On restricted projections to planes in
Abstract.
Let be a non-degenerate curve in , that is to say, . For each , let and let be the orthogonal projections. We prove that if is a Borel set, then for a.e. we have .
More generally, we prove an exceptional set estimate. For and , define . We have . We also prove that if , then for a.e. we have .
Key words and phrases:
decoupling inequalities, superlevel set2020 Mathematics Subject Classification:
42B15, 42B201. Introduction
Let denote the unit sphere. Let be a curve. We say that is non-degenerate if
(1) |
for every . A model example for the non-degenerate curve is .
For a given , let denote the orthogonal complement of in , and let denote the orthogonal projection onto . For and a Borel set , we will use to denote the -dimensional Hausdorff measure of . Moreover, we use to denote the Hausdorff dimension of a set .
Theorem 1.
Suppose is a Borel set of Hausdorff dimension . For , define the exceptional set
Then we have
As an immediate corollary, we have:
Corollary 1.
Suppose is a Borel set of Hausdorff dimension . Then we have
Theorem 2.
Suppose is a Borel set of Hausdorff dimension greater than . Then
(2) |
for almost every .
1.1. Background of the problems
The projection theory dates back to Marstrand [18], 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 [19], who showed that if is a Borel set in , then the projection of onto almost every -plane through the origin has Hausdorff dimension . It is more general to consider the projection problem when the direction set is restricted to some submanifold of the Grassmannian. To have a better understanding of this restricted projection problem, the first step is to study the problem in . Fässler and Orponen made a conjecture about restricted projections to lines and planes (see Conjecture 1.6 in [5]), and there has been much related research (see for example [5], [2], [16], [15], [17], [21], [22], [23], [13], [14]). For more of an introduction to this problem, we refer to [13]. Recently, Käenmäki-Orponen-Venieri [17] and Pramanik-Yang-Zahl [25] proved one half of the conjecture: restricted projections to lines. In this paper, we resolve another half of the conjecture: restricted projections to planes.
1.2. An overview of the high-low method
The high-low method is a powerful tool developed recently in Fourier analysis. There are many applications of the high-low method, see for example [11], [3], [12], [10], [6], [7].
In this subsection, we briefly discuss how the high-low method can be used to study projection theory. As a warm-up, we study Marstrand’s projection theorem from another point of view, using the high-low method.
Theorem 3 (Marstrand’s projection theorem).
For each , define and let be the projection. Suppose is a Borel set, then we have for a.e. .
We will frequently use the following definition.
Definition 1.
For a number and any set , we use to denote the maximal number of -separated points in .
Marstrand’s projection theorem can be reduced to the following discretized version. We do not show how reduction works in this section, but we will give the full details in the later section when we prove our main theorems.
Proposition 1.
Fix . Fix a small scale and set . For each , define and let be the projection. Suppose is a union of disjoint -balls with measure , or equivalently . We also assume there is a subset with , such that for any , (which is a union of line segments of length in ) satisfies the -dimensional condition: For each and line segment of length , we have
(3) |
Then,
(4) |
Proof.
For each , let be a set of tubes that cover and hence cover . We can also assume that satisfies a similar -dimensional condition that is inherited from . For each , let be a -tube centered at the origin whose longest direction forms an angle with -axis. We see that is dual to the tubes in . Now, for each we choose a bump function satisfying the following properties: on , decays rapidly outside , and .
Define functions
By definition, for any , so we simply have
We can do better by performing a high-low decomposition for .
Let be a large number that will be determined later. Let be a smooth bump function adapted to and let . We have the following high-low decomposition for :
where and . We will show that the high part dominates on . Actually, for , we have
(5) |
By definition, . Note that is morally an -normalized bump function at . By the -dimensional condition of , we have
In the last step, we use the condition (3) in Proposition 1. As a result, we have
Since , by choosing large enough (depending on ) and plugging into (5), we see that
We obtain that
(6) |
Next, we will use a strong separation property for the high part. Note that , and is at most -overlapped. We have
In the last inequality, we used the -dimensional condition of .
As a result, we have
which implies . ∎
Notation.
-
(1)
For two positive real numbers and , we say that if there exists a large constant , depending on relevant parameters, such that ; we say that if .
-
(2)
We use for the Hausdorff dimension of .
-
(3)
For a given Borel measure supported on and the projection , the pushforward measure , supported on , is defined by
for every Borel .
-
(4)
Let be a compactly supported Borel measure on . Take . Define
where is the ball of radius centered at .
