Keywords

2010 Mathematics Subject Classification

1 Truth and the Foundations of Mathematics

Theories of comprehension and satisfaction are closely intertwined. To say that a is an element of the class \(\{x:\, \varphi (x)\}\) for a formula \(\varphi ( x )\) seems tantamount to saying that the formula \(\varphi (x)\) is satisfied by a or that the formula \(\varphi ( x )\) is true of a. Using this observation, one can reduce class theories to theories of satisfaction or those of truth: Quantification over classes is translated to quantification over formulae or propositional functions. The basic strategy can be traced back at least to Russell [51] (see Schindler [53]).

In the good old days of logical positivism, however, the concept truth was considered with suspicion. It had been stained by too many dodgy philosophical theories. So the reduction of mathematical theories to theories of satisfaction or truth didn’t look too attractive. Why should one want to replace a respectable mathematical theory with a ‘philosophical’ theory of truth or satisfaction? As foundational concepts membership and classes seemed much more suitable than truth and satisfaction.

More desirable was a reduction in the other direction, that is, the reduction of truth and satisfaction to a theory of classes. Consequently many philosophers were much more interested in Tarski’s [58] definition of truth in a type theory over a theory of syntax. This reduction vindicated truth as a respectable notion for philosophers like Popper (see Leitgeb [43]).Footnote 1 Tarski added further reasons not to treat truth or satisfaction as a primitive notions to the neopositivist qualms: He thought that either the theory of truth will lack ‘deductive power’, if truth is axiomatized by the typed T-sentences, or the axioms for truth will have an ‘accidental character’ (see Halbach [33, Sect. 7]). Hence mathematical theories with comprehension axioms retained their conceptual priority over axiomatic theories of truth in foundational discussion. Of course some work on truth was done, but truth wasn’t much used in the foundations of mathematics.

It took a long time for truth to recover from the neopositivist qualms and Tarski’s verdict that any axiomatization of truth is either weak or arbitrary. Even philosophers beyond any suspicion of neopositivist convictions have endorsed the conceptual priority of some class theory over a theory of truth or satisfaction. For instance, the type-free truth theories advanced by Kripke [42], Gupta [30] or Herzberger [35] follow the pattern of Tarski’s theory in a crucial aspect. On their accounts set theory is employed as the basic framework; then a semantics for a language with the truth predicate is defined within set theory. The concept of truth these authors are interested in is more worrisome than Tarski’s, because type-free truth is prone to paradox. Consequently it is not a surprise that these authors rely on set theory as ultimate framework and show that semantics for languages with a type-free truth predicate can be developed within set theory. Again a theory of classes serves as the bedrock foundational framework on which a theory of truth can rest.

Truth, however, conceived as a primitive, undefined notion, does have a potential for use in the foundations of mathematics and the formal sciences. One use where the need for a truth predicate is obvious are the proof-theoretic reflection principles. As Gödel [29] had shown, a system \(\mathsf {S}\) cannot prove even very weak consequences of the soundness of the theory. In particular, \(\mathsf {S}\) cannot prove the consistency of \(\mathsf {S}\), if the latter is expressed in a natural way. However, by endorsing or accepting a system \(\mathsf {S}\), one is also committed to the soundness of \(\mathsf {S}\) and thus to all consequences of its soundness such as the consistency of \(\mathsf {S}\). One can try to add statements expressing the soundness of \(\mathsf {S}\) to the system \(\mathsf {S}\) in order to obtain a system that features commitments implicit in the acceptance of \(\mathsf {S}\) as explicit theorems. As a very early reaction to Gödel’s incompleteness theorems, Turing [59] had tried to add such principles to \(\mathsf {S}\) and to iterate this procedure even along transfinite ordinals.

Kreisel and Lévy [41, p. 98] describe a reflection principle for a system \(\mathsf {S}\) as ‘the formal statement stating the soundness of \(\mathsf {S}\)’. As a soundness principle the consistency statement is very weak. More powerful are the local reflection principle \({\text {Bew}}_{\mathsf {S}}(\ulcorner \varphi \urcorner )\rightarrow \varphi \) and the uniform reflection principle \(\forall x\, ({\text {Bew}}_{\mathsf {S}} (\ulcorner \varphi (\dot{x})\urcorner )\rightarrow \varphi (x ))\). But even these schemata fall short of expressing full soundness.

Full soundness is the statement that all (closed) theorems of \(\mathsf {S}\) are true. Kreisel and Lévy [41, p. 98] call a version of this the \({global\, reflection\, principle}\). They write:

Literally speaking, the intended reflection principle cannot be formulated in \(\mathsf {S}\) itself by means of a single statement. This would require a truth definition \(T_\mathsf {S}\) [...]

They go on to point out that such a truth definition doesn’t exist because of Tarski’s theorem on the undefinability of truth. Hence, under the usual assumptions, the ‘intended reflection’ cannot be expressed in the language of \(\mathsf {S}\). It is also hardly an option to pass from \(\mathsf {S}\) to another stronger system containing \(\mathsf {S}\) with the resources for defining a predicate \(T_{\mathsf {S}}\), because it is only soundness that is to be added to the system and not, for instance, more comprehension axioms or other axioms sufficient for defining \(T_{\mathsf {S}}\).

In order to avoid the addition of new mathematical resources for the definition of \(T_{\mathsf {S}}\), one can add a primitive symbol for truth in \(\mathsf {S}\). This approach goes directly against the neopositivist qualms against truth. Thus it doesn’t come as a surprise that initially a primitive formal truth predicate wasn’t used in the discussion about reflection principles, reflective closure, recursive progressions and related topics. Turing’s early work was continued without the use of a truth predicate in the object language. Of course, Feferman is the single most influential author in the wake of Turing’s approach. In particular, Feferman [9] contained amazing results on the iterated addition of proof-theoretical reflection principles in the sense of Kreisel and Lévy [41] to Peano arithmetic. These progressions of theories were all formulated in the language of arithmetic without any additions.

In the 1970s more logicians were less restrained by worries about truth and started working on formal truth theories with truth as a predicate in the object language. They also became bolder and turned to type-free notions of truth. They explored other ways of solving or blocking the paradoxes than Tarski’s restrictive method involving the distinction between an object and a metalanguage. Especially after the publication of Kripke [42], more philosophers and logicians turned their attention to truth. Feferman recognized their potential for adding strong soundness claims and making explicit assumptions implicit in the acceptance of a theory.

2 Ramified Truth and Reflection

In 1979 Feferman gave a talk entitled Gödel’s incompleteness theorems and the reflective closure of theories at the meeting of the Association of Symbolic Logic in San Diego in 1979 ([45] and [17, p. 3]). This paper, which was to become the paper [17], had been circulated for quite some time and other logicians including Reinhardt [50] and Cantini [5] had published on Feferman’s ideas. Feferman’s talk and the subsequent papers by him and others mark the return of truth as an undefined, primitive notion of foundational significance. Let \(\mathcal {L}_{\mathsf {PA}}\) be the language of first-order arithmetic.

The starting point is the ‘compositional’ axioms for truth. Given a system satisfying certain natural conditions, axioms for a truth predicate \(T\) can be added that correspond to the clauses in Tarski’s definition of satisfaction. When one works in a theory that can encode every object as a closed term within it one can dispense with satisfaction and use a unary truth predicate – at least if the theory satisfies certain further assumptions. Feferman [17] gave fairly general account that applies to a variety of theories such as set theory for instance (see Fujimoto [28]). This generality is significant and not just a trivial extension of the special case of Peano arithmetic as starting theory.

However, here we use Peano arithmetic as our base theory. This will allow us to keep the presentation sufficiently simple. Let \({\mathcal {L}_T}\) be \({\mathcal {L}_{\mathsf {PA}}} \cup {T}\). For \(\mathsf {PA}\) the compositional axioms can be chosen as follows:

  1. T1

    \(\, \forall s \, \forall t \, \big ( T ( s \mathop {=}\limits _{\cdot } t)\leftrightarrow {s}^\circ \! =\!{t}^\circ )\big ) \) and similarly for other predicates other than \(=\), except for the special predicate T

  2. T2

    \(\, \forall x\, \big ({{\mathrm{{\text {Sent}}_{\mathsf {PA}}}}}(x)\rightarrow (T (\mathop {\lnot }\limits _{\cdot }x)\leftrightarrow \lnot T (x))\big )\)

  3. T3

    \(\, \forall x\,\forall y\, \big ( {{\mathrm{{\text {Sent}}_{\mathsf {PA}}}}}(x\mathop {\wedge }\limits _{\cdot }y)\rightarrow (T (x\mathop {\wedge }\limits _{\cdot }y)\leftrightarrow T (x)\wedge T (y)) \big )\)

  4. T4

    \(\, \forall v\,\forall x\, \big ( {{\mathrm{{\text {Sent}}_{\mathsf {PA}}}}}(\mathop {\forall }\limits _{\cdot }vx)\rightarrow (T (\mathop {\forall }\limits _{\cdot }vx)\leftrightarrow \forall t\, T (x[t/v])) \big )\)

Here and in what follows we use quantifiers \(\forall s\) and \(\forall t\) to range over the codes (or Gödel numbers) of closed \(\mathcal {L}_{\mathsf {PA}}\)-terms; namely, the expression \(\forall t\) is short for \(\forall x\, (\text {ClTerm}(x)\rightarrow \ldots )\), where \( \text {ClTerm}(x)\) represents the set of the codes of closed \(\mathcal {L}_{\mathsf {PA}}\)-terms. The symbol \({}^\circ \) is a representation in \(\mathsf {PA}\) of a recursive function that takes a code of a closed \(\mathcal {L}_{\mathsf {PA}}\)-term and returns its value in the standard model; e.g., \({\ulcorner 0 + 0\urcorner }^\circ = 0\). Hence the axiom T1 expresses that a closed equation \(s \! =\!t\) is true iff the values of the closed terms s and t coincide. There won’t be a function symbol \({}^\circ \) in the language, but the function can be expressed using suitable formulae. The formula \({{\mathrm{{\text {Sent}}_{\mathsf {PA}}}}}(x)\) represents that x is a code of a sentence of the language of \(\mathsf {PA}\). The symbol \(\mathop {\lnot }\limits _{\cdot }\) is a representation in \(\mathsf {PA}\) of a recursive function that takes a code of a sentence and returns the code of its negation; e.g., \(\mathop {\lnot }\limits _{\cdot }\, \ulcorner 0 = 0\urcorner = \ulcorner 0 \ne 0\urcorner \). Hence the axiom T2 expresses that the negation of a sentence of \(\mathsf {PA}\) is true iff the sentence is not true. The symbol \(\mathop {\wedge }\limits _{\cdot }\) is a representation in \(\mathsf {PA}\) of a recursive function that takes two codes of sentences and returns the code of their conjunction; e.g., \(\ulcorner 0 = 0\urcorner \, \mathop {\wedge }\limits _{\cdot }\, \ulcorner 0 \ne 0\urcorner = \ulcorner 0 = 0 \wedge 0 \ne 0\urcorner \). Hence the axiom T3 expresses that the conjunction of two sentences of \(\mathsf {PA}\) is true iff both of the conjuncts are true. The symbol \(\mathop {\forall }\limits _{\cdot }\) is a representation in \(\mathsf {PA}\) of a recursive function that takes a code of a variable and a code of a formula and returns the universal quantification of the formula with respect to the variable; e.g., \(\mathop {\forall }\limits _{\cdot }\, ( \ulcorner v\urcorner , \ulcorner \varphi \urcorner ) = \ulcorner \forall v \varphi \urcorner \) for a variable v and a formula \(\varphi \). The operation x[y / z] yields the code of the result of substituting a term encoded by y for a variable encoded by z in a formula encoded by x; e.g., \(\ulcorner \varphi ( v )\urcorner [ \ulcorner t\urcorner / \ulcorner v\urcorner ] = \ulcorner \varphi ( t )\urcorner \) for a formula \(\varphi \), a term t, and a variable v. Hence axiom T4 expresses that a universally quantified sentence of \(\mathsf {PA}\) is true iff all its substitution instances (with closed terms) are true. A more detailed explanation of the notation can be found in Halbach [33]. But we hope that our notation should be largely self-explanatory. There are also some minor deviations from Feferman’s definition of this theory in [17, p. 14].

