You are currently browsing the tag archive for the ‘Gowers uniformity norms’ tag.
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.
Let be a finite set of order
; in applications
will be typically something like a finite abelian group, such as the cyclic group
. Let us define a
-bounded function to be a function
such that
for all
. There are many seminorms
of interest that one places on functions
that are bounded by
on
-bounded functions, such as the Gowers uniformity seminorms
for
(which are genuine norms for
). All seminorms in this post will be implicitly assumed to obey this property.
In additive combinatorics, a significant role is played by inverse theorems, which abstractly take the following form for certain choices of seminorm , some parameters
, and some class
of
-bounded functions:
Theorem 1 (Inverse theorem template) Ifis a
-bounded function with
, then there exists
such that
, where
denotes the usual inner product
Informally, one should think of as being somewhat small but fixed independently of
,
as being somewhat smaller but depending only on
(and on the seminorm), and
as representing the “structured functions” for these choices of parameters. There is some flexibility in exactly how to choose the class
of structured functions, but intuitively an inverse theorem should become more powerful when this class is small. Accordingly, let us define the
-entropy of the seminorm
to be the least cardinality of
for which such an inverse theorem holds. Seminorms with low entropy are ones for which inverse theorems can be expected to be a useful tool. This concept arose in some discussions I had with Ben Green many years ago, but never appeared in print, so I decided to record some observations we had on this concept here on this blog.
Lebesgue norms for
have exponentially large entropy (and so inverse theorems are not expected to be useful in this case):
Proposition 2 (norm has exponentially large inverse entropy) Let
and
. Then the
-entropy of
is at most
. Conversely, for any
, the
-entropy of
is at least
for some absolute constant
.
Proof: If is
-bounded with
, then we have
Now suppose that there is an -inverse theorem for some
of cardinality
. If we let
be a random sign function (so the
are independent random variables taking values in
with equal probability), then there is a random
such that
Most seminorms of interest in additive combinatorics, such as the Gowers uniformity norms, are bounded by some finite norm thanks to Hölder’s inequality, so from the above proposition and the obvious monotonicity properties of entropy, we conclude that all Gowers norms on finite abelian groups
have at most exponential inverse theorem entropy. But we can do significantly better than this:
- For the
seminorm
, one can simply take
to consist of the constant function
, and the
-entropy is clearly equal to
for any
.
- For the
norm, the standard Fourier-analytic inverse theorem asserts that if
then
for some Fourier character
. Thus the
-entropy is at most
.
- For the
norm on cyclic groups for
, the inverse theorem proved by Green, Ziegler, and myself gives an
-inverse theorem for some
and
consisting of nilsequences
for some filtered nilmanifold
of degree
in a finite collection of cardinality
, some polynomial sequence
(which was subsequently observed by Candela-Sisask (see also Manners) that one can choose to be
-periodic), and some Lipschitz function
of Lipschitz norm
. By the Arzela-Ascoli theorem, the number of possible
(up to uniform errors of size at most
, say) is
. By standard arguments one can also ensure that the coefficients of the polynomial
are
, and then by periodicity there are only
such polynomials. As a consequence, the
-entropy is of polynomial size
(a fact that seems to have first been implicitly observed in Lemma 6.2 of this paper of Frantzikinakis; thanks to Ben Green for this reference). One can obtain more precise dependence on
using the quantitative version of this inverse theorem due to Manners; back of the envelope calculations using Section 5 of that paper suggest to me that one can take
to be polynomial in
and the entropy to be of the order
, or alternatively one can reduce the entropy to
at the cost of degrading
to
.
- If one replaces the cyclic group
by a vector space
over some fixed finite field
of prime order (so that
), then the inverse theorem of Ziegler and myself (available in both high and low characteristic) allows one to obtain an
-inverse theorem for some
and
the collection of non-classical degree
polynomial phases from
to
, which one can normalize to equal
at the origin, and then by the classification of such polynomials one can calculate that the
entropy is of quasipolynomial size
in
. By using the recent work of Gowers and Milicevic, one can make the dependence on
here more precise, but we will not perform these calcualtions here.
- For the
norm on an arbitrary finite abelian group, the recent inverse theorem of Jamneshan and myself gives (after some calculations) a bound of the polynomial form
on the
-entropy for some
, which one can improve slightly to
if one degrades
to
, where
is the maximal order of an element of
, and
is the rank (the number of elements needed to generate
). This bound is polynomial in
in the cyclic group case and quasipolynomial in general.
For general finite abelian groups , we do not yet have an inverse theorem of comparable power to the ones mentioned above that give polynomial or quasipolynomial upper bounds on the entropy. However, there is a cheap argument that at least gives some subexponential bounds:
Proposition 3 (Cheap subexponential bound) Letand
, and suppose that
is a finite abelian group of order
for some sufficiently large
. Then the
-complexity of
is at most
.
Proof: (Sketch) We use a standard random sampling argument, of the type used for instance by Croot-Sisask or Briet-Gopi (thanks to Ben Green for this latter reference). We can assume that for some sufficiently large
, since otherwise the claim follows from Proposition 2.
Let be a random subset of
with the events
being iid with probability
to be chosen later, conditioned to the event
. Let
be a
-bounded function. By a standard second moment calculation, we see that with probability at least
, we have
If we then let be
rounded to the nearest Gaussian integer multiple of
in the unit disk, one has from the triangle inequality that
Now we remove the failure probability by independent resampling. By rounding to the nearest Gaussian integer multiple of in the unit disk for a sufficiently small
, one can find a family
of cardinality
consisting of
-bounded functions
of
norm at least
such that for every
-bounded
with
there exists
such that
One way to obtain lower bounds on the inverse theorem entropy is to produce a collection of almost orthogonal functions with large norm. More precisely:
Proposition 4 Letbe a seminorm, let
, and suppose that one has a collection
of
-bounded functions such that for all
,
one has
for all but at most
choices of
for all distinct
. Then the
-entropy of
is at least
.
