o⊑_obs\IfValueT#1^#1 \NewDocumentCommand\satleo≤_sat\IfValueT#1^#1 \NewDocumentCommand\behleo⊑_beh\IfValueT#1^#1 \NewDocumentCommand\opensO \NewDocumentCommand\JonmJon: #1 \NewDocumentCommand\LYmLY: #1 \NewDocumentCommand\AxiomSQCISQCI \NewDocumentCommand\PrintAxiomSQCI
Axiom (\AxiomSQCI).
is stably quasi-coherent, i.e. every finitary quotient of is quasi-coherent.
SQCP \NewDocumentCommand\PrintAxiomSQCP
Axiom (\AxiomSQCP).
The polynomial -algebra is quasi-coherent.
NT \NewDocumentCommand\PrintAxiomNT
Axiom (\AxiomNT).
We have in .
L \NewDocumentCommand\PrintAxiomL
Axiom (\AxiomL).
is local, i.e. , and for .
cL \NewDocumentCommand\PrintAxiomCL
Axiom (\AxiomCL).
is colocal, i.e. and for all .
SL \NewDocumentCommand\PrintAxiomSL
Axiom (\AxiomSL).
is a strict linear order, i.e. and for all .
1cS \NewDocumentCommand\PrintAxiomOneCS
Axiom (\AxiomOneCS).
is 1-coskeletal, i.e. (\AxiomSL) holds and for all .
SQCF \NewDocumentCommand\PrintAxiomSQCF
Axiom (\AxiomSQCF).
All finitely presented -algebras are quasi-coherent, or (equivalently) all finitely generated free -algebras are stably quasi-coherent.
SQCC \NewDocumentCommand\PrintAxiomSQCC
Axiom (\AxiomSQCC).
All countably presented -algebras are quasi-coherent.
Domains and classifying topoi
Abstract.
We explore a new connection between synthetic ___domain theory and Grothendieck topoi related to the distributive lattice classifier. In particular, all the axioms of synthetic ___domain theory (including the inductive fixed point object and the chain completeness of the dominance) emanate from a countable version of the synthetic quasi-coherence principle that has emerged as a central feature in the unification of synthetic algebraic geometry, synthetic Stone duality, and synthetic category theory. The duality between quasi-coherent algebras and affine spaces in a topos with a distributive lattice object provides a new set of techniques for reasoning synthetically about ___domain-like structures, and reveals a broad class of (higher) sheaf models for synthetic ___domain theory.
Contents
- 1 Introduction
- 2 Quasi-coherence and affine spaces
- 3 Open propositions and the dominance property
- 4 Partial map classifier
- 5 Distributive lattices and locality
- 6 Order-theoretic structure on algebras and spaces
- 7 Phoa’s principle, finitary quasi-coherence, and homotopy
- 8 Chain completeness and infinitary ___domain theory
- 9 Local properties for the interval
- 10 New models for synthetic ___domain theory
- 11 Future directions
1. Introduction
1.1. Synthetic ___domain theory
The proposal to develop a synthetic theory of domains was raised by Dana Scott in the early 1980s. The thesis is that (pre)domains should be viewed simply as some special sets in a suitable universe, and any function between them should automatically respect the computational data associated to domains. Later Hyland, [16] gave an extensive list of the properties such a universe might satisfy.
Firstly, a universe for synthetic domains should contain an interval object , which induces an observational preorder (or specialisation preorder from a topological perspective) on domains for which function is automatically monotone. An internal axiom that characterises the synthetic nature of this fact is the so-called Phoa principle: The function space should classify the order on . In other words, the functions from to itself are completely determined by the order structure on .
Additionally, should also form a subuniverse of propositions closed under dependent sums, so as to form a dominance in the sense of Rosolini, [26]. The dominance structure is used to parametrise partial functions through a partial map classifier functor constructed from .
Besides these elementary axioms, the main additional axiom of synthetic ___domain theory is the chain completeness of the interval:
Let and be the carriers of the internal initial algebra and final coalgebra to the partial map classifier functor respectively. There is a canonical inclusion , and the chain completeness axiom states that is right orthogonal to this inclusion in the sense that the canonically induced restriction map should be an equivalence.
This can be viewed as the synthetic version of the -completeness of the Sierpiński space in traditional ___domain theory.
1.2. Two viewpoints on synthetic domains
The most studied models of synthetic ___domain theory are those arising from realisability—in which the interval is taken to be Rosolini’s dominance of semidecidable propositions in the Effective Topos [16, 23], which happens to coincide there with the initial -frame by virtue of countable choice. The model of synthetic ___domain theory in ultimately interprets ___domain theoretic concepts in terms of Turing machines, which makes them applicable to the classical theory of computation [3].
On the other hand, realisability models do not provide a direct connection to classical theories of domains. For the latter, researchers returned to the original suggestion of Scott and devised sheaf models of synthetic ___domain theory in which the site is obtained from some existing category of domains or a dense generator thereof [12, 13, 14].
Much effort has been put toward identifying a common axiomatics that accounts for both realisability and sheaf models of synthetic ___domain theory; cf. the work of Reus and Streicher, [24], Simpson, [27]. Beyond these basic axioms, the realisability and sheaf models of synthetic ___domain theory behave very differently, as exemplified in the failure of the initial -algebra to be inductive in the former, as van Oosten and Simpson, [32] pointed out.
The goal of the present paper is to identify a number of axioms emerging from the theory of classifying topoi that suffice to explain the behaviour of sheaf models of synthetic ___domain theory all at once, including not only the dominance property and Phoa’s principle, but also the chain completeness of the interval and the inductive construction of the initial -algebra. We stress that the strongest of these axioms cannot hold in the realisability model of synthetic ___domain theory, but they can be used to generate new and exotic sheaf models of synthetic domains.
1.3. Classifying topoi and Phoa’s principle
As a preliminary observation, there is a connection between Phoa’s principle and classifying topoi (cf. Gratzer et al., [15, Lem 3.8]). A posteriori, the technical elements required for this observation are already contained in the work of Blass, [5].
To recap, let be a (finitary) Horn theory. Recall that the classifying topos of is given by the following presheaf category (cf. Johnstone, [18, D3.1]),
where is the category of finitely presented -algebras. In particular, there is a generic -model , such that any -model in a topos is realised as the inverse image of under an essentially unique geometric morphism . The observation of Blass, [5] is that, internally in , the function space of is completely characterised by polynomials,
where denotes the free -model generated by an additional element, which is a polynomial algebra on .
Gratzer et al., [15] have shown that when is taken to be the theory of bounded distributive lattices and is the generic -model, the equivalence implies Phoa’s principle—because for bounded distributive lattices, the polynomial -algebra always classifies the partial order on . This is a striking example of how the properties of a theory impact the internal logic of its classifying topos. We will see many more examples in this paper.
1.4. Synthetic quasi-coherence
Recently, the work of Blass has been greatly generalised by Blechschmidt in his PhD thesis [7] (also see the unpublished note [6] by the same author), which identifies a stronger property satisfied by the generic model in , termed (synthetic) quasi-coherence. The terminology is motivated by its application in algebraic geometry.
The main algebraic objects of study within the classifying topos will be the -algebras, i.e. -models equipped with a homomorphism . Of course, the identity homomorphism exhibits as a -algebra, and given any object , we may define
to be the -fold product of -algebras.
Conversely, given any -algebra we may define its spectrum to be the type of -algebra homomorphisms
The quasi-coherence principle in proposed by Blechschmidt, [7] is, then, that for any finitely presented -algebra , the canonical homomorphism
obtained on elements by transposing the evaluation map is an equivalence. Types of the form for such are called affine by Blechschmidt, [7]. This in fact induces an internal duality between finitely presented -algebras and affine types. In particular, it implies Blass’s result.
The appearance of finite presentation in the above formulation of quasi-coherence is perhaps misleading. Following the development of Blechschmidt, [7, 6], it is clear that the quasi-coherence principle for finitely presented algebras can be generalised to larger cardinality, if we enlarge the site of the classifying topos . For instance, one can consider
where denotes the category of countably presented -models. In this case, countably presented -algebras will be quasi-coherent in . This is the theoretical basis in a recent approach for synthetic topology suggested by Cherubini et al., 2024a [10], who consider a variation on the quasi-coherence principle for countably presented Boolean algebras. Horn theories whose operations have countable arity can also be considered in this case, so long as the base topos satisfies countable choice. Of course, these examples will not fit directly into the discipline of classifying toposes because geometric theories cannot have infinitary operations: but a suitable doctrine may arise when considering geometric morphisms whose inverse image functors preserve countable rather than finite limits.
Blechschmidt also points out that quasi-coherence descends from the classifying (presheaf) topos to any subtopos containing the generic model ; see Blechschmidt, [6, Cor. 7.7]. From a logical perspective, a subtopos can be identified as the classifying topos for a geometric quotient of the Horn theory . If is a sheaf for , then in the subtopos will be the generic -model. Hence, in such a subtopos, will validates more geometric sequents as specified by the topology, which should be viewed as certain local properties. For instance, Blechschmidt, [7], Cherubini et al., 2024b [11] work with local rings, where the additional geometric properties are validated by the Zariski topology. In the case of synthetic Stone duality, the Boolean algebra used by Cherubini et al., 2024a [10] is the set 2, which is again only true in a suitably chosen topology.
1.5. Summary of contributions
We study an emerging connection between synthetic ___domain theory and the quasi-coherence axioms for bounded distributive lattices and -frames. As a first indication of why quasi-coherence for these theories might be useful for ___domain theory, we have observed that by taking the interval to be a bounded distributive lattice or -frame internal to some topos, many important objects for ___domain theory can be written as spectra. This include the cubes (Example 2.5), the simplices (Example 3.5), and the final coalgebra (Example 8.3). Here we outline the main results of this paper, and discuss some of their potential applications.
