Models Set Theory
Models Set Theory
STEFAN GESCHKE
Contents
1. First order logic and the axioms of set theory 2
1.1. Syntax 2
1.2. Semantics 2
1.3. Completeness, compactness and consistency 3
1.4. Foundations of mathematics and the incompleteness theorems 3
1.5. The axioms 4
2. Review of basic set theory 5
2.1. Classes 5
2.2. Well-founded relations and recursion 5
2.3. Ordinals, cardinals and arithmetic 6
3. The consistency of the Axiom of Foundation 8
4. Elementary Submodels, the Reflection Principle, and the Mostowski
Collapse 10
5. Constructibiltiy 13
5.1. Definability 13
5.2. The definition of L and its elementary properties 14
5.3. ZF in L 15
5.4. V = L? 15
5.5. AC and GCH in L 17
6. Forcing 20
6.1. Partial orderings, dense sets and filters 20
6.2. Generic extensions 22
6.3. The forcing relation 24
6.4. ZFC in M [G] 28
7. CH is independent of ZFC 30
7.1. Forcing CH 30
8. Forcing ¬CH 31
8.1. The countable chain condition and preservation of cardinals 32
8.2. Nice names and the size of 2ℵ0 33
9. Martin’s Axiom 35
9.1. Martin’s Axiom and Souslin’s Hypothesis 35
9.2. The Baire Category Theorem and Martin’s Axiom 37
9.3. More consequences of MA 40
9.4. Iterated forcing 42
9.5. Long iterations 45
References 47
1
2 STEFAN GESCHKE
(3) is a very precise statement. In particular, if (3) holds for a specific system of
axioms, as a true mathematical statement it should be provable from the axioms.
However, if the system of axioms was inconsistent, then it would prove its own
consistency. It follows that a proof of (3) would actually not increase our believe
in the consistency of the system of axioms.
In fact, Gödel has shown that for every system of axioms satisfying (1) and
(2) (in a precise sense), (3) is not provable from the axioms. This is the Second
Incompleteness Theorem.
Gödels First Incompleteness Theorem says that if ZFC (or any similar theory)
is consistent, then there are sentences ϕ such that neither ϕ nor ¬ϕ follow from it.
Such a sentence ϕ is called independent over ZFC. The main goal of this course is
to show that the Continuum Hypothesis (CH) is independent over ZFC.
1.5. The axioms. The axioms of ZFC are the following (we use some obvious
abreviations):
(1) Set Existence.
∃x(x = ∅)
(2) Extensionality.
∀x∀y(∀z(z ∈ x ↔ z ∈ y) → x = y)
(3) Foundation.
∀x(x 6= ∅ → ∃y ∈ x∀z ∈ x(z 6∈ y))
(4) Separation (really a scheme of axioms). For every formula
ϕ(x, y1 , . . . , yn ) without a free occurrence of y,
∀y1 . . . ∀yn ∀z∃y∀x(x ∈ y ↔ x ∈ z ∧ ϕ(x, y1 , . . . , yn ))
is an axiom.
(5) Pairing.
∀x∀y∃z(x ∈ z ∧ y ∈ z)
(6) Union.
∀F ∃A∀Y ∈ F (Y ⊆ A)
(7) Replacement (again a scheme of axioms). For every formula
ϕ(x, y, y1 , . . . , yn ), without a free occurence of y,
∀y1 . . . ∀yn ∀A(∀x ∈ A∃!yϕ(x, y, y1 , . . . , yn )
→ ∃Y ∀x ∈ A∃y ∈ Y ϕ(x, y, y1 , . . . , yn ))
is an axiom. Here ∃! means “there is exactly one”.
(8) Infinity.
∃x(∅ ∈ x ∧ ∀y ∈ x(y ∪ {y} ∈ x))
(9) Power set.
∀x∃y∀z(z ⊆ x → z ∈ y)
(10) Choice.
∀x∃R(R is a well-ordering of x)
For the Axiom of Choice recall that a binary relation ≤ on a set X is well-
ordering if ≤ is a linear order on X and every non-empty set Y ⊆ X has a minimal
element with respect to ≤. (See also Definition .)
Exercise 1.7. Write out the formulas ∃!xϕ, x = ∅, and x = y ∪ z.
Exercise 1.8. Write out the formula “R is a well-ordering of x”.
Note that since Separation and Replacement are schemes and not just single
axioms, ZFC consists of infinitely many axioms.
MODELS OF SET THEORY 5
least ordinal that is of the same size as a. It follows from the Axiom of Choice that
every set has a cardinality, and the cardinality of a set is always a cardinal. Every
infinite cardinal is a limit ordinal.
If κ and λ are cardinals, then κ + λ is the size of ({0} × κ) ∪ ({1} × λ) and κ · λ
is the size of κ × λ. κλ is the size of the set of all functions from λ to κ.
Theorem 2.7. If κ and λ are cardinals and at least one of them is infinite, then
κ × λ = κ + λ = max(κ, λ).
The finite ordinals happen to be cardinals and are denoted by 0, 1, 2, . . . . Recall
that 0 = ∅, 1 = {0}, 2 = {0, 1} and so on. The set of all finite ordinals is ω, the
first infinite ordinal. For every ordinal α the α-th infinite cardinal is denoted by
ℵα . In other words, the first infinite cardinals are ω = ℵ0 , ℵ1 , ℵ2 and so on.
ℵ0 is the size of the set N = ω. 2ℵ0 is the size of the set R of real numbers which
is the same as the size of P(ω).
Theorem 2.8. For every cardinal κ, 2κ > κ.
By this theorem, the first candidate for 2ℵ0 is ℵ1 . The statement 2ℵ0 = ℵ1 is
known as the Continuum Hypothesis (CH). The main goal of this course is to show
that CH can be neither proved nor refuted from the axioms of ZFC.
Note that Theorem 2.8 implies that for every cardinal there is a larger cardinal.
If κ is a cardinal, then the smallest cardinal (strictly) larger than κ is denoted by
κ+ . The Generalized Continuum Hypothesis (GCH) is the statement that for every
infinite cardinal κ we have 2κ = κ+ .
Let α be an ordinal. A ⊆ α is cofinal (in α) iff for all β < α there is γ ∈ A such
that β ≤ γ. The cofinality cf(α) of an ordinal α is the least size of a cofinal subset
of α. A cardinal κ is regular if κ = cf(κ), otherwise it is singular. Cofinalities of
ordinals turn out to be regular cardinals.
Exercise 2.9. Show that ℵ1 is regular S and ℵω is singular.
Hint: If A Sis cofinal in ℵ1 , then ℵ1 = A. (Why?) If A ⊆ ℵ1 is countable, what
is the size of A? For the singularity of ℵω use the fact that the union of a set of
cardinals, i.e., the supremum of the set of cardinals, is again a cardinal.
The only limitation to the possible values of 2ℵ0 is the following strengthening
of Theorem 2.8.
Theorem 2.10. Let κ be an infinite cardinal. Then κ < cf(2κ ).
8 STEFAN GESCHKE
Proof. Assume that ZF− is consistent. Then this theory has a model (V, ∈). We
use (V, ∈) in order to construct a model of ZF. To do this, we pretend to live inside
V . The ordinals are the ordinals of V , the cardinals are the cardinals of V , and so
on. Note that we should adjust our definition of ordinals so that we arrive at the
same notion as in the context of Foundation: an ordinal is a transitive set that is
well-ordered (rather than just linearly ordered) by ∈. It can be shown without the
Axiom of Foundation that the class Ord of ordinals is well-ordered by ∈.
For α ∈ Ord let Vα be defined by
(i) V0 := ∅
S P(Vα )
(ii) Vα+1 :=
(iii) Vα := β<α Vβ , if α is a limit ordinal.
Using transfinite induction we show that (Vα )α∈Ord is a strictly ⊆-increasing se-
quence of transitive
S sets.
Let WF = α∈Ord Vα , i.e., let WF be the class consisting of all sets a such that
for some ordinal α, a ∈ Vα . For every a ∈ WF let rk(a) denote the least ordinal
such that a ∈ Vα+1 . The ordinal rk(a) is the rank of a. Note that every ordinal α
is an element of W F with rk(α) = α.
