Abstract Introduction 1 Stone duality 2 Overtly discrete spaces 3 Stone spaces 4 Compact Hausdorff spaces 5 Cohomology References

A Foundation for Synthetic Stone Duality

Felix Cherubini ORCID University of Gothenburg, Sweden
Chalmers University of Technology, Gothenburg, Sweden
Thierry Coquand ORCID University of Gothenburg, Sweden
Chalmers University of Technology, Gothenburg, Sweden
Freek Geerligs ORCID University of Gothenburg, Sweden
Chalmers University of Technology, Gothenburg, Sweden
Hugo Moeneclaey ORCID University of Gothenburg, Sweden
Chalmers University of Technology, Gothenburg, Sweden
Abstract

The language of homotopy type theory has proved to be an appropriate internal language for various higher toposes, for example for the Zariski topos in Synthetic Algebraic Geometry. This paper aims to do the same for the higher topos of light condensed anima of Dustin Clausen and Peter Scholze. This seems to be an appropriate setting for synthetic topology in the style of Martín Escardó.

We use homotopy type theory extended with 4 axioms. We prove Markov’s principle, LLPO and the negation of WLPO. Then we define a type of open propositions, inducing a topology on any type such that any map is continuous. We give a synthetic definition of second countable Stone and compact Hausdorff spaces, and show that their induced topologies are as expected. This means that any map from e.g. the unit interval 𝕀 to itself is continuous in the usual epsilon-delta sense.

With the usual definition of cohomology in homotopy type theory, we show that H1(S,)=0 for S Stone and that H1(X,) for X compact Hausdorff can be computed using Čech cohomology. We use this to prove H1(𝕀1,)=0 and H1(𝕊1,)= where 𝕊1 is the set /. As an application, we give a synthetic proof of Brouwer’s fixed-point theorem.

Keywords and phrases:
Homotopy Type Theory, Synthetic Topology, Cohomology
Copyright and License:
[Uncaptioned image] © Felix Cherubini, Thierry Coquand, Freek Geerligs, and Hugo Moeneclaey; licensed under Creative Commons License CC-BY 4.0
2012 ACM Subject Classification:
Theory of computation Type theory
Acknowledgements:
The idea of proving Theorem 4.18 by using the topological characterization of Stone spaces as totally disconnected compact Hausdorff spaces was explained to us by Martín Escardó. We profited a lot from a discussion with Reid Barton and Johann Commelin. The fact that Markov’s principle (Corollary 1.16) holds was noticed by David Wärn. We had an interesting discussion with Bas Spitters on the topic of the article at TYPES 2024.
Funding:
ForCUTT project, ERC advanced grant 101053291.
Editors:
Rasmus Ejlers Møgelberg and Benno van den Berg

Introduction

The language of homotopy type theory consists of dependent type theory enriched with the univalence axiom and higher inductive types. It has proven exceptionnally well-suited to a synthetic development of homotopy theory [13]. It also provides a framework precise enough to analyze categorical models of type theory [20]. Moreover, arguments in this language can be represented in proof assistants rather directly. In this article we use homotopy type theory to give a synthetic development of topology, analogous to the synthetic development of algebraic geometry in [4].

We introduce four axioms inspired by the light condensed sets introduced in [5]. Interestingly, our axioms have strong connections with constructive mathematics [3], in particular constructive reverse mathematics [11, 7]. Indeed they imply several of Brouwer’s principles (e.g. any real function on the unit interval is continuous, the celebrated fan theorem), as well as not intuitionistically valid principles (Markov’s Principle, the so-called Lesser Limited Principle of Omniscience).

Our axioms also closely align with the program of Synthetic Topology [9, 12, 19, 10, 21]. Indeed we have a dominance of open propositions, so that any type comes with an induced topology. Using this induced topology, we manage to capture synthetically the notion of second-countable compact Hausdorff spaces. While working on our axioms, we learnt about [2], which provides a similar axiomatisation in extensional type theory. We show that some of their axioms are consequences of ours. For example111We can actually prove all of their axioms, from which their directed univalence follows. This will be presented in a following paper., we can define in our setting the notion of overtly discrete types, which is dual to the notion of compact Hausdorff spaces.

A central theme of homotopy type theory is that the notion of type is more general than the notion of set. We illustrate this theme in this work. Indeed we can form the types of Stone spaces and of compact Hausdorff spaces, which are not sets but rather a groupoids. Moreover these spaces are closed under Σ-type types, which would be impossible to formulate in the traditional setting. Additionally, we can leverage higher types by using the elegant definition of cohomology groups in homotopy type theory [13]. We then prove a special case of a theorem of Dyckhoff [8] describing the cohomology of compact Hausdorff spaces. As an application, we give a synthetic proof of Brouwer’s fixed point theorem, similar to the proof of an approximated form in [16].

We expect our axioms to be validated by the interpretation of homotopy type theory into the higher topos of light condensed anima [17], although checking this rigorously is still work in progress. We even expect this to be valid in a constructive metatheory, using [6]. It is important to stress that our axioms only capture the properties of light condensed anima that are internally valid. Since David Wärn [23] has proved that an important property of condensed abelian groups is not valid internally, this means that we cannot prove it in our setting. We also conjecture that the present axiom system is complete for the properties that are internally valid.

1 Stone duality

1.1 Preliminaries

 Remark 1.1.

For X any type, a subtype U of X is a family of propositions over X. We write UX. If X is a set, we call U a subset. Given x:X we sometimes write xU instead of U(x). For subtypes A,BX, we write AB for pointwise implication. We will freely switch between a subtype UX and the corresponding embedding x:XU(x)X. In particular, if we write x:U we mean x:X such that U(x).

Definition 1.2.

A type is countable if and only if it is merely equal to some decidable subset of .

Definition 1.3.

For I a set we write 2[I] for the free Boolean algebra on I. A Boolean algebra B is countably presented if there exist countable sets I,J with generators g:IB and relations f:J2[I] such that g induces an equivalence between 2[I]/(fj)j:J and B.

 Remark 1.4.

Any countably presented algebra is merely of the form 2[]/(rn)n:.

 Remark 1.5.

We denote the type of countably presented Boolean algebras by 𝖡𝗈𝗈𝗅𝖾ω. This type does not depend on a choice of universe. Moreover 𝖡𝗈𝗈𝗅𝖾ω has a natural category structure.

Example 1.6.

If both the set of generators and relations are empty, we get the Boolean algebra 2. Its underlying set is {0,1} with 01. We have that 2 is initial in 𝖡𝗈𝗈𝗅𝖾ω.

Definition 1.7.

For B a countably presented Boolean algebra, we define the spectrum Sp(B) as the set Hom(B,2) of Boolean morphisms from B to 2. Any type which is merely equivalent to some spectrum is called a Stone space.

Example 1.8.
  1. (i)

    There is only one Boolean morphism from 2 to 2, thus Sp(2) is the singleton type .

  2. (ii)

    The trivial Boolean algebra is presented as 2/(1). We have 0=1 in the trivial Boolean algebra, so there cannot be a map from it into 2 preserving both 0 and 1. Therefore the corresponding Stone space is the empty type .

  3. (iii)

    The type Sp(2[]) is called the Cantor space. It is equivalent to the set of binary sequences 2. Given α:Sp(2[]) and n:, we write αn for α(gn), the n-th bit of the corresponding binary sequence.

