Groups Graphs and Trees An Introduction To The Geometry of Infinite Groups
Groups Graphs and Trees An Introduction To The Geometry of Infinite Groups
L O N D O N M AT H E M AT I C A L S O C I E T Y S T U D E N T T E X T S
JOHN MEIER
Lafayette College
CAMBRIDGE UNIVERSITY PRESS
Cambridge, New York, Melbourne, Madrid, Cape Town, Singapore, São Paulo
© J. Meier 2008
Cambridge University Press has no responsibility for the persistence or accuracy of urls
for external or third-party internet websites referred to in this publication, and does not
guarantee that any content on such websites is, or will remain, accurate or appropriate.
For my driver, Piotr
Contents
Preface page ix
1 Cayley’s Theorems 1
1.1 Cayley’s Basic Theorem 1
1.2 Graphs 6
1.3 Symmetry Groups of Graphs 10
1.4 Orbits and Stabilizers 15
1.5 Generating Sets and Cayley Graphs 17
1.6 More Cayley Graphs 22
1.7 Symmetries of Cayley Graphs 29
1.8 Fundamental Domains and Generating Sets 30
1.9 Words and Paths 37
2 Groups Generated by Reflections 44
3 Groups Acting on Trees 54
3.1 Free Groups 54
3.2 F3 is a Subgroup of F2 65
3.3 Free Group Homomorphisms and
Group Presentations 67
3.4 Free Groups and Actions on Trees 70
3.5 The Group Z3 ∗ Z4 73
3.6 Free Products of Groups 79
3.7 Free Products of Finite Groups are Virtually Free 83
3.8 A Geometric View of Theorem 3.35 87
3.9 Finite Groups Acting on Trees 89
3.10 Serre’s Property FA and Infinite Groups 90
4 Baumslag–Solitar Groups 100
vii
viii Contents
ix
x Preface
an instructor can cover much of the rest of this book, if for example the
material in Chapter 7 or Chapter 11 is presented more as a colloquium
than as course material. My own hope is that various classes will find
the space in their semester to pursue tangents of interest to them, and
then let me know the results of their exploration.
I have many people to thank. My wife Trisha and son Robert were
unreasonably supportive of this project. Many students provided import-
ant feedback as I fumbled through the process of presenting this bit of
advanced mathematics at an elementary level: George Armagh, Kari
Barkley, Jenna Bratz, Jacob Carson, Joellen Cope, Joe Dudek, Josh
Goldstein, Ekaterina Jager, Brian Kronenthal, Rob McEwen, and
Zachary Reiter. I also benefited from extensive feedback given by my
colleagues Ethan Berkove and Jon McCammond. Finally, a number of
anonymous referees provided comments on various draft chapters. I was
impressed by the fact that there was no intersection between the com-
ments provided by students, the comments provided by colleagues, and
the anonymous referees!
1
Cayley’s Theorems
1
2 Cayley’s Theorems
Because the general theme of this book is to study groups via actions,
we need a bit of notation and a formal definition.
ful in using our convention, comes from the integers. If the integers are
thought of as simply a set, containing infinitely many elements, then
Sym(Z) is an infinite permutation group, which contains Symn for any
n. On the other hand, if Z denotes the group of integers under addition,
then Sym(Z) ≈ Z2 . (The only non-trivial automorphism of the group of
integers sends n to −n for all n ∈ Z.)
Definition 1.2. An action of a group G on a mathematical object X
is a group homomorphism from G to Sym(X). Equivalently, it is a map
from G × X → X such that
Proof. The objects that G permutes are the elements of G. In this proof
we use “SymG ” to denote Sym(G), to emphasize that “G” denotes the
underlying set of elements, not the group. The permutation associated
to g ∈ G is defined by left multiplication by g. That is, g → πg ∈ SymG
where πg (h) = g · h for all h ∈ G. This is a permutation of the elements
of G, since if g · h = g · h , then by left cancellation, h = h . Denote the
map taking the element g to the permutation πg by π : G → SymG .
To check that π is a group homomorphism we need to verify that
π(gh) = π(g) · π(h). In other words, we need to show that πgh = πg · πh .
We do this by evaluating what each side does to an arbitrary element of
G. We denote the arbitrary element by “x”, thinking of it as a variable.
The permutation πgh takes x → (gh) · x, and successively applying πh
then πg sends x → h · x → g · (h · x). Thus checking that φ is a homo-
morphism amounts to verifying the associative law: (gh) · x = g · (h · x).
As this is part of the definition of a group, the equation holds.
In order to see that the map is faithful it suffices to show that no
non-identity element is mapped to the trivial permutation. One can do
this by simply noting that if g ∈ G \ {e}, then g · e = g, hence πg (e) = g,
and so πg is not the identity (or trivial) permutation.
Example 1.7. The group Sym3 has six elements, shown as disjoint
vertices in Figure 1.1. The permutations described by Cayley’s Basic
Theorem – for the elements (12) and (123) – are also shown.
(12) (12)
I I
Fig. 1.1. The permutation of Sym3 induced by (12) is shown on the left, and
the permutation induced by (123) is shown on the right.
1.2 Graphs
One of the key insights into the study of groups is that they can be
viewed as symmetry groups of graphs. We refer to this as “Cayley’s
Better Theorem,” which we prove in Section 1.5.2. In this section we
establish some terminology from graph theory, and in the following
section we discuss groups acting on graphs.
Graphs are often visualized by making the vertices points on paper and
edges arcs connecting the appropriate vertices. Two simple graphs are
shown in Figure 1.3; a graph which is not simple is shown in Figure 1.4.
Fig. 1.3. The complete graph on five vertices, K5 , and the complete bipartite
graph K3,4 .
vertices – such that, for every e ∈ E(Γ), Ends(e) contains one black
vertex and one white vertex. The complete bipartite graphs are simple
graphs whose vertex sets have been partitioned into two collections, V◦
and V• , with edges joining each vertex in V◦ with each vertex in V• . If
|V◦ | = n and |V• | = m then the corresponding complete bipartite graph
is denoted Kn,m .
The valence or degree of a vertex is the number of edges that contain
it. For example, the valence of any vertex in Kn is n − 1. If a vertex
v is the vertex for a loop, that is an edge e where Ends(e) = {v, v},
then this loop contributes twice to the computation of the valence of v.
For example, the valence of the leftmost vertex in the graph shown in
Figure 1.4 is six.
A graph is locally finite if each vertex is contained in a finite number
of edges, that is, if the valence of every vertex is finite.
An edge path, or more simply a path, in a graph consists of an alter-
nating sequence of vertices and edges, {v0 , e1 , v1 , . . . , vn−1 , en , vn } where
Ends(ei ) = {vi−1 , vi } (for each i). A graph is connected if any two ver-
tices can be joined by an edge path. In Figure 1.4 we have indicated an
b
c i
a
1 d 2 3
h j
e 4 f 5
g
Fig. 1.4. On top is a graph which is not simple, with its vertices labelled by
numbers and its edges labelled by letters. Below is the set of edges traversed in
an edge path, joining the vertex labelled 1 to the vertex labelled 3, is indicated.
edge path from the leftmost vertex to the rightmost vertex. If vi is the
vertex labelled i and eα is the edge labelled α, then this path is:
{v1 , ea , v1 , ee , v4 , eg , v5 , eh , v2 , eb , v1 , ed , v2 , ei , v3 }
8 Cayley’s Theorems
valence n. Thus, for example, the complete bipartite graphs are all bireg-
ular. Given the valences, there is a unique biregular tree, which we denote
by Tm,n . You can see an example in Figure 1.5.
Example 1.16. The symmetry group of the graph Γ shown in Figure 1.6
is isomorphic to Sym3 ⊕ Z2 (or Sym3 ⊕ Sym2 if you prefer the symmetry
in notation). The symmetries come from permuting the multiple edges
joining the leftmost vertices of Γ (which gives the Sym3 factor) and from
a reflection through a horizontal axis (which gives the Sym2 factor). The
fact that these two types of symmetries commute with each other leads
to the direct product structure.
Definition 1.17. Let G be a subgroup of Sym(Γ). Then G is vertex
transitive if, given any two vertices, v and v , there is an α ∈ G where
α(v) = v . The symmetry groups of the complete graphs are all vertex
transitive, while the symmetry group of a complete bipartite graph Kn,m
is not vertex transitive when n = m. (Why?)
A group G < Sym(Γ) is edge transitive if, given any two edges, e and
e , there is a symmetry α ∈ G where α(e) = e .
A flag consists of a pair (v, e), where v ∈ Ends(e). The group G is flag
transitive if, given any two flags, (v, e) and (v , e ), there is a symmetry
α ∈ G where α(v) = v and α(e) = e .
The symmetry group G acts simply transitively on the vertices (edges,
flags, etc.) if it is vertex transitive, and given any two vertices v and v
2 Graph theorists usually refer to automorphisms of a graph, and call the collection
of all automorphisms the automorphism group of the graph.
12 Cayley’s Theorems
one can exchange the front–back pair in order to have the image of u
be v.
The subgroup H described above is in fact the entire group of sym-
metries. To show this, let α ∈ Sym(Γ) be an arbitrary symmetry. Once
we know where α moves the vertices u, v, and w, we have determined α.
Since the action of H is vertex transitive, there is an a ∈ H that takes
u to α(u). Thus a−1 α is a symmetry that fixes u. If a−1 α exchanges the
front–back pair of vertices with the left–right pair of vertices, then there
is a b ∈ H that fixes the top–bottom pair of vertices and also exchanges
the front–back pair of vertices with the left–right pair of vertices. Taking
b to be the identity if a−1 α did not make this exchange, one sees that
b−1 a−1 α fixes u and takes the front–back pair of vertices to itself, and
similarly with the left–right pair. Thus there is a c ∈ (Z2 )3 < H such
that c−1 b−1 a−1 α fixes u, v, and w. But then c−1 b−1 a−1 α = e, hence
α = abc ∈ H.3
Once one has this description of Sym(Γ), it is not difficult to argue
that not only is Sym(Γ) vertex transitive, it is flag transitive.
Fig. 1.8. This is the Petersen graph. A 4-vertex subgraph, a null graph, has
been highlighted.
white vertices to white vertices and black vertices to black vertices. (This
is described more symbolically by writing g · V◦ = V◦ and g · V• = V• .)
Prove that Sym+ (Kn,m ) = Symn ⊕ Symm . What can you say about the
full symmetry group Sym(Kn,m )?
(Hint: It matters whether or not n = m.)
Fig. 1.9. The dihedral group D6 acts on the regular hexagon. The stabilizer
of any grey vertex is isomorphic to Z2 . The stabilizer of the center point is all
of D6 . Any point which is not indicated and which is not on one of the dotted
lines, has a trivial stabilizer.
Notice, however, that if x and x are two vertices that are not directly
opposite each other, then Stab(x) = Stab(x ). If x is the center of the
n-gon then Stab(x) ≈ Dn . If x is any point that is not a vertex, the
center of an edge, or on a line connecting the center of the polygon to a
vertex or the center of an edge, then Stab(x) = {e}.
Orb(x) = {g · x | g ∈ G}.
Proof. The order of G is the product of the order of Stab(x) and the
index of Stab(x) in G. But the index of Stab(x) is the number of left
cosets of Stab(x), which by Theorem 1.33 is the size of the orbit of x.
1.5.1 Generators
Definition 1.37. If G is a group and S is a subset of elements, then
S generates G if every element of G can be expressed as a product of
elements from S and inverses of elements of S. A group G is finitely
generated if it has a finite generating set.
As you can see in the examples above, a given group can have dif-
ferent, interesting, and fairly distinct generating sets. Although it is a
guideline that is frequently broken, it is best to not refer to a particular
set of generators as the generators of a group. Authors often circumvent
this issue using phrases such as “the natural generating set” or “the
generating set we are considering.”
We end this list of examples with a group where it is easy to show
that the group admits no finite set of generators.
an associated path:
1s 2 3s n s s
• −→• −→ • −→ · · · −→ •
e s1 s1 s2 s1 s2 ···sn =g
Here we have been sloppy, and have not bothered to take orientation into
account. If any of the letters si in the expression g = s1 s2 · · · sn is the
inverse of an element of S, then one would be travelling opposite to the
orientation of the directed edge joining s1 s2 · · · si−1 to s1 s2 · · · si−1 si .
To avoid confusion between elements of G and vertices in ΓG,S , let
the vertex in ΓG,S associated to g be denoted vg .
The graph ΓG,S is locally finite because the set S is finite. In fact,
each and every vertex will be incident with exactly 2|S| edges. Then
vg is the initial vertex of exactly one directed edge labelled s for each
s ∈ S (going from vg to vg·s ) and it is the terminal vertex of exactly one
directed edge labelled s for each s ∈ S (going from vg·s−1 to vg ).
The proof of Cayley’s Basic Theorem describes the left action of G
on the vertices of this graph: the element g ∈ G sends the vertex vh
to vg·h .
Does this action on vertices extend to an action on the edges of ΓG,S ?
Notice that the vertex vh is joined to vh·s by a directed edge labelled s.
The element g ∈ G takes the vertex vh to vg·h and the vertex vh·s to
vg·h·s . Thus we may define the action of g ∈ G on the edges by stating
the edge labelled s joining vh to vhs is sent to the edge labelled s joining
vgh to vghs . (See Figure 1.11.)
hs ghs
s s
h gh
Example 1.44. In Figure 1.12 we show the Cayley graph of Sym3 with
respect to the generating set {(12), (123)}. The version on the left follows
the definition precisely. However, we notice that when a generator is an
involution (an element of order 2) there is little need to have a “parallel”
pair of directed edges. It is a standard convention to replace the two
directed edges with a single undirected edge, as is done in the figure on
the right. While this is not what is described by the definition of a Cayley
graph, it is often more aesthetically pleasing and easier to understand.
When this convention is employed, the claim that every vertex is of
degree 2|S| is no longer true. Further, the graph is then a strange hybrid
of directed and undirected edges.
(12) (12)
e e
Fig. 1.12. The Cayley graph of Sym3 with respect to {(12), (123)}, drawn
using two different conventions on how to treat generators of order 2.
Exercise 1.45. You may have noticed that Figures 1.1 and 1.12 are
quite similar. However, the vertices corresponding to group elements
have been slightly shifted. Figure out what is similar and what is different
22 Cayley’s Theorems
in these two figures. Along the way, determine what the action of Sym3
on its Cayley graph looks like.
Fig. 1.13. Two Cayley graphs of D4 . On the left is the graph with respect
to a generating set consisting of a rotation and a reflection; on the right the
generating set consists of two adjacent reflections.
One can construct these Cayley graphs directly from the action of
Dn on a regular n-gon. In Figure 1.14, on the left, we have chosen a
1.6 More Cayley Graphs 23
point whose stabilizer is trivial, so that the elements in the orbit of this
point correspond to the six elements of D3 (Corollary 1.36). We have
also indicated two particular reflections to be used as the generating
set of D3 . One can then connect the points in the orbit to form the
Cayley graph with respect to the indicated generators. We start with
our original point, the only non-solid point in the figure, and apply the
two generators to it, in order to get the intermediate figure. If ◦ denotes
our original point, then the top-right point is a · ◦. By definition, in the
Cayley graph this point is joined by a dashed edge to the vertex ab · ◦.
Noticing that ab is a clockwise rotation through an angle of 120◦ , one
can begin to verify that the picture shown on the right in Figure 1.14 is
indeed accurate.
Fig. 1.14. The dihedral group D3 is generated by the reflections a and b. The
associated Cayley graph can be constructed by connecting points in an orbit
of a point in “general position.”
Fig. 1.15. The Cayley graph of the symmetric group Sym4 with respect to the
generating set S = {(12), (23), (34)}.
If you have studied the Archimedean solids you might notice that this
Cayley graph looks like the edges of a truncated octahedron. This is no
accident. The following argument – similar to the one that embedded
Cayley graphs of dihedral groups into the interior of regular n-gons –
explains why.
First, represent Sym4 as the group of symmetries of the regular tetra-
hedron T. To do this, identify the four vertices of T with the numbers
[4] = {1, 2, 3, 4}. Any symmetry of T results in a permutation of [4],
which is how one constructs a homomorphism φ : Sym(T) → Sym4 .
One can prove that φ is onto by showing that the three transpositions
(12), (23) and (34) are in the image. The reflection across the plane
spanned by the edge joining vertex 3 to vertex 4, and the midpoint
of the edge joining 1 to 2, is taken by φ to the transposition (12). (See
1.6 More Cayley Graphs 25
1 1
4 4
3 3
2 2
Figure 1.16.) Similar reflections yield the generating set for Sym4 .
Because the two groups both have order 24, and φ is onto, it must
be the case that φ is an isomorphism.
To draw the Cayley graph with respect to S = {(12), (23), (34)} we
pick a point in T whose stabilizer is trivial. By drawing in some of the
fixed sets for the reflections corresponding to these transpositions, we
see a triangular region on the exterior of the tetrahedron, shown on the
right in Figure 1.16, any of whose points fit the bill. Picking a point, and
applying the reflections corresponding to (12) and (23) we see a hexag-
onal portion of this Cayley graph. (This subgraph illustrates the fact
that this subgroup is isomorphic to D3 .) Using the same point and the
reflections corresponding to (12) and (34) we see a rectangular portion
of the Cayley graph. These are highlighted in Figure 1.17. The hexagon
corresponding to the generators (23) and (34) is not shown in this figure,
but it sits just a bit below the vertex labelled 1. From these facts it is
not hard to establish that the Cayley graph we are constructing has its
underlying structure coming from the edges of an Archimedean solid.
4 4
3 3
2 2
Fig. 1.17. Parts of the Cayley graph of Sym4 shown embedded in the tetra-
hedron.
Fig. 1.18. A cube with the fixed planes for reflections indicated.
To find the Cayley graph, we use the Drawing Trick, using any point
inside our chosen triangle. The reader should check the following claims:
The portion of the Cayley graph around each vertex of forms a
hexagon; each edge of has a corresponding square in the Cayley graph;
and each face of has an octagon corresponding to the D4 stabilizer of
that face. From this one sees that the Cayley graph is the edge set of
the great rhombicuboctahedron (see Figure 1.19).
they stay well away from the more exotic exhibits. We will encounter
some of these later in the text but, for now, let’s just look at one of the
basic beasts.
Fig. 1.20. Two Cayley graphs of Z. The top is with respect to S = {1}, the
bottom with respect to S = {2, 3}.
The Cayley graph of Z with respect to the single cyclic generator {1}
is a combinatorial version of the real line. If one uses a generating set
such as {2, 3} the Cayley graph is a bit more complicated, but somehow
it still seems to keep the basic “long and skinny” shape of the real line.
(See Figure 1.20.)
If you generate the free abelian group Zn using “the standard basis
vectors”
edges connect points that are exactly a distance 1 apart. The case where
n = 2 is shown in Figure 1.21.