-
(1)
Quasi-coherence produces synthetic ___domain theory. We show the principle of quasi-coherence suffices to account for all essential axioms for synthetic ___domain theory. Specifically:
-
(a)
In the case of both bounded distributive lattices and -frames, the quasi-coherence for finitely presented algebras (in fact a weaker assumption; cf. Section 3) implies that the generic interval and its dual form a dominance (Proposition 3.14 and Corollary 5.2). It also leads to an explicit computation of the partial map classifiers of quasi-coherent -algebras (Proposition 4.2), and spectra (Proposition 4.5).
-
(b)
The quasi-coherence principle allows an explicit computation of the observational preorder (Definition 6.6) on affine spaces (6.12) and quasi-coherent -algebras (Proposition 7.20).
-
(c)
A generalisation of Phoa’s principle that applies to an arbitrary -algebra is stated (Definition 7.1), and compared with quasi-coherence for polynomial algebras over (Theorem 7.3). As a corollary, we see that quasi-coherence of finitely generated free -algebras is equivalent to Phoa’s principle (Corollary 7.5).
-
(d)
In the case of -frames, quasi-coherence for countably presented algebras implies chain completeness of (Theorem 8.9) assuming is non-trivial. This is the only place where it is important to work with -frames and not bounded distributive lattices (cf. Remark 8.12). This gives a full account of the axioms for synthetic ___domain theory as specified by Hyland, [16].
-
(a)
-
(2)
Local properties of spectra. Though in general we do not assume the generic interval to satisfy any local property such as linearity, we do discuss many instances of them and investigate their consequences.
-
(a)
Section 5 introduces various local conditions for , all of which will be compatible with quasi-coherence. This showcases the flexibility of this framework to encompass different flavours of ___domain theory.
-
(b)
More interestingly, without assuming the local properties globally, one can still show that will be right orthogonal to the maps that classifies these local properties (Propositions 9.1, 9.3 and 9.5). In general, any limiting diagram of quasi-coherent algebras will induce a localisation class containing ; cf. Remark 8.13. This exhibits a new type of techniques in reasoning about domains in this framework.
-
(c)
As a special case of locality, we also connect to the recent approach of synthetic (higher) category theory [25, 8, 15]. In particular, we show spectra will be synthetic categories (Theorem 9.13). As another example, we also show is a synthetic category, in fact it satisfies all the orthogonality conditions discussed in this paper except chain completeness (Theorem 9.16).
-
(a)
-
(3)
New models for synthetic ___domain theory. Consequently, we uncover a large family of new models for synthetic ___domain theory based on sites induced by countably presented -frames. We will discuss these models in Section 10, and compare them to the existing sheaf models for synthetic ___domain theory [13] in Section 11.4.
1.6. Style and notation of the paper
Our motivation for ___domain theory encourages us to work in a context as general as possible, so as to allow future developments that connect more specific flavours of ___domain theory existing in the literature. This means that besides Section 10 where we discuss models, throughout the paper we will work constructively in a type theory enriched with a generic model . The base type system we work with is intensional type theory with a universe satisfying function extensionality, which can be interpreted in any (-)topos.
We do not assume any additional assumption globally. Whenever we introduce an assumption under the Axiom environment, it should be viewed as introducing the content of that assumption, rather than assuming it directly afterwards. In particular, the development is completely modular, and any additional assumption will be explicitly mentioned in each result.
In fact, we even take a step further. For the first half of this paper we will not work with bounded distributive lattices specifically. Instead, we assume to work with an arbitrary propositionally stable theory in the sense of Definition 3.1. In particular, the construction of the dominance and the computation of partial map classifier mentioned in Section 1.5 (Item 1a) works in this generality. Only from Section 5 onward will we then work more specifically with theories based on bounded distributive lattices.
Notation and preliminaries
We work informally in the vernacular of univalent foundations [31]. By proposition and set, we will always use them in the sense of the HoTT Book [31], i.e. -types and -types. The subuniverse of propositions and sets will be denoted as and , respectively. Propositions are used to define subtypes , and we also write the dependent sum suggestively as . In this case, for we also write to denote the proposition . Notice that function extensionality implies and are both closed under dependent product, and furthermore is closed under dependent sums. For emphasis, we will also use to denote the dependent product when is a family of propositions; in this case, we shall write for the propositional truncation . By existence we will always mean the latter truncated version, and we may emphasise this by referring to mere existence.111Here we deviate from the conventions of the HoTT Book [31]. Similarly, we use to mean the truncated proposition . We will write for the type of natural numbers. For , we will also abuse notation and treat as the finite type of elements.
Given an -algebra , we will abuse notation by identifying elements of as elements of via the structural map . For an algebra and two lists of terms , we will write to indicate the quotient algebra identifying with for all . We also write for the free algebra over generated from , or we also explicitly mention the generators . We write the coproduct of -algebras as .
2. Quasi-coherence and affine spaces
We will work in intensional type theory extended by a model of some Horn theory . In this section we will first describe the quasi-coherence property, and briefly recap some of its elementary consequences. The results in this section applies to any Horn theory , thus at this stage we do not assume to satisfy any additional properties.
As mentioned in Section 1.4, an -algebra is a -model equipped with a homomorphism . Here by a -model we always mean a set (in the sense of univalent foundations) equipped with a family of operations satisfying the axioms specified by . will denote the category of -algebras. We have a construction of spectra as a contravariant functor:
Definition 2.1 (Spectrum).
The spectrum of an -algebra is defined to be the set of -algebra homomorphisms from into :
This gives us a contravariant functor
acting on morphisms by pre-composition.
Definition 2.2 (Observation algebra).
Given a set , its algebra of observations is the product of following -algebras:
This gives us a functor
taking any set to its algebra of observations.
Proposition 2.3.
The spectrum construction is right adjoint to the functor taking a set to its algebra of observations:
Proof.
For any -algebra and set , we have a natural equivalence
Convention 2.4 (Unit and counit).
Following Taylor, [30], we shall write and for the unit and counit of the adjunction respectively. Due to the contravariance of the adjunction, the components of the counit are written in the category of -algebras. Both and are the evident evaluation maps.
Many sets can be naturally viewed as the spectrum of a certain -algebra. For example, we have because is the initial -algebra. Similarly, itself it the spectrum of the polynomial -algebra , as we have . More generally:
Example 2.5 (Cubes).
For any set , let be the free -algebra generated by . Then the set is a spectrum, because we have
Notice is the -fold coproduct of the polynomial -algebra , and its spectrum is the -fold product of . In fact, all the spectra can be obtained as equalisers of powers of :
Proposition 2.6.
Every spectrum is an equaliser of powers of .
To verify Proposition 2.6, we recall some results concerning adjunctions of descent type. This notion was first given in Barr and Wells, [2], and the following characterisation appeared in Kelly and Power, [19]
Definition 2.7.
An adjunction is said to be of descent type when either of the equivalent conditions hold:
-
(1)
The comparison functor into the Eilenberg–Moore category of the monad on is fully faithful.
-
(2)
For each , the counit is a coequaliser in :
Proposition 2.8.
For a Horn theory , the adjunction is of descent type.
Proof.
Consider a presentation where is a signature and is a set of Horn clauses over this signature; we may of course regard as an algebraic theory without any equations. We may then decompose the adjunction into the composite adjunction:
The existence and reflectivity of is established by Remark 3.19 (1) of Adámek and Rosicky, [1], as is a “quasivariety” in the sense of op. cit. The right-hand adjunction is monadic (i.e. of effective descent type) because is algebraic. Corollary 3.3 of Kelly and Power, [19] implies that the composition of a reflective adjunction followed by an adjunction of descent type is of descent type. ∎
We are now prepared to verify Proposition 2.6.
Proof of Proposition 2.6.
We wish to show that every spectrum an equaliser of powers of the interval. To that end, we note that the adjunction is of descent type by Proposition 2.8, since the theory of -algebras is Horn. Hence, the following is a coequaliser in :
Because is a right adjoint (Proposition 2.3), it sends colimits of algebras to limits of sets; hence, the following is an equaliser:
Remark 2.9 (Spectra are replete).
Recall the notion of replete type given by Hyland, [16]: is said to replete when it is right orthogonal to any map that is right orthogonal to. Throughout the paper, by right orthogonality we always mean the internal notion: is right orthogonal to a map iff the restriction map is an equivalence. Thus, an object is replete iff it belongs to the smallest internal localisation class containing . In particular, replete objects form an exponential ideal. This implies a spectrum as an equaliser of powers of will also be replete.
We will now define quasi-coherent algebras and affine spaces as the fixed points of the adjunction in Proposition 2.3:
Definition 2.10 (Quasi-coherent algebra).
An -algebra is said to be quasi-coherent when it is a fixed point of the adjunction , i.e. the component of the counit is an equivalence of -algebras.
We emphasise that the equivalence is not only an equivalence of types, but an equivalence of -algebras. This implies that the algebraic structure on can be viewed as pointwise induced by that on .
Remark 2.11 (Quasi-coherent algebras are replete).
The underlying set of any quasi-coherent algebra is also replete, as we have and hence the underlying set of is .
Example 2.12.
itself by definition is quasi-coherent. We have seen that . Under this equivalence, the canonical map
is exactly the identity on .
Definition 2.13 (Affine sets).
We say a set is affine if it is a fixed point of the adjunction , i.e. the unit map is an equivalence.
In the future whenever we indicate is affine with , we will assume is the quasi-coherent -algebra .
Remark 2.14 (Quasi-coherence and affineness as general properties).
Our notion of being quasi-coherent and affine already indicates an important difference between our approach and the ones taken in the related works [11, 10]. There, the notion is directly connected to the size of presentation, i.e. they are spectra of finitely presented or countably presented algebras. Since we aim for a completely modular development, our notion is more general, which a priori does not favour a class of size-related algebras.
