You are currently browsing the category archive for the ‘paper’ category.
Asgar Jamneshan, Or Shalom, and myself have just uploaded to the arXiv our preprints “A Host–Kra -system of order 5 that is not Abramov of order 5, and non-measurability of the inverse theorem for the
norm” and “The structure of totally disconnected Host–Kra–Ziegler factors, and the inverse theorem for the
Gowers uniformity norms on finite abelian groups of bounded torsion“. These two papers are both concerned with advancing the inverse theory for the Gowers norms and Gowers-Host-Kra seminorms; the first paper provides a counterexample in this theory (in particular disproving a conjecture of Bergelson, Ziegler and myself), and the second paper gives new positive results in the case when the underlying group is bounded torsion, or the ergodic system is totally disconnected. I discuss the two papers more below the fold.
Tamar Ziegler and I have just uploaded to the arXiv our paper “Infinite partial sumsets in the primes“. This is a short paper inspired by a recent result of Kra, Moreira, Richter, and Robertson (discussed for instance in this Quanta article from last December) showing that for any set of natural numbers of positive upper density, there exists a sequence
of natural numbers and a shift
such that
for all
this answers a question of Erdős). In view of the “transference principle“, it is then plausible to ask whether the same result holds if
is replaced by the primes. We can show the following results:
Theorem 1
- (i) If the Hardy-Littlewood prime tuples conjecture (or the weaker conjecture of Dickson) is true, then there exists an increasing sequence
of primes such that
is prime for all
.
- (ii) Unconditionally, there exist increasing sequences
and
of natural numbers such that
is prime for all
.
- (iii) These conclusions fail if “prime” is replaced by “positive (relative) density subset of the primes” (even if the density is equal to 1).
We remark that it was shown by Balog that there (unconditionally) exist arbitrarily long but finite sequences of primes such that
is prime for all
. (This result can also be recovered from the later results of Ben Green, myself, and Tamar Ziegler.) Also, it had previously been shown by Granville that on the Hardy-Littlewood prime tuples conjecture, there existed increasing sequences
and
of natural numbers such that
is prime for all
.
The conclusion of (i) is stronger than that of (ii) (which is of course consistent with the former being conditional and the latter unconditional). The conclusion (ii) also implies the well-known theorem of Maynard that for any given , there exist infinitely many
-tuples of primes of bounded diameter, and indeed our proof of (ii) uses the same “Maynard sieve” that powers the proof of that theorem (though we use a formulation of that sieve closer to that in this blog post of mine). Indeed, the failure of (iii) basically arises from the failure of Maynard’s theorem for dense subsets of primes, simply by removing those clusters of primes that are unusually closely spaced.
Our proof of (i) was initially inspired by the topological dynamics methods used by Kra, Moreira, Richter, and Robertson, but we managed to condense it to a purely elementary argument (taking up only half a page) that makes no reference to topological dynamics and builds up the sequence recursively by repeated application of the prime tuples conjecture.
The proof of (ii) takes up the majority of the paper. It is easiest to phrase the argument in terms of “prime-producing tuples” – tuples for which there are infinitely many
with
all prime. Maynard’s theorem is equivalent to the existence of arbitrarily long prime-producing tuples; our theorem is equivalent to the stronger assertion that there exist an infinite sequence
such that every initial segment
is prime-producing. The main new tool for achieving this is the following cute measure-theoretic lemma of Bergelson:
Lemma 2 (Bergelson intersectivity lemma) Letbe subsets of a probability space
of measure uniformly bounded away from zero, thus
. Then there exists a subsequence
such that
for all
.
This lemma has a short proof, though not an entirely obvious one. Firstly, by deleting a null set from , one can assume that all finite intersections
are either positive measure or empty. Secondly, a routine application of Fatou’s lemma shows that the maximal function
has a positive integral, hence must be positive at some point
. Thus there is a subsequence
whose finite intersections all contain
, thus have positive measure as desired by the previous reduction.
