This paper mainly discusses the problems raised in Kalmbach's book: When are the p-ideals of an irreducible orthomodular lattice well ordered under set inclusion? We give three classes of orthomodular lattices whose p-ideals set is a chain under set inclusion. Furthermore, this article also provides a sufficient and necessary condition for the p-ideals set of orthomodular lattices to be a chain under set inclusion from the perspective of L-algebras, which gives a new point to solve the questions. Moveover, we also give some characterizations about central elements of orthomodular lattice.
Citation: Ziteng Zhao, Jing Wang, Yali Wu. Note on p-ideals set of orthomodular lattices[J]. AIMS Mathematics, 2024, 9(11): 31947-31961. doi: 10.3934/math.20241535
[1] | Chun Ge Hu, Xiao Guang Li, Xiao Long Xin . Dual ideal theory on L-algebras. AIMS Mathematics, 2024, 9(1): 122-139. doi: 10.3934/math.2024008 |
[2] | Jie Qiong Shi, Xiao Long Xin . Ideal theory on EQ-algebras. AIMS Mathematics, 2021, 6(11): 11686-11707. doi: 10.3934/math.2021679 |
[3] | Fengxia Gao, Jialei Chen . Conjugacy classes of left ideals of Sweedler's four-dimensional algebra $ H_{4} $. AIMS Mathematics, 2022, 7(5): 7720-7727. doi: 10.3934/math.2022433 |
[4] | Ahmed Ramadan, Anwar Fawakhreh, Enas Elkordy . Novel categorical relations between $ \mathcal{L} $-fuzzy co-topologies and $ \mathcal{L} $-fuzzy ideals. AIMS Mathematics, 2024, 9(8): 20572-20587. doi: 10.3934/math.2024999 |
[5] | Xin Zhou, Xiao Long Xin . Ideals on neutrosophic extended triplet groups. AIMS Mathematics, 2022, 7(3): 4767-4777. doi: 10.3934/math.2022264 |
[6] | Jehan A. Al-Bar, T. Asir, K. Mano, Wafaa M. Fakieh . Class of crosscap two graphs arising from lattices-Ⅱ. AIMS Mathematics, 2023, 8(10): 24802-24824. doi: 10.3934/math.20231265 |
[7] | Pakorn Palakawong na Ayutthaya, Bundit Pibaljommee . On $ n $-ary ring congruences of $ n $-ary semirings. AIMS Mathematics, 2022, 7(10): 18553-18564. doi: 10.3934/math.20221019 |
[8] | Hui Yang, Yi Shi . $L$-biconvex sets on some fuzzy algebraic substructures. AIMS Mathematics, 2020, 5(5): 4311-4321. doi: 10.3934/math.2020275 |
[9] | Seok-Zun Song, Hee Sik Kim, Young Bae Jun . Commutative ideals of BCK-algebras and BCI-algebras based on soju structures. AIMS Mathematics, 2021, 6(8): 8567-8584. doi: 10.3934/math.2021497 |
[10] | M. Mohseni Takallo, Rajab Ali Borzooei, Seok-Zun Song, Young Bae Jun . Implicative ideals of BCK-algebras based on MBJ-neutrosophic sets. AIMS Mathematics, 2021, 6(10): 11029-11045. doi: 10.3934/math.2021640 |
This paper mainly discusses the problems raised in Kalmbach's book: When are the p-ideals of an irreducible orthomodular lattice well ordered under set inclusion? We give three classes of orthomodular lattices whose p-ideals set is a chain under set inclusion. Furthermore, this article also provides a sufficient and necessary condition for the p-ideals set of orthomodular lattices to be a chain under set inclusion from the perspective of L-algebras, which gives a new point to solve the questions. Moveover, we also give some characterizations about central elements of orthomodular lattice.
The concept of orthomodular lattices (referred to as OML for convenience) originated from von Neumann algebraic theory. In 1936, Birkhoff and von Neumann proposed using lattices of closed subspaces in Hilbert spaces as the fundamental mathematical model for studying quantum mechanics. However, the projection lattice of a Hilbert space is modular only if the Hilbert space is finite dimensional. However, the projection lattice of type Ⅱ von Neumann algebras related to continuous geometries—especially those favoured by John von Neumann—is modular. Therefore, if the projection lattice is required to be a modular lattice, the case of infinite-dimensional Hilbert spaces is excluded. As a remedy, Husimi proposed to study the following orthomodular law instead:
p≤q⇒q=p∨(p′∧q), |
this law, indeed, holds in arbitrary Hilbert spaces.
Finch [1] shows that a congruence relation in an OML L is completely determined by its kernel and that an ideal I of L is called a p-ideal or orthomodular ideal. Let IP(L) denote the set of all p-ideals of an orthomoular lattice. Then, Ip(L) is a complete distributive lattice. In the literature [2], it is raised: when are p-ideals of an irreducible orthomodular lattice well ordered under set inclusion? In the paper, we exhibit three big classes of orthomodular lattices where the answer is positive.
During the past decade, L-algebras occurred in various contexts, including algebraic logic, combinatorial group theory, operator algebras, and number theory (see, e.g., [3,4,5,6]). Wu and Yang provide a succinct description of orthomodular lattices [7], which is called OM-L-algebra. The ideals of an OM-L-algebra L are equivalent to p-filters of orthomodular L. The paper provides a sufficient and necessary condition for the p-ideals set of orthomodular lattices to be a chain under set inclusion from the perspective of L-algebra.
The paper is organized as follows: In Section 2, we recall some basic definitions and facts related to OML L and we exhibit three big classes of OMLs where its p-ideals are set to be a chain under the set inclusion relation. In Section 3, we focus on the OM-L-algebras. A sufficient and necessary condition for the L-ideals set of OM-L-algebra to be a chain is given.
At the beginning of this section, we introduce our basic terminology and present some results.