Being the fixed points of an adjunction, quasi-coherent -algebras and affine spaces induces certain duality results. The first half of the following result appears in Prop. 2.2.1 of Cherubini et al., 2024b [11] (for the more restrictive definition of quasi-coherence). We include the result here for completeness since our assumption is slightly more general.
Proposition 2.15.
Let be quasi-coherent. For any -algebra , the canonical map is an equivalence. Similarly, if is affine, for any set the canonical map is an equivalence.
Proof.
This holds generally for fixed points of an adjunction. ∎
3. Open propositions and the dominance property
Notice that up till this point we have not used any special property of rather than the fact that it is a Horn theory. However, to move closer to the intended application in ___domain theory, we start by assuming our theory is propositionally stable in the following sense:
Definition 3.1 (Propositionally stable theory).
We say a Horn theory is propositionally stable, if it extends the theory of bounded meet-semi-lattices (1, ), and truth of an element is computed by slicing: For any -model and element , the quotient is given by the restriction map
where the slice by definition is , with the partial order on induced by the bounded meet-semi-lattice structure.
Remark 3.2 (Examples of propositionally stable theories).
of bounded meet-semi-lattices, of bounded distributive lattices, of Heyting algebras, and of Boolean algebras are all examples of propositionally stable theories. In fact, all finitary quotients of or will again be propositionally stable. More generally, for any propositionally stable theory and any -model , the theory of -algebras will again be so, because quotients of -algebras are computed the same as quotients of models of .
Remark 3.3 (The theory of -frames).
A -frame is a lattice with finite meets and countable joins satisfying the distributivity axiom
for any in and . We can axiomatise -frames as an infinitary algebraic theory , extending the theory of bounded meet-semi-lattices with a constant and a function symbol with countable arity. (Such an axiomatisation can be found e.g. in Adámek and Rosicky, [1, Exa. 3.26].) It is easy to see that is propositional.
A notable fact about -frames is that their finitary behaviours are exactly the same as bounded distributive lattices, in the sense that for any -frame , the finitely presented -frame coincides with the finitely presented bounded distributive lattice . Thus, there will be no difference between working with -frames or bounded distributive lattice when we work with finitely presented -algebras. The salience of countable joins will emerge only in Section 8 when we prove chain completeness of .
Remark 3.4 (Propositionally op-stable theories).
There is an evident “opposite” notion of being propositionally stable, which requires to extend the theory of bounded join-semi-lattices and the quotient for any is computed by . Every propositionally stable theory corresponds to a propositionally op-stable theory by dualising: The theories of join-semi-lattices and coHeyting algebras are propositionally op-stable; Boolean algebras and distributive lattices dualise to themselves; the theory of -frames is also propositionally op-stable.
For a propositionally stable theory , we think of the generic model as certain interval object, since it is equipped with a partial order. In this case, more types can be realised as spectra. The important examples are simplices:
Example 3.5.
For any , let be the -simplex
This type is indeed a spectrum, since by definition we have
To simplify later discussion, we might also introduce the following types isomorphic to simplices above,
The constant , which is the top element in , induces a predicate
which takes to the proposition . Using , we can think of the spectrum of an -algebra as its set of models, with a satisfaction relation as follows:
Definition 3.6 (Satisfaction).
For any -algebra and elements and , we say satisfies , written , if .
For instance, the satisfaction relation can be used to carve out affine subspaces consisting of models satisfying some element in :
Example 3.7.
For any -algebra and , the subtype defined below is again a spectrum:
The first observation is that takes to a spectrum:
Lemma 3.8.
The spectrum for is a proposition equivalent to .
Proof.
We have by definition. Since is the initial -algebra and is a quotient of , any homomorphism will be unique, hence is a proposition. It thus suffices to show implies . Because is propositionally stable, an -algebra morphism is a commutative diagram as follows,
In this case, we have
The final identity holds since preserves the top element. ∎
The goal of this section is to show that, under a suitable quasi-coherence assumption, makes a dominance in the sense of Rosolini, [26]. It turns out that it suffices to require all (finitely generated) principle quotients of to be quasi-coherent:
Definition 3.9 (Stable quasi-coherence and stably affine).
An -algebra is stably quasi-coherent, if for any and , the quotient is again quasi-coherent. An affine is stably affine, if is stably quasi-coherent.
As a first consequence, we show the following important lemma:
Lemma 3.10 (\AxiomSQCI).
The interval is conservative: For any
Proof.
Thus, under (\AxiomSQCI), we may view the generic algebra as a subuniverse of open propositions via the embedding :
Definition 3.11 (Open proposition).
We say that a proposition is open when is equivalent to for some in .
Under (\AxiomSQCI), by Lemma 3.10 the that will be unique, thus being open is itself a proposition. Also, open propositions are evidently closed under finite conjunctions, since preserves them.
More generally, we may define the notion of open subtypes:
Definition 3.12 (Open subtype).
A subtype of is open if for any the proposition is open.
If is affine, open subtypes are easy to classify:
Lemma 3.13 (\AxiomSQCI).
Let be affine. Then any open subset of is of the form for some .
Proof.
Let be an open subtype of , so that for any there exists some (by conservativity, necessarily unique) element that . Hence we have a map with . Because is affine, we may identify with an element , where . This way,
and thus is the subtype . ∎
Proposition 3.14 (\AxiomSQCI).
The type of opens forms a dominance.
Proof.
Since by definition is closed under finite meets, it suffices to show that an open subtype of an open proposition is again an open proposition. Suppose is an open proposition and is an open subtype of . By definition for some . Since is affine, Lemma 3.13 implies that is for some . Since is propositionally stable, equivalently can be viewed as an element with . This way,
which implies is also an open proposition. ∎
Remark 3.15 (Dominance without choice).
The related work of Cherubini et al., 2024b [11] contains a similar construction of a dominance from a generic ring. However, the proof there relies on certain classical axiom, which they call Zariski local choice. The reason that the construction here does not require a similar assumption is precisely due to conservativity of : the dominance is itself, rather than an image of . In this way, we can transform an open subtype of to a map , which by quasi-coherence can be further identified as an element of the algebra if is affine.
4. Partial map classifier
Given then dominance structure , we can construct internally the partial map classifier. For any type , it is given by
The functoriality is easy to express: For any , we have
There is an evident unit , where
The unit then classifies the partial maps out of with an open ___domain. The dominance also provides a multiplication , where takes any with and first to , where is the dependent sum
and is the partial element such that for and
Example 4.1.
By definition, it is easy to see that
For synthetic ___domain theory, the object of particular importance is the partial map classifier of the interval itself. The following computation works more generally for all stably quasi-coherent algebras:
Proposition 4.2.
If is stably quasi-coherent, then we have
Proof.
By assumption for , the quotient is again quasi-coherent. Now notice that we do have
By Proposition 2.3 and quasi-coherence of and ,
This way, it follows that
In other words, an element is simply a pair and that . This way, we can easily compute :
Corollary 4.3 (\AxiomSQCI).
.
Proof.
By Proposition 4.2 and the assumption (\AxiomSQCI),
The second equivalence is again due to being propositionally stable. ∎
Remark 4.4 (Algebra vs. geometry).
One interesting fact to note here is that, though by computation is a dependent sum of algebras, it is naturally equivalent to a spectrum. In some sense the source is the assumption that is propositionally stable, which allows us to identify the algebraic object as a subset . There will be more examples of such nature once we work more specifically with bounded distributive lattices; cf. Proposition 7.23.
More generally, for ___domain theoretic applications we would want to compute the partial map classifier of , or for any . For this purpose, we observe that we can more generally compute them for any spectra (not only affine spaces):
Proposition 4.5 (\AxiomSQCI).
For an -algebra , we have
Proof.
By (\AxiomSQCI) and Lemma 3.8, is affine for any . By the duality result in Proposition 2.15, we have
Corollary 4.6 (\AxiomSQCI).
For any , we have
and the unit takes in to in .
Proof.
By Proposition 4.5, for the spectrum ,
Again the third steps holds since is propositionally stable. ∎
5. Distributive lattices and locality
One important finitary axiom for synthetic ___domain theory is Phoa’s principle, which we have mentioned in Section 1.3 is a consequence of the quasi-coherence principle for bounded distributive lattices. Thus, we now work with the theory of bounded distributive lattices, or more generally the theory of -algebra for some bounded distributive lattice . As indicated in Remark 3.3, computation for finitely presented algebras will not change if we replace bounded distributive lattices by -frames. Hence, the results here and in Section 6 will be applicable to -frames as well, where only the properties of finitely presented -algebras matter.
As mentioned in Remark 3.4, the theory of bounded distributive lattices is also propositionally op-stable, and similarly for -frames. Hence, dual versions of the previous results will also hold by symmetry, when we exchange for and for . Thus, we first record some simple corollaries for the dual structure.
5.1. The dual dominance and co-partial map classifier
In this case, has a minimal element . Notice that the constant also induces a predicate on ,
which takes any to .
Corollary 5.1 (\AxiomSQCI).
is an embedding.
We will now call propositions in the image of closed, and a subtype classified by a closed subtype. Analogously to the previous Proposition 3.14, closed propositions also form a dominance:
Corollary 5.2 (\AxiomSQCI).
forms a dominance.
There is an accompanying “co-partial map classifier” for partial maps with closed domains, as Hyland, [16] pointed out. Concretely, for any type it is defined by
And analogously to Proposition 4.2, under (\AxiomSQCI) we can explicitly compute
where now the unit takes a sequence to in .
The remaining part of this section will introduce various locality axioms for the interval , and discuss some of their consequences with the quasi-coherence axiom.
