An exceptional set estimate for restricted projections to lines in
Abstract.
Let be a non-degenerate curve in , that is to say, . For each , let and be the orthogonal projections. We prove an exceptional set estimate. For any Borel set and , define . We have .
Key words and phrases:
decoupling inequalities, superlevel set2020 Mathematics Subject Classification:
42B15, 42B201. Introduction
If is a curve that satisfies the non-degenerate condition
then we call a non-degenerate curve. A model example for the non-degenerate curve is .
In this paper, we study the the projections in whose directions are determined by . For each , let be the -dimensional subspace that is orthogonal to and let be the -dimensional subspace spanned by . We also define to be the orthogonal projection onto , and define to be the orthogonal projection onto . We use to denote the Hausdorff dimension of set . Let us state our main results.
Theorem 1.
Suppose is a Borel set of Hausdorff dimension . For , define the exceptional set
| (1) |
Then we have
| (2) |
As a corollary, we have
Corollary 1.
Suppose is a Borel set of Hausdorff dimension . Then we have
Remark 1.
The proof of Theorem 1 relies on the small cap decoupling for the general cone. We also remark that, for the set of directions determined by the model curve , Käenmäki, Orponen and Venieri can prove the exceptional set estimate with upper bound when (see [6] Theorem 1.3). The novelty of our paper is that we prove a Falconer-type exceptional set estimate for general non-degenerate curve, hence Corollary 1.
Remark 2.
Pramanik, Yang and Zahl [10] have also recently proved Corollary 1 with an exceptional set estimate of the form , compared to (2). Their proof is based on some incidence estimates for curves in the spirit of Wolff’s circular maximal function estimate. The estimates in [10] hold for curves that are only , which requires a very different proof from earlier work of Wolff and others on these problems.
Remark 3.
It is also an interesting question to ask for the estimate of the set
which consists of directions to which the projection of has zero measure. We notice that recently Harris [5] proved that
| (3) |
Intuitively, one may think of as ( is defined in (1)). The main result of this paper (2) yields which is better than the bound . This shows that (3) cannot imply (2).
Now we briefly discuss the history of projection theory. Projection theory dates back to Marstrand[8], who showed that if is a Borel set in , then the projection of onto almost every line through the origin has Hausdorff dimension . This was generalized to higher dimensions by Mattila[9], who showed that if is a Borel set in , then the projection of onto almost every -plane through the origin has Hausdorff dimension . More recently, Fässler and Orponen [2] started to consider the projection problems when the direction set is restricted to some submanifold of Grassmannian. Such problems are known as the restricted projection problem. Fässler and Orponen made conjectures about restricted projections to lines and planes in (see Conjecture 1.6 in [2]). In this paper, we give an answer to the conjecture about the projections to lines.
2. Projection to one dimensional family of lines
In this section, we prove Theorem 1. Theorem 1 will be a result of an incidence estimate that we are going to state later. Recall that a non-degenerate curve.
Definition 1.
For a number and any set , we use to denote the maximal number of -separated points in .
Definition 2 (-set).
Let be a bounded set. Let be a dyadic number, and let . We say that is a -set if
for any being a ball of radius with .
Let denote the -dimensional Hausdorff content which is defined as
We recall the following result (see [2] Lemma 3.13).
Lemma 1.
Let , and with . Then there exists a -set with cardinality .
Next, we state a useful lemma. It is proved in [3], but we still provide the full details here. The lemma roughly says that given a set of Hausdorff dimension less than , then we can find a covering of by squares of dyadic lengths which satisfy a certain -dimensional condition. Let us use to denote the lattice squares of length in .
Lemma 2.
Suppose with . Then for any , there exist dyadic squares so that
-
(1)
-
(2)
,
-
(3)
satisfies the -dimensional condition: For and any , we have .
Proof of Lemma 2.
Consider all the covering of by dyadic lattice squares that satisfy condition (1), (2) in Lemma 2, i.e., , and . We also assume all the dyadic squares in are disjoint. We will define an order “” between any two of such coverings . First, we define the -th covering number of by
which is the number of -squares in the covering .
We say , if they satisfy: (1) There is a such that (), and ; (2) For any , the square in that covers contains the square in that covers . It is not hard to check the transitivity: If and , then .
Suppose is a covering that is maximal with respect to the order . Then we can show that satisfies condition in Lemma 2. Suppose by contradiction, there exist and so that
| (4) |
We define another covering by adding to and deleting from . It is easy to check that is still a covering of . By (4), we can also check , so satisfies in Lemma 2. However, which contradicts the maximality of .
Now, it suffices to find a maximal element among all the coverings that satisfy condition in Lemma 2. First of all, such covering exists by the definition of Hausdorff dimension and . By Zorn’s lemma, it suffices to find an upper bound for any ascending chain.
Let be an infinite chain of coverings of . Define
We show that is a maximal element of the chain. First, we show that covers . For , let be the largest dyadic square in containing . By the definition of the partial order and the fact that chains are totally ordered, is independent of , and thus . This shows that is a covering of . Let . Choose such that for all and all . Then
Letting gives
So, satisfies condition . By definition, it is easy to check for every in the initial chain. This proves that is a maximal element. ∎
Remark 4.
