Movatterモバイル変換


[0]ホーム

URL:


Jump to content
WikipediaThe Free Encyclopedia
Search

Forcing (mathematics)

From Wikipedia, the free encyclopedia
Technique invented by Paul Cohen for proving consistency and independence results
For the use of forcing inrecursion theory, seeForcing (computability).

In the mathematical discipline ofset theory,forcing is a technique for provingconsistency andindependence results. Intuitively, forcing can be thought of as a technique to expand the set theoreticaluniverseV{\displaystyle V} to a larger universeV[G]{\displaystyle V[G]} by introducing a new "generic" objectG{\displaystyle G}.

Forcing was first used byPaul Cohen in 1963, to prove the independence of theaxiom of choice and thecontinuum hypothesis fromZermelo–Fraenkel set theory. It has been considerably reworked and simplified in the following years, and has since served as a powerful technique, both in set theory and in areas ofmathematical logic such asrecursion theory.Descriptive set theory uses the notions of forcing from both recursion theory and set theory. Forcing has also been used inmodel theory, but it is common in model theory to definegenericity directly without mention of forcing.

Intuition

[edit]

Forcing is usually used to construct an expanded universe that satisfies some desired property. For example, the expanded universe might contain many new real numbers (at least2{\displaystyle \aleph _{2}} of them), identified withsubsets of the setN{\displaystyle \mathbb {N} } of natural numbers, that were not there in the old universe, and thereby violate thecontinuum hypothesis.

In order to intuitively justify such an expansion, it is best to think of the "old universe" as amodelM{\displaystyle M} of the set theory, which is itself a set in the "real universe"V{\displaystyle V}. By theLöwenheim–Skolem theorem,M{\displaystyle M} can be chosen to be a "bare bones" model that isexternally countable, which guarantees that there will be many subsets (inV{\displaystyle V}) ofN{\displaystyle \mathbb {N} } that are not inM{\displaystyle M}. Specifically, there is anordinal2M{\displaystyle \aleph _{2}^{M}} that "plays the role of thecardinal2{\displaystyle \aleph _{2}}" inM{\displaystyle M}, but is actually countable inV{\displaystyle V}. Working inV{\displaystyle V}, it should be easy to find one distinct subset ofN{\displaystyle \mathbb {N} } per each element of2M{\displaystyle \aleph _{2}^{M}}. (For simplicity, this family of subsets can be characterized with a single subsetX2M×N{\displaystyle X\subseteq \aleph _{2}^{M}\times \mathbb {N} }.)

However, in some sense, it may be desirable to "construct the expanded modelM[X]{\displaystyle M[X]} withinM{\displaystyle M}". This would help ensure thatM[X]{\displaystyle M[X]} "resembles"M{\displaystyle M} in certain aspects, such as2M[X]{\displaystyle \aleph _{2}^{M[X]}} being the same as2M{\displaystyle \aleph _{2}^{M}} (more generally, thatcardinal collapse does not occur), and allow fine control over the properties ofM[X]{\displaystyle M[X]}. More precisely, every member ofM[X]{\displaystyle M[X]} should be given a (non-unique)name inM{\displaystyle M}. The name can be thought as an expression in terms ofX{\displaystyle X}, just like in asimple field extensionL=K(θ){\displaystyle L=K(\theta )} every element ofL{\displaystyle L} can be expressed in terms ofθ{\displaystyle \theta }. A major component of forcing is manipulating those names withinM{\displaystyle M}, so sometimes it may help to directly think ofM{\displaystyle M} as "the universe", knowing that the theory of forcing guarantees thatM[X]{\displaystyle M[X]} will correspond to an actual model.

A subtle point of forcing is that, ifX{\displaystyle X} is taken to be anarbitrary "missing subset" of some set inM{\displaystyle M}, then theM[X]{\displaystyle M[X]} constructed "withinM{\displaystyle M}" may not even be a model. This is becauseX{\displaystyle X} may encode "special" information aboutM{\displaystyle M} that is invisible withinM{\displaystyle M} (e.g. thecountability ofM{\displaystyle M}), and thus prove the existence of sets that are "too complex forM{\displaystyle M} to describe".[1][2]

Forcing avoids such problems by requiring the newly introduced setX{\displaystyle X} to be ageneric set relative toM{\displaystyle M}.[1] Some statements are "forced" to hold for any genericX{\displaystyle X}: For example, a genericX{\displaystyle X} is "forced" to be infinite. Furthermore, any property (describable inM{\displaystyle M}) of a generic set is "forced" to hold under someforcing condition. The concept of "forcing" can be defined withinM{\displaystyle M}, and it givesM{\displaystyle M} enough reasoning power to prove thatM[X]{\displaystyle M[X]} is indeed a model that satisfies the desired properties.

Cohen's original technique, now calledramified forcing, is slightly different from theunramified forcing expounded here. Forcing is also equivalent to the method ofBoolean-valued models, which some feel is conceptually more natural and intuitive, but usually much more difficult to apply.[3]

The role of the model

[edit]

In order for the above approach to work smoothly,M{\displaystyle M} must in fact be astandard transitive model inV{\displaystyle V}, so that membership and other elementary notions can be handled intuitively in bothM{\displaystyle M} andV{\displaystyle V}. A standard transitive model can be obtained from any standard model through theMostowski collapse lemma, but the existence of any standard model ofZFC{\displaystyle {\mathsf {ZFC}}} (or any variant thereof) is in itself a stronger assumption than the consistency ofZFC{\displaystyle {\mathsf {ZFC}}}.

To get around this issue, a standard technique is to letM{\displaystyle M} be a standard transitive model of an arbitrary finite subset ofZFC{\displaystyle {\mathsf {ZFC}}} (any axiomatization ofZFC{\displaystyle {\mathsf {ZFC}}} has at least oneaxiom schema, and thus an infinite number of axioms), the existence of which is guaranteed by thereflection principle. As the goal of a forcing argument is to proveconsistency results, this is enough since any inconsistency in a theory must manifest with a derivation of a finite length, and thus involve only a finite number of axioms.

Forcing conditions and forcing posets

[edit]

Each forcing condition can be regarded as afinite piece of information regarding the objectX{\displaystyle X} adjoined to the model. There are many different ways of providing information about an object, which give rise to differentforcing notions. A general approach to formalizing forcing notions is to regard forcing conditions as abstract objects with aposet structure.

Aforcing poset is an ordered triple,(P,,1){\displaystyle (\mathbb {P} ,\leq ,\mathbf {1} )}, where{\displaystyle \leq } is apreorder onP{\displaystyle \mathbb {P} }, and1{\displaystyle \mathbf {1} } is the largest element. Members ofP{\displaystyle \mathbb {P} } are theforcing conditions (or justconditions). The order relationpq{\displaystyle p\leq q} means "p{\displaystyle p} isstronger thanq{\displaystyle q}". (Intuitively, the "smaller" condition provides "more" information, just as the smaller interval[3.1415926,3.1415927]{\displaystyle [3.1415926,3.1415927]} provides more information about the numberπ than the interval[3.1,3.2]{\displaystyle [3.1,3.2]} does.) Furthermore, the preorder{\displaystyle \leq } must beatomless, meaning that it must satisfy thesplitting condition:

In other words, it must be possible to strengthen any forcing conditionp{\displaystyle p} in at least two incompatible directions. Intuitively, this is becausep{\displaystyle p} is only a finite piece of information, whereas an infinite piece of information is needed to determineX{\displaystyle X}.

