such that for all and . We know from before that , and , so and . (This type of dyadic pigeonholing will also be used later. We also remark that dyadic pigeonholing was not necessary to achieve this step; simply let be the set of satisfying for some large , and use the bound on to get a lower bound for .)
Now, we claim that is a -set. Fix and let be a -plate. We first bound . Letting be the -plate that is a dilate of with the same center, we have
Thus, since , we have .
Finally, since and , by dyadic pigeonholing there exists a subset with such that each contains many -balls in . Now since , , and for all , by dyadic pigeonholing we can find with such that each lies in many tubes in .
Now, we are in good shape to apply multilinear Kakeya. For , let be the tubes in through . By a bush argument, contains many -balls in . Since is a -set, there are many -tuples of points such that and lie on some tube and (where is the unit vector in the direction of tube ). Thus, there is a choice of such that there are many valid choices for . But this leads to a contradiction by the following argument. Let be the tubes of through , ; then by a rescaled version of Multilinear Kakeya (Theorem 2.28), the number of valid choices for is , which (using ) is much smaller than provided that are sufficiently small in terms of the parameters. This contradiction completes the proof. ∎
3.3 From Furstenberg to weak slicing
Theorem 3.6.
For , , and , there exists such that the following holds for sufficiently small . Let be a -set of -tubes each making angle with the plane with . Let be a probability measure on such that for all , we have . Then there exists such that the slice of at has -covering number .
Proof.
We use to denote , where may depend on .
Let ; without loss of generality, assume is closed. Let and ; then since .
Let and ; we are given that . On the other hand, since each passes through many elements in and many elements in , we get that
so in fact, and .
Let ; then for small enough.
Let ; note that on , we have that is -bilipschitz, and for all .
The problem condition literally states for ; since , we can divide through by to get
Now pick an arbitrary . In particular, we get , so by Lemma 2.23, we have for all ,
In addition, since , the Plünnecke-Rusza inequality (Lemma 2.24) gives .
Define , the pushforward of ; then (like ) is -bilipschitz on , so also satisfies a non-concentration condition for . Now pick , and assume is chosen sufficiently small in terms of . By the iterated sum-product Theorem 2.26, we can find an integer such that for ,
where and for any integer , define .
By applying the ring structure Lemma 2.25 many times, we see that and since , that . By definition of and Lemma 2.23, we get for ,
In other words, for all , the slice has -covering number . Since is -bilipschitz and , we have .
Now, we seek a contradiction to Theorem 3.5. For every , we let be the -balls on with -coordinate in . We observe the following:
-
•
Recall our assumption that is a -set of -tubes with .
-
•
.
-
•
.
Thus, if are sufficiently small in terms of , then we contradict Theorem 3.5.
∎
3.4 An intermediate slicing result
This subsection contains ideas from [10].
Let be the set of exceptional slices,
Just like in [10, Proposition 25], we will prove a weaker version of Theorem 3.3; the stronger version follows from a formal exhaustion argument which we present in the next subsection.
Theorem 3.7.
With assumptions of Theorem 3.3, there exists such that .
Proof.
We use to denote , where may depend on . Let be the projection onto the plane orthogonal to the -axis. For a tube , let , and for a set of tubes , let denote the slice .
We follow the argument in [10, Proof of Proposition 7]. Suppose Theorem 3.7 is false. We can find and a subset with such that . For this we have , hence by the non-concentration property of . Thus, we can find with and with such that and . Since every passes through a point in and a point in , and since there are many tubes through given points and , we can find with such that for every , there is at most one tube in through . In particular,
and so , .
For this we have , so defining , we have .
Claim 1. For , we have . Furthermore, there exists , , and with such that and for each , we have and .
Proof. The first claim is evident by definition of . For the second claim, since , there exists such that and . Now notice that for each there is at most one tube passing through . Let be the set of with exactly one tube passing through . So . We also observe that , and so . Thus, by the Balog-Szemerédi-Gowers theorem 2.27, we can find , , and such that , , and for each , we have and . Then and , proving the Claim. ∎
Now, we apply Lemma 2.22 to the sets , the measure , and for a sufficiently large . The result, after applying Fubini’s theorem, is that we can find , , , and a subset with and such that for all , we have
Since and , we have in fact . In particular, for all .
The next leg of the proof is to show:
Claim 2. For all , if we write , then .
Proof. Note that Claim 1 tells us . Combining this with the Rusza triangle inequality (Lemma 2.23), , and for any subset of the doubling metric space , we have
The same argument shows (where ):
Thus, by Lemma 2.23 again, we have
Similarly, we have . A final application of Lemma 2.23 gives
This proves Claim 2. ∎
Finally, we seek a contradiction by applying Theorem 3.6 to and . We satisfy the condition (if is sufficiently small) because Claim 1 and tell us that is a -set with . But we violate the conclusion (if is sufficiently small) because Claim 2 tells us that . This contradiction finishes the proof of Theorem 3.7.
∎
3.5 Formal exhaustion argument
Using Theorem 3.7, we prove the following proposition, which implies Theorem 3.3 with a different value for .
Proposition 3.8.
For , , and , there exists such that the following holds for sufficiently small . Let be a -set of -tubes each making angle with the plane with . Let be a probability measure on such that for all , we have . Then .
The idea is the following. A first application of Theorem 3.7 gives a subset with . Either is large enough in which case we are done or we can cut out of and apply Theorem 1.8 again. This will give us another subset . Then we iterate until the union of these sets is large enough.
Proof.
