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

The Kakeya Set conjecture over /N\mathbb{Z}/N\mathbb{Z} for general NN

Manik Dhar
Abstract

We prove the Kakeya set conjecture for /N\mathbb{Z}/N\mathbb{Z} for general NN as stated by Hickman and Wright [15]. This entails extending and combining the techniques of Arsovski [1] for N=pkN=p^{k} and the author and Dvir [6] for the case of square-free NN. We also prove stronger lower bounds for the size of (m,ϵ)(m,\epsilon)-Kakeya sets over /pk{\mathbb{Z}}/p^{k}{\mathbb{Z}} by extending the techniques of [1] using multiplicities as was done in [17, 10]. In addition, we show our bounds are almost sharp by providing a new construction for Kakeya sets over /pk{\mathbb{Z}}/p^{k}{\mathbb{Z}} and /N{\mathbb{Z}}/N{\mathbb{Z}}.

\aicAUTHORdetails

title = The Kakeya Set conjecture over /N\mathbb{Z}/N\mathbb{Z} for general NN, author = Manik Dhar, plaintextauthor = Manik Dhar, plaintexttitle = The Kakeya Set conjecture over Z/N Z for general N, keywords = Kakeya conjecture, polynomial method, \aicEDITORdetailsyear=2024, number=2, received=18 January 2022, revised=4 June 2023, published=26 January 2024, doi=10.19086/aic.2024.2,

[classification=text]

1 Introduction

We are interested in proving lower bounds for the sizes of sets in (/N)n({\mathbb{Z}}/N{\mathbb{Z}})^{n} which have large intersections with lines in many directions. We first define the set of possible directions a line can take in (/N)n({\mathbb{Z}}/N{\mathbb{Z}})^{n}.

Definition 1.1 (Projective space (/N)n1{\mathbb{P}}({\mathbb{Z}}/N{\mathbb{Z}})^{n-1}).

Let N=p1k1prkrN=p_{1}^{k_{1}}\ldots p_{r}^{k_{r}} where p1,,prp_{1},\ldots,p_{r} are distinct primes. The Projective space (/N)n1{\mathbb{P}}({\mathbb{Z}}/N{\mathbb{Z}})^{n-1} consists of vectors u(/N)nu\in({\mathbb{Z}}/N{\mathbb{Z}})^{n} up to unit multiples of each other such that uu (mod pikip_{i}^{k_{i}}) has at least one unit co-ordinate for every i=1,,ri=1,\ldots,r.

For each direction in (/N)n1{\mathbb{P}}({\mathbb{Z}}/N{\mathbb{Z}})^{n-1} we pick a representative in (/N)n({\mathbb{Z}}/N{\mathbb{Z}})^{n}. This allows us to treat (/N)n1{\mathbb{P}}({\mathbb{Z}}/N{\mathbb{Z}})^{n-1} as a subset of (/N)n({\mathbb{Z}}/N{\mathbb{Z}})^{n}.

Definition 1.2 (mm-rich lines).

Let N,nN,n\in{\mathbb{N}}. For a subset S(/N)nS\subseteq({\mathbb{Z}}/N{\mathbb{Z}})^{n}, we say a line L(/N)nL\subseteq({\mathbb{Z}}/N{\mathbb{Z}})^{n} is mm-rich with respect to SS if |SL|m|S\cap L|\geq m.

We now define sets that have large intersections with lines in many directions.

Definition 1.3 ((m,ϵ)(m,\epsilon)-Kakeya Sets).

Let n,Nn,N\in{\mathbb{N}}. A set S(/N)nS\subseteq({\mathbb{Z}}/N{\mathbb{Z}})^{n} is said to be (m,ϵ)(m,\epsilon)-Kakeya if for at least an ϵ\epsilon fraction of directions u(/N)n1u\in{\mathbb{P}}({\mathbb{Z}}/N{\mathbb{Z}})^{n-1} there exists a line Lu={a+λu|λ/N}L_{u}=\{a+\lambda u|\lambda\in{\mathbb{Z}}/N{\mathbb{Z}}\} in the direction uu which is mm-rich with respect to SS.

An (N,1)(N,1)-Kakeya set in (/N)n({\mathbb{Z}}/N{\mathbb{Z}})^{n} is simply called a Kakeya set. In other words, a Kakeya set is a set that contains a line in every direction.

In this paper, we resolve the following conjecture of Hickman and Wright [15].

Conjecture 1.4 (Kakeya set conjecture over /N{\mathbb{Z}}/N{\mathbb{Z}}).

For all ϵ>0\epsilon>0 and nn\in{\mathbb{N}}, there exists a constant Cn,ϵC_{n,\epsilon} such that any Kakeya set S(/N)nS\subset({\mathbb{Z}}/N{\mathbb{Z}})^{n} satisfies

|S|Cn,ϵNnϵ.|S|\geq C_{n,\epsilon}N^{n-\epsilon}.

Wolff in [18] first posed Conjecture 1.4 with /N{\mathbb{Z}}/N{\mathbb{Z}} replaced by a finite field as a possible problem whose resolution might help in proving the Euclidean Kakeya conjecture. Wolff’s conjecture was proven by Dvir in [9] with Cn=1/n!C_{n}=1/n! (Over finite fields of size NN the ϵ\epsilon dependence in Conjecture 1.4 is not needed. For composite NN it is known that the ϵ\epsilon dependence is essential [15, 6].) Using the method of multiplicities and its extensions [17, 10, 3] the constant was improved to Cn=2n+1C_{n}=2^{-n+1}, which is known to be tight.

Ellenberg, Oberlin, and Tao in [11] proposed studying the size of Kakeya sets over the rings /pk{\mathbb{Z}}/p^{k}{\mathbb{Z}} and 𝔽q[x]/xk{\mathbb{F}}_{q}[x]/\langle x^{k}\rangle. They were motivated by the fact that these rings have ‘scales’ and hence are closer to the Euclidean version of the problem. Hickman and Wright posed Conjecture 1.4 for /N{\mathbb{Z}}/N{\mathbb{Z}} with arbitrary NN and considered connections between the problem and the Kakeya conjecture over the pp-adics in [15]. Indeed, resolving Conjecture 1.4 for the rings /pk{\mathbb{Z}}/p^{k}{\mathbb{Z}} and 𝔽q[x]/xk{\mathbb{F}}_{q}[x]/\langle x^{k}\rangle resolves the Minkowski dimension Kakeya set conjecture for the pp-adic integers and the power series ring 𝔽q[[x]]{\mathbb{F}}_{q}[[x]] respectively [11, 8, 15]. The Kakeya problem over these rings is interesting as, similarly to the Euclidean case, one can construct Kakeya sets of Haar measure 0 for the pp-adic integers and the power series ring 𝔽q[[x]]{\mathbb{F}}_{q}[[x]]. These constructions and generalizations to other settings can be found in [8, 12, 4, 15].

For n=2n=2 and N=pkN=p^{k}, Conjecture 1.4 was resolved by Dummit and Hablicsek in [8] by proving that sizes of Kakeya sets are lower bounded by p2k/2kp^{2k}/2k. The author and Dvir [6] resolved Conjecture 1.4 for the case of NN square-free by proving the following bound, which implies Conjecture 1.4 using Observation 1.10 at the end of this section.

Theorem 1.5 (Kakeya set bounds for square-free NN [6, Theorem 1.3]).

Let NN\in{\mathbb{N}} where N=p1prN=p_{1}\ldots p_{r} for distinct primes p1,,prp_{1},\ldots,p_{r}. Any Kakeya set SS in (/N)n({\mathbb{Z}}/N{\mathbb{Z}})^{n} for nn\in{\mathbb{N}} satisfies,

|S|2rnNn.|S|\geq 2^{-rn}N^{n}.

In [6] a reduction was also proven which lower bounds the size of Kakeya sets in (/pk)n({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n} by the 𝔽p{\mathbb{F}}_{p}-rank of the point-hyperplane incidence matrix of (/pk)n({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n}. Building on the ideas behind this reduction, Arsovski proved Conjecture 1.4 for N=pkN=p^{k} for general nn.

Theorem 1.6 (Kakeya Set bounds over /pk{\mathbb{Z}}/p^{k}{\mathbb{Z}},[1, Theorem 2]).

For pp prime and k,nk,n\in{\mathbb{N}}, every Kakeya set SS in (/pk)n({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n} satisfies,

|S|(kn)npkn.|S|\geq(kn)^{-n}p^{kn}.

As mentioned before this also resolves the Minkowski dimension Kakeya Set conjecture over the pp-adics. In [2] (which is the arxiv version 2 of the paper [1]) a different approach inspired by the polynomial method proofs in [9, 17, 10] is used to give the following bound for (m,ϵ)(m,\epsilon)-Kakeya sets.

Theorem 1.7 ((m,ϵ)(m,\epsilon)-Kakeya Set bounds over /pk{\mathbb{Z}}/p^{k}{\mathbb{Z}},[2, Theorem 3]).

For pp prime and k,nk,n\in{\mathbb{N}}, every (m,ϵ)(m,\epsilon)-Kakeya set SS in (/pk)n({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n} satisfies,

|S|(Ckn2p)nϵnmn,|S|\geq(Ckn^{2}p)^{-n}\epsilon^{n}m^{n},

where C>1C>1 is some universal constant.

As can be seen from the bounds in the Theorems 1.6 and 1.7, the analysis in [2] is looser than the analysis in [1] for the case of (pk,1)(p^{k},1)-Kakeya sets and leads to worse bounds in that setting.

In [1, 2] tools from pp-adic analysis are used to develop the techniques which prove Theorems 1.6 and 1.7. Both the proofs can be thought of as trying to develop a polynomial method on roots of unity in {\mathbb{C}}. Our first result extends the ideas in [1] using multiplicities as in [17, 10] to bound the size of (m,ϵ)(m,\epsilon)-Kakeya sets over /pk{\mathbb{Z}}/p^{k}{\mathbb{Z}} with improved constants. Another advantage of our proof is that it is more elementary and does not require tools from pp-adic analysis.

Theorem 1.8 (Stronger (m,ϵ)(m,\epsilon)-Kakeya Set bounds over /pk{\mathbb{Z}}/p^{k}{\mathbb{Z}}***We also note that the techniques presented here can also be used to prove norm bounds for functions f:(/N)nf:({\mathbb{Z}}/N{\mathbb{Z}})^{n}\rightarrow{\mathbb{C}} which have rich lines in many directions as was done for NN prime (and in general for finite fields) in Theorem 19 of [7]. A line LL is mm-rich with respect to ff in this setting if xL|f(x)|m\sum_{x\in L}|f(x)|\geq m. ).

Let k,n,pk,n,p\in{\mathbb{N}} with pp prime. Any (m,ϵ)(m,\epsilon)-Kakeya set S(/pk)nS\subseteq({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n} satisfies the following bound,

|S|ϵ(mn(2(k+logp(n)))n).|S|\geq\epsilon\cdot\left(\frac{m^{n}}{(2(k+\lceil\log_{p}(n)\rceil))^{n}}\right).

When p>np>n, we also get the following stronger bound for (m,ϵ)(m,\epsilon)-Kakeya set SS in (/pk)n({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n},

|S|ϵ(mn(k+1)n)(1+n/p)n.|S|\geq\epsilon\cdot\left(\frac{m^{n}}{(k+1)^{n}}\right)(1+n/p)^{-n}.

We note that for fields of prime size, the bound above recovers the result of [10] as pp tends to infinity. This gives us a new proof for the result with new techniques.

Combining techniques developed for /pk{\mathbb{Z}}/p^{k}{\mathbb{Z}} with techniques from [6], we prove the following lower bound for sizes of Kakeya sets in /N{\mathbb{Z}}/N{\mathbb{Z}} for general NN.

Theorem 1.9 (Kakeya set bounds for /N{\mathbb{Z}}/N{\mathbb{Z}}).

Let nn\in{\mathbb{N}} and R=/NR={\mathbb{Z}}/N{\mathbb{Z}} where N=p1k1prkrN=p_{1}^{k_{1}}\ldots p_{r}^{k_{r}} with distinct primes p1,,prp_{1},\ldots,p_{r} and k1,,krk_{1},\ldots,k_{r}\in{\mathbb{N}}. Any Kakeya set SS in RnR^{n} satisfies,

|S|(i=1r(2(ki+logpi(n)))n)Nn.|S|\geq\left(\prod\limits_{i=1}^{r}(2(k_{i}+\lceil\log_{p_{i}}(n)\rceil))^{-n}\right)\cdot N^{n}.

When p1,,prnp_{1},\ldots,p_{r}\geq n, we also get the following stronger lower bound for the size of a Kakeya set SS in (/N)n,N=p1k1prkr({\mathbb{Z}}/N{\mathbb{Z}})^{n},N=p_{1}^{k_{1}}\ldots p_{r}^{k_{r}},

|S|Nni=1r(ki+1)n(1+n/pi)n.|S|\geq N^{n}\prod\limits_{i=1}^{r}(k_{i}+1)^{-n}(1+n/p_{i})^{-n}.

The bound above recovers Theorem 1.5 as the size of the divisors pip_{i} grows towards infinity.

Note, the techniques here do not naively prove (m,ϵ)(m,\epsilon)-Kakeya bounds over /N{\mathbb{Z}}/N{\mathbb{Z}} for general NN as the techniques in [6] also do not seem to extend to this setting (even for square-free NN). In a follow-up paper [5] we solve this problem by using probabilistic arguments on top of the techniques in this paper and also give Maximal Kakeya bounds over /N{\mathbb{Z}}/N{\mathbb{Z}} for general NN.

Observation 1.10.

The number of prime divisors of NN satisfies r=O(logN/loglogN)r=O(\log N/\log\log N). The expression i=1rki\prod_{i=1}^{r}k_{i} is upper bounded by the number of divisors τ(N)\tau(N) of NN which satisfies log(τ(N))=O(logN/loglogN)\log(\tau(N))=O(\log N/\log\log N). The proof for the bound on rr can be found in [13] and the bound for τ(n)\tau(n) is Theorem 317 in [14]. We now see that the expression in Theorem 1.9 is lower bounded by CnNnO(n/loglogN)C_{n}N^{n-O(n/\log\log N)} and so it indeed proves Conjecture 1.4 for all NN.

Note, we could also get Kakeya Set bounds for /N{\mathbb{Z}}/N{\mathbb{Z}} for general NN by combining the techniques from [1] and [6]. This would also resolve Conjecture 1.4 by proving a Kakeya set lower bound of CnNnO(nlog(n)/loglogN)C_{n}N^{n-O(n\log(n)/\log\log N)} with a worse dependence on the dimension nn.

We also construct Kakeya sets over /pk{\mathbb{Z}}/p^{k}{\mathbb{Z}} showing that the bound in Theorem 1.8 is close to being sharp.

Theorem 1.11 (Small Kakeya sets over /pk{\mathbb{Z}}/p^{k}{\mathbb{Z}}).

Let s,ns,n\in{\mathbb{N}}, p3p\geq 3 be a prime and k=(ps+11)/(p1)k=(p^{s+1}-1)/(p-1). There exists a Kakeya set SS in (/pk)n({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n} such that,

|S|pknkn1(11/p)n.|S|\leq\frac{p^{kn}}{k^{n-1}}(1-1/p)^{-n}.

The construction here uses ideas from the earlier constructions in [15, 12, 4] but is quantitatively stronger.

Using the Chinese remainder theorem we also get a construction for NN with multiple prime factors showing that the bounds in Theorem 1.9 are also close to being sharp.

Corollary 1.12 (Small Kakeya sets over /N{\mathbb{Z}}/N{\mathbb{Z}}).

Let n,s1,,srn,s_{1},\ldots,s_{r}\in{\mathbb{N}}, p1,,pr3p_{1},\ldots,p_{r}\geq 3 be primes, ki=(pisi+11)/(pi1),i=1,,rk_{i}=(p_{i}^{s_{i}+1}-1)/(p_{i}-1),i=1,\ldots,r and N=p1k1prkrN=p_{1}^{k_{1}}\ldots p_{r}^{k_{r}}. There exists a Kakeya set SS in (/N)n({\mathbb{Z}}/N{\mathbb{Z}})^{n} such that,

|S|Nnk1n1krn1i=1r(11/pi)n|S|\leq\frac{N^{n}}{k_{1}^{n-1}\ldots k_{r}^{n-1}}\prod\limits_{i=1}^{r}(1-1/p_{i})^{-n}
Proof.

