Vous êtes sur la page 1sur 73

Abstract Algebraic Logic

An Introductory Chapter

Josep Maria Font

1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
2 Some preliminary issues . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3 Bare algebraizability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
3.1 The equational consequence relative to a class of algebras . . . . . . . . . . . . . 11
3.2 Translating formulas into equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.3 Translating equations into formulas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.4 Putting it all together . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
4 The origins of algebraizability: the Lindenbaum-Tarski process . . . . . . . . . . . . . . . . . . 18
4.1 The process for classical logic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
4.2 The process for algebraizable logics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
4.3 The universal Lindenbaum-Tarski process: matrix semantics . . . . . . . . . . . 23
4.4 Algebraizability and matrix semantics: definability . . . . . . . . . . . . . . . . . . . 27
4.5 Implicative logics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
5 Modes of algebraizability, and non-algebraizability . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
6 Beyond algebraizability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
6.1 The Leibniz hierarchy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
6.2 The general definition of the algebraic counterpart of a logic . . . . . . . . . . . 45
6.3 The Frege hierarchy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
7 Exploiting algebraizability: bridge theorems and transfer theorems . . . . . . . . . . . . . . . 55
8 Algebraizability at a more abstract level . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

1 Introduction

This chapter has been conceived as a brief introduction to (my personal view of)
abstract algebraic logic. It is organized around the central notion of algebraizabil-
ity, with particular emphasis on its connections with the techniques of traditional
algebraic logic and especially with the so-called Lindenbaum-Tarski process. It

Josep Maria Font


Departament of Mathematics and Computer Engineering, Universitat de Barcelona (UB), Gran Via
de les Corts Catalanes 585, E-08007 Barcelona, Spain. e-mail: jmfont@ub.edu

1
2 Josep Maria Font

goes beyond algebraizability to offer a general overview of several classifications


of (sentential) logics—basically, the Leibniz hierarchy and the Frege hierarchy—
which have emerged in recent decades, and to show how the classification of a logic
into any of them provides some knowledge regarding its algebraic or its metalogical
properties. In the final section, a more abstract view of algebraizability is introduced.
When attempting to describe the essential ideas of abstract algebraic logic for
readers who are new to the topic, the first issue to be addressed is: What is ab-
stract algebraic logic? If in general algebraic logic can be described as the branch
of mathematics that studies the connections between logics and their algebra-based
semantics, then abstract algebraic logic can be described as the more general or
abstract side of that study.
The expression algebra-based semantics is used in a deliberately ambiguous and
informal sense.1 It is intended to refer to any kind of semantics where the mathemat-
ical objects where the formulas of the language are interpreted are either just plain
algebras, or algebras endowed with some additional structure (a particular element
or subset, a family of subsets, or an order relation), and where the interpretations are
given by the homomorphisms from the formula algebra to the algebras that underlie
the models.
The study of these connections may have different goals.
• To describe them: It is natural that this was, historically, the first aspect of the
subject to be developed, starting with Boole’s pioneering work (1847). The old-
est connections to be described take the form of completeness theorems; later on
other, stronger connections were discovered, such as the notion of algebraizabil-
ity. The task of finding meaningful descriptions (preferably algebraic descrip-
tions) of the reduced models of a logic and of its algebraic counterpart (in what-
ever sense) also belongs to this aspect of the subject.
• To explain them: This involves developing general theories regarding the connec-
tions and exploring them in many different ways, to obtain a deeper understand-
ing of how and why the different kinds of connections work, of the relations be-
tween them, of how large their domain of application is, etc. Ideally, each general
theory should provide criteria that justify the choice of a specific algebra-based
semantics as the algebraic counterpart of a logic (either for arbitrary logics, or
only for those of a certain class which the theory is intended to explain).
• To exploit them: Once a connection has been described, the most usual way of
exploiting it is to use powerful and well-developed algebraic techniques to prove
properties of the logic; but there are also cases of the reverse process. This oc-
curred at each stage in the evolution of algebraic logic, in different ways. One of
the distinctive features of the abstract algebraic logic way of doing this, in con-
trast to the more traditional way, is that its results concern not just an individual
1 The term algebraic semantics, which has been much used for years in informal comments in the
same vague sense as my “algebra-based semantics”, has acquired a formal meaning in the modern
theory (Definition 3.4). I consider it to be good practice to limit its usage to this strictly technical
sense.
Abstract Algebraic Logic 3

logic, but classes of logics that are treated in a uniform way;2 often these classes
are defined by some abstract characteristic (such as the classes in the hierarchies
reviewed in Section 6) or constitute a set of extensions of a particular logic (as in
the applications of Theorem 4.2 mentioned at the end of Section 7).
Naturally, these three aspects are inextricably intertwined. In particular, the gen-
eral notions need to be tested against many examples, either natural or ad hoc, not
just to obtain properties of particular logics, but in order to gain insight into the
general notions themselves, their relations, their applicability, their scope and their
limits. This empirical work is an important guide for the more abstract work.
The framework where the connections we are interested in here have been found
to be tighter and more fruitful, and where the algebra-based semantics is exclusively
algebraic, is undoubtedly that of algebraizable logics. This justifies devoting more
than half of this chapter to describing that framework, its variants and extensions,
and its roots in the earliest milestone in modern algebraic logic: the construction
of the Lindenbaum-Tarski algebras in order to obtain the first general algebraic
completeness theorems.
After the preliminary Section 2, algebraizability is gradually described in Sec-
tions 3 to 5 by relating it to the progressive generalization of the Lindenbaum-
Tarski construction, starting from its well-known application to classical proposi-
tional logic and extending it to increasingly larger classes of logics. These sections,
which are decidedly introductory and discuss the details of (some of) the key proofs,
should be accessible to any reader with a general background in mathematical logic.
A certain tendency to identify abstract algebraic logic with the theory of alge-
braizability initiated by Blok and Pigozzi has been observed in the literature; how-
ever, this is misguided, as algebraizability is only the core of the subject. More-
over, the full development of the theory of algebraizability itself does indeed need
the more general approach to algebraic logic provided by the classical theory of
matrices (the next kind of algebra-based semantics); the results concerning alge-
braizability that appear in this context form the crowning stage of a very elaborated
framework, where the classical theory is extended in newer and more abstract di-
rections. Some of those directions were developed by Blok and Pigozzi themselves
(the theory of protoalgebraic logics), some earlier on by Czelakowski (the theory
of equivalential logics), and some later on by others (such as Herrmann’s exten-
sion to non-finitary logics, or Raftery’s theory of truth-equational logics). Abstract
algebraic logic has led to the development of tools and techniques together with
a strong and convoluted mathematical construction with deep results that provide
the whole construction with unity. Moreover, it turns out that some less standard
logics do show some algebraic behaviour, but it is not naturally captured by the the-
ory of matrices; this brings generalized matrices (a more complicated algebra-based
semantics) into the picture.
2 This feature has also been highlighted as distinctive of (some of) the modern studies of non-
classical logics, notably of modal logics; see for instance Chagrov and Zakharyashev (1997,
p. 109).
4 Josep Maria Font

The essentials of all these issues, in a brief overview or survey style, are described
in Section 6 (which is anticipated in Section 4.3). In particular, that survey includes a
description of several of the classification criteria3 that have arisen, and which have
given rise to the so-called Leibniz and Frege hierarchies. Section 6.3 also reviews
the main relations between the two hierarchies and mentions a few recent results
and some open problems.
Revealing a general notion behind the Lindenbaum-Tarski process is only part of
the interest in the notion of algebraizability. It is certainly true that the general notion
helped to make it clear that some obscure logics are actually amenable to algebraic
treatment, or to provide a sound foundation for claims of the form “such-and-such
class of algebras is the algebraic counterpart of such-and-such logic”. However, the
importance of the notion also lies in its consequences, that is, in how it can be ap-
plied to obtain a precise formulation of the equivalences, informally called bridge
theorems, that hold between certain metalogical properties (such as the deduction
theorem or the interpolation property) and certain purely algebraic properties (such
as having equationally definable principal congruences or several forms of amal-
gamation); in some cases, these equivalences had previously been reported in the
literature for restricted groups of particular logics. Although most of this would be
the natural continuation of Section 5, it is treated in Section 7, after the exposition of
the hierarchies, because it has been discovered that some of the equivalences men-
tioned are best dealt with within the framework of a class in the Leibniz hierarchy
that is larger than the class of algebraizable logics.
Finally, in Section 8, one of the new directions along which the theory is expand-
ing, a rather abstract one, is just hinted at. It is impossible to describe all the topics
and directions of research that have been pursued in the last two decades. Among
these4 are: categorical abstract algebraic logic (Voutsadakis, 2003); applications to
many-sorted logics and in particular to behavioural logics (Caleiro and Gonçalves,
2013); the algebraization of Gentzen systems, hypersequent systems and other more
general formalisms (Raftery, 2006a); the use of generalized matrices as models of
Gentzen systems, and the interplay with their rôle as generalized models of sen-
tential logics as described in Section 6.2 (Font, Jansana, and Pigozzi, 2001); the
interplay with equality-free model theory (Nurakunov and Stronkowski, 2013); the
implicational hierarchy (Cintula and Noguera, 2010); the algebraic study of log-
ics that preserve degrees of truth (Bou et al., 2009); the algebraization of logics by
classes of algebras of a different signature (Russo, 2013); etc. Something similar
happens when considering applications to the study of particular metalogical prop-
erties, and of particular classes of logics and their algebraic counterparts: the choice
of the topics to be discussed in detail was difficult and often dictated—besides per-
sonal taste—by the space constraints of a chapter of this kind.
This Introduction opens with an attempt to describe what abstract algebraic logic
is. An alternative way of grasping this is to disregard the introductory discussion
3 Looking back at history, it is easy to see that finding useful classifications has been one major
driving force in the development of several areas of mathematics, and of all sciences.
4 The references given here for each case are just pointers to facilitate contact with the topic, rather

than the earliest or the main work on it.


Abstract Algebraic Logic 5

and just to go ahead and read the technical contents of the chapter. I hope that, as a
sample of the topic, it speaks for itself, and sufficiently supports the statement that,
simply: abstract algebraic logic is the algebraic logic of the twenty-first century.

Further reading

Only proofs of basic results in the first sections are given or sketched; detailed proofs
of all the statements can be found in the monumental monograph by Czelakowski
(2001), or in my textbook (Font, 2016); or also in the references given for selected
results. The Handbook of Mathematical Fuzzy Logic contains a long chapter (Cin-
tula and Noguera, 2011) whose aim is, in the author’s own words, “to present a
marriage of Mathematical Fuzzy Logic and (Abstract) Algebraic Logic”; it also
contains proofs and its (more specialized) point of view may appeal particularly to
readers of the present volume. As a matter of fact, the seminal monograph by Blok
and Pigozzi (1989) still makes a good introduction to the core of the subject, as
does the unpublished (but available) Blok and Pigozzi (2001). Some survey papers
that may complement this chapter with different views are Andréka, Németi, and
Sain (2001), Font, Jansana, and Pigozzi (2003) and Raftery (2011b), as well as the
encyclopædia articles by Pigozzi (2001) and by Jansana (2011).
For very precise historical information on technical points, see the notes at the
end of each chapter of Czelakowski (2001); in general, Anellis and Houser (1991),
Blok and Pigozzi (1988), Burris and Legris (2015), Czelakowski and Malinowski
(1985) and Surma (1982) provide a wider historical context.

2 Some preliminary issues

The simplest, neatest kind of algebra-based models are just plain algebras. The no-
tion of algebraizability expresses a relation between a (sentential) logic and a class
of algebras. Before presenting that notion, we have to agree on what we understand
by each of these two components, and this entails fixing some terminology and no-
tations.
Let L be an algebraic language, that is, a similarity type with no relational sym-
bols. The set of terms of L is defined in the usual way, starting from a certain count-
ably infinite set V of variables. In our context, the terms are also called formulas
and constitute an algebra of type L denoted by Fm (or FmL if necessary); it is the
absolutely free algebra of type L generated by V. The algebraic operations of L are
called connectives when their logical function is privileged; the usual such opera-
tions are ¬ , ∧ , ∨ , →, etc. Substitutions of formulas for variables are, algebraically
speaking, the endomorphisms of the formula algebra.
Besides formulas or terms, we can form equations, which are just pairs hα , β i of
formulas, denoted by α ≈ β , for α , β ∈ Fm; and quasi-equations, which are first-
6 Josep Maria Font

order formulas of the form5



i<k αi ≈ βi → α ≈ β . The set of equations, which
V

is Fm × Fm, is denoted by Eq.


There can be no confusion about what an algebra (of a certain type) is. Algebras
are denoted by A, B, etc., and their universes by A, B, etc. Classes of algebras are
always understood to be of the same similarity type, which coincides with that of
the logic under consideration, when it is present.
However, there can be some confusion about what a logic is. It is not uncommon
to consider certain sets of formulas that are closed under certain operations (at least,
substitutions) as “logics”. Let me be clear from the start: algebraizability makes
sense only when a (sentential) logic is understood as a consequence relation; more
precisely:

Definition 2.1. A (sentential) logic of type L is a pair L = hL, `L i where L is an


algebraic language and `L ⊆ P(Fm) × Fm is a relation that satisfies the following
properties, for all Γ ∪ {ϕ} ⊆ Fm:

If ϕ ∈ Γ , then Γ `L ϕ. (I)
If Γ `L ϕ, then ∆ `L ϕ for any ∆ ⊆ Fm such that Γ ⊆ ∆ . (M)
If Γ `L ϕ, then ∆ `L ϕ for any ∆ ⊆ Fm such that ∆ `L ψ for all ψ ∈ Γ . (C)
If Γ `L ϕ, then σ (Γ ) `L σ (ϕ) for every substitution σ . (S)

Ordinary infix notation for the binary relation `L has been used.
The labels are self-explanatory: (I) is for “identity”, (M) is for “monotonicity”,
(C) is for “cut” and (S) stands for structurality or “substitution invariance”. A re-
lation satisfying (I), (M) and (C) is a closure relation. Thus, it is safe to identify a
logic with a substitution-invariant closure relation on the algebra6 of formulas.
The relation `L is called the derivability relation or the consequence relation of
the logic. A theorem of the logic is a formula ϕ such that 0/ `L ϕ (one usually writes
just `L ϕ ); in semantically defined logics, the term tautology is used as a synonym
of “theorem”. A theory is a set of formulas that is closed under consequence; that
is, a set Γ ⊆ Fm such that Γ `L ϕ implies ϕ ∈ Γ . The set of all theories of L is
denoted by ThL .

Thus, in some contexts you have to be precise as to which consequence relation


you are talking about, lest you (or your readers) fall into some unexpected misun-
derstandings. Probably the best known, paradigmatic example of this situation is
that of modal logics based on classical (non-modal) logic. Let us agree that a set
S of formulas (in a language expanding the classical language with a unary con-
nective  representing “necessity”, “provability”, or an analogous modal notion)
is called a normal modal system whenever it contains all instances of theorems of
classical, non-modal logic and all formulas (α → β ) → (α → β ), and is closed
under substitutions (if α ∈ S then σ (α) ∈ S for all substitutions σ ), under modus
5 See the comment on usage of ∧ and → on page 9.
6 Set-theoretically, it is a relation on the set of formulas, but since condition (S) takes the algebraic
structure of formulas into account, it makes sense to say it is a relation on the algebra of formulas.
Abstract Algebraic Logic 7

ponens (if α ∈ S and α → β ∈ S then β ∈ S) and under necessitation (if α ∈ S


then α ∈ S). Then, two consequence relations are naturally associated with S,
both having all the formulas of S as axioms: the local consequence ``S has modus
ponens as its only rule, and the global consequence `gS has both modus ponens and
necessitation as its rules of inference; the terms “local” and “global” refer to the
kind of completeness they satisfy, with respect to classes of Kripke-style frames,
when these are available. It is clear that both consequences have the same set of
theorems, namely S, and that the theories of `gS are just the theories of ``S that are
closed under necessitation. While the two logics have the same algebraic counter-
part,7 they show quite different behaviour in several respects; for instance, as far
as their algebraizability and their classification in the hierarchies of abstract alge-
braic logic are concerned (except for trivial cases, `gS is algebraizable while ``S is
only equivalential, and ``S is fully selfextensional while `gS is not even selfexten-
sional). Thus, when speaking about modal logic in the context of abstract algebraic
logic one has to clarify whether one refers to the local or to the global consequence.
Other logical contexts, where consequence relations do not usually play a prominent
rôle, demand a similar clarification before discussing algebraizability or, in general,
for their study with the tools of abstract algebraic logic.
In some cases, the need to deal with a consequence relation rather than with a set
of formulas may introduce a certain degree of artificiality. For instance, in the do-
main of relevance logic, reaching a consensus to discuss axiomatic systems is easier
when these are intended only to generate the theorems of some logic (i.e., when
only proofs from axioms are considered) than when they are intended to define a
consequence relation (i.e., when proofs from arbitrary assumptions are considered).
This is a price one has to pay for entering the universe of abstract algebraic logic.
A reasonable compromise to reconcile the general theory with everyday practice
in some fields, which is useful when dealing with a large family of particular logics,
is to take the sets of formulas that are the theorems of the axiomatic extensions of
a base logic as the “logics” under study; in this way there is always a consequence
relation associated with each such “logic”, albeit hidden in the background.8
It is not uncommon to find work in the literature that states that the notion of al-
gebraizability, and the whole theory, concerns only syntactically defined logics; this
is an important and most regrettable misunderstanding. In fact, the notion does not
depend on the way the logic (that is, its consequence relation) is presented: it can be
syntactic (defined by an axiomatic system in the Frege-Hilbert style, a Gentzen cal-
culus, a tableaux system, a labelled deductive system, etc.) or semantic (and there is
also a large variety of semantic devices: truth-tables and logical matrices, algebras,
ordered algebras, topological models, relational semantics, etc.); but once defined,
7 This exemplifies why it is imperative that algebra-based semantics that are more general than just
plain algebras be considered in some domains, and hence in the general theory.
8 Two modern treatises where this strategy is explicitly adopted, and where algebraizability plays

some rôle, are Kracht (2007) and Galatos et al. (2007, pp. 72,88).
8 Josep Maria Font

the fact that a logic is or is not algebraizable depends only on the consequence rela-
tion itself, not on its presentation.9
In contrast, in Galatos, Jipsen, Kowalski, and Ono (2007, p. 9), we can read that
“one and the same logic can be algebraizable in one presentation and not algebraiz-
able in another”. This seems to contradict the previous paragraph, but the contra-
diction is only apparent, because the terms “logic” and “presentation” are used in
different senses: in the previous paragraph the term “logic” was used in the techni-
cal sense of Definition 2.1, and by different “presentations” I meant different ways
of defining the consequence relation; in the sentence just quoted, the term “logic”
is used in a liberal, informal way, so that several formal systems of incomparable
frameworks are viewed as different “presentations” (or “formalizations”) of “the
same logic”, in the sense we say, for instance, that classical logic can be formalized
or presented both as a sentential logic (in the technical sense of Definition 2.1) and
as a two-sided, multiple-conclusion sequent calculus.
In this chapter, algebraizability is described primarily as concerning sentential
logics in the sense of Definition 2.1; but in Section 8, we see that it can be applied
to Gentzen systems as well, mutatis mutandis, so that the situation hinted at in the
sentence quoted applies. To mention just a very simple example: the conjunction-
disjunction fragment of classical logic is not algebraizable when it is formalized as a
sentential logic (Font and Verdú, 1991), but it is algebraizable when it is formalized
as a Gentzen system (Rebagliato and Verdú, 1993).
Logics defined by syntactic means are normally finitary; that is, they satisfy, for
all Γ ∪ {ϕ} ⊆ Fm,

Γ `L ϕ if and only if there is a finite Γ0 ⊆ Γ such that Γ0 `L ϕ . (F)

Logics defined semantically need not have this property, and accordingly in this
chapter it is not considered as part of the definition; but notice that many works (in
particular, those by Blok and Pigozzi themselves) do consider it so.
Definition 2.1 establishes a crisp boundary delimiting which formalizations of
logic may be covered by the algebraizability paradigm, and more generally by ab-
stract algebraic logic as evolved from the algebraizability paradigm and related stud-
ies. Before going on, let me say a word about what is not covered by Definition 2.1,
and hence by the algebraizability paradigm; notice that almost every alternative for-
malization of the notion of a logic has been studied algebraically to some degree,
but such studies either do not fit into the framework of this chapter, or have not
attained a comparable degree of sophistication. Technically, logics understood as
Gentzen-style systems of several kinds (one-conclusion, multiple-conclusion, one-
sided, many-sided, etc.) do not fall under Definition 2.1, but as already said, many
9 Quite a different issue is whether one is able or not, or whether it is easy, to prove that a given

logic is algebraizable, or to disprove this. This indeed may depend heavily on its presentation.
Indeed, there are better tools to achieve this for syntactically presented logics (see Theorems 4.1
and 5.3) than for semantically presented ones (see Theorem 4.14); but, more generally, Moraschini
(2016) has proved that the general classification problem, either in the Leibniz hierarchy or in the
Frege hierarchy, is undecidable for axiomatically presented logics (hence a fortiori for arbitrary
logics), while it is decidable for logics defined by a finite set of finite matrices of finite type.
Abstract Algebraic Logic 9

of the basic ideas can be adapted to fit these formalisms, since they define conse-
quence relations on sets of “formulas” of more complicated grammatical structure.
More radical is the difference in the case of “symmetric” consequence relations, and
of non-monotonic logics. But probably the most important large area that may be
missing here is the algebraization of predicate logics. As Blok and Pigozzi (1989,
Appendix C) say: “The problem of algebraizing predicate logic is of a different
character than the problem for propositional logics because the standard deductive
systems for predicate logic are not structural [. . . ]”. In fact, the algebraic study of
predicate logics has a long tradition10 and has produced a vast literature as well
as many remarkable results, some of great technical difficulty; but the approach is
based on another general definition of the notion of a logic, which incorporates a
semantic component. The chapter by Andréka, Németi, and Sain (2001) is a good
guide to and place to start the study of this area.

Some more points on notation

Note the usage of ≈ for formal equations, and of ∧ and → for the first-order con-
junction and implication needed to write quasi-equations; these two symbols should
not be confused with the sentential connectives of conjunction (∧) and implication
(→), which may or may not be present in the language L of a particular sentential
logic under consideration.
ConA is the set of congruences of an algebra A. Hom(A, B) is the set of all
homomorphisms from algebra A to algebra B, and End(A) := Hom(A, A) is the set
of all endomorphisms of A.
Sequences of the form ha1 , . . . , an i or han : n ∈ ωi are denoted by ~a, while “~a ∈
~A” is shorthand for “ai ∈ A for every i”; this is used for variables, formulas and
elements of arbitrary algebras. If δ denotes a formula, writing it as δ (z,~z) expresses
the assumption that the variables occurring in it belong to the set {z, z1 , . . . , zn }, and
that a special rôle is assigned to z (the remaining variables are then referred to
as parameters); this notation is mainly used so as to be able to write δ (α , ~β ) to
denote the result of a substitution (defined by z 7→ α and zi 7→ βi ) on δ (z,~z), and to
describe the interpretations of this formula in a practical way: if h ∈ Hom(Fm, A)
satisfies h(z) = a and h(zi ) = ci , then h(δ ) is also denoted by δ A (a,~c). The same
notation (more familiar in the interpretation of terms in universal algebra and model
theory) is extended to sets of formulas and of equations in the natural way; recall
10 To the best of my knowledge, the term “abstract algebraic logic” first appears in the literature
in the title of Section 5.6 of Henkin, Monk, and Tarski (1985), which is devoted to a general study
of the connections between theories of classical first-order logic and classes of cylindric algebras.
A related approach, through polyadic algebras, is offered by Halmos (1962), who is generally
credited for having coined the term “algebraic logic”. A third, essentially different approach to
the algebraic study of predicate logics exploits Mostowski’s idea of interpreting quantifiers as
infinite meets and joins in ordered structures (rather than as independent algebraic operations);
over time this has become the most popular choice for the algebraic study of particular first-order
non-classical logics for which a successful algebraic study of their sentential fragment exists: see
Rasiowa (1974), Cintula and Noguera (2015) and Hájek (1998).
10 Josep Maria Font

that equations are just pairs of formulas, so that if E(x) ⊆ Eq is a set of equations
A
 a single variable x and a ∈AA, then E A(a) ⊆ AA × A; more precisely, if E(x) =
in
αi (x) ≈ βi (x) : i ∈ I , then E (a) = hαi (a), βi (a)i : i ∈ I .
The consequence `L of a logic relates sets of formulas to single formulas. This
relation is extended, with the same notation, to a relation between two sets of for-
mulas, by putting, for Γ , Γ 0 ⊆ Fm,
def
Γ `L Γ 0 ⇐⇒ Γ `L ϕ for all ϕ ∈ Γ 0 .
 

