Sie sind auf Seite 1von 294

Radim Bělohlávek, Vilém Vychodil

Fuzzy Equational Logic


Studies in Fuzziness and Soft Computing, Volume 186
Editor-in-chief
Prof. Janusz Kacprzyk
Systems Research Institute
Polish Academy of Sciences
ul. Newelska 6
01-447 Warsaw
Poland
E-mail: kacprzyk@ibspan.waw.pl

Further volumes of this series Vol. 178. Claude Ghaoui, Mitu Jain,
can be found on our homepage: Vivek Bannore, Lakhmi C. Jain (Eds.)
Knowledge-Based Virtual Education, 2005
springeronline.com ISBN 3-540-25045-X
Vol. 179. Mircea Negoita,
Vol. 170. Martin Pelikan
Bernd Reusch (Eds.)
Hierarchical Bayesian Optimization
Real World Applications of Computational
Algorithm, 2005
Intelligence, 2005
ISBN 3-540-23774-7
ISBN 3-540-25006-9
Vol. 171. James J. Buckley
Vol. 180. Wesley Chu,
Simulating Fuzzy Systems, 2005
Tsau Young Lin (Eds.)
ISBN 3-540-24116-7
Foundations and Advances in Data Mining,
Vol. 172. Patricia Melin, Oscar Castillo 2005
Hybrid Intelligent Systems for Pattern ISBN 3-540-25057-3
Recognition Using Soft Computing, 2005
Vol. 181. Nadia Nedjah,
ISBN 3-540-24121-3
Luiza de Macedo Mourelle
Vol. 173. Bogdan Gabrys, Kauko Leiviskä, Fuzzy Systems Engineering, 2005
Jens Strackeljan (Eds.) ISBN 3-540-25322-X
Do Smart Adaptive Systems Exist?, 2005
Vol. 182. John N. Mordeson,
ISBN 3-540-24077-2
Kiran R. Bhutani, Azriel Rosenfeld
Vol. 174. Mircea Negoita, Daniel Neagu, Fuzzy Group Theory, 2005
Vasile Palade ISBN 3-540-25072-7
Computational Intelligence: Engineering of
Vol. 183. Larry Bull, Tim Kovacs (Eds.)
Hybrid Systems, 2005
Foundations of Learning Classifier Systems,
ISBN 3-540-23219-2
2005
Vol. 175. Anna Maria Gil-Lafuente ISBN 3-540-25073-5
Fuzzy Logic in Financial Analysis, 2005
Vol. 184. Barry G. Silverman, Ashlesha Jain,
ISBN 3-540-23213-3
Ajita Ichalkaranje, Lakhmi C. Jain (Eds.)
Vol. 176. Udo Seiffert, Lakhmi C. Jain, Intelligent Paradigms for Healthcare
Patric Schweizer (Eds.) Enterprises, 2005
Bioinformatics Using Computational ISBN 3-540-22903-5
Intelligence Paradigms, 2005
Vol. 185. Dr. Spiros Sirmakessis (Ed.)
ISBN 3-540-22901-9
Knowledge Mining, 2005
Vol. 177. Lipo Wang (Ed.) ISBN 3-540-25070-0
Support Vector Machines: Theory and
Vol. 186. Radim Bělohlávek, Vilém
Applications, 2005
Vychodil
ISBN 3-540-24388-7
Fuzzy Equational Logic, 2005
ISBN 3-540-26254-7
Radim Bělohlávek
Vilém Vychodil

Fuzzy Equational Logic

ABC
Radim Bělohlávek Vilém Vychodil
Department of Computer Science Department of Computer Science
Palacky University Palacky University
Tomkova 40 Tomkova 40
CZ-779 00 Olomouc CZ-779 00 Olomouc
Czech Republic Czech Republic
E-mail: radim.belohlavek@upol.cz E-mail: vilem.vychodil@upol.cz

Library of Congress Control Number: 2005927316

ISSN print edition: 1434-9922


ISSN electronic edition: 1860-0808
ISBN-10 3-540-26254-7 Springer Berlin Heidelberg New York
ISBN-13 978-3-540-26254-7 Springer Berlin Heidelberg New York

This work is subject to copyright. All rights are reserved, whether the whole or part of the material is
concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting,
reproduction on microfilm or in any other way, and storage in data banks. Duplication of this publication
or parts thereof is permitted only under the provisions of the German Copyright Law of September 9,
1965, in its current version, and permission for use must always be obtained from Springer. Violations are
liable for prosecution under the German Copyright Law.
Springer is a part of Springer Science+Business Media
springeronline.com
c Springer-Verlag Berlin Heidelberg 2005
Printed in The Netherlands
The use of general descriptive names, registered names, trademarks, etc. in this publication does not imply,
even in the absence of a specific statement, that such names are exempt from the relevant protective laws
and regulations and therefore free for general use.
Typesetting: by the authors and TechBooks using a Springer LATEX macro package
Printed on acid-free paper SPIN: 11376422 55/TechBooks 543210
To Jana

Radim

To my parents

Vilém
Preface

The present book deals with algebras, congruences, morphisms, reasoning


about identities, classes of algebras and their axiomatizability, etc., from the
point of view of fuzzy logic. We therefore deal with topics traditionally stud-
ied in universal algebra. Our approach is the following. In classical universal
algebra, one works in bivalent setting. That is, one works with a two-element
Boolean algebra 2 as the underlying structure of truth degrees. As an exam-
ple, any two elements of an algebra either are congruent or not with respect
to a given congruence relation. In our approach, we allow for more general
structures of truth degrees than 2. We consider a structure of truth degrees as
a parameter for which we can take any suitable structure L of truth degrees.
For us, being suitable means that L is a complete residuated lattice. That is, L
consists of a set L of truth degrees equipped with (suitable truth functions of)
logical connectives. In this sense, we parameterize concepts and results known
from classical universal algebra. If we put L = 2, we get just the concepts and
results of classical universal algebra in which case the set L of truth degrees
contains just two elements, namely L = {0, 1}. This is a boundary choice and
we have other possibilities, e.g. L = [0, 1] (real unit interval). Returning to
our example, a practical consequence is that, in general, congruence is a fuzzy
relation. Two elements of an algebra may thus be congruent in degree, say,
0.8. This seems to be natural. Namely, the idea of congruence is to group to-
gether elements which are similar from a certain point of view (example from
crisp setting: congruent numbers in the sense of having the same remainder
when dividing by 3). Congruence in degrees thus corresponds to the fact that
we may consider elements similar to degrees not necessarily equal to 0 or 1
(example: two sets might be similar in degree proportional to the extent of
their overlap). Thus, allowing more general structures than the two-element
Boolean algebra is not just a mathematical exercise.
We thus deal with a kind of development of universal algebra and related
topics from the point of view of fuzzy logic. Compared to other approaches,
our approach is quite modest. Namely, equality of elements of algebras is the
only component which becomes fuzzy. Although we could, we deliberately use
VIII Preface

neither “fuzzy subalgebras” nor “fuzzy functions”. Our “picture of the world”
is the following. Classical algebras are structures of predicate logic with a
language containing function symbols and a symbol of equality as a single
relation symbol. We are thus interested in structures of predicate fuzzy logic
with a language containing function symbols and just one relation symbol – a
symbol of equality. The function symbols are interpreted by ordinary functions
and the symbol of equality is interpreted by a fuzzy equality relation. We
require all functions to be compatible with the fuzzy equality relation. As a
consequence, if the fuzzy equality is understood as a similarity relation, our
structures can be seen as sets equipped with “functions mapping similar to
similar”. We call them algebras with fuzzy equality and in our setting, they
play the same role as universal algebras in classical setting.
From the point of view of fuzzy logic, our framework fits into that of
Pavelka’s abstract fuzzy logic. We fix a particular structure L of truth de-
grees and develop our fuzzy logic. This gives a whole family of fuzzy logics
parameterized by L. Picking a particular L, we deal with formulas, L-fuzzy
sets of formulas, degrees of truth from L, and degrees of provability from L.
Our formulas can be seen as generalized (since we sometimes use infinite con-
junctions) formulas of predicate fuzzy logic with truth constants in language.
For convenience, however, we use a particular way of writing our formulas. As
a result, our formulas can be translated into formulas of predicate fuzzy logic
with truth constants but are shorter. We prefer shorter formulas although not
directly fitting into the framework of predicate fuzzy logic but this is a matter
of taste and the reader may choose the other way.
We develop a small fragment of fuzzy logic – we deal with a restricted
language and restricted formulas. From this point of view, we contribute to
already existing examples of restricted systems of fuzzy logic. On the one hand,
they have less expressive power. On the other hand, they are more powerful
than the more expressive systems since they obey desirable properties for a
larger class of structures of truth degrees. For instance, compared to first-order
fuzzy logic with evaluated syntax which is not syntactico-semantically com-
plete for structures of truth degrees other than standard L  ukasiewicz algebra
(and its isomorphic copies), we have completeness also for other structures of
truth degrees. Developing special fragments of fuzzy logic for various applica-
tion domains is interesting and important: with a small fragment tailored for
a particular application domain, one can hope to be able to go farther than
with a full-fledged fuzzy logic. Although we do not think that our fragment
will enjoy practical applications, we claim that it shows the effect of being
able to go farther within a smaller system.
The book is organized as follows. Chapter 1 contains preliminary no-
tions and results. Chapter 2 deals with algebras with fuzzy equalities. Chap-
ters 3 and 4 develop equational logic and Horn logic in fuzzy setting, their
proof systems, completeness results, and results concerning definability by
means of particular formulas which involve identities. At the close of each
chapter is a section “Bibliographical Remarks” which contains references to
Preface IX

bibliographical items and also comments on alternative approaches. The book


contains a list of references, a table of notation, and an index of terms.
We gratefully acknowledge support by grant no. B1137301 of the Grant
Agency of the Czech Academy of Sciences and institutional support, research
plan MSM 6198959214.

April 2005 Radim Bělohlávek


Vilém Vychodil
Contents

1 Introduction to Fuzzy Sets and Fuzzy Logic . . . . . . . . . . . . . . . 1


1.1 Sets and Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Structures of Truth Degrees . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.3 Fuzzy Sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
1.4 Pavelka-Style Fuzzy Logic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
1.5 Bibliographical Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

2 Algebras with Fuzzy Equalities . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59


2.1 Definition and Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
2.2 Subalgebras, Congruences, and Morphisms . . . . . . . . . . . . . . . . . . 66
2.3 Direct and Subdirect Products . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
2.4 Terms, Term L-Algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
2.5 Direct Unions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
2.6 Direct Limits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
2.7 Reduced Products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120
2.8 Class Operators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 127
2.9 Related Approaches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
2.10 Bibliographical Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136

3 Fuzzy Equational Logic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139


3.1 Syntax and Semantics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 140
3.2 Completeness of Fuzzy Equational Logic . . . . . . . . . . . . . . . . . . . . 146
3.3 Varieties, Free L-Algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
3.4 Equational Classes of L-Algebras . . . . . . . . . . . . . . . . . . . . . . . . . . 158
3.5 Properties of Varieties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
3.6 Bibliographical Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170

4 Fuzzy Horn Logic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171


4.1 Syntax and Semantics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
4.2 Semantic Entailment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
4.3 Completeness of Fuzzy Horn Logic . . . . . . . . . . . . . . . . . . . . . . . . . 204
XII Contents

4.4 Fuzzy Equational Logic Revisited . . . . . . . . . . . . . . . . . . . . . . . . . . 221


4.5 Implicationally Defined Classes of L-Algebras . . . . . . . . . . . . . . . 222
4.6 Sur-Reflections and Sur-Reflective Classes . . . . . . . . . . . . . . . . . . 235
4.7 Semivarieties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
4.8 Quasivarieties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
4.9 Bibliographical Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267

Table of Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 273

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
1
Introduction to Fuzzy Sets and Fuzzy Logic

This chapter surveys preliminary notions and results that we use in subsequent
chapters. We pay more attention to concepts which might not be well known.
Well-known notions and results are just recalled.

1.1 Sets and Structures


Sets, Relations, Mappings

We assume that the reader is familiar with the notion of a set as used in the
intuitive set theory, i.e. as a collection of objects. Each object either belongs
or does not belong to a given set. Objects which belong to a given set are also
called its elements. We write a = b to denote that a and b are identical. The
fact that an object a belongs to a set A is denoted by a ∈ A, the opposite
by a ∈ A. Given a property ϕ, we denote by {a | ϕ(a )} the set of all objects
having ϕ. The empty set (a set containing no elements) is denoted by ∅. A set
A is a subset of a set B (denoted by A ⊆ B) if each element of A belongs to B.
We write A ⊂ B if A ⊆ B but A = B. Given sets A and B, the intersection of
A and B (denoted by A∩B) is the set containing the elements which belong to
both A and B, the union of A and B (denoted by A ∪ B) is the set containing
the elements which belong to A or to B (or both), the difference between A
and B (denoted by A − B) is the set containing the elements belonging to A
but not to B. In some situations, only elements of some given set U are taken
into account. U is called a universe of discourse (universal set, universe). If
U is a universe and A ⊆ U , then the complement of A (w.r.t. U , denoted by
A) is the set U − A.
An ordered n-tuple of objects u1 , . . . , un is denoted by u1 , . . . , un . We
have u1 , . . . , un = v1 , . . . , vm iff n = m and u1 = v1 , . . . , un = vn . A direct
product (or Cartesian product) of sets U1 , . . . , Un is the set U1 × · · · × Un =
{u1 , . . . , un | u1 ∈ U1 , . . . , un ∈ Un }, i.e. the set of all n-tuples with ui from
Ui . If U1 = · · · = Un = U , we denote the direct product by U n .

R. Bělohlávek and V. Vychodil: Fuzzy Equational Logic, StudFuzz 186, 1–58 (2005)
www.springerlink.com 
c Springer-Verlag Berlin Heidelberg 2005
2 1 Introduction to Fuzzy Sets and Fuzzy Logic

A relation between sets U1 , . . . , Un is any subset r of the direct product


of these sets, i.e. r ⊆ U1 × · · · × Un . The number n is called the arity of r
(r is said to be n-ary). For n = 1, 2, 3, . . . we use the terms unary, binary,
ternary, . . . If U1 = · · · = Un = U , r is called a relation in U (or, on U ). The
inverse relation of a binary relation r ⊆ U × V is a relation r−1 ⊆ V × U
defined by r−1 = {v , u | u , v ∈ r}. For r ⊆ U × V and A ⊆ U we put
r(A) = {v | u , v ∈ r for some u ∈ A}. r({u }) is also denoted by [u ]r .
A mapping (or function) of a set U into a set V is a binary relation
f between U and V such that for each u ∈ U there is v ∈ V such that
u , v ∈ f , and if u , v1 ∈ f and u , v2 ∈ f then v1 = v2 . A mapping f
may be thought of as representing an assignment of elements of V to elements
of U . We write f : U → V if f is a mapping of U to V . If u , v ∈ f for a
mapping f , we write f (u ) = v . The set of all mappings of U to V is denoted
by V U . A mapping f : V n → V is also called an n-ary operation in V . A
mapping f is called injective (or injection) if u1 = u2 implies f (u1 ) = f (u2 );
surjective (or surjection) if for each v ∈ V there is u ∈ U such that f (u ) = v ;
bijective (or bijection) if it is both injective and surjective. If f : U → V ,
g : V → W , then a composition of f and g is a mapping f ◦ g of U to W
defined by (f ◦ g)(u ) = g(f (u )).
Putting 2 = {0, 1} (there is no danger of confusion with the integer 2),
2U denotes the set of all mappings of U to {0, 1}. Mappings χ : X → {0, 1}
are called characteristic functions of subsets of U . Namely, there is a natural
bijection between 2U and the set of all subsets of U : each χ ∈ 2U induces a
subset Aχ = {u ∈ U | χ(u ) = 1} of U ; each subset A of U induces a mapping
χA : U → {0, 1} by χA (u ) = 1 if u ∈ A and χA (u ) = 0 if u ∈ A; χ is called
a characteristic function of the corresponding subset Aχ ; we have AχA = A
and χAχ = χ. For convenience, we usually do not distinguish between subsets
of U and their characteristic functions and denote the set of all subsets of U
also by 2U .
If I is a set and for each i ∈ I we have a set Ui , we say that {Ui | i ∈ I}
is a system (or family) of sets indexed by I. In a natural  way, intersection
and union is extended to a system of sets by putting i∈I Ui = {u | u ∈
Uj for all j ∈ I}, i∈I Ui = Q {u | u ∈ Uj for some j ∈ I}. A direct product of
a system {Ui | i ∈ I} is a set i∈I UQi = {f : I → i∈I Ui | f (j) ∈ Uj for all j ∈
I}. In particular, if I = ∅, then i∈I Ui = {∅} (one-element set containing 
the empty set) since Q in this case, ∅ is the only mapping f : I → i∈I Ui .
For I = {1, . . . , n}, i∈I Ui coincides in an obvious way with the notion of a
Q The j-th projection (j ∈ I)
direct product of U1 ,Q. . . , Un , as introduced above.
of a Q direct product i∈I Ui is a mapping πj : i∈I Ui → Uj defined for any
a ∈ i∈I Ui by πj (a) = a(j) (i.e. πj selects the j-th component).
N, Z, Q, and R denote the set of all positive integers, the set of all integers,
the set of all rationals, and the set of all reals, respectively. We put N0 =
N ∪ {0} (non-negative integers).
A set A is finite iff A is empty or for some n ∈ N there is a bijection
f : A → {k ∈ N | 1 ≤ k ≤ n}. In the latter case, elements of A can be assigned
1.1 Sets and Structures 3

natural numbers 1, . . . , n in a one-one way and so, n is called the number of


elements of A. If a set A is not finite, it is called infinite. A set A is called
denumerable (or countably infinite) if there is a bijection f : A → N. Cardinal
and ordinal numbers are denoted by Greek letters α, . . . , κ, etc. Particularly,
ω denotes the cardinality of N, i.e. the least infinite cardinal number.
A set F ⊆ 2U is called a filter over U if X ∩Y ∈ F for every X, Y ∈ F , and
if X ∈ F and X ⊆ Y imply Y ∈ F . A filter F is called proper if there is X ⊆ U
such that X ∈ F . Filter F is called improper if ∅ ∈ F (in this case, F = 2U ).
If {u } ∈ F for some u ∈ U , then F is said to be a trivial filter. If for every
X ⊆ F we either have X ∈ F or U −X ∈ F , then F is said to be an ultrafilter.
For any infinite universe set U , a filter F = {X ⊆ U | U − X is finite} is called
a Fréchet filter over U .

Equivalences and Partial Orders

A binary relation r in a set U is called reflexive if for each u ∈ U we have


u , u ∈ r; symmetric if for each u , v ∈ U , u , v ∈ r implies v , u ∈ r;
antisymmetric if for each u , v ∈ U , u , v ∈ r and v , u ∈ r imply u = v ;
transitive if for each u , v , w ∈ U , u , v ∈ r and v , w ∈ r imply u , w ∈ r.
A binary relation in a set is called an equivalence if it is reflexive, sym-
metric, and transitive. For an equivalence E in a set U and u ∈ U , the set
[u ]E = {v | u , v ∈ E} is called a class of E (given by u ). For each set U ,
there is a bijective correspondence between equivalences in U and so-called
partitions of U . A partition of a set U is a system Π of non-empty subsets of
U (called classes of Π) such that for each u ∈ U there is A ∈ Π with u ∈ A
(classes cover U ), and for each A, B ∈ Π with A = B we have A ∩ B = ∅
(classes do not overlap). The correspondence: For each equivalence E in U ,
the set ΠE = {[u ]E | u ∈ U } is a partition of U . For each partition Π of U ,
the relation EΠ = {u , v | u , v ∈ A for some A ∈ Π} is an equivalence in
U . Moreover, we have E = EΠE and Π = ΠEΠ . The set ΠE of all classes of
E is also called a factor set of U by E and is denoted by U/E.
A binary relation in a set is called a partial order if it is reflexive, anti-
symmetric, and transitive. If ≤ is a partial order, we write u ≤ v instead of
u , v ∈ ≤, and write u < v if u ≤ v and u = v . If ≤ is a partial order
in U , the pair U = U, ≤ is called a partially ordered set. A partially or-
dered set U = U, ≤ with finite U may be visualized using so-called Hasse
diagram: Each element u ∈ U is depicted as a node labeled by “u ” in such
a way that if u is covered by v , i.e. u < v and there is no w ∈ U such
that u < w and w < v , then we connect the nodes of u and v , and put
the node of u below the node of v . For a partially odered set U, ≤ , a func-
tion f (. . . , x, . . . ) : U n → U is called non-decreasing (isotone) in x if we
have f (. . . , u , . . . ) ≤ f (. . . , v , . . . ) whenever u ≤ v ; f is called non-increasing
(antitone) in x if we have f (. . . , u , . . . ) ≥ f (. . . , v , . . . ) whenever u ≤ v .
Let U, ≤ be a partially ordered set and A ⊆ U . An element u ∈ A is
called a minimal element of A if for each v ∈ A such that v ≤ u we have
4 1 Introduction to Fuzzy Sets and Fuzzy Logic

u = v ; a maximal element of A if for each v ∈ A such that u ≤ v we have


u = v ; a least element of A if for each v ∈ A we have u ≤ v ; a greatest
element of A if for each v ∈ A we have v ≤ u . If there exists a least (greatest)
element of U , it is denoted by 0 (1). A lower cone of A is a set L(A) denoted
by L(A) = {u ∈ U | u ≤ v for all v ∈ A}, an upper cone of A is a set U(A)
denoted by U(A) = {u ∈ U | v ≤ u for all v ∈ A}. An element u ∈ L(A)
is called a lower bound of A, an element u ∈ U(A) is called an upper bound
of A. If L(A) has agreatest element u , then u is called an infimum of A in
U, ≤ , denoted by A or inf A. Dually, if U(A) has  a least element u , then
u is called a supremum of A inU, ≤ , denoted
 by A or sup A. Note that if
the least
 element0 exists then U = 0 = ∅ by definition (dually, if 1 exists
then ∅ = 1 = U ).
A partially ordered set U, ≤ is said to be lattice ordered (or a lattice)
if infimum and supremum exist for any two-element (and hence for any non-
empty finite) subset of U . U, ≤ is said to be completely lattice ordered (or
a complete lattice) if infimum and supremum exist for any subset of U . A
partially ordered set U, ≤ is said to be linearly ordered (or a chain) if for
every u , v ∈ U we have u ≤ v or v ≤ u , i.e. every two elements are com-
parable. In this case, ≤ is called a linear order. Each chain is a lattice with
inf{u , v } = min{u , v } and sup{u , v } = max{u , v } (where min{u , v } and
max{u , v } denote the smaller and the bigger one of u and v , respectively).
 an element u ∈ U
If U, ≤ is a lattice ordered set,  is called compact if for
each A ⊆ U we have that  if A exists and u ≤ A then there is a finite
B ⊆ A such that u ≤ B still holds. A lattice is called compactly generated
if each of its elements is a supremum of compact elements. An algebraic lat-
tice is a lattice which is both complete and compactly generated. A lattice
is called Noetherian if it is complete and every its non-empty subset has at
least one maximal element. Therefore, in a Noetherian chain, each non-empty
subset has a greatest element. If U, ≤ is a Noetherian lattice then for any
A ⊆ U there is some finite F ⊆ A such  that sup A = sup F . An element  u of
a lattice ordered set U, ≤ is called -irreducible
 if whenever u = A for
 a
non-empty A ⊆ U , then u ∈ A; u is called -irreducible if whenever u = A
for a non-empty A ⊆ U , then u ∈ A.
We assume Zorn lemma which says that if each chain A of a partially
ordered set U (i.e. a subset A ⊆ U with pairwise comparable elements) has
an upper bound (i.e. U(A) = ∅) then for each u ∈ U there is a maximal
element v of U such that u ≤ v .

Example 1.1. (1) The genuine order ≤ on the set R of all real numbers is a
partial order. Moreover, R, ≤ is a chain and thus a lattice (which is not
complete).  
(2) For any set U , 2U , ⊆ is a complete lattice (which is not
 linear iff
U has at least two elements)
 with inf{Ai ⊆ U | i ∈ I} = i∈I Ai and
sup{Ai ⊆ U | i ∈ I} = i∈I Ai .
1.1 Sets and Structures 5

1
g

c e f
b

a d
0
Fig. 1.1. Hasse diagram of a lattice from Example 1.1 (4)

(3) More generally, if S ⊆ 2U is a non-empty


 system of subsets of U that
is closed w.r.t. arbitrary intersections (i.e. i∈I Ai ∈ S whenever Ai ∈ S)
then S, ⊆ is a complete lattice where  infima coincide
 with intersections and
suprema are given by supi∈I Ai = {A ∈ S | i∈I Ai ⊆ A}. A system S
of subsets of U that is closed w.r.t. arbitrary intersections is called a closure
system
 in U . For a closure system S and a subset A ⊆ U , the set [A]S =
{B ∈ S | A ⊆ B} is called the closure of A; it is the least (w.r.t. ⊆) subset
of U that belongs to S and contains A.
(4) Figure 1.1 shows a Hasse diagram of a nine-element lattice U, ≤ with
U = {0, a, . . . , g, 1}. We can see from the diagram that, for instance, a ≤ c,
a ≤ g, a and d are not comparable, sup{a, d} = e, inf{b, c} = inf{b, c, e, g} =
a, etc.

Let U = U, ≤U be a complete lattice. A mapping c : U → U is called a


closure operator in U if (i) u ≤U c(u ); (ii) u1 ≤U u2 implies c(u1 ) ≤U c(u2 );
(iii) c(u ) = c(c(u )) (for any u , ui ∈ U ). If (i) is replaced by c(u ) ≤U u , c is
called an interior operator in U. A subset S of U is called a closure system
in U if S is closed under arbitrary infima. For a closure operator c in U, the
set Sc = {u ∈ U | c(u ) = u } of all fixed points of c is a closure system in
U. Conversely, for a closure system S in U, a mapping cS : U → U defined
by c(u ) = {v ∈ S | u ≤ v } is a closure operator in U. Moreover, we have
cSc = c and ScS = S.  
Let U = U, ≤U and V = V, ≤V be complete lattices. A pair ↑ , ↓ of
mappings ↑ : U → V and ↓ : V → U is said to form a Galois connection
between U and V if (i) u1 ≤U u2 implies u2↑ ≤V u1↑ ; (ii) v1 ≤V v2 implies
v2↓ ≤U v1↓ ; (iii) u ≤U u ↑↓ ↓↑
 ; (iv) v ≤V v (for u , ui↑↓∈ U↑◦↓ , v , vi ∈ V ). For
↑ ↓
a Galois connection , , the composed mappings = and ↓↑ = ↓◦↑
are closure operators in U and V, respectively. Furthermore, u ↑ = u ↑↓↑ and
v ↓ = v ↓↑↓ .    
For U, ≤U = 2X , ⊆ and V, ≤V = 2Y , ⊆ , one obtains from the above
notions the notions of a closure operator in a set X, a closure system in X, cf.
also Example 1.1 (3), and that of a Galois connection between sets X and Y .
6 1 Introduction to Fuzzy Sets and Fuzzy Logic

Algebras and Structures

A structure is a set equipped with relations and functions with prescribed


arities. Arities of relations and functions are given by a so-called type. For-
mally, a type is a triplet R, F, σ consisting of a set R (of symbols of rela-
tions), a set F (of symbols of functions) with R ∩ F = ∅, and a mapping
σ of R ∪ F into a set N0 of non-negative integers. For s ∈ R ∪ F , σ(s)
is called
 the arity of  a symbol s. A structure of type R, F, σ is a triplet
M = M, RM , F M where M is a non-empty set (universe of M, support
of M), RM = {rM ⊆ M σ(r) | r ∈ R} is a set of relations rM in M corre-
sponding to symbols r ∈ R, and F M = {f M : M σ(r) → M | f ∈ F } is a
set of functions f M in M corresponding to symbols f ∈ F . Each rM ∈ RM
is σ(r)-ary and each f M ∈ F M is σ(f )-ary. If there is no danger of confu-
sion, we omit superscripts and write just r and f for rM and f M . An algebra
is a structure M of some type R, F, σ with R = ∅ (no relation  symbols).

In that case, we write only F, σ and M = M, F M . If M = M, F M is
 {f1 , . . . , fn }, we write only M, f1 , . . . , fn instead of
M M
an algebra and F =
M, {f1 , . . . , fn } , and say that M is of type σ(f1 ), . . . , σ(fn ) .
M M

Example 1.2. (1) Let R = {r}, F = ∅, and σ(r) = 2.


A structure M = M, RM , F M of type R, F, σ consists of a non-empty set
M , a set RM = {rM } consisting of a single binary relation rM in M , and
an empty set F M . That is, a structure of type R, F, σ consists of a set M
equipped with a single binary relation rM . In particular, a partially ordered
set may be considered a structure of the above type R, F, σ .
(2) A monoid is an algebra of type F, σ with F = {◦, e}, σ(◦) = 2,
and σ(e) = 0, satisfying conditions of associativity
 and neutral element. That
is, a monoid is an algebra M = M, ◦M , eM where M is a non-empty set,
◦M is a binary operation in M , and eM is a (constant) element of M (0-
ary operations are constants) such that (a ◦M b ) ◦M c = a ◦M (b ◦M c )
(associativity) and a ◦M eM = eM ◦M a (eM is a neutral element) for each
a , b , c ∈ M . N, +, 0 (positive integers with addition), [0, 1], ·, 1 (real unit
interval with multiplication), [0, 1], min, 1 (real unit interval with minimum)
are examples of monoids.
(3) We introduced lattices as partially ordered sets in which infima and
suprema exist for arbitrary two-element subsets. Alternatively, lattices can be
described as algebras. Call a lattice an algebra M = M, ∧, ∨ of type 2, 2 ,
i.e. ∧ and ∨ are binary operations in M , satisfying
u ∨v =v ∨u u ∧v =v ∧u (commutativity) ,
u ∨ (v ∨ w ) = (u ∨ v ) ∨ w u ∧ (v ∧ w ) = (u ∧ v ) ∧ w (associativity) ,
u ∨ (u ∧ v ) = u u ∧ (u ∨ v ) = u (absorption) ,
for each u , v , w ∈ M . If U = U, ≤ is a lattice ordered set, put u ∧ v =
inf{u , v } and u ∨ v = sup{u , v }. Then U, ∧, ∨ is a lattice. Conversely, if
U = U, ∧, ∨ is a lattice, put u ≤ v iff u ∧ v = u (or, which is equivalent, iff
1.2 Structures of Truth Degrees 7

u ∨ v = v ). Then U, ≤ is a lattice ordered set such that inf{u , v } = u ∧ v


and sup{u , v } = u ∨ v . Moreover, these constructions are mutually inverse.
A bounded lattice is an algebra U, ∧, ∨, 0, 1 where U, ∧, ∨ is a lattice and
0 and 1 satisfy 0 ∨ u = u and 1 ∧ u = u for each u ∈ U . This means that
0 and 1 are the least and the greatest element in the corresponding lattice
ordered set.

We now recall basic structural notions related to algebras. If M = M, F M


is an algebra, a subset N of M is a subuniverse of M if it is closed under any
f M ∈ F M , i.e. if for any f ∈ F , n = σ(f ), and every a1 , . . . , an ∈ M , we
have f M (a1 , . . . , an ) ∈ N whenever a1 , . . . , an ∈ N . Any subuniverse N = ∅
equipped with restrictions of operations of M to N is again analgebra of the
same type as M; it is called a subalgebra of M. If M = M, F M is an algebra,
a binary relation θ in M is said to be compatible (with operations of M) if for
each f ∈ F with σ(f ) = n and every  a1 , b1 , . . . , an , bn ∈ M , we have  that if
a1 , b1 ∈ θ, . . . , an , bn ∈ θ then f M (a1 , . . . , an ), f M (b1 , . . . , bn ) ∈ θ. If θ
is, moreover, an equivalence, it is called a congruence on M. Con(M)  denotes

the set of all congruence relations on M. Given an algebra M = M, F M and
a congruence  θ ∈ Con(M), a factor algebra of M by θ is an algebra M/θ =
M/θ, F M/θ defined by putting f M/θ ([a1 ]θ , . . . , [an ]θ ) = [f M (a1 , . . . , an )]θ
for each n-ary f ∈ F . Let M and N be algebras of the same type F, σ .
A mapping h : M → N is called a morphism (or homomorphism) if for each
f ∈ F with σ(f ) = n and every a1 , . . . , an ∈ M , we have h(f M (a1 , . . . , an )) =
f N (h(a1 ), . . . , h(an )). In such a case we write h : M → N. A morphism
h is called a monomorphism (epimorphism, isomorphism) if h is injective
(surjective, bijective). If there is an isomorphism of M to N, we write M ∼ = N.
A direct Q product of asystem
Q {M iQ| i ∈ I}
 of algebras of the same
Q type F,
Q σ an
algebra i∈IQMi = M , F i∈I Mi with operations f i∈I Mi on M
i∈I i i∈I i
defined by f i∈I Mi (a1 , . . . , an )(i) = f Mi (a1 (i), . . . , an (i)).

1.2 Structures of Truth Degrees


Fuzzy Logic, Graded Truth,
and Why Complete Residuated Lattices

The main idea of fuzzy logic is that of graded truth. In classical logic, each
proposition is either assigned truth degree 1 (true) or truth degree 0 (false).
Classical logic is bivalent. In fuzzy logic, each proposition is assigned a truth
degree taken from some scale L of truth degrees. Scale L is partially ordered
and contains 0 and 1 as its boundary elements. That is, L is equipped with
a partial order ≤ and we have 0 ≤ a ≤ 1 for each a ∈ L. Elements a from L
are called truth degrees. If propositions ϕ and ψ are assigned truth degrees a
and b, which we denote by ||ϕ|| = a and ||ψ|| = b, then a ≤ b means that we
consider ϕ less true than ψ. That is why we want L to be partially ordered.
8 1 Introduction to Fuzzy Sets and Fuzzy Logic

For instance, if ϕ and ψ denote “Bern is a large city” and “New York is a
large city”, we might assign ||ϕ|| = 0.3 and ||ψ|| = 1 to express the fact that
we consider Bern somewhat large but consider New York large without any
reservations. It is in this sense that fuzzy logic is a logic of graded truth. A
favorite example of L is the real unit interval [0, 1] but L need not be ordered
linearly. For instance, a truth degree of “person x is financially well” might
be assigned a two-dimensional truth degree a, b ∈ [0, 1] × [0, 1] with a and b
being truth degrees of “investment of x is high” and “liquidity of x’s assets is
high”. Notice that from this point of view, bivalent logic is a special case of
fuzzy logic – just take L = {0, 1} with 0 ≤ 1.
Scale L needs to be equipped with (truth functions of) logical connectives.
That is, as in classical logic, we need binary functions ⊗ : L × L → L of
conjunction, → : L × L → L of implication, etc. Then we get a structure
L = L, ≤, . . . , ⊗, →, . . . of truth degrees. Like classical logic, fuzzy logic is
truth-functional. Truth-functionality means that a truth degree of a composed
proposition is computed from truth degrees of its constituent propositions
using logical connectives. For instance, if propositions ϕ and ψ are assigned
truth degrees ||ϕ|| and ||ψ||, then the truth degree of proposition ϕ o ψ (“ϕ
and ψ”) is ||ϕ|| ⊗ ||ψ||, i.e. ⊗ applied to ||ϕ|| and ||ψ||.
Contrary to classical logic, logical connectives are not uniquely given in
fuzzy logic. That is, even if we fix a scale L, there is no “the right” conjunction
⊗ for L. The reason for this: In classical logic, conjunction is given by how
“and” is used in natural language. We consider “ϕ and ψ” true if and only
if both ϕ and ψ are true. Whence the definition of conjunction in classical
logic (1 ⊗ 1 = 1, 1 ⊗ 0 = 0 ⊗ 1 = 0 ⊗ 0 = 0). Intuition behind the usage
of “and” in natural language fails us if more truth degrees come into play.
If ||ϕ|| = 0.7 and ||ψ|| = 0.8, what is the truth degree of “ϕ and ψ”, i.e.
what is 0.7 ⊗ 0.8? There is no obvious answer. In fuzzy logic, rather than
picking one particular connective, it is useful to postulate desirable properties
of connectives and consider as a “good connective” any one satisfying the
properties. For instance, a desirable property of conjunction is monotony – the
more true the propositions, the more true their conjunction. This translates
to requiring a1 ⊗ b1 ≤ a2 ⊗ b2 whenever a1 ≤ a2 and b1 ≤ b2 . In the following,
we proceed this way to justify a particular structure of truth degrees. The
structure will be called a complete residuated lattice and will play the role
of a basic structure of truth degrees in our investigation. Our justification of
complete residuated lattices as suitable structures of truth degrees is due to
Goguen [45].
We already agreed on a set L of truth degrees equipped with a partial order
≤ and bounded by 0 and 1. Furthermore, we want L to have arbitrary infima
and suprema, i.e. we want L, ≤, 0, 1 to be a complete residuated lattice.
Why infima and suprema? Let X be a collection of persons of our interest
and let ϕ(x) denote a proposition “person x is tall”. The truth degree of this
proposition is ||ϕ(x)||. What then is the truth degree of “there is a person x
from X who is tall”? We argue that a good choice is the supremum of ||ϕ(x)||
1.2 Structures of Truth Degrees 9

over x ∈ X, i.e. ||“there is a person x from X who is tall”|| = x∈X ||ϕ(x)||.
Intuitively, evaluating “there is a person x from X who is tall” means going
through x from X and take the best, i.e. the highest, ||ϕ(x)|| and this is
exactly what supremum does. That is why we need suprema in L. The need
of infima can be justified dually (as a truth degree of propositions of the type
“each person x from X is tall”).
We now turn to (truth function of) conjunction. We denote it by ⊗. Basic
requirements are that L, ⊗, 1 is a commutative monoid. This means that ⊗ is
a binary operation on L which is commutative, associative, and 1 is a neutral
element of ⊗. Commutativity of ⊗ means that we have a ⊗ b = b ⊗ a for each
a, b ∈ L. If a and b are truth degrees of propositions ϕ and ψ, respectively, then
commutativity of ⊗ means ||ϕ o ψ|| = ||ϕ|| ⊗ ||ψ|| = ||ψ|| ⊗ ||ϕ|| = ||ψ o ϕ||,
i.e. the truth degree of “ϕ and ψ” equals the truth degree of “ψ and ϕ”. This
is an intuitively desirable property justifying commutativity. Associativity of
⊗ means a ⊗ (b ⊗ c) = (a ⊗ b) ⊗ c for each a, b, c ∈ L. Associativity results from
requiring that the truth degree of “ϕ and (ψ and χ)” equals the truth degree
“(ϕ and ψ) and χ” which we consider a desirable property as well. That 1 is
a neutral element of ⊗ says a ⊗ 1 = a for each a ∈ L. This results from the
following requirement. If ψ is a proposition which is fully true, i.e. ||ψ|| = 1,
then for any proposition ϕ we have ||ϕ o ψ|| = ||ϕ|| ⊗ ||ψ|| = ||ϕ|| ⊗ 1 = ||ϕ||,
i.e. the truth degree of a conjunction of ϕ with a fully true proposition equals
the truth degree of ϕ. We held this property desirable as well.
We proceed by (truth function of) implication which we denote by →. Sev-
eral desirable properties of implication follow from a simple condition called
adjointness. Adjointness itself says that for each a, b, c ∈ L we have a⊗b ≤ c if
and only if a ≤ b → c. We show that adjointness follows from how modus po-
nens should behave in fuzzy setting. Recall first that in classical logic, modus
ponens is an inference rule saying: if ϕ is valid and ϕ ⇒ ψ (“ϕ implies ψ”) is
valid then we may infer that ψ is valid. An appropriate formulation of modus
ponens in fuzzy setting is the following: if ϕ is valid in degree at least a and
ϕ ⇒ ψ is valid in degree at least b then we may infer that ψ is valid in de-
gree at least a ⊗ b. Observe first that if “a formula is valid” is understood
as a shorthand for “a formula is valid in degree 1” then the formulation of
modus ponens in fuzzy setting is equivalent to that in classical setting. This
follows by a moment’s reflection since in classical logic we have only 0 and 1
as truth degrees and any formula is always valid in degree at least 0. Using
modus ponens, we get a ⊗ b as a lower estimation of degree of validity of ψ.
Now, we want modus ponens to satisfy two points: it should be sound and, at
the same time, it should yield the highest possible estimation of validity of ψ.
Soundness: The requirement of soundness says that when evaluating formu-
las, if the truth degree of ϕ is at least a (a ≤ ||ϕ||) and the truth degree of
ϕ ⇒ ψ is at least b (b ≤ ||ϕ ⇒ ψ||) then the truth degree of ψ is at least as
high as the degree obtained by modus ponens (a ⊗ b ≤ ||ψ||). In words, from
lower estimations of ϕ and ϕ ⇒ ψ, modus ponens yields a lower estimation
of ψ. This ensures that we do not get more than the actual truth degree of ψ.
10 1 Introduction to Fuzzy Sets and Fuzzy Logic

Particularly, if ||ϕ|| = a and ||ψ|| = c, then since ||ϕ ⇒ ψ|| = a → c, sound-


ness says that b ≤ a → c implies a ⊗ b ≤ c. Highest possible estimation of
validity of ψ: Let ||ϕ|| = a and ||ψ|| = c. Then ||ϕ ⇒ ψ|| = a → c and since
we require soundness, a ⊗ (a → c) needs to be a lower estimation of c, i.e.
a ⊗ (a → c) ≤ c. Since ⊗ is non-decreasing, the lower estimation a ⊗ (a → c)
is the higher, the higher a → c. Since we want a ⊗ (a → c) to be the high-
est possible estimation of c, a → c needs to be the highest degree for which
a ⊗ (a → c) ≤ c. That is, if b is any truth degree for which a ⊗ b ≤ c then
b ≤ a → c. In other words, the requirement of highest estimation of validity
of ψ yields that a ⊗ b ≤ c implies b ≤ a → c. Putting the two conditions
together we get that ⊗ and → should satisfy
a ⊗ b ≤ c if and only if b ≤ a → c
for any truth degrees a, b, c ∈ L. A structure consisting of a set L of truth
degrees, a partial order ≤ bounded by 0 and 1, and functions ⊗ and → which
satisfy the above conditions is called a complete residuated lattice and plays
the role of a basic structure of truth degrees in our investigation. This structure
will be furthermore equipped with a unary operation ∗ called a truth stresser.
A truth stresser ∗ serves as (truth function of) unary connective “very true”
and will be introduced later.
Note that we might impose additional conditions on the structure of truth
degrees if desirable for our purpose. In such a case, we restrict our attention
only to structures satisfying additional conditions. For instance, we might
want conjunction ⊗ to be idempotent, i.e. to satisfy a ⊗ a = a for any a ∈ L.
Or, in the light of the 7 ± 2 phenomenon well-known from psychology, we
might restrict our attention to structures with up to 7 ± 2 elements. Recall
according to Miller’s 7 ± 2 phenomenon [65], humans are able to assign de-
grees to elements of some universe in a consistent manner if the scale L of
truth degrees contains up to 7 ± 2 elements. With more than 7 ± 2 degrees,
the assignment becomes inconsistent. For instance, when the assignment is
performed twice, the same person may assign different truth degrees to the
same element of universe.

Complete Residuated Lattices with Truth Stressers

We first introduce residuated lattices, their examples, and their basic proper-
ties. Then we add truth stressers, their examples, and properties.

Definition 1.3. A complete residuated lattice is an algebra


L = L, ∧, ∨, ⊗, →, 0, 1 where
(i) L, ∧, ∨, 0, 1 is a complete lattice with the least element 0 and the greatest
element 1,
(ii) L, ⊗, 1 is a commutative monoid, i.e. ⊗ is associative, commutative,
and a ⊗ 1 = a for each a ∈ L,
1.2 Structures of Truth Degrees 11

(iii) ⊗ and → satisfy adjointness, i.e.


a⊗b≤c iff a≤b→c (1.1)
for each a, b, c ∈ L (≤ denotes the lattice ordering).

Remark 1.4. (1) If we weaken (i) so that it requires L, ∧, ∨, 0, 1 to be a


bounded lattice, not necessarily a complete one, L is called a residutated lat-
tice.
(2) The operations ⊗ and → are called multiplication and residuum, re-
spectively, ⊗, → is called an adjoint pair. For a, b ∈ L, a → b is called the
residuum of b by a. For a given ⊗, there is at most one operation → satisfying
(1.1). For if → is another one, we get a ≤ b → c iff a ⊗ b ≤ c iff a ≤ b → c
for any a, from which it clearly follows b → c = b → c. Similarly, → uniquely
determines ⊗. n
(3) Since ⊗ is commutative and associative, we may write i=1 ai to
denote any ⊗-product of a1 , . . . , an . For instance, a1 ⊗(a2 ⊗a3 ), (a1 ⊗a2 )⊗a3 ,
a2 ⊗ (a1 ⊗ a3 ), . . . , (a3 ⊗ a2 ) ⊗ a1 , all have the same value and this value is
3
denoted by i=1 ai .

Most important examples of complete residuated lattices are those with


the universe L being a real unit interval [0, 1] or its subchain.

Example 1.5. Take L = [0, 1] (the interval of all reals between 0 and 1). The
natural ordering of [0, 1] makes L, ∧, ∨, 0, 1 a complete lattice where a ∧ b =
min(a, b), a ∨ b = max(a, b). Each of the following pairs of operations makes
L, ∧, ∨, ⊗, →, 0, 1 a complete residuated lattice:
a ⊗ b = max(a + b − 1, 0) ,
L
 ukasiewicz operations: (1.2)
a → b = min(1 − a + b, 1) ,
a ⊗ b = min(a, b) ,

Gödel operations: 1 if a ≤ b , (1.3)
a→b =
b otherwise ,

a⊗b = a·b,

Goguen (product) operations: 1 if a ≤ b , (1.4)
a→b = b
a otherwise .
The corresponding algebras L are called standard L
 ukasiewicz algebra, stan-
dard Gödel algebra, and standard Goguen (product) algebra on [0, 1].

Example 1.6. Take L = {a0 , a1 , . . . , an } with an order given by 0 = a0 < · · · <


an = 1. We have ai ∧ aj = amin(i,j) and ai ∨ aj = amax(i,j) . The following two
pairs of operations determine a complete residuated lattice structure on L:
ak ⊗ al = amax(k+l−n,0) ,
L
 ukasiewicz operations: (1.5)
ak → al = amin(n−k+l,n) ,
12 1 Introduction to Fuzzy Sets and Fuzzy Logic

ak ⊗ al = amin(k,l) ,

Gödel operations: 1 if k ≤ l , (1.6)
ak → al =
al otherwise .
If {a0 , . . . , an } ⊆ [0, 1] and ai = ni , then all the operations are restrictions of
 ukasiewicz operations on [0, 1] to {a0 , . . . , an }, i.e. {a0 , . . . , an } is a subuni-
L
verse of the standard L  ukasiewicz algebra. If {a0 , . . . , an } ⊆ [0, 1] and a0 = 0,
an = 1, then all the operations are restrictions of the Gödel operations on
[0, 1] to {a0 , . . . , an }, i.e. {a0 , . . . , an } is a subuniverse of the standard Gödel
algebra.
Remark 1.7. If L = {0, 1} then both of the structures from Example 1.6 yield
the same, namely, the two-element Boolean algebra, i.e. the structure of truth
degrees of classical two-valued logic. We denote this algebra by 2 and write
2 = {0, 1}. There is no other residuated lattice structure on {0, 1} (easy to
check). Anything we establish for a general complete residuated lattice L
(definitions, theorems, etc.), has its special “instance” for L = 2. This is a
basic way of generalization from ordinary (bivalent) setting to fuzzy setting
in our approach. For instance, if we prove TheoremL true for an arbitrary
complete residuated lattice L and the instance Theorem2 of TheoremL for L =
2 is known from ordinary setting, we say that TheoremL is a generalization
of Theorem2 to fuzzy setting. We will see particular examples later on.
Example 1.8. (1) Residuated lattices from Example 1.5 are particular cases of
those induced by so-called left-continuous t-norms (see Definition 1.27 and the
subsequent paragraphs for details). A t-norm is a binary operation ⊗ on [0, 1]
which is associative, commutative, has 1 as its neutral element, and is non-
decreasing. A t-norm ⊗ is left-continuous if limn→∞ (an ⊗b) = (limn→∞ an )⊗b
for any non-decreasing sequence {an ∈ [0, 1] | n = 1, 2, 3, . . . }. For a left-
continuous t-norm ⊗, put
a → b = sup{c | a ⊗ c ≤ b} . (1.7)
Then [0, 1], min, max, ⊗, →, 0, 1 is a complete residuated lattice. It is easy to
see that each of the three operations ⊗ from Example 1.5 is even a continuous
t-norm and that the corresponding residua are obtained by (1.7).
(2) Residuated lattices on L = {a0 , a1 , . . . , an } from Example 1.6 are par-
ticular cases of the following one. Take I = {i1 = 0, . . . , im = n} ⊆ {0, . . . , n}
with i0 < · · · < im and define

amax(k+l−ij+1 ,ij ) if k, l ∈ [ij , ij+1 ]
ak ⊗ al =
amin(k,l) otherwise


 1 if k ≤ l

amin(ij+1 −k+l,ij+1 ) if k > l and
ak → al =

 k, l ∈ [ij , ij+1 ]

al otherwise .
Then L, ∧, ∨, ⊗, →, 0, 1 is a complete residuated lattice. For I = {0, 1} we
get (1.5) and for I = {0, 1, . . . , n} we get (1.6).
1.2 Structures of Truth Degrees 13

1 ⊗ 0 a b c d e f g 1 → 0 a b c d e f g 1
0 0 0 0 0 0 0 0 0 0 0 1 1 1 1 1 1 1 1 1
g a 0 a a a 0 a 0 a a a f 1 1 1 f 1 f 1 1
b 0 a a a 0 a 0 a b b f g 1 g f g f 1 1
e c 0 a a a 0 a 0 a c c f g g 1 f g f 1 1
b c f d 0 0 0 0 0 0 0 0 d d g g g g 1 1 1 1 1
e 0 a a a 0 a 0 a e e f g g g f 1 f 1 1
a d f 0 0 0 0 0 0 0 0 f f g g g g g g 1 1 1
g 0 a a a 0 a 0 a g g f g g g f g f 1 1
0 1 0 a b c d e f g 1 1 0 a b c d e f g 1

Fig. 1.2. Non-linear residuated lattice from Example 1.9.

Example 1.9. Figure 1.2 shows a non-linear residuated lattice L with elements
0, a, . . . , g, 1. The lattice part L, ≤ of L is shown by its Hasse diagram (left).
The adjoint operations ⊗ and → are shown by their tables (middle and right).

The following are important operations derived in terms of basic operations


of residuated lattices.
Definition 1.10. For a complete residuated lattice L we define
a ↔ b = (a → b) ∧ (b → a) , (1.8)
¬a = a → 0 , (1.9)
a0 = 1 and an+1 = an ⊗ a , (1.10)
for each a, b ∈ L and a non-negative integer n. The operations ↔, ¬, and n

are called biresiduum, negation, and n-th power in L.


One can easily check that for the biresidua corresponding to standard
L
 ukasiewicz, Gödel, and product algebras on [0, 1] we have
a ↔ b = 1 − |a − b| ,

1 for a = b
a↔b=
min(a, b) otherwise ,

1 for a = b = 0
a ↔ b = min(a,b)
max(a,b) otherwise ,
respectively. For negations of standard L
 ukasiewicz, Gödel, and product alge-
bras on [0, 1] we have
¬a = 1 − a
for L
 ukasiewicz and

1 for a = 0
¬a =
0 for a > 0
for both Gödel and product. For powers of standard L
 ukasiewicz, Gödel, and
product algebras on [0, 1] we have
14 1 Introduction to Fuzzy Sets and Fuzzy Logic

an = max(0, 1 − n(1 − a)) ,


an = a ,
an = the usual n-th power of a ,
respectively.
Several important structures of truth degrees are just residuated lattices
satisfying additional conditions. Some of these are introduced by the following
definition.

Definition 1.11. A Heyting algebra is a residuated lattice satisfying a⊗b =


a ∧ b. A BL-algebra is a residuated lattice which satisfies a ∧ b = a ⊗ (a → b)
(divisibility) and (a → b) ∨ (b → a) = 1 (prelinearity). An MV-algebra
is a BL-algebra satisfying a = ¬¬a (law of double negation). A Π-algebra
(product algebra) is a BL-algebra satisfying (c → 0) → 0 ≤ ((a ⊗ c) →
(b ⊗ c)) → (a → b) and a ∧ (a → 0) = 0. A G-algebra (Gödel algebra) is a
BL-algebra which satisfies a ⊗ a = a (idempotency). A Boolean algebra is
a residuated lattice which is both a Heyting algebra and an MV-algebra.

Remark 1.12. (1) Consider residuated lattices from Examples 1.5 and 1.6.
Those with L  ukasiewicz structure are MV-algebras, those with Gödel struc-
ture are G-algebras, and that with product structure is a Π-algebra.
(2) Recall that a Boolean algebra is usually defined as a bounded lattice
which is distributive (i.e., satisfies a ∧ (b ∨ c) = (a ∧ b) ∨ (a ∧ c)) and comple-
mented (i.e., a unary operation  exists such that a ∧ a = 0 and a ∨ a = 1
for each a). In such a case, putting a ⊗ b = a ∧ b and a → b = a ∨ b, we get a
Boolean algebra in the sense of Definition 1.11. Conversely, having a Boolean
algebra in the sense of Definition 1.11, we get a Boolean algebra according to
the usual definition if we put a = a → 0.
(3) If L is a Heyting algebra, a → b is also called a relative pseudocomple-
ment of a in b.

We now turn our attention to truth stressers – particular unary operations


on complete residuated lattices.

Definition 1.13. Let L = L, ∧, ∨, ⊗, →, 0, 1 be a complete residuated lat-


tice. A unary operation ∗ : L → L satisfying
1∗ = 1 , (1.11)
a∗ ≤ a , (1.12)

(a → b) ≤ a∗ → b∗ , (1.13)
for every a, b ∈ L, is called a truth stresser for L.
The algebra L, ∧, ∨, ⊗, →, ∗ , 0, 1 is then called a complete residuated lat-
tice with truth stresser.
1.2 Structures of Truth Degrees 15


Example 1.14. (1) Let be the identity on L, i.e.
a∗ = a , (1.14)

a ∈ L. Then is a truth stresser for L.
(2) Let ∗ be defined by

1 if a = 1 ,
a∗ = (1.15)
0 otherwise .
Then ∗ , called a globalization in [83], is a truth stresser for L.
(3) Denote the two above truth stressers by ∗1 (identity) and ∗2 (globaliza-
tion). Trivially, for each truth stresser ∗ , if a < 1 then a∗2 = 0 ≤ a∗ ≤ a = a∗1 ,
and 1 = 1∗1 = 1∗ = 1∗2 . Therefore, truth stressers are bounded by ∗1 and ∗2 .
Remark 1.15. (1) If L is a complete residuated lattice and ∗ a truth stresser
for L then the resulting complete residuated lattice with truth stresser, i.e.
the algebra L, ∧, ∨, ⊗, →, ∗ , 0, 1 , will be denoted by L∗ .
(2) A truth stresser ∗ may be thought of as a (truth function of) unary
connective “very true”. That is, if ϕ is a proposition then a truth degree

of “ϕ is very true” is ||ϕ|| . This interpretation of truth stressers is due to
Hájek [51]. In particular, identity may be thought of as connective “· · · is
true”. Globalization may be thought of as connective “· · · is fully true”. As
we will see, a ≤ b is equivalent to a → b = 1. Therefore, ||ϕ|| ≤ ||ψ|| means that
||ϕiψ|| = 1 (truth degree of “ϕ implies ψ” is 1). Taking this into account, we
can see that conditions (1.11)–(1.13) are natural properties of truth function of
“very true”. Namely, (1.11) ensures that if ϕ is true in degree 1 then “ϕ is very
true” has degree 1 (succinctly: if ϕ is fully true then ϕ is very true). Equation
(1.12) says that if ϕ is very true then ϕ is true. Finally, using adjointness,

(1.13) is equivalent to a∗ ⊗ (a → b) ≤ b∗ , and so it says that if ϕ is very true
and if it is very true that ϕ implies ψ then ψ is very true.

(3) Note that for technical reasons Hájek requires also (a ∨ b) ≤ a∗ ∨ b∗

in [51]. Note also that a truth stresser satisfying (a ∨ b) ≤ a∗ ∨ b∗ and

a∗ ∨ ¬a∗ = 1 is just a truth function of a unary connective  satisfying


Baaz’s axioms for  from [4]. We will see later that if L is a linearly ordered
residuated lattice then ∗ is a truth stresser for L satisfying the two additional
conditions iff ∗ is globalization.
(4) It is an obvious but useful fact that if we prove a theorem true for a
particular complete residuated lattice L, then this theorem remains true for
any L∗ resulting from L by an arbitrary truth stresser ∗ for L.
We continue with further examples of truth stressers.
Example 1.16. (1) For any L, c1 , . . . , ck ∈ L, and non-negative integers n1 , . . . ,
nk , let

1 if a = 1 ,
a∗ = k (1.16)
i=1 c i ⊗ ani
if a<1.

Then is a truth stresser, see Lemma 1.49.
16 1 Introduction to Fuzzy Sets and Fuzzy Logic

Fig. 1.3. Truth stressers for residuated lattice from Fig. 1.2

Fig. 1.4. All truth stressers on five-element Lukasiewicz chain

Fig. 1.5. Truth stressers on five-element Gödel chain

(2) Consider the following special cases of (1.16). For k = 1, we have 1∗ = 1


and a∗ = c ⊗ an for a < 1. In particular, for c = 0 we get globalization while
for c = 1 and n = 1, we get identity. For L = [0, 1] and ⊗ = min, (1.16) yields
a∗ = 1 for a = 1, and for a < 1 we get a∗ = c if a ≥ c and a∗ = a if a < c.
(3) For a residuated lattice L from Fig. 1.2 (see Example 1.2), there exist
56 truth stressers (generated by a computer program). Three of them are
depicted in Fig. 1.3.
(4) Figure 1.4 shows all truth stressers for a five-element chain L equipped
with L ukasiewicz operations (see Example 1.6). For a five-element chain L
equipped with Gödel operations (see Example 1.6), each unary function from
Fig. 1.4 is a truth stresser for L. In addition to that, there are four more truth
stressers for L and these are depicted in Fig. 1.5.
In our investigation, we will need truth stressers satisfying additional con-
ditions. Here are two collections of additional conditions.
Definition 1.17. Let ∗ be a truth stresser for L. ∗ is called an implicational
truth stresser if it satisfies
1.2 Structures of Truth Degrees 17

a∗∗ = a∗ , (1.17)
∗ ∗ ∗
a ⊗a =a , (1.18)
 ∗
 ∗
i∈I ai = i∈I ai , (1.19)
for every a ∈ L, ai ∈ L for all i ∈ I. ∗ is called a Horn truth stresser if it
is an implicational truth stresser which satisfies
a ⊗ b ∗ = a ∧ b∗ , (1.20)
 ∗  ∗
(a → i∈I bi ) = i∈I (a → bi ) , (1.21)
for every non-empty index set I and a, b, bi ∈ L.

Remark 1.18. (1) Since a∗∗ ≤ a∗ for each truth stresser ∗ , (1.17) is equivalent
to a∗∗ ≥ a∗ . (1.18) says that each a∗ is an idempotent element  of ⊗.∗ As
we
 will see, truth stressers are monotone w.r.t. ≤ and so (a →  bi ) ≥∗
i∈I


i∈I (a → b i ) is always true. Therefore, (1.21) is equivalent to (a → i∈I bi )

≤ i∈I (a → bi ) .
(2) Condition (1.20) is not too restrictive for implicational truth stressers.
Namely, if L satisfies divisibility, i.e. a ∧ b = a ⊗ (a → b) for each a, b ∈ L,
then if b is an idempotent element, i.e. b ⊗ b = b, we have a ⊗ b = a ∧ b for
each a ∈ L, see e.g. [10, Theorem 2.39]. Since b∗∗ = b∗ for an implicational
truth stresser, (1.20) is implied by divisibility. Therefore, in particular, (1.20)
is satisfied by implicational truth stressers on  every BL-algebra.

(3) Condition (1.21) is similar to a → i∈I bi = i∈I (a → bi ) (left-
continuity of → in the second argument, see Lemma 1.32). However, our con-
dition seems to be more restrictive. For instance, in the standard L  ukasiewicz
algebra on [0, 1], we have a → b = min(1 − a + b, 1), and so → satisfies left-
continuity in the second argument. But for ∗ being globalization, condition
(1.21) is not true. Take I = [0, 0.5), bi = i for every i ∈ I, so for a = 0.5 we
have
∗   ∗
(0.5 → 0.5) = 1∗ = 1  0 = i ∈[0,0.5) 0∗ = i ∈[0,0.5) (0.5 → i) .
(4) IfL is a Noetherian chain (i.e. L, ≤ is a Noetherian
 lattice which is a
∗ ∗
chain), i∈I bi = bi0 for some i0 ∈ I. Thus, (a → bi0 ) ≤ i∈I (a → bi ) , and
so (1.21) is always satisfied.
(5) Let us mention that (1.21) is required for non-empty index sets. For

empty index set I = ∅, (1.21) yields (a → 0) ≤ 0, but for a = 0 we would

have (0 → 0) = 1∗ = 1  0 which is never satisfied.
(6) We will see later that for condition (1.21) simplifies in some particular
cases.

Example 1.19. The following are examples of implicational truth stressers.


(1) Identity is an implicational truth stresser for any complete Heyting
algebra.
(2) Globalization is an implicational truth stresser for every complete resid-
uated lattice L. Obviously, (1.17), (1.18) are satisfied. For every index set I
18 1 Introduction to Fuzzy Sets and Fuzzy Logic

and ai ∈ L (i ∈ I) we have i∈I a∗i = 1 iff for every i ∈ I we have a∗i = 1
  ∗ 
iff a = 1 iff i∈I ai = 1 iff i∈I ai = 1. Consequently, i∈I a∗i = 0 iff
 i ∗
i∈I ai = 0. Hence, condition (1.19) holds as well.
(3) The left-most truth stresser in Fig. 1.3 is an implicational truth stresser.

Properties of Complete Residuated Lattices

We now prove basic properties of complete residuated lattices.

Theorem 1.20. Each complete residuated lattice satisfies


a ⊗ (a → b) ≤ b, b ≤ a → (a ⊗ b), a ≤ (a → b) → b , (1.22)
a≤b iff a→b=1, (1.23)
a → a = 1, a → 1 = 1, 0→a=1, (1.24)
1→a=a, (1.25)
a⊗0=0, (1.26)
a ⊗ b ≤ a, a ≤ b → a , (1.27)
a⊗b≤a∧b, (1.28)
(a ⊗ b) → c = a → (b → c) = b → (a → c) , (1.29)
(a → b) ⊗ (b → c) ≤ a → c , (1.30)
a1 ≤ a2 and b1 ≤ b2 implies a1 ⊗ b1 ≤ a2 ⊗ b2 , (1.31)
b1 ≤ b2 implies a → b1 ≤ a → b2 , (1.32)
a1 ≤ a2 implies a2 → b ≤ a1 → b , (1.33)
(a → b) ⊗ (c → d) ≤ (a ⊗ c) → (b ⊗ d) , (1.34)
(a → b)n ≤ an → bn , (1.35)
a → b is the greatest element of {c | a ⊗ c ≤ b} , (1.36)
a ⊗ b is the least element of {c | a ≤ b → c} . (1.37)

Proof. (1.22): Follows directly by adjointness and commutativity of ⊗. For


instance, we have (a → b) ⊗ a = a ⊗ (a → b) ≤ b iff a → b ≤ a → b, which is
true by reflexivity of ≤.
(1.23): By adjointness, a ≤ b iff 1 ⊗ a ≤ b iff 1 ≤ a → b iff 1 = a → b.
(1.24): Direct consequences of (1.23).
(1.25): We check both inequalities. a ≤ 1 → a iff a = 1 ⊗ a ≤ a which is true.
Conversely, 1 → a ≤ 1 → a yields 1 → a = 1 ⊗ (1 → a) ≤ a.
(1.26) is equivalent to a ⊗ 0 ≤ 0 which is, by adjointness, equivalent to 0 ≤
a → 0 which is true.
(1.27): By (1.24), a → a = 1, hence b ≤ a → a, and so a⊗b ≤ a by adjointness.
Applying adjointness to the last inequality once more, we get a ≤ b → a.
(1.28): By (1.27), a ⊗ b ≤ a and a ⊗ b ≤ b from which (1.28) immediately
follows.
1.2 Structures of Truth Degrees 19

(1.29): It is sufficient to prove (a ⊗ b) → c = a → (b → c) since then,


(a ⊗ b) → c = b → (a → c) follows from commutativity of ⊗. We check both
inequalities of (a ⊗ b) → c = a → (b → c). On the one hand, (a ⊗ b) → c ≤
a → (b → c) iff (applying adjointness twice) (a ⊗ b) ⊗ ((a ⊗ b) → c) ≤ c
which is true by (1.22). Conversely, a → (b → c) ≤ (a ⊗ b) → c iff (by
adjointness) b ⊗ a ⊗ (a → (b → c)) ≤ c which is true. Indeed, (1.22) yields
b ⊗ a ⊗ (a → (b → c)) ≤ b ⊗ (b → c) ≤ c.
(1.30) is equivalent to a ⊗ (a → b) ⊗ (b → c) ≤ c which is true by (1.22).
(1.31): By (1.22), b2 ≤ a1 → (a1 ⊗ b2 ), and so b1 ≤ a1 → (a1 ⊗ b2 ). Using
adjointness we get a1 ⊗ b1 ≤ a1 ⊗ b2 . Due to commutativity of ⊗ we also have
a1 ⊗ b2 ≤ a2 ⊗ b2 . Putting together, we get a1 ⊗ b1 ≤ a2 ⊗ b2 .
(1.32): By (1.22), a ⊗ (a → b1 ) ≤ b1 , and thus a ⊗ (a → b1 ) ≤ b2 . Applying
adjointness we get a → b1 ≤ a → b2 .
(1.33): a2 → b ≤ a1 → b is equivalent to a1 ⊗ (a2 → b) ≤ b which is true.
Indeed, from a1 ≤ a2 (assumption) we have by (1.31) and (1.22) that a1 ⊗
(a2 → b) ≤ a2 ⊗ (a2 → b) ≤ b.
(1.34): By adjointness, (a → b) ⊗ (c → d) ≤ (a ⊗ c) → (b ⊗ d) is equivalent to
(a ⊗ c) ⊗ (a → b) ⊗ (c → d) ≤ (b ⊗ d) which is true by (1.22) and (1.31) since
(a ⊗ c) ⊗ (a → b) ⊗ (c → d) = a ⊗ (a → b) ⊗ c ⊗ (c → d) ≤ (b ⊗ d).
(1.35): By induction: the assertion is trivial for n = 1. If the assertion is true
for n, we have (a → b)n+1 ≤ an+1 → bn+1 iff an+1 ⊗ (a → b)n+1 ≤ bn+1 which
is true since an+1 ⊗ (a → b)n+1 = a ⊗ (a → b) ⊗ an ⊗ (a → b)n ≤ b ⊗ bn = bn+1
is a consequence of 1.31 and the induction hypothesis.
(1.36): We have a ⊗ (a → b) ≤ b by (1.22), whence a → b ∈ {c | a ⊗ c ≤ b}. If
a ⊗ c ≤ b then by adjointness, c ≤ a → b.
(1.37): We have a ≤ b → (a ⊗ b) by (1.22), thus a ⊗ b ∈ {c | a ≤ b → c}. If
a ≤ b → c then by adjointness, a ⊗ b ≤ c. 

Theorem 1.21. Each complete residuated lattice satisfies
 
a ⊗ i∈I bi = i∈I (a ⊗ bi ) , (1.38)
 
a → i∈I bi = i∈I (a → bi ) , (1.39)
 
i∈I ai → b = (ai → b) , (1.40)
  i∈I
a ⊗ i∈I bi ≤ i∈I (a ⊗ bi ) , (1.41)
 
(a → bi ) ≤ a → i∈I bi , (1.42)
i∈I 
(ai → b) ≤ i∈I ai → b , (1.43)
i∈I  
i∈I (ai → bi ) ≤ i∈I ai → i∈I bi . (1.44)

Proof. (1.38):Isotony of ⊗ gives a ⊗ bi ≤ a ⊗ i∈I bi for each i ∈ I from
which
 we get i∈I (a ⊗ bi ) ≤ a⊗ i∈I bi . On the other hand, we have a ⊗ bi ≤
i∈I ⊗ bi ), i.e.
(a  bi ≤ a → i∈I (a ⊗ bi )for each i ∈ I from which we get
b
i∈I i ≤ a → i∈I (a ⊗ bi ) and thus a ⊗ i∈I bi ≤ i∈I (a ⊗ bi ).
(1.39): Isotony of → in the second argument  yields a
 → i∈I bi ≤ a → bi
for each i ∈ I from which
 we get a → b
i∈I i ≤ i∈I → bi ). On the
(a
other hand, we have i∈I (a → bi ) ≤ a → bi , i.e. a ⊗ i∈I (a → bi ) ≤ bi
20 1 Introduction to Fuzzy Sets and Fuzzy Logic
 
 each i ∈ I, fromwhich we get a ⊗ i∈I (a → bi ) ≤ i∈I bi and thus
for
i∈I (a → bi ) ≤ a → i∈I bi . 
(1.40): Antitony of → inthe first argument  yields i∈I ai → b ≤ ai → b for
each
 i ∈ I fromwhich i∈I ai → b ≤ i∈I (ai → b).
 On the other hand,
(a
i∈I i → b) ≤ a
i∈I i → b
 is equivalent
 to i∈I i  i∈I (ai → b) ≤ b which
a ⊗
is true since for any i ∈ I, i∈I ai ⊗ i∈I (ai → b) ≤ i∈I ai ⊗ (ai → b) ≤ b.
(1.41): Follows from monotony of ⊗ while (1.42) and (1.43) follow from an-
titony and isotony of →, respectively.  
(1.44):
 By adjointness,(1.44) isequivalent to i∈I ai ⊗ i∈I (ai → bi ) ≤
i∈I bi which is true iff  i∈I ai ⊗ i∈I (ai → bi ) ≤ bi for each i ∈ I. The latter
inequality is true since i∈I ai ⊗ i∈I (ai → bi ) ≤ ai ⊗ (ai → bi ) ≤ bi . 

The following theorem shows some conditions equivalent to adjointness.
Theorem 1.22 (adjointness). Let L, ∧, ∨, ⊗, 0, 1 be a structure satisfying
(i) and (ii) of Definition 1.3. Then the following conditions are equivalent.
There exists → satisfying adjointness (with ⊗) . (1.45)
For any a, b, {c | a ⊗ c ≤ b} has a greatest element . (1.46)
 
a ⊗ i∈I bi = i∈I (a ⊗ bi ) holds in L . (1.47)

For a → b = {c | a ⊗ c ≤ b} , ⊗ and → satisfy adjointness . (1.48)
Proof. Trivially, (1.48) implies (1.45). Since, by Theorems 1.20 and 1.21, (1.45)
implies (1.46) and (1.45) implies (1.47), it suffices to prove that (1.46) implies
(1.48) and (1.47) implies (1.48).
Assume (1.46) and define → as in (1.48). Then a → b is the greatest
element of {c | a ⊗ c ≤ b}. Therefore, if a ⊗ c ≤ b then c ≤ a → b. If c ≤ a → b
then, by isotony of ⊗ and by the property of a → b, a ⊗ c ≤ a ⊗ (a → b) ≤ b.
Hence, ⊗ and → defined in (1.48) satisfy adjointness, i.e. (1.48) holds true.
Assume (1.47) and define → as in (1.48). If a⊗c ≤ b then clearly c ≤ a → b
(a → b is the supremumof such c’s). If c ≤ a → b then a ⊗ c ≤ a ⊗ (a → b) =
a ⊗ {d | a ⊗ d ≤ b} = {a ⊗ d | a ⊗ d ≤ b} ≤ b. Therefore, (1.48) is true.  
In what follows, we introduce a particular construction of complete resid-
uated lattices. The construction is called an ordinal sum and will be used
later.
Definition 1.23. Suppose I, ≤ is a chain with the least element 0 and the
greatest element 1. For i ∈ I, put
i+ = inf{j ∈ I | i < j} .
Thus either i+ covers i (i.e., i < i+ and there is no j with i < j < i+ ) or
i+ = i (if there is no j which covers i). Let {Li | i ∈ I} be a family of complete
residuated lattices Li = Li , ∧i , ∨i , ⊗i , →i , 0i , 1i . Suppose that for every i ∈ I
we have 1i = 0i+ and Li ∩Li+ = {0i+ }, and for every i, j ∈ I such that i+ < j,
we have Li ∩ Lj = ∅. Thus if i < i+ then the top element of Li coincides with
the bottom element of Li+ , and if i = i+ then Li is a singleton.
1.2 Structures of Truth Degrees 21

An ordinal sum i∈I Li of the family {Li  | i ∈ I} is a complete residu-
ated lattice L = L, ∧, ∨, ⊗, →, 0, 1 , where L = i∈I Li , 0 = 00 , 1 = 11 , and
a ≤ b iff either a, b ∈ Li , a ≤i b, or a ∈ Li , b ∈ Lj and i < j. Operations ∧, ∨
are the infimum and supremum w.r.t. ≤, and ⊗, → are defined by

a ⊗i b if a, b ∈ Li ,
a⊗b=
a∧b otherwise ,

1 if a ≤ b ,
a→b= b if a ∈ Lj − Li , b ∈ Li − Lj , i < j ,

a →i b if a  b, a, b ∈ Li .

Remark 1.24. Given a chain I, ≤ and complete residuated lattices Li which
satisfy the assumption of Defintion 1.23, we say that Li can be ordinally
added.

If we write “consider an ordinal sum i∈I Li ” or the like, we always
assume that Li are complete residuated lattices which can be ordinally added
and I, ≤ is known from the context.
 In order to show that Definition 1.23 is
correct, we need to show that i∈I Li is indeed a complete residuated lattice.

Lemma 1.25. An ordinal sum i∈I Li is a complete residuated lattice.

Proof. Clearly, L, ∧, ∨, 0, 1 is a complete lattice and L, ⊗, 1 is a commuta-


tive monoid. Thus, it suffices to check the adjointness property, i.e. a ⊗ b ≤ c
iff a ≤ b → c.
“⇒”: Let a ⊗ b ≤ c. If b → c = 1, then clearly a ≤ b → c. If b → c = c,
b ∈ Lj − Li , c ∈ Li − Lj , i < j, then since a ⊗ b ≤ c, we have a ∈ Lj and a < b.
This yields a = a ∧ b = a ⊗ b ≤ c = b → c. If b, c ∈ Li , b  c then, by (1.36),
b →i c is equal to the greatest element d ∈ Li such that d ⊗ b = d ⊗i b ≤ c,
thus a ≤ b →i c = b → c.
“⇐”: Let a ≤ b → c. If b → c = 1, then b ≤ c, which yields a ⊗ b ≤ c. If
b → c = c, we have a ≤ c, and thus a ⊗ b ≤ c. If b, c ∈ Li then b → c = b →i c.
Therefore, for a ∈ Li , we have a ∈ Lj for some j < i and so a ⊗ b ≤ c. For
a ∈ Li , we have a ⊗ b ≤ c using adjointness property of →i . 


Example 1.26. In particular, we may consider an ordinal sum L = L ⊕ 2 of


any complete residuated lattice L and a two-element Boolean algebra 2. That
is, we put I = {0, 1}, L0 = L, L1 = 2 = {1, 1}, ∧, ∨, ⊗, →, 1, 1 , and 1 is a
greatest element of L. Then we have
  
 a for b = 1 , 1 for a ≤L b ,
L L
a⊗ b= b for a = 1 , a→ b= b for a = 1 ,
 
a ⊗L b otherwise . a →L b otherwise .
Constructing L⊕2 corresponds to adding a new top element to L, see Fig. 1.6.
22 1 Introduction to Fuzzy Sets and Fuzzy Logic

1
1

Fig. 1.6. Complete residuated lattice with new top element

T-norms and Residuated Lattices

Examples 1.5 and 1.8 show that some examples of complete residuated lat-
tices can be obtained from so-called t-norms. In fact, these examples are very
important. In this section, we elaborate more on this topic. First, we show
that complete residuated lattices on [0, 1] with its genuine ordering correspond
to so-called left-continuous t-norms. Then, we show that continuous t-norms
correspond to those complete residuated lattices on [0, 1] which satisfy divis-
ibility. We conclude by listing some further properties of t-norms.
Definition 1.27. A t-norm is a binary operation ⊗ on the real unit interval
[0, 1] satisfying
(a ⊗ b) ⊗ c = a ⊗ (b ⊗ c) ,
a⊗b=b⊗a,
a⊗1=a,
a1 ≤ a2 , b1 ≤ b2 implies a1 ⊗ b1 ≤ a2 ⊗ b2 .
Algebraically speaking, ⊗ is a t-norm iff [0, 1], ⊗, 1, ≤ is a partially or-
dered monoid.
Definition 1.28. A t-norm ⊗ is called left-continuous if
 
lim (an ⊗ b) = lim an ⊗ b (1.49)
n→∞ n→∞

holds true for any non-decreasing sequence {an ∈ [0, 1] | n = 1, 2, 3, . . . }.


The reader might wonder how left-continuity of a t-norm from (1.49) re-
lates to the usual left-continuity of function of a real variable. We are now
going to present the relationship. Recall first the usual definition of left-
continuity.
Definition 1.29. A function g : [0, 1] → [0, 1] is called left-continuous at
c ∈ [0, 1] iff for each ε > 0 there exists δ > 0 such that for each a ∈ [0, c] such
that |a − c| < δ we have |g(a) − g(c)| < ε. g is called left-continuous (in [0, 1])
if g is left-continuous at c for each c ∈ [0, 1].
We will need the following lemma.
1.2 Structures of Truth Degrees 23

Lemma 1.30. A function g : [0, 1] → [0, 1] is left-continuous iff we have


 
lim g(an ) = g lim an (1.50)
n→∞ n→∞

for any non-decreasing sequence {an ∈ [0, 1] | n = 1, 2, 3, . . . }.


Proof. We make use of the following fact well-known from basic calculus.
Claim. g is left-continuous at c iff limn→∞ an = c implies limn→∞ g(an ) =
g(c) for each sequence {an | n = 1, 2, . . . } of elements an ∈ [0, c].
“⇒”: Take any non-decreasing sequence {an }n = {an ∈ [0, 1] | n = 1,
2, . . . }. Since {an }n is non-decreasing and bounded, it has a limit, say limn→∞
an = c. Since an ≤ c, left-continuity of g and Claim give limn→∞ g(an ) =
g(limn→∞ an ).
“⇐”: Consider any sequence {an }n of elements an ∈ [0, c] such that
limn→∞ an = c. By Claim, we need to show limn→∞ g(an ) = g(c). Suppose
by contradiction that limn→∞ g(an ) = g(c) is not the case. By definition of
a limit we have that (*) there exists ε > 0 such that for each n ∈ N there is
m ≥ n such that |g(am ) − g(c)| ≥ ε. Using (*) for n = 1, there exists m1 ≥ n
such that |g(am1 ) − g(c)| ≥ ε; denote this am1 by b1 . Using (*) for n = m1 + 1,
there exists m2 ≥ n such that |g(am2 ) − g(c)| ≥ ε; denote this am2 by b2 .
Repeating this over and over, we get a subsequence {bn }n of {an }n . Since
{bn }n is bounded (namely, bn ≤ c for each n), it contains a non-decreasing
subsequence {cn }n of elements cn ∈ [0, c]. Notice that for each cn we have
|g(cn ) − g(c)| ≥ ε . (1.51)
As {cn }n is a subsequence of {an }n , we have limn→∞ cn = limn→∞ an = c.
Applying assumption (1.50) to {cn }n we get limn→∞ g(cn ) = g(limn→∞ cn ) =
g(c) which is impossible due to (1.51). The proof is complete. 

Lemma 1.30 has the following corollary relating left-continuity of a t-norm
⊗ to left-continuity of ⊗ as a function of a real variable.
Corollary 1.31. For a t-norm ⊗, the following claims are equivalent.
(i) ⊗ is a left-continuous t-norm.
(ii) for each b ∈ [0, 1], function g(x) = x ⊗ b is left-continuous (in the sense
of Definition 1.29). 

Recall now that a function f (x, y) : [0, 1] × [0, 1] → [0, 1] is called non-
decreasing (non-increasing) in x if for any a1 , a2 , b ∈ [0, 1], a1 ≤ a2 implies
f (a1 , b) ≤ f (a2 , b) (f (a1 , b) ≥ f (a2 , b)). The notions non-decreasing (non-
increasing) in y are defined analogously. The following lemma translates alge-
braic properties of ⊗ and → on [0, 1] in terms of infima and suprema to their
continuity properties known from basic calculus.
Lemma 1.32. Let a function f (x, y) : [0, 1] × [0, 1] → [0, 1] be non-decreasing
in x. Then f is left-continuous in x iff for any {aj | j ∈ J} ⊆ [0, 1] and each
b ∈ [0, 1], we have
24 1 Introduction to Fuzzy Sets and Fuzzy Logic
 
f ( j∈J aj , b) = j∈J f (aj , b) . (1.52)
f is right-continuous in x iff for any {aj | j ∈ J} ⊆ [0, 1] and each b ∈ [0, 1],
we have
 
f ( j∈J aj , b) = j∈J f (aj , b) . (1.53)
Therefore, f is continuous in x iff both (1.52) and (1.53) hold true. We obtain
analogous statements if one replaces “in x” by “in y”; a dual statement holds
true for f non-increasing.

Proof. We prove the assertion concerning the left-continuity of a non-decrea-


sing function. The case of right-continuity is symmetric, the case of continuity
is clear, and for non-increasing f , the proof is symmetric.
Let fbe both non-decreasing and left-continuous in x. We  verify (1.52).
For a = {a j | j ∈ J}, there are two possibilities.
 Either a = ({aj | j ∈ J} −
{a}) or a > ({aj | j ∈ J} − {a}). If a = ({aj | j ∈ J} − {a}) then for each
n ∈ N there exists some aj(n) ∈ {aj | j ∈ J} − {a} such that a − aj(n) < n1 .
Clearly, we may safely assume aj(n) ≤ aj(n+1) . Then we have limn→∞ aj(n) =
a and aj(n) < a (n ∈ N). A moment’s reflection shows that

f ( j∈J aj , b) = f ( lim aj(n) , b) = lim f (aj(n) , b) =
n→∞ n→∞
 
= n∈N f (aj(n) , b) ≤ j∈J f (aj , b)

by the definition of left-continuity in x. If a > ({aj | j ∈ J} − {a}) then
there is some n ∈ N such that a − ak >n for each ak ∈ {aj | j∈ J} − {a}.
1

But then a = al for some l ∈ J,hence f ( j∈J aj , b) = f (al , b) ≤ j∈J f (aj , b).
The converse inequality, i.e. f ( j∈J aj , b) ≥ j∈J f (aj , b), holds since f is non-
decreasing in x.
Conversely, assume (1.52). By Lemma 1.30, to show that f is left-
continuous in x it suffices to show that for any a ∈ [0, 1] and any non-
decreasing sequence {an ∈ [0, 1] | n = 1, 2, . . . } such that limn→∞ an = a
we have limn→∞ f (an , b) = f (limn→∞ an , b). Using (1.52),this readily follows
from limn→∞ f (an , b) = n∈N f (an , b) and limn→∞ an = n∈N an . 


Theorem 1.33 (left-continuous t-norms and residuated lattices).


A binary operation ⊗ in [0, 1] is a left-continuous t-norm iff a structure
[0, 1], min, max, ⊗, →, 0, 1 with → defined by

a → b = {c | a ⊗ c ≤ b} , (1.54)
is a residuated lattice.

Proof. By Corollary
  and Lemma 1.32, ⊗ is a left-continuous t-norm
1.31  iff
we
 have a ⊗ i∈I b i = i∈I (a ⊗ bi ). By Theorem 1.22, we have a ⊗ i∈I b i =
i∈I (a⊗bi ) iff (1.54) makes [0, 1], min, max, ⊗, →, 0, 1 a complete residuated
lattice. 

1.2 Structures of Truth Degrees 25

Example 1.34. (1) An example of a left-continuous t-norm which is not con-


tinuous is “nilpotent minimum” [40, 60] T0 defined by

min(x, y) for x + y > 1
T0 (x, y) =
0 for x + y ≤ 1 .
(2) An example of a t-norm that is even not left-continuous is the “drastic
product” [60] defined by

x · y for max(x, y) = 1
Td (x, y) =
0 otherwise .
(3) A moment’s reflection shows that for each t-norm T we have
Td (x, y) ≤ T (x, y) ≤ min(x, y)
for every x, y ∈ [0, 1].
The following property of left-continuous t-norms will be used in the sequel.
Lemma 1.35. For a left-continuous t-norm ⊗, let a, b ∈ [0, 1].
(i) If a ≥ b ⊗ c for each c ∈ [0, 1) then a ≥ b.
(ii) If a < b then there is c ∈ [0, 1) such that a < b ⊗ c.

Proof. (i): If a ≥ b ⊗ c for each c = 1 then a ≥
 {b ⊗ c | c = 1} = b ⊗
{c | c = 1} = b ⊗ 1 = b.
(ii) is a restatement of (i). 

We now turn our attention to continuous t-norms and the corresponding
residuated lattices.
Definition 1.36. A t-norm ⊗ is called continuous if it is continuous as a
real function of two variables.
Lemma 1.37. Let f : [0, 1]2 → [0, 1] be a function such that for each a ∈ [0, 1],
f (x, a) : [0, 1] → [0, 1] and f (a, x) : [0, 1] → [0, 1] are continuous. Then f is
continuous.
Proof. We have to show that for every a, b ∈ [0, 1], for each ε > 0 there is
some δ > 0 such that |f (x, y) − f (a, b)| < ε for every x ∈ (a − δ, a + δ),
y ∈ (b − δ, b + δ) (if, e.g., x = 1 then “x ∈ (a − δ, a + δ)” is to be replaced by
“x ∈ (a − δ, a]”; similarly for the other boundary cases). Since for any A, B ∈
[0, 1], both f (A, x) and f (x, B) are continuous functions on a compact set
(closed interval), they are uniformly continuous. Therefore, there are δ1 , δ2 > 0
such that |f (x1 , B) − f (x2 , B)| < 2ε and |f (A, y1 ) − f (A, y2 )| < 2ε for any
x1 , x2 , y1 , y2 ∈ [0, 1] such that |x1 −x2 | < δ1 , |y1 −y2 | < δ2 . Put δ = min(δ1 , δ2 )
and take any x ∈ (a − δ, a + δ), y ∈ (b − δ, b + δ). We have
|f (x, y) − f (a, b)| = |f (x, y) − f (a, y) + f (a, y) − f (a, b)| ≤
≤ |f (x, y) − f (a, y)| + |f (a, y) − f (a, b)| < ε
2 + ε
2 =ε
completing the proof. 

26 1 Introduction to Fuzzy Sets and Fuzzy Logic

Due to Lemma 1.37, asserting that a t-norm is continuous (as a real func-
tion of two arguments) is tantamount to asserting that it is continuous in both
the first and the second variable.
Corollary 1.31 and Lemma 1.32 yield the following corollary.

Corollary 1.38. A left-continuous t-norm ⊗ is continuous iff it satisfies


 
a ⊗ i∈I bi = i∈I (a ⊗ bi ) . 


Lemma 1.39. A binary operation ⊗ in [0, 1] is a continuous t-norm iff the


structure [0, 1], min, max, ⊗, →, 0, 1 with → defined by (1.54) is a divisible
residuated lattice.

Proof. First, we show that a residuated lattice L satisfies divisibility iff for
each a ≤ b there exists c such that a = b ⊗ c. Denote this alternate condition
by (D). On the one hand, if L satisfies divisibility and a ≤ b, we have a =
b ∧ a = b ⊗ (b → a), i.e. one can take c = b → a in (D). On the other hand,
if L satisfies (D), then for any a, b there exists c such that a ∧ b = a ⊗ c. By
adjointness, c ≤ a → (a ∧ b) ≤ a → b. Therefore, a ∧ b = a ⊗ c ≤ a ⊗ (a → b).
Since a ⊗ (a → b) ≤ a ∧ b is always the case, (D) implies divisibility.
Now, let ⊗ be continuous. We verify (D). Let a, b ∈ [0, 1], a ≤ b and
consider a function f (z) : [0, 1] → [0, b] defined by f (z) = b ⊗ z. Since f is a
non-decreasing function and since 0 ≤ a ≤ b, continuity yields some c ∈ [0, 1]
such that a = f (c), i.e. a = b ⊗ c.
Conversely, let (D) hold. For an arbitrary b ∈ [0, 1], let f (z) = b ⊗ z. By
(D), f is a surjection from [0, 1] onto [0, b] and since it is non-decreasing, it is
continuous. The proof is complete. 


Theorem 1.40 (continuous t-norms and residuated lattices).


A binary operation ⊗ in [0, 1] is a continuous t-norm iff the structure
[0, 1], min, max, ⊗, →, 0, 1 with → defined by (1.54) is a BL-algebra.

Proof. The assertion follows from Lemma 1.39 since prelinearity is always
satisfied in a linearly ordered residuated lattice. 


The following definition summarizes further properties of t-norms which


we will need.

Definition 1.41. A t-norm ⊗ is called


(i) Archimedean iff for each a, b ∈ (0, 1) there exists n ∈ N such that an < b,
(ii) strictly monotone iff for each a > 0 and b < c we have a ⊗ b < a ⊗ c,
(iii) strict iff it is continuous and strictly monotone.

Continuous Archimedean t-norms can be characterized by a simple condi-


tion, see e.g. [47, 60].

Lemma 1.42. A continuous t-norm ⊗ is Archimedean iff we have a ⊗ a < a


for each a ∈ (0, 1).
1.2 Structures of Truth Degrees 27

Furthermore, continuous Archimedean t-norms can be represented using


so-called additive generators, see e.g. [47, 60].
Theorem 1.43 (additive generators). A mapping ⊗ : [0, 1]2 → [0, 1] is a
continuous Archimedean t-norm iff there is a continuous additive gener-
ator f such that
x ⊗ y = f (−1) (f (x) + f (y)) ,
i.e. f is a strictly decreasing continuous mapping f : [0, 1] → [0, ∞] with
f (1) = 0 and f (−1) is the pseudoinverse of f defined by f (−1) (x) = f −1 (x)
if x ≤ f (0) and f (−1) (x) = 0 otherwise.
Example 1.44. L ukasiewicz as well as product t-norms are both continuous
and Archimedean. f (x) = 1 − x and f (−1) (x) = max(1 − x, 0) are an additive
generator and its pseudoinverse of L ukasiewicz t-norm. f (x) = − log(x) and
f (−1) (x) = e−x are an additive generator and its pseudoinverse of product
t-norm. Min is continuous but not Archimedean.
The following theorem presents a well-known representation of continuous
t-norms by ordinal sums.
Theorem 1.45 (Mostert-Shields representation). Let ⊗ be a continuous
t-norm. Then the corresponding complete residuated lattice L on [0, 1] is iso-
morphic to an ordinal sum i∈I Li of complete residuated lattices such that
I is the set of all idempotents of ⊗, and each Li is a standard L  ukasiewicz
algebra on [0, 1], a standard product algebra on [0, 1], or a one-element algebra.
Proof. The proof is technically involved, see e.g. [47, 49, 60]. 

Remark 1.46. (1) The following claim is easy to check. Let I ⊆ [0, 1] be a
closed set (w.r.t. standard metric d on [0, 1] given by d(x, y) = |x − y|).
Consider the operation + w.r.t. I, ≤ with ≤ being a natural ordering of I
(see Definition 1.23). Let Li be a one-element algebra if i = i+ , and let Li be a
standard L  ukasiewicz algebra on [0, 1] or a standard product algebra on [0, 1]
if i < i+ . Then there is a continuous ⊗ t-norm such that the corresponding

complete residuated lattice on [0, 1] is isomorphic to the ordinal sum i∈I Li .
(2) Theorem 1.45 and (1) give a precise meaning of the claim that con-
tinuous t-norms are just ordinal sums of L  ukasiewicz t-norms and product
t-norms.

Properties of Truth Stressers

Theorem 1.47 (basic properties of truth stressers). Each truth stresser


on a complete residuated lattice satisfies
0∗ = 0 , (1.55)
monotony, i.e. a ≤ b implies a∗ ≤ b∗ , (1.56)

a∗ ⊗ b∗ ≤ (a ⊗ b) . (1.57)
28 1 Introduction to Fuzzy Sets and Fuzzy Logic

Proof. (1.55): By (1.12), 0∗ ≤ 0, i.e. 0∗ = 0.


(1.56): If a ≤ b then a → b = 1. Therefore, (1.11) and (1.13) imply 1 = 1∗ =

(a → b) ≤ a∗ → b∗ from which we get a∗ ≤ b∗ .

(1.57): By (1.22), a ≤ b → (a ⊗ b). Using (1.56) we get a∗ ≤ (b → (a ⊗ b)) .
∗ ∗ ∗ ∗ ∗
Furthermore, (b → (a ⊗ b)) ≤ b → (a ⊗ b) by (1.13). Therefore, a ≤ b →
∗ ∗
(a ⊗ b) from which we get a∗ ⊗ b∗ ≤ (a ⊗ b) by adjointness. 


Remark 1.48. Note that (1.12),(1.56), and(1.17) say that is an interior

operator on L, ≤ . Therefore, i∈I ai ∗ = ( i∈I ai ∗ ) .

We are now ready to prove that (1.16) defines a truth stresser.



Lemma 1.49. For each complete residuated lattice, defined by (1.16) is a
truth stresser.

Proof. (1.11) and (1.12) are obvious. For (1.13), distinguish the following
cases. First, if a = 1, then (1.13) becomes b∗ ≤ b∗ which is true. Second,

if 1 = a ≤ b, then a → b = 1 and so (1.13) says 1 = (a → b) ≤ a∗ → b∗ which
∗ ∗ ∗ ∗
is equivalent to a ≤ b . Now observe that a ≤ b is true both for b = 1 (since
a∗ ≤ 1 = b∗ ) as well as for b < 1 (by monotony of ⊗). Third, if 1 = a ≤ b = 1,
k k k
then (1.13) becomes i=1 ci ⊗ (a → b)ni ≤ i=1 ci ⊗ ani → i=1 ci ⊗ bni .
By adjointness, this inequalityk is true iff for each p, q = 1, . . . , k we have
cp ⊗ (a → b)np ⊗ cq ⊗ anq ≤ i=1 ci ⊗ bni which is true. Indeed, for np ≤ nq ,
cp ⊗ (a → b)np ⊗ cq ⊗ anq ≤ cp ⊗ (a → b)np ⊗ cq ⊗ anp ≤
k
≤ cp ⊗ cq ⊗ anp ⊗ (a → b)np ≤ cp ⊗ cq ⊗ bnp ≤ cp ⊗ bnp ≤ i=1 ci ⊗ bni ,
and analogously for np > nq . 


Theorem 1.50 (implicational truth stressers). Each implicational truth


stresser on a complete residuated lattice satisfies

(a ⊗ b) = a∗ ⊗ b∗ . (1.58)
∗ ∗
Proof. Due to (1.57), it is enough to see (a ⊗ b) ≤ a∗ ⊗b∗ . Since (a ⊗ b) ≤ a∗
∗ ∗ ∗
and (a ⊗ b) ≤ b∗ , (1.18) and monotony of ∗ yield (a ⊗ b) = (a ⊗ b) ⊗

(a ⊗ b) ≤ a∗ ⊗ b∗ . 


Remark 1.51. Looking at (1.58) one can ask under what conditions L∗ satisfies

(a → b) = a∗ → b∗ . It turns out that this condition is too restrictive. Indeed,

if ∗ is an implicational truth stresser satisfying (a → b) = a∗ → b∗ , we get

a ≤ (a → a∗ ) → a∗ ≤ (a → a∗ ) → a∗ = (a∗ → a∗∗ ) → a∗ = 1 → a∗ = a∗ , i.e.
a ≤ a∗ which means a = a∗ . Therefore, if ∗ is an implicational truth stresser

satisfying (a → b) = a∗ → b∗ , ∗ must be identity.

The following lemma shows how to obtain implicational truth stressers in


BL-chains (i.e. linearly ordered BL-algebras) which satisfy certain additional
condition.
1.2 Structures of Truth Degrees 29

Lemma 1.52. Let L be a complete BL-chain satisfying


 
a ⊗ i∈I bi = i∈I (a ⊗ bi ) . (1.59)

Then a unary operation defined by

a∗ = n∈N0 an (1.60)

for all a ∈ L is an implicational truth stresser. Moreover, a is the greatest
idempotent less or equal a.

Proof. (1.11) is obvious. It remains to check conditions (1.12), (1.13), (1.17)–


(1.19). 
(1.12): For n = 1 we have a1 ≤ a, thus a∗ = n∈N0 an ≤ a.
(1.13): From (1.44) and (1.35) it follows
∗  n 
(a → b) = n∈N0 (a → b) ≤ n∈N0 (an → bn ) ≤
 
≤ n∈N0 an → n∈N0 bn = a∗ → b∗ .
(1.17): Due to (1.59), we have
  
n m  m 
a∗∗ = m∈N0 n∈N0 a = m∈N0 i=1 ni ∈N0 ani =
  m   m
= m∈N0 n1 ,...,nm ∈N0 i=1 ani = m∈N0 n1 ,...,nm ∈N0 a i=1 ni =
  
= m∈N0 n∈N0 an = n∈N0 an = a∗ .
1.18: Analogously as for (1.17), we have
     
a ⊗ a∗ =

n∈N0 a
n
⊗ m∈N0 a
m
= m,n∈N0 an+m = k∈N0 ak = a∗ .
    n
(1.19): We have to show i∈I n∈N0 ani = n∈N0 i∈I ai . The “≥”-part
   n
is evident since for each i ∈ I we have n∈N0 ani ≥ n∈N0 i∈I ai . For the
“≤”-part, observe that since ani ≤ ai1 ⊗ · · · ⊗ ain for ai = min {ai1 , . . . , ain }
holds for each n (recall that L is supposed to be a chain), we have
     n
n∈N0 ai ≤ i∈I ai ≤ i1 ,...,in ∈I (ai1 ⊗ · · · ⊗ ain ) =
n n
i∈I i∈I ai .
Hence, the required inequality follows immediately.
Altogether, ∗ is an implicational truth stresser for a complete BL-chain
satisfying (1.59). It remains to show that a∗ is the greatest idempotent which
 or equal toa. Indeed, let b ∈ L be an idempotent such that b ≤ a. Then
is less
b = n∈N0 bn ≤ n∈N0 an = a∗ . 


Remark 1.53. (1) Since each BL-algebra satisfies a ⊗ (b ∧ c) = (a ⊗ b) ∧ (a ⊗


c) [49], every finite BL-chain satisfies (1.59).
(2) If L = [0, 1] , max, min, ⊗, →, 0, 1 is a residuated lattice with ⊗ being
a continuous t-norm, then L is a BL-algebra and (1.59) is a consequence of
right-continuity of ⊗.

Condition (1.21) of Horn truth stressers follows from two simpler condi-
tions.
30 1 Introduction to Fuzzy Sets and Fuzzy Logic

Lemma 1.54. If L∗ is a complete residuated lattice with a truth stresser ∗


such that
 
a → i∈I bi ≤ i∈I (a → bi ) for every non-empty index set I , (1.61)
 ∗  ∗
i∈I bi ≤ i∈I bi for every index set I , (1.62)
then L∗ satisfies (1.21).
Proof. Applying monotony of ∗ to (1.61) and using 1.62 we get
  ∗  ∗  ∗
a → i∈I bi ≤ i∈I (a → bi ) ≤ i∈I (a → bi ) ,
which is the non-trivial inequality of (1.21). 

Remark 1.55. (1) Observe that the converse inequality
 of (1.62) is always sat-
isfied. Indeed, for every i ∈ I, we have bi ≤ i∈I bi , so by monotony of ∗ we
 ∗   ∗
have b∗i ≤ i∈I bi . Thus, i∈I b∗i ≤ i∈I bi .
(2) Condition (1.62) does not hold for a general implicational truth
stresser. For instance, for L being a four-element Boolean algebra L =
{0, a, b, 1}, ∨, ∧, ⊗, →, 0, 1 (i.e. a, b ∈ L are incomparable), globalization ∗

is an implicational truth stresser. We have a ∨ b = 1, i.e. (a ∨ b) = 1. On the
∗ ∗ ∗ ∗
other hand a = b = 0, thus a ∨ b = 0.
In each residuated lattice, prelinearity is equivalent to a → (b ∨ c) =
(a → b) ∨ (a → c) [10, Theorem 2.34]. Therefore, we have the following
consequence of Lemma 1.54.
Corollary 1.56. Let L be a finite residuated lattice, which satisfies the con-
dition of prelinearity. Then every implicational truth stresser for L satisfying
(1.20) and (1.62) is a Horn truth stresser.
We now present two examples of implicational truth stressers satisfying
(1.20) and (1.62).
Example 1.57. (1) Let L be any complete Heyting algebra (i.e. a complete
residuated lattice, where ⊗ = ∧, the residuum is then a relative pseudo-
complement). The implicational truth stresser ∗ defined by a∗ = a, for all
a ∈ L, satisfies (1.20) and (1.62) trivially.
(2) Let L be an arbitrary complete residuated lattice equipped with glob-
alization. Evidently, ∗ satisfies (1.20) trivially because 0 ⊗ a = 0 ∧ a, and
1 ⊗ a = 1 ∧ a for all a ∈ L. Furthermore, ∗ satisfies (1.62) iff 1 (the greatest
element of L) is ∨-irreducible, i.e. iff for every family {bi < 1 | i ∈ I} we have

i∈I bi < 1.

Globalization on chains can be characterized by an additional identity.


Lemma 1.58 (globalization on chains). Let L be a linearly ordered com-
plete residuated lattice. A truth stresser ∗ is a globalization for L iff it satisfies
a∗ ∨ ¬a∗ = 1 . (1.63)
1.2 Structures of Truth Degrees 31

Proof. Evidently, globalization satisfies (1.63). Let ∗ satisfy (1.63). We have


1∗ = 1 due to (1.11). If a < 1 then a∗ ≤ a < 1. Since L is linearly ordered,
we have max(a∗ , ¬a∗ ) = a∗ ∨ ¬a∗ = 1, and so ¬a∗ = 1 from which we get
a∗ = 0 (indeed, if ¬a∗ = 1 then 1 ≤ ¬a∗ , i.e. 1 ≤ a∗ → 0, whence a∗ ≤ 0 by
adjointness). Therefore, ∗ is globalization. 


Remark 1.59. A characterization of truth stressers which are globalizations


for general complete residuated lattices by means of identities or inequalities
is not possible. First, note that each inequality can be equivalently replaced
by identity. Namely, x ≤ y can be replaced by x ∨ y = y. Now, if L∗11 and
L∗22 are two complete residuated lattices with globalizations ∗ 1 and ∗ 2 , their
direct product L∗ = L∗11 × L∗22 is a complete residuated lattice with a truth
stresser ∗ which is not a globalization since for 0, 1 ∈ L1 × L2 we have

0, 1 = 0∗1 , 1∗2 = 0, 1 = 0, 0 although 0, 1 < 1, 1 . On the other
hand, L∗11 × L∗22 satisfies each identity satisfied by both L∗11 and L∗22 . This
shows that globalizations cannot be characterized by identities.

Truth Stressers on Ordinal Sums

We now present some ways to define truth stressers on ordinal sums of com-
plete residuated lattices.

Theorem 1.60. Let I, ≤ be a finite chain with the least element 0 and the
greatest element 1. Let {L∗ii | i ∈ I} be a family of complete residuated lattices
Li = Li , ∨i , ∧i , ⊗i , →i , 0i , 1i with implicational truth stressers ∗i satisfying
(1.20) and (1.62).
 Suppose {Li | i ∈ I} can be ordinally added. Then the op-
eration ∗ on i∈I Li defined by a∗ = a∗i for a ∈ Li is an implicational truth
stresser on i∈I Li satisfying (1.20) and (1.62).

Proof. Clearly, conditions (1.11), (1.12), (1.17),


 and (1.18) follow directly from
the definition. Denote the operations on i∈I Li by ∧, ∨, ⊗, →, 0, 1.
(1.13): We consider three separate cases. If a ≤ b then clearly a∗ ≤ b∗ ,
∗ ∗
i.e. (a → b) = 1 = a∗ → b∗ . If a → b = b, then (a → b) = b∗ ≤ a∗ → b∗ .
Finally, when a  b and a → b = a →i b, i.e. a, b ∈ Li , then we have
∗ ∗
(a → b) = (a →i b) = (a →i b)∗i ≤ a∗i →i b∗i ≤ a∗ → b∗ .
(1.19): Take {aj | j ∈ J}.
Let k = min {i ∈ I | there is j ∈ J such that aj ∈ Li }. For J  = {j | aj ∈ Lk },
     ∗ k
we have j∈J aj = j∈J  aj . Thus, j∈J a∗j = j∈J  a∗j k = j∈J  aj =
 ∗
j∈J ja .
(1.20): If a, b∗ ∈ Li , the claim is trivial. If b∗ ∈ Li and a ∈ Li , we have
a ⊗ b∗ = a ∧ b∗ by definition.
1.62: For {bj | j ∈ J}, take l = max{i ∈ I | there is j ∈ J such that bj ∈
 ∗   ∗l
L }, put J  = {j | bj ∈ Ll }, and observe that j∈J bj = j∈J  bj =
i ∗l  ∗

j∈J j b = j∈J b j . 