There are various conventions in use. Some authors require{\displaystyle \leq } to also beantisymmetric, so that the relation is apartial order. Some use the termpartial order anyway, conflicting with standard terminology, while some use the termpreorder. The largest element can be dispensed with. The reverse ordering is also used, most notably bySaharon Shelah and his co-authors.

Examples

[edit]

LetS{\displaystyle S} be any infinite set (such asN{\displaystyle \mathbb {N} }), and let the generic object in question be a new subsetXS{\displaystyle X\subseteq S}. In Cohen's original formulation of forcing, each forcing condition is afinite set of sentences, either of the formaX{\displaystyle a\in X} oraX{\displaystyle a\notin X}, that are self-consistent (i.e.aX{\displaystyle a\in X}andaX{\displaystyle a\notin X} for the same value ofa{\displaystyle a} do not appear in the same condition). This forcing notion is usually calledCohen forcing.

The forcing poset for Cohen forcing can be formally written as(Fin(S,2),,0){\displaystyle (\operatorname {Fin} (S,2),\supseteq ,0)}, the finite partial functions fromS{\displaystyle S} to2 =df {0,1}{\displaystyle 2~{\stackrel {\text{df}}{=}}~\{0,1\}} underreverse inclusion. Cohen forcing satisfies the splitting condition because given any conditionp{\displaystyle p}, one can always find an elementaS{\displaystyle a\in S} not mentioned inp{\displaystyle p}, and add either the sentenceaX{\displaystyle a\in X} oraX{\displaystyle a\notin X} top{\displaystyle p} to get two new forcing conditions, incompatible with each other.

Another instructive example of a forcing poset is(Bor(I),,I){\displaystyle (\operatorname {Bor} (I),\subseteq ,I)}, whereI=[0,1]{\displaystyle I=[0,1]} andBor(I){\displaystyle \operatorname {Bor} (I)} is the collection ofBorel subsets ofI{\displaystyle I} having non-zeroLebesgue measure. The generic object associated with this forcing poset is arandom real numberr[0,1]{\displaystyle r\in [0,1]}. It can be shown thatr{\displaystyle r} falls in every Borel subset of[0,1]{\displaystyle [0,1]} with measure 1, provided that the Borel subset is "described" in the original unexpanded universe (this can be formalized with the concept ofBorel codes). Each forcing condition can be regarded as a random event with probability equal to its measure. Due to the ready intuition this example can provide, probabilistic language is sometimes used with other divergent forcing posets.

Generic filters

[edit]

Even though each individual forcing conditionp{\displaystyle p} cannot fully determine the generic objectX{\displaystyle X}, the setGP{\displaystyle G\subseteq \mathbb {P} } of all true forcing conditions does determineX{\displaystyle X}. In fact, without loss of generality,G{\displaystyle G} is commonly considered tobe the generic object adjoined toM{\displaystyle M}, so the expanded model is calledM[G]{\displaystyle M[G]}. It is usually easy enough to show that the originally desired objectX{\displaystyle X} is indeed in the modelM[G]{\displaystyle M[G]}.

Under this convention, the concept of "generic object" can be described in a general way. Specifically, the setG{\displaystyle G} should be ageneric filter onP{\displaystyle \mathbb {P} } relative toM{\displaystyle M}. The "filter" condition means that it makes sense thatG{\displaystyle G} is a set of all true forcing conditions:

ForG{\displaystyle G} to be "generic relative toM{\displaystyle M}" means:

Given thatM{\displaystyle M} is a countable model, the existence of a generic filterG{\displaystyle G} follows from theRasiowa–Sikorski lemma. In fact, slightly more is true: Given a conditionpP{\displaystyle p\in \mathbb {P} }, one can find a generic filterG{\displaystyle G} such thatpG{\displaystyle p\in G}. Due to the splitting condition onP{\displaystyle \mathbb {P} }, ifG{\displaystyle G} is a filter, thenPG{\displaystyle \mathbb {P} \setminus G} is dense. IfGM{\displaystyle G\in M}, thenPGM{\displaystyle \mathbb {P} \setminus G\in M} becauseM{\displaystyle M} is a model ofZFC{\displaystyle {\mathsf {ZFC}}}. For this reason, a generic filter is never inM{\displaystyle M}.

P-names and interpretations

[edit]

Associated with a forcing posetP{\displaystyle \mathbb {P} } is theclassV(P){\displaystyle V^{(\mathbb {P} )}} ofP{\displaystyle \mathbb {P} }-names. AP{\displaystyle \mathbb {P} }-name is a setA{\displaystyle A} of the form

A{(u,p)u is a P-name and pP}.{\displaystyle A\subseteq \{(u,p)\mid u~{\text{is a}}~\mathbb {P} {\text{-name and}}~p\in \mathbb {P} \}.}

Given any filterG{\displaystyle G} onP{\displaystyle \mathbb {P} }, theinterpretation orvaluation map fromP{\displaystyle \mathbb {P} }-names is given by

val(u,G)={val(v,G)pG: (v,p)u}.{\displaystyle \operatorname {val} (u,G)=\{\operatorname {val} (v,G)\mid \exists p\in G:~(v,p)\in u\}.}

TheP{\displaystyle \mathbb {P} }-names are, in fact, an expansion of theuniverse. GivenxV{\displaystyle x\in V}, one definesxˇ{\displaystyle {\check {x}}} to be theP{\displaystyle \mathbb {P} }-name

xˇ={(yˇ,1)yx}.{\displaystyle {\check {x}}=\{({\check {y}},\mathbf {1} )\mid y\in x\}.}

Since1G{\displaystyle \mathbf {1} \in G}, it follows thatval(xˇ,G)=x{\displaystyle \operatorname {val} ({\check {x}},G)=x}. In a sense,xˇ{\displaystyle {\check {x}}} is a "name forx{\displaystyle x}" that does not depend on the specific choice ofG{\displaystyle G}.

This also allows defining a "name forG{\displaystyle G}" without explicitly referring toG{\displaystyle G}:

G_={(pˇ,p)pP}{\displaystyle {\underline {G}}=\{({\check {p}},p)\mid p\in \mathbb {P} \}}

so thatval(G_,G)={val(pˇ,G)pG}=G{\displaystyle \operatorname {val} ({\underline {G}},G)=\{\operatorname {val} ({\check {p}},G)\mid p\in G\}=G}.