The relation `L induces an equivalence relation, the interderivability relation of


L , defined as
def  
α a`L β ⇐⇒ α `L β and β `L α ,
which is also denoted by Λ L from Section 6.3 on. This relation is naturally ex-
tended to sets of formulas:
def
Γ a`L Γ 0 ⇐⇒ Γ `L Γ 0 and Γ 0 `L Γ .
 

More details of the necessary background in logic can be read in Chapter 1 of


Galatos et al. (2007); moreover, that chapter and the one by Raftery in this vol-
ume contain the basic background in universal algebra needed in abstract algebraic
logic.

3 Bare algebraizability

There is a host of logics L for which there exists a class of algebras K such that
the following algebraic completeness theorem holds. For all Γ ∪ {ϕ} ⊆ Fm,

Γ `L ϕ ⇐⇒ for all A ∈ K and all h ∈ Hom(Fm, A),


(1)
h(γ) = 1 for all γ ∈ Γ implies h(ϕ) = 1.

Here, 1 denotes a special element in the algebras of the class K. When this element
is an algebraic constant of the class K, a logic L satisfying (1) is called the asser-
tional logic of K. Table 1 recaps some of the best-known examples, but you most
probably know of other cases.
For many years, the algebraic study of many logics rested mainly upon this kind
of algebraic completeness and its consequences. The genius of Blok and Pigozzi
was to realize that this completeness result can be generalized and enhanced in
three directions:
• Expressing (1) as a relation between two substitution-invariant closure relations:
the consequence relation `L of the logic and another relation K , associated
with the class of algebras K.
Abstract Algebraic Logic 11

L K
Classical (sentential) logic Boolean algebras
Intuitionistic logic Heyting algebras
Dummett-Gödel logic Linear Heyting algebras
Łukasiewicz’s infinitely-valued logic MV-algebras
Global consequence of modal system K Normal modal algebras
Global consequence of modal system S4 Closure algebras
Global consequence of modal system S5 Monadic algebras
BCK logic BCK algebras
Table 1 Some logics and their algebraic counterparts obtained by the traditional method.

• Expressing (1) in terms of a translation of formulas into equations (namely, the


translation α 7−→ α ≈ 1), and generalizing this to a suitable notion of a structural
transformer.
• Observing that this translation can be, so to speak, inverted, in the sense that
there is another translation that turns every equation α ≈ β into a set of formulas
∆ (α , β ), that a kind of completeness dual to (1) holds, and that moreover the
two translations are mutually inverse in a precise, technical sense.
The three directions are discussed separately, before reaching the real, general defi-
nition.

3.1 The equational consequence relative to a class of algebras

The key concept here can be expressed as follows.

Definition 3.1. Let K be a class of algebras (of a common similarity type). The
equational consequence relative to K is the relation between sets of equations and
equations defined, for any Θ ∪ {ε ≈ δ } ⊆ Eq, as
def
Θ K ε ≈ δ ⇐⇒ for all A ∈ K and all h ∈ Hom(Fm, A) ,
(2)
h(α) = h(β ) for all α ≈ β ∈ Θ implies h(ε) = h(δ ).
T
For an algebra A, A means {A} , so that K = A∈K A . Sometimes writing A 
α ≈ β [[h]] instead of h(α) = h(β ) helps to emphasize the model-theoretic aspect of
this relation.

An important point is not to confuse11 this relation with Birkhoff’s equational


consequence (denoted, as usual, by ). The latter does not depend on a particular
class of algebras K (i.e., it is absolute) and is defined as
11 This warning complements the very pertinent one in Galatos et al. (2007, pp. 68–69) concerning
the relation between Birkhoff’s consequence and first-order consequence.
12 Josep Maria Font
def
Θ  ε ≈ δ ⇐⇒ for all algebras A of the similarity type ,
(3)
A  α ≈ β for all α ≈ β ∈ Θ implies A  ε ≈ δ .

As usual, A  α ≈ β means that h(α) = h(β ) for all h ∈ Hom(Fm, A). These two
consequences are really different, as may be guessed from the different scope of the
quantifiers “for all h” present in (2) and implicit in (3). Both consequences  and
K satisfy the replacement rule

α ≈β
for each formula δ (z,~z), (4)
δ (α,~z) ≈ δ (β ,~z)

and while Birkhoff’s consequence  satisfies the substitution rule

α ≈ β  σ (α) ≈ σ (β ) for each substitution σ ,

the consequences K are invariant under substitutions in a sense similar to condition


(S) in Definition 2.1, namely that

if Θ K ε ≈ δ then σ (Θ ) K σ (ε) ≈ σ (δ ), for each substitution σ .

In contrast,  is not invariant under substitutions (example: x ≈ y  x ≈ z, but x ≈


x 6 x ≈ z) and in general K does not satisfy the substitution rule (example: x ≈
y 2K x ≈ z, when K contains at least one non-trivial algebra).
The sets of equations that are closed under Birkhoff’s consequence are called
equational theories, and correspond to varieties, the classes of algebras that are
closed under isomorphisms, subalgebras, products12 and homomorphic images;
these are actually the classes of algebras defined by a set of equations. When K
is a quasivariety (a class of algebras defined by quasi-equations, or, equivalently, a
class that is closed under isomorphisms, subalgebras, products and ultraproducts),
then K is finitary, in a sense similar to condition (F) on page 8. Note that when Θ
is finite, the definition (2) can be expressed in terms of modelling quasi-equations
by  V 

{αi ≈ βi : i < k} K ε ≈ δ ⇐⇒ K  ∀~x αi ≈ βi → ε ≈ δ
i<k

where ~x denotes the sequence of the variables that occur in the enclosed formula.
The algebraic completeness theorem (1) can be expressed, in terms of K , as

Γ `L ϕ ⇐⇒ {γ ≈ 1 : γ ∈ Γ } K ϕ ≈ 1. (5)

In light of this, the completeness result acquires a more algebraic rather than seman-
tic tone, and at the same time one that makes several generalizations possible and
natural, as discussed next.
12For simplicity, closure under the product operator includes the case of the product of an empty
family, which is defined to be a trivial algebra.
Abstract Algebraic Logic 13

As in the case of logics, initially the consequence K is a relation between a set


of equations and a single equation. This relation is extended, keeping the symbol, to
a relation between two sets of equations, by putting, for Θ ,Θ 0 ⊆ Eq,
def
Θ K Θ 0 ⇐⇒ Θ K ε ≈ δ for all ε ≈ δ ∈ Θ 0 ;
 

the associated equivalence relation =||=K and its extension to sets are defined simi-
larly to a`L :
def
Θ =||=K Θ 0 ⇐⇒ Θ K Θ 0 and Θ 0 K Θ
 

3.2 Translating formulas into equations

The use of the equational consequence of the relevant class of algebras in order
to convert the algebraic completeness theorem (1) into the equivalence (5) was
made possible by “rewriting” each formula α as the equation α ≈ 1. For this to
make sense, this 1 should be a symbol of the language; either a primitive con-
stant or a constant term in the relevant class of algebras. We often say that we
have “translated” formulas into equations, but technically, the function is called a
transformer;13 moreover, in general we take it that it turns a formula into a set of
equations rather than a single equation (you will see why later on). Observe also
that the translation has been achieved by taking a “basic” equation x ≈ 1 and then
replacing x by α in it. Putting all these features together, we arrive at the following
notion.

Definition 3.2. A transformer from formulas to equations is a function τ : Fm →


P(Eq). A transformer is structural when it commutes with substitutions
 in the
sense that for all substitutions σ and all ϕ ∈ Fm , σ τ (ϕ) = τ σ (ϕ) .

For this definition to make sense, substitutions are extended to equations in the
natural way, i.e., by defining σ (α ≈ β ) :=Sσ (α) ≈ σ (β ), and to sets of equations
by taking  unions, i.e., by defining σ (Θ ) := α≈β ∈Θ σ (α ≈ β ); then, the expression
σ τ (ϕ) makes sense.
Meanwhile, transformers are also extended to sets of formulas by taking unions,
S
i.e., τ (Γ ) := ϕ∈Γ τ (ϕ). A function between the two power sets satisfies this if
and only if it commutes with unions; thus, it is equally useful to think of a structural
transformer as a function τ : P(Fm) → P(Eq) that commutes with unions14 and
with substitutions.
It is easy to see that:
13 The term “translation” is much overused in the literature.
14 Considering the natural lattice structure of these power sets (with set inclusion as the order rela-
tion), a function that commutes with unions is just a residuated function from P(Fm) to P(Eq).
The residuation view is the key to the more abstract approaches to the notion of algebraizability
that are briefly touched upon in Section 8.
14 Josep Maria Font

Lemma 3.3. A transformer τ : Fm → P(Eq) is structural if and only if it is defin-


able from a set of equations in a single variable E(x) ⊆ Eq by putting τ (ϕ) := E(ϕ)
for all ϕ ∈ Fm.

In many cases, the set τ (ϕ) is actually unitary, as in the previous example, where
τ (α) = {α ≈ 1}, obtained from the equation x ≈ 1. In many cases where there is no
constant 1 in the language, the equation x ≈ x → x works as well; but there are more
complicated examples; for instance, Blok and Pigozzi (1989) show that relevance
logic R satisfies a result like (5) with x ∧ (x → x) ≈ x → x playing the rôle of x ≈ 1.
Those authors gave the general form of (5) a name:

Definition 3.4. Let L be a logic, K a class of algebras, and τ a structural trans-


former. The class K is a τ -algebraic semantics for the logic L , and L is the
τ -assertional logic of the class K, when for all Γ ∪ {ϕ} ⊆ Fm,

Γ `L ϕ ⇐⇒ τ (Γ ) K τ (ϕ). (ALG1)

A logic has an algebraic semantics when it has a τ -algebraic semantics for some
structural transformer, τ . The equations in a single variable of the set E(x) that
defines τ are called the defining equations of the algebraic semantics.

The algebraic completeness theorem (5) can be rephrased as saying that

“K is an algebraic semantics for L with defining equation x ≈ 1”,

and we can view Table 1 as listing algebraic semantics of a few well-known logics.
In this way, the term “algebraic semantics” has acquired a very precise, technical
meaning. Note, however, that the term is very often used in a more ambiguous sense:
as indicating that there is some kind of relation between a certain logic and a certain
class of algebras, but not necessarily in the sense of Definition 3.4; as an example
of this consider the local consequence associated with modal system S4: it is also
related to the class of closure algebras, as we see on page 27, but this class is not,
technically, an algebraic semantics for that logic.
Observe that if K is an algebraic semantics for a logic L , then any other class
K0 such that K = K0 is one as well, with the same defining equations. Thus, for
instance, the single two-element Boolean algebra 2, or the class of all finite Boolean
algebras, are also algebraic semantics for classical logic (together with the class
of all Boolean algebras). The results of Blok and Rebagliato (2003) show that an
algebraic semantics for a logic can be quite weird; the bare notion has received little
attention and seems to be of limited interest by itself. Only when coupled with the
inverse translation to be discussed in the next section does it give rise to a more
interesting notion.
Abstract Algebraic Logic 15

3.3 Translating equations into formulas

The perspective from which we view an algebraic semantics as a relation between


`L and K expressed by a structural transformer makes it all the more natural to
consider the symmetric situation: if ∆ (x, y) ⊆ Fm is a set of formulas in two vari-
ables, then by putting ρ (α ≈ β ) := ∆ (α , β ), and taking unions, one obtains a func-
tion ρ : P(Eq) → P(Fm) which commutes with unions and with substitutions in
the obvious sense; that is, a structural transformer from equations to formulas.
Then the dual relation to (ALG1) would be that, for all Θ ∪ {ε ≈ δ } ⊆ Eq,

Θ K ε ≈ δ ⇐⇒ ρ (Θ ) `L ρ (ε ≈ δ ). (ALG2)

There is no name for logics that satisfy this property alone; it holds in all equivalen-
tial logics (one of the classes in the Leibniz hierarchy), and it actually characterizes
them if supplemented with a rather natural property of the set ∆ (x, y); see Theo-
rem 6.2.
The “reverse algebraic completeness” expressed by (ALG2) is certainly found in
classical logic and Boolean algebras, or in intuitionistic logic and Heyting algebras,
although it is not usually expressed through K . Those logics (and many others) sat-
isfy (ALG2) relative to the corresponding class of algebras with ∆ (x, y) = {x ↔ y};
the case where Θ is the empty set may look more familiar: it says that an equation
ε ≈ δ holds in all Boolean (resp. Heyting) algebras if and only if the formula ε ↔ δ
is a theorem of classical (resp. intuitionistic) logic, and these are well-known facts.
It is easy to see that if a logic L satisfies (ALG2) with respect to a class K
through a transformer ρ , then any fragment of L whose language contains the
language of the formulas that define ρ also satisfies it, with respect to the appro-
priate class of algebraic reducts of the algebras in K. In the examples of classical
and intuitionistic logic, this applies to all their fragments containing ↔ (obvious);
but also to all those containing implication →, if instead of {x ↔ y} we consider
∆ (x, y) = {x → y , y → x}, which performs the same function. This explains why, in
the general case, it is natural to take transformers as mapping an equation to a set of
formulas, and symmetrically a formula to a set of equations.

3.4 Putting it all together

Algebraizability adds, to the existence of two transformers going back and forth
between the logic and the equational consequence of the class of algebras, the re-
quirement that these transformers are mutually inverse modulo the consequences.
This is reflected in the two additional conditions found in the real definition:15

15 This is actually an extension of the original notion, due to Blok and Pigozzi (1989); see Section 5
for details.
16 Josep Maria Font

Definition 3.5. A logic L is algebraizable when there are a class K of algebras and
structural tranformers τ : P(Fm) → P(Eq) and ρ : P(Eq) → P(Fm) such that
for all Γ ∪ {ϕ} ⊆ Fm and all Θ ∪ {ε ≈ δ } ⊆ Eq, the following conditions hold:

Γ `L ϕ ⇐⇒ τ (Γ ) K τ (ϕ) (ALG1)
Θ K ε ≈ δ ⇐⇒ ρ (Θ ) `L ρ (ε ≈ δ ) (ALG2)

ϕ a`L ρ τ (ϕ) (ALG3)

ε ≈ δ =||=K τ ρ (ε ≈ δ ) (ALG4)

The equations E(x) that define the transformer τ are called the defining equations,
and the formulas ∆ (x, y) that define the transformer ρ are called the equivalence
formulas. The equivalent algebraic semantics of an algebraizable logic L is the
largest class K satisfying the above properties.16
Observe that, by substitution invariance, it is enough to require that conditions
(ALG3) and (ALG4) hold just for variables (instead of arbitrary formulas):

x a`L ρ τ (x) (ALG3)

x ≈ y =||=K τ ρ (x ≈ y) (ALG4)

The definition is highly symmetric, at the price of being redundant, because

(ALG1) + (ALG4) ⇐⇒ (ALG2) + (ALG3).

The more direct approach to algebraizability is through (ALG1) and (ALG4): while
(ALG1) is a natural completeness theorem of the logic L relative to the class
of algebras K, the condition (ALG4)  concerns the class K alone. Moreover, the
consequence x ≈ y K τ ρ (x ≈ y) is trivially equivalent to the statement that
K  τ ρ (x ≈ x) . Thus, a logic is algebraizable if and only if it has a τ -algebraic se-
mantics K and a reverse transformer ρ such thatin each A ∈ K, the equations of the
set τ ρ (x ≈ x) and the entailment τ ρ (x ≈ y) A x ≈ y hold. The latter property
is a kind of “rule of thumb” for identifying algebraizability: in the candidate class of
algebras, identity should be characterized by some set of equations that collectively
has the form of the equations appearing in the algebraic completeness, applied to
a set of two-variable formulas. For instance, there is a very large class of algebras
associated with non-classical logics where the quasi-equation

(x → y ≈ 1)∧∧ (y → x ≈ 1) → x ≈ y

holds; namely, the class of implicative algebras introduced by Rasiowa (1974).


There is a large group of logics (the so-called implicative logics of Section 4.5)
16 In the original notion, any class K satisfying the four conditions was called an equivalent alge-
braic semantics of L . Raftery (2006a) started the practice of reserving the name for the largest of
all such classes, which is hence unique; see the observations on its existence and character at the
end of this section.
Abstract Algebraic Logic 17

that have an implication connective → and which satisfy an algebraic complete-


ness theorem like (5) with respect to a class of implicative algebras; the preceding
considerations show that all these logics are, indeed, algebraizable. This group is al-
ready very large, but there are also many logics that are algebraizable without being
implicative:
• The equivalence fragments of classical logic and of intuitionistic logic are alge-
braizable, with E(x) = {x ≈ x ↔ x} and ∆ (x, y) = {x ↔ y}.
• All the substructural logics associated with varieties of residuated lattices in the
book by Galatos et al. (2007) are algebraizable, with E(x) = {x ∧ 1 ≈ 1} and
∆ (x, y) = {x\y, y\x}. This constitutes a very large group, including many logics
in the linear logic family, most fuzzy logics, Łukasiewicz many-valued logics,
and so on. Only some of the logics in this large group are also implicative.
• The relevance logic R mentioned above is algebraizable, with E(x) = {x ∧ (x →
x) ≈ x → x} and ∆ (x, y) = {x → y, y → x}. Technically, this logic does not belong
to the preceding group, because it does not contain a constant that plays the rôle
of 1, but is very close to it.
In Section 5, some less standard examples of algebraizable logics are listed.
As a consequence of Theorem 4.1 below, it is easy to see that any extension17
of an algebraizable logic is algebraizable as well, with the same transformers; any
fragment18 whose language contains the sets E(x) and ∆ (x, y) is also algebraiz-
able; and any expansion,18 provided it satisfies a simple condition on ∆ , is again
algebraizable. Thus, from each of the preceding examples, a host of new ones is
automatically produced; for instance, a host of logics that build on classical logic by
expanding its language, such as many logics of the modal family (dynamic, tempo-
ral, etc.), are automatically algebraizable if the connectives added satisfy condition
(Re∆ ) of Theorem 4.1.
A few final observations are in order here concerning the class called in Defi-
nition 3.5 the equivalent algebraic semantics of L . Observe that the class K that
appears in conditions (ALG1)–(ALG4) does so only through its relative equational
consequence K ; therefore, any other class that generates the same consequence
would satisfy the same conditions and might thus replace it. It is not difficult to
prove that there is always a largest one among all such classes, and that it is a gen-
eralized quasivariety19 ; that is, a class of algebras that can be axomatized by a set of
17 A logic L 0 is an extension of a logic L when they have the same language and `L ⊆ `L 0 .
18 If L and L 0 are logics in languages L and L0 respectively, with L ⊆ L0 , then L 0 is an
expansion of L when `L ⊆ `L 0 . A conservative expansion is an expansion such that `L =
`L 0 ∩ P(FmL ) × FmL ; in this case, it is also said that L is the L-fragment of L 0 .
19 Generalized quasivarieties are also called “σ -quasivarieties” in the literature. Due to the fact

that we have assumed a fixed set V of variables, generalized quasivarieties as defined above are
characterized as the classes of algebras that are closed under isomorphisms, subalgebras, products,
and the operation U introduced in Blok and Jónsson (2006):

U(K) := A : If B is a subalgebra of A generated by a set of cardinality 6 |V |, then B ∈ K .
Thus, in particular, generalized quasivarieties are “SP-classes” or prevarieties: these are the classes
that are closed under just isomorphisms, subalgebras and products, and can be characterized as
18 Josep Maria Font

equations and generalized quasi-equations (infinitary formulas that are like quasi-
equations but with a possibly infinite conjunction of equations in the antecedent of
the implication). In fact, such a class is the only generalized quasivariety that sat-
isfies the algebraizability conditions for a given logic. Moreover, it is also easy to
show20 that this class is independent of the transformers. Thus, it is a uniquely de-
termined algebraic object associated with an algebraizable logic. Later on, in the
proof of Theorem 4.1, we see how to construct it explicitly, starting from the logic
and any pair of transformers that show algebraizability; see also footnote 22.

4 The origins of algebraizability: the Lindenbaum-Tarski


process

It is common to hear or read that the notion of algebraizability arises from a gener-
alization of the Lindenbaum-Tarski procedure of proving the completeness of clas-
sical logic with respect to the class of Boolean algebras, in the sense of (1) or (5),
and similar statements. In order to understand why this is so, let us review how that
original process works.

4.1 The process for classical logic

In this subsection, L denotes classical logic (or another logic with similar prop-
erties; see below), and we assume it is defined by some axiomatic system that
has modus ponens for the implication connective (MP→ ) x , x → y `L y among
its inference rules; while K denotes the class of Boolean algebras (or the corre-
sponding class for other logics). The proof of the completeness theorem (1) by the
Lindenbaum-Tarski method has two separate halves which have very different the-
oretical importance:
(⇒) This is called the Soundness Theorem, and is proved by what is commonly
called routine checking; i.e., by checking that if h ∈ Hom(Fm, A) for a Boolean
algebra A, then h(ϕ) = 1 for all axioms ϕ , and that all inference rules preserve the
property of “being evaluated to 1”. In the case of (MP→ ) this means checking that
if h(ϕ) = 1 and h(ϕ → ψ) = 1 then h(ψ) = 1. All this work is purely algebraic and
uses the properties of Boolean algebras in an essential way; for instance, checking
(MP→ ) amounts to checking that in any Boolean algebra, if 1 → a = 1, then a = 1
(and this holds simply because 1 → a = ¬1 ∨ a = 0 ∨ a = a). Then, induction on the

those defined by (a possibly proper class of) generalized quasi-equations in a language having a
proper class of variables (see for instance Hodges, 1993, § 9.2). Note that the term implicational
(or implicative) class has been used in the literature for ordinary quasivarieties, for generalized
quasivarieties, and for prevarieties.
20 This can be proved directly, just from the definition of algebraizability; but see also Corol-

lary 4.15.
Abstract Algebraic Logic 19

length of proofs in the axiomatic system defining L completes the proof.


(⇐) This is proved by contraposition: one assumes that Γ 0L ϕ , and one falsi-
fies the right-hand side of (1) by constructing a particular Boolean algebra A and a
particular h ∈ Hom(Fm, A) such that h(Γ ) ⊆ {1} while h(ϕ) 6= 1. This is strictly
speaking the so-called Lindenbaum-Tarski process, and it can be broken down into
several different steps:
(LT1) One starts from the assumption that Γ 0L ϕ , and one considers the theory
Γ 0 of L generated by Γ . Thus, Γ ⊆ Γ 0 while ϕ ∈ / Γ 0.
0 0
(LT2) One defines a relation, denoted here by Ω  Γ , by putting α ≡ β (Ω Γ ) if
and only if α → β ∈ Γ 0 and β → α ∈ Γ 0 .
(LT3) One shows that Ω Γ 0 is a congruence of Fm. This requires usage of specific
axioms or theorems of L ; for instance, that (α → β ) → (¬β → ¬α) is a
theorem, plus (MP→ ), shows that Ω Γ 0 is a congruence with respect to ¬.
(LT4) One shows that the quotient algebra Fm/Ω Γ 0 is a Boolean algebra. Again,
this depends on using very particular axioms and theorems of the logic. A
key point in this step is to realize that the order relation of the quotient
algebra is determined in the following way: α/Ω Γ 0 6 β /Ω Γ 0 if and only
if α → β ∈ Γ 0 .
(LT5) One shows that for any α ∈ Fm , α ∈ Γ 0 if and only if α/Ω Γ 0 ∈ Γ 0 /Ω Γ 0 .
It is easy to see that what is required for this to hold is just that if β ∈ Γ 0 and
β ≡ α (Ω Γ 0 ) then α ∈ Γ 0 , and this in turn follows from (MP→ ), given the
definition of Ω Γ 0 in (LT2).
(LT6) Finally, one shows that all formulas in Γ 0 are mutually equivalent, so that
Γ 0 constitutes a single element in the quotient algebra, and that this element
is the top element 1 ∈ Fm/Ω Γ 0 . Both facts follow from applying (MP→ )
to the formula α → (β → α), which is a theorem of classical logic (and of
many similar logics).
Note that after this, the property in (LT5) becomes: For any α ∈ Fm , α ∈ Γ 0
if and only if α/Ω Γ 0 = 1.
The “process” is now finished. The quotient algebra Fm/Ω Γ 0 is the desired particu-
lar Boolean algebra, by (LT4), and the canonical projection π ∈ Hom(Fm, Fm/Ω Γ 0 )
defined by π(α) := α/Ω Γ 0 for all α ∈ Fm is the desired particular homomor-
phism. Then, by (LT1) and (LT5) as rewritten after (LT6), π(Γ ) ⊆ π(Γ 0 ) = {1}
while π(ϕ) 6= 1. This completes the proof. 
Exactly the same process works for many other logics, including those in Table 1.
One class of logics for which the process works with absolutely no modifications
(for the corresponding class of algebras) is described in Section 4.5.
You may realize that the proof works even if Γ 0 is not exactly the theory gener-
ated by Γ ; actually, taking any theory containing Γ and not containing ϕ works.
This leaves room for suitable modifications of the choice of Γ 0 in (LT1), which
may produce an algebra Fm/Ω Γ 0 with particular properties, and leads to different
20 Josep Maria Font

completeness theorems, with respect to restricted classes of algebras.21 An extreme


and popular choice for classical logic is to take Γ 0 to be a maximally consistent
theory containing Γ but not ϕ (the property that guarantees the existence of such
a theory is commonly called “Lindenbaum’s Lemma”); then the quotient algebra
is (isomorphic to) the two-element Boolean algebra 2, and we obtain completeness
with respect to just this algebra; this proof is virtually equivalent to (one of) the
usual textbook completeness proofs for classical logic with respect to two-valued
truth tables.
In Section 4.3, we see that in fact the process can be generalized in such a way
that it applies to absolutely every logic. In some sense, one may say that this is a
starting point of abstract algebraic logic; in particular, we see that it is not necessary
for the logic to be presented as an axiomatic system, as the previous exposition
might suggest.
But before that, in Section 4.2, we will see that the process can be performed,
mutatis mutandis, for algebraizable logics, in order to find their equivalent algebraic
semantics. Then in Section 4.4, we see that this adapted process can actually be con-
sidered as a version of the universal process, by introducing appropriate definability
conditions.