-
(5)
For and a rectangular box , we use to mean the dilation of by with respect to the center of , unless otherwise stated. For and , we use to mean .
-
(6)
We often use the Lebesgue measure of the set ; that is, .
-
(7)
By a measure we always assume that it is non-negative, unless stated otherwise.
Acknowledgement. S. Guo is partly supported by NSF-1800274 and NSF-2044828. L. Guth is supported by a Simons Investigator Award. H. Wang is supported by NSF Grant DMS-2055544. D. Maldague is supported by the NSF under Award No. 2103249.
2. Proof of Theorem 1
In this section we prove Theorem 1. First, we make some remarks on . We can cut into several small pieces and work on each of them. From now on, we assume that is and non-degenerate, and satisfies
(7) |
Here is sufficiently small depending on . Since the parameter does not play any role here, we may pretend . We need some notation.
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
Here, each in the covering is a cube and is the length of the cube. We recall the following result (see [5] Lemma 3.13).
Lemma 1.
Let , and with . Then there exists a -separated -set with cardinality .
Our main effort will be devoted to the proof of the following theorem.
Theorem 4.
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 -tubes pointing in direction . Each satisfies the -dimensional condition:
-
(1)
,
-
(2)
, for any being a ball of radius .
We also assume that each -ball contained in intersects many tubes from . Then
2.1. -discretization of the projection problem
In this subsection we show how Theorem 4 implies Theorem 1. Before starting the proof, we state a very useful lemma. We use the following notation. For any (), let denote the lattice of -squares in . For technical reasons, we remove the top edge and the right edge of each -square so that they are disjoint.
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 the lemma.
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 maximal 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
(8) |
We define another covering by adding to and deleting from . It is easy to check that is still a covering of . By (8), 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 an upper bound of the chain. First, we show that covers . For and , 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 . It also shows that the squares in are disjoint. 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 an upper bound of the chain. ∎
Remark 1.
Besides , this lemma holds for other compact metric spaces, for example or . The proof is exactly the same.
Proof that Theorem 4 implies Theorem 1.
Suppose is a Borel set of Hausdorff dimension . We may assume . Recall the definition of the exceptional set
Recall the definition of the -dimensional Hausdorff content is given by
A property for the Hausdorff content 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 may assume , otherwise . 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 disks , each of which has radius for some integer . We define . Lemma 2 yields the following properties:
(9) |
For each and -ball with , we have
(10) |
For each , we can find such a . We also define the tube sets , . Each tube in has dimensions and direction . One easily sees that . By pigeonholing, there exists such that
(11) |
For each , define . Then we obtain a partition of :
By pigeonholing again, there exists such that
(12) |
In the rest of the poof, we fix this . We also set . By Lemma 1, there exists a -separated -set with cardinality .
Next, we consider the set . We also use to denote the counting measure on . Define the section of :
By (11) and Fubini, we have
(13) |
This implies
(14) |
since
(15) |
By (14), we have
(16) |
We are ready to apply Theorem 4. Recall and . By (16), 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 tubes from that intersect . We can now apply Theorem 4 to obtain
Letting (and hence ) and then , we obtain .
∎
2.2. Proof of Theorem 4
The proof of Theorem 4 is base on the decoupling inequality for cone which is well-understood. For convenience, we will prove the following version of Theorem 4 after rescaling .
Theorem 5.
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 -tubes pointing in 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 tubes from . Then
We first discuss the geometry of . Let be the non-degenerate curve as discussed in the beginning of this section. We have . For convenience, we define
(17) |
We see that form a Frenet coordinate along . Define the corresponding conical surface .
We first show that satisfies the same non-degenerate condition as the standard cone. Note that we have the following formulae for the Frenet coordinate:
(18) | |||
(19) | |||
(20) |
where .
First, we show that is a surface. We will do this by finding a reparametrization so that is a function of . Choose
and then . We have
Since is , we have is with respect to , and therefore is with respect to . Moreover, by the above, which is nonvanishing since is nonvanishing.
For any large scale , there is a standard partition of into planks of dimensions :
For any Schwartz function , we define as usual. We have the following -decoupling inequality for these planks.
Theorem 6 (Bourgain-Demeter [1]).
For any Schwartz with , we have
(21) |
Remark 2.
We will actually apply Theorem 6 to a slightly different cone
(22) |
for some . Compared with , we see that is at distance from the origin, but we still have a similar decoupling inequality. Instead of (21), we have
(23) |
The idea is to partition into many parts, each of which is roughly a cone for which we can apply Theorem 6. By triangle inequality, this results in an additional factor . It turns out that this factor is not harmful, since we will set which can be absorbed into .
