You are currently browsing the tag archive for the ‘Hilbert’s fifth problem’ tag.
Emmanuel Breuillard, Ben Green, and I have just uploaded to the arXiv our survey “Small doubling in groups“, for the proceedings of the upcoming Erdos Centennial. This is a short survey of the known results on classifying finite subsets of an (abelian) additive group
or a (not necessarily abelian) multiplicative group
that have small doubling in the sense that the sum set
or product set
is small. Such sets behave approximately like finite subgroups of
(and there is a closely related notion of an approximate group in which the analogy is even tighter) , and so this subject can be viewed as a sort of approximate version of finite group theory. (Unfortunately, thus far the theory does not have much new to say about the classification of actual finite groups; progress has been largely made instead on classifying the (highly restricted) number of ways in which approximate groups can differ from a genuine group.)
In the classical case when is the integers
, these sets were classified (in a qualitative sense, at least) by a celebrated theorem of Freiman, which roughly speaking says that such sets
are necessarily “commensurate” in some sense with a (generalised) arithmetic progression
of bounded rank. There are a number of essentially equivalent ways to define what “commensurate” means here; for instance, in the original formulation of the theorem, one asks that
be a dense subset of
, but in modern formulations it is often more convenient to require instead that
be of comparable size to
and be covered by a bounded number of translates of
, or that
and
have an intersection that is of comparable size to both
and
(cf. the notion of commensurability in group theory).
Freiman’s original theorem was extended to more general abelian groups in a sequence of papers culminating in the paper of Green and Ruzsa that handled arbitrary abelian groups. As such groups now contain non-trivial finite subgroups, the conclusion of the theorem must be modified by allowing for “coset progressions” , which can be viewed as “extensions” of generalized arithmetic progressions
by genuine finite groups
.
The proof methods in these abelian results were Fourier-analytic in nature (except in the cases of sets of very small doubling, in which more combinatorial approaches can be applied, and there were also some geometric or combinatorial methods that gave some weaker structural results). As such, it was a challenge to extend these results to nonabelian groups, although for various important special types of ambient group (such as an linear group over a finite or infinite field) it turns out that one can use tools exploiting the special structure of those groups (e.g. for linear groups one would use tools from Lie theory and algebraic geometry) to obtain quite satisfactory results; see e.g. this survey of Pyber and Szabo for the linear case. When the ambient group
is completely arbitrary, it turns out the problem is closely related to the classical Hilbert’s fifth problem of determining the minimal requirements of a topological group in order for such groups to have Lie structure; this connection was first observed and exploited by Hrushovski, and then used by Breuillard, Green, and myself to obtain the analogue of Freiman’s theorem for an arbitrary nonabelian group.
This survey is too short to discuss in much detail the proof techniques used in these results (although the abelian case is discussed in this book of mine with Vu, and the nonabelian case discussed in this more recent book of mine), but instead focuses on the statements of the various known results, as well as some remaining open questions in the subject (in particular, there is substantial work left to be done in making the estimates more quantitative, particularly in the nonabelian setting).
In the previous notes, we established the Gleason-Yamabe theorem:
Theorem 1 (Gleason-Yamabe theorem) Let
be a locally compact group. Then, for any open neighbourhood
of the identity, there exists an open subgroup
of
and a compact normal subgroup
of
in
such that
is isomorphic to a Lie group.
Roughly speaking, this theorem asserts the “mesoscopic” structure of a locally compact group (after restricting to an open subgroup to remove the macroscopic structure, and quotienting out by
to remove the microscopic structure) is always of Lie type.
In this post, we combine the Gleason-Yamabe theorem with some additional tools from point-set topology to improve the description of locally compact groups in various situations.
We first record some easy special cases of this. If the locally compact group has the no small subgroups property, then one can take
to be trivial; thus
is Lie, which implies that
is locally Lie and thus Lie as well. Thus the assertion that all locally compact NSS groups are Lie (Theorem 10 from Notes 4) is a special case of the Gleason-Yamabe theorem.
In a similar spirit, if the locally compact group is connected, then the only open subgroup
of
is the full group
; in particular, by arguing as in the treatment of the compact case (Exercise 19 of Notes 3), we conclude that any connected locally compact Hausdorff group is the inverse limit of Lie groups.
Now we return to the general case, in which need not be connected or NSS. One slight defect of Theorem 1 is that the group
can depend on the open neighbourhood
. However, by using a basic result from the theory of totally disconnected groups known as van Dantzig’s theorem, one can make
independent of
:
Theorem 2 (Gleason-Yamabe theorem, stronger version) Let
be a locally compact group. Then there exists an open subgoup
of
such that, for any open neighbourhood
of the identity in
, there exists a compact normal subgroup
of
in
such that
is isomorphic to a Lie group.
We prove this theorem below the fold. As in previous notes, if is Hausdorff, the group
is thus an inverse limit of Lie groups (and if
(and hence
) is first countable, it is the inverse limit of a sequence of Lie groups).
It remains to analyse inverse limits of Lie groups. To do this, it helps to have some control on the dimensions of the Lie groups involved. A basic tool for this purpose is the invariance of domain theorem:
Theorem 3 (Brouwer invariance of domain theorem) Let
be an open subset of
, and let
be a continuous injective map. Then
is also open.
We prove this theorem below the fold. It has an important corollary:
Corollary 4 (Topological invariance of dimension) If
, and
is a non-empty open subset of
, then there is no continuous injective mapping from
to
. In particular,
and
are not homeomorphic.
Exercise 1 (Uniqueness of dimension) Let
be a non-empty topological space. If
is a manifold of dimension
, and also a manifold of dimension
, show that
. Thus, we may define the dimension
of a non-empty manifold in a well-defined manner.
If
are non-empty manifolds, and there is a continuous injection from
to
, show that
.