4.2 The process for algebraizable logics

It may at first seem that in order to establish the algebraizability of a logic, one
must previously know of a candidate class of algebras K that satisfies one of the
two equivalent pairs of conditions in the definition (notice that K appears in three of
the four conditions). The following fundamental result, usually called the syntactic
characterization, offers a way to establish algebraizability using conditions exclu-
sively on the logic; moreover it gives a way to find the corresponding equivalent
algebraic semantics. It has been used very often in particular cases, and its proof
is particularly instructive, as it clearly shows that the five conditions it contains are
exactly those that are needed for the Lindenbaum-Tarski process to work in this
generalized way.

Theorem 4.1. A logic L is algebraizable if and only if there are sets of equations
E(x) ⊆ Eq and of formulas ∆ (x, y) ⊆ Fm such that L satisfies the following five
conditions:

`L ∆ (x, x) (R∆ )
∆ (x, y) `L ∆ (y, x) (Sym∆ )
∆ (x, y) ∪ ∆ (y, z) `L ∆ (x, z) (Trans∆ )
21 Examples of this strategy are the completeness of the basic fuzzy logic BL with respect to
linear BL-algebras (Hájek, 1998) and its generalization to the so-called implicational semilinear
logics (Cintula and Noguera, 2010), which include all the usual fuzzy logics.
Abstract Algebraic Logic 21
n
[
∆ (xi , yi ) `L ∆ (λ x1 . . . xn , λ y1 . . . yn ) for all λ ∈ L, with n = arλ (Re∆ )
i=1

x a`L ∆ E(x) (ALG3)

Then, E(x) is the set of defining equations and ∆ (x, y) is the set of equivalence
formulas; and the equivalent algebraic semantics of L is the class

ρ ) := A : τ (Γ ) A τ (ϕ) whenever Γ `L ϕ ,
K(L ,ττ ,ρ
 (6)
and E ∆ (x, y) A x ≈ y .

Recall that τ (α) := E(α) and ρ (α ≈ β ) := ∆ (α , β ), and that these functions


are extended to sets of formulas by taking unions. There is no harm in mixing the
two notations, as in (6). Indeed, sometimes one is more illustrative than the other;
for instance, the five conditions were displayed in terms of the sets of formulas to
highlight their character of syntactic conditions on the logic alone.
Sketch of the proof of Theorem 4.1. Notice that the four conditions (R∆ )–(Re∆ )
are the “ρρ -transforms” of some basic, universal properties of the consequence K .
Therefore, if L is algebraizable, those conditions follow from these properties by
(ALG2); for instance, x ≈ y K y ≈ x implies (Sym∆ ), the replacement rule (4)
implies (Re∆ ), and so forth. As for condition (ALG3), it already appears in the def-
inition of algebraizability. Thus, the five conditions hold.
Now assume that L satisfies the five conditions. We are going to review the
proof that in this case, L satisfies conditions (ALG1) and (ALG4) with respect
to the class K(L ,ττ ,ρ ρ ) defined in (6). Notice that by (ALG1) and (ALG4), if L
is algebraizable with respect to some class K by the transformers τ and ρ , then
K ⊆ K(L ,ττ ,ρ ρ ); therefore, after showing that L is indeed algebraizable with re-
spect to K(L ,ττ ,ρ ρ ), this class turns out to be the largest one,22 that is, it is the
equivalent algebraic semantics of L .
To lighten notation, let K denote K(L ,ττ ,ρ ρ ) in the rest of this proof. Condition
(ALG4) concerns only the class K, and one of its halves appears explicitly as the
second condition on A in definition (6) of K; as for the other half, it amounts to
showing that for each A ∈ K , A  E ∆ (x, x) , but this is a consequence of applying
the first condition in (6) to (R∆ ).
Thus, what we have to see is that (ALG1) holds. Again, its (⇒) half is just a con-
sequence of the first point in definition (6) of K, and we are left with the proof of
its (⇐) half. And this is what an appropriate generalization of the Lindenbaum-
Tarski process achieves. Let us review it, without entering into the details, keeping
as parallel with the steps in Section 4.1 as possible.
22 This argument only shows that K(L ,τ τ ,ρ
ρ ) is the largest class for τ and ρ ; Corollary 4.15
shows that this class is actually independent of the transformers, so it is the absolute largest. Notice
that (6) describes K(L ,ττ ,ρρ ) as the class of all algebras that satisfy a certain set of generalized
quasi-equations; thus, this class is indeed a generalized quasivariety, as claimed on page 17.
22 Josep Maria Font

(LT1) One starts from the assumption that Γ 0L ϕ , and one considers the theory
Γ 0 of L generated by Γ . Thus, Γ ⊆ Γ 0 while ϕ ∈ / Γ 0.
(LT2) One defines the relation α ≡ β (Ω Γ 0 ) if and only if ∆ (α , β ) ⊆ Γ 0 .
(LT3) One shows that Ω Γ 0 is a congruence of Fm; this results from mechanical
application of conditions (R∆ )–(Re∆ ). Observe that each formalizes one of
the properties required: reflexivity, symmetry, transitivity, and compatibility
with the operations, respectively.
(LT4) One shows that the quotient of the formula algebra under this congru-
ence belongs to the target class, i.e., that Fm/Ω Γ 0 ∈ K. To this end, one
has to show that Fm/Ω Γ 0 satisfies the two conditions in (6). Both are
shown by using the property in the next step (LT5) and the trick that any
h ∈ Hom(Fm, Fm/Ω Γ 0 ) has the form h = π ◦ σ for some substitution σ ,
where π : Fm → Fm/Ω Γ 0 is the canonical projection.
(LT5) One shows that for any α ∈ Fm , α ∈ Γ 0 if and only if Fm/Ω Γ 0  τ (α) [[π]].
This is easily shown by using exclusively (ALG3); as a matter of fact, this
property is equivalent to the statement that the first sentence in this point
holds for all theories Γ 0 of L .
The “process” stops here. The quotient algebra Fm/Ω Γ 0 is the desired particular
algebra in K, by (LT4), and the canonical projection π ∈ Hom(Fm, Fm/Ω Γ 0 ) is
the desired particular homomorphism. Applying (LT5) to the facts of (LT1), we can
conclude that Fm/Ω Γ 0  τ (Γ ) [[π]] while Fm/Ω Γ 0 6 τ (ϕ) [[π]]. This shows that
τ (Γ ) 6K τ (ϕ) and completes the proof of (ALG1) by contraposition. 
The parallelism between this proof and the particular one given above for clas-
sical logic is clear; observe that the present (LT5) condenses steps (LT5) and (LT6)
of the case of classical logic, because the equality α/Ω Γ 0 = 1 appearing there in
step (LT6) can be rewritten as the condition Fm/Ω Γ 0  α ≈ 1 [[π]]. In the case of
an arbitrary algebraizable logic, there is no need to show that all formulas in Γ 0 are
equivalent under Ω Γ 0 , and it makes no sense to speak of the top element of the
algebras in K, as they might not have one. Among the examples of substructural
logics mentioned near the end of Section 3.4, there are many where these additional
properties do not hold.
Since every algebraizable logic satisfies the five conditions, the second part of
the proof confirms the common claim that algebraizable logics satisfy an algebraic
completeness theorem proved by a natural generalization of the Lindenbaum-Tarski
process. Moreover, the same proof clearly shows that the completeness also holds
(as does the algebraizability) for just the class of algebras

Fm/Ω Γ 0 : Γ 0 ∈ ThL .

(7)

After Theorem 4.11, we see that these algebras deserve the name Lindenbaum-
Tarski algebras of the logic L , and they are uniquely determined by it (i.e., they do
not depend on the particular set of formulas ∆ used to define Ω Γ 0 ).
The properties mentioned at the end of Section 3.4 follow immediately from
Theorem 4.1: since all the conditions (R∆ )–(Re∆ ) are formulated in terms of the
Abstract Algebraic Logic 23

consequence `L , they are preserved by extensions; they are preserved by fragments,


provided that the language contains the connectives that occur in the sets E and ∆ ;
and finally, they are preserved by expansions, whenever the new connectives satisfy
(Re∆ ).
The “canonical” character of the association of a class K(L ,ττ ,ρ
ρ ) with each al-
gebraizable logic L , coupled with the first fact just mentioned about the algebraiz-
ability of the extensions of an algebraizable logic, produces an important property:

Theorem 4.2. Let L be an algebraizable logic with equivalent algebraic semantics


the generalized quasivariety K. Then there is a dual isomorphism between the lat-
tice of all the extensions of L and the lattice of all sub-generalized quasivarieties
of K; all these logics are algebraizable and have the corresponding generalized
quasivariety as their equivalent algebraic semantics.

Proof. Clearly, if L 0 is an extension of L , then K(L 0 ,ττ ,ρ ρ ) ⊆ K(L ,ττ ,ρ


ρ ) = K,
and both are generalized quasivarieties. If L is algebraizable, then by Theo-
rem 4.1, L 0 is also algebraizable, with the same transformers. The function L 0 7→
K(L 0 ,ττ ,ρ
ρ ) is order-reversing. Now let K0 ⊆ K be a generalized quasivariety. Since
K satisfies (ALG4), K0 also satisfies it (it is a property satisfied individually by each
algebra in the class), for the same transformers. Therefore, if we define a logic L 0
from K0 by condition (ALG1) with the same transformers, L 0 is automatically al-
gebraizable with respect to K0 , and its equivalent algebraic semantics, which is the
unique generalized quasivariety that satisfies (ALG1) and (ALG4), must perforce be
K0 . Finally, that L 0 is an extension of L is a consequence of (ALG1) plus K0 ⊆ K.
Thus, the function K0 7→ L 0 also reverses order. The two functions are clearly in-
verses to one another. 

Thus, once an algebraizable logic is found and its equivalent algebraic semantics
is identified, one can deal with the lattice of its extensions, in a uniform way, by
studying the lattice of sub-generalized quasivarieties of the base class of algebras.
This generalizes a number of well-known situations, where attention is restricted to
the axiomatic extensions of a particular logic and to the lattice of subvarieties of a
particular variety. For more applications of Theorem 4.2, see the end of Section 7.

4.3 The universal Lindenbaum-Tarski process: matrix semantics

After having reviewed the Lindenbaum-Tarski process for classical logic and for
algebraizable logics, it is now time to encounter its generalization to absolutely
every logic; as is to be expected, this does not rely on the existence of certain sets of
formulas or equations with particular properties, but on a more abstract construction
which requires a few basic notions of matrix semantics for sentential logics. Logical
matrices are constituted by an algebra together with a subset of its universe, and in
this sense they are clearly “algebra-based” objects.
24 Josep Maria Font

Definition 4.3. Let L be an algebraic language. An (L-)matrix is a pair hA, Fi


where A is an algebra (of type L) and F ⊆ A; this subset is called the filter or
truth filter of the matrix. A matrix is a model of a logic L (in the same language)
when the following holds, for all Γ ∪ {ϕ} ⊆ Fm:

Γ `L ϕ =⇒ for all h ∈ Hom(Fm, A),


(8)
h(γ) ∈ F for all γ ∈ Γ implies h(ϕ) ∈ F .

The class of all models of L is denoted by ModL . A set F ⊆ A is an L -filter


when hA, Fi ∈ ModL , and the set of all L -filters of A is denoted by FiL A.
Comparison with (1) shows that the Soundness Theorem, i.e., the (⇒) half of (1),
in the case of classical logic and Boolean algebras, can be rephrased by saying that
in each Boolean algebra the set {1} is a filter of classical logic. In fact, every lattice
filter of a Boolean algebra is a filter of classical logic; every lattice filter of a Heyting
algebra is a filter of intuitionistic logic; and so on.23
Filters and models can be considered on any algebra, and in particular on the
algebra of formulas. Since the endomorphisms of the formula algebra are the sub-
stitutions, a quick application of property (S) in Definition 2.1 proves a crucial fact:
Lemma 4.4. The filters of a logic on the formula algebra are its theories.
And this quickly gives:
Theorem 4.5 (Wójcicki, 1969). Every logic is complete with respect to the class of
all its models. That is, for all Γ ∪ {ϕ} ⊆ Fm:

Γ `L ϕ ⇐⇒ for all hA, Fi ∈ ModL and all h ∈ Hom(Fm, A),


(9)
h(γ) ∈ F for all γ ∈ Γ implies h(ϕ) ∈ F.

Proof. Part (⇒) is just the definition of model. Part (⇐) is proved by a radical
simplification of the Lindenbaum-Tarski process, as follows. Assume that Γ 6`L ϕ ;
just do step (LT1) and obtain the theory Γ 0 generated by Γ ; take hFm,Γ 0 i as the
required model (it is, by Lemma 4.4) and the identity function from the formula
algebra into itself as the required homomorphism; this falsifies the right-hand side
of (9). 
The real generalization of the process requires, of course, a factorization step,
and the choice of the right class of algebras (actually, of matrices) so that part (⇒)
and step (LT4) work. There is a common misunderstanding, according to which the
factorization can only be performed for logics where there is some formula or set of
formulas that defines a congruence, as in the two versions of step (LT2) previously
shown, or else in those logics where the interderivability relation a`L is itself a
congruence. But this is not the case: there is always a natural, canonical way of
factorizing every model of every logic. To this end, the following notions are crucial:
23 The logical usage of the term “filter” is undoubtedly inspired by its algebraic and topological
origins. Sometimes the term “deductive filter” is used to emphasize the difference.
Abstract Algebraic Logic 25

Definition 4.6. Let hA, Fi be an L-matrix. A congruence θ ∈ ConA is compatible


with F when for all a, b ∈ A, if a ∈ F and a ≡ b (θ ), then b ∈ F . The Leibniz
congruence of an L-matrix hA, Fi is defined as

Ω A F := max{θ ∈ ConA : θ is compatible with F}.

A matrix is reduced when its Leibniz congruence is the identity relation. The class
of all reduced matrices that are models of a logic L is denoted by Mod∗L .

The largest compatible congruence always exists, because the ordinary supre-
mum of a family of compatible congruences (which exists because ConA is a com-
plete lattice) is also compatible. Notice that Ω A F is a purely algebraic object, and
does not depend on any logic. The following characterization, due independently to
Shoesmith and Smiley (1978) and to Czelakowski (1980), extends an idea that is
already present in Łoś (1949):

Theorem 4.7. For any L-matrix hA, Fi  and any a, b ∈ A , a ≡ b (Ω AF) if and only
~
if for all δ (x,~z) ∈ Fm and all ~c ∈ A , δ (a,~c) ∈ F ⇔ δ A (b,~c) ∈ F .
A

Though the notion is older than its namesake, it was named after Leibniz by
Blok and Pigozzi (1989), based on their reading of this characterization24 as stating
that the Leibniz congruence is a first-order analogue of the definition of identity in
second-order logic. The Leibniz congruence has also been called the indiscernibility
relation or, in a more linguistic-oriented view and for the particular case of the
formula algebra, the synonymity relation. The main basic properties we need are:

Lemma 4.8. Let hA, Fi be an L-matrix and L a logic. Then:


1. For all a ∈ A , a ∈ F if and only if a/Ω A F ∈ F/Ω A F .
2. The matrix hA/Ω A F , F/Ω A Fi is reduced.
3. hA, Fi ∈ ModL if and only if hA/Ω A F , F/Ω A Fi ∈ ModL .

Actually, the first point is equivalent to the property that Ω A F is compatible


with F . The other two points show that any model of a logic produces a reduced
model by factorization through the Leibniz congruence. In particular, when applying
this to the models of theories of the logic (Lemma 4.4), one finds the so-called
Lindenbaum-Tarski models of a logic L :

LTMod∗L := hFm/Ω FmΓ, Γ /Ω FmΓ i : Γ ∈ ThL ⊆ Mod∗L .



(10)

The inclusion follows from Lemma 4.4 and points 2 and 3 of Lemma 4.8. Then:

Theorem 4.9 (Wójcicki, 1973). Every logic L is complete with respect to any class
of matrices M such that LTMod∗L ⊆ M ⊆ Mod∗L . That is, for all Γ ∪ {ϕ} ⊆ Fm,
24 Actually, of a stronger one, formulated in the framework where L-matrices are considered as

structures for a first-order language whose constants and function symbols are those in L, and
which has only one relation symbol, a unary one, interpreted as the filter of the matrix. This deeply
influential idea, due to Bloom (1975), is further explained on page 44.
26 Josep Maria Font

Γ `L ϕ ⇐⇒ for all hA, Fi ∈ M and all h ∈ Hom(Fm, A),


(11)
h(γ) ∈ F for all γ ∈ Γ implies h(ϕ) ∈ F.

In particular, every logic is complete with respect to the class of its Lindenbaum-
Tarski models and with respect to the class of all its reduced models.

Proof. Part (⇒) follows from the assumption that M ⊆ Mod∗L , and part (⇐) is
proved by the real generalization of the Lindenbaum-Tarski process:
(LT1) One starts from the assumption that Γ 0L ϕ , and one considers Γ 0 , the
theory of L generated by Γ . Thus, Γ ⊆ Γ 0 while ϕ ∈ / Γ 0.
(LT2) One considers the Leibniz congruence Ω FmΓ 0 given by Definition 4.6.
(LT3) According to its own definition, Ω FmΓ 0 is a congruence of Fm.
(LT4) The quotient matrix hFm/Ω FmΓ 0 , Γ 0 /Ω FmΓ 0 i belongs to M by the as-
sumption that LTMod∗L ⊆ M.
(LT5) For any α ∈ Fm , α ∈ Γ 0 if and only if α/Ω FmΓ 0 ∈ Γ 0 /Ω FmΓ 0 , by point 1
of Lemma 4.8.
In particular, by (LT1) plus (LT5), the canonical projection π ∈ Hom(Fm, Fm/Ω FmΓ 0 )
defined by π(α) := α/Ω Γ 0 for all α ∈ Fm satisfies that π(Γ ) ⊆ π(Γ 0 ) = Γ 0 /Ω FmΓ 0
while π(ϕ) ∈ / Γ 0 /Ω FmΓ 0 . Taking (LT4) into account, this falsifies the right-hand
side of (11), and finishes the proof. 

In Theorem 4.11 we see that the congruences denoted by Ω Γ 0 in steps (LT2)


on page 19 and on page 22 actually coincide with what has just been denoted by
Ω FmΓ 0 , following Definition 4.6; this justifies the choice of the previous notation.
You may have observed that, as in the case of an arbitrary algebraizable logic, the
general step (LT5) is enough to finish the proof, and there is no need for step (LT6),
which only makes sense in particular cases, such as that of classical and similar
logics.
Matrix semantics provides a first notion of the algebraic counterpart of a logic
L (a second, more general notion is found in Section 6.2): it is the class25

Alg∗L := the algebraic reducts of the matrices in Mod∗L .




The Completeness Theorem 4.9, for M = Mod∗L , can be restated in terms of the
class of algebras as follows:

Γ `L ϕ ⇐⇒ h(Γ ) ⊆ F implies h(ϕ) ∈ F ,


for all A ∈ Alg∗L , all h ∈ Hom(Fm, A), (12)
and all F ∈ FiL A such that Ω F is the identity relation.
A

In the best behaved cases, the conditions in the last row can be formulated men-
tioning only the class of algebras; for instance, if L is classical logic, they amount
25 Sometimes the algebras in Alg∗L are called the Leibniz-reduced algebras of L .
Abstract Algebraic Logic 27

to F = {1}, so that Mod∗L = hA, {1}i : A is a Boolean algebra , and (12) be-


comes (1), our starting point. Similarly for the other logics in Table 1 on page 11.
Corollary 4.15 shows that this simplification to use just algebras is actually possible
whenever L is algebraizable, but it is not possible in general; the class Alg∗L does
not characterize the logic, and matrices cannot be dispensed with. The paradigmatic
case of the two consequences associated with any normal system S of modal logic
has already been mentioned: for them, Alg∗ (``S ) = Alg∗ (`gS ); for S4, this is the
variety of closure algebras. In Lemma 6.3, we see that there is a one-to-one nat-
ural correspondence between Mod∗L and Alg∗L for a large class of logics that
includes the algebraizable ones as well as many others.

4.4 Algebraizability and matrix semantics: definability

In this section, we see how the Lindenbaum-Tarski process for algebraizable log-
ics matches the universal process, and we check what was announced before: the
algebraizable version arises from the general one by plugging two definability con-
ditions into the two crucial steps of the process. To do this, we need to consider
the general version of a property that was seen to play a key rôle in the process for
classical logic. Consider the following generalization of modus ponens:

x , ∆ (x, y) `L y (MP∆ )

Lemma 4.10. Every algebraizable logic L satisfies (MP∆ ) for any of its sets of
equivalence formulas ∆ .

Proof. By (ALG1), condition (MP∆ ) holds if and only if E(x), E ∆ (x, y) K E(y).
But by condition (ALG4), this holds if and only if E(x), x ≈ y K E(y), and this is
trivially true. 

It turns out that coupling (MP∆ ) with four of the five conditions of Theorem 4.1
amounts to one of the definability conditions we are after:

Theorem 4.11 (Czelakowski, 1981). Let L be a logic and let ∆ (x, y) ⊆ Fm. The
following conditions are equivalent.
(i) L satisfies conditions (R∆ )–(Re∆ ) and (MP∆ ).
(ii) The set ∆ defines the Leibniz congruence in every model of L , in the sense
that for every hA, Fi ∈ ModL and every a, b ∈ A , a ≡ b (Ω A F) if and only if
∆ A (a, b) ⊆ F .
(iii) The set ∆ defines equality in every reduced model of L , in the sense that for
every hA, Fi ∈ Mod∗L and every a, b ∈ A , a = b if and only if ∆ A (a, b) ⊆ F .
When these conditions hold, ∆ is called a set of equivalence formulas26 for L .
26 They are also called “congruence formulas”, for obvious reasons.
28 Josep Maria Font

The logics that satisfy any of these equivalent conditions are considered in Def-
inition 6.1. It is possible to show that conditions (Sym∆ ) and (Trans∆ ) can be dis-
pensed with in point (i) of this result.
By Theorem 4.1 and Lemma 4.10, algebraizable logics satisfy point (i) of Theo-
rem 4.11; this confirms that the congruence denoted simply as Ω Γ 0 in steps (LT2)
on pages 19 and 22 coincides with Ω FmΓ 0 (thus justifying the initial choice of
notation), and that the rôle played by conditions (R∆ )–(Re∆ ) in step (LT3) of the
algebraizable case corresponds to the fact that in the general case we are factoring
out by a congruence. Thus, from this point onwards, I write Ω instead of Ω Fm ,
even in the general case. Moreover, this coincidence also shows that the congruence
Ω Γ 0 defined above does not depend on the particular set ∆ used to define it; and in
particular that the algebras in (7) are the algebraic reducts of the Lindenbaum-Tarski
models of (10).
The other definability condition concerns truth; the set F in a matrix hA, Fi is
called the truth filter because when using matrices as semantics for logics as in (8)
or (9), the truth condition is “to belong to F ”. Using Theorems 4.5 and 4.9, one
finds the following.
Theorem 4.12 (Herrmann, 1993). Let L be a logic having a set ∆ of equivalence
formulas, and let E(x) ⊆ Eq. The following conditions are equivalent.
(i) L satisfies condition (ALG3) for ∆ and E .
(ii) The set E defines truth from the Leibniz congruence in every model of L , in
the sense that for every hA, Fi ∈ ModL and every a ∈ A , a ∈ F if and only if
E A (a) ⊆ Ω A F .
(iii) The set E defines truth from equality in every reduced model of L , in the
sense that for every hA, Fi ∈ Mod∗L and every a ∈ A , a ∈ F if and only if
A  E(x) [[a]].
Note that the set ∆ does not appear in conditions (ii) and (iii); actually, the equiv-
alence between these two is independent of the assumption about ∆ made in the
theorem, and the logics that satisfy either of these two conditions are considered in
Definition 6.1.
Again, comparison between steps (LT5) in the two versions of the process on
pages 22 and 26 shows that in the algebraizable case the truth condition has been
made definable by satisfaction of equations E . Thus, coupling Theorem 4.1 with the
two preceding ones we obtain:
Corollary 4.13. A logic is algebraizable if and only if, in all its reduced models,
equality is definable from the truth filter through a set ∆ (x, y) in the sense of Theo-
rem 4.11(iii) and the truth filter is definable from equality through a set E(x) in the
sense of Theorem 4.12(iii).
This view of algebraizability as mutual interdefinability of equality and the truth
filter allows us to obtain one of the few general results that characterize algebraiz-
ability for semantically defined logics (it is, however, of limited application). A finite
algebra is primal when for all n ∈ ω , all functions f : An → A are definable by a
term in n variables. Using just the definition, it is not difficult to show the following.
Abstract Algebraic Logic 29

Theorem 4.14. Let L be a logic defined by a finite matrix hA, Fi such that A is a
/ Then L is algebraizable.27
primal algebra and F 6= 0.
To obtain a more graphical rendering of the next property, it is useful to consider
the set
SolτA := a ∈ A : A  E(x) [[a]] ,


that is, the set of solutions of the equations E(x) that define the transformer τ ;
thus, it is what we could call, with a geometrical analogy, an “algebraic set”. Now,
condition 4.12(iii) says that F = SolτA for each hA, Fi ∈ Mod∗L ; from this it is
easy to prove that the classes of algebras obtained by the algebraizability approach
and by the universal matrix semantics coincide for algebraizable logics.
Corollary 4.15. If L is an algebraizable logic, with transformers τ and ρ , then
Mod∗L = hA, SolτA i : A ∈ Alg∗L and Alg∗L = K(L ,ττ ,ρ

ρ ), the equivalent al-
gebraic semantics of L .
Since Mod∗L and Alg∗L are intrinsic28 to L , this is one way of showing that
the notion of the equivalent algebraic semantics and the class K(L ,ττ ,ρ ρ ) are inde-
pendent of the actual transformers. In this case, there is a one-to-one correspondence
between reduced models and algebras of the equivalent algebraic semantics of L ;
and the algebraic completeness theorem (12) can be expressed solely in terms of the
class of algebras Alg∗L .

