Weak hypergraph regularity and applications to geometric Ramsey theory

By Neil Lyall and Ákos Magyar

Abstract

Let , where with each a non-degenerate simplex of points. We prove that any set , with of positive upper Banach density necessarily contains an isometric copy of all sufficiently large dilates of the configuration . In particular any such set contains a -dimensional cube of side length , for all . We also prove analogous results with the underlying space being the integer lattice. The proof is based on a weak hypergraph regularity lemma and an associated counting lemma developed in the context of Euclidean spaces and the integer lattice.

1. Introduction

1.1. Existing results I: Distances and simplices in subsets of

Recall that the upper Banach density of a measurable set is defined by

where denotes Lebesgue measure on and denotes the cube .

A result of Furstenberg, Katznelson, and Weiss Reference 6 states that if has positive upper Banach density, then its distance set contains all sufficiently large numbers. Note that the distance set of any set of positive Lebesgue measure in automatically contains all sufficiently small numbers (by the Lebesgue density theorem) and that it is easy to construct a set of positive upper density which does not contain a fixed distance by placing small balls centered on an appropriate square grid.

Theorem A (Furstenberg, Katznelson, and Weiss Reference 6).

If with , then there exists a such that is guaranteed to contain pairs of points with for all .

This result was later reproved using Fourier analytic techniques by Bourgain in Reference 1 where he established the following more general result for all configurations of points in whose affine span is dimensional, namely for all non-degenerate simplices.

Theorem B (Bourgain Reference 1).

Let be a non-degenerate simplex of points. If with , then there exists a threshold such that contains an isometric copy of for all .

Recall that a finite point configuration is said to be an isometric copy of if there exists a bijection such that for all , i.e. if is obtained from (the dilation of by a factor ) via a rotation and translation.

Bourgain deduced Theorem B as an immediate consequence of the following stronger quantitative result for measurable subsets of the unit cube of positive measure. In Proposition C, and throughout this article, we shall refer to a decreasing sequence as lacunary if for all .

Proposition C (Bourgain Reference 1).

Let be a non-degenerate simplex of points. For any there exists a constant such that if is any lacunary sequence and with , then there exists such that contains an isometric copy of for all .

In Reference 12 the authors provided a short direct proof of Theorem B without using Proposition C. It is based on the observation that uniformly distributed sets contain the expected “number” of isometric copies of dilates and that all sets of positive upper density become uniformly distributed at sufficiently large scales. However, for the purposes of this paper it will be important to recall Bourgain’s indirect approach.

To see that Proposition C implies Theorem B notice that if Theorem B were not to hold for some set of upper Banach density , then there must exist a lacunary sequence , with the constant in Proposition C, such that does not contain an isometric copy of for any . Taking a sufficiently large cube with side length and and scaling back contradicts Proposition C.

We further note that by taking in Proposition C we obtain the following “Falconer-type” result for subsets of of positive Lebesgue measure.

Corollary D.

If is a non-degenerate simplex of points, then any with will necessarily contain an isometric copy of for all in some interval of length at least .

Bourgain further demonstrated in Reference 1 that no result along the lines of Theorem B can hold for configurations that contain any three points in arithmetic progression along a line, specifically showing that for any there are sets of positive upper Banach density in which do not contain an isometric copy of configurations of the form with for all sufficiently large . This should be contrasted with the following remarkable result of Tamar Ziegler.

Theorem E (Ziegler Reference 25).

Let be any configuration of points in with .

If has positive upper density, then there exists a threshold such that contains an isometric copy of for all and any , where denotes the -neighborhood of .

Bourgain’s example was later generalized by Graham Reference 9 to establish that the condition that in Theorem E is necessary and cannot be strengthened to for any given non-spherical configuration in for any , that is for any finite configuration of points that cannot be inscribed in some sphere. We note that the sets constructed by Bourgain and Graham have the property that for any their -neighborhoods will contain arbitrarily large cubes and hence trivially satisfy Theorem E with .

It is natural to ask if any spherical configuration , beyond the known example of simplices, has the property that every positive upper Banach density subset of , for some sufficiently large , contains an isometric copy of for all sufficiently large , and even to conjecture that this ought to hold for all spherical configurations. The first breakthrough in this direction came in Reference 12 when the authors established this for configurations of four points forming a -dimensional rectangle in and more generally for any configuration that is the direct product of two non-degenerate simplices in for suitably large .

The purpose of this article is to present a strengthening of the results in Reference 12 and to extend them to cover configurations with a higher dimensional product structure in both the Euclidean and discrete settings.

1.2. New results I: Rectangles and products of simplices in subsets of

The first main result of this article is the following

Theorem 1.1.

Let be points forming the vertices of a fixed -dimensional rectangle in .

(i)

If has positive upper Banach density, then there exists a threshold such that contains an isometric copy of for all .

(ii)

For any there exists a constant such that any with is guaranteed to contain an isometric copy of for all in some interval of length at least .

Moreover, if has sidelengths given by , …, , then the isometric copies of in both (i) and (ii) above can all be realized in the special form with each .

The multi-dimensional extension of Szemerédi’s theorem on arithmetic progressions in sets of positive density due to Furstenberg and Katznelson Reference 5 implies, and is equivalent to the fact, that there are isometric copies of in for arbitrarily large , with sides parallel to the coordinate axis. Theorem 1.1 states that there is an isometric copy of in for every sufficiently large , but only with sides parallel to given 2-dimensional coordinate subspaces which provides an extra degree of freedom for each side vector of the rectangle .

A weaker version of Theorem 1.1, with replaced with , was later established by Durcik and Kovač in Reference 4 using an adaptation of arguments of the second author with Cook and Pramanik in Reference 3. This approach also makes direct use of the full strength of the multi-dimensional Szemerédi theorem and as such leads to quantitatively weaker results.

Our arguments work for more general patterns where -dimensional rectangles are replaced with direct products of non-degenerate simplices.

Theorem 1.2.

Let , where and each is a non-degenerate simplex of points.

(i)

If has positive upper Banach density, then there exists a threshold such that contains an isometric copy of for all .

(ii)

For any there exists a constant such that any with is guaranteed to contain an isometric copy of for all in some interval of length at least .

Moreover the isometric copies of in both (i) and (ii) above can all be realized in the special form with each an isometric copy of .

Quantitative Remark. A careful analysis of our proof reveals that the constant can be taken greater than where is a tower of exponentials defined by and for .

1.3. Existing results II: Distances and simplices in subsets of

The problem of counting isometric copies of a given non-degenerate simplex in (with one vertex fixed) has been extensively studied via its equivalent formulation as the number of ways a quadratic form can be represented as a sum of squares of linear forms, see Reference 11 and Reference 19. This was exploited by the second author in Reference 16 and Reference 17 to establish analogous results to those described in Section 1.1 for subsets of the integer lattice of positive upper density.

Recall that the upper Banach density of a set is analogously defined by

where now denotes counting measure on and the discrete cube .

In light of the fact that any pairs of distinct points in have the property that the square of the distance between them is always a positive integer we introduce the convenient notation

Theorem A (Magyar Reference 16).

Let .

If has upper Banach density at least , then there exists an integer and such that contains pairs of points with for all with .

Theorem B (Magyar Reference 17).

Let and be a non-degenerate simplex of points.

(i)

If has upper Banach density at least , then there exists an integer and such that contains an isometric copy of for all with .

(ii)

If , then any ,…, with cardinality will necessarily contain an isometric copy of for some with .

Note that the fact that could fall entirely into a fixed congruence class of some integer ensures that the that appears in Theorems A and B must be divisible by the least common multiple of all integers . Indeed if with then has upper Banach density at least , however the distance between any two points is of the form for some .

However, in both Theorems A and Part (i) of Theorem B, one can take if the sets are assumed to be suitably uniformly distributed on congruence classes of small modulus. This leads via an easy density increment strategy to short new proofs, see Reference 14 for Theorem A and Section 8 for Part (i) of Theorem B.

The original argument in Reference 17 deduced Theorem B from the following discrete analogue of Proposition C.

Proposition C (Magyar Reference 17).

Let be a non-degenerate simplex of points.

For any there exist constants and such that if is any lacunary sequence in and , …, with cardinality , then will necessarily contain an isometric copy of for some .

To see that Proposition Cimplies Theorem B notice that if Part (i) of Theorem B were not to hold for some set of upper Banach density with from Proposition C, then there must exist a lacunary sequence in , with the constant from Proposition C, such that does not contain an isometric copy of for any . Since we can find a sufficiently large cube with integer side length that is divisible by and greater than such that , this contradicts Proposition C. Part (ii) of Theorem B follows from Proposition C by taking .

1.4. New results II: Rectangles and products of simplices in subsets of

We will also establish the following discrete analogues of Theorem 1.1 and 1.2.

Theorem 1.3.

Let and be points forming the vertices of a -dimensional rectangle in .

(i)

If has upper Banach density at least , then there exist integers and such that contains an isometric copy of for all with .

(ii)

There exists a constant such that if , then any , …, with cardinality will necessarily contain an isometric copy of for some with .

If has side lengths given by , …, , then each of the isometric copies in (i) and (ii) above can be realized in the form with each and , respectively.

Our arguments again work for more general patterns where -dimensional rectangles are replaced with direct products of non-degenerate simplices.

Theorem 1.4.

Let and , where and each is a non-degenerate simplex of points.

(i)

If has upper Banach density at least , then there exist integers and such that contains an isometric copy of for all with .

(ii)

There exists a constant such that if , then any , …, with cardinality will necessarily contain an isometric copy of for some with .

Moreover, each of the isometric copies in (i) and (ii) above can be realized in the special form with each an isometric copy of and , respectively.

Quantitative Remark. A careful analysis of our proof reveals that the constant (and consequently also ) can be taken less than where is a tower of exponentials defined by and for .

1.5. Notations and outline

We will consider the parameters , , …, fixed and will not indicate the dependence on them. Thus we will write if , …, . If the implicit constants in our estimates depend on additional parameters , , , …then we will write . We will use the notation to indicate that for some constant sufficiently small for our purposes.

Given an and a (finite or infinite) sequence , we will say that the sequence is -admissible if and for all . Moreover, if is given and for all , then we will call the sequence -admissible if in addition . Such sequences of scales will often appear in our statements both in the continuous and the discrete case.

Our proofs are based on a weak hypergraph regularity lemma and an associated counting lemma developed in the context of Euclidean spaces and the integer lattice. In Section 2 we introduce our approach in the model case of finite fields and prove an analogue of Theorem 1.1 in this setting. In Section 3 we review Theorem 1.2 for a single simplex and ultimately establish the base case of our general inductive approach to Theorem 1.2. In Section 4 we address Theorem 1.2 for the direct product of two simplices, this provides a new proof (and strengthening) of the main result of Reference 12 and serves as a gentle preparation for the more complicated general case which we present in the Section 5. The proof of Theorem 1.4 is outlined in Sections 6 and 7, while a short direct proof of Part (i) of Theorem B is presented in Section 8.