We show that (W F , ∈) is a model of ZF. Set Existence is obviously satisfied. It
follows from the transitivity of the Vα that W F is transitive. Hence, W F satisfies
the Axiom of Extensionality. The Axiom of Separation follows directly from the
construction of the Vα . Namely, if a is an element of Vα , then so is every subset of
a. The Axioms of Infinity, Union, and Pairing are easily checked.
For the Axiom of Replacement let F be a function, i.e., a class of pairs in W F
such that for all a ∈ W F there is at most one b ∈ W F such that (a, b) ∈ F . Now
consider the function rk ◦F . For every a ∈ W F , F [a] und (rk ◦F )[a] are sets in
V since (V, ∈) satisfies Replacement. We have to find a superset of F [a] in W F .
It follows from Separation the F [a] is a set in W F . Let α := sup((rk ◦F )[a]).
Now F [a] ⊆ Vα+1 and Vα+1 ∈ Vα+2 ⊆ W F . It follows that (W F , ∈) satisfies
replacement.
It remains to show that (W F , ∈) satisfies the Axiom of Foundation. Let a ∈ W F .
Let b ∈ a be of minimal rank. Then b is an ∈-minimal element of a since for all
c, d ∈ W F with c ∈ d we have rk(c) < rk(d).
While this proof looks convincing at first sight, it does have a couple of problems.
First of all, it is a bit confusing to denote a model of ZF− by (V, ∈), since we don’t
assume ∈ to be the real ∈-relation and also the real V is not a set, while structures
for the language of set theory are assumed to be sets. On the other hand, for
someone living inside the structure, the elements of the structure do form the class
of all sets and the binary relation of the structure is simply the ∈-relation. Looking
at the structure from the outside, it should be denoted by M = (X, E) or something
similar. This is what we will do from now on.
The class W F of M , really a formula with a single free variable defining the
class, has been tacitly identified with the subset of X the consists of all a ∈ X that
MODELS OF SET THEORY 9
satisfy this formula in the structure M . Then we showed that this subset of X with
the restriction of E to it is a model of ZF.
It is possible to completely ingnore the world outside M in the proof of Theorem
3.1. In this case the notation (V, ∈) for M is again appropriate. We again define
the class W F . But now the statement “W F is a model of ZF” does not make sense
anymore, since W F is a class and not a set. Still, it is possible to formalize that
(W F , ∈) is a model of a certain sentence ϕ.
Let C be a class and ϕ a formula. The relativization of ϕ to C is the formula
ϕC that is obtained by replacing every quantifier ∃x that occurs in ϕ by ∃x ∈ C.
(Recall that ∀x is just an abbreviation and therefore we do not have to consider
this quantifier here.) Of course, here ∃x ∈ Cψ is just an abbreviation for a more
complicated formula that depends on the formula that defines C. Now “(W F , ∈)
is a model of ZF” can be expressed by saying that for every axiom ϕ of ZF, ϕW F
holds (in V ).
There is one subtle problem if we want to show ϕW F for every axiom ϕ of ZF.
Consider for example the axiom Pairing. It looks like we have to show for all
a, b ∈ W F that {a, b} is an element of W F . This is easy since by Pairing in V ,
{a, b} is a set in V as well and it easy to check that in fact, {a, b} ∈ W F . Something
similar is true for Power Set and Union.
Exercise
S 3.2. Let a, b ∈ W F . Show that {a, b} ∈ W F . Also, show that P(a) and
a are elements of W F .
It is, however, not totally obvious that the set that satisfies the definition of
{a, b} in V also satisfies this definition
S in W F . Similarly, we have to show that the
sets that satisfy the definition of a and of P(a) in V satisfy the same definition in
W F . Luckily, these definitions are simple enough that an element of W F satisfies
the respective definition in V if and only if it satisfies that definition relativized to
WF.
We introduce absoluteness in order to deal with this problem explicitly.
Definition 3.3. Let C be a class and ϕ(x1 , . . . , xn ) a formula. The formula ϕ is
absolute over C if the following holds:
∀x1 , . . . , xn ∈ C(ϕ(x1 , . . . , xn ) ↔ ϕC (x1 , . . . , xn ))
A function (possibly a proper class) is absolute over C if the formula that defines
it is absolute over C.
Obviously, every formula that does not have any quantifiers (i.e., Boolean com-
binations of atomic formulas) of is absolute over every class. Quantifiers of the form
∃x ∈ y are bounded. A formula ϕ is ∆0 if all quantifiers of ϕ are bounded.
Lemma 3.4. ∆0 -formulas are absolute over transitive classes.
Let C be a class that satisfies a certain fragment ZF∗ of ZF, i.e., assume that
ϕ holds for every ϕ ∈ ZF∗ . Let ϕ be a formula and suppose that ZF∗ implies that
C
Observe that the function α 7→ Vα satisfies all the assumptions of Theorem 4.3.
In particular, if V satisfies ZFC, then for every finite list ϕ1 , . . . , ϕn of axiom in
ZFC there are arbitrarily large α such that (Vα , ∈) satisfies ϕ1 , . . . , ϕn .
A proof very similar to the proof of Theorem 4.3 yields
Theorem 4.4 (Löwenheim-Skolem Theorem, downward). Let (M, E) be a struc-
ture for the language of set theory. For every X ⊆ M there is an elementary
submodel N ⊆ M of M such that X ⊆ N and |N|≤|X| +ℵ0 .
Proof. As in the proof of Theorem 4.3 it is enough to find N ⊆ M with X ⊆ N
and |N |=|X | +ℵ0 such that for every existential formula ϕ(x, y1 , . . . , yn ) and all
b1 , . . . , bn ∈ N with
M |= ∃xϕ(b1 , . . . , bn )
there is a ∈ N such that
M |= ϕ(a, b1 , . . . , bn ).
For every formula ϕ(x, y1 , . . . , yn ) we define a function fϕ : M n → M such that
for all b1 , . . . , bn ∈ M we have
M |= ϕ(fϕ (b1 , . . . , bn ), b1 , . . . , bn )
provided
M |= ∃xϕ(b1 , . . . , bn ).
The fϕ are called Skolem functions.
For every X ⊆ M let sk(X) be the Skolem hull of X, i.e., the closure of X
under all the functions fϕ . Since there are only countably many formulas and
hence only countably many Skolem functions, for every X ⊆ M , |sk(X)|≤|X| +ℵ0 .
Clearly, N = sk(X) satisfies all existential statements with parameters in N that
are satisfied in M . Hence, N is an elementary submodel of M .
Now, if V satisfies all axioms of ZFC, we can carry out the following construc-
tion. Let Σ be a finite collection of axioms of ZFC. By the Reflection Principle
there is α such that all ϕ ∈ Σ are absolute over Vα . Since (V, ∈) satisfies Σ, (Vα , ∈)
is a model of Σ as well. By the Löwenheim-Skolem Theorem, Vα has a count-
able elementary submodel N . This shows that every finite collection of axioms of
ZFC has a countable model (assuming that V satisfies ZFC). We will proceed by
constructing countable transitive models of finite parts of ZFC.
Theorem 4.5 (Mostowski Collapse). Let C be a class and R a well-founded re-
lation on C (recall that we assume well-founded relations to be set-like). If R is
extensional, i.e., if two elements a and b of C agree iff extR (a) = extR (b), then
there are a transitive class D and an isomorphism µ : (C, R) → (D, ∈).
12 STEFAN GESCHKE
5. Constructibiltiy
In ZF we will define a class L that satisfies ZFC+GCH. Recall that GCH is the
statement
∀κ ∈ Card(κ ≥ ℵ0 → 2κ = κ+ ).
L is Gödel’s universe of constructible sets. Since we want to show the consistency
of the Axiom of Choice (AC) with ZF, we have to be careful not to use AC in the
following arguments. We will explicit indicate uses of AC in certain places that are
not essential for the theory.
Using the formula sat we can define a function D that assigns to each set M the
set D(M ) of subsets of M that are definable in (M, ∈). For every set M and every
real-world formula ϕ(x, y1 , . . . , yn ) we can show that for all b1 , . . . , bn ∈ M we have
{a ∈ M : ϕM (a, b1 , . . . , bn )} ∈ D(M ).