Remark 1 Note that the analogue of the above exercise for surjections is false: the existence of a continuous surjection from one non-empty manifold
to another
does not imply that
, thanks to the existence of space-filling curves. Thus we see that invariance of domain, while intuitively plausible, is not an entirely trivial observation.
As we shall see, we can use Corollary 4 to bound the dimension of the Lie groups in an inverse limit
by the “dimension” of the inverse limit
. Among other things, this can be used to obtain a positive resolution to Hilbert’s fifth problem:
Theorem 5 (Hilbert’s fifth problem) Every locally Euclidean group is isomorphic to a Lie group.
Again, this will be shown below the fold.
Another application of this machinery is the following variant of Hilbert’s fifth problem, which was used in Gromov’s original proof of Gromov’s theorem on groups of polynomial growth, although we will not actually need it this course:
Proposition 6 Let
be a locally compact
-compact group that acts transitively, faithfully, and continuously on a connected manifold
. Then
is isomorphic to a Lie group.
Recall that a continuous action of a topological group on a topological space
is a continuous map
which obeys the associativity law
for
and
, and the identity law
for all
. The action is transitive if, for every
, there is a
with
, and faithful if, whenever
are distinct, one has
for at least one
.
The -compact hypothesis is a technical one, and can likely be dropped, but we retain it for this discussion (as in most applications we can reduce to this case).
Remark 2 It is conjectured that the transitivity hypothesis in Proposition 6 can be dropped; this is known as the Hilbert-Smith conjecture. It remains open; the key difficulty is to figure out a way to eliminate the possibility that
is a
-adic group
. See this previous blog post for further discussion.
In this set of notes we will be able to finally prove the Gleason-Yamabe theorem from Notes 0, which we restate here:
Theorem 1 (Gleason-Yamabe theorem) Let
be a locally compact group. Then, for any open neighbourhood
of the identity, there exists an open subgroup
of
and a compact normal subgroup
of
in
such that
is isomorphic to a Lie group.
In the next set of notes, we will combine the Gleason-Yamabe theorem with some topological analysis (and in particular, using the invariance of domain theorem) to establish some further control on locally compact groups, and in particular obtaining a solution to Hilbert’s fifth problem.
To prove the Gleason-Yamabe theorem, we will use three major tools developed in previous notes. The first (from Notes 2) is a criterion for Lie structure in terms of a special type of metric, which we will call a Gleason metric:
Definition 2 Let
be a topological group. A Gleason metric on
is a left-invariant metric
which generates the topology on
and obeys the following properties for some constant
, writing
for
:
- (Escape property) If
and
is such that
, then
.
- (Commutator estimate) If
are such that
, then
where
is the commutator of
and
.
Theorem 3 (Building Lie structure from Gleason metrics) Let
be a locally compact group that has a Gleason metric. Then
is isomorphic to a Lie group.
The second tool is the existence of a left-invariant Haar measure on any locally compact group; see Theorem 3 from Notes 3. Finally, we will also need the compact case of the Gleason-Yamabe theorem (Theorem 8 from Notes 3), which was proven via the Peter-Weyl theorem:
Theorem 4 (Gleason-Yamabe theorem for compact groups) Let
be a compact Hausdorff group, and let
be a neighbourhood of the identity. Then there exists a compact normal subgroup
of
contained in
such that
is isomorphic to a linear group (i.e. a closed subgroup of a general linear group
).
To finish the proof of the Gleason-Yamabe theorem, we have to somehow use the available structures on locally compact groups (such as Haar measure) to build good metrics on those groups (or on suitable subgroups or quotient groups). The basic construction is as follows:
Definition 5 (Building metrics out of test functions) Let
be a topological group, and let
be a bounded non-negative function. Then we define the pseudometric
by the formula
and the semi-norm
by the formula
Note that one can also write
where is the “derivative” of
in the direction
.
Exercise 6 Let the notation and assumptions be as in the above definition. For any
, establish the metric-like properties
- (Identity)
, with equality when
.
- (Symmetry)
.
- (Triangle inequality)
.
- (Continuity) If
, then the map
is continuous.
- (Boundedness) One has
. If
is supported in a set
, then equality occurs unless
.
- (Left-invariance)
. In particular,
.
In particular, we have the norm-like properties
- (Identity)
, with equality when
.
- (Symmetry)
.
- (Triangle inequality)
.
- (Continuity) If
, then the map
is continuous.
- (Boundedness) One has
. If
is supported in a set
, then equality occurs unless
.
We remark that the first three properties of in the above exercise ensure that
is indeed a pseudometric.
To get good metrics (such as Gleason metrics) on groups , it thus suffices to obtain test functions
that obey suitably good “regularity” properties. We will achieve this primarily by means of two tricks. The first trick is to obtain high-regularity test functions by convolving together two low-regularity test functions, taking advantage of the existence of a left-invariant Haar measure
on
. The second trick is to obtain low-regularity test functions by means of a metric-like object on
. This latter trick may seem circular, as our whole objective is to get a metric on
in the first place, but the key point is that the metric one starts with does not need to have as many “good properties” as the metric one ends up with, thanks to the regularity-improving properties of convolution. As such, one can use a “bootstrap argument” (or induction argument) to create a good metric out of almost nothing. It is this bootstrap miracle which is at the heart of the proof of the Gleason-Yamabe theorem (and hence to the solution of Hilbert’s fifth problem).
The arguments here are based on the nonstandard analysis arguments used to establish Hilbert’s fifth problem by Hirschfeld and by Goldbring (and also some unpublished lecture notes of Goldbring and van den Dries). However, we will not explicitly use any nonstandard analysis in this post.
Hilbert’s fifth problem concerns the minimal hypotheses one needs to place on a topological group to ensure that it is actually a Lie group. In the previous set of notes, we saw that one could reduce the regularity hypothesis imposed on
to a “
” condition, namely that there was an open neighbourhood of
that was isomorphic (as a local group) to an open subset
of a Euclidean space
with identity element
, and with group operation
obeying the asymptotic
for sufficiently small . We will call such local groups
local groups.