The axioms T1–T4 are adjoined to those of Peano arithmetic. Crucially, the induction schema is expanded to the new language with the truth predicate. The resulting theory \(\mathsf {T(\mathsf {PA})}\) or also called \(\mathsf {CT}\) (for ‘compositional truth’) proves the soundness of Peano arithmetic, that is, it proves the global reflection principle

$$\begin{aligned} \forall x\, \big ({{\mathrm{{\text {Sent}}_{\mathsf {PA}}}}}(x)\wedge {\text {Bew}}_{\mathsf {PA}} (x)\rightarrow T(x)\big ) \end{aligned}$$
(2.1)

Thus the expressive resources needed for stating the soundness of \(\mathsf {PA}\) already imply the soundness of \(\mathsf {PA}\).Footnote 2 The global reflection principle doesn’t have to be added as an additional axiom. As we have discussed, the uniform reflection principle is derived from (2.1) with the help of the axioms T1–T4; furthermore, \(\mathsf {CT}\) is strong enough to derive the iterated reflection principle along any constructive ordinal \(\alpha \) provably well-founded in \(\mathsf {CT}\) in the sense that all the \(\mathcal {L}_{T}\)-instances of transfinite induction along \(\alpha \) are derivable in \(\mathsf {CT}\) .

If the commitment to the soundness of \(\mathsf {PA}\) is implicit in the acceptance of \(\mathsf {PA}\) , then also the resources needed for expressing the soundness claim are implicit in the acceptance of \(\mathsf {PA}\). Thus the acceptance of \(\mathsf {PA}\) commits one to \(\mathsf {CT}\). Thus \(\mathsf {CT}\) makes explicit some implicit commitments of the acceptance of \(\mathsf {PA}\). However, the implicit commitment in the acceptance of \(\mathsf {PA}\) is not exhausted by \(\mathsf {CT}\) and (2.1), and the iteration procedure further continues. For, once \(\mathsf {CT}\) is explicitly accepted, one is then committed to the soundness of \(\mathsf {CT}\) and thus to a truth predicate for \(\mathsf {CT}\), which is needed to express the soundness of \(\mathsf {CT}\). To this end one can add a further truth predicate \(T_1\) that applies to all sentences formulated in the language of arithmetic expanded by \(T\). The new predicate \(T_1\) is then axiomatized in the same way as \(T\) except that \(T\) is treated as one of the non-special predicate symbols in T1. Moreover in T2–T4 quantification over sentences of the arithmetical language is replaced with quantification over sentences of the arithmetical language with \(T\). This procedure can be iterated and an axiomatization of Tarski’s hierarchy of languages is obtained. The exact specification of the procedure requires some more detail; but a general recipe for the transition from one level to the next can be specified. The procedure fits Feferman’s [9, p. 274] description of a reflection principle:

By a reflection principle we understand a description of a procedure for adding to any set of axioms A certain new axioms whose validity follow from the validity of the axioms A and which formally express, within the language of A, evident consequences of the assumption that all the theorems of A are valid.

This characterization of reflection principles precedes Feferman’s work on truth by more than one and a half decades.

The addition of truth predicates and accompanying axioms can be iterated into the transfinite. A truth predicate \(T_{\lambda }\) at a limit level \(\lambda \) is axiomatized in the same way as the truth predicates at successor level with the exception that an axiom is added that says that a sentence is true iff it is true at one of the previous levels. For details see Halbach [33, Sect. 9.1] and Fujimoto [28].

The need for an ordinal notation system is obvious: The truth predicates need to be indexed and all levels of the hierarchy need to be axiomatized. Technically it is no problem to define theories up to any recursive ordinal. To this end one can pick a path through Kleene’s \(\mathcal {O}\). However, this would betray the original motivation for considering these iterated theories of truth: their purpose is to make explicit assumptions implicit in the acceptance of \(\mathsf {PA}\) . Using our general mathematical machinery we can prove that there is a well-founded ordering of natural numbers whose order-type is any ordinal below the first non recursive ordinal \(\omega _1^{\text {CK}}\). But this theorem is not implicit in the acceptance of \(\mathsf {PA}\) in any way. In particular, \(\mathsf {PA}\) doesn’t prove that these orderings are well-founded. Kreisel [40] suggested to employ autonomous iterations in situations of this kind. That is, one iterates a procedure of this kind in unfolding the implicit commitments in accepting a theory only if the theory in question can prove transfinite induction up to that ordinal level.

Since \(\mathsf {PA}\) proves transfinite induction for any ordinal up to \(\varepsilon _0\), the truth theories are iterated up to that point. This new theory, however, proves transfinite induction for longer wellorderings. Hence the truth theories are iterated even further until a point is reached where the hierarchy of truth theories is iterated to a point \(\Gamma _0\) that proves transfinite induction for all ordinals smaller than \(\Gamma _0\). This ordinal \(\Gamma _0\) is the so-called Feferman–Schütte ordinal that had figured prominently in Feferman’s earlier work [10] on predicativity.

The iterated truth theories very much resemble the systems of predicative analysis, which had been studied thoroughly by Feferman [10] and Schütte [54] in the 1960s. Thus, in a sense, the results on iterated truth theories are formally not extremely exciting. Perhaps this is the reason why they figure less prominently in the published paper [17] than in the draft version [16].

From a foundational point of view, however, we think that the iterated truth theories are significant. They are a very convincing way of carrying out the programme of determining the reflective closure of \(\mathsf {PA}\) , that is, of characterizing the theory that makes explicit what is implicit in the acceptance of \(\mathsf {PA}\).

The formulation of the systems of iterated truth is technically awkward. The specification of the language already requires an ordinal notation system. Then the motivation of the terminal ordinal \(\varepsilon _0\) or \(\Gamma _0\) relies on some deeper results. Moreover, it is highly specific to \(\mathsf {PA}\).

Feferman has made various attempts at characterizing the reflective closure of theories in a more elegant way. The reasons for seeking a more succinct characterization are not only of an aesthetic nature. A method of defining the reflective closure of a theory that is less reliant on ordinal notation systems and an explicit appeal to proof-theoretic techniques and notions, should also be more generally applicable; moreover, it would also be philosophically less prone to the objection that it depends on arbitrary stipulation; a more elegant system would depend on a ‘natural’ ordinal notation system and arithmetization.

Feferman has tried various methods for characterizing the reflective closure of a theory. The various approaches should not be seen so much as competing but rather as different characterization of the same concept. The situation is similar to that in recursion theory: The different characterizations of computability do not exclude each another, rather their equivalence assures us that we have found a stable concept.

3 Kripke–Feferman

The first method of characterizing the reflective closure of \(\mathsf {PA}\) is fairly close to the iterated truth theories. But it shuns already the need for an ordinal notation system. Somewhat metaphorically speaking, the ordinals emerge from the theory itself and are not imposed on it from the outside.

The theory has been dubbed the Kripke–Feferman theory or \(\mathsf {KF}\) for short. Presumably Reinhardt was the first to publish on the theory in [49, 50] and by the time Feferman published his paper [17], the label \(\mathsf {KF}\) had already been established. Unfortunately, since no authoritative version had been published by Feferman, different authors formulated \(\mathsf {KF}\) in slightly different ways. Here we try to stick to the formulation chosen by Feferman; but we use a slightly different notation.

Feferman [17] formulated \(\mathsf {KF}\) in the language of arithmetic augmented with two new unary predicates \(T\) and \(F\) for truth and falsity. Let \(\mathcal {L}_{\mathsf {KF}}:= \mathcal {L}_{\mathsf {PA}}\cup \{ T, F\}\) denotes the language of \(\mathsf {KF}\). As the axiom K4 below indicates, the falsity predicate \(F\) is actually not necessary and could be understood as defined notion, because the falsity of a sentence coincides with the truth of its negation. Hence for the sake of simplicity, we will identify \(\mathcal {L}_{\mathsf {KF}}\) with \(\mathcal {L}_{T}\). The axioms of \(\mathsf {KF}\) comprise those of \(\mathsf {PA}\) and the following truth-theoretic axioms:

  1. K1

    \(\forall s\, \forall t\, \big (\big ( T ( s \mathop {=} \limits _{\cdot } t)\leftrightarrow {s}^\circ \! =\!{t}^\circ ) \big ) \wedge \big ( F( s \mathop {=} \limits _{\cdot } t)\leftrightarrow {s}^\circ \! \ne \! {t}^\circ )\big ) \), and similarly for other predicates other than \(=\), except for the special predicate T;

  2. K2

    \(\forall s \, \big ( ( T (\mathop {T}\limits _{\cdot }s) \leftrightarrow T ({s}^\circ ) ) \wedge ( F(\mathop {T}\limits _{\cdot }s) \leftrightarrow F({s}^\circ ) ) \big )\);

  3. K3

    \(\forall s \, \big ( ( T (\mathop {F}\limits _{\cdot }s) \leftrightarrow F ({s}^\circ ) ) \wedge ( F(\mathop {F}\limits _{\cdot }s) \leftrightarrow T ({s}^\circ ) ) \big )\);

  4. K4

    \(\forall x \, \big ( {{\mathrm{{\text {Sent}}_{\mathsf {KF}}}}}( x ) \rightarrow ( T (\mathop {\lnot }\limits _{\cdot }x) \leftrightarrow F (x) ) \wedge ( F(\mathop {\lnot }\limits _{\cdot }x )\leftrightarrow T (x) ) \big )\);

  5. K5

    \(\forall x\,\forall y\, \big ( {{\mathrm{{\text {Sent}}_{\mathsf {KF}}}}}( x \mathop {\wedge }\limits _{\cdot }y )\rightarrow (T (x \mathop {\wedge }\limits _{\cdot }y)\leftrightarrow T (x) \wedge T (y)) \wedge (F(x \mathop {\wedge }\limits _{\cdot }y)\leftrightarrow F(x) \vee F(y)) \big )\);

  6. K6

    \( \forall v\,\forall x\, \big ( {{\mathrm{{\text {Sent}}_{\mathsf {KF}}}}}(\mathop {\forall }\limits _{\cdot }vx)\rightarrow (T (\mathop {\forall }\limits _{\cdot }vx)\leftrightarrow \forall t\, T (x[t/v])) \wedge (F(\mathop {\forall }\limits _{\cdot }vx)\leftrightarrow \exists t\, F(x[t/v])) \big )\).

