Anda di halaman 1dari 36

THE METHOD OF FORCING

JUSTIN TATCH MOORE


Abstract. The purpose of this article is to give a presentation of
the method of forcing aimed at someone with a minimal knowledge
of set theory and logic. The emphasis will be on how the method
can be used in proving theorems in ZFC.

1. Introduction
Let us begin with two thought experiments. First consider the following paradox in probability: if Z is a continuous random variable,
then for any possible outcome z in R, the event Z 6= z occurs almost
surely (i.e. with probability 1). How does one reconcile this with the
fact that, in a truly random outcome, every event having probability
1 should occur? Recasting this in more formal language we have that,
for all z R, almost surely Z 6= z, while almost surely there exists
a z R, Z = z.
Next suppose that, for some index set I, hZi : i Ii is a family of
independent continuous random variables. It is a trivial matter that
for each pair i 6= j, the inequality Zi 6= Zj holds with probability 1.
For large index sets I, however,
|{Zi : i I}| = |I|
holds with probability 0; in fact this event contains no outcomes if I
is larger in cardinality than R. In terms of the formal logic, we have
that, for all i 6= j in I, almost surely the event Zi 6= Zj occurs, while
almost surely it is false that for all i 6= j I, the event Zi 6= Zj
occurrs.
It is natural to ask whether it is possible to revise the notion of almost surely so that its meaning remains unchanged for simple logical
assertions such as Zi 6= Zj but such that it interacts more naturally
with quantification. For instance one might reasonably ask that, in
the second example, |{Zi : i I}| = |I| should occur almost surely
regardless of the cardinality of the index set. Such a formalism would
describe the logical properties of a necessarily larger model of mathematics, in which there are new outcomes to the random experiment
which did not exist before the experiment was performed.
1

JUSTIN TATCH MOORE

The method of forcing, which was introduced by Paul Cohen to establish that the Continuum Hypothesis was not provable from the standard axioms of mathematics,1 addresses issues precisely of this kind.
From a modern perspective, forcing provides a formalism for examining what occurs almost surely not only in this situation but also in a
much more general setting than what is provided by our conventional
notion of randomness. Forcing has proved extremely useful in developing an understanding of models of set theory and in determining
what can and cannot be proved within the standard axiomatization of
mathematics. In fact it is a heuristic of modern set theory that if a
statement arises naturally in mathematics and is consistent, then its
consistency can be established using forcing, possibly utilizing a large
cardinal assumption.
The focus of this article, however, is of a different nature: the aim is
to demonstrate how the method of forcing can be used to prove theorems as opposed to establish consistency results. Forcing itself concerns
the study of adding generic objects to a model of set theory, resulting
in a larger model of set theory. One of the key aspects of forcing is that
it provides a formalism for studying what happens almost surely as the
result of introducing the generic object. An analysis of this formalism
sometimes leads to new results concerning the original model itself
results which are in fact independent of the model entirely. This can be
likened to how the probabilistic method is used in finite combinatorics
in settings where more constructive methods fail.
In this article, I will present several examples of how forcing can be
used to prove theorems. Even though our goals and examples are somewhat unconventional, the forcings themselves and the development of
the theory is much the same as one would encounter in a more standard
treatment of forcing. I will only assume a minimal knowledge of set
theory and logic, similar to what a graduate or advanced undergraduate student might encounter in their core curriculum. In particular, no
prior exposure to forcing will be assumed.
The topics which will be covered include the following: genericity,
names, the forcing relation, absoluteness, the countable chain condition,
countable closure, and homogeneity arguments. These concepts will be
put to use though several case studies:
the intersection properties of uncountable families of events in
a probability space;
1In this article,

we will take the axioms of Zermelo-Frankel set theory with Choice


(ZFC) as the foundations for mathematics. It will be assumed throughout that ZFC
is consistent.

THE METHOD OF FORCING

the existence of points with countable neighborhood bases in


Rosenthal compacta;
the existence of dense metrizable subspaces of Rosenthal compacta;
the Halpern-Lauchli Theorem concerning partitions of finitely
branching trees;
the necessary presence of irregular behavior in marker sequences
for Bernoulli shifts.
While I will generally avoid proving consistency results, the temptation to establish the consistency of the Continuum Hypothesis and its
negation along the way is too great. The proofs of these results will
almost be for free given what needs to be developed.
For those interested in supplementing the material in this article
with a more conventional approach to forcing, Kunens [8] is widely
considered to be the standard treatment. It also provides a complete
introduction to combinatorial set theory and independence results; the
reader looking for further background on set theory is referred there.
None of the results presented below are my own. Since the emphasis
is on developing the mathematics and on not developing the history,
I have not always given attributions to the lemmas and propositions
which play a supporting role.
2. Preliminaries
Before beginning, I will fix some notational conventions and review
some of the basic concepts from set theory which will be needed. Further reading and background can be found in [8]. A set x is transitive
if whenever z y and y x, then z is in x. Equivalently, x is transitive
if every element of x is also a subset of x. An ordinal is a set which
is transitive and well ordered by . Every wellorder is isomorphic to
an ordinal; moreover this ordinal and the isomorphism are unique. If
and are two ordinals, then exactly one of the following is true: ,
, or = .
The least ordinal is the emptyset, which is denoted 0. If is an
ordinal, then + 1 is the least ordinal greater than ; this coincides
with the set {}. The finite ordinals therefore coincide with the
natural numbers: n = {0, . . . , n 1}. The least infinite ordinal is
= {0, 1, . . .}, which coincides with the set of natural numbers. (In the
text below, there are places where N seems more natural; in other cases
is the clear choice. For the sake of uniformity, is used throughout.)
If x is a set, then the rank of x is defined recursively: the rank
of the emptyset is 0 and the rank of x is the least ordinal which is

JUSTIN TATCH MOORE

strictly greater than the ranks of its elements. This is always a well
defined quantity and it will sometimes be necessary to give definitions
by recursion on rank. We recall that formally an ordered pair (x, y) is
defined to be {x, {x, y}}; this is only relevant in that the rank of (x, y)
is greater than the rank of either x or y.
An ordinal is a cardinal if whenever , || < ||. If is an
ordinal which is not a successor, then we say that is a limit ordinal.
In this case, the cofinality of is the minimum cardinality of a cofinal
subset of . A cardinal is regular if its cofinality is . The regularity
of a cardinal is equivalent to the assertion that if is partitioned into
fewer than sets, then one of these sets has cardinality . If is a
cardinal, then + denotes the least cardinal greater than . Cardinals
of the form + are called successor cardinals and are always regular.
Since every set can be well ordered, every set has the same cardinality
as some (unique) cardinal; we will adopt the convention that |x| is the
cardinal such that |x| = ||. If is an ordinal, then we define to
be the th infinite cardinal, starting with 0 = ; +1 = ( )+ . The
Greek letters , , , , and will be used to refer to ordinals; the
letters , , , and will be reserved for cardinals.
If A and B are sets, then B A will be used to denote the collection of
all functions from A into B. For us, a function is simply a set of ordered
pairs. Thus if B C, then B A C A and if f and g are functions,
f g means that f is a restriction of g. There is one exception to
this notation worth noting. I will follow the custom of writing for
in situations where the underlying order structure is unimportant
(formally equals ). Arithmetic expressions involving s will be
used to refer to the cardinality of the resulting set. For instance 21 is
a collection of functions whose cardinality is 21 , which is a cardinal.
A sequence is a function whose domain is an ordinal; the domain of
a sequence will be referred to as its length. If is a sequence, I will
write lh() to denote the length of . If X is a set and is an ordinal,
X < will be used to denote the collection of all sequences of elements
of X of length less than .
If (X, d) is a metric space, then we define its completion to be the set
of equivalence classes of Cauchy sequences, where {xn }
n=0 is equivalent
to {yn }
if
d(x
,
y
)

0.
In
particular,
we
will
regard
the set of real
n
n
n=0
numbers R as being the completion of the rational numbers Q with
the usual metric d(p, q) = |p q|. Notice that even if X is complete,
the completion is not literally equal to X, even though it is canonically
isometric to X. This will serve as a minor annoyance when we define
names for complete metric spaces in Section 5.