2. Model case: Vector spaces over finite fields

In this section we will illustrate our general method by giving a complete proof of Theorem 1.1 in the model setting of where denotes the finite field of elements. We do this as the notation and arguments are more transparent in this setting yet many of the main ideas are still present.

We say that two vectors are orthogonal, if , where stands for the usual dot product. A rectangle in is then a set with side vectors being pairwise orthogonal.

The finite field analogue of Theorem 1.1 is the following

Proposition 2.1.

For any there exists an integer with the following property:

If and , …, , then any with will contain points

where we have written with pairwise orthogonal coordinate subspaces.

2.1. Overview of the proof of Proposition 2.1

Write with pairwise orthogonal coordinate subspaces. For any , …, and we define

where we used the shorthand notation for each and the averaging notation:

for a finite set . We have also used the notation

for each . Note that the function may be viewed as the discrete analogue of the normalized surface area measure on the sphere of radius . It is well-known, see Reference 10, that

and for all one has

Note that if , then this implies that contains a rectangle of the form with and for .

Our approach to Proposition 2.1 in fact establishes the following quantitatively stronger result.

Proposition 2.2.

For any there exists an integer with the following property:

If , then for any and , …, one has

where we have written with pairwise orthogonal coordinate subspaces.

A crucial observation in the proof of Proposition 2.2 is that the averages can be compared to ones which can be easily estimated from below. We define, for any , the (unrestricted) count

It is easy to see, by carefully applying Cauchy-Schwarz times to , that

Our approach to Proposition 2.2 therefore reduces to establishing that for any one has

The validity of Equation 2.2 will follow immediately from the case of Proposition 2.3. However, before we can state this counting lemma we need to introduce some further notation from the theory of hypergraphs, notation that we shall ultimately make use of throughout the paper.

2.2. Hypergraph notation and a counting lemma

In order to streamline our notation we will make use the language of hypergraphs. For , …, and , we let denote the full -regular hypergraph on the vertex set . For we define the projection as and use this in turn to define the hypergraph bundle

using the shorthand notation , , …, to indicate that for all .

Notice when then consists of one element, the set , …, , and

Let and with pairwise orthogonal coordinate subspaces. For a given , , …, , with and a given edge , …, , we write

Note that the map defines a projection . With this notation, we can clearly now write

Now for any and any edge , i.e. , …, , , we let . For every and , we define where is the natural projection map.

Our key counting lemma, Proposition 2.3, which we will establish by induction on below, is then the statement that given a family of functions , the averages (generalizing those discussed above) which are defined by

are approximately equal. Specifically, one has

Proposition 2.3 (Counting lemma).

Let and . For any collection of functions

one has

If we apply this proposition with and for all , then Theorem 2.1 clearly follows given the lower bound Equation 2.1.

2.3. Proof of Proposition 2.3

We will establish Proposition 2.3 by inducting on .

For the result follows from the basic observation that if and let , then

by the properties of the function given above.

To see how this implies Proposition 2.3 for we note that since it follows that

The induction step has two main ingredients, the first is an estimate of the type which is often referred to as a generalized von-Neumann inequality, namely

Lemma 2.1.

Let . For any collection of functions with one has

where for any and we define

The corresponding inequality for the multi-linear expression , namely the fact that

is well-known and is referred to as the Gowers-Cauchy-Schwarz inequality Reference 8.

The second and main ingredient is an approximate decomposition of a graph to simpler ones, and is essentially the so-called weak (hypergraph) regularity lemma of Frieze and Kannan Reference 7. We choose to state this from a somewhat more abstract/probabilistic point of view, a perspective that will be particularly helpful when we consider our general results in the continuous and discrete settings.

We will first introduce this in the case . A bipartite graph with (finite) vertex sets , is a set and a function may be viewed as weighted bipartite graph with weights on the edges . If and are partitions of and respectively then is a partition and we let denote the function that is constant and equal to on each atom of . The weak regularity lemma states that for any and for any weighted graph there exist partitions of with for , so that

for all and . Informally this means that the graph can be approximated with precision with the “low complexity” graph . If we consider the -algebras generated by the partitions and the -algebra generated by then we have , the so-called conditional expectation function of . Moreover it is easy to see, using Cauchy-Schwarz, that estimate Equation 2.9 follows from

With this more probabilistic point of view the weak regularity lemma says that the function can be approximated with precision by a low complexity function , corresponding to -algebras on generated by sets. This formulation is also referred to as a Koopman von Neumann type decomposition, see Corollary 6.3 in Reference 23.

We will need a natural extension to -regular hypergraphs. See Reference 8Reference 22, and also Reference 2 for extension to sparse hypergraphs. Given an edge of elements we define its boundary . For each let be a -algebra on and denote its pull-back over the space . The -algebra is the smallest -algebra on containing for all . Note that the atoms of are of the form where is an atom of . We say that the complexity of a -algebra is at most , and write , if it is generated by sets.

Lemma 2.2 (Weak hypergraph regularity lemma).

Let and be a given function for each . For any there exists -algebras on for each such that

and

The proof of Lemmas 2.1 and 2.2 are presented in Section 2.4. We close this subsection by demonstrating how these lemmas can be combined to establish Proposition 2.3.

Proof of Proposition 2.3.

Let , and assume that the lemma holds for . It follows from Lemma 2.2 that there exists -algebras of complexity on for each for which Equation 2.12 holds for all . For each we let and write . By Lemma 2.1 and multi-linearity we have that

and also by the Gowers-Cauchy-Schwarz inequality

The conditional expectation functions are linear combinations of the indicator functions of the atoms of the -algebras . Since the number of terms in this linear combination is at most , with coefficients at most 1 in modulus, plugging these into the multi-linear expressions and one obtains a linear combination of expressions of the form and respectively with each being an atoms of for all .

The key observation is that these expressions are at level instead of . Indeed, where , with being an atom of when . If , …, , …, , let , …, , obtained from by removing the -entry. Then we have since , and hence

It therefore follows that

and similarly that

It then follows from the induction hypotheses that

for any . If we choose , with sufficiently large, then and it follows that

This, together with Equation 2.13 and Equation 2.14, establishes that Equation 2.5 hold for as required.

2.4. Proof of Lemmas 2.1 and 2.2

Proof of Lemma 2.1.

We start by observing the following consequence of Equation 2.6, namely that

for any and .

Now, fix an edge, say , , …, . Partition the edges into three groups: the first group consisting of edges for which , the second where and write with and the third when , using the notation , …, . Accordingly we can write

If for given and , , …, , we define

then we can write

By Equation 2.15 we can estimate the inner sum in Equation 2.17 by the square root of

Thus by Cauchy-Schwarz, and the fact that for all , we can conclude that

The expression on the right hand side of the inequality above is similar to that in Equation 2.16 except for the following changes. The functions for are eliminated i.e. replaced by 1, as well as the factor . The functions are replaced by for all . Repeating the same procedure for , …, one eliminates all the factors for , moreover all the functions for edges such that for some , which leaves only the edges so that , , …, , moreover for such edges the functions are eventually replaced by . The factors are not changed for however as the function does not depend on the variables for , averaging over these variables gives rise to a factor of . Thus one obtains the following final estimate

This proves the lemma, as it is clear that the above procedure can be applied to any edge in place of , , …, .

Proof of Lemma 2.2.

For a function and a -algebra on define the energy of with respect to as

and for a family of functions and -algebras , its total energy as

We will show that if Equation 2.12 does not hold for a family of -algebras , then the -algebras can be refined so that the total energy of the system increases by a quantity depending only on . Since the functions are bounded the total energy of the system is , the energy increment process must stop in steps, and Equation 2.12 must hold. The idea of this procedure appears already in the proof of Szemerédi’s regularity lemma Reference 20, and has been used since in various places Reference 7Reference 8Reference 22.

Initially set and hence to be the trivial -algebras. Assume that in general Equation 2.12 does not hold for a family of -algebras , with . Then there exists an edge so that , with . Let , …, for simplicity of notation, hence , …, . Then, with notation , …, , one has

for some functions that are bounded by 1 in magnitude. Indeed if and edge , …, then does not depend at least one of the variables . Thus there must be an for which the inner sum in the above expression is at least . Fix such an . Decomposing the functions into their positive and negative parts and then writing them as an average of indicator functions, one obtains that there sets such that

which can be written more succinctly, using the inner product notation, as

For let be the -algebra generated by and the set and let . Since the functions are measurable with respect to the -algebra for all , we have that

and hence, by Cauchy-Schwarz, that

Note that the first equality above follows from the fact that conditional expectation function is the orthogonal projection of to the subspace of -measurable functions in . This also implies that energy of a function is always increasing when the underlying -algebra is refined, and Equation 2.22 tells us that the energy of is increased by at least .

For we set . Then the total energy of the family with respect to the system , is also increased by at least .

It is clear that the complexity of the -algebras is increased by at most 1, hence, as explained above, the lemma follows by applying this energy increment process at most times.

3. The base case of an inductive strategy to establish Theorem 1.2

In this section we will ultimately establish the base case of our more general inductive argument. We however start by giving a quick review of the proof of Theorem 1.2 when (which contains both Theorem B and Corollary D as stated in Section 1.1), namely the case of a single simplex. This was originally addressed in Reference 1 and revisited in Reference 12 and Reference 13.

3.1. A single simplex in

Let be a fixed cube and let denotes its side length.

Let , , …, be a fixed non-degenerate simplex and define for where is the dot product on . Given , a simplex , , …, is isometric to if and only if for all . Thus the configuration space of isometric copies of is a non-singular real variety given by the above equations. Let be natural normalized surface area measure on , described in Reference 1, Reference 12, and Reference 13. It is clear that the variable can be replaced by any of the variables by redefining the constants .

For any family of functions , …, and we define the multi-linear expression

We note that all of our functions are 1-bounded and both integrals, in fact all integrals in this paper are normalized. Recall that we are using the normalized integral notation . Since the normalized measure is supported on we will not indicate the support of the variables , …, explicitly.

Note also that if is a measurable set and , …, then must contain an isometric copy of . Proposition 3.1 (with ) is a quantitatively stronger version of Proposition C that appeared in Section 1.1 and hence immediately establishes Theorem 1.2 for .

Proposition 3.1.

For any there exists an integer with the following property:

Given any lacunary sequence and , there is some such that

for all .

Our approach to establishing Proposition 3.1 is to compare the above expressions to simpler ones for which it is easy to obtain lower bounds. Given a scale we define the multi-linear expression

where and is the shift of the cube by the vector . Note that if is a set of measure for some , then for a given , Hölder implies

for all scales .

