1 Introduction

In 1968, the theory of relative annihilators was introduced in lattices by Mark Mandelker [9] and he characterized distributive lattices in terms of their relative annihilators. Later, many authors introduced the concept of annihilators in the structures of rings as well as lattices and characterized several algebraic structures in terms of annihilators. Speed [8] and Cornish [4, 5] made an extensive study of annihilators in distributive lattices. The main aim of this paper is to study some further properties of annihilators in the form of disjunctive ideals of distributive lattices.

In this note, the concepts of disjunctive ideals and strongly disjunctive ideals are introduced in terms of annihilators. Normal lattices are once again characterized in terms of annihilators. A set of equivalent conditions are derived for every ideal of a lattice to become a disjunctive ideal. The notion of normal prime ideals is introduced and proved that every normal prime ideal is a disjunctive ideal as well as a minimal prime ideal. Some properties of disjunctive ideals are observed with respect to inverse homomorphic images and cartesian products. The notion of weakly normal lattices is introduced and established a set of equivalent conditions for every weakly normal lattice to become a normal lattice. A set of equivalent conditions is derived for every ideal of a lattice to become a strongly disjunctive ideal. Finally, a set of equivalent conditions is deduced for the class of all strongly disjunctive ideals of a distributive lattice to become a sublattice of the ideal lattice.

2 Preliminaries

The reader is referred to [2] and [3] for the elementary notions and notations of distributive lattices. However, some of the preliminary definitions and results are presented for the ready reference of the reader.

For any element a of a distributive lattice (L, ∨, ∧, 0) with zero, we define the annihilator of a by the set (a)={xL | xa = 0} and the principle ideal generated by a is written as (a]. The set \(\mathcal {I}(L)\) of all ideals of (L, ∨,∧,0) forms a complete distributive lattice. A proper ideal P of a lattice L is called prime if for any x, yL, xyP implies xP or yP. A prime ideal is called minimal if it is the minimal element in the partial ordered set of all prime ideals. A prime ideal P is minimal if and only if to each xP there exists yP such that xy = 0. A proper ideal M of a lattice is called maximal if there exists no proper ideal N such that MN. The properties of minimal prime ideals are exclusively studied by J. Kist [7] in commutative semi-groups and later these ideals are used by Cornish [4].

A distributive lattice with 0 is called normal [4] if every prime ideal contains a unique minimal prime ideal. A distributive lattice L with 0 is normal if and only if for all x, yL, (x) ∨ (y) = L whenever xy = 0. Normal lattices were extensively studied by W.H. Cornish [4]. In his study, he characterized normal lattices with the help of annihilators and minimal prime ideals. Some examples of normal lattices are as follows:(1) the lattice of all zero sets in a completely regular space, and (2) the generalized stone lattices. Example (1) is considered in [1, pp. 107, 108] and quoted by Cornish in [4, pp. 214]. Example (2) is considered by Katrinak [6] and it is proved in [5, Proposition 5.4]. Throughout this note, L stands for a bounded distributive lattice with 0 and 1, unless otherwise mentioned.

3 Disjunctive ideals

In this section, the concept of disjunctive ideals is introduced in lattices. Normal lattices are characterized. A set of equivalent conditions is derived for every ideal of a lattice to become a disjunctive ideal.

Definition 3.1

For any non-empty subset A of L, define

$$A^{\circ } = \{x\in L~|~(a)^{*}\vee (x)^{*} = L\text{ for all }a\in A\}. $$

Clearly {0} = L and L ={0}. For any aL, we denote ({a}) by (a). Then it is obvious that (0) = L and (1)={0}.

Proposition 3.2

For any non-empty subset A of L, A is an ideal in L.

Proof

Clearly 0 ∈ A . Let x, yA . For any aA, we get (xy) ∨ (a)={(x)∩(y)}∨(a)={(x) ∨ (a)}∩{(y) ∨ (a)}=LL = L. Hence xyA . Again, let xA and yx. Then we get (x) ∨ (a) = L for any aA and (x)⊆(y). For any cA, we then get L = (x) ∨ (a)⊆(y) ∨ (a). Hence yA . Therefore A is an ideal in L. □

The following lemma is a direct consequence of the above definition.