Remark 1.50. Cayley restricted his attention to the Cayley graphs
of finite groups, and this restriction also occurs in many elementary
presentations of Cayley graphs. In 1912 Max Dehn demonstrated the
utility of Cayley graphs, and a related construction called Gruppenbilder,
in the study of finitely generated infinite groups [De12]. This work of
Max Dehn is often cited as the start of the combinatorial and geometric
study of finitely generated infinite groups.
does not immediately imply anything about the group G. On the other
hand, this result also shows that viewing groups via their actions on
graphs is a very flexible approach to the study of group theory.
Theorem 1.51. Let ΓG,S be the Cayley graph of a group G with respect
to a finite generating set S. Consider ΓG,S to be decorated with directions
on its edges and labellings of its edges, corresponding to the generating
set S. Then Sym+ (ΓG,S ) ≈ G.
1. F is closed;
2. the set {g · F | g ∈ G} covers the graph Γ; and
3. no subset of F satisfies properties (1) and (2).
Fig. 1.22. A wheel graph Γ and two examples of fundamental domains for the
action of Z6 on Γ, the action being by rotations. Any fundamental domain
for this action must include the central vertex but, after that, there are many
choices as to how to construct a fundamental domain.
1. v ∈ C; and
2. if x and y are distinct vertices in C then there is no element g ∈ G
such that g · x = y.
The vertex v by itself satisfies these two conditions, so there are such
subgraphs in Γ. Let C0 ⊂ C1 ⊂ C2 ⊂ · · · be a sequence of subgraphs
satisfying our conditions, one properly contained in the next. If there
are only finitely many orbits of vertices under the action of G, and the
graph Γ is locally finite, then at some point one of the Ci ’s in the sequence
above will be a maximal subgraph satisfying the conditions above. In this
32 Cayley’s Theorems
case, call this subgraph the Core. Otherwise form a maximal subgraph
satisfying our conditions by defining
Core = Ci
1. v ∈ C0 ⊂ Core; and
2. if x and y are distinct vertices in Core then there must be some
Ci that contains them. Thus there is no g ∈ G taking x to y.
Claim: The image of Core under the action of G contains all the vertices
of Γ.
Assume to the contrary that there is a vertex v ∈ Γ that is not con-
tained in some g · Core, g ∈ G. Let {v = v0 , v1 , . . . , vn } be the vertices
in a minimal-length edge path joining v to a vertex in G · Core. It must
be the case that vn−1 is not in G · Core (since otherwise there would
be a shorter path to G · Core), which implies that vn−1 is a vertex
outside of G · Core which is joined to G · Core by a single edge. So
we may assume that v ∈ Γ \ G · Core and there is an edge e joining
v to some vertex in g · Core for some g ∈ G. (In other words, we can
now assume “vn−1 ” is “v.”) But then we could add g −1 · v and g −1 · e to
Core to create a larger subgraph satisfying the properties listed above,
contradicting the assumption that Core is a maximal subgraph with
these properties. Thus if v ∈ Γ then v ∈ g · Core for some g ∈ G.
S = {g ∈ G | g = e and g · F ∩ F = ∅}
Fig. 1.23. A bipartite graph Γ where there are two orbits of vertices under
the action of Sym(Γ). A fundamental domain for the action Sym(Γ) Γ is
shown on the right.
Exercise 1.56. Let Γ be the graph shown in Figure 1.23 and let F be
the indicated fundamental domain for the action of Sym(Γ) ≈ D3 . Show
that the generating set given by Theorem 1.55 consists of two adjacent
reflections.
1 4
2 3
Fig. 1.24. A fundamental domain for the action of Sym5 on the complete
graph with five vertices.
The idea of finding fundamental domains for group actions, and using
them to find generating sets, is much more general than what we have
1.9 Words and Paths 37
ω −1 = x−1 −1 −1 −1
k xk−1 · · · x2 x1 ∈ {S ∪ S
−1 ∗
} .
End of path
Identity
Fig. 1.26. The edge path pω associated to the word ω = xxy −1 x−1 yyyxxx.
1. finite edge paths in the Cayley graph ΓG,S which begin at the
vertex corresponding to the identity and
2. words in the free monoid {S ∪ S −1 }∗ .
Exercises 39
Because of this connection between words and paths, the Cayley graph
of a group G is often thought of as a calculator for G. Let g and h be
elements of G and let ωg and ωh be words in {S ∪ S −1 }∗ representing
these elements. To compute the product g · h, one can follow the edge
path pg described by ωg and then, starting at the vertex corresponding to
g instead of the vertex corresponding to the identity, trace out the edge
path described by the word ωh . Continuing with the Z × Z example, if
g = y 2 x3 and h = y −3 x then g·h = y 2 x3 y −3 x, as is shown in Figure 1.27.
From this picture it is clear that g · h can be expressed as x4 y −1 .
Identity
gh
Exercises
(1) Enumerate all of the cycles in the graph of Figure 1.4.
(2) Construct infinitely many distinct biregular graphs where the white
vertices have degree 2 and the black vertices have degree 3.
(3) Construct a graph Γ where Sym(Γ) is vertex transitive but not
edge transitive.
(4) Construct a graph whose symmetry group is the cyclic group Zn
for each n ≥ 3.
(5) There are regular graphs whose symmetry groups are edge transi-
tive but not vertex transitive. (The smallest example is the “Folk-
man graph,” pictures of which are easily found on the Internet.)
Prove that any such graph must be bipartite. (A graph is regular
if the valence of every vertex is the same.)
40 Cayley’s Theorems
(6) Let Pn be the graph formed by the edges of an n-prism (for n ≥ 3).
The graphs P3 , P4 and P5 are shown in Figure 1.28. Prove that
Sym(Pn ) ≈ Dn ⊕ Z2 when n = 4. What goes wrong when n = 4?
(15) Let n be the n-cube formed by taking the convex hull of the
points in Rn whose coordinates are (±1, . . . , ±1). Let Sym(n ) be
the associated symmetry group.
a. What is the order of Sym(n )?
b. Find a set of generators for Sym(n ).
c. Draw the Cayley graph of Sym(3 ) with respect to your
chosen set of generators.
(16) The alternating group A4 consists of the twelve even permutations
of {1, 2, 3, 4}. The permutations (123) and (12)(34) generate A4 .
Draw the associated Cayley graph.
(17) Let G = Sym(K3,3 ).
a. Show that G is vertex transitive.
b. Show that G is flag transitive.
c. Show that G is 2-transitive. That is, elements of G can take
any embedded arc of combinatorial length 2 to any other
such arc.
d. Show that G is also 3-transitive.
e. Is G 4-transitive?
(18) Let Γ be a graph and let G Γ.
a. If the action of G on Γ is vertex transitive, show that, for
any v and w in V (Γ), Stab(v) ≈ Stab(w).
b. If you didn’t already show it above, show that Stab(v) is
conjugate to Stab(w).
(19) Given an edge e in a graph Γ there are two reasonable ways to
define the stabilizer of e. One is to say
Stab(e) = {g ∈ G | g · e = e}
Fig. 1.29. The Doyle graph, which is vertex and edge transitive, but not flag
transitive.
(20) Prove the following Corollary to Theorem 1.51: For every finitely
generated group G, there is a graph Γ, neither directed nor edge
labelled, where Sym(Γ) ≈ G.
(21) Let H be the subset of permutations of the set of integers, SymZ ,
such that h ∈ H if and only if there is a finite set of integers
C ⊂ Z and a number k such that h(n) = n + k if n ∈ C. In other
words, the elements of H all look like translations outside of a finite
set of numbers. (Note: the subset C and the number k are both
dependent on which element h you are examining.)
a b
Fig. 2.1. The reflections a (across the solid line) and b (across the dashed line)
generate an infinite group.
44
Groups Generated by Reflections 45
a b
(The reader who finds Proposition 2.2 interesting should work through
Exercise 3 at the end of this chapter.)
tiling, and consider the group generated by the three reflections whose
axes of reflection are the (extended) sides of T. In Figure 2.3 we have
g b
Fig. 2.3. A tiling of the Euclidean plane where the edges in the tiling have
one of three types associated to them.
highlighted one such triangle; we refer to the reflections across the sides
of this triangle as r, g and b, as is indicated. As a quick exercise, the
reader should verify that r, rgr and rbr are the reflections in the sides
of the triangle immediately above T in this tiling. For example, in rgr
the rightmost “r” flips the triangle above T down to T’s position; the
middle “g” flips T to the left; then the leftmost r flips g · T up to the
correct position.
{T = T0 , T1 , T2 , · · · , Tn = T }
48 Groups Generated by Reflections
ρ1 ρ3
ρ2
r
g b
Fig. 2.4. A corridor joining two triangles in the tiling of the Euclidean plane.
"r"
b "g"
Fig. 2.5. The element brg moves T to the triangle highlighted on the right.
the vertices of the Cayley graph. When the edges of the Cayley graph
are added, one sees that the Cayley graph is the vertices and edges of
the regular hexagonal tiling of plane that is dual to the original tiling
by equilateral triangles. This is shown in Figure 2.6.
There are two other tilings of the Euclidean plane that lead naturally
to infinite groups generated by three reflections. One can tile the plane
by triangles with interior angles {π/2, π/4, π/4} and one can also tile
the plane by triangles with interior angles {π/2, π/3, π/6}. The groups
generated by reflecting in the sides of such triangles are similar to the
one we have discussed above, but they have their individual quirks.
At this point we have a reasonably long list of “groups generated by
reflections.” There are the finite dihedral groups as well as D∞ and the
groups generated by reflections in the sides of certain Euclidean triangles.
The symmetry groups of all five Platonic solids are also groups generated
by reflections, as are the symmetry groups of cubes in all dimensions.
There is a well-developed theory that contains these groups and many
Exercises 51
The reader who is interested in Coxeter groups can learn the basics by
reading [Hu90]. An excellent book introducing the deep combinatorial
information associated to Coxeter groups should read [BjBr05]. A reader
with a stronger background in mathematics than we are assuming in this
text, who is interested in the geometry and topology of Coxeter groups,
should read [Da08].
Exercises
(1) Prove Proposition 2.1. This is perhaps easiest to do if you view
a as the function a(x) = −x and b as the function b(x) = 2 − x,
remembering that the group operation is function composition.
52 Groups Generated by Reflections
Fig. 2.7. The group W333 contains an index 4 subgroup that is isomorphic
to W333 .
(2) Recall that T2,2 is the regular 2-valent, bipartite tree. Prove that
D∞ ≈ Sym+ (T2,2 ).
(3) Let the two reflections generating D∞ be denoted a and b, and use
the same conventions for the action D∞ R as in Proposition 2.1.
a. Prove that the subgroup a, babab is isomorphic to D∞ and
is of index 3 in the original group.
b. Prove that for any n ≥ 1 the group D∞ contains a subgroup
that is isomorphic to D∞ and of index n.
c. Prove that if α is any automorphism of D∞ then α takes a
and b to reflections fixing x = n and x = n + 1, for some
n ∈ Z.
d. Prove that the center of D∞ = {e} and conclude that D∞ ≈
Inn(D∞ ) < Aut(D∞ ).
e. Show that the index of Inn(D∞ ) in Aut(D∞ ) is 2. (Hint:
There is an automorphism given by φ(a) = b and φ(b) = a,
which is not an inner automorphism.)
f. Prove that Aut(D∞ ) ≈ D∞ .
(4) Determine how the element rgb acts on the tiling of the Euclidean
plane by regular triangles. What about gbr? or brg?
(5) Tile the Euclidean plane by squares of side length 1. Let W be
the group generated by the four reflections in the (extended) sides
of any one square. Draw the Cayley graph of W and prove that
W ≈ D∞ ⊕ D∞ .
(6) Prove that Z ⊕ Z is a subgroup of W333 .
(7) Draw the Cayley graphs of W244 and W236 .
Exercises 53
54
3.1 Free Groups 55
a1 · · · ai−1 ai a−1
i ai+1 · · · ak ∼ a1 · · · ai−1 ai+1 · · · ak
Exercise 3.3. Show that every equivalence class contains exactly one
freely reduced word from {x1 , . . . , xn , x−1 −1 ∗
1 , . . . , xn } .
The reason for introducing the equivalence relation, and not simply
stating that freely reduced words are the elements of Fn , is that the con-
catenation of two freely reduced words is not necessarily freely reduced.
y
x x
y
y y
x x v x x
y y
y
x x
Since every group acts on its Cayley graph (Theorem 1.51) we imme-
diately get:
The group F2 acts on T4 on the left. So, for example, the generator x
takes the vertex corresponding to a freely reduced word ω to the vertex
corresponding to the freely reduced word equivalent to x · ω. One can
3.1 Free Groups 57
write a formula for this action. Let ω be a freely reduced word, and let
ω = α · ω , where α is a single letter. Then the action of x is given by
xω if x−1 = α
x · ω = x · α · ω =
ω if x−1 = α.
The action of x ∈ F2 on T4 is something of a “shift to the right.” To see
this, consider the subgraph of T4 induced by the vertices corresponding
to {xn | n ∈ Z}. Left multiplication by x shifts this combinatorial line
one unit to the right, thereby shifting the entire Cayley graph, as is
illustrated in Figure 3.2. The action of y is similar, the main difference
being that the combinatorial line induced by vertices corresponding to
{y n | n ∈ Z} is shifted one unit up.
Fig. 3.2. Left multiplying by x moves the vertical florets to the right, as in
the left-hand side of the picture; left multiplying by y moves the horizontal
florets up.
v
w
Fig. 3.3. The edge path connecting v and w in the Cayley graph of F2 .
x –1
y y
v v
x –1 x –1
y y
v v
y –1 y –1
x w
w · Xy ⊂ Xw for y = w−1 ,
Remark 3.8. The fact that SL2 (Z) contains a free subgroup is not an
isolated fact. There is a celebrated theorem of Jacques Tits [Ti72] that
says: Let G be a finitely generated group that can be faithfully represented
as a group of matrices.1 Then either G contains a finite index subgroup
that is solvable, or G contains a subgroup isomorphic to F2 . The reader
unfamiliar with solvable groups can interpret this result as saying: linear
groups contain free subgroups unless they obviously do not.
Fig. 3.6. The graph of the function f , which takes points in the open intervals
{i + (1/5, 4/5) | i ∈ Z} to points in the open intervals {i + (4/5, 1) | i ∈ Z}.
in the grid formed by Xf in the domain and in the range. In this case,
if x ∈ (i − 4/5, i − 1/5) then f −1 (x) ∈ (i − 1, i − 4/5), so f −1 (x) ∈ Xf .
Further, since f −1 takes (i − 1, i − 4/5) into (i − 1, i − 4/5), f −n (x) ∈ Xf
for any n ∈ N and x ∈ Xf .
Our second function is just a shifted copy of f . Define g(x) to be
f (x − 1/2) + 1/2 and let
Xg = [i + 1/2 − 1/5, i + 1/2 + 1/5].
i∈Z
Then the same arguments as are given above show that, for any non-zero
integer n, positive or negative, g n (x) ∈ Xg if x ∈ Xg .
We can now establish our claim that the subgroup of Homeo(R) gen-
erated by f and g is a free group, using an argument that is similar to
the proof of Proposition 3.7.
The method used to establish Propositions 3.7 and 3.9 is often referred
to as the Ping Pong Lemma. The version of the Ping Pong Lemma that
was used to establish 3.9 is stated as:
interested in the diverse places where free groups arise, [Gl92] is a nice
expository article on this topic.
3.2 F3 is a Subgroup of F2
It is easy to see that a free group of rank 2 sits as a subgroup inside
a free group of rank 3 (or rank 4, 5, ...). If {x1 , x2 , . . . , xn } is a basis
for Fn , then no freely reduced word in {x1 , x2 , x−1 −1 ∗
1 , x2 } can equal the
identity. Thus the subgroup generated by {x1 , x2 } is a free group of rank
2. On the other hand, the following result is surprising.
This notion of length has the useful property that |g| = |g −1 |, since
the freely reduced expression for g −1 is the formal inverse of the freely
reduced expression for g.
Let H be the subset of F2 consisting of elements of even length:
H = {g ∈ F2 | |g| is even}.
Since |g| = |g −1 |, this set is closed under taking inverses. Because cancel-
lation in non-freely reduced words occurs in pairs, if |a| and |b| are both
even, |ab| will also be even, although it could be less than |a| + |b|. Thus
H is closed under products and inverses, and therefore it is a subgroup
of F2 . We call H the even subgroup of F2 . Finally, because the length of
every element in F2 is either even or odd, [F2 : H] = 2.
Notice that in this example, while there is cancellation in c−1 a and b−1 a,
there is no cancellation when one rewrites ac−1 in terms of the original
generators.
Converting ω from a freely reduced word in {a, b, c, a−1 , b−1 , c−1 }∗ to
a word in {x, y, x−1 , y −1 }∗ is done by replacing each wi by αi βi where
αi , βi ∈ {x, y, x−1 , y −1 }. In the example above, with the word c−1 ac−1 b,
we had β1 = x−1 = α2−1 and β3 = x−1 = α4−1 . Because of this there
was a small amount of internal cancellation, but it ends up that such
constrained cancellations are the worst that can happen.
−1
1. αi = βi+1 ;
−1
2. βi+1 = αi+2 ; and
−1
3. βi−1 = αi .
We can now show that the even subgroup is a free group with basis
{a, b, c}. We first convert a given freely reduced word
|α1 β1 · α2 β2 · · · αn βn | ≥ n,
hence w = e ∈ F2 .
3.3 Group Presentations 67
Thus the element φ(ω) ∈ G does not depend on which word is used to
represent ω. It follows immediately from the definition that φ(ω −1 ) =
−1
[φ(ω)] and φ(ω · ω ) = φ(ω)φ(ω ). So the function φ is indeed a group
homomorphism.
There is also the following result, which is the foundation for what are
called group presentations.
Remark 3.18. Because relations are words that represent the identity,
there is a one-to-one correspondence between relations and paths in the
Cayley graph that begin and end at the vertex associated to the identity.
G = g1 , . . . , gn | ω1 , . . . , ωm
D∞ = a, b | a2 , b2
Dn = a, b | a2 , b2 , (ab)n .
n ≡ 0, 2, 9 or 11(mod 12);
Fig. 3.7. Two triangles of hexagons are shown: On the left is T4 and on the
top right is T2 . On the bottom right is the line of hexagons L3 .
1. v ∈ C; and
2. if x and y are distinct vertices in C then there is no element g ∈ G
such that g · x = y.
Fig. 3.8. A fundamental domain for the action of F2 on its Cayley graph is a
cross of four half-edges. For each of these half-edges, there is a corresponding
subset Te ⊂ T . One such subtree is highlighted above.
Fig. 3.9. A Core and fundamental domain for the action of the even subgroup
on the Cayley graph of F2 .
After removing inverses we see that {x2 , xy, xy −1 } is a basis for the even
subgroup of F2 .
We end this section with a striking result about free groups.
Corollary 3.23 (Nielsen–Schreier Theorem). Every subgroup of a free
group is free.
Proof. Let H be a subgroup of a free group F. The group F acts freely
on its Cayley graph, and therefore H acts freely on the Cayley graph of
F. Thus, by the characterization of free groups given by Theorem 3.20,
H is a free group.