We are ready to prove Theorem 5.
Proof of Theorem 5.
Recall that is a collection of -tubes pointing to direction . We consider the dual of each in the frequency space. For each , we define to be a slab centered at the origin that has dimensions , and its shortest direction is parallel to . We see that is the dual rectangle 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
(24) |
for any . For our purpose, we just choose , so we have
(25) |
Our goal is to find an upper bound for the right hand side of (25). We will decompose into pieces and estimate the contribution of from each piece.
Let us discuss the decomposition for . Recall that is a -slab centered at the origin with normal direction . Recall (17), we can write .
Definition 3.
See Figure 1. Let be a large number which we will choose later. (Actually, we will choose ) Define the high part of as
Define the low part of as
For dyadic numbers , define
In particular, we define
Remark 3.
We obtain a partition of as
We see that consists of four planks of dimensions whose longest side is along direction . Here plays a role of angular parameter in the sense that are roughly those points in so that the lines connecting them with the origin form an angle with .
We choose a smooth partition of unity adapted to this covering which we denote by , so that
(26) |
on . Since , we also obtain a decomposition of
(27) |
where Similarly, we have a decomposition of
(28) |
where
Recalling (25) and using triangle inequality, we have
(29) |
We will discuss three cases depending on which term on the right hand side of (29) dominates.
Case 1: Low case
If the first term on the right hand side of (29) dominates, we say we are in the low case. Actually, we will see that we are never in the low case by showing
(30) |
for some large constant . This means the low term on the right hand side of (29) will not dominate. By properly choosing , we can show a pointwise bound for :
(31) |
Recall that . Since is a bump function at , we see that is a bump function essentially supported in the dual of . Denote the dual of by which is a -tube parallel to . One has
Here is a bump function on and decays rapidly outside .
By definition, is the sum of bump function of tubes. We have
(32) |
If we ignore the rapidly decaying tails, we have
(33) |
Recalling the condition (2) in Theorem 5, we have
This implies
(34) |
Since , by choosing , we obtain (31).
In the rest of the proof, we may pretend is a large constant, since any -loss is allowable (see Remark 2).
Case 2: High case
If the second term on the right hand side of (29) dominates, we say we are in the high case.
Since for any there is at most tube in pass through , we have (here we use instead of to take care of the rapidly decaying tail). Recalling the definition and noting that each is bounded, we have
We see that
(35) |
Next we will show that are finitely overlapping, i.e., are finitely overlapping. (Actually they are -overlapping. But since the -loss are acceptable, we may just pretend . See also Remark 2.) If this is true, then we have
(36) |
Since We obtain
(37) |
which implies
(38) |
Now we prove that are finitely overlapping. First, recall that
We see that is contained in the -neighborhood of the plane
To show the finitely overlapping property, we just need to show: For any and with (for some bounded to be determined later), if , then
Write , where and . Since the normal direction of is , it suffices to prove
(39) |
By Taylor’s expansion, we have . We see the left hand side of (39) is , if is large enough (depending on ).
Case 3: -middle case If the term on the right hand side of (29) dominates, we say we are in the -middle case. We remark that when is close to , can be estimated in a similar way as in the High case. We will be interested in the cone
Recall Remark 2 that we still have the decoupling inequality for this cone.
We first discuss the case that .
Case 3.1:
When , we have , where each is supported in . Note that consists of two pieces: One is
the other is
We note that lies in the -neighborhood of . Symmetrically, lies in the -neighborhood of , which is the reflection of with respect to the origin. We can write , so that and . We also write and , and then . We have
(40) |
By symmetry, we only estimate .
Note that and are essentially the same when ; and are essentially distinct when . We can choose a partition of by finitely overlapping planks of dimensions , denoted by . We attach each to one of the and denote by , if . We see that for each , there are many attached to it. We define . The Fourier support of is contained in , by Theorem 6, we have
By pigeonholing, we may pass to a subset of so that are all comparable. For simplicity, we write as , and write the number of after pigeonholing as . For each , we have by triangle inequality:
We obtain that
Note that , (by the -spacing of ), and . We obtain
Similarly, we have
As a result, we obtain
(41) |
Combined with , we have which is even better than what we aimed.
Case 3.2:
For being a dyadic scale in , we see that is supported in which consists of four separated planks ( only consists of two planks). As in the proof of case , we will write as the sum of four functions each of which has Fourier support in one of the planks of . We will estimate for one of the planks. We define
(42) |
Roughly speaking, is the top-right plank of and the distance between and the line is . For simplicity, we may assume has Fourier support in .
