953
Views
6
CrossRef citations to date
0
Altmetric
Main Articles

Proofs of the Compactness Theorem

Pages 73-98 | Received 02 Feb 2009, Accepted 16 Nov 2009, Published online: 22 Feb 2010
 

Abstract

In this study, several proofs of the compactness theorem for propositional logic with countably many atomic sentences are compared. Thereby some steps are taken towards a systematic philosophical study of the compactness theorem. In addition, some related data and morals for the theory of mathematical explanation are presented.

Acknowledgements

I am grateful to Alan Baker, Andrew Arana, Wilfrid Hodges and two HPL referees for their helpful comments.

Notes

For an excellent introduction to intrinsicness or purity as an ideal of proof taking off from Aristotle see Detflesen 2008. Intrinsicness is semantic purity in the sense of Arana (Citation 2008 ).

Some brief history and bibliography (see Dawson Citation 1993 for more on the history of the compactness theorem): Henkin Citation 1949 is the basis for the so-called Henkin proof. Gödel Citation 1930 contains the first published statement of the compactness theorem. The topological proof is found in various versions in the literature, e.g. in Burris and Sankappanavar Citation 1981 , which also presents the ultraproduct proof offered here. The topological proof is an instance of a more general proof due to Marshall Stone. So far as I am aware, the combinatorial proof is folklore and not due to any particular author. The ultraproduct proof is a corollary of Łoś's theorem, due in essence to Jerzy Łoś. Needless to say, there are more proofs of PL(ω)'s compactness than those mentioned here.

A more constructive argument for the existence of Γmax goes as follows. Let {Ai } i ∊ ω be a (denumerable) enumeration of the set of sentences of PL(ω). Given a finitely satisfiable set of sentences Γ, define the (denumerable) sequence of sets {Γ i } i ∊ ω recursively as follows: Γ0 = Γ; Γ n + 1 = Γ n ∪{An } if every finite subset of Γ n ∪{An } is satisfiable, and Γ n + 1 = Γ n if some finite subset of Γ n ∪{An } is unsatisfiable. We set Γmax = ∪ i ∊ ω i }.

The falsum may either be a primitive symbol of the language or abbreviate an arbitrary contradiction.

Compare the first paragraph of Henkin Citation 1949 for the motivation for the analogous proof of completeness.

Where the variable X is k-adic. A k-ary property of the domain D is a subset of Dk .

By ‘indefinable’ we mean indefinable in the augmented language.

Here, I disagree with Resnik and Kushner (Citation 1987 , pp. 149–150), but this is not the place to engage with their compressed discussion.

A cover of a topological space T is a set of open sets {Oi } i ∊ I such that T = ∪ i ∊ I {Oi }.

A homeomorphism between topological spaces is a continuous bijection with continuous inverse.

For a proof of Tychonoff's Theorem see e.g. Royden Citation 1988 , pp. 196–198.

The proper class {m ∊ ℳ1: m⊧φ} is often written ‘Mod(φ)’. This is a topology since U1) ∩ U2) = {m ∊ ℳ1: m⊧φ1} ∩ {m ∊ ℳ1: m⊧φ2} = {m ∊ ℳ1: m⊧φ1∧φ2} = U1∧φ2).

By definition, [m]⊧φ iff ∃m∗ ∊ [m] s.t. m∗⊧φ. The topology is a projection onto ℳ1/⋍ of the topology on ℳ1.

Of course, we have also used the fact that sentences of the logic form a Boolean algebra.

The domain of ℳ2/⋍2 is the domain ℳ2 of models of second-order logic quotiented by second-order equivalence, symbolised here by ‘⋍2’ Two models in ℳ2 are in the same ⋍2-equivalence class iff they satisfy the same second-order sentences.

Since Na  ∩ Nb  = {U ∊ B∗: a ∊ U} ∩ {U ∊ B∗: b ∊ U} = {U ∊ B∗: ab ∊ U}, where ab is the meet of a and b in B, the Na are a base for this topology.

If Γ⊧ L ⊥ then Γ contains a sentence γ whose interpretation in the standard model of arithmetic is false, hence γ⊧ L ⊥.

We go round full circle by noting that the BPI for the algebra of PL(κ)'s sentences is equivalent to the completeness of PL(κ). Given a deductive system or an equivalence relation ‘≡’ on the set of sentences, we define one in terms of the other by:

We may then verify that the set of sentences quotiented by ≡ is a nontrivial algebra iff ⊢ is consistent.

Following the analogy with topology: a topological space such that any cover for it has a countable subcover is known as a Lindelöf space.

Consider the tree consisting of λ-many disjoint branches of size κi < κ where sup {κ i : i < λ} = κ, for λ = cf(κ) < κ.

For the argument, see e.g. Jech Citation 2003 , pp. 116–117.

For more details of the proof of Cowen's Lemma, which uses Tukey's Lemma, see Cowen Citation 1977 .

There are many other combinatorial proofs of compactness, for instance another by Cowen (Citation 1970 ) based on a combinatorial lemma due to Rado (Citation 1949 ).

Keisler had a vested interest, given that he is responsible for a high proportion of the model-theoretic applications of ultraproducts. The use of ultraproducts is now less fashionable in model theory.

Reflexivity and symmetry are immediate, and transitivity follows from the conjunction and containment properties.

