1. Hyperidentitis and Coidentities
By definition, a Boolean algebra is a complemented bounded distributive lattice. In the nineteenth century, George Boole’s [1] attempt to formalize propositional logic led to the concept of Boolean algebra. Historically, lattice theory started with Boolean algebras which have played a fundamental role in the development of the lattice theory.
Boolean functions, i.e.
-valued functions of a finite number on
-valued variables, are among the most fundamental objects that have been investigated in pure and applied mathematics, especially in discrete mathematics, algebra, mathematical logic and computer science. The Boolean algebra of Boolean functions of n variables is a free Boolean algebra on n free generators. The bounded lattice of monotone Boolean functions of n variables is a free bounded distributive lattice on n free generators (R. Dedekind [2]). These are the first Boolean structures in both mathematics and applied mathematics, and they are used extensively in discrete mathematics.
Algebra and model theory study the connections between formal languages and their interpretations in models and algebras. The simplest and most widespread formal language is the first-order language (A. Church [3], A. I. Mal’tsev [4]-[7], G. Grätzer [8], C. Chang with H. Keisler [9], S. Burris with H. P. Sankappanavar [10], B. I. Plotkin [11]). The founders of the first-order language (logic) are Löwenheim, Skolem, Gödel, Łoś, Tarski, Mal’tsev and Birkhoff.
However, there exist very commonly encountered, classical algebraic structures that are not axiomatizable by the first-order formulae (logic). For example, rings, associative rings, commutative rings, associative-commutative rings, fields, or fields of fixed characteristics are axiomatized by first-order formulae, but their multiplicative groupoids, multiplicative semigroups and multiplicative groups are not, because these classes of groupoids, semigroups and groups are not closed under elementary equivalence (A. I. Mal’tsev, S. Kogalovskii [12], G. Sabbagh [13]). The situation is analogous for near-fields (M. Hall [14]), Grätzer algebras (G-fields) [15]-[18], topological rings and topological fields (L. S. Pontryagin [19]). However, if some class
of lattices is axiomatized by the first-order formulae, then the class of its multiplicative semigroups is also axiomatized by the first-order formulae [20].
Characterizations of such semigroups and groups are the most important problems in modern algebra, logic and topology. L. Fuchs ([21], p. 324) called the characterization of multiplicative groups of fields a big problem.
This is why it is necessary to widen the formal language to allow to express phenomena that the first-order logic can not capture.
An important extension of the first-order logic (language) is the second-order logic (language), described in detail in [3] [5] [6] [9] (also see [22]). The second-order formulae consist of the same logical symbols of
of individual and functional (predicate) variables, which are used in the first-order formulae. The difference is that in the second-order formulae, the quantifiers
can be applied not only to individual variables, but also to functional (or to predicate) variables. Investigations of the second-order formulae (logic) go back to L. Henkin, A. I. Mal’tsev, A. Church, S. Kleene, A. Tarski. Many important mathematical concepts can be written in the second-order language. Consequently, investigation of the theories of the second-order languages (logic) is one of the central problems of algebra and mathematical logic (Mal’tsev [5]). For a compactness theorem for quasi-universal second-order formulae see [5] [22].
Starting with the 1960’s the following second-order formulae were studied in various domains of algebra and its applications (see [5] [18] [20] [23]-[63])
(1.1)
(1.2)
(1.3)
(1.4)
(1.5)
where
are words (hyperterms) in the functional variables
and the individual (object) variables
. The first formula is called hyperidentity or
-identity (see [20] [27] [34] [35] [39] [40] [43] [50] [54] [64]-[69], and also [70]); the second (third, fourth, fifth) formula is called an
-identity (
-identity,
-identity,
-identity). Sometimes the
-identity is called a generalized identity [27] [71], the
-identity is called a coidentity [30] [39] (also see [70]) and
-identity is called a hybrid identity [32] [50]. The satisfiability of these second-order formulae in an algebra
is understood by functional quantifiers
and
, meaning: “for every value
of the corresponding arity” and “there exists a value
of the corresponding arity”. It is assumed that such a replacement is possible, that is
where
is the arity of S, and
is called the arithmetic type of
. A T-algebra is an algebra with arithmetic type T. We say that a hyperidentity (coidentity) is called a T-hyperidentity (T-coidentity), if the set of arities of all functional variables accuring in it is contained in T.
Hyperidentities and coidentities are usually presented without quantifiers:
. The coidentity
is said to be satisfied in the algebra
, if there exist values for object variables
from Q, such that the equality
holds when every functional variable in it is replaced by any operation of the corresponding arity from Σ (a possibility of such replacement is supposed, too). In addition, the object variables in the notation of the coidentity
are replaced by corresponding fixed values from Q.
Coidentities can be defined as second-order formulas of the following form, as well:
Example 1.1. In any multioperator Ω-group the following coidentity is valid:
for all
, where all object variables are replaced by the zero element of the Ω-group [70] [72] [73].
Theorem 1.1 (J. von Neumann). Let
be a modular lattice and
. The sublattice of L, generated by the elements
, is distributive iff the following coidentity of left distributivity
holds in
.
Second-order formulae with analogous predicative quantifiers in models and algebraic systems are also often used in mathematical logic. For example, finiteness, the axiom of well-ordering, the continuum hypothesis, the property of being countable and others can be formulated within the second-order logic. Every algebra can be embedded into an algebra of functions whose identities are equivalent to hyperidentities or
-identities (A. Cayley, P. M. Cohn, unpublished).
The above semantics is different from the standard definition of the semantics of second-order formulas (logics) ([3] [5] [6]). Studying logic via semantics is a well-established and very active branch of mathematical logic with many applications in computer science and elsewhere.