Besides , this lemma holds for other compact metric spaces, for example or . The proof is exactly the same.
Our main effort will be devoted to the proof of the following theorem.
Theorem 2.
Fix . For each , there exists so that the following holds. Let . Let be a union of disjoint -balls and we use to denote the number of -balls in . Let be a -separated subset of such that is a -set and for some . Assume for each , we have a collection of -slabs with normal direction . satisfies the -dimensional condition:
-
(1)
,
-
(2)
, for any being a ball of radius .
We also assume that each -ball contained in intersects many slabs from . Then
2.1. -discretization of the projection problem
Proof of Theorem 1 assuming Theorem 2.
Suppose is a Borel set of Hausdorff dimension . We may assume . Recall the definition of the exceptional set
If , then there is nothing to prove. Therefore, we assume . Recall the definition of the -dimensional Hausdorff content is given by
A property for the Hausdorff dimension is that
We choose . Then , and by Frostman’s lemma there exists a probability measure supported on satisfying for any being a ball of radius . We only need to prove
since then we can send and . As and are fixed, we may assume is a constant.
Fix a . By definition we have . We also fix a small number which we will later send to . By Lemma 2, we can find a covering of by intervals , each of which has length for some integer . We define (Here denotes the length of ). Lemma 2 yields the following properties:
| (5) |
For each and -interval , we have
| (6) |
For each , we can find such a . We also define the slab sets , . Each slab in has dimensions and normal direction . One easily sees that . By pigeonholing, there exists such that
| (7) |
For each , define . Then we obtain a partition of :
By pigeonholing again, there exists such that
| (8) |
In the rest of the poof, we fix this . We also set . By Lemma 1, there exists a -set with cardinality .
Next, we consider the set . We also use to denote the counting measure on (note that is a finite set). Define the section of :
By (7) and Fubini, we have
| (9) |
This implies
| (10) |
since
| (11) |
By (10), we have
| (12) |
We are ready to apply Theorem 2. Recall and . By (12) and noting that , we can find a -separated subset of with cardinality . We denote the -neighborhood of this set by , which is a union of -balls. For each -ball contained in , we see that there are many slabs from that intersect . We can now apply Theorem 2 to obtain
Letting (and hence ) and then , we obtain . ∎
2.2. Proof of Theorem 2
For convenience, we will prove the following version of Theorem 2 after rescaling .
Theorem 3.
Fix . For each , there exists so that the following holds. Let . Let be a union of many disjoint unit balls so that has measure . Let be a -separated subset of so that is a -set and . Assume for each , we have a collection of -slabs with normal direction . satisfies the -dimensional condition:
-
(1)
,
-
(2)
, for any being a ball of radius .
We also assume that each unit ball contained in intersects many slabs from . Then
We define the cone
| (13) |
For any large scale , there is a standard partition of into planks of dimensions :
Here, the subscript of denotes its angular size. For any function and plank , we define as usual. The main tool we need is the following fractal small cap decoupling for the cone .
Theorem 4 (fractal small cap decoupling).
Suppose , where each is a -cap. Given a function , we say is -spacing if , where is a set of -caps from the partition of and satisfies:
| (14) |
If is -spacing, then we have
| (15) |
Small cap decoupling for the cone was studied by the second and third authors in [7], where they proved amplitude-dependent versions of the wave envelope estimates (Theorem 1.3) of [4]. Wave envelope estimates are a more refined version of square function estimates, and sharp small cap decoupling is a straightforward corollary. For certain choices of conical small caps, the critical exponent is (as is the case in our Theorem 4). When , the sharp small cap decoupling inequalities follow already from the wave envelope estimates of [4]. A version of this was first observed in Theorem 3.6 of [1] and was later thoroughly explained in §10 of [7]. To prove Theorem 4 above, we repeat the derivation of small cap decoupling from the wave envelope estimates of [4] but incorporate the extra ingredient of -spacing.
Remark 5.
We will actually apply Theorem 4 to a slightly different cone
| (16) |
Compared with , we see that is at distance from the origin, but we still have a similar fractal small cap decoupling for . Instead of (15), we have
| (17) |
The idea is to partition into many parts, each of which is roughly a cone that we can apply Theorem 4 to. By triangle inequality, it gives (17) with an additional factor . It turns out that this factor is not harmful, since we will set which can be absorbed into .
Proof of Theorem 3 assuming Theorem 4.
We consider the dual of each in the frequency space. For each , we define to be a tube centered at the origin that has dimensions , and its direction is . We see that is the dual of each . Now, for each , we choose a bump function satisfying the following properties: on , decays rapidly outside , and .
Define functions
From our definitions, we see that for any , we have . Therefore, we obtain
| (18) |
where “” means “”.
Next, we will do a high-low decomposition for each .
Definition 3.