Lemma 3.3

For any two non-empty subsets A and B of L, we have the following:

  1. 1.

    \(A^{\circ } = \bigcap \limits _{a\in A}(a)^{\circ },\),

  2. 2.

    AA = {0},,

  3. 3.

    AB implies B ⊆A ,,

  4. 4.

    AA ∘∘,,

  5. 5.

    A ∘∘∘ = A ,,

  6. 6.

    A = L if and only if A = {0},.

In case of ideals, we have the following result.

Proposition 3.4

For any two ideals I, J of L, (I∨J) = I ∩J .

Proof

Clearly (IJ)I J . Conversely, let xI J . Let cIJ be an arbitrary element. Then c = ij for some iI and jJ. Now (x) ∨ (c) = (x) ∨ (ij) = (x)∨{(i)∩(j)}={(x) ∨ (i)}∩{(x) ∨ (j)}=LL = L. Therefore x ∈ (IJ). □

The following corollary is a direct consequence of the above results.

Corollary 3.5

For any a, b ∈ L, we have the following:

  1. 1.

    ab implies (b) ⊆ (a) ,

  2. 2.

    (ab) = (a) ∩ (b) ,

  3. 3.

    (a) = L if and only if a = 0.

For any ideal I of a distributive lattice L, it can be easily observed that I I . However, we derive a set of equivalent conditions for every ideal to satisfy the reverse inclusion which is not true in general.

Theorem 3.6

The following conditions are equivalent in a lattice L.

  1. 1.

    L is normal,

  2. 2.

    For any ideals I, J of L, I ∩ J = {0} if and only if I ⊆ J ,

  3. 3.

    For any ideal I of L, I = I ,

  4. 4.

    For any a ∈ L, (a) = (a) .

Proof

  1. (1)⇒(2):

    Assume that L is normal. Suppose IJ = {0}. Let xI. For any aJ, we get xaIJ = {0}. Hence xa = 0. Since L is normal, we get (x) ∨ (a) = L. Thus it concludes that xJ . Therefore IJ . Conversely, suppose that IJ . Let xIJ. Then xIJ . Hence xJJ ={0}. Therefore IJ = {0}.

  2. (2)⇒(3):

    Assume the condition (2). Let I be an ideal of L. Clearly I I . Conversely, let xI . Then, for any aI, we have the following consequence:

    $$\begin{array}{@{}rcl@{}} x\wedge a = 0 & \Rightarrow & (x]\cap (a] = (0]\\[.5pt] & \Rightarrow & (x]\subseteq (a)^{\circ }\hspace{1 cm}\text{ by condition (2) }\\[.5pt] & \Rightarrow & (x]\subseteq \bigcap \limits_{a\in I}(a)^{\circ }\\[.5pt] & \Rightarrow & x\in I^{\circ }. \end{array} $$

    Hence it concludes I I . Therefore I = I .

  3. (3)⇒(4):

    It is obvious.

  4. (4)⇒(1):

    Let x, yL be such that xy = 0. Then x ∈ (y) = (y). Hence (x) ∨ (y) = L. Therefore L is normal.

The concept of disjunctive ideals is now introduced in lattices.

Definition 3.7

An ideal I of a lattice L is called disjunctive if for all x, yL, (x) = (y) and xI imply that yI.

Clearly, each (x), xL is a disjunctive ideal. It is evident that any ideal I is a disjunctive ideal if it satisfies (x)∘∘I for all xI. By Theorem 3.6, it can be easily observed that the disjunctive ideals are equivalent to α-ideals of Cornish [4] in the case of normal lattices.

Theorem 3.8

The following conditions are equivalent in a lattice L:

  1. 1.

    Every ideal is a disjunctive ideal.

  2. 2.