A variety (or equational class) is a class of algebras (all of the same similarity type or signature, or Ω-algebras) closed under the formation of products, subalgebras and homomorphic images. Equivalently, a variety is a class of algebras defined by a set of equations (identities). A hypervariety is a class of algebras (all of the same arithmetic type or the same signature) defined by a set of hyperidentities.
Note that even in the classical algebras (groups, semigroups, quasigroups, rings, etc) the characterization of varieties in which every finitely generated algebra is finite is an open problem (see S. Adian [74] [75], E. Zel’manov [76]-[79]).
A variety of algebras is called Burnside variety (W. Burnside) [80] if its any finitely generated algebra is finite.
In 1902, Burnside proposed the following problem: Is a finitely generated group satisfying an identity of the form
necessarily finite? The answer in yes for
and 6. The case
is trivial, the case
was settled by Burnside [81], the case
by Sanov [82] and the case
by M. Hall [14]. The original problem finally received a negative answer by Novikov and Adian in 1968 [83] (see also [74] [84]).
The numbers m and n in hyperidentity (1.1) are called the functional and object rank, respectively. A hyperidentity is said to be nontrivial if its functional rank is >1, and it is called trivial otherwise (
). A hyperidentity is called n-ary, if its functional variables are n-ary. For
the n-ary hyperidentity is called unary, binary, ternary. A formula (hyperidentity, coidentity, ...) is called a formula (hyperidentity, coidentity, ...) of algebra
, if it is satisfied in algebra
. Let V be a variety or another class of algebras. A hyperidentity (coidentity, ...)
is called a hyperidentity (coidentity, ...) of V if it is a hyperidentity (coidentity, ...) for any algebra
.
The concept of hyperidentity is present in many well known notions. For example, an algebra
is said to be Abelian (A.G.Kurosh [85]) or entropic (medial) if the following nontrivial hyperidentity
is valid for all
. An algebra
is said to be idempotent if any its operation is idempotent, i.e. if the following hyperidentity of idempotency
is valid for all
.
A mode is an idempotent and entropic algebra (studied in monographs [86] [87]). A distributive bisemilattice (multisemilattice) [88] is a binary algebra with semilattice operations satisfying the following (nontrivial) hyperidentity of distributivity
A doppelsemigroup (see [89]-[97]) is an algebra with two binary operations satisfying the following non-trivial hyperidentity of associativity
Binary algebras satisfying the hyperidentity of associativity
we call hyperassociative algebras. These algebras under the name of Γ-Semigroups, Γ-Semirings (or gamma-Semigroups, gamma-Semirings) also were considered by various authors [98]-[112].
Example 1.2. Let A and B be non empty sets, Σ be the set of all mappings from B to A and Q be the set of all mappings from A to B. Then we can consider every element
as a binary operation on Q:
where
and
is the usual superposition of mappings. So we obtain a binary algebra
, which is hyperassociative.
The investigation of hyperidentities (coidentities) is a relatively new, actively developing field of pure and applied algebra. The concept of hyperidentity offers a high-level approach to algebraic questions, leading to new results, applications and problems. In particular, the investigation of hyperidentities is useful from the point of view of new technologies too, via optimization problems of block diagrams [40]. For applications of hyperidentities in discrete mathematics and topology see [62] [113]-[116]. For characterization of Sheffer functions and primal algebras via hyperidentities see (K. Denecke, R. Pöschel [113] [114]).
Hyperidentities are also “identities” of algebras in the category of bihomomorphisms (
), where
which were studied in the monograph [39]. Thus, we obtain the categorical definition of hyperidentities. More about the application of such morphisms in the cryptography can be found in [117]. For the categorical definition of
-identities and Birkhoff style theorem for
-identities see [29] (Belousov’s problem).
Hyperidentities for binary algebras with quasigroup operations were first considered by V. D. Belousov [27] (as a special case of
-identities which earlier is considered by R. Schauffler ([23] [24]) in coding theory) and then J. Aczel [118], about the classification of associative and distributive hyperidentities in binary algebras with quasigroup operations. Currently, more general results about these and other classifications of hyperidentities may be found in [20] [39] and [40]. Observe that in algebras with quasigroup operations many
-identities are equivalent to hyperidentities (see [20]). Coidentities for algebras were first considered in [30] [39].
The multiplicative groups of fields have been characterized in [119] via hyperidentities. On the base of these results the concept of binary G-spaces is developed in topology [115]. The hyperidentities of varieties of lattices, modular lattices, distributive lattices, Boolean algebras, De Morgan algebras and weakly idempotent lattices have been characterized in the works [63] [120]-[122].
Theorem 1.2. The variety of lattices satisfies the following hyperidentities:
(1.6)
(1.7)
(1.8)
(1.9)
And conversely, every hyperidentity of the variety of lattices is a consequence of the hyperidentities: (1.6)-(1.9) (see [63]).
Theorem 1.3. The variety of modular lattices satisfies the following hyperidentities: (1.6)-(1.9) and
(1.10)
And conversely, every hyperidentity of the variety of modular lattices is a consequence of the hyperidentities: (1.6)-(1.10) (see [63]).
Theorem 1.4. The variety of distributive lattices satisfies the following hyperidentities: (1.6)-(1.8) and
(1.11)
And conversely, every hyperidentity of the variety of distributive lattices is a consequence of the hyperidentities: (1.6)-(1.8), (1.11) (see [63]).
Theorem 1.5. The variety of Boolean algebras satisfies the following hyperidentities: (1.6)-(1.8), (1.11) and
(1.12)
(1.13)
(1.14)
And conversely, every hyperidentity of the variety of Boolean algebras is a consequence of the hyperidentities: (1.6)-(1.8), (1.11)-(1.14).
