Model Theory
Model Theory
Model Theory
Course Notes
Chapter 2. Compactness 13
2.1. Filter 14
2.2. Ultrafilter 14
2.3. Ultraproducts 16
2.4. The Theorem of Łoś 16
2.5. Proof of the Compactness Theorem 18
2.6. Proving the Ultrafilter Lemma with Compactness 18
Chapter 5. Types 35
5.1. Stone Spaces 37
5.2. Saturated Structures 37
5.3. Omitting Types 40
Chapter 8. Stability 65
8.1. The Order Property 66
8.2. Indiscernible Sequences 68
8.3. ω-stable Theories 71
8.4. The Classification Picture 71
First-order Logic
There are many excellent text books about model theory. A classic is Hodges’
model theory [18]; a short version is also available [19]. Another introduction is given
by Marker [22]. A more recent book has been written by Tent and Ziegler [29].
In other words, two points are linked by an edge if they have distance one. What is
the chromatic number of this graph? 4
The following statement follows from the compactness theorem of first-order logic,
as we will see later in this chapter.
Proposition 1.1.1. Let G be a graph such that all finite subgraphs of G are
k-colourable. Then G is k-colourable.
The problem to determine the chromatic number χ of this graph is known as the
Hadwiger-Nelson problem. It is known that χ ≤ 7. We have seen that 4 ≤ χ. In
April 11, 2018, Aubrey de Grey announced a proof that 5 ≤ χ. The precise value of
χ ∈ {5, 6, 7} is not known.
1.1.6. Products. Let A and B be τ -structures. Then the (direct, or categorical )
product C = A×B is the τ -structure with domain A×B, which has for each k-ary R ∈
τ the relation that contains a tuple ((a1 , b1 ), . . . , (ak , bk )) if and only if R(a1 , . . . , ak )
holds in A and R(b1 , . . . , bk ) holds in B. For each k-ary f ∈ τ the structure C has
the operation that maps ((a1 , b1 ), . . . , (ak , bk )) to (f (a1 , . . . , ak ), f (b1 , . . . , bk )). The
direct product A × A is also denoted by A2 , and the k-fold product A × · · · × A,
defined analogously, by Ak .
Note that the product of two groups, viewed as structures, is precisely the well-
known product of groups.
We generalise the definition of products to infinite products Q as follows. For a
sequence of τ -structures
Q (A )
i i∈I , the direct product B = i∈I i is the τ -structure
A
on the domain i∈I Ai such that for R ∈ τ of arity k
(a1i )i∈I , . . . , (aki )i∈I ∈ RB iff (a1i , . . . , aki ) ∈ RAi for each i ∈ I ,
If Ai = A for all i ∈ I, we also write AI instead of i∈I Ai , and call it the I-th power
Q
of A.
Exercises.
(1) Show that there is a homomorphism from A2 to A. Find an example of a
structure A where all such homomorphisms are strong, and another example
where all such homomorphisms are not strong.
(2) Find a signature τ for vector spaces and describe how a vector space may
be viewed as a τ -structure. Your definition should have the property that
homomorphisms between the structures you consider correspond precisely
to linear maps.
8 1. FIRST-ORDER LOGIC
Note that item 1 is a special case of item 3. The function described by t is also called
the term function of t (with respect to A).
Exercises.
(6) Let B be a τ -structure and G ⊆ B. Let A be the substructure of B generated
by G. Show that for every element a ∈ A there exists a τ -term t(x1 , . . . , xn )
and elements g1 , . . . , gn ∈ G such that tB (g1 , . . . , gn ) = a.
1.2.3. Formulas and sentences. Let τ be a signature. The relation symbols
in the signature τ did not play any role when defining τ -terms, but they become
important when defining τ -formulas. Moreover, the equality symbol = is ‘hard-wired’
into first-order logic; we can use it to create formulas by equating terms. Finally, we
can combine formulas using Boolean connectives, and quantify over variables.
Definition 1.2.2. An atomic (τ -)formula is an expression of the form
• t1 = t2 where t1 and t2 are τ -terms;
1.2. FORMULAS, SENTENCES, THEORIES 9
A A
• if φ equals φ1 ∧ φ2 then φA := φ1 ∩ φ2 ;
• if φ equals ¬ψ then φA := An \ ψ A ;
• if φ equals ∃x. ψ(x, x1 , . . . , xn ) then
φA := {(a1 , . . . , an ) | there exists a ∈ A s.t. (a, a1 , . . . , an ) ∈ ψ A }.
A relation R ⊆ Ak is called (first-order) definable in A if there exists a τ -formula
φ(x1 , . . . , xk ) such that R = φA ; we also say that φ defines R over A. An function
f : S → T , for S ⊆ Ak and T ⊆ Al , is called definable if the relation
{(a1 , . . . , ak , b1 , . . . , bl ) | f (a1 , . . . , ak ) = (b1 , . . . , bl )}
is definable in A.
We freely use brackets to avoid ambiguities when writing down terms. That is,
we write ¬(x ∧ y) to distinguish it from (¬x) ∧ y. If brackets are omitted, there is the
convention that negation ¬ binds stronger than conjunction ∧; that is, ¬x ∧ y would
stand for (¬x) ∧ y.
For φ(x1 , . . . , xn ) we write
A |= φ(a1 , . . . , an )
A
instead of (a1 , . . . , an ) ∈ φ . In particular, if φ is a sentence, i.e., if n = 0, we write
A |= φ, and say that A satisfies φ, if () ∈ φA (that is, if φA 6= ∅).
10 1. FIRST-ORDER LOGIC
Shortcuts:
• Disjunction: φ ∨ ψ is an abbreviation for ¬(¬φ ∧ ¬ψ)
• Implication: φ ⇒ ψ is an abbreviation for ¬φ ∨ ψ
• Equivalence: φ ⇔ ψ is an abbreviation for φ ⇒ ψ ∧ ψ ⇒ φ
• Universal quantification: ∀x. φ(x) is an abbreviation for ¬∃x. ¬φ(x)
• Inequality: x 6= y is an abbreviation for ¬(x = y)
• False: ⊥ is an abbreviation for ∃x.x 6= x.
• True: > is an abbreviation for ¬⊥ (the same as ∀x.x = x).
Moreover, when A is a unary relation symbol, then we may write ∃x ∈ A. φ instead
of ∃x (A(x) ∧ φ) and ∀x ∈ A. φ instead of ∀x (A(x) ⇒ φ).
Example 10. The following statements about well-known structures follow straight-
forwardly from the definitions.
• (Z; <) |= 0 < 1
• (Q; <) |= ∀x, y x < y ⇒ ∃z (x < z ∧ z < y) (density)
• (Z; <) |6 = ∀x, y x < y ⇒ ∃z (x < z ∧ z < y) 4
∃x∀y. ¬(y ∈ x)
stating the existence of the empty set. Note that ZF is not complete: for example the
Axiom of Choice (C) is independent from ZF in the sense that both ZFC := ZF ∪ {C}
and ZF ∪ {¬C} have a model. 4
Example 12. The theory TAGroup of abelian groups is over the signature τAGroup =
{+, −, 0} and contains the following axioms:
∀x, y, z. (x + y) + z = x + (y + z) (associativity)
∀x. 0 + x = x + 0 = y (neutral element)
∀x. x + (−x) = 0 (inverse elements)
∀x, y. x + y = y + x (Abelian) 4
Example 13. The theory TCRing of commutative rings is over the signature τRing ,
contains TAGroup and the following additional axioms:
Example 14. The theory TField of fields is over the signature τRing , contains
TCRing and the following additional axioms:
¬(0 = 1)
∀x (¬(x = 0) ⇒ ∃y.xy = 1) 4
If S and T are τ -theories then we write
S |= T
if every model of S is also a model of T . We also write S |= φ instead of S |= {φ}.
Exercises.
(7) Let A and B be τ -structures. If a ∈ An and b ∈ B n satisfy the same
quantifier-free formulas, then the substructure of A generated by the entries
of a is isomorphic to the substructure of B generated by the entries of b.
(8) Show that for every first-order formula φ(x) and for every ` ∈ N there exists
a first-order sentence ψ such that a structure A satisfies ψ if and only if
there are precisely ` elements a ∈ A such that A |= φ(a). The sentence ψ is
often denoted by ∃x=` .φ.
(9) Write down the axioms of algebraically closed fields in first-order logic.
(10) A formula is in prenex normal form if it is of the form Q1 x1 . . . Qn xn . φ where
Qi is either ∃ or ∀ and φ is without quantifiers. Show that every formula
φ(y1 , . . . , yn ) is equivalent to a formula ψ(y1 , . . . , yn ) in prenex normal form,
that is, for every structure M we have M |= ∀ȳ(φ(ȳ) ⇔ ψ(ȳ)).
(11) Show that every quantifier-free formula is equivalent to a formula in conjunc-
tive normal form, i.e., to a conjunction of disjunctions of atomic formulas
and negated atomic formulas.
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; <). 4
Vaught’s theorem states that for any theory T the spectrum IT (ℵ0 ) cannot be
two. Morley showed that if IT (ℵ0 ) is infinite, then
IT (ℵ0 ) ∈ {ℵ0 , ℵ1 , 2ℵ0 }
Compactness
A special case of the compactness theorem was already proved by Gödel; the
general case below was proved by A. Maltsev in 1936.
Theorem 2.0.1. A first-order theory T is satisfiable if and only if all finite subsets
of T are satisfiable.
Remark 2.0.2. The name compactness theorem comes from the fact that the
compactness theorem is equivalent to the statement that the following natural topo-
logical space is compact: the space is the set T (τ ) of all complete τ -theories, and the
basic open sets are the sets Tφ of the form {T ∈ T (τ ) | φ ∈ T }.
To see the equivalence, let C be Sa covering of T (τ ) by open subsets of T (τ ). This
means that T (τ ) can be written as φ∈S Tφ for some set S of τ -sentences. Consider
the set S 0 of all τ -sentences that are not contained in S. If S 0 has a model B, then
the theory Th(B) is not covered by C. The compactness theorem implies that there
is a finite subset F of S 0 which is unsatisfiable. But then {T¬φ | φ ∈ F } is a finite
subset of C covering T (τ ), showing that T (τ ) is compact.
Conversely, suppose that T (τ ) is compact, and that the τ -theory S is unsatisfiable.
Then {T¬φ | φ ∈ S} is an open covering of T (τ ). So by compactness
S it has a finite
subcovering, i.e., there is a finite subset F of S such that {T¬φ | φ ∈ F } = T (τ ).
Hence, F is unsatisfiable, which is the statement of the compactness theorem.
Exercises.
(17) Show that if a first-order theory has infinite models, then it also has arbi-
trarily large models (for every set X there is a model M with |M | ≥ |X|).
(18) Let K be a field. The characteristic of K is the smallest n ∈ N such that
1 + ··· + 1 = 0
| {z }
n times
2.1. Filter
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 2.1.1. 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 2.1.2. 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.
2.2. Ultrafilter
A filter F is called a ultrafilter if F is maximal, that is for every filter G ⊇ F we
have G = F.
Lemma 2.2.1. 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 2.1.2
there is a filter that contains this set, and hence F was not maximal. 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 2.2.2. Let F be a filter on a set X. Then the following are equivalent.
(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.
2.2. ULTRAFILTER 15
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 2.2.1 (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 2.2.1 (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.
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 (Theorem A.2.1), 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 2.2.4. An ultrafilter is free if and only if it contains the Fréchet filter.
Now let U be a principal ultrafilter, i.e., there is x ∈ X with {x} ∈ U (Lemma 2.2.2).