32 1 Introduction to Fuzzy Sets and Fuzzy Logic

11 = 1
11
L1
1i = 0 1
1i
Li

0i
10
10

L0

00 = 0

Fig. 1.7. Ordinal sum of {Li ⊕ 2 | i ∈ I}

Lemma 1.61. For any complete residuated lattice L, globalization defined on


L ⊕ 2 (see Example 1.26) is an implicational truth stresser satisfying (1.20)
and (1.62).
Proof. As seen in Example 1.19, globalization is an implicational truth stresser
for any complete residuated lattice, thus in particular for L ⊕ 2. 

The above-described structures Li ⊕2 can be ordinally added, yielding new
structures with non-trivial implicational truth stressers satisfying (1.20) and
(1.62). This construction is illustrated in Fig. 1.7 and is justified by the next
assertion which is a direct consequence of Theorem 1.60 and Lemma 1.61.
Corollary 1.62. Let I, ≤ be a finite chain with the least element 0 and the
greatest element 1. Let {Li | i ∈ I} be a family of complete residuated lattices
Li = Li , ∨i , ∧i , ⊗i , →i , 0i , 1i , where L  i ∩ Lj = ∅ for all i, j ∈ I, i = j. More-
     
over, let
 Li = L i , ∨i , ∧ i , ⊗i , → i , 0i , 1 i be Li ⊕ 2. 
For the complete residuated
lattice i∈I Li we define a unary operation ∗ on i∈I Li by

∗ 1i for a = 1i ,
a = (1.64)
0i for a ∈ Li , a = 1i .

Then ∗ is an implicational truth stresser for 
i∈I Li satisfying (1.20) and
(1.62). 


1.3 Fuzzy Sets


Fuzzy Sets and Fuzzy Relations
The concepts of a fuzzy set and a fuzzy relation are the basic ones in fuzzy set
theory and its applications. Fuzzy sets and relations are mathematical models
1.3 Fuzzy Sets 33

of vaguely delineated collections of objects and relationships between objects.


Unlike the case of an ordinary set, an element may belong to a fuzzy set in a
degree different from 0 and 1. In general, the degree is taken from a suitable
scale L of truth degrees. We assume that L is a support set of some complete
residuated lattice L. For instance, if L = [0, 1], an element u may belong to a
fuzzy set A in degree, say 0.9. We denote this fact by A(u ) = 0.9. This may
be interpreted as “u almost belongs to A” since 0.9 is a high truth degree.
A(u ) = 1 means that u fully belongs to A, A(u ) = 0 means that u does not
belong to A at all. For example, we might say that a person who is 183 cm tall
belongs to a fuzzy set of tall people in degree 0.9 while a person 190 cm tall
belongs to that fuzzy set in degree 1. From this point of view, ordinary sets
correspond to fuzzy sets over the scale L = {0, 1} – each element either belongs
(i.e., belongs in degree 1) or does not belong (i.e., belongs in degree 0) to an
ordinary set. Fuzzy sets are mathematical objects. The aim of this section is
to introduce basic notions and facts concerning fuzzy sets.

Definition 1.63. An L-set (or fuzzy set with truth degrees in L) in a


universe set U is a mapping A : U → L.

If A is an L-set in U , A(u ) ∈ L is interpreted as the truth degree of “u


belongs to A”. Here and in the sequel, truth degrees (i.e. the elements of L)
will be denoted by a, b, c, . . . while elements of universes of L-sets will be
denoted by a , b , c , . . . , u , v , . . . , both possibly with indices.

Remark 1.64. (1) 2-sets coincide with characteristic functions of ordinary sets.
This is a basic way the concept of an L-set generalizes the concept of an
ordinary set.
(2) Note that a characteristic function χA of an ordinary set A ⊆ U also
can be seen as an L-set (for any L) such that for each element u we have
χA (u ) ∈ {0, 1}.

Example 1.65. (1) Let X = {circle, square, hexagon}, L be the standard Gödel
algebra on [0, 1]. Let A : X → [0, 1] be given by A(circle) = 1, A(square) =
0, A(hexagon) = 0.5. Then A is an L-set in X. A may be thought of as
representing the concept of “being circle-shaped” in the universe X.
(2) Let X be the set of all real numbers, L be the standard L  ukasiewicz
algebra on [0, 1]. Define an L-set A ∈ LX by

 x − 4 for 4 ≤ x ≤ 5
A(x) = 6 − x for 5 < x ≤ 6

0 otherwise .
Then A is a fuzzy set that represents the concept “approximately 5”. A is
depicted in Fig. 1.8.
(3) Both medical doctors (experts) and their patients (non-experts) un-
derstand and use the term “normal blood pressure”. The left part of Fig. 1.9
shows a fuzzy set which represents the meaning of this term. If one would like
34 1 Introduction to Fuzzy Sets and Fuzzy Logic

1 2 3 4 5 6 7
Fig. 1.8. Fuzzy set representing “approximately 5”

to represent the meaning of “normal blood pressure” by an ordinary set, one


would have to come up with a set like the one depicted by its membership
function in the right part of Fig. 1.9. That is, one would have to pick some
interval around a prototypical value of a normal blood pressure. Everybody
can see that this is not appropriate – a very small difference in blood pres-
sure can cause a transition from being normal to not being normal. This is
in contradiction to how experts and non-experts understand “normal blood
pressure”.

1 1

0 0
100 120 140 100 120 140
Fig. 1.9. Normal blood pressure as a fuzzy set (left) and as an ordinary set (right)

Note that A : U → L is sometimes called a membership function of a fuzzy


set. We do not follow this approach since it implies a distinction between a
fuzzy set and its membership function, and this causes unnecessary complica-
tions. For us, a fuzzy set is a function A : U → L (which might be occasionally
called a membership function).
For every L-set A : U → L, we define an ordinary set Supp(A) by
Supp(A) = {u | A(u ) > 0}. Supp(A) is called a support set of A. An L-set A in
U is said to be finite if Supp(A) is a finite set. A finite L-set A in U such that
Supp(A) = {u1 , . . . , un } is also denoted by A = {A(u1 )/u1 , . . . , A(un )/un }. An
L-set A : U → L such that Supp(A) = {u } is called a singleton. For every L-
set A : U → L and a ∈ L, we define an ordinary set aA by aA = {u | A(u ) ≥ a}.
a
A is called an a-cut of A.
1.3 Fuzzy Sets 35

The set of all L-sets in U is denoted by LU or LU . An empty L-set in U ,


denoted by ∅U , is a mapping ∅U : U → L where ∅U (u ) = 0 for all u ∈ U . If
the universe of discourse U is clear from context, we denote ∅U by ∅. A full
L-set in U , denoted by 1U , is a mapping 1U : U → L, where 1U (u ) = 1 for
every u ∈ U . If there is no danger of confusion, we write U instead of 1U .
An L-set A in U is called crisp if A(u ) ∈ {0, 1} for each u ∈ U . Crisp
L-sets in U correspond in an obvious way to ordinary subsets of U (crisp
L-sets are the characteristic functions of ordinary subsets). Therefore, we will
sometimes identify crisp L-sets with the corresponding ordinary subsets.
An n-ary L-relation (or fuzzy relation with truth degrees in L) in a universe
set U is an L-set in the universe set U n . That is, a binary L-relation R in U
is a mapping R : U × U → L. A binary L-relation R : U × U → L is called
a restriction of R : U × U → L, if R (u , v ) > 0 implies R (u , v ) = R(u , v )
for every elements u , v ∈ U . If a restriction R of R is a finite L-relation, it is
called a finite restriction of R.
For a binary L-relation R, we define an inverse L-relation R−1 of R by
putting R−1 (u , v ) = R(v , u ) for all u , v ∈ U . For binary L-relations R1 , R2
in U , the ◦-composition of R1 and R2 is a binary L-relation on U defined by

(R1 ◦ R2 )(u , v ) = w ∈ U (R1 (u , w ) ⊗ R2 (w , v )) . (1.65)
for all u , v ∈ U .
For L-sets A and B in U we define
  
S(A, B) = u ∈ U A(u ) → B(u ) , (1.66)
  
A ≈ B = u ∈ U A(u ) ↔ B(u ) . (1.67)
S(A, B) is called a degree of subsethood of A in B, and A ≈ B is called a
degree of equality of A and B. Note that S(A, B) can be thought of as a truth
degree of “for each u ∈ U : if u belongs to A then u belongs to B.” Likewise,
A ≈ B can be thought of as a truth degree of “for each u ∈ U : u belongs to
A iff u belongs to B.”
Example 1.66. For standard L
 ukasiewicz, Gödel, and product structures on
[0, 1], we have
S(A, B) = inf{1 − A(u ) + B(u ) | u ∈ U, A(u ) > B(u )} (Lukasiewicz) ,
S(A, B) = inf{B(u ) | u ∈ U, A(u ) > B(u )} (Gödel) ,
S(A, B) = inf{B(u )/A(u ) | u ∈ U, A(u ) > B(u )} (product) ,
and
A ≈ B = inf{1 − |A(u ) − B(u )| | u ∈ U } (Lukasiewicz) ,
A ≈ B = inf{min(A(u ), B(u )) | u ∈ U, A(u ) = B(u )} (Gödel) ,


 0 if for some u ∈ U :

 A(u ) = 0 = B(u ) or
A≈B= A( u ) = 0 = B(u ) (Goguen) .

  

 inf u ∈U min A( u ) B( u )
B(u ), A(u ) otherwise
36 1 Introduction to Fuzzy Sets and Fuzzy Logic

From the properties of residuated lattices it follows that A ≈ B =


S(A, B) ∧ S(B, A). Furthermore, we write A ⊆ B (A is a subset of B)
if S(A, B) = 1. Since S(A, B) = 1 means u ∈U (A(u ) → B(u )) = 1,
S(A, B) = 1 is equivalent to A(u ) → B(u ) = 1 for each u ∈ U which is,
due to (1.23), equivalent to A(u ) ≤ B(u ) for each u ∈ U . That is, A ⊆ B
means that for each u ∈ U we have A(u ) ≤ B(u ).
Common operations with L-sets which generalize the ordinary operations
with sets result by componentwise extension of operations on L. That is, any
operation o (possibly infinitary) on L induces in a componentwise manner
an operation O on LU by putting O(A, . . . )(u ) = o(A(u ), . . . ) for arbitrary
L-sets A, . . . and arbitrary u ∈ U . Arguments of O are L-sets in U and the
result of O applied to A, . . . is an L-set O(A, . . . ) in U to which an element
u ∈ U belongs to a degree obtained by applying o to A(u ), . . . In particular,
for L-sets A, B, Ai (i ∈ I) in U we define
(A ∧ B)(u ) = A(u ) ∧ B(u ) ,
(A ∨ B)(u ) = A(u ) ∨ B(u ) ,
  
i∈I Ai (u ) = i∈I Ai (u ) ,
  
i∈I Ai (u ) = i∈I Ai (u ) ,
 
 A ∧ B, A ∨ B, i∈I Ai , and i∈I Ai
etc. Following commonusage we denote
also by A ∩ B, A ∪ B, i∈I Ai , and i∈I Ai , respectively.
 
The componentwise extensions make LU = LU , ∧, ∨,Q⊗, →, ∅U , 1U a
complete residuated lattice. Namely, LU is a direct product u ∈U Lu of com-
plete residuated lattices Lu = L.
Theorem 1.67. For any complete residuated lattice L and any set U =  ∅,
LU isa complete residuated lattice where infima and suprema are given by
and , respectively, and the corresponding lattice order is ⊆. Moreover, all
identities and inequalities satisfied by L are valid also in LU . 

The following proposition summarizes the basic properties of the degree
of subsethood.
Theorem 1.68. For L-sets A, B, and C in U , we have
S(A, A) = 1 , (1.68)
S(A, B) ⊗ S(B, C) ≤ S(A, C) , (1.69)
S(A, B) = 1 iff A ⊆ B , (1.70)
S(∅, A) = 1 and S(A, U ) = 1 , (1.71)
 
S(A, i∈I Bi ) = i∈I S(A, Bi ) , (1.72)
 
S( i∈I Ai , B) = i∈I S(Ai , B) , (1.73)
S(A, A ∩ B) = S(A, B) = S(A ∪ B, B) . (1.74)
Proof. We will  1.67 and the fact that if au ≤ bu (au , bu ∈ L,
 use Theorem
u ∈ U ) then u ∈U au ≤ u ∈U bu .
1.3 Fuzzy Sets 37
 
(1.68): S(A, A) = u ∈U (A(u )→ A(u )) = u ∈U 1 = 1.
(1.69):
 S(A, B) ⊗ S(B, C) = u ∈U (A(u ) → B(u )) ⊗  v ∈U (B(v ) → C(v )) ≤
u ∈U v ∈U ((A(u ) → B( u )) ⊗ (B(v ) → C(v ))) ≤ u ∈U ((A(u ) → B(u )) ⊗
(B(u ) → C(u ))) ≤ u ∈U (A(u ) → C(u )) = S(A, C).
(1.70): If S(A, B) = 1 iff for each u ∈ U we have A(u ) → B(u ) = 1 iff for
each u ∈ U we have  A(u ) ≤ B(u ) iff A ⊆ B. 
(1.71): S(∅, A) = u ∈U (0→ A(u )) = 1 and S(A, U ) = u∈U (A(u ) → 1) = 1.
(1.72): S(A, i∈I Bi ) = u ∈U (A(u ) →  i∈I Bi (u )) = u ∈U i∈I (A(u ) →
Bi (u )) =  i∈I u ∈U (A(u ) → Bi (u )) = i∈I S(A, Bi ).  
(1.73): S( i∈I Ai , B) = u ∈U ( i∈I Ai ( u ) → B(u )) = u ∈U i∈I (Ai (u ) →
B(u )) = i∈I u ∈U (Ai (u ) → B(u )) = i∈I S(Ai , B)
(1.74): By (1.68), (1.72), and (1.73), S(A, A ∩ B) = S(A, A) ∧ S(A, B) =
1 ∧ S(A, B) = S(A, B), and S(A ∪ B, B) = S(A, B) ∧ S(B, B) = S(A, B) ∧ 1 =
S(A, B). 


Analogous assertions can be proved for the degree of equality. Notice how
the assertions of Theorem 1.68 can be described verbally. For instance, take
(1.69). If L = 2 (bivalent case), (1.69) expresses exactly that if a set A is a
subset of B, and B is a subset of C, then A is a subset of C. For a general L,
the meaning of (1.69) is the same, only interpreted in many-valued setting.
The same is true of other assertions.

Fuzzy Equivalences and Fuzzy Equalities

The concept of a fuzzy equivalence relation results by carrying over the con-
cept of an ordinary equivalence relation to fuzzy setting. Fuzzy equivalences
belong to the most studied fuzzy relations. The reason for this is that a fuzzy
equivalence can be interpreted as similarity – a degree to which two object
are equivalent is understood as a degree of their similarity. Similarity is one of
crucial phenomena accompanying human reasoning and perception. With the
concept of fuzzy equivalence, fuzzy logic provides a simple model of similarity.

Definition 1.69. An L-equivalence (fuzzy equivalence) relation θ on a


set U is a mapping θ : U × U → L satisfying
θ(u , u ) = 1 , (1.75)
θ(u , v ) = θ(v , u ) , (1.76)
θ(u , v ) ⊗ θ(v , w ) ≤ θ(u , w ) , (1.77)
for every u , v , w ∈ U . An L-equivalence θ on U where
θ(u , v ) = 1 implies u=v (1.78)
will be called an L-equality (fuzzy equality).

Remark 1.70. (1) There are various other terms used for L-equivalences: the
most common are similarity (or fuzzy similarity) and indistinguishability.
38 1 Introduction to Fuzzy Sets and Fuzzy Logic

These terms, however, suggest an epistemic interpretation that may be dis-


putable.
(2) Conditions (1.75), (1.76), and (1.77) are called reflexivity, symme-
try, and transitivity, respectively. These conditions generalize their bivalent
counterparts. Indeed, consider L = 2. Then (1.75) and (1.76) obviously ex-
press reflexivity and symmetry (in bivalent sense). Equation (1.77) implies
that if u , v and v , w are in θ (i.e. θ(u , v ) = 1 and θ(v , w ) = 1) then
1 = 1 ⊗ 1 = θ(u , v ) ⊗ θ(v , w ) ≤ θ(u , v ). Therefore, θ(u , w ) = 1, i.e. u , w is
in θ, i.e. θ is transitive (in bivalent sense). Conversely, let θ be transitive (in
bivalent sense). There are two possibilities: Either θ(u , v ) = 0 or θ(v , w ) = 0
and then (1.77) is true; or both θ(u , v ) = 1 and θ(v , w ) = 1 in which case
u , v and v , w are in θ, and so also u , w is in θ, i.e. θ(u , w ) = 1, i.e.
(1.77) holds true.
(3) Equations (1.75)–(1.77) are equivalent to saying that first-order logical
formulas which express reflexivity, symmetry, and transitivity, are true in
degree 1. Note that denoting conjunction and implication by o and i, the
formulas are
(∀x)(x ∼ x) ,
(∀x, y)(x ∼ y i y ∼ x) ,
(∀x, y, z)(((x ∼ y) o (y ∼ z)) i x ∼ z) .
(4) There has been a lot of debates of whether reflexivity, symmetry, and
transitivity as described verbally are appropriate properties of similarity, see
e.g. [82, 86]. By verbal description we mean e.g. “if u and v are similar then v
and u are similar” in case of similarity. Reflexivity seems to be mostly agreed
upon. There is an argument against symmetry. It is being pointed out that
“u is similar to v ” is not the same as “v is similar to u ” (e.g. u being a father
and v being his son). However, if one is interested in relation describing “u
and v are similar to each other”, symmetry seems to be an obvious property.
Transitivity of similarity has been a point of disagreement. One usually argues
against transitivity as follows. If similarity were transitive then any two colors
would be similar. For we may suppose that two colors with sufficiently close
wave lengths are similar. Now, for any two colors u and v we may find a
chain u = u1 , u2 , . . . , un = v , of colors such that ui and ui+1 are similar.
Using transitivity, u and v are similar. On the other hand, if the transitivity
condition is formulated verbally (i.e. “if u and v are similar, and if v and w
are similar then u and w are similar”), it seems acceptable. The solution to
this puzzle lies in the fact that similarity, by its nature, is a graded (fuzzy)
notion. If we look at the meaning of transitivity in fuzzy setting, we find it
quite natural. For example, if θ(u , v ) = 0.8 (u and v are similar in degree 0.8)
and θ(v , w ) = 0.8 (v and w are similar in degree 0.8) then u and w have to be
similar at least in degree 0.8 ⊗ 0.8. Thus, in case of the product conjunction,
transitivity forces θ(u , w ) ≥ 0.8 ⊗ 0.8 = 0.64 which seems to reflect intuition
in a reasonable way.
1.3 Fuzzy Sets 39

(5) L-equivalences will usually be denoted by θ. L-equalities will usually


be denoted by ≈ and we will use an infix notation, i.e. we write u ≈ v instead
of ≈ (u , v ). In order to facilitate reading, we use also (u ≈ v ) instead of u ≈ v .
Example 1.71. (1) Figure 1.10 describes an L-equivalence relation θ in U =
{u1 , u2 , u3 , u4 } with L being the standard L
 ukasiewicz algebra on [0, 1]. That

Fig. 1.10. L-equivalence θ

is, θ(u1 , u1 ) = 1, θ(u1 , u2 ) = 0.9, θ(u1 , u3 ) = 0.8, etc. Note that θ is not an
L-equality. Indeed, θ(u1 , u4 ) = 1 but u1 = u4 . On the other hand, restriction
of θ to U − {u4 } is an L-equality on {u1 , u2 , u3 }.
(2) If L is the standard Gödel algebra on [0, 1] (i.e. a⊗b = min(a, b)), θ from
Fig. 1.10 is not an L-equivalence. Indeed, 0.9 ⊗ 0.8 = θ(u2 , u1 ) ⊗ θ(u1 , u3 ) =
min(0.9, 0.8) ≤ θ(u2 , u3 ) = 0.7, and so θ is not transitive.
We now list basic properties of fuzzy equivalences.
Theorem 1.72. Let θ be an L-equivalence in U .
(i) For each a ∈ L, a θ is a reflexive and symmetric relation. If a ⊗ a = a then
a
θ is an equivalence relation.
(ii) 1 θ is an equivalence relation. If θ is an L-equality then 1 θ is the identity
in U .
Proof. (i): Reflexivity and symmetry of a θ is obvious. Let a⊗a = a. If u , v ∈
a
θ and v , w ∈ a θ then a ≤ θ(u , v ) and a ≤ θ(v , w ) and thus a = a ⊗ a ≤
θ(u , v ) ⊗ θ(v , w ) ≤ θ(u , w ), i.e. u , w ∈ a θ and so a θ is also transitive.
(ii): Since 1 ⊗ 1 = 1, the fact that 1 θ is an L-equivalence follows from (i).
If θ is an L-equality then 1 θ is identity in U by definition. 

∗ ∗ ∗
In classical setting, equivalence relations correspond uniquely to parti-
tions, see Sect. 1.1. We are going to show that an analogous correspondence
is available also in fuzzy setting.
For an L-equivalence θ in U and u ∈ U , denote by [u ]θ an L-set in U
defined by
[u ]θ (v ) = θ(u , v ) (1.79)
for any v ∈ U . We call [u ]θ a class of θ given by u . Therefore, a degree to
which v belongs to a class given by u is the degree to which u and v are
equivalent.
40 1 Introduction to Fuzzy Sets and Fuzzy Logic

Definition 1.73. An L-partition of a set U is a set Π of L-sets in U such


that
(i) for each u ∈ U there is A ∈ Π such that A(u ) = 1;
(ii) for each A ∈ Π there is u ∈ U such that A(u ) = 1;
(iii) for every A, B ∈ Π, u ∈ U we have A(u ) ⊗ B(u ) ≤ (A ≈ B).

Remark 1.74. (1) For L = 2, the concept of an L-partition coincides with the
concept of a partition. (i) says that Π covers U ; (ii) says that each A ∈ Π is
non-empty; (iii) says that if u belongs to A and B then A and B are equal.
(2) Since A ≈ B denotes the equality degree defined by (1.67), (iii) implies
that if A(u ) = 1 and B(u ) = 1 for some A, B ∈ Π, then A = B.

Theorem 1.75 (fuzzy equivalences and partitions). Let θ be an L-


equivalence in U , Π be an L-partition of U . Define a binary L-relation θΠ in
U by
θΠ (u , v ) = Au (v )
where Au ∈ Π is such that Au (u ) = 1, and put Πθ = {[u ]θ | u ∈ U }.
Then (1) Πθ is an L-partition of U ; (2) θΠ is an L-equivalence in U ; and (3)
θ = θΠθ and Π = ΠθΠ .

Proof. (1): We verify that Πθ is an L-partition of U . By reflexivity of θ,


[u ]θ (u ) = 1 verifying (i) and (ii). To verify (iii) we have to show that for
every u , v , w ∈ U we have [u ]θ (w ) ⊗ [v ]θ (w ) ≤ ([u ]θ ≈ [v ]θ ) which holds iff
for each z ∈ U we have both
[u ]θ (w ) ⊗ [v ]θ (w ) ≤ [u ]θ (z ) → [v ]θ (z )
and
[u ]θ (w ) ⊗ [v ]θ (w ) ≤ [v ]θ (z ) → [u ]θ (z ) .
Due to symmetry of both of the cases we check only the first inequality which
is equivalent to
[u ]θ (z ) ⊗ [u ]θ (w ) ⊗ [v ]θ (w ) ≤ [v ]θ (z ) ,
i.e. to
θ(z , u ) ⊗ θ(u , w ) ⊗ θ(w , v ) ≤ θ(z , v ) ,
which is true by transitivity of θ. Therefore, (1) is proved.
(2): We verify that θΠ is an L-equivalence in U . First, notice that θΠ is
defined correctly. Indeed, by Remark 1.74 (2), there is just one Au ∈ Π such
that Au (u ) = 1. Reflexivity of θΠ is obvious. Symmetry: We have to show
that if A(u ) = 1 and B(v ) = 1 for u , v ∈ U and A, B ∈ Π, then A(v ) = B(u ).
By definition we have A(v ) ⊗ B(v ) ≤ (A ≈ B), hence A(v ) = A(v ) ⊗ 1 =
A(v ) ⊗ B(v ) ≤ A(u ) → B(u ) = 1 → B(u ) = B(u ), i.e. A(v ) ≤ B(u ).
Similarly, one gets B(u ) ≤ A(v ), proving the required equality. Transitivity:
We have to show Au (v ) ⊗ Av (w ) ≤ Au (w ), i.e. Au (v ) ≤ Av (w ) → Au (w ).
1.3 Fuzzy Sets 41

The last inequality is true since Au (v ) = Au (v ) ⊗ Av (v ) ≤ (Au ≈ Av ) ≤


Av (w ) → Au (w ).
(3): We show θ = θΠθ : θΠθ (u , v ) = A(v ) for A ∈ Πθ such that A(u ) = 1.
Since A = [u ]θ , we have θΠθ (u , v ) = [u ]θ (v ) = θ(u , v ). We show Π = ΠθΠ :
We have A ∈ ΠθΠ iff A = [u ]θΠ . Now, for Au ∈ Π such that Au (u ) = 1 we
have A(v ) = [u ]θΠ (v ) = θΠ (u , v ) = Au (v ). Thus, A = Au ∈ Π. 


∗ ∗ ∗
Similarity of objects is often related to a collection of attributes of these
objects. Two objects are considered similar if they are similar according to
having some attributes. The following assertion says that fuzzy equivalences
are just fuzzy relations which result by the criterion of having the same at-
tributes.

Theorem 1.76 (Leibniz equivalence). For each system S ⊆ LU of L-sets


in U , the L-relation θS in U defined by

θS (u , v ) = A∈S (A(u ) ↔ A(v )) (1.80)
is an L-equivalence in U . For each L-equivalence θ in U , there is some S ⊆ LU
such that θ = θS . Moreover, θS is an L-equality iff for every distinct u and v
there is some A ∈ S such that A(u ) = A(v ).

Proof. First, we show that θS isan L-equivalence. Reflexivity: Since A(u ) ↔


A(u ) = 1, we have θS (u , u ) = A∈S (A(u ) ↔ A(u )) = 1. Symmetry follows
from A(u ) ↔ A(v ) = A(v ) ↔ A(u ). Transitivity:
θS (u , v ) ⊗ θS (v , w ) =
 
= A∈S (A(u ) ↔ A(v )) ⊗ B∈S (B(v ) ↔ B(w )) ≤
 
≤ ((A( u ) ↔ A(v )) ⊗ (B(v ) ↔ B(w ))) ≤
A∈S B∈S
≤ ((A(u ) ↔ A(v )) ⊗ (A(v ) ↔ A(w ))) ≤
A∈S
≤ A∈S (A(u ) ↔ A(w )) = θS (u , w ) .

Conversely, let θ be an L-equivalence in U . Put S = Πθ , i.e. S =


{[u ]θ | u ∈ U }. Due to Theorem 1.75,  it suffices to show that θS (u , v ) =
[u ]θ (v ). On the one hand, θS (u , v ) = w ∈U ([w ]θ (u ) ↔ [w ]θ (v )) ≤ [u ]θ (u ) ↔
[u ]θ (v ) = [u ]θ (v ). On the other hand, [u ]θ (v ) ≤ θS (u , v ) iff for each w ∈ U we
have [u ]θ (v ) ≤ [w ]θ (u ) ↔ [w ]θ (v ). Due to symmetry, we verify only [u ]θ (v ) ≤
[w ]θ (u ) → [w ]θ (v ) which is equivalent to [w ]θ (u ) ⊗ [u ]θ (v ) ≤ [w ]θ (v ) which
follows from transitivity: [w ]θ (u ) ⊗ [u ]θ (v ) = θ(w , u ) ⊗ θ(u , v ) ≤ θ(w , v ) =
[w ]θ (v ).
Finally, let u and v be distinct (i.e. u = v ). Then θS (u , v ) = 1 iff there is
some A ∈ S such that A(u ) ↔ A(v ) = 1, i.e. A(u ) = A(v ). This proves the
last claim. 


Remark 1.77. (1) The set S from Theorem 1.76 can be considered a set of
attributes. Each A ∈ S can be thought of as a (fuzzy) attribute, A(u ) being
42 1 Introduction to Fuzzy Sets and Fuzzy Logic

a degree to which u ∈ U has attribute A. Then θS (u , v ) is a truth degree of


“for each attribute A: u has A iff v has A” or, more succinctly, a truth degree
of “u and v have the same attributes (of S)”. Having the same attributes
has been considered by Leibniz [3] a criterion for being the same. Therefore,
(1.80) can be considered a fuzzy counterpart of Leibniz definition.
(2) Equality degree ≈ defined by (1.67) is an L-relation in LU . It is easy
to see that ≈ is an L-equality. This fact can be proved directly but it also
follows from Theorem 1.76. Namely, having a set V , put U = LV , S = {Av ∈
LU | v ∈ V, Av (B) = B(v) for B ∈ U }. Then by Theorem 1.76, θS is just ≈
defined by (1.67) and it is an L-equality in LV .

∗ ∗ ∗
Next, we recall a relationship between fuzzy equivalences and generalized
pseudometrics. A generalized pseudometric on a non-empty set U is a mapping
δ : U × U → [0, ∞] satisfying
δ(u , u ) = 0 ,
δ(u , v ) = δ(v , u ) ,
δ(u , w ) ≤ δ(u , v ) + δ(v , w ) .
If, in addition to the first condition, we require that δ(u , v ) = 0 implies u = v ,
we speak of a generalized metric. A metric (pseudometric) is a generalized
metric (pseudometric) δ such that δ(u , v ) ∈ [0, ∞), i.e. δ(u , v ) < ∞, for each
u , v ∈ U . Note that a generalized (pseudo)metric is sometimes called simply
a (pseudo)metric.

Theorem 1.78 (fuzzy equivalences vs. pseudometrics). Let ⊗ be a con-


tinuous Archimedean t-norm with an additive generator f , L be a residuated
lattice on [0, 1] given by ⊗, ≈ be an L-equivalence on X, δ be a generalized
pseudometric on U . Then
(i) δ≈ : [0, 1]2 → [0, ∞] defined by
δ≈ (u , v ) = f (u ≈ v ) (1.81)
is a generalized pseudometric which is a generalized metric iff ≈ is an
L-equality;
(ii) ≈δ : [0, 1]2 → [0, 1] defined by
(u ≈δ v ) = f (−1) (δ(u , v )) (1.82)
is an L-equivalence on U which is an L-equality iff δ is a generalized
metric;
(iii) ≈ equals ≈δ≈ and if δ(U, U ) ⊆ [0, f (0)] then δ equals δ≈δ .

Proof. (i): δ≈ (u , u ) = 0 and δ≈ (u , v ) = δ≈ (v , u ) follow from (u ≈ u ) = 1,


(u ≈ v ) = (v ≈ u ), and f (1) = 0. Triangle inequality for δ≈ can be obtained
as follows: Transitivity of ≈ yields f (−1) (f (u ≈ v ) + f (v ≈ w )) ≤ (u ≈ w ).
1.3 Fuzzy Sets 43

Since f is decreasing, we have f (u ≈ w ) ≤ f (f (−1) (f (u ≈ v ) + f (v ≈ w ))).


Now, there are two possibilities: either f (u ≈ v ) + f (v ≈ w ) > f (0) and then
f (u ≈ w ) ≤ f (f (−1) (f (u ≈ v ) + f (v ≈ w ))) = f (0) < f (u ≈ v ) + f (v ≈ w ),
or f (u ≈ v ) + f (v ≈ w ) ≤ f (0) and then f (u ≈ w ) ≤ f (f (−1) (f (u ≈
v ) + f (v ≈ w ))) = f (u ≈ v ) + f (v ≈ w ). In both of the cases we have f (u ≈
w ) ≤ f (u ≈ v ) + f (v ≈ w ) which means δ≈ (u , w ) ≤ δ≈ (u , v ) + δ≈ (v , w ),
i.e. the required triangle inequality. That δ≈ is a metric metric iff ≈ is an
L-equality follows easily from the fact that f is strictly decreasing.
(ii): (u ≈δ u ) = 1 and (u ≈δ v ) = (v ≈δ u ) follow from δ(u , u ) =
0, δ(u , v ) = δ(v , u ), and f (−1) (0) = 1. Transitivity of ≈δ : Note first that
f (−1) (u ) = f −1 (min(f (0), u )), f (u ≈δ v ) = min(f (0), δ(u , v )), and that f −1
is decreasing. Using triangle inequality for δ, we have
(u ≈δ w ) = f (−1) (δ(u , w )) = f −1 (min(f (0), δ(u , w ))) ≥
≥ f −1 (min(f (0), δ(u , v ) + δ(v , w ))) =
= f −1 (min(f (0), f (u ≈δ v ) + f (v ≈δ w ))) = (u ≈ v ) ⊗ (v ≈ w ) ,
verifying transitivity of ≈δ . Since f (−1) is strictly decreasing, ≈δ is an L-
equality iff δ is a generalized metric.
(iii): (u ≈δ≈ v ) = f (−1) (δ≈ (u , v )) = f (−1) (f (u ≈ v )) = (u ≈ v ).
If δ(u , v ) ≤ f (0) then δ≈δ (u , v ) = f (u ≈δ v ) = f (f (−1) (δ(u , v ))) =
f (f −1 (min(f (0), δ(u , v )))) = f (f −1 (δ(u , v ))) = δ(u , v ). 


Sets with Fuzzy Equalities

In the ordinary setting (of naive set theory), a set U is always considered to
be available with the identity idU in U , though idU is mostly not explicitly
mentioned. Intuitively, idU provides trivial information about which elements
are distinct and which are not and this information is prior to the concept of
a set. Nevertheless, identity is used in formulation of various properties. For
instance, the condition of antisymmetry of a partial order ≤ in U says that
for each u , v ∈ U , if u ≤ v and v ≤ u then u equals v , i.e. u , v ∈ idU .
The concept of a fuzzy equality is an extension of the concept of ordinary
identity. First, an L-equality ≈ in U carries information about the ordinary
identity in U since, due to (1.75) and (1.78), we have u = v iff (u ≈ v ) =
1. Second, ≈ can be thought of as carrying information about similarity of
elements of U since we may have (u ≈ v ) > 0 even if u = v . Note, however,
that we can always have an ordinary identity in fuzzy setting as well since an
L-relation ≈ in U defined by

1 for u = v ,
(u ≈ v ) =
0 for u = v ,
is an L-equality. That is, an L-equality in U can be thought of as providing
information about a kind of underlying similarity of elements of U such that
being similar in degree 1 means being equal. The underlying similarity needs
44 1 Introduction to Fuzzy Sets and Fuzzy Logic

to be taken into account when considering additional functions and relations


in U . Taking the underlying L-equality ≈ into account means that additional
functions and (fuzzy) relations in U should be, in a sense, compatible with
≈. For functions, compatibility means that for any n-ary function f in U ,
if u1 is similar to v1 and · · · and un is similar to vn , then f (u1 , . . . , un ) is
similar to f (v1 , . . . , vn ). In words, similar arguments are mapped to similar
results. For (fuzzy) relations, compatibility means that for any n-ary (fuzzy)
relation r in U , if u1 is similar to v1 and · · · and un is similar to vn , and
if u1 , . . . , un are in r, then v1 , . . . , vn are in r. In words, similar arguments
cannot be distinguished by being or being not in a relation.
Therefore, in analogy to ordinary setting, where we start by a set U and
consider functions and relations in U , in fuzzy setting, a natural situation is
the following. We start by a set U equipped by a fuzzy equality relation ≈.
Fuzzy equality ≈ imposes the above-described compatibility constraints on
additionally considered functions and (fuzzy) relations in U . This leads to the
following definitions.

Definition 1.79. A set with L-equality (set with fuzzy equality ) is a


pair U, ≈ where ≈ is an L-equality in U .

Compatibility of functions and relations with an L-equality is captured by


the following definition.

Definition 1.80. An n-ary function f : U n → U in U is said to be compat-


ible with an L-equality ≈ in U if for arbitrary u1 , v1 , . . . , un , vn ∈ U we have
(u1 ≈ v1 ) ⊗ · · · ⊗ (un ≈ vn ) ≤ (f (u1 , . . . , un ) ≈ f (v1 , . . . , vn )) . (1.83)
n
An n-ary L-relation r ∈ LU in U is said to be compatible with an L-equality
≈ in U if for arbitrary u1 , v1 , . . . , un , vn ∈ U we have
(u1 ≈ v1 ) ⊗ · · · ⊗ (un ≈ vn ) ⊗ r(u1 , . . . , un ) ≤ r(v1 , . . . , vn ) . (1.84)

Remark 1.81. (1) The above compatibility conditions make sense for arbitrary
binary L-relation ≈ in U . Particularly, it is often used for ≈ being an L-
equivalence relation.
(2) Verbal descriptions of compatibility conditions are the well-known
equality axioms (sometimes called congruence axioms or compatibility ax-
ioms). For functions, compatibility condition says “if u1 and v1 are in ≈
and · · · un and vn are in ≈ then f (u1 , . . . , un ) and f (v1 , . . . , vn ) are in ≈”.
For fuzzy relations, compatibility condition says “if u1 and v1 are in ≈ and
· · · un and vn are in ≈, and u1 , . . . , un are in r then v1 , . . . , vn are in r”.
It is almost immediate that (1.83) and (1.84) are satisfied iff the first-order
formulas expressing the equality axioms are true in degree 1.
(3) Compatibility of fuzzy relations has useful equivalent formulations.
Using adjointness, (1.83) is equivalent to
(u1 ≈ v1 ) ⊗ · · · ⊗ (un ≈ vn ) ≤ r(u1 , . . . , un ) → r(v1 , . . . , vn ) ,
1.3 Fuzzy Sets 45

which is due to symmetry of ≈ equivalent to


(u1 ≈ v1 ) ⊗ · · · ⊗ (un ≈ vn ) ≤ r(u1 , . . . , un ) ↔ r(v1 , . . . , vn ) .
(4) If ≈ is a (characteristic function of) ordinary identity, both conditions
(1.83) and (1.84) are satisfied for free.
(5) For an arbitrary n-ary L-relation r in U , the least L-relation C≈ (r)
containing r which is compatible with ≈ is given by

C≈ (r)(u1 , . . . , un ) = v1 ,...,vn ∈U r(v1 , . . . , vn ) ⊗ (u1 ≈ v1 ) ⊗ · · · ⊗ (un ≈ vn ) .
C≈ (r) is called the ≈-closure (extensional closure) of r. C≈ is a particular
n
closure operator in LU , see in [12, 59].

Lemma 1.82. An L-equivalence θ is compatible with an L-equality ≈ on U


iff ≈ ⊆ θ.

Proof. Suppose θ is compatible with ≈, i.e.


θ(a , b ) ⊗ (a ≈ a  ) ⊗ (b ≈ b  ) ≤ θ(a  , b  ) , (1.85)
for all a , a  , b , b  ∈ M . 1.85 and reflexivity of θ and ≈M yield (a ≈ b ) =
θ(a , a ) ⊗ (a ≈ a ) ⊗ (a ≈ b ) ≤ θ(a , b ). Conversely, if ≈ ⊆ θ, we get θ(a , b ) ⊗
(a ≈ a  ) ⊗ (b ≈ b  ) ≤ θ(a , b ) ⊗ θ(a , a  ) ⊗ θ(b , b  ) ≤ θ(a  , b  ), by transitivity
and symmetry of θ. 

   
Given two sets with fuzzy equalities, U, ≈U and V, ≈V , it might be
desirable to consider a mapping h of U to V . Such a mapping h should again
be compatible with the underlying fuzzy equality. In terms of the following
definition, h should be an ≈-morphism.
   
Definition 1.83. For sets with L-equalities U, ≈U and V, ≈V , a mapping
h : U → V is called an ≈-morphism if
u1 ≈U u2 ≤ h(u1 ) ≈V h(u2 ) (1.86)
for all u1 , u2 ∈ U . For a mapping h : U → V denote by θh a binary L-relation
on U for which we have θh (u1 , u2 ) = h(u1 ) ≈V h(u2 ). θh is called a kernel
of the ≈-morphism h.
   
Remark 1.84. (1) If h : U → V is an ≈-morphism
  of
 U, ≈ U
 to V, ≈V
, we
usually denote this fact explicitly by h : U, ≈ → V, ≈ .
U V

(2) Condition (1.86) says “if u1 and u2 are in ≈U then h(u1 ) and h(u2 )
are in ≈V ”. Since we interpret ≈U and ≈V as underlying similarities, (1.86)
in fact says “if u1 and u2 are similar then h(u1 ) and h(u2 ) are similar”, i.e.
an ≈-morphism is required to map similar elements to similar ones.
(3) If ≈U is a (characteristic function of) ordinary identity then (1.86) is
satisfied for free.

The following assertions summarize basic properties of ≈-morphisms.


46 1 Introduction to Fuzzy Sets and Fuzzy Logic
   
Theorem 1.85. Let h : U → V be a mapping between U, ≈U and V, ≈V .
Then h is an ≈-morphism if and only if ≈U ⊆ θh .

Proof. Follows directly from Definition 1.83. 



   
Theorem 1.86. Let us have ≈-morphisms g : U, ≈U → U  , ≈U and
    
h : U  , ≈U → U  , ≈U . Then
(i) the composed mapping (g ◦ h) : U → U  is an ≈-morphism.
(ii) If g is a bijection, then g −1 is an ≈-morphism iff

u1 ≈U u2 = g(u1 ) ≈U g(u2 ) for all u1 , u2 ∈ U .

Proof. (i): The mapping g ◦ h, i.e. (g ◦ h)(u ) = h(g(u )) for every u ∈ U , is


evidently an ≈-morphism.
(ii): “⇒”: Suppose g is a bijection and g −1 is an ≈-morphism. Then u1 ≈U

u2 ≤ g(u1 ) ≈U g(u2 ) ≤ g −1 (g(u1 )) ≈U g −1 (g(u2 )) = u1 ≈U u2 , which

implies the requested condition u1 ≈U u2 = g(u1 ) ≈U g(u2 ).

“⇐”: If g is a bijection satisfying u1 ≈ u2 = g(u1 ) ≈U g(u2 ) for all
U
−1
u1 , u2 ∈ U , then g is evidently an ≈-morphism. 

   
Theorem 1.87. Let h : U, ≈U → V, ≈V be an ≈-morphism. Then
(i) θh is an L-equivalence compatible with ≈U ;
(ii) θh is an L-equality iff h is injective.

Proof. (i): Since θh (u1 , u2 ) = h(u1 ) ≈V h(u2 ) and since ≈V is an L-


equivalence, θh is an L-equivalence. By Theorem 1.85, ≈U ⊆ θh , and by
Lemma 1.82, θh is compatible with ≈U .
(ii): “⇒”: For an L-equality θh and h(u1 ) = h(u2 ) we have θh (u1 , u2 ) = 1
which implies u1 = u2 . Thus, h is injective.
“⇐”: Let h be an injective ≈-morphism. For θh (u1 , u2 ) = 1 we have
h(u1 ) ≈V h(u2 ) = 1, and so h(u1 ) = h(u2 ), that is u1 = u2 , i.e. θh is an
L-equality. 


Let us note that if u1 ≈U u2 = h(u1 ) ≈V h(u2 ) for all u1 , u2 ∈ U , then


θh = ≈U , i.e. θh is an L-equality. By applying Theorem 1.87 (ii), we obtain
the following corollary.
   
Corollary 1.88. If for an ≈-morphism h : U, ≈U → V, ≈V we have
u1 ≈U u2 = h(u1 ) ≈V h(u2 ) for every u1 , u2 ∈ U , then h is injective. 


Factor Sets by Fuzzy Equivalences

One of the basic constructions in the ordinary setting is that of formation of


a factor set U/θ of a set U by an equivalence relation θ in U , see Sect. 1.1.
Theorem 1.75 gives a way to proceed analogously in fuzzy setting. That is,
given an L-equivalence θ in U , one may define a factor set U/θ of U by θ by
U/θ = Πθ , i.e. U/θ = {[u ]θ | u ∈ U }. In the light of the previous section,
1.3 Fuzzy Sets 47

however, one might ask what happens if we start by a set with fuzzy equality
rather than just by a set, i.e. by U, ≈ rather than just by U .
First, note that if starting by U, ≈ , i.e. a set U equipped with an L-
equality ≈, we require the L-equivalence θ to be compatible with ≈, i.e.
(u1 ≈ v1 ) ⊗ (u2 ≈ v2 ) ⊗ θ(u1 , u2 ) ≤ θ(v1 , v2 ) ,
cf. also Lemma 1.82.
Second, note that Theorem 1.75 can be carried over to the more general
setting when we start by U, ≈ . In more detail, one can proceed as follows.
Modify the definition of an L-partition (Definition 1.73) to that of an L-
partition compatible with ≈ by requiring that each A ∈ Π be compatible
with ≈. Then the corresponding modification of Theorem 1.75 remains true
(now, it concerns a bijective correspondence between L-equivalences θ in U
compatible with an L-equivalence ≈, and L-partitions Π of U compatible with
≈). To see this, it is sufficient to check that starting from an L-equivalence
θ in U compatible with ≈, each [u ]θ is compatible with ≈, and that starting
from an L-partition Π of U compatible with ≈, the induced L-equivalence θΠ
is compatible with ≈. Both claims are true. Indeed, since
[u ]θ (v ) ⊗ (v ≈ w ) = θ(u , v ) ⊗ (v ≈ w ) ≤ θ(u , w ) = [u ]θ (w ) ,
each [u ]θ is compatible w.r.t. ≈. Furthermore, compatibility of θΠ with ≈
means
(u1 ≈ v1 ) ⊗ (u2 ≈ v2 ) ⊗ θΠ (u1 , u2 ) ≤ θΠ (v1 , v2 ) ,
i.e.
(u1 ≈ v1 ) ⊗ (u2 ≈ v2 ) ⊗ Au1 (u2 ) ≤ Av1 (v2 ) ,
where Au1 , Au2 ∈ Π are (uniquely determined) elements of Π such that
Au1 (u1 ) = 1 and Av1 (v1 ) = 1. Now, the latter inequality is true since
(u1 ≈ v1 ) ⊗ (u2 ≈ v2 ) ⊗ Au1 (u2 ) ≤
≤ (u1 ≈ v1 ) ⊗ Au1 (v2 ) =
= Av1 (v1 ) ⊗ Au1 (u1 ) ⊗ (u1 ≈ v1 ) ⊗ Au1 (v2 ) ≤
≤ Av1 (v1 ) ⊗ Au1 (v1 ) ⊗ Au1 (v2 ) ≤ Av1 (v2 ) ,
the last inequality being true due to condition (iii) of the definition of L-
partition. Now, since we work in the framework of sets with fuzzy equalities,
a factor set U, ≈ /θ of U, ≈ by an L-equivalence θ in U compatible with
≈ should be again a set with fuzzy equality. A natural choice is to define
U, ≈ /θ = U/θ, ≈U/θ where U/θ = Πθ and ≈U/θ is an L-equality in U/θ
defined by
([u ]θ ≈U/θ [v ]θ ) = θ(u , v ) .

Remark 1.89. In order to check that the above definition of U, ≈ /θ is sound,
we need to verify that ≈U/θ is correctly defined and that it is indeed an
L-equality. To see that the definition of ≈U/θ is correct, observe first the
48 1 Introduction to Fuzzy Sets and Fuzzy Logic

following Claim: For any u , u  ∈ U , [u ]θ = [u  ]θ iff θ(u , u  ) = 1. Indeed, if


[u ]θ = [u  ]θ then θ(u , u  ) = [u ]θ (u  ) = [u  ]θ (u  ) = 1; if θ(u , u  ) = 1 then by
symmetry and transitivity of θ, for any v ∈ U we have [u ]θ (v ) = θ(u , v ) =
θ(u  , u ) ⊗ θ(u , v ) ≤ θ(u  , v ) = [u  ]θ (v ) and similarly, [u  ]θ (v ) ≤ [u ]θ (v ),
whence [u ]θ = [u  ]θ . Now, if [u ]θ = [u  ]θ and [v ]θ = [v  ]θ , we have ([u ]θ ≈U/θ
[v ]θ ) = θ(u , v ) = [u ]θ (v ) = [u  ]θ (v ) = θ(u  , v ) = θ(v , u  ) = [v ]θ (u  ) =
[v  ]θ (u  ) = θ(v  , u  ) = θ(u  , v  ) which means that ([u ]θ ≈U/θ [v ]θ ) does not
depend on the particular u and v and so ≈U/θ is defined correctly. Now,
the fact that θ is reflexive, symmetric, and transitive immediately implies the
corresponding properties of ≈U/θ . By definition, ([u ]θ ≈U/θ [v ]θ ) = 1 means
θ(u , v ) = 1, from which we get [u ]θ = [v ]θ by Claim. Therefore, ≈U/θ is an
L-equality in U/θ.
By the above generalization of Theorem 1.75, U/θ is a set of L-sets in U which
are compatible with ≈. Moreover,
 θ can be reconstructed from U/θ (namely,
U/θ = Πθ and θ = θΠθ ). U/θ, ≈U/θ is thus a good candidate for being a set
with L-equality which results by factorization of U, ≈ by θ.
But now, notice that such a definition  of a factor
 set of U, ≈ by θ can
be simplified. The point is that using U/θ, ≈U/θ as a factor set with fuzzy
equality, a complete information about θ is present in both U/θ and ≈U/θ .
Consider, instead of U/θ, ≈U/θ , an alternate definition of a factor set with
fuzzy equality. By Theorem 1.72 (ii), 1 θ is an ordinary equivalence relation.
 1  1
One may thus consider a pair U/1 θ, ≈U/ θ where ≈U/ θ is an L-relation in
the ordinary factor set U/1 θ defined by
1
([u ]1 θ ≈U/ θ
[v ]1 θ ) = θ(u , v ) . (1.87)
By Claim of Remark 1.89, [u ]θ = [u  ]θ iff θ(u , u  ) = 1 iff u , u  ∈ 1 θ iff
[u ]1 θ = [u  ]1 θ . Therefore, a mapping sending [u ]θ to [u ]1 θ is a bijection be-
tween U/θ and U/1 θ. As a direct consequence of the above definitions and
considerations we have the following lemma.

Lemma 1.90. Let U, ≈ be a set with L-equality, θ be an L-equivalence in


U compatible with ≈. Then
 
(i) U/θ, ≈U/θ is a set with L-equality,
 1 1 
(ii) U/ θ, ≈U/ θ is a set with L-equality,
(iii) a mapping h : U/θ → U/1 θ defined by h([u ]θ ) = [u ]1 θ is a bijective ≈-
   1 
morphism between U/θ, ≈U/θ and U/1 θ, ≈U/ θ for which
1
([u ]θ ≈U/θ [v ]θ ) = ([u ]1 θ ≈U/ θ [v ]1 θ ).

Lemma 1.90 thus says that we have two almost equivalent ways to define a
factor set with fuzzy equality of U, ≈ by θ. Since by Claim of Remark 1.89,
 1 
[u ]1 θ = 1 [u ]θ , i.e. [u ]1 θ is a subset of [u ]θ . That is, U/1 θ, ≈U/ θ is more
 
simple a concept than U/θ, ≈U/θ . Therefore the following definition.
1.3 Fuzzy Sets 49

Definition 1.91. For a set U, ≈ with L-equality and an L-equivalence θ in


 1 
U which is compatible with ≈, U/1 θ, ≈U/ θ is called a factor set with
fuzzy equality of U, ≈ by θ.
 1 
Remark 1.92. To simplify notation, a factor set U/1 θ, ≈U/ θ with fuzzy
 
equality will be denoted just by U/θ, ≈U/θ and classes [u ]1 θ just by [u ]θ .
This means that, for the sake
 of simplicity, we sacrifice uniqueness of no-
tation. Namely, U/θ, ≈U/θ introduced above is now redefined to denote
 1 1   
U/ θ, ≈U/ θ . But since the above-introduced meaning of U/θ, ≈U/θ will
not be used any more, there is no danger of confusion involved here.

Further Concepts from Fuzzy Sets

Here we recall a concept of a particular closure operator in LU . For a truth


stresser ∗ , an L∗ -closure operator (fuzzy closure operator with truth stresser)
on U [14] is a mapping C : LU → LU satisfying
A ⊆ C(A) , (1.88)

S(A, B) ≤ S(C(A), C(B)) , (1.89)
C(A) = C(C(A)) , (1.90)
for every L-sets A, B ∈ LU .

Remark 1.93. (1) Conditions (1.88) and (1.90) are the usual conditions of
extensivity and idempotency. Condition (1.89) is a particular form of graded
monotony. It says “if is is very true that A is a subset of B then C(A) is a
subset of C(B)” where the interpretation of “very true” is given by ∗ .
(2) It is easy to see that if ∗ is globalization, (1.89) says that A ⊆ B
implies C(A) ⊆ C(B). This is a commonly required condition of fuzzy closure
operators, see e.g. [42]. If ∗ is identity, (1.89) says S(A, B) ≤ S(C(A), C(B)).
This form of monotony is considered e.g. in [7]. This is a stronger form of
monotony than the one for ∗ being globalization. The corresponding L∗ -closure
operators are called L-closure operators.
(3) If L = 2, the only truth stresser is identity (which coincides with
globalization in this case). In this case, the concept of an L∗ -closure operator
coincides with the concept of an ordinary closure operator on U , see Sect. 1.1.

We will make use of the following characterization of L∗ -closure operators.

Lemma 1.94. A mapping C : LU → LU is an L∗ -closure operator if it satis-


fies (1.88) and

S(A1 , C(A2 )) ≤ S(C(A1 ), C(A2 )) . (1.91)

Proof. Assume (1.88)–(1.90). We get S(A1 , C(A2 )) ≤ S(C(A1 ), C(C(A2 ))) =
S(C(A1 ), C(A2 )), verifying (1.91). Conversely, assume (1.88) and (1.91). By
(1.88), A2 ⊆ C(A2 ), whence S(A1 , A2 ) ≤ S(A1 , C(A2 )). Furthermore,
50 1 Introduction to Fuzzy Sets and Fuzzy Logic
∗ ∗
S(A1 , A2 ) ≤ S(A1 , C(A2 )) ≤ S(C(A1 ), C(A2 )) ,

proving (1.89). By (1.91), 1 = S(C(A), C(A)) ≤ S(C(C(A)), C(A)), i.e. we
have C(C(A)) ⊆ C(A). Since the converse inclusion holds by (1.88), we con-
clude (1.90). 


1.4 Pavelka-Style Fuzzy Logic


In this section we recall an approach developed in Pavelka’s seminal [74].
Particularly, we introduce what Pavelka calls an abstract fuzzy logic. Our main
goal is to show what has become known as Pavelka-style fuzzy logic.
The basic setting in logic can be seen as follows. We work with formulas,
i.e. particular sequences of symbols of a particular language. From an abstract
point of view, we might consider formulas just as elements of some (abstract)
set Fml of all formulas. Furthermore, we want to evaluate formulas, i.e. to as-
sign truth degrees to formulas. In a general setting, a set of all truth degrees
may be supposed to form a complete lattice L (i.e., not necessarily equipped
with ⊗ and →) with support L. A truth degree from L may be interpreted as
a degree of truth in the sense of fuzzy logic, as a probability of a given formula
(event), but any other meaningful interpretation is possible. The process of
assignment of truth degrees to formulas has two “inputs” and one “output”.
The first input is a formula ϕ ∈ Fml . The second input is a semantic com-
ponent S in which formulas can be evaluated. The output is a truth degree
||ϕ||S ∈ L of ϕ in S. This way, a mapping E : Fml → L, i.e. an L-set E in
Fml , is induced by E(ϕ) = ||ϕ||S . From an abstract point of view, we may
forget about the semantic components S and see the induced mappings E
as primitive concepts which describe the semantics of our logic. This way we
come to the following definition.
Definition 1.95. An L-semantics for a set Fml of formulas is a set S of
L-sets in Fml , i.e. S ⊆ LFml .
Elements E ∈ S are called evaluations. For ϕ ∈ Fml and E ∈ S, E(ϕ) is
called a truth degree of ϕ in E. To retain the intended meaning of E (E as a
semantic component of evaluation), we use also ||ϕ||E instead of E(ϕ).
Example 1.96. (1) Propositional logic. Let Fml be a set of formulas of classical
propositional logic. That is, Fml contains propositional variables p, q, . . . , and
if ϕ, ψ ∈ Fml then nϕ ∈ Fml and (ϕ i ψ) ∈ Fml (n and i are symbols
of negation and implication). Let L be a two-element chain, (i.e. L = {0, 1},
0 ≤ 1). Each mapping e (called elementary evaluation) assigning truth degrees
e(p), e(q), . . . from L to propositional variables p, q, . . . induces a mapping
|| · · · ||e assigning truth degrees ||ϕ||e from L to formulas ϕ ∈ Fml as follows:
For a propositional variable p, put ||p||e = e(p); for formulas ϕ, ψ ∈ Fml , put
||nϕ||e = ¬||ϕ||e and ||ϕ i ψ||e = ||ϕ||e → ||ψ||e where ¬ and → are truth
functions of classical negation and implication, respectively. Then
1.4 Pavelka-Style Fuzzy Logic 51

S = {E ∈ LFml | for some e : E(ϕ) = ||ϕ||e for each ϕ ∈ Fml }


is an L-semantics. This shows how the usual semantics of ordinary proposi-
tional logic can be seen in the framework of Pavelka’s abstract approach.
(2) Predicate fuzzy logic with truth constants in language. Let L be a com-
plete residuated lattice and Fml be a set of all formulas of predicate fuzzy logic
with (some) truth constants in a language. That is, we start with a language
consisting of a non-empty set R of relation symbols (each r ∈ R with its arity
σ(r) ∈ N0 ), a set F of function symbols (each f ∈ F with its arity σ(f ) ∈ N0 ),
(object) variables x, y, . . . , symbols o, i, c, d of connectives, symbols ∀, ∃ of
quantifiers, symbols a of truth constants for each a ∈ K for some K ⊆ L, and
possibly auxiliary symbols. Then, in a usual way, we define terms (each vari-
able x is a term; if t1 , . . . , tn are terms and f ∈ F is n-ary then f (t1 , . . . , tn ) is
a term) and formulas (if a ∈ K then a is a formula; if t1 , . . . , tn are terms and
r ∈ R is n-ary then r(t1 , . . . , tn ) is an (atomic) formula; if ϕ, ψ are formulas
and x is a variable then (ϕ o ψ), (ϕ i ψ), (ϕ c ψ), (ϕ d ψ), (∀x)ϕ, (∃x)ϕ are
formulas). Denote Fml the set of all such  formulas. Then, we define a structure
of a given language as a triplet M = M, RM , F M where M = ∅ is a universe,
σ(r)
RM = {rM ∈ LM | r ∈ R} is a set of L-relations corresponding to relation
symbols, and F M = {f M : M σ(f ) → M | f ∈ F } is a set of (ordinary) func-
tions corresponding to function symbols. Given a structure M and a valuation
v assigning elements v(x), v(y), · · · ∈ M to variables x, y, . . . , we define eval-
uation of terms and evaluation of formulas. A value ||t||M,v ∈ M of term t is
defined as follows: for variable x, put ||x||M,v = v(x); for a term f (t1 , . . . , tn ),
put ||f (t1 , . . . , tn )||M,v = f M (||t1 ||M,v , . . . , ||tn ||M,v ). A value (truth degree)
||ϕ||M,v ∈ L of formula ϕ is defined as follows: ||a||M,v = a for each a ∈ K;
for formula r(t1 , . . . , tn ), put ||r(t1 , . . . , tn )||M,v = rM (||t1 ||M,v , . . . , ||tn ||M,v );
for formulas ϕ, ψ, and variable x, put ||ϕ o ψ||M,v = ||ϕ||M,v ⊗ ||ψ||M,v ;
||ϕ i ψ||M,v = ||ϕ||M,v → ||ψ||M,v ; ||ϕ c ψ||M,v= ||ϕ||M,v ∧ ||ψ||M,v ;
||ϕ d ψ||M,v = ||ϕ||M,v ∨ ||ψ||M,v ; ||(∀x)ϕ||M,v = w≡x v ||ϕ||M,w with in-
fimum takenover all valuations w which differ  from v at most in x;
||(∃x)ϕ||M,v = w≡x v ||ϕ||M,w with supremum taken over all valuations
w which differ from v at most in x. Then
S = {E ∈ LFml | for some M, v : E(ϕ) = ||ϕ||M,v for each ϕ ∈ Fml }
is an L-semantics. Another L-semantics can be obtained as follows. Define a
truth degree of a formula ϕ in a structure M by

||ϕ||M = v:X→M ||ϕ||M,v
where the infimum ranges over all valuations v in M . Then,
S = {E ∈ LFml | for some M : E(ϕ) = ||ϕ||M for each ϕ ∈ Fml } (1.92)
is an L-semantics. This shows how semantics of predicate fuzzy logic can be
seen in the framework of Pavelka’s abstract approach.
(3) Probability in finite sets. Let Ω = ∅ be a finite set, ω ∈ Ω be called
elementary events. Let L be the real unit interval [0, 1] with its natural order
52 1 Introduction to Fuzzy Sets and Fuzzy Logic

≤, let Fml be the set of all events over Ω, i.e. Fml = 2 Ω


. A probability
distribution on Ω is a function p : Ω → [0, 1] satisfying ω∈Ω p(ω) = 1.
Probability
 distribution induces a probability measure Pp on Ω by putting
Pp (A) = ω∈A p(ω) for each event A ∈ 2Ω . Then
S = {E ∈ LFml | for some p : E(ϕ) = Pp (ϕ)}
is an L-semantics.

The concept of an L-semantics is rather simple, namely, an L-semantics is


but an L-set in Fml . Nevertheless, it serves as a good starting point to further
considerations interesting from the logical point of view.
In classical logic, a theory is a set T of formulas, i.e. T ⊆ Fml . Theories
are of crucial importance in logic. We might have a set Th(M) of all for-
mulas which are true in some structure M as a theory. Or, a theory T can
be understood as representing somebody’s knowledge about some domain.
For instance, a theory T = {husband (george), . . . , husband (x)imale(x), . . . }
contains formulas which assert that (an individual referred to as) george is
somebody’s husband and that if somebody is a husband then he is a male.
From the point of view of fuzzy approach, however, it is quite appealing to
consider a theory as an L-set of formulas. For instance, theory
T = {0.8/young(john), 0.9/baby(joe), . . . , 1/baby(x) i young(x)}
asserts that john is young in degree (at least) 0.8, that joe is young in degree
(at least) 0.9, that the fact that somebody is a baby implies that he/she is
young in degree 1, etc. This leads to the following definition.

Definition 1.97. A theory over Fml and L (L-theory over Fml ) is any L-set
T ∈ LFml . Given an L-semantics S for Fml , we say that E ∈ S is a model
of a theory T over Fml and L if T ⊆ E.

That is, E is a model of T is for each formula ϕ ∈ Fml we have T (ϕ) ≤


E(ϕ). This says that for each formula ϕ, the truth degree E(ϕ) of ϕ in E is at
least as high as the truth degree prescribed to ϕ by theory T . For instance, for
the L-semantics of predicate fuzzy logic from Example 1.96 (2), an evaluation
E which corresponds to a structure M and a valuation v is a model of a theory
T if for each formula ϕ, the truth degree ||ϕ||M,v of ϕ in M under v is greater
than or equal to T (ϕ). It is easily seen that this is a natural generalization of
the classical concept of a model (just consider L = 2). In Example 1.96 (3),
let Ω = {1, . . . , 6} be a set of elementary events which corresponds to tossing
a die. If the die is fair, we have p(1) = · · · = p(6) = 16 for the corresponding
probability distribution p. However, for a biased die we might have p(1) =
· · · = p(4) = 16 , p(5) = 12 1 3
, p(6) = 12 (6 is more likely to appear than 5).
Then we might consider a theory T = { 12 1 1
/{1}, . . . , 12 /{6}} which can be seen
as requiring for each possible result i ∈ {1, . . . , 6}, the probability of i is at
1
least 12 , i.e. the die cannot be biased too much. For instance, an evaluation E
corresponding to the above distribution p of a biased die, is a model of T . For
1.4 Pavelka-Style Fuzzy Logic 53

a fake die containing only even numbers, the corresponding E is not a model
of T since for i ∈ {1, 2, 3} we have T ({i}) ≤ 0 = E({i}).
Now, we can define the degree of semantic entailment in our abstract
setting.

Definition 1.98. For an L-semantics S for Fml , a theory T , and a formula


ϕ ∈ Fml , the degree ||ϕ||S T to which ϕ semantically follows from T
(denoted simply by ||ϕ||T ) is defined by

||ϕ||ST = {||ϕ||E | E ∈ S is a model of T } .

Recall that the completeness theorem (for classical logic) says that (for
suitable proof systems), formula ϕ follows semantically from a theory T iff ϕ
is provable from T . Therefore, an interesting problem arises of whether the
degree ||ϕ||T can be obtained by syntactic means, i.e. whether ||ϕ||T is equal
or at least can be approximated by a suitably defined “degree of provability”
of ϕ from T . However strange the concept of degree of provability may sound
at the first encounter, it turns out that it is quite natural. A way to go, showed
by Pavelka [74] and inspired by Goguen [45], is the following.
We need to prove formulas from L-sets of formulas. The first step is to
modify the concept of a deduction rule. An ordinary deduction rule takes
formulas as inputs and yields a formula as its output. A deduction rule in our
setting takes formulas with their truth-weights from L as inputs and yields a
formula with a truth-weight as its output. Then, a truth-weighted proof is a
finite sequence ϕ1 , a1 , . . . , ϕn , an of formulas ϕi with their truth-weights
ai ∈ L which results from a theory T (L-set of formulas) by a repeated
application of deduction rules. The truth-weight an of the last formula ϕn
in a proof is considered as a degree of the proof for ϕn . But different proofs
can yield different degrees, i.e. we might have a proof ϕ1 , b1 , . . . , ϕm , bm
with ϕn = ϕm but an = bm . Therefore, the degree |ϕ|T of provability of ϕ
from T is defined as  the supremum of degrees of proofs of ϕ over all possible
proofs, i.e. |ϕ|T = {a | there exists a proof . . . , ϕ, a from T }. The details
are captured by the following definitions.

Definition 1.99. An n-ary deduction rule for Fml and L is a pair R =


Rsyn , Rsem consisting of a partial function Rsyn : Fml n → Fml (syntatic
part) and a function Rsem : Ln → L (semantic part).

Analogously as in the ordinary case, a deduction rule R may be visualized by


ϕ1 , a1 , . . . , ϕn , an
.
Rsyn (ϕ1 , . . . , ϕn ), Rsem (a1 , . . . , an )
A rule R enables us to infer that formula Rsyn (ϕ1 , . . . , ϕn ) is valid in degree
(at least) Rsem (a1 , . . . , an ) if we know that each ϕi is valid in degree at least
ai (i = 1, . . . , n).
54 1 Introduction to Fuzzy Sets and Fuzzy Logic

Example 1.100. (1) Modus ponens for predicate fuzzy logic and for a complete
residuated lattice L, see Example 1.96 (2), is a deduction rule
ϕ, a , ϕ i ψ, b
.
ψ, a ⊗ b
Returning to the above example, modus ponens enables us to infer young(joe),
0.9 from baby(joe), 0.9 and baby(joe) i young(joe), 1 .
(2) Generalization for predicate fuzzy logic is a deduction rule
ϕ, a
.
(∀x)ϕ, a
Note that a truth-weight a of the inferred formula is the same as of the input
formula.
(3) As an example of a deduction rule which is degenerate in the ordinary
case, consider
ϕ, a , ϕ, b
.
ϕ, a ∨ b
Here, the inferred formula is the same as both the input formulas but we infer
the least upper bound a ∨ b of the truth-weights of the input formulas.

In addition to deduction rules, a (Hilbert-style) logical system contains a


collection of (logical) axioms. In our setting, we may consider an arbitrary
L-set A of formulas from Fml , i.e. a theory over Fml and L, as a collection
of logical axioms.

Definition 1.101. Let R be a set of deduction rules for Fml and L, A (logical
axioms) and T (non-logical axioms) be theories over Fml and L, let ϕ ∈ Fml
and a ∈ L. An (L-weighted ) proof of ϕ, a from T using A and R is
a sequence ϕ1 , a1 , . . . , ϕn , an such that ϕn is ϕ, an = a, and for each
i = 1, . . . , n, we have ai = T (ϕi ), or ai = A(ϕi ), or ϕi , ai follows from some
ϕj , aj ’s, j < i, by some deduction rule R ∈ R. The number n is called a
length of the proof. In such a case, we write T A,R ϕ, a and call ϕ, a
provable from Σ using A and R. If T A,R ϕ, a , ϕ is called provable in
degree (at least) a from T using A and R. A degree of provability of ϕ
from T using A and R, denoted by |ϕ|A,R T , is defined by

|ϕ|A,R
T = {a | T A,R ϕ, a } .

Remark 1.102. (1) The fact that ϕi , ai follows from some ϕj , aj ’s, j < i,
by a k-ary deduction rule R ∈ R means that ϕi = Rsyn (ϕi1 , . . . , ϕik ) and
ai = Rsem (ai1 , . . . , aik ) for some i1 , . . . , ik ≤ i.
(2) We say also “A, R-provable” instead of “provable using A and R”. If
A and R are obvious from the context, we omit them, e.g. we write T  ϕ, a
instead of T A,R ϕ, a , |ϕ|T instead of |ϕ|A,R T , etc.
1.4 Pavelka-Style Fuzzy Logic 55

It is usually the case that formulas from Fml are particular strings of
symbols of a so-called language. Then, it is sometimes useful to have symbols
a of truth degrees a ∈ L as particular symbols of the language. One may then
consider the concept of an L-weighted formula.
Definition 1.103. An L-weighted formula is a pair ϕ, a where ϕ is a
formula and a is a symbol of a truth degree a ∈ L.
Remark 1.104. (1) If formulas are strings of symbols of a language, then L-
weighted formulas are as well. Therefore, the concept of an L-weighted formula
is a concept of syntax. If there is no danger of misunderstanding, we usually
do not distinguish between a and a. Therefore, we also write ϕ, a instead of
ϕ, a .
(2) If L is obvious from the context, we also say a truth-weighted formula
or just weighted formula. Terminology is not consistent here. For instance, [71]
uses adjective “evaluated” while [49] prefers “weighted”. We use “weighted”
since “evaluated” suggest a connection to valuations E and truth degrees
||ϕ||E which is not proper here.
(3) There is a correspondence between L-sets (theories) Σ ∈ LFml of
formulas and (particular) sets of L-weighted formulas. Namely, for Σ ∈ LFml ,
one may consider a set TΣ = {ϕ, a | ϕ ∈ Fml , a = Σ(ϕ)} of L-weighted
formulas. For a set T of L-weighted formulas such that for each ϕ there is
at most one a with ϕ, a ∈ T one may consider an L-set ΣT defined by
ΣT (ϕ) = a if ϕ, a ∈ T and ΣT (ϕ) = 0 if there is no a ∈ L with ϕ, a ∈ T .
(4) The main reason we introduce the concept of an L-weighted formula
is that we might want to have some concepts in a usual fashion from the
point of view of syntax and semantics. Namely, a proof is usually defined as
a certain sequence of formulas (well-formed strings of symbols) of a language
of the corresponding logic. Without the concept of an L-weighted formula, an
L-weighted proof is a sequence of pairs ϕ, a where ϕ is a formula and a ∈ L
is a truth degree. Truth degrees come from the metalevel (of a corresponding
logic) and are not a part of language. That is, without the concept of an L-
weighted formula, the concept of an L-weighted proof is not a sequence of well-
formed strings of a language. Having the concept of an L-weighted formula as
a concept of syntax, L-weighted proofs are sequences of well-formed strings of
symbols of language, as usual. In a similar manner, ordinary proofs go from
sets of formulas while in the setting of abstract logic, proofs go from L-sets of
formulas. However, as we have seen in (3), an L-set of formulas can be seen
as an ordinary set of L-weighted formulas. Thus, again, the concept of an
L-weighted formula enables us to have a concept of theory in a fashion usual
in logic.
(5) Note that we could define a truth degree ||ϕ, a ||E of L-weighted
formula ϕ, a by
||ϕ, a ||E = a → ||ϕ||E .
This way, we have ||ϕ, a ||E = 1 iff a ≤ ||ϕ||E . Furthermore, ||ϕ, 1 ||E =
||ϕ||E , and so, formulas correspond to L-weighted formulas with weight a = 1.
56 1 Introduction to Fuzzy Sets and Fuzzy Logic

A degree |ϕ|T of provability is thus a least upper bound of all a ∈ L such


that there exists a proof of ϕ, a from T . However, it may happen that |ϕ|T
is not attained by any proof, i.e. there does not exist any proof of ϕ, a for
a = |ϕ|T . Therefore, if we find a proof of ϕ, a from T , we have just a ≤ |ϕ|T ,
i.e. we have only a lower approximation a of the degree |ϕ|T of provability.
Note that this is different from the classical notion of a proof (!) since in
classical setting, all proofs of a formula are equally important. That is, any
proof does the job.
Then, we can define the following.

Definition 1.105. An abstract logic is a tuple L = Fml , L, S, A, R where


Fml is a set of formulas, L is a complete lattice, S is an L-semantics for Fml ,
A is a theory (of logical axioms) over Fml and L, and R is a set of deduction
rules. L is (Pavelka-)sound if for each theory T over Fml and L and each
formula ϕ ∈ Fml we have |ϕ|A,R T ≤ ||ϕ||ST . L is (Pavelka-)complete if for
each theory T over Fml and L and each formula ϕ ∈ Fml we have |ϕ|A,R T =
||ϕ||ST .

Remark 1.106. One can see that the ordinary notions of soundness and com-
pleteness are particular cases of those from Definition 1.105 for L being a
two-element chain.

Another issue of a standard logical agenda which can be introduced in


terms of our abstract setting is a pair of mappings Md and Th. Introduce
mappings Md : LFml → 2S and Th : 2S → LFml as follows. For an L-set T of
formulas Fml define a set Md(T ) of evaluations from S by
Md(T ) = {E ∈ S | E is a model of T } . (1.93)
For a set K of evaluations from S, define an L-set Th(K) of formulas by

(Th(K))(ϕ) = {E(ϕ) | E ∈ K} . (1.94)
That is, Md(T ) is a set of all models of T while Th(K) is a theory of K. It is
 S to see that Md and Th form a Galois connection between L , ⊆ and
Fml
easy
2 , ⊆ . That is, we have
T1 ⊆ T2 implies Md(T2 ) ⊆ Md(T1 ) , (1.95)
K1 ⊆ K2 implies Th(K2 ) ⊆ Th(K1 ) , (1.96)
K ⊆ Md(Th(K)) , (1.97)
T ⊆ Th(Md(T )) , (1.98)
for any K, K1 , K2 ⊆ S and any T, T1 , T2 ∈ LFml . As a consequence, we also
have
Md(T ) = Md(Th(Md(T ))) , (1.99)
Th(K) = Th(Md(Th(K))) . (1.100)
1.5 Bibliographical Remarks 57

Remark 1.107. Note that using Md and Th, a degree ||ϕ||ST to which ϕ seman-
tically follows from T can be expressed by
||ϕ||ST = (Th(Md(T )))(ϕ) .

1.5 Bibliographical Remarks


Sets and Structures

More information related to sets, mappings, relations, and structures can be


found in many textbooks, e.g. in [25, 75, 88].

Structures of Truth Degrees

Fuzzy sets and fuzzy logic were introduced by Lotfi A. Zadeh [100]. The set of
truth degrees proposed by Zadeh is the real unit interval [0, 1] equipped with
min as conjunction. The idea to use complete residuated lattices as structures
of truth degrees is due to Goguen [44, 45]. Particularly, our justification of
adjointness property is basically due to Goguen (cf. also beginning of Chap. 2
in [49]). Residuated lattices have been introduced by Ward and Dilworth [94].
For the point of view of fuzzy logic, residuated lattices have been thorouhly
investigated by Höhle, see e.g. [55]. Several useful results concerning residuated
lattices can be found e.g. in [10, 47, 49, 71]. The notion of a t-norm goes back
to the study of probabilistic metric spaces an particularly to Schweizer and
Sklar, see [78] and [79]. A good book on t-norms is [60]. For representation
of continuous t-norms by ordinal sums see [67, 63], for definition of ordinal
sums of residuated lattices see [48]. The concept of a truth stresser is due to
Hájek [51], see also [4, 83] for related papers.

Fuzzy Sets

The first paper on fuzzy sets is Zadeh’s [100]. Fuzzy sets and their applications
are well covered e.g. in [35, 46, 47, 61]. The set of truth degrees is usually
[0, 1]. The concept of an L-set is due to Goguen [44] (Goguen uses the term
L-fuzzy set). Fuzzy sets and fuzzy relations with truth degrees in complete
residuated lattices were investigated by Höhle, see e.g. a survey paper [56].
The notion of a fuzzy equivalence is due to Zadeh [99]. Fuzzy equivalences
and fuzzy equalities with truth degeees in a complete residuated lattice were
studied by Höhle [54, 56, 57]. There are many papers on various aspects of
fuzzy equivalence relations, see e.g. [21, 30, 36, 37, 49, 58, 59, 72, 85, 87].
58 1 Introduction to Fuzzy Sets and Fuzzy Logic

Pavelka-style Fuzzy Logic

The concept of an abstrat fuzzy logic is due to Pavelka [74] (it was inspired
by Goguen’s [44, 45]). A thorough investigation of abstract fuzzy logic is
provided by [42]. Several particular logical calculi fitting the framework of
abstract fuzzy logic (most noticeably: obeying Pavelka-completeness) can be
found e.g. in [42, 49, 71, 74].
2
Algebras with Fuzzy Equalities

The present chapter studies algebras with fuzzy equalities. Briefly speaking,
an algebra with fuzzy equality is a set equipped with a fuzzy equality and with
functions which are compatible with this fuzzy equality. If we interpret fuzzy
equality as similarity, compatibility of functions with fuzzy equality says that
the functions map similar (tuples of) elements to similar elements. From this
point of view, an algebra with fuzzy equality can be seen as a collection of
similarity-preserving functions on a set.
Functions operating on a set so that close (similar) elements are mapped
to close elements have traditionally been the subject of study of calculus
and functional analysis. The concept of closeness has been almost exclusively
formalized using the notion of a metric. However, the very essence of the prob-
lem calls for a logical treatment. Namely, formulated verbally, the condition
of mapping similar elements to similar ones reads “if arguments of a func-
tion are pairwise similar then the results are similar as well”. From a logical
point of view, this condition can be described by a logical formula called a
compatibility axiom (or congruence axiom). Equivalence relations satisfying
compatibility axiom with some given functions are called congruence relations.
In bivalent setting, congruence relations do not constitute appropriate frame-
work for studying similarity-preserving mappings. Namely, ordinary congru-
ence relations are bivalent while similarity is a graded notion. In fuzzy setting,
however, one can use fuzzy equivalence/equality relations for modeling of sim-
ilarity. Nevertheless, the axiom expressing preservation of similarity by func-
tions remains the same as in bivalent setting, namely, the above-mentioned
compatibility axiom. While this axiom retains its clear verbal description, its
meaning depends on the choice of a conjunction operation (usually a t-norm)
and if truth degrees are numbers, it has a numerical significance. As we will
see, the meaning of compatibility axiom in fuzzy setting is rich enough to
capture interesting aspects of the problem of similarity-preserving functions.
The notion of an algebra with fuzzy equality generalizes that of a uni-
versal algebra. Namely, with fuzzy equality being the identity relation, the
concept of an algebra with fuzzy equality coincides with that of a universal

R. Bělohlávek and V. Vychodil: Fuzzy Equational Logic, StudFuzz 186, 59–137 (2005)
www.springerlink.com 
c Springer-Verlag Berlin Heidelberg 2005
60 2 Algebras with Fuzzy Equalities

algebra. This fact delineates the basic way of our investigation. In principle,
we study problems traditionally studied in universal algebra. Algebraic prop-
erties of algebras with fuzzy equalities are our main concern in this chapter.
Our generalized setting leaves many of the results obtained in universal al-
gebra particular cases of more general results. In addition to that, there are
several new aspects in our setting which are hidden in the ordinary bivalent
setting.
Analogously to the ordinary case, algebras with fuzzy equalities are struc-
tures of a fragment of predicate fuzzy logic. Several aspects interesting from
this point of view will be investigated in subsequent chapters.
∗ ∗ ∗
Section 2.1 introduces the concept on an algebra with fuzzy equality and
presents examples. Basic structural notions and constructions for algebras
with fuzzy equalities are introduced and studied in Sect. 2.2. Section 2.3 stud-
ies direct and subdirect products of algebras with fuzzy equalities. Section 2.4
deals with terms, term functions, and related issues. Sections 2.5, 2.6, and 2.7
study more advanced constructions of direct unions, direct limits, and reduced
products. Section 2.8 deals with classes of algebras with fuzzy equalities and
with operators over these classes. Section 2.9 presents selected approaches to
study of algebras from the point of view of fuzzy logic with comments.

2.1 Definition and Examples


A type is a triplet ≈, F, σ where ≈ ∈ F and σ is a mapping σ : F ∪ {≈} → N0
with σ(≈) = 2. Each f ∈ F is called a function symbol, ≈ is a relation symbol
called a symbol for fuzzy equality. Mapping σ assigns an arity σ(f ) to every
function symbol f ∈ F . Symbol ≈ stands always for a binary relation symbol.
If there is no danger of confusion, a type will be denoted simply by F .
Definition
 2.1. An  algebra with
 L-equality
 of type ≈, F, σ is a triplet
M = M, ≈M , F M such that M, F M is an algebra of type F, σ and ≈M
on M such that each f ∈ F is compatible with ≈ . The
M M M
is an L-equality
 M
algebra M, F , denoted by ske(M), is called a skeleton of M .
 
The fact that M, F M is an algebra of type F, σ means that F M =
{f M : M σ(f ) → M | f ∈ F }, i.e., F M contains a σ(f )-ary function f M on M
for each f ∈ F . Recall from Definition 1.80 that the compatibility condition
says that
(a1 ≈M b1 ) ⊗ · · · ⊗ (an ≈M bn ) ≤ f M (a1 , . . . , an ) ≈M f M (b1 , . . . , bn )
for each n-ary f ∈ F and every a1 , b1 , . . . , an , bn ∈ M . An algebra with
L-equality will also be simply called an L-algebra. If L is obvious, we say
also an algebra with fuzzy equality. Following common usage, we some-
times identify an L-algebra M with its support set M . Moreover, we write
2.1 Definition and Examples 61
   
M, ≈M , f1M , . . . , fnM instead of M, ≈M , {f1M , . . . , fnM } and sometimes also
omit the superscript M from ≈M , F M , and f M . In the following, we often
call the degree (a ≈M b ) the degree of similarity between a and b which due
to the intended interpretation of ≈M .
The concept of an L-algebra is not artificial. It is obvious that the com-
patibility axiom expresses a natural constraint on the operations requiring
that each f M maps similar arguments to similar results. If one takes, e.g.,
L = [0, 1], this constraint has a numerical character. Moreover, this con-
straint is expressed in a simple fragment of first-order fuzzy logic. Namely,
f M is compatible with ≈M if and only if a first-order formula saying that “if
a1 and b1 are similar and · · · and an and bn are similar then f M (a1 , . . . , an )
and f M (b1 , . . . , bn ) are similar” is true in M. Therefore, unlike ordinary al-
gebras, L-algebras M have two non-trivial  parts. First, the “functional part”,
which is an ordinary algebra M, F M . Second,  the “relational part” which
is a set with a fuzzy equality, namely, M, ≈M .

Remark 2.2. (1) For L = 2 (two-element Boolean algebra), the only L-equality
≈ on M is the ordinary identity relation, i.e. (a ≈ b ) = 1 for a = b , (a ≈
b ) = 0 for a = b . It is therefore easy to see that 2-algebras coincide with
ordinary algebras. This is the first way the notion of an L-algebra generalizes
the notion of an (ordinary) algebra.
(2) Taking this point of view, one may consider a structure L of truth
degrees as a parameter and think of the theory of L-algebras as a parametrized
theory. Setting L = 2, we get the theory of ordinary universal algebras. As
we will see in the following, several results are valid for each L (each complete
residuated lattice). However, there are results which are true only for L’s
satisfying special properties and so, additional properties of L are important.
(3) Obviously, taking
 arbitrary L and a crisp L-equality ≈M , L-algebras
M = M, ≈ , F M M
can be identified with ordinary algebras. This is the
second way the notion of an L-algebra generalizes the notion of an algebra.

The following are some examples of L-algebras.

Example 2.3. Take a four-element set L = {0, a, b, 1} of truth degrees linearly


ordered by 0 < a < b < 1. Consider a complete residuated lattice L on L given
by ⊗ defined by a ⊗ b = min(a, b). Fig. 2.1 and Fig. 2.2 show an example of
an L-algebra M. In particular, M = {0 , u , . . . , z , 1 } is the support of M. A
fuzzy equality ≈M on M is defined by the first table of Fig. 2.1. We have
0 ≈M u = 0, 0 ≈M v = a, etc. There are two operations on M, denoted by
∧M and ∨M , which are defined in the second and the third table of Fig. 2.1.
 that both ∧ and ∨ are compatible with ≈ . Moreover,
M M M
One
 can check
M, ∧ , ∨
M M
is a lattice with its Hasse diagram depicted in the left part of
Fig. 2.2. Therefore, M is a lattice with fuzzy equality. M (both the lattice
part and the fuzzy equality part) is depicted in the right part of Fig. 2.2.
Fuzzy equality ≈M is visualized using gray color. A degree of fuzzy equality
corresponds to darkness of gray. White corresponds to 0, light gray to a, dark
62 2 Algebras with Fuzzy Equalities

Fig. 2.1. Lattice with L-equality (definition)

1 1

z z

w x y w x y

u v u v

0 0

Fig. 2.2. Lattice with L-equality (diagrams)

gray to b, and black to 1. For instance, 0 ≈M v = a since 0 and v are contained


in a light gray box, v ≈M y = b since v and y are contained in a dark gray
box, etc. Clearly, black and white boxes can be omitted.

Example 2.4. Let U be a set equipped with an L-equality ≈U . Let M = S(U )


be the set of all permutations of U compatible with ≈U , i.e. the set of all
bijective mappings π onMU Mfor which we have (u ≈U v ) ≤ (π(u ) ≈U π(v )).
The triplet M = M, ≈ , ◦ where
  
π ≈ σ = u ,v (u ≈U v ) → (π(u ) ≈U σ(v ))
M
(2.1)
and ◦M denotes the composition of permutations, is an L-algebra. Note that
π ≈M σ is the degree to which it is true that similar elements are mapped to
similar ones. Moreover, the restriction of M to bijective mappings satisfying
(u ≈U v ) = (π(u ) ≈U π(v )), is a group with fuzzy equality. To prove this
assertion, we first show that

π ≈M σ = u (π(u ) ≈U σ(u )) . (2.2)
On the one hand, we clearly have
2.1 Definition and Examples 63
  
π ≈M σ = u ,v (u ≈U v ) → (π(u ) ≈U σ(v )) ≤
  
≤ u (u ≈U u ) → (π(u ) ≈U σ(u )) =
   
= u 1 → (π(u ) ≈U σ(u )) = u (π(u ) ≈U σ(u )) .
On the other hand,

u (π(u ) ≈U σ(u )) ≤ π ≈M σ
holds true iff for each u , v ∈ U we have
  
u (π(u ) ≈ σ(u )) ≤ (u ≈ v ) → (π(u ) ≈ σ(v )) ,
U U U

i.e. iff

(u ≈U v ) ⊗ u (π(u ) ≈U σ(u )) ≤ π(u ) ≈U σ(v )
which is true. Indeed, compatibility of σ yields

(u ≈U v ) ⊗ u (π(u ) ≈U σ(u )) ≤ (σ(u ) ≈U σ(v )) ⊗ (π(u ) ≈U σ(u ))
≤ π(u ) ≈U σ(v ) .
It is now routine to check that ≈M is an L-equality on M : reflexivity and sym-
metry follow directly from definition, and for transitivity we have on account
of (2.2) for each u ∈ U that
(π ≈M σ) ⊗ (σ ≈M θ) ≤ (π(u ) ≈U σ(u )) ⊗ (σ(u ) ≈U θ(u )) ≤ π(u ) ≈U θ(u ) ,
that is,

(π ≈M σ) ⊗ (σ ≈M θ) ≤ u (π(u ) ≈U θ(u )) = π ≈M θ ,
proving transitivity of ≈M . If (π ≈M σ) = 1 then, using (2.2), (π(u ) ≈U
σ(u )) = 1 for each u ∈ U . Since ≈U is an L-equality, we have that π(u ) =
σ(u ) for each u ∈ U , whence π = σ. Therefore, ≈M is an L-equality on S(U ).
To verify that ◦M is compatible with ≈M take any π, π  , ,  ∈ M . We have
(π ≈M π  ) ⊗ ( ≈M  ) ≤ (π ◦  ≈M π  ◦  ) iff for each u ∈ U we have
(π ≈M π  ) ⊗ ( ≈M  ) ≤ ((π(u )) ≈U  (π  (u ))) which is true:
(π ≈M π  ) ⊗ ( ≈M  ) ≤
 
≤ (π(u ) ≈U π  (u )) ⊗ (π  (u )) ≈U  (π  (u )) ≤
   
≤ (π(u )) ≈U (π  (u )) ⊗ (π  (u )) ≈U  (π  (u )) ≤
≤ (π(u )) ≈U  (π  (u )) .
 
To sum up, M = M, ≈M , ◦M is an algebra with fuzzy equality.

Example 2.5. Let L be the standard L  ukasiewicz algebra on the unit inter-
val. Consider U = {a , . . . , f }, and let ≈U be given by Fig. 2.3. Following
Example 2.4, there are the four compatible permutations on U :
     
π1 = idU , π2 = aa bb cc df ee df , π3 = ba ba cc dd ee ff , π4 = ba ba cc df ee df .
64 2 Algebras with Fuzzy Equalities

Fig. 2.3. Example of compatible permutations

 
The resulting L-algebra M = {π1 , . . . , π4 }, ≈M , ◦M of compatible permu-
tations is depicted in Fig. 2.3. Note that in this particular case, M can be
seen an a group with L-equality, because ske(M) is a group (so-called Klein’s
group: π1 is the neutral element and each permutation is the inverse for itself).

Example 2.6. Let C : LU → LU be an L-closure operator on X, see Re-


mark 1.93 (2). Furthermore, let SC = {A ∈ LX | C(A) = A} be the system of
all closed fuzzy setsof C, see [7]. Then SC is a complete lattice with respect
to
 ⊆ where infima
 coincide with intersections and for suprema we have
i∈I Ai = C( i∈I Ai ). It is easy to verify that
    
i∈I (Ai ≈ Bi ) ≤ ( i∈I Ai ≈ i∈I Bi ) = ( i∈I Ai ≈ i∈I Bi ) , (2.3)
  
i∈I (Ai ≈ Bi ) ≤ ( i∈I Ai ≈ i∈I Bi ) . (2.4)
Hence, using (1.89), (2.4) it follows that
   
i∈I (Ai ≈ Bi ) ≤ i∈I Ai ≈ i∈I Bi =
     
= S i∈I Ai , i∈I Bi ∧ S i∈I Bi , i∈I Ai ≤
         
≤ S C i∈I Ai , C i∈I Bi ∧ S C i∈I Bi , C i∈I Ai =
       
= C i∈I Ai ≈ C i∈I Bi = i∈I Ai ≈ i∈I Bi .
For I = {1, 2} we can use (1.28), (2.3) and the previous inequality to get
(A1 ≈ B1 ) ⊗ (A2 ≈ B2 ) ≤ (A1 ∧ A2 ) ≈ (B1 ∧ B2 ) ,
(A1 ≈ B1 ) ⊗ (A2 ≈ B2 ) ≤ (A1 ∨ A2 ) ≈ (B1 ∨ B2 ) .
Therefore, SC , ≈, ∧, ∨ is an L-algebra. Particularly, since the identity map-
 operator C (sending A to A) for which SC = L , we get
X
ping is an L-closure
that L , ≈, ∩, ∪ is an L-algebra, i.e. a lattice with fuzzy equality.
X

Example 2.7. Let X and Y be a set of objects and a set of attributes, respec-
tively, and I be an L-relation between X and Y . Let I(x , y ) be interpreted as
a degree to which the object x has the attribute y . Furthermore, let us define
for all L-sets A ∈ LX , B ∈ LY the L-sets A↑ ∈ LY , B ↓ ∈ LX by
  
A↑ (y ) = x ∈X A(x ) → I(x , y ) ,
  
B ↓ (x ) = y ∈Y B(y ) → I(x , y ) .
2.1 Definition and Examples 65

Now, put B(X, Y, I) = {A, B ∈ LX × LY | A↑ = B, B ↓ = A} and define for


A1 , B1 , A2 , B2 ∈ B(X, Y, I) a binary relation ≤ by A1 , B1 ≤ A2 , B2
iff A1 ⊆ A2 (or, iff B2 ⊆ B1 ; both ways are equivalent). The structure
B(X, Y, I), ≤ is called a fuzzy concept lattice induced by X, Y , and I. The
elements of A, B of B(X, Y, I) are naturally interpreted as concepts hidden
in (the input data represented by) I. Namely, A↑ = B and B ↓ = A says that
B is the collection of all attributes shared by all objects from A, and A is
the collection of all objects sharing all attributes from B. Note that these
conditions represent exactly the definition of a concept as developed in the
so-called Port-Royal logic. A and B are called the extent and the intent of
the concept A, B , respectively, and represent the collection of all objects
and all attributes covered by the particular concept. Furthermore, ≤ models
a subconcept-superconcept hierarchy – concept A1 , B1 is a subconcept of
A2 , B2 iff each object from A1 belongs to A2 (dually for attributes).
Put Ext(X, Y, I) = {A ∈ LX | A, B ∈ B(X, Y, I) for some B ∈ LY }, i.e.
Ext(X, Y, I) consists of all extents of the concepts from B(X, Y, I). Now, it can
be shown (see [7, 8]) that for a fuzzy concept lattice B(X, Y, I), Ext(X, Y, I)
is an L-closure system in X and that for each L-closure system S in X
there are some X, Y , and I, such that S = Ext(X, Y, I). This means that
B(X, Y, I), ≈, ∧, ∨ , where (A1 , B1 ≈ A2 , B2 ) = (A1 ≈ A2 ) and ∧ and ∨
are the infimum and supremum induced by the corresponding operations in
Ext(X, Y, I), is an L-algebra (see Example 2.6).

Example 2.8. The following is a particular example of a fuzzy concept lattice.


Consider a complete residuated lattice on L = {0, 12 , 1} given by L  ukasiewicz
structure. Figure 2.4 contains a data table with fuzzy attributes describing a
fuzzy relation I between sets X (planets) and Y (fuzzy attributes of planets).
The corresponding fuzzy concept lattice contains 38 concepts. These are de-
picted in Fig. 2.6. The corresponding fuzzy concept lattice B(X, Y, I), ≤ is
shown in the left part of Fig. 2.5. Fuzzy equality ≈ on B(X, Y, I) (see Exam-
ple 2.7) is depicted in the right part of Fig. 2.5. Truth degrees 0, 12 , and 1 are

size distance
small (s) large (l) far (f) near (n)
Mercury (Me) 1 0 0 1
Venus (Ve) 1 0 0 1
Earth (Ea) 1 0 0 1
Mars (Ma) 1 0 1/2 1
Jupiter (Ju) 0 1 1 1/2

Saturn (Sa) 0 1 1 1/2

Uranus (Ur) 1/2 1/2 1 0


Neptune (Ne) 1/2 1/2 1 0
Pluto (Pl) 1 0 1 0

Fig. 2.4. Data table with fuzzy attributes


66 2 Algebras with Fuzzy Equalities
38

10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
1.
2.
3.
4.
5.
6.
7.
8.
9.
1.
2.
28 36 3.
4.
5.
6.
25 7.
8.
9.
37 10.
11.
27 13 33 12.

35 34 13.
14.
15.
22 16.
26 24 21 31 17.
18.
12 10 32 19.
30 20.
21.
11 20 22.
23 9 18 23.
24.
7 25.
17 29 26.
8 19 6 27.
28.
29.
30.
5 16 4 15 31.
32.
33.
34.
2 35.
3 14 36.
37.
38.

Fig. 2.5. Fuzzy concept lattice with L-equality

denoted by white, gray, and black boxes. Denote a concept no. i by Ci . Then,
for instance, C8 ≈ C2 = 0, C8 ≈ C3 = 12 , etc.
Note that [6] shows how the above-discussed fuzzy equality on fuzzy con-
cept lattice B(X, Y, I) can be used to form a factor lattice of B(X, Y, I). Ele-
ments of the factor lattice can be seen as similarity-based clusters of concepts
from B(X, Y, I). The factor lattice itself thus represents an approximate ver-
sion of B(X, Y, I). The clusters in our example are apparent from the right
part of Fig. 2.5. This is important from the application point of view since
B(X, Y, I) can be large and its factor lattice, which is smaller, is more com-
prehensible for a user. For more information see the last paragraph of Re-
mark 2.13.

Example 4.18 on page 183 shows a possible application of algebras with


fuzzy equalities in formal specification of humanistic systems. Further exam-
ples illustrating notions related to algebras with fuzzy equalities will be shown
in the course of the following sections.

2.2 Subalgebras, Congruences, and Morphisms

The goal of this section is to investigate subalgebras, morphisms, congru-


ences, and factor algebras, i.e. the very fundamental algebraic constructions.
All principal properties of such constructions, being developed in universal
algebra, will generalize for every complete residuated lattice L as a structure
of truth degrees.

Remark 2.9. Before delving into the basic structural notions, an important
remark is in order. As mentioned above, each algebra M = M, ≈M , F M
can be thought of as having two parts, a skeleton ske(M) (functional part)
2.2 Subalgebras, Congruences, and Morphisms 67

extent intent
no. Me Ve Ea Ma Ju Sa Ur Ne Pl s l f n
1. 0 0 0 0 0 0 0 0 0 1 1 1 1
2. 0 0 0 0 0 0 1/2 1/2 0 1 1 1 1/2
3. 0 0 0 1/2 0 0 0 0 0 1 1/2 1 1

4. 0 0 0 1/2 0 0 1/2 1/2 1/2 1 1/2 1 1/2

5. 1/2 1/2 1/2 1/2 0 0 0 0 0 1 1/2 1/2 1

6. 1/2 1/2 1/2 1/2 0 0 1/2 1/2 1/2 1 1/2 1/2 1/2

7. 0 0 0 1/2 0 0 1/2 1/2 1 1 0 1 0


8. 1/2 1/2 1/2 1 0 0 0 0 0 1 0 1/2 1
9. 1/2 1/2 1/2 1 0 0 1/2 1/2 1/2 1 0 1/2 1/2
10. 1/2 1/2 1/2 1 0 0 1/2 1/2 1 1 0 1/2 0
11. 1 1 1 1 0 0 0 0 0 1 0 0 1
12. 1 1 1 1 0 0 1/2 1/2 1/2 1 0 0 1/2
13. 1 1 1 1 0 0 1/2 1/2 1 1 0 0 0
14. 0 0 0 0 1/2 1/2 0 0 0 1/2 1 1 1
15. 0 0 0 0 1/2 1/2 1/2 1/2 0 1/2 1 1 1/2
16. 0 0 0 1/2 1/2 1/2 0 0 0 1/2 1/2 1 1

17. 0 0 0 1/2 1/2 1/2 1/2 1/2 1/2 1/2 1/2 1 1/2

18. 0 0 0 1/2 1/2 1/2 1 1 1/2 1/2 1/2 1 0

19. 1/2 1/2 1/2 1/2 1/2 1/2 0 0 0 1/2 1/2 1/2 1

20. 1/2 1/2 1/2 1/2 1/2 1/2 1/2 1/2 1/2 1/2 1/2 1/2 1/2

21. 1/2 1/2 1/2 1/2 1/2 1/2 1 1 1/2 1/2 1/2 1/2 0

22. 0 0 0 1/2 1/2 1/2 1 1 1 1/2 0 1 0


23. 1/2 1/2 1/2 1 1/2 1/2 0 0 0 1/2 0 1/2 1
24. 1/2 1/2 1/2 1 1/2 1/2 1/2 1/2 1/2 1/2 0 1/2 1/2
25. 1/2 1/2 1/2 1 1/2 1/2 1 1 1 1/2 0 1/2 0
26. 1 1 1 1 1/2 1/2 0 0 0 1/2 0 0 1
27. 1 1 1 1 1/2 1/2 1/2 1/2 1/2 1/2 0 0 1/2
28. 1 1 1 1 1/2 1/2 1 1 1 1/2 0 0 0
29. 0 0 0 0 1 1 1/2 1/2 0 0 1 1 1/2
30. 0 0 0 1/2 1 1 1/2 1/2 1/2 0 1/2 1 1/2

31. 0 0 0 1/2 1 1 1 1 1/2 0 1/2 1 0

32. 1/2 1/2 1/2 1/2 1 1 1/2 1/2 1/2 0 1/2 1/2 1/2

33. 1/2 1/2 1/2 1/2 1 1 1 1 1/2 0 1/2 1/2 0

34. 0 0 0 1/2 1 1 1 1 1 0 0 1 0
35. 1/2 1/2 1/2 1 1 1 1/2 1/2 1/2 0 0 1/2 1/2
36. 1/2 1/2 1/2 1 1 1 1 1 1 0 0 1/2 0
37. 1 1 1 1 1 1 1/2 1/2 1/2 0 0 0 1/2
38. 1 1 1 1 1 1 1 1 1 0 0 0 0

Fig. 2.6. Extracted concepts

 
and a set with L-equality M, ≈M (relational part) which are connected
via the compatibility condition. Traditional universal algebra deals with the
functional part only. However, it is obvious that when developing structural
properties of L-algebras, we will face situations where (1) only the functional
part is important, (2) only the relational part is important, (3) both the
68 2 Algebras with Fuzzy Equalities

functional and the relational parts are important. For (1), we can obviously
use well-known results established in universal algebra. However, since we
want our text to be self-contained, we omit the parts of proofs dealing with
the functional part only if they are easy to see.
 
Definition 2.10.
 Let M =
 M, ≈ M
, F M
be an L-algebra of type F . An L-
algebra N = N, ≈ , F
N N
is called a subalgebra of M if ∅ = N ⊆ M , each
function f N ∈ F N is a restriction of f M to N , and ≈N is a restriction of
≈M to N . A subuniverse of M is any subset N ⊆ M which is closed under
all functions of M.

That is, N ⊆ M is a subuniverse of M iff for each n-ary f M ∈ F M


and every a1 , . . . , an ∈ N we have f M (a1 , . . . , an ) ∈ N . The collection of
all subuniverses of M will be denoted by Sub(M). If ≈, F, σ is a type with
nullary function symbols (i.e., symbols of constants) and M is an L-algebra of
that type then each subuniverse N  ∈ Sub(M) is non-empty because for each
nullary function symbol c ∈ F , the corresponding constant cM ∈ M belongs
to N  . On the contrary, if ≈, F, σ does not contain nullary function symbols
one can easily see that ∅ ∈ Sub(M). Thus, ∅ ∈ Sub(M) if and only if M is an
L-algebra of a type without nullary function
 symbols.
An L-algebra M = M, ≈M , F M is said to be trivial if M = {a }. Since
the universe of a trivial L-algebra contains only one element, all mappings
{a }n → {a } are trivially compatible with ≈M . In the case of trivial L-
algebras, the L-equality ≈M does not represent any constraint for functions
f M and vice versa. An L-algebra which is not trivial is called non-trivial .
For convenience, we write N ∈ Sub(M) if N is a subalgebra of M. It is
almost immediate that there is a bijective correspondence between subalgebras
and non-empty subuniverses of M. Namely, a support set of a subalgebra of
M is a non-empty subuniverse and, conversely, a non-empty subuniverse of
M equipped with the restrictions of f M ’s and the restriction of ≈M is a
subalgebra of M.
It follows from the ordinary case that Sub(M) is closed under arbitrary
intersections and, therefore, it is a closure system under the subsethood rela-
tion ⊆. We denote the corresponding closure operator by [ ]M , i.e. for N ⊆ M ,
[N ]M is the least subuniverse of M containing N and we have

[N ]M = {N  | N  ∈ Sub(M) and N ⊆ N  } . (2.5)
If [N ]M is non-empty then the corresponding subalgebra of M is called the
subalgebra of M generated by N . If |N | < κ for an infinite cardinal κ then the
corresponding subalgebra of M is called κ-generated. If N is finite then the
corresponding subalgebra of M is called finitely generated. Hence, a subalgebra
is finitely generated if and only if it is ω-generated (ω is the least infinite
cardinal).
∗ ∗ ∗
2.2 Subalgebras, Congruences, and Morphisms 69

In ordinary case, congruences correspond to abstract views on algebraic


systems: one can put congruent (i.e. equivalent from a certain abstract point
of view) elements together and form a factor algebra which provides a view on
the original algebra under which congruent elements are indistinguishable. In
fuzzy setting, it is natural to consider graded indistinguishability of elements.
This leads to the notion of a congruence on an L-algebra.

