A common theme in mathematical analysis (particularly in analysis of a “geometric” or “statistical” flavour) is the interplay between “macroscopic” and “microscopic” scales. These terms are somewhat vague and imprecise, and their interpretation depends on the context and also on one’s choice of normalisations, but if one uses a “macroscopic” normalisation, “macroscopic” scales correspond to scales that are comparable to unit size (i.e. bounded above and below by absolute constants), while “microscopic” scales are much smaller, being the minimal scale at which nontrivial behaviour occurs. (Other normalisations are possible, e.g. making the microscopic scale a unit scale, and letting the macroscopic scale go off to infinity; for instance, such a normalisation is often used, at least initially, in the study of groups of polynomial growth. However, for the theory of approximate groups, a macroscopic scale normalisation is more convenient.)
One can also consider “mesoscopic” scales which are intermediate between microscopic and macroscopic scales, or large-scale behaviour at scales that go off to infinity (and in particular are larger than the macroscopic range of scales), although the behaviour of these scales will not be the main focus of this post. Finally, one can divide the macroscopic scales into “local” macroscopic scales (less than for some small but fixed
) and “global” macroscopic scales (scales that are allowed to be larger than a given large absolute constant
). For instance, given a finite approximate group
:
- Sets such as
for some fixed
(e.g.
) can be considered to be sets at a global macroscopic scale. Sending
to infinity, one enters the large-scale regime.
- Sets such as the sets
that appear in the Sanders lemma from the previous set of notes (thus
for some fixed
, e.g.
) can be considered to be sets at a local macroscopic scale. Sending
to infinity, one enters the mesoscopic regime.
- The non-identity element
of
that is “closest” to the identity in some suitable metric (cf. the proof of Jordan’s theorem from Notes 0) would be an element associated to the microscopic scale. The orbit
starts out at microscopic scales, and (assuming some suitable “escape” axioms) will pass through mesoscopic scales and finally entering the macroscopic regime. (Beyond this point, the orbit may exhibit a variety of behaviours, such as periodically returning back to the smaller scales, diverging off to ever larger scales, or filling out a dense subset of some macroscopic set; the escape axioms we will use do not exclude any of these possibilities.)
For comparison, in the theory of locally compact groups, properties about small neighbourhoods of the identity (e.g. local compactness, or the NSS property) would be properties at the local macroscopic scale, whereas the space of one-parameter subgroups can be interpreted as an object at the microscopic scale. The exponential map then provides a bridge connecting the microscopic and macroscopic scales.
We return now to approximate groups. The macroscopic structure of these objects is well described by the Hrushovski Lie model theorem from the previous set of notes, which informally asserts that the macroscopic structure of an (ultra) approximate group can be modeled by a Lie group. This is already an important piece of information about general approximate groups, but it does not directly reveal the full structure of such approximate groups, because these Lie models are unable to see the microscopic behaviour of these approximate groups.
To illustrate this, let us review one of the examples of a Lie model of an ultra approximate group, namely Exercise 28 from Notes 7. In this example one studied a “nilbox” from a Heisenberg group, which we rewrite here in slightly different notation. Specifically, let be the Heisenberg group
and let , where
is the box
thus is the nonstandard box
where . As the above exercise establishes,
is an ultra approximate group with a Lie model
given by the formula
for and
. Note how the nonabelian nature of
(arising from the
term in the group law (1)) has been lost in the model
, because the effect of that nonabelian term on
is only
which is infinitesimal and thus does not contribute to the standard part. In particular, if we replace
with the abelian group
with the additive group law
and let and
be defined exactly as with
and
, but placed inside the group structure of
rather than
, then
and
are essentially “indistinguishable” as far as their models by
are concerned, even though the latter approximate group is abelian and the former is not. The problem is that the nonabelian-ness in the former example is so microscopic that it falls entirely inside the kernel of
and is thus not detected at all by the model.
The problem of not being able to “see” the microscopic structure of a group (or approximate group) also was a key difficulty in the theory surrounding Hilbert’s fifth problem that was discussed in previous notes. A key tool in being able to resolve such structure was to build left-invariant metrics (or equivalently, norms
) on one’s group, which obeyed useful “Gleason axioms” such as the commutator axiom
for sufficiently small , or the escape axiom
when was sufficiently small. Such axioms have important and non-trivial content even in the microscopic regime where
or
are extremely close to the identity. For instance, in the proof of Jordan’s theorem from Notes 0, which showed that any finite unitary group
was boundedly virtually abelian, a key step was to apply the commutator axiom (2) (for the distance to the identity in operator norm) to the most “microscopic” element of
, or more precisely a non-identity element of
of minimal norm. The key point was that this microscopic element was virtually central in
, and as such it restricted much of
to a lower-dimensional subgroup of the unitary group, at which point one could argue using an induction-on-dimension argument. As we shall see, a similar argument can be used to place “virtually nilpotent” structure on finite approximate groups. For instance, in the Heisenberg-type approximate groups
and
discussed earlier, the element
will be “closest to the origin” in a suitable sense to be defined later, and is centralised by both approximate groups; quotienting out (the orbit of) that central element and iterating the process two more times, we shall see that one can express both
and
as a tower of central cyclic extensions, which in particular establishes the nilpotency of both groups.