    Every principal ideal is a disjunctive ideal.

  3. 3.

    Every prime ideal is a disjunctive ideal.

  4. 4.

    For a, b ∈ L,(a) = (b) implies (a] = (b].

Proof

  1. (1)⇒(2):

    It is clear.

  2. (2)⇒(3):

    Assume that every principal ideal is a disjunctive ideal. Let P be a prime ideal of L. Suppose (a) = (b) and aP. Then clearly (a]⊆P. Since (a) = (b) and (a] is a disjunctive ideal, we get that b ∈ (a]⊆P. Therefore, P is a disjunctive ideal.

  3. (3)⇒(4):

    Assume that every prime ideal of A is a disjunctive ideal. Let a, bA such that (a) = (b). Suppose (a]≠(b]. Without loss of generality assume that \((a]\nsubseteq (b]\). Let \({\Sigma } = \{~I\in \mathcal {I}(L)~|~ a\wedge b\in I\text { and } a\notin I~\}\). It is clear that (ab] ∈ Σ. Let {I i } i∈Δ be a chain in Σ. Then clearly \(\bigcup _{i\in {\Delta }} I_{i}\) is an ideal, \(a\wedge b\in \bigcup _{i\in {\Delta }} I_{i}\) and \(a\notin \bigcup _{i\in {\Delta }} I_{i}\). Hence, \(\bigcup _{i\in {\Delta }} I_{i}\) is an upper bound for {I i } i∈Δ in Σ. Therefore, by Zorn’s Lemma, Σ has a maximal element, say P. We now prove that P is a prime ideal in L. Let x, yL such that xP and yP. Hence PP∨(x] and PP∨(y]. Therefore, by the maximality of P, P∨(x] and P∨(y] are not in Σ. Hence aP∨(x] and aP∨(y]. So we have the following consequence

    $$\begin{array}{@{}rcl@{}} a & \in & \{~P\vee (x]~\}\cap \{~P\vee (y]~\}\\ & = & P\vee \{~(x]\cap (y]~\}\\ & = & P\vee (x\wedge y] \end{array} $$

    If xyP, then aP∨(xy]=P, which is a contradiction to that aP. Thus we get xyP. Hence P is a prime ideal. Therefore, by hypothesis (3), we can get that P is a disjunctive ideal of L. Since P ∈ Σ, we get that abP and aP. Since P is prime, we get bP. Since bP and P is disjunctive, we get aP, which is a contradiction to aP. Therefore, it concludes that (a] = (b].

  4. (4)⇒(1):

    Assume the condition (4). Let I be an arbitrary ideal of L. Suppose a, bL are such that (a) = (b). Then by condition (4), we get (a] = (b]. Suppose aI. Then b ∈ (b] = (a]⊆I. Therefore I is a disjunctive ideal of L.

The notion of normal prime ideal is now introduced in a lattice.

Definition 3.9

A prime ideal P of a lattice L is called a normal prime ideal if to each xP, there exists x P such that (x) ∨ (x ) = L.

Proposition 3.10

Every normal prime ideal is a minimal prime ideal.

Proof

Let P be a normal prime ideal of a lattice L. Suppose xP. Since P is normal, there exists x P such that (x) ∨ (x ) = L. Hence, we get L = (x) ∨ (x )⊆(xx ). Thus by Corollary 3.5(3), we get that xx =0. Therefore, P is a minimal prime ideal of L. □

In general, the converse of the above Proposition 3.10 is not true. It can be seen in the following example.

Example 3.11

Consider the following distributive lattice L = {0, a, b, c, 1} whose Hasse diagram is given by:

figure a

Consider the prime ideal P = {0, a}. It can be easily observed that P is a minimal prime ideal but not a normal prime ideal.

However, in the following, we derive a sufficient condition for every minimal prime ideal to become a normal prime ideal.

Proposition 3.12

If L is a normal lattice, then every minimal prime ideal of L is a normal prime ideal.