5.2. Non-triviality
As a start, a minimal requirement for synthetic ___domain theory to model divergent computation is for the dominance to be closed under falsum. For this, we introduce the following minimal amount of non-triviality:
Geometrically, the proposition can be identified with the spectrum . Hence, (\AxiomNT) states that the unique map is an equivalence. This way, now will be affine, and it is both an open and closed proposition. As mentioned in introduction, (\AxiomNT) together with a strong enough quasi-coherence principle will imply all of Hyland’s axioms for synthetic ___domain theory. Before we discuss that in Section 8, a minimal amount of quasi-coherence provided by (\AxiomSQCI) already implies a lot of elementary properties. As a first consequence of (\AxiomNT) with (\AxiomSQCI), we can show that is almost the Boolean algebra in the following sense:
Proposition 5.3 (\AxiomNT, \AxiomSQCI).
For any we have:
In particular, the embedding induced by is -dense:
Proof.
If then by (\AxiomNT). On the other hand, by the conservativity result in Lemma 3.10 and (\AxiomNT), if then since . The dual case also follows by symmetry. Now given , if , which implies , equivalently , contradictory to (\AxiomNT). Hence, . ∎
This allows us to observe that the open and closed propositions have a bijective correspondence between them:
Corollary 5.4 (\AxiomNT, \AxiomSQCI).
For any proposition , is open iff is closed and vice versa. Furthermore, open and closed propositions are -stable.
On the other hand, Section 6 will show that under the axiom (\AxiomNT), quasi-coherence implies the interval will never be isomorphic to .
5.3. Locality
A slightly stronger axiom which is common in the practice of ___domain theory is that the dominance should furthermore be closed under all finite disjunctions. For this, we consider the following axiom:
Geometrically, the second condition states that the following two inclusions are jointly surjective.
Under (\AxiomL), the dominance will be closed under finite disjunctions. Furthermore, it also implies more elementary properties about :
Lemma 5.5 (L, \AxiomSQCI).
has no non-trivial complemented elements, i.e. if is complemented, then .
Proof.
Suppose has a complement . Then by (\AxiomL) we have
and similarly,
It follows that , and by Proposition 5.3, . This way, . ∎
Lemma 5.5 above allows us to realise 2 as a spectrum:
Example 5.6.
Consider the algebra . By construction, classifies complemented elements in -algebras. Thus, assuming (\AxiomL) and (\AxiomSQCI), by Lemma 5.5
In particular, this means that for any affine space , will be equivalent to the set of complemented elements in by Proposition 2.15.
As another example of a new spectrum (\AxiomL) allows us to define:
Example 5.7.
Consider the right outer horn , which we may describe by the following pushout:
By viewing as a subspace of , we may identify it as follows:
Assuming (\AxiomL), , it follows that
Axiom (\AxiomL) clearly has a dual counterpart:
Similarly to Example 5.7, this allows one to identify the left outer horn as a spectrum. Of course we can combine the axioms (\AxiomL) and (\AxiomCL).
5.4. Linearity and coskeletality
An axiom even stronger than both (\AxiomL) and (\AxiomCL) is the simplicial axiom (\AxiomSL):
Geometrically, the simplicial axiom states that the two simplices covers the square . The strongest locality axiom states that the simplicial structure is truncated to level 1:
Geometrically, (\AxiomOneCS) additionally requires the canonical inclusion from the boundary to the 2-simplex is an equivalence, where type-theoretically is simply
In general we will at most work with the weakest locality axiom (\AxiomNT). However, we will show in Section 9 that , hence all quasi-coherent algebras and spectra, will be right orthogonal to the maps that define these local principles as shown above. We will see in Section 10 that these axioms corresponds to various coverages we can put on the presheaf classifying topos, and the orthogonality result particularly implies that will belong to these subtopoi.
6. Order-theoretic structure on algebras and spaces
In this section, we describe several important local orderings on algebras and spaces, summarised in Table 1.
(canonical order) | ||||
(behavioural preorder) | ||||
(observational preorder) | ||||
(satisfaction order) |
6.1. Local ordering of algebras
Every -algebras has a built-in “canonical” partial ordering.
Definition 6.1 (Canonical partial order).
We shall write for the canonical partial ordering of an -algebra defined equivalently by meets or joins:
Example 6.2.
The canonical partial ordering of an algebra of observations is given pointwise because is the product of algebras :
In addition to the canonical partial order, there is a local preordering of algebras that mirrors the observational order for spaces.
Definition 6.3 (Behavioural preorder).
For any -algebra , we define the behavioural preorder on as follows:
To relate the behavioural preorder to the canonical partial ordering, we first define a slight weakening of the quasi-coherence principle.
Definition 6.4.
We say that an -algebra is pre-quasi-coherent when the counit component is an embedding.
Proposition 6.5.
The canonical partial order of any pre-quasi-coherent -algebra is its behavioural preorder.
Proof.
For we have the following sequence of equivalences:
The first equivalence requires that is pre-quasi-coherent. ∎
6.2. Local ordering of spaces
The observational preorder is a classical notion in synthetic ___domain theory (cf. Phoa, [23], Hyland, [16]):
Definition 6.6 (Observational preorder).
For any type , we define the observational preorder on as follows:
By definition, the observational preorder is reflexive and transitive. As already observed in loc. cit., one important property of the observational preorder is that every map is monotone w.r.t. this order:
Lemma 6.7.
For , in implies in .
Proof.
This simply follows from compositionality of functions. ∎
We can very easily classify the observational preorder on a set with decidable equality.
Lemma 6.8 (\AxiomNT).
If has decidable equality, then the observational preorder on is discrete, in the sense that for ,
Proof.
Since has decidable equality, we can define functions by case distinction. If , we can construct a function with , contradicting with . Hence, , and with decidable equality this implies . ∎
Corollary 6.9 (\AxiomNT).
If the observational preorder on has a least or greatest element, then is internally connected in the sense that is right orthogonal to for any set with decidable equality.
Proof.
Since is inhabited, the restriction map is an embedding (here we use the fact that is a set). It suffices to show that this map is also surjective. Fixing , we must show that there exists such that for all . Let be the least element of in its observational preorder, so that we have ; by Lemma 6.7 we have which implies by Lemma 6.8. Therefore, we may choose . The case for a maximal element is analogous. ∎
Remark 6.10.
With greater efforts, the above proof would also work with the weaker assumption that the observational preorder on is connected. This generalises a similar result given by Hyland, [16, Prop. 4.4.1].
6.3. The satisfaction order of a spectrum
Corollary 6.9 demonstrated the usefulness of the observational preorder. We would like to use it to show e.g. that is internally connected. For this, we need to characterise the observational preorder on affine spaces. It turns out that for any affine space, this is simply its satisfaction order in the sense defined below:
Definition 6.11 (Satisfaction order).
We shall write for the satisfaction order on the spectrum of an -algebra as defined below:
Observation 6.12.
When is a quasi-coherent -algebra (and thus is affine), the observational and satisfaction orders on coincide.
Proof.
For any we have the following chain of equivalences:
Observation 6.13.
The satisfaction order on the spectrum of a free -algebra is induced by the canonical order on the observation algebra via the canonical bijection .
In the above, is playing two roles simultaneously: it is both a spectrum of and the observational algebra of .
Proof.
For any we have the following equivalences:
Lemma 6.14.
For any quotient of a free algebra , the satisfaction order on is induced by the satisfaction order on via the inclusion .
Proof.
Let . We have the following equivalences:
The second equivalence holds because is surjective. ∎
7. Phoa’s principle, finitary quasi-coherence, and homotopy
The interval is playing multiple roles so far:
-
(1)
Mapping a space into computes its algebra of observations .
-
(2)
Homomorphically mapping an algebra into computes its spectrum .
In both cases above, is viewed as an algebra of observations and not as a geometrical figure. For the latter, we may consider mappings from into either an algebra or a space as defining a notion of (directed) homotopy. In particular, the function space classifies paths drawn in ; this geometrical role of as a figure shape for paths is the primary one in applications to synthetic category theory [25], and it is also upon the geometry of the interval that the notion of chain completeness in synthetic ___domain theory is founded (cf. Section 8).
7.1. The generalised Phoa principle and quasi-coherence for polynomial algebras
As soon as we have path spaces , we immediately wish to begin characterising them for specific . For example, the path space of a function space is given pointwise in terms of the path space of , which follows immediately from the laws of exponentials. To characterise paths in spaces like and , however, we need additional axioms. The simplest such axiom asserts that the polynomial algebra is quasi-coherent—which, in the case of bounded distributive lattices, is equivalent to Phoa’s principle.
Definition 7.1.
We say that an -algebra satisfies the generalised Phoa principle when either of the following equivalent conditions hold:
-
(1)
The path space classifies the canonical order on .
-
(2)
For any function , we have .
Proof.
Suppose that classifies the canonical order on ; we must show that for any we have . By our assumption that classifies the canonical order on , we know that that any such function must be uniquely determined by its values on and and moreover that ; therefore, it suffices to observe for arbitrary such that
Conversely, suppose that every path is of the form . We must show that holds if and only if there exists a unique path sending and to and respectively.
-
(1)
If holds, we define ; this path is unique with the required property by our assumption.
-
(2)
If there exists a path from to , we know that this takes the form . Therefore we have
The standard Phoa principle is the generalised Phoa principle for itself.
Theorem 7.2.
If satisfies the generalised Phoa principle, then so does the polynomial -algebra .
Proof.
Fixing , we must check that . The normal form theorem for polynomials of bounded distributive lattices [20, Ch. 1, Thm. 10.11] implies that any function must be determined by functions such that in . Meanwhile, Phoa’s principle for characterises and as follows:
Hence we have . On the other hand, we can use the normal form theorem to compute and as polynomials in :
Hence . Using elementary lattice algebra we deduce that . ∎
Theorem 7.3.
Let be a quasi-coherent -algebra. Then the following are equivalent:
-
(1)
The polynomial -algebra is quasi-coherent.
-
(2)
The generalised Phoa principle holds for .
Proof.