Rigorous definitions

[edit]

The concepts ofP{\displaystyle \mathbb {P} }-names, interpretations, andxˇ{\displaystyle {\check {x}}} may be defined bytransfinite recursion. With{\displaystyle \varnothing } theempty set,α+1{\displaystyle \alpha +1} thesuccessor ordinal to ordinalα{\displaystyle \alpha },P{\displaystyle {\mathcal {P}}} thepower-set operator, andλ{\displaystyle \lambda } alimit ordinal, define the following hierarchy:

Name()=,Name(α+1)=P(Name(α)×P),Name(λ)={Name(α)α<λ}.{\displaystyle {\begin{aligned}\operatorname {Name} (\varnothing )&=\varnothing ,\\\operatorname {Name} (\alpha +1)&={\mathcal {P}}(\operatorname {Name} (\alpha )\times \mathbb {P} ),\\\operatorname {Name} (\lambda )&=\bigcup \{\operatorname {Name} (\alpha )\mid \alpha <\lambda \}.\end{aligned}}}

Then the class ofP{\displaystyle \mathbb {P} }-names is defined as

V(P)={Name(α) | α is an ordinal}.{\displaystyle V^{(\mathbb {P} )}=\bigcup \{\operatorname {Name} (\alpha )~|~\alpha ~{\text{is an ordinal}}\}.}

The interpretation map and the mapxxˇ{\displaystyle x\mapsto {\check {x}}} can similarly be defined with a hierarchical construction.

Forcing

[edit]

Given a generic filterGP{\displaystyle G\subseteq \mathbb {P} }, one proceeds as follows. The subclass ofP{\displaystyle \mathbb {P} }-names inM{\displaystyle M} is denotedM(P){\displaystyle M^{(\mathbb {P} )}}. Let

M[G]={val(u,G) | uM(P)}.{\displaystyle M[G]=\left\{\operatorname {val} (u,G)~{\Big |}~u\in M^{(\mathbb {P} )}\right\}.}

To reduce the study of the set theory ofM[G]{\displaystyle M[G]} to that ofM{\displaystyle M}, one works with the "forcing language", which is built up like ordinaryfirst-order logic, with membership as the binary relation and all theP{\displaystyle \mathbb {P} }-names as constants.

DefinepM,Pφ(u1,,un){\displaystyle p\Vdash _{M,\mathbb {P} }\varphi (u_{1},\ldots ,u_{n})} (to be read as "p{\displaystyle p} forcesφ{\displaystyle \varphi } in the modelM{\displaystyle M} with posetP{\displaystyle \mathbb {P} }"), wherep{\displaystyle p} is a condition,φ{\displaystyle \varphi } is a formula in the forcing language, and theui{\displaystyle u_{i}}'s areP{\displaystyle \mathbb {P} }-names, to mean that ifG{\displaystyle G} is a generic filter containingp{\displaystyle p}, thenM[G]φ(val(u1,G),,val(un,G)){\displaystyle M[G]\models \varphi (\operatorname {val} (u_{1},G),\ldots ,\operatorname {val} (u_{n},G))}. The special case1M,Pφ{\displaystyle \mathbf {1} \Vdash _{M,\mathbb {P} }\varphi } is often written as "PM,Pφ{\displaystyle \mathbb {P} \Vdash _{M,\mathbb {P} }\varphi }" or simply "M,Pφ{\displaystyle \Vdash _{M,\mathbb {P} }\varphi }". Such statements are true inM[G]{\displaystyle M[G]}, no matter whatG{\displaystyle G} is.

What is important is that thisexternal definition of the forcing relationpM,Pφ{\displaystyle p\Vdash _{M,\mathbb {P} }\varphi } is equivalent to aninternal definition withinM{\displaystyle M}, defined bytransfinite induction (specifically{\displaystyle \in }-induction) over theP{\displaystyle \mathbb {P} }-names on instances ofuv{\displaystyle u\in v} andu=v{\displaystyle u=v}, and then by ordinary induction over the complexity of formulae. This has the effect that all the properties ofM[G]{\displaystyle M[G]} are really properties ofM{\displaystyle M}, and the verification ofZFC{\displaystyle {\mathsf {ZFC}}} inM[G]{\displaystyle M[G]} becomes straightforward. This is usually summarized as the following three key properties:

Internal definition

[edit]

There are many different but equivalent ways to define the forcing relationM,P{\displaystyle \Vdash _{M,\mathbb {P} }} inM{\displaystyle M}.[4] One way to simplify the definition is to first define a modified forcing relationM,P{\displaystyle \Vdash _{M,\mathbb {P} }^{*}} that is strictly stronger thanM,P{\displaystyle \Vdash _{M,\mathbb {P} }}. The modified relationM,P{\displaystyle \Vdash _{M,\mathbb {P} }^{*}} still satisfies the three key properties of forcing, butpM,Pφ{\displaystyle p\Vdash _{M,\mathbb {P} }^{*}\varphi } andpM,Pφ{\displaystyle p\Vdash _{M,\mathbb {P} }^{*}\varphi '} are not necessarily equivalent even if the first-order formulaeφ{\displaystyle \varphi } andφ{\displaystyle \varphi '} are equivalent. The unmodified forcing relation can then be defined aspM,PφpM,P¬¬φ.{\displaystyle p\Vdash _{M,\mathbb {P} }\varphi \iff p\Vdash _{M,\mathbb {P} }^{*}\neg \neg \varphi .}In fact, Cohen's original concept of forcing is essentiallyM,P{\displaystyle \Vdash _{M,\mathbb {P} }^{*}} rather thanM,P{\displaystyle \Vdash _{M,\mathbb {P} }}.[3]

The modified forcing relationM,P{\displaystyle \Vdash _{M,\mathbb {P} }^{*}} can be defined recursively as follows:

  1. pM,Puv{\displaystyle p\Vdash _{M,\mathbb {P} }^{*}u\in v} means((w,q)v)(qppM,Pw=u).{\displaystyle (\exists (w,q)\in v)(q\geq p\wedge p\Vdash _{M,\mathbb {P} }^{*}w=u).}
  2. pM,Puv{\displaystyle p\Vdash _{M,\mathbb {P} }^{*}u\neq v} means((w,q)v)(qppM,Pwu)((w,q)u)(qppM,Pwv).{\displaystyle (\exists (w,q)\in v)(q\geq p\wedge p\Vdash _{M,\mathbb {P} }^{*}w\notin u)\vee (\exists (w,q)\in u)(q\geq p\wedge p\Vdash _{M,\mathbb {P} }^{*}w\notin v).}
  3. pM,P¬φ{\displaystyle p\Vdash _{M,\mathbb {P} }^{*}\neg \varphi } means¬(qp)(qM,Pφ).{\displaystyle \neg (\exists q\leq p)(q\Vdash _{M,\mathbb {P} }^{*}\varphi ).}
  4. pM,P(φψ){\displaystyle p\Vdash _{M,\mathbb {P} }^{*}(\varphi \vee \psi )} means(pM,Pφ)(pM,Pψ).{\displaystyle (p\Vdash _{M,\mathbb {P} }^{*}\varphi )\vee (p\Vdash _{M,\mathbb {P} }^{*}\psi ).}
  5. pM,Pxφ(x){\displaystyle p\Vdash _{M,\mathbb {P} }^{*}\exists x\,\varphi (x)} means(uM(P))(pM,Pφ(u)).{\displaystyle (\exists u\in M^{(\mathbb {P} )})(p\Vdash _{M,\mathbb {P} }^{*}\varphi (u)).}

Other symbols of the forcing language can be defined in terms of these symbols: For example,u=v{\displaystyle u=v} means¬(uv){\displaystyle \neg (u\neq v)},xφ(x){\displaystyle \forall x\,\varphi (x)} means¬x¬φ(x){\displaystyle \neg \exists x\,\neg \varphi (x)}, etc. Cases 1 and 2 depend on each other and on case 3, but the recursion always refers toP{\displaystyle \mathbb {P} }-names with lesserranks, so transfinite induction allows the definition to go through.

