Hostname: page-component-848d4c4894-x5gtn Total loading time: 0 Render date: 2024-05-18T11:10:30.478Z Has data issue: false hasContentIssue false

THE JACOBSON RADICAL OF A PROPOSITIONAL THEORY

Published online by Cambridge University Press:  02 December 2021

GIULIO FELLIN
Affiliation:
DIPARTIMENTO DI MATEMATICA UNIVERSITÀ DEGLI STUDI DI TRENTO VIA SOMMARIVE 14, 38123 POVO (TRENTO), ITALY and DIPARTIMENTO DI INFORMATICA UNIVERSITÀ DEGLI STUDI DI VERONA STRADA LE GRAZIE 15, 37134VERONA, ITALY and DEPARTMENT OF PHILOSOPHY, HISTORY AND ART STUDIES UNIVERSITY OF HELSINKI HELSINKI, FINLAND E-mail: giulio.fellin@univr.it
PETER SCHUSTER
Affiliation:
DIPARTIMENTO DI INFORMATICA UNIVERSITÀ DEGLI STUDI DI VERONA STRADA LE GRAZIE 15, 37134VERONA, ITALYE-mail: peter.schuster@univr.it
DANIEL WESSEL
Affiliation:
DIPARTIMENTO DI INFORMATICA UNIVERSITÀ DEGLI STUDI DI VERONA STRADA LE GRAZIE 15, 37134VERONA, ITALY and MATHEMATISCHES INSTITUT DER UNIVERSITÄT MÜNCHEN THERESIENSTR. 39, D-80333 MÜNCHEN, GERMANYE-mail: daniel.wessel@univr.it
Rights & Permissions [Opens in a new window]

Abstract

Alongside the analogy between maximal ideals and complete theories, the Jacobson radical carries over from ideals of commutative rings to theories of propositional calculi. This prompts a variant of Lindenbaum’s Lemma that relates classical validity and intuitionistic provability, and the syntactical counterpart of which is Glivenko’s Theorem. The Jacobson radical in fact turns out to coincide with the classical deductive closure. As a by-product we obtain a possible interpretation in logic of the axioms-as-rules conservation criterion for a multi-conclusion Scott-style entailment relation over a single-conclusion one.

Type
Articles
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
© The Author(s), 2021. Published by Cambridge University Press on behalf of The Association for Symbolic Logic

1 Introduction

Glivenko’s theorem from 1929 says that if a propositional formula ${\varphi }$ is provable in classical logic, then its double negation $\neg \neg {\varphi }$ is provable in intuitionistic logic. In 1933 Gödel extended this to predicate logic, which move required to admit on the intuitionistic side the scheme of double negation shift. With Gödel’s and Gentzen’s negative translation in place of double negation, both from 1933, one can even get by with minimal logic in place of intuitionistic logic. More than one related proof translation saw the light of the day, e.g., Kolmogorov’s (1925) and Kuroda’s (1951).

Glivenko’s theorem thus stood right at the beginning of a fundamental change of perspective: that classical logic can be embedded into intuitionistic or even minimal logic, rather than the latter being a limited version of the former. Together with the revision of Hilbert’s Programme ascribed to Kreisel and Feferman, this has led to the much broader quest for the computational content of classical proofs, today culminating in agile areas such as dynamical algebra, formal topology, program extraction from proofs, proof analysis, proof mining and proof translations. The growing success of these approaches suggests that customary mathematics, with classical logic and set theory, might eventually prove to be much more constructive than widely thought.

In 1930 Tarski ascribed to Lindenbaum the theorem that in classical logic any given theory T equals the intersection of all the complete theories containing T. Its typical use for Gödel’s Completeness Theorem aside, this Lindenbaum Lemma is one of several theorems from that period which describe the intersection of all the ideal objects extending a given concrete object. Those intersection theorems, in their full generality recognised as forms of the Axiom of Choice (AC), are often put by contraposition as extension or separation theorems. Apart from Lindenbaum’s, prominent cases are known by the names of Artin–Schreier, Hahn–Banach, Krull and Szpilrajn. The case in algebra closest to Lindenbaum’s Lemma, however, gained prominence only in 1945, when Jacobson pointed out the relevance of the intersection of all the maximal ideals of a given ring, i.e., of what is now known as the Jacobson radical.

In the present note we follow the analogy between maximal (proper) ideals and complete (consistent) theories to carry over the Jacobson radical from ideals of commutative rings to theories of propositional calculi (Section 5.2), where it turns out to coincide with the stable closure or with the closure with respect to classical logic (Proposition 3 and Corollary 1). This prompts a variant of Lindenbaum’s Lemma that relates classical validity and intuitionistic provability (Proposition 2), and the syntactical counterpart of which happens to be Glivenko’s Theorem in the form recalled above (Theorem 2).

As a by-product we obtain a possible interpretation in logic (Theorem 3) of the axioms-as-rules conservation criterion (Theorem 1) for a multi-conclusion Scott-style entailment relation $\vdash $ over a single-conclusion one $\rhd $ . This criterion has proved to be the common core of many a syntactical counterpart of a semantic conservation theorem corresponding to one of the aforementioned intersection theorems. Typically any such case of conservation means reduction to a special case characterised by additional axioms with (possibly empty) disjunctions in positive position. Applying the criterion means to eliminate the additional axioms for $\vdash $ by way of the corresponding disjunction elimination rules for $\rhd $ . The latter equally suffice for proof practice, and have proved admissible in all mathematical instances yet considered.

Our interpretation of the conservation criterion in propositional logic (Theorem 3) is tantamount to Glivenko’s Theorem (Theorem 2). As for the latter, disjunction elimination plays a central role in the proof of the former, together with some notorious features of (double) negation in intuitionistic logic and of provability in classical propositional logic (Lemmas 1 and 2).

2 Preliminaries

Unless specified otherwise, we work in a suitable fragment of Aczel’s Constructive Zermelo–Fraenkel Set Theory ( $\operatorname {\textbf {CZF}}$ ) [Reference Aczel1Reference Aczel and Rathjen5] based on intuitionistic first-order predicate logic. While in general the concepts of this paper are elementary and the proofs are direct anyway, we still pin down $\operatorname {\textbf {CZF}}$ as metatheory if only rather for convenience’s sake; in fact much less might suffice. Likewise, when we occasionally need to invoke a fragment of the principle of Excluded Middle or even a form of the $\operatorname {\textrm {AC}}$ , and thus go beyond $\operatorname {\textbf {CZF}}$ , we simply switch to $\operatorname {\textbf {ZF}}$ and $\operatorname {\textbf {ZFC}}$ , respectively, and indicate this accordingly.

For example, the Restricted Law of Excluded Middle (REM) is not a principle of $\operatorname {\textbf {CZF}}$ . This REM means ${\varphi }\,\lor \,\neg {\varphi }$ for every set-theoretic formula ${\varphi }$ that is, bounded in the sense that only set-bounded quantifiers of the types $\forall x\in y$ and $\exists x\in y$ occur in ${\varphi }$ . As is common in this context, negation is a defined connective: $\lnot {\varphi }\equiv {\varphi }\to \bot $ .

By a finite set we understand a set that can be written as ${a_1, \dots , a_n}$ for some $n \geq 0$ . Given any set S, let $\operatorname {\textrm {Pow}}(S)$ (respectively, $\operatorname {\textrm {Fin}}(S)$ ) consist of the (finite) subsets of S. We refer to [Reference Rinaldi, Schuster and Wessel92] for further provisos to carry over to the present note.Footnote 1

Convention. By an intermediate logic we mean an intermediate propositional calculus obtained by adding to the axioms of intuitionistic logic some classically valid propositional formulas [37].

We write $\rhd $ to denote (deducibility in) any such intermediate logic in a propositional language S.

The subsequent properties of (double) negation are due to Brouwer for intuitionistic logic [Reference Brouwer11, Reference Brouwer12, Reference Glivenko45, Reference Troelstra109, Reference van Atten and Sundholm110] and carry over to an arbitrary $\rhd $ :

Lemma 1. For any given intermediate logic $\rhd $ in a propositional language S,

for every $\Gamma \in \operatorname {\mathop {Pow}}(S)$ and all ${\varphi },\psi \in S$ .

We refer to [Reference Ishihara55] and [Reference Negri and von Plato77, p. 27] for a deeper discussion with earlier references of the following:

Lemma 2. Let $\vdash $ and $\rhd $ stand for classical logic and an intermediate logic, respectively, in a propositional language S. If $\Gamma \in \operatorname {\mathop {Pow}}(S)$ and $\psi \in S$ , then $\Gamma \vdash \psi $ if and only if $\Gamma ,\Delta \rhd \psi $ for a suitable finite subset $\Delta $ of

$$ \begin{align*} \operatorname{\mathop{TND}}_0(\Gamma,\psi)=\{{\varphi}\lor\lnot{\varphi}: {\varphi} \ propositional\ variable\ occurring\ in\ \Gamma\ or\ \psi \}, \end{align*} $$

i.e., the set of relevant instances of tertium non datur for propositional variables.