Then the element X \ {x} of the Fréchet filters is not in U.
Exercises.
(19) Show that a set of subsets of a set X can be extended to an ultrafilter if and
only if it has the FIP.
(20) 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.
(21) Show that every filter F on a set X is the intersection of all ultrafilters on
X that extend F.
(22) Show that if U is a free ultrafilter on X, S ∈ U, and T ⊆ X is such that the
symmetric difference S∆T is finite, then S ∈ U.
|X|
(23) 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= ∅.
(24) 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
2.3. Ultraproducts
Let τ be a signature, let U be an ultrafilter on X, and for each a ∈ X let M a be
a τ -structure. The idea of ultraproducts is to define an “average” structure of all the
structures M a .
Definition 2.3.1. We write
Y
M a /U
a∈X
for the τ -structure M whose domain are the equivalence classes of the equivalence
relation ∼ defined on the set
Y [
Ma := {g : X → Ma | for all x ∈ X : g(a) ∈ Ma }
a∈X a∈X
as follows:
g ∼ g 0 :⇔ {a ∈ X | g(a) = g 0 (a)} ∈ U.
The equivalence class of a function g with respect to the equivalence relation ∼ will
be denoted by [g]. The relations and functions of M are defined as follows:
• for constant symbols c ∈ τ :
cM := [a 7→ cM a ]
• for relation symbols R ∈ τ of arity k:
RM ([g1 ], . . . , [gk ]) :⇔ {a ∈ X | RM a (g1 (a), . . . , gk (a))} ∈ U
• for function symbols f ∈ τ of arity k:
f M ([g1 ], . . . , [gk ]) = [g0 ] :⇔ {a ∈ X | f M a (g1 (a), . . . , gk (a)) = g0 (a)} ∈ U
It is straightforward (but tedious) to verify that this is indeed well defined, i.e.,
the interpretation of function and relation symbols is independent from the choice of
the representatives2.
Exercises.
(25) Show that Definition 2.3.1 is well-defined.
(26) Let n ∈ N and let M be a ultraproduct of finite structures each of which
has at most n elements. Then M has at most n elements, too.
(27) (Exercise 1.31 in [17]). Let X be an index set and U an ultrafilter on X. Let
(Aa )a∈X and (B a )a∈X be families
Q of τ -structures. If Aa can be Q embedded in
B a for all a ∈ X, show that a∈X Aa /U can be embedded in a∈X B a /U.
for example
x = [(1, 1/2, 1/3, 1/4, . . . )].
Abraham Robinson used this idea to develop ‘nonstandard analysis’ which gives a for-
mal interpretation to the reasoning with infinitely small positive entities à la Leibniz,
Euler, and Cauchy. 4
The theorem of Löwenheim and Skolem implies that if T is a theory with infinite
models, then T has models of arbitrary infinite cardinality κ, unless the signature is
larger than κ. The proof of the theorem of Löwenheim and Skolem and has two parts,
upwards and downwards (Theorem 3.2.1). Going upwards is essentially an application
of the compactness theorem. For going downwards, we need to introduce elementary
chains.
3.1. Chains
We will need another important operation to build structures, namely the for-
mation of limits of chains. Chains of τ -structures are a fundamental concept from
model theory. Let I be a linearly ordered index set. Let (Ai )i∈I be a sequence of
τ -structures. Then (Ai )i∈I is called a chain if Ai is a substructure of Aj for all i < j.
Definition 3.1.1. The union of a chain (Ai )i∈I is a τ -structure B = limi∈I Ai
defined as follows.
S
• The domain of B is B := i∈I Ai .
• For each relation symbol R ∈ τ put a ∈ RB if a ∈ RAi for some i ∈ I.
• For each function symbol f ∈ τ put f B (a) := b if f Ai (a) = b for some i ∈ I.
Example 20. For each n ∈ N, let
An := {−n, −n + 1, . . . , 0, 1, 2, . . . } ⊆ Z
and let An := (An ; ≤) be the substructure of (Z; ≤) induced by An . Then Th(An ) =
Th(A0 ) for all n ∈ N, but Th(limi∈I Ai ) is different (why?). 4
Let A and B be τ -structures. A function f : A → B preserves a first-order τ -
formula φ(x1 , . . . , xn ) if and only if for all a1 , . . . , an ∈ A
A |= φ(a1 , . . . , an ) implies B |= φ(f (a1 ), . . . , f (an )).
Remark 3.1.2. This definition is compatible with the notion of preservation as
introduced in Section 1.1.5, in the following sense: if A0 is the τ ∪ {R}-expansion of
0
A by the relation defined by φ in A (i.e., RA = φA ), and B 0 is the τ ∪ {R} expansion
of B by the relation defined by φ in B, then f preserves φ if and only if f is a
homomorphism from A0 to B 0 .
Functions that preserve all first-order formulas are called elementary (they are
necessarily embeddings). If B is an extension of A such that the identity map from
A to B is an elementary embedding, we say that B is an elementary extension of A,
and that A is an elementary substructure of B, and write A ≺ B. Note that A ≺ B
if and only if A is a substructure of B and Th(B A ) = Th(AA ).
Example 21. The structure An+1 from Example 20 is not an elementary exten-
sion of An (why?). 4
21
22 3. THE LÖWENHEIM-SKOLEM THEOREM
Exercises.
(28) Suppose that A is an elementary substructure of B, and B is an elementary
extension of C, and A ⊆ C. Is A an elementary substructure of C?1
(29) Let A, B sets with A ⊆ B and suppose that A is infinite. Show that A :=
(A; =) is an elementary substructure of B := (B; =).
Solution. We have to show that idA : A → B is an elementary embed-
ding of A into B, i.e., preserves all first-order formulas: if φ(x1 , . . . , xn ) is a
first-order formula and a1 , . . . , an ∈ A, then A |= φ(a1 , . . . , an ), if and only if
B |= ψ(a1 , . . . , an ). We show this by induction on the size of φ. If φ is atomic,
it must be of the form x = y, and in this case the statement is clear. The
inductive step is clear if φ is of the form ¬ψ or of the form φ1 ∧φ2 . The inter-
esting case is that φ is of the form ∃y.ψ(y, x1 , . . . , xn ). If A |= φ(a1 , . . . , an )
then clearly B |= φ(a1 , . . . , an ). Conversely, suppose that B |= φ(a1 , . . . , an ).
Then there exists b ∈ B such that B |= ψ(b, a1 , . . . , an ). Since A is infinite
there exists c ∈ A \ {a1 , . . . , an }. Note that (B, a1 , . . . , an ) has an automor-
phism that exchanges b to c and fixes all other elements. The automorphism
shows that A |= ψ(c, a1 , . . . , an ), and hence A |= φ(a1 , . . . , an ). This con-
cludes the induction.
(30) Is (Q; <) an elementary substructure of (R; <)?
(31) Is (Z; <) an elementary substructure of (Q; <)?
(32) (Exercise 2.2.4 in [29]) Let A = (R; 0, <, f A ) where f is a unary function
symbol. Let B be an elementary extension of A. Call b ∈ B an infinitesimal
if − n1 < x < 21 for all n ∈ N. Show that if f A (0) = 0, then f A is continuous
in 0 if and only if for every elementary extension B of A the map f B maps
infinitesimals to infinitesimals.
(33) A formula is called existential if it is built from a quantifier-free formula by
existential quantification (universal quantification is forbidden). Show that
the class of all models of an existential sentence is closed under extensions.
(34) A formula is called ∀∃ if it is built from an existential formula by universal
quantification. Show that the class of all models of a ∀∃ sentence is closed
under unions of chains.
(35) Let φ be a formula in prenex normal form (see Exercise 11) whose quantifier-
free part is written in conjunctive normal form (see Exercise 11). Then φ
is called a Horn formula if each conjunct has at most one disjunct which
is positive, i.e., an atomic formula. In other words, all but one of the dis-
juncts of each conjunct must be of the form ¬ψ for some atomic formula ψ.
Prove that the class of all models of a Horn sentence is closed under taking
products.
Exercises.
(36) Is (Z; <) ℵ0 -categorical?
CHAPTER 4
Fraïssé Amalgamation
27
28 4. FRAÏSSÉ AMALGAMATION
B 1 and B 2 and all common elements of B 1 and B 2 are elements of A; note that in
this case A = B 1 ∩ B 2 . Then we call B 1 ∪ B 2 the free amalgam of B 1 , B 2 (over A).
More generally, a τ -structure C is an amalgam of B 1 and B 2 over A if for i ∈ {1, 2}
there are embeddings fi of B i to C such that f1 (a) = f2 (a) for all a ∈ A. We refer
to (A, B 1 , B 2 ) as an amalgamation diagram.
Definition 4.2.1. An isomorphism-closed class C of τ -structures
• has the free amalgamation property if for any B 1 , B 2 ∈ C the free amalgam
of B 1 and B 2 is contained in C;
• has the amalgamation property if every amalgamation diagram (A, B 1 , B 2 )
has an amalgam C ∈ C;
• is an amalgamation class if C is countable up to isomorphism, has the amal-
gamation property, and is closed under taking substructures.
Note that since we only look at relational structures here (and since we allow
structures to have an empty domain), the amalgamation property of C implies the
joint embedding property.
Example 24. The class of all finite graphs has AP: an amalgam of B 1 and B 2 is
C := B 1 ∪ B 2 (and fi : B i ,→ C the identity). Similarly, for every relational signature
τ , the class of all finite τ -structures has AP. 4
A systematic source of amalgamation classes is the following proposition.
Proposition 4.2.2. Let B be a homogeneous relational structure. Then Age(B)
is an amalgamation class.
Proof. Let A, B 1 , B 2 ∈ Age(B) so that for i ∈ {1, 2} there are embeddings
ei : A ,→ B i . By definition, there are embeddings gi : B i ,→ B. Let A0 be the sub-
structure of B with domain g1 (e1 (A)). Then the restriction of g2 ◦e2 ◦e−1 −1
1 ◦g1 to A is
0
an embedding of A into B, and by the homogeneity of B can be extended to an auto-
morphism a of B. But then the substructure C of B with domain a(g1 (B1 ))∪g2 (B2 )
is Age(B) and a ◦ g1 : B 1 → B and g2 : B 2 → B are embeddings of B 1 and B 2 into C
satisfying a ◦ g1 (e1 (a)) = g2 (e2 (a)) for all a ∈ A, showing AP.
Since we have already seen that (Q; <) is homogeneous, the age of (Q; <), which
is the class of all finite linear orders, has AP; this can also bee seen directly.
(e) Show that Age(B; E) equals the class C from Exercise 42.
(f) Show that (B; E) is isomorphic to the tournament whose vertices are
a countable dense subset S ⊆ R2 of the unit circle without antipodal
points, and where the edges are oriented in clockwise order, i.e., put
((u1 , u2 ), (v1 , v2 )) ∈ E if and only if u1 v2 − u2 v1 > 0.
(44) Show that there are permutation groups G1 , G2 on a countably infinite set
such that both G1 and G2 are isomorphic (as permutation groups) to (Q; <),
but G1 ∩ G2 = {id}.
(45) Construct a permutation group G on a set X with precisely n! orbits of
n-element subsets.
(46) Show that the random graph can be partitioned into two subsets so that
both parts are isomorphic to the random graph. Show that the same is not
true for all partitions of the random graph into two infinite subsets.
(47) Show that for every partition of the vertices of the Rado graph into two
subsets, one of the two subsets induces a subgraph of the Rado graph which
is isomorphic to the Rado graph.