THE METHOD OF FORCING

Finally, we will need some notation from first order logic. The language of set theory is the first order language with a single binary
relation .
If is a formula in the language of set theory, then (v0 , . . . , vn1 )
will be used to indicate that vi is the ith occurrence of a distinct free
variable in and that every free variable in is vi for some i < n. If xi
(i < n) are constants, then (x0 , . . . , xn1 ) is the result of substituting
xi for vi . If is a formula and v is variable and x is a term, then [x/v]
is the result of substituting x for every free occurrence of v in (of
which there may be none). If has no free occurrences of a variable,
we say that is a sentence.
A formula in the language of set theory is a 0 -formula if every quantifier in is bounded it is of the form x y or x y
for some variables x and y. Many assertions can
S be expressed by 0 formulas: for instance the assertions A =
B and (Q, ) is a
partially ordered set are expressible by 0 -formulas. We note the following propositions, which justifies our emphasis on transitive models
of set theory below.
Proposition 2.1. If M is a transitive set, (v) is a 0 -formula, and
a is in M , then (M, ) |= (a) if and only if (a) is true.
Thus, for example, if M is a transitive set and Q is a partial order
in M , then (M, ) satisfies that Q is a partial order.
3. What is forcing?
Forcing is the procedure of adjoining to a model M of set theory
a new generic object G in order to create a larger model M [G]. In
this context, M is referred to as the ground model and M [G] is a
generic extension of M . For us, the generic object will always be a
new subset of some partially ordered set Q in M , known as a forcing.
This procedure has the following desirable properties:
M [G] is also a model of set theory and is the minimal model
of set theory which has as members all the elements of M and
also the generic object G.
The truth of mathematical statements in M [G] can be determined by a formalism within M , known as the forcing relation,
which is completely specified by Q. The workings of this formalism are purely internal to M .
While it will generally not concern us in this paper, the key metamathematical feature of forcing is that it is often the case that it is
easier to determine truth in the generic extension M [G] than in the

JUSTIN TATCH MOORE

ground model M . For instance Cohen specified the description of a


forcing Q with the property that if M [G] is any generic extension created by forcing with Q, then M [G] necessarily satisfies that the Continuum Hypothesis is false (see Section 7 below). In fact the thought
experiment presented at the beginning of the introduction is derived
from a variation of this forcing. It is also not difficult to specify different forcings which always produce generic extensions satisfying the
Continuum Hypothesis (see Section 13 below).
There are two perspectives one can have of forcing: one which is primarily semantic and one which is primarily syntactic. Each has its own
advantages and disadvantages. The semantic approach makes certain
properties of the forcing relation and the generic extension intuitive
and transparent. On the other hand, it is fraught with metamathematical issues and philosophical hangups. The syntactic approach is
less intuitive but more tangible and makes certain other features of
forcing constructions more transparent.
We will now fix some terminology.
Definition 3.1 (forcing). A forcing is a set Q equipped with a transitive relation Q which contains a greatest element 1Q . If Q is clear
from the context, the subscripts are usually suppressed.
Our prototypical example of a forcing is R, the collection of all Borel
subsets of [0, 1] having positive Lebesgue measure, ordered by containment. Elements of a forcing are often referred to as conditions and are
regarded as being approximations to a desired generic object. In the
analogy with randomness, the conditions correspond to the events. If
q p, then we sometime say that q is stronger than p or that q extends
p. We think of q as carrying more information or, equivalently, providing a better approximation to the generic object. It will be helpful to
abstract the notion of an outcome in terms of a collection of mutually
compatible outcomes. A set G Q is a filter if G is nonempty, upward
closed, and downward directed in Q:
if q is in G, p is in Q and q p, then p is in G;
if p and q are in G, then there is an r in G such that r p, q.
If p and q are in Q, then we say that p and q are compatible if there is a r
in Q such that r p, q. Otherwise we say that p and q are incompatible
(p q). Notice that two conditions are compatible exactly when there
is a filter which contains both of them. Of course two events in R are
compatible exactly when they intersect in a set of positive measure.
A forcing Q is separative if whenever p 6 q, there is an r p such
that r and q are incompatible. Notice that every separative transitive

THE METHOD OF FORCING

relation is antisymmetric and hence a partial ordering. Notice that if


Q is any forcing, we can define an equivalence relation on Q by q p
if
{r Q : r is compatible with p} = {r Q : r is compatible with q}.
The relation induces an ordering on the -classes which is separative.
The forcing R is not separative; in this example p q if p and q
differ by a measure 0 set. It is often convenient to assume forcings are
separative and we will often pass to the separative quotient without
further mention (just often writes equality in analysis when they really
mean equality modulo a measure 0 set).
A family E of conditions is said to be exhaustive if whenever p is an
element of Q, there is an element q of E which is compatible with p.
Observe that if E R is exhaustive, then its union has full measure.
If A [0, 1] has measure 1, then {q R : q A} is exhaustive. The
following definition will play a central role in all that follows.
Definition 3.2 (generic). If M is a collection of sets and Q is a forcing,
then we say that a filter G Q is M -generic if whenever E Q is an
exhaustive set in M , E G 6= .
Observe that if Q is any forcing and p is in Q, then
{q Q : (q p) (q p)}
is exhaustive. In particular, if M includes all of these sets, then M generic filters are necessarily ultrafilters.
There are two other order-theoretic notions closely related to being
exhaustive which it will be useful to define. A family of pairwise incompatible conditions is said to be an antichain (notice that this differs
from the usual notion of an antichain in a poset, where antichain would
mean pairwise incomparable). Notice that any maximal antichain is exhaustive but that in general exhaustive families need not be pairwise
incompatible. A family D of conditions is dense if every element of
Q has an extension in D. Notice that the collection of all elements of
R which are compact is dense in R. Observe that, by Zorns Lemma,
every dense set contains a maximal antichain. We leave it as an exercise to the reader to show that the property of being M -generic is
unchanged if one replaces exhaustive with either dense or maximal antichain in the definition. Two forcings are regarded as being equivalent
if they have dense suborders which are isomorphic. The reason for this
is that such forcings generate the same generic extensions.

JUSTIN TATCH MOORE

4. The formalism of the forcing relation


In this section we will develop the forcing relation and the forcing
language, treating the notion of a Q-name and the forcing relation
as undefined concepts; the definitions are postponed until Section 5.
The advantage of this more axiomatic approach is that emphasizes the
aspects of the formalism which are actually used in practice.
Let Q be a fixed (separative) forcing for the duration of the section.
As we stated earlier, one can view Q as providing the collection of
events with respect to some abstract notion of randomness. In this
analogy, a Q-name would correspond to a set-valued random variable.
It is conventional to denote Q-names by letters with a dot over them.
There are two examples of Q-names which deserve special mention.
The first is the check name. For each set x, there is a Q-name
x. This correspondes to a random variable which is constant. The
other is the Q-name G for the generic filter. This corresponds to the
principle random variable corresponding to the outcome of the random
experiment.
The forcing language associated to Q is the collection of all first
order formulas in the language of set theory augmented by adding a
constant symbol for each Q-name. If q is in Q and is a formula in the
forcing language, then informally the forcing relation q asserts that
if the event corresponding to q occurrs, then almost surely will be
true. In the absence of the definitions of Q-name and , the following
properties can be regarded as axioms which govern the behaviour of
these primitive concepts. They can be proved from the definitions of
Q-names and the forcing relation which will be given in Section 5.
Property 1. For any p Q and any sets x and y, p x y if and
only if x y. For any p and q in Q, p q G if and only if p q.
It is useful to define the following terminology: if there is a z such
that q y = z, then we say that q decides y (to be z). Similarly, if
p or p , then we say that p decides .
Property 2. For any x, any Q-name y,
and p Q, if p y x,
then there is a q p which decides y.
If is a Q-name, p Q, and
p is an ordinal, then there is a q p which decides .

Property 3. For any q Q, the class of formulas in the forcing


language which are forced by q contains the ZFC axioms, the axioms of
first order logic, and is closed under modus ponens.
Property 4. If x is a Q-name, then the collection of all Q-names y
such that 1 y x forms a set.

THE METHOD OF FORCING

Property 5. If p Q and is a formula in the forcing language, then


p if and only if there is no q p such that q . In particular,
if p and q p, then q .
Property 6. If p Q, then p v if and only if there is an x such
that p [x/v].

If 1 Q , then we will sometimes say that Q forces or, if Q is


clear from the context, that is forced. Similarly, we will write x
is a name for... to mean Q forces x is....
In order to demonstrate how these properties can be used, we will
prove the following useful proposition.
Proposition 4.1. Suppose that x is a set and (v) is a formula in the
forcing language. If for all y x, p (
y ), then p y x(y).
Proof. We will prove the contrapositive. Toward this end, suppose
that p does not force y x(y). It follows from Property 5 there is a
q p such that q y x(y). By Property 3, this is equivalent to
q y x(y). By Property 6, there is a Q-name y such that
q (y x) ((y)).

By Property 2, there is a r q and a z in x such that r y = z. But


now, by Property ??, r (
z ) and hence p does not force (
z ). 
There is a special class of forcings for which there is a more conceptual picture of the forcing relation. We begin by stating a general fact
about forcings.
Theorem 4.2. For every forcing Q, Q is isomorphic to a dense suborder the positive elements of a complete Boolean algebra.
Here we recall that a Boolean algebra is complete if every subset has a
least upper bound.
Suppose now that Q is the positive elements of some complete Boolean
algebra B. If is a formula in the forcing language, then define the
truth value [[]] of to be the join of all b B such that b ; if no
condition forces , then define [[]] = 0. The rules which govern the
logical connectives now take a particularly nice form:
^
[[]] = [[]]c
[[ ]] = [[]] [[]]
[[v]] = [[[x/v]]]

Notice that while x ranges over all Q-names in the last equation a
proper class the collection of all possible values of [[[x/v]]]

is a set
and therefore the last item is meaningful.