It turns out that one cannot quite combine the standard Maynard sieve with the intersectivity lemma because the events that show up (which roughly correspond to the event that
is prime for some random number
(with a well-chosen probability distribution) and some shift
) have their probability going to zero, rather than being uniformly bounded from below. To get around this, we borrow an idea from a paper of Banks, Freiberg, and Maynard, and group the shifts
into various clusters
, chosen in such a way that the probability that at least one of
is prime is bounded uniformly from below. One then applies the Bergelson intersectivity lemma to those events and uses many applications of the pigeonhole principle to conclude.
Rachel Greenfeld and I have just uploaded to the arXiv our paper “A counterexample to the periodic tiling conjecture“. This is the full version of the result I announced on this blog a few months ago, in which we disprove the periodic tiling conjecture of Grünbaum-Shephard and Lagarias-Wang. The paper took a little longer than expected to finish, due to a technical issue that we did not realize at the time of the announcement that required a workaround.
In more detail: the original strategy, as described in the announcement, was to build a “tiling language” that was capable of encoding a certain “-adic Sudoku puzzle”, and then show that the latter type of puzzle had only non-periodic solutions if
was a sufficiently large prime. As it turns out, the second half of this strategy worked out, but there was an issue in the first part: our tiling language was able (using
-group-valued functions) to encode arbitrary boolean relationships between boolean functions, and was also able (using
-valued functions) to encode “clock” functions such as
that were part of our
-adic Sudoku puzzle, but we were not able to make these two types of functions “talk” to each other in the way that was needed to encode the
-adic Sudoku puzzle (the basic problem being that if
is a finite abelian
-group then there are no non-trivial subgroups of
that are not contained in
or trivial in the
direction). As a consequence, we had to replace our “
-adic Sudoku puzzle” by a “
-adic Sudoku puzzle” which basically amounts to replacing the prime
by a sufficiently large power of
(we believe
will suffice). This solved the encoding issue, but the analysis of the
-adic Sudoku puzzles was a little bit more complicated than the
-adic case, for the following reason. The following is a nice exercise in analysis:
Theorem 1 (Linearity in three directions implies full linearity) Letbe a smooth function which is affine-linear on every horizontal line, diagonal (line of slope
), and anti-diagonal (line of slope
). In other words, for any
, the functions
,
, and
are each affine functions on
. Then
is an affine function on
.
Indeed, the property of being affine in three directions shows that the quadratic form associated to the Hessian at any given point vanishes at
,
, and
, and thus must vanish everywhere. In fact the smoothness hypothesis is not necessary; we leave this as an exercise to the interested reader. The same statement turns out to be true if one replaces
with the cyclic group
as long as
is odd; this is the key for us to showing that our
-adic Sudoku puzzles have an (approximate) two-dimensional affine structure, which on further analysis can then be used to show that it is in fact non-periodic. However, it turns out that the corresponding claim for cyclic groups
can fail when
is a sufficiently large power of
! In fact the general form of functions
that are affine on every horizontal line, diagonal, and anti-diagonal takes the form
During the writing process we also discovered that the encoding part of the proof becomes more modular and conceptual once one introduces two new definitions, that of an “expressible property” and a “weakly expressible property”. These concepts are somewhat analogous to that of sentences and
sentences in the arithmetic hierarchy, or to algebraic sets and semi-algebraic sets in real algebraic geometry. Roughly speaking, an expressible property is a property of a tuple of functions
,
from an abelian group
to finite abelian groups
, such that the property can be expressed in terms of one or more tiling equations on the graph
Rachel Greenfeld and I have just uploaded to the arXiv our announcement “A counterexample to the periodic tiling conjecture“. This is an announcement of a longer paper that we are currently in the process of writing up (and hope to release in a few weeks), in which we disprove the periodic tiling conjecture of Grünbaum-Shephard and Lagarias-Wang. This conjecture can be formulated in both discrete and continuous settings:
Conjecture 1 (Discrete periodic tiling conjecture) Suppose thatis a finite set that tiles
by translations (i.e.,
can be partitioned into translates of
). Then
also tiles
by translations periodically (i.e., the set of translations can be taken to be a periodic subset of
).
Conjecture 2 (Continuous periodic tiling conjecture) Suppose thatis a bounded measurable set of positive measure that tiles
by translations up to null sets. Then
also tiles
by translations periodically up to null sets.