A theory of an intermediate logic $\rhd $ is a subset T of the underlying propositional language S that is, deductively closed with respect to $\rhd $ :

$$ \begin{align*} \forall {\varphi} \in S(T\rhd {\varphi} \Rightarrow T\ni {\varphi}). \end{align*} $$

As usual, a theory T of $\rhd $ is

  • consistent if $\bot \notin T$ , which is to say that $T\neq S$ ;

  • complete if

    $$ \begin{align*} \forall {\varphi} \in S(T\ni {\varphi} \lor \lnot T\ni {\varphi}); \end{align*} $$
  • stable if

    $$ \begin{align*} \forall {\varphi}\in S(T\ni \lnot\lnot{\varphi}\Rightarrow T\ni {\varphi}). \end{align*} $$

As an immediate consequence of Lemmas 1 and 2 we have the following:

Lemma 3. Let $\rhd $ be an intermediate logic in a propositional language S. The following statements are equivalent for any given subset T of S:

  1. 1. T is deductively closed with respect to classical logic.

  2. 2. T is a stable theory of $\rhd $ .

  3. 3. T is a theory of $\rhd $ that contains all instances of excluded middle ${\varphi }\lor \neg {\varphi }$ with ${\varphi }\in S$ .

  4. 4. T is a theory of $\rhd $ that contains all ${\varphi }\lor \neg {\varphi }$ where ${\varphi }$ is a propositional variable of S.

In particular, if a theory T of an intermediate logic $\rhd $ is complete, then T is stable.

3 Entailment relations

Entailment relations, both in their single- and multi-conclusion variant, are at the heart of this note. We briefly recall the basic notions, to which end we closely follow [Reference Rinaldi, Schuster and Wessel91, Reference Rinaldi, Schuster and Wessel92].

3.1 Consequence

Let S be a set and $\rhd \subseteq \operatorname {\textrm {Pow}}(S)\times S$ . Once abstracted from the context of logical formulas, all but one of Tarski’s axioms of consequence [Reference Tarski106] can be put as

where $U,V\subseteq S$ and $a\in S$ . These axioms also characterise a finitary covering or Stone covering in formal topology [Reference Rinaldi and Wessel95];Footnote 2 see further [Reference Ciraulo, Maietti and Sambin17, Reference Ciraulo and Sambin19, Reference Negri72, Reference Negri73, Reference Sambin97, Reference Sambin98]. The notion of consequence has allegedly been described first by Hertz [Reference Hertz49Reference Hertz51]; see also [Reference Béziau7, Reference Legris59].

We do not employ the one of Tarski’s axioms by which he required that S be countable. This aside, Tarski has rather characterised the set of consequences of a set of propositions, which corresponds to the algebraic closure operator $U\mapsto U^{\rhd }$ on $\operatorname {\textrm {Pow}}(S)$ of a relation $\rhd $ as above where

$$\begin{align*}U^{\rhd}\equiv\{a\in S:U\rhd a\}. \end{align*}$$

3.2 Single-conclusion entailment

Rather than with Tarski’s notion, we henceforth work with its (tantamount) restriction to finite subsets, i.e., a single-conclusion entailment relation. This is, a relation $\rhd \subseteq \operatorname {\textrm {Fin}}(S)\times S$ such that

for all finite $U,U',V,V'\subseteq S$ and $a,b\in S$ , where as usual $U,V\equiv U\cup V$ and $V,b\equiv V\cup \{b\}$ . Our focus thus is on finite subsets of S, for which we henceforth reserve the letters $U, V, W, \ldots $ ; we also sometimes write $a_1,\ldots ,a_n$ in place of $\{a_1,\ldots ,a_n\}$ even if $n=0$ . Redefining

(1) $$ \begin{align} T^{\rhd}\equiv\{a\in S:\exists U\in\operatorname{\textrm{Fin}}(T)(U\rhd a)\}, \end{align} $$

for arbitrary subsets T of S gives back an algebraic closure operator on $\operatorname {\textrm {Pow}}(S)$ . By writing $T \rhd a$ in place of $a \in T^{\rhd }$ , the single-conclusion entailment relations thus correspond exactly to the relations satisfying Tarski’s axioms above.

3.3 Multi-conclusion entailment

Let S be a set and ${\vdash } \subseteq \operatorname {\textrm {Fin}}(S)\times \operatorname {\textrm {Fin}}(S)$ . Scott’s [Reference Scott102] axioms of entailment can be put as

for finite $U,U',V,V',W,W'\subseteq S$ and $b\in S$ , where $U\between W$ means that U and W have an element in common [Reference Sambin97]. Any such $\vdash $ is a multi-conclusion entailment relation, where ‘multi’ includes ‘empty’. In practice, $\rhd $ and $\vdash $ are inductively generated by the axioms of the intended models, which procedure we here take for granted [Reference Cederquist and Coquand14, Reference Coquand, Sambin, Smith and Valentini31]; see also [Reference Aczel3, Reference Rathjen87, Reference Rinaldi, Schuster and Wessel92, Reference Rinaldi and Wessel94].

This fairly general notion of entailment has been introduced by Scott [Reference Scott101Reference Scott103], building on Hertz’s and Tarski’s work (see above), and of course on Gentzen’s sequent calculus [Reference Gentzen43, Reference Gentzen44]. Shoesmith and Smiley [Reference Shoesmith and Smiley104] trace multi-conclusion entailment relations back to Carnap [Reference Carnap13], and Popper apparently had related ideas [Reference Popper85, Reference Popper86].Footnote 3 Before Scott, Lorenzen had developed analogous concepts formally [Reference Lorenzen65Reference Lorenzen68]; he even listed [Reference Lorenzen66, pp. 84–85] counterparts of the axioms (R), (T) and (M) for single- and multi-conclusion entailment [Reference Coquand, Lombardi and Neuwirth27, Reference Neuwirth81].Footnote 4 As compared with Gentzen’s and Lorenzen’s approaches, Scott’s entailment relation follow the contexts-as-sets paradigm, which has caused reservations [Reference Negri and von Plato78, Reference Negri and von Plato79]. The relevance of the notion of entailment relation to point-free topology and constructive algebra has been pointed out in [Reference Cederquist and Coquand14], and has been used very widely, e.g., in [Reference Coquand20Reference Coquand22, Reference Coquand24, Reference Coquand and Lombardi25, Reference Coquand and Persson29, Reference Coquand and Zhang32, Reference Negri, von Plato and Coquand80, Reference Rinaldi89, Reference Rinaldi and Wessel93, Reference Schlagbauer, Schuster and Wessel100, Reference Wessel114, Reference Wessel115]. Consequence and entailment have further caught interest from various other angles [Reference Avron6, Reference Došen35, Reference Gabbay41, Reference Humberstone52Reference Iemhoff54, Reference Payette and Schotch83, Reference Sandqvist99, Reference Shoesmith and Smiley104, Reference Wójcicki117].

4 Conservation

Again following [Reference Rinaldi, Schuster and Wessel91, Reference Rinaldi, Schuster and Wessel92], we sketch the concept of conservative extension of a multi-conclusion entailment relation $\vdash $ over a single-conclusion entailment relation $\rhd $ on the same set S. After that we extract from [Reference Rinaldi and Schuster90]—based on [Reference Ciraulo, Rinaldi and Schuster18]—possible interpretations limited to classical logic.

4.1 Conservation in syntax and semantics

Let S be a set, and let $a, b, c, \ldots $ and $U, V, W, \ldots $ range over the elements of S and $\operatorname {\textrm {Fin}}(S)$ , respectively. Given a multi-conclusion entailment relation $\vdash $ and a single-conclusion entailment relation $\rhd $ on the same set S, we throughout assume Extension:

$$\begin{align*}\operatorname{\textrm{Ext}}\qquad \frac{U\rhd a}{U\vdash a} \end{align*}$$

Of major interest to us is the converse, alias Conservation:

$$\begin{align*}\operatorname{\textrm{Con}}\qquad \frac{U\vdash a}{U\rhd a} \end{align*}$$

The trace of any given $\vdash $ is the single-conclusion entailment relation $\rhd _{\vdash }$ defined by

$$\begin{align*}U\rhd_{\vdash}a\equiv U\vdash a,\end{align*}$$

for which $\operatorname {\textrm {Ext}}$ and $\operatorname {\textrm {Con}}$ are tantamount to $\rhd \subseteq \rhd _{\vdash }$ and $\rhd \supseteq \rhd _{\vdash }$ , respectively.

An arbitrary subset P of S is a model of $\vdash $ if

$$\begin{align*}P\supseteq U\Rightarrow V\between P\quad\text{whenever } U\vdash V. \end{align*}$$

The notion of model carries over to single-conclusion $\rhd $ in the apparent manner, such that the models of $\rhd $ are exactly the $P\in \operatorname {\textrm {Pow}}(S)$ which are closed under $\rhd $ , i.e., for which $P^\rhd =P$ . Let $\operatorname {\textrm {Mod}}(\vdash )$ and $\operatorname {\textrm {Mod}}(\rhd )$ consist of the models of $\vdash $ and $\rhd $ , respectively. By Extension, $\operatorname {\textrm {Mod}}(\vdash )\subseteq \operatorname {\textrm {Mod}}(\rhd )$ , which in $\operatorname {\textbf {ZFC}}$ is equivalent to Extension [Reference Rinaldi, Schuster and Wessel92, Lemma 9].

