Hostname: page-component-745bb68f8f-f46jp Total loading time: 0 Render date: 2025-02-05T23:55:14.636Z Has data issue: false hasContentIssue false

HOMOTOPY MODEL THEORY

Published online by Cambridge University Press:  05 October 2020

BRICE HALIMI*
Affiliation:
DÉPARTEMENT D’HISTOIRE ET DE PHILOSOPHIE DES SCIENCES & SPHERE UNIVERSITÉ DE PARISPARIS, FRANCEE-mail: brice.halimi@u-paris.fr
Rights & Permissions [Opens in a new window]

Abstract

Drawing on the analogy between any unary first-order quantifier and a “face operator,” this paper establishes several connections between model theory and homotopy theory. The concept of simplicial set is brought into play to describe the formulae of any first-order language L, the definable subsets of any L-structure, as well as the type spaces of any theory expressed in L. An adjunction result is then proved between the category of o-minimal structures and a subcategory of the category of linearly ordered simplicial sets with distinguished vertices.

Type
Article
Copyright
© Association for Symbolic Logic 2020

1 Simplicial ideas

1.1 Formulae as chains

Homotopy theory dramatically entered the scene of logic through the connections that have been made between Martin-Löf type theory and model categories, since Hofmann and Streicher’s seminal paper.Footnote 1 This paper aims at focusing on another connection of homotopy theory with logic, based on the following remark: any structure for a first-order language can be turned into a simplicial set. Hence, whereas “Homotopy Type Theory” connects logic with homotopy theory through type theory, “Homotopy Model Theory” is the proposal to connect logic with homotopy theory through model theory. This is the task taken up here.Footnote 2

An intuitive motivation of such perspective is that the notion of boundary can easily be transposed to the context of first-order logic, formulae being conceived of as chains (in the sense of a formal sum of faces). The boundary of a given formula $\phi (v_0, \ldots , v_n)$ with exactly $v_0, \ldots , v_n$ as free variables can be defined as follows:

$$ \begin{align*} \partial\phi := \bigwedge\limits_{i=0}^{n-1}\neg^i\forall x\phi(v_0, \ldots, v_{i-1}, x, v_{i+1}, \ldots, v_{n-1}), \end{align*} $$

where ‘ $\neg^i$ ’ indicates i consecutive occurrences of the negation symbol. The boundary operator $\partial$ is not stable under variable renaming, however. To avoid any problem of that kind, we will consider, throughout this paper, a single fixed first-order language L with equality, containing the existential quantifier as a primitive, whose free variable symbols are exactly ‘ $v_i$ ’, $i\geq 0$ , and whose bound variable symbols are exactly ‘x’, ‘y’, ‘z’, and so on. Each L-formula will be taken up to bound variable renaming, but not up to free variable renaming.

Definition 1.1. An L-formula $\phi $ is called a well-formed L-formula iff the indices of all the free variables occurring in $\phi $ , regardless of the order of the occurrences of those variables, make up an initial segment of $\mathbb {N}$ . From now on, “L-formula” will always refer to a well-formed L-formula only.

Starting with an L-formula $\phi (v_0, v_1, v_2)$ , one finds that $\partial (\partial \phi )\equiv \bot $ (since universal quantification commutes with conjunction). This prompts a comparison of $\partial $ with a genuine boundary operator, in a sense pointing to homotopy theory. To that end, a few explanatory remarks are in order about the concept of simplicial set, given the role that it will play in what follows. CW complexes were introduced in algebraic topology in order to analyze nonpathological topological spaces as reconstructible by glueing elementary “cells” along other cells of smaller dimension: the building pattern of CW complexes can be used to compute the homology groups and homotopy groups of that space. Designed to formalize this kind of decomposition of a topological space into cells of increasing dimension, glued together along a common face, a simplicial set X is a sequence $(X_n)_{n\in \mathbb {N}}$ of sets, together with maps $d_i : X_n\rightarrow X_{n-1}$ ( $0\leq i\leq n$ ) and $s_j : X_n\rightarrow X_{n+1}$ ( $0\leq j\leq n$ ), for each $n\in \mathbb {N}$ , satisfying the following simplicial identities:

$$ \begin{align*}\begin{cases} d_id_j= d_{j-1}d_i\ \ \ \ \text{if}\ i<j,\\ d_is_j = s_{j-1}d_i\ \ \ \ \text{if}\ i<j,\\ d_js_j = d_{j+1}s_j = \text{id},\\ d_is_j = s_jd_{i-1}\ \ \ \ \text{if}\ i>j+1,\\ s_is_j = s_{j+1}s_i\ \ \ \ \text{if}\ i\leq j. \end{cases}\end{align*} $$