Definition 2.1. [2] An ortholattice is a structure (L, ≤, ', 0, 1), where (L,≤,′,0,1) is a lattice with maximum 1 and minimum 0, and ′ is a unary operation on L such that the following conditions are satisfied:}
(Involutivelaw)a=a″;(Antitony)ifa≤b,thenb′≤a′;(Excludemiddlelaw)a∨a′=1;(Non−contradictionlaw)a∧a′=0. |
An orthomodular lattice (OML) is an ortholattice (L,≤,′,0,1), that satisfies the following condition for ∀a,b∈L:
(Orthomodularidentity)ifa≤b,thenb=a∨(a′∧b). | (2.1) |
In an ortholattice L, we write aCb (in words, a commutes with b) if a=(a∧b)∨(a∧b′),a,b∈L. Let L be an OML. It follows that aCb, aCb′, a′Cb and a′Cb′ are equivalent. In particular, aCb is characterized by a=(a∨b)∧(a∨b′). Let L be a lattice. For x,y,z∈L, we call {x,y,z} is a distributive triple, if the elements of a triple {x,y,z} can generate a distributive sublattice. In any orthomodular lattice L, let aCb and aCc. Then {a,b,c} generate a distributive sublattice of L [8]. If M is a subset of an orthomodular lattice L, the set C(M)={x∈L∣xCb,∀b∈M} is called the commutant of M in L. The elements of C(L) are called central and C(L) the centre of L. For every subset M of L, the set C(M) is a subalgebra of L containing arbitrary joins and meets, provided they exist in L. In an OML L, x∈L, then x has a unique complement if and only if x∈C(L) [2].
Definition 2.2. Let a,b be elements of an orthomodular lattice L [2].
(ⅰ). a,b∈L, a is said to be perspective to b, a∼b, if a and b have a common complement, i.e., if there exists c in L such that a∨c=1=b∨c, a∧c=0=b∧c hold; elements a,b are said to be strongly perspective, a∼sb, if a and b have a common complement c in the sublattice [0,a∨b], the element c then is called a relative complement of a and b; it is easy to see that strong perspectivity implies perspectivity. Note, by [2] (the parallelogram law for ∼), we have ∀a,b∈L, a∧(a′∨b)∼b∧(b′∨a).
(ⅱ). x is projective to y, denoted x≈y, if there exists a sequence of elements x1,...,xn such that x1=x, xn=y, and xi∼xi+1 hold for 1≤i≤n−1.
(ⅲ). A p-ideal in L is an ideal I closed under perspectivity, i.e., a∈I and a∼b imply b∈I. Dually, a p-filter in L is a filter F closed under perspectivity, i.e., a∈F and a∼b imply b∈F.
Let L be an OML; a,b,c∈L. [2] presents: If c≤a and a∼b, then there exists d∈L such that d≤b and c∼d holds. Dual, we have the following result:
Lemma 2.1. Let L be an orthomodular lattice. Assume that a,b,c∈L satisfies c∼a and a≤b. Then, there exists d such that c≤d and d∼b holds.
Proof. Let y be the common complement of c and a, satisfy y∨c=y∨a=1,y∧c=y∧a=0. Define z=y∧(b′∨y′)≤y, d=c∨(c′∧z′)≥c. Then, z is the common complement of d and b, i.e., d∨z=b∨z=1,d∧z=b∧z=0. In fact, d∨z=(c∨(c′∧z′))∨z=(c∨z)∨(c′∧z′)=(c∨z)∨(c∨z)′=1. Since b′,y′≤b′∨y′, {b,y,b′∨y′} is a distributive triple. We have b∨z=b∨(y∧(b′∨y′)) = (b∨y)∧(b′∨y′∨b) = b∨y≥a∨y=1, so b∨z=1. Since c′∧z′≤c′,z′, {c,z,c′∧z′} is a distributive triple. We have d∧z=(c∨(c′∧z′))∧ z=(c∧z)∨(c′∧z′∧z) = (c∧z)∨0 = c∧z≤c∧y=0, b∧z=b∧(y∧(b′∨y′)) = (b∧y)∧(b′∨y′) = (b∧y)∧(b∧y)′=0.
Corollary 2.1. Let L be an orthomodular lattice. For a,b,c∈L with a≈b≤c. Then, there exists d such that a≤d≈c holds.
Proof. Assume first that a≈b≤c. There exist z1,z2,⋯,zn such that a=z1∼⋯∼zn−2∼zn−1∼zn=b≤c. For zn−1∼zn=b≤c. By Lemma 2.1, ∃dn∈L, we have a=z1∼⋯∼zn−2∼zn−1≤dn∼c. Repeat to use the lemma 2.1, ∃dn−1,dn−2,⋯,d2, such that a=z1≤d2⋯∼dn−2∼dn−1∼dn∼c, i.e., ∃d=d2, a≤d≈c.
Let L be an orthomodular lattice. In [1], Finch shows that Con(L) is isomorphic to Ip(L), where Con(L) represents the set of all congruence relations on L, Ip(L) represents the set of all p-ideals on L, i.e., congruence relations in an OML L is completely determined by its kernel, and an ideal I of L is the kernel of a congruence relation if and only if it is p-ideal (or orthomodular ideal). There are a lot of other conditions that can be used to define p-ideal. More precisely, we have:
Lemma 2.2. [2] If I is an ideal of orthomodular latticeL, then, the following statements are equivalent:
(ⅰ) I is a p-ideal;
(ⅱ) a∈I implies x∧(x′∨a)∈I for all x∈L.
Dual to Lemma 2.2, we have the following result:
Lemma 2.3. [9] If F is a filter of orthomodular lattice L, then, the following statement are equivalent:
(ⅰ) F is a p-filter;
(ⅱ) a∈F implies x∨(x′∧a)∈F for all x∈L.
Remark 2.1. Let L be an orthomodular lattice. If I be an ideal of L, I′={x′∈L∣x∈I}, by Lemmas 2.2 and 2.3, we can obtain I is a p-ideal of L if and only if I′ is a p-filter of L, and vice versa.
Lemma 2.4. Let L be an orthomodular lattice, y∈L. Then, [0,y] is a p-ideal if and only if y∈C(L).
Proof. Let y∈C(L). Clearly, [0,y] is the principle ideal of L. Set a∈[0,y], i.e., 0≤a≤y. Since y∈C(L), then ∀x∈L, xCy. Then, 0≤x∧(x′∨a) ≤x∧(x′∨y) ≤(x∨y)∧(x′∨y)=y. Consequently, x∧(x′∨a)∈[0,y]. By Lemma 2.2, we have [0,y] is a p-ideal.
Conversely, if [0,y] is a p-ideal, ∀x∈L, since x′,y≤x′∨y, then x′C(x′∨y),yC(x′∨y). Then, we have (x′∨y)Cx,(x′∨y)Cy, so {x,y,x′∨y} is a distributive triple; hence, (x′∨y)∧(x∨y)=((x′∨y)∧x)∨((x′∨y)∧y)=((x′∨y)∧x)∨y. Since [0,y] is a p-ideal and y∈[0,y], by Lemma 2.2, we have x∧(x′∨y)∈[0,y], i.e., x∧(x′∨y)≤y. So, (x′∨y)∧(x∨y)=((x′∨y)∧x)∨y=y, i.e., y∈C(L).