  4. (iv)

    We denote by B the Boolean algebra generated by (gn)n: quotiented by the relations gmgn=0 for nm. A morphism B2 corresponds to a function 2 that hits 1 at most once. We denote Sp(B) by . For α: and n: we write αn for α(gn). For n:, we define n: as the unique α: such that αn=1. We define : as the unique α: such that αn=0 for all n:.

    By conjunctive normal form, any element of B can be written uniquely as i:Ign or as i:I¬gn for some finite I.

Lemma 1.9.

Given α:2, we have an equivalence of propositions:

(n:αn=0)Sp(2/(αn)n:).

Proof.

There is only one Boolean morphism x:22, and it satisfies x(αn)=0 for all n: if and only if αn=0 for all n:.

1.2 Axioms

Axiom 1 (Stone duality).

For all B:𝖡𝗈𝗈𝗅𝖾ω, the evaluation map B2Sp(B) is an isomorphism.

Axiom 2 (Surjections are formal surjections).

For all morphism g:BC in 𝖡𝗈𝗈𝗅𝖾ω, we have that g is injective if and only if ()g:Sp(C)Sp(B) is surjective.

Axiom 3 (Local choice).

For all B:𝖡𝗈𝗈𝗅𝖾ω and type family P over Sp(B) such that Πs:Sp(B)P(s), there merely exists some C:𝖡𝗈𝗈𝗅𝖾ω and surjection q:Sp(C)Sp(B) such that Πt:Sp(C)P(q(t)).

Axiom 4 (Dependent choice).

For all types (En)n: with surjections En+1En for all n:, the projection from the sequential limit limkEk to E0 is surjective.

1.3 Anti-equivalence of 𝖡𝗈𝗈𝗅𝖾ω and 𝖲𝗍𝗈𝗇𝖾

By Axiom 1, the map Sp is an embedding of 𝖡𝗈𝗈𝗅𝖾ω into any universe of types. We denote its image by 𝖲𝗍𝗈𝗇𝖾.

 Remark 1.10.

Stone spaces will take over the role of the affine schemes from [4], so let us repeat some results here. Analogously to Lemma 3.1.2 of [4], for X:𝖲𝗍𝗈𝗇𝖾, Axiom 1 tells us that X=Sp(2X). Proposition 2.2.1 of [4] now says that Sp gives a natural equivalence

Hom(A,B)=(Sp(B)Sp(A))

By the above and Lemma 9.4.5 of [13], the map Sp defines a dual equivalence of categories between 𝖡𝗈𝗈𝗅𝖾ω and 𝖲𝗍𝗈𝗇𝖾. In particular the spectrum of any colimit in 𝖡𝗈𝗈𝗅𝖾ω is the limit of the spectrum of the opposite diagram.

 Remark 1.11.

Axiom 3 can also be formulated as follows: Given S:𝖲𝗍𝗈𝗇𝖾 with E,F arbitrary types, a map f:SF and a surjection e:EF, there exists a Stone space T, a surjective map TS and an arrow TE making the following diagram commute:

Lemma 1.12.

For B:𝖡𝗈𝗈𝗅𝖾ω, we have 0=B1 if and only if ¬Sp(B).

Proof.

If 0=B1, there is no map in B2 preserving both 0 and 1, thus ¬Sp(B). Conversely, if ¬Sp(B) then Sp(B)=. Since is the spectrum of the trivial Boolean algebra and Sp is an embedding, we conclude that B is the trivial Boolean algebra, hence 0=B1.

Corollary 1.13.

For S:𝖲𝗍𝗈𝗇𝖾, we have that ¬¬SS

Proof.

Let B:𝖡𝗈𝗈𝗅𝖾ω and suppose ¬¬Sp(B). By Lemma 1.12 we have that 0B1, therefore the morphism 2B is injective. By Axiom 2 the map Sp(B)Sp(2) is surjective, thus Sp(B) is merely inhabited.

1.4 Principles of omniscience

The so-called principles of omniscience are all weaker than the law of excluded middle (LEM), and help measure how close a logical system is to satisfying LEM [7, 11]. In this section, we will show that two such principles hold (MP and LLPO), and that another one fails (WLPO).

Theorem 1.14 (The negation of the weak lesser principle of omniscience (¬WLPO)).
¬α:2((n:αn=0)¬(n:αn=0))

Proof.

We will prove that any decidable property of binary sequences is determined by a finite prefix of fixed length, contradicting n:αn=0 being decidable for all α. Indeed assume f:22 such that f(α)=0 if and only if n:αn=0. By Axiom 1, there is some c:2[] with f(α)=0 if and only if α(c)=0. There exists k: such that c is expressed in terms of the generators (gn)nk. Now consider β,γ:2 given by β(gn)=0 for all n: and γ(gn)=0 if and only if nk. As β and γ are equal on (gn)nk, we have β(c)=γ(c). However, f(β)=0 and f(γ)=1, giving a contradiction.

Theorem 1.15.

For all α:, we have that

(¬(n:αn=0))Σn:αn=1

Proof.

By Lemma 1.9, we have that ¬(n:αn=0) implies that Sp(2/(αn)n:) is empty. Hence 2/(αn)n: is trivial by Lemma 1.12. Then there exists k: such that ikαi=1. As αi=1 for at most one i:, there exists a unique n: with αn=1.

Corollary 1.16 (Markov’s principle (MP)).

For all α:2, we have that

(¬(n:αn=0))Σn:αn=1

Proof.

Given α:2, consider the sequence α: satisfying αn=1 if and only if n is minimal with αn=1. Then apply the above theorem.

Theorem 1.17 (The lesser limited principle of omniscience (LLPO)).

For all α:, we have that

(k:α2k=0)(k:α2k+1=0)

Proof.

Define f:BB×B on generators as follows

f(gn)={(gk,0) if n=2k(0,gk) if n=2k+1

Note that f is a well-defined morphism in 𝖡𝗈𝗈𝗅𝖾ω as f(gn)f(gm)=0 whenever mn. We claim that f is injective. If I, write I0={k| 2kI},I1={k| 2k+1I}. Recall that any x:B is of the form iIgi or iI¬gi for some finite set I.

  • If x=iIgi, then f(x)=(iI0gi,iI1gi). So if f(x)=0, then I0=I1=I= and x=0.

  • Suppose x=iI¬gi. Then f(x)=(iI0¬gi,iI1¬gi), so f(x)0.

By Axiom 2, we have that f corresponds to a surjection s:+. Thus for α:, there exists some x:+ such that s(x)=α. If x=inl(β), then for any k: we have that

α2k+1=s(x)2k+1=x(f(g2k+1))=inl(β)(0,gk)=β(0)=0.

Similarly, if x=inr(β), we have that α2k=0 for all k:. The surjection s:+ above does not have a section. Indeed:

Lemma 1.18.

The function f defined above does not have a retraction.

Proof.

Suppose r:B×BB is a retraction of f. Then r(0,gk)=g2k+1 and r(gk,0)=g2k. Note that r(0,1)r(0,gk)=g2k+1 for all k:. As a consequence, r(0,1) is of the form iI¬gi for some finite set I. By similar reasoning so is r(1,0). But then

r(0,1)r(1,0)=r((1,0)(0,1))=r(0,0)=0.

This is a contradiction.

1.5 Open and closed propositions

Open (resp. closed) propositions are defined as countable disjunctions (resp. conjunctions) of decidable propositions. In this section we will study their logical properties.

Definition 1.19.

A proposition P is open (resp. closed) if there exists some α:2 such that Pn:αn=0 (resp. Pn:αn=0). We denote by 𝖮𝗉𝖾𝗇 and 𝖢𝗅𝗈𝗌𝖾𝖽 the types of open and closed propositions.