Proof: Suppose we have an -inverse theorem with some family
. Then for each
there is
such that
. By the pigeonhole principle, there is thus
such that
for all
in a subset
of
of cardinality at least
:
Thus for instance:
- For the
norm, one can take
to be the family of linear exponential phases
with
and
, and obtain a linear lower bound of
for the
-entropy, thus matching the upper bound of
up to constants when
is fixed.
- For the
norm, a similar calculation using polynomial phases of degree
, combined with the Weyl sum estimates, gives a lower bound of
for the
-entropy for any fixed
; by considering nilsequences as well, together with nilsequence equidistribution theory, one can replace the exponent
here by some quantity that goes to infinity as
, though I have not attempted to calculate the exact rate.
- For the
norm, another similar calculation using polynomial phases of degree
should give a lower bound of
for the
-entropy, though I have not fully performed the calculation.
We close with one final example. Suppose is a product
of two sets
of cardinality
, and we consider the Gowers box norm
In the modern theory of higher order Fourier analysis, a key role are played by the Gowers uniformity norms for
. For finitely supported functions
, one can define the (non-normalised) Gowers norm
by the formula
The significance of the Gowers norms is that they control other multilinear forms that show up in additive combinatorics. Given any polynomials and functions
, we define the multilinear form
-
and
have true complexity
;
-
has true complexity
;
-
has true complexity
;
- The form
(which among other things could be used to count twin primes) has infinite true complexity (which is quite unfortunate for applications).
Gowers and Wolf formulated a conjecture on what this complexity should be, at least for linear polynomials ; Ben Green and I thought we had resolved this conjecture back in 2010, though it turned out there was a subtle gap in our arguments and we were only able to resolve the conjecture in a partial range of cases. However, the full conjecture was recently resolved by Daniel Altman.
The (semi-)norm is so weak that it barely controls any averages at all. For instance the average
Because of this, I propose inserting an additional norm in the Gowers uniformity norm hierarchy between the and
norms, which I will call the
(or “profinite
“) norm:
The norm recently appeared implicitly in work of Peluse and Prendiville, who showed that the form
had true complexity
in this notation (with polynomially strong bounds). [Actually, strictly speaking this control was only shown for the third function
; for the first two functions
one needs to localize the
norm to intervals of length
. But I will ignore this technical point to keep the exposition simple.] The weaker claim that
has true complexity
is substantially easier to prove (one can apply the circle method together with Gauss sum estimates).
The well known inverse theorem for the norm tells us that if a
-bounded function
has
norm at least
for some
, then there is a Fourier phase
such that
For one has a trivial inverse theorem; by definition, the
norm of
is at least
if and only if
For one has the intermediate situation in which the frequency
is not taken to be zero, but is instead major arc. Indeed, suppose that
is
-bounded with
, thus
Here is a diagram showing some of the control relationships between various Gowers norms, multilinear forms, and duals of classes of functions (where each class of functions
induces a dual norm
:
Here I have included the three classes of functions that one can choose from for the inverse theorem, namely degree two nilsequences, bracket quadratic phases, and local quadratic phases, as well as the more narrow class of globally quadratic phases.
The Gowers norms have counterparts for measure-preserving systems , known as Host-Kra seminorms. The
norm can be defined for
as
Joni Teräväinen and myself have just uploaded to the arXiv our preprint “Quantitative bounds for Gowers uniformity of the Möbius and von Mangoldt functions“. This paper makes quantitative the Gowers uniformity estimates on the Möbius function and the von Mangoldt function
.
To discuss the results we first discuss the situation of the Möbius function, which is technically simpler in some (though not all) ways. We assume familiarity with Gowers norms and standard notations around these norms, such as the averaging notation and the exponential notation
. The prime number theorem in qualitative form asserts that
Once one restricts to arithmetic progressions, the situation gets worse: the Siegel-Walfisz theorem gives the bound
for any residue classIn 1937, Davenport was able to show the discorrelation estimate
For the situation with the norm the previously known results were much weaker. Ben Green and I showed that
For higher norms , the situation is even worse, because the quantitative inverse theory for these norms is poorer, and indeed it was only with the recent work of Manners that any such bound is available at all (at least for
). Basically, Manners establishes if
Our first result gives an effective decay bound:
Theorem 1 For any, we have
for some
. The implied constants are effective.
This is off by a logarithm from the best effective bound (2) in the case. In the
case there is some hope to remove this logarithm based on the improved quantitative inverse theory currently available in this case, but there is a technical obstruction to doing so which we will discuss later in this post. For
the above bound is the best one could hope to achieve purely using the quantitative inverse theory of Manners.
We have analogues of all the above results for the von Mangoldt function . Here a complication arises that
does not have mean close to zero, and one has to subtract off some suitable approximant
to
before one would expect good Gowers norms bounds. For the prime number theorem one can just use the approximant
, giving
Theorem 2 For any, we have
for some
. The implied constants are effective.
By standard methods, this result also gives quantitative asymptotics for counting solutions to various systems of linear equations in primes, with error terms that gain a factor of with respect to the main term.