The properties of commute listed above, together with Lemma 2.4, imply.
Corollary 2.2. Let L be an orthomodular lattice, [y,1] is a p-filter if and only if y∈C(L).
Proposition 2.1. Let L be an orthomodular lattice. Then, c∈C(L) if and only if ∀a∈L,c∼a implies a=c.
Proof. Let c∈C(L), c∼a. Then, a and c have a common complement c′∈L. Since aCc⇔a=(c∨a)∧(c′∨a)=(c∨a)∧1=c∨a, then c≤a. Similarly, cCa⇔c=(c∨a)∧(c∨a′)=(c∨a)∧(c′∧a)′=(c∨a)∧0′=(c∨a)∧1=c∨a, which implies a≤c. Thus, a=c.
Conversely, assuming c∉C(L), we have c′∉C(L). Then, c′ has at least two different complements, a,c. Obviously, if c∼a, which is contrary to a=c. Hence, c∈C(L).
Corollary 2.3. Let L be an orthomodular lattice. Then, c∈C(L) if and only if ∀a∈L,c≈a implies a=c.
Proof. Let c∈C(L) and c≈a. Then, ∃d1,d2,...,dn∈L, c=d1∼d2∼...∼dn=a. Since c∼d2, By proposition 2.1, we have c=d2, and repeat to use proposition 2.1, we have c=d1=d2=...=dn=a.
Conversely, Let ∀a∈L,c≈a implies a=c and c∉C(L). proposition 2.1 yields ∃a∈L,c∼a and a≠c. Clearly, c∼a implies c≈a, it can be inferred that a=c, which is a contradiction. Hence, c∈C(L).
An order ideal in a lattice L is a subset I, such that x∈L,y∈I, and x≤y imply x∈I.
Lemma 2.5. [2] Let L be an orthomodular lattice and M be an order ideal of L that is closed under perspectivity:
(ⅰ) The smallest p-ideal containing M is the set B={c∈L∣, there exists a finite subset C of M with c = ⋁C},
(ⅱ) The set Na={d∈L∣d≈e,e≤a} is an order ideal and is closed under perspectivity. Pa denotes the smallest p-ideal containing Na.
Proposition 2.2. Let L be an orthomodular lattice. Then, c∈C(L) if and only if Nc=[0,c].
Proof. Let c∈C(L). Lemma 2.5 gives [0,c]⊆Nc. Let ∀d∈Nc. Then, ∃e≤c, d≈e. By Corollary 2.1, we have ∃d1∈L, d≤d1 and d1≈c. By Corollary 2.3, we have d1=c. Hence, d≤d1=c, and then d∈[0,c], hence Nc⊆[0,c], so Nc=[0,c].
If c∉C(L), we have c′∉C(L). Then, c′ has at least two different complements a,c, whence a∈Nc=[0,c]. Thus, a≤c, whence aCc. Then, a=(a∨c)∧(a∨c′)=(a∨c)∧1=c∨a=c, which is a contradiction. Hence, c∈C(L).
Proposition 2.3. Let L be an orthomodular lattice. Then, Nc∩Nc′={0} if and only if c∈C(L).
Proof. Let c∼a. Then, c′∼a′. Since a∼c∈Nc,a′∼c′∈Nc′. By Lemma 2.5, we have a∈Nc,a′∈Nc′. Then, c∧(c′∨a′)∼a′∧(a∨c), since a′∧(a∨c)≤a′∈Nc′, then c∧(c′∨a′)∈Nc′. c∧(c′∨a′)≤c yields c∧(c′∨a′)∈Nc. Thus, c∧(c′∨a′)∈Nc∩Nc′. Similarly, a∧(a′∨c′)∈Nc∩Nc′={0}. Then, c∧(c′∨a′)=c∧(c∧a)′=0, since Nc∩Nc′={0}. Then, c≤a. Similarly, a≤c, i.e., a=c. Thus, proposition 2.1 implies that c∈C(L).
Conversely, let c∈C(L). By Proposition 2.2, we get Nc=[0,c],Nc′=[0,c′]. Assume that d∈Nc∩Nc′, then d≤c,c′⇔d≤c∧c′=0, i.e., d=0. Hence, Nc∩Nc′={0}.
Proposition 2.4. Let L be an orthomodular lattice. M and N are order ideals and closed under perspectivity; PM and PN are the smallest p-ideals containing M and N, respectively. Then, M∩N={0}⇔PM∩PN={0}.
Proof. Let d∈PM∩PN; we have d∈PM,PN. Then, Lemma 2.5 yields that there exists a finite subset C of M, with d=∨C. Let ei∈C,∀i∈I, then ei≤∨C=d∈PM∩PN. Since PM∩PN is an ideal, then ei∈PM∩PN⊆PN. By Lemma 2.5, there exists a finite subset Di of N, with ei=∨Di. Let fij∈Di,∀j∈J, we have fij∈Di⊆N and fij≤∨Di=ei∈C⊆M. Then, fij∈M∩N={0}, i.e., fij=0. Hence, d=∨C=∨i∈Iei=∨i∈I∨j∈Jfij=0. Therefore, PM∩PN={0}.
Conversely, since M⊆PM, N⊆PN, M∩N⊆PM∩PN={0}. So, M∩N={0}.
Recall that a poset P satisfies the ascending condition (ACC) and is called Noetherian when every nonvoid subset of P has a maximal element. It satisfies the descending chain condition (DCC) when its dual satisfies the ACC. Nonvoid chains that satisfy the DCC are called well-ordered sets. The length l(P) of a poset P is defined as the least upper bound of the lengths of the chains in P. When l(P) is finite, P is said to be finite length. A poset P is chain-finite if all chains in L are finite. It is easy to show that a finite length lattice must be chain-finite; a chain-finite lattice must be Noetherian. The dual of a poset P is that poset ˜P defined by the converse partial ordering relation on the same elements.
Lemma 2.6. [8] The following conditions about a poset P are equivalent:
(ⅰ) P is Noetherian,
(ⅱ) every chain in the dual ˜P of P is well-ordered,
(ⅲ) every well-ordered subset of P is finite.