(48) A permutation group G on a set D is called n-transitive if the componentwise
action of G on Dn has precisely one orbit. Construct for every n ∈ N a
permutation group on D which is n-transitive, but not (n + 1)-transitive.
(49) Give an example of a homogeneous structure with a transitive automorphism
group whose age does not have strong amalgamation.
(50) Let A be a homogeneous structure with finite relational signature. Show
that the following are equivalent.
(a) There exists a structure B with finite relational signature such that
Aut(B) = Aut(A).
(b) There exists a relation R ⊆ An such that Aut(A) = Aut(A, R).
(c) There exists a structure A with finite relational signature such that
G = Aut(A), and all relations in A have pairwise distinct entries.
(d) There exists a relation R ⊆ An such that Aut(B) = Aut(A, R) and R
has pairwise distinct entries.
Which implications between these items require the homogeneity of A?
B1
(B1)1 (B1)2
u v
u v
(B2)1 (B2)2
B2
First consider the case that there is a vertex u that lies in both B 11 and B 12 , and a
vertex v that lies in both B 12 and B 21 (see Figure 4.1 for an illustration). We claim that
in this case no vertex w from B 22 can lie inside B 1 : for otherwise, w is either in B 11 ,
in which case we have uw|v in B 1 , or in B 21 , in which case we have vw|u in B 1 . But
since u, v, w are in A, this is in contradiction to the fact that uv|w holds in B 2 . Let
C 0 ∈ C be the amalgam of B 1 and B 12 over A, which exists by inductive assumption,
and let T 0 ∈ T be its underlying tree. Now let T be the tree with root r and T 0 as
a left subtree, and the underlying tree of B 22 as a right subtree. It is straightforward
to verify that the leaf structure of T is in C, and that it is an amalgam of B 1 and B 2
over A (via the identity embeddings).
Up to symmetry, the only remaining essentially different case we have to consider
is that B11 ∪ B21 and B12 ∪ B22 are disjoint. In this case it is similarly straightforward
to first amalgamate B 11 with B 12 and B 21 with B 22 to obtain the amalgam of B 1 and
B 2 ; the details are left to the reader.
Let B be the Fraïssé-limit of C. The relation | in B is closely related to so-
called C-relations, following the terminology of [1]. C-relations became an important
concept in model theory, see e.g. [15]. They are given axiomatically; the presentation
here follows [1, 4].
A ternary relation C is said to be a C-relation on a set L if for all a, b, c, d ∈ L
the following conditions hold:
C1 C(a; b, c) ⇒ C(a; c, b);
C2 C(a; b, c) ⇒ ¬C(b; a, c);
C3 C(a; b, c) ⇒ C(a; d, c) ∨ C(d; b, c);
C4 a 6= b ⇒ C(a; b, b).
A C-relation is called proper if it satisfies
34 4. FRAÏSSÉ AMALGAMATION
Exercises.
(51) Show that the class of all finite forests (i.e., undirected graphs without cycles)
is not an amalgamation class.
(52) Prove Observation 4.5.1.
(53) Prove Proposition 4.5.3.
CHAPTER 5
Types
35
36 5. TYPES
Exercises.
(54) Let κ be an infinite cardinal. Show that a structure A realises all 1-types
over all B ⊆ A with |B| < κ if and only if A realises all n-types over all
B ⊆ A with |B| < κ.
(55) Let A, B be a τ -structures and f an isomorphism between A and B. Then
every tuple ā ∈ An has the same type in A as f (ā) in B.
(56) Prove or disprove:
• The set {x > n | n ∈ N} is a type of (Z, <)N .
• The set {x > 0} ∪ x < n | n ∈ {1, 2, 3, . . . } is a type of (Z, <)N .
• The set {x > n | n ∈ N} is realised (Q, <)N .
• The set {x > 0} ∪ x < 1/n | n ∈ {1, 2, 3, . . . } is a type of (Q, <)Q .
(57) Show that (R; 0, +) has exactly two complete 1-types and ℵ0 many complete
2-types, and that (R; 0, +, <) has exactly three complete 1-types and 2ℵ0
many complete 2-types.
If pu,v , for u ∈ Z and v ∈ N\ {0}, also contains this formula, then s·u = r ·v,
so r/s = u/v. This implies that there are at least |Q| = ℵ0 many complete
2-types.
Then there is the complete 2-type of (0, 0) and the complete 2-type of
(0, 1); they are clearly distinct and distinct from all the types of the form
pr,s for some r ∈ Z and s ∈ N \ {0}. I claim that there is exactly one more
complete 2-type of (R; 0; +), namely the type of (1, i) where i is any irrational
number. To see this, view R as a Q-vector space (of dimension 2ℵ0 ); we may
choose a basis B1 that contains the element i. Let p be a complete 2-type
(R; 0; +) which does not contain the formula x1 = 0, does not contain the
formula x2 = 0, and does not contain (1) for any r, s ∈ N. By Lemma 5.0.2,
p is realised by an element a of an elementary extension R of (R; 0; +); by
the theorem of Löwenheim-Skolem we may choose an elementary extension
of cardinality 2ℵ0 . The elementary extension can be viewed as a Q-vector
space of dimension 2ℵ0 ; using the axiom of choice, we may choose a basis B2
that contains a. Then there exists a bijection between the basis elements B1
and B2 that extends to a vector space isomorphism between (R; 0; +) and
R which shows that i satisfies p.
(c) The structure (R; 0, +, <) has the three complete 1-types realised by −1,
0, and 1, and the argument that there are all the complete 1-types can be
shown similarly as in part (a) of this solution.
(d) There are clearly at most 2ℵ0 many 2-types in (R; 0, +, <) because the sig-
nature is countable. For a ∈ R, let pa be the 2-type of the pair (a, 1) in
(R; 0, +, <). I claim that if a, b ∈ R are such that a 6= b, then pa 6= pb . It
then follows that there are |R| = 2ℵ0 many complete 2-types in (R; 0, +, <).
5.2. SATURATED STRUCTURES 37
Exercises.
(59) Spell out the proof of Lemma 5.3.4.
(60) (Exercise 13.1.5 in [25]) Show that every countable model M of Peano Arith-
metic has an elementary extension N such that for all a ∈ M and b ∈ N \ M
we have a < b (such extensions are called end extensions).
CHAPTER 6
(2) ⇒ (3). If all types of B are principal then B is atomic; the same applies to
all models of Th(B).
(3) ⇒ (1). All countable atomic models with the same theory are isomorphic by
Lemma 6.0.1.
(2) ⇒ (4). Suppose that all types are principal and let n ≥ 1. Then there exists
a sequence of formulas (φi )i∈I such that every n-type is isolated by one of those
formulas. Then Th(B) ∪ {¬φi | i ∈ I} is unsatisfiable and hence by the compactness
theorem (Theorem 2.0.1) there exists a finite F ⊆ I such that Th(B) ∪ {¬φi | i ∈ F }
is unsatisfiable. That is, in every model of Th(B), every n-tuple satisfies φi for some
i ∈ F , which shows that there are finitely many complete n-types in B.
(4) ⇒ (5). Every n-type is described by the complete n-types that contain the
n-type, so if there are finitely many complete n-types, there are finitely many n-types
in B. And this provides a finite upper bound for the number of formulas with free
variables x1 , . . . , xn .
(5) ⇒ (6). Let A be a model of Th(B), let a ∈ An , and let p be a com-
plete 1-type of (A, a). If there is a finite number of inequivalent first-order formulas
φ(x1 , x2 , . . . , xn+1 ), the conjunction over all formulas such that φ(x1 , ā) ∈ p isolates
p. So p is realised in (A, a). This shows that A is ω-saturated.
(6) ⇒ (1) follows from the fact that all countable ω-saturated structures with the
same theory are isomorphic (Theorem 5.2.2).
(2)∧(3)∧(5) ⇒ (7). Let R be an n-ary relation that is preserved by Aut(B). The
relation R is a union of orbits of n-tuples of Aut(B). It suffices to show that orbits
are first-order definable: by assumption (5), there are only finitely many inequivalent
first-order formulas, we can then define R by forming a finite disjunction. Since B is
atomic (3), if two n-tuples have the same type, then there is an automorphism that
maps one to the other, by Lemma 6.0.1. So types define orbits of n-tuples. Since
n-types of B are principal (2), it follows that the orbits of n-tuples are first-order
definable in B.
(7) ⇒ (8). Suppose that Aut(B) are infinitely many orbits of n-tuples, for some
n. Then the union of any subset of the set of all orbits of n-tuples is preserved by all
automorphisms of B; but there are only countably many first-order formulas over a
countable language, so not all the invariant sets of n-tuples can be first-order definable
in B.
(8) ⇒ (5) is immediate since automorphisms preserve first-order formulas.
The second condition in Theorem 6.0.3 provides another possibility to verify that
a structure is ω-categorical. We again illustrate this with the structure (Q; <). The
orbit of an n-tuple (t1 , . . . , tn ) from Qn with respect to the automorphism group of
(Q; <) is determined by the weak linear order induced by (t1 , . . . , tn ) in (Q; <). We
write weak linear order, and not linear order, because some of the elements t1 , . . . , tn
might be equal (that is, a weak linear order is a total quasiorder). The number of
weak linear orders on n elements is bounded by nn , and hence the automorphism
group of (Q; <) has a finite number of orbits of n-tuples, for all n ≥ 1.
Two structures A and B with the same domain A = B are called interdefinable
if every relation and operation of A is first-order definable in B, and vice versa.
Corollary 6.0.4. Two ω-categorical structures A and B with the same domain
A = B have the same automorphism group if and only if they are interdefinable.
Exercises.
(61) Is the structure (Z; {(x, y) : |x − y| ≤ 2}) ω-categorical?
6.1. ALGEBRAICITY 45
6.1. Algebraicity
The model-theoretic notion of algebraic closure generalises the notion of algebraic
closure in fields (see Chapter B).
Definition 6.1.1. Let A be a structure and let B ⊆ A. We say that a ∈ A is
algebraic over B if there exists a (τ ∪{B})-formula φ(x) such that {b ∈ A | AB |= φ(b)}
is finite and contains a. The algebraic closure of B in A is the set of elements of A
that are algebraic over B, and is denoted by aclA (B). We say that A has algebraicity
if aclA (B) 6= B for some finite B ⊆ A.
Remark 6.1.2. Let A be a countable ω-categorical structure and let B ⊆ A be
finite. Then the algebraic closure of B in A is the set of elements of A that lie in
finite orbits in Aut(A)(B) ; this is a direct consequence of Theorem 6.0.3 (also see
Exercise 64). Also note that for all finite subsets B the set aclA (B) is finite (see
Exercise 69).
Theorem 6.1.3 (See (2.15) in [7]). Let C be a homogeneous structure. Then the
age of C has strong amalgamation if and only if C has no algebraicity.
Proof. We first show that strong amalgamation of Age(C) implies no algebraic-
ity of C. Let A ⊆ C be finite, and u ∈ C \ A. We want to show that the orbit of u
in Aut(C)(A) is infinite. Let n ∈ N, A := C[A], and B := C[A ∪ {u}]. Then there
exists a strong amalgam B 0 ∈ Age(C) of B with B over A. We iterate this, taking
a strong amalgam of B with B 0 over A, showing that, because of homogeneity, there
are n distinct elements in C \ A that lie in the same orbit as u in Aut(C)(A) . Since
n ∈ N and u ∈ C \ A were chosen arbitrarily, the group Aut(C)(A) has no finite orbits
outside A.
For the other direction, we rely on the following lemma of Peter Neumann.