All hyperidentities of the variety of Boolean algebras are consequences of one of its hyperidentities, i.e. the hyperequational theory of the variety of Boolean algebras is one-based (see [63]).
Theorem 1.6. The variety of De Morgan algebras satisfies the following hyperidentities: (1.6)-(1.8), (1.11), (1.12) and
(1.15)
(1.16)
(1.17)
(1.18)
(1.19)
And conversely, every hyperidentity of the variety of De Morgan algebras is a consequence of the hyperidentities: (1.6)-(1.8), (1.11), (1.12), (1.15)-(1.19).
The hyperequational theory of the variety of De Morgan algebras is not one-based (see [63]).
A hyperidentity
is called termal or t-hyperidentity of the algebra
if it is valid in the term algebra
. Let V be a variety. A hyperidentity
is called a termal hyperidentity of V if it is a termal hyperidentity for any algebra
. Termal hyperidentities for varieties were first considered by W. Taylor ([34]) (as a special case of Mal’tsev conditions for varieties) for characterization of classes of varieties which are closed under formation of equivalent varieties, products of varieties, reducts of varieties and subvarieties. Since the operations of an algebra are included in the set of term operations (clone) of the algebra, the concept of termal hyperidentity of a variety is stronger than the concept of hyperidentity. In particular, the variety of rings (even commutative rings) does not have termal hyperidentities except
, but has hyperidentities.
Termal hyperidentities of varieties of groups and semigroups have been characterized by G. Bergman [35] (also see [36]). Termal hyperidentities of the variety of lattices and of the variety of semilattices were studied by R. Padmanabhan and P. Penner ([37] [44] [123]).
Hyperidentities and coidentities in algebras as an individual direction of investigations, were first presented in the monographs [20] [39] (see also [124]-[126]).
2. Bihomomorphisms of T-Algebras
Let
be a T-algebra and
be a T’-algebra, where
. The pair
of maps
and
is called a bihomomorphism from T-algebra
into T’-algebra
and is denoted by
or
, if the map
preserves the arity of operations, i.e.,
for any operation
, and the equality
holds for all n-ary operations
and for all
.
Now we define the concepts of subalgebra, congruence relation and quotient algebra for T-algebras.
An algebra
is called a subalgebra of the algebra
if
and every operation from
is the restriction of some operation from Σ (to the subset
). For example, the additive semigroup of all natural numbers is a subalgebra of the ring of integers, every Abelian group is a subalgebra of some ring, and, a fortiori, every groupoid is a subalgebra of some ring. Every semi-lattice is a subalgebra of some distributive lattice and every distributive lattice is a subalgebra of some Boolean algebra, and so on. When we wish to specify the arithmetic type of subalgebras, we call them T-subalgebras.
An intermediate notion here is that of a subsystem. Let
be an arbitrary T-algebra, and let
,
, where
and
. A pair
is called a subsystem of the T-algebra
if
is closed with respect to all the operations of
. Subsystems of the form
are called principal.
A class of T-algebras is called hereditary if it contains all the T-subalgebras of any T-algebra from itself.
Let
be a T-algebra and r and
be equivalence relations defined respectively on the sets A and Σ. The pair
is called a congruence relation (or simply congruence) of the T-algebra
if the following two conditions hold:
1)
preserves the arity of the operations, i.e., implies
, for any two operations
;
2) the relations r and
are compatible in the following sense:
for any elements
and any n-ary operations
.
The class of all congruences
of an T-algebra
is a lattice and denoted by
.
If r is a congruence relation of
considered as an Ω-algebra for some type Ω, then
is obviously a congruence of
considered as a T-algebra and vice versa, where
is the identity relation of Σ, i.e.,
.
If
is a congruence of a T-algebra
then every element
of the quotient set
defines an operation on the quotient set
in the following way:
where
,
,
.
The definition of a congruence implies that the operation
is well defined. As a result we shall obtain a quotient algebra of the T-algebra
modulo the congruence
, denoted by
.
A congruence
of a T-algebra
is said to be nuclear if different elements of
define different operations on
, that is
where
,
.
For a congruence
of a T-algebra
the natural bihomomorphism
is defined by
,
,
,
.
3. Biautomorphisms of T-Algebras
A bihomomorphism
is called a biepimorphism (bimonomorphism, biisomorphism) if the maps
and
are surjective (injective, bijective). A bihomomorphism
is called a semiepimorphism, (semimonomorphism, semiisomorphism) if the map
is surjective (injective, bijective). A bihomomorphism (biisomorphism) of a T-algebra into itself is called a biendomorphism, (biautomorphism). Any semimonomorphism
of T-algebras is a bimonomorphism:
where
. We denote
(or
) for biisomorphic T-algebras
and
, i.e. if there exists a biisomorphism
.
A congruence
of a T-algebra
is said to be fully invariant if it preserves any biendomorphism
of this algebra, that is,
and
where
and
.
A T-algebra
is called a bihomomorphic image of the T-algebra
if there exists a biepimorphism
. A class of T-algebras is said to be bihomomorphically closed if along with each T-algebra it contains every bihomomorphic image of that algebra under any bihomomorphism.
The pair
of identical maps
and
is an biautomorphism of T-algebra
.
The set of all biendomorphisms (biautomorphisms) of the same T-algebra forms a semigroup with identity (group) under the componentwise multiplication of pairs:
We will denote the group of all biautomorphisms
of the same T-algebra
by
. The biautomorphisms of the form
, where
is the identical mapping, are ordinary automorphisms and form a group denoted by
. Obviously
, i.e.,
is an invariant subgroup of the group
.