By construction,M,P{\displaystyle \Vdash _{M,\mathbb {P} }^{*}} (and thusM,P{\displaystyle \Vdash _{M,\mathbb {P} }}) automatically satisfiesDefinability. The proof thatM,P{\displaystyle \Vdash _{M,\mathbb {P} }^{*}} also satisfiesTruth andCoherence is by inductively inspecting each of the five cases above. Cases 4 and 5 are trivial (thanks to the choice of{\displaystyle \vee } and{\displaystyle \exists } as the elementary symbols[5]), cases 1 and 2 relies only on the assumption thatG{\displaystyle G} is a filter, and only case 3 requiresG{\displaystyle G} to be ageneric filter.[3]

Formally, an internal definition of the forcing relation (such as the one presented above) is actually a transformation of an arbitrary formulaφ(x1,,xn){\displaystyle \varphi (x_{1},\dots ,x_{n})} to another formulapPφ(u1,,un){\displaystyle p\Vdash _{\mathbb {P} }\varphi (u_{1},\dots ,u_{n})} wherep{\displaystyle p} andP{\displaystyle \mathbb {P} } are additional variables. The modelM{\displaystyle M} does not explicitly appear in the transformation (note that withinM{\displaystyle M},uM(P){\displaystyle u\in M^{(\mathbb {P} )}} just means "u{\displaystyle u} is aP{\displaystyle \mathbb {P} }-name"), and indeed one may take this transformation as a "syntactic" definition of the forcing relation in the universeV{\displaystyle V} of all sets regardless of any countable transitive model. However, if one wants to force over some countable transitive modelM{\displaystyle M}, then the latter formula should be interpreted underM{\displaystyle M} (i.e. with all quantifiers ranging only overM{\displaystyle M}), in which case it is equivalent to the external "semantic" definition ofM,P{\displaystyle \Vdash _{M,\mathbb {P} }} described at the top of this section:

For any formulaφ(x1,,xn){\displaystyle \varphi (x_{1},\dots ,x_{n})} there is a theoremT{\displaystyle T} of the theoryZFC{\displaystyle {\mathsf {ZFC}}} (for example conjunction of finite number of axioms) such that for any countable transitive modelM{\displaystyle M} such thatMT{\displaystyle M\models T} and any atomless partial orderPM{\displaystyle \mathbb {P} \in M} and anyP{\displaystyle \mathbb {P} }-generic filterG{\displaystyle G} overM{\displaystyle M}(a1,,anMP)(pP)(pM,Pφ(a1,,an)MpPφ(a1,,an)).{\displaystyle (\forall a_{1},\ldots ,a_{n}\in M^{\mathbb {P} })(\forall p\in \mathbb {P} )(p\Vdash _{M,\mathbb {P} }\varphi (a_{1},\dots ,a_{n})\,\Leftrightarrow \,M\models p\Vdash _{\mathbb {P} }\varphi (a_{1},\dots ,a_{n})).}

This the sense under which the forcing relation is indeed "definable inM{\displaystyle M}".

Consistency

[edit]

The discussion above can be summarized by the fundamental consistency result that, given a forcing posetP{\displaystyle \mathbb {P} }, we may assume the existence of a generic filterG{\displaystyle G}, not belonging to the universeV{\displaystyle V}, such thatV[G]{\displaystyle V[G]} is again a set-theoretic universe that modelsZFC{\displaystyle {\mathsf {ZFC}}}. Furthermore, all truths inV[G]{\displaystyle V[G]} may be reduced to truths inV{\displaystyle V} involving the forcing relation.

Both styles, adjoiningG{\displaystyle G} to either a countable transitive modelM{\displaystyle M} or the whole universeV{\displaystyle V}, are commonly used. Less commonly seen is the approach using the "internal" definition of forcing, in which no mention of set or class models is made. This was Cohen's original method, and in one elaboration, it becomes the method of Boolean-valued analysis.

Cohen forcing

[edit]

The simplest nontrivial forcing poset is(Fin(ω,2),,0){\displaystyle (\operatorname {Fin} (\omega ,2),\supseteq ,0)}, the finite partial functions fromω{\displaystyle \omega } to2 =df {0,1}{\displaystyle 2~{\stackrel {\text{df}}{=}}~\{0,1\}} underreverse inclusion. That is, a conditionp{\displaystyle p} is essentially two disjoint finite subsetsp1[1]{\displaystyle {p^{-1}}[1]} andp1[0]{\displaystyle {p^{-1}}[0]} ofω{\displaystyle \omega }, to be thought of as the "yes" and "no" parts ofp{\displaystyle p}, with no information provided on values outside the domain ofp{\displaystyle p}. "q{\displaystyle q} is stronger thanp{\displaystyle p}" means thatqp{\displaystyle q\supseteq p}, in other words, the "yes" and "no" parts ofq{\displaystyle q} are supersets of the "yes" and "no" parts ofp{\displaystyle p}, and in that sense, provide more information.

LetG{\displaystyle G} be a generic filter for this poset. Ifp{\displaystyle p} andq{\displaystyle q} are both inG{\displaystyle G}, thenpq{\displaystyle p\cup q} is a condition becauseG{\displaystyle G} is a filter. This means thatg=G{\displaystyle g=\bigcup G} is a well-defined partial function fromω{\displaystyle \omega } to2{\displaystyle 2} because any two conditions inG{\displaystyle G} agree on their common domain.

In fact,g{\displaystyle g} is a total function. Givennω{\displaystyle n\in \omega }, letDn={pp(n) is defined}{\displaystyle D_{n}=\{p\mid p(n)~{\text{is defined}}\}}. ThenDn{\displaystyle D_{n}} is dense. (Given anyp{\displaystyle p}, ifn{\displaystyle n} is not inp{\displaystyle p}'s domain, adjoin a value forn{\displaystyle n}—the result is inDn{\displaystyle D_{n}}.) A conditionpGDn{\displaystyle p\in G\cap D_{n}} hasn{\displaystyle n} in its domain, and sincepg{\displaystyle p\subseteq g}, we find thatg(n){\displaystyle g(n)} is defined.

LetX=g1[1]{\displaystyle X={g^{-1}}[1]}, the set of all "yes" members of the generic conditions. It is possible to give a name forX{\displaystyle X} directly. Let

X_={(nˇ,p)p(n)=1}.{\displaystyle {\underline {X}}=\left\{\left({\check {n}},p\right)\mid p(n)=1\right\}.}

Thenval(X_,G)=X.{\displaystyle \operatorname {val} ({\underline {X}},G)=X.} Now suppose thatAω{\displaystyle A\subseteq \omega } inV{\displaystyle V}. We claim thatXA{\displaystyle X\neq A}. Let

DA={p(n)(nDom(p)(p(n)=1nA))}.{\displaystyle D_{A}=\{p\mid (\exists n)(n\in \operatorname {Dom} (p)\land (p(n)=1\iff n\notin A))\}.}

ThenDA{\displaystyle D_{A}} is dense. (Given anyp{\displaystyle p}, findn{\displaystyle n} that is not in its domain, and adjoin a value forn{\displaystyle n} contrary to the status of "nA{\displaystyle n\in A}".) Then anypGDA{\displaystyle p\in G\cap D_{A}} witnessesXA{\displaystyle X\neq A}. To summarize,X{\displaystyle X} is a "new" subset ofω{\displaystyle \omega }, necessarily infinite.

