Intersection patterns and incidence theorems
Abstract.
Let and be sets in a finite vector space. In this paper, we study the magnitude of the set , where runs through a set of transformations. More precisely, we will focus on the cases that the set of transformations is given by orthogonal matrices or orthogonal projections. One of the most important contributions of this paper is to show that if satisfy some natural conditions then, for almost every , there are at least elements such that
This infers that for almost every . In the flavor of expanding functions, with , we also show that the image grows exponentially. In two dimensions, the result simply says that if and , as long as , then for almost every , we can always find such that . To prove these results, we need to develop new and robust incidence bounds between points and rigid motions by using a number of techniques including algebraic methods and discrete Fourier analysis. Our results are essentially sharp in odd dimensions.
Key words and phrases:
Intersection, Group action, Rigid motion, Incidences, Distances2020 Mathematics Subject Classification:
52C10, 42B05, 11T231. Introduction
Let and be compact sets in . One of the fundamental problems in Geometric Measure Theory is to study the relations between the Hausdorff dimensions of , , and , where runs through a set of transformations.
This study has a long history in the literature. A classical theorem due to Mattila [18, Theorem 13.11] or [19, Theorem 7.4] states that for Borel sets and in of Hausdorff dimension and with
and assume in addition that the Hausdorff measures satisfy and , then, for almost every , one has
This result means that for almost every , the set of s such that has positive Lebesgue measure. This has been extended for other sets of transformations, for instance, the group generated by the orthogonal group and the homotheties [12, 17], the set of translations restricted on Cantor sets [1, Chapter 1], and the set of orthogonal projections [21]. A number of variants and applications can be found in a series of papers [1, 4, 5, 7, 22] and references therein.
Let be a finite field of order , where is a prime power. In this paper, we introduce the finite field analog of this type of questions and study the primary properties with an emphasis on the group of orthogonal matrices and the set of orthogonal projections. More precisely, we consider the following three main questions in this paper.
Question 1.1.
Given and , under what conditions on , , and can we have
(1.1) |
or a stronger form
(1.2) |
for almost every ?
Question 1.2.
Given and , under what conditions on and can we have
Question 1.3.
Let and be a positive integer.
-
(1)
If , then under what conditions on and can we have
for almost every ?
-
(2)
If , then under what conditions on and can we have
for almost every ?
Here denotes the orthogonal projection of onto .
Main ideas (sketch): This paper presents more than twenty new theorems on these three questions, for the reader’s convenience, we want to briefly explain the main steps at the beginning. Our study of the three above questions relies mainly on incidence theorems. While an incidence bound between points and affine subspaces due to Pham, Phuong, and Vinh [25] is sufficient to prove sharp results on 1.3, we need to develop incidence theorems between points and rigid-motions for the first two questions. In , such an incidence structure has been studied intensively in the breakthrough solution of the Erdős distinct distances problem [6, 8]. In this paper, we present the reverse direction, namely, from the distance problem to incidence theorems. This strategy allows us to take advantage of recent developments on the distance topic, as a consequence, we are able to establish a complete picture in any dimensions over finite fields. Our paper provides two types of incidence results: over arbitrary fields and over prime fields . The method we use is the discrete Fourier analysis in which estimates on the following sum
play the crucial role. While we use results from the Restriction theory due to Chapman, Erdogan, Hart, Iosevich, and Koh [2], and Iosevich, Koh, Lee, Pham, and Shen [10] to bound this sum effectively over arbitrary finite fields, the proofs of better estimates over prime fields are based on the recent distance estimate due to Murphy, Petridis, Pham, Rudnev, and Stevens [24] which was proved by using algebraic methods and Rudnev’s point-plane incidence bound [26]. This presents surprising applications of the Erdős-Falconer distance problem. We would like to emphasize here that the approach we use to bound the above sum over prime fields is also one of the novelties of this paper, and that sum will have more applications in other topics.
1.1. Intersection patterns I
Let us recall the first question. See 1.1
We note that for any , we have
Thus, there always exists such that .
Let us start with some examples.
Example 1: Let be a subspace of dimension with , , and be the set of matrices such that . It is well-known that . For any , we have
Example 2: In with odd, given , for each , there exist with and . To see this, let where is an arithmetic progression of size and . It is clear that the number of such that is at most .
