Automorphism Groups
Automorphism Groups
Lecture Notes
Bibliography 131
Appendix A. Background Material 137
A.1. Ultrafilter 137
A.2. The Axiom of Choice and its Weaker Versions 139
CONTENTS 5
Disclaimer: this is a draft and probably contains many typos and mistakes.
Please report them to [email protected].
Recommendations. Chapters 1-4 are suitable for a 3rd year bachelor course or the
beginning of a master course. Chapters 5-9 are for a master course.
Notes concerning text book literature. We use material from the following text
books:
• Cameron’s Permutation groups [40] and Oligomorphic permutation groups [39],
• Dixon and Mortimer’s Permutation groups [48],
• The collection Notes on Infinite Permutation Groups [12] by Macpherson,
Möller, and Neumann,
• Kechris’ Classical descriptive set theory [88],
• Gao’s Invariant descriptive set theory [60], and
• Hodges’ Model theory [72].
Oligomorphic permutation groups are the topic of the short and stimulating book by
Peter Cameron [39]. The book on permutation groups by the same author [40] is
mostly about finite permutation groups, but Section 5 covers oligomorphic permuta-
tion groups. There is even less material on infinite permutation groups in [48]. The
notes in [12] are on infinite permutation groups, but they neglect the topological as-
pect of the topic. Hodges [73] is titled ‘model theory’, but covers many fundamental
things for permutation groups on the way, including topological aspects. Relevant
facts about Polish groups can be found in [60, 88].
Prerequisites. The text is essentially self-contained. Signatures, structures, sub-
structures, etc. are introduced, even though many readers may be familiar with these
concepts. However, we do not introduce first-order logic, but rather refer to logic
bachelor course notes [18]; the same applies for axiomatic set theory (we assume
familiarity for instance with Zorn’s lemma).
The present notes contains a few theorems that are just stated but not proved,
for instance
• the theorem of Ryll-Nardzewski (Theorem 3.2.3); however, we have ex-
tracted all the consequences of the (proof of the) Ryll-Nardzewski theorem
that we need in this text into Theorem 3.1.1 which we do prove. The full
proof of Theorem 3.2.3 can be found in many text books on model theory.
• the theorem of Birkhoff-Kakutani (Theorem 4.2.8); this is covered for ex-
ample in [60, 88].
• a consequence of a theorem of Lusin-Sierpiński (Theorem 6.3.12); this is
again covered in [88] (21.6).
6 CONTENTS
Exercises.
The text contains 153 exercises;
some of them are graded using the
Mandala scale.
Acknowledgements.
The author thanks the participants
of the course on automorphism groups
at TU Dresden
in the winter semester 2022/23.
CHAPTER 1
Permutation Groups
Examples.
• The set of all permutations of X, denoted by Sym(X). We use Sn as a
shortcut for Sym({1, . . . , n}) and Sω as a shortcut for Sym(N).
• The set of functions {ta : Z → Z | ta (x) = x + a, a ∈ Z} ⊂ Sym(Z).
• The set of all order-preserving permutations of Q.
All of the three examples of permutation groups given above are transitive, i.e., for
any a, b ∈ X there exists u ∈ G such that u(a) = b.
Inspiration “Erlanger Programm” of Felix Klein: Understanding structure (in his
case, geometry) by understanding its symmetry (via permutation groups).
Definition 1.0.1. Let G be a permutation group on a set A and H a permutation
group on a set B. Then a bijection i between A and B is called a (permutation group)
isomorphism if there exists a bijection ξ between G and H such that for all g ∈ G and
a ∈ A we have
i(g(a)) = ξ(g)(i(a)).
1.1. Structures
A signature τ is a set of relation and function symbols. Each symbol is equipped
with an arity k ∈ N. A τ -structure A is a set A (the domain of A) together with
• a relation RA ⊆ Ak for each k-ary relation symbol in τ , and
• a function f A : Ak → A for each k-ary function symbol in τ ; here we allow
the case k = 0 to model constant symbols.
Unless stated otherwise, A, B, C, . . . denote the domains of the structures A, B, C,
A A A A
. . . , respectively. We sometimes write (A; R1 , R2 , . . . , f1 , f2 , . . . ) for the relational
A A A A
structure A with relations R1 , R2 , . . . and functions f1 , f2 , . . . When there is no
danger of confusion, we use the same symbol for a function and its function symbol,
and for a relation and its relation symbol. We say that a structure is infinite if its
domain is infinite.
7
8 1. PERMUTATION GROUPS
Moreover, note that injective homomorphisms are embeddings (this is not true for
structures whose signature contains relation symbols! Find a counterexample!). 4
Note that if G and H are isomorphic (as permutation groups), then the corre-
sponding abstract groups are isomorphic as well, but the converse need not be true
(see Corollary 5.1.2 for a characterisation of permutation groups that are isomorphic
as abstract groups).
Example 6. For digraphs (see Example 1), connectivity in the sense we have
just introduced corresponds to weak connectivity in graph theory. The definition of
strong connectivity for digraphs can be found in Exercise 5. For undirected graphs,
connectivity in the sense introduced above coincides with the notion of connectivity
from graph theory. 4
Exercises.
(1) Show that a function f : A → A preserves h : A → A if and only if
f preserves the graph of h, i.e., the binary relation
{(a, h(a)) | a ∈ A}.
(2) Show that isomorphic structures have isomorphic automorphism
groups, but that the converse is false.
(3) Prove Proposition 1.1.1.
(2) |G| ≤ ω.
(3) |G| < 2ω .
Proof. For the implication from (1) to (2), let b̄1 , b̄2 , . . . be an enumeration of
the tuples b̄ ∈ Nn such that there exists gb̄ ∈ G with gb̄ (ā) = b̄. Then every g ∈ G can
be written as gb̄i ◦ h for some i ∈ N and some h ∈ Gā .
The implication from (2) to (3) is trivial.
For the implication from (3) to (1), suppose that ¬(1). We define inductively
sequences a0 , a1 , a2 , . . . and b0 , b1 , b2 , . . . of tuples of elements of N and a sequence
g0 , g1 , g2 , . . . of elements of G such that for every i ∈ N
(1) gi (bi ) = bi ,
(2) gi (ai ) 6= ai ,
(3) ai contains i as an entry,
(4) bi+1 is the concatenation of all sequences ki ◦ · · · ◦ k0 (a0 , . . . , ai ) where kj ∈
{gj , 1} for all j ∈ {0, . . . , k}.
(
gi i∈S
giS :=
1 i∈
/S
fiS := giS ◦ · · · ◦ g1S ◦ g0S .
For each j ≥ i we have fjS (ai ) = fiS (ai ) by the properties (1) and (4). In particular,
(3) implies that fjS (i) = fiS (i). Define hS : N → N by hS (i) := fiS (i) for every i ∈ N.
To see that hS is surjective, let i ∈ N and put j := (fiS )−1 (i). If j ≤ i then
If j > i, then
We claim that for every finite F ⊆ N there exists g ∈ G such that hS (x) = g(x) for all
x ∈ F . Let i := max (F ). Then note that for every x ∈ F we have hS (x) = fxS (x) =
fiS (x) and fiS ∈ G. In particular, hS is injective. Since G is closed in Sω , we have
that hS ∈ G.
It remains to show that hS 6= hT whenever S 6= T . Let i > 0 be smallest which
S
is, say, in S but not in T . Let j ≥ i be larger than all the entries in (fi−1 )−1 (ai ).
1.2. AUTOMORPHISM GROUPS 13
Then
hS ((fi−1
S
)−1 (ai )) = fjS ((fi−1
S
)−1 (ai ))
= gjS ◦ · · · ◦ gi+1
S
◦ giS (ai )
= giS (ai ) (by (4) and (1)
= gi (ai ) (since i ∈ S)
6= ai (by (2)
= gjT ◦ · · · ◦ gi+1
T
◦ giT (ai ) (since i ∈
/ T)
= fjT (fi−1
T
)−1 (ai )
= fjT (fi−1
S
)−1 (ai ) (by the choice of i)
= hT ((fi−1
S
)−1 (ai )).
1.2.2. Aut-sInv. Recall that the automorphism group of a relational structure
A, i.e., the set of all automorphisms of A, is denoted by Aut(A). In the following it
will be convenient to define the operator Aut also on sets R of relations over the same
domain A, in which case Aut(R) denotes the set of all permutations p of A such that
p and its inverse p−1 preserve all relations form R.
For P ⊆ Sym(A), and sets R of relations over the domain A, we present a
description of the closure operator P 7→ Aut(sInv(P )); the closure operator R 7→
sInv(Aut(R)) will be described in Section 3.1.
Definition 1.2.7. For P ⊆ Sym(A), we define
• hP i, the permutation group generated by P , to be the smallest permutation
group on A that contains P .
• P , the closure of P in Sym(A), to be the smallest closed subset of Sym(A)
that contains P .
Example 9. Let P be the set of permutations f of N that have finite support,
that is, the set {i ∈ N | f (i) 6= i} is finite. Then P ( P = Sym(N). 4
Proposition 1.2.8. Let P ⊆ Sym(A) be arbitrary. Then Aut(sInv(P )) = hP i
equals the smallest permutation group that contains P and is closed in Sym(A).
Proof. Let P 0 be the smallest permutation group that contains P and is closed
in Sym(A). Since P ⊆ P 0 and P 0 is a permutation group, we must have hP i ⊆ P 0 , and
therefore also hP i ⊆ P 0 since P 0 is closed in Sym(A). To show the converse inclusion
P 0 ⊆ hP i, it suffices to verify that hP i is a closed subgroup of Sym(A). Since hP i is
clearly closed in Sym(A) we only have to show that hP i contains compositions and
inverses. We do the verification for closure under compositions on finite subsets F of
A. Indeed, when f, g ∈ hP i, then there are f 0 , g 0 ∈ hP i such that f (x) = f 0 (x) for all
x ∈ F and g(x) = g 0 (x) for all x ∈ f (F ). We therefore have g(f (x)) = g 0 (f 0 (x)) for
all x ∈ F , and hence g ◦ f ∈ hP i, as desired.
We now show that hP i ⊆ Aut(sInv(P )). Let p ∈ hP i be arbitrary, and let R be
from sInv(P ). We have to show that p and p−1 preserve R. Let t ∈ R; we have that
p(t) = q1 ◦ · · · ◦ qk (t) for some permutations q1 , . . . , qk ∈ P ∪ P −1 . Since q1 , . . . , qk
preserve R, we have that q(t) ∈ R. The argument for p−1 is analogous.
Finally, we show Aut(sInv(P )) ⊆ hP i. Let p be from Aut(sInv(P )). It suffices
to show that for every finite subset {a1 , . . . , an } of A there is a q ∈ hP i such that
p(ai ) = q(ai ) for all i ≤ n. Consider the relation {(q(a1 ), . . . , q(an )) | q ∈ hP i}. It is
preserved by all permutations in P . Therefore, p preserves this relation, and so there
exists q ∈ hP i as required.
14 1. PERMUTATION GROUPS
Exercises.
(7) Let G be the permutation group on Z that consists of all shift
operations {x 7→ x + c | c ∈ Z}. Is G closed?
(8) Let G be the permutation group on Z that is generated by the
transpositions τi := (i, −i), for i ∈ Z. What is the cardinality
of G, and what is the cardinality of G?
(9) Let P = {f, g} ⊆ Sym(Z) where f is a transposition and g is
x 7→ x + 1. Determine the cardinalities of hP i, P , and hP i.
(10) The finitary alternating group A on N is the set of all permutations
of N that can be written as a composition of an even number
of transpositions. Determine A.
(11) Let G, H ≤ Sym(A). Show that if H is closed, then G ≤ H
if and only if every relation R ⊆ An which is preserved by H
is also preserved by G.
1.3. Group Actions
We now consider abstract groups, that is, algebraic structures G over a set G
of group elements, with a function symbol for multiplication of group elements, a
function for the inverse of a group element, and the constant for the identity (see
Example 3). The link to permutation groups is given by the concept of an action of
such a group on a set, which is described below.
Definition 1.3.1. Let G be a group and X a set. An action of G on X is a
homomorphism φ from G to Sym(X). An action φ is called faithful if φ is injective.
Example 10 (The componentwise action). If G is a permutation group on a set
X and n ∈ N, then the componentwise action of G on X n is given by
ξ(g)(x1 , . . . , xn ) := (g(x1 ), . . . , g(xn )).
Note that this action is faithful unless n = 0. 4
Example 11. If G is a permutation group on a set X and n ∈ N, then the setwise
action of G on X
n is given by
ξ(g)({x1 , . . . , xn }) := {g(x1 ), . . . , g(xn )}.
If and n > 0, then this action is faithful; this follows e.g. from the argument given in
Example 76 in Chapter 5. 4
Clearly, to every action of G on X we can associate a permutation group as
considered before, namely the image of the action in Sym(X). Conversely, to every
permutation group G on a set X we can associate an abstract group G whose domain
is G (the permutations), where composition and inverse are defined in the obvious
way, and which acts on X faithfully by φ(g) := g.
In this way we can also use other terminology introduced for permutation groups
(such a transitivity, congruences, primitivity, etc.) for group actions. For instance,
we say that an action ξ : G → Sym(X) is transitive if the permutation group ξ(G) ≤
Sym(X) is transitive. We give an alternative characterisation of action which in many
texts is taken to be the official definition.
Proposition 1.3.2. Let G be a group and X a set. The φ : G → Sym(X) is an
action of G on X if and only if the map · : G × X → X defined by g · x := φ(g)(x)
satisfies
• (gh) · x = g · (h · x) for all g, h ∈ G and x ∈ X, and
• 1 · x = x for every x ∈ X.
1.3. GROUP ACTIONS 15
The action φ is faithful if and only if for any two distinct g, h ∈ G there exists an
x ∈ X such that g · x 6= h · x.
Proof. The proof is just moving symbols.
Exercises.
(12) Show that if (a1 a2 . . . )(b1 b2 . . . ) . . . is the cycle representation of g ∈ Sn , and
f ∈ Sn , then (f (a1 )f (a2 ) . . . )(f (b1 )f (b2 ) . . . ) . . . is the cycle representation
of f −1 gf .
16 1. PERMUTATION GROUPS
(13) Let H 1 and H 2 be subgroups of Sym(X). Show that H 1 and H 2 are iso-
morphic as permutation groups if and only if there exists f ∈ Sym(X) such
that
H1 = {f hf −1 | h ∈ H2 }.
(14) Let G be a group with an action φ on A, and let A be
the corresponding G-set, i.e., the structure with domain A
which contains a unary operation for every permutation
in the image of φ. Show that the domains of the substructures
of A are precisely the orbits of the action of G on A.
(15) (Exercise 1 on page 9 of [39]) Let H a subgroup of G.
When is the the action φ of G on G/HT by left translation faithful?
Hint. Show that the kernel of φ is g∈G g −1 Hg.
(16) Let k ≥ 1 and n > k. Show that the setwise action of Sn
on k-element subsets of {1, . . . , n} is faithful.
In the following we review the classical theory how permutation groups can be
built from simpler permutation groups forming various forms of products. The direct
product of a sequence of groups (Gi )i∈I is the product of this sequence as defined in
general in Section 1.1.6; note that the product is again a group. Products appear in
several ways when studying permutation groups; the first is when we want to describe
the relation between a permutation group and its ‘transitive constituents’, described
in the following.
Exercises.
18 1. PERMUTATION GROUPS
Exercises.
(21) Let G ≤ Sn be primitive. Show that if G contains a transposition, then
G = Sn .
(22) Show that if B is a block of the permutation group G and a ∈ B, then Ga
is contained in GB .
(23) Show that if B is a block of a transitive permutation group G, then
{g|B : g ∈ GB }
is transitive as well.
(24) (from [48], Exercise 1.5.8) Let G ≤ S6 be the group generated by
{(123456), (26)(35)}.
Find all blocks that contain 1.
Find all subgroups of G that contain G1 .
(25) Give examples of permutation groups G ≤ S2n which cannot
be generated by fewer than n elements.
(26) (from [48], Exercise 1.5.14) Suppose that G ≤ Sn has r orbits.
Show that G can be generated by a subset of size n − r
(in particular, every permutation group on n elements
can be generated by n − 1 elements).
Proof. (1) ⇒ (2): to verify that E is a congruence, we have to show that for
all (a1 , b1 ), (a2 , b2 ) ∈ E, (a1 a2 , b1 b2 ) ∈ E. Indeed, (a1 a2 )(b1 b2 )−1 = a1 (a2 b−1 −1
2 )b1 ∈
−1 −1
a1 N b1 = N a1 b1 ⊆ N N = N .
(2) ⇒ (3): g 7→ gN is a group homomorphism from G to G/N .
(3) ⇒ (4): For g ∈ G and v ∈ h−1 (0), we must show that gvg −1 ∈ h−1 (0).
Indeed, h(gvg −1 ) = h(g)h(v)h(g)−1 = h(g)0h(g)−1 = 0.
(4) ⇒ (1): assume that gN g −1 ⊆ N for all g ∈ G. Let a ∈ G be arbitrary.
Applying the assumption for g = a we find that aN ⊆ N a. Applying the assumption
for g = a−1 we find that a−1 N (a−1 )−1 = a−1 N a ⊆ N , and hence N a ⊆ aN . We
conclude that aN = N a.
Example 16. The alternating group of degree n is the subgroup An of S n which
consists of all even permutations, i.e., the permutations that can be written as a
composition of an even number of transpositions. Then the map sgn that sends
g ∈ S n to 0 if g ∈ An , and to 1 otherwise, is a homomorphism from S n to Z2 and An
is a normal subgroup of Sn . 4
Non-trivial groups without non-trivial proper normal subgroups are called simple.
Exercises.
(27) Show that the group of permutations of N with finite support is a normal
subgroup of Sym(N).
(28) Show that AT5 has no proper non-trivial normal subgroups.
(29) Show that g∈G g −1 Hg is the largest normal subgroup of G which is con-
tained in H (also see Exercise 15).
1.5.2. Semidirect products: motivation. Let G be a group and let K and
H be subgroups of G. We have already defined the set-product KH := {kh | h ∈
K, k ∈ H} in Exercise 19. Note that KH might not be a subgroup (Exercise 30).
However, if H or K is a normal subgroup, then KH is a subgroup. For instance, if
K is a normal subgroup, then
(kh)(k 0 h0 ) = (khkh−1 )(hh0 ) ∈ KH (1)
−1 −1 −1 −1 −1 −1
and (kh) =h k = (h k h)h ∈ KH.
Example 17. Let G = S n , for n ≥ 3, let N be the normal subgroup An of S n
(see Example 16), and let H be the subgroup of G generated by the transposition
(12), i.e., H = {id, (12)}. Then G = N H, because every element g ∈ G is either in
An or can be written as (g(12))(12), which is in HN since g(12) ∈ An and (12) ∈ H.
Clearly, N ∩ H = {id}. However, S n is not isomorphic to An × Zn : for n ≥ 3, we
have
(123) ◦ (12) = (132) 6= (32) = (12) ◦ (123),
while
((123), 1)(1, (12)) = ((123), (12)) = (1, (12))((123), 1)
(see property (c) in Exercise 19). 4
Note the appearance of hkh−1 in (1), which defines a group action of H on K
(generalising Example 13). This group action is in fact a homomorphism from H to
Aut(K). Such homomorphisms from H to Aut(K) will be the starting point of our
first definition of semidirect products in the next section, which is in the setting where
we do not require that K and H are subgroups of the same group G (and which are
therefore called outer semidirect products).
Exercises.
22 1. PERMUTATION GROUPS
1.5.3. The outer semidirect product. Let H and N be groups and let θ : H →
Aut(N ) be a homomorphism.
