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.
— 1. Van Dantzig’s theorem —
Recall that a (non-empty) topological space is connected if the only clopen (i.e. closed and open) subsets of
are the whole space
and the empty set
; a non-empty topological space is disconnected if it is not connected. (By convention, the empty set is considered to be neither connected nor disconnected, somewhat analogously to how the natural number
is neither considered prime nor composite.)
At the opposite extreme to connectedness is the property of being a totally disconnected space. This is a space whose only connected subsets are the singleton sets. Typical examples of totally disconnected spaces include discrete spaces (e.g. the integers with the discrete topology) and Cantor spaces (such as the standard Cantor set).
Most topological spaces are neither connected nor totally disconnected, but some intermediate combination of both. In the case of topological groups , this rather vague assertion can be formalised as follows.
Exercise 3
- Define a connected component of a topological space
to be a maxmial connected set. Show that the connected components of
form a partition of
, thus every point in
belongs to exactly one connected component.
- Let
be a topological group, and let
be the connected component of the identity. Show that
is a closed normal subgroup of
, and that
is a totally disconnected subgroup of
. Thus, one has a short exact sequence
of topological groups that describes
as an extension of a totally disconnected group by a connected group.
- Conversely, if one has a short exact sequence
of topological groups, with
connected and
totally disconnected, show that
is isomorphic to
, and
is isomorphic to
.
- If
is locally compact, show that
and
are also locally compact.
In principle at least, the study of locally compact groups thus splits into the study of connected locally compact groups, and the study of totally disconnected locally compact groups. (Note however that even if one has a complete understanding of the factors of a short exact sequence
, it may still be a non-trivial issue to fully understand the combined group
, due to the possible presence of non-trivial group cohomology. See for instance this previous blog post for more discussion.)
For totally disconnected locally compact groups, one has the following fundamental theorem of van Dantzig’s:
Theorem 7 (Van Danztig’s theorem) Every totally disconnected locally compact group
contains a compact open subgroup
(which will of course still be totally disconnected).
Example 1 Let
be a prime. Then the
-adic field
(with the usual
-adic valuation) is totally disconnected locally compact, and the
-adic integers
are a compact open subgroup.
Of course, this situation is the polar opposite of what occurs in the connected case, in which the only open subgroup is the whole group.
To prove van Dantzig’s theorem, we first need a lemma from point set topology, which shows that totally disconnected spaces contain enough clopen sets to separate points:
Lemma 8 Let
be a totally disconnected compact Hausdorff space, and let
be distinct points in
. Then there exists a clopen set that contains
but not
.
Proof: Let be the intersection of all the clopen sets that contain
(note that
is obviously clopen). Clearly
is closed and contains
. Our objective is to show that
consists solely of
. As
is totally disconnected, it will suffice to show that
is connected.
Suppose this is not the case, then we can split where
are disjoint non-empty closed sets; without loss of generality, we may assume that
lies in
. As all compact Hausdorff spaces are normal, we can thus enclose
in disjoint open subsets
of
. In particular, the topological boundary
is compact and lies outside of
. By definition of
, we thus see that for every
, we can find a clopen neighbourhood of
that avoids
; by compactness of
(and the fact that finite intersections of clopen sets are clopen), we can thus find a clopen neighbourhood
of
that is disjoint from
. One then verifies that
is a clopen neighbourhood of
that is disjoint from
, contradicting the definition of
, and the claim follows.
Now we can prove van Dantzig’s theorem. We will use an argument from the book of Hewitt and Ross. Let be totally disconnected locally compact (and thus Hausdorff). Then we can find a compact neighbourhood
of the identity. By Lemma 8, for every
, we can find a clopen neighbourhood of the identity that avoids
; by compactness of
, we may thus find a clopen neighbourhood of the identity that avoids
. By intersecting this neighbourhood with
, we may thus find a compact clopen neighbourhood
of the identity. As
is both compact and open, we may then the continuity of the group operations find a symmetric neighbourhood
of the identity such that
. In particular, if we let
be the group generated by
, then
is an open subgroup of
contained in
and is thus compact as required.
Remark 3 The same argument shows that a totally disconnected locally compact group contains arbitrarily small compact open subgroups, or in other words the compact open subgroups form a neighbourhood base for the identity.