Definition 2.11. Let M be an L-algebra of type F . An L-relation θ on M is


called a congruence on M if
(i) θ is an L-equivalence relation on M ,
(ii) θ is compatible with ≈M ,
(iii) all functions f M ∈ F M are compatible with θ, see Definition 1.80 and
Remark 1.81.

Remark 2.12. (1) In what follows, the (ordinary) sets of all L-equivalences
and congruences on an L-algebra M are denoted by EqL (M ) and ConL (M).
Evidently, ConL (M) ⊆ EqL (M ).
(2) It is immediate that congruences on 2-algebras correspond to ordi-
nary congruences on algebras since condition (ii) of Definition 2.11 is trivially
satisfied by ordinary congruence relations, generalizing thus the well-known
concept of a congruence.
(3) Note that due to Lemma 1.82, condition (ii) may be equivalently re-
placed by requiring ≈M ⊆ θ, i.e. (a ≈M b ) ≤ θ(a , b ) for arbitrary a , b ∈ M .

Remark 2.13. Let L be a complete residuated lattice with ⊗ being ∧ (i.e.


L is a complete Heyting algebra). If L = [0, 1], this means that ⊗ is the
minimum which is probably the most common choice of “fuzzy conjunction”.
It is straightforward to verify that a binary L-relation on M is an L-equality
if and only if each a-cut a θ, a ∈ L is an ordinary equivalence relation and 1 θ
is the identity relation.
Furthermore, for an L-algebra M, θ satisfies the compatibility condition
w.r.t. the operations of M if and only if each a θ satisfies the compatibility  
condition w.r.t. operations of the corresponding ordinary algebra M, F M .
Indeed, for a1 , b1 ∈ a θ, . . . , an , bn ∈ a θ we have a ≤ θ(ai , bi ) for every
i = 1, . . . , n, thus
 
a ≤ θ(a1 , b1 ) ⊗ · · · ⊗ θ(an , bn ) ≤ θ f M (a1 , . . . , an ), f M (b1 , . . . , bn )
 
by the assumption ⊗ = ∧, i.e. f M (a1 , . . . , an ), f M (b1 , . . . , bn ) ∈ a θ. Con-
versely, put a = θ(a1 , b1 ) ⊗  ·M · · ⊗ θ(an , bn ). Then we have ai , bi ∈ a θ for
every i = 1, . . . , n, and so f (a1 , . . . , an ), f M (b1, . . . , bn ) ∈ a θ. Hence, it
follows that a ≤ θ f M (a1 , . . . , an ), f M (b1 , . . . , bn ) .
Therefore, for a congruence θ on an L-algebra M, Sθ = {a θ | a ∈ L} is a
system of binary relations on M indexed by L which satisfies: (i) a ≤ b implies
b
θ ⊆ a θ; (ii) for each a , b ∈ M there exists a greatesta ∈ L with a , b ∈ a θ;
(iii) for each a ∈ L, a θ is a congruence on M, F M containing a ≈M . One
70 2 Algebras with Fuzzy Equalities

may verify that conversely, if S = {θa | a ∈ L} is a system of binary relations


on M satisfying (i)–(iii) then putting θS (a , b ) = {a | a , b ∈ θa }, θS is a
congruence on M. Furthermore, θ = θSθ and S = SθS .
This way of looking at congruences of L-algebras (but only when ⊗ is ∧!)
gives the following interpretation: An ordinary congruence on an ordinary
algebra may be thought of as representing an abstract view of the algebra – a
view under which one does not distinguish congruent elements. A congruence θ
on an L-algebra may be thought of as a hierarchic system (the hierarchy 
supplied by L) of abstract views a θ. Since each

a
θ is a congruence on M, F M ,
one may form the ordinary factor algebra M, F M /a θ. Clearly, the smaller a,
the bigger a θ, and so the coarser the factorization.
This fact has non-trivial applications. In [6], a method of factorization of
concept lattices is presented. As shown in Example 2.7, a fuzzy concept lat-
tice B(X, Y, I) can be naturally turned into an L-algebra B(X, Y, I), ≈, ∧, ∨ .
A fuzzy concept lattice B(X, Y, I) represents a set of natural clusters (con-
cepts) contained in the input data X, Y, I . If there are too many concepts
in B(X, Y, I), the fuzzy concept lattice is hardly graspable by human mind.
If ⊗ is ∧, one can form the factor lattice B(X, Y, I)/a ≈ for any a. The factor
lattice B(X, Y, I)/a ≈ may be thought of as a simplified version of B(X, Y, I).
The choice of a controls the coarseness of the factorization. The point of [6] is
that, in fact, one may form the factor lattice even for general ⊗ (not necessar-
ily ∧) in which case a ≈ is a tolerance relation (i.e., reflexive and symmetric)
which is compatible with the lattice structure of B(X, Y, I). Since (complete)
lattices can be factorized even by compatible tolerance relations, a method for
general ⊗ is available, see [6]. It is also worth to note that the factor lattice
can be computed directly from the data table, without the need to compute
the whole B(X, Y, I). For details we refer to [13].

Example 2.14. Consider a fuzzy concept lattice B(X, Y, I) with fuzzy equal-
ity from Example 2.8 (see Fig. 2.5). Figure 2.7 depicts six congruences on
B(X, Y, I). Again, truth degrees 0, 12 , and 1 are denoted by white, gray, and
black boxes. Notice that the congruence from part (a) is a fuzzy equality
while those from (b)–(f) are not. Intuitively, one can see that each congruence
induces clusters of concepts from B(X, Y, I).

By definition, ConL (M) ⊆ EqL (M ). The following theorem shows that


ConL (M) inherits even the complete lattice structure of EqL (M ).

Theorem 2.15. ConL (M), ⊆ is a complete sublattice of EqL (M ), ⊆ .

 set I and a family {θi | θi ∈ ConL (M),


Proof. Take an index  i ∈ I}  of congru-
ences. Obviously, i∈I θi ∈ EqL (M ) and (a ≈M b ) ≤ i∈I θi (a , b ). For
every f M ∈ F M and arbitrary a1 , b1 , . . . , an , bn ∈ M , we have,
n    n 
j=1 i∈I θi (aj , bj ) ≤ i∈I j=1 θi (aj , bj ) ≤
  
≤ i∈I θi f M (a1 , . . . , an ), f M (b1 , . . . , bn ) ,
2.2 Subalgebras, Congruences, and Morphisms 71

10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.

10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
1.
2.
3.
4.
5.
6.
7.
8.
9.

1.
2.
3.
4.
5.
6.
7.
8.
9.
1. 1.
2. 2.
3.
4.
5.
(a) 3.
4.
5.
(b)
6. 6.
7. 7.
8. 8.
9. 9.
10. 10.
11. 11.
12. 12.
13. 13.
14. 14.
15. 15.
16. 16.
17. 17.
18. 18.
19. 19.
20. 20.
21. 21.
22. 22.
23. 23.
24. 24.
25. 25.
26. 26.
27. 27.
28. 28.
29. 29.
30. 30.
31. 31.
32. 32.
33. 33.
34. 34.
35. 35.
36. 36.
37. 37.
38. 38.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.

10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
1.
2.
3.
4.
5.
6.
7.
8.
9.

1.
2.
3.
4.
5.
6.
7.
8.
9.
1. 1.
2. 2.
3.
4.
5.
(c) 3.
4.
5.
(d)
6. 6.
7. 7.
8. 8.
9. 9.
10. 10.
11. 11.
12. 12.
13. 13.
14. 14.
15. 15.
16. 16.
17. 17.
18. 18.
19. 19.
20. 20.
21. 21.
22. 22.
23. 23.
24. 24.
25. 25.
26. 26.
27. 27.
28. 28.
29. 29.
30. 30.
31. 31.
32. 32.
33. 33.
34. 34.
35. 35.
36. 36.
37. 37.
38. 38.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.

10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
1.
2.
3.
4.
5.
6.
7.
8.
9.

1.
2.
3.
4.
5.
6.
7.
8.
9.

1. 1.
2. 2.
3.
4.
5.
(e) 3.
4.
5.
(f)
6. 6.
7. 7.
8. 8.
9. 9.
10. 10.
11.
12.
Fig. 2.7. Examples of congruences
11.
12.
13. 13.
14. 14.
15. 15.
16. 16.
17. 17.
18. 18.
19. 19.
20. 20.
21. 21.
22. 22.
23. 23.
24. 24.
25. 25.
26. 26.
27. 27.
28. 28.
29. 29.
30. 30.
31. 31.
32. 32.
33. 33.
34. 34.
35. 35.
36. 36.
37. 37.
38. 38.

Fig. 2.7. Examples of congruences


72 2 Algebras with Fuzzy Equalities

thus every f M ∈ F M is compatible with i∈I θi . Hence, ConL (M), ⊆ is
closed under arbitrary infima showing that infima in ConL (M), ⊆ coincide
with infima in EqL (M ), ⊆ .
To complete the proof we show that ConL (M), ⊆ and EqL (M ), ⊆ agree
on
 their suprema. First note that the supremum of θi (i ∈ I) in EqL (M ) equals
θ
i1 ,...,ik ∈I i1 ◦ · · · ◦ θ ik . Indeed, denote i1 ,...,ik ∈I θi1 ◦ · · · ◦ θik by θ. One can
see that each θi , i ∈ I is contained in θ (i.e. θi ⊆ θ), θ is contained in any L-
equivalence containing all θi ’s, and that θ is reflexive (due to reflexivity of θi )
and symmetric (due to symmetry of θi and commutativity of ⊗). Moreover,
θ(a , b ) ⊗ θ(b , c ) =
    
= i1 ,...,ik ∈I θi1 ◦ · · · ◦ θik (a , b ) ⊗ θj1 ◦ · · · ◦ θjl (b , c ) ≤
j1 ,...,jl ∈I
    
≤ i1 ,...,ik ∈I θi1 ◦ · · · ◦ θik ◦ θj1 ◦ · · · ◦ θjl (a , c ) ≤
j1 ,...,jl ∈I
  
≤ i1 ,...,im ∈I θi1 ◦ · · · ◦ θim (a , c ) = θ(a , c ) ,
thus θ is also transitive showing that θ is the supremum of θi ’s in EqL (M ).
Now it suffices to show that θ ∈ ConL (M). By Lemma 1.82, condition (ii) of
Definition 2.11 is satisfied trivially. For (i) take any n-ary function f M ∈ F M
and arbitrary a1 , b1 , . . . , an , bn ∈ M . We have
θ(a1 , b1 ) ⊗ · · · ⊗ θ(an , bn ) =
n     
= j=1 ij1 ,...,ijk ∈I θij1 ◦ · · · ◦ θijkj (aj , bj ) =
  n 
j

= i11, ...,i1k1 ∈I c11, ...,c1(k1 −1) ∈M j=1 θij1 (aj , cj1 ) ⊗ θij2 (cj1 , cj2 ) ⊗ · · ·
.. .. .. ..
. .
in1,...,inkn ∈I
. .
cn1,...,cn(kn −1) ∈M 
· · · ⊗ θijkj (cj(kj −1) , bj ) .
Now for any of θij1 (aj , cj1 ), θij2 (cj1 , cj2 ), . . . , θijkj (cj(kj −1) , bj ), j ∈ J we have
 
θi11 (a1 , c11 ) ≤ θi11 f M (a1 , . . . , an ), f M (c11 , a2 , . . . , an ) ,
..
.  
θin1 (an , cn1 ) ≤ θin1 f M (c11 , c21 . . . , c(n−1)1 , an ), f M (c11 , . . . , cn1 ) ,
 
θi12 (c11 , c12 ) ≤ θi12 f M (c11 , . . . , cn1 ), f M (c12 , c21 , . . . , cn1 ) ,
..
.  
θinkn (cn(kn −1) , bn ) ≤ θinkn f M (b1 , . . . , b(n−1) , cn(kn −1) ), f M (b1 , . . . , bn ) .
Hence, we finish the proof with
  n 
i11, ...,i1k1 ∈I c11, ...,c1(k1 −1) ∈M j=1 θij1 (aj , cj1 ) ⊗ θij2 (cj1 , cj2 ) ⊗ · · ·
.. .. .. ..
. .
in1,...,inkn ∈I
. .
cn1,...,cn(kn −1) ∈M 
· · · ⊗ θijkj (cj(kj −1) , bj ) ≤
   
≤ l1 ,...,lm ∈I θl1 ◦ · · · ◦ θlm f (a1 , . . . , an ), f (b1 , . . . , bn )
M M
=
 M 
= θ f (a1 , . . . , an ), f M (b1 , . . . , bn ) .
Altogether, θ ∈ ConL (M). 

2.2 Subalgebras, Congruences, and Morphisms 73

Lemma 2.16. For congruences θ1 , θ2 ∈ ConL (M) the following conditions


are equivalent:
(i) θ1 ◦ θ2 = θ2 ◦ θ1 ,
(ii) θ1 ∨ θ2 = θ1 ◦ θ2 ,
(iii) θ2 ◦ θ1 ⊆ θ1 ◦ θ2 .

Proof. “(i) ⇒ (iii)” is trivial.


“(iii) ⇒ (ii)”:We only need to show θ1 ∨ θ2 ⊆ θ1 ◦ θ2 which follows easily
from θ1 ∨ θ2 = i1 ,...,ik ∈{1,2} θi1 ◦ · · · ◦ θik and from θi ◦ θi ⊆ θi , the latter
relationship is a reformulation of transitivity.
“(ii) ⇒ (i)”: It is easily seen that θ1 ◦ θ2 = θ1 ∨ θ2 = (θ1 ∨ θ2 )−1 =
(θ1 ◦ θ2 )−1 = θ2−1 ◦ θ1−1 = θ2 ◦ θ1 . 


An important notion is that of a congruence generated by a finite collec-


tion of pairs. A natural meaning of this notion is the following: We are given a
collection of pairs ai , bi of elements of M and are looking for the least con-
gruence containing all ai , bi ’s. In other words, we are looking for the least
abstract view on M under which all ai and bi are indistinguishable. In fuzzy
setting, it is natural to provide additionally the degrees of required indistin-
guishability of ai and bi . In what follows, we present tractable descriptions
of congruences generated by prescribed pairs of elements and their required
indistinguishability degrees.

Definition 2.17. Let M be an L-algebra. For a binary L-relation R in M


we denote by θ(R) the least congruence on M containing R. Particularly,
for any b1 , c1 , . . . , bk , ck ∈ M and arbitrary truth degrees a1 , . . . , ak ∈ L
we denote θ({a1/b1 , c1 ,. . ., ak/bk , ck }) by θ(a1/b1 , c1 , . . . , ak/bk , ck ). It
is the least congruence θ ∈ ConL (M) such that θ(b1 , c1 ) ≥ a1 , . . . , θ(bk , ck ) ≥
ak , i.e. θ(a1/b1 , c1 , . . . , ak/bk , ck ) is generated by the finite L-relation R =
{a1/b1 , c1 , . . . , ak/bk , ck }. A congruence θ(a/b , c ) ∈ ConL (M) is called a
principal congruence on M.

Remark 2.18. Note that due toTheorem 2.15, θ(R) exists for every L-relation
R ∈ LM ×M . Namely, θ(R) = {θ ∈ ConL (M) | R ⊆ θ}.

Example 2.19. Consider the congruences from Example 2.14, see Fig. 2.7. (a) is
θ(0.5/C1 , C38 ), (b) is θ(1/C15 , C29 ), (c) is θ(1/C8 , C12 ), (d) is θ(1/C2 , C7 ),
(e) is θ(1/C1 , C14 ), and (f) is θ(1/C1 , C11 ).

Lemma 2.20. The following are properties of principal congruences:


(i) θ(a/b , c ) = θ(a/c , b ),
b1 , c1 , . . . , /bk , ck ) =θ( /b1 , c1 ) ∨ · · · ∨ θ( /bk , ck ),
(ii) θ(a1/ ak a1 ak

(iii) θ = b ,c ∈ M θ(θ(b ,c )/b , c ) = b ,c ∈ M θ(θ(b ,c )/b , c ).


74 2 Algebras with Fuzzy Equalities

Proof. (i) follows directly from the definition.


(ii): It is easily seen that θ(a1/b1 , c1 , . . . , ak/bk , ck ) is always contained
in θ(a1/b1 , c1 ) ∨ · · · ∨ θ(ak/bk , ck ). For the converse inequality observe
that θ(ai/bi , ci ) ⊆ θ(a1/b1 , c1 , . . . , ak/bk , ck ) for every i = 1, . . . , k. Thus,
property (ii) holds.
(iii): We have
 
θ(b ,c )/b , c ) ⊆ θ(b ,c )/b , c ) ⊆ θ ⊆
b ,c ∈M θ( b ,c ∈M θ(

⊆ b ,c ∈M θ(θ(b ,c )/b , c )
proving the claim. 

Remark 2.21. The so-called Mal’cev lemma provides a description of principal
congruences. Mal’cev lemma can be generalized to fuzzy setting. However, it
involves notions concerning terms, term functions and term algebras which
are discussed in Sect. 2.4. Therefore, we postpone the generalized Mal’cev
lemma and present it on page 94 as Theorem 2.82.
We are going to define the concept of a factor L-algebra by a congruence θ.
For this purpose, we utilize the concept of a factor set with fuzzy equality by
a fuzzy equivalence (Definition 1.91).
Definition 2.22. Let θ be a congruence on an L-algebra  M of type F. A
factor L-algebra of M by θ is an L-algebra M/θ = M/θ, ≈M/θ , F M/θ of
type F such that
   
(i) M/θ, ≈M/θ is the factor set with L-equality of M, ≈M by θ,
see Definition 1.91 and Remark 1.92;
   
(ii) f M/θ [a1 ]θ , . . . , [an ]θ = f M (a1 , . . . , an ) θ
for every n-ary function f M/θ ∈ F M/θ and arbitrary a1 , . . . , an ∈ M .

Remark 2.23. (1) Condition (i) says that elements of M/θ are ordinary equiv-
alence classes of 1 θ. By convention, these classes are denoted by [a ]θ instead
of [a ]1 θ . That is, M/θ = {[a ]θ | a ∈ M } where [a ]θ = {a  | θ(a , a  ) = 1}. Fur-
thermore, condition (i) says that ≈M/θ is defined by [a ]θ ≈M/θ [b ]θ = θ(a , b ).
(2) A factor L-algebra is well-defined. First, ≈M/θ is a well-defined L-
equality on factor
 set M/θ. Namely, since θ is a congruence on M =
M, ≈M , F M , θ is compatible with ≈M , see Remark 2.12 (3). Therefore,
≈M/θ is a well-defined L-equality on M/θ by Lemma 1.90. Second, each
f M/θ ∈ F M/θ is a well-defined function on M/θ which is compatible with
≈M/θ . Namely, for each n-ary function f M/θ ∈ F M/θ and a1 , b1 , . . . , an , bn ∈
M , we have
[a1 ]θ ≈M/θ [b1 ]θ ⊗ · · · ⊗ [an ]θ ≈M/θ [bn ]θ = θ(a1 , b1 ) ⊗ · · · ⊗ θ(an , bn ) ≤
 
≤ θ f M (a1 , . . . , an ), f M (b1 , . . . , bn ) =
   
= f M (a1 , . . . , an ) θ ≈M/θ f M (b1 , . . . , bn ) θ =
   
= f M/θ [a1 ]θ , . . . , [an ]θ ≈M/θ f M/θ [b1 ]θ , . . . , [bn ]θ .
2.2 Subalgebras, Congruences, and Morphisms 75

Applying this inequality for [a1 ]θ ≈M/θ [b1 ]θ = 1, . . . , [an ]θ ≈M/θ [bn ]θ = 1
yields that if [a1 ]θ = [b1 ]θ , . . . , [an ]θ = [bn ]θ , then f M/θ ([a1 ]θ , . . . , [an ]θ ) =
f M/θ ([b1 ]θ , . . . , [bn ]θ ). Hence, every function f M/θ ∈ F M/θ is well defined.
Furthermore, the above inequality itself says that every f M/θ ∈ F M/θ is
compatible with ≈M/θ .

Remark 2.24. An important role in the definition of a factor L-algebra is


played by the 1-cut 1 θ of a congruence θ on M. It is important to note that
in general, a congruence θ on an L-algebra is not determined by 1 θ, not even
if θ is an L-equality (in fact, all L-equalities on M have a common 1-cut of
the form {a , a | a ∈ M }). Furthermore, one can easily verify that for a con-

gruence θ ∈ ConL (M), 1 θ is an ordinary congruence relation on M, F M .
Therefore, the functional part of M/θ coincides with M, F M /1 θ. Clearly,
this issue is completely degenerate in the ordinary case since if L = 2 then
θ and 1 θ coincide (modulo the relationship between crisp fuzzy relations and
the corresponding ordinary relations).

Example 2.25. Fig. 2.8 contains the skeletons of factor lattices with fuzzy
equalities which result by factorization of a fuzzy concept lattice with fuzzy
equality from Example 2.8 by congruences (a)–(f) from Example 2.14. Con-
sider in more detail case (e) and denote the corresponding congruence by θ.
The elements of the factor L-algebra (factor lattice with fuzzy equality) are
the following classes of θ:

(a) (b) (c)

(d) (e) (f)

Fig. 2.8. Examples of factor L-algebras


76 2 Algebras with Fuzzy Equalities

Fig. 2.9. Factor lattice with L-equality

[C1 ]θ = {C1 , C14 } , [C2 ]θ = {C2 , C15 } ,


[C3 ]θ = {C3 , C16 } , [C4 ]θ = {C4 , C7 , C17 , C18 , C22 } ,
[C5 ]θ = {C5 , C8 , C19 , C23 } , [C6 ]θ = {C6 , C9 , C10 , C20 , C21 , C24 , C25 } ,
[C11 ]θ = {C11 , C26 } , [C12 ]θ = {C12 , C13 , C27 , C28 } ,
[C29 ]θ = {C29 } , [C30 ]θ = {C30 , C31 , C34 } ,
[C32 ]θ = {C32 , C33 , C35 , C36 } , [C37 ]θ = {C37 , C38 } .
The corresponding factor lattice is shown (once again) in the left part of
Fig. 2.9. The fuzzy equality on the factor lattice is shown in the right part of
Fig. 2.9.

∗ ∗ ∗
In the sequel, a morphism of L-algebras is defined as a mapping that
preserves both their operations f M and their fuzzy equality ≈M . All properties
well-known from ordinary case generalize in full scope.

Definition 2.26. Let M and N be L-algebras of type F . A mapping h : M →


N is called a morphism (or homomorphism) of M to N if
(i) h(f M (a1 , . . . , an )) = f N (h(a1 ), . . . , h(an ))
for every n-ary f ∈ F and arbitrary   a1 , . . . , an ∈ M ;
(ii) h is a ≈-morphism of M, ≈M to N, ≈N (Definition 1.83 ).
The fact that h : M → N is a morphism is denoted by h : M → N.
Furthermore,
• an injective morphism is called a monomorphism,
• a morphism such that
a ≈M b = h(a ) ≈N h(b ) for all a , b ∈ M (2.6)
is called an embedding,
• a surjective morphism is called an epimorphism,
2.2 Subalgebras, Congruences, and Morphisms 77

• an epimorphism which is an embedding is called an isomorphism,


• a morphism h : M → M is called an endomorphism,
• an isomorphism h : M → M is called an automorphism.
For an epimorphism h : M → N, the L-algebra N is called an image of M.
We say M is isomorphic to N, written M ∼ = N, if there exists an isomor-
phism h : M → N. Let idM denote the identity mapping on M (we use idM
to emphasize that M is a universe of M).

Remark 2.27. Note that condition (ii) requires a ≈M b ≤ h(a ) ≈N h(b )


which can be verbally described as “if a and b are similar, then h(a ) and
h(b ) are similar”. Namely, a ≈M b ≤ h(a ) ≈N h(b ) is true for all a , b ∈ M
iff formula
 
(∀x)(∀y) x ≈ y i h(x) ≈ h(y)
is fully true (i.e. has truth degree 1) in an appropriate structure encompass-
ing both M and N. Condition h(f M (a1 , . . . , an )) = f N (h(a1 ), . . . , h(an )) is
equivalent to saying that h is an (ordinary) morphism of the skeletons ske(M)
and ske(N).

Theorem 2.28. Let h : M → N be a morphism. Then we have


(i) if M  ∈ Sub(M) then h(M  ) ∈ Sub(N),
(ii) if N  ∈ Sub(N) then h−1 (N  ) = {a ∈ M | h(a ) ∈ N  } ∈ Sub(M),
(iii) if h : M → N is an embedding then M ∼ = h(M),
where h(M) is a subalgebra of N induced by h(M ).

Proof. (i): Take an n-ary f N ∈ F M . For arbitrary b1 , . . . , bn ∈ h(M  ) there


are c1 , . . . , cn ∈ M  such that b1 = h(c1 ), . . . , bn = h( 
n ). Since M ∈
 cM  Sub(M),
we get f (b1 , . . . , b1 ) = f h(c1 ), . . . , h(c1 ) = h f (c1 , . . . , cn ) ∈ h(M  ).
N N

Furthermore, each f N is trivially compatible with the restriction of ≈N on


h(M  ). Hence, h(M  ) ∈ Sub(N).
(ii): Take an n-ary f M ∈ F M . For each a1 , . . . , an ∈ h−1 (N  ) we have
   
h f M (a1 , . . . , an ) = f N h(a1 ), . . . , h(an ) ∈ N 
because h(a1 ), . . . , h(an ) ∈ N  . Therefore, f M (a1 , . . . , an ) ∈ h−1 (N  ). Evi-
dently, each f M is compatible with the restriction of ≈M on h−1 (N  ).
(iii): Suppose h : M → N is an embedding. (i) yields that h(M ) ∈ Sub(N).
Thus, g : M → h(M), where g(a ) = h(a ) for every a ∈ M is a surjective
embedding, i.e. M ∼ = h(M). 


Theorem 2.29. Suppose g : M → M and h : M → M are morphisms. The


composition g ◦ h is a morphism (g ◦ h) : M → M .

Proof. Theorem 1.86 yields that h is an ≈-morphism. The rest follows from
the ordinary case. 

78 2 Algebras with Fuzzy Equalities

Theorem 2.30. Let h : M → M , g : M → M be morphisms such that


h ◦ g = idM , g ◦ h = idM . Then M ∼
= M .

Proof. Evidently, h is bijective since for every a ∈ M  we have h(g(a )) = a ,


and from h(a ) = h(b ) it follows that a = g(h(a )) = g(h(b )) = b . Clearly,
h−1 = g since for h(a ) = b we have g(b ) = g(h(a )) = a = h−1 (b ). Hence,

we can apply Theorem 1.86 (ii) to see that a ≈M b = h(a ) ≈M h(b ) for all
a , b ∈ M . Thus, h is an isomorphism. 


Theorem 2.31. Let h : M → N be a morphism. Then θh ∈ ConL (M).

Proof. From Theorem 1.85 and Theorem 1.87 it follows that a ≈M b ≤


θh (a , b ) for all a , b ∈ M and θh ∈ EqL (M ). Now it is sufficient to show that
θh is also compatible with functions f M ∈ F M . For n-ary f M ∈ F M and
arbitrary a1 , b1 , . . . , an , bn ∈ M , the following inequality holds:
n n
i=1 h(ai ) ≈ h(bi ) ≤
N
i=1 θh (ai , bi ) =
   
≤ f N h(a1 ), . . . , h(an ) ≈N f N h(b1 ), . . . , h(bn ) =
   
= h f M (a1 , . . . , an ) ≈N h f M (b1 , . . . , bn ) =
 
= θh f M (a1 , . . . , an ), f M (b1 , . . . , bn ) .
Thus, θh ∈ ConL (M). 


Definition 2.32. For every L-algebra M and θ ∈ ConL (M) a mapping hθ :


M → M/θ, where hθ (a ) = [a ]θ for all a ∈ M , is called a natural mapping.

Theorem 2.33. A natural mapping hθ from an L-algebra M to a factor L-


algebra M/θ is an onto morphism (i.e. an epimorphism).

Proof. For any a , b ∈ M we have


a ≈M b ≤ θ(a , b ) = [a ]θ ≈M/θ [b ]θ = hθ (a ) ≈M/θ hθ (b ) .
Furthermore, for n-ary f M ∈ F M and arbitrary a1 , . . . , an ∈ M we have
   
hθ f M (a1 , . . . , an ) = f M (a1 , . . . , an ) θ =
   
= f M/θ [a1 ]θ , . . . , [an ]θ = f M/θ hθ (a1 ), . . . , hθ (an ) .
Surjectivity of hθ is evident. 


Remark 2.34. In what follows we call hθ a natural morphism.

In the rest of this section we present some of the basic results on relation-
ships between morphisms and congruence relations in fuzzy setting. Note that
in a more general setting, some of the results are presented in [10].

Theorem 2.35 (first isomorphism theorem). Let h : M → N be a sur-


jective morphism. Then there is an isomorphism g : M/θh → N such that
hθh ◦ g = h.
2.2 Subalgebras, Congruences, and Morphisms 79

Fig. 2.10. First isomorphism theorem

g1
M N1

g2 d g1

N2 M
g2

Fig. 2.11. Diagonal fill-in

 
Proof. Introduce g by putting g [a ]θh = h(a ) for all elements a ∈ M . Evi-
dently, hθh ◦ g = h. Furthermore,
   
[a ]θh ≈M/θh [b ]θh = θh (a , b ) = h(a ) ≈N h(b ) = g [a ]θh ≈N g [b ]θh .
 
The mapping g is surjective as h is surjective and g [a ]θh = h(a ). Take any
n-ary f M/θh ∈ F M/θh and arbitrary [a1 ]θh , . . . , [an ]θh ∈ M/θh . We have
      
g f M/θh [a1 ]θh , . . . , [an ]θh = g f M a1 , . . . , an θ =
   h

= h f M (a1 , . . . , an ) = f N h(a1 ), . . . , h(an ) =
    
= f N g [a1 ]θh , . . . , g [an ]θh .
Thus g : M/θh → N is an isomorphism and hθh ◦ g = h, see Fig. 2.10. 


An easy extension of Theorem 2.35 says that every morphism h : M → M


can be expressed as h = g ◦ g  , where g : M → N is an epimorphism and
g  : N → M is an embedding. The following theorem deals in more detail
with this kind of decomposition.

Theorem 2.36. Let h : M → M be a morphism. Moreover, let g1 : M → N1 ,


g2 : M → N2 be surjective morphisms and let g1 : N1 → M , g2 : N2 → M
be embeddings such that h = g1 ◦ g1 = g2 ◦ g2 . Then there exists a unique
morphism d : N1 → N2 such that g1 = d ◦ g2 , g2 = g1 ◦ d.

Proof. Let us introduce a mapping d : N1 → N2 defined for a ∈ N1 by


d(a ) = g2 (b ), where b ∈ M, g1 (b ) = a . (2.7)
In ordinary case, d is called a diagonal fill-in [97]. First, we prove that d is a
well-defined morphism. Due to the surjectivity of g1 , for every a ∈ N1 there
80 2 Algebras with Fuzzy Equalities

is some b ∈ M such that g1 (b ) = a . Clearly, for g1 (b ) = g1 (b  ) = a we


have g1 (g1 (b )) = g1 (g1 (b  )), thus it follows that g2 (g2 (b )) = g2 (g2 (b  )) since
g1 ◦ g1 = g2 ◦ g2 . But g2 is an embedding, that is g2 (b ) = g2 (b  ), i.e. d is a
well-defined mapping.
For every n-ary f N1 ∈ F N1 and arbitrary elements a1 , . . . , an ∈ N1 there
are b1 , . . . , bn ∈ M such that g1 (bi ) = ai for i = 1, . . . , n. Hence,
   
f N2 d(a1 ), . . . , d(an ) = f N2 g2 (b1 ), . . . , g2 (bn ) =
   
= g2 f M (b1 , . . . , bn ) = d f N1 (a1 , . . . , an )
   
since g1 f M (b1 , . . . , bn ) = f N1 g1 (b1 ), . . . , gn (bn ) = f N1 (a1 , . . . , an ), i.e.
d is compatible with f N1 ∈ F N1 .
Take a1 , a2 ∈ N1 . Thus, there are b1 , b2 ∈ M such that g1 (b1 ) = a1 ,
g1 (b2 ) = a2 . It readily follows that

a1 ≈N1 a2 = g1 (b1 ) ≈N1 g1 (b2 ) = g1 (g1 (b1 )) ≈M g1 (g1 (b2 )) =

= g2 (g2 (b1 )) ≈M g2 (g2 (b2 )) = g2 (b1 ) ≈N2 g2 (b2 ) =
= d(a1 ) ≈N2 d(a1 ) ,
i.e. d is an embedding. Furthermore, for a , a  ∈ N1 , b , b  ∈ M such that
a = g1 (b ), a  = g1 (b  ) we have
g1 (a ) = g1 (g1 (b )) = g2 (g2 (b )) = g2 (d(a )) = (d ◦ g2 )(a ) ,
g2 (b  ) = d(a  ) = d(g1 (b  )) = (g1 ◦ d)(b  ) ,
i.e. d is a morphism satisfying the required conditions. The situation is de-
picted in Fig. 2.7.
It suffices to check the uniqueness of d. So, let d : N1 → N2 be a morphism
such that g1 = d ◦ g2 , g2 = g1 ◦ d . Clearly, d ◦ g2 = g1 = d ◦ g2 . Since g2 is
an embedding, we have d = d . 

Definition 2.37. Suppose M is an L-algebra and φ, θ ∈ ConL (M), θ ⊆ φ.
Then we let φ/θ denote an L-relation on M/θ defined by
 
(φ/θ) [a ]θ , [b ]θ = φ(a , b ) (2.8)
for all a , b ∈ M .
Theorem 2.38. Let φ, θ ∈ ConL (M), θ ⊆ φ. Then φ/θ ∈ ConL (M/θ).
Proof. Clearly, φ/θ is an L-equivalence. For every a , b ∈ M we have
 
[a ]θ ≈M [b ]θ = θ(a , b ) ≤ φ(a , b ) = (φ/θ) [a ]θ , [b ]θ .
Take n-ary f M/θ ∈ F M/θ and [a1 ]θ , [b1 ]θ . . . , [an ]θ , [bn ]θ ∈ M/θ. We have
n   n
i=1 (φ/θ) [ai ]θ , [bi ]θ = i=1 φ(ai , bi ) ≤
 M 
≤ φ f (a1 , . . . , an ), f M (b1 , . . . , bn ) =
    
= (φ/θ) f M (a1 , . . . , an ) θ , f M (b1 , . . . , bn ) θ =
 
= (φ/θ) f M/θ ([a1 ]θ , . . . , [an ]θ ), f M/θ ([b1 ]θ , . . . , [bn ]θ ) .
Altogether, φ/θ ∈ ConL (M/θ). 

2.2 Subalgebras, Congruences, and Morphisms 81

Theorem 2.39 (second isomorphism theorem). Suppose M is an L-


 θ ⊆ φ. Then the mapping h : (M/θ)/(φ/θ) →
algebra and φ, θ ∈ ConL (M),
M/φ defined by h [[a ]θ ]φ/θ = [a ]φ is an isomorphism.

Proof. For every [[a ]θ ]φ/θ , [[b ]θ ]φ/θ ∈ (M/θ)/(φ/θ) we have,


 
[[a ]θ ]φ/θ ≈(M/θ)/(φ/θ) [[b ]θ ]φ/θ = (φ/θ) [a ]θ , [b ]θ =
= φ(a , b ) = [a ]φ ≈M/φ [b ]φ .
 
Since for all a ∈ M we have h [[a ]θ ]φ/θ = [a ]φ , h is a surjective ≈-morphism.
Take any n-ary function f (M/θ)/(φ/θ) ∈ F (M/θ)/(φ/θ) and arbitrary elements
[[a1 ]θ ]φ/θ , . . . , [[an ]θ ]φ/θ ∈ (M/θ)/(φ/θ).
      
h f (M/θ)/(φ/θ) [[a1 ]θ ]φ/θ , . . . ,[[an ]θ ]φ/θ = h f M/θ [a1 ]θ , . . . ,[an ]θ φ/θ =
    
= h [f M (a1 , . . . , an )]θ φ/θ = f M (a1 , . . . , an ) φ =
      
= f M/φ [a1 ]φ , . . . , [an ]φ = f M/φ h [[a1 ]θ ]φ/θ , . . . , h [[an ]θ ]φ/θ .
Thus h : (M/θ)/(φ/θ) → M/φ is an isomorphism. 


Definition 2.40. Suppose M is an L-algebra,  N ⊆ M and θ ∈ ConL (M).


Let N θ = {a ∈ M | [a ]θ ∩ N = ∅}, put Nθ = N θ M , and let θ|N denote the
restriction of θ to N .

In other words, N θ is a union of congruence classes being incident with N .

Lemma 2.41. Let N be a subalgebra of M, θ ⊆ ConL (M). Then


(i) N θ is the universe of Nθ ,
(ii) θ|N ∈ ConL (N).

Proof. (i): Take any n-ary f N ∈ F N and a1 , . . . , an ∈ N θ . By Definition 2.40


there are elements b1 , . . . , bn ∈ N for which we have a1 ∈ [b1 ]θ , . . . , an ∈
[bn ]θ , in other words θ(a1 , b1 ) = · · · = θ(an , bn ) = 1. Hence, it follows that
θ f M (a1 , . . . , an ), f M (b1 , . . . , bn ) = 1. Since the result f M (b1 , . . . , bn ) is in
N , we have f M (a1 , . . . , an ) ∈ N θ . (ii) is easy to check. 


Theorem 2.42 (third isomorphism theorem). Let N be a subalgebra of


an L-algebra M, θ ⊆ ConL (M). Then N/(θ|N ) ∼
= Nθ /(θ|N θ ).

Proof. Take a mapping h : N/(θ|N ) → N θ /(θ|N θ ), where h([a ]θ|N ) = [a ]θ|N θ


for all a ∈ M . It is routine to verify that h is the required isomorphism.  

Lemma 2.43. Let M and N be L-algebras such that M ∈ Sub(N) and let
θ ∈ ConL (M), φ ∈ ConL (N), θ ⊆ φ|M . Then a mapping h : M/θ → N/φ
defined by h([a ]θ ) = [a ]φ is a morphism.
82 2 Algebras with Fuzzy Equalities

Proof. The claim is a consequence of Theorem 2.39 and Theorem 2.42. Indeed,
let h(φ|M )/θ : M/θ → (M/θ)/((φ|M )/θ) be the natural morphism, where
(φ|M )/θ ∈ ConL (M/θ) is defined by (2.8). From Theorem 2.39 it follows that
there is an isomorphism g : (M/θ)/((φ|M )/θ) → M/(φ|M ) such that we have
g(h(φ|M )/θ ([a ]θ )) = [a ]φ|M for every a ∈ M . Due to Theorem 2.42, there is
an isomorphism g  : M/(φ|M ) → Mφ /(φ|M φ ), where g  ([a ]φ|M ) = [a ]φ|M φ .
Obviously, Mφ /(φ|M φ ) ∈ Sub(N/φ). Put h = h(φ|M )/θ ◦ g ◦ g  . Clearly, we
have h([a ]θ ) = g  ([a ]φ|M ) = [a ]φ|M φ = [a ]φ for all a ∈ M . 


Lemma 2.44. Let h : M → N be a morphism and let φ ∈ ConL (M) such


that φ ⊆ θh . Then h = hφ ◦ g, where g : M/φ → N is a uniquely determined
morphism.

Proof. Since φ ⊆ θh from Lemma   2.43 it follows that there is a morphism


g  : M/φ → M/θh , where g  [a ]φ = [a ]θh , a ∈ M . Moreover, Theorem 2.35
 
yields that there is an embedding g  : M/θh → N, where g  [a ]θh = h(a ),
[a ]θh ∈ M/θh . Thus, we can put g = g  ◦ g  . Clearly,
      
h(a ) = g  [a ]θh = g  g  [a ]φ = g [a ]φ = g(hφ (a )) = (hφ ◦ g)(a )
for every a ∈ M , i.e. h = hφ ◦ g. It remains to check the uniqueness of g. Since
hφ is surjective, hφ ◦ g = h = hφ ◦ k implies g = k. 


For θ1 , θ2 ∈ ConL (M) with θ1 ⊆ θ2 we put


[θ1 , θ2 ] = {φ ∈ ConL (M) | θ1 ⊆ φ ⊆ θ2 } . (2.9)
It is easily seen that [θ1 , θ2 ] is a complete sublattice of ConL (M).

Theorem 2.45. Let an L-algebra M and θ ∈ ConL (M) be given. Then the
mapping h : [θ, M ×M ] → ConL (M/θ) defined by h(φ) = φ/θ is a lattice
isomorphism.

Proof. For every φ1 , φ2 ∈ [θ, M ×M ] we have φ1 ⊆ φ2 iff φ1 /θ ⊆ φ2 /θ iff


h(φ1 ) ⊆ h(φ2 ). Now it is sufficient to prove
 thath is a bijective mapping.
Suppose that (φ1 /θ) [a ]θ , [b ]θ = (φ2 /θ) [a ]θ , [b ]θ for every a , b ∈ M . Then
also φ1 (a , b ) = φ2 (a , b ) for all a , b ∈ M . Thus, the mapping h is injective.
To show surjectivity, first  take any ϕ ∈ ConL (M/θ) and define an L-relation
φ on M by φ(a , b ) = ϕ [a ]θ , [b ]θ . We will show that φ is an L-equivalence
on M for which θ ⊆ φ and h(φ) = ϕ.
First, we check conditions (i), (ii) of Definition 2.11. Clearly, φ is an L-
equivalence relation and for every a , b ∈ M we have
 
a ≈M b ≤ [a ]θ ≈M/θ [b ]θ ≤ ϕ [a ]θ , [b ]θ = φ(a , b ) ,
which is equivalent to (ii), see Lemma 1.82. To check condition (i), take any
n-ary f M ∈ F M and a1 , b1 , . . . , an , bn ∈ M . Now the following inequality
holds,
2.3 Direct and Subdirect Products 83
n n  
i=1 φ(ai , bi ) = i=1 ϕ [ai ]θ , [bi ]θ ≤
 M/θ    
≤ϕ f [a1 ]θ , . . . , [an ]θ , f M/θ [b1 ]θ , . . . , [bn ]θ =
    
= ϕ f M (a1 , . . . , an ) θ , f M (b1 , . . . , bn ) θ =
 
= φ f M (a1 , . . . , an ), f M (b1 , . . . , bn ) .
Thus, φ ∈ [θ, M ×M ] ⊆ ConL (M). Clearly, h(φ) = φ/θ = ϕ. To sum up, h is
a surjective mapping. Hence, h is a lattice isomorphism. 


2.3 Direct and Subdirect Products

Direct product is a basic construction yielding larger L-algebras from collec-


tions of input L-algebras. Subdirect product can be thought of as a derived
construction since it is a special subalgebra of a direct product. This section
introduces basic properties of both constructions. It is worth to note that some
results presented in this section will generalize only for certain subclasses of
residuated lattices as the structures of truth degrees.

Definition 2.46. Let I be an index  set. A direct


 product of aQfamily
{Mi | i ∈ I} of L-algebras Mi = Mi , ≈Mi , F Mi is an L-algebraQ i∈I Mi
Q
ofQtype F such that M = i∈I Mi and for every n-ary function f i∈I Mi ∈
F i∈I Mi and a1 , . . . , an ∈ M we have
Q
f i∈I Mi (a1 , . . . , an )(i) = f Mi (a1 (i), . . . , an (i)), (2.10)
Q
for all i ∈ I and the L-equality ≈ i∈I Mi is defined by
Q   
(a ≈ i∈I Mi b) = i∈I a (i) ≈
Mi
b (i) (2.11)
for all a , b ∈ M .
Q
Remark 2.47. (1) Note that if I = ∅, then | i∈I M i | = 1.
Q (2) A direct product as defined above is a well-defined L-algebra. Suppose
i∈I Mi is a direct product of a family {Mi | i ∈ I} of L-algebras. L-relation
Q Q
≈ i∈I Mi is evidently reflexive and symmetric. Now for every a , b , c ∈ i∈I M i
we have
 Q   Q 
a ≈ i∈I Mi b ⊗ b ≈ i∈I Mi c =
   
i∈I a (i) ≈ b (i) ⊗ j∈I b (j) ≈Mj c (j) ≤
Mi
=
  
≤ i,j∈I a (i) ≈Mi b (i) ⊗ b (j) ≈Mj c (j) ≤
  
≤ i∈I a (i) ≈Mi b (i) ⊗ b (i) ≈Mi c (i) ≤
   Q
≤ i∈I a (i) ≈Mi c (i) = a ≈ i∈I Mi c .
Q Q
Q For a ≈
Hence, ≈ i∈I Mi is transitive. i∈I Mi b = 1 we have a (j) ≈Mj b (j) =
Q
1 for every j ∈ I, thus a ≈ i∈I M i
b = 1 implies a = b . Altogether, ≈ i∈I Mi
84 2 Algebras with Fuzzy Equalities

Fig. 2.12. Example of direct product

Q
is an L-equality relation. The compatibility ofQfunctions with Q ≈ i∈I Mi follows
from the definition. Q Indeed, for every n-ary f i∈I M i
∈ F i∈I i and arbitrary
M

a1 , b1 , . . . , an , bn ∈ i∈I M i , we have
 Q   Q   
a1 ≈ i∈I Mi b1 ⊗ · · · ⊗ an ≈ i∈I Mi bn = nj=1 i∈I aj (i) ≈Mi bj (i) ≤
 n
≤ i∈I j=1 aj (i) ≈Mi bj (i) ≤

≤ i∈I f Mi (a1 (i), . . . , an (i)) ≈Mi f Mi (b1 (i), . . . , bn (i)) =
 Q Q Q 
= f i∈I Mi (a1 , . . . , an ) ≈ i∈I Mi f i∈I Mi (b1 , . . . , bn ) .
Q Q
Hence, ≈ i∈I Mi is an L-equality compatible with all functions of i∈I Mi .

Example 2.48. Let L be a three-element L  ukasiewicz chain. The direct product


of two linearly ordered lattices with L-equality is depicted in Fig. 2.12. The
L-equalities are depicted the same way as in Example 2.3.
Q
Definition 2.49. Let i∈I Mi be a direct product Q of a family {Mi | i ∈ I} of
L-algebras. For every j ∈ I a mapping πj : i∈I Mi → Q Mj , where πj (a ) =
a (j) is called a projection
Q map on the j-th coordinate of i∈I Mi or, shortly,
a j-th projection of i∈I Mi .
Q
Theorem 2.50. For a direct product i∈I Mi of a family {Mi | i ∈ I} of L-
algebras, the j-th projection πj is an epimorphism for every j ∈ I.

Q Take any j ∈ I. For arbitrary a ∈ Mj we have πj (b ) = a for every


Proof.
b ∈ i∈I Mi , where b (j) = a . Thus, πj is a surjective mapping. Moreover,
Q 
a ≈ i∈I Mi b = i∈I a (i) ≈Mi b (i) ≤ a (j) ≈Mj b (j) = πj (a ) ≈Mj πj (b ) .
Q Q Q
For any n-ary f i∈I Mi ∈ F i∈I Mi and a1 , . . . , an ∈ i∈I Mi we have,
 Q  Q
πj f i∈I Mi (a1 , . . . , an ) = f i∈I Mi (a1 , . . . , an )(j) =
   
= f Mj a1 (j), . . . , an (j) = f Mj πj (a1 ), . . . , πj (an ) .
Hence, πj is an epimorphism. 

2.3 Direct and Subdirect Products 85

h
M i∈I Mi

hi πi

Mi
Fig. 2.13. Direct product property

Theorem 2.51. For every family of morphisms Q {hi : M → Mi | i ∈ I} there


is a uniquely determined morphism h : M → i∈I Mi such that h ◦ πi = hi for
every i ∈ I.
Q
Proof. Let us have a mapping h : M → i∈I M i defined by h(a )(i) = hi (a )
for every i ∈ I, a ∈ M . We have a ≈M b ≤ hi (a ) ≈Mi hi (b ) since hi is
supposed to be a morphism for every i ∈ I. Thus,
 Q
a ≈M b ≤ i∈I hi (a ) ≈Mi hi (b ) = h(a ) ≈ i∈I Mi h(b ) .
Furthermore, for n-ary f M ∈ F M , and a1 , . . . , an ∈ M we have
     
h f M (a1 , . . . , an ) (i) = hi f M (a1 , . . . , an ) = f Mi hi (a1 ), . . . , hi (an ) =
  Q  
= f Mi h(a1 )(i), . . . , h(an )(i) = f i∈I Mi h(a1 ), . . . , h(an ) (i)
for every i ∈ I. Hence h is a morphism. Clearly, h ◦Qπi = hi for every i ∈ I.
It suffices to check the uniqueness. Let g : M → i∈I Mi be a morphism
satisfying g ◦ πi = hi for all i ∈ I. Hence, g ◦ πi = h ◦ πi for every i ∈ I. Thus,
for arbitrary a ∈ M , we have g(a )(i) = h(a )(i) for all i ∈ I which implies
g(a ) = h(a ). Hence, h is determined uniquely. 

Remark 2.52. The unique existence of morphism h described by Theorem 2.51
Q occasionally referred to as the direct product property of family {πj :
is
i∈I Mi → Mj | j ∈ I}, cf. [97], and is illustrated in in Fig. 2.13.

Lemma 2.53. For a direct product M = M1 × M2 , we have


(i) θπ1 ∧ θπ2 = ≈M ,
(ii) θπ1 ∨ θπ2 = M ×M ,
(iii) θπ1 ◦ θπ2 = θπ2 ◦ θπ1 .
 
Proof.
 For
 (i), we have θ π 1
∧ θ π 2
(a , b ) = θπ1 (a , b ) ∧ θπ2 (a , b ). That is
θπ1 ∧ θπ2 (a , bQ) = a (1) ≈ M1
b (1) ∧ a (2) ≈M2 b (2), but this is exactly the
definition of ≈ i∈I M i
for I = {1, 2}. Thus, (i) holds.
(ii): For any a , b ∈ M1 × M2 , we have
θπ1 (a , a (1), b (2) ) = 1, and θπ2 (a (1), b (2) , b ) = 1 . (2.12)
By transitivity, for each congruence θ ∈ ConL (M) containing both θπ1 and
θπ2 we have θ(a , b ) = 1, whence θπ1 ∨ θπ2 = M ×M proving (ii).
(iii): It is immediate that (θπ1 ◦ θπ2 )(a , b ) = 1 for every a , b ∈ M . Thus,
θπ1 ◦ θπ2 = θπ1 ∨ θπ2 from which we get (iii). 

86 2 Algebras with Fuzzy Equalities

Definition 2.54. A congruence θ ∈ ConL (M) is called a factor congru-


ence, if there is a congruence θ∗ ∈ ConL (M) such that
(i) θ ∧ θ∗ = ≈M ,
(ii) [a ]θ ∩ [b ]θ∗ = ∅ for all a , b ∈ M .
The pair θ, θ∗ is called a pair of factor congruences on M.

Remark 2.55. The proof of Lemma 2.53 yields that for every M = M1 × M2
the couple θπ1 , θπ2 is a pair of factor congruences on M. Moreover, condition
(ii) of Definition 2.54 implies that for every a , b ∈ M there is some c ∈ M
such that θ∗ (a , c ) ⊗ θ(c , b ) = 1. Hence, from (ii) it follows that θ ◦ θ∗ =
θ∗ ◦ θ = θ ∨ θ∗ = M ×M , this is easy to check.

Theorem 2.56. If θ, θ∗ is a pair of factor congruences on an L-algebra M,


then M ∼
= M/θ × M/θ∗ .

Proof. Put h(a ) = [a ]θ , [a ]θ∗ for all a ∈ M . For every a , b ∈ M , we have
a ≈M b = (θ ∧ θ∗ )(a , b ) = θ(a , b ) ∧ θ∗ (a , b ) =

= [a ]θ ≈M/θ [b ]θ ∧ [a ]θ∗ ≈M/θ [b ]θ∗ =
∗ ∗
= [a ]θ , [a ]θ∗ ≈M/θ×M/θ [b ]θ , [b ]θ∗ = h(a ) ≈M/θ×M/θ h(b ) .
Condition (ii) of Definition 2.54 yields that for every a , b ∈ M there is some
c ∈ M such that c ∈ [a ]θ ∩ [b ]θ∗ . That is, c ∈ [a ]θ and c ∈ [b ]θ∗ , i.e. [a ]θ =
[c ]θ , [b ]θ∗ = [c ]θ∗ . Therefore, we obtain h(c ) = [c ]θ , [c ]θ∗ = [a ]θ , [b ]θ∗ , i.e.
h is surjective.
Finally, we have to check the compatibility with functions. Take any n-ary
f M ∈ F M and arbitrary a1 , . . . , an ∈ M . We have,
      
h f M (a1 , . . . , an ) = f M (a1 , . . . , an ) θ , f M (a1 , . . . , an ) θ∗ =
 ∗ 
= f M/θ ([a1 ]θ , . . . , [an ]θ ), f M/θ ([a1 ]θ∗ , . . . , [an ]θ∗ ) =
∗ 
= f M/θ×M/θ [a1 ]θ , [a1 ]θ∗ , . . . , [an ]θ , [an ]θ∗ =
∗ 
= f M/θ×M/θ h(a1 ), . . . , h(an ) .
Hence, h is an isomorphism, M ∼
= M/θ × M/θ∗ . 


Remark 2.57. The previous theorem  yields that every L-algebra is isomorphic
to a direct product. Namely, since ≈M , M ×M is a pair of factor congruences
on M, whence M ∼ = M/≈M × M/(M ×M ).

Definition 2.58. An L-algebra M is said to be directly indecomposable if


M is not isomorphic to a direct product of two non-trivial L-algebras.
 
Theorem 2.59. An L-algebra is directly indecomposable iff ≈M , M ×M is
the only one pair of its factor congruences. Moreover, every finite L-algebra
is isomorphic to a direct product of directly indecomposable L-algebras.
2.3 Direct and Subdirect Products 87

Proof. Remark 2.55 and Theorem 2.56 yield that M = ∼ N × N iff there is a
pair of factor congruences θ, θ on M, such that N ∼

= M/θ  ∼
 and N = M/θ .

So clearly, M is directly indecomposable iff ≈ , M ×M is the only one pair


M

of factor congruences of M.
Trivial L-algebras are directly indecomposable. Any finite non-trivial L-
algebra M either is directly indecomposable or M ∼ = N × N where |N | < n

and |N | < n. Thus, by induction one can prove that M is isomorphic to a
direct product of directly indecomposable L-algebras. 

   
Definition 2.60. If a , b ∈ M and h : M, ≈M → N, ≈N is a mapping, we
 b if h(a ) ≈ h(b ) < 1. A family of mappings
N
say  h separates
 that   a and
hi : M, ≈ M
→ Ni , ≈ Ni
| i ∈ I separates pointsiff for every
 a , b ∈ M
with a ≈M b < 1 there is an index j ∈ I such that hj : M, ≈M → Nj , ≈Nj
separates a and b .
Lemma 2.61. For a family of morphisms {hi : M → Mi | i ∈ I}, the follow-
ing conditions are equivalent:
(i) the family {hi : M → M Qi | i ∈ I} separates points,
(ii) a mapping h : M → i∈I Mi , where h(a )(i) = hi (a ) for every i ∈ I,
∈ M is an injective morphism,
a
(iii) 1 i∈I θhi = {a , a | a ∈ M }.
Proof. “(i) ⇒ (ii)”: Theorem 2.51 yields that h is a morphism. For a = b
we have a ≈M b < 1. Thus, if {hi : M → Mi | i ∈ I} separates points,
then there
Q is some i0 ∈ I such that hi0 (a ) ≈Mi0 hi0 (b ) < 1, which implies
h(a ) ≈ i∈I Mi h(b ) < 1, i.e. h is injective.

“(ii) ⇒ (iii)”: Evidently, i∈I θ hi
(a , a ) = 1, thus {a , a | a ∈ M } ⊆
i∈I θhi . Let us assume that (ii) holds. Then for a = b we obtain
1
Q 
1 > h(a ) ≈ i∈I Mi h(b ) = i∈I hi (a ) ≈Mi hi (b ) =
  
= i∈I θhi (a , b ) = i∈I θhi (a , b ) ,
 
that is a , b ∈ 1 i∈I θhi . Hence,
 (ii) implies
 (iii).
“(iii) ⇒ (i)”: Let a = b , i.e. i∈I hiθ (a , b ) < 1 by (iii). This gives

i∈I hi (a ) ≈
Mi
hi (b ) < 1 ,
i.e., there is i0 ∈ I such that hi0 (a ) ≈Mi0 hi0 (b ) < 1, showing that hi0
separates a and b . Therefore, {hi : M → Mi | i ∈ I} separates points. 

Lemma Q 2.62. Let {hi : M → Mi | i ∈ I} be a family of morphisms and let
h : M → i∈I Mi denote a morphism, where h(a )(i) = hi (a ) for every i ∈ I,
a ∈ M . The morphism h is an embedding iff
 
i∈I θhi (a , b ) = a ≈ b .
M

Q  
Proof. Recall that h(a ) ≈ i∈I Mi h(b ) = i∈I θhi (a , b ) for all a , b ∈ M .
The rest is evident. 

88 2 Algebras with Fuzzy Equalities

∗ ∗ ∗
In bivalent case, every algebra can be represented by a subdirect product
of subdirectly irreducible algebras. In fuzzy case, this is not true in general. In
the subsequent development, we introduce a sufficient condition for subdirect
representation. However, unlike the bivalent case, we will also show that there
are L-algebras, which are not subdirectly representable.
Definition 2.63. Let M be an L-algebra of type F . The L-algebra M is said
to be a subdirect product of a family {Mi | i ∈ I} of L-algebras of type F if
Q
(i) M is a subalgebra of i∈I Mi ,
(ii) πi (M ) = Mi for every i ∈ I.
Q
An embedding h : M → i∈I Mi is called subdirect if h(M) is a subdirect
product of the family {Mi | i ∈ I}.

Lemma 2.64. If θQi ∈ ConL (M) for every i ∈ I and i∈I θi = ≈M , then the
mapping g : M Q → i∈I M/θi , where g(a )(i) = [a ]θi is a subdirect embedding.
If h : M → i∈I Mi is a subdirect embedding,
 then there is a family of con-
gruences {θi ∈ ConL (M) | i ∈ I} such that i∈I θi = ≈M and Mi ∼ = M/θi for
every i ∈ I.
Proof. Since g defined by g(a )(i) = [a ]θi is a morphism, using Lemma 2.62 we
can deduce that g is an embedding. Moreover, g(M )(i) = {[a ]θi | a ∈ M } =
M/θi for every i ∈ I, thus g isQa subdirect embedding.
Furthermore, let h : M → i∈I Mi be a subdirect embedding. Put θi = θhi
for every i ∈ I (recall that hi : M → Mi , where hi (a ) = h(a )(i) is a 
surjective
morphism). Now by applying Lemma 2.62 one can conclude that i∈I θi =
≈M . Since every hi is surjective, Theorem 2.35 yields that Mi ∼ = M/θi for
every i ∈ I. 

Definition 2.65. An L-algebra QM is subdirectly irreducible if for every
subdirect embedding h : M → i∈I Mi there is an index j ∈ I such that
h ◦ πj : M → Mj is an isomorphism.
Theorem 2.66. An L-algebra M is subdirectly irreducible iff M either is triv-
congruence in ConL (M) − {≈ }. In the latter case
M
ial,
  or there is a least
ConL (M) − {≈ } is a principal congruence and ConL (M), ⊆ contains
M

exactly one atom.


Proof. “⇒”: Suppose, by contradiction, that M is not a trivial L-algebra and
ConL (M) − {≈M } does not have the least element. Then
 
ConL (M) − {≈M } = ≈M .

QL (M)−{≈ }. Using Lemma 2.64, there is a subdirect embedding


M
Put I = Con
h : M → θ∈I M/θ, where h(a )(θ) = [a ]θ for every a ∈ M and θ ∈ I.
Moreover, for every congruence θ ∈ I we have θ ⊃ ≈M . Thus, there are
a , b ∈ M such that a ≈M b < θ(a , b ), i.e.
2.3 Direct and Subdirect Products 89

a ≈M b < θ(a , b ) = [a ]θ ≈M/θ [b ]θ =


= hθ (a ) ≈M/θ hθ (b ) = (h ◦ πθ )(a ) ≈M/θ (h ◦ πθ )(b ) .
Hence, for every θ ∈ I the mapping h ◦ πθ : M → M is not an isomorphism.
Consequently, the L-algebra M is not subdirectly irreducible.
“⇐”: A trivial L-algebra MQis subdirectly irreducible. Indeed, for every
subdirect embedding h : M → i∈I Mi , each L-algebra Mi must be trivial,
because πi (h(M )) = πi (h({a })) = Mi holds for every i ∈ I by virtue of
the assumption. Hence, h ◦ πi : M → Mi is an isomorphism for all i ∈ I.
Altogether,
 M is subdirectlyM irreducible.
 So suppose M is non-trivial and
let θ = ConL (M) − {≈ } > ≈M , i.e. θ is the least congruence in
ConL (M) − {≈Q }. Then there are a , b ∈ M such that θ(a , b ) > a ≈ b .
M M

Let h : M → i∈I Mi be a subdirect embedding. Now we have,



a ≈M b = i∈I h(a )(i) ≈Mi h(b )(i) < θ(a , b ) .
Therefore, it follows that there is an index i ∈ I such that h(a )(i) ≈Mi
h(b )(i)  θ(a , b ), that is, (h ◦ πi )(a ) ≈Mi (h ◦ πi )(b )  θ(a , b ), so θh◦πi 
θ. Since θ is the least congruence in ConL (M) − {≈M }, we readily obtain
θh◦πi = ≈M . Thus, h ◦ πi : M → Mi is an isomorphism and the L-algebra M
is subdirectly irreducible.
If θ is the least congruence in ConL (M) − {≈M }, then there are elements
a , b ∈ M such that θ(a , b ) > a ≈M b . Obviously, we have θ(θ(a ,b )/a , b ) ⊆
θ. The converse inclusion holds since θ(θ(a ,b )/a , b ) > a ≈M b and θ is the
least congruence with θ(a , b ) > a ≈M b . Thus, we have θ(θ(a ,b )/a , b ) = θ,
i.e. θ is a principal congruence. 


Theorem 2.67 (representation theorem). Let M be a non-trivial L-


algebra such that for every distinct a , b ∈ M there exists θa ,b ∈ ConL (M),
where θa ,b (a , b ) = a ≈M b , and θa ,b is ∧-irreducible in ConL (M). Then
M is isomorphic to a subdirect product of a family of subdirectly irreducible
L-algebras.

Proof. Let θa ,b be ∧-irreducible in ConL (M). As a result, [θa ,b , M ×M ] −


{θa ,b } has the least element. Now from Theorem 2.45 andTheorem 2.66
it follows that M/θa ,b is subdirectly irreducible. We have a ,b ∈I θa ,b =
≈M , where I = {a , b | a , b ∈ M and a = b }. Thus, Q Lemma 2.64 yields
that there is a subdirect embedding h : M → a ,b ∈I M/θa ,b , where
h(c )(a , b ) = [c ]θa ,b for all a , b , c ∈ M , a = b . Hence, M is isomorphic
to a subdirect product of subdirectly irreducible L-algebras. 


The previous representation theorem is sort of abstract. It only delimits a


condition under which an L-algebra M has a subdirect representation. In the
following we will focus on the existence of suitable ∧-irreducible elements in
ConL (M). First, we prove a technical lemma.
90 2 Algebras with Fuzzy Equalities

Lemma 2.68. Let {θi | i ∈ I} ⊆ ConL (M) be a directed system of congru-


ences, i.e. for any finite number θi1 , . . . , θik  ∈ {θi | i ∈  I}, there is θi , i ∈ I
such that θij ⊆ θi , for all j = 1, . . . , k. Then i∈I θi = i∈I θi .
 
Proof. Clearly, i∈I θi ⊆ i∈I θi . We have to show “⊇”:
    
i∈I θi (a , b ) = i1 ,...,ik ∈I θi1 (a , c1 ) ⊗ · · · ⊗ θik (ck−1 , b ) ≤
c1 ,...,ck−1 ∈M
  
≤ i∈I θi (a , c1 ) ⊗ · · · ⊗ θi (ck−1 , b ) ≤
c1 ,...,ck−1 ∈M
  
≤ i∈I θi (a , b ) = i∈I θi (a , b )

holds for every a , b ∈ M . 




Lemma 2.69. Suppose M is a non-trivial L-algebra. Then for every distinct


elements a , b ∈ M :
(i) there is a maximal θa ,b ∈ ConL (M) such that θa ,b (a , b ) = a ≈M b ;
(ii) if a ≈M b is ∧-irreducible in L then θa ,b is ∧-irreducible in ConL (M).
 
Proof. (i): Let Ia ,b = θ ∈ ConL (M) | θ(a , b ) = a ≈M b . It is easy to ob-
serve that Ia ,b , ⊆ is a partially ordered set, Ia ,b ⊆ ConL (M). Moreover,
≈ ∈ Ia,b implies I a ,b = ∅. I ⊆ Ia ,b of congruences we have
M
 For every chain
 i∈I θ i (a , b ) = i∈I θ i (a , b ) = a ≈ M
b due to Lemma 2.68. That is,
i∈I θ i ∈ Ia , b . Hence, every chain I ⊆ Ia , b is bounded from above. Now using
Zorn Lemma it readily follows that there is a maximal element θa ,b ∈ Ia ,b .
(ii): Put J = [θa ,b , M ×M ] − {θa ,b } and suppose a ≈M b to be an ∧-
irreducible element of L. By maximality of θa ,b , θ(a , b ) > θa ,b (a , b ) for each
θ ∈ J. Thus, it follows that
  
J (a , b ) = θ∈J θ(a , b ) > a ≈M b .

As a consequence, J ∈ J, i.e. θa ,b is ∧-irreducible in ConL (M). 


If the lattice part of L is a finite chain then each 1 = a ∈ L is ∧-irreducible.


Thus, for finite linearly ordered residuated lattices we can use Theorem 2.67
and Lemma 2.69 to obtain the following consequence.

Corollary 2.70. If L is a finite chain then every non-trivial L-algebra is


isomorphic to a subdirect product of subdirectly irreducible L-algebras. 


Remark 2.71. (1) In [16] we have presented more general criterion for subdirect
representation. This criterion is, however, more technical than the one given
by Lemma 2.69. Let M be an L-algebra. If for every distinct b , b  ∈ M we
have
 
b ≈M b  <  
a > b ≈M b  θ( /b , b ) (b , b ),
a (2.13)
then M is isomorphic to a subdirect product of subdirectly irreducible L-
algebras. Indeed, take a maximal θb ,b  ∈ ConL (M) with θb ,b  (b , b  ) = b ≈M
2.4 Terms, Term L-Algebras 91


. Clearly, θb ,b  is ∧-irreducible
b

 in ConL (M) due to (2.13). Namely, θb ,b  ∨
a > b ≈M b  θ(a/ b , b ) is the least congruence in [θb ,b  , M ×M ] − {θb ,b  }.
Now apply Theorem 2.67. It is easily seen that if b ≈M b  is ∧-irreducible in
L, then (2.13) holds trivially.
(2) The subdirect representation does not pass for every L. In other words,
for certain structures of truth degrees, there are still L-algebras which cannot
be isomorphic to a subdirect product of subdirectly irreducible L-algebras.
An example follows.

Example 2.72. Take a complete residuated  lattice


 L on the real unit interval
[0, 1]. Let us have an L-algebra M = M, ≈M , ∅ of type F = ∅, where M =
{a , b }, and a ≈M b = 0. It is easy to see that every reflexive and symmetric
binary L-relation θ on M is a congruence on M since M is a two-element set
with a ≈M b = 0, and F M = ∅. Obviously, every congruence θ ∈ ConL (M)
is uniquely determined by the truth degree θ(a , b ) ∈ L. Thus, there is a
one-to-one correspondence between congruences from ConL (M) and truth
degrees from L. Moreover, ConL (M) is isomorphic to the lattice part of L.
For θ ∈ ConL (M), Theorem 2.45 yields ConL (M/θ) ∼ = [θ, M ×M ] ∼ = [c, 1],
where θ(a , b ) = c. Since [c, 1] − {c} does not have theQleast element, M/θ is
subdirectly reducible. As a consequence, if h : M → i∈I Mi is a subdirect
embedding, then every Mi ∼ = M/θh◦πi is subdirectly reducible, i.e. M is not
isomorphic to a subdirect product of subdirectly irreducible L-algebras.

2.4 Terms, Term L-Algebras


The notion of a term is defined as usual. As we will see, results known from
the ordinary case generalize naturally to fuzzy case. In addition to that, there
are new natural properties in fuzzy case which are degenerate in the ordinary
case (cf. Theorem 2.76 and Remark 2.77).

Definition 2.73. Let X be a set of variables. Let F be a type such that X ∩


F = ∅. The set T (X) of terms of type F is the smallest set such that
(i) X ⊆ T (X),
(ii) If f ∈ F , f is n-ary and t1 , . . . , tn ∈ T (X), then f (t1 , . . . , tn ) ∈ T (X).
Every t ∈ T (X) is called a term of type F over X.

Definition 2.74. Let t be a term of type F over X. For a variable x ∈ X


let |t|x denote the number of occurrences of x in term t. That is, if t is a
variable, then

1 for t = x ,
|t|x =
0 otherwise ,
if t is of the form f (t1 , . . . , tn ), then |t|x = |t1 |x +· · ·+|tn |x . Variable x ∈ X has
an occurrence in t ∈ T (X) if |t|x > 0. For t ∈ T (X) we write t(x1 , . . . , xk )
92 2 Algebras with Fuzzy Equalities

instead of t to indicate that the variables occurring in t are among x1 , . . . , xk ∈


X. A term t is n-ary if the number of variables appearing explicitly in t is at
most n. For t ∈ T (X) we define var(t) ⊆ X by var(t) = {x ∈ X | |t|x > 0}.
Definition 2.75. Given a term t(x1 , . . . , xn ) of type F over X and given an
L-algebra M of type F let tM denote a mapping tM: M n → M such that
(i) if t is a variable xi then tM (a1 , . . . , an ) = ai ,
(ii) if t is of the form f (t1 , . . . , tk ) then 
tM (a1 , . . . , an ) = f M tM M
1 (a1 , . . . , an ), . . . , tk (a1 , . . . , an ) .

Mapping tM is called a term function on M corresponding to term t.


Theorem 2.76. Let t(x1 , . . . , xn ) be a term of type F over X. Then for every
L-algebra M we have the following properties:
 
(i) θ(a1 , b1 )|t|x1 ⊗ · · · ⊗ θ(an , bn )|t|xn ≤ θ tM (a1 , . . . , an ), tM (b1 , . . . , bn )
 every θ ∈ ConL (M) and arbitrary a1 ,b1 , . . . , an , bn ∈ M ;
for
(ii) h tM (a1 , . . . , an ) = tN h(a1 ), . . . , h(an )
for every morphism h : M → N and arbitrary a1 , . . . , an ∈ M .
Proof. The first statement can be proved using structural induction. Take any
t ∈ T (X) and θ ∈ ConL (M). If t is a variable xi , 1 ≤ i ≤ n, then (i) holds
trivially. Suppose that t is of the form f (t1 , . . . , tk ) and (i) holds for every
t1 , . . . , tk . Now it follows that
θ(a1 , b1 )|t|x1 ⊗ · · · ⊗ θ(an , bn )|t|xn =
k  
= i=1 θ(a1 , b1 )|ti |x1 ⊗ · · · ⊗ θ(an , bn )|ti |xn ≤
k  
≤ i=1 θ tM i (a1 , . . . , an ), ti (b1 , . . . , bn ) ≤
M
 
≤ θ tM (a1 , . . . , an ), tM (b1 , . . . , bn ) .
Hence, (i) holds. (ii) follows from the ordinary case. 

Remark 2.77. (1) Estimation (i) of Theorem 2.76 is degenerate and has no
interesting meaning in the ordinary case. In fuzzy case, however, (i) provides
a non-trivial (possibly numerical, when truth degrees are numbers) estimation
of similarity (degree of equivalence) of results of term functions which can be
interpreted as describing sensitivity of a term function.
(2) Since ≈M is a congruence relation, by applying (i) of Theorem 2.76 to
≈ we obtain
M

(a1 ≈M b1 )|t|x1 ⊗ · · · ⊗ (an ≈M bn )|t|xn ≤


≤ tM (a1 , . . . , an ) ≈M tM (b1 , . . . , bn ) . (2.14)
|t|xi
(3) If ⊗ is ∧, the powers |t|xi can be removed from θ(ai , bi ) and
(ai ≈M bi )|t|xi . This has the following consequence. In general, the estimation
describing sensitivity of a term function depends on the structure of the cor-
responding term, hence the number of occurrences of variables is important.
If ⊗ = ∧, the structure of the term does not play any role.
2.4 Terms, Term L-Algebras 93

Definition 2.78. Given a type F and a set of variables X, if T (X) = ∅


then
 the term L-algebra
 of type F over X, is an L-algebra T(X) =
T (X), ≈T(X) , F T(X) of type F with functions defined by
f T(X) (t1 , . . . , tn ) = f (t1 , . . . , tn ) (2.15)
for any n-ary f T(X)
∈F T(X)
and t1 , . . . , tn ∈ T (X), and L-equality ≈T(X)
defined by

T(X)  1 for t = t ,
t≈ t = (2.16)
0 otherwise .
for all t, t ∈ T (X). The set X is called a set of free generators of T(X).
Theorem 2.79. Let T(X) be a term L-algebra  of type F in variables X.
Every mapping h : X → M , where M = M, ≈M , F M is an L-algebra of
type F , can be extended to a morphism h : T(X) → M, where h(x) = h (x)
for all x ∈ X.
Proof. The homomorphic extension h : T(X) → M can be defined induc-
 for every variable x ∈ X and h (f (t1 , . . . , tn )) =
 
  by h (x) = h(x)
tively
f h (t1 ), . . . , h (tn ) for every n-ary f ∈ F and terms t1 , . . . , tn ∈ T (X).
M

Since ≈T(X) is a crisp identity relation, h is evidently an ≈-morphism, the


rest is evident. Hence, h is a homomorphic extension of h. 

Example 2.80. Homomorphic extension h described by Theorem 2.79 can be
used to define important structural notions which involve terms. In the fol-
lowing we show examples.
(1) Let M be an L-algebra, a1 , . . . , an ∈ M . For X = {x1 , . . . , xn }, put
h(xi ) = ai (i = 1, . . . , n). One can see that h : T (X) → M is a mapping
such that h (t) = tM (a1 , . . . , an ) for any term t(x1 , . . . , xn ). Hence, h (t) is
the value of term function tM applied to a1 , . . . , an .
(2) For each term L-algebra T(X) we can consider an L-algebra M of the
same type such that M = N0 , ≈M is the crisp L-equality on M , and for each
n-ary f M ∈ F M we have f M (a1 , . . . , an ) = a1 + · · · + an (a1 , . . . , an ∈ M ).
Now introduce h : X → M by putting h(x) = 1, h(y) = 0 if x = y. Then
h (t) = |t|x for each t ∈ T (X).
(3) Analogously as in (2), for T(X) consider M with M = 2X , ≈M being
the crisp L-equality on M , f M (a1 , . . . , an ) = a1 ∪ · · · ∪ an (a1 , . . . , an ∈ M ).
For the homomorphic extension h of mapping h : X → M where h(x) = {x}
(x ∈ X), we have h (t) = var(t).
∗ ∗ ∗
In the rest of this section, we illustrate the role of term functions by pro-
viding a description of two notions introduced above. In the first case, the
theorem we prove results by a straightforward generalization of the corre-
sponding theorem from the ordinary case. In the second case, however, the
theorem and its proof in fuzzy setting is technically considerably more com-
plicated compared to the ordinary case.
94 2 Algebras with Fuzzy Equalities

Theorem 2.81. Let M be an L-algebra and N ⊆ M . Then for any set X of


variables satisfying |N | ≤ |X| we have
 
[N ]M = tM (a1 , . . . , ak ) | t ∈ T (X), a1 , . . . , ak ∈ N . (2.17)

Proof. Denote the right hand side of (2.17) by N . Take a variable x ∈ T (X).
For each a ∈ N we have xM (a ) = a ∈ N , i.e. N ⊆ N . Furthermore, we
show that N is closed under all functions f M ∈ F M . Indeed, take an n-ary
f M ∈ F M and tM 1 (a11 , . . . , a1k1 ), . . . , tn (an1 , . . . , ankn ) ∈ N , where t1 , . . . , tn
M

are terms t1 (x11 , . . . , x1k1 ), . . . , tn (xn1 , . . . , xnkn ). We may safely assume that
aij = akl iff xij = xkl . Now we clearly have that
f M (tM M M
1 (a11 , . . . , a1k1 ), . . . , tn (an1 , . . . , ankn )) = t (a1 , . . . , am )

for some a1 , . . . , am ∈ N and t = f (t1 , . . . , tn ) ∈ T (X). Thus, N is a sub-


universe of M which contains N . Since [N ]M is the least subuniverse of M
containing N , we have [N ]M ⊆ N . Conversely, each a ∈ N can be obtained
by applying finitely many functions on finitely many elements from N , i.e.,
N ⊆ [N ]M . Hence, [N ]M = N , proving (2.17). 


The following theorem generalizes the well-known Mal’cev lemma describ-


ing principal congruences. Note that although on verbal level (and for the
truth degree a = 1) the theorem reads the same way as in the ordinary case,
the proof is technically much more involved in fuzzy setting.

Theorem 2.82. Suppose M is an L-algebra of type F and let c , d ∈ M .


For every set of variables X = {x, y1 , . . . , yk }, terms p1 , . . . , pn ∈ T (X) and
arbitrary elements s1 , t1 , . . . , sn , tn ∈ M such that {sj , tj } = {c , d }, 1 ≤ j ≤
n and b , b  , e1 , . . . , ek ∈ M let Π (b , b  , p1 , . . . , pn , s , t , e ) denote
n−1  M  M 
b ≈M pM 1 (s1 , e ) ⊗ i=1 pi (ti , e ) ≈
M M
pi+1 (si+1 , e ) ⊗ pM n (tn , e ) ≈ b ,

where e is an abbreviation for e1 , . . . , ek , s is an abbreviation for s1 , . . . , sn


and t is an abbreviation for t1 , . . . , tn . Then for θ(a/c , d ) ∈ ConL (M) and
every b , b  ∈ M we have
  n 
θ(a/c , d )(b , b  ) = a i=1 |pi |x ⊗ Π (b , b  , p1 , . . . , pn , s , t , e ) . (2.18)
X={x,y1 ,...,yk }; e1 ,...,ek ∈M
p1 ,...,pn ∈ T (X); {sj ,tj } ={c ,d }, 1 ≤ j ≤ n

Proof. First, let θ∗ be an L-relation on M defined by the right side of (2.18).


We have to check that θ(a/c , d ) = θ∗ . In what follows we use the fact that
j with ai ≤ bj and for every bi there is aj such that
if for every aithere is b
bi ≤ aj then i∈I ai = j∈J bj .
“θ(a/c , d ) ⊆ θ∗ ”: It is sufficient to prove that θ∗ (c , d ) ≥ a and

θ ∈ ConL (M). The first condition is obvious. Indeed, take a term p(x) = x,
elements s1 = c , t1 = d , and observe that
   
θ∗ (c , d ) ≥ a|p|x ⊗ c ≈M pM (c ) ⊗ pM (d ) ≈M d = a1 ⊗ 1 ⊗ 1 = a .
2.4 Terms, Term L-Algebras 95

Thus, θ∗ (c , d ) ≥ a. Moreover, take b , b  ∈ M and a binary term p(x, y) = y.


It follows that
   
θ∗ (b , b  ) ≥ a|p|x ⊗ b ≈M pM (c , b  ) ⊗ pM (d , b  ) ≈M b  =
= a0 ⊗ (b ≈M b  ) ⊗ 1 = b ≈M b  ,
i.e. b ≈M b  ≤ θ∗ (b , b  ). Clearly, we can put b = b  , thus the foregoing
inequality yields 1 = b ≈M b ≤ θ∗ (b , b ). Hence, θ∗ is reflexive.
Furthermore, for every set of variables X = {x, y1 , . . . , yk }, arbitrary terms
p1 , . . . , pn ∈ T (X) and elements s1 , t1 , . . . , sn , tn ∈ M such that {sj , tj } =
{c , d }, 1 ≤ j ≤ n and b , b  , e1 , . . . , ek ∈ M it is easy to see that
n−1  M  M 
b ≈M pM 1 (s1 , e ) ⊗ i=1 pi (ti , e ) ≈ pi+1 (si+1 , e ) ⊗ pM
M M
n (tn , e ) ≈ b =
n−2  M 
= b  ≈M pM n (tn , e ) ⊗ i=0 pn−i (sn−i , e ) ≈
M M
pn−i−1 (tn−i−1 , e ) ⊗
1 (s1 , e ) ≈ b .
⊗ pM M

Hence, the following equality holds,


n
|pi |x
a i=1 ⊗ Π (b , b  , p1 , . . . , pn , s , t , e ) =
n
|pi |x
=a i=1 ⊗ Π (b  , b , pn , . . . , p1 , t , s , e ) ,
where s is an abbreviation for sn . . . , s1 and t is an abbreviation for tn , . . . , t1 .
This equality ensures that θ∗ (b , b  ) = θ∗ (b  , b ), i.e. θ∗ is symmetric.
Consider b , b  , b  ∈ M , two sets of variables X1 = {x1 , y11 , . . . , y1k1 },
X2 = {x2 , y21 , . . . , y2k2 }, terms p11 , . . . , p1n1 ∈ T (X1 ), p21 , . . . , p2n2 ∈ T (X2 )
and elements s11 , t11 , . . . , s1n1 , t1n1 ∈ M , s21 , t21 , . . . , s2n2 , t2n2 ∈ M abbre-
viated by s1 , t1 , s2 , t2 such that {s1j , t1j } = {c , d }, {s2l , t2l } = {c , d },
1 ≤ j ≤ n1 , 1 ≤ l ≤ n2 . Moreover, let us have e11 , . . . , e1k1 ∈ M ,
e21 , . . . , e2k2 ∈ M abbreviated by e1 and e2 . Without loss of generality, we
can assume that x1 = x2 and (X1 − {x1 }) ∩ (X2 − {x2 }) = ∅ (if this is not so,
one can rename variables in X1 and X2 accordingly). In other words, all terms
p11 , . . . , p1n1 , p21 , . . . , p2n2 can be thought of as terms in variables X1 ∪ X2 .
Using transitivity of ≈M we have
 n1 −1  M 
b ≈M pM 11 (s11 , e1 ) ⊗ i=1 p1i (t1i , e1 ) ≈M pM 1(i+1) (s1(i+1) , e1 ) ⊗
M 
 
⊗ pM 1n1 (t1n1 , e1 ) ≈ b ⊗ b  ≈M pM 21 (s21 , e2 ) ⊗

n2 −1   M 

⊗ i=1 2i (t2i , e2 ) ≈
pM p2(i+1) (s2(i+1) , e2 ) ⊗ pM
M M
2n2 (t2n2 , e2 ) ≈ b ≤
n1 −1  M 
≤ b ≈M 11 (s11 , e1 ) ⊗
pM i=1 p1i (t1i , e1 ) ≈M pM
1(i+1) (s1(i+1) , e1 ) ⊗

⊗ pM 1n1 (t1n1 , e1 ) ≈ p21 (s21 , e2 ) ⊗


M M
n2 −1  M  M 
⊗ i=1 p2i (t2i , e2 ) ≈M pM 2(i+1) (s2(i+1) , e2 ) ⊗ p2n2 (t2n2 , e2 ) ≈ b .
M

Hence, it follows that


96 2 Algebras with Fuzzy Equalities
 n1
|p1i |x1
a i=1 ⊗ Π (b , b  , p11 , . . . , p1n1 , s1 , t1 , e1 ) ⊗
 n2
|p2i |x2
⊗a i=1 ⊗ Π (b  , b  , p21 , . . . , p2n2 , s2 , t2 , e2 ) ≤
 n1  n2
≤ a i=1 |p1i |x1 ⊗ a i=1 |p2i |x2 ⊗ Π (b , b  , p11 , . . . , p1n1 , p21 , . . . , p2n2 , s , t , e ) =
 n  n2 
= a i=1 |p1i |x1 + i=1 |p2i |x2 ⊗ Π (b , b  , p11 , . . . , p1n1 , p21 , . . . , p2n2 , s , t , e ) ,
1

where e is an abbreviation for elements e11 , . . . , e1k1 , e21 , . . . , e2k2 , s denotes


s11 , . . . , s1n1 , s21 , . . . , s2n2 , analogously for t . Taking into account the defini-
tion of θ∗ , we can conclude that θ∗ (b , b  ) ⊗ θ∗ (b  , b  ) ≤ θ∗ (b , b  ), i.e. θ∗ is
transitive.
Now we will check compatibility with operations of M. Take any m-ary
f M ∈ F M and elements a1 , b1 , . . . , am , bm ∈ M . For every 1 ≤ i ≤ m let us
have a set Xi = {xi , yi1 , . . . , yiki } of variables, terms pi1 , . . . , pini ∈ T (Xi ),
arbitrary elements si1 , ti1 , . . . , sini , tini ∈ M such that {sij , tij } = {c , d }
(1 ≤ j ≤ ni ) denoted by si , ti and elements ei1 , . . . , eiki ∈ M denoted by ei .
Again, we can assume that Xi , Xj are pairwise disjoint for every 1 ≤ i, j ≤ m,
i = j, i.e. for every 1 ≤ i ≤ m, 1 ≤ l ≤ ni we have pil ∈ T (X  ), where
X  = X1 ∪ · · · ∪ Xm .
Put n = max(n1 , . . . , nm ). We are going to show that we can enlarge every
sequence pi1 , . . . , pini of terms by defining new terms and suitable elements
to obtain a sequence pi1 , . . . , pini , pi(ni +1) , . . . , pin which satisfies certain con-
ditions. After that, all sequences have the same length, so it will be possible
to apply the compatibility of f M with ≈M simultaneously. First of all, put
X = X  ∪{z1 , . . . , zm }, where z1 , . . . , zm ∈ X  . All terms pij can be thought as
terms of the form pij (x1 , y11 , . . . , y1k1 , x2 , y21 , . . . , ymkm , z1 , . . . , zm ). For every
sequence pi1 , . . . , pini such that ni < n, let pij = zi for all ni < j ≤ n. We can
conclude that
 ni
|pij |xi
a j=1 ⊗ Π (ai , bi , pi1 , . . . , pini , si , ti , ei ) =
n
|pij |xi
=a j=1 ⊗ Π (ai , bi , pi1 , . . . , pin , si , ti , e ) ,
where e denotes e11 , . . . , emkm , b1 , . . . , bm , furthermore, si denotes si1 , . . . , sin ,
and ti denotes elements ti1 , . . . , tin such that {sij , tij } = {c , d }, ni < j ≤ n.
Indeed,
 ni
since thenew terms pij = zi do not have any occurrence of xi , we have
a j=1 |pij |xi = a j=1 |pij |xi and the values of pM
n
ij ’s do not depend on newly
defined elements sij , tij for j > ni . Thus,
 M 
ini (tini , ei ) ≈ bi = pini (tini , ei ) ≈ bi ⊗ 
pM 1 ⊗ 1 ⊗ ··· ⊗ 1 =
M M
 
(n−ni )-times
 
= pM ini (tini , e ) ≈ pi(ni +1) (si(ni +1) , e ) ⊗
M M
   
⊗ pM
i(ni +1) (ti(ni +1) , e ) ≈
M M
pi(ni +2) (si(ni +2) , e ) ⊗ · · · ⊗ pM in (tin , e ) ≈ bi .
M

Hence, it is possible to state


2.4 Terms, Term L-Algebras 97
m  n−1  
j=1aj ≈M pM j1 (sj1 , e ) ⊗ i=1 pji (tji , e ) ≈
M M
j(i+1) (sj(i+1) , e ) ⊗
pM

⊗ pMjn (tjn , e ) ≈
M
bj ≤
 
≤ f M (a1 , . . . , am ) ≈M f M pM 11 (s11 , e ), . . . , pm1 (sm1 , e ) ⊗
M
n−1    M
⊗ i=1 f M pM 1i (t1i , e ), . . . , pmi (tmi , e ) ≈
M
 
≈M f M pM M
1(i+1) (s1(i+1) , e ), . . . , pm(i+1) (sm(i+1) , e ) ⊗
  M M
⊗ f p1n (t1n , e ), . . . , pmn (tmn , e ) ≈ f (b1 , . . . , bm ) .
M M M

 
Moreover, the value of every f M pM M
1i (s1i , e ), . . . , pmi (smi , e ) can be expressed 
as the resulting value of the term f p1i (x1 , y), p2i (x2i , y) . . . , pmi (xmi , y) in
variables (X ∪ {xji | j ≥ 2}) − {x2 , . . . , xm  } for elements eMtogether with ele-
ments {tji , sji | j ≥ 2}. In the case of f M pM 1i (t1i , e ), . . . , pmi (tmi , e ) , we can
proceed analogously. Thus, we can claim,
m  ni |pij |x
i ⊗
i=1 a Π (ai , bi , pi1 , . . . , pini , si , ti , ei ) =
j=1

  
m
= a i=1 j=1 |pij |xi ⊗ i=1 Π (ai , bi , pi1 , . . . , pin , si , ti , e ) ≤
m n

n m
≤ a j=1 |p1j |x1 ⊗ i=1 Π (ai , bi , pi1 , . . . , pin , si , ti , e ) ≤
n
|p1j |x1
≤a j=1 ⊗
⊗ Π(f (a1 , . . . , am ), f M (b1 , . . . , bm ) ,
M

f (p11 , . . . , pm1 ), . . . , f (p1n , . . . , pmn ), s1 , t1 , e ) .


n
Clearly, f (p11 , . . . , pm1 ), . . . , f (p1n , . . . , pmn ) are terms and j=1 |p1j |x1 is the
number of occurrences of the variable x1 in these terms. This observation
yields the compatibility of f M with θ∗ . Together with the foregoing results,
θ∗ ∈ ConL (M) and θ∗ (c , d ) ≥ a. Consequently, θ(a/c , d ) ⊆ θ∗ .
“θ(a/c , d ) ⊇ θ∗ ”: Take a set X = {x, y1 , . . . , yk } of variables, terms
p1 , . . . , pn ∈ T (X) and arbitrary elements s1 , t1 , . . . , sn , tn ∈ M abbreviated
by s , t such that {sj , tj } = {c , d }, 1 ≤ j ≤ n and b , b  , e1 , . . . , ek ∈ M ,
where e is an abbreviation for e1 , . . . , ek . It is easy to see that
n
|pi |x
a i=1 = a|p1 |x ⊗ · · · ⊗ a|pn |x ≤
≤ θ(a/c , d )(c , d )|p1 |x ⊗ · · · ⊗ θ(a/c , d )(c , d )|pn |x .
 
Theorem 2.76 yields θ(a/c , d )(c , d )|p|x ≤ θ(a/c , d ) pM (c, e ), pM (d , e ) for
every term p(x, y) and any b , b  , e1 , . . . , ek ∈ M . Thus, we have
98 2 Algebras with Fuzzy Equalities
n n
|pi |x
a i=1 ⊗ Π (b , b  , p1 , . . . , pn , s , t , e ) = a i=1 |pi |x ⊗ b ≈M pM 1 (s1 , e ) ⊗
n−1  M  M 
⊗ i=1 pi (ti , e ) ≈M pM i+1 (si+1 , e ) ⊗ pn (tn , e ) ≈
M
b ≤
n
|pi |x
 
≤a i=1 ⊗ θ( /c , d ) b , p1 (s1 , e ) ⊗
a M
n−1   
⊗ i=1 θ(a/c , d ) pM M
i (ti , e ), pi+1 (si+1 , e ) ⊗
 M 

⊗ θ(a/c , d ) pn (tn , e ), b ≤
  |p1 |x
≤ θ(a/c , d ) b , pM
1 (s1 , e ) ⊗ θ( /c , d )(c , d )
a ⊗
n−1   M 
⊗ i=1 θ(a/c , d ) pi (ti , e ), pM i+1 (si+1 , e ) ⊗
|pi+1 |x
  

⊗ θ( /c , d )(c , d )
a ⊗ θ( /c , d ) pM
a
n (tn , e ), b ≤
   M 
≤ θ(a/c , d ) b , p1 (s1 , e ) ⊗ θ(a/c , d ) p1 (s1 , e ), p1 (t1 , e ) ⊗
M M
n−1   
⊗ i=1 θ(a/c , d ) pM i (ti , e ), pi+1 (si+1 , e ) ⊗
M
   

⊗ θ(a/c , d ) pM M
i+1 (si+1 , e ), pi+1 (ti+1 , e ) ⊗ θ(a/c , d ) pM n (tn , e ), b .
Hence, we can apply the transitivity of θ(a/c , d ) repeatedly to obtain
n
|pi |x
a i=1 ⊗ Π (b , b  , p1 , . . . , pn , s , t , e ) ≤ θ(a/c , d )(b , b  ) .
Since X = {x, y1 , . . . , yk }, p1 , . . . , pn ∈ T (X) and b , b  , s , s1 , t1 , . . . , sn , tn ∈
M , e1 , . . . , ek ∈ M have been chosen arbitrarily, we have proved the desired
inequality.
Altogether, θ∗ = θ(a/c , d ). 


2.5 Direct Unions



In general, there is no obvious way to naturally equip the union i∈I Mi of
universes of L-algebras Mi with functions so that it became  an L-algebra. In
some
 cases, however, it is possible to define functions on i∈I Mi such that
i∈I M i equipped with such functions and suitable L-equality turns into an
L-algebra. For instance, when {Mi | i ∈ I} is a directed family of L-algebras,
we can define a direct union of {Mi | i ∈ I} which is a well-defined L-algebra.

Definition 2.83. A partially ordered index set I, ≤ is called directed, if


I = ∅ and for every i, j ∈ I there is k ∈ I such that i, j ≤ k. A family
{Mi | i ∈ I} of L-algebras of type F , where I, ≤ is a directed index set and
Mi ∈ Sub(Mj ) whenever i ≤ j is called a directed family of L-algebras.
Let κ be an infinite cardinal. A family {Mi | i ∈ I} = ∅ of L-algebras of
type F is called a κ-directed family, if for every J ⊆ I, |J| < κ there exists
an index i ∈ I such that Mj ∈ Sub(Mi ) for all j ∈ J.

Lemma 2.84. Let κ be an infinite cardinal. Then


(i) every κ-directed family {Mi | i ∈ I} is a directed family for some partial
order ≤ on I;
(ii) every directed family is an ω-directed family.
2.5 Direct Unions 99

Proof. (i): Let {Mi | i ∈ I} be a κ-directed family of L-algebras. It is easily


seen that we can define a partial order ≤ on I by
i≤j iff Mi ∈ Sub(Mj ) . (2.19)
Moreover, for J = {i, j} ⊆ I we have |J| = 2 < κ, i.e. there is some k ∈ I such
that Mi , Mj ∈ Sub(Mk ), that is i, j ≤ k. Hence, I, ≤ is a directed index set
and {Mi | i ∈ I} is a directed family of L-algebras.
(ii): Let {Mi | i ∈ I} be a directed family of L-algebras. Let J ⊆ I, |J| < ω,
i.e. J = {i1 , . . . , in }. Using the induction on elements of J, it follows that there
is an index i ∈ I such that Mi1 , . . . , Min ∈ Sub(Mi ). Thus, {Mi | i ∈ I} is a
ω-directed family of L-algebras. 


 {Mi | i ∈ I} M
Definition 2.85. For a directed family of L-algebras
  of type F
we define an L-algebra i∈I Mi = i∈I M i , ≈ i∈I

i
, F i∈I Mi such that
for every n-ary f ∈ F , and arbitrary a1 , . . . , an ∈ i∈I Mi we put

Mi
f i∈I (a1 , . . . , an ) = f Mj (a1 , . . . , an ) , (2.20)
where a1 ∈ Mi1 , . . . , an ∈ Min , j ∈ I, and i1 , . . . , in ≤ j. For every elements
a , b ∈ i∈I Mi such that a ∈ Mi , b ∈ Mj we define a ≈ i∈I Mi b by

(a ≈ b ) = (a ≈Mk b ) ,
i∈I Mi
(2.21)

where k ∈ I, i, j ≤ k. The L-algebra i∈I Mi is called a direct union of a
directed family {Mi | i ∈ I} of L-algebras.
Given a κ-directed family {Mi | i ∈ I}, the direct union of a directed family
{Mi | i ∈ I} with a partial order ≤ on I definedκ by (2.19) is called the κ-direct
union of {Mi | i ∈ I} and is denoted by i∈I Mi .

Remark 2.86. (1) The direct union i∈I Mi of a directed family {Mi | i ∈ I}
is a well-defined L-algebra. First observe, that for finitely many i1 , . . . , in ∈ I
there is always an index j ∈ I such that i1 , . . . , in ≤ j (this follows from the
definition of directed index set). Moreover,{Mi | i ∈ I} is a directed family
so we have Mi1, . . . , Min ∈ Sub(Mj ), i.e. f i∈I Mi (a1 , . . . , an ) always
 exists.
The fact that i∈I Mi equipped with functions f i∈I Mi ∈ F i∈I Mi is an
algebra follows from the  ordinary case.
The
 L-relation ≈ i∈I Mi is well-defined. Indeed, it suffices to check that

a ≈ i∈I Mi b is independent on the choice of indices i, j, k ∈ I used in (2.21).


Let a ∈ Mi , a ∈ Mi , b ∈ Mj , b ∈ Mj  . Take k ≥ i, j and k  ≥ i , j  . Since
I, ≤ is directed, there is an index l ∈ I such that l ≥ k, k  . As a consequence,
Mk , Mk ∈ Sub(Ml ). Thus, it follows that
(a ≈Mk b ) = (a ≈Ml b ) = (a ≈Mk b ) .

Therefore, a ≈ i∈I Mi b is  independent on the choice of i, j, k ∈ I.
Furthermore, for a ∈ i∈I Mi we 
have a ∈ Mj for some j ∈ I. Since
a ≈Mj a = 1, it follows that a ≈ i∈I Mi a = 1 (reflexivity). Symmetry
follows from the symmetry of all ≈Mi ’s. For every a , b , c ∈ M there is an
index j ∈ I such that a , b , c ∈ Mj , thus
100 2 Algebras with Fuzzy Equalities
 
(a ≈ i∈I Mi
b ) ⊗ (b ≈ i∈I Mi
c ) = (a ≈Mj b ) ⊗ (b ≈Mj c ) ≤

≤ (a ≈Mj c ) = (a ≈ i∈I Mi
c) ,
 
i.e. ≈ i∈I is transitive. Moreover, if a ≈ i∈I
Mi
b = 1 then for j ∈ I such
Mi

that a , b ∈ Mj we have ( a ≈Mj


b ) = 1. Thus, (a ≈ i∈I Mi b ) = 1 implies
a = b . Altogether, ≈ i∈I i is an L-equality.
M
 
i∈I Mi ∈ F i∈I Mi is compatible with
 It suffices to verify, that every f 
≈ i∈I i . For every n-ary f ∈ F and arbitrary a1 , b1 , . . . , an , bn ∈ i∈I Mi
M

there is an index j ∈ I such that a1 , b1 , . . . , an , bn ∈ Mj . Thus,


 
(a1 ≈ i∈I Mi
b1 ) ⊗ · · · ⊗ (an ≈ i∈I Mi
bn ) =
= (a1 ≈ Mj
b1 ) ⊗ · · · ⊗ (an ≈Mj bn ) ≤
≤ f Mj (a1 , . . . , an ) ≈Mj f Mj (b1 , . . . , bn ) =
  
=f i∈I Mi
(a1 , . . . , an ) ≈ i∈I Mi
f i∈I Mi
(b1 , . . . , bn ) .

Hence, i∈I Mi is a well-defined L-algebra.
(2) Clearly,every L-algebra M is isomorphic to a trivial direct union.
Namely, M ∼ = i∈I Mi , where I = {1} and M1 is M.
(3) If {Mi | i ∈ I} is a directed family of L-algebras, where I is finite, then
evidently I has the greatest element, let us denote it by k. Clearly, for every
i ∈ I we have Mi ∈ Sub(Mk ), that is i∈I Mi ∼ = Mk .

The following theorem presents a representation of L-algebras as κ-direct


unions of κ-directed families of subalgebras generated by less than κ genera-
tors. In particular, the theorem says that every L-algebra can be composed
from its finitely generated subalgebras. Recall that by a κ-generated L-algebra
we mean M such that M = [M  ]M for some M  ⊆ M with |M  | < κ.

Theorem 2.87. Let κ be any infinite cardinal. Every L-algebra is isomorphic


to a κ-direct union of a κ-directed family of its κ-generated subalgebras.

Proof. Let M be an L-algebra. Put IM = {M  ⊆ M | κ > |M  |}. For each


M  ∈ IM we identify [M  ]M with the subalgebra of M with universe [M  ]M .
Obviously, each [M  ]M is κ-generated. Furthermore,  for J ⊆  IM2 such that
|J| < κ, we have |M  | < κ for every M  ∈ J. Thus,  M  ∈J M 
 < κ = κ. As a
 
  
 
consequence, M  ∈J M ∈ IM . It is evident that [M ]M ∈ Sub M  ∈J M M
for every M  ∈ J. So, {[M  ]M | M  ∈ IM } is a κ-directed family of κ-generated
subalgebras of M.
Now observe that a ∈ M is contained in[{a }]M and where {a } ∈ IM . It is
thus possible to define a mapping h : M → {[M  ]M | M  ∈ IM } by h(a ) = a
(a ∈ M ). For all a , b ∈ M we have

{[M  ]M | M  ∈IM }
a ≈M b = a ≈[{a ,b }]M b = h(a ) ≈ h(b ) ,
i.e. h is an ≈-morphism. For any n-ary f ∈ F and a1 , . . . , an ∈ M we have,
2.5 Direct Unions 101
   
h f M (a1 , . . . , an ) = h f [{a1 ,...,an }]M (a1 , . . . , an ) =
    
= f [{a1 ,...,an }]M (a1 , . . . , an ) = f {[M ]M | M ∈IM } h(a1 ), . . . , h(an ) .
Since for each finite M  ⊆ M ,[M  ]M is a subalgebra of M, h is a surjective
embedding which yields M ∼ = {[M  ]M | M  ∈ IM }. 

Theorem 2.87 and Lemma 2.84 give the following
Corollary 2.88. Every L-algebra is isomorphic to the direct union of its fi-
nitely generated subalgebras. 

The following asserion says that any direct union can be thought of as a
derived construction since it is isomorphic to a factorization of certain subal-
gebra of a direct product.
Lemma  2.89. Let {Mi | i ∈ I} be a directed family of L-algebras of type F .
Then i∈I Mi ∼ = N/θ, where
Q
(i) N ∈ Sub(
 M ), where 
Qi∈I i
N = a ∈ i∈I M i | there is k ∈ I  that a (k) = a (i) for all i ≥ k ,
such 
(ii) θ ∈ ConL (N) such that θ(a , b ) = i∈I k≥i a (k) ≈Mk b (k).
Proof. For the sake of brevity, let a ∈ N(j) denote the fact that a (j) = a (i)
for all i ≥ j. Clearly, if a ∈ N(j) , then a ∈ N(l)Qfor every l ≥ j. First, we
will check that N is a non-empty subuniverse of i∈I Mi . Evidently, N = ∅.
Furthermore, for any n-ary f ∈ F and arbitrary elements a1 , . . . , an ∈ N there
is an index j ∈ I, such that a1 ∈ N(j) , . . . , an ∈ N(j) . Since Mj ∈ Sub(Mi )
for all i ≥ j, it follows that
Q  
f i∈I Mi (a1 , . . . , an )(j) = f Mj a1 (j), . . . , an (j) =
  Q
= f Mi a1 (i), . . . , an (i) = f i∈I Mi (a1 , . . . , an )(i)
Q Q
for every i ≥ j, i.e. f i∈I Mi (a1 , . . . , an ) ∈ N(j) . Hence, N ∈ Sub( i∈I Mi ).
In the the rest of the proof, we will take advantage of the following claim:
for every elements a ∈ N(j) , b ∈ N(j  ) we have
θ(a , b ) = a (m) ≈Mm b (m) for every m ≥ j, j  . (2.22)

“≤” of (2.22): Take m ≥ j, j , where a ∈ N(j) , b ∈ N(j  ) . For arbitrary
index i ∈ I there is some m ∈ I such that i, m ≤ m . Moreover, from
a (m ) = a (m), b (m ) = b (m), Mm ∈ Sub(Mm ) it follows that

k≥i a (k) ≈
Mk
b (k) ≤ a (m ) ≈Mm b (m ) = a (m) ≈Mm b (m) .
 