The discrete periodic tiling conjecture can be easily established for by the pigeonhole principle (as first observed by Newman), and was proven for
by Bhattacharya (with a new proof given by Greenfeld and myself). The continuous periodic tiling conjecture was established for
by Lagarias and Wang. By an old observation of Hao Wang, one of the consequences of the (discrete) periodic tiling conjecture is that the problem of determining whether a given finite set
tiles by translations is (algorithmically and logically) decidable.
On the other hand, once one allows tilings by more than one tile, it is well known that aperiodic tile sets exist, even in dimension two – finite collections of discrete or continuous tiles that can tile the given domain by translations, but not periodically. Perhaps the most famous examples of such aperiodic tilings are the Penrose tilings, but there are many other constructions; for instance, there is a construction of Ammann, Grümbaum, and Shephard of eight tiles in which tile aperiodically. Recently, Rachel and I constructed a pair of tiles in
that tiled a periodic subset of
aperiodically (in fact we could even make the tiling question logically undecidable in ZFC).
Our main result is then
Theorem 3 Both the discrete and continuous periodic tiling conjectures fail for sufficiently large. Also, there is a finite abelian group
such that the analogue of the discrete periodic tiling conjecture for
is false.
This suggests that the techniques used to prove the discrete periodic conjecture in are already close to the limit of their applicability, as they cannot handle even virtually two-dimensional discrete abelian groups such as
. The main difficulty is in constructing the counterexample in the
setting.
The approach starts by adapting some of the methods of a previous paper of Rachel and myself. The first step is make the problem easier to solve by disproving a “multiple periodic tiling conjecture” instead of the traditional periodic tiling conjecture. At present, Theorem 3 asserts the existence of a “tiling equation” (where one should think of
and
as given, and the tiling set
is known), which admits solutions, all of which are non-periodic. It turns out that it is enough to instead assert the existence of a system
It is convenient to replace sets by functions, so that this tiling language can be translated to a more familiar language, namely the language of (certain types of) functional equations. The key point here is that the tiling equation
The non-periodic behaviour that we ended up trying to capture was that of a certain “-adically structured function”
associated to some fixed and sufficiently large prime
(in fact for our arguments any prime larger than
, e.g.,
, would suffice), defined by the formula
It turns out that we cannot describe this one-dimensional non-periodic function directly via tiling equations. However, we can describe two-dimensional non-periodic functions such as for some coefficients
via a suitable system of tiling equations. A typical such function looks like this:
A feature of this function is that when one restricts to a row or diagonal of such a function, the resulting one-dimensional function exhibits “-adic structure” in the sense that it behaves like a rescaled version of
; see the announcement for a precise version of this statement. It turns out that the converse is essentially true: after excluding some degenerate solutions in which the function is constant along one or more of the columns, all two-dimensional functions which exhibit
-adic structure along (non-vertical) lines must behave like one of the functions
mentioned earlier, and in particular is non-periodic. The proof of this result is strongly reminiscent of the type of reasoning needed to solve a Sudoku puzzle, and so we have adopted some Sudoku-like terminology in our arguments to provide intuition and visuals. One key step is to perform a shear transformation to the puzzle so that many of the rows become constant, as displayed in this example,
and then perform a “Tetris” move of eliminating the constant rows to arrive at a secondary Sudoku puzzle which one then analyzes in turn:
It is the iteration of this procedure that ultimately generates the non-periodic -adic structure.
Kaisa Matomäki, Xuancheng Shao, Joni Teräväinen, and myself have just uploaded to the arXiv our preprint “Higher uniformity of arithmetic functions in short intervals I. All intervals“. This paper investigates the higher order (Gowers) uniformity of standard arithmetic functions in analytic number theory (and specifically, the Möbius function , the von Mangoldt function
, and the generalised divisor functions
) in short intervals
, where
is large and
lies in the range
for a fixed constant
(that one would like to be as small as possible). If we let
denote one of the functions
, then there is extensive literature on the estimation of short sums
Traditionally, asymptotics for such sums are expressed in terms of a “main term” of some arithmetic nature, plus an error term that is estimated in magnitude. For instance, a sum such as would be approximated in terms of a main term that vanished (or is negligible) if
is “minor arc”, but would be expressible in terms of something like a Ramanujan sum if
was “major arc”, together with an error term. We found it convenient to cancel off such main terms by subtracting an approximant
from each of the arithmetic functions
and then getting upper bounds on remainder correlations such as
- For the Möbius function
, we simply set
, as per the Möbius pseudorandomness conjecture. (One could choose a more sophisticated approximant in the presence of a Siegel zero, as I did with Joni in this recent paper, but we do not do so here.)