In view of van Dantzig’s theorem, we see that the “local” behaviour of totally disconnected locally compact groups can be modeled by the compact totally disconnected groups, which are better understood. Thanks to the Gleason-Yamabe theorem for compact groups, such groups are the inverse limits of compact totally disconnected Lie groups. But it is easy to see that a compact totally disconnected Lie group must be finite, and so compact totally disconnected groups are necessarily profinite. The global behaviour however remains more complicated, in part because the compact open subgroup given by van Dantzig’s theorem need not be normal, and so does not necessarily induce a splitting of into compact and discrete factors.
Example 2 Let
be a prime, and let
be the semi-direct product
, where the integers
act on
by the map
, and we give
the product of the discrete topology of
and the
-adic topology on
. One easily verifies that
is a totally disconnected locally compact group. It certainly has compact open subgroups, such as
. However, it is easy to show that
has no non-trivial compact normal subgroups (the problem is that the conjugation action of
on
has all non-trivial orbits unbounded).
We can pull van Dantzig’s theorem back to more general locally compact groups:
Exercise 4 Let
be a locally compact group.
- Show that
contains an open subgroup
which is “compact-by-connected” in the sense that
is compact. (Hint: apply van Dantzig’s theorem to
.)
- If
is compact-by-connected, and
is an open neighbourhood of the identity, show that there exists a compact subgroup
of
in
such that
is isomorphic to a Lie group. (Hint: use Theorem 1, and observe that any open subgroup of the compact-by-connected group
has finite index and thus has only finitely many conjugates.) Conclude Theorem 2.
- Show that any locally compact Hausdorff group
contains an open subgroup
that is isomorphic to an inverse limit of Lie groups
, which each Lie group
has at most finitely many connected components. Furthermore, each
is isomorphic to
for some compact normal subgroup
of
, with
for
. If
is first countable, show that this inverse limit can be taken to be a sequence (so that the index set
is simply the natural numbers
with the usual ordering), and the
then shrink to zero in the sense that they lie inside any given open neighbourhood of the identity for
large enough.
Exercise 5 Let
be a totally disconnected locally compact group. Show that every compact subgroup
of
is contained in a compact open subgroup. (Hint: van Dantzig’s theorem provides a compact open subgroup, but it need not contain
. But is there a way to modify it so that it is normalised by
? Why would being normalised by
be useful?)
— 2. The invariance of domain theorem —
In this section we give a proof of the invariance of domain theorem. The main topological tool for this is Brouwer’s famous fixed point theorem:
Theorem 9 (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. We quickly sketch one of these proofs as follows:
Exercise 6 For this exercise, suppose for sake of contradiction that Theorem 9 is false, thus there is a continuous map from
to
with no fixed point.
- Show that there exists a smooth map from
to
with no fixed point.
- Show that there exists a smooth map from
to the unit sphere
, which equals the identity function on
.
- Show that there exists a smooth map
from
to the unit sphere
, which equals the map
on a neighbourhood of
.
- By computing the integral
in two different ways (one by using Stokes’ theorem, and the other by using the
-dimensional nature of the sphere
), establish a contradiction.
Now we prove Theorem 3. By rescaling and translation invariance, it will suffice to show the following claim:
Theorem 10 (Invariance of domain, again) Let
be an continuous injective map. Then
lies in the interior of
.
Let be as in Theorem 10. The map
is a continuous bijection between compact Hausdorff spaces and is thus a homeomorphism. In particular, the inverse map
is continuous. Using the Tietze extension theorem, we can find a continuous function
that extends
.
The function has a zero on
, namely at
. We can use the Brouwer fixed point theorem to show that this zero is stable:
Lemma 11 (Stability of zero) Let
be a continuous function such that
for all
. Then
has at least one zero (i.e. there is a
such that
).
Proof: Apply Theorem 9 to the function
Now suppose that Theorem 10 failed, so that is not an interior point of
. We will use this to locate a small perturbation of
that no longer has a zero on
, contradicting Lemma 11.
We turn to the details. Let be a small number. By continuity of
, we see (if
is chosen small enough) that we have
whenever
and
.