Lemma 6.1.4. Let G be a permutation group on D without finite orbits, and let
A, B ⊂ D be finite. Then there exists a g ∈ G with g(A) ∩ B = ∅.
Proof. The proof here is from Cameron [7], and is a nested induction. The
outer induction is on |A|. We assume the result for any set A0 with |A0 | < |A|. The
induction base A = ∅ is trivial. Suppose for contradiction that no g ∈ G with the
required property exists.
Claim. For any C with |C| ≤ |A|, there are only finitely many translates g(A)
of A that contain C. The proof of the claim is by induction on |A| − |C|. When
|A| − |C| = 0 then the only translate of A that contains C is C, and the statement
holds. So suppose that |C| < |A| and that the claim holds for all C 0 with |C 0 | > |C|.
By the outer induction hypothesis, we may assume that C ∩ B = ∅. By the inner
induction hypothesis, for each of the finitely many points b ∈ B, only finitely many
translates of A contain C ∪ {b}. So only finitely many translates of A contain C and
46 6. COUNTABLY CATEGORICAL STRUCTURES
have non-empty intersection with B. Since we assumed that every translate of A has
non-empty intersection with B, we have shown the claim.
For C = ∅, the claim implies that A has only finitely many translates, a contra-
diction to the assumption that G has no finite orbits.
We now continue with the reverse implication of Theorem 6.1.3. Let A be with-
out algebraicity, and let (A0 , B 1 , B 2 ) be an amalgamation diagram with A0 , B 1 , B 2 ∈
Age(A). By homogeneity of A we can furthermore assume that A0 , B 1 , B 2 are sub-
structures of A; that is, the structure induced by B1 ∪B2 in A is an amalgam, but pos-
sibly not a strong one. Since A has no algebraicity, Aut(A)(A0 ) has no finite orbits out-
side A0 . By the lemma, there exists g ∈ Aut(A)(A0 ) such that (B1 \A0 )∩(B2 \A0 ) = ∅.
Then the structure induced by g(B1 ∪ B2 ) is a strong amalgam of B 1 and B 2 over
A0 .
Exercises.
(67) Find an example of a countable ω-categorical structure with only one 1-type
and algebraicity.
(68) Show that if a A is structure and B ⊆ A is finite such that aclA (B) is infinite,
then A is not ω-categorical.
(69) Show that the converse of the implication in the previous exercise is false.
Exercises.
(70) Prove that the theory T 0 constructed in Remark 6.2.1 indeed has quantifier
elimination.
(71) Which of the following ω-categorical structures has quantifier elimination:
• the structure (N × {0, 1}; E) where E = ((a, b), (a, c)) | b 6= c .
• the structure (N × {0, 1, 2}; E) where
E = ((a, b), (a, c)) | (b, c) ∈ {(0, 1), (1, 0), (1, 2), (2, 1)} .
• the structure (Q+ +
0 ; <) where Q0 := {x ∈ Q | x ≥ 0}.
• (Q; O) where O := {(x, y, z) ∈ Q3 | x < y ∨ x < z}.
• (Q; B) where B := {(x, y, z) ∈ Q3 | x < y < z ∨ z < y < x}.
Exercises.
(72) Show the claim from Example 38 that (N2 ; {(x, y), (u, v) | x = u}) and
(N; =) are mutually interpretable.
(73) Show that the structure (N2 ; {((x, y), (y, z)) | x, y, z ∈ N}) is bi-interpretable
with (N; =).
(74) Show that (N2 ; {((x, y), (x, z)) | x, y, z ∈ N}, {((x, y), (z, y)) | x, y, z ∈ N})
and (N; =) are not bi-interpretable.
(75) (*) Show that two finite structures A and B have isomorphic automorphism
groups (as abstract groups) if and only if A and B are bi-interpretable.
CHAPTER 7
Definition 7.1.1. When T is a first-order theory and φ(x̄) and ψ(x̄) are formulas,
we say that φ and ψ are equivalent modulo T if T |= ∀x̄(φ(x̄) ⇔ ψ(x̄)).
Note that here the assumption that ⊥ is always part of first-order logic is impor-
tant: the first-order formula ∃x. x 6= x is preserved by all homomorphisms between
models of T , but without ⊥ it might not be equivalent to an existential positive
formula modulo T (for instance when T is the empty theory).
In the proof of Theorem 7.1.2 we need the following property of saturated struc-
tures.
53
54 7. MODEL COMPLETENESS AND QUANTIFIER-ELIMINATION
Corollary 7.1.4 (Łos-Tarski; see e.g. Corollary in 5.4.5 of [19]). Let T be a first-
order theory. A first-order formula φ is equivalent to an existential formula modulo
T if and only if φ is preserved by all embeddings between models of T .
Proof. Add for each atomic formula ψ a new relation symbol Nψ to the signature
of T , and add the sentence ∀x̄(Nψ (x̄) ⇔ ¬ψ(x̄)); let T 0 be the resulting theory. Then
every existential formula φ is equivalent to an existential positive formula in T 0 , which
can be obtained from φ by replacing negative literals ¬ψ(x̄) in φ by Nψ (x̄). Similarly,
homomorphisms between models of T 0 must be embeddings. Hence, the statement
follows from Theorem 7.1.2.
7.1.2. The Theorem of Chang-Łos-Suszko. Another important preserva-
tion theorem concerns definability by so-called ∀∃-formulas; theories consisting of
∀∃-sentences are also called inductive for reasons that will soon become clear. Such
theories will play an important role in Section 7.5.
Definition 7.1.5. A formula φ is called ∀∃ if it is of the form ∀y1 , . . . , ym . ψ
where ψ is existential.
Example 39. The property of the random graph which is described in Exam-
ple 26 can be formulated as a set of ∀∃-sentences. Note that this property described
the random graph up to isomorphism. It follows that the theory of the random graph
is equivalent to a ∀∃-theory. 4
Our next goal will be a preservation theorem that characterises whether a formula
is equivalent to a ∀∃ formula modulo a theory T (Theorem 7.1.8 below). We first need
the following simple but important and frequently used lemma.
Lemma 7.1.6. Let A and B be τ -structures and suppose that every existential
sentence true in A is also true in B. Then there exists an elementary extension C of
B and an embedding g : A → C.
Proof. Let T be the set of all quantifier-free sentences in Th(AA ). It suffices to
show that T ∪ Th(B B ) has a model C 0 , since then
0
• the map cAA 7→ cC embeds A into the τ -reduct C of C 0 , and
• C is an elementary extension of B.
If T ∪ Th(B B ) has no model, then by the compactness theorem there is a τ -formula
φ such that φ(c̄) ∈ T and {φ(c̄)} ∪ Th(B B ) has no model, and in particular B |=
¬∃ȳ.φ(ȳ). Since ∃ȳ.φ(ȳ) is existential, the assumptions imply that A |= ¬∃ȳ.φ(ȳ).
This contradicts that AA |= φ(c̄).
Similarly, one can show the following.
Lemma 7.1.7. Let T be a first-order theory, and let A be a model of the ∀∃-
consequences of T . Then A can be extended to a model B of T such that every
existential formula that holds on a tuple ā in B also holds on ā in A.
Proof. Let A0 be an expansion of A by constants such that each element of A0
is denoted by a constant symbol. It suffices to prove that T ∪ Th(A0 )qf ∪ Th(A0 )∀ has
a model B. Suppose for contradiction that it were inconsistent; then by compactness,
there exists a finite subset U of Th(A0 )∀ ∪ Th(A0 )qf such that T ∪ U is inconsistent.
Let φ be the conjunction over U where all new constant symbols are existentially
quantified. Then T ∪ {φ} is inconsistent as well. But ¬φ is equivalent to a ∀∃
formula, and a consequence of T . Hence, A |= ¬φ, a contradiction.
Theorem 7.1.8 (Chang-Łoś-Suszko Theorem). A theory T is equivalent to a ∀∃-
theory if and only if the class of models of T is closed under forming unions of chains.
56 7. MODEL COMPLETENESS AND QUANTIFIER-ELIMINATION
Exercises.
(76) Show that Th(Q; 0, 1, +, ·) is not model complete. You may use the following
facts:
• There exists a first-order formula φ(x) that defines Z in (Q; 0, 1, +, ·)
(the formula can even be chosen to be a ∀∃-formula; Poonen [24]).
• Every recursively enumerable subset of Z has an existential definition
in (Z; 0, 1, +, ·) [23].
• There are recursively enumerable subsets of Z whose complement is not
recursively enumerable (see, e.g., [3]).
of all pairs (φ, ā) where φ(x1 , . . . , xn ) is existential and ā is an n-tuple from Bi . We
construct a chain (Bji )j∈N of countable models of T as follows.
Set B0i = Bi−1 , and let j ∈ N. If there is a model C of T and an embedding
e : Bji → C such that C |= φj (e(āj )), then by the theorem of Löwenheim-Skolem
there is also a countable model C 0 of T and an embedding e0 : Bji → C 0 such that
C 0 |= φj (e0 (āj )). Identify the elements of B ji with elements of C 0 along e0 , and set
Bj+1
i := C 0 . Otherwise, if there is no such model C, we set Bj+1 i := Bji .
j
Let B i be limj∈N B i . Clearly, B := limi<ω B i is a countable model of T∀∃ . To
verify that B is T -ec, first note that B embeds into a model of T by Lemma 7.1.6.
Now suppose that g is an embedding from B to a model C of T , and suppose that
there is a k-tuple b̄ over B and an existential formula φ such that C |= φ(g(b̄)). There
is an i ∈ N such that b̄ ∈ Bik . By construction we have that B i+1 |= φ(ā). Thus, we
also have that B |= φ(b̄), which is what we had to show.
Corollary 7.4.3. Let K be a field. Then K has an algebraically closed field
extension.
Proof. Let T be the theory of fields. By Theorem 7.4.2, K embeds into a T -ec
structure F which is a model of T . Note that the theory of fields is ∀∃, so F is a field,
too. The observations in Example 44 show that F is algebraically closed.
Exercises.
(77) Prove that the model companion of the theory of equivalence relations is the
theory of equivalence relations with infinitely many infinite classes.
(78) Prove that the model companion of the theory of loopless undirected graphs
is the theory of the countably infinite random graph.
7.6. QUANTIFIER ELIMINATION 61
(79) Let us call a partial order (P ; ≤) tree-like if for all a, b, c ∈ P such that a ≤ b
and a ≤ c, then b and c must be comparable in (P ; ≤). Show that the theory
of tree-like partial orders has an ω-categorical model companion.
(80) (*) Does the theory of monoids have a model companion?
(81) (*) Does the theory of rings have a model companion?
(82) (*) Does the theory of abelian groups have a model companion?
Lemma 7.6.2. A theory T has quantifier elimination if and only if every formula
∃x.φ where φ is quantifier-free is equivalent modulo T to a quantifier-free formula.
Proof. For the interesting direction, let φ be a first-order formula. We can
assume that φ is of the form Q1 x1 . . . Qn xn .ψ for Q1 , . . . , Qn ∈ {∃, ∀} and ψ quantifier-
free. We show the statement by induction on n. For n = 0 there is nothing to be
shown. First consider the case that Qn = ∃. By assumption, ∃xn .ψ is equivalent to
a quantifier-free formula ψ 0 , so φ is equivalent to Q1 x1 . . . Qn−1 xn−1 .ψ 0 , which is in
turn equivalent to a quantifier-free formula by induction.
Now consider the case that Qn = ∀. Then ∃xn .¬ψ is by assumption equivalent
to a quantifier-free formula ψ 0 , and φ is equivalent to Q1 x1 . . . Qn−1 xn−1 .¬ψ 0 , which
is equivalent to a quantifier-free formula by induction.