These two examples suggest that if we want
for almost every , then and cannot be small and cannot be chosen arbitrary in . Our first theorem describes this phenomenon in detail.
Theorem 1.4.
Let and be sets in . Then there exists with
such that for any , there are at least elements satisfying
This theorem is valid in the range , since
The condition is sharp in odd dimensions for comparable sets and . A construction will be provided in Section 5. When one set is of small size, we can hope for a better estimate, and the next theorem presents such a result.
Theorem 1.5.
Let and be sets in . Assume in addition either ( odd) or (). Then there exists such that for any , there are at least elements satisfying
In particular,
-
(1)
If , then one has .
-
(2)
If , then one has .
In Theorem 1.5, the roles of and are symmetric. In a reduced form, the theorem says that if , , and , then
Our proof involves a number of results from Restriction theory in which the conditions on and are necessary. We do not know if it is still true for the case and , so it is left as an open question.
The sharpness construction of Theorem 1.4 can be modified to show that these two statements are also optimal in odd dimensions. In even dimensions, there is no evidence to believe that the above theorems are sharp. In the next theorem, we present an improvement in two dimensions.
Theorem 1.6.
Assume that . Let and be sets in with . Then there exists with
such that for any , there are at least elements satisfying
Remark 1.7.
Note that this theorem is stronger than Theorem 1.5 in the range . When we can switch between the roles of and to obtain a similar result, namely, there exists with
such that for any , there are at least elements satisfying
If our sets lie on the plane over a prime field , then further improvements can be made.
Theorem 1.8 ().
Assume that . Let and be sets in with . Then there exists such that for any , there are at least elements satisfying
In particular,
-
(1)
If and , then
-
(2)
If and , then
Theorem 1.9 ().
Assume that . Let and be sets in with . Then there exists such that for any , there are at least elements satisfying
In particular,
-
(1)
If and , then
-
(2)
If and , then
Remark 1.10.
We do not believe the results in even dimensions are optimal, but proving improvements is outside the realm of methods of this paper.
Remark 1.11.
In the above two theorems, the sets and cannot be both small since, otherwise, it might lead to a contradiction from the inequalities that . To compare to the theorems over arbitrary finite fields, we include the following table.
1.2. Intersection patterns II
For , define the map by
The results in the previous subsection say that if with and is larger than a certain threshold, then for almost every , one has .
In this subsection, we present a result for general sets . With the same approach, we have the following theorem.
Theorem 1.12.
Given , there exists with such that, for all , we have .
Remark 1.13.
We always have , since for each and for each , there is at most one such that . Thus, for all .
1.3. Incidences between points and rigid motions
We now move to incidence theorems.
Let be a set of points in and be a set of rigid motions in , i.e. maps of the form with and . We define the incidence as follows:
We first provide a universal incidence bound.
Theorem 1.14.
Let . Then we have
In this theorem and the next ones, the quantities and are referred to the main and error terms, respectively.
Under some additional conditions on and , if one set is of pretty small size compared to the other, then we can prove stronger incidence bounds.
Theorem 1.15.
Let for . Assume in addition that either ( odd) or ( and ).
-
(1)
If , then
-
(2)
If , then
In terms of applications (Theorem 1.4 and Theorem 1.5), one can expect that these two incidence theorems are sharp in odd dimensions. However, this is not true for even dimensions, and the next theorems present improvements in two dimensions.
Theorem 1.16.
Assume that . Let for . Then we have
A direct computation shows that this incidence theorem is better than the previous in the range .
In the plane over prime fields, we have the following three major improvements corresponding to three cases: , , and , respectively.
Theorem 1.17 ().
Assume that . Let for with . The following hold.
-
(1)
If and , then
-
(2)
If and , then
-
(3)
If and , then
Theorem 1.18 ().
Assume that . Let for with . The following hold.
-
(1)
If and , then
-
(2)
If and , then
-
(3)
If and , then
Theorem 1.19 ().
Assume that . Let for with . The following hold.
-
(1)
If and , then
-
(2)
If and , then
We now include a comparison to the results in offered by Theorem 1.15 and Theorem 1.16.
Remark 1.20.