We note that unconditionally, and so the algebra of observations is actually the path space of ; taking quasi-coherence of into account, the counit component is the evaluation map , and we wish to prove that the latter is an equivalence if and only if for each we have . This follows immediately from the normal form theorem for polynomials of bounded distributive lattices [20, Ch. 1, Thm. 10.11]: any polynomial is uniquely of the form with . ∎
Corollary 7.4.
If both and are quasi-coherent, then so is .
Proof.
By Theorem 7.3, the generalised Phoa principle holds for because and are quasi-coherent; by Theorem 7.2, the generalised Phoa principle holds for because it holds for . By Theorem 7.2 again, the generalised Phoa principle holds for because it holds for . ∎
Corollary 7.5.
The following are equivalent:
-
(1)
The polynomial -algebra is quasi-coherent.
-
(2)
Phoa’s principle holds.
Proof.
By Theorem 7.3, because is unconditionally quasi-coherent. ∎
We therefore identify the axiom (\AxiomSQCP) below accordingly, which we have seen to be equivalent to Phoa’s principle.
The quasi-coherence axiom for polynomials in one variable is surprisingly strong:
Lemma 7.6 (\AxiomSQCP).
Every finitely generated free -algebra is quasi-coherent.
Proof.
The free -algebra on generators is just , which is automatically quasi-coherent. The free -algebra on one generator is quasi-coherent by assumption. In the case of we recall Corollary 7.4 and see that quasi-coherent implies quasi-coherent, which implies quasi-coherent and so on. ∎
Theorem 7.7 (\AxiomSQCP).
For any -algebra , the path space classifies the satisfaction order on .
Proof.
By assumption is quasi-coherent, so from Proposition 2.15 may characterise the path space of algebraically as follows:
By the normalisation theorem for , the pair of evaluation maps
classifies the canonical order on . Thus every homomorphism is canonically induced by elements with such that . ∎
Theorem 7.8 (\AxiomSQCP).
If both and are quasi-coherent -algebras, then the path space classifies the canonical order on .
Proof.
By assumption and Proposition 2.3, since , we have
We have already seen that classifies the canonical order of by the normalisation result for polynomial -algebras. ∎
Remark 7.9 (Algebraic properties in classifying topoi).
We emphasise that the above proofs are purely the consequences of the algebraic fact that classifies the canonical order of any -algebra , plus quasi-coherence. This is a perfect example of how an algebraic property of a theory has a non-trivial effect on the internal logic of its classifying topos.
Corollary 7.10 (\AxiomSQCP).
Each cube is affine.
Proof.
The -cube is the spectrum of the free -algebra , which is quasi-coherent by Lemma 7.6. ∎
Corollary 7.11 (\AxiomNT, \AxiomSQCP).
The -cube is internally connected.
In particular, is not the boolean set .
Proof.
By 6.12 and 6.14, the observational preorder space is induced by the satisfaction order on because is affine (Corollary 7.10); the satisfaction order is simply the pointwise order, so it has a top element. Thus by Corollary 6.9, is internally connected. So are because they are retracts of . ∎
7.2. Comparing the different orders
An immediate application of Phoa principle is to compare path structure with the observational preorder.
Observation 7.12 (\AxiomSQCP).
For any type , the boundary evaluation map factors through the observational preorder relation as displayed below:
Proof.
Fix a path . We have by Phoa’s principle; because every function is monotone in the observation preorder, we conclude . ∎
Definition 7.13.
A type for which is an surjection is called linked (following Phoa, [23]); when the map is an equivalence, is called strongly linked.
Lemma 7.14 (\AxiomSQCP).
Any affine space is strongly linked.
Proof.
Let be the spectrum of a quasi-coherent algebra . By Theorem 7.7, the path space classifies the satisfaction order on , which is to say that there is a (necessarily unique) path from to if and only if
Because is quasi-coherent, we have and there is a unique path from to if and only if
but this is the observational order of . ∎
Lemma 7.15 (\AxiomSQCP).
The interval is strongly linked.
Proof.
The interval is affine because it is the spectrum of , which is quasi-coherent by (\AxiomSQCP). ∎
Lemma 7.16.
Being strongly linked is an exponential ideal.
Proof.
Phoa, [23, Prop. 5.4.4] implies that being linked is an exponential ideal, because an exponential is an internal limit. A type is strongly linked if and only if it is linked and it is -separated in the sense that is an embedding; but -separated types also form an exponential ideal (in fact, a reflective exponential ideal). ∎
Lemma 7.17 (\AxiomSQCP).
Any algebra of observations is strongly linked.
Proof.
We assume that for some not necessarily affine ; then is strongly linked by Lemmas 7.16 and 7.15. ∎
Corollary 7.18.
When is strongly linked, the observational preorder on is induced pointwise by the observational preorder on .
Proof.
Let be strongly linked; by Lemma 7.16, so is . Thus is precisely . ∎
Proposition 7.19 (\AxiomSQCP).
On any observation algebra , the observational preorder coincides with the canonical order.
Proof.
By Phoa’s principle, is strongly linked; thus by Corollary 7.18, the observational preorder on is pointwise induced by that of . In particular, we have the following equivalences involving the function space :
The second equivalence uses Phoa’s principle and Lemma 7.17 to identify the canonical order on with its observational preorder.
Proposition 7.20 (\AxiomSQCP).
For a quasi-coherent -algebra , the observational preorder on the underlying set of is induced by the canonical order on by the counit isomorphism . Hence the observational order and canonical order of coincide.
Proof.
Because is quasi-coherent, the counit component gives us an order-isomorphism . By Proposition 7.19, this is an equivalently an order-isomorphism . Meanwhile, any bijective function tracks an isomorphism of observational preorders, so the inverse function tracks an order-isomorphism . ∎
Corollary 7.21 (\AxiomSQCP).
The canonical partial order, observational preorder, and behavioural preorder of a quasi-coherent -algebra all coincide.
Proof.
By Propositions 7.20 and 6.5. ∎
7.3. Quasi-coherence for finitely presented algebras
The (\AxiomSQCP) axiom does not imply that the simplices are affine; for this, we need a stronger quasi-coherence principle that applies to all finitely presented algebras.
It is immediate that is affine under (\AxiomSQCF).
Corollary 7.22 (\AxiomSQCF).
Each -simplex is internally connected.
Proof.
This follows in the same way as Corollary 7.11. ∎
7.4. Classification of simplices by algebras
At the end of this section, we describe another interesting perspective arising from the proof of Theorem 7.7. We have seen that the dualising object has a double role: It is both an algebra and a spectrum. The mixture of algebraic and geometric object has also been observed in Remark 4.4. The proof of Theorem 7.7 gives us many more such examples. For instance, classifies the order on , which by definition is the spectrum . In fact, all the simplices are classified by some algebra.
Just as we described in terms of descending sequences in , we can do the same replacing with another algebra . For any and -algebra , we define to be the type of lists of decreasing elements of length in , namely
Of course is simply . We have a general algebraic description of .
Proposition 7.23.
For any and -algebra , there is an equivalence
sending any in to the polynomial
We omit the proof here, which again follows from a general normal form for polynomials with finitely many variables for bounded distributive lattices; see Lausch and Nobauer, [20, Ch. 1, Thm. 10.21]. We only note here that the above polynomial is well-defined, in the sense that the expression is associative and its value does not depend on how one add parenthesises. The reason for this is that for any expression to have a definite meaning in a distributive lattice, i.e. the two value and coincide, it suffices to have . Then the condition for the parameters, and the fact that , makes sure the above polynomial has a definite meaning in the quotient.
8. Chain completeness and infinitary ___domain theory
Until this point, we have seen that elementary axioms for synthetic ___domain theory follow from (SQCF) for bounded distributive lattices and thus -frames. In this section we show that the final axiom of synthetic ___domain theory, viz. chain completeness of the interval , is also a consequence of quasi-coherence for -frames. In fact, it implies the main infinitary axiom of synthetic ___domain theory in Fiore and Plotkin, [12]—the initial algebra for is inductive.
8.1. Chain completeness and inductivity of the initial algebra
Chain completeness is again specified as an orthogonality condition. Let denote the initial algebra and final coalgebra for the partial map classifier , respectively. Intuitively, behaves as an infinite chain, and as an infinite chain with a top element. The chain completeness condition is thus intuitively stating a type has joins for infinite sequences internally:
Definition 8.1 (Chain completeness).
A type is chain complete if it is right orthogonal to the inclusion .
The importance for chain completeness of for ___domain theory is that, as observed in Hyland, [16], it produces fixed points of endo-morphisms on a suitably defined class of objects, which could be viewed as domains.
As by construction is a polynomial functor, it preserves connected limits. This implies the final coalgebra can always be constructed as a sequential limit as follows,
However, the dual statement fails for the initial algebra , i.e. it is not always the following sequential colimit,
Definition 8.2.
is inductive if it is the sequential colimit above.
The failure of inductivity of has been observed in various realisability models for synthetic ___domain theory [32]. As shown by Fiore and Plotkin, [12], the inductivity of is indeed a desirable property: If it holds, then there will be a much closer correspondence between models for synthetic ___domain theory and models for axiomatic ___domain theory. Specifically, for a model of synthetic ___domain theory where is inductive, one can construct a corresponding model of axiomatic ___domain theory, where the domains are the well-complete types222A type is well-complete if is chain complete.. The inductivity of makes an inductive fixed point object in the category of domains in the sense of loc. cit., which is part of their requirement for an axiomatic model of ___domain theory.
The remaining part of this section is dedicated to showing that under a suitable quasi-coherence assumption, we can indeed show that is chain complete (Theorem 8.9) and is inductive (Proposition 8.7). In fact, the proof of Theorem 8.9 is the only place we where have used the essential properties of -frames, and we will also indicate why it fails for bounded distributive lattices in Remark 8.12.
