Movatterモバイル変換


[0]ホーム

URL:


SEP home page
Stanford Encyclopedia of Philosophy

Second-order and Higher-order Logic

First published Thu Aug 1, 2019; substantive revision Sat Aug 31, 2024

Second-order logic has a subtle role in the philosophy of mathematics.It is stronger than first order logic in that it incorporates“for all properties” into the syntax, while first orderlogic can only say “for all elements”. At the same time itis arguably weaker than set theory in that its quantifiers range overone limited domain at a time, while set theory has the universalistapproach in that its quantifiers range over all possible domains. Thisstronger-than-first-order-logic/weaker-than-set-theory duality is thesource of lively debate, not least because set theory is usuallyconstrued as based on first order logic. How can second-order logic beat the same time stronger and weaker? To make the situation even morecomplex, it was suggested early on that in order that the strength ofsecond-order logic can be exploited to its full extent and in orderthat “for all properties” can be given an exactinterpretation, a first order set-theoretical background has to beassumed. This seemed to undermine the claimed strength of second-orderlogic as well as its role as the primary foundation of mathematics. Italso seemed to attach second-order logic to aspects of set theorywhich second-order logic might have wanted to bypass: the higherinfinite, the independence results, and the difficulties in findingnew convincing axioms. Setting aside philosophical questions, it isundeniable and manifested by a continued stream of interestingresults, that second-order logic is part and parcel of alogician’s toolbox.


1. Introduction

Second-order logic[1] was introduced by Frege in his Begriffsschrift (1879) who also coinedthe term “second order” (“zweiterOrdnung”) in (1884: §53). It was widely used in logicuntil the 1930s, when set theory started to take over as a foundationof mathematics. It is difficult to say exactly why this happened, butset theory has certain simplicity in being based on one single binarypredicate \(x\in y\), compared to second- and higher-order logics,including type theory. Discussion among philosophers of the merits ofsecond-order logic especially in comparison to set theory continues tothis day (Shapiro 1991). Meanwhile second-order logic has become astandard tool of computer science logic, finite model theory inparticular. Central questions of theoretical computer science, such asthe P = NP? question, can be seen as questions about second-orderlogic in the finite context.

Ultimately second-order logic is just another formal language. It canbe given the standard treatment making syntax, semantics and prooftheory exact by working in a mathematical metatheory which we chooseto be set theory. It would be possible to use second-order logicitself as metatheory, but it would be more complicated, simply becausesecond-order logic is less developed than set theory.

The distinction between first order logic and second-order logic iseasiest to understand in elementary mathematics. Let us considernumber theory. The objects of our study are the natural numbers 0, 1,2,… and their arithmetic. With first order logic we canformulate statements about number theory by using atomic expressions\(x = y,\) \(x+y = z\) and \(x\times y = z\) combined with thepropositional operations \(\land,\) \(\neg,\) \(\lor,\) \(\to\) andthe quantifiers \(\forall x\) and \(\exists x\). Here the variablesx,y,z,… are thought to range over thenatural numbers. With second-order logic we can express more: we havevariablesx,y,z,… as in first orderlogic for numbers. In addition we have variablesX,Y,Z,… for properties of numbers and relations betweennumbers as well as quantifiers \(\forall X\) and \(\exists X\) forthese variables. We have the atomic expressions of first order logicand also new atomic expressions of the form \(X(y_1,\ldots,y_n)\).

An interesting and much studied question arises: what properties ofnatural numbers can be expressed in first order logic and what can beexpressed in second-order logic? The question is equally interestingif “expressed” is replaced by “proved”, once asuitable proof system is given in both cases. We can replace naturalnumbers by some other mathematical context, for example real numbers,complex numbers, and so on.

With just the constant \(0\) and the unary function \(n \mapsto n^+\)(where \(n^+\) means \(n+1\)) we can express in second-order logic theInduction Axiom

\[\tag{1}\label{ind} \forall X\left(\left[X(0) \land \forally\left(X(y) \to X(y^+)\right)\right] \to \forall y\,X(y)\right), \]

of natural numbers. This, together with the axioms \(\forallx\,\neg(x^+=0)\) and \(\forall x\,\forall y(x^+=y^+\to x=y)\)characterizes the natural numbers with their successor operation up toisomorphism. In first order logic any theory which has a countablyinfinite model has also an uncountable model (by the UpwardLöwenheim Skolem Theorem). Hence (\ref{ind}) cannot be expressedin first order logic. Another typical second-order expression is theCompleteness Axiom of the linear order \(\le\) of the realnumbers:

\[\tag{2}\label{comp} \begin{align}&\forall X\left(\left[\exists y\,X(y) \land \exists z\,\forall y(X(y) \to y \le z)\right]\right. \to\\ &\qquad\left.\exists z\,\forall y\left(\exists u\left(X(u) \land y \le u\right) \lor z \le y\right)\right),\end{align}\]

This, together with the axioms of ordered fields characterizes theordered field of real numbers up to isomorphism. In first order logicany countable theory which has an infinite model has also a countablemodel (by the Downward Löwenheim Skolem Theorem). Hence, again,(\ref{comp}) cannot be expressed in first order logic.

To early logicians it seemed natural to use second-order logic on apar with first order logic, as if there were not much difference. Thisincluded Russell, Löwenheim, Hilbert, and Zermelo. Essentiallyonly Skolem emphasized the difference (see Moore 1988 for theemergence of first order logic). In 1929 Gödel proved hisCompleteness Theorem and the next year his Incompleteness Theorem. Itbecame gradually obvious that second-order logic is very differentfrom first order logic. One of the first indicators was theincompleteness phenomenon. Gödel showed that any effectiveaxiomatization of number theory is incomplete. On the other hand,there was a simple finite categorical—hence complete (§10)—axiomatization of the structure \((\oN,\) \(+,\) \(\times)\) in second-order logic(see also the discussion related to (\ref{ind})). This showed thatthere cannot be such a complete axiomatization of second-order logicas there was for first order logic. What became known in the case offirst order logic as Gödel’s Completeness Theorem simplycannot hold for second-order logic.

The situation changed somewhat when Henkin proved the CompletenessTheorem for second-order logic with respect to so-called generalmodels (§9). Now it became possible to use second-order logic in the same way asfirst order logic, if only one keeps in mind that the semantics isbased on general models. The ability of second-order logic tocharacterize mathematical structures up to isomorphism does not extendto general models (except in the form of internalcategoricity—see§10). Any second-order theory, in a countable vocabulary, with an infinitegeneral model has general models ofall infinitecardinalities, and therefore cannot be categorical in the context ofgeneral models. We shall discuss general models and the related Henkinmodels in§9.

2. The Syntax of Second-Order Logic

Avocabulary in second-order logic is just as a vocabulary infirst order logic, that is, a setL ofrelation,function andconstant symbols. Each relation andfunction symbol has anarity, which is a positive naturalnumber.

Second-order logic has several kinds ofvariables. It hasindividualvariables denoted by lower case letters\(x, y, z, \ldots\) possibly with subscripts. It haspropertyandrelationvariables denoted by upper case letters\(X,Y,Z,\ldots\) possibly with subscripts. Finally, it hasfunction variables denoted by upper case letters \(F, G, H,\ldots\) possibly with subscripts. Sometimes function variables areomitted as, after all, functions are special kinds of relations. Eachrelation and function variable has an arity, which is a positivenatural number. We identify property variables with 1-ary relationvariables.

It is noteworthy that although we have property variables we do nothave variables for properties of properties. Such variables would bepart of the formalism of third order logic, see§12.

Theterms of second-order logic are defined recursively asfollows: Constant symbols and individual variables are terms. If\(t_1,\ldots,t_n\) are terms,U is ann-ary functionsymbol andF is ann-ary function variable, then\(U(t_1,\ldots,t_n)\) and \(F(t_1,\ldots,t_n)\) are terms. Note thatterms denote individuals, not relations or properties. ThusXalone is not a term butx is.

Theatomic formulas of second-order logic are defined fromterms as follows: Ift and \(t'\) are terms, then \(t = t'\) isan atomic formula. IfR is ann-ary relation symbol and\(t_1,\ldots,t_n\) are terms, then \(R(t_1,\ldots,t_n)\) is an atomicformula. IfX is ann-ary relation variable, then also\(X(t_1,\ldots,t_n)\) is an atomic formula. The intuitive meaning of\(X(t_1,\ldots,t_n)\) is that the elements \(t_1,\ldots, t_n\) are inthe relationX or arepredicated byX. We donot allow atomic formulas of the form \(X = Y\), although they wouldhave an obvious meaning. We achieve the same effect by usingquantifiers.

Definition 1 Theformulas of second-orderlogic are defined as follows: Atomic formulas are formulas. If\(\phi\) and \(\psi\) are formulas, then \(\neg\phi\), \(\phi\land\psi\), \(\phi\lor \psi\), \(\phi \to \psi\) and \(\phi\leftrightarrow \psi\) are formulas. If \(\phi\) is a formula,x an individual variable,X a relation variable andF a function variable, then \(\exists x\phi\), \(\forallx\phi\), \(\exists X\phi\), \(\forall X\phi, \exists F\phi\) and\(\forall F\phi\) are formulas.

It is interesting to note that in second order logic we can actuallydefine the identity \(t=t'\) as \(\forallX(X(t)\leftrightarrow X(t'))\) and prove the familiar axioms ofidentity from properties of the implication.

An important special case ismonadic second-order logic whereno function variables are allowed and the relation variables arerequired to be monadic (a.k.a. unary), i.e., of arity one.

We did not take \(X=Y\) as an atomic formula (although we could have)but having introduced the quantifiers we can use

\[\tag{3}\label{XY} \forall x_1\ldots \forallx_n(X(x_1,\ldots,x_n)\leftrightarrow Y(x_1,\ldots,x_n)) \]

as a substitute for \(X=Y\). The advantage of taking \(X=Y\) as abasic atomic formula would be that using (\ref{XY}) brings along extraquantifiers. Sometimes it is interesting to minimize the number ofquantifiers in a formula. Also, (\ref{XY}) gives the identity \(X=Y\)anextensional flavour in contrast to a possibly differentintensional construal.

The concepts of afree andboundoccurrenceof a variable in a formula are defined in the usual way. A formula iscalled asentence if it has no free variables. The concept ofa term beingfreefor a variable in a formula isdefined as in first order logic.

3. The Semantics of Second-Order Logic

In order to define the semantics of second-order logic we have toagree what ourmetalanguage is. This fact has nothing to dowith second-order logic but is rather a general feature of semantics(Tarski 1933 [1956]). The most commonly used metalanguage forsecond-order logic isset theory. We thus give aset-theoretical interpretation of second-order logic, interpreting“properties” as sets in a domain which itself is a set.This is the most common choice and brings out the main features ofsecond-order logic. We cannot interpretall properties by sets,e.g. the property of being identical to oneself. But if the domainthat the individual variables range over is taken to be a set, then wecan meaningfully interpret properties of individuals in that domain assets. If we want to interpret second-order logic in a domain which istoo large to be a set, we can use the set-theoretical concept of aclass (see entry onset theory for more on classes).

We use the same concept of anL-structure (or equivalently, anL-model) as in first order logic. That is, anL-structure \(\mm\) has adomainM,which is any non-empty set, aninterpretation \(c^\mm \in M\)of any constant symbolc ofL, an interpretation \(R^\mm\subseteq M^n\) of anyn-ary relation symbolR ofL, as well as an interpretation \(H^\mm : M^n \to M\) of anyn-ary function symbolH ofL. In principle, thedomain could be a class but we disregard this possibility here. For adiscussion on the set/class distinction in this context, we refer toShapiro (1991, p. 16).

Convention: Whenever we use upper case Fraktur letters, such as\(\mm,\mn\) etc, to name a structure, we use the upper case Italicversions of those letters, such as \(M,N\) etc, to name the domain ofthe structure.

Given anL-structure \(\mm\), anassignment is afunctions from variables to the domainM of \(\mm\)such that: ifx is an individual variable, then \(s(x) \in M\);ifX is relation variable of arityn, then \(s(X)\subseteq M^n\); ifF is function variable of arityn,then \(s(F) : M^n \to M\). We use \(s(P/X)\) to denote the assignmentwhich is otherwise ass except that the value atX hasbeen changed toP. Similarly \(s(a/x)\) and \(s(f/F)\).

Note that \(0\)-ary relation variables are essentially propositions.Their interpretations under an assignment are the truth values (true,false). \(0\)-ary function variables are essentially individualvariables as an assignment maps a \(0\)-ary function symbol simply toan element of \(M\). A \(0\)-ary relation symbol or variable is at thesame time an atomic formula and a non-logical symbol. We disregards inthe current presentation this potential notational confusion.

The value \(t^\mm\langle s\rangle\) of a termt in a model\(\mm\) under the interpretations is defined as in first orderlogic.

The concept \(\mm\models\phi\) of thetruth of the sentence\(\phi\) in \(\mm\) can now be defined by first defining the auxiliaryconcept \(\mm\models_s\phi\) of the assignment \(s\)satisfying\(\phi\) in \(\mm\) (see entry onTarski’s truth definitions for more details):

Definition 2 (Tarski’s Truth Definition) Thetruth definition for second-order logic extends the respective truthdefinition for first order logic by the clauses:

  1. \(\mm \models_s X(t_1,\ldots,t_n)\) iff \((t^\mm_1 \langles\rangle,\ldots,t^\mm_n \langle s\rangle) \in s(X)\), if \(X\) is\(n\)-ary.
  2. \(\mm\models_s \exists X\,\phi\) iff \(\mm\models_{s(P/X)} \phi\)for some \(P \subseteq M^n\), if \(X\) is \(n\)-ary.
  3. \(\mm\models_s \exists F\,\phi\) iff \(\mm\models_{s(f/F)} \phi\)for some \(f:M^n \to M\), if \(F\) is \(n\)-ary.

and similarly for the universal quantifiers. For a sentence \(\phi\)we define \(\mm\models\phi\) to mean \(\mm\models_s\phi\) for all(equivalently some)s, and then say \(\phi\) istruein \(\mm\).

In clause 1 of the above definition we follow the common practice ofgiving a set-theoretical interpretation to the predication\(X(t_1,\ldots,t_n)\). Having interpretedX as a subset of\(M^n\), we interpret predication \(X(t_1,\ldots,t_n)\) as membership\((t^\mm_1 \langle s\rangle,\ldots,t^\mm_n \langle s\rangle) \ins(X)\). In clause 2 we interpret quantification over properties andrelations as quantification over subsets of the cartesian product\(M^n\). Respectively, in clause 3 we interpret quantification overfunctions as quantification over sets that are functions \(M^n\to M\)in the set-theoretical sense.

Now that the semantics of second-order logic is defined we can definewhat it means for a formula \(\phi\) to bevalid and for twoformulas \(\phi\) and \(\psi\) to belogically equivalent. Asin logic in general, we say that \(\phi\) is (logically) valid if\(\mm\models_s\phi\) holds for all \(\mm\) and alls. Likewise,we define \(\phi\) and \(\psi\) to be logically equivalent,\(\phi\equiv\psi\), if \(\phi\leftrightarrow\psi\) is valid. Twomodels \(\mm\) and \(\mn\) are said to besecond-orderequivalent, in symbols \(\mm\equiv_{L^2}\mn\) if for allsentences \(\phi\) we have \(\mm\models\phi\iff\mn\models\psi\).

The truth definition involves the concepts “for some \(P\subseteq M^n\)” and “for some \(f:M^n \to M\)”.Since we assume set theory as our metalanguage, we can interpret thesein the sense of set theory. Thus we interpret “for some \(P\subseteq M^n\)” as “for someset \(P \subseteqM^n\)” and “for some \(f:M^n\to M\)” as “forsomeset which is a function \(f:M^n\to M\)”.

For infiniteM the collection of subsets of \(M^n\) and the setof functions \(M^n\to M\) are famously complex. Currently set theory(ZFC) cannot even decide howmany subsets of an infinite setthere are. The situation is quite different with first order logic.The respective concept “for some \(a \in M\)” isunproblematic, relatively speaking. Of course, depending on whatM is, “for some \(a \in M\)” can be quiteproblematic too. To reflect the difficulties involved with finding a\(P \subseteq M^n\) or an \(f:M^n \to M\) we sometimes say we“guess” a \(P \subseteq M^n\) or an \(f:M^n \to M\).

When we use set theory as the metatheory for the semantics offirst order logic, the reliance on the metatheory is of alower degree than when we use set theory as the metatheory for thesemantics ofsecond order logic. The central concept, whichexplains the difference, is that ofabsoluteness. Fordetails, see§6.