The elements of each set $X_n$ , called n-simplices, represent the n-cells involved in the construction of the topological space to be built, each map $d_i$ the projection of an n-cell onto one of its faces, and each $d_j$ the upgrading of an n-cell as a degenerate $(n+1)$ -cell, while the simplicial identities encode the purely combinatorial properties of the building pattern of the space.

Despite being deprived of any topology, simplicial sets are sufficient to capture most features relevant to homotopy theory, since the homotopy category of the category of simplicial sets is a model of homotopy types. As will be seen presently, simplicial sets also supply an interesting connection between the combinatorial aspect of logical syntax and the use of topological methods in model theory.

Definition 1.2. Let $F_n$ be the set of L-formulae with exactly $v_0, \ldots , v_{n}$ as free variables. ( $F_{-1}$ may be defined as the set of all L-sentences.) For each $n\in \mathbb {N}$ , two systems of maps, $\exists _i : F_n\rightarrow F_{n-1}\ (0\leq i\leq n)$ and $E_j : F_n\rightarrow F_{n+1}\ (0\leq j\leq n)$ are, respectively:

$$ \begin{align*}\exists_i(\phi(v_0, \ldots, v_{n}))& = \exists x\phi(v_0, \ldots, v_{i-1}, x, v_i, \ldots, v_{n-1}),\\E_j(\phi(v_0, \ldots, v_{n}))& = ((v_j = v_{j+1})\rightarrow \phi(v_0, \ldots, v_{j-1}, v_{j+1}, \ldots, v_{n+1})).\end{align*} $$

Logical equivalence between L-formulae is considered here as an equivalence relation always restricted to some $F_n$ , in other words as a graded logical equivalence. On this condition only, the operators $\exists _i$ , $E_j$ and $\partial $ can be defined up to logical equivalence.Footnote 3 Then, up to graded logical equivalence, the maps $\exists _i$ and $E_j$ satisfy the above simplicial identities.

Definition 1.3. To ensure that these simplicial identities be genuine equalities, every L-formula will henceforth be considered up to graded logical equivalence.

Theorem 1.4. $F_* = \langle F_n, (\exists _i)_{0\leq i\leq n}, (E_j)_{0\leq j\leq n}\rangle _{n\in \mathbb {N}}$ is a simplicial set.

In this perspective, the existential quantifier can be likened to a “face operator,” while the maps $E_j$ are the corresponding “degeneracy operators.” In fact, if any L-formula in $F_n$ is seen as the presentation of some n-dimensional subset, $\exists _i$ is exactly a projection, and $E_j$ the converse operation of cylindrification.

Remark 1.5. In the definition of $\exists _i$ , the existential quantifier can be replaced with any first-order unary quantifier Q satisfying $Qy Qx \phi (y, x, \vec {u})\equiv Qx Qy \phi (x, y, \vec {u})$ and $Qx ((x=y)\rightarrow \phi [x/y])\equiv \phi (y)$ , for every L-formula $\phi $ . The associated face operators are written $d_i^Q$ and the ensuing simplicial set is written $F_*^Q$ (so $\exists _i = d_i^{\exists }$ and $F_* = F_*^{\exists }$ ).

Given two simplicial sets X and Y, a simplicial map $f : X\rightarrow Y$ consists of a system of maps $f_n : X_n\rightarrow Y_n$ ( $n\in \mathbb {N}$ ) such that both diagrams below