Example 18. In our running example, H is Z2 and N is An for n ≥ 3 (see
Example 16). Note that for every t ∈ Sn the map αt : N → N given by g 7→ tgt−1
(conjugation) is from Aut(N ). Pick any t ∈ Sn \ An such that t2 = 1. Then the
map that sends 1 ∈ Z2 to αt and that sends 0 to idN is a homomorphism from H to
Aut(N ). 4
Definition 1.5.2. The semidirect product of N by H with respect to θ, de-
noted by N oθ H (or H nθ N ), is the group G with the elements N × H and group
multiplication defined by
(u, x)(v, y) := (uθ(x)(v), xy)
for all (u, x), (v, y) ∈ G. If the reference to θ is clear, we use o without the subscript.
Note that if θ is the trivial homomorphism that maps every element of H to
idN ∈ Aut(N ), then the semidirect product equals the direct product (we will see in
Exercise 31 that a converse of this statement is true as well). Definition 1.5.2 contains
some claims that we still have to verify. Multiplication is indeed associative:
((u, x)(v, y))(w, z) = (uθ(x)(v), xy)(w, z)
= (uθ(x)(v)θ(xy)(w), (xy)z)
= (uθ(x)(v)θ(x)(θ(y)(w)), x(yz)) (since θ is a homomorphism)
= (uθ(x)(vθ(y)(w)), x(yz)) (since θ(x) ∈ Aut(N ))
= (u, x)(vθ(y)(w), yz)
= (u, x)((v, y)(w, z)).
In the following, we write x(v) instead of θ(x)(v) for better readability. Clearly,
(1, 1) is a neutral element, and the inverse of (u, x) is (x−1 (u−1 ), x−1 ):
(u, x)(x−1 (u−1 ), x−1 ) = (ux(x−1 (u−1 )), xx−1 )
= (uu−1 , 1) = (1, 1)
Note that H ∗ := {(1, x) | x ∈ H} is a subgroup of G that is isomorphic to H, and
that N ∗ := {(u, 1) | u ∈ N } is a subgroup of G isomorphic to N . The next proposition
collects some further important properties of semidirect products (compare them with
the properties of direct products in Exercise 19!).
Proposition 1.5.3. Let G = N o H. Then
• G = N ∗H ∗,
• N ∗ ∩ H ∗ = {(1, 1)}, and
• N ∗ / G,
Proof. To see that G = N ∗ H ∗ it suffices to observe that (u, x) can be written as
(u, 1)(1, x), and obviously N ∗ ∩ H ∗ = {(1, 1)}. Finally, for (u, x) ∈ N ∗ and (v, y) ∈ G
we have
(u, x)(v, 1)(x−1 (u−1 ), x−1 ) = (ux(v), x)(x−1 (u−1 , x−1 )
= (ux(v)x(x−1 (u−1 )), xx−1 )
= (ux(v)u−1 , 1) ∈ N
1.5. SEMIDIRECT PRODUCTS 23
Proof. (1) implies (2): Suppose that n1 , n2 ∈ N and h1 , h2 are such that n1 h1 =
n2 h2 . Then in particular, n1 H = n2 H, which is the case if and only if n−1 2 n1 ∈
N ∩ H = {1}, and hence n1 = n2 . Similarly we deduce that h1 = h2 .
(2) implies (3). Let µ : G → H be the function that maps g ∈ G to the unique
h ∈ H such that g = nh for some n ∈ N . Then µ is a homomorphism: if g1 = n1 h1
24 1. PERMUTATION GROUPS
and g2 = n2 h2
µ(g1 g2 ) = µ(n1 h1 n2 h2 )
= µ(n1 h1 n2 h1−1 h1 h2 )
= h1 h2 (since h1 n2 h−1
1 ∈ N)
= µ(n1 h1 )µ(n2 h2 )
= µ(g1 )µ(g2 ).
−1
Then µ (1) = N and for any h ∈ H we have µ(h) = µ(1h) = h.
(3) implies (4): The restriction of σ to H is a homomorphism from H to G/N .
It is injective since for u ∈ H we have σ(u) = 1G/N if and only if u ∈ N if and only if
µ(u) = 1H if and only if u = 1H . It is surjective since for every [g]N ∈ G/N we have
that µ(g) = µ(µ(g)), so [g]N = [µ(g)]N = σ(µ(g)).
(4) implies (5): let τ : H → G/N be the restriction of the factor map σ which
is an isomorphism by (4). Then β := τ −1 σ : G → H is a surjective homomorphism
whose kernel is N . Choosing ρ : H → G to be the inclusion map we obtain β ◦ ρ =
τ −1 σρ = idH .
(5) implies (6): we may assume that α and ρ are inclusion maps. Define θ : H →
Aut(N ) as n 7→ hnh−1 . We claim that N oθ H is isomorphic to G, the isomorphism
ξ being (n, h) 7→ nh. We verify that ξ is a homomorphism:
ξ((n1 , h1 )(n2 , h2 )) = ξ(n1 h1 n2 h−1
1 , h1 h2 )
= n1 h1 n2 h−1
1 h1 h2
= n1 h1 n2 h2 = ξ(n1 , h1 )ξ(n2 , h2 ).
The homomorphism ξ is injective: if ξ(n1 , h1 ) = n1 h1 = 1 then
1 = β(n1 h1 ) = β(n1 )β(h1 ) = 1 · β(h1 ),
and hence h1 = 1 since β is injective. Since n1 h1 = 1 this implies that n1 = 1, too.
To show that ξ is surjective let g ∈ G. We claim that N g = N β(g). It suffices to
show that gβ(g)−1 ∈ N , i.e., lies in the kernel of β. And indeed,
β(gβ(g)−1 ) = β(g)β(β(g))−1 ) = β(g)β(g)−1 = 1.
Hence, there exists n ∈ N such that g = nβ(g). Then ξ(n, β(g)) = nβ(g) = g which
shows that ξ is surjective.
(6) implies (1): this is Proposition 1.5.3.
If the equivalent conditions in Proposition 1.5.4 apply, then G is called a split
extension of N (by H). We also say that G splits over N .
Example 20. Revisiting Example 19, we note that the short exact sequence
sgn
1 → An → Sn → Z2 → 0 splits: any homomorphism ρ from Z2 to Sn that maps 1
to an element t ∈ Sn \ An such that t2 = 1 satisfies sgn ◦ρ = idZ2 . Proposition 1.5.4
then implies that S n is isomorphic to An o Z2 . 4
Here is an example of a short exact sequence that does not split.
Example 21. Let G := (Z6 ; +). Then h : G → (Z2 ; +) given by h(g) := g
mod 2 is a surjective homomorphism, and there is an isomorphism i between Z3 and
the kernel N of h. We then have the short exact sequence
i h
1 −→ N −→ G −→ Z2 −→ 1.
However, there is no homomorphism r : Z2 → Z6 such that h ◦ r = idH since any non-
constant homomorphism s : Z2 → Z6 would have to map 1 to 3 since s(1) + s(1) = 0
1.5. SEMIDIRECT PRODUCTS 25
implies that s(1) = 3, but then h ◦ s(1) = h(3) = 0 6= 1. So the sequence does not
split, and the equivalent conditions from Proposition 1.5.4 do not apply. 4
Exercises.
(31) Show that in a semidirect product N oθ H, the subgroup H is normal if and
only if θ : H → Aut(N ) is trivial (in the sense that it maps every h ∈ H to
idN ), and in this case N oθ H = N × H.
1.5.5. Application: the wreath product. We will now describe a natural op-
eration to construct new structures from known structures, and then describe how the
semidirect product helps to explain the automorphism groups of the new structures.
We start with a simple example.
Example 22. Let A the disjoint union of two copies of the 5-element clique
K 5 = ({1, 2, . . . , 5}; 6=). Note that G = Aut(A) has a normal subgroup N which is
isomorphic to S 5 × S 5 . Also note that G/N is isomorphic to Z2 , and that G is in fact
isomorphic to (S 5 )2 o Z2 . 4
Generalising Example 22, we may start from any two structures A and B with
disjoint relational signatures σ and τ . We will define a new structure A[B]; the idea
is that we replace the elements of A by copies of B.
Formally, we create a copy B a of B for every element a of A such that all the B a
have pairwise disjoint domains. Let E be a binary relation symbol that is not already
in σ ∪ τ . Then A[B] is the σ ∪ τ ∪ {E}-structure C defined as follows. The τ -reduct
of C equals the disjoint union of the B a . The relation E C is the equivalence relation
such that E(x, y) holds for x, y ∈ C if and only if x and y lie in the same copy of B
in C. For every relation symbol R ∈ σ of arity k we set
RC := {(c1 , . . . , ck ) | there is (a1 , . . . , ak ) ∈ RA and ci ∈ B ai for i ∈ {1, . . . , k}}.
In order to describe Aut(C) we need the following definition.
Definition 1.5.5 (Wreath product). Let G be a group and let H be a group
acting on a set A. Let N := GA . Note that for every h ∈ H and n ∈ N the map
(na )a∈A 7→ (nh−1 (a) )a∈A is an automorphism of N , and that the map θ that sends
h ∈ H to this automorphism is a homomorphism from H to Aut(N ). Define
G Wr H := N oθ H.
Proposition 1.5.6. Let G be a group acting on a set B and H be a group acting
on a set A. Then G Wr H has the following action on B × A:
(n, h) · (b, a) := (nh(a) (b), h(a)).
Proof. We verify the two conditions from Proposition 1.3.2. Let (b, a) ∈ B × A.
First note that
A
1G Wr H (b, a) = (1G , 1H ) · (b, a) = (1G (b), 1H (a)) = (b, a).
If (n, h), (n0 , h0 ) ∈ G Wr H, then
(n0 , h0 ) · ((n, h) · (b, a)) = (n0 , h0 ) · (nh(a) (b), h(a))
= (n0h0 (h(a)) (nh(a) (b)), h0 h(a))
= (n0h0 (h(a)) nh(a) (b), h0 h(a))
= (n0h0 (h(a)) θ(h0 )(n)h0 h(a) (b), h0 h(a))
= ((n0 θ(h0 )(n))h0 h(a) (b), h0 h(a))
= (n0 θ(h0 )(n), h0 h) · (b, a) = (n0 , h0 )(n, h) · (b, a).
26 1. PERMUTATION GROUPS
If G and H are permutation groups, then we also use G Wr H for the permutation
group on B × A induced by this action. Note that if H acts on B with |B| > 1 and
|A| > 1, then the permutation group G Wr H is imprimitive, with block {(b, a) | b ∈
B} for every a ∈ A.
Proposition 1.5.7. For any two structures A and B we have
Aut(A[B]) = Wr(Aut(A), Aut(B)).
Exercises.
(32) Give an example of two structures A and B that illustrates why we need the
extra equivalence relation in the definition of the structure A[B].
(33) Describe the automorphism group of the digraph depicted in Figure 1.2.
(34) Show that there is no tree whose automorphism group is isomorphic to
(Z3 ; +). Hints.
• Show that every tree has a center, i.e., a vertex or an edge that is fixed
by every automorphism.
• Find an explicit description of point stabiliser of the automorphism
group of a tree.
CHAPTER 2
Counting Orbits
27
28 2. COUNTING ORBITS
Exercises.
(35) Show that Proposition 2.1.1 is false if G
is a permutation group on a finite set.
(36) Prove that (k + 1)-transitivity implies k-transitivity, for all k ≥ 1.
∗
(37) Show that if G is a permutation group on an infinite set, then fG (k) ≤
∗
fG (k + 1), for all k ≥ 1.
∗ ∗ ∗
(38) Show that if there exists a k such that fG (k) = fG (k+1), then fG (k+1) = 1.
(39) Show that a permutation group G on a set A is highly transitive if and only
if G = Sym(A) = Aut(A; =).
(40) Let (A; E) be a countably infinite structure where E denotes an equivalence
relation with infinitely many infinite classes. Describe the automorphism
group Aut(A; E). How many orbits of n-subsets are there?
(41) (Exercise 3 on page 57 in [39]) Let (A; E2 ) be a countably infinite
structure where E2 denotes an equivalence relation with infinitely
many classes of size two, and let (A; E 2 ) be a structure where E 2
denotes an equivalence relation with two infinite classes.
Show that Aut(A; E2 ) and Aut(A; E 2 ) have the same number
of orbits of n-subsets, for all n.
Figure 2.1. A tree with arbitrarily long paths, but no infinite paths.
We now state Ramsey’s theorem in it’s full strength; the proof is similar to the
proof of Theorem 2.2.3 shown above.
Theorem 2.2.4 (Ramsey’s theorem). Let s, c ∈ N. Then N → (ω)sc .
30 2. COUNTING ORBITS
A proof of Theorem 2.2.4 can be found in [73] (Theorem 5.6.1); for a broader
introduction to Ramsey theory see [63]. It is easy to derive the following finite version
of Ramsey’s theorem from Theorem 2.2.4 via König’s tree lemma.
Theorem 2.2.5 (Finite version of Ramsey’s theorem). For all c, m, s ∈ N there
is an l ∈ N such that [l] → (m)sc .
Proof. A proof by contradiction: suppose that there are positive integers c, m, s
such that for all l ∈ N there is a χ : [l]
s → [c] such that (∗)[l] for all m-subsets M of [l]
the mapping χ is not constant on M s . We construct a tree as follows. The vertices
are the maps χ : s → [c] that satisfy (∗)[l] . We make the vertex χ : [l]
[l]
s → [c]
adjacent to χ : [l+1] 0
s → [c] if χ is a restriction of χ . Clearly, every vertex in the
tree has finite degree. By assumption, there are arbitrarily long paths that start in
the vertex χ0 where χ0 is the map with the empty domain. By Lemma 2.2.1, the
tree contains an infinite path χ0 , χ1 , . . . We use this to define a map χN : Ns → [c] as
follows. For every n ∈ N, there exists a c0 ∈ [c] and an i0 ∈ N such that χi (S) = c0
for all S ∈ [n] and i ≥ i0 . Define χN (S) := c0 for all S ∈ [n]
s s . Then χN satisfies
(∗)N , a contradiction to Theorem 2.2.4.
Here comes a variant of Ramsey’s theorem.
Lemma 2.2.6. Let X be an infinite set. Suppose that χ : Xs → [c] is surjective.
Then there exist infinite sets X1 , . . . , Xc ⊆ X and k1 , . . . , kc ∈ [c] such that ki ∈ χ Xsi
Exercises.
(42) Let (X; <) be a partially ordered set on a countably infinite set X.
Show that (X; <) contains an infinite chain, or an infinite antichain.
(43) Show that an infinite sequence of elements of a totally ordered set
contains one of the following:
• a constant subsequence;
• a strictly increasing subsequence;
• a strictly decreasing subsequence.
Derive the Bolzano-Weierstrass theorem (every bounded sequence in Rn has
a convergent subsequence), using the completeness property of Rn .
(44) Show that for every permutation group on an infinite set A
with finitely many orbitals there are pairwise distinct x, y, z ∈ A
such that (x, y), (y, z), (x, z) lie in the same orbital.
Sym(Q)
Aut(Q;Sep)
Aut(Q;Betw) Aut(Q;Cycl)
Aut(Q;<)
Exercises.
(45) Show that Aut(Q; Cycl) strictly contains Aut(Q; <).
(46) Label each edge in Figure 2.2 by its index (that is, the index
of the group at the bottom in the group at the top of the edge).
CHAPTER 3
3.1. sInv-Aut
There is a surprising link between oligomorphicity of permutation groups and first-
order logic. Let A be a τ -structure and let φ(x1 , . . . , xk ) be a first-order formula with
free variables from x1 , . . . , xk , i.e., a formula built in the usual way with existential and
universal quantifiers, conjunction, disjunction, negation, equality, relation symbols
from τ , terms build from function symbols in τ , variables x1 , . . . , xk , and the quantified
variables. If a1 , . . . , ak ∈ A then we write A |= φ(a1 , . . . , ak ) if the formula φ evaluates
in A to true when instantiating xi with ai . For a rigorous definition, we refer to
course notes (e.g., [18]) or text books (e.g., [71]) in mathematical logic, or text books
in model theory (e.g., [72, 116, 152]). We say that φ(x1 , . . . , xk ) defines over A the
relation
{(a1 , . . . , ak ) ∈ Ak | A |= φ(a1 , . . . , ak )}.
Example 23. The formula x1 < x2 < x3 ∨ x3 < x2 < x1 defines the relation
Betw over (Q; <), and the formula ∀y(x1 < y ∨ x1 = y) defines the relation {0} over
(Q≥0 ; <). 4
Theorem 3.1.1. Let A be a structure such that Aut(A) is oligomorphic. Then
R ∈ sInv(Aut(A)) if and only if R is first-order definable over A.
The proof of Theorem 3.1.1 can be found below. One direction of the equivalence
is easy to show, and holds for general relational structures.
Proposition 3.1.2. Let A be a structure. If R is first-order definable in A, then
R ∈ sInv(Aut(A)).
Proof. Straightforward induction over the syntactic structure of first-order for-
mulas and their semantics.
33
34 3. OLIGOMORPHIC PERMUTATION GROUPS
Corollary 3.1.7. Let B and C be structures on the same domain such that
Aut(B) and Aut(C) are oligomorphic. Then Aut(B) = Aut(C) if and only if B and
C are (first-order) interdefinable in the sense that all relations of B have a first-order
definition in C and vice-versa.
If two countable structures have oligomorphic automorphism groups that are
isomorphic as permutation groups (Definition 1.0.1), then this corresponds to a model-
theoretic relation between the structures.
Definition 3.1.8. Two structures A and B are called bi-definable if there exists
a bijection f : A → B such that every R ⊆ An is definable in A if and only if f (A) is
definable in B.
Corollary 3.1.9. Let A and B be countable structures such that Aut(A) and
Aut(B) are oligomorphic. Then Aut(A) and Aut(B) are isomorphic as permutation
groups if and only if A and B are bi-definable.
Proof. Exercise 49.
Corollary 3.1.11. Let A and B be structures with the same domain and oligo-
morphic automorphism groups. Then Aut(A) ≤ Aut(B) if and only if every relation
that is first-order definable in B is first-order definable in A.
Proof. Exercise 50.
The following example demonstrates that Theorem 3.1.1 fails if we keep the as-
sumption that Aut(A) has only finitely many orbits of n-tuples, for every n ∈ N, but
do not require that the domain of A is countable (which is part of the definition of
oligomorphicity).
Example 25. Consider the structure (R; ≺) where ≺ is the binary relation
Q × (R \ Q) ∪ {(x, y) | x < y, x, y ∈ Q or x, y ∈ R \ Q}.
Let a ∈ Q and b ∈ R \ Q. First note that every formula φ(x) that holds on a
also holds on b; this can be shown by induction over the shape of formulas, using
that we can extend isomorphisms between finite substructures (R; ≺) by one more
point in the image or pre-image. However, there is no automorphism of (R; ≺) that
maps a to b, because the set {x | x ≺ a} is countable, but the set {x | x ≺ b} is
uncountable. Hence, the unary relation consisting of all elements of Q is preserved by
all automorphisms of (R; ≺) but not first-order definable in (R; ≺). It is also easy to
see that the automorphism group of (R; ≺) has finitely many orbits of n-tuples, for
all n ∈ N (the expansion with the unary relation Q is homogeneous). 4
Remark 3.1.12. Note that if G1 and G2 act oligomorphically on A and B, re-
spectively, then the natural intransitive action of G1 × G2 is also oligomorphic: when
a(n) is the number of orbits of the componentwise action of G1 on An , and b(n) is
the number of orbits of the componentwise action ofP G2 on B, then the number of
orbits of the componentwise of G1 × G2 on A ∪ B is 0≤i≤n a(i)b(n − i), and hence
finite for all n.
If A and B have the same signature τ , then the automorphism group of the τ -
structure A × B (see Definition 1.1.6) contains the image of the product action of
Aut(A) × Aut(B) on A × B. The number of orbits of n-tuples of this action can
be bounded by an bn where an is the number of orbits of n-tuples in A and bn is the
number of orbits on n-tuples in B. Hence Aut(A × B) is oligomorphic.