Hence, we have θ(a , b ) = i∈I k≥i a (k) ≈Mk b (k) ≤ a (m) ≈Mm b (m).
“≥” of (2.22): Let us have a ∈ N(j) , b ∈ N(j  ) and let m ≥ j, j  . We
can take i = m. Since  Mm ∈ Sub(Mk ) for all k ≥ m, we readily obtain
a (m) ≈Mm b (m) = k≥m a (k) ≈Mk b (k). That is,
 
a (m) ≈Mm b (m) ≤ i∈I k≥i a (k) ≈Mk b (k) = θ(a , b ) .
102 2 Algebras with Fuzzy Equalities

Altogether, claim (2.22) holds true.


Now it is easy to check that θ ∈ ConL (N). Reflexivity and symmetry of
θ are obvious. For every a , b , c ∈ N , we have θ(a , b ) = a (m) ≈Mm b (m),
θ(b , c ) = b (m ) ≈Mm c (m ) for certain m, m ∈ I due to (2.22). Evidently,
a , b ∈ N(m) , b , c ∈ N(m ) . Thus, for m ≥ m, m , we have a , b , c ∈ N(m ) .
Using this observation, it follows that
   
θ(a , b ) ⊗ θ(b , c ) = a (m) ≈Mm b (m) ⊗ b (m ) ≈Mm c (m ) =
   
= a (m ) ≈Mm b (m ) ⊗ b (m ) ≈Mm c (m ) ≤
≤ a (m ) ≈Mm c (m ) = θ(a , c ) .
Hence, θ is transitive. In an analogous way it is possible to prove that every
n-ary f N ∈ F N is compatible with θ. For a1 , b1 ∈ N(m1 ) , . . . , an , bn ∈ N(mn ) ,
and m ∈ I such that m ≥ m1 , . . . , mn we have
n     
i=1 ai (mi ) ≈
M mi
bi (mi ) = ni=1 ai (m) ≈Mm bi (m) ≤
   
≤ f Mm a1 (m), . . . , an (m) ≈Mm f Mm b1 (m), . . . , bn (m) =
= f N (a1 , . . . , an )(m) ≈Mm f N (b1 , . . . , bn )(m) =
 
= θ f N (a1 , . . . , an ), f N (b1 , . . . , bn ) .
That is, we have θ ∈ ConL (N).
Let us introduce mappings hi : Mi → N/θ (i ∈ I) defined by hi (a ) = [a  ]θ ,
where a  ∈ N such that for every j ≥ i we have a  (j) = a . Every hi is a
well-defined mapping since for a  , a  ∈ N , where a  (j) = a  (j) = a for all
j ≥ i we have θ(a  , a  ) = 1, i.e. [a  ]θ = [a  ]θ . Furthermore, for any n-ary
f Mi ∈ F Mi and arbitrary a1 , . . . , an ∈ Mi it is obvious that
   
f N/θ hi (a1 ), . . . , hi (an ) = f N/θ [a1 ]θ , . . . , [an ]θ =
   
= f N (a1 , . . . , an ) θ = hi f Mi (a1 , . . . , an ) .
Hence, every hi : Mi → N/θ is compatible with operations. Moreover, hi
is a restriction of hjon Mi whenever i ≤ j. Now it is possible to de-
fine a mapping h : i∈I Mi → N/θ by h(a ) = hi (a ) for all a ∈ Mi ,
   
i ∈ I. Clearly,
 f
N/θ
h(a1 ), . . . , h(an ) = h f Mi (a1 , . . . , an ) . Furthermore,
for a , b ∈ i∈I Mi , a ∈ Mi , b ∈ Mj we can take k ≥ i, j, thus

a≈ i∈I Mi
b = a ≈Mk b = a  (k) ≈Mk b  (k) ,
where a  , b  ∈ N such that for every l ≥ k we have a  (l) = a , b  (l) = b , i.e.
a  , b  ∈ N(k) . Moreover, we can apply (2.22) to obtain
a  (k) ≈Mk b  (k) = θ(a  , b  ) = [a  ]θ ≈N/θ [b  ]θ =
= hk (a ) ≈N/θ hk (b ) = hi (a ) ≈N/θ hj (b ) = h(a ) ≈N/θ h(b ) .
Consequently, h is an embedding. Furthermore, for any [b ]θ ∈ N/θ there is
an index i ∈ I such that b (j) = a for some a ∈ Miand every j ≥ i. Thus,
h(a ) = hi (a ) = [b ]θ , i.e. h is surjective. Altogether, i∈I Mi ∼
= N/θ. 

2.6 Direct Limits 103

2.6 Direct Limits


This section focuses on the construction of a direct limit. In ordinary case,
every algebra is isomorphic to a direct limit of finitely presented algebras. In
the subsequent development we will present an analogous characterization for
L-algebras. Furthermore, direct limits are important from the point of view
of quasivariety theory, see Sect. 4.8.
We begin our development with the definition of a (weak ) direct family.
This notion is similar to that one from ordinary case, but we have to postulate
an additional condition (which holds true automatically in ordinary case) to
be able to generalize some required properties of direct limits.

Definition 2.90. A weak direct family of L-algebras of type F consists of:


(i) a directed index set I, ≤ ,
(ii) a family {Mi | i ∈ I} of pairwise disjoint L-algebras of type F ,
(iii) a family {hij : Mi → Mj | i ≤ j} of morphisms, where
hii = idMi for every i ∈ I , (2.23)
hik = hij ◦ hjk for all i, j, k ∈ I, where i ≤ j ≤ k . (2.24)
A weak direct family is called a direct family if for every a ∈ Mi , b ∈ Mj
there exists k ∈ I, i, j ≤ k such that for each l ∈ I, k ≤ l we have
hik (a ) ≈Mk hjk (b ) = hil (a ) ≈Ml hjl (b ) . (2.25)

Remark 2.91. (1) A (weak) direct family of L-algebras is usually denoted sim-
ply by {Mi | i ∈ I}. If there is no danger of confusion, we will not mention the
morphisms hij : Mi → Mj explicitly.
(2) In general, there are weak direct families which do not satisfy (2.25).
Take L with L = [0, 1] as a structure of truth degrees and a family {Mi | i ∈
[0, 1)} of L-algebras, where Mi = {ai , bi }, ≈Mi , ∅ , and ai ≈Mi bi = i.
Furthermore, morphisms hij : Mi → Mj (i ≤ j) defined by hij (ai ) = aj ,
hij (bi ) = bj evidently satisfy (2.23) and (2.24). Therefore, [0, 1), ≤ together
with {M i | i ∈ [0, 1)} and hij ’s is a weak direct family. On the other hand, for
ai , bj ∈ m∈[0,1) Mm , and every k ≥ i, j there is l > k, i.e.

hik (ai ) ≈Mk hjk (bj ) = ak ≈Mk bk = k <


< l = al ≈Ml bl = hil (ai ) ≈Ml hjl (bj ) ,
showing that {Mi | i ∈ [0, 1)} is not a direct family.

Lemma 2.92. Let {Mi | i ∈ I} be a weak direct family of L-algebras. For


every a ∈ Mi , b ∈ Mj and arbitrary k, l ∈ I such that i, j ≤ k ≤ l we
have
hik (a ) ≈Mk hjk (b ) ≤ hil (a ) ≈Ml hjl (b ), (2.26)
 
k≥i,j hik (a ) ≈ hjk (b ) = m≥l him (a ) ≈Mm hjm (b ) .
Mk
(2.27)
104 2 Algebras with Fuzzy Equalities

Proof. Equation (2.26): Using (2.24) we have


hik (a ) ≈Mk hjk (b ) ≤ hkl (hik (a )) ≈Ml hkl (hjk (b )) =
= (hik ◦ hkl )(a ) ≈Ml (hik ◦ hkl )(b ) = hil (a ) ≈Ml hjl (b ) .
Equation (2.27): Take an index k0 ≥ i, j. For m0 ∈ I such that m0 ≥ k0 , l
we can use (2.26) to get
hik0 (a ) ≈Mk0 hjk0 (b ) ≤ him0 (a ) ≈Mm0 hjm0 (b ) ≤

≤ m≥l him (a ) ≈Mm hjm (b ) ,
by which follows the “≤”-part of (2.27). The converse inequality is trivial. 


Remark 2.93. If L is a Noetherian residuated lattice then every weak direct


family is a direct family. Indeed, for any a ∈ Mi and b ∈ Mj there are indices
k1 , . . . , kn ≥ i, j such that
 n
k≥i,j hik (a ) ≈ hjk (b ) = m=1 hikm (a ) ≈Mkm hjkm (b )
Mk

Thus, we can take k ∈ I with k ≥ k1 , . . . , kn . Now (2.26) gives


hikm (a ) ≈Mkm hjkm (b ) ≤ hik (a ) ≈Mk hjk (b )
for each m = 1, . . . , n. Therefore, hik (a ) ≈Mk hjk (b ) is the greatest one of all
hik (a ) ≈Mk hjk (b ) for k  ≥ i, j. Since for each l ≥ k we have
hik (a ) ≈Mk hjk (b ) ≤ hil (a ) ≈Ml hjl (b )
by (2.26), it follows that in fact hik (a ) ≈Mk hjk (b ) = hil (a ) ≈Ml hjl (b ) for
all l ≥ k proving (2.25).

Definition 2.94. For every weak  direct family of L-algebras {Mi | i ∈ I} let
θ∞ denote a binary L-relation on i∈I Mi defined by

θ∞ (a , b ) = k≥i,j hik (a ) ≈Mk hjk (b ) (2.28)
for all a ∈ Mi , b ∈ Mj .

Remark 2.95. If {Mi | i ∈ I} is a direct family, then (2.28) can be equivalently


expressed without using the general suprema. Indeed, taking into account
(2.25) and (2.27), for any a ∈ Mi , b ∈ Mj there is k0 ≥ i, j such that
 
θ∞ (a , b ) = k≥i,j hik (a ) ≈Mk hjk (b ) = m≥k0 him (a ) ≈Mm hjm (b ) =
= hik0 (a ) ≈Mk0 hjk0 (b ) .

Lemma 2.96. Let F be a type of L-algebras and let {Mi | i ∈ I} be a weak


direct family of L-algebras of type F . The following are properties of θ∞ :
 
(i) θ∞ (a , b ) = θ∞ hil (a ), hjl (b) for all a ∈ Mi , b ∈ Mj , l ≥ i, j;
(ii) θ∞ is an L-equivalence
 on i∈I Mi ;
(iii) θ∞ a , hik (a ) = 1 for every a ∈ Mi , k ≥ i;
2.6 Direct Limits 105

(iv) for every n-ary f ∈ F , a1 ∈ Mi1 , b1 ∈ Mj1 , . . . , an ∈ Min , bn ∈ Mjn , and


k ≥ i1 , j1 , . . . , in , jn we have
θ∞ (a1 , b1 ) ⊗ · · · ⊗ θ∞ (an , bn ) ≤
 
≤ θ∞ f Mk (hi1 k (a1 ), . . . , hin k (an )), f Mk (hj1 k (b1 ), . . . , hjn k (bn )) .
Proof. (i): Clearly, using (2.27) we have
 
θ∞ (a , b ) = k≥i,j hik (a ) ≈Mk hjk (b ) = m≥l him (a ) ≈Mm hjm (b ) =

= m≥l hlm (hil (a )) ≈Mm hlm (hjl (b )) = θ∞ (hil (a ), hjl (b )) .

(ii): We have θ∞ (a , a ) = k≥i hik (a ) ≈Mk hik (a ) = 1 for every a ∈ Mi ,
i.e. θ∞ is reflexive. Symmetry is obvious. Hence, it suffices to check transitivity.
Let a ∈ Mi , b ∈ Mj , c ∈ Mk , and let l ≥ i, j, k. Furthermore, using (2.26)
and (i) together with monotony of ⊗ it follows that
   
θ∞ (a , b ) ⊗ θ∞ (b , c ) = θ∞ (hil a ), hjl (b ) ⊗ θ∞ (hjl b ), hkl (c ) =
 
= m≥l him (a ) ≈Mm hjm (b ) ⊗ n≥l hjn (b ) ≈Mn hkn (c ) =
  
= m,n≥l him (a ) ≈Mm hjm (b ) ⊗ hjn (b ) ≈Mn hkn (c ) =
  
= n≥l hin (a ) ≈Mn hjn (b ) ⊗ hjn (b ) ≈Mn hkn (c ) ≤

≤ n≥l hin (a ) ≈Mn hkn (c ) = θ∞ (a , c ) .
Hence, θ∞ is an L-equivalence.
(iii): Let us have a ∈ Mi , k ≥ i. Take l ∈ I such that l ≥ k, i. From
reflexivity of θ∞ together with (i) it follows that
 
θ∞ (a , hik (a )) = θ∞ hil (a ), hkl (hik (a )) = θ∞ (hil (a ), hil (a )) = 1 .
(iv): For an n-ary f ∈ F , arbitrary am ∈ Mim , bm ∈ Mjm , m = 1, . . . , n,
and k ≥ i1 , j1 , . . . , in , jn , we can use the compatibility of f Ml with ≈Ml
assumed for every l ∈ I to get
θ∞ (a1 , b1 ) ⊗ · · · ⊗ θ∞ (an , bn ) =
n 
= m=1 km ≥k him km (am ) ≈Mkm hjm km (bm ) =
 n
= k1 ,...,km ≥k m=1 him km (am ) ≈Mkm hjm km (bm ) =
 n
= l≥k m=1 him l (am ) ≈Ml hjm l (bm ) ≤
    
≤ l≥k f Ml hi1 l (a1 ), . . . , hin l (an ) ≈Ml f Ml hj1 l (b1 ), . . . , hjn l (bn ) =
  
= l≥k f Ml hkl (hi1 k (a1 )), . . . , hkl (hin k (an )) ≈Ml
 
f Ml hkl (hj1 k (b1 )), . . . , hkl (hjn k (bn )) =
  
= l≥k hkl f Mk (hi1 k (a1 ), . . . , hin k (an )) ≈Ml
 
hkl f Mk (hj1 k (b1 ), . . . , hjn k (bn )) =
 
= θ∞ f Mk (hi1 k (a1 ), . . . , hin k (an )), f Mk (hj1 k (b1 ), . . . , hjn k (bn )) ,
which is the desired inequality. 

106 2 Algebras with Fuzzy Equalities

Condition (iv) of Lemma 2.96 is similar to that of compatibility, but in


this case, (iv) expresses a compatibility with respect to homomorphic
 images.
Now we define suitable operations on the factorization of i∈I Mi by θ∞ .

Definition 2.97. Let {Mi | i ∈ I} be a (weak  ) direct  family of L-algebras 


of type F . Then the L-algebra lim Mi = i∈I M i /θ ∞ , ≈ lim Mi
, F lim Mi ,
where
  
(i) i∈I M
 i /θ∞ is a factorization  of i∈I Mi by θ∞ , 
(ii) f lim Mi [a1 ]θ∞ , . . . , [an ]θ∞ = f Mk (hi1 k (a1 ), . . . , hin k (an )) θ
 ∞ 
for every n-ary f ∈ F and arbitrary [a1 ]θ∞ , . . . , [an ]θ∞ ∈ i∈I Mi /θ∞
such that a1 ∈ Mi1 , . . . , an ∈ Min , and k ∈ I, k ≥ i1 , .. . , in , 
(iii) [a ]θ∞ ≈lim Mi [b ]θ∞ = θ∞ (a , b ) for all [a ]θ∞ , [b ]θ∞ ∈ i∈I Mi /θ∞

is called a direct limit of a (weak ) direct family {Mi | i ∈ I}.

Remark 2.98. A direct limit lim Mi of a weak direct family {Mi | i ∈ I} is a


well-defined L-algebra. Obviously, ≈lim Mi is an L-equality, see Definition 1.91
and Remark 1.92. It remains to show that each f lim Mi is well defined and 
compatible with ≈lim Mi . First, we show that f Mk (hi1 k (a1 ), . . . , hin k (an )) θ

given by (iii) does not depend on the chosen k ∈ I. Thus, take k  ∈ I with
k  ≥ i1 , . . . , in , and arbitrary l ≥ k, k  . Lemma 2.96 gives
 
θ∞ f Mk (hi1 k (a1 ), . . . , hin k (an )), f Ml (hi1 l (a1 ), . . . , hin l (an )) =
  
= θ∞ f Mk (hi1 k (a1 ), . . . , hin k (an )), hkl f Mk (hi1 k (a1 ), . . . , hin k (an )) = 1 .
   
That is, f Mk (hi1 k (a1 ), . . . , hin k (an )) θ = f Ml (hi1 l (a1 ), . . . , hin l (an )) θ
∞ ∞
and analogously for k  . Hence,
 M   
f k (hi1 k (a1 ), . . . , hin k (an )) θ∞ = f Mk (hi1 k (a1 ), . . . , hin k (an )) θ∞ .
 
Moreover, f lim Mi [a1 ]θ∞ , . . . , [an ]θ∞ does not depend on a1 , . . . , an chosen
from classes [a1 ]θ∞ , . . . , [an ]θ∞ , because for bm ∈ [am ]θ∞ , bm ∈ Mjm (m =
1, . . . , n), and k ≥ i1 , j1 , . . . , in , jn we have
1 = θ∞ (a1 , b1 ) ⊗ · · · ⊗ θ∞ (an , bn ) ≤
 
≤ θ∞ f Mk (hi1 k (a1 ), . . . , hin k (an )), f Mk (hj1 k (b1 ), . . . , hjn k (bn )) =
   
= f Mk (hi1 k (a1 ), . . . , hin k (an )) θ∞ ≈lim Mi f Mk (hj1 k (b1 ), . . . , hjn k (bn )) θ∞ .
   
Hence, f Mk (hi1 k (a1 ), . . . , hin k (an )) θ = f Mk (hj1 k (b1 ), . . . , hjn k (bn )) θ ,
∞ ∞
i.e. f lim Mi is well defined.
It remains to check the compatibility.  Takef
lim Mi
∈ F lim Mi and arbitrary
[a1 ]θ∞ , [b1 ]θ∞ , . . . , [an ]θ∞ , [bn ]θ∞ ∈ i∈I Mi /θ∞ , where am ∈ Mim , bm ∈
Mjm (m = 1, . . . , n). For k ∈ I such that k ≥ i1 , j1 , . . . , in , jn , Lemma 2.96
together with the definition of ≈lim Mi yield
2.6 Direct Limits 107

[a1 ]θ∞ ≈lim Mi [b1 ]θ∞ ⊗ · · · ⊗ [an ]θ∞ ≈lim Mi [bn ]θ∞ =
= θ∞ (a1 , b1 ) ⊗ · · · ⊗ θ∞ (an , bn ) ≤
 
≤ θ∞ f Mk (hi1 k (a1 ), . . . , hin k (an )), f Mk (hj1 k (b1 ), . . . , hjn k (bn )) =
 
= f Mk (hi1 k (a1 ), . . . , hin k (an )) θ ≈lim Mi

 M 
f k (hj1 k (b1 ), . . . , hjn k (bn )) θ =

  lim Mi lim Mi  
=f lim Mi
[a1 ]θ∞ , . . . , [an ]θ∞ ≈ f [b1 ]θ∞ , . . . , [bn ]θ∞ .
Hence, lim Mi is a well-defined L-algebra.

Lemma 2.99. Let lim Mi be a weak direct limit  of {M  i | i ∈ I}. Then for
every n-ary f lim Mi , and [a1 ]θ∞ , . . . , [an ]θ∞ ∈ i∈I M i /θ∞ we have
   
f lim Mi [a1 ]θ∞ , . . . , [an ]θ∞ = f Mk (a1 , . . . , an ) θ , (2.29)

where a1 , . . . , an ∈ Mk , and 


am ∈ [am ]θ∞ for each m = 1, . . . , n.

Proof. Since θ∞ (am , am ) = 1 for each m = 1, . . . , n, we can take an index
l ∈ I such that l ≥ k, i1 , . . . , in and apply Definition 2.97 and Lemma 2.96:
 M 
f k (a1 , . . . , an ) θ =
  M ∞   
= hkl f k (a1 , . . . , an ) θ = f Ml (hkl (a1 ), . . . , hkl (an )) θ =
∞ ∞
   
= f Ml (hi1 l (a1 ), . . . , hin l (an )) θ = f lim Mi [a1 ]θ∞ , . . . , [an ]θ∞ ,

proving (2.29). 


Definition 2.100. Let {Mi | i ∈ I} be a weak direct family of L-algebras of


type F . A family {hi : Mi → lim Mi | i ∈ I} of morphisms hi : Mi → lim Mi ,
where hi (a ) = [a ]θ∞ for every i ∈ I, a ∈ Mi is called a limit cone of the
weak direct family {Mi | i ∈ I}.
Let N be an L-algebra of type F . A family {gi : Mi → N | i ∈ I} of mor-
phisms is said to satisfy a direct limit property (DLP) with respect to a
weak direct family {Mi | i ∈ I} if gi = hij ◦ gj for all i ≤ j and for every
family {gi : Mi → N | i ∈ I} of morphisms with gi = hij ◦ gj for all i ≤ j,
there exists a unique morphism g : N → N such that gi = gi ◦ g for every
i ∈ I, see Fig. 2.14.

hij hij
Mi Mj Mi Mj
gi gi
gi gj gj
N N g N

Fig. 2.14. Direct limit property


108 2 Algebras with Fuzzy Equalities

Remark 2.101. Every hi : Mi → lim Mi of a limit cone of {Mi | i ∈ I} is indeed


a morphism. Clearly, for all i ∈ I, a , b ∈ Mi we have

a ≈Mi b ≤ k≥i hik (a ) ≈Mk hik (b ) = θ∞ (a , b ) =
= [a ]θ∞ ≈lim Mi [b ]θ∞ = hi (a ) ≈lim Mi hi (b ) .
Furthermore, for an n-ary f ∈ F , and a1 , . . . , an ∈ Mi :
   
hi f Mi (a1 , . . . , an ) = f Mi (a1 , . . . , an ) θ =

 Mi  lim Mi
 
= f (hii (a1 ), . . . , hii (an )) θ = f [a1 ]θ∞ , . . . , [an ]θ∞ =
 ∞
= f lim Mi hi (a1 ), . . . , hi (an ) ,
i.e. hi is a morphism. Moreover, Lemma 2.96 yields θ∞ (hij (a ), a ) = 1, that
is hi (a ) = [a ]θ∞ = [hij (a )]θ∞ = hj (hij (a )), i.e. hi = hij ◦ hj .

Remark 2.102. If I, ≤ is a finite directed index set, then I has the greatest
element. In consequence, every weak direct family {Mi | i ∈ I} is a direct
family since for every a ∈ Mi , b ∈ Mj (2.25) is satisfied trivially for k being
the greatest element of I. Consequently, θ∞ (a , b ) = hik (a ) ≈Mk hjk (b ).
Moreover, for hk : Mk → lim Mi we have
a ≈Mk b = hkk (a ) ≈Mk hkk (b ) = θ∞ (a , b ) = hk (a ) ≈lim Mi hk (b ) ,
 
and for every [c ]θ∞ ∈ i∈I Mi /θ∞ , c ∈ Mi we have

hk (hik (c )) = [hik (c )]θ∞ = [c ]θ∞ .


Hence, hk is an isomorphism (hk is compatible with operations since it is a
part of the limit cone), Mk ∼
= lim Mi . In other words, the direct limit is trivial
for finite I, ≤ .

Theorem 2.103. Let {Mi | i ∈ I} be a weak direct family of L-algebras. Then


the limit cone {hi : Mi → lim Mi | i ∈ I} of {Mi | i ∈ I} satisfies the direct
limit property with respect to {Mi | i ∈ I}.

Proof. Let us have a family {gi : Mi → N | i ∈ I} of morphisms such that gi =


hij ◦ gj for all i ≤ j. We have to check the existence and uniqueness of a
morphism h : lim Mi → N, where  gi = hi ◦ h for every i ∈ I.
Existence of h: Every
 a ∈  i∈I Mi belongs to some Mi , hence we can
define a mapping h : i∈I M i /θ∞ → N by
 
h [a ]θ∞ ) = gi (a , where a ∈ Mi . (2.30)
For every a ∈ Mi , b ∈ Mj we have

[a ]θ∞ ≈lim Mi [b ]θ∞ = θ∞ (a , b ) = k≥i,j hik (a ) ≈Mk hjk (b ) ≤

≤ k≥i,j gk (hik (a )) ≈N gk (hjk (b )) =

= k≥i,j gi (a ) ≈N gj (b ) = gi (a ) ≈N gj (b )
2.6 Direct Limits 109

Thus, [a ]θ∞ = [b ]θ∞ implies gi (a ) = gj (b ). Hence, h defined by (2.30) is a


well-defined ≈-morphism.  
Now for every n-ary f ∈ F , and [a1 ]θ∞ , . . . , [an ]θ∞ ∈ i∈I Mi /θ∞ there
are indices i1 , . . . , in ≤ l such that a1 ∈ Mi1 , . . . , an ∈ Min . Hence, using
Lemma 2.96 it follows that
     
h f lim Mi ([a1 ]θ∞ , . . . , [an ]θ∞ ) = h f Ml hi1 l (a1 ), . . . , hin l (an ) θ∞ =
   
= gl f Ml (hi1 l (a1 ), . . . , hin l (an )) = f N gl (hi1 l (a1 )), . . . , gl (hin l (an )) =
   
= f N gi1 (a1 ), . . . , gin (an ) = f N h([a1 ]θ∞ ), . . . , h([an ]θ∞ ) .
That is, h : lim Mi → N is the required morphism satisfying gi (a ) =
h([a ]θ∞ ) = h(hi (a )) = (hi ◦ h)(a ) for all a ∈ Mi , i ∈ I.
 
 of h: Let h : lim Mi → N be a morphism satisfying gi = hi ◦h .
Uniqueness
Every a ∈ i∈I Mi belongs to Mi for a suitable i ∈ I, that is
   
h [a ]θ∞ = h (hi (a )) = gi (a ) = h(hi (a )) = h [a ]θ∞ .
Therefore, h is determined uniquely. 


Theorem 2.104. Let {Mi | i ∈ I} be a weak direct family of L-algebras of type


F and let N be an L-algebra of type F . There is a family {gi : Mi → N | i ∈ I}
of morphisms satisfying DLP w.r.t. {Mi | i ∈ I} iff N ∼ = lim Mi .
Proof. “⇒”: For N being lim Mi , the limit cone {hi : Mi → lim Mi | i ∈ I}
satisfies direct limit property w.r.t. {Mi | i ∈ I}. Hence, it is sufficient to prove
that for any families {gi : Mi → N | i ∈ I}, {gi : Mi → N | i ∈ I} satisfying
DLP w.r.t. {Mi | i ∈ I} we have N ∼ = N .
Thus, if both {gi : Mi → N | i ∈ I} and {gi : Mi → N | i ∈ I} satisfy DLP
w.r.t. {Mi | i ∈ I}, there are uniquely determined morphisms g : N → N ,
g  : N → N, where gi = gi ◦ g  , gi = gi ◦ g, see Fig. 2.15. Consequently
gi = gi ◦ g  = (gi ◦ g) ◦ g  = gi ◦ (g ◦ g  ) ,
thus g ◦ g  = idN , similarly g  ◦ g = idN . So, we can apply Theorem 2.30 to
claim that g, g  are mutually inverse isomorphisms between N and N . Thus,
for N being lim Mi it follows N ∼ = lim Mi .
“⇐”: Obviously, for N ∼ = lim Mi we can take morphisms gi : Mi → N
such that gi = hi ◦ h, where h : lim Mi → N is an isomorphism. It is routine
to check that {gi : Mi → N} satisfies DLP w.r.t. {Mi | i ∈ I}. 


Mi Mi
gi gi gi gi

N N N N
g g
Fig. 2.15. Morphisms between N and N
110 2 Algebras with Fuzzy Equalities

Definition 2.105. Every L-algebra M such that M = ∼ T(X)/θ(R), where


R ∈ LT (X)×T (X) is said to be presented by X, R . If both X and R are
finite, then M is said to be finitely presented by X, R .

Remark 2.106. Note that every L-algebra of type F is isomorphic to T(X)/θ,


where T(X) is a term L-algebra of type F , X is a suitable set of variables, and
θ ∈ ConL (T(X)). Indeed, we can take |X| ≥ |M |, and a surjective mapping
h : X → M . Due to Theorem 2.79, there exists a surjective morphism h :
T(X) → M. Finally, from Theorem 2.35 it follows that M ∼ = T(X)/θh .
Consequently, M ∼ = T(X)/θ(R) for R = θ h . From this point of view, every
L-algebra is presented by some X, R . Let us stress that such X and R are
not finite in general.

Theorem 2.107. Every L-algebra is isomorphic to a direct limit of a direct


family of finitely presented L-algebras.

Proof. Let M be an L-algebra. Due to Remark 2.106, we can identify M with


T(X)/θ(R) for some set of variables X and R ∈ LT (X)×T (X) . Recall that for
every Y ⊆ X we can consider a restriction θ(R)|T(Y ) of θ(R) on T(Y ). Now
let us assume an index set

I = Y, S | Y ⊆ X, Y is finite,

S is a finite restriction of θ(R)|T(Y ) .
We can define a partial order ≤ on I by
Yi , Si ≤ Yj , Sj iff Yi ⊆ Yj and Si ⊆ Sj .
It is easily seen that I, ≤ is directed. For the sake of brevity, we will denote
indices of the form Yi , Si , Yj , Sj , . . . simply by i, j, . . .
We introduce
 morphisms
 hij : T(Yi )/θ(Si ) → T(Yj )/θ(Sj ) (i ≤ j) de-
fined by hij [t]θ(Si ) = [t]θ(Sj ) , see Lemma 2.43. Clearly, every hij satisfies
(2.23) and (2.24). Thus, I, ≤ , {T(Yi )/θ(Si ) | i ∈ I} together with hij ’s is a
weak direct family. Moreover, it is even a direct family. Indeed, take elements
[ti ]θ(Si ) ∈ T(Yi )/θ(Si ), [tj ]θ(Sj ) ∈ T(Yj )/θ(Sj ). There is k ≥ i, j such that
Yk = Yi ∪ Yj and Sk (ti , tj ) = θ(R)(ti , tj ). Clearly, for every l ≥ k we have
   
hik [ti ]θ(Si ) ≈T(Yk )/θ(Sk ) hjk [tj ]θ(Sj ) = [ti ]θ(Sk ) ≈T(Yk )/θ(Sk ) [tj ]θ(Sk ) =
= θ(Sk )(ti , tj ) = θ(R)(ti , tj ) = θ(Sl )(ti , tj ) = [ti ]θ(Sl ) ≈T(Yl )/θ(Sl ) [tj ]θ(Sl ) =
   
= hil [ti ]θ(Si ) ≈T(Yl )/θ(Sl ) hjl [tj ]θ(Sj ) ,
showing that I, ≤ , {T(Yi )/θ(Si ) | i ∈ I} together with hij ’s is a direct family.
Now we show that there is a family {hi : T(Yi )/θ(Si ) → T(X)/θ(R) | i ∈ I}
of morphisms satisfying DLP w.r.t. {T(Yi )/θ(Si ) | i ∈ I}. Then T(X)/θ(R) ∼ =
lim T(Yi )/θ(S
 i ) on
 account of Theorem 2.104.
Put hi [t]θ(Si ) = [t]θ(R) for all t ∈ T (Yi ). Due to Lemma 2.43, every hi is
a morphism. Evidently, hi = hij ◦ hj for all i, j ∈ I with i ≤ j. Take a family
2.6 Direct Limits 111

{gi : T(Yi )/θ(Si ) → N | i ∈ I} of morphisms with gi = hij ◦ gj (i ≤ j), and


define h : T(X)/θ(R) → N by
   
h [t]θ(R) = gi [t]θ(Si ) , (2.31)
where t ∈ T (X), and i ∈ I such that var(t) ⊆ Yi . Let us note that the value
of h [t]θ(R) does not depend on the chosen index i ∈ I since for every i, j ∈ I
such that var(t) ⊆ Yi , Yj we can take l ∈ I, i, j ≤ l. Thus,
         
gi [t]θ(Si ) ≈N gj [t]θ(Sj ) = gl hil [t]θ(Si ) ≈N gl hjl [t]θ(Sj ) =
   
= gl [t]θ(Sl ) ≈N gl [t]θ(Sl ) = 1 ,
   
i.e. gi [t]θ(Si ) = gj [t]θ(Sj ) . Moreover, h is a well-defined ≈-morphism. In-
deed, take ti , tj ∈ T (X) such that var(ti ) ⊆ Yi , and var(tj ) ⊆ Yj . For k ∈ I,
k ≥ i, j such that θ(R)(ti , tj ) = θ(Sk )(ti , tj ), it follows that
[ti ]θ(R) ≈T(X)/θ(R) [tj ]θ(R) = θ(R)(ti , tj ) = θ(Sk )(ti , tj ) =
   
= [ti ]θ(Sk ) ≈T(Yk )/θ(Sk ) [tj ]θ(Sk ) ≤ gk [ti ]θ(Sk ) ≈N gk [tj ]θ(Sk ) =
     
= gk hik [ti ]θ(Si ) ≈N gk hjk [tj ]θ(Sj ) =
   
= gi [ti ]θ(Si ) ≈N gj [tj ]θ(Sj ) .
   
Hence, [ti ]θ(R) = [tj ]θ(R) implies gi [ti ]θ(Si ) = gj [tj ]θ(Sj ) . Altogehter, h is
a well-defined ≈-morphism. Furthermore, take any n-ary f ∈ F and arbi-
trary elements [t1 ]θ(R) , . . . , [tn ]θ(R) . We can consider an index k ∈ I for which
var(ti ) ⊆ Yk for each i = 1, . . . , n. Thus,
   
h f T(X)/θ(R) ([t1 ]θ(R) , . . . , [tn ]θ(R) ) = h [f (t1 , . . . , tn )]θ(R) =
   
= gk [f (t1 , . . . , tn )]θ(Sk ) = gk f T(Yk )/θ(Sk ) ([t1 ]θ(Sk ) , . . . , [tn ]θ(Sk ) ) =
         
= f N gk [t1 ]θ(Sk ) , . . . , gk [t1 ]θ(Sk ) = f N h [t1 ]θ(R) , . . . , h [tn ]θ(R) .
Hence, h is a morphism. In addition to that,
      
gi [t]θ(Si ) = h [t]θ(R) = h hi [t]θ(Si ) ,
for every t ∈ T (Yi ), i.e. gi = hi ◦ h.
Finally, we will check the uniqueness of h. Let h : T(X)/θ(R) → N be
a morphism satisfying gi = hi ◦ h for all i ∈ I. It is immediate, that for
t ∈ T (X), var(t) ⊆ Yi we have
           
h [t]θ(R) = h hi [t]θ(Si ) = gi [t]θ(Si ) = h hi [t]θ(Si ) = h [t]θ(R) .
Altogether, h : T(X)/θ(R) → N is a uniquely determined morphism such that
gi = hi ◦ h for every i ≤ j. Consequently, M ∼
= lim T(Yi )/θ(Si ). 


Theorem 2.108. Let {Mi | ∈ I} be a direct family of L-algebras. For a mor-


phism h : N → lim Mi from a finitely presented L-algebra N to lim Mi , there
exists k ∈ I and a morphism g : N → Mk such that h = g ◦ hk .
112 2 Algebras with Fuzzy Equalities

Fig. 2.16. Morphism g from T(X)/θ(R) to some Mk

Proof. Since N is supposed to be finitely presented, we can identify N with


some T(X)/θ(R), where X, R are finite. Thus, let us assume a morphism
h : T(X)/θ(R) → lim Mi is given.  
It is obvious that T(X)/θ(R) is generated by [x]θ(R) | x ∈ X . For every
 
variable x ∈ X there is an index ix ∈ I such that h [x]θ(R) ∈ hix (Mix ). Since
there are only finitely many variables in X, we can choose an index j ∈ I such
that j ≥ ix for each x ∈ X. Clearly,
 
h [x]θ(R) ∈ hix (Mix ) = hj (hix j (Mix )) ⊆ hj (Mj )
for each x ∈ X. Therefore, h(hθ(R) (X)) ⊆ hj (Mj ), where hθ(R) is the natural
morphism sending elements of T (X) to T (X)/θ(R).  Following
 this observa-
tion, for each x ∈ X there is ax ∈ Mj such that h [x]θ(R) = hj (ax ) ∈ hj (Mj ).
Hence, we introduce a mapping  v: X  → Mj by putting v(x) = ax (x ∈ X).
By definition, hj (v(x)) = h [x]θ(R) for each x ∈ X. Since for v  we have
 
hj (v  (t)) = h [t]θ(R) for all t ∈ T (X), it follows that v  ◦ hj = hθ(R) ◦ h. The
situation is depicted in Fig. 2.16.
Recall that R is finite, i.e. Supp(R) = {t1 , t1 , . . . , tm , tm }. Since
{Mi | ∈ I} is a direct family of L-algebras, it follows from (2.25) that for
each i = 1, . . . , m there is an index ki ∈ I such that θ∞ (v  (ti ), v  (ti )) =
hjki (v  (ti )) ≈Mki hjki (v  (ti )), see Remark 2.95. Hence, for an index k ∈ I
such that k ≥ k1 , . . . , km we have
   
θ(R)(ti , ti ) = [ti ]θ(R) ≈T(X)/θ(R) [ti ]θ(R) ≤ h [ti ]θ(R) ≈lim Mi h [ti ]θ(R) =
= hj (v  (ti )) ≈lim Mi hj (v  (ti )) = θ∞ (v  (ti ), v  (ti )) =
= hjki (v  (ti )) ≈Mki hjki (v  (ti )) ≤ hjk (v  (ti )) ≈Mk hjk (v  (ti )) .
Thus, R(ti , ti ) ≤ hjk (v  (ti )) ≈Mk hjk (v  (ti )) = θv ◦hjk (ti , ti ) for each i =
1, . . . , m. Since θv ◦hjk ∈ ConL (Mk ), and θ(R) is generated by R, we readily
obtain θ(R) ⊆ θv ◦hjk .
 
Finally, put g [t]θ(R) = (v  ◦ hjk )(t) = hjk (v  (t)). For [t]θ(R) , [t ]θ(R) ∈
T (X)/θ(R) it follows that
[t]θ(R) ≈T(X)/θ(R) [t ]θ(R) = θ(R)(t, t ) ≤ θv ◦hjk (t, t ) =
   
= hjk (v  (t)) ≈Mk hjk (v  (t )) = g [t]θ(R) ≈Mk g [t ]θ(R) ,
2.6 Direct Limits 113

Fig. 2.17. Direct limit of a weak direct family

i.e. g is a well-defined ≈-morphism. For any n-ary f ∈ F and arbitrary ele-


ments [t1 ]θ(R) , . . . , [tn ]θ(R) ∈ T (X)/θ(R) we have
    
g f T(X)/θ(R) [t1 ]θ(R) , . . . , [tn ]θ(R) = g [f (t1 , . . . , tn )]θ(R) =
     
= hjk v  f (t1 , . . . , tn ) = hjk f Mj v  (t1 ), . . . , v  (tn ) =
   
= f Mk hjk (v  (t1 )), . . . , hjk (v  (tn )) = f Mk g([t1 ]θ(R) ), . . . , g([tn ]θ(R) ) .
Hence, g : T(X)/θ(R) → Mk is a morphism. In addition to that,
      
h [t]θ(R) = hj (v  (t)) = hk (hjk (v  (t))) = hk g [t]θ(R) = (g ◦ hk ) [t]θ(R)
holds for every [t]θ(R) ∈ T (X)/θ(R), i.e. h = g ◦ hk . 


Remark 2.109. Let us stress that the existence of a morphism described by


Theorem 2.108 is limited to direct families {Mi | ∈ I}. In bivalent case, every
weak direct family is a direct family, i.e. Theorem 2.108 then coincides with
the well-known image factorization theorem for ordinary algebras. It follows
from Remark 2.93 that for Noetherian residuated lattices, Theorem 2.108 is
true for all weak direct families. However, the following example illustrates
that postulating (2.25) is necessary for general residuated lattices.

Example 2.110.Take L = [0, 1]. Let us have a family {Mi | i ∈ N} of L-


algebras Mi = Mi , ≈Mi , ∅ such that Mi = {ai , bi }, and ai ≈Mi bi = 2i+1 i
.
That is, a1 ≈ b1 = 3 , a2 ≈ b2 = 5 , a3 ≈ b3 = 7 , . . . , see Fig. 2.17.
M1 1 M2 2 M3 3

Clearly, N, ≤ is a directed index set, the universe sets Mi (i ∈ N)


are pairwise disjoint, and {hij : Mi → Mj | i ≤ j}, where hij (ai ) = aj , and
hij (bi ) = bj is a family of morphisms satisfying (2.23) and (2.24). Altogether,
N, ≤ with {Mi | i ∈ N}, and {hij : Mi → Mj | i ≤ j} is a weak direct family.
On the other hand, it is not a direct family, because for i, k ∈ N with i ≤ k we
have hik (ai ) ≈Mk hik (bi ) < hi,k+1 (ai ) ≈Mk+1 hi,k+1 (bi ). Moreover, we have

θ∞ (ai , bi ) = k≥i hik (ai ) ≈Mk hik (bi ) = 12 ,
   
i.e. i∈N Mi /θ∞ = [ai ]θ∞ , [bi ]θ∞ , see Fig. 2.17. Since lim Mi is of the
empty type (F lim Mi = ∅), it readily follows that T (X) = X. Thus, for X =
{x, y}, and R ∈ LT (X)×T (X) , where R(x, y) = R(y, x) = 12 , and R(x, x) =
R(y, y) = 1, we have θ(R) = R. Therefore, T(X)/θ(R) is finitely presented.
114 2 Algebras with Fuzzy Equalities
 
Let h : T(X)/θ(R) → lim Mi be defined by h [x]θ(R) = [aj ]θ∞ and
 
h [y]θ(R) = [bj ]θ∞ . Clearly, h is an ≈-morphism and thus a morphism. Sup-
pose h = g ◦ hk , where g : T (X)/θ(R) → Mk and hk : Mk → lim Mi is a
morphism of the limit cone of lim Mi . In this case, h = g ◦ hk yields
   
g [x]θ(R) = ak and g [y]θ(R) = bk .

Thus, g cannot be an ≈-morphism since 1


2  k
2k+1 .

An important thing to stress is that our generalized direct limit has a


special property which is degenerated in ordinary case. If L is infinite and
{Mi | i ∈ I} is a weak direct family which is not a direct family, there can
be elements a ∈ Mi , b ∈ Mj the homomorphic images of which are distinct
in every Mk for k ≥ i, j. However, it can happen that θ∞ (a , b ) = 1, i.e.
[a ]θ∞ = [b ]θ∞ due to the general suprema used in (2.28). Such a situation
is apparently ill at least from the standpoint of compatibility with ordinary
direct limits. Indeed, the skeleton ske(lim Mi ) (i.e. an ordinary algebra being
the functional part of lim Mi ) is then not isomorphic to the ordinary direct
limit of skeletons ske(Mi ). On the other hand, such a situation cannot occur
for direct families of L-algebras.

Theorem 2.111. Let {Mi | i ∈ I} be a direct family of L-algebras. Then


ske(lim Mi ) ∼
= lim{ske(Mi ) | i ∈ I} . (2.32)

Proof. The claim is almost evident. It remains to show that for a ∈ Mi , b ∈


Mj we have θ∞ (a , b ) = 1 iff there exists k ≥ i, j such that hik (a ) = hjk (b )
since then i∈I Mi /θ∞ coincides with its crisp counterpart. So, assume
that θ∞ (a , b ) = 1. Since {Mi | i ∈ I} is a direct family, there is some index
k ≥ i, j such that hik (a ) ≈Mk hjk (b ) = 1, that is, hik (a ) = hjk (b ). The
converse implication holds trivially. Altogether, ske(lim Mi ) ∼ = lim ske(Mi )
since the corresponding functions on ske(lim Mi ) and lim ske(Mi ) are defined
the same way. 


Example 2.112. Consider L = [0, 1]. We can take a weak direct family from
Remark 2.91 (2) on page 103. It is evident that lim Mi is a trivial L-algebra
but there is not any j ∈ I such that hij (ai ) = hij (bi ). On the other
hand, lim ske(Mi ) is a two-element (ordinary) algebra. In consequence, The-
orem 2.111 is not true for general weak direct families of L-algebras.

∗ ∗ ∗
In the rest of this section, we show basic relationships between the notions
of a direct union, direct limit, and some of the notions introduced earlier. In
the literature, ordinary direct limit is often defined as a factorization of a spe-
cial subalgebra of a direct product. This generalization does not require the
algebras of a direct family to be pairwise disjoint. Since this approach uses
2.6 Direct Limits 115

only the notions we have already generalized for L-algebras (subalgebras, di-
rect products and congruences), we can introduce an alternative generalization
of a direct limit. In what follows, we will show that there is a natural rela-
tionship between both of these generalizations (the former approach and the
one introduced below).
For every directed index set I, ≤ and a family {hij : Mi → Mj | i ≤ j} of
morphisms satisfying (2.23), (2.24) we can define a set M by
 Q
M = a ∈ i∈I M i | there is i ∈ I such that for j, k ∈ I, i ≤ j ≤ k
 (2.33)
we have hjk (a (j)) = a (k) .
Q
In other words, M represents a subset of i∈I M i every element of which
respects morphisms {hij : Mi → Mj | i ≤ j}. Namely, for every a ∈ M there
is an index i ∈ I such that for j ∈ I, i ≤ j we have hij (a (i)) = a (j).
Furthermore, we define a binary L-relation θ on M by
 
θ (a , b ) = i∈I k≥i a (k) ≈Mk b (k) (2.34)
for every a , b ∈ M .

Theorem 2.113. Let I, ≤ be a directed index set, {hij : Mi → Mj | i ≤ j}


be a family
Q of morphisms satisfying conditions (2.23), (2.24). Then ∅ = M ∈
Sub( i∈I Mi ), and θ ∈ ConL (M ), where M is the subalgebra given by M .

Q
Proof. We will show that M is Q
a non-empty
Q
subuniverse of i∈I M i . Evi-
dently, M = ∅. For an n-ary f i∈I Mi ∈ F i∈I Mi and elements a1 , . . . , an ∈
M there exists j ∈ I such that hjk (ai (j)) = ai (k) for each i = 1, . . . , n and
k ≥ j. So, it follows that
 Q   
hjk f i∈I Mi (a1 , . . . , an )(j) = hjk f Mj (a1 (j), . . . , an (j)) =
   
= f Mk hjk (a1 (j)), . . . , hjk (an (j)) = f Mk a1 (k), . . . , an (k) =
Q
=f i∈I Mi (a1 , . . . , an )(k) .
Q Q
Hence, f i∈I Mi (a1 , . . . , an ) ∈ M . That is, ∅ = M ∈ Sub( i∈I M i ). There-
fore, we can consider the subalgebra M whose universe set is M .
Introduce binary L-relation θ defined on M by (2.34). Reflexivity and
symmetry of θ follows directly from reflexivity and symmetry of each ≈Mi .
Now, for a , b , c ∈ M we have
   
k≥i a (k) ≈ b (k) ⊗ j∈I l≥j b (l) ≈Ml c (l) ≤
Mk
i∈I
   
≤ i,j∈I k≥i, l≥j a (k) ≈Mk b (k) ⊗ b (l) ≈Ml c (l) ≤
   
≤ i,j∈I k≥i,j a (k) ≈Mk b (k) ⊗ b (k) ≈Mk c (k) ≤
   
≤ i,j∈I k≥i,j a (k) ≈Mk c (k) ≤ m∈I m ≥m a (m ) ≈Mm c (m ) ,

 i, j ∈ I we can choose
the last inequality follows from the fact that for every
an index m ∈ I such that m ≥ i, j, in which case k≥i,j a (k) ≈Mk c (k) ≤
116 2 Algebras with Fuzzy Equalities
  
m ≥m a (m ) ≈Mm c (m ) ≤ m∈I m ≥m a (m ) ≈Mm c (m ). Hence, θ is
transitive. Compatibility with operations can be checked analogously. Briefly,
 
for an n-ary f M ∈ F M and a1 , b1 , . . . , an , bn ∈ M we have
θ (a1 , b1 ) ⊗ · · · ⊗ θ (an , bn ) =
n  
= m=1 im ∈I km ≥im am (km ) ≈Mkm bm (km ) ≤
  n
≤ i1 ,...,in ∈I k≥i1 ,...,in m=1 am (k) ≈Mk bm (k) ≤
  n
≤ i∈I j≥i m=1 am (j) ≈Mj bm (j) ≤
 
≤ i∈I j≥i f Mj (a1 (j), . . . , an (j)) ≈Mj f Mj (b1 (j), . . . , bn (j)) =
   
= i∈I j≥i f M (a1 , . . . , an )(j) ≈Mj f M (b1 , . . . , bn )(j) =
   
= θ f M (a1 , . . . , an ), f M (b1 , . . . , bn ) .

Clearly, ≈M ⊆ θ . Altogether, θ ∈ ConL (M ). 


Lemma 2.114. Let {Mi | i ∈ I} be a weak direct family of L-algebras. Then


θ∞ (a , b ) = θ (a  , b  ), where a ∈ Mi , b ∈ Mj , a  , b  ∈ M , and hik (a ) =
a  (k), hjl (b ) = b  (l) for every k, l ∈ I, i ≤ k, j ≤ l.

Proof. Let us note that we have


   Ml   
k∈I l≥k a (l) ≈ b (l) = k≥i,j l≥k a  (l) ≈Ml b  (l) . (2.35)
Indeed, the “≥”-part of (2.35) is trivial. Conversely, for every k ∈ I there is
an index m ∈ I such that i, j, k ≤ m. Thus,
  Ml  
l≥k a (l) ≈ b (l) ≤ l≥m a  (l) ≈Ml b  (l)
implies the “≤”-part of (2.35). Now, using (2.26) and (2.35) we obtain
   
θ (a  , b  ) = k∈I l≥k a  (l) ≈Ml b  (l) = k≥i,j l≥k a  (l) ≈Ml b  (l) =
  
= k≥i,j l≥k hil (a ) ≈Ml hjl (b ) = k≥i,j hik (a ) ≈Mk hjk (b ) = θ∞ (a , b ) ,
which is the desired equality. 


Theorem 2.115. For every weak direct family {Mi | i ∈ I} of L-algebras we


have lim Mi ∼ = M /θ .
 
Proof. For a ∈ Mi put h [a ]θ∞ = [a  ]θ , where a  ∈ M satisfies a  (k) =
hik (a ) for each k ≥ i. It is easily seen that such an element a  ∈ M always
exists. Moreover, Lemma 2.114 yields
[a ]θ∞ ≈lim Mi [b ]θ∞ = θ∞ (a , b ) = θ (a  , b  ) = [a ]θ ≈M

/θ 
[b ]θ ,
  
where a ∈ Mi , b ∈ Mj and a , b ∈ M such that

hik (a ) = a (k), hjl (b ) =
b  (l) for every k, l ∈ I, i ≤ k, j ≤ l. Clearly, h : i∈I Mi /θ∞ → M is a

well-defined ≈-morphism. Furthermore, h is surjective since for every element


c ∈ M there
 is an index i ∈ I such that for all k ≥ i we have hik (c (i)) = c (k),
that is h [c (i)]θ∞ = [c ]θ .
2.6 Direct Limits 117
 
Take an n-ary f ∈ F and arbitrary elements a1 , . . . , an ∈ i∈I Mi /θ∞ .
We can take a1 , . . . , an ∈ M , where h ( a ) = a 
(k) for each m =

im k  m m
1, . . . , n, and k ≥ im . Since f M a1 , . . . , an ∈ M , there is j ∈ I such
     
that i1 , . . . , in ≤ j and hjk f M a1 , . . . , an (j) = f M a1 , . . . , an (k) for all
k ≥ j. Now, taking into account previous observations, we have
     
h f lim Mi [a1 ]θ∞ , . . . , [an ]θ∞ = h f Mj (hi1 j (a1 ), . . . , hin j (an )) θ =

 Mj      
= h f (a1 (j), . . . , an (j)) θ = h f M (a1 , . . . , an )(j) θ

=
∞ ∞
 M   
 
M /θ 
 

= f a1 , . . . , an θ = f [a1 ]θ , . . . , [an ]θ =
     
= f M /θ h [a1 ]θ∞ , . . . , h [an ]θ∞ .
Altogether h : lim Mi → M is an isomorphism, i.e. lim Mi ∼
= M . 


We established that every direct limit is isomorphic to M /θQ


, i.e. it can
be thought of as a special factorization of a suitable subalgebra of i∈I Mi . On
the other hand, the following theorem shows that every M /θ is isomorphic
to a direct limit of some weak direct family of L-algebras.

Theorem 2.116. Let I, ≤ be directed index set. Let {hij : Mi → Mj | i ≤ j}


be a family of morphisms satisfying (2.23) and (2.24). Then there is a weak
direct family {gij : Ni → Nj | i ≤ j} such that lim Ni ∼
= M /θ .
Proof. Suppose we have a family of morphisms {hij : Mi → Mj | i ≤ j} sat-
isfying conditions (2.23), (2.24). Since Mi ’s are not assumed to be pairwise-
disjoint, we will introduce a new family {Ni | i ∈ I} of L-algebras, where
Ni = {a , i | a ∈ Mi } ,
a , i ≈Ni b , i = a ≈Mi b ,
   
f Ni a1 , i , . . . , an , i = f Mi (a1 , . . . , an ), i ,
for every n-ary f ∈ F , a , b , a1 , . . . , an ∈ Mi , and i ∈ I. Evidently, Ni ∼
= Mi
and I, ≤ together with Ni ’s and {gij : Ni → Nj | i ≤ j}, where gij (a , i ) =
hij (a ), i is a weak direct family of L-algebras.
Moreover, Theorem 2.115 yields lim Ni ∼ = N /θN
, where θN ∈ ConL (N )

is defined by (2.34). Hence, it is sufficient to check that M /θ = N /θN
.

Evidently, M = N since every a ∈ M is in one-to-one correspondence


with a  ∈ N , where a  (i) = a (i), i for all i ∈ I. Clearly, h [a ]θ = [a  ]θ
. Thus, lim Ni ∼
N
defines an isomorphism h : M /θ → N /θN
= M /θ . 


Let us note that due to Lemma 2.89, every direct union i∈I Mi is iso-
morphic to N/θ, where N can be thought of as M for a special family of
morphisms. In consequence, it is possible to treat a direct union as a special
direct limit.

Theorem 2.117. Let {Mi | i ∈ I} be a directed family of L-algebras. Then


 ∼
i∈I Mi = lim Ni for some direct family {Ni | i ∈ I} of L-algebras.
118 2 Algebras with Fuzzy Equalities

Proof. For indices i, j ∈ I satisfying i ≤ j we put hij (a ) = a (a ∈ Mi ).


Clearly, {hij : Mi → Mj | i ≤ j}, is a family of morphisms satisfying (2.23),
(2.24), and (2.25). Due to Lemma 2.89, we have M ∼
i = N/θ. Hence,
 ∼
i∈I
i∈I M i = M /θ since N (see Lemma 2.89) is defined the same way as M

’s with hij (a ) = a . Trivially, θ = θ . Now we can use Theorem 2.116 to
for hij
claim i∈I Mi ∼ = M /θ ∼ = lim Ni , where {Ni | i ∈ I} is a weak direct family
of L-algebras defined as in the proof of Theorem 2.116. A moment’s reflection
shows that {Ni | i ∈ I} is a direct family since for a , i ∈ Ni , b , j ∈ Nj and
arbitrary k ≥ i, j we have
   
gik a , i ≈Nk gjk b , j = hik (a ), k ≈Nk hjk (b ), k =

= hik (a ) ≈Mk hjk (b ) = a ≈Mk b = a ≈ i∈I Mi
b,
i.e. {Ni | i ∈ I} is a direct family, completing the proof. 

The following assertion shows that direct limits can be constructed by
means of direct unions and factorizations.
Theorem 2.118. Every direct limit of a weak direct family of L-algebras is
isomorphic to a direct union of a directed family of L-algebras.
Proof. Let us have a weak direct family of L-algebras which consists of a
directed index set I, ≤ , a family of L-algebras {Mi | i ∈ I}, and a family of
morphisms {hij : Mi → Mj | i ≤ j}. Recall that for every hi : Mi → lim Mi
and the corresponding congruence θhi ∈ ConL (Mi ) we have

θhi (a , b ) = hi (a ) ≈lim Mi hi (b ) = θ∞ (a , b ) = k≥i hik (a ) ≈Mk hik (b ) ,
for all i ∈ I, a , b ∈ Mi . Hence, θhi is a restriction of θ∞ on Mi , i.e. θhi =
θ∞ |Mi . For a family {Mi /θhi | i ≤ j} of factor L-algebras we can define a
family of mappings gij : Mi /θhi → Mj /θhj such that gij ([a ]θhi ) = [hij (a )]θhj
for all i ≤ j. Thus, for every i, j ∈ I, i ≤ j and arbitrary a , b ∈ Mi we have
 
[a ]θhi ≈Mi /θhi [b ]θhi = θhi (a , b ) = θ∞ (a , b ) = θ∞ hij (a ), hij (b ) =
 
= θhj hij (a ), hij (b ) = [hij (a )]θhj ≈Mj /θhj [hij (b )]θhj =
   
= gij [a ]θhi ≈Mj /θhj gij [b ]θhi .
Therefore, every gij is a well-defined ≈-morphism. Moreover, for every n-ary
f Mi /θhi and arbitrary elements [a1 ]θhi , . . . , [an ]θhi ∈ Mi /θhi it follows that
     
gij f Mi /θhi [a1 ]θhi , . . . , [an ]θhi = gij f Mi (a1 , . . . , an ) θhi =
     
= hij f Mi (a1 , . . . , an ) θhj = f Mj hij (a1 ), . . . , hij (an ) θhj =
 
= f Mj /θhj [hij (a1 )]θhj , . . . , [hij (an )]θhj =
    
= f Mj /θhj gij [a1 ]θhi , . . . , gij [an ]θhi ,
i.e., gij (i ≤ j) are embeddings. We check that gij ’s satisfy conditions (2.23)
and (2.24). Equation (2.23) holds trivially and for i ≤ j ≤ k, [a ]θhi ∈ Mi /θhi
we have
2.6 Direct Limits 119
 
gik [a ]θhi = [hik (a )]θhk = [hjk (hij (a ))]θhk =
      
= gjk [hij (a )]θhj = gjk gij [a ]θhi = (gij ◦ gjk ) [a ]θhi ,
showing (2.24). Therefore, I, ≤ , {Mi /θhi | i ∈ I}, and morphisms gij (i ≤ j)
form a weak direct family. Now, we can consider a direct limit lim Mi /θhi .
In what follows, we will prove that lim Mi ∼ = lim Mi /θhi , and also that the
direct limit lim Mi /θhi is isomorphic to a direct union of a suitable directed
family of L-algebras.
First, we will check lim Mi ∼ = lim Mi /θhi . Note that every member gi of
the limit cone {gi : Mi /θhi → lim Mi /θhi | i ∈ I} associated with lim Mi /θhi is
an embedding. Indeed, due to Remark 2.101, every gi is a morphism satisfying
gi = gij ◦ gj . Moreover, for every k ≥ i, a , b ∈ Mi we have
[a ]θhi ≈Mi /θhi [b ]θhi = θhi (a , b ) = θ∞ (a , b ) =
   
= θ∞ hik (a ), hik (b ) = θhk hik (a ), hik (b ) .
Taking into account this observation, we readily obtain
  
[a ]θhi ≈Mi /θhi [b ]θhi = k≥i θhk hik (a ), hik (b ) =

= k≥i [hik (a )]θhk ≈Mk /θhk [hik (b )]θhk =
    
= k≥i gik [a ]θhi ≈Mk /θhk gik [b ]θhi =
   
= gi [a ]θhi ≈lim Mi /θhi gi [b ]θhi
for every gi : Mi /θhi → lim Mi /θhi and [a ]θhi , [b ]θhi ∈ Mi /θhi .
Let hi be the natural morphism hi : Mi → Mi /θhi (i ∈ I). Clearly, for the
composed morphism hi ◦ gi : Mi → lim Mi /θhi we have
      
gi (hi (a )) = gi [a ]θhi = gj gij [a ]θhi = gj [hij (a )]θhj = gj (hj (hij (a )))
i.e. hi ◦ gi = hij ◦ (hj ◦ gj ) for every i ≤ j. Since {hi : Mi → lim Mi } satisfies
DLP w.r.t. weak direct family {Mi | i ∈ I}, there is a uniquely determined
morphism g : lim Mi → lim Mi /θhi such that hi ◦ gi = hi ◦ g for every i ∈ I,
see Fig. 2.18 (left).
Moreover,
  to Theorem 2.35 a family {ki : Mi /θhi → lim Mi | i ∈ I},
due
where ki [a ]θhi = [a ]θ∞ = hi (a ) is a family of morphisms for which
      
ki [a ]θhi = hi (a ) = hj (hij (a )) = kj [hij (a )]θhj = kj gij [a ]θhi ,

Fig. 2.18. Morphisms between lim Mi and lim Mi /θhi


120 2 Algebras with Fuzzy Equalities

i.e. ki = gij ◦ kj holds for every i ≤ j. Since {gi : lim Mi → lim Mi /θhi | i ∈ I}
satisfies DLP w.r.t. weak direct family {Mi /θhi | i ∈ I}, there is a morphism
g  : lim Mi /θhi → lim Mi such that ki = gi ◦ g  for every i ∈ I, see Fig. 2.18
(right). Now, from
hi = hi ◦ ki = hi ◦ gi ◦ g  = hi ◦ g ◦ g  ,
hi ◦ gi = hi ◦ g = hi ◦ ki ◦ g = hi ◦ gi ◦ g  ◦ g ,
it readily follows that g ◦ g  = idlim Mi , g  ◦ g = idlim Mi /θhi . Thus, we can
apply Theorem 2.30 to get lim Mi ∼ = lim Mi /θhi .
Finally, it suffices to check the following claim:
   
lim Mi /θhi ∼ = gi Mi /θhi | i ∈ I , (2.36)
 
but this is almost evident. Clearly, gi Mi /θhi ∈ Sub(lim Mi /θhi ) by Theo- 
rem 2.28.
 Furthermore,
 for i
 ≤ j we have
 gi =
 g ij ◦ g j , i.e. g i M i /θ hi

Sub(gj Mj /θhj ). That is, gi Mi /θhi | i ∈ I is a directed family of L-

algebras. Hence, we consider h : lim Mi /θhi → gi Mi /θhi | i ∈ I to be
the identity automorphism on  lim M /θ
  
i hi
. Since every element  of lim Mi /θhi
can be expressed by some gi [a ]θhi ∈ gi Mi /θhi | i ∈ I , we have
   
gi [a ]θhi ≈lim Mi /θhi gj [b ]θhj =
     
= gk gik [a ]θhi ≈lim Mi /θhi gk gjk [b ]θhj =
     
= gk gik [a ]θhi ≈gk (Mk /θhk ) gk gjk [b ]θhj =
      
= gk gik [a ]θhi ≈ {gi (Mi /θhi ) | i∈I} gk gjk [b ]θhj =
    
= gi [a ]θhi ≈ {gi (Mi /θhi ) | i∈I} gj [b ]θhj
   
for every gi [a ]θhi , gj [b ]θhj , and k ≥ i, j. Analogously, for any n-ary f ∈ F
 
and arbitrary gim [am ]θhi , im ≤ k, m = 1, . . . , n it follows that
m
   
f lim Mi /θhi . . . , gim [am ]θhi , . . . =
    
m

= f lim Mi /θhi . . . , gk gim k [am ]θhi ,... =


    
m

= f gk (Mk /θhk ) . . . , gk gim k [am ]θhi ,... =


    
m

= f {gi (Mi /θhi ) | i∈I} . . . , gk gim k [am ]θhi ,... =
   
m

= f {gi (Mi /θhi ) | i∈I} . . . , gim [am ]θhi , . . . .
m
   
Altogether, lim Mi ∼ = lim Mi /θhi ∼ = gi Mi /θhi | i ∈ I . 


2.7 Reduced Products


In this section we propose a generalization of reduced products and show its
connection with previously introduced constructions. In addition to that, in
2.7 Reduced Products 121

Sect. 4.8 we show its non-trivial applications. Note that the generalization
presented below is probably not the only one possible. A survey on reduced
products and ultraproducts in fuzzy setting will be available in [19].
We define a reduced product of L-algebras by means of previously defined
constructions similarly Q
as in ordinary case. The key issue is how to define a
congruence relation on i∈I Mi with respect to a filter F over I (see Sect. 1.1).
Recall that in ordinary case we put
a , b ∈ θF iff {i ∈ I | a (i) = b (i)} ∈ F
Q
for every a , b ∈ i∈I M i . Thus, on the verbal level: “a , b ∈ θF iff the set of
indices on which a equals to b is large (i.e. belongs to a filter F ).” In what
follows, we will proceed in two steps. First, we try to generalize the notion
of “being equal on indices from X ∈ F ”. Then, using such a graded equality
with respect
Q to some index set, we define an L-relation representing for every
a , b ∈ i∈I M i a degree to which a equals to b over a large set of indices.
In the sequel, we use an ordinary filter. That is, we do not fuzzify the
notion of a filter itself. We denote a filter by F , and the elements of F will
be denoted X, Y, Z, . . . (there is no danger of confusion with the symbol of a
type of an L-algebra and with sets of variables, because we use a fixed type
and we do not use variables in the rest of this section).

Definition 2.119. Let {Mi | i ∈ I} be a family of L-algebras


Q of the same type
and let F be a filter over I. Then for every a , b ∈ i∈I M i and X ∈ F we
define a truth degree [[a ≈ b ]]X ∈ L as follows,

[[a ≈ b ]]X = i∈X a (i) ≈Mi b (i). (2.37)

Remark 2.120. (1) On the verbal level, [[a ≈ b ]]X expresses a truth degree to
which it is true that a is equal to b over all indices taken from X.
(2) An easy but important observation is that the truth degree [[a ≈ b ]]X
depends only on the truth degrees a (i) ≈Mi b (i) for i ∈ X.

Lemma 2.121. Let {Mi | i ∈ I} be a family of L-algebras of the same type


and let F be a filter over I. Then
(i) for every
Q
X, Y ∈ F , such that X ⊆ Y we have [[a ≈ b ]]Y ≤ [[a ≈ b ]]X ,
a ≈ i∈I Mi b ≤ [[a ≈
(ii)   b ]]X for every X ∈ F ,
(iii) X∈F [[a ≈ b ]]X = X1 ,...,Xn ∈F [[a ≈ b ]]X1 ∩···∩Xn .

Proof. (i) follows directly by Definition Q2.119.


(ii): Since X ⊆ I ∈ F , (i) yields a ≈ i∈I Mi b = [[a ≈ b ]]I ≤ [[a ≈ b ]]X .
(iii): The “≤”-part follows easily since for each X ∈ F , X = X ∩ · · · ∩ X.
Conversely, if X1 , . . . , Xn ∈ F then X1 ∩ · · · ∩ Xn ∈ F because F is closed
under finite intersections. Hence, the “≥”-part is also evident. 

Q
Now we use [[a ≈ b ]]X to define a suitable L-relation on i∈I M i .
122 2 Algebras with Fuzzy Equalities

Definition 2.122. Let {Mi | i ∈ I} be a family of L-algebras of the


Q same type
and let F be a filter over I. We define a binary L-relation θF on i∈I M i by

θF (a , b ) = X∈F [[a ≈ b ]]X (2.38)
Q
for all a , b ∈ i∈I M i .

Remark 2.123. Since X ∈ F are thought of as large subsets, θF (a , b ) can be


understood as a degree to which “there is a large X such that a equals b over
all indices from X”.

Theorem 2.124. Let {Mi | i ∈ I} be a familyQof L-algebras of the same type


and let F be a filter over I. Then θF ∈ ConL ( i∈I Mi ).
Q
Proof. From Lemma 2.121 it readily follows that ≈ i∈I Mi ⊆ θF . Moreover,
reflexivity and symmetry of θF follow directly from reflexivity and symmetry
of every ≈Mi , respectively. Q Thus, it suffices to check transitivity and compat-
Q
ibility with functions of i∈I Mi . Using Lemma 2.121, for a , b , c ∈ i∈I M i
we have
 
θF (a , b ) ⊗ θF (b , c ) = X∈F [[a ≈ b ]]X ⊗ Y ∈F [[b ≈ c ]]Y =
  
= X,Y ∈F [[a ≈ b ]]X ⊗ [[b ≈ c ]]Y =
   
= X,Y ∈F i∈X a (i) ≈Mi b (i) ⊗ j∈Y b (j) ≈Mj c (j) ≤
   
≤ X,Y ∈F i,j∈X∩Y a (i) ≈Mi b (i) ⊗ b (j) ≈Mj c (j) ≤
   
≤ X,Y ∈F i∈X∩Y a (i) ≈Mi b (i) ⊗ b (i) ≈Mi c (i) ≤
  
≤ X,Y ∈F i∈X∩Y a (i) ≈Mi c (i) = X,Y ∈F [[a ≈ c ]]X∩Y =

= X∈F [[a ≈ c ]]X = θF (a , c ) .
Q
Thus, θF is transitive. For compatibility withQfunctions take an n-ary f i∈I Mi

and arbitrary elements a1 , b1 , . . . , an , bn ∈ i∈I M i . We have


θF (a1 , b1 ) ⊗ · · · ⊗ θF (an , bn ) =
 n 
= X1 ,...,Xn ∈F i=1 ji ∈Xi a (ji ) ≈Mji b (ji ) ≤
  Q Q 
≤ X1 ,...,Xn ∈F f i∈I Mi (a1 , . . . , an ) ≈ f i∈I Mi (b1 , . . . , bn ) X1 ∩···∩Xn =
  Q Q 
= X∈F f i∈I Mi (a1 , . . . , an ) ≈ f i∈I Mi (b1 , . . . , bn ) X =
 Q Q 
= θF f i∈I Mi (a1 , . . . , an ), f i∈I Mi (b1 , . . . , bn ) .
Q
Altogether, θF is a congruence on i∈I Mi . 


Finally, we introduce reduced product of L-algebras.

Definition 2.125. Let {Mi | i ∈ I}be


Q a family
 of L-algebras of
Q the same type
and let F be a filter over I. Then i∈I Mi /θ F denoted by F Mi is called
a reduced product of {Mi | i ∈ I} modulo F .
2.7 Reduced Products 123
Q
Remark 2.126. Clearly, θF and the corresponding F Mi are determined by
{Mi | i ∈ I} and a filter F over I. In borderline cases, θF behaves the same way
as in ordinary case. Indeed, when
Q F is an improper
Q filter (i.e. ∅ ∈ F ), we have
θF (a , b ) = 1 for all a , b ∈ i∈I M i . Thus, F Mi is a trivial (one-element)
L-algebra. If F is a trivial filter (i.e. F is a proper filter and {i0Q } ∈ F for some
i0 ∈ I) it follows that θF = ≈Mi0 for Qcertain i0 ∈ I. Thus, F Mi ∼ = Mi0 .
Q Q
Finally, if F = {I} then clearly θF = ≈ i∈I Mi , i.e. F Mi ∼ = i∈I i M .

∗ ∗ ∗
In the ordinary case, reduced products are isomorphic to particular direct
limits and conversely, direct limits are isomorphic to particular subalgebras
of reduced products. In the subsequent development, we present analogous
characterizations.
Let {Mi | i ∈ I} be a family of L-algebras of the same type and
Q let F be a
filter over I. For every X ∈ F we can consider a direct product i∈X M Q i
. For
Q Q
brevity, let MX denote i∈X Mi . That is, MX = i∈X Q
M i , ≈MX
= ≈ i∈X Mi ,

and for an n-ary function symbol f let f MX denote f i∈X Mi . It readily follows
that

a ≈MX b = i∈X a (i) ≈Mi b (i) = [[a ≈ b ]]X . (2.39)
In addition to that, filter F can be partially ordered using the ordinary set
inclusion. Moreover, F, ⊇ can be thought of as a (downward) directed index
set. Indeed, for every X, Y ∈ F it follows that X, Y ⊇ X ∩ Y ∈ F , i.e. for
every X, Y ∈ F there is Z ∈ F such that X, Y ⊇ Z.
Clearly, MX , MY are disjoint for all X, Y ∈ F , X = Y . For every X, Y ∈ F ,
X ⊇ Y , we can consider a morphism hXY : MX → MY defined by
hXY (a )(i) = a (i) (2.40)
for every a ∈ MX and i ∈ Y . Due to Theorem 2.51, hXY is the uniquely deter-
mined morphism induced by family {πi : MX → Mi | i ∈ Y } of projections. It
is easily seen that {hXY : MX → MY | X ⊇ Y, X, Y ∈ F } satisfies conditions
(2.23) and (2.24). As a consequence, {MX | X ∈ F } is a weak direct family of
L-algebras.

Theorem 2.127. Let {Mi | i ∈ I}Qbe a family of L-algebras of the same type
and let F be a filter over I. Then F Mi ∼ = lim MX .
Q
Proof. We present a family {hX : MX → F Mi } of Q morphisms satisfying
DLP with respect to {MX | X ∈ F }. Then we get F Mi ∼ = lim MX as a
consequence of Theorem 2.104. Q Q
Recall that I ∈ F , and MI stands for i∈I Mi .QThus, hIX : i∈I Mi → MX
(X ∈ F ) are a surjective morphisms. For a , b ∈ i∈I M i we have

θhIX (a , b ) = hIX (a ) ≈MX hIX (b ) = i∈X hIX (a )(i) ≈Mi hIX (b )(i) =

= i∈X a (i) ≈Mi b (i) = [[a ≈ b ]]X ≤ θF (a , b ) .
124 2 Algebras with Fuzzy Equalities

hθ F hIX
i∈I Mi F Mi i∈I Mi MX
hθF gI
hIX gX
hX
MX F Mi g N
 Q 
Fig. 2.19. hX : MX → F Mi satisfies DLP w.r.t. {MX | X ∈ F }

Q Q
Therefore, θhIX ⊆ θF for every X ∈ F . Let hθF : i∈I Mi → F Mi denote
the natural morphism. Hence, from Lemma 2.44 it followsQ that for every
X ∈ F there is a uniquely determined morphism hX : MX → F Mi satisfying
hθF = hIX ◦ hX , see Fig. 2.19 (left). Moreover, we can apply (2.24) to obtain
hθF = hIY ◦ hY = hIX ◦ hXY ◦ hY , i.e. hIX ◦ hXY ◦ hY = hIX ◦ hX . Thus, the
surjectivity of hIX yields hX = hXY ◦ hY .
Let us have a family {gX : MX → N | X ∈ F } of morphisms satisfying
gX = hXY ◦ gY for every X, Y ∈ F , X Q ⊇ Y . It remains to show that there is
a uniquely determined morphism Q g : F Mi → N such that gX = hX ◦ g for
every X ∈ F . First, for a , b ∈ i∈I M i it follows that
  
θF (a , b ) = X∈F [[a ≈ b ]]X = X∈F i∈X a (i) ≈Mi b (i) =
  
= X∈F i∈X hIX (a )(i) ≈Mi hIX (b )(i) = X∈F hIX (a ) ≈MX hIX (b ) ≤
 
≤ X∈F gX (hIX (a )) ≈N gX (hIX (b )) = X∈F gI (a ) ≈N gI (b ) =
= gI (a ) ≈N gI (b ) = θgI (a , b ) .

Q θF ⊆ θgI . Due to Lemma 2.44 there is a uniquely determined morphism


Hence,
g : F Mi → N, where gI = hθF ◦ g, see Fig. 2.19 (right). Finally, gI =
hθF ◦g = hIX ◦hX ◦g and gI = hIX ◦gX by the assumption. AsQa consequence,
gX = hX ◦g due to the surjectivity of hIX . Thus, {hQ
X : MX → F Mi } satisfies
DLP with respect to {MX | X ∈ F }. Altogether, F Mi ∼ = lim MX . 


Theorem 2.128. Let {Mi | i ∈ I} be a weak direct family. ThenQ there is a


filter F over I such that lim Mi is isomorphic to a subalgebra of F Mi .

Proof. Introduce F by putting


X∈F iff there is i ∈ I such that {k ∈ I | k ≥ i} ⊆ X ⊆ I. (2.41)
Obviously, if X ∈ F and X ⊆ Y ⊆ I then Y ∈ F . If X, Y ∈ F then there
are indices i, j ∈ I such that {k ∈ I | k ≥ i} ⊆ X and {k ∈ I | k ≥ j} ⊆ Y , i.e.
{k ∈ I | k ≥ i, j} ⊆ X ∩ Y . Since I, ≤ is a directed index set there is l ≥ i, j.
Thus, {k ∈ I | k ≥ l} ⊆ X ∩ Y , showing X ∩ Y ∈ F . So, F defined by (2.41)
is a filter over I.
Consider M and θ defined by (2.33) and (2.34), respectively. We claim
that θ equals θF |M (recall that θF |M denotes the restriction of θF on M ).
It suffices to show that for any a , b ∈ M we have
2.7 Reduced Products 125
   
i∈I k≥i a (k) ≈Mk b (k) = X∈F k∈X a (k) ≈Mk b (k) .
“≤”: For each i ∈ I we have {k ∈ I | k ≥ i} ∈ F , i.e. the “≤”-part is obvious.
 X ∈ F . By definition
“≥”: Take  of F , there is i ∈ I with {k ∈ I | k ≥ i} ⊆ X.
Thus, k≥i a (k) ≈Mk b (k) ≥ k∈X a (k) ≈Mk b (k), showing “≥”.
Using the claim together with Theorem 2.115 and Theorem 2.42, we get
lim Mi ∼ = M /θ ∼ = M /(θF |M ) ∼
= (M )θF/(θF |(M )θF ) .
Q
Since (M )θF/(θF |(M )θF ) is Q
a subalgebra of F Mi we get that lim Mi is
isomorphic to a subalgebra of F Mi . 


As we have seen, for {Mi | i ∈ I} and a filter F over I, {MX | X ∈ F } is


a weak direct family of L-algebras. Suppose {MX | X ∈ F } is a direct family
of L-algebras and let us look whether the essential property (2.25) has some
natural translation in terms of the properties of F .

Definition 2.129. Let {Mi | i ∈ I} be a family of L-algebras of the same type.


Q filter F over I is safe with respect to {Mi | i ∈ I} if for every a , b ∈
A
i∈I M i there is X ∈ F such that θF (a , b ) = [[a ≈ b ]]X . If F is safe with
respect to any family of L-algebras of the same type then FQis called safe.
If F is safe with respect to a family {Mi | i ∈ I} then F Mi is called a
safe reduced product of {Mi | i ∈ I} modulo F .

Remark 2.130. Safeness of a filter F with respect to a given family of L-


algebras is a non-trivial property.
(1) If F = {I} then F is safe. Also every trivial and improper filterQis safe.
all a , b ∈ i∈I M i
(2) If θF (a , b ) is compact element of L (see Sect. 1.1) for Q
then F is safe w.r.t. {Mi | i ∈ I}. Indeed, n for any a , b ∈ i∈I M i there are
X1 , . . . , Xn ∈ F such that θF (a , b ) = i=1 [[a ≈ b ]]Xi . Since X1 ∩ · · · ∩ Xn ∈
F , it follows that θF (a , b ) ≤ [[a ≈ b ]]X1 ∩···∩Xn . The converse inequality holds
trivially. In particular, if L is a Noetherian residuated lattice then every filter
is safe.
(3) Take L = [0, 1] as the structure of truth degrees. Let us have an index
set N. We can consider a family {Mi | i ∈ N} of L-algebras of empty type,
where Mi = {a , b } and a ≈Mi b = 1 − 1i for every i ∈ N. Thus, a ≈M1 b = 0,
a ≈M2 bQ= 12 , a ≈M3 b = 23 , etc. Let F be the Fréchet filter over N. Take
a  , b  ∈ i∈N Mi , where a  (i) = a and b  (i) = b for all i ∈ N. Clearly, we
have θF (a  , b  ) = 1 but [[a  ≈ b  ]]X < 1 for every X ∈ F . Hence, F is not safe
w.r.t. {Mi | i ∈ N}.

Lemma 2.131. Let {Mi | i ∈ I} be a family of L-algebras of the same type


and let F be a filter over I. Then for a ∈ MX , b ∈ MY , and any Z ∈ F such
that X, Y ⊇ Z we have
hXZ (a ) ≈MZ hYZ (b ) = [[a  ≈ b  ]]Z , (2.42)
Q
where a  , b  ∈ i∈I M i satisfy hIX (a  ) = a and hIY (b  ) = b .
126 2 Algebras with Fuzzy Equalities

Proof. Clearly, we have



hXZ (a ) ≈MZ hYZ (b ) = i∈Z hXZ (a )(i) ≈Mi hYZ (b )(i) =

= i∈Z hXZ (hIX (a  ))(i) ≈Mi hYZ (hIY (b  ))(i) =
 
= i∈Z hIZ (a  )(i) ≈Mi hIZ (b  )(i) = i∈Z a  (i) ≈Mi b  (i) = [[a  ≈ b  ]]Z ,
which is the desired equality. 

Theorem 2.132. A filter F is safe w.r.t. a family {Mi | i ∈ I} of L-algebras
of the same type if and only if {MX | X ∈ F } is a direct family of L-algebras.
Proof. “⇒”: Let F be safe w.r.t. {Mi | i ∈ I}. Take a ∈ MX , b ∈ MY . We
have to show that there is Z ∈ F such that X, Y ⊇ Z and
hXZ (a ) ≈MZ hYZ (b ) = hXZ  (a ) ≈MZ  hYZ  (b )
Q
is satisfied for every Z  ∈ F , where Z ⊇ Z  . Let us have a  , b  ∈ i∈I M i such
that hIX (a  ) = a and hIY (b  ) = b . Since F is safe, we have θF (a  , b  ) =
[[a  ≈ b  ]]Z0 for certain Z0 ∈ F . Put Z = Z0 ∩ X ∩ Y . Clearly, for every Z  ∈ F
such that Z ⊇ Z  , it follows that
θF (a  , b  ) = [[a  ≈ b  ]]Z0 = [[a  ≈ b  ]]Z ≥ [[a  ≈ b  ]]Z  .
Moreover, Lemma 2.121 (i) yields [[a  ≈ b  ]]Z ≤ [[a  ≈ b  ]]Z  . Altogether, using
Lemma 2.131, we obtain
hXZ (a ) ≈MZ hYZ (b ) = [[a  ≈ b  ]]Z = [[a  ≈ b  ]]Z  = hXZ  (a ) ≈MZ  hYZ  (b ) .
Hence, {MX | X ∈ F } is a direct family of L-algebras.
“⇐”: Let {MX | X Q ∈ F } be a direct family of L-algebras. From (2.25) it
follows that for a , b ∈ i∈I M i there is some Z ∈ F such that
[[a ≈ b ]]Z = hIZ (a ) ≈MZ hIZ (b ) = hIZ  (a ) ≈MZ  hIZ  (b ) = [[a ≈ b ]]Z 
holds for every Z  ∈ F , Z ⊇ Z  . Thus, we have

θF (a , b ) = X∈F [[a ≈ b ]]X ≤
 
≤ X∈F [[a ≈ b ]]X∩Z = X∈F [[a ≈ b ]]Z = [[a ≈ b ]]Z .
The converse inequality follows from the definition of θF . That is, filter F is
safe with respect to a family {Mi | i ∈ I} of L-algebras. 

Remark 2.133. The construction of a safe reduced product is compatible with
its ordinary counterpart in the sense of preserving the skeletons:
Q Q
ske( Mi ) ∼
F = ske(lim MX ) ∼= lim{ske(MX ) | X ∈ F } ∼= ske(Mi ) . F

This is an immediate consequence of Theorem 2.111 and Theorem 2.127 and


Theorem
Q 2.132.QOn the contrary, we can use Remark 2.130 (3) to observe that
ske( F Mi ) ∼= F ske(Mi ) does not hold for general reduced products.
Reduced products will be used in Sect. 4.8 where we establish a connection
between safe reduced products and closure properties of certain implication-
ally definable classes of L-algebras.
2.8 Class Operators 127

2.8 Class Operators


In this section we present basic notions related to classes of L-algebras.
Definition 2.134. For a class K of L-algebras of the same type we define the
following operators:
H(K) = {h(M) | M ∈ K, h is a morphism} ,
I(K) = {M | M is isomorphic to some N ∈ K} ,
S(K) = {M | M is a subalgebra of some N ∈ K} ,
P(K) = {M | M is a direct product of a family P ⊆ K} ,
PS (K) = {M | M is a subdirect product of a family P ⊆ K} ,
U(K) = {M | M is a direct union of a directed family U ⊆ K} ,
L(K) = {M | M is a direct limit of a direct family L ⊆ K} ,
PR (K) = {M | M is a safe reduced product of a family P ⊆ K} .
That is, H(K) is the class of all homomorphic images of L-algebras from K.
I(K) is the class of all L-algebras isomorphic to some N ∈ K. S(K) is the class
of all subalgebras of L-algebras from K. P(K) and PS (K) denote the classes of
all direct and subdirect products of families of L-algebras from K, respectively.
U(K) denotes the class of all direct unions of directed families of L-algebras
from K. L(K) denotes the class of all direct limits of direct families of L-
algebras from K. Finally, PR (K) denotes the class of all safe reduced products
of families of L-algebras from K.
Remark 2.135. (1) Class operators may be composed. That is, we may have
HS, HHPHS, and so on.
(2) If O1 , O2 are two operators on classes of L-algebras of the same type
we put O1 ≤ O2 iff O1 (K) ⊆ O2 (K) for every class K of L-algebras of type F .
It is easy to see that ≤ is a partial order on operators on classes of L-algebras
of the same type.
Definition 2.136. A class K of L-algebras of type F is said to be closed
under an operator O if O(K) ⊆ K. An operator O is said to be idempotent
iff O = OO.
Definition 2.137. A class K of L-algebras closed under the operator I is
called an abstract class of L-algebras.
We adopt the following convention for class operators on abstract classes
of L-algebras. Let O be an operator on classes of algebras. If K is an ab-
stract class of L-algebras, then O(K) denotes an abstract class IO(K). Roughly
speaking, we add to O(K) all isomorphic copies of L-algebras from O(K). The
notion of an abstract class will be used mainly in Sect. 4.5 to avoid technical
complications. Let us note that for abstract classes we have several conve-
nient consequences, e.g. P is an idempotent operator on abstract classes of
L-algebras while this is not so for general classes of L-algebras.
128 2 Algebras with Fuzzy Equalities

Definition 2.138. Let K be a class of L-algebras of the same type. Let M


be an L-algebra generated by M  ⊆ M , i.e. M = [M  ]M . If for each N ∈ K
   
and every ≈-morphism h : M  , ≈M → N, ≈N there exists a homomorphic
extension h : M → N of h (that is h is an morphism and h (a ) = h(a ) for
every a ∈ M  ), we say that M has a universal mapping property (UMP)
for K over M  . In this case M  is said to be a set of free generators of M
over K.
Theorem 2.139. Suppose that M has UMP for KNover
 M  . Let N ∈ K.
 M
Then for every ≈-morphism h : M , ≈ → N, ≈ there exists a unique
morphism h : M → N extending h.
Proof. Suppose there are two homomorphic extensions, h , h : M → N. Since
M is generated by M  , Theorem 2.81 yields that for every a ∈ M there
is a term t ∈ T (X), |X| = |M  | such that tM (a1 , . . . , ak ) = a for some
a1 , . . . , ak ∈ M  . Thus,
   
h (a ) = h tM (a1 , . . . , ak ) = tN h (a1 ), . . . , h (ak ) =
   
= tN h (a1 ), . . . , h (ak ) = h tM (a1 , . . . , ak ) = h (a ) .
Hence, h admits a unique homomorphic extension h . 

 
Theorem
 2.140. Suppose that M has UMP for K over M and let N, N ∈ K.

If g : M  , ≈M → N, ≈N is an ≈-morphism and h : N → N is a morphism
then (g ◦ h) = g  ◦ h.
Proof. For each a ∈ M  we have (g  ◦h)(a ) = h(g  (a )) = h(g(a )) = (g ◦h)(a )
because g  is an extension of g. Thus, g  ◦ h is a morphism which extends g ◦ h.
Since g ◦ h is an ≈-morphism, Theorem 2.139 gives that (g ◦ h) = g  ◦ h.  
Theorem 2.141. For a type F , a set X of variables, and a class K of L-
algebras of type F , if T(X) exists, then it has a universal mapping property
for K over X.
Proof. Consequence of Theorem 2.79. 


2.9 Related Approaches


Algebras with fuzzy equalities represent but one particular approach to the
notion of an algebra from the point of view of fuzzy logic. This approach is
a modest one – at the beginning, the only basic concept of ordinary algebras
which becomes fuzzy is the equality on the universe. Doing so, our approach
to algebras fits the framework of predicate fuzzy logic. Namely, algebras with
fuzzy equalities are just the structures of predicate fuzzy logic with equality
symbol as the only relation symbol.
Needless to say, there can be many ways to approach algebraic concepts
from the point of view of fuzzy logic. The aim of this section is to present some
of them which appeared in the literature. If possible, we discuss relationships
to algebras with fuzzy equalities.
2.9 Related Approaches 129

Metric Algebras
Recall from Sect. 1.3 that a generalized metric on a set M is a mapping
 : M × M → [0, ∞] satisfying conditions listed on page 42. For the sake of
brevity, we call a generalized metric simply a metric throughout this section.
In [96] (see also [95]), Weaver studies so-called metric algebras. A metric
algebra is basically an algebra equipped with a metric on its support such that
operations of the algebra are in a particular sense compatible with the metric.
The metric can be seen as a constraint for the operations. Metric algebras thus
represent an approach to the idea of extending algebras by adding constraints.
Since the very idea of a metric is to describe closeness of elements of the
universe, metric algebras can be seen as providing an alternative to algebras
with fuzzy equalities. An obvious question is that of the relationship between
metric algebras and algebras with fuzzy equalities. The goal of this section is
to look at it.
We start with basic notions
 related to metric algebras. A metric algebra
of type F is a triplet M = M, M , F M , where
 
(i) M, F M is a classical algebra of type F ,
(ii) M is a metric on M .
Recall from the beginning of this section that we allow M (a , b ) = ∞. The
notion of a metric algebra itself does not include any constraint on functions.
In order to introduce a constraint, Weaver uses implications between so-called
atomic inequalities and  a notion of an equicontinuous satisfaction. Given a
metric algebra M = M, M , F M and a mapping v : X → M , we utilize the
usual notion of a value !t!M,v of a term t ∈ T (X) in M under a valuation
v: for variable x ∈ X, put !x!M,v = v(x); for a term f (t1 , . . . , tn ) ∈ T (X),
put !f (t1 , . . . , tn )!M,v = f M (!t1 !M,v , . . . , !tn !M,v ), cf. Remark 1.96 (2). An
atomic inequality is an expression of the form (t, t )  α, where t, t ∈ T (X) 
and α ∈ [0, ∞]. (t, t )  α is δ-true in M under v if M !t!M,v , !t !M,v ≤
α + δ. (t, t )  α is true in M under v if it is δ-true in M under v for δ = 0.
An implication (between atomic inequalities) is an expression of the form
(s1 , s1 )  α1 c · · · c (sn , sn )  αn i (t, t )  β . (2.43)
Let K be a class of metric algebras. K satisfies (2.43) if for each M ∈ K
and each valuation v we have: if (si , si )  αi is true in M under v for
each i = 1, . . . , n, then (t, t )  β is true in M under v. K satisfies (2.43)
equicontinuously, if for each ε > 0 there is δ > 0 such that for each M ∈ K
and each valuation v we have: if (si , si )  αi is δ-true in M under v for each
i = 1, . . . , n, then (t, t )  β is ε-true in M under v.
K has equicontinuous functions if for any n-ary f ∈ F , K satisfies
(x1 , y1 )  0 c · · · c (xn , yn )  0 i
i (f (x1 , . . . , xn ), f (y1 , . . . , yn ))  0 (2.44)
equicontinuously. M has equicontinuous functions if K = {M} has equicon-
tinuous functions.
130 2 Algebras with Fuzzy Equalities

Remark 2.142. (1) If K satisfies (2.43) equicontinuously then K satisfies (2.43).


In more detail, if each (si , si )  αi is true (i.e., 0-true) in M under v then it
is also δ-true. Therefore, if (si , si )  αi (i = 1, . . . , n) are true in M under
v then (t, t )  β is ε-true for each ε > 0 which immediately gives that
(t, t )  β is true (i.e., 0-true).
(2) Condition (2.44) can be seen as the classical compatibility axiom, see
Remark 1.81 (2), which is formulated in terms of metric algebras, i.e., as an
implication between atomic inequalities instead of the classical identities.
(3) In certain cases, equicontinuity of functions of a metric algebra M is
trivial. For example, if
{M (a , b ) | a , b ∈ M and a = b }
has a lower bound γ > 0 then M has equicontinuous functions since for each
ε > 0 one can put δ = γ2 . M then satisfies (2.44) equicontinuously, because
(xi , yi )  0 is δ-true ( γ2 -true) in M under v iff v(xi ) = v(yi ). Note that
such a situation occurs if M is finite, or more generally, if {M (a , b ) | a , b ∈
M and a = b } is finite.
(4) M has equicontinuous functions iff functions of M are uniformly con-
tinuous.
(5) Let us note that the existence of δ > 0 for each ε > 0 in the definition
of equicontinuous satisfaction is required for a whole class K. That is, the fact
that each M ∈ K has equicontinuous functions does not imply in general that
K has equicontinuous functions.
We are now going to look at the relationship between algebras with fuzzy
equalities and metric algebras with equicontinuous functions. As mentioned
above, algebras with fuzzy equalities and metric algebras can be seen as two
different approaches to functions on a set which are compatible with close-
ness/similarity of elements of this set. In addition to this, as described in
Theorem 1.78, there is a close relationship between metrics and fuzzy equal-
ities. We are going to show that L-algebras with L given by a continuous
Archimedean t-norm can be turned into metric algebras with equicontinuous
functions.
Lemma 2.143. Let L be a residuated lattice given by a continuous Archime-
dean t-norm ⊗, ≈M be an L-equivalence on M , f M : M n → M be a function
compatible with ≈M . Then
 
≈M f M (a1 , . . . , an ), f M (b1 , . . . , bn ) ≤ ≈M (a1 , b1 ) + · · · + ≈M (an , bn )
for each a1 , b1 , . . . , an , bn ∈ M .
Proof. Let g be the additive generator of ⊗. First, we claim that for each
a1 , . . . , an ∈ L we have
g(a1 ⊗ · · · ⊗ an ) ≤ g(a1 ) + · · · + g(an ) . (2.45)
For n = 1, the claim is trivial. By induction, suppose
n g(a2 ⊗
 · · · ⊗ an ) ≤
g(a2 ) + · · · + g(an ). By Theorem 1.43, we have g( i=1 ai ) = g g (−1) (g(a1 ) +
2.9 Related Approaches 131
n  n n
g(i=2 ai )) . Now,
 if g(a1 )+g( i=2 ai ) ≤ g(0), we have g( i=1 ai ) = g(a1 )+
n n n
g( i=2 ai ) ≤  i=1 g(a i ) by induction assumption.
n If g(a )
1n + g( i=2 ai ) 
n
g(0), we have g( i=1 ai ) = g(0) < g(a1 ) + g( i=2 ai ) ≤ i=1 g(ai ), showing
that (2.45) holds. Thus, applying this claim and the fact that g is decreasing,
we get
 
≈M f M (a1 , . . . , an ), f M (b1 , . . . , bn ) =
 
= g f M (a1 , . . . , an ) ≈M f M (b1 , . . . , bn ) ≤
≤ g(a1 ≈M b1 ⊗ · · · ⊗ an ≈M bn ) ≤ g(a1 ≈M b1 ) + · · · + g(an ≈M bn ) =
= ≈M (a1 , b1 ) + · · · + ≈M (an , bn ) ,
which is the desired inequality. 


Theorem 2.144. Let  L be a residuated


 lattice given by a continuous Archi-

medean t-norm ⊗, M, ≈M , F M be an L-algebra. Then M, ≈M , F M is a
metric algebra with equicontinuous functions.

Proof. Theorem 1.78 yields that ≈M is a metric since ≈ 


M
is an L-equality.

Thus, it remains to check that for each n-ary f ∈ F , M, ≈M , F M satis-
fies (2.44) equicontinuously. If f is a nullary function symbol, the claim is
trivial. So suppose n > 0. Given ε > 0, put δ = nε . Suppose ≈M (ai , bi ) ≤ δ
for each i = 1, . . . , n. Using Lemma 2.143 we have
 
≈M f M (a1 , . . . , an ), f M (b1 , . . . , bn ) ≤
≤ ≈M (a1 , b1 ) + · · · + ≈M (an , bn ) ≤ nδ = ε .
 
Hence, M, ≈M , F M satisfies (2.44) equicontinuously. 


Remark 2.142 (4) yields the following corollary.

Corollary 2.145. Let


 L be a residuated
 lattice given by a continuous Archi-
medean t-norm ⊗, M, ≈M , F M be an L-algebra. Then each f M ∈ F M is a
uniformly continuous function with respect to ≈M . 


Note that the nontrivial point of the above assertions is that if the func-
tions f M of the L-algebra are compatible with ≈M , then f M are uniformly
continuous w.r.t. ≈M .
The converse transformation is not possible in general. Namely, one cannot
use the pseudo-inverse g (−1) of g to transform a metric algebra with equicon-
tinuous functions into an L-algebra. An example follows.

Example 2.146. For a complete residuated lattice L given by an arbitrary


continuous Archimedean
 t-norm, one can find a finite metric algebra M =
M, M , F M with a single unary function f M ∈ F M such that f M is not
compatible with ≈M . For instance, put M = {a , b , c , d }, and let f M and
≈M be defined by the following tables.
132 2 Algebras with Fuzzy Equalities

fM ≈M a b c d
a c a 1 x 0 0
b d b x 1 0 0
c c c 0 0 1 y
d d d 0 0 y 1

Suppose 1 > x > y. It is easily seen that ≈M is an L-equality.


 Furthermore,

≈M equals ≈≈M by Theorem 1.78. Since M is finite, M, ≈M , F M is a
metric algebra with equicontinuous functions, see Remark 2.142 (3). On the
other hand, f M is not compatible with ≈≈M , because ≈≈M coincides with
 
≈M and a ≈M b = x  y = f (a ) ≈M f (b ). Therefore, M, ≈M , F M cannot
be turned into an L-algebra by using (1.82).

Therefore, although the idea behind equicontinuity of operations in metric


algebras is similar to that of compatibility with a given L-equality, metric
algebras deal with a different type of restrictions on operations. On the one
hand, L-algebras can be thought of as more general than metric algebras
because L need not be linearly ordered. On the other hand, if we restrict
ourselves only to L’s given by continuous Archimedean t-norms, compatibility
with a given L-equality is more restrictive than the equicontinuity of functions
as we have seen in Theorem 2.144 and Example 2.146.
For further information on metric algebras we refer to [95, 96] and [91].

Fuzzy Congruences of Murali and Samhan

In this section, we take a brief look at the approach developed in [68] and [77]
and its relationship to algebras with fuzzy equalities. The authors develop the
conceptof a congruence
 in fuzzy setting. Their approach goes as follows.
Let M, F M be an ordinary algebra. Consider the real unit interval [0, 1]
as a set of truth degrees. A binary [0, 1]-relation (i.e. a fuzzy relation
 with

truth degrees from [0, 1]) θ in M is called a fuzzy congruence on M, F M if
θ is a fuzzy equivalence with min as conjunction operation (i.e., transitivity
says min(θ(a , b ), θ(b , c )) ≤ θ(a , c )) which satisfies
min(θ(a1 , b1 ), . . . , θ(an , bn )) ≤ θ(f M (a1 , . . . , an ), f M (b1 , . . . , bn )) (2.46)
for each n-ary f M ∈ F M and every a1 , b1 , . . . , an , bn ∈ M .
 In M [77],
 the author presents the following
 M
definition
 of a factor algebra of
M, F by a fuzzy
 congruence
 θ on M, F . A fuzzy
 factor  algebra of an
ordinaryalgebra M, F M by a fuzzy congruence θ on M, F M is an ordinary
algebra Mθ , FθM such that
(i) Mθ = {a θ | a ∈ M }, where a θ is a fuzzy set in M defined by
(a θ)(b ) = θ(a , b ) for all a , b ∈ M .
 
(ii) f Mθ (a1 θ, . . . , an θ) = f M (a1 , . . . , an ) θ
for any n-ary f Mθ ∈ F Mθ and a1 θ, . . . , an θ ∈ Mθ .
2.9 Related Approaches 133

The following are basic remarks from the point of view of algebras with
fuzzy equalities.
 
Remark 2.147. (1) On the one hand, M, F M is an ordinary algebra, i.e.
it is equipped
 with crisp equality relation. On the other hand, congruences
on M, F M are considered as fuzzy relations. Proceeding this way, one
does not have some (usual) relationships between congruences and mor-
phisms. For instance, in the ordinary case we have that all congruences of
an algebra can be obtained as kernels of morphisms, i.e. are of the form
θh = {a , b | h(a ) = h(b )} where h is a morphism of the algebra into
some other algebra. For algebras with fuzzy equalities, this fact remains valid.
Namely, for a congruence θ on an L-algebra M we have θ = θhθ (see Defin-
ition 1.83, Definition 2.32, and Theorem 2.33). For the approach of [68, 77]
this is no longer true since each kernel θh is a crisp relation.
(2) From the point of view of the choice of a structure of truth degrees,
[68, 77] deal with a set L = [0, 1] of truth degrees equipped with ⊗ = min.
That is, they deal with a particular complete residuated lattice
 L given by a
t-norm ⊗ = min on [0, 1]. Recall that
 an ordinary algebra M, F M
coincides
with the L-algebra M, ≈M , F M with ≈M being a crisp equality. Further-
more, notice that with ⊗ = min, condition (2.46) expressing compatibility
of a fuzzy congruence coincides with our condition of compatibility, see De-
finition 1.80 and Remark 1.81 (1). From this point of view, the concept of a
fuzzy congruence on an ordinary algebra as considered in [68, 77] is a spe-
cial case of the concept of a congruence on an algebra with fuzzy equality.
Note also that in addition to the fact that [0, 1] equipped with min is only a
particular example of a complete residuated lattice, one may lose insight into
some important algebraic concepts. For instance, recall from Corollary 2.70
and Example 2.72 that the subdirect representation theorem for L-algebras is
true only for certain classes of complete residuated lattices. However, it is not
true for L = [0, 1]. As another example, the left side of compatibility condition
(2.46) depends only on the least θ(ai , bi ). This is a very particular property
which is not available (and may even be considered strange) in the general
case of ⊗, see also Remark 2.13.
(3) The concept of an algebra is one of the fundamental concepts in math-
ematics. When generalizing to fuzzy setting, one should be as general as pos-
sible. Namely, any theory of algebras developed in fuzzy setting can be seen
as useful only when one finds a rich class of examples for which the general
algebraic concepts and results will bring new insights and results. We showed
several such examples existing in the literature in Sect. 2.1. Considering only
[0, 1] with min is restrictive from this point of view since it rules out both
existing examples as well as examples which can appear in future work in
fuzzy logic.
Consider now the concept of a fuzzy factor algebra. The following two
remarks summarize basic observations and show an advantage of factorization
as presented in the framework of L-algebras.
134 2 Algebras with Fuzzy Equalities

Remark 2.148. Except for the fact that [77] consider ordinary algebras (with-
out fuzzy equalities), the main distinction between the concept of a fuzzy
factor algebra and that of a factor L-algebra (Definition 2.22) is the uni-
verses. While the elements of the universe Mθ of a fuzzy factor algebra are
fuzzy sets a θ in M defined by (a θ)(b ) = θ(a , b ), the elements of the universe
M/θ of a factor L-algebra M/θ are crisp sets [a ]θ in M (i.e. ordinary subsets
of M ) defined by [a ]θ = {b | a , b ∈ 1 θ}. Recall from Sect. 1.3 (page 46
and onwards) that of the two possibilities of what the elements of a factor
set of a set with fuzzy equality should be, whether a θ or [a ]θ , we decided
to take [a ]θ since it is more simple than a θ (namely, [a ]θ = 1 a θ) and since
both ways lead to isomorphic sets with fuzzy equality (Lemma 1.90). Now,
if we would
 take a θ and would define a factor structure Mθ of an L-algebra
M = M, ≈M , F M by a congruence θ on M in the spirit of [77], Mθ would
be isomorphic to M/θ. Therefore, the approach of [77], modified to the setting
of algebras with fuzzy equalities, yields a factor structure which is isomorphic
to but more complex than the factor L-algebra of Definition 2.22. In more
detail, for a congruence θ on an L-algebra M,let a fuzzy factor L-algebra be
defined as an L-algebra Mθ = Mθ , ≈M M
θ , Fθ with Mθ and FθM defined as
in (i) and (ii) of the definition of a fuzzy factor algebra on page 132 and with
≈M θ defined by

(iii) (a θ ≈M
θ b θ) = θ(a , b ).