Let be an integer. Suppose we have already constructed pairwise disjoint sets such that for every . Either we have
(3.4) |
in which case we stop, or the set satisfies the conditions of Theorem 3.3. In the latter case Theorem 3.3 gives us with . By construction, is disjoint with any of the , .
When this procedure ends write . Then (3.4) says . Moreover, since the ’s are disjoint, .
Set . We claim that
where the index set runs over subsets of with . Since for all , the desired upper bound then follows immediately from Markov’s inequality applied to the event (or [10, Lemma 20]).
We will now show the claim. Let , so there exists with and . Consider the index set defined as
We have
Hence . On the other hand, for all , since
we have for all . This finishes the proof of the claim. ∎
3.6 Proof of Proposition 3.1
This subsection is based on Section A.7 of [18].
We restate Proposition 3.1.
Proposition 3.9.
Given , , , there exist and such that the following holds for all .
Let be a -set, and for each , assume that is a -set with cardinality . Let
For every , assume that is a set of -tubes each making an angle with the plane with such that for all . Then , where .
Proof.
Let denote for some absolute constant . A -set stands for a -set.
First, without loss of generality, assume for each .
Suppose . Let
Since each tube in has angle with the plane , it only intersects many -balls for a given . Since for each , we get . With the counter-assumption , this forces for each . On the other hand, and so .
Now, we check that is a -set. Pick a -plane . We claim that either or is contained in a -plate. Indeed, if is not contained within a -plate, then is contained within the -neighborhood of the plane , which means that cannot contain any tubes of if is large enough (since the tubes of have angle with that plane). Thus, we may assume is contained within a -plate, which means
Since for each , there is a subset such that and each belongs to of the sets . We show is a -set. Indeed, given a -plate , we have
Finally, we refine further: since
we can find a subset with the property that for each . Also, is still a -set.
Now for each , the large subset has small covering number . On the other hand, . This contradicts Theorem 3.3 if is chosen sufficiently small in terms of the of the theorem. ∎
4 Improved incidence estimates for regular sets
In this section, we prove a version of Theorem 1.8 for regular sets.
Definition 4.1.
Let be a dyadic number. Let , and let . A non-empty set is called -regular if is a -set, and
Theorem 4.2.
For any , , , there exists such that the following holds for all small enough , depending only on . Let be a -regular set. Assume that for every , there exists a and -set with such that for all . Then .
4.1 Initial reductions
This subsection is based on Sections 6 and A.1-A.3 of [18].
In this section, let denote for some constant depending only on . Also, let .
The proof will be based on contradiction, so assume . Let’s rename to and to , reserving for the use of Proposition 2.13.
By Corollary 2.10, we have , so and . But is a -set, so . Finally, by Lemma 2.7, we may assume (passing to subsets will preserve the -regularity of ).
The next reduction will make the value uniform for different . Let . By -regularity of , we have . On the other hand, since is a -set, we have that for all ,
(4.5) |
This means . Hence, . Now using (4.5) again and , there exists with such that for each ,
(4.6) |
Using (4.6), we quickly check that is a -set. Indeed, for and , we have
(4.7) |
(The second inequality uses that is a -set.) Let ; then and for . Apply Proposition 2.13 to find , , , and . Let .
Claim. and .
Proof. By Proposition 2.13iii, we know that is -nice, so . Also, by Corollary 2.10, we have that
(4.8) |
Next, for any , we know that is -nice. Recall that
We also know and is a -set, so by a similar check to (4.7), we get that is a -set. Thus by Corollary 2.10, we have
But by our counterassumption , we get from (2.3) in Proposition 2.13 and ,
Thus, . Substitute into (4.8) to get
Thus, , so and , proving the Claim. ∎
Thus, we get the higher-dimensional analogues of properties (H1-2), (G1-4) of [18] except we only know and not . But this is not a limitation. We repeat and relabel these properties here:
-
(G1)
and for all .
-
(G2)
Every tube satisfies .
-
(G3)
For every square , there corresponds a -set and -set of cardinality such that for all .
-
(G4)
and .
-
(G5)
For , we have
Item 4 follows from and Claim.
4.2 Transferring angular non-concentration to ball non-concentration
This subsection is based on Section A.4 of [18].
We first recall some notation. For a unit vector , define to be the orthogonal projection to the orthogonal complement of . For a -tube , let denote the direction of .
In this subsection, we fix a . Our goal is to show that for many , the -rescaled version of contains a and -set for some . This is the content of the next Proposition 4.3, which is a higher-dimensional extension of Lemma A.6 of [18]. The proposition encodes the following principle: If we have a set of orthogonal projections in (which we view as ) that don’t concentrate around -planes, and we have a -dimensional set with , then many projections of will not concentrate around -planes.
Proposition 4.3.
Let , , and . Let be a -set in , and let be a -set and -set. There exists a subset with such that the following holds for all : if is an arbitrary subset of cardinality , then contains a and -set, where is absolute depending on .
Proof.
We will use a variation of the energy argument due to Kaufman [13] in the form used to prove [18, Lemma A.6]. An alternate proof can follow [10, Lemma 27], but this approach would give weaker bounds.
Let be the -discretized probability measure corresponding to ,
where is -dimensional Lebesgue measure. Since is a -set, we have for all , and it’s also true for since behaves like Lebesgue measure at small scales. We will choose a uniformly random and consider what happens to the energy of under projection by . By linearity of expectation and definition of energy,
Since is a -set, we have (c.f. [13]), and so by Lemma 2.30(a) and .