Replacingω{\displaystyle \omega } withω×ω2{\displaystyle \omega \times \omega _{2}}, that is, consider instead finite partial functions whose inputs are of the form(n,α){\displaystyle (n,\alpha )}, withn<ω{\displaystyle n<\omega } andα<ω2{\displaystyle \alpha <\omega _{2}}, and whose outputs are0{\displaystyle 0} or1{\displaystyle 1}, one getsω2{\displaystyle \omega _{2}} new subsets ofω{\displaystyle \omega }. They are all distinct, by a density argument: Givenα<β<ω2{\displaystyle \alpha <\beta <\omega _{2}}, let

Dα,β={p(n)(p(n,α)p(n,β))},{\displaystyle D_{\alpha ,\beta }=\{p\mid (\exists n)(p(n,\alpha )\neq p(n,\beta ))\},}

then eachDα,β{\displaystyle D_{\alpha ,\beta }} is dense, and a generic condition in it proves that the αth new set disagrees somewhere with theβ{\displaystyle \beta }th new set.

This is not yet the falsification of the continuum hypothesis. One must prove that no new maps have been introduced which mapω{\displaystyle \omega } ontoω1{\displaystyle \omega _{1}}, orω1{\displaystyle \omega _{1}} ontoω2{\displaystyle \omega _{2}}. For example, if one considers insteadFin(ω,ω1){\displaystyle \operatorname {Fin} (\omega ,\omega _{1})}, finite partial functions fromω{\displaystyle \omega } toω1{\displaystyle \omega _{1}}, thefirst uncountable ordinal, one gets inV[G]{\displaystyle V[G]} a bijection fromω{\displaystyle \omega } toω1{\displaystyle \omega _{1}}. In other words,ω1{\displaystyle \omega _{1}} hascollapsed, and in the forcing extension, is a countable ordinal.

The last step in showing the independence of the continuum hypothesis, then, is to show that Cohen forcing does not collapse cardinals. For this, a sufficient combinatorial property is that all of theantichains of the forcing poset are countable.

The countable chain condition

[edit]
Main article:Countable chain condition

A(strong) antichainA{\displaystyle A} ofP{\displaystyle \mathbb {P} } is a subset such that ifp,qA{\displaystyle p,q\in A} andpq{\displaystyle p\neq q}, thenp{\displaystyle p} andq{\displaystyle q} areincompatible (writtenpq{\displaystyle p\perp q}), meaning there is nor{\displaystyle r} inP{\displaystyle \mathbb {P} } such thatrp{\displaystyle r\leq p} andrq{\displaystyle r\leq q}. In the example on Borel sets, incompatibility means thatpq{\displaystyle p\cap q} has zero measure. In the example on finite partial functions, incompatibility means thatpq{\displaystyle p\cup q} is not a function, in other words,p{\displaystyle p} andq{\displaystyle q} assign different values to some domain input.

P{\displaystyle \mathbb {P} } satisfies thecountable chain condition (c.c.c.) if and only if every antichain inP{\displaystyle \mathbb {P} } is countable. (The name, which is obviously inappropriate, is a holdover from older terminology. Some mathematicians write "c.a.c." for "countable antichain condition".)

It is easy to see thatBor(I){\displaystyle \operatorname {Bor} (I)} satisfies the c.c.c. because the measures add up to at most1{\displaystyle 1}. Also,Fin(E,2){\displaystyle \operatorname {Fin} (E,2)} satisfies the c.c.c., but the proof is more difficult.

Given an uncountable subfamilyWFin(E,2){\displaystyle W\subseteq \operatorname {Fin} (E,2)}, shrinkW{\displaystyle W} to an uncountable subfamilyW0{\displaystyle W_{0}} of sets of size at mostn{\displaystyle n}, for somen<ω{\displaystyle n<\omega } (for somen{\displaystyle n} this is uncountable, since otherwiseW=n<ω{wW:|w|<n}{\displaystyle W=\bigcup _{n<\omega }\{w\in W:|w|<n\}} would be a countable union of countable sets, thus countable). Ifp(e1)=b1{\displaystyle p(e_{1})=b_{1}} for uncountably manypW0{\displaystyle p\in W_{0}}, shrink this to an uncountable subfamilyW1{\displaystyle W_{1}} and repeat, getting a finite set{(e1,b1),,(ek,bk)}W0{\displaystyle \{(e_{1},b_{1}),\ldots ,(e_{k},b_{k})\}\in W_{0}} and an uncountable familyWk{\displaystyle W_{k}} of incompatible conditions of sizenk{\displaystyle n-k} such that everye{\displaystyle e} is inDom(p){\displaystyle \operatorname {Dom} (p)} for at most countable manypWk{\displaystyle p\in W_{k}}. Now, pick an arbitrarypWk{\displaystyle p\in W_{k}}, and pick fromWk{\displaystyle W_{k}} anyq{\displaystyle q} that is not one of the countably many members that have a domain member in common withp{\displaystyle p}. Thenp{(e1,b1),,(ek,bk)}{\displaystyle p\cup \{(e_{1},b_{1}),\ldots ,(e_{k},b_{k})\}} andq{(e1,b1),,(ek,bk)}{\displaystyle q\cup \{(e_{1},b_{1}),\ldots ,(e_{k},b_{k})\}} are compatible, soW{\displaystyle W} is not an antichain. In other words,Fin(E,2){\displaystyle \operatorname {Fin} (E,2)}-antichains are countable.[6]

The importance of antichains in forcing is that for most purposes, dense sets and maximal antichains are equivalent. Amaximal antichainA{\displaystyle A} is one that cannot be extended to a larger antichain. This means that every elementpP{\displaystyle p\in \mathbb {P} } is compatible with some member ofA{\displaystyle A}. The existence of a maximal antichain follows fromZorn's Lemma. Given a maximal antichainA{\displaystyle A}, let

D={pP(qA)(pq)}.{\displaystyle D=\left\{p\in \mathbb {P} \mid (\exists q\in A)(p\leq q)\right\}.}

ThenD{\displaystyle D} is dense, andGD{\displaystyle G\cap D\neq \varnothing } if and only ifGA{\displaystyle G\cap A\neq \varnothing }. Conversely, given a dense setD{\displaystyle D}, Zorn's Lemma shows that there exists a maximal antichainAD{\displaystyle A\subseteq D}, and thenGD{\displaystyle G\cap D\neq \varnothing } if and only ifGA{\displaystyle G\cap A\neq \varnothing }.