Then, as an immediate consequence of Lemma 1.90 we get:


Theorem 2.149. Under the above notation, Mθ is isomorphic to M/θ with
a mapping h : Mθ → M/θ defined by h(a θ) = [a ]θ being an isomorphism.  
 
Remark 2.150. Consider  an M L-algebra
 M = M, ≈M , F M . Then ≈M is a
fuzzy congruence on M, F in the sense of [77] (without reservations if
L = [0, 1] and ⊗ = min, and with appropriate modification of the con-
cepts of
 [77] for general
 L). Therefore,
 we can consider a fuzzy factor al-
gebra  M≈M , F≈MM (facrotization of M, F M by ≈M in the sense of [77]).
Now, M≈M , F≈MM encompasses the same information as M, ≈M , F M since
the L-equality ≈M can be reconstructed from M≈M , and vice versa. Namely,
M≈M = {a ≈M | a ∈ M } and a ≈M b = (a ≈M )(b ). Now, for instance,
morphisms of L-algebras are required to preserve L-equalities, i.e. to be ≈-
morphisms (Definition 2.26), while morphisms of ordinary algebras (which are
considered
 in [77])are, of course, not. As a consequence,
 there are morphisms
from M≈M , F≈MM which are not morphisms from M, ≈M , F M since they
are not ≈-morphisms. This shows that due to “loss of similarity information”
the concept of an L-algebra with its ramifications cannot be appropriately
captured by framework provided by [68, 77].
We close our visit on fuzzy congruences by a remark on factor congruences.
Remark 2.151. In [77], instead of Definition 2.54 (ii), the author requires θ ◦
θ∗ = M ×M (denote this condition by (ii’)). As we know from Remark 2.55,
2.9 Related Approaches 135

(ii) implies (ii’). In the ordinary case, both (ii) and (ii’) are equivalent. This,

however,
 might not be ∗
 in general since we may have θ ◦ θ (a , b ) = 1,
the case
i.e. c ∈M θ(a , c )⊗θ (b , c ) = 1, although there is no c ∈ M with θ(a , c ) = 1
and θ∗ (b , c ) = 1, i.e. c ∈ [a ]θ and c ∈ [b ]θ∗ . Therefore, (ii) is stronger than
(ii’) in general. This fact reflects itself in Theorem 4.2 of [77] which is a
generalization of the well-known theorem on factor congruences. Contrary to
our approach, Theorem 4.2 of [77] needs an additional assumption.

Fuzzy Subalgebras

Let us briefly comment on so-called fuzzy (sub)algebras. Fuzzy subalgebras


started with Rosenfeld’s fuzzy groups [76]. Since then, many papers developing
this approach for general algebras but also for many special algebras appeared,
see e.g. [20, 29, 34, 66, 76, 81, 101] (this selection is probably not represen-
tative). Recall that a fuzzy subalgebra in an ordinary algebra M = M, F M
is an L-set A in M such that for each n-ary operation f M ∈ F M we have
A(a1 ) ⊗ · · · ⊗ A(an ) ≤ A(f M (a1 , . . . , an ))
for every a1 , . . . , an ∈ M . This condition is satisfied iff a first-order for-
mula saying that “if a1 belongs to A and · · · and an belongs to A then
f M (a1 , . . . , an ) belongs to A”. Therefore, from the point of view of predicate
fuzzy logic, this condition generalizes a condition required for an ordinary
subset A of M to be a subalgebra (to be precise: a subuniverse). Note that
that there is no fuzzy equality on M involved in the concept of a fuzzy sub-
algebra. On the other hand, a subalgebra of an algebra with fuzzy equality is
again an algebra with fuzzy equality (i.e. with a crisp set as a universe), see
Definition 2.10. In a sense, the distinction between the theory of fuzzy subal-
gebras and the theory of algebras with fuzzy equalities lies in what becomes
fuzzy: in fuzzy subalgebras, it is the universe set, while in algebras with fuzzy
equalities, it is the equality on the universe. While most parts of universal
algebra can be developed in the setting of algebras with fuzzy equalities, we
are not aware of a development of these parts in the setting of fuzzy subalge-
bras. Nevertheless, fuzzy subalgebras and algebras with fuzzy equalities are
complementary and it is thus an open problem to look at the combination of
both of these approaches, i.e. to consider fuzzy subalgebras A in an algebra
M with fuzzy equality ≈M , possibly such that A is compatible with ≈M , i.e.
A(a ) ⊗ (a ≈M b ) ≤ A(b ).
Recall also Head’s paper [53] which is one of few existing attempts which
can lead stop the boom of papers (very often without any motivations and
providing only easy-to-get results) on fuzzy subalgebras.

Fuzzy Functions

Another approach to universal algebras from the point of view of fuzzy sets
might concern the functions f M of an algebra. Namely, several approaches
136 2 Algebras with Fuzzy Equalities

have been developed to give a reasonable meaning to the concept “fuzzy func-
tion”, see e.g. [35]. In a series of papers, see e.g. [33], Demirci considers sets
equipped with fuzzy equivalence/equality relations and compatible fuzzy func-
tions. Fuzzy functions used by Demirci are fuzzy relations that need to satisfy
certain properties which make them behave like functions. Ordinary functions
can be seen as particular fuzzy functions. Our approach is thus a particular
approach of Demirci’s. In his papers, Demirci considers only binary fuzzy
functions. Moreover, he does not consider general structural algebraic notions
which we have developed for algebras with fuzzy equalities. A development of
universal algebraic results in the setting of Demirci is an open problem. Note,
however, that with fuzzy functions, things get technically more complicated.
For instance, it is not immediately obvious how terms and terms functions
should look like in the setting of fuzzy functions. More generally, structures
with fuzzy functions should possibly be studied in a framework of a suitable
predicate fuzzy logic. This would provide a unifying framework and make it
possible to use general results of predicate fuzzy logic. Also, this would perhaps
suggest a way to develop logical calculi for reasoning about fuzzy identities in
Demirci’s setting analogous the calculi developed in Chap. 3 and Chap. 4.
Note also that in the early development of predicate fuzzy logic by
Novák, function symbols were interpreted in fuzzy structures by so-called
fuzzy functions of type 2, see [70]. Given a fuzzy set A in a set M , a func-
tion f : M n → M is said to be an n-ary fuzzy function of type 2 in A if
A(a1 ) ⊗ · · · ⊗ A(an ) ≤ A(f (a1 , . . . , an )) for every a1 , . . . , an ∈ M which is is
just the condition used in the definition of a fuzzy subalgebra. Later on, fuzzy
functions of type 2 were replaced by ordinary functions in the semantics of
predicate fuzzy logic.

2.10 Bibliographical Remarks


The concept of a general/universal algebra goes back to the end of nineteenth
century. However, the birth of contemporary universal algebra is being at-
tributed to Birkhoff and his [22, 23, 24]. A good book on universal algebra
is [27]. Universal algebra as a research field deals with the general concept of
an algebra, i.e. a set with operations. It is concerned with topics related to
structural properties of algebras. It disregards problems which result as con-
sequences of special nature of particular algebras. Results of general algebra
are useful if one comes up with a new kind of algebra. Then, many properties
are immediately available as instances of results of universal algebra.
Except for that, universal algebras are semantical structures of a fragment
of predicate logic with an equality symbol as the only relational symbol. This
fragment is the framework of so-called equational reasoning which is studied
in theoretical computer science, see e.g. [1, 62].
The concept of an algebra with fuzzy equality appeared in [9]. It results by
putting together the concept of a set with equality (Definition 1.79) and that
2.10 Bibliographical Remarks 137

of a universal algebra and by connecting them via compatibility condition


(Definition 1.80). Bibliographical comments on these notions were presented
in Sect. 1.5. The present chapter contains results most of which are contained
in [16] and [92]. Some of the results on algebras with fuzzy equalities can be
obtained from more general results on fuzzy relational structures presented
in [10]. However, since we do not want to introduce concepts on the level of
generality as they appear in [10] (e.g., a degree to which a fuzzy relation is a
congruence), we do not refer to [10].
Note that reduced products and their special case of ultraproducts play a
crucial role in universal algebra and model theory. A thorough study on this
topic can be found in [32]. There is also a chapter [27] on reduced products and
their applications in universal algebra. In [31], the authors present a general-
ization of an ultraproduct for structures equipped with relations whose truth
degrees form a compact Hausdorff space. In fuzzy setting, however, there is
only a small effort in studying reduced products and ultraproducts. An excep-
tion is Gottwald’s [47] where the author presents an approach to ultraproducts
in fuzzy setting which is based on [31]. Reference [98] deals with ultraproducts
in first-order Pavelka-style logic (but with limiting assumptions). A further
study of reduced products and ultraproducts in fuzzy setting is an important
open problem.
The present chapter deals with basic results of universal algebra in fuzzy
setting. Needless to say, it might be interesting to develop other parts of
universal algebra and, in particular, to look at phenomena which are hidden
in crisp setting. This seems to be an open problem.
3
Fuzzy Equational Logic

Equational logic deals with identities (equations) like x+y ≈ y+x, x·(y+z) ≈
x · y + x · z, dec(inc(x)) ≈ x, etc. Identities are simple formulas which can be
interpreted in algebras. Thus, given an identity and an algebra, either the
identity is true or false in the algebra. For instance, x ◦ y ≈ y ◦ x is true in
the algebra Z of all integers if ◦ is interpreted by addition of integers, but
it is false in the algebra of all square real matrices (say, of dimension 5 × 5)
if ◦ is interpreted by matrix multiplication. The two most important aspects
dealt with in equational logic are reasoning over identities and definability
(specification of requirements) using sets of identities.
Reasoning: Deriving new identities from known ones follows some intuitive
rules. A trivial example is “derive x ≈ x” (more precisely, “derive x ≈ x from
an arbitrary set of known identities”, i.e. no specific identities are needed to
be known), “derive x + z ≈ x + z”, or generally, “derive t ≈ t” where t is
an arbitrary term. Another trivial example is “derive x + y ≈ y + x from
y + x ≈ x + y”. Although a bit complex, a rule “derive x · (y + z) ≈ x · z + x · y
from x · (y + z) ≈ x · y + x · z and x + y ≈ y + x” is also being used without
reservations. In his seminal paper, Birkhoff [23] showed that there are five
simple rules which are sufficient for deriving new identities from known ones.
More precisely, an identity t ≈ s can be derived from a set Σ of identities using
these rules if an only if t ≈ s is true in every algebra satisfying each identity
from Σ. This result is the well-known Birkhoff’s completeness theorem of
equational logic.
Definability: Interesting classes of algebras are often “definable by iden-
tities.” This is to say, there is a set Σ of identities such that an (interest-
ing) class K of algebras contains just algebras which satisfy each identity
from Σ. For instance, the class of all commutative semigroups is definable by
Σ = {x ◦ y ≈ y ◦ x, x ◦ (y ◦ z) ≈ (x ◦ y) ◦ z}. Birkhoff [23] showed that a class K
of algebras is definable by identities if and only if K is closed under formation
of homomorphic images, subalgebras, and direct products of algebras from K.
This result is the well-known Birkhoff’s variety theorem.

R. Bělohlávek and V. Vychodil: Fuzzy Equational Logic, StudFuzz 186, 139–170 (2005)
www.springerlink.com 
c Springer-Verlag Berlin Heidelberg 2005
140 3 Fuzzy Equational Logic

The aim of this chapter is to develop equational logic in fuzzy setting. We


focus mainly on the two above-mentioned Birkhoff’s results, i.e. completeness
and variety theorem. A logical calculus which we develop will be called a fuzzy
equational logic.
The basic features of fuzzy equational logic are the following. Like in case
of ordinary equational logic, formulas of fuzzy equational logic are identities.
However, identities will be interpreted in algebras with fuzzy equalities. That
is, algebras with fuzzy equalities are semantical structures of fuzzy equational
logic. From this point of view, Chap. 2 can be seen as developing model-
theoretical constructions of fuzzy equational logic. Furthermore, we develop
fuzzy equational logic in Pavelka style, see Sect. 1.4. In particular, we deal
with fuzzy sets of identities, a degree of provability, and a degree of semantic
entailment.
∗ ∗ ∗
The rest of this chapter is organized as follows. Section 3.1 introduces basic
concepts of syntax and semantics. In Sect. 3.2 we prove (Pavelka-) complete-
ness of fuzzy equational logic for arbitrary complete residuated lattice L.
Section 3.3 develops varieties and free algebras with fuzzy equalities. In
Sect. 3.4, we prove Birkhoff’s variety theorem in our setting. Section 3.5 we
deal with two topics in varieties in fuzzy setting which are degenerate in the
ordinary setting.

3.1 Syntax and Semantics


We start with basic concepts of syntax and semantics of fuzzy equational
logic. We define the concepts of a language, term, formula, structure, value of
a term, and truth degree of a formula.
Suppose ≈, F, σ is a type (page 60) and L is a complete residuated lattice.
A language of fuzzy equational logic of type ≈, F, σ for L consists of a binary
relation symbol ≈ called a symbol of (fuzzy) equality, function symbols f ∈ F
with their arities σ(f ) ∈ N0 , symbols a of truth degrees (a ∈ L), (at least
denumerable) set X of variables with X ∩ F = ∅, and auxiliary symbols
(parentheses, etc.). If ≈, F, σ and L are clear from the context, a language
of type ≈, F, σ for L is called shortly a language.
Remark 3.1. (1) Until otherwise mentioned, we will use a fixed type denoted
simply by F . T(X) always refers to a term L-algebra of type F where X is
(at least) denumerable set of variables with its universe T (X) denoting the
set of all terms of type F over X.
(2) For the sake of convenience and since there is no danger of misunder-
standing, we identify each a with a ∈ L, i.e. we identify each symbol of truth
degree with the truth degree itself.
Terms (of type F over X) of fuzzy equational logic are those defined in
Definition 2.73. Formulas of fuzzy equational logic are identities.
3.1 Syntax and Semantics 141

Definition 3.2. Let F be a type. An identity of type F over X is an expres-


sion t ≈ t , where t, t ∈ T (X).

Remark 3.3. (1) We can see that a language of fuzzy equational logic is a part
of a language of predicate fuzzy logic as introduced in Example 1.96 (2). In
more detail, if we put R = {≈} and K = L in Example 1.96 (2), a language
of fuzzy equational logic of type ≈, F, σ results from a language of predicate
fuzzy logic of Example 1.96 (2) by removing of symbols of connectives and
quantifiers.
(2) Also terms of fuzzy equational logic coincide with terms of predicate
fuzzy logic of Example 1.96 (2). Formulas of fuzzy equational logic (identities)
are just the atomic formulas of predicate fuzzy logic with R = {≈}.
(3) Up to now, we did not make use of symbols a of truth degrees in
developing syntactic notions. Symbols a will be used when introducing the
notions of a proof and provability. Note that we do not consider a as formulas
of fuzzy equational logic, cf. Example 1.96 (2).

We are now going to introduce basic concepts of semantics of fuzzy equa-


tional logic. Structures of fuzzy equational logic are algebras with fuzzy equali-
ties. As usual, given an algebra M with fuzzy equality, a valuation (of variables
from X) in M is a mapping v : X → M assigning to each variable x ∈ X some
element v(x) ∈ M . Then, given an L-algebra M and a valuation v : X → M ,
a value !t!M,v of a term t, and a truth degree !t ≈ t !M,v of an identity t ≈ t
are defined in a straightforward way. The details follow.

Definition 3.4. Let M be an L-algebra of type F , v : X → M be a valuation,


t be a term. A value tM,v of t in M under v is defined as follows:
(i) if t is a variable x, then !t!M,v = v(x),
 
(ii) if t is of the form f (t1 , . . . , tn ), then !t!M,v = f M !t1 !M,v , . . . , !tn !M,v .

Remark 3.5. (1) !t!M,v is thus defined in a usual way.


(2) It follows directly from Theorem 2.79 that for each term t(x1 , . . . , xn )
we have
(i) !t!M,v = tM (a1 , . . . , an ) for any valuation v : X → M satisfying v(x1 ) =
a1 , . . . , v(xn ) = an ,
(ii) !t!M,v = v  (t) where v  is the homomorphic extension of v.
By (i), !t!M,v can be seen as a value of a term function tM induced by t.
By (ii), !t!M,v can be seen as a value assigned to a term t by the (unique)
homomorphic extension v  of v.

Definition 3.6. Let M be an L-algebra of type F , v : X → M be a valuation,


t ≈ t be an identity. A truth degree t ≈ t M,v of t ≈ t in M under v
is defined by
!t ≈ t !M,v = !t!M,v ≈M !t !M,v . (3.1)
142 3 Fuzzy Equational Logic

A truth degree t ≈ t M of t ≈ t in M is defined by



!t ≈ t !M = v:X→M !t ≈ t !M,v . (3.2)

Remark 3.7. (1) A truth degree !t ≈ t !M,v is defined in the usual sense of
predicate fuzzy logic, see Example 1.96 (2). !t ≈ t !M,v can be seen as follows.
First, we evaluate terms t and t and get elements !t!M,v and !t !M,v of M .
Then, we take a fuzzy equality ≈M of M and let !t ≈ t !M,v be just the degree
to which !t!M,v and !t !M,v are ≈M -equal.
(2) A truth degree !t ≈ t !M can be seen as a degree to which t ≈ t is
true in M under all valuations v : X → M .

∗ ∗ ∗
As to the concepts developed so far, we have seen that fuzzy equational
logic can be considered as a certain fragment of predicate fuzzy logic as in-
troduced in Example 1.96 (2). Our aim is to develop further concepts of fuzzy
equational logic in Pavelka style. To this end, we proceed by presenting fuzzy
equational logic in terms of abstract logic as defined in Definition 1.105. This
gives us for free the concepts of a theory, model, degree of semantic entailment,
proof, degree of provability, soundness, and completeness in fuzzy equational
logic (see Sect. 1.4). Nevertheless, we recall all of these concepts for the par-
ticular setting of fuzzy equational logic.
Recall from Definition 1.105 (page 56) that an abstract logic is a tuple
L = Fml , L, S, A, R where Fml is a set of formulas, L is a structure of truth
degrees, S is an L-semantics for Fml , A is a theory of logical axioms, and R
is a set of deduction rules. In the following, we present fuzzy equational logic
as a particular abstract logic. That is, we define its components Fml , L, S,
A, and R.
For a set Fml of formulas, we take the set of all identities, i.e.
Fml = {t ≈ t | t, t ∈ T (X)}.
For L we take an arbitrary complete residuated lattice. Recall that ac-
cording to Definition 1.105, L is required to be a complete lattice. We assume
that L is, moreover, a complete residuated lattice. That is, L is a complete
lattice with an additional structure on it.
For S we take an L-semantics S for Fml defined by
S = {E ∈ LFml | for some L-algebra M : E(t ≈ t ) = !t ≈ t !M (3.3)
for each t ≈ t ∈ Fml },
cf. (1.92).
For A, we take an empty L-set of formulas, i.e. A(t ≈ t ) = 0 for each
identity t ≈ t .
For R we take a set of deduction rules (ERef)–(ESub) for Fml and L intro-
duced below. Before presenting the rules, introduce the following concepts. A
subterm of a term t is any substring of t which is a term itself. For a variable
3.1 Syntax and Semantics 143

x ∈ X and a term r ∈ T (X), the pair (x/r) is called a substitution of r for x.


Given a substitution (x/r) and a term t ∈ T (X), let t(x/r) denote the term
resulting from t by substitution of r for every occurrence of x in t. Now, R
consists of the following five deduction rules for Fml and L:
t ≈ t , a t ≈ t , a , t ≈ t , b
(ERef) : , (ESym) : , (ETra) : ,
t ≈ t, 1 t ≈ t, a t ≈ t , a ⊗ b

t ≈ t , a t ≈ t , a
(ERep) : , (ESub) : ,
s ≈ s , a t(x/r) ≈ t (x/r), a
where t, t , t , r, s ∈ T (X), a, b ∈ L are arbitrary, x ∈ X, term s has an
occurrence of t as a subterm and s is a term resulting from s by substitution
of one occurrence of t by t . These rules can be seen as allowing for the
following inference steps (in parentheses, we attach names of the rules which
will be used for the sake of brevity):
(ERef)(from the empty set of assumptions) infer t ≈ t, 1 (reflexivity);
(ESym) from t ≈ t , a infer t ≈ t, a (symmetry);
(ETra)from t ≈ t , a and t ≈ t , b infer t ≈ t , a ⊗ b (transitivity);
(ERep) from t ≈ t , a infer s ≈ s , a where s is a term containing t as a
subterm and s results from s by replacing one occurrence of t by t
(replacement);
(ESub) from t ≈ t , a infer t(x/r) ≈ t (x/r), a (substitution).

Remark 3.8. Rules (ERef)–(ESub) generalize the corresponding rules from the
ordinary case, see [23, 27]. Namely, the corresponding rules of ordinary equa-
tional logic are: infer t ≈ t (reflexivity); from t ≈ t infer t ≈ t (symmetry);
from t ≈ t and t ≈ t infer t ≈ t (transitivity); from t ≈ t infer s ≈ s
where s is a term containing t as a subterm and s results from s by replacing
one occurrence of t by t (replacement, instead of “one occurrence” we can
have “some occurrences” which is equivalent in the ordinary case); from t ≈ t
infer t(x/r) ≈ t (x/r) (substitution).

We have thus defined a particular abstract logic. This gives us automati-


cally the concepts defined in Sect. 1.4. However, in order for these concepts to
fit better into our desired framework, we need to extend them somewhat. In
particular, we want to deal with L-algebras instead of evaluations and with
classes of L-algebras instead of the corresponding sets of evaluations. The
following remark clarifies the basic relationships.

Remark 3.9. Due to (3.3), there is the following relationship between L-


algebras and evaluations from S. Each L-algebra M induces an evaluation
EM ∈ S by EM (t ≈ t ) = !t ≈ t !M for each identity t ≈ t . For an evalua-
tion E ∈ S, however, there might be whole class of L-algebras M such that
E = EM .
144 3 Fuzzy Equational Logic

Using mappings Md and Th defined by (1.93) and (1.94), we can introduce


mappings Mod and Id by
Mod(Σ) = {M | EM ∈ Md(Σ)} and Id(K) = Th({EM | M ∈ K})
where Σ ∈ L Fml
is a theory and K is a class of L-algebras. Now, we can say
that an L-algebra M is a model of a theory Σ if M ∈ Mod(Σ) (i.e., if EM
is a model of Σ, see Definition 1.97). It is easy to see that M is a model of
Σ ∈ LFml iff Σ(t ≈ t ) ≤ !t ≈ t !M for each identity t ≈ t ∈ Fml . Then,
Mod(Σ) is a class of L-algebras and Id(K) is an L-set of identities over X for
which we have
Mod(Σ) = {M | M is a model of Σ} , (3.4)

(Id(K))(t ≈ t ) = {!t ≈ t !M | M ∈ K} (3.5)

for each identity t ≈ t , cf. (1.93) and (1.94). Occasionally, if we want to make
the set X of variables expicit, we write IdX (K) instead of Id(K). Also, we
write Id(M) instead of Id({M}). Then, like Md and Th, Mod and Id form a
Galois connection and satisfy thus (1.95)–(1.100) with sets K’s of evaluations
replaced by classes K’s of L-algebras. As in Remark 1.107, we can define a
degree !t ≈ t !Σ to which t ≈ t semantically follows from Σ ∈ LFml by
!t ≈ t !Σ = (Id(Mod(Σ)))(t ≈ t ).
Putting fuzzy equational logic into a setting of abstract logic gives us thus
(with a small modification concerning evaluations vs. L-algebras) automati-
cally semantical concepts in Pavelka style. Notice that all the concepts of fuzzy
equational logic are uniquely determined by the choice of a type ≈, F, σ and
a complete residuated lattice L. Therefore, we might speak of a fuzzy equa-
tional logic given by ≈, F, σ and L. The following definition gives a summary
of semantical concepts of a fuzzy equational logic given by ≈, F, σ and L.
Definition 3.10. A theory is an L-set Σ of identities, i.e. Σ ∈ LFml . An
L-algebra M is a model of a theory Σ if M ∈ Mod(Σ). A degree t ≈ t Σ
to which t ≈ t semantically follows from a theory Σ is defined by
!t ≈ t !Σ = (Id(Mod(Σ)))(t ≈ t ) .
For a class K of L-algebras, a truth degree t ≈ t K of t ≈ t in K is
defined by
!t ≈ t !K = (Id(K))(t ≈ t ) .

Remark 3.11. (1) Let us return to some of the concepts of Definition 3.10
and rewrite the definitions in a way which might be more convenient for the
reader. We have
M is a model of Σ iff for each t ≈ t : Σ(t ≈ t ) ≤ !t ≈ t !M ,
  
!t ≈ t !K = M ∈ K !t ≈ t !M = M ∈ K v:X→M !t ≈ t !M,v ,

!t ≈ t !Σ = !t ≈ t !Mod(Σ) = M ∈ Mod(Σ) !t ≈ t !M =
 
= M ∈ Mod(Σ) v:X→M !t ≈ t !M,v .
3.1 Syntax and Semantics 145

(2) The above concepts are straightforward generalizations of ordinary


logical concepts. Namely, M is a model of Σ if each identity is true in M at
least in a degree to which it belongs to Σ. Furthermore, !t ≈ t !K can be seen
as a degree to which t ≈ t is true in all L-algebras from K. Finally, !t ≈ t !Σ
can be seen as a degree to which t ≈ t is true in each L-algebra which is a
model of Σ.
Definition 3.12. For a theory Σ ∈ LFml , Mod(Σ) is called an equational
class of Σ. For a class K of L-algebras, Id(K) is called an equational theory
of K. A class K of L-algebras is called an equational class if K = Mod(Σ)
for some theory Σ. An L-set Σ of identities is called an equational theory
if Σ = Id(K) for some class K of L-algebras.
Remark 3.13. Equational class of a theory Σ thus consists of all L-algebras
which are models of Σ. A theory of a class K of L-algebras contains identities,
each in a degree to which it is true in K. A class of L-algebras is equational
if it is definable by an L-set of identities.
We now turn to the concepts of a proof and provability. Again, we utilize
the notions of abstract logic, see Definition 1.101. In our setting, the concept
of an L-weighted formula (Definition 1.103) gives the following.
Definition 3.14. An L-weighted identity is a pair t ≈ t , a where t ≈ t
is an identity and a is a symbol of a truth degree a ∈ L.
Remark 3.15. Remark 1.104 applies here. In particular, an L-weighted identity
is a string of symbols of a language and is thus a part of syntax of fuzzy
equational logic; we also write t ≈ t , a instead of t ≈ t , a ; we also say a
truth-weighted identity or just weighted identity.
We can now define concepts related to provability of a fuzzy equational
logic given by ≈, F, σ and L.
Definition 3.16. Let R be a set of deduction rules for Fml and L, Σ ∈
LFml be a theory (L-set of identities), let t ≈ t ∈ Fml be an identity and
a ∈ L. An (L-weighted ) proof of t ≈ t , a from Σ using R is a sequence
t1 ≈ t1 , a1 , . . . , tn ≈ tn , an such that tn ≈ tn is t ≈ t , an = a, and for
each
 i = 1,. . . , n, we have ai = Σ(ti ≈ ti ) or ti ≈ ti , ai follows from some
tj ≈ tj , aj ’s, j < i, by some deduction rule R ∈ R. The number n is called
a length of the proof. In such a case, we write Σ R t ≈ t , a and call
t ≈ t , a provable from Σ using R. If Σ R t ≈ t , a , t ≈ t is called
provable in degree (at least) a from Σ using R. A degree of provability
of t ≈ t from Σ using R, denoted by |t ≈ t |R Σ , is defined by

|t ≈ t |R
Σ = {a | Σ R t ≈ t , a } .
Remark 3.17. The conventions introduced in Remark 1.102 apply. In particu-
lar, R will denote a fixed set consisting of deduction rules (ERef)–(ESub) and
will therefore be omitted.
146 3 Fuzzy Equational Logic

∗ ∗ ∗
Note that it is the Pavelka style what makes fuzzy equational logic inter-
esting. For instance, if we had considered only provability from crisp sets of
identities, fuzzy equational logic would, in a sense, collapse to ordinary equa-
tional logic. This would also be the case of semantic entailment for crisp sets
of identities.
First, consider provability from crisp L-sets Σ of identites. That is, for
any t ≈ t we either have Σ(t ≈ t ) = 1 or Σ(t ≈ t ) = 0. An easy inspection
of deduction rules (ERef)–(ESub) shows that if t ≈ t , a is a member of a
proof from crisp Σ then a ∈ {0, 1}. Hence, |t ≈ t |Σ ∈ {0, 1} for every t, t ∈
T (X). Moreover, one can easily see that when applied to weighted identities
ti ≈ ti , 1 ’s (i.e., with weights equal to 1), our rules (ERef)–(ESub) yield a
weighted identity t ≈ t , 1 if and only if t ≈ t can be obtained from ti ≈ ti ’s
by the corresponding rules of ordinary equational logic (see Remark 3.8).
Therefore, with crisp Σ, fuzzy equational logic in fact collapses to ordinary
equational logic. Once again, even if we allow for degrees of provability in
Pavelka style, degrees of provability from crisp theories can only be 0 or 1
and they coincide with “non-provable” or “provable” in ordinary equational
logic. Note that this is not the case of general predicate fuzzy logic (one might
have a degree of provability other than 0 and 1 even if one proves from a crisp
theory).
Second, consider semantic entailment for a crisp L-set Σ. Observe that an
identity t ≈ t is fully true in an L-algebra M (i.e., !t ≈ t !M = 1) iff for any
valuation v we have 1 = !t ≈ t !M,v = !t!M,v ≈M !t !M,v which is true iff for
any valuation v, !t!M,v equals !t !M,v . That is, the fact that t ≈ t is fully true
 
in M depends only on the functional part M, F M of M (does not depend
on the L-equality ≈M ). Therefore, as one can easily see, semantic entailment
from crisp L-sets of identities coincides with semantic entailment of ordinary
equational logic. As a consequence, for a crisp Σ, a degree !t ≈ t !Σ to which
t ≈ t semantically follows from Σ can only be 0 or 1 and it coincides with
“does not follow from Σ” and “follows from Σ” in ordinary equational logic.
That is, for crisp Σ’s, semantic entailment would collapse to the ordinary case
as well.

3.2 Completeness of Fuzzy Equational Logic


In this section we work out a proof of (Pavelka)-completeness of fuzzy equa-
tional logic. That is, we show that for any complete residuated lattice, an
L-set Σ of identities, and an identity t ≈ t the degree !t ≈ t !Σ to which
t ≈ t semantically follows from Σ equals the degree of provability of t ≈ t
from Σ.
First, we describe semantic entailment using appropriate algebraic means.
In particular, we define so-called fully invariant congruences, show that equa-
tional theories can be seen as fully invariant congruences on T(X), and use
3.2 Completeness of Fuzzy Equational Logic 147

fully invariant congruences to describe semantic entailment. We start by an


auxiliary lemma.
Lemma 3.18. Suppose that T(X) of type F exists and let M be an L-algebra
of type F . For every valuation v : X → M on M and for every endomorphism
h of T(X), there is a valuation w : X → M such that !h(t)!M,v = !t!M,w for
every term t ∈ T (X).
Proof. For v : X → T (X) and endomorphism h : T(X) → T(X), we can
put w(x) = !h(x)!M,v for every variable x ∈ X. By structural induction we
can easily prove that !h(t)!M,v = !t!M,w . Indeed, for x ∈ X the assertion is
trivial. For t of the form f (t1 , . . . , tn ) we have
 
!h(t)!M,v = !f (h(t1 ), . . . , h(tn ))!M,v = f M !h(t1 )!M,v , . . . , !h(tn )!M,v =
 
= f M !t1 !M,w , . . . , !tn !M,w = !f (t1 , . . . , tn )!M,w = !t!M,w .
Thus, !h(t)!M,v = !t!M,w for every t ∈ T (X). 

Definition 3.19. We call a congruence θ on M fully invariant if for any
endomorphism h : M → M we have θ(a , b ) ≤ θ(h(a ), h(b )) (a , b ∈ M ).
Note that theories Σ (i.e., L-sets of identities) can be identified with binary
L-relations in T (X). Namely, given a theory Σ ∈ LFml , we may consider a
corresponding binary L-relation θΣ in T (X) defined for each p, q ∈ T (X) ×
T (X) by θΣ (p, q) = Σ(p ≈ q). Conversely, given a binary L-relation θ in T (X),
we may consider a theory Σθ ∈ LFml defined for each identity p ≈ q ∈ Fml
by Σθ (p ≈ q) = θ(p, q). Then we clearly have Σ = ΣθΣ and θ = θΣθ . For
convenience, we make use of this possibility and identify L-sets of identities
with binary L-relations in T (X) in the subsequent text.
Theorem 3.20. IdX (K) is a fully invariant congruence on T(X).
  
Proof. For brevity, we abbreviate “ M∈K v:X→M ” by “ M,v ”. Reflexivity
and symmetry of IdX (K) are obvious. In more detail, we have
    
(IdX (K))(t, t) = M,v !t ≈ t!M,v = M,v !t!M,v ≈M !t!M,v = M,v 1 = 1
for each t ∈ T (X), i.e. IdX (K) is reflexive. Furthermore,
   
(IdX (K))(t, t ) = M,v !t ≈ t !M,v = M,v !t!M,v ≈M !t !M,v =
   
= M,v !t !M,v ≈M !t!M,v = M,v !t ≈ t!M,v =
= (IdX (K))(t , t) ,
i.e. IdX (K) is symmetric. Transitivity of IdX (K) follows by
(IdX (K))(t, t ) ⊗ (IdX (K))(t , t ) =
 
= M,v !t ≈ t !M,v ⊗ M,v !t ≈ t !M,v ≤
  
≤ M,v !t ≈ t !M,v ⊗ !t ≈ t !M,v =
  
= M,v !t!M,v ≈M !t !M,v ⊗ !t !M,v ≈M !t !M,v ≤
 
≤ M,v !t!M,v ≈M !t !M,v = M,v !t ≈ t !M,v = (IdX (K))(t, t ) .
148 3 Fuzzy Equational Logic

Since ≈T(X) is the identity relation, ≈T(X) ⊆ IdX (K) is trivial, showing com-
patibility of IdX (K) with ≈T(X) . For any n-ary function symbol f we have
( IdX (K))(t1 , t1 ) ⊗ · · · ⊗ (IdX (K))(tn , tn ) =
 
= M,v !t1 ≈ t1 !M,v ⊗ · · · ⊗ M,v !tn ≈ tn !M,v ≤
  
≤ M,v !t1 ≈ t1 !M,v ⊗ · · · ⊗ !tn ≈ tn !M,v =
  
= M,v !t1 !M,v ≈M !t1 !M,v ⊗ · · · ⊗ !tn !M,v ≈M !tn !M,v ≤
    
≤ M,v f M !t1 !M,v , . . . , !tn !M,v ≈M f M !t1 !M,v , . . . , !tn !M,v =

= M,v !f (t1 , . . . , tn )!M,v ≈M !f (t1 , . . . , tn )!M,v =

= M,v !f (t1 , . . . , tn ) ≈ f (t1 , . . . , tn )!M,v =
= (IdX (K))(f (t1 , . . . , tn ), f (t1 , . . . , tn )) .
Thus, IdX (K) ∈ ConL (T(X)). Finally, we check that IdX (K) is fully invariant.
Let h : T(X) → T(X) be an endomorphism. Take M ∈ K. For any valuation
v : X → M , Lemma 3.18 yields that there is w : X → M such that !h(t)!M,v =
!t!M,w for every t ∈ T (X). Therefore,

(IdX (K))(t, t ) ≤ !t ≈ t !M,w = !t!M,w ≈M !t !M,w =


= !h(t)!M,v ≈M !h(t )!M,v = !h(t) ≈ h(t )!M,v ,

yielding (IdX (K))(t, t ) ≤ M,v !h(t) ≈ h(t )!M,v = (IdX (K))(h(t), h(t )). To
sum up, IdX (K) is a fully invariant congruence of T(X). 


We need the following lemma.

Lemma 3.21. Suppose that T(X) of type F exists and let θ ∈ ConL (T(X)).
Let v : X → T (X)/θ be a valuation. Then there is an endomorphism h on
T(X) such that [h(t)]θ = !t!T(X)/θ,v .

Proof. Let v : X → T (X)/θ be a valuation. Take any hX : X → T (X) such


that hX (x) ∈ v(x) for all x ∈ X. Let h be the homomorphic extension of
hX . For a natural morphism hθ : T(X) → T(X)/θ we have (hX ◦ hθ )(x) =
[hX (x)]θ = v(x) for all x ∈ X. Thus, Theorem 2.139 and Theorem 2.140 give
(hX ◦ hθ ) = v  . Using Theorem 2.141 we get h ◦ hθ = (hX ◦ hθ ) = v  . That
is, for any t ∈ T (X) we have
[h(t)]θ = hθ (h(t)) = (h ◦ hθ )(t) = v  (t) = !t!T(X)/θ,v
for every t ∈ T (X) which is the desired equality. 


Now we show the converse assertion to Theorem 3.20

Theorem 3.22. Let θ be a fully invariant congruence on T(X). Then θ is an


equational theory such that IdX ({T(X)/θ}) = θ.
3.2 Completeness of Fuzzy Equational Logic 149

Proof. We prove the claim by showing !t ≈ t !T(X)/θ = θ(t, t ).


“≤”: For a valuation v : X → T (X)/θ such that v(x) = [x]θ (x ∈ X) we
have !t!T(X)/θ,v = [t]θ for all t ∈ T (X) (follows from Lemma 3.21 for h such
that h(t) = t for each term t). Therefore,
!t ≈ t !T(X)/θ ≤ !t ≈ t !T(X)/θ,v = !t!T(X)/θ,v ≈T(X)/θ !t !T(X)/θ,v =
= [t]θ ≈T(X)/θ [t ]θ = θ(t, t ) .
“≥”: Take any v : X → T (X)/θ. Lemma 3.21 yields that there is an
endomorphism h on T(X), such that [h(t)]θ = !t!T(X)/θ,v for every t ∈ T (X).
It follows that
!t ≈ t !T(X)/θ,v = !t!T(X)/θ,v ≈T(X)/θ !t !T(X)/θ,v =
= [h(t)]θ ≈T(X)/θ [h(t )]θ = θ(h(t), h(t )) ≥ θ(t, t )
for all t, t ∈ T (X). Since v is chosen arbitrarily, !t ≈ t !T(X)/θ ≥ θ(t, t ). 


Putting Theorem 3.20 and Theorem 3.22 together, we get the following
corollary.

Corollary 3.23. Let Σ be an L-set of identities. Then Σ is a fully invariant


congruence on T(X) iff Σ is an equational theory. 


Lemma 3.24. The system of all fully invariant congruences on T(X) is a


closure system.

Proof. Clearly, T (X)×T (X) is a fully invariant congruence on T(X). It is now


sufficient to show that for any non-empty  family {θi ∈ ConL (T(X)) | i ∈ I}
of fully invariant congruences, θ = {θi | i ∈ I} is a fully invariant con-
gruence on T(X). According to Theorem 2.15, it suffices to show that
θ(t, t ) ≤ θ(h(t), h(t )) for every endomorphism h on T(X) and t, t ∈ T (X).
However, this is true since
 
θ(t, t ) = i∈I θi (t, t ) ≤ i∈I θi (h(t), h(t )) = θ(h(t), h(t )) .
Thus, θ is a fully invariant congruence on T(X). 


For an L-set Σ of identities denote by Σ  the least fully invariant congru-


ence on T(X) such that Σ ⊆ Σ  (its existence follows from Lemma 3.24).

Theorem 3.25. For any L-set Σ of identities, and every terms t, t ∈ T (X)
we have !t ≈ t !Σ = Σ  (t, t ).

Proof. “≤”: Theorem 3.22 yields T(X)/Σ  ∈ Mod(Σ  ). Since Σ ⊆ Σ  , it


follows that Mod(Σ  ) ⊆ Mod(Σ). Thus, T(X)/Σ  is a model of Σ. Using
Theorem 3.22 we get

!t ≈ t !Σ = M∈Mod(Σ) !t ≈ t !M ≤ !t ≈ t !T(X)/Σ  =
= (IdX (T(X)/Σ  ))(t, t ) = Σ  (t, t ) .
150 3 Fuzzy Equational Logic

“≥”: Take any M ∈ Mod(Σ), i.e. !t ≈ t !M ≥ Σ(t ≈ t ) for all t, t ∈ T (X).


Thus, Σ ⊆ IdX (M). Theorem 3.20 gives that IdX (M) is a fully invariant
congruence containing Σ. Since Σ  is the least fully invariant congruence
containing Σ, we have Σ  ⊆ IdX (M). Therefore, Σ  (t, t ) ≤ !t ≈ t !M for
every M ∈ Mod(Σ) and arbitrary t, t ∈ T (X). This further yields

!t ≈ t !Σ = M∈Mod(Σ) !t ≈ t !M ≥ Σ  (t, t )
which is the desired inequality. 


Remark 3.26. Since !t ≈ t !Σ = IdX (Mod(Σ))(t ≈ t ), Theorem 3.25 says that


IdX (Mod(Σ)) = Σ  .

Definition 3.27. Let Σ be an L-set of identities. We define a deductive


closure Σ  of Σ by letting Σ  be the least L-set of identities such that
Σ ⊆ Σ (3.6)

Σ (t, t) = 1 , (3.7)
   
Σ (t, t ) ≤ Σ (t , t) , (3.8)
      
Σ (t, t ) ⊗ Σ (t , t ) ≤ Σ (t, t ) , (3.9)
   
Σ (t, t ) ≤ Σ (s, s ) , (3.10)
   
Σ (t, t ) ≤ Σ (t(x/r), t (x/r)) , (3.11)
 
where t, t , t , r, s ∈ T (X), x ∈ X, term s has an occurrence of t as a subterm
and s is a term resulting from s by substitution of one occurrence of t by t .

Remark 3.28. Note that the existence of Σ  follows from the fact that the set
of all L-sets of identities which contain Σ and satisfy (3.6)–(3.11) is non-empty
(it contains the full L-set) and is closed with respect to arbitrary intersections
(easy to check).

Now, the next theorem shows that the deductive closure Σ  is just the
least fully invariant congruence Σ  on T(X).

Theorem 3.29. Let Σ be an L-set of identities. Then Σ  = Σ  .

Proof. “⊆”: It suffices to check that Σ  is a fully invariant congruence con-


taining Σ. By (3.6)–(3.9), Σ  is an L-equivalence on T (X) which contains Σ.
To show that Σ  is a congruence on T(X), we repeatedly use (3.10). For
any t1 , s1 , . . . , tn , sn ∈ T (X) and any n-ary f ∈ F , (3.10) yields
Σ (ti , si ) ≤ Σ (f (s1 , . . . , si−1 , ti , ti+1 , . . . , tn ), f (s1 , . . . , si−1 , si , ti+1 , . . . , tn ))
for each i = 1, . . . , n. Recall that in every i-th step, we have applied (3.10)
correctly, because f (s1 , . . . , si−1 , ti , ti+1 , . . . , tn ) contains ti as a subterm and
f (s1 , . . . , si−1 , si , ti+1 , . . . , tn ) results from it by replacing ti by si . Hence, (3.9)
yields
3.2 Completeness of Fuzzy Equational Logic 151

Σ  (t1 , s1 ) ⊗ · · · ⊗ Σ  (tn , sn ) ≤
n  
≤ i=1 Σ  f (s1 , . . . , si−1 , ti , . . . , tn ), f (s1 , . . . , si , ti+1 , . . . , tn ) ≤
 
≤ Σ  f (t1 , . . . , tn ), f (s1 , . . . , sn ) ,
proving compatibility. Thus, Σ  is a congruence.
We show that Σ  is fully invariant. Take any endomorphism h on T(X),
and terms t, t ∈ T (X). Suppose var(t) ∪ var(t ) ⊆ {x1 , . . . , xn }. Since X is
denumerable, we can consider paiwise distrinct variables y1 , . . . , yn so that no
yi occurs in xj or h(xj ) (j = 1, . . . , n). Using (3.11), we get
Σ  (t(x1 , . . . , xn ), t (x1 , . . . , xn )) ≤ Σ  (t(y1 , . . . , yn ), t (y1 , . . . , yn )) ≤
≤ Σ  (t(h(x1 ), . . . , yn ), t (h(x1 ), . . . , yn )) ≤ · · · ≤
≤ Σ  (t(h(x1 ), . . . , h(xn )), t (h(x1 ), . . . , h(xn ))) =
= Σ  (h(t(x1 , . . . , xn )), h(t (x1 , . . . , xn ))) ,
completing the proof.
“⊇”: We check that Σ  satisfies conditions (3.6)–(3.11). Equation (3.6)
follows from definition of Σ  . Equations (3.7)–(3.9) are true because Σ  is an
L-equivalence.
Equation (3.10) follows from the compatibility of Σ  . Indeed, take t, t , s, s ∈
T (X) such that s has an occurrence of t as a subterm and s is a term re-
sulting from s by substitution of t by t . If s = f (t1 , . . . , tk−1 , t, tk+1 , . . . , tn )
and s = f (t1 , . . . , tk−1 , t , tk+1 , . . . , tn ), compatibility of Σ  with f ∈ F and
Σ  (ti , ti ) = 1 yield
k−1 n
Σ  (t, t ) = i = 1 Σ  (ti , ti ) ⊗ Σ  (t, t ) ⊗ j = k+1 Σ  (tj , tj ) ≤
 
≤ Σ  f (t1 , . . . , tk−1 , t, tk+1 , . . . , tn ), f (t1 , . . . , tk−1 , t , tk+1 , . . . , tn ) =
= Σ  (s, s ) .
This argument can be used to show Σ  (t, t ) ≤ Σ  (s, s ) even in general case
(one can proceed by structural induction over the rank of s).
(3.11): For any x ∈ X and r ∈ T (X) consider a mapping g : X → T (X)
defined by g(x) = r and g(y) = y if y = x. The homomorphic extension g  of
g is an endomorphism on T(X) satisfying g  (t) = t(x/r) for each t ∈ T (X).
Therefore,
Σ  (t, t ) ≤ Σ  (g  (t), g  (t )) = Σ  (t(x/r), t (x/r))
because Σ  is fully invariant. 


The next theorem shows that the deductive closure Σ  coincides with
provability from Σ.

Theorem 3.30. Let Σ be an L-set of identities. Then for every t, t ∈ T (X)


we have |t ≈ t |Σ = Σ  (t, t ).
152 3 Fuzzy Equational Logic

Proof. “≤”: Clearly, it suffices to check that if ti ≈ ti , ai is a member of


some L-weighted proof, then ai ≤ Σ  (ti , ti ). This is obvious if ai = Σ(ti ≈ ti ).
Otherwise (i.e., ti ≈ ti , ai is obtained by some inference rule), proceed by
induction over i and suppose that the assertion is true for j < i. The following
inference rules could have been used:
(ERef): ti = ti and ai = 1 = Σ  (ti , ti ) by (3.7).
(ESym): ti = tj , ti = tj , and ai = aj for some j < i. Using (3.8),
ai = aj ≤ Σ  (tj , tj ) ≤ Σ  (tj , tj ) = Σ  (ti , ti ) .
(ETra): Suppose ti ≈ ti , ai was obtained by (ETra) from tj ≈ tj , aj
and tk ≈ tk , ak (in this order) for some j, k < i That is, we have ti = tj ,
tj = tk , tk = ti , and ai = aj ⊗ ak . Then, (3.9) yields
ai = aj ⊗ ak ≤ Σ  (tj , tj ) ⊗ Σ  (tk , tk ) = Σ  (ti , tk ) ⊗ Σ  (tk , ti ) ≤ Σ  (ti , ti ) .
(ERep): ti is obtained from ti by replacement of one occurrence of tj by
tj , and ai = aj for some j < i, whence by (3.10) we get
ai = aj ≤ Σ  (tj , tj ) ≤ Σ  (ti , ti ) .
(ESub): ti = tj (x/r), ti = tj (x/r), and ai = aj for some r ∈ T (X) and
j < i. Thus, (3.11) gives
ai = aj ≤ Σ  (tj , tj ) ≤ Σ  (tj (x/r), tj (x/r)) = Σ  (ti , ti ) .
“≥”: It suffices to prove that an L-set D of identities defined by D(t, t ) =
|t ≈ t |Σ contains Σ and satisfies (3.6)–(3.11). The required inequality then
follows from the fact that Σ  is the least L-set with these properties.
(3.6): Since t ≈ t , Σ(t ≈ t ) is a proof from Σ, we have Σ ⊆ D.
(3.7) follows from the fact that t ≈ t, 1 is a proof, i.e. D(t, t) = 1.
(3.8) follows from the fact that each proof δ1 , . . . , δk , t ≈ t , a from Σ
can be extended to a sequence δ1 , . . . , δk , t ≈ t , a , t ≈ t, a which is a proof
from Σ as well. Hence, D(t, t ) = |t ≈ t |Σ ≤ |t ≈ t|Σ = D(t , t).
(3.9): Let δi1 , . . . , δini , t ≈ t , ai (i ∈ I) and δj1 , . . . , δjnj , t ≈ t , bj (j ∈
J) be all proofs of  the form . . . , t ≈ t , ai and . . . , t ≈ t , bj , respectively.
 
Thus, D(t, t ) = i∈I ai and D(t , t ) = j∈J bj . For each i ∈ I and j ∈ J we
can concatenate the corresponding proofs and use (ETra) we get a sequence
δi1 , . . . , δini ,
1: t ≈ t , ai , proof of t ≈ t , ai
δj1 , . . . , δjnj ,
2: t ≈ t , bj , proof of t ≈ t , bj
3: t ≈ t , ai ⊗ bj . by (ETra) on 1, 2
which is a proof of t ≈ t , ai ⊗ bj from Σ. Hence,
 
D(t, t ) ⊗ D(t , t ) = i∈I ai ⊗ j∈J bj =
 
= i∈I j∈J (ai ⊗ bj ) ≤ |t ≈ t |Σ = D(t, t ) ,
which proves (3.9).
3.2 Completeness of Fuzzy Equational Logic 153

Equations (3.10) and (3.11) follow from the fact that if δ1 , . . . , δk , t ≈ t , a


is a proof from Σ and r ≈ s, a is obtained from t ≈ t , a by (ERep) or
(ESub), then δ1 , . . . , δk , t ≈ t , a , r ≈ s, a is a proof from Σ as well. 


We thus have (Pavelka-)completeness for fuzzy equational logic.

Theorem 3.31 (completeness). Let L be any complete residuated lattice,


X be denumerable set of variables. Let Σ be an L-set of identities over X.
Then |t ≈ t |Σ = !t ≈ t !Σ for every t, t ∈ T (X).

Proof. Consequence of Theorem 3.25, Theorem 3.29, and Theorem 3.30. 




Remark 3.32. A degree of provability may be strictly greater than the weight
of any proof. Indeed, let F = {◦}, ◦ be binary, denote by xn the n-th power of
x with respect to ◦, i.e. x3 = (x◦x)◦x, etc. Let L be the standard L
 ukasiewicz
algebra on [0, 1]. Define Σ by Σ(x ◦ x ≈ x) = 1, Σ(xn ≈ y n ) = 1 − n1 , and
Σ(t ≈ t ) = 0 otherwise. Clearly,
1: x ◦ x ≈ x, 1 , follows from Σ
2: y ◦ y ≈ y, 1 , by (ESub) on 1
3: xn ≈ y n , 1 − n
1
, follows from Σ
4: x ≈ x
n n−1
, 1 , by (ERep) on 1
5: xn−1 ≈ xn , 1 , by (ESym) on 4
6: xn−1 ≈ y n , 1 − n
1
, by (ETra) on 5 and 3
7: y ≈ y
n n−1
, 1 , by (ERep) on 2
8: xn−1 ≈ y n−1 , 1 − n
1
, by (ETra) on 6 and 7
..
.
x ≈ y, 1 − n1 ,
is an L-weighted proof of x ≈ y from Σ for any n. Therefore, |x ≈ y|Σ = 1. On
the other hand, there is no proof of x ≈ y from Σ the weight of which is 1: by
contradiction, let δ1 , . . . , δk be a proof of x ≈ y, 1 from Σ. Thus, δ1 , . . . , δk
also is a proof of x ≈ y, 1 from some finite subset Σ  ⊆ Σ because δ1 , . . . , δk
contains only finitely many weighted identities of the form xn ≈ y n , 1 − n1 .
Consider now an L-algebra M = M, ≈M , ◦M such that M = {a , b }, a ◦M
a = a , b ◦M b = b , a ◦M b = b ◦M a = a , and ≈M is an L-equality on
M satisfying a ≈M b = 1 − m 1
for m = 1 + max {n ∈ N | Σ  (xn ◦ y n ) = 0}.
Obviously, M ∈ Mod(Σ ). Since δ1 , . . . , δk is a proof of x ≈ y, 1 from Σ  ,


we get !x ≈ y!M = 1 due to the soundness of fuzzy equational logic. On the


other hand, a ≈M b = 1 − m 1
< 1, a contradiction to !x ≈ y!M = 1.

Remark 3.33. For readers familiar with Pavelka-style fuzzy logic: As observed
by Pavelka [74], we cannot have graded style completeness for arbitrary com-
plete residuated lattice even in the case of propositional logic (the less so for
the first-order case [49, 71]). However, since (ERef)–(ESub) can be used in
Pavelka-style first-order fuzzy logic as derived rules (more precisely: derived
154 3 Fuzzy Equational Logic

rules in first-order fuzzy logic with the usual inference rules where the relation
symbol ≈ is confined in an obvious sense by axioms of reflexivity, symmetry,
transitivity, and compatibility), our result implies that the equational frag-
ment (i.e. restriction to formulas of the form of identities) of first-order fuzzy
logic is completely axiomatizable (in Pavelka style) using any complete resid-
uated lattice as the structure of truth degrees.

3.3 Varieties, Free L-Algebras


In this section, we show basic properties of classes of L-algebras closed under
formation of homomorphic images, subalgebras, and direct products. Follow-
ing the ordinary case, we call them varieties.

Definition 3.34. A class K of L-algebras of type F is called a variety if it


is closed under homomorphic images, subalgebras and direct products, i.e., if
H(K) ⊆ K, S(K) ⊆ K and P(K) ⊆ K. If K is a class of L-algebras of the same
type let V(K) denote the least variety containing K. We say that V(K) is a
variety generated by K.

Remark 3.35. (1) Each variety is non-empty since it contains a trivial L-


algebra (i.e., an L-algebra with a one-element universe). Namely, a trivial
L-algebra is a direct product of an empty family of L-algebras.
(2) Note that V(K) exists since for a given type F , the class of all varieties
is non-empty (it contains the largest variety – the class of all L-algebras of
type F ) and closed under intersections (easy to check).

In the following, we describe the operator V using a single operator con-


sisting of H, S, and P.

Theorem 3.36. We have SH ≤ HS, PS ≤ SP, PH ≤ HP. Furthermore, H,


S, and IP are idempotent.

Proof. Suppose M ∈ SH(K). Then for some M ∈ K and some epimorphism


h : M → N, M is a subalgebra of N. Theorem 2.28 yields that the inverse
image h−1 (M) is a subalgebra of M . Moreover, h(h−1 (M)) = M. Hence, it
follows that M ∈ HS(K). Q
If M ∈ PS(K),
Q then M = i∈I Mi for Q suitable subalgebras Mi of Mi ∈ K
(i ∈ I). Since i∈I Mi is a subalgebra of i∈I Mi , we have M ∈ SP(K).
Let M ∈ PH(K). Then thereQare L-algebras Mi andQ epimorphisms Q hi :

Mi → Mi (i ∈ I) such that M = i∈I MiQ . A mapping h : i∈I Mi → i∈I Mi
defined by Q h(b )(i) = hi (b (i)) for all b ∈ i∈I Mi is an epimorphism. Indeed,
for a , b ∈ i∈I Mi we have,
Q    
a ≈ i∈I Mi b = i∈I a (i) ≈Mi b (i) ≤ i∈I hi (a (i)) ≈Mi hi (b (i)) =
 Q
= i∈I h(a )(i) ≈Mi h(b )(i) = h(a ) ≈ i∈I Mi h(b ) .
3.3 Varieties, Free L-Algebras 155
Q  Q  Q
For any n-ary f i∈I Mi ∈ F i∈I Mi , a1 , . . . , an ∈ i∈I Mi , we have
 Q    Q  
h f i∈I Mi (a1 , . . . , an ) (i) = hi f i∈I Mi (a1 , . . . , an )(i) =
    
= hi f Mi a1 (i), . . . , an (i) = f Mi hi (a1 (i)), . . . , hi (an (i)) =
  Q  
= f Mi h(a1 )(i), . . . , h(an )(i) = f i∈I Mi h(a1 ), . . . , h(an ) (i)
 Q   Q  
for each i ∈ I, i.e. h f i∈I Mi (a1 , . . . , an ) = f i∈I Mi h(a1 ), . . . , h(aQ n) .
The idempotency of H, S is evident. Let M ∈ IPIP(K), i.e. M = ∼ 
Q i∈I Mi
 ∼
and Mi = j∈Ji Nij for all iQ∈ I, where every Nij belongs Q to K. Thus, there
are isomorphisms h : M → i∈I Mi and hi : Mi → Q j∈Ji Nij (i ∈ I). Put
K = {i, j | i ∈ I, j ∈ Ji }. Now, a mapping g : M → i,j ∈K Nij defined by
 
g(a )(i, j ) = hi h(a )(i) (j) is an isomorphism. For a , b ∈ M we have
Q   
a ≈M b = h(a ) ≈ i∈I Mi h(b ) = i∈I h(a )(i) ≈Mi h(b )(i) =
 Q
= i∈I hi (h(a )(i)) ≈ j∈Ji Nij hi (h(b )(i)) =
 
= i∈I j∈Ji hi (h(a )(i))(j) ≈Nij hi (h(b )(i))(j) =
 
= i∈I j∈Ji g(a )(i, j ) ≈Nij g(b )(i, j ) =
 Q
= i,j ∈K g(a )(i, j ) ≈Nij g(b )(i, j ) = g(a ) ≈ i,j∈K Nij g(b ) .
The rest follows from the ordinary case. Hence, IP is idempotent. 

The following theorem says that the variety operator V is just the compo-
sition of H, S, and P.
Theorem 3.37. For the class operators V and HSP we have V = HSP.
Proof. The proof is fully analogous to that one in the ordinary case. For every
class K of L-algebras of the same type we have HSP(K) ⊆ HSP(V(K)). Since
H(V(K)) ⊆ V(K), S(V(K)) ⊆ V(K) and P(V(K)) ⊆ V(K) we obtain
HSP(K) ⊆ HSP(V(K)) ⊆ HS(V(K)) ⊆ H(V(K)) ⊆ V(K) .
Hence, HSP(K) is contained in V(K). Now it suffices to show that HSP(K)
is a variety, i.e. HSP(K) is closed under H, S and P. Using Theorem 3.36 we
have H(HSP) = HHSP = HSP and S(HSP) = SHSP ≤ HSSP = HSP, thus
HSP(K) is closed under homomorphic images and subalgebras. Furthermore,
P(HSP) = PHSP ≤ HPSP ≤ HSPP. As P ≤ IP, I ≤ H and IP is idempotent,
it readily follows that
HSPP ≤ HSIPIP = HSIP ≤ HSHP ≤ HHSP = HSP ,
thus HSP(K) is closed under direct product. Altogether HSP(K) is a variety
and HSP = V. 

We now introduce so-called free L-algebras and show their properties and
their role. In particular, we show that each variety is fully determined by its
free L-algebras. Namely, a variety is just the class of all homomorphic images
of its free L-algebras (Corollary 3.42 and Remark 3.43) and, moreover, it is
generated by a single free L-algebra (Theorem 3.44).
156 3 Fuzzy Equational Logic

Definition 3.38. Let K be a class of L-algebras of the same type, X be a set


of variables. Put
ΦK (X) = {φ ∈ ConL (T(X)) | T(X)/φ ∈ IS(K)} , (3.12)
i.e. ΦK (X) is the set of all congruences φ on T(X) such that the factor L-
algebra T(X)/φ is isomorphic to some subalgebra of some M ∈ K. Let

θK (X) = ΦK (X) (3.13)
denote the intersection of all congruences from ΦK (X). Then
FK (X) = T(X)/θK (X) , (3.14)
 
where X = [x]θK (X) | x ∈ X is the set of generators of T(X)/θK (X), is
called a K-free L-algebra over X. For x ∈ X we write x instead of [x]θK (X) .
Theorem 3.39. Let K be a class of L-algebras of type F , X be a set of vari-
ables such that T(X) exists. Then
(i) any mapping g : X → M where M ∈ K can be extended to a morphism
from FK (X) to M;
(ii) FK (X) has the universal mapping property for K over X.
   
Proof. (i): Let M ∈ K and take a mapping g : X, ≈FK (X) → M, ≈M .
Furthermore, let n : T(X) → FK (X) denote the natural morphism, i.e. n(t) =
[t]θK (X) for all t ∈ T (X). Suppose nX is a restriction of n on X, i.e. nX is a
mapping nX : X → X. Thus, a universal mapping property of T(X) implies
that there is a morphism k : T(X) → M extending nX ◦g. Moreover, it follows
from Theorem 2.35 that T(X)/θk is isomorphic to a subalgebra of M ∈ K,
thus θk ∈ ΦK (X), θK (X) ⊆ θk . It is immediate that g is an ≈-morphism, since
the following inequality holds for every x, y ∈ X:
x ≈FK (X) y = θK (X)(x, y) ≤
≤ θk (x, y) = g(nX (x)) ≈M g(nX (y)) = g(x) ≈M g(y) .
Now using Theorem 2.39 we obtain FK (X)/(θk /θK (X)) ∼
= T(X)/θk . Hence,
there are mappings h, i and j,
h i j
FK (X) −−→ FK (X)/(θk /θK (X)) −−→ T(X)/θk −−→ M ,
where h : FK (X) → FK (X)/(θk /θK (X)) is a natural morphism, i is an iso-
morphism between FK (X)/(θk /θK (X)) and T(X)/θk . Finally, using Theo-
rem 2.35, j : T(X)/θk → M is defined by j([t]θk ) = k(t) for every t ∈ T (X).
Thus,
     
(h ◦ i ◦ j)(x) = j i h(x) = j i [x]θk /θK (X) = j [x]θk =
   
= k(x) = g nX (x) = g [x]θK (X) = g(x) .
Altogether h ◦ i ◦ j is a homomorphic extension of g.
(ii) is a consequence of (i). 

3.3 Varieties, Free L-Algebras 157

Theorem 3.40. Suppose T(X) exists. Then FK (X) ∈ ISP(K). Thus if K is


closed under I, S, P, in particular if K is a variety, then FK (X) ∈ K.

Proof. Theorem 2.45 and Lemma 2.64 yield that forQFK (X) = T(X)/θK (X)
there is a subdirect embedding h : T(X)/θK (X) → θ∈ΦK (X) T(X)/θ. Using
PS (K) ⊆ SP(K) and Theorem 3.36 we obtain
FK (X) ∈ IPS ({T(X)/θ | θ ∈ ΦK (X)}) =
= IPS ({T(X)/θ | θ ∈ ConL (T(X)) and T(X)/θ ∈ IS(K)}) ⊆
⊆ IPS IS(K) ⊆ ISPIS(K) = ISPS(K) ⊆ ISSP(K) = ISP(K)
completing the proof. 


Theorem 3.41. Let K denote a class of L-algebras of the same type, M ∈ K.


Then for some sufficiently large set of variables X we have M ∈ H(FK (X)).

Proof. Take a set of variables X such that |M | ≤ |X|. Every surjective map-
ping h : X → M has a surjective homomorphic extension h : FK (X) → M
due to Theorem 3.39 (i). Hence, we have M ∈ H(FK (X)). 


Thus, we have the following consequence.


 
Corollary 3.42. K ⊆ H FK (X) | X is a set of variables . 


Remark 3.43. If K is a variety then every FK (X) belongs to K. Thus, the
previous corollary yields K = H FK (X) | X is a set of variables , i.e. every
variety is determined by its K-free L-algebras.

The following theorem shows that every variety is closed under the for-
mation of direct unions. This fact suffices to prove that every variety K is
determined by the only one K-free L-algebra FK (X) over a denumerable set
X of variables.

Theorem 3.44. Every variety K of L-algebras is closed under direct unions.


Moreover, K = HSP({FK (X)}) for X being a denumerable set of variables.

Proof. Let {M  i ∈ K | i ∈ I} be a directed family of L-algebras.QLemma 2.89


yields that i∈I Mi ∼ = N/θ for a suitable L-algebra N ∈ Sub( i∈I Mi ) and
θ ∈ ConL (N). Since  K = HSP(K), we have N ∈ SP(K) = K and N/θ ∈
H({N}) ⊆ K, i.e. i∈I Mi ∈ I(K) = K. Thus, every variety is closed under
the formation of direct unions.
Let us have M ∈ K. Every finitely generated L-algebra [M  ]M , is an
image of FK (X), where X is a denumerable set of variables. Indeed, let h :
X → [M  ]M be any mapping such that h(X) = M  . This mapping can be
extended to a morphism h : FK (X) → [M  ]M . Take an element a ∈ [M  ]M .
Since [M  ]M is finitely generated, we can express a = tM (a1 , . . . , an ), where
M  = {a1 , . . . , an }, t ∈ T (X). Thus for xi1 , . . . , xin ∈ X such that h (xi1 ) =
a1 , . . . , h (xin ) = an we have
158 3 Fuzzy Equational Logic
   
h [t(xi1 , . . . , xin )]θK (X) = h tFK (X) (xi1 , . . . , xin ) =
 
= tM h (xi1 ), . . . , h (xin ) = a .
Hence, h : FK (X) → [M  ]M is a surjective morphism. Consequently, [M  ]M ∈
H({FK (X)}) ⊆ HSP({FK (X)}). Using Corollary 2.88 it follows that
  
M∼ = [M ] ∈ HSP({FK (X)}) | M  is a finite subset of M
M

for every M ∈ K. Since HSP({FK (X)}) is a variety, i.e. it is closed under


direct unions, we have M ∈ HSP({FK (X)}), i.e. K ⊆ HSP({FK (X)}). The
converse inclusion follows from Theorem 3.40. 

Theorem 3.45. Let K be a variety of L-algebras.
(i) K is closed under direct limits of weak direct families of L-algebras from K;
(ii) K is closed under reduced products of L-algebras from K.
Proof. (i): Let {Mi ∈ K | i ∈ I} be a weak direct family of L-algebras. It is eas-
ily seen that Theorem 2.118 yields lim Mi ∈ IUH({Mi | i ∈ I}) ⊆ IUH(K) ⊆ K
due to Theorem 3.44.Q Q 
(ii): Obviously, F Mi = i∈I Mi /θF ∈ HP(K) ⊆ K for every family
{Mi ∈ K | i ∈ I}. 


3.4 Equational Classes of L-Algebras


In this section, we prove Birkhoff’s variety theorem [23] in the setting of fuzzy
equational logic. The following lemma characterizes interpretations of terms
in homomorphic images, subalgebras, and direct products.
Lemma 3.46. Suppose that T(X) of type F exists and let M, Mi (i ∈ I) be
L-algebras of type F .
(i) For
 a morphism
 h : M → M , and a valuation v : X → M we have
h !t!M,v = !t!M ,v◦h for all t ∈ T (X);
(ii) for every valuation v : X → N , where N ∈ Sub(M) we have
!t!M,v = !t!N,v for all t ∈ T (X);
Q
(iii) for every valuation v : X → i∈I M i and every j ∈ I we have
!t!Q Mi ,v (j) = !t!Mj ,v◦πj for all t ∈ T (X).
i∈I

Proof. (i): Theorem 2.140 and Remark 3.5 (2) yield


 
h !t!M,v = h(v  (t)) = (v  ◦ h)(t) = (v ◦ h) (t) = !t!M ,v◦h
for any term t ∈ T (X).
(ii): Take an embedding h: N  → M defined by h(a ) = a for all a ∈ N .
By (i) we get !t!N,v = h !t!N,v = !t!M,v◦h Q = !t!M,v for any t ∈ T (X).
(iii): Since every j-th projection πj : i∈I Mi → Mj is a morphism, we
can apply (i) to get !t!Q Mi ,v (j) = πj !t!Q Mi ,v = !t!Mj ,v◦πj . 

i∈I i∈I
3.4 Equational Classes of L-Algebras 159

Lemma 3.46 (i) can be used for h being the natural morphism hθ : M →
M/θ. Then we obtain the following consequence.

Corollary 3.47. Suppose that T(X) of type F exists


 and
 M is an L-algebra
of type F . Then for every θ ∈ ConL (M) we have !t!M,v θ = !t!M/θ,w , where
w(x) = [v(x)]θ for all variables x ∈ X. 


Lemma 3.48. Let M be an L-algebra, v : X → M be a valuation. Then for


any terms t, t ∈ T (X) we have
!t ≈ t !M,v = !t!M,v ≈M !t !M,v = v  (t) ≈M v  (t ) = θv (t, t ) .

Proof. Follows by Definition 1.83 and Remark 3.5 (2). 




Lemma 3.49. Let M be an L-algebra. Then for any t, t ∈ T (X) we have


(i) !t ≈ t !M = !t ≈ t !N for every N ∼= M;
(ii) !t ≈ t !M ≤ !t ≈ t !h(M) for every epimorphism h;
(iii) !t ≈ t !M ≤ !t ≈ t !N for every N ∈ Sub(M);
(iv) !t ≈ t !{Mi |i∈I} ≤ !t ≈ t !Q Mi for every family {Mi | i ∈ I}.
i∈I

Proof. We postpone the proof of (i) after proving (ii).


(ii): Let h : M → N be an epimorphism, i.e. h(M) = N. Take any valuation
v : X → N . For v we can choose w : X → M such that h(w(x)) = v(x) for
each variable x ∈ X. Thus, w ◦ h = v. Applying Lemma 3.46 (i), we get
!t ≈ t !M ≤ !t ≈ t !M,w = !t!M,w ≈M !t !M,w ≤
   
≤ h !t!M,w ≈N h !t !M,w = !t!N,w◦h ≈N !t !N,w◦h =
= !t!N,v ≈N !t !N,v = !t ≈ t !N,v = !t ≈ t !h(M),v .
Therefore, !t ≈ t !M ≤ !t ≈ t !h(M),v for any v : X → h(M ) from which the
claim follows immediately.
(i): Consider M ∼= N. Thus, there is an isomorphism h : M → N. Since
both h and its inverse h−1 are epimorphisms, we can apply (ii) twice to get
!t ≈ t !M ≤ !t ≈ t !h(M) = !t ≈ t !N
and
!t ≈ t !N ≤ !t ≈ t !h−1 (N) = !t ≈ t !M .
Putting it together, !t ≈ t !M = !t ≈ t !N .
(iii): Following the idea of (ii), for any valuation v : X → N where N ∈
Sub(M), Lemma 3.46 (ii) gives
!t ≈ t !M ≤ !t ≈ t !M,v = !t!M,v ≈M !t !M,v =
= !t!N,v ≈N !t !N,v = !t ≈ t !N,v .
Hence, !t ≈ t !M ≤ !t ≈ t !N .
family {Mi | i ∈ I} of L-algebras of the same type. For any
(iv): Consider aQ
valuation v : X → i∈I M i , we can consider a family {vi : X → Mi | i ∈ I} of
160 3 Fuzzy Equational Logic

valuations such that vi (x) = πi (v(x)) for each i ∈ I and any variable x ∈ X.
That is, vi = v ◦ πi (i ∈ I). Now we can apply Lemma 3.46 (iii) to get
 
!t ≈ t !{Mi |i∈I} = i∈I !t ≈ t !Mi ≤ i∈I !t ≈ t !Mi ,vi =
  
= i∈I !t!Mi ,vi ≈Mi !t !Mi ,vi =
  
= i∈I !t!Mi ,v◦πi ≈Mi !t !Mi ,v◦πi =
  
= i∈I !t!Q Mi ,v (i) ≈Mi !t !Q Mi ,v (i) =
i∈I i∈I
Q
= !t!Q ≈ i∈I Mi !t !Q = !t ≈ t !Q .
i∈I Mi ,v i∈I Mi ,v i∈I Mi ,v

Since v was arbitrary, we get !t ≈ t !{Mi |i∈I} ≤ !t ≈ t !Q . 



i∈I Mi

The following lemma shows that the equational theory of a class K of L-


algebras coincides with those of classes of isomorphic copies, homomorphic
images, subalgebras, and direct products of K.

Lemma 3.50. For a class K of L-algebras we have


IdX (K) = IdX (I(K)) = IdX (H(K)) = IdX (S(K)) = IdX (P(K)) .

Proof. First we show IdX (K) = IdX (I(K)). Since K ⊆ I(K) we have IdX (K) ⊇
IdX (I(K)). Conversely, for any N ∈ I(K) there is M ∈ K such that N ∼ = M.
For such M, Lemma 3.49 (i) gives !t ≈ t !M = !t ≈ t !N for all t, t ∈ T (X).
This yields !t ≈ t !K ≤ !t ≈ t !N for any N ∈ I(K), i.e. IdX (K) ⊆ IdX (I(K)).
Next, since for O = H, O = S, or O = IP we have K ⊆ O(K), it follows that
IdX (K) ⊇ IdX (H(K)), IdX (K) ⊇ IdX (S(K)), and IdX (K) ⊇ IdX (IP(K)) which
yields IdX (K) ⊇ IdX (P(K)) since IdX (K) = IdX (I(K)). We thus need to estab-
lish the converse inclusions, i.e. to verify (IdX (K))(t ≈ t ) ≤ (IdX (O(K)))(t ≈
t ) for all t, t ∈ T (X). However, this is a consequence of Lemma 3.49 (ii)–(iv).
Indeed, we check the inequality for each operator separately.
“H(K)”: For any M ∈ H(K) there is an epimorphism h : M → M where
M ∈ K. Lemma 3.49 (ii) thus gives !t ≈ t !K ≤ !t ≈ t !M ≤ !t ≈ t !M for


all t, t ∈ T (X). This further yields



!t ≈ t !K ≤ M∈H(K) !t ≈ t !M = !t ≈ t !H(K)
for all t, t ∈ T (X). Therefore, IdX (K) ⊆ IdX (H(K)).
“S(K)”: Briefly, for any M ∈ S(K) there is M ∈ K with M ∈ Sub(M ).
Therefore, Lemma 3.49 (iii) gives !t ≈ t !K ≤ !t ≈ t !M ≤ !t ≈ t !M yielding
IdX (K) ⊆ IdX (S(K)).
Q
“P(K)”: Take i∈I Mi ∈ P(K), i.e. {Mi | i ∈ I} ⊆ K. By Lemma 3.49 (iv),
!t ≈ t !K ≤ !t ≈ t !{Mi |i∈I} ≤ !t ≈ t !Q
i∈I Mi

from which it follows that IdX (K) ⊆ IdX (P(K)). 




Theorem 3.51. Every equational class is a variety.


3.4 Equational Classes of L-Algebras 161

Proof. Let K = Mod(Σ). We show that K is closed under H, S, and P. Let O


denote any of H, S, or P. Take any M ∈ O(K). Lemma 3.50 yields
Σ ⊆ IdX (Mod(Σ)) = IdX (K) = IdX (O(K)) ⊆ IdX (M) ,
i.e. for any t ≈ t we have Σ(t ≈ t ) ≤ !t ≈ t !M whence M ∈ Mod(Σ). This
shows that K is closed under O. 

Lemma 3.52. For a class K of L-algebras and t, t ∈ T (X) we have
 
!t ≈ t !K = !t ≈ t !FK (X) = [t]θK (X) ≈FK (X) [t ]θK (X) = (θK (X))(t, t ) .
Proof. We prove the assertion by showing
!t ≈ t !K ≤ !t ≈ t !FK (X) ≤
 
≤ [t]θK (X) ≈FK (X) [t ]θK (X) ≤ (θK (X))(t, t ) ≤ !t ≈ t !K .
“!t ≈ t !K ≤ !t ≈ t !FK (X) ”: Theorem 3.40 yields FK (X) ∈ ISP(K). Thus,
using Lemma 3.50 we get
!t ≈ t !K = (IdX (K))(t ≈ t ) = (IdX (ISP(K)))(t ≈ t ) ≤ !t ≈ t !FK (X) .

“!t ≈ t !FK (X) ≤ ([t]θK (X) ≈FK (X) [t ]θK (X) )”: Take w : X → T (X)/θK (X)
sending x to [x]θK (X) . We have
  
!t ≈ t !FK (X) = v:X→T (X)/θK (X) !t!FK (X),v ≈FK (X) !t !FK (X),v ≤
≤ !t!FK (X),w ≈FK (X) !t !FK (X),w = [t]θK (X) ≈FK (X) [t ]θK (X) .

“([t]θK (X) ≈FK (X) [t ]θK (X) ) ≤ (θK (X))(t, t )” is true by definition since FK (X)
is T(X)/θK (X).
“(θK (X))(t, t ) ≤ !t ≈ t !K ”: We check that for any L-algebra M ∈ K and
any valuation v : X → M , (θK (X))(t, t ) ≤ !t ≈ t !M,v . Due to Lemma 3.48,
it suffices to prove θK (X) ⊆ θv . Clearly, θv ∈ ConL (T(X)). Furthermore,
Theorem 2.35 gives that T(X)/θv is isomorphic to a subalgebra of M which
gives T(X)/θv ∈ IS(K). Thus, by Definition 3.38, θv ∈ ΦK (X), i.e. θK (X) ⊆
θv . The required inequality now readily follows. 

Theorem 3.53. If K is a variety of L-algebras and X a denumerable set of
variables, then K = Mod(IdX (K)).
Proof. Denote K = Mod(IdX (K)). K is a variety by Theorem 3.51. Obviously,
K ⊆ K . It suffices to check the converse inclusion. Clearly, IdX (K ) ⊆ IdX (K)
because K ⊆ K . Conversely, IdX (K) ⊆ IdX (K ) is true iff for each M ∈ K we
have (IdX (K))(t ≈ t ) ≤ !t ≈ t !M which is true by definition of K . We thus
have IdX (K ) = IdX (K). By Lemma 3.52,
(θK (X))(t, t ) = !t ≈ t !K = !t ≈ t !K = (θK (X))(t, t )
for all t, t ∈ T (X), i.e. FK (X) coincides with FK (X). Hence, Theorem 3.44
yields K = HSP(FK (X)) = HSP(FK (X)) = K . 

162 3 Fuzzy Equational Logic

Theorem 3.54. Let K be a class of L-algebras of the same type, X be a


denumerable set of variables. Then the following are equivalent:
(i) K is an equational class,
(ii) K is closed under H, S, and P,
(iii) K is closed under HSP,
(iv) K = HSP(K),
(v) K = HSP(K ) for some class K of L-algebras,
(vi) K = Mod(IdX (K)),
(vii) K = Mod(IdX (K )) for some class K of L-algebras.

Proof. “(i) ⇒ (ii)”: By Theorem 3.51.


“(ii) ⇒ (iii)”: By Theorem 3.37.
“(iii) ⇒ (iv)”: (iv) is a restatement of (iii).
“(iv) ⇒ (v)”: Trivial.
“(v) ⇒ (vi)”: Apply Theorem 3.37 and Theorem 3.53.
“(vi) ⇒ (vii)”: Trivial.
“(vii) ⇒ (i)”: By definition. 


In the rest of this section, we show several examples of equational classes.

Example 3.55. Consider a type F = {·, −1 , 1}, where · is a binary function


symbol, −1 is a unary function symbol, and 1 is symbol of a constant (nullary
function symbol). For an arbitrary L and a, b, c ∈ L, consider weighted iden-
tities
G1: x · (y · z) ≈ (x · y) · z, a ,
G2: x · 1 ≈ x, b , 1 · x ≈ x, b ,
G3: x · x−1 ≈ 1, c , x−1 · x ≈ 1, c ,
and the corresponding theory ΣG (i.e., ΣG (x · (y · z) ≈ (x · y) · z) = a, etc.,
see Remark 3.15). A class K of all L-algebras satisfying G1–G3 is a variety.
It is easy to see that K contains, among other L-algebras, all groups endowed
with compatible L-equalities. In addition to that, K is determined by truth
degrees a, b, c ∈ L. Namely, for a = b = c = 0, K is the class of all L-algebras.
On the other hand, if a = b = c = 1, then K is the class of all groups with
L-equalities.

Example 3.56. Let L be the standard product algebra on [0, 1]. Consider an
L-equality depicted in the left table of Fig. 3.1. The operations described
in the middle and the right pair of tables of Fig. 3.1 are compatible with
the L-equality. Denote by M1 and M2 the L-algebras corresponding to the
operations of the middle and left part with 1M1 = b and 1M2 = a . Both M1
and M2 satisfy identities of commutativity and associativity in degree 1, i.e.
we have !x · y ≈ y · x!Mi = 1 and !x · (y · z) ≈
 (x · −1
y) · z!M
 i = 1 for i = 1, 2.
Furthermore, we have !x · 1 ≈ x!M1 = 1 and x · x ≈ 1M = 3/4 (M1 has

1
inverse elements w.r.t. b = 1M1 in degree 3/4), while !x · 1 ≈ x!M2 = 3/4
3.5 Properties of Varieties 163

Fig. 3.1. L-equality and operations of Example 3.56

Fig. 3.2. “Almost semilattice” of Example 3.57

 
(a = 1M2 is a neutral element of M2 in degree 3/4) and x · x−1 ≈ 1M = 1.
2
Thus, for instance M1 is a model of ΣG with a = 1, b = 1, c = 0.5, but is not
a model of ΣG for a = 1, b = 1, c = 0.8.

Example 3.57. Let L be the standard L  ukasiewicz algebra on [0, 1]. Consider
a type which consists of a single binary function symbol ◦. We define an L-
equality ≈M and a function ◦M on the universe set M = {a , b , c , d , e , f } by
tables in Fig. 3.2.  
One can check that ◦M is compatible with ≈M , i.e. M = M, ≈M , ◦M
is an L-algebra. Note that ◦M on ske(M) is idempotent but it is neither as-
sociative nor commutative. On the other hand, we have !x ◦ y ≈ y ◦ x!M =
!x ◦ (y ◦ z) ≈ (x ◦ y) ◦ z!M = 7/8. This can be read: “L-algebra M is com-
mutative in degree 7/8 and associative in degree 7/8”. As a consequence, M
is a member of some variety of L-algebras given by a theory to which the
identity x ◦ x ≈ x of idempotency belongs in degree 1, and both the identities
of associativity and commutativity belong in degree 7/8.

3.5 Properties of Varieties


When approaching an area (a well-developed theory, for instance) from a
point of view of fuzzy logic, one can (should, perhaps) ask the following two
“methodological questions”: First, how to proceed in order to obtain an ap-
propriate generalization of the area such that the ordinary things (definitions,
theorems, algorithms, methods) become special cases of their “fuzzy coun-
terparts”? Second, what new phenomena come up in fuzzy setting (“new”
since they might be completely hidden, and thus not interesting, in the ordi-
nary case)? Concerning equational logic, the preceding sections of this chap-
ter showed a feasible answer to the first question. In this section, we give
164 3 Fuzzy Equational Logic

two examples which may be supplied as (examples of) answers to the second
question.

Crisply Generated Varieties

The following class operator comes up naturally in a fuzzy setting.

Definition 3.58. For a class K of L-algebras of the same type, a new class
F(K) is defined by
F(K) = {M | M is an L-algebra such that
for some N ∈ K : ske(M) = ske(N)} .

Operator F is called fuzzification. We have M ∈ F(K) iff there is N ∈ K


such that the functional parts (i.e. ordinary algebras M, F M and N, F N )
of M and N coincide.

Remark 3.59. For L = 2 we have F(K) = K. Therefore, F is trivial and thus


not interesting in the ordinary case. We show, that in general, F can be used
to characterize equational classes given by crisp L-sets of identities.

Lemma 3.60. Let Σ be a crisp L-set of identities. Then for any L-algebras
M, N with ske(M) = ske(N) we have M ∈ Mod(Σ) iff N ∈ Mod(Σ).

Proof. Take M and N such that ske(M) = ske(N), i.e. the functional parts of
M and N coincide. Observe that for any valuation v : X → M and t ∈ T (X)
we have
!t!M,v = !t!ske(M),v = !t!ske(N),v = !t!N,v .

Thus, !t ≈ t !M,v = 1 iff !t!M,v ≈M !t !M,v = 1 iff !t!M,v = !t !M,v iff
!t!N,v = !t !N,v iff !t!N,v ≈N !t !N,v = 1 iff !t ≈ t !N,v = 1. Therefore, M ∈
Mod(Σ) iff for each t ≈ t such that Σ(t ≈ t ) = 1 we have !t ≈ t !M,v = 1
for any valuation v : X → M iff for each t ≈ t such that Σ(t ≈ t ) = 1 we
have !t ≈ t !N,v = 1 for any valuation v : X → N iff N ∈ Mod(Σ), proving
the claim. 


The following assertion shows that varieties closed under the fuzzification
operator are exactly the equational classes of crisp L-sets of identities.

Theorem 3.61. Let K be a variety of L-algebras. Then K = F(K) iff there is


a crisp L-set Σ of identities such that K = Mod(Σ).

Proof. “⇒”: Let K be a variety such that K = F(K). Consider the subclass
K ⊆ K such that K = {M ∈ K | ≈M is crisp}. Put Σ = IdX (K ). Evi-
dently, Σ is a crisp L-set of identities. We show K = Mod(Σ) by proving both
inclusions.
3.5 Properties of Varieties 165

“⊆”: Let N ∈ K. It remains to show that N is a model of Σ. Consider


M such that ≈M is crisp and ske(M) = ske(N). Since K is closed under F,
we have M ∈ K. Therefore, M ∈ K , i.e. M ∈ Mod(Σ). Lemma 3.60 yields
N ∈ Mod(Σ).
“⊇”: Clearly, K = Mod(IdX (K)) ⊇ Mod(IdX (K )) = Mod(Σ).
“⇐”: Let Σ be a crisp L-set of identities such that K = Mod(Σ). Take
an L-algebra N ∈ K. For any L-algebra M such that ske(M) = ske(N) we
have M ∈ K by Lemma 3.60. Therefore, F(K) ⊆ K. The converse inclusion is
trivial. 


Remark 3.62. One can further show that if K is a variety closed under F then
there exists a variety Kc of classical algebras such that K can be reconstructed
from Kc by means of fuzzification. In fact, K = {M ∈ K | ≈M is crisp} is
a class of L-algebras with crisp equalities which is closed under isomorphic
images, subalgebras and direct products. Moreover, if h(M) is an image of
M ∈ K which has crisp L-equality then h(M) ∈ K . Thus, one can verify
that Kc = {ske(M) | M ∈ K} is a (classical) variety of ordinary algebras.
In addition to that, K = {M | M is L-algebra such that ske(M) ∈ Kc }. See
also [15, 52].

Graded Permutability of Varieties

Permutability of congruences of ordinary algebras is one of the most studied


issues in universal algebra. Namely, permutability simplifies several proper-
ties and conditions of algebras. Note that an ordinary algebra M is called
permutable (has permuting congruences) if θ ◦ φ = φ ◦ θ for any two con-
gruences θ, φ of M. A class K of ordinary algebras is called permutable if
each of its members is a permutable algebra. For instance, groups, rings, and
Boolean algebras are examples of permutable algebras. Permutability of a vari-
ety of algebras was characterized using a simple term condition by Mal’cev [64]
in 1954. Mal’cev’s ingenious result, left without any interest for almost two
decades, started a long series of results known as Mal’cev-type characteriza-
tion of properties of algebras and classes of algebras. Mal’cev showed that a
variety is permutable if and only if there is a ternary term p(x, y, z) such that
each algebra of the variety satisfies both p(x, y, y) ≈ x and p(x, x, y) ≈ y.
In fuzzy setting, permutability of congruences can be introduced as a
graded property. That is, we might speak of a degree to which two congruences
are permutable. We are going to show that even in this case, a Mal’cev-type
characterization is possible.
In fuzzy set theory, there are several sound definitions of a composition of
fuzzy relations, all of which coincide in the ordinary case, see e.g. [10, 47]. For
our purposes, we present another generalization of one of the most common
compositions of L-relations. For binary L-relations R1 , R2 on U , we define
binary L-relations R1 ◦n R2 , R1 ◦ω R2 on U by
166 3 Fuzzy Equational Logic

(R1 ◦n R2 )(a , b ) = c ∈U (R1 (a , c ) ⊗ R2 (c , b )) ,
n
(3.15)
 
(R1 ◦ω R2 )(a , b ) = c ∈U n∈N (R1 (a , c ) ⊗ R2 (c , b ))n , (3.16)
where n ∈ N and a , b ∈ U . Observe that for n = 1 we obtain the usual
notion of composition of L-relations. Let us stress that in general we can have
(R1 ◦ R2 )(a , b ) = 1, although there is no c ∈ U such that R1 (a , c ) = 1 and
R2 (c , b ) = 1. On the other hand, such a situation cannot happen if 1 is an
∨-irreducible element of L.
Now we present the notion of a graded permutability. Recall that in the
ordinary case, congruences θ, φ are called permutable iff θ ◦ φ = φ ◦ θ, which is
equivalent to θ ◦ φ ⊆ φ ◦ θ. We are going to generalize classical permutability
using the subsethood degree “S” which generalizes the classical inclusion “⊆”,
see (1.66). For θ, φ ∈ ConL (M), we define a degree Per(θ, φ) to which θ, φ are
permutable by
Per(θ, φ) = S(θ ◦ω φ, φ ◦ θ) .
For an L-algebra M we put

Per(M) = θ,φ∈ConL (M) Per(θ, φ) .
For a variety K of L-algebras we define a degree Per(K) by

Per(K) = M∈K Per(M) . (3.17)
Thus, Per(K) represents a degree to which a variety K is permutable. Observe
that for L = 2, Per(K) = 1 iff K is permutable in the classical sense. The rela-
tionship between the classical permutability and the degree given by Per(K)
will be commented later on.
Now we turn our attention to a term characterization of Per(K). Re-
call that according to [64], congruences of any algebra of a variety K are
permutable iff there exists a ternary term p(x, y, z) such that identities
p(x, y, y) ≈ x and p(x, x, y) ≈ y are true in K. In fuzzy case, however, we deal
with a degree of permutability Per(K) ∈ L and with truth degrees !· · ·!K ∈ L
of identities. We start by three auxiliary results.
Lemma 3.63. Let K be a variety of L-algebras and let X = {x1 , . . . , xn } be
a set of variables. Then for every M ∈ K and θ ∈ ConL (M) we have
 
θK (X) p(x1 , . . . , xn ), q(x1 , . . . , xn ) ≤
 
≤ θ pM (a1 , . . . , an ), q M (a1 , . . . , an ) , (3.18)
where p, q ∈ T (X) and a1 , . . . , an ∈ M .
Proof. Let v : X → M be a valuation, where v(xi ) = ai for i = 1, . . . , n.
Clearly, v induces a valuation w : X → M/θ such that w(xi ) = [v(xi )]θ = [ai ]θ
for all i = 1, . . . , n. Thus,
 
θK (X) p(x1 , . . . , xn ), q(x1 , . . . , xn ) = !p ≈ q!K ≤ !p ≈ q!M/θ,w =
   
= !p!M/θ,w ≈M/θ !q!M/θ,w = !p!M,v θ ≈M/θ !q!M,v θ =
   
= θ !p!M,v , !q!M,v = θ pM (a1 , . . . , an ), q M (a1 , . . . , an ) ,
which is the desired inequality. 

3.5 Properties of Varieties 167

Lemma 3.64. Let K be a variety of L-algebras, X = {x1 , . . . , xm , y1 , . . . , yn }


and Y = {x1 , . . . , xm , y} be sets of variables. Then for every binary L-relation
R on T (X)/θK (X) such that Supp(R) = {yi , yj | 1 ≤ i, j ≤ n} we have
 
θ(R) pFK (X) (x1 , . . . , xm , y1 , . . . , yn ), q FK (X) (x1 , . . . , xm , y1 , . . . , yn ) ≤
≤ !p(x1 , . . . , xm , y, . . . , y) ≈ q(x1 , . . . , xm , y, . . . , y)!K (3.19)
for arbitrary terms p, q ∈ T (X).
Proof. Take a mapping h : X → T (Y )/θK (Y ), where h(xi ) = xi (i =
1, . . . , m), and h(yj ) = y (j = 1, . . . , n). We can assume the variable y to
be among y1 , . . . , yn . For every z1 , z2 ∈ X let z1 , z2 ∈ Y with h(z1 ) = z1
and h(z2 ) = z2 . By Theorem 3.39, h admits the uniquely determined homo-
morphic extension h : FK (X) → FK (Y ). Moreover,
 
θh yi , yj = h (yi ) ≈FK (Y ) h (yj ) = y ≈FK (Y ) y = 1
for 1 ≤ i, j ≤ n. Therefore, R ⊆ θh . As a consequence, θ(R) ⊆ θh . Thus,
 
θ(R) pFK (X) (x1 , . . . , xm , y1 , . . . , yn ), q FK (X) (x1 , . . . , xm , y1 , . . . , yn ) ≤
   
≤ h pFK (X) (x1 , . . . , xm , y1 , . . . ) ≈FK (Y ) h q FK (X) (x1 , . . . , xm , y1 , . . . ) =
= pFK (Y ) (x1 , . . . , xm , y, . . . , y) ≈FK (Y ) q FK (Y ) (x1 , . . . , xm , y, . . . , y) =
 
= θK (Y ) p(x1 , . . . , xm , y, . . . , y), q(x1 , . . . , xm , y, . . . , y) =
= !p(x1 , . . . , xm , y, . . . , y) ≈ q(x1 , . . . , xm , y, . . . , y)!K
is true for arbitrary terms p, q ∈ T (X). 

Lemma 3.65. Let K be a variety of L-algebras and let M ∈ K. Then we have
   
φ a , pM (a , b , b ) ⊗ θ pM (c , c , b ), b ≤
  n
≤ n∈N θ(a , c ) ⊗ φ(c , b ) →
    
→ φ a , pM (a , c , b ) ⊗ θ pM (a , c , b ), b (3.20)
for every term p(x, y, z) ∈ T (X), elements a , b , c ∈ M , and θ, φ ∈ ConL (M).
  n
Proof. For brevity, put d = n∈N θ(a , c ) ⊗ φ(c , b ) . Since every term has
only finitely many occurrences of a variable, it follows that
  
d = n∈N θ(a , c )n ⊗ φ(c , b )n ≤ θ(a , c )|p|x ⊗ φ(c , b )|p|y .
Moreover,
   
φ a , pM (a , b , b ) ⊗ θ pM (c , c , b ), b ≤
    
d → θ(a , c )|p|x ⊗ φ(c , b )|p|y ⊗ φ a , pM (a , b , b ) ⊗ θ pM (c , c , b ), b ≤
    
d → θ pM (a , c , b ), pM (c , c , b ) ⊗ θ pM (c , c , b ), b ⊗
   
⊗ φ a , pM (a , b , b ) ⊗ φ pM (a , b , b ), pM (a , c , b ) ≤
    
d → φ a , pM (a , c , b ) ⊗ θ pM (a , c , b ), b .
Hence, the inequality holds. 

168 3 Fuzzy Equational Logic

The following assertion presents a connection between graded permutabil-


ity of a variety and graded validity of Mal’cev identities in a variety. A degree
of permutability of a variety K is shown to be equal to a supremum of truth
degrees of identities connected in a conjunctive manner which ranges over all
ternary terms.
Theorem 3.66 (characterization of graded permutability). Let K be a
variety of L-algebras and let X = {x, y, z}. Then
  
Per(K) = p∈T (X) !p(x, y, y) ≈ x!K ⊗ !p(x, x, y) ≈ y!K . (3.21)
Proof. “≤”: Obviously, Per(K) ≤ Per(M) for every M ∈ K. Take principal
congruences θ(1/x, y ), θ(1/y, z ) ∈ ConL (FK (X)). Now using (3.19), we get
Per(K) ≤ Per(FK (X)) ≤ Per(θ(1/x, y ), θ(1/y, z )) =
= S(θ(1/x, y ) ◦ω θ(1/y, z ), θ(1/y, z ) ◦ θ(1/x, y )) ≤
   
≤ θ(1/x, y ) ◦ω θ(1/y, z ) (x, z) → θ(1/y, z ) ◦ θ(1/x, y ) (x, z) ≤
  n
≤ n∈N θ(1/x, y )(x, y) ⊗ θ(1/y, z )(y, z) →
   
→ p∈T (X) θ(1/y, z ) x, pFK (X) (x, y, z) ⊗
 
⊗ θ(1/x, y ) pFK (X) (x, y, z), z =
   
= p∈T (X) θ(1/y, z ) x, pFK (X) (x, y, z) ⊗
 
⊗ θ(1/x, y ) pFK (X) (x, y, z), z ≤
  
≤ p∈T (X) !p(x, y, y) ≈ x!K ⊗ !p(x, x, y) ≈ y!K .
“≥”: Take arbitrary M ∈ K, a , b , c ∈ M , and θ, φ ∈ ConL (M). Properties of
free L-algebras together with (3.18) and (3.20) yield
  
p∈T (X) !p(x, y, y) ≈ x!K ⊗ !p(x, x, y) ≈ y!K =
  
= p∈T (X) θK (X)(x, p(x, y, y)) ⊗ θK (X)(p(x, x, y), y) ≤
     
≤ p∈T (X) φ a , pM (a , b , b ) ⊗ θ pM (c , c , b ), b ≤
   n
≤ p∈T (X) n∈N θ(a , c ) ⊗ φ(c , b ) →
    
→ φ a , pM (a , c , b ) ⊗ θ pM (a , c , b ), b ≤
  n
≤ θ(a , c ) ⊗ φ(c , b ) →
n∈N
     
→ p∈T (X) φ a , pM (a , c , b ) ⊗ θ pM (a , c , b ), b ≤
  n
≤ n∈N θ(a , c ) ⊗ φ(c , b ) → (φ ◦ θ)(a , b ) .
Thus, we have
  
p∈T (X) !p(x, y, y) ≈ x!K ⊗ !p(x, x, y) ≈ y!K ≤
    
≤ M∈K θ,φ∈ConL (M) a ,b ∈M (θ ◦ω φ)(a , b ) → (φ ◦ θ)(a , b ) =
= Per(K) ,
3.5 Properties of Varieties 169

proving the claim. 




Remark 3.67. For L being the two-element Boolean algebra, Theorem 3.66
gives exactly the above-mentioned Mal’cev characterization of congruence per-
mutability for varieties of algebras [64]. Indeed, we have that θ ◦ φ = φ ◦ θ
iff θ ◦ φ = θ ◦ω φ ⊆ φ ◦ θ iff S(θ ◦ω φ, φ ◦ θ) = 1. Hence, K is congruence
permutable iff Per(K) = 1, i.e. iff there is a ternary term p(x, y, z) such that
!p(x, y, y) ≈ x!K = 1 and !p(x, x, y) ≈ y!K = 1 on account of (3.21).
Suppose L = L, ∧, ∨, ⊗, →, 0, 1 is a complete residuated lattice, where
⊗ is ∧ (L is a complete Heyting algebra). Then, since ⊗ is idempotent, the
definition of Per(θ, φ) simplifies to Per(θ, φ) = S(θ ◦ φ, φ ◦ θ), which is a
generalization of θ ◦ φ ⊆ φ ◦ θ in fuzzy setting. Hence, the degree of equality
θ ◦ φ ≈ φ ◦ θ is equal to Per(θ, φ) ∧ Per(φ, θ). Per(K) can then be interpreted
as a lower estimation of a degree to which θ ◦ φ and φ ◦ θ are equal for all
θ, φ ∈ M, M ∈ K. Clearly, if Per(K) = 1, then θ ◦ φ = φ ◦ θ for all congruences
θ, φ ∈ ConL (M), where M ∈ K. Hence, for ⊗ = ∧ the meaning of Per(K)
corresponds well with the classical permutability.
The situation for non-idempotent ⊗ is not so straightforward. Note that
the ◦ω -composition has been defined to avoid problems with sensitivity of
term functions, see Theorem 2.76 and Remark 2.77. The interpretation of a
truth degree (R1 ◦ωR2 )(a , b ) is interesting for L being a BL-algebra (prelinear
and divisible residuated lattice, see [10, 49]) on the unit interval [0, 1] with
∧ and ∨ being the minimum and the maximum, respectively. In  such a case,
⊗ is a continuous t-norm (Theorem 1.40) and for each a ∈ L, n∈N an is an
idempotent (namely, the greatest idempotent which is less or equal to a ∈ L,
see Lemma 1.52). Thus, the idempotents of L still play an important role,
because (R1 ◦ω R2 )(a , b ) is a supremum of idempotents. For ⊗ being a con-
tinuous Archimedean t-norm (0, 1 are its only idempotents), (3.16) simplifies
as follows:

1 if there is c ∈ U such that R1 (a , c ) = R2 (c , b ) = 1 ,
(R1 ◦ R2 )(a , b ) =
ω
0 otherwise .
That is, R1 ◦ω R2 corresponds with the bivalent relation 1 R1 ◦ 1 R2 . As a
consequence, Per(K) = 1 iff for every M ∈ K and θ, φ ∈ ConL (M) we have
1
θ ◦ 1 φ ⊆ 1 (φ ◦ θ).

Example 3.68. Consider Example 3.55 again. We show how fuzzy equational
logic can be used to estimate the permutability degree of a given variety. For
a ternary term x · (y −1 · z), we can use deduction rules of fuzzy equational
logic to estimate the provability degree of identities x · (y −1 · y) ≈ x and
x · (x−1 · y) ≈ y from G1–G3. We have
1. y −1 · y ≈ 1, c G3, substitution
2. x · (y −1 · y) ≈ x · 1, c by replacement on 1.
3. x · 1 ≈ x, b G2
4. x · (y −1 · y) ≈ x, c ⊗ b by transitivity on 2., 3.
170 3 Fuzzy Equational Logic

and
1. x · x−1 ≈ 1, c G3
2. (x · x−1 ) · y ≈ 1 · y, c by replacement on 1.
3. 1 · y ≈ y, b G2, substitution
4. (x · x−1 ) · y ≈ y, c ⊗ b by transitivity on 2., 3.
5. x · (x−1 · y) ≈ (x · x−1 ) · y, a G1, substitution
6. x · (x−1 · y) ≈ y, a ⊗ c ⊗ b by transitivity on 5., 4.
As a consequence
 of completeness
 of fuzzy equational logic (Theorem 3.31),
we have x · (y −1 · y) ≈ xK ≥ b ⊗ c and x · (x−1 · y) ≈ y K ≥ a ⊗ b ⊗ c.
Hence, K is congruence permutable in degree at least a ⊗ b2 ⊗ c2 . Now, let
L be the standard product algebra on [0, 1]. Then, for the varieties given by
G1–G3 with a = 1, b = 1, c = 0.75 a a = 1, b = 0.75, c = 1, the above
lower estimation a ⊗ b2 ⊗ c2 says that both of the varieties are permutable in
a degree at least 0.5625.

One can proceed in a similar way to obtain further graded properties of


classes of L-algebras (modularity, distributivity, regularity of congruences,
etc.) and to establish graded Mal’cev-like characterizations. An interesting
question is whether it is possible to characterize a large scale of graded prop-
erties of varieties at once by generalizing the well-known result of Taylor [84],
see also related paper [69]. This seems to be an interesting open problem.

3.6 Bibliographical Remarks

Equational logic is perhaps the most known part of universal algebra which
is known outside of universal algebra. Namely, equational logic serves as a
basis for methods of formal specification developed in theoretical computer
science [1, 62]. A very good textbook (focused on computer scientists) on
ordinary equational logic is [97].
Ordinary equational logic and ordinary variety theorem were developed by
Birkhoff [23]. Since Birkhoff’s seminal contribution, both equational logic and
the theory of varieties of universal algebras were investigated in hundreds of
papers. A representative selection can be found in journal Algebra Universalis,
published since 1971. From this point of view, the present chapter deals only
with the basic notions and results of equational logic and theory of varieties in
fuzzy setting. Developing further results in this direction is an open problem
which might bring new insight to both fuzzy logic (especially to its model
theory) and universal algebra.
This chapter is based mainly on [9, 11, 16, 89]. Further results on fuzzy
equational logic can be found in [93].
4
Fuzzy Horn Logic

Implications between identities have been widely studied in universal algebra.


