Hilbert's Fifth Problem and Related Topics
Hilbert's Fifth Problem and Related Topics
Topics
Terence Tao
This is a preliminary version of the book Hilberts Fifth Problem and Related Topics published by the
American Mathematical Society (AMS). This preliminary version is made available with the permission of
the AMS and may not be changed, edited, or reposted at any other website without explicit written
permission from the author and the AMS.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
Contents
Preface
ix
Notation
Acknowledgments
xi
Chapter 1.
1.1.
Introduction
1.2.
22
1.3.
48
1.4.
67
1.5.
1.6.
114
1.7.
126
1.8.
152
1.9.
181
1.10.
Chapter 2.
Related articles
213
2.1.
214
2.2.
217
2.3.
Ados theorem
228
2.4.
237
vii
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
viii
Contents
2.5.
Local groups
243
2.6.
254
2.7.
265
2.8.
271
2.9.
280
2.10.
283
2.11.
294
Bibliography
299
Index
305
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
Preface
ix
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
Preface
This text is based on the lecture notes from that course, as well as from
some additional posts on my blog at terrytao.wordpress.com on further
topics related to Hilberts fifth problem. The first chapter of this text can
thus serve as the basis for a one-quarter or one-semester advanced graduate course, depending on how much of the optional material one wishes to
cover. The material here assumes familiarity with basic graduate real analysis (such as measure theory and point set topology), as covered for instance
in my texts [Ta2011], [Ta2010], and including topics such as the Riesz representation theorem, the Arzela-Ascoli theorem, Tychonoffs theorem, and
Urysohns lemma. A basic understanding of linear algebra (including, for
instance, the spectral theorem for unitary matrices) is also assumed.
The core of the text is Chapter 1. The first part of this chapter is devoted
to the theory surrounding Hilberts fifth problem, and in particular in fleshing out the long road from locally compact groups to Lie groups. First, the
theory of Lie groups and Lie algebras is reviewed, and it is shown that a Lie
group structure can be built from a special type of metric known as a Gleason metric, thanks to tools such as the Baker-Campbell-Hausdorff formula.
Some representation theory (and in particular, the Peter-Weyl theorem) is
introduced next, in order to classify compact groups. The two tools are
then combined to prove the fundamental Gleason-Yamabe theorem, which
among other things leads to a positive solution to Hilberts fifth problem.
After this, the focus turns from the soft analysis of locally compact
groups to the hard analysis of approximate groups, with the useful tool of
ultraproducts serving as the key bridge between the two topics. By using this
bridge, one can start imposing approximate Lie structure on approximate
groups, which ultimately leads to a satisfactory classification of approximate
groups as well. Finally, Chapter 1 ends with applications of this classification
to geometric group theory and the geometry of manifolds, and in particular
in reproving Gromovs theorem on groups of polynomial growth.
Chapter 2 contains a variety of additional material that is related to one
or more of the topics covered in Chapter 1, but which can be omitted for
the purposes of teaching a graduate course on the subject.
Notation
For reasons of space, we will not be able to define every single mathematical
term that we use in this book. If a term is italicised for reasons other than
emphasis or for definition, then it denotes a standard mathematical object,
result, or concept, which can be easily looked up in any number of references.
(In the blog version of the book, many of these terms were linked to their
Wikipedia pages, or other on-line reference pages.)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
Acknowledgments
xi
Acknowledgments
I am greatly indebted to my students of the course on which this text was
based, as well as many further commenters on my blog, including Marius
Buliga, Tony Carbery, Nick Cook, Alin Galatan, Pierre de la Harpe, Ben
Hayes, Richard Hevener, Vitali Kapovitch, E. Mehmet Kiral, Allen Knutson,
Mateusz Kwasnicki, Fred Lunnon, Peter McNamara, William Meyerson,
Joel Moreira, John Pardon, Ravi Raghunathan, David Roberts, David Ross,
Olof Sisask, David Speyer, Benjamin Steinberg, Neil Strickland, Lou van den
Dries, Joshua Zelinsky, Pavel Zorin, and several anonymous commenters.
These comments can be viewed online at
terrytao.wordpress.com/category/teaching/254a-hilberts-fifth-problem/
The author was supported by a grant from the MacArthur Foundation,
by NSF grant DMS-0649473, and by the NSF Waterman award. Last, but
not least, I thank Emmanuel Breuillard and Ben Green for introducing me
to the beautiful interplay between geometric group theory, additive combinatorics, and topological group theory that arises in this text.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
Chapter 1
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
1.1. Introduction
This text focuses on three related topics:
Hilberts 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, see Theorem 1.1.17);
Approximate groups in nonabelian groups, and their classification
[Hr2012], [BrGrTa2011] via the Gleason-Yamabe theorem; and
Gromovs theorem [Gr1981] on groups of polynomial growth, as
proven via the classification of approximate groups (as well as some
consequences to fundamental groups of Riemannian manifolds).
These three families of results exemplify two broad principles (part of
what I like to call the the dichotomy between structure and randomness
[Ta2008]):
(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) Furthermore, 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 us illustrate 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 n n complex matrix A Mn (C), define the
matrix exponential exp(A) of A by the formula
exp(A) :=
X
Ak
k=0
k!
=1+A+
1 2
1
A + A3 + . . .
2!
3!
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
1.1. Introduction
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
for all real t, which gives existence of the representation and also real analyticity and smoothness. Finally, uniqueness of the representation (t) =
exp(tA) follows from the identity
d
A=
exp(tA)|t=0 .
dt
Exercise 1.1.2. Generalise Proposition 1.1.1 by replacing the hypothesis
that is continuous with the hypothesis that is Lebesgue measurable.
(Hint: use the Steinhaus theorem, see e.g. [Ta2011, Exercise 1.6.8].) 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 R and range Mn (C); 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
Jordans theorem [Jo1878], which uses a discreteness hypothesis to upgrade
Lie type structure to nilpotent (and in this case, abelian) structure. We
shall formulate Jordans theorem in a slightly stilted fashion in order to
emphasise the adherence to the above-mentioned principles.
Theorem 1.1.2 (Jordans theorem). Let G be an object with the following
properties:
(1) (Group-like object) G is a group.
(2) (Discreteness) G is finite.
(3) (Lie-type structure) G is a subgroup of Un (C) (the group of unitary
n n matrices) for some n.
Then there is a subgroup G0 of G such that
(i) (G0 is close to G) The index |G/G0 | of G0 in G is On (1) (i.e.
bounded by Cn for some quantity Cn depending only on n).
(ii) (Nilpotent-type structure) G0 is abelian.
A key observation in the proof of Jordans theorem is that if two unitary
elements g, h Un (C) are close to the identity, then their commutator
[g, h] = g 1 h1 gh is even closer to the identity (in, say, the operator norm
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
1.1. Introduction
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
not necessarily unitary, and thus do not necessarily preserve the standard
Hilbert inner product of Cn . However, if one averages that inner product by
the finite group G, one obtains a new inner product on Cn that is preserved
by G, which allows one to conjugate G to a subgroup of Un (C). This
averaging trick is (a small) part of Weyls unitary trick in representation
theory.)
Remark 1.1.3. We remark that one can strengthen Jordans theorem further by relaxing the finiteness assumption on G to a periodicity assumption;
see Section 2.1.
Exercise 1.1.4 (Inability to discretise nonabelian Lie groups). Show that
if n 3, then the orthogonal group On (R) cannot contain arbitrarily dense
finite subgroups, in the sense that there exists an = n > 0 depending
only on n such that for every finite subgroup G of On (R), there exists a ball
of radius in On (R) (with, say, the operator norm metric) that is disjoint
from G. What happens in the n = 2 case?
Remark 1.1.4. More precise classifications of the finite subgroups of Un (C)
are known, particularly in low dimensions. For instance, it is a classical
result that the only finite subgroups of SO3 (R) (which SU2 (C) 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.
1.1.1. Hilberts fifth problem. One of the fundamental categories of
objects in modern mathematics is the category of Lie groups, which are
rich in both algebraic and analytic structure. Let us now briefly recall the
precise definition of what a Lie group is.
Definition 1.1.5 (Smooth manifold). Let d 0 be a natural number. A ddimensional topological manifold is a Hausdorff topological space M which
is locally Euclidean, thus every point in M has a neighbourhood which is
homeomorphic to an open subset of Rd .
A smooth atlas on an d-dimensional topological manifold M is a family
( )A of homeomorphisms : U V from open subsets U of M to
open subsets V of Rd , such that the U form an open cover of M , and for
any , A, the map 1
is smooth (i.e. infinitely differentiable) on
the domain of definition (U U ). Two smooth atlases are equivalent if
their union is also a smooth atlas; this is easily seen to be an equivalence
relation. An equivalence class of smooth atlases is a smooth structure. A
smooth manifold is a topological manifold equipped with a smooth structure.
A map : M M 0 from one smooth manifold to another is said to be
smooth if 0 1
is a smooth function on the domain of definition V
1
1
(U (U )) for any smooth charts , 0 in any the smooth atlases
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
1.1. Introduction
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
the topology induced from Rd ) and are thus not Lie groups; similarly, quotients such as Rd /Qd are not Lie groups either (they are not even Hausdorff).
Also, infinite-dimensional topological vector spaces (such as RN with the
product topology) will not be Lie groups.
Example 1.1.11. The general linear group GLn (C) of invertible n n
complex matrices is a Lie group. A theorem of Cartan (Theorem 1.3.2)
asserts that any closed subgroup of a Lie group is a smooth submanifold of
that Lie group and is in particular also a Lie group. In particular, closed
linear groups (i.e. closed subgroups of a general linear group) are Lie groups;
examples include the real general linear group GLn (R), the unitary group
Un (C), the special unitary group SUn (C), the orthogonal group On (R), the
special orthogonal group SOn (R), and the Heisenberg group
1 R R
0 1 R
0 0 1
of unipotent upper triangular 3 3 real matrices. Many Lie groups are
isomorphic to closed linear groups; for instance, the additive group R can
be identified with the closed linear group
1 R
,
0 1
the circle R/Z can be identified with SO2 (R) (or U1 (C)), and so forth.
However, not all Lie groups are isomorphic to closed linear groups. A somewhat trivial example is that of a discrete group with cardinality larger than
the continuum, which is simply too large to fit inside any linear group. A
less pathological example is provided by the Weil-Heisenberg group
1 R R/Z
1 R R
1 0 Z
R := 0 1 R / 0 1 0
(1.3)
G := 0 1
0 0
1
0 0 1
0 0 1
which is isomorphic to the image of the Heisenberg group under the Weil
representation, or equivalently the group of isometries of L2 (R) generated
by translations and modulations. Despite this, though, it is helpful to think
of closed linear groups and Lie groups as being almost the same concept
as a first approximation. For instance, one can show using Ados theorem
(Theorem 2.3.2) that every Lie group is locally isomorphic to a linear local
group (a concept we will discuss in Section 1.2.1).
An important subclass of the closed linear groups are the linear algebraic
groups, in which the group is also a real or complex algebraic variety (or at
least an algebraically constructible set). All of the examples of closed linear
groups given above are linear algebraic groups, although there exist closed
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
1.1. Introduction
linear groups that are not isomorphic to any algebraic group; see Proposition
2.11.2.
Exercise 1.1.5 (Weil-Heisenberg group is not linear). Show that there is
no injective homomorphism : G GLn (C) from the Weil-Heisenberg
group (1.3) to a general linear group GLn (C) for any finite n. (Hint: The
centre [G, G] maps via to a circle subgroup of GLn (C); diagonalise this
subgroup and reduce to the case when the image of the centre consists of
multiples of the identity. Now, use the fact that commutators in GLn (C)
have determinant one.) This fact was first observed by Birkhoff.
Hilberts fifth problem, like many of Hilberts problems, does not have a
unique interpretation, but one of the most commonly accepted interpretations of the question posed by Hilbert is to determine if the requirement of
smoothness in the definition of a Lie group is redundant. (There is also an
analogue of Hilberts fifth problem for group actions, known as the HilbertSmith conjecture; see Section 2.7.) To answer this question, we need to relax
the notion of a Lie group to that of a topological group.
Definition 1.1.12 (Topological group). A topological group is a group G =
(G, ) that is also a topological space, in such a way that the group operations
: GG G and ()1 : G G are continuous. (As before, we also consider
additive topological groups G = (G, +) provided that they are abelian.)
Clearly, every Lie group is a topological group if one simply forgets the
smooth structure, retaining only the topological and group structures. Furthermore, such topological groups remain locally Euclidean. It was established by Montgomery-Zippin [MoZi1952] and Gleason [Gl1952] that the
converse statement holds, thus solving at least one formulation of Hilberts
fifth problem:
Theorem 1.1.13 (Hilberts fifth problem). Let G be an object with the
following properties:
(1) (Group-like object) G is a topological group.
(2) (Weak regularity) G is locally Euclidean.
Then
(i) (Lie-type structure) G is isomorphic to a Lie group.
Exercise 1.1.6. Show that a locally Euclidean topological group is necessarily Hausdorff (without invoking Theorem 1.1.13).
We will prove this theorem in Section 1.6. As it turns out, Theorem
1.1.13 is not directly useful for many applications, because it is often difficult
to verify that a given topological group is locally Euclidean. On the other
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
10
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
1.1. Introduction
11
(although we will see later that one can do so by restricting to the locally
connected case).
We have now seen several examples of locally compact groups that are
not Lie groups. However, all of these examples are almost Lie groups in
that they can be turned into Lie groups by quotienting out a small compact
normal subgroup. (It is easy to see that the quotient of a locally compact
group by a compact normal subgroup is again a locally compact group.) For
instance, a group with the trivial topology becomes Lie after quotienting out
the entire group (which is small in the sense that it is contained in every
open neighbourhood of the origin). The infinite-dimensional torus (R/Z)N
can be quotiented into a finite-dimensional torus (R/Z)d (which is of course
a Lie group) by quotienting out the compact subgroup {0}d (R/Z)N ; note
from the definition of the product topology that these compact subgroups
shrink to zero in the sense that every neighbourhood of the group identity
contains at least one (and in fact all but finitely many) of these subgroups.
Similarly, with the p-adic group Zp , one can quotient out by the compact
(and open) subgroups pj Zp (which also shrink to zero, as discussed above) to
obtain the cyclic groups Z/pj Z, which are discrete and thus Lie. Quotienting
out Qp by the same compact open subgroups pj Zp also leads to discrete
(hence Lie) quotients; similarly for algebraic groups defined over Qp , such
as GLn (Qp ). Finally, with the solenoid group G := (Zp R)/Z , one can
quotient out the copy of pj Zp {0} in G for j = 0, 1, 2, . . . (which are
another sequence of compact subgroups shrinking to zero) to obtain the
quotient group (Z/pj Z R)/Z , which is isomorphic to a (highly twisted)
circle R/Z and is thus Lie.
Inspired by these examples, we might be led to the following conjecture:
if G is a locally compact group, and U is a neighbourhood of the identity,
then there exists a compact normal subgroup K of G contained in U such
that G/K is a Lie group. In the event that G is Hausdorff, this is equivalent
to asserting that G is the projective limit (or inverse limit) of Lie groups.
This conjecture is true in several cases; for instance, one can show using
the Peter-Weyl theorem (which we will discuss in Section 1.4) that it is true
for compact groups, and we will later see that it is also true for connected
locally compact groups (see Theorem 1.6.1). However, it is not quite true
in general, as the following example shows.
Exercise 1.1.8. Let p be a prime, and let T : Qp Qp be the automorphism T x := px. Let G := Qp oT Z be the semidirect product of Qp and Z
twisted by T ; more precisely, G is the Cartesian product Qp Z with the
product topology and the group law
(x, n)(y, m) := (x + T n y, n + m).
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
12
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
1.1. Introduction
13
difficult to then conclude that any locally compact NSS group (regardless of
connectedness) is Lie. Conversely, this claim (which we isolate as Corollary
1.5.8) turns out to be a key step in the proof of Theorem 1.1.17, as we shall
see later. (It is also not difficult to show that all Lie groups are NSS; see
Exercise 1.5.6.)
The proof of the Gleason-Yamabe theorem proceeds in a somewhat
lengthy series of steps in which the initial regularity (local compactness)
on the group G is gradually upgraded to increasingly stronger regularity
(e.g. metrisability, the NSS property, or the locally Euclidean property) until one eventually obtains Lie structure; see Figure 1. A key turning point
in the argument will be the construction of a metric (which we call a Gleason metric) on (a large portion of) G which obeys a commutator estimate
similar to (1.2).
While the Gleason-Yamabe theorem does not completely classify all locally compact groups (as mentioned earlier, it primarily controls the mediumscale behaviour, and not the very fine-scale or very coarse-scale behaviour,
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
14
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
1.1. Introduction
15
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
16
In this text we will focus on the latter question, as this allows us to bring
in qualitative tools such as the Gleason-Yamabe theorem to bear on the
problem.
In the abelian case when the ambient group G is additive, approximate
groups are classified by Freimans theorem for abelian groups 3 [GrRu2007].
As before, we phrase this theorem in a slightly stilted fashion (and in a
qualitative, rather than quantitative, manner) in order to demonstrate its
alignment with the general principles stated in the introduction.
Theorem 1.1.24 (Freimans theorem in an abelian group). Let A be an
object with the following properties:
(1) (Group-like object) A is a subset of an additive group G.
(2) (Weak regularity) A is a K-approximate group.
(3) (Discreteness) A is finite.
Then there exists a finite subgroup H of G, and a subset P of G/H, with
the following properties:
(i) (P is close to A/H) 1 (P ) is contained in 4A := A + A + A + A,
where : G G/H is the quotient map, and |P | K |A|/|H|.
(ii) (Nilpotent type structure) P is a symmetric generalised arithmetic
progression of rank OK (1) (see Example 1.1.20).
Informally, this theorem asserts that in the abelian setting, discrete
approximate groups are essentially bounded rank symmetric generalised
arithmetic progressions, extended by finite groups (such extensions are also
known as coset progressions). The theorem has a simpler conclusion (and is
simpler to prove) in the case when G is a torsion-free abelian group (such
as Z), since in this case H is trivial.
We will not discuss the proof of Theorem 1.1.24 from [GrRu2007] here,
save to say that it relies heavily on Fourier-analytic methods, and as such,
does not seem to easily extend to a general non-abelian setting. To state the
non-abelian analogue of Theorem 1.1.24, one needs multiplicative analogues
of the concept of a generalised arithmetic progression. An ordinary (symmetric) arithmetic progression {N v, . . . , N v} has an obvious multiplicative
analogue, namely a (symmetric) geometric progression {aN , . . . , aN } for
some generator a G. In a similar vein, if one has r commuting generators
a1 , . . . , ar and some dimensions N1 , . . . , Nr > 0, one can form a symmetric
generalised geometric progression
(1.4)
3The original theorem of Freiman [Fr1973] obtained an analogous classification in the case
when G was a torsion-free abelian group, such as the integers Z.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
1.1. Introduction
17
1 Z Z
(1.5)
G = 0 1 Z
0 0 1
and let
1 1 0
1 0 0
e1 := 0 1 0 , e2 := 0 1 1
0 0 1
0 0 1
be the two generators of G. Let N 1 be a sufficiently large natural number.
Show that the noncommutative
progression P (e1 , e2 ; N, N ) contains all the
1 a c
group elements 0 1 b of G with |a|, |b| N and |c| N 2 for a
0 0 1
sufficiently small absolute constant > 0; conversely,
show that all elements
1 a c
of P (e1 , e2 ; N, N ) are of the form 0 1 b with |a|, |b| CN and |c|
0 0 1
2
CN for some sufficiently large absolute constant C > 0. Thus, informally,
we have
1 O(N ) O(N 2 )
1
O(N ) .
P (e1 , e2 ; N, N ) = 0
0
0
1
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
18
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
1.1. Introduction
19
1.1.3. Gromovs theorem. The final topic of this chapter will be Gromovs theorem on groups of polynomial growth [Gr1981]. This theorem
is analogous to Theorem 1.1.17 or Theorem 1.1.29, but in the category of
finitely generated groups rather than locally compact groups or approximate
groups.
Let G be a group that is generated by a finite set S of generators;
for notational simplicity we will assume that S is symmetric and contains
the origin. Then S defines a (right-invariant) word metric on G, defined
by setting d(x, y) for x, y G to be the least natural number n such that
x S n y. One easily verifies that this is indeed a metric that is right-invariant
(thus d(xg, yg) = d(x, y) for all x, y, g G). Geometrically, this metric
describes the geometry of the Cayley graph on G formed by connecting x to
sx for each x G and s S. (See [Ta2011c, 2.3] for more discussion of
using Cayley graphs to study groups geometrically.)
Let us now consider the growth of the balls B(1, R) = S bRc as R ,
where bRc is the integer part of R. On the one hand, we have the trivial
upper bound
|B(1, R)| |S|R
that shows that such balls can grow at most exponentially. And for typical
non-abelian groups, this exponential growth actually occurs; consider the
case for instance when S consists of the generators of a free group (together
with their inverses, and the group identity). However, there are some groups
for which the balls grow at a much slower rate. A somewhat trivial example
is that of a finite group G, since clearly |B(1, R)| will top out at |G| (when
R reaches the diameter of the Cayley graph) and stop growing after that
point. Another key example is the abelian case:
Exercise 1.1.11. If G is an abelian group generated by a finite symmetric
set S containing the identity, show that
|B(1, R)| (1 + R)|S| .
In particular, B(1, R) grows at a polynomial rate in R.
Let us say that a finitely group G is a group of polynomial growth if one
has |B(1, R)| CRd for all R 1 and some constants C, d > 0.
Exercise 1.1.12. Show that the notion of a group of polynomial growth
(as well as the rate d of growth) does not depend on the choice of generators
S; thus if S 0 is another set of generators for G, show that G has polynomial
growth with respect to S with rate d if and only if it has polynomial growth
with respect to S 0 with rate d.
Exercise 1.1.13. Let G be a finitely generated group, and let G0 be a finite
index subgroup of G.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
20
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
1.1. Introduction
21
In Section 1.9 we will use this connection to deduce Theorem 1.1.30 from
Theorem 1.1.29. From a historical perspective, this was not the first proof
of Gromovs theorem; Gromovs original proof in [Gr1981] relied instead on
a variant of Theorem 1.1.13 (as did some subsequent variants of Gromovs
argument, such as the nonstandard analysis variant in [vdDrWi1984]), and
a subsequent proof of Kleiner [Kl2010] went by a rather different route,
based on earlier work of Colding and Minicozzi [CoMi1997] on harmonic
functions of polynomial growth. (This latter proof is discussed in [Ta2009,
1.2] and [Ta2011c, 2.5].) The proof we will give in this text is more
recent, based on an argument of Hrushovski [Hr2012]. We remark that
the strategy used to prove Theorem 1.1.29 - namely taking an ultralimit
of a sequence of approximate groups - also appears in Gromovs original
argument5. We will discuss these sorts of limits more carefully in Section
1.7, but an informal example to keep in mind for now is the following: if
one takes a discrete group (such as Zd ) and rescales it (say to N1 Zd for a
large parameter N ), then intuitively this rescaled group converges to a
continuous group (in this case Rd ). More generally, one can generate locally
compact groups (or at least locally compact spaces) out of the limits of
(suitably normalised) groups of polynomial growth or approximate groups,
which is one of the basic observations that tie the three different topics
discussed above together.
As we shall see in Section 1.10, finitely generated groups arise naturally
as the fundamental groups of compact manifolds. Using the tools of Riemannian geometry (such as the Bishop-Gromov inequality), one can relate
the growth of such groups to the curvature of a metric on such a manifold.
As a consequence, Gromovs theorem and its variants can lead to some nontrivial conclusions about the relationship between the topology of a manifold
and its geometry. The following simple consequence is typical:
Proposition 1.1.31. Let M be a compact Riemannian manifold of nonnegative Ricci curvature. Then the fundamental group 1 (M ) of M is virtually nilpotent.
We will discuss this result and some related results (such as a relaxation
of the non-negative curvature hypothesis to an almost non-negative curvature hypothesis) in Section 1.10. We also remark that the above proposition
can also be proven (with stronger conclusions) by more geometric means,
but there are some results of the above type which currently have no known
proof that does not employ some version of Gromovs theorem at some point.
5Strictly speaking, he uses Gromov-Hausdorff limits instead of ultralimits, but the two types
of limits are closely related, as we shall see in Section 1.7.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
22
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
23
1.2.1. Local groups. The connection between Lie groups and Lie algebras
will be local in nature - the only portion of the Lie group that will be of
importance will be the portion that is close to the group identity 1. To
formalise this locality, it is convenient to introduce the notion of a local
group and a local Lie group, which are local versions of the concept of a
topological group and a Lie group respectively. We will only set up the
barest bones of the theory of local groups here; a more detailed discussion
is given in Section 2.5.
Definition 1.2.2 (Local group). A local topological group G = (G, , , 1, , ()1 ),
or local group for short, is a topological space G equipped with an identity
element 1 G, a partially defined but continuous multiplication operation
: G for some domain G G, and a partially defined but continuous inversion operation ()1 : G, where G, obeying the following
axioms:
(1) (Local closure) is an open neighbourhood of G {1} {1} G,
and is an open neighbourhood of 1.
(2) (Local associativity) If g, h, k G are such that (gh)k and g(hk)
are both well-defined in G, then they are equal. (Note however that
it may be possible for one of these products to be defined but not
the other.)
(3) (Identity) For all g G, g 1 = 1 g = g.
(4) (Local inverse) If g G and g 1 are well-defined in G, then6 g
g 1 = g 1 g = 1. (In particular this, together with the other
axioms, forces 11 = 1.)
We will sometimes use additive notation for local groups if the groups
are abelian (by which we mean the statement that if g + h is defined, then
h + g is also defined and equal to g + h.)
A local group is said to be symmetric if = G, i.e. if every element g
in G has an inverse g 1 that is also in G.
A local Lie group is a local group that is also a smooth manifold, in such
a fashion that the partially defined group operations , ()1 are smooth on
their domain of definition.
Clearly, every topological group is a local group, and every Lie group is
a local Lie group. We will sometimes refer to the former concepts as global
topological groups and global Lie groups in order to distinguish them from
their local counterparts. One could also consider local discrete groups, in
6Here we adopt the convention that any mathematical sentence involving an undefined operation is automatically false, thus for instance g g 1 = 1 is false unless g g 1 is well-defined,
so that (g, g 1 ) .
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
24
which the topological structure is just the discrete topology, but we will not
need to study such objects in here.
A model class of examples of a local (Lie) group comes from restricting a
global (Lie) group to an open neighbourhood of the identity. Let us formalise
this concept:
Definition 1.2.3 (Restriction). If G is a local group, and U is an open
neighbourhood of the identity in G, then we define the restriction G U
of G to U to be the topological space U with domains U := {(g, h)
: g, h, g h U } and U := {g : g, g 1 U }, and with the group
operations , ()1 being the restriction of the group operations of G to U ,
U respectively. If U is symmetric (in the sense that g 1 is well-defined
and lies in U for all g U ), then this restriction G U will also be symmetric.
If G is a global or local Lie group, then G U will also be a local Lie group.
We will sometimes abuse notation and refer to the local group G U simply
as U .
Thus, for instance, one can take the Euclidean space Rd , and restrict it
to a ball B centred at the origin, to obtain an additive local group Rd B . In
this group, two elements x, y in B have a well-defined sum x + y only when
their sum in Rd stays inside B. Intuitively, this local group behaves like the
global group Rd as long as one is close enough to the identity element 0,
but as one gets closer to the boundary of B, the group structure begins to
break down.
It is natural to ask the question as to whether every local group arises as
the restriction of a global group. The answer to this question is somewhat
complicated, and can be summarised as essentially yes in certain circumstances, but not in general; see Section 2.5.
A key example of a local Lie group arises from pushing forward a Lie
group via a coordinate chart near the origin:
Example 1.2.4. Let G be a global or local Lie group of some dimension d,
and let : U V be a smooth coordinate chart from a neighbourhood U
of the identity 1 in G to a neighbourhood V of the origin 0 in Rd , such that
maps 1 to 0. Then we can define a local group G U which is the set
V (viewed as a smooth submanifold of Rd ) with the local group identity 0,
the local group multiplication law defined by the formula
x y := (1 (x) 1 (y))
defined whenever 1 (x), 1 (y), 1 (x) 1 (y) are well-defined and lie in
U , and the local group inversion law ()1 defined by the formula
x1 := (1 (x)1 )
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
25
defined whenever 1 (x), 1 (x)1 are well-defined and lie in U . One easily
verifies that G U is a local Lie group. We will sometimes denote this local
Lie group as (V, ), to distinguish it from the additive local Lie group (V, +)
arising by restriction of (Rd , +) to V . The precise distinction between the
two local Lie groups will in fact be a major focus of this section.
Example 1.2.5. Let G be the Lie group GLn (R), and let U be the ball
U := {g GLn (R) : kg 1kop < 1}. If we then let V Mn (R) be the ball
V := {x Mn (R) : kxkop < 1} and be the map (g) := g 1, then
is a smooth coordinate chart (after identifying Mn (R) with Rnn ), and by
the construction in the preceding exercise, V = G U becomes a local Lie
group with the operations
x y := x + y + xy
(defined whenever x, y, x + y + xy all lie in V ) and
x1 := (1 + x)1 1 = x x2 + x3 . . .
(defined whenever x and (1+x)1 1 both lie in V ). Note that this Lie group
structure is not equal to the additive structure (V, +) on V , nor is it equal
to the multiplicative structure (V, ) on V given by matrix multiplication,
which is one of the reasons why we use the symbol instead of + or for
such structures.
Many (though not all) of the familiar constructions in group theory can
be generalised to the local setting, though often with some slight additional
subtleties. We will not systematically do so here, but we give a single such
generalisation for now:
Definition 1.2.6 (Homomorphism). A continuous homomorphism : G
H between two local groups G, H is a continuous map from G to H with
the following properties:
(i) maps the identity 1G of G to the identity 1H of H: (1G ) = 1H .
(ii) If g G is such that g 1 is well-defined in G, then (g)1 is welldefined in H and is equal to (g 1 ).
(iii) If g, h G are such that g h is well-defined in G, then (g) (h)
is well-defined and equal to (g h).
A smooth homomorphism : G H between two local Lie groups G, H is
a continuous homomorphism that is also smooth.
A (continuous) local homomorphism : U H between two local
groups G, H is a continuous homomorphism from an open neighbourhood
U of the identity in G to H. Two local homomorphisms are said to hbe
equivalent if they agree on a (possibly smaller) open neighbourhood of the
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
26
identity. One can of course define the notion of a smooth local homomorphism similarly.
It is easy to see that the composition of two continuous homomorphisms
is again a continuous homomorphism, and that the identity map on a local
group is automatically a continuous homomorphism; this gives the class of
local groups the structure of a category. Similarly, the class of local Lie
groups with their smooth homomorphisms is also a category.
Example 1.2.7. With the notation of Example 1.2.4, : U V is a
smooth homomorphism from the local Lie group G U to the local Lie group
G U . In fact, it is a smooth isomorphism, since 1 : V U provides
the inverse homomorphism.
Let us say that a word g1 . . . gn in a local group G is well-defined in
G (or well-defined, for short) if every possible way of associating this word
using parentheses is well-defined from applying the product operation. For
instance, in order for abcd to be well-defined, ((ab)c)d, (a(bc))d, (ab)(cd),
a(b(cd)), and a((bc)d) must all be well-defined. For instance, in the additive
local group {9, . . . , 9} (with the group structure restricted from that of
the integers Z), 2 + 6 + 5 is not well-defined because one of the ways of
associating this sum, namely 2 + (6 + 5), is not well-defined (even though
(2 + 6) + 5 is well-defined).
Exercise 1.2.1 (Iterating the associative law).
(i) Show that if a word g1 . . . gn in a local group G is well-defined, then
all ways of associating this word give the same answer, and so we
can uniquely evaluate g1 . . . gn as an element in G.
(ii) Give an example of a word g1 . . . gn in a local group G which has
two ways of being associated that are both well-defined, but give
different answers. (Hint: the local associativity axiom prevents
this from happening for n 3, so try n = 4. A small discrete local
group will already suffice to give a counterexample; verifying the
local group axioms are easier if one makes the domain of definition
of the group operations as small as one can get away with while
still having the counterexample.)
Exercise 1.2.2. Show that the number of ways to associate a word g1 . . . gn
is given by the Catalan number Cn1 := n1 2n2
n1 .
Exercise 1.2.3. Let G be a local group, and let m 1 be an integer. Show
that there exists a symmetric open neighbourhood Um of the identity such
that every word of length m in Um is well-defined in G (or more succinctly,
m is well-defined). (Note though that these words will usually only take
Um
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
27
values in G, rather than in Um , and also the sets Um tend to become smaller
as m increases.)
1.2.2. Some differential geometry. To define the Lie algebra of a Lie
group, we must first quickly recall some basic notions from differential geometry associated to smooth manifolds (which are not necessarily embedded
in some larger Euclidean space, but instead exist intrinsically as abstract
geometric structures). This requires a certain amount of abstract formalism
in order to define things rigorously, though for the purposes of visualisation,
it is more intuitive to view these concepts from a more informal geometric
perspective.
We begin with the concept of the tangent space and related structures.
Definition 1.2.8 (Tangent space). Let M be a smooth d-dimensional manifold. At every point x of this manifold, we can define the tangent space
Tx M of M at x. Formally, this tangent space can be defined as the space of
all continuously differentiable curves : I G defined on an open interval
I containing 0 with (0) = x, modulo the relation that two curves 1 , 2 are
considered equivalent if they have the same derivative at 0, in the sense that
d
d
(1 (t))|t=0 = (2 (t))|t=0
dt
dt
where : U V is a coordinate chart of G defined in a neighbourhood of
x; it is easy to see from the chain rule that this equivalence is independent
of the actual choice of . Using such a coordinate chart, one can identify
the tangent space Tx M with the Euclidean space Rd , by identifying with
d
dt ((t))|t=0 . One easily verifies that this gives Tx M the structure of a ddimensional vector space, in a manner which is independent of the choice
of coordinate chart . Elements of Tx M are called tangent vectors of M at
x. If : I G is a continuously differentiable curve with (0) = x, the
equivalence class of in Tx M will be denoted 0 (0).
S
The space T M := xM ({x} Tx M ) of pairs (x, v), where x is a point
in M and v is a tangent vector of M at x, is called the tangent bundle.
If : M N is a smooth map between two manifolds, we define the
derivative map D : T M T N to be the map defined by setting
D((x, 0 (0))) := ((x), ( )0 (0))
for all continously differentiable curves : I G with (0) = x for
some x M ; one can check that this map is well-defined. We also write
((x), D(x)(v)) for D(x, v), so that for each x M , D(x) is a map
from Tx M to T(x) N . One can easily verify that this latter map is linear.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
28
D( ) = (D) (D)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
29
(although, in most texts, the latter definition would be circular, because the
Lie derivative is usually defined using the directional derivative).