8.2. Quasi-coherent for countably presented algebras and infinitary ___domain theory
The connection between the infinitary aspect of synthetic ___domain theory and quasi-coherence starts from the observation that the final coalgebra for the functor can be described as a spectrum:
Example 8.3 ( is a spectrum).
As mentioned, the final coalgebra can be characterised as the sequential limit,
where under the isomorphisms, the transition map takes the sequence to the final segment . As observed by Hyland, [16, Sec. 5.2], it also has an equivalent type-theoretic description as the object of infinite descending sequences in ,
This way, we have a natural equivalence
Here now is a countably presented -algebra.
In general, by a countably presented -algebra we mean one of the form for some -indexed lists of terms . In particular, all finitely presented -algebras will also be countably presented. Motivated by the above characterisation of , we naturally consider the following stronger quasi-coherence principle:
Remark 8.4.
Our modular development in the previous sections can be now immediately applied to if we assume (\AxiomSQCC). For instance, the description of the observational preorder for affine spaces in 6.12 now applies to , which shows this again coincides with its pointwise order as a subspace of . In particular, it also has both a top and bottom element, thus Corollary 6.9 now implies is also internally connected.
(\AxiomSQCC) combined with the non-triviality axiom (\AxiomNT) has many logical consequences. The crucial observation is that (\AxiomNT) implies a weak form of Nullstellensatz result, as already noted by several authors [7, 6, 11]:
Lemma 8.5 (\AxiomNT).
For an affine space , iff is trivial, viz. in .
Proof.
The backward direction holds since because is non-trivial (\AxiomNT), thus there is no homomorphism from a trivial algebra to . For the forward direction: by assumption , which implies is trivial. ∎
Together with (\AxiomSQCC), this implies the following form of Markov’s principle; a similar result is shown by Cherubini et al., 2024a [10]:
Lemma 8.6 (\AxiomNT, \AxiomSQCC).
For any , we have
Proof.
Equipped with the above result, we can now proceed to study the initial algebra for . Jibladze, [17] has given a beautiful formula for a type-theoretic description of as the following subset of ,
For another proof, see e.g. that of van Oosten and Simpson, [32]. In the presence of Lemma 8.6, this description can be greatly simplified:
Proposition 8.7 (\AxiomNT, \AxiomSQCC).
is equivalent to the following subset of ,
In particular, is the colimit of the following sequence, thus is inductive:
The transition maps above are given by appending to the end of a descending sequence. Under the identification , the corresponding map is the unit component .
Proof.
Let . It suffices to show that
Assume the premise. We can instantiate to . By assumption, we have , which by Corollary 5.4 is equivalent to . Then Lemma 8.6 implies this is . The fact that this makes into the above sequential colimit follows from van Oosten and Simpson, [32, Cor. 1.10]. ∎
Using this colimit description of and the fact that is affine, we can show is chain complete. This again crucially depends on a normal form result for a countably presented -frame, which generalises the finitary version for bounded distributive lattices given in Proposition 7.23:
Lemma 8.8.
For the countably presented -frame , an element can be uniquely written as
with for all . In other words, is isomorphic to the following -frame with the pointwise order induced by ,
Proof.
We directly prove that satisfies the universal property of the countably presented -frame . We pick the generators in as follows,
For any -frame with , we define a map sending to
By construction it is easy to see is a -frame morphism. Evidently for any ,
Furthermore, for any -frame map with , we must have because any in can be written as
which implies that
This completes the proof. ∎
Theorem 8.9 (\AxiomNT, \AxiomSQCC).
is right orthogonal to the inclusion .
Proof.
Since is now affine, we have
On the other hand, since is the colimit of and they are affine, we have
Note that the transition maps induced by under quasi-coherence gives us the following maps on algebras:
Hence, it suffices to show that is indeed the sequential limit of the above -algebras,
where the map takes to itself for , and takes to for . We can more directly compute the above sequential limit via Proposition 7.23,
where forgets the last entry, i.e. it takes to . Hence, the sequential limit is given by
Now the desired result follows from Lemma 8.8. ∎
Remark 8.10.
As indicated in Remarks 2.11 and 2.9, both spectra and quasi-coherent -algebras are replete. The above result then implies that they are also chain complete, i.e. orthogonal to .
Corollary 8.11 (\AxiomNT, \AxiomSQCC).
. In particular, is not affine.
Proof.
By the above proof, we have . ∎
Remark 8.12 (-frames vs. bounded distributive lattices).
Note that the inductivity of as shown in Proposition 8.7 is independent of working with distributive lattices or -frames. The completeness result shown in Theorem 8.9 is the only one in this paper that works for -frames and not for bounded distributive lattices more generally. The reason is exactly because the normal form in Lemma 8.8 for the countably presented bounded distributive lattice will not be isomorphic to . More specifically, for the countably presented bounded distributive lattice , since it is a sequential colimit of the following finitely presented bounded distributive lattices
and, due to the fact that the theory of bounded distributive lattices is finitary, this sequential colimit of algebras is computed the same as their underlying sets. Via Proposition 7.23, the result can be identified as the subtype of ,
where in some sense is the order-dual to the inclusion . The fact that we have all infinite increasing sequences in the case for -frames is clearly due to the fact that we have countable disjunctions in -frames, as when identifying in with .
Remark 8.13 (Limits of algebras induce locality for ).
By inspecting the proof of Theorem 8.9, it is clear that is chain complete precisely because there is a specific limiting diagram of quasi-coherent -algebras. In general, any such limit diagram will induce an orthogonality property satisfied by the all spectra, and we will see many more examples in Section 9. From Remark 8.12 it is clear whether a diagram of -algebras is a limit heavily relies on the underlying algebraic theory. This is again a perfect example of how the algebraic properties of a theory has significant consequences for the internal logic of its classifying topos.
9. Local properties for the interval
In this section we review some of the locality axioms we have introduced in Section 5. As mentioned in the introduction, the observation is that even if we do not assume them to be true globally for the interval , we can still show the maps representing the locality axiom to be left orthogonal to . Unlike chain completeness in Theorem 8.9, the local conditions in Section 5 are all induced by limits of f.p. -algebras. Thus in this section, it does not matter whether we work with bounded distributive lattices or -frames.
Starting from the simplest example, let us recall the local property (\AxiomNT). In general it is not necessarily , if we do not assume (\AxiomNT). However, we can look at the localisation class that thinks this map is invertible. The following fact shows that, from the perspective of , (\AxiomNT) indeed holds:
Proposition 9.1 (\AxiomSQCI).
is right orthogonal to .
Proof.
Observe the proposition is propositionally equivalent to the (affine) spectrum . By Proposition 2.15 we have:
The last isomorphism is due to being the terminal -algebra, since its underlying set is the singleton. ∎
As another example, consider the simplicial axiom (\AxiomSL). We skip the discussion of (\AxiomL) and (\AxiomCL), because (\AxiomSL) is strictly stronger, and it has a better-known geometric meaning. Recall from Section 5.4 that (\AxiomSL) can be represented geometrically as an embedding .
Definition 9.2 (Simplicial type).
A type is said to be simplicial when it is right orthogonal to the inclusion .
Again, (\AxiomSL) holds globally iff the above embedding is an equivalence, thus every type in this case will be simplicial. When it does not hold globally, we can still show:
Proposition 9.3 (\AxiomSQCF).
is simplicial.
Proof.
To show that is right orthogonal to , it is equivalent to verify that for any indicated below making the solid diagram below commute, there exists a unique lift making the whole diagram commute:
Now by (\AxiomSQCF), since the vertices of the left square are all affine, equivalently it suffices to show we have a pullback of -algebras,
Recall from Proposition 7.23 that the normal form of elements in . An element there can be viewed as the polynomial
with . And similarly for in ,
They agree in iff
This then gives us a polynomial in , which we write as follows,
where and . By the general normal form for for any -algebra , if we view as , the above exactly corresponds to the normal form of polynomials in ; see also Lausch and Nobauer, [20, Ch. 1, Thm. 10.21]. This means the above is a pullback of -algebras. ∎
Finally, we discuss the even stronger locality condition (\AxiomOneCS). As mentioned in Section 5.4, the additional property of (\AxiomOneCS) is characterised by the embedding .
Definition 9.4 (1-coskeletal type).
A type is said to be 1-coskeletal when it is both simplicial and right orthogonal to the boundary inclusion .
Proposition 9.5 (\AxiomSQCF).
is 1-coskeletal.
Proof.
Completely similar to the proof of Proposition 9.3, it suffices to show that we have a limit diagram of -algebras as follows,
Now by the normalisation theorem, an element in the limit consists of in , in and in , such that
This exactly corresponds to a normal form in the algebra with as follows (cf. Proposition 7.23)
Hence, the above is a limit diagram of -algebras. ∎
Besides the locality principles discussed in Section 5, we also consider the orthogonality classes emerging from synthetic category theory, as introduced by Riehl and Shulman, [25]. A synthetic category will be a type that satisfies the internal orthogonality condition of being simplicial, Segal complete, and Rezk complete.
We start with the Segal completeness condition. Besides the outer horn discussed in Example 5.7, we can also define the inner horn as a pushout,
As a subtype of , it can be identified as .
Definition 9.6 (Segal complete types).
is called Segal complete when it is right orthogonal to .
Remark 9.7 (Path transitivity).
The Segal completeness condition has also been studied independently by Fiore and Rosolini, [14] under the name of path transitivity.
Proposition 9.8 (\AxiomSQCF).
is Segal complete.
Proof.
In this case, we need to show the following diagram is a pullback of -algebras,
Again by the normal form theorem, an element in the pullback is given by in and in with . This way, it again corresponds to the following normal form in by Proposition 7.23,
Hence, the above is again a pullback. ∎
Next we consider the Rezk completeness condition. Following Buchholtz and Weinberger, [8], we can define the type classifying categorical equivalences as the colimit of the following diagram,
Definition 9.9 (Rezk types).