We now discuss the methods of proof, focusing first on the case of the Möbius function. Suppose first that there is no “Siegel zero”, by which we mean a quadratic character of some conductor
with a zero
with
for some small absolute constant
. In this case the Siegel-Walfisz bound (1) improves to a quasipolynomial bound
Now suppose we have a Siegel zero . In this case the bound (5) will not hold in general, and hence also (6) will not hold either. Here, the usual way out (while still maintaining effective estimates) is to approximate
not by
, but rather by a more complicated approximant
that takes the Siegel zero into account, and in particular is such that one has the (effective) pseudopolynomial bound
For the analogous problem with the von Mangoldt function (assuming a Siegel zero for sake of discussion), the approximant is simpler; we ended up using
In principle, the above results can be improved for due to the stronger quantitative inverse theorems in the
setting. However, there is a bottleneck that prevents us from achieving this, namely that the equidistribution theory of two-step nilmanifolds has exponents which are exponential in the dimension rather than polynomial in the dimension, and as a consequence we were unable to improve upon the doubly logarithmic results. Specifically, if one is given a sequence of bracket quadratics such as
that fails to be
-equidistributed, one would need to establish a nontrivial linear relationship modulo 1 between the
(up to errors of
), where the coefficients are of size
; current methods only give coefficient bounds of the form
. An old result of Schmidt demonstrates proof of concept that these sorts of polynomial dependencies on exponents is possible in principle, but actually implementing Schmidt’s methods here seems to be a quite non-trivial task. There is also another possible route to removing a logarithm, which is to strengthen the inverse
theorem to make the dimension of the nilmanifold logarithmic in the uniformity parameter
rather than polynomial. Again, the Freiman-Bilu theorem (see for instance this paper of Ben and myself) demonstrates proof of concept that such an improvement in dimension is possible, but some work would be needed to implement it.
Ben Green and I have updated our paper “An arithmetic regularity lemma, an associated counting lemma, and applications” to account for a somewhat serious issue with the paper that was pointed out to us recently by Daniel Altman. This paper contains two core theorems:
- An “arithmetic regularity lemma” that, roughly speaking, decomposes an arbitrary bounded sequence
on an interval
as an “irrational nilsequence”
of controlled complexity, plus some “negligible” errors (where one uses the Gowers uniformity norm as the main norm to control the neglibility of the error); and
- An “arithmetic counting lemma” that gives an asymptotic formula for counting various averages
for various affine-linear forms
when the functions
are given by irrational nilsequences.
The combination of the two theorems is then used to address various questions in additive combinatorics.
There are no direct issues with the arithmetic regularity lemma. However, it turns out that the arithmetic counting lemma is only true if one imposes an additional property (which we call the “flag property”) on the affine-linear forms . Without this property, there does not appear to be a clean asymptotic formula for these averages if the only hypothesis one places on the underlying nilsequences is irrationality. Thus when trying to understand the asymptotics of averages involving linear forms that do not obey the flag property, the paradigm of understanding these averages via a combination of the regularity lemma and a counting lemma seems to require some significant revision (in particular, one would probably have to replace the existing regularity lemma with some variant, despite the fact that the lemma is still technically true in this setting). Fortunately, for most applications studied to date (including the important subclass of translation-invariant affine forms), the flag property holds; however our claim in the paper to have resolved a conjecture of Gowers and Wolf on the true complexity of systems of affine forms must now be narrowed, as our methods only verify this conjecture under the assumption of the flag property.
In a bit more detail: the asymptotic formula for our counting lemma involved some finite-dimensional vector spaces for various natural numbers
, defined as the linear span of the vectors
as
ranges over the parameter space
. Roughly speaking, these spaces encode some constraints one would expect to see amongst the forms
. For instance, in the case of length four arithmetic progressions when
,
, and
The arguments in our paper turn out to be perfectly correct under the assumption of the “flag property” that for all
. The problem is that the flag property turns out to not always hold. A counterexample, provided by Daniel Altman, involves the four linear forms
Fortunately, the flag property does hold in several key cases, most notably the translation invariant case when contains
, as well as “complexity one” cases. Nevertheless non-flag property systems of affine forms do exist, thus limiting the range of applicability of the techniques in this paper. In particular, the conjecture of Gowers and Wolf (Theorem 1.13 in the paper) is now open again in the non-flag property case.
Kaisa Matomäki, Maksym Radziwill, Joni Teräväinen, Tamar Ziegler and I have uploaded to the arXiv our paper Higher uniformity of bounded multiplicative functions in short intervals on average. This paper (which originated from a working group at an AIM workshop on Sarnak’s conjecture) focuses on the local Fourier uniformity conjecture for bounded multiplicative functions such as the Liouville function . One form of this conjecture is the assertion that
The conjecture gets more difficult as increases, and also becomes more difficult the more slowly
grows with
. The
conjecture is equivalent to the assertion
For , the conjecture is equivalent to the assertion
Now we apply the same strategy to (4). For abelian the claim follows easily from (3), so we focus on the non-abelian case. One now has a polynomial sequence
attached to many
, and after a somewhat complicated adaptation of the above arguments one again ends up with an approximate functional equation
We give two applications of this higher order Fourier uniformity. One regards the growth of the number
The second application is to obtain cancellation for various polynomial averages involving the Liouville function or von Mangoldt function
, such as
In the modern theory of additive combinatorics, a large role is played by the Gowers uniformity norms , where
,
is a finite abelian group, and
is a function (one can also consider these norms in finite approximate groups such as
instead of finite groups, but we will focus on the group case here for simplicity). These norms can be defined by the formula
where we use the averaging notation
for any non-empty finite set (with
denoting the cardinality of
), and
is the multiplicative discrete derivative operator
One reason why these norms play an important role is that they control various multilinear averages. We give two sample examples here:
We establish these claims a little later in this post.
