arXiv:2010.06763v2 [math.LO] 14 Mar 2021 ∗ Kentarˆo Yamamoto and McDonald Joseph spaces spectral of means by orthocomplemented lattices for Choice-free h aeaaShlrhpFudto o hi upr duri support their for Foundation paper. th Scholarship of Informatics Takenaka of the Institute Methodology pre the and the Berkeley, Logic during California, in of support Group the its thanks for researcher Amsterdam second of University the at orsodn author. Corresponding Keywords. Classification. Subject lattices. orthocomplemented for Bo duality for to Bimb´o’s orthosp Stone and choice-dependent Goldblatt to with approach bras space spectral Bezhanishvili choice-free catego combines co duality our Our the orthospaces. how and Vietoris categor lattices shown of orthocomplemented equivalence then dual of is choice-free category a It to rise representation. give our tions of spaces in spectral isomorphi used the (and filters reducts) orthospace their to to up respect characterize call to we order which then in spaces We spectral principles. of subclass choice new nonconstructive Theorem, the Subbase associated in Alexander’s its choice-free of spaces; use spectral avoids of representation a subclass give special to we lat a work, known orthocomplemented of this well of Within representation is Choice. topological which of free Lemma, Axiom the Zorn’s to upon equivalent Ultrafilte depends the of proof consequence easy whose an is This subcover. finite Abstract. h rtrsace hnsteIsiuefrLgc Languag Logic, for Institute the thanks researcher first The ed pnAeadrsSbaeTerm hc that asserts which Ston Theorem, a space Subbase of logical subsets Alexander’s orthoregular upon clopen pends the via mented rhsae opc pnotoeua ler,Veoi h Vietoris choice. algebra, of orthoregular Axiom open Compact Orthospace, h xsigtplgclrpeetto fa orthocomple an of representation topological existing The oooia ult,Otoopeetdltie Spectra lattice, Orthocomplemented duality, Topological X scmati vr ubscoe oe of cover open subbasic every if compact is 61,0E5 03E25. 06E15, 06C15, pe itrsorthospaces Vietoris upper zc cdm fSine,and Sciences, of Academy Czech e ∗ aaino hsppr The paper. this of paration gtepeaaino this of preparation the ng fSinea h University the at Science of e ewe the between ies ie ymeans by tices n Holliday’s and es htour that sense X ,adComputation and e, rprlattice proper c approach ace Theorem— r nrdc a introduce yo upper of ry diso a of admits pc de- space e la alge- olean ln with along yperspace, mwith sm topo- a nstruc- choice- space, l be - 2 Joseph McDonald and KentarˆoYamamoto

1. Introduction Assuming Alexander’s Subbase Theorem—which asserts that a X is compact if every subbasic open cover of X admits of a finite subcover - Goldblatt demonstrated in [13] that given an orthocomplemented ± lattice L, the orthospace XL generated by the collection of all proper lattice filters F(L) of L (with its associated patch topology) equipped with a binary orthogonal relation ⊥L⊆ F(L) × F(L) which is irreflexive and symmetric, gives rise to a ; namely, a zero-dimensional compact T0 space (or equivalently, a compact Hausdorff space generated by a basis of clopens). In addition, Goldblatt proved that (up to isomorphism) every orthocomple- ± ± mented lattice L arises via the clopen orthoregular subsets of XL = (XL , ⊥L) ordered by set-theoretic inclusion. Much later, Bimb´oin [5] topologized the class of orthospaces as a means to characterize (up to homeomorphism and ± isomorphism with respect to ⊥) the dual space of XL and used this to prove that the category of orthocomplemented lattices along with their associated lattice homomorphisms is dually equivalent to the category of orthospaces along with their associated continuous p-morphisms; namely, continuous func- tions which are p-morphisms with respect to the set-theoretic complements of the orthogonality relations. The presence of a p-morphism associated to this class of continuous functions can be more naturally understood through an emedding of so-called orthologic into the modal logic B. Refer to Goldblatt in [14] for more details. Note that the topological representation just described depends on the Axiom of Choice, as the proof of Alexander’s Subbase Theorem assumes the Ultrafilter Theorem, whose proof depends upon Zorn’s Lemma, which is equivalent to the Axiom of Choice. We refer to [15, 25] for an in-depth exposition concerning how the above choice-principles hang together. The indispensability of the Axiom of Choice within Goldblatt’s representation is a common facet among related topological representation theorems of var- ious classes of ordered algebraic structures. Indeed, Stone’s representation of Boolean algebras via Stone spaces in [26], Priestley’s representation of distributive lattices via Priestley spaces in [24], Esakia’s representation of Heyting algebras via Esakia spaces in [11, 12], and J´onsson and Tarski’s rep- resentation of modal algebras via modal spaces in [19], all depend upon some nonconstructive choice principle. It was however recently demonstrated by Bezhanishvili and Holliday in [4] that a choice-free topological representation of Boolean algebras is achievable, one which is independent of the Boolean Prime Theorem. Whereas Stone’s choice-dependent representation for Boolean algebras shows that any Boolean algebra B be can represented via the clopen sets of a Stone space X, Bezhanishvili and Holliday demonstrated independently of the Boolean Prime Ideal Theorem that every Boolean algebra B arises via the compact open subsets of a X, which are also regular open in the Alexandroff topology UP(X, 6) where 6⊆ X × X is the specialization order over X. In addition, they established a choice-free categorical dual equivalence Choice-free duality for orthocomplemented lattices 3 between the category of Boolean algebras and Boolean homomomorphisms and a the dual category of upper Vietoris spaces and spectral p-morphisms. Their techniques stemmed from Stone’s observation in [27] that distribu- tive lattices can be represented via the compact open subsets of a subclass of spectral spaces as well as Tarski’s discovery in [28, 29] that the regular open subsets of a spectral space give rise to a Boolean algebra. In addition, they incorporated techniques developed by Vietoris in [30] as the subclass of spectral spaces they employ can also be shown as arising as the hyperspace of closed non-empty subsets of a Stone space that comes equipped with the upper Vietoris topology. Lastly, the duality established in [4] can be viewed as a special case of Jipsen and Moshier’s duality for arbitrary lattices devel- oped in [22] modulo the fact that the Jipsen-Moshier duality makes use of the nonconstructive Prime Ideal Theorem. Within this work, we combine Bezhanishvili and Holliday’s choice-free spectral space approach to Stone duality for Boolean algebras with Gold- blatt and Bimb´o’s choice-dependent orthospace approach to Stone duality for orthocomplemented lattices as a means to prove a choice-free topological representation theorem for the class of orthocomplemented lattices by means of a special subclass of spectral spaces, independently of Alexander’s Sub- base Theorem and its associated nonconstructive choice principles. We then introduce a new subclass of spectral spaces which we call upper Vietoris or- thospaces as a means to characterize (up to homeomorphism and isomorphism with respect to ⊥) the spectral space of proper lattice filters used in our rep- resentation. We then prove that the category induced by this class of spectral spaces, along with their associated spectral p-morphisms, is dually equivalent to the category of orthocomplemented lattices, along with their associated lattice homomorphisms. In light of this duality theorem, we then proceed to develop a “duality dictionary” which establishes how various lattice-theoretic concepts (as applied to orthocomplemented lattices) can be translated into their corresponding dual upper Vietoris orthospace counterparts. Throughout the present paper, we assume the general motivations dis- cussed by Herrlich in [15] of investigating mathematical structures based on ZF instead of ZFC and also assume the motivations in [4] of applying this gen- eral constructive (or choice-free) approach to mathematics to the topological duality theory of ordered algebraic structures. Our motivations for studying orthocomplemented lattices is two-fold: First, orthocomplemented lattices, in comparison to Boolean algebras, Heyt- ing algebras, distributive lattices, etc., are a relatively understudied class of lattice structures within duality theory. Second, the class of all orthocomple- mented lattices contains various subclasses of lattice structures that behave as algebraic models for various quantum logics. For instance, the algebraic model for quantum logics of a finite dimensional Hilbert space is a modular lattice and the algebraic model for quantum logics of an infinite dimensional Hilbert space is an orthomodular lattice, both of which are the most im- portant subclasses of the class of orthocomplemented lattices. These insights 4 Joseph McDonald and KentarˆoYamamoto arose, in part, from the discoveries made by Birkhoff and Von Neumann in [6]. The contents of this paper are organized in the following manner: In the second section, we establish the basic algebra of orthocomplemented lattices and discuss some important examples. In the third section, we investigate or- thospaces, spectral spaces, and give the promised choice-free topological rep- resentation theorem for orthcomplemented lattices. In the fourth section, we characterize the choice-free duals of the spectral spaces used in our represen- tation. In the fifth section, we prove the promised choice-free categorical dual equivalence theorem. In light of our duality theorem, in the sixth section we develop a “duality dictionary” which establishes how various lattice-theoretic concepts (as applied to orthocomplemented lattices) can be translated into their corresponding dual UVO-space counterparts.

2. Orthocomplemented lattices The aim of this section is to establish some basic algebraic foundations for the class of orthocomplemented lattices. For a more detailed treatment of orthocomplemented lattices and important subclasses of these lattices, refer to MacLaren in [21], Bruns and Harding in [8], and Kalmbach in [20]. 2.1. Foundations Let L be a set and let ≤⊆ L × L be a partial-ordering over L; namely a binary relation over L which is reflexive, transitive, and anti-symmetric. We will often conflate the induced partially-ordered set (L, ≤) with its underlying carrier set L. We begin however by defining the class of orthocomplemented lattices as a variety (presentable in possibly many distinct signatures) characterized by satisfying finitely many equations. Definition 2.1.1. If L = (L; ∧, ∨,⊥ , 0, 1) is an algebra of type (2, 2, 1, 0, 0), then L is an orthocomplemented lattice (henceforth, an ortholattice) when the following equations are satisfied:

(1) a ∧ (b ∧ c) = (a ∧ b) ∧ c (8) 0 ∨ a = a (2) a ∨ (b ∨ c) = (a ∨ b) ∨ c (9) (a ∧ b)⊥ = a⊥ ∨ b⊥ (3) a ∧ b = b ∧ c (10) (a ∨ b)⊥ = a⊥ ∧ b⊥ (4) a ∨ b = b ∨ c (11) (a⊥ ∧ b⊥)⊥ = a ∨ b (5) a ∧ (b ∨ a)= a (12) (a⊥ ∨ b⊥)⊥ = a ∧ b (6) a ∨ (b ∧ a)= a (13) a ∧ a⊥ =0 (7) 1 ∧ a = a (14) a ∨ a⊥ =1

Observe that the above formulation guarantees that every ortholattice is a bounded complemented lattice satisfying De Morgan’s distribution laws for orthocomplements over meets and joins, so that they are interdefinable lattice operations with respect to orthocomplements. Choice-free duality for orthocomplemented lattices 5

Definition 2.1.2. If L = (L; ∧,⊥ , 0) is an algebra of type (2, 1, 0) whose joins are defined by a ∨ b := (a⊥ ∧ b⊥)⊥ and whose top universal bound is defined by 1 := 0⊥, then L is an ortholattice if (L; ∧, ∨) is a lattice and the following conditions are satisfied: (1) a ∧ a⊥ =0 (2) a ≤ b =⇒ b⊥ ≤ a⊥ (3) a⊥⊥ = a

Definitions 2.1.2.2 and 2.1.2.3 guarantee that the orthocomplement op- erator is an involutive poset isomorphism onto the order-theoretic dual of its domain. That the above two formulations of an ortholattice coincide can be easily verified. Although the equations within Definition 2.1.1 include some redundan- cies, they make explicit the fact that an ortholattice can simply be viewed as a variety which drops the general distributive property of meets over joins and joins over meets; a property characteristic of Boolean algebras. In fact, an algebra B = (A; ∧, ∨, ⊥, 0, 1) of type (2, 2, 1, 0, 0) is a Boolean algebra when B satisfies equations within definition 2.1.1 and in addition, satisfies the following distribution laws: a ∧ (b ∨ c) = (a ∧ b) ∨ (a ∧ c), a ∨ (b ∧ c) = (a ∨ b) ∧ (a ∨ c) (2.1) Given that Definitions 2.1.1 and 2.1.2 of an ortholattice are equivalent, we will adopt the latter for the sake of simplicity. The Hasse diagrams depicted in Figure 1 are examples of ortholattices.

