In Notes 5, we saw that the Gowers uniformity norms on vector spaces in high characteristic were controlled by classical polynomial phases
.
Now we study the analogous situation on cyclic groups . Here, there is an unexpected surprise: the polynomial phases (classical or otherwise) are no longer sufficient to control the Gowers norms
once
exceeds
. To resolve this problem, one must enlarge the space of polynomials to a larger class. It turns out that there are at least three closely related options for this class: the local polynomials, the bracket polynomials, and the nilsequences. Each of the three classes has its own strengths and weaknesses, but in my opinion the nilsequences seem to be the most natural class, due to the rich algebraic and dynamical structure coming from the nilpotent Lie group undergirding such sequences. For reasons of space we shall focus primarily on the nilsequence viewpoint here.
Traditionally, nilsequences have been defined in terms of linear orbits on nilmanifolds
; however, in recent years it has been realised that it is convenient for technical reasons (particularly for the quantitative “single-scale” theory) to generalise this setup to that of polynomial orbits
, and this is the perspective we will take here.
A polynomial phase on a finite abelian group
is formed by starting with a polynomial
to the unit circle, and then composing it with the exponential function
. To create a nilsequence
, we generalise this construction by starting with a polynomial
into a nilmanifold
, and then composing this with a Lipschitz function
. (The Lipschitz regularity class is convenient for minor technical reasons, but one could also use other regularity classes here if desired.) These classes of sequences certainly include the polynomial phases, but are somewhat more general; for instance, they almost include bracket polynomial phases such as
. (The “almost” here is because the relevant functions
involved are only piecewise Lipschitz rather than Lipschitz, but this is primarily a technical issue and one should view bracket polynomial phases as “morally” being nilsequences.)
In these notes we set out the basic theory for these nilsequences, including their equidistribution theory (which generalises the equidistribution theory of polynomial flows on tori from Notes 1) and show that they are indeed obstructions to the Gowers norm being small. This leads to the inverse conjecture for the Gowers norms that shows that the Gowers norms on cyclic groups are indeed controlled by these sequences.
— 1. General theory of polynomial maps —
In previous notes, we defined the notion of a (non-classical) polynomial map of degree at most
between two additive groups
, to be a map
obeying the identity
for all , where
is the additive discrete derivative operator.
There is another way to view this concept. For any , define the Host-Kra group
of
of dimension
and degree
to be the subgroup of
consisting of all tuples
obeying the constraints
for all faces of the unit cube
of dimension at least
, where
. (These constraints are of course trivial if
.) A
-dimensional face of the unit cube
is of course formed by freezing
of the coordinates to a fixed value in
, and letting the remaining
coordinates vary freely in
.
Thus for instance is (essentially) space of parallelograms
in
, while
is the diagonal group
, and
is all of
.
Exercise 1 Let
be a map between additive groups, and let
. Show that
is a (non-classical) polynomial of degree at most
if it maps
to
, i.e. that
whenever
.
It turns out (somewhat remarkably) that these notions can be satisfactorily generalised to non-abelian setting, this was first observed by Leibman (in these papers, and also later by personal communication, in which the role of the Host-Kra group was emphasised). The (now multiplicative) groups need to be equipped with an additional structure, namely that of a filtration.
Definition 1 (Filtration) A filtration on a multiplicative group
is a family
of subgroups of
obeying the nesting property
and the filtration property
for all
, where
is the group generated by
, where
is the commutator of
and
. We will refer to the pair
as a filtered group. We say that an element
of
has degree
if it belongs to
, thus for instance a degree
and degree
element will commute modulo
errors.
In practice we usually have . As such, we see that
for all
, and so all the
are normal subgroups of
.
Exercise 2 Define the lower central series
of a group
by setting
and
for
. Show that the lower central series
is a filtration of
. Furthermore, show that the lower central series is the minimal filtration that starts at
, in the sense that if
is any other filtration with
, then
for all
.
Example 1 If
is an abelian group, and
, we define the degree
filtration
on
by setting
if
and
for
.
Example 2 If
is a filtered group, and
, we define the shifted filtered group
; this is clearly again a filtered group.
Definition 2 (Host-Kra groups) Let
be a filtered group, and let
be an integer. The Host-Kra group
is the subgroup of
generated by the elements
with
an arbitrary face in
and
an element of
, where
is the element of
whose coordinate at
is equal to
when
and equal to
otherwise.
