Effect algebras, which were introduced by Foulis and Bennett [8], are partially ordered partial algebras closely related to the logical foundations of quantum mechanics. The standard example is the structure of self-adjoint operators between zero and identity on a Hilbert space, the so-called effects. Also some commonly known (total) algebras, such as orthomodular lattices and MV-algebras, may be regarded as particular cases of effect algebras. We were not the first who observed that lattice-ordered effect algebras can in a natural way be made into total algebras 〈A;⊕, ,0,1〉 of type 〈2,1,0,0〉; see [2, 3]. We proved that the class \(\mathcal {E}\) of such algebras is a variety containing both the variety \(\mathcal {M}\mathcal {V}\) of MV-algebras and the variety \(\mathcal {OM}\) (term equivalent to the variety) of orthomodular lattices; relative to \(\mathcal {E}\), \(\mathcal {M}\mathcal {V}\) and \(\mathcal {OM}\) can be axiomatized by the identities xy=yx and xx=x, respectively. We also observed that \(\mathcal {E}\) is not the join of \(\mathcal {M}\mathcal {V}\) and \(\mathcal {OM}\) in the lattice of subvarieties of \(\mathcal {E}\). The aim of the present paper is to axiomatize the join of the varieties \(\mathcal {M}\mathcal {V}\) and \(\mathcal {OM}\), but we actually axiomatize the join of \(\mathcal {M}\mathcal {V}\) with any finitely based subvariety of \(\mathcal {E}\).

First, we recall some basic facts about lattice effect algebras; for more information we refer the reader to the book [7]. An effect algebra is a structure 〈E;+,0,1〉, where + is a partial binary operation and 0,1 are two constants, satisfying the following conditions:

  1. (i)

    x+y=y+x if one side is defined;

  2. (ii)

    (x+y)+z=x+(y+z) if one side is defined;

  3. (iii)

    for every xE there exists a unique x E such that x +x=1;

  4. (iv)

    if x+1 is defined, then x=0.

This is the definition by Foulis and Bennett [8], but we should mention that effect algebras are essentially the same as weak orthoalgebras introduced by Giuntini and Greuling [9] who, however, attribute their introduction to Foulis and Randall. Besides, effect algebras are equivalent to D-posets which were introduced by Kôpka and Chovanec [14].

Every effect algebra is naturally ordered by stipulating that xy iff y=x+z for some zE; such an element z is unique and we may denote it by yx. Thus, if xy, then yx is the only element such that y=x+(yx), or more directly, yx=(x+y ). The structure 〈E;≤,−,0,1〉 so obtained is a D-poset in the sense of [14].

It is worth noticing that x+y is defined iff xy , in which case x+y=(y x).

An effect algebra (or D-poset) which is a lattice with respect to the natural order ≤ is called a lattice effect algebra (or D-lattice).

Two examples of lattice effect algebras that are of particular interest to us are orthomodular lattices and MV-algebras. We refer the reader to [13] for orthomodular lattices, and to [6] for MV-algebras.

Orthomodular lattices are equivalent to lattice effect algebras satisfying the condition that x+x is defined only if x=0. Indeed, if 〈L;∨,∧, ,0,1〉 is an orthomodular lattice, then the structure 〈L;+,0,1〉, where x+y is defined and equals xy iff xy , is a lattice effect algebra satisfying the condition; and conversely, if 〈E;+,0,1〉 is such a lattice effect algebra with induced lattice operations ∨ and ∧, then x+y=xy if x+y is defined, and 〈E;∨,∧, ,0,1〉 is an orthomodular lattice.

MV-algebras are equivalent to MV-effect algebras, i.e., lattice effect algebras satisfying (xy)−y=x−(xy) for all x,y. Though we do not want to go into details here, we need to mention that every MV-algebra 〈A;⊕, ,0,1〉 bears a lattice order which is given by xy iff x y=1 iff y=xz for some zA. Now, if 〈A;⊕, ,0,1〉 is an MV-algebra, then 〈A;+,0,1〉, where x+y is defined and equals xy iff xy , is an MV-effect algebra; and conversely, if 〈E;+,0,1〉 is an MV-effect algebra, then by setting

$$ x\oplus y=(x\wedge y^{\prime})+y $$
(1)

we obtain an MV-algebra 〈E;⊕, ,0,1〉 with the same lattice order as 〈E;+,0,1〉.