10

JUSTIN TATCH MOORE

In spite of the usefulness of complete Boolean algebras in understanding forcing and also in some of the development of the abstract theory
of forcing, forcings of interest rarely present themselves as complete
Boolean algebras. While Theorem 4.2 allows us to represent any forcing inside a complete Boolean algebra, defining forcing strictly in terms
of complete Boolean algebras would prove cumbersome in practice.
5. Names, interpretation, and semantics
In this section we will turn to the task of giving a formal definition of
what is meant by a Q-name and q . This will in turn be used to give
a semantic perspective of forcing. The definitions in this section are
not essential for understanding most forcing arguments and the reader
may wish to skip this section on their first reading of the material. Let
Q be a fixed forcing for the duration of the section.
Definition 5.1 (name). A set x is a Q-name if every element of x is
of the form (y,
q) where y is a Q-name and q is in Q.
Notice that this apparently implicit definition is actually a definition
by induction on rank, as discussed in Section 2. The following provide
two important examples of Q-names.
Definition 5.2 (check names). If x is a set, x is defined recursively by
{(
y , q) : (y x) and (q Q)}.
Definition 5.3 (name for the generic filter). G = {(
q , q) : q Q}.
As mentioned in the previous section, the notion of a Q-name is
intended to describe a proceedure for building a new set from a given
filter G Q. This proceedure is formally described as follows.
Definition 5.4 (interpretation). If G is any filter and x is any Q-name,
define xG recursively by
xG = {y G : p G ((y, p) x)}
Again, this is a definition by induction on rank. In the analogy with
randomness, x G corresponds to evaluating a random variable at a given
outcome.

The following gives the motivation for the definitions of x and G.


Proposition 5.5. If G is any filter and x is any set, then xG = x.
Proposition 5.6. If G is any filter, then G G = G.
We now turn to the formal definition of the forcing relation. The
main complexity of the definition of the forcing relation is tied up in
the formal definition of p x y.

THE METHOD OF FORCING

11

Definition 5.7 (forcing relation: atomic formulae). If Q is a forcing


and x and y are Q-names, then we define the meaning of p x = y
and p x y as follows (the definition is by simultaneous induction
on rank):
p x = y if and only if for all z and p0 p,
(p0 z x)
(p0 z y).

p x y if and only if for every p0 p there is a p00 p0 and


a (z,
q) in y such that p00 q and p00 x = z.

Notice that the definition of p x = y is precisely to ensure that the


Axiom of Extensionality is forced by any condition. The definition of
the forcing relation for nonatomic formulas is straightforward and essentially spelled out in the properties of the forcing relation mentioned
already in Section 4.
Definition 5.8 (forcing relation: logical connectives). Suppose that
p Q and and are formulas in the forcing language.
p if there does not exist a q p such that q .
p if and only if p and p .
p v if and only if for all x,
p [x/v].

(The logical symbols and and handled via de Morgans law.)


The following theorem is one of the fundamental results about forcing. It connects the syntactic properties of the forcing relation with
truth in generic extensions of models of set theory. If M is a countable transitive model of ZFC, Q is a forcing in M , and G Q is an
M -generic filter, define
M [G] = {x G : x M and x is a Q-name}.
In this context, M [G] is the generic extension of M by G and M is
referred to as the ground model.
Theorem 5.9. Suppose that M is a countable transitive model of ZFC
and that Q is a forcing which is in M . If q is in Q, (v1 , . . . , vn ) is
a formula in the language of set theory, and x 1 , . . . , x n are in M , then
the following are equivalent:
q (x 1 , . . . , x n ).
M [G] |= (x G
G
1 ,...,x
n ) whenever G Q is an M -generic filter
and q is in G.
While we will generally not work with the semantics of forcing, let
us note here that it is conventional to use x to denote a Q-name for
an element x of a generic extension M [G]. While such names are

12

JUSTIN TATCH MOORE

not unique, the choice generally does not matter and this informal
convention affords a great deal of notational transparency and economy.
We will now finish this section with some further discussion and
notational conventions concerning names. In some cases, checks are
suppressed in writing certain check names for ease of readability. Typically this is done for , Q, and for specific elements of these sets such
as 0, 1, 1/2.
It is frequently the case in a forcing construction that one encounters
a Q-name for a function f whose domain is forced by some condition
to be a ground model set; that is, for some set x, p dom(f) = x. A
particularly common occurrence is when x = or, more generally, some
ordinal. Under these circumstances, it is common to abuse notation
and regard f as a function defined on x, whose values are themselves
names: f(x) is a Q-name y such that it is forced that f(
x) = y.
Notice

that if, for some sets x and y, p f : x y, it is not the case that
f(x) is of the form z for some z in y i.e. p need not decide the value
of f(x).
In most cases, names are not constructed explicitly. Rather a procedure is described for how to build the object to which the name is
referring and then Properties ?? and 6 are invoked. For example, if x
is a Q-name, x is the Q-name for the unique set which is forced to
be equal to the union of x.
(Notice that there is an abuse of notation
at work here: formally, x is a set which has a union y. It need not be
the case that y is even a Q-name and certainly one should not expect
1 x = y. This is one of the reasons for using to dot notation: it
emphasizes the role of the object as a name.)
A more typical example of is 1 , the least uncountable ordinal. Since
ZFC proves there is a unique set 1 such that 1 is an ordinal, 1 is
uncountable, and every element of 1 is countable, it follows that if Q
is any forcing, 1 Q x(x), where (x) is the above description of 1 .
In particular there is a name x such that 1 Q (x).
Unless readability
dictates otherwise, such names are denoted by adding a dot above
the usual notation (e.g. 1 ).
Another example is R. Recall that R is the completion of Q with
respect to its metric formally the collection of all equivalence classes
of Cauchy sequences of rationals. We use this same formal definition
if Q is a forcing, R is the collection of all Q-names for
of R to define R:
equivalence classes of Cauchy sequences of rational numbers. Notice
and, more to the point, we need not even
that R is not the same as R

have that 1 R = R.

THE METHOD OF FORCING

13

This construction also readily generalizes to define X if X is a complete metric space. The Q-name X is then the collection of all Q-names
x such that 1 forces x is an equivalence class of Cauchy sequences of
That is, X is a Q-name for the completion of X.

elements of X.
6. The cast
We will now introduce the examples which we will put to work
throughout the rest of the paper. The first class of examples provides
the justification for viewing forcings as providing abstract notions of
randomness.
Example 6.1 (Random forcing). Define R to be the collection of all
measurable subsets of [0, 1]. If I is any index set, let RI denote the
collection of all measurable subsets q of [0, 1]I which have positive measure. Here [0, 1] is equipped with Lebesgue measure and [0, 1]I is given
the product measure. Define q p to mean q p modulo a measure
0 set (the order defined by true containment is not separative). Notice
that every element of RI contains a compact set in RI . Furthermore,
any two elements of RI are compatible if their intersection has positive
measure.
Whenever one is considering a forcing Q, one is rarely interested in
the generic filter itself but rather in some generic object which can be
derived in some natural way from the generic filter. For instance, in
RI , it is forced that
\

{cl(q) : q G}
contains a unique element. We will let r be a fixed RI -name for this
element and let ri be a fixed RI -name for the ith coordinate of r.
Also
observe that for all i 6= j in I,
Di,j = {q RI : (x q) (x(i) 6= x(j))}
is dense and therefore 1 i 6= j I (ri 6= rj ). In particular, it
|I|.
Notice however, that we have not
is forced by RI that |R|
2 . This will be
established that if, e.g., I = 2 , then 1 RI 2 =
established in Section 7.
In the context
of R, we will use r to denote a R-name for the unique
T
If M is a transitive model of ZFC, then
element of {cl(q) : q G}.
r [0, 1] is in every measure 1 Borel set coded in M if and only if
{q R M : r q} is a M -generic filter; such an r is commonly
referred to as a random real over M .