From construction we see that the Host-Kra group is symmetric with respect to the symmetry group of the unit cube
. We will use these symmetries implicitly in the sequel without further comment.
Example 3 Let us parameterise an element of
as
. Then
is generated by elements of the form
for
,
and
, and
for
. (This does not cover all the possible faces of
, but it is easy to see that the remaining faces are redundant.) In other words,
consists of all group elements of the form
, where
,
, and
. This example is generalised in the exercise below.
Exercise 3 Define a lower face to be a face of a discrete cube
in which all the frozen coefficients are equal to
. Let us order the lower faces as
in such a way that
whenever
is a subface of
. Let
be a filtered group. Show that every element of
has a unique representation of the form
, where
and the product is taken from left to right (say).
Exercise 4 If
is an abelian group, show that the group
defined in Definition 2 agrees with the group defined at the beginning of this section for additive groups (after transcribing the former to multiplicative notation).
Exercise 5 Let
be a filtered group. Let
be an
-dimensional face of
. Identifying
with
in an obvious manner, we then obtain a restriction homomorphism from
with
. Show that the restriction of any element of
to
then lies in
.
Exercise 6 Let
be a filtered group, let
and
be integers, and let
and
be elements of
. Let
be the element of
defined by setting
for
to equal
for
, and equal to
otherwise. Show that
if and only if
and
, where
is defined in Example 2. (Hint: use Exercises 3, 5.)
Exercise 7 Let
be a filtered group, let
, and let
be an element of
. We define the derivative
in the first variable to be the tuple
. Show that
if and only if the restriction of
to
lies in
and
lies in
, where
is defined in Example 2.
Remark 1 The the Host-Kra groups of a filtered group in fact form a cubic complex, a concept used in topology; but we will not pursue this connection here.
In analogy with Exercise 1, we can now define the general notion of a polynomial map:
Definition 3 A map
between two filtered groups
is said to be polynomial if it maps
to
for each
. The space of all such maps is denoted
.
Since are groups, we immediately obtain
Theorem 4 (Lazard-Leibman theorem)
forms a group under pointwise multiplication.
(From our choice of definitions, this theorem is a triviality, but the theorem is less trivial when using an alternate but non-trivially equivalent definition of a polynomial, which we will give shortly.) In a similar spirit, we have
Theorem 5 (Filtered groups and polynomial maps form a category) If
and
are polynomial maps between filtered groups
, then
is also a polynomial map.
We can also give some basic examples of polynomial maps. Any constant map from to
taking values in
is polynomial, as is any map
which is a filtered homomorphism in the sense that it is a homomorphism from
to
for any
.
Now we turn to an alternate definition of a polynomial map. For any and any map
Define the multiplicative derivative
by the formula
.
Theorem 6 (Alternate description of polynomials) Let
be a map between two filtered groups
. Then
is polynomial if and only if, for any
,
, and
for
, one has
.
In particular, from Exercise 1, we see that a non-classical polynomial of degree from one additive group
to another
is the same thing as a polynomial map from
to
. More generally, a
map from
to a filtered group
is polynomial if and only if
for all and
.
Proof: We first prove the “only if” direction. It is clear (by using -dimensional cubes) that a polynomial map must map
to
. To obtain the remaining cases, it suffices by induction on
to show that if
is polynomial from
to
, and
for some
, then
is polynomial from
to
. But this is easily seen from Exercise 7.
Now we establish the “if” direction. We need to show that maps
to
for each
. We establish this by induction on
. The case
is trivial, so suppose that
and that the claim has already been estabilshed for all smaller values of
.
Let . We split
as
. From Exercise 7 we see that we can write
where
and
, thus
(extending
to act on
or
in the obvious manner). By induction hypothesis,
, so by Exercise 7, it suffices to show that
.
By telescoping series, it suffices to establish this when for some face
of some dimension
in
and some
, as these elements generate
. But then
vanishes outside of
and is equal to
on
, so by Exercise 6 it will suffice to show that
, where
is
restricted to
(which one then identifies with
). But by the induction hypothesis,
maps
to
, and the claim then follows from Exercise 5.