Using the Chinese Remainder Theorem (see Lemma 5.1) we see that the product of Kakeya sets over /piki{\mathbb{Z}}/p_{i}^{k_{i}}{\mathbb{Z}} for i=1,,ri=1,\ldots,r will be a Kakeya set over /N{\mathbb{Z}}/N{\mathbb{Z}}. We are then done using Theorem 1.11. ∎

1.1 Proof Overview

We first start with a brief overview of the approach introduced in [6]. Given a Kakeya set SS in (/N)n({\mathbb{Z}}/N{\mathbb{Z}})^{n}, we construct a matrix KSK_{S} whose columns are indexed by points in (/N)n({\mathbb{Z}}/N{\mathbb{Z}})^{n} and its rows are indexed by a direction in (/N)n1{\mathbb{P}}({\mathbb{Z}}/N{\mathbb{Z}})^{n-1} where the ddth row would be supported on the line in direction dd contained in SS. This ensures that the non-zero columns of KSK_{S} correspond to points in SS, which implies that the rank of KSK_{S} then lower bounds the size of SS. The goal is to construct a suitable KSK_{S} and find a matrix EE such that KSEK_{S}E is a matrix independent of SS whose rank can be analyzed easily. For prime N=pN=p, EE can be a matrix whose columns contain the evaluation of a monomial on 𝔽pn{\mathbb{F}}_{p}^{n}. In this case KSK_{S} ends up being a ‘decoder’ matrix where each row outputs the evaluation of a monomial on a direction given its evaluations on a line in that direction. This turns out to be a reformulation of Dvir’s polynomial method proof [9]. When N=pqN=pq (or in general is square-free) as /N𝔽p×𝔽q{\mathbb{Z}}/N{\mathbb{Z}}\cong{\mathbb{F}}_{p}\times{\mathbb{F}}_{q} we define EE as a tensor of matrices one acting on the 𝔽p{\mathbb{F}}_{p} part and the other on the 𝔽q{\mathbb{F}}_{q} part. This allows for an inductive argument to give Kakeya set lower bounds for square-free NN. To get quantitatively stronger bounds we can use the evaluations of (Hasse) derivatives and multiplicities as was done in [17, 10].

In [1] for prime power NN points in 𝔽pn{\mathbb{F}}_{p}^{n} are embedded in pn{\mathbb{C}}_{p}^{n} (the pp-adic complex numbers which is isomorphic to {\mathbb{C}} as a field) using pkp^{k}th roots of unity. The proof then implements the strategy of [6] using matrices with polynomial entries. UU is a matrix whose columns are the evaluations of monomials over the embedding of (/pk)n({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n} in the torus. Let LL be a line in direction dd. The key statement is that some linear combination (with polynomial coefficients) of the rows in UU corresponding to points in LL can generate the ddth row of a ‘Vandermonde’ matrix MM after applying a mod pp operation (using the pp-adic structure). The rank of MM can be easily lower bounded. The linear combination being taken here corresponds to a row of KSK_{S} for a Kakeya set S(/pk)nS\subseteq({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n} containing LL in direction dd. The key statement can be reformulated as saying that

KSUmodp=MK_{S}U\mod p=M (1)

The argument can be completed by saying that the non-zero columns of SS correspond to points in SS, alternatively we note that we only use rows in EE which correspond to points in SS. This shows the rank of MM over 𝔽p{\mathbb{F}}_{p} lower bounds the size of SS. The argument in [1] is only for (pk,1)(p^{k},1)-Kakeya sets. This is because, apriori it is not clear how to define a suitable KSK_{S} (equivalently how to decode from lines with only some points contained in SS). The revised version of the paper [2] has a very different approach to give (m,ϵ)(m,\epsilon)-Kakeya bounds for the prime power case (the bounds there are quantitatively weaker than the ones here).

We now discuss how we improve the prime power bound quantitatively. Using simple linear algebraic arguments and the Chinese remainder theorem for a suitable polynomial ring we show that given mm points on a line LL in direction dd we can decode the ddth row of MM up to degree mm (see Corollary 3.16). We use multiplicities and (Hasse) derivatives to get stronger quantitative bounds (which Theorem 1.11 shows are almost sharp). We also develop these ideas without using the theory of pp-adic numbers. We make this precise in Section 3.

To get bounds for general NN we follow an inductive style argument as was done in [6] with some small technical improvements to get better constants.

1.2 Organization of the paper

In Section 2 we state definitions and results we need from [6]. In Section 3 we state and extend results from [1]. In Section 4 we prove Theorem 1.8. In Section 5 we prove Theorem 1.9. Finally, in section 6 we prove Theorem 1.11.

1.3 Acknowledgements

The author would like to thank Zeev Dvir for helpful comments and discussion. The author would also like to thank the reviewers for their helpful suggestions. This work was done while the author was supported by the NSF grant DMS-1953807.

2 Rank and crank of matrices with polynomial entries

Definition 2.1 (Rank of matrices with entries in 𝔽[z]/f(z)\mathbb{F}[z]/\langle f(z)\rangle).

Given a field 𝔽{\mathbb{F}} and a matrix MM with entries in 𝔽[z]/f(z){\mathbb{F}}[z]/\langle f(z)\rangle where f(z)f(z) is a non-constant polynomial in 𝔽[z]{\mathbb{F}}[z], we define the 𝔽{\mathbb{F}}-rank of MM denoted rank𝔽M\textsf{rank}_{\mathbb{F}}M as the maximum number of 𝔽{\mathbb{F}}-linearly independent columns of MM.

In our proof, it will also be convenient to work with the following extension of rank for sets of matrices.

Definition 2.2 (crank of a set of matrices).

Let 𝔽{\mathbb{F}} be a field and f(z)f(z) a non-constant polynomial in 𝔽[z]{\mathbb{F}}[z]. Given a finite set T={A1,,An}T=\{A_{1},\ldots,A_{n}\} of matrices over the ring 𝔽[z]/f(z){\mathbb{F}}[z]/\langle f(z)\rangle having the same number of columns we let crank𝔽(T)\textsf{crank}_{\mathbb{F}}(T) be the 𝔽{\mathbb{F}}-rank of the matrix obtained by concatenating all the elements AiA_{i} in TT along their columns. In [6] crank was defined using the rows of the matrices. The two definitions are equivalent when the entries of the matrices are from the field 𝔽{\mathbb{F}} following from simple linear algebra. In our setting, the column and row definitions are not necessarily identical and the column version is better for our purposes.

We will use a simple lemma which follows from the definition of crank.

Lemma 2.3 (crank bound for multiplying matrices).

Let 𝔽{\mathbb{F}} be a field and f(z)f(z) a non-constant polynomial in 𝔽[z]{\mathbb{F}}[z]. Given matrices A1,,AnA_{1},\ldots,A_{n} of size a×ba\times b and matrices H1,,HnH_{1},\ldots,H_{n} of size c×ac\times a with entries in 𝔽[x]/f(z){\mathbb{F}}[x]/\langle f(z)\rangle we have

crank𝔽{Ai}i=1ncrank𝔽{HiAi}i=1n.\textsf{crank}_{\mathbb{F}}\{A_{i}\}_{i=1}^{n}\geq\textsf{crank}_{\mathbb{F}}\{H_{i}\cdot A_{i}\}_{i=1}^{n}.
Proof.

A dependence on the columns of the matrix obtained by concatenating AiA_{i}s will be represented by a non-zero vector w𝔽bw\in{\mathbb{F}}^{b} such that Aiw=0A_{i}w=0 for all i=1,,ni=1,\ldots,n. This would imply (HiAi)w=0(H_{i}\cdot A_{i})w=0 for all ii as well, completing the proof. ∎

Given a matrix AA with entries in 𝔽[z]/f(z){\mathbb{F}}[z]/\langle f(z)\rangle we want to construct a new matrix with entries only in 𝔽{\mathbb{F}} such that their 𝔽{\mathbb{F}}-ranks are the same. First, we state a simple fact about 𝔽[z]/f(z){\mathbb{F}}[z]/\langle f(z)\rangle.

Lemma 2.4 (Unique representation of elements in 𝔽[z]/f(z){\mathbb{F}}[z\text{]}/\langle f(z)\rangle).

Let 𝔽{\mathbb{F}} be a field and f(z)f(z) a non-constant polynomial in 𝔽[z]{\mathbb{F}}[z] of degree d>0d>0. Every element in 𝔽[z]/f(z){\mathbb{F}}[z]/\langle f(z)\rangle is uniquely represented by a polynomial in 𝔽[z]{\mathbb{F}}[z] with degree strictly less than dd and conversely every degree strictly less than dd polynomial in 𝔽[z]{\mathbb{F}}[z] is a unique element in 𝔽[z]/f(z){\mathbb{F}}[z]/\langle f(z)\rangle. When we refer to an element h(z)𝔽[z]/f(z)h(z)\in{\mathbb{F}}[z]/\langle f(z)\rangle we also let it refer to the unique degree strictly less than dd polynomial it equals.

Definition 2.5 (Coefficient matrix of AA).

Let 𝔽{\mathbb{F}} be a field and f(z)f(z) a non-constant polynomial in 𝔽[z]{\mathbb{F}}[z] of degree d>0d>0. Given any matrix AA of size n1×n2n_{1}\times n_{2} with entries in 𝔽[z]/f(z){\mathbb{F}}[z]/\langle f(z)\rangle we can construct the coefficient matrix of AA denoted by Coeff(A)\text{Coeff}(A) with entries in 𝔽{\mathbb{F}} which will be of size dn1×n2dn_{1}\times n_{2} whose rows are labeled by elements in {0,,d1}×[n1]\{0,\ldots,d-1\}\times[n_{1}] such that its (i,j)(i,j)’th row is formed by the coefficients of ziz^{i} of the polynomial entries of the jj’th row of AA.

The key property of the coefficient matrix immediately follows from its definition.

Lemma 2.6.

Let 𝔽{\mathbb{F}} be a field and f(z)f(z) a non-constant polynomial in 𝔽[z]{\mathbb{F}}[z] of degree d>0d>0. Given any matrix AA with entries in 𝔽[z]/f(z){\mathbb{F}}[z]/\langle f(z)\rangle and its coefficient matrix Coeff(A)\text{Coeff}(A) it is the case that an 𝔽{\mathbb{F}}-linear combination of a subset of columns of AA is 0 if and only if the corresponding 𝔽{\mathbb{F}}-linear combination of the same subset of columns of Coeff(A)\text{Coeff}(A) is also 0.

In particular, the 𝔽{\mathbb{F}}-rank of AA equals the 𝔽{\mathbb{F}}-rank of Coeff(A)\text{Coeff}(A).

We now need some simple properties related to the crank of tensor products. To that end, we first define the tensor/Kronecker product of matrices. Let [r]={1,,r}[r]=\{1,\ldots,r\}.

Definition 2.7 (Kronecker Product of two matrices).

Given a commutative ring RR and two matrices MAM_{A} and MBM_{B} of sizes n1×m1n_{1}\times m_{1} and n2×m2n_{2}\times m_{2} corresponding to RR-linear maps A:Rn1Rm1A:R^{n_{1}}\rightarrow R^{m_{1}} and B:Rn2Rm2B:R^{n_{2}}\rightarrow R^{m_{2}} respectively, we define the Kronecker product MAMBM_{A}\otimes M_{B} as a matrix of size n1n2×m1m2n_{1}n_{2}\times m_{1}m_{2} with its rows indexed by elements in [n1]×[n2][n_{1}]\times[n_{2}] and its columns indexed by elements in [m1]×[m2][m_{1}]\times[m_{2}] such that

MAMB((r1,r2),(c1,c2))=MA(r1,c1)MB(r2,c2),M_{A}\otimes M_{B}((r_{1},r_{2}),(c_{1},c_{2}))=M_{A}(r_{1},c_{1})M_{B}(r_{2},c_{2}),

where r1[n1],r2[n2],c1[m1]r_{1}\in[n_{1}],r_{2}\in[n_{2}],c_{1}\in[m_{1}] and c2[m2]c_{2}\in[m_{2}]. MAMBM_{A}\otimes M_{B} corresponds to the matrix of the RR-linear map AB:Rn1Rn2Rn1n2Rm1Rm2Rm1m2A\otimes B:R^{n_{1}}\otimes R^{n_{2}}\cong R^{n_{1}n_{2}}\rightarrow R^{m_{1}}\otimes R^{m_{2}}\cong R^{m_{1}m_{2}}.

We will need the following simple property of Kronecker products which follows from the corresponding property of the tensor product of linear maps.

Lemma 2.8 (Multiplication of Kronecker products).

Given matrices A1,A2,B1A_{1},A_{2},B_{1} and B2B_{2} of sizes a1×n1a_{1}\times n_{1}, a2×n2a_{2}\times n_{2}, n1×b1n_{1}\times b_{1} and n2×b2n_{2}\times b_{2} we have the following identity,

(A1A2)(B1B2)=(A1B1)(A2B2).(A_{1}\otimes A_{2})\cdot(B_{1}\otimes B_{2})=(A_{1}\cdot B_{1})\otimes(A_{2}\cdot B_{2}).

We want to prove a crank bound for a family of matrices with a tensor product structure. This statement is an analog of Lemma 4.8 in [6] for our setting. Indeed we will prove it using Lemma 4.8 in [6] which we now state.

Lemma 2.9 (Lemma 4.8 in [6]).

Given matrices A1,,AnA_{1},\ldots,A_{n} of size a1×a2a_{1}\times a_{2} over a field 𝔽{\mathbb{F}} such that

crank{Ai}i=1nr1\textsf{crank}\{A_{i}\}_{i=1}^{n}\geq r_{1}

and matrices Bi,jB_{i,j} over the field 𝔽{\mathbb{F}} for i[n]i\in[n] and j[m]j\in[m] of size b1×b2b_{1}\times b_{2} such that

crank{Bi,j}j=1mr2\textsf{crank}\{B_{i,j}\}_{j=1}^{m}\geq r_{2}

for all i[n]i\in[n] we have,

crank{AiBi,j|i[n],j[m]}r1r2.\textsf{crank}\{A_{i}\otimes B_{i,j}|i\in[n],j\in[m]\}\geq r_{1}r_{2}.

The above lemma follows easily from simple properties of the tensor product and a proof can be found in [6]. We now prove a generalization for our setting.

Lemma 2.10 (crank bound for tensor products).

Let 𝔽{\mathbb{F}} be a field and f(z)𝔽[z]f(z)\in{\mathbb{F}}[z] a non-constant polynomial. Given matrices A1,,AnA_{1},\ldots,A_{n} of size a1×a2a_{1}\times a_{2} over the ring 𝔽[z]/f(z){\mathbb{F}}[z]/\langle f(z)\rangle such that

crank𝔽{Ai}i=1nr1\textsf{crank}_{\mathbb{F}}\{A_{i}\}_{i=1}^{n}\geq r_{1}

and matrices Bi,jB_{i,j} over the field 𝔽{\mathbb{F}} for i[n]i\in[n] and j[m]j\in[m] of size b1×b2b_{1}\times b_{2} such that

crank𝔽{Bi,j}j=1mr2\textsf{crank}_{\mathbb{F}}\{B_{i,j}\}_{j=1}^{m}\geq r_{2}

for all i[n]i\in[n] we have,

crank𝔽{AiBi,j|i[n],j[m]}r1r2.\textsf{crank}_{\mathbb{F}}\{A_{i}\otimes B_{i,j}|i\in[n],j\in[m]\}\geq r_{1}r_{2}.

Note the asymmetry between the matrices AiA_{i} having entries in 𝔽[z]/f(z){\mathbb{F}}[z]/\langle f(z)\rangle and the matrices Bi,jB_{i,j} only having entries in 𝔽{\mathbb{F}}. This is important for the proof to work.

Proof.

Consider the coefficient matrices Coeff(Ai)\text{Coeff}(A_{i}). Using Lemma 2.6 it follows that

crank𝔽{Ai}i=1n=crank𝔽{Coeff(Ai)}i=1n.\displaystyle\textsf{crank}_{\mathbb{F}}\{A_{i}\}_{i=1}^{n}=\textsf{crank}_{\mathbb{F}}\{\text{Coeff}(A_{i})\}_{i=1}^{n}. (2)

As Bi,jB_{i,j} only has entries in 𝔽{\mathbb{F}} it is easy to see that Coeff(Ai)Bi,j=Coeff(AiBi,j)\text{Coeff}(A_{i})\otimes B_{i,j}=\text{Coeff}(A_{i}\otimes B_{i,j}) (Note this would not be true if Bi,jB_{i,j} had entries in the ring 𝔽[z]/f(z){\mathbb{F}}[z]/\langle f(z)\rangle). Using Lemma 2.6 now gives us,

crank𝔽{AiBi,j|i[n],j[m]}=crank𝔽{Coeff(Ai)Bi,j|i[n],j[m]}.\displaystyle\textsf{crank}_{\mathbb{F}}\{A_{i}\otimes B_{i,j}|i\in[n],j\in[m]\}=\textsf{crank}_{\mathbb{F}}\{\text{Coeff}(A_{i})\otimes B_{i,j}|i\in[n],j\in[m]\}. (3)

We apply Lemma 2.9 on the family Coeff(Ai)Bi,j,i[n],j[m]\text{Coeff}(A_{i})\otimes B_{i,j},i\in[n],j\in[m] and use equations (2) and (3) to complete the proof. ∎

3 Polynomial Method on the complex torus

As mentioned in the introduction, the techniques presented here are an extension of the techniques used in [1] and are developed without invoking the tools and the language of pp-adic analysis.