This result was first proved by Jakob Nielsen for finitely generated
free groups in 1921. His result was extended to all free groups by Otto
Schreier a few years later.
Fig. 3.10. The element x2 can be chosen to be part of a basis for the even
subgroup of F2 since x2 · F intersects F at the rightmost point of F. Because
xy·F intersects F at the highest point highlighted above, and xy −1 ·F intersect
F at the lowest point highlighted above, we know that {x2 , xy, xy −1 } is a basis.
a
e
b
NLN NRS
NLW NRW
NL NR
N
WRW NLS NRN ELW
WRN ELS
WR EL
WRS ELN
W E
WLN ERS
WL ER
WLS ERN
WLW SRN SLS ERW
S
SR SL
SRW SLW
SRS SLN
Fig. 3.12. Labelling the vertices of T3,4 using left/right and the four cardinal
directions.
76 Groups Acting on Trees
The element a fixes the origin, and otherwise it applies the permuta-
tion N ⇒ E ⇒ S ⇒ W ⇒ N to the first letter labelling a vertex v. So,
as examples, under the action of a the vertex labelled NRW is taken to
the vertex labelled ERW and the vertex labelled WLS is taken to the
vertex labelled NLS.
The action of b is a bit more complicated to describe in terms of these
coordinates. In order to avoid repeatedly writing “the vertex labelled by
” we conflate vertices with their labels. The vertex E is fixed by b. The
origin is taken to EL; the vertex EL is taken to ER; and ER is taken to
the origin. The rules for determining the action of b on the remaining
vertices are:
1. If a vertex has “EL” as a prefix, it is taken to the vertex obtained
by replacing this prefix by “ER.”
2. If a vertex has “ER” as a prefix, strip off this prefix.
3. If v has any other prefix, then add EL as a prefix. (The vertex
WL is taken to ELWL, for example.)
To get a little bit of practice with this group action, you should show:
1. The element aba−1 is a rotation about the vertex labelled S in
Figure 3.12.
2. The element a2 ba−2 is a rotation about the vertex W.
3. The element a3 ba−3 is a rotation about the vertex N.
in Figure 3.12. By reading a word ω = ai1 bj1 · · · ain bjn left-to-right one
can easily determine the image of e under the action of ω.
Assuming that i1 = 0, one rotates e about the white vertex fixed by
a as far as i1 dictates. Denote the resulting edge e . Rotate e around
its black vertex as far as is dictated by j1 . Denote the resulting edge
e , and then rotate e about its white vertex as far as is dictated by
i2 , and so on. Similarly, if i1 = 0, and so ω really should be expressed
as ω = bj1 ai2 · · · ain bjn , then one applies the same process, starting by
rotating about the black vertex of e as far as j1 dictates, then rotating
about the resulting white vertex as far as i2 dictates, and so on. For
example, if ω = a2 ba then the image of e under the action of ω is the
edge f shown in Figure 3.13.
e
b
a2
a
f
Fig. 3.13. The element a2 ba in Z3 ∗ Z4 takes the edge e to the edge f . The
arrows are labelled using the “two wrongs make a right” principle.
But b actually rotates about the vertex labelled E. These two errors
together give the right answer by the “two wrongs make a right” principle
that we previously encountered in our discussion of reflection groups
(Chapter 2). You can see why this works if you notice that ω can be
rewritten as a product of conjugates:
ω = a2 ba = (a2 bab−1 a−2 )(a2 ba−2 )a2
The rightmost a2 gives the rotation about the unlabelled vertex; the term
(a2 ba−2 ) is the rotation falsely labelled “b” in Figure 3.13; the final ex-
pression (a2 bab−1 a−2 ) is the rotation falsely labelled as “a”
Figure 3.13.
This “two wrongs make a right” principle is actually just an appli-
cation of a broader principle, namely, “conjugation is doing the same
thing somewhere else.” If g ∈ Z3 ∗ Z4 takes the unlabelled vertex fixed
by a to some vertex v, then gaj g −1 will be a rotation about v. Thus
gaj = (gaj g −1 )g will take the unlabelled vertex to v (via g) and then
rotate about v (via gaj g −1 ).
This description of how to determine where the edge e goes under
the action of ω = ai1 bj1 · · · ain bjn is the key insight for establishing the
following lemma.
Lemma 3.24. The group Z3 ∗ Z4 acts transitively on the edges of T3,4 .
Further, let e be the edge joining the vertex stablized by a to the
vertex stabilized by b, and let g = ai1 bj1 · · · ain bjn (as prescribed above)
with n = ||ai1 bj1 · · · aik bjk ||. Then the number of vertices between e and
g · e is n.
Proof. Given any edge f we can use the method described above to write
down an element of gf ∈ Z3 ∗ Z4 that will take e to f . Thus the product
gf · gf−1
will take an edge f to f , so the action is edge-transitive. The
claim about the norm determining the distance is immediate from the
description of how one finds the image of e under the action of g.
Proposition 3.25. The action of Z3 ∗ Z4 on T3,4 is simply transitive
on the edges. Further, any g ∈ Z3 ∗ Z4 can be uniquely expressed as
a word σ s1 τ t1 · · · σ sn τ tn where s1 and/or tn can be zero, but otherwise
1 ≤ si ≤ 2 and 1 ≤ ti ≤ 3.
Proof. That the action is transitive is established by Lemma 3.24. To
establish that it is simply transitive, it suffices to show that no element
of Z3 ∗ Z4 can have two such expressions:
g = si1 tj1 · · · sin tjn = sk1 tl1 · · · skm tlm .
3.6 Free Products of Groups 79
But then by Lemma 3.24 the norm of this product must be zero. This is
only possible if tjn · t−lm = e, that is, jn = lm , and then sin · s−km = e,
hence in = km , and so on until there has been complete cancellation.
This contradicts the assumption that g has two distinct expressions.
then πω1 = πω2 , as these permutations do not agree on the reduced word
∅. So no two distinct, reduced words can represent the same element
of A ∗ B.
Corollary 3.27. The groups A and B inject into A ∗ B.
Proof. The elements of A and B form reduced words (of length 1).
Theorem 3.28. Every free product of groups A ∗ B can be realized as
a group of symmetries of a biregular tree T , and a fundamental domain
for this action consists of a single edge and its two vertices. Further, if
A and B are both finite, this tree is T|A|,|B| .
Proof. Define V (T ) to be the the left cosets gA and gB. In order to have
a bit of notation, we let vgA denote the vertex associated to the coset
gA and similarly vgB is the vertex associated to gB.
The edges of the tree T correspond to the elements of A ∗ B, and we
use eg to denote the edge associated to g ∈ A ∗ B. This does have the
unfortunate consequence that the edge associated to the identity is “ee .”
We define Ends(eg ) = {vgA , vgB }. The bipartite structure of T comes
from considering the vertices associated to cosets of A as one “part” and
the vertices associated to cosets of B as the other “part.”
Note that two edges eg and eh intersect if and only if g −1 h ∈ A or B.
For example, if eg ∩ eh = vkA then g ∈ kA and h ∈ kA, hence g −1 h ∈ A.
Conversely, if g −1 h ∈ A then gA = hA and so vgA = vhA . The same
argument holds mutatis mutandis for B.
The action of A ∗ B on T is induced by the standard left action of
a group on itself and on its left cosets. An element g ∈ A ∗ B moves
vertices by g : vhA → vghA and g : vhB → vghB ; the edges move by
g : eh → egh . As in the case of the action of a group on any of its Cayley
graphs, we get an action G T primarily because the group is per-
muting labels by left multiplication, while incidence is defined by right
multiplication. It follows from the description of this action that the edge
ee , which joins vA to the vB , is a fundamental domain for the action of
A ∗ B on T .
To prove that the graph T is a tree we need to show that it is
connected and contains no cycles. If g ∈ A ∗ B then g = [ω] for a
reduced word ω = x1 x2 · · · xn . Assume for convenience that x1 ∈ A,
x2 ∈ B, and so on. Then ee ∩ ex1 = vA , ex1 ∩ ex1 x2 = vx1 B , and so
on. Hence {ee , ex1 , ex1 x2 , . . . , ex1 x2 ···xn } is an edge path from ee to eg . It
follows that any edge eg can be joined to ee by an edge path, hence T is
connected.
82 Groups Acting on Trees
E → e, S → a, W → a2 , and N → a3
along with
L → b, and R → b2 .
The edge joining the origin to the vertex associated to E is the edge ee
from Theorem 3.28; the edge joining the origin to S is ea ; and the edge
joining W L to W LS is ea2 ba , as is shown in Figure 3.13. The action of a
and b on T3,4 , described in Section 3.5 in terms of adjusting labels, can
be reconstructed by simply applying the above translation scheme with
left multiplication.
Exercise 3.29. Pick two finite groups, A and B, that you are comfort-
able with. Construct the tree T , and the action A ∗ B T , given by
Theorem 3.28. Be as concrete as possible.
φ(g) = (a1 a2 a3 , b1 b2 ) ∈ A ⊕ B.
This map cannot be injective (if A and B are non-trivial). Take any two
non-identity elements a ∈ Z \ {eA } and b ∈ B \ {eB }, and form their
commutator, which is defined as
c. A ≈ A and B ≈ B.
(ω) = x1 · · · xk an bn a−1 −1
n+1 bn+1 an+1 bn+1 ,
ω ∼ x1 · · · xk an bn b−1 −1
n+1 an+1 bn+1 an+1 .
(ω) = x1 · · · xk an ba−1
n+1 bn+1 an+1 ,
a = an a−1
n+1 ∈ A \ {eA }
and then
(ω) = x1 · · · xk abn+1 an+1 ,
Theorem 3.20 also gives a concrete way to find a basis for K using a
fundamental domain. In describing a fundamental domain for the action
K T3,4 we use the notation introduced in Section 3.5. In particular, a
is the generator of order 4 and b is the generator of order 3. To construct
a Core for K T3,4 we begin with the vertex we called “the origin” and
we add edges and vertices, until we can no longer add vertices coming
from distinct orbits.
Fig. 3.14. A Core and fundamental domain for the action of the kernel K on
T3,4 . The dotted line indicates two half-edges that are connected by the action
of K.
One possible Core is shown in Figure 3.14. At first glance, our Core
may appear to be too small. All of the vertices in Core are contained
in ai · ee or bj · ee , where ee is the edge joining the vertex fixed by a to
the vertex fixed by b. The vertex labelled NL in Figure 3.12 is an end of
a3 b · ee , and the image of a3 b in Z3 ⊕ Z4 is not the same as the image
of a single power of a or b. However, the vertex labelled NL is also an
end of a3 ba · ee , and the image of a3 ba is the same as the image of b in
the quotient. One further check that the Core, and the corresponding
fundamental domain formed by adding half-edges, is correct comes from
3.9 Finite Groups Acting on Trees 89
are called leaves of the tree.) The action of G must permute the leaves
of TO(v) , hence it also permutes all the non-leaf vertices of TO(v) . Let
T be the tree formed by removing the leaves of TO(v) and the edges
incident to these leaves. We now have an action of G on T . If T is not
an edge or a single vertex, then we can remove the leaves of T to form
a smaller subtree T on which G acts.
Because G is finite, O(v) and TO(v) are finite as well. Each time we
trim off leaves, we create even smaller subtrees, and so by continuing in
this manner we eventually reduce our original subtree to either a single
vertex or a single edge. If it is an edge, then either all of the elements
of G fix the edge, or some elements of G invert the edge. In either case,
the center point of the edge is fixed.
Proof. The vertex stabilizers are conjugates of A and B for the action
of A ∗ B on its associated tree, by Corollary 3.30. By Corollary 3.47, a
finite subgroup of A ∗ B must fix a vertex, hence H < gAg −1 or gBg −1
for some g ∈ A ∗ B.
Fix(g) = {x ∈ X | g · x = x}.
3.10 Serre’s Property FA and Infinite Groups 91
Nous nous intéressons aux groupes G ayant la propriété: (FA) – Quel que soit
l’arbre X sur lequel opère G, on a X G = ∅.
Proof. Assume Fix(G) is not empty. If v and w are two distinct fixed
points in Fix(G), then there is exactly one minimal-length edge path
containing v and w. Since this edge path is unique, it must also be fixed
(pointwise) by G. Thus Fix(G) is a connected subgraph of T , hence it
is a subtree.
Lemma 3.51. Let G act on a tree T and let H and K be two subgroups
of G. Then
Fix(H, K) = Fix(H) ∩ Fix(K).
For the rest of this section we will explore the group of automorphisms
of a free group Fn , with the goal of establishing that Aut(Fn ), for n ≥ 3,
3.10 Serre’s Property FA and Infinite Groups 93
has Serre’s property FA. In order not to take up too many pages, our
exposition will be fairly terse. Any reader who finds this discussion in-
teresting is encouraged to fill in the missing details, and to consult the
various references mentioned.
Let S = {x1 , . . . , xn } be a basis for Fn and let S = S∪S −1 be the union
of this basis and its inverses. There are a number of automorphisms of
Fn that come from permutations of S. For example, if σ ∈ Symn is a per-
mutation of {1, . . . , n} then one can define an associated automorphism
σ ∈ Aut(Fn ) induced by requiring
σ(xi ) = xσ(i)
(where we are abusing the use of the symbol “σ”). To verify that this
is indeed an automorphism, notice that σ must be surjective, since the
image of S is S. Further, it must be injective, since freely reduced words
are mapped to freely reduced words, hence no non-identity element is
mapped to the identity. The subgroup of Aut(Fn ) generated by these
permutation automorphisms is isomorphic to Symn .
Similarly one can exchange generators and their inverses. That is, for
each i ∈ {1, . . . , n} there is an automorphism ιi ∈ Aut(Fn ) induced by
requiring
−1
xj if i = j
ιi (xj ) =
xj otherwise.
The subgroup generated by these inversion automorphisms is isomorphic
to (Z2 )n ≈ Z2 ⊕ · · · ⊕ Z2 . The subgroup of Aut(Fn ) generated by the
permutation and inversion automorphisms injects into Sym(S), hence
it must be finite. It is, in fact, a group of order 2n · n!. We denote this
subgroup of Aut(Fn ) by Ωn .
Jakob Nielsen introduced another important collection of automor-
phisms, which we call Nielsen automorphisms. These are the automor-
phisms {ρij } induced by
xi xj if k = i
ρij (xk ) =
xk otherwise.
That this defines a homomorphism from Fn to Fn is an application of
Theorem 3.15. To see that it is a bijection, one needs to establish that the
image of S = {x1 , . . . , xn } under ρij is a basis for Fn . This is essentially
part (a) of Exercise 6.
The following theorem (first proved by Nielsen in 1924 [Ni24]) is not
too difficult, and the interested reader can find a good exposition of it
94 Groups Acting on Trees
in [MKS66]. The details, however, take too many pages for us to include
here.
Proof. First, we have to show that {H1 ∪H2 ∪H3 } is a generating set for
Aut(Fn ). Let σij denote the permutation automorphism that exchanges
xi and xj . Conjugating the automorphism σ3n ∈ H1 by τ ∈ H3 we get
3.10 Serre’s Property FA and Infinite Groups 95
→ x−1 x−1
τ σ3n τ
x1 1 → 1 → x1
τ σ3n τ
x2 → x3 → xn → xn
τ σ3n τ
x3 → x2 → x2 → x3
τ σ3n τ
xn → xn → x3 → x2
(Every other xi is fixed by both τ and σ3n .) Conjugating ιn ∈ H1 by σ2n
results in ι2 ; conjugating σ3n by σ2n creates σ23 . We then also have ι1 ,
as ι1 = σ23 τ , and σ12 , which is ηι1 2 . Thus the subgroup generated by
the Hi contains all the “transposition automorphisms” σij and all the
inversions ιi , hence it contains Ωn . Since θ ∈ H2 , and ι2 ∈ Ωn , it follows
that ρ12 = θι2 ∈ H1 , H2 , H3 . Thus H1 , H2 , H3 contains Ωn and ρ12 .
Thus, by Corollary 3.55, H1 , H2 , H3 = Aut(Fn ).
As we already know each Hi is finite, it remains to verify that each
Hij = Hi , Hj is finite. This we do, case by case.
The group H13 is generated by elements of Ωn , hence H13 is a subgroup
of Ωn , and is therefore finite.
The group H23 is generated by two elements of order 2, θ and τ . Their
product is the automorphism that takes x1 to x−1 −1
2 x1 , x2 to x3 and x3 to
x−1
2 . Direct computation then shows that θτ has order 4, and therefore
H23 ≈ D4 .
The subgroup generated by Symn−2 and ιn is Ωn−2 , that is, it consists
of all elements of Ωn that fix x1 and x2 . The subgroup H12 is generated
by this Ωn−2 and the elements θ and η. The elements of Ωn−2 commute
with both θ and η so H12 ≈ Ωn−2 ⊕ θ, η. Further, θ and η are both of
order 2, and their product θη has order 3. Thus θ, η ≈ D3 , and H12 is
therefore the finite group Ωn−2 ⊕ D3 .
Lemma 3.58. If G has Property FA, then any quotient of G also has
Property FA.
Exercises
(1) Define the distance between two vertices in a tree to be the number
of edges that occur in the unique reduced path between them.
a. Prove that the number of vertices a distance n from a fixed
vertex v in Tm is m · (m − 1)n−1 .
b. Find a formula for the number of vertices a distance n from
a fixed vertex v in T{m,n} . (You will actually need to find two
formulas, one for the black vertices and one for the white
vertices.)
(2) Prove that Zn is a subgroup of Homeo(R), for all n ∈ N.
(3) In Remark 3.11 it is stated that the subgroup of Homeo(R) gener-
ated by f (x) = x3 and g(x) = x + 1 is a free group.
a. What are f −1 and g −1 ?
b. Show that no freely reduced word in {f, g, f −1 , g −1 }∗ , of
length at most 3, represents the identity.
(4) Prove the Ping Pong Lemma (Lemma 3.10).
Exercises 97
Fig. 3.15. The hyperbolic isometry whose axis is the horizontal line and the
hyperbolic isometry whose axis is the vertical line generate a free subgroup of
the isometry group of the hyperbolic plane.
(5) (This problem is for people who have studied hyperbolic geome-
try.) Consider the hyperbolic plane with six specified geodesics, as
shown in Figure 3.15. Let x be the hyperbolic isometry of the hy-
perbolic plane whose axis is the horizontal line, and which takes
the geodesic on the left to the geodesic on the right. Let y be the
hyperbolic isometry with axis the vertical line, and which takes the
lower geodesic to the upper geodesic. Show that {x, y} generate a
free group.
(6) Let F2 be the free group with basis {x, y}.
(8) Let F2 be a free group with basis {x, y}. For any natural number
n let βn = {y −i xy i | 1 ≤ i ≤ n}. Show that the subgroup of F2
generated by βn is free of rank n.
(9) Show that F2 has exactly three distinct subgroups of index 2.