3.1 The Ehrenfeucht-Fraïssé game of second-order logic

TheEhrenfeucht-Fraïssé game is a game-theoretictool for investigating to what extent two models are similar (seeentry onlogic and games for a general introduction to Ehrenfeucht-Fraïssé games).Two isomorphic models would be very similar but normally we areinterested in similarity of models that are not isomorphic. TheEhrenfeucht-Fraïssé game of second-order logiccharacterizes \(\ma\equiv_{L^2}\mn\), i.e. the proposition thatexactly the same second-order sentences are true in \(\ma\) and\(\mb\), just as the Ehrenfeucht-Fraïssé game of firstorder logic characterizes the first order elementary equivalence\(\ma\equiv\mn\). See the entry onfirst-order model theory for more on elementary equivalence.

For simplicity we disallow function and constant symbols as well asfunction variables in this section. Suppose \(\ma\) and \(\mb\) aretwo models of the same finite relational vocabulary. In the game whichwe denote by \(G^2_n(\ma,\mb)\) two players I and II pick one at atime subsets (or elements) ofA or subsets (or elements) ofB. During roundi of the game player I can pick arelation \(A_i\) onA (or an element \(a_i\) ofA) andthen player II has to pick a relation \(B_i\) onB of the samearity as \(A_i\) (or an element \(b_i\) of onB) and viceversa: Player I can instead pick a relation \(B_i\) onB (or anelement \(b_i\) ofB) and then II picks a relation \(A_i\) onA of the same arity as \(B_i\) (or an element \(a_i\)) ofA. Aftern rounds the pairs of played elements\((a_i,b_i)\) form a binary relationR on \(A\times B\). Ifthis relation is a partial isomorphism of the structures \(\ma\) and\(\mb\) expanded by the played relations \(A_i\) and \(B_i\), i.e., itpreserves atomic formulas and their negations, we say that II has won.The models \(\ma\) and \(\mb\) satisfy the same second-order sentencesif and only if for each \(n\in \oN\) Player II has a winning strategyin \(G^2_n(\ma,\mb)\). A model class (i.e. a class of models of afixed vocabulary, closed under isomorphisms)K is definable insecond-order logic if and only if there is ann such that if\(\ma\in K\) and Player II has a winning strategy in the\(G^2_n(\ma,\mb)\), then \(\mb\in K\).

The first order version \(G^1_n(\ma,\mb)\), where players play onlyindividual elements, i.e., no relations are played, is very useful inworking with first order logic. Unfortunately the game\(G^2_n(\ma,\mb)\) is much more complex. It is very difficult toinvent winning strategies for Player II except in the trivial casewhen \(\ma\cong\mb\). If Player I plays a binary relationR onA, Player II should be able to play a binary relation \(R'\) onB such that whatever challenges Player I makes in the rest ofthe game, even involving new binary relations, the relationsRand \(R'\) look similar. If \(V=L\), then countable second-orderequivalent models (with a finite vocabulary) are actually isomorphic(Ajtai 1979). For details, see§10. Thus consistently in countable models there is no other winningstrategy for Player II than an isomorphism. Perhaps this explains whythe game is not as useful in second-order logic as it is in firstorder logic. However, if we restrict to monadic second-order logic,which in terms of the Ehrenfeucht-Fraïssé game meansrestricting the game tounary predicates, the situationchanges. A unary predicate just divides the model into two parts. IfPlayer I dividesA into two parts, Player II should find asimilar division inB. This is more reasonable and thereactually are useful strategies for Player II. For examples in thefinite context see§16.

4. Properties of Second-Order Formulas

Many syntactic operations that we are used to in first order logicwork equally well in second-order logic. One such is the operation ofrelativization, in which we want to limit what a formulaexpresses to a fixed part of the universe. For example, we mayconsider models where a certain unary predicateU is used forthe set of natural numbers. Then we relativize our axioms ofarithmetic to the predicateU. Some other part of the model maybe used for the power set of the set of natural numbers. Then werelativize our axioms of power set (see§5.3) to that part, and so on.

If \(\mm\) is a structure and \(A\subseteq M\), then \(\mm^{A}\) isobtained by restricting the relations \(R^\mm\) and functions\(F^\mm\) toA. This does not always result in a structure butif it does, it is called therelativization of \(\mm\) toA. Suppose\(\phi=\phi(x_1,\ldots,x_n,X_1,\ldots,X_m,F_1,\ldots,F_k)\) is asecond-order formula andU is a unary relation variable. Thereis a second-order formula \(\phi^{U}\), the relativization of \(\phi\)toU, such that for all \(\mm\) for which \(\mm^{U^\mm}\) is astructure:

\[\mm\models\phi^{U}\iff \mm^{U^\mm}\models\phi.\]

Intuitively, \(\phi^{U}\) says about \(U^\mm\) what \(\phi\) saysaboutM. To obtain \(\phi^{U}\) from \(\phi\) one restricts allthe first order quantifiers to elements ofU and second-orderquantifiers to subsets, relations and functions onU.

Hierarchies based on quantifier structure are a useful method tocompare the definability of concepts in different systems. In firstorder logic it was observed very early on that formulas can be broughtinto a logically equivalentPrenex Normal Form in which allquantifiers occur in the beginning of the formula. This is possiblealso in second-order logic and the proof is essentially the same.Moreover, using the equivalence (which is reminiscent of a principleof set theory called Axiom of Choice, to which we return in Section5.4., since in the below formula the relation \(Y\) in a sense choosesone \(X\) for each \(x\) and puts them together into one relation\(Y\) – see entry onAxiom of Choice for more on the topic.)

\[\forall x\,\exists X\,\phi\equiv \exists Y\,\forall x\,\phi',\]

where the arity ofY is one higher than the arity ofX,and \(\phi'\) is obtained from \(\phi\) by replacing everywhere\(X(t_1,\ldots,t_n)\) by \(Y(x,t_1,\ldots,t_n)\), one can bring everysecond-order logic formula to a logically equivalent normal form

\[\tag{4}\label{qqq}Q_1X_1\ldots Q_nX_n\phi,\]

where each \(Q_i\) is either \(\exists\) or \(\forall\) and \(\phi\)is first order. That is, all second order quantifiers of the formulaare in the front.

NotationThe \(Q_i\) in (\ref{qqq}) are:
\(\Sigma^1_1\)All existential
\(\Pi^1_1\)All universal
\(\Sigma^1_2\)First existential, then universal
\(\Pi^1_2\)First universal, then existential
\(\vdots\)\(\vdots\)
NotationThe formula is logically equivalent to:
\(\Delta^1_1\)Both \(\Sigma^1_1\) and \(\Pi^1_1\) formulas
\(\Delta^1_2\)Both \(\Sigma^1_2\) and \(\Pi^1_2\) formulas
\(\vdots\)\(\vdots\)

Table 1: The hierarchy of second-orderformulas.

By restricting what kind of quantifiers the sequence \(Q_1X_1\ldotsQ_nX_n\) of the Prenex Normal Form (\ref{qqq}) can contain we obtainthe classes of \(\Sigma^1_n\)-, \(\Pi^1_n\)- and\(\Delta^1_n\)-formulas, as indicated inTable 1. For example, the second-order formulas (1) and (2) are\(\Pi^1_1\)-formulas.

We have considered above only quantifiers that bind relation variablesbut the situation is exactly the same if we had bound functionvariables.

These classes of formulas have some obvious closure properties: Theyare all closed, up to logical equivalence, under \(\land , \lor\), andfirst order quantifiers. Moreover, the hierarchy is proper in thesense that for all \(n\ge 1\)

\[\tag{5}\label{kurat}\Sigma^1_n \nsubseteq \Pi^1_n\text{ and } \Pi^1_n \nsubseteq \Sigma^1_n,\]

which can be proved with a diagonalisation argument. The classes\(\Delta^1_n\) are closed under \(\land , \lor, \neg\) and first orderquantifiers. By Craig’s Interpolation Theorem,\(\Delta^1_1\)-formulas are logically equivalent to first orderformulas[2].

There is a sense in which \(\Pi^1_1\), or equivalently \(\Sigma^1_1\),already has the full power of second-order logic, although the abovehierarchy result (\ref{kurat}) shows that this cannot literally betrue. To see this, i.e., to see what the reduction of second-orderlogic to its \(\Pi^1_1\)-part means, letL be a vocabulary,E a binary andU a unary relation symbol, both not inL. Let \(\theta\) be the second-order\(\Pi^1_1\)-formula

\[\tag{6}\label{ex6}\forall X(\phi_1\land\phi_2\land\phi_3),\]

where

\[\begin{align}\phi_1&\text{ is }\quad \forall x\,\forall y(\forall z(E(z,x)\leftrightarrow E(z,y))\to x=y)\\\phi_2&\text{ is }\quad \forall x\,\forall y(E(x,y)\to (U(x)\land\neg U(y)))\\\phi_3&\text{ is }\quad \forall x (X(x)\to U(x))\to\exists y\,\forall z(X(z)\leftrightarrow E(z,y)).\end{align}\]

It is easy to see[3] that models of \(\theta\) are, up to isomorphism, models \(\mm\)where \(M=U^\mm\cup\P(U^\mm)\), \(U^\mm\cap\P(U^\mm)=\emptyset\) andfor all \(a,b\in M\), \(E^\mm(a,b)\leftrightarrow a\in b\). In such amodel monadic second-order formulas \(\phi\) over \(U^\mm\) can bereduced to first order formulas \(\phi^*\) overM. Thisreduction, due to Hintikka (1955) and further developed by Montague(1963), shows that \(\Pi^1_1\)-formulas alone are enough to expressany second-order property with extra predicates. The idea can beextended to third and higher order logic, and also to non-monadicsecond-order logic.

As a consequence, checking the validity of a second-order sentence\(\phi\) can be recursively reduced to checking the validity of the\(\Sigma^1_1\)-sentence \(\theta\to\phi^*\). Likewise checking whethera second-order sentence \(\phi\) has a model of cardinality \(\kappa\)can be reduced to asking whether the \(\Pi^1_1\)-sentence\(\theta\land\phi^*\) has a model of cardinality \(2^\kappa\). Thismeans that the Löwenheim number[4] and the Hanf number[5] of the entire second-order logic are the same as those of thefragment \(\Pi^1_1\).

Summing up, upon first inspection the levels \(\Sigma^1_n\) and\(\Pi^1_n\) of the hierarchy of second-order formulas grow strictly inexpressive power asn increases, but a more careful analysisreveals that already the first level \(\Sigma^1_1\cup \Pi^1_1\) hasthe power of all the levels \(\Sigma^1_n, \Pi^1_n\) even if the poweris somewhat implicit and needs to be brought to light.

In set theory there is theLevy-hierarchy of \(\Sigma_n\)-and \(\Pi_n\)-formulas (Lévy 1965). Formulas where allquantifiers are bounded, i.e., of the form \(\forall x\in y\) or ofthe form \(\exists x\in y\) for somex andy, are called\(\Sigma_0\) or equivalently \(\Pi_0\). Formulas of the form \(\forallx\,\phi\), where \(\phi\) is \(\Sigma_n\), are called \(\Pi_{n+1}\).Formulas of the form \(\exists x\,\phi\), where \(\phi\) is \(\Pi_n\),are called \(\Sigma_{n+1}\). This is a strict hierarchy roughly forthe same reason why the hierarchy of second-order \(\Sigma^1_n\)- and\(\Pi^1_n\)-formulas is strict. But there is no known method to reducethe truth of an arbitrary formula to the truth of a formula on the\(\Sigma_1\cup\Pi_1\) level. In fact, if the decision problem,Löwenheim-Skolem number and Hanf number are suitably defined for\(\Sigma_n\cup \Pi_n\)-formulas of set theory, the decision problemgets more and more complicated, and the numbers obtained get biggerand bigger asn increases (seeTheorem 4; Väänänen 1979).

The hierarchy \(\Sigma_n^1\cup \Pi_n^1\) inside second-order logic andthe hierarchy \(\Sigma_n\cup\Pi_n\) in set theory, are furthercompared to each other in§7.

5. The Infamous Power of Second-Order Logic

5.1 Putting distance between second- and first-order logic

Second-order logic has some remarkable qualities which we will nowuncover in detail. Above we characterized the structure of the naturalnumbers by means of (\ref{ind}). This application of second-orderlogic is perhaps the most famous and most important. We nowreformulate this in a slightly more general form in order to get morepower out of it. SupposeU is a unary relation variable(“the set of natural numbers”),G a unary functionvariable (“the successor function”), andz anindividual variable (“the zero”). Let\(\theta_{\textrm{PA}}(U,G,z)\) be the sentence

\[\tag{7}\begin{align}&U(z)\land{}\\&\forall x(U(x)\to(U(G(x))\land\neg G(x)=z))\land{}\\&\forall x\,\forall y((U(x)\land U(y)\land G(x)=G(y))\to x=y)\land{}\\&\forall X([X(z) \land \forall x((U(x)\land X(x)) \to X(G(x)))] \to {}\\&\forall x(U(x)\to X(x))),\end{align}\]

Then \((N,A,g,a)\models\theta_{\textrm{PA}}(U,G,z)\) if and only if\((N,A,g,a)\) is isomorphic to a structure \((M,\oN,s,0)\) where\(\oN\subseteq M\) and \(s(n)=n+1\). Thus we have used second-orderlogic to say that theU-part of any model of\(\theta_{\textrm{PA}}(U,G,z)\), together with the functionGand the individualz are isomorphic to the natural numbers\(\oN\) with their successor function and zero. There are many ways tosee that no first order sentence can have this property. For example,by the Compactness Theorem such a sentence would have also anon-standard model where some element is different from eachin the sequencez, \(G(z)\), \(G(G(z))\), \(G(G(G(z)))\),\(\ldots\). Such a model cannot be isomorphic to the natural numbers\(\oN\) with their successor function and zero. Hereupon we mayconclude that the Compactness Theorem does not hold for second-orderlogic in the form that it holds for first order logic.

In models of \(\theta_{\textrm{PA}}(U,G,z)\) the formula\(\theta_+(U,G,z,H)\)

\[\left\{\begin{array}{l}\forall x(U(x)\to H(x,z)=x)\land\\\forall x\,\forall y(U(x)\to (U(y)\to H(x,G(y))=G(H(x,y))))\\\end{array}\right.\]

and the formula \(\theta_\times(U,G,z,H,J)\)

\[\left\{\begin{array}{l}\forall x(U(x)\to J(x,z)=z)\land\\\forall x\,\forall y(U(x)\to (U(y)\to J(x,G(y))=H(J(x,y),x)))\\\end{array}\right.\]

define unique functions \(m+n=H(m,n)\) and \(m\cdot n=J(m,n)\). Afirst order sentence \(\phi(+,\cdot,0)\) of arithmetic is true in\((\oN,+,\cdot,0)\) if and only if

\[\begin{align}&\forall U\,\forall G\,\forall z\,\forall H\,\forall J \\&\ \left(\left(\theta_{\textrm{PA}} (U,G,z) \land \theta_+(U,G,z,H) \land \theta_\times(U,G,z,H,J)\right)\to (\phi(H,J,z))^{U}\right)\end{align}\]

is valid. We know that truth of first order sentences in\((\oN,+,\cdot,0)\) is undecidable, and even non-arithmetical byresults of Gödel (1931 [1986: 144–195]) and Tarski (1933[1956]). This shows that second-order logic is not completelyaxiomatizable by effective means, or decidable even in the emptyvocabulary. That is, there is no decidable second-order theory with aninfinite model. This is in sharp contrast to first order logic wherethe following theories are decidable: first order logic with emptyvocabulary, first order logic with monadic predicates, dense linearorder, Pressburger arithmetic (first order theory of \((\oN,+,0)\)),real closed fields (first order theory of \((\oR,+,\cdot,0,1)\)),theory of algebraically closed fields, and the first order theory of\((\oC,+,\cdot,0,1)\). In second-order logic none of the respectivesecond-order theories are decidable. However, if we restrictsecond-order logic tomonadic second-order logic, we obtainmany important decidable theories (see§8).

5.2 The collapse of the Completeness Theorem

The Completeness Theorem is one of the cornerstones of ourunderstanding of first order logic. Also many extensions of firstorder logic have a Completeness Theorem, most notably the extension offirst order logic by thegeneralized quantifier “There are uncountably many” and theinfinitary language \(L_{\omega_1\omega}\). We shall now see that there is no hope of aCompleteness Theorem for second-order logic. Later in§9.1 we shall modify the semantics so that a Completeness Theorem can beproved.