There are numerous results on implicationally defined classes of algebras and
provability of implications from theories in form of implications. A survey
on implications in the context of universal algebra can be found in mono-
graph [97]. In this chapter, we suggest a generalization of the concept of an
implication between identities in fuzzy setting and show that the results on
(classical) Horn logic have their analogies in fuzzy setting. As in Chap. 3, we
develop our logical calculus in Pavelka style.
∗ ∗ ∗
In Sect. 4.1, we introduce the concept of an implication and basic con-
cepts of syntax and semantics. Section 4.2 deals with semantic entailment. In
Sect. 4.3, we elaborate a general completeness theorem and discuss some of
its special cases. In Sect. 4.4 we revisit fuzzy equation logic from the point of
view of fuzzy Horn logic. Sections 4.5, 4.6, 4.7, and 4.8 deal with definability
of classes of algebras with fuzzy equalities by means of implications between
identities.

4.1 Syntax and Semantics


We start by recalling the concept of an implication (between identities) as it
is usually used in universal algebra. Given terms s1 , s1 , . . . , sn , sn , t, t of type
F over variables X, an implication is a formula
s1 ≈ s1 c · · · c sn ≈ sn i t ≈ t , (4.1)
with the intended meaning “if s1 equals s1
and · · · and sn equals then sn ,
t equals t ”. Such a formula is, in fact, an (ordinary) formula of predicate
logic with a language which contains a binary relation symbol ≈, function
symbols f ∈ F , and variables x ∈ X. The identities on the left-hand side of
(4.1) are called premises, t ≈ t is called a consequent. In the literature on

R. Bělohlávek and V. Vychodil: Fuzzy Equational Logic, StudFuzz 186, 171–266 (2005)
www.springerlink.com 
c Springer-Verlag Berlin Heidelberg 2005
172 4 Fuzzy Horn Logic

universal algebra, formulas of this form are alternatively called quasi-identities


or universal positive Horn clauses. It is sometimes desirable to deal with these
formulas as with sentences. That is, instead of (4.1), one considers a formula
(∀x1 ) · · · (∀xk )(s1 ≈ s1 c · · · c sn ≈ sn i t ≈ t ) , (4.2)
such that all variables occurring in terms s1 , s1 , . . . , sn , sn , t, t
are among vari-
ables x1 , . . . , xk . An important generalization of (4.1) is obtained by allowing
infinite premises. Namely, given an index set I and terms si , si (i ∈ I), t, t ,
one can consider a generalized formula
^
si ≈ si i t ≈ t , (4.3)
i∈I
V
where “ ” denotes “generalized conjunction”. The intuitive meaning of (4.3)
is “if si equals si for any i ∈ I, then t equals t ”. Obviously, (4.3) is not a
first-order formula if I is an infinite set – in such a case (4.3) is an “infinite
expression”. On the other hand, if I is finite, (4.1) and (4.3) can be identified.
In particular, if I = ∅, both (4.1) and (4.3) simplify to
t ≈ t . (4.4)
Therefore, identities can be thought of as implications without premises. This
corresponds well with the intuitive meaning of identities and implications
between identities.
Formulas (4.1) are long expressions. To save space, one usually introduces
the following set-theoretical convention for writing premises. Instead of (4.1)
or (4.3), one writes
P i (t ≈ t ) (4.5)
where P is a binary relation on T (X) such that s, s ∈ P iff identity s ≈ s
is one of the premises of (4.3). Thus, P is a binary relation on terms, called
a set of premises, which uniquely describes the conjunction of premises up
to their order. Realizing that the order of identities is not important because
of associativity and commutativity of conjunction, P fully describes premises
of an implication. For instance, implication (4.1) with finite premises can be
written as P i (t ≈ t ) where P = {s1 , s1 , . . . , sn , sn }. This notation is
handy especially if we deal with implications on a general level – we suppose
that P is “some” set of premises, but we do not specify it in more detail.
Therefore, in general, we deal with implications whose sets P of premises
are possibly infinite. Results concerning implications between identities have
shown that it is important which sets P of premises we consider permissi-
ble. For instance, model classes of implications with possibly infinite num-
ber of premises are closed under isomorphic images, subalgebras, and direct
products. Simpler premises of implications naturally lead to further closure
properties of model classes. Notice that two extreme cases of implications be-
tween identities are those of identities (P is empty) and implications with
unrestricted premises (P is arbitrary). Other types of implications result by
imposing additional constraints on sets of premises (e.g., sets of premises are
4.1 Syntax and Semantics 173

required to be finite). The following summary gives an overview of types of


implications which can be found in the literature [80, 97].

(i) Implications: P is arbitrary, model classes (sur-reflective classes) are ab-


stract classes closed under subalgebras, and direct products.
(ii) Finitary implications: P may be infinite, but X is a finite set of variables.
In other words, finitary implications contain only finitely many variables.
Model classes (semivarieties) are abstract classes closed under formations
of subalgebras, direct products, and direct unions.
(iii) Horn clauses: P is finite, model classes (quasivarieties) are abstract
classes closed under formations of subalgebras, direct products, and di-
rect limits. Alternatively, quasivarieties can be characterized as abstract
classes closed under subalgebras and reduced products or as abstract
classes closed under subalgebras, direct product, and ultraproducts.
(iv) Equation implications: P either is a singleton, i.e. P = {s, s } for some
s, s ∈ T (X), or P = ∅. Equation implications are particular Horn clauses.
(v) Identities: P = ∅, model classes (varieties) are closed under formations
of homomorphic images, subalgebras, and direct products.

∗ ∗ ∗
We are now going to introduce a logical calculus for dealing with impli-
cations between identities in a fuzzy setting. We call our logic a fuzzy Horn
logic. We start by introducing basic concepts of syntax and semantics. Fuzzy
Horn logic can be seen as an extension of fuzzy equational logic developed
in Chap. 3. Compared to it, fuzzy Horn logic has a richer language and has
more general formulas called implications. Identities, i.e. formulas of fuzzy
equational logic, can be seen as implications with empty premises. As in case
of fuzzy equational logic, we develop fuzzy Horn logic in Pavelka style.
We start by a language. Suppose ≈, F, σ is a type (page 60) and L is a
complete residuated lattice. A language of fuzzy Horn logic of type ≈, F, σ
for L consists of a binary relation symbol ≈ called a symbol of (fuzzy) equal-
ity, function symbols f ∈ F with their arities σ(f ) ∈ N0 , symbols a of truth
degrees (a ∈ L), (at least denumerable) set X of variables with X V ∩ F = ∅,
symbols of logical connectives i (implication), c (conjunction), (general-
ized conjunction), and auxiliary symbols (parentheses, etc.). If ≈, F, σ and L
are clear from the context, a language of type ≈, F, σ for L is called shortly
a language. Conventions of Remark 3.1 apply also here.
Terms (of type F over X) of fuzzy Horn logic are the usual ones introduced
in Definition 2.73. Recall that T (X) denotes the set of all terms of (a given)
type F over X. Formulas of fuzzy Horn logic are called (P-)implications and
are defined in what follows.
∗ ∗ ∗
Before going to the definition of implications and their interpretation, con-
sider the following comments.
174 4 Fuzzy Horn Logic

First, when evaluating an implication ϕ i ψ in a standard way (call it the


first way), one takes truth degrees !ϕ! and !ψ!, and a truth function → of
implication and defines the truth degree !ϕ i ψ! of ϕ i ψ to be !ϕ! → !ψ!.
There is, however, a second way to look at evaluating implications in bivalent
case. Namely, one takes ϕ and tests whether it is true (!ϕ! equals 1). If not,
!ϕ i ψ! equals 1; if yes, !ϕ i ψ! equals !ψ!. Note that in bivalent case,
both of the ways yield the same truth degree of ϕ i ψ. While the first way
can be directly adopted for fuzzy setting (one just allows !ϕ! and !ψ! to take
also intermediate truth degrees and takes a suitable “fuzzy implication” →),
the second way is not so straightforward. A general way to go is the following:
Pick a threshold a ∈ L such that ϕ is considered sufficiently true iff its truth
degree is at least a. In evaluating ϕ i ψ, one takes !ϕ! and compares it
to a. If !ϕ! does not exceed the threshold (i.e. a ≤ !ϕ!), set !ϕ i ψ! to 1;
otherwise (i.e. a ≤ !ϕ!) set !ϕ i ψ! to !ψ!. In bivalent case, there is only
one nontrivial choice of a, namely, a = 1 which yields the usual interpretation
of implications. In fuzzy setting, however, the choice of a is not unique.
Second, interestingly enough, both of the above ways of interpreting impli-
cations are particular cases of a more general approach via truth stressers: let
 represent a unary logical connective “very true” (i.e.  ϕ reads “ϕ is very
true”) which is interpreted by a truth stresser ∗ defined on L, see Sect. 1.2.
Consider a formula (a i ϕ) i ψ with a a truth constant interpreted by
a ∈ L. If ∗ is the identity on L and a = 1, the truth degree of (a i ϕ) i ψ
is just the truth degree defined in the first way. If ∗ is the globalization (i.e.
a∗ = 1 for a = 1 and a∗ = 0 otherwise), the truth degree of (a i ϕ) i ψ is
just the truth degree defined in the second way with a being the threshold.
Third, as we are interested in implications between identities with possibly
several identities in the premise connected in a conjunctive manner and want
to allow each identity to have its own threshold, we deal with formulas
^ 
 (ai i ϕi ) i ψ (4.6)
i∈I
V
where ϕi ’s and ψ are identities, ai ’s are the thresholds of ϕi ’s, and is a
“generalized conjunction” interpreted by infimum in L. For convenience and
simplicity, we write only
^
ϕi , ai i ψ (4.7)
i∈I
instead of (4.6). That is, we omit  since its placement is fixed, we write
ϕi , ai instead of ai i ϕi , and do not distinguish between constants ai for
truth degrees and truth degrees ai themselves. Note that the second con-
vention follows the identification of weighted formulas ϕ, a introduced in
Sect. 1.4 with ordinary formulas a i ϕ, see e.g. [49, Sect. 3.3]. Doing so, we
deal with formulas which can be thought of as implications
P iψ (4.8)
where P is a fuzzy set of identities ϕi with P (ϕi ) = ai . For our purpose, this
is a convenient and sufficiently general way.
4.1 Syntax and Semantics 175

∗ ∗ ∗
It will sometimes be desirable to have only particular implications which
are “interesting” from some point of view. For instance, we might be interested
in implications P i ψ where P is a finite and/or crisp L-set, etc. Therefore,
we should always consider implications whose sets of premises belong to a
given set P of all “interesting sets of premises”. Another way of looking at
P is that P is a constraint which we supply if we need to restrict ourselves
only to particular formulas – we allow only those P i ψ where P ∈ P. What
properties of P should we require? P should be non-empty in order to have any
interesting formulas at all. Another natural requirement on P is the following:
if P ∈ P, i.e. P is an interesting set of premises, then each particularization
of P should belong to P (should also be interesting). By a particularization of
P we mean any set of premises which is more specific than P . This leads to
a requirement for P to be closed under substitutions because by substituting
terms for variables in formulas we get, in fact, more specific formulas from
the original ones. The following definition formalizes our requirements on P.
Definition 4.1. For P ∈ LT (X)×T (X) , and endomorphism h : T(X) → T(X)
we define an endomorphic image h(P ) ∈ LT (X)×T (X) by

h(P )(t, t ) = h(s) = t P (s, s ) (4.9)
h(s ) = t

for any t, t ∈ T (X). For P ∈ LT (X)×T (X) we define a set var(P ) ⊆ X by



var(P ) = {var(t) ∪ var(t ) | P (t, t ) > 0} . (4.10)
A non-empty subset P ⊆ LT (X)×T (X) is called a proper family of premises
of type F (in variables X) if for every P ∈ P and any endomorphism h on
T(X) we have h(P ) ∈ P. Every P ∈ P is then called an L-set of premises
of type F (in variables X).
Remark 4.2. (1) Given P ∈ P, the degree P (s, s ) ∈ L can be read “the degree
to which identity s ≈ s belongs to P ”. var(P ) can be seen as a set of variables
occurring in identities which belong to P in some nonzero degree.
(2) Returning to the motivation for our need to define P, the closedness
under endomorphic images corresponds with our requirement of closedness
under particularizations because each endomorphism h can be seen as a “sub-
stitution”: each variable x ∈ X is replaced by term h(x).
(3) If P ∈ LT (X)×T (X) is a crisp L-relation then the notion of an endo-
morphic image as defined above coincides with the classical one. That is,
h(P ) = {h(t), h(t ) | t, t ∈ P } .
As a trivial consequence, h(∅) = ∅.
Example 4.3. The following are examples of proper families of premises.
(1) P = LT (X)×T (X) is a proper family of premises (obvious). This fam-
ily does not represent any constraints on premises. Implications with such
176 4 Fuzzy Horn Logic

premises will represent the most general type of formulas used in our investi-
gation.  
(2) Let P = P ∈ LT (X)×T (X) | var(P ) is finite . Then P is a proper fam-
ily of premises. Indeed, every endomorphism h on T(X) is determined by
its restriction on the set of generators X. Since  var(P ) is finite, we can put
Y = var(P ) ∪ Z, where Z is defined by Z = {var(h(x)) | x ∈ var(P )}. Now
we can define g : Y → T (Y ) by letting g(y) = h(y) for each y ∈ Y . Thus,
g  (t) = h(t) for all t ∈ T (Y ). For t, t ∈ T (X) such that t, t ∈ T (Y ) we can
see that h(s) = t, h(s ) = t implies s, s ∈ T (var(P )), thus P (s, s ) = 0, i.e.
h(P )(t, t ) = 0. That is, var(h(P )) = var(g  (P )). Moreover, Y is finite since
both sets var(P ) and Z are finite. Thus, var(h(P )) ⊆ Y is finite as well. P
of this form will be called a family of all finitary premises. Note that if X is
finite, then P = LT (X)×T 
(X)
. 
(3) A family P = P ∈ LT (X)×T (X) | P is finite is a special subfamily of
that of (2). It is easy to observe that endomorphic image of every finite L-
relation is finite. Trivially, var(P ) is finite. P of this form is called a proper
family of all finite premises.
(4) Families defined in (1)–(3) have their “crisp variants”. Clearly, if P is
crisp then h(P ) is crisp. Hence, the following families
P = {P ∈ LT (X)×T (X) | P is crisp},
P = {P ∈ LT (X)×T (X) | var(P ) is finite and P is crisp},
P = {P ∈ LT (X)×T (X) | P is finite and crisp},
are proper families of premises. Such families will be used to define implica-
tions with crisp premises.
(5) For each a ∈ L, let Pa denote a subset of LT (X)×T (X) such that
Pa = {P | for any t, t ∈ T (X) : P (t, t ) > 0 implies P (t, t ) ≥ a} .
It is easy to see that Pa is a proper family of premises since for every P ∈ P,
h(P )(t, t ) either is zero or h(P )(t, t ) ≥ a. In fact, we have P0 = LT (X)×T (X) ,
and P1 denotes the proper family of all crisp premises.
(6) P = {∅} is a proper family of premises trivially. As we will see later
on, implications with empty premises can be identified with identities.

Now we can define the general notion of a P-implication. We define also


(L-)weighted implications, see Definition 1.103 and Remark 1.104, and also
Definition 3.14 and Remark 3.15.

Definition 4.4. Let L be a complete residuated lattice, X be a set of variables,


F be a type, P be a proper family of premises of type F in variables X.
A P -implication is an expression of the form
^

s ≈ s , P (s, s ) i (t ≈ t ) , (4.11)
P (s,s )>0

where P ∈ P and t, t ∈ T (X). For a P-implication ϕ and a truth degree a ∈ L,
a couple ϕ, a is called an L-weighted P-implication.
4.1 Syntax and Semantics 177

Remark 4.5. (1) P-implications are in general “infinite expressions”. Following


the previous discussion, we consider (4.11) a shorthand for
^  
 
P (s, s ) i (s ≈ s ) i (t ≈ t ) , (4.12)
P (s,s )>0

a formula in language containing a unary connective V , symbols a for truth


degrees a ∈ L, and a “generalized conjunction” . Equation (4.12) says “if
it is very true that all identities s ≈ s from P are true then t ≈ t is true”,
which is the intended meaning of (4.11).
(2) As in Remark 1.104 (3), we can identify L-sets of P-implications with
ordinary sets of weighted P-implications. Recall that for an L-set Σ of P-
implications we put TΣ = {ϕ, a | Σ(ϕ) = a}.
(3) For simplicity, P-implication (4.11) will be denoted by P i (t ≈ t ). If
P is finite, P i (t ≈ t ) will be occasionally denoted by
t1 ≈ t1 , a1 c t2 ≈ t2 , a2 c · · · c tn ≈ tn , an i t ≈ t ,
where P (ti , ti ) = ai for i = 1, . . . , n, and Supp(P ) ⊆ {t1 , t1 , . . . , tn , tn }.
In the sequel, we will be interested in general proper families of premises
and families which result from given families by restrictions. For instance, we
are going to study Pavelka-style completeness of fuzzy Horn logic which uses
formulas with finitely many premises. That is, given a family P, we will be
interested in those P-implications where P ∈ P is a finite L-relation. For any
proper family P of premises we thus define the following subsets of P:
PFin = {P ∈ P | P is finite} , (4.13)
Pω = {P ∈ P | var(P ) is finite} , (4.14)
Pκ = {P ∈ P | κ > | var(P )|} , (4.15)
where κ is any infinite cardinal.
Remark 4.6. (1) PFin is a subset of P which consists of P ∈ P such that P
is finite (i.e., Supp(P ) is a finite set). Pω is a subset of P which consists of
P ∈ P such that the set of all variables occurring in identities which belong to
P in some nonzero degree is finite. Finally, Pκ is a subset of P which consists
of P ∈ P such that the set of variables occurring in P ∈ P is strictly lesser
than κ. Notice that Pω is a particular case of Pκ for κ = ω.
(2) Observe that PFin and Pκ are closed under endomorphic images. Indeed,
since each P ∈ PFin is finite, we have that h(P ) is finite, i.e. h(P ) ∈ PFin . In
case of Pκ , we have | var(h(P ))| ≤ | var(h(var(P )))| < κ for any P ∈ Pκ (notice
that var(h(x)) is finite) from which the claim immediately follows.
(3) PFin and Pκ can be empty. Thus, PFin and Pκ need not be proper
families of premises. For instance, consider a type F = {c, d, f }, where c, d are
constants, and f is a unary function symbol. Put t0 = c, t0 = d, ti = f (ti−1 ),
ti = f (ti−1 ) for each i ∈ N. Define P ∈ LT (X)×T (X) so that P (ti , ti ) = 1
(i ∈ N0 ), and P (· · ·) = 0 otherwise. Obviously, P = {P } is a proper family of
premises because for any endomorphism h on T(X), h(c) = c and h(d) = d
178 4 Fuzzy Horn Logic

which further gives h(ti ) = ti and h(ti ) = ti (i ∈ N0 ). Clearly, PFin = ∅, i.e.
PFin is not a proper family of premises.

According to Remark 4.6, PFin and Pκ need not be proper families of


premises because it can happen that PFin = ∅ and Pκ = ∅. On the other hand,
in the sequel we will be interested only in cases where PFin and Pκ are non-
empty in order to have any PFin -implications (Pκ -implications). Thus, from
this moment on, we assume the following convention: if we consider PFin or
Pκ , we automatically assume that PFin and Pκ are proper families of premises,
i.e. we assume that P is a proper family of premises such that PFin or Pκ
are non-empty. The following definition introduces a terminology that will be
used in the subsequent development.

Definition 4.7. PFin defined by (4.13) is called a Horn restriction of P.


Pω defined by (4.14) is called a finitary restriction of P. PFin -implications
will be called P-Horn clauses. Pω -implications will be called P-finitary im-
plications.

Remark 4.8. (1) Note that PFin ⊆ Pω ⊆ Pκ ⊆ P for any infinite κ. Hence, each
P-Horn clause is a P-finitary implication, and each P-finitary implication is a
P-implication.
(2) If P ⊆ PFin , we have PFin = Pω = Pκ = P for any infinite κ.
(3) If P is clear from the context, a P-implication (P-Horn clause) can be
called a finitary implication (Horn clause).

∗ ∗ ∗
We are now going to introduce basic concepts of semantics of fuzzy Horn
logic. Structures of truth degrees of fuzzy Horn logic will be complete resid-
uated lattices L with truth stressers ∗ . Structures for interpretation of terms
and formulas of fuzzy Horn logic are, as in case of fuzzy equational logic,
algebras with fuzzy equalities. The concepts of a valuation (i.e. a mapping
v : X → M ) and value !t!M,v of term t in an L-algebra M under valuation v
are defined as usual (see Definition 3.4).
Recall (Definition 3.6) that given terms t, t ∈ T (X), a truth degree
!t ≈ t !M,v of an identity t ≈ t in M under a valuation v : X → M is de-
fined by !t ≈ t !M,v = !t!M,v ≈M !t !M,v . A truth degree !P i (t ≈ t )!M,v
of implication (4.11) is defined in a straightforward way, taking into account
that (4.11) is a shorthand for (4.12) and that connective  is interpreted by
a truth-stresser ∗ on L.

Definition 4.9. Let P i (t ≈ t ) be a P-implication and let ∗ be a truth


stresser for a complete residuated lattice L. For an L-algebra M and a val-
uation v : X → M , we define a truth degree P i (t ≈ t )M,v of
P i (t ≈ t ) in M under v with respect to ∗ by
!P i (t ≈ t )!M,v = !P !M,v → !t ≈ t !M,v , (4.16)
4.1 Syntax and Semantics 179

where
  ∗
!P !M,v = s,s ∈T (X) P (s, s ) → !s ≈ s !M,v . (4.17)
A truth degree P i (t ≈ t )M of P i (t ≈ t ) in M with respect to ∗
 

is defined by

!P i (t ≈ t )!M = v:X→M !P i (t ≈ t )!M,v . (4.18)
For a weighted P-implication P i (t ≈ t ), a we define a truth degree
P i (t ≈ t ), aM,v of P i (t ≈ t ), a in M under v w.r.t. ∗ by
a → !P i (t ≈ t )!M,v . (4.19)

Therefore, we consider ∗ as a parameter controlling the interpretation of


(4.11). For the boundary truth stressers of Example 1.14, we have the follow-
ing. First, for ∗ being the identity we get
!P i (t ≈ t )!M,v =
   

= s,s ∈T (X) P (s, s ) → !s ≈ s !M,v → !t ≈ t !M,v . (4.20)

Second, for being globalization we get

 !t ≈ t !M,v if P (s, s ) ≤ !s ≈ s !M,v

!P i (t ≈ t )!M,v = for all s, s ∈ T (X) , (4.21)

1 otherwise .
Denoting by a and b the truth degree of P i (t ≈ t ) as defined by (4.20) and
(4.21), respectively, it is easily seen that a ≤ b. Equations (4.20) and (4.21) are
thus the boundary cases of (4.16). As we will see, both types of interpretations
are reasonable (cf. also discussion in the beginning of this section). Note that
in the bivalent case, both (4.20) and (4.21) coincide.

Remark 4.10. (1) We will use only one structure of truth degrees and one truth
stresser at a time, so there is no danger of confusion if the degree is denoted

simply by !P i (t ≈ t )!M,v . Sometimes, we will use !P i (t ≈ t )!L M,v to
point out L with ∗ explicitly.
(2) It is easily seen that (4.19) is equal to !P !M,v → (a → !t ≈ t !M,v ).
This corresponds well to the intuitive meaning of a weighted implication and
also justifies a possible notation
t1 ≈ t1 , a1 c t2 ≈ t2 , a2 c · · · c tn ≈ tn , an i t ≈ t , a
for a weighted Horn clause
t1 ≈ t1 , a1 c t2 ≈ t2 , a2 c · · · c tn ≈ tn , an i t ≈ t , a .
(3) Evidently, !P i (t ≈ t )!M,v ≥ b iff !P i (t ≈ t ), b !M,v = 1.
(4) For P = ∅, we have
  
∗  ∗
!P !M,v = s,s ∈T (X) 0 → !s ≈ s !M,v = s,s ∈T (X) 1 =1.

Therefore, for any truth stresser and arbitrary t, t ∈ T (X),
180 4 Fuzzy Horn Logic

!∅ i (t ≈ t )!M,v = !∅!M,v → !t ≈ t !M,v = !t ≈ t !M,v .


Thus, from the point of view of semantics, identities can be identified with
P-implications with empty premises, and the truth degree !∅ i (t ≈ t )!M,v
does not depend on ∗ .

∗ ∗ ∗
Fuzzy Horn logic as developed so far can be seen as a certain frag-
ment of predicate fuzzy logic as introduced in Example 1.96 (2), but with
infinite conjunction. As in Chap. 3, we now present fuzzy Horn logic in
a setting of abstract logic. That is, we present a particular abstract logic
L = Fml , L, S, A, R , see Definition 1.105, which gives us then automatically
further concepts we need (theory, semantic entailment, provability, etc.), see
Sect. 1.4.
For a set Fml of formulas we take the set of all P-implications, where P is
a given proper family of premises (of type F in variables X), i.e.
Fml = {P i (t ≈ t ) | P ∈ P, t, t ∈ T (X)}.
For L we take an arbitrary complete residuated lattice equipped with a
truth stresser ∗ . That is, as required by Definition 1.105, L is a complete
lattice, but we require an additional structure on it.
For S we take an L-semantics S for Fml defined by
S = {E ∈ LFml | for some L-algebra M : E(ϕ) = !ϕ!M (4.22)
for each P-implication ϕ ∈ Fml } ,
cf. (1.92).
For A, we take an empty L-set of formulas, i.e. A(ϕ) = 0 for each P-
implication ϕ.
For R we take a set of deduction rules defined later on.
Up to R which we specify later, we have defined a particular abstract logic.
This gives us automatically all the notions of abstract logic defined in Sect. 1.4.
However, as in Chap. 3, in order for these concepts to fit better into our
desired framework, we extend them. We want to deal with L-algebras instead
of evaluations and with classes of L-algebras instead of the corresponding sets
of evaluations.
Using mappings Md and Th defined by (1.93) and (1.94), we can introduce
mappings Mod and Impl by
Mod(Σ) = {M | EM ∈ Md(Σ)} and Impl(K) = Th({EM | M ∈ K})
where Σ ∈ LFml is a theory (L-set of P-implications) and K is a class of
L-algebras (of type F ).
Again, we can then say that an L-algebra M is a model of a theory Σ if
M ∈ Mod(Σ). It is easy to see that M is a model of Σ ∈ LFml iff Σ(ϕ) ≤
!ϕ!M for each P-implication ϕ ∈ Fml . Then, Mod(Σ) is a class of L-algebras
and Impl(K) is an L-set of P-implications for which we have
4.1 Syntax and Semantics 181

Mod(Σ) = {M | M is a model of Σ} , (4.23)



(Impl(K))(ϕ) = {!ϕ!M | M ∈ K} (4.24)
for each P-implication ϕ.

Remark 4.11. (1) Occasionally, if we want to make L∗ and P explicit, we use


∗ ∗
also ModL (Σ), and also ImplP(K) and ImplL P (K).
(2) We also write Implκ(K), Horn(K) instead of ImplPκ(K), ImplPFin(K),
respectively. If K = {M}, we write Impl(M) instead of Impl({M}). Since
PFin ⊆ Pω ⊆ P, Horn(K) is a restriction of ImplP(K) on P-Horn clauses;
Implω(K) is a restriction of ImplP(K) on P-finitary implications.
(3) ImplP(K) can be seen as an L-set of P-implications such that a P-impli-
cation P i (t ≈ t ) belongs to ImplP(K) in the degree to which P i (t ≈ t )
is true in K.
(4) If P and Q are proper families of premises such that Q ⊆ P, we some-
times look at ImplQ(K) as it were an L-set of P-implications, where

 (ImplP(K))(P i (t ≈ t )) if P ∈ Q ,
(ImplQ(K))(P i (t ≈ t )) =
0 otherwise ,
for any P-implication P i (t ≈ t ). For instance, Horn(K) can be seen as an
L-set of P-implications such that (Horn(K))(P i (t ≈ t )) = 0 if P is infinite.
(5) For a proper family P = {∅} and for any class K of L-algebras of a given

  
type we get (ImplL P (K))(∅ i (t ≈ t )) = (Id(K))(t ≈ t ) for any t, t ∈ T (X).
This is an immediate consequence of Remark 4.10 (4).

Like Md and Th, Mod and Impl form a Galois connection and satisfy
thus (1.95)–(1.100) with sets K’s of evaluations replaced by classes K’s of
L-algebras. As in Remark 1.107, we can define a degree !P i (t ≈ t )!Σ to
which a P-implication P i (t ≈ t ) semantically follows from an L-set Σ of
P-implications by !P i (t ≈ t )!Σ = (Impl(Mod(Σ)))(P i (t ≈ t )).
Notice that all the concepts of fuzzy Horn logic are uniquely determined
by the choice of a type ≈, F, σ , a complete residuated lattice L with truth
stresser ∗ , and a proper family P of premises. The following definition sum-
marizes semantical concepts of fuzzy Horn logic given by a particular choice
of ≈, F, σ , L, ∗ , and P.

Definition 4.12. A theory is an L-set Σ of P-implications. An L-algebra M


is a model of a theory Σ if M ∈ Mod(Σ). A degree P i (t ≈ t )Σ to
which P i (t ≈ t ) semantically follows from a theory Σ is defined by
!P i (t ≈ t )!Σ = (Impl(Mod(Σ)))(P i (t ≈ t )).
For a class K of L-algebras, a truth degree P i (t ≈ t )K of
P i (t ≈ t ) in K is defined by
!P i (t ≈ t )!K = (Impl(K))(P i (t ≈ t )).
182 4 Fuzzy Horn Logic

Remark 4.13. (1) For reader’s convenience, let us rewrite the concepts of De-
finition 4.12:
M is a model of Σ iff for each P i (t ≈ t ) :
Σ(P i (t ≈ t )) ≤ !P i (t ≈ t )!M ,

!P i (t ≈ t )!K = M ∈ K !P i (t ≈ t )!M =
 
= M ∈ K v:X→M !P i (t ≈ t )!M,v ,

!P i (t ≈ t )!Σ = !P i (t ≈ t )!Mod(Σ) =

= M ∈ Mod(Σ) !P i (t ≈ t )!M =
 
= M ∈ Mod(Σ) v:X→M !P i (t ≈ t )!M,v .


(2) If we need to make L and explicit, we also use !P i (t ≈ t )!L
Σ
instead of !P i (t ≈ t )!Σ .

Definition 4.14. For a theory Σ ∈ LFml , Mod(Σ) is called an P -implica-


tional class of Σ. A class K of L-algebras is called a P -implicational
class if K = Mod(Σ) for some theory Σ. K is called a P -finitary implica-
tional class (PP-Horn class) if K = Mod(Σ) for some L-set Σ of P-finitary
implications (P-Horn clauses).
For a class K of L-algebras, Impl(K) is called an L∗ -implicational P -
theory of K. An L-set Σ of P-implications is called an L∗ -implicational
P -theory if Σ = Impl(K) for some class K of L-algebras.

Remark 4.15. (1) Since PFin ⊆ Pω ⊆ P, every P-Horn class is a P-finitary class
and every P-finitary class is a P-implicational class.
(2) Remark 4.10 (4) yields that for P = {∅} and an L-set Σ of P-im-
plications, Mod(Σ) is the equational class of Σ  (see Definition 3.12) where
Σ  is the L-set of identities such that Σ  (t ≈ t ) = Σ(∅ i (t ≈ t )) for any
t, t ∈ T (X).

Utilizing the notions of abstract logic, we can now define concepts related
to provability of a fuzzy Horn logic given by ≈, F, σ , L, ∗ , and P. How-
ever, since provability for general systems P of proper premises would require
infinitary deduction rules, we restrict ourselves to P-Horn caluses.

Definition 4.16. Let R be a set of deduction rules for P-Horn clauses, Γ ∈


LFml be a theory (L-set of P-Horn clauses), let P i (t ≈ t ) ∈ Fml be a P-
Horn clause and a ∈ L. An (L-weighted ) proof of P i (t ≈ t ), a from
Γ using R is a sequence ϕ1 , a1 , . . . , ϕn , an of L-weighted P-Horn clauses
ϕi , ai such that ϕn is P i (t ≈ t ), an = a, and for each i = 1, . . . , n, we
have ai = Γ (ϕi ) or ϕi , ai follows from some ϕj , aj ’s, j < i, by some
deduction rule R ∈ R. The number n is called a length of the proof. In such
a case, we write Γ R P i (t ≈ t ), a and call P i (t ≈ t ), a provable
from Γ using R. If Γ R P i (t ≈ t ), a , P i (t ≈ t ) is called provable in
4.1 Syntax and Semantics 183

degree (at least) a from Γ using R. A degree of provability of P i (t ≈ t )


from Γ using R, denoted by |P i (t ≈ t )|R
Γ , is defined by

|P i (t ≈ t )|R
Γ = {a | Γ R P i (t ≈ t ), a }.

Remark 4.17. The conventions introduced in Remark 1.102 apply.

∗ ∗ ∗

Example 4.18. (1) Consider a language with a single binary function symbol
◦ and take P = LT (X)×T (X) . A P-Horn clause x ◦ y ≈ x ◦ z, a i (y ≈ z) can
be seen as a type of a graded cancellation rule, saying “if x ◦ y equals x ◦ z
in degree (at least) a ∈ L, then y equals z”. Note that in the ordinary case,
cancellation cannot be expressed by identities (classes of cancellative algebras
are not closed under the formation of homomorphic images).
(2) The apparatus of weighted implications can be seen as a tool in formal
specification in presence of vagueness. In particular, applications of weighted
implications lie mainly in the field of so-called humanistic systems, where
the description of a system behavior is influenced by human judgment or
perceptions, and is therefore inherently vague. In the following, we present a
way to describe approximate knowledge about a simple function-based system.
We deal with the problem of human perception of colors and the related
problem of color mixture. Needles to say, the problems in question are hardly
graspable by bivalent logic (crisp structures) since the notions of “color simi-
larity” and “color indistinguishability” that naturally appear in the problem
domain are vague. The color perception itself is a complex neuro-chemical
process with a psychological feedback. Denote the set of all colors by M .
Equip M with an L-equality relation ≈M the meaning of which is to repre-
sent similarity of colors from M . Note that ≈M is a nontrivial L-relation for
which all properties of an L-equality seem to be justified.
The (additive) mixture of colors can be thought of as an operation on
M . Thus, suppose we have a language F = {f }, where f is a binary function
symbol and a term f (t, t ) of type F represents a color resulting by the mixture
of colors represented by terms t, t . It is a well-known fact [39] that assuming
sufficiently high light intensities, if x is indistinguishable from x , and y is
indistinguishable from y  , then f (x, y) is indistinguishable from f (x , y  ). This
rule immediately translates into a compatibility
 condition
 for f M .
To sum up, an L-algebra M = M, ≈ , f M M
of type F seems to be a
suitable semantical structure enabling us to study color mixture. Weighted
implications can be used to define additional constraints on our perception of
color mixture. For instance, the weighted P-Horn clause
x ≈ x , a c f (x, y) ≈ f (x , y  ), b i (y ≈ y  ), c (4.25)
can be read as: “if colors x, x are similar in degree a and if mixtures f (x, y),
f (x , y  ) are similar in degree b, then y, y  are similar (at least) in degree c”.
184 4 Fuzzy Horn Logic

∗ ∗ ∗
When describing the abstract logic for fuzzy Horn logic, we left a set R
of deduction rules unspecified. Deduction rules of fuzzy Horn logic will be
described now.
Note in advance that our deduction rules do not conform to the notion of
a deduction rule as introduced in Sect. 1.4. Recall that according to Pavelka,
a deduction rule is a pair R = Rsyn , Rsem , where Rsyn : Fml n → Fml is
a partial mapping on the set of formulas and Rsem : Ln → L is a mapping
on the set of truth degrees. Weighted formula ϕ, a inferred by R is of the
form ϕ = Rsyn (ϕ1 , . . . , ϕn ) and a = Rsem (a1 , . . . , an ), meaning that one infers
validity of ϕ in degree (at least) a ∈ L given formulas ϕi valid in degree (at
least) ai (i = 1, . . . , n). Contrary to that, some of our rules compute a ∈ L
in the inferred weighted formula ϕ, a not only from ai ’s but also from truth
degrees (represented by constants) which are present in ϕi ’s. This is, however,
only for the sake of convenience. Namely, as we will see in Remark 4.23,
all of our deduction rules are in fact derived rules in a suitably extended
Pavelka-style first-order fuzzy logic with ordinary deduction rules of the form
R = Rsyn , Rsem .
Therefore, our deduction rules are partial mappings
n
R : (Fml × L) → Fml × L . (4.26)
A set R of deduction rules will also be called a deductive system.
Remark 4.19. (1) Instead of R(ϕ1 , a1 , . . . , ϕn , an ) = ϕ, a , we again use
the common notation
ϕ1 , a1 , . . . , ϕn , an
. (4.27)
ϕ, a
Described verbally, the deduction rule (4.27) should be read as “From ϕ1 in
degree a1 , and · · · and ϕn in degree an infer ϕ in degree a”.
(2) An axiom can be thought of as nullary deduction rule, i.e. a mapping
A : {∅} → Fml × L. Hence, an axiom is a weighted formula from Fml ×L.
In accordance with Remark 4.5, we can denote an axiom P i (t ≈ t ), a
also by P i t ≈ t , a . All axioms introduced below are considered as nullary
deduction rules, i.e. unlike the abstract approach described in Sect. 1.4, we
do not introduce an additional set A of (logical) axioms.
In order to introduce our basic set R of deduction rules for fuzzy Horn
logic, we define an application of a substitution to L-sets of premises as follows.
Definition 4.20. For P ∈ P, let P (x/r) ∈ LT (X)×T (X) denote a binary L-
relation defined by
  
P (x/r) (t, t ) = s(x/r) = t P (s, s ) (4.28)
s (x/r) = t

for all terms t, t ∈ T (X).
4.1 Syntax and Semantics 185

Remark 4.21. Evidently, every substitution (x/r) can be expressed as an en-


domorphism h on T(X), which is a homomorphic extension of a mapping
g : X → T (X), where g(x) = r, and g(y) = y for y ∈ X, y = x. Thus, we have
t(x/r) = h(t) for each term t ∈ T (X), i.e. P (x/r) = h(P ). This fact can be
easily proved by structural induction. Thus, for a proper family of premises
P and any (x/r), we have P (x/r) ∈ P for every P ∈ P. As a consequence, if
P i (· · ·) is a P-Horn clause then so is P (x/r) i (· · ·).

In what follows, we use a system R of deduction rules (Ref)–(Mon) intro-


duced below. Therefore, if there is no danger of confusion, we omit “R” from
R
“provable from Γ using R”, “|P i (t ≈ t )|Γ ”, etc. and write simply “prov-

able from Γ ”, “|P i (t ≈ t )|Γ ”, etc. Until otherwise mentioned, R always
stands for the set of deduction rules which consists of (Ref)–(Mon).
The first group of rules are the rules of congruence:

(Ref) : ,
P i (t ≈ t), 1

P i (t ≈ t ), a
(Sym) : ,
P i (t ≈ t), a

P i (t ≈ t ), a , P i (t ≈ t ), b
(Tra) : ,
P i (t ≈ t ), a ⊗ b

P i (t ≈ t ), a
(Rep) : ,
P i (s ≈ s ), a
where P ∈ PFin , a, b ∈ L, t, t , t , s, s ∈ T (X), and s contains t as a subterm
and s results from s by substitution of one occurrence of t in s by t .
The second group are the rules of extensivity, substitution, and monotony:

(Ext) : ,
P i (t ≈ t ), P (t, t )

P i (t ≈ t ), a
(Sub) :    ,
P (x/r) i t(x/r) ≈ t (x/r) , a

{Q i (ti ≈ ti ), ai ; i = 1, . . . , n} , P i (t ≈ t ), b


(Mon) :  n ∗ ,
Q i (t ≈ t ), b ⊗ i=1 (P (ti , ti ) → ai )
where P, Q ∈ PFin such that Supp(P ) = {t1 , t1 , . . . ,tn , tn }, t, t , r ∈ T (X),
x ∈ X, and a1 , . . . , an , a, b ∈ L.

Remark 4.22. (1) Deduction rules (Ref)–(Rep) are generalizations of the rules
(ERef)–(ERep) presented in Sect. 3.1. Rules (Ref) and (Ext) are nullary, i.e.
they can be thought of as axioms. The rule of extensivity (Ext) expresses
the relationship between sets of weighted premises and provability degrees.
186 4 Fuzzy Horn Logic

In other words, using (Ext), every P i (t ≈ t ) is provable in degree at least


P (t, t ). The rule of substitution is also similar to (ESub) but for technical
reasons we use P (x/r) defined by (4.28).
(2) On the verbal level, the rule of monotony (Mon) can be read: “if P
implies t ≈ t and each premise ti ≈ ti from P is implied by Q, then also t ≈ t
is implied by Q”. A finer reading of (Mon) is: “Q implies t ≈ t (at least) in
degree to which P implies t ≈ t and to which it is very true that each ti ≈ ti

is implied by Q at least in degreen to which ti ≈ ti ∗belongs to P ”.
(3) The truth degree b ⊗ i=1 (P (ti , ti ) → ai ) as used in (Mon) is com-
puted using truth degrees of input formulas together with P (ti , ti )’s. In other
words, the degree P (ti , ti ) to which the identity ti ≈ ti belongs to premises
of P i (t ≈ t ) has an influence on the resulting truth degree. Consider the
following example. Take a residuated lattice L = {0, a, 1}, ∨, ∧, ⊗, →, 0, 1 ,
where 0 < a < 1, and a ⊗ a = 0, the rest is determined uniquely. Furthermore,
we equip L by globalization which is Horn truth stresser on L. Now, (Mon)
gives
r ≈ r , 1 i s ≈ s , a , s ≈ s , a i t ≈ t , 1
.
r ≈ r , 1 i t ≈ t , 1
On the other hand, increasing the “threshold” (given by the degree a in the
formula s ≈ s , a i t ≈ t , 1 ) from a to 1, we get
r ≈ r , 1 i s ≈ s , a , s ≈ s , 1 i t ≈ t , 1
.
r ≈ r , 1 i t ≈ t , 0
(4) One way to have deduction rules in Pavelka-style (separate syntactic
and semantic parts) is to split the (Mon) rule into separate rules (MonP ) for
every set of premises P ∈ P. In this case, it would be possible to distinguish
the syntactic and semantic part of a rule by two independent mappings as
usual. For the above example, we would then use two different deduction
rules (MonP1 ), (MonP2 ).

Remark 4.23. We are going to show that (Ref)–(Mon) are derived rules in a
natural Pavelka-style first-order fuzzy logic. To that purpose, we assume that
some (weighted) formulas are provable in the Pavelka-style logic we work with
(we mention them in the course of our demonstration). In particular, we as-
sume that we have formulas guaranteeing reflexivity, symmetry, transitivity,
and compatibility of ≈, and formulas guaranteeing the required properties of
logical connectives as axioms. As we will work with truth constants (for every
a ∈ L we consider a truth constant a), we need to assume appropriate “book-
keeping axioms” for the constants [49]. Namely, for  we assume   a e a∗ .
We use the following deduction rules: modus ponens (MP; from ϕ, a and
ϕ i ψ, b infer ψ, a ⊗ b ), logical constant introduction [71] (from ϕ, a in-
fer a i ϕ, 1 ), and truth confirmation [49, 51] (from ϕ, 1 infer  ϕ, 1 ).
For convenience, we write  ϕ instead of  ϕ, 1 .
Let Supp(P ) = {t1 , t1 , . . . , tn , tn } and put P (ti , ti ) = pi (i = 1, . . . , n).
4.1 Syntax and Semantics 187

(Ext): Let P i (t ≈ t ). We show  p i (P i (t ≈ t )) for p = P (t, t ).


Observe that for P (t, t ) = 0, the claim is trivial. Thus, suppose there is some
j ∈ {1, . . . , n} with t = tj and t = tj . Therefore, p = P (tj , tj ) = pj . We have
Vn Vn
  i=1 (pi i (ti ≈ ti )) i i=1 (pi i (ti ≈ ti )),
[instance of  ϕ i ϕ,]
Vn
 i=1 (pi i (ti ≈ ti )) i (pj i (tj ≈ tj )),
[using  (ϕ c ψ) i ϕ,]
Vn
 i=1 (pi i (ti ≈ ti )) i (pj i (tj ≈ tj )),
[by  (ϕ i ψ) i ((ψ i χ) i (ϕ i χ)), MP,]
 Vn 
 pj i  i=1 (pi i (ti ≈ ti )) i (tj ≈ tj ) ,
[by  (ϕ i (ψ i χ)) i (ψ i (ϕ i χ)), MP,]
showing that  p i (P i (t ≈ t )).
(Mon): We show that if   b iN(P i (t ≈ t )) and 
  ai i (Q i (ti ≈ ti ))
n
for each i = 1, . . . , n then  b o i=1 (pi i ai ) i (Q i (t ≈ t )). First,
 Q i (ai i (ti ≈ ti )),
[by  (ϕ i (ψ i χ)) i (ψ i (ϕ i χ)), MP, i = 1, . . . , n,]
Vn
Qi i=1 (ai i (ti ≈ ti )),
[using  (ϕ i ψ) i ((ϕ i χ) i (ϕ i (ψ c χ))), MP,]
 Vn 
  Q i i=1 (ai i (ti ≈ ti )) ,
[truth confirmation,]
Vn
 Q i  i=1 (ai i (ti ≈ ti )),
[by  (ϕ i ψ) i ( ϕ i  ψ), MP,]
 Q i  Q,
[instance of  ϕ i   ϕ (Q is of the form  ϕ),]
Vn
Qi i=1 (ai i (ti ≈ ti )),
[by  (ϕ i ψ) i ((ψ i χ) i (ϕ i χ)), MP.]

Now  (ψ i χ) i ((ϕ o ψ) i (ϕ o χ)) and  ( ϕ o  ψ) i (ϕ o ψ) give


Nn   Nn Vn 

 i=1 (pi i ai ) o Q i i=1 (pi i ai ) o  i=1 (ai i (ti ≈ ti )) ,
Nn Vn 
 i=1 (pi i ai ) o  i=1 (ai i(t ≈ t )) i 
Ni n i Vn 
i i=1 (pi i ai ) o i=1 (ai i (ti ≈ ti )) .

By  (ϕ o (ψ c χ)) i ((ϕ o ψ) c (ϕ o χ)),


 (ϕ o ψ) i ϕ,  ((ϕ i ψ) o (ψ i χ)) i (ϕ i χ), and using truth confir-
mation with isotony of o and  (ϕ i ψ) i ( ϕ i  ψ):
Nn Vn 

 i=1 (pi i ai ) o i=1 (ai i (ti ≈ tiV)) i 
i  i=1 ((pi i ai ) o (ai i (ti ≈ ti ))) ,
n
Vn  V n 
  i=1 ((pi i ai ) o (ai i (ti ≈ ti ))) i  i=1 (pi i (ti ≈ ti )) .
Using transitivity applied on previous formulas, isotony, and MP:
188 4 Fuzzy Horn Logic
Nn  Vn 
 (pi i ai ) o Q i  i=1 (pi i (ti ≈ ti )) ,
i=1
 Nn   V n 
 b o i=1 (pi i ai ) o Q i b o  i=1 (pi i (ti ≈ ti )) .
Taking into account  b i (P i (t ≈ t )), we have
 Vn 
 b o  i=1 (pi i (ti ≈ ti )) i (t ≈ t ),
[by  (ϕ i (ψ i χ)) i ((ϕ o ψ) i χ), MP,]
 Nn 
 b o i=1 (pi i ai ) o Q i (t ≈ t ),
[using transitivity, MP,]
 Nn 
 b o i=1 (pi i ai ) i (Q i (t ≈ t )),
[by  ((ϕ o ψ) i χ) i (ϕ i (ψ i χ)), MP,]
which is the desired weighted formula. One can proceed similarly for the rest
of the deduction rules. Note that the rule of truth confirmation and axioms
ϕ i ϕ ,
(ϕ i ψ) i ( ϕ i  ψ) ,
ϕ i ϕ ,
( ϕ o  ψ) i (ϕ o ψ)
being used in the previous demonstration naturally correspond to properties
of implicational truth stressers, see (1.11), (1.12), (1.13), (1.17), and (1.57).

∗ ∗ ∗
In the remaining sections we focus on two problems. The first one is com-
pleteness of fuzzy Horn logic. The second one is definability of classes of L-
algebras by implicational theories and characterization of definable classes by
their closedness under suitable class operators.

4.2 Semantic Entailment

In this section we are going to characterize a degree to which a P-implication


semantically follows from a given L-set of P-implications. Recall that in fuzzy
equational logic, we identified L-sets of identities (over variables X) with
binary L-relations on T (X), see Sect. 3.2. Thus, if Σ was an L-set of identities
over X, we sometimes denoted Σ(t ≈ t ) by Σ(t, t ) and worked with Σ as
with an L-relation and vice versa. This allowed us to identify semantically
closed L-sets of identities with particular binary L-relations (fully invariant
congruences). In this section, we give an analogous characterization for L-sets
of P-implications. The subsequent results are, in fact, generalizations of those
of fuzzy equational logic. We will comment on this later on.
Let us stress a complication which arises when dealing with general P-
implications: an L-set Σ of P-implications cannot be in general represented
by a single binary L-relation on T (X). For instance, we can have P, Q ∈ P,
4.2 Semantic Entailment 189

P = Q, such that Σ(P i (t ≈ t )) = a = b = Σ(Q i (t ≈ t )). So, repre-


senting Σ by an L-relation to which t, t belongs in degree Σ(P i (t ≈ t ))
does not work any more for P-implications. We resolve the problem of repre-
senting Σ as follows. Considering any P-implication P i (t ≈ t ), the L-set
P of premises can be seen as an index. Thus, for each P ∈ P we can intro-
duce a separate binary L-relation SP on T (X). Now we can put SP (t, t ) =
Σ(P i (t ≈ t )) for any P ∈ P and t, t ∈ T (X). Clearly, a system of all such
SP ’s contains full information about the original L-set Σ of P-implications.
Returning to the previous example, we have SP (t, t ) = a and SQ (t, t ) = b.
Furthermore, given a system {SP | P ∈ P} of binary L-relations on T (X),
there is a unique L-set Σ of P-implications with Σ(P i (t ≈ t )) = SP (t, t )
for any P i (t ≈ t ). This particular representation of L-sets of P-implications
is introduced in the following definition.

Definition 4.24. Let P be a proper family of premises. A system


 
S = SP ∈ LT (X)×T (X) | P ∈ P (4.29)
is called a P -indexed system of L-relations. For a P-indexed system S of
L-relations we define an L-set ΣS of P-implications by
ΣS (P i (t ≈ t )) = SP (t, t ), for every P ∈ P and t, t ∈ T (X). (4.30)
For every L-set Σ of P-implications we define a P-indexed system SΣ of L-
relations as follows

SΣ = SP ∈ LT (X)×T (X) | SP (t, t ) = Σ(P i (t ≈ t ))
 (4.31)
for every P ∈ P and t, t ∈ T (X) .
For convenience, if Σ is an L-set of P-implications, then SP ∈ SΣ will be
alternatively denoted by ΣP .

Remark 4.25. (1) It is immediate that ΣSΣ = Σ, SΣS = S. That is, there is an
obvious bijective correspondence between P-indexed systems S of L-relations
and L-sets Σ of P-implications, and we can go from S to the corresponding
Σ and vice versa.
(2) For P-indexed systems S, S  of L-implications, we put S ≤ S  iff for
every P ∈ P we have SP ⊆ SP , i.e. iff ΣS ⊆ ΣS  . Consequently, S = S  iff
SP = SP for every P ∈ P iff ΣS = ΣS  .
(3) An L-set Σ of identities can be thought of as an L-set of P-implications
for P = {∅}. Thus, the corresponding {∅}-indexed system SΣ consists of a
single L-relation denoted by Σ∅ .
(4) SP denotes an element of a system S. Notice that we use S to denote
the subsethood degree, see (1.66). There is no danger of confusion here since
SP is a fuzzy relation on terms while S is a fuzzy relation on fuzzy sets.
Moreover, elements of S are always used with subscripts (SP , Si , etc.).

The following definition introduces particular P-indexed systems of L-


relations that will be important in the subsequent development.
190 4 Fuzzy Horn Logic

Definition 4.26. Suppose L∗ is a complete residuated lattice with implica-


tional truth stresser ∗ . Let Σ be an L-set of P-implications. If the correspond-
ing SΣ satisfies
P ⊆ ΣP , (4.32)
ΣP (t, t ) ≤ Σh(P ) (h(t), h(t )) , (4.33)

S(P, ΣQ ) ≤ S(ΣP , ΣQ ) , (4.34)

for every P, Q ∈ P, t, t ∈ T (X), and every endomorphism h on T(X),
then SΣ is called an L∗ -implicational P -indexed system of L-relations.
Moreover, if ΣP is a congruence for every P ∈ P, then SΣ is called an L∗ -
implicational P -indexed system of congruences.
Remark 4.27. (1) Since P is supposed to be proper, h(P ) ∈ P for every P ∈ P
as required by Definition 4.1. That is, (4.33) is defined correctly. Also note
that for Σ, ΣP (t ≈ t ) = Σ(P i (t ≈ t )).
(2) Condition (4.34) has equivalent formulations. For instance, (4.34) holds
for every P, Q ∈ P iff
  ∗
ΣP (t, t ) ⊗  
s,s ∈T (X) P (s, s ) → ΣQ (s, s ) ≤ ΣQ (t, t ) (4.35)
for all t, t ∈ T (X). This equivalent formulation will be used later on.
(3) Suppose Σ∅ is a congruence.  Then if (4.33) holds
 for P = ∅, we can
claim Σ∅ (t, t ) ≤ Σh(∅) h(t), h(t ) = Σ∅ h(t), h(t ) . Hence, Σ∅ is a fully
invariant congruence, see Sect. 3.2.
(4) For particular proper families of premises, some of the conditions
(4.32)–(4.34) will simplify or hold trivially.
Remark 4.28. Let us comment on the intuitive meaning of (4.32)–(4.34). Sup-
pose Σ = Impl(M) for some L-algebra M. So, think of Σ(P i (t ≈ t )) as
of a degree to which P i (t ≈ t ) is true in M. Then (4.32), called exten-
sivity, says that the validity degree of P i (t ≈ t ) is at least as high as the
degree to which t, t (recall that t, t stands for identity t ≈ t ) belongs
to P . Roughly speaking, if t, t is in P then P i (t ≈ t ) is always true
which is an obvious property. (4.33), called stability, says, roughly speaking,
that if P i (t ≈ t ) is true then h(P ) i (h(t) ≈ h(t )) is true. Thinking of en-
domorphisms as of substitutions, (4.33) says that validity is preserved under
substitutions. Finally, (4.34), called ∗ -monotony, says that a mapping sending
P to ΣP obeys one of characteristic properties of an L∗ -closure operator, see
Sect. 1.3. If P = LT (X)×T (X) then Lemma 1.94 yields that a mapping sending
P to ΣP is an L∗ -closure operator. Note, however, that if P = LT (X)×T (X) ,
then our mapping sending P to ΣP is in fact only a partial L∗ -closure operator
since it is a mapping of P to LT (X)×T (X) , not of LT (X)×T (X) to LT (X)×T (X)
(consider, for instance, families of crisp premises). Nevertheless, (4.32)–(4.34)
say that sending P to ΣP has closure properties and that Σ is closed under
substitutions. To sum up, L∗ -implicational P-indexed systems of L-relations
(congruences) seem to be promising candidates for being algebraic represen-
tatives of L∗ -implicational P-theories.
4.2 Semantic Entailment 191

The following auxiliary lemma shows that (4.33) and (4.34) can be equiv-
alently replaced by a single condition.

Lemma 4.29. Suppose we have a P-indexed system SΣ of L-relations. Then


SΣ is an L∗ -implicational P-indexed system of L-relations if and only if SΣ
satisfies (4.32) together with
ΣP (t, t ) ≤
  
 
∗  
≤ s,s ∈T (X) P (s, s ) → ΣQ h(s), h(s ) → ΣQ h(t), h(t ) (4.36)
for every P, Q ∈ P, t, t ∈ T (X), and every endomorphism h on T(X).

Proof. “⇒”: Let us suppose (4.32), (4.33), and (4.34) hold. Take P, Q ∈ P,
t, t ∈ T (X), and an endomorphism h : T(X) → T(X). Using (1.40), (4.9),
and the adjointness property it follows that
 
h(P )(r, r ) → ΣQ (r, r ) =  
h(s) = r P (s, s ) → ΣQ (r, r ) =
h(s ) = r 
   
    
= h(s) = r P (s, s ) → ΣQ (r, r ) = h(s) = r P (s, s ) → ΣQ h(s), h(s ) ,
h(s ) = r  h(s ) = r 

for every r, r ∈ T (X). Thus, using (4.33), (4.35) we obtain


 
ΣP (t, t ) ≤ Σh(P ) h(t), h(t ) ≤
   
∗  
≤ r,r  ∈T (X) h(P )(r, r ) → ΣQ (r, r ) → ΣQ h(t), h(t ) =
     ∗  
= r,r  ∈T (X) h(s) = r P (s, s ) → ΣQ h(s), h(s ) → ΣQ h(t), h(t ) =
h(s ) = r 
   ∗  
= s,s ∈T (X) P (s, s ) → ΣQ h(s), h(s ) → ΣQ h(t), h(t ) .
Hence, the inequality (4.36) is satisfied.
“⇐”: Assume that conditions (4.32), (4.36) hold.
(4.33): Take any endomorphism h : T(X) → T(X) and put Q = h(P ).
From Definition 4.1 it follows that Q ∈ P. Now, using (4.36) we obtain
ΣP (t, t ) ≤
  
 
∗  
≤ r,r  ∈T (X) P (r, r ) → Σh(P ) h(r), h(r ) → Σh(P ) h(t), h(t ) .
Moreover, applying (4.9), (4.32), we have
    
P (r, r ) ≤ h(s) = h(r) P (s, s ) = h(P ) h(r), h(r ) ≤ Σh(P ) h(r), h(r ) ,
h(s ) = h(r  )
 
i.e. P (r, r ) → Σh(P ) h(r), h(r ) = 1 for all r, r ∈ T (X). Now, (1.11) gives
   
ΣP (t, t ) ≤ 1∗ → Σh(P ) h(t), h(t ) = Σh(P ) h(t), h(t ) ,
proving (4.33).
(4.34): Applying (4.36) to h being the identical morphism (i.e. h(t) = t)
we get (4.35), a condition equivalent to (4.34). 

192 4 Fuzzy Horn Logic

Remark 4.30. In the classical case, a concept of a semantically closed set of im-
plications corresponds to so-called fully invariant closure operators and fully
invariant closure systems in T (X) × T (X), see [97]. Recall that the condition
of full invariance of a closure operator cl means that h(cl (P )) ⊆ cl (h(P )) for
every P ⊆ T (X)×T (X). In case of implications with finite premises, semanti-
cally closed sets of implications correspond to algebraic (i.e. finitely generated)
fully invariant closure operators. The following assertions show that if we re-
strict ourselves to P = LT (X)×T (X) , our L∗ -implicational P-indexed systems of
L-relations can be seen as particular L∗ -closure operators which have proper-
ties analogous to the fully invariant closure operators known from the classical
case (cf. Remark 4.28).

Theorem 4.31. Let P = LT (X)×T (X) and let SΣ be an L∗ -implicational P-


indexed system of L-relations. Then an operator cl on LT (X)×T (X) defined by
cl (P ) = ΣP is an L∗ -closure operator and
   
h cl (P ) ⊆ cl h(P ) (4.37)
holds for every P ∈ LT (X)×T (X) , and every endomorphism h : T(X) → T(X).

Proof. Due to Lemma 1.94, we only check (4.37). Take P ∈ LT (X)×T (X) and
let h be an endomorphism on T(X). We have
 
h(ΣP )(t, t ) = h(s) = t ΣP (s, s ) ≤ h(s) = t Σh(P ) (h(s), h(s )) =
h(s ) = t h(s ) = t

= h(s) = t Σh(P ) (t, t ) = Σh(P ) (t, t )


h(s ) = t
   
for all t, t ∈ T (X). Thus, h cl (P ) ⊆ cl h(P ) . 


Theorem 4.32. Let P = LT (X)×T (X) and suppose ∗ is an implicational truth


stresser. Let cl be an L∗ -closure operator on LT (X)×T (X) satisfying (4.37) for
every P ∈ LT (X)×T (X) and every endomorphism h : T(X) → T(X). Then the
P-indexed system SΣ = {ΣP | ΣP = cl (P )} is an L∗ -implicational P-indexed
system of L-relations.

Proof. (4.33): Using (4.37) we get


  
ΣP (t, t ) = cl (P )(t, t ) ≤ h(s) = h(t) cl (P )(s, s ) = h cl (P ) (h(t), h(t )) ≤
h(s ) = h(t )
 
≤ cl h(P ) (h(t), h(t )) = Σh(P ) (h(t), h(t ))
for all terms t, t ∈ T (X). The rest follows from Lemma 1.94. 


Remark 4.33. Note that the correspondences described by Theorem 4.31 and
Theorem 4.32 are in fact mutually inverse. Therefore, for P = LT (X)×T (X) (no
restriction on premises) there is a natural bijective correspondence between
L∗ -implicational P-indexed system of binary L-relations and fully invariant
L∗ -closure operators in T (X), generalizing the ordinary case.
4.2 Semantic Entailment 193

∗ ∗ ∗
We are now going to show that L∗ -implicational P-indexed systems of con-
gruences are in one-to-one correspondence with L∗ -implicational P-theories,
i.e. theories (composed of P-implications) of classes of L-algebras with respect
to a given implicational truth stresser.
 if K is a class of L-algebras,
For brevity,we adopt the following convention:

we denote by “ M,v · · · ” the infimum “ M∈K v:X→M · · · ” which ranges over
all L-algebras M ∈ K and all valuations v : X → M .
Theorem 4.34. Let Σ be an L∗ -implicational P-theory. Then ΣP is a con-
gruence on T(X) for every P ∈ P.
Proof. Since Σ is an L∗ -implicational P-theory, there is a class K of L-algebras,
such that Σ = Impl(K). That is, we have
  
Σ P i (t ≈ t ) = !P i (t ≈ t )!K = M∈K !P i (t ≈ t )!M =
 
= M∈K v:X→M !P i (t ≈ t )!M,v .
  
Thus, denoting Σ P i (t ≈ t ) by M,v !P i (t ≈ t )!M,v , we check the
conditions of Definition 2.11.
Reflexivity and symmetry of ΣP follow from reflexivity and symmetry
of every L-equality. For transitivity, let P ∈ P and t, t , t ∈ T (X). Using
(1.34), (1.41), (1.18), properties of ≈M ’s, and the isotony of → in the second
argument, we have
ΣP (t, t ) ⊗ ΣP (t , t ) =
 
   

= M,v !P i (t ≈ t )!M,v ⊗ M,v !P i (t ≈ t )!M,v ≤
  
≤ M,v !P i (t ≈ t )!M,v ⊗ !P i (t ≈ t )!M,v =
    
= M,v !P !M,v → !t ≈ t !M,v ⊗ !P !M,v → !t ≈ t !M,v ≤
    
≤ M,v !P !M,v ⊗ !P !M,v → !t ≈ t !M,v ⊗ !t ≈ t !M,v =
   
= M,v !P !M,v → !t ≈ t !M,v ⊗ !t ≈ t !M,v ≤ ΣP (t, t ) .
Hence, ΣP is transitive.
It suffices to check the compatibility with functions since ≈T(X) ⊆ ΣP
holds trivially. Take P ∈ P, an n-ary f ∈ F , and terms t1 , t1 , . . . , tn , tn . Since
n  n
i=1 !ti ≈ ti !M,v = i=1 !ti !M,v ≈
M
!ti !M,v ≤
   
≤ f M !t1 !M,v , . . . , !tn !M,v ≈M f M !t1 !M,v , . . . , !tn !M,v =
= !f (t1 , . . . , tn )!M,v ≈M !f (t1 , . . . , tn )!M,v =
= !f (t1 , . . . , tn ) ≈ f (t1 , . . . , tn )!M,v ,
we get
n 
ΣP (t1 , t1 ) ⊗ · · · ⊗ ΣP (tn , tn ) = i=1 M,v !P i (ti ≈ ti )!M,v ≤
 n
≤ M,v i=1 !P i (ti ≈ ti )!M,v =
194 4 Fuzzy Horn Logic
 n  
= M,v i=1 !P !M,v → !ti ≈ ti !M,v ≤
  n n 
≤ M,v !P !M,v → i=1 !ti ≈ ti !M,v ≤
  
≤ M,v !P !M,v → !f (t1 , . . . , tn ) ≈ f (t1 , . . . , tn )!M,v =
 
= ΣP f (t1 , . . . , tn ), f (t1 , . . . , tn ) .
Hence, ΣP is compatible. Altogether, ΣP is a congruence for any P ∈ P. 


Remark 4.35. In the proof of Theorem 4.34 we used (1.18), i.e. a∗ ⊗ a∗ = a∗ ,


which is required for implicational truth stressers. Without postulating (1.18),
we are not able to prove that ΣP is transitive and compatible. When every
ΣP is a congruence, we can easily define a class of models as factorizations of
T(X), see Theorem 4.38.

We need the following lemma.

Lemma 4.36. Suppose P is a proper set of premises and ∗ is an implicational


truth stresser. Then for every P, Q ∈ P, for every L-algebra M, and for every
valuation v : X → M , we have
    
∗
s,s ∈T (X) P (s, s ) → M,v !Q i (s ≈ s )!M,v ≤
  
≤ M,v !Q!M,v → !P !M,v . (4.38)

Proof. First, we can use properties of ∧, ∨ together with (1.29), (1.39) to get
    

s,s ∈T (X) P (s, s ) → M,v !Q i (s ≈ s )!M,v =
    
= s,s ∈T (X) P (s, s ) → M,v !Q!M,v → !s ≈ s !M,v =
    
= s,s ∈T (X) M,v P (s, s ) → !Q!M,v → !s ≈ s !M,v =
    
= M,v s,s ∈T (X) !Q!M,v → P (s, s ) → !s ≈ s !M,v =
    
= M,v !Q!M,v → s,s ∈T (X) P (s, s ) → !s ≈ s !M,v .
Now using (1.17), (1.13), and (1.19), we have
    
∗
s,s ∈T (X) P (s, s ) → M,v !Q i (s ≈ s )!M,v =
     
∗
= M,v !Q!M,v → s,s ∈T (X) P (s, s ) → !s ≈ s !M,v =
     
∗
= M,v !Q!M,v → s,s ∈T (X) P (s, s ) → !s ≈ s !M,v ≤
  ∗    
∗ 
≤ M,v !Q!M,v → s,s ∈T (X) P (s, s ) → !s ≈ s !M,v =
     
∗ 
= M,v !Q!M,v → s,s ∈T (X) P (s, s ) → !s ≈ s !M,v =
  
= M,v !Q!M,v → !P !M,v ,
proving the inequality (4.38). 


Theorem 4.37. Let Σ be an L∗ -implicational P-theory. Then the correspond-


ing SΣ is an L∗ -implicational P-indexed system of congruences.
4.2 Semantic Entailment 195

Proof. Let us have a class K of L-algebras such that Σ = Impl(K). We have


to check extensivity, stability and ∗ -monotony of SΣ . The rest follows from
Theorem 4.34.
(4.32): Suppose we have P ∈ P and terms t, t ∈ T (X). For every M ∈ K
and a valuation v : X → M , we have
 
P (t, t ) ≤ P (t, t ) → !t ≈ t !M,v → !t ≈ t !M,v ,
which follows from the fact that the inequality a ≤ (a → b) → b holds in every
residuated lattice. Moreover, we can use (1.12) and the antitony of → in the
first argument to state that
 ∗
P (t, t ) ≤ P (t, t ) → !t ≈ t !M,v → !t ≈ t !M,v .

Hence, by (1.19), and by properties of , we obtain the following inequality,
  ∗
P (t, t ) ≤ s,s ∈T (X) P (s, s ) → !s ≈ s !M,v → !t ≈ t !M,v =
   
∗
= s,s ∈T (X) P (s, s ) → !s ≈ s !M,v → !t ≈ t !M,v =
= !P i (t ≈ t )!M,v ,

showing that P (t, t ) ≤ M,v !P i (t ≈ t )!M,v = ΣP (t, t ), i.e. P ⊆ ΣP .
(4.33): First, we can use Lemma 3.18 to observe that for all terms
r, r ∈ T (X), a truth degree !h(r) ≈ h(r )!M,v equals to !r ≈ r !M,w for some
valuation w : X → M . Moreover, w is determined uniquely by the endomor-
phism h. Hence, it follows that
   ∗ 
ΣP (t, t ) = M,v  
s,s ∈T (X) P (s, s ) → !s ≈ s !M,v → !t ≈ t !M,v ≤
    
∗ 
≤ M,v s,s ∈T (X) P (s, s ) → !h(s) ≈ h(s )!M,v → !h(t) ≈ h(t )!M,v
for any endomorphism h on T(X). Furthermore,
   

s,s ∈T (X) P (s, s ) → !h(s) ≈ h(s )!M,v =
   
= s,s ∈T (X) h(r) = h(s) P (r, r ) → !h(s) ≈ h(s )!M,v =
h(r  ) = h(s )
   
= s,s ∈T (X) h(r) = h(s) P (r, r ) → !h(s) ≈ h(s )!M,v =
h(r  ) = h(s )
  
= h(P )(h(s), h(s )) → !h(s) ≈ h(s )!M,v =
s,s ∈T (X)
  
= s,s ∈T (X) h(P )(s, s ) → !s ≈ s !M,v .
Putting both facts together, we obtain
   ∗
ΣP (t, t ) ≤ M,v  
s,s ∈T (X) P (s, s ) → !h(s) ≈ h(s )!M,v →


→ !h(t) ≈ h(t )!M,v =
    
∗ 
= M,v s,s ∈T (X) h(P )(s, s ) → !s ≈ s !M,v → !h(t) ≈ h(t )!M,v =

= M,v !h(P ) i (h(t) ≈ h(t ))!M,v = Σh(P ) (h(t), h(t )) .
(4.34): We can use a consequence of Lemma 4.36 together with (1.30),
(1.41), and well-known properties of to prove the following inequality
196 4 Fuzzy Horn Logic
  ∗
ΣP (t, t ) ⊗  
s,s ∈T (X) P (s, s ) → ΣQ (s, s ) =
  

= M,v !P !M,v → !t ≈ t !M,v ⊗
   ∗
⊗ s,s ∈T (X) P (s, s ) → M,v !Q i (s ≈ s )!M,v


  
   
≤ M,v !P !M,v → !t ≈ t !M,v ⊗ M,v !Q!M,v → !P !M,v ≤
  
  
≤ M,v !P !M,v → !t ≈ t !M,v ⊗ !Q!M,v → !P !M,v ≤
  
≤ M,v !Q!M,v → !t ≈ t !M,v = ΣQ (t, t ) .
The previous inequality is equivalent to (4.34). The proof is complete. 


Now we turn our attention to the converse problem. Having given an L∗ -


implicational P-indexed system of congruences SΣ , we construct a suitable
class of L-algebras K, such that Σ = Impl(K).

Theorem 4.38. Let SΣ be an L∗ -implicational P-indexed system of congru-


ences. Then Σ is an L∗ -implicational P-theory. Namely, Σ = Impl(K) for a
class K defined by K = {T(X)/ΣP | P ∈ P}.

Proof. We have to check that ΣP (t, t ) = !P i (t ≈ t )!K , where



!P i (t ≈ t )!K = Q ∈ P !P i (t ≈ t )!T(X)/ΣQ ,v .
v: X→T (X)/ΣQ

“≤”: Take T(X)/ΣQ ∈ K and v : X → T (X)/ΣQ . Lemma 3.21 yields that


there is an endomorphism h on T(X), such that [h(s)]ΣQ = !s!T(X)/ΣQ ,v for
each s ∈ T (X). Consequently,
!s ≈ s !T(X)/ΣQ ,v = !s!T(X)/ΣQ ,v ≈T(X)/ΣQ !s !T(X)/ΣQ ,v =
 
= [h(s)]ΣQ ≈T(X)/ΣQ [h(s )]ΣQ = ΣQ h(s), h(s )
for every s, s ∈ T (X). Using (4.36) we get
   ∗  
ΣP (t, t ) ≤  
s,s ∈T (X) P (s, s ) → ΣQ h(s), h(s ) → ΣQ h(t), h(t ) =
   
∗
= s,s ∈T (X) P (s, s ) → !s ≈ s !T(X)/ΣQ ,v → !t ≈ t !T(X)/ΣQ ,v =
= !P i (t ≈ t )!T(X)/ΣQ ,v .
That is, ΣP (t, t ) ≤ !P i (t ≈ t )!T(X)/ΣQ ,v holds for arbitrary ΣQ , and
every valuation v : X → T (X)/ΣQ . Hence, we have proved the “≤” inequality.
“≥”: For every P ∈ P and a valuation v : X → T (X)/ΣP , where v(x) =
[x]ΣP for all x ∈ X, we have !t!T(X)/ΣP ,v = [t]ΣP for each t ∈ T (X) (easy
proof by structural induction). Thus, it follows that
ΣP (s, s ) = [s]ΣP ≈T(X)/ΣP [s ]ΣP =
= !s!T(X)/ΣP ,v ≈T(X)/ΣP !s !T(X)/ΣP ,v = !s ≈ s !T(X)/ΣP ,v
for every s, s ∈ T (X). Now we can use property (1.11) of ∗
and (4.32) to
obtain the desired inequality:
4.2 Semantic Entailment 197

ΣP (t, t ) = !t ≈ t !T(X)/ΣP ,v = 1∗ → !t ≈ t !T(X)/ΣP ,v =


   
∗
= s,s ∈ T (X) P (s, s ) → ΣP (s, s ) → !t ≈ t !T(X)/ΣP ,v =
   
∗
= s,s ∈ T (X) P (s, s ) → !s ≈ s !T(X)/ΣP ,v → !t ≈ t !T(X)/ΣP ,v =

= !P i (t ≈ t )!T(X)/ΣP ,v ≥ Q ∈ P !P i (t ≈ t )!T(X)/ΣQ ,v .
v: X→T (X)/ΣQ

Putting all together, we obtain Σ = Impl(K). 




A summary of the previous observations follows.

Corollary 4.39. Let L∗ be a complete residuated lattice with an implicational


truth stresser ∗ . Let P be a proper family of premises, Σ be an L-set of P-
implications. Then SΣ is an L∗ -implicational P-indexed system of congruences
iff Σ is an L∗ -implicational P-theory. 


∗ ∗ ∗
From this moment on, we focus on a semantic entailment from L-sets of P-
implications. First, we check that a system of all L∗ -implicational P-indexed
systems of congruences is a closure system itself. Then, given an L-set Σ
of P-implications, we introduce its semantic closure and show that a degree
!P i (t ≈ t )!Σ of semantic entailment equals a degree to which P i (t ≈ t )
belongs to the semantic closure of Σ.
Let N = {Si | i ∈ I} be a system of P-indexed systems Si of(X)
L-relations.

That is, each Si ∈ N is a P-indexed system
Si = Si,P ∈ LT (X)×T
 | P ∈ P .
 
We define a P-indexed system i∈I Si = i∈I Si P | P ∈ P of L-relations
as the intersection N , i.e.
  
i∈I Si P = i∈I Si,P . (4.39)
for every P ∈ P. Hence,
  
i∈I Si P
(t, t ) = i∈I Si,P (t, t )
for all t, t ∈ T (X) and P ∈ P.

Theorem 4.40. Let S ∗ denote a system of all L∗ -implicational P-indexed sys-


tems of congruences. Then S ∗ is a closure system.

Proof. We have to check, that S ∗ is non-empty and closed under arbitrary


intersections. It is easy to observe that the P-indexed system of L-relations
Smax = {SP | SP (t, t ) = 1 for every P ∈ P, t , t ∈ T (X)}
satisfies conditions (4.32)–(4.34) trivially, and every SP ∈ Smax , P ∈ P is a
congruence relation. Hence, Smax ∈ S ∗ and so S ∗ is non-empty.
 
Let N = {Si ∈ S ∗ | i ∈ I}. We have to show that N = i∈I Si ∈ S ∗ .
 
Theorem 2.15 yields that i∈I Si P = i∈I Si,P ∈ ConL (T(X)) for every
198 4 Fuzzy Horn Logic

P ∈ P,  since Si,P is a congruence for each i ∈ I. Hence, it remains to check


that i∈I Si satisfies conditions (4.32)–(4.34).
(4.32): Let P ∈ P, t, t ∈ T (X). We have 
 P (t, t ) ≤ S i,P (t,
 t ) for all
 i ∈ I

since every Si is extensive. Thus, P (t, t ) ≤ i∈I Si,P (t, t ) = i∈I i P (t, t ),
S
i.e. (4.32) is satisfied.
(4.33): We can proceed analogously, let P ∈ P, t,t ∈ T (X)and let h be an
endomorphism on T(X). Thus, Si,P (t, t ) ≤ Si,h(P ) h(t), h(t ) for each i ∈ I.
That is,
    
i∈I Si P (t, t ) = i∈I Si,P (t, t ) ≤
      
≤ i∈I Si,h(P ) h(t), h(t ) = 
i∈I Si h(P ) h(t), h(t ) .

(4.34): Take any P, Q ∈ P, and terms t, t ∈ T (X). Taking into account


(1.19) and ∗ -monotony of every Si , it follows that
  
  
  
∗
i∈I Si P (t, t ) ⊗ s,s ∈T (X) P (s, s ) → i∈I Si Q (s, s ) =
     ∗
= i∈I Si,P (t, t ) ⊗ s,s ∈T (X) P (s, s ) → j∈I Sj,Q (s, s ) =
 
    
∗
= i∈I Si,P (t, t ) ⊗ j∈I s,s ∈T (X) P (s, s ) → Sj,Q (s, s ) =
    ∗
= i∈I Si,P (t, t ) ⊗ j∈I 
s,s ∈T (X) P (s, s ) → Sj,Q (s, s )


  
    
 ∗
≤ i,j∈I Si,P (t, t ) ⊗ s,s ∈T (X) P (s, s ) → Sj,Q (s, s ) ≤
      ∗ 
≤ i∈I Si,P (t, t ) ⊗ s,s ∈T (X) P (s, s ) → Si,Q (s, s ) ≤ Sk,Q (t, t )
for each k ∈ I. Hence,
  
  
  
∗
i∈I Si P (t, t ) ⊗ s,s ∈T (X) P (s, s ) → i∈I Si Q (s, s ) ≤
 
  
≤ k∈I Sk,Q (t, t ) = i∈I Si Q (t, t ) ,

verifying (4.35) which is equivalent to (4.34). 




As a direct consequence of Theorem 4.40, we obtain the following corollary.

Corollary 4.41. For every L-set Σ of P-implications, and every implicational


truth stresser ∗ , there ∗
  exists a least L -implicational P-indexed system of con-
gruences SΣ  = ΣP | P ∈ P , such that ΣP ⊆ ΣP for every P ∈ P. 


Thus, we define a semantic closure as follows.

Definition 4.42. Let L∗ be a complete residuated lattice with implicational



truth stresser, Σ be an L-set of P-implications. Then SΣ  = ΣP | P ∈ P
(see Corollary 4.41) is called a semantic closure of SΣ . The corresponding
L-set Σ  of P-implications is called a semantic closure of Σ.

Finally, recalling (Impl(Mod(Σ)))(P i (t ≈ t )) = !P i (t ≈ t )!Σ , the


following assertion shows that Σ  = Impl(Mod(Σ)).
4.2 Semantic Entailment 199


Theorem 4.43. Suppose Σ is an L-set of P-implications, and is an impli-
cational truth stresser. Then
!P i (t ≈ t )!Σ = ΣP (t, t ) (4.40)

for every P ∈ P, and for all terms t, t ∈ T (X).

Proof. “≤”: Previous results yield that Σ can be closed to Σ  . Since Σ ⊆ Σ  ,


we have Mod(Σ  ) ⊆ Mod(Σ). Furthermore, SΣ  is an L∗-implicational P-

indexed system of congruences, thus Theorem 4.38 yields T(X)/ΣQ |Q ∈
    
P ⊆ Mod(Σ ) since ΣP (t, t ) ≤ !P i (t ≈ t )!T(X)/Σ  for every P, Q ∈ P,
Q
and t, t ∈ T (X). Thus, we have

!P i (t ≈ t )!Σ = M ∈ Mod(Σ) !P i (t ≈ t )!M ≤

≤ M ∈ Mod(Σ  ) !P i (t ≈ t )!M ≤

≤ Q ∈ P !P i (t ≈ t )!T(X)/Σ  = ΣP (t, t ) .
Q

“≥”: Take an arbitrary L-algebra M ∈ Mod(Σ). That is, for every


P-implication P i (t ≈ t ) we have Σ(P i (t ≈ t )) ≤ !P i (t ≈ t )!M ,
whence Σ ⊆ Impl({M}). Moreover, Impl({M}) is an L∗ -implicational P-
theory (defined by the one element class K = {M}), and so SImpl({M}) is an
L∗ -implicational P-indexed system of congruences (due to Theorem 4.37), con-
taining SΣ . As SΣ  is the least one containing SΣ , we get SΣ  ≤ SImpl({M}) .
As a consequence,
 
ΣP (t, t ) ≤ Impl({M}) (P i (t ≈ t )) = !P i (t ≈ t )!M .
Since M ∈ Mod(Σ) is arbitrary, we obtain

ΣP (t, t ) ≤ M ∈ Mod(Σ) !P i (t ≈ t )!M = !P i (t ≈ t )!Σ

proving “≥”. Thus, !P i (t ≈ t )!Σ = ΣP (t, t ) for each P i (t ≈ t ). 




Remark 4.44. Theorem 4.43 provides an algebraic characterization of the de-


gree of semantic entailment from L-sets of P-implications. In Sect. 4.3, we are
going to shows under what conditions the degree of semantic entailment has
a syntactic characterization.

∗ ∗ ∗
We are now going to investigate a relationship between L∗ -implicational P-
theories and their special subtheories determined by restrictions on P. Namely,
we will be interested in a restriction on finiteness of every P ∈ P. Recall that
for any proper family P of premises we can consider its restriction on finite
premises PFin . PFin -implications, called P-Horn clauses, are P-implications
which have finite L-sets of premises. The following definition introduces par-
ticular subtheories resulting from L∗ -implicational P-theories by a restriction
of finite premises.
200 4 Fuzzy Horn Logic

Definition 4.45. Let Σ be an L∗ -implicational P-theory. An L-set ΣFin of P-


Horn clauses, where ΣFin (P i (t ≈ t )) = Σ(P i (t ≈ t )) for each P-Horn
clause P i (t ≈ t ), is called an L∗ -implicational Horn subtheory of Σ.

Remark 4.46. (1) For an L∗ -implicational P-theory Σ, the corresponding L∗ -


implicational Horn subtheory ΣFin is an L∗ -implicational PFin -theory.
(2) In case of finite premises, the equivalent formulation (4.35) of condition
(4.34) simplifies. Namely, if P = PFin , (4.35) simplifies to
n  ∗
ΣP (t, t ) ⊗  
i=1 P (ti , ti ) → ΣQ (ti , ti ) ≤ ΣQ (t, t ) , (4.41)
for every t, t ∈ T (X), P, Q ∈ P such that Supp(P ) ⊆ {t1 , t1 , . . . , tn , tn }.

Now, for any L∗ -implicational P-theory Σ we have an L∗ -implicational


PFin -theory ΣFin which is a restriction of Σ on P-implications with finitely
many premises. Conversely, we may ask a question, whether for a given proper
family of premises P and an L∗ -implicational PFin -theory Γ there is an L∗ -
implicational P-theory Σ such that Γ = ΣFin . The subsequent theorems show
that Σ exists for any Γ provided that P and L∗ satisfy additional conditions.
Before presenting the results, recall that by a finite restriction of P ∈ P
we mean any finite L-relation P  ∈ LT (X)×T (X) , where P  (t, t ) > 0 implies
P  (t, t ) = P (t, t ) for all t, t ∈ T (X). The following definition introduces the
closedness under finite restrictions.

Definition 4.47. For any proper family P of premises and P ∈ P let Fin(P )
denote the set of all finite restrictions of P . P is said to be closed under
finite restrictions if Fin(P ) ⊆ P for each P ∈ P.

The following characterization is restricted to Horn truth stressers.

Theorem 4.48. Let L∗ be a complete residuated lattice with a Horn truth


stresser ∗ , let Γ be an L∗ -implicational PFin -theory, where P is a proper family
of premises closed under finite restrictions. Then putting

ΣP = P  ∈ Fin(P ) ΓP  , for P ∈ P , (4.42)
Σ is an L∗ -implicational P-theory. Moreover, Σ is the least L∗ -implicational
P-theory with Γ = ΣFin .

Proof. Using Theorem 4.38, we need to check conditions (4.32)–(4.34) of SΣ ,


the fact that every ΣP is a congruence, and the fact that Σ is the least one
with Γ = ΣFin .
ΣP is a congruence: This is easy to see since the system of congruences
{ΓP  | P  ∈ Fin(P )} is directed. Indeed, for every P1 , P2 ∈ Fin(P ), we have
P1 , P2 ⊆ P1 ∪ P2 , and P1 ∪ P2 ∈ Fin(P ), thus by extensivity of SΓ , it follows
that P1 , P2 ⊆ ΓP1 ∪ P2 , whence, by (4.34) ΓP1 , ΓP2 ⊆ ΓP1 ∪ P2 . This idea general-
izes easily for finitely many congruences, i.e. {ΓP  | P  ∈ Fin(P  )} is a directed
system of congruences. Hence, Lemma 2.68 yields that ΣP = P  ∈ Fin(P ) ΓP 
is a congruence for every P ∈ P.
4.2 Semantic Entailment 201

(4.32): Since P = P  , extensivity of SΓ gives
P  ∈ Fin(P )
 
P = P  ∈ Fin(P ) P  ⊆ P  ∈ Fin(P ) ΓP  = ΣP ,
verifying (4.32) for every P ∈ P.
(4.33): For P  ∈ Fin(P ), P ∈ P, and an endomorphism h on T(X) we
have h(P  ) ⊆ h(P ) and h(P  ) is finite, i.e. we can take P  ∈ Fin(h(P )) such
that Supp(h(P  )) = Supp(P  ). Obviously, h(P  ) ⊆ P  . Thus, using (4.32)
and (4.34) of SΓ we have Γh(P  ) ⊆ ΓP  . Hence,
   
ΣP (t, t ) = P  ∈ Fin(P ) ΓP  (t, t ) ≤ P  ∈ Fin(P ) Γh(P  ) h(t), h(t ) ≤
    
≤ P  ∈ Fin(h(P )) ΓP  h(t), h(t ) = Σh(P ) h(t), h(t )
proves (4.33) for SΣ .
(4.34): Let P  ∈ Fin(P ) with Supp(P  ) = {ti , ti | i = 1, . . . , k}, P ∈ P,
and t, t ∈ T (X). Since SΓ satisfies (4.34) equivalently formulated by (4.35),
we have
k   ∗
ΓP  (t, t ) ⊗  
i=1 P (ti , ti ) → ΓQ (ti , ti ) ≤ ΓQ (t, t )
for every Q ∈ Fin(Q), Q ∈ P. Hence,
 k   ∗
ΓP  (t, t ) ⊗ Q ∈ Fin(Q)  
i=1 P (ti , ti ) → ΓQ (ti , ti )
 ≤
  
≤ Q ∈ Fin(Q) ΓQ (t, t ) = ΣQ (t, t ) .
Using (1.19), (1.20) we get
 k  ∗
ΓP  (t, t ) ⊗ Q ∈ Fin(Q) i=1 P  (ti , ti ) → ΓQ (ti , ti ) ≤ ΣQ (t, t ) . (4.43)
Since {ΓQ | Q ∈ Fin(Q)} is a directed system, we have
k      ∗

i=1 Q ∈ Fin(Q) P (ti , ti ) → ΓQ (ti , ti ) =
 k   ∗
= Q ,...,Q ∈ Fin(Q) i=1 P (ti , ti ) → ΓQi (ti , ti ) =

1
k  ∗
k

= Q ∈ Fin(Q) i=1 P  (ti , ti ) → ΓQ (ti , ti ) .
Using (1.19), (1.20), (1.21), (4.43), and the previous equality we further get
k   ∗
ΓP  (t, t ) ⊗  
i=1 P (ti , ti ) → ΣQ (ti , ti ) =
 k     ∗

= ΓP  (t, t ) ⊗ i=1 P (ti , ti ) → ΣQ (ti , ti ) =
k   ∗
= ΓP  (t, t ) ⊗ i=1 P  (ti , ti ) → Q ∈ Fin(Q) ΓQ (ti , ti ) =
k   ∗
= ΓP  (t, t ) ⊗ i=1 Q ∈ Fin(Q) P  (ti , ti ) → ΓQ (ti , ti ) =
 k  ∗
= ΓP  (t, t ) ⊗ Q ∈ Fin(Q) i=1 P  (ti , ti ) → ΓQ (ti , ti ) ≤ ΣQ (t, t ) .
Note that (1.21) can be used properly because Fin(Q) is always a non-empty
set (e.g. ∅T (X)×T (X) ∈ Fin(Q)). Moreover, for every P  ∈ Fin(P ), we have
P  (s, s ) = P (s, s ) for every s, s ∈ Supp(P  ). Thus, using monotony of ∗ ,
we have
202 4 Fuzzy Horn Logic
  ∗
ΓP  (t, t ) ⊗ P (s, s ) → ΣQ (s, s )
s,s ∈ T (X) ≤
   ∗
≤ ΓP  (t, t ) ⊗  
≤ ΣQ (t, t )
k
i=1 P (ti , ti ) → ΣQ (ti , ti )

for any P  ∈ Fin(P ). Hence, we have


  ∗
ΣP (t, t ) ⊗ 
s,s ∈ T (X) P (s, s ) → ΣQ (s, s )

=
 
   
∗
= P  ∈ Fin(P ) ΓP  (t, t ) ⊗ s,s ∈ T (X) P (s, s ) → ΣQ (s, s ) ≤ ΣQ (t, t ) ,
verifying (4.34). Altogether, SΣ is an L∗ -implicational P-indexed system of
congruences, thus the corresponding Σ is an L∗ -implicational P-theory by
Theorem 4.38.
Σ is the least one with Γ = ΣFin : First, we check Γ = ΣFin . Take any
P ∈ PFin . For every P  ∈ Fin(P ), using (4.32), (4.34), we have P  ⊆ ΓP ,
and consequently ΓP  ⊆ ΓP . On the other hand, ΓP ⊆ P  ∈ Fin(P ) ΓP  holds

trivially due to the fact P ∈ Fin(P ). Hence, ΣP = P  ∈ Fin(P ) ΓP  = ΓP for
every P ∈ PFin , i.e. Γ = ΣFin .
Suppose Σ  is an L∗ -implicational P-theory such that Γ = Σ  Fin . Take
any P ∈ P. We will show that ΣP ⊆ ΣP . If P ∈ PFin , we are done, since
ΣP = ΓP = ΣP as we have shown in the previous paragraph. For P ∈ PFin ,
and arbitrary P  ∈ Fin(P ), it follows that P  ⊆ P . Using (4.32) and (4.34)
for SΣ  , we have P  ⊆ ΣP , and 
ΣP  ⊆ ΣP . Thus, ΣP  = ΓP  = ΣP  ⊆ ΣP for
all P ∈ Fin(P ). Hence, ΣP = P  ∈ Fin(P ) ΓP  ⊆ ΣP . That is, Σ is the least


L∗ -implicational P-theory for which Γ = ΣFin . 



If L is a Noetherian residuated lattice, we can avoid the usage of (1.21):
Theorem 4.49. Let L∗ be a Noetherian residuated lattice with an implica-
tional truth stresser ∗ satisfying (1.20), Γ be an L∗ -implicational PFin -theory,
where P is a proper family of premises closed under finite restrictions, ΣP
(P ∈ P) be defined by (4.42). Then Σ is an L∗ -implicational P-theory. More-
over, Σ is the least L∗ -implicational P-theory with Γ = ΣFin .
Proof. We show only the critical part where (1.21) is used. Since L is a
Noetherian residuated lattice, for each i = 1, . . . , k, there are {Q1 , . . . , Qm } ⊆
Fin(Q) such that

ΣQ (ti , ti ) = Q ∈ Fin(Q) ΓQ (ti , ti ) = ΓQ1 (ti , ti ) ∨ · · · ∨ ΓQm (ti , ti ) .
Furthermore, {ΓQ | Q ∈ Fin(Q)} is directed due to (4.32) and (4.34). Hence,
there is Qi ∈ Fin(Q) such that ΓQ1 ∪ · · · ∪ ΓQm ⊆ ΓQi . Thus, ΣQ (ti , ti ) ≤
ΓQi (ti , ti ), i.e. ΣQ (ti , ti ) = ΓQi (ti , ti ) because the “≥”-part follows from (4.42).
Therefore, for each i = 1, . . . , k there is Qi ∈ Fin(Q) where ΣQ (ti , ti ) =
ΓQi (ti , ti ). This further yields
k   
 ∗ k    ∗
i=1 (P (ti , ti ) → ΣQ (ti , ti )) = i=1 (P (ti , ti ) → ΓQi (ti , ti )) ≤
k  ∗
≤ i=1 Q ∈ Fin(Q) (P  (ti , ti ) → ΓQ (ti , ti )) .
The rest follows from the proof of Theorem 4.48. 

4.2 Semantic Entailment 203

Remark 4.50. In the ordinary case [97], fully invariant algebraic closure sys-
tems of congruences form the algebraic counterparts of Horn theories. Theo-
rem 4.48 shows a condition analogous to algebraicity in our framework. Thus,
we could define algebraic P-indexed systems of congruences as those SΣ , where
Σ satisfies (4.42). For P = LT (X)×T (X) , this could yield fully invariant alge-
braic L∗ -closure systems of congruences. For L = 2, this would further yield
fully invariant algebraic closure systems of congruences – the classical case.

∗ ∗ ∗
Let us close this section with a remark on the role of L∗ in the interpretation
of P-implications. Notice that the truth degree !P i (t ≈ t )!M,v depends on
both ∗ and →. Using → is clear (we deal with implications). We saw that the
use of ∗ naturally unifies two possible meanings of P i (t ≈ t ). Clearly, if ∗ is
globalization then !P i (t ≈ t )!M,v does not depend on →. The effect of →
is completely displaced by ∗ . Surprisingly, an analogy applies also to general
implicational truth stresser satisfying (1.20).
Since !P i (t ≈ t )!M,v equals a∗ → !t ≈ t !M,v for
  
a = s,s ∈T (X) P (s, s ) → !s ≈ s !M,v ,
we are interested in truth degrees of a∗ → b. For a∗ ≤ b we have a∗ → b = 1.
If a∗  b, a∗ → b is the greatest element of {c ∈ L | a∗ ⊗ c ≤ b}. Due to (1.20),
a∗ → b is the greatest element of {c ∈ L | a∗ ∧ c ≤ b}. That is, a∗ → b is the
relative pseudocomplement of a∗ to b. For  instance, when L is a chain, then
a∗ → b = b for a∗ > b. Note that even a = s,s ∈T (X) P (s, s ) → !s ≈ s !M,v
is defined using →. However, the following theorem shows that a∗ = !P !M,v
is not influenced by the definition of →. As a consequence, we get that the
truth degree !P i (t ≈ t )!M,v is (for given M, v) fully determined by the
lattice structure of L and the implicational truth stresser ∗ satisfying (1.20).

Theorem 4.51. Let L1 = L, ∨, ∧, ⊗1 , →1 , 0, 1 , L2 = L, ∨, ∧, ⊗2 , →2 , 0, 1


be complete residuated lattices. Let ∗ be an implicational truth stresser satis-
∗ ∗
fying (1.20) for both L1 and L2 . Then (a →1 b) = (a →2 b) for all a, b ∈ L.

Proof. For each a ∈ L, let H(a) = {c∗ | c ∈ L and c∗ ≤ a}. We claim that
H(a →1 b) = H(a →2 b) (4.44)
for all a, b ∈ L. Indeed, for any a, b, c ∈ L we have c∗ ∈ H(a →1 b) iff
c∗ ≤ a →1 b iff a ⊗1 c∗ ≤ b by adjointness, iff a ⊗2 c∗ = a ∧ c∗ = a ⊗1 c∗ ≤ b by
(1.20), iff c∗ ≤ a →2 b by adjointness, iff c∗ ∈ H(a →2 b). Hence, (4.44) holds

true for all a, b ∈ L. Clearly, (a →1 b) ≤ (a →1 b) due to (1.12). Thus, (4.44)
∗ ∗
yields (a →1 b) ∈ H(a →1 b) = H(a →2 b). That is, (a →1 b) ≤ a →2 b.

Analogously, we have (a →2 b) ≤ a →1 b. Now using monotony of ∗ together
∗ ∗∗ ∗ ∗
with (1.17) we have (a →1 b) = (a →1 b) ≤ (a →2 b) and (a →2 b) =
∗∗ ∗ ∗ ∗
(a →2 b) ≤ (a →1 b) . Hence, (a →1 b) = (a →2 b) for any a, b ∈ L. 

204 4 Fuzzy Horn Logic

Now an easy inspection of (4.16) and (4.17) shows that !P i (t ≈ t )!M,v


is not influenced by the definition of →. This is an immediate consequence of
previous observations. A summary follows.
Corollary 4.52. Let L1 = L, ∨, ∧, ⊗1 , →1 , 0, 1 , L2 = L, ∨, ∧, ⊗2 , →2 , 0, 1
be complete residuated lattices. Let ∗ be an implicational truth stresser satis-
fying (1.20) for both L1 and L2 . Let M be an Li -algebra (for both i = 1, 2)
and v be a valuation on M. Then,
L∗ L∗
!P i (t ≈ t )!M,v
1
= !P i (t ≈ t )!M,v
2

for every P-implication P i (t ≈ t ). 



Remark 4.53. One should not be mislead. Even though the effect of residuum
is eliminated as shown in Corollary 4.52, multiplication and residuum in L∗
are still relevant. For instance, compatibility (1.83) is based on ⊗. Generally,
there are L1 -algebras which are not L2 -algebras and vice versa, i.e. semantic
∗ ∗
entailments given by !· · ·!L1 and !· · ·!L2 do not coincide.

4.3 Completeness of Fuzzy Horn Logic


In this section we are interested in soundness and completeness of fuzzy Horn
logic. The soundness and completeness theorems will be proved as follows:
We define the notion of a deductive closure of an L-set of P-Horn clauses and
then we show that a deductive closure equals a semantic closure, which has
already been introduced in Sect. 4.2. Then we focus on the relationship of a
syntactic (deductive) closure and a provability degree.
We prove that fuzzy Horn logic is Pavelka-sound and in many important
cases it is Pavelka-complete. On the other hand, we show that fuzzy Horn logic
is not Pavelka-complete for arbitrary L, ∗ , and P. To demonstrate the influence
of the choice of L, ∗ , and P on completeness, we will pay attention to partic-
ular fuzzy Horn logics which arise by restrictions to particular subclasses of
structures of truth degrees (Noetherian residuated lattices, residuated lattices
on the real unit interval), or to particular families of proper premises (unre-
stricted premises, K-valent premises, crisp premises). We start with general
results which cover all the important cases.
Until otherwise mentioned, P denotes a proper family of premises satisfying
P = PFin , i.e. each P ∈ P is finite. For such a P, any P-implication is a P-Horn
clause. An L-set of P-Horn clauses will be denoted by Γ .
Definition 4.54. Suppose Γ is an L-set of P-Horn clauses. A deductive
closure of SΓ is the least P-indexed system SΓ  of L-relations satisfying
ΓP ⊆ ΓP , (4.45)
ΓP (t, t) =1, (4.46)
ΓP (t, t ) ≤ ΓP (t , t) , (4.47)
4.3 Completeness of Fuzzy Horn Logic 205

ΓP (t, t )X ⊗ ΓP (t , t ) ≤ ΓP (t, t ) , (4.48)


ΓP (t, t ) ≤ ΓP (s, s ) , (4.49)

P (t, t ) ≤ ΓP (t, t ) , (4.50)
 
ΓP (t, t ) ≤ ΓP(x/r) t(x/r), t (x/r) , (4.51)
k  ∗
ΓP (t, t ) ⊗ 
i=1 P (ti , ti ) → ΓQ (ti , ti ) ≤ ΓQ (t, t ) , (4.52)
for every P, Q ∈ P, Supp(P ) = {ti , ti | i = 1, . . . , k}, x ∈ X, terms
t, t , t , r, s, s ∈ T (X), where s contains t as a subterm and s results from s
by replacing one occurrence of t by t . The corresponding L-set Γ  of P-Horn
clauses is called a deductive closure of Γ .

Remark 4.55. (1) Observe that for a Horn truth stresser ∗ and finite P ∈ P,
condition (4.52) coincides with (4.41). Indeed, this is a consequence of (1.19)
and (1.20).
(2) The deductive closure Γ  of an L-set Γ of P-Horn clauses always exists
since the system of all P-indexed systems of L-relations satisfying conditions
(4.45)–(4.52) is non-empty and closed under arbitrary intersections. The proof
is analogous to that of Theorem 4.40 and therefore omitted.

We now present an auxiliary lemma that will be used in the sequel. For
the sake of brevity, we denote substitutions by τ , τ1 , τ2 , . . . , and so on. Fur-
thermore, instead of writing (· · · ((tτ1 )τ2 ) · · · )τn , we simply write tτ1 τ2 · · · τn .
Similarly, we denote (· · · ((P τ1 )τ2 ) · · · )τn by P τ1 τ2 · · · τn .

Lemma 4.56. Suppose P is a proper family of premises, and let us have sub-
stitutions τ1 , . . . , τk . Let τ denote τ1 · · · τk . We have,

P τ (t, t ) = sτ = t  P (s, s ) (4.53)
s τ =t

for all terms t, t ∈ T (X).

Proof. We can prove the claim using induction on the number of substitutions.
If k = 1, i.e. τ = τ1 , the claim follows directly from (4.28). Suppose that the
claim holds for each k − 1 substitutions and let us denote τ = τ1 · · · τk−1 .
Using the induction hypothesis, we have
  
P τ τk (t, t ) = uτk = t P τ (u, u ) = uτk = t 
sτ = u P (s, s ) =

u τk = t 
u τk = t s τ = u

= sτ τk = t P (s, s )
s τ τk = t

proving (4.53). 


The following assertion shows that the semantic closure of Γ and the
deductive closure of Γ coincide provided that we consider implicational truth
stressers which satisfy (1.20).
206 4 Fuzzy Horn Logic

Theorem 4.57. Let L∗ be a complete residuated lattice with an implicational


truth stresser ∗ satisfying (1.20). Then for any L-set Γ of P-Horn clauses we
have Γ  = Γ  .