The incidence theorems in two dimensions offer both the upper and lower bounds which depend simultaneously on the exponents of , , and . Hence, it is very difficult to come up with a conjecture that is sharp for most ranges.
1.4. Growth estimates under orthogonal matrices
For , we have seen that there exists a set such that for all , one has . In this subsection, we represent this type of results in the language of expanding functions, namely, assume , we want to have a weaker conclusion of the form for a given . Note that in this setting, we can obtain non-trivial results for small sets and . With the same proof, we have the following theorems.
Theorem 1.21.
Let . Given with and , there exists with
such that for all , we have
Theorem 1.22.
Let . Assume either ( odd) or (). Given with and , there exists such that for all , we have
In particular,
-
(1)
If , then one has .
-
(2)
If , then one has .
The two above theorems say that when the sizes of and belong to certain ranges, then the image grows exponentially for almost every . However, in two dimensions, the statement is much more beautiful: if and , as long as , then for almost every , we can always find such that .
Theorem 1.23.
Let . Assume that . Given with and , there exists with
such that for all , we have
In this paper, we do not compute the exponent explicitly in terms of and , but it can be improved when we replace by . The ranges of improvements are the same as those indicated in Table 2.
Theorem 1.24 (Small ).
Let . Assume that . Given with and , there exists such that for all , we have
In particular,
-
(1)
If and , then .
-
(2)
If and , then .
-
(3)
If and , then .
Theorem 1.25 (Medium ).
Let . Assume that . Given with and , there exists such that for all , we have
In particular,
-
(1)
If and , then .
-
(2)
If and , then .
-
(3)
If and , then .
Theorem 1.26 (Large ).
Let . Assume that . Given with and , there exists such that for all , we have
In particular,
-
(1)
If and , then .
-
(2)
If and , then .
1.5. Intersection pattern III
For a finite set and a subspace of , the orthogonal projection of onto is defined by
(1.3) |
where denotes the orthogonal complement of .
In or vector spaces over arbitrary finite fields, due to the fact that there exist null-vectors, i.e. vectors with , the orthogonal projection of onto is defined by
The elements of are -dimensional affine planes of when dim. We also note that as in the Euclidean we have the property that
for all subspaces .
Chen [3] proved the following result.
Theorem 1.27.
[3, Theorem 1.2.] Let .
-
(1)
For any ,
-
(2)
For any ,
Corollary 1.28.
[3, Corollary 1.3.] Let with .
-
(1)
If and , then
-
(2)
If , then
-
(3)
If , then
As mentioned earlier, we study the following question.
Question 1.29.
Let and be a positive integer.
-
(1)
If , then under what conditions on and can we have
for almost every ?
-
(2)
If , then under what conditions on and can we have
for almost every ?
The next theorem provides a partial optimal solution to this question.
Theorem 1.30.
Let and be a positive integer.
-
(1)
If , then there are at least subspaces such that
-
(2)
If , then there there are at least subspaces such that
-
(3)
Let . If , , and , then there there are at least subspaces such that
We now discuss the sharpness of this theorem. In the first statement, it is clear that the condition can not be replaced by for any . The second statement is sharp when and . To see this, let be a Kakeya set in of size , i.e. a set contains a full line in all directions, such an example can be found in [23]. Set . It is clear that for all directions . The third statement is also sharp in in the following sense: for any and any positive constant , there exist , , and there are at most subspaces such that . This is a long construction, so we omit it here and present it in detail in Section 8. When , we do not have any examples for its sharpness.
To keep this paper not too long, we do not want to make a full comparison between the results in this paper and those in the fractal. There is one crucial point we have to mention here that while Theorem 1.4, Theorem 1.5, and Theorem 1.12 are directly in line with Mattila’s results in [18, Theorem 13.11] and [20], we are not aware of any results in the continuous setting that are similar to those in or . This suggests that there might be room for improvements in .
2. Preliminaries-key lemmas
Let be a complex valued function. The Fourier transform of is defined by
here, we denote by a non-trivial additive character of . Note that satisfies the following orthogonality property
We also have the Fourier inversion formula as follows
With these notations in hand, the Plancherel theorem states that
In this paper, we denote the quadratic character of by , precisely, for , if is a square and otherwise. The convention that will be also used in this paper.