4.5 Implicative logics

Although the proof of Theorem 4.1 was presented as a natural generalization of the
Lindenbaum-Tarski process for classical logic and other similar logics, historically
it took a long time before this possibility was recognized and the general notion of
an algebraizable logic was isolated.
A first, much earlier generalization isolated the rôle of the implication connective
as used in Section 4.1; as a consequence, the class of logics it applies to is more
restricted than that of the algebraizable ones, but is still very large, and forms the
most popular kind of algebraizable logics.
Definition 4.16. An implicative logic is a logic L in a language L having a binary
term, represented here by x → y, such that the following conditions are satisfied.
(IL1) `L x → x.
(IL2) x → y , y → z `L x → z.
(IL3) x1 → y1 , . . . , xn → yn , y1 → x1 , . . . , yn → xn `L λ x1 . . . xn → λ y1 . . . yn , for
each λ ∈ L, of arity n > 0.
27Actually, the logic is BP-algebraizable, in the sense of Definition 5.1.
28In the sense that they are uniquely determined by L . In the reverse sense, only Mod∗L deter-
mines L , by Theorem 4.9; Alg∗L does not, as already mentioned at the end of Section 4.3.
30 Josep Maria Font

(IL4) x , x → y `L y.
(IL5) x `L y → x.

Theorem 4.17. All implicative logics are algebraizable, with defining equation
E(x) := {x ≈ x → x} and equivalence formulas ∆ (x, y) := {x → y , y → x}.

Proof. It is enough to check that the mentioned sets E and ∆ satisfy the conditions
of Theorem 4.1. For this ∆ , (Re∆ ) is exactly (IL1); (Sym∆ ) is contained in the very
definition of ∆ ; (Trans∆ ) and (Re∆ ) are straightforward consequences of (IL2) and
(IL3), respectively; and finally, in this case (ALG3) is x a`L {x → (x → x) , (x →
x) → x}, and this is easily proved using (IL1), (IL4) and (IL5). 

The converse of Theorem 4.17 does not hold. The class of implicative logics is
larger than that of the logics that are algebraizable in that way: the five conditions
defining implicative logics are close, but clearly stronger than those in Theorem 4.1;
this is apparent, for instance, if we compare (IL4), which is (MP→ ), with (MP∆ )
for ∆ (x, y) = {x → y , y → x}, which is the condition x , x → y , y → x `L y. Both
sets of conditions are designed to obtain a smooth and natural generalization of
the Lindenbaum-Tarski process; but while in algebraizable logics this is done by
generalizing the properties of the equivalence connective, in implicative logics it
is based on properties of implication, expressed through the binary relation 6Γ 0
def
defined by: α 6Γ 0 β ⇐⇒ Γ 0 `L α → β . It is easy to check that:
• (IL1) is equivalent to requiring that for any Γ 0 ∈ ThL , the relation 6Γ 0 is re-
flexive.
• (IL2) is equivalent to requiring that for any Γ 0 ∈ ThL , the relation 6Γ 0 is tran-
sitive.
Under these two conditions, 6Γ 0 is a quasi-order, and it is well known that its sym-
metrization, which coincides with Ω Γ 0 for ∆ (x, y) = {x → y , y → x}, is an equiva-
lence relation compatible with 6Γ 0 ; so that in the quotient, the quasi-order induces
an order 6 defined by
def
α/Ω Γ 0 6 β /Ω Γ 0 ⇐⇒ α 6Γ 0 β ⇐⇒ Γ 0 `L α → β .

Taking this into account, we can go on as follows.


• (IL3) is equivalent to requiring that for any Γ 0 ∈ ThL , the relation Ω Γ 0 is com-
patible with the algebraic operations of L; that is, it is a congruence of the for-
mula algebra. Thus, the quotient becomes an algebra of type L.
• (IL4) is equivalent to requiring that each Γ 0 ∈ ThL is an up-set of its associated
quasi-order 6Γ 0 (that is, α ∈ Γ 0 and α 6Γ 0 β imply β ∈ Γ 0 ).
• (IL5) is equivalent to requiring that each Γ 0 ∈ ThL constitutes a single equiv-
alence class (i.e., any two formulas in Γ 0 are equivalent modulo Ω Γ 0 ) and this
class is the maximum of the order 6 of the quotient. Note that the first of these
two properties amounts to requiring the condition:

x , y `L ∆ (x, y) (G∆ )
Abstract Algebraic Logic 31

for the particular ∆ we are dealing with here; this condition is a consequence of
(IL5) but is important in its own right, and reappears in Section 6.1.
(In each step, the properties of the preceding steps are assumed to hold.)
If we review the Lindenbaum-Tarski process again, we see that steps (LT1)–
(LT3) are obtained from conditions (IL1)–(IL3); (LT5) is a consequence of (IL4),
which implies that Ω Γ 0 is compatible with Γ 0 ; and in the discussion on page 19 it
was already stressed that in the case of classical logic this step is essentially due to
(MP→ ), here (IL4). Step (LT4) depends on the wise definition of the class K with
respect to which the completeness is to be proved, and in the case of implicative
logics this is actually the class K(L ,ττ ,ρρ ) for the transformers specified in The-
orem 4.17. By the facts pointed out in the previous paragraph, these algebras are
ordered. The stronger form of (IL4) and the additional property (IL5) allow us to
obtain here, unlike in the general case of algebraizable logics, step (LT6); that is,
that the theory Γ 0 constitutes a single element in the quotient Fm/Ω Γ 0 and that
this element is the top element of its order.
Thus, in this case the algebraic completeness is totally parallel to that of classical
or intuitionistic logic. Actually, Henkin (1950) observed that, after changing only
the class of algebras, the process works for the implication fragment of intuition-
istic logic. In parallel, Rasiowa and Sikorski (1953) did the same for its positive
fragment (i.e., admitting also conjunction and disjunction into the language). The
general theory was finally refined29 by Rasiowa (1974) by isolating the conditions
of Definition 4.16, which are somehow weaker and are, as discussed before, exactly
equivalent to what is required for the process to work in this “implicative version”.
Recently, this notion has been weakened and generalized in several natural ways
by Cintula and Noguera (2010), as shown in Definition 5.2, where the logics of
Definition 4.16 are called precisely Rasiowa implicative.
The class of implicative logics is certainly large. Besides all the fragments of
classical or intuitionistic logic containing implication (writing x → x instead of 1 if
necessary), it contains Johansson’s minimal logic, the global consequences of nor-
mal modal logics, Nelson’s constructive logic with strong negation and Rousseau’s
version of Post’s finitely-valued logics. Classical predicate logic, when formalized
as a deductive system satisfying Definition 2.1, as in Appendix C of Blok and
Pigozzi (1989), is also implicative. It is clear that any extension of an implicative
logic is also implicative, and that an expansion is implicative if and only if the new
connectives satisfy condition (IL3).
29 Rasiowa’s notion assumes two inessential requirements for the language: that it contains no
connectives of arity greater than 2, and that → is a primitive connective (rather than an arbitrary
term in two variables). Moreover, she considers finitarity as part of the definition of a logic. All
these restrictions are usually removed in later studies.
32 Josep Maria Font

5 Modes of algebraizability, and non-algebraizability

The notion of algebraizability contained in Definition 3.5 is a result of the original


definition in Blok and Pigozzi (1989) after removing some restrictions; namely, that
it applied only to finitary logics and to quasivarieties, and that the two transformers
were assumed finite. This move was due to Herrmann (1993) and has been adopted
in the literature; while the original notion is sometimes called algebraizability in the
sense of Blok and Pigozzi.30 The refinements of the notion mostly considered in the
literature are the following.

Definition 5.1. An algebraizable logic L is:


- finitely algebraizable when the set ∆ (x, y) of equivalence formulas is finite.
- regularly algebraizable when it moreover satisfies condition (G∆ ).
- finitely regularly algebraizable when it is both finitely and regularly algebraizable.
- BP-algebraizable when it is finitary and finitely algebraizable.

It may seem strange that the term “finitely” does not mean that the two transform-
ers are finite. This is due to historical reasons (see footnote 33) and to the (straight-
forward) fact that for finitary algebraizable logics, the set E(x) can always be taken
as finite; therefore, for finitary and finitely algebraizable logics both transformers
can be taken as finite, as in Blok and Pigozzi’s original definition. This explains the
name “BP-algebraizable”. Similarly, when K is finitary, the set ∆ (x, y) can always
be taken as finite.
Regularly algebraizable logics have a single defining equation, of the form x ≈ >,
where > is any theorem of the logic with at most one variable x. Since this equation
defines the truth filter SolτA of reduced models, it turns out that for these logics the
truth filter of reduced models is a one-element set.
Clearly, all implicative logics are in fact finitely regularly algebraizable. It is
instructive to review some less standard cases, which show the flexibility of these
general notions and the diversity of logics that fall under their scope.
• Łukasiewicz’s infinitely-valued logic, defined semantically from the unit real in-
terval with {1} as the truth filter, is known to be non-finitary, but it is still im-
plicative, thus falling under Theorem 4.17. Therefore, in this case, the two trans-
formers are finite; but both the logic and its associated equational consequence
(relative to the generalized quasivariety generated by the algebra on the unit real
interval, which is known not to be a quasivariety) are non-finitary.
• The so-called “Last Judgement” logic of Herrmann (1996) is a finitary modal
logic, very close to the local consequence of the minimal normal modal system
K , but without the Necessitation Rule even in its weak version (concerning only
theorems). It is algebraizable, with the single defining equation ¬x ≈ ¬(x → x)
and with the infinite set of equivalence formulas n (x → y), n (y → x) : n > 0 .


30 This phrase is also used to refer, in a more general way, to the idea of characterizing algebraiz-
ability as a relation between the consequence of a logic and the equational consequence of a class
of algebras through transformers, when compared with more distant approaches.
Abstract Algebraic Logic 33

One can show that no finite algebraization is possible, and that the equivalent
equational consequence is not finitary.
• Raftery (2010) constructed an example of a (non-finitary) logic that is finitely
algebraizable but needs an infinite set of defining equations, while its equivalent
algebraic semantics is a variety, and hence has a finitary equational consequence.
• All the substructural logics determined by a variety of residuated lattices studied
by Galatos et al. (2007) are finitely algebraizable, with (x\y) ∧ (y\x) as equiva-
lence formula, and with x ∧ 1 ≈ 1 as defining equation; therefore, the truth filter
of their reduced models has the form {a ∈ A : 1 6 a}. If the variety is integral,
then 1 (the unit of the monoid operation) is the maximum of the lattice order, and
this set reduces to {1}, which means the logic is finitely regularly algebraizable.
If the variety is not integral, then the set is not always unitary, and the logic is not
regularly algebraizable.
A related, intriguing OPEN PROBLEM is that to date, no algebraizable logic is known
whose set of defining equations can be finite but cannot consist of a single equation.
The dual situation is known to be possible: there are algebraizable logics that admit
a finite set of equivalence formulas but not a one-element set (Cintula and Noguera,
2010).
Observe that the Lindenbaum-Tarski process, when performed for a regularly al-
gebraizable logic, takes us to the first point in step (LT6) on page 19; namely, that
all formulas in Γ 0 are mutually equivalent and hence form a single point in the quo-
tient; this is an immediate consequence of condition (G∆ ). Thus, in all these cases
the algebraic models have a single designated element. It need not be, however,
the top element. Actually, in general, the algebras need not even be naturally or-
dered, although this certainly happens in implicative logics. Paradigmatic examples
of finitely regularly algebraizable logics that are not implicative are the equivalence
fragments of classical and of intuitionitic logic.
Actually, there is a qualitative break between implicative logics and algebraizable
logics in general, which you may have already noticed: in the former, the equiva-
lence set {x → y , y → x} has the special feature of being the “symmetrization” of
a simpler set of formulas, namely {x → y}, to which the notion of an implicative
logic assigns definite properties (moreover, this set is a singleton). Extending this
feature, Cintula and Noguera (2010) have introduced further modes of algebraiz-
ability31 which arise as a consequence of generalized modes of being implicative;
briefly, they are as follows.

Definition 5.2. Let ∇(x, y) ⊆ Fm be a set of formulas in two variables, and put
∇s(x, y) := ∇(x, y) ∪ ∇(y, x). Let L be a logic, and consider the following proper-
ties:

`L ∇(x, x) (R∇ )
∇(x, y) ∪ ∇(y, z) `L ∇(x, z) (Trans∇ )
31This approach gave rise to other classes of (non-algebraizable) logics, forming the so-called
implicational hierarchy, which is not described here.
34 Josep Maria Font
n
[
∇s(xi , yi ) `L ∇(λ x1 . . . xn , λ y1 . . . yn ) for all λ ∈ L , n = arλ > 0 (SRe∇ )
i=1

x , ∇(x, y) `L y (MP∇ )
x a`L ∇s E(x) for some E(x) ⊆ Eq

(ALG3)
x , y `L ∇(x, y) (G∇ )
x `L ∇(y, x) (W∇ )

A logic L is algebraically implicational when it satisfies the conditions (R∇ )–


(MP∇ ) and (ALG3) for some set ∇ of formulas and some set E of equations. If
condition (ALG3) is replaced by (G∇ ), L is regularly implicational; and if it is
replaced by (W∇ ), it is Rasiowa implicational. When the set ∇ is finite, the adjective
finitely is prepended to the name, and when it is a singleton, “implicational” is
replaced by implicative.

It is clear how these conditions generalize those previously considered; note,


however, the subtle difference between (Re∆ ) and (SRe∇ ): while the former has ∆ in
the assumption, the latter has the symmetrized set ∇s rather than ∇ itself (the “S” in
the label is for “symmetrized”). Obviously, (W∇ ) implies (G∇ ), and, using (MP∇ ),
it is easy to see that (ALG3) follows from (G∇ ), with E(x) := {x ≈ >}, where >
represents any theorem of the logic in at most the variable x. Thus, informally:

Rasiowa =⇒ regularly =⇒ algebraically

It is also easy to check, using Theorem 4.1, that “algebraically implicational” is the
same as “algebraizable”, with ∆ := ∇s as the set of equivalence formulas (hence
the same holds for its subclasses), and that “Rasiowa implicative” is the same as
“implicative” (Definition 4.16), with → as the single member of ∇. Thus, in all, we
have nine classes, of which four are new; they form the poset depicted in Figure 1.
The necessary counterexamples establish that they are all different and that no other
inclusion relations hold between them (besides the ones in the diagram and the ones
implied by it). For instance, the example of the equivalence fragment of classical
logic already mentioned is regularly implicative but not Rasiowa implicational; and
so on.
Although its primary motivation is exclusively syntactical, this approach has
some semantic consequences, concerning the relation defined by the set ∇ on the
matrix models hA, Fi of the logic by putting a 6A∇ b if and only if ∇A (a, b) ⊆ F .
In the present context, this relation is a quasi-order and the set F is an up-set with
respect to it. The quasi-order turns out to be an order if and only if the matrix is
reduced. Then the regularity condition (G∇ ) corresponds to the truth filter F of re-
duced models being a singleton, while the “Rasiowa” condition (W∇ ) corresponds
to this single element being the maximum of the order.
Note, however, that some conditions in Definition 5.2 are stronger than strictly
needed in order to show that the mentioned relation 6A∇ is an order in reduced
models of the logic. A finer analysis of this issue, (Raftery, 2013a), shows that a set
Abstract Algebraic Logic 35

implicative

 
regularly finitely Rasiowa
implicative implicational

   
algebraically finitely regularly Rasiowa
implicative algebraizable implicational

   
finitely regularly
algebraizable algebraizable

 
algebraizable

Fig. 1 Nine of the ten modes of algebraizability considered so far; the “BP-algebraizable” logics,
not shown here, are the finitely algebraizable ones that are finitary. The arrow means “included
in” (for the classes) or “implies” (for the corresponding properties). Thus, the diagram is “upside
down”, with larger classes in lower position; joins in the graph are in fact class intersections.

∇(x, y) has this property if and only if it satisfies the properties (R∇ ), (Trans∇ ) and

∇s(x, y) , δ (x,~z) `L δ (y,~z) for all δ (x,~z) ∈ Fm.

These three conditions together with the condition (ALG3) for ∇ turn out to char-
acterize syntactically the class of order algebraizable logics. These are logics that
enjoy a tight connection between their consequence relation and a relative “inequa-
tional” consequence (i.e., one similar to K but where equality is replaced by order),
namely a connection expressed by structural transformers (between formulas and
inequations) satisfying properties completely analogous to the (ALG1)–(ALG4) of
algebraizability. Roughly speaking, these logics are “algebraized” by a class of or-
dered algebras, i.e., algebras with an independent order relation,32 which explains
the name, and seem to be one solution to Pigozzi’s problem of finding an abstract
characterization of the notion of an implication; see Raftery (2013a) for a thorough
discussion and more references.
The semantic approach is essential when one wants to find counterexamples.
While Theorem 4.1 provides a very useful tool for proving that a given logic is alge-
braizable, its existential character (“there are transformers such that . . . ”) make it in
32 Although this is clearly one kind of algebra-based semantics, it departs from the general frame-
work of this chapter, and its exposition would exceed current space constraints. Let me just add that
order algebraizable logics in general need not be algebraizable; and that they are so if and only if
the order relation in their equivalent “ordered algebraic semantics” is equationally definable. They
are, though, stronger than equivalential, a significant class in the Leibniz hierarchy that appears in
Section 6.1.
36 Josep Maria Font

general useless (like the definition itself) when one wants to prove that a logic is not
algebraizable. To do this, lattice-theoretic characterizations of a universal character
have been developed with great success. They exploit the properties of the Leibniz
operator, the function F 7→ Ω A F restricted to the family FiL A of the L -filters of
an algebra A; i.e.,
Ω A : FiL A → Con Alg∗L A
(13)
F 7−→ Ω A F
where Con Alg∗L A := {θ ∈ ConA : A/θ ∈ Alg∗L }. Recall that by Corollary 4.15,
the equivalent algebraic semantics of an algebraizable logic L coincides with the
class Alg∗L of the Leibniz-reduced algebras of L .
It turns out that it is possible to characterize the very fact that a logic is algebraiz-
able by the behaviour of this operator. To reduce the burden of definitions and aux-
iliary properties, only the characterization concerning BP-algebraizability, which
is particularly simple, is reproduced here; other isomorphism theorems appear in
Table 2 on page 41 and in Theorem 8.2.
Theorem 5.3. Let L be a finitary logic and let K be a quasivariety. The logic L
is BP-algebraizable and K is its equivalent algebraic semantics if and only if for
each algebra A the Leibniz operator Ω A is an isomorphism between the lattice
FiL A of L -filters of A and the lattice Con K A of congruences of A relative to the
quasivariety K.
In general, for any class K of algebras and any algebra A , Con K A is the complete
lattice of congruences θ of A such that A/θ ∈ K; if K is a variety and A ∈ K, then
simply Con K A = ConA. The isomorphism and its inverse are definable as suggested
by Theorems 4.11 and 4.12:

F 7−→ Ω A F = ha, bi ∈ A × A : ∆ A (a, b) ⊆ F



−1 (14)
θ 7−→ Ω A θ = a ∈ A : E A (a) ⊆ θ


This result generalizes the well-known isomorphism between filters and congru-
ences in Boolean algebras or in Heyting algebras, as well as that between normal
subgroups and congruences in groups, or between ideals and congruences in rings.
Perhaps the most noteworthy feature of this result is that the isomorphism holds
for every algebra, and not just for algebras in K. This means that there is a lot of
freedom when constructing counterexamples that show the non-algebraizability of
certain logics by contradicting this isomorphism; that is, when constructing (finite,
small) algebras A such that Ω A is for instance non-monotonic, or non-injective, on
the L -filters of A; by Theorem 5.3, this would signify a failure of algebraizability,
no matter which class K one may have in mind. This is how the non-algebraizability
of ``S5 was first proved (Blok and Pigozzi, 1989, § 5.2.1).
From the consequences of Theorem 4.1 stated on page 17, it follows that a non-
algebraizability result entails the non-algebraizability of a number of other logics
related to the initial one; for instance, from the S5 case, the non-algebraizability of
virtually all local consequences of modal logics follows as well.
Abstract Algebraic Logic 37

This result can also be used to show that a given class of algebras cannot be logi-
fied by the paradigm of algebraizability; that is, that the class is not the equivalent
algebraic semantics of any BP-algebraizable logic. This is usually trickier, but is
based on the same idea, though in dual perspective: no matter which logic one has
in mind, if it should be algebraizable then the lattice structure of the family of its
filters on a given algebra would have to be isomorphic, through the Leibniz operator,
to that of its relative congruences (if the class is a variety and the chosen algebra be-
longs to it, this is just its congruence lattice); meanwhile, since the Leibniz operator
(and its inverse) is a purely algebraic object independent of the logic, knowledge
of the particular algebra and of its congruences may imply that such a situation is
impossible. This was done for the first time by Blok for the variety of distributive
lattices (see Font and Verdú 1991, page 397).

6 Beyond algebraizability

As explained in the Introduction, abstract algebraic logic goes far beyond alge-
braizability, and develops a richer framework. Within that framework, several larger
classes of logics have been identified and characterized in several ways, and have
been seen to be relevant for the study of the correspondences between logical prop-
erties and algebraic properties. In some sense, this study is the ultimate goal of
the subject. This very sketchy (but not short) section concentrates on the first is-
sue (description of the hierarchies), while the second (consequences, for a logic, of
belonging to a certain class) is touched upon in Section 7.

6.1 The Leibniz hierarchy

Strictly speaking, the term “Leibniz hierarchy” refers to a classification of logics


according to properties of the Leibniz operator (13). However, this should be taken
with a pinch of salt: there is not a single unifying scheme that uses the Leibniz op-
erator and encompasses all the classes of logics usually considered in the hierarchy.
The classification includes classes defined by related properties, mostly concerning
their classes of reduced models (the notion of a reduced model being itself defined
in terms of the Leibniz congruence), and for whose study the Leibniz operator is
still a useful tool.
The core of the hierarchy is of course the class of algebraizable logics. Corol-
lary 4.13 presents algebraizability as the conjunction of two definability conditions:
the definability of the Leibniz congruence in terms of the truth filter, and the defin-
ability of the truth filter in terms of the Leibniz congruence. Each of these conditions
defines per se a new, larger class of logics.
Definition 6.1. A logic is equivalential when there is a set of formulas ∆ (x, y) ⊆ Fm
that defines the Leibniz congruence in every model of the logic in the sense of
38 Josep Maria Font

Theorem 4.11(ii). If the set is finite, then the logic is called finitely equivalential.
A logic is truth-equational when there is a set of equations E(x) ⊆ Eq that defines
truth in every model of the logic in the sense of Theorem 4.12(ii).