Remark 1.2.12. If V is an open subset of Rd , a smooth vector field on
V can be identified with a smooth map X : V Rd from V to Rd . If
X : M T M is a smooth vector field on M and : U V is a coordinate
chart of M , then the pushforward X := D X 1 : V T V of X by
is a smooth vector field of V . Thus, in coordinates, one can view vector
fields as maps from open subsets of Rd to Rd . This perspective is convenient
for quick and dirty calculations; for instance, in coordinates, the directional
derivative X f is the same as the familiar directional derivative X f from
several variable calculus. If however one wishes to perform several changes
of variable, then the more intrinsically geometric (and coordinate-free)
perspective outlined above can be more helpful.
There is a fundamental link between smooth vector fields and derivations
of C (M ):
Exercise 1.2.4 (Correspondence between smooth vector fields and derivations). Let M be a smooth manifold.
(i) If X (T M ) is a smooth vector field, show that X : C (M )
C (M ) is a derivation on the (real) algebra C (M ), i.e. a (real)
linear map that obeys the Leibniz rule
(1.7)
X (f g) = f X g + (X f )g
for all f, g C (M ).
(ii) Conversely, if d : C (M ) C (M ) is a derivation on C (M ),
show that there exists a unique smooth vector field X such that
d = X .
We see from the above exercise that smooth vector fields can be interpreted as a purely algebraic construction associated to the real algebra
C (M ), namely as the space of derivations on that vector space. This can
be useful for analysing the algebraic structure of such vector fields. Indeed,
we have the following basic algebraic observation:
Exercise 1.2.5 (Commutator of derivations is a derivation). Let d1 , d2 :
A A be two derivations on an algebra A. Show that the commutator
[d1 , d2 ] := d1 d2 d2 d1 is also a derivation on A.
From the preceding two exercises, we can define the Lie bracket [X, Y ]
of two vector fields X, Y (T M ) by the formula
[X,Y ] := [X , Y ].
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
30
for all X, Y, Z V .
Exercise 1.2.6. If M is a smooth manifold, show that (T M ) (equipped
with the Lie bracket) is a Lie algebra.
Remark 1.2.14. This is the abstract definition of a Lie algebra. A more
concrete definition would be to let V be a subspace of an algebra of operators,
and to define the Lie bracket as the commutator. The relation between the
two notions of a Lie algebra is explored in Section 2.3.
1.2.3. The Lie algebra of a Lie group. Let G be a (global) Lie group.
By definition, G is then a smooth manifolds, so we can thus define the
tangent bundle T G and smooth vector fields X (T G) as in the preceding
section. In particular, we can define the tangent space T1 G of G at the
identity element 1.
If g G, then the left multiplication operation left
g : x 7 gx is, by definition of a Lie group, a smooth map from G to G. This creates a derivative
map Dleft
g : T G T G from the tangent bundle T G to itself. We say that
a vector field X (T G) is left-invariant if one has (left
g ) X = X for all
left
left
g G, or equivalently if (Dg ) X = X g for all g G.
Exercise 1.2.7. Let G be a (global) Lie group.
(i) Show that for every element x of T1 G there is a unique left-invariant
vector field X (T G) such that X(1) = x.
(ii) Show that the commutator [X, Y ] of two left-invariant vector fields
is again a left-invariant vector field.
From the above exercise, we can identify the tangent space T1 G with
the left-invariant vector fields on T G, and the Lie bracket structure on the
latter then induces a Lie bracket (which we also call [, ]) on T1 G. The vector
space T1 G together with this Lie bracket is then a (finite-dimensional) Lie
algebra, which we call the Lie algebra of the Lie group G, and we write as
g.
Remark 1.2.15. Informally, an element x of the Lie algebra g is associated
with an infinitesimal perturbation 1 + x + O(2 ) of the identity in the Lie
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
31
group G. This intuition can be formalised fairly easily in the case of matrix
Lie groups such as GLn (C); for more abstract Lie groups, one can still
formalise things using nonstandard analysis, but we will not do so here.
Exercise 1.2.8.
(i) Show that the Lie algebra gln (C) of the general linear group GLn (C)
can be identified with the space Mn (C) of n n complex matrices,
with the Lie bracket [A, B] := AB BA.
(ii) Describe the Lie algebra un (C) of the unitary group Un (C).
(iii) Describe the Lie algebra sun (C) of the special unitary group SUn (C).
(iv) Describe the Lie algebra on (R) of the orthogonal On (R).
(v) Describe the Lie algebra son (R) of the special orthogonal SOn (R).
1 R R
(vi) Describe the Lie algebra of the Heisenberg group 0 1 R.
0 0 1
Exercise 1.2.9. Let : G H be a smooth homomorphism between
(global) Lie groups. Show that the derivative map D(1G ) at the identity
element 1G is then a Lie algebra homomorphism from the Lie algebra g
of G to the Lie algebra h of H (thus this map is linear and preserves the
Lie bracket). (From this and the chain rule (1.6), we see that the map
7 D(1G ) creates a covariant functor from the category of Lie groups to
the category of Lie algebras.)
We have seen that every global Lie group gives rise to a Lie algebra.
One can also associate Lie algebras to local Lie groups as follows:
Exercise 1.2.10. Let G be a local Lie group. Let U be a symmetric neighbourhood of the identity in G. (It is not difficult to see that least one
such neighbourhood exists.) Call a vector field X (T U ) left-invariant
left is the leftif, for every g U , one has (left
g ) X(g) = X(g), where g
multiplication map x 7 gx, defined on the open set {x U : gx U }
(where we adopt the convention that gx U is shorthand for g x is
well-defined and lies in U ).
(i) Establish the analogue of Exercise 1.2.7 in this setting. Conclude
that one can give T1 G the structure of a Lie algebra, which is
independent of the choice of U .
(ii) Establish the analogue of Exercise 1.2.9 in this setting.
Remark 1.2.16. In the converse direction, it is also true that every finitedimensional Lie algebra can be associated to either a local or a global Lie
group; this is known as Lies third theorem. However, this theorem is somewhat tricky to prove (particularly if one wants to associate the Lie algebra
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
32
with a global Lie group), requiring the non-trivial algebraic tool of Ados
theorem (discussed in Section 2.3); see Exercise 1.2.23 below.
1.2.4. The exponential map. The exponential map x 7 exp(x) on the
reals R (or its extension to the complex numbers C) is of course fundamental
to modern analysis. It can be defined in a variety of ways, such as the
following:
(i) exp : R R is the differentiable map obeying the ODE
exp(x) and the initial condition exp(0) = 1.
d
dx
exp(x) =
(ii) exp : R R is the differentiable map obeying the homomorphism property exp(x+y) = exp(x) exp(y) and the initial condition
d
dx exp(x)|x=0 = 1.
(iii) exp : R R is the limit of the functions x 7 (1 + nx )n as n .
P
xn
(iv) exp : R R is the limit of the infinite series x 7
n=0 n! .
We will need to generalise this map to arbitrary Lie algebras and Lie
groups. In the case of matrix Lie groups (and matrix Lie algebras), one can
use the matrix exponential, which can be defined efficiently by modifying
definition (iv) above, and which was already discussed in Section 1.1. It
is however difficult to use this definition for abstract Lie algebras and Lie
groups. The definition based on (ii) will ultimately be the best one to use
for the purposes of this text, but for foundational purposes (i) or (iii) is
initially easier to work with. In most of the foundational literature on Lie
groups and Lie algebras, one uses (i), in which case the existence and basic
properties of the exponential map can be provided by the Picard existence
theorem from the theory of ordinary differential equations. However, we
will use (iii), because it relies less heavily on the smooth structure of the Lie
group, and will therefore be more aligned with the spirit of Hilberts fifth
problem (which seeks to minimise the reliance of smoothness hypotheses
whenever possible). Actually, for minor technical reasons it is slightly more
n
convenient to work with the limit of (1 + 2xn )2 rather than (1 + nx )n .
We turn to the details. It will be convenient to work in local coordinates,
and for applications to Hilberts fifth problem it will be useful to forget
almost all of the smooth structure. We make the following definition:
Definition 1.2.17 (C 1,1 local group). A C 1,1 local group is a local group
V that is an open neighbourhood of the origin 0 in a Euclidean space Rd ,
with group identity 0, and whose group operation obeys the estimate
(1.9)
x y = x + y + O(|x||y|)
for all sufficiently small x, y, where the implied constant in the O() notation
can depend on V but is uniform in x, y.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
33
Example 1.2.18. Let G be a local Lie group of some dimension d, and let
: U V be a smooth coordinate chart that maps a neighbourhood U
of the group identity 1 to a neighbourhood V of the origin 0 in Rd , with
(1) = 0. Then, as explained in Example 1.2.4, V = (V, ) = G U is a
local Lie group with identity 0; in particular, one has
0 x = x 0 = x.
From Taylor expansion (using the smoothness of ) we thus have (1.9) for
sufficiently small x, y. Thus we see that every local Lie group generates a
C 1,1 local group when viewed in coordinates.
Remark 1.2.19. In real analysis, a (locally) C 1,1 function is a function
f : U Rm on a domain U Rn which is continuously differentiable (i.e.
in the regularity class C 1 ), and whose first derivatives f are (locally) Lipschitz (i.e. in the regularity class C 0,1 ) the C 1,1 regularity class is slightly
weaker (i.e. larger) than the class C 2 of twice continuously differentiable
functions, but much stronger than the class C 1 of singly continuously differentiable functions. See [Ta2010, 1.14] for more on these sorts of regularity
classes. The reason for the terminology C 1,1 in the above definition is that
C 1,1 regularity is essentially the minimal regularity for which one has the
Taylor expansion
f (x) = f (x0 ) + f (x0 ) (x x0 ) + O(|x x0 |2 )
for any x0 in the domain of f , and any x sufficiently close to x0 ; note that
the asymptotic (1.9) is of this form.
We now estimate various expressions in a C 1,1 local group.
Exercise 1.2.11. Let V be a C 1,1 local group. Throughout this exercise,
the implied constants in the O() notation can depend on V , but not on
parameters such as x, y, , k, n.
(i) Show that there exists an > 0 such that one has
X
(1.10)
x1 xk = x1 + + xk + O
|xi ||xj |
1i<jk
P
whenever k 1 and x1 , . . . , xk V are such that ki=1 |xi | ,
and the implied constant is uniform in k. Here and in the sequel we
adopt the convention that a statement such as (1.10) is automatically false unless all expressions in that statement are well-defined.
(Hint: induct on k using (1.9). It is best to replace the asymptotic
O() notation by explicit constants C in order to ensure that such
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
34
k
X
!
|xi yi |
i=1
Pk
i=1 |xi |,
Pk
j=1 |yi |
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
35
We can now define the exponential map exp : V 0 V on this C 1,1 local
group by defining
2n
1
(1.11)
exp(x) := lim
x
n 2n
for any x in a sufficiently small neighbourhood V 0 of the origin in V .
Exercise 1.2.12. Let V be a local C 1,1 group.
(i) Show that if V 0 is a sufficiently small neighbourhood of the origin
in V , then the limit in (1.11) exists for all x V 0 . (Hint: use the
n
previous exercise to estimate the distance between ( 21n x)2 and
n+1
1
x)2 .) Establish the additional estimate
( 2n+1
exp(x) = x + O(|x|2 ).
(1.12)
(iii) Show that for all sufficiently small x, y, one has the bilipschitz property
1
|(exp(x) exp(y)) (x y)| |x y|.
2
0
Conclude in particular that for V sufficiently small, exp is a homeomorphism between V 0 and an open neighbourhood exp(V 0 ) of the
origin. (Hint: To show that exp(V 0 ) contains a neighbourhood of
the origin, use (1.12) and the contraction mapping theorem.)
(iv) Show that
(1.13)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
36
Let us say that a C 1,1 local group is radially homogeneous if one has
sx tx = (s + t)x
(1.16)
(1.18)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
37
for all x g and t R. If x and t are sufficiently small, and one uses a
coordinate chart near the origin, the function f (t, x) := (exp(tx)) then
satisfies an ODE of the form
d
f (t, x) = F (f (t, x), x)
dt
for some smooth function F , with initial condition f (0, x) = 0; thus by the
fundamental theorem of calculus we have
Z t
(1.19)
f (t, x) =
F (f (t0 , x), x) dt0 .
0
(exp(t
1 X1 + + td Xd )) := (t1 , . . . , td )
for sufficiently small t1 , . . . , td R. These are known as exponential coordinates of the first kind. Although we will not use them much here,
we also note that there are exponential coordinates of the second kind, in
which the expression exp(t1 X1 + + td Xd ) is replaced by the slight variant
exp(t1 X1 ) . . . exp(td Xd ).
Using exponential coordinates of the first kind, we see that we may identify a local piece U of the Lie group G with the radially homogeneous C 1,1
local group V . In the next section, we will analyse such radially homogeneous C 1,1 groups further. For now, let us record some easy consequences of
the existence of exponential coordinates. Define a one-parameter subgroup
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
38
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
39
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
40
1.2.5. The Baker-Campbell-Hausdorff formula. We now study radially homogeneous C 1,1 local groups in more detail, in particular filling in
some of the last few steps in the program in Figure 1. We will show
Theorem 1.2.24 (Baker-Campbell-Hausdorff formula, qualitative version).
Let V Rd be a radially homogeneous C 1,1 local group. Then the group
operation is real analytic near the origin. In particular, after restricting
V to a sufficiently small neighbourhood of the origin, one obtains a local Lie
group.
We will in fact give a more precise formula for , known as the BakerCampbell-Haudorff-Dynkin formula, in the course of proving Theorem 1.2.24.
This formula is usually proven just for Lie groups, but it turns out that
the proof of the formula extends without much difficulty to the C 1,1 local
group setting (the main difference being that continuous operations, such
as Riemann integrals, have to be replaced by discrete counterparts, such as
Riemann sums).
Remark 1.2.25. In the case where V comes from viewing a general linear
group GLn (C) in local exponential coordinates, the group operation is
given by x y = log(exp(x) exp(y)) for sufficiently small x, y Mn (C).
Thus, a corollary of Theorem 1.2.24 is that this map is real analytic.
We begin the proof of Theorem 1.2.24. Throughout this section, V Rd
is a fixed radially homogeneous C 1,1 local group. We will need some variants
of the basic bound (1.9).
Exercise 1.2.18 (Lipschitz bounds). If x, y, z V are sufficiently small,
establish the bounds
(1.20)
x y = x + y + O(|x + y||y|)
(1.21)
x y = x + y + O(|x + y||x|)
(1.22)
x y = x z + O(|y z|)
and
(1.23)
y x = z x + O(|y z|).
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
41
Remark 1.2.27. Using the matrix example from Remark 1.2.25, we are
asserting here that
exp(x) exp(y) exp(x) = exp(Adx (y))
for some linear transform Adx (y) of y, and all sufficiently small x, y. Indeed,
using the basic matrix identity exp(AxA1 ) = A exp(x)A1 for invertible A
(coming from the fact that the conjugation map x 7 AxA1 is a continuous
ring homomorphism) we see that we may take Ad(x) = exp(x)y exp(x)
here.
Proof. Fix x. The map y 7 x y (x) is continuous near the origin, so
it will suffice to establish additivity, in the sense that
x (y + z) (x) = (x y (x)) + (x z (x))
for y, z sufficiently close to the origin.
Let n be a large natural number. Then from (1.16) we have
1
1 n
(y + z) =
y+ z
.
n
n
Conjugating this by x, we see that
n
1
1
x (y + z) (x) = x
y + z (x)
n
n
1
1
y + z (x) .
=n x
n
n
But from (1.9) we have
1
1
1
1
y+ z = y z+O
n
n
n
n
1
n2
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
42
for all x, y sufficiently small. Combining these two properties (and using
(1.20)) we conclude in particular that
(1.25)
adx y = [x, y]
z log z
.
z1
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
43
(1.27)
1 exp( adx )
z + O(|z|2 );
adx
inverting
1 exp( adx )
Adx 1
=
adx
Adx log Adx
we obtain the claim.
It remains to verify (1.27). Let n be a large natural number. We can
expand the left-hand side of (1.27) as a telescoping series
n1
X j + 1 j + 1
j+1
j
j
j
(1.28)
x
x+
z x
x+ z .
n
n
n
n
n
n
j=0
x+ z .
n
n
n n
n
n
From (1.20) one has nx nx + nz = nz + O( n|z|2 ), so by (1.22), (1.23) we
can write the preceding expression as
z
j
j
|z|
j
x+ z +O
x
n
n
n
n
n2
which by definition of Ad can be rewritten as
z
j
j
j
|z|
(1.29)
Ad j x
x
x+ z +O
.
n
n
n
n
n
n2
From (1.20) one has
j
j
j
x
x + z = O(|z|)
n
n
n
while from (1.25) one has Ad j x nz = O(|z|/n), hence from (1.9) we can
n
rewrite (1.29) as
2
z
j
j
j
|z|
|z|
Ad j x + x
x+ z +O
+O
.
n
n
n
n
n
n
n2
Inserting this back into (1.28), we can thus write the left-hand side of (1.27)
as
n1
X
z
|z|
2
Ad j x + O(|z| ) + O
.
n
n
n
j=0
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
44
Writing Ad j x = exp nj adx , and then letting n , we conclude
n
(from the convergence of the Riemann sum to the Riemann integral) that
Z 1
exp(t adx )z dt + O(|z|2 )
(x) (x + z) =
0
Remark 1.2.29. In the matrix case, the key computation is to show that
exp(x) exp(x + z) = 1 +
1 exp( adx )
z + O(|z|2 ).
adx
To see this, we can use the fundamental theorem of calculus to write the
left-hand side as
Z 1
d
1+
(exp(tx) exp(t(x + z))) dt.
0 dt
d
d
Since dt
exp(tx) = exp(tx)(x) and dt
exp(t(x + z)) = (x + z) exp(t(x +
z)), we can rewrite this as
Z 1
1+
exp(tx)z exp(t(x + z)) dt.
0
since exp(tx)z exp(tx) = exp(t adx )z, we obtain the desired claim.
We can integrate the above formula to obtain an exact formula for :
Corollary 1.2.30 (Baker-Campbell-Hausdorff-Dynkin formula). For x, y
sufficiently small, one has
Z 1
xy =x+
F (Adx Adty )y dt.
0
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
45
+ ...
2
3
4
to obtain the explicit expansion
F (z) = 1
xy =x+
X
(1)m
n=0
n+1
(ady )r1 (adx )s1 . . . (ady )rn (adx )sn
y
r1 !s1 ! . . . rn !sn !(r1 + + rn + 1)
X
ri ,si 0
X
(1)n
n=0
n+1
X
ri ,si 0
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
46
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
47
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
48
group (which is the same concept as a C 1,1 local group, but without the
asymptotic (1.9)). Thus we have reduced Hilberts fifth problem to the task
of boosting C 0 regularity to C 1,1 regularity, rather than that of boosting C 0
regularity to C regularity.
Exercise 1.2.24. Let G be a Lie group with Lie algebra g. For any X, Y
g, show that
2
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
49
Theorem 1.3.2 (Cartans theorem). If H is a (topologically) closed subgroup of a Lie group G, then H is a smooth submanifold of G, and is thus
also a Lie group.
Note that the hypothesis that H is closed is essential; for instance, the
rationals Q are a subgroup of the (additive) group of reals R, but the former
is not a Lie group even though the latter is.
Exercise 1.3.1. Let H be a subgroup of a locally compact group G. Show
that H is closed in G 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 1.3.3. If V is a finite-dimensional vector space, and W is another
vector space with an injective linear transformation : W V from W to
V , then W is also a finite-dimensional vector space.
Here is the non-linear version:
Theorem 1.3.4 (von Neumanns theorem). If G is a Lie group, and H is a
locally compact group with an injective continuous homomorphism : H
G, then H 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 Neumanns theorem localises to
the case when H is a local group rather a group. The requirement that H be
locally compact is necessary, for much the same reason that the requirement
that H be closed was necessary in Cartans theorem.
Example 1.3.5. Let G = (R/Z)2 be the two-dimensional torus, let H = R,
and let : H G be the map (x) := (x, x), where R 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 1.3.4 is
consistent with the fact that H is a Lie group. On the other hand, note that
when is irrational, then (H) is not closed; and so Theorem 1.3.4 does not
follow immediately from Theorem 1.3.2 in this case. (We will see, though,
that Theorem 1.3.4 follows from a local version of Theorem 1.3.2.)
As a corollary of Theorem 1.3.4, we observe that any locally compact
Hausdorff group H with a faithful linear representation, i.e. a continuous
injective homomorphism from H into a linear group such as GLn (R) or
GLn (C), is necessarily a Lie group. This suggests a representation-theoretic
approach to Hilberts 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
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
50
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 sections.
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 1.3.6. Let V be a locally compact Hausdorff topological vector
space. Then V is isomorphic (as a topological vector space) to Rd for some
finite d.
Remark 1.3.7. The Banach-Alaoglu theorem asserts that in a normed vector space V , the closed unit ball in the dual space V is always compact in the
weak-* topology. Of course, this dual space V 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 GleasonYamabe theorem, which we are not yet ready to prove in this section. However, by using methods similar to that used to prove Cartans theorem and
von Neumanns 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 1.3.8. Let G be a topological group. A Gleason metric on G is
a left-invariant metric d : G G R+ which generates the topology on G
and obeys the following properties for some constant C > 0, writing kgk for
d(g, id):
(1) (Escape property) If g G and n 1 is such that nkgk
kg n k C1 nkgk.
1
C,
then
1
C,
Exercise 1.3.2. Let G be a topological group that contains a neighbourhood of the identity isomorphic to a C 1,1 local group. Show that G admits
at least one Gleason metric.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
51
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
52
Since H is a group, we see that (exp(tX/2n ) exp(tY /2n ))2 lies in H. Since
H is closed, we conclude that exp(t(X + Y )) H for all t R, which
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
53
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
54
the identity. Thus we may write exp(Xn ) = exp(Yn ) exp(Zn ) for sufficiently
large n, where Yn h and Zn k both go to zero as n . Since Xn 6 V ,
we see that Zn is non-zero for n sufficiently large.
We arbitrarily place a norm on k. As before, we may pass to a subsequence and assume that Zn /kZn k converges to some limit in the unit
sphere of k; in particular, 6 h.
Since exp(Xn ) and exp(Yn ) both lie in H, exp(Zn ) does also. By arguing
as in the proof of Lemma 1.3.13 we conclude that exp(t) lies in H for all
t R, and so h, yielding the desired contradiction.
From the above lemma we see that H locally agrees with exp(V ) near the
identity, and thus locally agrees with exp(V )h near h for every h H. This
implies that H is a smooth submanifold of G; since it is also a topological
group, it is thus a Lie group. This establishes Cartans theorem.
Remark 1.3.15. Observe a posteriori that h is the Lie algebra of H, and
in particular is closed with respect to Lie brackets. This fact can also be
established directly using Exercise 1.2.24.
There is a local version of Cartans theorem, in which groups are replaced
by local groups:
Theorem 1.3.16 (Local Cartans theorem). If H is a locally compact local
subgroup of a local Lie group G, then there is an open neighbourhood H 0 of
the identity in H that is a smooth submanifold of G, and is thus also a local
Lie group.
The proof of this theorem follows the lines of the global Cartans theorem, with some minor technical changes, and we set this proof out in the
following exercise.
Exercise 1.3.4. Define a local one-parameter subgroup of a local group H
to be a continuous homomorphism : (, ) H from the (additive)
local group (, ) to H. Call two local one-parameter subgroups equivalent
if they agree on a neighbourhood of the origin, and let L(H) be the set
of all equivalence classes of local one-parameter subgroups. Establish the
following claims:
(i) If H is a global group, then there is a canonical one-to-one correspondence that identifies this definition of L(H) with the definition
of L(H) given previously.
(ii) In the situation of Theorem 1.3.16, show that L(H) can be identified with a linear subspace h of g, namely
h := {X g : exp(tX) H for all sufficiently small t}.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
55
(iii) Let the notation and assumptions be as in (ii). For any neighbourhood H 0 of the identity in H, there is a neighbourhood V of the
origin in h such that exp(V ) H 0 .
(iv) Let the notation and assumptions be as in (ii). There exists a
neighbourhood U of the identity in H, and a neighbourhood V of
the origin in h, such that exp : V U is a homeomorphism.
(v) Prove Theorem 1.3.16.
One can then use Theorem 1.3.16 to establish von Neumanns theorem,
as follows. Suppose that H is a locally compact group with an injective
continuous homomorphism : H G into a Lie group G. As H is locally
compact, there is an open neighbourhood U of the origin in H whose closure
U is compact. The map from U to (U ) is a continuous bijection from a
compact set to a Hausdorff set, and is therefore a homeomorphism (since it
maps closed (and hence compact) subsets of U to compact (and hence closed)
subsets of (U )). The set (U ) is then a locally compact local subgroup of
G and thus has a neighbourhood of the identity which is a local Lie group,
by Theorem 1.3.16. Pulling this back by , we see that some neighbourhood
of the identity in H is a local Lie group, and thus H is a global Lie group
by Exercise 1.2.17.
Exercise 1.3.5. State and prove a local version of von Neumanns theorem,
in which G and H are local groups rather than global groups, and the global
injectivity condition is similarly replaced by local injectivity.
1.3.2. Locally compact vector spaces. We will now turn to the study
of topological vector spaces, which we will need to establish Theorem 1.3.9.
We begin by recalling the definition of a topological vector space.
Definition 1.3.17 (Topological vector space). A topological vector space
is a (real) vector space V equipped with a topology that makes the vector
space operations + : V V V and : R V R (jointly) continuous.
(In particular, (V, +) is necessarily a topological group.)
One can also consider complex topological vector spaces, but the theory
for such spaces is almost identical to the real case, and we will only need the
real case for what follows. In the literature, it is often common to restrict
attention to Hausdorff topological vector spaces, although this is not a severe
restriction in practice, as the following exercise shows:
Exercise 1.3.6. Let V be a topological vector space. Show that the closure
W := {0} of the origin is a closed subspace of V , and the quotient space
V /W is a Hausdorff topological vector space. Furthermore, show that a set
is open in V if and only if it is the preimage of an open set in V /W under
the quotient map : V V /W .
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
56
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
57
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
58
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
59
From the commutator estimate (1.30) and the triangle inequality we also
obtain a conjugation estimate
kghg 1 k khk
whenever kgk, khk . Since left-invariance gives
d(g, h) = kg 1 hk
we then conclude an approximate right invariance
d(gk, hk) d(g, h)
whenever kgk, khk, kkk . In a similar spirit, the commutator estimate
(1.30) also gives
d(gh, hg) kgkkhk
(1.31)
whenever kgk, khk .
This has the following useful consequence, which asserts that the power
maps g 7 g n behave like dilations:
Lemma 1.3.23. If n 1 and kgk, khk /n, then
d(g n hn , (gh)n ) n2 kgkkhk
and
d(g n , hn ) nd(g, h).
Proof. We begin with the first inequality. By the triangle inequality, it
suffices to show that
(1.32)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
60
uniformly for all 0 i < n. By left-invariance and approximate rightinvariance, the left-hand side is comparable to
d(g ni1 h, hg ni1 ),
which by (1.31) is bounded above by
kg ni1 kkhk nkgkkhk
as required.
Now we prove the second estimate. Write g = hk, then kkk = d(g, h)
2/n. We have
d(hn k n , hn ) = kk n k nkkk
thanks to the escape property (shrinking if necessary). On the other hand,
from the first inequality, we have
d(g n , hn k n ) n2 khkkkk.
If is small enough, the claim now follows from the triangle inequality.
Remark 1.3.24. Lemma 1.3.23 implies (by a standard covering argument)
that the group G is locally of bounded doubling, though we will not use
this fact here. The bounds above should be compared with the bounds in
Exercise 1.2.11. Indeed, just as the bounds in that exercise were used in the
previous sections to build the exponential map for Lie groups, the bounds
in Lemma 1.3.23 are crucial for controlling the exponential function on the
locally compact group G equipped with the Gleason metric d.
Now we bring in the space L(G) of one-parameter subgroups. We give
this space the compact-open topology, thus the topology is generated by balls
of the form
{ L(G) : sup d((t), 0 (t)) < r}
tI
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
61
(1.33)
cf. Exercise 1.2.13. In view of this, we would like to define the sum +
of two one-parameter subgroups , L(G) by the formula
( + )(t) := lim ((t/n)(t/n))n .
n
Observe from continuity of multiplication that to prove this claim for a given
t, it suffices to do so for t/2; thus we may assume without loss of generality
that t is small.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
62
= lim ((t/n)(t/n))an
n
and similarly for ( + )(bt) and ( + )((a + b)t), whence the claim. To
prove the second claim, we see that
( + )(t)1 = lim ((t/n)(t/n))n
n
= lim ((t/n)(t/n))n ,
n
((t/n)(t/n))n
but
is ((t/n)(t/n))n conjugated by (t/n), which goes
to the identity; and the claim follows.
L(G) also has an obvious zero element, namely the trivial one-parameter
subgroup t 7 id.
Lemma 1.3.27. L(G) is a topological vector space.
Proof. We first show that L(G) is a vector space. It is clear that the zero
element 0 of L(G) is an additive and scalar multiplication identity, and
that scalar multiplication is associative. To show that addition is commutative, we again use the observation that ((t/n)(t/n))n is ((t/n)(t/n))n
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
63
exists and defines a norm on L(G) that generates the topology on L(G).
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
64
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
65
kknNn k
khn k.
Nn
Nn
Nn
Since exp(n ) = hn , we conclude from the triangle inequality and leftinvariance that
1
1
1
d exp n +
, gn
khn k + d kn , exp
.
Nn
Nn
Nn
But from Lemma 1.3.23 again, one has
1
1
d kn , exp
d knNn , exp() = o(1/Nn )
Nn
Nn
9The author thanks Lou van den Dries and Isaac Goldbring for bringing this argument to
his attention.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
66
and thus
1
d exp n +
, gn = o(1/Nn ).
Nn
But for n large enough, n + N1n lies in K, and so the distance from gn to
K is o(1/Nn ) = o(d(gn , hn )). But this contradicts the minimality of hn for
n large enough, and the claim follows.
If K is a sufficiently small compact neighbourhood of the identity in
L(G), then exp : K exp(K) is bijective by Lemma 1.3.9; since it is also
continuous, K is compact, and exp(K) is Hausdorff, we conclude that exp :
K exp(K) is a homeomorphism. The local group structure G exp(K) on
exp(K) then pulls back to a local group structure on K.
Exercise 1.3.10. If we identify L(G) with Rd for some d, show that the
exponential map exp : K exp(K) is bilipschitz.
Proposition 1.3.30. K is a radially homogeneous C 1,1 local group (as defined in Definition 1.2.17 and (1.16)), after identifying L(G) with Rd for
some finite d.
Proof. The radial homogeneity is clear from (1.33) and the homomorphism
property, so the main task is to establish the C 1,1 property
x y = x + y + O(|x||y|)
for the local group law on K. By Exercise 1.3.10, this is equivalent to the
assertion that
d((1)(1), ( + )(1)) k(1)kk(1)k
for , sufficiently close to the identity in L(G). By definition of + , it
suffices to show that
d((1)(1), ((1/n)(1/n))n ) k(1)kk(1)k
for all n; but this follows from Lemma 1.3.23 (and the observation, from the
escape property, that k(1/n)k k(1)k/n and k(1/n)k k(1)k/n).
Combining this proposition with Lemma 1.2.34, we obtain Theorem
1.3.9.
Exercise 1.3.11. State and prove a version of Theorem 1.3.9 for local
groups. (In order to do this, you must first decide how to define an analogue
of a Gleason metric on a local group.)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
67
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
68
Fourier analysis on general compact abelian groups. With this and some
additional (largely combinatorial) arguments, we will also be able to obtain
satisfactory structural control on locally compact abelian groups as well.
The link between Haar measure and useful metrics on G is a little more
complicated. Firstly, once one has the regular representation : G
U (L2 (G, d)), and given a suitable test function : G C, one can then
embed G into L2 (G, d) (or into other function spaces on G, such as Cc (G)
or L (G)) by mapping a group element g G to the translate (g) of
in that function space. (This map might not actually be an embedding if
enjoys a non-trivial translation symmetry (g) = , but let us ignore this
possibility for now.) One can then pull the metric structure on the function
space back to a metric on G, for instance defining an L2 (G, d)-based metric
d(g, h) := k (g) (h)kL2 (G,d)
if is square-integrable, or perhaps a Cc (G)-based metric
(1.34)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
69
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
70
Exercise 1.4.2. Let G be a locally compact group. Show that there exists
an open subgroup G0 which is locally compact and -compact. (Hint: take
the group generated by a compact neighbourhood of the identity.)
Exercise 1.4.3. Let G be a locally compact group. Let H = {id} be the
topological closure of the identity element.
(i) Show that given any open neighbourhood U of a point x in G,
there exists a neighbourhood V of x whose closure lies in U . (Hint:
translate x to the identity and select V so that V 2 U .) In other
words, G is a regular space.
(ii) Show that for any group element g G, that the sets gH and H
are either equal or disjoint.
(iii) Show that H is a compact normal subgroup of G.
(iv) Show that the quotient group G/H (equipped with the quotient
topology) is a locally compact Hausdorff group.
(v) Show that a subset of G is open if and only if it is the preimage of
an open set in G/H.
Now that we have restricted attention to the -compact Hausdorff case,
we can now define the notion of a Haar measure.
Definition 1.4.1 (Radon measure). Let X be a -compact locally compact
Hausdorff topological space. The Borel -algebra B[X] on X is the -algebra
generated by the open subsets of X. A Borel measure is a countably additive
non-negative measure : B[X] [0, +] on the Borel -algebra. A Radon
measure is a Borel measure obeying three additional axioms:
(i) (Local finiteness) One has (K) < for every compact set K.
(ii) (Inner regularity) One has (E) = supKE,K compact (K) for
every Borel measurable set E.
(iii) (Outer regularity) One has (E) = inf U E,U open (U ) for every
Borel measurable set E.
Definition 1.4.2 (Haar measure). Let G = (G, ) be a -compact locally
compact Hausdorff group. A Radon measure is left-invariant (resp. rightinvariant) if one has (gE) = (E) (resp. (Eg) = (E)) for all g G
and Borel measurable sets E. A left-invariant Haar measure is a non-zero
Radon measure which is left-invariant; a right-invariant Haar measure is
defined similarly. A bi-invariant Haar measure is a Haar measure which is
both left-invariant and right-invariant.