Exercise 8 Let
be integers. If
is a filtered group, define
to be the subgroup of
generated by the elements
, where
ranges over all faces of
and
, where
are the coordinates of
that are frozen. This generalises the Host-Kra groups
, which correspond to the case
. Show that if
is a polynomial map from
to
, then
maps
to
.
Exercise 9 Suppose that
is a non-classical polynomial of degree
from one additive group to another. Show that
is a polynomial map from
to
for every
. Conclude in particular that the composition of a non-classical polynomial of degree
and a non-classical polynomial of degree
is a non-classical polynomial of degree
.
Exercise 10 Let
,
be non-classical polynomials of degrees
,
respectively between additive groups
, and let
be a bihomomorphism to another additive group (i.e.
is a homomorphism in each variable separately). Show that
is a non-classical polynomial of degree
.
— 2. Nilsequences —
We now specialise the above theory of polynomial maps to the case when
is just the integers
(viewed additively) and
is a nilpotent group. Recall that a group
is nilpotent of step at most
if the
group
in the lower central series vanishes; thus for instance a group is nilpotent of step at most
if and only if it is abelian. Analogously, let us call a filtered group
nilpotent of degree at most
if
is nilpotent and
vanishes. Note that if
and
is nilpotent of degree at most
, then
is nilpotent of step at most
. On the other hand, the degree of a filtered group can exceed the step; for instance, given an additive group
and an integer
,
has degree
but step
. The step is the traditional measure of nilpotency for groups, but the degree seems to be a more suitable measure in the filtered group category. One is primarily interested in the case when
, but for technical reasons it is occasionally convenient to allow
to be strictly less than
, although this does not add much generality (see Exercise 18 below).
We refer to sequences which are polynomial maps from
to
as polynomial sequences or Hall-Petresco sequences adapted to
. The space of all such sequences is denoted
; by the machinery of the previous section, this is a multiplicative group. These sequences can be described explicitly:
Exercise 11 Let
be an integer, and let
be a filtered group which is nilpotent of degree
. Show that a sequence
is a Hall-Petresco sequence if and only if one has
for all
and some
for
, where
. Furthermore, show that the
are unique. We refer to the
as the Taylor coefficients of
at the origin.
Exercise 12 In a degree
nilpotent group
, establish the formula
for all
and
. This is the first non-trivial case of the Hall-Petresco formula, a discrete analogue of the Baker-Campbell-Hausdorff formula that expresses the polynomial sequence
explicitly in the form (1).
Define a nilpotent filtered Lie group of degree to be a nilpotent filtered group of degree
, in which
and all of the
are connected, simply connected finite-dimensional Lie groups. A model example here is the Heisenberg group, which is the degree
nilpotent filtered Lie group
(i.e. the group of upper-triangular unipotent matrices with arbitrary real entries in the upper triangular positions) with
and trivial for
(so in this case,
is also the lower central series).
Exercise 13 Show that a sequence
from
to the Heisenberg group
is a polynomial sequence if and only if
are linear polynomials and
is a quadratic polynomial.
It is a standard fact in the theory of Lie groups that a connected, simply connected nilpotent Lie group is topologically equivalent to its Lie algebra
, with the homeomorphism given by the exponential map
(or its inverse, the logarithm function
. Indeed, the Baker-Campbell-Hausdorff formula lets one use the nilpotent Lie algebra
to build a connected, simply connected Lie group with that Lie algebra, which is then necessarily isomorphic to
. One can thus classify filtered nilpotent Lie groups in terms of filtered nilpotent Lie algebras, i.e. a nilpotent Lie algebras
together with a nested family of sub-Lie algebras
with the inclusions (in which the bracket is now the Lie bracket rather than the commutator). One can describe such filtered nilpotent Lie algebras even more precisely using Mal’cev bases; see these papers of Mal’cev and of Leibman. For instance, in the case of the Heisenberg group, one has
and
From the filtration property, we see that for , each
is a normal closed subgroup of
, and for
, the quotient group
is connected, simply connected abelian Lie group (with Lie algebra
), and is thus isomorphic to a vector space (with the additive group law). Related to this, one can view
as a group extension of the quotient group
(with the degree
filtration
) by the central vector space
. Thus one can view degree
filtered nilpotent groups as an
-fold iterated tower of central extensions by finite-dimensional vector spaces starting from the base space
(which is a point in the most important case
); for instance, the Heisenberg group is an extension of
by
.