Exercises.
(49) Prove Corollary 3.1.9.
(50) Prove Corollary 3.1.11. Hint: Exercise 11 and Theorem 3.1.1.
(51) Show that the assumption of Example 24 that Aut(A) oligomorphic
is necessary (there is even a directed graph A such that C(Aut(A))
contains automorphisms that are not definable in A).
(52) Show that (Z2 )N is isomorphic to the automorphism group
of a countable structure.
(53) Show that (Z2 )N is not isomorphic to the automorphism group
of an ω-categorical structure.
Cantor [43] proved that the linear order of the rational numbers (Q; <), which
we will use as a running example in this section. We will see many more examples
of ω-categorical structures later. One of the standard approaches to verify that a
structure is ω-categorical is via a so-called back-and-forth argument. To illustrate, we
give the back-and-forth argument that shows that (Q; <) is ω-categorical; much more
about this important concept in model theory can be found in [73, 132].
Proposition 3.2.1. The structure (Q; <) is ω-categorical.
Proof. Let A be a countable model of the first-order theory T of (Q; <). It is
easy to verify that T contains (and, as this argument will show, is uniquely given by)
• ∃x. x = x (no empty model)
• ∀x, y, z ((x < y ∧ y < z) ⇒ x < z) (transitivity)
• ∀x. ¬(x < x) (irreflexivity)
• ∀x, y (x < y ∨ y < x ∨ x = y) (totality)
• ∀x ∃y. x < y (no largest element)
• ∀x ∃y. y < x (no smallest element)
• ∀x, z ∃y (x < y ∧ y < z) (density).
An isomorphism between A and (Q; <) can be defined inductively as follows.
Suppose that we have already defined f on a finite subset S of Q and that f is an
embedding of the structure induced by S in (Q; <) into A. Since <A is dense and
unbounded, we can extend f to any other element of Q such that the extension is
still an embedding from a substructure of Q into A (going forth). Symmetrically, for
every element v of A we can find an element u ∈ Q such that the extension of f that
maps u to v is also an embedding (going back ). We now alternate between going forth
and going back; when going forth, we extend the domain of f by the next element of
Q, according to some fixed enumeration of the elements in Q. When going back, we
extend f such that the image of A contains the next element of A, according to some
fixed enumeration of the elements of A. If we continue in this way, we have defined
the value of f on all elements of Q. Moreover, f will be surjective, and an embedding,
and hence an isomorphism between A and (Q; <).
A second important running example of this section is the countable random
graph (V; E). This (simple and undirected) graph with a countably infinite number
of vertices has the following extension property: for all finite disjoint subsets U, U 0 of
V there exists a vertex v ∈ V \ (U ∪ U 0 ) such that v is adjacent to all vertices in U
and to no vertex in U 0 . The existence of such a graph will be show in Example 29.
Proposition 3.2.2. The random graph (V; E) is ω-categorical.
Proof. Note that the extension property of (V; E) given above is a first-order
property; a back-and-forth argument similar to the one given in the proof of Proposi-
tion 3.2.1 shows that every countably infinite graph with this property is isomorphic
to (V; E).
The reason why we treat ω-categoricity in this course is the following theorem.
An accessible proof can be found in Hodges’ book (Theorem 6.3.1 in [73]).
Theorem 3.2.3 (Engeler, Ryll-Nardzewski, Svenonius). A countable structure B
is ω-categorical if and only if Aut(B) is oligomorphic.
If the signature of B is countable, there is another characterisation of ω-categoricity
of B via the property of Theorem 3.1.1.
Corollary 3.2.4. Let B be a structure with a countable signature and a countable
domain. Then B is ω-categorical if and only if sInv(Aut(B)) equals the set of relations
with a first-order definition over B.
38 3. OLIGOMORPHIC PERMUTATION GROUPS
Exercises.
(54) Show that in every ω-categorical structure B there exists a unique
finest equivalence relation which is definable in B and has finitely
many classes, and a unique coarsest equivalence which is definable
in B and has finite classes.
The following exercises are taken from Peter Cameron’s book “Oligomorphic permu-
tation groups”.
(55) Write down sentences φn , ψn (over the signature {=}) such that
(a) any model of φn has at least n elements;
(b) any model of ψn has exactly n elements.
(56) Write down a sentence, using equality and one binary relation
symbol, all of whose models are infinite.
Is this possible with equality alone?
1The entire theory can be adapted to general signatures that might also contain function sym-
bols; to keep the exposition simple, we restrict our focus to relational signatures in this section.
3.3. HOMOGENEOUS STRUCTURES AND AMALGAMATION CLASSES 39
Exercises.
(57) Recall that a tournament is a directed graph without self-loops such that
for all pairs x, y of distinct vertices exactly one of the pairs (x, y), (y, x) is
an arc in the graph. Show that the class of all finite tournaments has the
amalgamation property.
Example 29. Let C be the class of all finite graphs. It is even easier than in the
previous examples to verify that C is an amalgamation class, since here the free amal-
gam itself shows the amalgamation property. The Fraı̈ssé-limit of C is the countable
random graph (V; E) (also called the Rado graph) which we already encountered in
Section 3.2. 4
Proposition 3.3.6. For every k, there is a permutation group which is k-transitive
but not (k + 1)-transitive.
Proof. Let R be a relation symbol of arity k + 1. Let C be the class of all finite
{R}-structures where R denotes a relation that only contains tuples with pairwise
distinct entries. The class C is clearly closed under isomorphism, substructures, and
3.3. HOMOGENEOUS STRUCTURES AND AMALGAMATION CLASSES 41
T3 T4 T5
has only countably many isomorphism classes of structures. It also has the free
amalgamation property. Note that any two structures with at most k elements are
isomorphic, since R denotes the empty relation in those structures. Since the Fraı̈ssé-
limit is homogeneous, its automorphism group is therefore k-transitive. On the other
hand, the class contains non-isomorphic structures of size k+1 (e.g., a structure where
R denotes the empty relation and a structure where R is non-empty), and hence the
automorphism of Flim(C) is not k + 1-transitive.
Example 30. Let C be the class of all finite triangle-free graphs, that is, all
graphs that do not contain K3 as a subgraph. Again, we have the free amalgamation
property. The Fraı̈ssé-limit is up to isomorphism uniquely described as the triangle-
free graph A such that for any finite S, T ⊂ A such that S is stable (i.e., induces a
graph with no edges; such a vertex subset is sometimes also called an independent
set) there exists v ∈ A \ (S ∪ T ) which is connected to all points in S, but to no point
in T . 4
We now introduce a convenient tool to describe classes of finite τ -structures. If
N is a class of τ -structures, we say that a structure A is N -free if no B ∈ N embeds
into A. The class of all finite N -free structures we denote by Forb(N ).
Example 31. Henson [65] used Fraı̈ssé limits to construct 2ω many homogeneous
directed graphs. Note that for all classes N of finite tournaments, Forb(N ) is an
amalgamation class, because if A1 and A2 are directed graphs in Forb(N ) such that
A = A1 ∩ A2 is an induced substructure of both A1 and A2 , then the free amalgam
A1 ∪ A2 is also in Forb(N ).
Henson in his proof specified an infinite set T of tournaments T 3 , T 4 , . . . with the
property that T i does not embed into T j if i 6= j. The tournament T n , for n ≥ 3, in
Henson’s set T has vertices 0, . . . , n + 1, and the following edges (see Figure 3.1):
• (i, i + 1) for 0 ≤ i ≤ n;
• (0, n + 1);
• (j, i) for j > i + 1 and (i, j) 6= (0, n + 1).
Note that this property implies that for two distinct subsets N1 and N2 of T the
two sets Forb(N1 ) and Forb(N2 ) are distinct as well. Since there are 2ω many subsets
42 3. OLIGOMORPHIC PERMUTATION GROUPS
of the infinite set T , there are also that many distinct homogeneous directed graphs;
they are often referred to as Henson digraphs. 4
The structures from Example 31 can be used to prove various negative results
about homogeneous structures with finite signature. A better behaved class of homo-
geneous structures are those whose age is finitely bounded (this is the same terminology
as in [110]).
Definition 3.3.7. A class C of finite relational τ -structures (or a structure with
age C) is called finitely bounded if τ is finite and there exists a finite set of finite
τ -structures N such that C = Forb(N ).
Fraı̈ssé’s theorem (Theorem 3.3.5) has a converse.
Proposition 3.3.8. The age of every homogeneous relational structure has the
amalgamation property.
Proof. Let A be a homogeneous structure and let (B 1 , B 2 ) be an amalgamation
diagram with B 1 , B 2 ∈ Age(A). Then there are embeddings ei B 1 ,→ A, for i ∈ {1, 2},
and by the homogeneity of A there exists an automorphism α ∈ Aut(A) such that
α(e1 (x)) = e2 (x) for every x ∈ B1 ∩ B2 . Let C be the substructure of A with
domain α(e1 (B1 )) ∪ e2 (B2 ). Then the embedding f1 : B1 → C given by α ◦ e1 and
the embedding e2 show that C is an amalgam of (B 1 , B 2 ).
Exercises.
(58) Is the class of finite forests (i.e., simple acyclic graphs)
an amalgamation class?
(59) Let C be the class of all finite graphs G such that there is no
embedding from the 5-cycle C 5 into G.
Is C an amalgamation class?
(63) Let P be a unary relation symbol. Let D be the class of all finite
{P, <}-structures A such that <A is a linear order.
(a) Show that D is an amalgamation class.
3.3. HOMOGENEOUS STRUCTURES AND AMALGAMATION CLASSES 43
V1 := {u} ∪ {v ∈ V | (u, v) ∈ E}
B A
Exercises.
(79) Prove that the age of (Z; {(x, y) | x = y + 1}) does not have the AP,
but the WAP.
(80) Prove that the age of (Q; <, {0})
does not have the AP, but the WAP.
(81) Prove that the age of (Q; {(x, y, z) | x < y, z 6= x, z 6= y})
does not have the AP, but the WAP.
(82) A countable ω-categorical τ -structure B is called model-complete if
every first-order τ -formula is over B equivalent to an existential
τ -formula. Show that B is model-complete if and only if Aut(B)
is dense in Emb(B) ⊆ B B , where Emb(B) is the set of all
self-embeddings of B.
(83) Let A be an ω-categorical structure. The a model companion of A
is a structure B with the same age as A which is model-complete
(see the previous exercise). Show that every ω-categorical structure
has a model companion, and that the model companion is unique
up to isomorphism. We refer to B as the model companion of A.
(84) (Todor Tsankov, personal communication 2012) Prove that the age
of every ω-categorical structure A has the WAP.
(85) A structure A with a finite relational signature is called homogenizable if
there exists a definable expansion B of A by finitely many new relations
such that B is homogeneous. We say that A is boundedly homogenizable
if it is homogenizable and for every a ∈ An , n ∈ N, there exists b ∈ Am ,
m ∈ N, such that the type of (a, b) in A is equivalent to a quantifier-free
formula. Find a structure which is model-complete, homogenizable, but not
boundedly homogenizable.
More on the WAP can be found in Example 81.
that shows that I2 ◦ I1 is homotopic to the identity interpretation. That is, θ defines
in B the (d1 d2 + 1)-ary relation that contains a tuple (a, b1,1 , . . . , bd1 ,d2 ) iff
a = h2 (h1 (b1,1 , . . . , b1,d2 ), . . . , h1 (bd1 ,1 , . . . , bd1 ,d2 )) .
Let φ be an atomic τ -formula with k free variables x1 , . . . , xk . We will specify a
σ-formula that is equivalent to φ over B 0 .
^
1
∃y1,1 , . . . , ydk1 ,d2 i
θ(xi , y1,1 , . . . , ydi 1 ,d2 )
i≤k
1 k 1 k
, . . . , ydk1 ,d2 )
∧ φI1 I2 (y1,1 , . . . , y1,d 2
, y2,d2
, . . . , y2,d 2
Topological Groups
We start with a very brief introduction to concepts from topology that will be
relevant in what follows.
Example 38. On every set S, there is the trivial topology where the only open
sets are S and the empty set. 4
Example 39. Every set S can be equipped with the discrete topology: in this
topology, every subset of S is open (and hence also closed). 4
Example 42. Every set S can be equipped with the cofinite topology: in this
topology, the open sets are the empty set and every cofinite subset of S. The only
closed subsets in this topology are the finite sets and the entire set S. 4
For E ⊆ S, the closure of E, denoted by Ē, is the intersection over all closed sets
over S that contain E. A subset E of S is called dense (in S) if its closure is the full
space S. The interiour of E is the largest open set contained in E; and is denoted
by Int(E). Equivalently, Int(E) := S \ S \ E.
The subspace of S induced on E is the topological space on E where the open
sets are exactly the intersections of E with the open sets of S. When we work with
permutation groups G ⊆ Sym(X) then we will always work with the topology on G
inherited from Sym(X) in this way.
53
54 4. TOPOLOGICAL GROUPS
We have seen that R and Sym(N) share some countability and separation prop-
erties. But in some other respects, these two spaces are very different. A topological
space S is called disconnected if it is the union of two or more disjoint nonempty open
subsets, and connected otherwise. A subset of S is said to be connected if it is con-
nected under its subspace topology. The inclusion-wise maximal connected subsets
of a non-empty topological space are called the connected components of that space.
Picture in R2 !
are open, and for any irrational π they partition Q. On the other hand, R and Rd are
connected1. 4
A topological space S is totally disconnected if all connected subsets of X are
one-element sets.
Example 47. The topology of pointwise convergence on Sym(N) is totally dis-
connected: if f, g ∈ Sym(N) are distinct, there exists an i ∈ N such that S f (i) 6= g(i).
Then S(i, f (i)) is an open set that contains f , and Sym(N)\S(i, f (i)) = j6=f (i) S(i, j)
is a disjoint set that contains g, and it is open as a union of basic open sets. Hence,
no set that contains more than one element is connected. 4
Exercises.
(86) Show the claim from Example 46 that the standard topology on R
is connected.
(87) On which sets X is the cofinite topology separable?
(88) Show that R with the standard topology is not homeomorphic
to a closed subgroup of Sym(N).
(89) Show that the cofinite topology on an uncountably infinite set
is not first-countable.
4.1.2. Continuity and convergence. A map between two topological spaces
is called continuous if the pre-images of open sets are open, and open if images of
open sets are open. A bijective open and continuous map is called a homeomorphism.
There are equivalent characterisations of continuity of maps from a first-countable
space S to a topological space T that are often easier to work with. For a sequence
(sn )n≥1 of elements of S, we say that sn converges against s if for every open set U
of S that contains s there exists an n0 such that sn ∈ U for all n ≥ n0 . Note that if
T is Hausdorff, then s is unique, and called the limit of (sn )n≥1 , and we write
lim sn = s.
n→∞
For x ∈ S, we say that f is continuous at x if for every open V ⊆ T containing f (x)
there is an open U ⊆ S containing x whose image f (U ) is contained in V .
Proposition 4.1.1. Let S be a first-countable and T an arbitrary topological
space. Then for every f : S → T the following are equivalent.
(1) f is continuous.
(2) For all sequences sn , if sn converges against s, then f (sn ) converges against
f (s).
(3) f is continuous at every x ∈ S.
Proof. The implication from (1) to (2) is true even without the assumption that
S is first-countable. Let (sn )n≥1 be such that limn→∞ sn = s, and let V be open so
that f (s) ∈ V . Then U := f −1 (V ) is open, and s ∈ U . So there exists an n with
sn ∈ U . For this n we have f (sn ) ∈ V . So limn→∞ f (sn ) = f (s).
For the implication from (2) to (3), we show the contraposition. Suppose that f is
not continuous at some s ∈ S. That is, there exists an open set V containing f (s) such
that all open sets U that contain s have an image that is not contained in V . Since S is
first-countable, there exists a countable collection Un of open sets containing s so that
any open set that contains s also contains some Un . Replacing Un by ∩nk=1 Uk where
1This relies on the fact that the real numbers are (by definition!) Dedekind-complete: every
non-empty subset S of R with an upper bound in R has a least upper bound.
56 4. TOPOLOGICAL GROUPS
(4) the right end point of I(s,n) is the left end point of I(s,n+1) ,
(5) {I(s,n) | n ∈ Z} covers Is ∩ P,
(6) the n-th rational qn is an endpoint of Is for some s ∈ Zk with k ≤ n + 1.
T
Define the function f : ZN → P as follows. Given x ∈ ZN the set n∈N Ix1 ...xn must
consist of a singleton irrational:
• it is nonempty because Ix1 ...xn xn+1 ⊆ Ix1 ...xn ;
• it is a singleton because the length of Ix1 ...xn tends to zero for n → ∞.
So we can define f by
\
{f (x)} := Ix1 ...xn .
n∈N
The function f is injective because if s and t are not prefixes of each other then Is
and It are disjoint, and f is surjective because for every u ∈ P and k ∈ N there is a
unique s ∈ Zk with u ∈ Is . Finally, f is a homeomorphism because
f ({x ∈ ZN | x1 . . . xk = s}) = Is ∩ P
and the sets of the form Is ∩ P form a basis for P.
4.1.4. Metric spaces. Important examples of topologies come from metric spaces.
A pseudometric space is a pair (M, d) where M is a set and d is a pseudometric on
M , i.e., a function
d: M × M → R
such that for any x, y, z ∈ M , the following holds:
(1) d(x, y) ≥ 0 (non-negativity)
(2) d(x, y) = d(y, x) (symmetry)
(3) d(x, z) ≤ d(x, y) + d(y, z) (subadditivity or triangle inequality)
If d additionally satisfies
(4) d(x, y) = 0 ⇔ x = y (indiscernibility)
then d is called a metric, and (M, d) is called a metric space. If M 0 ⊆ M then the
restriction of d to M 0 is clearly a metric, too. Every metric on M gives rise to a
topology on M , namely the topology with the basis
{y ∈ M | d(x, y) < } | 0 < ∈ R, x ∈ M .
A topological space S is metrisable if there exists a metric d on S which is compatible
with the topology, i.e., the topology equals the topology that arises from the metric
as described above.
Example 49. The discrete metric ρ on X is defined by
1 if x 6= y,
ρ(x, y) =
0 if x = y
for any x, y ∈ X. In this case (X, ρ) is called a discrete metric space or a space of
isolated points. 4
Example 50. The distance function d(x, y) = |x−y| (absolute difference) defines
a metric on R, Rd , and on Q. The topology that arises from this metric is precisely
the standard topology on R, Rd , and on Q. 4
also surjective. To show that b ∈ D lies in the image of g, let j be such that b = bj .
Then there exists i0 ∈ N such that c := gr−1 (b) = gs−1 (b) for all r, s ≥ i, again by the
assumption that (gm ) is Cauchy with respect to d0 . Then g(c) = b by the definition
of g. Since Sym(D) is also separable, we have that Sym(D) is Polish. 4
Example 55. The Hilbert cube [0, 1]N . Here, the interval [0, 1] is equipped with
the usual topology inherited from R, and [0, 1]N carries the product topology. 4
Exercises.
(90) Every metrisable space X has a compatible metric d
which satisfies d(x, y) ≤ 1 for all x, y ∈ X.
(91) Let (X, d) be a complete metric space and S ⊆ X. Then S is
closed in X if and only if (S, d|S 2 ) is complete.
(92) Show that every metrisable space is Hausdorff.
(93) Show that every metrisable space X is regular : for any x ∈ X
and any open set U that contains x there is an open set O
that contains x such that Ō ⊆ U .
(94) Show that every metrisable space X is normal : for any
disjoint closed C, F ⊆ X there are disjoint open O, U ⊆ X
such that C ⊆ O and F ⊆ U .