1 1

b c a a⊥ a d d⊥ a⊥ c⊥ b⊥

0 0

Figure 1. the lattices 2 × 2 and O10

Clearly, the 2 × 2 lattice is an example of an ortholattice which is also a Boolean algebra and hence a . The fact that an ortholattice however in general drops the distributive property is easily exhibited within the O10 ortholattice which admits of a subset A ⊆ O10 which is isomorphic to the M3 and N5 lattices, depicted in Figure 2. Note that this implies that the ortholattice O10 is non-distributive, and hence, not a Boolean algebra. This is a consequence of the following well known characterization theorem of distributive lattices. 6 Joseph McDonald and KentarˆoYamamoto

1 1

a

a b c b

c

0 0

Figure 2. The lattices M3 and N5

Theorem 2.1.3 (Birkhoff and Dedekind [10]). A lattice L is distributive if and only if there exists no sublattice A ⊆ L such that either h: A → M3 or i: A → N5 are lattice isomorphisms.

If L and L′ are ortholattices, then h: L → L′ is an ortholattice homo- morphism if h preserves the ortholattice operations from L to L′. An ortho- lattice homomorphism h: L → L′ is an isomorphism if h is bijective and ′ there exists an ortholattice homomorphism i: L → L such that i ◦ h = 1L ′ ′ and h ◦ i = 1L′ where 1L : L → L and 1L′ : L → L are the trivial identity ortholattice homomorphisms on L and L′ respectively.

2.2. Examples We conclude our analysis of orthocomplemented lattices by considering two standard examples.

Example 2.2.1. Every Boolean algebra B with Boolean complements taken to be orthocomplements is an ortholattice.

Hence, although ortholattices are not in general distributive, there are many examples of distributive ortholattices.

Example 2.2.2. Let H be a Hilbert space over a field F (such as R or C); namely a real or complex valued inner product space which is also a com- plete metric space with respect to the metric induced by the inner product h·, ·i: H×H → F associated with H. The collection L(H) of closed linear subspaces of H ordered by subspace inclusion gives rise to an ortholattice in which each closed linear subspace X ⊆ H admits of an orthogonal comple- ment defined by X⊥ = {x ∈H|∀y ∈ X : hx, yi =0}.

Remark 2.2.3. Note that in particular, if H is a finite dimensional Hilbert space, then L(H) is a modular lattice and if H is an infinite dimensional Hilbert space, then L(H) is an orthomodular lattice.

Refer to [3] and [6] for more details pertaining to the modular and orthomodular lattices induced by the lattice of closed linear subspaces of H. Choice-free duality for orthocomplemented lattices 7

3. Representation of ortholattices via spectral spaces We proceed by examining orthospaces and spectral spaces. We then demon- strate how certain spectral spaces give rise to the promised choice-free rep- resentation of ortholattices. Refer to Bell in [3] for an in-depth exposition of the general theory of orthospaces and Dickmann, Tressl, and Schwartz in [9] for an in-depth exposition of the general theory of spectral spaces.

3.1. Orthospaces and orthoregularity Definition 3.1.1. An orthospace is pair (X, ⊥) such that X is a set and ⊥⊆ X2 is a binary orthogonality relation which is irreflexive (i.e., ∀x ∈ X, x 6⊥ x) and symmetric (i.e., ∀x, y ∈ X, if x ⊥ y, then y ⊥ x). Moreover: (1) For every x ∈ X and Y ⊆ X, we define x ⊥ Y ⇐⇒ x ⊥ y, ∀y ∈ Y . (2) Given any Y ⊆ X, we define Y ∗ = {x | ∀y ∈ Y : x ⊥ y}. Informally, Y ∗ can be thought of as the set which is orthogonal to that of Y . The first example of an orthogonality relation we consider can be easily seen as arising via the dot product over a vector space. Example 3.1.2. Let Rn be the n-dimensional Euclidean space. Given non-zero n vectors x = [x1,...,xn], y = [y1,...,yn] ∈ R , we have x ⊥ y if and only if n x · y = xiyi =0. i=1 X Orthogonality relations can also be seen as arising from the space of integrable functions that form a vector space equipped with some inner product. Example 3.1.3. Define a continuous weight function w over some real closed interval [a,b]. Then, two continuous functions f,g : R → R are orthogonal if b hf,giw = f(x)g(x)w(x)dx =0. Za For instance, the functions f(x) = 1 and g(x)= x are orthogonal if 1 hf,giw = f(x)g(x)dx. Z−1 We now turn to a construction from modal logic which is not an example of an orthogonality relation, but rather an example of taking the set-theoretic complement of the orthogonality relation. Example 3.1.4. A modal B-frame F = (X, R) is a Kripke frame (or state space) such that X is a set of states and R ⊆ X × X is a binary accessibility relation (proximity relation) over X which is reflexive and symmetric. Since for arbitrary states x, y ∈ X, we have that x ⊥ y if and only if not xRy, it follows that (X, R) is a B-frame if and only if (X, ⊥) is an orthospace. Note that the modal axiom system KTB is characterized by this class of frames. 8 Joseph McDonald and KentarˆoYamamoto

Remark 3.1.5. Example 3.1.4 will appear later in section 5 when we construct morphisms in the category of the subclass of spectral spaces which character- ize (up to homeomorphism) the spectral spaces used in our representation. As the class of orthospaces (orthoframes) turn out to be reducts of the class of spectral spaces we introduce, we will require that their associated morphisms be spectral maps as well as p-morphisms with respect to the set-theoretic complements of their orthogonality relations. Definition 3.1.6. Let (X, ⊥) be an orthospace. A subset Y ⊆ X is orthoregular (or ⊥-regular) if and only if Y = Y ∗∗ = {z | ∀x ∈ Y ∗ : z ⊥ x}. Example 3.1.7. Any closed linear subspace X ⊆ Rn is orthoregular in the sense that X⊥⊥ = X since Rn = X ⊕ X⊥ meaning that any vector x = n [x1,...,xn] ∈ R can be uniquely written as x = y +z with y = [y1,...,yn] ∈ ⊥ X and z = [z1,...,zn] ∈ X , which implies that 0 = x · z = (y + z) · z = y · z + z · z = z · z and thus z = 0 and x = y. 3.2. Spectral spaces It will be useful to fix the following notation for important subsets of rela- tional topological spaces that will be studied throughout this work. Notation 3.2.1. Given a topologized orthospace (X, 6, ⊥, T ) where T ⊆ P(X) is some topology and 6⊆ X2 is the specialization order over X, we define the following collections of subsets of X as follows: (1) C(X) is the collection of sets that are compact in X; (2) O(X) is the collection of sets that are open in X; (3) R(X) is the collection of sets that are orthoregular in X; (4) UP(X) is the collection of sets that are open in the upset topology (i.e., the upwards closed or upper set topology) on X; (5) RO(X) is the collection of subsets that are regular open in the upset topology; UP(X 6) where 6 is the specialization order over X; (6) CLOP(X) is the collection of sets that are clopen in X; (7) CO(X)= C(X) ∩ O(X); (8) COR(X)= CO(X) ∩ R(X); (9) CORO(X)= CO(X) ∩ RO(X); (10) CLOPR(X) = CLOP(X) ∩ R(X). We will demonstrate that every ortholattice L can be represented as COR(X) for some spectral space X. Recall that a space X is a T0 space if X satisfies the weakest separation axiom for topological spaces; namely, for points x, y ∈ X, if x 6= y, then there exists an U ∈ O(X) such that x ∈ U and y 6∈ U. A space X is a if every basic open cover of X admit of a finite subcover. A space X is coherent if CO(X) be closed under intersection and forms a basis for the topology over X. Lastly, a space X is sober if every completely prime filter in the lattice O(X) of open sets of X is of the form:

OX (x)= {U ∈ O(X) | ∃x ∈ X : x ∈ U} Choice-free duality for orthocomplemented lattices 9

We now recall the definition of a spectral space and then a classical instance of how spectral spaces arise. Definition 3.2.2. A topological space X is a spectral space if:

(1) X is a T0 space; (2) X is a compact space; (3) X is a coherent space; and (4) X is a . Theorem 3.2.3 (Hochster [16]). A topological space X is a spectral space if and only if X is homeomorphic to the spectrum of a commutative ring R. The following results highlight the importance of spectral spaces for the purposes of the present article. Theorem 3.2.4 (Stone [27]). Every distributive lattice can be represented (up to isomorphism) as CO(X) for some spectral space X. Theorem 3.2.5 (Bezhanishvili and Holliday [4]). Every Boolean algebra can be represented (up to isomorphism) as CORO(X) for some spectral space X. Our representation theorem for ortholattices (in Theorem 3.4.2) is very much an analogue of the above representation theorem for Boolean algebras. Definition 3.2.6. Let L be an ortholattice, let F(L) be the collection of all proper lattice filters of L, and define a = {x ∈ F(L) | a ∈ x}. Moreover, let ⊥L⊆ F(L) × F(L) be an orthogonality relation defined by: ⊥ x ⊥L y ⇐⇒ ∃a ∈ Lb : a ∈ x & a ∈ y. Then, we define the following topological spaces: + + (1) XL = (XL , ⊥L) is the space of proper lattice filters of L whose topology + is generated by {a | a ∈ L}, known as the spectral topology over XL . ± ± (2) XL = (XL , ⊥L) is the space of proper lattice filters of L whose topology is generated by {ba | a ∈ L}∪{∁a | a ∈ L} (where ∁ is the set-theoretic ± complement operator) known as the patch topology over XL . b b Note that a ∩ b = a[∧ b and so the basis {a | a ∈ L} of the spectral + topology for the space XL is closed under binary intersections. Moreover, note that since ⊥b L isb an orthogonality relation overb F(L), ⊥L is symmetric so for x, y ∈ F(L), we can alternatively define x ⊥L y if and only if there exists some a ∈ L such that a ∈ x and a⊥ ∈ y. 3.3. The Stone space of an ortholattice Assuming Alexander’s Subbase Theorem, it was shown in [13] that the space ± XL with its associated patch topology is a Stone space that represents (up to isomorphism) the original ortholattice L. As demonstrated in the following proposition, the use of some choice principle in this claim is essential. Proposition 3.3.1. The following are equivalent: (1) PIT, the Prime Ideal Theorem for Boolean algebras. 10 Joseph McDonald and KentarˆoYamamoto

± (2) The space XL is compact for all Boolean algebras L.