Lemma 2.7. [10] In every infinite boolean algebra, there is an infinite pairwise disjoint family, a strictly increasing infinite sequence is and a strictly decreasing infinite sequence.
Next, we give a characterization of Noetherian orthomodular lattice and boolean lattice.
Proposition 2.5. Let L be a boolean lattice. Then, L is Noetherian if and only if it is finite.
Proof. Let L be a Noetherian boolean lattice. Assume that L is infinite. By Lemma 2.7, L has a strictly increasing sequence A. Since L is a Noetherian lattice, then A′={x∈L|x′∈A} is also a Noetherian lattice. Let B be any subset of A, then B′ is a subset of A′, so B′ has a maximal element b′, which is equivalent to B having a minimal element, i.e., A satisfies the descending chain condition, thus A is well-ordered. By Lemma 2.6, A is finite, which is a contradiction. Thus, L is finite. Conversely, any well ordered subset of a finite lattice is finite. By Lemma 2.6, a finite lattice is a Noetherian lattice.
Proposition 2.6. Let L be an orthomodular lattice. Then, L is Noetherian if and only if it is chain-finite.
Proof. Assume that L has an infinite chain, say A. Then A generates a boolean subalgebra B of L; B is also Noetherian. By Proposition 2.5, B is finite; it follows that A is finite, which is a contradiction. Then L is chain-finite. The converse is trivial.
Corollary 2.4. Let L be an orthomodular lattice. Then, L is Noetherian ⇔ L satisfies DCC.
Proof. L is Noetherian ⇔ L is chain-finite ⇔ ˜L is chain-finite ⇔ ˜L is Noetherian ⇔ L satisfies DCC.
Let (P;≤) be a partially ordered set, a,b∈P and a≤b. If a<b and [a,b]={a,b}, then b covers a (or a is covered by b). We use the symbol b≻a (or a≺b)to indicate that b covers a (or a is covered by b). Let P have a minimum element of 0. We call a∈L an atom of P if 0≺a. Let L be a lattice. if for all x∈L∖{0}, there exists an atom a of L such that a≤x. Then, L is called an atomic lattice. Dually, let L be a lattice with a maximum element 1. And we call b∈L is a coatom of L if b≺1. For all y∈L∖{1}, there exists a coatom b of L such that y≤b. Then, L is called a coatomic lattice. Let L be an atomic orthomodular lattice. For ∀x∈L−{1}, we have x=⋁{a∈L∣a≤x,aisanatom}. Dually, let L be a coatomic orthomodular lattice. For ∀y∈L−{0}, we have x=⋀{b∈L∣b≥x,bisacoatom}. An element b of an atomic orthomodular lattice is finite if it is the join of finitely many atoms.
Corollary 2.5. Let L be a Noetherian orthomodular lattice. Then, L is atomic.
Proof. L is Noetherian. Corollary 2.4 implies that L satisfies DCC (Descending Chain Condition), i.e., every nonvoid subset of P has a minimal element. Thus, for ∀x∈L and 0<x, (0,x] has a minimal element, which is an atom. This proves that L is atomic.
We are now ready to have a look at the p-ideals set of irreducible OML. First, however, we shall need some additional machinery.
Definition 2.3. [2] Every orthomodular lattice L is irreducible, i.e., L is isomorphism to a product M×N of orthomodular lattices implies that |M|=1 or |N|=1, where |M| denotes cardinality of M.
Remark 2.2. It is easy to see that C(L)={0,1} is equivalent to the irreducibility of L.
Lemma 2.8. [2] Let L be an orthomodular lattice. If L is chain-finite. Then, Con(L) is isomorphic to C(L). In particular, Con(L) is a boolean algebra.
Theorem 2.1. Let L be an irreducible orthomodular lattice and Noetherian. Then, the p-ideals are well ordered under set inclusion.
Proof. Let L be Noetherian. Proposition 2.6 implies that L is chain-finite. Since L is irreducible, Remark 2.2 implies that C(L)={0,1}. By Lemma 2.8 and the fact that Con(L) and Ip(L) are isomorphic, we have Ip(L)={{0},L}, Hence, the p-ideals set is well ordered under set inclusion.
Next, we introduce our basic terminology and notations, which are borrowed from [2]:
Let L be an OML. The binary relation ∇,S0,S1,S∞ are defined in an orthomodular lattice L by:
a▽bif(a∨x)∧b=x∧bfor allx∈L.aS0bifc≤a,d≤bandc∼sdimplyc=0=d.aSbifa∧b=0anda∈C([0,a∨b]).aS1bifc≤a,d≤bandc∼dimplyc=0=d.aS∞bifc≤a,d≤bandc≈dimplyc=0=d. |
Immediately, we can obtain aS∞b⇒aS1b⇒aS0b. Furthermore, we have the following results:
Lemma 2.9. [2] Let L be an orthomodular lattice. For ∀a,b∈L, aS0b⇔aSb.
Lemma 2.10. [2] For elements a,b of an orthomodular lattice L, the following statements are equivalent:
(ⅰ) a∇b,
(ⅱ) aS1b,
(ⅲ) a∨x=1 implies b≤x,
(ⅳ) (a∨x)∧(b∨x)=x,∀x∈L.
Proposition 2.7. Let L be an orthomodular lattice, a,b∈L. Then, Na∩Nb={0} if and only if aS∞b.
Proof. For c,d∈L, let c≤a,d≤b,c≈d, then c,d∈Na∩Nb={0}, whence c=0=d. Hence, aS∞b.
Conversely, let d∈Na∩Nb. There exists c,e, such that d≈c≤a,d≈e≤b. Then, c≈e. Since aS∞b, we have c=0=e. Thus, d=0, which proves that Na∩Nb={0}.
Corollary 2.6. Let L be an orthomodular lattice, a∈L. Then, a∇a′⇔aS1a′⇔aS0a′⇔aSa′⇔aS∞a′⇔a∈C(L).
Proof. By Lemma 2.9, we have aS0b⇔aSb. By Lemma 2.10, this gives a∇a′⇔aS1a′⇔a∈C(L) and aS∞b⇒aS1b⇒aS0b. It suffices to prove a∈C(L)⇒aS∞a′ and aSa′⇒a∈C(L). For a∈C(L)⇒aS∞a′, by Propositions 2.3 and 2.7. We have Na∩Na′={0} and aS∞a′, then a∈C(L)⇒aS∞a′. Let aSa′, we obtain a∈C([0,a∨a′])=C([0,1])=C(L). So, aS∞a′⇒aS0a′⇒aSa′⇒a∈C(L).