The escape axiom (3) is a particularly important axiom in connecting the microscopic structure of a group to its macroscopic structure; for instance, as shown in Notes 2, this axiom (in conjunction with the closely related commutator axiom) tends to imply dilation estimates such as
that allow one to understand the microscopic geometry of points
close to the identity in terms of the (local) macroscopic geometry of points
that are significantly further away from the identity.
It is thus of interest to build some notion of a norm (or left-invariant metrics) on an approximate group that obeys the escape and commutator axioms (while being non-degenerate enough to adequately capture the geometry of
in some sense), in a fashion analogous to the Gleason metrics that played such a key role in the theory of Hilbert’s fifth problem. It is tempting to use the Lie model theorem to do this, since Lie groups certainly come with Gleason metrics. However, if one does this, one ends up, roughly speaking, with a norm on
that only obeys the escape and commutator estimates macroscopically; roughly speaking, this means that one has a macroscopic commutator inequality
and a macroscopic escape property
but such axioms are too weak for analysis at the microscopic scale, and in particular in establishing centrality of the element closest to the identity.
Another way to proceed is to build a norm that is specifically designed to obey the crucial escape property. Given an approximate group in a group
, and an element
of
, we can define the escape norm
of
by the formula
Thus, equals
if
lies outside of
, equals
if
lies in
but
lies outside of
, and so forth. Such norms had already appeared in Notes 4, in the context of analysing NSS groups.
As it turns out, this expression will obey an escape axiom, as long as we place some additional hypotheses on which we will present shortly. However, it need not actually be a norm; in particular, the triangle inequality
is not necessarily true. Fortunately, it turns out that by a (slightly more complicated) version of the Gleason machinery from Notes 4 we can establish a usable substitute for this inequality, namely the quasi-triangle inequality
where is a constant independent of
. As we shall see, these estimates can then be used to obtain a commutator estimate (2).
However, to do all this, it is not enough for to be an approximate group; it must obey two additional “trapping” axioms that improve the properties of the escape norm. We formalise these axioms (somewhat arbitrarily) as follows:
Definition 1 (Strong approximate group) Let
. A strong
-approximate group is a finite
-approximate group
in a group
with a symmetric subset
obeying the following axioms:
An ultra strong
-approximate group is an ultraproduct
of strong
-approximate groups.
The first trapping condition can be rewritten as
and the second trapping condition can similarly be rewritten as
This makes the escape norms of , and
comparable to each other, which will be needed for a number of reasons (and in particular to close a certain bootstrap argument properly). Compare this with equation (12) from Notes 4, which used the NSS hypothesis to obtain similar conclusions. Thus, one can view the strong approximate group axioms as being a sort of proxy for the NSS property.
Example 1 Let
be a large natural number. Then the interval
in the integers is a
-approximate group, which is also a strong
-approximate group (setting
, for instance). On the other hand, if one places
in
rather than in the integers, then the first trapping condition is lost and one is no longer a strong
-approximate group. Also, if one remains in the integers, but deletes a few elements from
, e.g. deleting
from
), then one is still a
-approximate group, but is no longer a strong
-approximate group, again because the first trapping condition is lost.
A key consequence of the Hrushovski Lie model theorem is that it allows one to replace approximate groups by strong approximate groups:
Exercise 1 (Finding strong approximate groups)
- (i) Let
be an ultra approximate group with a good Lie model
, and let
be a symmetric convex body (i.e. a convex open bounded subset) in the Lie algebra
. Show that if
is a sufficiently small standard number, then there exists a strong ultra approximate group
with
and with
can be covered by finitely many left translates of
. Furthermore,
is also a good model for
.
- (ii) If
is a finite
-approximate group, show that there is a strong
-approximate group
inside
with the property that
can be covered by
left translates of
. (Hint: use (i), Hrushovski’s Lie model theorem, and a compactness and contradiction argument.)
The need to compare the strong approximate group to an exponentiated small ball will be convenient later, as it allows one to easily use the geometry of
to track various aspects of the strong approximate group.
As mentioned previously, strong approximate groups exhibit some of the features of NSS locally compact groups. In Notes 4, we saw that the escape norm for NSS locally compact groups was comparable to a Gleason metric. The following theorem is an analogue of that result:
Theorem 2 (Gleason lemma) Let
be a strong
-approximate group in a group
.
- (Symmetry) For any
, one has
.