Theorem 3.1 ([39]). For any group G there exists a binary algebra
with the hyperidentity of associativity
such that
and
is singletone.
Theorem 3.2 ([39]). For any group G there exists a binary algebra
with the hyperidentity of distributivity
such that
and
.
Below we will formulate a more general result.
Theorem 3.3. Let G be an arbitrary group with an invariant subgroup H. Then there exists an algebra
such that
and
, where the last isomorphism is induced by the first isomorphism.
Let
and
be T-algebras. With each bihomomorphism
two congruencies are associated, called the kernel and the weak kernel of that bihomomorphism and denoted by
and
respectively. They are defined as follows:
where
Clearly
and if
is a surjection then
.
But in general
. For example, let
be an endomorphism of a non-zero ring, where
is the zero map and
is the identical. Then
.
In general we have
, where the partial order between two congruencies is defined in the natural way:
The class of all congruencies of the same T-algebra forms a complete algebraic lattice with respect to this partial order “≤”.
4. Superproducts and Filtered Products of T-Algebras
Now we define the concepts of direct and subdirect products for T-algebras.
Let
be T-algebras of the same arithmetic type. We form the Cartesian product
. In addition we form the Cartesian product
and define the subset
as the set of all possible functions
satisfying the following two conditions:
1)
for all
;
2)
for all
,
that is, for fixed F the arity of an operation
does not depend on
.
The set
can be identified with the set of all possible systems
of operations of the same arity, with one taken from each set
. We write
. If
and
, then F defines an n-ary operation on the set
defined componentwise:
where
.
As result we obtain a T-algebra
, which is called the (direct) product of the T-algebras
, or their direct T-product and is written
.
The direct product of T-algebras
is also called their superproduct. The superproduct of two T-algebras
and
is denoted by
. For example, the superproduct
of the two lattices
and
is an algebra
with four binary operations:
which operate component-wise.
The semisubalgebra (reduct) of direct product is called a semidirect product.
The projection maps
for each
are defined by
These maps give surjective bihomomorphisms
Definition 4.1. A T-algebra
is called a subdirect product of the T-algebras
, if
is a subalgebra of the direct product
and the projection maps
and
are surjective for all
.
Definition 4.2. A T-algebra
is called subdirectly irreducible if
and the (componentwise) intersection of all nonzero congruencies of
is nonzero.
If r is a congruence of an algebra
considered as an Ω-algebra, then
is a congruence of
considered as a T-algebra. Further, if
is a congruence of an algebra
considered as a T-algebra, then
is also a congruence of
considered as a T-algebra, i.e., r is a congruence of an algebra
considered as an Ω-algebra. From these facts we conclude that an algebra
is subdirectly irreducible as an Ω-algebra if and only if it is subdirectly irreducible as a T-algebra.
Theorem 4.1 (see Birkhoff’s subdirect representation theorem [127]). Every T-algebra
can be represented as a subdirect product of subdirectly irreducible T-algebras, which are bihomomorphic images of
. Every finite algebra can be subdirectly embedded into a product of finitely many finite subdirectly irreducible algebras.
A class of T-algebras is said to be multiplicatively closed if it contains the direct product of any family of T-algebras from this class.
Let
,
, be T-algebras of the same arithmetic type, and let
be a filter1 I [6]. on I. We introduce the relation
defined on
and the relation
defined on by setting
where
, and
where ,
.
According to the definition of a filter the relations
and
will be equivalence relations. In addition, it is easy to prove that the pair of equivalence relations
is a congruence of the product of the T-algebras
,
. The corresponding quotient algebra is denoted by
and is called the filtered product (with respect to the filter
) of the T-algebras
,
. A filtered product of T-algebras with respect to an ultrafilter is called an ultraproduct of T-algebras.
Moreover, the congruence
is nuclear. Indeed, for any
we have
As a product of T-algebras the superproduct of two non-zero rings (lattices), where
, is an algebra with four binary operations, and does not a ring (lattice). However it follows from the following Proposition that an ultraproduct of a finite number of finite rings (lattices), as an ultraproduct of T-algebras, is a ring (lattice).
Proposition 4.1. Let
be pairwise nonisomorphic finite T-algebras with finite sets of operations, and let
,
, be a non-empty family of T-algebras, each of which is one of the algebras
. If
is an ultrafilter on I, then there exists a number j,
, such that we have the isomorphism:
Proof. We define the sets
and notice that
are pairwise disjoint and
. Since the filter
is maximal and
, there exists a unique
such that
. Since the congruences
are nuclear, we have
where
is an ultrafilter on
.
For
we define
with the property
for all
. For
we define with the property
for all
. If
and
, then
Hence
and the map
, where
is an equivalence class modulo
, will be an injection. In an analogous way we show that the map
, where
is an equivalence class modulo
, is an injection. Moreover, starting from the equality
, we verify that, for the pair
,
It remains to verify that
and
are surjective.
Let
and
. Then, for each
and
we define:
and
Since the
are pairwise disjoint and
, we conclude that there exists exactly one
such that
. Now
hence,
and
is onto. The similar proof is valid for
. Therefore,
□
A set of T-hyperidentities is said to be satisfiable if there exists a non-trivial T-algebra in which every T-hyperidentity of this set is true. The compactness theorem for first-order languages extends to hyperidentities using ultraproducts of T-algebras.
Theorem 4.2 (The Godel-Mal’tsev type compactness theorem for hyperidentities). If every finite part of an infinite set of T-hyperidentities is satisfiable, then the whole set of T-hyperidentities is also satisfiable.