(10) Let F2 be the free group with basis {x, y}. Define a map φ : F2
Z2 by φ(x) = (1, 0) and φ(y) = (0, 1). Prove that the kernel of φ is
a free group generated by
X = {[xk , y l ] | k, l ∈ Z and k · l = 0}.
Conclude that this kernel is a free group of infinite rank.
(11) Show that Fn ≈ Fm if n = m.
(12) Show that F2 ≈ Z ∗ Z.
(13) Show that A ∗ B is a free group if and only if A and B are free.
(14) Let F2 be a free group with basis {x, y}, and let K be the kernel
of the homomorphism φ : F2 → Z3 induced by φ(x) = φ(y) = 1.
a. Find a fundamental domain for the action of K on the
Cayley graph of F2 .
b. Find a basis for the kernel K.
(15) Let F2 be a free group of rank 2. Show that, for any n ≥ 2, F2
contains a finite-index subgroup isomorphic to Fn .
(16) Let N be an infinite-index normal subgroup of a free group Fn
(n ≥ 2). Show that N is not finitely generated.
(17) Let G = A ∗ B and let g = a1 b1 · · · an bn with the right-hand side
being a reduced expression. Prove that the order of g is infinite.
(18) Let A and B be non-trivial groups. Prove that the center of A ∗ B
is trivial.
(19) Prove Theorem 3.33. You may want to use the proof of Theo-
rem 3.15 as a model for your argument.
(20) The alternating group A4 is generated by (12)(34) and (123). Let
s generate a cyclic group of order 2, let t generate a cyclic group
of order 3, and let Z2 ∗ Z3 be their free product. Finally, let φ :
Z2 ∗ Z3 → A4 be the map induced by s → (12)(34) and t → (123).
a. Show that K = Ker(φ) is a free group.
b. Find a basis for K.
c. If you are ambitious, replace “3” by “n.”
(21) The centralizer of an element g ∈ G, C(g), is the subgroup of G
consisting of all elements that commute with g. Let g ∈ A ∗ B.
a. Show that if g is in a conjugate of A or B, then C(g) is a
subgroup of the same conjugate of A or B.
Exercises 99
The first time I saw the group a, b | ab = b2 a was in a paper of Graham
Higman’s around 1950, in the Journal of the London Mathematical Society.
Actually the group may have been known to others. Anyway, the reference
to Graham Higman is the reason I often complain about this group being
attributed to Donald and myself.
. –Gilbert Baumslag
100
Baumslag–Solitar Groups 101
Thus the orbit of 0, or in fact any element of Z[1/2], is Z[1/2]. This fact
also follows from the following description of this group.
Outline of the proof. We ask the reader to verify the following steps as
a useful exercise.
then
g(x) = a−k bm ak an .
By its construction we know that BS(1, 2) acts on the real line. But
this action is neither discrete nor proper, and the tools developed in pre-
vious chapters seem to rely on, or at least prefer, discrete and proper ac-
tions. For example, the Drawing Trick (Remark 1.49) provides a method
for constructing a Cayley graph of a group G, by using the orbit of a
point which is moved freely by G as the vertices of the Cayley graph.
This method worked nicely in Chapter 2, where we used it to draw Cay-
ley graphs of some groups generated by reflections, but it does not work
well for the action BS(1, 2) R. Because of this, we delay discussing
the Cayley graph of BS(1, 2) with respect to {a, b} until Section 5.4.
Instead, we introduce an alternative set on which BS(1, 2) acts.
Corollary 4.2. The orbit of the closed interval [0, 1] under the action
BS(1, 2) R is the set of closed intervals
Further, if g([0, 1]) = [0, 1] then g = e ∈ BS(1, 2). Thus g ∼ g([0, 1]) is a
bijection between I and the elements of BS(1, 2).
The group BS(1, 2) will show up from time to time in the remainder
of this text. In order to give an example of the importance of Baumslag–
Solitar groups within geometric group theory, we state a result that we
will not prove:
Theorem 4.3 (See [Me72]). There is a surjective group homomorphism
φ : BS(2, 3) → BS(2, 3) (where φ(a) = a and φ(b) = b2 ) that is not an
isomorphism.
Exercises
(1) Fill in the outline of the proof of Proposition 4.1.
(2) Show that ab = bn a in BS(1, n).
(3) Show that Stab(0) = a for BS(1, 2) R.
(4) Let x ∈ Z[1/2] be any dyadic rational. What is Stab(x) under the
action BS(1, 2) R?
(5) State and prove the analog of Proposition 4.1 for BS(1, n).
(6) Prove that the cyclic subgroup of BS(1, 2) generated by the element
b is not a normal subgroup.
(7) Prove that the smallest normal subgroup of BS(1, 2) that contains
b is isomorphic to the dyadic rationals.
(8) Prove that
BS(1, 2) is isomorphic
to the group of matrices gener-
2 0 1 1
ated by and .
0 1 0 1
5
Words and Dehn’s Word Problem
Irgend ein Element der Gruppe ist durch seine Zusammensetzung aus den
Erzeugenden gegeben. Man soll eine Methode angeben, um mit einer endlichen
Anzahl von Schritten zu entscheiden, ob dies Element der Identität gleich ist
oder nicht.
. –Max Dehn
105
106 Words and Dehn’s Word Problem
η : G → {S ∪ S −1 }∗
NF1 = {ai bj | i, j ∈ Z}
Fig. 5.1. Three paths corresponding to the three normal forms NF1 , NF2
and NF3 . Each path joins the vertex associated to the identity to the vertex
associated to a3 b2 = b2 a3 = (ab)3 b−1 ∈ Z ⊕ Z.
words in NF2 move vertically then horizontally. Finally the paths corre-
sponding to the words in NF3 move along the diagonal, and then they
either move directly up or down.
Remark 5.3. When thought of as a collection of paths in a Cayley
graph, a normal form is occasionally referred to as a “combing,” al-
though that word is often restricted to sets of normal forms where the
associated paths satisfy certain geometric and/or language theoretic re-
strictions. The terminology is appropriate, at least for “nice” normal
forms. Consider for example G = Z ⊕ Z4 with generators x = (1, 0) and
y = (0, 1). There is a normal form
N (G) = {xi y j | i ∈ Z and 0 ≤ j ≤ 3} .
If you trace out the paths associated to this normal form, you get some-
thing that looks “combed” (Figure 5.2).
D∞ -rule: If ω ∈ {a, b}∗ and ω contains a subword of the form aa or bb, then
remove this subword to make a new word ω .
This new word represents the same element of D∞ as the word we started
with, so determining if π(ω ) = e is sufficient for determining if the
original word represents the identity. There are, potentially, a number
of choices that could be made in applying this rule, as is illustrated in
Figure 5.2. Any word that results from such a process is equivalent to
w – in the sense that its image under the evaluation map is π(w) – and
the final word one gets is either empty or an alternating sequence of a’s
and b’s. But Proposition 2.1 shows that all alternating sequences of a’s
and b’s yield non-identity elements of D∞ . Thus the word w represents
the identity in D∞ if and only if w reduces to the empty word by repeated
application of the D∞ -rule.
There is a variant of Dehn’s word problem that asks not only if one
can determine when a word represents the identity, but: given a group
G, a finite generating set S, and two words ω, ω ∈ {S ∪ S − }∗ , can one
determine if ω and ω represent the same element of G? In other words,
can you determine if π(ω) = π(ω )? This is referred to as the “equality
problem.”
110 Words and Dehn’s Word Problem
baaaababb
baababb baaaaba
bbabb baaba
abb bba
Fig. 5.3. There are a variety of choices that can be made in showing that
baaaababb = a ∈ D∞ . Thicker arrows with solid arrowheads indicate when
multiple choices result in the same word.
Proof. Since the identity is represented by the empty word, if one can
solve the equality problem then one can determine when a given word ω
gives the same element as the empty word, hence one can solve Dehn’s
word problem.
Conversely, ω and ω represent the same element in G if and only if
ω · ω −1 represents the identity element. Thus if one can solve Dehn’s
1. the identity;
2. an element in the center Z(G);
3. an element in the commutator subgroup;
4. an element of finite order.
We will not construct examples of such groups in this book, but they
are of great importance in a variety of fields and they have practical
applications to areas such as coding theory. For more information, see
the survey article [Mi92].
Example 5.8. The word problem for Dn (with respect to the generating
set {f, ρ} where f is a reflection and ρ a rotation through an angle of
2π/n) can be solved. To determine if a word ω represents the identity,
simply trace out the corresponding edge path pω in the Cayley graph Γ.
The word ω represents the identity if and only if the end of the path pω
is at the vertex in Γ that corresponds to the identity.
This example indicates that the word problem can be solved in any
group where the Cayley graph can be constructed. Since our primary
focus is on infinite groups, we need an appropriate notion of what it
means to say a Cayley graph “can be constructed” if we hope to be able
to convert this example to a theorem. A reasonable answer is that one
can build arbitrarily large pieces of the Cayley graph, where the idea of
“arbitrarily large pieces” is codified by the notion of balls in a Cayley
graph.
112 Words and Dehn’s Word Problem
Fig. 5.4. The sphere of radius 3 in the Cayley graph of the reflection group
W333 (see Chapter 2) consists of all the vertices with a dark halo; the ball of
radius 3 is indicated by thicker edges.
Fig. 5.5. The ball of radius 2 in the Cayley graph of Z ⊕ Z consists of all the
dark vertices and solid arrows.
Fig. 5.6. Part of the Cayley graph of Z3 ∗ Z4 with respect to the generating
set {a, b}.
composing the appropriate linear functions. In the end one has g(x) = x
if and only if π(ω) = e. For example, consider the word
ω = a2 ba−2 ba2 b−1 a−2 b−1 ,
and let g = π(ω). Then
Given any vertex in the Cayley graph, one can form a combinato-
rial line by right multiplying by positive and negative powers of b. Call
such a line a b-line in the Cayley graph of BS(1, 2). Each such line
will have one set of edges labelled a coming into it and two sets of
edges labelled a pointing away from it, corresponding to whether the
exponent of b is odd or even. Figure 5.9 illustrates the fact that two
distinct sheets come together at each b-line. These sheets merge below
the b-line (the portion corresponding to a-edges which point into the b-
line) but are distinct above the b-line. This is why, informally speaking,
a side view of the Cayley graph of BS(1, 2) looks like a trivalent tree
(Figure 5.10).
At this stage the reader is well-poised to think about our characteriza-
tion of BS(1, 2) given in Corollary 4.2. There it is shown that there is a
one-to-one correspondence between elements of BS(1, 2) and the images
of the unit interval [0, 1] under the action of BS(1, 2). Let φ : Γ → R be
the map induced by sending the vertex associated to bn to n ∈ R and
by requiring that the image of any edge labelled a is a single point, as
indicated in Figure 5.11. We leave the proof of the following insight to
the exercises.
118 Words and Dehn’s Word Problem
Fig. 5.10. The Cayley graph of BS(1, 2) projects onto a trivalent tree.
–2 –1 0 1 2
Exercises
(1) Three different normal forms for Z ⊕ Z are given in Example 5.2.
Draw the Cayley graph of Z ⊕ Z and the paths corresponding to
the words in NF1 , NF2 and NF3 for the elements:
a. (3, 5)
b. (−2, −1)
c. (1, 6)
(2) Find a normal form for F2 ⊕ F2 .
(3) Use colored pencils to draw spheres of radius n, for n = 1, 2 and 3,
in the Cayley graph of:
a. the cyclic group Z6 (with respect to a cyclic generator).
b. the symmetric group Sym4 (with respect to the adjacent
transpositions {(12), (23), (34)}).
c. the reflection group W333 from Chapter 2 (with respect to
the three reflections in the sides of the equilateral triangle).
d. BS(1, 2) (with respect to {a, b}).
(4) Prove that if you can solve the word problem for groups A and B
then you can solve it for G = A ⊕ B.
(5) Prove that if you can solve the word problem for groups A and B
then you can solve it for G = A ∗ B.
(6) Solve the word problem for the reflection group W333 from Chap-
ter 2.
(7) Show that π(a3 b2 a−3 ba3 b−2 a−3 b−1 ) = e ∈ BS(1, 2).
(8) Prove Proposition 5.13.
(9) Construct the Cayley graph of BS(1, 3).
(10) In addition to the word problem, Max Dehn also introduced the
following conjugacy problem in his 1912 paper.
Let S be a finite generating set for a group G. Let ω and ω be words in
{S ∪ S −1 }∗ . Can one determine if π(ω) and π(ω ) are conjugate?
Solve the conjugacy problem in the infinite dihedral group D∞ .
(11) Show that you can solve the conjugacy problem (see Exercise 10)
in A ∗ B if you can solve it in A and B.
6
A Finitely Generated, Infinite Torsion
Group
Today this is known as the Weak Burnside Problem and, while it appears
to be an elementary question, it took over sixty years to resolve. On the
face of it, one would probably believe the answer is “yes.” The infinite
groups we have encountered so far all have elements of infinite order.
This is true for Zn , Fn , A ∗ B where A and B are non-trivial groups, and
BS(1, 2). However, in this chapter we construct an example that shows
that the answer is actually “no.”
Z2 ⊕ Z2 ⊕ Z2 ⊕ · · ·
and the factor group Q/Z is also an infinite torsion group. But in nei-
ther of these examples is the group finitely generated. (Establishing these
claims is your Exercise 1.) To construct our example of a finitely gener-
ated, infinite torsion group we use a variation on the theme of Chapter 3
120
A Finitely Generated, Infinite Torsion Group 121
0 1 2
00 01 02 10 11 12 20 21 22
We let T denote the rooted ternary tree for the remainder of this chap-
ter. We label the vertices of this tree using finite sequences n1 n2 · · · nk
where each ni ∈ Z3 . The root corresponds to the empty sequence; the
vertex associated to n1 · · · nk nk+1 is a child of the vertex associated to
n1 · · · nk ; and the vertex associated to n1 · · · nk nk+1 · · · nl is a descendant
122 A Finitely Generated, Infinite Torsion Group
let ωn1 ···nk denote a waterfall based at vn1 ···nk . More formally, ωn1 ···nk
can be written as an infinite product of shuffles, where one adds n1 · · · nk
as a prefix to all the indices used in describing ω. For example, ω21
would be
1 −1 1 −1 1 −1
ω21 = σ210 σ211 σ2120 σ2121 σ21220 σ21221 ···
Using this notation we notice the following formulas, which the reader
can verify:
ω = σ0 σ1−1 ω2 ,
σωσ −1 = ω0 σ1 σ2−1 , (6.1)
2
σ ωσ −2
= σ0−1 ω1 σ2 .
The following is another set of formulas that are useful in the proofs
of Lemmas 6.4 and 6.5:
ωσ = σ(σ 2 ωσ −2 ),
(σωσ −1 )σ = σω, (6.2)
2 −2 −1
(σ ωσ )σ = σ(σωσ ).
124 A Finitely Generated, Infinite Torsion Group
yk−1 yk = ωσ = σ(σ 2 ωσ −2 ).
We will need to refer to the process described above in proofs that oc-
cur below. To facilitate this, we define an operation Σ on words x1 · · · xn ∈
{ω, σωσ −1 , σ 2 ωσ −2 } by first defining Σ on the three basic elements
Σ(ω) = (σ 2 ωσ −2 ),
Σ(σωσ −1 ) = ω,
Σ(σ 2 ωσ −2 ) = (σωσ −1 ),
ωωσωσωσωωσ → σΣ(ω)Σ(ω)ωσωσωωσ
→ σ 2 Σ2 (ω)Σ2 (ω)Σ(ω)ωσωωσ
→ σ 3 Σ3 (ω)Σ3 (ω)Σ2 (ω)Σ(ω)ωωσ
→ σ 4 Σ4 (ω)Σ4 (ω)Σ3 (ω)Σ2 (ω)Σ(ω)Σ(ω)
= σ(σ 2 ωσ −2 )2 (ω)(σωσ −1 )(σ 2 ωσ −2 )2
A Finitely Generated, Infinite Torsion Group 125
where the final equality comes from applying Σ and using the fact that
the group element σ and the operation Σ both have order 3.
Proof. Since σ and ω send the vertices of level k to the vertices of level k
(for any k ≥ 0), the group U sends vertices at level k to vertices at level
k. Thus for a fixed k there is a map from U to Sym3k . In this proof we
focus on the action of U on the three vertices {v0 , v1 , v2 }. The symmetry
σ induces the permutation (123) (on the indices) while ω induces the
trivial permutation. Thus the only permutations that occur in the image
of φ are
Image of φ = {e, (123), (132)} = A3 ;
g = (g0 , g1 , g2 ) ∈ U0 × U1 × U2 .
gi = σij x1 x2 · · · xmi
The construction given above was first presented by Gupta and Sidki
[GuSi83]; it is inspired by a construction of Grigorchuk [Gr80]. The first
examples of finitely generated, infinite torsion groups were found in the
1960s; a discussion of the first known example can be found in [Ol95].
The example presented here does not completely settle all questions that
128 A Finitely Generated, Infinite Torsion Group
can be derived from the problem Burnside posed in 1902. In fact, there
is a stronger version, often called the Burnside Problem, which asks:
If n is a fixed positive integer and G is a finitely generated group where every
element in G satisfies g n = e, is G finite?
Burnside noted that this problem is quite tractable for n = 2, where the
answer is “yes,” and established that the claim is also true for n = 3.
Burnside’s question can be further refined by asking if for given in-
tegers k and n (both ≥ 2) there is an infinite group with a generating
set of size k where every element satisfies g n = e. In fact, there is a
“free Burnside group” B(k, n) which is the quotient of the free group of
rank k, Fk , by the smallest normal subgroup containing the nth powers:
{xn | x ∈ Fk }. There is an infinite group with generating set of size k
where every element satisfies g n = e if and only if B(k, n) is infinite.
Burnside proved that B(k, 3) is finite, for all k, but did not determine
the order of this group. In 1933 Levi and van der Waerden proved that
the order of B(k, 3) is 3n where
k k
n=k+ + .
2 3
(Question: Why does the fact that the group B(2, 3) has order 33 not
contradict our claim that the group U < Sym(T ), generated by σ and
ω, is infinite?)
In 1940 Sanov showed that B(k, 4) is always finite, and in 1958 Mar-
shall Hall showed that the groups B(k, 6) are all finite and was able to
give a concrete formula for the size of B(k, 6), similar to the formula
above. It is still an open question if B(2, 5) is finite or infinite.
On the other hand, for large values of n there are finitely generated
infinite groups where every element satisfies g n = e. This was first shown
by Novikov and Adian in 1968, who showed that B(k, n) is infinite when
k > 2 and n is an odd number ≥ 665.
There is yet another refinement, formulated by Wilhelm Magnus in
the 1930s:
For a fixed choice of k and n, is there is a largest, finite k generator group in
which every element satisfies g n = e?
A rephrasing of this question asks if B(k, n) has only finitely many finite
quotients. In 1991 Efim Zelmanov showed that this so-called restricted
Burnside Problem is true; this is work that earned him a Fields Medal
in 1994.