Recall that for any we call a sequence -admissible if and for all . Note that given any lacunary sequence with , one can always finds an -admissible sequence of scales such that for each the interval contains at least two consecutive elements from the original lacunary sequence.

In light of this observation, and the one above regarding a lower bound for , …, , our proof of Proposition 3.1 reduces to establishing the following “counting lemma”.

Proposition 3.2.

Let . There exists an integer such that for any -admissible sequence of scales and there is some such that

for all .

There are two main ingredients in the proof of Proposition 3.2, this will be typical to all of our arguments. The first ingredient is a result which establishes that the multi-linear forms , …, are controlled by an appropriate norm which measures the uniformity of distribution of functions with respect to particular scales . This is analogous to estimates in additive combinatorics Reference 8 which are often referred to as generalized von-Neumann inequalities.

The result below was proved in Reference 12 for , however a simple scaling of the variables transfers the result to an arbitrary cube .

Lemma 3.1 (A generalized von-Neumann inequality Reference 12).

Let , , and .

For any collections of functions , …, we have

where for any we define

with denoting the shift of the cube by the vector .

The corresponding inequality for the multi-linear expression , …, , namely the fact that

whenever follows easily from Cauchy-Schwarz together with the simple observation that

whenever .

The second key ingredient, proved in Reference 13 and generalized in Lemma 3.3, is a Koopman-von Neumann type decomposition of functions where the underlying -algebras are generated by cubes of a fixed length. To recall it, let be a cube, be scale that divides , , and denote the collection of cubes partitioning the cube and denote the grids corresponding to the centers of the cubes. By a slightly abuse of notation we also write for the -algebra generated by the grid. Recall that the conditional expectation function is constant and equal to on each cube .

Lemma 3.2 (A Koopman-von Neumann type decomposition Reference 13).

Let and be a cube.

There exists an integer such that for any -admissible sequence and function there is some such that

Proof of Proposition 3.2.

Let be the grid obtained from Lemma 3.2 for the functions for some fixed . Let , then by Equation 3.6 and multi-linearity, we have

and also

provided for . Thus in showing Equation 3.5 one can replace the functions with . If we make the additional assumption that then it is easy to see, using the fact that the function is constant on the cubes , that

Since the condition can be replaced with if one passes to a subsequence of scales, for example , this completes the proof of Proposition 3.2.

3.2. The base case of a general inductive strategy

In this section, as preparation to handle the case of products of simplices, we prove a parametric version of Proposition 3.2, namely Proposition 3.3, which will serve as the base case for later inductive arguments.

Let with be cubes of equal side length . Let be a scale dividing and for each , …, let and . Note that . Here and for each .

Let , …, be a non-degenerate simplex for each .

Proposition 3.3 (Parametric counting lemma on for simplices).

Let and . There exists an integer such that for any -admissible sequence of scales with the property that divides and collection of functions

there exists and a set of size such that

for all and uniformly in and .

The proof of Proposition 3.3 will follow from Lemma 3.1 and the following generalization of Lemma 3.2 in which we simultaneously consider a family of functions supported on the subcubes in a partition of an original cube .

Lemma 3.3 (A simultaneous Koopman-von Neumann type decomposition).

Let , , and be a cube. There exists an integer such that for any -admissible sequence with the property that divides , and collection of functions

defined for each , there is some and a set of size such that

for all and .

Proof of Proposition 3.3.

Fix . For and , …, , we will abuse notation and write

for , …, .

If we apply Lemma 3.3 to the family of functions on for , , and , with , then this produces a grid for some , and a set of size , such that

uniformly for , and for .

Since , …, for , …, it is easy to see that

Let , then by Lemma 3.1, one has

and

for all provided . Finally, if we also have then it is easy to see that

as the functions are constant on cubes of , which are of size .

Passing first to a subsequence of scales, for example , the condition can be replaced with so this completes the proof of the proposition.

We conclude this section with a sketch of the proof of Lemma 3.3. These arguments are standard, see for example the proof of Lemma 3.2 given in Reference 12.

Proof of Lemma 3.3.

First we make an observation about the -norm. Suppose with dividing . If and then and hence

for any function . Moreover, since the cube is partitioned into the smaller cubes , we have by Cauchy-Schwarz

From these observations it is easy to see that

and we note that the right side of the above expression is since the conditional expectation function is constant and equal to on the cubes .

Suppose that Equation 3.10 does not hold for some for every in some set of size . If we apply the above observation to , for every , we obtain by orthogonality that

for some constant .

If we now define such that , for , average over , and use the fact , we obtain

It is clear that the sums in the above expressions are bounded by for all , thus Equation 3.11 cannot hold for some for . This implies that Equation 3.10 must hold for some , for all and all for a set of size .

4. Product of two simplices in

Although not strictly necessary, we discuss in this section the special case of Theorem 1.2. This already gives an improvement of the main results of Reference 12, but more importantly serves as a gentle preparation for the more complicated general case, presented in the Section 5, which involve both a plethora of different scales and the hypergraph bundle notation introduced in Section 2.2.

4.1. Proof of Theorem 1.2 with

Let with and be cubes of equal side length and with , …, and , …, two non-degenerate simplices.

In order to “count” configurations of the form with and isometric copies of and respectively for some in a set we introduce the multi-linear expression

for any family of functions with and .

Indeed, if for all and then the above expression is 0 unless there exists a configuration of the form with and isometric copies of and respectively.

The short argument presented in Section 1.1 demonstrating how both Theorem B and Corollary D follow from Proposition B, and hence from Proposition 3.1, applies equally well to each of our main theorems. This reduces our main theorems to analogous quantitative results involving an arbitrary lacunary sequence of scales. In the case of Theorem 1.2 this stronger quantitative result takes the following form:

Proposition 4.1.

For any there exists an integer with the following property:

Given any lacunary sequence and , there is some such that

for all .

Our approach to establishing Proposition 4.1 is again to compare the above expressions to simpler ones for which it is easy to obtain lower bounds. For any and family of functions with and we consider

where , , …, and for .

Note that if is a set of measure for some , then careful applications of Hölder’s inequality give

for all scales .

In light of the observation above, and the discussion preceding Proposition 3.2, we see that Proposition 4.1, and hence Theorem 1.2 when , will follow as a consequence of the following

Proposition 4.2.

Let . There exists an integer such that for any -admissible sequence of scales and there is some such that

for all .

There are again two main ingredients in the proof of Proposition 4.2. The first establishes that the our multi-linear forms are controlled by an appropriate box-type norm attached to a scale .

Let be a cube. For any scale and function we define its local box norm at scale to be

where and

for any cube of the form with for .

Lemma 4.1 (A generalized von-Neumann inequality Reference 12).

Let , , and .

For any collections of functions with and we have both

The result above was essentially proved in Reference 12 for the multi-linear forms when , however a simple scaling argument transfers the result to an arbitrary cube . For completeness we include its short proof in Section 4.2.

The second and main ingredient is an analogue of a weak form of Szemerédi’s regularity lemma due to Frieze and Kannan Reference 7. The more probabilistic formulation, we will use below, can be found for example in Reference 21, Reference 22, and Reference 23, and is also sometimes referred to as a Koopman-von Neumann type decomposition.

For any cube and scale that divides we will let and denote the collection of cubes partitioning the cube and let denote grid corresponding to the centers of these cubes. We will say that a finite -algebra on is of scale if it contains and for simplicity of notation will write for .

Recall that if we have two -algebras on a cube and on then by we mean the -algebra on generated by the sets with and . Recall also that we say the complexity of a -algebra is at most , and write , if it is generated by sets.

Lemma 4.2 (Weak regularity lemma in ).

Let and with and be cubes of equal side length .

There exists an integer such that for any -admissible sequence and function there is some and a -algebra of scale on such that

which has the additional local structure that for each there exist -algebras on and on with for such that .

Comparing the above statement to Lemma 2.2 for , i.e to the weak regularity lemma, note that the -algebra of scale has a direct product structure only locally, inside each cube . Moreover this product structure varies with , however the “local complexity” remains uniformly bounded.

Assuming for now the validity of Lemmas 4.1 and 4.2 we prove Proposition 4.2. We will make crucial use of Proposition 3.3, namely our parametric counting lemma on for simplices.

Proof of Proposition 4.2.

Let , for some , and be an -admissible sequence of scales. Set and be the parameter appearing in Proposition 3.3, noting that .

For write if . We now choose a subsequence so that and . Applying Lemma 4.2, with for all and , guarantees the existence of a -algebra of scale on such that

for some . Moreover, we know that has the additional local structure that for each there exist -algebras on and on with for such that . Thus, if we let and denote the number of atoms in and respectively, then we can assume, by formally adding the empty set to these collections of atoms if necessary, that for all .

If we let , then by Lemma 4.1 and multi-linearity we have

provided for . For a given write for the restriction of to the cube . By localization, one then has

and

provided one also insists that .

For given , the functions are linear combinations of functions of the form , where and are the collections of the atoms of the -algebras and defined on the cubes and . Thus for each one has

where . Plugging these linear expansions into the multi-linear expressions in above one obtains

using the notations , . Notice that the product

is nonzero only if , that is if for all , as the atoms are all disjoint. Similarly, one has that for all . Thus, in fact

and similarly

Note, that indices are running through the index set of size at most if .

The key observation is that

and

Let and , . Writing and , one may apply Proposition 3.3 for the families of functions , where and , with respect to the -admissible sequence of scales

This is possible as . Then there is a scale with so that

and

for all uniformly in and , for a set of size . Then, by Equation 4.14Equation 4.15 and Equation 4.12Equation 4.13, we have

for , as and . Finally, since , by averaging in , one has

using Equation 4.10Equation 4.11 and the Proposition follows by Equation 4.9 with an index .

4.2. Proof of Lemmas 4.1 and 4.2

Proof of Lemma 4.1.

First we note that if and , then

and hence for any function , with being a cube of side length , one has

Write , …, and let . Then one may write

Using estimate Equation 3.6, the above observation, and Cauchy-Schwarz one has

provided and where for . Applying the same procedure again ultimately gives

The same estimate can of course be given for any function in place of . This establishes Equation 4.5. Estimate Equation 4.6 is established similarly.

Proof of Lemma 4.2.

For each we will let and , in other words the trivial -algebras on and respectively. If Equation 4.8 holds with , noting that in this case, then we are done.

We now assume that we have developed, for each , -algebras on and on with for . Let be the -algebra such that for all and assume that Equation 4.8 does not hold, namely that

where . By the definition of the local box norm this means that

and hence, as , it is easy to see that

This implies that there is a set of size such that for all , one has that . It therefore follows, as is well-known see for example Reference 12 or Reference 23, that there exist sets and such that

For a given there is a unique such that . Let and noting that for , as the complexity of a -algebra does not increase when restricted to a set. If, for , we let denote the -algebra generated by and the set if and let otherwise, then clearly is at most . We now define to the the sigma algebra of scale with the property that for all .