This section is devoted to proving upper bounds of the following sum
which is the key step in our proofs of incidence theorems.
2.1. Results over arbitrary finite fields
We first start with a direct application of the Plancherel theorem.
Theorem 2.1.
Let be sets in . Then we have
To improve this result, we need to recall a number of lemmas in the literature.
For any , let be the sphere centered at the origin of radius defined as follows:
The next lemma provides the precise form of the Fourier decay of for any . A proof can be found in [9] or [13].
Lemma 2.2.
For any , we have
where is the quadratic character, and if and otherwise.
Moreover, for , we have
For , define
We recall the following result from [13], which is known as the finite field analog of the spherical average in the classical Falconer distance problem [16, Chapter 3].
Theorem 2.3.
Let . We have
-
(1)
If , then .
-
(2)
If even, then .
-
(3)
If odd, then .
In some specific dimensions, a stronger estimate was proved in [10] for the sphere of zero radius.
Theorem 2.4.
Let . Assume and , then we have
We are now ready to improve Theorem 2.1.
Theorem 2.5.
Let be sets in . Assume that either ( odd) or ( and ), then the following hold.
-
(1)
If , then
-
(2)
If , then
Proof.
The proof follows directly from Theorem 2.3 and Theorem 2.4. More precisely,
This completes the proof. ∎
In two dimensions, we can obtain a better estimate as follows.
Theorem 2.6.
Let be sets in . Assume in addition that , then we have
Proof.
We note that when , the circle of radius zero contains only one point which is , so
Applying Theorem 2.3, as above, one has
and
Thus, the theorem follows. ∎
2.2. Results over prime fields
To improve Theorem 2.6 over prime fields, we need to introduce the following notation.
For . Define
The following picture describes the case .
![[Uncaptioned image]](https://cdn.awesomepapers.org/papers/07be8e81-0edf-4eb8-acb9-7bcb7f77d4e3/x1.png)
We note that counts the number of pairs in and of the same distance. This quantity is not the same as the distance estimate of the set .
While the sum can be bounded directly over arbitrary finite fields, the strategy over prime fields is different. More precisely, we will use a double counting argument to bound . The first bound is proved in the next theorem which presents a connection between the sum and the magnitude of . The second bound (Theorem 2.8, Corollary 2.9) for is due to Murphy, Petridis, Pham, Rudnev, and Stevens [24] which was proved by using algebraic methods and Rudnev’s point-plane incidence bound [26].
Theorem 2.7.
Let for . Then we have
Proof.
For any and , we define
Therefore, for any , we have
Thus,
Lemma 2.2 tells us that
Therefore,
This completes the proof. ∎
Theorem 2.8.
For , with , and , we have
for some large constant .
Corollary 2.9.
For , with , and , there exists a large constant such that the following hold.
-
(1)
If , then .
-
(2)
If , then .
-
(3)
If , then .
Remark 2.10.
We want to add a remark here that [24, Theorem 4] presents a bound on the number of isosceles triangles in . This implies the bound for as stated in Theorem 2.8 since is at most the number of isosceles triangles times the size of by the Cauchy-Schwarz inequality.
Theorem 2.11.
Let with and . Define
Then the following hold.
-
(1)
If and , then .
-
(2)
If and , .
-
(3)
If and , then .
-
(4)
If and , then .
-
(5)
If and , then .
-
(6)
If and , then
-
(7)
If and , then .
-
(8)
If and , then .
-
(9)
If and , then .
To see how good this theorem is, we need to compare with the results obtained by Theorem 2.5 and Theorem 2.6. More precisely, the two theorems give
The following table gives the information we need.
Proof.
We have
From Theorem 2.7, one has
and
By the Plancherel, we know that
Hence,
-
(1)
If , then Corollary 2.9 gives
-
(2)
If , then Corollary 2.9 gives
-
(3)
If , then Corollary 2.9 gives .
The sum is estimated in the same way, namely,
-
(1)
If , then Corollary 2.9 gives
-
(2)
If , then Corollary 2.9 gives
-
(3)
If , then Corollary 2.9 gives .
When , we use Theorem 2.6 to get that .
Combining these estimates gives us the desired result. ∎
2.3. An extension for general sets
In this subsection, we want to bound the sum
(2.1) |
where is a general set in .