Using (1), an arbitrary lattice effect algebra can be made into a total algebra. So let \(\mathcal {E}\) be the class of all (total) algebras A=〈A;⊕, ,0,1〉 of type 〈2,1,0,0〉 that arise from lattice effect algebras by means of (1). It is useful to define two more total operations, as follows:

$$ x\oslash y=(x^{\prime}\oplus y)^{\prime} \quad\text{and}\quad x\ominus y=(y\oplus x^{\prime})^{\prime}. $$
(2)

In the language of lattice effect algebras we have

$$ x\oslash y=(x\vee y)-y \quad\text{and}\quad x\ominus y=x-(x\wedge y), $$
(3)

and it is straightforward to show that all algebras in \(\mathcal {E}\) satisfy the identities

$$\begin{array}{@{}rcl@{}} &&x\oplus 0 = x, \end{array} $$
(4)
$$\begin{array}{@{}rcl@{}} &&x^{\prime\prime} = x, \end{array} $$
(5)
$$\begin{array}{@{}rcl@{}} &&(x\oslash y)\oplus y = (y\oslash x)\oplus x, \end{array} $$
(6)
$$\begin{array}{@{}rcl@{}} &&(((x\oplus y)\oslash y)\oplus z)^{\prime}\oplus (x\oplus z) = 1, \end{array} $$
(7)

and the quasi-identity

$$ x\oplus y\leq z^{\prime} \quad \Rightarrow\quad (x\oplus y)\oplus z = x\oplus (z\oplus y). $$
(8)

On the other hand, we proved in [3] that if an algebra A=〈A;⊕, ,0,1〉 satisfies the identities (4)–(7), then the rule xy iff x y=1 (equivalently, xy=0 or xy=0) defines a bounded lattice, with bounds 0 and 1, where

$$x\vee y=(x\oslash y)\oplus y \quad\text{and}\quad x\wedge y=(x^{\prime}\vee y^{\prime})^{\prime}=x\ominus (x\ominus y). $$

In [3], as well as in some other papers, we called these algebras “basic algebras”. Moreover, if A satisfies also (8), then the partial algebra 〈A;+,0,1〉 obtained by restricting ⊕ as in the case of MV-algebras, i.e., x+y is defined and equals xy iff xy , is a lattice effect algebra with the same natural order as A, which entails that the total algebra associated with 〈A;+,0,1〉 via (1) is the initial algebra A.

The quasi-identity (8) may be replaced with an identity; for example, it suffices to write (xy)z instead of z. Hence the class \(\mathcal {E}\) is a variety.

Recalling that MV-algebras correspond to MV-effect algebras, it is evident by (3) that the variety of MV-algebras \(\mathcal {M}\mathcal {V}\) is the subvariety of \(\mathcal {E}\) axiomatized by the identity xy=xy, which is in view of (2) equivalent to xy=yx. MV-algebras are usually defined as algebras A=〈A;⊕, ,0,1〉 such that 〈A;⊕,0〉 is a commutative monoid, satisfying the identities (5), (6) and 1⊕x=1=0; see [6].

We have seen that orthomodular lattices correspond to lattice effect algebras satisfying the condition that xx implies x=0. In the language of \(\mathcal {E}\), this is equivalent to the identity xx =0, and in turn to xx=x. We let \(\mathcal {OM}\) denote the subvariety of \(\mathcal {E}\) defined by xx =0 or xx=x. Note that the total addition in orthomodular lattices is given by xy=(xy )∨y, which is not the same as xy.

The variety \(\mathcal {E}\) is congruence distributive, because its members are lattice based algebras. In fact, \(\mathcal {E}\) is an arithmetical variety; see [3]. Congruence distributivity implies that for any two subvarieties \(\mathcal {V}_{1},\mathcal {V}_{2}\) of \(\mathcal {E}\) one has \(\text {Si}({\mathcal {V}_{1} \vee \mathcal {V}_{2}}) = \text {Si}({\mathcal {V}_{1}}) \cup \text {Si}({\mathcal {V}_{2}})\), where \(\text {Si}(\mathcal {K})\) denotes the class of subdirectly irreducible algebras in the respective class \(\mathcal {K}\). Since there exist subdirectly irreducible algebras in \(\mathcal {E}\) which are neither in \(\mathcal {M}\mathcal {V}\) nor in \(\mathcal {OM}\), it follows that \(\mathcal {E}\) is not the join \(\mathcal {M}\mathcal {V}\vee \mathcal {OM}\). For instance, a non-trivial horizontal sum of non-Boolean MV-algebras is a simple algebra in \(\mathcal {E}\), but it is neither an MV-algebra nor an orthomodular lattice.