The axiom K1, exactly like T1, says that a closed equation \(s \! =\!t\) is true (false) iff the value of the terms s and t agree (disagree, resp.). The formula \({{\mathrm{{\text {Sent}}_{\mathsf {KF}}}}}(x)\) expresses that x is a sentence of the language of (i.e. \(\mathcal {L}_T\)); hence the axiom K5 and K6 express essentially the same compositional axioms as T3 and T4 but extended to sentences of the larger language \(\mathcal {L}_{T}\). The axioms K2 and K3 describes the iterative self-applicative characteristic of the truth and falsity predicates in \(\mathsf {KF}\); K2 says that it is true (false) that a sentence is true iff the sentence is true (false, resp.) and K3 says its dual. Finally, K4 defines the falsity of a sentence to be the truth of its negation. There are different ways to motivate the axioms of \(\mathsf {KF}\) and Halbach [33] develops a fuller picture. \(\mathsf {KF}\) can be seen as a generalization of \(\mathsf {CT}\) , which is a subtheory of \(\mathsf {KF}\), or, more naturally, as a generalization of a theory \(\mathsf {PT}\) of a positive inductive definition of truth and falsity [33, Sect. 8.7].

Although \(\mathsf {KF}\) derives transfinitely iterated uniform reflection principles for its base theory (i.e., \(\mathsf {PA}\) in the current setting), \(\mathsf {KF}\) cannot derive its own soundness due to Gödel’s incompleteness theorem: namely, there is an \(\mathcal {L}_{T}\)-sentence \(\varphi \) (e.g., \(0 = 1\)) such that

$$ \mathsf {KF}\not \vdash \bigl ( {\text {Bew}}_{\mathsf {KF}}( \ulcorner \varphi \urcorner ) \rightarrow \varphi \bigr ). $$

Hence \(\mathsf {KF}\) doesn’t derive the global reflection principle for itself:

$$ \mathsf {KF}\not \vdash \forall x\, \bigl ({{\mathrm{{\text {Sent}}_{\mathsf {KF}}}}}(x)\wedge {\text {Bew}}_{\mathsf {KF}}(x) \rightarrow T(x) \bigr ). $$

This fact may suggest one to iterate \(\mathsf {KF}\)-truth as in the case of \(\mathsf {RT}_{\alpha }\); c.f., [27]. However, Feferman [17] gave an argument against such iteration, and thereby explain why \(\mathsf {KF}\) (or \( Ref ^{*} ( \mathsf {PA}(P) )\) defined below in Sect. 3.2.3) is to be called reflective closure, that is, why \(\mathsf {KF}\) is to be seen as exhausting ‘what notions and principles one ought to accept if one accepts the basic notions and principles of the theory’ [18, p.205]. Iterating \(\mathsf {KF}\) would mean to adopt principles that go beyond what is implicit in the acceptance of the base theory, that is, \(\mathsf {PA}\) in the case considered here.

We sketch Feferman’s argument against using iterations of \(\mathsf {KF}\) in order to define the reflective closure of \(\mathsf {PA}\). If one were to add axioms for a further truth predicate \(T'\), that is, a truth predicate for the language of \(\mathsf {KF}\) including the truth predicate \(T\), one would specify axioms for \(T'\) analogous to those for \(T\). To this end, one would now quantify over sentences in the full language with \(T'\) in axioms K4–K6; moreover, one would treat \(T\) as just like another predicate of the base language and therefore add the following axiom in analogy to K1 for the predicate \(T\) [17, p.40]:

$$\begin{aligned} \forall s\, \bigl ( ( T' (\mathop {T}\limits _{\cdot }s)\leftrightarrow T({s}^\circ )) \wedge ( F'( \mathop {T}\limits _{\cdot }s)\leftrightarrow \lnot T({s}^\circ )) \bigr ) \end{aligned}$$
(3.1)

From the logical truth \(\forall s\, (T({s}^\circ )\vee \lnot T({s}^\circ )\) and this axiom, we can derive the following ‘totality’ claim:

$$\begin{aligned} \forall s\, \bigl ( T' (\mathop {T}\limits _{\cdot }s) \vee F'( \mathop {T}\limits _{\cdot }s)\bigr ) \end{aligned}$$
(3.2)

Therefore, by endorsing the new axiom (3.1), we would treat \(T\) as a ‘total’ or ‘determined’ predicate, just like the predicates of the base language. Feferman [17, p.40] thereby concludes that ‘when iterating reflective closure we thus vitiate the informal idea behind the use of partial predicates of truth and falsity.’ Hence, for Feferman, \(\mathsf {KF}\) is closed under the process of making explicit what are implicitly accepted in accepting the basic notions and principles.

\(\mathsf {KF}\) is a quite rich theory. Indeed, it is intimately related to an approach to predicativity, to the theme of unfolding (see Feferman and Strahm [20], Feferman and Strahm [21]), and it can be regarded as characteristic of a fruitful interaction between the logical development of non-extensional concepts (classification, operation) and semantical investigations. Further, it has a sort of mathematical appeal: it hinges upon well-known lattice theoretical facts (Knaster–Tarski theorem), and it naturally implies the existence of fixed points of arithmetically definable monotone operators. One can show that \(\mathsf {KF}\) and the standard fixed point theory \(\widehat{\mathsf {ID}}_{1}\) (see Feferman [14]) are mutually interpretable, thus reducing the classification of proof theoretic strength of \(\mathsf {KF}\) to that of \(\widehat{\mathsf {ID}}_{1}\) . \(\widehat{\mathsf {ID}}_{1}\) leads towards the foundations of intuitionistic type theory and predicatively reducible subsystems of analysis. Below we summarize a few important facts and results on \(\mathsf {KF}\).

3.1 Inner Logic and Outer Logic

The theory \(\mathsf {KF}\) is sometimes criticized for yielding a discrepancy of its outer and inner logic. Let us call \(\{ \varphi \mid \varphi \text { is an } \mathcal {L}_{T}\text {--sentence } \text {and }\mathsf {KF}\vdash \varphi \}\) the outer theory of \(\mathsf {KF}\) and \(\{ \varphi \mid \varphi \text { is an } \mathcal {L}_{T} \text { -sentence } \text {and } \mathsf {KF}\vdash T(\ulcorner \varphi \urcorner ) \}\) the inner theory of \(\mathsf {KF}\), and also call the logic governing the former the outer logic of \(\mathsf {KF}\) and the logic governing the latter the inner logic of \(\mathsf {KF}\).Footnote 3 The inner logic of \(\mathsf {KF}\) is strong Kleene logic whereas its outer logic is classical logic; also the inner theory of \(\mathsf {KF}\) fails to coincide with the outer theory of \(\mathsf {KF}\). Hence, \(\mathsf {KF}\) does not meet one of the desiderata that Leitgeb [44] suggests for a theory of truth.

In order to develop a theory that is an axiomatization of Kripke’s theory with strong Kleene logic that avoids this discrepancy, Halbach and Horsten [34] presented a system based on strong Kleene logic. The resulting theory \(\mathsf {PKF}\) has exactly the same inner theories and outer theories. Halbach and Horsten [34] showed that \(\mathsf {PKF}\) proves significantly less arithmetical statements than the classical system \(\mathsf {KF}\). They suggested that thus \(\mathsf {KF}\) should not be seen as a formal device for generating theorems in strong Kleene logic: the use of the classical outer logic is indispensable for proving certain arithmetical theorems. Hence it is unlikely that a purely ‘instrumentalist’ interpretation of \(\mathsf {KF}\) in the spirit of Reinhardt [50] is a viable option.

The difference between inner and outer logic is adumbrated in the Sect. 6.1.2 of Feferman [17, p.40] on the informal interpretation of partial-self-applicable truth predicate. There, considering the question as to whether Kripke’s [42] construction corresponds to a (more or less) clear informal notion, Feferman finds it ‘reasonably convincing’ that in the formulation of \(\mathsf {KF}\), \(T(\ulcorner \varphi \urcorner )\) expresses that \(\varphi \) is a grounded truth, that is, that the denotation of \(T\) is given by Kripke’s least fixed-point construction, understood in its full generality.Footnote 4 As a consequence, the analogy with partial computable predicates becomes helpful: Truth should be relative to given rules of computation or inductive rules, and ought to be distinguished from our everyday informal notion of truth with which classical logic is justified.

On the formal side, this leads to define for an \(\mathcal {L}_{T}\)-sentence \(\varphi \) to have a determined truth-value, when \(\mathsf {KF}\vdash T(\ulcorner \varphi \urcorner ) \leftrightarrow \lnot F(\ulcorner \varphi \urcorner )\) holds, and to choose D to be the set of all sentences with a determined truth-value, that is the set of all sentences that are true or false, but not both (see the next subsection for a formal elaboration). This set D is interpreted to apply to those sentences whose truth-value is determinable by the rules of internal truth and falsity, while the complement of D is interpreted to apply to those whose truth-value is not determinable by those rules, but still definite in the sense that classical logic is justified with it. Hence, for instance, the statement \(\lambda \vee \lnot \lambda \) is not an internal truth but true in our informal and, say, naïve platonistic sense.

Those sentences with determined truth-value in this sense enjoy some nice properties. First, the laws of classical logic are provably true for all determined \(\mathcal {L}_{T}\)-sentencesFootnote 5: namely, if \(\varphi _{1}, \ldots , \varphi _{k}\) have determined truth-value, then each logical axiom of classical logic involving these sentences is provably true in \(\mathsf {KF}\). Second, more importantly, the T-schema holds when restricted to those sentences even with the untyped truth predicate of \(\mathsf {KF}\): namely, it holds that, for all \(\mathcal {L}_{T}\)-sentence \(\varphi \), \(\mathsf {KF}\vdash D ( \ulcorner \varphi \urcorner ) \rightarrow \bigl ( T(\ulcorner \varphi \urcorner ) \leftrightarrow \varphi \bigr )\).