Note that we do not consider the zero measure to be a Haar measure.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
71
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
72
where the implied constant in the O() notation can depend on , , f, g but
not on . But by the left-invariance of , the left-hand side is also
Z Z
f (y) (x1 y) d(y)d(x)
G
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
73
which simplifies to I (f )
G
1
G (x )
Z
I (f ) = I (f )
and similarly
Z
I (g) = I (g)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
74
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
75
and thus
Z X
K
ck (yk )(x) d =
G k=1
K
X
ck
k=1
k=1
of a given function f Cc
(G)+
by , we have
Z
f d [f : ].
G
[f : ] [f : f0 ][f0 : ]
we also obtain the uniform bound axiom, and from the infimal nature of
[f : ] we also easily obtain the subadditivity property
I(f + g) I(f ) + I(g).
To finish the construction, it thus suffices to show that
I(fi + fj ) I(fi ) + I(fj )
for each 1 i, j n, if > 0 is chosen sufficiently small depending on
, f0 , f1 , . . . , fn .
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
76
(1.36)
K
X
ck (yk )(x)
k=1
(1.37)
k=1
[f0 : ].
ck I(fi + fj ) +
2
fi (yk ) +
fi (yk ) + fj (yk ) + 2
c00k := ck
fj (yk ) +
fi (yk ) + fj (yk ) + 2
and
fi (x)
(1.38)
k=1
and
fj (x)
(1.39)
K
X
k=1
K
X
k=1
k=1
ck (yk1 x)
fi (yk ) +
.
fi (yk ) + fj (yk ) + 2
(yk1 x)
If
is non-zero, then by the construction of and U , one has |fi (yk )
fi (x)| and |fj (yk ) fj (x)| , which implies that
fi (yk ) +
fi (x)
=
.
fi (yk ) + fj (yk ) + 2
fi (x) + fj (x) + 4
Using (1.36) we thus have
K
X
k=1
fi (x)
(fi (x) + fj (x)) + 4
fi (x) + fj (x) + 4
which gives (1.38); a similar argument gives (1.39). From the subadditivity
(and monotonicity) of I, we conclude that
PK 0
c
I(fi ) k=1 k + 4I(g)
[f0 : ]
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
77
and
PK
I(fj )
00
k=1 ck
[f0 : ]
+ 4I(g)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
78
Z
G
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
79
(E) =
(E)
(ii) Verify the cocycle equation gh (x) = g (x)h (gx) for all g, h, x G.
(iii) Show that the measure defined by
Z
(E) :=
x (id)1 d(x)
E
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
80
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
81
f (x)g(x) dx.
G
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
82
g(x1 ) = g(x)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
83
kT kop =
sup
|hT x, xi|
W :kxk1
valid for all self-adjoint operators T (see Exercise 1.4.16 below). Thus, we
may find a sequence xn of vectors of norm at most 1 such that
hT xn , xn i
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
84
(1.42)
applying T we conclude that
T (T xn ) T xn 0.
By compactness of T , we may pass to a subsequence so that T xn converges
to a limit y, and thus T y y = 0. As T has no eigenvectors, y must be
trivial; but then hT xn , xn i converges to zero, a contradiction.
Exercise 1.4.16. Establish (1.42) whenever T : W W is a bounded
self-adjoint operator on W . (Hint: Bound |hT x, yi| by the right-hand side
of (1.41) whenever x, y are vectors of norm at most 1, by playing with
hT (ax + by), (ax + by)i for various choices of scalars a, b, in the spirit of the
proof of the Cauchy-Schwarz inequality.)
This leads to the consequence that we can find non-trivial finite-dimensional
representations on at least a single non-identity element:
Theorem 1.4.12 (Baby Peter-Weyl theorem). Let G be a compact Hausdorff group with Haar measure , and let y G be a non-identity element
of G. Then there exists a finite-dimensional invariant subspace of L2 (G) on
which (y) is not the identity.
Proof. Suppose for contradiction that (y) is the identity on every finitedimensional invariant subspace of L2 (G), thus (y)1 annihilates every such
subspace. By Theorem 1.4.11, we conclude that (y) 1 has range in the
kernel of every convolution operator Tg with g L2 , thus Tg ( (y) 1)f = 0
for any f, g L2 (G) with g obeying (1.40), i.e.
(y)(f g) = (f g)
for any such f, g. But one may easily construct f, g such that f g is nonzero at the identity and vanishing at y (e.g. one can set f = g = 1U where
U is an open symmetric neighbourhood of the identity, small enough that y
lies outside U 2 ). This gives the desired contradiction.
Remark 1.4.13. The full Peter-Weyl theorem describes rather precisely
all the invariant subspaces of L2 (G). Roughly speaking, the theorem asserts
that for each irreducible finite-dimensional representation : G U (V )
of G, dim(V ) different copies of V (viewed as an invariant G-space) appear
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
85
in L2 (G), and that they are all orthogonal and make up all of L2 (G); thus,
one has an orthogonal decomposition
M dim(V )
L2 (G)
V
of G-spaces. Actually, this is not the sharpest form of the theorem, as it only
describes the left G-action and not the right G-action; see Section 2.8 for a
precise statement and proof of the Peter-Weyl theorem in its strongest form.
This form is of importance in Fourier analysis and representation theory, but
in this text we will only need the baby form of the theorem (Theorem 1.4.12),
which is an easy consequence of the full Peter-Weyl theorem (since, if g is
not the identity, then (g) is clearly non-trivial on L2 (G) and hence on at
least one of the V factors).
The Peter-Weyl theorem leads to the following structural theorem for
compact groups:
Theorem 1.4.14 (Gleason-Yamabe theorem for compact groups). Let G
be a compact Hausdorff group, and let U be a neighbourhood of the identity.
Then there exists a compact normal subgroup H of G contained in U such
that G/H is isomorphic to a linear group (i.e. a closed subgroup of a general
linear group GLn (C)).
Note from Cartans theorem (Theorem 1.3.2) that every linear group is
Lie; thus, compact Hausdorff groups are almost Lie in some sense.
Proof. Let g be an element of G\U . By the baby Peter-Weyl theorem, we
can find a finite-dimensional invariant subspace V of L2 (G) on which (g)
is non-trivial. Identifying such a subspace with Cn for some finite n, we
thus have a continuous homomorphism : G GLn (C) such that (g)
is non-trivial. By continuity, (g) will also be non-trivial for some open
neighbourhood of g. Using the compactness of G\U , one can then find
a finite number 1 , . . . , k of such continuous homomorphisms i : G
GLni (C) such that for each g G\U , at least one of 1 (g), . . . , k (g) is
non-trivial. If we then form the direct sum
k
k
M
M
:=
i : G
GLni (C) GLn1 ++nk (C)
i=1
i=1
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
86
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
87
Hausdorff group is metrisable, show that one can take the inverse limit to
be indexed instead by the natural numbers with the usual ordering.
Exercise 1.4.22. Let G be an abelian group with a homomorphism :
G 7 U (V ) into the unitary group of a finite-dimensional space V . Show
that V can be decomposed as the vector space sum of one-dimensional Ginvariant spaces. (Hint: By the spectral theorem for unitary matrices, any
unitary operator T on V decomposes V into eigenspaces, and any operator
commuting with T must preserve each of these eigenspaces. Now induct on
the dimension of V .)
Exercise 1.4.23 (Fourier analysis on compact abelian groups). Let G be a
compact abelian Hausdorff group with Haar probability measure . Define
a character to be a continuous homomorphism : G 7 S 1 to the unit circle
be the collection of all such characters.
S 1 := {z C : |z| = 1}, and let G
(i) Show that for every g G not equal to the identity, there exists a
character such that (g) 6= 1. (Hint: combine the baby PeterWeyl theorem with the preceding exercise.)
(ii) Show that every function in C(G) is the limit in the uniform topology of finite linear combinations of characters. (Hint: use the
Stone-Weierstrass theorem.)
form an orthonormal basis
(iii) Show that the characters for G
2
of L (G, d).
1.4.3. The structure of locally compact abelian groups. We now use
the above machinery to analyse locally compact abelian groups. We follow
some combinatorial arguments of Pontryagin, as presented in the text of
Montgomery and Zippin [MoZi1974].
We first make a general observation that locally compact groups contain
open subgroups that are finitely generated modulo a compact set. Call
a subgroup of a topological group G cocompact if the quotient space is
compact.
Lemma 1.4.16. Let G be a locally compact group. Then there exists an
open subgroup G0 of G which has a cocompact finitely generated subgroup .
Proof. Let K be a compact neighbourhood of the identity. Then K 2 is also
compact and can thus be covered by finitely many copies of K, thus
K 2 KS
for some finite set S, which we may assume without loss of generality to be
contained in K 1 K 2 . In particular, if is the group generated by S, then
K 2 K.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
88
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
89
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
90
Exercise 1.4.26. Show that every locally compact abelian Hausdorff group
is isomorphic to the inverse limit of abelian Lie groups.
Thus, in principle at least, the study of locally compact abelian group is
reduced to that of abelian Lie groups, which are more or less easy to classify:
Exercise 1.4.27.
(i) Show that every discrete subgroup of Rd is iso0
morphic to Zd for some 0 d0 d.
(ii) Show that every connected abelian Lie group G is isomorphic to
0
Rd (R/Z)d for some natural numbers d, d0 . (Hint: first show
that the kernel of the exponential map is a discrete subgroup of the
Lie algebra.) Conclude in particular the divisibility property that
if g G and n 1 then there exists h G with hn = g.
(iii) Show that every compact abelian Lie group G is isomorphic to
(R/Z)d H for some natural number d and a H which is a finite
product of finite cyclic groups. (You may need the classification of
finitely generated abelian groups, and will also need the divisibility
property to lift a certain finite group from a certain quotient space
back to G.)
(iv) Show that every abelian Lie group contains an open subgroup that
0
00
is isomorphic to Rd (R/Z)d Zd H for some natural numbers
d, d0 , d00 and a finite product H of finite cyclic groups.
Remark 1.4.20. Despite the quite explicit description of (most) abelian
Lie groups, some interesting behaviour can still occur in locally compact
abelian groups after taking inverse limits; consider for instance the solenoid
example (Exercise 1.1.7).
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
91
Finally, we will also need the compact case of the Gleason-Yamabe theorem
(Theorem 1.4.14), which was proven via the Peter-Weyl theorem.
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 1.5.1 (Building metrics out of test functions). Let G be a topological group, and let : G R+ be a bounded non-negative function.
Then we define the pseudometric d : G G R+ by the formula
d (g, h) := sup | (g)(x) (h)(x)|
xG
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
92
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
93
Vn = {id}.
n=1
Using the continuity of the group operations, we can recursively find a sequence of nested open neighbourhoods of the identity
(1.43)
such that each U1/2n is symmetric (i.e. g U1/2n if and only if g 1 U1/2n ),
is contained in Vn , and is such that U1/2n+1 U1/2n+1 U1/2n for each n 0.
In particular the U1/2n are also a neighbourhood base of the identity with
(1.44)
U1/2n = {id}.
n=1
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
94
For every dyadic rational a/2n in (0, 1), we can now define the open sets
Ua/2n by setting
Ua/2n := U1/2nk U1/2n1
where a/2n = 2n1 + + 2nk is the binary expansion of a/2n with 1
n1 < < nk . By repeated use of the hypothesis U1/2n+1 U1/2n+1 U1/2n
we see that the Ua/2n are increasing in a/2n ; indeed, we have the inclusion
(1.45)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
95
kg n k
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
96
(1.49)
(1.50)
thus
(h )(y)(y 1 x) d(y).
h (x) =
G
We would like to similarly move the g operator over to the second factor,
but we run into a difficulty due to the non-abelian nature of G. Nevertheless,
we can still do this provided that we twist that operator by a conjugation.
More precisely, we have
Z
(1.51)
g h (x) = (h )(y)(gy )(y 1 x) d(y)
G
gy
y 1 gy
:=
where
is g conjugated by y. If h B(0, ), the integrand is only
non-zero when y B(0, 2). Applying (1.47), we obtain the bound
kg h kCc (g) khk
sup
kg y k.
yB(0,2)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
97
This implies that and gn have overlapping support, and hence g n lies in
B(0, 4). By the escape property (1.46), this implies (if is small enough)
that kgk n1 , and the claim follows.
Combining Claim 2 with (1.49) we see that
kg h kCc (G) kgk khk
whenever kgk , khk are small enough. Now we use the identity
k[g, h]k = k ([g, h]) kCc (G)
= k (g) (h) (h) (g)kCc (G)
= kg h h g kCc (G)
and the triangle inequality to conclude that
k[g, h]k kgk khk
whenever kgk , khk are small enough. Theorem 1.5.5 then follows from
Claim 1 and Claim 2.
1.5.3. Building metrics on NSS groups. We will now be able to build
metrics on groups using a set of hypotheses that do not explicitly involve
any metric at all. The key hypothesis will be the no small subgroups (NSS)
property:
Definition 1.5.6 (No small subgroups). A topological group G has the no
small subgroups (or NSS) property if there exists an open neighbourhood
U of the identity which does not contain any subgroup of G other than the
trivial group.
Exercise 1.5.6. Show that any Lie group is NSS.
Exercise 1.5.7. Show that any group with a weak Gleason metric is NSS.
For an example of a group which is not NSS, consider the infinitedimensional torus (R/Z)N . From the definition of the product topology,
we see that any neighbourhood of the identity in this torus contains an
infinite-dimensional subtorus, and so this group is not NSS.
Exercise 1.5.8. Show that for any prime p, the p-adic groups Zp and Qp
are not NSS. What about the solenoid group R Zp /Z ?
Exercise 1.5.9. Show that an NSS group is automatically Hausdorff. (Hint:
use Exercise 1.4.3.)
Exercise 1.5.10. Show that an NSS locally compact group is automatically
metrisable. (Hint: use Exercise 1.5.4.)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
98
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
99
n
X
kgi ke,U0
i=1
kg1 . . . gn ke,U0 M
n
X
kgi ke,U0
i=1
for all n, g1 , . . . , gn and some huge constant M ; we will then deduce the same
estimate with a smaller value of M . Afterwards we will show how to remove
the hypothesis (1.54).
Now suppose we have (1.54) for some M . Motivated by the argument
in the previous section, we now try to convolve together two Lipschitz
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
100
functions. For this, we will need some metric-like functions. Define the
modified escape norm kgk,U0 by the formula
)
( n
X
kgk,U0 := inf
kgi ke,U0 : g = g1 . . . gn
i=1
where the infimum is over all possible ways to split g as a finite product of
group elements. From (1.54), we have
1
kgke,U0 kgk,U0 kgke,U0
M
and we have the triangle inequality
(1.55)
|g (x)| M kgke,U0
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
101
The functions , need not be continuous, but they are compactly supported, bounded, and Borel measurable, and so one can still form their convolution := , which will then be continuous and compactly supported;
indeed, is supported in U04 .
We have a lower bound on how big is, since
(0) (U0 ) 1
(where we allow implied constants to depend on , U0 , but remain independent of L, U1 , or M ). This gives us a way to compare kk with kke,U0 .
Indeed, if nkgk < (0), then (as in the proof of Claim 1 in the previous
section) we have g n U08 ; this implies that
kgke,U08 kgk
for all g G, and hence by (1.53) we have
kgke,U0 kgk
(1.58)
(1.59)
M kgke,U0
thanks to (1.56). But we can do better than this, as follows. For any
g, h G, we have the analogue of (1.51), namely
Z
g h (x) = (h )(y)(gy )(y 1 x) d(y)
G
gn = ng +
n1
X
g g i
i=0
for any g G and natural number n, and thus by the triangle inequality
!
n1
X
n
(1.60)
kg k = nkgk + O
kg gi kCc (G) .
i=0
We conclude that
M
kg k = nkgk + O n kgke,U0
L
n
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
102
n
X
kgi k .
i=1
n
X
(kgi ke,U0 + )
i=1
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
103
elements with zero escape norm, and hence no non-identity elements with
zero modified escape norm either; thus d,U0 is a genuine metric.
We now claim that d,U0 generates the topology of G. Given the leftinvariance of d,U0 , it suffices to establish two things: firstly, that any open
neighbourhood of the identity contains a ball around the identity in the
d,U0 metric; and conversely, any such ball contains an open neighbourhood
around the identity.
To prove the first claim, let U be an open neighbourhood around the
identity, and let U 0 U be a smaller neighbourhood of the identity. From
(1.53) we see (if U 0 is small enough) that kk,U0 is comparable to kke,U 0 , and
U 0 contains a small ball around the origin in the d,U0 metric, giving the
claim. To prove the second claim, consider a ball B(0, r) in the d,U0 metric.
For any positive integer m, we can find an open neighbourhood Um of the
1
m U , and hence kgk
identity such that Um
0
e,U0 m for all g Um . For m
large enough, this implies that Um B(0, r), and the claim follows.
To finish the proof of Theorem 1.5.7, we need to verify the escape property (1.46). Thus, we need to show that if g G, n 1 are such that
nkgk,U0 is sufficiently small, then we have kg n k,U0 nkgk,U0 . We may of
course assume that g is not the identity, as the claim is trivial otherwise. As
kk,U0 is comparable to kke,U0 , we know that there exists a natural number
m 1/kgk,U0 such that g m 6 U0 .
Let U1 be a neighbourhood of the identity small enough that U12 U0 .
We have kg i k,U0 nkgk,U0 for all i = 1, . . . , n, so g i U1 and hence
m > n. Let m + i be the first multiple of n larger than n, then i n and so
g i U1 . Since g m 6 U0 , this implies g m+i 6 U1 . Since m + i is divisible by
n
nkgk,U0 , and the claim follows from
n, we conclude that kg n ke,U1 m+i
(1.53).
1.5.4. NSS from subgroup trapping. In view of Theorem 1.5.7, the only
remaining task in the proof of the Gleason-Yamabe theorem is to locate big
subquotients G0 /H of a locally compact group G with the NSS property. We
will need some further notation. Given a neighbourhood V of the identity
in a topological group G, let Q[V ] denote the union of all the subgroups of
G that are contained in V . Thus, a group is NSS if Q[V ] is trivial for all
sufficiently small V .
We will need a property that is weaker than NSS:
Definition 1.5.10 (Subgroup trapping). A topological group has the subgroup trapping property if, for every open neighbourhood U of the identity,
there exists another open neighbourhood V of the identity such that Q[V ]
generates a subgroup hQ[V ]i contained in U .
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
104
Clearly, every NSS group has the subgroup trapping property. Informally, groups with the latter property do have small subgroups, but one
cannot get very far away from the origin just by combining together such
subgroups.
Example 1.5.11. The infinite-dimensional torus (R/Z)N does not have the
NSS property, but it does have the subgroup trapping property.
It is difficult to produce an example of a group that does not have the
subgroup trapping property; the reason for this will be made clear in the
next section. For now, we establish the following key result (another arrow
of Figure 1):
Proposition 1.5.12 (From subgroup trapping to NSS). Let G be a locally
compact group with the subgroup trapping property, and let U be an open
neighbourhood of the identity in G. Then there exists an open subgroup G0 of
G, and a compact subgroup N of G0 contained in U , such that G0 /N is locally
compact and NSS. In particular, by Theorem 1.5.7, G0 /N is isomorphic to
a Lie group.
Intuitively, the idea is to use the subgroup trapping property to find
a small compact normal subgroup N that contains Q[V ] for some small
V , and then quotient this group out to get an NSS group. Unfortunately,
because N is not necessarily contained in V , this quotienting operation
may create some additional small subgroups. To fix this, we need to pass
from the compact subgroup N to a smaller one. In order to understand
the subgroups of compact groups, the main tool will be Gleason-Yamabe
theorem for compact groups (Theorem 1.4.14).
For us, the main reason why we need the compact case of the GleasonYamabe theorem is that Lie groups automatically have the NSS property,
even though G need not. Thus, one can view Theorem 1.4.14 as giving the
compact case of Proposition 1.5.12.
We now prove Proposition 1.5.12, using an argument of Yamabe [Ya1953].
Let G be a locally compact group with the subgroup trapping property, and
let U be an open neighbourhood of the identity. We may find a smaller
neighbourhood U1 of the identity with U12 U , which in particular implies that U1 U ; by shrinking U1 if necessary, we may assume that U1 is
compact. By the subgroup trapping property, one can find an open neighbourhood U2 of the identity such that hQ(U2 )i is contained in U1 , and thus
H := hQ(U2 )i is a compact subgroup of G contained in U . By shrinking U2
if necessary we may assume U2 U1 .
Ideally, if H were normal and contained in U2 , then the quotient group
hU2 i/H would have the NSS property. Unfortunately H need not be normal,
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
105
and need not be contained in U2 , but we can fix this as follows. Applying
Theorem 1.4.14, we can find a compact normal subgroup N of H contained
in U2 H such that H/N is isomorphic to a Lie group, and in particular
is NSS. In particular, we can find an open symmetric neighbourhood U3
of the identity in G such that U3 N U3 U2 and that the quotient space
(U3 N U3 H) has no non-trivial subgroups in H/N , where : H H/N
is the quotient map.
We now claim that N is normalised by U3 . Indeed, if g U3 , then the
conjugate N g := g 1 N g of N is contained in U3 N U3 and hence in U2 . As
N g is a group, it must thus be contained in Q(U2 ) and hence in H. But
then (N g ) is a subgroup of H/N that is contained in (U3 N U3 H), and
is hence trivial by construction. Thus N g N , and so N is normalised by
U3 . If we then let G0 be the subgroup of G generated by N and U3 , we see
that G0 is an open subgroup of G, with N a compact normal subgroup of
G0 .
To finish the job, we need to show that G0 /N has the NSS property. It
suffices to show that U3 N U3 /N has no nontrivial subgroups. But any subgroup in U3 N U3 /N pulls back to a subgroup in U3 N U3 , hence in U2 , hence
in Q(U2 ), hence in H; since (U3 N U3 H)/N has no nontrivial subgroups,
the claim follows. This concludes the proof of Proposition 1.5.12.
1.5.5. The subgroup trapping property. In view of Theorem 1.5.7,
Proposition 1.5.12, and Exercise 1.5.4, we see that the Gleason-Yamabe
theorem (Theorem 1.1.17) now reduces to the following claim.
Proposition 1.5.13. Every locally compact metrisable group has the subgroup trapping property.
This proposition represents the final two arrows of Figure 1.
We now prove Proposition 1.5.13, which is the hardest step of the entire
proof and uses almost all the tools already developed. In particular, it
requires both Theorem 1.4.14 and Gleasons convolution trick, as well as
some of the basic theory of Hausdorff distance; as such, this is perhaps the
most infinitary of all the steps in the argument.
The Gleason-type arguments can be encapsulated in the following proposition, which is a weak version of the subgroup trapping property:
Proposition 1.5.14 (Finite trapping). Let G be a locally compact group,
let U be an open precompact neighbourhood of the identity, and let m 1
be an integer. Then there exists an open neighbourhood V of the identity
with the following property: if Q Q[V ] is a symmetric set containing the
identity, and n 1 is such that Qn U , then Qmn U 8 .
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
106
for all g V U .
Now let := . Then is supported on U 4 and kkCc (G) 1 (where
implied constants can depend on U , ). As before, we conclude that g U 8
whenever kgk is sufficiently small.
Now suppose that q Q[V ]; we will estimate kqk . From (1.60) one has
kqk
1 n
kq k + sup kqi q kCc (G)
n
0in
(note that qi and q commute). For the first term, we can compute
kq n k = sup |qn ( )(x)|
x
and
Z
qn ( )(x) =
(y)(qn )y (y 1 x)d(y).
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
107
1
.
Mn
yF
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
108
E of U . Since the Q[Vi ] are symmetric and contain the identity, E is also
symmetric and contains the identity. For any fixed m, we have Q[Vi ]mni
U 8 for all sufficiently large i, which on taking Hausdorff limits implies that
E m U 8 . In particular, the group H := hEi is a compact subgroup of G
contained in U 8 .
Let U1 be a small neighbourhood of the identity in G to be chosen later.
By Theorem 1.4.14, we can find a normal subgroup N of H contained in
U1 H such that H/N is NSS. Let B be a neigbourhood of the identity in
H/N so small that B 10 has no small subgroups. A compactness argument
then shows that there exists a natural number k such that for any g H/N
that is not in B, at least one of g, . . . , g k must lie outside of B 10 .
Now let > 0 be a small parameter. Since Q[Vi ]ni +1 6 U , we see
that Q[Vi ]ni +1 does not lie in the -neighbourhood 1 (B) of 1 (B) if is
small enough, where : H H/N is the projection map. Let n0i be the first
0
integer for which Q[Vi ]ni does not lie in 1 (B) , then n0i ni + 1 and n0i
0
as i (for fixed ). On the other hand, as Q[Vi ]ni 1 1 (B) , we see
0
from another application of Proposition 1.5.14 that Q[Vi ]kni ( 1 (B) )8
if i is sufficiently large depending on .
On the other hand, since Q[Vi ]ni converges to a subset of H in the
Hausdorff distance, we know that for i large enough, Q[Vi ]2ni and hence
0
Q[Vi ]ni is contained in the -neighbourhood of H. Thus we can find an
0
element gi of Q[Vi ]ni that lies within of a group element hi of H, but
does not lie in 1 (B) ; thus hi lies inside H\ 1 (B). By construction
of B, we can find 1 ji k such that hji i lies in H\ 1 (B 10 ). But hji i
0
also lies within o(1) of giji , which lies in Q[Vi ]kni and hence in ( 1 (B) )8 ,
where o(1) denotes a quantity depending on that goes to zero as 0.
We conclude that H\ 1 (B 10 ) and 1 (B 8 ) are separated by o(1), which
leads to a contradiction if is sufficiently small (note that 1 (B 8 ) and
H\ 1 (B 10 ) are compact and disjoint, and hence separated by a positive
distance), and the claim follows.
Exercise 1.5.14. Let X be a compact metric space, Kc (X) denote the
space of non-empty closed and connected subsets of X. Show that Kc (X)
with the Hausdorff metric is also a compact metric space.
Exercise 1.5.15. Show that the metrisability condition in Proposition 1.5.13
can be dropped; in other words, show that every locally compact group has
the subgroup trapping property.
1.5.6. The local group case. In [Go2009], [Go2010], [vdDrGo2010],
the above theory was extended to the setting of local groups. In fact, there
is relatively little difficulty (other than some notational difficulties) in doing
so, because the analysis in the previous sections can be made to take place
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
109
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
110
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
111
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
112
Theorem 1.5.24 (Local Gleason-Yamabe theorem). Let G be a locally compact local group. Then there exists an open symmetric neighbourhood G0 of
the identity, and a compact global group H in G0 that is normalised by G0 ,
such that G0 /H is well-defined and isomorphic to a local Lie group.
The proofs of this theorem by Goldbring and Goldbring-van den Dries
were phrased in the language of nonstandard analysis. However, it is possible
to translate those arguments to standard analysis arguments, which closely
follow the arguments given in previous sections11. We briefly sketch the
main points here.
As in the global case, the route to obtaining (local) Lie structure is via
Gleason metrics. On a local group G, we define a local Gleason metric to be
a metric d : U U R+ defined on some symmetric open neighbourhood
U of the identity with (say) U 100 well-defined (to avoid technical issues),
which generates the topology of U , and which obeys the following version
of the left-invariance, escape and commutator properties:
(1) (Left-invariance) If g, h, k U are such that gh, gk U , then
d(h, k) = d(gh, gk).
(2) (Escape property) If g U and nkgk
defined in U and kg n k C1 nkgk.
1
C,
1
C,
One can then verify (by localisation of the arguments in Section 1.3)
that any locally compact local Lie group with a local Gleason metric is locally Lie (i.e. some neighbourhood of the identity is isomorphic to a local
Lie group); see Exercise 1.3.11. Next, one can define the notion of a weak
local Gleason metric by dropping the commutator estimate, and one can
verify an analogue of Theorem 1.5.5, namely that any weak local Gleason
metric is automatically a local Gleason metric, after possibly shrinking the
neighbourhood U and adjusting the constant C as necessary. The proof of
this statement is essentially the same as that in Theorem 1.5.5 (which is
already localised to small neighbourhoods of the identity), but uses a local
Haar measure instead of a global Haar measure, and requires some preliminary shrinking of the neighbourhood U to ensure that all group-theoretic
operations (and convolutions) are well-defined. We omit the (rather tedious)
details.
Now we define the concept of an NSS local group as a local group which
has an open neighbourhood of the identity that contains no non-trivial global
11Actually, our arguments are not a verbatim translation of those in Goldbring and
Goldbring-van den Dries, as we have made a few simplifications in which the role of Gleason
metrics is much more strongly emphasised.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
113
subgroups. The proof of Theorem 1.5.7 is already localised to small neighbourhoods of the identity, and it is possible (after being sufficiently careful
with the notation) to translate that argument to the local setting, and conclude that any NSS local group admits a weak Gleason metric on some open
neighbourhood of the identity, and is hence locally Lie. (A typical example
of being sufficiently careful with the notation: to define the escape norm
(1.52), one adopts the convention that a statement such as g, . . . , g n U
is automatically false if g, . . . , g n are not all well-defined. The induction
hypothesis (1.54) will play a key role in ensuring that all expressions involved are well-defined and localised to a suitably small neighbourhood of
the identity.) Again, we omit the details.
The next step is to obtain a local version of Proposition 1.5.12. Here we
encounter a slight difficulty because in a general local group G, we do not
have a good notion of the group hAi generated by a set A of generators in G.
As such, the subgroup trapping property does not automatically translate
to the local group setting as defined in Definition 1.5.10. However, this
difficulty can be easily avoided by rewording the definition:
Definition 1.5.25 (Subgroup trapping). A local group has the subgroup
trapping property if, for every open neighbourhood U of the identity, there
exists another open neighbourhood V of the identity such that Q[V ] is contained in a global subgroup H that is in turn contained in U . (Here, Q[V ]
is, as before, the union of all the global subgroups contained in V .)
Because Q[V ] is now contained in a global group H, the group hQ[V ]i
generated by H is well-defined. As H is in the open neighbourhood U , one
can then also form the closure hQ[V ]i; if we choose U small enough to be
precompact, then this is a compact global group (and thus describable by the
Gleason-Yamabe theorem for such groups, Theorem 1.4.14). Because of this,
it is possible to adapt Proposition 1.5.12 without much difficulty to the local
setting to conclude that given any locally compact local group G with the
subgroup trapping property, there exists an open symmetric neighbourhood
G0 of the identity, and a compact global group H in G0 that is normalised
by G0 , such that G0 /H is well-defined and NSS (and thus locally isomorphic
to a local Lie group).
Finally, to finish the proof of Theorem 1.5.24, one has to establish the
analogue of Proposition 1.5.13, namely that one has to show that every
locally compact metrisable local group has the subgroup trapping property.
(It is not difficult to adapt Exercise 1.5.4 to the local group setting to reduce
to the metrisable case.) The first step is to prove the local group analogue of
Proposition 1.5.14 (again adopting the obvious convention that a statement
such as Qn U is only considered true if Qn is well-defined, and adding the
additional hypothesis that U is sufficiently small in order to ensure that all
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
114
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
115
1 Qp Qp Qp
0 1 Qp Qp
,
G=
0 0
1 Qp
0 0
0
1
where Qp is a p-adic group, show that one cannot demand in Theorem 1.6.1
that the open subgroup G0 be normal.
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 1.6.2 (Brouwer invariance of domain theorem). Let U be an open
subset of Rn , and let f : U Rn be a continuous injective map. Then f (U )
is also open.
We prove this theorem later in this section also. It has an important
corollary:
Corollary 1.6.3 (Topological invariance of dimension). If n > m, and U
is a non-empty open subset of Rn , then there is no continuous injective
mapping from U to Rm . In particular, Rn and Rm are not homeomorphic.
Exercise 1.6.2 (Uniqueness of dimension). Let X be a non-empty topological space. If X is a manifold of dimension d1 , and also a manifold of
dimension d2 , show that d1 = d2 . Thus, we may define the dimension
dim(X) of a non-empty manifold in a well-defined manner.
If X, Y are non-empty manifolds, and there is a continuous injection
from X to Y , show that dim(X) dim(Y ).
Remark 1.6.4. Note that the analogue of the above exercise for surjections is false: the existence of a continuous surjection from one non-empty
manifold X to another Y does not imply that dim(X) dim(Y ), 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 1.6.3 to bound the dimension of
the Lie groups Ln in an inverse limit G = limn Ln by the dimension
of the inverse limit G. Among other things, this can be used to obtain a
positive resolution to Hilberts fifth problem, Theorem 1.1.13. Again, this
theorem will be proven later in this section.
Another application of this machinery is the following variant of Hilberts
fifth problem, which was used in Gromovs original proof of Gromovs theorem on groups of polynomial growth, although we will not actually need it
here:
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
116
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
117
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
118
Proof. Let K be the intersection of all the clopen sets that contain x (note
that X is obviously clopen). Clearly K is closed and contains x. Our objective is to show that K consists solely of {x}. As X is totally disconnected,
it will suffice to show that K is connected.
Suppose this is not the case, then we can split K = K1 K2 where
K1 , K2 are disjoint non-empty closed sets; without loss of generality, we may
assume that x lies in K1 . As all compact Hausdorff spaces are normal, we can
thus enclose K1 , K2 in disjoint open subsets U1 , U2 of X. In particular, the
topological boundary U2 is compact and lies outside of K. By definition of
K, we thus see that for every y U2 , we can find a clopen neighbourhood of
x that avoids y; by compactness of U2 (and the fact that finite intersections
of clopen sets are clopen), we can thus find a clopen neighbourhood L of x
that is disjoint from U2 . One then verifies that L\U2 = L\U2 is a clopen
neighbourhood of x that is disjoint from K2 , contradicting the definition of
K, and the claim follows.
Now we can prove van Dantzigs theorem. We will use an argument
from [HeRo1979]. Let G be totally disconnected locally compact (and thus
Hausdorff). Then we can find a compact neighbourhood K of the identity.
By Lemma 1.6.9, for every y K, we can find a clopen neighbourhood
of the identity that avoids y; by compactness of K, we may thus find a
clopen neighbourhood of the identity that avoids K. By intersecting this
neighbourhood with K, we may thus find a compact clopen neighbourhood F
of the identity. As F is both compact and open, we may then the continuity
of the group operations find a symmetric neighbourhood U of the identity
such that U F F . In particular, if we let G0 be the group generated by U ,
then G0 is an open subgroup of G contained in F and is thus compact as
required.
Remark 1.6.10. 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 Dantzigs 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 (i.e. the inverse limit
of finite groups). The global behaviour however remains more complicated,
in part because the compact open subgroup given by van Dantzigs theorem
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
119
need not be normal, and so does not necessarily induce a splitting of G into
compact and discrete factors.
Example 1.6.11. Let p be a prime, and let G be the semi-direct product
Z n Qp , where the integers Z act on Qp by the map m : x 7 pm x, and we
give G the product of the discrete topology of Z and the p-adic topology
on Qp . One easily verifies that G is a totally disconnected locally compact
group. It certainly has compact open subgroups, such as {0}Zp . However,
it is easy to show that G has no non-trivial compact normal subgroups (the
problem is that the conjugation action of Z on Qp has all non-trivial orbits
unbounded).