The concept that plays a central role in our proof is compatibility (see [7], Sect. 1.10). In general, two elements x,y in an effect algebra are said to be compatible, in symbols xy, if there exist x 1,y 1,z such that x=x 1+z, y=y 1+z and x 1+z+y 1 is defined. It is obvious that 0⇔x⇔1 and xx for any x. Also, if xy or yx, then xy. Another important fact that we will use repeatedly is that xy iff xy . In lattice effect algebras we have

$$x\leftrightarrow y \quad\text{iff}\quad (x\vee y)-y=x-(x\wedge y) \quad\text{iff}\quad (x\vee y)-y\leq x \quad\text{iff}\quad x-(x\wedge y)\leq y^{\prime}, $$

and hence, in the language of the variety \(\mathcal {E}\),

$$x\leftrightarrow y \quad\text{iff}\quad x\oslash y=x\ominus y \quad\text{iff}\quad x\oslash y\leq x \quad\text{iff}\quad x\ominus y\leq y^{\prime}. $$

Since xy iff xy , it is also true that

$$x\leftrightarrow y \quad\text{iff}\quad x\oplus y=y\oplus x \quad\text{iff}\quad x\leq x\oplus y. $$

Consequently, since MV-effect algebras are lattice effect algebras where xy for all x,y, the variety \(\mathcal {M}\mathcal {V}\) may be axiomatized, relative to \(\mathcal {E}\), by any of the identities xy=xy, xyx, xyy , xy=yx or xxy.

A block of an effect algebra is a maximal set of mutually compatible elements. For lattice effect algebras Riečanová [17] proved that if x,y are compatible with a given z, then so are xy, xy, x+y (when defined) and xy (when defined). It follows that every block B of any algebra \(\boldsymbol {A}\in \mathcal {E}\) is a subuniverse of A, and obviously, the subalgebra B is an MV-algebra. The intersection of the blocks of A is called the compatibility centre and we denote it by K(A). Clearly, K(A)={aAax for all xA} and K(A) is a subuniverse of A.

Though the addition ⊕ is neither commutative nor associative,Footnote 1 we may unambiguously write nx=x⊕⋯⊕x (with n occurrences of x) for any positive integer \(n\in \mathbb {N}\), because every element x belongs to a block which is in fact an MV-algebra. We will mostly write just nx instead of nx.

By [3], the variety \(\mathcal {E}\) is congruence regular, i.e., any congruence 𝜃 of an algebra \(\boldsymbol {A}\in \mathcal {E}\) is determined by each of its classes [a] 𝜃 , and in particular, by [0] 𝜃 . We say that IA is an ideal Footnote 2 of A if I=[0] 𝜃 for some congruence 𝜃 of A. Pulmannová and Vinceková [16] proved that IA is an ideal if and only if

  1. (i)

    xyI for all x,yI;

  2. (ii)

    xyI for all xI and yA.

An alternative characterization of ideals can be found in [5]. Of course, if \(\boldsymbol {A} \in \mathcal {M}\mathcal {V}\), then the condition (ii) amounts to saying that I is downwards closed.

For any ideal I of an algebra \(\boldsymbol {A}\in \mathcal {E}\), the only congruence 𝜃(I) with the property that [0] 𝜃(I)=I is given by

$$\langle x,y\rangle\in\theta(I) \quad\text{iff}\quad x\oslash y,y\oslash x\in I \quad\text{iff}\quad x\ominus y,y\ominus x\in I. $$

When ordered by set inclusion, the ideals of \(\boldsymbol {A}\in \mathcal {E}\) form a distributive lattice that is isomorphic to the congruence lattice of A under the mutually inverse assignments I𝜃(I) and 𝜃↦[0] 𝜃 .

In general, we do not have a reasonable description of the ideal Ig(X) generated by a given subset XA, but if X is an ideal of the compatibility centre or X={a} where aK(A), then Ig(X) can be described easily:

FormalPara Lemma 1 (cf. [12], Prop. 1)

Let \(\boldsymbol {A}\in \mathcal {E}\) . If J is an ideal of the MV-algebra K(A), then Ig(J)={x∈A∣x≤a for some a∈J}. In particular, for any a∈K(A), \(\text {Ig}({a})=\{x\in A\mid x\leq na \text { for some}\,\, n\in \mathbb {N}\}\).