3.2 Kripke–Feferman: The Proof Theoretic Side

The main result which is proven in Feferman’s paper on reflective closure is the following:

Theorem 1

$$\begin{aligned} \mathsf {KF}\equiv (\Pi ^0_1{\text {-CA}})_{< \varepsilon _0} \end{aligned}$$
(3.3)

In general \((\Pi ^0_1\text {-CA})_{< \alpha }\) is the theory of iterated jump up to any \(\beta <\alpha \).Footnote 6

Let us outline the proof of the theorem, possibly following different routes.

3.2.1 Lower Bound

We exploit the foundational side of truth, as a way of interpreting non-extensional predicate application, in order to show the main result of Feferman [17].

In order to enhance readability, we adopt the following shortenings:

  1. (i)

    a(x) stands for the the term representing the operation of substituting the first free variable of the expression encoded by a with the xth-numeral.

  2. (ii)

    \(D(a):= {{\mathrm{{\text {Sent}}_{\mathsf {KF}}}}}(a)\wedge ( T(a)\vee T(\mathop {\lnot }\limits _{\cdot }a))\wedge \lnot ( T(a)\wedge T(\mathop {\lnot }\limits _{\cdot }a))\); D(a) means ‘a is determined’;

  3. (iii)

    non-extensional membership: \(x\in a := T(a(x))\wedge \lnot T(\mathop {\lnot }\limits _{\cdot }a(x)))\); non-extensional co-membership: \(x \bar{\in }a := T(\mathop {\lnot }\limits _{\cdot }a(x))\wedge \lnot T(a(x)) \)

  4. (iv)

    the class(ification) predicate \(Cl(a):= \forall x D(a(x))\).

Then it is easily seen that the language \(\mathcal {L}_2\) of second order arithmetic can be regarded as a sublanguage of \(\mathcal {L}_{T}\).

In order to relate \(\mathsf {KF}\) with standard subsystems of second order arithmetic, we need a few additional definitions:

  1. (i)

    A formula \(\varphi (x, \vec {z})\) of \(\mathcal {L}_{T}\) is elementary in \(\vec {z}:= z_1, \ldots z_n\) iff it is inductively generated from atoms of the form \(t=s\), \(t\in z_i\), (\(1\le i \le n\)) by means of \(\wedge \), \(\lnot \), \(\forall u\), with \(u\notin \lbrace z_1, \ldots z_n \rbrace \).

  2. (ii)

    We say that Cl is closed under elementary comprehension iff for every \(\varphi (x, \vec {z})\) elementary in \(\vec {z}\), then there exists a primitive recursive term \(s_\varphi (\vec {z})\) such that, provably in \(\mathsf {KF}\),

    $$\begin{aligned} Cl(\vec {z})\rightarrow Cl(s_\varphi (\vec {z}))\wedge \forall x ( \varphi (x, \vec {z})\leftrightarrow x\in s_\varphi (\vec {z})). \end{aligned}$$
    (3.4)
  3. (iii)

    A family g of classes indexed by a class a is simply an index g of a partial recursive function \(\lambda n.\lbrace g\rbrace (n)\),Footnote 7 such that \( \forall y(y\in a \rightarrow Cl( \lbrace g\rbrace (y))\).

  4. (iv)

    Finally, we say that Cl is closed under join iff, whenever f is a family of classes indexed by a class a, there exists a term j(af) whose elements are exactly those ordered pairs (uv) such that that \(u\in a\) and \(v\in \lbrace f \rbrace (u)\).

Lemma 1

The collection of classes is closed – provably in \(\mathsf {KF}\)- under elementary comprehension and join.Footnote 8

Proposition 1

$$\begin{aligned} (\Pi ^0_1\text {-CA})_{<\varepsilon _0}\le \mathsf {KF}\end{aligned}$$
(3.5)

Proof

First of all, \(\mathsf {KF}\) proves, for each \(\alpha < \varepsilon _0\), the transfinite induction scheme \(TI(\varphi , <\alpha )\), for arbitrary formulas \(\varphi \) in the full language. This is well-known (Gentzen–Schütte–Feferman) and it follows insofar as \(\mathsf {KF}\) has the full number theoretic induction schema (see Lemma 4.3.2 in Feferman [17]). Then apply \(TI(\varphi , <\alpha )\) in conjunction with the previous lemma. \(\square \)

The fact that the jump hierarchy is available in \(\mathsf {KF}\) up to any \(\alpha < \varepsilon _0\) is crucial in order to carry out a wellordering proof for each initial segment of the standard wellordering of \(\phi \varepsilon _0 0\) (see Feferman [17], appendix and Feferman [11]).

Lastly, it is not difficult to prove that \(\mathsf {KF}\) can directly interpret the fixed point theory \(\widehat{\mathsf {ID}}_{1}\) (see Cantini [5, Sect. 3.11] and Halbach [33, Sect. 19.5]).

3.2.2 Upper Bound

We sketch two strategies for proving the upper bound direction of the theorem.

The first route: Kripke–Feferman as a fixed point theory. This route is essentially Feferman’s route in Feferman [17] and consists of a reduction of \(\mathsf {KF}\) to classical subsystems of known strength.

The theory \(\widehat{\mathsf {ID}}_{1}\) contains– besides the usual axioms for Peano arithmetic and induction schema for the whole language– fixed point axioms FP asserting the existence of fixed points \(I_\varphi \) for arbitrary elementary positive operators \(\varphi (x, P)\)Footnote 9

$$\begin{aligned} \forall x (\varphi (x,I_\varphi )\leftrightarrow I_\varphi (x)) \end{aligned}$$
(3.6)

but \(\widehat{\mathsf {ID}}_{1}\) has no minimality schema.

By Aczel [2] it is known that \(\widehat{\mathsf {ID}}_{1}\) has the same arithmetical theorems as \(\mathsf {PA}\) with the schema of transfinite induction for each initial segment of the canonical primitive recursive wellordering of type \(\phi \varepsilon _0 0\). The idea is to show that there are (not necessarily minimal) \(\Sigma ^1_1\)-solutions to the fixed point Eq. (3.6) by standard diagonalization, provably in the subsystem \(\mathsf {\Sigma ^1_1{\text{- }AC_0}}\) .Footnote 10 The argument works in second order arithmetic with arithmetical comprehension except that \(\mathsf {\Sigma ^1_1{\text{- }AC_0}}\) is also required in order to show that the result of replacing the parameter P by means of a \(\Sigma ^1_1\)-predicate in a positive elementary operator can be still made equivalent to a \(\Sigma ^1_1\)-formula (see Feferman [17], 4.2). The argument also holds when induction is restricted to arithmetical formulae. For any system \(\mathsf {S}\), let \(\mathsf {S}\!\!\upharpoonright \) be the system with the induction axioms restricted to the language of arithmetic.

Theorem 2

  1. (i)

    \(\mathsf {KF}\le \mathsf {\Sigma ^1_1{\text{- }AC_0}}\)

  2. (ii)

    \(\mathsf {KF}\!\!\upharpoonright \le \mathsf {\Sigma ^1_1{\text{- }AC_0}}\!\!\upharpoonright \)

This is sufficient for calibrating \(\mathsf {KF}\) and its subsystems.

The second route. The previous proof is simple but it has a disadvantage: it cannot be adapted to deal with truth consistency (or minimality), i.e. if we want to consider \(\mathsf {KF}\) plus \(\mathsf {CONS}\):

$$\begin{aligned} \forall x({{\mathrm{{\text {Sent}}_{\mathsf {KF}}}}}(x)\rightarrow \lnot T(x)\vee \lnot F(x)) \qquad {(\mathsf{{CONS}})} \end{aligned}$$

Of course, Axiom CONS rules out truth value gluts. That is, the axiom excludes models where the extension and the antiextension of the truth predicate overlap and sentences are simultaneously true and false. This restriction to consistent fixed points is in line with Kripke’s [42] original account.

In order to analyze \(\mathsf {KF}\) plus \(\mathsf {CONS}\) , we can easily apply methods from predicative proof theory. Thus we devise a suitable sequent calculus version of \(\mathsf {KF}\), \(\mathsf {KF}\!\!\upharpoonright \), ..., choosing to rephrase \(\mathsf {KF}\!\!\upharpoonright \) as a fixed point theory with consistency. This means that we can easily find a formula \(\mathcal {T}(x,P)\), formalizing the closure properties of the truth predicate \(T\) (see Halbach [33], p. 281), i.e. such that \(\forall x( \mathcal {T}(x,T)\leftrightarrow T(x))\).

Since we like to formalize \(\mathsf {KF}\) and \(\mathsf {KF}\!\!\upharpoonright \) as Tait calculi, the basic positive atoms have the form: \(t=s\), \(T(t)\), and the negative atoms are obtained by negating the positive ones. An atom is simply a positive or a negative atom and we stipulate that \(\lnot \lnot \varphi :=\varphi \) (\(\varphi \) atom). Formulas are inductively generated from atoms by closing under disjunction, conjunction, unbounded quantification. If \(\varphi \) is an arbitrary formula, \(\lnot \varphi \) is the formula which results from the negation normal form of \(\lnot \varphi \) by erasing each even sequence of occurrences of negation in front of atoms.

Let us expand \(\mathcal {L}_{T}\) with a new predicate symbol P. If \(Q:=P, T\), a formula \(\varphi \) of \(\mathcal {L}_{T}\) is Q-positive (Q-negative) if every occurrence of Q in \(\varphi \) occurs within positive (negative) atoms of the form Q(t) (\(\lnot Q(t)\)). A formula \(\varphi \) is Q-separated if \(\varphi \) is Q-positive or Q-negative. A formula \(\varphi \) is Q-free if Q does not occur in \(\varphi \). A Q-free formula can be regarded as both Q-positive and Q-negative.