We embed (/pk)n({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n} in the complex torus. As mentioned in the proof overview (Section 1.1) we will be working with matrices with polynomial entries. We first define the rings where the entries come from.

Definition 3.1 (Rings T¯\overline{T}_{\ell} and TkT_{\ell}^{k}).

Let

T¯=𝔽p[z]/zp1\overline{T}_{\ell}={\mathbb{F}}_{p}[z]/\langle z^{p^{\ell}}-1\rangle

and

Tk=(ζpk)[z]/zp1T_{\ell}^{k}={\mathbb{Z}}(\zeta_{p^{k}})[z]/\langle z^{p^{\ell}}-1\rangle

where ζpk\zeta_{p^{k}} is a primitive complex pkp^{k}’th root of unity.

As stated in the proof overview (Section 1.1) the goal is decode the ddth row of a ‘Vandermonde’ matrix starting from the evaluations of monomials on a line LL in direction dd. We set \ell according to the number of points in LL and to what order derivatives we are working with and it controls the size of our ‘Vandermonde’ matrix as well. If we were not using any multiplicities/derivatives then \ell could be set to logp(m)\log_{p}(m) to lower bound the size of (m,ϵ)(m,\epsilon)-Kakeya sets with weaker constants.

We suppress pp in the notation as it will be fixed to a single value throughout our proofs. We also let ζ=ζpk\zeta=\zeta_{p^{k}} throughout this section for ease of notation.

Note that (ζ){\mathbb{Z}}(\zeta) is the ring [x]/ϕpk(x){\mathbb{Z}}[x]/\langle\phi_{p^{k}}(x)\rangle where,

ϕpk(x)=xpk1xpk11=i=0p1xipk1,\displaystyle\phi_{p^{k}}(x)=\frac{x^{p^{k}}-1}{x^{p^{k-1}}-1}=\sum\limits_{i=0}^{p-1}x^{ip^{k-1}}, (4)

is the pkp^{k} Cyclotomic polynomial. We will also work with the field (ζ)=[x]/ϕpk(x){\mathbb{Q}}(\zeta)={\mathbb{Q}}[x]/\langle\phi_{p^{k}}(x)\rangle.

We need a simple lemma connecting (ζ){\mathbb{Z}}(\zeta) to 𝔽p{\mathbb{F}}_{p}.

Lemma 3.2 (Quotient map ψpk\psi_{p^{k}} from (ζ){\mathbb{Z}}(\zeta) to 𝔽p{\mathbb{F}}_{p}).

The field 𝔽p{\mathbb{F}}_{p} is isomorphic to (ζ)/p,ζ1{\mathbb{Z}}(\zeta)/\langle p,\zeta-1\rangle. In particular, the map ψpk\psi_{p^{k}} from (ζ){\mathbb{Z}}(\zeta) to 𝔽p{\mathbb{F}}_{p} which maps {\mathbb{Z}} to 𝔽p{\mathbb{F}}_{p} via the mod pp map and ζ\zeta to 11 is a ring homomorphism.

Proof.

ζ\zeta is a root of the Cyclotomic polynomial ϕpk\phi_{p^{k}} defined in (4). As mentioned earlier, this allows us to write (ζ){\mathbb{Z}}(\zeta) as [ζ]/ϕpk(ζ){\mathbb{Z}}[\zeta]/\langle\phi_{p^{k}}(\zeta)\rangle. Notice that using (4) we have ϕpk(1) (mod p)=0\phi_{p^{k}}(1)\text{ (mod }p\text{)}=0 which implies that ϕpk(ζ)\phi_{p^{k}}(\zeta) is divisible by ζ1\zeta-1 over 𝔽p{\mathbb{F}}_{p}.

If we quotient the ring (ζ){\mathbb{Z}}(\zeta) by p,ζ1\langle p,\zeta-1\rangle we get

[ζ]ζ1,p,ϕpk(ζ)=𝔽p[ζ]ζ1,ϕpk(ζ)=𝔽p[ζ]ζ1=𝔽p.\frac{{\mathbb{Z}}[\zeta]}{\langle\zeta-1,p,\phi_{p^{k}}(\zeta)\rangle}=\frac{{\mathbb{F}}_{p}[\zeta]}{\langle\zeta-1,\phi_{p^{k}}(\zeta)\rangle}=\frac{{\mathbb{F}}_{p}[\zeta]}{\langle\zeta-1\rangle}={\mathbb{F}}_{p}.

Therefore, the map ψpk\psi_{p^{k}} is the map which quotients the ring (ζ){\mathbb{Z}}(\zeta) by the ideal p,ζ1\langle p,\zeta-1\rangle. ∎

We note ψpk\psi_{p^{k}} can be extended to the rings (ζ)[z]/h(z){\mathbb{Z}}(\zeta)[z]/\langle h(z)\rangle for any h(x)(ζ)[x]h(x)\in{\mathbb{Z}}(\zeta)[x] by mapping zz to zz. The proof above immediately generalizes to the following corollary.

Corollary 3.3 (Extending ψpk\psi_{p^{k}}).

Let h(x)(ζ)[x]h(x)\in{\mathbb{Z}}(\zeta)[x]. ψpk\psi_{p^{k}} is a ring homomorphism from (ζ)[z]/h(z){\mathbb{Z}}(\zeta)[z]/\langle h(z)\rangle to (ζ)[z]/h(z),p,ζ1=𝔽p[z]/ψpk(h(z)){\mathbb{Z}}(\zeta)[z]/\langle h(z),p,\zeta-1\rangle={\mathbb{F}}_{p}[z]/\langle\psi_{p^{k}}(h(z))\rangle.

Tk/p,ζ1T_{\ell}^{k}/\langle p,\zeta-1\rangle being isomorphic to T¯\overline{T}_{\ell} is a special case of the corollary. The ψpk\psi_{p^{k}} will correspond to the mod pp operation from the overview.

The operation ψpk\psi_{p^{k}} corresponds to the mod pp operation in (1) in the proof overview (Section 1.1).

We now prove that the rank of a matrix can only decrease under the quotient map ψpk\psi_{p^{k}}.

Lemma 3.4.

If AA is a matrix with entries in TkT_{\ell}^{k}, then we have the following bound,

rank(ζ)Arank𝔽pψpk(A),\textsf{rank}_{{\mathbb{Q}}(\zeta)}A\geq\textsf{rank}_{{\mathbb{F}}_{p}}\psi_{p^{k}}(A),

where ψpk(A)\psi_{p^{k}}(A) is the matrix with entries in T¯\overline{T}_{\ell} obtained by applying ψpk\psi_{p^{k}} to each entry of AA.

In [1] the above lemma is implicit in the proof of Proposition 4, Arsovski’s proof uses tools from pp-adic analysis. We provide an alternate proof without using those techniques.

Proof of Lemma 3.4.

Let AA be a matrix with entries in Tk=(ζ)[z]/zp1T_{\ell}^{k}={\mathbb{Z}}(\zeta)[z]/\langle z^{p^{\ell}}-1\rangle. TkT^{k}_{\ell} is a sub-ring of the ring (ζ)[z]/zp1{\mathbb{Q}}(\zeta)[z]/\langle z^{p^{\ell}}-1\rangle. As (ζ){\mathbb{Q}}(\zeta) is a field we can define Coeff(A)\text{Coeff}(A) with entries in (ζ){\mathbb{Q}}(\zeta).

This also means ψpk\psi_{p^{k}} can be applied on Coeff(A)\text{Coeff}(A). On the other hand ψpk\psi_{p^{k}} can be directly applied on AA as it has entries in TkT_{\ell}^{k} and ψpk(A)\psi_{p^{k}}(A) will have entries in T¯=𝔽p[z]/z1\overline{T}_{\ell}={\mathbb{F}}_{p}[z]/\langle z^{\ell}-1\rangle. This means we can define Coeff(ψpk(A))\text{Coeff}(\psi_{p^{k}}(A)) with entries in 𝔽p{\mathbb{F}}_{p}. Notice that ψpk(Coeff(A))=Coeff(ψpk(A))\psi_{p^{k}}(\text{Coeff}(A))=\text{Coeff}(\psi_{p^{k}}(A)).

Let rank(ζ)A=r\textsf{rank}_{{\mathbb{Q}}(\zeta)}A=r. By Lemma 2.6 and simple properties of the determinant, we have that every r+1×r+1r+1\times r+1 sub-matrix MM of Coeff(A)\text{Coeff}(A) has 0 determinant. As ψpk\psi_{p^{k}} is a ring homomorphism it immediately follows that the corresponding r+1×r+1r+1\times r+1 sub-matrix ψpk(M)\psi_{p^{k}}(M) in ψpk(Coeff(A))=Coeff(ψpk(A))\psi_{p^{k}}(\text{Coeff}(A))=\text{Coeff}(\psi_{p^{k}}(A)) will have 0 determinant too.

This implies that Coeff(ψpk(A))\text{Coeff}(\psi_{p^{k}}(A)) has rank at most rr. The statement now follows from applying Lemma 2.6 on ψpk(A)\psi_{p^{k}}(A). ∎

We now define the ‘Vandermonde’ matrix (corresponding to MM in Section 1.1) with entries in T¯\overline{T}_{\ell} which was defined by Arsovski in [1] and whose 𝔽p{\mathbb{F}}_{p}-rank will help us lower bound the size of Kakeya Sets.

Definition 3.5 (The matrix Mp,nM_{p^{\ell},n}).

The matrix Mp,nM_{p^{\ell},n} is a matrix over T¯\overline{T}_{\ell} with its rows and columns indexed by points in (/p)n({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n}. The (u,v)(/p)n×(/p)n(u,v)\in({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n}\times({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n} entry is

Mp,n(u,v)=zu,v.M_{p^{\ell},n}(u,v)=z^{\langle u,v\rangle}.
Lemma 3.6 (Rank of Mp,nM_{p^{\ell},n}).

The 𝔽p{\mathbb{F}}_{p}-rank of Mp,nM_{p^{\ell},n} is at least

rank𝔽pMp,n(p1+nn).\text{rank}_{{\mathbb{F}}_{p}}M_{p^{\ell},n}\geq\binom{\lceil p^{\ell}\ell^{-1}\rceil+n}{n}.

A lower bound of pn(n)np^{\ell n}(\ell n)^{-n} is proven within Lemma 5 and Theorem 2 in [1]. The same argument can also give the bound above with a slightly more careful analysis. For completeness, we give a proof in Appendix A. For convenience, we now prove that removing the rows in Mp,nM_{p^{\ell},n} corresponding to elements u(/p)n(/p)n1u\in({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n}\setminus{\mathbb{P}}({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n-1} does not change the 𝔽p{\mathbb{F}}_{p}-rank of the matrix.

For a given set of elements V(/p)nV\subseteq({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n} let Mp,n(V)M_{p^{\ell},n}(V) refer to the sub-matrix obtained by restricting to rows of Mp,nM_{p^{\ell},n} corresponding to elements in VV. In particular, for any given uRnu\in R^{n} we let Mpk,n(u)M_{p^{k},n}(u) refer to the uu’th row of the matrix.

Lemma 3.7.
rank𝔽pMp,n((/p)n1)=rank𝔽pMp,n.\text{rank}_{{\mathbb{F}}_{p}}M_{p^{\ell},n}({\mathbb{P}}({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n-1})=\text{rank}_{{\mathbb{F}}_{p}}M_{p^{\ell},n}.
Proof.

Consider the matrix Coeff(Mp,n)\text{Coeff}(M_{p^{\ell},n}). For any u(/p)nu\in({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n}, the uu’th row in Mp,nM_{p^{\ell},n} will correspond to a pp^{\ell} block of rows BuB_{u} in Coeff(Mp,n)\text{Coeff}(M_{p^{\ell},n}). Say uu doesn’t have a unit coordinate. We can find an element u(/p)n1u^{\prime}\in{\mathbb{P}}({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n-1} such that for some ii, piu=up^{i}u^{\prime}=u. We claim that the block BuB_{u} can be generated by the block BuB_{u^{\prime}} via a linear map. This follows because given the coefficient vector of a polynomial f(z)f(z) in T¯\overline{T}_{\ell} the coefficient vector of the polynomial f(zp)T¯f(z^{p})\in\overline{T}_{\ell} can be obtained via a 𝔽p{\mathbb{F}}_{p}-linear map. This shows that the span of the rows of Coeff(Mp,n)\text{Coeff}(M_{p^{\ell},n}) is identical to the span of the rows in Coeff(Mp,n(V)\text{Coeff}(M_{p^{\ell},n}(V) where VV is the set of vectors in (/pk)n({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n} with at least one unit co-ordinate.

Now for any element uVu\in V we can find an element u(/p)n1u^{\prime}\in{\mathbb{P}}({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n-1} such that there exists a λ(/pk)×\lambda\in({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{\times} for which u=λuu=\lambda u^{\prime}. We now note that the block BuB_{u} can be generated by the block BuB_{u^{\prime}} via a linear map as the coefficient vector of a polynomial f(zλ)T¯f(z^{\lambda})\in\overline{T}_{\ell} can be linearly computed from f(z)T¯f(z)\in\overline{T}_{\ell}. This shows that the span of the rows of Coeff(Mp,n(V))\text{Coeff}(M_{p^{\ell},n}(V)) is identical to the span of the rows in Coeff(Mp,n((/pk)n1)\text{Coeff}(M_{p^{\ell},n}({\mathbb{P}}({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n-1}) completing the proof using Lemma 2.6. ∎

In [1] a row vector which encodes the evaluation of monomials over points in a line in direction u(/pk)n1u\in{\mathbb{P}}({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n-1} is used to decode the uu’th row of Mpk,nM_{p^{k},n}. We extend this method by using row vectors which encode the evaluations of Hasse derivative of a monomial over a given point. We first define Hasse derivative.

Definition 3.8 (Hasse Derivatives).

Given a polynomial f𝔽[x1,,xn]f\in{\mathbb{F}}[x_{1},\ldots,x_{n}] for any field 𝔽{\mathbb{F}} and an 𝛂0n{\bm{\alpha}}\in{\mathbb{Z}}_{\geq 0}^{n} the 𝛂{\bm{\alpha}}’th Hasse derivative of ff is the polynomial f(𝛂)f^{({\bm{\alpha}})} in the expansion f(x+z)=𝛃0nf(𝛃)(x)z𝛃f(x+z)=\sum_{{\bm{\beta}}\in{\mathbb{Z}}_{\geq 0}^{n}}f^{({\bm{\beta}})}(x)z^{{\bm{\beta}}} where x=(x1,,xn)x=(x_{1},...,x_{n}), z=(z1,,zn)z=(z_{1},...,z_{n}) and z𝛃=k=1nzkβkz^{{\bm{\beta}}}=\prod_{k=1}^{n}z_{k}^{\beta_{k}}.

Definition 3.9 (The evaluation vector Ud(𝜶)(y)U_{d}^{({\bm{\alpha}})}(y)).

Let 𝔽{\mathbb{F}} be a field, n,dn,d\in{\mathbb{Z}} and 𝛂0n{\bm{\alpha}}\in{\mathbb{Z}}_{\geq 0}^{n}. For any given point y𝔽ny\in{\mathbb{F}}^{n} we define Ud(𝛂)(y)U_{d}^{({\bm{\alpha}})}(y) to be a row vector of size dnd^{n} whose columns are indexed by monomials m=x1j1x2j2xnjn𝔽[x1,,xn]m=x_{1}^{j_{1}}x_{2}^{j_{2}}\ldots x_{n}^{j_{n}}\in{\mathbb{F}}[x_{1},\ldots,x_{n}] for jk{0,,d1},k[n]j_{k}\in\{0,\ldots,d-1\},k\in[n] such that its mm’th column is m(𝛂)(y)m^{({\bm{\alpha}})}(y).