In fact, Dom(Π i ∊ I (Ai )/U) has cardinality 20 (see Bell and Slomson 1969, pp.127–129).

Ded (F) = defxyz((F(x) = zF(y) = z) →x = y) ∧∃xy¬(x = F(y)).

Suppose φ = ¬ψ and that the theorem holds for ψ. Then π U (φ) = π U (¬ψ) = 1−π U (ψ), and π U (ψ) = 1 iff {i ∊ I: Ai (ψ) = 1} ∊ U by the inductive hypothesis. The result follows from the fact that {i ∊ I: Ai (ψ) = 1}∉U iff I/{i ∊ I: A i(ψ) = 1} ∊ U since U is an ultrafilter. This is the only step, either in this proof or in that of Łoś’s theorem for the first-order case, that requires U to be an ultrafilter rather than just a filter. If φ = ψ1∧ψ2 then π U (φ) = 1 iff π U 1) = π U 2) = 1 iff {i ∊ I: Ai 1) = 1} ∊ U and {i ∊ I: Ai 2) = 1} ∊ U, by the inductive hypothesis. If {i ∊ I: Ai 1) = 1} ∊ U and {i ∊ I: Ai 2) = 1} ∊ U then {i ∊ I: Ai 1∧ψ2) = 1} ∊ U, since {i ∊ I: Ai 1∧ψ2) = 1} = {i ∊ I: Ai 1) = 1} ∩ {i ∊ I: Ai 2) = 1} and U is closed under intersection. The other direction follows from the fact that if {i ∊ I: Ai 1∧ψ2) = 1} ∊ U then {i ∊ I: Ai 1) = 1} ∊ U and {i ∊ I: Ai 2) = 1} ∊ U, since U is closed under upward containment and {i ∊ I: Ai 1∧ψ2) = 1} is contained in both {i ∊ I: Ai 1) = 1} and {i ∊ I: Ai 2) = 1}.

Usually with the condition κ ≥ λ.

No extension of first-order logic with a quantifier of the form ‘Q α ’ or ‘Q >ℵα ’ or ‘Q <ℵα or ‘Q ≤ℵα ’ or ‘Q ≥ℵα ’ can be compact (where ‘Q α xφ’ is interpreted as ‘there are ℵα xs satisfying φ’, and so on). However, such logics tend to satisfy weaker notions of compactness, e.g. the logic obtained by adding ‘Q >ℵ0 ’ to first-order logic is (ℵ1, ℵ0)-compact. The completeness proof generalises to this logic: as Keisler (Citation 1970 ) proved, this logic is ℵ1-complete, that is to say, there is a deductive system such that if Γ is a set of sentences of cardinality < ℵ1 then Γ⊧δ iff Γ⊢δ.

See e.g. Jech Citation 2003 , pp. 120–121 and pp. 292–295. Such cardinals are also known as weakly compact, that is, they satisfy the partition property κ→ (κ)2. This is the usual definition of weak compactness; it then comes by way of a theorem that a weakly compact cardinal satisfies the Weak Compactness theorem.

See Jech Citation 2003 , pp. 365–366, for the proof.

For an interesting discussion of some of the issues here, and some others touched on in relation to the compactness theorem, see Dawson Citation 2006 .

The importance of visualisation in this regard has been noted by many writers, e.g. Steiner Citation 1978 , p. 143.

For the importance of memorability as a goodness factor in proofs, see Gowers Citation 2007 .

See Hafner and Mancosu Citation 2008 , pp. 170–171, for the criticism of Philip Kitcher's theory of mathematical explanation that generalisability and unification have a qualitative as well as quantitative dimension.

See especially Kitcher Citation 1981 and 1989. The latter, pp. 423–426 and p. 437, discusses the case of mathematics. The further details of Kitcher's theory will not matter for our discussion.

This is not to say that the notion of compactness is more natural than others in its vicinity. For example, there are many theorems of the form ‘if X is a compact and Hausdorff space then …’. It is therefore arguable which of compact and compact and Hausdorff is the mathematically more natural notion. Indeed, different mathematical cultures go their different ways on this point: for instance Bourbaki calls compactness ‘pseudo-compactness’ and reserves the term ‘compact’ for spaces that are compact and Hausdorff. This suggests that the metaphor of a concept carving mathematical reality at the joints is infelicitous. There are better and worse ways to carve up mathematical reality. But that is not to say that there is one best way of doing it, a unique carving that ‘follows the joints’.

This is not to say that they are correct as accounts of mathematical value. Indeed they cannot be if intrinsicness is an aspect of explanatoriness.

Compare the first, logical, sense of purity in Arana Citation 2008 .

Sandborg Citation 1998 is an outline of a theory of mathematical explanation that takes context-dependence seriously, applying van Fraassen's account of explanation to the mathematical case.

Log in via your institution

Log in to Taylor & Francis Online

PDF download + Online access

  • 48 hours access to article PDF & online version
  • Article PDF can be downloaded
  • Article PDF can be printed
USD 53.00 Add to cart

Issue Purchase

  • 30 days online access to complete issue
  • Article PDFs can be downloaded
  • Article PDFs can be printed
USD 490.00 Add to cart

* Local tax will be added as applicable

Related Research

People also read lists articles that other readers of this article have read.

Recommended articles lists articles that we recommend and is powered by our AI driven recommendation engine.

Cited by lists all citing articles based on Crossref citations.
Articles with the Crossref icon will open in a new tab.