 Remark 1.20.

The negation of an open proposition is closed, and by MP (Corollary 1.16), the negation of a closed proposition is open. Moreover both open and closed propositions are ¬¬-stable. By ¬WLPO (Theorem 1.14), not every closed proposition is decidable. Therefore, not every open proposition is decidable. Every decidable proposition is both open and closed.

Lemma 1.21.

We have the following:

  • Closed propositions are stable under countable conjunctions and finite disjunctions.

  • Open propositions are stable under countable disjunctions and finite conjunctions.

Proof.

All statements but the one about finite disjunctions have similar proofs, so we only present the proof that closed propositions are stable under countable conjunctions. Let (Pn)n: be a countable family of closed propositions. By countable choice, for each n: we have an αn:2 such that Pnm:αn,m=0. Consider a surjection s:×, and let βk=αs(k). Note that k:βk=0 if and only if n:Pn.

To prove that closed propositions are closed under finite disjunctions, we use the known fact that LLPO (Theorem 1.17) is equivalent to the statement that for P and Q open, we have that (¬P¬Q)¬(PQ). We conclude using that closed propositions are negations of open propositions, and that the conjunction of two open propositions is open.

From now on we will use the above properties silently.

Corollary 1.22.

If a proposition is both open and closed, then it is decidable.

Proof.

If P is open and closed, then P¬P is open. So it is ¬¬-stable and we conclude from ¬¬(P¬P).

Lemma 1.23.

For (Pn)n: a sequence of closed propositions, we have ¬n:Pnn:¬Pn.

Proof.

Both ¬n:Pn and n:¬Pn are open, hence ¬¬-stable. The equivalence follows.

Lemma 1.24.

If P is open and Q is closed then PQ is closed. If P is closed and Q open, then PQ is open.

Proof.

Note that ¬PQ is closed. Using ¬¬-stability we conclude that (PQ)(¬PQ). The other proof is similar.

1.6 Types as spaces

The subset 𝖮𝗉𝖾𝗇 of the set of propositions induces a topology on every type. This is the viewpoint taken in synthetic topology, from which we borrow terminology [9, 12].

Definition 1.25.

Let T be a type, and let AT be a subtype. We call AT open (resp. closed) if A(t) is open (resp. closed) for all t:T.

 Remark 1.26.

It follows immediately that the pre-image of an open by any map is open, so that any map is continuous. In Theorem 3.11, we will see that the resulting topology is as expected for Stone spaces. In Lemma 4.27, we will see that the same holds for the unit interval.

2 Overtly discrete spaces

Definition 2.1.

We call a type overtly discrete if it is a sequential colimit of finite sets.

 Remark 2.2.

It follows from Corollary 7.7 of [18] that overtly discrete types are sets, and that the sequential colimit can be defined as in set theory. We write 𝖮𝖣𝗂𝗌𝖼 for the type of overtly discrete types.

Using dependent choice, we have the following results:

Lemma 2.3.

A map between overtly discrete sets is a sequential colimit of maps between finite sets.

Lemma 2.4.

For f:AB a sequential colimit of maps of finite sets fn:AnBn, we have that the factorisation AIm(f)B is the sequential colimit of the factorisations AnIm(fn)Bn.

Corollary 2.5.

An injective (resp. surjective) map between overtly discrete types is a sequential colimit of injective (resp. surjective) maps between finite sets.

2.1 Closure properties of 𝖮𝖣𝗂𝗌𝖼

We can get the following result using Lemma 2.3 and dependent choice.

Lemma 2.6.

Overtly discrete types are stable under sequential colimits.

We have that Σ-types, identity types and propositional truncation commute with sequential colimits (Theorem 5.1, Theorem 7.4 and Corollary 7.7 in [18]). Then by closure of finite sets under these constructors, we can get the following:

Lemma 2.7.

Overtly discrete types are stable under Σ-types, identity types and propositional truncations.

2.2 𝖮𝗉𝖾𝗇 and 𝖮𝖣𝗂𝗌𝖼

Lemma 2.8.

A proposition is open if and only if it is overtly discrete.

Proof.

If P is overtly discrete, then Pn:Fn with Fn finite sets. But a finite set being inhabited is decidable, hence P is a countable disjunction of decidable propositions, so it is open. Suppose Pn:αn=1. Let Pn=nk(αn=1), which is a decidable proposition, hence a finite set. Then the colimit of Pn is P.

Corollary 2.9.

Open propositions are stable under Σ-types.

Corollary 2.10 (transitivity of openness).

Let T be a type, let VT open and let WV open. Then WT is open as well.

 Remark 2.11.

It follows from Proposition 2.25 of [12] that 𝖮𝗉𝖾𝗇 is a dominance in the setting of Synthetic Topology.

Lemma 2.12.

A type B is overtly discrete if and only if it is the quotient of a countable set by an open equivalence relation.

Proof.

If B:𝖮𝖣𝗂𝗌𝖼 is the sequential colimit of finite sets Bn, then B is an open quotient of (Σn:Bn). Conversely, assume B=D/R with D decidable and R open. By dependent choice we get α:DD2 such that R(x,y)k:αx,y(k)=1. Define Dn=(Dn), and define Rn:DnDn2 as the equivalence relation generated by the relation knαx,y(k)=1. Then the Bn=Dn/Rn are finite sets, and their colimit is B.

2.3 Relating 𝖮𝖣𝗂𝗌𝖼 and 𝖡𝗈𝗈𝗅𝖾ω

Lemma 2.13.

Every countably presented Boolean algebra is a sequential colimit of finite Boolean algebras.

Proof.

Consider a countably presented Boolean algebra of the form B=2[]/(rn)n:. For each n:, let Gn be the union of {gi|in} and the finite set of generators occurring in ri for some in. Denote Bn=2[Gn]/(ri)in. Each Bn is a finite Boolean algebra, and there are canonical maps BnBn+1. Then B is the colimit of this sequence.

Corollary 2.14.

A Boolean algebra B is overtly discrete if and only if it is countably presented.

Proof.

Assume B:𝖮𝖣𝗂𝗌𝖼. By Lemma 2.12, we get a surjection B and that B has open equality. Consider the induced surjective morphism f:2[]B. By countable choice, we get for each b:2[] a sequence αb:2 such that (f(b)=0)k:(αb,k=1). Consider r:2[]2[] given by

r(b,k)={b if αb(k)=10 if αb(k)=0

Then B=2[]/(r(b,k))b:2[],k:. The converse comes from Lemma 2.13.

 Remark 2.15.

By Lemma 2.7 and Corollary 2.14, it follows that any g:BC in 𝖡𝗈𝗈𝗅𝖾ω has an overtly discrete kernel. As a consequence, the kernel is enumerable and B/Ker(g) is in 𝖡𝗈𝗈𝗅𝖾ω. By uniqueness of epi-mono factorizations and Axiom 2, the factorization BB/Ker(g)C corresponds to Sp(C)Sp(B/Ker(g))Sp(B).

 Remark 2.16.

Similarly to Lemma 2.3 and Lemma 2.4, a (resp. surjective, injective) morphism in 𝖡𝗈𝗈𝗅𝖾ω is a sequential colimit of (resp. surjective, injective) morphisms between finite Boolean algebras.

3 Stone spaces

3.1 Stone spaces as profinite sets

Here we present Stone spaces as sequential limits of finite sets. This is the perspective taken in Condensed Mathematics [15, 1, 5]. Some of the results in this section are versions of the axioms used in [2]. A full proof of all these axioms is part of future work.