Proof. To see that Proposition 3.3.1.1 implies Proposition 3.3.1.2, note that ± the PIT proves the compactness of XL for any Boolean algebra L as the only choice principle used in Goldblatt [13] Alexander’s Subbase Theorem, which is equivalent to PIT. ± To see that Proposition 3.3.1.2 implies 3.3.1.1, assume that XL is com- pact for all Boolean algebras L. To show PIT, it suffices [18] to prove: (1) the existence of a choice function for an arbitrary family of nonempty finite sets, and (2) the Weak Rado Selection Lemma (whose statement can be found below). For the proof of the first statement, let S := (Si)i∈I be a family of nonempty finite sets. Let L be the Boolean algebra presented by

h Si |{a ∧ b =0 | a 6= b ∈ Si,i ∈ I}i i I G∈

± ′ ′ ± ′ Consider XL . For I ⊆fin I, let FI = {u ∈ XL | ∀i ∈ I ∃a ∈ Si a ∈ u}. It can ′ ′ ± ± be shown that F := (FI )I ∈Pfin(I) is a filter basis of XL . Since XL is compact, + + F has a cluster point u . We show that f := {(i,a) | i ∈ I,a ∈ Si,a ∈ u } + is a choice function for S. Since u is a proper filter of L, at most one a ∈ Si can belong to u+ by the construction of L. This shows that f is a function. We now show that dom f = I. Let i ∈ I be arbitrary. Suppose by way of + + contradiction that Si ∩ u = ∅. Then ∁a is a neighborhood of u for a ∈ Si, + and so is U := a∈Si ∁a, which is open as Si is finite. Since u is a cluster point, U ∩ F{i} is nonempty, i.e., ∀a ∈ Si ∃u ∈ F{i} a 6∈ u, contradicting the T b definition of F{i}. b For the proof of the second statement, we will prove the Weak Rado Selection Lemma by showing the following: Suppose that for a set Λ there is a family of functions (γS)S∈Pfin(Λ) such that γS : S → {±1}. Then there is f :Λ → {±1} such that for all S ⊆fin λ there exists T ⊆ Λ with S ⊆ T and f ↾ S = γT ↾ S. + − + − To that end, let (γS)S be given. Let L = hλ , λ | λ = ¬λ iλ∈Λ. For ± S ⊆fin λ, let uS be the filter of L generated by {λ | λ ∈ Λ,γS(λ)= ±1}. It ± can be shown that uS is proper so uS ∈ XL . Consider the net (uS)S∈Pfin(Λ), ± where the indices are ordered by inclusion. Since XL is compact, the net has a cluster point u∞. Now we have

± ± ± ± ∀λ ∈ Λ ∀S ⊆fin Λ ∃T ⊇ S[u∞ ∈ λ ⇒ uT ∈ λ and u∞ ∈ ∁λ ⇒ uT ∈ ∁λ ], i.e., c c c c ± ± ∀λ ∈ Λ ∀S ⊆fin Λ ∃T ⊇ S[λ ∈ u∞ ⇐⇒ λ ∈ uT ] (3.1)

± Let f = {(λ, ±1) | λ ∈ u∞}. By a similar argument as before, f is a function Λ → {±1}. Also, by 3.1, ∀S ⊆fin Λ ∃T ⊇ Sf ↾ T = γT (a fortiori, f ↾ S = γT ↾ S).  Choice-free duality for orthocomplemented lattices 11

3.4. The representation theorem In contrast, we will demonstrate independently of Alexander’s Subbase The- orem (along with its associated nonconstructive choice-principles) that the + space XL with its associated spectral topology is a spectral space that rep- resents (up to isomorphism) the original ortholattice L. We first verify that + for every ortholattice L, the space XL gives rise to a spectral space. + Proposition 3.4.1. For every ortholattice L, the space XL is a spectral space whose specialization order 6 is given by set-theoretic inclusion. + + Proof. To see that XL is a T0 space, assume that x, y ∈ XL are such that x 6= y. If we then suppose without loss of generality that a ∈ x\y, then a ∈ x + and a 6∈ y which implies that x ∈ a and y 6∈ a where a ∈ O(XL ). + We now show that in fact a is compact for each a ∈ L, whence 1= XL + is also compact. Since by definitionb of the spaceb XL ,b sets of the form a are a + b basis for XL , it suffices to showb that if a ⊆ i∈I bi, then there exists a finite subcover. With that in mind, assume that a ⊆ b , then the principalb S i∈I i filter ↑a = {b ∈ L | a ≤ b} contains oneb of the bisb, which by the definition of S principal filters implies that a ≤ bi which meansb a ⊆ bib, so bi is itself a finite subcover. + To see that XL is a coherent space, first observeb b that by definition of + + + XL , it immediately follows that CO(XL ) forms a basis. To show that CO(XL ) + is closed under binary intersections, let U, V ∈ CO(XL ). Then, observe that for finite index sets I and K, we have U = i∈I ai and V = k∈K bk so \ U ∩ V = (ai ∩ bk)=S ai ∧ bk S b b i I,k K i I,k K ∈ [∈ ∈ [∈ and thus U ∩ V is a finite union ofb compactb open sets and therefore we have + U ∩ V ∈ CO(XL ). + To show that XL is a sober space, it will be sufficient to show that every + completely prime filter xp ⊆ O(XL ) is of the form + O + (x)= {U ∈ O(X ) | x ∈ U} XL L + for some x ∈ XL . Hence, let x be the filter in L generated by the set {a ∈ L | a ∈ xp}. The equality xp = O + (x) is can be seen by first observing that XL the inclusion O + (x) ⊆ xp is immediate by the definition of x and xp being XL a filter.b For the converse inclusion xp ⊆ O + (x), assume that ai ∈ xp. XL i∈I Since by hypothesis, x is a completely prime filter, there exists some a such p S i that ai ∈ xp, which means that ai ∈ x, hence x ∈ ai. Therefore, we haveb that + ai ∈ O + (x) so in particular, we have ai ∈ O + (x). Therefore, X is XL i∈I XL L a spectralb space. b + S + b Finally, note that since XL is a T0 space,b we have that for x, y ∈ XL , x 6⊆ y implies that x 6 y. For the converse direction, suppose x ⊆ y. Then for each basic open a, if x ∈ a i.e., a ∈ x, then a ∈ y i.e., y ∈ a, which implies that x 6 y.  b b b 12 Joseph McDonald and KentarˆoYamamoto

Now that we have seen that given an ortholattice L, spaces of the form + XL are a subclass of spectral spaces, we proceed to the promised choice-free representation theorem for ortholattices.

+ Theorem 3.4.2. Given an ortholattice L, the map •: L → COR(XL ) is an + isomorphism where COR(XL ) is an ortholattice ordered by set-theoretic in- clusion, whose operation for meet is ∩, whose operationb for orthocomplements is ∗, and whose bottom universal bound is ∅.

Proof. We first show that the mapping • is an ortholattice homomorphism + L → R(XL ), where the is known to be an ortholattice under ∗ the operations ∩, , and ∅ [13]. We firstb check that • preserves meets by demonstrating a[∧ b = a ∩ b. For the a[∧ b ⊆ a ∩ b inclusion, assume that a ∈ a[∧ b so that a ∧ b ∈ x. Then, since a ∧ b ≤ a and ab∧ b ≤ b, we have that a ∈ x and b ∈ x as x is ab filter.b Hence, we find xb∈ ab and x ∈ b, so x ∈ a ∩ b. For the a ∩ b ⊆ a[∧ b inclusion, assume that x ∈ a ∩ b. Then, a ∈ a and x ∈ b so a ∈ x and b ∈ x. Since x is a filter, we find that ab∧ b ∈ x andb so a ∈ ba[∧ b, as required.b b Hence, the function • is a homomorphismb b for ∧. b b We now verify that • preserves orthocomplements by demonstrating a⊥ = (a)∗. For the a⊥ ⊆ (a)∗ inclusion,b suppose x ∈ a⊥. Then a⊥ ∈ x which ∗ ∗ ⊥ implies that x ⊥L y for everyb y ∈ a so x ∈ (a) . For the inclusion (a) ⊆ a , ∗ supposec b that x ∈ (a)cso xb⊥L y for every y ∈ a. Let y c= ↑ a = {b ∈ L | b ≥ a} be the principal filter generated byba ∈ L. Then,b we have y =↑a ∈ a sob x ⊥Lcy. ⊥ Hence, we have thatb there exists some b ∈ L suchb that b ∈ x and b ∈ y i.e., ⊥ ⊥ a ≤ b which by Definition 2.1.2.2 implies that b ≤ a . Therefore,b we have that a⊥ ∈ x i.e., x ∈ a⊥, as required. Lastly, note that the equality 0= ∅ is obvious. To show that • isc an injection, let a,b ∈ L such that a 6= b. If a 6≤bb, then + ↑a ∈ a\b which means a 6= b. To see that the range of • is COR(XL ), suppose + n that U ∈ COR(XLb). Since U is compact open, we have that U = i=1 ai for a1b,...ab n ∈ L, thatb is, Ub is a finite union of basicb opens. Since U is also S ⊥-regular, we calculate b [n n ∗∗ ∗∗ ai = ai = U = U i=1 i=1 _  [  so U is in the of •. Since CORb is the image of an ortholattice homo- + ∗ morphism, (COR(XL ), ∩, , ∅) is an ortholattice.  b

4. The dual space of an ortholattice In describing topological spaces throughout this work, we will denote a gen- eral topological space by X = (X, T ) where X is a set and T ⊆P(X) is some topology over X. Just as in our discussion of lattices, we will often conflate Choice-free duality for orthocomplemented lattices 13 a topological space with its underlying carrier set. We proceed by character- + izing the class of spectral spaces which are homeomorphic to the space XL for some ortholattice L. 4.1. UVO-spaces The following definition is an analogue of the construction given in [4] of the + class of spectral spaces which are homeomorphic to the space XB for some Boolean algebra B. Definition 4.1.1. Let X = (X, 6, ⊥, T ) be an ordered topological space en- dowed with an orthogonal binary relation ⊥⊆ X2 and whose specialization order is 6, then X is an upper Vietoris orthospace (henceforth, a UVO-space) whenever the following conditions are satisfied:

(1) X is a T0 space; (2) COR(X) is closed under ∩ and ∗; (3) COR(X) is a basis for X (4) Every proper filter in COR(X) is of the form:

CORX (x)= {U ∈ COR(X) | ∃x ∈ X : x ∈ U}; and (5) x ⊥ y =⇒ ∃U ∈ COR(X): x ∈ U & y ∈ U ∗ Note that given a UVO-space X, the requirement that COR(X) form a basis for X implies the following analogue of the Priestly separation axiom: x 6 y =⇒ ∃U ∈ COR(X): x ∈ U & y 6∈ U. Notice that if we replace the compact open ⊥-regular subsets of X by the clopen upsets of X, then we arrive exactly at Priestley’s seperation for the dual space of a distributive lattice. Moreover, note that the fourth condition is an analogue of the sobriety condition of a spectral space. The construction which associates to each UVO-space X an ortholattice L is provided to us by the following lemma. Lemma 4.1.2. If X is a UVO-space, then L = (COR(X), ∩,∗ , ∅) is an ortho- lattice. Proof. Here, we define the joins of L by De Morgan’s distribution laws for complements over meets and set 1 = ∅∗. We first verify that COR(X) is closed under the relevant operations. Clearly ∅∈CO(X) and since ∅ = ∅∗∗, we have that ∅ ∈ COR(X). By Definition 4.1.1.2, if U ∈ COR(X) then U ∗ ∈ COR(X) and if U, V ∈ COR(X), then U ∩ V ∈ COR(X). To see that the algebra induced by COR(X) is an ortholattice, first observe that by the irreflexivity of ⊥, we have that U ∩ U ∗ = ∅ for every U ∈ COR(X). If on the other hand there was some y ∈ U ∩ U ∗, then by definition of U ∗, we would have y ∈ {x | ∀y ∈ U : x⊥y} which contradicts the fact that ⊥ is irreflexive. Hence Definition 2.1.2.1 is satisfied. Given the definition of the ∗ operator, the symmetry of ⊥ guarantees that ∗ is an order- reversing function, so Definition 2.1.2.2 is satisfied. By the ⊥-regularity of COR(X), if U ∈ COR(X), then U = U ∗∗ so Definition 2.1.2.3 is satisfied.  14 Joseph McDonald and KentarˆoYamamoto

We now must conversely verify that every ortholattice gives rise to a UVO-space.

+ + Lemma 4.1.3. If L is an ortholattice, then XL = (XL , ⊥L) is a UVO-space.