Analogously, we have (let ):
Observe that
where is the distance from to the plane spanned by through . (The first equality follows since is orthogonal to each . The second equality follows since is multilinear and . The third equality follows by the geometric definition of wedge product as a volume of a parallelepiped.) Thus, since is a -set and , we have
and so by Lemma 2.30(a) and .
Consequently, by Markov’s inequality we can find with such that for each , we have and . For any with , we have and , where is the renormalized restriction of to . Then Lemma 2.30(b) gives the desired conclusion. ∎
4.3 Finding a special -tube
This subsection is based on Section A.4 of [18].
Apply Proposition 4.3 to , which is a -set using 1 and the fact that is a -set. Define
where is the set of good directions of cardinality (since for a given direction, there are many -tubes in that direction that intersect ). Then remain -sets of cardinality , and so the properties 1-5 remain valid upon replacing with . (We leave unchanged, so only 3 and 5 are affected.) Thus, for and their large subsets have nice projections in the sense of Proposition 4.3 in every direction orthogonal to the tubes . We keep the symbol “” as a reminder of this fact.
The next goal is to find a tube with the following properties:
-
(P1)
The set contains a -subset, which we denote .
-
(P2)
.
-
(P3)
For each , there exists a subset such that
-
(P4)
Let be the direction of . Then contains a -set with cardinality , where .
To get 1- 3, we will mostly follow Section A.4 of [18]. (We have used the fact that is a -set, by converting it into ball concentration near -planes in Proposition 4.3; the rest of the argument will only use the fact that is a -set.) First, we refine the sets and further to ensure that the family will be -sets for . Indeed, we have
The first inequality uses the fact that is a -set of tubes with , and the second inequality uses the fact that is a -set with .
Thus, by Markov’s inequality, for a fixed absolute large constant , we have
(4.9) |
can only hold for many tubes .
Claim 2. If is sufficiently large, then there exists a subset with such that for all , at most half of the tubes satisfy (4.9).
Proof. Suppose this is not true: there exists a set such that for , at least many tubes satisfy (4.9). Then apply Corollary 2.10 to and the bad parts of , which are still )-sets. By Corollary 2.10, we have many -tubes in that satisfy (4.9). But we observed before that (4.9) only holds for many tubes . By choosing large enough (and small enough), we obtain a contradiction. ∎
In what follows, the in Claim 2 will be absorbed into the notation. Replace by and by their good subsets without changing notation. All of the properties 1-5 remain valid, and
(4.10) |
Now, we will find satisfying
(4.11) |
Indeed, the average tube works, because of the following: since , and (by 4, 1, 3 respectively), we have
Now, we show that using (4.10) and (4.11), the family contains a -set, which proves item 1. Indeed, rewrite (4.10) as
(4.12) |
Let
(4.13) |
By Markov’s inequality on (4.12), we have . Hence, if is chosen large enough, we have by (4.11), . By Markov’s inequality on (4.13), we have that for all and ,
Thus, is a -set, which proves 1.
4.4 Product-like structure
This subsection is based on Section A.6 of [18].
Our goal is to find a product-type structure and apply Proposition 3.1. Choose coordinates such that the -axis is in the direction of , and let denote the orthogonal projection to the orthogonal complement of the -axis. Define the function . If , then is roughly a -tube: it is contained in some -tube and contains a -tube for some universal constants . This technicality will not cause issues in what follows.
For each , let be a point such that the plane intersects . By 1, we know that is a -set. By 4, we know that for each that contains a -set with cardinality . Let that is rounded to the nearest multiple of .
5 Improved incidence estimates for general sets
In this section, we will prove the following refinement of Theorem 1.8, following Sections 7-9 of [18].
Theorem 5.1.
For any , , , , there exist and such that the following holds for all small enough , depending only on . Let be a -set with , and let be a family of -tubes. Assume that for every , there exists a and -set with such that for all . Then .
The original theorem follows from taking and pigeonholing, since .
Proof.
Before anything else, we state the dependencies of the parameters: , .
First, choose such that . By Lemma 2.17 we may find a subset with that is -uniform for with associated sequence . Thus, is a -set. Replacing with and with , we may assume from the start that is -uniform.
Let be the corresponding branching function. Since is a -set, we have for all .
Let be the intervals from Proposition 2.14 applied with parameters , corresponding to a sequence . We can partition , “structured” and “bad” scales such that:
-
•
for all , and ;
-
•
For each and , the set is either
-
(i)
an -regular set, where ;
-
(ii)
a -set.
-
(i)
-
•
.
Apply Proposition 2.14 and to get a family of tubes with the property that is a -nice configuration for some and
Let and . Then
6 Sets contained in an -plate
We restate Theorem 1.11.
Theorem 6.1.
For any , , , , , there exists such that the following holds for all small enough , with depending only on . Let be a -plate, be a -set with , and let be a family of -tubes. Assume that for every , there exists a set such that:
-
•
for all ;
-
•
is a -set down from scale ;
-
•
is a -set.
Then .
6.1 Multiscale analysis
We will use Theorem 1.8 to prove Theorem 1.11. Let be the dilation sending to . Then , and become deformed under , but they satisfy the following statistics assumptions for :
(6.15) | ||||
(6.16) | ||||
(6.17) |
To prove (6.15), observe that is contained in an -ball, and then we use that is a -set.
To prove (6.16), observe that is contained in a box with sides of length and sides of length . This box can be covered by many -balls. Finally, use that is a -set.
To prove (6.17), observe that is contained in a -plate.
Using these observations, we obtain the following refinement of Proposition 2.13. We use -nice configuration down from scale to indicate that is a -set down from scale .
Proposition 6.2.