We thus see that nilpotent filtered Lie groups are generalisations of vector spaces (which correspond to the degree case). We now turn to filtered nilmanifolds, which are generalisations of tori. A degree
filtered nilmanifold
is a filtered degree
nilpotent Lie group
, together with a discrete subgroup
of
, such that all the subgroups
in the filtration are rational relative to
, which means that the subgroup
is a cocompact subgroup of
(i.e. the quotient space
is cocompact, or equivalently one can write
for some compact subset
of
. Note that the subgroups
give
the structure of a degree
filtered nilpotent group
.
Exercise 14 Let
and
, and let
. Show that the subgroup
of
is rational relative to
if and only if
is a rational number; this may help explain the terminology “rational”.
By hypothesis, the quotient space is a smooth compact manifold. The space
is a compact connected abelian Lie group, and is thus a torus; the degree
filtered nilmanifold
can then be viewed as a principal torus bundle over the degree
filtered nilmanifold
with
as the structure group; thus one can view degree
filtered nilmanifolds as an
-fold iterated tower of torus extensions starting from
, which is a point in the most important case
. For instance, the Heisenberg nilmanifold
is an extension of the two-dimensional torus by the circle
.
Every torus of some dimension can be viewed as a unit cube
with opposite faces glued together; up to measure zero sets, the cube then serves as a fundamental domain for the nilmanifold. Nilmanifolds can be viewed the same way, but the gluing can be somewhat “twisted”:
Exercise 15 Let
be the Heisenberg nilmanifold. If we abbreviate
for all
, show that for almost all
, that
has exactly one representation of the form
with
, which is given by the identity
where
is the greatest integer part of
, and
is the fractional part function. Conclude that
is topologically equivalent to the unit cube
quotiented by the identifications
between opposite faces.
Note that by using the projection
, we can view the Heisenberg nilmanifold
as a twisted circle bundle over
, with the fibers being isomorphic to the unit circle
. Show that
is not homeomorphic to
. (Hint: show that there are some non-trivial homotopies between loops that force the fundamental group of
to be smaller than
.)
The logarithm of the discrete cocompact subgroup
can be shown to be a lattice of the Lie algebra
. After a change of basis, one can thus view the latter algebra as a standard vector space
and the lattice as
. Denoting the standard generators of the lattice (and the standard basis of
) as
, we then see that the Lie bracket
of two such generators must be an integer combination of more generators:
The structure constants describe completely the Lie group structure of
and
. The rational subgroups
can also be described by picking some generators for
, which are integer combinations of the
. We say that the filtered nilmanifold has complexity at most
if the dimension and degree is at most
, and the structure constants and coefficients of the generators also have magnitude at most
. This is an admittedly artificial definition, but for quantitative applications it is necessary to have some means to quantify the complexity of a nilmanifold.
A polynomial orbit in a filtered nilmanifold is a map
of the form
, where
is a polynomial sequence. For instance, any linear orbit
, where
and
, is a polynomial orbit. The space of
Exercise 16 For any
, show that the sequence
(using the notation from Exercise 15) is a polynomial sequence in the Heisenberg nilmaniofold.
With the above example, we see the emergence of bracket polynomials when representing polynomial orbits in a fundamental domain. Indeed, one can view the entire machinery of orbits in nilmanifolds as a means of efficiently capturing such polynomials in an algebraically tractable framework (namely, that of polynomial sequences in nilpotent groups). The piecewise continuous nature of the bracket polynomials is then ultimately tied to the twisted gluing needed to identify the fundamental domain with the nilmanifold.
Finally, we can define the notion of a (basic Lipschitz) nilsequence of degree . This is a sequence
of the form
, where
is a polynomial orbit in a filtered nilmanifold of degree
, and
is a Lipschitz function. (One needs a metric on
to define the Lipschitz constant, but this can be done for instance by using a basis
of
to identify
with a fundamental domain
, and using this to construct some (artificial) metric on
. The details of such a construction will not be important here.) We say that the nilsequence has complexity at most
if the filtered nilmanifold has complexity at most
, and the (inhomogeneous Lipschitz norm) of
is also at most
.