Proof. We first verify that ⊥L⊆ F(L) × F(L) is indeed an orthogonality re- lation over the proper filters of L. For irreflexivity, assume by contradiction ⊥ that there exists x ∈ F(L) such that x ⊥L x. Then, there exists a ∈ x such that a ∈ x. Since x is a filter, we have that a ∧ a⊥ ∈ x which by Defini- tion 2.1.2.1 implies that 0 ∈ x which contradicts the fact that x is a proper lattice filter over L. Therefore, ⊥L is irreflexive. For symmetry assume that ⊥ x, y ∈ F(L) are such that x ⊥L y. Then by definition, there exists a ∈ x such that a ∈ y. By Definition 2.1.2.3, we have that a⊥⊥ = a and so a⊥⊥ ∈ y ⊥ but since a ∈ x, we have that y ⊥L x by the definition of ⊥L. Hence, we conclude that ⊥L is symmetric. + We already know that XL is a T0 space from Proposition 3.4.1. Note + that by Theorem 3.4.2, if U, V ∈ COR(XL ), then U = a and V = b for some a,b ∈ L. Moreover, we saw that a ∩ b = a[∧ b and (a)∗ = a⊥ with [ + ⊥ + b b a ∧ b ∈ COR(XL ) and a ∈ COR(XL ). Since by definition, sets of the form b + c a for some a ∈ L form a basis for the spaceb XL , it follows thatb the second and third conditions arec satisfied. For the fourth condition, let x be a proper + bfilter in COR(XL ). Then y = {a ∈ L | a ∈ x} is a proper filter in L. It is easy to verify that COR + (y)= x. Finally, for the fifth condition, let x, y ∈ F(L) XL ⊥ such that x ⊥L y. Then there exists someb a ∈ L such that a ∈ x and a ∈ y. By the definition of a, we have that x ∈ a and that y ∈ a⊥, but since • is a homomorphism for ⊥, we have that y ∈ (a)∗. Again, by Theorem 3.4.2, for + U ∈ COR(XL ), web have U = a for someb a ∈ L, whichc means thatb there + ∗ exists some U ∈ COR(XL ) such that x ∈ bU and y ∈ U , as desired.  b + 4.2. The characterization theorem for XL We now proceed by demonstrating that the class of UVO-spaces provides us with the desired topological characterization of the class of spectral spaces used in our representation.

+ Theorem 4.2.1. For each UVO-space X, the map X → XCOR(X) is a home- omorphism and an isomorphism (i.e., a bijective embedding) with respect to + the orthospace reducts (X,⊥) and (XCOR(X), ⊥).

Proof. We will show that the map g : x 7→ CORX (x) gives the desired home- + omorphism from X to XCOR(X). To see that g is an injective function, let x, y ∈ X and assume that x 6= y. Since X is a T0 space, we have that either x 6 y or y 6 x. If x 6 y, then from Definition 4.1.1.3 (which, as already mentioned, implies our analogue of the Priestly separation axiom), we have that there exists some U ∈ COR(X) such that x ∈ U and y 6∈ U, which implies that U ∈ CORX (x) and U 6∈ CORX (y) so we have the desired in- equality CORX (x) 6= CORX (y). If on the other hand, we have that y 6 x, Choice-free duality for orthocomplemented lattices 15 then we similarly find that there exists some U ∈ COR(X) such that y ∈ U but x 6∈ U, which implies that CORX (x) 6= CORX (y). As the surjectivity of g is immediate from Definition 4.1.1.4, we have established that g is a bijective function. To see that g is continuous, it will suffice to demonstrate that the inverse + image of each basic open set in XCOR(X) is an open set in X. Note that each + + basic open set in XCOR(X) is of the form U for some U ∈ XCOR(X). The continuity of g can then be proved by observing the following calculation: b −1 g [U]= {x ∈ X | CORX (x) ∈ U}

= {x ∈ X | U ∈ CORX (x)} b b = {x ∈ X | x ∈ U} = U

The continuity of g−1 is established by calculating the image of g as follows:

g[U]= {CORX (x) | x ∈ U}

= {CORX (x) | U ∈ CORX (x)} b = U

Now that we have establishedb that g is a homeomorphism of topological spaces, we proceed by verifying that g is an isomorphism with respect to the orthospace reducts. Suppose for x, y ∈ X, we have g(x) ⊥ g(y). Then by the definition of g, we have that CORX (x) ⊥ CORX (y). By the definition of ⊥, this implies that there exists some U ∈ COR(X) such that U ∈ CORX (x) ∗ ∗ and U ∈ CORX (y) which means that x ∈ U and y ∈ U . By univer- sal instantiation and the definition of the ∗ operator, we have that x ⊥ y. Conversely, let x, y ∈ X and suppose that x ⊥ y. By hypothesis, X is a UVO-space and so by Definition 4.1.1.5, there exists some U ∈ COR(X) such that x ∈ U and y ∈ U ∗. By the definition of g, this means that ∗ U ∈ CORX (x) and U ∈ CORX (y). Hence, by the definition of ⊥, we have that CORX (x) ⊥ CORX (y) i.e., g(x) ⊥ g(y). 

Corollary 4.2.2. Let X be a UVO-space. Then: (1) X is a spectral space. (2) Every element in CO(X) is a finite union of elements in COR(X).

Proof. For part 1, note that by Theorem 4.2.1, we have that every UVO- + space X is homeomorphic to the space XCOR(X), which is a spectral space by Proposition 3.4.1, since COR(X) is an ortholattice whenever X is a UVO- space by Lemma 4.1.2. For part 2, let X be a UVO-space and let U ∈ CO(X). Then by Definition 4.1.1.3, U is a finite union of elements from COR(X).  16 Joseph McDonald and KentarˆoYamamoto

5. The category of UVO-spaces We now proceed by investigating the abstract category-theoretic structure underlying the constructions and results achieved in the previous two sec- tions. For an in-depth exposition of pure , refer to [1]. Definition 5.0.1. Let OrthLatt be the category whose collection of objects are given by the class of ortholattices and whose collection of morphisms are given by the class of ortholattice homomorphisms between them. It is clear that isomorphisms in the category OrthLatt are given exactly by those ortholattice homomorphisms which are isomorphisms.

5.1. UVO-mappings Just as in the categorical dual equiavelnce result in [4] between the category BoolAlg of Boolean algebras and Boolean homomorphisms and the category UV of UV-spaces and UV-mappings, our conception of an appropriately de- fined between UVO-spaces depends upon the notions of a spectral mapping and a p-morphism. Definition 5.1.1. Given spectral spaces X and X′, a map f : X → X′ is a spectral map if f −1[U] ∈ CO(X) for every U ∈ CO(X′). Example 5.1.2. If X is a spectral space and Y is a Stone space, then the map f : X → Y is a spectral map if and only if f is a continuous function. Clearly, any spectral map is a continuous function but the converse is not in general true, as can be easily seen through the following example. Example 5.1.3. Let X be an infinite Stone space and let Y be the Sierpi´nsky space; namely a topological space whose carrier set is the two-element set {0, 1} endowed with the fine lower topology over the total order 0 < 1. Since X has a non-isoloated point x ∈ X, the map f : X → Y is continuous but not spectral since ∁{x} has no finite subcover. Definition 5.1.4. Let (X, R) and (X′, R′) be Kripke frames, a function f : X → X′ is a p-morphism if: (1) xRy =⇒ f(x)R′f(y) (2) f(x)R′y′ =⇒ ∃y ∈ X : xRy & f(y)= y′ Example 5.1.5. Let (N,<) be a Kripke frame determined by a strict linear ordering over the natural numbers and let (X, R) be a Kripke frame given by X = {x} and R = {hx, xi}. Then the frame homomorphism f : (N,<) → (X, R) as defined in Figure 3 is a p-morphism. Example 5.1.6. Let (Z, R) and (Z′, R′) be Kripke frames defined by Z = {x, y} with R = {hx, xi, hy,yi, hx, yi, hy, xi} and Z′ = {x′} with R′ = {hx, xi}. Clearly, (Z, R) and (Z′, R′) are B-frames (i.e., R and R′ are reflexive and symmetric frame relations over Z and Z′ respectively) and the frame homo- morphism g : (Z, R) → (Z′, R′) as defined in Figure 3 is a p-morphism. Choice-free duality for orthocomplemented lattices 17

. . <

2 R x < g R f 1 ′ ′ R yg x R < f 0 x R f

Figure 3. p-morphisms from (N,<) to (X, R) and (Z, R) to (Z′, R′)

A p-morphism is a special case of a bisimulation between two relational structures. Refer to [7] for more details pertaining to the modal logic of p- morphisms and bisimulations. With the notions of a spectral map and p-morphism in mind, we arrive at the notion of a UVO-map. Definition 5.1.7. If X and X′ are UVO-spaces, then a map f : X → X′ is a UVO-map if f is a spectral map and p-morphism with respect to the set- theoretic complement of the orthogonality relations 6⊥ and 6⊥′ of X and X′ respectively, so it satisfies the following conditions: (1) x 6⊥ y =⇒ f(x) 6⊥′ f(y) (2) f(x) 6⊥′ y′ =⇒ ∃y ∈ X : x 6⊥ y & f(y)= y′ The above construction of a UVO-map between UVO-spaces is highly reminscent to the construction of a continuous map between two Stone spaces of an ortholattice, as defined in [5]. The p-morphism condition within Definition 5.1.7 can be viewed dia- grammatically in Figure 4.

y′ ∃y y′

′ ′ 6⊥ =⇒ 6⊥ 6⊥

x f(x) x f(x)

Figure 4. p-morphism condition of a UVO-map

It is important to notice that unlike the category UV of UV-spaces and UV-mappings, we do not require that UVO-maps be p-morphisms with respect to the specialization order of UVO-spaces. 18 Joseph McDonald and KentarˆoYamamoto

Definition 5.1.8. If X and Y are UVO-spaces, then a UVO-map f : X → Y is a homeomorphism if f is a homeomorphism as a continuous map and an isomorphism (or bijective embedding) with respect to the orthospace (or- thoframe) reducts (X, ⊥) and (Y, ⊥) of X and Y respectively. Definition 5.1.9. Let UVO be the category whose collection of objects are given by the class of UVO-spaces and whose collection of morphisms are given by the class UVO-mappings between them. Note that the isomorphisms in UVO are given exactly by those UVO- maps which are as described in Definition 5.1.8.

5.2. Basic results about UVO-mappings The following results will be useful in our proof of the categorical dual equiv- alence between OrthLatt and UVO. Proposition 5.2.1. If X and X′ are UVO-spaces and f : X → X′ is a UVO- map, then f −1[U] ∈ COR(X) for each U ∈ COR(X′). Proof. Let f +(U) be the inverse f-image of U. We see that f +(U) is ⊥- ′ ′ regular as well. Indeed, since U is ⊥′-regular, we have ′♦′U = U ⊥ ⊥ = U, where ′ and ♦′ are the modal operators corresponding to 6⊥′ [13]. Since f is p-morphic with respect to 6⊥′, the map f + preserves modal operators. Thus we have f +(U)⊥⊥ = ♦f +(U) = f +(′♦′U) = f +(U), where  and ♦ are defined from 6⊥ likewise. 

Proposition 5.2.2. If X and X′ are UVO-spaces and f : X → X′ is a map such that f −1[U] ∈ CO(X) for every U ∈ COR(X′), then f is a spectral map. Proof. Suppose that X and X′ are UVO-spaces and that f : X → X′ is n a UVO-map. Then by Corollary 4.2.2, we find that U = i=1 Ui for Ui ∈ COR(X′), which yields the following equalities: S n n −1 −1 −1 f [U]= f Ui = f [Ui] " i=1 # i=1 [ [ −1 −1 By hypothesis, we have that f [Ui] ∈ CO(X) which implies that f [U] is a finite union of compact opens and thus f is a spectral map. 