On the other hand, since is not an interior point of
, there exists a point
with
that lies outside
. By translating
if necessary, we may take
; thus
avoids zero,
, and we have
Let denote the set
, where
and
By construction, is compact but does not contain
. Crucially, there is a continuous map
defined by setting
Note that is continuous and well-defined since
avoids zero. Informally,
is a perturbation of
caused by pushing
out a small distance away from the origin
(and hence also away from
), with
being the “pushing” map.
By construction, is non-zero on
; since
is compact,
is bounded from below on
by some
. By shrinking
if necessary we may assume that
.
By the Weierstrass approximation theorem, we can find a polynomial such that
for all ; in particular,
does not vanish on
. At present, it is possible that
vanishes on
. But as
is smooth and
has measure zero,
also has measure zero; so by shifting
by a small generic constant we may assume without loss of generality that
also does not vanish on
. (If one wishes, one can use an algebraic geometry argument here instead of a measure-theoretic one, noting that
lies in an algebraic hypersurface and can thus be generically avoided by perturbation. A purely topological way to avoid zeroes in
is also given in Kulpa’s paper.)
Now consider the function defined by
This is a continuous function that is never zero. From (3), (2) we have
whenever is such that
. On the other hand, if
, then from (2), (1) we have
and hence by (3) and the triangle inequality
Thus in all cases we have
for all . But this, combined with the non-vanishing nature of
, contradicts Lemma 11.
— 3. Hilbert’s fifth problem —
We now establish Theorem 5. Let be a locally Euclidean group. By Exercise 5 of Notes 0,
is Hausdorff; it is also locally compact and first countable. Thus, by Exercise 4, such a group contains an open subgroup
which is isomorphic to the inverse limit
of Lie groups
, each of which has only finitely many components. Clearly,
is also locally Euclidean. If it is Lie, then
is locally Lie and thus Lie. Thus, by replacing
with
if necessary, we may assume without loss of generality that
is the inverse limit
, each of which has only finitely many components.
By Exercise 4, each is isomorphic to the quotient of
by some compact normal subgroup
with
. In particular,
is isomorphic to the quotient of
by a compact normal subgroup
. By Cartan’s theorem (Theorem 2 of Notes 2),
is also a Lie group. Among other things, this implies that the quotient homomorphism from the Lie algebra
of
to the Lie algebra
of
is surjective; indeed, it is the quotient map by the Lie algebra
of
. This implies that there is a continuous map from
to
that inverts the quotient map; in other words, we have a continuous map
from the one-parameter subgroups
of
to the one-parameter subgroups
of
, such that
for all
.
Exercise 7 By iterating these maps and passing to the inverse limit, conclude that for each
, there is a continuous map
such that
for all
.
Because is a Lie group, the exponential map
is a homeomorphism from a neighbourhood of the origin in
to a neighbourhood of the identity in
. We can thus obtain a continuous map
from a neighbourhood of the identity in
to
. Since
, this map is injective.
Now we use the hypothesis that is locally Euclidean (and in particular, has a well-defined dimension
). By Exercise 1, we have
for all . On the other hand, since each
is a quotient of the next Lie group
, one has
Since there are only finitely many possible values for the (necessarily integral) dimension between
and
, we conclude that the dimension must eventually stabilise, i.e. one has
for all sufficiently large . By discarding the first few terms in the sequence and relabeling, we may thus assume that the dimension is constant for all
. Since
, this implies that the Lie groups
have dimension zero for all
. As the
are also compact, they are thus finite. Thus each
is a finite extension of
. As
is the inverse limit of the
as
, we conclude that
is a profinite group, i.e. the inverse limit of finite groups. In particular,
is totally disconnected.
We now study the short exact squence
playing off the locally connected nature of the Lie group against the totally disconnected nature of
.
As discussed earlier, we have a continuous injective map from a neighbourhood
of the identity in
to
that partially inverts the quotient map. By translation, we may normalise
. As
is locally connected, we can find a connected neighborhood
of the identity in
such that
.
Now consider the set . On the one hand, this set is contained in
and contains
; on the other hand, it is connected. As
is totally disconnected, this set must equal
, thus
for all
. A similar argument based on consideration of the set
shows that
for all
. Thus
is a homomorphism from the local group
to
.