Let L be an orthomodular lattice, a∈L, M={z∈C(L)∣a≤z}. If the infimum of M exists, we denote v(a)=⋀M and say v(a) is the center cover of a. If for ∀a∈L, C([0,a])={a∧c∣c∈C(L)}, we say that the relative centre property holds.
Lemma 2.11. [2] Assume that the orthomodular lattice L has the relative centre property. For elements a,b of L, the following statements are equivalent:
(ⅰ) There exists c∈C(L) with a≤c and b≤c′,
(ⅱ) aS1b,
(ⅲ) aS0b.
Lemma 2.12. [2] Let L be a complete orthomodular lattice. For ∀a,b∈L, there exists c,d,e,f∈L such that the following conditions hold:
(ⅰ) a=c∨d,b=e∨f,
(ⅱ) c≤d′,e≤f′,
(ⅲ) c∼se,
(ⅳ) dS0f.
Theorem 2.2. Let L be a complete irreducible orthomodular lattice with the relative centre property. Then, the p-ideals set Ip(L) is a chain under set inclusion.
Proof. If L is irreducible, then by Remark 2.2, we get C(L)={0,1}. If L has the relative centre property, by Lemma 2.11, we have aS0b⇔a=0 or b=0. Next, we prove ∀a,b≠0,1, Pa⊆Pb or Pb⊆Pa. In fact, Lemma 2.12 implies that there exists c=a,d=0,b=e∨f,e≤f′,c∼se, or b=e,f=0,a=c∨d,c≤d′,c∼se. In the first case, a=c∼e≤b implies a∈Nb and thus Na⊆Nb. Similarly, Nb⊆Na, which yields Pa⊆Pb or Pb⊆Pa. Let I,J∈Ip(L),I⊈J,J⊈I. Then, ∃x∈I−J,y∈J−I. If Px⊆Py, we have x∈Py⊆J, thus x∈J, which is a contradiction; if Py⊆Px, we obtain y∈Px⊆I, which is impossible. Then, Ip(L) is a chain.
Now, we consider the OML with exchange axiom, which states the following:
We say that the atomic orthomodular lattice L satisfies the exchange axiom if one of the equivalent conditions of the following conditions holds for L:
(ⅰ) For atoms p,q of L, the conditions p≤q∨a and p∧a=0 imply q≤p∨a;
(ⅱ) If for an atom p of L holds p≠a, then p∨a covers a;
(ⅲ) If a covers a∧b, then a∨b covers b.
Lemma 2.13. [2] Let L be an atomic complete orthomodular lattice satisfying the exchange axiom. Then, L is irreducible if and only if every two atoms of L are perspective.
Theorem 2.3. Let L be an atomic irreducible complete orthomodular lattice satisfying the exchange axiom. Then 1 is finite ⇔Pr=L⇔∀x∈L,x is finite, where ∀r∈A(L), A(L) is the set of all atoms of L. In particular, in the case, Ip(L) is well ordered under set inclusion relations.
Proof. First, we show that p is an atom, p∼q, then q is an atom. In fact, let a be a common complement of p,q. Then, p∧a=0≺p⇒ a≺p∨a=1=q∨a⇒ 0=q∧a≺q, i.e., q is an atom. Then, ∀ r∈A(L), Nr={x|x≈d≤r}={0}∪A(L). By Lemma 2.5, we have Pr={x=q1∨q2∨⋯∨qn|qi∈A(L),i=1,2,⋯,n}. Then, 1 is finite ⇒1∈Pr⇒Pr=L⇒∀x∈L,x∈Pr⇒x is finite ⇒1 is finite. For ∀0≠a∈L,∀ r∈A(L), we have r∈Na. In fact, since 0<a, then ∃s∈A(L), s≤a. Thus, r∼s≤a implies r∈Na. It is a routine to show Pr⊆Pa,∀0≠a∈L,∀r∈A(L). So, Ip(L)={{0},L} is well ordered, as desired.
Definition 3.1. [7] An L-algebra is an algebra (L,→,1) of type (2,0) satisfying:
x→x=x→1=1,1→x=x. | (3.1) |
(x→y)→(x→z)=(y→x)→(y→z). | (3.2) |
x→y=y→x=1⇒x=y. | (3.3) |
For ∀x,y,z∈L, condition (3.1) states that 1 is a logical unit. Note that a logical unit is always unique. There is a partial ordering by [7,P. 2332,Prop.2].
x≤y⇔x→y=1. | (3.4) |
Let L be an L-algebra. By [3], x≤y implies that z→x≤z→y for ∀x,y,z∈L. An L-algebra L is a KL-algebra. If it satisfies:
x≤y→x. | (3.5) |
Definition 3.2. [11] Let L be an L-algebra. A subset Y of L is said to be an L-subalgebra if ∀x,y∈Y implies x→y∈Y. If x→y∈Y holds for ∀x∈L,y∈Y, we call Y an invariant L-subalgebra.
Definition 3.3. [3] Let (L,→,1) be an L-algebra. We call I⊆L an ideal if the following holds for all x,y∈L:
1∈I, | (I0) |
x,x→y∈I⇒y∈I, | (I1) |
x∈I⇒(x→y)→y∈I, | (I2) |
x∈I⇒y→x∈I,y→(x→y)∈I. | (I3) |
We call the ideal of L-algebra is an L-ideal. For an element x∈L, ⟨x⟩ represents the L-ideal generated by x. We will use I(L) to denote the set of all L-ideals of an L-algebra L.
Regarding the relevant background of L-algebra, we refer to [3].
We call an L-algebra (L,→,1) self-similar if, for ∀x∈L, the left multiplication εx:y↦x→y induces a bijection ↓x→L. The inverse of εx gives rise to a multiplication xy:=ε−1y(x) that happens to be associative. This allows to describe a self-similar L-algebra equationally as a monoid with an operation → satisfying:
a→ba=b, | (3.6) |
ab→c=a→(b→c), | (3.7) |
(a→b)a=(b→a)b. | (3.8) |
Axiom (3.6) implies that as a monoid; a self-similar L-algebra is right cancellative. Let L be a self-similar L-algebra. For a,b∈L, we define
a∧b:=(a→b)a. | (3.9) |
Then, (3.9) is an infimum, and for a,b,c∈L, we have
a→(b∧c)=(a→b)∧(a→c). | (3.10) |
Let L be a self-similar L-algebra. Then L satisfies
a→bc=((c→a)→b)(a→c). | (3.11) |
Every L-algebra L admits a self-similar closure, a self-similar L-algebra S(L) with L as an L-subalgebra such that as a monoid, S(L) is generated by L. Up to isomorphism of left hoops, S(L) is the unique self-similar closure of L.