14

JUSTIN TATCH MOORE

The next class of examples includes Cohens original forcing. Just as


Random forcing is rooted in the measure theory of [0, 1], Cohen forcing
is rooted in the notion of Baire category in [0, 1].
Example 6.2 (Cohen forcing). Let C denote the collection of all finite
partial functions from to 2: that is all functions q such that the
domain of q is a finite subset of and the range of q is contained in
2 = {0, 1}. We order C by q p if q extends p as a function. If I is a
set, let CI denote the collection of all finite partial functions from I
to 2, similarly ordered by extension. It is not difficult to show that CI
is isomorphic to a dense suborder of the collection of all nonempty open
subsets of [0, 1]I , ordered by containment.
It is very often the case that forcings consist of a collection of partial
functions ordered by extension. By this we mean that q p means
that p is the restriction of q to the domain of p. A filter in the forcing
is then a collection of functions which is directed under containment
and whose union is therefore also a function. This union is the generic
object derived from the generic filter.
In the case of CI , observe that for each i 6= j in I and n , both
{q CI : (i, n) dom(q)}
{q CI : m ({(i, m), (j, m)} dom(q)) (q(i, m) 6= q(j, m))}
are dense. In particular, the generic object will be a function from
I into 2. As in the case of RI , such a generic object naturally
corresponds to an indexed family hri : i Ii of elements of 2 and
genericity ensures that these elements are all distinct.
If M is a transitive model of ZFC, then r [0, 1] is in every dense
open set coded in M if and only if the finite restrictions of the binary
expansion of r is a M -generic filter for the forcing C; such an r is commonly referred to as a Cohen real over M (the binary expansion itself is
also referred to as a Cohen real since it carries the same information).
Notice that since [0, 1] is a union of a measure 0 set and a set of first
category, no element of [0, 1] is both a Cohen real and a random real
over a transitive model of ZFC.
The next example of a forcing appears similar at first to the random
forcing, but in fact it is quite different in nature.
Example 6.3 (Ameoba forcing). If 1 > > 0, then define A to be
the collection of all elements of R of measure greater than . This is
regarded as a forcing with the order induced from R.

THE METHOD OF FORCING

15

Notice that the compatibility relation on A differs from that inherited from R: two conditions in A are compatible in A if and only if
their intersection has measure greater than .
The previous forcings all introduce a new subset of to the ground
model. The next example adds a new ultrafilter on but, as we will
later see, does not introduce a new subset of .
Example 6.4. The collection [] of all infinite subsets of . If x and
y are in [] , define x = y to mean the symmetric difference of x and
y is finite and x y to mean x \ y is finite. The relation is not
antisymmetric; it is once we identify elements which are = -equivalent.
We will abuse notation and denote this quotient by [] as well.
The next forcing is useful in analyzing infinite dimensional generalizations of Ramseys theorem.
Example 6.5. (Mathias forcing; Prikry forcing) Let M denote the
collection of all pairs p = (ap , Ap ) such that Ap is in [] and ap is a
finite initial part of Ap . Define q p to mean ap aq and Aq Ap .
Note in particular that in this situation ap is an initial segment of aq
and aq \ap is contained in Ap . This forcing is known as Mathias forcing.
A related example, known as Prikry forcing, requires that the second
coordinates are elements of some specified nonprincipal ultrafilter U ,
upon which the forcing construction depends. This forcing is denoted
PU .
The final example is an illustration of the potential raw power of
forcing. Typically the phenomenon of collapsing cardinals to 0 is
something one wishes to avoid.
Example 6.6. (collapsing to 0 ) If X is a set, consider the collection
X < of all finite sequences of elements of X, ordered by extension.
Observe that if x is in X, then the collection of all elements of X <
which contain x in their range is dense. As we will see later, this forcing
has the effect of making X countable.
7. The countable chain condition
Something which is often an important consideration in the analysis
of a forcing is whether uncountability is preserved. That is, does 1 Q
1 = 1 ? More generally, one can ask whether cardinals are preserved

by forcing with Q: if X and Y are sets such that |X| < |Y |, then does
< |Y |?
1 Q |X|
In general, this can be a very subtle matter. One way to demonstrate
that a forcing preserves cardinals is to verify that it satisfies the countable chain condition (c.c.c.). A forcing Q satisfies the c.c.c. if every

16

JUSTIN TATCH MOORE

family of pairwise incompatible elements of Q is at most countable.


The following proposition dates to Cohens proof that the Continuum
Hypothesis is independent of ZFC.
Proposition 7.1. Suppose that Q is a c.c.c. forcing. If is a regular
cardinal, then 1 Q
is a regular cardinal. In particular, if is an
.
ordinal, then 1 Q =
Proof. The second conclusion follows from the first since every cardinal
is the supremum of a set of successor cardinals and every supremum of
a set of successor cardinals is a cardinal. Let be a regular cardinal
and Q be a given forcing. Suppose that f and are Q-names and that
p is an element of Q such that
p Q ( )
(f : )

By extending p if necessary, we may assume without loss of generality


for some . It is sufficient to show that p forces f is
that =
not a surjection. If is countable, then is finite and it is possible to
decide f (this does not require that Q is c.c.c.). Thus we will assume
that is uncountable.
For each , define

F () = { : q p(q f(
) = )}.
Notice that if 6= 0 are in F () and q forces f(
) = and q 0 forces
f(
) = 0 , then q and q 0 are incompatible (otherwise any extension q
would force = f(
) = 0 ). Since Q is c.c.c., F () is countable and
has an upper bound g() .
Since is regular, the range of g is bounded. We are therefore
finished once we show that
(f(
p
) F (
)).
By Proposition 4.1, this is equivalent to showing that for all in ,
p f(
) F (
). Suppose for contradiction that this is not the case.
Then there is a q p and an in such that q f(
) 6 F (
). By
0
0

Proposition 2, there is a q q and a such that q f (


) = .
But now F (), a contradiction.

Proposition 7.2. Suppose that Q is a c.c.c. forcing and that hq : <
1 i is a sequence of conditions in Q. Then there is a p such that
is uncountable.
p {
1 : q G}
Remark 7.3. Notice that this characterizes the c.c.c.: if A is an uncountable antichain in a forcing Q, then any condition forces A G
contains at most one element.

THE METHOD OF FORCING

17

Proof. Suppose that this is not the case. Then


1 Q
1
1 (q G ).
By Property 6 of the forcing relation, there is a Q-name for an element
of 1 such that

1 Q
1 (q G ).
As in the proof of Proposition 7.1, the set of 1 such that, for
some q Q, q =
is countable and therefore bounded by some .
a
That is 1
1 (q G ). But now q forces q is in G,
contradiction.

We will now return to the examples introduced in Section 6.
Proposition 7.4. For any index set I, both RI and CI are c.c.c. forcings.
Proof. In the case of RI , this is just a reformulation of the obvious
statement that if F is a family of measurable subsets of [0, 1]I , each
having positive measure, then there are two elements of F which intersect in a set of positive measure. The same argument applies to CI ,
with the observation that it is possible to regard CI as a dense suborder of the collection of all nonempty open subsets of [0, 1]I , ordered by
containment.

Proposition 7.5. The forcing A satisfies the c.c.c. for every > 0.
Remark 7.6. The reader may wonder why we have not bothered to
generalize A to a larger index set, given that we did this for R. The
reason is that, for uncountable index sets, the analog of A is not a
c.c.c. and in fact collapses the cardinality of I to become countable.
Proof. Let D denote the collection of all elements of A which are finite
unions of rational intervals. Notice that D is countable. For each p in
D, let Fp denote the collection of all elements q of A such that q p
and
(p)
(p) (q) <
.
2
S
Notice that A = pD Fp and that any two elements of Fp intersect in
a set of measure greater than and hence have a common lower bound
in A . Since any antichain in A can meet each family Fp in at most
a single element, A is c.c.c.

We finish this section by demonstrating that the Continuum Hypothesis isnt provable within ZFC. By Theorem 7.1,
1 ) ( 2 =
2 ).
1 R ( 1 =
2

18

JUSTIN TATCH MOORE

On the other hand, it is easy to see that for all 6= 2 ,


1 r 6= r .
2 = 2 . Since the set of formulas which are forced by

Thus 1 |R|
1 is a consistent theory extending ZFC, this establishes that that ZFC
cannot prove the Continuum Hypothesis. The same argument shows
that C2 forces that CH is false.
uchli theorem
8. The Halpern-La
The Halpern-Lauchli Theorem is a Ramsey-theoretic result concerning colorings of products of finitely branching trees. Before stating the
theorem, we need to first define some terminology. Recall that a subset
T of < is a tree if it is closed under initial segments: whenever t is in
T and s is an initial part of t, it follows that s is in T . A tree T <
comes equipped with a natural partial order: s t if and only if s is
an initial part of t. If T < is a tree and l , the lth level of
T consists of those elements of T with exactly l predecessors. The lth
level of T is denoted (T )l ; notice that this coincides with the set of
elements of T of length l.
All trees considered in this section will be assumed to be pruned
without further mention: every element will have at least one immediate successor. A tree T < is finitely branching if every element of
T has only finitely many immediate successors in T . If S T <
are trees and J is infinite, then we say that S is a strong subtree
of T based on J if whenever s is in S with lh(s) J, every immediate
successor of s in T is in S. The Halpern-Lauchli Theorem can now be
stated as follows.
Theorem 8.1. [4] If hTi : i < di is a sequence of finitely branching
subtrees of < and
Y
[
f:
(Ti )l k
l=0 i<d

then there exists an infinite set J and strong subtrees Si Ti


based
on J for each i < d such that f is constant when restricted to
S
Q
lH
i<d (Si )l .
Unlike essentially all other Ramsey-theoretic statements concerning
the countably infinite, the full form of the Halpern-Lauchli Theorem
at least at present can not be derived from the machinery of the
semigroup dynamics of and related structures (see [6], [12]). Note
that the special case of the Halpern-Lauchli Theorem for n-ary trees
is a consequnce of a form of the Hales-Jewett Theorem (see [12]). The