Now $\operatorname {\textrm {Con}}$ follows from the Generalised Krull–Lindenbaum (GKL) Lemma, viz.

$$\begin{align*}\operatorname{\textrm{GKL}}\qquad \forall P \in \operatorname{\textrm{Mod}}(\vdash)(P \supseteq U \Rightarrow a \in P) \implies U \rhd a, \end{align*}$$

the converse of which holds by Extension. Again by Extension, $\operatorname {\textrm {GKL}}$ implies the Trace Completeness Theorem (TCT), viz.

$$\begin{align*}\operatorname{\textrm{TCT}} \qquad \forall P \in \operatorname{\textrm{Mod}}(\vdash)(P \supseteq U \Rightarrow a \in P) \implies U \vdash a, \end{align*}$$

the converse of which holds by the definition of a model of $\vdash $ . This $\operatorname {\textrm {TCT}} $ is a fragment of $\operatorname {\textrm {AC}}$ that implies REM [Reference Rinaldi, Schuster and Wessel92, Corollary 5].Footnote 5

In $\operatorname {\textbf {ZFC}}$ , $\operatorname {\textrm {GKL}}$ and $\operatorname {\textrm {Con}}$ are equivalent [Reference Rinaldi, Schuster and Wessel92, Theorem 6]. In $\operatorname {\textbf {CZF}}$ we can make this more precise:

Remark 1. In the presence of $\operatorname {\textrm {Ext}}$ , $\operatorname {\textrm {GKL}}$ is equivalent to the conjunction of $\operatorname {\textrm {Con}}$ and $\textrm{TCT}$ .

In all, $\operatorname {\textrm {GKL}}$ is semantic conservation, and Con is its syntactical counterpart.

4.2 Conservation in proof practice

In proof practice, $\operatorname {\textrm {GKL}}$ is useful for reductions to special cases, by making possible to use $\vdash $ in proofs about $\rhd $ , but $\operatorname {\textrm {GKL}}$ is of semantic nature, entails REM and requires some $\operatorname {\textrm {AC}}$ . In comparison, $\operatorname {\textrm {Con}}$ is equally sufficient for that kind of reduction, is syntactical and has elementary proofs. Many such cases are known in point-free topology such as locale theory and formal topology [Reference Cederquist and Coquand14, Reference Cederquist, Coquand and Negri15, Reference Coquand20, Reference Coquand23, Reference Mulvey and Wick-Pelletier70, Reference Mulvey and Wick-Pelletier71]; in constructive algebra, especially with dynamical methods [Reference Coquand and Lombardi26, Reference Coste, Lombardi and Roy33, Reference Lombardi61Reference Lombardi and Quitté64, Reference Wessel116, Reference Yengui118, Reference Yengui119]; and in the proof theory of order relations [Reference Negri and von Plato78, Reference Negri, von Plato and Coquand80]. Most of those cases concern algebra at large. But what about logic? One may think of Gentzen’s classical multi-succedent sequent calculus as extending his intuitionistic single-succedent variant [Reference Gentzen43, Reference Gentzen44, Reference Negri and von Plato77, Reference Takeuti105]. As we will see, this thought goes in the right direction.

A typical situation is as follows: Let the single-conclusion entailment relation $\rhd $ on a set S be generated by axioms. Then the multi-conclusion entailment relation $\vdash $ on the same set S is generated by the axioms of $\rhd $ , of course with $\vdash $ in place of $\rhd $ , and by additional axioms

$$ \begin{align*} \quad a_1,\dots,a_k\vdash b_1,\ldots,b_\ell, \end{align*} $$

where $k,\ell \ge 0$ . In any such situation we say that $\vdash $ extends $\rhd $ , and list the additional axioms if needed. This is legitimate inasmuch as if $\vdash $ extends $\rhd $ , then $\operatorname {\textrm {Ext}}$ is satisfied. What about $\operatorname {\textrm {Con}}$ ?

The following most versatile conservation criterion [Reference Rinaldi, Schuster and Wessel91, Reference Rinaldi, Schuster and Wessel92], which in fact gathers together many of the cases of Con mentioned before, will also help to understand Con for logic:

Theorem 1. Let $\vdash $ extend $\rhd $ with certain additional axioms of the form

(2) $$ \begin{align} \quad a_1,\dots,a_k\vdash b_1,\ldots,b_\ell, \end{align} $$

where $k,\ell \ge 0$ . Then $\vdash $ and $\rhd $ satisfy $\operatorname {\mathop {Con}}$ if and only if

(3) $$ \begin{align} \frac{W,b_1\rhd c \quad \cdots \quad W,b_\ell\rhd c }{W,a_1,\dots,a_k\rhd c} \end{align} $$

for every additional axiom (2), all $c\in S$ and every $W\in \operatorname {\mathop {Fin}}(S)$ .

This swiftly follows [Reference Rinaldi, Schuster and Wessel92, Theorem 2] from a sandwich criterion for conservation given by Scott [Reference Scott102], and also is a corollary of cut elimination for entailment relations [Reference Rinaldi and Wessel94] as related to cut elimination in the presence of axioms [Reference Negri and von Plato76].

Quite a few instances of $\operatorname {\textrm {GKL}}$ can be classified by the two cases named Universal Krull (UK) and Universal Lindenbaum (UL) in [Reference Rinaldi, Schuster and Wessel92], for which S is a set with

$$\begin{align*}& \textrm{UK}: \quad \text{a distinguished element}\ e\ \text{of}\ S\ \text{and a binary operation}\ \ast\ \text{on}\ S.\\ &\textrm{UL}: \quad \text{a unary operation}\ \sim\ \text{on}\ S. \end{align*}$$

The additional axioms for $\vdash $ extending $\rhd $ are

$$\begin{align*}&\textrm{UK}: \quad e \vdash \ \quad a \ast b \vdash a,b, \\ &\textrm{UL}: \quad a, {\mathop{\sim}} a \vdash \ \quad \ \vdash a, {\mathop{\sim}} a, \end{align*}$$

where $a,b\in S$ . The corresponding conservation criteria (Theorem 1) read

$$\begin{align*}&\textrm{UK}: \quad \frac{}{W, e \rhd c} \quad \frac{ W, a \rhd c \quad W, b \rhd c }{ W, a \ast b \rhd c }\\[5pt] &\textrm{UL}: \quad \frac{}{ W, a, {\mathop{\sim}} a \rhd c } \quad \frac{ W, a \rhd c \quad W, {\mathop{\sim}} a \rhd c }{ W \rhd c } \end{align*}$$

where $W\in \operatorname {\textrm {Fin}}(S)$ and $a,b,c\in S$ .Footnote 6 We refer to [Reference Rinaldi, Schuster and Wessel91, Reference Rinaldi, Schuster and Wessel92] for details and references.

4.3 The case of classical logic

Building upon [Reference Ciraulo, Rinaldi and Schuster18], in [Reference Rinaldi and Schuster90] the instances of $\operatorname {\textrm {GKL}}$ in the cases UK and UL have been considered for the following data: S consists of the sentences of a logical language, $\rhd $ stands for deducibility with classical logic, e is absurdity $\bot $ , the operator $\ast $ is disjunction $\lor $ , and ${\mathop {\sim }}$ is negation $\neg $ . While the models of $\rhd $ are the stable theories of $\rhd $ , the models of $\vdash $ are the complete consistent theories in S. Hence $\operatorname {\textrm {GKL}}$ is Lindenbaum’s Lemma [Reference Tarski106], and Con is provable but little interesting, simply because $\rhd $ is classical logic already. Let’s try to get more by relativising $\rhd $ .

Now let $\rhd $ denote (deducibility in) an intermediate logic in a propositional language S; whence the models of $\rhd $ are the theories of $\rhd $ .Footnote 7 Let $\vdash $ extend $\rhd $ with the following additional axioms:

$$\begin{align*}\bot\vdash\quad \qquad \vdash {\varphi}, \lnot {\varphi}\quad ({\varphi} \in S). \end{align*}$$

The models of $\vdash $ are exactly the complete consistent theories of $\rhd $ , and the corresponding conservation criteria (Theorem 1) read

$$ \begin{align*} \frac{}{\Gamma,\bot\rhd \psi}\qquad \frac{\Gamma, {\varphi}\rhd \psi \quad \Gamma, \lnot{\varphi}\rhd \psi}{\Gamma \rhd \psi} \end{align*} $$

with $\Gamma \in \operatorname {\textrm {Fin}}(S)$ and ${\varphi }, \psi \in S$ . While the first criterion holds for any given intermediate logic $\rhd $ , the second one amounts to $\rhd $ satisfying $\rhd {\varphi }\lor \lnot {\varphi }$ , which is to say that $\rhd $ be classical logic. Hence $\operatorname {\textrm {Con}}$ in this case simply means that conservatively adding $\vdash {\varphi }, \lnot {\varphi }$ is equivalent to requiring $\rhd {\varphi }\lor \lnot {\varphi }$ . This of course is well known and of relatively little interest either. Can’t we do better?