FormalPara Proof

The first part is but a translation of [12], Prop. 1, into the language of the variety \(\mathcal {E}\). Since K(A) is an MV-algebra, it is obvious that for any aK(A), the ideal of K(A) generated by a is \(\{x\in K(\boldsymbol {A})\mid x\leq na \text { for some}\,\, n\in \mathbb {N}\}\), and consequently, the ideal of A generated by a is \(\{x\in A\mid x\leq na \text { for some}\,\, n\in \mathbb {N}\}\). □

We should notice that in any algebra \(\boldsymbol {A}\in \mathcal {E}\) we have

$$ x\leq y \quad\Rightarrow\quad x\oplus z\leq y\oplus z,\quad x\oslash z\leq y\oslash z \quad\text{and}\quad z\ominus y\leq z\ominus x, $$
(9)

which follows directly from (1) and (3). In particular, we have yxy, xyy and xyx. We will also need the identity

$$ (x\wedge y)\oplus z=(x\oplus z)\wedge (y\oplus z), $$
(10)

which is over \(\mathcal {E}\) equivalent to the identities (xy)⊘z=(xz)∨(yz) and x⊖(yz)=(xy)∨(xz). It is not hard to show (cf. [7], Prop. 1.8.6) that any lattice effect algebra satisfies the equality (xy)+z=(x+z)∧(y+z) provided x+z,y+z are defined, whence it follows that the identity (10) holds in every algebra in \(\mathcal {E}\). In fact, (9) and (10) hold in all “basic algebras” (i.e., algebras satisfying (4)–(7), see [15]).

FormalPara Lemma 2

Let \(\boldsymbol {A}\in \mathcal {E}\) . If a 1 ,…,a n ,b∈A are such that a i ⇔b and a i ∧b=0 for each i=1,…,n, then s(a 1 ,…,a n )⇔b and s(a 1 ,…,a n )∧b=0 for every additive term Footnote 3 s(x 1 ,…,x n ).

FormalPara Proof

We know that if a i b for every i=1,…,n, then also any “sum” of the a i ’s is compatible with b. The rest is an easy induction: Suppose that s=pq where p,q are additive terms for which the statement holds true. Let c=p(a 1,…,a n ) and d=q(a 1,…,a n ). Then bc=bd=0 and, since bd and using (10), we have bs(a 1,…,a n )=b∧(cd)=b∧(bd)∧(cd)=b∧((bc)⊕d)=bd=0. □

FormalPara Lemma 3

Let \(\boldsymbol {A}\in \mathcal {E}\) and a∈A. Suppose that (x⊘(x⊕y))∧a=0 for all x,y∈A. Then a∈K(A) and the polar a ={x∈A∣x∧a=0} is an ideal of A such that Ig(a)∩a ={0}.

FormalPara Proof

We have xy=yx whenever xy . Hence, by (8), x ya⊕(x y)=(a ⊘(x y)) implies

$$\begin{array}{@{}rcl@{}} ((x\oslash y)\wedge a)^{\prime} = (x^{\prime}\oplus y)\vee a^{\prime} = [a^{\prime}\oslash (x^{\prime}\oplus y)]\oplus (x^{\prime}\oplus y) \\ =(x^{\prime}\oplus y)\oplus [a^{\prime}\oslash (x^{\prime}\oplus y)] = x^{\prime}\oplus ([a^{\prime}\oslash (x^{\prime}\oplus y)]\oplus y), \end{array} $$

whence ((xy)∧a)⊖x=x ⊘((xy)∧a)=x ⊘(x z) where z=(a ⊘(x y))⊕y. But then ((xy)∧a)⊖xa yields ((xy)∧a)⊖x=(x ⊘(x z))∧a=0 by the assumption about a. Thus (xy)∧ax for all x,yA.

Then xa =(xa )∧ax for any xA, because we have xa ≤1⊘a =a. This shows that a K(A), whence also aK(A). Now, (i) if x,ya , then xya by Lemma 2; and (ii) if xa , then (xy)∧axa=0, so xya . Hence a is an ideal of A.