A basic example of a degree nilsequence is a polynomial phase
, where
is a polynomial of degree
. A bit more generally,
is a degree
sequence, whenever
is a Lipschitz function. In view of Exercises 15, 16, we also see that
or more generally
are also degree nilsequences, where
is a Lipschitz function that vanishes near
and
. The
factor is not needed (as there is no twisting in the
coordinate in Exercise 15), but the
factor is (unfortunately) necessary, as otherwise one encounters the discontinuity inherent in the
term (and one would merely have a piecewise Lipschitz nilsequence rather than a genuinely Lipschitz nilsequence). Because of this discontinuity, bracket polynomial phases
cannot quite be viewed as Lipschitz nilsequences, but from a heuristic viewpoint it is often helpful to pretend as if bracket polynomial phases are model instances of nilsequences.
The only degree nilsequences are the constants. The degree
nilsequences are essentially the quasiperiodic functions:
Exercise 17 Show that a degree
nilsequence of complexity
is Fourier-measurable with growth function
depending only on
, where Fourier measurability was defined in Notes 2.
Exercise 18 Show that the class of nilsequences of degree
does not change if we drop the condition
, or if we add the additional condition
.
Remark 2 The space of nilsequences is also unchanged if one insists that the polynomial orbit be linear, and that the filtration be the lower central series filtration; and this is in fact the original definition of a nilsequence. The proof of this equivalence is a little tricky, though, and will appear in a forthcoming paper of Green, Ziegler, and myself.
— 3. Connection with the Gowers norms —
We define the Gowers norm of a function
by the formula
where is any integer greater than
,
is embedded inside
, and
is extended by zero outside of
. It is easy to see that this definition is independent of the choice of
. Note also that the normalisation factor
is comparable to
when
is fixed and
is comparable to
.
One of the main reasons why nilsequences are relevant to the theory of the Gowers norms is that they are an obstruction to that norm being small. More precisely, we have
Theorem 7 (Converse to the inverse conjecture for the Gowers norms) Let
be such that
and
for some degree
nilsequence of complexity at most
. Then
.
We now prove this theorem, following an argument of Green, Ziegler, and myself. It is convenient to introduce a few more notions. Define a vertical character of a degree filtered nilmanifold
to be a continuous homomorphism
that annihilates
, or equivalently an element of the Pontryagin dual
of the torus
. A function
is said to have vertical frequency
if
obeys the equation
for all and
. A degree
nilsequence is said to have a vertical frequency if it can be represented in the form
for some Lipschitz
with a vertical frequency.
For instance, a polynomial phase , where
is a polynomial of degree
, is a degree
nilsequence with a vertical frequency. Any nilsequence of degree
is trivially a nilsequence of degree
with a vertical frequency of
. Finally, observe that the space of degree
nilsequences with a vertical frequency is closed under multiplication and complex conjugation.
Exercise 19 Show that a degree
nilsequence with a vertical frequency necessarily takes the form
for some
and
(and conversely, all such sequences are degree
nilsequences with a vertical frequency). Thus, up to constants, degree
nilsequences with a vertical frequency are the same as Fourier characters.
A basic fact (generalising the invertibility of the Fourier transform in the degree case) is that the nilsequences with vertical frequency generate all the other nilsequences:
Exercise 20 Show that any degree
nilsequence can be approximated to arbitrary accuracy in the uniform norm by a linear combination of nilsequences with a vertical frequency. (Hint: use the Stone-Weierstrass theorem.)
More quantitatively, show that a degree
nilsequence of complexity
can be approximated uniformly to error
by a sum of
nilsequences, each with a representation with a vertical frequency that is of complexity
. (Hint: this can be deduced from the qualitative result by a compactness argument using the Arzelá-Ascoli theorem.)
A derivative of a polynomial phase is a polynomial phase of one lower degree. There is an analogous fact for nilsequences with a vertical frequency:
Lemma 8 (Differentiating nilsequences with a vertical frequency) Let
, and let
be a degree
nilsequence with a vertical frequency. Then for any
,
is a degree
nilsequence. Furthermore, if
has complexity
(with a vertical frequency representation), then
has complexity
.
Proof: We just prove the first claim, as the second claim follows by refining the argument.