Fix dyadic numbers . Let be a -nice configuration down from scale , and assume for some -plate . Then there exist refinements , , and such that denoting and the following hold:
-
(i)
and for all .
-
(ii)
We have for all .
-
(iii)
is -nice for some , , and .
-
(iv)
For all , we have
-
(v)
For each , there exist , , , a subset with and a family of tubes such that is -nice down from scale .
Furthermore, the families can be chosen so that
(6.18) |
Proof.
The proof will involve many dyadic pigeonholing steps.
Step 1: construct . For a given , we claim that we can find a subset with and a family of dyadic -tubes intersecting such that the following holds:
-
(T1)
is a -set and -set for some .
-
(T2)
there exists a constant such that
This claim generalizes [18, Proposition 4.1] and relies on the same dyadic pigeonholing steps; for brevity, we only state these steps and refer the reader to [18] for the detailed proof. (We essentially follow the same proof for 2, and we introduce a nice shortcut to derive 1 from 2.) Let be a minimal finitely overlapping cover of by -tubes. For , define
Since and , we in fact have
Thus, by dyadic pigeonholing, there exists such that . Another dyadic pigeonholing allows us to find such that is constant for . This is the desired refinement of . Finally, let
Then by a similar dyadic pigeonholing (for calculations, see [18, Proposition 4.1]), there is such that
(6.19) |
Finally, we define , which is the desired refinement of .
We check 2 holds with , which satisfies by (6.19) and . With this choice of , fix and note that
To check 1, we first pick a -tube with . Then by 2 and (6.16),
Thus, is a -set with . Doing the same calculation with an -plank instead of an -tube, we get that is a -set with . This proves 1 and thus the claim.
Step 2: uniformity of . By the pigeonhole principle, we can find and a subset with such that for all . Write
Next, by another dyadic pigeonholing, we can find a subset such that and for all . Also, for all . Thus, we can find with , and for each a subset of cardinality , such that . In other words,
Thus, we obtain item ii.
(6.20) |
Reduce the families such that their cardinality is . By 1, remains a and -set with .
For , , define
Thus, is a -nice configuration, establishing item iii. To summarize, in this step, we refined and for , so 2/iv still holds (with same and a weaker implied constant).
Step 3: uniformity of and construct . This step will be devoted to verifying v and (6.18). We will not change , or .
Fix , and let . Define
By dyadic pigeonholing and 2, we can find a -comparable subset of (which we keep denoting ) such that
Next,
(6.21) |
For a given , we consider the tube packet (discarding duplicate tubelets). Each tubelet lies in at most many tubes of , so by dyadic pigeonholing, we can refine by a factor to ensure that each tubelet lies in many tubes of , and there are many distinct tubelets through . By refining by a -factor, we may assume for each . Now, define
Since tubelets are essentially distinct and each tubelet in any corresponds to many tubes in , we obtain:
(6.22) |
Then (6.18) will follow by combining (6.20), (6.21), and (6.22).
We finally check is a -nice configuration down from scale . First, for any , we have for any -plank in ,
Thus, is a -set down from scale with . Similarly, is a -set with . This shows item v and thus the proof of the Proposition.
∎
6.2 Good multiscale decomposition
The idea is to apply Proposition 6.2, then apply Theorem 1.8 to bound and Corollary 6.5 to bound . Unfortunately, while we use pigeonholing to ensure that is a -set, we don’t know that is a -set. In fact, we won’t show this statement, but rather a slightly weaker statement that is good enough. For this, a good choice of based on the branching structure of is needed.
First, we explain the pigeonholing preliminaries.
Lemma 6.3.
Given , a -plane, there is a subset with such that is constant for all .
Proof.
Let . Then . For each , either or . Hence, if is the largest for which , we get . Thus, we have
Thus, by dyadic pigeonholing, there exists such that . ∎
The next step is to make satisfy a -dimensional spacing condition with just slightly less than , for all at a certain scale . To do so, we need the following lemma.
Lemma 6.4.
Fix , and let and be a -set in , a -plane. Let . If , then there exists and a subset with such that for any , , and , we have
(6.23) |
and for and a ball , we have
(6.24) |
Proof.
Throughout this proof we will not distinguish between and .
First, we will make uniform at scales . By Lemmas 6.3 and 2.17, we can find such that there is a sequence with for all and .
Let be the largest index such that for one (equivalently all) . Certainly is a valid index since .
Now, we will check the given conditions. By maximality of , we have for ,
Noticing that and likewise for , this proves (6.23).
Finally, we will need the following variant of Corollary 2.10.
Corollary 6.5.
Let , , and let . Let be a set contained in an -plate satisfying the following conditions:
-
•
For all and balls , we have
(6.25) -
•
For all and balls , we have
(6.26)
Assume that for every there exists a family of dyadic -tubes satisfying the following conditions:
-
•
for all ;
-
•
for all -tubes with .
Further assume that for some . If , then
Proof.
Let
We have the following:
Lemma 6.6.
For all , we have .
Proof.
We count by first choosing a dyadic , then counting the number of with , then finally counting the number of that pass through .
If , we claim that if and some tube through lies in , then , so .
To prove this, we may assume , as otherwise we can use the simple fact . Now choose a coordinate system such that the first axes correspond to the long sides of , and the remaining axes correspond to the short sides of . Let . Then . Furthermore, we have , otherwise any tube through would be roughly orthogonal to and intersect in a subtube with length , contradiction. Thus, we have .
Using the claim and condition (6.25), we see that there are many choices for . For each , the set of tubes passing through lies in a -tube, so by the tube non-concentration condition (and noting that ), we have choices for .