Lemma 3.1.

Any S:𝖲𝗍𝗈𝗇𝖾 is a sequential limit of finite sets.

Proof.

Assume B:𝖡𝗈𝗈𝗅𝖾ω. By Remark 1.10 and Lemma 2.13, we have that Sp(B) is a sequential limit of spectra of finite Boolean algebras, which are finite sets.

Lemma 3.2.

A sequential limit of finite sets is a Stone space.

Proof.

By Remark 1.10 and Lemma 2.6, we have that 𝖲𝗍𝗈𝗇𝖾 is closed under sequential limits, and finite sets are Stone.

Corollary 3.3.

Stone spaces are stable under finite limits.

 Remark 3.4.

By Remark 2.16 and Axiom 2, maps (resp. surjections, injections) of Stone spaces are sequential limits of maps (resp. surjections, injections) of finite sets.

Lemma 3.5.

For (Sn)n: a sequence of finite types with S=limnSn and k:, we have that Fin(k)S is the sequential colimit of Fin(k)Sn.

Proof.

By Remark 1.10 we have Fin(k)S=Hom(2k,2S). Since 2k is finite, we have that Hom(2k,_) commutes with sequential colimits, therefore Hom(2k,2S) is the sequential colimit of Hom(2k,2Sn). By applying Remark 1.10 again, the latter type is Fin(k)Sn.

Lemma 3.6.

For S:𝖲𝗍𝗈𝗇𝖾 and f:S, there exists some k: such that f factors through Fin(k).

Proof.

For each n:, the fiber of f over n is a decidable subset fn:S2. We must have that Sp(2S/(fn)n:)=, hence there exists some k: with nkfn=2S1. It follows that f(s)k for all s:S as required.

Corollary 3.7.

For (Sn)n: a sequence of finite types with S=limnSn, we have that S is the sequential colimit of Sn.

Proof.

By Lemma 3.6 we have that S is the sequential colimit of Fin(k)S. By Lemma 3.5, Fin(k)S is the sequential colimit of the Fin(k)Sn and we can swap the sequential colimits to conclude.

3.2 𝖢𝗅𝗈𝗌𝖾𝖽 and 𝖲𝗍𝗈𝗇𝖾

Corollary 3.8.

For all S:𝖲𝗍𝗈𝗇𝖾, the proposition S is closed.

Proof.

By Lemma 1.12, ¬S is equivalent to 0=2S1, which is open by Lemma 2.13 and Lemma 2.12. Hence ¬¬S is a closed proposition, and by Corollary 1.13, so is S.

Corollary 3.9.

A proposition P is closed if and only if it is a Stone space.

Proof.

By the above, if S is both a Stone space and a proposition, it is closed. By Lemma 1.9, any closed proposition is Stone.

Lemma 3.10.

For all S:𝖲𝗍𝗈𝗇𝖾 and s,t:S, the proposition s=t is closed.

Proof.

Suppose S=Sp(B) and let G be a countable set of generators for B. Then s=t if and only if s(g)=t(g) for all g:G. So s=t is a countable conjunction of decidable propositions, hence closed.

3.3 The topology on Stone spaces

Theorem 3.11.

Let AS be a subset of a Stone space. The following are equivalent:

  1. (i)

    There exists a map α:S2 such that A(x)n:αx,n=0 for any x:S.

  2. (ii)

    There exists a family (Dn)n: of decidable subsets of S such that A=n:Dn.

  3. (iii)

    There exists a Stone space T and some embedding TS whose image is A.

  4. (iv)

    There exists a Stone space T and some map TS whose image is A.

  5. (v)

    A is closed.

Proof.

  • (i)(ii). Dn and α can be defined from each other by Dn(x)(αx,n=0). Then observe that

    xn:Dnn:(αx,n=0).
  • (ii)(iii). Let S=Sp(B). By Axiom 1, we have (dn)n: in B such that Dn={x:S|x(dn)=0}. Let C=B/(dn)n:. Then Sp(C)S is as desired because:

    Sp(C)={x:S|n:x(dn)=0}=n:Dn.
  • (iii)(iv). Immediate.

  • (iv)(ii). Assume f:TS corresponds to g:BC in 𝖡𝗈𝗈𝗅𝖾ω. By Remark 2.15, f(T)=Sp(B/Ker(g)) and there exists a surjection d:Ker(g). For n:, we denote by Dn the decidable subset of S corresponding to dn. Then we have that Sp(B/Ker(g))=n:Dn.

  • (i)(v). By definition.

  • (v)(iv). We have a surjection 2𝖢𝗅𝗈𝗌𝖾𝖽 defined by αn:αn=0. Remark 1.11 gives us that there merely exists T,e,β as follows:

    Define B(x)n:βx,n=0. As (i)(iii) by the above, B is the image of some Stone space. Note that A is the image of B, thus A is the image of some Stone space.

Corollary 3.12.

Closed subtypes of Stone spaces are Stone.

Corollary 3.13.

For S:𝖲𝗍𝗈𝗇𝖾 and AS closed, we have that x:SA(x) is closed.

Proof.

By Corollary 3.12, we have that Σx:SA(x) is Stone, so its truncation is closed by Corollary 3.8.

Corollary 3.14.

Closed propositions are closed under sigma types.

Proof.

Let P:𝖢𝗅𝗈𝗌𝖾𝖽 and Q:P𝖢𝗅𝗈𝗌𝖾𝖽. Then Σp:PQ(p)p:PQ(p). As P is Stone by Corollary 3.9, Corollary 3.13 gives that Σp:PQ(p) is closed.

 Remark 3.15.

Analogously to Corollary 2.10 and Remark 2.11, it follows that closedness is transitive and 𝖢𝗅𝗈𝗌𝖾𝖽 forms a dominance.

Lemma 3.16.

Assume S:𝖲𝗍𝗈𝗇𝖾 with F,G:S𝖢𝗅𝗈𝗌𝖾𝖽 such that FG=. Then there exists a decidable subset D:S2 such FD,G¬D.

Proof.

Assume S=Sp(B). By Theorem 3.11, for all n: there is fn,gn:B such that xF if and only if n:x(fn)=0 and yG if and only if n:y(gn)=0. Denote by h the sequence defined by h2k=fk and h2k+1=gk. Then Sp(B/(hk)k:)=FG=, so by Lemma 1.12 there exists finite sets I,J such that 1=B((i:Ifi)(j:Jgj)). If yF, then y(fi)=0 for all i:I, hence y(j:Jgj)=1 If xG, we have x(j:Jgj)=0. Thus we can define the required D by D(x)x(j:Jgj)=1.

4 Compact Hausdorff spaces

Definition 4.1.

A type X is called a compact Hausdorff space if its identity types are closed propositions and there exists some S:𝖲𝗍𝗈𝗇𝖾 with a surjection SX. We write 𝖢𝖧𝖺𝗎𝗌 for the type of compact Hausdorff spaces.

4.1 Topology on compact Hausdorff spaces

Lemma 4.2.

Let X:𝖢𝖧𝖺𝗎𝗌, S:𝖲𝗍𝗈𝗇𝖾 and q:SX surjective. Then AX is closed if and only if it is the image of a closed subset of S by q.

Proof.

As q is surjective, we have q(q1(A))=A. If A is closed, so is q1(A) and hence A is the image of a closed subset of S. Conversely, let BS be closed. Then xq(B) if and only if

s:S(B(s)q(s)=x).

Hence by Corollary 3.13, q(B) is closed.