Definition 1

The system \(\mathsf {TKF^c}\) consists of:

  • logical axioms of the form

    $$\begin{aligned}&\Gamma ,\lnot \varphi , \varphi \\&\Gamma , \lnot t=s, \varphi [x:=t], \varphi [x:=s] \end{aligned}$$

    where \(\varphi \) is an atom (according to the previous definitions);

  • axioms of the form \(\Gamma , \Delta \) where \(\Delta \) is an e-atom or a finite set of e-atoms; \(\Delta \) formalizes the standard axioms for zero, successor, or the defining equations for the function symbols of \(\mathcal {L}_{T}\);

  • standard logical rules (see Schwichtenberg and Wainer [55]) for introducing \(\wedge \), \(\vee \), \(\forall \), \(\exists \) and the cut rule:

    $$\begin{aligned} \frac{\Gamma , \varphi \quad \Gamma , \lnot \varphi }{\Gamma } \end{aligned}$$
  • T-consistency:

    $$\begin{aligned} \Gamma , \lnot {{\mathrm{{\text {Sent}}_{\mathsf {KF}}}}}(t), \lnot T(t), \lnot T(\mathop {\lnot }\limits _{\cdot }t) \end{aligned}$$
  • induction rule for classifications: if t is an arbitrary term,

    $$\begin{aligned} \frac{\Gamma , Cl(a) \quad \Gamma , 0\in a\quad \Gamma , \forall x(x\in a\rightarrow (x+1)\in a)}{\Gamma , t\in a} \end{aligned}$$
  • \(T\)-closure:

    $$\begin{aligned} \frac{\Gamma , \mathcal {T}(t,T)}{\Gamma , T(t)} \end{aligned}$$
  • \(T\)-soundness:

    $$\begin{aligned} \frac{\Gamma , \lnot \mathcal {T}(t,T)}{\Gamma , \lnot T(t)} \end{aligned}$$

\(\mathsf {TKF}\) is the calculus which is obtained from \(\mathsf {TKF^c}\) by replacing the induction rule for classes by full number-theoretic induction rule, that is, if \(\varphi (x)\) is an arbitrary formula of \(\mathcal {L}_{T}\) an arbitrary term,

$$\begin{aligned} \frac{\Gamma , \varphi (0)\quad \Gamma , \forall x(\varphi (x)\rightarrow \varphi (x+1))}{\Gamma , \varphi (t)} \end{aligned}$$

If \(S:=\mathsf {TKF^c},\mathsf {TKF}\), we can inductively define a derivability relation with explicit tree height and suitable cut complexity, so that the following holds:

Lemma 2

(Partial Cut-elimination) Every \(\mathsf {TKF^c}\)-derivation \(\mathcal {D}\) can be effectively transformed into a \(\mathsf {TKF^c}\)-derivation of the same end-sequent, where cut–formulas are T-separated.

The argument is standard; it essentially depends on the fact that the active formulas in the axioms and in the conclusions of the mathematical inferences (in particular, number theoretic induction) are \(T\)-separated. It is also obvious that the system \(\mathsf {TKF^c}\) (\(\mathsf {TKF}\)) proves the sequents correspondingFootnote 11 to the theorems of \(\mathsf {KF}\!\!\upharpoonright \)(\(\mathsf {KF}\)) plus \(\mathsf {CONS}\) .

Approximating truth by its finite levels. By Lemma 2 it is then possible to approximate truth by its finite levels and hence to eliminate truth from \(\mathcal {L}_{T}\).

Definition 2

Let \(\bot = 0=1\); then

$$\begin{aligned} T^0(t)= & {} 0=1\end{aligned}$$
(3.7)
$$\begin{aligned} T^{m+1}(t)= & {} \mathcal {T}(t, T^m)) \end{aligned}$$
(3.8)

Clearly each formula in the sequence belongs to the language \(\mathcal {L}_{\mathsf {PA}}\).

If \(\varphi \) is any formula in negation normal form, let \(\varphi [m,n]\) be obtained from \(\varphi \) by replacing each atom of the form \(T(t)\) (\(\lnot T(t)\)) by \(T^n(t)\) (\(\lnot T^m(t)\)).

Finally, a derivation \(\mathcal {D}\) of \(\mathsf {TKF^c}\) is quasi-normal provided all cut–formulas in \(\mathcal {D}\) are T-separated.

Theorem 3

Let \(\mathcal {D}\) be a quasi-normal \(\mathsf {TKF^c}\)-derivation of \(\Gamma \) with height k. Then, provably in \(\mathsf {PA}\), for every \(m>0\),Footnote 12 if \(H(m):=m+2^k\):

$$\begin{aligned} \Gamma [m,H(m)] \end{aligned}$$
(3.9)

The proof is by induction on the height k of the given derivation. The soundness of the asymmetric interpretation – with respect to \(\mathsf {PA}\)-provability, relies on the fact that cut formulas are always T-separated and on standard persistence properties.Footnote 13 On the other hand, restriction to classes has the effect that number theoretic induction for \(\mathcal {L}_{\mathsf {PA}}\)-formulas is enough, as a consequence of the fact that for each given m, \(T^m(t)\) is an arithmetical formulaFootnote 14 As to the verification of consistency, it reduces to verify by outer induction on m

$$\begin{aligned} \forall x( {{\mathrm{{\text {Sent}}_{\mathsf {KF}}}}}(x)\rightarrow \lnot T^m(x)\vee \lnot T^m(\mathop {\lnot }\limits _{\cdot }x)) \end{aligned}$$
(3.10)

Since the transform \(\varphi \mapsto \varphi [m,n]\) is the identity function on formulas of \(\mathcal {L}_{\mathsf {PA}}\), we eventually conclude that

Corollary 1

If \(\mathsf {TKF^c}\vdash \varphi \) and \(\varphi \in \mathcal {L}_{\mathsf {PA}}\), then \(\mathsf {PA}\vdash \varphi \).

Theorem 4

  1. (i)

    \(\mathsf {PA}\ \equiv \mathsf {KF}\!\!\upharpoonright \ +\mathsf {CONS}\)

  2. (ii)

    \((\Pi ^0_1\text {-CA})_{< \varepsilon _0}\equiv \mathsf {KF}+\mathsf {CONS}\)

As to \(\mathsf {KF}+\mathsf {CONS}\le (\Pi ^0_1\text {-CA})_{< \varepsilon _0}\): by lifting the previous method to the case of systems with full number theoretic induction. This can be carried out by standard embedding into systems with \(\omega \)-rule (see Schwichtenberg and Wainer [55]).

3.2.3 From Ordinary Reflective Closure to Schematic Reflective Closure

Equation 3.4 means that the ordinary reflective closure of arithmetic is in fact equivalent to predicative analysis of any level up to the first \(\epsilon \)-number \(\varepsilon _0\) i.e. roughly to the fragment of second order arithmetic, which is based upon transfinite iteration up to \(\varepsilon _0\) of the standard arithmetical comprehension

$$ (\exists X)(\forall u)(u \in X\leftrightarrow \varphi (u)), $$

\(\varphi \) being a formula containing only number theoretic quantifiers and possibly free second order variables.

The notion of ordinary reflective closure can then be extended by a suitable substitution rule in order to answer the problem of which schemata can be regarded as implicit in accepting a given list of schematic axioms and rules. The substitution rule has the form

$$\begin{aligned} \frac{\varphi (P)}{\varphi (\hat{x}\psi (x))} \end{aligned}$$

where \(\varphi \) is a formula of \(\mathcal {L}_{\mathsf {PA}}\) with an additional predicate symbol P, \(\psi \) is a formula of the language \(\mathcal {L}_{T}\) expanded with P, and \(\varphi (\hat{x}\psi (x))\) is the formula obtained by replacing the atoms of the form P(t) in \(\varphi (P)\) by \(\psi (t)\).Footnote 15 Informally, the rule allows us to make inferences from schemata accepted in the original arithmetical language to schemata of the language with reflective means (i.e. self-referential truth). It turns out that the resulting notion of schematic reflective closure \( Ref ^*(\mathsf {PA}(P))\) yields an alternative unramified characterization of predicative analysis (see Feferman [17]):

$$ Ref ^*(\mathsf {PA}(P))\equiv (\Pi ^0_1\text {-CA})_{<\Gamma _0}; $$

this proof-theoreticFootnote 16 equivalence still holds even when we add consistency (CONS) to \( Ref ^{*}(\mathsf {PA}(P))\). The theorem witnesses that investigations deriving from formal semantics and paradoxes have reached a high level of integration with different areas of foundational investigations; for instance, while the lower bound result is actually a refinement of typical predicative well-ordering techniques, the upper bound theorem can be achieved by techniques and results from reductive proof theory (see Feferman [12]).

Remarks

  1. (i)

    The main result about reflective closure (either ordinary or schematic) still holds once consistency (CONS) is replaced by the completeness axiom, i.e.

    $$ (\forall x) ({{\mathrm{{\text {Sent}}_{\mathsf {KF}}}}}(x) \rightarrow ( T(x) \vee F(x)). $$

    This axiom, which is true in the greatest fixed point of the Kripkean operator for self-referential truth, rules out truth value gaps in the same way as the consistency axiom (CONS) rules out truth value gluts. Kripke [42] considered only consistent fixed point (that is, fixed points without overlapping extension and antiextension of truth). But later Visser [60] and others generalized the approach to fixed points where extension and antiextension are allowed to overlap and thus a sentence can be simultaneously true and false.

  2. (ii)

    It turns out that suitable variants of ‘reflective closure’ can be profitably applied as tools for proof-theoretic investigations: for instance, as an intermediate step for computing the proof theoretic strength of transfinitely iterated fixed point theories (see Jäger et al. [38]).

3.2.4 Digression: \(\mathsf {KF}\) with Minimality

Many philosophers think that the minimal fixed point model of Kripke’s [42] theory is the most natural. It gives a picture of grounded truth: The truth and falsity of any sentence ultimately depends on the truth and falsity of non-semantic sentences, that is, sentences that do not contain the truth or falsity predicates. In other fixed points that are not minimal ungrounded sentences such as a truth teller sentence can be true and false as well.

\(\mathsf {KF}\) is the theory of all fixed point models of the monotone operator specifying the clauses for reflective truth. Consequently, \(\mathsf {KF}\) doesn’t decide the truth teller sentence. If one aims at a theory of the minimal fixed point and thus of grounded truth, one can try to add axioms that exclude fixed point models that are not minimal. Of course, this is conceptually relevant for the whole enterprise of truth: think of the important distinctions that arise from considering the plurality of truth predicates (grounded sentence, paradoxical sentence, intrinsic, etc. see Kripke [42]).