Thus, the contribution to for a given dyadic is , and summing over dyadic gives .
If , then by condition 6.26 we see that there are many choices for . For each , the set of tubes passing through lies in a -tube . We note that , so the tube non-concentration doesn’t apply directly, but luckily we note that can be covered by many -tubes. Thus, by using tube non-concentration at scale , we have choices for .
Thus, the contribution to for a given dyadic is (after some manipulation)
Since , the sum is .
Adding up both and contributions, we prove the Lemma. ∎
Proof of Theorem 6.1.
A small reduction: we would like to assume for all . To assume this, we first observe that for all . On the other hand, if for at least half of the (call them ) we have , then we are immediately done by Corollary 6.5 applied to and . Thus, by reducing if necessary, we may assume . Then by reducing further by a factor, we may assume for some . Finally, we may remove some tubes from each to make . Then is a -nice configuration.
Pick such that , and let . Pick such that . Find such that the conclusion of Lemma 6.4 holds. Now by Proposition 6.2, we have
If , where is the parameter in Theorem 5.1, we have .
Pick . Then satisfies the conditions of Corollary 6.5 with . Thus, we have . Using these two bounds and , we get
It remains to choose and also for small enough, we have and . Thus, and we are done. ∎
7 Power decay around -planes
In this section, we will roughly deal with the following situation:
-
•
are -Frostman measures with ;
-
•
gives mass to any -plate.
In other words, does not concentrate around -plates. We would like to understand the -mass of -plates for much smaller than . A result of Shmerkin [25, Proposition B.1] says that there exist , a subset with , and for each , a subset with such that for all and -plates through . Thus, we do obtain a power decay for sufficiently small . But what is the optimal starting point of the power decay? Can we hope for a power decay for all -plates through ? The answer is yes, and indeed we shall prove it by making small but meaningful tweaks to Shmerkin’s argument. But before stating our result, we shall introduce some convenient notation. We define thin -plates, a generalization of thin tubes, as follows.
Definition 7.1.
Let , , and . Let supported on . Fix . We say has -thin -plates on down from scale if
(7.27) |
Remark 7.2.
In this paper, we will choose where is small. (The complement is taken with respect to .) In this case, the equation (7.27) becomes
Now, we can state the main proposition, which generalizes and extends Proposition B.1 of [25]. It may be of independent interest.
Proposition 7.3.
Let and . There exist and with the following property. Fix and . Suppose that are positive measures with and for any -plate , we have
Let be the pairs of points that lie in some -concentrated -plate. Then there exists with such that have -thin -plates on . (The complement is taken with respect to .)
Remark 7.4.
(a) We can apply Proposition 7.3 in case are -dimensional with .
To prove Proposition 7.3, we need the following two lemmas. Fix . The first says that there are few dense -plates, and the second says that for most , the dense -plates through lie in some -plate.
Lemma 7.5.
There is such that the following holds: let be a measure with mass such that for all -plates , . Let be a set of -plates such that every -plate contains many -plates of (as in Section 2.2). Let . Then .
Lemma 7.5 follows from the condition on and the following generalization of [25, Lemma B.3]. In the case , the resulting bound is stronger but the assumption is also stronger.
Lemma 7.6.
Suppose for all -plates , . Then there exists a family of many -plates such that every -plate with is contained in some plate .
Proof.
Choose a maximal set of -plates such that
-
1.
,
-
2.
for .
We claim . Indeed, if and , then, then
(7.28) |
Now, , so . Combining with (7.28) gives , a contradiction.
Let be the -plates with same central -plane as . We show the problem condition. Given an -plate with , by maximality there exists such that . Thus, if is the largest principal angle between the central planes of and , then is contained in a box of dimensions
Thus, is contained in a -plate, so . Thus, , so is contained in . ∎
Remark 7.7.
We would like to present an alternative proof of Lemma 7.5, which was the original one found by the author. It gives slightly worse bounds but we believe it is slightly more motivated.
If then , so assume . Let . By induction, for each , there exist such that and that lie in at least many elements of .
The base case is trivial. For the inductive step, suppose are found. Let be the -neighborhood of the span of . Then since for every , we have . Thus, there is such that lie in at least many elements of , and by construction, . This completes the inductive step and thus the proof of the claim.
Finally, the set of -plates through must lie in a -plate, so at most many -plates of can lie in it. Thus, .
The following lemma is in the same spirit as [25, Proposition B.2].
Lemma 7.8.
Let be a collection of -plates, and suppose for all -plates , . Then for all except a set of -measure , there exists an -plate that contains every -plate in that passes through .
Proof.
The exceptional set is contained in the set of that lies in two plates of with “angle” . The intersection of two such plates is contained in a box with dimensions , which in turn is contained in a -plate (since ). Thus, by assumption on , this box has mass . Finally, there are pairs of plates in . ∎
Proof of Proposition 7.3.
Fix , and let , where is the constant in Lemma 7.5. We may assume . By Lemmas 7.5 and 7.8, we can find a set with and, for each , a set that is either empty or a -plate through such that for every intersecting .
Now, let and . We claim that and if , then . Then if , then for ; for any , we have for any -plate ,
Then have -thin -plates relative to .
To prove the first claim, we observe that if is sufficiently large in terms of .
Next, by definition of , we have . We also have the bound from the given condition (note that is a -plate). Thus,
if is chosen large enough. ∎
8 Radial projection estimates
In this section, we will first prove a key special case, and then the general case of Theorem 1.13.
8.1 Maximal plate concentration case
This subsection is based on ideas from [19].