We can pull van Dantzigs theorem back to more general locally compact
groups:
Exercise 1.6.5. Let G be a locally compact group.
(i) Show that G contains an open subgroup G0 which is compact-byconnected in the sense that G0 /(G0 ) is compact. (Hint: apply
van Dantzigs theorem to G/G .)
(ii) If G is compact-by-connected, and U is an open neighbourhood of
the identity, show that there exists a compact subgroup K of G in U
such that G/K is isomorphic to a Lie group. (Hint: use Theorem
1.1.17, and observe that any open subgroup of the compact-byconnected group G has finite index and thus has only finitely many
conjugates.) Conclude Theorem 1.6.1.
(iii) Show that any locally compact Hausdorff group G contains an open
subgroup G0 that is isomorphic to an inverse limit of Lie groups
(L )A , which each Lie group L has at most finitely many connected components. Furthermore, each L is isomorphic to G0 /K
for some compact normal subgroup K of G0 , with K K for
< . If G is first countable, show that this inverse limit can be
taken to be a sequence (so that the index set A is simply the natural numbers N with the usual ordering), and the Kn then shrink to
zero in the sense that they lie inside any given open neighbourhood
of the identity for n large enough.
Exercise 1.6.6. Let G be a totally disconnected locally compact group.
Show that every compact subgroup K of G is contained in a compact open
subgroup. (Hint: van Dantzigs theorem provides a compact open subgroup,
but it need not contain K. But is there a way to modify it so that it is
normalised by K? Why would being normalised by K be useful?)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
120
has at
function such that kG(y) G(y)k
1 for all y f (B n ). Then G
n
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
121
(y) := max(
, 1)y.
kyk
Note that is continuous and well-defined since f (B n ) avoids zero. Informally, is a perturbation of f (B n ) caused by pushing f (B n ) out a small
distance away from the origin 0 (and hence also away from f (0)), with
being the pushing map.
By construction, G is non-zero on 1 ; since 1 is compact, G is bounded
from below on 1 by some > 0. By shrinking if necessary we may assume
that < 0.1.
By the Weierstrass approximation theorem, we can find a polynomial
P : Rn Rn such that
(1.65)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
122
:= P ((y)).
G(y)
This is a continuous function that is never zero. From (1.65), (1.64) we have
kG(y) G(y)k
<
whenever y f (B n ) is such that kyk > . On the other hand, if kyk ,
then from (1.64), (1.63) we have
kG(y)k, kG((y))k 0.1
and hence by (1.65) and the triangle inequality
kG(y) G(y)k
0.2 + .
Thus in all cases we have
kG(y) G(y)k
0.2 + 0.3
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
123
a continuous map nn+1 : L(Ln ) L(Ln+1 ) from the one-parameter subgroups n : R Ln of Ln to the one-parameter subgroups n+1 : R Ln+1
of Ln+1 , such that nn+1 (n ) mod Hn = n for all n L(Ln ).
Exercise 1.6.8. By iterating these maps and passing to the inverse limit,
conclude that for each n N, there is a continuous map n : L(Ln ) L(G)
such that n (n ) mod Kn = n for all n L(Ln ).
Because Ln is a Lie group, the exponential map n 7 n (1) is a
homeomorphism from a neighbourhood of the origin in L(Ln ) to a neighbourhood of the identity in Ln . We can thus obtain a continuous map
n (1) 7 n (n )(1) from a neighbourhood of the identity in Ln to G. Since
n (n )(1) mod Kn = n (1), this map is injective.
Now we use the hypothesis that G is locally Euclidean (and in particular,
has a well-defined dimension dim(G)). By Exercise 1.6.2, we have
dim(Ln ) dim(G)
for all n. On the other hand, since each Ln is a quotient of the next Lie
group Ln+1 , one has
dim(Ln ) dim(Ln+1 ).
Since there are only finitely many possible values for the (necessarily integral) dimension dim(Ln ) between 0 and dim(G), we conclude that the
dimension must eventually stabilise, i.e. one has
dim(Ln ) = dim(Ln+1 )
for all sufficiently large n. By discarding the first few terms in the sequence
and relabeling, we may thus assume that the dimension is constant for all n.
Since Ln Ln+1 /Hn , this implies that the Lie groups Hn have dimension
zero for all n. As the Hn are also compact, they are thus finite. Thus each
Kn+1 is a finite extension of Kn . As Kn is the inverse limit of the Kn /Km
as m , we conclude that Kn is a profinite group, i.e. the inverse limit
of finite groups. In particular, Kn is totally disconnected.
We now study the short exact squence
0 Kn G Gn 0,
playing off the locally connected nature of the Lie group Gn against the
totally disconnected nature of Kn .
As discussed earlier, we have a continuous injective map n from a neighbourhood Un of the identity in Gn to G that partially inverts the quotient
map. By translation, we may normalise n (1) = 1. As Gn is locally connected, we can find a connected neighborhood Vn of the identity in Gn such
that (Vn Vn1 )2 Un .
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
124
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
125
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
126
finite. On the other hand, G/H is locally connected, which implies that
the Kn /Kn+1 are eventually trivial. Indeed, if we pick a simply connected
neighbbourhood U1 of the identity in G1 /H1 , then by local connectedness
of G/H, there exists a connected neighbourhood U of the identity in G/H
whose projection to G1 /H1 is contained in U1 . Being open, U must contain
one of the Kn . If Kn /Km is non-trivial for any m > n, then the projection of
U to Gm /Hm will then be disconnected (as this projection will be contained
in a neighbourhood with the topological structure of U K1 /Km , and its
intersection with the latter fibre is at least as large as Kn /Km . We conclude
that Kn is trivial for n large enough, and so G = Gn is a Lie group as
required.
Note that while the manifold X in Proposition 1.6.5 is initially only
required a priori to be a topological manifold, it automatically acquires a
smooth structure also:
Exercise 1.6.12. Let G and X be as in Proposition 1.6.5. Show that one
can endow X with the structure of a smooth manifold, such that the action
of G is also smooth. (Hint: apply Cartans theorem to the stabiliser of a
point in X.)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
127
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
128
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
129
Remark 1.7.3. Note that the above deduction of (ii) from (i) is ineffective
in that it gives no explicit bound on the uniform bound M in (ii). Without
any further information on how the qualitative bound (i) is proven, this
is the best one can do in general (and this is one of the most significant
weaknesses of infinitary methods when used to solve finitary problems); but
if one has access to the proof of (i), one can often finitise or proof mine that
argument to extract an effective bound for M , although often the bound
one obtains in the process is quite poor (particularly if the proof of (i) relied
extensively on infinitary tools, such as limits). See [Ta2008, 1.2] for some
related discussion.
The above simple example illustrates that in order to get from an infinitary statement such as (i) to a finitary statement such as (ii), a key step
is to be able to take a sequence (xn )nN (or in some cases, a more general
net (x )A ) of finitary objects and extract a suitable infinitary limit object
x. In the literature, there are three main ways in which one can extract such
a limit:
(1) (Topological limit) If the xn are all elements of some topological
space S (e.g. an incomplete function space) which has a suitable
compactification or completion X (e.g. a Banach space), then
(after passing to a subsequence if necessary) one can often ensure
the xn converge in a topological sense (or in a metrical sense) to
a limit x. The use of this type of limit to pass between quantitative/finitary and qualitative/infinitary results is particularly common in the more analytical areas of mathematics (such as ergodic
theory, asymptotic combinatorics, or PDE), due to the abundance
of useful compactness results in analysis such as the (sequential)
Banach-Alaoglu theorem, Prokhorovs theorem, the Helly selection
theorem, the Arzel
a-Ascoli theorem, or even the humble BolzanoWeierstrass theorem. However, one often has to take care with the
nature of convergence, as many compactness theorems only guarantee convergence in a weak sense rather than in a strong one.
(2) (Categorical limit) If the xn are all objects in some category (e.g.
metric spaces, groups, fields, etc.) with a number of morphisms between the xn (e.g. morphisms from xn+1 to xn , or vice versa), then
one can often form a direct limit lim xn or inverse limit lim xn of
these objects to form a limiting object x. The use of these types of
limits to connect quantitative and qualitative results is common in
subjects such as algebraic geometry that are particularly amenable
to categorical ways of thinking. (We already have seen inverse limits appear in the discussion of Hilberts fifth problem, although in
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
130
that context they were not really used to connect quantitative and
qualitative results together.)
(3) (Logical limit) If the xn are all distinct spaces (or elements or subsets of distinct spaces), with few morphisms connecting them together, then topological and categorical limits are often unavailable
or unhelpful. In such cases, however, one can still tie together such
objects using an ultraproduct construction (or similar
Q device) to
create a limiting object limn xn or limiting space n xn that
is a logical limit of the xn , in the sense that various properties of
the xn (particularly those that can be phrased using the language
of first-order logic) are preserved in the limit. As such, logical limits are often very well suited for the task of connecting finitary and
infinitary mathematics together. Ultralimit type constructions are
of course used extensively in logic (particularly in model theory),
but are also popular in metric geometry. They can also be used in
many of the previously mentioned areas of mathematics, such as
algebraic geometry (as discussed in [Ta2011b, 2.1]).
The three types of limits are analogous in many ways, with a number
of connections between them. For instance, in the study of groups of polynomial growth, both topological limits (using the metric notion of GromovHausdorff convergence) and logical limits (using the ultralimit construction)
are commonly used, and to some extent the two constructions are at least
partially interchangeable in this setting. (See also [Ta2011c, 4.4-4.5] for
the use of ultralimits as a substitute for topological limits.) In the theory of
approximate groups, though, it was observed by Hrushovski [Hr2012] that
logical limits (and in particular, ultraproducts) are the most useful type of
limit to connect finitary approximate groups to their infinitary counterparts.
One reason for this is that one is often interested in obtaining results on approximate groups A that are uniform in the choice of ambient group G.
As such, one often seeks to take a limit of approximate groups An that lie
in completely unrelated ambient groups Gn , with no obvious morphisms or
metrics tying the Gn to each other. As such, the topological and categorical
limits are not easily usable, whereas the logical limits can still be employed
without much difficulty.
Logical limits are closely tied with non-standard analysis. Indeed, by
applying an ultraproduct construction to standard number systems such as
the natural numbers N or the reals R, one can obtain nonstandard number
systems such as the nonstandard natural numbers N or the nonstandard
real numbers (or hyperreals) R. These nonstandard number systems behave
very similarly to their standard counterparts, but also enjoy the advantage
of containing the standard number systems as proper subsystems (e.g. R is
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
131
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
132
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
133
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
134
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
135
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
136
Given a sequence
(Xn )nN of standard spaces Xn U, we define the
Q
ultraproduct n Xn to be the space of all ultralimits limn xn , where
xn Xn for each n. Such spaces will be called nonstandard spaces (also
known as nonstandard sets or internal sets).
If X is a standard
Q space, we define the ultrapower X of X to be the
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
137
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
138
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
139
whenever xn Xn . One can easily verify that this does indeed define
a function, which we refer to as a nonstandard function (also called an
internal function in the literature). The ultralimit f := limn f of a
single standard function is thus a nonstandard function from X to Y that
extends f ; by abuse of notation we shall also refer to f as f (analogously
to how we identify limn x with x rather than giving it a different name
such as x).
Similarly, for a (standard) natural number k N, define a standard kary relation (or standard k-ary predicate) to be a standard function R : X1
Xk {true, false} from the product of k standard spaces to the Boolean
space {true, false} (which we will treat as part of the standard universe U).
By the preceding construction (and Exercise 1.7.7), the ultralimit limn Rn
of a sequence of standard k-ary relations Rn : X1,n Xk,n {true, false}
(with k independent of n) is another k-ary relation
lim Rn : lim X1,n lim Xk,n {true, false},
which we will call a nonstandard k-ary operator (also known as an internal k-ary operator ). Again, we identify each standard operator O with its
nonstandard extension O := limn O.
Example 1.7.17. Let a = limn an and b = limn bn be two nonstandard natural numbers (thus an , bn are sequences of standard natural numbers). Then, by definition,
(1) a + b is the nonstandard natural number a + b := limn an + bn .
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
140
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
141
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
142
(2) Let GQ
n = (Gn , n ) be a sequence of groups; then the ultraproduct
G := n Gn (with the ultralimit group operation := limn n ,
and similarly for the group identity and inverse operations) is also
a group, because the group axioms can be phrased in first-order
logic using the indicated structures. If, for each n, gn is an element
of Gn , then the ultralimit limn gn is a central element of G if and
only if gn is a central element of Gn for all n sufficiently close to .
Similarly, if for each n, Hn is a subset of Gn , then the ultraproduct
Q
n Hn is a normal subgroup of G if and only if Hn is a normal
subgroup of Gn for all n sufficiently close to .
(3) The standard natural numbers N obey the axiom of induction: if
P (n) is a predicate definable in the language of Peano arithmetic,
and P (0) is true, and P (n) implies P (n + 1) for all n N, then
P (n) is true for all n N. As a consequence, the same axiom of
induction also holds for the nonstandard natural numbers N. Note
however it is important for the nonstandard axiom of induction that
the predicate be definable in the language of Peano arithmetic. For
instance, the following argument is fallacious: 0 N, and for any
nonstandard natural number n, n N implies n + 1 N. Hence,
by induction, n N for all nonstandard natural numbers n. The
reason is that the predicate n N is not formalisable in the
language of Peano arithmetic.
(4) Exercise 1.7.9 can be interpreted as a special case of Corollary
1.7.19.
Exercise 1.7.14.
(i) Show that N (viewed as a subset of N) is not a nonstandard space.
(Hint: if it were, the fallacious induction mentioned earlier could
now be made valid.)
(ii) Establish the overspill principle: if a nonstandard predicate P (n)
on the nonstandard natural numbers is true for all standard natural
numbers, then it is true for at least one strictly nonstandard natural
number as well.
(iii) Show that if a nonstandard subset of N consists only of standard
natural numbers, then it is finite.
(iv) Show that the nonstandard natural numbers N are not well-ordered.
Why does this not contradict Loss theorem?
Exercise 1.7.15.
Q For each n, let GnQbe a group, and let Sn be a subset of
Gn . Write G := n Gn and S := n Sn .
(i) Give an example in which each Sn generates Gn as a group, but S
does not generate G as a group.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
143
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
144
and (x X : P (x1 , . . . , xk , x)) are predicates in x1 , . . . , xk . (Similarly for any permutation in the ordering of x1 , . . . , xk , x.)
In practice, of course, many of the parentheses in the above constructions
can be removed without causing any ambiguity, and one usually does so in
order to improve readability, for instance abbreviating ((P Q) R) as
P Q R. However, for the purposes of studying formal logic, it can be
convenient to insist on these parentheses.
Given a formal predicate involving L and one or more free variables
x1 , . . . , xn , we may quantify (or interpret) the predicate by associating each
formal object in L to an actual object, which may be a standard object,
a nonstandard object, or something else entirely, and similarly associating
an actual space, relation, or operator to each formal space, relation, or
operator, making sure that all the relevant domains and ranges are respected.
When one does so, the formal predicate P becomes a concrete predicate
P : X1 Xk {true, false} on the appropriate quantified spaces by
interpreting all the symbols in P in the obvious fashion. For instance, if the
primitive terms consist of a formal space X and a formal binary operation
: XX X, then P (x) := (y X : xy = yx) is a formal unary predicate,
and if one quantifies this predicate over a concrete group G = (G, ) (which
may be standard, nonstandard, or neither), then one obtains a concrete
unary predicate P : G {true, false}, with P (g) true precisely when g is a
central element of G.
Exercise 1.7.16. With the above definitions, prove Theorem 1.7.18.
Exercise 1.7.17 (Compactness theorem). Let L be an at most countable
collection of formal objects, spaces, operators and relations. Let S1 , S2 , S3 , . . .
be a sequence of sentences involving the symbols in L. Suppose that any
finite collection S1 , . . . , Sn of these sentences is satisfiable in the standard
universe, thus there exists an assignment of standard objects, spaces, operators, or relations to each element of L for which S1 , . . . , Sn are all true when
quantified over these assignments. Show that the entire collection S1 , S2 , . . .
are then satisfiable in the nonstandard universe, thus there exists an assignment of nonstandard objects. (This is the countable case of the compactness
theorem in logic. One can also use ultraproducts over larger sets than N to
prove the general case of the compactness theorem, but we will not do so
here.)
1.7.3. Nonstandard finite sets and nonstandard finite sums. It will
be convenient to extend the standard machinery of finite sets and finite sums
to the nonstandard
setting. Define a nonstandard finite set to be an ultraQ
product A = n An of finite sets, and a nonstandard finite sequence of
reals (xm )mA to be a nonstandard function m 7 xm from a nonstandard
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
145
finite set A to the nonstandard reals, or equivalently an ultralimit of standard finite sequences (xm,n )mAn of reals. We can define the (nonstandard)
cardinality |A| of a nonstandard finite set in the usual manner:
Y
An | := lim |An |.
|
n
Thus, for instance, if N is a nonstandard natural number, then the nonstandard finite set {n N : 1 n N } has nonstandard cardinality
N.
Similarly, if (xm )mA = limn (xm,n )mAn P
is a nonstandard finite sequence of reals, we define the nonstandard sum mA xm as
X
X
xm := lim
xm,n .
mA
mAn
Thus for instance mA 1 = |A|. By using Loss theorem, all the basic laws
of algebra for manipulating standard finite sums carry over to nonstandard
finite sums. For instance, we may interchange summations
X X
X X
X
xm1 ,m2 =
xm1 ,m2 =
xm1 ,m2
m1 A1 m2 A2
m2 A2 m1 A1
for any nonstandard finite sequence (xm1 ,m2 )(m1 ,m2 )A1 A2 of reals indexed
by a product set, simply because the same assertion is obviously true for
standard finite sequences.
1.7.4. Asymptotic notation. The ultrapower and ultralimit constructions are extremely general, being basically applicable to any collection of
standard objects, spaces, objects, and relations. However, when one applies
these constructions to standard number systems, such as the natural numbers N, the integers Z, the real numbers R, or the complex numbers C to
obtain nonstandard number systems such as N, Z, R, C, then one can
gain an additional tool of use in analysis, namely a clean asymptotic notation
which is closely related to standard asymptotic notation, but has a slightly
different arrangement of quantifiers that make the nonstandard asymptotic
notation better suited to algebraic constructions (such as the quotient space
construction) than standard asymptotic notation.
Definition 1.7.20 (Asymptotic notation). Let R be one of the standard
number systems (N, Z, Q, R, C). A nonstandard number x in R is said
to be bounded if one has |x| C for some standard real number C (or
equivalently, |x| n some standard natural number n), and unbounded
otherwise. If y is a non-negative nonstandard real, we write x = O(y) if
|x| Cy for some standard real number C, thus a nonstandard number is
bounded if and only if it is O(1). We say that x is infinitesimal if |x| for
every standard real number > 0 (or equivalently, if |x| n1 for all standard
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
146
positive natural numbers n), and write x = o(y) if |x| y for all standard
real numbers > 0, thus x is infinitesimal if and only if it is o(1).
Example 1.7.21. All standard numbers are bounded, and all non-zero standard numbers are non-infinitesimal. The nonstandard natural number N :=
limn n is unbounded, and the nonstandard real number N1 = limn n1 is
non-zero but infinitesimal.
Remark 1.7.22. One can also develop analogous nonstandard asymptotic
notation on other spaces, such as normed vector spaces or locally compact
groups; see for instance [Ta2011c, 4.5] for an example of the former, and
Hirschfelds nonstandard proof [Hi1990] of Hilberts fifth problem for an
example of the latter. We will however not need to deploy nonstandard
asymptotic notation in such generality here.
Note that we can apply asymptotic notation to individual nonstandard
numbers, in contrast to standard asymptotic notation in which one needs
to have the numbers involved to depend on some additional parameter if
one wishes to prevent the notation from degenerating into triviality. In
particular, we can form well-defined sets such as the bounded nonstandard
reals
O(R) := {x R : x = O(1)}
and the infinitesimal nonstandard reals
o(R) := {x R : x = o(1)}.
Exercise 1.7.18 (Standard part). Show that o(R) is an ideal of the commutative ring O(R), and one has the decomposition O(R) = R o(R), thus
every bounded real x = O(R) has a unique decomposition x = st(x) + (x
st(x)) into a standard part st(x) R and an infinitesimal part x st(x)
o(R). Show that the map x 7 st(x) is a ring homomorphism from O(R) to
R.
Exercise 1.7.19. Let N be an unbounded nonstandard natural number.
Write OZ (N ) := {n Z : n = O(N )} and oZ (N ) := {n Z : n = o(N )}.
Show that oZ (N ) is a subgroup of the additive group OZ (N ), and that
the quotient group OZ (N )/oZ (N ) is isomorphic (as an additive group) to
the standard real numbers R. (One could in fact use this construction as
a definition of the real number system, although this would be a rather
idiosyncratic choice of construction.)
Remark 1.7.23. In some texts, the standard part x = st limn xn of
an ultralimit of a bounded sequence of numbers is denoted by limn xn
(with the ultralimit itself not having a specific notation assigned to it). The
assertion x = st limn xn is also sometimes referred to as xn converges to
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
147
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
148
(xn )1nN
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
149
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
150
(ii) (Infinitary statement) For any ultra approximate group A in a rtorsion nonstandard group G, one can find a nonstandard finite
subgroup H of G such that A4 contains H, and A can be covered
by finitely many left-translates of H.
Q
Proof. Let us first assume (i) and establish (ii). Let A = nQ
An be an
ultra approximate group in an r-torsion nonstandard group G = n Gn .
Since G is r-torsion, and the property of being r-torsion is a first-order
statement in the language of groups, we see from Loss theorem that Gn is
a r-torsion group for all n sufficiently close to . Since the An are all finite
K-approximate groups for some K independent of n, we conclude from (i)
that for all n sufficiently close to , we can find a finite subgroup Hn of
and An can be covered by at most CK,r leftGn such that A4n contains Hn , Q
translates of Hn . Taking H := n Hn and applying Loss theorem again,
we cnoclude that H is a nonstandard finite subgroup of G, that A4 contains
H, and A can be covered by at most CK,r left-translates of H, giving (ii) as
desired.
Now we assume (ii) and establish (i). Suppose for contradiction that (i)
failed. Carefully negating the quantifiers, we conclude that there exists K
such that for each natural number n, one can find a finite K-approximate
group An in an r-torsion group Gn for which there does not exist any finite
subgroup Hn of Gn such that A4n contains Hn and An can be covered
Q by at
most n left-translates
of Hn . We then form the ultraproducts A := n An
Q
and G := n Gn . By Loss theorem, G is an r-torsion nonstandard
group, and A is an ultra approximate
there is
Q group in G. Thus, by (ii),
4
a nonstandard finite subgroup H = n Hn of G such that A contains
H and A is covered by M left-translates of H for some (standard) natural
number M . By Loss theorem, we conclude that for all n sufficiently close to
, Hn is a finite subgroup of Gn , A4n contains Hn , and An is covered by M
left-translates of Hn . In particular, as is nonprincipal, this assertion holds
for at least one n > M . But this contradicts the construction of An .
Note how the infinitary statement in the above proposition is a qualitative version of the more quantitative finitary statement. In particular,
parameters such as K and CK,r have been efficiently concealed in the infinitary formulation, whereas they must be made explicit in the finitary
formulation. As such, the infinitary approach leads to a reduction in the
amount of epsilon management that one often has to perform in finitary
settings.
In the next section we will use the Gleason-Yamabe theorem to establish
(ii) and hence conclude (i) as a corollary (this argument is essentially due to
Hrushovski [Hr2012]). Interestingly, no purely finitary proof of (i) in full
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
151
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
152
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
153
mean by a model later, but as a first approximation one can view a model
as a representation of the ultra approximate group A (or of hAi) that is
macroscopically faithful in that it accurately describes the large scale
behaviour of A (or equivalently, that the kernel of the representation is
microscopic in some sense). In the next section we will see how one can
use Gleason lemma technology to convert this macroscopic control of an
ultra approximate group into microscopic control, which will be the key to
classifying approximate groups.
Models of ultra approximate groups can be viewed as the multiplicative
combinatorics analogue of the more well known concept of an ultralimit of
metric spaces, which we briefly review below the fold as motivation.
The crucial observation is that ultra approximate groups enjoy a local
compactness property which allows them to be usefully modeled by locally
compact groups (and hence, through Theorem 1.1.17, by Lie groups also).
As per the Heine-Borel theorem, the local compactness will come from a
combination of a completeness property and a local total boundedness property. The completeness property turns out to be a direct consequence of
the countable saturation property of ultraproducts, thus illustrating one of
the key advantages of the ultraproduct setting. The local total boundedness property is more interesting. Roughly speaking, it asserts that large
bounded sets (such as A or A100 ) can be covered by finitely many translates of small bounded sets S, where small is a topological group sense,
implying in particular that large powers S m of S lie inside a set such as A
or A4 . The easiest way to obtain such a property comes from the following
lemma of Sanders [Sa2009]:
Lemma 1.8.1 (Sanders lemma). Let A be a finite K-approximate group in
a (global) group G, and let m 1. Then there exists a symmetric subset S
of A4 with |S| K,m |A| containing the identity such that S m A4 .
This lemma has an elementary combinatorial proof, and is the key to
endowing an ultra approximate group with locally compact structure. There
is also a closely related lemma of Croot and Sisask [CrSi2010] which can
achieve similar results, and which will also be discussed below. (The locally
compact structure can also be established more abstractly using the much
more general methods of definability theory, as was first done by Hrushovski
[Hr2012], but we will not discuss this approach here.)
By combining the locally compact structure of ultra approximate groups
A with the Gleason-Yamabe theorem, one ends up being able to model a
large ultra approximate subgroup A0 of A by a Lie group L. Such Lie
models serve a number of important purposes in the structure theory of
approximate groups. Firstly, as all Lie groups have a dimension which is
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
154
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
155
yF
Thus, if dH (E, F ) < r, then every point in E lies within a distance less than
r of a point in F , and every point in F lies within a distance less r of a point
in E (thus E Nr (F ) and F Nr (E), where Nr (F ) is the r-neighbourhood
of E); and conversely, if the latter claim holds, then dH (E, F ) r.
This distance is always symmetric, non-negative, and obeys the triangle
inequality. If one restricts attention to non-empty compact sets E, F , then
one easily verifies that dH (E, F ) = 0 if and only if E = F , so that the
Hausdorff distance becomes a metric. In particular, it becomes meaningful
to discuss the concept of a sequence En of non-empty compact subsets of X
converging in the Hausdorff distance to a (unique) limit non-empty compact
set E. Note that this concept captures the intuitive example (iii) given
above, but not any of the others. Nevertheless, we will discuss it first as it
is slightly simpler than the more general notions we will be using.
Exercise 1.8.1 (Hausdorff convergence and connectedness). Let En be a
sequence of non-empty compact subsets of a metric space X converging to
another non-empty compact set E.
(i) If the En are all connected, show that E is connected also.
(ii) If the En are all path-connected, does this imply that E is pathconnected also? Support your answer with a proof or counterexample.
(iii) If the En are all disconnected, does this imply that E is disconnected also? Support your answer with a proof or counterexample.
One of the key properties of Hausdorff distance in a compact set is that
it is itself compact:
Lemma 1.8.3 (Compactness of the Hausdorff metric). Let En be a sequence
of compact subsets of a compact metric space X. Then there is a subsequence
of the En that is convergent in the Hausdorff metric.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
156
The proof of this lemma was set as Exercise 1.5.13. It can be proven by
conventional means (relying in particular on the Heine-Borel theorem),
but we will now sketch how one can establish this result using ultralimits
instead. As in Section 1.7, we fix a standard universe U and a non-principal
ultrafilter in order to define ultrapowers.
If X = (X, d) is a (standard) metric space with metric d : X X R+ ,
then the ultrapower X comes with a nonstandard metric d : X X
R+ that extends d. (In the previous sections, we referred to this metric
as d rather than d, but here it will be convenient to use a different symbol
for this extension of d to reduce confusion.) Let us say that two elements
x, y of X are infinitesimally close if d(x, y) = o(1). (The set of points
infinitesimally close to a standard point x is sometimes known as the monad
of x, although we will not use this terminology.)
Exercise 1.8.2. Let (X, d) be a (standard) compact metric space. Show
that for any nonstandard point x X there exists a unique standard point
st(x) X which is infinitesimally close to x, with
d(st(x), st(y)) = st( d(x, y)).
Exercise 1.8.3 (Automatic completeness).
Let (X, d) be a (standard) comQ
pact metric space, and let E = n En be a nonstandard subset of X
(i.e. an ultraproduct of standard subsets En of X), and let st : X X be
the standard part function as defined in the previous exercise. Show that
st(E) is complete (and thus compact, by the Heine-Borel theorem). (Hint:
take advantage of the countable saturation property from Section 1.7.)
Exercise 1.8.4. Let (X, d) be a compact metric space, and let En for
Q nN
be a sequence of non-empty compact subsets of X. Write E := n En
for the ultraproduct of the En , and let st : X X be the standard part
function as defined in the previous exercises. Show that st(E) is a nonempty compact subset of X, and that limn dH (En , st(E)) = o(1). Use
this to give an alternate proof of Lemma 1.8.3.
One can extend some of the above theory from compact metric spaces
to locally compact metric spaces.
Exercise 1.8.5. Let (X, d) be a (standard) locally compact metric space.
Let X O(X) X denote the set of ultralimits limn xn of precompact
sequences xn in X.
(i) Show that O(X) = X if and only if X is compact.
(ii) Show that there is a unique function st : O(X) X such that x is
infinitesimally close to st(x) for all x O(X).
(iii) If E is a nonstandard subset of X, show that st(E O(X)) is
complete.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
157
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
158
C balls of radius in the dn metric. (Of course, this implies that each Xn
is individually totally bounded.)
Proposition 1.8.5 (Uniformly total bounded spaces are Gromov-Hausdorff precompact). Let Xn be a sequence of uniformly totally bounded metric
spaces. Then there exists a subsequence Xnj which converges in the GromovHausdorff space to a compact limit X .
We will prove this proposition using ultrafilters. Let Xn = (Xn , dn ) be a
sequence of uniformly totally bounded metric spaces, which we will assume
to be standard (by defining the standardQuniverse in a suitable fashion).
Then we can form the ultraproduct X := n Xn and the ultralimit d :=
limn dn , thus d : X X R+ is a nonstandard metric (obeying the
nonstandard symmetry, triangle inequality, and positivity properties). As
the Xn are uniformly bounded, d has bounded range. If we then take the
standard part st(d) : X X R+ of d, then st(d) is a pseudometric on X
(i.e. it obeys all the axioms of a metric except possibly for positivity.) We
can then construct a quotient metric space X := X/ in the usual manner,
by declaring two points x, y in X to be equivalent, x y, if st(d)(x, y) = 0
(or equivalently if d(x, y) = o(1)). The pseudometric st(d) then descends to
a genuine metric d : X X R+ on X . The space (X , d ) is
known as a metric ultralimit (or ultralimit for short) of the (Xn , dn ) (note
that this is a slightly different usage of the term ultralimit from what we
have been using previously).
One can easily establish compactness:
Exercise 1.8.8 (Compactness).
(i) Show that (X , d ) is totally bounded. (Hint: use the uniform
total boundedness of the Xn together with Loss theorem.)
(ii) Show that (X , d ) is complete. (Hint: use countable saturation).
In particular, by the Heine-Borel theorem, X is compact.
Now we establish Gromov-Hausdorff convergence, in the sense that for
every standard > 0, one has dGH (Xn , X ) < for all n sufficiently close
to . From this it is easy to extract a subsequence Xnj that converges in
the Gromov-Hausdorff sense to X as required.
Fix a standard > 0. As X is totally bounded, we can cover it
by at most M balls B (x1, , /10), . . . , B (xM, , /10) of radius /10
(say) in the metric d for some standard natural number M . Lifting
back to the ultraproduct X, we conclude that we may cover X by M balls
B(x1 , 2/10), . . . , B(xM , 2/10) in the nonstandard metric d.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
159
Note that d (xi, , xj, ) is the standard part of d(xi , xj ), and in particular
d (xi, , xj, ) /10 < d(xi , xj ) < d (xi, , xj, ) + /10.
Each ball centre xi is an ultralimit xi = limn xi,n of points xi,n in Xn . By
Loss theorem, we conclude that for n sufficiently close to , Xn is covered
by the M balls Bn (x1,n , 2/10), . . . , Bn (xM,n , 2/10) in the metric dn , and
d (xi, , xj, ) /10 < dn (xi,n , xj,n ) < d (xi, , xj, ) + /10
for all 1 i, j M .
Because of this, we can embed both Xn and X in the disjoint union
Xn ] X , with a metric d(n) extending the metrics on Xn and X , with
the cross distances d(n) (x, y) between points x Xn and y X defined by
the formula
d(n) (x, y) = d(n) (y, x) := inf{d (xi, , y) + d (xi,n , x) + /2 : 1 i M };
informally, this connects each ball center xi, to its counterpart xi,n by a
path of length /2. It is a routine matter to verify that d(n) is indeed a
metric, and that Xn and X are separated by a Hausdorff distance less
than in Xn ] X , and the claim follows.
Exercise 1.8.9. Prove Proposition 1.8.5 without using ultrafilters.
Exercise 1.8.10 (Completeness). Let Xn be a sequence of bounded metric
spaces which is Cauchy in the Gromov-Hausdorff sense (i.e. dGH (Xn , Xm )
0 as n, m ). Show that Xn converges in the Gromov-Hausdorff sense
to some limit X.
Now we generalise Gromov-Hausdorff convergence to the setting of unbounded metric spaces. To motivate the definition, let us first give an equivalent form of Gromov-Hausdorff convergence:
Exercise 1.8.11. Let (Xn , dn ) be a sequence of bounded metric spaces, and
let (X , d ) be another bounded metric space. Show that the following are
equivalent:
(i) (Xn , dn ) converges in the Gromov-Hausdorff sense to (X , d ).
(ii) There exist maps n : X Xn which are asymptotically isometric isomorphisms in the sense that supx,yX |dn (n (x), n (y))
d(x, y)| 0 and supxn Xn distn (xn , n (X )) 0 as n .
Define a pointed metric space to be a triplet (X, d, p), where (X, d) is a
metric space and p is a point in X.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
160
x,yB (p ,R)
as n ;
(3) (Asymptotic surjectivity) For each R0 > R > 0, one has
sup
as n .
Exercise 1.8.12. Verify that all the examples (i)-(v) given at the start of
the section are examples of pointed Gromov-Hausdorff convergence, once
one selects a suitable point in each space.