Theorem 2.12.
Let . We have
To prove this result, as in the prime field case, we use a double counting argument to bound . To prove a connection between and the sum 2.1, a number of results on exponential sums are needed.
For each , the Gauss sum is defined by
The next lemma presents the explicit form the Gauss sum which can be found in [15, Theorem 5.15].
Lemma 2.13.
Let be a finite field of order , where is an odd prime and We have
We also need the following simple lemma, its proof can be found in [14].
Lemma 2.14.
For and , we have
Let be the variety defined by
The Fourier transform of can be computed explicitly in the following lemma.
Lemma 2.15.
Let .
-
(1)
If , then
-
(2)
If and , then
-
(3)
If and , then
Proof.
By Lemma 2.14, we have
By Lemma 2.13, we have . Thus, the lemma follows from the orthogonality of the character . ∎
In the following, we compute explicitly which is helpful to estimate the sum 2.1.
Lemma 2.16.
For , we have
Proof.
We now bound by a different argument.
Theorem 2.17.
For . We have
Remark 2.18.
Proof.
To prove this lemma, we start with the following observation that
can be written as
We now write as follows
here We now estimate the term .
In other words, we obtain
This completes the proof. ∎
With Lemma 2.16 and Theorem 2.17 in hand, we are ready to prove Theorem 2.12.
Proof of Theorem 2.12.
Indeed, one has
By Plancherel theorem, we have
So, the theorem follows directly from Theorem 2.17. ∎
3. Warm-up incidence theorems
In this section, we present direct incidence bounds which can be proved by using the Cauchy-Schwarz inequality and the results on from the previous setion.
Theorem 3.1.
Let be a set of points in and be a set of rigid motions in . Then we have
Proof.
For each , denote by . Then, it is clear that
(3.1) |
We observe that, for each , counts the number of pairs on . This infers . Thus, the sum can be bounded by
where we used the fact that the stabilizer of a non-zero element is at most , and the term comes from pairs with and . Therefore,
Using Theorem 2.17, the theorem follows. ∎
If we use the trivial bound , then the next theorem is obtained.
Theorem 3.2.
Let be a set of points in and be a set of rigid motions in . Then we have
Compared to Theorem 1.14 and Theorem 1.15, these two incidence theorems only give weaker upper bounds and tell us nothing about the lower bounds.
In two dimensions over prime fields, if with , then Corollary 2.9 says that . As above, the next theorem is a direct consequence.
Theorem 3.3.
Let with and . Assume that , then we have
In particular, if then
4. Incidence theorems: proofs
Let us present a framework that will work for most cases.
We have
where . We next bound the second term. By the Cauchy-Schwarz inequality, we have
We now consider two cases.
Case : If is a general set in , then we have
where we used the fact that the stabilizer of a non-zero element in is at most . From here, we apply Theorem 2.12 to obtain Theorem 1.14.
Case : If is of the structure , where , then
Thus,
where we again used the fact that the stabilizer of a non-zero element in is at most .
From here, we apply Theorem 2.5, Theorem 2.6, and Theorem 2.11 to obtain Theorem 1.15, Theorem 1.16, Theorem 1.17, Theorem 1.18, and Theorem 1.19, except Theorem 1.17 (2) and (3).
To prove these two statements, we need to bound the sum in a different way. More precisely,
Moreover,
The first approach is equivalent to bound by using Theorem 2.7 and Theorem 2.11.
If and , then Theorem 2.11 tells us that
As a consequence, one has
However, when and , by the Cauchy-Schwarz inequality, a better upper bound can be obtained. Indeed, using Corollary 2.9, we have
Together with the above estimates, we obtain
This gives
Similarly, if and , we have
and
Hence,
Sharpness of Theorem 1.14 and Theorem 1.15 in odd dimensions:
We first show that Theorem 1.14 is sharp up to a constant factor.
Let be an arithmetic progression in , and let be vectors in such that for all . The existence of such vectors can be found in Lemma 5.1 in [9] when () or ( with ). Define
here . Set . The number of quadruples such that , is at least a constant times , say, . For each , let . Define . So, .
We call , , if the rank of the system
is .
For any pair , where is , the number of such that is at most .