Proof. “⊆”: It is sufficient to show that SΓ  is an L∗ -implicational P-indexed


system of congruences. Then Γ  ⊆ Γ  since Γ  is the least L∗ -implicational
P-indexed system of congruences containing Γ .
First, we check that every ΓP (P ∈ P) is a congruence on T(X). Reflexiv-
ity, symmetry, and transitivity of ΓP follows from (4.46)–(4.48). For compati-
bility with functions, we repeatedly use (4.49). Namely, for t1 , s1 , . . . , tn , sn ∈
T (X) and any n-ary f ∈ F (4.49) yields
 
ΓP (ti , si ) ≤ ΓP f (s1 , . . . , si−1 , ti , ti+1 , . . . , tn ), f (s1 , . . . , si−1 , si , ti+1 , . . . , tn )
for each i = 1, . . . , n. Recall that in each i-th step, we have applied (4.49)
correctly, because f (s1 , . . . , si−1 , ti , ti+1 , . . . , tn ) contains ti as a subterm and
f (s1 , . . . , si−1 , si , ti+1 , . . . , tn ) results from it by replacing ti by si . Hence, by
(4.48) and monotony of ⊗,
ΓP (t1 , s1 ) ⊗ · · · ⊗ ΓP (tn , sn ) ≤
n  
≤ i=1 ΓP f (s1 , . . . , si−1 , ti , . . . , tn ), f (s1 , . . . , si , ti+1 , . . . , tn ) ≤
 
≤ ΓP f (t1 , . . . , tn ), f (s1 , . . . , sn )
which is the desired compatibility with functions. Clearly, ≈T(X) ⊆ ΓP . Thus,
every ΓP is a congruence on T(X).
Now we check (4.32)–(4.34).
(4.32): Trivial, because of (4.50).
(4.33): Take t, t ∈ T (X), P ∈ P, where Supp(P ) = {ti , ti | i = 1, . . . , k}.
Since P is finite, a set of variables X  = var(P )∪var(t)∪var(t ) = {x1 , . . . , xn }
is finite as well. For arbitrary endomorphism h on T(X) we can take a set vari-
ables Y = {y1 , . . . , yn } such that yi ∈ X  , yi ∈ var(h(x)) for each i = 1, . . . , n,
and each variable x ∈ X  . That is, each yi is different from variables occur-
ring in terms t1 , t1 , . . . , tn , tn , t, t , and yi does not occur in any endomorphic
image h(x) of a variable x ∈ X  . Since X is assumed to be denumerable, we
can always pick such yi ’s.
Put τ = (x1 /y1 ) · · · (xn /yn )(y1 /h(x1 )) · · · (yn /h(xn )). First of all, we will
show that for a term r ∈ T (X  ) we have rτ = h(r). Indeed, we have
r(x1 , . . . , xn )(x1 /y1 ) · · · (xn /yn ) = r(y1 , . . . , yn )
since none of yi ’s is in X  . Moreover, since none of yi ’s has an occurrence in
any endomorphic image h(xi ), i = 1, . . . , n, we directly obtain
rτ = r(y1 , . . . , yn )(y1 /h(x1 )) · · · (yn /h(xn )) =
   
= r h(x1 ), . . . , h(xn ) = h r(x1 , . . . , xn ) = h(r) .
Applying this fact to t, t ∈ T (X  ) and using (4.51) we obtain
4.3 Completeness of Fuzzy Horn Logic 207
 
ΓP (t, t ) ≤ ΓP(x1 /y1 ) t(x1 /y1 ), t (x1 /y1 ) ≤
 
≤ ΓP(x1 /y1 )(x2 /y2 ) t(x1 /y1 )(x2 /y2 ), t (x1 /y1 )(x2 /y2 ) ≤ · · · ≤
 
≤ ΓPτ (tτ, t τ ) = ΓPτ h(t), h(t ) .
Clearly, now it suffices to show that h(P ) = P τ . Observe that P (r, r ) > 0
implies r, r ∈ T (X  ) (X  consists of all the variables occurring in couples of
terms which belong to P in some nonzero degree). Thus, using Lemma 4.56,
and the fact rτ = h(r) for r ∈ T (X  ), it follows that
   
P τ (s, s ) = rτ = s  P (r, r ) = h(r) = s P (r, r ) = h(P ) (s, s )
r τ =s
r,r  ∈T (X  ) h(r  ) = s
r,r  ∈T (X  )
 
for every s, s ∈ T (X). Hence, h(P ) = P τ , and so ΓP (t, t ) ≤ Γh(P  
) h(t), h(t ) ,
proving (4.33).
(4.34): For finite P ∈ P, we can express (4.34) equivalently by (4.41).
Hence, using (1.19), (1.20), and (4.52) we have
k  ∗
ΓP (t, t ) ⊗   
i=1 P (ti , ti ) → ΓQ (ti , ti ) =
  ∗
= ΓP (t, t ) ⊗ i=1 P (ti , ti ) → ΓQ (ti , ti ) =
k

k  ∗
= ΓP (t, t ) ⊗ i=1 P (ti , ti ) → ΓQ (ti , ti ) ≤ ΓQ (t, t )
for every P, Q ∈ P, Supp(P ) = {ti , ti | i = 1, . . . , k} and any terms t, t ∈
T (X). Altogether, SΓ  is an L∗ -implicational P-indexed system of congru-
ences, showing “⊆”.
“⊇”: We check that SΓ  satisfies conditions (4.45)–(4.52). Since SΓ  is
the least P-indexed system of L-relations satisfying (4.45)–(4.52), we obtain
Γ  ⊆ Γ .
Equation (4.45) holds trivially. Since every ΓP (P ∈ P) is a congruence,
conditions (4.46)–(4.48) are satisfied obviously.
(4.49): This property will be proved using the compatibility of every ΓP
with functions. Let us have terms t, t , s, s ∈ T (X), where s has an occurrence
of t as a subterm and s is a term resulting from s by substitution of t by t .
If s = f (t1 , . . . , tk−1 , t, tk+1 , . . . , tn ) and s = f (t1 , . . . , tk−1 , t , tk+1 , . . . , tn ),
compatibility of ΓP with f ∈ F and ΓP (ti , ti ) = 1 yield
k−1 n
ΓP (t, t ) = i = 1 ΓP (ti , ti ) ⊗ ΓP (t, t ) ⊗ j = k+1 ΓP (tj , tj ) ≤
 
≤ ΓP f (t1 , . . . , tk−1 , t, tk+1 , . . . , tn ), f (t1 , . . . , tk−1 , t , tk+1 , . . . , tn ) =
= ΓP (s, s ) .
This argument can be used to show ΓP (t, t ) ≤ ΓP (s, s ) even in general case
(one can proceed by structural induction over the rank of s).
(4.50): Holds trivially because of (4.32).
(4.51): Take a substitution (x/r) and a mapping g : X → T (X) defined
by g(x) = r and g(y) = y for each y ∈ X with y = x. As in Remark 4.21,
208 4 Fuzzy Horn Logic

g has a homomorphic extension h : T(X) → T(X), i.e. h = g  . Evidently,


h(t) = t(x/r) for all terms t ∈ T (X). Moreover,
     
P (x/r) (t, t ) = s(x/r) = t P (s, s ) = h(s) = t P (s, s ) = h(P ) (t, t ) .
s (x/r) = t h(s ) = t

Thus, (4.33) gives


   
ΓP (t, t ) ≤ Γh(P
   
) h(t), h(t ) = ΓP (x/r) t(x/r), t (x/r)

which is the required inequality.


(4.52): Since SΓ  satisfies (4.34), we can use the equivalent formulation
for finite sets of premises (4.41) together with (1.19), (1.20). Hence,
k  ∗
ΓP (t, t ) ⊗ i=1 P (ti , ti ) → ΓQ (ti , ti ) =
k  ∗
= ΓP (t, t ) ⊗ i=1 P (ti , ti ) → ΓQ (ti , ti ) =
k  ∗
= ΓP (t, t ) ⊗  
i=1 P (ti , ti ) → ΓQ (ti , ti )

≤ ΓQ (t, t )
holds for every P, Q ∈ P, where Supp(P ) = {ti , ti | i = 1, . . . , k}, and for all
t, t ∈ T (X). 


The following lemma shows that given any complete residuated lattice
with truth stresser, a provability degree of a P-Horn clause from an L-set Γ
of P-Horn clauses is smaller than or equal to a degree to which the P-Horn
clause belongs to a deductive closure of Γ .

Lemma 4.58. Let Γ be an L-set of P-Horn clauses. Then


|P i (t ≈ t )|Γ ≤ ΓP (t, t ) (4.54)

for every P ∈ P and t, t ∈ T (X).

Proof. It is sufficient to check that for every member P i (t ≈ t ), a of an


L∗ -weighted proof from Γ , we have a ≤ ΓP (t, t ). We proceed by induction
over the length of an L∗ -weighted proof.
Suppose P i (t ≈ t ), a is a member of an L∗ -weighted proof from Γ
where a = Γ (P i (t ≈ t )) = ΓP (t, t ). Using (4.45), we have a = ΓP (t, t ) ≤
ΓP (t, t ). Otherwise, P i (t ≈ t ), a was derived using one of (Ref)–(Mon).
We check “≤” for all of these rules separately.
(Ref): If P i (t ≈ t), 1 was inferred by (Ref), we directly obtain 1 ≤
ΓP (t, t) = 1 due to (4.46).
(Sym): If P i (t ≈ t), a was inferred from P i (t ≈ t ), a , then induc-
tion hypothesis together with (4.47) give a ≤ ΓP (t, t ) ≤ ΓP (t , t).
(Tra): Suppose P i (t ≈ t ), a ⊗ b was inferred from P i (t ≈ t ), a
and P i (t ≈ t ), b . By induction hypothesis, a ≤ ΓP (t, t ), b ≤ ΓP (t , t ).
Hence, (4.48) yields a ⊗ b ≤ ΓP (t, t ) ⊗ ΓP (t , t ) ≤ ΓP (t, t ).
(Rep): If P i (s ≈ s ), a was inferred by (Rep) from P i (t ≈ t ), a ,
induction hypothesis and (4.49) give a ≤ ΓP (t, t ) ≤ ΓP (s, s ).
(Ext): Apply (4.50).
4.3 Completeness of Fuzzy Horn Logic 209

(Sub): If P (x/r) i (t(x/r) ≈ t (x/r)), a results from P 


 i (t ≈ t ), a ,

  
induction hypothesis and (4.51) yield a ≤ ΓP (t, t ) ≤ ΓP (x/r) t(x/r), t (x/r) .
(Mon): Suppose Q i (t ≈ t ), c was inferred from P i (t ≈ t ), b and
Q i (ti ≈ ti ), ai ’s with Supp(P ) = {ti , ti | i = 1, . . . , k}. Thus, we have
k  ∗
c = b ⊗ i=1 P (ti , ti ) → ai . Using the induction hypothesis, we have b ≤
ΓP (t, t ), and ai ≤ ΓQ (ti , ti ) for each i = 1, . . . , k, from which we obtain
k  ∗
c = b ⊗ i=1 P (ti , ti ) → ai ≤
k  ∗
≤ ΓP (t, t ) ⊗ i=1 P (ti , ti ) → ΓQ (ti , ti ) ≤ ΓQ (t, t )
by (1.32), (1.56), and (4.52). 


We are now ready to prove soundness of fuzzy Horn logic.

Theorem 4.59 (soundness). Let L∗ be a complete residuated lattice with an


implicational truth stresser ∗ satisfying (1.20). Let Γ be an L-set of P-Horn
clauses. Then
|P i (t ≈ t )|Γ ≤ ΓP (t, t ) = ΓP (t, t ) = !P i (t ≈ t )!Γ (4.55)
for every P ∈ P and t, t ∈ T (X).

Proof. Follows from Theorem 4.43, Theorem 4.57, and Lemma 4.58. 


Remark 4.60. Notice that Theorem 4.59 yields that for any complete residu-
ated lattice L we have at least one Pavelka-sound fuzzy Horn logic because any
L can be endowed with a globalization on L which is an implicational truth
stresser satisfying (1.20). Hence, soundness of fuzzy Horn logic is established
without any restrictions on P or L.

∗ ∗ ∗
We now turn our attention to completeness. Due to the previous observa-
tions, we are interested in equality |P i (t ≈ t )|Γ = ΓP (t, t ). We first prove
some properties of the provability degree |· · ·|Γ .

Lemma 4.61. Let Γ be an L-set of P-Horn clauses. For any P-Horn clause
P i (t ≈ t ) put DP (t, t ) = |P i (t ≈ t )|Γ . Then D = {DP | P ∈ P} is a
P-indexed system of L-relations satisfying (4.45)–(4.51).

Proof. (4.45): Since P i (t ≈ t ), ΓP (t, t ) is an L∗ -weighted proof of length 1,


we have DP (t, t ) = |P i (t ≈ t )|Γ ≥ ΓP (t, t ), verifying (4.45).
(4.46): Evidently, DP (t, t) = 1 since P i (t ≈ t), 1 is a proof by (Ref).
(4.47): Using (Sym), each proof δi1 , . . . , δini, P i (t ≈ t ), ai can be ex-
tended to a proof
δi1 , . . . , δini ,
1: P i (t ≈ t ), ai , proof of P i (t ≈ t ), ai
2: P i (t ≈ t), ai . by (Sym) on 1
210 4 Fuzzy Horn Logic

Hence, DP (t, t ) = |P i (t ≈   
 t )|Γ ≤ |P i (t  ≈t)|Γ =DP (t , t).

(4.48): Let DP (t, t ) = i∈I ai and DP (t , t ) = j∈J bj where for each
i ∈ I and j ∈ J there are proofs
δi1 , . . . , δini, P i (t ≈ t ), ai and δj1 , . . . , δjnj, P i (t ≈ t ), bj .

Concatenating the proofs and using (Tra) we get a proof


δi1 , . . . , δini ,
1: P i (t ≈ t ), ai , proof of P i (t ≈ t ), ai
δj1 , . . . , δjnj ,
2: P i (t ≈ t ), bj , proof of P i (t ≈ t ), bj
3: P i (t ≈ t ), ai ⊗ bj . by (Tra) on 1, 2
Hence, P i (t ≈ t ) is provable in degree at least ai ⊗ bj and so
   
DP (t, t ) ⊗ DP (t , t ) = i∈I ai ⊗ j∈J bj = i∈I j∈J (ai ⊗ bj ) ≤
≤ |P i (t ≈ t )|Γ = DP (t, t ) .
That is, D satisfies (4.48).
(4.49): For a term s resulting from s by substitution of one occurrence
of t in s by t , every proof δi1 , . . . , δini, P i (t ≈ t ), ai can be extended by
(Rep) to a proof
δi1 , . . . , δini ,
1: P i (t ≈ t ), ai , proof of P i (t ≈ t ), ai
2: P i (s ≈ s ), ai . by (Rep) on 1
It follows that DP (t, t ) = |P i (t ≈ t )|Γ ≤ |P i (s ≈ s )|Γ = DP (s, s ).
(4.50): P (t, t ) ≤ |P i (t ≈ t )|Γ = DP (t, t ) since P i (t ≈ t ), P (t, t )
is a proof of length 1. 
(4.51): Let DP (t, t ) = i∈I ai where for each i ∈ I there is a proof
δi1 , . . . , δini, P i (t ≈ t ), ai . Using (Sub), such proofs can be extended to
δi1 , . . . , δini ,
1: P i (t ≈ t ), ai , proof of P i (t ≈ t ), ai
   
2: P (x/r) i t(x/r) ≈ t (x/r) , ai . by (Sub) on 1
 
Hence, DP (t, t ) ≤ DP (x/r) t(x/r), t (x/r) . 


Theorem 4.62. Let L∗ be a complete residuated lattice with a truth stresser ∗ ,


P be a proper family of premises where each P ∈ P is finite and let
  ∗  ∗
P (s, s ) → i∈I bi = i∈I (P (s, s ) → bi ) (4.56)
for any bi ∈ L (i ∈ I = ∅), P ∈ P, and s, s ∈ T (X). Then
|P i (t ≈ t )|Γ = ΓP (t, t ) (4.57)
for every P ∈ P and t, t ∈ T (X).
4.3 Completeness of Fuzzy Horn Logic 211

Proof. “≤”: See Lemma 4.58.


“≥”: Consider D = {DP | P ∈ P}, where DP (t, t ) = |P i (t ≈ t )|Γ for
any P ∈ P, and t, t ∈ T (X). It suffices to show that D satisfies (4.45)–(4.52)
from which SΓ  ≤ D follows immediately by definition of Γ  . Observing
that SΓ  ≤ D means ΓP (t, t ) ≤ |P i (t ≈ t )|Γ , then finishes the proof. By
Lemma 4.61, D satisfies conditions (4.45)–(4.51); it remains to check (4.52).
 
Let P, Q ∈ P and Supp(P ) = {tk , tk | k = 1, . . . , n}. For any t, t ∈ T (X),

we can write DP (t, t ) = i∈I bi , where for each i ∈ I there is a proof
δi1 , . . . , δini, P i (t ≈ t ), bi .

Furthermore, we can write DQ (tk , tk ) = jk ∈Jk ak,jk (k = 1, . . . , n), where for
each jk ∈ Jk , there is a proof
δjk ,1 , . . . , δjk ,njk, Q i (tk ≈ tk ), ak,jk .
Now, we can take any i ∈ I, j1 ∈ J1 , . . . , jn ∈ Jn , concatenate the proofs, and
apply (Mon) to get a proof
δj1 ,1 , . . . , δj1 ,nj1 ,
1: Q i (t1 ≈ t1 ), a1,j1 , proof of Q i (t1 ≈ t1 ), a1,j1
.. .. .. ..
. . . .
δjn ,1 , . . . , δjn ,njn ,
n: Q i (tn ≈ tn ), an,jn , proof of Q i (tn ≈ tn ), an,jn
δi1 , . . . , δini ,
n+1: P i (t ≈ t ), bi , proof of P i (t ≈ t ), bi
 n  ∗ 
n+2: Q i (t ≈ t ), bi ⊗ k=1 P (tk , tk ) → ak,jk

. (Mon) on 1, . . . , n + 1
Hence, we have
n ∗
DQ (t, t ) = |Q i (t ≈ t )|Γ ≥ bi ⊗ k=1 (P (tk , tk ) → ak,jk ) .
Since Q i (tk ≈ tk ), ΓQ (tk , tk ) is a proof, every Jk is non-empty. Thus, ap-
plying (4.56) we get
n  ∗
DP (t, t ) ⊗ k=1 P (tk , tk ) → DQ (tk , tk ) =
 n   ∗
= i∈I bi ⊗ k=1 P (tk , tk ) → jk ∈Jk ak,jk =
 n   ∗
= i∈I bi ⊗ k=1 jk ∈Jk P (tk , tk ) → ak,jk =
  n  ∗
= i∈I bi ⊗ j1 ∈J1 ,...,jn ∈Jn k=1 P (tk , tk ) → ak,jk =
  n  ∗ 
= i∈I bi ⊗ k=1 P (tk , tk ) → ak,jk ≤ DQ (t, t ) ,
j1 ∈J1 ,...,jn ∈Jn

which gives (4.52). The proof is finished. 




Recall that the notions of a semantic and syntactic entailment are de-
termined by L∗ and P. In the sequel, we present results on completeness
dependent on various families P of premises, truth stressers ∗ , and complete
residuated lattices L.
212 4 Fuzzy Horn Logic

General Families of Premises


In this section we focus on P-Horn clauses for general families of premises.
That is, we do not assume any special properties of P. In order to prove
completeness, we restrict ourselves to particular structures of truth degrees.
Theorem 4.63 (completeness for Horn truth stressers). Let L∗ be a
complete residuated lattice with a Horn truth stresser ∗ , P be a proper family
of premises. Then for any L-set Γ of P-Horn clauses we have
|P i (t ≈ t )|Γ = ΓP (t, t ) = ΓP (t, t ) = !P i (t ≈ t )!Γ (4.58)

for any P-Horn clause P i (t ≈ t ).
Proof. Equation (4.56) is satisfied because it is a particularization of (1.21).
Therefore, (4.58) follows from Theorem 4.43, Theorem 4.57, and Theo-
rem 4.62. 

If L is a Noetherian residuated lattice, we can prove completeness of fuzzy
Horn logic for any implicational truth stresser satisfying (1.20). That is, in
this particular case we can avoid (1.21) which was required in Theorem 4.63.
For technical reasons we introduce an additional deduction rule:
P i (t ≈ t ), a , P i (t ≈ t ), b
(Sup) : ,
P i (t ≈ t ), a ∨ b
where P ∈ P, a, b ∈ L, and t, t ∈ T (X). Let R∨ denote the deductive system
which results from the original system R by adding (Sup). The following
lemma shows that if L∗ is a Noetherian residuated lattice then each degree of
R
provability |· · ·|Γ ∨ is fully described by a single proof from R∨ .
Lemma 4.64. Let L∗ be a Noetherian residuated lattice with a truth stresser ∗ .
Then for any L-set Γ of P-Horn clauses,
 R 
Γ R∨ P i (t ≈ t ), |P i (t ≈ t )|Γ ∨
for any P-Horn clause P i (t ≈ t ).
Proof. Recall that {a ∈ L | Γ R∨ P i (t ≈ t ), a } is non-empty. Since L is a
Noetherian residuated lattice, there are truth degrees a1 , . . . , ak ∈ L such that
R
Γ R∨ P i (t ≈ t ), ai (i = 1, . . . , k), and |P i (t ≈ t )|Γ ∨ = a1 ∨ · · · ∨ ak .
Thus, for each i = 1, . . . , k there is a proof δi,1 , . . . , δi,ni , P i (t ≈ t ), ai
from Γ . We can concatenate these proofs and repeatedly use (Sup) to get a
 R 
proof of P i (t ≈ t ), a1 ∨ · · · ∨ ak = |P i (t ≈ t )|Γ ∨ from Γ . 

Theorem 4.65 (completeness for Noetherian residuated lattices). Let
L∗ be a Noetherian residuated lattice with an implicational truth stresser ∗
satisfying (1.20), P be a proper family of premises. Then for any L-set Γ of
P-Horn clauses we have
R
|P i (t ≈ t )|Γ ∨ = !P i (t ≈ t )!Γ
for any P-Horn clause P i (t ≈ t ).
4.3 Completeness of Fuzzy Horn Logic 213

Proof. Using Theorem 4.43 and Theorem 4.57, we only check that
R
|P i (t ≈ t )|Γ ∨ = ΓP (t, t ) .
“≤”: If P i (t ≈ t ), a ∨ b was inferred from weighted P-Horn clauses
P i (t ≈ t ), a and P i (t ≈ t ), b using (Sup), then assuming a ≤ ΓP (t, t )
and b ≤ ΓP (t, t ), we have a ∨ b ≤ ΓP (t, t ). The rest follows from Lemma 4.58.
R
“≥”: Put DP (t, t ) = |P i (t ≈ t )|Γ ∨ for any P-Horn clause P i (t ≈ t ).
We check that D = {DP | P ∈ PFin } satisfies (4.52). Suppose P, Q ∈ PFin and
Supp(P ) = {tk , tk | k = 1, . . . , n}. Take t, t ∈ T (X). Let DP (t, t ) = b, and
DQ (tk , tk ) = ak (k = 1, . . . , n). Lemma 4.64 yields that there are proofs
δ1 , . . . , δn ,P i (t ≈ t ), b , and δk,1 , . . . , δk,nk ,Q i (tk ≈ tk ), ak
for each k = 1, . . . , n. Concatenating the proofs and applying (Mon), we get
n  ∗
DP (t, t ) ⊗ k=1 P (tk , tk ) → DQ (tk , tk ) =
n  ∗ R
= b ⊗ k=1 P (tk , tk ) → ak ≤ |Q i (t ≈ t )|Γ ∨ = DQ (t, t ) .
Hence, D satisfies (4.52); Lemma 4.61 gives that D satisfies (4.45)–(4.52). This
R
yields ΓP (t, t ) ≤ DP (t, t ) = |P i (t ≈ t )|Γ ∨ . 

Example 4.66. The following are examples of L∗ for which we have a complete
fuzzy Horn logic.
(1) If L is a Noetherian residuated lattice then globalization defined on
L is an implicational truth stresser satisfying (1.20) since 1 ⊗ a = 1 ∧ a, and
0 ⊗ a = 0 ∧ a. The usage of globalization as a truth stresser has an important
influence on the deduction rule (Mon). If P (ti , ti )  ai for some i ∈ I, then
the resulting formula Q i (t ≈ t ) is inferred in degree 0 (not interesting). On
the other hand, when P (ti , ti ) ≤ ai for all i ∈ I, Q i (t ≈ t ) is inferred in
degree b. To sum up, for ∗ being the globalization, (Mon) simplifies to
{Q i (ti ≈ ti ), ai ; i = 1, . . . , n} , P i (t ≈ t ), b
(BMon) :  
Q i (t ≈ t ), b
if Supp(P ) = {t1 , t1 , . . . ,tn , tn }, and P (ti , ti ) ≤ ai for each i = 1, . . . , n.
(2) Consider a subalgebra L of the standard product algebra with universe
L = { 21n | n ∈ N0 } ∪ {0}. Since

1 if 21n ≤ 21m ,
2
1
n ⊗ 2
1
m = 2
1
m+n , 2
1
n → 2
1
m = 1
2m−n otherwise ,
L is closed under ⊗ and →. L is a Noetherian residuated lattice which can be
endowed with globalization – a particular case of (1).
(3) Take a Noetherian residuated lattice L such that ⊗ = ∧. That is, L
is a complete Heyting algebra such that the lattice part of L is a Noetherian
lattice. We can define ∗ by a∗ = a (a ∈ L). Trivially, ∗ is an implicational
truth stresser; (1.20) is satisfied since ⊗ = ∧.
(4) Consider L = { n1 | n ∈ N} ∪ {0} endowed with ∧ and ∨ being minimum
and maximum, respectively; ⊗ = ∧, a → b = 1 if a ≤ b, a → b = b else. L =
214 4 Fuzzy Horn Logic

L, ∧, ∨, ⊗, →, 0, 1 is a Noetherian residuated lattice which is a subalgebra


of the standard Gödel algebra. Such an L with globalization or identity is a
particular case of (1) or (3), respectively.
(5) Globalization and identity (on a Heyting algebra) can be seen as two
extreme cases of implicational truth stressers satisfying (1.20). There are,
of course, other examples of implicational truth stressers with this property.
Suppose L is a finite BL-chain. Prelinearity and finiteness of L imply (1.59).
Due to Theorem 1.52, we can define ∗ on L by (1.60), i.e. ∗ sends each element
a ∈ L to the greatest idempotent of L which is less or equal to a. In addition
to that, (1.20) follows from divisibility. Hence ∗ is a Horn truth stresser for L.
(6) Take L1 from example (2) and L2 from example (4). Let ∗1 be glob-
alization on L1 and ∗2 be identity on L2 . Let L∗ be L∗11 × L∗22 . Then ∗ is an
implicational truth stresser on L which satisfies (1.20). Furthermore, L∗11 ×L∗22
is a Noetherian lattice: take {ai , bi ∈ L1 × L2 | i ∈ I} = ∅ since both L1 and
L2 are Noetherian
 lattices, there are a1 , . . . , am ∈ L1 , and b1 , . . . , bn ∈ L2
such that i∈I ai = a1 ∨ · · · ∨ am , and i∈I bi = b1 ∨ · · · ∨ bn . Thus, for a finite
set {ai , bj | i = 1, . . . , m and j = 1, . . . , n} we have
   
i∈I ai , bi = i∈I ai , i∈I bi =
= a1 ∨ · · · ∨ am , b1 ∨ · · · ∨ bn = a1 , b1 ∨ · · · ∨ am , bn .
To sum up, L is a non-linear infinite Noetherian lattice with implicational
truth stresser ∗ satisfying (1.20). Notice that ∗ is neither globalization nor
identity.
(7) Consider a non-linear residuated lattice from Example 1.9 (page 13),
see Fig. 1.2. The left-most truth stresser in Fig. 1.3 (page 16) is an implica-
tional truth stresser satisfying (1.20).

Continuous Fuzzy Horn Logic

In the previous section, we proved completeness over structures of truth de-


grees equipped with Horn truth stressers and over all Noetherian residuated
lattices. These classes of structures of truth degrees do not include, in gen-
eral, structures defined on the real unit interval. For instance, for L being the
standard L  ukasiewicz algebra, there is no Horn truth stresser. In addition to
that, the following example shows that in case of L = [0, 1], fuzzy Horn logic
is not Pavelka-style complete for general proper families of premises even if
we use globalization as the truth stresser. In this section we are going to show
that, considering only particular theories which obey some form of continuity,
fuzzy Horn logic is Pavelka-style complete for any complete residuated lattice
on [0, 1] given by a left-continuous t-norm.

Example 4.67. Let L∗ be a complete residuated lattice on [0, 1] given by a left-


continuous t-norm, ∗ be the globalization on [0, 1]. Take P = LT (X)×T (X) and
consider an L-set Γ of P-Horn clauses such that
4.3 Completeness of Fuzzy Horn Logic 215

Γ (x ≈ y, 0.5 i (x ≈ y)) = 1 for fixed x, y ∈ X ,


Γ (∅ i (xn ≈ yn )) = 2n+1
n
for each n ∈ N ,
Γ (· · ·) = 0 otherwise .
Described verbally, Γ contains identities xn ≈ yn in degrees 2n+1
n
, and a single
P-Horn clause with non-empty premises: x ≈ y, 0.5 i (x ≈ y) (this P-Horn
clause belongs to Γ in degree 1, i.e. it fully belongs to Γ ). Take any model
M ∈ Mod(Γ ). Clearly,
!x ≈ y!M = !xn ≈ yn !M ≥ n
2n+1

for each n ∈ N, i.e. !x ≈ y!M ≥ 0.5. Since !x ≈ y, 0.5 i (x ≈ y)!M = 1, for
any valuation v : X → M we get

!x ≈ y!M,v = 1 → !x ≈ y!M,v = (0.5 → !x ≈ y!M,v ) → !x ≈ y!M,v =
= !x ≈ y, 0.5 i (x ≈ y)!M,v = 1 .
Hence, !x ≈ y!M = 1 for any model M ∈ Mod(Γ ). This yields !x ≈ y!Γ = 1.
Theorem 4.43 and Theorem 4.57 thus give Γ∅ (x, y) = 1.
We now show that the provability degree |x ≈ y|Γ is strictly lower than 1.
Evidently, |x ≈ y|Γ ≥ 0.5 due to (Sub) applied on xn ≈ yn , 2n+1
n
(n ∈ N). On
the other hand, if Γ  x ≈ y, a then a < 0.5. Indeed, if Γ R x ≈ y, a
R

then there is a finite Γ  ⊆ Γ such that Γ  R x ≈ y, a because in any


weighted proof from R one can use only finitely many weighted P-Horn clauses
of the form Pi (t ≈ t ), Γ(P i (t ≈ t )) . For Γ  we can consider an L-
algebra M = M, ≈M , F M with M = {a , b }; functions f M ∈ F M are
defined by f M (· · ·) = a ; a ≈M b = 2m+1
m
for some m ∈ N satisfying Γ  (xn ≈
yn ) = 0 for all n ≥ m (m always exists since Γ  is finite). Then
!xn ≈ yn !M = m
2m+1 ≥ Γ  (xn ≈ yn )
for each n ∈ N, and !x ≈ y, 0.5 i (x ≈ y)!M,v = 1 for any v because either
!x!M,v = !y!M,v , or !x ≈ y!M,v = 2m+1m
< 0.5 and so

!x ≈ y, 0.5 i (x ≈ y)!M,v = (0.5 → !x ≈ y!M,v ) → !x ≈ y!M,v =

= (0.5 → m
2m+1 ) → !x ≈ y!M,v =
= 0 → !x ≈ y!M,v = 1

because is the globalization on [0, 1]. Thus,
1 = !x ≈ y, 0.5 i (x ≈ y)!M ≥ Γ  (x ≈ y, 0.5 i (x ≈ y)) .
As a consequence, M ∈ Mod(Γ  ), and !x ≈ y!M = 2m+1 m
< 0.5. Since fuzzy
Horn logic is sound, we get a ≤ |x ≈ y|Γ  < 0.5. So, we have shown that if
Γ R x ≈ y, a then a < 0.5. Using this observation, we readily obtain
 
|x ≈ y|Γ = {a ∈ L | Γ R x ≈ y, a } ≤ {a ∈ L | a < 0.5} = 0.5 .
Therefore, |x ≈ y|Γ = 0.5 < Γ∅ (x, y) = !x ≈ y!Γ = 1. So, in this particular
case, fuzzy Horn logic is Pavelka-sound but it is not Pavelka-complete.
216 4 Fuzzy Horn Logic

For technical reasons, we define a type of continuity of an L-set of P-Horn


clauses using an arbitrary but fixed strict continuous Archimedean t-norm.

Definition 4.68. Let L be a complete residuated lattice on [0, 1] given by a


left-continuous t-norm ⊗. Let " be a strict continuous Archimedean t-norm.
For P ∈ LT (X)×T (X) and d ∈ [0, 1) we define d"P ∈ LT (X)×T (X) by
(d"P )(r, r ) = d " P (r, r ) (4.59)
for all r, r ∈ T (X).

Remark 4.69. (1) Note that in general, " has nothing to do with ⊗. We just
use two conjunctions simultaneously: the left-continuous t-norm ⊗ determines
the structure of truth degrees L while the strict continuous t-norm " deter-
mines the notion of "-continuity.
(2) Note that if P i (t ≈ t ) is a P-Horn clause for P = LT (X)×T (X) then
so is d"P i (t ≈ t ) because d"P ∈ LT (X)×T (X) is finite. On the other hand,
there are proper families of premises such that for some P ∈ P and d ∈ [0, 1),
we have d"P ∈ P. This is the case of e.g. families of crisp premises.

From this moment on, we assume P = LT (X)×T (X) ; L is a complete resid-


uated lattice on the real unit interval given by a left-continuous t-norm; ∗ is
the globalization on [0, 1].

Definition 4.70. An L-set Γ of P-Horn clauses is called "-continuous if for


each P-Horn clause P i (t ≈ t ) and each e ∈ [0, 1) there is d ∈ [0, 1) such
that
IF Γ R P i (t ≈ t ), b ,
THEN there is be ∈ [e " b, 1] such that Γ R d"P i (t ≈ t ), be .

For "-continuous L-sets we have the following characterization.

Theorem 4.71 (completeness of continuous FHL). Let L∗ be a complete


residuated lattice on [0, 1] given by a left-continuous t-norm ⊗. Let ∗ be the
globalization on [0, 1], " be a strict continuous Archimedean t-norm, P =
LT (X)×T (X) . Then for any "-continuous L-set Γ of P-Horn clauses,
|P i (t ≈ t )|Γ = !P i (t ≈ t )!Γ
for each P-Horn clause P i (t ≈ t ).

Proof. According to Theorem 4.59, we check that ΓP (t, t ) ≤ |P i (t ≈ t )|Γ .


Put DP (t, t ) = |P i (t ≈ t )|Γ for every P ∈ P, and t, t ∈ T (X). We check
that D = {DP | P ∈ P} satisfies (4.45)–(4.52). We focus only on (4.52), the
rest follows from Lemma 4.61. Since ∗ is globalization, it suffices to show that
IF P (s, s ) ≤ DQ (s, s ) for all s, s ∈ T (X) such that P (s, s ) > 0,
THEN DP (t, t ) ≤ DQ (t, t ) .
4.3 Completeness of Fuzzy Horn Logic 217

So, let P (s, s ) ≤ DQ (s, s ) for any s, s ∈ T (X) such that P (s, s ) > 0 and let
Γ R P i (t ≈ t ), b . Since Γ is "-continuous, for each e ∈ [0, 1) there is
d ∈ [0, 1) such that we have Γ R d"P i (t ≈ t ), be for some be ≥ e " b.
Furthermore, if P (s, s ) > 0 then
(d"P )(s, s ) < P (s, s ) ≤ DQ (s, s ) = |Q i (s ≈ s )|Γ .
Therefore, for any s, s ∈ T (X) satisfying P (s, s ) > 0 we have
Γ R Q i (s ≈ s ), as,s ,
where as,s ≥ (d"P )(s, s ). Thus, applying (Mon) on
{Q i (s ≈ s ), as,s | P (s, s ) > 0} and d"P i (t ≈ t ), be ,
we get
Γ R Q i (t ≈ t ), be .
That is, for each e ∈ [0, 1), we have Γ R Q i (t ≈ t ), be where b " e ≤ be .
So, for each e ∈ [0, 1), b " e ≤ |Q i (t ≈ t )|Γ = DQ (t, t ), i.e. b ≤ DQ (t, t ) by
Lemma 1.35 (i). Hence, we immediately obtain DP (t, t ) ≤ DQ (t, t ). 


Example 4.72. (1) Let L be a residuated lattice on [0, 1] with ⊗ = ∧. One


can prove by induction on the length of a proof that if for each d ∈ [0, 1) and
P i (t ≈ t ) we have d " Γ (P i (t ≈ t )) ≤ Γ (d"P i (t ≈ t )), then Γ is
"-continuous. Indeed, by induction on length of a proof we can prove
IF Γ R P i (t ≈ t ), b ,
THEN there is bd ∈ [d " b, 1] such that Γ R d"P i (t ≈ t ), bd
for each d ∈ [0, 1) from which the "-continuity of Γ follows immedi-
ately. Let P i (t ≈ t ), b be a member of a proof from Γ using R. If
b = Γ (P i (t ≈ t )) then Γ R d"P i (t ≈ t ), bd for
bd = Γ (d"P i (t ≈ t )) ≥ d " Γ (P i (t ≈ t )) = d " b ,
i.e. bd ∈ [d"b, 1]. The claim is routine to check if P i (t ≈ t ), b was inferred
using (Ref), (Sym), (Rep), (Ext), or (Mon). If P i (t ≈ t ), b was inferred
from P i (t ≈ s), a and P i (s ≈ t ), a by (Tra) then b = a⊗a ; by induc-
tion hypothesis, Γ R d"P i (t ≈ s), ad and Γ R d"P i (s ≈ t ), ad
where ad ≥ d " a and ad ≥ d " a . Therefore, one can use (Tra) to conclude
that Γ R d"P i (t ≈ t ), ad ⊗ ad where
ad ⊗ ad = ad ∧ ad ≥ (d " a) ∧ (d " a ) = d " (a ∧ a ) = d " (a ⊗ a ) ,
i.e. ad ⊗ ad ∈ [d " (a ⊗ a ), 1]. If P i (t ≈ t ), b results from Q i (s ≈ s ), b
by (Sub) then there is a substitution (x/r) such that P = Q(x/r), t = s(x/r),
and t = s (x/r). Observe that
   
(d"Q)(x/r) (t, t ) = s(x/r) = t (d"Q)(s, s ) = d " s(x/r) = t Q(s, s ) =
s (x/r) = t s (x/r) = t
  
= d " Q(x/r)(t, t )
218 4 Fuzzy Horn Logic

for any t, t ∈ T (X). Therefore, we have (d"Q)(x/r) = d"(Q(x/r)). By in-


duction hypothesis, Γ R d"Q i (s ≈ s ), bd , where bd ∈ [d " b, 1]. Hence,
using (Sub) we get Γ R (d"Q)(x/r) i (s(x/r) ≈ s (x/r)), bd from which it
follows that Γ R d"(Q(x/r)) i (s(x/r) ≈ s (x/r)), bd . As a consequence,
Γ is "-continuous.
(2) If ⊗ = ∧ then we can use (1) to get that each L-set Γ of P-Horn clauses
where Γ (P i (t ≈ t )) = 0 if P = ∅ is "-continuous.
(3) Let F be a type which consists of a single binary function symbol f ,
Γ be an L-set of P-Horn clauses, where Γ (x ◦ z ≈ y ◦ z, a i x ≈ y) = a for
each a ∈ [0, 1], and Γ (P i (t ≈ t )) = 0 otherwise. If ⊗ = ∧, (1) gives that
Γ is "-continuous. Γ can be seen as a theory of grupoids (understood as
L-algebras) satisfying a particular type of continuous cancellation.
(4) There are, of course, L-sets of P-Horn clauses which are not "-
continuous. For instance, consider Γ such that Γ (x ≈ y, 0.5 i (x ≈ y)) = 1,
and Γ (P i (t ≈ t )) = 0 otherwise. For e = 0.9 there is not any d ∈ [0, 1)
such that Γ R x ≈ y, d " 0.5 i (x ≈ y), be for be ≥ e" 1 = 0.9, because
for each d ∈ [0, 1) we can consider an L-algebra M = M, ≈M , ∅ (of the
empty type) with M = {a , b } and a ≈M b = d " 0.5 which is a model of
such Γ but for a valuation v : X → M with v(x) = a and v(y) = b we
have !x ≈ y, d " 0.5 i (x ≈ y)!M,v = d " 0.5  0.9 = be . Therefore, the
soundness of fuzzy Horn logic (see Theorem 4.59) gives
|x ≈ y, d " 0.5 i (x ≈ y)|Σ < be ,
meaning that x ≈ y, d " 0.5 i (x ≈ y) is not provable from Γ using R in
degree at least be . As a consequence, Γ is not "-continuous.

Families with Restricted Premises

So far, we were considering P-Horn clauses without additional restrictions on P


(sometimes, we used P = LT (X)×T (X) ). From now on, we will restrict ourselves
to P-Horn clauses with P being a particular (restricted) proper family of
premises. Our motivation is basically to show that restricting ourselves to
simpler families of premises we can obtain Pavelka-style complete fuzzy Horn
logic for wider classes of structures of truth degrees. We will be interested
mainly in families of crisp premises, i.e. families where every P ∈ P is crisp.
Notice that crisp premises naturally correspond with premises of ordinary
Horn clauses. Later on, we comment on a way to generalize this result for
P-Horn clauses with K-valent premises.
In the sequel, we let P be defined by
P = {P ∈ LT (X)×T (X) | P is finite and crisp} .
Recall that in such a case, the notion of an endomorphic image coincides with
the classical one, i.e. h(P ) = {h(t), h(t ) | t, t ∈ P } for every P ∈ P and
arbitrary endomorphism h on T(X). Since every P ∈ P is crisp, we either
4.3 Completeness of Fuzzy Horn Logic 219

have P (s, s ) = 1, or P (s, s ) = 0. Thus, the degree to which a P-Horn clause


P i (t ≈ t ) is true in an L-algebra M under a valuation v simplifies to
 ∗
!P i (t ≈ t )!M,v = 
s,s ∈P !s ≈ s !M,v → !t ≈ t !M,v .
Using Corollary 4.41, and Theorem 4.43, we can describe a semantic clo-
sure of SΓ , where Γ is any L-set of P-Horn clauses with crisp premises. As
we have shown before, the semantic closure SΓ  of SΓ is the least P-indexed
system of congruences satisfying (4.32)–(4.34) such that SΓ ≤ SΓ  . For crisp
premises, conditions (4.32) and (4.34) simplify to
t, t ∈ P implies ΓP (t, t ) = 1 , (4.60)
 
 ∗
s,s ∈P ΓQ (s, s ) ≤ S(ΓP , ΓQ ) . (4.61)
Moreover, (4.61) is equivalent to
 
 ∗
ΓP (t, t ) ⊗ s,s ∈P ΓQ (s, s ) ≤ ΓQ (t, t ) (4.62)
which is required to be true for any P, Q ∈ P and t, t ∈ T (X).
Now we turn our attention to the notion of a provability degree. In the
case of crisp premises, properties (4.50) and (4.52) of a deductive closure SΓ 
can be equivalently reformulated:
t, t ∈ P implies ΓP (t, t ) = 1 , (4.63)
 ∗
ΓP (t, t ) ⊗ s,s ∈P ΓQ (s, s ) ≤ ΓQ (t, t ) , (4.64)
where P, Q ∈ P, and t, t ∈ T (X). In order to prove Γ  = Γ  , it is sufficient to
assume L∗ to be a complete residuated lattice equipped with an implicational
truth stresser ∗ satisfying (1.20), see Theorem 4.57. So, it remains to connect
the deductive closure with degree of provability. Consider deductive system
R as it has been introduced in Sect. 4.1. (Ext) and (Mon) simplify to
(CExt) : P i (t ≈ t ), 1 for every t, t ∈ P ,

{Q i (ti ≈ ti ), ai ; i = 1, . . . , n}, P i (t ≈ t ), b


(CMon) : ,
Q i (t ≈ t ), b ⊗ a∗1 ⊗ · · · ⊗ a∗n
where P, Q are crisp sets of premises, P = {t1 , t1 , . . . , tn , tn }, t, t ∈ T (X),
and a1 , . . . , an , b ∈ L. The following assertion establishes a connection between
a deductive closure and a degree of provability.

Theorem 4.73. Let Γ be an L-set of P-Horn clauses such that each P ∈ P


is crisp. If ∗ a truth stresser satisfying (1.62) then
|P i (t ≈ t )|Γ = ΓP (t, t ) (4.65)

for each P-Horn clause P i (t ≈ t ).

Proof. “≤”: Apply Lemma 4.58.


“≥”: Since each P ∈ P is crisp, P (s, s ) ∈ {0, 1} for any s, s ∈ T (X). If
P (s, s ) = 0 then (4.56) is satisfied trivially. If P (s, s ) = 1, we get
220 4 Fuzzy Horn Logic
  ∗  ∗   ∗
P (s, s ) → i∈I bi = i∈I bi = i∈I b∗i = i∈I (P (s, s ) → bi )
on account of (1.62). Now apply Theorem 4.62. 


Summarizing previous observations, we get the following

Corollary 4.74 (completeness for crisp premises). Let L∗ be a complete


residuated lattice equipped with an implicational truth stresser ∗ satisfying
(1.20), (1.62), P be a proper family of premises such that each P ∈ P is crisp.
For every L-set Γ of P-Horn clauses,
|P i (t ≈ t )|Γ = ΓP (t, t ) = ΓP (t, t ) = !P i (t ≈ t )!Γ
for any P-Horn clause P i (t ≈ t ) 


Example 4.75. The following are examples of L∗ for which we have a complete
fuzzy Horn logic with crisp premises.
(1) Let L be any complete residuated lattice such that 1 (the greatest ele-
ment of L) is ∨-irreducible. Then, globalization defined on a complete resid-
uated lattice with ∨-irreducible 1 is an implicational truth stresser satisfying
(1.20) and (1.62), see Example 1.57 (2).
(2) For any complete residuated lattice L we can consider L ⊕ 2. In fact,
L ⊕ 2 results from L by equipping it with a new top element, see Fig. 1.6,
which is then ∨-irreducible. That is, L ⊕ 2 with globalization is a particular
case of (1). The new complete residuated lattice L ⊕ 2 is “rather similar” to
the starting one.
(3) Complete residuated lattices with implicational truth stressers satis-
fying (1.20) and (1.62) can be ordinally added yielding new structures with
nontrivial implicational truth stressers satisfying (1.20) and (1.62), see Theo-
rem 1.60 and Fig. 1.7.

Remark 4.76. Let us discuss some epistemic impacts of adding a new top
element to a complete residuated lattice. First, we may ask about the nature
of the structures with ∨-irreducible greatest element. The top element of a
complete residuated lattice represents full truth, while other elements can be
thought of as degrees of partial truths, but not the full truth. From this point of
view, it might be not natural to allow the fully true statement to be obtained
from partially true statements. On the level of the structure of truth degrees,
this corresponds to ∨-irreducibility of the greatest element, i.e. 1 cannot be
obtained as a supremum of truth degrees which differ from 1.
From the point of view of Pavelka-style logic, if we use L with ∨-irreducible
1 as the structure of truth degrees, we have to find a weighted proof of ϕ, 1
in order to show that the provability degree of ϕ equals 1. In other words,
|ϕ|R
Γ = 1 implies that there is a weighted proof of ϕ, 1 from Γ using R.

Remark 4.77. Obviously, if each P ∈ P is crisp and if Γ is a crisp set of P-Horn


clauses then Γ can be seen as an ordinary set of the classical Horn clauses.
Furthermore, (Ref)–(Mon) collapse to the deduction rules of the classical Horn
4.4 Fuzzy Equational Logic Revisited 221

logic [97]. It is then easily seen that Γ R P i (t ≈ t ), a yields a ∈ {0, 1}


and, more importantly, |P i (t ≈ t )|Γ = 1 iff P i (t ≈ t ) is provable from Γ
in the classical sense, see [97]. So, analogously as in the comment on page 146,
in case of crisp sets of P-Horn clauses with crisp premises, the notion of graded
provability can be fully described by the classical one.

∗ ∗ ∗
The idea of having crisp premises can be naturally generalized. Let L∗ be a
complete residuated lattice with implicational truth stresser satisfying (1.20).
Take K ⊆ L such that {0, 1} ⊆ K. A proper family of premises P can be
called K-valent if {P (s, s ) | s, s ∈ T (X)} ⊆ K for each P ∈ P. Obviously, if
P is {0, 1}-valent then each P ∈ P is crisp. Now, if L∗ satisfies (1.21) for any
a ∈ K, then |P i (t ≈ t )|Γ = !P i (t ≈ t )!Γ for any P-Horn  clause due to
Theorem 4.43, Theorem 4.57, and Theorem 4.62. For instance, i∈I (Li ⊕ 2)
(see Fig. 1.7), equipped with ∗ which is an ordinal sum of globalizations
 on

Li ⊕ 2 (i ∈ I), satisfies (1.21) for any a ∈ K, where K = {00 } ∪ 1i | i ∈ I .
So, for i∈I (Li ⊕ 2) we have Pavelka-style complete fuzzy Horn logic which
uses P-Horn clauses with K-valent premises.

4.4 Fuzzy Equational Logic Revisited


The main aim of this section is to look again at fuzzy equational logic which
has been introduced in Chap. 3. We are going to show how the results of fuzzy
Horn logic generalize the results of fuzzy equational logic.
First, we show that the completeness of fuzzy equational logic can be
easily derived from our results on fuzzy Horn logic. Recall that if we deal
with identities, we can put P = {∅}. In this case, P-implications are exactly
formulas of the form ∅ i (t ≈ t ), i.e. P-implications with empty premises. We
have !∅ i (t ≈ t )!M,v = !t ≈ t !M,v for any M and v. Hence, from the point
of view of interpretation of formulas, we can identify t ≈ t with ∅ i (t ≈ t ),
see Remark 4.10 (4). In Sect. 3.2, we introduced fully invariant congruences as
algebraic counterparts of semantically closed L-sets of identities. Now, in the
context of fuzzy Horn logic, we can consider SΣ , which is a one-element system
SΣ = {Σ∅ }, where Σ∅ = Σ. Hence, due to Corollary 4.41 and  Theorem
 4.43,
the semantic closure of SΣ is a one-element system SΣ  = Σ∅ , where Σ∅
is the least congruence on T(X) which contains Σ and SΣ  satisfies (4.32)–
(4.34). Equation (4.32), i.e. ∅ ⊆ Σ∅ , is a trivial property. Equation (4.34)
holds trivially as well because S(Σ∅ , Σ∅ ) = 1. Thus, the only one nontrivial
property of SΣ  is that of stability, i.e. Σ∅ (t, t ) ≤ Σ∅ (h(t), h(t )), which is but
the full-invariance of Σ∅ . To sum up, Σ∅ is the least fully invariant congruence
on T(X) which contains Σ. Hence, up to a slightly different formalism, the
representations of semantic entailment from Sect. 3.2 and Sect. 4.2 coincide
if P = {∅}. As a consequence, !t ≈ t !Σ = !t ≈ t !T(X)/Σ  . As to the notion

222 4 Fuzzy Horn Logic

of provability, deduction rules (ERef)–(ESub), introduced in Sect. 3.1, are


instances of (Ref)–(Sub) for P = ∅. Furthermore, for P = {∅}, both rules (Ext)
and (Mon) can be omitted. Namely, (Ext) yields t ≈ t , 0 and (Mon) derives
t ≈ t , b from t ≈ t , b . It is thus easily seen that the degree of provability
of t ≈ t from an L-set of identities as defined in Sect. 3.1 coincides with
the degree of provability of ∅ i (t ≈ t ) from the L-set Σ  of P-implications
where Σ  (∅ i (t, t )) = Σ(t, t ). Analogously as for semantic closure, P = {∅}
leads to a simplified notion of a deductive closure Σ  which in fact coincides
with that presented in Sect. 3.2. Lemma 4.58 gives |t ≈ t |Σ ≤ Σ∅ (t, t ). The
converse inequality follows from Theorem 4.62 because in case of P = ∅,
(4.56) is satisfied trivially. We thus have |t ≈ t |Σ = Σ∅ (t, t ) for any complete
residuated lattice L because any L can be equipped with an implicational
truth stresser satisfying (1.20), e.g. globalization on L. Finally,
|t ≈ t |Σ = Σ∅ (t, t ) = !t ≈ t !T(X)/Σ  = !t ≈ t !Σ

is a consequence of Theorem 4.57. This is just the same completeness result


as presented in Sect. 3.2.
Another way of looking at fuzzy equational logic inside fuzzy Horn logic is
the following. Consider any proper family P with ∅ ∈ P, and let Σ be an L-set
of P-Horn clauses such that for any ∅ = P ∈ P we have Σ(P i (t ≈ t )) = 0.
Thus, Σ can be seen as an L-set of identities. Using Theorem 3.31, |t ≈ t |Σ =
!t ≈ t !Σ . Soundness of fuzzy Horn logic gives
R
|∅ i (t ≈ t )|Σ ≤ !∅ i (t ≈ t )!Σ = !t ≈ t !Σ = |t ≈ t |Σ ,
where R denotes the deductive system of fuzzy Horn logic. Since R contains
all instances of (ERef)–(ESub) for P = ∅, we get |∅ i (t ≈ t )|R 
Σ ≥ |t ≈ t |Σ .
 R 
So, we have |∅ i (t ≈ t )|Σ = |t ≈ t |Σ which has the following interpretation:
the degree to which t ≈ t is provable using the original deductive system of
fuzzy equational logic equals to the degree to which t ≈ t is provable using
the deductive system of fuzzy Horn logic. From this point of view, fuzzy Horn
logic is a kind of a conservative extension of fuzzy equational logic – if we
focus on provability (using rules of fuzzy Horn logic) of identities from L-
sets of identities, we cannot infer more that can be inferred by the deductive
system of fuzzy equational logic.

4.5 Implicationally Defined Classes of L-Algebras


In this and further sections we turn our attention to the problem of closure
properties of implicationally defined classes of L-algebras. Our development
is motivated by the following question: “How does a choice of L, ∗ , and P
affect closure properties of the model classes?” Such a question in nontrivial
from the very beginning. We already know that a generalization of Birkhoff’s
variety theorem, which was presented in Sect. 3.4, can be established for any
complete residuated lattice taken as the structure of truth degrees. If we take
4.5 Implicationally Defined Classes of L-Algebras 223

into consideration P-implications as our basic formulas, the situation gets


more complicated.
In the classical case, closure properties of implicational classes are influ-
enced by the structure (complexity) of premises of the implications in ques-
tion. It is not surprising that in fuzzy setting the structure of premises will
play an analogous role. We are therefore going to investigate closure proper-
ties in dependence on particular proper families of premises. However, unlike
the classical case, we have to take into account also a structure L of truth
degrees equipped with a truth stresser ∗ . The subsequent development will
show how the choice of L∗ and P influences the closure properties of model
classes and under what conditions the appropriate closure properties imply
the definability by L-sets of P-implications.
In the sequel, we will use the notion of a P-implication which has been
introduced in Sect. 4.1, see Definition 4.4. Recall that the interpretation of a
P-implication is always considered with respect to a particular truth stresser ∗
for L, see Definition 4.9.
We start by investigating closure properties of implicationally defined
classes of L-algebras. Note again that in the ordinary case, closure proper-
ties depend on the complexity of sets of premises we take into consideration.
For instance, universal Horn classes (determined by formulas with finite sets
of premises) are known to be closed under direct limits while general implica-
tional classes (determined by generalized formulas with infinite sets premises)
are not, see [97]. In what follows, we present an analogy to those closure
properties in dependence on a given proper family of premises.

Lemma 4.78. Let L∗ be a complete residuated lattice with a truth stresser ∗ .


For M ∼
= N we have
!P i (t ≈ t )!M = !P i (t ≈ t )!N (4.66)

for every P-implication P i (t ≈ t ).

Proof. Let M ∼ = N. Thus, there is an isomorphism h : M → N. For any


valuation v in M we can take valuation v ◦ h in N. We have !t ≈ t !M,v =
!t ≈ t !N,v◦h for all t, t ∈ T (X). Conversely, for any valuation v in N we can
take v ◦ h−1 so that !t ≈ t !N,v = !t ≈ t !M,v◦h−1 . This immediately gives

!P i (t ≈ t )!M = v:X→M !P i (t ≈ t )!M,v =

= v:X→N !P i (t ≈ t )!N,v = !P i (t ≈ t )!N ,
proving (4.66). 


Lemma 4.79. Let L∗ be a complete residuated lattice with a truth stresser ∗ .


If N is a subalgebra of an L-algebra M then
!P i (t ≈ t )!M ≤ !P i (t ≈ t )!N (4.67)
for every P-implication P i (t ≈ t ).
224 4 Fuzzy Horn Logic

Proof. Since N ∈ Sub(M), every valuation v in N is a valuation in M as


well. Moreover, Lemma 3.46 yields !P i (t ≈ t )!M,v = !P i (t ≈ t )!N,v
for every valuation v : X → N . Thus,

!P i (t ≈ t )!M = v:X→M !P i (t ≈ t )!M,v ≤

≤ v:X→N !P i (t ≈ t )!M,v =

= v:X→N !P i (t ≈ t )!N,v = !P i (t ≈ t )!N ,
which is the desired inequality. 


Lemma 4.80. Let LQ be a complete residuated lattice with an implicational

truth stresser . Let i∈I Mi be the direct product of {Mi | i ∈ I}. Then
!P i (t ≈ t )!{Mi | i∈I} ≤ !P i (t ≈ t )!Q (4.68)
i∈I Mi

for every P-implication P i (t ≈ t ).


Q
Proof. First, using Lemma 3.46 (iii), for every valuation v of X in i∈I Mi
and arbitrary terms r, r ∈ T (X) it follows that
Q
!r ≈ r !Q Mi ,v = !r!Q Mi ,v ≈ i∈I Mi !r !Q Mi ,v =
  
i∈I i∈I i∈I
 Q
= i∈I !r! Q (i) ≈ !r !
Mi
(i) =
i∈I Mi ,v i∈I Mi ,v
   
= i∈I !r!Mi ,v◦πi ≈Mi !r !Mi ,v◦πi = i∈I !r ≈ r !Mi ,v◦πi .
Thus, using this fact together with property (1.19) of implicational truth
stressers and (1.39), we have

i∈I !P !Mi ,v◦πi =
    
∗
= i∈I s,s ∈T (X) P (s, s ) → !s ≈ s !Mi ,v◦πi =
    
 ∗
= i∈I s,s ∈T (X) P (s, s ) → !s ≈ s !Mi ,v◦πi =
    
∗
= s,s ∈T (X) i∈I P (s, s ) → !s ≈ s !Mi ,v◦πi =
    
 ∗
= s,s ∈T (X) P (s, s ) → i∈I !s ≈ s !Mi ,v◦πi =
    Q
∗
= s,s ∈T (X) P (s, s ) → !s ≈ s !i∈I Mi ,v = !P !Q Mi ,v .
i∈I

Now, Qit is essential to observe that the i-th projection v ◦ πi of a valuation v :


X → i∈I M i is a valuation in Mi . Conversely,
Q for any valuation vi : X → Mi
there is at least one valuation v : X → i∈I M i such that v  ◦ πi = vi . Finally,


using (1.44) it follows that



!P i (t ≈ t )!{Mi | i∈I} = i∈I !P i (t ≈ t )!Mi =
 
= i∈I vi :X→Mi !P i (t ≈ t )!Mi ,vi =
 
= v:X→Q Mi i∈I !P i (t ≈ t )!Mi ,v◦πi =
   
i∈I

= v:X→Q Mi i∈I !P !Mi ,v◦πi → !t ≈ t !Mi ,v◦πi ≤



i∈I
  
≤ v:X→Q Mi i∈I !P !Mi ,v◦πi → i∈I !t ≈ t !Mi ,v◦πi =
i∈I
4.5 Implicationally Defined Classes of L-Algebras 225
  
= Q
v:X→ i∈I Mi !P !Q → !t ≈ t !Q =
i∈I Mi ,v i∈I Mi ,v

= Q
v:X→ i∈I Mi !P i (t ≈ t )!Q = !P i (t ≈ t )!Q ,
i∈I Mi ,v i∈I Mi

i.e. we have proved (4.68). 



Hence, we have the following consequence for complete residuated lattices
equipped with implicational truth stressers.
Corollary 4.81. Let L∗ be a complete
Q residuated lattice with an implicational
truth stresser ∗ . Let M ∈ Sub( i∈I Mi ) be a subdirect product of a family
{Mi | i ∈ I} of L-algebras. Then
!P i (t ≈ t )!{Mi | i∈I} ≤ !P i (t ≈ t )!M (4.69)
for every P-implication P i (t ≈ t ). 

Recall that abstract classes of L-algebras are classes closed under isomor-
phic images, see Definition 2.137 on page 127.
Theorem 4.82. Let L∗ be a complete residuated lattice with an implicational
truth stresser ∗ . For an L-set Σ of P-implications, Mod(Σ) is an abstract class
of L-algebras closed under the formations of subalgebras and direct products.
Proof. For any M ∈ I(Mod(Σ)) there is M ∈ Mod(Σ) such that M ∼ = M .
  
By (4.66), Σ(P i (t ≈ t )) ≤ !P i (t ≈ t )!M = !P i (t ≈ t )!M for every
P-implication P i (t ≈ t ). Thus, Mod(Σ) is closed under isomorphic images,
i.e. Mod(Σ) is an abstract class of L-algebras.
“S(Mod(Σ)) ⊆ Mod(Σ)”: Let N ∈ S(Mod(Σ)), i.e. there is M ∈ Mod(Σ)
where N ∈ Sub(M). Using (4.67) we have
Σ(P i (t ≈ t )) ≤ !P i (t ≈ t )!M ≤ !P i (t ≈ t )!N
for all P-implications P i (t ≈ t ), whence N ∈QMod(Σ).
“P(Mod(Σ)) ⊆ Mod(Σ)”: Analogously, for i∈I Mi ∈ P(Mod(Σ)), where
{Mi | i ∈ I} ⊆ Mod(Σ), we can use (4.68) to get
Σ(P i (t ≈ t )) ≤ !P i (t ≈ t )!{Mi | i∈I} ≤ !P i (t ≈ t )!Q Mi
i∈I
Q
for all P-implications P i (t ≈ t ), i.e. i∈I Mi ∈ Mod(Σ). 

Theorem 4.83. Let L∗ be a complete residuated lattice with an implicational
truth stresser ∗ . For a class K of L-algebras we have
Impl(K) = Impl(I(K)) = Impl(S(K)) = Impl(P(K)) .
Proof. Put Σ = Impl(K). Theorem 4.82 yields O(Mod(Σ)) ⊆ Mod(Σ) for O
being any of I, S, or P. Since O(K) ⊆ O(Mod(Σ)) and Σ = Impl(Mod(Σ)),
Impl(K) = Σ = Impl(Mod(Σ)) ⊆ Impl(O(Mod(Σ))) ⊆ Impl(O(K)) .
Conversely, K ⊆ O(K) for O being I, S, or IP. Thus, Impl(O(K)) ⊆ Impl(K)
for O being I, S, or IP. Furthermore, Impl(P(K)) = Impl(IP(K)) ⊆ Impl(K).


226 4 Fuzzy Horn Logic

In much the same way as in the ordinary case, one may easily show that
P-implicational classes are not closed under (arbitrary) homomorphic images.
We will demonstrate this later on.

Lemma
 4.84. Let L∗ be a complete residuated lattice with a truth stresser ∗ .
Let i∈I Mi be a direct union of a directed family {Mi | i ∈ I} of L-algebras.
Then
!P i (t ≈ t )!{Mi |i∈I} ≤ !P i (t ≈ t )! Mi (4.70)
i∈I

for every P-finitary implication P i (t ≈ t ).

Proof. Since P i (t ≈ t ) is supposed to be a P-finitary implication, Y =


var(P
 ) ∪ var(t) ∪ var(t ) is a finite set. Thus, for every valuation v of X in
 
i∈I Mi , we canconsider a finite set M = {v(x) | x ∈ Y }. Moreover, M is a

finite subset of i∈I Mi , i.e. there is some index k ∈ I such that M ⊆ Mk .
Hence the restriction vY of v on Y can be thought of as a valuation in Mk ∈
Sub( i∈I Mi ). Thus, Lemma 3.46 yields
!r ≈ r ! Mi ,vY = !r ≈ r !Mk ,vY
i∈I

for every r, r ∈ Supp(P ) ∪ {t, t }. Furthermore,


!P i (t ≈ t )!{Mi | i∈I} ≤ !P i (t ≈ t )!Mk ,vY =
= !P i (t ≈ t )! Mi ,vY = !P i (t ≈ t )! Mi ,v .
i∈I i∈I

As a consequence, !P i (t ≈ t )!{Mi | i∈I} ≤ !P i (t ≈ t )! Mi . 



i∈I

Theorem 4.85. Let L∗ be a complete residuated lattice with an implicational


truth stresser ∗ . For an L-set Σ of P-finitary implications, Mod(Σ) is an
abstract class of L-algebras closed under the formations of subalgebras, direct
products, and direct unions.

Proof. Let i∈I Mi ∈ U(Mod(Σ)), where {Mi ∈ Mod(Σ) | i ∈ I} is a directed
family of L-algebras. Now, using (4.70) it follows that
Σ(P i (t ≈ t )) ≤ !P i (t ≈ t )!{Mi |i∈I} ≤ !P i (t ≈ t )! Mi ,
i∈I



for every P-finitary implication P i (t ≈ t ). Hence, i∈I Mi ∈ Mod(Σ). The
rest follows from Theorem 4.82. 


Theorem 4.86. Let L∗ be a complete residuated lattice with an implicational


truth stresser ∗ . For a class K of L-algebras we have
Implω(K) = Implω(I(K)) = Implω(S(K)) = Implω(P(K)) = Implω(U(K)) .

Proof. Since K ⊆ U(K), we have Implω(U(K)) ⊆ Implω(K). Conversely, The-


orem 4.85 yields U(K) ⊆ U(Mod(Implω(K))) ⊆ Mod(Implω(K)) This gives
Implω(K) ⊆ Implω(U(K)). The rest follows from Theorem 4.83. 

4.5 Implicationally Defined Classes of L-Algebras 227

Universal Horn classes of ordinary algebras are closed under direct limits
and reduced products. In the sequel we present analogous closure properties of
P-Horn classes of L-algebras. However, the situation is not so straightforward
as in the classical case. On the one hand, we show that P-Horn classes are
closed under direct limits of direct families and safe reduced products. On
the other hand, we show that in general P-Horn classes are not closed under
direct limits and reduced products of arbitrary families of L-algebras.
Lemma 4.87. Let L∗ be a complete residuated lattice with a truth stresser ∗ .
Let lim Mi be a direct limit of a direct family {Mi | i ∈ I} of L-algebras. Then
!P i (t ≈ t )!{Mi | i∈I} ≤ !P i (t ≈ t )!lim Mi (4.71)

for every P-Horn clause P i (t ≈ t ).
  
Proof. Let us have a valuation v : X → i∈I Mi /θ∞ and let P i (t ≈ t )

be a P-Horn clause, where Supp(P ) = {tm , tm | m = 1, . . . , n}. Take Y =
var(P ) ∪ var(t) ∪ var(t ) and consider a restriction vY of v on Y and the
homomorphic extension vY : T(Y ) → lim Mi . Since Y is finite, T(Y ) is finitely
presented. Due to Theorem 2.108 there is an index k ∈ I and a morphism
g : T(Y ) → Mk such that vY = g ◦ hk . Let gY denote a restriction of g on Y .
We have
 
!r!lim Mi ,v = !r!lim Mi ,vY = vY (r) = hk (g(r)) = hk !r!Mk ,gY
for any r ∈ T (Y ). Moreover for r, r ∈ T (Y ) it follows that
!r ≈ r !lim Mi ,v = !r!lim Mi ,v ≈lim Mi !r !lim Mi ,v =
   
= hk !r!Mk ,gY ≈lim Mi hk !r !Mk ,gY =
   
= !r!Mk ,gY θ∞ ≈lim Mi !r !Mk ,gY θ∞ =
 
= θ∞ !r!Mk ,gY , !r !Mk ,gY .
Taking into account previous observations and Remark 2.95, for every r, r ∈
T (Y ) there is an index l ∈ I, k ≤ l such that
 
!r ≈ r !lim Mi ,v = θ∞ !r!Mk ,gY , !r !Mk ,gY =
   
= hkl !r!Mk ,gY ≈Ml hkl !r !Mk ,gY =
= !r!Ml ,gY ◦hkl ≈Ml !r !Ml ,gY ◦hkl = !r ≈ r !Ml ,gY ◦hkl .
The previous idea yields that there are indices j0 , j1 , . . . , jn ≥ k such that
!tm ≈ tm !lim Mi ,v = !tm ≈ tm !Mj ◦hkjm for each m = 1, . . . , n,
m,gY
 
!t ≈ t !lim Mi ,v = !t ≈ t !Mj ,gY ◦hkj0 .
0

Moreover, I is directed, i.e. there is an index j ∈ I with j0 , j1 , . . . , jn ≤ j.


Using (2.25) it follows that
!tm ≈ tm !lim Mi ,v = !tm ≈ tm !Mj ,(gY ◦hkj for each m = 1, . . . , n,
m )◦hjm j
 
!t ≈ t !lim Mi ,v = !t ≈ t !Mj ,(gY ◦hkj )◦hj0 j .
0
228 4 Fuzzy Horn Logic

Now observe that for each m = 0, . . . , n we have


(gY ◦ hkjm ) ◦ hjm j = gY ◦ (hkjm ◦ hjm j ) = gY ◦ hkj .
Denoting gY ◦ hkj by w, we get
!P i (t ≈ t )!{Mi | i∈I} ≤ !P i (t ≈ t )!Mj ,w = !P i (t ≈ t )!lim Mi ,v .
Hence, !P i (t ≈ t )!{Mi | i∈I} ≤ !P i (t ≈ t )!lim Mi . 

Theorem 4.88. Let L∗ be a complete residuated lattice with an implicational
truth stresser ∗ . For an L-set Σ of P-Horn clauses, Mod(Σ) is an abstract
class of L-algebras closed under the formations of subalgebras, direct products,
direct unions, direct limits, and safe reduced products.
Proof. The closedness under I, S, P, and U follows from Theorem 4.85. Let
lim Mi be the direct limit of direct family {Mi ∈ Mod(Σ) | i ∈ I}. Equation
(4.71) gives
Σ(P i (t ≈ t )) ≤ !P i (t ≈ t )!{Mi | i∈I} ≤ !P i (t ≈ t )!lim Mi ,
for every P-Horn clause P i (t ≈ t ). Hence, lim Mi ∈ Mod(Σ).
Take a filer F which is safe w.r.t {Mi ∈ Mod(Σ) | i ∈ I}. Using Theo-
rem 4.82, we have MZ ∈ P(Mod(Σ)) ⊆ Mod(Σ) for every Z ∈ F (see defin-
Q MZ on page 123). Furthermore, Theorem 2.127 and Theorem 2.132
ition of
yield F Mi ∼ = lim MZ , where {MZ ∈ Mod(Σ) | Z ∈ F } is a direct family of
L-algebras. Thus,
Σ(P i (t ≈ t )) ≤ !P i (t ≈ t )!{Mi | i∈I} ≤ !P i (t ≈ t )!{MZ | Z∈F } ≤
≤ !P i (t ≈ t )!lim MZ = !P i (t ≈ t )!Q Mi
Q
F

is a consequence of (4.66), (4.68), and (4.71). Hence, F Mi ∈ Mod(Σ). 



Theorem 4.89. Let L∗ be a complete residuated lattice with an implicational
truth stresser ∗ . For a class K of L-algebras we have
Horn(K) = Horn(I(K)) = Horn(S(K)) = Horn(P(K)) =
= Horn(U(K)) = Horn(L(K)) = Horn(PR (K)) .
Proof. Follows from Theorem 4.86 and Theorem 4.88 using analogous argu-
ments as in the proof of Theorem 4.86. 

Remark 4.90. Theorem 4.88 and Theorem 4.89 are formulated for implica-
tional truth stressers, although they can be reformulated for ordinary truth
stressers satisfying (1.19) since it is the only nontrivial property of implica-
tional truth stressers being used in the proofs.
Theorem 4.88 showed that P-Horn classes are closed under direct limits of
direct families and safe reduced products. A natural question arises whether P-
Horn classes are closed also under arbitrary direct limits and reduced products.
The following examples show a negative answer: general P-Horn classes are
not closed under direct limits of weak direct families and arbitrary reduced
products.
4.5 Implicationally Defined Classes of L-Algebras 229

Fig. 4.1. L-equality from Example 4.91 and Example 4.92

Example 4.91. Take L∗ , where L is a structure of truth degrees on the unit


interval [0, 1], and ∗ is globalization. Let F = {f1 , f2 , g1 , g2 } be a type of
L-algebras such that f1 , f2 , g1 , g2 are nullary function symbols (constants).
Consider a proper family of premises P = LT (∅)×T (∅) and P-Horn clause
f1 ≈ f2 , 1 i g1 ≈ g2 . (4.72)
Moreover, let Σ be an L-set of P-Horn clauses such that
Supp(Σ) = {f1 ≈ f2 , 1 i g1 ≈ g2 } .
Thus, Mod(Σ) is a P-Horn class of L-algebras.
Now we introduce a weak direct family of L-algebras. Let I, ≤ with
I = [0, 1) be a directed index set and let Mi = Mi , ≈Mi , f1Mi , f2Mi , g1Mi , g2Mi ,
i ∈ I be L-algebras of type F such that Mi = {ai , ai , bi , bi }, f1Mi = ai , f2Mi =
ai , g1Mi = bi , g2Mi = bi , and ≈Mi is an L-equality defined in Fig. 4.1 (left). In
addition to that, we define a family {hij : Mi → Mj | i ≤ j} of morphisms by
hij (ai ) = aj , hij (ai ) = aj , hij (bi ) = bj , hij (bi ) = bj . It is immediate that
this defines a weak direct family of L-algebras which is not a direct family.
Clearly, !f1 ≈ f2 !Mi = i < 1 for every i ∈ I, thus we have
!f1 ≈ f2 , 1 i g1 ≈ g2 !Mi = 1 . (4.73)
Hence, Mi ∈ Mod(Σ) for all i ∈ I. On the other hand, for the direct limit
lim Mi it follows that θ∞ (ai , ai ) = 1, and θ∞ (bi , bi ) = 0. That is,
f1lim Mi = [ai ]θ∞ = [ai ]θ∞ = f2lim Mi ,
g1lim Mi = [bi ]θ∞ = [bi ]θ∞ = g2lim Mi ,

i.e. f1lim Mi ≈lim Mi f2lim Mi = 1 while g1lim Mi ≈lim Mi g2lim Mi = 0. Thus,


!f1 ≈ f2 , 1 i g1 ≈ g2 !lim Mi = !g1 ≈ g2 !lim Mi = 0 .
As a consequence, lim Mi ∈ Mod(Σ) since Σ(f1 ≈ f2 , 1 i g1 ≈ g2 ) > 0.
In other words, a P-Horn class Mod(Σ) is not closed under direct limits of
arbitrary weak direct families.

Example 4.92. Take L∗ such that L = [0, 1], and ∗ is globalization. Consider
the same type of L-algebras and Σ as in Example 4.91. Let N be an index
set and for every i ∈ N let Mi = Mi , ≈Mi , f1Mi , f2Mi , g1Mi , g2Mi be defined
the same way as in Example 4.91 except that for ai , ai we put ai ≈Mi ai =
230 4 Fuzzy Horn Logic

ai ≈Mi ai = 1 − 1i , see Fig. 4.1 (right). Clearly,


Q Mi ∈ Mod(Σ) for each i ∈ N.
Let F be the Fréchet filter over N. Put M = i∈N Mi . For any X ∈ F ,

[[f1M ≈ f2M ]]X = i∈X f1M (i) ≈Mi f2M (i) =
 
= i∈X f1Mi ≈Mi f2Mi = i∈X ai ≈Mi ai < 1 .
Thus, F is not safe w.r.t. {Mi | i ∈ N} since

θF (f1M , f2M ) = X∈F [[f1M ≈ f2M ]]X = 1 .
As a consequence,
!f1 ≈ f2 , 1 i g1 ≈ g2 !Q Mi = !g1 ≈ g2 !Q Mi = θF (g1M , g2M ) =
  
F F

= X∈F [[g1M ≈ g2M ]]X = X∈F i∈X bi ≈Mi bi = 0 ,


Q
showing F Mi ∈ Mod(Σ).

In certain cases, P-Horn classes are closed even under weak direct limits
and arbitrary reduced products. Obviously, this is the case if L is a Noetherian
residuated lattice since then closedness under direct limits (safe reduced prod-
ucts) implies closedness under weak direct limits (arbitrary reduced products),
see Remark 2.93 and Remark 2.130 (3). Previous examples showed that consid-
ering a complete residuated lattice on [0, 1] given by left-continuous t-norm,
there are P-Horn classes which are not closed under such constructions. In
what follows we introduce P-Horn classes which obey some form of continu-
ity. For these classes we can prove the closedness under the constructions in
question even if we work with structures of truth degrees on the real unit
interval.

Definition 4.93. Let L∗ be a complete residuated lattice on [0, 1] given by a


left-continuous t-norm ⊗. Let ∗ be the globalization on [0, 1]. Let " be a strict
continuous Archimedean t-norm.
Let P = LT (X)×T (X) , K be a class of L-algebras, P i (t ≈ t ) be a P-Horn
clause. K satisfies P i (t ≈ t ) "-continuously if for each e ∈ [0, 1) there is
d ∈ [0, 1) such that
e " !P i (t ≈ t )!K ≤ !d"P i (t ≈ t )!K . (4.74)
K is called -continuous if K satisfies all P-Horn clauses "-continuously.
If K is a P-Horn class which is "-continuous then K is called a "-continuous
P-Horn class.

Theorem 4.94. Let L∗ be a complete residuated lattice on [0, 1] given by ⊗.


Let ∗ be the globalization on [0, 1]. If K is a "-continuous P-Horn class for
some ", then K is an abstract class of L-algebras closed under subalgebras, di-
rect products, direct unions, direct limits of weak direct families, and arbitrary
reduced products.
4.5 Implicationally Defined Classes of L-Algebras 231

Proof. Let K be a "-continuous P-Horn class. The closedness under I, S,


P, and U follows from Theorem 4.88. We now show that K is closed under
arbitrary reduced products.
Take a family {Mi ∈ K | i ∈ I} of L-algebras and a filter F over I. Since
K = Mod(Σ), it suffices to check !P i (t ≈ t )!Q Mi ≥ !P i (t ≈ t )!K for
F
each P i (t ≈ t ). Assuming !t ≈ t !Q Mi ,w < !P i (t ≈ t )!K for a valua-
Q  F
 Q
tion w : X → i∈I M i /θF , we show that !P i (t ≈ t )! F Mi ,w = 1. So, let
!t ≈ t !Q Mi ,w < !P i (t ≈ t )!K . Using Lemma 1.35 (ii), there is e ∈ [0, 1)
F
such that !t ≈ t !Q Mi ,w < e " !P i (t ≈ t )!K . Since K is "-continuous, we
F
have !t ≈ t !Q Mi ,w < e " !P i (t ≈ t )!K ≤ !d"P i (t ≈ t )!K for some
F
d ∈ [0, 1).
We prove the following claim: there are s, s ∈ T (X) with P (s, s ) = 0 such
that for each Z ∈ F we have
  
i∈Z !s ≈ s !Mi ,vi < (d"P )(s, s ) , (4.75)
where
Q vi : X → Mi is a valuation defined by vi (x) = v(x)(i), and v : X →
i∈I M i is a mapping with v(x) ∈ w(x) for all x ∈ X. This claim can be
proved by contradiction: assumethat for all s, s ∈ T (X) with P (s, s ) =
0 there is Zs,s ∈ F such that i∈Zs,s !s ≈ s !Mi ,vi ≥ (d"P )(s, s ). Since

P is finite, for Z  = P (s,s )=0 Zs,s we have Z  ∈ F . As a consequence,
    
i∈Z  !s ≈ s !Mi ,vi ≥ (d"P )(s, s ) for all s, s ∈ T (X) with P (s, s ) = 0, i.e.
   
(d"P )(s, s ) ≤ !s ≈ s !Mi ,vi for all s, s ∈ T (X), and any i ∈ Z . Therefore,
one can conclude
   
i∈Z  !t ≈ t !Mi ,vi = i∈Z  !d"P i (t ≈ t )!Mi ,vi ≥
≥ !d"P i (t ≈ t )!K ≥ e " !P i (t ≈ t )!K .
This immediately gives
 
!t ≈ t !Q = !t ≈ t !Mi ,vi ≥ e " !P i (t ≈ t )!K
F Mi ,w Z∈F i∈Z

which violates !t ≈ t !Q Mi ,w < e " !P i (t ≈ t )!K . Hence, (4.75) is true.


F
Using this claim, we get
 
!s ≈ s !Q Mi ,w = Z∈F i∈Z !s ≈ s !Mi ,vi ≤ (d"P )(s, s ) < P (s, s )
F

showing that P (s, s )  !s ≈ s !Q Mi ,w , i.e. !P i (t ≈ t )!Q Mi ,w = 1. We


F F
have shown that K is closed under reduced products.
Consider a weak direct family {Mi ∈QK | i ∈ I}. Due to Theorem 2.128,
lim Mi is isomorphic to a subalgebra of F Mi . Since K is an abstract class
closed under subalgebras and reduced products, we get that lim Mi ∈ K,
proving that K is closed under direct limits of weak direct families. 


Remark 4.95. Recall the P-Horn classes Mod(Σ) introduced in Example 4.91
and Example 4.92. Neither of these two classes is "-continuous due to Theo-
rem 4.94. Examples of "-continuous classes will be shown in a further section.
232 4 Fuzzy Horn Logic

∗ ∗ ∗
In the ordinary case, validity of an implication can be expressed using the
notion of injectivity. Namely, an algebra M is injective w.r.t. an implication
P i (t ≈ t ) iff P i (t ≈ t ) is valid in M. This criterion is known as the
Banaschewski-Herrlich criterion, see [97]. Such a criterion can be used to prove
that a P-Horn class is closed under L. In what follows, we present an analogy
to Banaschewski-Herrlich criterion.

Definition 4.96. Let L∗ be a complete residuated lattice with a truth stresser.


For a weighted P-implication P i (t ≈ t ), a we put Q = P ∪ {a/t, t }. Let
hP Q : T(X)/θ(P ) → T(X)/θ(Q) be a morphism defined by
 
hP Q [t]θ(P ) = [t]θ(Q) (4.76)
for all t ∈ T (X). An L-algebra M is called injective w.r.t. P i (t ≈ t ), a
if for every morphism h : T(X)/θ(P ) → M, there exists a morphism g :
T(X)/θ(Q) → M such that h = hP Q ◦ g.

Remark 4.97. (1) For L-set Q introduced in the previous definition we have

P (t, t ) ∨ a if s = t, and s = t ,
Q(s, s ) =  (4.77)
P (s, s ) otherwise .
for every s, s ∈ T (X).
(2) Due to Lemma 2.43 mapping hP Q sending elements of T (X)/θ(P ) to
T (X)/θ(Q) defined by hP Q [t]θ(P ) = [t]θ(Q) for all t ∈ T (X) is a well-defined
morphism. The injectivity of M w.r.t. P i (t ≈ t ), a is depicted in Fig. 4.2.

Theorem 4.98. Let L∗ be a complete residuated lattice with a truth stresser



and let P i (t ≈ t ), a be a weighted P-implication.
(i) If a ≤ !P i (t ≈ t )!M , then M is injective w.r.t. P i (t ≈ t ), a .
(ii) If ∗ is globalization and M is injective w.r.t. P i (t ≈ t ), a ,
then a ≤ !P i (t ≈ t )!M .

Proof. (i): By the assumption, a ≤ !P i (t ≈ t )!M,v holds for every val-


uation v : X → M . It suffices to check that M is injective w.r.t. weighted
P-implication P i (t ≈ t ), a .

Fig. 4.2. Injectivity of M w.r.t. P i (t ≈ t ), a


4.5 Implicationally Defined Classes of L-Algebras 233

Fig. 4.3. Scheme for the proof of Theorem 4.98

Consider a morphism
  h : T(X)/θ(P ) → M and a valuation v : X → M ,
where v(x) = h [x]θ(P ) for each x ∈ X. Hence, for a homomorphic extension
v  of v we have v  = hθ(P ) ◦ h. Furthermore, !s ≈ s !M,v = v  (s) ≈M v  (s ) =
θv (s, s ) = θhθ(P ) ◦h (s, s ) for all terms s, s ∈ T (X). Therefore,
P (s, s ) ≤ θ(P )(s, s ) = [s]θ(P ) ≈T(X)/θ(P ) [s ]θ(P ) ≤
   
≤ h [s]θ(P ) ≈M h [s ]θ(P ) = θhθ(P ) ◦h (s, s ) = θv (s, s ) .
for all s, s ∈ T (X). By (1.11),
a ≤ !P i (t ≈ t )!M,v =
   
∗
= s,s ∈T (X) P (s, s ) → !s ≈ s !M,v → !t ≈ t !M,v =
   
 ∗
= s,s ∈T (X) P (s, s ) → θv  (s, s ) → θv (t, t ) =
= 1∗ → θv (t, t ) = 1 → θv (t, t ) = θv (t, t ) .
We thus have P (t, t ) ≤ θv (t, t ) and a ≤ θv (t, t ), i.e. Q(t, t ) = P (t, t ) ∨
a ≤ θv (t, t ). Since θ(Q) ∈ ConL (T(X)) is generated by Q, it readily
follows that θ(Q) ⊆ θv . Now, Lemma 2.44 yields that there is a mor-
phism g : T(X)/θ(Q) → M such that v  = hθ(Q) ◦ g, see Fig. 4.3. Hence,
hθ(P ) ◦ h = hθ(Q) ◦ g, that is hθ(P ) ◦ h = (hθ(P ) ◦ hP Q ) ◦ g. Surjectivity of hθ(P )
implies h = hP Q ◦ g. Hence, M is injective w.r.t. P i (t ≈ t ), a .
(ii): Let ∗ be globalization and let M be injective w.r.t. P i (t ≈ t ), a .
We have to show a ≤ !P i (t ≈ t )!M,v for every v. Take v : X → M . If there
are terms s, s ∈ T (X) such that P (s, s )  !s ≈ s !M,v , then !P !M,v = 0,
i.e. a ≤ !P i (t ≈ t )!M,v = 1. If P (s, s ) ≤ !s ≈ s !M,v for all s, s ∈ T (X),
we have !P !M,v = 1. Hence, we check a ≤ !t ≈ t !M,v . For the homomorphic
extension v  of v we have P ⊆ θv . That is, θ(P ) ⊆ θv . Furthermore, from
Lemma 2.44 it follows that there is a morphism g  : T(X)/θ(P ) → M such
that v  = hθ(P ) ◦ g  . Since M is injective w.r.t. P i (t ≈ t ), a , there is a
morphism g : T(X)/θ(Q) → M with g  = hP Q ◦ g. Thus,
v  = hθ(P ) ◦ g  = hθ(P ) ◦ hP Q ◦ g = hθ(Q) ◦ g .
As a consequence,
   
a ≤ θ(Q)(t, t ) = [t]θ(Q) ≈T(X)/θ(Q) [t ]θ(Q) ≤ g [t]θ(Q) ≈M g [t]θ(Q) =
= v  (t) ≈M v  (t ) = !t!M,v ≈M !t !M,v = !t ≈ t !M,v .
234 4 Fuzzy Horn Logic

By (1.11), we obtain
a ≤ !t ≈ t !M,v = 1∗ → !t ≈ t !M,v = !P i (t ≈ t )!M,v
showing a ≤ !P i (t ≈ t )!M . 


Remark 4.99. (1) For a globalization, Theorem 4.98 gives an “if and only if”
criterion for a P-implication to be true in M in degree at least a.
(2) For P = ∅, Theorem 4.98 also gives an “if and only if” criterion because
the truth stresser does not influence the degree !∅ i (t ≈ t )!M,v which equals
to !t ≈ t !M,v . Thus, for any identity t ≈ t we have a ≤ !t ≈ t !M iff M is
injective w.r.t. ∅ i (t ≈ t ), a .
(3) Theorem 4.98 can be used to prove that, considering a globalization
as the truth stresser on L, !P i (t ≈ t )!{Mi | i∈I} ≤ !P i (t ≈ t )!lim Mi is
true for every direct family {Mi | i ∈ I} of L-algebras (this is already covered
by Theorem 4.71):
Put a = !P i (t ≈ t )!{Mi | i∈I} , i.e. a ≤ !P i (t ≈ t )!Mi for all i ∈ I.
That is, every Mi is injective w.r.t. P i (t ≈ t ), a . It remains to show
that lim Mi is injective w.r.t. P i (t ≈ t ), a as well. Consider a morphism
h : T(X)/θ(P ) → lim Mi . Since P i (t ≈ t ) is a P-Horn clause, T(X)/θ(P ),
where X = var(P ) ∪ var(t) ∪ var(t ) is a finitely presented L-algebra. Due
to Theorem 2.108, for some index k ∈ I the mapping h factorizes through
some component of lim Mi , i.e. h = h ◦ hk , where h : T(X)/θ(P ) → Mk is a
morphism, see Fig. 4.4. By assumption, Mk is injective w.r.t. P i (t ≈ t ), a ,
thus there is a morphism g : T(X)/θ(Q) → Mk such that h = hP Q ◦ g. As a
consequence, h = h ◦ hk = hP Q ◦ (g ◦ hk ), i.e. g ◦ hk is the desired morphism.
Thus, lim Mi is injective with respect to P i (t ≈ t ), a .
To sum up, this section has shown that every Mod(Σ) is an abstract class of
L-algebras closed under subalgebras and direct products. In addition to that,
for certain P-implicational classes with additional constraints on their families
of premises, the model classes are closed under direct unions, (weak) direct
limits, and (safe) reduced products. In the following sections we show that the
converse assertions to those of Theorem 4.82, Theorem 4.85, Theorem 4.88,
and Theorem 4.94 are true as well. That is, we show that closedness of an
abstract class K under the corresponding operators implies that K is a P-
implicational (P-finitary implicational, P-Horn, "-continuous P-Horn) class.

Fig. 4.4. Injectivity of lim Mi with respect to P i (t ≈ t ), a


4.6 Sur-Reflections and Sur-Reflective Classes 235

K
r R

M h

h N

Fig. 4.5. Sur-reflection of M in K

4.6 Sur-Reflections and Sur-Reflective Classes


In this section, we characterize P-implicational classes as abstract classes of
L-algebras closed under subalgebras and direct products. We start by sur-
reflections which play an analogous role to that of K-free L-algebras. Later
on, we will show that every abstract class of L-algebras admits sur-reflections
if and only if it satisfies the desired closure properties. The last result in
this section establishes a connection between P-implicational classes and the
so-called sur-reflective classes.

Definition 4.100. Let K be an abstract class of L-algebras of type F , M be


an L-algebra of type F . A morphism r : M → R, where R ∈ K, is called a
reflection of M in K, if for every morphism h : M → N, where N ∈ K,
there exists uniquely determined morphism h : R → N such that h = r ◦ h .
Moreover, if r is an epimorphism (surjective morphism), then r is called a
sur-reflection of M in K, see Fig. 4.5.
An abstract class K of L-algebras of type F is called sur-reflective if
every L-algebra M of type F admits a sur-reflection r : M → R in K.

Example 4.101. Let K be a an abstract class of L-algebras. If FK (X) ∈ K then


the natural mapping hθK (X) : T(X) → FK (X) is a sur-reflection of the term
L-algebra T(X) in K. Indeed, hθK (X) is an epimorphism. Moreover, for every
morphism h : T(X) → N, where N ∈ K, we can consider a mapping g : X → N
defined by g(x) = h(x) for all x ∈ X. Due to Theorem 2.139 and Theorem 3.39,
g has a uniquely
 determined
 homomorphic extension g  : FK (X) → N. Hence,
h(t) = g [t]θK (X) = (hθK (X) ◦g  )(t) holds for all t ∈ T (X), i.e. h = hθK (X) ◦g  .


To sum up, hθK (X): T(X) → FK (X) is a sur-reflection of T(X) in K.

Theorem 4.102. Let K be an abstract class of L-algebras and let r1 : M →


R1 , r2 : M → R2 be sur-reflections of M in K. Then R1 ∼
= R2 .

Proof. Since R1 , R2 ∈ K by the definition of sur-reflections, there are uniquely


determined morphisms r1 : R2 → R1 , r2 : R1 → R2 such that r1 = r2 ◦r1 , and
r2 = r1 ◦ r2 . Thus, r1 = (r1 ◦ r2 ) ◦ r1 , and r2 = (r2 ◦ r1 ) ◦ r2 . As a consequence,
r2 ◦ r1 = idR1 , r1 ◦ r2 = idR2 . Hence, using Theorem 2.30 we readily obtain
R1 ∼ = R2 . 

236 4 Fuzzy Horn Logic

Remark 4.103. According to Theorem 4.102, a sur-reflection of M in K is


determined up to an isomorphism. This observation enables us to denote a
sur-reflection of M in K by rM : M → RK (M). Moreover, when considering
sur-reflections, we sometimes omit the surjective mapping rM and use the
term “sur-reflection” for RK (M) instead. In such a case we assume that rM
is the corresponding mapping.
A sur-reflection of M in an abstract class K can be thought of as the
greatest image of M in K. The notion of a greatest image can be defined as
follows. An L-algebra M ∈ K is said to be the greatest image of M in K if
there is an epimorphism h : M → M and for every epimorphism g : M → N,
N ∈ K we have θh ⊆ θg . Obviously, M ∼ = M/θh , and M/θh is a “greater
factor L-algebra” than M/θg ∼ = N. That is, the definition of the greatest
image corresponds well to the intuition. The following theorem characterizes
the relationship between sur-reflections and greatest images in more detail.
Theorem 4.104. Let K be an abstract class of L-algebras such that S(K) ⊆
K. Then an epimorphism r : M → R is a sur-reflection of M in K iff R is
the greatest image of M in K.
Proof. “⇒”: Let r : M → R be a sur-reflection of M in K. That is, for every
epimorphism g : M → N there is a morphism g  : R → N such that g = r ◦ g  .
Thus, θr ⊆ θr◦g = θg , i.e. R is the greatest image of M in K.
“⇐”: Let R be the greatest image of M in K. Take arbitrary N ∈ K and
a morphism h : M → N. Note that h is not supposed to be surjective. On
the other hand, Theorem 2.35 yields h = h ◦ g, where h : M → M/θh is an
epimorphism, and g : M/θh → N is an embedding. Since K is closed under
subalgebras, we have M/θh ∈ IS(K) = K. Moreover, R is supposed to be
the greatest image, i.e. θr ⊆ θh for some epimorphism r : M → R. Since
R∼ = M/θr , from Lemma 2.44 it follows that there is a uniquely determined
morphism g  : R → M/θh such that h = r ◦ g  , see Fig. 4.6. Therefore,
h = h ◦ g = (r ◦ g  ) ◦ g = r ◦ (g  ◦ g). Altogether, r : M → R is a sur-reflection
of M in K. 

Theorem 4.105. An abstract class of L-algebras is sur-reflective iff it is
closed under the formations of subalgebras and direct products.

Fig. 4.6. Sur-reflection as the greatest image in K


4.6 Sur-Reflections and Sur-Reflective Classes 237

Proof. “⇒”: Let K be a sur-reflective class of L-algebras. It remains to check


closedness under S and P. Take M ∈ K, and N ∈ Sub(M). We show N ∈ K
by checking that N is isomorphic to its sur-reflection RK (N) with rN : N →
RK (N). Consider an embedding h : N → M. Then h = rN ◦ h for some
morphism h : RK (N) → M. Since h is an embedding, we have
a ≈N b ≤ rN (a ) ≈RK (N) rN (b ) ≤ h (rN (a )) ≈M h (rN (b )) =
= h(a ) ≈M h(b ) = a ≈N b .
for every a , b ∈ N . Thus, rN is an embedding. Since rN is a sur-reflection it is
also an epimorphism. Hence, N ∼ = RK (N) ∈ K proving Q N ∈ K, i.e. S(K) ⊆ K.
Q Take a family
Q {M i | i ∈ I} ⊆ K.
Q We will show Q i∈I Mi ∈ K by proving
M ∼ R ( M ), where r : M → R (
Q
i∈I i = K i∈I i i∈I Qi K i∈I Mi ) is a sur-reflection
of i∈I Mi in K. Every projection πj : i∈I Mi → Mj is Q an epimorphism.
Hence, for every j ∈ I there exists a morphism pj : RK ( i∈I Mi ) → Mj
suchQthat πj = Q r ◦ pj . Theorem 2.51 yields that there is a morphism h :
RK ( i∈I Mi ) → i∈I Mi such that h◦πj = pj for every j ∈ I. Thus, r◦h◦πj =
r ◦ pj = πj for every j ∈ I and so r ◦ h = idQi∈I Mi . Now, we have
Q Q
a≈ i∈I Mi b ≤ r(a ) ≈RK ( i∈I Mi ) r(b ) ≤
Q Q
≤ h(r(a )) ≈ h(r(b )) = a ≈ i∈I Mi b
i∈I Mi

Q Q Q
Q a , b ∈ i∈I M i . Hence, r : i∈I Mi → RK ( i∈I Mi ) is an isomorphism,
for all
i.e. i∈I Mi ∈ K.
“⇐”: Suppose K is an abstract class of L-algebras of type F and let S(K) ⊆
K and P(K) ⊆ K. We show that every L-algebra M of type F has a sur-
reflection in K. Let
HK (M) = {θ ∈ ConL (M) | M/θ ∈ K} . (4.78)
HK (M) is non-empty since K is closed under P. Putting,
Q
PK (M) = θ∈HK (M) M/θ , (4.79)
P(K) ⊆ K implies PK (M) ∈ K.
For a family {hθ : M → M/θ | θ ∈ HK (M)} of natural morphisms we can
apply Theorem 2.51 to get a uniquely determined morphism p : M → PK (M)
with p ◦ πθ = hθ . Finally, from Theorem 2.35 it follows that p = r ◦ s, where
r : M → R is an epimorphism, and s : R → PK (M) is an embedding, i.e.
R ∈ IS(K) ⊆ K.
We claim that r : M → R is a sur-reflection of M in K. Take a morphism
h : M → N where N ∈ K. Using Theorem 2.35 we have h = hθh ◦ g, where
hθh : M → M/θh is a natural morphism, and g : M/θh → N is an embedding,
see Fig. 4.7. Thus, M/θh ∈ K, i.e. θh ∈ HK (M). We have,
h = hθh ◦ g = p ◦ πθh ◦ g = r ◦ s ◦ πθh ◦ g .
Hence, for h : R → N being s ◦ πθh ◦ g we have h = r ◦ h . Since r is surjective,


the uniqueness of h is immediate. Altogether, r : M → R is a sur-reflection


of M in K, i.e. K is sur-reflective. 

238 4 Fuzzy Horn Logic

Fig. 4.7. Construction of a sur-reflection of M in K

Remark 4.106. In fact, the construction presented in Theorem 3.40 can be


thought of as a special case to that one presented in “⇐” of Theorem 4.105.
Indeed,
Q for M being T(X), we have HK (M) = ΦK (X), and PK (M) stands for
θ∈ΦK (X) T(X)/θ. Using this we obtain FK (X) ∈ K.

Theorem 4.107. Let L∗ be a complete residuated lattice with an implicational


truth stresser ∗ . Then every P-implicational class is sur-reflective.

Proof. Consequence of Theorem 4.82 and Theorem 4.105. 




∗ ∗ ∗
Now we present a closure characterization of P-implicational classes in
terms of closedness under S and P. Unlike the ordinary case, we do not have an
“if and only if” criterion for any P and any L∗ . On the other hand, the results
presented below show the importance of P-implications with P = LT (X)×T (X)
as well as the importance of complete residuated lattices with globalization.
 
For an L-algebra M = M, ≈M , F M we can consider a set X of variables
with |X| = |M |. For the sake of convenience, we can assume X = M . Then
T(M ) is a term L-algebra of type F . The terms of type F over M are de-
noted by a, b, f (a1 , . . . , an ), and so on while the elements of M are denoted
by a , b , f M (a1 , . . . , an ).
Evidently, for the identical mapping idM : M → M and the corresponding
homomorphic extension idM : T(M ) → M we have
idM (f (a1 , . . . , an )) = f M (a1 , . . . , an ) . (4.80)
For technical reasons we introduce particular L-sets of P-implications:

Definition 4.108. Suppose K is a sur-reflective class of L-algebras of type F .


For every L-algebra M of type F let PM = LT (M )×T (M ) . Moreover, we define
an L-set PM ∈ PM of premises by
PM (s, s ) = idM (s) ≈M idM (s ) (4.81)
 K
for all terms s, s ∈ T (M ). A PM -theory of M over K is an L-set ΣM of
PM -implications defined by
4.6 Sur-Reflections and Sur-Reflective Classes 239

K
  rM (idM (t)) ≈RK (M) rM (idM (t )) if P = PM ,
ΣM P i (t ≈ t ) =
0 otherwise ,
(4.82)
where rM : M → RK (M) is a sur-reflection of M in K.

Lemma 4.109. Let L∗ be a complete residuated lattice with globalization.


 Let
K
K be a sur-reflective class of L-algebras of type F . Then K = M Mod(ΣM )
where M ranges over all L-algebras of type F .
 K K
Proof. “K ⊆ M Mod(ΣM )”: It suffices to show K ⊆ Mod(ΣM ) for every
L-algebra M of type F , i.e. to check that
K
ΣM (P i (t ≈ t )) ≤ !P i (t ≈ t )!N
holds for every PM -implication P i (t ≈ t ) and every N ∈ K.
So, let us have N ∈ K, and let v : M → N be a valuation of M in N with its
homomorphic extension v  : T(M ) → N. Theorem 2.35 yields that v  = g ◦ g  ,
where g : T(M ) → T(M )/θv is a natural morphism and g  : T(M )/θv → N
is an embedding.
Let P i (t ≈ t ) be a PM -implication such that P = PM . If P (s, s ) 
v (s) ≈N v  (s ) for some s, s ∈ T (M ), then obviously !P i (t ≈ t )!N,v = 1.


Thus, let P (s, s ) ≤ v  (s) ≈N v  (s ) for all s, s ∈ T (M ).


Consider a mapping h : M → T (M )/θv defined by h(a ) = [a]θ  for every
v
a ∈ M . Since
idM (s) ≈M idM (s ) = P (s, s ) ≤ v  (s) ≈N v  (s ) ,
for all s, s ∈ T (M ), it follows that
a ≈M b = idM (a) ≈M idM (b) ≤
≤ v  (a) ≈N v  (b) = θv (a, b) = [a]θ ≈T(M )/θv [b]θ
v v

for all a , b ∈ M . That is, h is an ≈-morphism. Consider an n-ary f M ∈ F M


and a1 , . . . , an ∈ M . Putting f M (a1 , . . . , an ) = b , we have
1 = f M (a1 , . . . , an ) ≈M b = idM (f (a1 , . . . , an )) ≈M idM (b) ≤
≤ [f (a1 , . . . , an )]θ ≈T(M )/θv [b]θ .
v v

Then, [f (a1 , . . . , an )]θ  = [b]θ  . This further gives


v v
 M 
h f (a1 , . . . , an ) = h(b ) = [b]θ  = [f (a1 , . . . , an )]θ  =
   
v v
T(M )/θv T(M )/θv
=f [a1 ]θ  , . . . , [an ]θ  = f h(a1 ), . . . , h(an ) .
v v

Altogether, h : M → T(M )/θv is a morphism. Clearly, g = idM ◦ h.


Since K is sur-reflective, we have h ◦ g  = rM ◦ h , where h : RK (M) → N,
see Fig. 4.8. Thus, v  = g ◦ g  = idM ◦ h ◦ g  = idM ◦ rM ◦ h , which implies
240 4 Fuzzy Horn Logic

Fig. 4.8. Scheme for the proof of Lemma 4.109

K
   
ΣM (P i (t ≈ t )) = rM idM (t) ≈RK (M) rM idM (t ) ≤
     
≤ h rM idM (t) ≈N h rM idM (t ) =
= v  (t) ≈N v  (t ) = !t ≈ t !N,v = !P i (t ≈ t )!N,v .
K
Therefore,ΣM (P i (t ≈ t )) ≤ !P i(t ≈ t )!N , i.e. N ∈ Mod(ΣM K
).
K K
“K ⊇ M Mod(ΣM )”: Let N ∈ M Mod(ΣM ). It suffices to show that
the sur-reflection rN : N → RK (N) is an embedding, since then N ∼ = RK (N),
K K
i.e. N ∈ K. Evidently, N ∈ M Mod(ΣM ) implies N ∈ Mod(ΣN ). Hence,
we can consider a valuation idN : N → N and its homomorphic extension
K
idN : T(N ) → N. Taking into account N ∈ Mod(ΣN ), it follows that
   RK (N)    
rN idN (t) ≈ rN idN (t ) = ΣN (PN i (t ≈ t )) ≤
K

≤ !PN i (t ≈ t )!N,idN = !t ≈ t !N,idN = idN (t) ≈N idN (t ) .


Thus, rN : N → RK (N) is an embedding, i.e. N ∈ K. 


Now we face the following problem. Given a sur-reflective class K, we


K
have shown that there is a class of L-sets ΣM of PM -implications such that
K
K = M Mod(ΣM ). For every M we use a separate proper family of premises
PM . In addition to that, we deal with a proper class of L-sets since M ranges
over a proper class of all L-algebras of type F . Thus, Lemma 4.109 itself does
not yield that K is a P-implicational class.
In the ordinary case, the above-mentioned problem of definability by sets of
implications has been solved by J. Adámek, see [2]. In what follows, we adopt
Adámek’s approach to get the desired result. The key point of [2] is that one
can show that every sur-reflective class (in [2] called a quasivariety) is definable
by a set of implications assuming the so-called Vopěnka’s Principle. Moreover,
it has been shown that considering the negation of Vopěnka’s Principle, there
is always a sur-reflective class which cannot be defined by implications. For
our purpose, we use the following principle concerning classes of L-algebras.

Principle 4.110. Given any proper class K of L-algebras of the same type,
there are distinct L-algebras M, N ∈ K such that M can be embedded into N.

Theorem 4.111. Vopěnka’s Principle implies Principle 4.110.