Using the inner product notation we can rewrite Equation 4.19 as

for all . Since the function is measurable with respect to one clearly has

and hence

It then follows from Cauchy-Schwarz and orthogonality that

Since averaging over all gives

Trivially both sides are at most thus the process must stop at a step where Equation 4.8 holds for a -algebra of “local complexity” at most . This proves the lemma.

5. Proof of Theorem 1.2: The general case

After these preparations we will now consider the general case of Theorem 1.2. Let with cubes of equal side length and with each a non-degenerate simplex of points for .

We will use a generalized version of the hypergraph terminology introduced in Section 2. In particular, for a vertex set , , …, and set we will let denote the projection defined by . As before we will let denote the complete -regular hypergraph with vertex set , and for the multi-index , …, define the hypergraph bundle

noting that for all .

In order to parameterize the vertices of direct products of simplices, i.e. sets of the form with , we consider points , …, with , …, for each . Now for any and any edge we will write , and for every and we define , where is the natural projection map. Writing , …, we have that since every edge is of the form , …, . We can therefore identify points with configurations of the form .

For any the measures , introduced in Section 3.1, are supported on points , …, for which the simplex , , …, is isometric to . For simplicity of notation we will write

Note that the support of the measure is the set of points so that the simplex , …, is isometric to and , moreover the measure is normalized. Thus if is a set then the density of configurations in of the form with each an isometric copy of is given by the expression

The proof of Theorem 1.2 reduces to establishing the following stronger quantitative result.

Proposition 5.1.

For any there exists an integer with the following property:

Given any lacunary sequence and , there is some such that

for all .

Quantitative remark.

A careful analysis of our proof reveals that there is a choice of which is less than , where is again the tower-exponential function defined by and for .

For any and set we define the expression:

where and

for any cube of the form with for . Note that if is a set of measure for some , then careful applications of Hölder’s inequality give

for all scales .

In light of the discussion above, and that preceding Proposition 3.2, we see that Proposition 5.1, and hence Theorem 1.2 in general, will follow as a consequence of the following

Proposition 5.2.

Let . There exists an integer such that for any -admissible sequence of scales and there is some such that

for all .

The validity of Proposition 5.2 will follow immediately from the case of Proposition 5.3.

5.1. Reduction of Proposition 5.2 to a more general “local” counting lemma

For any given and collection of functions with we define the following multi-linear expressions

and

where and

for any cube of the form with for .

Our strategy to proving Proposition 5.2 is the same as illustrated in the finite field settings, that is we would like to compare averages to those of , at certain scales , inductively for . However in the Euclidean case, an extra complication emerges due to the fact the (hypergraph) regularity lemma, the analogue of Lemma 2.2, does not produce -algebras , for , on the cubes . In a similar manner to the case for discussed in the previous section, we will only obtain -algebras “local” on cubes at some scale . This will have the effect that the functions will be replaced by a family of functions , where runs through a grid .

To be more precise, let be a scale dividing the side-length . For and we will use to denote the projection of onto and to denote the projection of the cube centered at onto . It is then easy to see that for any we have

and

provided where denotes the restriction of a function to the cube .

At this point the proof of Proposition 5.2 reduces to showing that the expressions in Equation 5.9 and Equation 5.10 only differ by at some scales , given an -admissible sequence , for any collection of bounded functions , , . Indeed, our crucial result will be the following

Proposition 5.3 (Local counting lemma).

Let and . There exists an integer such that for any -admissible sequence of scales with the property that divides , and collection of functions

there exists and a set of size such that

for all and uniformly in and .

5.2. Proof of Proposition 5.3

We will prove Proposition 5.3 by induction on . For this is basically Proposition 3.3.

Indeed, in this case for a given , …, and edge we have that with and hence both

By Proposition 3.3 there exists an and an exceptional set of size , such that uniformly for and for , one has

hence

as the all factors are trivially bounded by 1 in magnitude. This implies Equation 5.11 for .

For the induction step we again need two main ingredients. The first establishes that the our multi-linear forms are controlled by an appropriate box-type norm attached to a scale .

Let and . For any scale and function with we define its local box norm at scale by

where

for any cube of the form .

Lemma 5.1 (Generalized von-Neumann inequality).

Let , and .

For any and collection of functions with we have both

The crucial ingredient is the following analogue of the weak hypergraph regularity lemma.

Lemma 5.2 (Parametric weak hypergraph regularity lemma for ).

Let , , and .

There exists such that for any -admissible sequence with the property that divides and collection of functions

there is some and -algebras of scale on for each and such that

uniformly for all , , and , where with .

Moreover, the -algebras have the additional local structure that the exist -algebras on with for each , , and such that if , then

Lemma 5.2 is the parametric and simultaneous version of the extension of Lemma 4.2 to the product of simplices. The difference is that in the general case one has to deal with a parametric family of functions as is running through a grid . The essential new content of Lemma 5.2 is that one can develop -algebras on the cubes with respect to the family of functions such that the local structure described above and Equation 5.16 hold simultaneously for almost all .

Proof of Proposition 5.3.

Assume the proposition holds for .

Let , for some large constant , and be an -admissible sequence of scales. Set with .

For we again write if . We now choose a subsequence so that and . Lemma 5.2 then guarantees the existence of -algebras of scale on for each and , with the local structure described above, such that

uniformly for all , , and , for some , where with . Let for and . If , then by Equation 5.14, Equation 5.15, and Equation 5.16 we have both

provided . For given one may write for the restriction of on the cube , as uniquely determines . By localization, provided , we then have both

For a fixed cube we have that

where is the family of atoms of the -algebra restricted to the cube . Note that and . By adding the empty set to the collection of atoms one may assume for all and . Then, by multi-linearity, using the notations and , one has both

The key observation is that these expressions in the sum above are all at level instead of . To see this let , …, , …, so , …, , …, . If then recall that the edge , …, is obtained from by removing the -entry. Thus, for any atom of we have by Equation 5.17, that

where is an atom of the -algebra . Thus

It follows that

and hence that

and similarly

Note that number of index vectors is with and hence if .

Writing and it then follows from our inductive hypothesis functions, applied with respect to the -admissible sequence of scales

which is possible as , that there is a scale with so that

for all uniformly in for , where is a set of size .

Since the cubes form a partition of as runs through the grid the relative density of the set can substantially increase only of a few cubes . Indeed, it is easy to see that for the set

We claim that Equation 5.11 holds for uniformly in , , and . Indeed, from Equation 5.29, Equation 5.30, and Equation 5.31 and the fact that , it follows

for since . Finally, the fact that together with localization, namely Equation 5.21 and Equation 5.22, ensures that averaging over gives

which in light of Equation 5.19, Equation 5.20, and the fact that complete the proof.

5.3. Proof of Lemmas 5.1 and 5.2

Proof of Lemma 5.1.

The argument is similar to that of Lemma 2.1. Fix an edge, say , , …, , and partition the edges in to as follows. Let be the set of those edges for which , and for , …, let denote the collection of edges of the form , , …, , in other words if for some edge , …, . Accordingly write

For and , …, with , define

Then one may write

For the inner integrals we have, using Equation 3.6, the estimate

provided , where as in the proof of Lemma 4.1 we use the notation

with for . By Cauchy-Schwarz we then have

where with for .

The expression we have obtained above is similar to the one in Equation 5.6 except for the following changes. The variable is replaced by and the measure by . The functions are replaced by , for , while the functions for all such that are eliminated, that is replaced by 1. Repeating the same procedure for , …, replaces all variables with variables as well as the measures with . The procedure eliminates all functions when is an edge such that for some ; for the remaining edges, when , …, , it replaces the functions with . For the variables and the measures are not changed, however integrating in these variables will have no contribution as the measures are normalized. Thus one obtains the following final estimate

noting that these integrals are not normalized. Thus, one may write the expression in Equation 5.34, using a change of variables , , as

where the last equality follows from the facts that the function is supported on the cube and hence the integration in is restricted to the cube , giving rise an error of . Estimate Equation 5.14 follows from Equation 5.34 and Equation 5.35 noting that the above procedure can be applied to any in place of . Estimate Equation 5.15 is established similarly.

Proof of Lemma 5.2.

For we set and for , , and . We will develop -algebras of scale such that Equation 5.17 holds with .

We define the total energy of a family of functions with respect to a family of -algebras as

Since for all , , and it follows that the total energy is bounded by . Our strategy will be to show that if Equation 5.16 does not hold then there exist a family of -algebras such that the total energy of the family of functions is increased by at least with respect to this new family of -algebras, and at the same time ensuring that Equation 5.17 remains valid with . This iterative process must stop at some proving the lemma.

Assume that we have developed -algebras and of scale such that Equation 5.17 holds with . If Equation 5.16 does not hold then for the set

Fix and let and be such that

and write . Consider the partition of the cube into small cubes where . By the localization properties of the -norm, and the fact that we have that

for any function . Thus there exists a set of size

such that

for all .

For a given cube and functions , define the normalized inner product of and as

Then by the well-known property of the -norm, see for example Reference 23 or the proof of Lemma 2.2, it follows from Equation 5.37 that there exists sets

for such that

If then there is a unique such that . If and then we define the -algebras on as follows. Write where and let be the -algebra generated by the set and the -algebra restricted to where is the unique element so that . Note that that the complexity of the -algebra is at most one larger than the complexity of the -algebra as restricting a -algebra to a set does not increase its complexity. If or then let be simply the restriction of to the cube , or equivalently define the sets . Finally, let

be the corresponding -algebra on the cube .

Since the cubes partition the cube as runs through the grid , these -algebras define a -algebra on , such that its restriction to the cubes is equal to the -algebras .

Since the function is measurable with respect to the -algebra restricted to the cube one clearly has

and hence, by Equation 5.38, that

It then follows from Cauchy-Schwarz and orthogonality, using the fact that the -algebra is a refinement of , that

for . Since averaging over implies

At this point we have shown that if then there exists an edge , , and -algebras of scale on , with , such that Equation 5.43 holds.

For all with let be the restriction of the -algebra to the cube , where is such that . By Equation 5.39 this implies that is also the restriction of to the cube , and hence the -algebra is generated by the grid and the -algebra .

We have therefore defined a family of the -algebras for , satisfying

Using the fact that and averaging over it follows using the notations of Equation 5.36 that

As the total energy is bounded by , the process must stop at a step where Equation 5.16 holds for a -algebra of “local complexity” at most , completing the proof of Lemma 5.2.

6. The base case of an inductive strategy to establish Theorem 1.4

In this section we will ultimately establish the base case of our more general inductive argument. We will however start by giving a (new) proof of Theorem B, namely the case of Theorem 1.4.

6.1. A single simplex in

Let , , …, be a fixed non-degenerate simplex of points in with and define for . Recall, see Reference 17, that a simplex , …, is isometric to if and only if for all .