In some more recent literature (e.g., this paper of Conlon, Fox, and Zhao), the role of Gowers norms have been replaced by (generalisations) of the cut norm, a concept originating from graph theory. In this blog post, it will be convenient to define these cut norms in the language of probability theory (using boldface to denote random variables).
Definition 2 (Cut norm) Let
be independent random variables with
; to avoid minor technicalities we assume that these random variables are discrete and take values in a finite set. Given a random variable
of these independent random variables, we define the cut norm
where the supremum ranges over all choices
of random variables
that are
-bounded (thus
surely), and such that
does not depend on
.
If
, we abbreviate
as
.
Strictly speaking, the cut norm is only a cut semi-norm when , but we will abuse notation by referring to it as a norm nevertheless.
Example 3 If
is a bipartite graph, and
,
are independent random variables chosen uniformly from
respectively, then
where the supremum ranges over all
-bounded functions
,
. The right hand side is essentially the cut norm of the graph
, as defined for instance by Frieze and Kannan.
The cut norm is basically an expectation when :
Example 4 If
, we see from definition that
If
, one easily checks that
where
is the conditional expectation of
to the
-algebra generated by all the variables other than
, i.e., the
-algebra generated by
. In particular, if
are independent random variables drawn uniformly from
respectively, then
Here are some basic properties of the cut norm:
Lemma 5 (Basic properties of cut norm) Let
be independent discrete random variables, and
a function of these variables.
- (i) (Permutation invariance) The cut norm
is invariant with respect to permutations of the
, or permutations of the
.
- (ii) (Conditioning) One has
where on the right-hand side we view, for each realisation
of
,
as a function
of the random variables
alone, thus the right-hand side may be expanded as
- (iii) (Monotonicity) If
, we have
- (iv) (Multiplicative invariances) If
is a
-bounded function that does not depend on one of the
, then
In particular, if we additionally assume
, then
- (v) (Cauchy-Schwarz) If
, one has
where
is a copy of
that is independent of
and
is the random variable
- (vi) (Averaging) If
and
, where
is another random variable independent of
, and
is a random variable depending on both
and
, then
Proof: The claims (i), (ii) are clear from expanding out all the definitions. The claim (iii) also easily follows from the definitions (the left-hand side involves a supremum over a more general class of multipliers , while the right-hand side omits the
multiplier), as does (iv) (the multiplier
can be absorbed into one of the multipliers in the definition of the cut norm). The claim (vi) follows by expanding out the definitions, and observing that all of the terms in the supremum appearing in the left-hand side also appear as terms in the supremum on the right-hand side. It remains to prove (v). By definition, the left-hand side is the supremum over all quantities of the form
where the are
-bounded functions of
that do not depend on
. We average out in the
direction (that is, we condition out the variables
), and pull out the factor
(which does not depend on
), to write this as
which by Cauchy-Schwarz is bounded by
which can be expanded using the copy as
Expanding
and noting that each is
-bounded and independent of
for
, we obtain the claim.
Now we can relate the cut norm to Gowers uniformity norms:
Lemma 6 Let
be a finite abelian group, let
be independent random variables uniformly drawn from
for some
, and let
. Then
If
is additionally assumed to be
-bounded, we have the converse inequalities
Proof: Applying Lemma 5(v) times, we can bound
where are independent copies of
that are also independent of
. The expression inside the norm can also be written as
so by Example 4 one can write (6) as
which after some change of variables simplifies to
which by Cauchy-Schwarz is bounded by
which one can rearrange as
giving (2). A similar argument bounds
by
which gives (3).
For (4), we can reverse the above steps and expand as
which we can write as
for some -bounded function
. This can in turn be expanded as
for some -bounded functions
that do not depend on
. By Example 4, this can be written as
which by several applications of Theorem 5(iii) and then Theorem 5(iv) can be bounded by
giving (4). A similar argument gives (5).
Now we can prove Proposition 1. We begin with part (i). By permutation we may assume , then by translation we may assume
. Replacing
by
and
by
, we can write the left-hand side of (1) as
where
is a -bounded function that does not depend on
. Taking
to be independent random variables drawn uniformly from
, the left-hand side of (1) can then be written as
which by Example 4 is bounded in magnitude by
After many applications of Lemma 5(iii), (iv), this is bounded by
By Lemma 5(ii) we may drop the variable, and then the claim follows from Lemma 6.
For part (ii), we replace by
and
by
to write the left-hand side as
the point here is that the first factor does not involve , the second factor does not involve
, and the third factor has no quadratic terms in
. Letting
be independent variables drawn uniformly from
, we can use Example 4 to bound this in magnitude by
which by Lemma 5(i),(iii),(iv) is bounded by
and then by Lemma 5(v) we may bound this by
which by Example 4 is
Now the expression inside the expectation is the product of four factors, each of which is or
applied to an affine form
where
depends on
and
is one of
,
,
,
. With probability
, the four different values of
are distinct, and then by part (i) we have
When they are not distinct, we can instead bound this quantity by . Taking expectations in
, we obtain the claim.
The analogue of the inverse theorem for cut norms is the following claim (which I learned from Ben Green):
Lemma 7 (
-type inverse theorem) Let
be independent random variables drawn from a finite abelian group
, and let
be
-bounded. Then we have
where
is the group of homomorphisms
is a homomorphism from
to
, and
.