Exercises 129
Exercises
(1) Prove that the following groups are infinite torsion groups, but
neither is finitely generated.
a. A direct sum of a countable number of copies of Z2 :
Z2 ⊕ Z2 ⊕ Z2 ⊕ · · · .
The elements of this group consist of all N-tuples of elements
of Z2 , where only finitely many entries are non-zero.
b. The factor group Q/Z.
(2) Show that any finitely generated group G, where every element
satisfies g 2 = e, is a finite group.
(3) Prove Case 2 of Lemma 6.7.
(4) Show that ωσ has order 9.
(5) Let T be the rooted, infinite binary tree and let Sym(T ) be the
group of all symmetries of T .
a. Show that any α ∈ Sym(T ) sends the root to the root, even
if you just view T as an unrooted tree.
b. Show that Sym(T ) contains an index 2 subgroup isomorphic
to Sym(T ) ⊕ Sym(T ).
c. Show that Sym(T ) is uncountable. (Hence, by Exercise 11
in Chapter 1, Sym(T ) is not finitely generated.)
(6) Does Sym(T ) contain any element of infinite order?
(7) Let Tk be the regular, unrooted, k-valent tree, where k ≥ 3. Show
that Sym(Tk ) is not finitely generated. (Hint: See Exercise 5.)
(8) Let U be the group constructed in this chapter.
a. Prove that for any g ∈ U, g = e, there is a finite group H
and a homomorphism φ : U → H such that φ(g) = e ∈ H.
(Hint: The group U acts on the set of vertices at level k.)
b. Prove that the intersection of all finite index subgroups of
U is the trivial subgroup {e}.
7
Regular Languages and Normal Forms
130
7.1 Regular Languages and Automata 131
S a a
b b
a S a
b b
a a a
S b a S b a
b b b
Fig. 7.3. Two automata whose languages are the reduced words describing
elements in D∞ .
on the left. In the FSA on the right, words containing aa and bb do not
correspond to directed paths.
b
a
a S ε
a
c
The assumption that there is exactly one start state is not always part
of the definition of DFAs. As we will see in Theorem 7.12, in terms of
the languages accepted by DFAs, the two conventions are equivalent.
Similarly, the following result indicates that one can work with DFAs
or complete DFAs and, in terms of the resulting languages, there is no
difference. The proof of Lemma 7.9 is left as an end-of-chapter exercise.
Proof of Theorem 7.11. (1) By Lemma 7.9 we may assume that the lan-
guage K = L(M) where M is a complete deterministic automaton. The
language S ∗ − K is then accepted by the machine M◦ formed by taking
M and making all non-accept states into accept states, and making all
accept states non-accept states.
(2) Let MK and ML be deterministic automata whose languages are
K and L respectively. The disjoint union of MK and ML is a non-
deterministic automaton (with two start states) that accepts K ∪ L. By
Theorem 7.12, K ∪ L is then regular.
136 Regular Languages and Normal Forms
X x
y y y
X x
X S x
Y Y Y
x
X
1. |u| < n;
2. uv i w ∈ L for all i ≥ 0.
the paths is how many times they wind around the closed path from z
to z.
v
w
u
S Z
Example 7.15. The Pumping Lemma can be used to prove that the
language L = {an bn | n ≥ 1} is not a regular language. Assume to
the contrary that L is regular and let n be the number of states of an
automaton M whose language is L. Then the path corresponding to
an+1 contains a closed subpath, of some length k. It follows that the
word an+1+k bn must also be accepted by M, contradicting the claim
that L = L(M).
Definition 7.16. Let L be a language in S ∗ and let ω ∈ S ∗ . The cone
type of ω is the set of all words ω ∈ S ∗ such that ω · ω ∈ L. Denote the
cone type of ω by Cone(ω). The collection
Cone(L) = {Cone(ω) | ω ∈ S ∗ }
is the cone type of the language L.
Example 7.17. Let S = {a} and let L be the language of all words of
even length. Then Cone(a) = Cone(a3 ) = Cone(a5 ) = · · · and they are
all equal to the set of words of odd length, while Cone() = Cone(a2 ) =
Cone(a4 ) = · · · are all the words of even length. Thus Cone(L) consists
of exactly two sets, the set of words of even length and the set of words
of odd length.
Theorem 7.18 (Myhill–Nerode Theorem). A language L is regular if
and only if its set of cone types is finite.
Proof. We first consider the forward direction. If L is regular then it is
accepted by a complete DFA M. For each vertex v ∈ V (M) let Mv be
7.2 Not All Languages are Regular 139
the automaton formed by making v the only start state. The cone type
of a given word ω is then determined by the end vertex of the associated
path pω . If this vertex is denoted vω then Cone(ω) = L(Mvω ). Thus the
number of cone types is at most the number of vertices in M.
To establish the converse, assume that Cone(L) is finite. We construct
an automaton M starting with declaring the states of M to be the cone
types of L. Designate Cone(ε) to be the unique start state, where ε
represents the empty word. The accept states of M are the cone types
that contain ε.
We add an edge labelled x from the vertex given by Cone(ω) to
Cone(ωx) for each cone type and each x ∈ S. There is a possibility
that this definition depends on the word ω, not on the cone type of ω.
That is, perhaps Cone(ω) = Cone( ω ), but Cone(ω · x) = Cone(
ω · x).
∗
This cannot happen, because if x ∈ S and ω ∈ S then
Cone(ωx) = {ω ∈ S ∗ | xω ∈ Cone(ω)} .
ω ) then
Thus if Cone(ω) = Cone(
Cone(ωx) = {ω ∈ S ∗ | xω ∈ Cone(ω)}
= {ω ∈ S ∗ | xω ∈ Cone(
ω )} = Cone(
ω x)
L = {ωω | ω ∈ S ∗ }
Thus WP(G, S) is the set of all words in the generators and their inverses
that evaluates to the identity in G. This set is often colloquially referred
to as “the word problem” for G, since being able to solve Dehn’s word
problem for G (and generating set S) amounts to being able to determine
if a given word ω ∈ {S ∪ S −1 }∗ is in WP(G, S).
Since WP(G, S) ⊂ {S ∪ S −1 }∗ , it is a language, and one can ask:
When is WP(G, S) a regular language? Surprisingly, it is not easy to
come up with examples where the word problem is a regular language.
For example, let G be an infinite cyclic group with generator x. Then a
word in {x, x−1 }∗ represents the identity if and only if the total exponent
sum is zero. But then x−i ∈ Cone(xi ) but x−i ∈ Cone(xj ) for any i and
j ∈ Z where j = i. So WP(G, x) has infinitely many cone types, and the
Myhill–Nerode Theorem (Theorem 7.18) implies that WP(G, x) is not a
regular language. Faced with this non-example, you might start looking
at free groups, free products, the Baumslag–Solitar group BS(1, 2), etc.,
but you would only be wasting your time.
a a τ
S b a S b a
b τ b a
Fig. 7.7. Converting the finite-state automaton giving a regular normal form
for D∞ with respect to {a, b} to one using {a, τ, τ −1 } as its generating set.
π(uv i w) = π(uv j w)
⇒ π(u)π(v i )π(w) = π(u)π(v j )π(w)
⇒ π(v)i = π(v)j
Proof. The given set is in the kernal, and the quotient by the normal sub-
group containing these elements is isomorphic to Z. So it suffices to show
that the subgroup of F2 generated by S is normal. But the following com-
putations show that the conjugate of each element in S is in the subgroup
144 Regular Languages and Normal Forms
generated by S
x · xi y −i · x−1 = xi+1 y −(i+1) · yx−1 = xi+1 y −(i+1) · (xy −1 )−1 ,
x−1 · xi y −i · x = xi−1 y −(i−1) · y −1 x = xi−1 y −(i−1) · (x−1 y)−1 ,
y · xi y −i y −1 = yx−1 · xi+1 y −(i+1) = (xy −1 )−1 · xi+1 y −(i+1) ,
y −1 · xi y −i · y = y −1 x · xi−1 y −(i−1) = (x−1 y)−1 · xi−1 y −(i−1) .
So the subgroup generated by S is normal. Exercise 7 asks you to show
that S forms a basis.
The following example, combined with Theorem 7.32 below, indicates
that determining which subgroups of a given group G are finitely gen-
erated can be rather subtle.
Example 7.31. The intersection of two finitely generated subgroups of
a group G is not necessarily a finitely generated subgroup. Consider the
group F2 ⊕ Z with {x, y} being the basis for F2 and z a cyclic generator
of Z. Then the kernel K of the map ϕ : F2 ⊕ Z → Z induced by
ϕ(x) = ϕ(y) = ϕ(z) = 1 ∈ Z
is a free group generated by {x−1 z, y −1 z}. The proof of this claim is
similar to the proof of Lemma 7.30: show that the subgroup generated
by this set is normal and note that the quotient is indeed Z. Normality
is established by equations of the form:
y · x−1 z · y −1 = yz −1 · x−1 z · y −1 z.
Working out the details is left to Exercise 13.
The intersection of K with direct factor F2 is the kernel of the map
φ : F2 → Z considered in Lemma 7.30 above:
ker(ϕ) ∩ F2 = ker(φ).
The subgroups K = ker(ϕ) and F2 are both finitely generated but, as
we have seen, ker(φ) is not finitely generated.
Theorem 7.32 (Howson’s Theorem). The intersection of finitely gen-
erated subgroups of a free group is a finitely generated free group.
Our proof of Howson’s Theorem will exploit our understanding of
regular languages. If G is a group with finite generating set S, then a
subgroup H < G is the image of a regular language over S ∪ S −1 if
there is a regular language L ⊂ {S ∪ S −1 }∗ such that π(L) = H. This
condition is independent of one’s choice of finite generating set (see the
analogous argument in the proof of Proposition 7.24). Moreover, there
7.5 Finitely Generated Subgroups of Free Groups 145
z(e)
b (e)
S e
a (e)
Fig. 7.8. The spanning tree and other featured elements of the automaton.
empty. It follows from the definitions that α(e1 ) = g0 and more generally
α(ei+1 ) = β(ei )gi . Thus we may write
π[ω] = π[α(e1 )h1 β(e1 )−1 · α(e2 )h2 β(e2 )−1 · · · α(en )hn β(en )−1 · z].
If we let
(e) denote the label of an edge e ∈ M then the image of L is
contained in the subgroup generated by
E = {π(z), π[α(e)
(e)β(e)−1 ] | e is an edge in M \ T }.
It then remains to be shown that each of these elements is contained in
H, for then E will be a finite set of generators for H. Since z ends in an
accept state, π(z) ∈ H. If e is an edge in M \ T let z(e) be a path from
the target vertex of e to the accept state. Then
π[α(e)eβ(e)−1 ] = π[α(e)ez(e) · (β(e)z(e))−1 ].
But π[α(e)·e·z(e)] and π[(β(e)·z(e))] are both in H, so π[α(e)·e·β(e)−1 ]
is in H.
The proof given above, when restricted to the case of subgroups of
free groups, establishes the following result:
Corollary 7.34. Finitely generated subgroups of free groups are free.
(This result is, of course, an immediate corollary of the Nielsen–Schreier
Theorem 3.23, but the formal language argument given below is useful
none-the-less.)
Proof. If H is a finitely generated subgroup of Fn then one direction of
Theorem 7.33 implies that H is the image of a regular language L. The
7.5 Finitely Generated Subgroups of Free Groups 147
E = {π(z), π[α(e)
(e)β(e)−1 ] | e is an edge in M \ T }
and their inverses. But any such word which is freely reduced when
written as a product of elements of E ∪E −1 cannot represent the identity
in Fn . Consider for example the product
−1
α(e1 )
(e1 )β(e1 )−1 · α(e2 )
(e2 )β(e2 )−1
= α(e1 )
(e1 )β(e1 )−1 · β(e2 )
(e2 )−1 α(e2 ).
If β(e1 )−1 ·β(e2 ) cancels completely, that means that the initial vertex of
e2 is the initial vertex of e1 . But since this expression was freely reduced
when written in E ∪ E −1 , e1 = e2 . Thus no more cancellation occurs. An
induction argument along these lines establishes that a freely reduced
word written with n letters from E ∪E − is equivalent to a freely reduced
word of length at least n, when written in terms of the generators of Fn .
Thus
E = {π(z), π[α(e)
(e)β(e)−1 ] | e is an edge in M \ T }
is free means they consist of all freely reduced words representing ele-
ments of H and K. The intersection LH ∩ LK is then a regular language
consisting of all freely reduced words which represent elements in H and
in K, hence LH ∩ LK maps onto H ∩ K. So H ∩ K is finitely generated
by Theorem 7.33.
In the argument above, being able to claim that the languages consist
of freely reduced words is critical. In general, if LH and LK are two
regular languages whose images are H and K respectively, then LH ∩LK
maps to H ∩ K, but it does not have to map onto H ∩ K. There may be
elements in the intersection that have different words representing them
in LH and LK .
Howson proved Theorem 7.32 in 1954. The topological techniques he
used were quite different than the language-theoretic approach given
above. In addition to knowing that H ∩ K is finitely generated it would
be nice to have a method for bounding the rank of H ∩ K. In his paper
Howson established:
rk(H ∩ K) ≤ 2rk(H)rk(K) − rk(H) − rk(K) + 1.
In papers published a few years later Hanna Neumann improved How-
son’s bound, and conjectured that the rank of H ∩ K is at most
(rk(H) − 1)(rk(K) − 1) + 1.
This question is still open, and is referred to as the Hanna Neumann
Conjecture. Hanna Neumann’s son, Walter Neumann, has suggested a
strengthened version of the conjecture [Ne90], and this strengthened
version can be formulated as a purely graph theoretic conjecture [Di94].
Remark 7.36. The reader who would like to continue exploring con-
nections between formal language theory and the study of infinite groups
might begin by consulting Bob Gilman’s survey article [Gi05].
Exercises
(1) Show that the language {ak | k ≡ 0 or 2 (mod 5)} is a regular
language.
(2) Prove Lemma 7.9: If L is the language associated to a determi-
nistic automaton, then L is the language associated to a complete,
deterministic automaton.
(3) Let L be a language. Define its reverse language to be
Rev(L) = {xn · · · x1 | x1 · · · xn ∈ L}.
Exercises 149
g = σ k x1 x2 · · · xn
(13) Let G = F2 ⊕ Z with {x, y} being the basis for F2 and z a cyclic
generator of Z. Prove that the kernel K of the map φ : F2 ⊕Z → Z,
induced by φ(x) = φ(y) = φ(z) = 1, is a free group generated by
{x−1 z, y −1 z}.
(14) The even subgroup of F2 is a free group of rank 3. In Exercise 14
in Chapter 3 you showed that the kernel of φ : F2 → Z3 is a free
group of rank 4. What is the rank of their intersection?
8
The Lamplighter Group
A short drive from Lafayette College (in Easton, Pennsylvania) one can
find barns decorated in a classic Pennsylvania Dutch style with Hex
diagrams placed around on the sides of the barn. In Figure 8.1 there is
a simplified version of this, where copies of a single Hex diagram have
been placed at the corners of a square.
151
152 The Lamplighter Group
The exercise above shows that G is not the direct product of (Z6 )4
and Sym4 ; G is, however, a “semi-direct product” of (Z6 )4 and Sym4 ,
denoted (Z6 )4 Sym4 . To see the distinction between direct and semi-
direct products, we should examine how the binary operation works. In
order to emphasize that we are not working with a direct product, we
denote elements of G using square brackets:
Thus we have:
The key insight is that, in moving the permutation (34) past the ro-
tations (0, 0, 2, 1), we actually apply the permutation to the 4-tuple of
rotations. This example, we hope, provides some insight into and moti-
vation for the following construction.
154 The Lamplighter Group
The action of H on the direct sum defined is via the standard Cayley
action of H on itself. Thus, given g ∈ h∈H G, the element h ∈ H
permutes the entries of g by taking the entry in position h to the position
h · h (for every h ∈ H).
As an elementary example, consider Z2 Z3 . This group is a semi-direct
product (Z2 )3 Z3 where the associated automorphism
φ : Z3 → Aut(Z2 ⊕ Z2 ⊕ Z2 )
is given by declaring that φ(1) ∈ Aut(Z2 ⊕ Z2 ⊕ Z2 ) is the cyclic permu-
tation α(a, b, c) = (b, c, a).
Exercise 8.8. Verify the following equation:
[(0, 1, 1), 1] · [(1, 0, 0), 1] = [(0, 1, 0), 2] ∈ Z2 Z3
The remainder of this chapter is devoted to understanding a single
example of this wreath product construction, Z2 Z. This group is a
semi-direct product:
Z2 Z = Z2 Z = (· · · ⊕ Z2 ⊕ Z2 ⊕ Z2 ⊕ · · · ) Z.
i∈Z
integers. The empty set would then correspond to the identity element,
0 = (. . . , 0, 0, 0, . . .) ∈ A. The binary operation in A is coordinate-wise
addition in Z2 . Viewing elements of A as subsets of Z, we see that this
corresponds to the symmetric difference
S T = {n ∈ Z | n ∈ S or n ∈ T but n
∈ S ∩ T }.
In particular, {−2, 0, 1} {−3, 0, 4} = {−3, −2, 1, 4}. Thus every element
in L2 can be represented by an ordered pair [S, n] where S is a subset
of integers, and n ∈ Z. The binary operation is given by
[S, n] · [T, m] = [S (T + n), n + m]
where T + n = {t + n | t ∈ T }. The identity element is [∅, 0].
As the subgroup A < L2 is an infinite direct sum, the following lemma
may at first seem a bit surprising.
Lemma 8.9. The lamplighter group L2 can be generated by two ele-
ments: one of order 2 and the other of infinite order.
Proof. Let t be the element corresponding to [∅, 1] ∈ L2 and let a be the
element [{0}, 0] ∈ L2 . The action of Z on A can be described by noticing
that φ(1) ∈ Aut(A) has the effect of adding 1 to each coordinate. Thus
ta = [∅, 1] · [{0}, 0] = [∅ {1}, 1 + 0] = [{1}, 1]
and more generally
tn a = [∅, n] · [{0}, 0] = [{n}, n]
for any n ∈ Z. Thus
tn at−n = [{n}, 0].
It follows that an arbitrary element of the subgroup A can be expressed
as:
[{n1 , n2 , . . . , nm }, 0] = tn1 at−n1 · tn2 at−n2 · · · tnm at−nm .
Further, an arbitrary element of L2 can be expressed as:
[{n1 , n2 , . . . , nm }, k] = tn1 at−n1 · tn2 at−n2 · · · tnm at−nm · tk .
Therefore, the set {a, t} is a generating set for the lamplighter group.
The perspective of this book is geometric, and it is this perspec-
tive that illuminates L2 and explains why it is called the lamplighter
group. We know that an arbitrary element of L2 can be expressed as
[{n1 , n2 , . . . , nm }, k]. To think of this geometrically, take the Cayley
158 The Lamplighter Group
This procedure shows that this element can be expressed as t3 at−2 at. Of
course there are other ways one could express this particular element.