- For the von Mangoldt function
, we eventually went with the Cramér-Granville approximant
, where
and
.
- For the divisor functions
, we used a somewhat complicated-looking approximant
for some explicit polynomials
, chosen so that
and
have almost exactly the same sums along arithmetic progressions (see the paper for details).
The objective is then to obtain bounds on sums such as (1) that improve upon the “trivial bound” that one can get with the triangle inequality and standard number theory bounds such as the Brun-Titchmarsh inequality. For and
, the Siegel-Walfisz theorem suggests that it is reasonable to expect error terms that have “strongly logarithmic savings” in the sense that they gain a factor of
over the trivial bound for any
; for
, the Dirichlet hyperbola method suggests instead that one has “power savings” in that one should gain a factor of
over the trivial bound for some
. In the case of the Möbius function
, there is an additional trick (introduced by Matomäki and Teräväinen) that allows one to lower the exponent
somewhat at the cost of only obtaining “weakly logarithmic savings” of shape
for some small
.
Our main estimates on sums of the form (1) work in the following ranges:
- For
, one can obtain strongly logarithmic savings on (1) for
, and power savings for
.
- For
, one can obtain weakly logarithmic savings for
.
- For
, one can obtain power savings for
.
- For
, one can obtain power savings for
.
Conjecturally, one should be able to obtain power savings in all cases, and lower down to zero, but the ranges of exponents and savings given here seem to be the limit of current methods unless one assumes additional hypotheses, such as GRH. The
result for correlation against Fourier phases
was established previously by Zhan, and the
result for such phases and
was established previously by by Matomäki and Teräväinen.
By combining these results with tools from additive combinatorics, one can obtain a number of applications:
- Direct insertion of our bounds in the recent work of Kanigowski, Lemanczyk, and Radziwill on the prime number theorem on dynamical systems that are analytic skew products gives some improvements in the exponents there.
- We can obtain a “short interval” version of a multiple ergodic theorem along primes established by Frantzikinakis-Host-Kra and Wooley-Ziegler, in which we average over intervals of the form
rather than
.
- We can obtain a “short interval” version of the “linear equations in primes” asymptotics obtained by Ben Green, Tamar Ziegler, and myself in this sequence of papers, where the variables in these equations lie in short intervals
rather than long intervals such as
.
We now briefly discuss some of the ingredients of proof of our main results. The first step is standard, using combinatorial decompositions (based on the Heath-Brown identity and (for the result) the Ramaré identity) to decompose
into more tractable sums of the following types:
- Type
sums, which are basically of the form
for some weights
of controlled size and some cutoff
that is not too large;
- Type
sums, which are basically of the form
for some weights
,
of controlled size and some cutoffs
that are not too close to
or to
;
- Type
sums, which are basically of the form
for some weights
of controlled size and some cutoff
that is not too large.
The precise ranges of the cutoffs depend on the choice of
; our methods fail once these cutoffs pass a certain threshold, and this is the reason for the exponents
being what they are in our main results.
The Type sums involving nilsequences can be treated by methods similar to those in this previous paper of Ben Green and myself; the main innovations are in the treatment of the Type
and Type
sums.
For the Type sums, one can split into the “abelian” case in which (after some Fourier decomposition) the nilsequence
is basically of the form
, and the “non-abelian” case in which
is non-abelian and
exhibits non-trivial oscillation in a central direction. In the abelian case we can adapt arguments of Matomaki and Shao, which uses Cauchy-Schwarz and the equidistribution properties of polynomials to obtain good bounds unless
is “major arc” in the sense that it resembles (or “pretends to be”)
for some Dirichlet character
and some frequency
, but in this case one can use classical multiplicative methods to control the correlation. It turns out that the non-abelian case can be treated similarly. After applying Cauchy-Schwarz, one ends up analyzing the equidistribution of the four-variable polynomial sequence
For the type sum, a model sum to study is
In a sequel to this paper (currently in preparation), we will obtain analogous results for almost all intervals with
in the range
, in which we will be able to lower
all the way to
.