Proof

Assume that L is normal and P a minimal prime ideal of L. Let xP. Then there exists x P such that xx =0. Since L is normal, we get (x) ∨ (x ) = (x) ∨ (x ) = L. Therefore, P is normal in L. □

Proposition 3.13

Let P be a normal prime ideal of a lattice L. Then for each x ∈ L, we have the following property:

$$x\notin P~\Leftrightarrow ~(x)^{\circ }\subseteq P.$$

Proof

Let P be a normal prime ideal of L and xL. Suppose xP. Then clearly (x)P. Conversely, assume that (x)P. Suppose xP. Since P is normal, there exists x P such that (x) ∨ (x ) = L. Hence L = (x) ∨ (x )⊆(x) ∨ (x ). Then x ∈(x)P, which is a contradiction. Therefore xP. □

Theorem 3.14

Every normal prime ideal is a disjunctive ideal.

Proof

Let P be a normal prime ideal of L. Suppose x, yL are such that (x) = (y) and xP. Since P is normal, there exists x P such that (x) ∨ (x ) = L. Hence L = (x) ∨ (x ) = (y) ∨ (x )⊆(yx ). Hence by Corollary 3.5(3), we get yx = 0 ∈ P. Since P is prime and x P, it yields that yP. Therefore, P is a disjunctive ideal. □

In what follows, we prove a necessary and sufficient condition for the inverse image of a disjunctive ideal to become again a disjunctive ideal.

Theorem 3.15

Let f be a homomorphism of a lattice (L, ∨, ∧) onto another lattice (L , ∨, ∧). If J is a disjunctive ideal of L , then the following conditions are equivalent:

  1. 1.

    f −1 (J) is a disjunctive ideal in L,

  2. 2.

    For each x ∈ L , f −1 ((x) ) is a disjunctive ideal in L.

Proof

  1. (1)⇒(2):

    Assume that f −1(J) is a disjunctive ideal in L for each disjunctive ideal J of L . Since (x) is a disjunctive ideal in L for each xL , we get from (1) that f −1((x)) is a disjunctive ideal in L.

  2. (2)⇒(1):

    Assume that f −1((x)) is a disjunctive ideal in L for each xL . Let J be a disjunctive ideal of L . Then clearly f −1(J) is an ideal in L. Let x, yL be such that (x) = (y) and xf −1(J). Then f(x)∈J. For any aL , we get the following consequence:

    $$\begin{array}{@{}rcl@{}} a\in (f(x))^{\circ } & \Leftrightarrow & f(x)\in (a)^{\circ }\\ & \Leftrightarrow & x\in f^{-1}((a)^{\circ })\\ & \Leftrightarrow & y\in f^{-1}((a)^{\circ })\hspace{0.5 cm}\text{ since }f^{-1}((a)^{\circ })\text{ is disjunctive in }L \\ & \Leftrightarrow & f(y)\in (a)^{\circ }\\ & \Leftrightarrow & a\in (f(y))^{\circ }. \end{array} $$

    Hence, it concludes (f(x)) = (f(y)). Since f(x)∈J and J is a disjunctive ideal, we get f(y)∈J. Hence yf −1(J). Therefore f −1(J) is a disjunctive ideal in L.

We now discuss some properties of direct products of disjunctive ideals of lattices. First, we need the following lemma whose proof is routine.

Lemma 3.16

Let L 1 and L 2 be two distributive lattices. For any a∈L 1 , b ∈ L 2 and (a,b)∈L 1 × L 2 , we have the following:

  1. 1.

    (a, b) = (a) ×(b) ,

  2. 2.

    (a,b) ∨(c, d) = (a∨c, b∨d) ,

  3. 3.

    (a,b) = (a) × (b) .

Theorem 3.17

Let L=L 1 × L 2 be the product of the distributive lattices L 1 and L 2 . If I 1 and I 2 are disjunctive ideals of L 1 and L 2 , respectively, then I 1 ×I 2 is a disjunctive ideal of the product lattice L 1 × L 2 . Conversely, every disjunctive ideal of L 1 × L 2 can be expressed as I = I 1 ×I 2 where I 1 and I 2 are disjunctive ideals of L 1 and L 2 , respectively.