For example, one could follow a slightly different process:
From this we see that the same element can be expressed as tat2 at−1 .
In other words, t3 at−2 at = tat2 at−1 in L2 .
In Figure 8.5 we show a cycle in the Cayley graph of L2 (with respect
to {a, t}) that corresponds to the relation atat−1 = tat−1 a. We have
denoted the vertex corresponding to g ∈ L2 by its picture.
Exercises
(1) Prove Theorem 8.3.
(2) Prove that the symmetry group of a cube (Section 1.6.3) is isomor-
phic to (Z2 ⊕ Z2 ⊕ Z2 ) Sym3 .
(3) A finite subgroup of Aut(Fn ), Ωn , was introduced in Section 3.10.
Prove that Ωn ≈ (Z2 ⊕ · · · ⊕ Z2 ) Symn . (What does this have to
n copies
do with Exercise 2?)
(4) Prove that Symn ≈ An Z2 .
(5) Prove Hol(Z2 ⊕ Z2 ) ≈ Sym4 .
(6) Verify that Figure 8.2 is correct.
160 The Lamplighter Group
Fig. 8.5. A cycle in the Cayley graph of the lamplighter group, where
undirected edges correspond to the generator a and directed edges to the
generator t.
(7) There are five distinct groups of order 12. You have probably en-
countered four of them: Z12 ≈ Z3 ⊕Z4 , Z6 ⊕Z2 , D6 ≈ D3 ⊕Z2 , and
A4 . The fifth group of order 12 can be described using semi-direct
products. Let H = Z3 and let K = Z4 , where H is generated by
h and K is generated by k. Let φ : K → Aut(Z3 ) by φ(k) = α
where α(h) = h2 . Show that G = H φ K is a group of order 12
that is not isomorphic to any of the four previously listed groups of
order 12.
(8) Prove that D4 ≈ Z2 Z2 .
(9) Let G and H be two finite groups. Show that the order of G H is
|G||H| · |H|.
(10) In this exercise we examine the symmetries of rooted binary trees.
(See Chapter 6.)
a. Let T denote the finite, rooted binary tree with eight leaves
shown in Figure 8.6. Prove that Sym(T ) ≈ (Z2 Z2 ) Z2 .
b. Let T denote the infinite, rooted binary tree where every
vertex has two descendants. Prove that Sym(T ) ≈ Sym(T )
Z2 .
Exercises 161
(11) Prove that the wreath product of two finitely generated groups is
finitely generated.
(12) Show that atn at−n = tn at−n a ∈ L2 , for any n ∈ N.
(13) The Cayley graph Γ of L2 with respect to {a, t} is fairly compli-
cated. You can gain some intuition for this by drawing the ball
of radius 3, B(e, 3) ⊂ Γ. You should find that this is a tree, but
you should also realize that B(e, 4) is not a tree, as is shown by
Figure 8.5.
9
The Geometry of Infinite Groups
[E]ssentially all geometric constructs which are global in nature, such as paths
of shortest length, global manifestations of curvature, planes, half-spaces, rates
of growth, which are studied in differential geometry have manifestations in
combinatorial approximations to that geometry.
. –James Cannon
1. d(x, y) ≥ 0,
2. d(x, y) = 0 if and only if x = y,
162
9.1 Gromov’s Corollary, aka the Word Metric 163
where π(ω) = g. Denote this value by |g|. Notice, in particular, that |g|
represents the minimal length of a path from the vertex representing e
to the vertex representing g in the Cayley graph ΓG,S .
As a simple example, if G = Z⊕Z, S = {(1, 0), (0, 1)} is the generating
set, and g = (m, n) ∈ Z ⊕ Z, then the length of G is just |m| + |n|.
This metric on the group G is often referred to as the word metric
and it does depend on the choice of generators. This is the reason that
S appears as a subscript on the distance function. We reserve the right to
be sloppy and drop this subscript from time to time, when the fixed set
of generators is either clear or not of central importance to the argument.
Proof of Gromov’s Corollary. Conditions 1 and 2 have already been es-
tablished, so it only remains to show that the distance function is sym-
metric and that the triangle inequality holds.
The function dS is symmetric: if dS (g, h) = n then there is a word
ω of length n such that g −1 h = π(ω). But then h−1 g = π(ω −1 ) and
so dS (h, g) is at most n. But if there were a shorter word representing
h−1 g, then we could take its formal inverse and form a shorter word
representing g −1 h. Thus dS (h, g) = dS (g, h) for any g, h ∈ G.
Our function dS also satisfies the triangle inequality:
dS (g, k) + dS (k, h) ≥ dS (g, h)
for all g, h and k ∈ G. To see this, let ωgk and ωkh be minimal-length
words such that g · ωgk = k and k · ωkh = h. Thus g · π(ωgk ωkh ) = h,
hence
dS (g, h) ≤ |ωgk | + |ωkh | = dS (g, k) + dS (k, h).
The comments above show that the group G can be viewed as a metric
space. Cayley’s Theorem shows that left multiplication gives an action
of G on itself, and this action preserves distances. That is,
dS (h, k) = |h−1 k| = |h−1 g −1 gk| = dS (gh, gk)
for any g, h, k ∈ G.
Fig. 9.1. The distance between the lower left-hand vertex and the upper right-
hand vertex is 7.
these vertices consists of seven letters: four x’s and three y’s. Such a
word is determined by the placement of the x’s (or by the placement
of the y’s). For example, the edge path going across the bottom of the
figure and then the right edge corresponds to xxxxyyy while a diago-
nal path is given by xyxyxyx. In fact, the union of all edges that occur
in minimal-length paths between these two vertices forms the rectangle
that is the focus of this figure.
R = Max{S ∪ {0}},
L = min{S ∪ {0}}.
The descriptions above are not completely clear when applied to ele-
ments like g = [{1, 3}, 2] ∈ L2 , which is shown in Figure 8.4. The picture
for this element contains no lit lamps to the left of zero. The right-first
normal form for such an element simply moves the lamplighter right,
lighting lamps along the way, and then moves the pointer to the correct
ending position. The left-first normal form is the exact same.
Exercise 9.8. Find the right-first and left-first normal forms of the
element g = [{−3, 1, 2}, 4] ∈ L2 .
S(z) = 1 + 2z + 2z 2 + 2z 3 + · · · .
S(z) = 1 + 2z + 2z 2 + 2z 3 + · · ·
1
= 2(1 + z + z 2 + z 3 + · · · ) − 1 = 2 −1
1−z
2 1−z 1+z
= − = .
1−z 1−z 1−z
If Z is generated by {2, 3} then one would expect a fairly different
growth function and associated series. (The Cayley graphs for these
two generating sets are shown in Figure 1.20.) Let S{2,3} (n) be the
sphere of radius n, with respect to {2, 3}, and let σ{2,3} (n) be the size
of S{2,3} (n). By inspection we find that σ{2,3} (0) = 1 and σ{2,3} (1) = 4;
170 The Geometry of Infinite Groups
a short computation shows S{2,3} (2) = {±1, ±4, ±5, ±6} and therefore
σ{2,3} (2) = 8. After this, the pattern becomes simple. Because 2 and 3
are relatively prime, any integer can be expressed as n = l·2+m·3 where
l and m are integers. But is this a geodesic expression? That is, is |l|+|m|
as small as it can be in order to produce n? Not if |l| ≥ 3, for then the
substitution 3 · 2 = 2 · 3 would reduce the length (where our generators
have been underlined). Thus a minimal-length word representing n ∈ Z
contains at most two instances of 2. So the sphere of radius n contains
at most ten elements: ±n · 3, ±(n − 1) · 3 ± 2 and ±(n − 2) · 3 ± 2 · 2.
However, for n ≥ 3, (n − 1) · 3 − 2 = (n − 3) · 3 + 2 · 2, so (n − 1) · 3 − 2 is
actually in S{2,3} (n − 1). Further, (n − 2) · 3 − 2 · 2 = (n − 4) · 3 + 2, so
(n − 2) · 3 − 2 · 2 ∈ S{2,3} (n − 1). So for n ≥ 3, the sphere of radius n is
and therefore
1 n=0
4 n=1
σ{2,3} (n) =
8 n=2
6 n ≥ 3.
This formula is correct, but it takes some care to actually prove it. One
can also show that the corresponding growth series is the formal power
series associated to the rational function
1 + z + z2
S(z) = .
(1 − z)2
n
σG⊕H (n) = σG (i)σH (n − i),
i=0
Example 9.17. Pick cyclic generators for Z3 and Z4 and the corre-
sponding growth “series” for these finite groups are 1+2z and 1+2z+z 2 .
Applying Theorem 9.16 we see that the growth series for Z3 ∗ Z4 is
(1 + 2z)(1 + 2z + z 2 ) 1 + 4z + 5z 2 + 2z 3
SZ3 ∗Z4 (z) = =
1 − (2z)(2z + z 2 ) 1 − 4z 2 − 2z 3
= 1 + 4z + 9z 2 + 20z 3 + 44z 4 + 98z 5 + 216z 6 + · · · .
Remark 9.18. Not all growth series of finitely generated groups are
rational. For example, the growth series of Z2 F2 , a “lamplighter” group
based on lighting lamps on the Cayley graph of F2 instead of Z, is
not rational. (See [Pa92] for the details on this and related examples.)
The finitely generated, infinite torsion group U, which we constructed in
Chapter 6, is another example. Perhaps a bit more worrisome is the fact
that the growth series can be rational with respect to one generating set
but not rational with respect to another [St96].
Proof. This is what matrix multiplication is all about! (If that doesn’t
make you happy, prove it by induction.)
Lemma 9.21. Let Sij (z) = n≥0 Aij (n)z n . Then Sij (z) is the series
corresponding to the rational function
(−1)i+j det (I − zA : j, i)
Sij (z) =
det (I − zA)
where (I − zA : j, i) denotes the matrix formed by subtracting zA from
the identity matrix, and then removing the jth row and ith column.
Thus the formula stated is simply the standard formula for the inverse
of a matrix.
This may look daunting, but a program such as Mathematica can do the
computation almost instantly, giving the answer:
1+z
= 1 + 4z + 12z 2 + 36z 3 + 108z 4 + · · · .
1 − 3z
Since σ(n) = 4 · 3n−1 for F2 with respect to a basis, the reader should
be able to independently verify that this formula is indeed correct.
−1 1
(I − zA) = M
det [I − zA]
where M is a matrix with polynomial entries. It follows that
−1 p(z)
S · (I − zA) ·A=
det [I − zA]
9.4 Cannon Pairs 175
a + bz + cz 2 + dz 3 + ez 4
= a + (4a + b)z
(z − 1)2 (1 + z)(1 − 3z)
+ (14a + 4b + c)z 2 + (44a + 14b + 4c + d)z 3
+ (135a + 44b + 14c + 4d + e)z 4 + · · · ,
Remark 9.23. It is known that the growth series for BS(1, 2), with
respect to the standard generators a and b, is rational, but where the
rational function has degree 14. That the growth series is rational was
established by Brazil in [Bz94] and independently by Collins, Edjvet
and Gill [CEG94].2 However, it has also been proved that BS(1, 2) does
not admit a regular normal form consisting of geodesic words [Gv96]. In
particular, it is possible to have a rational growth series without it being
computed via the method described in this section.
All three of the papers cited above are accessible to anyone who has
read this far in this book, and are recommended for further study.
Example 9.24. Consider Z ⊕ Z with generators {(1, 0), (0, 1)}. If n and
m are two positive integers,
then the number of geodesics expressing the
element (m, n) is n+mn . The number of geodesics of length n that occur
in the first quadrant is then
n
n n n n n
= + + ··· + + = 2n .
i=0
i 0 1 n − 1 n
Adding the four quadrants, and subtracting 4 for the elements on the
axes that have been counted twice shows γ(n) = 4(2n − 1) for n ≥ 1.
Thus
Γ(z) = 1 + 4z + 12z 2 + 28z 3 + 60z 4 + · · ·
= 1 + [8z + 16z 2 + 32z 3 + · · · ] − [4z + 4z 2 + 4z 3 + · · · ]
2z z 1 + z + 2z 2
=1+4 −4 =
1 − 2z 1−z (1 − z)(1 − 2z)
This example indicates that the geodesic growth series of a group,
much like the ordinary growth series of a group, may be rational. An
argument similar to that given in Section 9.3, using the transfer-matrix
method, establishes:
Theorem 9.25. Let (G, S) be a group paired with a finite generating
set. If the language of all geodesic words is a regular language, then the
geodesic growth series is rational.
In 1984, James Cannon published a foundational paper for the study
of infinite groups from a geometric perspective [Ca84]. In particular,
he imported the idea of cone types, which we have encountered in our
discussion of formal languages, to the study of geodesics in the Cayley
graphs of infinite groups.
Definition 9.26. Fix a group G and a finite generating set S, and let
Γ = ΓG,S be the associated Cayley graph. Then, for a given g ∈ G,
Fig. 9.2. The lower-left vertex is the vertex associated to (0, 0) ∈ Z ⊕ Z while
the highlighted vertex is associated to (2, 1) ∈ Z ⊕ Z. The cone of this element
is the highlighted subgraph, which looks much like a quadrant in the Euclidean
plane.
Example 9.27. Consider Z ⊕ Z with the generators {(1, 0), (0, 1)}. The
cone of the element (2, 1) consists of the subgraph of the Cayley graph
that is induced by {(m, n) | m ≥ 2 and n ≥ 1}. This is illustrated in
Figure 9.2.
To avoid confusing the different uses of this phrase, we refer to the cone
types of L as L-Cone types.
Let L be the language of all geodesic words in {S ∪ S −1 }∗ . If L is
regular, then the Myhill–Nerode Theorem (7.18) states that there are
finitely many L-cone types. Given an element g ∈ G, the subgraph of
the Cayley graph Cone(g) is just the union of paths, based at g, corre-
sponding to words in L-Cone(ω) for all geodesic words ω with π(ω) = g.
As there are finitely many sets of L-Cones, there can only be finitely
many graphs arising as Cone(g) for g ∈ G.
Conversely, assume the Cayley graph ΓG,S has finitely many Cones.
Let M be the automaton whose vertex set corresponds to the finitely
many cone types, where the cone type of the identity is the only start
state, and every vertex is an accept state. For each generator s ∈ S add
a directed edge labelled s from Cone(g) to Cone(gs) when the vertex
associated to gs is in Cone(g). The language accepted by this machine
is L, hence L is regular.
Definition 9.29. If the pair (G, S) has finitely many cone types, or
equivalently if the language of all geodesics is regular, then (G, S) is a
Cannon pair.
and similarly
Thus the subgraph containing [e, α], whose edges are all labelled x and
y, is a copy of the Z ⊕ Z Cayley graph with the roles of x and y reversed.
9.5 Cannon’s Almost Convexity 179
Exercise 13 asks you to verify that this Cayley graph has finitely many
cone types.
We now switch our generating set. Define d = xy and z = x2 and
consider the set of elements S = {x, d, z, α} ∈ G. Since x and d are in S,
y = x−1 d ∈ {S∪ S−1 }∗ . Thus S is a generating set for G. Exercise 14 asks
you to prove that, in this generating set, the word αz n αz m is geodesic
if and only if m < n. Thus the language of geodesics is not regular with
respect to this generating set by an application of the Pumping Lemma
(Theorem 7.14).
Because of examples like this, it is important to remember that having
the language of all geodesics be regular is a property of a pair consisting
of a group and a fixed generating set. It is not a property of the group
independent of the choice of generators.
Example 9.31. Free abelian groups are almost convex as are free groups
and free products of finite groups. Each of these claims is true regard-
less of the set of generators that is chosen, but these claims are easiest
to verify using “standard” generators. Showing this is true for any set
of generators is actually somewhat difficult. There are, in fact, groups
whose Cayley graphs are almost convex with respect to one set of gen-
erators, but not with respect to another.
Note: In the ball of radius n you can always join two vertices by a path
of length ≤ 2n. You simply move from one vertex to the identity, and
then back out to the other vertex. The power of the almost convexity
condition is that there is a single, uniform constant that works for all n.
180 The Geometry of Infinite Groups
Proof. Let K = K(2). We argue by induction, the base case being our
hypothesis that Γ is almost convex. Assume then that v, w ∈ SΓ (n), and
dΓ (v, w) = m, where m ≥ 3. Let M be the maximum distance from the
vertex associated to the identity to a vertex on the chosen edge path
joining v to w. (Thus the chosen path is contained in BΓ (M + 1).) Note
that M must be less than n +
m+1 2 .
Let z be a vertex on this path where the distance from z to the vertex
associated to the identity is M . The neighboring vertices on this path,
call them z1 and z2 , are then either closer to the identity, or at least one of
them is the same distance from the identity as z. If both z1 and z2 are in
SΓ (M − 1) then they may be joined by a path inside of BΓ (M − 1) whose
length is at most K. This removes a vertex in SΓ (M ). In the other case,
z and z1 (or z2 ) can be joined by a path inside of BΓ (M ). This removes
an edge that was outside of BΓ (M ). In a worse-than-is-possible scenario,
this process would have to be applied m times, rerouting the original
path from its original vertices or edges, resulting in a path joining v to
w of length mK. This new path is contained in BΓ (M ).
Repeating this process results in an edge path of length less than
(mK)K = mK 2 , which is contained in BΓ (M − 1). Continuing this
process we eventually convert the original path to one which is contained
in BΓ (n). Since (M + 1) − n ≤ 1 +
m+1 2 , if we set l = 1 +
2 then
m+1
this process stops in at most l steps. The length of the path at the end
can be no more than mK l , hence we may take K(m) = mK l .
that the algorithm exists, but we do not actually provide it! The argu-
ment begins by mentioning that BΓ (K +2) exists, where K is the almost
convexity constant. Let C be any circuit in Γ of length ≤ K + 2, and
let vg be a vertex in this circuit. Then by multiplying by g −1 we may
translate this circuit into BΓ (K + 2). Thus all circuits of length ≤ K + 2
in Γ can be realized as translates of circuits in BΓ (K + 2).
Assuming you have built BΓ (n), you can make a list of “missing edges,”
that is, edges which are attached to vertices in SΓ (n) but which are
not contained in BΓ (n). (See the proof of Theorem 5.10.) We need to
determine how to attach these missing edges in order to form BΓ (n + 1).
If a missing edge joins two vertices in SΓ (n) then those vertices must be
joined by a path of length ≤ K in BΓ (n). Similarly, if two missing edges
meet at a vertex in SΓ (n + 1) then their associated vertices in SΓ (n)
are joined by a path of length ≤ K in BΓ (n). In either event, there is
then a circuit formed by the missing edge, or pair of edges, and an edge
path of length ≤ K. If there is such a circuit, a copy of it can be found
in BΓ (K + 2). If no such circuit appears in BΓ (K + 2) then the missing
edges go to distinct vertices in SΓ (n + 1).
Since there is an algorithm which constructs the Cayley graph Γ, then
by Theorem 5.10 there is an algorithm that solves the word problem.