Lemma 7.6.3. A theory T has quantifier elimination if and only if for all models
B 1 and B 2 of T with a common substructure A, for all a ∈ An , and quantifier-free
formula ψ(y, x1 , . . . , xn ), if B 1 |= ∃y.ψ(y, a) then B 2 |= ∃y.ψ(y, a).
62 7. MODEL COMPLETENESS AND QUANTIFIER-ELIMINATION
Proof. The forward implication is trivial. For the backward implication it suf-
fices to show that every formula φ(x1 , . . . , xn ) of the form ∃y.ψ, for ψ quantifier-free,
is equivalent modulo T to a quantifier-free formula (Lemma 7.6.2). By Lemma 7.6.1,
it suffices to show that φ is preserved by isomorphisms between substructures of mod-
els of T . So let B 1 , B 2 |= T and a ∈ B1n , and let h be an isomorphism between a
substructure that contains a and a substructure of B 2 . We identify the elements
of the substructures along h. Clearly, a quantifier-free formula holds on a in B 1 if
and only if it holds on h(a) in B 2 . The assumption then implies that B 1 |= φ(a)
if and only if B 2 |= φ(h(a)). Hence, φ is equivalent to a quantifier-free formula by
Lemma 7.6.1.
Example 46 (Ferrante and Rackoff [11], Section 3, Theorem 1). Let L be the
structure with domain Q and the signature {+, 1, ≤} where
• + is a binary relation symbols that denotes in L the usual addition over Q;
• 1 is a constant symbol that denotes 1 ∈ Q in L.
• ≤ is a binary relation symbol that denotes the usual linear order of the
rationals.
The structure L has quantifier elimination: TODO. 4
Example 47 (taken from [5]). Let H be the structure with domain Q and the
signature {1, <} ∪ {c·}c∈Q where
• 1 is a constant symbol that denotes 1 ∈ Q in L.
• < is a binary relation symbol that denotes the usual strict linear order of
the rationals.
• ·c, for c ∈ C, is a unary function symbols that denotes in L the function
x 7→ cx (multiplication by c).
The structure L has quantifier elimination: TODO. 4
Exercises.
(83) Find quantifier-free formulas that define the relations of H in L.
(84) Show that (Z; s) has quantifier elimination, where s : Z → Z is given by
x 7→ x + 1.
(85) Show that (Z; <, s) has quantifier elimination.
Exercises
(86) Find a quantifier-free formula which is equivalent to ∃x(ax2 + bx + c = 0)
over (C; +, ∗).
(87) Find a quantifier-free formula which is equivalent to the formula of the pre-
vious exercise over (R; +, ∗).
(88) Write down a first-order φ(a, b, c, d) in the language of fields which
formula
a b
states that the matrix is invertible. Find a quantifier-free formula
c d
which is equivalent to φ for any field F .
(89) (Exercise 2.3.3 in [29]) Show that a ∀∃-sentence which holds in all finite
fields is true in all algebraically closed fields.
(90) Show that in every algebraically closed field K, every injective polynomial
map of a definable subset of K n to K is surjective.
CHAPTER 8
Stability
Stability theory grew out of the attempt to understand the possible spectra of
first-order theories (Section 1.3). Today, stability theory might be thought of as a
way to classify definable sets in a structure whenever this is possible. The starting
point is that instead of counting the number of models of a first-order theory T , we
count the number of types in models of T over parameter sets of some cardinality.
This chapter presents some of the basic and fundamental facts of stability theory, and
is not yet finished.
Definition 8.0.1. Let κ be an infinite cardinal. A theory T is κ-stable if for
M
every M |= T and for every A ⊆ M , if |A| ≤ κ then |S1 (A)| ≤ κ. If T is κ-stable
for some κ, then T is called stable, and it is called unstable otherwise. A structure is
called κ-stable if its first-order theory is κ-stable.
Example 48. Let E be an equivalence relation on N with infinitely many infinite
classes. Then the structure (N; E) is stable. 4
Example 49. Let p be a prime or 0. Algebraically closed fields of characteristic p
are κ-stable for all κ. Recall that ACFp is κ-categorical for all κ > ω (Theorem 7.7.2).
We show in Theorem 8.3.2 below that uncountable categoricity implies ω-stability.
Then we show that ω-stability implies κ-stability for all κ in Proposition 8.3.3. Alter-
natively, it is easy to prove κ-stability directly, using quantifier elimination in ACF
(Theorem 7.7.1). 4
The following lemma shows that it does not matter whether we define stabil-
ity with respect to 1-types or with respect to n-types (similarly as with saturation,
Exercise 54).
Lemma 8.0.2. Let κ be an infinite cardinal and let T be κ-stable. Then T is
κ-stable for n-types, for every n ∈ N, i.e., for every model M of T and A ⊆ M of
M
cardinality at most κ we have |Sn (A)| ≤ κ.
Proof. Our proof is by induction on n. For n = 1 the statement holds by
assumption, so suppose that n > 1. Let M be a model of T and A ⊆ M of cardinality
at most κ. By Lemma 5.2.3, M has an elementary extension N that realises all
1-types over A. For an n-type p of M over A, let p0 be the 1-type that contains
all sentences of the form ∃x2 , . . . , xn .φ for φ ∈ p. By assumption, there are only
countably many 1-types of M over A. Therefore, it suffices to show that every n-type
N
p, the set Q : {q ∈ Sn (A) | q 0 = p0 } has cardinality at most κ, because ω · κ = κ.
By assumption, p0 is realised by an element b ∈ N in N . Let p00 be the (n − 1)-
type obtained from p by replacing x1 by the constant b. Then the (n − 1)-types of
M
N over A ∪ {b} are in bijective correspondence to the set of all q ∈ Sn (A) such that
0 0
q = p . By inductive assumption, there are at most κ many (n − 1)-types of N over
A ∪ {b}.
Stability has many equivalent characterisations; one is based on the binary tree
property and another one on the order property; they will be presented in Section 8.1.
65
66 8. STABILITY
Stability has another beautiful reformulation, based on item (2) of Theorem 8.1.5.
Theorem 8.1.6. T is unstable if and only if there exists a τ -formula φ(x̄, ȳ), a
model M of T , and a sequence (c̄i )i∈ω such that M |= φ(c̄i , c̄j ) ⇔ i < j.
Proof. Let ψ(x̄, ȳ) be a τ -formula so that there are a model M of T and se-
quences (āi )i∈ω and (b̄i )i∈ω of tuples from M such that M |= φ(āi , b̄j ) if and only if
i < j. Let ψ(x̄1 , x̄2 , ȳ1 , ȳ2 ) be the formula φ(x̄1 , ȳ2 ). For i ∈ ω, define ci := (ai , bi ).
Then M |= ψ(c̄i , c̄j ) if and only if M |= φ(ai , bj ) if and only if i < j. The backwards
direction is immediate, since we may take (bi )i∈ω = (ai )i∈ω = (ci )i∈ω .
Exercises
(91) Let T be the theory ACF0 of algebraically closed fields of characteristic 0.
Find an order indiscernible sequence (ai )i∈ω of distinct elements in a model
of T .
Indiscernible sequences can also be used to characterise stability.
Theorem 8.2.5. T is stable if and only if every order indiscernible sequence in
every model M of T is totally indiscernible.
Proof. If T is unstable, then by Theorem 8.1.6 there exists a τ -formula φ(x̄, ȳ),
a model M of T , and a sequence (āi )i∈ω of tuples of elements of M such that for all
i, j ∈ ω we have M |= φ(āi , āj ) ⇔ i < j. Then (āi )i∈ω is not totally indiscernible.
Conversely, suppose that T has a model A with an order indiscernible sequence
(āi )i∈ω which is not totally indiscernible. By Lemma 8.2.3 T has a model B with
an order indiscernible sequence (b̄i )i∈Q of the same Ehrenfeucht-Mostowski type as
(ai )i∈ω ; note that (b̄i )i∈Q is in particular not totally indiscernible. This means that
70 8. STABILITY
there exists a formula φ(x̄1 , . . . , x̄k ), some indices i1 , . . . , ik ∈ ω with i1 < · · · < ik and
some permutation σ of {1, . . . , k} such that B |= φ(b̄i1 , . . . , b̄ik )∧¬φ(b̄σ(i1 ) , . . . , b̄σ(ik ) ).
Any permutation on {1, . . . , k} is a product of transpositions of consecutive elements
of {1, . . . , k}, and hence we may assume that σ is a transposition that exchanges j, j +
1 ∈ {1, . . . , k} and fixes all elements of {1, . . . , k}. Then the formula φ(x1 , x2 ) given
by φ(ā1 , . . . , āj , x1 , x2 , āj+2 , . . . , āk ) orders the infinite sequence (āi ){i∈I|ij <i<ij+1 } .
Hence, Theorem 8.1.6 implies that T is unstable.
Lemma 8.2.6. Let τ be a countable signature and let M be a τ -structure generated
by a well-ordered set I of indiscernibles. Then M realises only countably many types
over every countable subset of M .
Proof. First consider the simpler case where the signature is relational. Let
M
S ⊆ M be countable. Suppose that a realises p ∈ S1 (S). Then p is determined
by the position of a with respect to the well-ordered subset S, and there are only
countably many possibilities. The general case with function symbols can be shown
similarly.
We conclude this section by showing that every countable theory has a model
that realises only countably many types over every countable subset.
Definition 8.2.7. Let τ be a signature. A Skolem theory for τ is a theory S in
a bigger signature σ such that
(1) |σ| ≤ max(|τ |, ω).
(2) S is universal.
(3) Every τ -structure can be expanded to a model of S.
(4) S has quantifier elimination.
Theorem 8.2.8. For every signature τ there exists a Skolem theory S.
Proof. We define an ascending sequence of signatures τ = σ0 ⊆ σ1 ⊆ · · · by
introducing for every quantifier-free {σi }-formula φ(x1 , . . . , xn , y) a new n-ary Skolem
function fφ and
S defining σi+1 as the union of σi and the set of these function symbols.
Define σ := i σi . Now define
S := {∀x̄(∃yφ(x̄, y) ⇒ φ(x̄, fφ (x̄)) | φ(x̄, y) quantifier-free σ-formula}.
The first two items are clearly satisfied. For the third item, let A be a τ -structure.
To define the expansion B |= S of A, suppose f ∈ σ is n-ary, and let a ∈ An .
If A |= ∃yφ(a, y) for φ quantifier-free, then choose for f B (a) any b ∈ A such that
A |= φ(a, b). Clearly, B |= S. We leave the proof that S has quantifier elimination as
an exercise.
Lemma 8.2.9. Let T be a countable τ -theory with an infinite model and let κ be
an infinite cardinal. Then T has a model of cardinality κ which realises only countably
many types over every countable subset.
Proof. Let τ be the signature of T , and consider the theory T ∗ := T ∪ S where
S is a Skolem theory for τ with signature σ (see Theorem 8.2.8). Then T ∗ is count-
able, has an infinite model and quantifier elimination. Since S is universal and has
quantifier-elimination, T ∗ is equivalent to a universal theory. By Lemma 8.2.3, T ∗
has a model N ∗ with a set of order indiscernibles I = {ai | i ∈ κ}. Then M ∗ := N ∗ [I]
is a model of T ∗ of cardinality κ. Since T ∗ has quantifier elimination, M ∗ is an ele-
mentary substructure of N ∗ and I is order indiscernible in M ∗ . By Lemma 8.2.6 the
structure M ∗ realises only countably many types over every countable subset of M ,
and the same is true for the τ -reduct M of M ∗ .