We now reduce the regularity hypothesis further, to one in which there is no explicit Euclidean space that is initially attached to . Of course, Lie groups are still locally Euclidean, so if the hypotheses on
do not involve any explicit Euclidean spaces, then one must somehow build such spaces from other structures. One way to do so is to exploit an ambient space with Euclidean or Lie structure that
is embedded or immersed in. A trivial example of this is provided by the following basic fact from linear algebra:
Lemma 1 If
is a finite-dimensional vector space (i.e. it is isomorphic to
for some
), and
is a linear subspace of
, then
is also a finite-dimensional vector space.
We will establish a non-linear version of this statement, known as Cartan’s theorem. Recall that a subset of a
-dimensional smooth manifold
is a
-dimensional smooth (embedded) submanifold of
for some
if for every point
there is a smooth coordinate chart
of a neighbourhood
of
in
that maps
to
, such that
, where we identify
with a subspace of
. Informally,
locally sits inside
the same way that
sits inside
.
Theorem 2 (Cartan’s theorem) If
is a (topologically) closed subgroup of a Lie group
, then
is a smooth submanifold of
, and is thus also a Lie group.
Note that the hypothesis that is closed is essential; for instance, the rationals
are a subgroup of the (additive) group of reals
, but the former is not a Lie group even though the latter is.
Exercise 1 Let
be a subgroup of a locally compact group
. Show that
is closed in
if and only if it is locally compact.
A variant of the above results is provided by using (faithful) representations instead of embeddings. Again, the linear version is trivial:
Lemma 3 If
is a finite-dimensional vector space, and
is another vector space with an injective linear transformation
from
to
, then
is also a finite-dimensional vector space.
Here is the non-linear version:
Theorem 4 (von Neumann’s theorem) If
is a Lie group, and
is a locally compact group with an injective continuous homomorphism
, then
also has the structure of a Lie group.
Actually, it will suffice for the homomorphism to be locally injective rather than injective; related to this, von Neumann’s theorem localises to the case when
is a local group rather a group. The requirement that
be locally compact is necessary, for much the same reason that the requirement that
be closed was necessary in Cartan’s theorem.
Example 1 Let
be the two-dimensional torus, let
, and let
be the map
, where
is a fixed real number. Then
is a continuous homomorphism which is locally injective, and is even globally injective if
is irrational, and so Theorem 4 is consistent with the fact that
is a Lie group. On the other hand, note that when
is irrational, then
is not closed; and so Theorem 4 does not follow immediately from Theorem 2 in this case. (We will see, though, that Theorem 4 follows from a local version of Theorem 2.)
As a corollary of Theorem 4, we observe that any locally compact Hausdorff group with a faithful linear representation, i.e. a continuous injective homomorphism from
into a linear group such as
or
, is necessarily a Lie group. This suggests a representation-theoretic approach to Hilbert’s fifth problem. While this approach does not seem to readily solve the entire problem, it can be used to establish a number of important special cases with a well-understood representation theory, such as the compact case or the abelian case (for which the requisite representation theory is given by the Peter-Weyl theorem and Pontryagin duality respectively). We will discuss these cases further in later notes.
In all of these cases, one is not really building up Euclidean or Lie structure completely from scratch, because there is already a Euclidean or Lie structure present in another object in the hypotheses. Now we turn to results that can create such structure assuming only what is ostensibly a weaker amount of structure. In the linear case, one example of this is is the following classical result in the theory of topological vector spaces.
Theorem 5 Let
be a locally compact Hausdorff topological vector space. Then
is isomorphic (as a topological vector space) to
for some finite
.
Remark 1 The Banach-Alaoglu theorem asserts that in a normed vector space
, the closed unit ball in the dual space
is always compact in the weak-* topology. Of course, this dual space
may be infinite-dimensional. This however does not contradict the above theorem, because the closed unit ball is not a neighbourhood of the origin in the weak-* topology (it is only a neighbourhood with respect to the strong topology).
The full non-linear analogue of this theorem would be the Gleason-Yamabe theorem, which we are not yet ready to prove in this set of notes. However, by using methods similar to that used to prove Cartan’s theorem and von Neumann’s theorem, one can obtain a partial non-linear analogue which requires an additional hypothesis of a special type of metric, which we will call a Gleason metric:
Definition 6 Let
be a topological group. A Gleason metric on
is a left-invariant metric
which generates the topology on
and obeys the following properties for some constant
, writing
for
:
- (Escape property) If
and
is such that
, then
.
- (Commutator estimate) If
are such that
, then
where
is the commutator of
and
.
Exercise 2 Let
be a topological group that contains a neighbourhood of the identity isomorphic to a
local group. Show that
admits at least one Gleason metric.
Theorem 7 (Building Lie structure from Gleason metrics) Let
be a locally compact group that has a Gleason metric. Then
is isomorphic to a Lie group.
We will rely on Theorem 7 to solve Hilbert’s fifth problem; this theorem reduces the task of establishing Lie structure on a locally compact group to that of building a metric with suitable properties. Thus, much of the remainder of the solution of Hilbert’s fifth problem will now be focused on the problem of how to construct good metrics on a locally compact group.
In all of the above results, a key idea is to use one-parameter subgroups to convert from the nonlinear setting to the linear setting. Recall from the previous notes that in a Lie group , the one-parameter subgroups are in one-to-one correspondence with the elements of the Lie algebra
, which is a vector space. In a general topological group
, the concept of a one-parameter subgroup (i.e. a continuous homomorphism from
to
) still makes sense; the main difficulties are then to show that the space of such subgroups continues to form a vector space, and that the associated exponential map
is still a local homeomorphism near the origin.
Exercise 3 The purpose of this exercise is to illustrate the perspective that a topological group can be viewed as a non-linear analogue of a vector space. Let
be locally compact groups. For technical reasons we assume that
are both
-compact and metrisable.