Theorem 4.11 establishes equivalent conditions for the definition of an equiv-


alential logic; notably the purely syntactic property that the logic satisfies condi-
tions (R∆ )–(Re∆ ) and (MP∆ ). It is also easy to see that conditions (ii) and (iii) of
Theorem 4.12 are equivalent in general, without the assumption regarding ∆ present
there; so this establishes an equivalent definition of truth-equational logics. Thus, for
both properties, the definability requirement can be restricted to the reduced models
of the logic; in this case, since the Leibniz congruence becomes the equality relation
in these models, it is equality that is definable from the truth filter (for equivalen-
tial logics) and dually the truth filter is definable from equality (for truth-equational
logics).
Observe that Corollary 4.13 can now be rephrased as saying that a logic is
(finitely) algebraizable if and only if it is both (finitely) equivalential and truth-
equational.33
Equivalential logics were introduced by Prucnal and Wroński (1974), and the
first in-depth study of them is Czelakowski (1981). One can obtain the following
characterization, announced on page 15.

Theorem 6.2. A logic L is equivalential if and only if there is a set of formulas


∆ (x, y) ⊆ Fm and a class of algebras K such that L satisfies (MP∆ ) and the struc-
tural transformer ρ defined by ∆ satisfies condition (ALG2) with respect to K. In
this situation, one can always take K := Alg∗L .

The view of the equational definition of truth in models of a logic as a version


of Beth’s definability theorem is an idea of Herrmann (1993). Truth-equational log-
ics were formally introduced and first studied by Raftery (2006b); from the defini-
tions, it is easy to see that they enjoy a natural one-to-one correspondence between
Mod∗L and Alg∗L :

Lemma 6.3. Let L be a truth-equational logic, and let τ be the transformer defined
by the set of equations E(x) of Definition 6.1. Then:
1. hA, Fi ∈ Mod∗L if and only if A ∈ Alg∗L and F = SolτA .
2. A ∈ Alg∗L if and only if hA, Sol A i ∈ Mod∗L .
τ

Note that we are moving from more restricted classes of logics to larger, weaker
ones. Truth-equational logics form one of the very large classes of the hierarchy; in
fact, this class includes all assertional logics, but it does not include all equivalential
logics.
33 Finitely equivalential logics appeared earlier than finitely algebraizable ones. Thus, “finitely”
applied to algebraizable logics was adopted to indicate the finiteness only of ∆ , in order to obtain
the mentioned equivalence. A dual notion of a “finitely truth-equational logic” is not considered in
the literature.
Abstract Algebraic Logic 39

Assertional logics are defined at the beginning of Section 3 (page 10). Using
Proposition 39 of Raftery (2006b), it is easy to show that a logic L is assertional
if and only if its reduced models are unital; that is, the truth filter of the matrices
in Mod∗L is a one-element set. From this it follows that they are truth-equational,
with a truth definition of the form x ≈ >, where > is any theorem of L with at
most the variable x, and that if L is an assertional logic (i.e., the assertional logic
of some class K), then it is the assertional logic of the class Alg∗L .
The other very large class in the hierarchy is that of protoalgebraic logics, intro-
duced by Blok and Pigozzi (1986). This class can be described in so many ways that
it is difficult to choose one to begin with; here is the first, original definition.

Definition 6.4. A logic L is protoalgebraic when for each theory Γ ∈ ThL , in-
discernibility modulo Γ implies interderivability modulo Γ ; that is, when for any
α , β ∈ Fm, if α ≡ β (Ω Γ ), then Γ , α `L β and Γ , β `L α .

All the other classes of logics considered up to now, except assertional and truth-
equational logics, are contained in the class of protoalgebraic logics. The importance
of this class is due to its many different characterizations from quite different points
of view, and to the pleasant algebraic properties of their matrix models. One of the
key points is a matrix version of the correspondence theorem of universal algebra,
which means that the filters of the logic present, in a certain sense, nice behaviour
similar to that of the congruences of algebras.
Only two classes of logics in the Leibniz hierarchy remain to be introduced.
They were defined in a different framework, but thanks to the discovery of truth-
equational logics, they are now easy to describe.

Definition 6.5. A logic is weakly algebraizable when it is both protoalgebraic and


truth-equational; and it is regularly weakly algebraizable when it is weakly alge-
braizable and assertional, or equivalently, when it is protoalgebraic and assertional.

Weakly algebraizable logics were introduced34 in Font and Jansana (1996), and
studied thoroughly in Czelakowski and Jansana (2000), and also in Czelakowski
(2001), where regularly weakly algebraizable logics appeared. There are few proper
(i.e., non-algebraizable) examples of logics in these classes in the literature. The
best known, the logic of ortholattices described in Czelakowski and Jansana (2000),
belongs to the latter class, as do all orthologics that are not orthomodular; while the
“logic of Andréka and Németi” that appears in Blok and Pigozzi (1989, Appendix
B) belongs to the former but not to the latter. Neither of those logics is algebraizable.
Figure 2 depicts the organization of the eleven main classes in the hierarchy;35
most of the relations between them follow easily from the definitions given, and
34 They were first considered by Czelakowski, in unpublished lectures in 1993, under the name
“algebraizable in the weak sense”; a term also used in Czelakowski (2001).
35 Other classes might be considered in the hierarchy, though in a looser sense. These include

those obtained by restricting all the classes to their finitary members (among them is the class of
BP-algebraizable logics), and some of the classes of Definition 5.2, for instance the “Rasiowa”
classes.
40 Josep Maria Font

finitely regularly
algebraizable

 
finitely regularly
algebraizable algebraizable

   
finitely regularly weakly
algebraizable
equivalential algebraizable

   
weakly 
equivalential assertional
algebraizable

   
protoalgebraic truth-equational

Fig. 2 The main classes in the Leibniz hierarchy and their relations. The arrow means “included
in” or “implies”. Intersections of classes correspond to joins in the graph.

there are counterexamples showing that all inclusions are proper and that there are
no inclusions other than those shown in the figure and those implicit in it (by tran-
sitivity of inclusion). Moreover, the graphical joins correspond to intersections of
classes; for instance, we see that an assertional logic is equivalential if and only if it
is algebraizable, and if and only if it is regularly algebraizable,36 and so on.
Now it is time to describe the characterizations that give unity to the hierarchy,
and best explain both its structure and the relations between the classes. They are
displayed in several tables, in a rather informal and compact way.

Order-theoretic characterizations

These are the properties that give the core of the hierarchy its distinctive ab-
stract character, and concern properties of the Leibniz operator Ω A as a function,
mostly relating the order structure of its domain FiL A with that of its codomain
Con Alg∗L A. These sets are ordered under set inclusion, and one must take into ac-
count that for any algebra A, the function Ω A is always onto Con Alg∗L A. Table 2
summarizes the relevant results.
The properties in the table with a not-so-obvious meaning are defined as follows:
• Ω A commutes with endomorphisms when for any h ∈ End(A) and any F ∈
FiL A , Ω A h−1 F = h−1 Ω A F . The name of the property is shorthand for “com-
mutes with inverse images under endomorphisms”, which is certainly more ac-
36 And also if and only if it is order algebraizable (see page 35).
Abstract Algebraic Logic 41

L is . . . if and only if for every A, Ω A is . . . (and: if and only if over Fm , Ω is . . . )


Protoalgebraic monotone
Equivalential monotone and commutes with endomorphisms
Finitely equivalential continuous
Truth-equational completely order-reflecting
Weakly algebraizable monotone and injective (i.e., an isomorphism)
Algebraizable an isomorphism that commutes with endomorphisms
Finitely algebraizable a continuous isomorphism
Table 2 The order-theoretic characterizations.

curate; besides being more practical, there is a technical reason for using this
shorter name (by analogy with the result in Lemma 8.1).
• Ω A is continuous when, for any family {Fi : i ∈ I} ⊆ Fi A that is upwards-
L
directed and such that i∈I Fi ∈ FiL A, it holds that Ω A i∈I Fi = i∈I Ω A F i .
S S S

Notice that when L is finitary, the condition that the union of the family is a
filter is automatically satisfied. Clearly, continuity implies monotonicity.
• Ω A is completely order-reflecting when, for any family {Fi : i ∈ I} ∪ {G} ⊆
FiL A, if i∈I Ω A F i ⊆ Ω A G, then i∈I Fi ⊆ G.
T T

One important feature of the sets of properties of the Leibniz operator in each
row of Table 2 is that each set holds as stated (i.e., for all algebras) if and only
if it holds for just the operator as considered on the formula algebra Ω : ThL →
Con Alg∗L Fm; in this case, the endomorphisms are the substitutions. Results of this
kind (i.e., asserting that some property holds in the formula algebra if and only if
it holds, mutatis mutandis, in all algebras) are called transfer theorems in abstract
algebraic logic. They are often far from trivial, and they have an important impact
in the theory. Other examples of transfer theorems appear in Section 7.
Actually, Theorem 5.3 is a specialized version of the result in the last row of
Table 2: observe that by the result in the last row of Table 5, when the class Alg∗L
is a quasivariety, if the Leibniz operator is an isomorphism, then it is automatically
continuous.
The four classes not present in Table 2 are characterized by adding another prop-
erty of the Leibniz operator, of a different character. It is easy to see that37 a logic
L is assertional if and only if it has theorems and x ≡ y (Ω Γ ) for every Γ ∈ ThL
such that x, y ∈ Γ . By adding to this condition those for being protoalgebraic, equiv-
alential or finitely equivalential, we obtain characterizations, all in terms of the
Leibniz operator, for the three classes with “regularly” in their name; moreover,
once protoalgebracity is assumed, the condition can be simplified by writing just
x ≡ y Ω CL {x, y} , where CL {x, y} is the theory of L generated by the set {x, y}.
37 According to Czelakowski (1981), this fact was first stated, essentially, by Suszko in unpublished

lectures (here it is expressed in modern terms). Using the Suszko operator ΩL , to be
 defined in (17)

on page 47, the condition can be more compactly written as x ≡ y ΩL CL {x, y} .
42 Josep Maria Font

Definability characterizations

Some definability characterizations have already been encountered in previous def-


initions or results. The basic ones are summarized in Table 3.

L is . . . if and only if . . .
Protoalgebraic Ω A F is definable from F ∈ FiL A by some ∆ (x, y,~z) with pa-
rameters, for every A
Equivalential Ω A F is definable from F ∈ FiL A by some ∆ (x, y) without pa-
rameters, for every A
Truth-equational The truth filter is definable in Mod∗L by some E(x) ⊆ Eq
Assertional The truth filter is definable in Mod∗L by x ≈ >, where > is an
algebraic constant of Alg∗L (38 )
Table 3 The main definability characterizations.

It should be understood that in each case, the definability condition has a common
form in all the relevant algebras or matrices (i.e., this is about uniform definability).
By putting several of these conditions together, we obtain characterizations of four
more classes; while by requiring the non-parameterized set ∆ (x, y) to be finite, we
obtain the three remaining classes (those with “finitely” in their name). The only
new notion here appears in the first row (the parameterized case):
• If ∆ (x, y,~z) ⊆ Fm is a set of formulas in two variables x, y and possibly other
variables ~z, called parameters, then Ω A F is definable from S F ∈ FiL A by
∆ (x, y,~
z) when, for any a, b ∈ A, a ≡ b (Ω A F) if and only if ∆ A (a, b,~c) :
~

~c ∈ A ⊆ F . When parameters are absent, this property reduces to that of Theo-
rem 4.11.

Syntactic characterizations

Theorems 4.11 and 4.12 show that under certain assumptions the definability condi-
tions are equivalent to the logic satisfying certain properties for the sets of formulas
or equations involved. The simplest to state are summarized in Table 4; their main
interest is that they concern just the consequence relation of the logic.
By demanding that the set ∆ be finite, we obtain parallel characterizations for the
classes obtained by prepending “finitely” to the three39 last ones. The second and
38 This condition is equivalent to saying that the truth filter in Mod∗L is a singleton; but as a
“definability” property this formulation looks weaker.
39 There is no class of “finitely protoalgebraic” logics in the literature. One partial reason may be

that for finitary logics, the set ∆ (x, y) of Table 4 can always be chosen as finite in the protoalge-
braic case, but not necessarily in the other cases. There is also the technical issue mentioned in
footnote 40.
Abstract Algebraic Logic 43

L is . . . if and only if there is ∆ (x, y) ⊆ Fm such that `L satisfies . . .


Protoalgebraic (R∆ ) + (MP∆ )
Equivalential (R∆ ) + (MP∆ ) + (Re∆ )
Algebraizable (R∆ ) + (MP∆ ) + (Re∆ ) + (ALG3) for some E(x) ⊆ Eq
Regularly algebraizable (R∆ ) + (MP∆ ) + (Re∆ ) + (G∆ )
Table 4 Some of the syntactic characterizations.

third row correspond to Theorems 4.11 and 4.1, respectively; I already commented
that the conditions (Sym∆ ) and (Trans∆ ) can be dispensed with in these results.
More surprising is the fact that (ALG3) can also be dispensed with in the presence
of (G∆ ), as shown in the table.
The first row of the table deserves comment. It says that a logic L is protoal-
gebraic if and only if there is a set ∆ (x, y) of formulas in at most two variables
such that the logic satisfies the “law of identity” or “reflexivity” `L ∆ (x, x) and
the rule of “modus ponens” x, ∆ (x, y) `L y. In particular, any logic L in a lan-
guage containing a binary connective → such that `L x → x and x, x → y `L y is
protoalgebraic; this explains why the class of protoalgebraic logics is so huge!
In fact, most of the examples of non-protoalgebraic logics in the literature are
implication-less, such as the implication-less fragment of intuitionistic logic (Blok
and Pigozzi, 1989), the fragment of classical logic with just conjunction and dis-
junction (Font and Verdú, 1991), Dunn-Belnap’s four-valued logic (Font, 1997), or
positive modal logics (Jansana, 2002). Other examples of non-protoalgebraic log-
ics have a very weak implication, such as some weak subintuitionistic logics (Suzuki
et al., 1998; Celani and Jansana, 2001), Wójcicki’s “weak relevance” logic WR (Font
and Rodrı́guez, 1994) or a large number of logics that preserve degrees of truth with
respect to certain classes of residuated lattices (Bou et al., 2009).
Notice that there is nothing in conditions (R∆ ) and (MP∆ ) that is specific to
implication. In fact, the equivalence fragments of classical logic and of intuitionistic
logic are finitely regularly algebraizable, hence protoalgebraic, and satisfy these
conditions with ∆ (x, y) = {x ↔ y}.
There are some relations between the sets ∆ satisfying the conditions in Table 3
and those satisfying the conditions in Table 4. By Theorem 4.11, any set ∆ (x, y)
satisfying the conditions in the second row of Table 3 itself satisfies the second row
of Table 4, and conversely. As for the first row of each (the protoalgebraic case),
the relations are more complicated. If ∆ (x, y,~z) defines the Leibniz congruence as
in the first row of Table 3, then the set ∆ 0 (x, y) := ∆ (x, y,~x) satisfies the first row of
Table 4, i.e., (R∆ ) and (MP∆ ). The converse process requires a more complicated40
p
transformation:
S  y) satisfies the first row of Table 4, then the set ∆ (x, y,~z) :=
if ∆ (x,
∆ δ (x,~z), δ (y,~z) : δ (x,~z) ∈ Fm satisfies the first row of Table 3. For a related,
simpler relation, under stronger assumptions, see Lemma 6.17.
40 Observe that the set ∆ p is always infinite, irrespective of whether ∆ is finite or not. Even if a
protoalgebraic logic is finitary, the existence of a finite set defining the Leibniz congruence with
parameters cannot be guaranteed. Compare this with the fact mentioned in footnote 39.
44 Josep Maria Font

Model-theoretic characterizations

These classify a logic by closure properties of certain classes of matrices or algebras


related to it under some model-theoretic operations, and are summarized in Table 5.

L is . . . if and only if . . .
Protoalgebraic Mod∗L is closed under PSD
Equivalential Mod∗L is closed under S and P
Finitely equivalential Mod∗L is closed under S, P and PU , i.e., it is a “quasivariety”
and finitary of matrices

L is . . . if and only if L is truth-equational and . . .


Weakly algebraizable Alg∗L is closed under PSD
Algebraizable Alg∗L is closed under S and P
Finitely algebraizable Alg∗L is closed under S, P and PU , i.e., it is a quasivariety

Table 5 Two groups of model-theoretic characterizations. The classes Mod∗L and Alg∗L always
contain the trivial matrices (resp., algebras) and are closed under isomorphisms.

This is a good place to observe that matrices are just relational structures hA, Fi,
where besides the algebraic reduct A there is only one relation, a unary one, inter-
preted by the subset F ; thus, they are structures for a very simple first-order lan-
guage, and it is all the more natural that many tools and results of model theory can
be successfully used. This observation was first made by Bloom (1975) and besides
its application to the study of matrices for sentential logics, has also prompted a
certain amount of work on the model theory of equality-free sentences, which has
generalized techniques and benefited from intuitions coming from algebraic logic.
A recent contribution to this programme is Nurakunov and Stronkowski (2013); for
more details and older references see Font et al. (2003, Section 4.3).
The operators S, P, PU and PSD appearing in Table 5 are those of taking sub-
matrices/subalgebras, products, ultraproducts and subdirect products, respectively,
of a given class of matrices/algebras. The characterizations in the second group
may seem slightly unsatisfactory, as they need an extra assumption with no model-
theoretic character (that the logic is truth-equational); but they highlight precisely
the fact that for truth-equational logics the characterizations need not involve matri-
ces, but just plain algebras, and this is also an interesting feature. Note the presence
of finitarity in the last row of the upper half of the table, and its absence in the last
row of the lower half.
The results summarized in the tables in this section are, in fact, the outcome
of several important and technically involved theorems that belong to the estab-
lished core of abstract algebraic logic. It is clear that they cover interesting cases
not falling under the algebraizability paradigm. In the case of algebraizability, the
order-theoretic and the syntactic characterizations (Tables 2 and 4) are sometimes
called intrinsic (see Blok and Pigozzi, 1989, Chapter 4), meaning that they establish
Abstract Algebraic Logic 45

algebraizability by properties of the logic alone, without previous knowledge of the


class of algebras that constitute the algebraic counterpart of the logic. In this sense,
the syntactic characterizations (Table 4) are less intrinsic, as they still need knowl-
edge of the transformers; while the order-theoretic ones (Table 2) are truly intrinsic,
and their versions limited to the formula algebra are even more so, as they involve
only properties of the Leibniz operator (which is a purely algebraic object) on the
theories of the logic, without any additional component.

6.2 The general definition of the algebraic counterpart of a logic

Research on the Leibniz hierarchy has fully confirmed that the theory of matrices is
an adequate tool for the algebraic study of logics in this hierarchy. Recall that this
theory defines (page 26) the class

Alg∗L := A : there is F ∈ FiL A such that the model hA, Fi is reduced




as the algebraic counterpart of the logic L . This fits particularly well with the the-
ory of algebraizable logics, where it produces the equivalent algebraic semantics
(Corollary 4.15). However, it is not clear that Alg∗L is always the class of algebras
most naturally associated with an arbitrary logic L ; particularly when it is non-
protoalgebraic. This was first noticed by Font et al. (1991), in the study of CPC∧∨ ,
the fragment of classical logic with only conjunction and disjunction. Those authors
proved that Alg∗CPC∧∨ is neither the class of distributive lattices, nor that of its
bounded members, as one might expect, but a strange subclass with no other logical
or algebraic significance (and which is not even a quasivariety). Soon, other exam-
ples of a similar situation arose.41 This prompted the introduction of a more general
definition of the algebraic counterpart of a logic, which was achieved by consider-
ing a more general kind of algebra-based models, already introduced by Wójcicki
(1969) in essence.
A generalized matrix is a pair hA, C i where A is an algebra and C is a clo-
sure system42 of subsets of A, the universe of A. Observe that ordinary matrices
hA, Fi can be viewed as generalized matrices of the form hA, {F, A}i, and that a
logic L can also be viewed as the generalized matrix hFm, ThL i. This suggests
that generalized matrices may be a very flexible, convenient tool. In particular, they
41 This includes almost all the non-protoalgebraic logics mentioned on page 43. A strikingly simple
example is that of CPC∧ , the fragment of classical logic with only conjunction: while this logic
is naturally associated with the variety of semilattices, it is not difficult to show that Alg∗CPC∧
contains just the one- and two-element semilattices (Font and Moraschini, 2014, Corollary 5.3).
42 A closure system on a set A is a family C of subsets of A that contains A and is closed under in-

tersections of arbitrary non-empty families. The sets FiL A and ThL are always closure systems.
Originally, the C in a generalized matrix was not assumed to be a closure system, but an arbitrary
(non-empty) family of subsets; in this form they were rediscovered by Dunn and Hardegree (2001),
who called them “atlases”. There is no essential difference between the two alternatives as far as
their rôle as models of logics is concerned.
46 Josep Maria Font

incorporate a general semantic notion in a natural way: it is best expressed through


the closure operator 43 C that is associated with the closure system C by putting, for

X ⊆ A , C(X) := {F ∈ C : X ⊆ F}; then hA, C i is a generalized model (g-model


T

for short) of a logic L when for all Γ ∪ {ϕ} ⊆ Fm,



Γ `L ϕ ⇒ for all h ∈ Hom(Fm, A) , h(ϕ) ∈ C h(Γ ) . (15)

Thus, g-models incorporate a semantics for the consequence relation of a logic. This
notion of a model appears as strikingly natural if one wants to privilege the view of
logics as consequence relations. Moreover, it is more neutral as to the meaning of the
objects in the model, since it does not depend on the designation of some particular
subset of the algebra as representing “the truth”.
The comparison with matrix semantics is straightforward, because trivially hA, C i
is a g-model of L if and only if C ⊆ FiL A. Thus, the finest g-model of L on a
given algebra A is the generalized matrix hA, FiL Ai; g-models of this form (there
is exactly one on each algebra) are called basic full g-models. Recall (Lemma 4.4)
that FiL Fm = ThL ; therefore, each logic L , when viewed as the generalized
matrix hFm, ThL i, is its own basic full g-model on the formula algebra. The basic
full g-models hA, FiL Ai can be viewed as an “algebraic image” of the logic, and
a good deal of the research in this branch of abstract algebraic logic is concerned
with the study of which properties of the logic (expressible as properties of a gen-
eralized matrix) also hold for its basic full g-models; see the discussion on “transfer
theorems” and the sample of results in Section 7.
The technical tools needed to speak about the use of generalized matrices in
abstract algebraic logic start with associating a congruence with each of them in a
natural way. The Tarski congruence of a generalized matrix hA, C i is:

Ω A C := Ω AG : G ∈ C .
\
(16)

It is easy to see that this is the largest congruence of A that is compatible with all the
G ∈ C . A generalized matrix is reduced when its Tarski congruence is the identity
∼ ∼
relation. The reduction of hA, C i is the generalized matrix hA/Ω A C , C /Ω A C i,
∼A ∼A
where C /Ω C := {F/Ω C : F ∈ C }; as is to be expected, it is always reduced
(compare with Lemma 4.8). A generalized matrix is a full g-model of a logic L
when its reduction is a basic full g-model of L .
The corresponding completeness theorems, parallel to Theorems 4.5 and 4.9,
are formulated by putting an “ ⇐⇒ ” in (15) and requiring its right-hand side to
hold for all generalized matrices in a certain class. It is easy to see that every logic
is complete with respect to the following classes: all its g-models; all its reduced
43 A closure operator on a set A is a function C : P(A) → P(A) that satisfies, for all X ,Y ⊆ A,
that X ⊆ C(X) = C C(X) ⊆ C(X ∪ Y ). Several notational shortcuts are popular, such as writing
 
C(a) for C {a} , or C(X , a) for C X ∪ {a} , and so on; recalling that the argument of C should
always be a subset of A helps to avoid misunderstandings. The closure operator associated with the
closure system FiL A is denoted by FgL A ; thus, for any X ⊆ A , FgA (X) is the smallest L -filter of
L
A containing X . The closure operator associated with ThL is denoted by CL . A closure operator
C is finitary when for any X ⊆ A , C(X) = {C(Y ) : Y ⊆ X , Y finite}.
S
Abstract Algebraic Logic 47

g-models; all its basic full g-models; all its full g-models; and all its reduced full
g-models, which coincide with the reduced basic full g-models (it is easy to see that
a reduced full g-model must be basic).
Here I am concerned only with the use of g-models to define the algebraic coun-
terpart of a logic in the most general case, and to set up another hierarchy of logics,
the so-called Frege hierarchy.
The Tarski operator can be used to define another operator on filters of a logic

on each algebra A, the Suszko operator: the function ΩLA : FiL A → ConA that
assigns to each F ∈ FiL A the congruence
∼ ∼
ΩLA F := Ω A {G ∈ FiL A : F ⊆ G} = {Ω A G : G ∈ FiL A , F ⊆ G}.
\
(17)

This operator can also be considered on the formula algebra, i.e., defined on the
theories Γ of the logic; in this case, the superscript indicating the algebra is omitted,

and we write simply ΩL Γ .
In contrast with the Tarski and the Leibniz operators, which are purely algebraic
objects, the Suszko operator is strictly relative to the logic L , as reflected in the

notation. A matrix hA, Fi is Suszko-reduced when ΩLA F is the identity relation.
The Suszko operator makes sense because in some contexts (particularly when
working in parallel with the Leibniz operator) it seems desirable to have operators
on L -filters rather than on closure systems of L -filters. It was thoroughly studied
in general by Czelakowski (2003)44 and it was one of the main tools in the study of
truth-equational logics by Raftery (2006b).
The more general notion of the algebraic counterpart of a logic can be defined
with either the Tarski operator or the Suszko operator. It is the class of L -algebras,
defined (among others) in any of the following equivalent ways:

AlgL := A : there is C ⊆ FiL A such that the g-model hA, C i is reduced




= A : the basic full g-model hA, FiL Ai is reduced




= A : there is F ∈ FiL A such that the model hA, Fi is Suszko-reduced .