commute for every i and every j. The category ${\mathbf { sSets}}$ of simplicial sets is the category whose objects are all simplicial sets and whose arrows are all simplicial maps. Within that category, certain simplicial sets stand out: the fibrant ones. Given two n-simplices $x, y\in X_n$ of a simplicial set X, a shared face is an $(n-1)$ -simplex of X which is a common face of x and y. An n-horn in X is a system of n distinct $(n-1)$ -simplices of X with a maximal number of $n(n-1)/2$ shared faces. For instance, a $2$ -horn in X can be represented as a pair of two edges with a shared vertex, in other words as a triangle with one side removed; a $3$ -horn in X, as a triple of full triangles making up the surface of a tetrahedron with one face removed. A simplicial set X is said to be fibrant if, for every $n\geq 1$ , any n-horn in X can be extended into an n-simplex of X. This amounts to saying, in combinatorial terms, that if $x_0, \ldots , \widehat {x_k}, \ldots , x_n$ is an n-tuple of $(n-1)$ -simplices of X (‘ $\widehat {x_k}$ ’ meaning that $x_k$ is omitted) such that $d_ix_j = d_{j-1}x_i$ for any $i, j\neq k$ with $i<j$ , then there is an n-simplex x of X such that $d_ix = x_i$ for any $i\neq k$ .

Proposition 1.6. $F_*$ is fibrant.

Proof The proof derives as a special case from the proof of Theorem 2.10. ⊣

1.2 Syntactic extensions

Following on from this first comparison, it is quite natural to ask about connectives. It turns out that $\wedge $ is characterized, among all symmetric binary connectives, by the condition that $\forall x (\phi x\wedge \psi x)\equiv (\forall x \phi x\wedge \forall x \psi x)$ , and $\vee $ in an analogous way. Because binary connectives are at stake, two-dimensional simplicial sets need to be introduced, which is straightforward as soon as it is noticed that a simplicial set can be equivalently defined as a functor $X : {\mathbf {\Delta ^{\scriptsize op}}}\rightarrow {\mathbf { Sets}}$ , where ${\mathbf {\Delta }}$ is the category whose objects are all the finite ordinals and whose morphisms are all order-preserving maps, and where ${\mathbf { Sets}}$ is the category of sets and maps between sets. Following this alternative view, one can define a bisimplicial set as a functor from ${\mathbf {\Delta ^{\scriptsize op}\times \Delta ^{\scriptsize op}}}$ to ${\mathbf { Sets}}$ .

The usual unary and binary connectives can then be described in simplicial and bisimplicial terms. Indeed, the commutative diagram

is a way to express that $Q''x(\phi c \psi )\equiv (Qv_i\phi ) c' (Q'v_i\psi )$ for any L-formulae $\phi \in F_m$ , $\psi \in F_p$ (here $n = {max}(m, p)$ ). One has that $\neg $ , $\wedge $ and $\vee $ are characterized, respectively, by the commutativity of the three following diagrams:

There might seem to be a small problem with conjunction of L-formulae that share the same variables. For instance, how to regiment the L-formula $(R(v_0, v_1)\wedge S(v_0, v_2))$ ? The L-formula has in fact to forego some transformation, but within its logical equivalence class (although not in the graded sense):

$$ \begin{align*} & (R(v_0, v_1)\wedge S(v_0, v_2))\\ &\equiv (R(v_0, v_1)\wedge S(v_2, v_3)\wedge (v_0 = v_2))\\ &\equiv ((R(v_0, v_1)\wedge (v_0 = v_2))\wedge ((v_0 = v_2)\wedge(v_1 = v_1)\wedge S(v_2, v_3))). \end{align*} $$

Besides bisimplicial sets, the definition of a simplicial set as a functor $X : {\mathbf {\Delta ^{\scriptsize op}}}\rightarrow \mathbf { Sets}$ lends itself to another kind of generalization: a simplicial object in an Abelian category $\mathbf { C}$ is a functor $X : {\mathbf {\Delta ^{\scriptsize op}}}\rightarrow \mathbf { C}$ . A simplicial object in the category of Abelian groups is called a simplicial group. Any simplicial object X in any Abelian category gives rise, in a canonical way, to an associated chain complex:Footnote 4 , where $d^n := \sum_{i = 0}^n (-1)^i d_i$ and $d^n\circ d^{n+1} = 0$ . Each $F_*$ turns out to be a simplicial Boolean algebra, with sum and product corresponding to disjunction and conjunction. The category of Boolean algebras, however, is not an Abelian category. An option is to consider each $F_n$ as naturally equipped with a group structure, namely $\langle F_n, \nleftrightarrow , \bot \rangle $ , where $\bot $ is any antilogy in $F_n$ and $(\phi \nleftrightarrow \psi ) := ((\phi \wedge \neg \psi )\vee (\psi \wedge \neg \phi ))$ . However, $F_*$ ’s being a simplicial group requires more than all the $F_n$ ’s being groups: it also requires all the $\exists _i$ ’s and $E_j$ ’s being morphisms of Abelian groups.