Lemma 5.1. Let M be a set. Then the following hold:
(1) D(M ) ⊆ P(M )
(2) If M is transitive, then D(M ) and M ⊆ D(M ).
(3) ∀X ⊆ M (|X|< ω → X ∈ D(M ))
(4) (AC) |M|≥ ω →|D(M )|=|M|
Proof. (1) is obvious. For (2) let M be transitive and a ∈ M . Then a = {x ∈
M : x ∈ a}. Hence a is a definable subset of M and thus a ∈ D(M ). Since a was
arbitrary, this shows M ⊆ D(M ). Now let a ∈ D(M ) and let b ∈ a. Since a ⊆ M ,
b ∈ M . But since M ⊆ D(M ), b ∈ D(M ). This shows that D(M ) is transitive.
(3) In V we use induction over the size of X. Clearly, the empty set is definable.
Now let n ∈ ω and let f : n + 1 → X be a bijection. By our inductive hypothesis,
f [n] is a definable subset of M , for instance f [n] = {x ∈ M : M |= ϕ(x, b1 , . . . , bm )}
for some b1 , . . . , bm ∈ M . Now
f [n + 1] = {x ∈ M : M |= ϕ(x, b1 , . . . , bm ) ∨ x = bm+1 },
where we choose bm+1 = f (n).
(4) easily follows from the facts that the language of set theory has only countably
many formulas and that for every n ∈ ω there are only |M| n-tuples of parameters
from M .
5.2. The definition of L and its elementary properties.
Definition 5.2. For each ordinal α we define Lα recursively as follows:
(i) L0 := ∅
S D(Lα )
(ii) Lα+1 :=
(iii) Lα := β<α Lβ if α is a limit ordinal.
S
Let L := α∈Ord Lα be the class of constructible sets.
Using Lemma 5.1 by transfinite induction we easily get
Lemma 5.3. For all α ∈ Ord the set Lα is transitive. For all α, β ∈ Ord with
α ≤ β we have Lα ⊆ Lβ .
In particular, the Reflection Principle relativized to L applies to the function
α 7→ Lα .
Definition 5.4. For every a ∈ L we define the L-rank ρ(a) of a to be the least
α ∈ Ord such that a ∈ Lα+1 .
Lemma 5.5. For all α ∈ Ord we have the following:
(1) Lα = {a ∈ L : ρ(a) < α}
(2) Lα ∩ Ord = α
(3) α ∈ L and ρ(α) = α
(4) Lα ∈ Lα+1
(5) Every finite subset of Lα is an element of Lα+1 .
MODELS OF SET THEORY 15
Now let α = β + 1. Then β = Lβ ∩ Ord. But the definition of the class of ordinals
is ∆0 and hence absolute over the transitive class Lβ . Thus, β is a definable subset
of Lβ , namely β = {a ∈ Lβ : (Lβ , ∈) |= a is an ordinal}. Hence β ∈ Lβ+1 . On the
other hand, α is not a subset of Lβ , simply because β 6∈ Lβ . The same holds for
every ordinal ≥ α. Hence α = Lα ∩ Ord.
(3) follows immediately from (2). (4) is obvious.
Proof. We already know that L contains all the ordinals. Hence, in order to show
(V = L)L , it suffices to show that for all α ∈ Ord we have Lα = LL α . Clearly
L0 = ∅ = LL 0 and the limit stages of this inductive proof are easy.
Now let α ∈ Ord and assume Lα = LL L
α . Now Lα+1 = Lα+1 is equivalent to
L
D(Lα ) = D (Lα ). S The definition of D is a formula that uses the parameter ω,
respectively ω <ω = n∈ω ω n , because that is the set that we use to code formulas
as sets. Moreover, the definition of D uses recursion over ω, but the individual
steps in the recursion only use ∆0 -formulas.
It is easily checked that recursively defined functions in which the definitions of
the individual steps of the recursion are absolute are again absolute over models
of sufficiently large fragments of ZF. This shows that D is absolute over transitive
classes that satisfies ZF without Power Set. (We need some fragment of ZF that
allows us to prove the recursion theorem. ZF without Power Set is convenient,
works for our purposes and is satisfied by many sets.)
This shows that L and V agree on whether a subset of Lα is definable or not.
Hence D(Lα ) = DL (Lα ). This finishes the proof of the theorem.
These absoluteness considerations show this: If M is a transitive class that
contains all the ordinals and satisfies ZF, then L ⊆ M and L = LM . In other
words, L is the smallest transitive class that contains all the ordinals and satisfies
ZF.
We will see later that ZF is consistent with V 6= L.
MODELS OF SET THEORY 17
sequence than b̄ or for some n ∈ ω, ā, b̄ ∈ X n and ā @n b̄. Then @<ω is a well-
ordering of X <ω .
Exercise 5.11. Prove Lemma 5.10.
Proof of Theorem 5.9. We show a statement stronger than AC. We show that V =
L implies that there is a definable binary relation / that is a well-ordering of L.
This clearly implies that every constructible set X can be well-ordered, namely by
the restriction of / to X.
Recursively, for all α ∈ Ord we define well-orderings Cα of Lα such that for
α < β, (Lα , Cα ) is an initial segment of (Lβ , Cβ ), i.e., Cα =Cβ Lα and for all
a ∈ Lα and b ∈ Lβ with b Cβ a we have b ∈ Lα . S
Obviously, we have to put C0 = ∅. If α is a limit ordinal, let Cα = β<α Cβ .
S
Similarly, let C= α∈Ord Cα . The only difficult step is the definition of Cα if
α = β + 1 for some ordinal β.
Assume Cβ is a wellordering of Lβ . Since there are only countably many formulas
in the language of set theory, the set of all formulas can be well-ordered. Let ≺ be
a well-ordering of the set of formulas.
Now let a, b ∈ Lα = Lβ+1 . If a, b ∈ Lβ , let a Cα b iff a Cβ b. If a ∈ Lβ and
b ∈ Lα \ Lβ , let a C b. If a, b ∈ Lα \ Lβ , let ϕa (x1 , . . . , xn ) and ϕb (y1 , . . . , ym ) be
≺-minimal formulas defining a, respectively b in (Lβ , ∈) with suitable parameters.
Let (a1 , . . . , an ) ∈ Lnβ be Cnβ -minimal with
a = {c ∈ Lβ : (Lβ , ∈) |= ϕa [c, a1 , . . . , an ]}
and let (b1 , . . . , bm ) ∈ Lm m
β be Cβ -minimal with
6. Forcing
In this section we will find a way to enlarge the universe of all sets.
S
What do dense subsets of O look like? If D ⊆ O is any set, then X = D is
a union of open sets and therefore open. If D is dense, then for every non-empty
open set U ⊆ R there is V ∈ D such that V ⊆ U . In particular, X ∩ U 6= ∅. It
follows that X is dense in R in the topological sense.
On the other hand, if X ⊆ R is dense and open, consider D = {U ⊆ R :
U ⊆ X is open}. Now for every V ∈ O, X ∩ V is non-empty and open and thus
U = X ∩ V ∈ D. It follows that D is dense in O in the order-theoretic sense.
Note however that not every dense subset of O is the collection of all non-empty
open subsets of a dense open set X ⊆ R. An example is the set of all open intervals
with rational endpoints. Another example is for ε > 0 the set Dε of all open
intervals of length < ε.
Now let x be a real number and consider Gx := {U ∈ O : x ∈ U }. It is easily
checked that Gx is a filter in O. Moreover, Gx is (Dε )ε>0 -generic. On the other
hand, if G ⊆ OTis a (Dε )ε>0 -generic filter, then thereTis exactly one real number x
such that x ∈ U ∈G cl(U ). Note, however, that x ∈ G can fail.
Exercise 6.3. Consider the partial order Fn(ω, 2). Show:
S
(1) If G ⊆ Fn(ω, 2) is a filter, then G is a function.
(2) For all n ∈ ω the setDn := {p ∈ Fn(ω, 2) : n ∈ dom(p)}S is dense Fn(ω, 2).
(3) If G ⊆ Fn(ω, 2) is a {Dn : n ∈ ω}-generic filter, then G is a function from
ω to 2.