- (Conjugacy bound) For any
, one has
.
- (Triangle inequality) For any
, one has
.
- (Escape property) One has
whenever
.
- (Commutator inequality) For any
, one has
.
The proof of this theorem will occupy a large part of the current set of notes. We then aim to use this theorem to classify strong approximate groups. The basic strategy (temporarily ignoring a key technical issue) follows the Bieberbach-Frobenius proof of Jordan’s theorem, as given in Notes 0, is as follows.
- Start with an (ultra) strong approximate group
.
- From the Gleason lemma, the elements with zero escape norm form a normal subgroup of
. Quotient these elements out. Show that all non-identity elements will have positive escape norm.
- Find the non-identity element
in (the quotient of)
of minimal escape norm. Use the commutator estimate (assuming it is inherited by the quotient) to show that
will centralise (most of) this quotient. In particular, the orbit
is (essentially) a central subgroup of
.
- Quotient this orbit out; then find the next non-identity element
in this new quotient of
. Again, show that
is essentially a central subgroup of this quotient.
- Repeat this process until
becomes entirely trivial. Undoing all the quotients, this should demonstrate that
is virtually nilpotent, and that
is essentially a coset nilprogression.
There are two main technical issues to resolve to make this strategy work. The first is to show that the iterative step in the argument terminates in finite time. This we do by returning to the Lie model theorem. It turns out that each time one quotients out by an orbit of an element that escapes, the dimension of the Lie model drops by at least one. This will ensure termination of the argument in finite time.
The other technical issue is that while the quotienting out all the elements of zero escape norm eliminates all “torsion” from (in the sense that the quotient of
has no non-trivial elements of zero escape norm), further quotienting operations can inadvertently re-introduce such torsion. This torsion can be re-eradicated by further quotienting, but the price one pays for this is that the final structural description of
is no longer as strong as “virtually nilpotent”, but is instead a more complicated tower alternating between (ultra) finite extensions and central extensions.
Example 2 Consider the strong
-approximate group
in the integers, where
is a large natural number not divisible by
. As
is torsion-free, all non-zero elements of
have positive escape norm, and the nonzero element of minimal escape norm here is
(or
). But if one quotients by
,
projects down to
, which now has torsion (and all elements in this quotient have zero escape norm). Thus torsion has been re-introduced by the quotienting operation. (A related observation is that the intersection of
with
is not a simple progression, but is a more complicated object, namely a generalised arithmetic progression of rank two.)
To deal with this issue, we will not quotient out by the entire cyclic group generated by the element
of minimal escape norm, but rather by an arithmetic progression
, where
is a natural number comparable to the reciprocal
of the escape norm, as this will be enough to cut the dimension of the Lie model down by one without introducing any further torsion. Of course, this cannot be done in the category of global groups, since the arithmetic progression
will not, in general, be a group. However, it is still a local group, and it turns out that there is an analogue of the quotient space construction in local groups. This fixes the problem, but at a cost: in order to make the inductive portion of the argument work smoothly, it is now more natural to place the entire argument inside the category of local groups rather than global groups, even though the primary interest in approximate groups
is in the global case when
lies inside a global group. This necessitates some technical modification to some of the preceding discussion (for instance, the Gleason-Yamabe theorem must be replaced by the local version of this theorem, due to Goldbring); details can be found in this recent paper of Emmanuel Breuillard, Ben Green, and myself, but will only be sketched here.
— 1. Gleason’s lemma —
Throughout this section, is a strong
-approximate group in a global group
. We will prove the various estimates in Theorem 2. The arguments will be very close to those in Notes 4; indeed, it is possible to unify the results here with the results in those notes by a suitable modification of the notation, but we will not do so here.
We begin with the easy estimates. The symmetry property is immediate from the symmetry of . Now we turn to the escape property. By symmetry, we may take
to be positive (the
case is trivial). We may of course assume that
is strictly positive, say equal to
; thus
and
, and
. By the first trapping property, this implies that
for some
.
Let be the first multiple of
larger than or equal to
, then
. Since
is less than
, we have
; since
, we conclude that
. In particular this shows that
, and the claim follows.
The escape property implies the conjugacy bound:
Exercise 2 Establish the conjugacy bound. (Hint: one can mimic the arguments establishing a nearly identical bound in Section 2 of Notes 4.)
Now we turn to the triangle inequality, which (as in Notes 4) is the most difficult property to establish. Our arguments will closely resemble the proof of Proposition 11 from these notes, with and
playing the roles of
and
from that argument. As in that theorem, we will initially assume that we have an a priori bound of the form
for all , and some (large)
independent of
, and remove this hypothesis later. We then introduce the norm
then is symmetric, obeys the triangle inequality, and is comparable to
in the sense that
for all .