Finally, for any , a consideration of the set
reveals that
commutes with
. As a consequence, we see that the preimage
of
under the quotient map
is isomorphic as a local group to
, after identifying
with
for any
and
.
On the other hand, is locally Euclidean, and hence
is locally Euclidean also, and in particular locally connected. This implies that
is locally connected; but as
is also totally disconnected, it must be discrete. This
is now locally isomorphic to
and hence to
, and is thus locally Lie and hence Lie as required. (Here, we say that two groups are locally isomorphic if they have neighbourhoods of the identity which are isomorphic to each other as local groups.)
Exercise 8 Let
be a locally compact Hausdorff first-countable group which is “finite-dimensional” in the sense that it does not contain continuous injective images of non-trivial open sets of Euclidean spaces
of arbitrarily large dimension. Show that
is locally isomorphic to the direct product
of a Lie group
and a totally disconnected compact group
. (Note that this local isomorphism does not necessarily extend to a global isomorphism, as the example of the solenoid group
shows.)
Remark 4 Of course, it is possible for locally compact groups to be infinite-dimensional; a simple example is the infinite-dimensional torus
, which is compact, abelian, metrisable, and locally connected, but infinite dimensional. (It will still be an inverse limit of Lie groups, though.)
Exercise 9 Show that a topological group is Lie if and only if it is locally compact, Hausdorff, first-countable, locally connected, and finite-dimensional.
Remark 5 It is interesting to note that this characterisation barely uses the real numbers
, which are of course fundamental in defining the smooth structure of a Lie group; the only remaining reference to
comes through the notion of finite dimensionality. It is also possible, using dimension theory, to obtain alternate characterisations of finite dimensionality (e.g. finite Lebesgue covering dimension) that avoid explicit mention of the real line, thus capturing the concept of a Lie group using only the concepts of point-set topology (and the concept of a group, of course).
— 4. Transitive actions —
We now prove Proposition 6. As this is a stronger statement than Theorem 5, it will not be surprising that we will be using a very similar argument to prove the result.
Let be a locally compact
-compact group that acts transitively, faithfully, and continuously on a connected manifold
. The advantage of transitivity is that one can now view
as a homogeneous space
of
, where
is the stabiliser of a point
(and is thus a closed subgroup of
). Note that a priori, we only know that
and
are identifiable as sets, with the identification map
defined by setting
being continuous; but thanks to the
-compact hypothesis, we can upgrade
to a homeomorphism. Indeed, as
is
-compact,
is also; and so given any compact neighbourhood of the identity
in
,
can be covered by countably many translates of
. By the Baire category theorem, one of these translates
has an image
in
with non-empty interior, which implies that
has
as an interior point. From this it is not hard to see that the map
is open; as it is also a continuous bijection, it is therefore a homeomorphism.
By the Gleason-Yamabe theorem (Theorem 2), has an open subgroup
that is the inverse limit of Lie groups. (Note that
is Hausdorff because it acts faithfully on the Hausdorff space
.)
acts transitively on
, which is an open subset of
and thus also a manifold. Thus, we may assume without loss of generality that
is itself the inverse limit of Lie groups.
As is
-compact, the manifold
is also. As
acts faithfully on
, this makes
first countable; and so (by Exercise 4)
is the inverse limit of a sequence of Lie groups
, with each
projecting surjectively onto
, and with the
shrinking to the identity.
Let be the projection of
onto
; this is a closed subgroup of the Lie group
, and each
projects surjectively onto
. Then
are manifolds, and
is the inverse limit of the
.
Exercise 10 Show that the dimensions of the
are be non-decreasing, and bounded above by the dimension of
. (Hint: repeat the arguments of the previous section. The
need no longer be compact, but they are still closed, and this still suffices to make the preceding arguments go through.)
Thus, for large enough, the dimensions of
must be constant; by renumbering, we may assume that all the
have the same dimension. As each
is a cover of
with structure group
, we conclude that the
are zero-dimensional and compact, and thus finite. On the other hand,
is locally connected, which implies that the
are eventually trivial. Indeed, if we pick a simply connected neighbbourhood
of the identity in
, then by local connectedness of
, there exists a connected neighbourhood
of the identity in
whose projection to
is contained in
. Being open,
must contain one of the
. If
is non-trivial for any
, then the projection of
to
will then be disconnected (as this projection will be contained in a neighbourhood with the topological structure of
, and its intersection with the latter fibre is at least as large as
. We conclude that
is trivial for
large enough, and so
is a Lie group as required.