There are at least three groups of reasons why this more general notion is more
relevant for arbitrary logics than that of Alg∗L .
I The first is that in the cases where the old definition works, the two coincide:

Theorem 6.6.
1. A logic L is protoalgebraic if and only if the Leibniz and the Suszko operators

coincide on its filters, i.e., ΩLA F = Ω A F for all F ∈ FiL A and all A.
2. If L is protoalgebraic, then Alg∗L = AlgL .
3. If L is algebraizable, then AlgL is the equivalent algebraic semantics of L .
44That author attributes the definition and first characterization of this operator to Suszko, in
unpublished lectures.
48 Josep Maria Font

In particular, for implicative logics, the class AlgL coincides with the class
Alg∗L as originally defined by Rasiowa; in fact, the name “L -algebras” was coined
by her, and we now see that it can be used in general without risking confusion.
Notice that the converses of the implications in points 2 and 3 of Theorem 6.6 do
not hold: there are many non-protoalgebraic (hence non-algebraizable) logics L for
which Alg∗L and AlgL coincide; among them are all those where the former is a
quasivariety, in particular a variety.45 This fact follows from the most basic relations
between the two classes, which are summarized in the following result.

Theorem 6.7. For every logic L , AlgL = IPSD (Alg∗L ). Moreover:


1. The two classes generate the same quasivariety and the same variety. This variety
is called 46 the intrinsic variety of L , and is denoted by VL .
2. Alg∗L ⊆ AlgL ⊆ VL .
3. Alg∗L is a variety if and only if it coincides with VL . In such a case, it also
coincides with AlgL .

4. The variety VL is the variety generated by the quotient algebra Fm/Ω L , where

Ω L :=
T
Ω Γ : Γ ∈ ThL is the Tarski congruence of the logic viewed as a
generalized matrix.

The significance of VL as an algebraic semantics for L is in general weak,


because there is no general theory47 that asserts that the algebraic counterpart of a
logic should be a variety. However, if one insists on having a variety associated with
a logic, then VL is the natural choice. This is reinforced by the following facts (the
first is a consequence of Theorem 4.7, and the second is proved using the first).

Lemma 6.8. Let L be a logic. Then:


1. VL  α ≈ β if and only if for all δ (x,~z) ∈ Fm , δ (α ,~z) a`L δ (β ,~z).
2. If L is complete with respect to a class M of reduced matrices (or of reduced
generalized matrices), then the variety generated by the class of algebraic reducts
of the (generalized) matrices in M is VL .

I The second group of reasons is of an empirical character: in all particular cases


examined in the literature where the two classes differ, it has been found that the
class of algebras naturally associated with L according to other criteria or moti-
vated in other ways, even if only intuitively, is AlgL and not Alg∗L . For instance,
AlgCPC∧∨ is the variety of distributive lattices, AlgCPC∧ is the variety of semilat-
tices, and so on. Incidentally, concerning Theorem 6.7.2, neither of the inclusions is
an equality in general, and there are even logics for which the three classes in it are
different.
45 To the best of my knowledge, the earliest example of this kind in the literature is Wójcicki’s
“weak relevance” logic WR (Font and Rodrı́guez, 1994).
46 This name was first used by Pynko (1999).
47 Recall that the equivalent algebraic semantics of a BP-algebraizable logic is in general a quasiva-

riety. The best-known example of a BP-algebraizable logic whose equivalent algebraic semantics
is not a variety is BCK logic (Wroński, 1983).
Abstract Algebraic Logic 49

I The third group of reasons is that the approach based on generalized matrices,
and in particular the study, begun by Font and Jansana (1996), of the notion of a full
g-model and of the structure of the set of full g-models of a logic on a given algebra
has generated a deep and rich theory. Moreover, that theory establishes connections
between areas of abstract algebraic logic that are in principle unrelated. The first
important result in this area is yet another “isomorphism theorem”:
Theorem 6.9. For any logic L and any algebra A, the Tarski operator, that is, the

function C 7→ Ω A C , establishes a dual isomorphism between the complete lattice
of all full g-models of L on A and that of all congruences of A relative to the class
AlgL (both ordered under set inclusion).
One of the remarkable things about this result is its generality (it holds for all
logics whatsoever), which contrasts with the algebraizability assumptions needed
for Theorem 5.3 and for the isomorphisms in Table 2. There is not enough space
here to enter into this general theory; some of the results in Sections 6.3 and 7
actually belong to it.

6.3 The Frege hierarchy

This hierarchy is organized around several replacement properties that a logic and
its basic full g-models may have. These properties can be formulated in general for
an arbitrary generalized matrix; however, here it is better to go directly from the first
general definitions to the relevant particular cases (but see also Definition 7.3). The
Frege relation ΛA C of a generalized matrix hA, C i is defined, for any a, b ∈ A, as
 def
a ≡ b ΛA C ⇐⇒ C(a) = C(b).

In general, this relation is an equivalence relation, but it need not be a congruence



of A; the Tarski congruence Ω A C is the largest congruence of A below ΛA C (this
could be an alternative definition of the Tarski congruence). Moreover, this idea
produces the Frege operator: the function ΛAC that assigns, to every F ∈ C , the
Frege relation of the closure system {G ∈ C : F ⊆ G}. That is,

ΛAC F := ΛA {G ∈ C : F ⊆ G}.

Each logic L defines natural specializations of these constructions to the formula


algebra and to arbitrary algebras. If we consider the generalized matrix given by all
the theories ThL , then we obtain the relations

Λ L := ΛFm ThL , and


ΛL Γ := ΛFm {Γ 0 ∈ ThL : Γ ⊆ Γ 0 } for each Γ ∈ ThL .

Observe that Λ L is the interderivability relation of L , also denoted by a`L . Sim-


ilarly, given any algebra A and considering the closure system FiL A, we obtain
50 Josep Maria Font

ΛAL := ΛA FiL A, and


A
ΛL F := ΛA{G ∈ FiL A : F ⊆ G} for each F ∈ FiL A.

Thus, ΛAL identifies the elements of the algebra that generate the same L -filter.
In general, all these relations are equivalence relations, but need not be congru-
∼ ∼
ences; Ω L is the largest congruence below Λ L , ΩL Γ is the largest congruence
∼A ∼
below ΛL Γ , Ω FiL A is the largest congruence below ΛAL , and ΩLA F is the
largest congruence below ΛL A F.

Now, the four classes in the hierarchy can be introduced in the following compact
way.

Definition 6.10. Let L be a logic.



1. L is selfextensional when Λ L is a congruence; that is, when Λ L = Ω L .
2. L is Fregean when, for each Γ ∈ ThL , ΛL Γ is a congruence; that is, when

ΛL Γ = ΩL Γ .
3. L is fully selfextensional when, for each algebra A, ΛAL is a congruence; that

is, when ΛAL = Ω A FiL A.
4. L is fully Fregean when for each algebra A and each F ∈ FiL A, ΛL A F is a
A ∼A
congruence; that is, when ΛL F = ΩL F .

It is easy to use the characterization of the Leibniz congruence in Theorem 4.7


and the expressions (16) and (17) to obtain what could be an alternative, more “log-
ical” definition.

Lemma 6.11. A logic L is selfextensional if and only if for all α , β ∈ Fm,

if α a`L β then δ (α ,~z) a`L δ (β ,~z) for all δ (x,~z) ∈ Fm.

A logic L is Fregean if and only if for all Γ ∈ ThL and all α , β ∈ Fm,

if Γ , α a`L Γ , β then Γ , δ (α ,~z) a`L Γ , δ (β ,~z) for all δ (x,~z) ∈ Fm.

These properties may be viewed as formal counterparts of the idea of substitu-


tivity of equivalents. As discussed more extensively by Czelakowski and Pigozzi
(2004), these properties can be considered, to a certain extent, as abstract coun-
terparts of Frege’s principle of compositionality of denotation (i.e., of truth-value,
according to Church) when the notion of logical equivalence is understood as iden-
tity of truth-value. The abstract algebraic logic approach to the Fregean principles
was pioneered by Pigozzi (1991), building on the seminal ideas of Suszko (1975).
Under this view, selfextensional logics would also deserve to be called “Fregean”;
but the term “selfextensional” had already been coined by Wójcicki (1979). More-
over, starting with Font (1993), the term “Fregean” has been ascribed to the more
restricted class of logics having the property in its stronger form;48 the existence of
48In a few early papers the term “Fregean” means the same as “protoalgebraic and Fregean” in the
present terminology.
Abstract Algebraic Logic 51

strong relations with classes of algebras also called “Fregean” in the literature (see
Theorem 6.18) has consolidated this choice.
The other two kinds of logics are defined by requiring that these replacement
properties transfer to all basic full g-models.49 Fully selfextensional logics have also
been called congruential in a few recent papers; but be aware that the latter term has
also been used in the past, either for Fregean logics, or for finitely equivalential
logics. As a matter of fact, the four classes in this hierarchy deserve to be consid-
ered as “congruential” or “Fregean” in different degrees; these considerations justify
naming the hierarchy after Frege. The obvious relations between the four classes of
logics are those displayed in Figure 3.

fully Fregean

 
fully Fregean
selfextensional

 
selfextensional

Fig. 3 The poset of the classes in the Frege hierarchy. The arrow means “included in” or “implies”.
In this case, the graphical join is not class intersection (but see Corollary 6.20).

The following examples show that the four classes are distinct and no relations
other than those shown hold in general:
• All two-valued logics are Fregean. Classical and intuitionistic logics, as well
as many of their fragments (for instance, all those with either implication or
equivalence, but also the one with only conjunction and disjunction), are fully
Fregean.
• Logics that are fully selfextensional (hence selfextensional) but not Fregean
(hence, not fully Fregean either): Among the protoalgebraic logics, the local con-
sequences of normal modal logics; among the non-protoalgebraic ones, Dunn-
Belnap’s four-valued logic (Font, 1997); Wójcicki’s “weak relevance” logic WR
(Font and Rodrı́guez, 1994); the positive modal logics (Jansana, 2002); and the
local consequence of the weakest subintuitionistic logic (Celani and Jansana,
2001).
• A logic that is Fregean (hence selfextensional) but not fully selfextensional
(hence not fully Fregean either) was constructed by Babyonyshev (2003).
• Two ad hoc examples of logics that are selfextensional but neither Fregean nor
fully selfextensional were constructed by Albuquerque et al. (2016a). Interest-
ingly, one of these logics is implicative.
49 Equivalently, to all full g-models; hence the “fully” in the names.
52 Josep Maria Font

• Logics that are not selfextensional, and hence are totally outside the Frege hi-
erarchy: There are lots of them, even among algebraizable logics, for instance
the global consequences of normal modal logics, and the Łukasiewicz-style
many-valued logics. At the other end of the spectrum, the paraconsistent weak
Kleene logic (Bonzio et al., 2016) and Priest’s logic of paradox (Albuquerque
et al., 2016b) are non-protoalgebraic (but truth-equational) examples of non-
selfextensional logics.
• Logics that are neither selfextensional nor protoalgebraic and are not truth-
equational either (hence, logics outside the two hierarchies) are harder to find.
One example is the “logic of distributive bilattices” studied by Bou and Riviec-
cio (2011).
Thus, the two hierarchies are, in some sense, orthogonal, as these examples show:
there are logics located in a very high position in one of them and totally outside the
other. Some relations have been found, however, as can be seen below.
The replacement property of selfextensional logics, though weak, is still useful
when one is interested in the intrinsic variety of the logic; for in this case, this class
does have a clear logical meaning:

Lemma 6.12. A logic L is selfextensional if and only if there is a class K of alge-


bras such that for all α , β ∈ Fm , α a`L β if and only if K  α ≈ β . In this case,
VL is the largest class K, and the only variety with this property; in particular,
VL is the variety generated by any such class K.

This rather weak replacement property becomes surprisingly strong when it is


coupled with other properties (see Definition 7.3 for their exact meaning):

Theorem 6.13 (Font and Jansana, 1996). Let L be a finitary logic that has either a
conjunction, or a uniterm deduction-detachment theorem (DDT). If L is selfexten-
sional, then it is fully selfextensional. Moreover, in this situation, the class AlgL is
a variety, namely AlgL = VL .

This result is an interesting step in the analysis of THE VARIETY PROBLEM,


towards explaining why the algebraic counterpart of so many “real” logics is indeed
a variety. The problem50 is perhaps more striking in the context of BP-algebraizable
logics, where it was originally formulated by Blok and Pigozzi, because while the
general theory guarantees that the equivalent algebraic semantics of those logics
is a quasivariety, there are few examples of “real” BP-algebraizable logics whose
equivalent algebraic semantics is not a variety.51
Fully selfextensional logics have nicer properties. To start with, the following
reformulation of their definition, in terms of the class AlgL , reveals a particularly
interesting view.
50 Of course, this is not exactly a technical problem that admits a definite answer, but rather one
of methodology, of understanding the algebraic behaviour of logics and the limits of existing tech-
niques and results. Often these problems are more interesting than other, purely technical ones.
51 The algebraizable logics whose equivalent algebraic semantics is a variety are sometimes called

strongly algebraizable in the literature.


Abstract Algebraic Logic 53

Lemma 6.14. For any logic L , the following conditions are equivalent.
(i) L is fully selfextensional.
(ii) For each A ∈ AlgL , the relation ΛAL is the identity relation.
(iii) For each A ∈ AlgL , the relation 6AL , defined as a 6AL b if and only if b ∈
A {a}, for any a, b ∈ A, is an order relation.
FgL

This “logically defined” ordering relation in the algebras of AlgL has been ex-
ploited to develop an abstract version of duality theory (Esteban, 2013; Gehrke et al.,
2010). Examples of the nice behaviour of these logics are the following results.

Theorem 6.15 (Font and Jansana, 1996). Let L be a finitary and fully selfexten-
sional logic. Then:
1. AlgL is a quasivariety.
2. L is weakly algebraizable if and only if it is BP-algebraizable.

Some general consequences of the stronger replacement property of Fregean


logics are easy to obtain. For instance, since always x ≡ y (ΛL CL (x, y)), Defi-

nition 6.10.2 implies that these logics satisfy the condition x ≡ y (ΩL CL (x, y)) that
appears in footnote 37. Therefore, truth filters of reduced models of these logics on
non-trivial algebras are singletons, which offers a practical way to disprove that a
given logic is Fregean. Moreover, it turns out that in an important part of the Frege
hierarchy, the difference between the assertional logics and the truth-equational ones
disappears.

Theorem 6.16 (Albuquerque et al., 2016a). If L is a logic that is Fregean or fully


selfextensional, then L is assertional if and only if it is truth-equational. If L is
Fregean, then these conditions hold if and only if L has theorems.

In the protoalgebraic scenario, to be Fregean becomes a surprisingly strong prop-


erty. To begin with:

Lemma 6.17. Let L be a protoalgebraic logic, and let ∆ (x, y) satisfy properties
(R∆ ) and (MP∆ ) (as in Table 4 on page 43). If L is Fregean, then it is equivalen-
tial, with the symmetrized set ∆ s(x, y) := ∆ (x, y) ∪ ∆ (y, x) as a set of equivalence
formulas.

But the strength of Fregeanity for protoalgebraic logics goes farther: it places
them significantly higher up in the Leibniz hierarchy, and at the absolute top of the
Frege hierarchy.

Theorem 6.18. Let L be a protoalgebraic and Fregean logic. Then:


1. If L has theorems,52 then L is regularly algebraizable (Font and Jansana, 1996).
52 There is a protoalgebraic logic without theorems, but only one (in each language) and it is
rather trivial: it is the so-called almost inconsistent logic, i.e., the logic without theorems such that
α `L β for all α , β ∈ Fm. This logic is protoalgebraic and Fregean in a trivial way, but is not
algebraizable.
54 Josep Maria Font

2. If L is finitary, then it is fully Fregean (Czelakowski and Pigozzi, 2004).


3. If L is finitary and has theorems, then AlgL a Fregean quasivariety (Czela-
kowski and Pigozzi, 2004).

Different points in this result are proved using quite different techniques. In-
terestingly, the proof of point 2 is obtained by expressing the strong replacement
property as an infinite set of Gentzen-style rules and working with generalized ma-
trices as models53 of rules of this kind. Point 3 involves the notion of a Fregean
quasivariety, a quasivariety K that is pointed (there is a term, represented as >,
that is constant in K), relatively point-regular (for all A ∈ K and all θ , θ 0 ∈ Con K A,
if >A/θ = >A/θ 0 then θ = θ 0 ) and congruence-orderable (for all A ∈ K and all
a, b ∈ A, if ΘAK (a, >A ) = ΘAK (b, >A ), then a = b); here ΘAK (c, d) denotes the small-
est congruence of Con K A that identifies c and d . One can also prove the following
converse: if K is a Fregean quasivariety, then K = AlgL for a finitary, protoalge-
braic and Fregean logic L with theorems; namely, the assertional logic of K. So the
study of this class of logics amounts to the study of Fregean quasivarieties. Fregean
varieties have been extensively studied in universal algebra with logical motiva-
tions (Idziak, Słomczyńska, and Wroński, 2009); they were first incorporated into
the abstract algebraic logic landscape by Pigozzi (1991).
To close the section, let me mention a recent result on the location of truth-
equational logics in the Frege hierarchy, and three corollaries that follow by com-
bining this result with some previous ones; the second contains alternative charac-
terizations of the logics in Theorem 6.18.3.

Theorem 6.19 (Albuquerque et al., 2016a). A truth-equational logic is fully selfex-


tensional if and only if it is fully Fregean.

Corollary 6.20. If L is a logic with theorems, then L is fully Fregean if and only
if it is both Fregean and fully selfextensional.

Corollary 6.21. If L is a finitary logic, then the following conditions are equiva-
lent.
(i) L is protoalgebraic and Fregean and has theorems.
(ii) L is algebraizable and fully selfextensional.
(iii) L is weakly algebraizable and fully selfextensional.
(iv) L is regularly algebraizable and fully Fregean.

Corollary 6.22. If L is a finitary and weakly algebraizable logic, then L is


Fregean if and only if it is fully selfextensional, and if and only if it is fully Fregean.

Thus, for logics with theorems, the graphical join in Figure 3 on page 51 is in-
deed a class intersection. Moreover, in significant parts of the Leibniz hierarchy, the
53 This dual character of generalized matrices, as g-models of sentential logics and as models of
Gentzen-style rules or systems, is one of the features that give generalized matrices their special
interest.
Abstract Algebraic Logic 55

Frege hierarchy becomes simplified: inside truth-equational logics (Theorem 6.19),


it consists of three classes (the selfextensional, the Fregean, and the fully Fregean);
and inside finitary and weakly algebraizable logics (Corollary 6.22), it is dramat-
ically reduced to just two classes (the selfextensional and the fully Fregean). Al-
though some of this was known earlier, the fact that in each case the classes are
really different has been shown only recently (Albuquerque et al., 2016a).
Some OPEN PROBLEMS concerning the relations between the hierarchies that
come naturally to the mind are:
• Are fully selfextensional and Fregean logics always fully Fregean ? The answer
is affirmative for logics with theorems (Corollary 6.20), but neither a proof nor a
counterexample in the theorem-less case is known.
• Are protoalgebraic and Fregean logics always fully Fregean ? The answer is af-
firmative for finitary logics (Corollary 6.21), but neither a proof nor a counterex-
ample in the non-finitary case is known.
Research around these and similar questions is expected to shed more light on the
structure of the hierarchies and the relative strength of each of the classes.

7 Exploiting algebraizability: bridge theorems and transfer


theorems

In the title of this section, “algebraizability” is used in an extended, informal way,


to refer to any framework within which a general procedure or criterion has been
adopted in order to associate, with each logic L (perhaps only for logics of a cer-
tain kind), a class of algebras K as the algebraic counterpart of that logic, or, more
generally, a class of algebra-based structures, such as matrices or generalized ma-
trices. You are in such a situation when you restrict your attention to algebraizable
logics; for each of them, you get its equivalent algebraic semantics. You can also
consider matrix semantics, which produces, for an arbitrary logic L , the class of
matrices Mod∗L and the classes of algebras Alg∗L and AlgL ; or generalized
matrix semantics, which definitely points to AlgL but also to the classes of full
g-models, etc. To simplify this introductory discussion, we consider only classes of
algebras.
In a given framework, a bridge theorem is a mathematical result stating that for
a certain property P concerning a logic and a certain property P 0 concerning a class
of algebras, if L is a logic (in the class, if applicable) and K is the corresponding
algebraic counterpart, then:

L satisfies P if and only if K satisfies P 0 . (18)

The term “bridge theorem” was coined by Andréka, Németi, and Sain long ago (see
2001, pages 135–136 and 186–188); according to them, results of this kind establish
a bridge between two different lands, Logic and Algebra, and allow us to transform
56 Josep Maria Font

problems about a logic into problems about a class of algebras, so that we can use
the powerful tools of algebra, a much more intensively studied discipline, to solve
the problem, and then go back (crossing the bridge again) and obtain a solution to
the original logical problem. In particular, they emphasize that this methodology
allows us to cope with an ever-increasing forest of new logics using a more uniform
and better known toolbox, that of Algebra.
In the first, paradigmatic bridge theorems, P is a typically logical property (such
as interpolation) and P 0 is typically algebraic, i.e., it is a global property of a class of
algebras (such as amalgamation). In other bridge theorems, the property P 0 has the
form “every algebra in K satisfies property Q”, where Q is a property concerning a
single algebra (such as having a distributive congruence lattice); in yet other cases,
it is the property Q that refers to the class K, while the restriction “in K” in the
theorem can be deleted, and a stronger result is obtained.
A particular class of bridge theorems are those where both P and P 0 (or P and
Q) are essentially the same property, suitably interpreted on each side (such as “to
be finitely axiomatizable”). These results have been called transfer theorems, as
they can be phrased as “the property P transfers from L to K” (respectively, “to all
algebras in K”, or even “to all algebras”, in the stronger cases).
An important subclass of transfer theorems, where Q is just “P interpreted in
an algebra”, are those where P is a property of a closure system (or of a closure
operator), so that it applies to a logic L when we view it as a generalized matrix,
through the closure system ThL of its theories or its associated closure operator
CL 54 ; and it applies to an algebra A through the closure system FiL A of all the
L -filters of A or its associated closure operator FgL A . Thus, transfer theorems of

this kind state that

ThL (or CL ) satisfies P


if and only if (19)
for each A ∈ K , FiL A (or A
FgL ) satisfies P,

and, as already said, in many cases the limitation “∈ K” can be removed, so that the
property on the algebraic side is asserted to hold for all algebras, unconditionally.
Bridge and transfer theorems are not limited to frameworks where a class of
algebras is associated with a logic. Results with a larger scope refer to properties
of classes of matrices (the class of all models of the logic or, more often, the class
Mod∗L of its reduced models) or classes of g-matrices; for instance, observe that
a result like (19) in its strong version (i.e., without the “∈ K”) would be a transfer
theorem to all basic full g-models of the logic.
One is even tempted to say that bridge theorems and transfer theorems are the
ultimate justification of abstract algebraic logic. There are bridge theorems and
54 The operator CL is defined as ϕ ∈ CL (Γ ) if and only if Γ `L ϕ for any Γ ∪ {ϕ} ⊆ Fm.
Thus, essentially, the closure operator associated with a logic is just another way of expressing the
consequence relation. The properties formulated in its terms, such as those in Definition 7.3, are
more naturally stated in terms of the relation; but the general definition uses a closure operator to
facilitate application to both the logic and the algebras.
Abstract Algebraic Logic 57

transfer theorems of a wide range of degrees of difficulty, and some of them need
extensive development of the general theory, or complicated algebraic technicalities
(or both). Only a few are reproduced here, as each of them would need several
definitions, both on the logical and on the algebraic side, and space is limited. We
find results of all the different kinds described above.
A large group of bridge theorems are unconditional, that is, they hold for all
logics and all algebras. The following result is considered one of the oldest folk-
lore55 results of its kind, and contains both a bridge theorem and a transfer theorem,
concerning the property of being finitary.