5. The Hypervariety of T-Algebras
Let
be a system of T-hyperidentities. Denote by
the class of all T-algebras each of which satisfies all T-hyperidentities from
. The class of T-algebras
is called a hypervariety of T-algebras if there exists a system of T-hyperidentities
such that
In this case the class of T-algebras
is called the hypervariety of T-algebras defined by the system of T-hyperidentities
. Similarly, a variety of Ω-algebras is called a hypervariety of Ω-algebras if it is defined by
-hyperidentities [40], where
.
A T-hyperidentity
is called a consequence of
if for every T-algebra
,
(the notation
means satisfiability of all hyperidentities of
in the algebra
). The set
of T-hyperidentities is called a consequence of
and denoted by
if every hyperidentity of
is a consequence of
.
Every T-hyperidentity is a T’-hyperidentity for every set of positive integers
. It is obvious that if
is a consequence of
in the meaning of T-hyperidentities, then
is a consequence of
in the meaning of T’-hyperidentities, for
.
Let us define the concept of formal consequence of
.
(i) Any T-hyperidentity of the form
and any hyperidentity from
are formal consequences of
;
(ii) If T-hyperidentity
is a formal consequence of
, then the hyperidentity
is a formal consequence of
as well;
(iii) If T-hyperidentities
and
are formal consequences of
, then the hyperidentity
is also formal consequence of
;
(iv) If T-hyperidentity
is a formal consequence of
for
, then for any n-ary functional variable Y (where
) the T-hyperidentity
is a formal consequence of
as well;
(v) If T-hyperidentity
is a formal consequence of
, then T-hyperidentity
is also a formal consequence of
, where
and
are obtained from
and
, respectively, by replacement of all occurrences of the functional variable X (the object variable x) by any functional variable Z with the same arity (respectively by any T-hyperterm
).
A T-hyperidentity
is a formal consequence of a system of T-hyperidentities
, if it is a formal consequence of
according to (i)-(v).
Theorem 5.1 (completeness for hyperidentities). The T-hyperidentity
is a consequence of the system of T-hyperidentities
iff it is a formal consequence of
([39] p. 169, [20] p. 43).
Theorem 5.2. A class of T-algebras is a hypervariety of T-algebras if and only if it is hereditary, homomorphicaly and multiplicatively closed.
6. Varieties of Boolean Algebras and Distributive Lattices
Let us start from the definitions of Boolean algebras.
Definition 6.1. An algebra
with two binary, one unary and two nullary operations is called a Boolean algebra if
is a bounded distributive lattice with least element 0 and greatest element 1 and the algebra
satisfies the following identities:
For a lattice
its partial order ≤ is defined in the following way:
For the definition and existence of free algebras
of the given variety V with the set of free generators X also see [8] [10] [11] [36] [72] [128]-[133]. If V is the variety of Boolean algebras then the free algebra
is called the free Boolean algebra with the set of free generators X. For the variety V of distributive lattices (bounded distributive lattices) the free algebra
is called the free distributive lattice (free bounded distributive lattice) with the set of free generators X. The problem of determining the cardinality of free distributive lattices goes back to R.Dedekind [2] (for a modern survey of the field, see [134]).
Definition 6.2. A Boolean algebra (distributive lattice)
is called a free Boolean algebra (free distributive lattice with the system of free generators
if the algebra
is generated by the subset
and for every Boolean algebra (distributive lattice)
and for every mapping
there exists a unique homomorphism:
with
.
Let
. Define the operations
on B in the following way:
,
,
,
,
,
. We get the two-element Boolean algebra
.
The two-element Boolean algebra is the only subdirectly irreducible Boolean algebra up to isomorphism, the two-element lattice is the only subdirectly irreducible distributive lattice up to isomorphism, the two-element semilattice is the only subdirectly irreducible semilattice up to isomorphism [72] [130] [131].
For a set X denote the set of all its subsets by
or
. If we consider subsets of a given set X then for a subset
we denote
.
A function
is called a Boolean function of n variables, where
is the set of all n-element sequences of B. The following result is well known.
Theorem 6.1 ([135]). For every Boolean function
there exists a unique set
such that
where the operations on the right hand side are the operations of Boolean algebra
.
Note that those terms are called disjunctive normal forms for Boolean functions.
For
we define:
iff
for all
. Here and afterwards
is a positive integer.
Definition 6.3. A Boolean function
is called monotone if
where
.
If
then we will say
if there exists k (
) such that
for all
and
,
. A Boolean function
is monotone iff
where
.
Denote the set of all monotone Boolean functions of n variables by
.
We can define
and
for any two Boolean functions of n variables by the standard way. It is obvious that if f and g are monotone Boolean functions, then
and
are monotone, too. Thus, we get the Boolean algebra of Boolean functions of n variables, and the algebra
which obviously is a bounded distributive lattice. Also let
be the number of monotone Boolean functions of n variables. (Note that the numbers
are called Dedekind’s numbers.) For instance,
,
,
,
,
,
([134]).
It is commonly known that the free Boolean algebra on n free generators is isomorphic to the Boolean algebra of Boolean functions of n variables ([2] [11] [130] [136]). The free bounded distributive lattice on n free generators is isomorphic to the bounded lattice of monotone Boolean functions of n variables ([2] [11] [130] [136]). Let us present the last result in detailed.
Now let
be an antichain (or Sperner set [2]) with respect to the order
. It means that S consists of subsets of
, none of which is contained in any other subset from S. Note that the empty set is also considered an antichain. For an antichain
define the following monotone Boolean function:
(6.1)
For
we set
, and for
we set
. Notice that
does not depend on the order of the elements in the set S. It is easy to see that if
are two antichains, then
. To see this without loss of generality suppose that there exists
such that
. We can also suppose that there does not exist
with
. Otherwise, we would take
instead of s (in that case
, because
is an antichain). Take the following values of the variables:
For that values of variables we have:
and
.