(95) Let X be a metrisable space and Y ⊆ X be closed. Then Y is
a countable intersection of open sets.
(96) True or false: the closure of an open ball of radius r in
a metric space is the closed ball of radius r
in that metric space.
(97) Show that a countable product of metrisable spaces
is metrisable.
(98) Every separable metrisable space is homeomorphic to a subspace
of the Hilbert cube [0, 1]N .
For a metric space (X, d) and A ⊆ X, define diamd (A) := supb,c∈A d(b, c). Note
that for every > 0 and every open subset U of X we find an open V ⊆ U such that
diamd (V ) < .
2
Definition 4.1.5. Countable intersections of open sets are called Gδ .
In the proof of the following lemma we use the axiom of dependent choice (Ap-
pendix A.2).
Lemma 4.1.6. Every Polish subspace Y of a Polish space X is Gδ .
Proof. Let dX be a complete compatible metric on X and let dY be a complete
compatible metric on Y . Define Vn ⊆ X as the union of all open sets U ⊆ X that
satisfy
(1) U ∩ Y 6= ∅;
(2) diamdX (U ) < n1 ;
(3) diamdY (U ∩ Y ) < n1 ;
T
It suffices to show that Y = n∈N Vn . Let x ∈ Y and n ∈ N; we want to show that
x ∈ Vn . Let U1 ⊆ Y be an open set that contains x such that diamdY (U1 ) < n1 . By the
definition of the relative topology, there is an open set U2 in X such that U2 ∩Y = U1 .
Let U3 be an open subset of X that contains x such that diamdX (U3 ) < n1 . Then
U := U2 ∩ U3 satisfies all of the three conditions above. Hence, x ∈ Vn .
2The origin of the notation is G for the German word Gebiet and δ for intersection.
60 4. TOPOLOGICAL GROUPS
T
Conversely, suppose that x ∈ n∈N Vn . Then for every n ∈ N, there exists an
open subset Un of X that contains x and satisfies the three conditions given above.
By the first condition, Un ∩ Y 6= ∅. Choose xn ∈ Un ∩ Y . Then the second condition
implies that limn→∞ xn = x. The third condition implies that (xn )n∈N is Cauchy
with respect to dY . Since dY is complete, limn→∞ xn = x ∈ Y .
Remark 4.1.7. The converse of Lemma 4.1.6 is true as well: every Gδ set is
Polish. We do not need this statement in the following and therefore omit the proof.
We finally state an important property of completely metrizable spaces.
Theorem 4.1.8 (The Baire Category Theorem). Every Polish3 space S is Baire,
i.e., has the property that countable intersections of dense open sets are dense.
Proof.
T Let (Un )n∈N be a sequence of open dense sets. We want to show that
U := Un is dense. It is sufficient to show that any non-empty open set W in S
contains an element of U . Since U1 is dense, there is x1 ∈ U1 ∩ W . Hence, there is
an r1 with 0 < r1 < 1 such that {z ∈ S | d(x1 , z) ≤ r1 } ⊆ U1 ∩ W where d is the
compatible complete metric. We can continue recursively to find a sequence (xn )n∈N
of elements of S and a sequence (rn )n∈N of elements of R such that
{z ∈ S | d(z, xn ) ≤ rn } ⊆ Un ∩ Brn−1 (xn−1 ) (4)
as follows: if we have defined x1 , . . . , xn and r1 , . . . , rn satisfying (4), then density of
Un+1 guarantees that Brn (xn ) ∩ Un+1 is non-empty. Using the axiom of dependent
choices, we may choose an element xn+1 from this set such that
there is an rn+1 ∈ R with {z ∈ S | d(z, xn+1 ) ≤ rn+1 } ⊆ Un+1 ∩ Brn (xn ). (5)
Instead of using this axiom we can instead use the assumption that there exists a
countable set {u1 , u2 , . . . } which is dense in S: we may then choose xn+1 := uk for k
smallest possible so that (5) is satisfied. Since xm ∈ Brn (xn ) for all m > n, we have
that (xn )n∈N is Cauchy, and hence converges to some limit x by the completeness of d.
For any n, the set {z ∈ S | d(z, xn ) ≤ rn } is closed and hence contains x. Therefore,
x ∈ W and x ∈ Un for all n.
4.1.6. Metric Completions. Let (M1 , d1 ) and (M2 , d2 ) be two metric spaces.
An isometry between (M1 , d1 ) and (M2 , d2 ) is a function i : M1 → M2 such that
d1 (x1 , x2 ) = d2 (i(x1 ), i(x2 )) (note that i must be injective, but it is not required to
be surjective). Two metrics are called isometric if there exists a bijective isometry
between them.
Metric spaces have the advantage that we can use Cauchy sequences to talk about
points that aren’t really there. More formally:
Definition 4.1.9. A completion of a metric space (M, d) is a complete metric
space (M ∗ , d∗ ) together with an isometry i : M → M ∗ such that i(M ) is dense in M ∗ .
Proposition 4.1.10. Every metric space has a completion.
Proof. Let (M, d) be a metric space. Let C be the collection of all Cauchy
sequences in M . Define an equivalence relation ∼ on C by setting (xn ) ∼ (yn ) if
limn→∞ d(xn , yn ) = 0 for (xn ), (yn ) ∈ C. Define
• M ∗ to be the set of all equivalence classes of ∼,
X ∗ := [(xn )] | (xn ) ∈ C ,
3Using the axiom of dependent choices (DC), this assumption can be weakened from Polish to
completely metrisable (in fact, the modified statement is then equivalent to DC; see Appendix A.2).
4.1. TOPOLOGICAL SPACES 61
• i : M → M ∗ by setting for x ∈ M
φ(x) := [(x, x, . . . )].
Claim 1. d∗ is well-defined. Let (x0n ) and (yn0 ) be two Cauchy sequences such that
(x0n ) ∼ (xn ) and (yn0 ) ∼ (yn ). By the triangle inequality
d(xn , yn ) ≤ d(xn , x0n ) + d(x0n , yn0 ) + d(yn0 , yn )
and thus
|d(xn , yn ) − d(x0n , yn0 )| ≤ |d(xn , x0n ) − d(yn0 , yn )|
which tends to 0 for n → ∞, and proves that limn→∞ d(xn , yn ) = limn→∞ d(x0n , yn0 ).
Claim 2. d∗ is a metric on M ∗ . This is straightforward.
Claim 3. i is an isometry:
d∗ (i(x), i(y)) = lim d(x, y) = d(x, y) .
n→∞
Claim 4. i(M ) is dense in M ∗ . Let [(xn )] ∈ M ∗ and > 0. Since (xn ) is Cauchy,
there exists an n0 ∈ N such that d(xm , xn ) < for all m, n ≥ n0 . For z := i(xn0 ) we
have
d∗ ([(xn )], i(z)) = lim d(xn , xn0 ) ≤ .
n→∞
d∗ ([(xm n m n
n )], [(xk(n) )]) = lim d(xn , xk(n) )
n→∞
= lim sup d(xm m m n
n , xk(m) ) + lim sup d(xk(m) , xk(n) )
n→∞ n→∞
≤ 1/n0 + /2 <
Proof. Let x ∈ M1∗ . Since i1 (M ) is dense in M1∗ for each n ∈ N there exists
xn ∈ i1 (M ) with d∗1 (xn , x) ≤ n1 . Let yn := i2 (i−1
1 (xn )). Since i1 and i2 are isometries,
we have d2 (yn , ym ) = d1 (xn , xm ) for all m, n ∈ N. The sequence (xn )n∈N converges
against x, so it is Cauchy, and it follows that (yn )n∈N is Cauchy, too. Since M2∗ is
complete the sequence (yn )n∈N must converge to some y ∈ M2∗ .
Claim 1. The map f : M1∗ → M2∗ defined by f (x) := y is well-defined. Suppose
that (x0n )n∈N is another sequence of elements of M1 that converges to x. For n ∈ N,
let yn0 := i2 (i−1 0
1 (xn ); we have to show that limn→∞ yn = y.
0
Let > 0. Since limn∈N yn = y there exists m ∈ N such that d∗2 (yn , y) < /2 for
all n ≥ m. There is also a k ∈ N such that for all n ≥ k we have d∗1 (xn , x) < /4 and
d∗1 (x0n , x) < /4. Hence, d1 (xn , x0n ) ≤ d∗1 (xn , x) + d∗1 (x, x0n ) < /4 + /4 = /2. Since i1
and i2 are isometries we have d2 (yn , yn0 ) = d1 (xn , x0n ). Hence, d2 (yn , yn0 ) < /2 for all
n ≥ k. So for n ≥ max (k, m) we have d∗2 (yn0 , y) ≤ d2 (yn0 , yn )+d2 (yn , y) < /2+/2 = ,
showing that limn→∞ yn0 = y.
Claim 2. f is an isometry. Let x, x0 ∈ M1∗ and let (xn )n∈N and (x0n )n∈N be
sequences of elements of i1 (M ) that converge to x and x0 , respectively. Define yn :=
i2 (i−1 0 −1 0
1 (xn )) and yn := i2 (i1 (xn )), and we have seen that limn→∞ yn = f (x) and
limn∈N yn0 = f (x0 ). Then
d∗2 (f (x), f (x0 )) = lim d2 (yn , yn0 ) = lim d1 (xn , x0n ) = d(x, x0 ).
n∈N n∈N
there exists an y0 ∈ Y such that for every open U ⊆ X that contains x0 and every
open V ⊆ Y that contains y0 , the set U × V × Z is not finitely coveredQ by U.
We finally prove that if X1 , X2 , . . . are compact, then X = i∈N Xi is compact.
Let U be a family of open sets that that no finite subset of U covers X. We will
construct an element x = (x1 , x2 , . . . ) of X that is not covered by U. Note first
that there is an x1 ∈ X1 such that for every open U1 ⊆ X1 that contains x1 the set
U1 × X2 × X2 × · · · is not finitely covered; the proof is as the proof of Claim 1, with
X2 × X3 × · · · playing the role of Y . Next, we can find x2 ∈ X2 such that such that
for every open U1 ⊆ X1 that contains x1 and every open U2 ⊆ X2 that contains x2
the set U1 × U2 × X3 × X4 × · · · is not covered by finitely many elements of U; this
follows from Claim 3 applied to X1 ×X2 ×(X3 ×X4 ×· · · ). Continuing in this way, we
inductively define x1 , x2 , x3 , . . . such that for each n and all open Ui ⊆ Xi for i ≤ n
such that Ui contains xi , the set U1 × · · · × Un × Xn+1 × · · · is not covered by finitely
many elements of U. The element (x1 , x2 , . . . ) ∈ X is then not covered by U.
Exercises.
(99) Prove that a finite union of compact sets is compact.
In order to discuss which subsets of R and of Sym(N) are compact (with respect
to the subspace topology), we need the following definition for metric spaces.
Definition 4.1.15. A subset S of a metric space (M, d) is bounded if it is con-
tained in an open ball of finite radius, i.e., if there exists x ∈ M and a real > 0 such
that for all s ∈ S, we have d(x, s) < .
The open ball of radius and center x will be denoted by Bx () in the following.
Example 60. A subset of Rd is compact if and only if it is closed and bounded
– this is the theorem of Heine-Borel. 4
Which subsets of Sym(N) are compact?
Proposition 4.1.16. Any compact subset S of a Hausdorff topological space X
is closed in X.
Proof. If S is compact but not closed then there exists a ∈ S̄ \ S. For each
x ∈ S there exists an open set Ux that contains x but does not intersect an open set
Vx that contains a, because X is Hausdorff. Then U := {Ux | x ∈ S} is an open cover
of S, and by compactness of S there exists a finite subcover {Ux1 , . . . , Uxn } of U. But
then V := Vx1 ∩ · · · ∩ Vxn is open and contains a, and hence contains a point b in S
since a ∈ S̄. Since V is disjoint from each of Ux1 , . . . , Uxn , we have b ∈
/ Ux1 ∪ · · · ∪ Uxn ,
in contradiction to {Ux1 , . . . , Uxn } being a cover of S.
The set S is totally bounded if for every real > 0 there exists a finite collection
of open balls in M of radius whose union contains S. Clearly, a totally bounded
space is bounded, but the converse is not true: the discrete metric is bounded, but
not totally bounded: for = 1/2, we need infinitely many open -balls (points!) to
cover the infinite set.
Proposition 4.1.18. If a metric space is totally bounded, then it is separable.
Proof. If X is totally
S bounded then for each positive
S n ∈ N there exists a finite
An ⊆ X such that X = x∈An Bx (1/n). Let A := n≥0 An . Clearly A is countable.
We claim that Ā = X. Let x ∈ X. For any n ∈ N there is some yn ∈ An such that
x ∈ Byn (1/n). This gives a sequence (yn ) with d(x, yn ) < 1/n. Thus lim yn = x
which proves the claim, and separability of X.
In the proof of the following theorem, we assume the axiom of countable choice
(see Appendix A.2).
Theorem 4.1.19. For a metric space (X, d), the following are equivalent.
(1) X is compact;
(2) Every collection of closed sets in X with the finite intersection property
(every finite subcollection has a nonempty intersection) has a nonempty in-
tersection;
(3) If F1 ⊇ FT2 ⊇ F3 ⊇ · · · is a decreasing sequence of nonempty closed sets in
X, then n≥1 Fn is nonempty;
(4) X is sequentially compact, that is, every sequence in X has a convergent
subsequence;
(5) X is totally bounded and complete.
Proof. (1) ⇒ (2). Suppose that C is a collection of closed sets with empty
intersection. Then U := {X \ C | C ∈ C} is an open cover of X, and hence contains
a finite subcover of X. The complements of the members of the subcover in X give
the collection with the finite intersection property.
(2) ⇒ (3). A decreasing sequence of non-empty closed sets obviously has the
finite intersection property.
(3) ⇒ (4). Let (xn )n∈N be a sequence of points in X, and let Fn be the closure of
the set {xn , xn+1 , . . . }. Then F1 ⊇ F2 T ⊇ F3 ⊇ · · · and all the sets Fn are nonempty
and closed. Therefore, by (3), the set n≥1 Fn contains at least one point a. Then
(xn )n∈N contains a subsequence converging to a: to see this, let d be the compatible
metric, and set n0 = 1. Now suppose that nk has already been defined for k ∈ N. Since
a is in the closure of {xnk +1 , xnk +2 , . . . } there exists an n ∈ {nk + 1, nk + 2, . . . } such
that d(xn , a) < 1/(k + 1). Let nk+1 the the smallest such n. Then limk→∞ xnk = a.
(4) ⇒ (5). To prove that X is complete, let (xn ) be any Cauchy sequence in X.
By (4), there is a subsequence converging to some point a ∈ X. But then the whole
sequence (xn ) converges to a. This shows that X is complete.
Now suppose that X is not totally bounded, i.e., there exists a number > 0 such
that X has no finite covering by open balls of radius . Then we can define a sequence
(xn )n≥1 of points in X having d(xi , xj ) ≥ for all i 6= j, by the following inductive
construction: First let x1 be any point in X. Then, supposing that x1 , . . . , xn−1 have
been chosen, we know Bx1 () ∪ · · · ∪ Bxn−1 () is not the whole space. Hence we can
choose a point xn satisfying d(xi , xn ) ≥ for all i < n. On the other hand, the
sequence (xn ) cannot have any convergent subsequence; for if it had a subsequence
(xnk ) converging to a, then there would exist an integer k0 such that d(xnk , a) < /2
for all k ≥ k0 , and hence by the triangle inequality d(xnk , xnk0 ) < for all k, k 0 ≥ k0 ,
contrary to the definition of the sequence (xn ).
66 4. TOPOLOGICAL GROUPS
(5) ⇒ (4). Let (xi )i∈N be a sequence of elements from X. Let S = {xn | n ∈ N}.
If S is finite then the statement is trivial so assume that S is infinite. Since X is
totally bounded, there exists a finite cover of X with open balls of radius 1 := 1.
One of those balls, call it B1 , must contain infinitely many elements from S. Again by
total boundedness, there exists a finite cover of X with open balls of radius 2 := 1/2,
and again, one of those balls must contain infinitely many elements from B1 ∩ S; this
ball we call B2 . We continue this process, producing a sequence of balls (Bk ) of radius
1/k so that Bk ∩ Bk−1 ∩ · · · ∩ B1 contains infinitely many elements of the sequence
xi . Pick now indices n1 < n2 < n3 < · · · such that yk := xnk ∈ Bk . It is easy to
see that (yi ) is Cauchy and so by the completeness assumption on X it must have a
convergent subsequence.
(4) ⇒ (1). Let U be an open cover of X. From the implication (4) ⇒ (5) we have
that X is totally bounded, and Proposition 4.1.18 implies that X is second-countable.
By Proposition 4.1.17 (Lindelöf) we can therefore assume that U is countable, U =
{U1 , U2 , . . . }. Suppose for contradiction that U does not have a finite subcover. Pick
xn ∈ X \ (U1 ∪ · · · ∪ Un ) arbitrarily. Then by assumption, the sequence (xn ) has a
subsequence (yn ) which converges to some y0 ∈ X. Since U is a cover of X there is
some m ∈ N with y ∈ Um . But then yj ∈ / Um for all j ≥ m, which is a contradiction.
Example 62. The discrete space on S is locally compact, but only compact if S
is finite. 4
Example 63. The groups (R, +) and (Q, +) are topological groups with respect
to their standard topology. (Why?) 4
U is
{(f, h) ∈ G2 | f ◦ h ∈ S(ā, c̄)} = {(f, h) ∈ G2 | ∃b̄ (h ∈ S(ā, b̄) and f ∈ S(b̄, c̄))}
[
= S(b̄, c̄) × S(ā, b̄)
b̄∈Nn
which is open as a union of open sets. The preimage of S(ā, b̄) under the inverse map
is S(b̄, ā), which is open, too. 4
Proposition 4.2.1. Let G be a topological group, g ∈ G, and U ⊆ G open. Then
gU is open, too. If U is an open subgroup, then it is also closed.
Proof. As a consequence of Proposition 4.1.2, for every g ∈ G the function
tg : G → G defined by tg (x) := gx is continuous. The pre-image of U under the
function tg−1 is gU . Therefore, this set is open as the pre-image of an open set under
Scontinuous function. The second part follows since the complement of U in G equals
a
g∈G\U gU , a union of open sets, hence open.
Remark 4.2.2. Proposition 4.2.1 also implies that the topology on G is given by a
basis at 1G : if B is a basis of open sets at the identity, and g ∈ G, then {gU | U ∈ B}
is a basis at g.
Exercises.
(100) Let G be a topological group and let A, B ⊆ G. If A is open,
then so is AB := {ab | a ∈ A, b ∈ B}.
(101) Show that a group G with a topology on G is a topological group
if and only if the map (x, y) 7→ xy −1 is continuous from G2 to G.
(102) Show that for all n ∈ N the groups GL(n, R) and GL(n, C)
of invertible real or complex matrices are topological groups
with respect to the standard topology.
4.2.1. Continuous group actions. Recall from Section 1.3 that an action of
a group G on a set S is a homomorphism from G to Sym(S).
Definition 4.2.3. An action ξ of a topological group G on a topological space S
is called continuous if (g, s) 7→ ξ(g)(s) is continuous as a map from G × S → S.
Example 65. Recall the faithful action of G on G by left multiplication from the
proof of Cayley’s theorem, Theorem 1.3.3). This is the special case of Example 12
where H = {1}. This action is continuous since it equals the group composition which
is continuous by definition. 4
If S is a topological space, then Homeo(S) ⊆ Sym(S) denotes the set of all
homeomorphisms of S. We view Homeo(S) as a topological space with the subspace
topology inherited from S S which carries the product topology4.
4Note that it is not clear (and depends on S) whether Homeo(S) with this topology is a topo-
logical group.
68 4. TOPOLOGICAL GROUPS
Hence, there exists i ∈ {1, . . . , k} such that {ai } ∈ U (see Lemma A.1.3) and U is
principal.