We suppress nn in the notation as there will be no ambiguity over it in this paper. We define the weight of 𝜶0n{\bm{\alpha}}\in{\mathbb{Z}}_{\geq 0}^{n} as wt(𝜶)=k=1nαk\text{wt}({\bm{\alpha}})=\sum_{k=1}^{n}\alpha_{k}. The following simple fact about uni-variate polynomials illustrates one use of Hasse derivatives and how they correspond to our intuition with regular derivatives in fields of characteristic 0.

Lemma 3.10.

Let 𝔽{\mathbb{F}} be a field, y𝔽y\in{\mathbb{F}} and ww\in{\mathbb{N}}. For any uni-variate polynomial f𝔽[x]f\in{\mathbb{F}}[x] all its Hasse derivatives at the point yy of weight strictly less than ww are 0 if and only if f(xy)wf\in\langle(x-y)^{w}\rangle or in other words ff is divisible by (xy)w(x-y)^{w}.

Let L(/pk)nL\subseteq({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n} be a line in direction u(/pk)n1u\in{\mathbb{P}}({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n-1}. A special case of the next lemma proves that, for any polynomial f[x1,,xn]f\in{\mathbb{Z}}[x_{1},\ldots,x_{n}], we can evaluate f(zu)T¯k=𝔽p[z]/zpk1,zu=(zu1,,zun)f(z^{u})\in\overline{T}_{k}={\mathbb{F}}_{p}[z]/\langle z^{p^{k}}-1\rangle,z^{u}=(z^{u_{1}},\ldots,z^{u_{n}}) from the evaluation of ff on the points f(ζx),xLf(\zeta^{x}),x\in L. This is implicit in the proof of Proposition 4 in [1]. We generalize this statement. Let π:L0\pi:L\rightarrow{\mathbb{Z}}_{\geq 0} be a function on the line such that xLπ(x)p\sum_{x\in L}\pi(x)\geq p^{\ell}. We prove that we can decode f(zu)T¯=𝔽p[z]/zp1f(z^{u^{\prime}})\in\overline{T}_{\ell}={\mathbb{F}}_{p}[z]/\langle z^{p^{\ell}}-1\rangle from the evaluations of weight at most π(x)\pi(x) Hasse derivatives of ff at ζx,xL\zeta^{x},x\in L for any uu^{\prime} in (/p)n({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n} such that u (mod pk)=uu^{\prime}\text{ (mod }p^{k}\text{)}=u.

The lemma below can be thought of as analogous to how over finite fields the evaluation of a polynomial and its Hasse derivatives with high enough weight along a line in direction uu can be used to decode the evaluation of that polynomial at the point at infinity along uu [10].

Lemma 3.11 (Decoding from evaluations on rich lines).

Let L={a+λu|λ/pk}(/pk)nL=\{a+\lambda u|\lambda\in{\mathbb{Z}}/p^{k}{\mathbb{Z}}\}\subset({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n} with a(/pk)n,u(/pk)n1a\in({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n},u\in{\mathbb{P}}({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n-1}, u(/p)nu^{\prime}\in({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n} be such that u (mod pk)=uu^{\prime}\text{ (mod }p^{k}\text{)}=u and π:L0\pi:L\rightarrow{\mathbb{Z}}_{\geq 0} be a function which satisfies xLπ(x)p\sum_{x\in L}\pi(x)\geq p^{\ell}.

Then, there exist elements cλ,𝛂(ζ)[z]c_{\lambda,{\bm{\alpha}}}\in{\mathbb{Q}}(\zeta)[z] (depending on π,L\pi,L and uu^{\prime}) for λ/pk\lambda\in{\mathbb{Z}}/p^{k}{\mathbb{Z}} and 𝛂0n{\bm{\alpha}}\in{\mathbb{Z}}_{\geq 0}^{n} with wt(𝛂)<π(a+λu)\text{wt}({\bm{\alpha}})<\pi(a+\lambda u) such that the following holds for all polynomials f[x1,,xn]f\in{\mathbb{Z}}[x_{1},\ldots,x_{n}],

ψpk(λ=0pk1wt(𝜶)<π(a+λu)cλ,𝜶f(𝜶)(ζa+λu))=f(zu)T¯.\psi_{p^{k}}\left(\sum\limits_{\lambda=0}^{p^{k}-1}\sum\limits_{\text{wt}({\bm{\alpha}})<\pi(a+\lambda u)}c_{\lambda,{\bm{\alpha}}}f^{({\bm{\alpha}})}(\zeta^{a+\lambda u})\right)=f(z^{u^{\prime}})\in\overline{T}_{\ell}.

For the application of ψpk\psi_{p^{k}} in the statement of the lemma to make sense we must have its input be an element in TkT_{\ell}^{k}. In other words, we need the input of ψpk\psi_{p^{k}} to be a polynomial in zz with coefficients in (ζ){\mathbb{Z}}(\zeta). This is indeed the case and will be part of the proof of the lemma. We need two simple facts for the proof.

Lemma 3.12 (Hasse Derivatives of composition of two functions).

Let 𝔽{\mathbb{F}} be a field, nn\in{\mathbb{N}}. Given a tuple of polynomials C(y)=(C1(y),C2(y),,Cn(y))(𝔽[y])nC(y)=(C_{1}(y),C_{2}(y),\ldots,C_{n}(y))\in({\mathbb{F}}[y])^{n}, ww\in{\mathbb{N}} and γ𝔽\gamma\in{\mathbb{F}} there exists a set of coefficients bw,𝛂𝔽b_{w,{\bm{\alpha}}}\in{\mathbb{F}} (which depend on CC and γ\gamma) where 𝛂0n{\bm{\alpha}}\in{\mathbb{Z}}^{n}_{\geq 0} such that for any f𝔽[x1,,xn]f\in{\mathbb{F}}[x_{1},\ldots,x_{n}] we have,

h(w)(γ)=wt(𝜶)wbw,𝜶f(𝜶)(C1(γ),,Cn(γ)),h^{(w)}(\gamma)=\sum\limits_{\text{wt}({\bm{\alpha}})\leq w}b_{w,{\bm{\alpha}}}f^{({\bm{\alpha}})}(C_{1}(\gamma),\ldots,C_{n}(\gamma)),

where h(y)=f(C1(y),,Cn(y))h(y)=f(C_{1}(y),\ldots,C_{n}(y)).

This fact follows easily from the definition of the Hasse derivative. A proof can also be found in Proposition 6 of [10].The idea is that Ci(γ+z)C_{i}(\gamma+z) can be expanded in terms of its Hasse derivatives. We use that in f(C1(γ+z),,Cn(γ+z))f(C_{1}(\gamma+z),\ldots,C_{n}(\gamma+z)) and expand it again using the Hasse derivative expansion of ff at (C1(γ),,Cn(γ))(C_{1}(\gamma),\ldots,C_{n}(\gamma)). The coefficient of zwz^{w} is the Hasse derivative h(w)(γ)h^{(w)}(\gamma) and it will only get contributions from f(𝜶)f^{({\bm{\alpha}})} with wt(𝜶)w\text{wt}({\bm{\alpha}})\leq w. We also need another fact about the isomorphism between polynomials and the evaluations of their derivatives at a sufficiently large set of points.

Lemma 3.13 (Computing polynomial coefficients from polynomial evaluations).

Let 𝔽{\mathbb{F}} be a field and nn\in{\mathbb{N}}. Given distinct ai𝔽a_{i}\in{\mathbb{F}} and mi0m_{i}\in{\mathbb{Z}}_{\geq 0}, let h(y)=i=1n(yai)mi𝔽[y]h(y)=\prod_{i=1}^{n}(y-a_{i})^{m_{i}}\in{\mathbb{F}}[y]. We have an isomorphism between

𝔽[z]h(z)𝔽i=1nmi,\frac{{\mathbb{F}}[z]}{\left\langle h(z)\right\rangle}\longleftrightarrow{\mathbb{F}}^{\sum\limits_{i=1}^{n}m_{i}},

which maps every polynomial f𝔽[z]/h(z)f\in{\mathbb{F}}[z]/\langle h(z)\rangle to the evaluations (f(ji)(ai))i,ji(f^{(j_{i})}(a_{i}))_{i,j_{i}} where i{1,,n}i\in\{1,\ldots,n\} and ji{0,,mi1}j_{i}\in\{0,\ldots,m_{i}-1\}.

This is a simple generalization of Lemma 3.10. It can be proven in several ways, for example it can be proven using Lemma 3.10 and the Chinese remainder theorem for the ring 𝔽[y]{\mathbb{F}}[y]. To prove Lemma 3.11 we will need the following corollary of the fact above.

Corollary 3.14 (Computing a polynomial from its evaluations).

Let 𝔽{\mathbb{F}} be a field and nn\in{\mathbb{N}}. Given distinct ai𝔽a_{i}\in{\mathbb{F}} and mi0m_{i}\in{\mathbb{Z}}_{\geq 0}, let h(y)=i=1n(yai)mi𝔽[y]h(y)=\prod_{i=1}^{n}(y-a_{i})^{m_{i}}\in{\mathbb{F}}[y]. Then there exists coefficients ti,j𝔽[z]t_{i,j}\in{\mathbb{F}}[z] (depending on hh) for i{1,,n},j{0,,mi1}i\in\{1,\ldots,n\},j\in\{0,\ldots,m_{i}-1\} such that for any f(y)𝔽[y]f(y)\in{\mathbb{F}}[y] we have,

i=1nj=0mi1ti,jf(j)(ai)=f(z) (mod h(z))𝔽[z]/h(z).\sum\limits_{i=1}^{n}\sum\limits_{j=0}^{m_{i}-1}t_{i,j}f^{(j)}(a_{i})=f(z)\text{ {\em(mod }}h(z)\text{{\em)}}\in{\mathbb{F}}[z]/\langle h(z)\rangle.
Proof.

Lemma 3.13 implies that there exists a 𝔽{\mathbb{F}}-linear map which can compute the coefficients of 1,z,z2,,zi=1nmi11,z,z^{2},\ldots,z^{\sum_{i=1}^{n}m_{i}-1} of f(z)𝔽[z]/h(z)f(z)\in{\mathbb{F}}[z]/\langle h(z)\rangle from the evaluations f(ji)(ai)f^{(j_{i})}(a_{i}) for i{1,,n}i\in\{1,\ldots,n\} and ji{0,,mi1}j_{i}\in~{}\{0,\ldots,m_{i}-1\}. Multiplying these coefficients with 1,,zi=1nmi11,\ldots,z^{\sum_{i=1}^{n}m_{i}-1} computes f(z)f(z) in 𝔽[z]/h(z){\mathbb{F}}[z]/\langle h(z)\rangle. ∎

We are now ready to prove Lemma 3.11.

Proof of Lemma 3.11.

As the statement we are trying to prove is linear over {\mathbb{Z}} we see that it suffices to prove the lemma in the case of when ff equals a monomial. Given v0nv\in{\mathbb{Z}}_{\geq 0}^{n} we let mv(x)=mv(x1,,xn)=x1v1xnvnm_{v}(x)=m_{v}(x_{1},\ldots,x_{n})=x_{1}^{v_{1}}\ldots x_{n}^{v_{n}} be a monomial in 𝔽[x1,,xn]{\mathbb{F}}[x_{1},\ldots,x_{n}] where 𝔽{\mathbb{F}} is an arbitrary field (we will be working with 𝔽=(ζ){\mathbb{F}}={\mathbb{Q}}(\zeta) and 𝔽=𝔽p{\mathbb{F}}={\mathbb{F}}_{p}). Let C(y)=yu=(yu1,yu2,,yun)(𝔽[y])nC(y)=y^{u^{\prime}}=(y^{u^{\prime}_{1}},y^{u^{\prime}_{2}},\ldots,y^{u^{\prime}_{n}})\in({\mathbb{F}}[y])^{n} where u=(u1,,un)(/p)n,u=(u1,,un)(/pk)n1u^{\prime}=(u^{\prime}_{1},\ldots,u^{\prime}_{n})\in({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n},u=(u_{1},\ldots,u_{n})\in{\mathbb{P}}({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n-1} and u (mod pk)=uu^{\prime}\text{ (mod }p^{k}\text{)}=u. For this proof we use the elements in {0,,p1}\{0,\ldots,p^{\ell}-1\} to represent the set /p{\mathbb{Z}}/p^{\ell}{\mathbb{Z}} (similarly for (/pk)n({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n}).

We first prove the following claim.

Claim 3.15.

Let w,λ/pkw\in{\mathbb{N}},\lambda\in{\mathbb{Z}}/p^{k}{\mathbb{Z}}. There exists coefficients bw,𝛂(λ)(ζ)b^{\prime}_{w,{\bm{\alpha}}}(\lambda)\in{\mathbb{Q}}(\zeta) (depending on w,λw,\lambda and CC) for 𝛂0n{\bm{\alpha}}\in{\mathbb{Z}}_{\geq 0}^{n} with wt(𝛂)w\text{wt}({\bm{\alpha}})\leq~{}w such that for all monomials mv(x)[x1,,xn],v0nm_{v}(x)\in{\mathbb{Z}}[x_{1},\ldots,x_{n}],v\in{\mathbb{Z}}_{\geq 0}^{n} we have,

wt(𝜶)wbw,𝜶(λ)mv(𝜶)(ζa+λu)=ζa,v(mvC)(w)(ζλ).\sum\limits_{\text{wt}({\bm{\alpha}})\leq w}b^{\prime}_{w,{\bm{\alpha}}}(\lambda)m_{v}^{({\bm{\alpha}})}(\zeta^{a+\lambda u})=\zeta^{\langle a,v\rangle}(m_{v}\circ C)^{(w)}(\zeta^{\lambda}).
Proof.

For every λ/pk\lambda\in{\mathbb{Z}}/p^{k}{\mathbb{Z}} and ww\in{\mathbb{N}}, using Lemma 3.12 we can find coefficients bw,𝜶(λ)(ζ)b_{w,{\bm{\alpha}}}(\lambda)\in{\mathbb{Q}}(\zeta) such that,

(fC)(w)(ζλ)=wt(𝜶)wbw,𝜶(λ)f(𝜶)(C(ζλ))=wt(𝜶)wbw,𝜶(λ)f(𝜶)(ζλu),\displaystyle(f\circ C)^{(w)}(\zeta^{\lambda})=\sum\limits_{\text{wt}({\bm{\alpha}})\leq w}b_{w,{\bm{\alpha}}}(\lambda)f^{({\bm{\alpha}})}(C(\zeta^{\lambda}))=\sum\limits_{\text{wt}({\bm{\alpha}})\leq w}b_{w,{\bm{\alpha}}}(\lambda)f^{({\bm{\alpha}})}(\zeta^{\lambda u^{\prime}}), (5)

for every polynomial f(ζ)[x1,,xn]f\in{\mathbb{Q}}(\zeta)[x_{1},\ldots,x_{n}] where (fC)(y)=f(yu)(ζ)[y](f\circ C)(y)=f(y^{u^{\prime}})\in{\mathbb{Q}}(\zeta)[y]. As u (mod pk)=uu^{\prime}\text{ (mod }p^{k}\text{)}=u and ζ\zeta is a primitive pkp^{k}’th root of unity in {\mathbb{C}} we note that

ζa+λu=ζa+λu.\zeta^{a+\lambda u}=\zeta^{a+\lambda u^{\prime}}.

We now make the simple observation that for any 𝜶0n{\bm{\alpha}}\in{\mathbb{Z}}_{\geq 0}^{n} and v0nv\in{\mathbb{Z}}_{\geq 0}^{n} we have

mv(𝜶)(x)=i=1n(viαi)xiviαi,m^{({\bm{\alpha}})}_{v}(x)=\prod\limits_{i=1}^{n}\binom{v_{i}}{\alpha_{i}}x_{i}^{v_{i}-\alpha_{i}},

which implies

mv(𝜶)(ζa+λu)=ζa,vζ𝜶,amv(𝜶)(ζλu)=ζa,vζ𝜶,amv(𝜶)(ζλu).m^{({\bm{\alpha}})}_{v}(\zeta^{a+\lambda u})=\zeta^{\langle a,v\rangle}\zeta^{-\langle{\bm{\alpha}},a\rangle}m^{({\bm{\alpha}})}_{v}(\zeta^{\lambda u})=\zeta^{\langle a,v\rangle}\zeta^{-\langle{\bm{\alpha}},a\rangle}m^{({\bm{\alpha}})}_{v}(\zeta^{\lambda u^{\prime}}).