The next two corollaries mean that compact Hausdorff spaces are compact in the sense of Synthetic Topology.

Corollary 4.3.

Assume given X:𝖢𝖧𝖺𝗎𝗌 with AX closed. Then x:XA(x) is closed, and equivalent to A.

Proof.

From Lemma 4.2 and Theorem 3.11, it follows that AX is closed if and only if it is the image of a map TX for some T:𝖲𝗍𝗈𝗇𝖾. Then x:XA(x) if and only T, which is closed by Corollary 3.8. Therefore x:XA(x) is ¬¬-stable and equivalent to A.

Corollary 4.4.

Assume given X:𝖢𝖧𝖺𝗎𝗌 with UX open. Then x:XU(x) is open.

The next lemma means that compact Hausdorff spaces are not too far from being compact in the classical sense.

Lemma 4.5.

Given X:𝖢𝖧𝖺𝗎𝗌 and Cn:X𝖢𝗅𝗈𝗌𝖾𝖽 closed subsets such that n:Cn=, there is some k: with nkCn=.

Proof.

By Lemma 4.2 it is enough to prove the result when X is Stone, and by Theorem 3.11 we can assume Cn decidable. So assume X=Sp(B) and cn:B such that

Cn={x:B2|x(cn)=0}.

Then we have that

Sp(B/(cn)n:)n:Cn=.

Hence 0=1 in B/(cn)n: and there is some k: with nkcn=1, which means that

=Sp(B/(cn)nk)nkCn

as required.

Corollary 4.6.

Let X,Y:𝖢𝖧𝖺𝗎𝗌 and f:XY. Suppose (Gn)n: is a decreasing sequence of closed subsets of X. Then f(n:Gn)=n:f(Gn).

Proof.

It is always the case that f(n:Gn)n:f(Gn). For the converse direction, suppose that yf(Gn) for all n:. We define FX closed by F=f1(y). Then for all n: we have that FGn is non-empty. By Lemma 4.5 this implies that n:(FGn). By Corollary 4.3, we have that n:(FGn) is merely inhabited. Thus yf(n:Gn) as required.

Corollary 4.7.

Let AX be a subset of a compact Hausdorff space and p:SX be a surjective map with S:𝖲𝗍𝗈𝗇𝖾. Then A is closed (resp. open) if and only if there exists a sequence (Dn)n: of decidable subsets of S such that A=n:p(Dn) (resp. A=n:¬p(Dn)).

Proof.

The characterization of closed subsets follows from characterization (ii) in Theorem 3.11, Lemma 4.2 and Corollary 4.6. To deduce the characterization of open subsets we use Remark 1.20 and Lemma 1.23.

 Remark 4.8.

For S:𝖲𝗍𝗈𝗇𝖾, there is a surjection 2S. It follows that for any X:𝖢𝖧𝖺𝗎𝗌 there is a surjection from to a basis of X. Classically this means that X is second countable.

The next lemma means that compact Hausdorff spaces are normal.

Lemma 4.9.

Assume X:𝖢𝖧𝖺𝗎𝗌 and A,BX closed such that AB=. Then there exist U,VX open such that AU, BV and UV=.

Proof.

Let q:SX be a surjective map with S:𝖲𝗍𝗈𝗇𝖾. As q1(A) and q1(B) are closed, by Lemma 3.16, there is some D:S2 such that q1(A)D and q1(B)¬D. Note that q(D) and q(¬D) are closed by Lemma 4.2. As q1(A)¬D=, we have that A¬q(¬D):=U. Similarly B¬q(D):=V. Then U and V are disjoint because ¬q(D)¬q(¬D)=¬(q(D)q(¬D))=¬X=.

4.2 Compact Hausdorff spaces are stable under sigma types

Lemma 4.10.

A type X is Stone if and only if it is merely a closed subset of 2.

Proof.

By Remark 1.4, any B:𝖡𝗈𝗈𝗅𝖾ω can be written as 2[]/(rn)n:. By Remark 2.15, the quotient map induces an embedding Sp(B)Sp(2[])=2, which is closed by Theorem 3.11.

Lemma 4.11.

Compact Hausdorff spaces are stable under Σ-types.

Proof.

Assume X:𝖢𝖧𝖺𝗎𝗌 and Y:X𝖢𝖧𝖺𝗎𝗌. By Corollary 3.14 we have that identity types in Σx:XY(x) are closed. By Lemma 4.10 we know that for any x:X there merely exists a closed C2 with a surjection Σα:2C(α)Y(x). By local choice we merely get S:𝖲𝗍𝗈𝗇𝖾 with a surjection p:SX such that for all s:S we have Cs2 closed and a surjection Σ2CsY(p(s)). This gives a surjection Σs:S,α:2Cs(α)Σx:XYx and the source is Stone by Remark 3.4 and Corollary 3.12.

4.3 Stone spaces are stable under sigma types

We will show that Stone spaces are precisely totally disconnected compact Hausdorff spaces. We will use this to prove that a sigma type of Stone spaces is Stone.

Lemma 4.12.

Assume X:𝖢𝖧𝖺𝗎𝗌, then 2X is countably presented.

Proof.

There is some surjection q:SX with S:𝖲𝗍𝗈𝗇𝖾. This induces an injection of Boolean algebras 2X2S. Note that a:S2 lies in 2X if and only if:

s,t:Sq(s)=Xq(t)a(s)=a(t).

As equality in X is closed and equality in 2 is decidable, the implication is open for every s,t:S. By Corollary 4.4, we conclude that 2X is an open subalgebra of 2S. Therefore, it is in 𝖮𝖣𝗂𝗌𝖼 by Lemma 2.8 and Lemma 2.7 and in 𝖡𝗈𝗈𝗅𝖾ω by Corollary 2.14.

Definition 4.13.

For all X:𝖢𝖧𝖺𝗎𝗌 and x:X, we define Qx the connected component of x as the intersection of all DX decidable such that xD.

Lemma 4.14.

For all X:𝖢𝖧𝖺𝗎𝗌 with x:X, we have that Qx is a countable intersection of decidable subsets of X.

Proof.

By Lemma 4.12, we can enumerate the elements of 2X, say as (Dn)n:. For n: we define En as Dn if xDn and X otherwise. Then n:En=Qx.

Lemma 4.15.

Assume X:𝖢𝖧𝖺𝗎𝗌 with x:X and suppose UX open with QxU. Then we have some decidable EX with xE and EU.

Proof.

By Lemma 4.14, we have Qx=n:Dn with DnX decidable. If QxU, then

Qx¬U=n:(Dn¬U)=.

By Lemma 4.5 there is some k: with

(nkDn)¬U=nk(Dn¬U)=.

Therefore nkDn¬¬U. As U is open, ¬¬U=U and E:=nkDn is as desired.

Lemma 4.16.

Assume X:𝖢𝖧𝖺𝗎𝗌 with x:X. Then any map in Qx2 is constant.

Proof.

Assume Qx=AB with A,B decidable and disjoint subsets of Qx. Assume xA. By Lemma 4.14, QxX is closed. Using Remark 3.15, it follows that A,BX are closed and disjoint. By Lemma 4.9 there exist U,V disjoint open such that AU and BV. By Lemma 4.15 we have a decidable D such that QxDUV. Note that E:=DU=D(¬V) is clopen, hence decidable by Corollary 1.22. But xE, hence BQxE but BE=, hence B=.

Lemma 4.17.

Let X:𝖢𝖧𝖺𝗎𝗌, then X is Stone if and only x:XQx={x}.