We discuss some geometric properties for the planks . First of all, there is a canonical finitely overlapping covering of by planks of dimensions . More precisely, we choose to be a set of -lattice points. For each , define
where is a large constant. We see that form a finitely overlapping covering of . We have the following three properties:
Lemma 3.
For defined above, we have
-
(1)
If , then is contained in .
-
(2)
If , then and are essentially the same.
-
(3)
If , then and are disjoint.
Before proving the lemma, we see how it can be used to finish the proof of Theorem 5. Motivated by Property (2) and (3), we define , and for each define
where is a large constant but much smaller than . Note that has the same dimensions as up to a -dilation.
Now we have three subsets of :
We will define a relationship between their elements. For any , we attach it to a such that , which we denote by . For any , we attach it to a such that , which we denote by . We also write if there is a such that and .
By property (1), if , then . By property (2), for a given , all the planks in are essentially the same and contained in . By property (3), if lie in different , then are disjoint.
Before estimating , we may apply a pigeonhole argument to pass to subsets of (still denoted by ), so that are comparable for , and are comparable for . For convinience, we write as , and write as .
We define
By decoupling for , we have
(43) |
By the trivial decoupling for , we have
(44) |
By Hölder’s inequality, we have
(45) |
Combining the three inequalities, we obtain
(46) |
Note that , (by the -spacing of ), and (by the -spacing of ). We also note that . We obtain
(47) |
∎
It remains to prove Lemma 3. Before proving the lemma, we give some intuition on why the lemma should be true. See Figure (2). We first cover by gray planks of dimensions . Fix a , we draw all the black slabs of dimensions whose corresponding is contained in . Morally speaking, which is a -plank. One tricky thing is that different may have essentially same , which is the reason to introduce (the thick-black planks in the Figure of dimensions ). Suppose we have a partition . We see that each is contained in one of the . If so, then we define We can talk about the intuition on the numerology of these planks.
-
(1)
,
-
(2)
,
-
(3)
,
-
(4)
.
By (2), we see Property (1) in Lemma 3 should be true. By (4), we see Property (2) and (3) in Lemma 3 should be true.
Proof of Lemma 3.
Recall . Defining
we have
(48) | |||
(49) | |||
(50) |
To prove Property (1), write with . Any can be written as with . By Taylor’s expansion, we have
(51) | ||||
One can easily check . The Property (2) can also be proved by using (51). For the Property (3), we have proved a special case in the High case, but here we need to do more work. We may assume . Consider the plane
We see is the -neighborhood of . We just need to show: For any and with (for some bounded to be determined later), if , then
Write , where and .
We consider two scenarios: (1) , (2) . If we are in the first scenario, since the normal direction of is , it suffices to prove
(52) |
By Taylor’s expansion, we have . We see the left hand side of (52) is , when and . If we are in the second scenario, we show that
(53) |
By Taylor’s expansion, we have . We see the left hand side of (53) is if the constant is big enough. ∎
3. Proof of Theorem 2
For a small positive number and , we use to denote a maximal -separated subset of . By definition, . If , then we abbreviate as and just choose it to be the -lattice points in . A rectangular box of dimensions will be referred to as a -tube. For each , there is a set of finitely overlapping collection of -tubes that cover whose long sides are parallel to .
In order to prove Theorem 2, we need the following result about incidence estimate.
3.1. An incidence estimate
Theorem 7.
Let be a -net of for some . Given a small constant and , let be a finite nonzero Borel measure supported in the unit ball in with . Suppose that is a set of -tubes, with directions in , such that each is disjoint, where we use to denote the subset of tubes in that points to direction . Suppose also that
(54) |
Then
(55) |
where the constant is allowed to depend on and but not on , and can be taken to be . We also remark that for .
Proof of Theorem 7.
The main argument of the proof is similar to that of Theorem 4, except for the “low case”.
For a given -tube , we denote the dual slab of by which is of dimensions and centered at the origin. Recalling (17), we can write
Remark 4.
The slab here is just the -dilation of that in the proof of Theorem 4. So are the tubes and that we will define right now.
Let be a non-negative function with supported on and for every . Set
(56) |
The assumption (54) implies that
(57) |
for every .
Next we will do the frequency decomposition for . Similar to Definition 3, we make the following definitions.
Definition 4.