Proof: Suppose first that for some
, then by definition
for some -bounded
. By Fourier expansion, the left-hand side is also
where . From Plancherel’s theorem we have
hence by Hölder’s inequality one has for some
, and hence
Conversely, suppose (7) holds. Then there is such that
which on substitution and Example 4 implies
The term splits into the product of a factor
not depending on
, and a factor
not depending on
. Applying Lemma 5(iii), (iv) we conclude that
The claim follows.
The higher order inverse theorems are much less trivial (and the optimal quantitative bounds are not currently known). However, there is a useful degree lowering argument, due to Peluse and Prendiville, that can allow one to lower the order of a uniformity norm in some cases. We give a simple version of this argument here:
Lemma 8 (Degree lowering argument, special case) Let
be a finite abelian group, let
be a non-empty finite set, and let
be a function of the form
for some
-bounded functions
indexed by
. Suppose that
for some
and
. Then one of the following claims hold (with implied constants allowed to depend on
):
- (i) (Degree lowering) one has
.
- (ii) (Non-zero frequency) There exist
and non-zero
such that
There are more sophisticated versions of this argument in which the frequency is “minor arc” rather than “zero frequency”, and then the Gowers norms are localised to suitable large arithmetic progressions; this is implicit in the above-mentioned paper of Peluse and Prendiville.
Proof: One can write
and hence we conclude that
for a set of tuples
of density
. Applying Lemma 6 and Lemma 7, we see that for each such tuple, there exists
such that
where is drawn uniformly from
.
Let us adopt the convention that vanishes for
not in
, then from Lemma 5(ii) we have
where are independent random variables drawn uniformly from
and also independent of
. By repeated application of Lemma 5(iii) we then have
Expanding out and using Lemma 5(iv) repeatedly we conclude that
From definition of we then have
By Lemma 5(vi), we see that the left-hand side is less than
where is drawn uniformly from
, independently of
. By repeated application of Lemma 5(i), (v) repeatedly, we conclude that
where are independent copies of
that are also independent of
,
. By Lemma 5(ii) and Example 4 we conclude that
with probability .
The left-hand side can be rewritten as
where is the additive version of
, thus
Translating , we can simplify this a little to
If the frequency is ever non-vanishing in the event (9) then conclusion (ii) applies. We conclude that
with probability . In particular, by the pigeonhole principle, there exist
such that
with probability . Expanding this out, we obtain a representation of the form
holding with probability , where the
are functions that do not depend on the
coordinate. From (8) we conclude that
for of the tuples
. Thus by Lemma 5(ii)
By repeated application of Lemma 5(iii) we then have
and then by repeated application of Lemma 5(iv)
and then the conclusion (i) follows from Lemma 6.
As an application of degree lowering, we give an inverse theorem for the average in Proposition 1(ii), first established by Bourgain-Chang and later reproved by Peluse (by different methods from those given here):
Proposition 9 Let
be a cyclic group of prime order. Suppose that one has
-bounded functions
such that
for some
. Then either
, or one has
We remark that a modification of the arguments below also give .
Proof: The left-hand side of (10) can be written as
where is the dual function
By Cauchy-Schwarz one thus has
and hence by Proposition 1, we either have (in which case we are done) or
Writing with
, we conclude that either
, or that
for some and non-zero
. The left-hand side can be rewritten as
where and
. We can rewrite this in turn as
which is bounded by
where are independent random variables drawn uniformly from
. Applying Lemma 5(v), we conclude that
However, a routine Gauss sum calculation reveals that the left-hand side is for some absolute constant
because
is non-zero, so that
. The only remaining case to consider is when
Repeating the above arguments we then conclude that
and then
The left-hand side can be computed to equal , and the claim follows.
This argument was given for the cyclic group setting, but the argument can also be applied to the integers (see Peluse-Prendiville) and can also be used to establish an analogue over the reals (that was first obtained by Bourgain).
Tamar Ziegler and I have just uploaded to the arXiv two related papers: “Concatenation theorems for anti-Gowers-uniform functions and Host-Kra characteoristic factors” and “polynomial patterns in primes“, with the former developing a “quantitative Bessel inequality” for local Gowers norms that is crucial in the latter.
We use the term “concatenation theorem” to denote results in which structural control of a function in two or more “directions” can be “concatenated” into structural control in a joint direction. A trivial example of such a concatenation theorem is the following: if a function is constant in the first variable (thus
is constant for each
), and also constant in the second variable (thus
is constant for each
), then it is constant in the joint variable
. A slightly less trivial example: if a function
is affine-linear in the first variable (thus, for each
, there exist
such that
for all
) and affine-linear in the second variable (thus, for each
, there exist
such that
for all
) then
is a quadratic polynomial in
; in fact it must take the form
for some real numbers . (This can be seen for instance by using the affine linearity in
to show that the coefficients
are also affine linear.)
The same phenomenon extends to higher degree polynomials. Given a function from one additive group
to another, we say that
is of degree less than
along a subgroup
of
if all the
-fold iterated differences of
along directions in
vanish, that is to say
for all and
, where
is the difference operator
(We adopt the convention that the only of degree less than
is the zero function.)
We then have the following simple proposition:
Proposition 1 (Concatenation of polynomiality) Let
be of degree less than
along one subgroup
of
, and of degree less than
along another subgroup
of
, for some
. Then
is of degree less than
along the subgroup
of
.
Note the previous example was basically the case when ,
,
,
, and
.
Proof: The claim is trivial for or
(in which
is constant along
or
respectively), so suppose inductively
and the claim has already been proven for smaller values of
.