- (i) (Open mapping theorem) Show that if
is a continuous homomorphism which is surjective, then it is open (i.e. the image of open sets is open). (Hint: mimic the proof of the open mapping theorem for Banach spaces, as discussed for instance in these notes. In particular, take advantage of the Baire category theorem.)
- (ii) (Closed graph theorem) Show that if a homomorphism
is closed (i.e. its graph
is a closed subset of
), then it is continuous. (Hint: mimic the derivation of the closed graph theorem from the open mapping theorem in the Banach space case, as again discussed in these notes.)
- (iii) Let
be a homomorphism, and let
be a continuous injective homomorphism into another Hausdorff topological group
. Show that
is continuous if and only if
is continuous.
- (iv) Relax the condition of metrisability to that of being Hausdorff. (Hint: Now one cannot use the Baire category theorem for metric spaces; but there is an analogue of this theorem for locally compact Hausdorff spaces.)
This fall (starting Monday, September 26), I will be teaching a graduate topics course which I have entitled “Hilbert’s fifth problem and related topics.” The course is going to focus on three related topics:
- Hilbert’s fifth problem on the topological description of Lie groups, as well as the closely related (local) classification of locally compact groups (the Gleason-Yamabe theorem).
- Approximate groups in nonabelian groups, and their classification via the Gleason-Yamabe theorem (this is very recent work of Emmanuel Breuillard, Ben Green, Tom Sanders, and myself, building upon earlier work of Hrushovski);
- Gromov’s theorem on groups of polynomial growth, as proven via the classification of approximate groups (as well as some consequences to fundamental groups of Riemannian manifolds).
I have already blogged about these topics repeatedly in the past (particularly with regard to Hilbert’s fifth problem), and I intend to recycle some of that material in the lecture notes for this course.
The above three families of results exemplify two broad principles (part of what I like to call “the dichotomy between structure and randomness“):
- (Rigidity) If a group-like object exhibits a weak amount of regularity, then it (or a large portion thereof) often automatically exhibits a strong amount of regularity as well;
- (Structure) This strong regularity manifests itself either as Lie type structure (in continuous settings) or nilpotent type structure (in discrete settings). (In some cases, “nilpotent” should be replaced by sister properties such as “abelian“, “solvable“, or “polycyclic“.)
Let me illustrate what I mean by these two principles with two simple examples, one in the continuous setting and one in the discrete setting. We begin with a continuous example. Given an complex matrix
, define the matrix exponential
of
by the formula
which can easily be verified to be an absolutely convergent series.
Exercise 1 Show that the map
is a real analytic (and even complex analytic) map from
to
, and obeys the restricted homomorphism property
for all
and
.
Proposition 1 (Rigidity and structure of matrix homomorphisms) Let
be a natural number. Let
be the group of invertible
complex matrices. Let
be a map obeying two properties:
- (Group-like object)
is a homomorphism, thus
for all
.
- (Weak regularity) The map
is continuous.
Then:
- (Strong regularity) The map
is smooth (i.e. infinitely differentiable). In fact it is even real analytic.
- (Lie-type structure) There exists a (unique) complex
matrix
such that
for all
.
Proof: Let be as above. Let
be a small number (depending only on
). By the homomorphism property,
(where we use
here to denote the identity element of
), and so by continuity we may find a small
such that
for all
(we use some arbitrary norm here on the space of
matrices, and allow implied constants in the
notation to depend on
).
The map is real analytic and (by the inverse function theorem) is a diffeomorphism near
. Thus, by the inverse function theorem, we can (if
is small enough) find a matrix
of size
such that
. By the homomorphism property and (1), we thus have
On the other hand, by another application of the inverse function theorem we see that the squaring map is a diffeomorphism near
in
, and thus (if
is small enough)
We may iterate this argument (for a fixed, but small, value of ) and conclude that
for all . By the homomorphism property and (1) we thus have
whenever is a dyadic rational, i.e. a rational of the form
for some integer
and natural number
. By continuity we thus have
for all real . Setting
we conclude that
for all real , which gives existence of the representation and also real analyticity and smoothness. Finally, uniqueness of the representation
follows from the identity
Exercise 2 Generalise Proposition 1 by replacing the hypothesis that
is continuous with the hypothesis that
is Lebesgue measurable (Hint: use the Steinhaus theorem.). Show that the proposition fails (assuming the axiom of choice) if this hypothesis is omitted entirely.
Note how one needs both the group-like structure and the weak regularity in combination in order to ensure the strong regularity; neither is sufficient on its own. We will see variants of the above basic argument throughout the course. Here, the task of obtaining smooth (or real analytic structure) was relatively easy, because we could borrow the smooth (or real analytic) structure of the domain and range
; but, somewhat remarkably, we shall see that one can still build such smooth or analytic structures even when none of the original objects have any such structure to begin with.
Now we turn to a second illustration of the above principles, namely Jordan’s theorem, which uses a discreteness hypothesis to upgrade Lie type structure to nilpotent (and in this case, abelian) structure. We shall formulate Jordan’s theorem in a slightly stilted fashion in order to emphasise the adherence to the above-mentioned principles.
Theorem 2 (Jordan’s theorem) Let
be an object with the following properties:
- (Group-like object)
is a group.
- (Discreteness)
is finite.
- (Lie-type structure)
is contained in
(the group of unitary
matrices) for some
.
Then there is a subgroup
of
such that
- (
is close to
) The index
of
in
is
(i.e. bounded by
for some quantity
depending only on
).
- (Nilpotent-type structure)
is abelian.
A key observation in the proof of Jordan’s theorem is that if two unitary elements are close to the identity, then their commutator
is even closer to the identity (in, say, the operator norm
). Indeed, since multiplication on the left or right by unitary elements does not affect the operator norm, we have
and so by the triangle inequality
Now we can prove Jordan’s theorem.