(See Figure 1.) Let . Define the high part of as
Define the low part of as
For dyadic numbers , define
In particular, we define
Since , there are two cases:
Case 1: High case We can find a Borel set satisfying
and
(60) |
Case 2: Low case We can find a Borel set satisfying
and
(61) |
Assume first that we are in the high case. We raise both sides of (60) to the sixth power, integrate with respect to , and obtain
(62) |
Since the functions on the right hand side are locally constant on -balls, together with the upper density condition on , we obtain
(63) |
We can just use (37) and (47) with and , noting there is a scaling difference, so that the right hand side above is bounded by
It follows that
This finishes the proof if we are in the high case (60).
Now we assume that we are in the low case (61). For each , the support of is essentially a thickened tube of with dimensions where . We use to denote this thickened tube. Let be the collection of these thickened tubes obtained from , and we only keep those essentially distinct tubes (each tube intersects other tubes of those whose angle is within of its own, and every is contained in some from ).
Write as a disjoint union
(64) |
where is the collection of thickened tubes that contain tubes from . Here is a large universal constant which is much larger than the implicit (universal) constant in (61). Note that
(65) |
As a consequence, we see that (61) can be upgraded to
(66) |
Next, by (66) and the fact that if then , we conclude that
(67) |
Write
(68) |
The tubes in satisfy the induction hypothesis at the scale with the new parameter . Hence
Elementary computation shows that
(69) |
Therefore,
(70) |
Since , the induction closes. ∎
We have the following corollary of Theorem 7.
Corollary 2.
Let and . Let be a finite non-zero Borel measure supported on the unit ball in with . Let be a small number. Let be a -net of . For each , let be a disjoint collection of at most balls of radius in . Then there exists , depending only on and , such that
(71) |
where the constant depends on , but not on .
Proof of Corollary 2.
We argue by contradiction. Suppose that for every , there exist and , a disjoint collection of at most balls of radius in for each , such that (71) fails. Note that for each , is a -tube. We denote these tubes by
(72) |
and write . Define
(73) |
Note that by our contradiction assumption, we have
(74) |
This further implies
(75) |
Note that by definition, for every , it holds that
(76) |
We apply Theorem 7 to the measure restricted to and obtain
(77) |
By pigeonholing, this further implies that there exists such that
(78) |
This is a contradiction to the assumption that when is chosen to be small enough. ∎
The corollary below is a special case of Corollary 2. It is recorded below for a later application.
Corollary 3.
Let and . Let be a finite non-zero Borel measure supported on the unit ball in with . Let be a small number. Let be a -net of . For each , let be a disjoint collection of at most rectangles of dimension in whose long sides point in the direction. Then there exists , depending only on and , such that
(79) |
where the constant depends on , but not on .
Proof of Corollary 3.
For each and recalling (17), define
(80) |
which is a -plank. By a simple geometric observation, we have that
(81) |
for every with , where is a constant depending only on For each we have a set of -planks
which are essentially the translates of . For each , we choose a with . We partition each plank in into -tubes with direction and denote all these tubes by . We also define which is a collection of at most balls of radius in .
By Frostman’s Lemma (see for instance [20, page 112]), Theorem 2 is an immediate consequence of the following theorem.
Theorem 8.
Let be and non-degenerate. If is a compactly supported Borel measure on such that for some , then is absolutely continuous with respect to , for a.e. .
To prove Theorem 8, we will cut into finitely many pieces with a number depending only on , and work on one piece. From now on, we assume that is and non-degenerate, and satisfy
(82) |
Here is sufficiently small depending on .
3.2. Decomposition of the frequency space
In this subsection, we discuss the decomposition of the frequency space . Recall the cone that we are considering:
For any , there are three directions that we would like to specify: the normal direction ; the tangent direction ; and the flat (or radial) direction . We want to decompose into regions according to the distance from the origin, the distance from the cone , and the angular parameter. We give the precise definition below.
For a given integer , define
(83) |
Fix . If and , then we define the plank in the forward/backward light cone
(86) |
Here is some large constant that depends only on . It is chosen such that the distance from to cone is comparable to . Let us digest a little bit about the plank : is the distance from the origin; is the distance from the cone ; and is the angular parameter of the plank.
If , then for we define
(89) |
Let
(90) |
and
(91) |
Roughly speaking, for , forms a canonical covering of the part of outside the cone by -planks. Each forms a covering of the -neighbourhood of .