8.4. THE CLASSIFICATION PICTURE 71
Example 55. A typical theory with the SOP is the theory of (Q; <). 4
Proposition 8.4.2. If a theory T has the SOP, then it is unstable.
Proof. Let φ(x̄, y1 , . . . , ym ) be a formula, let A be a model of T , and let (bi )i∈ω
be a sequence of elements of Am that witness that T has the SOP. Let ψ(x̄, x̄0 ) be the
formula ¬∃ȳ. φ(ȳ, x̄) ∧ ∃ȳ. φ(ȳ, x̄). Then A |= ψ(bi , bj ) if and only if i < j, and hence
Theorem 8.1.6 implies that T is unstable.
Proof. If φ has the IP, then by Lemma 8.2.3 we can choose an indiscernible
sequence (bi )i∈ω of elements in Am such that {bi | i ∈ ω} is shattered by φ. Let
s ⊆ ω be the set of even integers. Then φ(x̄, bs ) satisfies the statement of the lemma:
TODO.
Theorem 8.4.7. A theory T is unstable if and only if it has the IP or the SOP.
Proof. We have already proved the backwards direction in Proposition 8.4.2
and Proposition 8.4.5.
By Theorem 8.1.6 there exists a formula φ(x̄, ȳ), a model M |= T , and a sequence
(c̄i )i∈ω such that M |= φ(c̄i , c̄j ) ⇔ i < j. By Lemma 8.2.3 we may assume that there
exists a order indiscernible sequence (b̄i )i∈Q such that TODO.
Since T is NIP, Proposition TODO
CHAPTER 9
NIP Theories
UNDER CONSTRUCTION.
9.1. VC Dimension
We have already mentioned in Remark 8.4.4 that if φ(x1 , . . . , xn ; ȳ) is dependent
in T , then by compactness there is some integer k such that no subset A of M n of
size k is shattered by φ. The maximal k for which there is some A of size k that is
shattered by φ is called the VC-dimension of φ (called after Vapnik-Chervonenkis [30]
and discovered in the context of machine learning). If φ is independent, we say that
the VC-dimension of φ is infinite.
Example 59. The formula x ≤ y is dependent in Th(Q; ≤), and has VC-
dimension 1. indeed, A = {0} is shattered by φ because we can choose bA := 0
and b∅ := −1. On the other hand, for A = {0, 1} we cannot find a b1 such that
(Q; ≤) |= ¬(0 ≤ b1 ) ∧ 1 ≤ b1 . 4
Example 60. We have seen in Example 56 that the formula φ(x; y) := ∃z. x·z = y
is independent in T := Th(N; ·), and hence has infinite VC dimension. 4
Example 61. In the countable random graph, every finite subset of vertices is
scattered by the formula E(x, y), so this formulas has infinite VC dimension. 4
If φ(x̄; ȳ) is a formula, we write φ− (ȳ; x̄) for φ(x̄; ȳ) (so it is the same formula,
but we change the role of the variables and parameters).
Lemma 9.1.1. The formula φ(x̄; ȳ) is NIP in T if and only if the formula φ− (ȳ; x̄)
is NIP in T .
Proof. Let M be an uncountable model of T . If φ(x̄; ȳ) is independent then by
compactness, there is an uncountable set A = {aJ | J ⊆ ω} which is shattered by φ,
as witnessed by tuples bI for I ⊆ A. For j ∈ ω, let Ij := {X ⊆ ω | j ∈ X}. We will
show that B := {bIj | j ∈ ω} is shattered by φ(ȳ; x̄). Let K ⊆ ω. Then we have
T |= φ− (bIj , aK ) if and only if T |= φ(aK ; bIj )
if and only if K ∈ Ij
if and only if j ∈ K .
−
Therefore φ (ȳ; x̄) is independent.
Exercises.
(92) Show that the VC dimension of φ(x̄; ȳ) can be different from the VC dimen-
sion of φ− (ȳ, x̄).
Lemma 9.1.2. The formula φ(x; y) is independent in T if and only if there is an
indiscernible sequence (ai | i ∈ ω) and a tuple b such that T |= φ(ai ; b) if and only if
i is even.
75
76 9. NIP THEORIES
family in which all sets contain x or all sets do not contain x. Some of the shattered
sets may be shattered by both subfamilies. When a set S is shattered by only one
of the two subfamilies then it contributes both to the number of shattered sets of
the subfamily and to the number of shattered sets of F. When a set S is shattered
by both subfamilies, then both S and S ∪ {x} are shattered by F, so S contributes
twice to the number of shattered sets of the two subfamilies and to the number of
shattered sets of F. Therefore, the number of shattered sets of F is at least equal to
the number shattered by the two subfamilies of F, which is at least F.
Pd−1 n
Proof of Theorem 9.2.1. Only i=0 i subsets of {1, . . . , n} ∈ O(nd ) have
cardinality less than d, so if |F| 6∈ O(nd ) then by Lemma 9.2.2 there must be a set of
cardinality at least k that is shattered, which implies that the VC dimension is larger
than d.
Corollary 9.2.3. Let T be ω-categorical. Then T is NIP if and only if (orbit
growth description).
CHAPTER 10
One goal of geometric model theory (or geometric stability theory) is the classi-
fication of structures in terms of dimension-like quantities that can be axiomatised
using ideas from combinatorial geometry, such as matroids. Key guiding examples
include vector spaces and algebraically closed fields. This chapter is not yet finished.
10.1. Pregeometries
Pregeometries and geometries are the key axiomatic notion of geometric model
theory; geometries are also known as matroids in combinatorics.
Definition 10.1.1. Let X be a set. A closure operator on X is a function
C : 2X → 2X satisfying the following conditions:
(1) C is extensive (or reflexive), i.e., U ⊆ C(U ) for every U ⊆ X;
(2) C is increasing, i.e., U ⊆ V ⇒ C(U ) ⊆ C(V ) for all U, V ⊆ X;
(3) C is idempotent (or transitive), i.e., C(C(U )) = C(U ) for every U ⊆ X.
If C(U ) = U then U ⊆ X is called closed.
We sometimes write C(x1 , . . . , xn ) instead of C({x1 , . . . , xn }).
Proposition 10.1.2. Algebraic closure in a structure A (see Definition 6.1.1) is
a closure operator on A.
Proof. It is obvious that aclA is extensive and increasing. To show that it is
idempotent, suppose that b1 , . . . , bk ∈ aclA (C) and that A |= ψ(a, b1 , . . . , bk ) where
{y ∈ A | A |= ψ(y, b1 , . . . , bk )} has exactly m ∈ N elements. For i ∈ {1, . . . , k}, let
φi (x, c̄i ), for a tuple c̄i of elements from C, be a formula witnessing that bi ∈ aclA (C).
Let φm,ψ (x1 , . . . , xk ) be a formula that says that {y ∈ A | A |= ψ(y, x1 , . . . , xk )} has
at most m elements. Then
k
^
∃x1 , . . . , xk (ψ(y, x1 , . . . , xk ) ∧ φm,ψ (x1 , . . . , xk ) ∧ φi (xi , c̄i ))
i=1
is a formula with parameters in C that witnesses that c ∈ aclA (C), proving idempo-
tency of aclA .
Example 65. In algebraically closed fields K, we have a ∈ aclK (A) precisely if a
is algebraic (in the field-theoretic sense) over the subfield F of K generated by A (see
Appendix B.4.3). This follows from quantifier elimination of ACF (Theorem 7.7.1):
• if a is algebraic in the field-theoretic sense over F , then there exists a poly-
nomial p ∈ F [x] such that p(a) = 0. Since {b ∈ K | K |= p(b) = 0} is finite
and contains a. This shows that a is algebraic in the model-theoretic sense.
• if a ∈ aclK (A), there exists a first-order formula φ(x) such that {b ∈ K |
K |= φ(b)} is finite and contains a. Since ACF has quantifier elimination,
φ can be written as a finite disjunction of finite conjunctions of equalities
p(x) = 0 and inequalities p(x) 6= 0. So a must satisfy at least one of the
disjuncts. This disjunct cannot consist entirely of inequalities since in this
79
80 10. GEOMETRIC MODEL THEORY
Example 67. Let V be a vector space over a field F . For A ⊆ V , let C(A)
be the linear span of A, i.e., the set of all elements that can be written as linear
combinations u1 a1 + · · · + un an for a1 , . . . , an ∈ A and u1 , . . . , un ∈ F . (The linear
span is the smallest linear subspace of V that contains A.) Then C is an example of
a pregeometry. It is not a geometry since 0 ∈ C(∅). 4
Example 68. Let V be a vector space over a field F . For A ⊆ V , let C(A) be the
affine hull of A, i.e., the set of all elements that can be written as u1 a1 + · · · + un an
where a1 , . . . , an ∈ A and u1 + · · · + un = 1 u1 , . . . , un ∈ F . The affine hull is the
smallest affine subspace of V that contains A. Then C is a geometry. 4
Example 69. Let V be a vector space over a field F , and let X = P (V ) be the
projective space over V , i.e, the set of equivalence classes of the equivalence relation
∼ defined on V \ {0} by
Example 70. Let F be an algebraically closed field. Then aclF (recall that
model-theoretic and field-theoretic algebraic closure coincide, see Example 65) is a
pregeometry on F . To prove the exchange property, let A ⊆ F be a subset, and let K
be the subfield of F with domain aclF (A). Let a, b ∈ F be such that a ∈ aclF ({b}∪A).
So there is a non-zero polynomial p ∈ K[x, y] with p(a, b) = 0. This implies that
b ∈ aclF ({a} ∪ A). The operator aclF is not a geometry since 1 ∈ C(∅). 4
10.1. PREGEOMETRIES 81
Lemma 10.1.9. C is locally modular if and only if for all p ∈ X \ C(∅) the
relativised pregeometry C{p} is modular.
Example 74. Suppose that K |= ACF has transcendence degree at least five
over the prime field. Then aclK is not locally modular. By Lemma 10.1.9 it suffices
to show that for some p ∈ X \ aclK (∅) the relativised pregeometry C := (aclK ){p}
is not modular. Choose an independent set {p, a, b, c, d}, set x := a−cb−d , and let A :=
C(a, b) and B := C(x, a − b). Then C(A ∪ B) = C(a, b, x) has dimension 3, while
dim(A) = dim(B) = 2. To show that C is not modular it is sufficient to prove that
dim(A ∩ B) 6= 1.
Claim. A ∩ B = C(∅). Let z ∈ A ∩ B. We claim that there exists an α ∈ Aut(K)
that maps (a, b, c, d) to (c, d, a, b) and that fixes B pointwise. First note that any α
that maps (a, b, c, d) to (c, d, a, b) satisfies α(x) = x and
cb − cd − da + dc ab − da − ba + bc
α(a − bx) = c − dx = = = a − bx.
b−d b−d
Hence, α fixes {x, a−bx}, and can be extended to to an automorphism of K that fixes
B = C(x, a − bx) pointwise. In particular, α(z) = z. Furthermore, z ∈ A = C(a, b)
implies z = α(z) ∈ C(c, d). Consequently, z ∈ C(a, b) ∩ C(c, d) = C(∅). 4
The sets A and B are independent over C if for all sets A0 ⊆ A and B 0 ⊆ B, if
each A0 and B 0 is independent over C, then A0 and B 0 are disjoint and their union is
independent over C. Note that this relation is symmetric. The following is then easy
to see.
Exercises.
(94) Show that every pregeometry is submodular, i.e.,
(95) The set of all closed subsets of a closure operator forms a lattice, where the
infimum ∧ is intersection and the supremum ∨ of X and Y is C(X ∪ Y ).