We introduce the function by
where . Then
takes values between
and
, equals
on
, is supported on
, and obeys the Lipschitz bound
for all , thanks to the triangle inequality for
and (6). We also introduce the function
by
then also takes values between
and
, equals
on
, is supported on
, and obeys the Lipschitz bound
for all and
.
Now we form the convolution by the formula
By construction, is supported on
and at least
at the identity. As
or
is supported in
, which has cardinality at most
, we have the uniform bound
Similarly, from the identity
and (7) we have the Lipschitz bound
Finally, from the identity
and restricting to
(so that
is supported on
, which has cardinality at most
) we see from (7), (8) that
for and
.
We can use this to improve the bound (10). Indeed, using the telescoping identity
we see that
and thus
whenever . Using the second trapping property, this implies that
In the converse direction, if , then
and thus from the support of
, for all
. But then by the first trapping property, this implies that
for all
. We conclude that
The triangle inequality for then implies a triangle inequality for
,
which is (5) with replaced by
. If we knew (5) for some large but finite
, we could iterate this argument and conclude that (5) held with
replaced by
, which would give the triangle inequality. Now it is not immediate that (5) holds for any finite
, but we can avoid this problem with the usual regularisation trick of replacing
with
throughout the argument for some small
, which makes (5) automatically true with
, run the above iteration argument, and then finally send
to zero.
Exercise 3 Verify that the modifications to the above argument sketched above actually do establish the triangle inequality.
A final application of the Gleason convolution machinery then gives the final estimate in Gleason’s lemma:
Exercise 4 Use the properties of the escape norm already established (and in particular, the escape property and the triangle inequality) to establish the commutator inequality. (Hint: adapt the argument from Section 2 of Notes 4.)
The proof of Theorem 2 is now complete.
Exercise 5 Generalise Theorem 2 to the setting where
is not necessarily finite, but is instead an open precompact subset of a locally compact group
. (Hint: replace cardinality by left-invariant Haar measure and follow the arguments in Notes 4 closely.) Note that this already gives most of one of the key results from that set of notes, namely that any NSS group admits a Gleason metric, since it is not difficult to show that NSS groups contain open precompact strong approximate groups.
— 2. A cheap version of the structure theorem —
In this section we use Theorem 2 to establish a “cheap” version of the structure theorem for ultra approximate groups. We begin by eliminating the elements of zero escape norm. Let us say that an approximate group is NSS if it contains no non-trivial subgroups of the ambient group, or equivalently if every non-identity element of
has a positive escape norm. We say that an ultra approximate group is NSS if it is the ultralimit of NSS approximate groups.
Using the Gleason lemma, we can easily reduce to the NSS case:
Exercise 6 (Reduction to the NSS case) Let
be a connected Lie group with Lie algebra
, let
be a bounded symmetric convex body in
, let
be a sufficiently small standard real. Let
, and let
be an ultra strong approximate group which has a good model
with
Let
be the set of all elements in
of zero (nonstandard) escape norm. Show that
is a normal nonstandard finite subgroup of
that lies in
. If
is the quotient map, and
is the map
factored through
, show that there exists an ultra strong NSS approximate group
in
which has
as a good model with
and such that
is covered by finitely many left-translates of
.
Let us now analyse the NSS case. Let be a connected Lie group, with Lie algebra
, let
be a bounded symmetric convex body in
, let
be a sufficiently small standard real. Let
be an ultra strong NSS approximate group which has a good model
with
If is zero-dimensional, then by connectedness it is trivial, and then (by the properties of a good model)
is a nonstandard finite group; since it is NSS, it is also trivial. Now suppose that
is not zero-dimensional. Then
contains non-identity elements whose image under
is arbitrarily close to the identity of
; in particular,
does not consist solely of the identity element, and thus contains elements of positive escape norm by the NSS assumption. Let
be a non-identity element of
with minimal escape norm
, then
must be the identity (so in particular
is infinitesimal). (Note that any non-trivial NSS finite approximate group will contain non-identity elements of minimal escape norm, and the extension of this claim to the ultra approximate group case follows from Los’s theorem.) From Theorem 2 one has
for all . (Here we are now using the nonstandard asymptotic notation, thus
means that
for some standard
.) In particular, from the minimality of
, we see that there is a standard
such that
commutes with all elements
with
. In particular, if
is a sufficiently small standard real number, we can find an ultra approximate subgroup
of
with
which is centralised by .
Now we show that generates a one-parameter subgroup of the model Lie group
.
Exercise 7 (One-parameter subgroups from orbits) Let the notation be as above. Let
be such that
is infinitesimal but non-zero.
- Show that
whenever
.
- Show that the map
is a one-parameter subgroup in
(i.e. a continuous homomorphism from
to
).
- Show that there exists an element
of
such that
for all
.
Similar statements hold with
,
replaced by
.