Proof: We induct on , the case
being trivial. Suppose first that
contains a central element
which is not a multiple of the identity. Then, by definition,
is contained in the centraliser
of
, which by the spectral theorem is isomorphic to a product
of smaller unitary groups. Projecting
to each of these factor groups and applying the induction hypothesis, we obtain the claim.
Thus we may assume that contains no central elements other than multiples of the identity. Now pick a small
(one could take
in fact) and consider the subgroup
of
generated by those elements of
that are within
of the identity (in the operator norm). By considering a maximal
-net of
we see that
has index at most
in
. By arguing as before, we may assume that
has no central elements other than multiples of the identity.
If consists only of multiples of the identity, then we are done. If not, take an element
of
that is not a multiple of the identity, and which is as close as possible to the identity (here is where we crucially use that
is finite). By (2), we see that if
is sufficiently small depending on
, and if
is one of the generators of
, then
lies in
and is closer to the identity than
, and is thus a multiple of the identity. On the other hand,
has determinant
. Given that it is so close to the identity, it must therefore be the identity (if
is small enough). In other words,
is central in
, and is thus a multiple of the identity. But this contradicts the hypothesis that there are no central elements other than multiples of the identity, and we are done.
Commutator estimates such as (2) will play a fundamental role in many of the arguments we will see in this course; as we saw above, such estimates combine very well with a discreteness hypothesis, but will also be very useful in the continuous setting.
Exercise 3 Generalise Jordan’s theorem to the case when
is a finite subgroup of
rather than of
. (Hint: The elements of
are not necessarily unitary, and thus do not necessarily preserve the standard Hilbert inner product of
. However, if one averages that inner product by the finite group
, one obtains a new inner product on
that is preserved by
, which allows one to conjugate
to a subgroup of
. This averaging trick is (a small) part of Weyl’s unitary trick in representation theory.)
Exercise 4 (Inability to discretise nonabelian Lie groups) Show that if
, then the orthogonal group
cannot contain arbitrarily dense finite subgroups, in the sense that there exists an
depending only on
such that for every finite subgroup
of
, there exists a ball of radius
in
(with, say, the operator norm metric) that is disjoint from
. What happens in the
case?
Remark 1 More precise classifications of the finite subgroups of
are known, particularly in low dimensions. For instance, one can show that the only finite subgroups of
(which
is a double cover of) are isomorphic to either a cyclic group, a dihedral group, or the symmetry group of one of the Platonic solids.
The classical formulation of Hilbert’s fifth problem asks whether topological groups that have the topological structure of a manifold, are necessarily Lie groups. This is indeed, the case, thanks to following theorem of Gleason and Montgomery-Zippin:
Theorem 1 (Hilbert’s fifth problem) Let
be a topological group which is locally Euclidean. Then
is isomorphic to a Lie group.
We have discussed the proof of this result, and of related results, in previous posts. There is however a generalisation of Hilbert’s fifth problem which remains open, namely the Hilbert-Smith conjecture, in which it is a space acted on by the group which has the manifold structure, rather than the group itself:
Conjecture 2 (Hilbert-Smith conjecture) Let
be a locally compact topological group which acts continuously and faithfully (or effectively) on a connected finite-dimensional manifold
. Then
is isomorphic to a Lie group.
Note that Conjecture 2 easily implies Theorem 1 as one can pass to the connected component of a locally Euclidean group (which is clearly locally compact), and then look at the action of
on itself by left-multiplication.
The hypothesis that the action is faithful (i.e. each non-identity group element acts non-trivially on
) cannot be completely eliminated, as any group
will have a trivial action on any space
. The requirement that
be locally compact is similarly necessary: consider for instance the diffeomorphism group
of, say, the unit circle
, which acts on
but is infinite dimensional and is not locally compact (with, say, the uniform topology). Finally, the connectedness of
is also important: the infinite torus
(with the product topology) acts faithfully on the disconnected manifold
by the action
The conjecture in full generality remains open. However, there are a number of partial results. For instance, it was observed by Montgomery and Zippin that the conjecture is true for transitive actions, by a modification of the argument used to establish Theorem 1. This special case of the Hilbert-Smith conjecture (or more precisely, a generalisation thereof in which “finite-dimensional manifold” was replaced by “locally connected locally compact finite-dimensional”) was used in Gromov’s proof of his famous theorem on groups of polynomial growth. I record the argument of Montgomery and Zippin below the fold.
Another partial result is the reduction of the Hilbert-Smith conjecture to the -adic case. Indeed, it is known that Conjecture 2 is equivalent to
Conjecture 3 (Hilbert-Smith conjecture for
-adic actions) It is not possible for a
-adic group
to act continuously and effectively on a connected finite-dimensional manifold
.
The reduction to the -adic case follows from the structural theory of locally compact groups (specifically, the Gleason-Yamabe theorem discussed in previous posts) and some results of Newman that sharply restrict the ability of periodic actions on a manifold
to be close to the identity. I record this argument (which appears for instance in this paper of Lee) below the fold also.
This is another installment of my my series of posts on Hilbert’s fifth problem. One formulation of this problem is answered by the following theorem of Gleason and Montgomery-Zippin:
Theorem 1 (Hilbert’s fifth problem) Let
be a topological group which is locally Euclidean. Then
is isomorphic to a Lie group.
Theorem 1 is deep and difficult result, but the discussion in the previous posts has reduced the proof of this Theorem to that of establishing two simpler results, involving the concepts of a no small subgroups (NSS) subgroup, and that of a Gleason metric. We briefly recall the relevant definitions:
Definition 2 (NSS) A topological group
is said to have no small subgroups, or is NSS for short, if there is an open neighbourhood
of the identity in
that contains no subgroups of
other than the trivial subgroup
.
Definition 3 (Gleason metric) Let
be a topological group. A Gleason metric on
is a left-invariant metric
which generates the topology on
and obeys the following properties for some constant
, writing
for
:
- (Escape property) If
and
is such that
, then
- (Commutator estimate) If
are such that
, then
where
is the commutator of
and
.