Theorem 7.1. Let L be a logic. The following conditions are equivalent:


(i) L is finitary.
(ii) The class ModL is closed under ultraproducts.
A is finitary.
(iii) For every algebra A, the closure operator FgL

The characterizations in Table 2 (page 41) and in Table 5 (page 44) can be viewed
as bridge theorems, because they establish that certain properties of an arbitrary
logic (namely, the property of belonging to a certain class of the Leibniz hierarchy)
is equivalent to an algebraic or lattice-theoretic property of a class of matrices or of
algebras (the property may or may not involve the Leibniz operator). Moreover, as
already said there, some of these characterizations still hold when the properties are
required to hold just in the formula algebra, so that in fact they also involve transfer
theorems.

Theorem 7.2. For arbitrary logics, the following properties, if stated for the for-
mula algebra, transfer to all algebras.
• The order-theoretic properties of the Leibniz operator in Table 2 (page 41).
• The definability properties56 of the Leibniz operator in Table 3 (page 42).

Note that the transfer only holds for the properties when grouped as in the tables;
for instance, while injectivity does transfer when coupled with monotonicity, injec-
tivity alone is known to transfer when the language is countable, but not in general
(Moraschini, 2016).
Other (almost) universal transfer properties concern properties linking logical
connectives with the closure operator. Properties of this kind have been called
Tarski-style conditions by Wójcicki (1988); for simplicity I mention just a few of
them, and, for some, not the most general conceivable version.
55 This qualification certainly applies to the (easy) equivalence between (i) and (iii); the equiva-
lence between (i) and (ii) seems to have first been published by Zygmunt (1974).
56 It does not make sense to say, literally, that the definability of the truth filter in reduced models

of the logic (which appears in the third and fourth rows of Table 3) holds in the formula algebra,
because in general there are no reduced models on the formula algebra itself. But it can be refor-
mulated so that there is a transfer result for it: by Theorem 25 of Raftery (2006a), it holds in all
models that are reductions of arbitrary models of the logic (that is, in all its reduced models) if and
only if it holds in the models that are reductions of models on the formula algebra.
58 Josep Maria Font

Definition 7.3. Let C be a closure operator on the universe A of an algebra A, with


associated closure system C .
• A binary term x ∧ y is a conjunction for C when for all a, b ∈ A ,C(a ∧ b) =
C(a, b).
• C satisfies a deduction-detachment theorem (DDT)57 when there is a set I(x, y)
of binary terms such that for all X ∪ {a, b} ⊆ A , I A (a, b) ⊆ C(X) if and only if
b ∈ C(X , a). When the set I(x, y) consists of a single formula, the DDT is called
uniterm.
• A binary term x ∨ y is a disjunction58 for C when for all X ∪ {a, b} ⊆ A it holds
that C(X , a ∨ b) = C(X , a) ∩C(X , b).
• C satisfies an inconsistency lemma when there is a sequence hΨn : n > 1i of finite
sets of, respectively, n-ary terms, such that for all X ∪ {a1 , . . . , an } ⊆ A, the set
X ∪ {a1 , . . . , an } is C-inconsistent59 if and only if ΨnA (a1 , . . . , an ) ⊆ C(X).
• C satisfies the congruence property when its Frege relation ΛA C is a congruence

of A; i.e., when ΛA C = Ω A C .
• C satisfies the strong congruence property when for all F ∈ C , the Frege relation

ΛAC F is a congruence of A; i.e., when ΛAC F = ΩLA F for all F ∈ C .

It is trivial to see that the property of having a conjunction transfers to all algebras
for arbitrary logics. Among more interesting results we have the following.

Theorem 7.4.
1. The property of satisfying a DDT transfers to all algebras, for all logics. (Czela-
kowski, 1985; Raftery, 2011a; see footnote 61)
2. The property of having a disjunction transfers to all algebras, for finitary logics.
(Czelakowski, 1983)

The classical proofs, for the finitary case, are based on a common underlying
technique, arising from Hilbert-Bernays’ “Præmissentheorem” (1939), and which
can be generalized60 as follows.
57 Other, weaker versions of a DDT have been considered in the literature: The parameterized
one, which allows the terms of the set I to have parameters; the local one, which deals with a
(possibly infinite) family of sets of terms in such a way that only one set is used in each particular
application of the theorem; and obviously, the one that is both local and parameterized, which
turns out to characterize protoalgebraic logics. A fortiori, this implies that any logic satisfying a
DDT is protoalgebraic. The newest kinds of DDT are the graded ones (Font et al., 2001) and the
contextual ones (Raftery, 2011a); they have either local or parameterized variants as well.
58 An obvious generalization of this property is obtained when the single binary term is replaced

by a set of binary terms (as in the formulation given for the DDT); even more general is the
parameterized version, where a finite number of parameters are allowed in the set of terms. In both
cases the property is basically the same.
59 A set X is C-inconsistent when C(X) = A. The inconsistency lemma is a generalization of the

intuitionistic form of reductio ad absurdum, where Ψn (x1 , . . . , xn ) := {¬(x1 ∧ · · · ∧ xn )}.


60 This generalization corresponds, essentially, to the “operation on rules” considered, at the proof-

theoretic level, by Pogorzelski (1981); see also Pogorzelski and Wojtylak (2008).
Abstract Algebraic Logic 59

Lemma 7.5. Consider the following property of a closure operator C on some al-
gebra A:
There is a finite Γ (x,~y) ⊆ Fm such that for all ~a ∈ ~A and all n ∈ ω , if an ∈
C(a0 , . . . , an−1 ) then for all ~b ∈ ~A , Γ A (an ,~b) ∈ C Γ A (a0 ,~b) ∪ · · · ∪ Γ A (an−1 ,~b) .


For finitary logics, this property transfers to all algebras.

The restriction to finitary logics is very common in this area;61 other properties
have more restrictions, either on the logics they apply to, or on the algebras the
property is transferred to. For instance:

Theorem 7.6 (Raftery, 2013b). The property of satisfying an inconsistency lemma


transfers to all algebras, for finitary and protoalgebraic logics.

Meanwhile, the transfer of the congruence properties is not universal, not even
for finitary logics. Observe that a logic L is selfextensional if and only if it has the
congruence property; and it is Fregean if and only if it has the strong congruence
property. Some results in the previous section can now be rephrased as follows.
• The congruence property transfers to all algebras for finitary logics that have
either a conjunction or a uniterm DDT (Theorem 6.13).
• The strong congruence property transfers to all algebras for finitary and protoal-
gebraic logics (Theorem 6.18.2), and for fully selfextensional logics with theo-
rems (Corollary 6.20).
The two congruence properties have been known not to transfer in general since Baby-
onyshev (2003). Recently (Albuquerque et al., 2016a) the congruence property has
been shown not to transfer in any class of the Leibniz hierarchy, as it fails for a
finitely regularly algebraizable logic. The two OPEN PROBLEMS mentioned at the
end of Section 6.3 can also be rephrased as transfer problems:
• Does the strong congruence property transfer for theorem-less fully selfexten-
sional logics ?
• Does the strong congruence property transfer for non-finitary protoalgebraic log-
ics ?
The following bridge theorems concern, for an arbitrary algebra A, the family of
the finitely generated L -filters of A:

FiL A := F ∈ FiL A : F = FgLA


ω

(X) , X ⊆ A , X finite .

This family, ordered under inclusion, is a join-semilattice (it is in fact a join-sub-


semilattice of the filter lattice FiL A); this applies in particular to the family ThωL
61 In some cases where by tradition only finitary logics were considered, the properties may turn
out to hold in general. One example is the DDT (Theorem 7.4): Czelakowski’s original proof needs
finitarity, but Theorem 3.5 of Raftery (2011a) in particular shows the transfer for arbitrary logics.
In other cases finitarity seems to be unavoidable; the removal of finitarity from some transfer or
bridge theorems where it appears is an interesting OPEN PROBLEM.
60 Josep Maria Font

of the finitely generated theories of L . When L is finitary, by Theorem 7.1 the lat-
tice FiL A is algebraic, and the join-semilattice FiLω A has more algebraic content,

because in this case “finitely generated” is the same as “compact”.

Theorem 7.7 (Czelakowski, 1985). Let L be a finitary and protoalgebraic logic.


The following conditions are equivalent.
(i) L satisfies a DDT (with respect to some set I(x, y) of binary terms).
(ii) For all A, the join-semilattice FiL
ω A is dually residuated.62

(iii) The join-semilattice ThωL is dually residuated.

The equivalence between (i) and (ii) is a bridge theorem; while the implication
from (iii) to (ii) can be viewed as a transfer theorem. The bridge theorem has, as is to
be expected, more algebraic consequences, which in turn have logical applications
(crossing back over the bridge). For instance, since any algebraic lattice is isomor-
phic to the lattice of ideals of the join-semilattice of its compact elements, and taking
into account that satisfying a DDT implies protoalgebraicity (see footnote 57) we
obtain:

Corollary 7.8. If L is a finitary logic satisfying a DDT, then for every algebra A
the lattice FiL A is distributive.

Filter distributive logics (logics satisfying the conclusion of the corollary) have
been extensively studied63 in abstract algebraic logic. They enjoy some really in-
teresting features, resulting from the generalization of some well-known universal
algebraic properties to their matrix models. It turns out that protoalgebraicity greatly
enhances the possibilities for such generalizations. Just as a sample of what can be
achieved in this line of research, here is one:

Theorem 7.9 (Pigozzi, 1988; Czelakowski and Dziobiak, 1992). Let L be a finitary
logic, in a finite language, that is weakly complete64 with respect to a finite matrix.
If L is protoalgebraic and filter distributive, then L is finitely axiomatizable, and
it is so with a finite number of axioms and a single proper inference rule.

In parallel to Theorem 7.7, and with the same twofold character of a bridge and
a transfer theorem, we also have:

Theorem 7.10 (Raftery, 2013b). Let L be a finitary and protoalgebraic logic. The
following conditions are equivalent:
(i) L satisfies an inconsistency lemma.
(ii) For all A, the join-semilattice FiL
ω A is dually pseudo-complemented.

62 Dually residuated join-semilattices are sometimes called Brouwerian semilattices.


63 Not only in the context of protoalgebraic logics: any finitary logic having a disjunction is filter
distributive (Czelakowski, 1984).
64 The notion of weak completeness, as opposed to ordinary or strong completeness, refers to the

theorems of the logic rather than to the consequence relation.


Abstract Algebraic Logic 61

(iii) The join-semilattice ThωL is dually pseudo-complemented.

Up to now, each transfer theorem concerns a single, particular property. The fol-
lowing result brings together two of the few existing general transfer theorems; that
is, theorems regarding a potentially large group of properties, characterized in some
definite way:

Theorem 7.11. Let L be a finitary and protoalgebraic logic. The following groups
of properties transfer from L to all its full g-models on arbitrary algebras:65
1. All properties of the corresponding lattices of theories or of closed sets express-
ible as a universal sentence of the (first-order) language of lattices.
2. All properties that can be expressed as a set of accumulative66 Gentzen-style
rules.

Thus, when limited to basic full g-models, point 1 says that for a finitary and
protoalgebraic L , the lattice ThL satisfies a universal (first-order) property if and
only if for every algebra A, the lattice FiL A satisfies it. This includes some prop-
erties whose transfer had previously been shown by particular arguments, such as
distributivity (Corollary 7.8).
In some sense, the “best” bridge and transfer theorems are obtained for algebraiz-
able logics. In this case, the links between the logic and its equivalent algebraic se-
mantics are expressed syntactically, by means of two transformers, so that it is to be
expected that “metalogical” properties that can be expressed solely in terms of `L
can be easily “translated” into parallel properties of K . For instance, the following
should be immediately obvious:

Lemma 7.12. Let L be an algebraizable logic with finitary transformers, and let
K be its equivalent algebraic semantics. Then:
1. L is finitary if and only if K is a quasivariety.
If L is BP-algebraizable (thus, the conditions in point 1 hold), then:
2. L is finitely axiomatizable if and only if K is finitely based.
3. Theoremhood in L is decidable if and only if the equational theory of K is
decidable.
4. Finite consequences in L are decidable if and only if the quasi-equational theory
of K is decidable.

One of the keystones of bridge theorems for BP-algebraizable logics is Theo-


rem 5.3, which tells us that for each algebra, the lattice of filters of the logic is
65 Transfer to basic full g-models is due to both Czelakowski and Pigozzi; point 1 was published
in Czelakowski (2001), and point 2 in Czelakowski and Pigozzi (2004). The transfer to arbitrary
full g-models follows from the basic case and several results in Font and Jansana (1996).
66 Roughly speaking, a Gentzen-style rule is accumulative when an arbitrary set of “side assump-

tions” occurs (the same set) in the antecedent (left-hand side) of all the sequents in the rule.
62 Josep Maria Font

isomorphic to the lattice of congruences of the algebra relative to the equivalent al-
gebraic semantics of the logic, thus entering a more typically algebraic field. This
isomorphism can be restricted to the compact67 elements of the two lattices, and
so it maps FiL
ω A to Con ω A, the join-semilattice of compact relative congruences.
K
Then, Theorem 7.10 yields:

Theorem 7.13 (Raftery, 2013b). Let L be a BP-algebraizable logic, whose equiv-


alent algebraic semantics is a quasivariety K. The following conditions are equiva-
lent:
(i) L satisfies an inconsistency lemma.
(ii) For all A, the join-semilattice Con ω
K A is dually pseudo-complemented.
(iii) For all A ∈ K, the join-semilattice Con ωK A is dually pseudo-complemented.

The equivalence between (i) and (ii) is just the equivalence between the analo-
gous properties in Theorem 7.10, after considering the isomorphism of Theorem 5.3
(recall that every algebraizable logic is protoalgebraic). That the property in (ii) is
still equivalent if limited to algebras in K as in (iii) is a relevant fact; each of the two
properties has its strength in different applications, as commented on in footnote 72.
The most interesting results arise when the algebraic property is one that has been
studied due to independent interest in it within algebra. One of the most famous
theorems of this kind is the following.

Theorem 7.14 (Blok and Pigozzi, 2001, but originally proved before 1983). Let L
be a BP-algebraizable logic, whose equivalent algebraic semantics is a quasivariety
K. The following conditions are equivalent.
(i) L satisfies a DDT.
(ii) For all A, the join-semilattice Con ω
K A is dually residuated.
(iii) For all A ∈ K, the join-semilattice Con ωK A is dually residuated.
(iv) K has equationally definable principal relative congruences (EDPRC).68

The equivalence between (i) and (ii) follows from Theorems 7.7 and 5.3. Blok
and Pigozzi directly proved the equivalence between (i) and (iv) by showing that by
algebraizability the DDT for L is transferred to an analogous property for K , and
that this property is equivalent to (iv).
The equivalence between (iii) and (iv) seems a purely algebraic result, and in
fact it is: it holds for every quasivariety, regardless of whether it is the equivalent
algebraic semantics of an algebraizable logic or not. However, it is stimulating to
know that the first version of this equivalence, for the variety case, was obtained by
67 In this situation, the two lattices are lattices of closed sets of certain finitary closure operators,
therefore their compact elements are the finitely generated closed sets.
68 EDPRC is the analogue, for quasivarieties and relative congruences, of the property of hav-

ing equationally definable principal congruences (EDPC) for varieties, defined in Section 8 of
Raftery’s chapter in this volume. The property EDPC has been extensively studied in universal
algebra.
Abstract Algebraic Logic 63

Köhler and Pigozzi (1980) thanks to the intuitions that the link between the DDT and
EDPC suggests; with these intuitions, they also obtained an algebraic analogue of
Corollary 7.8, showing that EDPC implies congruence distributivity, which solved
a problem in universal algebra that had been open for some time. Actually, the very
notion of algebraizability was originally developed in order to build a framework
where the equivalence between the DDT and EDP(R)C could be rigorously formu-
lated and proved (Pigozzi, 1998). The work on the logical side made it apparent
that most of these purely algebraic investigations can be generalized from varieties
to quasivarieties if relative congruences enter into the picture (because relative con-
gruences are the algebraic analogue of the logical notions of a theory and a filter;
recall Theorem 5.3).
The other two most famous bridge theorems, proved under the same assumptions
as Theorems 7.13 and 7.14, can be roughly described as follows.
• L satisfies Craig’s interpolation property if and only if K satisfies the amalga-
mation property (Czelakowski and Pigozzi, 1999).
• L satisfies Beth’s definability property if and only if epimorphisms in K are
surjective (Blok and Hoogland, 2006).
I say “roughly” because each of the properties involved in the two results can be
formulated in several forms, of different strength; as a consequence, a plethora of
important theorems have been obtained. The cited references concern the exact for-
mulation of the results in the present framework, but the connections between these
pairs of properties had been observed and studied much earlier by several schol-
ars, and continue to be much extended, either restricted to classes of modal or su-
perintuitionistic logics (Maksimova), many-valued and fuzzy logics (Montagna) or
substructural logics (Wroński, Ono, Galatos, Kihara) or in more general contexts69
(Ono, Wroński, Németi, Sain, Madárasz). Accurate historical surveys are found in
the papers cited, as well as in Andréka, Németi, and Sain (2001, Section 6), but new
papers continue to appear.
The study of these connections under the perspective of abstract algebraic logic
has revealed that more general versions of these bridge theorems hold for equivalen-
tial logics (and hence apply to many more particular logics), provided that one deals
with matrices instead of plain algebras; in this case, the class of reduced models of
the logic plays the rôle of the equivalent algebraic semantics. Even more abstract
versions of the logical properties have been studied in later papers,70 where it turns
out that the categorial aspects of the algebraic properties play a prominent rôle;
69 The relations between (the different forms of) interpolation and amalgamation have also been
studied in other frameworks, such as equational logic, abstract model theory, and the theory of
institutions. The computational applications of interpolation seem to be partly responsible for such
interest in this property.
70 For instance: in the traditional versions of interpolation, the relation to be interpolated is either

consequence or a connective of implication. In some recent papers, it is equivalence that is inter-


polated, either understood as a connective or as the more abstract Leibniz congruence. Cabrer and
Gil-Férez (2014) study up to eleven different versions of interpolation of all these kinds.
64 Josep Maria Font

this makes it possible to study them not just in a category naturally associated with
algebras in K, but in equivalent categories.71
Besides their theoretical interest, bridge theorems are particularly useful when
applied to particular logics. Most often, they can be used to disprove that a certain
logic has a certain property by looking at the corresponding algebraic property, ei-
ther in general or inside the particular class of algebras or of matrices associated
with the logic. In some cases the task is to construct a particular counterexample; in
others, to search for known global properties of the class of algebras or matrices.72
These strategies have been successfully applied to several families of sentential log-
ics; the best known concern several versions of the DDT, of the interpolation prop-
erty, and of the Beth definability property, in fragments of classical or intuitionistic
logic, modal logics, many-valued logics, relevance logics, linear logics and sub-
structural logics, several versions of first-order logic (presented as a sentential-like
logic), etc. The reverse application is also possible; for instance, syntactic proofs of
the interpolation property have been used to infer amalgamation properties in the
associated classes of algebras.
Let me end this section by emphasizing that applications of algebraizability are
not limited to those of bridge theorems. There are logical problems for which there
are no precise, general bridge theorems available (yet) but which have been suc-
cessfully treated by algebraic means, i.e., after crossing the bridge on a less general
vehicle. One example of this kind of research, in the area of substructural logics,
is contained in the book by Galatos et al. (2007); while the general setting of the
book is explicitly the algebraizability paradigm, it takes advantage of several par-
ticular properties of the varieties of residuated lattices it deals with (for instance,
cut-elimination is proved thanks to several completion results, and in the reverse di-
rection semisimplicity of some free algebras is proved by proof-theoretic methods).
While the definition of algebraizability establishes a bridge between a single
logic and its algebraic counterpart, by Theorem 4.2 the bridge can be extended to
cover a large class of logics (its extensions) and their associated classes of algebras.
Its applications are often grounded on the following specialization.

Theorem 7.15. Let L be a BP-algebraizable logic whose equivalent algebraic se-


mantics is a variety V. Then there is a dual isomorphism between the lattice of all
the finitary extensions of L and the lattice of all sub-quasivarieties of V; this iso-
morphism restricts to one between the lattice of all the axiomatic extensions of L
and the lattice of all the sub-varieties of V.

This fact allows us to cross the bridge and place the study of some properties of a
logic in the land of universal algebra, where the study of lattices of (quasi)varieties
71 This idea was first pursued in Galatos and Raftery (2012), where, thanks to a category equiva-
lence between a class of integral residuated lattices and its non-integral counterpart, some metalog-
ical properties of certain logics having both a fuzzy and a substructural character were indirectly
obtained.
72 Observe that properties concerning the class of all algebras, such as those in points (ii) in The-

orems 7.13 and 7.14, offer more flexibility when it comes to attaining the first goal, while those
restricted to the class K , such as points (iii), may better support the second.
Abstract Algebraic Logic 65

has generated a number of very powerful results. Just as an example of a general


result that can be obtained in this way: Baker’s Theorem and Jónsson’s Lemma can
be used to show the following.

Theorem 7.16. If the equivalent algebraic semantics of a BP-algebraizable logic


is a congruence-distributive variety generated by a finite algebra, then the logic is
finitely axiomatizable and has only finitely many extensions.

It is, however, in the application to particular lattices of logics that this method-
ology has proved its strength.73 Of course, often the isomorphism can be (and has
been) established without appealing to the theory of algebraizability (but, in fact, by
using just a particular case of the general proof). However, knowing that the general
version holds whenever algebraizability is present should pave the way, and save
some work, for similar studies in other areas.

8 Algebraizability at a more abstract level

Soon after its invention, the idea of the algebraizability of a logic, in the sense of
its “equivalence” with the equational consequence relative to a class of algebras by
means of structural transformers, was extended to finite-dimensional deductive sys-
tems, to Gentzen systems, to logical systems based on hypersequents, etc. It was
clear that the whole framework would function wherever one could consider trans-
formers between the syntactic objects involved (formulas, m-formulas, sequents,
etc.) and equations, and back again, so that conditions parallel to (ALG1)–(ALG4)
would make sense; this would produce the parallel notions of algebraizable Gentzen
system (Rebagliato and Verdú, 1993), etc.
Moreover, it also became clear that this was an instance of a more general idea
of equivalence between two consequence relations, and that the fact that one of
them is a relative equational consequence was not essential: equivalence between
any of the kinds of consequences involved could be considered. This was done by
Raftery (2006a) for Gentzen systems, treating algebraizability as a special case of
the equivalence between two Gentzen systems of the most general kind. As a bonus,
the equivalence between two sentential logics,74 and that between a sentential logic
and a Gentzen system, were also studied; he applied this to shed new light on the
relations between different formalizations of some substructural logics.
73 The work done by Blok in the 1970s on several lattices of modal logics (see Rautenberg, Za-
kharyaschev, and Wolter, 2006, for a survey) is noteworthy, besides its intrinsic interest, for two
reasons: he was the first to develop this kind of application, and it was one of his main sources of
inspiration for the general theory of algebraizability.
74 In the literature, a relation between two sentential logics expressed by means of a transformer

in a way similar to (ALG1) has been named in a variety of ways; the term translation is popular,
sometimes with either conservative or faithful prepended to it, but these are not used in a uniform
way. Notice that the equivalence to which I refer here is stronger than this (the transformer is
“invertible”), and that some translations considered in the literature are not structural.
66 Josep Maria Font

The steps from the Lindenbaum-Tarski completeness of implicative logics to al-


gebraizability, and from algebraizability to the equivalence of deductive systems of
more general kinds, are generalizations. The next step in the research into alge-
braizability was an abstraction. To understand its basic idea we need to review the
characterizations of algebraizability in Theorem 5.3 and in the next-to-last row of
Table 2 on page 41. These characterizations mention the lattices FiL A and Con K A,
but almost nothing has been said about their lattice structure up to now. These sets
are easily seen to be closure systems, and hence complete lattices when ordered un-
der set inclusion, with intersection as the infimum operation. As to the supremum of
a family, it is computed by closing the union of the family under the closure operator
associated with the closure system; these closure operators are denoted by FgL A and

ΘK respectively; that is, FgL (X) is the smallest L -filter of A containing X ⊆ A,


A A

and ΘAK (θ ) is the smallest K-congruence of A containing θ ⊆ A × A. Thus, these


suprema are:
A
θi := ΘAK
_ [  _ [ 
Fi := FgL Fi and θi
i∈I i∈I i∈I i∈I

for {Fi : i ∈ I} ⊆ FiL A and {θi : i ∈ I} ⊆ Con K A.