For , if is , then, assume,
Let be the contribution to of pairs such that is . We have is at most the number of s times . For each , to count the number of s, we observe that , so . The number of elements of norm zero in is at most . So, the total number of s such that
is at most , which is, of course, larger than the number of satisfying
Summing over all and the corresponding s, the contribution to is at most . So, the pairs , where is and , contribute at most which is much smaller than . Thus, we can say that the contribution of mainly comes from s.
Let be the set of pairs such that and is .
Whenever , , by a direct computation, Theorem 1.14 shows that
for some positive constant . This gives . This matches the lower bound of up to a constant factor.
We note that this example can also be used to show the sharpness of Theorem 1.15(2) in the same way.
For the sharpness of Theorem 1.15(1), let with , . Set
where . Since any vector in is of the form
where , we have for all . Let be the set of such that . Then, we have . Theorem 1.15(1) gives
for some positive constant .
On the other hand, by the definitions of , , and , we have
This matches the incidence bound up to a constant factor.
5. Intersection pattern I: proofs
In this section, we prove Theorem 1.4, Theorem 1.5, and Theorem 1.6.
Proof of Theorem 1.4.
Set
for some .
We first show that . Indeed, let be the set of pairs with and . It is clear that . On the other hand, Theorem 1.14 also tells us that
for some positive constant . Thus, we have
Together this with implies the desired conclusion.
Similarly, for , let be the set of pairs with and . Then, . By Theorem 1.14 again, we obtain
and thus
The fact implies our desired result .
Next, note that for any , by setting , we have
implying that there at least elements satisfying
∎
Proof of Theorem 1.5.
We consider the case when , since other cases can be treated in the same way. We use the same notations as in Theorem 1.4. By Theorem 1.15, there exists such that
and
This, together with and , implies that
As similar to the proof of Theorem 1.4, the theorem follows. ∎
Proof of Theorem 1.6.
We use the same notations as in Theorem 1.4 for . By Theorem 1.16, there exists such that
and
This, together with and , gives that
As similar to the proof of Theorem 1.4, the theorem follows. ∎
Using Theorem 1.18 and Theorem 1.19 respectively, the proofs of Theorem 1.8 and Theorem 1.9 can be obtained by the same way.
Sharpness of Theorem 1.4 and Theorem 1.5:
The constructions we present here are similar to those in the previous section. For the reader’s convenience, we reproduce the details here.
It follows from the proof of Theorem 1.4 that for any , there exists such that if , then there are at least pairs such that
(5.1) |
We now show that this result is sharp in the sense that for small enough, say, , there exist with , , such that the number of pairs satisfying 5.1 is at most .
To be precise, let be an arithmetic progression in , and let be vectors in such that for all . The existence of such vectors can be found in Lemma 5.1 in [9] when () or ( with ). Define
here .
We first note that the distance between two points in or is of the form . By a direct computation and the fact that is an arithmetic progression, the number of quadruples such that , is at least a constant times , say, . For each , let . Define . So, .
We note that . If there were at least pairs satisfying 5.1, then we would bound in a different way, which leads to a bound that much smaller than , so we have a contradiction.
Choose pairs satisfying 5.1. The contribution of these pairs is at most to .
The number of remaining pairs is at most . We now compute the contribution of these pairs to . As before, we call , , if the rank of the system is .
For any pair , where is , the number of such that is at most . Thus, the contribution of pairs with s is at most .
As before, the contribution to of s, with , is at most . In other words, we have
when is large enough. By choosing small enough, we see that , a contradiction.
The second statement of Theorem 1.5 is valid in the range . This is also sharp in odd dimensions by the above construction, since we can choose and the conclusion fails.
The first statement of Theorem 1.5 is valid in the range . To see its sharpness, we construct two sets and with , , and the number of in such that is at most for any . Let are linearly independent vectors in . Let with . Set
where . Since any vector in is of the form
where , we have for all .
6. Intersection pattern II: proofs
In this section, we prove Theorem 1.12.
Proof of Theorem 1.12.
Let be the set of in such that and . By Theorem 1.14, we first observe that
Note that and . This infers
So,
as desired. ∎
7. Growth estimates under orthogonal matrices: proofs
In this section, we prove Theorem 1.21, Theorem 1.22, Theorem 1.23, Theorem 1.24, Theorem 1.25, and Theorem 1.26.