Behind the failure of the Completeness Theorem is the fact thatsecond-order logic is—almost by definition—able to saythat a set is the power set of another set. What this means is thefollowing: LetL be a vocabulary. LetE be a binary andU a unary relation symbol, both not inL. Let\(\theta_{\textrm{Pow}}(U,E)\) be the second-order formula(\ref{ex6}). Let us consider the conjunction

\[\tag{8}\label{papow}\theta_{\textrm{PA}}(U,G,z)\land\theta_{\textrm{Pow}}(U,E).\]

Models of this formula are, up to isomorphism, of the form\((M,\in,N,g,a)\), where \(\P(N)=M\), \(N\cap\P(N)=\emptyset\) and\(N\) is countably infinite. Thus the models are necessarilyuncountable. This shows that second-order logic, unlike first orderlogic, has sentences with models but only uncountable ones. Thus theDownward Löwenheim-Skolem Theorem fails for second-order logic,as we already noted above. Combining (\ref{papow}) with theobservations in§5.1 yields the following result (by validity we mean the set ofGödel numbers of valid formulas):

Theorem 3 (Hintikka 1955; Montague 1963) Validity insecond-order logic is not second-order definable over\((\oN,+,\cdot,0)\).

In consequence, validity in second-order logic is not \(\Sigma^1_n\)for anyn. Since building power sets in the style of\(\theta_{\textrm{Pow}}(U,E)\) can be iterated, we can see thatvalidity in second-order logic is not \(\Sigma^m_n\) for anymandn. We extend this result further below inTheorem 6.

5.3 “Set theory in sheep’s clothing”

Second-order logic hides in its semantics some of the most difficultproblems of set theory. In the words of Resnik (1988: 75):

In Philosophy of Logic [Quine 1970], W. V. Quine summed up a popularopinion among mathematical logicians by referring to second-orderlogic as “set theory in sheep’s clothing”.

Let us see where this opinion might come from. First we observe that avery simple, indeed a very basic, second-order formula can say thattwo sets have the same cardinality: SupposeP andR areunary relation variables. Let \(\theta_{\le}(P,R)\) be the formula

\[\exists F\left(\forall x\,\forall y\left((F(x)=F(y)\to x=y)\land(P(x)\to R(F(x))\right)\right).\]

Now \(\mm\models_s\theta_\le(P,R)\) if and only if \(|s(P)|\le|s(R)|\). Let \(\theta_{\textrm{EQ}}(P,R)\) be the formula\(\theta_{{\le}}(P,R)\land \theta_{{\le}}(R,P)\). Now\(\mm\models_s\theta_{\textrm{EQ}}(P,R)\) if and only if\(|s(P)|=|s(R)|\). Let \(\theta'_{\textrm{EC}}(R)\) be

\[\exists F\forall x\,\forall y((F(x)=F(y)\to x=y)\land R(F(x))).\]

Now \(\mm\models_s\theta'_{\textrm{EC}}(R)\) if and only if the setsM and \(s(R)\) have the same cardinality. We will use theseformulas to launch an attack on the Continuum Hypothesis, thenotorious problem of set theory, proved independent of ZFC by Cohen in1963 (Cohen 1966).

Let \(\theta_{\textrm{CH}}\) be the sentence

\[\tag{9}\begin{align}&\exists E\,\exists U\,\exists G\,\exists z \\&\ \left(\theta_{\textrm{Pow}}(E,U)\land\theta_{\textrm{PA}}(U,G,z) \land \forall Y(\theta'_{\textrm{EC}}(Y)\lor\theta_{\le}(Y,U))\right).\end{align}\]

Now \(\theta_{\textrm{CH}}\), which is a sentence of the emptyvocabulary, has a model if and only if the Continuum Hypothesis CHholds. Similarly, there is a sentence \(\theta_{\neg \textrm{CH}}\),which has a model if and only if the Continuum Hypothesis CH does nothold. This shows that the dependence of the semantics of second-orderlogic on the metatheoretic set theory is so deep that even questionsthat ZFC cannot solve can determine the truth or falsity of a sentencein a model.

The curious quality of second-order logic is that the truth of asentence such as \(\theta_{\textrm{CH}}\) in a big enough model of theempty vocabulary is equivalent to the truth of the ContinuumHypothesis in the set-theoretical universe. By the same techniquealmost any set-theoretical statement can be turned into a sentence ofsecond-order logic, the truth of which in a big enough model of theempty vocabulary is equivalent to the truth of the statement in theset-theoretical universe. Somehow second-order logic manages to readthe background set theory correctly. Does this mean that second-orderlogic is set theory in “sheep’s clothing”?

5.4 Does second-order logic depend on the Axiom of Choice?

The Axiom of Choice (AC) is a philosophical puzzle in the foundationsof mathematics. Roughly speaking, this axiom, that originally emergedin set theory, says that if we have a set \(A\) of non-empty disjointsets, then we can form a new set \(B\) which contains exactly oneelement from each set in \(A\). What can be seen as a problem is thefact that when \(A\) is infinite, forming the set \(B\) seems torequire making an infinite number of choices. See entry onAxiom of Choice for a thorough definition and discussion. On the one hand the Axiomof Choice seems magic, for example when it implies that the wildestsets have a well-ordering. On the other hand it seems innocuous, forexample when it implies that the Cartesian product of a family ofnon-empty sets is non-empty. The Axiom of Choice is typically used toconstruct odd “paradoxical” sets, such as a non-Lebesguemeasurable set of real numbers, a well-ordering of the real numbers, aparadoxical decomposition of the sphere (see entry onset theory), and so on. Such sets are usually in one sense or another undefinable.For example, a non-Lebesgue measurable set of real numbershas to be undefinable in second-order logic over\((\oN,+,\cdot)\), if there are sufficiently large cardinals (see theentry onlarge cardinals and determinacy).

The puzzling Axiom of Choice arises also in the middle of second-orderlogic: Let \(\theta\) be the sentence

\[\forall X(\forall x\,\exists y\,X(x,y)\to\exists F\,\forall x\,X(x,F(x))).\]

Then \(\oR\models\theta\) essentially says that if a binary relation\(X\) relates to each \(x\in\oR\) a non-empty set \(\{y\in\oR :X(x,y)\}\), then there is a function \(F\) which chooses for each\(x\in\oR\) some \(y=F(x)\) such that \(X(x,y)\). In set theory thisis one of many equivalent ways of saying that the Axiom of Choiceholds for (continuum size) families of subsets of \(\oR\). The basictenet of second-order logic is that “all” properties ofelements of a fixed domain exist and we can quantify over them. Inset-theoretical terms this means that all subsets, definable or not,of a set of any cardinality exist and we can quantify over them. Inthis spirit the Axiom of Choice seems innocent and is usually acceptedin the literature of second-order logic.

By Cohen’s results (1966) there areN and \(N'\)satisfying the axioms of ZF (without the Axiom of Choice) such that“\(\oN\models\theta\)” holds inN but does not holdin \(N'\). This is a further demonstration of the dependence of thesemantics of second-order logic on the metatheory.

6. Non-Absoluteness of Truth in Second-Order Logic

Absoluteness is one of the most important concepts of logic. It wasused already in 1939 by Gödel in his analysis of the hierarchy ofconstructible sets (1939 [1990: 46]) towards proving the consistencyof the Continuum Hypothesis. It is briefly mentioned by Gödelthree years earlier in connection with the notion of computability(1936 [1986: 397]).

Intuitively speaking, a concept is absolute if its meaning isindependent of the formalism used, or in other words, if its meaningin the formal sense is the same as its meaning in the “realworld”. This may sound very vague and inexact. However,absoluteness has a perfectly exact technical definition in settheory.

Recall that a seta is calledtransitive if everyelement of an element ofa is an element ofa. Atransitive model is a model \((M,\in)\) such thatM istransitive. By the Mostowski Collapsing Lemma (Mostowski 1949) everywell-founded model \((M,E)\) of the Axiom of Extensionality isisomorphic to a unique transitive model. A basic property oftransitive models \((M,\in)\) is that if \(a,b\in M\), then \(M\modelsa\subseteq b\) if and only if \(a\subseteq b\). (If all elements ofa that are inM are inb then all elements ofa are inb, as by transitivity ofM, elements ofa are necessarily elements ofM.) We could express thisby saying that the formula \(x\subseteq y\) is absolute in transitivemodels. More generally, we say that a formula \(\phi(x_1,\ldots,x_n)\)of the first order language of set theory isabsoluterelative to ZFC if there is a finite \(T\subseteq \ZFC\) such that forall transitive models \((M,\in)\) ofT and all\(a_1,\ldots,a_n\in M\) we have

\[\phi(a_1,\ldots,a_n)\text{ if and only if }(M,\in)\models\phi(a_1,\ldots,a_n).\]

This definition of absoluteness is equivalent to the following moresyntactic condition (Feferman & Kreisel 1966): there are a\(\Sigma_1\)-formula \(\exists y\,\psi(y,x_1,\ldots,x_n)\) and a\(\Pi_1\)-formula \(\forall y\,\theta(y,x_1,\ldots,x_n)\) such that

\[\ZFC\vdash\forall x_1\ldots\forall x_n(\phi(x_1,\ldots,x_n)\leftrightarrow \exists y\,\psi(y,x_1,\ldots,x_n))\]

and

\[\ZFC\vdash\forall x_1\ldots\forall x_n(\phi(x_1,\ldots,x_n)\leftrightarrow \forall y\,\theta(y,x_1,\ldots,x_n)).\]

In such a case we say that \(\phi(x_1,\ldots,x_n)\) is\(\Delta_1^{\text{ZFC}}\).

For first order logic it can be proved that“\(\mm\models_s\phi\)” is an absolute property of \(\mm\),s and \(\phi\) relative to ZFC. In fact, this is true even ifZFC is replaced by the weaker Kripke-Platek set theory KP (see Barwise1972a). The usual Tarski Truth Definition of first order logic can bewritten in \(\Delta_1^{\text{ZFC}}\) form. For second-order logic thepropositions “\(\phi\) is a second-order formula”,“\(\mm\) is anL-structure” and “s isan assignment” are all absolute relative to ZFC as well, but“\(\mm\models_s\phi\)” is not, as we shall now see. Thisis a crucial property of second-order logic. One may consider it aweakness, if one thinks of absoluteness as a desirable property, or astrength, if one takes non-absoluteness as a sign of expressive power.Whether a weakness or not, it is an important feature of second-orderlogic and a dominating topic in discussions about it. Second-orderlogic is, however, not the only non-absolute logic. For example,\(L(Q_1)\) (see the entry ongeneralized quantifiers) and \(L_{\omega_1\omega_1}\) (see then entry oninfinitary languages) are non-absolute, while \(L(Q_0)\) and \(L_{\omega_1\omega}\) areabsolute. The reason why \(L(Q_0)\) is absolute is essentially thatthe property of a set being infinite is absolute in transitive modelsof ZFC. Although the set of sentences of \(L_{\omega_1\omega}\) mayvary from one transitive model of ZFC to another, the truth of asentence of \(L_{\omega_1\omega}\) in a model is absolute intransitive models of ZFC essentially because the truth has a simpleinductive definition which only refers to subformulas of the sentenceand elements of the model.

Recall the definition of the sentence \(\theta_{\textrm{EC}}(P,R)\) in§5.3, which says that the interpretations of the predicatesP andR have the same cardinality. Cardinality is not an absoluteproperty. Two sets (even two ordinals) can be of different cardinalityin one transitive model of ZFC and of the same cardinality in atransitive extension. It is easy to see that“\(\mm\models_s\theta_{\textrm{EC}}(P,R)\)”is not absolute relative to ZFC. Essentially, the reason is that twosets have the same cardinality if there is a bijection between them.In a small transitive set it may happen that the sets are there butthe bijection manifesting their equicardinality is not. The bijectionmay be one obtained by the method of forcing. Or the lack of thebijection may be the result of applying the DownwardLöwenheim-Skolem Theorem[6].

A more serious case of non-absoluteness is the sentence\(\theta_{\textrm{CH}}\) of§5.3. The sentence \(\theta_{\textrm{CH}}\) of the empty vocabulary has amodel if and only if the Continuum Hypothesis is true. If \(T\subseteq\ZFC\) is finite, then there are countable transitive models\(M\subseteq M'\) such that one, sayM, satisfies CH and theother, in this case \(M'\), does not (by Cohen 1966). InM thesentence \(\theta_{\textrm{CH}}\) has a model \(\ma\), that is,\(M\models\textrm{“}\ma\models\theta_{\textrm{CH}}\textrm{”}\). In\(M'\) the same sentence has no models, in particular \(M'\not\models\textrm{“}\ma\models\theta_{\textrm{CH}}\textrm{”}\). Thesituation is similar if the Continuum Hypothesis is replaced by theAxiom of Choice (see§5.4). Thus the property of a model satisfying a second-order sentence isnot absolute relative to ZFC. Even the property of a second-ordersentencehaving a model is not absolute relative to ZFC,because inM the sentence \(\theta_{\textrm{CH}}\) has a modelbut in \(M'\) not. For first order logic the property of a sentence ofhaving a model is co-r.e. and therefore arithmetical. Arithmeticalproperties are always absolute relative to ZFC.

The CH and the AC are central questions of set theory and subjects ofcontinuing debate. While the AC is mostly accepted as an axiom, and ispart of the ZFC axiom system, the CH is widely considered an openproblem. There are even suggestions (e.g., Feferman 1999) that itcannot be solved by any axiom system with the kind of generalacceptance that ZFC has. The formalist position in the foundations ofmathematics goes further and maintains that the question, what thetruth value of CH is, is meaningless. Both CH and its negation areconsistent with ZFC, assuming ZFC itself is consistent. According tothe formalist position the status of CH has been solved with this. Thesemantics of second-order logic depending on these hard questions ofset theory puts second-order logic into the same “basket”with set theory. Criticism of set theory becomes criticism ofsecond-order logic and vice versa.

Another indication of the dependence of second-order logic on settheory as the metatheory is the result of Barwise to the effect thatthe existence of the Hanf number of second-order logic is not provablein set theory without a highly complex use of the Replacement Axiom(for details, see Barwise 1972b).

7. Model Theory of Second-Order Logic

Let us start by noting that there are some trivial consequences ofsecond-order logic being able to quantify over relations. For example,the Craig Interpolation Theorem holds for second-order logic fortrivial reasons: If the finite relational vocabulary of \(\phi\) isL, the finite relational vocabulary of \(\phi'\) is \(L'\) and\(\models\phi\to\phi',\) then there is \(\theta\) such that\(\models\phi\to\theta\text{ and } \models\theta\to\phi'\) and thevocabulary of \(\theta\) is \(L\cap L'\), because we can let\(\theta\) be the sentence \(\exists X_1\ldots\exists X_n\,\phi\),where \(L\setminus L'=\{X_1,\ldots, X_n\}\). We can additionallydemand that every predicate symbol occurring positively in \(\theta\)occurs positively in \(\phi\) and \(\psi\), and every predicate symboloccurring negatively in \(\theta\) occurs negatively in \(\phi\) and\(\psi\) (theLyndon Interpolation Theorem [Lyndon 1959]). Inthe same way the Beth Definability Theorem holds for trivial reasons,also some preservation results[7] have a trivial proof. For a negative result, see Craig 1965.

However, second-order logic does not have a model theory in the samesense as first order logic does. There is no Compactness Theorem toproduce non-standard models. One reason why first order logic has sucha rich model theory is that first order logic is relatively weak.Countable first order theories that have infinite models have modelsof all infinite cardinalities. This is a consequence of a weakness offirst order logic: its sentences cannot limit the size of the modelunless the size is finite. In model theory this weakness is turned toa strength: the structure of models of first order theories can beanalyzed in interesting ways. See the entry onmodel theory for examples of this.

Second-order theories can typically have just one model up toisomorphism. Thus if we investigate the class of models of asecond-order sentence, the class may very well consist of just onemodel, up to isomorphism. With just one model one cannot reallydevelop a model theory. One ends up investigating the one and onlymodel and there are no compelling reasons to use logic in the study ofit. If the one and only model is an algebraic structure, one shoulduse algebra to study it. If it is an ordered structure, one should usethe general theory of ordered structures to study it, and so on. It isunlikely that being second-order characterizable (see§7.1) tells much about the structure. It reveals more about second-orderlogic in general. For example, being able to characterize thestructure of natural numbers means that the Compactness Theorem andthe Completeness Theorem fail. Being able to characterize the orderedfield of real numbers means that both the Downward and the UpwardLöwenheim-Skolem Theorems fail. It is from thisperspective—to understand second-order logic itself—thatit is interesting to know exactly which structures are second-ordercharacterizable up to isomorphism.