In the same way it can be shown that for a non-principal ultrafilter U, the map
ξ2 ◦ ξ1 is a discontinuous group homomorphism from an oligomorphic permutation
group to Z2 . 4
Proof. Part 1: G/H is discrete if each left coset gH is open, which is the case
if and only if H is open.
Part 2: If G/H is Hausdorff, then every coset of H that is distinct from H is
contained in an open set O that does not intersect H. The union of all those open
sets is open, and defines the complement of H in G/H. Hence, H is closed.
Conversely, suppose that H is closed. Then the equivalence relation
[
R := {(x, y) | x−1 y ∈ H} = (gH)2
g∈G
{B ∈ XC | α : A ,→ B is an embedding}.
Note that this topology is compact because it is a closed subset of a product of finite
spaces (TODO: explain more, Proposition 4.1.13, Theorem 4.1.14).
We now define the so-called logic action of Sym(N) on XC : for g ∈ Sym(N)
and B ∈ XC , define g(B) to be the unique structure B 0 in XC such that g is an
isomorphism between B and B 0 . Clearly, this action is continuous. 4
Exercises.
(103) Show that the componentwise action of G ≤ Sym(X) on X n and
the setwise action of G on X
n are continuous.
4.2. TOPOLOGICAL GROUPS 71
Proof. Let > 0. Since ξ is continuous, there exists a δ > 0 such that for all
g ∈ G with d1 (1G , g) < δ we have d2 (1H , ξ(g)) < . Let g1 , g2 ∈ G be such that
d1 (g1 , g2 ) < δ. Then d1 (1G , g1−1 g2 ) < δ, and hence
d2 (ξ(g1 ), ξ(g2 )) = d2 (1H , ξ(g1 )−1 ξ(g2 )) = d2 (1H , ξ(g1−1 g2 )) <
which shows uniform continuity of ξ.
We have seen in Example 52 an example of a left-invariant metric d on Sym(D)
which is not complete.
Lemma 4.2.11. Let G be a topological group with a compatible left-invariant metric
d. Then
d0 (g, h) := d(g, h) + d(g −1 , h−1 )
is a compatible metric, too.
Proof. Clearly, d0 is non-negative, indiscernible, symmetric, and subadditive.
We have to show that d0 induces the same topology on G as d. Let > 0. The set
S := {g ∈ G | d(1, g) < } is open with respect to d and contains the identity 1 ∈ G.
This set contains the set S 0 := {g ∈ G | d0 (1, g) < } which is open with respect to d0
and also contains 1. Conversely, the set S 0 contains the set {g ∈ G | d(1, g) < /2}
(which is open with respect to d and contains 1): if g is such that d(1, g) < /2, the
by the left-invariance of d we have that d(g −1 , 1) < /2, and hence
d0 (1, g) ≤ d(1, g) + d(1, g −1 ) < /2 + /2 < ,
so g ∈ S2 . Since the topology on G is given by a basis of open sets at 1, the statement
follows.
Lemma 4.2.12. Let G be a topological group, let d be a compatible left-invariant
metric, and let d0 be the compatible metric defined in Lemma 4.2.11. Let (gi )i∈N and
(hi )i∈N be Cauchy sequences in (G, d0 ). Then (gi−1 hi )i∈N is Cauchy in (G, d) and in
(G, d0 ).
Proof. Let > 0. Then there exists an n0 ∈ N such that
d0 (hn , hm ) = d0 (h−1 G
m hn , 1 ) < /3 (6)
for all n, m ≥ n0 . By the continuity of the multiplication operation and since d is a
compatible metric there exists a δ > 0 such that for all k ∈ G with d(k, 1G ) < δ
d(h−1 G
n0 khn0 , 1 ) < /3.
Let n1 ≥ n0 be such that d0 (gn , gm ) < δ for all n, m ≥ n1 . Then for all n, m ≥ n1
d(gm gn−1 , 1G ) = d(gn−1 , gm
−1
) ≤ d0 (gn , gm ) < δ
and hence
d(h−1 −1 G
n0 gm gn hn0 , 1 ) < /3. (7)
Therefore
d(gn−1 hn , gm
−1
hm )
= d(h−1 −1 G
m gm gn hn , 1 )
≤ d(h−1 G −1 −1 G −1 G
m hn0 , 1 ) + d(hn0 gm gn hn0 , 1 ) + d(hn0 hn , 1 ) (Lemma 4.2.9)
≤ /3 + /3 + /3 = (by (6), (7)), and (6))
which proves that (gi−1 hi )i∈N
is Cauchy in (G, d). Note that by symmetry also the
sequence (h−1
i gi )i∈N is Cauchy in (G, d), and it follows that both sequences are also
Cauchy in (G, d0 ).
4.2. TOPOLOGICAL GROUPS 73
≤ d(h−1 G −1 −1 0 G −1 0 G
n hn0 , 1 ) + d(hn0 gn gn hn0 , 1 ) + d(hn0 hn , 1 )
≤ /3 + /3 + /3 = .
This shows that [(gn0 h0n )n∈N ] = [(gn hn )n∈N ] and hence the multiplication on G∗ is
indeed well-defined.
The multiplication operation defined on G∗ is associative and has the neutral
element [(1G )n∈N ]. The inverse of [(gn )n∈N ] is [(gn−1 )n∈N ] (Lemma 4.2.12 implies that
(gn−1 )n∈N is Cauchy). We use Proposition 4.1.1 to verify that the multiplication and
taking inverses in G∗ is continuous with respect to the topology induced by d∗ : if
limm→∞ limn→∞ gn,m = g and limm→∞ limn→∞ hn,m = h then
h i
−1 −1
lim [ lim gn,m ] · [ lim hn,m ] = lim lim gn,m · lim hn,m
m→∞ n→∞ n→∞ m→∞ n→∞ n→∞
h i
−1
= lim lim gn,m · lim hn,m
m→∞ n→∞ n→∞
= −1
lim gn,m hn,m = [g −1 · h]
n,m→∞
Proof. Let (G∗ , d∗ ) be the metric completion of (G, d0 ) (see Proposition 4.1.10).
By Lemma 4.2.13, G∗ can be viewed as a topological group G∗ with the compatible
complete metric d∗ , and since G∗ is also separable we conclude that G∗ is Polish. Since
G is Polish, too, Lemma 4.2.14 implies that G is closed in G∗ . Therefore, G = G∗ .
This shows that d0 is a compatible complete metric on G.
Exercises.
(104) Show that Sym(N) does not admit a compatible complete
and left-invariant metric.
(105) Show that a Polish group has a compatible and complete
left-invariant metric if and only if every compatible
left-invariant metric is complete.
(106) Show that no oligomorphic permutation group G on a countably
infinite set has a compatible complete and left-invariant metric.
d(hn , hm ) = d(g1 g2 . . . gn , g1 g2 . . . gm )
= d(1, gn+1 . . . gm ) ≤ 2−an ≤ 2−n
using the left invariance of d and that gn+1 . . . gm ∈ GAn . This shows that
(hn )n∈N is Cauchy with respect to d. But this sequence is not convergent
since the preimage of a1 after n − 1 steps is an which is a non-convergent
sequence.
4.3. CLOSED SUBGROUPS 75
If |B| = n0 is finite, we define the map ξ : G → Sym(N) similarly, but set ξ(g)(n) = n
for all n > n0 .
Claim 2. ξ is injective: when f, g ∈ G are distinct, then there are disjoint open
subsets U and V with f ∈ U and g ∈ V , because the topology is metrisable and
therefore Hausdorff; since B is a basis, we can assume that U = Un for some n ≥ 1.
If f Un = gUn , then g ∈ Un = U since f ∈ Un , contradicting the assumption that U
and V are disjoint. Hence, ξ(f )(n) 6= ξ(g)(n), and so ξ(f ) 6= ξ(g). So ξ is indeed an
isomorphism between G and a subgroup of Sym(N).
This set is a finite intersection of a union of open sets and thus open. Hence, ξ −1 (V )
is a union of open sets and therefore open as well, which concludes the proof that ξ
is continuous.
Claim 4. The map ξ is open and the image of ξ is closed in Sym(N). Let
• d1 be the left-invariant compatible metric on G (Theorem 4.2.8),
• d01 be the compatible complete metric on the Polish group G defined as
d01 (g, h) = d1 (g, h) + d1 (g −1 , h−1 ) (see Lemma 4.2.15), and
• d02 be the compatible complete metric on Sym(N) from Example 54.
We will show that ξ −1 is Cauchy-continuous as a map from (ξ(G), d02 ) to (G, d01 ). This
clearly implies both parts of the claim.
Let g1 , g2 , . . . be a sequence in G such that ξ(g1 ), ξ(g2 ), . . . converges against
h ∈ Sym(N). We have to show that g1 , g2 , . . . is d01 -Cauchy. Since d1 is left-invariant,
−1
limn,m→∞ d1 (gn , gm ) = 0 if and only if limn,m→∞ d1 (gm gn , 1) = 0. Let > 0 be
arbitrary. Since B is a basis, there exists Uk ∈ B such that
and Uk Uk−1 ⊆ {g ∈ G | d1 (g, 1) < }. Since limn→∞ ξ(gn ) = h, there exists an n0
such that ξ(gn )(k) = ξ(gm )(k) = h(k) for all n, m > n0 . Then gn Uk = gm Uk , and so
−1
gm gn ∈ Uk Uk−1 ⊆ {g ∈ G | d1 (g, 1) < } .
Exercises.
(107) Let A be a countable ω-categorical structure. Let B ⊆ A be finite and let
C := aclA (B). Then Aut(A){C} is open in Aut(A).
Subgroups of Sym(N)
compact
Proof. The forward implication follows from Proposition 4.5.1. For the other
direction, suppose that the orbits O1 , O2 , . . . of G are finite. We write G|Oi for
the permutation
Q group formed by the restrictions of G to the finite set Oi . Then
G = i G|Oi is a closed subset of a product of finite subgroups of G, and hence
compact by Tychonoff’s theorem (Theorem 4.1.14) and Proposition 4.1.13. We do
not use the entire strength of Tychonoff’s theorem, and show two alternative proofs.
Second Proof. Let {Ui }i∈A be an open cover of G. Since Sym(N) and hence
G are second-countable, we can assume that A = N (Proposition S 4.1.17). Suppose
for contradiction that for no finite B ⊆ A we have that G ⊆ i∈B Ui . Consider the
following rooted tree. The vertices on level n are the restrictions of the permutations
in G to {1, . . . , n}. Adjacency between a vertex on level n and vertices on level n + 1
is defined by restriction. Clearly, for every n there are finitely many vertices on level
n since the orbit of (1, . . . , n) with respect to the componentwise action of GSis finite.
A vertex on level n is good if it is the restriction of a function from G \ i≤n Un .
Clearly, the restriction of a good vertex is good. Moreover, by assumption there are
good vertices on all levels. By König’s tree lemma, there is an infinite branch of good
vertices in the tree. This branch defines an injection from N to N. In fact, it must
be a bijection since the finiteness of the orbits implies that the map is surjective onto
each orbit. This map is from G since G is closed, but it does not lie in any of the Ui ,
a contradiction.
Third proof. Our final proof uses Theorem 4.1.19. Clearly G is complete since
it is closed in Sym(N). To prove total boundedness of G, let > 0. Choose n ∈ N
such that 1/2n < . Since all orbits of G are finite, the orbit O of (1, . . . , n) with
respect to the componentwise action of G on Nn is finite too; choose f1 , . . . , fk ∈ G
so that O = {f1 (1, . . . , n), . . . , fk (1, . . . , n)}. Then by construction Bf1 (), . . . , Bfk ()
covers all of G.
It follows from Proposition 4.5.1 that a compact subgroup G of Sym(N) must have
infinitely many orbits, and in particular cannot be oligomorphic. In fact, oligomor-
phicity is already ruled out by local compactness; this can be seen from Lemma 3.0.1
and the following.
80 4. TOPOLOGICAL GROUPS
Also note that locally compact subgroups of Sym(N) contain all closed countable
subgroups of Sym(N) by Theorem 1.2.6; see Figure 4.1.
Exercises.
(108) Prove that every countable compact subgroup of Sym(N) is finite.
(109) Prove that every closed subgroup of Sym(N) which is not locally compact has
a homeomorphism ξ to Sym(N) such that ξ and ξ −1 are uniformly continuous
(with respect to the metric inherited from the Baire space).
Exercises.
(110) Let G ≤ Sym(N) be oligomorphic, and let G◦ be the intersection
of the closed subgroups of ∆ of finite index in G. Show that G◦ is
oligomorphic and that (G◦ )◦ = G◦ .
(111) Can the group Aut(A) from Example 67 be written
as a semidirect product N oθ ZN 2 ? Here, N is the normal
subgroup of Aut(A) consisting of all automorphisms that fix
all the equivalence classes of all the equivalence relations
E0 , E1 , . . . (see Proposition 4.6.1 and Proposition 1.5.4).
CHAPTER 5
Hence,
tB (b) = tB (ci (a)) = tS (ca )i
= rS (ca0 )i = rB (ci (a0 ))) = rB (b).
83
84 5. BIRKHOFF’S THEOREM AND PERMUTATION GROUPS
Corollary 5.1.2. Let G ≤ Sym(B) and H ≤ Sym(C), with |B|, |C| ≥ 2. Then
the following are equivalent:
(1) G and H are isomorphic as abstract groups;
(2) There exists a G-set B and an H-set C such that
HSP(B) = HSP(C).
5.2. TOPOLOGICAL BIRKHOFF 85
Choose F 0 to be smallest possible; note that this implies that F 0 contains at most
F
one element from each orbit of G. Let C be (F 0 ) and let m := |C| = |F 0 |k . Let
c1 , . . . , cm be the elements of C, and for j ≤ k define cj = (c1 (aj ), . . . , cm (aj )). Let S
be the substructure of B m generated by c1 , . . . , ck ; so the elements of S are precisely
those of the form tS (cj ) for a τ -term t and j ≤ k. Define a function µ : S → A by
setting
µ(tS (cj )) := ξ(tB )(aj ).
Claim 1. µ is well-defined. Suppose that tS (cj ) = rS (cl ) for j, l ≤ k. We first
show that tB |F 0 = rB |F 0 . Let b ∈ F 0 . Note that there is some i ≤ m such that
ci (aj ) = b and ci (al ) = b. Hence,
It also follows from tS (cj ) = rS (cl ) that for all i ≤ m the elements (cj )i and (cl )i lie
in the same orbit of G (here we use the assumption that |B| ≥ 2). By our assumption
on F 0 this means that l = j. Hence,
We also mention that Theorem 5.4.1 in combination with Proposition 3.6.5 shows
that for every ω-categorical structure B, having essentially infinite signature only
depends on the automorphism group of B viewed as a topological group.
Example 76. Let B be the graph with domain Nk and the edge relation
E B = (S, T ) | |S ∩ T | = 1 .
We claim that B and (N; =) are bi-interpretable. Let I be the k-dimensional inter-
pretation of B in (N; =}) whose domain is Nk6= and which maps (x1 , . . . , xk ) ∈ Nk6= to
{x1 , . . . , xk }. Clearly, the relations
I −1 (B) = Nk6=
I −1 (=B ) = (x, y) | x, y ∈ Nk6= and {x1 , . . . , xn } = {y1 , . . . , yn }
are definable in B.
Finally, note that for all (x1 , . . . , xk ), (y1 , . . . , yk ) ∈ Nk6= and z ∈ N we have
J I(x1 , . . . , xk ), I(y1 , . . . , yk ) = z
if and only if z = {x1 , . . . , xk } ∩ {y1 , . . . , yk }, which is definable in (N; =).
Moreover, for all (S1 , T1 ), . . . , (Sk , Tk ) ∈ E B and V ∈ B we have
I J(S1 , T1 ), . . . , J(Sk , Tk ) = V
B
if and only if there are U1 , . . . , Uk such that (Si , Ti , Ui , V ) ∈ R4 for all i ∈ [k] and
/ E B for all {i, j} ∈ [k]
(Ui , Uj ) ∈ 2 , and hence is definable in B. This concludes the
proof of the claim above.
Then Theorem 5.4.1 implies that there exists an injective (continuous)
homomor-
N
phism from Aut(N; =) to Aut(B). Thus, the action of Sω on k is faithful, as claimed
90 5. BIRKHOFF’S THEOREM AND PERMUTATION GROUPS
earlier in Example 11. The same argument works for any permutation group instead
of Sω . 4
Exercises.
(112) Show that the line graph of an undirected graph G
has an interpretation in G.
(113) Show that every finite structure has an interpretation
in every structure with at least two elements.
(114) Show that the automorphism group of infinitely many disjoint
copies of the 2-element clique K2 is not topologically
isomorphic to infinitely many disjoint copies of the
three-element clique K3 .
CHAPTER 6
The question what information about a structure can be recovered from its au-
tomorphism group when considered as an abstract group, has long attracted the
attention of research in model theory and the theory of infinite permutation groups;
a very incomplete list of references is [6, 47, 50, 52, 54, 60, 69, 74, 88, 104, 130, 138,
139, 158, 162].
Reconstruction
⇐ Automatic Homeomorphicity
⇐ Automatic Continuity (Proposition 6.3.18)
⇔ The Small Index Property (Proposition 6.2.1)
⇐ Ample Generics (Theorem 6.4.4)
⇐ EPPA / the Hrushovski property (Section 6.5)
For many automorphism groups (e.g., for the automorphism group of the Rado graph),
this chain of implications is to the best of my knowledge the only known way to prove
reconstruction.
countable intersection
of dense open
dense open
Exercises.
(115) Let S be a countable Polish space.
Is S meager or non-meager?
(116) Show that R \ Q is non-meager.
(117) Show that in non-empty Baire spaces
comeager sets are non-meager.3
(118) Prove Lemma 6.3.6.
(119) Let S1 , S2 , . . . be comeager subsets of Sω .
Show that ∩i∈N Si is non-empty.
(120) Show that a dense Gδ set (Definition 4.1.5) is comeager.
since the Cantor space is Baire; Example 79). For the same reason, U cannot be
comeager. Note that the symmetric difference of a set in the non-principal ultrafilter
U with a finite set must again be in U. Then Theorem 8.46 in [88] implies that U
cannot have the Baire property. 4
Let U, A ⊆ X. We say that A is meager in U if A ∩ U is meager in the subspace
U of X, and otherwise that A is non-meager in U . If U = A we then also say that
A is non-meager in its relative topology. Note that if U is open, then A is meager in
U if and only if A ∩ U is meager in X. The assumption that U is open is necessary,
as demonstrated by R ⊆ C, because R is non-meager in its relative topology, but R
is nowhere dense in C. Similarly, we say that A is comeager in U if U \ A is meager
in X.
Proposition 6.3.9 (‘Localisation’). Let X be a topological space and suppose that
A ⊆ X has the Baire property. Then A is meager or there is a non-empty open set
U ⊆ X such that A is comeager in U . If X is a Baire space, then the two alternatives
are mutually exclusive.
Proof. Let U ∆A = M where U is open and M is meager. If U is empty then
A = M is meager. So suppose that U is non-empty. Since U \ A ⊆ M is meager in
X, it is also meager in U , and hence A is comeager in U .
For the second statement, suppose that X is a Baire space, that A ⊆ X is meager,
and that U ⊆ X is open such that A is comeager in U . Then C := U \ A is meager in
X, and hence U ⊆ A ∪ C is meager. By item (3) of Proposition 6.3.7 we can therefore
conclude that U = ∅. This shows that the second alternative does not apply.