Definition 3

  1. (i)

    \(\mathsf {KF}\) + \(\mathsf {GID}\): see Cantini [5]. \(\mathsf {GID}\) is the schema

    $$\begin{aligned} \forall x( \varphi (x,\hat{x}\psi (x))\rightarrow \psi (x))\rightarrow \forall x (x\in I_\varphi \rightarrow \psi (x)) \end{aligned}$$
    (3.11)

    where \(\psi \) is an arbitrary formula, \(I_\varphi \) is obtained by diagonalization (Lemma 3.9, [5]), \(\varphi (x, P)\) is an elementary positive (in P) operator in the language of \(\mathcal {L}_T\) with an additional predicate variable, P positive.Footnote 17

  2. (ii)

    \({\mathsf {KF}_\mu }\) (Truth with minimality, see Burgess [4]): it is the fragment of \(\mathsf {KF}\!\!\upharpoonright \) with

    1. (i)

      only the composition principles, e.g. \(\forall x( \mathcal {T}(x, T)\rightarrow T(x))\);

    2. (ii)

      the schema: if \(\psi \) is an arbitrary formula,

      $$\begin{aligned} \forall x( \mathcal {T}(x,\hat{x}\psi (x))\rightarrow \psi (x))\rightarrow \forall x ( T(x)\rightarrow \psi (x)) \end{aligned}$$
      (3.12)

Then \({\mathsf {KF}_\mu }\) proves the decomposition axioms and the consistency axiom. Also, it explicitly refutes statements that fail in the least fixed point model. Clearly \({{\mathsf {KF}}_{\mu }}\!\!\upharpoonright \) has an inner model in \(\mathsf {KF}\!\!\upharpoonright +GID\) (see proposition 3.12, Cantini [5]).

As to the upper bound, we can apply the proof-theoretic methods of Cantini [7] to \({{\mathsf {KF}}_{\mu }}\!\!\upharpoonright \).

As to the lower bound, let \(\mathsf {ID^{acc}_1}\) be the theory of accessibility inductive definitions over \(\mathsf {PA}\), which is known to be proof-theoretically equivalent to the theory of elementary inductive definitions. Then we can lift to the present context a suitable version of the so-called bar-induction schema, which goes back to Kreisel:

Theorem 5

If \(\prec \) is a binary relation which is determined,Footnote 18 then the schema of transfinite induction on the largest well-founded part \(W(\prec )\) of \(\prec \) holds for arbitrary formulas \(\varphi \), provably in \({\mathsf {KF}_\mu }\) .

The proof takes inspiration from an analogous result of Friedman and Sheard [25]; it is also exploited by Burgess [4]. Hence we have by the previous theorem:

Corollary 2

\(\mathsf {ID^{acc}_1}\) is interpretable in \({\mathsf {KF}_\mu }\) .

3.3 Related Works Inspired by \(\mathsf {KF}\)

\(\mathsf {KF}\) is nowadays considered to have brought the birth to the subject called axiomatic theories of truth as an area of logical research in its own right. A variety of formal theories of truth not in a semantic form but in an axiomatic form have been presented since \(\mathsf {KF}\). Some try to axiomatize known semantic theory or construction of truth as \(\mathsf {KF}\) does for Kripke’s theory of truth; others take more purely axiomatic approach by considering what combinations of truth-theoretic principles are possible (consistent) and plausible. Friedman and Sheard [25] listed nine natural and naïvely correct postulates of truth, which are inconsistent altogether, and determined the maximal consistent combinations of them. The other two theories of Feferman’s own presented below would be counted in the latter ‘purely axiomatic’ type of axiomatic theories of truth as well. Cantini [6] presented a system that can be regarded as an axiomatization of Kripke’s least fixed-point truth with supervaluation schema. Horsten et al. [36] made an attempt to axiomatize Gupta–Belnap–Herzberger revision theoretic truth. There are far more examples that we could list here, and the subject is still lively developing.

4 Type-Free T-Schema with Reinterpreted Biconditional

The second theory of truth by Feferman [15] was originally intended to provide a uniform type-free treatment for set-comprehension and truth-predication, both of which notoriously yield a contradiction with their naïve formulations. Feferman [18] described the purpose of the theory as ‘pragmatic’; particularly, on the side of set-comprehension, the theory was presented for the purpose of suitably and consistently dealing with natural type-free statements that mathematicians would like to make, for instance, in category theory. However, he later gave it a reformation into a theory of deflationist truth in Feferman [19].

The core idea behind the theory is to formally interpret the informal biconditional ‘if and only if’ in the Comprehension Axiom and T-schema, not as the material biconditional \(\leftrightarrow \), but as another new primitive logical connective \(\equiv \) so that \(\equiv \) preserves certain essential connotations or characteristics of the informal biconditional but still avoids a contradiction. Let us restrict ourselves to his theory of truth. The T-schema is expressed in English as

$$\begin{aligned} \ulcorner \varphi \urcorner \text { is true if and only if } \varphi \end{aligned}$$
(4.1)

for arbitrary \(\varphi \in \mathcal {L}_{T}\). If we interpret ‘if and only if’ by the standard material biconditional \(\leftrightarrow \), then the resulting translation of (4.1) is inconsistent. Hence, Feferman introduces a new binary connective \(\equiv \) into the syntax of a theory, and add the following as a new axiom schema that meant to interpret (4.1):

$$ T(\ulcorner \varphi \urcorner ) \equiv \varphi , $$

for arbitrary \(\varphi \in \mathcal {L}_{T}\).

Since \(\equiv \) is a newly introduced connective separate and independent of any other standard connectives such as \(\rightarrow \) and \(\lnot \), a suitable axiomatic characterization should be provided for it, and also such a characterization should be so made that \(\equiv \) is seen as representing the informal biconditional ‘if and only if’ at least in the context of our discourse involving truth.

Feferman [15] initially gave a semantic characterization of \(\equiv \), where he describes how to define the extension of a truth predicate and the semantic evaluation rule for the new connective \(\equiv \) over a given model of a base theory. The construction goes as follows. We start with an arbitrary model \(\mathfrak {M}\) with a domain M of a base theory \(\mathsf {B}\) over a language \(\mathcal {L}\). For the simplicity of argument, we take \(\mathcal {L}_{\mathsf {PA}}\) and \(\mathsf {PA}\) to be \(\mathcal {L}\) and \(\mathsf {B}\) respectively in what follows. Recall that we defined \(\mathcal {L}_{T}= \mathcal {L}_{\mathsf {PA}}\cup \{ T\}\). Let \(\mathcal {L}_{T}( \equiv )\) be a language obtained by augmenting \(\mathcal {L}_{T}\) with the connective \(\equiv \). We first give the Kripkean least fixed-point construction of truth over \(\mathfrak {M}\) with a suitable partial logic such as strong Kleene or 3-valued Łukasiewicz logic, and let \(X \subset M\) be the extension of the predicate \(T\) in thus constructed fixed-point semantics over \(\mathcal {L}_{T}\) with the partial logic we have chosen. Now, for each \(\mathcal {L}_{T}( \equiv )\)-formula \(\varphi \), let us denote by \(\varphi ^{*}\) the \(\mathcal {L}_{T}\)-formula obtained by replacing all the occurrences of \(\equiv \) in \(\varphi \) with the material biconditional \(\leftrightarrow \). Then, we expand the base model \(\mathfrak {M}\) to an \(\mathcal {L}_{T}( \equiv )\)-structure \(\mathfrak {M} ( \equiv )\) by putting

  • \(\mathfrak {M} ( \equiv ) \models T(a)\), iff \(a = \ulcorner \sigma \urcorner \) for some \(\mathcal {L}_{T}( \equiv )\)-sentence \(\sigma \) and \(\sigma ^{*}\) is contained in thus constructed Kripkean truth, and

  • \(\mathfrak {M} ( \equiv ) \models \varphi \equiv \psi \), iff \(\varphi ^{*}\) and \(\psi ^{*}\) have the same semantic value in the fixed-point semantics;

note that \(\equiv \) is now changed to \(\leftrightarrow \) in \(\varphi ^{*}\) and \(\psi ^{*}\) and it is evaluated according to the partial logic we chose in the fixed-point semantics; the evaluation of \(\mathcal {L}_{\mathsf {PA}}\)-atomics in \(\mathfrak {M} ( \equiv )\) are the same as \(\mathfrak {M}\), and the other logical connectives and quantifiers are classically evaluated in \(\mathfrak {M} ( \equiv )\). Hence, an informal reading of \(\varphi \equiv \psi \) is ‘the semantic values of \(\varphi \) and \(\psi \) coincides in the Kripke’s least fixed-point semantics’, and thus since the Kripkean construction gives the same semantic value to \(T(\ulcorner \varphi \urcorner )\) and \(\varphi \) for all \(\mathcal {L}_{T}\)-sentences \(\varphi \) it holds that

$$\begin{aligned} \mathfrak {M} ( \equiv ) \models T(\ulcorner \varphi \urcorner ) \equiv \varphi , \text { for all } \mathcal {L}_{T}( \equiv )\text {--sentences } \varphi . \end{aligned}$$
(4.2)

Under this interpretation \(\equiv \) enjoys several desired properties. Let \(\mathbf {t} := 0 = 0\) and \(\mathbf {f} := 0 \ne 0\): namely, \(\mathbf {t}\) and \(\mathbf {f}\) respectively stand for definitely true and false sentences. Then we define \(\mathcal {D} ( \varphi )\) to abbreviate \(( \varphi \equiv \mathbf {t} ) \vee (\varphi \equiv \mathbf {f} )\), which informally expresses ‘\(\varphi \) is determinate.’ Then, we can show that \(\mathfrak {M} ( \equiv )\) is a model of the following.

  1. S1

    \(T \ulcorner \varphi \urcorner \equiv \varphi \) for all \(\mathcal {L}_{T} ( \equiv )\)-sentences \(\varphi \)

  2. S2

    \(\equiv \) is an equivalence relation

  3. S3

    \(\mathbf {t} \not \equiv \mathbf {f}\)

  4. S4

    \(\equiv \) preserves all the connectives and quantifiers including \(\equiv \) itself; e.g., \(( ( \varphi \equiv \psi ) \wedge ( \sigma \equiv \chi ) ) \rightarrow ( ( \varphi \equiv \sigma ) \equiv ( \psi \equiv \chi ) ) \wedge ( ( \varphi \vee \sigma ) \equiv ( \psi \vee \chi ) )\).

  5. S5

    \(\mathcal {D} ( \varphi )\) for all \(\mathcal {L}_{\mathsf {PA}}\)-sentences.

  6. S6

    \(\mathcal {D}\) is closed under \(\lnot \), \(\vee \), and \(\forall \).

  7. S7

    \(( \varphi \equiv \mathbf {t} \rightarrow \varphi ) \wedge ( \varphi \equiv \mathbf {f} \rightarrow \lnot \varphi )\).