Let us finally compare the two hierarchies, the model theoretichierarchy \(\Sigma_n^1\cup\Pi_n^1\) inside second-order logic and theLevy hierarchy \(\Sigma_n\cup\Pi_n\) in set theory.

Theorem 4

  1. The set of Gödel numbers of valid second-order sentences isthe complete \(\Pi_2\)-set of natural numbers. (Tharp 1973)
  2. The Löwenheim-Skolem number of second-order logic is thesupremum of all \(\Pi_2\)-definable ordinals. (Krawczyk & Marek1977)
  3. The Hanf number of second-order logic is the supremum of all\(\Sigma_2\)-definable ordinals. (Krawczyk & Marek 1977)
  4. Every model class which is definable in second-order logic is\(\Delta_2\)-definable in set theory. (Väänänen1979)

In the light of the above theorem second-order logic sits firmly onthe \(\Sigma_2\cup\Pi_2\)-level of set theoretic definability. This isa reason to consider second-order logicweaker than firstorder set theory. This sounds paradoxical. How can second-order logic,with all the power and non-first order properties it has, be weakerthan first order logic based set theory, when first order logic cannoteven express finiteness, countability, uncountability, and so on?Nevertheless the message ofTheorem 4 is undeniable. Perhaps this would not be so puzzling if we thought ofset theory as a very high order logic over the singleton of the emptyset. After all, set theory permits endless iterations of the power setoperation, while second-order logic permits only one iteration.

We start below with a subsection on models that can be characterizedin second-order logic by one sentence up to isomorphism. Anotherapproach to the model theory of second-order logic is to focus onmodels withlarge cardinal (inaccessible, measurable,supercompact cardinals) properties. This is the topic of§7.2. Finally, in§7.3 we consider the possibility of allowing the so-called general modelsintroduced in§9.1. In this case second-order logic becomes very much like first orderlogic. However, general models are mostly interesting when they areso-called Henkin models, i.e., they satisfy the so-calledComprehension Axioms. In Henkin models second-order logic satisfiesthe Compactness Theorem and other principles familiar from first ordermodel theory but no general structure- or classification theory existsyet, perhaps because the Comprehension Axioms bring in a lot ofcomplicated structure into the models.

7.1 Second-order characterizable structures

A structure \(\ma\) issecond-order characterizable if thereis a second-order sentence \(\theta_\ma\) such that\(\mb\models\theta_\ma\iff\mb\cong\ma\) for all structures \(\mb\) ofthe same vocabulary as \(\ma\).

Example 5: The following structures are second-ordercharacterizable:

  1. Natural numbers: \((\oN,+,\cdot)\).
  2. Real numbers: \((\oR,+,\cdot,0,1)\).
  3. Complex numbers: \((\oC,+,\cdot,0,1)\).
  4. The first uncountable ordinal \((\omega_1,<)\).
  5. The level \((V_\kappa,\in)\) of the cumulative hierarchy, where\(\kappa\) is the first strongly inaccessible cardinal\(>\omega\).
  6. The well-order \((\kappa,<)\) of the first weakly compactcardinal \(>\omega\).

Areall structures second-order characterizable? There areonly countably many second-order sentences, hence only countably many(up to isomorphism) second-order characterizable structures. Thereforethere are lots of structures of every infinite cardinality which arenot second-order characterizable. However, it is not easy to giveexamples. One example is \((\kappa,<)\), where \(\kappa\) is thefirst measurable cardinal (\(>\omega\)). See§7.2 for an explanation. Another example is \((\oN,<,A)\), whereA is the set of Gödel numbers of valid second-ordersentences in the vocabulary of one binary relation. For anexplanation, see§7.2. One may ask whether there is an example which arises frommathematical practice outside of mathematical logic. What about\((\oR,+,\times,<,A)\), whereA is a Hamel basis (a basisfor the vector space where the reals are the vectors, rationals arethe scalars and vector addition is the usual addition of reals)? It isconsistent, relative to the consistency of ZF that no such structure\((\oR,+,\times,<,A)\) is second-order characterizable (Hyttinen,Kangas, & Väänänen 2013).

A special property of second-order characterizable structures is thattheir reducts are also second-order characterizable, because we canuse existential second-order quantifiers to “guess” themissing relations and functions. Therefore it is interesting to findcharacterizable structures that have as many (but only finitely many)relations and functions as possible. We can endow \(\oN\) with anyfinite number of recursive functions \(f_1,\ldots, f_n \) andrelations \(R_1,\ldots, R_m\) obtaining the structure\((\oN,f_1,\ldots, f_n,R_1,\ldots, R_m)\) and this is second-ordercharacterizable. We can endow \(\oR\) with any familiar analyticfunctions such as trigonometric functions or any other functions givenby a convergent power series the coefficients of which are given by arecursive function, and the result is second-ordercharacterizable.

It is easy to see that the second-order theory of a second-ordercharacterizable structure \(\ma\) is Turing-reducible to thesecond-order theory of any bigger second-order characterizablestructure \(\mb\), as for all second-order \(\phi\):

\[\ma\models\phi\iff\mb\models\exists P(\theta_\ma^{P}\land\phi^{P}).\]

The bigger the characterizable structure is, the more complex is thesecond-order theory. That there is no largest characterizablestructure can be seen as follows: If \(\ma\) is second-ordercharacterizable, then so is the reduct of \(\ma\) to the empty vocabulary[8], that is, the cardinality \(|A|\) of \(\ma\) is characterizable. Suchcardinal numbers were studied by Garland (1974). For example, if\(\kappa\) is characterizable, then so are \(\kappa^+\) and\(2^\kappa\).

If \(\phi\) is a second-order sentence, we define

\[\Mod(\phi)=\{\mm: \mm\models\phi\}.\]

If \(\phi\) characterizes a model \(\ma\), then \(\Mod(\phi)\) is justthe class of models isomorphic to \(\ma\). If we look at\(\Mod(\phi)\) from the perspective of set theory, we know it is\(\Delta_2\) (Theorem 4). On the other hand, the set of Gödel numbers of validsecond-order sentences is \(\Pi_2\)-complete (Theorem 4), hence not \(\Sigma_2\) and in particular, not \(\Delta_2\). Thiscontemplation leads us to the following extension ofTheorem 3 above:

Theorem 6 (Väänänen 2012) Second-ordervalidity is not second-order definable on any second-ordercharacterizable structure.

This raises the question, how can we understand second-order validityfrom the second-order perspective? In view of the above theorem wecannot understand second-order validity as something that can beexpressed in second-order logic on some structure which itself can becharacterized in second-order logic. The reference to set theory seemsinevitable in understanding second-order validity. Let us discuss whatthis might mean from the point of view of structuralism. Parsonsdefines structuralism as follows:

By the “structuralist view” of mathematical objects, Imean the view that reference to mathematical objects is always in thecontext of some background structure, and that the objects involvedhave no more to them than can be expressed in terms of the basicrelations of the structure. (Parsons 1990: 303)

Structuralism in this sense (for other kinds of structuralism, see,e.g., Hellman 2001) seems to fit well into the second-order logicframework. Contrary to set theory, second-order logic always assumesan underlying (limited) structure which is supposed to reflect oneparticular aspect of mathematics such as the semi-ring of naturalnumbers, the ordered field of real numbers, the field of complexnumbers, the well-ordered structure \((\omega_1,<)\), and so on.Ideally, these underlying structures are second-order characterizable.This leads to a problem with universal truths such as validity, butalso with less complex ones such as the statement that every linearorder has a completion, or that every field has an algebraic closure.Such universal truths are not about any particular structure.

7.2 Second-order logic and large cardinals

We refer to the SEP entriesindependence and large cardinals andlarge cardinals and determinacy for the definition of concepts related to large cardinals.

The following early result on the connection between model theory ofsecond-order logic and large cardinals is a model for other moredelicate results. It is a kind of Downward Löwenheim-SkolemTheorem for second-order logic. It says that if a second-ordersentence has a model of the cardinality of a measurable cardinal, thenit has a smaller submodel:

Theorem 7 (Scott 1961) Suppose \(\kappa\) is ameasurable cardinal. If \(\phi\) is a second-order sentence and\(\phi\) has a model \(\mm\) of cardinality \(\kappa\), then for every\(X\subseteq M\) of cardinality \(<\kappa\) there is\(\mn\subseteq\mm\) such that \(X\subseteq N\), \(|N| < \kappa\),and \(\mn\models\phi\).[9]

A particular consequence is that the first measurable cardinal cannotbe second-order characterizable as a model of the empty vocabulary (afact that was referred to above in§7.1). Theorem 7 resembles remarkably the Downward Löwenheim-SkolemTheorem of first order logic. The only difference is that one has tostart with a big model, a model of the size of a measurable cardinal.Smaller cardinals need not work. For example, if \(\lambda\) is theleast weakly compact cardinal \(>\omega\), then there is a sentence\(\phi\) which has \(\lambda\) (with the empty vocabulary) as a modelbut no smaller models. The sentence[10] \(\phi\) says that \(\lambda\) is inaccessible (\(>\omega\)) andthat every \(\lambda\)-tree (a tree of height \(\lambda\) with alllevels of size \(<\lambda\)) has a branch of length\(\lambda\).

If we want to obtain a Downward Löwenheim-Skolem Theorem whichworks for any structure, not only for structures of a measurablecardinality, we have to look at even bigger cardinals: TheLST-number of second-order logic is defined to be the least(if any exist) \(\kappa\) such that if \(\ma\) is any structure and\(\phi\) is a second-order sentence true in \(\ma\), then there is asubstructure \(\mb\) of \(\ma\) of cardinality \(<\kappa\) whichalso satisfies \(\phi.\)

Theorem 8 (Magidor 1971) The LST-number ofsecond-order logic exists if and only if there are supercompactcardinals \(>\omega\) and then it is the smallest of them.

Second-order logic is said to satisfy strong\(\kappa\)-compactness if every second-order theory, everysubset of size \(<\kappa\) of which has a model, has itself amodel.

Theorem 9 (Magidor 1971) The least \(\kappa\) suchthat second-order logic satisfies the strong\(\kappa\)-compactnessis the least extendiblecardinal.

7.3 The model theory of general and Henkin models

The situation changes completely if we adopt so-called general models (§9.1). With general models second-order logic has similar model theoreticproperties as first order logic, as it can simply be thought of asmany sorted first order logic (see§9.1 and Manzano 1996). By and large, results of many-sorted first orderlogic translate immediately to second-order logic.

8. Decidability Results

The second-order theory of even the empty vocabulary is undecidable,as we noted in§5.1. In contrast,monadic second-order logic gives rise to manyimportant decidability results. Let us first observe that (not only inthe empty vocabulary but also) in a monadic vocabulary, i.e., oneconsisting of monadic predicates only, monadic second-order logic iscertainly decidable as proved already in 1915 by Löwenheim (1915)in one of the earliest papers on the mathematical aspects of formallogic. Löwenheim’s result can be proved easily byquantifier elimination, or alternatively by the more recent method ofEhrenfeucht-Fraïssé-games (see§3). However, monadicthird (and higher) order logic in monadicvocabulary, with quantifiers over not only subsets and also over setsof subsets of the domain, is undecidable (Tharp 1973). We now focus onvocabularies which are not merely monadic.

Compared to a monadic vocabulary, the other extreme can be describedas follows: LetP be a ternary predicate (a “pairingfunction”) on a non-empty setS. Suppose that for every\(x,y\in S\) there is \(z\in S\) with \((x, y, z) \in P\) and forevery \(z\in S\) there is at most one pair \((x, y)\) with \((x, y, z)\in P\). Then the true (full) second-order theory ofS isinterpretable in the monadic theory of \((S, P)\) (Gurevich 1985). Inother words, in a context where a pairing function is present, as inset theory and number theory, no special advantage arises fromlimiting second-order logic to its monadic fragment.

The next step from a vocabulary consisting of only monadic predicatesis a vocabulary with exactly one unary function. To describe somehighly non-trivial results in this case we introduce the followingconcept, which has independent interest: Thespectrum of afirst or second-order sentence is the set of cardinalities of itsfinite models. Spectra were introduced by Scholz, see Durand, Jones,Makowsky, and More (2012) for a history. A spectrum is of coursealways recursive. In fact, a set of natural numbers is a spectrum of afirst order sentence iff it is in NEXPTIME, that is, recognizable by anondeterministic Turing machine in exponential time (Jones &Selman 1974). In the case of sentences with just one unary functionsymbol there is a surprising characterization. Let us call a setS of natural numberseventually periodic if there arenatural numbersN andp with \(p > 0\) such that foreveryn with \(n > N\), we have \(n \in S\) iff \(n +p \inS\).

Theorem 10 (Durand, Fagin, & Loescher 1998; Shelah2004) Suppose the vocabulary has just one unary function.Then a set of natural numbers is the spectrum of a monadicsecond-order sentence if and only if it is the spectrum of a firstorder sentence if and only if it is eventually periodic.

In graphs monadic second-order logic can still express mostinteresting properties. For example, theconnectivity of thegraph (hereE is the edge relation) can be expressed by thesentence

\[\forall X((\exists x\,X(x)\land\forall x\forall y((X(x)\land xEy)\toX(y)))\to\forall x\,X(x)) \]

and the3-colorability of the graph by the sentence

\[\begin{align}&\exists X\,\exists Y\,\exists Z ( \forall x(X(x)\lor Y(x)\lor Z(x))\land{}\\&\quad\forall x\,\forall y (((X(x)\land X(y))\lor(Y(x)\land Y(y))\lor(Z(x)\land Z(y)))\to{}\\&\quad\neg xEy))\end{align}\]

There is a wealth of decidability results of monadic second-orderlogic on linear orders, trees and graphs.

Theorem 11 (Büchi 1962; Elgot 1961; Rabin 1968)The monadic second-order theory of \((\oN,s)\), i.e., the successorfunction on \(\oN\) and of the full infinite binary tree, i.e., twosuccessor functions on \(\oN\), are both decidable.

The proof goes via a reduction to automata. Anautomaton is arestricted form of a Turing machine. In an automaton the machine neverwrites on the tape, so the tape just contains the input (Rabin &Scott 1959). However, the computation may be infinite. The input isconsideredaccepted in the case of an infinite computation ifan accepting state is reached infinitely many times. (There are alsomore complicated criteria of acceptance). With monadic second-orderlogic we can guess an accepting computation of the automaton. It ispossible but more difficult to show that an automaton can check thetruth of a monadic second-order sentence. The monadic second-ordertheory of ordinals \((\alpha,<)\) have been studied extensivelyfrom the point of view of decidability. For example, the monadicsecond-order theory of all countable ordinals, the theory of\(\omega_1\), and the theory of all ordinals \(<\omega_2\) aredecidable (Büchi 1973), but ZFC cannot decide whether the monadicsecond-order theory of \(\omega_2\) is decidable (Gurevich, Magidor,& Shelah 1983).

9. Axioms of Second-Order Logic

Recall our convention: Whenever we use upper case Fraktur letters,such as \(\ma,\mb,\mm,\mn\) etc, to name a structure, we use the uppercase Italic versions of those letters, such as \(A,B,M,N\) etc, toname the domain of the structure.

The deductive system of second-order logic, presented first explicitlyin Hilbert-Ackermann (Hilbert & Ackermann 1938), is based on theobvious extension of axioms and rules of first order logic togetherwith theComprehension Axiom Schema, defined as follows:Suppose \(\phi(x_1,\ldots,x_n)\) is a second-order formula with\(x_1,\ldots, x_n\) among its free individual variables and thesecond-order variableR isnot free in \(\phi\). Thenthe following formula is an instance of the Comprehension AxiomSchema:

\[\tag{10}\label{CA} \exists R\,\forall x_1\ldots x_n(\phi(x_1,\ldots,x_n)\leftrightarrow R(x_1,\ldots, x_n)). \]

There is a similar schema for second-order formulas that define afunction but we omit it here.

An example of a very simple second-order inference in number theorywould be

\[\frac{\dfrac{2+5<9}{\exists R\ R(2+5,9)}}{\exists F\,\exists R\ R(F(2,5),9)}\]

A priori (\ref{CA}) might appear very strong as it stipulates theexistence of a relation. However, such anR seems reasonable toaccept because it is definable:

\[R=\{(a_1\ldots a_n) \in M^n: \mm\models \phi(a_1,\ldots, a_n)\}.\]

Our Comprehension Axiom Schema isimpredicative in the sensethat the boundn-ary relation variables which may potentiallyoccur in the formula \(\phi(a_1,\ldots, a_n)\) haveR in theirrange. In this sense second-order logic is utterly impredicative. TheComprehension Axiom Schema has weaker forms that are lessimpredicative. In theArithmetic Comprehension Schema weassume \(\phi(a_1,\ldots, a_n)\) is first order. In the\(\Pi^1_1\)-Comprehension Schema we assume \(\phi(a_1,\ldots, a_n)\)is \(\Pi^1_1\). In mathematical practice these two weaker principlesusually suffice, see Simpson (1999 [2009]) and§14.