5 Jacobson radicals

5.1 The Jacobson radical in algebra

Let $S=R$ be a commutative ring with $1$ , and let $\rhd $ stand for generation in R, i.e., $U\rhd a$ means that a is a linear combination with coefficients from R of the elements of U. A model of $\rhd $ is nothing but an ideal of R, i.e., a subset closed under linear combination. An ideal J of R is

  1. proper if $1\notin J$ , which is tantamount to $J\neq R$ and

  2. complete if modulo J any given $r\in R$ is either 0 or invertible, that is,

    (4) $$ \begin{align} \forall r\in R(J\ni r \lor J,r\rhd 1). \end{align} $$

Every proper complete ideal is a maximal ideal, i.e., maximal among the proper ideals, and vice versa in $\operatorname {\textbf {ZF}}$ . With the current notation, the Jacobson radical of an ideal J can be defined as

(5) $$ \begin{align} \operatorname{\textrm{Jac}}(J)=\{a\in R: \forall b\in R\,(a,b\rhd 1\Rightarrow J,b\rhd 1)\}. \end{align} $$

We thus carry over to commutative rings the first-order definition of the Jacobson radical for distributive lattices [Reference Blass10, Reference Coquand, Lombardi and Quitté28, Reference Johnstone58] rather than using the more common one for commutative rings present e.g., in [Reference Lombardi and Quitté64]. The latter reads

(6) $$ \begin{align} \operatorname{\textrm{Jac}}(J)=\{a\in R: \forall b\in R\,\exists c\in R\,(1-(1-ab)c\in J)\}, \end{align} $$

which is to say that any given $a\in R$ belongs to $\operatorname {\textrm {Jac}}(R)$ precisely when $1-ab$ is invertible modulo J for every $b\in R$ . We give precedence to (5) over (6) because the former, unlike the latter, can be transferred to logic without further ado (Section 5.2).

Just as (6), the first-order definition we employ (5) is anyway equivalent in $\operatorname {\textbf {ZFC}}$ to the following more customary second-order characterisation of the Jacobson radical [Reference Jacobson57]. Although the proof is of course similar to the one with (6) in place of (5) and for maximal rather than complete ideals, see e.g., [Reference Lombardi and Quitté64], we detail this one because it carries over to logic (Proposition 2).

Proposition 1 $\operatorname {\textbf {ZFC}}$

For every ideal J of a commutative ring R,

$$\begin{align*}\bigcap\operatorname{Com}_J(R)= \operatorname{\mathop{Jac}} (J) \end{align*}$$

where $\operatorname {Com}_J(R)$ consists of the complete ideals $\mathfrak {c}$ in R with $J\subseteq {\mathfrak {c}}$ .

Proof Let $a\in \operatorname {\textrm {Jac}}(J)$ , and let $\mathfrak {c}$ be a complete ideal such that $\mathfrak {c}\supseteq J$ . Either $\mathfrak {c}\ni a$ or $\mathfrak {c},a\rhd 1$ . In the former case we are done. In the latter case there is $b\in R$ such that $\mathfrak {c}\rhd b$ (in particular, $b\in \mathfrak {c}$ ) and $a,b\rhd 1$ . Since $a\in \operatorname {\textrm {Jac}}(J)$ , we get $J,b\rhd 1$ . As $J\subseteq \mathfrak {c}$ and $b\in \mathfrak {c}$ , this implies $\mathfrak {c}\rhd 1$ . Hence $\mathfrak {c}=R$ , by which again $\mathfrak {c}\ni a$ .

Conversely, if $a\notin \operatorname {\textrm {Jac}}(J)$ , then there is $b\in R$ for which $a,b\rhd 1$ holds but $J,b\rhd 1$ fails, and thus $(J,b)^\rhd $ lacks a. Zorn’s Lemma yields an ideal $\mathfrak {c}$ maximal among the ones that contain $(J,b)^\rhd $ yet miss a. Any such $\mathfrak {c}$ is complete: if $\mathfrak {c}\not \ni b$ , then $\mathfrak {c},b\rhd a$ by maximality, and thus $\mathfrak {c}, b\rhd 1$ by $a,b\rhd 1$ .

It obviously is irrelevant whether the intersection ranges over the only improper ideal R as well.

5.2 The Jacobson radical in logic

Let again $\rhd $ stand for (deducibility in) an intermediate logic in a propositional language S. That a (consistent) theory T of $\rhd $ be complete can equivalently be put as

(7) $$ \begin{align} \forall {\varphi}\in S(T\ni {\varphi}\lor T, {\varphi} \rhd \bot). \end{align} $$

This move makes fully evident the analogy between complete ideals (4) and complete theories (7), which rests upon the following brief (and necessarily incomprehensive) dictionary:

With this at hand we translate (5) into a definition of the Jacobson radical of a theory T:

$$ \begin{align*} \operatorname{\textrm{Jac}}(T)&=\{\alpha\in S: \forall \beta\in S\,(\alpha,\beta\rhd \bot\Rightarrow T,\beta\rhd \bot)\}. \end{align*} $$

This is obviously equivalent to the following characterisation:

$$ \begin{align*} \operatorname{\textrm{Jac}}(T)=\{\alpha\in S\colon\forall\beta\in S(\alpha\rhd\neg\beta\Rightarrow T\rhd\neg\beta)\}. \end{align*} $$

Mutatis mutandis the proof of Proposition 1 proves what we would like to provisionally call the Intermediate Lindenbaum Lemma:

Proposition 2 $\operatorname {\textbf {ZFC}}$

For every theory T of an intermediate logic $\rhd $ in a propositional languageS,

$$\begin{align*}\operatorname{\mathop{ILL}}\qquad \bigcap\operatorname{Com}_T(S)= \operatorname{\mathop{Jac}} (T), \end{align*}$$

where $\operatorname {Com}_T(S)$ consists of the complete (consistent) theories C in S with $T\subseteq C$ .

As for Proposition 1, it is irrelevant whether the intersection includes the only inconsistent theory S.

Since every complete theory is stable (Lemma 3), the left-hand side of $\operatorname {\textrm {ILL}}$ is as for the original Lindenbaum Lemma [Reference Tarski106] in the form

$$ \begin{align*} \bigcap\operatorname{Com}_T(S) = T, \end{align*} $$

for every stable theory T in S. Hence the left-hand side of $\operatorname {\textrm {ILL}}$ equals in $\operatorname {\textbf {ZFC}}$ the classical deductive closure of T. What about the right-hand side of $\operatorname {\textrm {ILL}}$ ?

Proposition 3. For every theory T of an intermediate logic $\rhd $ in a propositional language S,

$$\begin{align*}\operatorname{\mathop{Jac}}(T)=\{\alpha\in S: T\ni \lnot\lnot \alpha\}. \end{align*}$$

Proof Let $\alpha \in \operatorname {\textrm {Jac}}(T)$ . Since $\alpha \rhd \neg \neg \alpha $ , we get $T\rhd \neg \neg \alpha $ . Conversely, let $\alpha \in S$ be such that $T\ni \lnot \lnot \alpha $ . If $\beta \in S$ is such that $\alpha \rhd \neg \beta $ , then $\neg \neg \alpha \rhd \neg \beta $ and thus $T\rhd \neg \beta $ .

With Lemma 3 at hand we obtain the following:

Corollary 1. $\operatorname {\mathop {Jac}}(T)$ is the least stable theory of $\rhd $ which contains the given theory T of $\rhd $ ; in other words, $\operatorname {\mathop {Jac}}(T)$ equals the deductive closure of T with respect to classical logic.

So $\operatorname {\textrm {ILL}}$ for any intermediate logic $\rhd $ whatsoever is nothing but the original Lindenbaum Lemma!

Now let $\rhd $ be intuitionistic logic ${\rhd _i}$ , and write ${\rhd _c}$ for classical logic, always in the given propositional language S. In this case and by the above, Lindenbaum’s Lemma in the form of $\operatorname {\textrm {ILL}}$ is the semantics of Glivenko’s Theorem [Reference Glivenko46], which in turn is well known as purely syntactical:

Theorem 2 Glivenko 1929

Let S be a propositional language. For all $\Gamma \in \operatorname {\mathop {Fin}}(S)$ and ${\varphi }\in S$ ,

$$\begin{align*}\Gamma{\rhd_c} {\varphi}\Rightarrow \Gamma{\rhd_i} \lnot\lnot{\varphi}. \end{align*}$$

For example, this follows from Corollary 1. We hasten to add that the latter rests upon Lemmas 1 and 2, which of course are the main ingredients of a very common proof of Glivenko’s theorem. Recent literature about Glivenko’s Theorem includes [Reference Espíndola36, Reference Fellin and Schuster39, Reference Galatos and Ono42, Reference Guerrieri and Naibo48, Reference Ishihara and Schwichtenberg56, Reference Litak, Polzer and Rabenstein60, Reference Negri74, Reference Negri75, Reference Ono82, Reference Pereira and Haeusler84].Footnote 8

6 Glivenko’s theorem as syntactical conservation

Once more let ${\rhd _i}$ and ${\rhd _c}$ stand for intuitionistic and classical logic in a propositional language S. For $\Gamma , \Delta \in \operatorname {\textrm {Fin}}(S)$ and ${\varphi }\in S$ , set

$$\begin{align*}\Gamma{\rhd_g} {\varphi}\equiv\Gamma{\rhd_i} \lnot\lnot{\varphi}\qquad \text{and}\qquad \Gamma\vdash_c \Delta\equiv \Gamma{\rhd_c} \bigvee \Delta, \end{align*}$$

which defines a single- and a multi-conclusion entailment relation, respectively. Of course the trace of $\vdash _c$ is nothing but ${\rhd _c}$ ; so Glivenko’s Theorem (Theorem 2) can be rephrased as the following syntactical conservation:

Theorem 3. The extension $\vdash _c$ of ${\rhd _g}$ is conservative, that is,

$$\begin{align*}\operatorname{\mathop{Gli}}\qquad \Gamma{\rhd_c} {\varphi}\Rightarrow \Gamma{\rhd_g} {\varphi} , \end{align*}$$

for all $\Gamma \in \operatorname {\mathop {Fin}}(S)$ and ${\varphi }\in S$ .