For each , we define to be a set of planks of dual dimensions to (but scaled by in each direction where and ) and forming a finitely overlapping covering of . We will refer to as the wave packets determined by the plank . Now, we discuss the wave packet decomposition. For each , we can choose a smooth bump function supported in and choose a smooth bump function supported in the unit ball, so that we have the partition of the unity
on the union of the ’s. For each , we can choose a smooth function which is essentially supported in (with rapidly decaying tail outside of ), such that and
For any , we define the wave packet
Lemma 4.
For and , we have
for every .
Proof of Lemma 4.
Note that , where is an normalized function essentially supported in (the dual plank of ). So, we have . Therefore,
∎
Lemma 5.
For and with
(92) |
it holds that
(93) |
for every , and .
3.3. Proof of Theorem 8: good part and bad part
The main idea is to divide the wave packets into two parts, called the good part and the bad part. We will prove an estimate for the bad part and an estimate for the good part.
Let be as in Theorem 8. Let be a small number (note is different from ) and let be determined later (we will later let ). For and , define
Define
(96) |
We remark that the wave packets of are those that have heavy -mass and not too far away from the cone . We have
We remark that and depends on parameters , but for simplicity we just omit them.
Lemma 6.
Let and . Fix . For all Borel measures supported on the unit ball in with , it holds that
where is defined by (96), and the implicit constant depends on and .
Lemma 7.
Let . Then for sufficiently close to , and small enough depending on and , and ,
(97) |
where the implicit constant depends on and .
Proof of Lemma 6.
By definition, we first write
(98) |
By the triangle inequality, this is
(99) | ||||
(100) |
By Lemma 5, the contribution from (100) is
By choosing .
To estimate (99), we discretize the integration in and bound it by
(101) |
By Lemma 4 the contribution from (99) is
(102) |
where
For fixed and , let be a finitely overlapping cover of the unit ball in by balls of radius . For each and let
Let be the pushforward of under . Denote
(103) |
with the centre of , and define
(104) |
Then
(105) |
Note that for each , the set is contained in a union of planks of dimensions ; the number of planks is
(106) |
for some large constant , and each plank overlaps of the others. Moreover and is supported in a ball of radius 1. Therefore, by applying the triangle inequality and Corollary 3, we can find , depending only on and , such that
for some large constant , whose precise value is not important. Putting this into (105) yields
(109) |
Substituting this into (102) and then (98) gives
(110) |
Recall that depends only on and . We just need to pick to be sufficiently small, and will finish the proof. ∎
Proof of Lemma 7.
Take . Given , note that
(111) |
Fix the coordinate . Any can be written in this coordinate as . We can also rewrite (111) as
Doing the Fourier transfom in the -plane, we have
By Plancherel’s theorem,
(114) |
Roughly speaking,
where is roughly with rapidly decaying tail outside , is the sum of good wave packets which have controlled mass, and is the sum of wave packets which are far away from the cone . A formula for is
The above used that is empty when , and is sufficiently large, which follows from the Frostman condition on .
Claim 1.
Let with . If there exist and satisfying
(115) |
then it holds that and .
Proof of Claim 1.
Recalling (86), we may assume
If we discuss some geometric observations. Noting that in the definition of , we see that if ; we also note that if are too far apart, then . Therefore we must have . In this case, in order for , we must have and , which finishes the proof. ∎
By Claim 1,we see that (114) is bounded by
(118) | ||||
(119) |
For the first term, the change of variables
(120) |
has Jacobian
(121) | ||||
where in the last step we used
(122) |
Applying this change of variables to (118) gives
Here is the -energy of and we used the fact that and is sufficiently small. The last step is because .
It remains to bound the contribution from , in (119). By frequency disjointness,
(123) |
Consider the case and separately. In the former case, we apply the change of variables as in (120) and obtain
(128) |
When , we show that (128) holds as well. To see this, we first observe that for each fixed , in order for
(129) |
not to vanish, has to take values on an interval of length ; next, we apply the two dimensional Plancherel’s theorem in the and variables for every fixed , and (128) follows from the uncertainty principle.