11 comments
Comments feed for this article
10 October, 2011 at 10:43 am
Marius Buliga
A comment and a questions:
1. Re Remark 5, I would say that, according to the definition of “finite dimensional” from Exercise 7, there is a lot of structure entering by the back door, namely something akin to rectifiability (statements about continuous images of euclidean spaces). Maybe related, the problem of defining rectifiability in sub-riemannian spaces (which are main examples of spaces non euclidean at any scale) is open, in the sense that there is no satisfactory definition of rectifiability for these spaces.
2. Question: In your Theorem 12 (Freiman’s type theorem) from your Notes 0, if you take
to be the dyadic integers (or some noncommutative nilpotent like, but dyadic version), what could be the locally compact group generated by the said sequence of approximate groups, etc ?
Thanks!
10 October, 2011 at 11:35 am
Terence Tao
I’m not sure what the “group of dyadic integers” is in Q2; also, to specify an approximate group, one needs to not just specify the ambient group G, but also a subset A of that group G obeying the approximate group axioms. So I don’t have enough information to answer your question.
If you mean G to be the additive group
of rationals whose denominator is a power of two, and let
be the multiples of
between -1 and 1 for some large
, then this is an approximate group, and in the limit
, the associated ultraproduct can be modeled by the interval
in the real line. But I don’t know if this is the type of example you have in mind.
11 October, 2011 at 12:04 am
Marius Buliga
Thank you for the reply. I don’t know your proof, but I am thinking about taking a group G which is not locally euclidean (say Q2 or nilpotent versions, easy to construct, or even some strange group a la Grigorchuk). Inside this group, maybe, you can choose a (sequence of?) approximate group(s) with two properties: first, the approximate group is big enough to feel (??) all the group G, second, you can tell exactly what is the locally euclidean group (say G’) which has inside a progression which is roughly equivalent with the approximate group.
The question is: is there any relation between G and G’? (like approximate morphism, or a functorial correspondence G-G’…)
10 October, 2011 at 11:11 am
Neville
“Roughly speaking, this theorem … of a locally group …”
locally compact group?
[Corrected, thanks – T.]
27 October, 2011 at 8:37 pm
254A, Notes 7: Models of ultra approximate groups « What’s new
[…] precompact neighbourhood of the identity such that . By the Gleason-Yamabe theorem (Theorem 1 of Notes 5), there is an open subgroup of , and a compact normal subgroup of contained in , such that is […]
6 November, 2011 at 11:45 am
Ben Hayes
In Exercise 8, I’m a little confused by the concept of “local isomorphism.” Do you mean that there is an open subset
of
,a Lie Group
, a compact totally disconnected group
, and a homeomorphism
such that
,
are defined for all
and equal
,
? Or do you mean there are open subset
,
of
,
and a homeomorphism
which respects multiplication and inversion (when it is defined).
6 November, 2011 at 11:56 am
Terence Tao
I mean the latter; I’ll reword to clarify. (The example of the solenoid group from Notes 0 shows that the former is not true.)
6 November, 2011 at 12:11 pm
Ben Hayes
Okay, thanks It became clearer when I read the surrounding text.
4 October, 2014 at 8:43 am
Tomek Kania
Let me point out a minor typo: “theorem of van Dantzig’s theorem”.
[Corrected, thanks – T.]
29 June, 2017 at 8:45 am
Jeremy Brazas
I’m glad I found this post. Thank you! Do you know of any path-connected examples of inverse limits of Lie groups that are non-trivial in the sense that they are not isomorphic to a Lie group or an infinite product of Lie groups? Connected examples like solenoids are abundant but are not path connected.
11 September, 2021 at 10:20 am
Liam
You write “finite-dimensional in the sense that it does not contain continuous injective images of non-trivial open sets of Euclidean spaces Rn of arbitrarily large dimension.” Is it difficult to link this and the notions of dimensions of suitable topological spaces: lebesgue covering dimension, small inductive dimension or large inductive dimension ?