Proof

S1 immediately follows from (4.2). S2 and S4 are obvious from the evaluation for \(\equiv \). S3 and S5 hold since the fixed-point semantics does not change the evaluations of \(\mathcal {L}_{\mathsf {PA}}\)-sentences. S6 is due to the evaluation schema of the logic we chose. Finally we can show S7 by induction on \(\varphi \). \(\square \)

Later Feferman [19] adopted these seven properties of \(\equiv \) as the axioms of a deflationist theory of truth which he calls \(\mathsf {S}\). There he interprets deflationism in the form that the proposition \(T\ulcorner \varphi \urcorner \) is evaluated in the same way as \(\varphi \) and they are thereby equated even if they do lack a determined truth valueFootnote 19; hence the aforementioned informal reading of \(\equiv \) well fits in this interpretation. Furthermore, since any model of \(\mathsf {PA}\) can be expanded to a model of \(\mathsf {S}\), \(\mathsf {S}\) is conservative over \(\mathsf {B}\); it is even ‘semantically conservative’ over \(\mathsf {B}\) in the sense of Craig and Vaught [8]. Thus it satisfies the ‘conservativeness requirement’ for deflationist theories of truth.Footnote 20

Remark 1

Several extra axioms can be still conservatively added to \(\mathsf {S}\). For instance, we may consider the following further axioms for \(\mathsf {S}\).

  1. S8

    \(( \varphi \vee \psi ) \equiv \mathbf {t} \leftrightarrow \varphi \equiv \mathbf {t} \vee \psi \equiv \mathbf {t}\).

  2. S9

    \(\exists x \varphi ( x ) \equiv \mathbf {t} \leftrightarrow \exists x ( \varphi ( x ) \equiv \mathbf {t} )\).

  3. S10

    \(( \varphi \equiv \psi ) \equiv \mathbf {t} \leftrightarrow \varphi \equiv \psi \).

  4. S11

    \(( \varphi \equiv \psi ) \equiv \mathbf {t} \leftrightarrow \mathcal {D} ( \varphi ) \wedge \mathcal {D} ( \psi ) \wedge ( \varphi \equiv \psi )\).

  5. S12

    \(\mathcal {D}\) is strongly compositional with respect to the connectives and quantifiers except \(\equiv \); e.g., \(\mathcal {D} ( \varphi \vee \psi ) \leftrightarrow \mathcal {D} ( \varphi ) \wedge \mathcal {D} ( \psi )\).

Which of them can be conservatively added to \(\mathsf {S}\) depends on which partial logic we choose in the construction of the Kripkean fixed-point semantics over \(\mathfrak {M}\). For instance, if we choose 3-valued Łukasiewicz logic, then we can show that the resulting semantics \(\mathfrak {M} ( \equiv )\) is a model of S8–S10 and thus they can be conservatively added to \(\mathsf {S}\); if we choose strong Kleene logic instead, S8–S9 and S11 are satisfied in \(\mathfrak {M} ( \equiv )\) and thus can be conservatively added to \(\mathsf {S}\); by choosing weak Kleene logic, we can also see that S11–S12 can be conservatively added to \(\mathsf {S}\).

5 Determinate Truth

In the system \(\mathsf {S}\) of the previous section one has the unrestricted T-schema. However, the biconditional in the T-schema is no longer the classical biconditional \(\leftrightarrow \) but rather the nonclassical \(\equiv \). This is the price for the unrestricted T-schema. Instead of having the unrestricted T-schema and a nonclassical biconditional, one can restrict the T-schema to well-behaved sentences and then have the full classical T-schema and all other truth-theoretic axioms one might want to have. Feferman [18] pursued this strategy with his system \(\mathsf {{DT}}\) .

Of course the restriction needs to be well motivated. According to Russell [52], every predicate P has a domain of significance D and it makes sense to apply P only to objects in D and the principles which are supposed to characterize (or axiomatize) the concept expressed by P are meant to be only about the objects in D. Hence, in the case of truth, if such a domain is appropriately given and if no contradictory sentences, such as the liar sentence, are contained in D, the naïve truth-theoretic principles, such as Tarski’s T-schema, can be safely postulated. For Feferman such a domain D for the case of truth consists of the sentences that are meaningful and determinate,Footnote 21 that is, have a definite truth value, true of false.

Let us assume that our base theory is \(\mathsf {PA}\) as before. A theory of truth based on the above view have two predicate \(T\) and D representing truth and its domain of significance; namely, the language \(\mathcal {L}_{\mathsf {{DT}}}\) of \(\mathsf {{DT}}\) is \(\mathcal {L}_{\mathsf {PA}}\cup \{ T, D \}\). It is plausible for him to require that the condition on the domain D of significance for any predicate P should be prior to the conditions on P; in the current context, an axiomatic characterization of D should be given without appealing to \(T\). Feferman takes strong compositionality as the axiomatic characterization of the domain D of significance for \(T\). Namely, D is closed under the propositional operations and quantifiers, and, conversely, a sentence is meaningful only if all of (the substitution instances of) its parts are meaningful. Thus conceived strong compositionality can be axiomatized as follows:

  1. D1

    \(\forall x\, \big ({{\mathrm{{\text {AtSent}}_{\mathsf {PA}}}}}(x) \rightarrow D(x) \big )\)

  2. D2

    \(\forall s \,\bigl ( D ( \mathop {T}\limits _{\cdot }s ) \leftrightarrow D ({s}^\circ ) \bigr )\)

  3. D3

    \(\forall x\,\big ( {{\mathrm{{\text {Sent}}_{\mathsf {{DT}}}}}}(x)\rightarrow \bigl ( D(\mathop {\lnot }\limits _{\cdot }x) \leftrightarrow D(x) \bigr )\)

  4. D4

    \(\forall x \forall y \,\big ( {{\mathrm{{\text {Sent}}_{\mathsf {{DT}}}}}}(x\mathop {\vee }\limits _{\cdot }y)\rightarrow \big (D(x \mathop {\vee }\limits _{\cdot }y) \leftrightarrow (Dx \wedge Dy)\big ) \big )\)

  5. D5

    \(\forall v \forall x \, \big ( {{\mathrm{{\text {Sent}}_{\mathsf {{DT}}}}}}(\mathop {\forall }\limits _{\cdot }v. x) \rightarrow \bigl ( D(\mathop {\forall }\limits _{\cdot }v. x) \leftrightarrow \forall t D (x [t/v]) \bigr ) \big )\)

The formula \({{\mathrm{{\text {AtSent}}_{\mathsf {PA}}}}}\) stands for the set of all atomic \(\mathcal {L}_{\mathsf {PA}}\)-sentences, and \({{\mathrm{{\text {Sent}}_{\mathsf {{DT}}}}}}\) for the set of sentences of \(\mathcal {L}_{\mathsf {{DT}}}\). The axiom D1 does not come from the strong compositionality requirement in question but from the common and plausible view that the sentences of the base language, which contains no semantic predicates, should be non-paradoxical, determinate, and counted in the domain of the significance of \(T\).

The T-schema and other truth-theoretic principles are accordingly restricted to such sentences. For instance, the T-schema is formulated as follows:

$$\begin{aligned} D (\ulcorner \varphi \urcorner ) \rightarrow ( T (\ulcorner \varphi \urcorner ) \leftrightarrow \varphi ), \ \text { for all } \mathcal {L}_{\mathsf {{DT}}}\text {-sentences } \varphi . \end{aligned}$$
(5.1)

Feferman points out that thus formulated T-schema yields the following intimate connection of D and T as a special phenomenon particular to the case of truth:

$$\begin{aligned} D (\ulcorner \varphi \urcorner ) \leftrightarrow \big ( T (\ulcorner \varphi \urcorner ) \vee T (\ulcorner \lnot \varphi \urcorner ) \big ) \end{aligned}$$
(5.2)

Its proof goes as follows. Suppose \(D (\ulcorner \varphi \urcorner )\) for one direction. By D3 we have \(D (\ulcorner \lnot \varphi \urcorner )\). Hence, (5.1) yields \(T (\ulcorner \varphi \urcorner ) \leftrightarrow \varphi \) and \(T (\ulcorner \lnot \varphi \urcorner ) \leftrightarrow \lnot \varphi \). Finally, the law of excluded middle \(\varphi \vee \lnot \varphi \) entails \(T (\ulcorner \varphi \urcorner ) \vee T (\ulcorner \lnot \varphi \urcorner )\). Suppose \(( T (\ulcorner \varphi \urcorner ) \vee T (\ulcorner \lnot \varphi \urcorner ) )\) for the converse. Since T only applies to its domain of significance D, we have \(D (\ulcorner \varphi \urcorner ) \vee D (\ulcorner \lnot \varphi \urcorner )\). Hence \(D (\ulcorner \varphi \urcorner )\) follows from D3. Feferman then generalizes (5.2) to a universal statement and postulates

  1. D0

    \(\forall x \big ( D (x) \leftrightarrow ( T (x) \vee T (\mathop {\lnot }\limits _{\cdot }x) ) \big )\)

Thereby, specially in the case of truth, the domain predicate D becomes definable in terms of \(T\) and redundant. Indeed, Feferman [18] does not take a predicate D as a primitive predicate for this reason. Instead he formulates DT over \(\mathcal {L}_{T}\) and introduces Dx as a mere abbreviation of \(T (x) \vee T (\mathop {\lnot }\limits _{\cdot }x)\). Therefore let us assume \(\mathcal {L}_{\mathsf {{DT}}}= \mathcal {L}_{T}\) in what follows.

Given these characterizations of D, Feferman formulates the axioms for T as the the compositional Tarskian clauses restricted to D:

  1. D6

    \(\forall s\, \forall t\, \big ( T( s \mathop {=}\limits _{\cdot } t)\leftrightarrow {s}^\circ \! =\!{t}^\circ ) \big ) \), and similarly for other predicates of \(\mathcal {L}_{\mathsf {PA}}\)

  2. D7

    \(\forall s\, \bigl ( D ( {s}^\circ ) \rightarrow \bigl ( T(\mathop {T}\limits _{\cdot }s) \leftrightarrow T({s}^\circ ) ) \bigr ) \)

  3. D8

    \(\forall x\,\big ( {{\mathrm{{\text {Sent}}_{\mathsf {{DT}}}}}}(x)\wedge D(x)\rightarrow \bigl ( T(\mathop {\lnot }\limits _{\cdot }x) \leftrightarrow \lnot T(x) \bigr )\)

  4. D9

    \(\forall x \forall y \,\big ( {{\mathrm{{\text {Sent}}_{\mathsf {{DT}}}}}}(x\mathop {\vee }\limits _{\cdot }y)\wedge D(x \mathop {\vee }\limits _{\cdot }y)\rightarrow \bigl ( T( x \mathop {\vee }\limits _{\cdot }y ) \leftrightarrow (T(x) \vee T(y)) \bigr ) \big )\)

  5. D10

    \(\forall v \forall x \, \big ( {{\mathrm{{\text {Sent}}_{\mathsf {{DT}}}}}}(\mathop {\forall }\limits _{\cdot }v. x) \wedge D(\mathop {\forall }\limits _{\cdot }v. x) \rightarrow \bigl ( T( \mathop {\forall }\limits _{\cdot }v. x ) \leftrightarrow \forall t T(x [t/v]) \bigr )\big )\)

Let us denote by \(\mathsf {DT^{w}}\) the axiomatic system comprising \(\mathsf {PA}\) plus D0–D10 augmented by the expanded induction for \({{\mathrm{{\text {Sent}}_{\mathsf {{DT}}}}}}\). We can easily show by induction on formulae that (5.1) is a theorem of \(\mathsf {{DT}}\) . This \(\mathsf {DT^{w}}\) is a straightforward axiomatization of Feferman’s view described so far and is already proof-theoretically pretty strong; \(\mathsf {DT^{w}}\) is proof-theoretically equivalent to \(\mathsf {KF}\) and indeed identical as theories with \(\mathsf {WKF} + \mathsf {CONS}\) , the Kripke–Feferman system for weak Kleene logic plus the axiom of consistency; see [26] for more details. However, he sees some defects in \(\mathsf {DT^{w}}\) as a theory of truth and slightly amends it.