The remaining steps in the resolution of Hilbert’s fifth problem are then as follows:
Theorem 4 (Reduction to the NSS case) Let
be a locally compact group, and let
be an open neighbourhood of the identity in
. Then there exists an open subgroup
of
, and a compact subgroup
of
contained in
, such that
is NSS and locally compact.
Theorem 5 (Gleason’s lemma) Let
be a locally compact NSS group. Then
has a Gleason metric.
The purpose of this post is to establish these two results, using arguments that are originally due to Gleason. We will split this task into several subtasks, each of which improves the structure on the group by some amount:
Proposition 6 (From locally compact to metrisable) Let
be a locally compact group, and let
be an open neighbourhood of the identity in
. Then there exists an open subgroup
of
, and a compact subgroup
of
contained in
, such that
is locally compact and metrisable.
For any open neighbourhood of the identity in
, let
be the union of all the subgroups of
that are contained in
. (Thus, for instance,
is NSS if and only if
is trivial for all sufficiently small
.)
Proposition 7 (From metrisable to subgroup trapping) Let
be a locally compact metrisable group. Then
has the subgroup trapping property: for every open neighbourhood
of the identity, there exists another open neighbourhood
of the identity such that
generates a subgroup
contained in
.
Proposition 8 (From subgroup trapping to NSS) Let
be a locally compact group with the subgroup trapping property, and let
be an open neighbourhood of the identity in
. Then there exists an open subgroup
of
, and a compact subgroup
of
contained in
, such that
is locally compact and NSS.
Proposition 9 (From NSS to the escape property) Let
be a locally compact NSS group. Then there exists a left-invariant metric
on
generating the topology on
which obeys the escape property (1) for some constant
.
Proposition 10 (From escape to the commutator estimate) Let
be a locally compact group with a left-invariant metric
that obeys the escape property (1). Then
also obeys the commutator property (2).
It is clear that Propositions 6, 7, and 8 combine to give Theorem 4, and Propositions 9, 10 combine to give Theorem 5.
Propositions 6–10 are all proven separately, but their proofs share some common strategies and ideas. The first main idea is to construct metrics on a locally compact group by starting with a suitable “bump function”
(i.e. a continuous, compactly supported function from
to
) and pulling back the metric structure on
by using the translation action
, thus creating a (semi-)metric
One easily verifies that this is indeed a (semi-)metric (in that it is non-negative, symmetric, and obeys the triangle inequality); it is also left-invariant, and so we have , where
where is the difference operator
,
This construction was already seen in the proof of the Birkhoff-Kakutani theorem, which is the main tool used to establish Proposition 6. For the other propositions, the idea is to choose a bump function that is “smooth” enough that it creates a metric with good properties such as the commutator estimate (2). Roughly speaking, to get a bound of the form (2), one needs
to have “
regularity” with respect to the “right” smooth structure on
By
regularity, we mean here something like a bound of the form
for all . Here we use the usual asymptotic notation, writing
or
if
for some constant
(which can vary from line to line).
The following lemma illustrates how regularity can be used to build Gleason metrics.
Lemma 11 Suppose that
obeys (4). Then the (semi-)metric
(and associated (semi-)norm
) obey the escape property (1) and the commutator property (2).
Proof: We begin with the commutator property (2). Observe the identity
whence
From the triangle inequality (and translation-invariance of the norm) we thus see that (2) follows from (4). Similarly, to obtain the escape property (1), observe the telescoping identity
for any and natural number
, and thus by the triangle inequality
But from (4) (and the triangle inequality) we have
and thus we have the “Taylor expansion”
which gives (1).
It remains to obtain that have the desired
regularity property. In order to get such regular bump functions, we will use the trick of convolving together two lower regularity bump functions (such as two functions with “
regularity” in some sense to be determined later). In order to perform this convolution, we will use the fundamental tool of (left-invariant) Haar measure
on the locally compact group
. Here we exploit the basic fact that the convolution
of two functions tends to be smoother than either of the two factors
. This is easiest to see in the abelian case, since in this case we can distribute derivatives according to the law
which suggests that the order of “differentiability” of should be the sum of the orders of
and
separately.
These ideas are already sufficient to establish Proposition 10 directly, and also Proposition 9 when comined with an additional bootstrap argument. The proofs of Proposition 7 and Proposition 8 use similar techniques, but is more difficult due to the potential presence of small subgroups, which require an application of the Peter-Weyl theorem to properly control. Both of these theorems will be proven below the fold, thus (when combined with the preceding posts) completing the proof of Theorem 1.
The presentation here is based on some unpublished notes of van den Dries and Goldbring on Hilbert’s fifth problem. I am indebted to Emmanuel Breuillard, Ben Green, and Tom Sanders for many discussions related to these arguments.
Let be a Lie group with Lie algebra
. As is well known, the exponential map
is a local homeomorphism near the identity. As such, the group law on
can be locally pulled back to an operation
defined on a neighbourhood
of the identity in
, defined as
where is the local inverse of the exponential map. One can view
as the group law expressed in local exponential coordinates around the origin.
An asymptotic expansion for is provided by the Baker-Campbell-Hausdorff (BCH) formula
for all sufficiently small , where
is the Lie bracket. More explicitly, one has the Baker-Campbell-Hausdorff-Dynkin formula
for all sufficiently small , where
,
is the adjoint representation
, and
is the function
which is real analytic near and can thus be applied to linear operators sufficiently close to the identity. One corollary of this is that the multiplication operation
is real analytic in local coordinates, and so every smooth Lie group is in fact a real analytic Lie group.