Hilbert-Ackermann (1938) added to the proof system of second-orderlogic also two different schemas that they callAxioms ofChoice. We are used to thinking of the Axiom of Choice as aset-theoretical principle. But second-order logic has the samesituation with choice principles as set theory. It makes perfect sensein second-order logic to ask whether there is a well-order of thedomain. The existence of such a well-order does in general not followfrom the Comprehension Schema because the well-order may be---and islikely to be---undefinable. So the only way to get a well-order of,say, the real numbers, in second-order logic is to assume some form ofsecond-order Axiom of Choice. The first choice principle is

\[\tag{AC}\begin{align}&\forall x_1\ldots x_m\exists x_{m+1}\, R(x_1,\ldots, x_{m+1}) \rightarrow {}\\ &\ \exists F\,\forallx_1\ldots\forall x_m\, R(x_1,\ldots, x_m, F(x_1\ldots x_m))\end{align}\]

and the second is

\[\tag{AC$'$} \forall x_1\ldots x_m\exists F\,\varphi\rightarrow\existsF'\,\forall x_1\ldots x_m\varphi', \]

where the formula \(\varphi'\) is obtained from the formula\(\varphi\) by replacing everywhere \(F(t_1,\ldots, t_k)\) by\(F'(t_1\ldots t_k,x_1,\ldots, x_m)\).

The first Axiom of Choice (AC) says intuitively that if a set

\[\{a_{n+1}\in M : \mm\models R(a_1,\ldots,a_n, a_{n+1})\}\]

is non-empty, then there is a function which picks an element\(a_{n+1}\) from the set, using the parameters \(a_1,\ldots,a_{n}\) asarguments. The second Axiom of Choice (AC′) is a kind ofsecond-order choice: We have for all \(a_1,\ldots,a_n\) a functionF with the property \(\phi\), so in factF depends on\(a_1,\ldots, a_n\) and should be denoted \(F_{a_1\ldots a_n}\). What(AC′) says now is that we can collect the functions\(F_{a_1\ldots a_n}\) together to form just one function \(F'\) ofhigher arity such that

\[F'(a_1,\ldots,a_{n+1})=F_{a_1\ldots a_n}(a_{n+1}).\]

Although \(F'\) appears to be definable, this is not the case, as themapping

\[(a_1,\ldots,a_n)\mapsto F_{a_1\ldots a_n}\]

may be highly undefinable.

Naturally, we may also want to assume theWell-orderingPrinciple which states that there is a binary relation which is atotal order on the individuals and satisfies the condition that everynon-empty set of individuals has a least element in the order. It iseasy to see that this principle implies (AC′). In ZFC there arehundreds of equivalent formulations of the Axiom of Choice or weakerversions of it. Most of the formulations that are equivalent in ZFCare non-equivalent in second-order logic. This is because set theoryallows free movement between sets and their power sets, butsecond-order logic does not[11]. To obtain similar equivalence proofs as in set theory, one would haveto move between second-order logic and third order logic, seeGaßner (1994) and Siskind, Mancosu & Shapiro (2023).

9.1 General models and Henkin models

The concept of “general model” was introduced by Henkin(1950) in order to obtain a Completeness Theorem for second-orderlogic. General models are not the intended models of the axioms in thesense that bound relation and function variables do not range in thesemodels overall relations and functions, only over acollection of them. The generally accepted attitude towards them isthat they are not what one wants but they are necessary for obtaininga smooth theory. They are like irrational numbers in calculus: itwould be ideal if all lengths that we encounter in calculus (orgeometry) were commensurable, but that is not how things turned out.Some lengths are incommensurable, and we end up accepting irrationalnumbers in order to have a smooth general theory of lengths and theirproportions. General models can be also compared to transitive modelsof set theory. One transitive model may say one thing is true, anothersays just the opposite is true. But together the transitive modelshelp us understand the axioms of set theory. The current methods inthe metamathematical investigation of the axioms of set theory, namelyinner models and forcing extensions, all use transitive models of settheory in an essential way. Similarly general models of second-orderlogic help us investigate metamathematical properties of second-orderlogic and its axioms.

Definition 12 Ageneral model is a pair\((\mm,\G)\), where \(\mm\) is a usualL-structure and \(\G\)is a set of subsets, relations (of any arity) and functions (of anyarity) onM. In monadic second-order logic \(\G\) is assumed tocontain unary predicates only.

We can define the concept

\[(\mm,\G)\models_s\phi\]

for second-order \(\phi\) by induction on \(\phi\) by stipulating

\[\begin{align}(\mm,\G)\models_s\exists X\,\phi & \iff (\mm,\G)\models_{s(P/X)} \phi\text{ for some $P\in \G$}.\\(\mm,\G)\models_s\exists F\,\phi &\iff (\mm,\G)\models_{s(f/X)} \phi\text{ for some $f\in \G$}.\end{align}\]

The difference to the original truth definition of second-order logicis thus that the interpretations of the second-order variables cannotbe quite arbitrary, they have to be found in \(\G\). The smaller the\(\G\) is, the weaker the logic. On the other hand, if \(\G^\star\) isthe set ofall relations and functions onM, then thetruth definition for the general model \((\mm,\G^\star)\) is the sameas the original truth definition. We call such general modelsfull and indeed identify \((\mm,\G^\star)\) with \(\mm\). Theother extreme is that \(\G=\emptyset\). Then we are back in firstorder logic. To distinguish second-order logic with general modelsfrom the original second-order logic, we call the latterfull.

Definition 13 A general model which satisfies all theaxioms of second-order logic, including the Comprehension Axiom Schemaand the Axioms of Choice, is called a Henkin model.

Note that there isno sentence of second-order logic whichsays that the model is full. That is, there is no second-order\(\phi\) such that for all general models \((\mm,\G)\) it holds that

\[(\mm,\G)\models\phi\iff (\mm,\G) \text{ is full.}\]

If we attempt to use \(\forall X\,\exists Y\,\forallx(X(x)\leftrightarrow Y(x))\) as such a \(\phi\), we just obtain theresult that \(\phi\) is always true and says that the general model\((\mm,\G)\) is “full” in itsown sense of“full”, not in the sense of the metatheory.

A natural attempt to get a Henkin model \((\mm,\G)\) would be to let\(\G\) consist offirst order definable relations andfunctions on \(\mm\). However, in this way we get only ArithmeticComprehension, not the full Comprehension Axiom.

An easy way to obtain a Henkin model \((\mm,\G)\) is to take atransitive modelN of ZFC such that \(\mm\in N\) and let \(\G\)consist of all relations and functions onM that are elementsofN. It follows from \(N\models ZFC\) that \((\mm,\G)\) is aHenkin model.

Truth in a general model, namely the proposition“\((\mm,\G)\models_s\phi\)” is absolute relative to ZFC,contrary to the truth definition of (full) second-order logic. In thisrespect second-order logic with general models resembles first orderlogic. As we will see, it resembles first order logic in many otherrespects, too.

Second-order logic with general models can be thought of in terms ofmany sorted logic. Many sorted logic is first order logicwith several differentsorts of variables. Respectively,relations and functions may be between elements of different sorts.Here “sort” just means that variables of different“sorts” are kept apart by giving them different names.Otherwise the variables are just ordinary first order variables.Structures \(\mm\) of many-sorted logic have a separate domain \(M_s\)for each sorts. We may think of second-order logic as manysorted logic in which the individual variables are of a sort \(\si\),for eachn then-ary relation variables are of a sort\(\sr\), and then-ary function variables of a sort \(s_n^{\rmfu}\). In addition there is an \(n+1\)-ary relation symbol \(\sA\)betweenn-tuples of individuals andn-ary relationvariables indicating whichn-tuples are in a relation and whichare not. Finally there is an \(n+1\)-ary function symbol \(\sF\)betweenn-tuples of individual, individuals, andn-aryfunction variables indicating what value a function variable gives toeachn-tuple. There are obvious first order axioms whichguarantee that the many sorted logic just described faithfully mirrorssecond-order logic with general models.[12]

For more on the reduction of second-order logic to many-sorted logic,we refer to Manzano (1996). For example, the proof theory ofsecond-order logic can be understood via the translation to manysorted logic. For more on the proof theory of second-order logic, seeBuss (1998). There is a translation of many sorted logic further tosingle sorted first order logic due essentially to Herbrand (1930),see also Wang (1952) and Schmidt (1951). This can be used to obtainmany of the basic properties of first order logic first for manysorted logic and then further for second-order logic with generalmodels.

The most important application of general models is the CompletenessTheorem:

Theorem 14 (Henkin 1950) A second-order sentencefollows from the axioms if and only if it is true in all Henkinmodels.[13]

Respectively, we obtain the Compactness Theorem[14] as well as the downward and the upward versions of theLöwenheim-Skolem Theorem.[15]

One way to think about Henkin models is that they fill the gaps leftby full models, just as irrational numbers fill the gaps left byrational numbers. They are needed in order to obtain a smooth theoryof second-order logic. They manifest “paradoxical”phenomena that we do not see among full models. For example we obtainnon-standard models of number theory, countable models of the axiomsof real numbers, etc. The categorical sentences characterizingmathematical structures among full models now suddenly have also other“models”, namely Henkin models, and they come in allinfinite cardinalities.

If \(\phi\) is a second-order sentence we define

\[\tag{11}\label{henkin}\textit{Mod}_H(\phi)=\{(\mm,\G) : (\mm,\G)\models\phi\} \]

Obviously, \(\Mod(\phi)\subseteq \textit{Mod}_H(\phi)\). If \(\phi\)characterizes \(\mm\) up to isomorphism, then \(\mm\in \Mod(\phi) \).In all non-trivial cases[16] \(\textit{Mod}_H(\phi)\ne \{\mm\}\). For \(\mm\in\Mod(\phi)\) we canthink of \((\mm,\G)\in\textit{Mod}_H(\phi)\) as“approximations” of \(\mm\). One of the approximations isthe “real” \(\mm\), i.e. \((\mm,\G^*)\), but by means ofdeductions in second-order logic we cannot tell which. Because of theinherent weakness of formal systems, going back to Skolem andGödel, infinite structures are shrouded by Henkin models andcannot be gotten perfectly into focus by deductive means.

TheFraenkel-Mostowski method is a construction of Henkinmodels which was first used in set theory to obtain models where Axiomof Choice fails, and later used in second-order logic to obtainsimilar models before the method of forcing emerged. The method goesas follows: Suppose \(\ma\) is a structure andH is a group ofautomorphisms of \(\ma\). A subsetB of \(A^n\) is calledsymmetric if there is a finite set \(E\subseteq A\), called asupport ofB, such that for all \(\pi\in H\) which fixE pointwise[17],

\[\forall x_1\ldots\forall x_n((x_1,\ldots, x_n)\in B\to (\pi(x_1),\ldots, \pi(x_n))\in B).\]

The respective condition for a function \(f:A^n\to A\) is

\[\begin{align}&\forall x_1\ldots\forall x_n ((x_1,\ldots ,x_n) \in B\to f(\pi(x_1), \ldots ,\pi(x_n)) \in B \land {}\\&\qquad f(\pi(x_1),\ldots ,\pi(x_n)) = \pi (f (x_1,\ldots ,x_n)))\end{align}\]

Let \(\G\) be the set of symmetric relations and functions onA. The pair \((\ma,\G)\) is always a Henkin model.

Among the early applications of the Fraenkel-Mostowski method were theresult of Mostowski (1938) that one cannot prove from theComprehension Axiom Schema alone the equivalence of certain twodifferent definitions of finiteness and of Lindenbaum and Mostowski(1938) that Zermelo’s axioms for set theory are not enough toprove the Axiom of Choice, or even just that every infinite set is thedisjoint union of two infinite sets.

TheFraenkel model arises ifH is the group of allpermutations of an infinite setA. In the Fraenkel model thesetA is infinite but Dedekind-finite (i.e.,A has noone-one function into a proper subset) and cannot be well-ordered, noteven linearly ordered. Moreover, the setA is not the disjointunion of two infinite sets. There is no choice function for twoelement subsets ofA.

TheMostowski model arises if \(A=\oQ\) andH is thegroup of automorphisms of \((\oQ,<)\). In the Mostowski model thesetA can be linearly ordered but not well-ordered. The Axiomof Choice holds for (sets of) non-empty finite sets.

For more on Fraenkel-Mostowski models as Henkin models forsecond-order logic, see Gaßner (1994).

9.2 Axioms of infinity

Recall our convention: Whenever we use upper case Fraktur letters,such as \(\ma,\mb,\mm,\mn\) etc, to name a structure, we use the uppercase Italic versions of those letters, such as \(A,B,M,N\) etc, toname the domain of the structure.

The axioms of second-order logic are trivially true in a (full) modelwith just one element. This raises the question how to guarantee thatthere are more than one element in the domain. Of course, we can add,for everyn, an axiom which says that there are more thann elements. There are various ways to say in second-order logicwithone axiom that there areinfinitely manyindividuals. These are equivalent in full models but may benon-equivalent in Henkin models. Some are equivalent if the Axiom ofChoice is assumed. Let us call a second-order sentence \(\phi\) of theempty vocabulary anAxiom of Infinity if

\[\ma\models\phi\text{ if and only if $A$ is infinite}.\]

An axiom of infinity can say in second-order logic that a propersubset of the domain has the same cardinality as the entire domain(i.e., that the domain is not Dedekind-finite), or that there is apartial order without a maximal element, or that there is a set with aunary function and a constant which constitute a structure isomorphicto \((\oN,s,0)\), or that the domain is the union of two disjoint setswhich have the same cardinality as the domain, and so on. As is thecase in set theory without the Axiom of Choice, the differentformulations of infiniteness need not be equivalent. In second-orderlogic the situation is even more diffuse because of the variety ofdifferent formulations of the Axiom of Choice. We refer to Asser(1981) for a discussion of the different variants and to Hasenjaeger(1961) for a proof that the various non-equivalent forms of Axioms ofInfinity form in a sense a dense set. For a survey of differentconcepts of finiteness, see de la Cruz (2002).

10. Categoricity

A theory in second (or first) order logic is said to becategorical if any two of its models are isomorphic. A singlesentence (which has a model) is categorical if and only if itcharacterizes some model up to isomorphism (see§7.1). A nice property of a categorical theory \(\Gamma\) is (semantic)completeness: Any sentence \(\phi\) in the language in whichthe theory is written is decided by \(\Gamma\) in the following sense.Either every model of \(\Gamma\) satisfies \(\phi\) or every model of\(\Gamma\) satisfies \(\neg\phi\). In other words, either \(\phi\) or\(\neg\phi\) is a (semantic) logical consequence of \(\Gamma\). Thereason is very simple: \(\Gamma\) has, up to isomorphism, only onemodelM. Since isomorphism preserves truth in second-orderlogic, \(\phi\) or \(\neg\phi\) is a (semantic) logical consequence of\(\Gamma\) according to whether \(\phi\) is true inM or falseinM.

If \(\phi\) is a second-order sentence we defined above \(\Mod(\phi)\)as the class \(\{\mm : \mm\models\phi\}\). If \(\phi\) characterizes\(\mm\) up to isomorphism, then, up to isomorphism,

\[\Mod(\phi)=\{\mm\}.\]

The project of axiomatizing mathematical structures with second-ordersentences was so successful in the first quarter of the twentiethcentury that Carnap even proposed, inspired by Fraenkel (1919 [1928]),see Awodey and Reck 2002 for more on this, thateverymathematical structure is determined, up to isomorphism, by itssecond-order theory. There are trivial cardinality reasons why thiscould not possibly hold for all models of size \(2^\omega\) andlarger. There are simply not enough second-order theories incomparison with the number of non-isomorphic models. However, theproposal of Carnap is trivially true for finite models, and notentirely unreasonable for countable models. In fact Ajtai (1979)showed that it is consistent with ZFC, provided ZFC itself isconsistent, that any two countable structures in a finite vocabularythat satisfy the same second-order sentences are isomorphic. Ajtaishowed that it is also consistent with ZFC, provided ZFC itself isconsistent, that there are two countable structures in a finitevocabulary that satisfy the same second-order sentences without beingisomorphic. Thus Carnap’s conjecture (for countable models) isindependent of ZFC.