For any positive integer and we define , …, be the function whose value is 1 if with both and in for all and is equal to 0 otherwise. It is a well-known fact in number theory, see Reference 11 or Reference 17, that for we have that

for some absolute constant and some constant , the so-called singular series, which can be interpreted as the product of the densities of the solutions of the above system of equations among the -adics and among the reals. Thus if we define

then is normalized in so much that

for some absolute constant .

Let be a fixed cube and let denotes its side length. For any family of functions

and we define the following two multi-linear expressions

and

where . Note that if and , …, then must contain an isometric copy of , while if for some then as before Hölder implies that

for all scales with .

Recall that for any and positive integer we call a sequence -admissible if and for all and . Note that if is any lacunary sequence in with , one can always finds an -admissible sequence of scales with the property that for each the interval contains at least two consecutive elements from the original lacunary sequence.

In light of these observations we see that the following “counting lemma” ultimately establishes a quantitatively stronger version of Proposition B that appeared in Section 1.3 and hence immediately establishes Theorem 1.4 for .

Proposition 6.1.

Let and for with .

There exists such that for any -admissible sequence of scales and there is some such that

for all with .

As in the continuous setting the proof of Proposition 6.1 has two main ingredients, namely Lemmas 6.1 and 6.2. In these lemmas, and for the remainder Sections 6 and 7, we will continue to use the notation

for any given .

Lemma 6.1 (A generalized von Neumann inequality).

Let , with , and with and . For any collection of functions , …, we have

where for any function we define

with denoting the normalized characteristic function of the cubes .

For any cube of side length and satisfying with dividing , we shall now partition into cubic grids , with as usual. These grids form the atoms of a -algebra . Note that if and then .

Lemma 6.2 (A Koopman-von Neumann type decomposition).

Let and for all . There exists an integer such that any -admissible sequence of scales and function there is some such that

The reduction of Proposition 6.1 to these two lemmas is essentially identical to the analogous argument in the continuous setting as presented at the end of Section 3.1, we choose to omit the details.

Proof of Lemma 6.1.

We will rely on some prior exponential sum estimates, specifically Propositions 4.2 and 4.4 in Reference 17. First we deal with the case . By the change of variables for , one may write

We now write

where , , …, and for each , …, we are using denote the (essentially) normalized indicator function of the subset of that contains if and only if for all .

Using the fact that , together with Cauchy-Schwarz and Plancherel, one can then easily see that

with

It then follows by Propositions 4.2 and 4.4 in Reference 17, with and after rescaling by , that in addition to being non-negative and uniformly bounded in we in fact have

for all .

We note that the expression may be interpreted as the Fourier transform of the indicator function of the set of integer points on a certain variety, and estimate Equation 6.9 indicates that this concentrates near rational points of small denominator. It is this crucial fact from number theory which makes results like Theorem B possible.

Since

it is easy to see that for all and that there exists some absolute constant such that

for all and . It is then easy to see using our assumption that that

for some constant uniformly in provided . Substituting inequality Equation 6.7 into Equation 6.8, we obtain

provided . This proves Lemma 6.1 for , as it is clear that by re-indexing the above estimate holds for any of the functions in place of . For an easy modification of arguments in Reference 14, specifically the proof of Lemma 3 therein, establishes that

for provided .

Proof of Lemma 6.2.

Let such that , . The “modulo grids partition the cube with running through the set , …, , where denote the centers of the “integer” grids in an initial partition of . Let be positive integers so that , and . If and then and hence

for any function . Moreover, since the cube is partitioned into the smaller cubes , we have by Cauchy-Schwarz

From this it is easy to see that

and we note that the right side of the above expression is since the conditional expectation function is constant and equal to on the cubes .

Now suppose Equation 6.7 does not hold for some , that is

Since , , and we can apply the above observations to and obtain, by orthogonality, that

for some constant . Since the above expressions are clearly bounded by , the above procedure must stop in steps at which Equation 6.7 must hold for some with .

6.2. The base case of our general inductive strategy

Let with be cubes of equal side length and be a non-degenerate simplex of points for .

We note that for any and scale dividing if , …, , then the corresponding grids in the partition of take the form .

As in the continuous setting we will ultimately need a parametric version of Proposition 6.1, namely Proposition 6.2.

Proposition 6.2 (Parametric counting lemma on for simplices).

Let and .

There exists an integer such that for any -admissible sequence of scales with dividing and for with , and collection of functions

there exists and a set of size such that

for all with and uniformly in and .

This proposition follows, as the analogous result did in the continuous setting, from Lemma 6.1 and the following parametric version of Lemma 6.2.

Lemma 6.3 (A simultaneous Koopman-von Neumann type decomposition).

Let , , and be a cube. There exists an integer such that for any -admissible sequence with dividing and for with , and collection of functions

defined for each , there is some and a set of size such that

for all and .

Lemma 6.3 is of course the discrete analogue of Lemma 3.2. Since the proofs of Proposition 6.2 and Lemma 6.3 are almost identical to the arguments presented in Section 3.2 we choose to omit these details.

7. Proof of Theorem 1.4: The general case

After the preparations in Section 6 we can proceed very similarly as in Section 5 to prove our main result in the discrete case, namely Theorem 1.4. The main difference will be that given and , we construct a positive integer and assume that all our sequences of scales will be -admissible. The cubes will be naturally now be replaced by the grids of the form that already appear in Section 6 where we always assume .

Let with each a non-degenerate simplex of points for and with cubes of equal side length (taken much larger than the diameter of ). We will use the same parameterizations in terms of hypergraph bundles and corresponding notations as in Section 5 to count the configurations with each an isometric copy of for some .

Given any positive integer and we will make use of the notation

with as defined in the previous section and , …, .

Note that if then the density of configurations in , of the form with each an isometric copy of for some is given by the expression

More generally, for any given and a family of functions with we define the multi-linear expression

as well as

where with each and

for any cube of the form with for .

We note that it is easy to show, as in the continuous, that if with for some then

for all scales with . In light of this observation and the discussion preceding Proposition 6.1 the proof of Theorem 1.4 reduces, as it did in the continuous setting, to the following

Proposition 7.1.

Let . There exist positive integers and such that for any -admissible sequence of scales and there is some such that

for all with with .

Quantitative Remark. A careful analysis of our proof reveals that there exist choices of and which are less than and respectively where is again the tower-exponential function defined by and for .

The proof of Proposition 7.1 follows along the same lines as the analogous result in the continuous setting. As before we will compare the averages to those of , at certain scales and with with , inductively for . As the arguments closely follow those given in Section 5 we will be brief and emphasize mainly just the additional features.

7.1. Reduction of Proposition 7.1 to a more general “local” counting lemma

For any given and a family of functions with it is easy to see that for any , scale dividing the side-length , and we have

and

provided where denotes the restriction of a function to the cube .

Thus the proof of Proposition 7.1 reduces to showing that the expressions in Equation 7.8 and Equation 7.9 only differ by for all scales with , given an -admissible sequence , for any collection of bounded functions , , . Indeed, our crucial result will be the following

Proposition 7.2 (Local counting lemma in ).

Let and .

There exist positive integers and such that for any -admissible sequence of scales with dividing and for , and collection of functions

there exists and a set of size such that

for all with and uniformly in and .

Note that if , , , then , and moreover if for all for a set , then Proposition 7.2 reduces to precisely Proposition 7.1. In fact, Proposition 7.2 is a parametric, multi-linear and simultaneous extension of Proposition 7.1 which we need in the induction step, i.e. when going from level to level .

7.2. Proof of Proposition 7.2

We will prove Proposition 7.2 by induction on .

For this is basically Proposition 6.2, exactly as it was in the base case of the proof of Proposition 5.3.

For the induction step we will again need two main ingredients. The first establishes that the our multi-linear forms are controlled by a box-type norm attached to scales and .

Let with be cubes of equal side length and . For any scale and function with we define its local box norm at scales and by

where

for any cube of the form . We note that Equation 7.4 and Equation 7.5 are special cases of Equation 7.11 and Equation 7.12 with , , …, , and for all .

Lemma 7.1 (A generalized von-Neumann inequality on ).

Let .

Let , with , and with and . For any collection of functions with we have both

and

The proof of inequalities Equation 7.13 and Equation 7.14 follow exactly as in the continuous case, see Lemma 5.1, using Lemma 6.1 in place of Lemma 3.1. We omit the details.

The crucial ingredient is again a parametric weak hypergraph regularity lemma, i.e. Lemma 5.2 adapted to the discrete settings. The proof is essentially the same as in the continuous case, with exception that the -norms are replaced by -norms where is a given sequence of positive integers and is an -admissible sequence of scales. To state it we say that a -algebra on a cube is of scale if it is refinement of the grid , i.e. if its atoms partition each cube of the grid. We will always assume that and . Recall also that we say the complexity of a -algebra is at most , and write , if it is generated by sets.

Lemma 7.2 (Parametric weak hypergraph regularity lemma for ).

Let , , , and let for . There exists such that for any -admissible sequence with the property that divides and collection of functions

there is some and -algebras of scale on for each and such that

uniformly for all , , and , where with .

Moreover, the -algebras have the additional local structure that the exist -algebras on with for each , , and such that if , then

The proof of Lemma 7.2 follows exactly as the corresponding proof of Lemma 5.2 in the continuous setting, so we will omit the details. We will however provide some details of how one deduces Proposition 7.2, from Lemmas 7.1 and 7.2. The arguments are again very similar to those in the continuous setting, however one needs to make a careful choice of the integers , appearing in the statement of the Proposition.

Proof of Proposition 7.2.

Let and assume that the lemma holds for .

Let and for some large constant .

We then define recalling that and note that it is easy to see by induction that for and . We further define the function with and recall that for .

We now proceed exactly as in the proof of Proposition 5.3 but with being a -admissible sequence of scales, with . We again choose a subsequence so that and , but also now set , where . Lemma 7.2 then guarantees the existence of -algebras of scale on for each and , with the local structure described above, such that Equation 7.15 holds uniformly for all , , and , for some , where with .

Arguing as in the proof of Proposition 5.3 we can conclude from this that for each we have

and

provided with , where each and number of index vectors is with and hence if .

By induction, we apply Proposition 7.2 to the sequence of scales with and for where with respect to the family of functions . This is possible as and our sequence of scales is -admissible. Thus there exists an index such that for all with we have

uniformly in for , where is a set of size .

The remainder of the proof follows as just as it did for Proposition 5.3.

8. Appendix: A short direct proof of part (i) of Theorem B

We conclude by providing a short direct proof of Part (i) of Theorem B, namely the following

Theorem 8.1 (Magyar Reference 17).

Let and be a non-degenerate simplex of points.

If has upper Banach density at least , then there exists an integer and such that contains an isometric copy of for all with .