In this article, we regard (L,→,1,0) as OM-L-algebra; S(L) is the self-similar closure of L; for ∀a∈S(L), define a′=a→0. We call C(L)={a∈L∣ab=ba,∀b∈L} the centre of L.
Definition 3.4. [7] We call (L,→) an OM-L-algebra if it is an L-algebra with 0 (i.e., L admits the smallest element 0) and satisfies the following condition, where x′:=x→0.
x′≤y⇒y→x=x. | (3.12) |
Lemma 3.1. [7] Let L be an OM-L-algebra. Then, L is an orthomodular lattice and satisfies the following conditions, where x′:=x→0.
x∧y=(x→(x→y)′)′. | (3.13) |
x∨y=(x′→y′)→x. | (3.14) |
According to the proof process of [7], (3.12) is equivalent to x′≤y⇔x→y=y⇔y→x=x. In particular, ∀x,y∈L,x′≤x→y⇔x′→(x→y)=1, whence x→(x→y)=x→y, and (x→y)→x=x. This shows that OM-L-algebra is sharp [12].
Lemma 3.2. [7] Every orthomodular lattice L gives rise to an OM-L-algebra (L,→,1,0), where
x→y:=x′∨(x∧y). | (3.15) |
Lemma 3.3. [7] Let L be an OM-L-algebra. Then, 0≤a⇔a∈L.
Lemma 3.4. [7] Let L be an OM-L-algebra, a,b∈L. The equivalence aCb⇔ab=ba is valid.
Lemma 3.5. [7] Let L be an OM-L-algebra. a,b∈L. The equivalence ab=ba⇔a⩽b→a holds. In particular, a≤b implies ab=ba.
Lemma 3.6. [7] Let L be an OM-L-algebra. For all a,b∈L, the following are satisfied:
a≤b→a⇔b→a=a′→b′. | (3.16) |
a≤b→a⇔b≤a′→b. | (3.17) |
a≤b→a⇒(b→a)→a=a∨b=(a→b)→b. | (3.18) |
Remark 3.1. Let L be an OM-L-algebra. Then, for all a,x∈L,a≤x→aora∥x→a, where x∥y denotes x,y are incomparable, i.e., x≰y and y≰x. In fact, if not, then there exists b∈L such that b→a<a, i.e., b′∨(b∧a)<a, which implies b′<a. Thus, a>b′∨(b∧a)=a, which is impossible.
Proposition 3.1. Let L be an OM-L-algebra. For x,y,z∈L, we have the following properties:
(ⅰ) Let x≤y,z and y→x=z→x. Then, y=z.
(ⅱ) Let x≤y. Then, there exists a unique element z∈L such that x⩽z and y=z→x.
Proof. (ⅰ) Set x⩽y,z and y→x=z→x; we have (y→x)→x=(z→x)→x. Then, Lemma 3.5 and (3.18) imply that y=z.
(ⅱ) Set x≤y. By Lemma 3.5, we get xy=yx⇔x≤y→x. Then, ∃z=y→x, such that x≤z. (3.18) yields y=(y→x)→x=z→x. By (ⅰ), the uniqueness holds.
Definition 3.5. [13] We define a semibrace A to be a commutative monoid (A,∧) with an additional binary operation A×A˙→A such that
a⋅(b∧c)=(a⋅b)∧(a⋅c)a⋅1=1; | (3.19) |
(a∧b)⋅c=(a⋅b)⋅(a⋅c)1⋅a=a. | (3.20) |
Let L be an L-algebra. We define a ∧-closure of L to be an L-algebra θ(L), which is a semibrace such that L is an L-subalgebra of θ(L), and every a∈θ(L) is of the form a=x1∧⋅⋅⋅∧xn with xi∈L.
Definition 3.6. [14] We call an L-algebra L ∧-closed if L=θ(L).
Proposition 3.2. Every OM-L-algebra is a ∧-closed L-algebra.
Proof. It suffices to prove that the intersection in L is consistent with the intersection in S(L). For ∀x,y∈L, by (3.11), we have 0→(x→y)x = ((x→0)→(x→y))(0→x)=((0→x)→ (0→y))=1, then 0≤(x→y)x. By Lemma 3.3 and (3.9), we have x∧y=(x→y)x∈L. Thus, every OM-L-algebra is a ∧-closed L-algebra.
Lemma 3.7. [15] Let L be a ∧-closed L-algebra. x,y∈L. Then, the following are satisfied.
⟨x⟩∨⟨y⟩=⟨x∧y⟩. | (3.21) |
⟨x⟩∨⟨x→y⟩=⟨x∧y⟩. | (3.22) |
Corollary 3.1. Let L be a ∧-closed L-algebra. a,b∈L. If a∧b=0 and ⟨a⟩⊆⟨b⟩. Then, ⟨b⟩=L.
Proof. By (3.21), we have ⟨b⟩=⟨a⟩∨⟨b⟩=⟨a∧b⟩=⟨0⟩=L.
Proposition 3.3. Let L be an L-algebra. For ∀I∈I(L). Then, I=⋃i∈I0{⟨xi⟩∣∀xi∈I}, where I0 represents the index set.
Proof. For ∀t∈I, we have t∈⟨t⟩⊆⋃i∈I0{⟨xi⟩∣∀xi∈I}. Then, I⊆⋃i∈I0{⟨xi⟩∣∀xi∈I}. For ∀t∈⋃i∈I0{⟨xi⟩∣∀xi∈I}, ∃x0∈I, such that t∈⟨x0⟩⊆I, which yields ⋃i∈I0{⟨xi⟩∣∀xi∈I}⊆I. Thus, I=⋃i∈I0{⟨xi⟩∣∀xi⊆I}.
Definition 3.7. Let L be an L-algebra. L is irreducible if L is isomorphic to a product L1×L2 of L-algebra implies that |L1|=1 or |L2|=1. |L1|, |L2|, respectively, the cardinality of L-algebras L1 and L2.