Assume thatP{\displaystyle \mathbb {P} } satisfies the c.c.c. Givenx,yV{\displaystyle x,y\in V}, withf:xy{\displaystyle f:x\to y} a function inV[G]{\displaystyle V[G]}, one can approximatef{\displaystyle f} insideV{\displaystyle V} as follows. Letu{\displaystyle u} be a name forf{\displaystyle f} (by the definition ofV[G]{\displaystyle V[G]}) and letp{\displaystyle p} be a condition that forcesu{\displaystyle u} to be a function fromx{\displaystyle x} toy{\displaystyle y}. Define a functionF:xP(y){\displaystyle F:x\to {\mathcal {P}}(y)}, by

F(a)=df{b|(qP)[(qp)(q u(aˇ)=bˇ)]}.{\displaystyle F(a){\stackrel {\text{df}}{=}}\left\{b\left|(\exists q\in \mathbb {P} )\left[(q\leq p)\land \left(q\Vdash ~u\left({\check {a}}\right)={\check {b}}\right)\right]\right\}.\right.}

By the definability of forcing, this definition makes sense withinV{\displaystyle V}. By the coherence of forcing, a differentb{\displaystyle b} must come from an incompatiblep{\displaystyle p}. By c.c.c.,F(a){\displaystyle F(a)} is countable.

In summary,f{\displaystyle f} is unknown inV{\displaystyle V} as it depends onG{\displaystyle G}, but it is not wildly unknown for a c.c.c.-forcing. One can identify a countable set of guesses for what the value off{\displaystyle f} is at any input, independent ofG{\displaystyle G}.

This has the following very important consequence. If inV[G]{\displaystyle V[G]},f:αβ{\displaystyle f:\alpha \to \beta } is a surjection from one infinite ordinal onto another, then there is a surjectiong:ω×αβ{\displaystyle g:\omega \times \alpha \to \beta } inV{\displaystyle V}, and consequently, a surjectionh:αβ{\displaystyle h:\alpha \to \beta } inV{\displaystyle V}. In particular, cardinals cannot collapse. The conclusion is that202{\displaystyle 2^{\aleph _{0}}\geq \aleph _{2}} inV[G]{\displaystyle V[G]}.

Easton forcing

[edit]

The exact value of the continuum in the above Cohen model, and variants likeFin(ω×κ,2){\displaystyle \operatorname {Fin} (\omega \times \kappa ,2)} for cardinalsκ{\displaystyle \kappa } in general, was worked out byRobert M. Solovay, who also worked out how to violateGCH{\displaystyle {\mathsf {GCH}}} (thegeneralized continuum hypothesis), forregular cardinals only, a finite number of times. For example, in the above Cohen model, ifCH{\displaystyle {\mathsf {CH}}} holds inV{\displaystyle V}, then20=2{\displaystyle 2^{\aleph _{0}}=\aleph _{2}} holds inV[G]{\displaystyle V[G]}.

William B. Easton worked out the proper class version of violating theGCH{\displaystyle {\mathsf {GCH}}} for regular cardinals, basically showing that the known restrictions, (monotonicity,Cantor's Theorem andKönig's Theorem), were the onlyZFC{\displaystyle {\mathsf {ZFC}}}-provable restrictions (seeEaston's Theorem).

Easton's work was notable in that it involved forcing with a proper class of conditions. In general, the method of forcing with a proper class of conditions fails to give a model ofZFC{\displaystyle {\mathsf {ZFC}}}. For example, forcing withFin(ω×On,2){\displaystyle \operatorname {Fin} (\omega \times \mathbf {On} ,2)}, whereOn{\displaystyle \mathbf {On} } is the proper class of all ordinals, makes the continuum a proper class. On the other hand, forcing withFin(ω,On){\displaystyle \operatorname {Fin} (\omega ,\mathbf {On} )} introduces a countable enumeration of the ordinals. In both cases, the resultingV[G]{\displaystyle V[G]} is visibly not a model ofZFC{\displaystyle {\mathsf {ZFC}}}.

At one time, it was thought that more sophisticated forcing would also allow an arbitrary variation in the powers ofsingular cardinals. However, this has turned out to be a difficult, subtle and even surprising problem, with several morerestrictions provable inZFC{\displaystyle {\mathsf {ZFC}}} and with the forcing models depending on the consistency of variouslarge-cardinal properties. Many open problems remain.

Random reals

[edit]
Main article:Random algebra

Random forcing can be defined as forcing over the setP{\displaystyle P} of all compact subsets of[0,1]{\displaystyle [0,1]} of positive measure, ordered by the relation{\displaystyle \subseteq } (a smaller set in the context of inclusion is a smaller set in the ordering, and represents a condition with more information). There are two types of important dense sets:

  1. For any positive integern{\displaystyle n}, the setDn={pP:diam(p)<1n}{\displaystyle D_{n}=\left\{p\in P:\operatorname {diam} (p)<{\frac {1}{n}}\right\}} is dense, wherediam(p){\displaystyle \operatorname {diam} (p)} is thediameter of the setp{\displaystyle p}.
  2. For any Borel subsetB[0,1]{\displaystyle B\subseteq [0,1]} of measure 1, the setDB={pP:pB}{\displaystyle D_{B}=\{p\in P:p\subseteq B\}} is dense.

For any filterG{\displaystyle G} and any pair of elementsp1,p2G{\displaystyle p_{1},p_{2}\in G} there isqG{\displaystyle q\in G} such thatqp1,p2{\displaystyle q\leq p_{1},p_{2}}. In this ordering, this means that any filter is closed under finite intersection. Therefore, byCantor's intersection theorem, the intersection ofall the elements in any filter is nonempty. IfG{\displaystyle G} is a filter intersecting the dense setDn{\displaystyle D_{n}} for any positive integern{\displaystyle n}, then the filterG{\displaystyle G} contains conditions of arbitrarily small positive diameter. Therefore, the intersection of all conditions fromG{\displaystyle G} has diameter 0. But the only nonempty sets of diameter 0 are singletons. So there is exactly one real numberrG{\displaystyle r_{G}} such thatrGG{\displaystyle r_{G}\in \bigcap G}.

LetB[0,1]{\displaystyle B\subseteq [0,1]} be any Borel set of measure 1. IfG{\displaystyle G} intersectsDB{\displaystyle D_{B}}, thenrGB{\displaystyle r_{G}\in B}.

However, a generic filter over a countable transitive modelM{\displaystyle M} is not inM{\displaystyle M}. The realrG{\displaystyle r_{G}} defined byG{\displaystyle G} is provably not an element ofM{\displaystyle M}. One issue with this construction is that ifpP{\displaystyle p\in P}, thenM{\displaystyle M\models } "p{\displaystyle p} is compact", but from the viewpoint of some larger universeVM{\displaystyle V\supset M},p{\displaystyle p} can be non-compact and the intersection of all conditions from the generic filterG{\displaystyle G} can then be empty. To fix this, we consider the setC={p¯:pG}{\displaystyle C=\{{\bar {p}}:p\in G\}} of topologicalclosures of conditions fromG{\displaystyle G}. Becausep¯p{\displaystyle {\bar {p}}\supseteq p}, and becauseG{\displaystyle G} is closed under finite intersection, Cantor's intersection theorem applies and the intersection of the setC{\displaystyle C} is nonempty. Sincediam(p¯)=diam(p){\displaystyle \operatorname {diam} ({\bar {p}})=\operatorname {diam} (p)} and the ground modelM{\displaystyle M} inherits a metric from the universeV{\displaystyle V}, the setC{\displaystyle C} has elements of arbitrarily small diameter. Finally, there is exactly one real that belongs to all members of the setC{\displaystyle C}. The generic filterG{\displaystyle G} can be reconstructed fromrG{\displaystyle r_{G}} asG={pP:rGp¯}{\displaystyle G=\{p\in P:r_{G}\in {\bar {p}}\}}.