Lemma 5.2.3. Let X and X′ be spectral spaces and let f : X → X′ be a map. If for each set U in some subbasis of X′, we have f −1[U] ∈ CO(X), then f is a spectral map. Proof. By definition, every open set U ∈ O(X) is a union of finite intersec- tions of subbasic open sets so every compact open set U ∈ CO(X) is a finite n union i=1 Ui of finite intersections of subbasic sets. Then, since n n S −1 −1 −1 f [U]= f Ui = f [Ui] " i=1 # i=1 [ [ Choice-free duality for orthocomplemented lattices 19

−1 −1 it follows that f [U] ∈ CO(X) if every f [Ui] ∈ CO(X). Moreover, given m that Ui = j=1 Vj where each Vj is a subbasic set and given that T m m −1 −1 −1 f [Ui]= f Vj = f [Vj ] " j=1 # j=1 \ \ −1 −1 it similarly follows that f [Ui] ∈ CO(X) if every f [Vj ] ∈ CO(X). Finally, since by hypothesis, the inverse image of each Vj is compact open, we have that f is a spectral map, as desired. 

5.3. The main result We now proceed with the promised choice-free categorical dual equivalence result between the categories OrthLatt and UVO. Theorem 5.3.1. The category OrthLatt of ortholattices and ortholattice ho- momorphisms and the category UVO of UVO-spaces and UVO-mappings constitute a dual equivalence of categories. Proof. Let L and L′ be ortholattices and let h: L → L′ be an ortholattice + −1 homomorphism. Given x ∈ XL , define h+(x) = h [x]. Since h is an ortho- lattice homomorphism, h+(x) is a proper lattice filter in L. Hence, we have an + + induced map h+ : XL′ → XL . We want to show that h+ is a UVO-map. We first verify that h+ is a spectral map. By Lemma 5.2.3, it will suffice to show + −1 + that for each basic open a in the space XL , we have that h+ [a] ∈ CO(XL ). This is achieved by observing the following calculation: −b1 + b h+ [a]= {x ∈ XL′ | h+(x) ∈ a} + −1 = {x ∈ XL′ | h [x] ∈ a} b + −1 b = {x ∈ XL′ | a ∈ h [x]} + b = {x ∈ XL′ | h(a) ∈ x} = h(a), This is compact open. We now verify that h satisfies the p-morphism con- d + dition with respect to the complements of the orthogonality relations. To see that h+ satisfies the first condition of a p-morphism, suppose by way of con- + ′ ′ ′ tradiction that within the space XL′ , we have x 6⊥ y and that within the + ′ ′ space XL , we have h+(x ) ⊥ h+(y ). Then, there exists some a ∈ L such that ⊥ ′ ′ ′ a ∈ h+(x ) and a ∈ h+(y ). By definition of h+ it follows that h(a) ∈ x and h(a)⊥ = h(a⊥) ∈ y′, a contradiction. To see that h+ satisfies the second condition of a p-morphism, assume ′ ′ + + ′ that h+(x ) 6⊥ y for some x ∈ XL′ and y ∈ XL . Now let y be the filter ′ ′ ′ + generated by h[y]. Clearly, we have that y ∈ F(L ) and hence y ∈ XL′ with ′ ′ ′ h+(y ) = y. Now, if we assume by way of contradiction that x ⊥ y , then there exists some a′ ∈ L′ such that a′⊥ ∈ x′ and a′ ∈ y′. By definition, it then follows that there exists some a ∈ L such that h(a) ≤ a′ with a ∈ y ′ ′⊥ ⊥ ′ ⊥ ′ but this implies that a ≤ h(a ) ∈ x . This then implies that a ∈ h+(x ) 20 Joseph McDonald and KentarˆoYamamoto

′ which contradicts our hypothesis that h+(x ) 6⊥ y. Hence, each ortholattice ′ + + homomorphism h: L → L gives rise to the a UVO-map h+ : XL′ → XL . For the other direction, suppose that X and X′ are UVO-spaces and that f : X → X′ is a UVO-map. Given any U ∈ COR(X′), define f +(U ′)= f −1[U]. Note that by Proposition 5.2.1, we have that f +(U) = f −1[U] ∈ COR(X) since f is by hypothesis a UVO-map. To see that f + is an ortho- lattice homomorphism, we first verify that f +(U ∩ V ) = f +(U) ∩ f +(V ). Assume that x ∈ f +(U ∩ V ) i.e., x ∈ f −1[U ∩ V ]. Then f(x) ∈ U ∩ V so f(x) ∈ U and f(x) ∈ V . By definition of f −1, this implies that x ∈ f −1[U] and x ∈ f −1[V ] so x ∈ f −1[U] ∩ f −1[V ] i.e., x ∈ f +(U) ∩ f +(V ). Con- versely, if x ∈ f +(U) ∩ f +(V ) i.e., x ∈ f −1[U] ∩ f −1[V ], then x ∈ f −1[U] and x ∈ f −1[V ]. Hence, by definition of f −1, this implies that f(x) ∈ U and f(x) ∈ V , so f(x) ∈ U ∩ V . Therefore, x ∈ f −1[U ∩ V ] i.e., x ∈ f +(U ∩ V ). Hence, f + is a homomorphism for ∩. To see that f + is an ortholattice homomorphism for ∗ we must verify that f +(U ∗) = f +(U)∗. Assume (by contraposition) that x 6∈ f +(U)∗ i.e., x 6∈ f −1[U]∗. Then, there exists some y such that y 6⊥ x where y ∈ f −1[U]. As f is by hypothesis, a p-morphism with respect to 6⊥, it follows that f(y) 6⊥ f(x). Since we know that f(y) ∈ U as y ∈ f −1[U], we have f(x) 6∈ U ∗ so x 6∈ f −1[U ∗] i.e., x 6∈ f +(U ∗). Conversely, assume (by contraposition) that x 6∈ f +(U ∗) i.e., x 6∈ f −1[U ∗]. Then f(x) 6∈ U ∗ so by the definition of ∗ we find that there exists some y ∈ U such that f(x) 6⊥′ y. Since f is a p-morphism with respect to 6⊥, we find some z ∈ X such that x 6⊥ z and f(z) = y. It is easy to see that f(z) ∈ U and so z ∈ f −1[U] which by the definition of ∗ implies that x 6∈ f −1[U]∗ i.e., x 6∈ f +(U)∗, as required. Note that this can also be derived by translating into B as discussed in [14]. Lastly, we verify that f + is a lattice homomorphism for the bottom universal bound ∅ by verifying that f +(∅) = ∅. Assume that x ∈ f +(∅) i.e., x ∈ f −1[∅], so that f(x) ∈ ∅, a contradiction. Conversely, note that the ∅ ⊆ f +(∅) inclusion is trivial. Hence, we have established f(∅) = ∅, thereby demonstrating that every UVO-map f : X → X′ gives rise to the existence of an ortholattice homomorphism f + : COR(X′) → COR(X). Clearly (•)+ preserves identity maps and the composition structure. Hence (•)+, COR(•), along with Lemmas 4.1.3 and 4.1.2, give rise to con- travariant (•)+ : OrthLatt → UVO and COR(•): UVO → OrthLatt where (•)+ is defined on objects and morphisms by

+ ′ + + L 7→ XL , h: L → L 7→ h+ : XL′ → XL and COR(•) is defined on objects and morphisms by

X 7→ COR(X), f : X → X′ 7→ f + : COR(X′) → COR(X).

In light of Theorem 3.4.2 which established that every ortholattice L is iso- + morphic to COR(XL ), it is not difficult to verify that every ortholattice Choice-free duality for orthocomplemented lattices 21 homomorphism h: L → L′ makes the following diagram commute:

LLh ′

+ + ′ COR(XL ) + COR(XL ) (h+)

+ which implies that each component ηL : 1L(L) → COR(•) ◦ (•) (L) of the + natural transformation η : 1OrthLatt → COR(•) ◦ (•) is an isomorphism. Similarly, in light of Theorem 4.2.1, which established that every UVO- + space X is homeomorphic to XCOR(X) and order isomorphic with respect to the complements of the orthogonality relations, it is not difficult to verify that every UVO-map f : X → X′ makes the below diagram commute:

f X X′

+ + ′ XCOR(X) + XCOR(X ) (f )+

+ which implies that each component θX : 1X (X) → (•) ◦ COR(•)(X) of the natural transformation θ : 1UVO → (•)+ ◦ COR(•) is an isomorphism, which completes our proof that the contravariant functors COR(•) and (•)+ constitute a dual equivalence of categories. 

6. Duality dictionary In light of Theorem 5.3.1, we proceed by developing a “duality dictionary” (as depicted in the Figure 5) for the purposes of explicitly establishing how one can translate between various lattice-theoretic concepts (as applied to the category OrthLatt) and their corresponding dual topological concepts in the category UVO. For an analogous duality dictionary relating the category of Boolean algebras BoolAlg, the category of UV-spaces UV, and the category of Stone spaces Stone, refer to [4].

6.1. Complete lattices Definition 6.1.1. Let X be a UVO-space, then X is complete if for every open set U ∈ O(X), we have that U ∗◦∗ ∈ COR(X).1

We now verify that the classes of complete UVO-space and the duals of complete ortholattice coincide.

Proposition 6.1.2. Let L be an ortholattice and let X be its dual UVO-space. Then:

1We sometimes write (•)◦ for the interior of a set where appropriate in terms of typography. 22 Joseph McDonald and KentarˆoYamamoto

OrthLatt UVO ortholattice UVO-space homomorphism UVO-map complete UVO-space atom isolated point atomless lattice Xiso = ∅ atomic lattice Cl(Xiso)= X injective homomorphism surjective UVO-map surjective homomorphism UVO-embedding subalgebra image under UVO-map direct product UVO-sum MacNeille completion R(P(X)) canonical extension R(X)

Figure 5. Duality dictionary for Orthlatt and UVO

(1) A family {Ui}i∈I ⊆ COR(X) has a greatest lower bound in COR(X) iff Int i∈I Ui ∈ COR(X), in which case

T  Ui = Int Ui . i I i I ^∈  \∈  (2) A family {Ui}i∈I ⊆ COR(X) has a least upper bound in COR(X) iff ∗◦∗ ( i∈I Ui) ∈ COR(X), in which case ∗◦∗ S Ui = Ui . i∈I i∈I _ [  (3) L is a complete ortholattice iff X is a complete UVO-space.