We say that a type is Rezk complete when it is right orthogonal to .
Proposition 9.10 (\AxiomSQCP).
is Rezk complete.
Proof.
Notice by Theorem 7.7 classifies the canonical order on , which is antisymmetric by definition. Thus will be Rezk complete. ∎
In fact, is not only a synthetic category but in fact a synthetic poset; this property too can be expressed in terms of orthogonality. To that end, we define the “walking parallel pair” to be the following pushout:
Definition 9.11 (-separated types).
is called -separated when it is right orthogonal to .
Equivalently, is -separated iff is an embedding. It is an immediate consequence of (\AxiomSQCP) that is -separated.
The notion of synthetic categories and synthetic posets are formulated as these orthogonality classes:
Definition 9.12 (Synthetic categories and synthetic posets).
A type is a synthetic category, if it is Segal and Rezk complete. We say it is a synthetic poset, if it is also -separated.
Theorem 9.13 (\AxiomSQCF).
Any spectrum or quasi-coherent -algebra will be a synthetic poset.
Proof.
This follows from being a synthetic poset, and spectra and quasi-coherent algebras are replete as indicated in Remarks 2.11 and 2.9. ∎
At the end of this section, we also discuss the example of , which as we have seen is not affine. We first observe its observational preorder again coincides with its expected pointwise order viewed as a subspace of :
Lemma 9.14 (\AxiomNT, \AxiomSQCC).
The inclusion is an order embedding for the observational preorder. In particular, the observational preorder on is induced pointwise as a subspace of .
Proof.
The fact that is an order embedding for the observational preorder follows from chain completeness of in Theorem 8.9. The fact that the observational preorder on coincides with its satisfaction order as a spectrum follows from 6.12 by the fact that it is affine under (\AxiomSQCC). ∎
Furthermore, we can show that also satisfies a version of the Phoa principle, i.e. the path space again classifies its pointwise order. For this we need to compute the path space .
Recall from Proposition 8.7 that, under (\AxiomNT) and (\AxiomSQCC), is a colimit . As indicated in Remark 8.12, this does not depend on working with bounded distributive lattices or -frames. Type-theoretically, we have also shown that can be realised as the following subspace of ,
This way, the inclusion can be viewed as follows,
which implies is downward closed. This allows us to directly compute :
Proposition 9.15 (\AxiomNT, \AxiomSQCC).
We have a family of equivalences
and similarly by replacing with or . In particular, classifies the pointwise order on .
Proof.
We show the following canonical map is an equivalence,
It is evident this map is an embedding, hence it suffices to show it is surjective. Given . By assumption, there merely exists that factors through . Now the claim is that the entire map factors through . This indeed holds, since we have shown in Lemma 9.14 that the observational preorder on is pointwise. By monotonicity, for any we have , which implies belongs to as well, since is a downward closed. The same holds for cubes or simplices since they all have a top element. This way, classifies the pointwise order on since all simplices satisfy Phoa’s principle as shown in Theorem 7.7. ∎
As another consequence, we can also show the following general result establishing a large family of orthogonality conditions satisfied by :
Theorem 9.16 (\AxiomNT, \AxiomSQCC).
Let be a map where are finite colimits of cubes or simplices. If each is -local, then so is .
Proof.
Let be a finite colimit with each a simplex or a cube.
The second equivalence holds by Proposition 9.15; the third holds since finite limits commutes with sequential colimits. Thus, if each is -local, i.e. , then so is . ∎
Remark 9.17 (Localisation classes closed under sequential colimits).
Notice that in Proposition 9.15 we do not need to use any form of choice principle to show the exponential commutes with the sequential colimit , exactly because is downward closed. However, for general sequential colimits, the same proof still goes through if we assume:
Axiom (Choice principle for ).
For any type family over , we have
Furthermore, the same holds for cubes and simplices since the types satisfying the choice principle are closed under finite products and retracts. Assuming satisfies choice, following the proof of Theorem 9.16, one can show more generally that any orthogonality class specified by maps between finite colimits of simplices or cubes are always closed under sequential colimits. However, not in every model of quasi-coherence will satisfy the above choice principle. Hence, we do not include this result in the main text, as we tend to keep our assumptions as minimalistic as possible.
Quasi-coherence axioms: \PrintAxiomSQCI\PrintAxiomSQCP\PrintAxiomSQCF\PrintAxiomSQCC
Locality axioms: \PrintAxiomNT\PrintAxiomL\PrintAxiomCL\PrintAxiomSL\PrintAxiomOneCS
10. New models for synthetic ___domain theory
It is now instructive to discuss models for the axioms we have used in the previous sections. For this purpose, we now work externally under the assumption that the base topos satisfies the axiom of choice. Our axioms for synthetic ___domain theory can be organised into two classes:
-
(1)
The quasi-coherence principle (\AxiomSQCF), (\AxiomSQCC);
-
(2)
Various locality axioms discussed in Section 5.
For any Horn theory , we know from Blechschmidt, [6, 7] that the generic -model satisfies quasi-coherence for finitely presented -algebras, viz. (\AxiomSQCF). As mentioned in Section 1.4, the proof of quasi-coherence for finitely presented -algebras adapts readily to the countably presented case, provided we work with the larger site
where is the category of countably presented -models. The generic -model in is again defined to be the forgetful functor , and we also denote it as . In other words, (\AxiomSQCC) holds in the larger topos . And in this case, we can also allow to contain algebraic operations of countable arity.
These two topoi and are intimately related. There is a fully faithful and left exact inclusion of sites
inducing an adjoint triple (cf. Caramello, [9, Thm. 7.20]),
equivalently, a local geometric morphism .
More generally, it is already observed by Blechschmidt, [6, Thm. 4.11] that if we have a topology on where is a -sheaf, then (\AxiomSQCF) again holds in the sheaf subtopos . We refer to such a topology as -admissible, or simply admissible when no confusion could arise. For instance, since is representable, any subcanonical topology will in particular be admissible. In the same way we can define admissible topology for the larger presheaf topos , which again is a topology making the generic model a sheaf. In this case, the sheaf subtopos also models (\AxiomSQCC). As an example, this is the theoretical basis of the quasi-coherence principle for countably presented Boolean algebras in the topos of light condensed sets introduced by Clausen and Scholze, as shown by Cherubini et al., 2024a [10].
An admissible topology on restricts to an admissible one on . In this case, the adjoint triple mentioned above between the two presheaf topoi will restrict to the sheaf subtopoi,
which again identifies as a local topos over . These admissible topologies on or are exactly the required data to validate various local properties of the generic model .
More specifically for us, let be the category of -frames, i.e. whose objects are posets with finite meets and countable joins, where binary meets distribute over countable joins. We will also use to denote the full subcategory spanned by countably presented -frames, and to denote the finitely presented ones. Since finitely presented -frames are simply finitely presented bounded distributive lattices, we have an isomorphism .
It is well-known the dual category of is the category of finite posets. In this case, we can also have a fairly explicit description of the larger dual category of . The first observation is that any countably presented -frame is indeed a frame, i.e. it has all joins and finite meets, which distributes with each other. This is clear for all finitely presented -frames, since they are simply finite bounded distributive lattices. To see this for the countable case, consider the countably generated free -frame :
Lemma 10.1.
We have an isomorphism of -frames
where is the poset of finite subsets of .
Proof.
The free -frame can be generated by first freely adding finite meets to the discrete poset , and then freely adding all countable joins. The first step results in the poset . Now since is countable, freely adding all countable joins is equivalently freely adding all joins, which is achieved by the presheaf construction. This way,
Corollary 10.2.
There is a fully faithful embedding
preserving all countable colimits. This is again fully faithful when composed with , where is the category of topological spaces.
Proof.
Any countably presented -frame will be isomorphic to one of the form for some countably generated congruence . By Lemma 10.1, is a frame, so is any of its quotient. By Theorem 6.2.4 of Makkai and Reyes, [22], such frames are indeed spatial, hence they fully faithfully embed into the category of topological spaces via the functor . ∎
This way, we can view as a certain class of topological spaces. Notice that since all countably presented -frames will be quotients of , their dual spaces will be a subspace of , which we can compute quite easily:
Lemma 10.3.
The space of points of is Scott’s graph model , which is the countable product of the Sierpiński space .
Proof.
Note takes colimits in to limits in , since is a right adjoint. Since is the countable coproduct of the free frame on one generator, we have
Here follows from a simple computation. ∎
This way, if we write as the essential image of the fully faithful functor , its objects will all be subspaces of . We would then have the following diagram,
where here the inclusion simply takes each finite poset to its Alexandroff topological space. Hence, at the presheaf level, we have a local geometric morphism
In this case, the representable presheaf on the Sierpinski space will again be a -frame in the presheaf .
Below we discuss the corresponding admissible topologies on , modelling the various locality principles we have considered in Section 5. We encourage the readers to notice the connection between the topologies we discuss below, and the developments in Section 9.
Example 10.4 (\AxiomNT).
To model (\AxiomNT), we would want the empty sieve on to be a covering. Since is a strict initial object in , this is a subcanonical topology. Thus, this gives us a least topology that models (\AxiomNT), and we have
where is the full subcategory of excluding . The induced local geometric morphism now is given by
where, again is the full subcategory of consisting of non-empty posets, and is the classifying topos for bounded distributive lattices that are non-trivial.
Example 10.5 (\AxiomL).
The additional axiom for (\AxiomL) besides (\AxiomNT) is that
This means the dual embeddings of the following quotients need to form a covering family,
It is easy to see that is the space obtained by glueing the open points of two copies of Sierpinski spaces together, i.e. it is the following pushout:
The least topology is thus generated by the empty covering on , and the covering . Since we have the pushout above, this topology is again subcanonical.