We take a derivative in a direction along
to obtain
where is the shift of
by
. Then we take a further shift by a direction
to obtain
leading to the cocycle equation
Since has degree less than
along
and degree less than
along
,
has degree less than
along
and less than
along
, so is degree less than
along
by induction hypothesis. Similarly
is also of degree less than
along
. Combining this with the cocycle equation we see that
is of degree less than
along
for any
, and hence
is of degree less than
along
, as required.
While this proposition is simple, it already illustrates some basic principles regarding how one would go about proving a concatenation theorem:
- (i) One should perform induction on the degrees
involved, and take advantage of the recursive nature of degree (in this case, the fact that a function is of less than degree
along some subgroup
of directions iff all of its first derivatives along
are of degree less than
).
- (ii) Structure is preserved by operations such as addition, shifting, and taking derivatives. In particular, if a function
is of degree less than
along some subgroup
, then any derivative
of
is also of degree less than
along
, even if
does not belong to
.
Here is another simple example of a concatenation theorem. Suppose an at most countable additive group acts by measure-preserving shifts
on some probability space
; we call the pair
(or more precisely
) a
-system. We say that a function
is a generalised eigenfunction of degree less than
along some subgroup
of
and some
if one has
almost everywhere for all , and some functions
of degree less than
along
, with the convention that a function has degree less than
if and only if it is equal to
. Thus for instance, a function
is an generalised eigenfunction of degree less than
along
if it is constant on almost every
-ergodic component of
, and is a generalised function of degree less than
along
if it is an eigenfunction of the shift action on almost every
-ergodic component of
. A basic example of a higher order eigenfunction is the function
on the skew shift
with
action given by the generator
for some irrational
. One can check that
for every integer
, where
is a generalised eigenfunction of degree less than
along
, so
is of degree less than
along
.
We then have
Proposition 2 (Concatenation of higher order eigenfunctions) Let
be a
-system, and let
be a generalised eigenfunction of degree less than
along one subgroup
of
, and a generalised eigenfunction of degree less than
along another subgroup
of
, for some
. Then
is a generalised eigenfunction of degree less than
along the subgroup
of
.
The argument is almost identical to that of the previous proposition and is left as an exercise to the reader. The key point is the point (ii) identified earlier: the space of generalised eigenfunctions of degree less than along
is preserved by multiplication and shifts, as well as the operation of “taking derivatives”
even along directions
that do not lie in
. (To prove this latter claim, one should restrict to the region where
is non-zero, and then divide
by
to locate
.)
A typical example of this proposition in action is as follows: consider the -system given by the
-torus
with generating shifts
for some irrational , which can be checked to give a
action
The function can then be checked to be a generalised eigenfunction of degree less than
along
, and also less than
along
, and less than
along
. One can view this example as the dynamical systems translation of the example (1) (see this previous post for some more discussion of this sort of correspondence).
The main results of our concatenation paper are analogues of these propositions concerning a more complicated notion of “polynomial-like” structure that are of importance in additive combinatorics and in ergodic theory. On the ergodic theory side, the notion of structure is captured by the Host-Kra characteristic factors of a
-system
along a subgroup
. These factors can be defined in a number of ways. One is by duality, using the Gowers-Host-Kra uniformity seminorms (defined for instance here)
. Namely,
is the factor of
defined up to equivalence by the requirement that
An equivalent definition is in terms of the dual functions of
along
, which can be defined recursively by setting
and
where denotes the ergodic average along a Følner sequence in
(in fact one can also define these concepts in non-amenable abelian settings as per this previous post). The factor
can then be alternately defined as the factor generated by the dual functions
for
.
In the case when and
is
-ergodic, a deep theorem of Host and Kra shows that the factor
is equivalent to the inverse limit of nilsystems of step less than
. A similar statement holds with
replaced by any finitely generated group by Griesmer, while the case of an infinite vector space over a finite field was treated in this paper of Bergelson, Ziegler, and myself. The situation is more subtle when
is not
-ergodic, or when
is
-ergodic but
is a proper subgroup of
acting non-ergodically, when one has to start considering measurable families of directional nilsystems; see for instance this paper of Austin for some of the subtleties involved (for instance, higher order group cohomology begins to become relevant!).
One of our main theorems is then
Proposition 3 (Concatenation of characteristic factors) Let
be a
-system, and let
be measurable with respect to the factor
and with respect to the factor
for some
and some subgroups
of
. Then
is also measurable with respect to the factor
.
We give two proofs of this proposition in the paper; an ergodic-theoretic proof using the Host-Kra theory of “cocycles of type (along a subgroup
)”, which can be used to inductively describe the factors
, and a combinatorial proof based on a combinatorial analogue of this proposition which is harder to state (but which roughly speaking asserts that a function which is nearly orthogonal to all bounded functions of small
norm, and also to all bounded functions of small
norm, is also nearly orthogonal to alll bounded functions of small
norm). The combinatorial proof parallels the proof of Proposition 2. A key point is that dual functions
obey a property analogous to being a generalised eigenfunction, namely that
where and
is a “structured function of order
” along
. (In the language of this previous paper of mine, this is an assertion that dual functions are uniformly almost periodic of order
.) Again, the point (ii) above is crucial, and in particular it is key that any structure that
has is inherited by the associated functions
and
. This sort of inheritance is quite easy to accomplish in the ergodic setting, as there is a ready-made language of factors to encapsulate the concept of structure, and the shift-invariance and
-algebra properties of factors make it easy to show that just about any “natural” operation one performs on a function measurable with respect to a given factor, returns a function that is still measurable in that factor. In the finitary combinatorial setting, though, encoding the fact (ii) becomes a remarkably complicated notational nightmare, requiring a huge amount of “epsilon management” and “second-order epsilon management” (in which one manages not only scalar epsilons, but also function-valued epsilons that depend on other parameters). In order to avoid all this we were forced to utilise a nonstandard analysis framework for the combinatorial theorems, which made the arguments greatly resemble the ergodic arguments in many respects (though the two settings are still not equivalent, see this previous blog post for some comparisons between the two settings). Unfortunately the arguments are still rather complicated.