Proof. For part 1, observe that Int i∈I Ui = inf({Ui}i∈I ) for {Ui}i∈I ⊆ COR(X) immediately follows from our hypothesis that Int( Ui) ∈ COR(X). T  i∈I The for left to right implication of part 1, assume that i∈I Ui is defined in COR(X). Note that by Theorem 3.4.2, for every i ∈ I,T there exists some V + ai ∈ L such that Ui = ai, and since the map •: L → COR(XL ) defined by a 7→ a is an ortholattice isomorphism, we have the following equalities: b b [b U = a = a . b i i i i I i I i I ^∈ ^∈ ^∈ Hence, it suffices to show that b [ ai = Int ai . (6.1) i I i I ^∈  \∈  \ b \ We have i∈I ai ⊆ Int i∈I ai as clearly i∈I ai ⊆ i∈I ai such \   \ that ai isV an open set. ToT see that Int Vai ⊆ Tai, suppose i∈I b i∈I i∈I b   that Vx ∈ Int ai . Then there exists someT U ∈ CORV(X) such that i∈I b  T  b Choice-free duality for orthocomplemented lattices 23

x ∈ U ⊆ i∈I ai. Hence, by Theorem 3.4.2, we have that U = b for some b ∈ L. Moreover, since b ⊆ a , it follows that b ≤ a . Then, since T i∈I i i∈I i b \ b x ∈ b, we have b ∈ x so i∈ITai ∈ x, hence x ∈ i∈I ai. V b b For part 2, we first assume that i Ui exists. Let ai ∈ L be such that V ∗◦∗ V ai =bUi. One can show that u ∈ ( ai) if and only if i W S ⊥ b ∀v ∈ X[∃b ∈ v ∀i ∈ Iab i ≤ b =⇒ u ⊥ v]. (†) | {z } \ ∗◦∗ With this in mind, let u ∈ i Ui be arbitrary. We show u ∈ ( i ai) . Take ⊥ ⊥ an arbitrary v ∈ X with (†). Since ai ≤ b for all i ∈ I, we have ai ≤ b . W S i The left-hand side of this inequality is in u by assumption, as is theb right-hand W ∗◦∗ side. Recall b ∈ v to conclude u ⊥ v. We have established u ∈ ( i ai) . To ∗◦∗ [ show the other inclusion, we prove u 6∈ ( i ai) for u 6∈ i aSi. It suffices ⊥ to exhibit v ∈ X with the properties (†) and u 6⊥ v. Let b = ( ibai) and ⊥ T W v = ↑ b. Since b = i ai, the property (†)b is satisfied. Now, since v is a W principal filter, u ⊥ v if and only if b⊥ ∈ u, and we assumed otherwise. Now W ∗◦∗ ∗◦∗ we show that if ( i ai) ∈ COR(X), then ( i ai) is the least upper bound of {Ui}i. Clearly, it is an upper bound of the family, so it suffices to S S ∗◦∗ show that if c is an upperb bound of the family, thenb ( i ai) ⊆ c. Take an arbitrary u in the left-hand side of the inequality. Let b = c⊥ and v = ↑ b. ⊥ S Since c = b b, and ai ≤ c for all i ∈ I, the property (b†) is satisfied.b We conclude u ⊥ v, i.e., c ∈ u as desired. For part 3, we start by proving the left-to-right implication. Assume L is a complete ortholattice so that for each A ⊆ L, we have that A and A are defined. If U ∈ O(X), then by Definition 3.2.2, we have that V W U = {V ∈ COR(X) | V ⊆ U} [ Since by hypothesis, L is a complete ortholattice, by Theorem 3.4.2, so is the corresponding unique (up to isomorphism) ortholattice induced by COR(X) and hence {V ⊆ COR(X) | V ⊆ U} exists. By our proof of part 2, we have

W ∗◦∗ {V ⊆ COR(X) | V ⊆ U} = {V ⊆ COR(X) | V ⊆ U} , _ [  which implies that U ∗◦∗ ∈ COR(X) as desired, making X a complete UVO- space by definition. Conversely, suppose that X is a complete UVO-space.

Then for every family of subsets {Ui}i∈I ⊆ L, we have i∈I ai ∈ O(X). Since ∗◦∗ by hypothesis, X is a complete UVO-space and so we have that a ∈ S i∈I i COR(X). By part 2, it follows that ai exists. Finally,b recall that since i∈I S  the orthocomplement operation of L is an isomorphism between L andb the order dual of L, the ortholattice L isW complete if and only if arbitrary joins exist in L. We conclude that L is complete if X is complete.  24 Joseph McDonald and KentarˆoYamamoto

6.2. Atoms Notation 6.2.1. Let L be an ortholattice and let X be a UVO-space. We write At(L) to denote the set of all atoms of L and write Xiso to denote the set of all isolated points of X. Proposition 6.2.2. Given an ortholattice L and its dual UVO-space X, the mapping At(L) → Xiso defined by a 7→↑a is a .

+ Proof. Note that if a ∈ At(L), then a = {↑a} and since a ∈ O(XL ), it follows that ↑a is an isolated point. It immediately follows that the map is injective since clearly for all a,b ∈ L, if a 6=bb then without lossb of generality, there exists some c ∈↑a such that c 6∈↑b so ↑a 6=↑b. To see that the map is a surjection, note that if x is an isolated point, then {x} is an open set and since COR(X) forms a basis for a UVO-space + X, we have that {x} ∈COR(XL ). Then by Theorem 3.4.2, there exists some a ∈ L such that a = {x} which implies that a ∈ At(L). On the other hand, if a 6∈ At(L), then there exists some 0 6= b ∈ L such that b

Proposition 6.3.2. Let L be an ortholattice and let X be its dual UVO-space. Then, the following statements are equivalent: (1) L is atomic; (2) Cl(Xiso)= X. Proof. To show the forward implication, assume that L is complete, and take u ∈ X. It suffices to show that for every basic open neighborhood U ∋ u, the subset Xiso intersects with U nontrivially. Find a ∈ L \{0} such that U = a. By atomicity, there is an atom b ≤ a, i.e., ↑ b ∈ a, and ↑ b ∈ Xiso. To show the other implication, let a ∈ L \{0} be arbitrary. We will findb an atom b ≤ a. Consider ↑ a ∈ X and a neighborhoodb a. Since Xiso is dense, a intersects nontrivially with Xiso at, say, u. Recall that u is of the form ↑ b for some atom b. Since ↑ b ∈ a, we have b ≤ a. b  b b Choice-free duality for orthocomplemented lattices 25

6.4. Injective and surjective homomorphism Definition 6.4.1. Let X and Y be UVO-spaces. A UVO-map f : X → Y is a UVO-embedding if f is injective and for every U ∈ COR(X), there exists some V ∈ COR(Y ) such that f[U]= f[X] ∩ V . Proposition 6.4.2. Let L and L′ be ortholattices, let h: L → L′ be an or- + + tholattice homomorphism, and let h+ : XL′ → XL be the corresponding dual UVO-map of h. Then, h+ is a surjective UVO-map if h is an injective or- tholattice homomorphism. Moreover, if h+ is a UVO-embedding if h is a surjective ortholattice homomorphism. Proof. For the first part, assume that h: L → L′ is an injective ortholattice + homomorphism, and let x ∈ XL and y = {b ∈ L | ∃a ∈ h[x]: a ≤ b}. Clearly, y is the inverse h-filter of x. We want to show that y is a proper filter. To see this, note that if 0′ ∈ y, then 0′ ∈ h[x] which implies the existence of some a ∈ x such that h(a)=0′. By hypothesis, x is a proper filter, which implies that a 6= 0, but this contradicts the fact that h(0) = 0′ together with our hypothesis that h is injective. Hence, it follows that h+ is a surjective UVO-map. For the second part, let x, y ⊆ L be filters such that x 6= y. Without loss of generality, there exists some a ∈ L such that a ∈ x but a 6∈ y. By hypothesis, h is a surjective ortholattice homomorphism and, therefore, there exists some b ∈ L such that h(b) = a. It is easy to see that b ∈ h−1[x] and −1 −1 −1 b 6∈ h [y], which implies that h [x] 6= h [y]. Hence, h+ is an injective UVO-map. To see that h+ satisfies the UVO-embedding condition, first take an + arbitrary U ∈ COR(XL′ ). Note that by Theorem 3.4.2, U is of the form a for some a ∈ L′. Again, by our hypothesis that h is a surjective homomorphism, there exists some b ∈ L such that h(b)= a. It suffices to show that h+[ba]= + ′ h+[XL′ ] ∩ b. To show the left-to-right inclusion, take an arbitrary x ∈ h+[a], i.e., x = h−1[x′] for some x ∈ a. We now have b ∈ x, and we are done. Tob see + ′ + the other inclusion,b take an arbitrary x ∈ h+[XL′ ] ∩ b. There exists x ∈ XbL′ ′ −1 ′ for which x = h+(x ), i.e., x =bh [x ]. Since b ∈ x as well, we conclude that ′ x ∈ a. This shows that x ∈ h+[a]. b  Proposition 6.4.3. Let X and X′ be UVO-spaces, let f : X → X′ be a UVO- map,b and let f + : COR(X′) →b COR(X) be the corresponding ortholattice homomorphism dual to f. Then, f + is an injective ortholattice homomor- phism if f is a surjective UVO-map. Moreover, the map f + is a surjective ortholattice homomorphism if f is a UVO-embedding. Proof. For the first part, let X and X′ be UVO-spaces, and let f : X → X′ be a surjective UVO-map. Now suppose that U, V ∈ COR(X) are such that U 6= V . Without loss of generality, if y ∈ U \ V , then since f is surjective, there exists some x ∈ X such that f(x) = y so x ∈ f −1[U] and x 6∈ f −1[V ]. Since f −1[U] = f +(U) and f −1[V ] = f +(V ), we have f +(U) 6= f +(V ). Hence, f + is an injective ortholattice homomorphism. 26 Joseph McDonald and KentarˆoYamamoto

For the second part, let X and X′ be UVO-spaces and let f : X → X′ be a UVO-embedding. If U ∈ COR(X), then since f is a UVO-embedding, by Definition 6.4.1, there exists some V ∈ COR(X′) such that f[U]= f[X] ∩ V , which implies that f −1[f[U]] = f −1[f[X] ∩ V ]. Now observe that f −1[f[X] ∩ V ]= f −1[f[X]] ∩ f −1[V ]= X ∩ f −1[V ]= f −1[V ] By hypothesis, f is a UVO-embedding and therefore injective, which guaran- tees that f −1[f[U]] = U so f −1[V ]= U and since f −1[V ]= f +(V ), we have f +(V )= U, as desired.  6.5. Subalgebra Corollary 6.5.1. Let L be an ortholattice and let X be its dual UVO-space. Then, there exists a one-to-one correspondence between the subalgebras of L and the images via surjective UVO-maps of X. Proof. The result follows immediately by Theorem 5.3.1, the first part of Proposition 6.4.2 (i.e., that h+ is a surjective UVO-map if its dual ortholattice homomorphism h is injective), and the first part of Proposition 6.4.3 (i.e., that f + is an injective ortholattice homomorphism if its dual UVO-map f is surjective).  6.6. Direct product Definition 6.6.1. If X and Y are UVO-spaces, then their UVO-sum X + Y is the space whose underlying carrier set is of the following shape X + Y := X ∪ Y ∪ (X × Y ) and whose topology is generated by sets of the form U ∪ V ∪ (U × V ) for U ∈ COR(X) and V ∈ COR(Y ), together with the orthogonality relation ⊥X+Y , which is defined as the symmetric closure of:

⊥X ∪⊥Y ∪ (X × Y ) ′ ′ ′ ′ ∪ {hhx, yi, x i | x ⊥X x } ∪ {hhx, yi,y i | y ⊥Y y } ′ ′ ′ ′ ∪ {hx, yi, hx ,y i | x ⊥X x ,y ⊥Y y }. Proposition 6.6.2. Let X and Y be UVO-spaces whose specialization orders are 6X and 6Y respectively. Then, the specialization order 6X+Y of their UVO-sum X + Y is given by: ′ ′ ′ ′ Ω6 :=6X ∪ 6Y ∪{hhx, yi, x i | x 6X x } ∪ {hhx, yi,y i | y 6Y y }∪ ′ ′ ′ ′ {hx, yi, hx ,y i | x 6X x ,y 6Y y } ′ Proof. Assume that hz,z i ∈ Ω6 such that z ∈ W = U ∪ V ∪ (U × V ) ∈ O(X+Y ) for U ∈ COR(X) and V ∈ COR(Y ). We want to show that z′ ∈ W . ′ ′ In the case when z 6X z , we have z ∈ U ∈ COR(X) so z ∈ U ∈ COR(X), hence z′ ∈ W . In the case when z = hx, yi, we have hx, yi ∈ U × V with ′ x ∈ U ∈ COR(X) and y ∈ V ∈ COR(Y ). Thus, if x 6X z , it follows that ′ ′ ′ z ∈ U ∈ COR(X) and therefore, z ∈ W . The proof of the case for z 6Y z ′ ′ ′ ′ ′ and the case for z = hx ,y i, x 6X x , and y 6Y y run analogously, as does ′ the converse direction under the assumption that hz,z i 6∈ Ω6.  Choice-free duality for orthocomplemented lattices 27

The following result follows from Theorem 4.2.1 and is useful in di- agrammatically presenting examples of finite UVO-spaces in terms of the specialization ordering of their points. Proposition 6.6.3. If X is a finite UVO-space, then (X, 6) can be constructed from an ortholattice by deleting its bottom universal bound and taking its order-theoretic dual. Proof. First note that by Lemma 4.1.2, if X is a finite UVO-space, then COR(X) is a finite ortholattice. By Theorem 4.2.1, X is homeomorphic to + the space XCOR(X) of proper lattice filters of COR(X), which is a T0 space by + Lemma 4.1.3. Hence, (X, 6) is order isomorphic to the poset (XCOR(X), ⊆) of proper lattice filters of COR(X) ordered by set-theoretic inclusion. Since any filter of a finite ortholattice is a principal filter, we have that (X, 6) is isomorphic with respect to the ortholattice of proper principal filters of COR(X), which is isomorphic to the lattice (COR(X) \ {∅}, ⊇). 