(4) For every function f : ω → 2 the set Gf := {p ∈ Fn(ω, 2) : p ⊆ f } is a
{Dn : n ∈ ω}-generic filter.
The next theorem is a version of the Baire Category Theorem and guarantees
the existence of sufficiently generic filters.
Theorem 6.4. (Rasiowa-Sikorsky) Let (P, ≤) be a partial order and let D be a
countable family of dense subsets of P. Then there is a D generic filter G ⊆ P.
Proof. Let {Dn : n ∈ ω} be an enumeration of D. Choose a sequence (pn )n∈ω of
conditions in P as follows: Let p0 ∈ D0 . Suppose for n ∈ ω we have chosen pn
already. Choose pn+1 ≤ pn such that pn+1 ∈ Dn+1 . This can be done since Dn+1
is dense.
Now let
G := {p ∈ P : ∃n ∈ ω(pn ≤ p)}.
It is straight forward to check that G is a D-generic filter.
Exercise 6.5. Let (A, ≤A ) and (B, ≤B ) countably infinite, dense linear orders
without endpoints. (Recall that a linear order is dense if strictly between any two
elements there is another element of the linear order.) Show that (A, ≤A ) and
(B, ≤B ) are isomorphic.
Hint: Use the Rasiowa-Sikorski Theorem. Consider the partial P order of iso-
morphisms between finite subsets of A and B, ordered by reverse inclusion. Find a
S D of dense subsets of P such that for every D-generic filter G ⊆ P
countable family
the function G is an isomorphism from A to B. It might help to take another
look at Exercise 6.3.
22 STEFAN GESCHKE
7. CH is independent of ZFC
In this section we show that ZFC neither implies nor refutes CH. We already
know that ZFC does not refute CH, but we will give a forcing argument of this fact.
7.1. Forcing CH. Let M be a countable transitive model of ZFC. We define a
partial order P ∈ M such that 1P
CH.
Definition 7.1. In M let
P = {f : A → P(ω) : A is a countable subset of ℵ1 }.
P is ordered by ≤=⊇.
Let G be P-generic over M . In order to show that CH holds in M [G], we first
observe the following (in M [G]):
Lemma 7.2. Let fG = G. Then fG is a function from (ℵ1 )M onto P(ω) ∩ M .
S
S
Proof. Since any two elements of G have a common extension in G, G is a
function. To see that fG is defined on all of (ℵ1 )M and onto P(ω) ∩ M , in M
we define the following dense sets:
For every α < ℵ1 let Dα := {p ∈ P : α ∈ dom(p)}. For every A ⊆ ω let
DA := {p ∈ P : A ∈ rng(p)}. It is easily checked that the Dα and the DA are dense
in P. Since G is generic over M , G intersects all Dα . Hence dom(fG ) = (ℵ1 )M .
Since G also intersects all DA , we have rng(fG ) = (P(ω))M .
In order to show that M [G] satisfies CH, it is enough to show that (ℵ1 )M =
(ℵ1 )M [G] and (P(ω))M = (P(ω))M [G] . The first equality could fail since (ℵ1 )M
might be countable in M [G]. I.e., M [G] could contain a bijection between ω and
(ℵ1 )M . The second equality could fail since M [G] might contain new subsets of ω.
In both cases M [G] would have to contain new functions from ω into the ordinals.
We show that this cannot happen.
Lemma 7.3. Let f ∈ M [G] be a map from ω into the ordinals. Then f ∈ M .
Proof. Let f˙ ∈ M P be a name such that f˙G = f . Moreover, let p ∈ G be such that
p
f˙ is a function from ω to Ord .
In M let
D = {q ≤ p : ∃g : ω → Ord(q
f˙ = ǧ)}.
By the genericity of G it is enough to show that D is dense below p.
We first observe the following: Let q ≤ p and let F be P-generic over M with
q ∈ F . Then p ∈ F . By the choice of p, f˙F is a function from ω to the ordinals.
Let n ∈ ω. Then f˙F (n) = α for some ordinal α. Since M and M [F ] have the same
ordinals, α ∈ M . Hence there is r ∈ F such that r
f˙(ň) = α̌. (From now on
we will drop severalˇ’s in order to improve readability.) Since G is a filter, we can
choose r ≤ q. This shows that the set of all r ∈ P for which there is some α ∈ Ord
with r
f˙(n) = α in dense below p.
From now on we argue in M . (This saves us several M ’s.) A partial order Q
is σ-closed if for every descending sequence (qn )n∈ω of conditions in Q there is a
common extension q ∈ Q of the qn .
S P is σ-closed. Namely, let (pn )n∈ω be a descending sequence in P. Then p :=
n∈ω pn is a partial function from ℵ1 to P(ω) with countable domain and hence
an element of P. Clearly, p is a common extension of all the pn .
We are now in the position to show that D is dense below p. Let q ≤ p. Choose
q0 ≤ q and α0 ∈ Ord such that q0
f˙(0) = α0 . This is possible by the remark
at the beginning of this proof. Suppose we have chosen qn . As above, there is
MODELS OF SET THEORY 31
qn+1 ≤ qn and αn+1 ∈ Ord such that qn+1
f˙(n + 1) = αn+1 . By the σ-closedness
of P, there is a common extension r of all the qn . We show that r ∈ D.
Let g : ω → Ord; n 7→ αn . For all n ∈ ω we have r ≤ qn . In particular,
r
f˙(n) = αn for all n ∈ ω. Since ω is the same in all transitive models of set
theory, r
f˙ = ǧ. Hence r ∈ D. It follows that D is dense below p.
Proof. By Lemma 7.3 we have (P(ω))M = (P(ω))M [G] and (ℵ1 )M is uncountable in
M [G]. All ordinals below (ℵ1 )M are countable in M and hence in M [G]. It follows
that (ℵ1 )M = (ℵ1 )M [G] . By Lemma 7.2 in M [G] there is map from (ℵ1 )M [G] =
(ℵ1 )M onto (P(ω))M [G] = (P(ω))M . Hence M [G] |= |P(ω)| ≤ ℵ1 . This it is
provable in ZFC that P(ω) is uncountable and M [G] is a model of ZFC, we have
M [G] |= CH.
Exercise 7.5. Let P be a partial order. Consider the following two player game
that lasts ω many rounds:
Let p0 = q0 = 1P . In the n-th round the first player choses a condition pn+1 ≤ qn
and the second player replies by chosing a condition qn+1 ≤ pn+1 . The second player
wins the game iff there is a common extension p ∈ P of the pn .
Suppose the second player has a winning strategy for this game. Show that for
every G that is P-generic over M , M [G] does not contain any new function from ω
into the ordinals.
Exercise 7.6. Let P = Fn(ω, 2). Show that the first player has a winning strategy
in the game described above.
8. Forcing ¬CH
In order to force the failure of CH, we start from a countable transitive model
M of ZFC+CH and generically add at least (ℵ2 )M new subsets of ω. This is rather
easily accomplished. What requires work is to show that the ℵ2 of the ground
model M actually remains ℵ2 in the extension.
In M , let κ be a cardinal > ℵ1 and let
P = Fn(κ × ω, 2) = {p : A → 2 : A is a finite subset of κ × ω}
be ordered by reverse inclusion. S
Let G be P-generic over M . Then fG = G is a function from κ × ω to 2. For
every α ∈ κ let aα = {n ∈ ω : fG (α, n) = 1}.
Lemma 8.1. If α, β ∈ κ are different, then so are aα and aβ .
Proof. Consider the set
Dα,β = {p ∈ P : ∃n ∈ ω((α, n), (β, n) ∈ dom(p) ∧ p(α, n) 6= p(β, n))}.
Let p ∈ P. Since the domain of p is finite, there is n ∈ ω such that neither
(α, n) not (β, n) are in the domain of p. Extend p to a condition q such that
(α, n), (β, n) ∈ dom(q) and q(α, n) 6= q(β, n). Then clearly, q ∈ Dα,β .
It follows that Dα,β is dense in P. Hence there is p ∈ G ∩ Dα,β . Now p ⊆ fG
and hence fG (α, n) 6= fG (β, n). This implies aα 6= aβ .