THE METHOD OF FORCING

19

proof which is presented in this section is based on forcing and is an


inessential modification of an argument due to Leo Harrington (see [2]).
In order to prove the Halpern Lauchli Theorem, we will derive it
from the so-called dense set form of the theorem. If T < is a tree
and t is in S, then a set D T is (m, n)-dense in T above t if D (T )n
and whenever u is in (T )m with t u, there is a v in D such that u v.
If t is the null string, then we just say that D is (m, n)-dense in T .
Theorem 8.2. If hTi : i < di is a sequence of finitely branching subtrees
of < and
Y
[
f:
(Ti )l k
l=0 i<d
Q

then there is an l and a t in i<d (Ti )l such that for every m l there
is an n m and sets hDi : i < di such that for each i Q
< s, Di is
(m, n)-dense below ti in Ti and such that f is constant on i<d Di

The original form of the Halpern-Lauchli Theorem is an immediate


consequence of the dense set version and the following observation.
Observation 8.3. Let T < be a tree and t be an element of T . If
hDp : p i is a sequence of subsets of T such that for some increasing
sequence
hmp : p i, Dp is (mp , mp+1 )-dense in T above t, then
S
p=0 Dp contains a strong subtree of T which is based on {mp : p }.
It is also not difficult to see that, unlike the standard formulation
of the Halpern-Lauchli Theorem, the special case of Theorem 8.2 in
which each Ti is 2< is equivalent to the theorem in its full generality.
Harringtons proof of the Halpern-Lauchli theorem uses the forcing
relation to reduce the desired Ramsey-theoretic properties of trees to
Ramsey-theoretic properties of cardinals. In the proof we will need
some standard definitions and facts from combinatorial set theory (see,
e.g., [8, Ch.II]). If is a regular cardinal, a subset S of is stationary
if it intersects every closed and unbounded subset of . Clearly every
stationary subset of has cardinality . Furthermore, if is an infinite
cardinal less than , then the set of all ordinals in of cofinality is
stationary. We will need the following property of stationary sets.
Lemma 8.4 (Pressing Down Lemma; see [8]). Suppose that is a
regular cardinal and S is a stationary set. If r : S satisfies
that r() for all S, then r is constant on a stationary subset of
S. In particular if a stationary subset of is partitioned into fewer than
stationary sets, then one of the pieces of the partition is stationary.
We will need the following variant of the -System Lemma.

20

JUSTIN TATCH MOORE

Lemma 8.5. Suppose that X is a set, is the successor of a regular


cardinal, and hp : i is a sequence of partial functions from X to
| dom(p )|
2 such that for every
< . Then there exists a cofinal
S , 2
H such that H p is a function.
Proof. Let = + . Observe that by replacing X with the union of
the domains of the p s if necessary, we may assume that |X| . In
particular, we may assume that X . For each which is a limit
ordinal of cofinality , define d = dom(p ). Observe that | dom(p )|
must be less than for each and thus d is a bounded subset of
whenever cf() = . Let E consist of all such that if ,
then sup(dom(p )) . It is easily checked that E is a closed and
unbounded set. By Lemma 8.4, there is a stationary S E consisting
of ordinals of cofinality and a such that if is in S, sup d . By
the pigeonhole principle, there is a stationary set H S and partial
function r from X to 2 such that if is in H, then p  = r. Now if
are in H, then p p is a function. To see this, suppose that
is in dom(p ) dom(p ). Since is in E, it must be that . Thus
is in d = d = dom(r) and hence p () = r() = p ().

Next, we will need two closely related Ramsey-theoretic statements
which are relatives of the Erdos-Rado Theorem but which have simpler
proofs.
Lemma 8.6. Suppose that hi : i < di is a sequence
regular cardinals satisfying 2i < i+1 if i < d 1. If f
then there exist
Q cofinal sets Hi i for each i < d
constant on i<d Hi .

ofQuncountable
: i<d i ,
such that f is

Proof. The proof is by induction on d. If d is given and d1 , fix


Hi i for each i < d 1 such that f takes the constant value g() on
!
Y
Hi {}
i<d1

(if d = 1, then the product over the emptyset is the trivial map with
domain and this is vacuously true). By applying the pigeonhole
principle and our cardinal arithmetic assumption, there is an Hd1
d1 such that g is constant on Hd1 and Hi does not dependQ
on for
in Hd1 . It follows that f is constant when restricted to i<d Hi ,
where Hi = Hi for some (equivalently any) in Hd1 .

Lemma 8.7. Suppose that X is a set and hi : i < di is a sequence of
successors of infinite regular cardinals such that 2i < i+1 if i < d 1.

THE METHOD OF FORCING

21

Q
If hp : i<d i i is a sequence of finite partial functions from X
into a countable set, then there are Hi i of cardinality i such that
[
Y
{p :
Hi }
i<d

is a function.
Proof. The proof is by induction on d. The case dQ= 1 follows from
Lemma 8.5. Now suppose hi : i di and hp : id i i are given.
For each in d , find hHi : i < di such that Hi i and such that
[
Y
{p :
Hi }
i<d

is a function, which we will denote by q . Applying the pigeonhole


principle, find a cofinal d such that, for some hHi : i < di,
Hi = Hi if is in . Now apply Lemma 8.5 to hq : i to find
Hd of cardinality d such that
[
[
Y
q = {p :
Hi }
Hd

id

is a function.

Finally we turn to the task of proving the dense set form of the
Halpern-Lauchli Theorem.
Proof of Theorem 8.2. Let d be given and suppose that
Y
[
f:
2k 2
k=0 i<d

is given. Define hi : i < di by 0 = 1 and i+1 = (2i )++ . Set Q to


be the collection of all finite partial functions from d1 into 2< . The
order on Q is defined by q p if the domain of q contains the domain
of p and p() is an initial part q() whenever is in the domain of p.
Let r be a Q-name for the element of 2 defined by q forces q() is an
initial part of r .QFix a Q-name U for a nonprinciple ultrafilter on .
For each in i<d i , let p be a condition in Q and let e be such
that p forces
U = {m : f (r(0)  m, . . . , r(d1)  m) = e }
By extending p if necessary, we may assume that there is an
is in U.
l such that if is in the domain of p , p () has length l . Define
g() = (e , l , p ((0)), . . . , p ((d 1))). By Lemmas 8.6 and 8.7,
there are cofinal sets Hi i such that:

22

JUSTIN TATCH MOORE

Q
g is contantly (e, l, t0 , . . . , td1 ) on i<d Hi for some (e, l) and
t0 , . . . , td1 in 2l ;
Q
every finite subset of {p : i<d Hi } has a common lower
bound in Q.
Now let m be given. For each i < d, let Ai be a subset of Hi of
cardinality 2m and fix a bijection between Ai and 2m . Let q be a
condition in Q which is a common lower bound for
Y
{p :
Ai }
i<d

and such that if is the element of Ai which corresonds to u 2m


under the bijection, then q() has ti u as an inial part. That is, q
forces r  l + m = ti u.
Let q be an extension of q such that for some n > l + m, q forces
\
Y
n
{U :
Ai }.
i<d

By extending q if necessary, we may assume that for each i < d and


in Ai , q() has length at least n. Finally, set Di to be the set of all
w 2n such that for some in Ai , q()  n = w.
I now claim
Q that Di is (l + m, n) dense below ti for each i < d and
that f  i<d Di is constantly e. To see the former, let u be in 2m
and let be the corresponding element of Ai . By our choice of q,
q()  l + m = ti u and by our choice of q and the definition
Q of Di ,
there is Q
a w in Di such that q()  n = w. If hwi : i < di i<d Di ,
let i<d Ai be such that q(i )  n = wi . Clearly q forces
hr(i)  n : i < di = hwi : i < di.
Furthermore, by the definition of U and n, we have that
f (hr(i)  n : i < di) = f (hwi : i < di) = e.

9. Universally Baire sets and absoluteness
We have already encountered the issue of defining names for Borel
subsets of a Polish space. In this section, we will further develop this
concept, but in a more general setting. This will be useful is formalizing
absoluteness arguments later.
Let (X, d) be a (not necessarily separable) complete metric space.
Recall that the completion of a metric space is taken to be the collection
of all equivalence classes of its Cauchy sequences. Recall also that if Q

is a forcing, then X represents a Q-name for the completion of X.

THE METHOD OF FORCING

23

In this section we will be interested in interpreting names by filters


which are not fully generic. For this reason it is necessary to work with
names which have better properties with respect to such interpretations.
Definition 9.1. If Q is a forcing, then a nice Q-name for an element
of X is a Q-name x such that, for some countable collection of dense
subsets D of Q, x G is an element of (the completion of) (X, d) whenever
G is D-generic.
Remark 9.2. While the completion of a complete metric space is not
literally equal to the original space, there is a canonical isometry between the two and usually there is no need to distinguish them. The
point in the above definition is that the only meaningful way to define
Hence names for eleX is to be the name for the completion of X.

ments of X are equivalence classes of Cauchy sequences. When they


are interpreted by a sufficiently generic filter, they will typically result
in elements of the completion of X, not in elements of X.
We leave it to the reader to verify that if x is a Q-name and p Q
then there is a nice Q-name y for an element of X such that
x X,
p x = y.