To see how Theorem 1 applies in this context, we now prove Theorem 3 in detail. Clearly this proof will otherwise have the main ingredients of any proof of Glivenko’s Theorem (Theorem 2). By Lemma 2, $\vdash _c$ extends ${\rhd _i}$ , and thus ${\rhd _g}$ , with the following additional axioms:

(8) $$ \begin{align} \bot\vdash_c\quad \qquad \vdash_c {\varphi}, \lnot {\varphi}\quad ({\varphi} \in S). \end{align} $$

The corresponding conservation criteria (3) read

(9) $$ \begin{align} \frac{}{\Gamma,\bot{\rhd_g} \psi}\qquad \frac{\Gamma, {\varphi}{\rhd_g} \psi \quad \Gamma, \lnot{\varphi}{\rhd_g} \psi}{\Gamma {\rhd_g} \psi} \end{align} $$

with $\Gamma \in \operatorname {\textrm {Fin}}(S)$ and ${\varphi }, \psi \in S$ .

To prove Theorem 3, in view of Theorem 1 it thus is (necessary and) sufficient to verify (9). Writing ${\rhd _i} \lnot \lnot $ for ${\rhd _g}$ this goes as follows. Needless to say, $\Gamma ,\bot {\rhd _i} \lnot \lnot \psi $ . If both $\Gamma ,{\varphi }{\rhd _i} \lnot \lnot \psi $ and $\Gamma ,\neg {\varphi }{\rhd _i} \lnot \lnot \psi $ , then $\Gamma ,{\varphi }\lor \neg {\varphi }{\rhd _i} \lnot \lnot \psi $ by disjunction elimination. By Lemma 1 we get $\Gamma ,\neg \neg ({\varphi }\lor \neg {\varphi }){\rhd _i} \lnot \lnot \psi $ and thus $\Gamma {\rhd _i} \lnot \lnot \psi $ as desired.

As the models of $\vdash _c$ are exactly the complete consistent theories, ILL for $T\equiv \Gamma ^{{\rhd _i}}$ with $\Gamma \in \operatorname {\textrm {Fin}}(S)$ is to $\operatorname {\textrm {Gli}}$ just as $\operatorname {\textrm {GKL}}$ is to $\operatorname {\textrm {Con}}$ for $\vdash \,\equiv \,\vdash _c$ and $\rhd \equiv {\rhd _g}$ . Although $\vdash _c$ equally extends ${\rhd _i}$ with the same additional axioms (8), and the first conservation criterion of (9) also holds for ${\rhd _i}$ in place of ${\rhd _g}$ , this of course is not the case in general for the second one, e.g., if $\psi \equiv {\varphi }\lor \lnot {\varphi }$ .

We conclude by a relativised version of Glivenko’s Theorem as syntactical conservation. Let $\Gamma \subseteq \operatorname {\textrm {Fin}}(S)$ and $\psi \in S$ . With $\operatorname {\textrm {TND}}_0(\Gamma ,\psi )$ as in Lemma 2, for every propositional variable ${\varphi }\in S$ consider the conservation criterion from Theorem 1 for $\vdash _c {\varphi },\lnot {\varphi }$ over ${\rhd _i}$ :

$$ \begin{align*} \operatorname{Cri}_{\varphi}(\Gamma,\psi):&\quad\kern-6pt \frac{\Gamma,\Delta, {\varphi}{\rhd_i} \psi \quad \Gamma,\Delta, \lnot{\varphi}{\rhd_i} \psi}{\Gamma,\Delta {\rhd_i} \psi} \kern-6pt\quad \text{or, equivalently,}\quad \kern-6pt \frac{\Gamma, \Delta,{\varphi}\lor \lnot{\varphi}{\rhd_i} \psi}{\Gamma,\Delta {\rhd_i} \psi}\\ & \quad \text{for all finite subsets}\ \Delta\ \text{of}\ \operatorname{\textrm{TND}}_0(\Gamma,\psi). \end{align*} $$

Proposition 4. For arbitrary but fixed $\Gamma \subseteq \operatorname {\mathop {Fin}}(S)$ and $\psi \in S$ , the following items are equivalent:

  1. 1. $\operatorname {Cri}_{\varphi }(\Gamma ,\psi )$ for all propositional variables ${\varphi }\in S$ occurring in $\Gamma $ or $\psi $ .

  2. 2. ${\Gamma {\rhd _c} \psi }\Rightarrow {\Gamma {\rhd _i} \psi }$ , i.e., $\vdash _c$ is conservative over ${\rhd _i}$ for the given $\Gamma $ and $\psi $ .

While the first implies the second item by Lemma 2 and Theorem 1, the converse is evident.

Glivenko’s Theorem 2 is the case of the second item in which $\Gamma $ is arbitrary but $\psi $ is a negated formula, in which case the first item obtains by Lemma 1. Other cases include the one in which $\Gamma \cup \{\psi \}$ is made of negative formulas only; see e.g., [Reference David, Nour and Raffalli34, Reference Troelstra and Schwichtenberg107, Reference Troelstra and van Dalen108].

7 Complements

We briefly review some related observations recently made about double negation [Reference Fellin and Schuster39] in the more general context of a nucleus j in place of $\neg \neg $ .

First, for any given intermediate propositional logic $\rhd $ the following are equivalent:

  1. A. $\Gamma \vdash _c{\varphi }\Rightarrow \Gamma \rhd \neg \neg {\varphi }$ for all $\Gamma ,{\varphi }$ and

  2. B. ${\varphi }\to \neg \neg \psi \rhd \neg \neg ({\varphi }\to \psi )$ for all ${\varphi },\psi $ .

Now B is well-known to hold whenever $\rhd $ is intuitionistic logic $\rhd _i$ (see, e.g., [Reference van Dalen111, Lemma 6.2.2]), in which case A becomes Glivenko’s theorem. A posteriori B holds for any intermediate logic $\rhd $ whatsover, as any such $\rhd $ extends $\rhd _i$ .

Next, if $\rhd $ is an intermediate predicate logic, then A is equivalent to B in conjunction with the double negation shift for $\rhd $ :

  1. C. $\forall x\neg \neg {\varphi }\rhd \neg \neg \forall x{\varphi }$ for all formulae ${\varphi }$ .

Again if $\rhd $ is intuitionistic logic $\rhd _i$ , this yields Gödel’s extension of Glivenko’s theorem [37, Reference Gödel47].

Now C trivially holds for any existential logic, i.e., without $\forall $ altogether, for which A with $\rhd _i$ as $\rhd $ is [Reference Troelstra and Schwichtenberg107, Corollary to Proposition 2.3.8]. For A to hold it suffices to refrain from using the $\forall $ –introduction rule or right rule R $\forall $ , which in fact is the only rule that can cause issues in this setting [Reference Fellin and Schuster39]. To be able to avoid R $\forall $ it is enough that the sequent under consideration have no positive occurrences of $\forall $ , because derivations of such sequents by classical logic can be cleared from that rule (see, e.g., [Reference Negri74]).

Acknowledgments

The present study was carried out within the projects “A New Dawn of Intuitionism: Mathematical and Philosophical Advances” (ID 60842) funded by the John Templeton Foundation, and “Reducing complexity in algebra, logic, combinatorics—REDCOM” belonging to the programme “Ricerca Scientifica di Eccellenza 2018” of the Fondazione Cariverona. The opinions expressed in this paper are those of the authors and do not necessarily reflect the views of those foundations. An early version of the manuscript was conceived when the authors were visiting the Hausdorff Research Institute for Mathematics at Universität Bonn in 2018 during the Trimester Program “Types, Sets and Constructions.” Useful hints by Sara Negri are gratefully acknowledged. Schuster was glad for the feedback he received at the Scuola estiva di logica 2019 in Gargnano.

Footnotes

A Contribution to Ninety Years of Glivenko’s Theorem

This note emerged from the the first and third author’s M.Sc and Ph.D thesis, respectively [38, 113]. For a preprint superseded by the present note see [40].

1 For example, we deviate from the terminology prevalent in constructive mathematics and set theory [Reference Aczel and Rathjen4, Reference Aczel and Rathjen5, Reference Bishop8, Reference Bishop and Bridges9, Reference Lombardi and Quitté64, Reference Mines, Richman and Ruitenburg69]: to reserve the term ‘finite’ to sets which are in bijection with $\{1,\ldots ,n\}$ for a necessarily unique $n\ge 0$ . Those exactly are the sets which are finite in our sense and are discrete too, i.e., have decidable equality [Reference Mines, Richman and Ruitenburg69].

2 This is from where we have taken the symbol $\rhd $ , used also [Reference Cintula and Carles16, Reference Wang and Cintula112] to denote a ‘consecution’ [Reference Restall88].

3 David Binder has kindly hinted us at Popper’s work.

4 Stefan Neuwirth has kindly pointed this out to us.

5 The proof of [Reference Rinaldi, Schuster and Wessel92, Proposition 4] goes equally through with $\operatorname {\textrm {TCT}} $ in place of full $\textrm {CT}$ .

6 The criteria for UK have occurred [Reference Rinaldi and Schuster90] as ‘e is convincing for $\rhd $ ’ and ‘ $\rhd $ satisfies Encoding’.

7 For the related covering of formulas [Reference Coquand, Sadocco, Sambin and Smith30, Reference Sambin96], the saturated sets rather are the complements of the theories.