This shows that forcing with P adds κ new subsets of ω. In order to show that κ
remains large in M [G], we need to discuss a crucial property of the forcing notion
P.
32 STEFAN GESCHKE
We now have to show that P = Fn(κ×ω, 2) is actually c.c.c. This fact is also non-
trivial and can be shown using the ∆-System Lemma, which is pure combinatorics.
8.2. Nice names and the size of 2ℵ0 . Let M , κ, P and G be as before. We want
to compute the actual value of 2ℵ0 in M [G]. We do this by finding a small set of
P-names such that every a ∈ P(ω) ∩ M [G] has a name in that set.
Exercise 8.10. Let a ∈ M [G]. Show that in M there is a proper class of names σ
with σG = a.
34 STEFAN GESCHKE
Definition 8.11. Let Q be a partial order. A Q-name σ is a nice name for a subset
of ω if there is a family (An )n∈ω of antichains of Q such that
[
σ = {(ň, p) : p ∈ An } = ({ň} × An ).
n∈ω
Lemma 8.12. Let F be Q-generic over M . Then for every a ∈ P(ω) ∩ M [F ] there
is a nice name σ ∈ M for a subset of ω such that a = σF .
Proof. Fix a name τ ∈ M such that a = τF . For each n ∈ ω let An be a maximal
antichain in the set {p ∈ Q : p
n ∈ τ }. Clearly, σ = {(ň, p) : p ∈ An } is a nice
name for a subset of ω.
We have to verify that σF = a. Let n ∈ σF . Then for some p ∈ An , p ∈ F . By
the choice of An , p
n ∈ τ and hence n ∈ τF = a. This shows σF ⊆ a.
On the other hand, if n ∈ a, then there is (π, p) ∈ τ such that p ∈ F and πF = n.
For some q ∈ F , q
π = n. We can choose q ≤ p. Now q
n ∈ τ . Since An is a
maximal antichain in the set of conditions that force n ∈ τ , An is predense below
q. Since F is generic and q ∈ F , there is r ∈ F ∩ An . Now r
n ∈ σ and hence
n ∈ σF . This shows a ⊆ σF and hence a = σF .
Exercise 8.13. Recall the definition of the sets aα , α < κ, mentioned in Lemma
8.1. Write down explicitly a nice name for each aα .
Lemma 8.14. Let Q be c.c.c. Then there are at most |Q|ℵ0 nice Q-names for
subsets of ω.
Proof. Since Q is c.c.c., all antichains of Q are countable. Hence Q has at most
|Q|ℵ0 antichains. A nice name for a subset of ω is essentially just a function from
ω into the set of antichains of Q. It follows that there are at most
(|Q|ℵ0 )ℵ0 = |Q|ℵ0
nice names for subsets of ω.
Theorem 8.15. In M , let P = Fn(κ × ω, 2). Let G be P-generic over M . Then
M [G] |= 2ℵ0 = (κℵ0 )M .
Proof. Since κ is infinite, (κℵ0 )M ≥ (2ℵ0 )M . Since P is c.c.c., it preserves cardinals.
By Lemma 8.1, (2ℵ0 )M [G] is at least κ. Since (2ℵ0 )ℵ0 = 2ℵ0 , (2ℵ0 )M [G] is at least
(κℵ0 )M [G] . Clearly, (κℵ0 )M ≤ (κℵ0 )M [G] On the other hand, in M there are at most
κℵ0 nice names for subsets of ω and hence |P(ω)|M [G] ≤ (κℵ0 )M and the theorem
follows.
Corollary 8.16. Let M , G and κ be as above. If M satisfies GCH and cf(κ) >
ℵ0 , then in M [G], 2ℵ0 = κ. If M satisfies GCH but cf(κ) = ℵ0 , then in M [G],
2ℵ0 = κ+ .
Proof. The corollary follows from the previous theorem and the fact that under
GCH we have κℵ0 = κ if cf(κ) > ℵ0 and κℵ0 = κ+ if κ is an infinite cardinal of
countable cofinality.
Exercise 8.17. Prove the statement about the size of κℵ0 under GCH used in the
proof of the previous corollary.
MODELS OF SET THEORY 35
9. Martin’s Axiom
Having produced models of ZFC in which CH fails, there are several natural
questions.
1. Let κ < 2ℵ0 be a cardinal. Is 2κ ≤ 2ℵ0 ?
2. How many measure zero sets are needed to cover the real line?
3. Let X be a topological space. A ⊆ X is nowhere dense if the closure of A
has an empty interior. The Baire Category Theorem says (in particular)
that R is not the union of countably many nowhere dense sets. But how
many nowhere dense sets are needed to cover all of R?
4. A family A ⊆ P(ω) is almost disjoint if all A, B ∈ A with A 6= B have a
finite intersection. An easy application of Zorn’s Lemma shows that every
almost disjoint family of subsets of ω is contained in a maximal one. What
is the minimal size of an infinite maximal almost disjoint family? (An
easy argument shows that no countably infinite almost disjoint family is
maximal.)
All of these questions have an obvious answer under CH. Martin’s Axiom also
answers all these questions, but is consistent with ¬CH. In fact, MA is only inter-
esting if CH fails since under CH it does not say anything new as it follows from
CH.
Definition 9.1. Let κ be a cardinal. MAκ says that for every c.c.c. partial order P
and every family D of size κ of dense subsets of P there is a D-generic filter G ⊆ P.
Martin’s Axiom (MA) is the statement
∀κ < 2ℵ0 (MAκ ).
9.1. Martin’s Axiom and Souslin’s Hypothesis. The original motivation to
introduce Martin’s Axiom lies in Souslin’s Hypothesis. A linear order (L, ≤) is
c.c.c. if there is no uncountable family of pairwise disjoint open intervals of L. L
is separable if it has a countable dense subset. L is connected if it is not the union
of two disjoint open subsets.
It is relatively easy to show that every connected separable linear order without
endpoints is isomorphic to R. Suslins Hypothesis (SH) is the statement that every
connected c.c.c. linear order without endpoints is isomorphic to R. CH neither
implies nor refutes SH. (Both directions of this independence result are due to
Jensen.) SH fails in L. The consistency proof of MA+¬CH is a relatively straight
forward generalization generalization of Solovay’s and Tennenbaum’s proof of the
consistency of SH. Martin’s name is attached to the axiom because he observed that
the consistency proof of SH could be adapted to show the consistency of something
much stronger, namely of MA with an arbitrarily large size of 2ℵ0
36 STEFAN GESCHKE
We now show how MAℵ1 implies SH. A Souslin line is a connected linear order
without endpoints that is c.c.c. but not separable. Since every separable connected
linear order without endpoints is isomorphic to the real line, SH holds and only if
there is no Souslin line.
Definition 9.2. A partial order (T, ≤) is a tree if for all t ∈ T the set {s ∈ T : s ≤ t}
is well ordered. If T is a tree and t ∈ T , then the height htT (t) of t in T is the
order type of the set {s ∈ T : s < t}, i.e., the unique ordinal α such that (α, ∈) is
isomorphic to ({s ∈ T : s < t}, <). For an ordinal α, the α-th level Levα (T ) of T is
the set of all nodes t ∈ T of height α.
A branch of a tree T is a maximal chain in T . An antichain in a tree T is a family
of pairwise incomparable elements of the tree. A tree is Souslin if it is uncountable
and has neither uncountable chains nor antichains.
Lemma 9.3. If there is a Souslin line, then there is a Souslin tree.
Actually the existence of a c.c.c. linear order that is not separable implies the
existence of a Souslin line by roughly the same proof as below. Connectedness
simplifies the proof a tiny little bit. The main information that connectedness gives
us is that L has to be a dense linear order, i.e., if a, b ∈ L are such that a < b, then
(a, b) is nonempty (and in fact infinite).
Proof. Let (L, ≤) be a Souslin line. The Souslin tree T that we are going to con-
struct will consist of nonempty open intervals in L ordered by reverse inclusion. By
recursion on α ∈ Ord we define the α-th level of the tree T .