The above equation combined with (5) for f=mvf=m_{v} implies,

ζa,v(mvC)(w)(ζλ)=wt(𝜶)wζ𝜶,abw,𝜶(λ)mv(𝜶)(ζa+λu),\displaystyle\zeta^{\langle a,v\rangle}(m_{v}\circ C)^{(w)}(\zeta^{\lambda})=\sum\limits_{\text{wt}({\bm{\alpha}})\leq w}\zeta^{\langle{\bm{\alpha}},a\rangle}b_{w,{\bm{\alpha}}}(\lambda)m_{v}^{({\bm{\alpha}})}(\zeta^{a+\lambda u}),

for all ww\in{\mathbb{N}} and v0nv\in{\mathbb{Z}}_{\geq 0}^{n}. Setting bw,𝜶(λ)=ζ𝜶,abw,𝜶(λ)b^{\prime}_{w,{\bm{\alpha}}}(\lambda)=\zeta^{\langle{\bm{\alpha}},a\rangle}b_{w,{\bm{\alpha}}}(\lambda) we are done . ∎

Without loss of generality let us assume xLπ(x)=p\sum_{x\in L}\pi(x)=p^{\ell} (if it is greater we can reduce each of the π(x)\pi(x) until we reach equality - this would just mean that our computation was done by ignoring some higher order derivatives at some of the points).

Let h(y)(ζ)[y](ζ)[y]h(y)\in{\mathbb{Z}}(\zeta)[y]\subseteq{\mathbb{Q}}(\zeta)[y] be the polynomial,

h(y)=λ/pk(yζλ)π(a+λu).h(y)=\prod\limits_{\lambda\in{\mathbb{Z}}/p^{k}{\mathbb{Z}}}(y-\zeta^{\lambda})^{\pi(a+\lambda u)}.

Using Corollary 3.14 there is a (ζ)[z]{\mathbb{Q}}(\zeta)[z]-linear combination of the evaluations (mvC)(w)(ζλ)(m_{v}\circ C)^{(w)}(\zeta^{\lambda}) for λ/pk\lambda\in{\mathbb{Z}}/p^{k}{\mathbb{Z}} and w<π(a+λu)w<\pi(a+\lambda u) which can compute the element

(mvC)(z)=zv,u(ζ)[z]/h(z).(m_{v}\circ C)(z)=z^{\langle v,u^{\prime}\rangle}\in{\mathbb{Q}}(\zeta)[z]/\langle h(z)\rangle.

This statement along with Claim 3.15 leads to the following: there exist elements cλ,𝜶(ζ)[z]c_{\lambda,{\bm{\alpha}}}\in{\mathbb{Q}}(\zeta)[z] (depending on π,L\pi,L and uu^{\prime}) for λ/pk\lambda\in{\mathbb{Z}}/p^{k}{\mathbb{Z}} and 𝜶0n{\bm{\alpha}}\in{\mathbb{Z}}_{\geq 0}^{n} with wt(𝜶)<π(a+λu)\text{wt}({\bm{\alpha}})<\pi(a+\lambda u) such that the following holds for all monomials mv[x1,,xn],v0m_{v}\in{\mathbb{Z}}[x_{1},\ldots,x_{n}],v\in{\mathbb{Z}}_{\geq 0} we have,

λ=0pk1wt(𝜶)<π(a+λu)cλ,𝜶mv(𝜶)(ζa+λu)=ζa,v(mvC)(z)=ζa,vzv,u(ζ)/h(z).\sum\limits_{\lambda=0}^{p^{k}-1}\sum\limits_{\text{wt}({\bm{\alpha}})<\pi(a+\lambda u)}c_{\lambda,{\bm{\alpha}}}m_{v}^{({\bm{\alpha}})}(\zeta^{a+\lambda u})=\zeta^{\langle a,v\rangle}(m_{v}\circ C)(z)=\zeta^{\langle a,v\rangle}z^{\langle v,u^{\prime}\rangle}\in{\mathbb{Q}}(\zeta)/\langle h(z)\rangle.

We claim that these are coefficients we wanted to construct in the statement of this lemma.

All we need to show now is that ψpk\psi_{p^{k}} is a ring homomorphism from the ring (ζ)/h(z){\mathbb{Z}}(\zeta)/\langle h(z)\rangle to the ring T¯=𝔽p(ζ)/z1\overline{T}_{\ell}={\mathbb{F}}_{p}(\zeta)/\langle z^{\ell}-1\rangle and maps ζa,vzv,u\zeta^{\langle a,v\rangle}z^{\langle v,u^{\prime}\rangle} to zv,uz^{\langle v,u^{\prime}\rangle}. This follows from Corollary 3.3 and noting

ψpk(h(z))=(z1)λ/pkπ(a+λu)=(z1)p=zp1𝔽p[z].\psi_{p^{k}}(h(z))=(z-1)^{\sum\limits_{\lambda\in{\mathbb{Z}}/p^{k}{\mathbb{Z}}}\pi(a+\lambda u)}=(z-1)^{p^{\ell}}=z^{p^{\ell}}-1\in{\mathbb{F}}_{p}[z].

We will use the lemma above in the form of the following corollary.

Corollary 3.16 (Decoding Mp,nM_{p^{\ell},n} from Up(𝜶)U^{({\bm{\alpha}})}_{p^{\ell}}).

Let L={a+λu|λ/pk}(/pk)nL=\{a+\lambda u|\lambda\in{\mathbb{Z}}/p^{k}{\mathbb{Z}}\}\subset({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n} with a(/pk)n,u(/pk)n1a\in({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n},u\in{\mathbb{P}}({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n-1}, u(/p)nu^{\prime}\in({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n} be such that u (mod pk)=uu^{\prime}\text{ (mod }p^{k}\text{)}=u and π:L0\pi:L\rightarrow{\mathbb{Z}}_{\geq 0} be a function which satisfies xLπ(x)p\sum_{x\in L}\pi(x)\geq p^{\ell} then there exists a (ζ)[z]{\mathbb{Q}}(\zeta)[z]-linear combination (with coefficients depending on π,L\pi,L and uu^{\prime}) of the vectors Up(𝛂)(ζx)U_{p^{\ell}}^{({\bm{\alpha}})}(\zeta^{x}) for xLx\in L and 𝛂{\bm{\alpha}} of weight strictly less than π(x)\pi(x) which under the map ψpk\psi_{p^{k}} gives us the vector Mp,n(u)M_{p^{\ell},n}(u^{\prime}).

Proof.

Lemma 3.11 gives us the required linear combination. ∎

The coefficients defined by the above corollary correspond to the rows of the matrix KSK_{S} in the proof overview (Section 1.1).

4 Kakeya Set bounds over /pk{\mathbb{Z}}/p^{k}{\mathbb{Z}}

We first prove two helper lemmas. We recall that GLn(/pk)\text{GL}_{n}({\mathbb{Z}}/p^{k}{\mathbb{Z}}) is the set of linear isomorphisms over (/pk)n({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n}. They are represented by matrices whose determinants are units in /pk{\mathbb{Z}}/p^{k}{\mathbb{Z}}.

The first lemma implies that given a large subset D(/pk)n1D\subseteq{\mathbb{P}}({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n-1} there exists a WW in GLn(/pk)\text{GL}_{n}({\mathbb{Z}}/p^{k}{\mathbb{Z}}) such that Mp,n(WD)M_{p^{\ell},n}(W\cdot D) has high rank where

WD={Wu|uD}W\cdot D=\{W\cdot u|u\in D\}

that is the set of directions obtained by rotating the elements in DD by WW. In general, we also prove that there exists a WW such that Mp,n(DW)M_{p^{\ell},n}(D^{\prime}_{W}) has high rank where

DW={u(/p)n|u (mod pk)WD}D^{\prime}_{W}=\{u\in({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n}|u\text{ (mod }p^{k}\text{)}\in W\cdot D\}

that is the set of directions in (/p)n({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n} which under the mod pkp^{k} map give an element in WDW\cdot D. We prove the statement using a random rotation argument.

Lemma 4.1.

Let k,n,ϵ0k,n\in{\mathbb{N}},\epsilon\geq 0 and pp be a prime. Given a set D(/pk)n1D\subseteq{\mathbb{P}}({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n-1} containing at least an ϵ\epsilon fraction of elements in (/pk)n1{\mathbb{P}}({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n-1} then there exists a matrix WGLn(/pk)W\in\text{GL}_{n}({\mathbb{Z}}/p^{k}{\mathbb{Z}}) such that,

rank𝔽pMp,n(DW)ϵ(p1+nn).\text{rank}_{{\mathbb{F}}_{p}}M_{p^{\ell},n}(D^{\prime}_{W})\geq\epsilon\cdot\binom{p^{\ell}\ell^{-1}+n}{n}.
Proof.

From Lemma 3.7 we know that the 𝔽p{\mathbb{F}}_{p} rank of Mp,n((/p)n1)M_{p^{\ell},n}(({\mathbb{P}}{\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n-1}) is at least (pl1+nn)\binom{p^{\ell}l^{-1}+n}{n}. This means there exists a set of rows VV of size at least (pl1+nn)\binom{p^{\ell}l^{-1}+n}{n} in Coeff(Mp,n((/p)n1))\text{Coeff}(M_{p^{\ell},n}(({\mathbb{P}}{\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n-1})) which are 𝔽p{\mathbb{F}}_{p}-linearly independent.

We pick WGLn(/pk)W\in\text{GL}_{n}({\mathbb{Z}}/p^{k}{\mathbb{Z}}) uniformly at random and as in the statement of the lemma consider the set DW={u(/p)n|u (mod pk)WD}D^{\prime}_{W}=\{u\in({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n}|u\text{ (mod }p^{k}\text{)}\in W\cdot D\}. The key claim about DWD^{\prime}_{W} is the following.

Claim 4.2.

Any given row vVv\in V will appear in Coeff(Mp,n(DW))\text{Coeff}(M_{p^{\ell},n}(D^{\prime}_{W})) with probability at least ϵ\epsilon.

Proof.

As GLn(/pk)\text{GL}_{n}({\mathbb{Z}}/p^{k}{\mathbb{Z}}) acts transitively on the set (/pk)n1{\mathbb{P}}({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n-1} and |D|ϵ|(/pk)n1||D|\geq\epsilon|{\mathbb{P}}({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n-1}| we see that DWD^{\prime}_{W} will contain a particular u(/p)n1u\in{\mathbb{P}}({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n-1} with probability at least ϵ\epsilon. The row vVv\in V will be within Coeff(Mp,n(u))\text{Coeff}(M_{p^{\ell},n}(u)) for some u(/p)n1u\in{\mathbb{P}}({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n-1}. This means that vVv\in V will appear in Coeff(Mp,n(DW))\text{Coeff}(M_{p^{\ell},n}(D^{\prime}_{W})) with probability at least ϵ\epsilon. ∎

Now by linearity of expectation we see that the expected number of rows VV appearing in Coeff(Mp,n(DW))\text{Coeff}(M_{p^{\ell},n}(D^{\prime}_{W})) is at least ϵ|V|\epsilon|V|. This means there exists a choice of WW such that an ϵ\epsilon fraction of elements in VV do appear in Coeff(Mp,n(DW))\text{Coeff}(M_{p^{\ell},n}(D^{\prime}_{W})). As all these rows are linearly independent then by Lemma 2.6 we see that the 𝔽p{\mathbb{F}}_{p}-rank of Mp,n(DW)M_{p^{\ell},n}(D^{\prime}_{W}) must be at least ϵ|V|ϵ(pl1+nn)\epsilon|V|\geq\epsilon\binom{p^{\ell}l^{-1}+n}{n}. ∎

The second lemma lower bounds the size of (m,ϵ)(m,\epsilon)-Kakeya sets by the rank of a sub-matrix of Mp,nM_{p^{\ell},n} with rows corresponding to directions with rich lines.

Lemma 4.3.

Let k,nk,n\in{\mathbb{N}} and pp be a prime. Let S(/pk)nS\subseteq({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n} be a (m,ϵ)(m,\epsilon)-Kakeya set and D(/pk)n1,|D|ϵ|(/pk)n1|D\subseteq{\mathbb{P}}({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n-1},|D|\geq\epsilon|{\mathbb{P}}({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n-1}| such that for every uDu\in D we have a mm-rich line LuL_{u} with respect to SS in direction uu. Then we have the following bound,

|S|(pm1+n1n)rank𝔽pMp,n(D),|S|\binom{\lceil p^{\ell}m^{-1}\rceil+n-1}{n}\geq\text{rank}_{{\mathbb{F}}_{p}}M_{p^{\ell},n}(D^{\prime}),

for all logp(m)\ell\geq\log_{p}(m) where D=DI={u(/p)n|u (mod pk)D}.D^{\prime}=D^{\prime}_{I}=\{u\in({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n}|u\text{ (mod }p^{k}\text{)}\in D\}.

Proof.

We set

b=pm1.b=\lceil p^{\ell}m^{-1}\rceil.

Consider the family of row vectors, Up(𝜶)(ζx)U_{p^{\ell}}^{({\bm{\alpha}})}(\zeta^{x}) where xSx\in S and wt(𝜶)<b\text{wt}({\bm{\alpha}})<b. For a uDu\in D, let LuL_{u} be a mm-rich line with respect to SS in direction uu. Let 𝒰Lu,p{\cal U}_{L_{u},p^{\ell}} be the matrix constructed by concatenating the row vectors Up(𝜶)(ζx)U_{p^{\ell}}^{({\bm{\alpha}})}(\zeta^{x}) where xLux\in L_{u} and wt(𝜶)<b\text{wt}({\bm{\alpha}})<b. By construction as the rows in {𝒰Ld,p}uD\{{\cal U}_{L_{d},p^{\ell}}\}_{u\in D} correspond to tuples xSx\in S and 𝜶0n,wt(𝜶)<b{\bm{\alpha}}\in{\mathbb{Z}}_{\geq 0}^{n},\text{wt}({\bm{\alpha}})<b we have the following bound,

|S|(b+n1n)crank(ζ){𝒰Lu,p}uD.\displaystyle|S|\binom{b+n-1}{n}\geq\text{crank}_{{\mathbb{Q}}(\zeta)}\{{\cal U}_{L_{u},p^{\ell}}\}_{u\in D}. (6)

Let πu:Lu0\pi_{u}:L_{u}\rightarrow{\mathbb{Z}}_{\geq 0} be a weight function which gives weight bb to points in SLuS\cap L_{u} and 0 elsewhere. Using Corollary 3.16 for the line LuL_{u} and function πu\pi_{u}, for all u(/p)nu^{\prime}\in({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n} such that u (mod pk)=uu^{\prime}\text{ (mod }p^{k}\text{)}=u, we can now construct row vectors CuC_{u^{\prime}} such that ψpk(Cu𝒰Lu,p)=Mp,n(u)\psi_{p^{k}}(C_{u^{\prime}}{\cal U}_{L_{u},p^{\ell}})=M_{p^{\ell},n}(u^{\prime}). For convenience, for every u(/pk)n1u\in{\mathbb{P}}({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n-1} we define 𝒞u{\cal C}_{u} and u\mathcal{M}_{u} as the matrices whose row vectors are CuC_{u^{\prime}} and Mp,n(u)M_{p^{\ell},n}(u^{\prime}) respectively for all u(/p)nu^{\prime}\in({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n} such that u (mod pk)=uu^{\prime}\text{ (mod }p^{k}\text{)}=u. We note, ψpk(𝒞u𝒰Lu,p)=p,n(u)\psi_{p^{k}}({\cal C}_{u}{\cal U}_{L_{u},p^{\ell}})=\mathcal{M}_{p^{\ell},n}(u).

We use Lemma 2.3 on {𝒰Lu,p}uD\{{\cal U}_{L_{u},p^{\ell}}\}_{u\in D} being multiplied by 𝒞u,uD{\cal C}_{u},u\in D and applying Lemma 3.4 next to get,

crank(ζ){𝒰Lu,p}uD\displaystyle\text{crank}_{{\mathbb{Q}}(\zeta)}\{{\cal U}_{L_{u},p^{\ell}}\}_{u\in D} crank(ζ){𝒞u𝒰Lu,p}uD\displaystyle\geq\text{crank}_{{\mathbb{Q}}(\zeta)}\{{\cal C}_{u}{\cal U}_{L_{u},p^{\ell}}\}_{u\in D}
crank𝔽p{p,n(u)}uD=rank𝔽pMp,n(D),\displaystyle\geq\text{crank}_{{\mathbb{F}}_{p}}\{\mathcal{M}_{p^{\ell},n}(u)\}_{u\in D}=\text{rank}_{{\mathbb{F}}_{p}}M_{p^{\ell},n}(D^{\prime}),

where we recall D=DI={u(/p)n|u (mod pk)D}.D^{\prime}=D^{\prime}_{I}=\{u\in({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n}|u\text{ (mod }p^{k}\text{)}\in D\}. Using the above inequality with (6) we have,

|S|(b+n1n)=|S|(pm1+n1n)rank𝔽pMp,n(D).\displaystyle|S|\binom{b+n-1}{n}=|S|\binom{\lceil p^{\ell}m^{-1}\rceil+n-1}{n}\geq\text{rank}_{{\mathbb{F}}_{p}}M_{p^{\ell},n}(D^{\prime}).