Proof

Let I 1 and I 2 be the disjunctive ideals of L 1 and L 2, respectively. Then clearly, I 1×I 2 is an ideal of L 1×L 2. Let a, cL 1 and b, dL 2 be such that (a, b) = (c, d) and (a, b)∈I 1×I 2. Then aI 1 and bI 2. Since (a, b) = (c, d), we get (a)×(b) = (c)×(d) and hence (a) = (c) and (b) = (d). Since I 1 is a disjunctive ideal and aI 1, we get that cI 1. Similarly, we get dI 2. Hence it concludes that (c, d)∈I 1×I 2. Therefore, I 1×I 2 is a disjunctive ideal in L 1×L 2.

Conversely, let I be a disjunctive ideal of L 1×L 2. Consider I 1={aL 1 | (a, b)∈I for some bL 2}. Clearly, I 1 is an ideal in L 1. Let x, yL 1 be such that (x) = (y) and xI 1. Then (x, a)∈I for some aL 2. Since (x) = (y), we get (x, a) = (x)×(a) = (y)×(a) = (y, a). Since I is a disjunctive ideal in L 1×L 2, we get (y, a)∈I. Hence it concludes that yI 1. Therefore, I 1 is a disjunctive ideal in L 1. Similarly, we can obtain that I 2 is a disjunctive ideal in L 2.

We now prove that I = I 1×I 2. Clearly, II 1×I 2. Conversely, let (a 1, a 2)∈I 1×I 2. Then a 1I 1 and a 2I 2. Hence (a 1, b 1)∈I and (b 2, a 2)∈I for some b 2L 1 and b 1L 2. Hence, we get (a 1, 0) = (1, 0)∧(a 1, b 1)∈I and also (0, a 2)=(0,1)∧(b 2, a 2)∈I. Thus (a 1, a 2)=(a 1,0)∨(0, a 2)∈I. Therefore I 1×I 2I. □

We now introduce the concept of weakly normal lattices.

Definition 3.18

A distributive lattice L with 0 is called weakly normal if it satisfies (x) ∨ (y) = (x) ∨ (y) for all x, yL.

It is evident that every normal lattice is weakly normal. In general, the converse is not true. However, in the following, a set of equivalent conditions is derived for every weakly normal lattice to become normal.

Theorem 3.19

Let L be a weakly normal lattice. Then the following conditions are equivalent:

  1. 1.

    L is normal,

  2. 2.

    For x, y ∈ L, (x) ∨(y) = (x ∧ y) ,

  3. 3.

    For x, y ∈ L, x ∧ y = 0 implies (x) ∨(y) = L.

Proof

  1. (1)⇒(2):

    Let L be a weakly normal lattice. Assume that L is a normal lattice. Let x, yL. Since L is a normal lattice, by Theorem 3.6, we get (x) ∨ (y) = (x) ∨ (y) = (xy) = (xy).

  2. (2)⇒(3):

    It is clear.

  3. (3)⇒(1):

    Assume that (3) is valid. Let x, yL be such that xy = 0. Then L = (x) ∨ (y)⊆(x) ∨ (y). Hence L is normal.

4 Strongly disjunctive ideals

In this section, a concept of strongly disjunctive ideals is introduced in lattices. A set of equivalent conditions is derived for the class of all strongly disjunctive ideals to become a sublattice of the ideal lattice.

Definition 4.1

For any ideal I of a lattice L, define β(I) as follows:

$$\beta (I) = \{x\in L~|~(x)^{\circ }\vee I = L\}.$$

The following lemma is an immediate consequence of the above definition.

Lemma 4.2

For any ideals I, J of L, it holds that

  1. 1.

    β (I) ⊆ I,

  2. 2.

    I⊆J implies β(I)⊆β(J),

  3. 3.

    β(I∩J)=β(I)∩β(J).