We write for some polynomial sequence
and some Lipschitz function
with a vertical frequency. We then express
where is the function
and is the sequence
Now we give a filtration on by setting
for , where
is the subgroup of
generated by
and the diagonal group
. One easily verifies that this is a filtration on
. The sequences
and
are both polynomial with respect to this filtration, and hence by the Lazard-Leibman theorem,
is polynomial also.
Next, we use the hypothesis that has a vertical frequency to conclude that
is invariant with respect to the action of the diagonal group
. If we then define
to be the Lie group
with filtration
, then
is a degree
filtered nilpotent Lie group; setting
, we conclude that
is a degree
nilmanifold and
where are the projections of
from
to
. The claim follows.
We now prove Theorem 7 by induction on . The claim is trivial for
, so we assume that
and that the claim has already been proven for smaller values of
.
Let be as in Theorem 7. From Exercise 20 we see (after modifying
) that we may assume that
has a vertical frequency. Next, we use the identity
(extending by zero outside of
, and extending
arbitrarily) to conclude that
for values of
. By induction hypothesis and Lemma 8, we conclude that
for values of
. Using the identity
we close the induction and obtain the claim.
In the other direction, we have
Theorem 9 (Inverse conjecture for the Gowers norms on
) Let
be such that
and
. Then
for some degree
nilsequence of complexity
.
This conjecture has recently been proven by Green, Ziegler, and myself; an announcement of this result, which will contain extensive heuristic discussion of how this conjecture is proven, will appear very shortly, and the paper itself soon after that. For a discussion of the history of the conjecture, including the cases , see our previous paper.
Exercise 21 (
inverse theorem)
- (Straightening an approximately linear function) Let
. Let
be a function such that
for all but
of all
with
. If
is sufficiently small, show that there exists an affine linear function
with
such that
for all but
values of
, where
as
. (Hint: One can take
to be small. First find a way to lift
in a nice manner from
to
.)
- Let
be such that
and
. Show that there exists a polynomial
of degree
such that
, where
as
(holding
fixed). Hint: Adapt the argument of the analogous finite field statement. One cannot exploit the discrete nature of polynomials any more; and so one must use the preceding part of the exercise as a substitute.
The inverse conjecture for the Gowers norms, when combined with the equidistribution theory for nilsequences that we will turn to next, has a number of consequences, analogous to the consequences for the finite field analogues of these facts; see this paper of Green and myself for further discussion.
— 4. Equidistribution of nilsequences —
In the subject of higher order Fourier analysis, and in particular in the proof of the inverse conjecture for the Gowers norms, as well as in several of the applications of this conjecture, it will be of importance to be able to compute statistics of nilsequences , such as their averages
for a large integer
; this generalises the computation of exponential sums such as
that occurred in Notes 1. This is closely related to the equidistribution of polynomial orbits
in nilmanifolds. Note that as
is a compact quotient of a locally compact group
, it comes endowed with a unique left-invariant Haar measure
(which is isomorphic to the Lebesgue measure on a fundamental domain
of that nilmanifold). By default, when we talk about equidistribution in a nilmanifold, we mean with respect to the Haar measure; thus
is asymptotically equidistributed if and only if
for all Lipschitz . One can also describe single-scale equidistribution (and non-standard equidistribution) in a similar fashion, but for sake of discussion let us restrict attention to the simpler and more classical situation of asymptotic equidistribution here (although it is the single-scale equidistribution theory which is ultimately relevant to questions relating to the Gowers norms).
When studying equidistribution of polynomial sequences in a torus , a key tool was the van der Corput lemma. This lemma asserts that if a sequence
is such that all derivatives
with
are asymptotically equidistributed, then
itself is also asymptotically equidistributed.
The notion of a derivative requires the ability to perform subtraction on the range space :
. When working in a higher degree nilmanifold
, which is not a torus, we do not have a notion of subtraction. However, such manifolds are still torus bundles with torus
. This gives a weaker notion of subtraction, namely the map
, where
is the diagonal action
of the torus
on the product space
. This leads to a generalisation of the van der Corput lemma:
Lemma 10 (Relative van der Corput lemma) Let
be a sequence in a degree
nilmanifold for some
. Suppose that the projection of
to the degree
filtered nilmanifold
is asymptotically equidistributed, and suppose also that for each non-zero
, the sequence
is asymptotically equidistributed with respect to some
-invariant measure
on
. Then
is asymptotically equidistributed in
.