Proof.

By Axiom 1, it is clear that for all x:S with S:𝖲𝗍𝗈𝗇𝖾 we have that Qx={x}. Conversely, assume X:𝖢𝖧𝖺𝗎𝗌 such that x:XQx={x}. We claim that the evaluation map e:XSp(2X) is both injective and surjective, hence an equivalence. Let x,y:X be such that e(x)=e(y), i.e. such that f(x)=f(y) for all f:2X. Then yQx, hence x=y by assumption. Thus e is injective. Let q:SX be a surjective map. It induces an injection 2X2S, which by Axiom 2 induces a surjection p:Sp(2S)Sp(2X). Note that eq is equal to p so e is surjective.

Theorem 4.18.

Assume S:𝖲𝗍𝗈𝗇𝖾 and T:S𝖲𝗍𝗈𝗇𝖾. Then Σx:ST(x) is Stone.

Proof.

By Lemma 4.11 we have that Σx:ST(x) is a compact Hausdorff space. By Lemma 4.17 it is enough to show that for all x:S and y:T(x) we have that Q(x,y) is a singleton. Assume (x,y)Q(x,y), then for any map f:S2 we have that:

f(x)=fπ1(x,y)=fπ1(x,y)=f(x)

so that xQx and since S is Stone, by Lemma 4.17 we have that x=x. Therefore we have Q(x,y){x}×T(x). Assume z,z:Q(x,y), then for any map g:T(x)2 we have that g(z)=g(z) by Lemma 4.16. Since T(x) is Stone, we conclude z=z by Lemma 4.17.

4.4 The unit interval as a compact Hausdorff space

Since we have dependent choice, the unit interval 𝕀=[0,1] can be defined using Cauchy reals or Dedekind reals. We can freely use results from constructive analysis [3]. As we have ¬WLPO, MP and LLPO, we can use the results from constructive reverse mathematics that follow from these principles [11, 7].

Definition 4.19.

We define for each n: the Stone space 2n of binary sequences of length n. And we define csn:2n by csn(α)=i<nα(i)2i+1. Finally we write n for the binary relation on 2n given by αnβ|csn(α)csn(β)|12n.

 Remark 4.20.

The inclusion Fin(n) induces a restriction _|n:22n for each n:.

Definition 4.21.

We define cs:2𝕀 as cs(α)=i=0α(i)2i+1.

Theorem 4.22.

The type 𝕀 is a compact Hausdorff space.

Proof.

By LLPO, we have that cs is surjective. Note that cs(α)=cs(β) if and only if for all n: we have α|nnβ|n. This is a countable conjunction of decidable propositions, so that equality in 𝕀 is closed.

The following is also given by Definitions 2.7 and 2.10 of [3].

Definition 4.23.

Assume given x,y:𝕀 and α,β:2 such that x=cs(α),y=cs(β). Then x<y is the proposition n:csn(α)+12n<csn(β), which is independent of the choice of α,β.

 Remark 4.24.

For all x,y:𝕀, we have that x<y is an open proposition and that xy is equivalent to (x<y)(y<x).

Lemma 4.25.

For all D2 decidable, we have that cs(D) is a finite union of closed intervals.

Proof.

If D is contains precisely the α:2 with a fixed initial segment, cs(D) is a closed interval. Any decidable subset of 2 is a finite union of such subsets.

Lemma 4.26.

The complement of a finite union of closed intervals is a finite union of open intervals.

By Corollary 4.7 we can thus conclude:

Lemma 4.27.

Every open U𝕀 can be written as a countable union of open intervals.

It follows that the topology of 𝕀 is generated by open intervals, which corresponds to the standard topology on 𝕀. Hence our notion of continuity agrees with the ϵ,δ-definition of continuity one would expect and we get the following:

Theorem 4.28.

Every function f:𝕀𝕀 is continuous in the ϵ,δ-sense.

5 Cohomology

In this section we compute H1(S,)=0 for all S Stone, and show that H1(X,) for X compact Hausdorff can be computed using Čech cohomology. We use this to compute H1(𝕀,)=0.

 Remark 5.1.

We only work with the first cohomology group with coefficients in as it is sufficient for the proof of Brouwer’s fixed-point theorem, but the results could be extended to Hn(X,A) for A any family of countably presented abelian groups indexed by X.

 Remark 5.2.

We write Ab for the type of abelian groups and if G:Ab we write BG for the delooping of G [13, 22]. This means that H1(X,G) is the set truncation of XBG.

5.1 Čech cohomology

Definition 5.3.

Given a type S, types Tx for x:S and A:SAb, we define Cˇ(S,T,A) as the chain complex

where the boundary maps are defined as

d0(α)x(u,v)= αx(v)αx(u)
d1(β)x(u,v,w)= βx(v,w)βx(u,w)+βx(u,v)
Definition 5.4.

Given a type S, types Tx for x:S and A:SAb, we define its Čech cohomology groups by

Hˇ0(S,T,A)=ker(d0)Hˇ1(S,T,A)=ker(d1)/im(d0)

We call elements of ker(d1) cocycles and elements of im(d0) coboundaries.

This means that Hˇ1(S,T,A)=0 if and only if Cˇ(S,T,A) is exact at the middle term. Now we give three general lemmas about Čech complexes.

Lemma 5.5.

Assume a type S, types Tx for x:S and A:SAb with t:x:STx. Then Hˇ1(S,T,A)=0.

Proof.

Assume given a cocycle, i.e. β:x:SAxTx2 such that for all x:S and u,v,w:Tx we have that βx(u,v)+βx(v,w)=βx(u,w). We define α:x:SAxTx by αx(u)=βx(tx,u). Then for all x:S and u,v:Tx we have that d0(α)x(u,v)=βx(tx,v)βx(tx,u)=βx(u,v) so that β is a coboundary.

Lemma 5.6.

Given a type S, types Tx for x:S and A:SAb, we have that Hˇ1(S,T,λx.AxTx)=0.

Proof.

Assume given a cocycle, i.e. β:x:SAxTx3 such that for all x:S and u,v,w,t:Tx we have that βx(u,v,t)+βx(v,w,t)=βx(u,w,t). We define α:x:SAxTx2 by αx(u,t)=βx(t,u,t). Then for all x:S and u,v,t:Tx we have that d0(α)x(u,v,t)=βx(t,v,t)βx(t,u,t)=βx(u,v,t) so that β is a coboundary.

Lemma 5.7.

Assume a type S and types Tx for x:S such that x:STx and A:SAb such that Hˇ1(S,T,A)=0. Then given α:x:SBAx with β:x:S(α(x)=)Tx, we can conclude α=.

Proof.

We define g:x:SAxTx2 by gx(u,v)=βx(v)βx(u). It is a cocycle in the Čech complex, so that by exactness there is f:x:SAxTx such that for all x:S and u,v:Tx we have that gx(u,v)=fx(v)fx(u). Then we define β:x:S(α(x)=)Tx by βx(u)=βx(u)fx(u) so that for all x:S and u,v:Tx we have that βx(u)=βx(v) is equivalent to fx(v)fx(u)=βx(v)βx(u), which holds by definition. So β is constant on each Tx and therefore gives x:S(α(x)=)Tx. By x:STx we conclude α=.

5.2 Cohomology of Stone spaces

Lemma 5.8.

Assume given S:𝖲𝗍𝗈𝗇𝖾 and T:S𝖲𝗍𝗈𝗇𝖾 such that x:ST(x). Then there exists a sequence of finite types (Sk)k: with limit S and a compatible sequence of families of finite types Tk over Sk with x:SkTk(x) and limk(x:SkTk(x))=x:ST(x).