Proposition 3.4. Let L be an irreducible OM-L-algebra. Assume that a∈L; a is an atom. Then, ⟨a⟩=L.
Proof. Let a be an atom, and L is irreducible. Lemma 3.5 and Remark 3.1 imply that there exists x0∈L and x0≠0,1, such that ax0≠x0a⇔a⪇x0→a⇔ (x0→a)∥a. Since a is an atom, then a∧(x0→a)=0. Since a,x0→a∈⟨a⟩, then 0=a∧(x0→a)∈⟨a⟩, thus ⟨a⟩=L.
Let L be an OM-L-algebra, a subset I⊆L. By [16] and [11], we obtain that I is an L-ideal if and only if I is an invariant L-subalgebra satisfying (I0). Next, we will prove the p-filters in L (L as an orthomodular lattice) are consistent with the L-ideals (L as OM-L-algebra).
Proposition 3.5. Let L be an OM-L-algebra. The p-filter of orthomodular lattice L is the same as the L-ideals of OM-L-algebra L.
Proof. Let F be a p-filter of L. We need to prove that F satisfies (I1) and is an invariant L-subalgebra of L. Let x,x→y∈F. Since x→y=x′∨(x∧y)∈F and F is a p-filter, we have x∧(x→y)∈F. Then, x∧(x→y)=(x→(x→y))x=(x→y)x=x∧y≤y∈F. So y∈F. For ∀y∈L,x∈F, we get y→x=y′∨(y∧x)∈F, whence F is an invariant L-subalgebra of L.
Conversely, let I be an L-ideal of an OM-L-algebra L. Set a,b∈I. We have a→b∈I. Then, (a→b)→(a∧b) = (a→b)→(a→b)a = ((a→(a→b))→(a→b))((a→b)→a) = ((a→b)→(a→b))a=1a = a∈I. By (I1), a∧b∈I. Let a∈I,b∈L,a≤b. Then a→b=1∈I. By (I1), we have b∈I. Then, I is a filter. If a∈I, for ∀x∈L, by (3.15), we obtain x∨(x′∧a)=x′→a∈I.
Proposition 3.6. Let L be an OM-L-algebra. a∈L, a∈C(L) if and only if a↑={x→a∣∀x∈X}.
Proof. Let a be the central element; ∀y∈a↑, a≤y. By Proposition 3.1, there exists a unique element z∈L such that a≤z and y=z→a. Then, y∈{x→a∣∀x∈L}. Hence, a↑⊆{x→a∣∀x∈L}. Set ∀y∈{x→a∣∀x∈L}, there exists x0∈L, such that y=x0→a. Since a is the central element, then Lemma 3.4 implies that ax0=x0a. Since x0a≤a⇔ax0≤a⇔a≤x0→a=y, we have y∈a↑, then {x→a∣∀x∈L}⊆a↑. This shows that a↑={x→a∣∀x∈L}.
Conversely, if a↑={x→a∣∀x∈L}, for ∀x∈L, we have a≤x→a. Lemma 3.5 implies ∀x∈L, a≤x→a⇔ax=xa. Thus, a∈C(L).
Proposition 3.7. Let L be an OM-L-algebra, a∈L. Then, a∈C(L) if and only if ⟨a⟩=a↑.
Proof. If a∈C(L), by Corollary 2.2 and Proposition 3.5, a↑=[a,1] is an L-ideal. Since a∈a↑, we have ⟨a⟩⊆a↑. For ∀y∈a↑, we obtain a≤y. Since ⟨a⟩ is an upper set, it follows that y∈⟨a⟩. Hence, a↑⊆⟨a⟩. Thus, we have ⟨a⟩=a↑.
Conversely, if a∉C(L), then ∃x0∈L, such that ax0≠x0a⇔a≰x0→a, which is impossible since x0→a∈⟨a⟩=a↑. Thus, a∈C(L).
Definition 3.8. [17] An element p of an L-algebra L is said to be prime. If p<1 and x≤p or x→p≤p for all x∈L.
Definition 3.9. [17] We call an L-algebra prime if all of its elements p<1 are prime.
Remark 3.2. Let Ω be a partially ordered set with the greatest element 1. By [3],
x→y:={1,ifx≤y,y,ifx≰y, | (3.23) |
makes Ω into an L-algebra which, satisfies the inequality (3.5). Thus, Ω is a prime KL-algebra. Note that for any KL-algebra, x→p≤p implies that x→p=p.
A prime KL-algebra L is completely determined by its underlying partial order. So, the above example shows that there is a unique prime KL-algebra for each partially ordered set with greatest element 1.
Definition 3.10. [11] We call an element x of an L-algebra L spatial if x=⋀{p∈P(L)∣x≤p} holds in θ(L). If all x∈L are spatial, we say that L has enough primes.
Definition 3.11. [11] We define a closure L-algebra to be an L-algebra L such that for each x∈L, the meet ˉx:=⋀{p∈P(L)∣p≥x} exists in θ(L) and belongs to L. We call an L-algebra L complete if L=θ(L) and its underlying ∧-semilattice is a complete lattice.
Definition 3.12. [11] We call an L-algebra L distributive if its lattice of ideals is distributed.
Lemma 3.8. [17] The lattice of ideals of an L-algebra L is distributive.
Definition 3.13. [18] A complete lattice is said to be a locale if it satisfies the infinite distributive law
a∧⋁i∈Iai=⋁i∈I(a∧ai). | (3.24) |
Lemma 3.9. [11] For the distributive L-algebra L. The ideal lattice I(L) is spatial locale, and the operation on I(L) is
I→J={x∈L∣⟨x⟩∩I⊆J}, | (3.25) |
where ⟨x⟩ represents the ideal generated by x. Then, ∀I,J,K∈I(L). We have
I∩J⊆K⇔∀x∈I:⟨x⟩∩J⊆K⇔I⊆J→K, | (3.26) |
and
I→(J∩K)=(I→J)∩(I→K). | (3.27) |
Theorem 3.1. Let L be an L-algebra. Then, I(L) is a chain under the set inclusion if and only if I(L) is a bounded prime KL-algebra.