It is sometimes convenient to use the following logical notation:
∀∗ x.A(x)
stands for A is comeager, and is pronounced ‘A holds for comeagerly many x’. Simi-
larly,
∃∗ x.A(x)
stands for A is non-meager, and pronounced ‘A holds for non-meagerly many x’. The
notation has the obvious extension to localised expressions of the form
∀∗ x ∈ U.A(x) and ∃∗ x ∈ U.A(x)
The following result is sometimes called the Fubini theorem for category because
of the analogy to Fubini’s theorem about the order of integration in double integrals
in analysis.
Theorem 6.3.10 (Kuratowski-Ulam). Let X, Y be second-countable and suppose
that A ⊆ X × Y has the Baire property. Then
∃∗ (x, y).A(x, y) ⇔ ∀∗ x∃∗ y.A(x, y) ⇔ ∀∗ y∃∗ x.A(x, y) (13)
∗ ∗ ∗ ∗ ∗
and ∀ (x, y).A(x, y) ⇔ ∀ x∀ y.A(x, y) ⇔ ∀ y∀ x.A(x, y) (14)
Proof. Clearly, (13) and (14) are equivalent by taking complements. Moreover,
the two equivalences in (13) are symmetric, so we only show the first equivalence.
If S ⊆ X × Y and x ∈ X, then Sx denotes the set {y ∈ Y | (x, y) ∈ S}. To show
the forward implication in (13), we first prove that if a set B ⊆ X × Y is nowhere
dense, then ∀∗ x(Bx is nowhere dense in Y ). We may assume that B is closed, since
replacing B by its closure can only increase the sets Bx ; hence, if we can prove that
for comeagerly many x ∈ X some supersets of the Bx are nowhere dense in Y , then
the same holds for the sets Bx . Let U := (X × Y ) \ B. It suffices to show that
∀∗ x(Ux is dense), because if Ux is dense then B x = Bx = X \ Ux does not contain a
6.3. CONSISTENCY OF AUTOMATIC CONTINUITY 97
non-empty open set and hence ∀∗ (Bx is nowhere dense). Let {V1 , V2 , . . . } be a basis
of non-empty open sets for Y . Then
Un := {x ∈ X | ∃y ∈ Vn .(x, y) ∈ U }
is the projection of an open set and hence open. Moreover, Un is dense in X: if
G ⊆ X is nonempty open, then U ∩ (G × Vn ) 6= ∅, because otherwise B would
T contain
the nonempty open set G × Vn , contrary to being nowhere dense. If x ∈ n∈N Un ,
T for every n ∈ N we have Ux ∩ Vn 6= ∅, i.e., Ux is dense. We are done because
then
n∈N Un is comeager by Corollary 6.3.6.
S
Let A be meager. Then A = n∈N Bn where Bn is nowhere dense. For each
n ∈ N, let Xn be the set of all x ∈ X such that (Bn )x is nowhere dense;
T by what
we have seen above, Xn is comeager. By Corollary 6.3.6, we have that n∈N Xn is
comeager. It follows that
∀∗ x ∀n ∈ N (Bn )x is nowhere dense
Remark 6.3.11. Note that the assumption that A has the Baire property was only
used in the proof of the converse implication in (13); the forward implication holds in
general.
The following is beyond the scope of this text, but we need it in the important
Proposition 6.3.18 below.
Theorem 6.3.12 (Lusin-Sierpiński; see Kechris [88] (21.6)). Let X be Polish, let
Y be a metric space, and let f : X → Y be continuous. Then the image of every open
set in X under f has the Baire property in Y .
The axiom of dependent choices (DC) is a weak form of the axiom of choice
(AC) that is still sufficient to develop most of real analysis (see Appendix A.2). Over
Zermelo-Fraenkel set theory (ZF), the Axiom of Dependent Choices is equivalent to
the version of the Baire Category Theorem (Theorem 4.1.8) where we only require
that S is completely metrisable (instead of Polish). The relevance of the following for
reconstruction has been pointed out by Lascar [104, Theorem 2.7].
98 6. RECONSTRUCTION OF TOPOLOGY AND AUTOMATIC CONTINUITY
The following lemma is sometimes called Pettis’ lemma; it is Lemma 2.6 in [104]
and a consequence of Theorem 2.3.2 in [60] or Lemma 9.9 in [88].
Lemma 6.3.16. Let G be a closed subgroup of Sym(N) and let H be a subgroup
of G. Then H is meager, open, or for every non-empty open U ⊆ G the set U \ H is
non-meager (in particular, G\H is dense). Consequently, if H has the Baire property
then H is either meager or open.
Proof. The second statement clearly follows from the first: if U has the Baire
Property then there exists an open U ⊆ Sym(N) such that H∆U is meager, and in
particular U \ H is meager, so either U is empty and H is meager, or H is open by
the first statement.
To show the first statement, let H be non-meager and suppose that there exists
a non-empty open U ⊆ G such that U \ H is meager. Then U contains f Ga for some
f ∈ G, a ∈ Nn , and n ∈ N.
Claim. The subgroup H ∩ Ga of Ga is comeager in Ga . The set f Ga is non-
meager. If f Ga ⊆ U \H, then U \H would not be meager, contrary to our assumptions.
Hence, there exists exists an h ∈ H ∩ f Ga . Since f Ga \ H ⊆ U \ H is meager, we
have that h−1 f Ga \ h−1 H is meager. But h−1 H = H and h−1 f Ga = Ga and hence
Ga \ H is meager, which proves the claim.
The claim implies that all cosets of H ∩Ga in Ga are comeager. Since intersections
of comeager sets are comeager (Corollary 6.3.6) and in particular non-empty, there can
be only one coset. Therefore, H contains Ga and hence is open by Lemma 4.4.1.
Exercises.
(121) Prove the statement in Remark 6.3.19.
and it follows that C is comeager. For the forward implication, recall that there is
at most one comeager orbit, which shows that the elements of C are the only generic
elements of Sω . 4
100 6. RECONSTRUCTION OF TOPOLOGY AND AUTOMATIC CONTINUITY
X n+1 , for any z ∈ G we have (x̄, z) ∈ C if and only if there exists g ∈ G such that
g · (x̄, z) = (x̄, y), i.e., g ∈ Gx̄ such that g · z = y. It follows that for any y ∈ Cx̄ the
set Gx̄ · y is comeager. Fix a ∈ A ∩ Cx̄ .
By Effros’ theorem (Theorem 6.4.2, (2)), the set (Gx̄ ∩ V ) · a is open in Gx̄ · a.
By assumption, A is non-meager in (Gx̄ ∩ V ) · a. Hence, (Gx̄ ∩ V ) · a ∩ B 6= ∅. Fix
b ∈ (Gx̄ ∩ V ) · a ∩ B. Then for some v ∈ Gx̄ ∩ V we have v · a = b.
Theorem 6.4.4 (Theorem 5.3 in [74], Theorem 6.9 in [92]). Let G be a closed
subgroup of Sω with ample generics. Then G has the small index property.
Proof. Suppose that H ≤ G has countable index. Then H is non-meager by
Lemma 6.3.15. If there exists a non-empty open set U such that U \ H is meager,
then H is open by Lemma 6.3.16.
Otherwise, G \ H is non-meager in every non-empty open set. In this case we will
reach a contradiction as follows. We will apply Lemma 6.4.3 to A = H and B = G\H
to construct for every a ∈ 2N an element ha ∈ G such that for all a, b ∈ 2N , if a 6= b
then ha and hb lie in different cosets of H in G, contradicting the assumption that H
has countable index.
Let a ∈ 2N . For every n ∈ N and s = (s0 , s1 , . . . , sn ) ∈ {0, 1}n+1 we inductively
define xs , fs , hs ∈ G such that
(1) xs is generic,
(2) xs ∈ H if sn = an and xs ∈ G \ H if sn = 1 − an ,
(3) fs = 1G if sn = 1 − an ,
(4) hs = f(s0 ) ◦ f(s0 ,s1 ) ◦ · · · ◦ f(s0 ,...,sn ) ,
(5) d(hs , hs f(s,1−an ) ) < 2−n (we use any compatible complete metric d on G,
e.g., the metric d0 from Lemma 4.2.15),
(6) if sn = (s0 , an ) for s0 ∈ Nn , then fs x(s0 ,an ) fs−1
0 = x(s0 ,1−an ) .
For n = 1, and if a0 = 0, we apply Lemma 6.4.3 for A := H, B := G \ H, and V = G
and set x0 := a, x1 := b, f1 := 1G , and f1 := v. If a0 = 1, we set x0 := b, x1 = a,
f1 := v −1 , and f0 := 1G .
For n > 1, suppose that xs and fs , for s ∈ {0, 1}n , are already defined. First
consider the case that an = 0. Again by Lemma 6.4.3, this time applied to V :=
{f ∈ G | d(hs , hs f(s,1) ) < 2−n }, there are x(s,0) ∈ H and x(s,1) ∈ G \ H such that
(xs , s(s,0) ) and (xs , s(s,1) ) are generic, and f(s,0) · x(s,0) = x(s,1) . The inductive step
for an = 1 is analogous. By (3) and (4), the sequence (hsn )n∈N is Cauchy, and by the
completeness of d converges to some ha ∈ G.
Claim 1. a 7→ ha is a continuous map from 2N to G.
Claim 2. If a, b ∈ 2N are such that for some n ∈ N we have a1 = b1 , . . . , an = bn ,
an+1 = 0, and bn+1 = 1, then ha · H ∩ hb · (G \ H) 6= ∅. Indeed, we have
ha · x(a1 ,...,an ,0) = h(a0 ,...,an ,0) · x(a1 ,...,an ,0)
= h(a0 ,...,an ) f(a0 ,...,an ,1) · x(a1 ,...,an ,0)
= h(a0 ,...,an ) · x(a1 ,...,an ,0) (as f(a0 ,...,an ,1) = 1G )
and
hb · x(a1 ,...,an ,1) = h(a0 ,...,an ,1) · x(a1 ,...,an ,1)
= h(a0 ,...,an ) f(a0 ,...,an ,0) · x(a1 ,...,an ,1)
= h(a0 ,...,an ) · x(a1 ,...,an ,0) .
So ha · x(a1 ,...,an ,0) = hb · x(a1 ,...,an ,1) which proves the claim since x(a1 ,...,an ,0) ∈ H
and x(a1 ,...,an ,1) ∈ G \ H.
102 6. RECONSTRUCTION OF TOPOLOGY AND AUTOMATIC CONTINUITY
and hence
h−1 −1
b ha Hha hb ∩ G \ H 6= ∅,
so h−1 ha ∈
/ H. This shows that ha and hb are in different cosets of H.
Claim 3 contradicts the assumption that H has countable index in G.
Exercises.
(122) Discuss whether and why the adjective generic in the name of
the concept of a generic superposition (Section 3.4.2) is appropriate.
6.4.1. Dense conjugacy classes and the JEP. We have seen in the previous
section that if G ≤ Sω has ample generics, then it has the small index property: but
how do we prove that G has ample generics? To this end, we present in this section and
the following sections an elegant characterisation of the existence of ample generics
of Kechris and Rosendal [92] which builds on ideas from [74] and [159].
If G has a generic element α, then the orbit of α is in particular dense (since G
is Polish and by Proposition 6.3.7). So we will first focus on understanding whether
G has a dense conjugacy class.
Definition 6.4.5. Let K be an amalgamation class. Then Kp denotes the class
of all pairs
(A, e : B → C)
such that A ∈ K and e is an isomorphism between substructures B and C of A.
An embedding of (A, e : B → C) ∈ Kp into (A0 , e0 : B 0 → C 0 ) ∈ Kp is an embed-
ding f : A ,→ A0 such that f (B) ⊆ B 0 , f (C) ⊆ C 0 , and f ◦ e = e0 ◦ (f |B ). Note that
the definition of the joint embedding property (JEP) and the amalgamation prop-
erty (AP) were purely categorical in the sense that their definition only requires a
notion of embedding; so JEP and AP are naturally defined not only for structures
and embeddings, but also for classes of the form Kp as introduced above.
We say that a continuous action ξ : G → Sym(X), for some set X, is topologically
transitive if for any two non-empty open subsets U, V ⊆ X there exists g ∈ G such
that g(U ) ∩ V 6= ∅. The implication from (1) to (2) in the following theorem can
already be found in [158]; the equivalence of (1) and (2) is Theorem 2.1 in [92].
Theorem 6.4.6. Let K be an amalgamation class and let L be its Fraı̈ssé limit.
Then the following are equivalent.
(1) G := Aut(L) has a dense conjugacy class.
(2) Kp has the JEP.
(3) The action of G on G by conjugation is topologically transitive.
6.4. AMPLE GENERICS 103
is clearly open, and dense since the conjugation action is topologically transitive.
Since G is a Baire space, the countable intersection C over all the Da,b is dense, and
in particular non-empty. Let f ∈ C; then the conjugacy class of f is dense, because
for every a, b ∈ Ln , n ∈ N, we have that f ∈ Da,b , and hence there is g ∈ G such that
g · f ∈ Sa,b .
Corollary 6.4.7. Sω , the automorphism group of the random graph, and more
generally the automorphism group of Fraı̈ssé-limits of classes with free amalgamation
have a dense conjugacy class.
Proof. Let K be the class of all finite structures with the empty signature so
that the automorphism group of the Fraı̈ssé-limit of K is isomorphic (as a permutation
group) to Sym(N). By Theorem 6.4.6, it suffices to verify that Kp has the JEP. So
let (Ai , ei : Bi → Ci ) ∈ Kp for i ∈ {1, 2}. We may assume that A1 ∩ A2 = ∅ and
define A := A1 ∪ A2 , B := B1 ∪ B2 , C := C1 ∪ C2 , and e : B → C as the common
extension of both e1 and e2 . Then the identity map fi : Ai → A is an embedding of
(Ai , e) into (A, e) showing the JEP. The same proof works for the other groups in the
statement.
Exercises.
(126) Let E 2 be the equivalence relation on N with two infinite classes,
and let K := Age(N; E 2 ). Show that Kp does not have the JEP.
(127) Directly show that Aut(N; E 2 ) does not have a dense conjugacy
class (without using Theorem 6.4.6).
(128) Show that Aut(Q; Cycl) (see Section 2.4) does not have a dense
conjugacy class.
(129) Let V be the countably infinite vector space over F2 .
Show that Aut(V ) has a dense conjugacy class.
104 6. RECONSTRUCTION OF TOPOLOGY AND AUTOMATIC CONTINUITY
Conversely, suppose that Kp has the JEP and the WAP. We will construct an
element α ∈ G such that G · α is dense in G and non-meager in its closure.
Let e1 , e2 , . . . be an enumeration of all isomorphisms e between finite substruc-
tures of L such that any two fi : (L[dom(e) ∪ im(e)], e) ,→ (B, g) can be amalgamated.
Claim. For every open subgroup V ≤ G TODO.
By Proposition 6.4.8 (4) ⇒ (1) we have that G · α is non-meager in G · α. Theo-
rem 6.4.2 (1) ⇒ (5) now implies that G · α is comeager in G · α = G.
6.4.3. WAP and Ample Generics. This section finally presents the charac-
terisation of those homogeneous structures L whose automorphism group has ample
generics. Let L be the Fraı̈ssé-limit of the amalgamation class K. We introduce the
class Kpn for n ≥ 1, which consists of tuples (A, e1 : B1 → C1 , . . . , en : Bn → Cn ) where
A ∈ K and ei an isomorphism between substructures B i and C i of A. Embeddings
between elements of Kpn are defined analogously as embeddings between elements of
Kp , and again properties like the JEP and the AP make sense.
Theorem 6.4.10 (Theorem 6.2 in [92]). Let K be an amalgamation class and let
L be its Fraı̈ssé limit. Then the following are equivalent.
• Aut(L) has an n-generic element.
• The class Kpn has the JEP and the WAP.
Proof. The proof is similar to the proof of Theorem 6.4.9.
Lemma 6.4.11. Sym(N) has ample generics.
Proof. Let K be the class of all finite structures over the empty signature. We
have already seen in Corollary 6.4.7 that Kp has the JEP; the proof that Kpn has the
JEP is analogous. Moreover, we have already seen in Example 88 that Kp has the
WAP; the proof that Kpn has the WAP is analogous: again we may find for every
(A, e) ∈ Kpn an extension (A0 , e0 ) ∈ Kpn which is determined on (A, e) by choosing e0
to be an automorphism of A0 . Now the statement follows from Theorem 6.4.10.
Corollary 6.4.12. Sym(N) has the small index property, automatic continuity,
and automatic homeomorphicity.
Proof. We have just seen that Sym(N) has ample generics, so it follows from
Theorem 6.4.4 that Sym(N) has the small index property. Automatic continuity
follows from Proposition 6.2.1, and automatic homeomorphicity follows from Propo-
sition 6.3.18.
Corollary 6.4.13. Aut(Q; <) does not have a 2-generic element.
Proof. Let K := Age(Q; <). By Theorem 6.4.10 it suffices to show that Kp2 does
not have the WAP. Let a1 , a2 ∈ Q be such that a1 < a2 , let A := (Q; <)[{a1 , a2 }],
let e1 (a1 ) = a2 and e2 (a1 ) = a2 . We claim that there is no (A0 , e01 , e02 ) ∈ Kp2 which
is determined on (A, e1 , e2 ). Note that in A0 , for i ∈ {1, 2} we have a1 < e0i (a1 ),
and if (e0i )k (a1 ) is defined for k ≥ 1, then (e0i )k−1 (a1 ) < (e0i )k (a1 ). Let k ≥ 1 be
smallest so that (e01 )k (a1 ) is undefined, and let b := (e01 )k−1 (a1 ). Let B 1 be the
extension of A0 by one new element c which is larger than all elements in A0 , and
let e1,1 be the extension of e01 given by e1,1 (b) = c, and let e1,2 be the extension
of e02 given by e1,2 (c) = c. Let B 2 be the extension of A0 by two new elements d1
and d2 such that d1 is larger than all elements in A0 and d2 is larger than d1 . Let
e2,1 be the extension of e01 given by e2,1 (b) = d1 and let e2,2 be the extension of e02
given by e2,2 (d1 ) = d2 . Note that (B 1 , e1,1 , e1,2 ), (B 2 , e2,1 , e2,2 ) ∈ Kp2 . Now suppose
for contradiction that there exists (C, e001 , e002 ) and embeddings fi : (B i , ei,1 , ei,2 ) ,→
6.5. THE EXTENSION PROPERTY FOR PARTIAL AUTOMORPHISMS 107
(C, e001 , e002 ) such that f1 |A = f2 |A . Then in particular f1 (a1 ) = f2 (a1 ), and hence
f1 (b) = f1 ((e1,1 )k (a1 )) = f2 ((e2,1 )k (a1 )) = f2 (b), and f1 (e1,1 (b)) = f2 (e2,1 (b)). How-
ever, e1,2 (e1,1 (b)) = e1,1 (b) = c, whereas e2,1 (b) = d1 < d2 = e2,2 (e2,1 (b)). Hence,
f1 (e1,1 (b)) = f2 (e2,1 (b)) < f2 (e2,2 (e2,1 (b)) = f1 (e1,2 (e1,1 (b)) = f1 (e1,1 (b))
a contradiction.
Exercises.
(130) Let K be the class of all finite partial orders. Show that the automorphism
group of the Fraı̈ssé-limit of K does not have ample generics.
and we may amalgamate embeddings from (D, e01 , . . . , e0n ) into other structures of Knp ,
which shows that Kpn has the WAP.
Corollary 6.5.4. The automorphism group of the random graph has ample
generics and the small index property.
Proof. Combine Theorem 6.5.2 and Theorem 6.5.3 to obtain ample generics,
and Theorem 6.4.4 to obtain the SIP.