4.6 Sur-Reflections and Sur-Reflective Classes 241

Proof. Let K be a proper class of L-algebras of type F . We use K to construct a


proper class Kc of ordinary first-order structures corresponding to L-algebras
from K and then apply Vopěnka’s Principle for K to show that there are
distinct L-algebras M, N ∈ K and an embedding h : M → N.
Let R = {≈a | a ∈ L} be a set of binary relation symbols. For each  M∈
K, we can consider a first-order structure Mc = Mc , RMc , F Mc of type
R, F, σ , where Mc = M , F Mc = F M (i.e. the functional parts of M and Mc
coincide), and each ≈M a
c
a-cut ofM≈
is an M
, i.e. u ≈a v iff u ≈ v ≥ a.
Mc M

Thus, we have u ≈ v = M
a | u ≈a c v (u , v ∈ M ). Clearly, Kc is a
proper class of first-order structures. Hence, by Vopěnka’s Principle, there
are Mc , Nc ∈ Kc such that Mc can be (isomorphically) embedded into Nc .
That is, there is a mapping  h : Mc → Nc such that h f Mc (u1 , . . . , un ) =
f Nc
h(u1 ), . . . , h(un ) (f ∈ F , u1 , . . . , un ∈ Mc ) and u ≈M a v iff h(u ) ≈a
c Nc

h(v ) (u , v ∈ Mc , a ∈ L). As a consequence,


   
u ≈M v = a | u ≈Ma v =
c
a | h(u ) ≈N a h(v ) = h(u ) ≈ h(v ) ,
c N

showing that h is an embedding of L-algebras M, N. 




In Sect. 2.5, we introduced κ-direct unions as particular direct unions. We


are going to use this construction to delimit cardinalities of sets of variables.
The following definition introduces the closedness under κ-direct unions.

Definition 4.112. An abstract class K of L-algebras is said to be closed


κκ-direct unions, if for every κ-directed family {Mi ∈ K | i ∈ I} we
under
have i∈I Mi ∈ K.

Lemma 4.113. Suppose K is an abstract class of L-algebras which is closed


under κ-direct unions. Then K is closed under κ -direct unions for all κ > κ.

Proof. Let {Mi ∈ K | i ∈ I} be a κ -directed family. Then for every J ⊆ I,


|J| < κ we have |J| < κ , i.e. there is some i ∈ I such that Mj ∈ Sub(Mi )
for all j ∈ J. Therefore, {Mi ∈ K | i ∈ I} is a κ-directed
 family. Since K is
supposed to be closed under κ-direct unions, we have i∈I Mi ∈ K, i.e. K is
closed under κ -direct unions. 


Remark 4.114. Since ω is the least infinite cardinal, a class being closed under
ω-direct unions is closed under arbitrary κ-direct unions. In other words, if
U(K) ⊆ K, then K is closed under κ-direct unions for any κ.

Lemma 4.115. Let L∗ be a complete residuated lattice with globalization.


Suppose K is an abstract class of L-algebras, κ is an infinite cardinal,
P = LT (X)×T (X) , |X| = κ. Then K = Mod(Σ) for some L-set Σ of Pκ -
implications iff K is a sur-reflective class which is closed under κ-direct unions.

Proof. “⇒”: Let K = Mod(Σ) for some L-set Σ of Pκ -implications. From


Theorem 4.82 and Theorem 4.105 it follows that K is a sur-reflective class.
Thus, it suffices to show that K is closed under κ-direct unions.
242 4 Fuzzy Horn Logic

Let {Mi ∈ K | i ∈ I} be a κ-directed family. We have to show that


Σ(P i (t ≈ t )) ≤ !P i (t ≈ t )! κ Mi
i∈I



for every Pκ -implication P i (t ≈ t ). Take
 a valuation v : X → i∈I Mi and
κ
its homomorphic extension v  : T(X) → i∈I Mi . Put Y = var(P ) ∪ var(t) ∪

var(t ). For each x ∈ Y we can choose an index ix ∈ I such that v(x) ∈ Mix .
Let us have an index set J = {ix | v(x) ∈ Mix and x ∈ Y }. Since |Y | < κ, it
follows that |J| < κ, i.e. there is i ∈ I such that v(x) ∈ Mi for each x ∈ Y .
Consequently,
!P i (t ≈ t )! κ Mi ,v = !P i (t ≈ t )!Mi ,v .
i∈I

This further gives i∈I Mi ∈ Mod(Σ) = K.
“⇐”: Let K be a sur-reflective class which is closed under κ-direct unions.
Put Σ = Implκ(K). We claim that K = Mod(Implκ(K)). Trivially, K ⊆
Mod(Implκ(K)). Thus, it remains to check the converse inequality. Doing so, it
is sufficient to show that every κ-generated L-algebra from Mod(Implκ(K)) be-
longs to K. Indeed, due to Theorem 2.87, every M ∈ Mod(Implκ(K)) is isomor-
phic to a κ-direct union of {[M  ]M | M  ⊆ M, |M  | < κ} ⊆ Mod(Implκ(K))
and K is assumed to be closed under κ-direct unions.
So, let us have a κ-generated M ∈ Mod(Implκ(K)). Since K is sur-
reflective, M has a sur-reflection rM : M → RK (M) in K. We will show
that rM is an embedding. By contradiction, suppose there are b , b  ∈ M such
that b ≈M b   rM (b ) ≈RK (M) rM (b  ).
Let M  such that |M  | < κ, denote the set of generators of M. For a subset
of variables Y ⊆ X such that |Y | = |M  | we can consider a surjective valuation
v : Y → M  and its surjective homomorphic extension v  : T(Y ) → M, see
Theorem 2.81 and Theorem 2.141. Define an L-set P ∈ LT (X)×T (X) by
 
 v (s) ≈M v  (s ) for s, s ∈ T (Y ) ,
P (s, s ) =
0 otherwise .
Since |Y | < κ, it follows that P ∈ Pκ . The surjectivity of v  yields that there
are terms t, t ∈ T (Y ), where v  (t) = b and v  (t ) = b  . Hence,
!P i (t ≈ t )!M,v = !t ≈ t !M,v = v  (t) ≈M v  (t ) =
= b ≈M b   rM (b ) ≈RK (M) rM (b  ) .
Since M ∈ Mod(Implκ(K)), we have
(Implκ(K))(P i (t ≈ t ))  rM (b ) ≈RK (M) rM (b  ) .
Thus, there is an L-algebra N ∈ K and a valuation w : Y → N , where
P (s, s ) ≤ !s ≈ s !N,w holds for all terms s, s ∈ T (Y ), and !t ≈ t !N,w 
rM (b ) ≈RK (M) rM (b  ).
On the other hand, we clearly have θv ⊆ θw . Thus, from Lemma 2.44 it
follows that there is a morphism g : M → N such that w = v  ◦g, see Fig. 4.9.
Since N ∈ K, there is a morphism g  : RK (M) → N, where g = rM ◦ g  . As a
consequence, w = v  ◦ rM ◦ g  . Moreover,
4.6 Sur-Reflections and Sur-Reflective Classes 243

v M rM

T(Y ) g RK (M)

w N g
Fig. 4.9. Scheme for the proof of Lemma 4.115

!t ≈ t !N,w = w (t) ≈N w (t ) = g  (rM (v  (t))) ≈N g  (rM (v  (t ))) =


= g  (rM (b )) ≈N g  (rM (b  )) ≥ rM (b ) ≈RK (M) rM (b  )
which is a contradiction. Altogether, K = Mod(Σ) for Σ = Implκ(K). 

Remark 4.116. Note that the “⇐” part of the original proof of Lemma 4.115
for ordinary algebras (see [2]) differs from that one presented above. In [2],
the author presents a direct construction of an algebra in K, while we proceed
by contradiction and use properties of sur-reflections. From the viewpoint of
theory of L-algebras, the original construction pertains only to trivial ≈M ’s
and is thus not applicable for general L-algebras.
Analogously as for L-sets of P-implications and accordingly to the rela-
tionship between L-set of P-implications and sets of weighted P-implications
we denote the class of all models of a class Σ of weighted P-implications by
Mod(Σ), i.e.
Mod(Σ) = {M | a ≤ !ϕ!M for each ϕ, a ∈ Σ} . (4.83)
Lemma 4.117. Let L∗ be a complete residuated lattice with globalization, K
be a sur-reflective class of L-algebras of type F . Then there is a class Σ of
weighted implications with weighted premises such that
 K
K = M Mod(ΣM ) = Mod(Σ) ,
where M ranges over all L-algebras of type F .
Proof. Let Σ be a class, where P i (t ≈ t ), a ∈ Σ K
i (t ≈ t )) = a
 iff ΣM (P K
for some L-algebra M of type F . We will check M Mod(ΣM ) = Mod(Σ).
The rest follows from Lemma 4.109.
K K
“⊆”: Let N ∈ M Mod(ΣM ), i.e. ΣM (P i (t ≈ t )) ≤ !P i (t ≈ t )!N
for every L-algebra M of type F and every PM -implication P i (t ≈ t ). Thus,
a ≤ !P i (t ≈ t )!N for every P i (t ≈ t ), a ∈ Σ, i.e. N ∈ Mod(Σ).
“⊇”: Let N ∈ Mod(Σ). For every L-algebra M of type F and arbi-
trary PM -implication P i (t ≈ t ) we have P i (t ≈ t ), a ∈ Σ, where a =
K
ΣM (P i (t ≈ t )). Since N ∈ Mod(Σ), it follows thata ≤ !P i (t ≈ t )!N .
K K
Thus, N ∈ Mod(ΣM ) for every M of type F , i.e. N ∈ M Mod(ΣM ). 

Theorem 4.118. Let L∗ be a complete residuated lattice with globalization.
Assuming Principle 4.110, for every sur-reflective class K of L-algebras there
exists an L-set Σ of P-implications such that K = Mod(Σ).
244 4 Fuzzy Horn Logic

Proof. We will show by contradiction that if K were not definable by any


L-set of P-implications then Principle 4.110 would be violated. Thus, assume
K to be a sur-reflective class such that K = Mod(Σ) for every L-set Σ of P-
implications, where P is an arbitrary proper family of premises. Lemma 4.117
yields that K is definable by a class Σ of weighted implications with weighted
premises.
For every infinite cardinal κ we can consider a proper family of premises Pκ
such that P = LT (X)×T (X) , |X| = κ. Clearly, for every Pκ there is only a set
of weighted Pκ -implications in Σ. Moreover, from Lemma 4.115 it follows that
K cannot be closed under κ-direct unions. That is, for every infinite cardinal
κ there is an L-algebra Mκ ∈ K such that Mκ is a κ-direct union of some
κ-directed family {Mκ,i ∈ K | i ∈ Iκ }. Let us define an ordinal sequence of L-
 Mκ ’s. Put N0 = Mω . For every ordinal α let Nα be
algebras formed of such
Mκ such that κ > | β<α Nβ |.
Observe that |Mκ | < κ implies Mκ,i ∈ K for some index i ∈ Iκ , which
contradicts Mκ,i ∈ K. Hence, for every Mκ , we have |Mκ | ≥ κ. It immediately
follows that |Nα | > |Nβ | for every β < α. Thus, for β < α there cannot be
an injective mapping sending elements of Nα to Nβ . Since every embedding
is injective, Nα cannot be embedded into Nβ . On the other hand, Nβ cannot
be embedded into Nα either. Indeed, suppose h : Nβ → Nα is an embedding.
Since Nα is a κ-direct union and |Nβ | < κ by definition, for every a ∈ Nβ there
is an index ia ∈ Iκ such that h(a ) ∈ Mκ,ia . Moreover, |{ia | a ∈ Nβ }| < κ, i.e.,
there is some k ∈ Iκ such that h(Nβ ) ⊆ Mκ,k . Thus, Nβ ∈ K is a subalgebra
of Mκ,k ∈ K, which is a contradiction.
To sum up, the class of all Nα ’s is proper and there is no Nα which can be
embedded into another Nβ (for α = β), i.e. Principle 4.110 is violated. 


Remark 4.119. In ordinary case, assuming the negation of Vopěnka’s Principle,


there are sur-reflective classes of algebras definable only by proper classes of
implications, see [2]. This immediately applies in fuzzy case for L being 2.

The following theorem summarizes previous observations.

Theorem 4.120. Let L∗ be a complete residuated lattice with globalization.


Assuming Principle 4.110, for any abstract class K of L-algebras the following
are equivalent:
(i) K is a P-implicational class,
(ii) K is closed under S and P,
(iii) K = SP(K),
(iv) K = SP(K ) for some abstract class K of L-algebras,
(v) K is a sur-reflective class,
(vi) K = Mod(Σ) for some class Σ of weighted implications,
(vii) K = Mod(Impl(K)),
(viii) K = Mod(Impl(K )) for some abstract class K of L-algebras.
4.7 Semivarieties 245

Proof. “(i) ⇒ (ii)”: Consequence of Theorem 4.82.


“(ii) ⇒ (iii)”: Clearly, SP(K) = S(K) = K.
“(iii) ⇒ (iv)”: Trivial.
“(iv) ⇒ (v)”: Evidently, S(K) = SSP(K ) = SP(K ) = K. Analogously, we
have P(K) = PSP(K ) ⊆ SPP(K ) = SP(K ) = K. Now apply Theorem 4.105.
“(v) ⇒ (vi)”: Apply Lemma 4.117.
“(vi) ⇒ (vii)”: Clearly, Mod(Σ) is sur-reflective since it is an abstract class
closed under S, P (routine to check as in Theorem 4.82). From Theorem 4.118
it follows that K = Mod(Σ  ) for some L-set Σ  of Pκ -implications. Hence,
using (1.99) on Mod(Σ  ) we obtain
K = Mod(Σ  ) = Mod(Impl(Mod(Σ  ))) = Mod(Impl(K)) .
“(vii) ⇒ (viii)”: Trivial.
“(viii) ⇒ (i)”: By definition. 


From this moment on, we shall assume Principle 4.110 when necessary.

Theorem 4.121. Let L∗ be a complete residuated lattice with an implicational


truth stresser ∗ , P be any proper family of premises. Then if a class K of L-
algebras is definable by P -implications whose interpretation is given by ∗ then
K is definable by P-implications whose interpretation is given by globalization
on L, and P = LT (X)×T (X) for some infinite X.

Proof. Take L∗ with an implicational truth stresser ∗ . Let K be definable by


P -implications whose interpretation is given by ∗ . Thus, there is an L-set Σ
of P -implications such that K = ModL (Σ), i.e. M ∈ K iff


!P i (t ≈ t )!L 
M ≥ Σ(P i (t ≈ t ))

for any P -implication P i (t ≈ t ). Theorem 4.82 yields that K is an abstract


class of L-algebras closed under S and P. By Theorem 4.105, K is sur-reflective.
Finally, Theorem 4.118 gives that for some sufficiently large X there is an L-set
Σ  of Pκ -implications whose interpretation is given by globalization and P =
LT (X)×T (X) such that K = Mod(Σ  ). Hence, K is definable by P-implications
(observe that Pκ ⊆ P) whose interpretation is given by globalization. 


Remark 4.122. Theorem 4.121 showed the importance of globalization and


proper families of premises of the form P = LT (X)×T (X) . The assertion says
that the definability by P-implications with P = LT (X)×T (X) and globalization
is, in fact, the most general one.

4.7 Semivarieties
The aim of this section is to investigate relationship between semivarieties
of L-algebras and P-finitary implicational classes of L-algebras. The corre-
spondence between semivarieties and P-finitary implicational classes can be
246 4 Fuzzy Horn Logic

easily obtained from the previous results. We start with definition of a semi-
variety and characterization of semivarieties by a single closure operator. Our
approach to semivarieties is based on that one presented in [97].

Definition 4.123. A class K of L-algebras is called a semivariety if it is


an abstract class closed under formations of subalgebras, direct products, and
direct unions. For an abstract class K of L-algebras let VS (K) denote the
semivariety generated by K.

Remark 4.124. By Theorem 4.105, semivarieties are sur-reflective classes which


are closed under direct unions.

Lemma 4.125. Let K be an abstract class of L-algebras of type F and let


M be an L-algebra of type F . Then M ∈ US(K) iff every finitely generated
subalgebra of M belongs to S(K).

Proof. “⇒”: Let M ∈ US(K), i.e. M is a direct union of {Mi ∈ S(K) | i ∈ I}.
Take a finitely generated subalgebra N = [{a1 , . . . , an }]M . Clearly, there is
an index i ∈ I such that {a1 , . . . , an } ⊆ Mi . Therefore, N ∈ Sub(Mi ). Hence,
N ∈ SS(K) = S(K), i.e. every finitely generated subalgebra of M is in S(K).
The “⇐”-part of the claim follows from Corollary 2.88. 


Lemma 4.126. For class operators U, S we have


SU(K) ⊆ US(K) , (4.84)
UUS(K) ⊆ US(K) (4.85)
for every abstract class K of L-algebras. Moreover, US is a closure operator
on abstract classes of L-algebras.

Proof. (4.84): Let N ∈ SU(K), i.e. N ∈ Sub( i∈I Mi ) for some directed
family {Mi ∈ K | i ∈ I}. A family {[Mi ∩ N ]M | i ∈ I} of L-algebras is also
directed because for any i, j ∈ I there is k ≥ i, j such that Mi , Mj ∈ Sub(Mk ).
Thus, [Mi ∩ N ]M , [Mj ∩ N ]M ∈ Sub([Mk ∩ N ]M ). Clearly, N is a direct union
of {[Mi ∩ N ]M | i ∈ I}. Hence, N ∈ US(K).
(4.85): Take arbitrary M ∈ UUS(K). That is, M is a direct union of
{Mi ∈ US(K) | i ∈ I}. Due to Lemma 4.125 it suffices to show that all finitely
generated subalgebras of M belong to S(K). Let N = [{a1 , . . . , an }]M . Clearly,
there is some index i ∈ I such that N ∈ Sub(Mi ). Moreover, N is finitely
generated and Mi ∈ US(K), i.e Lemma 4.125 yields N ∈ S(K). Therefore,
M ∈ US(K).
Finally, USUS(K) ⊆ UUSS(K) ⊆ UUS(K) ⊆ US(K), thus it follows that
US is idempotent. As a consequence, US is a closure operator (extensivity and
monotony of US is obvious) on abstract classes of L-algebras. 


Lemma 4.127. For class operators P, U we have PU(K) ⊆ UP(K) for every
abstract class K of L-algebras.
4.7 Semivarieties 247
Q
Proof. Let us have M ∈ PU(K), i.e. M is a direct product i∈I Mi , where
Mi ∈ U(K) for all i ∈ I. Thus, for every i ∈ I there is a directed  index set
Ji , ≤i and a directed family {Mi,j ∈ K | j ∈ Ji } such that Mi is j∈Ji Mi,j .
We Q will construct a suitable directed family of direct products. First, put
J = i∈I Ji . Moreover, J can be equipped with a partial order. For every
j1 , j2 ∈ J we put j1 ≤ j2 iff j1 (i) ≤i j2 (i) for all i ∈ I. Clearly, J, ≤
is a partially ordered set. Furthermore, since every Ji , ≤i is directed, for
j1 , j2 ∈ J there is some j ∈ J such that j1 (i), j2 (i) ≤i j(i) for all i ∈ I, thus
j1 , j2 ≤ j. Altogether,
Q J, ≤ is a directed index set.
Let Nj denote i∈I Mi,j(i) for every j ∈ J. We claim that {Nj | j ∈ J}
is a directed family of L-algebras from P(K). Indeed, for j1 ≤ j2 we have
Mi,j1 (i) ∈ Sub(Mi,j2 (i) ) for all i ∈ I, i.e. there is an embedding h : Nj1 → Nj2
defined by h(a )(i) = a (i). That is, Nj1 ∈ Sub(Nj2 ).
It follows from the ordinary case that the skeletons ske(M), ske(N) coin-
cide (it is easy to check). Hence, it suffices to show that ≈M , ≈N coincide as
well. For a , b ∈ M there is an index j ∈ J such that a (i), b (i) ∈ Mi,j(i) for
all i ∈ I. That is,

a ≈M b = i∈I a (i) ≈Mi b (i) =

= i∈I a (i) ≈Mi,j(i) b (i) = a ≈Nj b = a ≈N b .
Hence, M ∈ UP(K). 


Lemma 4.128. A class operator USP is a closure operator on abstract classes


of L-algebras.

Proof. Monotony and extensivity of USP are evident. It suffices to check idem-
potency. Thus, let K be an abstract class of L-algebras. Using Lemma 4.126
and Lemma 4.127 we have
USPUSP(K) ⊆ USUPSP(K) ⊆ USUSPP(K) = USPP(K) = USP(K)
since K is an abstract class of L-algebras. That is, USP is idempotent. 


Theorem 4.129. For the class operators VS and USP we have


VS (K) = USP(K) (4.86)
for every abstract class K of L-algebras.

Proof. “⊆”: Let K be an abstract class. We check that USP(K) is closed


under U, S, and P. Lemma 4.126 gives U(USP(K)) ⊆ USP(K), S(USP(K)) ⊆
USSP(K) ⊆ USP(K). Furthermore, Lemma 4.127 yields
P(USP(K)) ⊆ UPSP(K) ⊆ USPP(K) = USP(K) .
“⊇”: Clearly, USP(K) ⊆ USPVS (K) = USVS (K) = UVS (K) = VS (K).
Altogether, USP(K) is the semivariety generated by K. 

248 4 Fuzzy Horn Logic

The following characterization shows that semivarieties are in one-to-


one correspondence with P-finitary implicational classes the interpretation of
which is determined by globalization. In fact, Theorem 4.130 can be thought
of as a special case of Lemma 4.115.

Theorem 4.130. Let L∗ be a complete residuated lattice with globalization, K


be an abstract class of L-algebras. Then K is a P-finitary implicational class
for some proper family of premises P iff K is a semivariety.

Proof. “⇒”: Consequence of Theorem 4.85.


“⇐”: Since K is supposed to be a semivariety, K is closed under direct
unions and thus under κ-direct unions for κ being ω. Hence, put κ = ω
and apply Lemma 4.115. Then clearly K = Mod(Σ), where Σ is an L-set
of P-finitary implications, P = LT (X)×T (X) , and X is a denumerable set of
variables. 


A summary follows.

Theorem 4.131. Suppose L∗ is a complete residuated lattice with globaliza-


tion. Then for any abstract class K of L-algebras the following are equivalent:
(i) K is a P-finitary implicational class,
(ii) K is a P-implicational class closed under U,
(iii) K is closed under U, S, and P,
(iv) K is a sur-reflective class closed under arbitrary κ-direct unions,
(v) K = USP(K),
(vi) K = USP(K ) for some abstract class K of L-algebras,
(vii) K is a semivariety,
(viii) K = Mod(Implω(K)),
(ix) K = Mod(Implω(K )) for some abstract class K of L-algebras.

Proof. “(i) ⇒ (ii)”: Every P-finitary implicational class is a P-implicational


class, the rest follows from Theorem 4.85.
“(ii) ⇒ (iii)”: Follows from Theorem 4.82.
“(iii) ⇒ (iv)”: Consequence of Theorem 4.105 and Lemma 4.113.
“(iv) ⇒ (v)”: Apply Lemma 2.84 (ii) and Theorem 4.129 for K = VS (K).
“(v) ⇒ (vi)”: Trivial.
“(vi) ⇒ (vii)”: Consequence of Theorem 4.129.
“(vii) ⇒ (viii)”: Using Theorem 4.130, we have K = Mod(Σ), where Σ is
an L-set of P-finitary implications. Hence, using (1.99), K = Mod(Implω(K)).
“(viii) ⇒ (ix)”: Trivial.
“(ix) ⇒ (i)” Trivial, by definition. 


Remark 4.132. Analogously as in Theorem 4.121, one can show that the de-
finability by P-finitary implications with P = LT (X)×T (X) (X being denumer-
able) and globalization is the most general one. This is a direct consequence
of Theorem 4.85, Theorem 4.105, and Theorem 4.130.
4.8 Quasivarieties 249

4.8 Quasivarieties
In this section we concentrate on relationships between P-Horn classes of L-
algebras and quasivarieties, i.e. particular classes of L-algebras with desired
closure properties. Recall that in the previous section we study implicational
classes of L-algebras that were defined by implications with finitely many
variables. However, the number of identities in premises of such implications
was not limited in any way. In the sequel, we will look closer at implications
with finite premises.
Before we begin, let us recall an obstacle we observed earlier. In Sect. 4.5,
we showed that P-Horn classes are not closed under arbitrary direct limits
and reduced products in general. This observation opens a question what
properties of quasivarieties we should actually require. In what follows we
present three reasonable ways to deal with quasivarieties in fuzzy setting.
Realizing that we have three independent generalizations of a quasivariety
which all coincide in the classical case, we can claim that there is perhaps no
“the right” notion of a quasivariety of L-algebras. We are going to use the
following definition which will be common to all the subsequent approaches.

Definition 4.133. A class K of L-algebras is called a quasivariety if it is


an abstract class closed under formations of subalgebras, direct products, and
direct limits of direct families of L-algebras from K. For an abstract class K
of L-algebras let VQ (K) denote the quasivariety generated by K.

Remark 4.134. From Theorem 4.105 and Lemma 4.144 it follows that quasi-
varieties of L-algebras are sur-reflective classes of L-algebras which are closed
under direct limits of direct families.

Basic Quasivarieties

We now focus on classes of L-algebras which are closed on direct limits of direct
families and safe reduced products. For technical reasons, several assertions
will be restricted only to a particular subclass of complete residuated lattices
with truth stressers.
Since every quasivariety K is a sur-reflective class, we can consider a sur-
reflection of M in K. We will use mainly sur-reflections of finitely presented L-
algebras. Thus, we will adopt the following convention. For a finitely presented
L-algebra T(X)/θ(R) let RK (X, R) denote the sur-reflection of T(X)/θ(R)
in K, the corresponding epimorphism will be usually denoted simply by r.
In the ordinary case, sur-reflections of finitely presented algebras play an
important role in theory of quasivarieties since every quasivariety can be re-
constructed by direct limits of such sur-reflections. In what follows we focus
on this phenomenon in fuzzy setting.

Definition 4.135. Let K be an abstract class of L-algebras of type F . An


L-algebra M of type F is said to satisfy the QF condition w.r.t. K, if
250 4 Fuzzy Horn Logic

Fig. 4.10. M satisfies the QF condition w.r.t. K

every morphism h : T(X)/θ(R) → M, where T(X)/θ(R) is a finitely


presented L-algebra, factorizes through RSP(K) (X, R), i.e. h = r ◦ h
for a sur-reflection r : T(X)/θ(R) → RSP(K) (X, R) of T(X)/θ(R) in
SP(K) and a morphism h sending elements of RSP(K) (X, R) to M.
The situation is depicted in Fig. 4.10.

Lemma 4.136. Let L be a Noetherian residuated lattice, K be an abstract


class of L-algebras, M be an L-algebra satisfying QF w.r.t. K. Then M ∼ = R,
where R is a direct limit of a direct family which consists of sur-reflections
RSP(K) (X, R) of certain finitely presented L-algebras.

Proof. Let M be an L-algebra satisfying QF w.r.t. K.


Theorem 2.107 yields that, under the notation used in its proof, M is
isomorphic to a direct limit lim T(Yi )/θ(Si ) of a direct family
{hij : T(Yi )/θ(Si ) → T(Yj )/θ(Sj ) | i ≤ j}
of finitely presented L-algebras. Let
{hi : T(Yi )/θ(Si ) → lim T(Yi )/θ(Si ) | i ∈ I}
denote the associated limitcone. 
We consider a family ri : T(Yi )/θ(Si ) → RSP(K) (Yi , Si ) | i ∈ I of sur-
reflections. For each composed morphism
(hij ◦ rj ) : T(Yi )/θ(Si ) → RSP(K) (Yj , Sj )
there exists a morphism gij : RSP(K) (Yi , Si ) → RSP(K) (Yj , Sj ) such that
hij ◦ rj = ri ◦ gij . (4.87)
 
We claim that I, ≤ , RSP(K) (Yi , Si ) | i ∈ I together with such gij ’s is a
weak direct family of L-algebras. Indeed, for i ≤ j ≤ k it follows that
ri ◦ gik = hik ◦ rk = (hij ◦ hjk ) ◦ rk = hij ◦ (hjk ◦ rk ) = hij ◦ (rj ◦ gjk ) =
= (hij ◦ rj ) ◦ gjk = (ri ◦ gij ) ◦ gjk = ri ◦ (gij ◦ gjk ) .
Since ri is surjective, we have gik = gij ◦ gjk , i.e. ri ◦ gii = hii ◦ ri = ri for every

i ∈ I. Thus, every gii is an identity mapping. Hence, RSP(K) (Yi , Si ) | i ∈ I
with gij ’s is a weak direct family. Since L is a Noetherian residuated lattice,
4.8 Quasivarieties 251

Fig. 4.11. Morphism from lim T(Yi )/θ(Si ) to lim RSP(K) (Yi , Si )

this family is in fact a direct family, see Remark 2.93 on page 104. Now it
suffices to check that M ∼ = lim RSP(K) (Yi , Si ). 
Let gi : RSP(K) (Yi , Si ) → lim RSP(K) (Yi , Si ) | i ∈ I be the limit cone of
lim RSP(K) (Yi , Si ). We have,
ri ◦ gi = ri ◦ (gij ◦ gj ) = (ri ◦ gij ) ◦ gj = (hij ◦ rj ) ◦ gj = hij ◦ (rj ◦ gj ) .
Recall that {hi : T(Yi )/θ(Si ) → lim T(Yi )/θ(Si ) | i ∈ I} satisfies DLP with re-
spect to {T(Yi )/θ(Si ) | i ∈ I}. Hence, from definition of DLP it follows that
there is a unique morphism g : lim T(Yi )/θ(Si ) → lim RSP(K) (Yi , Si ) such that
ri ◦ gi = hi ◦ g for every i ∈ I, see Fig. 4.11.
On the other hand, M is supposed to satisfy QF w.r.t. K. That is, every hi :
T(Yi )/θ(Si ) → lim T(Yi )/θ(Si ) factorizes through RSP(K) (Yi , Si ), i.e. there is
a morphism gi : RSP(K) (Yi , Si ) → lim T(Yi )/θ(Si ) such that hi = ri ◦ gi for
every i ∈ I. Moreover, for every gi we have
ri ◦ (gij ◦ gj ) = (hij ◦ rj ) ◦ gj = hij ◦ hj = hi = ri ◦ gi .
Hence, the surjectivity of ri implies gi = gij ◦ gj . Since the family of all
 
gi ’s satisfies DLP with respect to RSP(K) (Yi , Si ) | i ∈ I , there is morphism
g  : lim RSP(K) (Yi , Si ) → lim T(Yi )/θ(Si ) such that gi = gi ◦ g  for every i ∈ I,
see Fig. 4.12. Now, observe that
hi = ri ◦ gi = ri ◦ gi ◦ g  = hi ◦ (g ◦ g  ) ,
ri ◦ gi = hi ◦ g = ri ◦ gi ◦ g = ri ◦ (gi ◦ g  ) ◦ g = (ri ◦ gi ) ◦ (g  ◦ g)
for each i ∈ I. Therefore, g ◦g  = idlim T(Yi )/θ(Si ) and g  ◦g = idlim RSP(K) (Yi ,Si ) .
Hence, M ∼ = lim T(Yi )/θ(Si ) ∼
= lim RSP(K) (Yi , Si ) by Theorem 2.30. 


Fig. 4.12. Morphism from lim RSP(K) (Yi , Si ) to lim T(Yi )/θ(Si )
252 4 Fuzzy Horn Logic

Remark 4.137. The previous assertion is formulated for Noetherian residuated


lattices. This is due to the fact that general weak direct families lack important
properties: Theorem 2.107 yields that every L-algebra is isomorphic to a direct
limit of a direct family of finitely presented L-algebras. However, for general L,
there is no evidence that sur-reflections of those finitely presented algebras
form a direct family. Restricting ourselves on Noetherian residuated lattices
eliminates this obstacle.
Every quasivariety K is a sur-reflective class by definition, thus if L is
a Noetherian residuated lattice, and M ∈ K, then M can be thought of as
a direct limit of some sur-reflections RK (X, R) ∈ K of finitely presented L-
algebras. This is an immediate consequence of Lemma 4.136 since M ∈ K
satisfies QF w.r.t K trivially due to the sur-reflectivity of K. In other words,
every quasivariety is then uniquely defined by sur-reflections of all finitely
presented L-algebras. We sum up this observation by the following corollary.
Corollary 4.138. Let L be a Noetherian residuated lattice, K be a quasiva-
riety of L-algebras. Then for any M ∈ K we have M ∼ = R, where R is a
direct limit of a direct family of certain sur-reflections RK (X, R) of finitely
presented algebras. Hence, given a class K = {RK (X, R) | X, R are finite} it
follows that K = IL(K ). 

The following assertion presents an “if and only if” condition for M to
be in LSP(K) over an arbitrary Noetherian residuated lattice taken as the
structure of truth degrees.
Lemma 4.139. Let L be a complete residuated lattice, K be an abstract class
of L-algebras of type F , M be an L-algebra of type F . Then
(i) if M ∈ LSP(K), then M satisfies QF w.r.t. K;
(ii) if L is a Noetherian residuated lattice and M satisfies QF w.r.t. K,
then M ∈ LSP(K).
Proof. (i): Let M ∈ LSP(K). Then M can be identified with a direct limit of
a direct family {Mi ∈ SP(K) | i ∈ I} of L-algebras. Let hij : Mi → Mj (i ≤ j)
be the corresponding morphisms. Take a morphism h : T(X)/θ(R) → lim Mi ,
where T(X)/θ(R) is finitely presented. Theorem 2.108 gives h = g ◦ hk for
some k ∈ I and a morphism g : T(X)/θ(R) → Mk . Since Mk ∈ SP(K), we
have g = r ◦ g  , where r : T(X)/θ(R) → RSP(K) (X, R) is a sur-reflection of
T(X)/θ(R) in SP(K), and g  : RSP(K) (X, R) → Mk is a morphism. Thus,
h = g ◦ hk = (r ◦ g  ) ◦ hk = r ◦ (g  ◦ hk ), i.e. M satisfies QF w.r.t. K.
(ii): Clearly, Lemma 4.136 yields that M ∼ = R, where R is a direct limit
of L-algebras from SP(K). Hence, M ∈ LSP(K). 

In what follows we present a single closure operator LSP on abstract classes
of L-algebras and we will prove that LSP(K) is the quasivariety generated
by K. Note that due to the limitation of Lemma 4.139 we restrict ourselves only
to Noetherian residuated lattices. Also note that the proof of Lemma 4.140 is
based on its crisp counterpart, see [97].
4.8 Quasivarieties 253

Lemma 4.140. Let L be a Noetherian residuated lattice. Then


LLSP(K) ⊆ LSP(K) , (4.88)
SLSP(K) ⊆ LSP(K) , (4.89)
PLSP(K) ⊆ LSP(K) (4.90)
for every abstract class K of L-algebras.

Proof. (4.88): Let M ∈ LLSP(K), i.e. M can be identified with a direct limit
lim Mi of a direct family {Mi ∈ LSP(K) | i ∈ I}. We will show that M satisfies
QF w.r.t. K since then M ∈ LSP(K) by Lemma 4.139 (ii).
Theorem 2.108 yields that every morphism h : T(X)/θ(R) → lim Mi
from finitely presented L-algebra T(X)/θ(R) factorizes through some Mk ,
k ∈ I, i.e. there is a morphism g : T(X)/θ(R) → Mk such that h = g ◦ hk ,
where hk : Mk → lim Mi belongs to the limit cone associated with the di-
rect limit lim Mi . Moreover, Mk ∈ LSP(K), i.e. there exists a morphism
g  : RSP(K) (X, R) → Mk such that g = r ◦ g  , where r : T(X)/θ(R) →
RSP(K) (X, R) is a sur-reflection of T(X)/θ(R) in SP(K). Thus, we have
h = g ◦ hk = r ◦ (g  ◦ hk ). That is, M satisfies QF w.r.t. K.
(4.89): Let us have M ∈ SLSP(K), i.e. there is some N ∈ LSP(K) such
that M ∈ Sub(N). We will show that M satisfies QF w.r.t. K. Let k : M → N
denote the inclusion of M to N. That is, k(a ) = a for all a ∈ M . Clearly, k
is an embedding.
Let h : T(X)/θ(R) → M be a morphism from a finitely presented L-
algebra T(X)/θ(R). We define g : T(X)/θ(R) → N by putting g = h◦k. Since
N ∈ LSP(K), it follows that g = r ◦ g  , where r : T(X)/θ(R) → RSP(K) (X, R)
is a sur-reflection of T(X)/θ(R) in SP(K), and g  : RSP(K) (X, R) → N is a
morphism. Due to Theorem 2.35 morphisms h, g  can be expressed as h =
h1 ◦ h2 , g  = g1 ◦ g2 , where h1 : T(X)/θ(R) → M , g1 : RSP(K) (X, R) → N are
epimorphisms, and h2 : M → M, g2 : N → N are embeddings, the situation
is depicted in Fig. 4.13. Now it follows that
g = h ◦ k = (h1 ◦ h2 ) ◦ k = h1 ◦ (h2 ◦ k),
g = r ◦ g  = r ◦ (g1 ◦ g2 ) = (r ◦ g1 ) ◦ g2 .

Fig. 4.13. SLSP(K) ⊆ LSP(K)


254 4 Fuzzy Horn Logic

Observe that h1 , r ◦ g1 are epimorphisms and h2 ◦ k, g2 are embeddings.


Thus, Theorem 2.36 yields that there is a morphism d : N → M such that
h1 = (r ◦ g1 ) ◦ d, and g2 = d ◦ (h2 ◦ k). Hence, h = h1 ◦ h2 = r ◦ (g1 ◦ d ◦ h2 ), i.e.
h factorizes through the sur-reflection RSP(K) (X, R). Therefore, M satisfies
QF w.r.t. K. Q
(4.90): Let M ∈ PLSP(K), i.e. M is a direct product i∈I Mi , where
Mi ∈ LSP(K) for all i ∈ I. We will show that Q M satisfies QF w.r.t. K.
Take a morphism h : T(X)/θ(R) → i∈I Mi from a finitely presented
L-algebra T(X)/θ(R). h induces a family
{hi : T(X)/θ(R) → Mi | hi = h ◦ πi , i ∈ I}
of morphisms. Since Mi ∈ LSP(K), it follows that for every index i ∈ I there
exists a morphism gi : RSP(K) (X, R) → Mi such that hi = r ◦ gi , where
r : T(X)/θ(R) → RSP(K) (X, R) is a sur-reflection of T(X)/θ(R) in SP(K).
Furthermore, by Theorem 2.51 on the family {gi | i ∈ I} Q of morphisms we
can conclude that there is a morphism g : RSP(K) (X, R) → i∈I Mi such that
gi = g ◦πi for all i ∈ I. Now it readily follows that h◦πi = hi = r◦gi = r◦g ◦πi
holds for all i ∈ I. Hence, h = r ◦ g, i.e. M satisfies QF w.r.t. K. 


The following theorem generalizes the result of T. Fujiwara, see [41].

Theorem 4.141. Let L be a Noetherian residuated lattice. Then


VQ (K) = LSP(K) (4.91)
for every abstract class K of L-algebras.

Proof. “⊆”: Consequence of Lemma 4.140.


“⊇”: Clearly, LSP(K) ⊆ LSPVQ (K) = LSVQ (K) = LVQ (K) = VQ (K). 


We now give the first characterization of quasivarieties as P-Horn classes:

Theorem 4.142. Let L∗ be a Noetherian residuated lattice with globalization.


If K is a quasivariety then K = Mod(Horn(K)).

Proof. Let P = LT (Y )×T (Y ) , where Y is a denumerable set of variables. K ⊆


Mod(Horn(K)) is obvious. It remains to show the converse inclusion.
For the sake of brevity, put K = Mod(Horn(K)). From Theorem 4.88 it
follows that K is a quasivariety. Thus, every M ∈ K is a isomorphic to direct
limit of some sur-reflections RK (X, R) of finitely presented L-algebras, see
Corollary 4.138. Hence, it suffices to show that every sur-reflection RK (X, R)
of a finitely presented L-algebra T(X)/θ(R) belongs to K. Then clearly, since
K is an abstract class closed under L, we obtain M ∈ K.
Let T(X)/θ(R), where X and R are finite (without loss of general-
ity, we can assume X ⊆ Y ), and let r : T(X)/θ(R) → RK (X, R), r :
T(X)/θ(R) → RK (X, R) be the sur-reflections of T(X)/θ(R) in K, K . Since
K ⊆ K it follows that RK (X, R) ∈ K . As a consequence, the sur-reflection
r : T(X)/θ(R) → RK (X, R) factorizes through RK (X, R), i.e. there is a
4.8 Quasivarieties 255

Fig. 4.14. Factorization of a sur-reflection in K

morphism g  : RK (X, R) → RK (X, R) such that r = r ◦ g  . The situation


is depicted in Fig. 4.14. We finish the proof by demonstrating that g  is an
embedding, since then RK (X, R) ∈ IS(K) = K.
We proceed by contradiction. So, let there be elements b , b  such that
b ≈RK (X,R) b   g  (b ) ≈RK (X,R) g  (b  ) .
Since both X and R are finite, we introduce a finite P ∈ LT (X)×T (X) by

 θ(R)(s, s ) for R(s, s ) > 0 ,
P (s, s ) =
0 otherwise .
 
Define a valuation v of X in RK (X, R) by v(x) = r [x]θ(R) for every x ∈
X. Hence, for the homomorphic extension v  : T(X) → RK (X, R) we have
v  (s) = r [s]θ(R) for all terms s ∈ T (X). Clearly,

P (s, s ) ≤ θ(R)(s, s ) = [s]θ(R) ≈T(X)/θ(R) [s ]θ(R) ≤


   
≤ r [s]θ(R) ≈RK (X,R) r [s ]θ(R) =
= v  (s) ≈RK (X,R) v  (s ) = !s ≈ s !RK (X,R),v
holds for alls, s ∈T (X). Since r is surjective,
 there are terms t, t ∈ T (X)
such that r [t]θ(R) = b , and r [t ]θ(R) = b  . Thus,

!P i (t ≈ t )!RK (X,R),v = !t ≈ t !RK (X,R),v = v  (t) ≈RK (X,R) v  (t ) =


   
= r [t]θ(R) ≈RK (X,R) r [t ]θ(R) =
= b ≈RK (X,R) b   g  (b ) ≈RK (X,R) g  (b  ) .
Therefore, !P i (t ≈ t )!RK (X,R)  g  (b ) ≈RK (X,R) g  (b  ). In addition to
that, RK (X, R) ∈ K = Mod(Horn(K)) yields
 
Horn(K) (P i (t ≈ t ))  g  (b ) ≈RK (X,R) g  (b  ) .
Hence, there is N ∈ K, and w : X → N such that P (s, s ) ≤ !s ≈ s !N,w for
all s, s ∈ T (X) but !t ≈ t !N,w  g  (b ) ≈RK (X,R) g  (b  ). Observe that for
the homomorphic extension w : T(X) → N we have R ⊆ P ⊆ θw , and thus
θ(R) ⊆ θw .
256 4 Fuzzy Horn Logic

Fig. 4.15. Scheme for the proof of Theorem 4.142

Furthermore, we can apply Lemma 2.44 on θ(R) ⊆ θw to conclude that


there is a morphism h : T(X)/θ(R) → N with w = hθ(R) ◦ h. Since N ∈ K,
h factorizes through RK (X, R), i.e. there is a morphism h : RK (X, R) → N
such that h = r ◦ h , see Fig 4.15. Now it readily follows that w = hθ(R) ◦ h =
hθ(R) ◦ r ◦ h = hθ(R) ◦ r ◦ g  ◦ h . Thus, we have
!t ≈ t !N,w = w (t) ≈N w (t ) =
= (hθ(R) ◦ r ◦ g  ◦ h )(t) ≈N (hθ(R) ◦ r ◦ g  ◦ h )(t ) =
   
= (r ◦ g  ◦ h ) [t]θ(R) ≈N (r ◦ g  ◦ h ) [t ]θ(R) ≥
   
≥ (r ◦ g  ) [t]θ(R) ≈RK (X,R) (r ◦ g  ) [t ]θ(R) =
= g  (b ) ≈RK (X,R) g  (b  )
which contradicts !t ≈ t !N,w  g  (b ) ≈RK (X,R) g  (b  ). Hence, g  is an em-
bedding, RK (X, R) ∈ K. 


Remark 4.143. The proof of Theorem 4.142 differs from the corresponding
one presented in [97] and it is not just for technical reasons that naturally ap-
pear in fuzzy approach (e.g. weighted premises, general structures of truth
degrees). In [97], it is claimed that if M is isomorphic to a direct limit
lim T(Yi )/θ(Si ) of finitely presented algebras, and if M ∈ Mod(Horn(K)),
then every T(Yi )/θ(Si ) ∈ Mod(Horn(K)). This is not true as it is demon-
strated by the following counterexample.
Let us have a term algebra T(X), where X is finite. Clearly, T(X) is fi-
nitely presented. Take M ∈ K, where K is a variety such that T(X) ∈ K.
Moreover, algebra M is isomorphic to a direct limit lim T(Yi )/θ(Si ) of a di-
rect family {T(Yi )/θ(Si ) | i ∈ I} of finitely presented algebras. We can assume
T(X) ∈ {T(Yi )/θ(Si ) | i ∈ I} (if T(X) ∈ {T(Yi )/θ(Si ) | i ∈ I}, T(X) can be
added to {T(Yi )/θ(Si ) | i ∈ I} using morphisms gi : T(X) → T(Yi )/θ(Si ) de-
fined by gi (t) = [t]θ(Si ) for every i ∈ I with X ⊆ Yi – this can be made
without loss of generality, see Theorem 2.107). Using the argument from [97],
one can conclude {T(Yi )/θ(Si ) | i ∈ I} ⊆ Mod(Horn(K)) = K. That is, the
term algebra T(X) would be a member of K – a contradiction.
4.8 Quasivarieties 257

The following assertion shows that assuming a Noetherian residuated lat-


tice as the structure of truth degrees quasivarieties can be equivalently defined
using reduced products instead of direct limits.

Lemma 4.144. Let L be a Noetherian residuated lattice, K be a class of L-


algebras. K is a quasivariety iff K is an abstract class which is closed under
formations of subalgebras and safe reduced products.

Proof. The “⇒”-part follows from Theorem 2.127 and Theorem 2.132 because
K is closed under isomorphic images, direct products, and direct limits. For
the “⇐”-part, observe that K is closed under direct products which are partic-
ular safe reduced products. Furthermore, since K is closed under isomorphic
images, subalgebras, and reduced products, Theorem 2.128 gives that K is
closed under direct limits of direct families of L-algebras. 


A summary follows.

Theorem 4.145. Let L∗ be a Noetherian residuated lattice with globalization.


Then for any abstract class K of L-algebras the following are equivalent:
(i) K is a P-Horn class,
(ii) K is a P-finitary implicational class closed under L,
(iii) K is a P-implicational class closed under L,
(iv) K is closed under L, S, and P,
(v) K is closed under PR and S,
(vi) K is closed under PR , L, U, S, and P,
(vii) K = LSP(K),
(viii) K = LSP(K ) for some abstract class K of L-algebras,
(ix) K is a quasivariety,
(x) K = Mod(Horn(K)),
(xi) K = Mod(Horn(K )) for some abstract class K of L-algebras.

Proof. “(i) ⇒ (ii)”: Every P-Horn class is a P-finitary implicational class, the
rest follows from Theorem 4.88.
“(ii) ⇒ (iii)”: Trivial, by definition.
“(iii) ⇒ (iv)”: By Theorem 4.85.
“(iv) ⇒ (v)”: Apply Lemma 4.144.
“(v) ⇒ (vi)”: Consequence of Lemma 4.144 and Theorem 2.117.
“(vi) ⇒ (vii)”: By Theorem 4.141.
“(vii) ⇒ (viii)”: Trivial.
“(viii) ⇒ (ix)”: By Theorem 4.141.
“(ix) ⇒ (x)”: By Theorem 4.142.
“(x) ⇒ (xi)”: Trivial.
“(xi) ⇒ (i)”: By definition. 

258 4 Fuzzy Horn Logic

Remark 4.146. It is well known that in the ordinary case, the collections of all
varieties, quasivarieties, semivarieties, and sur-reflective classes are pairwise
distinct. This applies to the fuzzy case as well. Namely, suppose L∗ is a com-
plete residuated lattice with globalization. Let Σ be an L-set of P-implications
given by
Σ = {x ≈ y, a i x ≈ y, 1 | a ∈ L, a = 0} .
Evidently, Mod(Σ) consists of all L-algebras (of the given type) with crisp
L-equalities. Thus, Mod(Σ) is a quasivariety – the closedness under direct
limits of (weak) direct families follows from the crispness of L-equalities. Fur-
thermore, K is not a variety since Mod(Σ) is not closed under homomorphic
images (an image of a crisp L-equality need not to be crisp). To see that
quasivarieties and semivarieties of L-algebras are distinct, take an ordinary
semivariety K which is not a quasivariety (such K exists, see [97]). Consider
a class K of L-algebras defined by
 
K = N | M ∈ K, ske(N) = M, and ≈N is crisp .
That is, K results from K so that each M ∈ K uniquely corresponds with
N ∈ K which results from M by equipping M with the crisp L-equality. K
is a semivariety of L-algebras which is not a quasivariety (observe that K is
closed under any of I, S, P, U, L iff K is closed under the corresponding crisp
operator). In a similar way one can get a sur-reflective class of L-algebras
which is not a semivariety.
In the classical universal algebra, universal Horn classes are usually charac-
terized as abstract classes closed under subalgebras and reduced products. An
analogy of this assertion in fuzzy setting is already covered by Theorem 4.145.
The following theorem shows the nontrivial part of that assertion even without
invoking any connection to direct limits.
Theorem 4.147. Let L∗ be a residuated lattice with globalization, K be an
abstract class of L-algebras which is closed under subalgebras and safe reduced
products. If every filter F is safe, then K = Mod(Σ) for an L-set Σ of P-Horn
clauses, where P = LT (X)×T (X) .
Proof. Observe that K is closed under direct products since
Q F = {I}
Q is safe
with respect to any family {Mi | i ∈ I} of L-algebras, and F Mi ∼
= i∈I Mi .
As a consequence, K is a sur-reflective class. Thus, K = Mod(Impl(K)) due to
Theorem 4.120. We claim that K = Mod(Σ), where Σ = Horn(K). Trivially,
K ⊆ Mod(Horn(K)), i.e. we have to check the converse inclusion. We will
proceed by contradiction.
Let M ∈ Mod(Horn(K)) and M ∈ Mod(Impl(K)). That is,
!P i (t ≈ t )!M  (Impl(K))(P i (t ≈ t )) (4.92)
 
for some P i (t ≈ t ). Such a P i (t ≈ t ) induces a family
{P  i (t ≈ t ) | P  ∈ Fin(P )}
4.8 Quasivarieties 259

of P-Horn clauses, where Fin(P ) denotes the set of all finite restrictions of P .
Take any P  ∈ Fin(P ). Clearly, !P  i (t ≈ t )!M ≤ !P i (t ≈ t )!M .
Thus, from (4.92) it follows that !P  i (t ≈ t )!M  (Impl(K))(P i (t ≈ t )).
Since M ∈ Mod(Horn(K)), we have
(Horn(K))(P  i (t ≈ t ))  (Impl(K))(P i (t ≈ t )) . (4.93)

That is, for every P ∈ Fin(P ) there is an L-algebra NP  ∈ K and a valuation
vP  of X in NP  such that P  (s, s ) ≤ !s ≈ s !NP  ,vP  for all terms s, s ∈
T (X) and !t ≈ t !NP  ,vP   (Impl(K))(P i (t ≈ t )). In the following, we
construct certain safe reduced product of family {NP  | P  ∈ Fin(P )} to obtain
a contradiction.
Let us introduce a proper filter over Fin(P ). First, we can consider a
family Ps,s = {P  ∈ Fin(P ) | P  (s, s ) = P (s, s )} for every s, s ∈ T (X). Evi-
dently, ∅ = Ps,s ⊆ Fin(P ) for all s, s ∈ T (X). Put J = {Ps,s | s, s ∈ T (X)}.
Obviously, for every s1 , s1 , . . . , sn , sn ∈ T (X) there is a finite restriction
P  ∈ Fin(P   
 ) such that P (si ,si ) = P (si , si ) for all i = 1, . . . , n. As a conse-
quence, Psi ,si | i = 1, . . . , n = ∅ showing that J has the finite intersection
property. This enables us to define a proper filter F over Fin(P ) to be the
filter generated by J.

Q {vP  : X → NP  | P ∈ Fin(P )} of valuations induces a valuation
Family v:
X → P  ∈Fin(P ) NP  such that v(x)(P ) = vP  (x) for all x ∈ X, P  ∈ Fin(P ).
Q
Now Corollary 3.47 yields that there is a valuation w of X in F NP  such
that w(x) = [v(x)]θF for every x ∈ X. Thus, for s, s ∈ T (X) we have
Q Q
!s ≈ s !Q = w (s) ≈ F NP  w (s ) = [v  (s)]θF ≈ F NP  [v  (s )]θF =
F NP  ,w
  
= θF v  (s), v  (s ) = Z∈F [[v  (s) ≈ v  (s )]]Z =
 
= Z∈F P  ∈Z v  (s)(P  ) ≈NP  v  (s )(P  ) =
 
= Z∈F P  ∈Z vP  (s) ≈NP  vP  (s ) .
Recall that P  (s, s ) ≤ !s ≈ s !NP  ,vP  = vP  (s) ≈NP  vP  (s ) holds for every
P  ∈ Fin(P ) and all s, s ∈ T (X). Since Ps,s ∈ F , we get

!s ≈ s !Q NP  ,w ≥ [[v  (s) ≈ v  (s )]]Ps,s = P  ∈Ps,s vP  (s) ≈NP  vP  (s ) ≥
 
F

≥ P  ∈Ps,s P  (s, s ) = P  ∈Ps,s P (s, s ) = P (s, s ) ,

i.e. P (s, s ) ≤ !s ≈ s !Q NP  ,w for all terms s, s ∈ T (X). Moreover, since F


F
is safe by the assumption, we have

!t ≈ t !Q NP  ,w = [[v  (t) ≈ v  (t )]]Z0 = P  ∈Z0 vP  (t) ≈NP  vP  (t )
F

for some Z0 ∈ F . In addition to that,


vP  (t) ≈NP  vP  (t ) = !t ≈ t !NP  ,vP   (Impl(K))(P i (t ≈ t ))
for all P  ∈ Z0 . Putting previous facts together, we have
260 4 Fuzzy Horn Logic

!P i (t ≈ t )!Q = !t ≈ t !Q = P  ∈Z0 !t ≈ t !NP  ,vP  
F NP  ,w F NP  ,w
(4.94)
 (Impl(K))(P i (t ≈ t )) .
Since every NP  belongs
Q to K which is supposed to be closed under safe
reduced products, F NP  belongs to K. As a consequence,
!P i (t ≈ t )!Q ≥ (Impl(K))(P i (t ≈ t ))
F NP  ,w

which contradicts (4.94). 




Remark 4.148. Note that an observation on the importance of globalization


and unrestricted proper families of premises, which is an analog to those of
Theorem 4.121 and Remark 4.132, pertains to quasivarieties as well.

Continuous Quasivarieties

So far, the characterization of quasivarieties was restricted to Noetherian


residuated lattices. The reason for doing so was purely technical: L∗ being
a Noetherian lattice ensured the desired closure properties. We now turn our
attention to quasivarieties which are closed under arbitrary reduced products
(direct limits of weak direct families). We shall be motivated by Theorem 4.94
showing that "-continuous P-Horn classes satisfy such closure properties. In
the sequel we prove the converse assertion to that given by Theorem 4.94:
any quasivariety closed under arbitrary reduced products (or direct limits)
is a "-continuous P-Horn classes provided that we use complete residuated
lattices on [0, 1] given by left-continuous t-norms.
Unless otherwise mentioned, L is a complete residuated lattice on [0, 1]
given by left-continuous t-norm, ∗ is the globalization on [0, 1], and " is a
strict continuous Archimedean t-norm. Furthermore, P = LT (X)×T (X) where
X is a denumerable set of variables.

Theorem 4.149. Let L∗ be a complete residuated lattice on [0, 1] given by a


left-continuous t-norm ⊗. Let ∗ be the globalization on [0, 1]. If K is a class of
L-algebras which is closed under isomorphic images, subalgebras, and arbitrary
reduced products, then K is a P-Horn class.

Proof. Take a denumerable X and let P = LT (X)×T (X) . Put Σ = Horn(K).


Obviously, K ⊆ Mod(Σ). We check the converse inclusion. Take M ∈ Mod(Σ)
and let M (the universe of M) be the set of variables. If M is trivial then
M ∈ K because K is closed under P. So let M be a nontrivial L-algebra. Put
I = {P i (t ≈ t ), c | !P i (t ≈ t )!M,idM < c} .
The nontriviality of M yields I = ∅. Since M ∈ Mod(Horn(K)), for each
P i (t ≈ t ), c ∈ I, abbreviated by ϕ, c , there is M ϕ,c ∈ K and a valua-
tion v ϕ,c : M → M ϕ,c with
P (s, s ) ≤ !s ≈ s !Mϕ,c ,vϕ,c (4.95)
4.8 Quasivarieties 261

for all s, s ∈ T (M ), and !t ≈ t !Mϕ,c ,vϕ,c < c. For r, r ∈ T (M ), consider


Zr,r ⊆ I defined by
 
Zr,r = P i (t ≈ t ), c ∈ I | P (r, r ) = !r ≈ r !M,idM . (4.96)
Clearly, Zr,r = ∅ and {Zr,r | r, r ∈ T (M )} has the finite intersection prop-

Q Let F be the proper filter over I generated by {Zr,r | r, r ∈ T (M )}. Let
erty.
M ϕ,c  be the reduced product of {M ϕ,c | ϕ, c ∈ I} modulo F and let
F Q
w: M → ϕ,c ∈I M ϕ,c /θF be the valuation induced by v ϕ,c ’s. That is,
Q
w(x) = [v(x)]θF , where v : M → ϕ,c ∈I M ϕ,c such that v(x)(ϕ, c ) =
v ϕ,c (x) for all x ∈ M , and ϕ, c ∈ I. We show that w is an embedding. One
can easily check that w is compatible with functions: for any n-ary f M , and
arbitrary a1 , . . . , an ∈ M , f (a1 , . . . , an ) ∈ T (M ), i.e.
   
w f M (a1 , . . . , an ) = w f (a1 , . . . , an ) =
    Q  
= v  f (a1 , . . . , an ) θ = f ϕ,c∈I Mϕ,c v  (a1 ), . . . , v  (an ) θ =
    
F F
Q Q
Mϕ,c  Mϕ,c
=f F [v (a1 )]θF , . . . , [v (an )]θF = f F w(a1 ), . . . , w(an ) =
Q  
= f F Mϕ,c w(a1 ), . . . , w(an ) .
We now prove that w is an ≈-morphism: for a , b ∈ M , consider Za,b ∈ F , i.e.
 
Za,b = P i (t ≈ t ), c ∈ I | P (a, b) = a ≈M b
is an instance of (4.96). Using (4.95) we get a ≈M b ≤ !a ≈ b!Mϕ,c ,vϕ,c for
each ϕ, c ∈ Za,b . Thus,

a ≈M b ≤ ϕ,c ∈Za,b !a ≈ b!Mϕ,c ,vϕ,c ,
yielding
 
a ≈M b ≤ Z∈F ϕ,c ∈Z !a ≈ b!Mϕ,c ,vϕ,c =
Q
= !a ≈ b!Q = w(a ) ≈ F Mϕ,c w(b ) .
F Mϕ,c ,w

Hence, w is a morphism. Now for a = b take c > a ≈M b . Then we get


!a ≈ b!M,idM < c. So, for Zr,r ∈ F we have that if P i (t ≈ t ), d ∈ Zr,r ,
then P i (a ≈ b), c ∈ Zr,r . Since F is generated by {Zr,r | r, r ∈ T (M )},
for each Z ∈ F there is P such that P i (a ≈ b), c ∈ Z. Therefore,

ϕ,d ∈Z !a ≈ b!Mϕ,d ,vϕ,d < c .

Thus, ϕ,d ∈Z !a ≈ b!Mϕ,d ,vϕ,d ≤ a ≈M b for all Z ∈ F because c was
chosen arbitrarily. Now we have
Q
w(a ) ≈ F Mϕ,c w(b ) = !a ≈ b!Q Mϕ,c ,w =
 
F

= Z∈F ϕ,c ∈Z !a ≈ b!Mϕ,c ,vϕ,c ≤ a ≈M b .


Altogether, w is an
Q embedding. Thus, M is isomorphic to a subalgebra of the
reduced product F M ϕ,c , i.e. M ∈ ISPR (K) ⊆ K. 

262 4 Fuzzy Horn Logic

Theorem 4.150. Let L∗ be a complete residuated lattice on [0, 1] given by a


left-continuous t-norm ⊗. Let ∗ be the globalization on [0, 1]. If K is a class
of L-algebras which is closed under reduced products then K is "-continuous
for any strict continuous Archimedean t-norm ".

Proof. By contradiction, let there be P i (t ≈ t ) and e ∈ [0, 1) such that for


all d ∈ [0, 1) we have e " !P i (t ≈ t )!K > !d"P i (t ≈ t )!K . Therefore,
for each d ∈ [0, 1) there is Md ∈ K and vd : X → Md with
e " !P i (t ≈ t )!K > !d"P i (t ≈ t )!Md ,vd .
Thus, (d"P )(s, s ) ≤ !s ≈ s !Md ,vd for all s, s ∈ T (X), and
e " !P i (t ≈ t )!K > !t ≈ t !Md ,vd .
 
Put I = Q1 − n1 | n ∈ N and let F be the Fréchet filter over I. For the reduced
product FQMd of {M  d | d ∈ I ⊆ [0, 1)} modulo F one can consider a valuation
Q
w: X → d∈I M d /θF such that w(x) = [v(x)]θF , where v : X → d∈I Md
is the induced mapping satisfying v(x)(d) = vd (x) for all x ∈ X, and d ∈ I.
Since " is a continuous t-norm,
     
!s ≈ s !Q Md ,w = Z∈F d∈Z !s ≈ s !Md ,vd ≥ Z∈F d∈Z d"P (s, s ) =
F
     
= Z∈F P (s, s ) " d∈Z d = P (s, s ) " Z∈F d∈Z d =
= P (s, s ) " 1 = P (s, s )
for all s, s ∈ T (X). In addition to that,
 
!t ≈ t !Q Md ,w = Z∈F d∈Z !t ≈ t !Md ,vd ≤
F

≤ e " !P i (t ≈ t )!K < !P i (t ≈ t )!K .


Hence, !P i (t ≈ t )!Q Md < !P i (t ≈ t )!K which contradicts the fact
Q F
that F Md ∈ K. 


The following theorem characterizes "-continuous P-Horn classes of L-


algebras as particular quasivarieties of L-algebras which are closed under ar-
bitrary reduced products (direct limits).

Theorem 4.151. Let L∗ be a complete residuated lattice on [0, 1] given by a


left-continuous t-norm ⊗. Let ∗ be the globalization on [0, 1]. Then for any
abstract class K the following are equivalent:
(i) K is a "-continuous P-Horn class,
(ii) K is closed under S, and arbitrary reduced products,
(iii) K is closed under S, P, and direct limits of weak direct families.

Proof. “(i) ⇒ (ii)” and “(i) ⇒ (iii)”: Apply Theorem 4.94.


“(ii) ⇒ (i)”: By Theorem 4.149 and Theorem 4.150.
“(iii) ⇒ (ii)”: Consequence of Theorem 2.127. 

4.8 Quasivarieties 263

Remark 4.152. As a consequence of Theorem 4.151 we get that a P-Horn class


is "1 -continuous iff it is "2 -continuous. In other words, the notion of "-
continuity of P-Horn classes does not depend on the chosen ".
∗ ∗ ∗
Recall that in Sect. 4.3 we introduced a notion of a "-continuity of theories
(i.e., L-sets of P-Horn clauses). We now look in more detail at the relationship
between "-continuous P-Horn theories (see Definition 4.70 on page 216) and
"-continuous P-Horn classes (Definition 4.93).
Lemma 4.153. Let L∗ be a complete residuated lattice on [0, 1] given by a
left-continuous t-norm ⊗. Let ∗ be the globalization on [0, 1]. Let Σ be an
L-set of P-Horn clauses.
(i) If Σ is "-continuous, then Mod(Σ) is "-continuous.
(ii) If Mod(Σ) is "-continuous, then Horn(Mod(Σ)) is "-continuous.
Proof. (i): Let Σ be "-continuous and take a P-Horn clause P i (t ≈ t ).
For each e ∈ [0, 1) there is d ∈ [0, 1) such that if Σ R P i (t ≈ t ), b for
b ∈ [0, 1], then Σ R d"P i (t ≈ t ), be for be ≥ b " e. This immediately
gives e " |P i (t ≈ t )|Σ ≤ |d"P i (t ≈ t )|Σ , thus
e " !P i (t ≈ t )!Σ ≤ !d"P i (t ≈ t )!Σ
by Theorem 4.71, i.e. e " !P i (t ≈ t )!Mod(Σ) ≤ !d"P i (t ≈ t )!Mod(Σ) .
That is, Mod(Σ) is "-continuous.
(ii): Let Mod(Σ) be "-continuous. Take P i (t ≈ t ) and e ∈ [0, 1). Using
the "-continuity of Mod(Σ), we can take d ∈ [0, 1) such that
e " !P i (t ≈ t )!Mod(Σ) ≤ !d"P i (t ≈ t )!Mod(Σ) .
Let Horn(Mod(Σ)) R P i (t ≈ t ), b . We have
b ≤ |P i (t ≈ t )|Horn(Mod(Σ)) ≤ !P i (t ≈ t )!Horn(Mod(Σ)) =
= !P i (t ≈ t )!Mod(Σ)
by the soundness of FHL and using Mod(Σ) = Mod(Horn(Mod(Σ))). More-
over, the "-continuity of Mod(Σ) yields
e " b ≤ e " !P i (t ≈ t )!Mod(Σ) ≤ !d"P i (t ≈ t )!Mod(Σ) =
= Horn(Mod(Σ))(d"P i (t ≈ t )) .
Put be = Horn(Mod(Σ))(d"P i (t ≈ t )). Thus, d"P i (t ≈ t ), be is prov-
able from Horn(Mod(Σ)) using R, i.e. Horn(Mod(Σ)) is "-continuous. 

The following assertion summarizes the relationship of "-continuous P-
Horn classes and "-continuous P-Horn theories.
Theorem 4.154. Let L∗ be a complete residuated lattice on [0, 1] given by a
left-continuous t-norm ⊗. Let ∗ be the globalization on [0, 1]. Let Σ be an L-set
of P-Horn clauses. The following are equivalent.
264 4 Fuzzy Horn Logic

(i) Σ  is "-continuous,
(ii) Σ  is "-continuous,
(iii) Mod(Σ) is "-continuous,
(iv) Mod(Σ) is closed under arbitrary reduced products.

Proof. “(i) ⇔ (ii)”: See Theorem 4.57.


“(ii) ⇒ (iii)”: Apply Theorem 4.43 and Lemma 4.153 (i).
“(iii) ⇒ (ii)”: Follows from Theorem 4.43 and Lemma 4.153 (ii).
“(iii) ⇔ (iv)”: Consequence of Theorem 4.82 and Theorem 4.151. 


Remark 4.155. As an immediate consequence of Theorem 4.154 we get that


the "-continuity of the semantic closure of Σ does not depend on the chosen ".

Crisply Generated Quasivarieties

We now present a characterization of quasivarieties closed under an addi-


tional operator of fuzzification which has already been introduced in Sect. 3.5.
Throughout the rest of this section, P denotes a proper family of crisp
premises. That is, P is defined by
P = {P ∈ LT (X)×T (X) | P is crisp} (4.97)
where X is a denumerable set of variables. In what follows we will be inter-
ested in crisp L-sets of P-Horn clauses, i.e. crisp L-sets of Horn clauses with
crisp premises, and show that model classes of such L-sets are exactly the
quasivarieties closed under fuzzification. This result will be established for
any complete residuated lattice taken as the structure of truth degrees.
The following auxiliary lemma is a generalization of Lemma 3.60.

Lemma 4.156. Let L∗ be a complete residuated lattice with globalization, P be


defined by (4.97), Σ be a crisp L-set of P-Horn clauses. Then for any L-
algebras M, N with ske(M) = ske(N) we have M ∈ Mod(Σ) iff N ∈ Mod(Σ).

Proof. Take M and N such that ske(M) = ske(N). Using the same arguments
as in the proof of Lemma 3.60, we have !r ≈ r !M,v = 1 iff !r ≈ r !N,v = 1 for
every r, r ∈ T (X). Since ∗ is globalization, for any P-Horn clause P i (t ≈ t ),
!P i (t ≈ t )!M,v = 1 iff
IF P (s, s ) ≤ !s ≈ s !M,v for all s, s ∈ T (X), THEN !t ≈ t !M,v = 1 iff
IF P (s, s ) = 1 implies !s ≈ s !M,v = 1, THEN !t ≈ t !M,v = 1 iff
IF P (s, s ) = 1 implies !s ≈ s !N,v = 1, THEN !t ≈ t !N,v = 1 iff
IF P (s, s ) ≤ !s ≈ s !N,v for all s, s ∈ T (X), THEN !t ≈ t !N,v = 1 iff
!P i (t ≈ t )!N,v = 1.
Therefore, M ∈ Mod(Σ) iff for each P-Horn clause P i (t ≈ t ) such that
Σ(P i (t ≈ t )) = 1 we have !P i (t ≈ t )!M,v = 1 for any v iff for each
P i (t ≈ t ) such that Σ(P i (t ≈ t )) = 1 we have !P i (t ≈ t )!N,v = 1
for any v iff N ∈ Mod(Σ). 

4.8 Quasivarieties 265

The following assertion characterizes quasivarieties which are closed under


fuzzification as the classes of L-algebras defined by crisp L-sets of P-Horn
clauses with crisp premises.

Theorem 4.157. Let L∗ be a complete residuated lattice with globalization,


P be defined by (4.97), K be a class of L-algebras. Then K is a quasivariety
which is closed under F iff there is a crisp L-set Σ of P-Horn clauses such
that K = Mod(Σ).

Proof. “⇒”: Suppose K is closed under I, S, P, L, and F. Take Kc ⊆ K where


Kc = {M ∈ K | ≈M is crisp}. Observe that Kc is a quasivariety. Indeed, since
Kc contains only L-algebras with crisp L-equalities and K is a quasivariety,
we have that Kc is closed under I, S, P, and L, i.e. Kc is a quasivariety.
Furthermore, any weak direct family of L-algebras taken from Kc is a direct
family, i.e. Kc is closed under arbitrary direct limits.
Furthermore, consider a class K2 of 2-algebras where M ∈ K2 iff there is an
L-algebra N ∈ Kc such that ske(N) = ske(M). Obviously, K2 is but a class of
2-algebras resulting from Kc so that the crisp L-equalities (on M ) have been
replaced by 2-equalities (on M ). Hence, K2 is a quasivariety of 2-algebras.
Since 2 endowed with ∗ (i.e., 0∗ = 0, 1∗ = 1) is a Noetherian residuated
lattice with globalization, from Theorem 4.145 it follows that there is a 2-set
Σ2 of P2 -Horn clauses such that P2 = 2T (X)×T (X) and K2 = Mod(Σ2 ).
Since P2 -Horn clauses can be identified with P-Horn clauses where P is
defined by (4.97), we can introduce a crisp L-set Σ of P-Horn clauses by

0 if Σ2 (P i (t ≈ t )) = 0 ,
Σ(P i (t ≈ t )) =
1 if Σ2 (P i (t ≈ t )) = 1 ,
for every P-Horn clause P i (t ≈ t ). Now we check that K = Mod(Σ).
Take any L-algebra M. Consider an L-algebra Mc with ske(M) = ske(Mc ),
and crisp ≈Mc . Furthermore, let M2 denote the 2-algebra with ske(M) =
ske(M2 ). Since K is closed under F, we have M ∈ K iff Mc ∈ Kc . Obvi-
ously, Mc ∈ Kc iff M2 ∈ K2 = Mod(Σ2 ). Since ske(M2 ) = ske(Mc ) and
Σ is crisp, we have that M2 ∈ Mod(Σ2 ) iff Mc ∈ Mod(Σ). Finally, since
ske(Mc ) = ske(M), we get Mc ∈ Mod(Σ) iff M ∈ Mod(Σ) by Lemma 4.156.
So, for every M it follows that M ∈ K iff M ∈ Mod(Σ), i.e. K = Mod(Σ).
“⇐”: Let Σ be a crisp L-set of P-Horn clauses such that K = Mod(Σ).
Theorem 4.88 yields that K is closed under I, S, P, and L. Thus, K is a
quasivariety by Definition 4.133. It remains to check that F(K) ⊆ K. Take
M ∈ F(K). Thus, there is N ∈ K such that ske(M) = ske(N). Lemma 4.156
gives M ∈ K. Altogether, K is a quasivariety closed under F. 


Remark 4.158. Note that Theorem 4.157 gives a characterization of particular


quasivarieties (crisply generated quasivarieties) for any complete residuated
lattice with globalization taken as the structure of truth degrees. This is an
important distinction from characterizations given by Theorem 4.145 and The-
orem 4.151. On the other hand, quasivarieties closed under fuzzification are
266 4 Fuzzy Horn Logic

relatively simple extensions of the classical quasivarieties because they can


be reconstructed from classical quasivarieties (quasivarieties of 2-algebras) by
means of fuzzification.

4.9 Bibliographical Remarks

There are numerous results on properties of implicationally defined classes


of algebras, see e.g. [5, 26, 41, 73], proofs from implicational theories have
been studied as well, see e.g. [80]. A survey on implications in the context of
universal algebra can be found in monograph [97].
This chapter is based mainly on [17, 18, 91]. Further results can be found
in [90, 93], see also [15].
The present chapter shows an approach to Horn logic in Pavelka style.
Note that this is not the only approach possible. For instance, one might
develop Horn logic in a setting of Hájek’s Basic Logic [49]. This seems to be
an open problem.
References

1. Abramsky S., Gabbay D. M., Maibaum T. S. E.: Handbook of Logic in Computer


Science. Volume 1, Oxford University Press, 1992.
2. Adámek J.: How many variables does a quasivariety need? Algebra Universalis
27(1990), 44–48.
3. Audi R. (Ed.): The Cambridge Dictionary of Philosophy, 2nd Ed. Cambridge
University Press, Cambridge, 1999.
4. Baaz M.: Infinite-valued Gödel logics with 0-1 projections and relativizations.
GÖDEL ’96 – Logical Foundations of Mathematics, Computer Science and
Physics, Lecture Notes in Logic vol. 6, Springer-Verlag 1996, 23–33.
5. Banaschewski B., Herrlich H.: Subcategories defined by implications. Houston
J. Math. 2(1976), 149–171.
6. Bělohlávek R.: Similarity relations in concept lattices. J. Logic and Computa-
tion Vol. 10 No. 6(2000), 823–845.
7. Bělohlávek R.: Fuzzy closure operators. J. Math. Anal. Appl. 262(2001), 473–
489.
8. Bělohlávek R.: Fuzzy closure operators II. Soft Computing 7(2002) 1, 53–64.
9. Bělohlávek R.: Fuzzy equational logic. Arch. Math. Log. 41(2002), 83–90.
10. Bělohlávek R.: Fuzzy Relational Systems: Foundations and Principles. Kluwer
Academic/Plenum Publishers, New York, 2002.
11. Bělohlávek R.: Birkhoff variety theorem and fuzzy logic. Arch. Math. Log.
42(2003), 781–790.
12. Bělohlávek R.: Fuzzy closure operators induced by similarity. Fundamenta In-
formaticae 58(2)(2003), 79–91.
13. Bělohlávek R., Dvořák J., Outrata J.: Fast factorization by similarity in formal
concept analysis of data with fuzzy attributes (submitted).
14. Bělohlávek R., Funioková T., Vychodil V.: Fuzzy closure operators with truth
stressers Logic J. of IGPL (to appear).
15. Bělohlávek R., Hájek P., Vychodil V.: Models of theories in classical and fuzzy
logic (in preparation).
16. Bělohlávek R., Vychodil V.: Algebras with fuzzy equalities (submitted).
17. Bělohlávek R., Vychodil V.: Fuzzy Horn logic I: proof theory. Arch. Math. Log.
(to appear).
18. Bělohlávek R., Vychodil V.: Fuzzy Horn logic II: implicationally defined classes.
Arch. Math. Log. (to appear).
268 References

19. Bělohlávek R., Vychodil V.: On ultraproduct in fuzzy logic (in preparation).
20. Bhakat S. K., Das P.: Fuzzy partially ordered fuzzy subgroups. Fuzzy Sets and
Systems 67(1994), 191–198.
21. Biacino L., Gerla G., Ying M.: Approximate reasoning based on similarity.
Math. Log. Quart. 46(2000), 77–86.
22. Birkhoff G.: On the combination of algebras. Proc. Camb. Philos. Soc.
29(1933), 441–464.
23. Birkhoff G.: On the structure of abstract algebras. Proc. Camb. Philos. Soc.
31(1935), 433–454.
24. Birkhoff G.: Subdirect unions in universal algebra. Bull. Amer. Math. Soc.
50(1944), 764–768.
25. Birkhoff G.: Lattice Theory (3-rd Ed.). Am. Math. Soc. Colloq. Publ., Vol.
XXV, Providence, Rhode Island, 1967.
26. Bloom S. L., Wright J. B.: Finitary quasi varieties. J. Pure and Appl. Algebra
25(1982), 121–154.
27. Burris S., Sankappanavar H. P.: A Course in Universal Algebra. Springer-
Verlag, New York, 1981.
28. Călugăreanu G.: Lattice Concepts of Module Theory. Kluwer, Dordrecht, 2000.
29. Chakraborty A. B., Khare S. S.: Fuzzy homomorphism and algebraic structures.
Fuzzy Sets and Systems 59(1993), 211–221.
30. Chakraborty M. K., Das M.: On fuzzy equivalence I, II. Fuzzy Sets and Systems
11(1983), 185–193, 299–307.
31. Chang C. C., Keisler H. J.: Continuous Model Theory. Princeton University
Press, Princeton, NJ, 1966.
32. Chang C. C., Keisler H. J.: Model Theory. North–Holland, Amsterdam, 1973.
33. Demirci M.: Foundations of fuzzy functions and vague algebra based on many-
valued equivalence relations, part I: fuzzy functions and their applications, part
II: vague algebraic notions, part III: constructions of vague algebraic notions
and vague arithmetic operations. Int. J. General Systems 32(3)(2003), 123–
155, 157–175, 177–201.
34. Di Nola A., Gerla G.: Lattice valued algebras. Stochastica 11(1987), 137–150.
35. Dubois D., Prade H.: Fuzzy Sets and Systems: Theory and Applications. Aca-
demic Press, New York, 1980.
36. Dubois D., Prade H., Esteva F., Garcia P., Godo L.: A logical approach to
interpolation based on similarity relations. Int. J. Approx. Reasoning 17(1997),
1–36.
37. Erceg M. A.: Functions, equivalence relations, quotient spaces and subsets in
fuzzy set theory. Fuzzy Sets and Systems 3(1980), 75–92.
38. Esteva F., Godo L., Hájek P., Montagna F.: Hoops and fuzzy logic. J. Logic
and Computation. 13(2003), 532–555.
39. Feynman R. P., Leighton R. B., Sands M.: The Feynman Lectures on Physics.
Addison Wesley, Reading, MA, 1963.
40. Fodor J.: On contrapositive symmetry of implications in fuzzy logic. In:
H. J. Zimmermann (Ed.): Proc. EUFIT’93, Aachen, 1993. Verlag Augustinus
Buchhandlung, Aachen, pp. 1342–1348.
41. Fujiwara T.: On the construction of the least universal Horn class containing
a given class. Osaka J. Math. 8(1971), 425–436.
42. Gerla G.: Fuzzy Logic. Mathematical Tools for Approximate Reasoning. Kluwer,
Dordrecht, 2001.
References 269

43. Gerla G., Tortora R.: Fuzzy natural deduction. Z. Math. Logik Grundlagen
Math. 36(1990), 67–77.
44. Goguen J. A.: L-fuzzy sets. J. Math. Anal. Appl. 18(1967), 145–174.
45. Goguen J. A.: The logic of inexact concepts. Synthese 18(1968–9), 325–373.
46. Gottwald S.: Fuzzy Sets and Fuzzy Logic. Foundations of Application – from a
Mathematical Ponit of View. Vieweg, Wiesbaden, 1993.
47. Gottwald S.: A Treatise on Many-Valued Logics. Research Studies Press,
Baldock, Hertfordshire, England, 2001.
48. Hájek P.: Basic fuzzy logic and BL-algebras. Soft Computing 2(1998), 124–128.
49. Hájek P.: Metamathematics of Fuzzy Logic. Kluwer, Dordrecht, 1998.
50. Hájek P.: Function symbols in fuzzy predicate logic. Proc. East-West Fuzzy
Logic Days, Zittau 2000.
51. Hájek P.: On very true. Fuzzy Sets and Systems 124(2001), 329–333.
52. Hájek P.: A note on Birkhoff variety theorem in fuzzy logic. Preprint,
ICS AV ČR, 2002.
53. Head T.: A metatheorem for deriving fuzzy theorems from crisp versions. Fuzzy
Sets and Systems 73(1995), 349–358.
54. Höhle U.: Quotients with respect to similarity relations. Fuzzy Sets and Systems
27(1988), 31–44.
55. Höhle U.: Commutative, residuated l-monoids. In: U. Höhle, E.P. Klement
(Eds.): Non-classical Logics and Their Applications to Fuzzy Subsets. Kluwer,
Dordrecht, 1995.
56. Höhle U.: On the fundamentals of fuzzy set theory. J. Math. Anal. Appl.
201(1996), 786–826.
57. Höhle U.: Many-valued equalities, singletons and fuzzy partitions. Soft Com-
puting 2(1998), 134–140.
58. Jacas J.: Similarity relations – the calculation of minimal generating families.
Fuzzy Sets and Systems 35(1990), 151–162.
59. Klawonn F., Castro J. L.: Similarity in fuzzy reasoning. Mathware & Soft Com-
puting 2(1995), 197–228.
60. Klement E. P., Mesiar R., Pap E.: Triangular Norms. Kluwer, Dordrecht, 2000.
61. Klir G. J., Yuan B.: Fuzzy Sets and Fuzzy Logic. Theory and Applications.
Prentice–Hall, Upper Saddle River, NJ, 1995.
62. Leeuwen J. van: Handbook of Theoretical Computer Science, Vol. B. Elsevier,
Amsterdam, 1990.
63. Ling C. H.: Representation of associative functions. Publ. Math. Debrecen
12(1965), 189–212.
64. Mal’cev A. I.: On the general theory of algebraic systems (in Russian) Math.
Sbornik (N.S.) 35(77)1954, 3-20. English translation: Amer. Math Soc. Trans-
lations (2)27(1963), 125–142.
65. Miller G. A.: The magical number plus seven, plus or minus two: Some limits
on our capacity to for processing information. Psychol. Rev. 63(1956), 81–97.
66. Mostafa S. M.: Fuzzy implicative ideals in BCK-algebras. Fuzzy Sets and Sys-
tems 87(1997), 361–368.
67. Mostert P. S., Shields A. L.: On the structure of semigroups on a compact
manifold with boundary. Ann. Math. 65(1957), 117–143.
68. Murali V.: Fuzzy congruence relations. Fuzzy Sets and Systems 41(1991), 359–
369.
69. Neuman W. D.: On Malcev conditions. Austral. Math. Soc. 17(1974), 376–384.
270 References

70. Novák V.: Fuzzy Sets and Their Applications. Adam–Hilger, Bristol, 1989.
71. Novák V., Perfilieva I., Močkoř J.: Mathematical Principles of Fuzzy Logic.
Kluwer, Boston, 1999.
72. Ovchinikov S. V.: Structure of fuzzy binary relations. Fuzzy Sets and Systems
6(2)(1981), 169–195.
73. Palasińska K., Pigozzi G.: Gentzen-style axiomatization in equational logic.
Algebra Universalis 34(1995), 128–143.
74. Pavelka J.: On fuzzy logic I, II, III. Z. Math. Logik Grundlagen Math. 25(1979),
45–52, 119–134, 447–464.
75. Preparata F. P., Yeh R. T.: Introduction to Discrete Structures. Addison-
Wesley, Reading, Massachusetts, 1973.
76. Rosenfeld A.: Fuzzy groups. J. Math. Anal. Appl. 35(3)(1971), 512–517.
77. Samhan M. A.: Fuzzy quotient algebras and fuzzy factor congruences. Fuzzy
Sets and Systems 73(1995), 269–277.
78. Schweizer B., Sklar A.: Statistical metrics. Pac. J. Math. 10(1960), 313–334.
79. Schweizer B., Sklar A.: Probabilistic Metric Spaces. North–Holland, Amster-
dam, 1983.
80. Selman A.: Completeness of calculii for axiomatically defined classes of alge-
bras. Algebra Universalis 2(1972), 20–32.
81. Šešelja B., Tepavčević A.: On a generalization of fuzzy algebras and congru-
ences. Fuzzy Sets and Systems 65(1994), 85–94.
82. Sloman S.A., Rips L.J.: Similarity and Symbols in Human Thinking. A Brad-
ford Book, MIT Press, Cambridge, MA, 1998.
83. Takeuti G., Titani S.: Globalization of intuitionistic set theory. Annals of Pure
and Applied Logic 33(1987), 195–211.
84. Taylor W.: Characterizing Mal’cev conditions. Algebra Universalis 3(1973),
351–397.
85. Trillas E., Valverde L.: An inquiry into indistinguishability operators. In:
Skala H. J., Termini S., Trillas E.: Aspects of Vagueness. Reidel, Dordrecht,
1984. , pp. 231–256.
86. Tversky A.: Features of similarity. Psych. Rev. 84(1977), 327–352.
87. Valverde L.: On the structure of F-indistinguishability operators. Fuzzy Sets
and Systems 17(1985), 313–328.
88. Vaught R. L.: Set Theory. An Introduction. Birkhäuser, Boston, 1985.
89. Vychodil V.: A note on congruence permutability and fuzzy logic. Soft Com-
puting (to appear).
90. Vychodil V.: Cut and weakening in fuzzy Horn logic. Logic J. of IGPL (to
appear).
91. Vychodil V.: Continuous fuzzy Horn logic (submitted).
92. Vychodil V.: Direct limits and reduced products of algebras with fuzzy equal-
ities (submitted).
93. Vychodil V.: Extended fuzzy equational logic (submitted).
94. Ward M., Dilworth R. P.: Residuated lattices. Trans. Amer. Math. Soc.
45(1939), 335–354.
95. Weaver N.: Generalized varieties. Algebra Universalis 30(1993), 27–52.
96. Weaver N.: Quasi-varieties of metric algebras. Algebra Universalis 33(1995),
1–9.
97. Wechler W.: Universal Algebra for Computer Scientists. Springer-Verlag,
Berlin Heidelberg, 1992.
References 271

98. Ying M.: The fundamental theorem of ultraproduct in Pavelka’s logic. Z. Math.
Logik Grundlagen Math. 38(1992), 197–201.
99. Zadeh L. A.: Similarity relations and fuzzy orderings. Inf. Sci. 3(1971), 159–
176.
100. Zadeh L. A.: Fuzzy sets. Inf. Control 8(3)(1965), 338–353.
101. Zhang C.: Fuzzy complete inner-unitary subsemigroups and fuzzy group con-
gruences on a regular semigroup. Fuzzy Sets and Systems 112(2000), 327–332
Table of Notation

Sets and Structures


A, B, . . . , U, . . . sets 1
a, b , . . . , u, . . . elements of sets 1
∈ membership in set 1
∅ empty set 1
A⊆B set A is subset of set B 1
A ∩ B, A ∪ B intersection and union of sets A and B 1
A−B difference of sets A and B 1
u1 , . . . , un ordered n-tupple of elements 1
U1 × · · · × Un direct product of sets U1 , . . . , Un 1
Un direct product U × · · · × U (n times) 1
r ⊆ U1 × · · · × Un r is relation between sets U1 , . . . , Un 1
r−1 relation inverse to r 2
f:A→B f is mapping of set A to set B 2
f ◦g composition of mappings f and g 2
U
2
Q set of all subsets of U 2
i∈I Ui direct product of system
Q of sets Ui 2
πj j-th projection of i∈I Ui to Uj 2
N positive integers 2
Z integers 2
Q rationals 2
R reals 2
N0 non-negative integers 2
α, . . . , κ ordinal and cardinal numbers 3
U/E factor set of set U by equivalence relation E 3

 partial order 3
 A, inf A infimum of A 4
A, sup A supremum of A 4
R, F, σ type of structure 6
σ(s)  arity of relation/function symbol 6
M, M, RM , F M structure 6
 
M, F M algebra 6
274 Table of Notation

Structures of Truth Degrees


||ϕ|| truth degree of formula ϕ 7
a, b, c, . . . truth degrees 7
L set of truth degrees 7
L complete residuated lattice 11
⊗ truth function of conjunction (multiplication) 11
→ truth function of implication (residuum) 11
↔ truth function of equivalence (biresiduum) 13
¬ truth function of negation 13
an n-th power of a 13

truth stresser 14

L
 complete residuated lattice with truth stresser 14
i∈I Li ordinal sum of Li ’s 21

Fuzzy Sets
A: U → L L-set A in universe U 33
a, b , c, . . . elements of universe 33
Supp(A) support set of A 34
A = {A(u1 )/u1 , . . . definition of L-set A 34
a
A a-cut of A 34
LU , LU set of all L-sets in U 35
∅U empty L-set in U 35
1U full L-set in U 35
R: U × U → L binary L-relation in U 35
R−1 inverse L-relation of R 35
R1 ◦ R2 ◦-composition of R1 and R2 35
S(A, B) subsethood degree of A in B 35
A≈B equality (similarity) degree of A and B 35
A⊆B crisp subsethood relation 36
[ u ]θ class of L-equivalence θ 39
U, ≈ set U with L-equality ≈ 44
θh kernel of ≈-morphism h 45
U/θ, ≈U/θ factor set with fuzzy equality 49

Pavelka-style Fuzzy Logic


Fml set Fml of all formulas 50
S L-semantics for Fml 50
E, E  , . . . evaluations 50
E(ϕ), ||ϕ||E truth degree of ϕ ∈ Fml in E ∈ S 50
T, T  , . . . theories over Fml and L 52
||ϕ||S
T , ||ϕ||T degree to which ϕ semantically follows from T 53
R = Rsyn , Rsem deduction rule for Fml and L 53
Rsyn syntactic part of R 53
Rsem semantic part of R 53
T A,R ϕ, a ϕ, a is provable from T 54
|ϕ|A,R
T , |ϕ|T degree of provability of ϕ from T 54
ϕ, a L-weighted formula 55
Md(T ) set of all models of T 56
Th(K) theory of K 56
Table of Notation 275

Algebras with Fuzzy Equalities


≈, F, σ type of algebra with fuzzy equality 60
≈ equality symbol 60
σ(f ) arity of function symbol f 60
M, M , N, . . . L-algebras 60
M, M  , N, . . . universes of L-algebras M, M , N, . . . 60

≈M , ≈M , ≈N , . . . relational parts (L-equalities) of M, M , N, . . . 60

F M, F M , F N, . . . functional parts of M, M , N, . . . 60
ske(M) skeleton of M 60
fM σ(f )-ary function f M ∈ F M 60
Sub(M) collection of all subuniverses of M 68
[N ]M subuniverse of M generated by N 68
θ congruence L-relation 69
EqL (M ) set of all L-equivalences on M 69
ConL (M) set of all congruences on M 69
θ(R) congruence generated by R 73
θ(a/b , c ) principal congruence 73
M/θ, M/1θ factor set of M by 1 θ 74
M/θ factor L-algebra of M by θ 74
[ a ]θ congruence class containing a 74
h: M → N morphism from M to N 76
M= ∼N isomorphic L-algebras 77
idM , idM identity mapping on M (universe of M) 77
h(M ) image of M 77
h−1 (N ) inverse image of N 77
h(M) subalgebra induced by h(M ) 77
hθ : M → M/θ natural mapping (morphism) 78
φ/θ congruence on M/θ induced by φ ∈ ConL (M) with θ ⊆ φ 80
Nθ union of certain congruence classes 81
θ|N restriction of θ to N 81
[θ1 , θ2 ]
Q interval of congruences 82
i∈I Mi direct product of L-algebras Mi 83
πi i-th projection 84
M1 × M2 direct product of M1 , M2 85
θ, θ∗ pair of factor congruences 86
x, x , y, y  , . . . variables 91
X, X  , Y, Y  , . . . sets of variables 91
t, t , s, s , . . . terms 91
T (X) set of all terms in variables X 91
|t|x number of occurrences of x in t 91
t(x1 , . . . , xk ) term t in variables x1 , . . . , xk 92
var(t) set of variables occurring in t 92
tM term function 92
T(X) term L-algebra over X 93
h homomorphic extension of h 93
I,
 ≤ partially ordered index set 98
i∈I Mi direct union of L-algebras Mi 99
κ
i∈I Mi κ-direct union of L-algebras Mi 99
θ∞ L-relation induced by a weak direct family 104
276 Table of Notation

lim
 Mi  direct limit of L-algebras Mi 106
i∈I Mi /θ∞ support set of lim Mi 106
X, R abbreviation
Q for T(X)/θ(R) 110
M subset of i∈I M i 115
θ congruence on M  115
M L-algebra generated by M  115
[[a ≈ b ]]X degree of equality over X 121
θQF congruence induced by filter F 122
F Mi reduced product Q of L-algebras Mi modulo F 122
MX abbreviation for i∈X Mi 123
MX support set of MX 123
hXY morphism MX to MY 123
K, K , . . . classes of L-algebras 127
H(K) homomorphic images of K 127
I(K) isomorphic images of K 127
S(K) subalgebras from K 127
P(K) direct products of families from K 127
PS (K) subdirect products of families from K 127
U(K) direct unions of directed families from K 127
L(K) direct limits of direct families from K 127
PR (K) safe reduced products of families from K 127
HS, SP, HPH, . . . composed class operators 127
O, O , O1 , O2 , . . . class operators 127
Fuzzy Equational Logic
t ≈ t identity 141
v, v  , w, w , . . . valuations 141
tM,v value of t in M under v 141
t ≈ t M,v truth degree of t ≈ t in M under v 141
t ≈ t M truth degree of t ≈ t in M 142
(x/r) substitution 143
t(x/r) term resulting from t by substitution of r for x 143
(ERef) equational rule of reflexivity 143
(ESym) equational rule of symmetry 143
(ETra) equational rule of transitivity 143
(ERep) equational rule of replacement 143
(ESub) equational rule of substitution 143
t ≈ t K truth degree of t ≈ t in K 144
t ≈ t Σ degree to which t ≈ t semantically follows from Σ 144
t ≈ t , a weighted identity 145
Σ A,R t ≈ t , a t ≈ t , a is provable from Σ using R 145
|t ≈ t |RΣ , |t ≈ t |Σ

degree of provability of t ≈ t from Σ 145
V(K) variety generated by K 154
ΦK (X) set of all φ ∈ ConL (T(X)) such that T(X)/φ ∈ IS(K) 156
θK (X) intersection of congruences from ΦK (X) 156
x abbreviation for [x]θK (X) 156
X generators of FK (X) 156
FK (X) K-free L-algebra over X 156
Table of Notation 277

Fuzzy Horn Logic


h(P ) endomorphic image of P ∈ LT (X)×T (X) 175
var(P ) set of variables occurring in P ∈ P 175
P, Q, . . . proper families of premises 175
P, P  , Q, Q , . . . L-sets of premises 175
Pa restriction of P induced by a ∈ L 176
P i (t ≈ t ) P-implication 177
P i (t ≈ t ), a weighted P-implication 177
PFin Horn restriction of P 178
Pω finitary restriction of P 178
Pκ restriction of P by infinite cardinal κ 178
P M,v truth degree of premises P ∈ P in M under v 179
P i (t ≈ t )M,v truth degree of P i (t ≈ t ) in M under v 179
ϕ, a M,v truth degree of ϕ, a in M under v 179
Mod(Σ) model class of Σ 181
Impl(K) theory of K (L-set of P-implications) 181

ModL (Σ) model class of Σ (explicit L∗ ) 181
ImplP(K) theory of K (explicit P) 181

ImplL P (K) theory of K (explicit P and L∗ ) 181
Implκ(K) abbreviation for ImplPκ(K) 181
Horn(K) abbreviation for ImplPFin(K) 181
P i (t ≈ t )K truth degree of P i (t ≈ t ) in K 181
P i (t ≈ t )Σ degree to which P i (t ≈ t ) semantically follows from Σ 181
Γ R ϕ, a ϕ, a is provable from Γ using R 182
R
|P i (t ≈ t )|Σ degree of provability of P i (t ≈ t ) from Γ using R 183
P (x/r) L-relation resulting from a substitution (x/r) 184
(Ref) rule of reflexivity 185
(Sym) rule of symmetry 185
(Tra) rule of transitivity 185
(Rep) rule of replacement 185
(Ext) rule of extensivity 185
(Sub) rule of substitution 185
(Mon) rule of monotony 185
Σ, Γ, . . . L-sets of P-implications 189
S, S  , . . . P-indexed systems of L-relations SP , SP , . . . 189
ΣS L-set of P-implications corresponding to S 189
SΣ P-indexed system of L-relations corresponding to Σ 189
ΣP , ΓP , . . . member of SΣ , SΓ , . . . 189
N
 system of P-indexed systems 197
i∈I S i  intersection of Si ’s 197
i∈I Si P member of i∈I Si determined by P ∈ P 197
Smax greatest P-indexed system of L-relations 197
SΣ  , SΓ  , . . . semantic closures of SΣ , SΓ 198
ΣP , ΓP , . . . members of SΣ  , SΓ  , . . . determined by P ∈ P 198
Σ , Γ  , . . . semantic closures of Σ, Γ, . . . 198
ΣFin Horn restriction of Σ 200
SΓ  deductive closure of SΓ 204
ΓP member of SΓ  204
278 Table of Notation

Γ deductive closure of Γ 205


τ, τ1 , τ2 , . . . (composed) substitutions 205
(Sup) rule of supremum 212
 strict continuous Archimedean t-norm 216
dP particular L-set of premises 216
(CExt) crisp version of extensivity 219
(CMon) crisp version of monotony 219
hP Q morphism induced by P i (t ≈ t ), a 232
RK (M) sur-reflection of M in K (L-algebra) 236
rM sur-reflection of M in K (morphism) 236
HK (M) set of all θ ∈ ConL (M) such that M/θ ∈ K 237
PK (M) direct product of M/θ over all θ ∈ HK (M) 237
PM proper family of premises LT (M )×T (M ) 238
PM particular member of PM 238
K
ΣM PM -theory of M over K 238
VS (K) semivariety generated by K 246
VQ (K) quasivariety generated by K 249
RK (X, R) sur-reflection of T(X)/θ(R) in K 249
Index

a-cut 34 complete
abstract lattice 4
class 127 residuated lattice 10
logic 56 with truth stresser 14
additive generator of a continuous completeness 212
Archimedean t-norm 27 in Pavelka’s sense 56
adjoint pair 11 ◦-composition 35
adjointness 9, 11 composition of mappings 2
algebra 6 congruence 7, 69
with L-equality 60 axiom 44
arity 2, 6, 60 class 74
atomic inequality 129 -continuity 216, 230
automorphism 77 -continuous
class 230
Banaschewski-Herrlich criterion 232 theory 216
bijection 2 crisp
binary L-relation 35
L-set 35
biresiduum 13
premises 218
BL-algebra 14
Boolean algebra 14
bound deduction rule 53
lower 4 deductive
upper 4 closure 150, 204
system 184
class degree
closed under O 127 of provability 54, 145, 183
of L-equivalence relation 39 of semantic entailment 53
of equivalence relation 3 of semantic entailment in fuzzy
operator 127 equational logic 144
closure of semantic entailment in fuzzy Horn
operator 5 logic 181
system 5 denumerable set 3
compact element 4 diagonal fill-in 79
compatibility 44, 60 direct
280 Index

family 103 restriction 178


indecomposability 86 finite
limit 106 L-set 34
property 107 premises 176
product 1, 7, 83 restriction 35, 200
union 99 set 2
directed finitely presented L-algebra 110
family 98 formula
index set 98 L-weighted 55
DLP 107 truth-weighted 55
weighted 55
element 1 free
embedding 76 generators 93, 128
empty L-set 35 L-algebra 156
endomorphic image 175 full L-set 35
endomorphism 77 fully invariant
epimorphism 76 L∗ -closure operator 192
≈-morphism 45 closure operator 192
equality closure system 192
axioms 44 congruence 147, 190
symbol 60 function 2
equation implication 173 antitone 3
equational characteristic 2
class 145 isotone 3
class of a theory 145 left-continuous 22
theory 145 monotone 3
theory of a class 145 non-decreasing 3
equicontinuous non-increasing 3
function 129 symbol 60
satisfaction of atomic inequalities fuzzification operator 164
129 fuzzy
equivalence 3 concept lattice 65
extensivity 185, 190, 219 congruence 132
extent 65 equality 37
equational logic 140
factor equivalence 37
algebra 7 Horn logic 171
congruence 86 logic
L-algebra 74 abstract 50
set 3 Pavelka-style 50
with fuzzy equality 49 partition 40
filter 3, 121 relation 35
Fréchet filter 3 set 33
proper 3
safe 125 G-algebra 14
trivial 3 Gödel algebra 14
finitary Galois connection 5
implication 173, 178 graded
premises 176 equality 35
Index 281

subsethood 35 L-equality 37
truth 7 L-equivalence 37
greatest L∗ -implicational
element 4 Horn subtheory 200
image 236 P-theory 182
P-theory of a class 182
Hasse diagram 3 P-indexed system 190
Heyting algebra 14 L-partition 40
homomorphism 76 L-relation 35
Horn L-semantics 50
clause 173, 178 L-set 33
restriction 178 L-weighted proof 54, 145, 182
subtheory 200 language
truth stresser 17 of fuzzy equational logic 140
of fuzzy Horn logic 173
idempotency 127 lattice 4, 6
identity 141, 173 algebraic 4
L-weighted 145 bounded 7
mapping 77 compactly generated 4
truth-weighted 145 Noetherian 4
weighted 145 residuated 11
image 77 complete 10
implication 173 least element 4
between atomic inequalities 129 length 54, 145, 182
L-weighted 176 limit cone 107
truth-weighted 176 linear order 4
weighted 176 logic
implicational truth stresser 16 complete (in Pavelka’s sense) 56
infimum 4 sound (in Pavelka’s sense) 56
infinite set 3 lower cone 4
injection 2
injectivity 232 mapping 2
intent 65 bijective 2
interior operator 5 injective 2
intersection 1 surjective 2
inverse maximal element 4
L-relation 35 membership function 34
relation 2 metric 42
isomorphic L-algebras 77 algebra 129
isomorphism 77 generalized 42
theorem 78, 81 minimal element 3
model 52
κ-direct union 99 of a theory of fuzzy equational logic
κ-directed family 98 144
K-valent premises 218 of a theory of fuzzy Horn logic 181
kernel 45 monoid 6
monomorphism 76
L-algebra 60 monotony 185
L∗ -closure operator 49 ∗
-monotony 190, 219
282 Index

morphism 7, 76 reflection 235


multiplication 11 relation 2
MV-algebra 14 antisymmetric 3
reflexive 3
n-ary term 92 symmetric 3
natural transitive 3
mapping 78 relative pseudocomplement 14, 203
morphism 78 residuated lattice 11
negation 13 residuum 11
Noetherian lattice 4 restriction 35, 68, 81
non-trivial L-algebra 68
safe
occurrence 91 filter 125
open problem 135–137, 170, 266 reduced product 125
operation 2 satisfaction of atomic inequalities 129
with L-sets 36 semantic closure 149, 198
ordered n-tuple 1 semivariety 173, 246
separation 87
partial order 3
set 1
partially ordered set 3
set with L-equality 44
partition 3
similarity 37
fuzzy 40
singleton 34
P-finitary
skeleton 60
implication 178
soundness (in Pavelka’s sense) 56
implicational class 182
stability 190
P-Horn
structure 6
class 182
clause 178 of truth degrees 8
Π-algebra 14 subalgebra 7, 68
P-implication 176 κ-generated 68
P-implicational class 182 finitely generated 68
P-indexed system 189 subdirect
power in residuated lattice 13 embedding 88
premise 171, 175 irreducibility 88
presentation 110 product 88
principal congruence 73 representation 89
product algebra 14 subset 1, 36
projection map 84 subsethood 35
proof (L-weighted) 54, 145, 182 substitution 143, 185
proper family 175 subterm 142
provability 54, 145, 182 subuniverse 7, 68
pseudoinverse 27 support 34
pseudometric 42 supremum 4
generalized 42 sur-reflection 235
P-theory 238 sur-reflective class 173, 235
surjection 2
QF condition 249
quasivariety 173, 249 t-norm 12, 22
Archimedean 26
reduced product 122 continuous 25
Index 283

left-continuous 22 type 6, 60
strict 26
strictly monotone 26 ultrafilter 3
term 91 UMP 128
L-algebra 93 union 1
function 92 universal algebra 6
theory 52, 144, 181 universal mapping property 128
universe 1, 33
trivial L-algebra 68
upper cone 4
truth
degree 7, 141, 142 valuation 141
degree of implication 178 value of term 141
stresser 14 variable 91
Horn 17 variety 154, 173
implicational 16
tuple 1 weak direct family 103

Das könnte Ihnen auch gefallen