The form (6.1) is uniquely determined by the antichain
. And conversely, every monotone Boolean function is obtained in that way.
Proposition 6.1. For every monotone Boolean function f of n variables, there exists a unique antichain
such that
.
Proof. For
let
. Consider the set
. Let S be the subset of A, consisting exactly of all minimal sets in A. Then
is an antichain. Notice that
iff for some
we have
for all
. The same is valid for
. Therefore,
for all
, and so
. The uniqueness follows from the argument stated above. □
Define the Boolean functions:
Theorem 6.2 (Functional representation theorem for free bounded distributive lattices (R.Dedekind [2])). The algebra
is a free bounded distributive lattice with the system of free generators:
. Hence, every n-generated free bounded distributive lattice is isomorphic to the distributive lattice
.
Proof. Let
be any bounded distributive lattice and
be a mapping. We show that there exists a unique homomorphism
such that
. For any
there exists a unique antichain S such that
Take
Obviously
for
and
is a homomorphism. The uniqueness of
is also obvious. □
Corollary 6.1. Every finitely generated free bounded distributive lattice is finite.
Corollary 6.2. Every finitely generated bounded distributive lattice is finite, i.e. the variety of bounded distributive lattices is a Burnside variety.
Theorem 6.3. (Functional representation theorem for free Boolean algebras (R.Dedekind [2])) The algebra
of Boolean function of n variables is a free Boolean algebra with the system of free generators:
. Hence, every n-generated free Boolean algebra is isomorphic to the Boolean algebra
.
Corollary 6.3. Every finitely generated free Boolean algebra is finite.
Corollary 6.4. Every finitely generated Boolean algebra is finite, i.e. the variety of Boolean algebras is a Burnside variety.
7. Bool-De Morgan Algebras
Definition 7.1. An algebra
with two binary, one unary and two nullary operations is called a De Morgan algebra if
is a bounded distributive lattice with least element 0 and greatest element 1 and the algebra
satisfies the following identities:
where
([136]-[148]).
Definition 7.2. An algebra
with two binary, two unary and two nullary operations is called a Boole-De Morgan algebra if
is a De Morgan algebra and
is a Boolean algebra and the two unary operations commute, i.e.,
. So, Boole-De Morgan algebras form a variety.
This concept is introduced in [149] [150] under the name of Boolean bisemigroup. It is easy to note that the last condition
of the definition follows from the other conditions:
and
in the Boolean algebra
.
Let us consider some natural examples of Boole-De Morgan algebras. First note that every Boolean algebra can be considered as a Boole-De Morgan algebra with two equal unary operations. In particular if
is the two-element Boolean algebra and
(i.e., the unary operations
and
are equal) then the algebra
is a Boole-De Morgan algebra and we will denote it by
.
Now we define a Boole-De Morgan algebra on the four-element set
which will be used in the proof of the main theorem of the next Section. Defining
,
and
,
and
,
for all
, and
,
,
,
,
,
,
,
,
,
we get the Boole-De Morgan algebra
.
For a Boolean algebra
consider the direct product
. Defining one more unary operation
on the set
by
we get the Boole-De Morgan algebra
.
The set
of binary term operations of the nontrivial lattice
is a Boole-De Morgan algebra (of order 4) where the operations are defined below. For any two binary terms
and
the binary operations are defined as the following binary superpositions:
The nullary operations are the terms y and x. The unary operations are the commutation and dualization. The commutation is defined by
and for a binary term
to get its dual term
we shall change all variables x by y and vice versa, and also change all operations + by ∙ and vice versa. So we obtain the Boole-De Morgan algebra
:
,
,
,
,
,
,
,
. Note, that every lattice identity of the Boole-De Morgan algebra
is equivalent to certain hyperidentity. For applications of mentioned binary superpositions see [18] [20] [27] [115] [119] [151]-[158].
Let
be a Boolean algebra. Denote its dual Boolean algebra by
, i.e.,
. Consider the direct product
where
,
,
for any
. Defining one more unary operation
on the set
by
we get the Boole-De Morgan algebra
. Every Boole-De Morgan algebra
is an embedding into
. We define an embedding of the Boole-De Morgan algebra
into Boole-De Morgan algebra
by the following rule:
Now we formulate a Stone-type representation theorem for Boole-De Morgan algebras.
Let
be the Boolean algebra of all subsets of the set I.
Theorem 7.1 ([149] [150]). Every Boole-De Morgan algebra
is isomorphic to a subalgebra of the Boole-De Morgan algebra
for some Boolean algebra
.
For a Boole-De Morgan algebra
we define one more unary operation
by the following way: . It is easy to see that
,
, ,
. Thus the mapping
is an automorphism of the Boole-De Morgan algebra
. Also it is easy to see that
if and only if
.
8. Quasi-De Morgan Functions
Recall that
and
. Let us construct a one-to-one correspondence between the sets D and
as follows:
We define the operations
on the set
as follows:
(here the operations on the right hand side are the operations of the Boole-De Morgan algebra
). We get the Boole-De Morgan algebra
, which is isomorphic to the algebra
(the one-to-one correspondence described above is an isomorphism). However, if the ordered pair
corresponds to
then we will write
(this causes no confusion).
For
let
The unary operation
can also be defined on
taking into account the isomorphism described above. In result we get
,
. It is clear that (which agrees with the notation from the previous section).