Proof: It suffices to show that, for each Lipschitz function , that
By Exercise 20, we may assume that has a vertical frequency. If this vertical frequency is non-zero, then
descends to a function on the degree
filtered nilmanifold
, and the claim then follows from the equidistribution hypothesis on this space. So suppose instead that
has a non-zero vertical frequency. By vertically rotating
(and using the
-invariance of
we conclude that
. Applying the van der Corput inequality (see Notes 1), we now see that it suffices to show that
for each non-zero . The function
on
is
-invariant (because of the vertical frequency hypothesis) and so descends to a function
on
. We thus have
The function has a non-zero vertical frequency with respect to the residual action of
(or more precisely, of
, which is isomorphic to
). As
is invariant with respect to this action, the integral thus vanishes, as required.
This gives a useful criterion for equidistribution of polynomial orbits. Define a horizontal character to be a continuous homomorphism from
to
that annihilates
(or equivalently, an element of the Pontryagin dual of the horizontal torus
). This is easily seen to be a torus. Let
be the projection map.
Theorem 11 (Leibman equidistribution criterion) Let
be a polynomial orbit on a degree
filtered nilmanifold
. Suppose that
. Then
is asymptotically equidistributed in
if and only if
is non-constant for each non-trivial horizontal character.
This theorem was first established by Leibman (by a slightly different method), and also follows from the above van der Corput lemma and some tedious additional computations; see this paper of Green and myself for details. For linear orbits, this result was established by Parry and by Leon Green. Using this criterion (together with more quantitative analogues for single-scale equidistribution), one can develop Ratner-type decompositions that generalise those in (Notes 1). Again, the details are technical and I refer to my paper with Green for details. We give a special case of Theorem 11 as an exercise:
Exercise 22 Use Lemma 10 to show that if
are two real numbers such that
are linearly independent modulo
over the integers, then the polynomial orbit
is asymptotically equidistributed in the Heisenberg nilmanifold
; note that this is a special case of Theorem 11. Conclude that the map
is asymptotically equidistributed in the unit circle.
Unfortunately Lemma 10 is not strong enough to cover all cases of Theorem 11; in particular, if are independent but
are not, then the hypotheses of Lemma 10 are not obeyed for any fixed non-zero
, although they are in some sense asymptotically obeyed in the limit when
is large. To obtain Theorem 11 in this case one either needs a quantitative (single-scale) version of Lemma 10, or else one has to invoke the ergodic theorem in a number of places. The former approach is the one taken in the above mentioned paper of Green and myself, and the latter in the paper of Leibman.
One application of this equidistribution theory is to show that bracket polynomial objects such as (2) have a negligible correlation with any genuinely quadratic phase (or more generally, with any genuinely polynomial phase of bounded degree); this result was first established by Haland. On the other hand, from Theorem 7 we know that (2) has a large
norm. This shows that even when
, one cannot invert the Gowers norm purely using polynomial phases. This observation first appeared in the work of Gowers (with a related observation due to Furstenberg and Weiss).
Exercise 23 Let the notation be as in Exercise 22. Show that
for any
. (You can either apply Theorem 11, or go back to Lemma 10.)
11 comments
Comments feed for this article
30 May, 2010 at 10:47 am
Bogdan
Dear Professor Tao
Am I correct that Theorem 9 (Inverse conjecture for the Gowers norms on Z) is exactly the same conjecture mentioned in the paper “Linear equations in primes”, and thus that results are now fully unconditional?
30 May, 2010 at 11:06 pm
Terence Tao
It is equivalent to the conjecture formulated there, yes. We’re preparing a formal announcement of the results which should appear here within a week or so, and the full paper should be ready shortly afterwards.
4 June, 2010 at 1:24 pm
Jingzheng
Dear Professor Tao
Inverse conjecture for Gowers norms roughly says for example that if a bounded function f does not resemble a nilsequence then the sum of terms f(a)g(a+b)h(a+2b)k(a+3b) (for a, b from suitable ranges and g,h,k bounded) is of smaller order than the magnitude resulting from estimating trivially. Are there any partial results or conjectures concerning nonlinear equations? For instance: is there any conjecture about sum of f(a)g(d)h(b)k(c) over quadruples (a,b,c,d) satisfying ad-bc=1? If this sum is as large as possible then what are the functions that f,g,h and k must resemble?