We can now quotient out by the centraliser of and reduce the dimension of the Lie model:
Exercise 8 Let
be the centraliser of
in
, and let
be the nonstandard cyclic group generated by
. (Thus, by the preceding discussion,
contains
, and
is a central subgroup of
containing
. It will be important for us that
and
are both nonstandard sets, i.e. ultraproducts of standard sets.)
- (i) Show that
is a compact subset of
for each standard
.
- (ii) Show that
is a central subgroup of
that contains
.
- (iii) Show that
is a central subgroup of
that is a Lie group of dimension at least one, and so the quotient group
is a Lie group of dimension strictly less than the dimension of
.
- (iv) Let
be the quotient map, and let
be the obvious quotient of
. Let
be a convex symmetric body in the Lie algebra
of
. Show that for sufficiently small standard
, there exists an ultra strong approximate group
with
as a good model, with
, and with
covered by finitely many left-translates of
.
Note that the quotient approximate group obtained by the above procedure is not necessarily NSS. However, it can be made NSS by Exercise 6. As such, one can iterate the above exercise until the dimension of the Lie model shrinks all the way to zero, at which point the NSS approximate group one is working with becomes trivial. This leads to a “cheap” structure theorem for approximate groups:
Exercise 9 (Cheap structure theorem) Let
be an ultra approximate group in a nonstandard group
.
- (i) Show that if
has a good model by a connected Lie group
, then
is nilpotent. (Hint: first use Exercise 1, and then induct on the dimension of
.)
- (ii) Show that
is covered by finitely many left translates of a nonstandard subgroup
of
which admits a normal series
for some standard
, where for every
,
is a normal nonstandard subgroup of
, and
is either a nonstandard finite group or a nonstandard central subgroup of
. Furthermore, if
is not central, then it is contained in the image of
in
. (Hint: first use the Lie model theorem and Exercise 1, and then induct on the dimension of
.)
Exercise 10 (Cheap structure theorem, finite version) Let
be a finite
-approximate group in a group
. Show that
is covered by
left-translates of a subgroup
of
which admits a normal series
for some
, where for every
,
is a normal subgroup of
, and
is either finite or central in
. Furthermore, if
is not central, then it is contained in the image of
in
.
One can push the cheap structure theorem a bit further by controlling the dimension of the nilpotent Lie group in terms of the covering number of the ultra approximate group, as laid out in the following exercise.
Exercise 11 (Nilpotent groups) A Lie algebra
is said to be nilpotent if the derived series
,
,
becomes trivial after a finite number of steps.
- (i) Show that a connected Lie group is nilpotent if and only if its Lie algebra is nilpotent.
- (ii) If
is a finite-dimensional nilpotent Lie algebra, show that there is a simply connected Lie group
with Lie algebra
, for which the exponential map
is a (global) homeomorphism. Furthermore, any other connected Lie group with Lie algebra
is a quotient of
by a discrete central subgroup of
.
- (iii) If
and
are as in (ii), show that the pushforward of a Haar measure (or Lebesgue measure) on
is a bi-invariant Haar measure on
. (Recall from Exercise 6 of Notes 3 that connected nilpotent Lie groups are unimodular.)
- (iv) If
and
are as in (ii), and
is a bi-invariant Haar measure on
, show that
for all open precompact
, where
is the dimension of
.
- (v) If
is a connected (but not necessarily simply connected) nilpotent Lie group, and
is the maximal compact normal subgroup of
(which exists by Exercise 32 of Notes 7), show that
is central, and
is simply connected. As a consequence, conclude that if
is a left-Haar measure of
, then
for all open precompact
, where
is the dimension of
.
- (vi) Show that if
is an ultra
-approximate group which has a Lie model
, and
is the maximal compact normal subgroup of
, then
has dimension at most
.
- (vii) Show that if
is an ultra
-approximate group, then there is an ultra
-approximate group
in
that is modeled by a Lie group
, such that
is covered by finitely many left-translates of
. (Hint:
has a good model
by a locally compact group
; by the Gleason-Yamabe theorem,
has an open subgroup
and a normal subgroup
of
inside
with
a Lie group. Set
.)
- (viii) Show that if
is an ultra
-approximate group, then there is an ultra
-approximate group
in
that is modeled by a nilpotent group of dimension
, such that
can be covered by finitely many left-translates of
.
— 3. Local groups —
The main weakness of the cheap structure theorem in the preceding section is the continual reintroduction of torsion whenever one quotients out by the centraliser , which can destroy the NSS property. We now address the issue of how to fix this, by moving to the context of local groups rather than global groups. We will omit some details, referring to this recent paper for details.
We need to extend many of the notions we have been considering to the local group setting. We begin by generalising the concept of an approximate group.
Definition 3 (Approximate groups) A (local)
-approximate group is a subset
of a local group
which is symmetric and contains the identity, such that
is well-defined in
, and for which
is covered by
left translates of
(by elements in
). An ultra approximate group is an ultraproduct
of
-approximate groups.