Now we can introduce the following two concepts of quasi-De Morgan and De Morgan functions.
Definition 8.1. A function
is called a quasi-De Morgan function if the following conditions hold:
(1) if
, then
,
(2) if
, then
,
In terms of clone theory, Condition (1) means that the function f preserves the unary relation
, and Condition (2) means that f preserves the binary relation
, which is the graph of the automorphism
.
For
we say that d is a permitted modification of c if for some
we have
for all
,
, and
Definition 8.2. A quasi-De Morgan function
is called a De Morgan function if it satisfies the following condition:
(3) if
with
and y is a permitted modification of x then
.
In terms of clone theory the Condition (3) means that f preserves the order relation
.
Notice that Condition (1) is a consequence of Condition (2), however it is more convenient to write it as a separate condition.
Note that it follows from Condition (1) that every quasi-De Morgan function is an extension of some Boolean function. Notice that the constant functions
and
are quasi-De Morgan functions, but the constant functions
and
are not. This means that 0 and 1 are the only constant quasi-De Morgan functions. Further examples of quasi-De Morgan functions are
,
,
,
,
, where the operations on the right hand side are the operations of the Boole-De Morgan algebra
. Also note that the function p is an example of quasi-De Morgan function which is not a De Morgan function.
Below, for
we denote by
the pair from
which corresponds to
, i.e.,
.
Theorem 8.1. A function
is a quasi-De Morgan function if and only if there exists a Boolean function
such that
(8.1)
for all
.
For a quasi-De Morgan function
there exists a unique Boolean function
which satisfies (8.1). To emphasize that
is the unique Boolean function corresponding to f, we denote it by
.
Corollary 8.1. There are exactly
quasi-De Morgan functions of n variables.
Denote the set of all quasi-De Morgan functions of n variables by
. For the functions
define
,
,
and
by the standard way, i.e.,
,
,
,
,
, where the operations on the right hand side are the operations of the Boole-De Morgan algebra
.
Theorem 8.2. The set
is closed under operations
, i.e., if
, then
.
Thus, we get an algebra:
(here 0 and 1 are the constant quasi-De Morgan functions), which obviously is a Boole-De Morgan algebra.
For a set
define the function
by the following way:
(8.2)
where the operations on the right hand side are the operations of
(cf. [120] [121]).
Notice that
does not depend on the order of the elements in the set S.
Also we set
and
.
Let us consider the projection functions
as functions
. Obviously,
is a quasi-De Morgan function for each i. According to (8.2), for any set
we have:
Hence,
, i.e.,
is a quasi-De Morgan function for any set
.
For
let
, and for
let
. In this way we give a one-to-one correspondence between the sets
and
.
Now, for any quasi-De Morgan function
from Theorem 1 we conclude that there exists a set
such that:
where S is the subset of
corresponding to
.
Moreover, the number of all quasi-De Morgan functions of n variables is the same as the number of all subsets of
. Hence, we get the following result.
Theorem 8.3. For any quasi-De Morgan function f of n variables there exists a unique set
such that
.
In particular,
if
.
Thus every quasi-De Morgan function can be uniquely presented in the form (8.2). This form is called the disjunctive normal form (or briefly—DNF) of quasi-De Morgan function f. Notice that from Theorems 1 and 3 we get an algorithm which, given a quasi-De Morgan function, gives its disjunctive normal form. We can also prove that every quasi-De Morgan function can be uniquely presented in conjunctive normal form (CNF), i.e., in the following form:
Below we will use the concept of essential variable (and essential dependence) of quasi-De Morgan functions. The definitions are the same as in case of Boolean functions and so we do not give them here.
Using arguments similar to those given above we can prove the following theorem.
Theorem 8.4. For a quasi-De Morgan function f the corresponding Boolean function
does not essentially depend on the last n variables if and only if f can be represented as a term function with functional symbols
, i.e., f is a term function of the Boolean algebra
.
Theorem 8.5. (Functional representation theorem for free Boole-De Morgan algebras). The algebra
is a free Boole-De Morgan algebra with the system of free generators
. Hence every free n-generated Boole-De Morgan algebra is isomorphic to the Boole-De Morgan algebra
.
Corollary 8.2. Every finitely generated free Boole-De Morgan algebra is finite.
Corollary 8.3. Every finitely generated Boole-De Morgan algebra is finite, i.e. the variety of Bool-De Morgan algebras is a Burnside variety.
Problem 8.1. Prove the Post style functional completeness theorem for quasi-De Morgan functions (see [63] [159]-[161]).
Now let us formulate the following characterization of quasi-De Morgan functions.
Theorem 8.6. Quasi-De Morgan functions are precisely the term functions of the Boole-De Morgan algebra
, i.e., the set of all quasi-De Morgan functions is the clone of term functions of the Boole-De Morgan algebra
.
9. De Morgan Algebras
Apart from mathematical logic ([65] [144] [162]-[169]) and algebra, De Morgan algebras (and De Morgan bisemilattices) have applications in multi-valued simulations of digital circuits too ([139] [140]).
For example, the standard fuzzy algebra
is a De Morgan algebra.
Note that in any De Morgan algebra
:
,
and:
A characterization of De Morgan algebras can be found in [138] [140].
Let
be a bounded distributive lattice. Denote its dual bounded distributive lattice by
, i.e.,
. Consider the direct product
where
,
, for any
. Defining one more operation
on the set
by
, we convert the bounded distributive lattice
into the De Morgan algebra
. Every De Morgan algebra
is an embedding into
. We define an embedding of the De Morgan algebra
into De Morgan algebra
by the following rule:
Now we formulate a Stone-type representation theorem for De Morgan algebras.