We continue to estimate (128) and do not distinguish and anymore. We have
(130) | ||||
(131) |
where
(132) |
and from (130) to (131) we applied Fubini and expanded the square. We cut the unit ball into small balls of radius and let be the restriction of to . By Cauchy-Schwarz,
(133) |
Let be a non-negative bump function such that for . By the Fourier support information of , we have
(134) |
By pigeonholing, we can find a subset
(135) |
such that is constant up to a factor of as varies over , and
(136) |
By pigeonholing again and by Hölder’s inequality, there is a disjoint union of balls of radius , such that
(137) |
and such that each intersects planks as varies over , for some dyadic number . By rescaling and then applying the refined decoupling inequality in Theorem 9 from the Appendix A, the first term in (137) satisfies
(138) |
For the second term in (137), the assumption that is finite implies that
(139) |
Hence by Hölder’s inequality and the definition of , we have
(140) | ||||
(141) |
Combining (138) and (141), we obtain
(142) |
Note that
(143) |
for every . Substituting into (142) and then into (130) yields
(144) | |||
(145) |
By Cauchy-Schwarz in the sum over , we obtain
(146) |
By substituting back into (130), we obtain
(147) |
In the end, we pick and finish the proof. ∎
Appendix A Refined decoupling inequality
The refined decoupling inequality stated here is a natural analogue of the refined decoupling inequality for the paraboloid from [9]. The shortest length of a plank dual to an -cap in the cone is , rather than in the case of the paraboloid, so the setup uses unit cubes instead of -cubes. The argument is similar to the paraboloid case, using induction and with Lorentz rescaling in place of parabolic rescaling, but the use of unit cubes requires the induction to be carried out over a finer sequence of scales (similarly to the induction setup in [4]). The refined decoupling inequalities in [13, 14] used tubes rather than planks, and the use of planks here is a significant reason for the improved result on the projection problem (at least in Theorem 2). Much of the proof is similar to [13, 14], with only the differences outlined above.
For each let For each , let
(148) |
If it is clear from the context which is used, then we often abbreviate to . Let For , denote . Let
(149) |
Moreover, we will use to denote the collection of translates of that cover . For a fixed small constant , denote , and . For , set .
Definition 5.
Fix . We say that a function is a -function if is supported on and
(150) |
Theorem 9.
Let be a curve with nonvanishing. Let be such that
(151) |
and
(152) |
Let and suppose that
where each is a -function and
Assume that for all ,
(153) |
Let be a disjoint union of unit balls in , each of which intersects at most sets with . Then for ,
Proof.
Assume that . Fix , , ,
and assume inductively that a (superficially) stronger version of the theorem holds with -cubes instead of unit cubes, where , for all scales smaller than , for all curves satisfying (151) and (152), and for all .
For each , let be the element of which minimises . For each , let
and
Let . Given any and corresponding ,
(158) |
and
(159) |
It follows that for each , there are sets with , and moreover whenever . For each such let be some choice such that , and let be the set of ’s associated to .
For each and , let be a finitely overlapping cover of by translates of the ellipsoid
Using Poisson summation, let be a smooth partition of unity such that on ,
and such that each satisfies
and
with supported in
By dyadic pigeonholing,
where, for each , is a union over a subset of the sets , and is the corresponding sum over , such that each intersects a number different sets with , up to a factor of 2. By pigeonholing again,
where and are constant over , up to a factor of 2. By one final pigeonholing step,
where is a union over -balls such that each ball intersects a number of the sets in a set of strictly positive Lebesgue measure, as varies over . Fix . By the decoupling theorem for generalised cones, followed by Hölder’s inequality,
Summing over gives
This will be bounded using the inductive assumption, following a Lorentz rescaling.
For each , define the Lorentz rescaling map at by
Let
Then for any ,
and
Hence
Let , and for fixed , let
(171) |
The assumption that yields
Similarly,
It follows that
on , and that
For each , given , let , where . Then
(172) |
The inequalities (158) and (159) imply that for each , the set is a equivalent (up to a factor 1.01) to a plank of length in its longest direction parallel to , of length in its medium direction, and of length in its shortest direction. The ellipsoids are rescaled to -balls . Moreover, it will be shown that
(175) |
To prove this, let
where
The vector is parallel to , since is orthogonal to and . The inequality
gives
(176) |
Moreover,
(177) |
For the direction ,
(178) |
Inductively applying the theorem at scale gives
for each . Hence
By the dyadically constant property of , this is
The second bracketed term is , since
It remains to show that . Let be any -ball. By definition of ,
By definition of and ,
(179) | ||||
The inequality (179) above follows from the observation that if , and if is such that , then . ∎
Appendix B Decoupling for cones
We have been using the decoupling inequality for cones in in a few places above, but it may have not been written down in the literature. In the appendix, we state it and sketch its proof. We start with the decoupling for curves on .
Theorem 10.
Let with be and satisfy for every . Then
(180) |
for every and . Here
(181) |
for an interval .
One can follow the same argument as in [8] to prove Theorem 10. We leave out the proof. Via the bootstrapping argument as in Bourgain and Demeter [1] and Pramanik and Seeger [24], one can prove the following decoupling for cones.
Theorem 11.