Let Lev0 (T ) = {L}. We consider L as a nonempty open interval of L. Suppose
we have constructed the β-th level of T and α = β+1. For each interval I ∈ Levβ (T )
let AI be a an infinite maximal disjoint family of open intervals contained in I. This
is possible since by connectedness every nonempty open interval of L has at least
two disjoint nonempty open subintervals. Let
[
Levα (T ) = {AI : I ∈ Levβ (T )}.
Now assume that α is a limit ordinal and for all β < α we have already defined
Levβ (T ). Let Levα (T ) be a maximal of pairwise disjoint, nonempty open intervals
I with the property that for every β < α there is J ∈ Levβ (T ) with I ⊆ J. This
finishes the definition of T . Obviously, the levels of T are eventually empty.
Every antichain of T is a family of pairwise disjoint, nonempty open intervals of
L. Since L is c.c.c., every antichain of T is countable. In particular, every level of
T is countable.
Now suppose that T has an uncountable chain. Since T is a tree, every chain
of T is wellordered. Hence, if T has an uncountable chain, it has a chain of the
form (Iα )α∈ℵ1 , where Iβ ⊆ Iα if α < β. Now the sequence of left endpoints of the
Iα has an uncountable increasing subsequence or the sequence of right endpoints
of the Iα has an uncountable decreasing subsequence. Without loss of generality
assume the former and let (aα )α∈ℵ1 be a strictly increasing sequence in L. Then
((aα , aα+1 ))α∈ℵ1 is an uncountable family of pairwise disjoint, nonempty open in-
tervals in L, a contradiction. It follows that T has no uncountable chains. Note
that this implies that Levℵ1 (T ) is empty.
It remains to show that T is uncountable. Let D be the collection of all the
endpoints of the intervals I ∈ T that are not +∞ or −∞. We claim that D is dense
in L. Let a, b ∈ L be such that a < b. It is enough to find some interval I ∈ T that
intersects (a, b) but is not a superset of (a, b) since in this case (a, b) has to contain
an endpoint of I.
Let α be minimal with the property that (a, b) is disjoint from every element of
Levα (T ).
MODELS OF SET THEORY 37
Our first step in the proof of Theorem 9.5 is to associate every Boolean algebra
with a compact space. This connection between Boolean algebras and certain
compact spaces is known as Stone Duality.
Definition 9.7. Let A be a Boolean algebra and F ⊆ A. We say that F is an
ultrafilter of A iff F is a maximal filter in the partial order A \ {0} or equivalently,
if F is a filter in A \ {0} and for each a ∈ A either a ∈ F or ¬a ∈ F .
Let Ult(A) denote the set of ultrafilters of A topologized by declaring the sets
of the form [a] = {F ∈ Ult(A) : a ∈ F }, a ∈ A, as open. I.e., a subset of Ult(A) is
open iff it is a union of sets of the form [a]. We call the sets [a] basic open.
Exercise 9.8. A set S ⊆ A has the finite intersection property if for all n and
all a1 , . . . , an ∈ A, a1 ∧ · · · ∧ an 6= 0. By Zorn’s Lemma, every set with the finite
intersection property is contained in a maximal set with the finite intersection
property. Show that a maximal set S ⊆ A with the finite intersection property is
an ultrafilter of A.
Lemma 9.9. For every Boolean algebra A, Ult(A) is a compact space, the Stone
space of A.
Proof. We first show that Stone spaces are Hausdorff, i.e., for two distinct points
x, y ∈ Ult(A) there are disjoint open sets U, V ⊆ Ult(A) such that x ∈ U and
y ∈V.
Let x, y ∈ Ult(A) be such that x 6= y. Then there is a ∈ A such that a ∈ x and
a 6∈ y or vice versa. Without loss of generality we assume the first. Since y is an
ultrafilter, ¬a ∈ y. Now [a] and [¬a] are disjoint open sets, the first containing x,
the second containing y. This shows Hausdorffness.
Now let O be an open cover of Ult(A). For every x ∈ Ult(A) choose S ax ∈ A such
that for some Ux ∈ O we have x ∈ [ax ] ⊆ Ux . Clearly, Ult(A) = x∈Ult(A) [ax ]. If
there are finitely many points x1 , . . . , xn ∈ Ult(A) such that Ult(A) = [ax1 ] ∪ · · · ∪
[axn ], then {Ux1 , . . . , Uxn } is a finite subcover of O and we are done.
Suppose there are not finitely many x such that the corresponding [ax ] union
up to Ult(A). In this case the family {Ult(A) \ [ax ] : x ∈ Ult(A)} has the finite
intersection property, i.e., no finite intersection of sets from the family is empty.
This easily translates to the finite intersection property of {¬ax : x ∈ Ult(A)}. By
the previous exercise there isSan ultrafilter y of A that extends {¬ax : x ∈ Ult(A)}.
It is easily checked that y 6∈ x∈Ult(A) [ax ], a contradiction.
Exercise 9.10. Let A be a Boolean algebra. For each a ∈ A let Da = {b ∈ A : b ≤
a ∨ b ⊥ a}. Then each Da is a dense subset of A. If F ⊆ G is a (Da )a∈A -generic
filter, then F is an ultrafilter.
Lemma 9.11. Let A be a Boolean algebra and S ⊆ Ult(A). If S is nowhere dense
in Ult(A), then the set
DS = {a ∈ A \ {0} : [a] ∩ S = ∅}
is dense in A.
On the other hand, if D is a dense subset of A, then the set
[
{x ∈ Ult(A) : x 6∈ {[a] : a ∈ D}}
is nowhere dense in Ult(A).
Proof. Let S be nowhere dense. Taking the closure of S we can assume that S is
closed. Let a ∈ A \ {0}. Since S is closed and nowhere dense, [a] 6⊆ S. Hence [a] \ S
is a nonempty open subset of Ult(A). Since the topology on Ult(A) is generated
by the basic open sets, there is b ∈ A \ {0} such that [b] ⊆ [a] \ S and thus b ∈ DS
and b ≤ a. This shows the density of DS .
MODELS OF SET THEORY 39
S
Now let D be a dense subset of A and let UD = {[a] : a ∈ D}. As a union
of open sets, UD is open. In order to show that Ult(A) \ UD is nowhere dense, it
is enough to show that UD is dense in Ult(A). Let V ⊆ Ult(A) be nonempty and
open. Then there is a ∈ A \ {0} such that [a] ⊆ V . By the density of D, there is
b ∈ D such that b ≤ a. Now clearly [b] ⊆ V ∩ UD . This shows that UD is dense in
Ult(A).
We need yet another observation for the proof of Theorem 9.5 that will also be
useful when we show the consistency of Martin’s Axiom with the negation of CH.
Lemma 9.12. Let κ be a infinite cardinal. Then MAκ holds iff MAκ holds re-
stricted to ccc partial orders of size κ.
Proof. We show only the nontrivial direction. Let (P, ≤) be a ccc partial order and
let D be a family of κ dense subsets of P.
Let λ be sufficiently large and choose an elementary submodel M of Vλ of size λ
such that D ⊆ M and (P, ≤) ∈ M . We consider the partial order Q = M ∩ P. Like
P, Q is ccc. For each q ∈ Q and each D ∈ D, M believes that there is some p ∈ D
such that p ≤ q. It follows that for all q ∈ Q and all D ∈ D there is p ∈ M ∩ D
such that p ≤ q. Hence the sets D ∩ M , D ∈ D, are dense in Q. By MAκ restricted
to partial orders of size κ, there is a filter F ⊆ P that is generic for the collection
{D ∩ M : D ∈ D} of dense subsets of Q. F extends to a filter G ⊆ P. G is D
generic.
Proof of Theorem 9.5. Assume MA and let X be a c.c.c. compact space. Let P
denote the collection of all nonempty open subsets of X ordered by inclusion. Since
X is c.c.c., so is P. Let F be a collection of fewer than 2ℵ0 nowhere dense subsets
of X. We may assume that each member of F is a closed set. For each S ∈ F let
DS = {U ∈ P : cl(U ) ∩ S = ∅}.
Let S ∈ F and let U ⊆ X be nonempty and open. Since S is closed and nowhere
dense, U \ S is nonempty and open. Let V ⊆ X be nonempty, open and such that
cl(V ) ⊆ U \ S. By the definition of DS , V ∈ DS . It follows thatTDS is dense in P.