Another, natural option is thus to introduce the free Abelian group $\mathbb {Z}F_n$ generated by the n-simplices, up to the identification $(\phi + \phi ) = \phi $ . The operators $\exists _i : \mathbb {Z}F_n\rightarrow \mathbb {Z}F_{n-1}$ are given by linear extension. The elements of $\mathbb {Z}F_n$ can be seen as sequents between L-formulae with $n+1$ free variables, up to the equivalences provided by the structural rules of the sequent calculus for first-order classical logic. Indeed, any element of $\mathbb {Z}F_n$ can be written as follows:

$$ \begin{align*} \sum_{k=1}^{n}n_k\phi_k(v_0, \ldots, v_n) &= \sum_{n_k<0}n_k\phi_k \vdash \sum_{n_l>0}n_l\phi_l\\&= \phi_{k_1}, \ldots, \phi_{k_p} \vdash \phi_{l_1}, \ldots, \phi_{l_q}. \end{align*} $$

Any sequent of the form $\phi (v_0, \ldots , v_n) \vdash \phi (v_0, \ldots , v_n)$ represents the identity element $0$ of $\mathbb {Z}F_n$ .

Definition 1.7. Given

$$\begin{align*}\widetilde {\exists _i}(\phi _1, \ldots , \phi _p \vdash \psi _1, \ldots , \psi _q) := \exists _i(\phi _1), \ldots , \exists _i(\phi _p) \vdash \exists _i(\psi _1), \ldots , \exists _i(\psi _q),\end{align*}$$

the n -th boundary map associated to $\exists $ is: $\partial ^{\exists }_n = \sum _{i=0}^n (-1)^i \widetilde {\exists _i} : \mathbb {Z}F_n\rightarrow \mathbb {Z}F_{n-1}$ .

Proposition 1.8. The sequence $\langle \mathbb {Z}F_n, \partial _n\rangle _{n\in \mathbb {N}}$ defines a chain complex:

$$ \begin{align*}\partial^{\exists}_{n-1}\circ \partial^{\exists}_n = 0\ \ \text{for all}\ n\geq 1.\end{align*} $$

Proof Let S be:

$$\begin{align*}\phi _{1}(v_0, \ldots , v_n), \ldots , \phi _{p}(v_0, \ldots , v_n) \vdash \psi _{1}(v_0, \ldots , v_n), \ldots , \psi _{q}(v_0, \ldots , v_n).\end{align*}$$

Writing $\phi $ for $\sum_{1\leq k\leq p}\phi _k$ and $\psi $ for $\sum_{1\leq l\leq q}\psi _l$ , one has, for n even:

$$ \begin{align*} & \partial^{\exists}_n(S)\\&= (\exists x\phi(x,v_{0},\ldots,v_{n-1})\vdash \exists x\psi(x,v_{0},\ldots,v_{n-1}))\\& - (\exists x\phi(v_0,x,v_{1},\ldots,v_{n-1})\vdash \exists x\psi(v_0,x,v_{1},\ldots,v_{n-1})) + \cdots \\& \cdots + (\exists x\phi(v_{0},\ldots,v_{n-1},x)\vdash \exists x\psi(v_{0},\ldots,v_{n-1},x))\\&= (\exists x\phi(x,v_{0},\ldots,v_{n-1}), \exists x\psi(v_0,x,v_{1},\ldots,v_{n-1}), \exists x\phi(v_0,v_1,x,\ldots,v_{n-1}), \end{align*} $$

$$ \begin{align*} & \ldots, \exists x\phi(v_{0},\ldots,v_{n-1},x)) \vdash (\exists x\psi(x,v_{0},\ldots,v_{n-1}), \exists x\phi(v_0,x,v_{1},\ldots,v_{n-1}),\\& \exists x\psi(v_0,v_1,x,\ldots,v_{n-1}), \ldots, \exists x\psi(v_{0},\ldots,v_{n-1}, x)). \end{align*} $$