Solovay has made a posting in FOM (2006 [Other Internet Resources]) in which he shows that the related statement that every completesecond-order sentence is categorical, is independent of ZFC. Saarinen,Väänänen and Woodin (2024 [Other Internet Resources]) show that large cardinals, at least up to a supercomapct cardinals,cannot decide whether every complete second-order sentence iscategorical,

There is a strong form of categoricity which holds for Henkinstructures in important cases and agrees with the usual concept ofcategoricity in the case of full Henkin models. It builds on theremarkable ability of second-order logic to express its owncategoricity. The isomorphism \((M,R)\cong(M',R')\) of two structures\((M,R)\) and \((M',R')\), where the predicatesR and \(R'\)are, say, binary, can be expressed in second-order logic as follows,letting \(\ma=(A,M,M',R,R')\), whereA is any set containing\(M\cup M'\), \(M=U^\ma\), \(M'={U'}^\ma\), \(R=P^\ma\) and\(R'=P'^\ma\). Now

\[(M,R)\cong(M',R')\iff\ma\models\isom(U,P,U',P'),\]

where \(\isom(U,P,U',P')\) is the sentence

\[\begin{align}&\exists F[\forall x(U(x)\to U'(F(x))) \land {}\\&\quad \forall x\,\exists y(U'(x)\to (U(y)\land x=F(y))) \land {}\\&\quad\forall x\,\forall y((U(x) \land U(y)) \to {}\\&\quad[(F(x)=F(y)\to x=y) \land {}\\&\quad(P(x,y)\leftrightarrow P'(F(x),F(y)))])].\end{align}\]

This suggests that the categoricity of a sentence \(\theta(P)\) with abinary predicate symbolP can be redefined equivalently,[18] letting \(\categ(\theta(P))\) be the sentence

\[\forall P\,\forall P'\,\forall U\,\forall U'\left(\left(\theta(P)^{U}\land\theta(P')^{U'}\right)\to\isom(U,P,U',P')\right),\]

as

\[\tag{12}\label{redef}\models\categ(\theta(P)).\]

In fact, this is how Tarski defines categoricity in Tarski (1956),(page 387). This leads us to define:

Definition 15 We say that \(\theta(P)\) isinternally categorical if

\[\tag{13}\label{internal}\vdash\categ(\theta(P)).\]

Internal categoricity could also be calledprovablecategoricity. The phrase ‘internal categoricity’ wasintroduced in the context of arithmetic in Walmsley (2002). It wasadvocated more generally in Väänänen (2012), with moredetails in Väänänen and Wang (2015). We refer to Buttonand Walsh (2016) as well as Maddy and Väänänen (2023) for a recentdiscussion on this concept.

Note that (\ref{internal}) is stronger than (\ref{redef}) becauseprovable sentences are certainly valid. Owing to the CompletenessTheorem we can rewrite (\ref{internal}) as

\[\tag{14}\label{internals}(\ma,\G)\models\categ(\theta(P))\text{ for all Henkin models $(\ma,\G)$}.\]

Thus internal categoricity means categoricity not only for ordinarymodels, i.e., full Henkin models, but categoricity inside an arbitraryHenkin model. Note that categoricity is a meta-theoretical concept asit refers to the semantics of second-order logic. In\(\categ(\theta(P))\) the categoricity of \(\theta(P)\) is written inthe language of second-order logic. That is, the categoricity, despitebeing a meta-theoretical concept, is written in the object language.It can be asked, whether this leads us to a different concept? Be thatas it may, when the meaning of \(\categ(\theta(P))\) is uncovered byconsidering its semantics in the meta-theory, the meaning is the sameas the meaning we give to the categoricity of \(\theta(P)\). In theconcept of internal categoricity we take advantage of this situationby switching from semantical meaning to proof-theoretic meaning: Weinsist that \(\categ(\theta(P))\) is (not only valid but even)provable. Proof-theory requires less meta-theory than semantics. Wecan investigate provability of second-order sentences even in numbertheory by using numbers to code sentences and proofs, as Gödeldid in his proof of the Incompleteness Theorem. Of course, theCompleteness Theorem brings us back to the semantics.

To verify the categoricity of a second-order sentence one has to gothrough infinite structures in a way which essentially calls for settheory. The situation with internal categoricity is different. Toverify the internal categoricity of a second-order sentence one justhas to produce a proof. So there is a dramatic difference. And stillinternal categoricity is stronger than categoricity. So it would befoolish to establish categoricity if one could establish even internalcategoricity. Fortunately, the classical examples of categoricalsentences are all internally categorical:

Theorem 16 (Väänänen & Wang 2015)The received second-order sentences characterizing the structures\((\oN,<)\) and \((\oR,<,+,\cdot,0,1)\) are internallycategorical.

The concept of internal categoricity provides a bridge between fullsemantics and Henkin semantics. It works in the same way in both casesand shows that the full semantics is a limit case of Henkin semanticsbut does not have a monopoly when it comes to categoricity.

11. Logics Between First and Second Order

First order logic and second-order logic are in a sense two oppositeextremes. There are many logics between them i.e., logics that extendproperly first order logic, and are properly contained in second-orderlogic. One example is the extension of first order logic by thegeneralized quantifier known as the Henkin quantifier:

\[\left( \begin{array}{cc}\forall x&\exists y\\\forall u&\exists v\end{array}\right)\phi(x,y,u,v,z_1,\ldots,z_n)\]

which has the meaning

\[\exists f\,\exists g\,\forall x\,\forall u\,\phi(x,f(x),u,g(u),z_1,\ldots,z_n).\]

The extension \(L(H)\) of first order logic by the Henkin quantifieris almost the same as second-order logic: they have the same\(\Delta\)-extension (Krynicki & Lachlan 1979).[19]

The equicardinality or Härtig quantifier

\[Ixy\,\phi(x,z_1,\ldots,z_n)\psi(x,z_1,\ldots,z_n)\]

has the meaning “there are as many elementsx satisfying\(\phi(x,z_1,\ldots,z_n)\) as there are elementsy satisfying\(\psi(y,z_1,\ldots,z_n)\)”, i.e.,

\[\begin{align}&\exists f[\forall x(\phi(x,z_1,\ldots,z_n)\leftrightarrow\psi(f(x),z_1,\ldots,z_n))\land {}\\&\quad\forall x\,\forall y((\phi(x,z_1,\ldots,z_n)\land \phi(y,z_1,\ldots,z_n)\land f(x)=f(y))\to x=y)\land {}\\&\quad\forall y\,\exists x(\psi(y,z_1,\ldots,z_n)\to (\phi(x,z_1,\ldots,z_n)\land f(x)=y))]\end{align}\]

The extension \(L(I)\) of first order logic by the Härtigquantifier is weaker[20] than second-order logic as its Löwenheim number can be\(<2^\omega\), but if \(V=L\), then it is \(\Delta\)-equivalentwith second-order logic (Väänänen 1980, Väänänen andWelch 2023). For more on \(L(I)\), see Herre, Krynicki, Pinus andVäänänen (1991) as well as Magidor and Väänänen (2011).

Here are some other generalized quantifiers that are clearlysecond-order definable:

\[\begin{align}Q_0x\,\phi(x,z_1,\ldots,z_n)& \iff\exists X(|X|\ge\aleph_0\land \forall x\in X\phi(x,z_1,\ldots,z_n))\\Q_1x\,\phi(x,z_1,\ldots,z_n)&\iff\exists X( |X|\ge\aleph_1\land{} \forall x\in X\phi(x,z_1,\ldots,z_n))\\Q^{MM}_1xy\,\phi(x,y,z_1,\ldots,z_n)&\iff\exists X(|X|\ge\aleph_1\land{} \forall x,y\in X\phi(x,y,z_1,\ldots,z_n))\\Q^{cof}_\omega xy\,\phi(x,y,z_1,\ldots,z_n)&\iff \{(x,y):\phi(x,y,z_1,\ldots,z_n)\} \\& \qquad\qquad\text{ is a linear order of cofinality } \omega\end{align}\]

Inweak second-order logic we have no function variables andthe relation variables range over finite relations only. The resultinglogic is in many ways similar to the extension of first order logic bythe generalized quantifier \(Q_0\). Another way to limit the power ofsecond-order logic is the following: Suppose \(\psi\) is a first orderformula with ann-ary relation variableX and nonon-logical symbols. For infiniteA, we define \(\ma\models_sQ_{\psi}X\phi\) if and only if there is a relation \(P\subseteq M^n\)such that \(\ma\models_{s(P/X)}\phi\land\psi\). The quantifier\(Q_{\psi}\) allows second-order quantification over relations thatsatisfies \(\psi\). If \(\psi\) is \(\forall x(x=x)\) we obtain theusual second-order quantifier which we denote \(Q_{II}\). If moreover\( n=1\) we obtain the monadic second-order quantifier which we denote\(Q_{mon}\). If \(n=1\) and \(\psi\) says thatX is asingleton, we obtain the usual first order quantifier which we nowdenote \(Q_I\). Finally, if \(\psi\) saysX is the graph of apermutation of the model, we denote this quantifier by \(Q_{1-1}\).Rather surprisingly, it turns out that with the right concept ofinterpretability the quantifiers \(Q_I,Q_{mon},Q_{1-1}\) and\(Q_{II}\) are, up to biinterpretability, theonlysecond-order quantifiers of the form \(Q_{\psi}\) (Shelah 1973).

12. Higher Order Logic vis-à-vis Type Theory

There are many ways to further extend second-order logic. The mostobvious is third, fourth, and so on order logic. The generalprinciple, already recognized by Tarski (1933 [1956]), is that inhigher order logic one can formalize the semantics—definetruth—of lower order logic. Therefore one certainly gainsexpressive power by using higher order logic. Let us take numbertheory as an example. With first order logic we can express propertiessuch as “n is a prime number” and propositions suchas “there are infinitely many prime numbers”,Fermat’s Last Theorem and Goldbach’s Conjecture. Firstorder properties of natural numbers fall into a natural hierarchycalled thearithmetical hierarchy the levels of which aredenoted \(\Sigma^0_n\) and \(\Pi^0_n\). With second-order logic we cantalk about properties of sets of natural numbers, or in other words,properties of real numbers. In particular, we can define therationals. A continuous function on the real numbers is determined byits values on rationals. Also, an open set is determined by whichrational points it contains. Using these observations we can talkabout continuous functions and open sets of reals. Basic facts incalculus, such as Bolzano’s Theorem stating that “If acontinuous function has values of opposite sign inside an interval,then it has a root in that interval” are expressible insecond-order logic over the structure \((\oN,+,\cdot)\) of naturalnumbers. Second-order properties of natural numbers fall into anatural hierarchy called theanalytical hierarchy and itslevels are denoted \(\Sigma^1_n\) and \(\Pi^1_n\). With third orderlogic we can talk aboutsets of real numbers. We can writethe Continuum Hypothesis in third order logic and ask the hardquestion, whether the model \((\oN,+,\cdot)\) satisfies this sentence.Third order properties of natural numbers have again a naturalhierarchy and its levels are denoted \(\Sigma^2_n\) and\(\Pi^2_n\).

It is obvious that if we keep a base model, such as \((\oN,+,\cdot)\)fixed, and move to higher and higher order logic, we can express moreand more complicated concepts and propositions. However,third order logic over

\[\mn_1=(\oN,+,\cdot)\]

is nothing more thansecond order logic over

\[\mn_2=(\P(\oN),E,\oN,+,\cdot),\]

where \(E\subseteq \oN\times\P(\oN)\) is the relation \(nEr\iff n\inr\), which again is nothing more thanfirst order logic over

\[\mn_3=(\P(\P(\oN))),E',\P(\oN),E,\oN,+,\cdot),\]

where \(E'\subseteq \P(\oN)\times\P(\P(\oN)\) is the relation\(rE'X\iff r\in X\). Moreover, second-order logic over \(\mn_1\) isnothing more than first order logic over \(\mn_2\). When this line ofthought is taken to its limit we can see that first order logic overthe cumulative hierarchy of sets covers higher order logic over anyindividual level \(V_\alpha\) of the hierarchy. In a sense, firstorder logic over the cumulative hierarchy of sets is a very very highorder logic over \(\mn_1\).

There is a marked similarity in the way \(\mn_2\) arises from\(\mn_1\) and in the way \(\mn_3\) arises from \(\mn_2\). In bothcases we simply add a power set construction on top of what the modelalready has. Second-order logic alone is very good with power sets aswe saw in§5.3. A consequence of this is that all the higher order logics (of anyfinite order and up to a point also of infinite order) fromsecond-order on have Turing-equivalent decision problems and the sameHanf- and Löwenheim numbers (see§4). Thus, even though higher than second-order logics are strictlyspeaking stronger than second-order logic no difference can be seen bythe usual criteria of strength of expressive power.

Type theory is a systematic approach to higher order logics.Variables are assigned types, just as in second-order logic we havevariables for individual type, relation type and function type. In ahigher order logic it would make sense to have variables for relationsbetween objects of any lower types as well as functions mappingobjects of lower type to objects of lower type. In set theory objectshave a rank. Elements of a set have a lower rank than the set itself.However, the difference between set-theoretical rank and typing intype theory is that in set theory the rank can be computed when theset is known while in type theory the type of an object can be seenfrom the type of the variable that the object is the value of. In asense type theory imposes from the syntax a type for each objectconsidered while in set theory a set has a rank given by thesemantics. Arguably set theory gained dominance over type theory afterthe 1930s because of its simpler syntax and semantics. On the otherhand type theory is used in computer science, especially in the theoryof programming languages (see SEP entry ontype theory).

13. Foundations of Mathematics

Mathematics can be based on set theory. This means that mathematicalobjects are construed as sets and their properties are derived fromthe axioms of set theory. The intuitive informal picture behind settheory is that there is a cumulative hierarchy \(V_\alpha\)(\(\alpha\) an ordinal) of sets and the axioms are intended todescribe the basic properties of such sets. This kind of settheoretical foundation of mathematics has become widely accepted amongworking mathematicians. We refer to the SEP entry onset theory for more details. Alternatives to set theory are at leastcategory theory,constructive mathematics, and second-order logic.

If one tries to follow as a logician the language mathematicians usein their research, one cannot help seeing that they use second-orderconcepts very freely. For example, the Induction Axiom is withoutdoubt used by mathematicians in the second-order form depicted in(\ref{ind}). In their natural language mathematicians do not hesitateto use higher order concepts, and indeed do not always even (need to)know the difference.

Second-order logic as a foundation for mathematics focuses onmathematical structures rather than mathematical“objects”, as set theory does. In principle everystructure has its own “foundation” based on the specialfeatures of the structure. The role of second-order logic is toprovide a common framework for this. The intuitive informal picturebehind second-order logic on a structure is that there is a domainA of individuals and then separate domains for alln-aryrelations for all \(n\in\oN\) and for alln-ary functions forall \(n\in\oN\), associated withA. In comparison to theinformal picture of set theory, we have only the domains of subsets,relations and functions, not iterations such as sets of sets, sets ofsets of sets and so on, as in set theory. Therefore the intuitivepicture behind second-order logic is simpler than the picture behindset theory. In particular, there is no transfinite iteration over allordinals. Most structures \(\ma\) used in mathematics have asecond-order characterization \(\theta_\ma\), see§10. Proving (second order) properties \(\phi\) of such structures meansderiving \(\phi\) from \(\theta_\ma\). Opinions differ as to whetherthis derivation should be a syntactic (formal) derivation using some(incomplete) axiom system of second-order logic or whether it shouldbe a semantic derivation (of the form “Every model of theassumptions is a model of the conclusion”). For a workingmathematician there is not much difference and probably a semanticderivation is usually favored.

Suppose a mathematician wants to convince someone about the truth ofBolzano’s Theorem “If a continuous function on the realshas negative values as well as positive values inside an interval,then it has a root in that interval”. They would start with theordered field of real numbers. It would be (thought to be) clear whatthis means because the second-order axiomatization of the field iscategorical. Then they would take an arbitrary continuous functionf on this field such that it has values of opposite sign insidea fixed interval \((a,b)\). It would be clear what this means becausesecond-order logic has variables for functions. Let \(c,d\in (a,b)\)such that \(f(c)<0\) and \(f(d)>0\). Without loss of generality,\(c<d\). Let \(X=\{e\in(a,b) : f(e)<0\}\). Since we haverelation variables for subsets of the domain, we can think ofXsimply as a value of such a relation variable. It is obvious thatX exists, but a priori we could have imposed contradictoryconditions onX (e.g., \(X=\{e : e\notin X\}\)) and then weshould not be able to claim that it exists. However, in this case theComprehension Axiom Schema implies thatX exists. Clearly,\(X\ne\emptyset\) andX is bounded from above byd. Oneof the second-order axioms characterizing the field of real numbers isthat every non-empty set which is bounded from above has a supremum.Thus we can letz to be the supremum ofX. Now itfollows from basic properties of continuous functions that\(f(z)=0\).