8 This list is by no means meant exhaustive.

References

Aczel, P., The type theoretic interpretation of constructive set theory , Logic Colloquium ’77 , Studies in Logic and the Foundations of Mathematics, vol. 96, North-Holland, Amsterdam, 1978, pp. 5566.10.1016/S0049-237X(08)71989-XCrossRefGoogle Scholar
Aczel, P., The type theoretic interpretation of constructive set theory: Choice principles , The L. E. J. Brouwer Centenary Symposium , Studies in Logic and the Foundations of Mathematics, vol. 110, North-Holland, Amsterdam, 1982, pp. 140.Google Scholar
Aczel, P., The type theoretic interpretation of constructive set theory: Inductive definitions , Logic, Methodology and Philosophy of Science, VII , Studies in Logic and the Foundations of Mathematics, vol. 114, North-Holland, Amsterdam, 1986, pp. 1749.Google Scholar
Aczel, P. and Rathjen, M., Notes on constructive set theory, Technical report no. 40, Institut Mittag–Leffler, 2000.Google Scholar
Aczel, P. and Rathjen, M., Constructive set theory, Book draft, 2010. Available at https://www1.maths.leeds.ac.uk/~rathjen/book.pdf (accessed 24 August, 2021).Google Scholar
Avron, A., Simple consequence relations . Information and Computation , vol. 92 (1991), pp. 105139.10.1016/0890-5401(91)90023-UCrossRefGoogle Scholar
Béziau, J.-Y., Les axiomes de Tarski , La philosophie en Pologne 1919–1939 (R. Pouivet and M. Resbuschi, editors), Librairie Philosophique J. VRIN, Paris, 2006.Google Scholar
Bishop, E., Foundations of Constructive Analysis , McGraw-Hill, New York, 1967.Google Scholar
Bishop, E. and Bridges, D., Constructive Analysis , Springer, Berlin–Heidelberg, 1985.10.1007/978-3-642-61667-9CrossRefGoogle Scholar
Blass, A., Prime ideals yield almost maximal ideals . Fundamenta Mathematicae , vol. 127 (1987), no. 1, pp. 5766.10.4064/fm-127-1-57-66CrossRefGoogle Scholar
Brouwer, L. E. J., De onbetrouwbaarheid der logische principes , Tijdschrift voor Wijsbegeerte , vol. 2 (1908), pp. 152158.Google Scholar
Brouwer, L. E. J., Intuitionistische Zerlegung mathematischer Grundbegriffe , Jahresbericht der Deutschen Mathematiker-Vereinigung , vol. 33 (1925), pp. 251256.Google Scholar
Carnap, R., Formalization of Logic , Harvard University Press, Cambridge, 1943.Google Scholar
Cederquist, J. and Coquand, T., Entailment relations and distributive lattices , Logic Colloquium ’98 (S. R. Buss, P. Hájek, and P. Pudlák, editors), Lecture Notes in Logic, vol. 13, A. K. Peters, Natick, 2000, pp. 127139.Google Scholar
Cederquist, J., Coquand, T., and Negri, S., The Hahn–Banach theorem in type theory , Twenty-Five Years of Constructive Type Theory (G. Sambin and J. M. Smith, editors), Oxford Logic Guides, vol. 36, Oxford University Press, New York, 1998, pp. 5772.Google Scholar
Cintula, P. and Carles, N., The proof by cases property and its variants in structural consequence relations . Studia Logica , vol. 101 (2013), no. 4, pp. 713747.10.1007/s11225-013-9496-1CrossRefGoogle Scholar
Ciraulo, F., Maietti, M. E., and Sambin, G., Convergence in formal topology: A unifying notion . Journal of Logic and Analysis , vol. 5 (2013), no. 2, pp. 145.10.4115/jla.2013.5.2CrossRefGoogle Scholar
Ciraulo, F., Rinaldi, D., and Schuster, P., Lindenbaum’s lemma via open induction , Advances in Proof Theory (R. Kahle, T. Strahm, and T. Studer, editors), Progress in Computer Science and Applied Logic, vol. 28, Springer, Cham, 2016, pp. 6577.10.1007/978-3-319-29198-7_3CrossRefGoogle Scholar
Ciraulo, F. and Sambin, G., Finitary formal topologies and Stone’s representation theorem . Theoretical Computer Science , vol. 405 (2008), nos. 1–2, pp. 1123.10.1016/j.tcs.2008.06.020CrossRefGoogle Scholar
Coquand, T., A direct proof of the localic Hahn–Banach theorem, 2000. Available at http://www.cse.chalmers.se/~coquand/formal.html (accessed 24 August, 2021).Google Scholar
Coquand, T., Lewis Carroll, Gentzen and entailment relations, 2000. Available at http://www.cse.chalmers.se/~coquand/formal.html.Google Scholar
Coquand, T., About Stone’s notion of spectrum . Journal of Pure and Applied Algebra , vol. 197 (2005), nos. 1–3, pp. 141158.10.1016/j.jpaa.2004.08.024CrossRefGoogle Scholar
Coquand, T., Geometric Hahn–Banach theorem . Mathematical Proceedings of the Cambridge Philosophical Society , vol. 140 (2006), pp. 313315.10.1017/S0305004105008935CrossRefGoogle Scholar
Coquand, T., Space of valuations . Annals of Pure and Applied Logic , vol. 157 (2009), 97109.10.1016/j.apal.2008.09.003CrossRefGoogle Scholar
Coquand, T. and Lombardi, H., Hidden constructions in abstract algebra: Krull dimension of distributive lattices and commutative rings , Commutative Ring Theory and Applications (M. Fontana, S.-E. Kabbaj, and S. Wiegand, editors), Lecture Notes in Pure and Applied Mathematics, vol. 231, Addison-Wesley, Reading, 2002, pp. 477499.Google Scholar
Coquand, T. and Lombardi, H., A logical approach to abstract algebra . Mathematical Structures in Computer Science , vol. 16 (2006), pp. 885900.10.1017/S0960129506005627CrossRefGoogle Scholar
Coquand, T., Lombardi, H., and Neuwirth, S., Lattice-ordered groups generated by an ordered group and regular systems of ideals . The Rocky Mountain Journal of Mathematics , vol. 49 (2019), no. 5, pp. 14491489.10.1216/RMJ-2019-49-5-1449CrossRefGoogle Scholar
Coquand, T., Lombardi, H., and Quitté, C., Dimension de Heitmann des treillis distributifs et des anneaux commutatifs . Publications Mathématiques de Besançon: Algèbre et Théorie des Nombres , 2006, pp. 57100.10.5802/pmb.a-113CrossRefGoogle Scholar
Coquand, T. and Persson, H., Valuations and Dedekind’s Prague theorem . Journal of Pure and Applied Algebra , vol. 155 (2001), nos. 2–3, pp. 121129.10.1016/S0022-4049(99)00095-XCrossRefGoogle Scholar
Coquand, T., Sadocco, S., Sambin, G., and Smith, J. M., Formal topologies on the set of first-order formulae . The Journal of Symbolic Logic , vol. 65 (2000), no. 3, pp. 11831192.10.2307/2586694CrossRefGoogle Scholar
Coquand, T., Sambin, G., Smith, J., and Valentini, S., Inductively generated formal topologies . Annals of Pure and Applied Logic , vol. 124 (2003), pp. 71106.10.1016/S0168-0072(03)00052-6CrossRefGoogle Scholar
Coquand, T. and Zhang, G.-Q., Sequents, frames, and completeness , Computer Science Logic (P. G. Clote and H. Schwichtenberg, editors), Lecture Notes in Computer Science, vol. 1862, Springer, Berlin, 2000, pp. 277291.10.1007/3-540-44622-2_18CrossRefGoogle Scholar
Coste, M., Lombardi, H., and Roy, M.-F., Dynamical method in algebra: Effective Nullstellensätze . Annals of Pure and Applied Logic , vol. 111 (2001), no. 3, pp. 203256.10.1016/S0168-0072(01)00026-4CrossRefGoogle Scholar
David, R., Nour, K., and Raffalli, C., Introduction à la Logique. Théorie de la démonstration , second ed., Dunod, Paris, 2003.Google Scholar
Došen, K., On passing from singular to plural consequences , Logic at Work: Essays Dedicated to the Memory of Helena Rasiowa (E. Orlowska, editor), Studies in Fuzziness and Soft Computing, vol. 24, Physica, Heidelberg, 1999, pp. 533547.Google Scholar
Espíndola, C., A short proof of Glivenko theorems for intermediate predicate logics . Archive for Mathematical Logic , vol. 52 (2013), nos. 7–8, pp. 823826.10.1007/s00153-013-0346-7CrossRefGoogle Scholar
Encyclopedia of Mathematics, Intermediate logic, 2016. Available at http://www.encyclopediaofmath.org/index.php?title=Intermediate_logic&oldid=39747 (accessed 13 November, 2016).Google Scholar
Fellin, G., The Jacobson Radical: From Algebra to Logic , Master’s thesis, Università di Verona, Dipartimento di Informatica, Verona, 2018.Google Scholar
Fellin, G. and Schuster, P., A general Glivenko–Gödel theorem for nuclei , Proceedings of the 37th Conference on the Mathematical Foundations of Programming Semantics, MFPS 2021 (A. Sokolova, editor), Electronic Notes in Theoretical Computer Science, Elsevier, 2021.Google Scholar