Jan Grebik, Rachel Greenfeld, Vaclav Rozhon and I have just uploaded to the arXiv our preprint “Measurable tilings by abelian group actions“. This paper is related to an earlier paper of Rachel Greenfeld and myself concerning tilings of lattices , but now we consider the more general situation of tiling a measure space
by a tile
shifted by a finite subset
of shifts of an abelian group
that acts in a measure-preserving (or at least quasi-measure-preserving) fashion on
. For instance,
could be a torus
,
could be a positive measure subset of that torus, and
could be the group
, acting on
by translation.
If is a finite subset of
with the property that the translates
,
of
partition
up to null sets, we write
, and refer to this as a measurable tiling of
by
(with tiling set
). For instance, if
is the torus
, we can create a measurable tiling with
and
. Our main results are the following:
- By modifying arguments from previous papers (including the one with Greenfeld mentioned above), we can establish the following “dilation lemma”: a measurable tiling
automatically implies further measurable tilings
, whenever
is an integer coprime to all primes up to the cardinality
of
.
- By averaging the above dilation lemma, we can also establish a “structure theorem” that decomposes the indicator function
of
into components, each of which are invariant with respect to a certain shift in
. We can establish this theorem in the case of measure-preserving actions on probability spaces via the ergodic theorem, but one can also generalize to other settings by using the device of “measurable medial means” (which relates to the concept of a universally measurable set).
- By applying this structure theorem, we can show that all measurable tilings
of the one-dimensional torus
are rational, in the sense that
lies in a coset of the rationals
. This answers a recent conjecture of Conley, Grebik, and Pikhurko; we also give an alternate proof of this conjecture using some previous results of Lagarias and Wang.
- For tilings
of higher-dimensional tori, the tiling need not be rational. However, we can show that we can “slide” the tiling to be rational by giving each translate
of
a “velocity”
, and for every time
, the translates
still form a partition of
modulo null sets, and at time
the tiling becomes rational. In particular, if a set
can tile a torus in an irrational fashion, then it must also be able to tile the torus in a rational fashion.
- In the two-dimensional case
one can arrange matters so that all the velocities
are parallel. If we furthermore assume that the tile
is connected, we can also show that the union of all the translates
with a common velocity
form a
-invariant subset of the torus.
- Finally, we show that tilings
of a finitely generated discrete group
, with
a finite group, cannot be constructed in a “local” fashion (we formalize this probabilistically using the notion of a “factor of iid process”) unless the tile
is contained in a single coset of
. (Nonabelian local tilings, for instance of the sphere by rotations, are of interest due to connections with the Banach-Tarski paradox; see the aforementioned paper of Conley, Grebik, and Pikhurko. Unfortunately, our methods seem to break down completely in the nonabelian case.)
I’ve just uploaded to the arXiv my preprint “Perfectly packing a square by squares of nearly harmonic sidelength“. This paper concerns a variant of an old problem of Meir and Moser, who asks whether it is possible to perfectly pack squares of sidelength for
into a single square or rectangle of area
. (The following variant problem, also posed by Meir and Moser and discussed for instance in this MathOverflow post, is perhaps even more well known: is it possible to perfectly pack rectangles of dimensions
for
into a single square of area
?) For the purposes of this paper, rectangles and squares are understood to have sides parallel to the axes, and a packing is perfect if it partitions the region being packed up to sets of measure zero. As one partial result towards these problems, it was shown by Paulhus that squares of sidelength
for
can be packed (not quite perfectly) into a single rectangle of area
, and rectangles of dimensions
for
can be packed (again not quite perfectly) into a single square of area
. (Paulhus’s paper had some gaps in it, but these were subsequently repaired by Grzegorek and Januszewski.)
Another direction in which partial progress has been made is to consider instead the problem of packing squares of sidelength ,
perfectly into a square or rectangle of total area
, for some fixed constant
(this lower bound is needed to make the total area
finite), with the aim being to get
as close to
as possible. Prior to this paper, the most recent advance in this direction was by Januszewski and Zielonka last year, who achieved such a packing in the range
.