Proof of Theorem 1.21.
Set . Let . Define
Set . We observe that . Applying Theorem 1.14, one has a constant such that
Using the fact that with , we have
This implies
This completes the proof. ∎
Proof of Theorem 1.22.
We use the same notations as in Theorem 1.21, and assume that , since the other case can be proved similarly. By Theorem 1.15, one has a constant such that
Using the fact that with , we have
This implies
as desired. ∎
Proof of Theorem 1.23.
We use the same notations as in Theorem 1.21 for . Applying Theorem 1.16, there exists a constant such that
Using the fact that with , we have
This implies
as desired. ∎
Theorem 1.24 (1) and (2), Theorem 1.25, and Theorem 1.26 are proved by the same approach using Theorem 1.17, Theorem 1.18, and Theorem 1.19, respectively.
For Theorem 1.24 (3), the same proof implies that if and , then . However, if we use Theorem 3.3, then we are able to get rid of the term .
8. Intersection pattern III: proofs
In this section, we prove Theorem 1.30. Let us start by introducing the necessary theorem.
Theorem 8.1.
[25, Theorem 1.6] Let be the set of -planes and let be the set of -planes in with . Then the number of incidences between and satisfies
where .
See 1.30
Proof of Theorem 1.30.
We proceed as follows.
-
(1)
By applying Theorem 1.27(2), we have
for any with and . Set . Then, the number of -dimensional subspaces such that is at most , where for some . The same happens for the set . This implies that there are at most -dimensional subspaces such that or , where . That is, there are at least -dimensional subspaces such that and . We note that
The second inequality is from since the dimension of is . Therefore, there are at least -dimensional subspaces such that
This completes the proof of Theorem 1.30(1).
-
(2)
To prove the second part, we need to count the number of such that
if . To do this, we first count the number of such that . Let and . By Corollary 1.28(3), we have that
for with . Similarly, for with . Thus, it suffices to show that is smaller than the total number of -subspaces. Since , we have
which gives our desired result.
-
(3)
Lastly, we prove the third part. By Corollary 1.28(1) and Corollary 1.28(2), we have
Since , we also have that
Thus, this, together with , implies that the number of -dimensional subspaces such that
is at least . From now on, we omit the term and consider that is sufficiently large. We denote the set of these -dimensional subspaces by . Then we have .
Let be the set of such that the number of -dimensional affine planes in such that each contains at least points from is at least . If , then we are done.
Otherwise, by abuse of notation, we can assume that for any , there are at least -dimensional affine planes in such that each contains at most points from . For each , let be the subset of such that each contains at most points from , and . It is clear that .
Using the incidence bound in Theorem 8.1, note that
Since , then one has
This means that there are at least subspaces such that
since is at most which is the total number of all -subspaces.
For each such , let be the number of -dimensional affine planes in such that each contains at least one point from . Note that
This implies that . Since , we have
This completes the proof.
∎
As mentioned in the introduction, the following lemma is on the sharpness of Theorem 1.30(3). This construction is similar to Example 4.1 in [21] in the continuous.
Lemma 8.2.
Assume . For any , there exist sets with and with such that
Proof.
Let be the union of disjoint cosets of . Then . Let .
Let defined by
Set
It is clear that . We now construct arbitrary large such that
We denote the line of the form
by .
Set . We first observe that if with and , then . This gives
For any on the unit circle with , we have
Set Then
Since , we have . Define
Here we identify each point on the unit circle with the line containing it and the origin. So . By a direct computation, one can check that
This completes the proof. ∎
9. Acknowledgements
T. Pham would like to thank the Vietnam Institute for Advanced Study in Mathematics (VIASM) for the hospitality and for the excellent working condition. S. Yoo was supported by the KIAS Individual Grant (CG082701) at Korea Institute for Advanced Study.
References
- [1] C. J. Bishop, and Y. Peres. Fractals in probability and analysis, Vol. 162. Cambridge University Press, 2017.
- [2] J. Chapman, M. B. Erdoğan, D Hart, A. Iosevich, and D. Koh. Pinned distance sets, k-simplices, Wolff’s exponent in finite fields and sum-product estimates, Mathematische Zeitschrift, 271(1-2)(2012): 63–93.