Let be a large number which we will choose later. Define the high part of as
Define the low part of as
We choose a smooth partition of unity adapted to the covering which we denote by , so that
on . The key observation is that are at most -overlapping and form a canonical covering of . (See the definition of in (16)).
Since , we also obtain a decomposition of
| (19) |
where Similarly, we have a decomposition of
| (20) |
where
Recall that for , we have
We will show that by properly choosing , we have
| (21) |
Recall that . Since is a bump function at , we see that is an -normalized bump function essentially supported in the dual of . Denote the dual of by which is a -slab whose normal direction is . One actually has
Here, is bump function on and decays rapidly outside . Ignoring the rapidly decaying tails, we have
Recalling the condition (2) in Theorem 3, we have
This implies
Since , by choosing , we obtain (21). This shows that for , we have
We define . By remark (5), we actually see that form a -overlapping covering of , and we have the decoupling inequality (17). By (18), we have
By (17), it is further bounded by
Since the slabs in are essentially disjoint, the above expression is bounded by
This implies .
∎
2.3. Proof of Theorem 4
The proof of Theorem 4 is based on an inequality of Guth, Wang and Zhang. Let us first introduce some notation from their paper [4]. Let denote the standard cone in :
We can partition the -neighborhood of into -planks :
More generally, for any dyadic in the range , we can partition the -neighborhood of into -planks :
For each and frequency plank , we define the box in the physical space to be a rectangle centered at the origin of dimensions whose edge of length (respectively , ) is parallel to the edge of with length (respectively , ). Note that for any , is just the dual rectangle of . Also, is the convex hull of .
If is a translated copy of , then we define by
| (22) |
We can think of as the wave envelope of localized in in the physical space and localized in in the frequency space. We have the following inequality of Guth, Wang and Zhang (see [4] Theorem 1.5):
Theorem 5 (Wave envelope estimate).
Suppose . Then
| (23) |
Although the theorem above is stated for the standard cone , it is also true for general cone (see (13)). The appendix of [7] shows how to adapt the inductive proof of Guth-Wang-Zhang for to the case of a general cone .
As we did for , we can also define the -planks and -planks , which form a partition of certain neighborhood of . We can similarly define the wave envelope for . We have the following estimate for general cone.
Theorem 6 (Wave envelope estimate for general cone).
Suppose . Then
| (24) |
We are ready to prove Theorem 4.
Proof of Theorem 4.
By pigeonholing, we can assume all the wave packet of have amplitude , so we have
| (25) |
Apply Theorem 6 to , we have
For fixed , let us analyze the quantity on the right hand side. By definition,
Note that has dimensions , so its dual has dimensions . We will apply local orthogonality to each on . Let be a set of -planks that form a finitely overlapping covering of . We see that and each have the same angular size . For reader’s convenience, we recall that we have defined three families of planks: of dimensions ; of dimensions ; of dimensions .
Since , we have the nested property for these planks: each is contained in -dilation of some and each is contained in -dilation of some . We simply denote this relationship by . We can write
Choose a smooth bump function at satisfying: , decays rapidly outside , and is supported in . We have
Since and by a geometric observation that are finitely overlapping, we have
Summing over , we get
| (26) | ||||
| (27) | ||||
| (28) | ||||
| (29) | ||||
| (30) |
Therefore, we have
| (31) |
The last inequality is by the orthogonality.
Noting (25), we have
| (32) |
∎
References
- [1] C. Demeter, L. Guth, and H. Wang. Small cap decouplings. Geometric and Functional Analysis, 30(4):989–1062, 2020.
- [2] K. Fässler and T. Orponen. On restricted families of projections in . Proceedings of the London Mathematical Society, 109(2):353–381, 2014.
- [3] S. Gan, S. Guo, L. Guth, T. L. Harris, D. Maldague, and H. Wang. On restricted projections to planes in . arXiv preprint arXiv:2207.13844, 2022.
- [4] L. Guth, H. Wang, and R. Zhang. A sharp square function estimate for the cone in . Annals of Mathematics, 192(2):551–581, 2020.
- [5] T. L. Harris. Length of sets under restricted families of projections onto lines. arXiv preprint arXiv:2208.06896, 2022.
- [6] A. Käenmäki, T. Orponen, and L. Venieri. A Marstrand-type restricted projection theorem in . arXiv preprint arXiv:1708.04859, 2017.
- [7] D. Maldague and L. Guth. Amplitude dependent wave envelope estimates for the cone in , 2022.
- [8] J. M. Marstrand. Some fundamental geometrical properties of plane sets of fractional dimensions. Proceedings of the London Mathematical Society, 3(1):257–302, 1954.
- [9] P. Mattila. Hausdorff dimension, orthogonal projections and intersections with planes. Ann. Acad. Sci. Fenn. Ser. AI Math, 1(2):227–244, 1975.
- [10] M. Pramanik, T. Yang, and J. Zahl. A Furstenberg-type problem for circles, and a Kaufman-type restricted projection theorem in . arXiv preprint arXiv:2207.02259, 2022.