In this paper we are able to get arbitrarily close to
(which turns out to be a “critical” value of this parameter), but at the expense of deleting the first few tiles:
Theorem 1 If, and
is sufficiently large depending on
, then one can pack squares of sidelength
,
perfectly into a square of area
.
As in previous works, the general strategy is to execute a greedy algorithm, which can be described somewhat incompletely as follows.
- Step 1: Suppose that one has already managed to perfectly pack a square
of area
by squares of sidelength
for
, together with a further finite collection
of rectangles with disjoint interiors. (Initially, we would have
and
, but these parameter will change over the course of the algorithm.)
- Step 2: Amongst all the rectangles in
, locate the rectangle
of the largest width (defined as the shorter of the two sidelengths of
).
- Step 3: Pack (as efficiently as one can) squares of sidelength
for
into
for some
, and decompose the portion of
not covered by this packing into rectangles
.
- Step 4: Replace
by
, replace
by
, and return to Step 1.
The main innovation of this paper is to perform Step 3 somewhat more efficiently than in previous papers.
The above algorithm can get stuck if one reaches a point where one has already packed squares of sidelength for
, but that all remaining rectangles
in
have width less than
, in which case there is no obvious way to fit in the next square. If we let
and
denote the width and height of these rectangles
, then the total area of the rectangles must be
In comparison, the perimeter of the squares that one has already packed is equal to
By choosing the parameter suitably large (and taking
sufficiently large depending on
), one can then prove the theorem. (In order to do some technical bookkeeping and to allow one to close an induction in the verification of the algorithm’s correctness, it is convenient to replace the perimeter
by a slightly weighted variant
for a small exponent
, but this is a somewhat artificial device that somewhat obscures the main ideas.)
Asgar Jamneshan and myself have just uploaded to the arXiv our preprint “The inverse theorem for the Gowers uniformity norm on arbitrary finite abelian groups: Fourier-analytic and ergodic approaches“. This paper, which is a companion to another recent paper of ourselves and Or Shalom, studies the inverse theory for the third Gowers uniformity norm
Theorem 1 (Inverse theorem for) Let
be a finite abelian group, and let
be a
-bounded function with
for some
. Then:
- (i) (Correlation with locally quadratic phase) There exists a regular Bohr set
with
and
, a locally quadratic function
, and a function
such that
- (ii) (Correlation with nilsequence) There exists an explicit degree two filtered nilmanifold
of dimension
, a polynomial map
, and a Lipschitz function
of constant
such that
Such a theorem was proven by Ben Green and myself in the case when was odd, and by Samorodnitsky in the
-torsion case
. In all cases one uses the “higher order Fourier analysis” techniques introduced by Gowers. After some now-standard manipulations (using for instance what is now known as the Balog-Szemerédi-Gowers lemma), one arrives (for arbitrary
) at an estimate that is roughly of the form
So the key step is to obtain a representation of the form (1), possibly after shrinking the Bohr set a little if needed. This has been done in the literature in two ways:
- When
is odd, one has the ability to divide by
, and on the set
one can establish (1) with
. (This is similar to how in single variable calculus the function
is a function whose second derivative is equal to
.)
- When
, then after a change of basis one can take the Bohr set
to be
for some
, and the bilinear form can be written in coordinates as
for somewith
. The diagonal terms
cause a problem, but by subtracting off the rank one form
one can write
on the orthogonal complement offor some coefficients
which now vanish on the diagonal:
. One can now obtain (1) on this complement by taking
In our paper we can now treat the case of arbitrary finite abelian groups , by means of the following two new ingredients:
- (i) Using some geometry of numbers, we can lift the group
to a larger (possibly infinite, but still finitely generated) abelian group
with a projection map
, and find a globally bilinear map
on the latter group, such that one has a representation
of the locally bilinear formby the globally bilinear form
when
are close enough to the origin.
- (ii) Using an explicit construction, one can show that every globally bilinear map
has a representation of the form (1) for some globally quadratic function
.