It turns out that one does not need the full force of the smoothness hypothesis to obtain these conclusions. It is, for instance, a classical result that regularity of the group operations is already enough to obtain the Baker-Campbell-Hausdorff formula. Actually, it turns out that we can weaken this a bit, and show that even
regularity (i.e. that the group operations are continuously differentiable, and the derivatives are locally Lipschitz) is enough to make the classical derivation of the Baker-Campbell-Hausdorff formula work. More precisely, we have
Theorem 1 (
Baker-Campbell-Hausdorff formula) Let
be a finite-dimensional vector space, and suppose one has a continuous operation
defined on a neighbourhood
around the origin, which obeys the following three axioms:
- (Approximate additivity) For
sufficiently close to the origin, one has
(In particular,
for
sufficiently close to the origin.)
- (Associativity) For
sufficiently close to the origin,
.
- (Radial homogeneity) For
sufficiently close to the origin, one has
for all
. (In particular,
for all
sufficiently close to the origin.)
Then
is real analytic (and in particular, smooth) near the origin. (In particular,
gives a neighbourhood of the origin the structure of a local Lie group.)
Indeed, we will recover the Baker-Campbell-Hausdorff-Dynkin formula (after defining appropriately) in this setting; see below the fold.
The reason that we call this a Baker-Campbell-Hausdorff formula is that if the group operation
has
regularity, and has
as an identity element, then Taylor expansion already gives (2), and in exponential coordinates (which, as it turns out, can be defined without much difficulty in the
category) one automatically has (3).
We will record the proof of Theorem 1 below the fold; it largely follows the classical derivation of the BCH formula, but due to the low regularity one will rely on tools such as telescoping series and Riemann sums rather than on the fundamental theorem of calculus. As an application of this theorem, we can give an alternate derivation of one of the components of the solution to Hilbert’s fifth problem, namely the construction of a Lie group structure from a Gleason metric, which was covered in the previous post; we discuss this at the end of this article. With this approach, one can avoid any appeal to von Neumann’s theorem and Cartan’s theorem (discussed in this post), or the Kuranishi-Gleason extension theorem (discussed in this post).
Hilbert’s fifth problem asks to clarify the extent that the assumption on a differentiable or smooth structure is actually needed in the theory of Lie groups and their actions. While this question is not precisely formulated and is thus open to some interpretation, the following result of Gleason and Montgomery-Zippin answers at least one aspect of this question:
Theorem 1 (Hilbert’s fifth problem) Let
be a topological group which is locally Euclidean (i.e. it is a topological manifold). Then
is isomorphic to a Lie group.
Theorem 1 can be viewed as an application of the more general structural theory of locally compact groups. In particular, Theorem 1 can be deduced from the following structural theorem of Gleason and Yamabe:
Theorem 2 (Gleason-Yamabe theorem) Let
be a locally compact group, and let
be an open neighbourhood of the identity in
. Then there exists an open subgroup
of
, and a compact subgroup
of
contained in
, such that
is isomorphic to a Lie group.
The deduction of Theorem 1 from Theorem 2 proceeds using the Brouwer invariance of domain theorem and is discussed in this previous post. In this post, I would like to discuss the proof of Theorem 2. We can split this proof into three parts, by introducing two additional concepts. The first is the property of having no small subgroups:
Definition 3 (NSS) A topological group
is said to have no small subgroups, or is NSS for short, if there is an open neighbourhood
of the identity in
that contains no subgroups of
other than the trivial subgroup
.
An equivalent definition of an NSS group is one which has an open neighbourhood of the identity that every non-identity element
escapes in finite time, in the sense that
for some positive integer
. It is easy to see that all Lie groups are NSS; we shall shortly see that the converse statement (in the locally compact case) is also true, though significantly harder to prove.
Another useful property is that of having what I will call a Gleason metric:
Definition 4 Let
be a topological group. A Gleason metric on
is a left-invariant metric
which generates the topology on
and obeys the following properties for some constant
, writing
for
:
- (Escape property) If
and
is such that
, then
.
- (Commutator estimate) If
are such that
, then
where
is the commutator of
and
.
For instance, the unitary group with the operator norm metric
can easily verified to be a Gleason metric, with the commutator estimate (1) coming from the inequality
Similarly, any left-invariant Riemannian metric on a (connected) Lie group can be verified to be a Gleason metric. From the escape property one easily sees that all groups with Gleason metrics are NSS; again, we shall see that there is a partial converse.
Remark 1 The escape and commutator properties are meant to capture “Euclidean-like” structure of the group. Other metrics, such as Carnot-Carathéodory metrics on Carnot Lie groups such as the Heisenberg group, usually fail one or both of these properties.
The proof of Theorem 2 can then be split into three subtheorems:
Theorem 5 (Reduction to the NSS case) Let
be a locally compact group, and let
be an open neighbourhood of the identity in
. Then there exists an open subgroup
of
, and a compact subgroup
of
contained in
, such that
is NSS, locally compact, and metrisable.
Theorem 6 (Gleason’s lemma) Let
be a locally compact metrisable NSS group. Then
has a Gleason metric.
Theorem 7 (Building a Lie structure) Let
be a locally compact group with a Gleason metric. Then
is isomorphic to a Lie group.
Clearly, by combining Theorem 5, Theorem 6, and Theorem 7 one obtains Theorem 2 (and hence Theorem 1).
Theorem 5 and Theorem 6 proceed by some elementary combinatorial analysis, together with the use of Haar measure (to build convolutions, and thence to build “smooth” bump functions with which to create a metric, in a variant of the analysis used to prove the Birkhoff-Kakutani theorem); Theorem 5 also requires Peter-Weyl theorem (to dispose of certain compact subgroups that arise en route to the reduction to the NSS case), which was discussed previously on this blog.
In this post I would like to detail the final component to the proof of Theorem 2, namely Theorem 7. (I plan to discuss the other two steps, Theorem 5 and Theorem 6, in a separate post.) The strategy is similar to that used to prove von Neumann’s theorem, as discussed in this previous post (and von Neumann’s theorem is also used in the proof), but with the Gleason metric serving as a substitute for the faithful linear representation. Namely, one first gives the space of one-parameter subgroups of
enough of a structure that it can serve as a proxy for the “Lie algebra” of
; specifically, it needs to be a vector space, and the “exponential map” needs to cover an open neighbourhood of the identity. This is enough to set up an “adjoint” representation of
, whose image is a Lie group by von Neumann’s theorem; the kernel is essentially the centre of
, which is abelian and can also be shown to be a Lie group by a similar analysis. To finish the job one needs to use arguments of Kuranishi and of Gleason, as discussed in this previous post.