- [3] C. Chen, Projections in vector spaces over finite fields, Annales Academiæ Scientiarum Fennicæ Mathematica, 43(2018), 171–185.
- [4] C. Donoven, and K. Falconer, Codimension formulae for the intersection of fractal subsets of Cantor spaces, Proceedings of the American Mathematical Society, 144(2)(2016), 651–663.
- [5] M. Elekes, T. Keleti, and A. Máthé, Self-similar and self-affine sets: measure of the intersection of two copies, Ergodic Theory and Dynamical Systems, 30(2)(2010), 399–440.
- [6] G. Elekes, and M. Sharir, Incidences in three dimensions and distinct distances in the plane, Combin. Probab. Comput., 20(4)(2011), 571–608.
- [7] S. Eswarathasan, A. Iosevich, and K. Taylor, Intersections of sets and Fourier analysis, Journal d’Analyse Mathématique, 128(2016): 159–178.
- [8] L. Guth, and N. H. Katz. On the Erdős distinct distances problem in the plane, Annals of mathematics (2015): 155–190.
- [9] D. Hart, A. Iosevich, D. Koh, and M. Rudnev, Averages over hyperplanes, sum-product theory in vector spaces over finite fields and the Erdős-Falconer distance conjecture, Transactions of the American Mathematical Society, 363(6)(2011), 3255–3275.
- [10] A. Iosevich, D. Koh, S. Lee, T. Pham, and C. Y. Shen, On restriction estimates for the zero radius sphere over finite fields, Canadian Journal of Mathematics, 73(3)(2021), 769–786.
- [11] A. Iosevich, D. Koh, and F. Rakhmonov, The quotient set of the quadratic distance set over finite fields, arXiv:2301.12021 (2023).
- [12] J.–P. Kahane, Sur la dimension des intersections [On the dimension of intersections] Aspects of Mathematics and its Applications, North-Holland Math. Library, 34(1986), 419–430.
- [13] D. Koh, and H.S. Sun, Distance sets of two subsets of vector spaces over finite fields, Proceedings of the American Mathematical Society, 143(4)(2015): 1679–1692.
- [14] D. Koh, S. Lee, and T. Pham, On the finite field cone restriction conjecture in four dimensions and applications in incidence geometry, International Mathematics Research Notices, 21(2022), 17079–17111.
- [15] R. Lidl and H. Niederreiter, Finite fields, Cambridge University Press, (1997).
- [16] J. M. Marstrand, Some fundamental geometrical properties of plane sets of fractional dimensions, Proceeding of London Mathematical Society, 4(3)(1954), 257–302.
- [17] P. Mattila, Hausdorff dimension and capacities of intersections of sets in n-space, Acta Math., 152(1984), 77–105.
- [18] P. Mattila, Geometry of sets and measures in Euclidean spaces, Cambridge University Press, Volume 44, 1995.
- [19] P. Mattila, Fourier analysis and Hausdorff dimension, Cambridge University Press, Volume 150, 2015.
- [20] P. Mattila, Hausdorff dimension and projections related to intersections, Publicacions Matemàtiques, 66(1)(2022), 305–323.
- [21] P. Mattila, and T. Orponen, Hausdorff dimension, intersections of projections and exceptional plane sections, Proceedings of the American Mathematical Society, 144(8)(2016), 3419–3430.
- [22] C. G. T. de A. Moreira, and J-C. Yoccoz. Stable intersections of regular Cantor sets with large Hausdorff dimensions, Annals of Mathematics (2001): 45–96.
- [23] A. Maschietti, Kakeya sets in finite affine spaces, Journal of Combinatorial Theory Series A, 118(1) (2011): 228–230.
- [24] B. Murphy, G. Petridis, T. Pham, M. Rudnev, and S. Stevens, On the pinned distances problem over finite fields, Journal of London Mathematical Society, 105(1)(2022): 469–499.
- [25] N. D. Phuong, T. Pham, and L. A. Vinh, Incidences between planes over finite fields, Proceedings of the American Mathematical Society, 147(5)(2019): 2185–2196.
- [26] M. Rudnev, On the number of incidences between points and planes in three dimensions, Combinatorica, 38(2018): 219–254.