The above definition is by no means the only definition of pointed
Gromov-Hausdorff convergence. Here is another (which is basically the original definition of Gromov [Gr1981]):
Exercise 1.8.13. Let (Xn , dn , pn ) be a sequence of pointed metric spaces,
and let (X , d , p ) be another pointed metric space. Show that the following are equivalent:
(1) (Xn , dn , pn ) converges in the pointed Gromov-Hausdorff sense to
(X , d , p ).
(2) There exist metrics dn on the disjoint union Xn X extending
the metrics dn , d such that for some sequence n 0, one has
dn (pn , p ) n , Bn (pn , 1/n ) Nn (X ), and B (p , 1/n )
Nn (Xn ).
Using this equivalence, construct a pseudometric between pointed metric
spaces that describes pointwise Gromov-Hausdorff convergence.
Exercise 1.8.14. Let (Xn , dn , pn ) be a sequence of pointed metric spaces
which converge in the pointed Gromov-Hausdorff sense to a limit (X , d , p ),
0 , d0 , p0 ). Suppose that these two limit spaces
and also to another limit (X
0 (p0 , R)
are proper, which means that the closed balls B (p , R) and B
are compact. Show that the two limit spaces are pointedly isometric, thus
0 that maps p
0
there is an isometric isomorphism : X X
to p .
In the compact case, Gromov-Hausdorff convergence and pointwise GromovHausdorff convergence are almost equivalent:
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
161
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
162
k
|B|
m
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
163
|A2 gA2 | (1
1
)|A2 |.
m
1
)|A2 |.
m
1 0
|A A|,
m
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
164
1
)f (t)|A|.
m
In particular, from (1.67) and the definition of f , we conclude that
|(A0 gA0 )A| (1
t2
|A|
2K
and using the Cauchy-Schwarz inequality and the bound |A2 | K|A|, we
conclude that
X X
t2
1aA0 (x)1bA0 (x) |A|3
K
2
xA a,bA
and thus
X
|aA0 bA0 |
a,bA
t2
|A|3 .
K
a1 b)
X
ga1 A
|A0 gA0 |
t2
|A|2 .
K
t2
2K |A|
for at
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
165
`2 (G)
1 |A|3/2
<
m K 1/2
)
.
Exercise 1.8.19. Show that S is symmetric, contains the origin, and that
S m A4 .
To finish the proof of Sanders lemma, it suffices to show that there are
lots of almost periods of 1A 1A in the sense that |S| K,m |A|.
The key observation here is that the translates (g)1A 1A of 1A 1A ,
as g varies in A, range in a totally bounded set. This in turn comes from
a compactness property of the convolution operator f 7 f 1A . when
f is supported on A2 . Croot and Sisask establish this by using a random
sampling argument to approximate this convolution operator by a bounded
rank operator. More precisely, let M 1 be an integer parameter to be
chosen later, and select M sample points y1 , . . . , yM from A2 independently
and uniformly at random (allowing repetitions). We will approximate the
operator
X
T f := f 1A =
f (y)1yA
yA2
|A2 | X
Sf :=
f (yi )1yi A .
M
i=1
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
166
and
EkSf T f k2`2 (G) K |A|3 /M.
In particular, with probability at least 1/2, one has
kSf T f k`2 (G) K |A|3/2 M 1/2 .
Thus, for any g A, we have
kS1gA T 1gA k`2 (G) K |A|3/2 M 1/2
with probability at least 1/2. By the pigeonhole principle (or the first moment method), we thus conclude that there exists a choice of sample points
y1 , . . . , yM for which
(1.69)
|
2
coefficients between 0 and 1. Each of these functions |A
M 1yi A has an ` (G)
norm of OK (|A|3/2 ). Thus, by the pigeonhole principle, one can find a subset
A00 of A0 of size |A00 | K,m,M |A| such that the functions S1gA for g A00
3/2
1 |A|
all lie within 2m
in `2 (G) norm of each other. If M is large enough
K 1/2
depending on K, m, we then conclude from (1.69) and the triangle inequality
3/2
1 |A|
that the functions T 1gA for g A00 all lie within m
in `2 (G) norm of
K 1/2
00
each other. Translating by some fixed element of A , we obtain the claim.
Remark 1.8.11. For future reference, we observe that the above argument
did not need the full strength of the hypothesis that A was an K-approximate
group; it would have sufficed for A to be finite and non-empty with |A2 |
K|A|.
Remark 1.8.12. The above version of the Croot-Sisask argument gives
a lower bound on S of the shape |S| exp(O(m2 K log K))|A|. As was
observed in a subsequent paper of Sanders [Sa2010], by optimising the argument (in particular, replacing `2 with `p for a large value of p, and replacing
1A 1A by 1A 1A2 ), one can improve this to |S| exp(O(m2 log2 K))|A|.
Remark 1.8.13. It is also possible to replace the random sampling argument above by a singular value decomposition of T (restricted to something
like A2 ) to split it as the sum of a bounded rank component and a small
operator norm component, after computing the Frobenius norm of T in order to limit the number of large singular values. (In the abelian setting,
this corresponds to a Fourier decomposition into large and small Fourier
coefficients, which is a fundamental tool in additive combinatorics.) We will
however not pursue this approach here.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
167
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
168
slightly better than this by looking at the left action of G on the k-element
quotient space G/H, which one can think of as a homomorphism from G to
the symmetric group Sk of k elements. The kernel N of this homomorphism
is then clearly a normal subgroup of G of index at most |Sk | = k! that is
contained in H. However, this slightly more efficient argument seems to
be more fragile than the more robust proof given above, in that it is not
obvious (to me, at least) how to adapt it to the approximate group setting
of the Sanders lemma. (This illustrates the advantages of knowing multiple
proofs for various basic facts in mathematics.)
Inspired by the above argument, we now prove Lemma 1.8.14. The main
difficulty is to find an approximate version of the claim that the intersection of two finite index subgroups is again a finite index subgroup. This is
provided by the following lemma:
Lemma 1.8.17. Let A be a K-approximate group, and let A1 , A2 A be
such that |A1 | 1 |A| and |A2 | 2 |A|. Then there exists a subset B of A
1
with BB 1 A1 A1
1 A2 A2 and |B| 1 2 |A|/K.
1
2
Proof. Since A1
1 A2 A , we have |A1 A2 | K|A|. It follows that there
is some x with at least 1 2 |A|/K representations as a1
1 a2 . Let B be the
set of all values of a2 that appear. Obviously BB 1 A2 A1
2 . Suppose
1
0
0
that a2 , a2 B. Then there are a1 , a1 such that x = a1 a2 = (a01 )1 a02 , and
1 lies in A A1 as well.
0 1
so a01 a1
1 1
1 = a2 a2 . Thus BB
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
169
2
(AA
m ) Am1
for all m 1, and such that A4 can be covered by finitely many lefttranslates of Am for each m. (Hint: you will need Lemma 1.8.14, Lemma
1.8.17, Exercise 1.8.21, and some sort of recursive construction.)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
170
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
171
if a is an element of hAi, then a will be small (in the sense that it lies
in A) if (a) is close to the identity, and large (in the sense that it lies
outside A) if (a) is far away from the identity. Thus is faithful in some
coarse or macroscopic sense. The inclusion U0 (A) is a sort of local
surjectivity condition, and ensures that L does not contain any excess or
redundant components. The approximation by nonstandard set axiom is
a technical measurability axiom, that ensures that the model of the ultra
approximate group actually has something nontrivial to say about the finite
approximate groups that were used to build that ultra approximate group
(as opposed to being some artefact of the ultrafilter itself).
Example 1.8.23. We continue the example from Remark 1.8.19. A good
model for A = [N, N ] is provided by the homomorphism : hAi R
given by the formula (x) := st(x/N ). The thick image and compact image
properties are clear. To illustrate the approximation by nonstandard set
property, take F = [r, r] and U = (s, s) for some (standard) real numbers
0 < r < s. The preimages 1 (F ) = {n Z : |n| rN + o(N )} and
1 (U ) = {n Z : |n| < (s )N for some standard > 0} are not
nonstandard finite sets (why? use the least upper bound axiom), but one
can find a nonstandard integer M such that r < st(M/N ) < s, and [M, M ]
will be a nonstandard finite set between F and U .
The following fundamental observation is essentially due to Hrushovski
[Hr2012]:
Exercise 1.8.24 (Existence of good models). Let A be an ultra approximate
group. Show that A4 has a good model by a locally compact Hausdorff
metrisable group, given by the construction discussed previously.
Exercise 1.8.25. The purpose of this exercise is to show why it is necessary
to model A4 , rather than A, in Exercise 1.8.24. Let F2 be the field of two
elements. For each positive integer n, let An be a random subset of F2n
formed by selecting one element uniformly at random from the set x, x + en
for each x F2n1 {0}, and also selecting the identity 0. Clearly, An is a
2-approximate group, since An is contained in F2n = An + {0, en }.
(i) Show that almost surely, for all but finitely many n, An does not
contain any set of the form B + B, with |B| 20.9n . (Hint: use
the Borel-Cantelli lemma.)
Q
(ii) Show that almost surely, the ultraproduct n An cannot be
modeled by any locally compact group.
Let us now give some further examples of good models, beyond that
given by Example 1.8.23.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
172
0 0 1
Q
Consider the ultraproduct A := n An ; this is a subset of
1the
nilpoxz
1 Z Z
tent (nonstandard) group 0 1 Z , consisting of all elements 0 1 y with
0 0
0 0 1
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
173
1 R R
0 1 R
0 0 1
defined by
(1.70)
1 x z
0 1 y
0 0 1
1 st
:=
0
0
x
N
1
0
z
N2
y
st N
st
.
0 0 1
and
hAi :=
1 O(N ) O(N 10 )
0 1
O(N )
0 0
1
,
x
y
z
= st , st , st 10 .
N
N
N
is an ultra approximate group.
1 x z
0 1 y
0 0 1
A1
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
174
n An .
an An
and
|A| := lim |An |,
n
and the infimum is over all F + for which F + (a) f ((a)) for all a A.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
175
where for each standard natural number n, a mod 2n {0, . . . , 2n1 } is the
remainder of a modulo 2n (this is well-defined in A) and the limit is in the
2-adic metric. Note that the image (A) of A is the entire group Z2 , and
conversely the preimage of Z2 in A8 = A is trivially all of Z/2N Z; as such,
one can quotient out Z2 in this model and recover the trivial model of A.
Example 1.8.28 (Nonstandard abelian 2-torsion group). In a similar spirit
to
example, the nonstandard 2-torsion group A := (Z/2Z)N =
Q the preceding
n
N by
n (Z/2Z) can be modeled by the compact abelian group (Z/2Z)
the formula
(a) := lim n (a)
n
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
176
kZ
kZ
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
177
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
178
be shown that after quotienting out the (unique) maximal compact normal
subgroup from the Lie model L, the resulting quotient group (which is also
a Lie group, and in some sense describes the large scale structure of L) is
unique up to isomorphism. The following exercise fleshes out the details of
this observation.
Exercise 1.8.32 (Large-scale uniqueness of the Lie model). Let L, L0 be
connected Lie groups, and let A be an ultra approximate group with good
models : hAi L and 0 : hA0 i L0 .
(i) Show that the centre Z(L) := {g L : gh = hg for all h L} is an
abelian Lie group.
(ii) Show that the connected component Z(L)0 of the identity in Z(L)
0
is isomorphic to Rd (R/Z)d for some d, d0 0.
(iii) Show that the quotient group Z(L)/Z(L)0 is a finitely generated
00
abelian group, and is isomorphic to Zd H for some finite group
H.
(iv) Show that the torsion points of Z(L) are contained in a compact
0
subgroup of Z(L) isomorphic to (R/Z)d H.
(v) Show that any finite normal subgroup of L is central, and thus
lies in the compact subgroup indicated above. (Hint: L will act
continuously by conjugation on this finite normal subgroup.)
(vi) Show that given any increasing sequence N1 N2 . . . of comS
pact normal subgroups of L, the upper bound n Nn is also a compact normal subgroup of L. (Hint: The dimensions of Nn (which
are well-defined by Cartans theorem) are monotone increasing but
bounded by the dimension of L. In particular, the connected components Nn0 must eventually stabilise. Quotient them out and then
use (v).)
(vii) Show that L contains a unique maximal compact normal subgroup
N . Similarly, L0 contains a unique maximal compact normal subgroup N 0 . Show that the quotient groups L/N, L0 /N 0 contain no
non-trivial compact normal subgroups.
(viii) Show that 1 (N ) = ( 0 )1 (N 0 ). (Hint: if g L, then g N iff
the group generated by g and its conjugates is bounded.)
(ix) Show that L/N is isomorphic to L0 /N 0 .
(x) Show that for sufficiently large standard m, Am can be modeled by
a Lie group with no non-trivial compact normal subgroups, which
is unique up to isomorphism.
To illustrate how this theorem is useful, let us apply it in the bounded
torsion case.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
179
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
180
(Hint: first find and prove an analogous statement for ultra approximate
groups, in which the function F is not present.)
There is a finitary formulation of Theorem 1.8.31, but it takes some
effort to state. Let L be a connected Lie group, with Lie algebra l which we
identify using some coordinate basis with Rd , thus giving a Euclidean norm
kk on l. We say that L with this basis has complexity at most M for some
M 1 if
(1) The dimension d of L is at most M ;
(2) The exponential map exp : l L is injective on the ball {x l :
kxk 1/M };
(3) One has k[x, y]k M kxkkyk for all x, y l.
We then define balls BR on L by the formula
BR := {exp(x) : x l; kxk < R}.
Exercise 1.8.36 (Hrushovskis Lie model theorem, finitary version). Let
F : N N be a function, and let A be a finite K-approximate group. Show
that there exists a natural number 1 M K,F 1, a connected Lie group L
of complexity at most M , a symmetric set A0 containing the identity with
(A0 )4 A4 , and a map : (A0 )F (M ) L obeying the following properties:
(i) (Large subgroup) A can be covered by M left-translates of A0 .
(ii) (Approximate homomorphism) One has (1) = 1, and for all a, b
(A0 )F (M ) with ab (A0 )F (M ) , one has
(ab)(b)1 (a)1 B1/F (M ) .
(iii) (Thick image) If a (A0 )F (M ) and (a) B1/M , then a A0 .
Conversely, if g B1/M , then (A0 ) intersects gB1/F (M ) .
M.
(iv) (Compact image) One has (A0 ) BM
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
181
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
182
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 L(G) 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 (Theorem 1.8.31), 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 1.8.28. In this example one
studied a nilbox from a Heisenberg group, which we rewrite here in slightly
different notation. Specifically, let G be the Heisenberg group
G := {(a, b, c) : a, b, c Z}
with group law
(a, b, c) (a0 , b0 , c0 ) := (a + a0 , b + b0 , c + c0 + ab0 )
Q
and let A = n An , where An G is the box
(1.72)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
183
kg n k |n|kgk
when |n|kgk was sufficiently small. Such axioms have important and nontrivial content even in the microscopic regime where g or h are extremely
close to the identity. For instance, in the proof of Jordans theorem (Theorem 1.1.2), a key step was to apply the commutator axiom (1.73) (for the
distance to the identity in operator norm) to the most microscopic element of G, or more precisely a non-identity element of G of minimal norm.
The key point was that this microscopic element was virtually central in
G, and as such it restricted much of G to a lower-dimensional subgroup of
the unitary group, at which point one could argue using an induction-ondimension 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 A A1 and A0 (A0 )1
discussed earlier, the element (0, 0, 1) 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 A A1
and A0 (A0 )1 as a tower of central cyclic extensions, which in particular
establishes the nilpotency of both groups.
The escape axiom (1.74) is a particularly important axiom in connecting the microscopic structure of a group G to its macroscopic structure;
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
184
for instance, as shown in Section 1.3, this axiom (in conjunction with the
closely related commutator axiom) tends to imply dilation estimates such as
d(g n , hn ) nd(g, h) that allow one to understand the microscopic geometry
of points g, h close to the identity in terms of the (local) macroscopic geometry of points g n , hn 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 A that obeys the escape and commutator axioms (while being non-degenerate enough to adequately capture the
geometry of A in some sense), in a fashion analogous to the Gleason metrics that played such a key role in the theory of Hilberts 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 A that only obeys the escape and
commutator estimates macroscopically; roughly speaking, this means that
one has a macroscopic commutator inequality
k[g, h]k kgkkhk + o(1)
and a macroscopic escape property
kg n k |n|kgk o(|n|)
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 A in a
group G, and an element g of G, we can define the escape norm kgke,A of g
by the formula
1
: n N; g, g 2 , . . . , g n A .
kgke,A := inf
n+1
Thus, kgke,A equals 1 if g lies outside of A, equals 1/2 if g lies in A but g 2
lies outside of A, and so forth. Such norms had already appeared in Section
1.5, 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 A which we will present shortly.
However, it need not actually be a norm; in particular, the triangle inequality
kghke,A kgke,A + khke,A
is not necessarily true. Fortunately, it turns out that by a (slightly more
complicated) version of the Gleason machinery from Section 1.5 we can
establish a usable substitute for this inequality, namely the quasi-triangle
inequality
kg1 . . . gk ke,A C(kg1 ke,A + + kgk ke,A ),
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
185
(S A )1000K A.
A, then g S.
Q
An ultra strong K-approximate group is an ultraproduct A = n An of
strong K-approximate groups.
The first trapping condition can be rewritten as
kgke,A 1000kgke,A100
and the second trapping condition can similarly be rewritten as
kgke,S 106 K 3 kgke,A .
This makes the escape norms of A, A100 , and S 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 (1.53), 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 1.9.2. Let N be a large natural number. Then the interval A =
[N, N ] in the integers is a 2-approximate group, which is also a strong
2-approximate group (setting S = [106 N, 106 N ], for instance). On the
other hand, if one places A in Z/5N Z rather than in the integers, then the
first trapping condition is lost and one is no longer a strong 2-approximate
group. Also, if one remains in the integers, but deletes a few elements from
A, e.g. deleting b1010 N c from A), then one is still a O(1)-approximate
group, but is no longer a strong O(1)-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:
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
186
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
187
(2) From the Gleason lemma, the elements with zero escape norm form
a normal subgroup of A. Quotient these elements out. Show that
all non-identity elements will have positive escape norm.
(3) Find the non-identity element g1 in (the quotient of) A of minimal escape norm. Use the commutator estimate (assuming it is
inherited by the quotient) to show that g1 will centralise (most of)
this quotient. In particular, the orbit hg1 i is (essentially) a central
subgroup of hAi.
(4) Quotient this orbit out; then find the next non-identity element g2
in this new quotient of A. Again, show that hg2 i is essentially a
central subgroup of this quotient.
(5) Repeat this process until A becomes entirely trivial. Undoing all
the quotients, this should demonstrate that hAi is virtually nilpotent, and that A 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 A (in the sense that
the quotient of A 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 hAi is no longer as strong
as virtually nilpotent, but is instead a more complicated tower alternating
between (ultra) finite extensions and central extensions.
Example 1.9.4. Consider the strong O(1)-approximate group
A := {aN 10 + 5b : |a| N ; |b| N 2 }
in the integers, where N is a large natural number not divisible by 5. As
Z is torsion-free, all non-zero elements of A have positive escape norm, and
the nonzero element of minimal escape norm here is g = 5 (or g = 5). But
if one quotients by hgi, A projects down to Z/5Z, 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 A with hgi = 5Z is not a simple progression, but is
a more complicated object, namely a generalised arithmetic progression of
rank two.)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
188
To deal with this issue, we will not quotient out by the entire cyclic
group hgi = {g n : n Z} generated by the element g of minimal escape
norm, but rather by an arithmetic progression P = {g n : |n| N }, where
N is a natural number comparable to the reciprocal 1/kgke,A 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 P
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 A is
in the global case when A 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, Theorem 1.5.24); details can be found in [BrGrTa2011], but will
only be sketched here.
1.9.1. Gleasons lemma. Throughout this section, A is a strong K-approximate
group in a global group G. We will prove the various estimates in Theorem
1.9.3. The arguments will be very close to those in Section 1.5; indeed, it
is possible to unify the results here with the results in that section 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 A. Now we turn to the escape property. By symmetry, we may take n to be positive (the n = 0 case is trivial). We may of
course assume that kgke,A is strictly positive, say equal to 1/(m + 1); thus
g, . . . , g m A and g m+1 6 A, and n m. By the first trapping property,
this implies that g j(m+1) 6 A100 for some 1 j 1000.
Let kn be the first multiple of n larger than or equal to j(m + 1), then
kn m + 1. Since kn j(m + 1) is less than m, we have g knj(m+1) A;
since g j(m+1) 6 A100 , we conclude that g kn 6 A99 . In particular this shows
that kg n ke,A 1/k n/(m + 1), and the claim follows.
The escape property implies the conjugacy bound:
Exercise 1.9.2. Establish the conjugacy bound. (Hint: one can mimic the
arguments establishing a nearly identical bound in Section 1.5.2.)
Now we turn to the triangle inequality, which (as in Section 1.5) is the
most difficult property to establish. Our arguments will closely resemble the
4
proof of Proposition 1.5.9, with S A and A playing the roles of U1 and U0
from that argument. As in that theorem, we will initially assume that we
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
189
(1.76)
1
kgke,A kgk,A kgke,A
M
for all g G.
We introduce the function : G R+ by
(x) := (1 M dist,A (x, A))+ ,
where dist,A (x, A) := inf yA kx1 yk,A . Then takes values between 0 and
1, equals 1 on A, is supported on A2 , and obeys the Lipschitz bound
kg k` (G) M kgke,A
(1.78)
for all g, thanks to the triangle inequality for kk,A and (1.77). We also
introduce the function : G R+ by
j
2
: x (S A )j A} {0},
104 K 3
then also takes values between 0 and 1, equals 1 on A, is supported on
A2 , and obeys the Lipschitz bound
(x) := sup{1
(1.79)
khy k` (G)
1
104 K 3
1 X
(xy)(y 1 ).
|A|
yG
kk` (G) K 2 .
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
190
1 X
g (y)(y 1 x)
|A|
yG
kg k` (G) K 2 M kgke,A .
1 X
g (y)hy (y 1 x)
|A|
yG
we see that
n1
kg k` (G)
1X
1
kgn k` (G) +
kgi g k` (G)
n
n
i=0
and thus
1
1
+ 4 M kgke,A
n 10
S. Using the second trapping property, this
kg k` (G)
whenever g, g 2 , . . . , g n1
implies that
1
M + OK (1))kgke,A
104
In the converse direction, if kg k` (G) < n1 , then
kg k` (G) (
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
191
1
which is (1.76) with M replaced by 10
M +OK (1). If we knew (1.76) for some
large but finite M , we could iterate this argument and conclude that (1.76)
held with M replaced by OK (1), which would give the triangle inequality.
Now it is not immediate that (1.76) holds for any finite M , but we can avoid
this problem with the usual regularisation trick of replacing kgke,A with
kgke,A + throughout the argument for some small > 0, which makes (1.76)
automatically true with M = O(1/), run the above iteration argument, and
then finally send to zero.
Exercise 1.9.3. 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 Gleasons lemma:
Exercise 1.9.4. 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
1.5.2.)
The proof of Theorem 1.9.3 is now complete.
Exercise 1.9.5. Generalise Theorem 1.9.3 to the setting where A is not
necessarily finite, but is instead an open precompact subset of a locally
compact group G. (Hint: replace cardinality by left-invariant Haar measure
and follow the arguments in Section 1.5 closely.) Note that this already
gives most of one of the key results from that section, 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.
1.9.2. A cheap version of the structure theorem. In this section we
use Theorem 1.9.3 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 A is NSS if it contains
no non-trivial subgroups of the ambient group, or equivalently if every nonidentity element of A 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 1.9.6 (Reduction to the NSS case). Let L be a connected Lie
group with Lie algebra l, let B be a bounded symmetric convex body in l,
let r > 0 be a sufficiently small standard real. Let 0 < r0 < r/2, and let A
be an ultra strong approximate group which has a good model : hAi L
with
1 (exp(rB)) A 1 (exp(1.1rB)).
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
192
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
193
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
194
(i) Show that if A has a good model by a connected Lie group L, then
L is nilpotent. (Hint: first use Exercise 1.9.1, and then induct on
the dimension of L.)
(ii) Show that A is covered by finitely many left translates of a nonstandard subgroup G0 of G which admits a normal series
G0 = G00 B G01 B G02 B B G0k = {1}
for some standard k, where for every 0 i < k, G0i+1 is a normal nonstandard subgroup of G0 and of G0i , and G0i /G0i+1 is either
a nonstandard finite group or a nonstandard central subgroup of
G0 /G0i+1 . Furthermore, if G0i /G0i+1 is not central, then it is contained in the image of A4 G0 in G0 /G0i+1 ; and G0 is generated (as
a nonstandard subgroup) by A4 G0 . (Hint: first use the Lie model
theorem and Exercise 1.9.1, and then induct on the dimension of
L.)
Exercise 1.9.10 (Cheap structure theorem, finite version). Let A be a
finite K-approximate group in a group G. Show that A is covered by OK (1)
left-translates of a subgroup G0 of G which admits a normal series
G0 = G00 B G01 B G02 B B G0k = {1}
for some k = OK (1), where for every 0 i < k, G0i+1 is a normal subgroup
of G0i and of G0 , and G0i /G0i+1 is either finite or central in G0 /G0i+1 . Furthermore, if G0i /G0i+1 is not central, then it is contained in the image of A4 G0
in G0 /G0i+1 , and G0 is generated as a group by A4 G.
Exercise 1.9.11.
(i) Show that any finite extension G of a virtually nilpotent group H
(thus there is a short exact sequence 0 K G H 0 with
K finite) is virtually nilpotent. (Hint: G acts on K by conjugation;
look at the stabiliser of this action.)
(ii) Conclude that the group G0 in the previous exercise is virtually
nilpotent.
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 K
of the ultra approximate group, as laid out in the following exercise.
Exercise 1.9.12 (Nilpotent groups). A Lie algebra g is said to be nilpotent
if the derived series g1 := g, g2 := [g1 , g], g3 := [g2 , g], . . . 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.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
195
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
196
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
197
for minor technical reasons to assume that the local group L 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 1.9.7 (Local Hrushovski Lie model theorem). Let A be a (local)
ultra approximate group. Then there is an ultra approximate subgroup A0
of A (thus (A0 )4 A4 ) with A covered by finitely many left-translates of A0
(by elements in A (A0 )1 ), which has a good model by a connected local Lie
group L.
The proof of this theorem is basically a localisation of the proof of the
global Lie model theorem from Section 1.8, and is omitted (see [BrGrTa2011]
for details). One key replacement is that if A is a local approximate group
rather than a global one, then the global Gleason-Yamabe theorem (Theorem 1.1.17) must be replaced by the local Gleason-Yamabe theorem (Theorem 1.5.24).
One can define the notion of a strong K-approximate group and ultra
strong approximate group in the local setting without much difficulty, since
strong approximate groups only need to work inside A100 , which is welldefined. Using the local Lie model theorem, one can obtain a local version
of Exercise 1.9.1. The Gleason lemma (Theorem 1.9.3) also localises without
much difficulty to local strong approximate groups, as does the reduction to
the NSS case in Exercise 1.9.6.
Now we once again analyse the NSS case. As before, let L be a connected
(local) Lie group, with Lie algebra l, let B be a bounded symmetric convex
body in l, let r > 0 be a sufficiently small standard real. Let A be a
(local) ultra strong NSS approximate group which has a (local) good model
: hAi L with
1 (exp(rB)) A 1 (exp(1.1rB)).
Again, we assume L has dimension at least 1, since A is trivial otherwise.
We let u be a non-identity element of minimal escape norm. As before,
u will have an infinitesimal escape norm and lie in the kernel of . If we
set N := kuke,A , then N is an unbounded natural number, and the map
: t 7 (g btN c ) will be a local one-parameter subgroup, i.e. a continuous
homomorphism from [1, 1] to L. This one-parameter subgroup will be
non-trivial and centralised by a neighbourhood of the identity in L.
In the global setting, we quotiented (the group generated by a large
portion of) A by the centraliser Z(u) of u. In the local setting, we perform
a more gentle quotienting, which roughly speaking arises by quotienting
A by the geometric progression P := {un : cN n cN }, where c > 0
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
198
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
199
(
g n ) = exp(st(n/M )r0 X)
whenever |n| 4N (say) for some Y 1.1B\B that is also parallel to 0 (0).
In particular, Y = X for some
1
1.1.
1.1
Since 1/N = kuke,A is the minimal escape norm of non-identity elements
of A, we have k
g ke,A 1/N , and thus gi A2 \A for some 1 i N ; in
i
particular, (
g ) 6 exp(rB). Comparing this with (1.82) we see that
st(i/M )r0 X 6 rB
and thus
st
N
M
1 r
,
1.1 r0
and hence
M
r0
1.3 .
N
r
By the Euclidean algorithm, we can thus find a nonstandard integer number
m such that the quantity
M r
N r0
lies in the interval [0.5, 0.5]. In particular
:= 1 + m
N r0
.
Mr
then (as u commutes with g) we see for all |n| M that
|m| 2
If we set h := gum
hn = gn umn
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
200
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
1.10. Applications
201
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
202
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
1.10. Applications
203
symmetric set S containing the origin with |S| K,M |A| such that
1
(1.84)
k1A 1A (g)1A 1A k`2 (G)
|A|3/2
M
for all g S.
Let A be as in Proposition 1.10.4. We apply Lemma 1.10.5 with some
large M depending on K to be chosen later. Then for any g S 100 one has
100 3/2
|A| .
k1A 1A (g)1A 1A k`2 (G)
M
Since 1A 1A has an `1 (G) norm of |A|2 and is supported on the set A2 ,
which has cardinality at most K|A|, we see from Cauchy-Schwarz that
k1A 1A k`2 (G) K |A|3/2
and hence (if M is large enough depending on K)
k (g)1A 1A k`2 (A2 ) K |A|3/2 .
In particular, we have |gA2 A2 | K |A|, thus every element of S 100 has
K |A| representations of the form xy 1 with x, y A2 . As there are at
most K 2 |A|2 pairs (x, y) with x, y A2 , we conclude that |S 100 | K |A|.
In particular, by the Ruzsa covering lemma (Exercise 1.8.21) we see that S 4
can be covered by OK,M (1) left-translates of S 2 , and hence S 2 is a OK,M (1)approximate group.
In view of Theorem 1.10.1, we thus see that to conclude the proof of
Proposition 1.10.4, it suffices to show that A can be covered by OK,M (1)
left-translates (or right-translates) of S 2 if M is sufficiently large depending
on K.
We will just prove the claim for left-translates, as the claim for righttranslates is similar. We will need the following useful inequality:
Lemma 1.10.6 (Ruzsa triangle inequality). Let A, B, C be finite non-empty
1 ||BC 1 |
subsets of a group G. Then |A C 1 | |AB |B|
.
Proof. Observe that if x is an element of A C 1 , so that x = ac1 for
some a A and c C, then x has at least |B| representations of the form
x = yz with y A B 1 and z B C 1 , since ac1 = (ab1 )(bc1 ) for all
b B. As there are only |A B 1 ||B C 1 | possible pairs (y, z) that could
form such representations, the claim follows.
Now we return to the proof of Proposition 1.10.4. From (1.84) and
Minkowskis inequality we see that
1
1
1A 1A
1S 1A 1A
|A|3/2 ,
|S|
M
2
` (G)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
204
K |A|
` (A2 )
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
1.10. Applications
205
This is a consequence of Theorem 1.10.1 combined with [Ta2008b, Theorem 4.6]; we omit the details. There is also a statistical variant (using
[Ta2008b, Theorem 5.4] instead), based on an additional tool, the (nonabelian) Balog-Szemeredi-Gowers theorem, which will not be discussed in
detail here:
Proposition 1.10.9. Let A, B be finite non-empty subsets of a (global)
group G such that
|{(a, b, a0 , b0 ) A B A B : ab = a0 b0 }| |A|3/2 |B|3/2 /K.
Then there exists a coset nilprogression P of rank and step OK (1) and cardinality |P | K |A| such that A intersects a left-translate of P in a set
of cardinality K |A|, and B intersects a right-translate of P in a set of
cardinality K |A|.
1.10.2. Polynomial growth. The above results show that finite approximate groups A (as well as related objects, such as finite sets of bounded
doubling) can be efficiently covered by virtually nilpotent groups. However,
they do not place all of A inside a virtually nilpotent group. Indeed, this is
not possible in general:
Exercise 1.10.1. Let G be the ax + b group, that is to say the group of
all affine transformations x 7 ax + b on the real line, with a R\{0} and
b R. Show that there exists an absolute constant K and arbitrarily large
finite K-approximate groups A in G that are not contained in any virtually
nilpotent group. (Hint: build a set A which is very long in the b direction
and very thin in the a direction.)
Such counterexamples have the feature of being thin in at least one
of the directions of G. However, this can be fixed by adding a thickness
assumption to the approximate group. In particular, we have the following
result:
Theorem 1.10.10 (Thick sets of bounded doubling are virtually nilpotent).
For every K 1 there exists M 1 such that the following statement holds:
whenever G is a group, S is a finite symmetric subset of G containing the
identity, and A is a finite set containing S M such that |A2 | K|A|, then S
generates a virtually nilpotent group.
Proof. Fix K, and let M be a sufficiently large natural number depending
on K to be chosen later. Let S, A, G be as in the theorem. By Proposition
1.10.4, there exists a virtually nilpotent group H such that A is covered by
OK (1) left-cosets of H. In particular, S m H consists of OK (1) left-cosets
of H for all 0 m M . On the other hand, as S contains the identity,
S m H is nondecreasing in m. If M is large enough, then by the pigeonhole
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
206
[
hSi hSiH =
S k H = S m H.
k=m
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
1.10. Applications
207
can take f (m) to be m(log log m) for some c > 0 and m sufficiently large
depending on the size of S, by the result of my paper with Shalom, but
this is unlikely to be best possible. (In the converse direction, Grigorchuks
construction [Gr1984] of a group of intermediate growth shows that f (m)
cannot grow faster than exp(m ) for some absolute constant < 1 (and it
is believed that one can take = 1/2).)
Exercise 1.10.4 (Infinite groups have at least linear growth). If G is an
infinite group generated by a finite symmetric set S containing the identity,
show that |S m | |S| + m 1 for all m 1.
Exercise 1.10.5 (Linear growth implies virtually cyclic). Let G be an infinite group generated by a finite symmetric set S containing the identity.
Suppose that G is of linear growth, in the sense that |S m | Cm for all
m 1 and some finite C.
(i) Place a left-invariant metric d on G by defining d(x, y) to be the
least natural number m for which x yS m . Define a geodesic to
be a finite or infinite sequence (gn )nI indexed by some discrete
interval I Z such that d(gn , gm ) = |n m| for all n, m I. Show
that there exist arbitrarily long finite geodesics.