It would be difficult to tell in the above proof whether it was asyntactical derivation from the axioms or a semantic argument. On thesurface it looks like a semantic argument. But every step can bederived from the axioms, so it could be considered a shorthand versionof a syntactic derivation.

Syntactic derivations in second-order logic based on the ComprehensionAxiom Schema and Axioms of Choice are very much like syntacticderivations in set theory. In neither case would a workingmathematicians write all the details of the argument but would resortto shorthand notation. In both cases—second-order logic and settheory—we can interpret the proofs as shorthand versions ofsyntactic formal proofs or as semantic proofs. First order logicsatisfies the Completeness Theorem with the help of which semanticproofs can be turned into syntactic formal proofs. Since theCompleteness Theorem uses the concept of a first order structure,sentences proved from the axioms are true in all models of the axioms,both countable models and non-standard models. Are they true in theintuitive model of cumulative hierarchy of sets? By the ReflectionTheorem such sentences are true in the class of all sets, but sincethe intuitive model is informal, we cannotprove that theclass of all sets is the same as the intuitive model.

On the other hand, in order for the Completeness Theorem to beapplicable, a semantic argument has to be such that it works whatevermodel of set theory we are using. This means that a semantic argumentin set theory should not use properties of individual models of settheory unless they are universal first order properties of all modelsof ZFC. In practice this means that if a property such as \(CH\),\(\Diamond\), \(2^{\aleph_0}=\aleph_2\), and so on, is used, it ismentioned separately as an assumption. Non-first order properties ofmodels cannot be used at all, or else the Completeness Theorem cannotbe used. Still, sometimes non-first order propertiesareused. For example, we may have an argument that everycountable model of ZFC satisfies \(\phi\). But then, by theDownward Löwenheim-Skolem Theorem,every model of ZFCsatisfies \(\phi\). Thus we have been able to eliminate the non-firstorder assumption of the countability of the model.

Second-order logic satisfies the Completeness Theorem if Henkin modelsare used. Thus semantic arguments in second-order logic can be turnedinto syntactic formal proofs, but as with set theory the semanticargument has to be valid inall Henkin models, not only inall of thefull Henkin models. Thus, as in set theory, thesemantic argument cannot use properties of the Henkin models that arenot shared by all Henkin models, In particular, fullness (see§9.1) cannot be assumed because in general models it is not a second-orderproperty. What happens to categoricity when fullness is abandoned? Westill have internal categoricity, so we can use categoricity in proofsas long as we make sure we work inside one model. One certain way tomake sure of this is that we base everything on the axioms, theComprehension Axiom Schema and the Axioms of Choice.

When second-order logic is used as a foundation for mathematics, asituation may emerge in which third order logic is needed. Forexample, atopology is usually defined as a family of setsand is therefore a third order object. It would be well in the spiritof second-order logic to simply include third (or higher) order logicwhen needed. Naturally, the Comprehension Axiom Schema and the Axiomsof Choice have to be then assumed for third order quantifiers.

We have devoted a fair amount of space in this entry to the power ofsecond-order logic to characterize structures (see§7.1). Putting it briefly, if a structure has mathematical interest, it issecond-order characterizable. Structures constructed by means of theAxiom of Choice may escape being second-order characterizable. But,what is the power of second-order logic to demonstrate theexistence of structures? The axioms, including theComprehension Axioms and the Axioms of Choice, can be satisfied in aone element domain. So we cannotprove from the axioms thatthere are any structures of size \(>1\). This underminesQuine’s characterization of second-order logic as “settheory in sheep’s clothing” (1970, section heading inchapter 5), meant probably to suggest that the ontological commitmentsof second-order logic are on the same level as those of settheory.

Perhaps the philosophy of second-order logic is that if a structurecan be characterized, it exists, echoing Hilbert’s formalistphilosophy. However, even in that sense we need to know that thecharacterizing sentence \(\theta_\ma\) is consistent. A trivialapproach is to observe that \(\ma\) itself certainly satisfies\(\theta_\ma\), but this clearly begs the question. Still, this is howHilbert argues for the consistency of his axioms for the real numbers(1900). If \(\ma=(\oN,+,\cdot)\), an Axiom of Infinity saying thatthat a proper subset of the domain has the same cardinality as theentire domain, is what is needed, as was shown by Dedekind. Incombination with the Comprehension Axiom Schema the existence of\(\ma\) follows. For \((\oR,+,\cdot,0,1)\) this is not enough, we needa stronger Axiom of Infinity, for example the existence of a denseorder which satisfies the Completeness Axiom: Every bounded non-emptyset has a supremum. It seems inevitable that when we move from astructure to a bigger structure we need to make a largenessassumption. Such assumptions are called Large Domain Assumptions inVäänänen (2012). In set theory this aspect is takencare of at the outset by assuming the Power Set Axiom and theReplacement Axioms which together make sure that there are largeenough sets. In second-order logic we have to assume new and new LargeDomain Assumptions, as we move along. This can be seen as a drawback,having to make new assumptions all the time. On the other hand it canbe seen as an asset, not having to make too ambitious assumptionsbefore they are really needed.

14. Second-Order Arithmetic

Owing to the central role that real numbers play in mathematics, thesecond-order theory of natural numbers, known assecond-orderarithmetic (Simpson 1999 [2009]), and denoted \(Z_2\), is animportant foundational theory. It is stronger than (first order) Peanoarithmetic but weaker than set theory. It has variables forindividuals thought of as natural numbers as well as variables forsets of natural numbers thought of as real numbers. In addition thereare \(+\) and \(\times\) for arithmetic operations on the individuals.As axioms \(Z_2\) has some rather obvious axioms about \(+\) and\(\times\), the Induction Axiom (\ref{ind}), and the axioms ofsecond-order logic, including the Comprehension Principles (\ref{CA}).A surprising amount of mathematics can be derived in \(Z_2\). We referto Simpson (1999 [2009]) for details. In a sense, \(Z_2\) is a greatsuccess story for second-order logic.

Reverse mathematics uses \(Z_2\) to isolate the exact axiomson which well-known theorems from mathematics rely. In a textbook suchtheorems are proved perhaps in an informal set theory, but how muchset theory is actually needed in each case? For example, we may askwhat is the weakest set of axioms from which the Bolzano-Weierstrass Theorem[21] can be proved? How much set theory, comprehension, choice, induction,etc is needed? Since \(Z_2\) is a natural and sufficient environmentfor many mathematical theorems, it is an appropriate framework foranswering questions raised by the reverse mathematics program. Themain (but not the only) distinctions that are made in reversemathematics concern the amount of the Comprehension Principle that isneeded in proving this or that mathematical result. In particular, therole of Arithmetic and \(\Pi^1_1\)-Comprehension Principles isclarified. The basic theory of real numbers can be developed on thebasis of Arithmetic Comprehension only, but proofs relying on thenotion of a countable ordinal require the \(\Pi^1_1\)-ComprehensionPrinciple. Again, we refer to Simpson (1999 [2009]) for details.

15. Second-Order Set Theory

We have up to now treated set theory (ZFC) as a first order theory.However, when Zermelo (1930) introduced the axioms which constitutethe modern ZFC axiom system, he formulated the axioms in second-orderlogic. In particular, hisSeparation Axiom is

\[\forall x\,\forall X\,\exists y\,\forall z(z\in Y\leftrightarrow (z\in x\land X(z)))\]

and theReplacement Axiom is

\[\forall x\,\forall F\,\exists y\,\forall z(z\in y\leftrightarrow\exists u(u\in x\land z=F(u))).\]

Second-order ZFC, \(\ZFC^2\), is simply the received first order ZFCwith the Separation Schema replaced by the above single SeparationAxiom, and the Replacement Schema replaced by the above singleReplacement Axiom. Accordingly, \(\ZFC^2\) is a finite axiom system.Zermelo proved that the models of \(\ZFC^2\) are, up to isomorphism,of the form \((V_\kappa,\in)\), where \(\kappa\) is (strongly)inaccessible \(>\omega\).

Kreisel (1967) has pointed out that second-order set theory in a sensedecides the CH, i.e., decides whether it is true or not, even if we donot know which way the decision goes. More exactly \(\ZFC^2\modelsCH\) or \(\ZFC^2\models \neg CH\), because \(CH\) is true if and onlyif \(V_\kappa\models CH\) for inaccessible \(\kappa\), i.e., if andonly if \(\ZFC^2\models CH\). Of course we can express CH in firstorder set theory, too, so the situation is not really different fromfirst order set theory. Many set theorists think that the concept ofset is definitive enough to decide eventually also CH even if ZFC doesnot decide it. Likewise, we may argue that the concept of second-ordersemantics is definitive enough to decide CH even if the current axiomsof second-order logic cannot do it.

The question arises, why do we not use second-order set theory\(\ZFC^2\) as the metatheory of second-order logic, as recommended inShapiro (1991). In fact we could use it. However, the question mightrise, what is the semantics of our metatheory? In principle suchquestions can lead to an infinite regress. By using first order settheory as the metatheory, the question about the semantics of themetatheory would simply be the question about the semantics of firstorder logic. We have pointed out in§6 that semantics of first order logic is absolute relative to ZFC. Thisgives some assurance that we need not continue asking, what themetatheory is.

16. Finite Model Theory

The rise of mathematical logic in the first half of the twentiethcentury was interwoven with different approaches to understanding theinfinite structures of the natural and the real numbers. With theemergence of computer science in the second half of the twentiethcentury logic went through a process of rebirth. The concepts of acomputation and of a database both emphasized the need to understandnot so much the finite/infinite distinction as the new measures ofdegrees of finiteness. For example, the question whether a problem canbe decided in polynomial time arose to challenge the well-developedtheory of what can be decided in finite time. The analogy betweenfinite relational structures and databases led to the emergence offinite model theory. For a good review we refer to Fagin(1993).

In the context of finite models second-order logic does not sufferfrom the same philosophical problems as in the context of infinitemodels. It is just one language among many others. One of the earlysuccesses was the result of Fagin (1974) to the effect that classes ofmodels definable in existential second-order logic, i.e., in the\(\Sigma^1_1\)-fragment, are exactly those that are NP, i.e., that canbe recognized by a non-deterministic Turing machine running in timewhich is polynomial in the size of the encoding of the model as abinary sequence. The question whether NP is closed under complementsis notoriously open. From the point of view of second-order logic thisis surprising because on all models, finite or infinite,\(\Sigma^1_1\) is of course not closed under complements. For example,the class of infinite models (in the empty vocabulary) is\(\Sigma^1_1\) but not \(\Pi^1_1\). Also, the class of well-orders is\(\Pi^1_1\) but not \(\Sigma^1_1\). Therefore the following earlyresult about second-order logic on finite models was remarkable. Here\(\mon\Sigma^1_1\) refers to \(\Sigma^1_1\) in monadic second-orderlogic, similarly \(\mon\Pi^1_1\).

Theorem 17 (Fagin 1975) Connectivity of graphs is notdefinable in \(\mon\Sigma^1_1\). Hence\(\mon\Sigma^1_1\ne\mon\Pi^1_1\).

This has subsequently been extended to graphs with certain otherstructure present, see Fagin, Stockmeyer, and Vardi (1995). The proofsuse the Ehrenfeucht-Fraïssé games (see§3.1) and their elaborations.

Let \(\Sigma^1_{1,m}\) denote the class of \(\Sigma^1_1\)-formulas\(\exists X_1\ldots\exists X_k\phi\), where the bound second-ordervariables \(X_i\) are at mostm-ary and \(\phi\) is firstorder. Similarly \(\Pi^1_{1,m}\) and \(\Delta^1_{1,m}\).

Theorem 18 (Ajtai 1983) The property of a finitestructure \((A,R)\), whereR is \(n+1\)-ary, of \(|R|\) beingeven, is \(\Delta^1_{1,n+1}\) but not \(\Sigma^1_{1,n}\) or\(\Pi^1_{1,n}\).

This highly non-trivial theorem is one of the corner stones of thestudy of second-order logic on finite structures. We do not knowwhether \(\Delta^1_1\) is different from \(\Sigma^1_1\) on finitestructures, but the above theorem gives an arity-based hierarchyresult inside \(\Delta^1_1\).