By MA, there is a (DS )S∈F -generic filter G ⊆ P. Let T = {cl(U ) : U ∈ G}.
Since G has the finite intersection property and X is compact, T 6= ∅. By the
genericity of G, T is disjoint from every S ∈ F, showing that F does not cover X.
On the other hand, assume that no c.c.c. compact space is the union of fewer
than 2ℵ0 nowhere dense sets. By Lemma 9.12, in order to prove Martin’s Axiom,
it is enough to prove Martin’s Axiom restricted to partial orders of size < 2ℵ0 . Let
P be a partial order of size < 2ℵ0 . Let A = ro(P) and X = Ult(A).
Claim 9.13. X is c.c.c.
Let A be a family of pairwise disjoint, nonempty open subsets of X. We may
assume that every element of A is of the form [a] for some a ∈ A. We may further
assume that every element of A is of the form [a] where a = reg p for some p ∈ P.
Since the elements of A are pairwise disjoint, the corresponding elements of P are
pairwise incompatible. Since P is c.c.c., A is countable. This shows the claim.
Now let D be a collection of fewer than 2ℵ0 dense subsets of P. Since P is of size
< 2ℵ0 , we may assume that for all p, q ∈ P, the set
Dp,q = {r ∈ P : r ≤ p, q or r ⊥ p or r ⊥ q}
is in D. Let e : P → ro(P) be the natural embedding. Since e[P] is dense in ro(P),
the images of the SD ∈ D in ro(P) are also dense. It follows that for each D ∈ D
the set SD = X \ {[e(p)] : p ∈ D} is nowhere dense in X.
40 STEFAN GESCHKE
We now show that for every infinite cardinal κ, MAκ implies 2κ = 2ℵ0 . In order
to do this, we have to code subsets of κ by subsets of ω. This is done by almost
disjoint coding, which is due to Solovay.
Fix an almost disjoint family A of size κ of subsets of ω consisting of infinite
sets. Instead of subsets of κ, we code subsets of A by subsets of ω, in the following
way:
Lemma 9.22. Assume MAκ . Then for each B ⊆ A there is C ⊆ ω such that for
all B ∈ B, B ∩ C is finite and for all A ∈ A \ B, C ∩ A is infinite. In other words,
C is almost disjoint from every element of B and not almost disjoint from every
element of A \ B.
42 STEFAN GESCHKE
Proof. We use almost disjoint forcing for the almost disjoint family B. Let for n ∈ ω
and A ∈ B let Dn and DA be the dense sets of almost disjoint forcing defined in
the proof of Theorem 9.20. For all A ∈ A \ B and all n ∈ ω let
n
EA = {p ∈ AD : ∃m ∈ A \ n(m ∈ dom(fp ) ∧ fp (m) = 1)}.
n
Then each of the sets EA is dense in AD.
Now let G ⊆ AD be a filter that intersects all the dense sets Dn , n ∈ ω, the sets
DA , A ∈ B, and EA , n ∈ ω and A ∈ A \ B. Let f = p∈G fp and let C = f −1 (1).
n
S
Then C has the desired properties.
Corollary 9.23. If κ is an infinite cardinal and MAκ holds, then 2ℵ0 = 2κ .
9.4. Iterated forcing. We are going to show the consistency of MA with ¬CH
by means of iterated forcing. The strategy is as follows: we start with a countable
transitive model M of GCH and construct an increasing sequence (Mα )α≤(ℵ2 )M
of models of set theory and a sequence (Gα )α<(ℵ2 )M such that M0 = M and for
all α < (ℵ2 )M , Gα is Qα -generic over Mα where Qα ∈ Mα is a partial order and
Mα+1 = Mα [Gα ]. For each α < (ℵ2 )M ,
Mα |= “Qα is a c.c.c. partial order of size ℵ1 ”.
Most of the Qα add new reals, all the Mα , α < (ℵ2 )M satisfy CH, the final model
M(ℵ2 )M satisfies 2ℵ0 = ℵ2 and no cardinals are collapsed in the process. We choose
Qα in such a way that in the final model the following holds: whenever P is a c.c.c.
forcing of size ℵ1 and D is a family of fewer than ℵ2 dense subsets of P, then there
is α < (ℵ2 )M = ℵ2 such that P = Qα and D ⊆ Mα . In particular, Gα , which is an
element of M(ℵ2 )M , intersects every D ∈ D, showing that there is a D-generic filter
of P. By Lemma 9.12, this is enough to show MA in M(ℵ2 )M .
Our strategy has one major problem: If α ≤ (ℵ2 )M is a limit ordinal, how do we
define Mα ? It is tempting to choose Mα simply as the union of the previous Mβ ,
β < α. However, there is no reason why this union should be a model of ZFC. A
strategy that works, however, is to define a single forcing notion Q in the ground
model M such that every Q-generic filter G over M codes a sequence (Gα )α<(ℵ2 )M
of filters and a sequence (Mα )α≤(ℵ2 )M of models of set theory as above.
Let us first consider the case where we want to iterate just two forcings. Let P
be a partial order in M and let G be P-generic over M . Let Q be a forcing notion
in M [G] and let F be Q-generic over M [G]. Then M [G, F ] = M [G][F ] is a model
of set theory. We want to represent M [G, F ] as a single generic extension of M
with respect to a certain forcing notion in M . The problem is that Q might not
be an element of M . Hence, from the point of view of M , we can access Q only in
terms of a P-name Q̇. For simplicity, we would like to assume that 1P forces that Q̇
is a partial order. Moreover, typically the partial order Q has a simple definition in
M [G] such as “Q is the partial order of all closed subsets of R of positive measure”.
We would like to have a name Q̇ that is forced by 1P to satisfy this definition.
All this is made possible by the Existential Completeness Lemma.
Lemma 9.24 (Existential Completeness Lemma, respectively Maximality Princi-
ple). Let ϕ(x, y1 , . . . , yn ) be a formula in the language of set theory. Let P be a
partial order, τ1 , . . . , τn P-names and p ∈ P. Suppose that
p
∃xϕ(x, τ1 , . . . , τn ).
Then there is a P-name σ such that
p
ϕ(σ, τ1 , . . . , τn ).
The proof of this lemma uses the following sublemma:
MODELS OF SET THEORY 43
Lemma 9.25. Let P be a partial order, A ⊆ P an antichain and (σp )p∈A a family
of P-names. Then there is a P-name σ such that for all p ∈ A, p
σ = σp .
Proof. Let
σ = {(η, r) : ∃p ∈ A∃q ∈ P(r ≤ p ∧ r ≤ q ∧ (η, q) ∈ σp )}.
Now, if p ∈ A, we claim that p
σ = σp .
Let G be generic over the ground model with p ∈ G. Let (η, r) ∈ σ be such
that r ∈ G and ηG = x. Since A is an antichain, p is the unique element of A that
is in G. Hence r ≤ p. By the definition of σ there is q ∈ P such that r ≤ q and
(η, q) ∈ σp . Now q ∈ G and hence x = ηG ∈ (σp )G . This shows σG ⊆ (σp )G .
On the other hand, let x ∈ (σp )G . Fix (η, q) ∈ σp such that q ∈ G and ηG = x.
since p and q are both in G, there is a common extension r ∈ G of both p and q. By
the definition of σ, (η, r) ∈ σ and hence x = ηG ∈ σG . This shows (σp )G ⊆ σG .
Proof of the Existential Completeness Lemma. By the definition of the truthvalue
[[∃xϕ(x, τ1 , . . . , τn )]] the set
D = {q ≤ p : ∃η ∈ V P (q
ϕ(η, τ1 , . . . , τn ))}
is predense below p. Since D is open, it is actually dense below p. Choose a maximal
antichain A in D. For each q ∈ A choose a name ηq such that q
ϕ(ηq , τ1 , . . . , τn ).
By the previous lemma there is a name σ such that for all q ∈ A, q
σ = ηq . Since
A is predense below p,
p
ϕ(σ, τ1 , . . . , τn ).