For any we define

with a (sufficiently) large absolute constant. Following Reference 14 we further define to be -uniformly distributed (modulo ) if its relative upper Banach density on any residue class modulo never exceeds times its density on , namely if

for all , …, . It turns out that this notion is closely related to the -norm introduced in Section 6. Recall that for any cube and function we define

with denoting the normalized characteristic function of the cubes . Note that the -norm measures the mean square oscillation of a function with respect to cubic grids of size and gap .

The following observation from Reference 14 (specifically Lemmas 1 and 2) is key to our short proof of Theorem 8.1.

Lemma 8.1.

Let . If be -uniformly distributed with , then there exists an integer and cubes of arbitrarily large side length with such that

Let , , …, be a fixed non-degenerate simplex of points in with and define for . We now define a function which counts isometric copies of .

Recall, see Reference 17, that a simplex , …, is isometric to if and only if for all . For any we define , …, be the function whose value is 1 if for all and is equal to 0 otherwise. It is a well-known fact in number theory, see Reference 11 or Reference 17, that for we have that

for some absolute constant and constant , the so-called singular series, which can be interpreted as the product of the densities of the solutions of the above system of equations among the -adics and among the reals. Thus if we define

then is normalized in so much that

for some absolute constant .

Let be a fixed cube and let denotes its side length. For any family of functions

and we define

It is clear that if restricted to , then the above expression is a normalized count of the isometric copies of in . Thus, Theorem 8.1 will follow from Lemma 8.1 and the following special case (with ) of Lemma 6.1.

Lemma 8.2 (A generalized von Neumann inequality).

Let .

If with and then for any collection of functions , …, we have

This compares with the purely number theoretic fact that the number of simplices , , …, isometric to is asymptotic to . Thus, under the same conditions as in Lemma 8.2, we have

provided one also has .

Proof of Theorem 8.1.

Let and be a set of upper Banach density .

We assume first that is -uniformly distributed. Select a scale and a sufficiently large cube so that the conclusion of Lemma 8.1 holds. For a given with and write and substitute this decomposition into the multi-linear expression , …, . Then by Lemma 8.2 and Equation 8.3Equation 8.4, we have that

and we can conclude that must contain an isometric copy of .

If is not -uniformly distributed, then its upper Banach density is increased to at least when restricted to a residue class . Identify with and simultaneously the set with a set , via the map . Note that if is -uniformly distributed then it contains an isometric copy of for all sufficiently large and hence contains an isometric copy of .

Repeating the above procedure one arrives to a set for some in steps which contains an isometric copy of for all sufficiently large .

Table of Contents

  1. Abstract
  2. 1. Introduction
    1. 1.1. Existing results I: Distances and simplices in subsets of
    2. Theorem A (Furstenberg, Katznelson, and Weiss 6).
    3. Theorem B (Bourgain 1).
    4. Proposition C (Bourgain 1).
    5. Corollary D.
    6. Theorem E (Ziegler 25).
    7. 1.2. New results I: Rectangles and products of simplices in subsets of
    8. Theorem 1.1.
    9. Theorem 1.2.
    10. 1.3. Existing results II: Distances and simplices in subsets of
    11. Theorem A (Magyar 16).
    12. Theorem B (Magyar 17).
    13. Proposition C (Magyar 17).
    14. 1.4. New results II: Rectangles and products of simplices in subsets of
    15. Theorem 1.3.
    16. Theorem 1.4.
    17. 1.5. Notations and outline
  3. 2. Model case: Vector spaces over finite fields
    1. Proposition 2.1.
    2. 2.1. Overview of the proof of Proposition 2.1
    3. Proposition 2.2.
    4. 2.2. Hypergraph notation and a counting lemma
    5. Proposition 2.3 (Counting lemma).
    6. 2.3. Proof of Proposition 2.3
    7. Lemma 2.1.
    8. Lemma 2.2 (Weak hypergraph regularity lemma).
    9. 2.4. Proof of Lemmas 2.1 and 2.2
  4. 3. The base case of an inductive strategy to establish Theorem 1.2
    1. 3.1. A single simplex in
    2. Proposition 3.1.
    3. Proposition 3.2.
    4. Lemma 3.1 (A generalized von-Neumann inequality 12).
    5. Lemma 3.2 (A Koopman-von Neumann type decomposition 13).
    6. 3.2. The base case of a general inductive strategy
    7. Proposition 3.3 (Parametric counting lemma on for simplices).
    8. Lemma 3.3 (A simultaneous Koopman-von Neumann type decomposition).
  5. 4. Product of two simplices in
    1. 4.1. Proof of Theorem 1.2 with
    2. Proposition 4.1.
    3. Proposition 4.2.
    4. Lemma 4.1 (A generalized von-Neumann inequality 12).
    5. Lemma 4.2 (Weak regularity lemma in ).
    6. 4.2. Proof of Lemmas 4.1 and 4.2
  6. 5. Proof of Theorem 1.2: The general case
    1. Proposition 5.1.
    2. Proposition 5.2.
    3. 5.1. Reduction of Proposition 5.2 to a more general “local” counting lemma
    4. Proposition 5.3 (Local counting lemma).
    5. 5.2. Proof of Proposition 5.3
    6. Lemma 5.1 (Generalized von-Neumann inequality).
    7. Lemma 5.2 (Parametric weak hypergraph regularity lemma for ).
    8. 5.3. Proof of Lemmas 5.1 and 5.2
  7. 6. The base case of an inductive strategy to establish Theorem 1.4
    1. 6.1. A single simplex in
    2. Proposition 6.1.
    3. Lemma 6.1 (A generalized von Neumann inequality).
    4. Lemma 6.2 (A Koopman-von Neumann type decomposition).
    5. 6.2. The base case of our general inductive strategy
    6. Proposition 6.2 (Parametric counting lemma on for simplices).
    7. Lemma 6.3 (A simultaneous Koopman-von Neumann type decomposition).
  8. 7. Proof of Theorem 1.4: The general case
    1. Proposition 7.1.
    2. 7.1. Reduction of Proposition 7.1 to a more general “local” counting lemma
    3. Proposition 7.2 (Local counting lemma in ).
    4. 7.2. Proof of Proposition 7.2
    5. Lemma 7.1 (A generalized von-Neumann inequality on ).
    6. Lemma 7.2 (Parametric weak hypergraph regularity lemma for ).
  9. 8. Appendix: A short direct proof of part (i) of Theorem B
    1. Theorem 8.1 (Magyar 17).
    2. Lemma 8.1.
    3. Lemma 8.2 (A generalized von Neumann inequality).

Mathematical Fragments

Theorem B (Bourgain Reference 1).

Let be a non-degenerate simplex of points. If with , then there exists a threshold such that contains an isometric copy of for all .

Proposition C (Bourgain Reference 1).

Let be a non-degenerate simplex of points. For any there exists a constant such that if is any lacunary sequence and with , then there exists such that contains an isometric copy of for all .

Corollary D.

If is a non-degenerate simplex of points, then any with will necessarily contain an isometric copy of for all in some interval of length at least .

Theorem E (Ziegler Reference 25).

Let be any configuration of points in with .

If has positive upper density, then there exists a threshold such that contains an isometric copy of for all and any , where denotes the -neighborhood of .

Theorem 1.1.

Let be points forming the vertices of a fixed -dimensional rectangle in .

(i)

If has positive upper Banach density, then there exists a threshold such that contains an isometric copy of for all .

(ii)

For any there exists a constant such that any with is guaranteed to contain an isometric copy of for all in some interval of length at least .

Moreover, if has sidelengths given by , …, , then the isometric copies of in both (i) and (ii) above can all be realized in the special form with each .

Theorem 1.2.

Let , where and each is a non-degenerate simplex of points.

(i)

If has positive upper Banach density, then there exists a threshold such that contains an isometric copy of for all .

(ii)

For any there exists a constant such that any with is guaranteed to contain an isometric copy of for all in some interval of length at least .

Moreover the isometric copies of in both (i) and (ii) above can all be realized in the special form with each an isometric copy of .

Theorem A (Magyar Reference 16).

Let .

If has upper Banach density at least , then there exists an integer and such that contains pairs of points with for all with .

Theorem B (Magyar Reference 17).

Let and be a non-degenerate simplex of points.

(i)

If has upper Banach density at least , then there exists an integer and such that contains an isometric copy of for all with .

(ii)

If , then any ,…, with cardinality will necessarily contain an isometric copy of for some with .

Proposition C (Magyar Reference 17).

Let be a non-degenerate simplex of points.

For any there exist constants and such that if is any lacunary sequence in and , …, with cardinality , then will necessarily contain an isometric copy of for some .

Theorem 1.4.

Let and , where and each is a non-degenerate simplex of points.

(i)

If has upper Banach density at least , then there exist integers and such that contains an isometric copy of for all with .

(ii)

There exists a constant such that if , then any , …, with cardinality will necessarily contain an isometric copy of for some with .

Moreover, each of the isometric copies in (i) and (ii) above can be realized in the special form with each an isometric copy of and , respectively.

Proposition 2.1.

For any there exists an integer with the following property:

If and , …, , then any with will contain points

where we have written with pairwise orthogonal coordinate subspaces.

Proposition 2.2.

For any there exists an integer with the following property:

If , then for any and , …, one has

where we have written with pairwise orthogonal coordinate subspaces.

Equation (2.1)
Equation (2.2)
Proposition 2.3 (Counting lemma).

Let and . For any collection of functions

one has

Equation (2.6)
Lemma 2.1.

Let . For any collection of functions with one has

where for any and we define

Equation (2.9)
Lemma 2.2 (Weak hypergraph regularity lemma).

Let and be a given function for each . For any there exists -algebras on for each such that

and

Equation (2.13)
Equation (2.14)
Equation (2.15)
Equation (2.16)
Equation (2.17)
Equation (2.22)
Proposition 3.1.

For any there exists an integer with the following property:

Given any lacunary sequence and , there is some such that

for all .

Proposition 3.2.

Let . There exists an integer such that for any -admissible sequence of scales and there is some such that

for all .

Lemma 3.1 (A generalized von-Neumann inequality Reference 12).

Let , , and .

For any collections of functions , …, we have

where for any we define

with denoting the shift of the cube by the vector .

Lemma 3.2 (A Koopman-von Neumann type decomposition Reference 13).

Let and be a cube.

There exists an integer such that for any -admissible sequence and function there is some such that

Proposition 3.3 (Parametric counting lemma on for simplices).

Let and . There exists an integer such that for any -admissible sequence of scales with the property that divides and collection of functions

there exists and a set of size such that

for all and uniformly in and .

Lemma 3.3 (A simultaneous Koopman-von Neumann type decomposition).

Let , , and be a cube. There exists an integer such that for any -admissible sequence with the property that divides , and collection of functions

defined for each , there is some and a set of size such that

for all and .