As we mention below, not all groups are almost convex. There are
a number of weaker versions of almost convexity that have attracted
interest. The weakest (useful) variation is minimal almost convexity. A
Cayley graph Γ is minimally almost convex if, given v, w ∈ SΓ (n), where
dΓ (v, w) ≤ 2, then dB(n) (v, w) ≤ 2n − 1. (Remember, dB(n) (v, w) is
always less than or equal to 2n.) A finitely generated group is then
minimally almost convex if it has a minimally almost convex Cayley
graph.
Example 9.34. It was shown in 1998 that BS(1, 2) is not almost convex
with respect to any finite generating set [MiS98]. (Exercise 16 gives a
hint as to how to prove that the Cayley graph of BS(1, 2), with respect
to the generators given in Chapter 4, is not almost convex.) However, in
2005 it was shown that BS(1, 2) is minimally almost convex [EH05].
There are groups that are not minimally almost convex. For example,
Cleary and Taback proved that the lamplighter group L2 is not mini-
mally almost convex [CT05]. Their approach is to consider the elements
[{−n, n}, 1] and [{−n, n}, −1]. By Proposition 9.7, the length of both of
these elements is 4n + 1. They are both adjacent to [{−n, n}, 0], hence
182 The Geometry of Infinite Groups
Proposition 9.35. The Cayley graph of the lamplighter group with re-
spect to the generators t = [∅, 1] and a = [{0}, 0] is not minimally almost
convex.
Exercises
(1) Let x generate a cyclic group of order 5. Show that the length of x2
and the length of x3 (with respect to {x, x−1 } are both equal to 2.
(2) Characterize the set of elements in L2 for whom the right-first
normal form is geodesic.
(3) Let G be a group with generating set S = S ∪S −1 . The correspond-
ing notion of length induces a function L : G → N given by L(g) =
|g|. These length functions are surprisingly easy to axiomatize. Let
: G → N be a function. Prove that is the same as L if and only if
a. (e) = 0;
b. |(gs) − (g)| ≤ 1 for all g ∈ G and s ∈ S; and
c. if g ∈ G \ {e}, then there is at least one s ∈ S such that
(gs) < (g).
Exercises 183
(4) Find a formula for the size of S(n) and B(n) for Z ⊕ Z with respect
to the standard generating set (1, 0) and (0, 1). Then find a formula
for the size of S(n) and B(n) with respect to the larger generating
set {(1, 0), (0, 1), (1, 1)}.
(5) Let W333 be the group generated by three reflections in the sides of
an equilateral triangle, discussed in Chapter 2. Prove that σ(n) =
|S(n)| is given by
1 n=0
σ(n) =
3 · n n > 0.
(6) Prove that the growth series for W333 is the formal power series
associated to the rational function
1 + z + z2
S(z) = .
(1 − z)2
(7) Theorem 9.16 gives a formula for the growth series of a free product
of two groups. Find a similar formula for G ≈ A ∗ B ∗ C.
(8) The growth series for finitely generated free groups, with respect
to a basis, are always rational. In Section 9.2 it is shown that the
growth series for F1 is the series associated to (1+z)/(1−z); in Sec-
tion 9.3 we saw that the growth series for F2 is the series associated
to (1 + z)/(1 − 3z). What is the rational function corresponding to
the growth series for Fn ?
(9) Let {a, b} be the generating set for BS(1, 2) described in Chapter 4.
Show that the length of bn (n ≥ 1) is obtained by writing n in bi-
nary: If n = bk bk−1 · · · b2 b1 b0 (in binary), with bk = 0, then |bn | is
k
2k + i=0 bi . For example, 14 = 8 + 4 + 2 so, written in binary, 14
is 1110. The claim is that |b14 | is 2 · 3 + 1 + 1 + 1 = 9.
(10) If σ(n) = |S(e, n)|, then the associated series S(z) = n≥0 σ(n)z n
is sometimes referred to as the spherical growth series. One can also
consider the function β(n) = |B(e, n)| which gives the number of
vertices inside the ball of radius n, and the associated ball growth
series B(z) = n≥0 β(n)z n . Show that the spherical growth series
is rational if and only if the ball growth series is rational.
(11) Given any regular language L over an alphabet A, one can form
the associated growth series z |w| . The proof of Theorem 9.19
w∈L
establishes that, if L is regular, this growth series is rational. Show
that the converse is false using the language L of all palindromes
over the alphabet A = {a, b}.
184 The Geometry of Infinite Groups
(12) Let G = Z ⊕ Z and let S = {(1, 0), (0, 1)} be the standard set of
generators. Show that (G, S) has exactly nine cone types.
(13) Let G = (Z ⊕ Z) Z2 and let S be the set of generators
S = {[(1, 0), e], [(0, 1), e], [(0, 0), α]}
(as discussed in Section 9.4). Prove that (G, S) has finitely many
cone types, that is, that (G, S) is a Cannon pair.
(14) Let G be as in Exercise 13 and define d = xy and z = x2 . Show
that in this generating set the word αz n αz m is geodesic if and only
if m < n.
(15) Let G and H be almost convex groups. Show that G⊕H and G∗H
are also almost convex groups.
Fig. 9.3. Two vertices in the sphere of radius 10 in the Cayley graph of BS(1, 2)
are highlighted. They are a distance 2 apart in the Cayley graph but are much
further apart inside the ball of radius 10.
(16) This exercise indicates how to prove that the Cayley graph of
BS(1, 2), with respect to the generators {a, b}, is not almost con-
vex. The reader may want to refer to the view of the Cayley graph
of BS(1, 2) shown in Figure 9.3 throughout this problem.
Exercises 185
a. Show that ab4 a−1 and bab4 are both in the sphere of radius
6, and that the distance between them in the Cayley graph
is 2.
b. What is the distance between ab4 a−1 and bab4 when re-
stricted to paths in the ball of radius 6?
c. Show that a2 b4 a−2 and ba2 b4 a−1 are both in the sphere of
radius 8, and that the distance between them in the Cayley
graph is 2.
d. What is the distance between a2 b4 a−2 and ba2 b4 a−1 when
restricted to paths in the ball of radius 8?
e. Show that a3 b4 a−3 and ba3 b4 a−2 are both in the sphere
of radius 10, and that the distance between them in the
Cayley graph is 2. (These are the vertices highlighted in
Figure 9.3.)
f. What is the distance between a3 b4 a−3 and ba3 b4 a−2 when
restricted to paths in the ball of radius 10?
g. Attempt to prove that the Cayley graph of BS(1, 2), with
respect to {a, b}, is not almost convex. If you get stuck, try
reading [MiS98] or [EH05].
(17) Given a group G and a finite generating set S, a dead end is an
element g such that |gs| ≤ |g| for all s ∈ {S ∪ S −1 }. Finite groups
always have dead end elements, but it is somewhat surprising that
infinite groups can have dead end elements as well.
Dead end elements are also sometimes referred to as pockets.
(If the Cayley graph Γ is thought of as a net, dangling down
from the vertex associated to the identity, then a dead end ele-
ment sits at the bottom of a “pocket.”) The depth of a dead end
element g is
(18) Let dn ∈ L2 be the element whose picture has lit lamps at each
integer from −n to n, with the lamplighter positioned at the origin.
Show that dn is a dead end element of depth n + 1, and conclude
that the Cayley graph of L2 with respect to the generators {a, t}
has deep pockets.
10
Thompson’s Group
187
188 Thompson’s Group
[0,1]
[0,1/2] [1/2,1]
the root of T is identified with [0, 1]. The standard dyadic intervals that
form the leaves of such a finite, rooted binary tree give a dyadic division
of [0, 1], and any dyadic subdivision generates a finite, rooted binary
tree. As we will frequently refer to “finite, rooted binary trees” in the
remainder of this chapter, we abbreviate our terminology to “frb-trees.”
Given an ordered pair of dyadic divisions of [0, 1], with the same
number of pieces, there is a corresponding piecewise linear function
f : [0, 1] → [0, 1]. If 0 < m1 < · · · < mk < 1 denote the first set of
chosen middles and 0 < µ1 < · · · < µk < 1 denote the second set of
chosen middles, then f is defined by requiring:
Thompson functions are continuous bijections of [0, 1], hence they can
be viewed as elements of the group Homeo([0, 1]), whose binary operation
is function composition. Thus, to show that this subset is a subgroup,
it suffices to show that it is closed under inverses and composition. To
do this, we introduce a bit of notation. Since dyadic divisions of [0, 1]
correspond with frb-trees, we may use an ordered pair of frb-trees to
describe elements of the group F . In particular, we will use [T2 ← T1 ]
to denote the Thompson function where the domain has been divided
according to T1 and the range according to T2 . The reader should verify
the following lemma, which exploits this notation.
Fig. 10.3. A finite, rooted binary tree T is shown above, in darker and thicker
lines. The lighter edges show the result of forming (T ∧ 2) ∧ 2.
of T , you can form unions T1 ∪T2 , and the result is another frb-tree. This
process can be thought of in terms of overhead transparencies: take two
transparencies, one with T1 on it and the other containing T2 ; lay one
transparency over the other, lining up the roots, and you have produced
T1 ∪ T2 . As each Ti (i = 1 or 2) is a subtree of the union, each Ti can be
expanded to T1 ∪ T2 by adding carets.
Given two pairs of frb-trees, [T2 ← T1 ] and [T4 ← T3 ], there is a
set of equivalent pairs of frb-trees, [T2 ← T1 ] and [T4 ← T3 ], where
T2 = T3 = T2 ∪ T3 . Thus, by Lemma 10.1, the composition of the as-
sociated functions is the function associated to [T4 ← T1 ]. Thus the set
of Thompson functions is closed under composition; as we already knew
they were closed under inversion, the set of Thompson functions forms
a group.
f = [S ← T ] = [S ← Tn ][Tn ← T ]
1. x1 xi0 x1 ,
2. x−1
1 x0 x1 ,
i
i+1 −1
3. x1 x0 x1 ,
4. x−1 i+1 −1
1 x0 x1 .
where i is any positive integer. Checking that this set of words is a regular
language is routine; proving that it is a normal form is not. It should
be noted that this normal form is not composed of geodesic words. This
was not due to a lack of insight on the part of the authors. Rather, it
was later shown by Cleary, Elder and Taback that it is impossible to
construct such a normal form.
This shows that one cannot hope to show that the growth series for
F is rational using the approach of Section 9.3. As of the time of this
writing, it is unknown if the growth series for Thompson’s group F is
rational or not.
Finally, the geometry of the Cayley graph of F , with respect to {x0 , x1 },
is quite complicated. In particular, Belk and Bux have shown:
Exercises
(1) Show that if you divide [0, 1] into dyadic intervals, then this sub-
division must be a dyadic subdivision.
(2) Prove that a continuous, bijective, piecewise linear function f :
[0, 1] → [0, 1] (with finitely many corners) is a Thompson function
if and only if:
a. all of the slopes (away from the corners) are powers of 2;
and
b. all of the corners have both coordinates in the dyadic
rationals.
(3) Prove Lemma 10.2.
(4) F is torsion-free.
Prove that
m m+1
(5) Let I = n , be a standard dyadic interval. Let FI be the
2 2n
subset of functions
FI = {f ∈ F | Supp(f ) ⊂ I}.
a. Prove that FI is a subgroup.
b. Prove that FI ≈ F .
196 Thompson’s Group
(6) Show that the set of positive elements of F is closed under com-
position.
(7) In this problem we discuss the standard normal form for elements
of Thompson’s group F using {x0 , x1 , x2 , . . .}.
2 0
0 1 0 0 0
0
3 0
Fig. 10.5. The exponents of the leaves in this frb-tree, T , are as indicated.
This implies the element [T ← T8 ] can be expressed as x20 x32 x6 .
What one really cares about are the inherent properties of the group, not the
artefacts of a particular presentation.
. –Martin Bridson
198
11.1 Changing Generators 199
Fig. 11.1. The Cayley graph of D3 with respect to two different generating
sets. The Cayley graph on the left corresponds to using a reflection and a
rotation; the Cayley graph on the right corresponds to using two reflections.
Proposition 11.2. Let S and T be two finite generating sets for a group
G and let ΓS and ΓT be the corresponding Cayley graphs. Then there are
maps φT ←S : ΓS → ΓT and φS←T : ΓT → ΓS such that:
Fig. 11.2. A map from a finite tree to the Cayley graph of Z with respect to
the generating set {2, 3}.
Proof. The map φT ←S is the identity map on the vertices. That is, if
vg denotes the vertex corresponding to g ∈ G in ΓS , and vg denotes
the vertex corresponding to g in ΓT , then φT ←S (vg ) = vg . The map
φS←T is defined in the same way on vertices, and the first claim is then
immediate.
For each generator s ∈ S choose a word ωs = t1 t2 · · · tk ∈ {T ∪ T −1 }∗
such that s = π(ws ) ∈ G. (One can always do this since T is a generating
set for G.) By the construction of Cayley graphs, if e is an edge of ΓS
then e is labelled by a generator s ∈ S and joints the vertex associated
to g to the vertex associated to g · s. The map φT ←S sends every such
edge to the edge path
g −→ g · t1 −→ gt1 · t2 −→ · · · −→ gt1 t2 · · · tk
in ΓT .
The map φS←T is defined similarly. For each generator t ∈ T choose
a word ωt ∈ {S ∪ S −1 }∗ representing t and use these words to describe
the images of the edges of ΓT .
Let k be the maximal length of the words ωt and ωs . It follows that
φT ←S (e) is an edge path of length ≤ k. The image φS←T of this path is
then an edge path of length ≤ k 2 . Hence the constant K in the second
claim of the theorem can be taken to be k 2 .
Proof. Let
Λ1 = Max{|ωt | | t ∈ T }
Fig. 11.3. The Cayley graph of D∞ with respect to the generating set {a, τ }.
back from their graphs, in both instances it would seem as if you are
looking at the real line.
Lemma 11.6. Let S and T be two finite generating sets for a fixed group
G. Then there is a constant λ ≥ 1 such that
1
βS n ≤ βT (n) ≤ βS (λn)
λ
for all n ∈ N.
Proof. Let |g|S denote the length of g ∈ G with respect to the generating
set S, and define |g|T similarly. By Corollary 11.3, there is a λ ≥ 1
such that |g|T ≤ λ · |g|S for all g ∈ G. Thus if g ∈ BT (n) then g ∈
BS (λn), which implies βT (n) ≤ βS (λn). Exchanging the roles of S and
T establishes the other inequality.
Lemma 11.6 is one motivation for the following definition; another
motivation arises in Section 11.7.
Definition 11.7. Define to be the relation on growth functions de-
fined by f g if there is a constant λ ≥ 1 such that
f (x) ≤ λg(λx + λ) + λ
for all x ∈ [0, ∞). If f g then we say g dominates f . (Notice that the
right-hand side is the result of pre- and post-composing g(x) with the
linear function y = λx + λ.)
If f g and g f then g strictly dominates f . We denote this by
f ≺ g. On the other hand, if f g and g f then f and g are said to
be equivalent, denoted f ∼ g. Two equivalent growth functions are said
to grow at the same rate.
Exercise 2 at the end of this chapter asks you to prove that “” is
a reflexive and transitive relation, and that f ∼ g is an equivalence
relation.
Example 11.8. Consider the following growth functions: f (n) = n2 ,
g(n) = 15n2 + 10n + 1 and h(n) = 2n . We have f (n) < g(n) for all
n ∈ N, so f g. On the other hand,
f (5n) + 5 = 25n2 + 5 = 15n2 + 10n2 + 5 > 15n2 + 10n + 1 = g(n)
so g f , hence f ∼ g.
One expects that f h, since exponential functions grow faster than
polynomial functions. There is a slight hitch in that f (3) = 9 > 8 = h(3)
but, for all other n ∈ N, f (n) ≤ h(n). Thus f (n) < h(2n) + 2 for all n
and so it is indeed the case that f h. On the other hand,
λ(λn + λ)2 + λ
lim =0
n→∞ 2n
204 The Large-Scale Geometry of Groups
Corollary 11.9. If S and T are two finite generating sets for a group
G then the associated growth functions are equivalent: βS (n) ∼ βT (n).
Not all growth functions of groups are polynomial. Consider the free
group of rank k, Fk , along with a fixed basis. The number of elements
in the sphere of radius n, S(n), is 2k · (2k − 1)n−1 (for n ≥ 1). Thus
(2k − 1)n < |S(n)| < |B(n)|. Hence the growth function of Fk satisfies
(2k − 1)n β(n). In particular, as long as k ≥ 2, the growth is at least
11.3 The Growth of Thompson’s Group 205
Lemma 11.12. Let a and b be two integers, both greater than 1, and
let α(n) = an and β(n) = bn be the corresponding exponential functions.
Then α(n) ∼ β(n).
Proof. We will assume that a < b so that α(n) β(n) trivially. Con-
versely, if λ ≥ loga (b), then β(n) ≤ α(λn) + λ, so β(n) α(n).
Thus the resulting pair would still satisfy π(ω0 ) = π(ω1 ), but their
combined length will have been reduced. Since |ω0 | + |ω1 | was minimal,
this is not possible, and so the two words end in distinct generators.
Without loss of generality we assume ω0 ends in x0 and ω1 ends in x1 .
If f is an element of F , then when restricted to some small interval
[0, ), f is linear with slope 2k . The function φ : F → Z that takes f
to this exponent k is a homomorphism. (If you compose a function with
slope 2k with a function of slope 2l then the result has slope 2k+l .) In
particular, φ(x0 ) = −1 and φ(x1 ) = 0. Since ω0 and ω1 represent the
same element of F , φ(ω0 ) = φ(ω1 ). Further, |φ(ω0 )| = |φ(ω1 )| is the
number of x0 ’s in these words. Call this number n. Note that n > 0
because otherwise ω0 and ω1 would simply be powers of x1 .
Thinking of x0 as a function from [0, 1] to [0, 1] we see x0 (3/4) = 1/2
and, more generally, xk0 (3/4) = 1/2k for any positive integer k. Since x1
is the identity when restricted to [0, 1/2], and ω0 ends in x0 , it follows
that π(ω0 ) takes 3/4 to 1/2n (where n is as defined above).
11.3 The Growth of Thompson’s Group 207
Now consider the action of π(ω1 ) on [0, 1]. There is some positive k
such that ω1 ends with x0 xk1 . The function xk1 takes 3/4 to some number
strictly smaller than 3/4, hence x0 xk1 takes 3/4 to some number strictly
smaller than 1/2. Any additional x1 ’s that appear in ω1 have no effect on
[0, 1/2], but there are n−1 more x0 ’s in ω1 . Thus, in the end, π(ω1 ) takes
3/4 to some number which is strictly less than 1/2n . So it is impossible
for ω0 and ω1 to represent the same element of Thompson’s group F .
Because there are 2n words of length n in {x0 , x1 }∗ , it follows that
2n ≤ β(n), and so Thompson’s group F has exponential growth.
for all d.