Definition 9.26. Let P be a partial order and let Q̇ be a P-name for a partial
order, i.e., a P-name such that 1P forces Q̇ to be a partial order. The two-step
iteration of P and Q̇ is the partial order P ∗ Q̇ consisting of all pairs (p, q̇) where p
is a condition in P and q̇ is a P-name such that p
q̇ ∈ Q̇.
Let ≤P denote the order on P and let ≤ ˙ Q be a P-name for the order on Q̇. We
define the order ≤ on P ∗ Q̇ as follows: (p1 , q̇1 ) ≤ (p2 , q̇2 ) iff p1 ≤P p2 and
p1
q̇1 ≤˙ Q q̇2 .
We will drop the subscripts P and Q if they are clear from the context.
Observe that with this definition of the iteration of P and Q̇, P ∗ Q̇ turns out to
be a proper class. Hence we redefine P ∗ Q̇ as
{(p, q̇) ∈ P ∗ Q̇ : rk(q̇) < rk(P) + ω, rk(Q̇) + ω}
even though this definition obviously lacks the elegance of the first definition. Now
whenever p ∈ P and q̇ is a P-name such that p
q̇ ∈ Q̇, then there is a P-name ṙ
such that (p, ṙ) ∈ P∗ Q̇ and p
ṙ = q̇. Whenever we choose a name for an element of
Q̇ we implicitly assume that we only consider names of rank < rk(P) + ω, rk(Q̇) + ω.
44 STEFAN GESCHKE
Now let 1̇Q be a name for the largest element of Q̇. Then obviously,
i : P → P ∗ Q̇ : p 7→ (p, 1̇P )
is an embedding of partial orders. If H is P ∗ Q̇-generic over the ground model M ,
We define G = i−1 [H] and F = {q̇G : ∃p ∈ G((p, q̇) ∈ H)}.
Lemma 9.27. G is P-generic over M and F is Q̇G -generic over M [G]
Proof. It is easily checked that G is a filter. Let D ∈ M be a dense subset of P.
Then the set D0 = {(p, q̇) ∈ P ∗ Q̇ : p ∈ D} is dense in P ∗ Q̇ and hence there is
(p, q̇) ∈ H ∩ D0 . Since (p, q̇) ≤ (p, 1̇Q ), (p, 1̇Q ) ∈ H and thus p ∈ G ∩ D.
We now show that F is a filter in Q̇G . Suppose that q̇G ∈ F and ṙG ∈ Q̇G is
such q̇G ≤ ṙG . Then there is p ∈ G such that p
q̇ ≤ ṙ. Since p ∈ G, (p, 1̇Q ) ∈ H.
Since q̇G ∈ F , there is p0 ∈ P such that (p0 , q̇) ∈ H. Since H is a filter, (p, 1̇Q ) and
(p0 , q̇) have a common extension (p00 , ṡ) in H. Now p00
q̇ ≤ ṙ and p00
ṡ ≤ q̇. It
follows that p00
ṡ ≤ ṙ. Since p00 ≤ p0 we have (p00 , ṡ) ≤ (p00 , ṙ) and thus ṙG ∈ F .
A similar argument shows that any two elements of F have a common extension in
F.
Now let D ∈ M [G] be a dense subset of Q̇G . There is a P-name Ḋ ∈ M for D.
By the Maximality Principle we may assume that 1P forces Ḋ to be a dense subset
of Q̇. Consider the set
D0 = {(p, q̇) ∈ P ∗ Q̇ : p
q̇ ∈ Ḋ}.
This set is dense in P ∗ Q̇:
Let (p, q̇) be an element of P ∗ Q̇. Since Ḋ is forced to be dense in Q̇ and by the
Maximality Principle, there is a name ṙ such that p forces ṙ to be an element of
Ḋ that is an extension of q̇. Now (p, ṙ) ≤ (p, q̇) and p ∈ D0 . It follows that D0 is
dense in P ∗ Q̇.
Since H is P ∗ Q̇-generic over M , there is (p, q̇) ∈ D0 ∩ H. Now (p, 1̇Q ) ∈ H and
thus p ∈ G. By the definition of D0 , p
q̇ ∈ Ḋ. By the definition of F , q̇G ∈ F . It
follows that D ∩ F 6= ∅. Therefore F is generic over M [G].
We now forget about the particular filter G again and choose a P-name β̇ ∈ M
for the supremum of σ. Since P is c.c.c., there is a countable set C ⊆ ω1 such that
1P forces β̇ to be an element Č.
Let γ be the supremum of C. By the choice of β̇, 1P
β̇ ≤ γ̌. But by the
definition of β̇ this means that no pα with α > γ can be an element of a P-generic
filter over M , a contradiction.
9.5. Long iterations. We will now define iterations of infinite length.
Definition 9.30. Let δ be an ordinal. ((Pα )α≤δ , (Q̇β )β<δ ) is a finite support iter-
ation length δ if the following conditions are satisfied:
(1) For every α ≤ δ, Pα is a partial order consisting of sequences of length α.
In particular, P0 is the trivial partial order {∅}.
(2) If α < β ≤ δ and p ∈ Pβ , then p α ∈ Pα .
(3) For every α < δ, Q̇α is a Pα -name for a partial order and Pα+1 = Pα ∗ Q̇α .
Since Pα consists of sequences of length α, it is natural to consider Pα ∗ Q̇α
as consisting of sequences of length α + 1.
(4) If β ≤ δ is a limit ordinal, then Pβ consists of all sequences p of length β
such that the support
supt(p) = {α < β : p(α) 6= 1̇Qα }
of p is finite and for all α < β, p α ∈ Pα .
(5) If ≤ denotes the order on Pβ and p, q ∈ Pβ , then p ≤ q if for all α < β,
p α
p(α)≤ ˙ α q(α) where ≤
˙ α is a name for the order on Q̇α .
Definition 9.31. Let P and Q be partial orders and let e : P → Q be an embedding.
I.e., let e be 1-1 and such that for all p0 , p1 ∈ P, p0 ≤ p1 iff e(p0 ) ≤ e(p1 ). Then e
is a complete embedding if it preserves ⊥ and for all q ∈ Q there is p ∈ P such that
whenever r ∈ P is an extension of p, then r is compatible with q.
Lemma 9.32. Let e : P → Q be an embedding of partial orders in the ground
model M . Then e is a complete embedding iff for every Q-generic filter F over M ,
G = e−1 [F ] is P-generic over M .
If e : P → Q is a complete embedding, then there is a P-name Q : P for a forcing
notion such that forcing with Q is equivalent to forcing with P ∗ (Q : P).
Lemma 9.33. Let ((Pα )α≤δ , (Q̇β )β<δ ) be a finite support iteration and let β < δ.
For every p ∈ Pβ let eβ (p) ∈ Pδ be such that eβ (p) β = p and for all α < δ with
α ≥ β, eβ (p)(α) = 1̇Qα . Then e is a complete embedding.
Now let G be Pδ generic over the ground model M . Then for every β < δ,
G β = e−1β [G] is Pβ -generic over M . Let α < δ. Then Pα+1 = Pα ∗ Q̇α and hence
there is a Q̇α -generic filter Gα over M [G α] such that G (α + 1) = (G α) ∗ Gα .
Lemma 9.34. Let ((Pα )α≤δ , (Q̇β )β<δ ) be a finite support iteration. If for each
α < δ,
Pα
“Q̇α is c.c.c.”,
then Pδ is c.c.c. In short, finite support iterations of c.c.c. forcings are c.c.c.
Theorem 9.35. Martin’s Axiom is consistent with the negation of CH.
Proof. Let M be a countable transitive model of GCH. In M we construct an
iteration
((Pα )α≤ℵ2 , (Q̇β )β<ℵ2 )
of c.c.c. forcing notions such that for all α < ℵ2 , Q̇α is forced to be of size ℵ1 and
for every Pℵ2 -generic filter G over M and every c.c.c. partial order Q of size ℵ1 in
46 STEFAN GESCHKE
M [G] there are cofinally many α < ℵ2 such that (Q̇α )Gα ∼
= Q. It can be shown
that M [G] |= MA + 2ℵ0 = ℵ2 .
MODELS OF SET THEORY 47
References
[1] T. Jech, Set Theory, Academic Press (1978)
[2] K. Kunen, Set Theory, An Introduction to Independence Proofs, North Holland (1980)