Theorem 8.1.
Let , , and fix . There exists and depending on such that the following holds. Fix and . Let be -separated -dimensional measures with constant supported on , which lie in an -plate . Assume that . Let be the pairs of that lie in some -concentrated -plate. Then there exists a set with such that for every and -tube through , we have
The implicit constant may depend on .
Theorem 8.1 is the special case of Theorem 1.13 where are concentrated in a -plate for some small (we call this the maximal plate concentration case). For this, we closely follow the bootstrapping approach of [19]. There are three ingredients.
Proposition 8.2.
Let and , then there exists such that the following holds. Fix . Then for any -dimensional measures with constant contained in the -neighborhood of a -plane and , there exists with such that has -thin tubes on down from scale .
Proof.
Let be the projected measures on the -plane. Then satisfy -dimensional Frostman conditions for . Let
The rest is a standard argument following [8, Proof of Lemma 3.6]. Define the radial projection . Orponen’s radial projection theorem [17, Equation (3.5)] can be written in the form (where ):
(8.29) |
To effectively use (8.29), we will show that is small for . Indeed, let be a minimal set of finitely overlapping -tubes through such that any -tube through with lies in a -tube in . Then each -tube in has -measure . Since , we conclude that . Therefore, since the Lebesgue measure for a -tube through , we obtain . Finally, we can use Holder’s inequality and (8.29) to upper bound :
Choose to finish (the implicit constant is dominated by if is large enough). ∎
The bootstrapping step is as follows:
Proposition 8.3.
Let , , , . There exist and such that the following holds. Fix and . Let be -separated -dimensional measures with constant supported on , which lie in an -plate . Let . Suppose that and have -thin tubes and -thin -plates on down from scale . Then there exists a set with such that and have -thin tubes on down from scale . Furthermore, is bounded away from zero on any compact subset of .
Remark 8.4.
The reader is advised to set in the following argument, in which case it is a straightforward modification of [19, Lemma 2.8], with one small technical exception in the proof of the concentrated case, where we improve upon the dyadic pigeonholing step. Also if , then the simpler Theorem 1.8 can be used instead of Theorem 1.11 in the proof.
Proof.
We are given that for all ,
(8.30) | |||
(8.31) | |||
(8.32) | |||
(8.33) |
For and , let denote the -tubes through such that
(8.34) |
Now, let denote a covering of by essentially distinct -tubes. Then for , since , we have that the tubes in have -overlap on , so . For a dyadic , let , where denotes the union of the tubes in .
Claim. There are and such that the following holds for . If , then . Furthermore, is bounded away from zero on any compact subset of .
We will be done if we show the claim. Indeed, let ; then for any dyadic and any -tube through some , we either have , which means , or the negation of (8.34) holds. In either case, we get
(8.35) |
We have (8.35) for dyadic , but it also holds for all at the cost of introducing a multiplicative factor of on the RHS of (8.35). Thus, have -thin tubes on down from scale . Now we move to upper-bounding . By (8.30) and (8.34), we have for all , and so if from Claim, then
Let be the maximum of the value of from Claim, , and . Since , we get and have -thin tubes on down from scale . We can analogously find with such that have -thin tubes on down from scale , and so would be a good choice. Now we turn to proving the Claim.
Proof of Claim. We will choose , where is obtained from Theorem 1.11. From Remark 1.12 and the continuity of the function , we see that is bounded away from zero on any compact subset of .
Suppose that Claim is false. Let . Then .
Recall that for , the fiber is covered by , which is a set of cardinality . Let
Then for all . Furthermore, for every , we have
(8.36) |
The upper bound follows from , (8.30), and . In fact, we have in general,
We also take the time to state the thin plates assumption:
Since covers , we get by the upper bound in (8.36), . Hence, is a -set and -set for each .
Let . Call a tube concentrated if there is a ball with radius such that
(8.37) |
Suppose that there is with such that for each , at least half the tubes of are non-concentrated. Since and is Frostman with constant , we can find a -set . For each , the set of non-concentrated tubes is a -set and -set. Let . By Lemma 2.8, since , we have that is contained in the -neighborhood of . Now, we apply Theorem 1.11 with . Since and , we still have that for each , the set of non-concentrated tubes is a -set and -set. At this point, let us remark that implicit constants are dominated by if is chosen large enough.
Then if , where is obtained from Theorem 1.11, then
In other words, we get a gain of , which means a two-ends argument gives an immediate contradiction. Specifically, by (8.36) and (8.37), we have for each non-concentrated , . Thus, by Fubini, there exists a pair with such that for many tubes . However, since , we have that can only lie in many essentially distinct -tubes. Since , we get a contradiction.
Now we focus on the concentrated case: assume there is a subset with such that at least half of the tubes in are concentrated for all . This case is where we use the fact that is a -dimensional measure. Let denote the concentrated tubes and denote the corresponding heavy -balls. Because the family has -overlap on , the set
has measure
Notice that if , then there is a tube containing such that
Thus, can’t be too concentrated near :
assuming and . (The relevant inequalities are and .)
Therefore, for each , we can choose a dyadic number such that
where the annulus . (One remark: [19] used dyadic pigeonholing at this step, but we can’t do this because then we would introduce a factor. Fortunately, we are allowed to introduce the decaying tail , which is summable in .)
Then, recalling that , we can further find such that
By Fubini, we can find such that . Then by construction, can be covered by a collection of tubes containing that satisfy
Finally, we claim that contains a subset whose directions are separated by , such that if and if if . Indeed, if , then any -tube containing has
If , then any -tube containing lies in the union of many -tubes, and so
Thus, if , then it takes many -tubes to cover , and perhaps even more to cover . We may now choose to be a maximal subset with -separated directions to prove the claim for . A similar argument holds for .