Herwig [68] showed EPPA for the class of all finite τ -structures, for any finite
relational signature τ (also see [70] and [81] for other proofs), properly generalising
Corollary 6.5.4. Herwig [69] showed EPPA for the class of all Kn -free graphs. More
generally, Lascar and Herwig [70, Theorem 3.2] showed EPPA for all classes that
are described by homomorphically forbidding finitely many structures, i.e., classes C
such that there exists a finite set of structures F such that A ∈ C if no structure in
F admits a homomorphism to A. Hodkinson and Otto [75] (also see Corollary 4.6
in [145]) obtained the following, also properly generalising Corollary 6.5.4.
Theorem 6.5.5. Let B be a homogeneous structure with finite relational signa-
ture whose age C has the free amalgamation property. Then C has the EPPA. Conse-
quently, Aut(B) has ample generics, the small index property, automatic continuity
and homeomorphicity, and reconstruction.
Exercises.
(131) Let B be a homogeneous structure with finite relational signature. Show
that the age of B has the EPPA if and only if Aut(B) is the closure of the
union of a countable chain G1 ≤ G2 ≤ · · · of compact subgroups of Aut(B).
(132) Let A be a finite substructure of a τ -structure B. We write A ≤homog B if
for every n ∈ N and any two a, b ∈ An we have that a and b lie in the same
orbit of (the componentwise action of) Aut(B) on B n if and only if they lie
in the same orbit of (the componentwise action of) Aut(A) on An .
(a) Prove that if A ≤homog B and B ≤homog C, then A ≤homog C.
(b) Provide a counterexample to the transitivity of ≤ from
the previous exercise if we replace the action of Aut(A)
on An by the action of Aut(B)A on B n .
Exercises.
(133) Let A be a countable ω-categorical structure. Show that A has no algebraic-
ity and weak elimination of imaginaries if and only if Aut(A) has no fixed
points and for all finite B1 , B2 ⊆ A we have that hGB1 ∪ GB2 i = GB1 ∩B2 .
(134) If G is an oligomorphic permutation group on a set B, then let’s write G∗ for
the closed normal subgroup of G consisting of all α ∈ G that fix all blocks
of congruences of the action of G on B n , for some n ∈ N, that have finitely
many classes. Show that an oligomorphic permutation group on a set B is
G-finite if and only if for every finite A ⊆ B, the index of (GA )∗ in GA is
finite.
(135) Show that there are uncountably many closed oligomorphic
permutation groups, up to isomorphism of abstract groups.
Figure 6.2. Some open problems in the context of this chapter; see
Section 6.7. If there is no reference then the result is trivial or can
be deduced from other entries in the table and/or the results of this
chapter.
(3) Does the automorphism group of the countable universal homogeneous poset
(P; ≤) have the small index property (see [110])?
(4) Does the automorphism group of the countable universal homogeneous per-
mutation (Q; <1 , <2 ) have the small index property (see [110])?
(5) Is there for every n ∈ N a countably infinite structure which has n-generic
automorphisms, but not (n + 1)-generic automorphisms (see [146])? We
only know that the answer is positive for n = 1 (Example 87).
(6) Is the automorphism group of every homogeneous structure with a finite
relational language G-finite (Macpherson [110])?
CHAPTER 7
This section is under construction. [53, 78, 79, 89, 90, 102, 119, 121–126, 148,
149, 153–155, 163, 164]
An active field of research studies the question for which Polish groups G the
universal minimal flow M (G) is metrizable. If M (G) is metrizable then it has a
comeagre orbit (Ben Yaacov, Melleray, Tsankov 2017): TODO.
Exercises.
(139) Show that if G is an oligomorphic permutation group on a set A,
then the closure of G in AA contains some non-surjective maps
(this is related to Exercise 106).
Remark 7.4.1. The concept of canonical functions has turned out useful in nu-
merous applications: for classifying first-order reducts they are used in [1, 2, 19,
28, 107, 128, 133], for complexity classification for constraint satisfaction problems
(CSPs) in [20, 21, 25, 31, 96], for decidability of meta-problems in the context of the
CSPs in [30], for lifting algorithmic results from finite-domain CSPs to CSPs over
infinite domains in [22], for lifting algorithmic results from finite-domain CSPs to
homomorphism problems from definable infinite structures to finite structures [94],
and for decidability questions in computations with atoms in [95].
As indicated above, the technique is available for a function f : A → B whenever A
is a Ramsey structure and B is ω-categorical, and the existence of canonical functions
in the set
{β f α | α ∈ Aut(A), β ∈ Aut(B)} ⊆ B A
was originally shown under these conditions by a combinatorial argument [23,24,30].
It is natural to ask for a perhaps more elegant proof of the existence of canonical func-
tions via topological dynamics, reminiscent of the numerous proofs of combinatorial
statements obtained in this fashion (cf. the survey [11] for ergodic Ramsey theory; [91]
mentions some applications of extreme amenability). In this section we present such
a proof, taken from [27].
7.4.1. Canonicity. Let G ≤ Sym(A) and H ≤ Sym(B). A function f : A → B
is called canonical with respect to (G, H) if for every k ≥ 1, t ∈ Ak , and α ∈ G
there exists β ∈ H such that f α(t) = β f (t). Hence, functions that are canonical
with respect to (G, H) induce for each integer k ≥ 1 a function from the orbits of
the componentwise action of G of Ak to the orbits of the componentwise action of
H on B k . For oligomorphic permutation groups we have the following equivalent
characterisations of canonicity.
Proposition 7.4.2. Let G y A and H y B be permutation groups, where H y
B is oligomorphic. Then for any function f : A → B the following are equivalent.
(1) f is canonical with respect to (G, H);
(2) for every α ∈ G we have f α ∈ Hf := {βf | β ∈ H};
(3) for every α ∈ G there are e1 , e2 ∈ H such that e1 f α = e2 f .
A stronger condition is to require that for all α ∈ G there is an e ∈ H such that
f α = ef . To illustrate that this is strictly stronger, already if G = H, we give an
explicit example.
Example 94 (Trung Van Pham). Let G := Aut(Q; <). Note that (Q; <) and
(Q \ {0}; <) are isomorphic, and let f be such an isomorphism. Then f , viewed as
a function from Q → Q, is clearly canonical with respect to (G, G). But f does not
satisfy the stronger condition above: there is no e ∈ G such that f α = ef . To see
this, choose b, c ∈ Q such that f (b) < 0 < f (c). By transitivity there exists an α ∈ G
such that α(b) = c. Note that 0 < f α(b) < f α(c). Morever, the image of f α equals
the image of f , and hence any e ∈ G such that f α = ef must fix 0. Since e must also
preserve <, it cannot map f (b) < 0 to f α(b) > 0. Hence, there is no e ∈ G such that
f α = ef . 4
In Proposition 7.4.2, the implications from (1) to (2) and from (3) to (1) follow
straightforwardly from the definitions. For the implication from (2) to (3) we need
a lift lemma, which is in essence from [29]. This lemma has been applied frequently
lately [8, 20, 22], in various slightly different forms.
Let H y B be a permutation group, and let f, g ∈ B A , for some A. We say that
f = g holds locally modulo H if for all finite F ⊆ A there exist β1 , β2 ∈ H such that
7.5. MODEL-COMPLETE CORES OF RAMSEY STRUCTURES 119
7.5.1. Model Complete Cores. The results from this section are from [14];
we follow the presentation in [16]. An ω-categorical structure C is called
• model complete if every embedding from C into C preserves all first-order
formulas.
• a core if every endomorphism of C is an embedding.
Proposition 7.5.1 (see [16]). Let C be an ω-categorical structure. Then the
following are equivalent.
(1) C is a model-complete core.
(2) Every endomorphism of C preserves all first-order formulas.
(3) For every n ∈ N, the orbits of n-tuples of Aut(C) are primitively positively
definable in C.
(4) For every e ∈ End(B), n ∈ N, and a ∈ C n there exists i ∈ End(C) such that
i(e(a)) = a.
Theorem 7.5.2 (from [14]; see [16]). For every ω-categorical structure B there
exists an model-complete core structure C which is homomorphically equivalent to B;
the structure C is unique up to isomorphism, and ω-categorical. We may assume that
C is an induced substructure of B.
Remark 7.5.3. Saracino’s theorem.
7.5.2. Range Rigid Functions. The results in this section are from [120]. Let
G be a permutation group on a set X. A function g : X → X is called range-rigid
with respect to G if for all β ∈ G we have
g ∈ {α ◦ g ◦ β ◦ g | α ∈ G}.
In particular, the identity map is range-rigid. Note that a function g : X → X
is range-rigid with respect to G if and only if for all t ∈ X n , n ∈ N, if there exists
s ∈ X n and α ∈ G such that αg(s) = t, then t and g(t) lie in the same orbit. In
other words, g preserves the orbits of n-tuples that have non-empty intersection with
g(X)n .
Example 95. Let U, V be unary relation symbols, and let B be a countably
infinite {U, V }-structure where U = {u}, V = {v}, and u 6= v. Let w ∈ B \ {u, v}.
Then the map g : B → B with g(w) = g(u) = u and g(x) = v for all x ∈ B \ {u, w}
is range-rigid with respect to Aut(B), but not canonical with respect to B. The
identity map is canonical with respect to Aut(B) but not range-rigid with respect to
Aut(B). 4
We can use canonisation to find range-rigid functions in sufficiently rich sets of
operations.
Theorem 7.5.4. Let G be a closed oligomorphic extremely amenable permutation
group on a countable set X and let M be a non-empty closed transformation semigroup
on X such that G ◦ M ◦ G ⊆ M . Then M contains a function which is range-rigid
with respect to G and canonical with respect to G.
Proof. Pick any f ∈ M . Applying Proposition 7.4.2 to f , we obtain a function
f 0 ∈ {αf β | α, β ∈ G} ⊆ M which is canonical with respect to G. Since G is oligo-
morphic, for every n ∈ N there are finitely many orbits of n-tuples for every n, so we
may compose f 0 with itself sufficiently many times to obtain a function such that for
all t ∈ X n whose orbit contain a tuple of the form g(s) for s ∈ X n we have that g(t)
lies in the same orbit as t. A standard compactness argument shows that there is one
function that does it for all n; note that the resulting map is range-rigid with respect
to G and canonical with respect to G.
7.5. MODEL-COMPLETE CORES OF RAMSEY STRUCTURES 121
of g, there exists β ∈ Aut(A) such that β ◦ g(M 0 ) = α(M 0 ); hence, the restriction of
g to M 0 is an embedding, and the claim follows.
00 00
Claim 2. χ is constant on MS . If S 00 ∈ MS , then χ(S 00 ) = χ0 (B[g −1 ◦
f −1 (S 00 )]) = c since g −1 ◦ f −1 (S 00 ) is a copy of S in M 0 .
7.5.3. Range-Rigidity and Model-Complete Cores. A subset S of a monoid
M is called a left ideal of M if M S = S. Note that for every f ∈ M , then set T := M f
is a (closed) left ideal, because M T = M M f ⊆ M f = T .
Lemma 7.5.10. Let M be a monoid and f ∈ M . Then the following are equivalent.
(1) f lies in an inclusion-wise minimal closed left-ideal of M ;
(2) M f is an inclusion-wise minimal closed left-ideal;
(3) f ∈ M ef for every e ∈ M .
Proof. (1) implies (2). Suppose that f ∈ S for some inclusion-wise minimal
closed left-ideal S of M . Then T := M f ⊆ M S = S = S, so T = S by the minimality
of S, and hence T is an inclusion-wise minimal closed left-ideal S of M .
(2) implies (1). We have f ∈ M f since M is a monoid and thus contains 1.
(2) implies (3). Let e ∈ M . Then M ef ⊆ M f is closed, non-empty, and a left-
ideal since M M ef = M ef . By the minimality of M f we have that M ef = M f , so
f ∈ M ef since 1 ∈ M .
(3) implies (2). It suffices to show minimality of T := M f . Suppose that there
exists a non-empty closed left-ideal I of M contained in M f . Let t ∈ I. Then t ∈ M f
so there are r1 , r2 , · · · ∈ M such that limi∈N ri f = t. By assumption, for every i ∈ N
we have that f ∈ M ri f , so there exist si,1 , si,2 , · · · ∈ M such that limj∈N si,j ri f = f .
We claim that limj∈N sij t = f . Indeed,
lim sij t = lim sij lim ri f
j∈N j∈N i∈N
= lim lim sij ri f = f.
i∈N j∈N
Proof. Since A[g(A)] and Ag have the same age and Ag is homogeneous, there
exists an embedding f : A[g(A)] → Ag . Then f ◦ g is a homomorphism of B into C,
because every relation of B has a quantifier-free definition in A, and the same formula
defines the corresponding relations of C in Ag . Since C is even a substructure of B,
they are homomorphically equivalent. So it suffices to prove that C is a model-
complete core. Let e ∈ End(C), n ∈ N, and t ∈ (Ag )n .
Claim 1. We may choose f so that there exists s ∈ An with (f ◦ g 2 )(s) = t.
Indeed, there exists s ∈ An such that g(s) satisfies the same atomic formulas as
t in A, because A[g(A)] and Ag have the same age. Since g is range-rigid with
respect to Aut(A), there exists an automorphism of A that maps g(g(s)) to g(s). The
homogeneity of Ag implies that there exists α ∈ Aut(Ag ) which maps (f ◦ g 2 )(s) to t.
Therefore, α ◦ f is an embedding of A[g(A)] into Ag which has the required property.
Claim 2. There exists h ∈ End(B) such that h◦(e◦f ◦g)(g(s)) = g(s). Otherwise,
S 0 := End(B) ◦ {e ◦ f ◦ g 2 } ⊆ S does not contain g. Since End(B) ◦ S 0 = S 0 , this
contradicts the minimality of S. Then
t = (f ◦ g 2 )(s) = (f ◦ g) ◦ h ◦ (e ◦ f ◦ g)g(s) (Claim 2)
= (f ◦ g ◦ h) ◦ e(t).
The restriction of f ◦ g ◦ g to Ag therefore proves condition (2) of Proposition 7.5.1
for C, and hence C is a model-complete core.
[2, 10, 32, 33, 37, 87, 107, 128, 133, 134, 156, 157]
Conjecture 8.1 (Thomas [156]). Let A be a homogeneous structure with a fi-
nite relational signature. Then Aut(A) has only finitely many closed supergroups in
Sym(A). Equivalently, A has only finitely many first-order reducts up to interdefin-
ability.
Cameron’s theorem for highly set-transitive permutation groups on a countably
infinite set, Thomas’ result about the closed supergroups of the automorphism group
of the random graph.
Exercises.
(140) Let (V ; T ) be the Fraı̈ssé-limit of the class of all finite tournaments (see
Exercise 57). Show that
Aut(V ; {(x, y, u, v) | T (x, y) ⇔ T (u, v)})
is isomorphic to a semidirect product of Z2 and Aut(V ; T ).
125
CHAPTER 9
127
CHAPTER 10
129
Bibliography
[1] L. Agarwal. The reducts of the generic digraph. Annals of Pure and Applied Logic, 167:370–391,
2016.
[2] L. Agarwal and M. Kompatscher. 2ℵ0 pairwise nonisomorphic maximal-closed subgroups of
Sym(N) via the classification of the reducts of the Henson digraphs. Journal of Symbolic Logic,
83(2):395–415, 2018.
[3] G. Ahlbrandt and M. Ziegler. Quasi-finitely axiomatizable totally categorical theories. Annals
of Pure and Applied Logic, 30(1):63–82, 1986.
[4] R. Akhtar and A. H. Lachlan. On countable homogeneous 3-hypergraphs. Arch. Math. Log.,
34(5):331–344, 1995.
[5] R. Baer. Die Kompositionsreihe der Gruppe aller eineindeutigen Abbildungen einer unendlichen
Menge auf sich. Studia Mathematica, 5:15–17, 1934.
[6] S. Barbina and D. Macpherson. Reconstruction of homogeneous relational structures. Journal
of Symbolic Logic, 72(3):792–802, 2007.
[7] L. Barto, M. Kompatscher, M. Olšák, T. V. Pham, and M. Pinsker. Equations in oligomor-
phic clones and the constraint satisfaction problem for ω-categorical structures. Journal of
Mathematical Logic, 19(2):#1950010, 2019.
[8] L. Barto and M. Pinsker. The algebraic dichotomy conjecture for infinite domain constraint sat-
isfaction problems. In Proceedings of the 31th Annual IEEE Symposium on Logic in Computer
Science – LICS’16, pages 615–622, 2016. Preprint arXiv:1602.04353.
[9] H. Becker and A. Kechris. The Descriptive Set Theory of Polish Group Actions. Number 232
in LMS Lecture Note Series. Cambridge University Press, 1996.
[10] J. H. Bennett. The reducts of some infinite homogeneous graphs and tournaments. PhD thesis,
Rutgers university, 1997.
[11] V. Bergelson. Ergodic Ramsey theory: a dynamical approach to static theorems. In Proceedings
of the International Congress of Mathematicians, volume II, pages 1655–1678, Zürich, 2006.
European Mathematical Society.
[12] M. Bhattacharjee, D. Macpherson, R. G. Möller, and P. M. Neumann. Notes on Infinite Per-
mutation Groups. Springer Lecture Notes in Mathematics, 1998.
[13] C. E. Blair. The baire category theorem implies the principle of dependent choices. Bull Acad.
Polon. Sci. Ser. Sci. Math. Astron. Phys., 25, 1977.
[14] M. Bodirsky. Cores of countably categorical structures. Logical Methods in Computer Science
(LMCS), 3(1):1–16, 2007.
[15] M. Bodirsky. Ramsey classes: Examples and constructions. In Surveys in Combinatorics. Lon-
don Mathematical Society Lecture Note Series 424. Cambridge University Press, 2015. Invited
survey article for the British Combinatorial Conference; ArXiv:1502.05146.
[16] M. Bodirsky. Complexity of Infinite-Domain Constraint Satisfaction. Lecture Notes in Logic
(52). Cambridge University Press, Cambridge, United Kingdom; New York, NY, 2021.
[17] M. Bodirsky. Model theory, 2022. Course Notes, TU Dresden, https://wwwpub.zih.
tu-dresden.de/~bodirsky/Model-theory.pdf.
[18] M. Bodirsky. Introduction to mathematical logic, 2023. Course notes, TU Dresden, https:
//wwwpub.zih.tu-dresden.de/~bodirsky/Logic.pdf.
[19] M. Bodirsky, P. Jonsson, and T. V. Pham. The reducts of the homogeneous binary branching
C-relation. Journal of Symbolic Logic, 81(4):1255–1297, 2016. Preprint arXiv:1408.2554.
[20] M. Bodirsky, P. Jonsson, and T. V. Pham. The Complexity of Phylogeny Constraint Satisfac-
tion Problems. ACM Transactions on Computational Logic (TOCL), 18(3), 2017. An extended
abstract appeared in the conference STACS 2016.
[21] M. Bodirsky, B. Martin, M. Pinsker, and A. Pongrácz. Constraint satisfaction problems for
reducts of homogeneous graphs. SIAM Journal on Computing, 48(4):1224–1264, 2019. A con-
ference version appeared in the Proceedings of the 43rd International Colloquium on Automata,
Languages, and Programming, ICALP 2016, pages 119:1-119:14.
131
132 BIBLIOGRAPHY
[22] M. Bodirsky and A. Mottet. Reducts of finitely bounded homogeneous structures, and lifting
tractability from finite-domain constraint satisfaction. In Proceedings of the 31th Annual IEEE
Symposium on Logic in Computer Science (LICS), pages 623–632, 2016. Preprint available at
ArXiv:1601.04520.
[23] M. Bodirsky and M. Pinsker. Reducts of Ramsey structures. AMS Contemporary Mathematics,
vol. 558 (Model Theoretic Methods in Finite Combinatorics), pages 489–519, 2011.
[24] M. Bodirsky and M. Pinsker. Minimal functions on the random graph. Israel Journal of Math-
ematics, 200(1):251–296, 2014.