Finally, if x∈Ig(a)∩a , then xn a for some \(n\in \mathbb {N}\) by Lemma 1, and xa=0, which entails xn a=0 by Lemma 2. Thus x=xn a=0. □

Now, let \(\mathcal {K}\) be a finitely based subvariety of \(\mathcal {E}\) incomparable with \(\mathcal {M}\mathcal {V}\), and suppose that \(\mathcal {K}\) is axiomatized, relative to \(\mathcal {E}\), by an identity which is of the form

$$t(x_{1},\dots,x_{k}) = 0. $$

Since \(\mathcal {K}\) is finitely based, it can always be axiomatized by a single identity of such a form. Indeed, any identity p=q can be replaced with pq=0 and qp=0 where ◇ is ⊖ or ⊘, and any finite family of identities t 1=0,…,t n =0 can be replaced with the identity t 1∨⋯∨t n =0.

FormalPara Theorem 4

If \(\mathcal {K}\) is as above, then the join of \(\mathcal {M}\mathcal {V}\) and \(\mathcal {K}\) is axiomatized, relative to \(\mathcal {E}\) , by the identity

$$ (x\oslash (x\oplus y))\wedge t(z_{1},\dots,z_{k}) = 0. $$
(11)
FormalPara Proof

Let \(\mathcal {K}^{\prime }\) be the subvariety of \(\mathcal {E}\) axiomatized by (11). Since x⊘(xy)=0 holds in \(\mathcal {M}\mathcal {V}\) and t(z 1,…,z k )=0 in \(\mathcal {K}\), it is obvious that (11) holds in \(\mathcal {M}\mathcal {V}\) as well as in \(\mathcal {K}\). Hence \(\mathcal {M}\mathcal {V}\vee \mathcal {K}\subseteq \mathcal {K}^{\prime }\).

Conversely, let \(\boldsymbol {A}\in \text {Si}({\mathcal {K}^{\prime }})\) and suppose that \(\boldsymbol {A}\notin \mathcal {K}\), i.e., there exist a 1,…,a k A such that t(a 1,…,a k )≠0. Let b=t(a 1,…,a k ). Then (x⊘(xy))∧b=0 for any x,yA, and so, by Lemma 3, b is an ideal such that Ig(b)∩b ={0}. Since A is a subdirectly irreducible algebra and Ig(b)≠{0}, we have b ={0}. But x⊘(xy)∈b for all x,yA, hence A satisfies the identity x⊘(xy)=0, i.e. xxy. In other words, \(\boldsymbol {A}\in \mathcal {M}\mathcal {V}\). Therefore, \(\text {Si}({\mathcal {K}^{\prime }}) \subseteq \text {Si}({\mathcal {M}\mathcal {V}}) \cup \text {Si}({\mathcal {K}})=\text {Si}({\mathcal {M}\mathcal {V}\vee \mathcal {K}})\), whence \(\mathcal {K}^{\prime }\subseteq \mathcal {M}\mathcal {V}\vee \mathcal {K}\). □

Let us return to orthomodular lattices. The variety \(\mathcal {OM}\) is axiomatized, relative to \(\mathcal {E}\), by the identity xx =0, i.e., the term t(x 1,…,x k ) from the above theorem is t(x)=xx , and hence by Theorem 4 we obtain:

FormalPara Corollary 5

The join of \(\mathcal {M}\mathcal {V}\) and \(\mathcal {OM}\) is axiomatized, relative to \(\mathcal {E}\) , by the identity

$$(x\oslash (x\oplus y))\wedge z\wedge z^{\prime} = 0. $$

Let 〈D;+,0,1〉 be the so-called distributive diamond, i.e., the lattice effect algebra with universe D={0,a,b,1} such that a+a=1=b+b, while a+b and b+a are not defined. This effect algebra has a prominent role for distributive lattice effect algebras because by [10], every finite distributive lattice effect algebra is isomorphic to a direct product of chains and diamonds. Let D be the corresponding algebra in \(\mathcal {E}\). Note that ab=b and ba=a. Relative to the variety \(\mathcal {DE}\) of distributive members of \(\mathcal {E}\), the variety V(D) generated by D was axiomatized by the identity (xy)⊖2z=(x⊖2z)⊖(y⊖2z) in [11], and by 2x=3x in [4]. Using Theorem 4 we can simplify the axiomatization of the join of \(\mathcal {M}\mathcal {V}\) and V(D) which was given in [1]. We may take the term t(x)=3x⊘2x.

FormalPara Corollary 6

The join of \(\mathcal {M}\mathcal {V}\) and V( D) is axiomatized, relative to \(\mathcal {DE}\) , by the identity

$$(x\oslash (x\oplus y))\wedge (3z\oslash 2z) = 0. $$