Example 6.6.4. Consider the ortholattice O2 depicted in Figure 6 and the + + ortholattice M3 in Figure 2. Their respective UVO-spaces XO2 and XM3 are depicted in Figure 6, the direct product O2 × M3 in Figure 7, and the UVO- + + sum XO2 + XM3 in Figure 7. For the purpose of visualizing these examples, we represent the specialization order between two points of a UVO-space by a dotted line and the underlying partial ordering of an ortholattice by a solid line. We omit the orthogonality relations of these UVO-spaces from these diagrams for the sake of simplicity, but still explicitly define them below.

⊥x+ = {hy1,y2i, hy2,y3i, hy1,y3i}, ⊥X+ = ∅ M3 O2

⊥X+ +X+ = {⊥X+ , ⊥X+ , hhz,y1i,y2i, hhz,y1i,y3i, O2 M3 M3 O2

hhz,y2i,y1i, hhz,y2i,y3i, hhz,y3i,y1i, hhz,y3i,y2i}

1 y1 y2 y3

0 x z

Figure 6. + + The lattice O2 and the UVO-spaces XM3 and XO2

′ + + Proposition 6.6.5. If L and L are ortholattices and XL and XL′ are their + respective dual UVO-spaces, then there is a homeomorphism f : XL×L′ → + + XL + XL′ that is an isomorphism with respect to their orthospace reducts. + ′ Proof. For each point x ∈ XL×L′ i.e., every proper filter x ∈ F(L × L ), let ′ ′ xL = {a ∈ L | ∃b ∈ L : ha,bi ∈ x}, xL′ = {b ∈ L | ∃a ∈ L : ha,bi ∈ x}, 28 Joseph McDonald and KentarˆoYamamoto

h1, 1i

h1,ai h1,bi h1,ci h0, 1i z y1 y2 y3

h1, 0i h0,ai h0,bi h0,ci hz,y1i hz,y2i hz,y3i x

h0, 0i hz, xi

Figure 7. The direct product ortholattice O2 × M3 and its + + dual UVO-sum XO2 + XM3

′ Clearly, we have that xL ∈ F(L) and xL′ ∈ F(L ). Now define f by

xL if xL′ is improper, f(x)= xL′ if xL is improper, hxL, xL′ i otherwise.

The injectivity of f follows easily from the fact that x = xL × xL′ +  for every filter x ∈ XL×L′ . To see that f is a surjective function, let y ∈ + + + XL + XL′ . In the case when y ∈ XL , we have that for the proper filter ′ + ′ x := y × L ∈ XL×L′ , it follows that y = xL and that xL is improper. + Therefore, we find that f(x) = y. The proof for the case when y ∈ XL′ L + L′ + runs analogously. Lastly, for y ∈ XL and y ∈ XL′ , in the case when L L′ L L′ L L L′ L′ y = hy ,y i, since (y × y )L = y and (y × y )L′ = y , it is easy to L L′ + L L′ see that y × y ∈ XL×L′ where f(y × y )= y. Hence, f is a bijection. We now verify that f is a continuous function. First observe that by + + Definition 6.6.1, each basic open set within XL + XL′ is of the following + + shape U ∪ V ∪ (U × V ) for U ∈ COR(XL ) and V ∈ COR(XL′ ). By Theorem + 3.4.2, each U ∈ COR(XL ) is of the form a for some a ∈ L, and so U ∪ V ∪ (U × V )= a ∪ b ∪ (a × b) b for a ∈ L and b ∈ L′. We now verify that the inverse image of each basic b + b b b open set is a union of basic open sets in XL×L′ by the following calculation: [ [ [ f −1[a ∪ b ∪ (a × b)] = f −1[a] ∪ f −1[b] ∪ f −1[a × b]= ha, 0i ∪ h0,bi ∪ ha,bi Hence, f −1[a ∪ b ∪ (a × b)] can be written as the union of basic open sets in b b + b b b b b b −1 the space XL×L′ , so f is a continuous function. To see that its inverse f b b b b Choice-free duality for orthocomplemented lattices 29

+ is a continuous function, note that for each basic open set ha,bi ∈ XL×L,

[ ′ ha,bi = {x ∈ F(L × L ) | ha,bi ∈ x, xL′ is improper} ′ ∪{x ∈ F(L × L ) | ha,bi ∈ x, xL is improper} ′ ′ ∪{x ∈ F(L × L ) | ha,bi ∈ x : xL ∈ F(L), xL′ ∈ F(L )}

[ [ which implies that f[ha,bi]= a ∪ b ∪ (a × b) so that f[ha,bi] is basic open in + + the space XL + XL′ , as required. Finally, we show that f isb an isomorphismb b b with respect to the orthospace reducts. Let ⊥s and ⊥ be the orthogonality relations of the codomain and the domain of f, respectively. The preceding argument shows that the inverse map f −1 of f is given by f −1(x)= x × L′, f −1(y)= L × y, and f −1(x, y)= + + + + x × y, where x ∈ XL and y ∈ XL′ . Let u, v ∈ XL + XL′ . An argument −1 −1 showing that u ⊥s v if and only if f (u) ⊥ f (v) involves a case analysis + + + + based on whether u and v belong to XL , XL′ , or XL × XL′ . We present an + ′ + + argument for the case u ∈ XL and v = hw, w i ∈ XL × XL′ as the other cases can be handled in similar ways. By the definition of ⊥s, we have that ⊥ u ⊥s v if and only if there exists a ∈ w such that a ∈ u. On the other hand,

f −1(u) ⊥ f −1(v) ⇐⇒ u × L′ ⊥ w × w′ ⇐⇒ ∃ha,a′i ∈ w × w′ : ha⊥,a′⊥i ∈ u × L′ ⇐⇒ ∃a ∈ w : a⊥ ∈ L, proving the claim for this particular case. 

Corollary 6.6.6. If X and Y are UVO-spaces, then their UVO-sum X +Y is a UVO-space. Moreover, the mapping f : COR(X + Y ) → COR(X) ×COR(Y ) is an ortholattice isomorphism.

+ + Proof. Clearly, by Theorem 4.2.1, the maps X → XCOR(X) and Y → XCOR(Y ) are homeomorphisms (and isomorphisms with respect to ⊥) and thus there is + + a homeomorphism X +Y → XCOR(X) +XCOR(Y ). Then by Proposition 6.6.5, + + + we find that XCOR(X) + XCOR(Y ) → XCOR(X)×COR(Y ) is a homeomorphism. The above homeomorphisms are sufficient in establishing the fact that the UVO-sum X +Y is a UVO-space if X and Y are UVO-spaces. For the second part, simply apply Theorem 5.3.1 and Proposition 6.6.5. 

It is easy to check that every UVO-sum X + Y comes equipped with canonical coprojections κ1 : X → X + Y and κ2 : Y → X + Y satisfying the universal mapping property for categorical coproducts that for any UVO- space Z and pair of UVO-maps f : X → Z and g : Y → Z, there exists a 30 Joseph McDonald and KentarˆoYamamoto unique UVO-map hf,gi: X +Y → Z making the following diagram commute:

Z

g f hf,gi

Xκ1 X + Yκ2 Y Hence, given any two UVO-spaces X and Y , their UVO-sum X + Y is a co- product in the category UVO. As a consequence, it follows that if X1,...,Xn are UVO-spaces, then X1 + ··· + Xn is a UVO-space.

6.7. Lattice completions Notation 6.7.1. Let L be a lattice and let A ⊆ L. Then: (1) Au is the collection of upper bounds of A, i.e., Au = {a ∈ L | ∀b ∈ A : b ≤ a} (2) Al is the collection of lower bounds of A, i.e., Al = {a ∈ L | ∀b ∈ A : a ≤ b} Definition 6.7.2. Given a lattice L, a subset A ⊆ L is normal iff A = Aul. We denote the collection of all normal subsets of L by Norm(L). It was demonstrated by MacLaren in [21] that the MacNeille completion of an ortholattice L can be constructed from Norm(L). The uniqueness of this construction follows from Banaschewski in [2]. We call a point u of a UVO- space X principal if there exists an open neighborhood U of u such that v 6∈ U for every v 6 u distinct from u. Proposition 6.7.3. Let L be an ortholattice and X its dual UVO-space. A point in X is principal in the sense above if and only if it is a principal filter. Proof. It is clear that if u ∈ X is a principal filter, then it is principal in the sense above. Suppose that u ∈ X is principal in our sense. Take a neigh- borhood U of u as in the definition of principality and then a basic open set a such that u ∈ a ⊆ U. Let v be the principal filter generated by a. Assume by way of contradiction that u is not a principal filter. Then, v 6 u andbv 6= u. By principality,b we have v 6∈ U and a fortiori v 6∈ a, which is a contradiction.  b For a UVO-space X, let P(X) be the orthoframe of principal points of X with the induced orthogonality relation. We then have the following UVO-space translation of the MacNeille completion of an ortholattice. Theorem 6.7.4. Let L be an ortholattice and let X be its dual UVO-space. Then, the lattice R(P(X)) is (up to isomorphism) the MacNeille completion of L. Choice-free duality for orthocomplemented lattices 31

Proof. MacLaren [21] showed that the MacNeille completion of L is isomor- phic to R(L, ‹), where ‹ is a binary relation on (the domain of) L defined by a ‹ b ⇐⇒ a ≤ b⊥. We see that R(L, ‹) is isomorphic to R(L−, ‹), where (L−, ‹) is the relational substructure of L with the domain L− = L \{0}. To see this, first observe that 0 ‹ a for all a ∈ L and that 0 ∈ U for all U ∈ R(L, ‹). From this, it follows that U 7→ U \{0} for U ∈ R(L, ‹) is the desired isomorphism. It now suffices to show that (L−, ‹) is isomorphic to P(X) = (P(X), ⊥). To see this, first note that for an arbitrary c ∈ L and u ∈ X, we have u ⊥↑ c, where ↑ c is the principal filter generated by c, if and only if c⊥ ∈ u. Hence, (↑ a) ⊥ (↑ b) if and only if b⊥ ∈↑ a, i.e., b⊥ ≥ a.  We proceed by contemplating the canonical extension of an ortholattice, as defined by Bruns and Harding [8, Theorem 8.7]within its dual UVO-space. Theorem 6.7.5. Let L be an ortholattice and let X be its dual UVO-space. Then L′ := R(X) is (up to isomorphism) the canonical extension of L.