Then, writing ‘ $\phi _{01x2\ldots. n}$ ’ as a shorthand for $\exists x\phi (v_0, v_1, x, v_2, \ldots , v_n)$ , one can see that each $\widetilde {\exists _i}$ , for $0\leq i\leq n-1$ , will produce, if i is even (resp. odd), $\phi _{x0\ldots i-1yi\ldots n-2}$ , $\psi _{0x1\ldots i-1y i\ldots n-2}$ , …, $\phi _{0\ldots i-1yi\ldots x}$ on the left side (resp. right side) of the sequent and $\psi _{x0\ldots i-1yi\ldots n-2}$ , $\psi _{0x1\ldots i-1yi\ldots n-2}$ , …, $\psi _{0\ldots i-1yi\ldots x}$ on the right side (resp. left side). So $\widetilde {\exists _i}$ and $\widetilde {\exists _{i+1}}$ , put together, will produce the same sequent. Since n is supposed to be even, the n indices i between $0$ and $n-1$ can all be considered pairwise, and so in the end the right and left sides of $\partial ^{\exists }_{n-1}(\partial ^{\exists }_n(S))$ match up. The proof in case n is odd is analogous. ⊣

Definition 1.9. For each $n\in \mathbb {N}$ , $Z_n(\exists ) := \text {Ker } \partial ^{\exists }_n$ is the set of all n -cycles and $B_n(\exists ) := \text {Im } \partial ^{\exists }_{n+1}$ is the set of all n -boundaries associated to $\exists $ . The quotient $H_n(\exists ) := Z_n(\exists )/B_n(\exists )$ is the n -th homology group associated to $\exists $ .

The homology groups $(H_n(\exists ))_{n\in \mathbb {N}}$ represent the simplicial homology of first-order logic.

2 Simplicial sets and structures

2.1 Elementary extensions

Definition 2.1. For any L-structure M,

$$ \begin{align*}F_*^M := \langle D_n(M), (\exists^{M}_i)_{0\leq i\leq n}, (E_j^{M})_{0\leq j\leq n}\rangle_{n\in\mathbb{N}},\ \ \text{where:}\end{align*} $$
  • $D_n(M)$ is the set of all definable subsets of $|M|^{n+1}$ with parameters;

  • $\exists ^{M}_i : D_n(M)\rightarrow D_{n-1}(M)),\ A=\{\vec {a}\in |M|^{n+1} : M\vDash \phi (v_0, \ldots , v_{n}) [\vec {a}]\}\mapsto \{\vec {a'}\in |M|^{n} : M\vDash \exists x \phi (v_0, \ldots , v_{i-1}, x, v_i, \ldots , v_{n-1}) [\vec {a'}]\}$ ;

  • $E_j^{M} : D_{n}(M)\rightarrow D_{n+1}(M),\ A\mapsto \{(\vec {x}, y) : \vec {x}\in A\ \text {and}\ y = x_j\}$ .

Proposition 2.2. For any L-structure M, $F_*^M$ is a simplicial set.

Proof The simplicial identities between the operators $\exists ^{M}_i$ and $E_j^{M}$ are the semantic counterpart of the simplicial identities which turn $F_*$ into a simplicial set. ⊣

Definition 2.3. A simplicial set with distinguished vertices is a simplicial set X equipped with a selection $X^{\prime}_0\subseteq X_0$ of vertices, called distinguished vertices.

A morphism $f : (X, X_0^{\prime })\rightarrow (Y, Y_0^{\prime })$ between simplicial sets with distinguished vertices, or simplicial morphism for short, is a simplicial map $f : X\rightarrow Y$ such that $f^{-1}(Y_0^{\prime }) = X_0^{\prime }$ .

Definition 2.4. Given any L-structure M, $F_*^M$ is canonically equipped with a set $M_0^{\prime }$ of distinguished vertices, namely the set of all nonempty definable subsets of $|M|$ without parameters. The corresponding simplicial set with distinguished vertices is written $M_*$ .

Theorem 2.5. Let M be a substructure of an L-structure N. The restriction $r_n$ , for each $n\in \mathbb {N}$ , being the map which sends $\phi ^N$ to $\phi ^N\cap |M|^{n+1}$ , for each L-formula $\phi \in F_n$ , M is an elementary substructure of N iff the family $(r_n)_{n\in \mathbb {N}}$ of restriction maps induces a simplicial morphism $r_* : N_*\rightarrow M_*$ .