(ii) Show that there exists a doubly infinite geodesic (gn )nZ with g0 =
1. (Hint: use (i) and a compactness argument.)
(iii) Show that |S 2m | 2|S m | 1 for all m 1. (Hint: study the balls
of radius m centred at gm and gm .) More generally, show that
|S km | 2k|S m | 2k + 1 for all m, k 1.
(iv) Show that |S m |/m converges to a finite non-zero limit as m ,
thus |S m | = m + o(m) for all m 1, where o(m) denotes a
quantity which, when divided by m, goes to zero as m .
(v) Show that for all m 1, then all elements of S m lie within a
distance at most o(m) of the geodesic (gm , . . . , gm ). (Hint: first
show that all but at most o(m) elements of S m lie within this
distance, using (iv) and the argument used to prove (iii).)
1 lies within distance o(m) of
(vi) Show that for sufficiently large m, gm
gm .
(vii) Show that for sufficiently large m, S m+1 lies within distance m of
1 }.
{1, gm , gm
(viii) Show that G is virtually cyclic (i.e. it has a cyclic subgroup of finite
index).
Exercise 1.10.6 (Nilpotent groups have polynomial growth). Let S be a
finite symmetric subset of a nilpotent group G containing the identity.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
208
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
1.10. Applications
209
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
210
is monotone non-increasing in r.
We will not prove this proposition here (as it requires, among other
things, a definition of Ricci curvature, which would be beyond the scope of
this text); but see for instance [Pe2006, Chapter 9] or [Ta2009b, 2.10] for
a proof. This inequality is consistent with the geometric intuition that an
increase in curvature on a manifold should correspond to a stunting of the
growth of the volume of balls. For instance, in the positively curved spheres,
the volume of balls eventually stabilises as a constant; in the zero curvature
Euclidean spaces, the volume of balls grows polynomially; and in the negative curvature hyperbolic spaces, the volume of balls grows exponentially.
.
Informally, the balls in M cannot grow any faster than the balls in M
Setting = 0, we conclude in particular that if M has non-negative Ricci
curvature and dimension d, then vol(B(p, r))/rd is non-decreasing in r; in
particular, any manifold of non-negative Ricci curvature is of polynomial
growth. Applying Exercise 1.10.7 and Gromovs theorem, we conclude that
any manifold of non-negative Ricci curvature has a fundamental group which
is virtually nilpotent. (In fact, such fundametal groups can be shown, using
the splitting theorem, to be virtually abelian; see [ChGr1972]. However,
this improvement seems to be beyond the combinatorial methods used here.)
The above monotonicity also shows that whenever M has Ricci curvature
at least zero, we have the doubling bound
vol(B(p, 2r)) 2d vol(B(p, r)).
A continuity argument then shows that for every R > 0 and > 0, there
exists > 0 such that if M has Ricci curvature at least , then one has
vol(B(p, 2r)) (2d + ) vol(B(p, r))
for all 0 < r R.
Exercise 1.10.8. By using the above observation combined with Exercise
1.10.7 and Exercise 1.10.2, show that for every dimension d there exists
> 0 such that if M is any compact Riemannian manifold of diameter at
most 1 and Ricci curvature at least everywhere, then 1 (M ) is virtually
nilpotent.
Remark 1.10.14. This result was first conjectured by Gromov, and was
proven by Cheeger-Colding [ChCo1996] and Kapovitch-Wilking [KaWi2011]
using deep Riemannian geometry tools (beyond just the Bishop-Gromov inequality). (See also the paper [KaPeTu2010], which established the analogous result assuming a lower bound on sectional curvature rather than Ricci
curvature.)
The arguments in these papers in fact give more precise information on
the fundamental group 1 (M ), namely that there is a nilpotent subgroup of
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
1.10. Applications
211
step and rank Od (1) and index Od (1). The methods here (based purely on
controlling the growth of balls in 1 (M )) can give the step and rank bounds,
but appear to be insufficient to obtain the index bound. The exercise cannot be immediately obtained via a compactness-and-contradiction argument
from the (easier) = 0 case mentioned previously, because of the problem
of collapsing: there is no lower bound assumed on the injectivity radius of
M , and as such the space of all manifolds with the indicated diameter is
non-compact even if one bounds the derivatives of the metric to all orders.
(An equivalent way of phrasing the problem is that the orbits of 1 (M ) in
the universal cover may be arbitrarily dense, and so a ball of bounded radius
may correspond to an arbitrarily large subset S of the fundamental
in M
group. For this application it is thus of importance that there is no upper
bound on the size of the sets S or A assumed in Exercise 1.10.2.)
Remark 1.10.15. One way to view the above results is as an assertion that
it is quite rare for a compact manifold to be equippable with a Riemannian
metric with (almost) non-negative Ricci curvature. Indeed, an application of
van Kampens theorem [Se1931], [vKa1833] shows that every fundamental group 1 (M ) of a compact manifold is finitely presented, and conversely
a gluing argument for four-manifolds shows that every finitely presented
group is the fundamental group of some (four-dimensional) manifold. Intuitively, most finitely presented groups are not virtually nilpotent, and so
most compact manifolds cannot have metrics with almost non-negative
Ricci curvature.
1.10.4. A Margulis-type lemma. In Exercise 1.10.7 we saw that the fundamental group 1 (M ) of a connected Riemannian manifold can be viewed
. The
as a discrete group of isometries acting on a Riemannian manifold M
,
curvature properties of M then give doubling properties of the balls in M
and hence of 1 (M ), allowing one to use tools such as Exercise 1.10.2.
It turns out that one can abstract this process by replacing the universal
by a more general metric space X:
cover M
Lemma 1.10.16 (Margulis-type lemma). Let K 1. Let X = (X, d) be
a metric space, with the property that every ball of radius 4 can be covered
by K balls of radius 1. Let be a group of isometries of X, which acts
discretely in the sense that {g : gx B} is finite for every x X
and every bounded set B X. Then if x X and is sufficiently small
depending on K, the set S := {g : d(gx, x) } generates a virtually
nilpotent group.
Proof. We can can cover B(x, 4) by K balls B(xi , 1). If g, h S4 and
gx, hx B(xi , 1), then (by the isometric action of ) we see that h1 g S2 .
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
212
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
Chapter 2
Related articles
213
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
214
2. Related articles
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
215
we may thus locate a subgroup G0S of GS of index at most Cd for all finite
subsets S of G, obeying the compatibility
condition G0T = G0S GT whenever
S
T S. If we then set G0 := S G0S , where S ranges over all finite subsets
of G, we then easily verify that G0 is abelian and has index at most Cd in
G, as required.
2.1.1. Proofs. We now give a proof of Schurs theorem, based on the proof
in [We1973]. We begin with a lemma of Burnside. Given a vector space
V , let End(V ) denote the ring of linear transformations from V to itself.
Lemma 2.1.4. Let V be a finite-dimensional complex vector space, and
let A be a complex algebra with identity in End(V ), i.e. a linear subspace
of End(V ) that is closed under multiplication and contains the identity operator. Then either A = End(V ), or there exists some proper subspace
{0} ( W ( V which is A-invariant, i.e. aW W for all a A.
Proof. Suppose that no such proper A-invariant subspace W exists. Then
for any non-zero v V , the vector space Av must equal all of V , since it
is a non-trivial A-invariant subspace. By duality, this implies that for any
non-zero dual vector u V , the vector space A u must equal all of V .
Let u, v be linearly independent elements of V . We claim that there
exists an element a of A such that au 6= 0 and av = 0. Suppose that this
is not the case; then by the Hahn-Banach theorem, there exists t End(V )
such that au = tav for all a A. In particular, setting a = 1 we obtain
u = tv, and thus atv = tav. Replacing a by ab for some b A, we conclude
that tabv = abtv = atbv, thus taat annihilates all of Av. Since Av = V , we
conclude that at = ta, thus A lies in the centraliser of t. But since u = tv
and u, v are linearly independent, t is not a multiple of the identity, and
thus by the spectral theorem, t has at least one proper eigenspace. But this
eigenspace is fixed by A, a contradiction.
Thus we can find a A such that au 6= 0 and av = 0 for any linearly
independent u, v. Iterating this, we see that for any non-zero u V and
any 0 k dim(V ) 1, we can find a A of corank at least k that does
not annihilate u. In particular, A contains a rank one transformation. Since
Av = V and A u = V for all v V and u V , this implies that A
contains all rank one transformations, and hence contains all of End(V ) by
linearity.
Corollary 2.1.5 (Burnsides theorem). Let V be a complex vector space
of some finite dimension d, and let G be a subgroup of GL(V ) where every
element has order at most r. Then G is finite with cardinality Od,r (1).
Proof. We induct on dimension, assuming the claim has already been proven
for smaller values of d. Let A End(V ) be the complex algebra generated
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
216
2. Related articles
by G (or equivalently, the complex linear span of G). Suppose first that there
is a proper A-invariant subspace W . Then G projects down to GL(W ) and
to GL(V /W ), and by induction hypothesis both of these projections are finite with cardinality Od,r (1). Thus there exists a subgroup G0 of G of index
Od,r (1) whose projections to GL(W ) and GL(V /W ) are trivial; in particular, all elements of G0 are unipotent. But as the complex numbers have zero
characteristic, the only unipotent element of finite order is the identity, and
so G0 is trivial, and the claim follows.
Hence we may assume that V has no proper A-invariant subspace. By
Lemma 2.1.4, A must be all of GL(V ). In particular, one can find d2 linearly
independent elements g1 , . . . , gd2 of G.
For any g G, the element ggi has order at most r, and thus all the
eigenvalues of ggi are roots of unity of order at most r. This means that
there are at most Or,d (1) possible values of the trace tr(ggi ), which is a linear
functional of g. Letting gi vary among the basis g1 , . . . , gd2 of End(V ), we
conclude that there are at most Or,d (1) possible values of g, and the claim
follows.
Remark 2.1.6. The question of whether any finite group with m generators in which all elements of order at most r is necessarily of order Om,r (1)
is known as the restricted Burnside problem, and was famously solved affirmatively by Zelmanov [Ze1990]. (Note however that for certain values of m
and r it is possible for the group to be infinite. Also, while any finite group
is trivially embedded in some linear group, one does not have any obvious
control on the dimension of that group in terms of m and r, so one cannot
immediately solve this problem just from Corollary 2.1.5.)
To prove Schurs theorem (Theorem 2.1.2), it thus suffices to establish
the following proposition:
Proposition 2.1.7. Let k be a finitely generated extension of the field of
rationals Q. Then every periodic element of GLd (k) has order at most
Od,k (1).
Indeed, to obtain Schurs theorem, one applies Proposition 2.1.7 with
k equal to the field generated by the coefficients of the generators of the
finitely generated periodic group G, and then applies Corollary 2.1.5.
Proof. Suppose first that k is a finite extension of Q. If a GLd (k) has
period n, then the field generated by the eigenvalues of a contains a primitive
nth root of unity, and thus contains the cyclotomic field of that order. On the
other hand, this field has degree Od (1) over k, and thus has degree Od,k (1)
over the rationals. Thus n = Od,k (1), and the claim follows. Note that the
bound on n depends only on the degree of k, and not on k itself.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
217
Now we extend from the finite degree case to the finitely generated case.
By using a transcendence basis, one can write k as a finite extension of
Q(z1 , . . . , zm ) for some algebraically independent z1 , . . . , zm over Q. By the
primitive element theorem, one can then write k = Q(z1 , . . . , zm )() where
is algebraic over Q(z1 , . . . , zm ) of some degree D.
Now suppose we have an element a GLd (k) of period n, thus a, . . . , an1 6=
1 and an = 1. Let R be the ring in k generated by the coefficients of a. We
can create a ring homomorphism : R k to a finite extension k of the
rationals by mapping each zi to a rational number qi , and then replace
with a root of the polynomial formed by replacing zi with qi in the minimal
polynomial of . As long as one chooses (q1 , . . . , qm ) generically (i.e. outside of a codimension one subset of Qm ), this operation is well-defined on
R (in that no division by zero issues arise in any of the coefficients of a).
Furthermore, generically one has (a), . . . , (a)n1 6= 1 and (a)n = 1, thus
a has period n. Furthermore, the degree of k over Q is at most the degree
D of over Q(z1 , . . . , zm ) and is thus bounded uniformly in q1 , . . . , qm . The
claim now follows from the finite extension case.
1 R R
H(R) := 0 1 R
0 0 1
over an arbitrary ring R (which need not be commutative), or more generally
any matrix group consisting of unipotent upper triangular matrices, and
view a general nilpotent group as being an abstract generalisation of such
concrete groups. (In the case of nilpotent Lie groups, at least, this is quite
an accurate intuition, thanks to Engels theorem.) Or, one can adopt a Lietheoretic viewpoint and try to think of nilpotent groups as somehow arising
from nilpotent Lie algebras; this intuition is rigorous when working with
nilpotent Lie groups (at least when the characteristic is large, in order to
avoid issues coming from the denominators in the Baker-Campbell-Hausdorff
formula), but also retains some conceptual value in the non-Lie setting. In
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
218
2. Related articles
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
219
g h := h1 gh
and
(2.2)
[g, h] := g 1 h1 gh.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
220
2. Related articles
thus one can pull g to the right of h at the cost of adding an additional
commutator factor [g, h] to the right. Finally, commutation is related to
conjugation by the identity
g h = g[g, h].
The commutator can be viewed as a nonlinear group-theoretic analogue
of the Lie bracket. For instance, in a matrix group G = GLn (C), we observe
that the commutator [1 + A, 1 + B] of two elements 1 + A and 1 + B
close to the identity is of the form 1 + 2 (AB BA) + O(3 ), thus linking
the group-theoretic commutator to the Lie bracket A, B 7 AB BA.
Because of this link, we expect the group-theoretic commutator to obey
some nonlinear analogues of the basic Lie bracket identities, and this is
indeed the case. For instance, one easily observes that the commutator is
antisymmetric in the sense that
(2.3)
[g 1 , h] = ([g, h]1 )g
and
(2.5)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
221
Exercise 2.2.1.
(i) If H, K are normal subgroups of G generated by
subsets A H, B K respectively, show that [H, K] is the normal
subgroup of G generated (as a normal subgroup) by the commutators [a, b] with a A, b B.
(ii) If N, H, K are normal subgroups of G, show that [[N, H], K] lies
in the normal subgroup generated by [[H, K], N ] and [[K, N ], H].
(Hint: use the Hall-Witt identity (2.6).)
Given an arbitrary group G, we define the lower central series G =
G1 C G2 . . . of G by setting G1 := G and Gi+1 := [G, Gi ] for all i 1.
Observe that all of the groups Gi in this series are characteristic and thus
normal. A group G is said to be nilpotent of step at most s if Gs+1 is trivial;
in particular, by (2.7), we see that G mod Gs+1 = G/Gs+1 is nilpotent of
step at most s for any group G. We have a basic inclusion:
Exercise 2.2.2 (Filtration property). We have [Gi , Gj ] Gi+j for all i, j
1. (Hint: use Exercise 2.2.1.)
The commutator structure can be clarified by passing to the top order
components of this commutator, as follows. (Strictly speaking, this analysis is not needed to study nilprogressions, but it is still conceptually useful
nonetheless.) Consider the subquotients Vi := Gi mod Gi+1 of the group
G for i = 1, 2, . . . . As Gi+1 contains [Gi , Gi ], we see that each group Vi is
abelian. To emphasise this, we will write Vi additively instead of multiplicatively, thus we shall denote the group operation on Vi as +.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
222
2. Related articles
for g Gi , h Gj .
Proof. We can quotient out by Gi+j+1 , and assume that Gi+j+1 is trivial.
In particular, by Lemma 2.2.2, Gi+1 commutes with Gj , and Gi commutes
with Gj+1 . As such, it is easy to see that if g Gi , h Gj , then [g, h] is
unchanged if one multiplies g (on the left or right) by an element of Gi+1 ,
and similarly for h and Gj+1 . This defines the map [, ]i,j as required.
We refer to the maps [, ]i,j as quotiented commutator maps. The identity
(2.8) asserts that these maps [, ]i,j capture the top order nonlinear behaviour of the group G. As the following exercise shows, these maps behave
like (graded components of) a Lie bracket.
Exercise 2.2.3. Let G be a group with lower central series G = G1 C G2 C
. . . , subquotients Vi , and quotiented commutator maps [, ]i,j : Vi Vj
Vi+j . Let i, j, k 1, let x, x0 Vi , y, y 0 Vj , and z Vk . Establish the
antisymmetry
[x, y]i,j = [y, x]j,i ,
the bihomomorphism properties
[x + x0 , y]i,j = [x, y]i,j + [x0 , y]i,j
[x, y]i,j = [x, y]i,j
[0, y]i,j = 0
and
[x, y + y 0 ]i,j = [x, y]i,j + [x, y 0 ]i,j
[x, y]i,j = [x, y]i,j
[x, 0]i,j = 0
and the Jacobi identity
[[x, y]i,j , z]i+j,k + [[y, z]j,k , x]j+k,i + [[z, x]k,i , y]k+i,j = 0.
Remark 2.2.2. If one wished, one could combine all of these subquotients
Vi and quotiented commutator maps [, ]i,j into a single graded object, namely
the additive group V := i Vi consisting of tuples (vi )
i=1 with vi Vi (and
all but finitely many of the vi trivial), with a bracket [, ] : V V V defined
by
X
[(vi )
[vi , wj ]i,j )
i=1 , (wj )j=1 ] := (
k=1 .
i+j=k
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
223
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
224
2. Related articles
where a01 , . . . , a0R consists of all group elements of the form w(ai1 , . . . , ail )
where w is a formal commutator word of length l i, 1 i1 , . . . , il r, and
each a0j = w(ai1 , . . . , ail ) is associated to the dimension Nj0 := Ni1 . . . Nil
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
225
(note that we allow some of the a0i to be equal, or to degenerate to the identity). Thus the Qi (a1 , . . . , ar ; N1 , . . . , Nr ) are non-increasing in i and become
trivial for i s. We will usually abbreviate Qti (a1 , . . . , ar ; N1 , . . . , Nr ) as Qti .
Trivially we also have
P (a1 , . . . , ar ; N1 , . . . , Nr ) Qt1 .
0
Observe also that the Qti are symmetric, contain the identity, and Qti Qti
0
Qt+t
.
i
O(1)
and h
O(1)
[g, h] Qi+j
and
O(1)
g h Qi
O(1)
O(1)
(a, b; n, m).
bm
O(1)
(a, b; 1, m),
or equivalently
O(1)
(a, b; 1, m).
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
226
2. Related articles
Now we handle the general case. After some relabeling, we may write
w = [u, v], where u, v are words of length l1 , l2 respectively for some l1 , l2 < l
adding up to l, with
nl
nl
and similarly
w(ai1 , . . . , ail ) = [u(ai1 , . . . , ail1 , v(ail1 +1 , . . . , ail )].
By induction hypothesis, one has
nl
O(1)
In particular we have
nl
O(1)
Similarly we have
nl
O(1)
O(1)
Using (2.2.5) and the approximate homomorphism properties of commutators, we conclude that
nl
nl
O(1)
[u(ani11 , . . . , ail 1 ), v(ail 1+1 , . . . , anill )] [u(ai1 , . . . , ail1 )n1 ...nl1 , v(ail1 +1 , . . . , ail )nli +1 ...nl ]Ql+1 ,
+1
[u(ai1 , . . . , ail1 )n1 ...nl1 , v(ail1 +1 , . . . , ail )nli +1 ...nl ] [u(ai1 , . . . , ail1 ), v(ail1 +1 , . . . , ail )]n1 ...nl Ql+1 .
The claim follows.
g n an [a, b]nm Q3
(a, b; n, m)
whenever
O(1)
g a[a, b]m Q3
(a, b; 1, m).
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
227
O(1)
QC
Ql+1
l P
whenever C = O(1).
A similar argument shows that Q1 P for a sufficiently small >
0 depending only on r, s, assuming that N1 , . . . , Nr are sufficiently large
depending on r, s. Since P 2 Q21 , it thus suffices to show that for any
O()
> 0, that Q21 can be covered by O (1) right-translates of Q1 . But this
then follows by iterating the following exercise:
Exercise 2.2.9. Let > 0, l 1, and C = O(1). Show that QC
l can be
O()
O(1)
covered by O (1) right-translates of Ql
Ql+1 . (Hint: this is similar to
into
words w(ai1 , . . . , ail ) of length
Exercise 2.2.5. Factor an element of QC
l
l, together with words of higher length. Gather all the words of length l into
O(1)
monomials w(ai1 , . . . , ail )n with n = O(Ni1 . . . Nil ), times a factor in Ql+1 .
Split these monomials into a monomial with exponent O(Ni1 . . . Nil ), times
a monomial which can take at most O (1) possible values. Then push the
O(1)
latter monomials (and the Ql+1 factor) to the right.)
Exercise 2.2.10 (Polynomial growth). Suppose that a1 , . . . , ar G generate a nilpotent group of step s, and suppose that N1 , . . . , Nr 1. Show
that
|P (a1 , . . . , ar ; N1 , . . . , Nr )| r,s (N1 . . . Nr )Or,s (1) .
Exercise 2.2.11. Suppose that a1 , . . . , ar G generate a nilpotent group
of step s, and suppose that N1 , . . . , Nr are sufficiently large depending on
r, s. Write P := |P (a1 , . . . , ar ; N1 , . . . , Nr )|. Let a01 , . . . , a0R , N10 , . . . , NR0 be
as in (2.10), arranged in non-decreasing order of the length of the associated
formal commutator words.
(i) Show that every element of P can be represented in the form
(a01 )n1 . . . (a0R )nR
for some integers ni = Or,s (Ni0 ). (We do not claim however that
this representation is unique, and indeed the generators a01 , . . . , a0R
are likely to contain quite a lot of redundancy.)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
228
2. Related articles
One can of course define Lie algebras over other fields than the complex
numbers C, but it will be convenieng in this section to work over an algebraically closed field. For most of this text it is preferable to work over the
reals R instead, but in many cases one can pass from the complex theory to
the real theory by complexifying the Lie algebra.
An important special case of the abstract Lie algebras are the concrete
Lie algebras, in which g End(V ) is a (complex) vector space of linear
transformations X : V V on a vector space V (which again can be either
finite or infinite dimensional), and the bilinear form is given by the usual
Lie bracket
[X, Y ] := XY Y X.
It is easy to verify that every concrete Lie algebra is an abstract Lie algebra.
In the converse direction, we have
Theorem 2.3.1. Every abstract Lie algebra is isomorphic to a concrete Lie
algebra.
To prove this theorem, we introduce the useful algebraic tool of the
universal enveloping algebra U (g) of the abstract Lie algebra g. This is
the free (associative, complex) algebra generated by g (viewed as a complex
vector space), subject to the constraints
(2.13)
[X, Y ] = XY Y X.
This algebra is described by the Poincare-Birkhoff-Witt theorem, which asserts that given an ordered basis (Xi )iI of g as a vector space, that a basis
of U (g) is given by monomials of the form
(2.14)
Xia11 . . . Xiamm
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
229
such monomials, one can express their product as a finite linear combination of further monomials of the form (2.14) after repeatedly applying (2.13)
(which we rewrite as XY = Y X + [X, Y ]) to reorder the terms in this product modulo lower order terms until one all monomials have their indices
in the required increasing order. It is then a routine exercise in basic abstract algebra (using all the axioms of an abstract Lie algebra) to verify that
this is multiplication rule on monomials does indeed define a complex associative algebra which has the universal properties required of the universal
enveloping algebra.
The abstract Lie algebra g acts on its universal enveloping algebra
U (g) by left-multiplication: X : M 7 XM , thus giving a map from g to
End(U (g)). It is easy to verify that this map is a Lie algebra homomorphism (so this is indeed an action (or representation) of the Lie algebra),
and this action is clearly faithful (i.e. the map from g to End(U g) is injective), since each element X of g maps the identity element 1 of U (g) to a
different element of U (g), namely X. Thus g is isomorphic to its image in
End(U (g)), proving Theorem 2.3.1.
In the converse direction, every representation : g End(V ) of a Lie
algebra factors through the universal enveloping algebra, in that it extends
to an algebra homomorphism from U (g) to End(V ), which by abuse of
notation we shall also call .
One drawback of Theorem 2.3.1 is that the space U (g) that the concrete
Lie algebra acts on will almost always be infinite-dimensional, even when
the original Lie algebra g is finite-dimensional. However, there is a useful
theorem of Ado that rectifies this:
Theorem 2.3.2 (Ados theorem). Every finite-dimensional abstract Lie algebra is isomorphic to a concrete Lie algebra over a finite-dimensional vector
space V .
Among other things, this theorem can be used (in conjunction with
the Baker-Campbell-Hausdorff formula) to show that every abstract (finitedimensional) Lie group (or abstract local Lie group) is locally isomorphic
to a linear group, a result also known as Lies third theorem; see Section
1.2. It is well-known, though, that abstract Lie groups are not necessarily
globally isomorphic to a linear group, see for instance Exercise 1.1.5 for a
counterexample.
Ados theorem is surprisingly tricky to prove in general, but some special
cases are easy. For instance, one can try using the adjoint representation
ad : g End(g) of g on itself, defined by the action X : Y 7 [X, Y ]; the Jacobi identity (1.8) ensures that this indeed a representation of g. The kernel
of this representation is the centre Z(g) := {X g : [X, Y ] = 0 for all Y
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
230
2. Related articles
g}. This already gives Ados theorem in the case when g is semisimple, in
which case the center is trivial.
The adjoint representation does not suffice, by itself, to prove Ados
theorem in the non-semisimple case. However, it does provide an important reduction in the proof, namely it reduces matters to showing that every finite-dimensional Lie algebra g has a finite-dimensional representation
: g End(V ) which is faithful on the centre Z(g). Indeed, if one has such
a representation, one can then take the direct sum of that representation
with the adjoint representation to obtain a new finite-dimensional representation which is now faithful on all of g, which then gives Ados theorem for
g.
It remains to find a finite-dimensional representation of g which is faithful on the centre Z(g). In the case when g is abelian, so that the centre
Z(g) is all of g, this is again easy, because g then acts faithfully on g C
by the infinitesimal shear maps X : (Y, t) 7 (tX, 0). In matrix form, this
representation identifies each X in this abelian Lie algebra with an uppertriangular matrix:
0 X
X
.
0 0
This construction gives a faithful finite-dimensional representation of
the centre Z(g) of any finite-dimensional Lie algebra. The standard proof of
Ados theorem (which I believe dates back to work of Harish-Chandra) then
proceeds by gradually extending this representation of the centre Z(g) to
larger and larger sub-algebras of g, while preserving the finite-dimensionality
of the representation and the faithfulness on Z(g), until one obtains a representation on the entire Lie algebra g with the required properties. (For
technical inductive reasons, one also needs to carry along an additional property of the representation, namely that it maps the nilradical to nilpotent
elements, but we will discuss this technicality later.)
This procedure is a little tricky to execute in general, but becomes simpler in the nilpotent case, in which the lower central series g1 := g; gn+1 :=
[g, gn ] becomes trivial for sufficiently large n:
Theorem 2.3.3 (Ados theorem for nilpotent Lie algebras). Let n be a
finite-dimensional nilpotent Lie algebra. Then there exists a finite-dimensional
faithful representation : n End(V ) of n. Furthermore, there exists a natural number k such that (n)k = {0}, i.e. one has (X1 ) . . . (Xk ) = 0 for
all X1 , . . . , Xk n.
The second conclusion of Ados theorem here is useful for induction
purposes. (By Engels theorem, this conclusion is also equivalent to the
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
231
assertion that every element of (n) is nilpotent, but we can prove Theorem
2.3.3 without explicitly invoking Engels theorem.)
Below the fold, we give a proof of Theorem 2.3.3, and then extend the
argument to cover the full strength of Ados theorem. The presentation here
is based on the one in in [FuHa1991].
2.3.1. The nilpotent case. We first prove Theorem 2.3.3. We achieve this
by an induction on the dimension of n. The claim is trivial for dimension
zero, so we assume inductively that n has positive dimension, and that
Theorem 2.3.3 has already been proven for all lower-dimensional nilpotent
Lie algebras.
As noted earlier, the adjoint representation already verifies the claim
except for the fact that it is not faithful on the centre Z(n). (The nilpotency
of n ensures the existence of some k for which ad(n)k = 0.) Thus, it will
suffice to find a finite-dimensional representation : n End(V ) that is
faithful on the center Z(n), and for which (n)k = {0} for some k.
We have already verified the claim when n is abelian (and can take k = 1
in this case), so suppose that n is not abelian; in particular, the centre Z(n)
has strictly smaller dimension. We then observe that n must contain a codimension one ideal a containing Z(n), which will of course again be a nilpotent
Lie algebra. This can be seen by passing to the quotient n0 := n/Z(n) (which
has positive dimension) and then to the abelianisation n0 /[n0 , n0 ] (which still
has positive dimension, by nilpotency), arbitrarily selecting a codimension
one subspace of that abelianisation, and then passing back to n.
If we let h be a one-dimensional complementary subspace of a, then h is
automatically an abelian Lie algebra, and we have the decomposition
n=ah
as vector spaces. This is not necessarily a direct sum of Lie algebras, though,
because a and h need not commute. However, as a is an ideal, we do know
that [h, a] a, thus there is an adjoint action ad : h End(a) of h on a.
By induction hypothesis, we know that there is a faithful representation
0 : a End(V0 ) on some finite-dimensional space V0 with 0 (a)k0 = {0}
for some k0 . We would like to somehow extend this representation to a
finite-dimensional representation : n End(V ) which is still faithful on a
(and hence on Z(n)), and with (n)k = {0} for some k.
To motivate the construction, let us begin not with 0 , but with the
universal enveloping algebra representation A : M 7 AM of a on U (a). The
idea is to somehow combine this action a with an action of h on the same
space U (a) to obtain an action of the full Lie algebra n on U (a). To do this,
recall that we have the adjoint action ad(H) : A 7 [H, A] of h on a. This
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
232
2. Related articles
[H, A1 . . . Am ] :=
m
X
i=1
one can easily verify that this is indeed an action (and furthermore that
the map ad(H) : M 7 [H, M ] is a derivation on U (a)). One can then
combine the multiplicative action A : M 7 AM of a and the adjoint action
H : M 7 [H, M ] of h into a joint action
(2.16)
(A + H) : M 7 AM + [H, M ]
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
233
From the Leibniz rule (2.15) we see that the adjoint action of h on U (a)
preserves I, and thus descends to an action on U (a)/I. We can combine this
action with the multiplicative action of a as before using (2.16) to create an
action of n on U (a)/I, which is now a finite-dimensional representation which
(as observed previously) is faithful on a.
The only remaining thing to show is that for some sufficiently large k,
that nk annihilates U (a)/I. By linearity, it suffices to show that
X1 . . . Xk A1 . . . Am = 0 mod I
whenever X1 , . . . , Xk lie in either a or h, and A1 . . . Am is a monomial in
U (a). But observe that if one multiplies a monomial A1 . . . Am by an element
A of a, then one gets a monomial AA1 . . . Am of one higher degree; and if
instead one multiplies a monomial A1 . . . Am by an element H of h, then from
(2.15) one gets a sum of monomials in which one of the terms Ai has been
replaced with the higher order term [H, Ai ]. Using the nilpotency of n
(which implies that all sufficiently long iterated commutators must vanish),
we thus see that if k is large enough, then X1 . . . Xk A1 . . . Am will consist
only of terms of degree at least k0 , which automatically lie in I, and the
claim follows.
Remark 2.3.5. The above argument gives an effective (though not particularly efficient) bound on the dimension of V and on the order k of (n) in
terms of the dimension of n. Similarly for the arguments we give below to
prove more general versions of Ados theorem.
2.3.2. The solvable case. To go beyond the nilpotent case one needs to
use more of the structural theory of Lie algebras. In particular, we will use
Engels theorem:
Theorem 2.3.6 (Engels theorem). Let g End(V ) be a concrete Lie
algebra on a finite-dimensional space V which is concretely nilpotent in the
sense that every element of g is a nilpotent operator on V . Then there exists
a basis of V such that g consists entirely of upper-triangular matrices. In
particular, if d is the dimension of V , then gd = {0} (i.e. X1 . . . Xd = 0 for
all X1 , . . . , Xd g).
The proof of Engels theorem is not too difficult, but we omit it here
as we have nothing to add to the textbook proofs of this theorem (such as
the one in [FuHa1991]). Note that this theorem, combined with Theorem
2.3.3, gives a correspondence between concretely nilpotent Lie algebras and
abstractly nilpotent Lie algebras.
Engels theorem has the following consequence:
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
234
2. Related articles
k1
X
i=1
But [A, B] lies in [a, b], and the claim now follows from Corollary 2.3.7.
We can apply this corollary to the radical r of an (abstract) finitedimensional Lie algebra g, defined as the maximal solvable ideal of the Lie
algebra. (One can easily verify that the vector space sum of two solvable
ideals is again a solvable ideal, which implies that the radical is well-defined.)
Theorem 2.3.9. Let g be a finite-dimensional (abstract) Lie algebra, and
let r be a solvable ideal in g. Then [g, g] r is an (abstractly) nilpotent ideal
of g.
Proof. Without loss of generality we may take r to be the radical of g.
Let us first verify the theorem in the case when g is a concrete Lie algebra
over a finite-dimensional vector space V . We let r(i+1) := [r(i) , r(i) ] be the
derived series of r. One easily verifies that each of the r(i) are ideals of g.
We will show by downward induction on i that the ideals [g, g] r(i) are
concretely nilpotent. As r is solvable, this claim is trivial for i large enough.
Now suppose that i is such that the ideal [g, g]r(i+1) is concretely nilpotent.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
235
This ideal contains [r(i) , [r(i) , g]], which is then also concretely nilpotent. By
Corollary 2.3.8, this implies that the ideal [r(i) , g] is concretely nilpotent,
and hence the smaller ideal [g, [g, g] r(i) ] is also concretely nilpotent. By
a second application of Corollary 2.3.8, we conclude that [g, g] r(i) is also
concretely nilpotent, closing the induction. This proves the theorem in the
concrete case.
To establish the claim in the abstract case, we simply use the adjoint
representation, which effectively quotients out the centre Z(g) of g (which
will also be an ideal of r) to convert an finite-dimensional abstract Lie algebra
into a concrete Lie algebra over a finite-dimensional space. We conclude that
the quotient of [g, g] r by the central ideal Z(g) is nilpotent, which implies
that [g, g] r is nilpotent as required.
This has the following corollary. Recall that a derivation D : g g on an
abstract Lie algebra g is a linear map such that D[X, Y ] = [DX, Y ]+[X, DY ]
for each X, Y g. Examples of derivations include the inner derivations
DX := [A, X] for some fixed A g, but not all derivations are inner.