Fellin, G., Schuster, P., and Wessel, D., The Jacobson radical of a propositional theory , Proof-Theoretic Semantics: Assessment and Future Perspectives (T. Piecha and P. Schroeder-Heister, editors), University of Tübingen, 2019, pp. 287299, http://doi.org/10.15496/publikation-35319.Google Scholar
Gabbay, D. M., Semantical Investigations in Heyting’s Intuitionistic Logic , Synthese Library, vol. 148, D. Reidel Publishing Co., Dordrecht–Boston, 1981.10.1007/978-94-017-2977-2CrossRefGoogle Scholar
Galatos, N. and Ono, H., Glivenko theorems for substructural logics over FL . The Journal of Symbolic Logic , vol. 71 (2006), no. 4, pp. 13531384.10.2178/jsl/1164060460CrossRefGoogle Scholar
Gentzen, G., Untersuchungen über das logische Schließen I . Mathematische Zeitschrift , vol. 39 (1934), pp. 176210.10.1007/BF01201353CrossRefGoogle Scholar
Gentzen, G., Untersuchungen über das logische Schließen II . Mathematische Zeitschrift , vol. 39 (1934), pp. 405431.10.1007/BF01201363CrossRefGoogle Scholar
Glivenko, V., Sur la logique de M. Brouwer . Academie Royale des Sciences des Lettres et des Beaux-Arts de Belgique. Bulletin de la Classe des Sciences Cinquième Série , vol. 14 (1928), pp. 225228.Google Scholar
Glivenko, V., Sur quelques points de la Logique de M. Brouwer . Academie Royale des Sciences des Lettres et des Beaux-Arts de Belgique. Bulletin de la Classe des Sciences Cinquième Série , vol. 15 (1929), 183188.Google Scholar
Gödel, K., Über eine bisher noch nicht benützte Erweiterung des finiten Standpunktes . Dialectica , vol. 12 (1958), pp. 280287.10.1111/j.1746-8361.1958.tb01464.xCrossRefGoogle Scholar
Guerrieri, G. and Naibo, A., Postponement of $\mathsf{raa}$ and Glivenko’s theorem, revisited . Studia Logica , vol. 107 (2019), no. 1, pp. 109144.10.1007/s11225-017-9781-5CrossRefGoogle Scholar
Hertz, P., Über Axiomensysteme für beliebige Satzsysteme. I. Teil. Sätze ersten Grades . Mathematische Annalen , vol. 87 (1922), no. 3, pp. 246269.10.1007/BF01459067CrossRefGoogle Scholar
Hertz, P., Über Axiomensysteme für beliebige Satzsysteme. II. Teil. Sätze höheren Grades . Mathematische Annalen , vol. 89 (1923), no. 1, pp. 76102.10.1007/BF01448090CrossRefGoogle Scholar
Hertz, P., Über Axiomensysteme für beliebige Satzsysteme . Mathematische Annalen , vol. 101 (1929), no. 1, pp. 457514.10.1007/BF01454856CrossRefGoogle Scholar
Humberstone, L., On a conservative extension argument of Dana Scott . Logic Journal of the IGPL , vol. 19 (2011), pp. 241288.10.1093/jigpal/jzq046CrossRefGoogle Scholar
Humberstone, L., Dana Scott’s work with generalized consequence relations , Universal Logic: An Anthology. From Paul Hertz to Dov Gabbay (J.-Y. Béziau, editor), Studies in Universal Logic, Birkhäuser, Basel, 2012, pp. 263279.10.1007/978-3-0346-0145-0_23CrossRefGoogle Scholar
Iemhoff, R., Consequence relations and admissible rules . Journal of Philosophical Logic , vol. 45 (2016), no. 3, pp. 327348.10.1007/s10992-015-9380-8CrossRefGoogle Scholar
Ishihara, H., Classical propositional logic and decidability of variables in intuitionistic propositional logic . Logical Methods in Computer Science , vol. 10 (2014), no. 3, pp. 3:13:7.10.2168/LMCS-10(3:1)2014CrossRefGoogle Scholar
Ishihara, H. and Schwichtenberg, H., Embedding classical in minimal implicational logic . Mathematical Logic Quarterly , vol. 62 (2016), nos. 1–2, pp. 94101.10.1002/malq.201400099CrossRefGoogle Scholar
Jacobson, N., The radical and semi-simplicity for arbitrary rings . American Journal of Mathematics , vol. 67 (1945), no. 2, pp. 300320.10.2307/2371731CrossRefGoogle Scholar
Johnstone, P., Almost maximal ideals . Fundamenta Mathematicae , vol. 123 (1984), no. 3, pp. 197209.10.4064/fm-123-3-197-209CrossRefGoogle Scholar
Legris, J., Paul Hertz and the origins of structural reasoning , Universal Logic: An Anthology. From Paul Hertz to Dov Gabbay (J.-Y. Béziau, editor), Studies in Universal Logic, Birkhäuser, Basel, 2012, pp. 310.10.1007/978-3-0346-0145-0_1CrossRefGoogle Scholar
Litak, T., Polzer, M., and Rabenstein, U., Negative translations and normal modality , 2nd International Conference on Formal Structures for Computation and Deduction , LIPIcs–Leibniz International Proceedings in Informatics, vol. 84, Schloss Dagstuhl – Leibniz-Zentrum für Informatik, Wadern, 2017, Article no. 27, pp. 27:127:18.Google Scholar
Lombardi, H., Le contenu constructif d’un principe local-global avec une application à la structure d’un module projectif de type fini, Publications Mathématiques de Besançon: Algèbre et Théorie des Nombres , 1997, Fascicule 94–95 & 95–96.1997 Google Scholar
Lombardi, H., Relecture constructive de la théorie d’Artin–Schreier . Annals of Pure and Applied Logic , vol. 91 (1998), pp. 5992.10.1016/S0168-0072(97)80700-2CrossRefGoogle Scholar
Lombardi, H., Algèbre dynamique, espaces topologiques sans points et programme de Hilbert . Annals of Pure and Applied Logic , vol. 137 (2006), pp. 256290.10.1016/j.apal.2005.05.023CrossRefGoogle Scholar
Lombardi, H. and Quitté, C., Commutative Algebra: Constructive Methods. Finite Projective Modules , Algebra and Applications, vol. 20, Springer Netherlands, Dordrecht, 2015.10.1007/978-94-017-9944-7CrossRefGoogle Scholar
Lorenzen, P., Über halbgeordnete Gruppen . Mathematische Zeitschrift , vol. 52 (1950), no. 1, pp. 483526.10.1007/BF02230707CrossRefGoogle Scholar
Lorenzen, P., Algebraische und logistische Untersuchungen über freie Verbände . The Journal of Symbolic Logic , vol. 16 (1951), no. 2, pp. 81106.10.2307/2266681CrossRefGoogle Scholar
Lorenzen, P., Teilbarkeitstheorie in Bereichen . Mathematische Zeitschrift , vol. 55 (1952), no. 3, pp. 269275.10.1007/BF01181123CrossRefGoogle Scholar
Lorenzen, P., Die Erweiterung halbgeordneter Gruppen zu Verbandsgruppen . Mathematische Zeitschrift , vol. 58 (1953), no. 1, pp. 1524.10.1007/BF01174126CrossRefGoogle Scholar
Mines, R., Richman, F., and Ruitenburg, W., A Course in Constructive Algebra , Universitext, Springer, New York, 1988.10.1007/978-1-4419-8640-5CrossRefGoogle Scholar
Mulvey, C. J. and Wick-Pelletier, J., The dual locale of a seminormed space . Cahiers de Topologie et Géométrie Différentielle Catégoriques , vol. 23 (1982), no. 1, pp. 7392.Google Scholar
Mulvey, C. J. and Wick-Pelletier, J., A globalization of the Hahn–Banach theorem . Advances in Mathematics , vol. 89 (1991), pp. 159.10.1016/0001-8708(91)90082-ICrossRefGoogle Scholar
Negri, S., Stone bases alias the constructive content of Stone representation , Logic and Algebra (A. Ursini and P. Aglianò, editors), Lecture Notes in Pure and Applied Mathematics, vol. 180, Marcel Dekker, New York, 1996, pp. 617636.Google Scholar
Negri, S., Continuous domains as formal spaces . Mathematical Structures in Computer Science , vol. 12 (2002), no. 1, pp. 1952.10.1017/S0960129501003450CrossRefGoogle Scholar
Negri, S., Proof analysis beyond geometric theories: From rule systems to systems of rules . Journal of Logic and Computation , vol. 26 (2014), no. 2, pp. 513537.10.1093/logcom/exu037CrossRefGoogle Scholar
Negri, S., Glivenko sequent classes in the light of structural proof theory . Archive for Mathematical Logic , vol. 55 (2016), nos. 3–4, pp. 461473.10.1007/s00153-016-0474-yCrossRefGoogle Scholar
Negri, S. and von Plato, J., Cut elimination in the presence of axioms , this Journal, vol. 4 (1998), no. 4, pp. 418435.Google Scholar
Negri, S. and von Plato, J., Structural Proof Theory , Cambridge University Press, Cambridge, 2001.10.1017/CBO9780511527340CrossRefGoogle Scholar
Negri, S. and von Plato, J., Proof Analysis: A Contribution to Hilbert’s Last Problem , Cambridge University Press, Cambridge, 2011.10.1017/CBO9781139003513CrossRefGoogle Scholar