Note that we make no topological requirements on or
in this definition; in particular, we may as well give the local group
the discrete topology. There are some minor technical advantages in requiring the local group to be symmetric (so that the inversion map is globally defined) and cancellative (so that
or
implies
), although these assumptions are essentially automatic in practice.
The exponent here is not terribly important in practice, thanks to the following variant of the Sanders lemma:
Exercise 12 Let
be a finite
-approximate group in a local group
, except with only
known to be well-defined rather than
. Let
. Show that there exists a finite
-approximate subgroup
in
with
well-defined and contained in
, and with
covered by
left-translates of
(by elements in
). (Hint: adapt the proof of Lemma 1 from Notes 7.)
Just as global approximate groups can be modeled by global locally compact groups (and in particular, global Lie groups), local approximate groups can be modeled by local locally compact groups:
Definition 4 (Good models) Let
be a (local) ultra approximate group. A (local) good model for
is a homomorphism
from
to a locally compact Hausdorff local group
that obeys the following axioms:
- (Thick image) There exists a neighbourhood
of the identity in
such that
and
.
- (Compact image)
is precompact.
- (Approximation by nonstandard sets) Suppose that
, where
is compact and
is open. Then there exists a nonstandard finite set
such that
.
We make the pedantic remark that with our conventions, a global good model of a global approximate group only becomes a local good model of
by
after restricting the domain of
to
. It is also convenient for minor technical reasons to assume that the local group
is symmetric (i.e. the inversion map is globally defined) but this hypothesis is not of major importance.
The Hrushovski Lie Model theorem can be localised:
Theorem 5 (Local Hrushovski Lie model theorem) Let
be a (local) ultra approximate group. Then there is an ultra approximate subgroup
of
(thus
) with
covered by finitely many left-translates of
(by elements in
), which has a good model by a connected local Lie group
.
The proof of this theorem is basically a localisation of the proof of the global Lie model theorem from Notes 7, and is omitted (see for details). One key replacement is that if is a local approximate group rather than a global one, then the global Gleason-Yamabe theorem (Theorem 1 from Notes 4) must be replaced by the local Gleason-Yamabe theorem of Goldbring, discussed in Section 6 of Notes 4.
One can define the notion of a strong -approximate group and ultra strong approximate group in the local setting without much difficulty, since strong approximate groups only need to work inside
, which is well-defined. Using the local Lie model theorem, one can obtain a local version of Exercise 1. The Gleason lemma (Theorem 2) also localises without much difficulty to local strong approximate groups, as does the reduction to the NSS case in Exercise 6.
Now we once again analyse the NSS case. As before, let be a connected (local) Lie group, with Lie algebra
, let
be a bounded symmetric convex body in
, let
be a sufficiently small standard real. Let
be a (local) ultra strong NSS approximate group which has a (local) good model
with
Again, we assume has dimension at least
, since
is trivial otherwise. We let
be a non-identity element of minimal escape norm. As before,
will have an infinitesimal escape norm and lie in the kernel of
. If we set
, then
is an unbounded natural number, and the map
will be a local one-parameter subgroup, i.e. a continuous homomorphism from
to
. This one-parameter subgroup will be non-trivial and centralised by a neighbourhood of the identity in
.
In the global setting, we quotiented (the group generated by a large portion of) by the centraliser
of
. In the local setting, we perform a more “gentle” quotienting, which roughly speaking arises by quotienting
by the geometric progression
, where
is a sufficiently small standard quantity to be chosen later. However,
is only a local group rather than a global one, and so we must now digress to introduce the notion of quotients of local groups. It is convenient to restrict attention to symmetric cancellative local groups:
Definition 6 (Cancellative local groups) A local group
is symmetric if the inversion operation is globally defined. It is said to be cancellative if the following assertions hold:
- (i) Whenever
are such that
and
are well-defined and equal to each other, then
. (Note that this implies in particular that
.)
- (ii) Whenever
are such that
and
are well-defined and equal to each other, then
.
- (iii) Whenever
are such that
and
are well-defined, then
. (In particular, if
is symmetric and
is well-defined in
for some
, then
is also symmetric.)
Exercise 13 Show that every local group contains an open neighbourhood of the identity which is also a symmetric cancellative local group.
Definition 7 (Sub-local groups) Given two symmetric local groups
and
, we say that
is a sub-local group of
if
is the restriction of
to a symmetric neighbourhood of the identity, and there exists an open neighbourhood
of
with the property that whenever
are such that
is defined in
, then
; we refer to
as an associated neighbourhood for
. If
is also a global group, we say that
is a subgroup of
.
If
is a sub-local group of
, we say that
is normal if there exists an associated neighbourhood
for
with the additional property that whenever
are such that
is well-defined and lies in
, then
. We call
a normalising neighbourhood of
.