We are now ready to prove Theorem 1.8.

Theorem 1.8.

Let k,n,pk,n,p\in{\mathbb{N}} with pp prime. Any (m,ϵ)(m,\epsilon)-Kakeya set S(/pk)nS\subseteq({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n} satisfies the following bound,

|S|ϵ(mn(2(k+logp(n)))n).|S|\geq\epsilon\cdot\left(\frac{m^{n}}{(2(k+\lceil\log_{p}(n)\rceil))^{n}}\right).

When p>np>n, we also get the following stronger bound for (m,ϵ)(m,\epsilon)-Kakeya set SS in (/pk)n({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n},

|S|ϵ(mn(k+1)n)(1+n/p)n.|S|\geq\epsilon\cdot\left(\frac{m^{n}}{(k+1)^{n}}\right)(1+n/p)^{-n}.
Proof.

Set =k+logp(n)\ell=k+\lceil\log_{p}(n)\rceil. Let DD be the set of directions in (/pk)n1{\mathbb{P}}({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n-1} which have mm-rich lines. We know DD contains at least an ϵ\epsilon fraction of directions. By Lemma 4.1 there exists a matrix WGLn(/pk)W\in\text{GL}_{n}({\mathbb{Z}}/p^{k}{\mathbb{Z}}) such that,

rank𝔽pMp,n(DW)ϵ(p1+nn),\text{rank}_{{\mathbb{F}}_{p}}M_{p^{\ell},n}(D^{\prime}_{W})\geq\epsilon\cdot\binom{p^{\ell}\ell^{-1}+n}{n},

where DW={u(/p)n|u (mod pk)WD}.D^{\prime}_{W}=\{u\in({\mathbb{Z}}/p^{\ell}{\mathbb{Z}})^{n}|u\text{ (mod }p^{k}\text{)}\in W\cdot D\}.

We now note WSW\cdot S is also an (m,ϵ)(m,\epsilon)-Kakeya set with WDW\cdot D as the set of directions with mm-rich lines. We now apply Lemma 4.3 with the inequality above to get,

|S|(pm1+nn)|S|(pm1+n1n)rank𝔽pMp,n(DW)ϵ(p1+nn).|S|\binom{p^{\ell}m^{-1}+n}{n}\geq|S|\binom{\lceil p^{\ell}m^{-1}\rceil+n-1}{n}\geq\text{rank}_{{\mathbb{F}}_{p}}M_{p^{\ell},n}(D^{\prime}_{W})\geq\epsilon\cdot\binom{p^{\ell}\ell^{-1}+n}{n}.

For convenience we set a=logp(n)a=\lceil\log_{p}(n)\rceil which implies panpa1p^{a}\geq n\geq p^{a-1}. The above inequality implies,

|S|\displaystyle|S| ϵi=1npkpa(k+a)1+ipkpam1+i\displaystyle\geq\epsilon\cdot\prod\limits_{i=1}^{n}\frac{p^{k}p^{a}(k+a)^{-1}+i}{p^{k}p^{a}m^{-1}+i}
ϵi=1nm(k+a)1+impkpa1+impkpa\displaystyle\geq\epsilon\cdot\prod\limits_{i=1}^{n}\frac{m(k+a)^{-1}+imp^{-k}p^{-a}}{1+imp^{-k}p^{-a}}
ϵmn(k+a)ni=1n(1+ipa)1.\displaystyle\geq\epsilon\cdot\frac{m^{n}}{(k+a)^{n}}\prod\limits_{i=1}^{n}\left(1+ip^{-a}\right)^{-1}.

For ini\leq n we have (1+ipa)12\left(1+ip^{-a}\right)^{-1}\leq 2 which completes the proof. Note, the second half of the theorem follows from observing that for p>np>n we have (1+ipa)11+n/p\left(1+ip^{-a}\right)^{-1}\leq 1+n/p and a=logp(n)=1a=\lceil\log_{p}(n)\rceil=1. ∎

5 Kakeya Set bounds over /N{\mathbb{Z}}/N{\mathbb{Z}} for general NN

We need some simple facts about /pkN{\mathbb{Z}}/p^{k}N{\mathbb{Z}} for pp prime and pp and NN co-prime which follow from the Chinese remainder Theorem.

Lemma 5.1 (Geometry of /pkN{\mathbb{Z}}/p^{k}N{\mathbb{Z}}).

Let p,N,n,k,R=/pkN,R1=/N,R0=/pkp,N,n,k\in{\mathbb{N}},R={\mathbb{Z}}/p^{k}N{\mathbb{Z}},R_{1}={\mathbb{Z}}/N{\mathbb{Z}},R_{0}={\mathbb{Z}}/p^{k}{\mathbb{Z}} with pp prime and co-prime to NN. Using the Chinese remainder theorem we know that RnR^{n} is isomorphic to R0n×R1nR_{0}^{n}\times R_{1}^{n} and Rn1{\mathbb{P}}R^{n-1} is isomorphic to R0n1×R1n1{\mathbb{P}}R_{0}^{n-1}\times{\mathbb{P}}R_{1}^{n-1}. Finally, any line L={a+λu|λR}L=\{a+\lambda u|\lambda\in R\} with direction u=(u0,u1)R0n×R1nu=(u_{0},u_{1})\in R_{0}^{n}\times R_{1}^{n} in RnR^{n} is equivalent to the product of a line L0R0nL_{0}\subset R_{0}^{n} in direction u0u_{0} and a line L1R1nL_{1}\subset R_{1}^{n} in direction u1u_{1}.

We briefly discuss the proof strategy first. We use ideas from [6] to prove Theorem 1.9. We give the idea for N=pk1qk2N=p^{k_{1}}q^{k_{2}} where pp and qq are distinct primes. Let SS be Kakeya set which contains lines LuL_{u} in the direction u(/N)n1u\in{\mathbb{P}}({\mathbb{Z}}/N{\mathbb{Z}})^{n-1}. Using Lemma 5.1 we know that the line LuL_{u}, u=(up,uq)(/pk1)n1×(/qk2)n1u=(u_{p},u_{q})\in{\mathbb{P}}({\mathbb{Z}}/p^{k_{1}}{\mathbb{Z}})^{n-1}\times{\mathbb{P}}({\mathbb{Z}}/q^{k_{2}}{\mathbb{Z}})^{n-1} can be decomposed as a product of lines Lp(up,uq)L_{p}(u_{p},u_{q}) over /pk1{\mathbb{Z}}/p^{k_{1}}{\mathbb{Z}} and Lq(up,uq)L_{q}(u_{p},u_{q}) over /qk2{\mathbb{Z}}/q^{k_{2}}{\mathbb{Z}} with directions up(/pk1)n1u_{p}\in{\mathbb{P}}({\mathbb{Z}}/p^{k_{1}}{\mathbb{Z}})^{n-1} and uq(/qk2)n1u_{q}\in{\mathbb{P}}({\mathbb{Z}}/q^{k_{2}}{\mathbb{Z}})^{n-1} respectively. Note that Lp(up,uq)L_{p}(u_{p},u_{q}) and Lq(up,uq)L_{q}(u_{p},u_{q}) depend on u=(up,uq)u=(u_{p},u_{q}), and not just on upu_{p} or uqu_{q} respectively.

Let ζ\zeta be a primitive pk1p^{k_{1}}’th root of unity in {\mathbb{C}} and 𝟏y{\mathbf{1}}_{y} be the indicator vector of the point y(/qk2)ny\in({\mathbb{Z}}/q^{k_{2}}{\mathbb{Z}})^{n}. We then examine the span of vectors

Upk1(ζx)𝟏y=Upk1((0,,0))(x)𝟏yU_{p^{k_{1}}}(\zeta^{x})\otimes{\mathbf{1}}_{y}=U^{((0,\ldots,0))}_{p^{k_{1}}}(x)\otimes{\mathbf{1}}_{y}

for xLp(up,uq),yLq(up,uq),(up,uq)(/pk1)n1×(/qk2)n1x\in L_{p}(u_{p},u_{q}),y\in L_{q}(u_{p},u_{q}),(u_{p},u_{q})\in{\mathbb{P}}({\mathbb{Z}}/p^{k_{1}}{\mathbb{Z}})^{n-1}\times{\mathbb{P}}({\mathbb{Z}}/q^{k_{2}}{\mathbb{Z}})^{n-1}. As there is one vector for each point in SS, the dimension of the space spanned by these vectors is at most |S||S|. We then use the decoding procedure for Upk1U_{p^{k_{1}}} from Corollary 3.16 to linearly generate vectors

Mpk1,n(up)𝟏yM_{p^{k_{1}},n}(u_{p})\otimes{\mathbf{1}}_{y}

for yLq(up,uq),(up,uq)(/pk1)n1×(/qk2)n1y\in L_{q}(u_{p},u_{q}),(u_{p},u_{q})\in{\mathbb{P}}({\mathbb{Z}}/p^{k_{1}}{\mathbb{Z}})^{n-1}\times{\mathbb{P}}({\mathbb{Z}}/q^{k_{2}}{\mathbb{Z}})^{n-1} from the vectors Upk1(ζx)𝟏yU_{p^{k_{1}}}(\zeta^{x})\otimes{\mathbf{1}}_{y} for xLp(up,uq),yLq(up,uq)x\in L_{p}(u_{p},u_{q}),y\in L_{q}(u_{p},u_{q}). This means the dimension of the span of Mpk1,n(up)yM_{p^{k_{1}},n}(u_{p})\otimes y for yLq(up,uq),(up,uq)(/pk1)n1×(/qk2)n1y\in L_{q}(u_{p},u_{q}),(u_{p},u_{q})\in{\mathbb{P}}({\mathbb{Z}}/p^{k_{1}}{\mathbb{Z}})^{n-1}\times{\mathbb{P}}({\mathbb{Z}}/q^{k_{2}}{\mathbb{Z}})^{n-1} lower bounds the size of SS. Note that Mpk1,n(up)M_{p^{k_{1}},n}(u_{p}) only depends on upu_{p} and not on uqu_{q}.

We pick the largest subset of rows VV in Coeff(Mpk1,n((/pk1)n1)\text{Coeff}(M_{p^{k_{1}},n}({\mathbb{P}}({\mathbb{Z}}/p^{k_{1}}{\mathbb{Z}})^{n-1}) which are linearly independent. By Lemma 3.6, VV has large size. Any vector vVv\in V will correspond to a row in Coeff(Mpk1,n(up))\text{Coeff}(M_{p^{k_{1}},n}(u_{p})) for some up(/pk)n1u_{p}\in{\mathbb{P}}({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n-1}. For that upu_{p}, the span of 𝟏y{\mathbf{1}}_{y} for yLq(up,uq)y\in L_{q}(u_{p},u_{q}) as uqu_{q} varies in (/qk2)n1{\mathbb{P}}({\mathbb{Z}}/q^{k_{2}}{\mathbb{Z}})^{n-1} will have dimension exactly equal to the size of the set

Sv=uq(/qk2)n1Lq(up,uq)S_{v}=\bigcup\limits_{u_{q}\in{\mathbb{P}}({\mathbb{Z}}/q^{k_{2}}{\mathbb{Z}})^{n-1}}L_{q}(u_{p},u_{q})

which is a Kakeya set in (/qk2)n({\mathbb{Z}}/q^{k_{2}}{\mathbb{Z}})^{n} and has large size by Theorem 1.8. Simple properties of the tensor product (Lemma 2.10) now imply that v𝟏yv\otimes{\mathbf{1}}_{y} for vVv\in V and ySvy\in S_{v} are linearly independent which gives us a rank bound and hence a lower bound on the size of SS

Using Up(𝜶)U^{({\bm{\alpha}})}_{p^{\ell}} for 𝜶0n,wt(𝜶)<pk{\bm{\alpha}}\in{\mathbb{Z}}_{\geq 0}^{n},\text{wt}({\bm{\alpha}})<p^{\ell-k} as in the proof of Theorem 1.8 gives us better constants. The proof for an arbitrary number of distinct prime factors applies the above argument inductively.

Theorem 1.9.

Let nn\in{\mathbb{N}} and R=/NR={\mathbb{Z}}/N{\mathbb{Z}} where N=p1k1prkrN=p_{1}^{k_{1}}\ldots p_{r}^{k_{r}} with distinct primes p1,,prp_{1},\ldots,p_{r} and k1,,krk_{1},\ldots,k_{r}\in{\mathbb{N}}. Any Kakeya set SS in RnR^{n} satisfies,

|S|Nni=1r(2(ki+logpi(n)))n.|S|\geq N^{n}\prod\limits_{i=1}^{r}(2(k_{i}+\lceil\log_{p_{i}}(n)\rceil))^{-n}.

When p1,,prnp_{1},\ldots,p_{r}\geq n, we also get the following stronger lower bound for the size of a Kakeya set SS in (/N)n,N=p1k1prkr({\mathbb{Z}}/N{\mathbb{Z}})^{n},N=p_{1}^{k_{1}}\ldots p_{r}^{k_{r}},

|S|Nni=1r(ki+1)n(1+n/pi)n.|S|\geq N^{n}\prod\limits_{i=1}^{r}(k_{i}+1)^{-n}(1+n/p_{i})^{-n}.
Proof.

Our proof will apply induction on rr. The case of r=1r=1 is Theorem 1.8. Let the theorem be known for rr distinct prime factors. Let N=p0k0p1k1prkrN=p_{0}^{k_{0}}p_{1}^{k_{1}}\ldots p_{r}^{k_{r}} with N1=p1k1prkrN_{1}=p_{1}^{k_{1}}\ldots p_{r}^{k_{r}}. Let R=/NR={\mathbb{Z}}/N{\mathbb{Z}}, R0=/p0k0R_{0}={\mathbb{Z}}/p_{0}^{k_{0}}{\mathbb{Z}} and R1=/N1R_{1}={\mathbb{Z}}/N_{1}{\mathbb{Z}}.

Consider SS a Kakeya set in RnR^{n} and a set of lines indexed by directions uRn1u\in{\mathbb{P}}R^{n-1} such that L(u)={a(u)+λu|λR}RnL(u)=\{a(u)+\lambda u|\lambda\in R\}\subset R^{n} is a line in the direction uu and is contained in SS. By Lemma 5.1 we know uu can be written as a tuple (u0,u1)R0n×R1n(u_{0},u_{1})\in{\mathbb{P}}R_{0}^{n}\times{\mathbb{P}}R_{1}^{n} and L(u)L(u) is a product of a line L0(u0,u1)R0nL_{0}(u_{0},u_{1})\subset R_{0}^{n} in direction u0u_{0} and line L1(u0,u1)R1nL_{1}(u_{0},u_{1})\subset R_{1}^{n} in direction u1u_{1}. We note L1(u0,u1)L_{1}(u_{0},u_{1}) can actually depend on u0u_{0} (similarly for L0(u0,u1)L_{0}(u_{0},u_{1}) and u1u_{1}). This will not be the case only when SS itself is a product of sets from R0nR_{0}^{n} and R1nR_{1}^{n}.

Let

=k0+logp0(n) and b=plogp0(n).\displaystyle\ell=k_{0}+\lceil\log_{p_{0}}(n)\rceil\text{ and }b=p^{\lceil\log_{p_{0}}(n)\rceil}. (7)

Let ζ0\zeta_{0} be a complex primitive p0k0p_{0}^{k_{0}}’th root of unity. We define 𝒰L0(u0,u1){\cal U}_{L_{0}(u_{0},u_{1})} as the matrix constructed by concatenating the row vectors Up0(𝜶)(ζ0x)U_{p_{0}^{\ell}}^{({\bm{\alpha}})}(\zeta_{0}^{x}) where xL0(u0,u1)x\in L_{0}(u_{0},u_{1}) and wt(𝜶)<b\text{wt}({\bm{\alpha}})<b.

Let Yu0,u1Y_{u_{0},u_{1}} be a matrix whose columns are labeled by points in R1nR_{1}^{n} and rows labeled by points in L1(u0,u1)L_{1}(u_{0},u_{1}) such that its yy’th row is 𝟏y{\mathbf{1}}_{y} where 𝟏y{\mathbf{1}}_{y} is the indicator vector of yR1ny\in R_{1}^{n}. We will work with the family of matrices 𝒰L0(u0,u1)Yu0,u1{\cal U}_{L_{0}(u_{0},u_{1})}\otimes Y_{u_{0},u_{1}} for (u0,u1)R0n1×R1n1(u_{0},u_{1})\in{\mathbb{P}}R_{0}^{n-1}\times{\mathbb{P}}R_{1}^{n-1}.