Equation (3.11)
Proposition 4.1.

For any there exists an integer with the following property:

Given any lacunary sequence and , there is some such that

for all .

Proposition 4.2.

Let . There exists an integer such that for any -admissible sequence of scales and there is some such that

for all .

Lemma 4.1 (A generalized von-Neumann inequality Reference 12).

Let , , and .

For any collections of functions with and we have both

Lemma 4.2 (Weak regularity lemma in ).

Let and with and be cubes of equal side length .

There exists an integer such that for any -admissible sequence and function there is some and a -algebra of scale on such that

which has the additional local structure that for each there exist -algebras on and on with for such that .

Equation (4.8)
Equation (4.9)
Equation (4.10)
Equation (4.11)
Equation (4.12)
Equation (4.13)
Equation (4.14)
Equation (4.15)
Equation (4.19)
Proposition 5.1.

For any there exists an integer with the following property:

Given any lacunary sequence and , there is some such that

for all .

Proposition 5.2.

Let . There exists an integer such that for any -admissible sequence of scales and there is some such that

for all .

Equation (5.6)
Equation (5.9)
Equation (5.10)
Proposition 5.3 (Local counting lemma).

Let and . There exists an integer such that for any -admissible sequence of scales with the property that divides , and collection of functions

there exists and a set of size such that

for all and uniformly in and .

Lemma 5.1 (Generalized von-Neumann inequality).

Let , and .

For any and collection of functions with we have both

Lemma 5.2 (Parametric weak hypergraph regularity lemma for ).

Let , , and .

There exists such that for any -admissible sequence with the property that divides and collection of functions

there is some and -algebras of scale on for each and such that

uniformly for all , , and , where with .

Moreover, the -algebras have the additional local structure that the exist -algebras on with for each , , and such that if , then

Equations (5.19), (5.20)
Equations (5.21), (5.22)
Equation (5.29)
Equation (5.30)
Equation (5.31)
Equation (5.34)
Equation (5.35)
Equation (5.36)
Equation (5.37)
Equation (5.38)
Equation (5.39)
Equation (5.43)
Proposition 6.1.

Let and for with .

There exists such that for any -admissible sequence of scales and there is some such that

for all with .

Lemma 6.1 (A generalized von Neumann inequality).

Let , with , and with and . For any collection of functions , …, we have

where for any function we define

with denoting the normalized characteristic function of the cubes .

Lemma 6.2 (A Koopman-von Neumann type decomposition).

Let and for all . There exists an integer such that any -admissible sequence of scales and function there is some such that

Equation (6.8)
Equation (6.9)
Proposition 6.2 (Parametric counting lemma on for simplices).

Let and .

There exists an integer such that for any -admissible sequence of scales with dividing and for with , and collection of functions

there exists and a set of size such that

for all with and uniformly in and .

Lemma 6.3 (A simultaneous Koopman-von Neumann type decomposition).

Let , , and be a cube. There exists an integer such that for any -admissible sequence with dividing and for with , and collection of functions

defined for each , there is some and a set of size such that

for all and .

Equation (7.4)
Equation (7.5)
Proposition 7.1.

Let . There exist positive integers and such that for any -admissible sequence of scales and there is some such that

for all with with .

Equation (7.8)
Equation (7.9)
Proposition 7.2 (Local counting lemma in ).

Let and .

There exist positive integers and such that for any -admissible sequence of scales with dividing and for , and collection of functions

there exists and a set of size such that

for all with and uniformly in and .

Equation (7.11)
Equation (7.12)
Lemma 7.1 (A generalized von-Neumann inequality on ).

Let .

Let , with , and with and . For any collection of functions with we have both

and

Lemma 7.2 (Parametric weak hypergraph regularity lemma for ).

Let , , , and let for . There exists such that for any -admissible sequence with the property that divides and collection of functions

there is some and -algebras of scale on for each and such that

uniformly for all , , and , where with .

Moreover, the -algebras have the additional local structure that the exist -algebras on with for each , , and such that if , then

Theorem 8.1 (Magyar Reference 17).

Let and be a non-degenerate simplex of points.

If has upper Banach density at least , then there exists an integer and such that contains an isometric copy of for all with .

Lemma 8.1.

Let . If be -uniformly distributed with , then there exists an integer and cubes of arbitrarily large side length with such that

Lemma 8.2 (A generalized von Neumann inequality).

Let .

If with and then for any collection of functions , …, we have

Equation (8.4)