Example 3 If
are the (additive) local groups
and
, then
is a sub-local group of
(with associated neighbourhood
). Note that this is despite
not being closed with respect to addition in
; thus we see why it is necessary to allow the associated neighbourhood
to be strictly smaller than
. In a similar vein, the open interval
is a sub-local group of
.
The interval
is also a sub-local group of
; here, one can take for instance
as the associated neighbourhood. As all these examples are abelian, they are clearly normal.
Example 4 Let
be a linear transformation on a finite-dimensional vector space
, and let
be the associated semi-direct product. Let
, where
is a subspace of
that is not preserved by
. Then
is not a normal subgroup of
, but it is a normal sub-local group of
, where one can take
as a normalising neighbourhood of
.
Observe that any sub-local group of a cancellative local group is again a cancellative local group.
One also easily verifies that if is a local homomorphism from
to
for some open neighbourhood
of the identity in
, then
is a normal sub-local group of
, and hence of
. Note that the kernel of a local morphism is well-defined up to local identity. If
is Hausdorff, then the kernel
will also be closed.
Conversely, normal sub-local groups give rise to local homomorphisms into quotient spaces.
Exercise 14 (Quotient spaces) Let
be a cancellative local group, and let
be a normal sub-local group with normalising neighbourhood
. Let
be a symmetric open neighbourhood of the identity such that
. Show that there exists a cancellative local group
and a surjective continuous homomorphism
such that, for any
, one has
if and only if
, and for any
, one has
open if and only if
is open.
Example 5 Let
be the additive local group
, and let
be the sub-local group
, with normalising neighbourhood
. If we then set
, then the hypotheses of Exercise 14 are obeyed, and
can be identified with
, with the projection map
.
Example 6 Let
be the torus
, and let
be the sub-local group
, where
is an irrational number, with normalising neighbourhood
. Set
. Then the hypotheses of Exercise 14 are again obeyed, and
can be identified with the interval
, with the projection map
for
. Note, in contrast, that if one quotiented
by the global group
generated by
, the quotient would be a non-Hausdorff space (and would also contain a dense set of torsion points, in contrast to the interval
which is “locally torsion free”). It is because of this pathological behaviour of quotienting by global groups that we need to work with local group quotients instead.
We now return to the analysis of the NSS ultra strong approximate group . We give the ambient local group
the discrete topology.
Exercise 15 If
is a standard real that is sufficiently small depending on
, show that there exists an ultra approximate group
with
such that
is a sub-local group of
with normalising neighbourhood
, that is also centralised by
.
By Exercise 14, we may now form the quotient set
. Show that this is an ultra approximate group that is modeled by
, where
is an open neighbourhood of the identity in
and
is the local one-parameter subgroup of
introduced earlier. In particular,
is modeled by a local Lie group of dimension one less than the dimension of
.
Now we come to a key observation, which is the main reason why we work in the local groups category in the first place:
We will in fact prove a stronger claim:
Lemma 9 (Lifting lemma) If
, then there exists
such that
and
, where
is the projection map.
Since is NSS, all non-identity elements
of
have non-zero escape norm, and so by the lifting lemma, all non-identity elements of
also have non-zero escape norm, giving Lemma 8.
Proof: (Proof of Lemma 9) We choose to be a lift of
(i.e. an element of
in
) that minimises the escape norm
. (Such a minimum exists since
is nonstandard finite, thanks to Los’s theorem.) If
is trivial, then so is
and there is nothing to prove. Therefore we may assume that
is not the identity and hence, since
is NSS, that it has positive escape norm. Suppose, by way of contradiction, that
. Our goal will be to reach a contradiction by finding another lift
of
with strictly smaller escape norm than
. We will do this by setting
for some suitably chosen
.
We may assume that is infinitesimal, since otherwise there is nothing to prove; in particular
lies in the kernel of the local model
. We may thus find a lift
of
in the kernel of
. In particular, we may assume that
has infinitesimal escape norm.
Set , then
is unbounded. By hypothesis,
; thus
whenever
. In particular, for every (standard) integer
,
. This implies that the group generated by
lies in
. In particular,
lies in the kernel of
, and hence
lies in
for all
.
By (an appropriate local version of) Exercise 7, we can find such that
whenever ; since
lies in
for
, we conclude that
must be parallel to the generator
of
. Similarly, we have
whenever (say) for some
that is also parallel to
. In particular,
for some
Since is the minimal escape norm of non-identity elements of
, we have
, and thus
for some
; in particular,
. Comparing this with (11) we see that
and thus
and hence
By the Euclidean algorithm, we can thus find a nonstandard integer number such that the quantity
lies in the interval . In particular
If we set then (as
commutes with
) we see for all
that
for all . In particular,
. Since
is also a lift of
, this contradicts the minimality of
, and the claim follows.