Proof.

By theorem Theorem 4.18 and the usual correspondence between surjections and families of inhabited types, a family of inhabited Stone spaces over S correspond to a Stone space T with a surjection TS. Then we conclude using Remark 3.4.

Lemma 5.9.

Assume given S:𝖲𝗍𝗈𝗇𝖾 with T:S𝖲𝗍𝗈𝗇𝖾 such that x:STx. Then we have that Hˇ1(S,T,)=0.

Proof.

We apply Lemma 5.8 to get Sk and Tk finite. Then by Corollary 3.7 we have that Cˇ(S,T,) is the sequential colimit of the Cˇ(Sk,Tk,). By Lemma 5.5 we have that each of the Cˇ(Sk,Tk,) is exact, and a sequential colimit of exact sequences is exact.

Lemma 5.10.

Given S:𝖲𝗍𝗈𝗇𝖾, we have that H1(S,)=0.

Proof.

Assume given a map α:SB. We use local choice to get T:S𝖲𝗍𝗈𝗇𝖾 such that x:STx with β:x:S(α(x)=)Tx. Then we conclude by Lemma 5.9 and Lemma 5.7.

Corollary 5.11.

For any S:𝖲𝗍𝗈𝗇𝖾 the canonical map B(S)(B)S is an equivalence.

Proof.

This map is always an embedding. To show it is surjective it is enough to prove that (B)S is connected, which is precisely Lemma 5.10.

5.3 Čech cohomology of compact Hausdorff spaces

Definition 5.12.

A Čech cover consists of X:𝖢𝖧𝖺𝗎𝗌 and S:X𝖲𝗍𝗈𝗇𝖾 such that x:XSx and x:XSx:𝖲𝗍𝗈𝗇𝖾.

By definition any compact Hausdorff space X is part of a Čech cover (X,S).

Lemma 5.13.

Given a Čech cover (X,S) and A:XAb, we have an isomorphism H0(X,A)=Hˇ0(X,S,A) natural in A.

Proof.

By definition an element in Hˇ0(X,S,A) is a map f:x:XAxSx such that for all u,v:Sx we have f(u)=f(v). Since Ax is a set and the Sx are merely inhabited, this is equivalent to x:XAx. Naturality in A is immediate.

Lemma 5.14.

Given a Čech cover (X,S) we have an exact sequence

H0(X,λx.Sx)H0(X,λx.Sx/)H1(X,)0

Proof.

We use the long exact cohomology sequence associated to

0SxSx/0

We just need H1(X,λx.Sx)=0 to conclude. But by Corollary 5.11 we have that H1(X,λx.Sx)=H1(x:XSx,) which vanishes by Lemma 5.10.

Lemma 5.15.

Given a Čech cover (X,S) we have an exact sequence

Hˇ0(X,S,λx.Sx)Hˇ0(X,S,λx.Sx/)Hˇ1(X,S,)0

Proof.

For n=1,2,3, we have that Σx:XSxn is Stone so that H1(Σx:XSxn,)=0 by Lemma 5.10, giving short exact sequences

0Πx:XSxnΠx:X(Sx)SxnΠx:X(Sx/)Sxn0

They fit together in a short exact sequence of complexes

0Cˇ(X,S,)Cˇ(X,S,λx.Sx)Cˇ(X,S,λx.Sx/)0

But since Hˇ1(X,λx.Sx)=0 by Lemma 5.6, we conclude using the associated long exact sequence.

Theorem 5.16.

Given a Čech cover (X,S), we have that H1(X,)=Hˇ1(X,S,)

Proof.

By applying Lemma 5.13, Lemma 5.14 and Lemma 5.15 we get that H1(X,) and Hˇ1(X,S,) are cokernels of isomorphic maps, so they are isomorphic.

This means that Čech cohomology does not depend on S.

5.4 Cohomology of the interval

 Remark 5.17.

Recall from Definition 4.19 that there is a binary relation n on 2n=:𝕀n such that (2n,n) is equivalent to (Fin(2n),λx,y.|xy|1) and for α,β:2 we have (cs(α)=cs(β))(n:α|nnβ|n).

We define 𝕀n2=Σx,y:𝕀nxny and 𝕀n3=Σx,y,z:𝕀nxnyynzxnz.

Lemma 5.18.

For any n: we have an exact sequence

0d0𝕀nd1𝕀n2d2𝕀n3

where d0(k)=(_k) and

d1(α)(u,v) = α(v)α(u)
d2(β)(u,v,w) = β(v,w)β(u,w)+β(u,v).

Proof.

It is clear that the map 𝕀n is injective as 𝕀n is inhabited, so the sequence is exact at . Assume a cocycle α:𝕀n, meaning that for all u,v:𝕀n, if unv then α(u)=α(v). Then by Remark 5.17 we see that α is constant, so the sequence is exact at 𝕀n.

Assume a cocycle β:𝕀n2, meaning that for all u,v,w:𝕀n such that unv, vnw and unw we have that β(u,v)+β(v,w)=β(u,w). Using Remark 5.17 to pass along the equivalence between 2n and Fin(2n), we define α(k)=β(0,1)++β(k1,k). We can check that β(k,l)=α(l)α(k), so that β is indeed a coboundary and the sequence is exact at 𝕀n2.

Proposition 5.19.

We have that H0(𝕀,)= and H1(𝕀,)=0.

Proof.

Consider cs:2𝕀 and the associated Čech cover T of 𝕀 defined by:

Tx=Σy:2(x=𝕀cs(y))

Then for l=2,3 we have that limn𝕀nl=x:𝕀Txl. By Lemma 5.18 and stability of exactness under sequential colimit, we have an exact sequence

0colimn𝕀ncolimn𝕀n2colimn𝕀n3

By Corollary 3.7 this sequence is equivalent to

0Πx:𝕀TxΠx:𝕀Tx2Πx:𝕀Tx3

So it being exact implies that Hˇ0(𝕀,T,)= and Hˇ1(𝕀,T,)=0. We conclude by Lemma 5.13 and Theorem 5.16.

 Remark 5.20.

We could carry a similar computation for 𝕊1, by approximating it with 2n with 0nn1n added. We would find H1(𝕊1,)=. We will give an alternative, more conceptual proof in the next section.

5.5 Brouwer’s fixed-point theorem

Here we consider the modality defined by localising at 𝕀 as explained in [14]. It is denoted by L𝕀. We say that X is 𝕀-local if L𝕀(X)=X and that it is 𝕀-contractible if L𝕀(X)=1.

Lemma 5.21.

and 2 are 𝕀-local.

Proof.

By Proposition 5.19, from H0(𝕀,)= we get that the map 𝕀 is an equivalence, so is 𝕀-local. We see that 2 is 𝕀-local as it is a retract of .

 Remark 5.22.

Since 2 is 𝕀-local, we have that any Stone space is 𝕀-local.

Lemma 5.23.

B is 𝕀-local.

Proof.

Any identity type in B is a -torsor, so it is 𝕀-local by Lemma 5.21. So the map BB𝕀 is an embedding. From H1(𝕀,)=0 we get that it is surjective, hence an equivalence.

Lemma 5.24.

Assume X a type with x:X such that for all y:X we have f:𝕀X such that f(0)=x and f(1)=y. Then X is 𝕀-contractible.

Proof.