Proof Let N be an extension of M. Then $r_n : D_n(N)\rightarrow N^M_n$ , where $N^M_n$ is the set of all $\phi (N, M) := \{\vec {a}\in |M|^{n+1} : N\vDash \phi (v_0, \ldots , v_n) [\vec {a}]\}$ for $\phi \in F_n$ . Furthermore, the sequence $(N^M_n)_{n\in \mathbb {N}}$ can be endowed with a simplicial structure $N^M_*$ , with $\exists ^{N^M}_i : \{\vec {b}\in |M|^{n+1} : N\vDash \phi (\vec {v}) [\vec {b}]\}\mapsto \{\vec {b^{\prime }}\in |M|^{n} : N\vDash \exists _i(\phi ) [\vec {b^{\prime }}]\}$ and $E^{N^M}_j : \{\vec {b}\in |M|^{n} : N\vDash \phi (\vec {v}) [\vec {b}]\}\mapsto \{(\vec {b}, b_j)\in |M|^{n+1} : N\vDash \phi (\vec {v}) [\vec {b}]\}$ , for each $\phi \in F_n$ .

Now, suppose that $r_*$ is a simplicial morphism:

Then it turns out that $N^M_*$ is in fact $M_*$ . This point is proved by induction on $\phi $ . Since M is a substructure of N, $\phi (N, M) = \phi ^M$ for each negatomic L-formula $\phi \in F_n$ . It is straightforward, by induction, that $(\phi \wedge \psi )(N, M) = (\phi (N, M)\wedge \psi (N, M)) = (\phi ^M\wedge \psi ^M) = (\phi \wedge \psi )^M$ . Finally:

$$ \begin{align*} (\exists x\phi(v_0, \ldots, v_{i-1}, x, v_i, \ldots, v_n))(N, M) &= r_n(\exists^{N}_i(\phi(v_0, \ldots, v_{n+1})))\\ &= \exists^{N^M}_i(r_{n+1}(\phi(v_0, \ldots, v_{n+1}))). \end{align*} $$

By induction hypothesis, $r_{n+1}(\phi )\in D_{n+1}(M)$ , so $\exists ^{N^M}_i(r_{n+1}(\phi )) = (\exists x\phi )(N, M)$ is in $D_n(M)$ . Now, the fact that the restriction maps $r_n$ make up a simplicial morphism from $N_*$ to $M_*$ is the direct expression that the Tarski–Vaught test is met. Indeed, suppose $N\vDash \exists x\phi (x, \overline {a_1}, \ldots , \overline {a_n})$ . Then $(\phi (v_0, \overline {a_1}, \ldots , \overline {a_n}))^N\in N_0'$ , so $r_0(\phi (v_0, \overline {a_1}, \ldots , \overline {a_n})^N) = \phi (v_0, \overline {a_1}, \ldots , \overline {a_n})(N, M)\in M_0'$ is nonempty, which means that there is some $a\in |M|$ such that $N\vDash \phi (v_0, \overline {a_1}, \ldots , \overline {a_n}) [a]$ . As a result, M is an elementary substructure of N.

Conversely, suppose that M is an elementary substructure of N. Then, it is routine verification to check that $r_*$ is a simplicial map, with codomain $M_*$ , and thus a simplicial morphism $r : N_*\rightarrow M_*$ . The elementwise commutativity of the diagram involving $r_n$ and $r_{n-1}$ follows from the Tarski–Vaught test. Indeed, let us consider $\phi \in F_n$ and let us suppose that $(a_1, \ldots , a_n)\in |M|^n$ . Then:

$N\vDash \exists x\phi (\overline {a_1}, \ldots , \overline {a_i}, x, \overline {a_{i+1}}, \ldots , \overline {a_n})$

iff there exists $a\in |M|$ such that $N\vDash \phi (\overline {a_1}, \ldots , \overline {a_i}, v_0, \overline {a_{i+1}}, \ldots , \overline {a_n}) [a]$

iff there exists $a\in |M|$ such that $(a_1, \ldots , a_i, a, a_{i+1}, \ldots , a_n)\in \phi ^N\cap |M|^{n+1}$

iff $(a_1, \ldots , a_n)\in \exists _i(\phi ^N\cap |M|^{n+1})$ .