For combinatorial applications, dual formulations of the concatenation theorem are more useful. A direct dualisation of the theorem yields the following decomposition theorem: a bounded function which is small in norm can be split into a component that is small in
norm, and a component that is small in
norm. (One may wish to understand this type of result by first proving the following baby version: any function that has mean zero on every coset of
, can be decomposed as the sum of a function that has mean zero on every
coset, and a function that has mean zero on every
coset. This is dual to the assertion that a function that is constant on every
coset and constant on every
coset, is constant on every
coset.) Combining this with some standard “almost orthogonality” arguments (i.e. Cauchy-Schwarz) give the following Bessel-type inequality: if one has a lot of subgroups
and a bounded function is small in
norm for most
, then it is also small in
norm for most
. (Here is a baby version one may wish to warm up on: if a function
has small mean on
for some large prime
, then it has small mean on most of the cosets of most of the one-dimensional subgroups of
.)
There is also a generalisation of the above Bessel inequality (as well as several of the other results mentioned above) in which the subgroups are replaced by more general coset progressions
(of bounded rank), so that one has a Bessel inequailty controlling “local” Gowers uniformity norms such as
by “global” Gowers uniformity norms such as
. This turns out to be particularly useful when attempting to compute polynomial averages such as
for various functions . After repeated use of the van der Corput lemma, one can control such averages by expressions such as
(actually one ends up with more complicated expressions than this, but let’s use this example for sake of discussion). This can be viewed as an average of various Gowers uniformity norms of
along arithmetic progressions of the form
for various
. Using the above Bessel inequality, this can be controlled in turn by an average of various
Gowers uniformity norms along rank two generalised arithmetic progressions of the form
for various
. But for generic
, this rank two progression is close in a certain technical sense to the “global” interval
(this is ultimately due to the basic fact that two randomly chosen large integers are likely to be coprime, or at least have a small gcd). As a consequence, one can use the concatenation theorems from our first paper to control expressions such as (2) in terms of global Gowers uniformity norms. This is important in number theoretic applications, when one is interested in computing sums such as
or
where and
are the Möbius and von Mangoldt functions respectively. This is because we are able to control global Gowers uniformity norms of such functions (thanks to results such as the proof of the inverse conjecture for the Gowers norms, the orthogonality of the Möbius function with nilsequences, and asymptotics for linear equations in primes), but much less control is currently available for local Gowers uniformity norms, even with the assistance of the generalised Riemann hypothesis (see this previous blog post for some further discussion).
By combining these tools and strategies with the “transference principle” approach from our previous paper (as improved using the recent “densification” technique of Conlon, Fox, and Zhao, discussed in this previous post), we are able in particular to establish the following result:
Theorem 4 (Polynomial patterns in the primes) Let
be polynomials of degree at most
, whose degree
coefficients are all distinct, for some
. Suppose that
is admissible in the sense that for every prime
, there are
such that
are all coprime to
. Then there exist infinitely many pairs
of natural numbers such that
are prime.
Furthermore, we obtain an asymptotic for the number of such pairs in the range
,
(actually for minor technical reasons we reduce the range of
to be very slightly less than
). In fact one could in principle obtain asymptotics for smaller values of
, and relax the requirement that the degree
coefficients be distinct with the requirement that no two of the
differ by a constant, provided one had good enough local uniformity results for the Möbius or von Mangoldt functions. For instance, we can obtain an asymptotic for triplets of the form
unconditionally for
, and conditionally on GRH for all
, using known results on primes in short intervals on average.
The case of this theorem was obtained in a previous paper of myself and Ben Green (using the aforementioned conjectures on the Gowers uniformity norm and the orthogonality of the Möbius function with nilsequences, both of which are now proven). For higher
, an older result of Tamar and myself was able to tackle the case when
(though our results there only give lower bounds on the number of pairs
, and no asymptotics). Both of these results generalise my older theorem with Ben Green on the primes containing arbitrarily long arithmetic progressions. The theorem also extends to multidimensional polynomials, in which case there are some additional previous results; see the paper for more details. We also get a technical refinement of our previous result on narrow polynomial progressions in (dense subsets of) the primes by making the progressions just a little bit narrower in the case of the density of the set one is using is small.
This week I have been at a Banff workshop “Combinatorics meets Ergodic theory“, focused on the combinatorics surrounding Szemerédi’s theorem and the Gowers uniformity norms on one hand, and the ergodic theory surrounding Furstenberg’s multiple recurrence theorem and the Host-Kra structure theory on the other. This was quite a fruitful workshop, and directly inspired the various posts this week on this blog. Incidentally, BIRS being as efficient as it is, videos for this week’s talks are already online.