4 June, 2010 at 1:46 pm
Terence Tao
This is an important question, but not much is known presently, even in the simplest example of nonlinearity, namely averages over polynomials. If one is working in a finite field, and all parameters range freely in this field, then one can usually use a version of the van der Corput inequality (combined with PET induction) to control things by Gowers norms. But over an interval of integers [N], the problem once one has nonlinearity is that some of the parameters now need to be restricted to smaller intervals, such as [N^{1/d}] for some d. In some cases one can still hope to control things by “local” Gowers norms – Tamar Ziegler and I did something like this for instance when finding polynomial progressions in primes. But the general situation is still quite murky. I suspect that some of the theory from sum-product estimates and expansion in linear groups may have to come into play, particularly for patterns such as ad-bc=1 that are clearly connected to linear groups such as SL_2.
4 June, 2010 at 2:06 pm
Jingzheng
Dear Professor Tao
Thank you for the answer. Your mentioning of the paper with dr Ziegler about polynomial progressions of primes reminded me of another interesting problem. Since the progressions of length at least 4 involve as complicated objects as nilsequences then in order to try to find an asymptotics for polynomial progressions maybe it is appropriate first to investigate asymptotics for polynomial progressions of length 3? The simplest nontrivial case would be: a, a+b^2, a+2b^2. A relevant question here would be: is it true that if the sum of terms f(a)g(a+b^2)h(a+2b^2) is of magnitude N^(3/2) where a is of order N and b is of order N^(1/2) then necessarily f has a large Fourier coefficient?
4 June, 2010 at 3:25 pm
Terence Tao
Yes, something like this should be true (but perhaps one has to deal with a local Fourier coefficient rather than a global one), but this has not been proved yet as far as I know (this would correspond to a polynomial version of a conjecture of Gowers and Wolf, which has only been proven so far in the linear case). There is an ergodic counterpart to your question to which the answer is affirmative, thanks to the work of Bergelson, Leibman, and Lesigne, so heuristically the same should be true for the finitary version of the question, but unfortunately the correspondence principle does not seem to yield this directly.
If one applies the Cauchy-Schwarz inequality enough times, I believe one can control the above sum by something like the (local) U^7 norm of f, which suggests that f will at least correlate with a 6-step nilsequence, but one presumably do better than this with more work.
4 June, 2010 at 3:31 pm
Anonymous
陶教授
我非常感谢您回答
晶正
25 October, 2012 at 10:11 am
Walsh’s ergodic theorem, metastability, and external Cauchy convergence « What’s new
[…] some (standard) finite number of group elements of ; see e.g. Exercise 11 of this previous blog post. In the abelian case, we used the largest for which was non-trivial as the degree of . This turns […]
24 February, 2013 at 1:52 am
pavel zorin
Dear Terry,
(these are linearly independent mod 1 over Z because
is irreducible over Z, say by Cohn’s criterion in base 3). Then
it seems that the relative van der Corput lemma 10 does not apply directly in Exercise 22 (Exercise 1.6.22 in the book draft). Consider the following example:
Hence
.
-invariant then also the closure of the orbit
would be
-invariant since
is central and multiplication by a constant is a homeomorphism. However, this closure is contained in the diagonal of
.
If the closure of this orbit was
best regards,
pavel
24 February, 2013 at 7:53 pm
Terence Tao
Gah, you’re right, this particular argument only works under the stronger hypothesis that
are linearly independent. (In the example you gave, one does not have equidistribution for any fixed h, though in some sense the orbits become asymptotically equidistributed as h becomes large, though it is hard to formalise this at the qualitative level.) It seems that to handle the general case one has to either use a quantitative (single-scale) version of the van der Corput lemma (as was done in my paper with Ben) or else rely on the ergodic theorem (this is the approach in Leibman).
27 March, 2013 at 7:34 pm
An informal version of the Furstenberg correspondence principle | What's new
[…] the cost of turning the observable into a slightly messy piecewise smooth function; see e.g. this blog post for details. More generally, any function arising from phases that are bracket polynomials (or […]