Bibliography

  • Ajtai, M., 1979, “Isomorphism and Higher OrderEquivalence”,Annals of Mathematical Logic, 16(3):181–203. doi:10.1016/0003-4843(79)90001-9
  • –––, 1983,“\(\Sigma_1^1\)-Formulae on Finite Structures”,Annalsof Pure and Applied Logic, 24(1): 1–48.doi:10.1016/0168-0072(83)90038-6
  • Asser, Günter, 1981,Einführung in DieMathematische Logik: Teil Ill: Prädikatenlogik HöhererStufe, Thun: Verlag Harri Deutsch.
  • Awodey, Steve and Erich H. Reck, 2002, “Completeness andcategoricity. I. Nineteenth-century axiomatics to twentieth-centurymetalogic”,History and Philosophy of Logic, 23(1):1–30.
  • Barwise, K. Jon, 1972a, “Absolute Logics and\(L_{\infty \omega }\)”,Annals of Mathematical Logic,4(3): 309–340. doi:10.1016/0003-4843(72)90002-2
  • –––, 1972b, “TheHanf Number of Second Order Logic”,Journal of SymbolicLogic, 37(3): 588–594. doi:10.2307/2272748
  • Büchi, J. Richard, 1962, “On a DecisionMethod in Restricted Second Order Arithmetic”, inLogic,Methodology and Philosophy of Science: Proceedings of the 1960International Congress, Ernest Nagel, Patrick Suppes, and AlfredTarski (eds.), Stanford, CA: Stanford University Press,1–11.
  • –––, 1973, “The MonadicSecond Order Theory of \(\omega _{i}\)”, inDecidableTheories II: The Monadic Second Order Theory of All CountableOrdinals, by J. Richard Büchi and Dirk Siefkes, edited by G.H. Müller and D. Siefkes, (Lecture Notes in Mathematics 328),Berlin, Heidelberg: Springer Berlin Heidelberg, 1–127.doi:10.1007/BFb0082721
  • Buss, Samuel R. (ed.), 1998,Handbook of ProofTheory, (Studies in Logic and the Foundations of Mathematics137), New York: Elsevier.
  • Button, Tim and Sean Walsh, 2016,“Structure and Categoricity: Determinacy of Reference and TruthValue in the Philosophy of Mathematics”,PhilosophiaMathematica, 24(3): 283–307.doi:10.1093/philmat/nkw007
  • Carnap, Rudolf, 1927/1930 [2000].Untersuchungen zurallgemeinen Axiomatik, edited and with a foreword by Thomas Bonkand Jesus Mosterin, Darmstadt: WissenschaftlicheBuchgesellschaft.
  • Church, Alonzo, 1956,Introduction toMathematical Logic, volume 1, Revised and enlarged edition,(Princeton Mathematical Series 17), Princeton, NJ: PrincetonUniversity Press.
  • Cohen, Paul J., 1966,Set Theory and theContinuum Hypothesis, New York-Amsterdam: W. A. Benjamin.
  • Craig, William, 1965, “Satisfaction for\(n\)-th order languages defined in \(n\)-th order languages”,Journal of Symbolic Logic, 30: 13–25.
  • de la Cruz, Omar, 2002, “Finiteness andChoice”,Fundamenta Mathematicae, 173(1): 57–76.doi:10.4064/fm173-1-4
  • Durand, Arnaud, Ronald Fagin, and Bernd Loescher,1998, “Spectra with Only Unary Function Symbols”, inComputer Science Logic, Mogens Nielsen and Wolfgang Thomas(eds.), (Lecture Notes in Computer Science 1414), Berlin, Heidelberg:Springer Berlin Heidelberg, 189–202. doi:10.1007/BFb0028015
  • Durand, Arnaud, Neil D. Jones, Johann A. Makowsky,and Malika More, 2012, “Fifty Years of the Spectrum Problem:Survey and New Results”,The Bulletin of SymbolicLogic, 18(4): 505–553. doi:10.2178/bsl.1804020
  • Elgot, Calvin C., 1961, “Decision Problems ofFinite Automata Design and Related Arithmetics”,Transactions of the American Mathematical Society, 98(1):21–21. doi:10.1090/S0002-9947-1961-0139530-9
  • Fagin, Ronald, 1974, “Generalized First-OrderSpectra and Polynomial-Time Recognizable Sets”, inComplexity of Computation, Richard M. Karp (ed.), (SIAM-AMSProceedings 7), Providence, RI: American Mathematical Society,43–73.
  • –––, 1975, “MonadicGeneralized Spectra”,Zeitschrift für mathematischeLogik und Grundlagen der Mathematik, 21(1): 89–96.doi:10.1002/malq.19750210112
  • –––, 1993, “Finite-ModelTheory—a Personal Perspective”,Theoretical ComputerScience, 116(1): 3–31.doi:10.1016/0304-3975(93)90218-I
  • Fagin, Ronald, Larry J. Stockmeyer, and Moshe Y.Vardi, 1995, “On Monadic NP vs Monadic Co-NP”,Information and Computation, 120(1): 78–92.doi:10.1006/inco.1995.1100
  • Feferman, Solomon, 1999, “Does MathematicsNeed New Axioms?”,The American Mathematical Monthly,106(2): 99–111. doi:10.1080/00029890.1999.12005017
  • Feferman, Solomon and Georg Kreisel, 1966,“Persistent and Invariant Formulas Relative to Theories ofHigher Order”,Bulletin of the American MathematicalSociety, 72(3): 480–486.doi:10.1090/S0002-9904-1966-11507-0
  • Fraenkel, Abraham, 1919 [1928],Einleitung in dieMengenlehre (Die Grundlehren der mathematischen Wissenschaften inEinseldarstellungen: Volume 9), Berlin: Springer, 3rd edition,1928.
  • Frege, Gottlob, 1879,Begriffsschrift:Eine Der Arithmetischen Nachgebildete Formelsprache Des ReinenDenkens, Halle.
  • –––, 1884,DieGrundlagen Der Arithmetik, Breslau: Verlage Wilhelm Koebner.
  • Garland, Stephen J., 1974, “Second-OrderCardinal Characterizability”, inAxiomatic Set Theory, Part2, T. J. Jech (ed.), (Proceedings of Symposia in PureMathematics, 13.2), Providence, RI: American Math Society,127–146. [Garland 1974 available online]
  • Gaßner, Christine, 1994, “The Axiom ofChoice in Second-Order Predicate Logic”,Mathematical LogicQuarterly, 40(4): 533–546.doi:10.1002/malq.19940400410
  • Gödel, Kurt, 1931 [1986],“Über formal unentscheidbare Sätze der PrincipiaMathematica und verwandter Systeme I”,Monatshefte fürMathematik und Physik, 38(1): 173–198. Reprinted andtranslated as “On Formally Undecidable Propositions ofPrincipia Mathematica and Related Systems” inGödel 1986: 144–195. doi:10.1007/BF01700692
  • –––, 1936 [1986],“Über Die Länge von Beweisen”,ErgebnisseEirtes Mathematischen Kolloquiums, 7: 23–24. Reprinted andtranslated as “On the Length of Proofs” in Gödel1986: 396–399.
  • –––, 1939 [1990],“Consistency-Proof for the GeneralizedContinuum-Hypothesis”,Proceedings of the National Academyof Sciences, 25(4): 220–224. Reprinted in Gödel 1990:28–32. doi:10.1073/pnas.25.4.220
  • –––, 1986,Collected Works,Volume 1: Publications 1929-1936, Solomon Feferman (ed.), NewYork: Oxford University Press.
  • –––, 1990,Collected Works,Volume 2: Publications 1938-1974, Solomon Feferman (ed.), NewYork: Oxford University Press.
  • Gurevich, Yuri, 1985, “Monadic Second-OrderTheories”, inModel-Theoretic Logics, Jon Barwise andSol Feferman (eds.) (Perspectives in Logic), New York:Springer-Verlag, 479–506. doi:10.1017/9781316717158.019
  • Gurevich, Yuri, Menachem Magidor, and SaharonShelah, 1983, “The Monadic Theory of \(\omega_2\)”,Journal of Symbolic Logic, 48(2): 387–398.doi:10.2307/2273556
  • Herre, Heinrich, Michał Krynicki, Alexandr Pinus, and JoukoVäänänen, 1991, “The Härtig-quantifier. Asurvey”,Journal of Symbolic Logic, 56(4):1153–1183.
  • Hasenjaeger, Gisbert, 1961,“Unabhängigkeitsbeweise in Mengenlehre Und Stufenlogik DerModelle.”,Jahresbericht Der DeutschenMathematiker-Vereinigung, 63: 141–162.
  • Hellman, Geoffrey, 2001, “ThreeVarieties of Mathematical Structuralism†”,Philosophia Mathematica, 9(2): 184–211.doi:10.1093/philmat/9.2.184
  • Henkin, Leon, 1950, “Completeness in theTheory of Types”,Journal of Symbolic Logic, 15(2):81–91. doi:10.2307/2266967
  • Herbrand, Jacques, 1930,Recherches sur lathéorie de la démonstration, (Thèses del’entre-deux-guerres 110), L’Université de Paris. [Herbrand 1930 available online]
  • Hilbert, David, 1900, “Über DenZahlbegriff.”,Jahresbericht Der DeutschenMathematiker-Vereinigung, 8: 180–183.
  • Hilbert, David and William Ackermann, 1928,Grundzüge Der Theoretischen Logik, (Grundlehren DerMathematischen Wissenschaften in Einzeldarstellungen Mit BesondererBerücksichtigung Der Anwendungsgebiete 27), Berlin: J.Springer.
  • –––, 1938,Grundzüge Der Theoretischen Logik, second edition,Berlin: Springer. doi:10.1007/978-3-662-41928-1
  • Hintikka, K. Jaakko J., 1955, “Reductions inthe Theory of Types”,Acta Philosophica Fennica, 8:57–115.
  • Hyttinen, Tapani, Kaisa Kangas, and JoukoVäänänen, 2013, “On Second-OrderCharacterizability”,Logic Journal of IGPL, 21(5):767–787. doi:10.1093/jigpal/jzs047
  • Jones, Neil D. and Alan L. Selman, 1974,“Turing Machines and the Spectra of First-Order Formulas”,Journal of Symbolic Logic, 39(1): 139–150.doi:10.2307/2272354
  • Krawczyk, A. and W. Marek, 1977, “On theRules of Proof Generated by Hierarchies”, inSet Theory andHierarchy Theory V, Alistair Lachlan, Marian Srebrny, and AndrzejZarach (eds.) (Lecture Notes in Mathematics 619), Berlin: SpringerBerlin Heidelberg, 227–239. doi:10.1007/BFb0067654
  • Kreisel, Georg, 1967, “InformalRigour and Completeness Proofs”, inProblems in thePhilosophy of Mathematics, Imre Lakatos (ed.), (Studies in Logicand the Foundations of Mathematics 47), Amsterdam: North-Holland,138–186. doi:10.1016/S0049-237X(08)71525-8
  • Krynicki, Michał and Alistair H. Lachlan, 1979,“On the Semantics of the Henkin Quantifier”,Journalof Symbolic Logic, 44(2): 184–200. doi:10.2307/2273726
  • Kunen, Kenneth (ed.), 1980,Set Theory: AnIntroduction to Independence Proofs, (Studies in Logic and theFoundations of Mathematics 102), Amsterdam: North-Holland.
  • Lévy, Azriel, 1965,A Hierarchy ofFormulas in Set Theory, (Memoirs of the American MathematicalSociety 57), Providence, RI: American Mathematical Society.
  • Lindenbaum, Adolf and Andrzej Mostowski, 1938,“Über Die Unabhängigkeit Des Auswahlaxioms Und EinigerSeiner Folgerungen”,Sprawozdania z PosiedzeńTowarzystwa Naukowego Warszawskiego, Wydział III NaukMatematyczno-Fizycznych (Comptes-Rendus Des Séances de LaSociété Des Sciences et Des Lettres de Varsovie,ClasseIII), 31: 27–32.
  • Löwenheim, Leopold, 1915,“Über Möglichkeiten im Relativkalkül”,Mathematische Annalen, 76(4): 447–470.doi:10.1007/BF01458217
  • Lyndon, Roger C., 1959, “An InterpolationTheorem in the Predicate Calculus.”,Pacific Journal ofMathematics, 9(1): 129–142.
  • Maddy, Penelope and Jouko Väänänen,2023,Philosophical Uses of Categoricity Arguments (Elementsin the Philosophy of Mathematics), Cambridge: Cambridge University Press.doi:10.1017/9781009432894
  • Magidor, Menachem, 1971, “On the Role ofSupercompact and Extendible Cardinals in Logic”,IsraelJournal of Mathematics, 10(2): 147–157.doi:10.1007/BF02771565
  • Magidor, Menachem and Jouko Väänänen, 2011, “OnL¨wenheim-Skolem-Tarski numbers for extensions of first orderlogic”,Journal of Mathematical Logic, 11(1): 87-113.doi:10.1142/S0219061311001018
  • Makowsky, J.A., Saharon Shelah, and Jonathan Stavi,1976, “Δ-Logics and Generalized Quantifiers”,Annals of Mathematical Logic, 10(2): 155–192.doi:10.1016/0003-4843(76)90021-8
  • Manzano, María, 1996,Extensions ofFirst Order Logic, (Cambridge Tracts in Theoretical ComputerScience 19), Cambridge: Cambridge University Press.
  • Montague, Richard, 1963, “Reductions ofHigher-Order Logic”, inThe Theory of Models: Proceedings ofthe 1963 International Symposium at Berkeley, J.W. Addison, LeonHenkin, and Alfred Tarski (eds.), Amsterdam: North-Holland,251–264. doi:10.1016/B978-0-7204-2233-7.50030-7
  • Moore, Gregory H., 1988, “The Emergence ofFirst-Order Logic”, inHistory and Philosophy of ModernMathematics, (Minnesota Studies in the Philosophy of Science 11),Minneapolis, MN: University of Minnesota Press, 95–135.
  • Mostowski, Andrzej, 1938, “Oniezależności definicji skończoności w systemielogiki” (“On the independence of definitions of finitenessin a system of logic”),Dodatek do Rocznika Towarzystwa.Matematycznego, XI: 1–54; English translation in Mostowski1979.
  • Mostowski, Andrzej, 1949, “An UndecidableArithmetical Statement”,Fundamenta Mathematicae,36(1): 143–164.
  • Mostowski, Andrzej, 1979,FoundationalStudies. Selected works (Volume II), Studies in Logic and theFoundations of Mathematics: Volume 93, Amsterdam-New York:North-Holland Publishing Co.; PWN-Polish Scientific Publishers,Warsaw, 1979, edited by Kazimierz Kuratowski, Wiktor Marek, LeszekPacholski, Helena Rasiowa, Czesław Ryll-Nardzewski and PawełZbierski.
  • Parsons, Charles, 1990, “TheStructuralist View of Mathematical Objects”,Synthese,84(3): 303–346. doi:10.1007/BF00485186
  • Quine, W. V. O., 1970,Philosophy ofLogic, Cambridge, MA: Harvard University Press.
  • Rabin, Michael O., 1968, “Decidability ofSecond-Order Theories and Automata on Infinite Trees”,Bulletin of the American Mathematical Society, 74(5):1025–1030. doi:10.1090/S0002-9904-1968-12122-6
  • Rabin, Michael O. and Dana Scott, 1959,“Finite Automata and Their Decision Problems”,IBMJournal of Research and Development, 3(2): 114–125.doi:10.1147/rd.32.0114
  • Resnik, Michael D., 1988, “Second-Order LogicStill Wild”,The Journal of Philosophy, 85(2):75–87. doi:10.2307/2026993
  • Schmidt, Arnold, 1951, “Die Zulässigkeitder Behandlung mehrsortiger Theorien mittels der üblicheneinsortigen Prädikatenlogik”,MathematischeAnnalen, 123(1): 187–200. doi:10.1007/BF02054948
  • Scott, Dana, 1961, “Measurable Cardinals andConstructible Sets”,Bulletin de l’AcadémiePolonaise des Sciences, Série des sciencesmathématiques, astronomiques et physiques, 9:521–524.
  • Shapiro, Stewart, 1991,Foundationswithout Foundationalism: A Case for Second-Order Logic, New York:Oxford University Press. doi:10.1093/0198250290.001.0001
  • Shelah, Saharon, 1973, “There Are Just FourSecond-Order Quantifiers”,Israel Journal ofMathematics, 15(3): 282–300. doi:10.1007/BF02787572
  • –––, 2004, “Spectra ofMonadic Second Order Sentences”,Scientiae MathematicaeJaponicae, (Special issue on set theory and algebraic modeltheory), 59(2): 351–355. [Shelah 2004 available online]
  • Simpson, Stephen G., 1999 [2009],Subsystems ofSecond Order Arithmetic, second edition, (Perspectives in Logic),Cambridge: Cambridge University Press. first edition in 1999, NewYork: Springer. doi:10.1017/CBO9780511581007
  • Siskind, Benjamin, Paolo Mancosu and Stewart Shapiro, 2023,“A note on choice principles in second-order logic”Reviewof Symbolic Logic, 16(2): 339–350. doi:10.1017/S1755020320000301
  • Tarski, Alfred, 1933 [1956],PojęciePrawdy w Językach Nauk Dedukcyjnych (Le concept devérité dans les langages formalisés), (PraceTowarzystwa Naukowego Warszawskiego, Wydział III NaukMatematyczno-fizycznych/Travaux de la Société desSciences et des Lettres de Varsovie, Classe III SciencesMathématiques et Physiques, 34), Warsaw. German translation seeTarski 1936. English translation as “The Concept of Truth inFormalized Languages” in Tarski 1956: chapter 8:152–278.
  • –––, 1936 [1956], “DerWahrheitsbegriff in Den Formalisierten Sprachen”,StudiaPhilosophica, 1: 261–405. Translated as “The Conceptof Truth in Formalized Languages” in Tarski 1956: chapter8.
  • –––, 1956,Logic, Semantics,Metamathematics; Papers from 1923 to 1938, J. H. Woodger(trans.), Oxford: Clarendon Press.
  • Tharp, Leslie H., 1973, “The Characterizationof Monadic Logic”,Journal of Symbolic Logic, 38(3):481–488. doi:10.2307/2273046
  • Väänänen, Jouko, 1979,“Abstract Logic and Set Theory. I. Definability”, inLogic Colloquium ’78: Proceedings of the Colloquium Held inMons, Maurice Boffa, Dirk van Dalen, and Kenneth Mcaloon (eds.),(Studies in Logic and the Foundations of Mathematics 97), Amsterdam:Elsevier, 391–421. doi:10.1016/S0049-237X(08)71637-9
  • –––, 1980, “Boolean ValuedModels and Generalized Quantifiers”,Annals of MathematicalLogic, 18(3): 193–225.doi:10.1016/0003-4843(80)90005-4
  • –––, 2012,“Second Order Logic or Set Theory?”,The Bulletin ofSymbolic Logic, 18(1): 91–121.doi:10.2178/bsl/1327328440
  • Väänänen, Jouko and Tong Wang, 2015,“Internal Categoricity in Arithmetic and Set Theory”,Notre Dame Journal of Formal Logic, 56(1): 121–134.doi:10.1215/00294527-2835038
  • Väänänen, Jouko and Philip Welch, 2023, “Whencardinals determine the power set: inner models and Härtig quantifierlogic”,Mathematical Logic Quarterly, 69(4): 460–471.doi:10.1002/malq.202200030
  • Walmsley, James, 2002,“Categoricity and Indefinite Extensibility”,Proceedings of the Aristotelian Society, 102(1):239–257. doi:10.1111/j.0066-7372.2003.00052.x
  • Wang, Hao, 1952, “Logic of Many-SortedTheories”,Journal of Symbolic Logic, 17(2):105–116. doi:10.2307/2266241
  • Zermelo, Ernst, 1930, “ÜberGrenzzahlen und Mengenbereiche: Neue Untersuchungen über dieGrundlagen der Mengenlehre”,FundamentaMathematicæ, 16: 29–47.

Other Internet Resources

Copyright © 2024 by
Jouko Väänänen<jouko.vaananen@helsinki.fi>

Open access to the SEP is made possible by a world-wide funding initiative.
The Encyclopedia Now Needs Your Support
Please Read How You Can Help Keep the Encyclopedia Free

Browse

About

Support SEP

Mirror Sites

View this site from another server:

USA (Main Site)Philosophy, Stanford University

The Stanford Encyclopedia of Philosophy iscopyright © 2024 byThe Metaphysics Research Lab, Department of Philosophy, Stanford University

Library of Congress Catalog Data: ISSN 1095-5054


[8]ページ先頭

©2009-2025 Movatter.jp