Finally, let’s first assume . Since has bounded overlap in , we obtain
We will obtain a contradiction if we show the opposite inequality holds, for . Since and , it suffices to check .
If , then we obtain
Again, since , it suffices to check .
This proves the result. ∎
Proof of Theorem 8.1.
By Propositions 7.3 (with for ) and 8.2, there exists a set with such that and have -thin tubes on down from scale , and and have -thin -plates on . Then iterate Proposition 8.3 applied to a uniform . So initially we have , and after each iteration, becomes . After iterating many times and letting be the union of the ’s outputted from the Proposition (so ), we find that and have -thin tubes on . Then we can take to be our desired set. ∎
8.2 Proof of Theorem 1.13, general case
We will prove Theorem 1.13, which we restate here.
Theorem 8.5.
Let , , and . There exist depending on , and (with ) such that the following holds. Fix , and . Let be -separated -dimensional measures with constant supported on , which lie in . Assume that . Let be the pairs of that lie in some -concentrated -plate. Then there exists a set with such that for every and -tube through , we have
The implicit constant may depend on .
Remark 8.6.
Note that in Theorem 8.1, we demand the stronger conclusion .
The idea is to apply Theorem 8.1 at different scales. As a start, if , then we can directly apply Theorem 8.1 with (and thus we may take ).
We may assume for some . Let be the large constant in Lemma 7.5, and let . For , let be the pairs of that lie in some -concentrated -plate. We remark that .
Lemma 8.7.
Fix . There exists a set with such that for every and -tube through that intersects , we have
Proof.
By Lemma 2.8, there exists an absolute constant such that every -tube through some lies in some -concentrated -plate. We can find a collection of essentially distinct -concentrated -plates such that each -concentrated -plate is contained within some element of . By Lemma 7.5, . By construction, every -tube through some is contained in some member of . Apply Theorem 8.1 to each with measures and , , and to obtain a set with . Let , and then since . ∎
9 Corollaries of Radial Projection Estimates
We prove a variant of Corollary 1.1.
Proposition 9.1.
Fix and . Let be measures with and -separated supports. Suppose that for each -plane . Then for -almost all , for all sets of positive -measure,
Proof.
The proof is standard and follows [27, Proof of Proposition 6.9]. By Lemma 2.30, by passing to subsets of nearly full measure and replacing by an arbitrary , we may assume that for all .
Fix . By a compactness argument, there exists such that for all -plates . In Theorem 8.1, we know that for sufficiently small, the set . Thus, there exists with such that for every and -tube through , we have
Thus, there is a set with such that if , then
Taking completes the proof. ∎
Using this, we prove Corollary 1.2.
Corollary 9.2.
Let , then there exists such that the following holds. Let be Borel probability measures on with disjoint supports that satisfy and . Further, assume that don’t simultaneously give full measure to any affine -plane . Then there exist restrictions of to subsets of positive measure (which we keep denoting ) such that the following holds. For almost every affine 2-plane (with respect to the natural measure on the affine Grassmanian), if the sliced measures , on is non-trivial, then they don’t simultaneously give full measure to any line. In other words,
where we parametrize affine 2-planes as , for and in the Grassmannian with the rotationally invariant Haar measure .
Proof.
First, if for some affine -plane , then where denotes the complement of in . By restricting to and to (and calling the results ), we see that the sliced measures and can’t give full mass to any line for any affine -plane , for the simple reason that forces , and forces . Likewise, we are done if for some affine -plane . Thus, assume for all affine -planes .
Finally, we can deduce Theorem 1.1 from either Proposition 9.1 or Proposition 9.2, see [19, Section 4] for details. The only case not yet considered in this paper is when either gives positive mass to a -plane. But this special case was considered in [19, Section 4] (briefly, if gives positive mass to some -plane, then radial projections become orthogonal projections and then we apply Kaufman’s projection theorem; if gives positive mass to some -plane , then for , we have .)
Appendix A Proof of Balog-Szemerédi-Gowers
By a standard covering argument (e.g. see Section 3 of [12]), Theorem 2.27 follows from the case , which we prove below.
Theorem A.1 (refined Theorem 4.1 of [30]).
Let be a parameter. Let be finite subsets of , and let satisfy . Suppose that , where . Then one can find subsets with such that and .
Proof.
We follow the exposition in [24].
Claim. There exist subsets with , such that for each , there are many pairs such that , , and .
Assuming the claim, we will see how the theorem follows. First, we get . Since and , we get and .
Next, for , we have
Thus, there are many solutions to with . Since , we get .
Now, we prove the claim. For convenience, we can prune to satisfy (this is not necessary but will make the proof look nicer). Treat as a bipartite graph with an edge between and if . Then we want to find such that there are many paths of length between any .
The average degree of a vertex in is . Thus, if we delete the vertices in with degree , then at least many edges remain. Let be the set of edges. For , let be the set of neighbors of .
Now pick a vertex . On average, it has many neighbors.
Now, we say is bad if . For , let be the set of bad pairs in . There are many pairs in , so (expectation is taken over uniformly chosen )
If is the set of vertices of that lie in at least many pairs of , then
Finally, let . Then by linearity of expectation,
Thus, there exists such that . Then, let and
Let be the number of edges between and . We first check that . Indeed, since every vertex of has degree , we have
On the other hand, every vertex in corresponds to fewer than many edges of , so . Hence, .