The arguments here can be phrased either in the standard analysis setting (using sequences, and passing to subsequences often) or in the nonstandard analysis setting (selecting an ultrafilter, and then working with infinitesimals). In my view, the two approaches have roughly the same level of complexity in this case, and I have elected for the standard analysis approach.
Remark 2 From Theorem 7 we see that a Gleason metric structure is a good enough substitute for smooth structure that it can actually be used to reconstruct the entire smooth structure; roughly speaking, the commutator estimate (1) allows for enough “Taylor expansion” of expressions such as
that one can simulate the fundamentals of Lie theory (in particular, construction of the Lie algebra and the exponential map, and its basic properties. The advantage of working with a Gleason metric rather than a smoother structure, though, is that it is relatively undemanding with regards to regularity; in particular, the commutator estimate (1) is roughly comparable to the imposition
structure on the group
, as this is the minimal regularity to get the type of Taylor approximation (with quadratic errors) that would be needed to obtain a bound of the form (1). We will return to this point in a later post.
We recall Brouwer’s famous fixed point theorem:
Theorem 1 (Brouwer fixed point theorem) Let
be a continuous function on the unit ball
in a Euclidean space
. Then
has at least one fixed point, thus there exists
with
.
This theorem has many proofs, most of which revolve (either explicitly or implicitly) around the notion of the degree of a continuous map of the unit sphere
to itself, and more precisely around the stability of degree with respect to homotopy. (Indeed, one can view the Brouwer fixed point theorem as an assertion that some non-trivial degree-like invariant must exist, or more abstractly that the homotopy group
is non-trivial.)
One of the many applications of this result is to prove Brouwer’s invariance of domain theorem:
Theorem 2 (Brouwer invariance of domain theorem) Let
be an open subset of
, and let
be a continuous injective map. Then
is also open.
This theorem in turn has an important corollary:
Corollary 3 (Topological invariance of dimension) If
, and
is a non-empty open subset of
, then there is no continuous injective mapping from
to
. In particular,
and
are not homeomorphic.
This corollary is intuitively obvious, but note that topological intuition is not always rigorous. For instance, it is intuitively plausible that there should be no continuous surjection from to
for
, but such surjections always exist, thanks to variants of the Peano curve construction.
Theorem 2 or Corollary 3 can be proven by simple ad hoc means for small values of or
(for instance, by noting that removing a point from
will disconnect
when
, but not for
), but I do not know of any proof of these results in general dimension that does not require algebraic topology machinery that is at least as sophisticated as the Brouwer fixed point theorem. (Lebesgue, for instance, famously failed to establish the above corollary rigorously, although he did end up discovering the important concept of Lebesgue covering dimension as a result of his efforts.)
Nowadays, the invariance of domain theorem is usually proven using the machinery of singular homology. In this post I would like to record a short proof of Theorem 2 using Theorem 1 that I discovered in a paper of Kulpa, which avoids any use of algebraic topology tools beyond the fixed point theorem, though it is more ad hoc in its approach than the systematic singular homology approach.
Remark 1 A heuristic explanation as to why the Brouwer fixed point theorem is more or less a necessary ingredient in the proof of the invariance of domain theorem is that a counterexample to the former result could conceivably be used to create a counterexample to the latter one. Indeed, if the Brouwer fixed point theorem failed, then (as is well known) one would be able to find a continuous function
that was the identity on
(indeed, one could take
to be the first point in which the ray from
through
hits
). If one then considered the function
defined by
, then this would be a continuous function which avoids the interior of
, but which maps the origin
to a point on the sphere
(and maps
to the dilate
). This could conceivably be a counterexample to Theorem 2, except that
is not necessarily injective. I do not know if there is a more rigorous way to formulate this connection.
The reason I was looking for a proof of the invariance of domain theorem was that it comes up in the very last stage of the solution to Hilbert’s fifth problem, namely to establish the following fact:
Theorem 4 (Hilbert’s fifth problem) Every locally Euclidean group is isomorphic to a Lie group.
Recall that a locally Euclidean group is a topological group which is locally homeomorphic to an open subset of a Euclidean space , i.e. it is a continuous manifold. Note in contrast that a Lie group is a topological group which is locally diffeomorphic to an open subset of
, it is a smooth manifold. Thus, Hilbert’s fifth problem is a manifestation of the “rigidity” of algebraic structure (in this case, group structure), which turns weak regularity (continuity) into strong regularity (smoothness).
It is plausible that something like Corollary 3 would need to be invoked in order to solve Hilbert’s fifth problem. After all, if Euclidean spaces ,
of different dimension were homeomorphic to each other, then the property of being locally Euclidean loses a lot of meaning, and would thus not be a particularly powerful hypothesis. Note also that it is clear that two Lie groups can only be isomorphic if they have the same dimension, so in view of Theorem 4, it becomes plausible that two Euclidean spaces can only be homeomorphic if they have the same dimension, although I do not know of a way to rigorously deduce this claim from Theorem 4.
Interestingly, Corollary 3 is the only place where algebraic topology enters into the solution of Hilbert’s fifth problem (although its cousin, point-set topology, is used all over the place). There are results closely related to Theorem 4, such as the Gleason-Yamabe theorem mentioned in a recent post, which do not use the notion of being locally Euclidean, and do not require algebraic topological methods in their proof. Indeed, one can deduce Theorem 4 from the Gleason-Yamabe theorem and invariance of domain; we sketch a proof of this (following Montgomery and Zippin) below the fold.
Recent Comments