References

Reference [1]
J. Bourgain, A Szemerédi type theorem for sets of positive density in , Israel J. Math. 54 (1986), no. 3, 307–316, DOI 10.1007/BF02764959. MR853455,
Show rawAMSref \bib{Bourg87}{article}{ author={Bourgain, J.}, title={A Szemer\'{e}di type theorem for sets of positive density in ${\mathbf {R}}^k$}, journal={Israel J. Math.}, volume={54}, date={1986}, number={3}, pages={307--316}, issn={0021-2172}, review={\MR {853455}}, doi={10.1007/BF02764959}, }
Reference [2]
David Conlon, Jacob Fox, and Yufei Zhao, A relative Szemerédi theorem, Geom. Funct. Anal. 25 (2015), no. 3, 733–762, DOI 10.1007/s00039-015-0324-9. MR3361771,
Show rawAMSref \bib{CFZ}{article}{ author={Conlon, David}, author={Fox, Jacob}, author={Zhao, Yufei}, title={A relative Szemer\'{e}di theorem}, journal={Geom. Funct. Anal.}, volume={25}, date={2015}, number={3}, pages={733--762}, issn={1016-443X}, review={\MR {3361771}}, doi={10.1007/s00039-015-0324-9}, }
Reference [3]
Brian Cook, Ákos Magyar, and Malabika Pramanik, A Roth-type theorem for dense subsets of , Bull. Lond. Math. Soc. 49 (2017), no. 4, 676–689, DOI 10.1112/blms.12043. MR3725488,
Show rawAMSref \bib{CMP}{article}{ author={Cook, Brian}, author={Magyar, \'{A}kos}, author={Pramanik, Malabika}, title={A Roth-type theorem for dense subsets of $\mathbb {R}^d$}, journal={Bull. Lond. Math. Soc.}, volume={49}, date={2017}, number={4}, pages={676--689}, issn={0024-6093}, review={\MR {3725488}}, doi={10.1112/blms.12043}, }
Reference [4]
Polona Durcik and Vjekoslav Kovač, Boxes, extended boxes and sets of positive upper density in the Euclidean space, Math. Proc. Cambridge Philos. Soc. 171 (2021), no. 3, 481–501, DOI 10.1017/S0305004120000316. MR4324955,
Show rawAMSref \bib{DK}{article}{ author={Durcik, Polona}, author={Kova\v {c}, Vjekoslav}, title={Boxes, extended boxes and sets of positive upper density in the Euclidean space}, journal={Math. Proc. Cambridge Philos. Soc.}, volume={171}, date={2021}, number={3}, pages={481--501}, issn={0305-0041}, review={\MR {4324955}}, doi={10.1017/S0305004120000316}, }
Reference [5]
H. Furstenberg and Y. Katznelson, An ergodic Szemerédi theorem for commuting transformations, J. Analyse Math. 34 (1978), 275–291 (1979), DOI 10.1007/BF02790016. MR531279,
Show rawAMSref \bib{FK78}{article}{ author={Furstenberg, H.}, author={Katznelson, Y.}, title={An ergodic Szemer\'{e}di theorem for commuting transformations}, journal={J. Analyse Math.}, volume={34}, date={1978}, pages={275--291 (1979)}, issn={0021-7670}, review={\MR {531279}}, doi={10.1007/BF02790016}, }
Reference [6]
Hillel Furstenberg, Yitzchak Katznelson, and Benjamin Weiss, Ergodic theory and configurations in sets of positive density, Mathematics of Ramsey theory, Algorithms Combin., vol. 5, Springer, Berlin, 1990, pp. 184–198, DOI 10.1007/978-3-642-72905-8_13. MR1083601,
Show rawAMSref \bib{FKW86}{article}{ author={Furstenberg, Hillel}, author={Katznelson, Yitzchak}, author={Weiss, Benjamin}, title={Ergodic theory and configurations in sets of positive density}, conference={ title={Mathematics of Ramsey theory}, }, book={ series={Algorithms Combin.}, volume={5}, publisher={Springer, Berlin}, }, date={1990}, pages={184--198}, review={\MR {1083601}}, doi={10.1007/978-3-642-72905-8\_13}, }
Reference [7]
Alan Frieze and Ravi Kannan, The regularity lemma and approximation schemes for dense problems, 37th Annual Symposium on Foundations of Computer Science (Burlington, VT, 1996), IEEE Comput. Soc. Press, Los Alamitos, CA, 1996, pp. 12–20, DOI 10.1109/SFCS.1996.548459. MR1450598,
Show rawAMSref \bib{FK96}{article}{ author={Frieze, Alan}, author={Kannan, Ravi}, title={The regularity lemma and approximation schemes for dense problems}, conference={ title={37th Annual Symposium on Foundations of Computer Science}, address={Burlington, VT}, date={1996}, }, book={ publisher={IEEE Comput. Soc. Press, Los Alamitos, CA}, }, date={1996}, pages={12--20}, review={\MR {1450598}}, doi={10.1109/SFCS.1996.548459}, }
Reference [8]
W. T. Gowers, Hypergraph regularity and the multidimensional Szemerédi theorem, Ann. of Math. (2) 166 (2007), no. 3, 897–946, DOI 10.4007/annals.2007.166.897. MR2373376,
Show rawAMSref \bib{Gow06}{article}{ author={Gowers, W. T.}, title={Hypergraph regularity and the multidimensional Szemer\'{e}di theorem}, journal={Ann. of Math. (2)}, volume={166}, date={2007}, number={3}, pages={897--946}, issn={0003-486X}, review={\MR {2373376}}, doi={10.4007/annals.2007.166.897}, }
Reference [9]
R. L. Graham, Recent trends in Euclidean Ramsey theory, Discrete Math. 136 (1994), no. 1-3, 119–127, DOI 10.1016/0012-365X(94)00110-5. Trends in discrete mathematics. MR1313284,
Show rawAMSref \bib{Graham90}{article}{ author={Graham, R. L.}, title={Recent trends in Euclidean Ramsey theory}, note={Trends in discrete mathematics}, journal={Discrete Math.}, volume={136}, date={1994}, number={1-3}, pages={119--127}, issn={0012-365X}, review={\MR {1313284}}, doi={10.1016/0012-365X(94)00110-5}, }
Reference [10]
A. Iosevich and M. Rudnev, Erdős distance problem in vector spaces over finite fields, Trans. Amer. Math. Soc. 359 (2007), no. 12, 6127–6142, DOI 10.1090/S0002-9947-07-04265-1. MR2336319,
Show rawAMSref \bib{IR07}{article}{ author={Iosevich, A.}, author={Rudnev, M.}, title={Erd\H {o}s distance problem in vector spaces over finite fields}, journal={Trans. Amer. Math. Soc.}, volume={359}, date={2007}, number={12}, pages={6127--6142}, issn={0002-9947}, review={\MR {2336319}}, doi={10.1090/S0002-9947-07-04265-1}, }
Reference [11]
Y. Kitaoka, Lectures on Siegel modular forms and representation by quadratic forms, Tata Institute of Fundamental Research Lectures on Mathematics and Physics, vol. 77, Published for the Tata Institute of Fundamental Research, Bombay; by Springer-Verlag, Berlin, 1986, DOI 10.1007/978-3-662-00779-2. MR843330,
Show rawAMSref \bib{Kitaoka}{book}{ author={Kitaoka, Y.}, title={Lectures on Siegel modular forms and representation by quadratic forms}, series={Tata Institute of Fundamental Research Lectures on Mathematics and Physics}, volume={77}, publisher={Published for the Tata Institute of Fundamental Research, Bombay; by Springer-Verlag, Berlin}, date={1986}, pages={vi+227}, isbn={3-540-16472-3}, review={\MR {843330}}, doi={10.1007/978-3-662-00779-2}, }
Reference [12]
Neil Lyall and Ákos Magyar, Product of simplices and sets of positive upper density in , Math. Proc. Cambridge Philos. Soc. 165 (2018), no. 1, 25–51, DOI 10.1017/S0305004117000184. MR3811544,
Show rawAMSref \bib{LM17}{article}{ author={Lyall, Neil}, author={Magyar, \'{A}kos}, title={Product of simplices and sets of positive upper density in $\mathbb {R}^d$}, journal={Math. Proc. Cambridge Philos. Soc.}, volume={165}, date={2018}, number={1}, pages={25--51}, issn={0305-0041}, review={\MR {3811544}}, doi={10.1017/S0305004117000184}, }
Reference [13]
Neil Lyall and Ákos Magyar, Distance graphs and sets of positive upper density in , Anal. PDE 13 (2020), no. 3, 685–700, DOI 10.2140/apde.2020.13.685. MR4085119,
Show rawAMSref \bib{LM18}{article}{ author={Lyall, Neil}, author={Magyar, \'{A}kos}, title={Distance graphs and sets of positive upper density in $\mathbb {R}^d$}, journal={Anal. PDE}, volume={13}, date={2020}, number={3}, pages={685--700}, issn={2157-5045}, review={\MR {4085119}}, doi={10.2140/apde.2020.13.685}, }
Reference [14]
Neil Lyall and Ákos Magyar, Distances and trees in dense subsets of , Israel J. Math. 240 (2020), no. 2, 769–790, DOI 10.1007/s11856-020-2079-8. MR4193150,
Show rawAMSref \bib{LM19}{article}{ author={Lyall, Neil}, author={Magyar, \'{A}kos}, title={Distances and trees in dense subsets of $\mathbb {Z}^d$}, journal={Israel J. Math.}, volume={240}, date={2020}, number={2}, pages={769--790}, issn={0021-2172}, review={\MR {4193150}}, doi={10.1007/s11856-020-2079-8}, }
[15]
Neil Lyall, Ákos Magyar, and Hans Parshall, Spherical configurations over finite fields, Amer. J. Math. 142 (2020), no. 2, 373–404, DOI 10.1353/ajm.2020.0010. MR4084158,
Show rawAMSref \bib{LMP}{article}{ author={Lyall, Neil}, author={Magyar, \'{A}kos}, author={Parshall, Hans}, title={Spherical configurations over finite fields}, journal={Amer. J. Math.}, volume={142}, date={2020}, number={2}, pages={373--404}, issn={0002-9327}, review={\MR {4084158}}, doi={10.1353/ajm.2020.0010}, }
Reference [16]
Ákos Magyar, On distance sets of large sets of integer points, Israel J. Math. 164 (2008), 251–263, DOI 10.1007/s11856-008-0028-z. MR2391148,
Show rawAMSref \bib{Magy08}{article}{ author={Magyar, \'{A}kos}, title={On distance sets of large sets of integer points}, journal={Israel J. Math.}, volume={164}, date={2008}, pages={251--263}, issn={0021-2172}, review={\MR {2391148}}, doi={10.1007/s11856-008-0028-z}, }
Reference [17]
Ákos Magyar, -point configurations in sets of positive density of , Duke Math. J. 146 (2009), no. 1, 1–34, DOI 10.1215/00127094-2008-060. MR2475398,
Show rawAMSref \bib{Magy09}{article}{ author={Magyar, \'{A}kos}, title={$k$-point configurations in sets of positive density of $\mathbb {Z}^n$}, journal={Duke Math. J.}, volume={146}, date={2009}, number={1}, pages={1--34}, issn={0012-7094}, review={\MR {2475398}}, doi={10.1215/00127094-2008-060}, }
[18]
Hans Parshall, Simplices over finite fields, Proc. Amer. Math. Soc. 145 (2017), no. 6, 2323–2334, DOI 10.1090/proc/13493. MR3626492,
Show rawAMSref \bib{Hans17}{article}{ author={Parshall, Hans}, title={Simplices over finite fields}, journal={Proc. Amer. Math. Soc.}, volume={145}, date={2017}, number={6}, pages={2323--2334}, issn={0002-9939}, review={\MR {3626492}}, doi={10.1090/proc/13493}, }
Reference [19]
Carl Ludwig Siegel, On the theory of indefinite quadratic forms, Ann. of Math. (2) 45 (1944), 577–622, DOI 10.2307/1969191. MR10574,
Show rawAMSref \bib{Siegel}{article}{ author={Siegel, Carl Ludwig}, title={On the theory of indefinite quadratic forms}, journal={Ann. of Math. (2)}, volume={45}, date={1944}, pages={577--622}, issn={0003-486X}, review={\MR {10574}}, doi={10.2307/1969191}, }
Reference [20]
E. Szemerédi, On sets of integers containing no elements in arithmetic progression, Acta Arith. 27 (1975), 199–245, DOI 10.4064/aa-27-1-199-245. MR369312,
Show rawAMSref \bib{Szem75}{article}{ author={Szemer\'{e}di, E.}, title={On sets of integers containing no $k$ elements in arithmetic progression}, journal={Acta Arith.}, volume={27}, date={1975}, pages={199--245}, issn={0065-1036}, review={\MR {369312}}, doi={10.4064/aa-27-1-199-245}, }
Reference [21]
Terence Tao, Szemerédi’s regularity lemma revisited, Contrib. Discrete Math. 1 (2006), no. 1, 8–28. MR2212136,
Show rawAMSref \bib{Tao05}{article}{ author={Tao, Terence}, title={Szemer\'{e}di's regularity lemma revisited}, journal={Contrib. Discrete Math.}, volume={1}, date={2006}, number={1}, pages={8--28}, review={\MR {2212136}}, }
Reference [22]
Terence Tao, A variant of the hypergraph removal lemma, J. Combin. Theory Ser. A 113 (2006), no. 7, 1257–1280, DOI 10.1016/j.jcta.2005.11.006. MR2259060,
Show rawAMSref \bib{Tao06}{article}{ author={Tao, Terence}, title={A variant of the hypergraph removal lemma}, journal={J. Combin. Theory Ser. A}, volume={113}, date={2006}, number={7}, pages={1257--1280}, issn={0097-3165}, review={\MR {2259060}}, doi={10.1016/j.jcta.2005.11.006}, }
Reference [23]
Terence Tao, The ergodic and combinatorial approaches to Szemerédi’s theorem, Additive combinatorics, CRM Proc. Lecture Notes, vol. 43, Amer. Math. Soc., Providence, RI, 2007, pp. 145–193, DOI 10.1090/crmp/043/08. MR2359471,
Show rawAMSref \bib{TaoMontreal}{article}{ author={Tao, Terence}, title={The ergodic and combinatorial approaches to Szemer\'{e}di's theorem}, conference={ title={Additive combinatorics}, }, book={ series={CRM Proc. Lecture Notes}, volume={43}, publisher={Amer. Math. Soc., Providence, RI}, }, date={2007}, pages={145--193}, review={\MR {2359471}}, doi={10.1090/crmp/043/08}, }
[24]
Terence Tao and Van Vu, Additive combinatorics, Cambridge Studies in Advanced Mathematics, vol. 105, Cambridge University Press, Cambridge, 2006, DOI 10.1017/CBO9780511755149. MR2289012,
Show rawAMSref \bib{TV}{book}{ author={Tao, Terence}, author={Vu, Van}, title={Additive combinatorics}, series={Cambridge Studies in Advanced Mathematics}, volume={105}, publisher={Cambridge University Press, Cambridge}, date={2006}, pages={xviii+512}, isbn={978-0-521-85386-6}, isbn={0-521-85386-9}, review={\MR {2289012}}, doi={10.1017/CBO9780511755149}, }
Reference [25]
Tamar Ziegler, Nilfactors of -actions and configurations in sets of positive upper density in , J. Anal. Math. 99 (2006), 249–266, DOI 10.1007/BF02789447. MR2279552,
Show rawAMSref \bib{Ziegler06}{article}{ author={Ziegler, Tamar}, title={Nilfactors of $\mathbb {R}^m$-actions and configurations in sets of positive upper density in $\mathbb {R}^m$}, journal={J. Anal. Math.}, volume={99}, date={2006}, pages={249--266}, issn={0021-7670}, review={\MR {2279552}}, doi={10.1007/BF02789447}, }

Article Information

MSC 2020
Primary: 11B30 (Arithmetic combinatorics; higher degree uniformity)
Author Information
Neil Lyall
Department of Mathematics, The University of Georgia, Athens, Georgia 30602
lyall@math.uga.edu
MathSciNet
Ákos Magyar
Department of Mathematics, The University of Georgia, Athens, Georgia 30602
magyar@math.uga.edu
ORCID
Additional Notes

The first and second authors were partially supported by grants NSF-DMS 1702411 and NSF-DMS 1600840, respectively.

Journal Information
Transactions of the American Mathematical Society, Series B, Volume 9, Issue 5, ISSN 2330-0000, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , and published on .
Copyright Information
Copyright 2022 by the authors under Creative Commons Attribution-Noncommercial 4.0 License (CC BY NC ND 4.0)
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/btran/61
  • MathSciNet Review: 4396041
  • Show rawAMSref \bib{4396041}{article}{ author={Lyall, Neil}, author={Magyar, \'Akos}, title={Weak hypergraph regularity and applications to geometric Ramsey theory}, journal={Trans. Amer. Math. Soc. Ser. B}, volume={9}, number={5}, date={2022}, pages={160-207}, issn={2330-0000}, review={4396041}, doi={10.1090/btran/61}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.