Let
be the bounded distributive lattice of all subsets of set I.
Theorem 9.1. ([149] [150]). Every De Morgan algebra
is isomorphic to a subalgebra of the De Morgan algebra
for some bounded distributive lattice
.
Let us consider the following De Morgan algebras:
,
, where
, and
, where
,
,
,
.
Let us remind the following result.
Theorem 9.2. ([143]). Every non-trivial subdirectly irreducible De Morgan algebra is isomorphic to one of the following algebras: 2, 3, 4, where 2 is the unique non-trivial subdirectly irreducible Boolean algebra.
10. Free De Morgan Algebras and De Morgan Functions
If V is the variety of De Morgan algebras then the free algebra
is called free De Morgan algebra with the set of free generators X. Thus, De Morgan algebra
is called a free De Morgan algebra with the system of free generators
if the algebra
is generated by the subset
and for every De Morgan algebra
and for every mapping
there exists a unique homomorphism:
with
.
Denote
, where
,
,
,
. Defining
,
,
,
,
,
,
,
,
we get the De Morgan algebra
. Notice that
,
and
(here the operations on the right hand side are the operations of the De Morgan algebra 2). For
let
Also for
we say that d is a permitted modification of c if for some
we have
for all
,
and
Definition 10.1. A function
is called a De Morgan function of n variables if the following conditions hold:
(1) if , then
,
(2) if then
,
(3) if
with
and y is a permitted modification of x then
.
Note that it follows from the first condition that every De Morgan function is an extension of some Boolean function. And notice that the constant functions
and
are De Morgan functions, but the constant functions
and
are not. This means that 0 and 1 are only constant De Morgan functions. Other examples of De Morgan functions are
,
,
,
, where the operations on the right hand side are the operations of the De Morgan algebra 4.
As Boolean functions, De Morgan functions can be given by tables. Also note that there exists an algorithm which for a given table of a function
determines whether f is a De Morgan function.
Below, for
by
we mean the couple from
which is equal to
. But often we will consider
as a subset of D (when it can cause no confusion).
The equivalency of the following two definitions follows from the Theorem 1.
For any quasi-De Morgan function
there exists a unique Boolean function
which satisfies (8.1) and denoted by
(see previous Chapter).
Theorem 10.1. The function
is a De Morgan function iff it is a quasi-De Morgan function and
is a monotone Boolean function.
Corollary 10.1. There are
De Morgan functions of n variables, where
is the number of monotone Boolean functions of k variables.
Denote the set of all of n-variable De Morgan functions by
. For the functions
define
,
and
by the standard way, i.e.
,
,
,
, where the operations on the right hand side are the operations of De Morgan algebra 4. Notice that
is closed under those operations, i.e. if
, then
. We can verify it straightforwardly, using the definition of De Morgan function.
Thus, we get the algebra:
, which obviously is a De Morgan algebra.
Let
. We say that
, if
and
. In this way, we get a partially ordered set
. For the antichain,
, define the function,
, by the following way:
(10.1)
Notice that
does not depend on the order of the elements in the set S ([120]).
Note that we set
and
.
Let us consider the functions
as functions
. Obviously,
is a De Morgan function. And according to (10.1), for any antichain
we have:
Hence,
, i.e.
is a De Morgan function for any antichain
.
For
let
, and for
let
. In this way we give a bijective mapping from the set of all antichains of
to the set of all antichains of
. And so the number of all antichains of the partially ordered set
is
.
Now, for any De Morgan function
from Proposition 1 we conclude that there exists an antichain
such that:

where S is the antichain of
, corresponding to
.
Moreover, the number of all De Morgan functions of n variables is the same as the number of all antichains of
. Hence, we get the following result.
Theorem 10.2. For any De Morgan function f of n variables there exists a unique antichain
such that
.
In particular,
if
.
Thus every nonconstant De Morgan function can be uniquely presented in the form (10.1). This form is called the canonical form (or disjunctive normal form (or briefly—DNF)) of De Morgan function f. Notice that from the proofs of Theorem 2 and Proposition 1 we get an algorithm which, for a De Morgan function, gives its disjunctive normal form. From this we conclude that every nonconstant De Morgan function can be uniquely presented in conjunctive normal form (CNF), i.e. in the following form:
For a De Morgan function CNF is unique (we can prove this analogously to DNF).
Theorem 10.3 (Functional representation theorem for free De Morgan algebras). The algebra
is a free De Morgan algebra with the system of free generators:
. Hence, every free n-generated De Morgan algebra is isomorphic to the De Morgan algebra
.
Corollary 10.2. Every finitely generated free De Morgan algebra is finite.
Corollary 10.3. Every finitely generated De Morgan algebra is finite, i.e. the variety of De Morgan algebras is a Burnside variety.
A similar result is valid for the finitely generated free algebras with two binary operations of the hypervariety defined by the system of hyperidentities of the variety of Boolean algebras ( distributive lattices, De Morgan algebras) ([120]-[122]).
Problem 10.1. Develop the De Morgan analogue of the theory of Boolean functions.
The idea of De Morgan functions is also applicable in quantum computation and the theory of quantum computers.
Acknowledgements
The author was partially supported by the State Committee of Science of the Republic of Armenia, grants: 10-3/1-41, 21T-1A213.
NOTES
1A filter on a non-empty set I is a non-empty set
of subsets of I satisfying the requirements: a) the intersection of any two subsets from
belongs to
; b) all the supersets of any subset belonging to
also belong to
; c) the empty set
does not belong to
. A maximal filter on I, that is, a filter on I that does not lie in any other filter on I, is usually called an ultrafilter on I.