Proof

The proof of the identities (2) and (3) is routine. It is sufficient to prove the identity (1). For that, let xβ(I). Then (x)I = L. Hence x = ab for some a ∈ (x)⊆(x) and bI. Then we get xa = 0 and xbI. Thus, we can conclude that x = xx = x∧(ab)=(xa)∨(xb)=0∨(xb)=xbI. Therefore β(I) ⊆ I. □

Proposition 4.3

For any ideal I of a lattice L, β(I) is an ideal in L.

Proof

Clearly 0 ∈ β(I). Let x, yβ(I). Then (x)I = (y)I = L. Hence (xy)I = {(x)∩(y)}∨I = {(x)I}∩{(y)I}=L. Hence xyβ(I). Again let xβ(I) and yx. Then L = (x)I⊆(y)I. Thus yβ(I). Therefore, β(I) is an ideal in L. □

Definition 4.4

An ideal I of a lattice L is called a strongly disjunctive ideal if I = β(I).

Proposition 4.5

Every strongly disjunctive ideal is a disjunctive ideal.

Proof

Let I be a strongly disjunctive ideal of a lattice L. Let x, yL be such that (x) = (y) and xI = β(I). Then it is clear that (x)I = L. Hence (y)I = L and so yβ(I)=I. Thus I is a disjunctive ideal of L. □

In general, the converse of the above proposition is not true. However, we can derive a set of equivalent conditions for every ideal of a lattice to become strongly disjunctive.

Theorem 4.6

Consider the following conditions in a lattice L:

  1. 1.

    Every prime ideal is normal.

  2. 2.

    Every ideal is strongly disjunctive.

  3. 3.

    Every prime ideal is strongly disjunctive.

Then (1)⇒(2)⇒(3). If L is a weakly normal lattice, then all the above conditions are equivalent.

Proof

  1. (1)⇒(2):

    Assume that every prime ideal is normal. Let I be an ideal of L. Clearly β(I) ⊆ I. Conversely, let xI. Suppose (x)IL. Then there exists a prime ideal P such that (x)IP. Hence (x)P and xIP. Since P is normal and (x)P, by Proposition 3.13, we get that xP, which is a contradiction to that xP. Hence (x)I = L. Thus it yields that xβ(I). Therefore, I is strongly disjunctive.

  2. (2)⇒(3):

    It is obvious.

  3. (3)⇒(1):

    Suppose that L is weakly normal. Assume that every prime ideal is strongly disjunctive. Let P be a prime ideal of L. Then by our assumption, β(P)=P. Let xP. Then (x)P = L. Hence ab = 1 for some a ∈ (x) and bP. Since a ∈ (x) and L is weakly normal, we get (x) ∨ (a) = (x) ∨ (a) = L. Suppose aP. Then 1=abP, which is a contradiction. Therefore, aP and hence P is a normal prime ideal.

Theorem 4.7

The following conditions are equivalent in a lattice L:

  1. 1.

    (x) ∨(x) ∘∘ = L for all x ∈ L,

  2. 2.

    Every ideal of the form I = I ∘∘ is strongly disjunctive,

  3. 3.

    For each x ∈ L, (x) ∘∘ is strongly disjunctive.

Proof

  1. (1)⇒(2):

    Assume that the condition (1) holds in L. Let I be an ideal in L such that I = I ∘∘. Clearly β(I) ⊆ I. Conversely, let xI. Clearly (x)∘∘I ∘∘. Hence we get L = (x) ∨ (x)∘∘⊆(x)I ∘∘ = (x)I. Thus it yields that xβ(I). Therefore, I is strongly disjunctive.

  2. (2)⇒(3):

    It is obvious.

  3. (3)⇒(1):

    Assume the condition (3). Then we get β((x)∘∘)=(x)∘∘. Since x ∈ (x)∘∘, we get (x) ∨ (x)∘∘ = L. Hence the proof is complete.