Thus $\exists _i(\phi ^N)\cap |M|^n = \exists _i(\phi ^N\cap |M|^{n+1})$ for every L-formula $\phi \in F_n$ .⊣

Corollary 2.6. The mapping $(-)_*$ defines a contravariant functor from the category of L-structures and elementary embeddings, to the category of simplicial sets with distinguished vertices and simplicial morphisms.

Proof Any elementary embedding $f : M\rightarrow N$ gives rise to a simplicial morphism $f_* : N_*\rightarrow M_*$ , in a functorial way, with $f_n : \phi ^N\mapsto \{\vec {a}\in |M|^{n+1} : N\vDash \phi [f(\vec {a})]\} = \phi ^M$ .⊣

Note that any two elementary equivalent L-structures M and N have isomorphic hierarchies of definable subsets without parameters, but that the corresponding isomorphism $\phi ^M\mapsto \phi ^N$ between $M_*$ and $N_*$ is not induced in general by any actual map between M and N.

Let M be an elementary substructure of an L-structure N, and let us suppose that $|M|$ is definable in N, by an L-formula $\underline {M}(v_0)$ . For any $\phi \in F_n$ , the L-formula $\phi ^{\underline {M}}$ is defined as $(\underline {M}(v_0)\wedge \cdots \wedge \underline {M}(v_n)\wedge \phi ^{(\underline {M})})$ , where $\phi ^{(\underline {M})}$ is the quantifier relativization of $\phi $ to $\underline {M}$ . This relativization allows one to define, for each $n \in\mathbb{N}$ , an extension map $e_n : \phi ^M\in D_n(M)\mapsto (\phi ^{\underline {M}})^N\in D_n(N)$ , the existence of which, as will be seen presently, has a natural simplicial counterpart, if one moves to $M_*$ and $N_*$ . Given two simplicial sets X and Y (resp. two simplicial sets X and Y with distinguished vertices), a retraction of Y over X consists of a pair $\langle f, g\rangle $ of simplicial maps (resp. of simplicial morphisms) $f : X\rightarrow Y$ and $g : Y\rightarrow X$ , such that $g\circ f = \text {id}_X$ .

Theorem 2.7. Let M be an elementary substructure of an L-structure N. Then the domain $|M|$ of M is definable in N iff the family $(e_n)_{n\in\mathbb{N}}$ of extension maps induces a simplicial morphism $e_* : M_*\rightarrow N_*$ and $\langle e_*, r_*\rangle $ is a retraction of $N_*$ over $M_*$ .

Proof Suppose $|M|$ is definable in N by some L-formula $\underline {M}(v_0)$ . Since $N\vDash \phi [\vec {b}]$ iff $\vec {b}\in |M|^{n+1}$ and $M\vDash \phi [\vec {b}]$ , for any $\phi \in F_n$ , the diagram

commutes. Indeed, using the Tarski–Vaught test, one has:

$$ \begin{align*} N\vDash \exists_i(\phi^{\underline{M}}) [\vec{b'}]\ & \text{iff}\ N\vDash \phi^{\underline{M}} [b'_0, \ldots, b'_{i-1}, c_i, b'_i, \ldots, b'_{n-1}]\ \ \text{for some}\ c_i\in |N|\\&\text{iff}\ N\vDash \phi^{\underline{M}} [b'_0, \ldots, b'_{i-1}, d_i, b'_i, \ldots, b'_{n-1}]\ \ \text{for some}\ d_i\in |M|\\&\text{iff}\ N\vDash \exists v_i\ (\underline{M}(v_i)\wedge \phi^{\underline{M}}) [\vec{b'}]\\&\text{iff}\ N\vDash (\exists_i(\phi))^{\underline{M}} [\vec{b'}]. \end{align*} $$

That is to say, the collection of maps $(e_n : D_{n}(M)\rightarrow D_{n}(N))_{n\in \mathbb {N}}$ defines a simplicial map. Now, one easily checks that $(e_n r_n)(\phi ^N) = \phi ^M$ and that $(r_n e_n)(\phi ^M) = \phi ^M$ for every $\phi \in F_n$ :

$$\begin{align*}(e_nr_n)(\phi ^N)& = e_n(\{\vec {a}\in |M|^{n+1} : N\vDash \phi [\vec {a}]\}) = e_n(\phi ^M) = (\phi ^{\underline {M}})^N\\&= \{\vec {b}\in |N|^{n+1} : N\vDash \phi ^{\underline {M}} [\vec {b}]\}= \{\vec {a}\in |M|^{n+1} : N\vDash \phi ^{\underline {M}} [\vec {a}]\}= \phi ^M, \\(r_ne_n)(\phi ^M) &= r_n((\phi ^{\underline {M}})^N) = \{\vec {a}\in |M|^{n+1} : N\vDash \phi ^{\underline {M}} [\vec {a}]\} = \phi ^M.\end{align*}$$