IfaM(P){\displaystyle a\in M^{(\mathbb {P} )}} is a name forrG{\displaystyle r_{G}} (i.e.,M[G]val(a,G)=rG{\displaystyle M[G]\models val(a,G)=r_{G}}), and forBM{\displaystyle B\in M} holdsM{\displaystyle M\models } "B{\displaystyle B} is a Borel set of measure 1", then by the truth property of forcing

pM,PaBˇ{\displaystyle p\Vdash _{M,\mathbb {P} }a\in {\check {B}}}

for somepG{\displaystyle p\in G}. There is a namea{\displaystyle a} that satisfies

val(a,G)pGp¯{\displaystyle \operatorname {val} (a,G)\in \bigcup _{p\in G}{\bar {p}}}

for any generic filterG{\displaystyle G}. For thata{\displaystyle a},

pM,PaBˇ{\displaystyle p\Vdash _{M,\mathbb {P} }a\in {\check {B}}}

holds for any conditionp{\displaystyle p}.

Every Borel set can be (non-uniquely) built up, starting from intervals with rational endpoints and applying the operations of complement and countable union, a countable number of times. The record of such a construction is called aBorel code. Given a Borel setB{\displaystyle B} inV{\displaystyle V}, one recovers a Borel code, and then applies the same construction sequence inM[G]{\displaystyle M[G]}, getting a Borel setB{\displaystyle B^{*}}. It can be proven that one gets the same set independent of the code chosen forB{\displaystyle B}, and that basic properties are preserved. For example, ifBC{\displaystyle B\subseteq C}, thenBC{\displaystyle B^{*}\subseteq C^{*}}. IfB{\displaystyle B} has measure zero, thenB{\displaystyle B^{*}} has measure zero. This mappingBB{\displaystyle B\mapsto B^{*}} is injective.

For any setB[0,1]{\displaystyle B\subseteq [0,1]} such thatBM{\displaystyle B\in M} andM{\displaystyle M\models } "B{\displaystyle B} is a Borel set of measure 1" one hasrGB{\displaystyle r_{G}\in B^{*}}.

This means thatrG{\displaystyle r_{G}} is an "infinite random sequence of 0s and 1s" from the viewpoint ofM{\displaystyle M}, which means that it satisfies all statistical tests from the ground modelM{\displaystyle M}.[clarification needed]

So givenrG{\displaystyle r_{G}}, a random real, one can show that

G={B (in M)rB (in M[G])}.{\displaystyle G=\left\{B~({\text{in }}M)\mid r\in B^{*}~({\text{in }}M[G])\right\}.}

Because of this mutual inter-definability betweenr{\displaystyle r} andG{\displaystyle G}, one generally writesM[r]{\displaystyle M[r]} forM[G]{\displaystyle M[G]}.

A different interpretation of reals inM[G]{\displaystyle M[G]} was provided byDana Scott. Rational numbers inM[G]{\displaystyle M[G]} have names that correspond to countably-many distinct rational values assigned to a maximal antichain of Borel sets – in other words, a certain rational-valued function onI=[0,1]{\displaystyle I=[0,1]}. Real numbers inM[G]{\displaystyle M[G]} then correspond toDedekind cuts of such functions, that is,measurable functions.

Boolean-valued models

[edit]
Main article:Boolean-valued model

Perhaps more clearly, the method can be explained in terms of Boolean-valued models. In these, any statement is assigned atruth value from some complete atomlessBoolean algebra, rather than just a true/false value. Then anultrafilter is picked in this Boolean algebra, which assigns values true/false to statements of our theory. The point is that the resulting theory has a model that contains this ultrafilter, which can be understood as a new model obtained by extending the old one with this ultrafilter. By picking a Boolean-valued model in an appropriate way, we can get a model that has the desired property. In it, only statements that must be true (are "forced" to be true) will be true, in a sense (since it has this extension/minimality property).

Meta-mathematical explanation

[edit]

In forcing, we usually seek to show that somesentence isconsistent withZFC{\displaystyle {\mathsf {ZFC}}} (or optionally some extension ofZFC{\displaystyle {\mathsf {ZFC}}}). One way to interpret the argument is to assume thatZFC{\displaystyle {\mathsf {ZFC}}} is consistent and then prove thatZFC{\displaystyle {\mathsf {ZFC}}} combined with the newsentence is also consistent.

Each "condition" is a finite piece of information – the idea is that only finite pieces are relevant for consistency, since, by thecompactness theorem, a theory is satisfiable if and only if every finite subset of its axioms is satisfiable. Then we can pick an infinite set of consistent conditions to extend our model. Therefore, assuming the consistency ofZFC{\displaystyle {\mathsf {ZFC}}}, we prove the consistency ofZFC{\displaystyle {\mathsf {ZFC}}} extended by this infinite set.

Logical explanation

[edit]

ByGödel's second incompleteness theorem, one cannot prove the consistency of any sufficiently strong formal theory, such asZFC{\displaystyle {\mathsf {ZFC}}}, using only the axioms of the theory itself, unless the theory is inconsistent. Consequently, mathematicians do not attempt to prove the consistency ofZFC{\displaystyle {\mathsf {ZFC}}} using only the axioms ofZFC{\displaystyle {\mathsf {ZFC}}}, or to prove thatZFC+H{\displaystyle {\mathsf {ZFC}}+H} is consistent for any hypothesisH{\displaystyle H} using onlyZFC+H{\displaystyle {\mathsf {ZFC}}+H}. For this reason, the aim of a consistency proof is to prove the consistency ofZFC+H{\displaystyle {\mathsf {ZFC}}+H} relative to the consistency ofZFC{\displaystyle {\mathsf {ZFC}}}. Such problems are known as problems ofrelative consistency, one of which proves

ZFCCon(ZFC)Con(ZFC+H).{\displaystyle {\mathsf {ZFC}}\vdash \operatorname {Con} ({\mathsf {ZFC}})\rightarrow \operatorname {Con} ({\mathsf {ZFC}}+H).}

The general schema of relative consistency proofs follows. As any proof is finite, it uses only a finite number of axioms:

ZFC+¬Con(ZFC+H)(T)(Fin(T)TZFC(T¬H)).{\displaystyle {\mathsf {ZFC}}+\lnot \operatorname {Con} ({\mathsf {ZFC}}+H)\vdash (\exists T)(\operatorname {Fin} (T)\land T\subseteq {\mathsf {ZFC}}\land (T\vdash \lnot H)).}

For any given proof,ZFC{\displaystyle {\mathsf {ZFC}}} can verify the validity of this proof. This is provable by induction on the length of the proof.