Corollary 2.3.10. Let g be a finite-dimensional (abstract) Lie algebra, and
let r be a solvable ideal in g. Then for any derivation D : g g, Dr is
nilpotent.
Proof. If D were an inner derivation, the claim would follow easily from
Theorem 2.3.9. In the non-inner case, the trick is simply to view D as an
inner derivation coming from an extension of g. Namely, we work in the
enlarged Lie algebra g oD C, which is the Cartesian product g C with Lie
bracket
[(X, s), (Y, t)] := ([X, Y ] + sDy tDx, 0).
One easily verifies that this is an (abstract) Lie algebra which contains g
g {0} as a subalgebra. The derivation D on g then arises from the inner
derivation on g oD C coming from (0, 1). The claim then easily follows by
applying Theorem 2.3.9 to this enlarged algebra.
Remark 2.3.11. I would be curious to know if there is a more direct proof
of Corollary 2.3.10 (which in particular did not need the full strength of
Engels theorem), as this would give a simpler proof of Ados theorem in the
solvable case (indeed, it seems almost equivalent to that theorem, especially
in view of Lies third theorem, Exercise 1.2.23).
Define the nilradical of an (abstract) finite-dimensional Lie algebra g to
be the maximal nilpotent ideal in g. One can verify that the sum of two
nilpotent ideals is again a nilpotent ideal, so the nilradical is well-defined.
One can verify that the radical r is characteristic (i.e. preserved by all
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
236
2. Related articles
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
237
Leibniz rule (cf. (2.15)), we thus see that these derivations preserve I, and
thus (by the Leibniz rule) also preserve I k . Thus this also gives an action
of h on U (a)/I k by derivations.
In analogy with (2.16), we may now combine the actions of a and h on
U (a)/I k to an action of r which will remain faithful and nilpotent on n, and
the claim follows.
2.3.3. The general case. Finally, we handle the general case. Actually
we can use basically the same argument as in preceding cases, but we need
one additional ingredient, namely Levis theorem:
Theorem 2.3.13 (Levis theorem). Let g be a finite-dimensional (abstract)
Lie algebra. Then there exists a splitting g = r h as vector spaces, where
r is the radical of g and h is another Lie algebra.
We remark that h necessarily has trivial radical and is thus semisimple.
It is easy to see that the quotient g/r is semisimple; the entire difficulty of
Levis theorem is to ensure that one can lift this quotient back up into the
original space g. This requires some knowledge of the structural theory of
semisimple Lie algebras. The proof will be omitted, as I have nothing to
add to the textbook proofs of this result (such as the one in [FuHa1991]).
Using Levis theorem and Theorem 2.3.12, one obtains the full Ado
theorem:
Theorem 2.3.14 (Ados theorem for general Lie algebras). Let g be a finitedimensional Lie algebra. Then there exists a finite-dimensional faithful representation : g End(V ) of g. Furthermore, if n is the nilradical of r,
one can ensure that (n) is nilpotent.
Exercise 2.3.1. Prove Theorem 2.3.14 (and thus also Theorem 2.3.2).
(Hint: deduce this theorem from Theorem 2.3.12 and Theorem 2.3.13 by
using the same argument used to deduce Theorem 2.3.12 from the induction
hypothesis, with the key point again being Corollary 2.3.10 that places the
adjoint action of h on r inside n.)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
238
2. Related articles
where Adx := exp(adx ), adx : g g is the adjoint map adx (y) := [x, y], and
log z
, which is analytic for z near 1. Similarly,
FR is the function FR (z) := zz1
define the left Baker-Campbell-Hausdorff-Dynkin law
Z 1
(2.18)
Lx (y) := y +
FL (Adtx Ady )x dt
0
z
where FL (z) := log
z1 . One easily verifies that these expressions are welldefined (and depend smoothly on x and y) when x and y are sufficiently
small.
x y := Ry (x) = Lx (y)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
239
The key is to observe that the right and left BCH laws commute with
each other:
Proposition 2.4.2 (Commutativity). Let g be a finite-dimensional Lie algebra. Then for sufficiently small x, y, z, one has
(2.20)
Note that this commutativity has to hold if (2.19) is to be both welldefined and associative. Assuming Proposition 2.4.2, we can set x = 0 in
(2.20) and use the easily verified identities Rz (0) = z, Ly (0) = y to conclude
that Ly (z) = Rz (y) for small y, z, ensuring that (2.19) is well-defined; and
then inserting (2.19) into (2.20) we obtain the desired (local) associativity.
It remains to prove Proposition 2.4.2. We first make a convenient observation. Thanks to the Jacobi identity, the adjoint representation ad : x 7
adx is a Lie algebra homomorphism from g to the Lie algebra gl(g). As this
latter Lie algebra is the Lie algebra of a Lie group, namely GL(g), the BakerCampbell-Hausdorff formula is valid for that Lie algebra. In particular, one
has
log(exp(adx ) exp(ady )) = Rady (adx ) = Ladx (ady )
for sufficiently small x, y. But as ad is a Lie algebra homomorphism, one
has
Rady (adx ) = adRy (x)
and similarly
Ladx (ady ) = adLx (y) .
Exponentiating, we conclude that
(2.21)
This would already give what we want if the adjoint representation was
faithful. We unfortunately cannot assume this (and this is the main reason,
by the way, why Ados theorem is so difficult), but we can at least use (2.21)
to rewrite the formulae (2.17), (2.18) as
Z 1
Ry (x) = x +
FR (AdRty (x) ) dt
0
and
Z
Lx (y) := y +
0
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
240
2. Related articles
for fixed small y, z, a large natural number n, and with the understanding
that the operations are only applied to sufficiently small elements x. From
n , and so a large
radial homogeneity we have Ly = Lny/n and Rz = Rz/n
number (O(n2 ), to be more precise) of iterations of (2.23) (using uniform
smoothness to control all errors) gives
Ly Rz = Rz Ly + O(1/n),
and the claim (2.20) then follows by sending n .
It remains to prove (2.22). When y = 0, then Ly is the identity map and
the claim is trivial; similarly if z = 0. By Taylor expansion, it thus suffices
to establish the infinitesimal commutativity law
Lay (Rbz (x))|a=b=0 =
Rbz (Lay (x))|a=b=0 .
a b
a b
(One can interpret this infinitesimal commutativity as a commutativity of
the vector fields corresponding to the infinitesimal generators of the left
and right BCH laws, although we will not explicitly adopt that perspective
here.) This is a simplification, because the infinitesimal versions of (2.17),
(2.18) are simpler than the non-infinitesimal versions. Indeed, from the
fundamental theorem of calculus one has
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
241
for any smoothly varying invertible matrix function A(t) of a real parameter
t. Using this identity, we can write the left-hand side of (2.24) as
1
F (AdRbz (x) )|b=0 )FL (Adx )y.
b L
If we write Y := FL (Adx )y and Z := FR (Adx )z, then from Taylor expansion
we have
Rbz (x) = x + bZ + O(|b|2 )
FL (Adx )(
1
F (Adx+bZ )|b=0 )Y.
b L
1
F (Adx+aY )|a=0 )Z.
a R
1
FL (Adx+bZ )|b=0 )Y = Adx ( FR1 (Adx+aY )|a=0 )Z.
b
a
Now, we write
Z 1
1
Adtx dt
FL (Adx ) =
(2.25)
and
FR1 (Adx )
Z
=
Adtx dt
0
(2.26)
( Adtx+tbZ )|b=0 Y dt = Adx
( AdtxtaY )|a=0 Z dt.
b
b
0
0
We write Ad as the exponential of ad. Using the Duhamel matrix identity
Z 1
d
exp(A(t)) =
exp(sA(t))A0 (t) exp((1 s)A(t)) dt
dt
0
for any smoothly varying matrix function A(t) of a real variable t, together
with the linearity of ad, we see that
Z 1
( Adtx+tbZ )|b=0 =
Adstx t adZ Ad(1s)tx ds
b
0
and similarly
( AdtxtaY )|a=0 =
b
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
242
2. Related articles
For any x g, the adjoint map adx : g g is a derivation in the sense that
adx [y, z] = [adx y, z] + [y, adx z],
thanks to the Jacobi identity. Exponentiating, we conclude that
Adx [y, z] = [Adx y, Adx z]
(thus each Adx is a Lie algebra homomorphism) and thus
Adx ady Ad1
x = adAdx y .
Using this, we can simplify (2.27) as
Z 1Z 1
Z
adAdstx Z Adtx Y tdsdt =
0
1Z 1
1Z 1
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
243
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
244
2. Related articles
one easily verifies that this gives U the structure of a local group (which
we will sometimes call G U to emphasise the original group G). If U is
symmetric (i.e. U 1 = U ), then we in fact have a symmetric local group.
One can also restrict local groups G to open neighbourhoods U to obtain a
smaller local group G U by the same procedure (adopting the convention
that statements such as g h U or g 1 U are considered false if the
left-hand side is undefined). (Note though that if one restricts to non-open
neighbourhoods of the identity, then one usually does not get a local group;
for instance [1, 1] is not a local group (why?).)
Finite subsets of (Hausdorff) groups containing the identity can be viewed
as local groups. This point of view turns out to be particularly useful for
studying approximate groups in additive combinatorics, a point which I
hope to expound more on later. Thus, for instance, the discrete interval
{9, . . . , 9} Z is an additive symmetric local group, which informally
might model an adding machine that can only handle (signed) one-digit
numbers. More generally, one can view a local group as an object that behaves like a group near the identity, but for which the group laws (and in
particular, the closure axiom) can start breaking down once one moves far
enough away from the identity.
In many situations (such as when one is investigating the local structure
of a global group) one is only interested in the local properties of a (local or
global) group. We can formalise this by the following definition. Let us call
two local groups G = (G, , , 1G , , ()1 ) and G0 = (G0 , 0 , 0 , 1G0 , , ()1 )
locally identical if they have a common restriction, thus there exists a set
U G G0 such that G U = G0 U (thus, 1G = 1G0 , and the topology
and group operations of G and G0 agree on U ). This is easily seen to be
an equivalence relation. We call an equivalence class [G] of local groups a
group germ.
Let P be a property of a local group (e.g. abelianness, connectedness,
compactness, etc.). We call a group germ locally P if every local group in
that germ has a restriction that obeys P; we call a local or global group G
locally P if its germ is locally P (or equivalently, every open neighbourhood
of the identity in G contains a further neighbourhood that obeys P). Thus,
the study of local properties of (local or global) groups is subsumed by the
study of group germs.
Exercise 2.5.1.
(i) Show that the above general definition is consistent with the usual
definitions of the properties connected and locally connected
from point-set topology.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
245
(ii) Strictly speaking, the above definition is not consistent with the
usual definitions of the properties compact and local compact
from point-set topology because in the definition of local compactness, the compact neighbourhoods are certainly not required to be
open. Show however that the point-set topology notion of locally
compact is equivalent, using the above conventions, to the notion
of locally precompact inside of an ambient local group. Of course,
this is a much more clumsy terminology, and so we shall abuse notation slightly and continue to use the standard terminology locally
compact even though it is, strictly speaking, not compatible with
the above general convention.
(iii) Show that a local group is locally discrete if and only if it is locally
trivial.
(iv) Show that a connected global group is abelian if and only if it is
locally abelian. (Hint: in a connected global group, the only open
subgroup is the whole group.)
(v) Show that a global topological group is first-countable if and only
if it is locally first countable. (By the Birkhoff-Kakutani theorem
(Theorem 1.5.2), this implies that such groups are metrisable if and
only if they are locally metrisable.)
Let p be a prime. Show that the solenoid group Zp R/Z , where
Zp is the p-adic integers and Z := {(n, n) : n Z} is the diagonal embedding of Z inside Zp R, is connected but not locally
connected.
Remark 2.5.1. One can also study the local properties of groups using
nonstandard analysis. Instead of group germs, one works (at least in the
case when G is first countable) with the monad o(G) of the identity element
1G of G, defined as the nonstandard group elements g = limn gn in G
that are infinitesimally close to the origin in the sense that they lie in every
standard neighbourhood of the identity. The monad o(G) is closely related
to the group germ [G], but has the advantage of being a genuine (global)
group, as opposed to an equivalence class of local groups. It is possible
to recast most of the results here in this nonstandard formulation; see e.g.
[Ro1966]. However, we will not adopt this perspective here.
A useful fact to know is that Lie structure is local; see Exercise 1.2.17.
As with so many other basic classes of objects in mathematics, it is of
fundamental importance to specify and study the morphisms between local
groups (and group germs). Given two local groups G, G0 , we can define the
notion of a (continuous) homomorphism : G G0 between them, defined
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
246
2. Related articles
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
247
Show that the additive global groups R/Z and R are locally isomorphic.
Show that every locally path-connected group G is locally isomorphic to a
path-connected, simply connected group.
2.5.1. Lies third theorem. Lies fundamental theorems of Lie theory
link the Lie group germs to Lie algebras. Observe that if [G] is a locally Lie
group germ, then the tangent space g := T1 G at the identity of this germ is
well-defined, and is a finite-dimensional vector space. If we choose G to be
symmetric, then g can also be identified with the left-invariant (say) vector
fields on G, which are first-order differential operators on C (M ). The Lie
bracket for vector fields then endows g with the structure of a Lie algebra. It
is easy to check that every morphism : [G] [H] of locally Lie germs gives
rise (via the derivative map at the identity) to a morphism D(1) : g h
of the associated Lie algebras. From the Baker-Campbell-Hausdorff formula
(which is valid for local Lie groups, as discussed in this previous post) we
conversely see that D(1) uniquely determines the germ homomorphism .
Thus the derivative map provides a covariant functor from the category of
locally Lie group germs to the category of (finite-dimensional) Lie algebras.
In fact, this functor is an isomorphism:
Exercise 2.5.4 (Lies third theorem). For this exercise, all Lie algebras are
understood to be finite dimensional (and over the reals).
(i) Show that every Lie algebra g is the Lie algebra of a local Lie group
germ [G], which is unique up to germ isomorphism (fixing g).
(ii) Show that every Lie algebra g is the Lie algebra of some global
connected, simply connected Lie group G, which is unique up to
Lie group isomorphism (fixing g).
(iii) Show that every homomorphism : g h between Lie algebras
is the derivative of a unique germ homomorphism : [G] [H]
between the associated local Lie group germs.
(iv) Show that every homomorphism : g h between Lie algebras
is the derivative of a unique Lie group homomorphism : G
H between the associated global connected, simply connected, Lie
groups.
(v) Show that every local Lie group germ is the germ of a global connected, simply connected Lie group G, which is unique up to Lie
group isomorphism. In particular, every local Lie group is locally
isomorphic to a global Lie group.
(Hint: use Exercise 1.2.23.)
Lies third theorem (which, actually, was proven in full generality by
Cartan) demonstrates the equivalence of three categories: the category of
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
248
2. Related articles
finite-dimensonal Lie algebras, the category of local Lie group germs, and
the category of connected, simply connected Lie groups.
2.5.2. Globalising a local group. Many properties of a local group improve after passing to a smaller neighbourhood of the identity. Here are
some simple examples:
Exercise 2.5.5. Let G be a local group.
(i) Give an example to show that G does not necessarily obey the
cancellation laws
(2.28)
gk = hk = g = h;
kg = kh = g = h
(gh)1 = h1 g 1
for any g, h G for which both sides are well-defined.
(iii) Repeat the previous part, but with the inversion law replaced by
the involution law
(2.30)
(g 1 )1 = g
for any g for which the left-hand side is well-defined.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
249
come from a global group. Indeed, it is not hard to see that if (1, 1)3 /
is the restriction of a global group G, then G must be a Lie group with
Lie algebra R3 (by Exercise 1.2.17), and so the connected component G
of G containing the identity is isomorphic to R3 / for some sublattice of
R3 that contains a1 , a2 , a3 , a4 ; but for generic a1 , a2 , a3 , a4 , there is no such
lattice, as the ai will generate a dense subset of R3 . (The situation here is
somewhat analogous to a number of famous Escher prints, such as Ascending and Descending, in which the geometry is locally consistent but globally
inconsistent.) We will give this sort of argument in more detail later, when
we prove Proposition 2.5.6.
Nevertheless, the space (1, 1)3 / is still locally isomorphic to a global
Lie group, namely R3 ; for instance, the open neighbourhood (0.5, 0.5)3 /
is isomorphic to (0.5, 0.5)3 , which is an open neighbourhood of R3 . More
generally, Lies third theorem tells us that any local Lie group is locally
isomorphic to a global Lie group.
Let us call a local group globalisable if it is locally isomorphic to a
global group; thus Lies third theorem tells us that every local Lie group
is globalisable. Thanks to Goldbrings solution [Go2010] to the local version of Hilberts fifth problem (Theorem 1.5.24), we also know that locally
Euclidean local groups are globalisable. A modification of this argument
[vdDrGo2010] shows in fact that every locally compact local group is globalisable.
In view of these results, it is tempting to conjecture that all local groups
are globalisable;; among other things, this would simplify the proof of Lies
third theorem (and of the local version of Hilberts fifth problem). Unfortunately, this claim as stated is false:
Theorem 2.5.2. There exists local groups G which are not globalisable.
The counterexamples used to establish Theorem 2.5.2 are remarkably
delicate; the first example I know of appears in [vEsKo1964]. One reason
for this, of course, is that the previous results prevents one from using any
local Lie group, or even a locally compact group as a counterexample. We
will present a (somewhat complicated) example shortly, based on the unit
ball in the infinite-dimensional Banach space ` (N2 ).
Despite such counterexamples, there are certainly many situations in
which we can globalise a local group. For instance, this is the case if one has
a locally faithful representation of that local group inside a global group:
Lemma 2.5.3 (Faithful representation implies globalisability). Let G be
a local group, and suppose there exists an injective local homomorphism
: U H from G into a global topological group H with U symmetric.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
250
2. Related articles
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
251
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
252
2. Related articles
each x F/N . By arguing as in the proof of Lemma 2.5.3 one can verify
that this generates a topology that makes F/N a topological group, and
makes G homeomorphic to (G); we leave the details as an exercise.
Exercise 2.5.6. Complete the proof of the above proposition.
Exercise 2.5.7. Use Proposition 2.5.5 to give an alternate proof of Proposition 2.5.3.
2.5.3. A non-globalisable group. We now prove Theorem 2.5.2. We
begin with a preliminary construction, which gives a local group that has a
fixed small neighbourhood that cannot arise from a global group.
Proposition 2.5.6 (Preliminary counterexample). For any m 1, there
exists an equivalence relation m on the open unit ball B(0, 1) of ` (N)
which gives B(0, 1)/ m the structure of a local group, but such that B(0, 1/m)/ m
is not isomorphic (even as a discrete local group) to a subset of a global
group.
Proof. We will use a probabilistic construction, mimicking the three-dimensional
example (1, 1)3 / from the introduction. Fix m, and let N be a large
integer (depending on m) to be chosen later. We identify the N -dimensional
cube (1, 1)N with the unit ball in ` ({1, . . . , N }), which embeds in the unit
ball in ` (N) via extension by zero. Let b1 , . . . , bN +1 {1, 1}N be randomly chosen corners of this cube. A simple application of the union bound
shows that with probability approaching 1 as N , we have bi 6= bj for
all i < j, but also that for any 1 i1 < < i100m N and any choice
of signs 1 , . . . , 100m , the vectors 1 bi1 , . . . , 100m bi100m agree on at least one
coordinate. As a corollary of this, we see that
k
N
+1
X
ni bi k` =
N
+1
X
i=1
|ni |
i=1
PN +1
i=1
|ni | 100m.
N
+1
X
i=1
ni ai k` =
N
+1
X
|ni | + O()
i=1
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
253
(2.32)
N +1
1 X
ni ai
2m
i=1
PN +1
i=1
i=1
1
We claim that 2m
(n1 a1 + + nN +1 aN +1 ) lies in the kernel of , contradicting the description (2.33) of that description (and the linear independence of the ai ). To see this, we observe for a sufficiently large natural
number M that the local homomorphism property (and the commutativity)
gives
NY
nN +1
+1
1
1
(n1 a1 + + nN +1 aN +1 ) =
aN +1
2mM
2mM
i=1
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
254
2. Related articles
.
(n1 a1 + + nN +1 aN +1 ) =
aN +1
2m
2mM
i=1
aN +1
=
aN +1 = 1,
2mM
2m
and the claim follows.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
255
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
256
2. Related articles
it; this basic trick is helpful elsewhere in the theory, resolving a number of
cohomological issues in topological group theory. The result can be generalised to show in fact that arbitrary (topological) extensions of Lie groups
by Lie groups remain Lie; this was shown in [Gl1950]. However, the above
special case of this result is already sufficient (in conjunction with the rest
of the theory, of course) to resolve Hilberts fifth problem.
Remark 2.6.2. We have shown in the above discussion that every connected Lie group is a central extension (by an abelian Lie group) of a Lie
group with a faithful continuous linear representation. It is natural to ask
whether this central extension is necessary. Unfortunately, not every connected Lie group admits a faithful continuous linear representation; see Exercise 1.1.5. (On the other hand, the group G in that example is certainly
isomorphic to the extension of the linear group R2 by the abelian group
R/Z.)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
257
The group G can then be described, as a set, as the disjoint union of its
fibres (h) + K for h H:
]
G=
(h) + K.
hH
Using the associative law for G, we see that must obey the cocycle equation
(2.35)
for all h1 , h2 , h3 K; we refer to functions that obey this equation as cocycles (the reason for this terminology being explained in [Ta2009, 1.13]).
2 (H, K), if we
The space of all such cocyles is denoted Z 2 (H, K) (or Zdisc
wish to emphasise that we are working for now in the discrete category,
as opposed to the measurable, topological, or smooth category); this is an
abelian group with respect to pointwise addition. Using (2.34), we also have
a description of the group identity 1G of G in coordinates:
(2.36)
1G = (1H ) (1H , 1H ).
Indeed, this can be seen by noting that 1G is the unique solution of the
group equation g g = g. Similarly, we can describe the inverse operation in
H K coordinates:
(2.37)
Thus we see that the cocycle , together with the group structures on
H and K, capture all the group-theoretic structure of G. Conversely, given
any cocycle Z 2 (H, K), one can place a group structure on the set H K
by declaring the multiplication law as
(h1 , k1 ) (h2 , k2 ) := (h1 h2 , k1 + k2 + (h1 , h2 )),
the identity element as (1H , (1H , 1H )), and the inversion law as
(h, k)1 := h1 , k (h, h1 ) (1H , 1H ) .
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
258
2. Related articles
Exercise 2.6.1. Using the cocycle equation (2.35), show that the above
operations do indeed yield group structure on H K (i.e. the group axioms
are obeyed). If arises as the cocycle associated to a section of a group
extension G, we then see from (2.34), (2.36), (2.37) that this group structure
we have just placed on H K is isomorphic to that on G.
Thus we see that once we select a section, we can describe a central
extension of K by H (up to group isomorphism) as a cocycle, and conversely
every cocycle arises in this manner. However, we have some freedom in
deciding how to select this section. Given one section : H G, any other
section 0 : H G takes the form
0 (h) = (h) + f (h)
for some function f : H K; conversely, every such function f can be used
to shift a section to a new section 0 . We refer to such functions f as
gauge functions. The cocycles , 0 associated to , 0 are related by the
gauge transformation
0 (h1 , h2 ) = (h1 , h2 ) + df (h1 , h2 )
where df : H H K is the function
df (h1 , h2 ) := f (h1 ) + f (h2 ) f (h1 h2 ).
We refer to functions of the form df as coboundaries, and denote the space
2 (H, K), if we want to emphaof all such coboundaries as B 2 (H, K) (or Bdisc
sise the discrete nature of these coboundaries). One easily verifies that all
coboundaries are cocycles, and so B 2 (H, K) is a subgroup of Z 2 (H, K). We
2 (H, K)) to be
then define the second group cohomology H 2 (H, K) (or Hdisc
the quotient group
H 2 (H, K) := Z 2 (H, K)/B 2 (H, K),
and refer to elements of H 2 (H, K) as cohomology classes. (There are higher
order group cohomologies, which also have some relevance for the extension
problem, but will not be needed here; see [Ta2009, 1.13] for further discussion. The first group cohomology H 1 (H, K) = Hom(H, K) is just the
space of homomorphisms from H to K; again, this has some relevance for
the extension problem but will not be needed here.)
We call two cocycles cohomologous if they differ by a coboundary (i.e.
they lie in the same cohomology class). Thus we see that different sections of
a single central group extension provide cohomologous cocycles. Conversely,
if two cocycles are cohomologous, then the group structures on H K given
by these cocycles are easily seen to be isomorphic; furthermore, if we restrict
group isomorphism to fix each fibre of , this is the only way in which group
structures generated by such cocycles are isomorphic. Thus, we see that up
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
259
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
260
2. Related articles
where
f (h) := Eh3 H (h, h3 ) =
1 X
(h, h3 )
|H|
h3 H
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
261
problem, which is instead more concerned with the local topological and Lie
structure of groups.
Because of this, we shall work with local sections and cocycles instead
of global ones. A local section in a topological group extension
0KGH0
is a continuous map : U G from an open neighbourhood U of the
identity 1H in H to G which is a right inverse of on U , thus (h) 1 (h)
for all h U . (It is not yet obvious why local sections exist at all - there
may still be local obstructions to trivialising the fibre bundle - but certainly
this task should be easier than that of locating global continuous sections.)
Similarly, a local cocycle is a continuous map : U U K defined
on some open neighbourhood U of the identity in H that obeys the cocycle
equation (2.35) whenever h1 , h2 , h3 U are such that h1 h2 , h2 h3 U . We
consider two local sections : U G, 0 : U 0 G to be locally identical if
there exists a neighbourhood U 00 of the identity contained in both U and U 0
such that and 0 agree on U 00 . Similarly, we consider two local cocycles
: U U G and 0 : U 0 U 0 G to be locally identical if there is a
neighbourhood U 00 of the identity in U U 0 such that and 0 agree on
2
U 00 U 00 . We let Ztop,loc
(H, K) be the space of local cocycles, modulo local
identity; this is easily seen to be an abelian group.
One easily verifies that every local section induces a local cocycle (perhaps after shrinking the open neighbourhood U slightly), which is welldefined up to local identity. Conversely, the computations used to show
Exercise 2.6.1 show that every local cocycle creates a local group structure
on H K.
If U is an open neighbourhood the identity of H, f : U K is a continuous gauge function, and U 0 is a smaller neighbourhood such that (U 0 )2 U ,
we can define the local coboundary df : U 0 U 0 K by the usual formula
df (h1 , h2 ) := f (h1 ) + f (h2 ) f (h1 h2 ).
2
This is well-defined up to local identity. We let Btop,loc
(H, K) be the space of
2
local coboundaries, modulo local identity; this is a subgroup of Ztop,loc
(H, K).
2
Thus, one can form the local topological group cohomology Htop,loc (H, K) :=
2
2
Ztop,loc
(H, K)/Btop,loc
(H, K). We say that two local cocycles are locally cohomologous if they differ by a local coboundary.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
262
2. Related articles
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
263
(k + k 0 ) = (k)(k 0 )
for all k, k 0 K, the extension F0 need not obey any analogous symmetry
in general. However, this can be rectified by an averaging argument. Define
2
the function F1 : G Cn by the formula
Z
F1 (g) :=
(k 0 )1 F0 (g + k 0 ) dK (k 0 ),
K
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
264
2. Related articles
the equivariance
F2 (g + k) = (k)F2 (g)
for g G and k K that are sufficiently close to the identity. From this, we
see that for each h H that is sufficiently close to the identity, there exists
a unique solution (h) in 1 (h) to the equation F2 (g) = I that is also close
to the identity; from the continuity of F2 we see that is continuous near
the identity and is thus a local section of G as required.
Now we argue in the general case, in which the abelian Lie group K need
not be compact. If K is connected, then it is isomorphic to a Euclidean space
quotiented by a discrete subgroup. In particular, it can be quotiented down
to a torus in a manner which is a local homeomorphism near the origin.
Embedding that torus in a unitary group U (n), we obtain an continuous
homomorphism : K U (n) which is still locally injective and has compact
image. One can then repeat the previous arguments to obtain a local section;
we omit the details.
Finally, if K is not connected, we can work with the connected component K of the identity, which is locally the same as K, and repeat the
previous argument again to obtain a local section.
Using this local section, we obtain a local cocycle : U U K; shifting
by a constant, we may assume that (1H , 1H ) = 0. Near the identity, the
abelian Lie group K can be locally identified with a Euclidean space Rn , so
(shrinking U if necessary) induces a Rn -valued local cocycle : U U
Rn .
We will smooth this cocycle by an averaging argument akin to the one
used in Lemma 2.6.4. Instead of a discrete averaging, though, we will now
use a non-trivial left-invariant Haar measure H on H (the existence of
which is guaranteed by Theorem 1.4.6). As we only have a local cocycle,
though, we will need to only average using a smooth, compactly supported
function : H R+ supported on a small Rneighbourhood U 0 of U (small
enough that (U 0 )3 U ), normalised so that H (h3 ) dH (h3 ) = 1. For any
h1 , h2 , h3 U 0 , we have the cocycle equation
1 , h2 ) + (h
1 h2 , h3 ) = (h
1 , h2 h3 ) + (h
2 , h3 );
(h
we average this against (h3 ) dH (h3 ) and conclude that
1 , h2 ) + F (h1 h2 ) = F (h1 ) + F (h2 ) + (h1 , h2 )
(h
for h1 , h2 U 0 , where F : (U 0 )2 Rn is the function
Z
h3 )(h3 ) dH (h3 )
F (h) :=
(h,
H
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
265
from which it becomes clear that is smooth in the h2 variable. A similar averaging argument (now against a bump function on a right-invariant
measure) then shows that is in turn locally cohomologous to another local
cocycle 0 : U 00 U 00 Rn which is now smooth in both the h1 and h2 variables. Pulling Rn back to K, we conclude that is locally cohomologous to
a smooth local cocycle, and Theorem 2.6.1 now follows from Lemma 2.6.5.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
266
2. Related articles
acts on S 1 but is infinite dimensional and is not locally compact (with, say,
the uniform topology). Finally, the connectedness of X is also important:
the infinite torus G = (R/Z)N (with the product topology) acts faithfully
on the disconnected manifold X := R/Z N by the action
(gn )nN (, m) := ( + gm , m).
Note that (R/Z)N contains the p-adic group Zp as an embedded subgroup
(identifying a p-adic integer m with ( pmn mod 1)nN ), so this also gives a
faithful action of Zp on X.
The conjecture in full generality remains open. However, there are a
number of partial results. For instance, it was observed by Montgomery
and Zippin [MoZi1974] that the conjecture is true for transitive actions;
see Section 1.6.4. Another partial result is the reduction of the HilbertSmith conjecture to the p-adic case. Indeed, it is known that Conjecture
2.7.2 is equivalent to
Conjecture 2.7.3 (Hilbert-Smith conjecture for p-adic actions). It is not
possible for a p-adic group Zp to act continuously and effectively on a connected finite-dimensional manifold X.
The reduction to the p-adic case follows from the structural theory of
locally compact groups (specifically, the Gleason-Yamabe theorem, Theorem
1.1.17) and some results of Newman [Ne1931] that sharply restrict the
ability of periodic actions on a manifold X to be close to the identity. This
argument will also be given below (following the presentation in [Le1997]).
Very recently, the three-dimensional case of Conjecture 2.7.3 (and hence
the three-dimensional case of Conjecture 2.7.2) was settled in [Pa2013], by
topological methods; however, we will not discuss the proof of this conjecture
here.
2.7.1. Periodic actions of prime order. We now study periodic actions
T : X X on a manifold X of some prime order p, thus T p = id.
The basic observation to exploit here is that of rigidity: a periodic action
(or more precisely, the orbits of this action) cannot be too close to the
identity, without actually being the identity. More precisely, we have the
following theorem of Newman [Ne1931]:
Theorem 2.7.4 (Newmans first theorem). Let U be an open subset of Rn
containing the closed unit ball B, and let T : U U be a homeomorphism
of some prime period p 1. Suppose that for every x U , the orbit
{T n x : n = 0, 1, . . . , p 1} has diameter strictly less than 1. Then T (0) = 0.
Note that some result like this must be needed in order to establish
the Hilbert-Smith conjecture. Suppose for instance that one could find a
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
267
for any volume form on U of total mass 1 (one can show that this definition
is independent of the choice of ). This definition is stable under uniform
convergence of , and thus can be used to also define the degree for maps
that are merely continuous rather than continuously differentiable.
Proof. Suppose for contradiction that T (0) 6= 0. We use an averaging
argument, combined with degree theory. Let : U Rn be the map
p1
(x) :=
1X n
T x.
p
n=0
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
268
2. Related articles
not a fixed point of T , and thus (as T has prime order) all elements in the
preimage 1 ({x}) are not fixed points of T either. Thus 1 ({x}) can be
partitioned into a finite number of disjoint orbits of T of cardinality p. As
T is orientation preserving and x is a regular point of , each of these orbits
contributes +p or p to the degree, giving the claim.
The case when is not continuously differentiable is trickier, as the
degree is not as easily computed in this case. One way to proceed is to
perturb T (or more precisely, the graph {(x, T x, . . . , T p1 x) : x U } in U p )
to be piecewise linear near the preimage of 0 (while preserving the periodicity
properties of the graph), so that degree can be computed by hand; this is
the approach taken in Newmans original paper [Ne1931]. Another is to
use the machinery of singular homology, which is more general and flexible
than degree theory; this is the approach taken in [Sm1941], [Dr1969].
Note that the above argument shows not only that T fixes the origin
0, but must also fix an open neighbourhood of the origin, by translating B
slightly. One can then extend this open neighbourhood to the entire space
by the following variant of Theorem 2.7.4.
Theorem 2.7.5 (Newmans second theorem). Let X be a connected manifold, and let T : X X be a homeomorphism of some prime order p that
fixes a non-empty open set U . Then T is the identity.
Proof. We need to show that T fixes all points in X, and not just U .
Suppose for sake of contradiction that T just fixes some of the points in X
and not others. By a continuity argument, and applying a homeomorphic
change of variables if necessary, we can find a coordinate chart containing
the ball B, where T fixes all x = (x1 , . . . , xn ) B with x1 0, but does
not fix all x B with x1 > 0. By shrinking B we may assume that the
entire orbit B, T B, . . . , T p1 B stays inside the coordinate chart (and can
thus be viewed as a subset of Rn ). Then the map can be defined as
before. This map is the identity on the left hemisphere {x B : x1 0}.
On the right hemisphere {x B : x1 > 0}, one observes for x small enough
that x, T x, . . . , T p1 x must all stay on the right-hemisphere (as they must
lie in B and cannot enter the left-hemisphere, where T is the identity) and
so stays on the right. This implies that has degree 1 near 0 on a small
ball around the origin, but as before one can argue that the degree must in
fact be divisible by p, leading again to a contradiction.