Finally, for any , , we know that has at least many neighbors in , and fewer than of those form a bad pair with . For the remaining vertices that do not form a bad pair with , there are many vertices that are common neighbors of . Thus, we get at least many paths between and . ∎
References
- [1] Jean Bourgain. The discretized sum-product and projection theorems. Journal d’Analyse Mathématique, 112(1):193–236, 2010.
- [2] A. Carbery and S.I. Valdimarsson. The endpoint multilinear Kakeya theorem via the Borsuk–Ulam theorem. Journal of Functional Analysis, 264(7):1643–1663, 2013.
- [3] X. Du, A. Iosevich, Y. Ou, H. Wang, and R. Zhang. An improved result for Falconer’s distance set problem in even dimensions. Mathematische Annalen, 380(3):1215–1231, 2021.
- [4] X. Du, Y. Ou, K. Ren, and R. Zhang. New improvement to Falconer distance set problem in higher dimensions. arXiv preprint, 2023.
- [5] X. Du, Y. Ou, K. Ren, and R. Zhang. Weighted refined decoupling estimates and application to Falconer distance set problem. arXiv preprint, 2023.
- [6] X. Du and R. Zhang. Sharp estimates of the Schrödinger maximal function in higher dimensions. Annals of Mathematics, 189(3):837–861, 2019.
- [7] L. Guth, A. Iosevich, Y. Ou, and H. Wang. On Falconer’s distance set problem in the plane. Inventiones Mathematicae, 219(3):779–830, 2020.
- [8] L. Guth, N. Solomon, and H. Wang. Incidence estimates for well spaced tubes. Geometric and Functional Analysis, 29(6):1844–1863, 2019.
- [9] T.L.J. Harris. Low-dimensional pinned distance sets via spherical averages. The Journal of Geometric Analysis, 31(11):11410–11416, 2021.
- [10] W. He. Orthogonal projections of discretized sets. Journal of Fractal Geometry, 7(3):271–317, 2020.
- [11] Weikun He. Discretized sum-product estimates in matrix algebras. Journal d’Analyse Mathématique, 139(2):637–676, 2019.
- [12] Nets Hawk Katz and Terence Tao. Some connections between falconer’s distance set conjecture and sets of furstenburg type. New York J. Math, 7:149–187, 2001.
- [13] R. Kaufman. On Hausdorff dimension of projections. Mathematika, 15(2):153–155, 1968.
- [14] Tamás Keleti and Pablo Shmerkin. New bounds on the dimensions of planar distance sets. Geometric and Functional Analysis, 29(6):1886–1948, 2019.
- [15] B. Liu. Hausdorff dimension of pinned distance sets and the -method. Proceedings of the American Mathematical Society, 148(1):333–341, 2020.
- [16] P. Mattila. Geometry of sets and measures in Euclidean spaces: fractals and rectifiability. Number 44. Cambridge university press, 1999.
- [17] T. Orponen. On the dimension and smoothness of radial projections. Analysis & PDE, 12(5):1273–1294, 2018.
- [18] T. Orponen and P. Shmerkin. On the Hausdorff dimension of Furstenberg sets and orthogonal projections in the plane. arXiv preprint arXiv:2106.03338, 2021.
- [19] T. Orponen, P. Shmerkin, and H. Wang. Kaufman and Falconer estimates for radial projections and a continuum version of Beck’s Theorem. arXiv preprint arXiv:2209.00348, 2022.
- [20] Tuomas Orponen. An improved bound on the packing dimension of furstenberg sets in the plane. J. Eur. Math. Soc.(JEMS), 22(3):797–831, 2020.
- [21] Tuomas Orponen and Pablo Shmerkin. Projections, Furstenberg sets, and the sum-product problem, 2023.
- [22] Kevin Ren and Hong Wang. Furstenberg sets estimate in the plane, 2023.
- [23] I.Z. Ruzsa. On the cardinality of and . In Combinatorics (Proc. Fifth Hungarian Colloq., Keszthely, 1976), volume 2, pages 933–938, 1978.
- [24] Adam Sheffer. The Balog-Szemerédi-Gowers theorem, 2016.
- [25] P. Shmerkin. A non-linear version of Bourgain’s projection theorem. Journal of the European Mathematical Society, 2022.
- [26] P. Shmerkin and H. Wang. On the distance sets spanned by sets of dimension in . arXiv preprint arXiv:2112.09044, 2021.
- [27] P. Shmerkin and H. Wang. Dimensions of Furstenberg sets and an extension of Bourgain’s projection theorem. arXiv preprint arXiv:2211.13363, 2022.
- [28] Pablo Shmerkin. Improved bounds for the dimensions of planar distance sets. J. Fractal Geom., 8(1):27–51, 2021.
- [29] D.M. Stull. Pinned distance sets using effective dimension. arXiv preprint arXiv:2207.12501, 2022.
- [30] B. Sudakov, E. Szemerédi, and V.H. Vu. On a question of Erdős and Moser. 2005.
- [31] T. Tao and V.H. Vu. Additive combinatorics, volume 105. Cambridge University Press, 2006.
- [32] Z. Wang and J. Zheng. An improvement of the pinned distance set problem in even dimensions. In Colloquium Mathematicum, volume 170, pages 171–191. Instytut Matematyczny Polskiej Akademii Nauk, 2022.
- [33] T. Wolff. Decay of circular means of Fourier transforms of measures. Int. Math. Res. Not., 10:547–567, 1999.
- [34] J. Zahl. Unions of lines in . arXiv preprint arXiv:2208.02913, 2022.