+ ∗∗ ′ Proof. For u ∈ XL , the set {u} ∈ L is a meet of elements of L: u = {a | a ⊇ {u}∗∗}. The inclusion ⊆ is clear. To show ⊇, take v 6∈ {u}∗∗. If u ≤ v, then {u}∗ ⊆ {v}∗, so v ∈ {u}∗∗; hence, u 6≤ v. Take a ∈ u \ v; V ′ thenb ubis in a, but v is not (• denotes the embedding L → L ). Note that u = {a | a ⊇ {u}∗∗}. We have seen that {u}∗∗ ∈ L′ is a meet of elements ′ of L. We nowb show that everyb element of L is a meet of joins of elements of T ′ ∗ ∗ L. In particular,b b we claim that for Y ∈ L we have Y = {{u} | Y ⊇{u} }. The inclusion ⊆ is clear. To show the inclusion in the other direction, we show the contrapositive: if {u}∗ ⊇ Y =⇒ {u}∗ ∋ v forW every u, then v ∈ Y . Assume the hypothesis; we show v ⊥ Y ∗. Take an arbitrary u ∈ Y ∗. Then {u}∗ ⊇ Y , so we have {u}∗ ∋ v, i.e., u ⊥ v. Lastly, we verify that the embedding L → L′ is compact. Assume that

ai ⊆ bj (6.2) i j ^ _ b for families A := {ai | i ∈ I},B b:= {bj | j ∈ J} of elements of L. We show that there exists finite subfamilies A′,B′ of A, B, respectively, such that A ≤ B. Note that the right-hand side of formula (6.2) is equal to ( b )∗∗. Now let u be the filter generated by A. Assume that u is improper. j Vj W Then there exists a finite A′ ⊆ A such that 0 = A′. Now we let B′ = ∅, and S we areb done. Suppose, therefore, that u is proper. By construction, u is in the left-hand side of formula (6.2) and thus in theV right-hand side. Observe ∗∗

u ∈ bj ⇐⇒ ∀v[∀w[∃jbj ∈ w =⇒ w ⊥ v] =⇒ u ⊥ v], (6.3)   j [ where the variablesb  v and w range over X. Let v be the filter generated by ∗ ′ {bj | j ∈ J}. Assume that v is improper. Then for some finite B ⊆ B, we have 0 = {b∗ | b ∈ B′}, i.e., 1 = B′. Therefore, by a similar reasoning as before, we may assume that v is proper. By formula (6.3), u ⊥ v. By V W 32 Joseph McDonald and KentarˆoYamamoto definition, there exists c ∈ u such that c⊥ ∈ v. By the construction of u and v, we have A′ ≤ c and c⊥ ≥ {b⊥ | b ∈ B′}. The latter implies c ≤ B′, so we have A′ ≤ B′ as desired.  V V W V W 6.8. Homomorphic images of orthomodular lattices We conclude this section by characterizing the notion of homomorphic images as applied to an orthomodular lattice, in UVO-spaces. We leave the character- ization of homomorphic images as applied to ortholattices (the more general case) as an open problem. Recall that a subset S′ of a relational structure (S, R) where R is binary is an inner substructure, or a generated subframe (S, R), if y ∈ S′ whenever x ∈ S′ and xRy. For the remainder of this subsection, upsets simpliciter mean sets upward closed with respect to the specialization order 6.

Proposition 6.8.1. Let L be an orthomodular lattice and X be its dual UVO- space. Let C(L) be the set of congruences on L and PUGS(L) the set of principal upsets of X that are generated subframes of (X, 6⊥). Then there is a one-to-one correspondence between C(L) and PUGS(L).

Proof. For θ ∈ C(L), it is well known that [1]θ is a filter. Let f(θ) = ⇑[1]θ, where ⇑ u for u ∈ X is the principal upset generated by u. We see that f(θ) ∈ PUGS(L) and that f is a map C(L) → PUGS(L). Indeed, it suffices to show that f(θ) is a generated subframe with respect to the complement of the orthogonality relation of X. Consider the canonical surjection π : L ։ L/θ. The dual map π+ is a UVO-map and a fortiori a homeomorphism onto a subspace of X. We claim that ran π+, the range of π+, is f(θ), whence it follows that f(θ) is a generated subframe as π+ is p-morphic with respect to the complement of the orthogonality relation. To see that ran π+ = f(θ), first recall that u ∈ ran π if and only if there exists u′ ∈ F(L/θ) such that −1 ′ ′ ′ + π [u ]= u. For every u ∈ F(L/θ), we have [1]θ ∈ u . Hence, if u ∈ ran π , ′ ′ then [1]θ ⊆ u. Conversely, if [1]θ ⊆ u, assume a ∈ u and (a,a ) ∈ θ for a,a ∈ L. We show that a′ ∈ u, i.e., u ∈ ran π+. Let → be the so-called Sasaki hook, i.e., x→y := x⊥ ∨(y∧x) (see, e.g., [23]). We have π(a→a′)= π(a)→π(a′)=1 ′ ′ by assumption. Therefore, a → a ∈ [1]θ ⊆ u. Since a ∧ (a → a ) ∈ u, we have a′ ∈ u as well. For S ∈ PUGS(X), let g(S) = {(a,b) ∈ L2 | a ∩ S = b ∩ S}. We show that g(S) is a congruence on L and that g is a map PUGS(X) → C(X). It ⊥ suffices to show that g(S) respects ∧ and (•) . Theb formerb case is evident. For the latter goal, it suffices to show that for a,b ∈ L if a ∩ S = b ∩ S, then a⊥ ∩ S = b⊥ ∩ S. This can be proved by the translation into the modal logic KTB. b b c It isc easy to show that f and g are the inverses of each other by noting

[1]g(⇑ u) = {a ∈ L | a ∩⇑ u = ⇑ u} = {a | a ⊆⇑ u} = {a | u ∈ a} = u, which completes the proof.  b b b Choice-free duality for orthocomplemented lattices 33

7. Future work We intend to investigate the following applications and related themes of the results and constructions achieved in this work: (1) Characterize the subclass of UVO-spaces which arise as the choice-free dual spaces of the modular and orthomodular lattices. (2) Develop a theory of topological models based on UVO-spaces for which various non-classical logics (e.g. orthologic and quantum logic) are com- plete. (Since the algebraic model for quantum logics of a finite dimen- sional Hilbert space is a modular lattice and the algebraic model for quantum logics of an infinite dimensional Hilbert space is an orthomod- ular lattice, the open problem of characterizing the dual UVO-spaces of the modular and orthomodular lattices must be accomplished before this can be fully addressed). (3) Investigate the connections between lattices of varieties of ortholattices and lattices of varieties of modal algebras corresponding to KTB (the normal modal logic of reflexive symmetric Kripke frames) and its vari- ants. As already mentioned, such frames can be seen as arising by tak- ing the set-theoretic complement of the orthogonality relation of an orthospace. We believe Goldblatt in [14] develops the first step in this direction.

Acknowledgements Both researchers thank their respective supervisors, Dr. Nick Bezhanishvili and Professor Wesley Holliday, for their expertise and guidance throughout the preparation of this paper. We also thank Dr. Tommaso Moraschini, as well as the Algebra/Coalgebra Seminar at the Institute for Logic, Language, and Computation at the University of Amsterdam for their useful comments and suggestions.

References [1] Awodey, S.: Category Theory, 2nd edn. Oxford Universtiy Press, Oxford (2006) [2] Banaschewski, B.: H¨ullensysteme und Erweiterungen von Quasi-Ordnungen. Zeitschrift fur Mathematische Logik und Grundlagen der Mathematik. 2, 35– 46 (1956) [3] Bell, J.: Orthospaces and quantum logic. Foundations of Physics. 15, 1179–1202 (1985) [4] Bezhanishvili, N., Holliday W.: Choice-free Stone duality. Journal of Symbolic Logic. 85, 109–148 (2020) [5] Bimb´o, K.: Functorial duality for ortholattices and De Morgan lattices. Logica Universalis. 1, 311–333 (2007) [6] Birkhoff, G., Von Neumann, J.: The logic of quantum mechanics. Annals of Mathematics. 37, 823–843 (1936) 34 Joseph McDonald and KentarˆoYamamoto

[7] Blackburn, P., De Rijke. M., Venema. Y.: Modal Logic. Cambridge University Press, Cambridge (2001) [8] Bruns G., Harding J.: Algebraic aspects of orthomodular lattices. In: Coecke, B., Moore, D., Wilce, A. (eds.) Current Research in Operational Quantum Logic. Fundamental Theories of Physics, vol. 111, 37–65 (2000) [9] Dickmann, M., Tressl, M., Schwartz, N.: Spectral Spaces. Cambridge University Press, Cambridge (2019) [10] Davey, B., Priestly, H.: Introduction to Lattices and Order, 2nd edn. Cambridge University Press, Cambridge (2002) [11] Esakia, L.: Topological Kripke models. Soviet Math. Dokl. 15, 147-151 (1974) [12] Esakia, L.: Heyting algebras, duality theory. Trends in Logic. Ed. Bezhanishvili G, Holliday W. (2019) [13] Goldblatt, R., The Stone space of an ortholattice. Bulletin of the London Math- ematical Society. 7, 45–48 (1975) [14] Goldblatt, R.I., A semantic analysis of orthologic. Journal of Philosophical Logic. 3, 19–35 (1974) [15] Herrlich, The Axiom of Choice. Lecture Notes in Mathematics Vol. 76, Springer, Berlin (2006) [16] Hochster, M.: Prime ideal structure in commutative rings. Transactions of the American Mathematical Society. 142, 43–60 (1969) [17] Holland, S.S., The current interest in orthomodular lattices. In: Hooker, C.A. (ed.) The Logico-Algebraic Approach to Quantum Mechanics. The University of Western Ontario Series in Philosophy of Science, vol. 1, 437-496 (1975) [18] Howard, P.: Variations of Rado’s Lemma. Quarterly 39 (1) 353–356 (1993) [19] J´onsson, B., Tarski, A.: Boolean algebras with operators, part 1. American Journal of Mathematics. 73, 891–939 (1951) [20] Kalmbach, G.: Orthomodular Lattices. Academic Press. London (1983) [21] MacLaren, M.: Atomic orthocomplemented lattices. Pacific Journal of Mathe- matics. 14, 597–612 (1964) [22] Moshier, M., Jipsen, P.: Topological duality and lattice expansions, I: a topo- logical construction of canonical extensions. Algebra Universalis. 71, 109–126 (2014) [23] Paviˇci´c, M.: Minimal quantum logic with merged implications. International Journal of Theoretical Physics. 26, 845–852 (1987) [24] Priestly, H.: Representation of distributive lattices by means of ordered Stone spaces. Bulletin of the London Mathematical Society. 2, 186–190 (1970) [25] Rubin, H., Scott, D.: Some topological theorems equivalent to the Boolean prime ideal theorem. Bulletin of the American Mathematical Society. 60, 389 (1954) [26] Stone, M.: The theory of representation for Boolean algebras. Transactions of the American Mathematical Society. 40, 37–111 (1936) [27] Stone, M.: Topological representations of distributive lattices and Brouwerian logics. Casopisˇ pro Pˇestov´an´ıMatematiky a Fysiky. 67, 1–25 (1937) Choice-free duality for orthocomplemented lattices 35

[28] Tarski, A.: Uber¨ additive und multiplikative Mengenk¨orper und Mengenfunk- tionen. Sprawozdania z Posiedze´nTowarzystwa Naukowego Warszawskiego, Wydzia l III, Nauk Matematyczno-fizycznych. 30, 151–181 (1937) [29] Tarski, A.: Der Aussagenkalku¨ul und die Topologie. Fundamenta Mathemati- cae. 31, 103–134 (1938) [30] Vietoris, L.: Bereiche zweiter Ordnung. Monatshefte f¨ur Mathematik und Physik. 32, 1151–1169 (1922)

Joseph McDonald Institute for Logic, Language, and Computation University of Amsterdam North Holland 1098 XG Netherlands e-mail: [email protected] KentarˆoYamamoto∗ Group in Logic and Methodology of Science University of California at Berkeley California 94720-3840 U.S.A. e-mail: [email protected]