Definition 9.3. (see [3]) Let (X, d) be a complete metric space. A


subset A of X is universally Baire if whenever Q is a forcing there is

a Q-name A such that for every nice Q-name x for an element of X,


there is a countable collection of dense subsets D of Q such that:
is in D;
{q Q : q decides x A}
whenever G is a D-generic filter in Q, x G is in (the completion
of) X and x G is in A if and only if there is a q in G such that

q x A.
The following proposition, while easy to establish, is important in
what follows.
Proposition 9.4. If A and A 0 are Q-names which both witness that A
is universally Baire with respect to Q, then 1 Q A = A 0 .
Proof. If this were not the case, then there would exist a nice Q-name

x for an element of Xand


a p in Q such that
A 0 ).
p x (A4
Suppose without loss of generality that p x (A \ A 0 ). If G is a
sufficiently generic filter containing p, then x G will be in A since p is
On the other hand, x G cant be in A since
in G and p forces x is in A.
p is in G and p forces x is not in A 0 , a contradiction.


24

JUSTIN TATCH MOORE

The following is also easy to establish; the proof is left to the interested reader.
Proposition 9.5. The universally Baire subsets of a complete metric
space form a -algebra which includes the Borel subsets of X.
Putting this all together, we have the following proposition which
will used in establishing absolutenss results. If (v1 , . . . , vn ) is a logical
formula and x1 , . . . , xn are sets, then we say that (x1 , . . . , xn ) is generically absolute if whenever Q is a forcing and q is in Q, q (
x1 , . . . , xn )
if and only if (x1 , . . . , xn ) is true.
Proposition 9.6. The assertion that a given countable Boolean combination of universally Baire subsets of a complete metric space is
nonempty is generically absolute.
10. Irregularity in marker sequences for Bernoulli shifts
In this section we will give an example of how homogeneity properties
of a forcing can be put to use. The goal of the section is to prove
a special case of a theorem of Gao, Jackson, and Seward concerning
marker sequences in Bernoulli shift actions. A decreasing sequence of
Borel subsets hAn : n i of X is a vanishing marker sequence
for the
T
A
action if each An intersects every orbit of the action and
n=0 n = .
Certainly a necessary requirement for such a sequence to exist is that
every orbit of is infinite. In fact this is also a sufficient condition;
this is the content of the so-called Marker Lemma (see, e.g., [?]). The
following result of Gao, Jackson, and Seward grew out of their analysis
2
of the Borel chomatic number of the free part of the shift graph on 2Z ;
see [7] for a definition and discussion of the undefined terminology.
Theorem 10.1. [?] Suppose that is a countable group, k is a natural
number, and hAn : n i is a vanishing marker sequence for the
free part of the shift action of on k . For every increasing sequence
hFn : n i of finite sets which covers , there is an x k such that:
the closure of the orbit of x is contained in the free part of the
action;
the closure of the orbit of x is a minimal non-empty closed subset of k which is invariant under the action;
there are infinitely many n such that, for some g in Fn , g x is
in An .
Our interest will be primarily in the last clause, although in this
generality, [?] represents the first proof of the first two clauses in the
theorem.

THE METHOD OF FORCING

25

We will focus our attention on the special case = Z and k = 2 (the


case Zn and k arbitrary is just notationally more complicated, but the
full generality of the theorem requires a different argument). In this
context, the above theorem can be rephrased as follows:
Theorem 10.2. [?] Suppose that hAn : n i is a vanishing marker
sequence for the free part of the action of Z on 2Z by shift. For every
f : such that limn f (n) = , there exists an x in 2Z such that
the closure of the orbit of x is contained in the free part of the
action;
the closure of the orbit of x is a minimal non-empty closed subset of 2 which is invariant under the action;
there are infinitely many n such that, for some i [f (n), f (n)],
x + i is in An .
It will be useful to have a more combinatorial way of formulating
the first two conclusions of the theorem. They are provided by the
following lemmas.
Lemma 10.3. (see [?]) If x is in 2Z , then the closure of the orbit of
x is contained in the free part of the action exactly when the following
condition holds: for all a Z \ {0}, there exists a finite interval B Z
such that for all c Z there is a b B with
x(a + b + c) 6= x(b + c).
Lemma 10.4. (see [?]) If x is in 2Z , then the closure of the orbit of x
is a minimal closed invariant subset of 2Z if and only if the following
condition holds: for all finite interval A Z, there is a finite interval
B Z such that for all c Z there is an b B such that for all a A
x(a + c) = x(a + b)
Define Q to consist of all finite partial functions q : Z 2 such that
the domain of q is an interval of integers, denoted Iq . If n is in Z and q
is in Q, then the translate of q by n is denoted q + n and is defined by
(q + n)(i) = q(i n) (with q + n having domain {i Z : i n Iq }). If
q is in Q, define q to be the bitwise complement of q: q(i) = 1 q(i). A
disjoint union of conditions which is again a condition will be referred
to as a concatenation.
The order on Q is defined by q p if q = p or else q is a concatenation
of a set S of conditions, each of which is a translate of p or p and
such that both p andSat least one translate of p are in S. If G is a
filter in Q such that G is a total function x from Z to 2, then it is
straightforward to check that x satisfies the conclusion of Lemma 10.4

26

JUSTIN TATCH MOORE

and thus that the closure of the orbit of x is a minimal invariant closed
subset of 2Z .
Under a mild genericity assumption on G, the closure of the orbit of
x will be contained in the free part of the action. For each p 6= 1 in
Q and i , define Dp to be the set of all q in Q such that either q
is incompatible with p or else q p, m + i is in the domain of q, and
q(m + i) 6= q(m), where m = max(Ip ).
Lemma 10.5. Each Dp,i is a dense subset of Q and if G Q is a filter
S
which intersects Dp,i for each p Q \ {1} and i , then x = G
satisfies that the closure of the orbit of x is contained in the free part
of the action.
S
Thus we have shown that 1Q forces that x = G satisfies that the
closure of the orbit of x is minimal and contained in the free part of
the Bornouli shift. The following proposition, when combined with
Proposition 9.6, implies Theorem 10.2.
Proposition 10.6. Suppose that hAn : n i is a vanishing sequence
of markers for the free part of the Bornouli shift Z y 2Z and that
f : is a function such that limn f (n) = . Every condition in
Q forces that for every m there is an n m such that x + k A n for
some k with f (n) k f (n).
Proof. Suppose for contradiction that this is not the case. Then there
is a p in Q such that p forces: there is an m such that for every n n,
if f (n) k f (n) then x + k 6 A m . By replacing p with a stronger
condition if necessary, may assume that there is an m such that p
forces for every n n, if f (n) k f (n) then x + k 6 A m . Let
q = p (
p + l) where l is the length of Ip . Observe taht q p and that
if r q and i Z, then there is a j with 0 j < 2l such that r i + j
is compatible with p; simply chose j such that i j is a multiple of 2l.
Let n m be such that f (n) is greater than 2l and find a r q
and an i Z such that r forces x + i is in A n . This is possible since
it is forced that A n is a meets every orbit (strictly speaking, we are
appealing to Proposition 9.6 here). Now, let j < 2|p| be such that r is
compatible with p + i + j.
We now have that r i + j forces x + j is in A n . This follows from
the fact that
1 Q x i + j is generic over the ground model.
(This follows from the observation that if E Q is exhaustive, then so
is any translate of E.) Recall that r i + j is compatible with p and

THE METHOD OF FORCING

27

let p be a common lower bound for r i + j and p. It follows that p


forces both x + j 6 A n and x + j A n , a contradiction.

11. An intersection property of families of sets of
positive measure
The purpose of this section is to use the tools which we have developed in order to prove the following intersection property of sets of
positive measure in [0, 1].
Proposition 11.1. If X R is uncountable and hBx : x Xi is
an indexed collection of Borel subsets of [0, 1], each having positive
measure, then there is a nonempty
set Y X such that Y has no
T
isolated points and such that yY By has positive measure.
Proof. By replacing each Bx with a subset if necessary, we may assume
that each Bx is compact. Similarly, by replacing X with a subset if
necessary, we may assume that there is an > 0 such that if x is in X,
then Bx has measure greater than . Consider the Amoeba forcing A
and let Z be the Q-name for the set
:B
x G}.