2.7.2. Reduction to the p-adic case. We are now ready to prove Conjecture 2.7.2 assuming Conjecture 2.7.3. Let G be a locally compact group
acting continuously and faithfully on a connected manifold X; we wish to
show that G is Lie.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
269
We first make some basic reductions. We will need the following fact:
Exercise 2.7.1. Show that the extension of a Lie group by another Lie
group is again isomorphic to a Lie group. (Hint: use Corollary 1.5.8.) This
result was first established in [Gl1951]. Note that this result supersedes
Theorem 2.6.1, though it uses more of the theory of Hilberts fifth problem
in its proof.
From the Gleason-Yamabe theorem (Theorem 1.1.17), every locally compact group G contains an open subgroup which is an extension of a Lie group
by a compact subgroup of G. Since a group with a Lie group as an open
subgroup is again Lie (because all outer automorphisms of Lie groups are
smooth), it thus suffices by Exercise 2.7.1 to prove the claim when G is
compact. In particular, all orbits of G on X are also compact.
Let B be a small ball in chart of X around some origin x0 . By continuity,
there is some neighbourhood U of the identity in G such that gB 3 x0 for
all g U . By the Peter-Weyl theorem, there is a compact normal subgroup
G0 of G in U with G/G0 linear (and hence Lie). The set G0 B is then a
G0 -invariant manifold, which is precompact and connected (because all the
shifts gB are connected and share a common point). If we let G00 be the
subgroup of G0 that fixes G0 B, then G00 is a compact normal subgroup of G0 ,
and G0 /G00 acts faithfully on G0 B. By Newmans first theorem (Theorem
2.7.4), we see that if U is small enough, then G0 /G00 cannot contain any
elements of prime order, and hence cannot contain any non-trivial periodic
elements whatsoever; by Conjecture 2.7.3, it also cannot contain a continuously embedded copy of Zp for any p. We claim that this forces G0 /G00 to
be trivial.
As G0 B is precompact, the space of C(G0 B G0 B) of continuous maps
from G0 B to itself (with the compact-open topology) is first countable, which
makes G0 /G00 first-countable as well (since G0 /G00 is homeomorphic to a
subspace of C(G0 B G0 B)). The claim now follows from
Lemma 2.7.6. Let G be a compact first-countable group which does not
contain any non-trivial periodic elements or a continuously embedded copy
of Zp for any p. Then G is trivial.
Proof. Every element g of G is contained in a compact abelian subgroup
of G, namely the closed group hgi generated by g. Thus we may assume
without loss of generality that G is abelian.
As G is both compact and first countable, it can be written (using
Exercise 1.4.21) as the inverse limit lim Gn of a countable sequence of
compact abelian Lie groups Gn , with surjective continuous projection homomorphisms n+1n : Gn+1 Gn between these Lie groups. (Note that
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
270
2. Related articles
without first countability, one might only be an inverse limit of a net of Lie
groups rather than a sequence, as one can see for instance with the example
(R/Z)R .)
Suppose for contradiction that at least one of the Gn , say G1 , is nontrivial. It is a standard fact that every compact abelian Lie group is isomorphic to the direct product of a torus and a finite group. (Indeed, in the
connected case one can inspect the kernel of the exponential map, and then
one can extend to the general case by viewing a compact Lie group as an
extension of a finite group by a connected compact Lie group.) in particular,
the periodic points (i.e. points of finite order) are dense, and so there exists
an element g1 of G1 of finite non-trivial order. By raising g1 to a suitable
power, we may assume that g1 has some prime order p.
We now claim inductively that for each n = 1, 2, . . . , g1 can be lifted
to an element gn Gn of some order pkn , where 1 = k1 k2 . . . are
a non-decreasing set of integers. Indeed, suppose inductively that we have
already lifted g1 up to gn Gn with order pkn . The preimage of gn in Gn+1
is then a dense subset of the preimage of hgn i, which is a compact abelian
0
Lie group, and thus contains an element gn+1
of some finite order. As gn
k
0
n
has order p , gn+1 must have an order divisible by pkn , and thus of the
0
form pkn+1 q for some q coprime to p and some kn+1 kn . By raising gn+1
k
to a multiple of q that equals 1 mod p n , we may eliminate q and obtain a
preimage gn+1 of order pkn+1 , and the claim follows.
There are now two cases, depending on whether kn goes to infinity or
not. If the kn stay bounded, then they converge to a limit k, and the inverse
limit of the gn is then an element g of G of finite order pk , a contradiction.
But if the kn are unbounded, then G contains a continuously embedded
copy of the inverse limit of the Z/pkn Z, which is Zp , and again we have a
contradiction.
Since G0 /G00 and G/G0 are both Lie groups, G/G00 is Lie too. Thus it
suffices to show that G00 is Lie. Without loss of generality, we may therefore
replace G by G00 and assume that B is fixed by G.
Now let be the closure of the interior of the set of fixed points of G.
Then is non-empty, and is also clearly closed. We claim that is open;
by connectedness of X, this implies that = X, which by faithfulness of
G implies that G is trivial, giving the claim. Indeed, let x , and let B
be a small ball around x. As before, GB is then a connected G-invariant
-compact manifold, and so if G0 is the subgroup of G that fixes GB, then
G/G0 is a compact Lie group that acts faithfully on GB fixing a nontrivial
open subset of GB. By Newmans theorem (Theorem 2.7.5), G/G0 cannot
contain any periodic elements; by Conjecture 2.7.3, it also cannot contain
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
271
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
272
2. Related articles
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
273
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
274
2. Related articles
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
275
Proof. Let V : G U (V ) be a subrepresentation of the left-regular representation that is isomorphic to . Thus, we have an equivariant isometry
: V L2 (G) whose image is V ; it has an adjoint : L2 (G) V .
Let v V and K L2 (G). The convolution
Z
(v) K(g) :=
(v)(gh)K(h1 ) d(h)
G
can be re-arranged as
Z
d(h)
(g 1 )((v))(h)K(h)
L2 (G)
= h (g 1 )((v)), Ki
L2 (G)
= h( (g 1 )v), Ki
V
= h (g 1 )v, Ki
= hv, (g) KiV
where
:= K(g 1 ).
K(g)
In particular, we see that (v) K L2 (G) for every K. Letting K be
a sequence (or net) of approximations to the identity, we conclude that
(v) L2 (G) as well, and so V L2 (G) , which is the first claim.
To prove the converse claim, write n := dim(V ), and let e1 , . . . , en be
an orthonormal basis for V . Observe that we may then decompose L2 (G)
as the direct sum of the spaces
L2 (G),ei := {f,v,ei : v V }
for i = 1, . . . , n. The claim follows.
are
From Corollary 2.8.3, the -isotypic components L2 (G) for G
2
pairwise orthogonal, and so we can form the direct sum G L (G)
dim(G)
G
, which is an invariant subspace of L2 (G) that contains all
the finite-dimensional irreducible subrepresentations (and hence also all the
finite-dimensional representations, period). The essence of the Peter-Weyl
theorem is then the assertion that this direct sum in fact occupies all of
L2 (G):
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
276
2. Related articles
Z
=
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
277
,
tr(db )IV , ca
dim(V )
HS(V )
which simplifies to
f() := f =
dim(V )kf()k2HS(V ) .
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
278
2. Related articles
surjective, and thus (g) HS(V ) must span all of HS(V ). Thus, any
bi-invariant subspace of HS(V ) must also be invariant with respect to left
and right multiplication by arbitrary elements of HS(V ), and in particular
by rank one operators; from this one easily sees that there are no non-trivial
bi-invariant subspaces. Thus we can view the Peter-Weyl theorem as also
describing the irreducible decomposition of L2 (G) into G G-irreducible
components.
as being the
Remark 2.8.9. In view of (2.39), it is natural to view G
occuring with multiplicity
spectrum of G, with each frequency G
dim(V ).
In the abelian case, any eigenspace of one unitary operator (g) is automatically an invariant subspace of all other (h), which quickly implies (from
the spectral theorem) that all irreducible finite-dimensional unitary representations must be one-dimensional, in which case we see that the above
formulae collapse to the usual Fourier inversion and Plancherel theorems for
compact abelian groups.
In the case of a finite group G, we can take dimensions in (2.39) to obtain
the identity
X
|G| =
dim(V )2 .
f\
1 f2 () = f1 ()f2 ()
for f1 , f2 L2 (G), thus partially diagonalising convolution into multiplication of linear operators on finite-dimensional vector spaces V . (Of course,
one cannot expect complete diagonalisation in the non-abelian case, since
convolution would then also be non-abelian, whereas diagonalised operators
must always commute with each other.)
Call a function f L2 (G) a class function if it is conjugation-invariant,
thus f (gxg 1 ) = f (x) for all x, g G. It is easy to see that this is equivalent to each of the Fourier coefficients f() also being conjugation-invariant:
(g)f() (g) = f(). By Lemma 2.8.5, this is in turn equivalent to f()
being equal to a multiple of the identity:
f() =
1
1
tr(f())IV =
hf, iL2 (G) IV
dim(V )
dim(V )
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
279
thus the form an orthonormal basis for the space L2 (G)G of class functions. Analogously to (2.39), we have
M
C.
L2 (G)G
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
280
2. Related articles
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
281
for all z
This shows that there is no solution to P1 (z) = = Pr (z) =
0 over C, as required.
Remark 2.9.2. Observe that in the above argument, one could replace Q
and C by any other pair of fields, with the latter containing the algebraic
closure of the former, and still obtain the same result.
The above lemma asserts that if a system of rational equations is solvable
at all, then it is solvable with some algebraic solution. But it gives no bound
on the complexity of that solution in terms of the complexity of the original
equation. Changs lemma provides such a bound. If H 1 is an integer,
let us say that an algebraic number has height at most H if its minimal
polynomial (after clearing denominators) consists of integers of magnitude
at most H.
Lemma 2.9.3 (Quantitative solvability). Let P1 , . . . , Pr : Cd C be a
finite number of polynomials of degree at most D with rational coefficients,
each of height at most H. If there is a complex solution z = (z1 , . . . , zd ) Cd
to the simultaneous system of equations
P1 (z) = = Pr (z) = 0,
d
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
282
2. Related articles
simply using it as a black box, thus providing a soft proof of Lemma 2.9.3
that is an alternative to the hard proofs mentioned above.
Heres how the proof works. Informally, the idea is that Lemma 2.9.3
should follow from Lemma 2.9.1 after replacing the field of rationals Q with
the field of rationals of polynomially bounded height. Unfortunately, the
latter object does not really make sense as a field in standard analysis;
nevertheless, it is a perfectly sensible object in nonstandard analysis, and
this allows the above informal argument to be made rigorous.
We turn to the details. As is common whenever one uses nonstandard
analysis to prove finitary results, we use a compactness and contradiction
argument (or more precisely, an ultralimit and contradiction argument).
Suppose for contradiction that Lemma 2.9.3 failed. Carefully negating the
quantifiers (and using the axiom of choice), we conclude that there exists
D, d, r such that for each natural number n, there is a positive integer H (n)
(n)
(n)
and a family P1 , . . . , Pr : Cd C of polynomials of degree at most D
and rational coefficients of height at most H (n) , such that there exist at least
one complex solution z (n) Cd to
(2.40)
(n)
P1 (z (n) ) = = Pr (z (n) ) = 0,
but such that there does not exist any such solution whose coefficients are
algebraic numbers of degree at most n and height at most n(H (n) )n .
Now we take ultralimits, as in Section 1.7. Let N\N be a nonprincipal ultrafilter. For each i = 1, . . . , r, the ultralimit
(n)
Pi := lim Pi
n
(n)
Pi
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
283
But for n larger than C, this contradicts the construction of the Pi , and
the claim follows. (Note that as p is non-principal, any neighbourhood of p
in N will contain arbitrarily large natural numbers.)
Remark 2.9.4. The same argument actually gives a slightly stronger version of Lemma 2.9.3, namely that the integer coefficients used to define the
algebraic solution z can be taken to be polynomials in the coefficients of
P1 , . . . , Pr , with degree and coefficients bounded by CD,d,r .
Remark 2.9.5. A related application of nonstandard analysis to quantitative algebraic geometry was given in [Sc1989]. (Thanks to Matthias
Aschenbrenner for this reference.)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
284
2. Related articles
A measure space is complete if every subset of a null set (i.e. a measurable set of measure zero) is also a null set. Not all measure spaces are
complete, but one can always form the completion (X, B, ) of a measure
space (X, B, ) by enlarging the -algebra B to the space of all sets which are
equal to a measurable set outside of a null set, and extending the measure
appropriately.
Given two (-finite) measure spaces (X, BX , X ) and (Y, BY , Y ), one
can form the product space (X Y, BX BY , X Y ). This is a measure
space whose domain is the Cartesian product X Y , the -algebra BX BY
is generated by the rectangles A B with A BX , B BY , and the
measure X Y is the unique measure on BX BY obeying the identity
X Y (A B) = X (A)Y (B).
See for instance [Ta2011, 1.7] for a formal construction of product measure3. One of the fundamental theorems concerning product measure is
Tonellis theorem (which is basically the unsigned version of the more wellknown Fubini theorem), which asserts that if f : X Y [0, +] is BX BY
measurable, then the integral expressions
Z Z
( f (x, y) dY (y)) dX (x)
X Y
Z Z
( f (x, y) dX (x)) dY (y)
Y
and
Z
f (x, y) dXY (x, y)
XY
all exist (thus all integrands are almost-everywhere well-defined and measurable with respect to the appropriate -algebras), and are all equal to each
other; see e.g. [Ta2011, Theorem 1.7.10].
Any finite non-empty set V can be turned into a probability space
(V, 2V , V ) by endowing it with the discrete -algebra 2V := {A : A V }
of all subsets of V , and the normalised counting measure
|A|
(A) :=
,
|V |
where |A| denotes the cardinality of A. In this discrete setting, the probability space is automatically complete, and every function f : V [0, +]
is measurable, with the integral simply being the average:
Z
1 X
f dV =
f (v).
|V |
V
vV
3There are technical difficulties with the theory when X or Y is not -finite, but in this
section we will only be dealing with probability spaces, which are clearly -finite, so this difficulty
will not concern us.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
285
Of course, Tonellis theorem is obvious for these discrete spaces; the deeper
content of that theorem is only apparent at the level of continuous measure
spaces.
Among other things, this probability space structure on finite sets can
be used to describe various statistics of dense graphs. Recall that a graph
G = (V, E) is a finite vertex set V , together with a set of edges E, which
we will think of as a symmetric subset of the Cartesian product V V .
(If one wishes, one can prohibit loops in E, so that E is disjoint from the
diagonal V := {(v, v) : v V } of V V , but this will not make much
difference for the discussion below.) Then, if V is non-empty, and ignoring
some minor errors coming from the diagonal V , the edge density of the
graph is essentially
Z
e(G) := V V (E) =
1E (v, w) dV V (v, w),
V V
and so forth.
In [Ru1978], Ruzsa and Szemeredi established the triangle removal
lemma concerning triangle densities, which informally asserts that a graph
with few triangles can be made completely triangle-free by removing a small
number of edges:
Lemma 2.10.1 (Triangle removal lemma). Let G = (V, E) be a graph on a
non-empty finite set V , such that t(G) for some > 0. Then there exists
a subgraph G0 = (V, E 0 ) of G with t(G0 ) = 0, such that e(G\G0 ) = o0 (1),
where o0 (1) denotes a quantity bounded by c() for some function c() of
that goes to zero as 0.
The original proof of the triangle removal lemma was a finitary one,
and proceeded via the Szemeredi regularity lemma [Sz1978]. It has a number of consequences; for instance, as already noted in that paper, the triangle
removal lemma implies as a corollary the famous theorem of Roth [Ro1953]
that subsets of Z of positive upper density contain infinitely many arithmetic
progressions of length three.
It is however also possible to establish this lemma by infinitary means.
There are at least three basic approaches for this. One is via a correspondence principle between questions about dense finite graphs, and questions
about exchangeable random infinite graphs, as was pursued in [Ta2007]. A
second (closely related to the first) is to use the machinery of graph limits, as developed in [LoSz2006], [BoChLoSoVe2008]. The third is via
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
286
2. Related articles
nonstandard analysis (or equivalently, by using ultraproducts), as was pursued in [ElSz2006]. These three approaches differ in the technical details
of their execution, but the net effect of all of these approaches is broadly
the same, in that they both convert statements about large dense graphs
(such as the triangle removal lemma) to measure theoretic statements on
infinitary measure spaces. (This is analogous to how the Furstenberg correspondence principle converts combinatorial statements about dense sets of
integers into ergodic-theoretic statements on measure-preserving systems, as
discussed for instance in [Ta2009].)
In this section we will illustrate the nonstandard analysis approach from
[ElSz2006] by providing a nonstandard proof of the triangle removal lemma.
The main technical tool used here (besides the basic machinery of nonstandard analysis) is that Loeb measure,[Lo1975] which gives a probability
Q space
structure (V, BV , V ) to nonstandard finite non-empty sets V = n Vn
that is an infinitary analogue of the discrete probability space structures
V = (V, 2V , V ) one has on standard finite non-empty sets. The nonstandard analogue of quantities such as triangle densities then become the integrals of various nonstandard functions with respect to Loeb measure. With
this approach, the epsilons and deltas that are so prevalent in the finitary
approach to these subjects disappear almost completely; but to compensate
for this, one now must pay much more attention to questions of measurability, which were automatic in the finitary setting but now require some care
in the infinitary one.
The nonstandard analysis approaches are also related to the regularity
lemma approach; see [Ta2011c, 4.4] for a proof of the regularity lemma
using Loeb measure.
As usual, the nonstandard approach offers a complexity tradeoff: there
is more effort expended in building the foundational mathematical structures of the argument (in this case, ultraproducts and Loeb measure), but
once these foundations are completed, the actual arguments are shorter than
their finitary counterparts. In the case of the triangle removal lemma, this
tradeoff does not lead to a particularly significant reduction in complexity
(and arguably leads in fact to an increase in the length of the arguments,
when written out in full), but the gain becomes more apparent when proving more complicated results, such as the hypergraph removal lemma, in
which the initial investment in foundations leads to a greater savings in net
complexity, as can be seen in [ElSz2006].
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
287
Consider
a nonstandard finite non-empty set V , i.e. an ultraproduct
Q
V = n Vn of standard finite non-empty setsQVn . Define an internal
subset of V to be a subset of V of the form A = n An , where each An
is a subset of Vn . It is easy to see that the collection AV of all internal
subsets of V is a boolean algebra. In general, though, AV will not be a algebra. For instance, suppose that the Vn are the standard discrete intervals
Vn := [1, n] := {i N : i n}, then V is the non-standard discrete interval
V = [1, N ] := {i N : i N }, where N is the unbounded nonstandard
natural number N := limn n. For any standard integer m, the subinterval
[1, N/m] is an internal subset of V ; but the intersection
\
[1, N/m] = {i N : i = o(N )}
[1, o(N )] :=
mN
is not an internal subset of V . (This can be seen, for instance, by noting that
all non-empty internal subsets of [1, N ] have a maximal element, whereas
[1, o(N )] does not.)
Q
Given any internal subset A = n An of V , we can define the cardinality |A| of A, which is the nonstandard natural number |A| := limn |An |.
|A|
We then have the nonstandard density |V
| , which is a nonstandard real
number between 0 and 1. By Exercise 1.7.18, this bounded nonstandard
|A|
|A|
real number |V
| has a unique standard part st( |V | ), which is a standard real
number in [0, 1] such that
|A|
|A|
= st(
) + o(1),
|V |
|V |
where o(1) denotes a nonstandard infinitesimal (i.e. a nonstandard number
which is smaller in magnitude than any standard > 0).
In [Lo1975], Loeb observed that this standard density can be extended
to a complete probability measure:
Theorem 2.10.2 (Construction of Loeb measure). Let V be a nonstandard finite non-empty set. Then there exists a complete probability space
(V, LV , V ), with the following properties:
(Internal sets are Loeb measurable) If A is an internal subset of V ,
then A LV and
V (A) = st(
|A|
).
|V |
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
288
2. Related articles
[
Bn
EA
n=1
and
V (Bn ) .
n=1
|A|
Proof. The map V : A 7 st( |V
| ) is a finitely additive probability measure
on AV . We claim that this map V is in fact a pre-measure on AV , thus
one has
X
(2.41)
V (A) =
V (An )
n=1
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
289
and
(V W, LV W , V W ),
(2.42)
(V W, LV LW , V W ).
It is then natural to ask how the two probability spaces (2.42) and (2.43)
are related. There is one easy relationship, which shows that (2.42) extends
(2.43):
Exercise 2.10.1. Show that (2.42) is a refinement of (2.43), thus LV LW ,
and V W extends V W . (Hint: first recall why the product of Lebesgue
measurable sets is Lebesgue measurable, and mimic that proof to show that
the product of a LV -measurable set and a LW -measurable set is LV W measurable, and that the two measures V W and V W agree in this
case.)
In the converse direction, (2.42) enjoys the type of Tonelli theorem that
(2.43) does:
Theorem 2.10.3 (Tonelli theorem for Loeb measure). Let V, W be two
nonstandard finite non-empty sets, and let f : V W [0, +] be an
unsigned LV W -measurable function. Then the expressions
Z Z
(2.44)
(
f (v, w) dW (w)) dV (v)
V
(2.45)
(
W
and
Z
(2.46)
f (v, w) dV W (v, w)
V W
are well-defined (thus all integrands are almost everywhere well-defined and
appropriately measurable) and equal to each other.
This result is sometines referred to as the Keisler-Fubini theorem.
Proof. By the monotone convergence theorem it suffices to verify this when
f is a simple function; by linearity we may then take f to be an indicator
function f = 1E . Using Theorem 2.10.2 and an approximation argument
(and many further applications of monotone convergence) we may assume
without loss of generality that E is an internal set. We then have
Z
|E|
f (v, w) dV W (v, w) = st(
)
|V ||W |
V W
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
290
2. Related articles
.
n
|W |
n
On each Vi , we have
Z
(2.47)
W
i
f (v, w) dW (w) = + O
n
1
n
and
(2.48)
|Ev |
i
= +O
|W |
n
1
.
n
R
W
f (v, w) dW (w)
i=1
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
291
Remark 2.10.4. It is well known that the product of two Lebesgue measure spaces Rn , Rm , upon completion, becomes the Lebesgue measure space
on Rn+m . Drawing the analogy between Loeb measure and Lebesgue measure, it is then natural to ask whether (2.42) is simply the completion of
(2.43). But while (2.42) certainly contains the completion of (2.43), it is a
significantly larger space
Q in general (a fact
Q first observed by Doug Hoover).
Indeed, suppose V = n Vn , W = n Wn , where the cardinality of
Vn , Wn goes to infinity at some reasonable rate, e.g. |Vn |, |Wn | n for all
n. For each n, let En be a random subset of Vn Wn , with each element of
Vn Wn having an independent probability of 1/2 of lying in En . Then, as
is well known, the sequence of sets En is almost surely asymptotically regular
in the sense that almost surely, we have the bound
sup
An Vn ,Bn Wn
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
292
2. Related articles
that
Z
(2.49)
V V V
Then for any standard > 0, there exists a internal subsets Fij V V for
ij = 12, 23, 31 with V V (Eij \Fij ) < , which are completely triangle-free
in the sense that
(2.50)
for all u, v, w V .
Let us first see why Lemma 2.10.5 implies Lemma 2.10.1. We use the
usual compactness and contradiction argument. Suppose for contradiction
that Lemma 1.2 failed. Carefully negating the quantifiers, we can find a
(standard) > 0, and a sequence Gn = (Vn , En ) of graphs with t(Gn ) 1/n,
such that for each n, there does not exist a subgraph G0n = (Vn , En0 ) of n
with |En \En0 | |Vn |2 with t(G0n ) = 0. Clearly we may assume the Vn are
non-empty.
Q
We form
the
ultraproduct
G
=
(V,
E)
of
the
G
,
thus
V
=
n
n Vn
Q
and E = n En . By construction, E is a symmetric internal subset of
V V and we have
Z
1E (u, v)1E (v, w)1E (w, u) dV V V (u, v, w) = st lim t(Gn ) = 0.
n
V V V
Thus, by Lemma 2.10.5, we may find internal subsets F12 , F23 , F31 of V V
with V V (E\Fij ) < /6 (say) for ij = 12, 23, 31 such that (2.50) holds for
all u, v, w V . By letting E 0 be the intersection of all E with all the Fij
and their reflections, we see that E 0 is a symmetric internal subset of E with
V V (E\E 0 ) < , and we still have
1E 0 (u, v)1E 0 (v, w)1E 0 (w, u) = 0
for all u, v, w V . If we write E 0 = limn En0 for some sets En0 , then for n
sufficiently close to p, one has En0 a symmetric subset of En with
Vn Vn (En \En0 ) <
and
1En0 (u, v)1En0 (v, w)1En0 (w, u) = 0.
G0n
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
293
lemma will see strong similarities between that argument and the one given
here (and, on some level, they are essentially the same argument).
To begin with, we suppose first that the Eij are all elementary sets, in
the sense that they are finite boolean combinations of products of internal
sets. (At the finitary level, this corresponds to graphs that are bounded
combinations of bipartite graphs.) This implies that there is an internal
partition V = V1 Vn of the vertex set V , such that each Eij is the
union of some of the Va Vb .
Let Fij be the union of all the Va Vb in Eij for which Va and Vb have
positive Loeb measure; then V V (Eij \Fij ) = 0. We claim that (2.50)
holds for all u, v, w V , which gives Theorem 2.10.5 in this case. Indeed, if
u Va , v Vb , w Vc were such that (2.50) failed, then E12 would contain
Va Vb , E23 would contain Vb Vc , and E31 would contain Vc Va . The
integrand in (2.49) is then equal to 1 on Va Vb Vc , which has Loeb
measure V (Va )V (Vb )V (Vc ) which is non-zero, contradicting (2.49). This
gives Theorem 2.10.5 in the elementary set case.
Next, we increase the level of generality by assuming that the Eij are
all LV LV -measurable. (The finitary equivalent of this is a little difficult
to pin down; roughly speaking, it is dealing with graphs that are not quite
bounded combinations of bounded graphs, but can be well approximated
by such bounded combinations; a good example is the half-graph, which is
a bipartite graph between two copies of {1, . . . , N }, which joins an edge
between the first copy of i and the second copy of j iff i < j.) Then each Eij
can be approximated to within an error of /3 in V V by elementary sets.
In particular, we can find a finite partition V = V1 Vn of V , and sets
0 that are unions of some of the V V , such that
0
Eij
a
V V (Eij Eij ) < /3.
b
0 such that V , V
Let Fij be the union of all the Va Vb contained in Eij
a
b
have positive Loeb measure, and such that
2
V V (Eij (Va Vb )) > V V (Va Vb ).
3
Then the Fij are internal subsets of V V , and V V (Eij \Fij ) < .
We now claim that the Fij obey (2.50) for all u, v, w, which gives Theorem 2.10.5 in this case. Indeed, if u Va , v Vb , w Vc were such that
(2.50) failed, then E12 occupies more than 23 of Va Vb , and thus
Z
2
1E12 (u, v) dV V V (u, v, w) > V V V (Va Vb Vc ).
3
Va Vb Vc
Similarly for 1E23 (v, w) and 1E31 (w, u). From the inclusion-exclusion formula, we conclude that
Z
1E12 (u, v)1E23 (v, w)1E31 (w, u) dV V V (u, v, w) > 0,
Va Vb Vc
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
294
2. Related articles
Also, we have
0
V V (Eij \Eij
)
Z
=
V V
1Eij (1 1Eij0 )
Z
=
V V
fij (1 1Eij0 )
/2.
Applying the already established cases of Theorem 2.10.5, we can find
0 \F ) < /2, and hence
internal sets Fij obeying (2.50) with V V (Eij
ij
V V (Eij \Fij ) < , and Theorem 2.10.5 follows.
Remark 2.10.6. The full hypergraph removal lemma can be proven using
similar techniques, but with a longer tower of generalisations than the three
cases given here; see [Ta2007] or [ElSz2006].
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
295
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
296
2. Related articles
The other basic fact we will present here concerns the algebraic nature
of Lie groups and Lie algebras. An important family of examples of Lie
groups are the algebraic groups - algebraic varieties with a group law given
by algebraic maps. Given that one can always automatically upgrade the
smooth structure on a Lie group to analytic structure (by using the BakerCampbell-Hausdorff formula), it is natural to ask whether one can upgrade
the structure further to an algebraic structure. Unfortunately, this is not always the case. A prototypical example of this is given by the one-parameter
subgroup
(2.51)
G :=
t 0
0 t
:tR
of GL2 (R). This is a Lie group for any exponent R, but if is irrational,
then the curve that G traces out is not an algebraic subset of GL2 (R) (as
one can see by playing around with Puiseux series).
This is not a true counterexample to the claim that every Lie group
can be given the structure of an algebraic group, because one can give G
a different algebraic structure than one inherited from the ambient group
GL2 (R). Indeed, G is clearly isomorphic to the additive group R, which is of
course an algebraic group. However, a modification of the above construction
works:
Proposition 2.11.2. There exists a Lie group G that cannot be given the
structure of an algebraic group.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
297
1 0 0
G := x t 0 : x, y R; t R+
y 0 t
of GL3 (R), with an irrational number. This is a three-dimensional (metabelian)
Lie group, whose Lie algebra g gl3 (R) is spanned by the elements
0 0 0
X := 0 1 0
0 0
0 0 0
Y := 1 0 0
0 0 0
0 0 0
Z := 0 0 0
0 0
with the Lie bracket given by
[Y, X] = Y ; [Z, X] = Z; [Y, Z] = 0.
As such, we see that if we use the basis X, Y, Z to identify g to R3 , then the
adjoint representation of G is the identity map.
If G is an algebraic group, it is easy to see that the adjoint representation
Ad : G GL(g) is also algebraic, and so Ad(G) = G is algebraic in GL(g).
Specialising to our specific example, in which adjoint representation is the
identity, we conclude that if G has any algebraic structure, then it must also
be an algebraic subgroup of GL3 (R); but G projects to the group (2.51)
which is not algebraic, a contradiction.
A slight modification of the same argument also shows that not every Lie
algebra is algebraic, in the sense that it is isomorphic to a Lie algebra of an
algebraic group. (However, there are important classes of Lie algebras that
are automatically algebraic, such as nilpotent or semisimple Lie algebras.)
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
Bibliography
299
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
300
Bibliography
Etudes
Sci. Publ. Math. No. 53 (1981), 5373.
[GuKa2010] L. Guth, N. Katz, On the Erdos distinct distance problem in the plane,
preprint.
[He1983] J. Heintz, Definability and fast quantifier elimination in algebraically closed
fields, Theoret. Comput. Sci. 24 (1983), no. 3, 239277.
[HeRo1979] E. Hewitt, K. Ross, Abstract harmonic analysis. Vol. I. Structure of topological groups, integration theory, group representations. Second edition. Grundlehren der
Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences],
115. Springer-Verlag, Berlin-New York, 1979.
[Hi1990] J. Hirschfeld, The nonstandard treatment of Hilberts fifth problem, Trans. Amer.
Math. Soc. 321 (1990), no. 1, 379400.
[Ho1991] B. Host, Mixing of all orders and pairwise independent joinings of systems with
singular spectrum, Israel J. Math. 76 (1991), no. 3, 289298.
[Hr2012] E. Hrushovski, Stable group theory and approximate subgroups, J. Amer. Math.
Soc. 25 (2012), no. 1, 189243.
[IoRoRu2011] A. Iosevich, O. Roche-Newton, M. Rudnev, On an application of Guth-Katz
theorem, Math. Res. Lett. 18 (2011), no. 4, 691697.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
Bibliography
301
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
302
Bibliography
[Pe2006] P. Petersen, Riemannian geometry. Second edition. Graduate Texts in Mathematics, 171. Springer, New York, 2006.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
Bibliography
303
[Ta2009b] T. Tao, Poincares Legacies: pages from year two of a mathematical blog, Vol.
II, American Mathematical Society, Providence RI, 2009.
[Ta2010] T. Tao, An epsilon of room, Vol. I, American Mathematical Society, Providence
RI, 2010.
[Ta2010b] T. Tao, An epsilon of room, Vol. II, American Mathematical Society, Providence RI, 2010.
[Ta2011] T. Tao, An introduction to measure theory, American Mathematical Society,
Providence RI, 2011.
[Ta2011b] T. Tao, Higher order Fourier analysis, American Mathematical Society, Providence RI, 2011.
[Ta2011c] T. Tao, Compactness and contradiction, in preparation.
[TaVu2006] T. Tao, V. Vu, Additive combinatorics. Cambridge Studies in Advanced
Mathematics, 105. Cambridge University Press, Cambridge, 2006.
[TaYu2005] P. Tauvel, R. Yu, Lie algebras and algebraic groups. Springer Monographs in
Mathematics. Springer-Verlag, Berlin, 2005.
[vDa1936] D. van Dantzig, Zur topologischen Algebra. III. Brouwersche und Cantorsche
Gruppen, Compositio Math. 3 (1936), 408426.
[vdDrGo2010] L. van den Dries, I. Goldbring, Globalizing locally compact local groups, J.
Lie Theory 20 (2010), no. 3, 519524.
[vKa1833] E. R. van Kampen, On the connection between the fundamental groups of some
related spaces, Amer. J. Math., 55 (1933), 261267.
[vdW1927] B.L. van der Waerden, Beweis einer Baudetschen Vermutung, Nieuw. Arch.
Wisk. 15 (1927), 212216.
[vdDrWi1984] L. van den Dries, A. J. Wilkie, Gromovs theorem on groups of polynomial
growth and elementary logic, J. Algebra, 89 (1984), 349374.
[vEsKo1964] W. T. van Est, Th. J.Korthagen, Non-enlargible Lie algebras, Nederl. Akad.
Wetensch. Proc. Ser. A 67, Indag. Math. 26 (1964), 1531.
[We1973] B. A. F. Wehrfritz, Infinite linear groups. An account of the group-theoretic
properties of infinite groups of matrices. Ergebnisse der Matematik und ihrer Grenzgebiete, Band 76. Springer-Verlag, New York-Heidelberg, 1973.
[Ya1953] H. Yamabe, On the conjecture of Iwasawa and Gleason, Ann. of Math. 58,
(1953), 4854.
[Ya1953b] H. Yamabe, A generalization of a theorem of Gleason, Ann. of Math. 58 (1953),
351365.
[Ze1990] E. I. Zelmanov, Solution of the restricted Burnside problem for groups of odd
exponent, Izv. Akad. Nauk SSSR Ser. Mat. 54 (1990), no. 1, 4259.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
Index
305
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
306
Index
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
Index
307
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.
308
Index
well-defined word, 26
well-formed formula, 143
word metric, 19
Author's preliminary version made available with permission of the publisher, the American Mathematical Society.