In this case, it is not hard to give an explicit description of a covering family in : a family of maps is a -covering on iff it contains a (finite) subfamily of upward closed subsets , where ; see a similar calculation for commutative rings in MacLane and Moerdijk, [21, VIII. 6]. Following the terminology for algebraic geometry, this can be denoted as the Zariski topology, and the local geometric morphism
is over the Zariski topos for bounded distributive lattices.
Example 10.6 (\AxiomSL).
An even stronger axiom is the linearity axiom,
which requires the dual embeddings of the following quotients to form a covering family,
We can compute that , and
Again we have a pushout diagram,
where maps to the end points of , and is the two -chains in . The least topology for the linearity axiom is thus generated by the empty covering on , and . Similarly, given this pushout, will be subcanonical.
The topos is closely related to simplicial sets. This is due to Joyal’s result that the category of simplicial sets is the classifying topos of strict bounded linear orders, and thus we have an equivalence
This makes a local topos over simplicial sets.
Example 10.7 (\AxiomOneCS).
Similarly to the case of (\AxiomSL), the 1-truncation axiom requires dual embeddings of the following quotient maps
to be a covering, which translates to the fact that the three inclusions
covers . Again we have a colimit diagram as follows,
which implies is subcanonical.
Similar to the case of (\AxiomSL), is closely related to the topos of truncated simplicial sets . This is again induced by the fact that the classifying topos for 1-coskeletal bounded distributive lattices is given by
Hence, this again gives us a local geometric morphism
11. Future directions
11.1. Domain theory with quasi-coherence
In this paper we have explained the axioms of synthetic ___domain theory using the quasi-coherence principle for bounded distributive lattices and -frames. As we have seen from various examples (Propositions 4.2 and 4.5, Example 5.6 etc.) the quasi-coherence principle moreover helps with the actual computation for operations on domains. Thus, the natural step next is to further develop ___domain theory within this framework.
11.2. Relationship with Taylor’s Abstract Stone Duality
We would be remiss in failing to mention Taylor’s framework of Abstract Stone Duality [30, 29, 28], which is based on a similar duality between “algebras” and “spaces”. One apparent difference in methodology is that in Taylor’s case, the duality is of the form and is moreover required to be monadic. In our setting, the spectrum of an algebra consists not in all functions from the algebra into the dualising object but rather only the ones that preserve the structure of observational algebras—the primes in Taylor’s terminology.
Our own adjunction shall not be monadic, but a natural step toward understanding the relationship between abstract Stone duality and synthetic ___domain theory in the language of quasi-coherence might be to restrict to an adjunction involving certain more well-behaved subuniverses of . We leave this investigation for future work.
11.3. Quasi-coherence for related synthetic mathematics
Both synthetic topology [4] and synthetic computability theory [3] involve similar structures as synthetic ___domain theory. In particular, an interval object consisting of a dominance seems to be crucial in all three cases. We have showcased in Section 3 that quasi-coherence for a wide variety of theories will produce such a structure, and it will be interesting to see connections with quasi-coherence to these two types of synthetic mathematics as well.
11.4. Connection with existing models
Finally, it will be instructive to compare the sheaf models for synthetic ___domain theory constructed by Fiore and Rosolini, [13] with the models discussed in Section 10. Recall the topos constructed in loc. cit. is the following sheaf topos,
where be the full subcategory of the category of -cpos consisting of retracts of the Scott’s graph model , and is the canonical topology on . An -cpo is simply a poset having joins of countable chains, and morphisms between them are maps preserving joins of countable chains.
Here we observe that the inclusion is fully faithful, because
The first isomorphism holds because is the free -cpo on ; the second holds by formal manipulation; the third is due to the fact that is the free -frame over the discrete poset ; the final one is given by the duality between countably presented -frames and .
This leads us to conjecture that will be equivalent to the sheaf topos . However, a detailed proof of this requires us to better understand the categorical properties of .
Acknowledgements
We are grateful to Marcelo Fiore for alerting us to the connection between synthetic quasi-coherence and Blass’s observations concerning functions on universal algebras [5].
References
- Adámek and Rosicky, [1994] Adámek, J. and Rosicky, J. (1994). Locally presentable and accessible categories, volume 189. Cambridge University Press.
- Barr and Wells, [1985] Barr, M. and Wells, C. (1985). Toposes, triples, and theories. Grundlehren der mathematischen Wissenschaften; 278. Springer, New York.
- Bauer, [2006] Bauer, A. (2006). First steps in synthetic computability theory. Electronic Notes in Theoretical Computer Science, 155:5–31.
- Bauer and Taylor, [2009] Bauer, A. and Taylor, P. (2009). The Dedekind reals in abstract Stone duality. Mathematical Structures in Computer Science, 19(4):757–838.
- Blass, [1986] Blass, A. (1986). Functions on universal algebras. Journal of Pure and Applied Algebra, 42(1):25–28.
- Blechschmidt, [2020] Blechschmidt, I. (2020). A general nullstellensatz for generalized spaces. Unpublished note.
- Blechschmidt, [2021] Blechschmidt, I. (2021). Using the internal language of toposes in algebraic geometry. arXiv preprint arXiv:2111.03685.
- Buchholtz and Weinberger, [2023] Buchholtz, U. and Weinberger, J. (2023). Synthetic fibered -category theory. Higher Structures, 7:74–165.
- Caramello, [2019] Caramello, O. (2019). Denseness conditions, morphisms and equivalences of toposes. arXiv preprint arXiv:1906.08737.
- [10] Cherubini, F., Coquand, T., Geerligs, F., and Moeneclaey, H. (2024a). A foundation for synthetic Stone duality. arXiv preprint arXiv:2412.03203.
- [11] Cherubini, F., Coquand, T., and Hutzler, M. (2024b). A foundation for synthetic algebraic geometry. Mathematical Structures in Computer Science, 34(9):1008–1053.
- Fiore and Plotkin, [1996] Fiore, M. P. and Plotkin, G. D. (1996). An extension of models of axiomatic ___domain theory to models of synthetic ___domain theory. In van Dalen, D. and Bezem, M., editors, Computer Science Logic, 10th International Workshop, CSL ’96, Annual Conference of the EACSL, Utrecht, The Netherlands, September 21-27, 1996, Selected Papers, volume 1258 of Lecture Notes in Computer Science, pages 129–149. Springer.
- Fiore and Rosolini, [1997] Fiore, M. P. and Rosolini, G. (1997). Two models of synthetic ___domain theory. Journal of Pure and Applied Algebra, 116(1):151–162.
- Fiore and Rosolini, [2001] Fiore, M. P. and Rosolini, G. (2001). Domains in h. Theoretical Computer Science, 264(2):171–193.
- Gratzer et al., [2024] Gratzer, D., Weinberger, J., and Buchholtz, U. (2024). Directed univalence in simplicial homotopy type theory. arXiv preprint arXiv:2407.09146.
- Hyland, [1990] Hyland, J. M. E. (1990). First steps in synthetic ___domain theory. In Category Theory: Proceedings of the International Conference held in Como, Italy, July 22–28, 1990, pages 131–156. Springer.
- Jibladze, [1997] Jibladze, M. (1997). A presentation of the initial lift-algebra. Journal of Pure and Applied Algebra, 116(1):185–198.
- Johnstone, [2002] Johnstone, P. T. (2002). Sketches of an Elephant: A Topos Theory Compendium, volume 1, 2. Oxford University Press.
- Kelly and Power, [1993] Kelly, G. M. and Power, A. J. (1993). Adjunctions whose counits are coequalizers, and presentations of finitary enriched monads. Journal of pure and applied algebra, 89(1-2):163–179.
- Lausch and Nobauer, [2000] Lausch, H. and Nobauer, W. (2000). Algebra of polynomials. Elsevier.
- MacLane and Moerdijk, [1992] MacLane, S. and Moerdijk, I. (1992). Sheaves in geometry and logic: A first introduction to topos theory. Springer Science & Business Media.
- Makkai and Reyes, [2006] Makkai, M. and Reyes, G. E. (2006). First order categorical logic: model-theoretical methods in the theory of topoi and related categories, volume 611. Springer.
- Phoa, [1991] Phoa, W. K.-S. (1991). Domain theory in Realizability Toposes. Ph.D. Dissertation, University of Cambridge.
- Reus and Streicher, [1999] Reus, B. and Streicher, T. (1999). General synthetic ___domain theory — a logical approach. Mathematical Structures in Computer Science, 9(2):177–223.
- Riehl and Shulman, [2017] Riehl, E. and Shulman, M. (2017). A type theory for synthetic -categories. Higher Structures, 1:147–224.
- Rosolini, [1986] Rosolini, G. (1986). Continuity and Effectiveness in Topoi. PhD thesis, University of Oxford.
- Simpson, [2004] Simpson, A. (2004). Computational adequacy for recursive types in models of intuitionistic set theory. Annals of Pure and Applied Logic, 130(1):207–275. Papers presented at the 2002 IEEE Symposium on Logic in Computer Science (LICS).
- Taylor, [2002] Taylor, P. (2002). Sober spaces and continuations. Theory and Applications of Categories, 10(12):248–299.
- Taylor, [2005] Taylor, P. (2005). Inside every model of Abstract Stone Duality lies an Arithmetic Universe. Electronic Notes in Theoretical Computer Science, 122:247–296.
- Taylor, [2011] Taylor, P. (2011). Foundations for computable topology. In Sommaruga, G., editor, Foundational Theories of Classical and Constructive Mathematics, pages 265–310. Springer Netherlands, Dordrecht.
- The Univalent Foundations Program, [2013] The Univalent Foundations Program (2013). Homotopy Type Theory: Univalent Foundations of Mathematics. https://homotopytypetheory.org, Institute for Advanced Study.
- van Oosten and Simpson, [2000] van Oosten, J. and Simpson, A. K. (2000). Axioms and (counter) examples in synthetic ___domain theory. Annals of Pure and Applied Logic, 104(1):233–278.