As mentioned in the previous two posts, Ben Green, Tamar Ziegler, and myself proved the following inverse theorem for the Gowers norms:
Theorem 1 (Inverse theorem for Gowers norms) Let
and
be integers, and let
. Suppose that
is a function supported on
such that
Then there exists a filtered nilmanifold
of degree
and complexity
, a polynomial sequence
, and a Lipschitz function
of Lipschitz constant
such that
There is a higher dimensional generalisation, which first appeared explicitly (in a more general form) in this preprint of Szegedy (which used a slightly different argument than the one of Ben, Tammy, and myself; see also this previous preprint of Szegedy with related results):
Theorem 2 (Inverse theorem for multidimensional Gowers norms) Let
and
be integers, and let
. Suppose that
is a function supported on
such that
Then there exists a filtered nilmanifold
of degree
and complexity
, a polynomial sequence
, and a Lipschitz function
of Lipschitz constant
such that
The case of this theorem was recently used by Wenbo Sun. One can replace the polynomial sequence with a linear sequence if desired by using a lifting trick (essentially due to Furstenberg, but which appears explicitly in Appendix C of my paper with Ben and Tammy).
In this post I would like to record a very neat and simple observation of Ben Green and Nikos Frantzikinakis, that uses the tool of Freiman isomorphisms to derive Theorem 2 as a corollary of the one-dimensional theorem. Namely, consider the linear map defined by
that is to say is the digit string base
that has digits
. This map is a linear map from
to a subset of
of density
. Furthermore it has the following “Freiman isomorphism” property: if
lie in
with
in the image set
of
for all
, then there exist (unique) lifts
such that
and
for all . Indeed, the injectivity of
on
uniquely determines the sum
for each
, and one can use base
arithmetic to verify that the alternating sum of these sums on any
-facet of the cube
vanishes, which gives the claim. (In the language of additive combinatorics, the point is that
is a Freiman isomorphism of order (say)
on
.)
Now let be the function defined by setting
whenever
, with
vanishing outside of
. If
obeys (1), then from the above Freiman isomorphism property we have
Applying the one-dimensional inverse theorem (Theorem 1), with reduced by a factor of
and
replaced by
, this implies the existence of a filtered nilmanifold
of degree
and complexity
, a polynomial sequence
, and a Lipschitz function
of Lipschitz constant
such that
which by the Freiman isomorphism property again implies that
But the map is clearly a polynomial map from
to
(the composition of two polynomial maps is polynomial, see e.g. Appendix B of my paper with Ben and Tammy), and the claim follows.
Remark 3 This trick appears to be largely restricted to the case of boundedly generated groups such as
; I do not see any easy way to deduce an inverse theorem for, say,
from the
-inverse theorem by this method.
Remark 4 By combining this argument with the one in the previous post, one can obtain a weak ergodic inverse theorem for
-actions. Interestingly, the Freiman isomorphism argument appears to be difficult to implement directly in the ergodic category; in particular, there does not appear to be an obvious direct way to derive the Host-Kra inverse theorem for
actions (a result first obtained in the PhD thesis of Griesmer) from the counterpart for
actions.
Note: this post is of a particularly technical nature, in particular presuming familiarity with nilsequences, nilsystems, characteristic factors, etc., and is primarily intended for experts.
As mentioned in the previous post, Ben Green, Tamar Ziegler, and myself proved the following inverse theorem for the Gowers norms:
Theorem 1 (Inverse theorem for Gowers norms) Let
and
be integers, and let
. Suppose that
is a function supported on
such that
Then there exists a filtered nilmanifold
of degree
and complexity
, a polynomial sequence
, and a Lipschitz function
of Lipschitz constant
such that
This result was conjectured earlier by Ben Green and myself; this conjecture was strongly motivated by an analogous inverse theorem in ergodic theory by Host and Kra, which we formulate here in a form designed to resemble Theorem 1 as closely as possible:
Theorem 2 (Inverse theorem for Gowers-Host-Kra seminorms) Let
be an integer, and let
be an ergodic, countably generated measure-preserving system. Suppose that one has
for all non-zero
(all
spaces are real-valued in this post). Then
is an inverse limit (in the category of measure-preserving systems, up to almost everywhere equivalence) of ergodic degree
nilsystems, that is to say systems of the form
for some degree
filtered nilmanifold
and a group element
that acts ergodically on
.
It is a natural question to ask if there is any logical relationship between the two theorems. In the finite field category, one can deduce the combinatorial inverse theorem from the ergodic inverse theorem by a variant of the Furstenberg correspondence principle, as worked out by Tamar Ziegler and myself, however in the current context of -actions, the connection is less clear.
One can split Theorem 2 into two components:
Theorem 3 (Weak inverse theorem for Gowers-Host-Kra seminorms) Let
be an integer, and let
be an ergodic, countably generated measure-preserving system. Suppose that one has
for all non-zero
, where
. Then
is a factor of an inverse limit of ergodic degree
nilsystems.
Theorem 4 (Pro-nilsystems closed under factors) Let
be an integer. Then any factor of an inverse limit of ergodic degree
nilsystems, is again an inverse limit of ergodic degree
nilsystems.
Indeed, it is clear that Theorem 2 implies both Theorem 3 and Theorem 4, and conversely that the two latter theorems jointly imply the former. Theorem 4 is, in principle, purely a fact about nilsystems, and should have an independent proof, but this is not known; the only known proofs go through the full machinery needed to prove Theorem 2 (or the closely related theorem of Ziegler). (However, the fact that a factor of a nilsystem is again a nilsystem was established previously by Parry.)
The purpose of this post is to record a partial implication in reverse direction to the correspondence principle:
As mentioned at the start of the post, a fair amount of familiarity with the area is presumed here, and some routine steps will be presented with only a fairly brief explanation.
Recent Comments