Let with be and satisfy for every . Then
(182) |
for every and . Here
(183) |
for an interval .
We give a sketch of the proof of Theorem 11. By the triangle inequality, we can assume that is supported on . The key observation in [1] is that the cone
(184) |
is in the -neighborhood of the cylinder
(185) |
for every . To see this, let us take one point from (184), and we will show that its distance to , which lies in (185), is . This amounts to proving
(186) |
Note that
(187) |
The desired bound (186) follows from Taylor’s expansion and mean value theorems.
References
- [1] J. Bourgain and C. Demeter. The proof of the decoupling conjecture. Ann. of Math., pages 351–389, 2015.
- [2] C. Chen. Restricted families of projections and random subspaces. Real Anal. Exchange, 43(2):347–358, 2018.
- [3] C. Demeter, L. Guth, and H. Wang. Small cap decouplings. Geom. Funct. Anal., 30(4):989–1062, 2020.
- [4] X. Du and R. Zhang. Sharp estimates of the Schrödinger maximal function in higher dimensions. Ann. of Math. (2), 189(3):837–861, 2019.
- [5] K. Fässler and T. Orponen. On restricted families of projections in . Proc. Lond. Math. Soc., 109(2):353–381, 2014.
- [6] Y. Fu, L. Guth, and D. Maldague. Sharp superlevel set estimates for small cap decouplings of the parabola. arXiv preprint arXiv:2107.13139, 2021.
- [7] S. Gan and S. Wu. Square function estimates for conical regions. arXiv preprint arXiv:2203.12155, 2022.
- [8] S. Guo, Z. K. Li, P.-L. Yung, and P. Zorin-Kranich. A short proof of decoupling for the moment curve. Amer. J. Math., 143(6):1983–1998, 2021.
- [9] L. Guth, A. Iosevich, Y. Ou, and H. Wang. On Falconer’s distance set problem in the plane. Invent. Math., 219(3):779–830, 2020.
- [10] L. Guth, D. Maldague, and H. Wang. Improved decoupling for the parabola. arXiv preprint arXiv:2009.07953, 2020.
- [11] L. Guth, N. Solomon, and H. Wang. Incidence estimates for well spaced tubes. Geom. Funct. Anal., 29(6):1844–1863, 2019.
- [12] L. Guth, H. Wang, and R. Zhang. A sharp square function estimate for the cone in . Ann. of Math., 192(2):551–581, 2020.
- [13] T. L. J. Harris. Improved bounds for restricted projection families via weighted fourier restriction. arXiv preprint arXiv:1911.00615, 2019.
- [14] T. L. J. Harris. Restricted families of projections onto planes: the general case of nonvanishing geodesic curvature. arXiv preprint arXiv:2107.14701, 2021.
- [15] E. Järvenpää, M. Järvenpää, and T. Keleti. Hausdorff dimension and non-degenerate families of projections. J. Geom. Anal., 24(4):2020–2034, 2014.
- [16] E. Järvenpää, M. Järvenpää, F. Ledrappier, and M. Leikas. One-dimensional families of projections. Nonlinearity, 21(3):453, 2008.
- [17] A. Käenmäki, T. Orponen, and L. Venieri. A Marstrand-type restricted projection theorem in . arXiv preprint arXiv:1708.04859, 2017.
- [18] J. M. Marstrand. Some fundamental geometrical properties of plane sets of fractional dimensions. Proc. Lond. Math. Soc., 3(1):257–302, 1954.
- [19] P. Mattila. Hausdorff dimension, orthogonal projections and intersections with planes. Ann. Acad. Sci. Fenn. Ser. AI Math, 1(2):227–244, 1975.
- [20] P. Mattila. Geometry of sets and measures in Euclidean spaces, volume 44 of Cambridge Stud. Adv. Math. Cambridge University Press, Cambridge, 1995. Fractals and rectifiability.
- [21] D. Oberlin and R. Oberlin. Application of a fourier restriction theorem to certain families of projections in . J. Geom. Anal., 25(3):1476–1491, 2015.
- [22] T. Orponen. Hausdorff dimension estimates for restricted families of projections in . Adv. Math., 275:147–183, 2015.
- [23] T. Orponen and L. Venieri. Improved bounds for restricted families of projections to planes in . Int. Math. Res. Not. IMRN, 2020(19):5797–5813, 2020.
- [24] M. Pramanik and A. Seeger. regularity of averages over curves and bounds for associated maximal operators. Amer. J. Math., 129(1):61–103, 2005.
- [25] 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.