To illustrate (i), consider the Bohr set in
(where
denotes the distance to the nearest integer), and consider a locally bilinear form
of the form
for some real number
and all integers
(which we identify with elements of
. For generic
, this form cannot be extended to a globally bilinear form on
; however if one lifts
to the finitely generated abelian group
To illustrate (ii), the key case turns out to be when is a cyclic group
, in which case
will take the form
This concludes the Fourier-analytic proof of Theorem 1. In this paper we also give an ergodic theory proof of (a qualitative version of) Theorem 1(ii), using a correspondence principle argument adapted from this previous paper of Ziegler, and myself. Basically, the idea is to randomly generate a dynamical system on the group , by selecting an infinite number of random shifts
, which induces an action of the infinitely generated free abelian group
on
by the formula
This transference principle approach seems to work well for the higher step cases (for instance, the stability of polynomials result is known in arbitrary degree); the main difficulty is to establish a suitable higher step inverse theorem in the ergodic theory setting, which we hope to do in future research.
Asgar Jamneshan, Or Shalom, and myself have just uploaded to the arXiv our preprint “The structure of arbitrary Conze–Lesigne systems“. As the title suggests, this paper is devoted to the structural classification of Conze-Lesigne systems, which are a type of measure-preserving system that are “quadratic” or of “complexity two” in a certain technical sense, and are of importance in the theory of multiple recurrence. There are multiple ways to define such systems; here is one. Take a countable abelian group acting in a measure-preserving fashion on a probability space
, thus each group element
gives rise to a measure-preserving map
. Define the third Gowers-Host-Kra seminorm
of a function
via the formula
The analogous theory in complexity one is well understood. Here, one replaces the norm by the
norm
We return now to the complexity two setting. The most famous examples of Conze-Lesigne systems are (order two) nilsystems, in which the space is a quotient
of a two-step nilpotent Lie group
by a lattice
(equipped with Haar probability measure), and the action is given by a translation
for some group homomorphism
. For instance, the Heisenberg
-nilsystem
Our main result is that even in the infinitely generated case, Conze-Lesigne systems are still inverse limits of a slight generalisation of the nilsystem concept, in which is a locally compact Polish group rather than a Lie group:
Theorem 1 (Classification of Conze-Lesigne systems) Letbe a countable abelian group, and
an ergodic measure-preserving
-system. Then
is a Conze-Lesigne system if and only if it is the inverse limit of translational systems
, where
is a nilpotent locally compact Polish group of nilpotency class two, and
is a lattice in
(and also a lattice in the commutator group
), with
equipped with the Haar probability measure and a translation action
for some homomorphism
.
In a forthcoming companion paper to this one, Asgar Jamneshan and I will use this theorem to derive an inverse theorem for the Gowers norm for an arbitrary finite abelian group
(with no restrictions on the order of
, in particular our result handles the case of even and odd
in a unified fashion). In principle, having a higher order version of this theorem will similarly allow us to derive inverse theorems for
norms for arbitrary
and finite abelian
; we hope to investigate this further in future work.
We sketch some of the main ideas used to prove the theorem. The existing machinery developed by Conze-Lesigne, Furstenberg-Weiss, Host-Kra, and others allows one to describe an arbitrary Conze-Lesigne system as a group extension , where
is a Kronecker system (a rotational system on a compact abelian group
and translation action
),
is another compact abelian group, and the cocycle
is a collection of measurable maps
obeying the cocycle equation
There is an additional technical issue worth pointing out here (which unfortunately was glossed over in some previous work in the area). Because the cocycle equation (1) and the Conze-Lesigne equation (3) are only valid almost everywhere instead of everywhere, the action of on
is technically only a near-action rather than a genuine action, and as such one cannot directly define
to be the stabiliser of a point without running into multiple problems. To fix this, one has to pass to a topological model of
in which the action becomes continuous, and the stabilizer becomes well defined, although one then has to work a little more to check that the action is still transitive. This can be done via Gelfand duality; we proceed using a mixture of a construction from this book of Host and Kra, and the machinery in this recent paper of Asgar and myself.