Definition 4.8

For any maximal ideal M of a lattice L, define \({\Omega } (M) \,=\, \{x\in L~|~(x)^{\circ }\!\nsubseteq \! M\}\).

Lemma 4.9

For any maximal ideal M,Ω(M) is an ideal contained in M.

Proof

Clearly 0 ∈ Ω(M). Let x, y ∈ Ω (M). Then \((x)^{\circ }\nsubseteq M\) and \((y)^{\circ }\nsubseteq M\). Since M is prime, we get \((x\vee y)^{\circ } = (x)^{\circ }\cap (y)^{\circ }\nsubseteq M\). Hence xy ∈ Ω (M). Let x ∈ Ω (M) and yx. Then \((x)^{\circ }\nsubseteq M\) and (x)⊆(y). Thus we get \((y)^{\circ }\nsubseteq M\) and hence y ∈ Ω (M). Therefore, Ω(M) is an ideal of L. Now, let x ∈ Ω (M). Then \((x)^{\circ }\nsubseteq M\). Hence there exists a ∈ (x) such that aM. Since a ∈ (x), we get L = (a) ∨ (x)⊆(ax). Thus ax = 0 ∈ M. Since M is prime and aM, it concludes that xM. Therefore Ω(M) ⊆ M. □

By μ we denote the set of all maximal ideals of a lattice L. For any ideal I of a lattice L, let μ(I)={Mμ | IM}.

Theorem 4.10

For any ideal I of a lattice \(L, \beta (I) = \bigcap _{M\in \mu (I)}{\Omega } (M)\).

Proof

Let xβ(I) and IM where Mμ. Then L = (x)I⊆(x)M. Suppose (x)M, then M = L, which is a contradiction. Hence \((x)^{\circ }\nsubseteq M\). Thus x ∈ Ω (M) for all Mμ(I). Therefore \(\beta (I)\subseteq \bigcap _{M\in \mu (I)}{\Omega } (M)\). Conversely, let \(x\in \bigcap _{M\in \mu (I)}{\Omega } (M)\). Then x ∈ Ω (M) for all Mμ(I). Suppose (x)IL. Then there exists a maximal ideal M 0 such that (x)IM 0. Hence (x)M 0 and IM. Since IM 0, by hypothesis, we get x ∈ Ω (M 0). Hence \((x)^{\circ }\nsubseteq M_{0}\), which is a contradiction. Therefore (x)I = L. Thus xβ(I). Hence \(\bigcap _{M\in \mu (I)}{\Omega } (M)\subseteq \beta (I)\). □

By the above Theorem 4.10, it can be easily observed that β(I)⊆Ω(M) for every Mμ(I). In what follows, a set of equivalent conditions is derived for the class of all strongly disjunctive ideals of a lattice to become a sublattice of the ideal lattice \(\mathcal {I}(L)\) of the lattice L.

Theorem 4.11

The following conditions are equivalent in a lattice L:

  1. 1.

    For any M ∈ μ, Ω(M) is maximal,

  2. 2.

    For any \(I, J\in \mathcal {I}(L)\) , I∨J = L implies β(I) ∨ β(J) = L,

  3. 3.

    For any \(I, J\in \mathcal {I}(L)\) , β(I)∨β(J)=β(I∨J),

  4. 4.

    For any two distinct maximal ideals M, N, Ω(M) ∨ Ω(N) = L,

  5. 5.

    For any M ∈ μ, M is the unique member of μ such that Ω(M) ⊆ M.