This limit exists, although it may equal ∞, since, by Lemma 11.18, the
sequence {||Γ \ B(n)||} is a non-decreasing sequence of integers.
be the largest value that occurs for all finite subgraphs C, or set ec (Γ) =
∞ if there is no n ∈ N such that n ≥ ||Γ \ C|| for all finite subgraphs
C ⊂ Γ. (If you have studied real analysis, feel free to replace “Max” with
“sup.”)
210 The Large-Scale Geometry of Groups
Lemma 11.21. Let Γ be a connected, locally finite graph and let e(Γ)
and ec (Γ) be as defined above. Then e(Γ) = ec (Γ).
Proof. Since metric balls B(n) are finite subgraphs in any locally finite
graph Γ, e(Γ) ≤ ec (Γ). Conversely, given any finite subgraph C, there is
an n ∈ N such that C ⊂ B(n). Thus ||Γ \ C|| ≤ ||Γ \ B(n)||, by the same
argument used to prove Lemma 11.18, and therefore ec (Γ) ≤ e(Γ).
Lemma 11.22. Let S and T be two finite generating sets for G, and
let BS (n) and BT (n) be the balls of radius n in ΓS and ΓT , respectively.
Then there is a constant µ ≥ 1 such that if vg and vh are vertices in ΓS
that can be joined by an edge path outside of BS (µn + µ) then vg and vh
in ΓT are outside BT (n) and can be joined by a path that stays outside
of BT (n).
Theorem 11.23. Let S and T be two finite generating sets for a group
G, and let ΓS and ΓT be the corresponding Cayley graphs. Then
e(ΓS ) = e(ΓT ).
11.5 The Freudenthal–Hopf Theorem 211
that is, e(ΓS ) ≥ e(ΓT ). The opposite inequality holds mutatis mutandis,
hence we are done.
Exercise 11.25. Show that a finitely generated group G has zero ends
if and only if G is finite.
Proof. Previous examples show that there are groups with zero, one, two
or infinitely many ends. Thus we need to establish that there is no group
G where e(G) is finite but e(G) > 2. Assume to the contrary that the
Cayley graph Γ of a finitely generated group G has k ≥ 3 ends. Note that
212 The Large-Scale Geometry of Groups
this implies that G must be an infinite group, and that there is a number
n ∈ N such that Γ \ B(n) has k unbounded, connected components.
Since G is infinite there is an element g ∈ G such that d(e, g) > 2n,
where vg is in an unbounded, connected component of Γ \ B(n). Then
g · B(n) ∩ B(n) = ∅. Further, since gB(n) is contained in an unbounded,
connected component of Γ \ B(n), it divides this component into at least
k connected pieces, at least (k − 1) of which are unbounded. Thus if
C = B(n) ∪ g · B(n) then C is a finite subgraph of Γ and ||Γ \ C|| ≥ 2k − 2
(see Figure 11.5). It follows that ec (Γ) – as defined in Lemma 11.21 – is
at least 2k − 2. Thus e(Γ) ≥ 2k − 2 > k, since k ≥ 3. But this contradicts
the claim that e(G) = k.
(n) g (n)
Fig. 11.5. If a group G has at least three ends, then it has at least four ends,
which means it has at least six ends, ...
Exercise 11.28. Copy and complete Table 11.1, filling in the values
of e(G). The formatting of this table is intended to help you guess the
correct value of e(G), even if you are not clear on how to actually prove
that your guess is correct.
G e(G) G e(G)
Sym5 Z⊕Z
Z7 Z3 W244
(D397 Z736 )8
G e(G) G e(G)
Z F2
(D397 Z736 )8 (D397 Z736 )8
D∞ Z3 ∗ Z4
Z ⊕ Sym7 (Z ⊕ Z) ∗ Z2
Table 11.1. Exercise 11.28 asks you to fill in the values e(G) in this
table.
mind are Z, D∞ , and any direct sum of a two-ended group with a finite
group. In each case the group has a finite index subgroup ≈ Z, and the
goal of this section is to establish that this is indeed always the case.
Recall that the symmetric difference of two sets is defined to be
AB = (A ∪ B) \ (A ∩ B) .
(The symmetric difference was previously mentioned in our discussion of
the lamplighter group in Chapter 8.) Since our discussion of two-ended
groups will rely heavily on this operation, you should verify the following
two formulas.
214 The Large-Scale Geometry of Groups
Since we have assume EgE and EhE are both finite, so is EghE.
In order to establish that the index of H in G is at most 2, assume
that H = G and that g, h ∈ G \ H. Then
Egh−1 E = (EgE) gEgh−1 E
= (EgE) g Eh−1 E
c
= (EgE) g Eh−1 E .
c
c
By Lemma 11.30, both (EgE) and Eh−1 E are finite, hence so
c
is Egh−1 E.
Let C be the finite subgraph that divided the Cayley graph Γ into
two unbounded, connected components, and let g ∈ H be an element
where gC ∩ C = ∅. Then either gC ⊂ E or gC ⊂ E c . From this and the
definition of H, one gets:
Lemma 11.32. Let G, H, C and E be as above, and let g ∈ H be an
element such that C ∩ gC = ∅. Then either
1. E ∩ gE c = ∅ and E c ∩ gE = ∅, or
2. E ∩ gE c = ∅ and E c ∩ gE = ∅.
We are now well-positioned to establish the main result of this section.
Theorem 11.33 below was first established by C. T. C. Wall in 1967,
using purely algebraic techniques. Our approach has been to build up
an intuitive notion of the “right side” of the Cayley graph Γ, which is
the part that contains E. On an intuitive level, an element g ∈ H moves
Γ “to the right” if gE ⊂ E. Further, g moves Γ “far to the right” if
E ∩ gE c is large. Surprisingly, this vague intuition can be made quite
precise.
Theorem 11.33. Let G and H be as above. Then there is a homomor-
phism φ : H → Z whose kernel is finite.
Proof. Since g ∈ H, EgE is finite, hence E ∩ gE c and E c ∩ gE are
also finite. We may then define a function φ : H → Z by
φ(g) = |E ∩ gE c | − |E c ∩ gE|.
Notice that E ∩ gE c is the disjoint union
E ∩ gE c = (E ∩ gE c ∩ ghE c ) ∪ (E ∩ gE c ∩ ghE)
and similarly
E c ∩ gE = (E c ∩ gE ∩ ghE c ) ∪ (E c ∩ gE ∩ ghE) .
216 The Large-Scale Geometry of Groups
Given two integral points we can measure the distance between them
in R2 using the standard metric:
dR2 = (m1 − m0 )2 + (n1 − n0 )2 .
We can also measure the distance between them in the Cayley graph of
Z ⊕ Z or, in other words, in the word metric associated to the generators
{(1, 0), (0, 1)}:
dZ2 = |m1 − m0 | + |n1 − n0 |.
How different are these two notions of distance? The distance between
two points in the plane is realized by a straight line segment and, given
any such line, there is an edge path in Γ ⊂ R2 that stays close to it (as in
Figure 11.6). Because straight lines in R2 can be roughly approximated
Fig. 11.6. Straight lines in the Euclidean plane can be roughly approximated
by geodesic edge paths in an embedded Cayley graph for Z ⊕ Z.
(See Definition 11.7, where similar inequalities were used to define equiv-
alent growth functions.)
The function φ is quasi-dense, which is also sometimes referred to as
quasi-onto, if there is a λ such that, for any point y ∈ Y , there is some
x ∈ X such that dY (φ(x), y) ≤ λ.
Finally, φ : X → Y is a quasi-isometry if it is both a quasi-isometric
embedding and is quasi-dense.
If φ is a quasi-isometry and λ is large enough to satisfy all the re-
quirements above, then λ is referred to as the quasi-isometry constant
of φ.
There are a number of elementary examples of quasi-isometric spaces.
Our discussion of the polka-dot embedding of Z2 in R2 shows that the
group Z2 with its standard word metric is quasi-isometric to the Eu-
clidean plane. The cyclic subgroup of Z2 generated by (1, 0) is quasi-
isometrically embedded in R2 , but it is not quasi-isometric to R2 , as it
is not quasi-dense. In addition to the specific case of Z2 R2 , we have
also encountered a large collection of quasi-isometric metric spaces. If
G is a finitely generated group, and S and T are two finite generating
sets, then the word metric on G induced by S and the word metric on
G induced by T yield quasi-isometric metric spaces; this follows imme-
diately from Corollary 11.3. As this is an important point, we record
it as:
Lemma 11.37. Let G be a finitely generated group and let S and T be
two finite generating sets. Then G with the word metric induced by S is
quasi-isometric to G with the word metric induced by T .
Let φ : X → Y be a quasi-isometry with associated constant λ. Define
a function ψ : Y → X by ψ(y) = x, where x ∈ X is chosen so that
dY (φ(x), y) ≤ λ. (Such an x always exists because φ is quasi-dense.) It
follows from the definition of ψ that
1. dY (φ ◦ ψ(y), y) ≤ λ for all y ∈ Y .
Because φ is a quasi-isometry, with quasi-isometry constant λ, we know
that
1
dX (x, x ) − λ ≤ dY (φ(x), φ(x )) ≤ λdX (x, x ) + λ.
λ
In particular, if x = ψ ◦ φ(x), then φ(x) = φ(x ), which implies that
dY (φ(x), φ(x )) = 0. This establishes two additional facts:
2. dX (x, ψ ◦ φ(x)) ≤ λ for all x ∈ X.
11.7 Commensurable Groups and Quasi-Isometry 221
dY (φ ◦ ψ(y), φ ◦ ψ(y ))
≤ dY (φ ◦ ψ(y), y) + dY (y, y ) + dY (y , φ ◦ ψ(y ))
≤ dY (y, y ) + 2λ.
just needs to find “reasonable” actions of two groups on the same “rea-
sonable” metric space, where the sizes of the associated fundamental
domains are not rational multiples of each other. (An account of the
Švark–Milnor Theorem and its consequences – including explanations of
“reasonable” – can be found in [BrHa99].)
We have previously established that the growth functions for a given
group G, with respect to two finite generating sets, are equivalent (Corol-
lary 11.9) and that the number of ends of a group is not dependent on
the choice of finite generating set (Theorem 11.23). These results also
hold for quasi-isometric groups. For example, assume φ : H → G is a
quasi-isometry with associated constant λ. If h ∈ H is in the ball of ra-
dius n about the identity, then φ(h) must be in the ball of radius λn + λ
about the identity (in G). Thus βH (n) ≤ βG (λn + λ), so βH βG . The
quasi-inverse ψ : G → H shows that βG βH , and therefore βG ∼ βH .
We leave the proof that the number of ends is a quasi-isometry invariant
as an exercise (Exercise 16 at the end of this chapter).
Theorem 11.46. Let G and H be finitely generated groups that are
quasi-isometric. Then
1. if βG and βH denote growth functions of G and H, then βG ∼ βH ;
2. e(G) = e(H).
Thus we know, for example, that Zn is not quasi-isometric to Zm (for
m = n) because they have inequivalent growth rates. Similarly, F2 and
the lamplighter group are not quasi-isometric because e(F2 ) = ∞ while
e(L2 ) = 1 (Exercise 9).
Exercises
(1) The dihedral group D3 can be generated by two reflections or a
reflection and a rotation, and the associated Cayley graphs are
shown in Figure 11.1. Construct a map from the Cayley graph on
the right to the Cayley graph on the left in this figure, along the
lines discussed in the proof of Proposition 11.2.
(2) Let be the relation on growth functions defined by f g if there
is a constant λ ≥ 1 such that
f (x) ≤ λg(λx + λ) + λ
for all x ∈ [0, ∞). Show that is symmetric and transitive, and
that the relation f ∼ g – defined to mean both f g and g f is
an equivalence relation
226 The Large-Scale Geometry of Groups
227
228 Bibliography
[CET06] S. Cleary, M. Elder, and J. Taback, Cone types and geodesic languages
for lamplighter groups and Thompson’s group F , J. Algebra 303 (2006)
476–500.
[CEG94] D. J. Collins, M. Edjvet, and C. P. Gill, Growth series for the group
x, y|x−1 yx = y l , Arch. Math. (Basel) 62 (1994) 1–11.
[CL90] J. H. Conway and J. C. Lagarias, Tiling with polyminoes and combina-
torial group theory, J. Comb. Theory, A 53 (1990) 183–208.
[CV96] M. Culler and K. Vogtmann, A group-theoretic criterion for property
FA. Proc. Am. Math. Soc. 124 (1996) 677–83.
[Da08] M. W. Davis, The Geometry and Topology of Coxeter Groups, London
Mathematical Society Monographs, Princeton University Press, 2008.
[De12] M. Dehn, Über unendliche diskontinuierliche gruppen, Math. Ann. 71
(1912) 116–44.
[Di94] W. Dicks, Equivalence of the strengthened Hanna Neumann conjecture
and the amalgamated graph conjecture, Invent. Math. 117 (1994) 373–89.
[EH05] M. Elder and S. Hermiller, Minimal almost convexity, J. Group Theory
8 (2005) 239–66.
[E+92] D. Epstein, J. W. Cannon, D. Holt, S. V. F. Levy, M. S. Paterson, and
W. P. Thurston, Word Processing in Groups, Jones and Bartlett, Boston,
1992.
[Ge08] R. Geoghegan, Topological Methods in Group Theory, Springer, 2008.
[Gi87] R. Gilman, Groups with a rational cross-section, in S. M. Gersten and
J. R. Stallings (eds), Combinatorial Group Theory and Topology, Prince-
ton University Press, 1987, pp. 175–83.
[Gi05] R. Gilman, Formal languages and their application to combinatorial
group theory, in Groups, Languages, Algorithms, Cont. Math. 378, Amer-
ican Mathematical Society, 2005, pp. 1–36.
[Gl92] A. M. W. Glass, The ubiquity of free groups, Math. Intelligencer 14
(1992) 54–7.
[Gr80] R. I. Grigorchuk, On Burside’s problem on periodic groups, Funkt-
sional. Anal. i Prilozhen 14 (1980) 53–4.
[GrPa07] R. I. Grigorchuk and I. Pak, Groups of intermediate growth: an in-
troduction for beginners, to appear, L’Enseignement Mathématique.
[Gr81] M. Gromov, Groups of polynomial growth and expanding maps, Publ.
math. de l’ I.H.E.S., 53 (1981) 53–78.
[Gr87] M. Gromov, Hyperbolic groups, in S. M. Gersten (ed.), Essays in Group
Theory, Math. Sci. Res. Inst. Publ. 8, Springer, 1987, pp. 75–263.
[GrMa64] I. Grossman and W. Magnus, Groups and their Graphs, Mathemat-
ical Association of America, 1964.
[Gv96] J. R. J. Groves, Mimimal length normal forms for some soluble groups,
J. Pure Appl. Algebra 114 (1996) 51–8.
[Gu04] V. S. Guba, On the properties of the Cayley graph of Richard Thomp-
son’s group F , Int. J. Algebra Comput. 14 (2004) 677–702.
[GuSa97] V. S. Guba and M. V. Sapir, The Dehn function and a regular set of
normal forms for R. Thompson’s group F , J. Austral. Math. Soc. Ser. A
62 (1997) 315–28.
[GuSi83] N. Gupta and S. Sidki, On the Burside problem for periodic groups,
Math. Z. 182 (1983) 385–8.
[Hi51] G. Higman, A finitely related group with an isomorphic proper factor
group, J. Lond. Math. Soc. 26 (1951) 59–61.
Bibliography 229
[Ho81] D. F. Holt, A graph which is edge transitive but not arc transitive,
J. Graph Theory 5 (1981) 201–4.
[HoMe04] P. Hotchkiss and J. Meier, The growth of trees, The College Mathe-
matics Journal 35 (2004) 143–51.
[Hu90] J. E. Humphreys, Reflection Groups and Coxeter Groups, Cambridge
Studies in Advanced Mathematics 29, Cambridge University Press, 1990.
[Jo97] D. L. Johnson, Presentations of Groups, 2nd edition, Cambridge
University Press, 1997.
[Ku33] A. Kurosch, Über freie Produkte von Gruppen, Math. Ann. 108 (1933)
26–36.
[MKS66] W. Magnus, A. Karrass, and D. Solitar, Combinatorial Group Theory,
John Wiley, 1966. (Reprinted by Dover)
[Me72] S. Meskin, Non-residually finite one-relator groups, Trans. Am. Math.
Soc. 64 (1972) 105–14.
[Mi92] C. F. Miller, Decision problems for groups – survey and reflections, in
G. Baumslag and C. F. Miller (eds), Algorithms and Classification in
Combinatorial Group Theory, Math. Sci. Res. Inst. Publ. 23, Springer,
1992 pp. 1–59.
[MiS98] C. F. Miller and M. Shapiro, Solvable Baumslag-Solitar groups are not
almost convex, Geom. Dedicata 72 (1998) 123–7.
[MuS83] D. Muller and P. Schupp, Groups, the theory of ends and context-free
languages, J. Computer and System Sciences 26 (1983) 295–310.
[Ne90] W. D. Neumann, On intersections of finitely generated subgroups of free
groups, in L. G. Kovacs (ed.), Groups: Canberra 1989, Springer, 1990,
pp. 161–70.
[Ni24] J. Nielsen, Die Isomorphismengruppe der freien Gruppen, Math. Ann.
91 169–209.
[Ol95] A.Yu. Ol’shanskii, A simplification of Golod’s example, in A. C. Kim
and D. L. Johnson (eds), Groups: Korea ’94 (Pusan), de Gruyter, 1995,
pp. 263–5.
[Pa92] W. Parry, Growth series of some wreath products, Trans. Am. Math.
Soc. 331 (1992) 751–9.
[Re98] S. Rees, Hairdressing in groups: a survey of combings and formal lan-
guages, in The Epstein Birthday Shrift, Geometry and Topology Mono-
graphs 1, 1998, pp. 493–509.
[Se77] J-P. Serre, Arbres, Amalgams, SL2 , Astérisque No. 46, Société
Mathématique de France, 1977.
[Se80] J-P. Serre Trees (trans. John Stilwell), Springer-Verlag, 1980.
[St68] J. Stallings, On torsion-free groups with infinitely many ends, Ann.
Math. 88 (1968) 312–34.
[St96] M. Stoll, Rational and transcendental growth series for the higher
Heisenberg groups, Invent. Math. 126 (1996) 85–109.
[Th90] W. P. Thurston, Conway’s tiling Groups, Am. Math. Monthly 97 (1990)
757–73.
[Ti72] J. Tits, Free subgroups in linear groups, J. Algebra 20 (1972) 250–270.
[Wa86] S. Wagon, The Banach-Tarski Paradox, Cambridge University Press,
1986.
[Wh88] S. White, The group generated by x → x + 1 and x → xp is free,
J. Algebra 118 (1988) 408–22.
[Wi06] H. S. Wilf, Generating Functionology, 3rd edition, A. K. Peters, 2006.
Index
230
Index 231
path, 7 word, 37
reduced, 8 associated edge path, 38
Ping Pong Lemma, 64 freely reduced, 54
presentation, 68 word metric, 164
Property FA, 89 word problem, 109
Pumping Lemma, 136 wreath product, 156