ZFC(T)((T¬H)(ZFC(T¬H))).{\displaystyle {\mathsf {ZFC}}\vdash (\forall T)((T\vdash \lnot H)\rightarrow ({\mathsf {ZFC}}\vdash (T\vdash \lnot H))).}

Then resolve

ZFC+¬Con(ZFC+H)(T)(Fin(T)TZFC(ZFC(T¬H))).{\displaystyle {\mathsf {ZFC}}+\lnot \operatorname {Con} ({\mathsf {ZFC}}+H)\vdash (\exists T)(\operatorname {Fin} (T)\land T\subseteq {\mathsf {ZFC}}\land ({\mathsf {ZFC}}\vdash (T\vdash \lnot H))).}

By proving the following

ZFC(T)(Fin(T)TZFC(ZFCCon(T+H))),{\displaystyle {\mathsf {ZFC}}\vdash (\forall T)(\operatorname {Fin} (T)\land T\subseteq {\mathsf {ZFC}}\rightarrow ({\mathsf {ZFC}}\vdash \operatorname {Con} (T+H))),}⁎⁎

it can be concluded that

ZFC+¬Con(ZFC+H)(T)(Fin(T)TZFC(ZFC(T¬H))(ZFCCon(T+H))),{\displaystyle {\mathsf {ZFC}}+\lnot \operatorname {Con} ({\mathsf {ZFC}}+H)\vdash (\exists T)(\operatorname {Fin} (T)\land T\subseteq {\mathsf {ZFC}}\land ({\mathsf {ZFC}}\vdash (T\vdash \lnot H))\land ({\mathsf {ZFC}}\vdash \operatorname {Con} (T+H))),}

which is equivalent to

ZFC+¬Con(ZFC+H)¬Con(ZFC),{\displaystyle {\mathsf {ZFC}}+\lnot \operatorname {Con} ({\mathsf {ZFC}}+H)\vdash \lnot \operatorname {Con} ({\mathsf {ZFC}}),}

which gives (*). The core of the relative consistency proof is proving (**). AZFC{\displaystyle {\mathsf {ZFC}}} proof ofCon(T+H){\displaystyle \operatorname {Con} (T+H)} can be constructed for any given finite subsetT{\displaystyle T} of theZFC{\displaystyle {\mathsf {ZFC}}} axioms (byZFC{\displaystyle {\mathsf {ZFC}}} instruments of course). (No universal proof ofCon(T+H){\displaystyle \operatorname {Con} (T+H)} of course.)

InZFC{\displaystyle {\mathsf {ZFC}}}, it is provable that for any conditionp{\displaystyle p}, the set of formulas (evaluated by names) forced byp{\displaystyle p} is deductively closed. Furthermore, for anyZFC{\displaystyle {\mathsf {ZFC}}} axiom,ZFC{\displaystyle {\mathsf {ZFC}}} proves that this axiom is forced by1{\displaystyle \mathbf {1} }. Then it suffices to prove that there is at least one condition that forcesH{\displaystyle H}.

In the case of Boolean-valued forcing, the procedure is similar: proving that the Boolean value ofH{\displaystyle H} is not0{\displaystyle \mathbf {0} }.

Another approach uses the Reflection Theorem. For any given finite set ofZFC{\displaystyle {\mathsf {ZFC}}} axioms, there is aZFC{\displaystyle {\mathsf {ZFC}}} proof that this set of axioms has a countable transitive model. For any given finite setT{\displaystyle T} ofZFC{\displaystyle {\mathsf {ZFC}}} axioms, there is a finite setT{\displaystyle T'} ofZFC{\displaystyle {\mathsf {ZFC}}} axioms such thatZFC{\displaystyle {\mathsf {ZFC}}} proves that if a countable transitive modelM{\displaystyle M} satisfiesT{\displaystyle T'}, thenM[G]{\displaystyle M[G]} satisfiesT{\displaystyle T}. By proving that there is finite setT{\displaystyle T''} ofZFC{\displaystyle {\mathsf {ZFC}}} axioms such that if a countable transitive modelM{\displaystyle M} satisfiesT{\displaystyle T''}, thenM[G]{\displaystyle M[G]} satisfies the hypothesisH{\displaystyle H}. Then, for any given finite setT{\displaystyle T} ofZFC{\displaystyle {\mathsf {ZFC}}} axioms,ZFC{\displaystyle {\mathsf {ZFC}}} provesCon(T+H){\displaystyle \operatorname {Con} (T+H)}.

Sometimes in (**), a stronger theoryS{\displaystyle S} thanZFC{\displaystyle {\mathsf {ZFC}}} is used for provingCon(T+H){\displaystyle \operatorname {Con} (T+H)}. Then we have proof of the consistency ofZFC+H{\displaystyle {\mathsf {ZFC}}+H} relative to the consistency ofS{\displaystyle S}. Note thatZFCCon(ZFC)Con(ZFL){\displaystyle {\mathsf {ZFC}}\vdash \operatorname {Con} ({\mathsf {ZFC}})\leftrightarrow \operatorname {Con} ({\mathsf {ZFL}})}, whereZFL{\displaystyle {\mathsf {ZFL}}} isZF+V=L{\displaystyle {\mathsf {ZF}}+V=L} (theaxiom of constructibility).

See also

[edit]

Notes

[edit]
  1. ^abcCohen 2008, p. 111.
  2. ^As a concrete example, note thatα0{\displaystyle \alpha _{0}}, theorder type of all ordinals inM{\displaystyle M}, is a countable ordinal (inV{\displaystyle V}) that is not inM{\displaystyle M}. IfX{\displaystyle X} is taken to be awell-ordering ofN{\displaystyle \mathbb {N} } (as arelation overN{\displaystyle \mathbb {N} }, i.e. a subset ofN×N{\displaystyle \mathbb {N} \times \mathbb {N} }), then anyZFC{\displaystyle {\mathsf {ZFC}}} universe containingX{\displaystyle X} must also containα0{\displaystyle \alpha _{0}} (thanks to theaxiom of replacement).[1] (Such a universe would also not resembleM{\displaystyle M} in the sense that it would collapseall infinite cardinals ofM{\displaystyle M}.)
  3. ^abcShoenfield 1971.
  4. ^Kunen 1980.
  5. ^Notably, if definingM,P{\displaystyle \Vdash _{M,\mathbb {P} }} directly instead ofM,P{\displaystyle \Vdash _{M,\mathbb {P} }^{*}}, one would need to replace the{\displaystyle \vee } with{\displaystyle \wedge } in case 4 and{\displaystyle \exists } with{\displaystyle \forall } in case 5 (in addition to making cases 1 and 2 more complicated) to make this internal definition agree with the external definition. However, then when trying to proveTruth inductively, case 4 will require the fact thatG{\displaystyle G}, as afilter, is downwarddirected, and case 5 will outright break down.
  6. ^Cohen 2008, Section IV.8, Lemma 2.

References

[edit]

Bibliography

[edit]
Overview
Venn diagram of set intersection
Axioms
Operations
  • Concepts
  • Methods
Set types
Theories
Set theorists
Retrieved from "https://en.wikipedia.org/w/index.php?title=Forcing_(mathematics)&oldid=1312995099"
Category:
Hidden categories:

[8]ページ先頭

©2009-2025 Movatter.jp