{x X
Notice that every condition forces
\
x )
cl(B
xZ

has measure at least . By Proposition 7.2, there is a q in A such that


q forces Z is uncountable. Since it is a theorem that any uncountable
subset of R contains a nonempty subset without isolated points, there
is a Y such that q forces Y Z is nonempty and without isolated
points.
Let X be the completely metrizable space in which we take the
product of the discrete topology on X. Observe that the collection
of all elements of X which enumerate subsets of X without isolated
points is a countable Boolean T
combination of open sets. Furthermore,

if hxk : k i is in X Tand
k=0 Bxk has measure less than , then
there is an n such that kn Bxk has measure less than . Thus the
collection of all elements hxk : k i ofT
X which enumerate a subset
of R with no isolated points such that
k=0 Bxk has measure at least
is a countable Boolean combination of opens sets. The proposition
now follows from Proposition 9.6.


28

JUSTIN TATCH MOORE

12. Todorcevics absoluteness theorem for Rosenthal


compacta
We will now use the results we derived in the previous section to
prove Todorcevics absoluteness theorem for Rosenthal compacta. Fix
a Polish space X. Recall that a real valued function defined on X is
a Baire class one if it is the limit of a pointwise convergent sequence
of continuous functions. Baire characterized functions f which are not
Baire class one as those for which there exist rational numbers p < q
and nonempty sets D0 , D1 x such that the closures of D0 and D1
coincide, have no isolated points, and
inf f (x) p < q sup f (x).

xD0

xD1

The collection of all Baire class one functions on a Polish space X


is denoted BC1 (X) and is equipped with the topology of pointwise
convergence.
A compact topological space which is homeomorphic to a subspace
of BC1 (X) is said to be a Rosenthal compactum. This class includes
all compact metric spaces and is closed under taking closed subspaces
and countable products. The following are typical examples.
Example 12.1 (Hellys space; the double arrow). The collection of all
nondecreasing functions from [0, 1] to [0, 1]. This example is convex as
a subset of R[0,1] . The extreme points of this set are the characteristic
functions of the intervals (r, 1] and [r, 1]. This subspace is homeomorphic to the so-called double arrow space: the set [0, 1] 2 equipped
with the order topology from the lexicographic order.
Example 12.2 (one point compactification). The constant 0 function,
together with the functions
(
1 if t = r
r (t) =
0 otherwise.
This is homeomorphic to the one point compactification of a discrete
set of cardinality 20 .
Rosenthal compacta enjoy a number of strong properties similar to
those of compact metric space. One which will play an important role
below is the following result of Rosenthal.
Theorem 12.3. [9] If X is Polish, A BC1 (X) and f BC1 (X) is
in the closure of A, then there is a countable A0 A such that f is in
the closure of A0 .

THE METHOD OF FORCING

29

In [11], Todorcevic derived a number of properties of Rosenthal compacta by showing that there is a natural way to reinterpret such spaces
as Rosenthal compacta in generic extensions. This result is in fact a
fairly routine consequence of the machinery which was developed in
Section 9.
First, one must verify that elements of BC1 (X) extend to elements
of BC1 (X) in the generic extension.
Lemma 12.4. [11] Suppose that {fn }
n=0 is a sequence of continuous
functions on a Polish space X. The assertion that {fn }
n=0 converges

pointwise is generically absolute. Furthermore, if {fn }


n=0 and {gn }n=0 ,
the assertion that fn gn 0 pointwise is generically absolute.
Proof. Let {fn }
n=0 be sequence of continuous functions. Observe that
[
[\
{x X : |fi (x) fj (x)| > }
>0 n=0 i,jn

specifies a countable Boolean combination of opens subsets of X which


is empty if and only if {fn }
n=0 converges pointwise. Thus the assertion
converges
pointwise
is generically absolute by Proposition
that {fn }
n=0
9.6. The second conclusion is verified in a similar manner.

Now suppose that Q is a forcing, X is a Polish space, and f is in
BC1 (X). By Lemma 12.4, 1 forces that there is a unique element
which extends f; fix a Q-name f for this extension. If
of BC1 (X)
K BC1 (X) is a Rosenthal compactum, then K is a Q-name for the
to X.
(Specifically, it
closure of the set of extensions of elements of K

is a Q-name for the closure of {(f, 1) : f K} in R X .) Todorcevics


absoluteness theorem can now be stated as follows.
Theorem 12.5. [11] Suppose that X is a Polish space and F is a family
of Baire class one functions. The assertion that every accumulation
point of F is Baire class one is generically absolute.
Proof. Let X and F be fixed and let Q be a forcing. It is sufficient to show that the assertion that F has an pointwise accumulation point which is not in BC1 (X) is equivalent to a certain countable
Boolean combination of open sets in a complelely metrizable space being nonempty. Let Z be the set of all sequences
hhfk,i : i i : n i
such that, for each k, hfk,i : i i converges pointwise to an element
of F. We will regard Z as being a product of discrete spaces, noting
that with this topology, Z is compeletely metrizable.

30

JUSTIN TATCH MOORE

Observe that if g is a limit point of F which is not in BC1 (X),


then by Baires characterization, there are rational numbers p < q, sets
A = {ak : k }, B = {bk : k }, and {fk : k } F such that:
A and B are contained in X, have no isolated points, and have
the same closures;
if k < l, then fl (ak ) < p < q < fl (bk ).
Moreover, one can select sequences hfk,i : i i of continuous functions
such that fk,i fk pointwise for each k. Thus we have that for every
k < l there is an n such that if n < j, then fl,j (ak ) < p < q < fl,j (bk ).
It follows that there exist
(hak : k i, hbk : k i, hhfk,i : i i : k i)
in X X Z specifying objects with the above properties if and only
if F has an accumulation point outside of BC1 (X). Notice however,
that these properties define a countable Boolean combination of open
subsets of Z and therefore the theorem follows from Proposition 9.6.

13. Countably closed forcings
There are two basic aspects of a forcing which of fundamental importance in understanding its properties: how large are its families of
pairwise incompatible elements and how frequently do directed families
have lower bounds. Properties of the former type are often referred to
loosely as chain conditions; we have already seen the most important
of these in Section 7. Properties of the later type are known as closure
properties of a forcing. In this section, we will discuss the simplest and
most important example of a closure property.
Definition 13.1 (-closed). A forcing Q is -closed if whenever hqn :
i is a -decreasing sequence of elements of Q, there is a q in Q such
that q qn for all n.
It is perhaps worth remarking that any forcing which is -closed and
atomless (i.e. every element has two incompatible extensions), necessarily has an antichain of cardinality continuum and so in particular is
not c.c.c.. Like c.c.c. forcings, however, -closed forcings also preserve
uncountability, although for a quite different reason.
Proposition 13.2. Suppose that Q is a -closed forcing. If f is a
Q-name and p Q forces that f is a function with domain , then
there is a q p and a function g such that q f = g. In particular
1 and 1 Q R = R.

1 Q 1 =

THE METHOD OF FORCING

31

Proof. Let p and f be given as in the statement of the proposition.


By repeatedly appealing to Fact 2, recursively construct a sequence of
conditions hpn : n i and the values g(n) of a function g defined on
such that for all n , pn+1 pn p and

pn f(
n) = g(n).
Since Q is countably closed, there is a q in Q such that q pn for all
n . Thus by Proposition 4.1, it follows that
q n (f(n) = g(n)).

We will now consider some examples.
Example 13.3. Recall that [] is the collection of all infinite subsets
of , with the ordering : a b if a \ b is finite. This is not actually
an antisymmetric relation, but it is once we identify sets a and b which
have a finite symmetric difference. In this example, we will break
with our convention and write U to denote the name for the generic
ultrafilter added by the forcing. If hAn : n i is a -decreasing
T
sequence of infinite subsets of , let nk be the least element of ik Ai
which is greater than ni for each i < k. Notice that B = {ni : i } is
an infinite set and that {ni : i k} is a subset of Ak . Thus B Ak
for all k. Notice that, by Ramseys theorem, if f : []d 2, then
{q [] : f  [q]d is constant}
is dense in [] (here [A]d denoted the d-element subsets of A). Since
[] is -closed, it does not add new subsets of . Thus 1 forces that
U is a Ramsey ultrafilter on : if f : []d 2 is a coloring of the
d-element subset of , there is a U in the ultrafilter such that f is
constant on the d-element subsets of U .
Example 13.4. Let Q denote the collection of all q such that for
every k, there is a m in q such that 2k divides m. Order q p if there
is a k such that if m is in q \p, then 2k does not divide m. This example
is similar to ([] , ) except that the notion of smallness provided by
finite has been replaced by something more generous: there is an
upper bound on the power of 2 in the prime factorization. We define
q p if every element of q is the sum of a finite subset of p. We
leave it to the reader to verify that Q is -closed. If G is the Q-name
for the generic ultrafilter, let U be the name for the collection of all
subsets U of such that, for some p in G, every sum of a nonempty
finite subset of p is in U . It follows from Hinmans theorem [5] that
U is forced to be an ultrafilter. Define A[d] in this setting to mean the

32

JUSTIN TATCH MOORE

set of all d-element subsets a of A with the property that if i < j are
in a, then for some k, i < 2k and 2k divides j. It can be shown that
1 forces that for all f : 2, there is a U in U such that f  U [d]
where + denotes the
is constant. Furthermore, 1 forces U + U = U,

extension of (, +) to the Cech-Stone compactification (see [12] for


further information).
The next forcing provides a means for forcing the Continuum Hypothesis over a given model of set theory, providing a counterpoint to
the example in Section ??.
Example 13.5. Let Q denote the collection of all countable partial
functions from 1 to R, ordered by extension. Let g be the Q-name for
the union of the generic filter. It is easily verified that 1 forces that g is
Furthermore, if hqn : n i
defined on all of
1 and maps
1 onto R.
S
is a descending sequence of conditions, then
n=0 qn is a condition: it
is a function and its domain is countable, being a countable union of

countable sets. Thus Q is -closed and hence 1 forces in Q that R = R


1 . Hence 1 forces in Q that |R|
= 1 (i.e. that CH is true).
and 1 =
We are now in a position to derive another property of Rosenthal
compacta. The proof below is a reproduction of Todorcevics proof
in [11]; the result itself was originally proved by classical methods by
Bourgain.
Theorem 13.6. [1] If K is a Rosenthal compactum, then K contains
a point with a countable neighborhood base.