The following claim connects this family to the size of SS.

Claim 5.2.
crank(ζ0){𝒰L0(u0,u1)Yu0,u1}u0R0n1,u1R1n1|S|(b+n1n).\textsf{crank}_{{\mathbb{Q}}(\zeta_{0})}\{{\cal U}_{L_{0}(u_{0},u_{1})}\otimes Y_{u_{0},u_{1}}\}_{u_{0}\in{\mathbb{P}}R_{0}^{n-1},u_{1}\in{\mathbb{P}}R_{1}^{n-1}}\leq|S|\binom{b+n-1}{n}.
Proof.

As we are working with matrices whose entries are in a field, row and column ranks of matrices are identical. The rows of 𝒰L0(u0,u1)Yu0,u1{\cal U}_{L_{0}(u_{0},u_{1})}\otimes Y_{u_{0},u_{1}} are the vectors Up0(𝜶)(ζ0x)𝟏yU_{p_{0}^{\ell}}^{({\bm{\alpha}})}(\zeta_{0}^{x})\otimes{\mathbf{1}}_{y} for xL0(u0,u1)x\in~{}L_{0}(u_{0},u_{1}), wt(𝜶)<b\text{wt}({\bm{\alpha}})<b and yL1(u0,u1)y\in L_{1}(u_{0},u_{1}). Hence there are at most |S|(b+n1n)|S|\binom{b+n-1}{n} distinct rows in the set of 𝒰L0(u0,u1)Yu0,u1{\cal U}_{L_{0}(u_{0},u_{1})}\otimes Y_{u_{0},u_{1}} for u0R0n1,u1R1n1u_{0}\in{\mathbb{P}}R_{0}^{n-1},u_{1}\in{\mathbb{P}}R_{1}^{n-1}. ∎

We now use Lemmas 2.3 and 3.4 and Corollary 3.16 to prove the following claim. For convenience we let R0()=/p0R_{0}(\ell)={\mathbb{Z}}/p_{0}^{\ell}{\mathbb{Z}} and let

J={(u,u0)R0()n1×R0n1|u (mod p0k0)=u0}.J=\{(u^{\prime},u_{0})\in{\mathbb{P}}R_{0}(\ell)^{n-1}\times{\mathbb{P}}R_{0}^{n-1}|u^{\prime}\text{ (mod }p_{0}^{k_{0}}\text{)}=u_{0}\}.
Claim 5.3.
crank𝔽p0{Mp0,n(u)Yu0,u1}(u,u0,u1)J×R1n1crank(ζ0){𝒰L0(u0,u1)Yu0,u1}u0R0n1,u1R1n1.\textsf{crank}_{{\mathbb{F}}_{p_{0}}}\{M_{p_{0}^{\ell},n}(u^{\prime})\otimes Y_{u_{0},u_{1}}\}_{(u^{\prime},u_{0},u_{1})\in J\times{\mathbb{P}}R_{1}^{n-1}}\leq\textsf{crank}_{{\mathbb{Q}}(\zeta_{0})}\{{\cal U}_{L_{0}(u_{0},u_{1})}\otimes Y_{u_{0},u_{1}}\}_{u_{0}\in{\mathbb{P}}R_{0}^{n-1},u_{1}\in{\mathbb{P}}R_{1}^{n-1}}.
Proof.

Let πu0,u1:L0(u0,u1)0\pi_{u_{0},u_{1}}:L_{0}(u_{0},u_{1})\rightarrow{\mathbb{Z}}_{\geq 0} for u0R0n1,u1R1n1u_{0}\in{\mathbb{P}}R_{0}^{n-1},u_{1}\in{\mathbb{P}}R_{1}^{n-1} be a family of functions which takes the constant value bb everywhere. Using Corollary 3.16 for the line L0(u0,u1)L_{0}(u_{0},u_{1}) and function πu0,u1\pi_{u_{0},u_{1}}, for any uR0()n1u^{\prime}\in{\mathbb{P}}R_{0}(\ell)^{n-1} such that (u,u0)J(u^{\prime},u_{0})\in J we can construct row vectors CuC_{u^{\prime}} such that

ψp0k0(Cu𝒰L0(u0,u1))=Mp0,n(u).\psi_{p_{0}^{k_{0}}}(C_{u^{\prime}}\cdot{\cal U}_{L_{0}(u_{0},u_{1})})=M_{p_{0}^{\ell},n}(u^{\prime}).

Now, Lemma 2.8 implies,

ψp0k0((CuIL1(u0,u1))(𝒰L0(u0,u1)Yu0,u1))\displaystyle\psi_{p_{0}^{k_{0}}}((C_{u^{\prime}}\otimes I_{L_{1}(u_{0},u_{1})})\cdot({\cal U}_{L_{0}(u_{0},u_{1})}\otimes Y_{u_{0},u_{1}})) =ψp0k0(Cu𝒰L0(u0,u1))Yu0,u1\displaystyle=\psi_{p_{0}^{k_{0}}}(C_{u^{\prime}}\cdot{\cal U}_{L_{0}(u_{0},u_{1})})\otimes Y_{u_{0},u_{1}}
=Mp0,n(u)Yu0,u1,\displaystyle=M_{p_{0}^{\ell},n}(u^{\prime})\otimes Y_{u_{0},u_{1}}, (8)

where IL1(u0,u1)I_{L_{1}(u_{0},u_{1})} is the identity matrix of size |L1(u0,u1)|=N1|L_{1}(u_{0},u_{1})|=N_{1}.

Applying Lemma 2.3 implies that

crank(ζ0){𝒰L0(u0,u1)Yu0,u1}(u0,u1)R0n1×R1n1\textsf{crank}_{{\mathbb{Q}}(\zeta_{0})}\{{\cal U}_{L_{0}(u_{0},u_{1})}\otimes Y_{u_{0},u_{1}}\}_{(u_{0},u_{1})\in{\mathbb{P}}R_{0}^{n-1}\times{\mathbb{P}}R_{1}^{n-1}}

is larger than

crank(ζ0){(CuIL1(u0,u1))(𝒰L0(u0,u1)Yu0,u1)}(u,u0,u1)J×R1n1.\textsf{crank}_{{\mathbb{Q}}(\zeta_{0})}\{(C_{u^{\prime}}\otimes I_{L_{1}(u_{0},u_{1})})\cdot({\cal U}_{L_{0}(u_{0},u_{1})}\otimes Y_{u_{0},u_{1}})\}_{(u^{\prime},u_{0},u_{1})\in J\times{\mathbb{P}}R_{1}^{n-1}}.

Applying Lemma 3.4 using (5) implies that

crank(ζ0){(Cu𝒰L0(u0,u1))Yu0,u1}(u,u0,u1)J×R1n1\textsf{crank}_{{\mathbb{Q}}(\zeta_{0})}\{(C_{u^{\prime}}\cdot{\cal U}_{L_{0}(u_{0},u_{1})})\otimes Y_{u_{0},u_{1}}\}_{(u^{\prime},u_{0},u_{1})\in J\times{\mathbb{P}}R_{1}^{n-1}}

is greater than

crank𝔽p0{Mp0,n(u)Yu0,u1}(u,u0,u1)J×R1n1.\textsf{crank}_{{\mathbb{F}}_{p_{0}}}\{M_{p_{0}^{\ell},n}(u^{\prime})\otimes Y_{u_{0},u_{1}}\}_{(u^{\prime},u_{0},u_{1})\in J\times{\mathbb{P}}R_{1}^{n-1}}.

Combining these two inequalities using Lemma 2.8 completes the proof. ∎

We want to apply Lemma 2.10 on crank𝔽p0{Mp0,n(u)Yu0,u1}(u,u0,u1)J×R1n1\textsf{crank}_{{\mathbb{F}}_{p_{0}}}\{M_{p_{0}^{\ell},n}(u^{\prime})\otimes Y_{u_{0},u_{1}}\}_{(u^{\prime},u_{0},u_{1})\in J\times{\mathbb{P}}R_{1}^{n-1}}. To that end, we need the following two claims.

Claim 5.4.

For a given u0R0n1u_{0}\in{\mathbb{P}}R_{0}^{n-1},

crank𝔽p0{Yu0,u1}u1R1n1N1ni=1r(2(ki+logpi(n)))n.\textsf{crank}_{{\mathbb{F}}_{p_{0}}}\{Y_{u_{0},u_{1}}\}_{u_{1}\in{\mathbb{P}}R_{1}^{n-1}}\geq N_{1}^{n}\prod\limits_{i=1}^{r}(2(k_{i}+\lceil\log_{p_{i}}(n)\rceil))^{-n}.
Proof.

As we are working with a matrix with entries in 𝔽p0{\mathbb{F}}_{p_{0}} the row space and column space of the matrix formed by concatenating {Yu0,u1}u1R1n1\{Y_{u_{0},u_{1}}\}_{u_{1}\in{\mathbb{P}}R_{1}^{n-1}} along the columns will be equal. The row space of Yu0,u1Y_{u_{0},u_{1}} is the space spanned by 𝟏y{\mathbf{1}}_{y} for yL1(u0,u1)y\in L_{1}(u_{0},u_{1}). If we consider all the rows from matrices in the set {Yu0,u1}d1R1n\{Y_{u_{0},u_{1}}\}_{d_{1}\in R_{1}^{n}} we will obtain vectors 𝟏y{\mathbf{1}}_{y} for all yd1R1n1L1(u0,u1)y\in\bigcup_{d_{1}\in{\mathbb{P}}R_{1}^{n-1}}L_{1}(u_{0},u_{1}). This means crank𝔽p0{Yu0,u1}u1R1n1\textsf{crank}_{{\mathbb{F}}_{p_{0}}}\{Y_{u_{0},u_{1}}\}_{u_{1}\in{\mathbb{P}}R_{1}^{n-1}} is exactly equal to the size of u1R1n1L1(u0,u1)\bigcup_{u_{1}\in{\mathbb{P}}R_{1}^{n-1}}L_{1}(u_{0},u_{1}) but that is a Kakeya set in (/N1)n({\mathbb{Z}}/N_{1}{\mathbb{Z}})^{n}. Finally, applying the induction hypothesis we are done. ∎

Claim 5.5.
crank𝔽p0{Mp0,n(u)}uR0()n1(p01+nn).\textsf{crank}_{{\mathbb{F}}_{p_{0}}}\{M_{p_{0}^{\ell},n}(u^{\prime})\}_{u^{\prime}\in{\mathbb{P}}R_{0}(\ell)^{n-1}}\geq\binom{p_{0}^{\ell}\ell^{-1}+n}{n}.
Proof.

This follows from Lemmas 3.6 and 3.7. ∎

Now applying Lemma 2.10 and the above two claims we see that

crank𝔽p0{Mp0,n(u)Yu0,u1}(u,u0,u1)J(N1ni=1r(2(ki+logpi(n)))n)(p01+nn).\textsf{crank}_{{\mathbb{F}}_{p_{0}}}\{M_{p_{0}^{\ell},n}(u^{\prime})\otimes Y_{u_{0},u_{1}}\}_{(u^{\prime},u_{0},u_{1})\in J}\geq\left(N_{1}^{n}\prod\limits_{i=1}^{r}(2(k_{i}+\lceil\log_{p_{i}}(n)\rceil))^{-n}\right)\binom{p_{0}^{\ell}\ell^{-1}+n}{n}.

Applying the above equation and Claims 5.2 and 5.3 we obtain the following bound on |S||S|.

|S|(b+n1n)(N1ni=1r(2(ki+logpi(n)))n)(p01+nn).|S|\binom{b+n-1}{n}\geq\left(N_{1}^{n}\prod\limits_{i=1}^{r}(2(k_{i}+\lceil\log_{p_{i}}(n)\rceil))^{-n}\right)\binom{p_{0}^{\ell}\ell^{-1}+n}{n}.

Recall, =k0+logp0(n)\ell=k_{0}+\lceil\log_{p_{0}}(n)\rceil and b=p0logp0(n)nb=p_{0}^{\lceil\log_{p_{0}}(n)\rceil}\geq n. Substituting these in the inequality above we get.

(N1ni=1r(2(ki+logpi(n)))n)1|S|\displaystyle\left(N_{1}^{n}\prod\limits_{i=1}^{r}(2(k_{i}+\lceil\log_{p_{i}}(n)\rceil))^{-n}\right)^{-1}|S| i=1np0k0p0logp0(n)(k0+logp0(n))1+ip0logp0(n)+i\displaystyle\geq\prod\limits_{i=1}^{n}\frac{p_{0}^{k_{0}}p_{0}^{\lceil\log_{p_{0}}(n)\rceil}(k_{0}+\lceil\log_{p_{0}}(n)\rceil)^{-1}+i}{p_{0}^{\lceil\log_{p_{0}}(n)\rceil}+i}
i=1np0k0(k0+logp0(n))1+ip0logp0(n)1+ip0logp0(n)\displaystyle\geq\prod\limits_{i=1}^{n}\frac{p_{0}^{k_{0}}(k_{0}+\lceil\log_{p_{0}}(n)\rceil)^{-1}+ip_{0}^{-\lceil\log_{p_{0}}(n)\rceil}}{1+ip_{0}^{-\lceil\log_{p_{0}}(n)\rceil}}
p0k0n(k0+logp0(n))ni=1n(1+ip0logp0(n))1.\displaystyle\geq\frac{p_{0}^{k_{0}n}}{(k_{0}+\lceil\log_{p_{0}}(n)\rceil)^{n}}\prod\limits_{i=1}^{n}\left(1+ip_{0}^{-\lceil\log_{p_{0}}(n)\rceil}\right)^{-1}.

For ini\leq n we have (1+ip0logp0(n))12\left(1+ip_{0}^{-\lceil\log_{p_{0}}(n)\rceil}\right)^{-1}\leq 2. This observation with the inequality above completes the proof.

Note for p0>np_{0}>n we have (1+ip0logp0(n))11+n/p0\left(1+ip_{0}^{-\lceil\log_{p_{0}}(n)\rceil}\right)^{-1}\leq 1+n/p_{0} and logp0(n)=1\lceil\log_{p_{0}}(n)\rceil=1. This with the inequality above and suitably modifying the induction hypothesis will give us the second half of this theorem. ∎

6 Constructing small Kakeya sets

Theorem 1.11.

Let s,ns,n\in{\mathbb{N}}, pp be a prime and k=(ps+11)/(p1)k=(p^{s+1}-1)/(p-1). There exists a Kakeya set SS in (/pk)n({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n} such that,

|S|i=1npkiki1(1p1)i1pknkn1(1p1)n.|S|\leq\sum\limits_{i=1}^{n}\frac{p^{ki}}{k^{i-1}(1-p^{-1})^{i-1}}\leq\frac{p^{kn}}{k^{n-1}(1-p^{-1})^{n}}.
Proof.

The construction will be inductive. For n=1n=1 it SS is the whole set. For a general nn we first construct a set which has lines with directions u={1,u2,,un}(/pk)n1u=\{1,u_{2},\ldots,u_{n}\}\in{\mathbb{P}}({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n-1}. We will need the following claim.

Claim 6.1.

There exists a function g:/pk/pkg:{\mathbb{Z}}/p^{k}{\mathbb{Z}}\rightarrow{\mathbb{Z}}/p^{k}{\mathbb{Z}} such that for any fixed t/pkt\in{\mathbb{Z}}/p^{k}{\mathbb{Z}} the function xtxg(x),x/pkx\rightarrow tx-g(x),x\in{\mathbb{Z}}/p^{k}{\mathbb{Z}} has an image of size at most pk/(k(1p1))p^{k}/(k(1-p^{-1})).

Proof.

We write an element u/pku\in{\mathbb{Z}}/p^{k}{\mathbb{Z}} in its pp-ary expansion a0+a1p+ak1pk1a_{0}+a_{1}p+\ldots a_{k-1}p^{k-1} where ai{0,,p1}a_{i}\in\{0,\ldots,p-1\}. gg will be of the form,

g(a0+a1p+ak1pk1)=j=0k1ajcjpj,g(a_{0}+a_{1}p+\ldots a_{k-1}p^{k-1})=\sum\limits_{j=0}^{k-1}a_{j}c_{j}p^{j},

where ci{0,,ps1}c_{i}\in\{0,\ldots,p^{s}-1\}. We write the number cic_{i} in its pp-ary form ci(0)+ci(1)p+ci(s1)ps1,ci(j){0,,p1}c_{i}(0)+c_{i}(1)p+\ldots c_{i}(s-1)p^{s-1},c_{i}(j)\in\{0,\ldots,p-1\} and represent cic_{i} with the tuple (ci(0),,ci(s1))(c_{i}(0),\ldots,c_{i}(s-1)).