Show that a pregeometry is modular if and only if the lattice of closed sets
is modular, i.e., for all closed A, B, C
A ⊆ C ⇒ A ∨ (B ∧ C) = (A ∨ B) ∧ C.
Set Theory
In this appendix chapter we freely follow Appendix A of the book of Tent and
Ziegler [29]. See [20] for a more detailed introduction to set theory.
A.3. Ordinals
A well-ordering of a set A is a linear order < on A such that any non-empty subset
of A contains a smallest element with respect to <. The usual ordering of the set N
is an example of a well-ordering, and the usual ordering of Z and the usual ordering
of the non-negative rational numbers are non-examples. For a ∈ A, the predecessors
of a are the elements b ∈ A with b < a.
Definition A.3.1. An ordinal is a well-ordered set in which every element equals
its set of predecessors.
The well-ordering of an ordinal α is given by the relation ∈: if β, γ ∈ α, then
β < α ⇔ β is a predecessor of α
⇔β∈α
Hence, an ordinal is uniquely given by the set of its elements. Suppose that α is an
ordinal, and γ ∈ β ∈ α. Then γ is a predecessor of β, and hence must be in α. That
is, ordinals α are transitive: if β ∈ α then β ⊆ α.
An element β of an ordinal α is again an ordinal number: Since β ⊆ α, we obtain
a well-ordering of β by restricting the well-ordering of α to β. Moreover, if γ, γ 0 ∈ β,
then γ < γ 0 iff γ ∈ γ 0 so β is indeed an ordinal.
Proposition A.3.2. Every structure (A; <) where A is well-ordered by < is iso-
morphic to (α; ∈) where α is an ordinal.
A.4. CARDINALS 89
A.4. Cardinals
Two sets A and B have the same cardinality (|A| = |B|) if there exists a bijection
between them. By the well-ordering theorem, every set has the same cardinality as
some ordinal. We call the smallest such ordinal the cardinality |A| of A. Ordinals
occurring in this way are called cardinals. An ordinal α is a cardinal if and only if all
smaller ordinals do not have the same cardinality.
Notes.
• All natural numbers and ω are cardinals.
90 A. SET THEORY
Note that
(κλ )µ = κλ·µ .
Cantor’s well-known diagonalization argument shows that
2κ > κ .
In particular, there is no largest cardinal. Cantor’s result also follows from König’s
theorem below for κi := 1 and λi := 2 for all i ∈ I := ω.
Theorem A.4.1 (König’s theorem). Let (κi )i∈I and (λi )i∈I be sequences of car-
dinals. If κi < λi for all i ∈ I, then
X Y
κi < λi .
i∈I i∈I
P Q
Proof. We first show that i∈I κi ≤ i∈I λi . Choose pairwise disjoint sets
(Ai )i∈I and (Bi )i∈I such thatS|Ai | = κi , Q|Bi | = λi , and Ai ⊂ Bi for all i ∈ I. We
will construct an injection f : i∈I Ai → i∈I Bi .SChoose di ∈ Bi \ Ai for each i ∈ I
(here we use the Axiom of Choice). For x ∈ A := i∈I Ai , define
(
x if x ∈ Ai
f (x) := (ai )i∈I where ai :=
di otherwise.
To show the injectivity of f , let x, y ∈ A be distinct. Let i ∈ I be such that x ∈ Ai .
If y ∈ Ai then f (x)i = x 6= y = f (y)i . If y ∈ / Ai then f (x)i = x 6= di = f (y)i since
x ∈ Ai but di ∈ Bi \ Ai . So in bothP Q6= f (y).
cases, f (x)
Suppose for contradiction that i∈I κi = i∈I λi . Then we can find sets (Xi )i∈I
with |Xi | = κi such that
Y [
B := Bi = Xi .
i∈I i∈I
For each i ∈ I, define
Yi := {ai | a ∈ Xi } .
For every i ∈ I there exists bi ∈ Bi \ Yi because |Yi | ≤ |Xi | = κi < λi = |Bi |. Now
define Y
b := (bi )i∈I ∈ Bi .
i∈I
Let j ∈ I. Then bj ∈
/SYj by the choice of bj , and hence b ∈
/ Xj by the definition of Yj .
This shows that b ∈
/ i∈I Xi , a contradiction.
A.4. CARDINALS 91
We write κ+ for the smallest cardinal greater than κ, the successor cardinal of κ.
Positive cardinals which are not successor cardinals are called limit cardinals. There
is an isomorphism between the class of ordinals and the class of all infinite cardinals,
which is denoted by
α 7→ ℵα
and can be defined inductively by
ω
if α = 0
ℵα := ℵ+ β if α = β + 1
S
ℵ
β<α β if α is a limit ordinal.
S
Note that if (κi )i∈I is a family of cardinals, then κ := i∈I κi is again a cardinal:
S
• if there is an i ∈ I such that κj ≤ κi for all j ∈ I, then i∈I κi = κi and
the statement is true;
• otherwise, for every i ∈ I there is a j ∈ I with κi < κj . For each ordinal α
with α < κ we have that α ∈ κ and hence α ∈ κi for some i ∈ I. By the
above, there is a j ∈ I such that |α| ≤ κi < κj ≤ |κ|. Thus, every ordinal
smaller than κ has smaller cardinality than κ, and κ is a cardinal.
Theorem A.4.2. Let κ be an infinite cardinal. Then
(1) κ · κ = κ.
(2) κ + λ = max(κ, λ).
(3) κκ = 2κ .
Proof. For ordinals α, β, α0 , β 0 , define (α, β) < (α0 , β 0 ) iff
(max(α, β), α, β) <lex (max(α0 , β 0 ), α0 , β 0 )
where lex is the lexicographical ordering on triples of ordinals. Since this is a well-
ordering, there is a unique order-preserving bijection f between pairs of ordinals and
ordinals by Proposition A.3.2.
Claim. If κ is an infinite cardinal, then f maps κ × κ to κ, and hence κ · κ = κ.
The proof of the claim is by induction on κ. For α, β ∈ κ let Pα,β be the set of
predecessors of (α, β). Note that:
• Pα,β is contained in δ × δ with δ = max(α, β) + 1.
• Since κ is infinite and α, β < κ, the cardinality of δ is smaller than κ.
IA
• By inductive assumption |Pα,β | ≤ |δ × δ| = |δ| · |δ| = |δ| < κ.
Hence, f (α, β) < κ since f is an order isomorphism and thus f (α, β) ∈ κ.
Now (2) and (3) are simple consequences. Let µ := max(κ, λ).
µ≤κ+λ≤µ+µ≤2·µ≤µ·µ=µ
2κ ≤ κκ ≤ (2κ )κ = 2κ·κ = 2κ
Corollary A.4.3. For a non-empty set A, the set A<ω := An has cardi-
S
n∈ω
nality max(|A|, ℵ0 ).
Proof. Clearly, |A| ≤ |A<ω | and ℵ0 ≤ |A<ω |. On the other hand,
X
|A<ω | = |A|n ≤ (sup |A|n ) · ℵ0 = max(|A|, ℵ0 ).
n∈ω
n∈ω
The final equality holds because clearly supn∈ω |A|n = 1 if |A| = 1, supn∈ω |A|n = ℵ0
if 2 ≤ |A| ≤ ℵ0 , and |A| if |A| ≥ ℵ0 by Theorem A.4.2 (1).
92 A. SET THEORY
The Continuum Hypothesis (CH) states that ℵ1 = 2ℵ0 , that is: there is no cardinal
lying strictly between ω and the cardinality |R| of the continuum. The Generalised
Continuum Hypothesis (GCH) states that κ+ = 2κ for all infinite cardinals κ. As
with CH, the GCH is known to be independent of ZFC, that is, there are models of
ZFC where GCH is true , and models of ZFC where GCH is false (assuming that ZFC
is consistent; see [20]).
Let A be a set that is linearly ordered by <. A subset B ⊆ A is called cofinal if for
every a ∈ A there is some b ∈ B with a ≤ b. Any linear order contains a well-ordered
cofinal subset.
Definition A.4.4. The cofinality cf(A) of A is the smallest ordinal that is order-
isomorphic to a well-ordered cofinal subset of A.
Examples:
• If A has a greatest element, then the cofinality is one since the set consisting
only of the greatest element is cofinal and must be contained in any other
cofinal subset of A.
• A subset S of N is cofinal if and only if S is infinite, and thus cf(ω) = ω.
• cf(2ω ) > ω; see Exercise 92.
Lemma A.4.5. For linearly ordered A we have cf(cf(A)) = cf(A).
Proof. If {aα | α < β} is cofinal in A and {α(ν) | ν < µ} is cofinal in β, then
{aα(ν) | ν < µ} is cofinal in A.
Lemma A.4.6. For linearly ordered A we have that cf(A) is a cardinal.
Proof. Suppose for contradiction that cf(A) is not a cardinal. Choose a surjec-
tive map f from | cf(A)| to cf(A). This maps provides a cofinal sequence in cf(A) of
length | cf(A)|, and therefore cf(cf(A)) ≤ | cf(A)| < cf(A). This is in contradiction to
cf(cf(X)) = cf(X) from Lemma A.4.5.
A cardinal κ is regular if cf(κ) = κ, and singular otherwise. As we have seen
above, ℵ0 is an example of a regular cardinal. Lemma A.4.6 and Lemma A.4.5 show
that cf(A) is a regular cardinal. Assuming the axiom of choice, we also have the
following.
Proposition A.4.7. Successor cardinals ℵα+1 are regular.
Proof. Suppose for contradiction that cf(ℵα+1 ) ≤ ℵα . Then ℵα+1 would be
the union of at most ℵα sets of cardinality at most ℵα , contradicting item (1) in
Theorem A.4.2.
Exercises.
(91) Show that ℵω is singular.
(92) Show that cf(2ω ) >Pω.
Hint: write 2ω as ω
ν<µ κν for µ := cf(2 ), and apply König’s theorem
(Theorem A.4.1) with (κν )ν<µ and (λν )ν<µ where λν := 2ω for all ν < µ.
APPENDIX B
In this section we collect some standard material from classical algebra that is
typically taught in the second year; we freely follow Lang [21] and Dummit and
Foote [10]. My impression is that Dummit and Foote is the more elementary and
detailed presentation; on the other hand, the proof of Steinitz’ theorem (Corol-
lary B.4.12) is omitted there, but can be found in Lang’s book, which is also the
reference of Tent and Ziegler.
A commutative ring is a structure R over the signature τRing = {+, −, 0, 1, ·}
such that the reduct (R; +, −, 0) of R is an Abelian group and such that the following
axioms hold:
∀x, y, z. (xy)z = x(yz)
∀x. 1x = x
∀x, y. xy = yx
∀x, y, z. x(y + z) = xy + xz
Let R be a ring. A nonzero element a of R is called a zero divisor if there is an
b ∈ R \ {0} such that ab = 0. An element u ∈ R is called a unit if there is some v ∈ R
such that uv = vu = 1. In this terminology, a field is a commutative ring such that
0 6= 1 in which every nonzero element is a unit. Observe that a zero divisor cannot
be a unit.
Lemma B.0.1. Let R be a ring and a, b, c ∈ R be such that a is not a zero divisor.
If ab = ac then a = 0 or b = c.
Proof. If ab = ac then a(b − c) = 0 so a = 0 or b − c = 0.
It is easy to verify that this is well defined, and that multiplication has a unit, namely
1
1 , and is associative. Addition is defined by the rule
a a0 s0 a + sa0
+ 0 := .
s s ss0
It is easy to verify that this is well defined, with neutral element 01 , and that addition
and multiplication yield a ring structure.