Proof. First we recall the following result of Cech


and Pospisil: if K is
a compact topological space of cardinality at most 1 , then K contains
a point with a countable neighborhood base. Let Q be the forcing from
= 1
the previous example. We have seen that 1 forces in Q that |R|
and hence that the collection of all real valued Borel functions on a
given Polish space has cardinality 1 . In particular, 1 forces in Q that
any Rosenthal compactum has cardinality 1 .
Now, let K be a Rosenthal compactum consisting of Baire class 1
functions on some Polish space X. By Theorem 12.5, 1 forces that the
inside of RX still consists only of Baire class one functions.
closure of K
is closed and hence
Since Q is -closed, it follows that 1 forces that K
a compact space of cardinality 1 . Therefore there are Q-names g and
and that
U n (n ) such that 1 forces that g is an element of K
{U n : n } is a countable neighborhood base for g consisting of
basic open sets. Since Q is -closed, there is a q in Q which decides
g to be some f and U n to be some Vn for each n . It follows that
{Vn : n } is a countable neighborhood base.


THE METHOD OF FORCING

33

14. When compacta have dense metrizable subspaces


Suppose that K is a compact Hausdorff space. In this section we
will reformulate the question of when K contains a dense metrizable
subspace in terms of the language of forcing. Recall that every compact Hausdoff space is homeomorphic to a closed subspace of [0, 1]I for
some index set I. In this section, when we reinterpret K in a generic
in [0, 1]I.
extension, we will take K to be the name for the closure of K
Recall that a regular pair in K is a pair (F, G) such that F and
G are disjoint closed G subsets of K. If is an ordered set and
h(F , G ) : i is a sequence of regular pairs, then we say that
h(F , G ) : i is a free sequence if whenever A, B are finite
and satisfy max(A) < min(B), it follows that
\
\
G
F .
A

Recall also that a collection B of nonempty open subsets of K is a


-base if every nonempty open set in K contains an element of B.
We note the following result of Todorcevic.
Theorem 14.1. [10] If K is any compact Hausdorff space, there is a
sequence h(F , G ) : i of regular pairs in K such that {int(G ) :
} forms a -base for K of minimum cardinality and such that
whenever satisfies that {G : } has the finite intersection
property, h(F , G ) : i is a free sequence.
The following result is implicit in [11] and is a component in Todorcevics proof that every Rosenthal compactum contains a dense metrizable subspace. Let QK denote the forcing consisting of all nonempty
regular open subsets of K ordered by containment (recall that a subset
U of K is a regular open set if U is the interior of its closure). We will
let x denote the Q-name for the unique element of the interection of
when regarded as a collection of open sets.
G,
Theorem 14.2. Suppose that K is a compact Hausdorff space and
h(F , G ) : i is a sequence satisfying the conclusion of Theorem
14.1. The following are equivalent:
(1) K has a -disjoint -base.
(2) Every condition in QK forces that x has a countable neighborhood base.
:G
G}|
0 .
(3) Every condition in QK forces that |{
S
Proof. To see that (1)(2), let U =
n=0 Un be such that U is a base for K and such that each Un is pairwise disjoint. Let V be the

34

JUSTIN TATCH MOORE

QK -name for the set {U U : x U }. Since Un is pairwise disjoint,


every condition forces that V Un contains at most one element. It is
therefore sufficient to show that if U is any -base for K, then
V = {U U : x U }
is forced to be a neighborhood base for x.
To see this, let p be an
is a Q-name for an open
arbitrary element of QK and suppose that W
is a
subset of x.
We may assume without loss of generality that W
I
basic open set with respect to the product [0, 1] and, by extending p
. Furthermore, observe
if necessary, we may assume that p decides W
that it must be that p W . Since U is a -base, there is a U in U
which is contained in p W and let q p be a regular open set such
. Notice that we
that q U . Now q forces U V and that U W
have actually showed that if U is a -base, then
The equivalence of (2) and (3) follows from the fact that
{G : ( ) and (x G )}
is forced to be a neighborhood base for x and that
{(F , G ) : ( ) and (x G )}
is a free sequence and hence no smaller neighborhood base can suffice.
:G
G}|
0 .
Finally, suppose that every condition forces |{
Let hn : n i be a sequence of QK -names such that every condition
of QK forces
: x G
}.
{n : n } = {
Let Un be a maximal
antichain in QK such that elements of Un decide
n and let U = S Un . Clearly U is -disjoint; it suffices to show that
n=0
it is a -base. To see this, suppose that V is a nonempty open subset
of K. Let p be a nonempty regular open subset of V and let n be such
V . Now let U be an element of U which decides n
that p forces G
n
to be . Notice that we must have that U G V .

Recall that a topological space X is countably tight if whenever
A X and x cl(A), there is a countable A0 A such that x cl(A0 ).
It is easy to show that continuous images of countably tight spaces. It
is well known that in the class of compact Hausdorff spaces, countable tightness is equivalent to the non existence of uncountable free
sequences or regular pairs. The above theorem implies in particular
that if
1 Q K is countably tight,
then K contains a dense metrizable subspace.

THE METHOD OF FORCING

35

Corollary 14.3. If K contains a dense first countable subspace and


1 P K is countably tight
then K contains a dense metrizable subspace.
An immediate consequence of the results we have developed so far is
the following result of Todorcevic. Previously it was not known whether
there were Rosenthal compacta which were c.c.c. but nonseparable
or whether certain specific Rosenthal compacta had dense metrizable
subspaces (see the discussion in [11]).
Theorem 14.4. [11] Rosenthal compacta contain dense metrizable subspaces.
Proof. Let K be a Rosenthal compacta, and let P be the Q-name for
the space
: f K}.

{f  X
it follows from Theorem
Since K is forced to be contained in BC1 (X),

12.3 that K and hence P are forced to be countably tight. The theorem
now follows from Theorem 13.6 and Corollary 14.3.

References
[1] J. Bourgain. Some remarks on compact sets of first Baire class. Bull. Soc.
Math. Belg., 30(1):310, 1978.
[2] Ilijas Farah and Stevo Todorchevich. Some applications of the method of forcing. Yenisei, Moscow, 1995.
[3] Qi Feng, Menachem Magidor, and Hugh Woodin. Universally Baire sets of
reals. In Set theory of the continuum (Berkeley, CA, 1989), volume 26 of Math.
Sci. Res. Inst. Publ., pages 203242. Springer, New York, 1992.
[4] J. D. Halpern and H. L
auchli. A partition theorem. Trans. Amer. Math. Soc.,
124:360367, 1966.
[5] Neil Hindman. Finite sums from sequences within cells of a partition of N . J.
Combinatorial Theory Ser. A, 17:111, 1974.

[6] Neil Hindman and Dona Strauss. Algebra in the Stone-Cech


compactification,
volume 27 of de Gruyter Expositions in Mathematics. Walter de Gruyter &
Co., Berlin, 1998. Theory and applications.
[7] A. Kechris, S. Solecki, and S. Todorcevic. Borel choromatic numbers. Adv.
Math., 141(1):144, 1999.
[8] Kenneth Kunen. An Introduction to Independence Proofs, volume 102 of Studies in Logic and the Foundations of Mathematics. North-Holland, 1983.
[9] Haskell P. Rosenthal. Some recent discoveries in the isomorphic theory of Banach spaces. Bull. Amer. Math. Soc., 84(5):803831, 1978.
[10] Stevo Todorcevic. Free sequences. Topology Appl., 35(2-3):235238, 1990.
[11] Stevo Todorcevic. Compact subsets of the first Baire class. J. Amer. Math.
Soc., 12(4):11791212, 1999.
[12] Stevo Todorcevic. Introduction to Ramsey spaces, volume 174 of Annals of
Mathematics Studies. Princeton University Press, Princeton, NJ, 2010.

36

JUSTIN TATCH MOORE

Department of Mathematics, Cornell University, Ithaca, NY 14853


4201, USA
E-mail address: justin@math.cornell.edu

Anda mungkin juga menyukai