For all y:X we get a map g:𝕀L𝕀(X) such that g(0)=[x] and g(1)=[y]. Since L𝕀(X) is 𝕀-local this means that y:X[x]=[y]. We conclude y:L𝕀(X)[x]=y by applying the elimination principle for the modality.

Corollary 5.25.

We have that and 𝔻2={(x,y):2|x2+y21} are 𝕀-contractible.

Proposition 5.26.

L𝕀(/)=B.

Proof.

As for any group quotient, the fibers of the map / are -torsors, so we have an induced pullback square

Now we check that the bottom map is an 𝕀-localisation. Since B is 𝕀-local by Lemma 5.23, it is enough to check that its fibers are 𝕀-contractible. Since B is connected it is enough to check that is 𝕀-contractible. This is Corollary 5.25.

 Remark 5.27.

By Lemma 5.23, for any X we have that H1(X,)=H1(L𝕀(X),), so that by Proposition 5.26 we have that H1(/,)=H1(B,)=.

We omit the proof that 𝕊1={(x,y):2|x2+y2=1} is equivalent to /. The equivalence can be constructed using trigonometric functions, which exist by Proposition 4.12 in [3].

Proposition 5.28.

The map 𝕊1𝔻2 has no retraction.

Proof.

By Corollary 5.25 and Proposition 5.26 we would get a retraction of B1, so B would be contractible.

Theorem 5.29 (Intermediate value theorem).

For any f:𝕀𝕀 and y:𝕀 such that f(0)y and yf(1), there exists x:𝕀 such that f(x)=y.

Proof.

By Corollary 4.3, the proposition x:𝕀f(x)=y is closed and therefore ¬¬-stable, so we can proceed with a proof by contradiction. If there is no such x:𝕀, we have f(x)y for all x:𝕀. By Remark 4.24 we have that a<b or b<a for all distinct numbers a,b:𝕀. So the following two sets cover 𝕀

U0:={x:𝕀f(x)<y}U1:={x:𝕀y<f(x)}

Since U0 and U1 are disjoint, we have 𝕀=U0+U1 which allows us to define a non-constant function 𝕀2, which contradicts Lemma 5.21.

Theorem 5.30 (Brouwer’s fixed-point theorem).

For all f:𝔻2𝔻2 there exists x:𝔻2 such that f(x)=x.

Proof.

As above, by Corollary 4.3, we can proceed with a proof by contradiction, so we assume f(x)x for all x:𝔻2. For any x:𝔻2 we set dx=xf(x), so we have that one of the coordinates of dx is invertible. Let Hx(t)=f(x)+tdx be the line through x and f(x). The intersections of Hx and 𝔻2=𝕊1 are given by the solutions of an equation quadratic in t. By invertibility of one of the coordinates of dx, there is exactly one solution with t>0. We denote this intersection by r(x) and the resulting map r:𝔻2𝕊1 has the property that it preserves 𝕊1. Then r is a retraction from 𝔻2 onto its boundary 𝕊1, which is a contradiction by Proposition 5.28.

 Remark 5.31.

In constructive reverse mathematics [7], it is known that both the intermediate value theorem and Brouwer’s fixed-point theorem are equivalent to LLPO. But LLPO does not hold in real cohesive homotopy type theory, so [16] prove a variant of the statement involving a double negation.

References

  • [1] Dagur Ásgeirsson. The foundations of condensed mathematics. Master’s thesis, Université de Paris, 2021. URL: https://dagur.sites.ku.dk/condensed-foundations/.
  • [2] Reid Barton and Johan Commelin. Lean-ctt-snapshot. URL: https://github.com/jcommelin/lean-ctt-snapshot.
  • [3] Errett Bishop and Douglas Bridges. Constructive analysis, volume 279 of Grundlehren der mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1985. doi:10.1007/978-3-642-61667-9.
  • [4] Felix Cherubini, Thierry Coquand, and Matthias Hutzler. A foundation for synthetic algebraic geometry, 2023. arXiv:2307.00073.
  • [5] Dustin Clausen and Peter Scholze. Analytic stacks, 2023-2024. Lecture series. URL: https://www.youtube.com/playlist?list=PLx5f8IelFRgGmu6gmL-Kf_Rl_6Mm7juZO.
  • [6] Thierry Coquand, Fabian Ruch, and Christian Sattler. Constructive sheaf models of type theory. Math. Struct. Comput. Sci., 31(9):979–1002, 2021. doi:10.1017/S0960129521000359.
  • [7] Hannes Diener. Constructive reverse mathematics : Habilitationsschrift, 2018. URL: https://dspace.ub.uni-siegen.de/handle/ubsi/1306.
  • [8] Roy Dyckhoff. Categorical methods in dimension theory. Categor. Topol., Proc. Conf. Mannheim 1975, Lect. Notes Math. 540, 220-242 (1976)., 1976.
  • [9] Martín Escardó. Synthetic topology: of data types and classical spaces. Electronic Notes in Theoretical Computer Science, 87:21–156, 2004. Proceedings of the Workshop on Domain Theoretic Methods for Probabilistic Processes. doi:10.1016/j.entcs.2004.09.017.
  • [10] Florian Faissole and Bas Spitters. Synthetic topology in homotopy type theory for probabilistic programming. In Proc, page 2017, 2017.
  • [11] Hajime Ishihara. Reverse Mathematics in Bishop’s Constructive Mathematics. Philosophia Scientiae, 6:43–59, September 2006. doi:10.4000/philosophiascientiae.406.
  • [12] Davorin Lešnik. Synthetic Topology and Constructive Metric Spaces. PhD thesis, University of Ljubljana, 2021. arXiv:2104.10399.
  • [13] The Univalent Foundations Program. Homotopy type theory: Univalent foundations of mathematics. https://homotopytypetheory.org/book, 2013.
  • [14] Egbert Rijke, Michael Shulman, and Bas Spitters. Modalities in homotopy type theory. Logical Methods in Computer Science, Volume 16, Issue 1, January 2020. doi:10.23638/LMCS-16(1:2)2020.
  • [15] Peter Scholze. Lectures on condensed mathematics, 2019. URL: https://people.mpim-bonn.mpg.de/scholze/Condensed.pdf.
  • [16] Michael Shulman. Brouwer’s fixed-point theorem in real-cohesive homotopy type theory. Mathematical Structures in Computer Science, 28(6):856–941, 2018. doi:10.1017/S0960129517000147.
  • [17] Michael Shulman. All (,1)-toposes have strict univalent universes. arXiv preprint, 2019. arXiv:1904.07004.
  • [18] Kristina Sojakova, Floris van Doorn, and Egbert Rijke. Sequential colimits in homotopy type theory. In Proceedings of the 35th Annual ACM/IEEE Symposium on Logic in Computer Science (LICS 2020), page 14. ACM, New York, [2020] ©2020. doi:10.1145/3373718.3394801.
  • [19] Paul Taylor. Abstract Stone duality, 2011. URL: https://www.paultaylor.eu/ASD/.
  • [20] Niels van der Weide. The internal languages of univalent categories, 2024. doi:10.48550/arXiv.2411.06636.
  • [21] Steven Vickers. Locales and toposes as spaces. In Handbook of spatial logics, pages 429–496. Springer, 2007. doi:10.1007/978-1-4020-5587-4_8.
  • [22] David Wärn. Eilenberg-Maclane spaces and stabilisation in homotopy type theory. J. Homotopy Relat. Struct., 18(2-3):357–368, 2023. doi:10.1007/s40062-023-00330-5.
  • [23] David Wärn. On internally projective sheaves of groups, 2024. arXiv:2409.12835.