Proof

  1. (1)⇒(2):

    Assume the condition (1). Then it is clear that Ω(M)=M for all Mμ. Let \(I, J\in \mathcal {I}(L)\) be such that IJ = L. Suppose β(I)∨β(J)≠L. Then there exists a maximal ideal M such that β(I)∨β(J) ⊆ M. Hence β(I) ⊆ M and β(J) ⊆ M. Now

    $$\begin{array}{@{}rcl@{}} \beta (I)\subseteq M & \Rightarrow & \bigcap \limits_{M\in \mu (I)} {\Omega} (M)\subseteq M\\ & \Rightarrow & {\Omega} (M_{i})\subseteq M\hspace{0.1 cm}\mathrm{ for\, some } \, M_{i}\in \mu (I)\hspace{0.1 cm}(\mathrm{since }\, M \, \mathrm{ is\, prime})\\ & \Rightarrow & M_{i}\subseteq M\hspace{0.8 cm}\mathrm{ by\, condition (1) }\\ & \Rightarrow & I\subseteq M. \end{array} $$

    Similarly, we can get JM. Hence L = IJM, which is a contradiction. Therefore β(I)∨β(J)=L.

  2. (2)⇒(3):

    Assume the condition (2). Let \(I, J\in \mathcal {I}(L)\). Clearly β(I)∨β(J) ⊆ β(IJ). Let xβ(IJ). Then ((x)I)∨((x)J)=(x)IJ = L. Hence by condition (2), we get β((x)J)∨β((x)J)=L. Thus xβ((x)I)∨β((x)J). Hence x = rs for some rβ((x)I) and sβ((x)J). Now

    $$\begin{array}{@{}rcl@{}} r\in \beta ((x)^{\circ }\vee I) & \Rightarrow & (r)^{\circ }\vee (x)^{\circ }\vee I = L\\ & \Rightarrow & L = ((r)^{\circ }\vee (x)^{\circ })\vee I\subseteq (r\wedge x)^{\circ }\vee I\\ & \Rightarrow & (r\wedge x)^{\circ }\vee I = L\\ & \Rightarrow & r\wedge x\in \beta (I). \end{array} $$

    Similarly, we can get sxβ(J). Hence

    $$\begin{array}{@{}rcl@{}} x & = & x\wedge x\\ & = & x\wedge (r\vee s)\\ & = & (x\wedge r)\vee (x\wedge s)\in \beta (I)\vee \beta (J). \end{array} $$

    Then β(IJ) ⊆ β(I)∨β(J). Therefore β(I)∨β(J)=β(IJ).

  3. (3)⇒(4):

    Assume the condition (3). Let M, N be two distinct maximal ideals of L. Choose xMN and yNM. Now

    $$\begin{array}{@{}rcl@{}} x\notin N & \Rightarrow & \mathrm{there\,\,\, exists }~x_{1}\in N\,\mathrm{ such \,\,\, that }\,x\vee x_{1} = 1,\\ y\notin M & \Rightarrow & \mathrm{there \,\,\,exists }~y_{1}\in M\,\mathrm{ such\,\,\, that }\,y\vee y_{1} = 1. \end{array} $$

    Then (xy 1)∨(yx 1)=(xx 1)∨(yy 1)=1. Now

    $$\begin{array}{@{}rcl@{}} L & = & \beta (L)\\ & = & \beta ((1])\\ & = & \beta (((x\vee y_{1})\vee (y\vee x_{1})])\\ & = & \beta ((x\vee y_{1}])\vee \beta ((y\vee x_{1}])\hspace{0.7 cm}\text{ by condition (3) }\\ & \subseteq & {\Omega} (M)\vee {\Omega} (N)\hspace{0.5 cm}\text{since} \, (x\vee y_{1}]\subseteq M, (y\vee x_{1}]\subseteq N. \end{array} $$

    Therefore Ω(M) ∨ Ω (N)=L.

  4. (4)⇒(5):

    Assume the condition (4). Let Mμ. Suppose Nμ is such that NM and Ω(N) ⊆ M. Since Ω(M) ⊆ M by our hypothesis, we get L = Ω(M) ∨ Ω (N)=M, which is a contradiction. Therefore M is the unique maximal ideal such that Ω(M) is contained in M.

  5. (5)⇒(1):

    Let Mμ. Suppose Ω(M) is not maximal. Let M 0 be a maximal ideal of L such that Ω(M) ⊆ M 0. We have always Ω(M 0)⊆M 0, which is a contradiction to the hypothesis.