Because the NSS property is preserved, it is possible to improve upon Exercise 9:
Exercise 16 Strengthen Exercise 9 by ensuring the final quotient
is nonstandard finite, and all the other quotients
are central in
.
As a consequence, one obtains a stronger structure theorem than Exercise 9. Call a symmetric subset containing the identity in a local group nilpotent of step at most
if every iterated commutator in
of length
is well-defined and trivial.
Exercise 17 (Helfgott-Lindenstrauss conjecture)
- (i) Let
be a (local) NSS ultra strong approximate group. Show that there is a symmetric subset
of
containing the identity which is nilpotent of some finite step, such that
is covered by a finite number of left translates of
.
- (ii) Let
be a global NSS ultra strong approximate group with ambient group
. Show that there is a nonstandard nilpotent subgroup
of
such that
is covered by a finite number of left translates of
.
- (iii) Let
be an NSS strong
-approximate group in a global group
. Show that there is a nilpotent subgroup
of
of step
such that
can be covered by a finite number of left translates of
.
- (iv) Let
be a
-approximate group in a global group
. Show that there exists a subgroup
of
and a normal subgroup
of
contained in
, such that
is covered by
left-translates of
, and
is nilpotent of step
.
In fact, a stronger statement is true, involving the nilprogressions defined in Notes 6:
- (i) If
is an NSS ultra strong approximate group, then there is an ultra nilprogression
in
such that
contains
, and
can be covered by finitely many left-translates of
.
- (ii) If
is an ultra approximate group, then there is an ultra coset nilprogression
in
such that
contains
, and
can be covered by finitely many left-translates of
.
- (iii) For all
, there exists
such that, given a finite
-approximate group
in a group
, one can find a coset nilprogression
in
of rank at most
and step at most
such that
contains
, and
can be covered by at most
left-translates of
.
This proposition is established in this paper. The key point is to use the lifting lemma to observe that if (with the notation of the preceding discussion) contains a large nilprogression, then
also contains a large nilprogression. One consequence of this proposition is that there is essentially no difference between local and global approximate groups, at the qualitative level at least:
Corollary 11 Let
be a local
-approximate group. Then there exists a
-approximate subgroup
of
, with
covered by
left-translates of
, such that
is isomorphic to a global
-approximate subgroup.
This is because coset nilprogressions (or large fractions thereof) can be embedded into global groups; again, see this paper for details.
For most applications, one does not need the full strength of Proposition 10; Exercise 17 will suffice. We will give some examples of this in the next set of notes.
8 comments
Comments feed for this article
8 November, 2011 at 11:36 am
Marius Buliga
Re: trapping conditions (definition 1) remind me the paper by David and Toro “Reifenberg Flat Metric Spaces, Snowballs and Embeddings”, Math. Ann., 315, 1999, 641-720. Indeed, in order to understand the long proof, I collected all constants appearing there and remarked that they concentrate around the same powers of 10 as in your definition 1. Notice that Reinfenberg flatness qualify as a kind of mesoscopic flatness.
10 November, 2011 at 7:07 am
Machfudz
Hmmm….
I confused..
XD
11 November, 2011 at 5:00 am
Eza Volturi
I.m not very very understand Mr. Tao :)
15 March, 2012 at 12:09 pm
Lou van den Dries
Something strange about exercise 8: part (iv) suggests that the centralizer of u is normal in star G, but I see no reason why that should be so. My guess is that star G should be replaced by the centralizer of u, so as to reduce to the case that u is in the center.
Next, instead of the centralizer of u one should take the internal subgroup
generated by u, which by the previous reduction is normal.
[Corrected, thanks – T.]
19 August, 2021 at 10:00 am
yang
Mr. Tao: I Recently your book Hilbert’s fifth problem and related topics.
This book is
consisting of Wide Knowledge of mathematics. particular, I have a question, and i don’t solve it through effort. The following is my Doubt.
For a lie group G, This question is that can we find ε>0,such that for any two enough small elements g and h in G with g≠h, there is a integer n>0 such that d( g^n, h^n ) >ε. d is the Gleason metric. This question is can regard as a stronger Phenomenon than escape property. particular,If G is Commutative group ,this question actually is can regard as escape property. I guess the question is true for some class group, but i can solve it. Thanks .
19 August, 2021 at 10:04 am
yang
i can’t solve it.
19 August, 2021 at 10:16 am
Terence Tao
This property is not true in general. Consider for instance the case where
is a compact Lie group (e.g., an orthogonal group) and
is
conjugated by an element of distance
to the identity. Then
is similarly a conjugate of
by the same element and so
for all
. Since one can take
arbitrarily small, there is no uniform bound of the type you request.
20 August, 2021 at 5:55 pm
Anonymous
To Terence Tao :Thank you. I just saw your reply. I’m glad you answered this question in your busy schedule