We will set c0,c1,,cpk1c_{0},c_{1},\ldots,c_{p^{k}-1} iteratively. c0=ps1=(p1,,p1)c_{0}=p^{s}-1=(p-1,\ldots,p-1). Once we have set cic_{i}, if ci(s1)>0c_{i}(s-1)>0 we set ci+1=(ci(0),,ci(s1)1)c_{i+1}=(c_{i}(0),\ldots,c_{i}(s-1)-1). If ci(s1)=0c_{i}(s-1)=0 and ci0c_{i}\neq 0 then let α\alpha be the largest index such that ci(α)=0c_{i}(\alpha)=0 and ci(α1)0c_{i}(\alpha-1)\neq 0. We then set ci+1==ci+sα=cic_{i+1}=\ldots=c_{i+s-\alpha}=c_{i} and ci+s+1α=(ci(0),,ci(α1)1,p1,,p1)c_{i+s+1-\alpha}=(c_{i}(0),\ldots,c_{i}(\alpha-1)-1,p-1,\ldots,p-1). If ci=0c_{i}=0 then all the later cj,j>ic_{j},j>i are also set to 0. We note in this sequence of cic_{i}s each number in {1,,ps1}\{1,\ldots,p^{s}-1\} only appears at most as many times as the number of trailing zeros plus 1 in its pp-ary expansion. This means the number of zero equals,

ki=0s1(i+1)psi1(p1)=ps+11p1i=0s1(i+1)psi1(p1)=s+1.k-\sum\limits_{i=0}^{s-1}(i+1)p^{s-i-1}(p-1)=\frac{p^{s+1}-1}{p-1}-\sum\limits_{i=0}^{s-1}(i+1)p^{s-i-1}(p-1)=s+1.

The following property of cic_{i} quickly follows by the construction.

Statement 6.2.

Given a fixed t{0,,ps1}t^{\prime}\in\{0,\ldots,p^{s}-1\}, let β\beta be the smallest index such that cβ=tc_{\beta}=t^{\prime}. This construction guarantees that cβ+it=0(mod psi)c_{\beta+i}-t^{\prime}=0\text{(mod }p^{s-i}\text{)} for i={0,,s1}i=\{0,\ldots,s-1\}.

For any fixed tt consider tug(u)tv+g(v)tu-g(u)-tv+g(v) where u=a0+a1p+ak1pk1u=a_{0}+a_{1}p+\ldots a_{k-1}p^{k-1} and v=b0+b1p+bk1pk1v=b_{0}+b_{1}p+\ldots b_{k-1}p^{k-1},

tug(u)tv+g(v)=j=0k1(tcj)(ajbj)pj.\displaystyle tu-g(u)-tv+g(v)=\sum\limits_{j=0}^{k-1}(t-c_{j})(a_{j}-b_{j})p^{j}. (9)

Let t=t(mod ps)t^{\prime}=t\text{(mod }p^{s}\text{)} and β\beta be the smallest index such that cβ=tc_{\beta}=t^{\prime}. If ui=viu_{i}=v_{i} for i<βi<\beta then using Statement 6.2 on (9) we have,

(tug(u)tv+g(v))(mod ps+β)=0.\left(tu-g(u)-tv+g(v)\right)\text{(mod }p^{s+\beta}\text{)}=0.

This means after fixing the first β\beta coordinates of uu, tug(u)tu-g(u) can take at most pksβp^{k-s-\beta} values. This means the function tug(u)tu-g(u) can take at most pkspk/(k(1p1))p^{k-s}\leq p^{k}/(k(1-p^{-1})) many values. ∎

The set

Sn={(t,tu2g(u2),,tung(un))|t,u2,,un/pk}S_{n}=\{(t,tu_{2}-g(u_{2}),\ldots,tu_{n}-g(u_{n}))|t,u_{2},\ldots,u_{n}\in{\mathbb{Z}}/p^{k}{\mathbb{Z}}\}

contains a line in every direction {1,u2,,un}(/pk)n1\{1,u_{2},\ldots,u_{n}\}\in{\mathbb{P}}({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n-1}. By the claim above, for a fixed tt the function utug(u)u\rightarrow tu-g(u) will have an image of size at most pk/(k(1p1))p^{k}/(k(1-p^{-1})). This ensures SnS_{n} is of size at most pkn/(kn(1p1)n)p^{kn}/(k^{n}(1-p^{-1})^{n}). To add points with lines in other directions (0,u2,,un)(/pk)n1(0,u_{2},\ldots,u_{n})\in{\mathbb{P}}({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n-1} we simply need to add a Kakeya set in (/pk)n1({\mathbb{Z}}/p^{k}{\mathbb{Z}})^{n-1} which we can do using the induction hypothesis, completing the construction. ∎

Comment 6.3.

Let qq be a prime power, s,ns,n\in{\mathbb{N}} and k=qs1/(q1)k=q^{s}-1/(q-1). The construction above can be adapted to find Kakeya sets in (𝔽q[x]/xk)n({\mathbb{F}}_{q}[x]/\langle x^{k}\rangle)^{n} of size,

i=1nqkiki1(1q1)i1qknkn1(1q1)n.\sum\limits_{i=1}^{n}\frac{q^{ki}}{k^{i-1}(1-q^{-1})^{i-1}}\leq\frac{q^{kn}}{k^{n-1}(1-q^{-1})^{n}}.

Appendix A Proof of Lemma 3.6

We slightly sharpen the analysis of [1] here. We first show Mpk,1M_{p^{k},1} has an explicit decomposition as a product of lower and upper triangular matrices.

Lemma A.1 (Lemma 5 in [1]).

Let VmV_{m} for mm\in{\mathbb{N}} be an m×mm\times m matrix whose row and columns are labelled by elements in {0,,m1}\{0,\ldots,m-1\} such that its i,ji,jth entry is zij[z]z^{ij}\in{\mathbb{Z}}[z].

In this setting, there exists a lower triangular matrix LmL_{m} over [z]{\mathbb{Z}}[z] with ones on the diagonal such that its inverse is also lower triangular with entries in [z]{\mathbb{Z}}[z] with ones on the diagonal, and an upper triangular matrix DmD_{m} over [z]{\mathbb{Z}}[z] whose rows and columns are indexed by points in {0,,m1}\{0,\ldots,m-1\} such that the jjth diagonal entry for j{0,,m1}j\in\{0,\ldots,m-1\} equals

Dm(j,j)=i=0j1(zjzi)\displaystyle D_{m}(j,j)=\prod_{i=0}^{j-1}(z^{j}-z^{i})

such that Vm=LmDmV_{m}=L_{m}D_{m}.

This statement is precisely Lemma 5 in [1]. We will also need Lucas’s theorem from [16].

Theorem A.2 (Lucas’s Theorem [16]).

Let pp be a prime and Given two natural numbers aa and bb with expansions akpk++a1p+a0a_{k}p^{k}+\ldots+a_{1}p+a_{0} and bkpk++b0b_{k}p^{k}+\ldots+b_{0} in base pp we have,

(ab)(mod p)=i=0k(aibi)(mod p).\binom{a}{b}\text{(mod }p\text{)}=\prod\limits_{i=0}^{k}\binom{a_{i}}{b_{i}}\text{(mod }p\text{)}.

A particular consequence is that (ab)modp\binom{a}{b}\mod p is non-zero if and only if every digit in base-pp of bb is at most as large as every digit in base-pp of aa.

Lemma 3.6.

The 𝔽p{\mathbb{F}}_{p}-rank of Mp,nM_{p^{\ell},n} is at least

rank𝔽pMp,n(p1+nn).\text{rank}_{{\mathbb{F}}_{p}}M_{p^{\ell},n}\geq\binom{\lceil p^{\ell}\ell^{-1}\rceil+n}{n}.
Proof.

Using the previous lemma we note that Vp=LpDpV_{p^{\ell}}=L_{p^{\ell}}D_{p^{\ell}}. Under the ring map ff from [z]{\mathbb{Z}}[z] to [z]/zp1{\mathbb{Z}}[z]/\langle z^{p^{\ell}}-1\rangle, Vp=LpDpV_{p^{\ell}}=L_{p^{\ell}}D_{p^{\ell}} becomes Mp,1=L¯pD¯pM_{p^{\ell},1}=\overline{L}_{p^{\ell}}\overline{D}_{p^{\ell}} where L¯p\overline{L}_{p^{\ell}} and D¯p\overline{D}_{p^{\ell}} are the matrices f(Lp)f(L_{p^{\ell}}) and f(Dp)f(D_{p^{\ell}}) respectively.

We next notice that Mp,nM_{p^{\ell},n} is Mp,1M_{p^{\ell},1} tensored with itself nn times which we denote as Mp,n=Mp,1nM_{p^{\ell},n}=M_{p^{\ell},1}^{\otimes n}. Using Mp,1=L¯pD¯pM_{p^{\ell},1}=\overline{L}_{p^{\ell}}\overline{D}_{p^{\ell}} and Fact 2.8 we have Mp,n=Mp,1n=L¯pnD¯pnM_{p^{\ell},n}=M_{p^{\ell},1}^{\otimes n}=\overline{L}_{p^{\ell}}^{\otimes n}\overline{D}_{p^{\ell}}^{\otimes n}. As LpL_{p^{\ell}} was invertible with its inverse also having entries in [z]{\mathbb{Z}}[z] we see that L¯p\overline{L}_{p^{\ell}} is also invertible and (L¯pn)1Mp,n=D¯pn\left(\overline{L}_{p^{\ell}}^{\otimes n}\right)^{-1}M_{p^{\ell},n}=\overline{D}_{p^{\ell}}^{\otimes n}. Using Lemma 2.3 we have that,

rank𝔽pMp,nrank𝔽pD¯pn.\textsf{rank}_{{\mathbb{F}}_{p}}M_{p^{\ell},n}\geq\textsf{rank}_{{\mathbb{F}}_{p}}\overline{D}_{p^{\ell}}^{\otimes n}.

As D¯p\overline{D}_{p^{\ell}} is upper triangular so will D¯pn\overline{D}_{p^{\ell}}^{\otimes n} be. Therefore, to lower bound the rank of D¯pn\overline{D}_{p^{\ell}}^{\otimes n} we can lower bound the number of non-zero diagonal elements.

The diagonal elements in D¯pn\overline{D}_{p^{\ell}}^{\otimes n} correspond to the product of diagonal elements chosen from nn copies of D¯p\overline{D}_{p^{\ell}}. Recall, the rows and columns of D¯p\overline{D}_{p^{\ell}} are labelled by j{0,,p1}j\in\{0,\ldots,p^{\ell}-1\} with

Dm(j,j)=i=0j1(zjzi)\displaystyle D_{m}(j,j)=\prod_{i=0}^{j-1}(z^{j}-z^{i})

Setting z1=wz-1=w we note that 𝔽p[z]/zp1{\mathbb{F}}_{p}[z]/\langle z^{p^{\ell}}-1\rangle is isomorphic to 𝔽p[w]/wp{\mathbb{F}}_{p}[w]/\langle w^{p^{\ell}}\rangle. Dp(j,j)D_{p^{\ell}}(j,j) can now be written as

Dm(j,j)=(1+w)j(j1)/2i=1j((1+w)i1)\displaystyle D_{m}(j,j)=(1+w)^{j(j-1)/2}\prod_{i=1}^{j}((1+w)^{i}-1)

Using Lucas’s Theorem (Theorem A.2) we see that the largest power of tt which divides (1+w)l1(1+w)^{l}-1 is the same as the largest power of pp which divides ll. For any jp1j\leq p^{\ell}-1 , therefore the largest power of ww which divides D¯p(j,j)\overline{D}_{p^{\ell}}(j,j) is at most

t=0logp(j)(jptjpt+1)pt\displaystyle\sum\limits_{t=0}^{\lfloor\log_{p}(j)\rfloor}\left(\left\lfloor\frac{j}{p^{t}}\right\rfloor-\left\lfloor\frac{j}{p^{t+1}}\right\rfloor\right)p^{t} =j+t=1logp(j)jptpt1(p1)\displaystyle=j+\sum\limits_{t=1}^{\lfloor\log_{p}(j)\rfloor}\left\lfloor\frac{j}{p^{t}}\right\rfloor p^{t-1}(p-1)
j(1+logp(j)(11/p))\displaystyle\leq j(1+\lfloor\log_{p}(j)\rfloor(1-1/p))
j((1)/p).\displaystyle\leq j(\ell-(\ell-1)/p). (10)

Consider the set of tuples (j1,,jn)(j_{1},\ldots,j_{n})\in{\mathbb{N}} such that j1++jnp/j_{1}+\ldots+j_{n}\leq\lceil p^{\ell}/\ell\rceil. Using (10) we see that the diagonal entry in D¯pn\overline{D}_{p^{\ell}}^{\otimes n} corresponding to the tuple will be divisible by at most wp/((1)/p)w^{\lceil p^{\ell}/\ell\rceil(\ell-(\ell-1)/p)}. It is easy to check that p/((1)/p)p1\lceil p^{\ell}/\ell\rceil(\ell-(\ell-1)/p)\leq p^{\ell}-1 which will guarantee that the (j1,,jn)(j_{1},\ldots,j_{n})’th diagonal entry of D¯pn\overline{D}_{p^{\ell}}^{\otimes n} is non-zero.

This gives us at least (p1+nn)\binom{\lceil p^{\ell}\ell^{-1}\rceil+n}{n} non-zero diagonal entries proving the desired rank bound. ∎

References

  • [1] Bodan Arsovski, The p-adic Kakeya conjecture, arXiv preprint 2108.03750v1, 08-08-2021.
  • [2]  , The p-adic Kakeya conjecture, Journal of the American Mathematical Society, 37 (2023), 69-80.
  • [3] Boris Bukh and Ting-Wei Chao, Sharp density bounds on the finite field Kakeya, Discrete Analysis 2021, no. 26.
  • [4] Xavier Caruso, Almost all non-archimedean Kakeya sets have measure zero, Confluentes Mathematici 10 (2018), no. 1, 3–40 (en).
  • [5] Manik Dhar, Maximal Kakeya and (m,ϵ)(m,\epsilon)-Kakeya bounds over /N\mathbb{Z}/\text{{N}}\mathbb{Z} for general N, preprint arXiv:2209.11443 (2022).
  • [6] Manik Dhar and Zeev Dvir, Proof of the Kakeya set conjecture over rings of integers modulo square-free N, Combinatorial Theory 1 (2021).
  • [7] Manik Dhar, Zeev Dvir, and Ben Lund, Simple proofs for furstenberg sets over finite fields, Discrete Analysis (2021), no. 22 (en).
  • [8] Evan P. Dummit and Márton Hablicsek, Kakeya sets over non-archimedean local rings, Mathematika 59 (2013), no. 2, 257–266.
  • [9] Zeev Dvir, On the size of Kakeya sets in finite fields, Journal of the American Mathematical Society 22 (2009), no. 4, 1093–1097.
  • [10] Zeev Dvir, Swastik Kopparty, Shubhangi Saraf, and Madhu Sudan, Extensions to the method of multiplicities, with applications to Kakeya sets and mergers, SIAM Journal on Computing 42 (2013), no. 6, 2305–2328.
  • [11] Jordan S Ellenberg, Richard Oberlin, and Terence Tao, The Kakeya set and maximal conjectures for algebraic varieties over finite fields, Mathematika 56 (2010), no. 1, 1–25.
  • [12] Robert Fraser, Kakeya-type sets in local fields with finite residue field, Mathematika 62 (2016), no. 2, 614–629.
  • [13] G. H. Hardy and S. Ramanujan, The normal number of prime factors of a number nn, Quart. J. 48 (1917), 76–92 (English).
  • [14] G.H. Hardy and E.M. Wright, An introduction to the theory of numbers, 4th ed., 1975.
  • [15] Jonathan Hickman and James Wright, The Fourier restriction and Kakeya problems over rings of integers modulo N, Discrete Analysis (2018), no. 11.
  • [16] Edouard Lucas, Théorie des fonctions numériques simplement périodiques, American Journal of Mathematics 1 (1878), no. 4, 289–321.
  • [17] Shubhangi Saraf and Madhu Sudan, An improved lower bound on the size of Kakeya sets over finite fields, Anal. PDE 1 (2008), no. 3, 375–379.
  • [18] Thomas Wolff, Recent work connected with the Kakeya problem, Prospects in mathematics (Princeton,NJ, 1996) (1999), 29–162.
{aicauthors}{authorinfo}

[pgom] Manik Dhar
Massachusetts Institute of Technology

Cambridge, MA, USA
dmanik\imageatmit\imagedotedu
https://dharmanik.github.io