We can standardly show that \(\mathsf {DT^{w}}\) proves the global reflection principle as \(\mathsf {KF}\) does: i.e.,

$$ \mathsf {DT^{w}}\vdash \forall x \bigl ( {\text {Bew}}_{\mathsf {PA}}( x ) \rightarrow T(x) \bigr ). $$

However, in contrast to \(\mathsf {KF}\), \(\mathsf {DT^{w}}\) does not prove the truth of the global reflection principle. Feferman requires for an adequate theory of truth to derive it and thereby suggests to amend \(\mathsf {DT^{w}}\) .

Let us first see how its proof goes in \(\mathsf {KF}\). In addition to the global reflection principle, both \(\mathsf {KF}\) and \(\mathsf {DT^{w}}\) derive \(\forall s \big ( T(\mathop {T}\limits _{\cdot }s) \leftrightarrow T({s}^\circ ) \big )\) as well as

$$ \forall x\, \Big ( \big ( T(\ulcorner {\text {Bew}}_{\mathsf {PA}}( \dot{x} )\urcorner ) \leftrightarrow {\text {Bew}}_{\mathsf {PA}}( x ) \big ) \, \wedge \, \big ( T(\ulcorner \lnot {\text {Bew}}_{\mathsf {PA}}( \dot{x} )\urcorner ) \leftrightarrow \lnot {\text {Bew}}_{\mathsf {PA}}( x ) \big ) \Big ). $$

Hence, it follows from these that both \(\mathsf {KF}\) and \(\mathsf {DT^{w}}\) derive

$$ \forall x\, \big ( T(\ulcorner {\text {Bew}}_{\mathsf {PA}}( \dot{x} )\urcorner ) \rightarrow T(\mathop {T}\limits _{\cdot }\dot{x}) \big ). $$

By using the equivalence of \(\varphi \rightarrow \psi \) and \(\lnot \varphi \vee \psi \), we obtain

$$\begin{aligned} \forall x\, \big ( T(\ulcorner \lnot {\text {Bew}}_{\mathsf {PA}}( \dot{x} )\urcorner ) \vee T(\mathop {T}\limits _{\cdot }\dot{x}) \big ). \end{aligned}$$
(5.3)

However, one has to verify in \(\mathsf {DT^{w}}\) that \(\forall x D ( T x )\), in order to proceed from (5.3) to \(\forall x ( T \ulcorner \lnot {\text {Bew}}_{\mathsf {PA}}( \dot{x} ) \vee T x\urcorner )\), but \(\forall x D ( T x )\) is incompatible with the view behind \(\mathsf {DT^{w}}\) and indeed underivable in \(\mathsf {DT^{w}}\) ; in contrast, \(\mathsf {KF}\) can immediately derive it via the axioms K4 and K5 and thereby the truth of the global reflection principle. Given this observation, Feferman proposes to treat \(\rightarrow \) as separate basic propositional connectives from \(\vee \) and \(\lnot \), not identifying \(\varphi \rightarrow \psi \) with \(\lnot \varphi \vee \psi \), and then postulates an independent axiom for the meaningfulness and determinateness of the sentences of the form \(\varphi \rightarrow \psi \). Namely, he postulate the following two axioms for the conditional:

  1. D11

    \(\forall x \forall y \,\Big ( {{\mathrm{{\text {Sent}}_{\mathsf {{DT}}}}}}(x\mathop {\rightarrow }\limits _{\cdot }y)\rightarrow \big (D(x \mathop {\rightarrow }\limits _{\cdot }y) \leftrightarrow \big ( D(x) \wedge (T(x) \rightarrow D(y)) \big )\big ) \Big )\)

  2. D12

    \(\forall x \forall y \,\Big ( {{\mathrm{{\text {Sent}}_{\mathsf {{DT}}}}}}(x\mathop {\rightarrow }\limits _{\cdot }y)\wedge D (x\mathop {\rightarrow }\limits _{\cdot }y)\rightarrow \bigl ( T( x \mathop {\rightarrow }\limits _{\cdot }y ) \leftrightarrow ( T(x) \rightarrow T(y) ) \bigr ) \Big )\)

It is true, as Feferman himself points out, strong compositionality in the aforementioned sense is not met with respect to the conditional ‘\(\rightarrow \)’, since D11 indicates that if \(\varphi \) is false then \(D (\ulcorner \varphi \rightarrow \psi \urcorner )\) may hold even when \(D (\ulcorner \psi \urcorner )\) fails, but he argues that it is rather natural since we do not care whether \(D (\ulcorner \varphi \rightarrow \psi \urcorner )\) holds when \(\varphi \) is definitely false. Thereby the system \(\mathsf {{DT}}\) of determinate truth is defined to be \(\mathsf {PA}\) plus D0–D12 augmented by the expanded induction schema for \(\mathcal {L}_{\mathsf {{DT}}}\).

A closer inspection reveals that \(\mathsf {{DT}}\) is very intimately related to \(\mathsf {KF}\). \(\mathsf {KF}\) is an “axiomatization” of Kripke’s fixed-point semantics with strong Kleene logic. Naturally we may consider a similar “axiomatization” of the Kripkean fixed-point semantics with other logics. We have already mentioned such a theory for weak Kleene logic, which is related to \(\mathsf {DT^{w}}\). As a matter of fact, \(\mathsf {{DT}}\) is identical with such a \(\mathsf {KF}\)-style axiomatization of the Kripkean fixed-point semantics with what we call Aczel–Feferman logic, plus the axiom of consistency \(\mathsf {CONS}\) ; namely, the inner logic of the \(\mathsf {KF}\)-style system is Aczel–Feferman logic; see Fujimoto [27] for details.Footnote 22 Here we mean by Aczel–Feferman logic a variant of weak Kleene logic with an independent evaluation for the conditional ‘\(\rightarrow \)’ than that for \(\vee \) and \(\lnot \); in Aczel–Feferman logic, ‘\(\rightarrow \)’ is not defined away in terms of ‘\(\lnot \)’ and ‘\(\vee \)’ and the evaluation of ‘\(\rightarrow \)’ is determined by the following separate truth table:

$$ \begin{array}{ c@{\quad }|@{\quad }c@{\quad }|@{\quad }c@{\quad }|@{\quad }c } \rightarrow &{} T &{} U &{} F \\ \hline T &{} T &{} U &{} F \\ \hline U &{} U &{} U &{} U \\ \hline F &{} T &{} T &{} T \\ \end{array} $$

Hence, from a purely technical point of view, Feferman’s theory of determinate truth essentially boils down to his own axiomatization of Kripkean fixed-point semantics, and therefore \(\mathsf {{DT}}\) is proof-theoretically equivalent to \(\mathsf {KF}\). Also, if we extend \(\mathsf {{DT}}\) in the same way that we extend the reflective closure of \(\mathsf {PA}\), i.e., \(\mathsf {KF}\), to its schematic reflective closure , the resulting theory has the strength of the predicative limit \(\Gamma _{0}\).

Remark 2

Feferman [18] raised two conjectures about the strength of the theories of determinate truth over \(\mathsf {PA}\) , and the conjectures have been first settled by constructing a direct interpretation of them in \(\mathsf {KF}\) and \( Ref ^{*} ( \mathsf {PA}( P ) )\) in Fujimoto [26], but indeed the construction of Sect. 3.2.2 can also be adapted to the system \(\mathsf {{DT}}\) (over \(\mathsf {PA}\)), in order to produce an alternative proof of the first conjecture.

6 The Impact of Feferman’s Work on Truth

We have mentioned already some further research inspired by Feferman’s work. However, we don’t aim at a survey of the impact of his work, because the impact is far too wide-ranging to be summarized in a few pages. Moreover, an assessment of the reception of Feferman’s work is made harder by the fact that his work has been studied by different communities: In mathematical logic proof theorists have worked most extensively on Feferman’s system and their variants; but set theorists have also availed themselves to Feferman’s ideas (see Koellner [39]).

Also in philosophy Feferman’s work has been fruitful not only for one topic. For once, Feferman advanced the use of truth as a primitive concept in the foundations of mathematics. Of course, this was his main motivation in the seminal Reflecting on Incompleteness. Generally Feferman didn’t put so much emphasis on his theories as attempts to resolve the liar paradox; rather he was after theories of truth (and comprehension) that prove useful for foundational purposes. Of course, the paradoxes have to be addressed in some way, but the analysis of the paradoxes themselves doesn’t figure very prominently in Feferman’s work. Philosophers – among them Reinhardt [50], McGee [46] and Field [24] – have discussed Feferman’s systems from the perspective of the theory of paradoxes. In this respect they also proved very fruitful.

One aspect of Feferman’s work that hasn’t been fully exploited is its relevance for the discussion about deflationism. Of course in some way, deflationists have employed ideas from Feferman’s papers. In particular, Field’s [24] recent research programme is very much motivated by deflationist or disquotationalist considerations and Field acknowledges Feferman’s influence.

In particular, much of the work of Field and the work inspired by it can be seen as a continuation of Feferman’s [15]: The challenge is to obtain the full T-schema in an extension of classical logic with a reasonably behaved conditional (or biconditional).

Deflationists have often seen truth as a device for expressing generalizations and infinite conjunctions (see Horwich [37] and Halbach [31]). Feferman’s proof-theoretic analysis – more specifically the proof of the lower bound in Feferman [17, Sects. 4.3 and 5.3] – proceeds in terms of infinite conjunctions. While the discussion about the purpose of truth as a device of expressing infinite conjunctions usually remains somewhat metaphorical in the literature on deflationist, Feferman’s analysis offers a precise conceptual analysis on the usability of truth for expressing infinite conjunctions. This is just one aspect of Feferman’s work that offers scope for future philosophical research.