Now we discuss how to establish the Conze-Lesigne equation (3) in the cyclic group case . As this group embeds into the torus
, it is easy to use existing methods obtain (3) but with the homomorphism
and the function
taking values in
rather than in
. The main task is then to fix up the homomorphism
so that it takes values in
, that is to say that
vanishes. This only needs to be done locally near the origin, because the claim is easy when
lies in the dense subgroup
of
, and also because the claim can be shown to be additive in
. Near the origin one can leverage the Steinhaus lemma to make
depend linearly (or more precisely, homomorphically) on
, and because the cocycle
already takes values in
,
vanishes and
must be an eigenvalue of the system
. But as
was assumed to be separable, there are only countably many eigenvalues, and by another application of Steinhaus and linearity one can then make
vanish on an open neighborhood of the identity, giving the claim.
Joni Teräväinen and I have just uploaded to the arXiv our preprint “The Hardy–Littlewood–Chowla conjecture in the presence of a Siegel zero“. This paper is a development of the theme that certain conjectures in analytic number theory become easier if one makes the hypothesis that Siegel zeroes exist; this places one in a presumably “illusory” universe, since the widely believed Generalised Riemann Hypothesis (GRH) precludes the existence of such zeroes, yet this illusory universe seems remarkably self-consistent and notoriously impossible to eliminate from one’s analysis.
For the purposes of this paper, a Siegel zero is a zero of a Dirichlet
-function
corresponding to a primitive quadratic character
of some conductor
, which is close to
in the sense that
One of the early influential results in this area was the following result of Heath-Brown, which I previously blogged about here:
Theorem 1 (Hardy-Littlewood assuming Siegel zero) Letbe a fixed natural number. Suppose one has a Siegel zero
associated to some conductor
. Then we have
for all
, where
is the von Mangoldt function and
is the singular series
In particular, Heath-Brown showed that if there are infinitely many Siegel zeroes, then there are also infinitely many twin primes, with the correct asymptotic predicted by the Hardy-Littlewood prime tuple conjecture at infinitely many scales.
Very recently, Chinis established an analogous result for the Chowla conjecture (building upon earlier work of Germán and Katai):
Theorem 2 (Chowla assuming Siegel zero) Letbe distinct fixed natural numbers. Suppose one has a Siegel zero
associated to some conductor
. Then one has
in the range
, where
is the Liouville function.
In our paper we unify these results and also improve the quantitative estimates and range of :
Theorem 3 (Hardy-Littlewood-Chowla assuming Siegel zero) Letbe distinct fixed natural numbers with
. Suppose one has a Siegel zero
associated to some conductor
. Then one has
for
for any fixed
.
Our argument proceeds by a series of steps in which we replace and
by more complicated looking, but also more tractable, approximations, until the correlation is one that can be computed in a tedious but straightforward fashion by known techniques. More precisely, the steps are as follows:
- (i) Replace the Liouville function
with an approximant
, which is a completely multiplicative function that agrees with
at small primes and agrees with
at large primes.
- (ii) Replace the von Mangoldt function
with an approximant
, which is the Dirichlet convolution
multiplied by a Selberg sieve weight
to essentially restrict that convolution to almost primes.
- (iii) Replace
with a more complicated truncation
which has the structure of a “Type I sum”, and which agrees with
on numbers that have a “typical” factorization.
- (iv) Replace the approximant
with a more complicated approximant
which has the structure of a “Type I sum”.
- (v) Now that all terms in the correlation have been replaced with tractable Type I sums, use standard Euler product calculations and Fourier analysis, similar in spirit to the proof of the pseudorandomness of the Selberg sieve majorant for the primes in this paper of Ben Green and myself, to evaluate the correlation to high accuracy.
Steps (i), (ii) proceed mainly through estimates such as (1) and standard sieve theory bounds. Step (iii) is based primarily on estimates on the number of smooth numbers of a certain size.
The restriction in our main theorem is needed only to execute step (iv) of this step. Roughly speaking, the Siegel approximant
to
is a twisted, sieved version of the divisor function
, and the types of correlation one is faced with at the start of step (iv) are a more complicated version of the divisor correlation sum
Step (v) is a tedious but straightforward sieve theoretic computation, similar in many ways to the correlation estimates of Goldston and Yildirim used in their work on small gaps between primes (as discussed for instance here), and then also used by Ben Green and myself to locate arithmetic progressions in primes.
Recent Comments