Sometimes, the term quotient field is used, but we try to avoid this to avoid
confusion with the quotients that will be introduced in the next section.
Note that the map fQ : R → Q given by f (a) := a1 is an (injective) ring homo-
morphism from R to Q. Every element of fQ (R \ {0}) is invertible in Q: the inverse
of 1s is 1s . We often identify the elements of R with elements of Q along fQ , and hence
view R as a subring of Q.
Proposition B.1.2. Let R and B be rings and h : R → B be a homomorphism
such that all elements of h(R) are invertible in B. Let Q be the field of fractions of
R. Then there exists a unique homomorphism u from Q to B such that u ◦ fQ = h.
Example 81. If R is an integral domain, then the ring R[x1 , . . . , xn ] of polyno-
mials over R is also an integral domain: this comes from the fact that if n and m are
the degrees of the polynomials f ∈ R[x] and g ∈ R[x], then the degree of f g ∈ R[x]
is n + m, so there are no nonzero zero divisors. If K is the field of fractions of R,
then the field of fractions of R[x1 , . . . , xn ] is denoted by K(x1 , . . . , xn ); its elements
are called rational functions. A rational function can be written as fg(x̄) (x̄)
where f and
g are polynomials. 4
Fields:
ℚ, ℂ, ℤp
Euclidean domains:
ℤ, ℚ[x], ℤ[i]
Commutative rings:
ℤxℤ, ℤ6
Rings:
ℚ#$#
Exercises.
(93) A unique factorisation domain is an integral domain in which every non-
zero non-unit element can be written as a product of irreducible elements,
uniquely up to order and units. Show that every principal ideal domain is a
unique factorisation domain. See Figure B.1.
(94) Show that every principal ideal domain is a Dedekind domain, i.e., an integral
domain in which every nonzero proper ideal factors into a product of prime
ideals.
(95) Show that Q[x, y] is an example of a unique factorisation domain which is
not a Dedekind domain.
(96) Show that a Dedekind domain is a unique factorisation domain if and only
if it is a principal ideal domain.
98 B. RINGS AND FIELDS
The proof of Lemma B.4.3 shows that there exists a unique monic polynomial in
F [x] that is irreducible and has a as a root, and which is called the minimal polynomial
of a over F .
Proposition B.4.4. Let E be a field extension of F . Then a ∈ E is algebraic
over F if and only if F (a) is a finite extension of F .
Proof. If a is algebraic then the degree of F (a) over F is the degree of the
minimal polynomial for a over F . Conversely, suppose that F (a) has degree n
over F . Then 1, a, a2 , . . . , an cannot be linearly independent over F . So there are
b0 , b1 , . . . , bn ∈ F , not all zero, such that b0 + b1 a + · · · + bn an = 0. Hence, a is the
root of the non-zero polynomial b0 + b1 x + · · · + bn xn .
A literal converse of Lemma B.4.5 is false: for example the subfield of C consisting
of all algebraic numbers is an infinite algebraic extension of Q. The following lemma
prepares an exact characterisation of finite field extensions (Proposition B.4.7 below).
Lemma B.4.6. Let F be a subfield of K, and K a subfield of L. Then [L : F ] =
[L : K][K : F ].
Proof. Suppose first that [L : K] = m and [K : F ] = n are finite. Let a1 , . . . , am
be a basis for L over K and let b1 , . . . , bn be a basis for K over F . Then every c ∈ L
can be written as a linear combination
u1 a1 + · · · + um am
100 B. RINGS AND FIELDS
for some coefficients vij ∈ F . Then defining the elements ui ∈ K by equation 3 above
this equation could be rewritten to u1 a1 + · · · + um am = 0, and since a1 , . . . , am form
a basis for L over K this implies that u1 = · · · = um = 0. So we have vi,1 b1 + · · · +
vi,n bn = 0, and since b1 , . . . , bn for a basis for K over F this implies that vi,j = 0 for
all i ≤ m and j ≤ n. Hence the ai bj are linearly independent over F , so they form a
basis for L over F and [L : F ] = mn as claimed.
If one of [L : K] or [K : F ] is infinite, then [L : F ] is infinite, too.
L
g
B
E
k(K) k
e(a) K
Ramsey’s theorem
We now state Ramsey’s theorem in it’s full strength; the proof is similar to the
proof of Theorem C.0.1 shown above.
Theorem C.0.2 (Ramsey’s theorem). Let s, c ∈ N. Then N → (ω)sc .
A proof of Theorem C.0.2 can be found in [19] (Theorem 5.6.1); for a broader
introduction to Ramsey theory see [14]. Compactness and Theorem C.0.2 imply the
following finite version of Ramsey’s theorem (exercise). In fact, full compactness is
not needed, as we will see in our proof which is based on Kőnig’s tree lemma.
A walk in a graph (V, E) (see Example 1) is a sequence x0 , x1 , . . . , xn ∈ V with the
property that (xi , xi+1 ) ∈ E for all i ∈ {1, . . . , n−1}. A walk is a path if all its vertices
are distinct. A cycle is a walk of length at least three of the form x0 , x1 , . . . , xn = x0
such that x1 , . . . , xn are pairwise distinct. A tree is a connected graph (V, E) (see
Section 1.1.6) without cycles. The degree of a vertex u ∈ V is the number of vertices
v ∈ V such that {u, v} ∈ E.
105
106 C. RAMSEY’S THEOREM
Lemma C.0.3 (Kőnig’s Tree Lemma). Let (V, E) be a tree such that every vertex
in V has finite degree, and let v0 ∈ V . If there are arbitrarily long paths that start in
v0 , then there is an infinitely long path that starts in v0 .
Proof. Since the degree of v0 is finite, there exists a neighbour v1 of v0 such
that arbitrarily long paths start in v0 and continue in v1 (by the infinite pigeon-
hole principle). We now construct the infinitely long path by induction. Suppose
we have already found a sequence v0 , v1 , . . . , vi that can be continued to arbitrarily
long paths in (V, E). Since the degree of vi is finite, vi must have a neighbour vi+1 in
V \{v0 , v1 , . . . , vi } such that v0 , v1 , . . . , vi+1 can be continued to arbitrarily long paths
in (V, E). In this way, we define an infinitely long path v0 , v1 , v2 , . . . in (V, E).
Exercises.
(97) Let (X; <) be a partially ordered set on a countably infinite set X. Show
that (X; <) contains an infinite chain, or an infinite antichain.
(98) 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 .
Ramsey’s theorem can be used to prove the Erdös-Makkai theorem which we need
in the proof of Theorem 8.1.5 about equivalent characterisations of stability.
Theorem C.0.5 (Erdős-Makkai). Let A be an infinite set and let S be a set of
subsets of A with |A| < |S|. Then there are a sequence (ai )i∈I of elements of A and
a sequence (Si )i∈I of elements of S such that one of the following holds:
(1) for all i, j ∈ ω ai ∈ Sj ⇔ i < j
(2) for all i, j ∈ ω ai ∈ Sj ⇔ j < i.
Proof. There are at most |A| many pairs of finite subsets of A. So we may
choose S 0 ⊆ S with |S 0 | = |A| such that any two finite subsets A0 and A1 of A that
can be separated by a set in S can also be separated by a set in S 0 . By assumption
there is some S ∗ ∈ S which is not a Boolean combination of elements of S 0 , because
there are at most |B| different Boolean combinations of elements of S 0 .
C. RAMSEY’S THEOREM 107
[1] S. A. Adeleke and P. M. Neumann. Relations related to betweenness: their structure and auto-
morphisms, volume 623 of Memoirs of the AMS. American Mathematical Society, 1998.
[2] G. Ahlbrandt and M. Ziegler. Quasi-finitely axiomatizable totally categorical theories. Annals
of Pure and Applied Logic, 30(1):63–82, 1986.
[3] M. Bodirsky. Introduction to mathematical logic, 2023. Course notes, TU Dresden, https:
//wwwpub.zih.tu-dresden.de/~bodirsky/Logic.pdf.
[4] 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.
[5] M. Bodirsky, M. Mamino, and C. Viola. Piecewise linear valued CSPs solvable by linear pro-
gramming relaxation. ACM Transactions of Computational Logic, 23(1):1–35, 2022. Preprint
arXiv:1912.09298.
[6] S. Braunfeld, A. Dawar, I. Eleftheriadis, and A. Papadopoulos. Monadic NIP in monotone
classes of relational structures. In 50th International Colloquium on Automata, Languages, and
Programming, ICALP 2023, July 10-14, 2023, Paderborn, Germany, pages 119:1–119:18, 2023.
[7] P. J. Cameron. Oligomorphic permutation groups. Cambridge University Press, Cambridge,
1990.
[8] P. J. Cameron. Homogeneous permutations. Electronic Journal of Combinatorics, 9(2), 2002.
[9] A. Chernikov. Lecture notes on stability theory. AMS Open Math Notes, 2019.
[10] Dummit and Foote. Abstract Algebra. Wiley, 2004. Third edition.
[11] J. Ferrante and C. Rackoff. A decision procedure for the first order theory of real addition with
order. SIAM Journal on Computing, 4(1):69–76, 1975.
[12] R. Fraïssé. Sur l’extension aux relations de quelques propriétés des ordres. Annales Scientifiques
de l’École Normale Supérieure, 71:363–388, 1954.
[13] R. Fraïssé. Theory of Relations. Elsevier Science Ltd, North-Holland, Amsterdam, 1986.
[14] R. L. Graham, B. L. Rothschild, and J. H. Spencer. Ramsey theory. Wiley-Interscience Series
in Discrete Mathematics and Optimization. John Wiley & Sons, Inc., New York, 1990. Second
edition.
[15] D. Haskell and D. Macpherson. Cell decompositions of C-minimal structures. Annals of Pure
and Applied Logic, 66:113–162, 1994.
[16] C. W. Henson. Countable homogeneous relational systems and categorical theories. Journal of
Symbolic Logic, 37:494–500, 1972.
[17] C. W. Henson. Model theory. Class Notes for Mathematics 571, University of Illinois, 2010.
[18] W. Hodges. Model theory. Cambridge University Press, Cambridge, 1993.
[19] W. Hodges. A shorter model theory. Cambridge University Press, Cambridge, 1997.
[20] T. Jech. Set theory. Springer Monographs in Mathematics. Springer-Verlag, Berlin, 2003. The
third millennium edition, revised and expanded.
[21] S. Lang. Algebra. Springer, 2002. Revised third edition.
[22] D. Marker. Model Theory: An Introduction. Springer, New York, 2002.
[23] Y. V. Matiyasevich. Hilbert’s Tenth Problem. MIT Press, Cambridge, Massachusetts, 1993.
[24] B. Poonen. Characterizing integers among rational numbers with a universal-existential formula.
American Journal of Mathematics, pages 675–682, 2009.
[25] P. Rothmaler. Introduction to Model Theory. CRC Press, 2000.
[26] D. Saracino. Model companions for ℵ0 -categorical theories. Proceedings of the American Math-
ematical Society, 39:591–598, 1973.
[27] J. Schreier and Stanisław Marcin Ulam. Über die Permutationsgruppe der natürlichen Zahlen-
folge. Studia Mathematica, 4:134–141, 1933.
[28] P. Simon. A Guide to NIP Theories. Lecture Notes in Logic. Cambridge University Press, 2015.
[29] K. Tent and M. Ziegler. A course in model theory. Lecture Notes in Logic. Cambridge University
Press, 2012.
109
110 BIBLIOGRAPHY