Consideration of these lattice structures facilitates a better understanding of the
property that the Leibniz operator “commutes with endomorphisms”. Every h ∈
End(A) induces a function h−1 : FiL A → FiL A, because the inverse image of an
L -filter is always an L -filter; then the commutativity property defined on page 41
asserts that the Leibniz operator commutes with this function: Ω A h−1 F = h−1 Ω A F .
But h also induces another function on the set of L -filters, obtained by lifting it
to the power set and closing the result; if we define hL F := FgL A (hF) for each

F ∈ FiL A, then hL : FiL A → FiL A. Similarly, by considering the extension of


h to the Cartesian product A × A, we obtain a function on Con K A, namely hK θ :=
ΘAK (hθ ) for each θ ∈ Con K A. Then one can prove the following.75
Lemma 8.1. Let Φ : FiL A → Con K A be a lattice isomorphism, and h ∈ End(A).
The following conditions are equivalent.
(i) Φ hL F = hK Φ F for all F ∈ FiL A.
(ii) Φ h−1 F = h−1 Φ F for all F ∈ FiL A.
When these properties hold for all h ∈ End(A), we say that Φ commutes with en-
domorphisms of A.
By condition (i), such a Φ is a homomorphism of the enriched structures ob-
tained when the lattices FiL A and Con K A are expanded with the functions hL
and hK , respectively, considered as unary operations, for all h ∈ End(A). All these
constructions, of course, can be carried out in the particular case of the formula
algebra.76 The characterization in the next-to-last row of Table 2 (page 41) can be
75 This is a particular case, tailored for this point of the exposition, of a more general and abstract
result, which should be easy to guess (Font, 2016, Theorem 1.56).
76 Recall that in this case, Fi Fm = ThL and hence that FgFm = C , and that the endomor-
L L L
phisms of Fm are the substitutions.
Abstract Algebraic Logic 67

rephrased by saying that L is algebraizable if and only if, for each algebra A, Ω A
is an isomorphism between these expanded lattices.
But there is much more. The mere existence of an isomorphism of this kind is
enough to guarantee algebraizability:
Theorem 8.2 (essentially, Blok and Pigozzi, 1989, 2001). Let L be a logic and let
K be a generalized quasivariety. The following conditions are equivalent:
(i) L is algebraizable with equivalent algebraic semantics K.
(ii) For

every algebra A there is an isomorphism between the expanded lattices
FiL A, hhL : h ∈ End(A)i and Con K A, hhK : h ∈ End(A)i .


(iii) There is an
isomorphism

between the expanded lattices ThL , hσL : σ ∈
End(Fm)i and Con K Fm, hσ K : σ ∈ End(Fm)i .
Observe that in condition (iii) “points” (formulas and equations) have almost dis-
appeared; the condition concerns the expanded lattices of closed sets77 of the two
closure operators (corresponding to the closure relations `L and K ) whose equiv-
alence through structural transformers constitutes the definition of algebraizability.
Another equally essential component should be highlighted before going on: the
two closure operators involved are structural, which is also a property that relates
them to the endomorphisms of the formula algebra, and which can equivalently be
formulated in terms of the closed sets. For instance, it is easy to see that property (S)
in Definition 2.1 is equivalent to the condition that σ −1Γ ∈ ThL for all Γ ∈ ThL
and all σ ∈ End(Fm); this property is then propagated to the L -filters of arbitrary
algebras: if F ∈ FiL A then h−1 F ∈ FiL A for all h ∈ End(A).
A result parallel to Theorem 8.2 for the notion of equivalence between two
Gentzen systems was obtained by Raftery (2006a, Theorem 6.8). It seems clear
that a similar result can be obtained if one develops a theory of equivalence between
hypersequent systems, or between finitely-sided sequents, and so on. However, for
this, a parallel theory of matrix models, filters, Leibniz operator, etc., has to be devel-
oped in each case, with a formalism heavily dependent on the grammatical structure
of the languages involved. The subject naturally called for a more abstract setting
that would free it from this dependence on the languages on which the deductive
systems are defined.
This move was made by Blok and Jónsson (2006), who considered this kind of
isomorphism theorem as capturing the abstract essence of algebraizability. They
jumped one abstraction level up, and placed the study in the landscape of structural
closure operators on M-sets, that is, sets endowed with the action of a monoid M
(the abstraction of the monoid of substitutions of the formula algebra, which acts on
the sets of formulas, of equations, of sequents, etc.). The structurality of a closure
operator on the universe of an M-set is defined in terms of these actions, and the
lattice of closed sets is also expanded by these actions.
The next level of abstraction was introduced by Galatos and Tsinakis (2009). It
consists in “getting rid of points” completely, by observing that most of the work
77 The relative congruences of the formula algebra, Con Fm, are the closed sets (the “theories”)
K
of the equational consequence K .
68 Josep Maria Font

is already done at the level of subsets (the consequences, the filters, etc.), and that
the endomorphisms and the transformers, which originally act on points, can be ex-
tended to sets by taking unions, so that all the relevant definitions (e.g., structurality,
or the four algebraizability conditions) can be equivalently formulated by referring
only to subsets. Since complementation plays no rôle here, the abstract structures
corresponding to power sets are complete lattices, and the abstract correlate of point
functions extended to sets by taking unions are residuated functions; some of these
functions (those corresponding to the substitutions) act on the elements of the lat-
tice, giving it the structure of a module. Thus, the theory is placed in the framework
of modules over complete residuated lattices and closure operators78 over them.
In the two abstract approaches just mentioned, the notion of equivalence between
structural closure operators appears naturally, and is defined in terms of the exis-
tence of an isomorphism between the associated expanded lattices of closed sets
or, in the second case, the associated modules of fixed points. A new problem thus
arises: that of identifying conditions under which any such isomorphism, assum-
ing it exists, is induced by transformers, as the Leibniz operator is in the original
setting. This leads to the problem of identifying the abstract correlate of the proper-
ties of the algebra of formulas that make the equivalence in Theorem 8.2 possible,
and has come to be known as THE ISOMORPHISM PROBLEM. Some sufficient, but
not necessary, conditions had already been found by Blok and Jónsson (2006); the
solutions obtained at different levels by Galatos and Tsinakis (2009), and by Font
and Moraschini (2015) are formulated in categorial terms, involving the notion of
projectivity.
These lines of research seem to be expanding further in several directions. Both
the idea of equivalence between deductive systems and the isomorphism problem
can be studied in other contexts. Let me just mention that Galatos and Gil-Férez
(2016) jump up again, to the categories of modules over quantaloids, which allows
them to merge this line of research with the already very fruitful one of categorical
abstract algebraic logic pursued by Voutsadakis (see 2003, and many others), cover-
ing many-sorted logics. Meanwhile, Russo (2013) addresses the issue of considering
the equivalence between deductive systems over different languages at these more
abstract levels. Furthermore, Moraschini (2016) develops a new concept of transla-
tion between equational consequences relative to generalized quasivarieties (again,
possibly over different languages) which characterizes the adjunctions between the
associated categories, and uses this to prove that every generalized quasivariety is
categorially equivalent to the equivalent algebraic semantics of a finitely algebraiz-
able logic (possibly over a different language); in particular, every quasivariety is
categorially equivalent in this sense to the equivalent algebraic semantics of a fini-
tary and finitely algebraizable (i.e., BP-algebraizable) logic.

I hope that there is more to come.


78 Now in the purely order-theoretic sense, as opposed to closure operators on power sets as de-
scribed in footnote 43.
Abstract Algebraic Logic 69

Acknowledgements I am most grateful to Tommaso Moraschini and to an anonymous referee,


whose many observations concerning details as well as more substantial remarks decidedly im-
proved the exposition presented here. While writing this chapter, I was partially funded by the
research project MTM2011-25747 from the government of Spain, which includes FEDER funds
from the European Union, and the research grant 2009SGR-1433 from the government of Catalo-
nia.

References

H. Albuquerque, J. M. Font, R. Jansana, and T. Moraschini. Assertional logics, truth-equational


logics, and the hierarchies of abstract algebraic logic. In J. Czelakowski, editor, Don Pigozzi
on Abstract Algebraic Logic and Universal Algebra, series Outstanding Contributions to Logic.
Springer-Verlag, 2016a. 25 pp. To appear.
H. Albuquerque, A. Přenosil, and U. Rivieccio. An algebraic view of super-Belnap logics. Sub-
mitted manuscript, 2016b.
H. Andréka, I. Németi, and I. Sain. Algebraic logic. In D. M. Gabbay and F. Guenthner, editors,
Handbook of philosophical logic, second edition, volume 2, pages 133–248. Kluwer Academic
Publishers, 2001.
I. H. Anellis and N. Houser. Nineteenth century roots of algebraic logic and universal algebra. In
H. Andréka, J. D. Monk, and I. Németi, editors, Algebraic logic, volume 54 of Colloquia Math.
Soc. Janos Bolyai, pages 1–36. North-Holland, Amsterdam, 1991.
S. Babyonyshev. Fully Fregean logics. Reports on Mathematical Logic, 37:59–78, 2003.
W. Blok and E. Hoogland. The Beth property in algebraic logic. Studia Logica, 83:49–90, 2006.
W. Blok and B. Jónsson. Equivalence of consequence operations. Studia Logica (Special issue in
memory of Willem Blok), 83:91–110, 2006.
W. Blok and D. Pigozzi. Protoalgebraic logics. Studia Logica, 45:337–369, 1986.
W. Blok and D. Pigozzi. Alfred Tarski’s work on general metamathematics. The Journal of
Symbolic Logic, 53:36–50, 1988.
W. Blok and D. Pigozzi. Algebraizable logics, volume 396 of Mem. Amer. Math. Soc. A.M.S.,
Providence, January 1989. Out of print. Scanned copy available from http://orion.math.
iastate.edu:80/dpigozzi/.
W. Blok and D. Pigozzi. Abstract algebraic logic and the deduction theorem. Manuscript, 2001.
Available from http://orion.math.iastate.edu:80/dpigozzi/.
W. Blok and J. Rebagliato. Algebraic semantics for deductive systems. Studia Logica (Special
issue on Abstract Algebraic Logic, Part II), 74:153–180, 2003.
S. L. Bloom. Some theorems on structural consequence operations. Studia Logica, 34:1–9, 1975.
S. Bonzio, J. Gil-Férez, F. Paoli, and L. Peruzzi. On paraconsistent weak Kleene logic. Submitted
manuscript (2016).
G. Boole. The mathematical analysis of logic. Being an essay towards a calculus of deductive
reasoning. Macmillan, Cambridge, 1847.
F. Bou, F. Esteva, J. M. Font, A. J. Gil, Ll. Godo, A. Torrens, and V. Verdú. Logics preserving
degrees of truth from varieties of residuated lattices. Journal of Logic and Computation, 19:
1031–1069, 2009.
F. Bou and U. Rivieccio. The logic of distributive bilattices. Logic Journal of the IGPL, 19:
183–216, 2011.
S. Burris and J. Legris. The algebra of logic tradition. In E. N. Zalta, editor, The Stanford En-
cyclopedia of Philosophy. Spring 2015 edition. Archived as: http://plato.stanford.edu/
archives/spr2015/entries/algebra-logic-tradition/.
L. M. Cabrer and J. Gil-Férez. Leibniz interpolation properties. Annals of Pure and Applied Logic,
165:933–962, 2014.
70 Josep Maria Font

C. Caleiro and R. Gonçalves. Abstract valuation semantics. Studia Logica (Special issue on
Abstract Algebraic Logic), 101:677–712, 2013.
S. Celani and R. Jansana. A closer look at some subintuitionistic logics. Notre Dame Journal of
Formal Logic, 42:225–255, 2001.
A. Chagrov and M. Zakharyashev. Modal Logic, volume 35 of Oxford Logic Guides. Oxford
University Press, 1997.
P. Cintula and C. Noguera. Implicational (semilinear) logics I: a new hierarchy. Archive for
Mathematical Logic, 49:417–446, 2010.
P. Cintula and C. Noguera. A general framework for Mathematical Fuzzy Logic. In P. Cintula,
P. Hájek, and C. Noguera, editors, Handbook of Mathematical Fuzzy Logic, vol. 1, volume 37
of Studies in Logic. College Publications, London, 2011.
P. Cintula and C. Noguera. A Henkin-style proof of completeness for first-order algebraizable
logics. The Journal of Symbolic Logic, 80:341–358, 2015.
J. Czelakowski. Reduced products of logical matrices. Studia Logica, 39:19–43, 1980.
J. Czelakowski. Equivalential logics, I, II. Studia Logica, 40:227–236 and 355–372, 1981.
J. Czelakowski. Logical matrices, primitive satisfaction and finitely based logics. Studia Logica,
42:89–104, 1983.
J. Czelakowski. Filter distributive logics. Studia Logica, 43:353–377, 1984.
J. Czelakowski. Algebraic aspects of deduction theorems. Studia Logica, 44:369–387, 1985.
J. Czelakowski. Protoalgebraic logics, volume 10 of Trends in Logic - Studia Logica Library.
Kluwer Academic Publishers, Dordrecht, 2001.
J. Czelakowski. The Suszko operator. Part I. Studia Logica (Special issue on Abstract Algebraic
Logic, Part II), 74:181–231, 2003.
J. Czelakowski and W. Dziobiak. A single quasi-identity for a quasivariety with the Fraser-Horn
property. Algebra Universalis, 29:10–15, 1992.
J. Czelakowski and R. Jansana. Weakly algebraizable logics. The Journal of Symbolic Logic, 65:
641–668, 2000.
J. Czelakowski and G. Malinowski. Key notions of Tarski’s methodology of deductive systems.
Studia Logica, 44:321–351, 1985.
J. Czelakowski and D. Pigozzi. Amalgamation and interpolation in abstract algebraic logic. In
X. Caicedo and C. H. Montenegro, editors, Models, algebras and proofs, volume 203 of Lecture
Notes in Pure and Applied Mathematics Series, pages 187–265. Marcel Dekker, New York and
Basel, 1999.
J. Czelakowski and D. Pigozzi. Fregean logics. Annals of Pure and Applied Logic, 127:17–76,
2004.
J. M. Dunn and G. M. Hardegree. Algebraic methods in philosophical logic, volume 41 of Oxford
Logic Guides. Oxford Science Publications, Oxford, 2001.
M. Esteban. Duality theory and abstract algebraic logic. Ph. D. Dissertation, University of
Barcelona, 2013.
J. M. Font. On the Leibniz congruences. In C. Rauszer, editor, Algebraic Methods in Logic and in
Computer Science, volume 28 of Banach Center Publications, pages 17–36. Polish Academy
of Sciences, Warszawa, 1993.
J. M. Font. Belnap’s four-valued logic and De Morgan lattices. Logic Journal of the IGPL, 5:
413–440, 1997.
J. M. Font. Abstract Algebraic Logic - An Introductory Textbook, volume 60 of Studies in Logic.
College Publications, London, 2016.
J. M. Font, F. Guzmán, and V. Verdú. Characterization of the reduced matrices for the conjunction-
disjunction fragment of classical logic. Bulletin of the Section of Logic, 20:124–128, 1991.
J. M. Font and R. Jansana. A general algebraic semantics for sentential logics, volume 7 of Lecture
Notes in Logic. Springer-Verlag, 1996. Out of print. Second revised edition published in
2009 by the Association for Symbolic Logic, freely available through Project Euclid at http:
//projecteuclid.org/euclid.lnl/1235416965.
J. M. Font, R. Jansana, and D. Pigozzi. Fully adequate Gentzen systems and the deduction theorem.
Reports on Mathematical Logic, 35:115–165, 2001.
Abstract Algebraic Logic 71

J. M. Font, R. Jansana, and D. Pigozzi. A survey of abstract algebraic logic. Studia Logica (Special
issue on Abstract Algebraic Logic, Part II), 74:13–97, 2003. With an update in vol. 91 (2009),
125–130.
J. M. Font and T. Moraschini. Logics of varieties, logics of semilattices, and conjunction. Logic
Journal of the IGPL, 22:818–843, 2014.
J. M. Font and T. Moraschini. M-sets and the representation problem. Studia Logica, 103:21–51,
2015.
J. M. Font and G. Rodrı́guez. Algebraic study of two deductive systems of relevance logic. Notre
Dame Journal of Formal Logic, 35:369–397, 1994.
J. M. Font and V. Verdú. Algebraic logic for classical conjunction and disjunction. Studia Logica
(Special issue on Algebraic Logic), 50:391–419, 1991.
N. Galatos and J. Gil-Férez. Modules over quantaloids: Applications to the isomorphism problem
in algebraic logic and π -institutions. Journal of Pure and Applied Algebra, 2016. To appear.
N. Galatos, P. Jipsen, T. Kowalski, and H. Ono. Residuated lattices: an algebraic glimpse at
substructural logics, volume 151 of Studies in Logic and the Foundations of Mathematics.
Elsevier, Amsterdam, 2007.
N. Galatos and J. Raftery. A category equivalence for odd Sugihara monoids and its applications.
Journal of Pure and Applied Algebra, 216:2177–2192, 2012.
N. Galatos and C. Tsinakis. Equivalence of consequence relations: an order-theoretic and categor-
ical perspective. The Journal of Symbolic Logic, 74:780–810, 2009.
M. Gehrke, R. Jansana, and A. Palmigiano. Canonical extensions for congruential logics with the
deduction theorem. Annals of Pure and Applied Logic, 161:1502–1519, 2010.
P. Hájek. Metamathematics of fuzzy logic, volume 4 of Trends in Logic - Studia Logica Library.
Kluwer Academic Publishers, Dordrecht, 1998.
P. R. Halmos. Algebraic logic. Chelsea Publ. Co., New York, 1962.
L. Henkin. An algebraic characterization of quantifiers. Fundamenta Mathematicae, 37:63–74,
1950.
L. Henkin, J. D. Monk, and A. Tarski. Cylindric algebras, part II. North-Holland, Amsterdam,
1985.
B. Herrmann. Equivalential logics and definability of truth. Ph. D. Dissertation, Freie Universität
Berlin, 1993.
B. Herrmann. Equivalential and algebraizable logics. Studia Logica, 57:419–436, 1996.
W. Hodges. Model theory, volume 42 of Encyclopedia of Mathematics and its Applications. Cam-
bridge University Press, Cambridge, 1993.
P. Idziak, K. Słomczyńska, and A. Wroński. Fregean varieties. International Journal of Algebra
and Computation, 19:595–645, 2009.
R. Jansana. Full models for positive modal logic. Mathematical Logic Quarterly, 48:427–445,
2002.
R. Jansana. Propositional consequence relations and algebraic logic. In E. N. Zalta, editor, The
Stanford Encyclopedia of Philosophy. Spring 2011 edition, 2011. Archived as: http://plato.
stanford.edu/archives/spr2011/entries/consequence-algebraic/.
P. Köhler and D. Pigozzi. Varieties with equationally definable principal congruences. Algebra
Universalis, 11:213–219, 1980.
M. Kracht. Modal consequence relations. Chapter 8 of P. Blackburn, J. van Benthem, and F. Wolter,
editors, Handbook of Modal Logic, volume 3 of Studies in Logic and Practical Reasoning,
pages 491–545. Elsevier, Amsterdam, 2007.
J. Łoś. O matrycach logicznych, volume 19 of Prace Wrocławskiego Towarzystwa Naukowege,
Ser. B. University of Wrocław, 1949.
T. Moraschini. Investigations into the role of translations in abstract algebraic logic. Ph. D.
Dissertation, University of Barcelona, June 2016.
A. Nurakunov and M. Stronkowski. Relation formulas for protoalgebraic equality free quasiva-
rieties: Pałasińska’s theorem revisited. Studia Logica (Special issue on Abstract Algebraic
Logic), 101:827–847, 2013.
72 Josep Maria Font

D. Pigozzi. Finite basis theorem for relatively congruence-distributive quasivarieties. Transactions


of the American Mathematical Society, 310:499–533, 1988.
D. Pigozzi. Fregean algebraic logic. In H. Andréka, J. D. Monk, and I. Németi, editors, Alge-
braic Logic, volume 54 of Colloq. Math. Soc. János Bolyai, pages 473–502. North-Holland,
Amsterdam, 1991.
D. Pigozzi. Abstract algebraic logic: past, present and future. A personal view. In J. M. Font,
R. Jansana, and D. Pigozzi, editors, Workshop on Abstract Algebraic Logic, volume 10 of
Quaderns, pages 122–138. Centre de Recerca Matemàtica, Bellaterra, January 1998.
D. Pigozzi. Abstract Algebraic Logic. In M. Hazewinkel, editor, Encyclopædia of Mathematics,
Supplement III, pages 2–13. Kluwer Academic Publishers, Dordrecht, 2001.
W. Pogorzelski. On Hilbert’s operation on logical rules I. Reports on Mathematical Logic, 12:
35–50, 1981.
W. Pogorzelski and P. Wojtylak. Completeness Theory for Propositional Logics. Studies in Uni-
versal Logic. Birkhäuser Verlag, Basel, 2008.
T. Prucnal and A. Wroński. An algebraic characterization of the notion of structural completeness.
Bulletin of the Section of Logic, 3:30–33, 1974.
A. Pynko. Definitional equivalence and algebraizability of generalized logical systems. Annals of
Pure and Applied Logic, 98:1–68, 1999.
J. Raftery. Correspondences between Gentzen and Hilbert systems. The Journal of Symbolic Logic,
71:903–957, 2006a.
J. Raftery. The equational definability of truth predicates. Reports on Mathematical Logic (Special
issue in memory of Willem Blok), 41:95–149, 2006b.
J. Raftery. A non-finitary sentential logic that is elementarily algebraizable. Journal of Logic and
Computation, 20:969–975, 2010.
J. Raftery. Contextual deduction theorems. Studia Logica (Special issue in honor of Ryszard
Wójcicki on the occasion of his 80th birthday), 99:279–319, 2011a.
J. Raftery. A perspective on the algebra of logic. Quaestiones Mathematicae, 34:275–325, 2011b.
J. Raftery. Order algebraizable logics. Annals of Pure and Applied Logic, 164:251–283, 2013a.
J. Raftery. Inconsistency lemmas in algebraic logic. Mathematical Logic Quarterly, 59:393–406,
2013b.
H. Rasiowa. An algebraic approach to non-classical logics, volume 78 of Studies in Logic and the
Foundations of Mathematics. North-Holland, Amsterdam, 1974.
H. Rasiowa and R. Sikorski. Algebraic treatment of the notion of satisfability. Fundamenta Math-
ematicae, 40:62–95, 1953.
W. Rautenberg, M. Zakharyaschev, and F. Wolter. Willem Blok and modal logic. Studia Logica,
83:15–30, 2006.
J. Rebagliato and V. Verdú. On the algebraization of some Gentzen systems. Fundamenta Infor-
maticae (Special issue on Algebraic Logic and its Applications), 18:319–338, 1993.
C. Russo. An order-theoretic analysis of interpretations among propositional deductive systems.
Annals of Pure and Applied Logic, 164:112–130, 2013.
D. J. Shoesmith and T. J. Smiley. Multiple-conclusion logic. Cambridge University Press, Cam-
bridge, 1978.
S. Surma. On the origin and subsequent applications of the concept of the Lindenbaum algebra.
In Logic, methodology and philosophy of science VI, Hannover 1979, pages 719–734. North-
Holland, Amsterdam, 1982.
R. Suszko. Abolition of the Fregean axiom. In R. Parikh, editor, Logic Colloquium 1972–73,
volume 453 of Lecture Notes in Mathematics, pages 169–239. Springer-Verlag, Berlin, 1975.
Y. Suzuki, F. Wolter, and M. Zacharyaschev. Speaking about transitive frames in propositional
languages. Journal of Logic, Language and Information, 7:317–339, 1998.
G. Voutsadakis. Categorical abstract algebraic logic: equivalent institutions. Studia Logica (Special
issue on Abstract Algebraic Logic, Part II), 74:275–311, 2003.
R. Wójcicki. Logical matrices strongly adequate for structural sentential calculi. Bulletin de
l’Académie Polonaise des Sciences, Classe III, XVII:333–335, 1969.
Abstract Algebraic Logic 73

R. Wójcicki. Matrix approach in the methodology of sentential calculi. Studia Logica, 32:7–37,
1973.
R. Wójcicki. Referential matrix semantics for propositional calculi. Bulletin of the Section of
Logic, 8:170–176, 1979.
R. Wójcicki. Theory of logical calculi. Basic theory of consequence operations, volume 199 of
Synthèse Library. Reidel, Dordrecht, 1988.
A. Wroński. BCK algebras do not form a variety. Mathematica Japonica, 28:211–213, 1983.
J. Zygmunt. A note on direct products and ultraproducts of logical matrices. Studia Logica, 33:
349–358, 1974.

Vous aimerez peut-être aussi