Negri, S. and von Plato, J., Cut elimination in sequent calculi with implicit contraction, with a conjecture on the origin of Gentzen’s altitude line construction , Concepts of Proof in Mathematics, Philosophy, and Computer Science (D. Probst and P. Schuster, editors), Ontos Mathematical Logic, vol. 6, Walter de Gruyter, Berlin, 2016, pp. 269290.10.1515/9781501502620-016CrossRefGoogle Scholar
Negri, S., von Plato, J., and Coquand, T., Proof-theoretical analysis of order relations . Archive for Mathematical Logic , vol. 43 (2004), pp. 297309.10.1007/s00153-003-0209-8CrossRefGoogle Scholar
Neuwirth, S., Lorenzen's theory of divisibility in monoid-preordered sets, preprint, 2018. Available at https://lmb.univ-fcomte.fr/IMG/pdf/lorenzen_s_theory_of_divisibility_in_monoid_preordered_sets.pdf.Google Scholar
Ono, H., Glivenko theorems revisited . Annals of Pure and Applied Logic , vol. 161 (2009), no. 2, pp. 246250.10.1016/j.apal.2009.05.006CrossRefGoogle Scholar
Payette, G. and Schotch, P. K., Remarks on the Scott–Lindenbaum theorem . Studia Logica , vol. 102 (2014), no. 5, pp. 10031020.10.1007/s11225-013-9519-yCrossRefGoogle Scholar
Pereira, L. C. and Haeusler, E. H., On constructive fragments of classical logic , Dag Prawitz on proofs and meaning , Outstanding Contributions to Logic, vol. 7, Springer, Cham, 2015, pp. 281292.Google Scholar
Popper, K. R., On the theory of deduction, Part I. Derivation and its generalizations . Proceedings of the Koninklijke Nederlandse Akademie van Wetenschappen , vol. 51 (1948), no. 2, pp. 173183.Google Scholar
Popper, K. R., On the theory of deduction, Part II. The definitions of classical and intuitionist negation . Proceedings of the Koninklijke Nederlandse Akademie van Wetenschappen , vol. 51 (1948), no. 3, pp. 322331.Google Scholar
Rathjen, M., Generalized inductive definitions in constructive set theory , From Sets and Types to Topology and Analysis: Towards Practicable Foundations for Constructive Mathematics (L. Crosilla and P. Schuster, editors), Oxford Logic Guides, vol. 48, Clarendon Press, Oxford, 2005, Chap. 16.Google Scholar
Restall, G., An Introduction to Substructural Logics , Routledge, London, 2000.10.4324/9780203252642CrossRefGoogle Scholar
Rinaldi, D., Formal Methods in the Theories of Rings and Domains . Doctoral dissertation, Universität München, 2014.Google Scholar
Rinaldi, D. and Schuster, P., A universal Krull–Lindenbaum theorem . Journal of Pure and Applied Algebra , vol. 220 (2016), pp. 32073232.10.1016/j.jpaa.2016.02.011CrossRefGoogle Scholar
Rinaldi, D., Schuster, P., and Wessel, D., Eliminating disjunctions by disjunction elimination . this Journal , vol. 23 (2017), no. 2, pp. 181200.Google Scholar
Rinaldi, D., Schuster, P., and Wessel, D., Eliminating disjunctions by disjunction elimination . Indagationes Mathematicae. New Series , vol. 29 (2018), no. 1, pp. 226259.10.1016/j.indag.2017.09.011CrossRefGoogle Scholar
Rinaldi, D. and Wessel, D., Extension by conservation. Sikorski’s theorem . Logical Methods in Computer Science , vol. 14 (2018), nos. 4:8, pp. 117.Google Scholar
Rinaldi, D. and Wessel, D., Cut elimination for entailment relations . Archive for Mathematical Logic , vol. 58 (2019), nos. 5–6, pp. 605625.10.1007/s00153-018-0653-0CrossRefGoogle Scholar
Sambin, G., Intuitionistic formal spaces—A first communication , Mathematical Logic and Its Applications (D. Skordev, editor), Plenum, New York, 1987, pp. 187204.10.1007/978-1-4613-0897-3_12CrossRefGoogle Scholar
Sambin, G., Pretopologies and completeness proofs . The Journal of Symbolic Logic , vol. 60 (1995), no. 3, pp. 861878.10.2307/2275761CrossRefGoogle Scholar
Sambin, G., Some points in formal topology . Theoretical Computer Science , vol. 305 (2003), nos. 1–3, pp. 347408.10.1016/S0304-3975(02)00704-1CrossRefGoogle Scholar
Sambin, G., The Basic Picture: Structures for Constructive Topology , Oxford Logic Guides, Clarendon Press, Oxford, 2018.Google Scholar
Sandqvist, T., Preservation of structural properties in intuitionistic extensions of an inference relation , this Journal, vol. 24 (2018), no. 3, pp. 291305.Google Scholar
Schlagbauer, K., Schuster, P., and Wessel, D., Der Satz von Hahn–Banach per Disjunktionselimination . Confluentes Mathematici , vol. 11 (2019), no. 1, pp. 7993.10.5802/cml.57CrossRefGoogle Scholar
Scott, D., On engendering an illusion of understanding . The Journal of Philosophy , vol. 68 (1971), pp. 787807.10.2307/2024952CrossRefGoogle Scholar
Scott, D., Completeness and axiomatizability in many-valued logic , Proceedings of the Tarski Symposium (L. Henkin, J. Addison, C. C. Chang, W. Craig, D. Scott, and R. Vaught, editors), Proceedings of Symposia in Pure Mathematics, Vol. XXV, American Mathematical Society, Providence, 1974, pp. 411435.10.1090/pspum/025/0363802CrossRefGoogle Scholar
Scott, D. S., Background to formalization , Truth, Syntax and Modality (H. Leblanc, editor), Studies in Logic and the Foundations of Mathematics, Vol. 68, North-Holland, Amsterdam, 1973, pp. 244273.10.1016/S0049-237X(08)71542-8CrossRefGoogle Scholar
Shoesmith, D. J. and Smiley, T. J., Multiple-Conclusion Logic , Cambridge University Press, Cambridge, 1978.10.1017/CBO9780511565687CrossRefGoogle Scholar
Takeuti, G., Proof Theory , second ed., Studies in Logic and the Foundations of Mathematics, vol. 81, North-Holland Publishing Co., Amsterdam, 1987.Google Scholar
Tarski, A., Fundamentale Begriffe der Methodologie der deduktiven Wissenschaften. I . Monatshefte für Mathematik Physik , vol. 37 (1930), pp. 361404.10.1007/BF01696782CrossRefGoogle Scholar
Troelstra, A. S. and Schwichtenberg, H., Basic Proof Theory , second edition, Cambridge University Press, Cambridge, 2000.10.1017/CBO9781139168717CrossRefGoogle Scholar
Troelstra, A. S. and van Dalen, D., Constructivism in Mathematics: An Introduction , Studies in Logic and the Foundations of Mathematics, vol. I, North-Holland, Amsterdam, 1988.Google Scholar
Troelstra, A. S., On the early history of intuitionistic logic , Mathematical Logic (P. P. Petkov, editor), Plenum Press, New York, 1990, pp. 317.10.1007/978-1-4613-0609-2_1CrossRefGoogle Scholar
van Atten, M. and Sundholm, G., L.E.J. Brouwer’s ‘Unreliability of the Logical Principles’: A new translation, with an introduction . History and Philosophy of Logic , vol. 38 (2017), no. 1, pp. 2447.10.1080/01445340.2016.1210986CrossRefGoogle Scholar
van Dalen, D., Logic and Structure , fifth ed., Universitext, Springer, London, 2013.10.1007/978-1-4471-4558-5CrossRefGoogle Scholar
Wang, S.-m. and Cintula, P., Logics with disjunction and proof by cases . Archive for Mathematical Logic , vol. 47 (2008), no. 5, pp. 435446.10.1007/s00153-008-0088-0CrossRefGoogle Scholar
Wessel, D., Choice, extension, conservation. From transfinite to finite proof methods in abstract algebra , Ph.D. thesis, Università degli Studi di Trento, 2018.Google Scholar
Wessel, D., Ordering groups constructively . Communications in Algebra , vol. 47 (2019), no. 12, pp. 48534873.10.1080/00927872.2018.1477947CrossRefGoogle Scholar
Wessel, D., Point-free spectra of linear spreads , Mathesis Universalis, Computability and Proof (S. Centrone, S. Negri, D. Sarikaya, and P. Schuster, editors), Synthese Library, Springer, Cham, 2019, pp. 353374.10.1007/978-3-030-20447-1_19CrossRefGoogle Scholar
Wessel, D., A note on connected reduced rings . Journal of Commutative Algebra , vol. 13 (2021), no. 4, pp. 583588.10.1216/jca.2021.13.583CrossRefGoogle Scholar
Wójcicki, R., Theory of Logical Calculi: Basic Theory of Consequence Operations , Synthese Library, vol. 199, Kluwer Academic Publishers Group, Dordrecht, 1988.10.1007/978-94-015-6942-2CrossRefGoogle Scholar
Yengui, I., Making the use of maximal ideals constructive . Theoretical Computer Science , vol. 392 (2008), pp. 174178.10.1016/j.tcs.2007.10.011CrossRefGoogle Scholar
Yengui, I., Constructive Commutative Algebra: Projective Modules over Polynomial Rings and Dynamical Gröbner Bases , Lecture Notes in Mathematics, vol. 2138, Springer, Cham, 2015.Google Scholar