Proof. Let I(L) be a chain, I∈I(L), J∈I(L). If J⊆I. Then, J→I=L. If J⊈I, since I(L) is a chain, then I⫋J. Let ∀x∈J→I, we have ⟨x⟩⊆J→I, by (3.26), ⟨x⟩⊆J→I⇔⟨x⟩∩J⊆I. Then, by Lemma 3.8, we can obtain I=I∨(⟨x⟩∩J)=(I∨⟨x⟩)∩(I∨J)=(I∨⟨x⟩)∩J. If J⊆J→I, by (3.26), we get J=J∩J⊆I, which is a contradiction. Then, x∈J→I⊆J. Obviously, x∈⟨x⟩⊆I∨⟨x⟩, we have x∈(I∨⟨x⟩)∩J=I, i.e., J→I⊆I. Since I(L) is a KL-algebra, we have I⊆J→I, then I=J→I. Remark 4 gives I(L) is a prime KL-algebra. Clearly, I(L) is bounded. Thus, I(L) is a bounded prime KL-algebra.
Conversely, assume that I,J∈I(L) is incomparable. Since I∥J, we obtain I≠I∩J, i.e., I⊈I∩J. Since I(L) is a bounded prime KL-algebra, by Remark 3.2, we have I→J=J and I→(I∩J)=I∩J. By (3.27), we have J=L→J=(I→I)→(I→J)=I→(I∩J)=I∩J⊆I, which is a contradiction. Thus, I(L) is a chain under the set inclusion.
It immediately, we have:
Corollary 3.2. Let L be an OML. Then, Ip(L) is a chain under the set inclusion relationship if and only if Ip(L) is a bounded prime KL-algebra.
Proof. By Lemma 3.2, Proposition 3.5, Remark 2.1, and Theorem 3.1, it clearly.
This paper aims to give an answer to the question raised in Kalmbach's book: When are the p-ideals of an irreducible orthomodular lattice well ordered under set inclusion? For irreducible orthomodular lattice L, we prove that if L is Noetherian or completely with the relative central properties or complete atomic satisfying the exchange axiom, then the p-ideal sets of L are a chain (or well ordered) under set inclusion. In Section 3, from the perspective of L-algebra, we obtain that L-ideals lattice is a chain under the set inclusion if and only if L-ideals lattice is a bounded prime KL-algebra. Remarkably, this paper only partially answers the issues mentioned above. Therefore, we look forward to a more in-depth study about it.
Z. T. Zhao and J. Wang are mainly responsible for writing the initial draft; Y. L. Wu is responsible for reviewing, commenting, and revising before and after publication. All authors have read and approved the final version of the manuscript for publication.
This research is supported by the following grant: Cultivation Fund of Qilu University of Technology (Shandong Academy of Sciences) (2023PX068).
All authors declare no conflicts of interest in this paper.
[1] |
P. D. Finch, Congruence relations in orthomodular lattices, J. Aust. Math. Soc., 6 (1966), 46–54. https://doi.org/10.1017/s1446788700003992 doi: 10.1017/s1446788700003992
![]() |
[2] | G. Kalmbach, Orthomodular lattices, Cambridge: Cambridge University Press, 1983. |
[3] |
W. Rump, L-Algebras, self-similarity, and l-groups, J. Algebra, 320 (2008), 2328–2348. https://doi.org/10.1016/j.jalgebra.2008.05.033 doi: 10.1016/j.jalgebra.2008.05.033
![]() |
[4] |
W. Rump, Right l-groups, geometric Garside groups, and solutions of the quantum Yang–Baxter equation, J. Algebra, 439 (2015), 470–510. https://doi.org/10.1016/j.jalgebra.2015.04.045 doi: 10.1016/j.jalgebra.2015.04.045
![]() |
[5] |
W. Rump, The L-algebra of Hurwitz primes, J. Number Theory, 190 (2018), 394–413. https://doi.org/10.1016/j.jnt.2018.03.004 doi: 10.1016/j.jnt.2018.03.004
![]() |
[6] |
W. Rump, Von Neumann algebras, L-algebras, Baer ∗-monoids, and Garside groups, Forum Math., 30 (2018), 973–995. https://doi.org/10.1515/forum-2017-0108 doi: 10.1515/forum-2017-0108
![]() |
[7] |
Y. L. Wu, Y. C. Yang, Orthomodular lattices as L-algebras, Soft Comput., 24 (2020), 14391–14400. https://doi.org/10.1007/s00500-020-05242-7 doi: 10.1007/s00500-020-05242-7
![]() |
[8] | G. Birkhoff, Lattice theory, Providence: American Mathematical Society, 1967. |
[9] | R. Cignoli, Deductive systems and congruence relations in ortholattices, Bulletin of the Section of Logic, 7 (1978), 87–88. |
[10] | J. D. Monk, B. Robert, Handbook of Boolean algebras, Netherlands: North-Holland, 1989. |
[11] |
W. Rump, L-Algebras and topology, J. Algebra Appl., 22 (2023), 2350034. https://doi.org/10.1142/S0219498823500342 doi: 10.1142/S0219498823500342
![]() |
[12] |
W. Rump, Symmetric quantum sets and L-algebras, Int. Math. Res. Notices, 2022 (2022), 1770–1810. https://doi.org/10.1093/imrn/rnaa135 doi: 10.1093/imrn/rnaa135
![]() |
[13] |
W. Rump, Semidirect products in algebraic logic and solutions of the quantum Yang-Baxter equation, J. Algebra Appl., 7 (2008), 471–490. https://doi.org/10.1142/S0219498808002904 doi: 10.1142/S0219498808002904
![]() |
[14] |
W. Rump, X. Zhang, L-Effect algebras, Stud. Logica, 108 (2020), 725–750. https://doi.org/10.1007/s11225-019-09873-2 doi: 10.1007/s11225-019-09873-2
![]() |
[15] |
W. Rump, L. Vendramin, The prime spectrum of an L-algebra, Proc. Amer. Math. Soc., 152 (2024), 3197–3207. https://doi.org/10.1090/proc/16802} doi: 10.1090/proc/16802
![]() |
[16] |
Y. L. Wu, J. Wang, Y. C. Yang, Lattice-ordered effect algebras and L-algebras, Fuzzy Set. Syst., 369 (2019), 103–113. https://doi.org/10.1016/j.fss.2018.08.013 doi: 10.1016/j.fss.2018.08.013
![]() |
[17] |
W. Rump, Prime L-algebras and right-angled Artin groups, Semigroup Forum, 106 (2023), 481–503. https://doi.org/10.1007/s00233-023-10343-4 doi: 10.1007/s00233-023-10343-4
![]() |
[18] | P. T. Johnstone, Stone spaces, Cambridge: Cambridge University Press, 1982. |