[25] M. Bodirsky and M. Pinsker. Schaefer’s theorem for graphs. Journal of the ACM, 62(3):52
pages (article number 19), 2015. A conference version appeared in the Proceedings of STOC
2011, pages 655-664.
[26] M. Bodirsky and M. Pinsker. Topological Birkhoff. Transactions of the American Mathematical
Society, 367:2527–2549, 2015.
[27] M. Bodirsky and M. Pinsker. Canonical functions: a proof via topological dynamics. Homoge-
neous Structures, A Workshop in Honour of Norbert Sauer’s 70th Birthday, Contributions to
Discrete Mathematics, 16(2):36–45, 2021.
[28] M. Bodirsky, M. Pinsker, and A. Pongrácz. The 42 reducts of the random ordered graph.
Proceedings of the LMS, 111(3):591–632, 2015. Preprint available from arXiv:1309.2165.
[29] M. Bodirsky, M. Pinsker, and A. Pongrácz. Projective clone homomorphisms. Journal of Sym-
bolic Logic, 86(1):148–161, 2021.
[30] M. Bodirsky, M. Pinsker, and T. Tsankov. Decidability of definability. Journal of Symbolic
Logic, 78(4):1036–1054, 2013. A conference version appeared in the Proceedings of the Twenty-
Sixth Annual IEEE Symposium on. Logic in Computer Science (LICS 2011), pages 321-328.
[31] M. Bodirsky and M. Wrona. Equivalence constraint satisfaction problems. In Proceedings of
Computer Science Logic, volume 16 of LIPICS, pages 122–136. Dagstuhl Publishing, September
2012.
[32] B. Bodor, P. J. Cameron, and C. Szabó. Infinitely many reducts of homogeneous structures.
Algebra Universalis, 79(2):43, 2018. arXiv:1609.07694.
[33] B. Bodor, K. Kalina, and C. Szabo. Permutation groups containing infinite linear groups and
reducts of infinite dimensional linear spaces over the two element field. Communications in
Algebra, 45(7):2942–2955, 2017. Preprint arXiv:1506.00220.
[34] J. Böttcher and J. Foniok. Ramsey properties of permutations. Electronic Journal of Combi-
natorics, 20(1), 2013.
[35] N. Bourbaki. General Topology, Volume 1. Springer, 1998.
[36] S. Braunfeld. Monadic stability and growth rates of ω-categorical structures. Proceedings of
the London Mathematical Society, Series B, 124(3), 2022. Preprint arXiv:1910.04380.
[37] P. J. Cameron. Transitivity of permutation groups on unordered sets. Mathematische
Zeitschrift, 148:127–139, 1976.
[38] P. J. Cameron. Orbits of permutation groups on unordered sets, II. Journal of the London
Mathematical Society, 2:249–264, 1981.
[39] P. J. Cameron. Oligomorphic permutation groups. Cambridge University Press, Cambridge,
1990.
[40] P. J. Cameron. Permutation Groups. LMS Student Text 45. Cambridge University Press, Cam-
bridge, 1999.
[41] P. J. Cameron. Some counting problems related to permutation groups. Discrete Mathematics,
225(1-3):77–92, 2000.
[42] P. J. Cameron. Homogeneous permutations. Electronic Journal of Combinatorics, 9(2), 2002.
[43] G. Cantor. Über unendliche, lineare Punktmannigfaltigkeiten. Mathematische Annalen,
23:453–488, 1884.
[44] G. Cherlin. Homogeneous digraphs I. The imprimitive case. Logic Colloquium 1985, 1987.
[45] G. L. Cherlin. Combinatorial problems connected with finite homogeneity. Contemporary
Mathematics, 131:3–30, 1993.
[46] G. L. Cherlin. The classification of countable homogeneous directed graphs and countable
homogeneous n-tournaments. AMS Memoir, 131(621), January 1998.
[47] J. Dixon, P. M. Neumann, and S. Thomas. Subgroups of small index in infinite symmetric
groups. Bulletin of the London Mathematical Society, 18(6):580–586, 1986.
[48] J. D. Dixon and B. Mortimer. Permutation Groups. Springer, New York, 1996.
[49] F. R. Drake. Set Theory, An Introduction to Large Cardinals. North-Holland Publishing Co.,
Amsterdam, 1974.
[50] M. Droste, C. W. Holland, and D. Macpherson. Automorphism groups of infinite semilinear
orders (I). Proceedings of the London Mathematical Society, 58:454 – 478, 1989.
BIBLIOGRAPHY 133
[83] T. Jech. Set theory. Springer Monographs in Mathematics. Springer-Verlag, Berlin, 2003. The
third millennium edition, revised and expanded.
[84] T. Jech and K. Hrbáček. Introduction to Set Theory, Third edition. Monographs and Textbooks
in Pure and Applied Mathematics. CRC Press, 1999.
[85] T. Jenkinson, J. K. Truss, and D. Seidel. Countable homogeneous multipartite graphs. Euro-
pean Journal of Combinatorics, 33:82–109, 2012.
[86] R. B. Jensen. Independence of the axiom of dependent choices from the countable axiom of
choice. Journal of Symbolic Logic, 31:294, 1966.
[87] M. Junker and M. Ziegler. The 116 reducts of (Q, <, a). Journal of Symbolic Logic, 74(3):861–
884, 2008.
[88] A. Kechris. Classical descriptive set theory, volume 156 of Graduate Texts in Mathematics.
Springer, 1995.
[89] A. Kechris, V. Pestov, and S. Todorčević. Fraı̈ssé limits, Ramsey theory, and topological dy-
namics of automorphism groups. Geometric and Functional Analysis, 15(1):106–189, 2005.
[90] A. Kechris and M. Sokić. Dynamical properties of the automorphism groups of the random
poset and random distributive lattice. Fund. Math., 218(1):69–94, 2012.
[91] A. S. Kechris. Dynamics of non-archimedean Polish groups. In Proceedings of the European
Congress of Mathematics, Krakow, pages 375–397. European Math. Society, 2014.
[92] A. S. Kechris and C. Rosendal. Turbulence, amalgamation and generic automorphisms of ho-
mogeneous structures. Proceedings of the London Mathematical Society, 94(3):302–350, 2007.
[93] J. L. Kelley. The tychonoff product theorem implies the axiom of choice. Fund. Math., 37:75–76,
1950.
[94] B. Klin, E. Kopczynski, J. Ochremiak, and S. Toruńczyk. Locally finite constraint satisfaction
problems. In 30th Annual ACM/IEEE Symposium on Logic in Computer Science, LICS 2015,
Kyoto, Japan, pages 475–486, 2015.
[95] B. Klin, S. Lasota, J. Ochremiak, and S. Torunczyk. Homomorphism Problems for First-Order
Definable Structures. In A. Lal, S. Akshay, S. Saurabh, and S. Sen, editors, 36th IARCS An-
nual Conference on Foundations of Software Technology and Theoretical Computer Science
(FSTTCS 2016), volume 65 of Leibniz International Proceedings in Informatics (LIPIcs),
pages 14:1–14:15, Dagstuhl, Germany, 2016. Schloss Dagstuhl–Leibniz-Zentrum fuer Infor-
matik.
[96] M. Kompatscher and T. V. Pham. A Complexity Dichotomy for Poset Constraint Satisfaction.
In 34th Symposium on Theoretical Aspects of Computer Science (STACS), volume 66 of Leibniz
International Proceedings in Informatics (LIPIcs), pages 47:1–47:12, 2017.
[97] M. Krom. Equivalents of a weak axiom of choice. Notre Dame Journal of Formal Logic, 22(3),
1981.
[98] A. Kwiatkowska and A. Panagiotopoulos. The automorphism group of the random poset does
not admit a generic pair, 2020. ArXiv 2012.04376.
[99] A. H. Lachlan. Countable homogeneous tournaments. Transactions of the American Mathe-
matical Society (TAMS), 284:431–461, 1984.
[100] A. H. Lachlan. Stable finitely homogeneous structures: A survey. In Algebraic Model Theory,
NATO ASI Series, volume 496, pages 145–159, 1996.
[101] A. H. Lachlan and R. E. Woodrow. Countable ultrahomogeneous undirected graphs. Transac-
tions of the AMS, 262(1):51–94, 1980.
[102] C. Laflamme, J. Jasinski, L. N. V. Thé, and R. Woodrow. Ramsey precompact expansions of
homogeneous directed graphs. Electron. J. Combin., 21(4), 2014.
[103] D. Lascar. On the category of models of a complete theory. Journal Symbolic Logic, 47(2):249–
266, 1982.
[104] D. Lascar. Autour de la propriété du petit indice. Proceedings of the London Mathematical
Society, 62(1):25–53, 1991.
[105] B. Latka. Finitely constrained classes of homogeneous directed graphs. Journal of Symbolic
Logic, 59(1):124 – 139, 1994.
[106] H. Läuchli. The independence of the ordering principle from a restricted axiom of choice.
Fundamenta Mathematicae, 54:31–43, 1964.
[107] J. Linman and M. Pinsker. Permutations on the random permutation. Electronic Journal of
Combinatorics, 22(2):1–22, 2015.
[108] D. C. Lockett and J. K. Truss. Homogeneous coloured multipartite graphs. European Journal
of Combinatorics, 42:217–242, 2014.
[109] G. Lolli. On Ramsey’s theorem and the axiom of choice. Notre Dame Journal of Formal Logic,
18:599–601, 1977.
BIBLIOGRAPHY 135
[141] J. Schreier and Stanislaw Marcin Ulam. Über die Permutationsgruppe der natürlichen Zahlen-
folge. Studia Mathematica, 4:134–141, 1933.
[142] S. W. Semmes. Endomorphisms of infinite symmetric groups. Abstracts of the American Math-
ematical Society, 2:426, 1981.
[143] S. Shelah. Can you take Solovay’s inaccessible away? Israel Journal of Mathematics, 48(1):1–
47, 1984.
[144] P. Simon. On ω-categorical structures with few finite substructures, 2018.
[145] D. Siniora and S. Solecki. Coherent extension of partial automorphisms, free amalgamation
and automorphism groups. J. Symb. Log., 85(1):199–223, 2020.
[146] D. N. Siniora. Automorphism groups of homogeneous structures. Ph.D. thesis, University of
Leeds, 2017.
[147] M. Sokić. Ramsey property of posets and related structures. PhD thesis, University of Toronto,
2010.
[148] M. Sokić. Ramsey property, ultrametric spaces, finite posets, and universal minimal flows.
Israel Journal of Mathematics, 194(2):609–640, 2013.
[149] M. Sokić. Directed graphs and Boron trees. Journal of Combinatorial Theory, Series A,
132:142–171, 2015.
[150] R. M. Solovay. A model of set theory in which every set of reals is Lebesgue measurable. Annals
of Mathematics, 92:1–56, 1970.
[151] A. Tarski. Prime ideal theorem for set algebras and ordering principles. Bulletin of the Amer-
ican Mathematical Society, 60:390–39, 1954.
[152] K. Tent and M. Ziegler. A course in model theory. Lecture Notes in Logic. Cambridge University
Press, 2012.
[153] L. N. V. Thé. More on the Kechris-Pestov-Todorcevic correspondence: precompact expansions.
Fund. Math., 222(1):19–47, 2013. Preprint arXiv:1201.1270.
[154] L. N. V. Thé. Universal flows of closed subgroups of S∞ and relative extreme amenability.
Asymptotic Geometric Analysis, Fields Institute Communications, 68:229–245, 2013.
[155] L. N. V. Thé. A survey on structural Ramsey theory and topological dynamics with the Kechris-
Pestov-Todorčević correspondence in mind. Accepted for publication in Zb. Rad. (Beogr.),
2014. Preprint arXiv:1412.3254v2.
[156] S. Thomas. Reducts of the random graph. Journal of Symbolic Logic, 56(1):176–181, 1991.
[157] S. Thomas. Reducts of random hypergraphs. Annals of Pure and Applied Logic, 80(2):165–193,
1996.
[158] J. K. Truss. Infinite permutation groups. II. Subgroups of small index. Journal of Algebra,
120(2):494–515, 1989.
[159] J. K. Truss. Generic automorphisms of homogeneous structures. Proc. London Math. Soc.,
3(65):121–141, 1992.
[160] J. K. Truss. On notions of genericity and mutual genericity. The Journal of Symbolic Logic,
72(3):755–766, 2007.
[161] A. N. Tychonoff. über die topologische Erweiterung von Räumen. Mathematische Annalen,
102(1):544–561, 1930.
[162] I. B. Yaacov and T. Tsankov. Weakly almost periodic functions, model-theoretic stability,
and minimality of topological groups. Transactions of the AMS, 368(11):8267–8294, 2016.
arXiv:1312.7757.
[163] A. Zucker. Amenability and unique ergodicity of automorphism groups of Fraı̈ssé structures.
Fund. Math., 841:41–62, 2014. Preprint, arXiv:1304.2839.
[164] A. Zucker. Topological dynamics of closed subgroups of Sω . Preprint, arXiv:1404.5057, 2014.
APPENDIX A
Background Material
A.1. Ultrafilter
Let X be a set. A filter on X is a certain set of subsets of X; the idea is that the
elements of F are (in some sense) ‘large’; it helps thinking of the elements F ∈ F as
being ‘almost all’ of X.
Definition A.1.1. A filter F on X is a set of subsets of X such that
(1) ∅ ∈/ F and X ∈ F;
(2) if F ∈ F and G ⊆ X contains F , then G ∈ F.
(3) if F1 , F2 ∈ F then F1 ∩ F2 ∈ F.
Note that filters have the finite intersection property:
A1 , . . . , An ∈ F ⇒ A1 ∩ · · · ∩ An 6= ∅ (FIP)
Lemma A.1.2. Every subset S ⊆ P(X) with the FIP is contained in a smallest
filter that contains S; this filter is called the filter generated by S.
Proof. First add finite intersections, and then all supersets to S.
Example 96. For a non-empty subset Y ⊆ X, the family
F := {Z ⊆ X | Y ⊆ Z}
is a filter, the filter generated by a {Y }; such filters are called principal. 4
Example 97. The Fréchet filter : for an infinite set X this is the filter F that
consists of all cofinite subsets of X, i.e.,
F := {Y ⊆ X | X \ Y is finite}. 4
A filter F is called a ultrafilter if F is maximal, that is for every filter G ⊇ F we
have G = F.
Lemma A.1.3. Let F be a filter. Then the following are equivalent.
(1) F is a ultrafilter.
(2) For all A ⊆ X either A ∈ F or X \ A ∈ F.
(3) For all A1 ∪ · · · ∪ An ∈ F there is an i ≤ n with Ai ∈ F.
Proof. (1) ⇐ (2): No A ⊆ X can be added to F. Hence, F is maximal.
(2) ⇐ (3): Note that A ∪ (X \ A) = X ∈ F.
(1) ⇒ (3): If there is an i ≤ n such that F ∪ {Ai } has the FIP, then by
Lemma A.1.2 there is a filter that contains this set, and hence F was not maxi-
mal. Otherwise, there are S1 , . . . , Sn ⊆ F with Ai ∩ Si = ∅. Then Si ⊆ X \ Ai and
thus S1 ∩ · · · ∩ Sn ⊆ X \ (A1 ∪ · · · ∪ An ) ∈
/ F, a contradiction.
A filter F is principal if it
T contains a inclusionwise minimal element. Note that
this is the case if and only if F ∈ F.
Lemma A.1.4. Let F be a filter on a set X. Then the following are equivalent.
137
138 A. BACKGROUND MATERIAL
(1) F
is a principal ultrafilter;
(2) F
contains {a} for some a ∈ X.
(3) F
is of the form {Y ⊆ X | a ∈ Y } for some a ∈ X.
(4) F
is an ultrafilter and contains a finite set.
T
Proof. (1) ⇒ (2): let A := F ∈ F. If |A| > 1 then we can write A = B1 ∪ B2
for B1 , B2 ⊆ X non-empty. But then Lemma A.1.3 (3) implies that B1 ∈ F or
B2 ∈ F, in contradiction to the definition of A. So A = {a} for some a ∈ X.
(2) ⇒ (3). Clearly, {Y ⊆ X | a ∈ Y } ⊆ F since F is closed under supersets, and
F ⊆ {Y ⊆ X | a ∈ Y } since F does not contain the empty set.
(3) ⇒ (4): Clearly F contains a finite set; use Lemma A.1.3 (2) to check that F
is an ultrafilter. T
(4) ⇒ (1): If A ∈ F is finite, then B := F is finite, and hence B is the
intersection of finitely many elements in F, and hence in F since F is a filter. This
shows that F is principal.
Are there non-principal ultrafilters?
Lemma A.1.5 (Ultrafilter Lemma). Every filter F is contained in a ultrafilter.
Proof. Let M be the set of all filters on X that contain F, partially ordered
by containment. Note that unions of chains of filters in this partial order are again
filters. By Zorn’s lemma, M contains a maximal filter.
Non-principal ultrafilters are also called free ultrafilters. In particular the Fréchet
filter is contained in an ultrafilter, which must be free:
Lemma A.1.6. An ultrafilter is free if and only if it contains the Fréchet filter.
Proof. Let U be a free ultrafilter on X and let x ∈ X. Either {x} ∈ U or
X \ {x} ∈ U. As U is free, {x} ∈
/ U (Lemma A.1.4). Hence, X \ {x} ∈ U for every
x ∈ X. Let F ⊆ X be finite. Then
\
X \F = (X \ {x}) ∈ U .
x∈F
Now let U be a principal ultrafilter, i.e., there is x ∈ X with {x} ∈ U (Lemma A.1.4).
Then the element X \ {x} of the Fréchet filters is not in U.
Exercises.
(143) Show that a set of subsets of a set X can be extended to an ultrafilter if and
only if it has the FIP.
(144) Show that a set F of subsets of a set X can be extended to a free ultrafilter
if and only if the intersection of every finite subset of F is infinite.
(145) Show that every filter F on a set X is the intersection of all ultrafilters on
X that extend F.
(146) Show that if U is a free ultrafilter on X, and S ∈ U and T ⊆ X are
such that the symmetric difference S∆T is finite, then S ∈ U.
|X|
(147) Show that there are 22 many ultrafilters on an infinite set X.
Hint: first show that there is a family F of 2|X| subsets of X such that for
any A1 , . . . , An , B1 , . . . , Bn ∈ F
A1 ∩ · · · ∩ An ∩ (X \ B1 ) ∩ · · · ∩ (X \ Bn ) 6= ∅.
(148) True or false: if U and V are free ultrafilters on an infinite set X, is there is
a permutation π of X such that S ∈ U if and only if π(S) ∈ V?1
1Thanks to Lukas Juhrich for the idea for this exercise.
A.2. THE AXIOM OF CHOICE AND ITS WEAKER VERSIONS 139
Ultrafilter Compactness
Lemma of first-order logic Axiom of Baire Category
Dependent Choices Theorem
Boolean Prime Ideal
Theorem
Tychonoff Theorem Tychonoff Theorem Tychonoff Theorem
for Hausdorff spaces for Finite Spaces for Countable Products
ZF
Countable Choice from Finite Sets is equivalent to the Axiom of Countable Choice,
or whether it is equivalent to the Axiom of Choice from Finite Sets.
Exercises.
(149) Show that the Order Extension Property implies the Ordering
Property.
(150) Show that the Ordering Property implies the Axiom of Choice
for families of non-empty finite sets.
(151) Show that the Axiom of Choice is equivalent to the statement
that every surjective function f : A → B has a right inverse,
i.e., a function g : B → A such that g(f (x)) = x for all x ∈ A.
(152) Show that the Axiom of Choice implies the Axiom
of Dependent Choices.
(153) (Exercise 5.7 in [83]) Show that the Axiom of Dependent
Choices implies the Axiom of Countable Choice
(without quoting the facts stated above).
Hint. Given (An )n∈N , consider the set A of all choice functions
on some Sn := {Ai | i ≤ n}, ordered by extensions.