Consequently, $r_* : N_*\rightarrow M_*$ defines a retraction of $N_*$ over $M_*$ . Conversely, the very condition $r_* e_* = \text {id}_{M_*}$ supposes that $e_*$ is well defined, and thus that M is a definable elementary substructure of N. As a matter of fact, there must be some L-formula $\chi (v_0)$ such that $e_0(|M|) = e_0((v_0 = v_0)^M) = \chi (v_0)^N$ .⊣

The existence of a retraction $r_* : N_*\rightarrow M_*$ implies in particular that $M_*$ and $N_*$ are homotopy equivalent.Footnote 5 The emerging analogy is thus between elementary equivalence and homotopy equivalence: as two L-structures are elementary equivalent iff they have isomorphic ultrapowers, two topological spaces are homotopy equivalent iff they are deformation retracts of a single (up to isomorphism) larger space, namely the “mapping cylinder” of their homotopy.

2.2 Types

Another application of the simplicial framework deserves attention, namely types. Given a complete theory T in L, an n-type is a complete theory in $\text {L}\cup \{v_0, \ldots, v_{n-1}\}$ which is consistent with T and whose sentences are considered up to graded logical equivalence. The set of all n-types is written $S_{n-1}(T)$ . If T is the theory of some L-structure M and $A\subseteq |M|$ , the set of all n-types over A (n-types in the language $\text {L}(A)$ ) is written $S_n(A)$ .

Definition 2.8. The natural extensions of the maps $\exists _i : F_n\rightarrow F_{n-1}$ and $E_j : F_n\rightarrow F_{n+1}$ to types as sets of L-formulae are also written $\exists _i : S_{n}(T)\rightarrow S_{n-1}(T)$ and $E_j : S_{n}(T)\rightarrow S_{n+1}(T)$ .

Proposition 2.9. The structure $S_*(T) = \langle S_n(T), (\exists _i)_{0\leq i\leq n}, (E_j)_{0\leq j\leq n} \rangle _{n\in \mathbb {N}}$ is a simplicial set.

Proof Given any type $q\in S_n(T)$ , there is a type $p\in S_{n+1}(T)$ such that $p\supseteq \{E_i(\psi ) : \psi \in q\}$ . Indeed, since every finite subset of q is consistent and realized in some model of T, this is obviously the case of $\{E_i(\psi ) : \psi \in q\}$ too. But then $\exists _ip\supseteq \{(\exists _iE_i)(\psi ) : \psi \in q\} = q$ . Since q is supposed to be complete, it follows that $\exists _ip = q$ . (However, all the L-formulae in q are not in an existentially quantified form: they are only so up to graded logical equivalence —see Definition 1.3.) As to the simplicial identities, they are verified by straightforward extension from L-formulae to types. ⊣

Theorem 2.10. $S_*(T)$ is fibrant.

Proof In combinatorial terms, one has to show, given n simplices $p_0$ , …, $\widehat {p_k}$ , …, $p_n\in S_{n-1}(T)$ such that $\exists _ip_j = \exists _{j-1}p_i$ for all $i, j\neq k$ with $i<j$ , the existence of $p\in S_n(T)$ such that $\exists _i p = p_i$ for all $i\neq k$ . But the complete ( $n+1$ )-type p generated by the family $\{(v_j = v_j\ \wedge \ \phi ^j(v_0, \ldots , v_{j-1}, v_{j+1}, \ldots , v_n)) : \phi ^j\in p_j,\ j\neq k\}$ matches that condition. A preliminary remark is in order: the hypothesis $\exists _ip_j = \exists _{j-1}p_i$ (for $i<j$ and $i,j\neq k$ ) means that, for any L-formula $\phi ^j(v_0, \ldots , v_{n-1})\in p_j$ , there exists an L-formula $\psi _{\phi }^i(v_0, \ldots , v_{n-1})\in p_i$ such that $\exists x \phi ^j(v_0, \ldots , v_{i-1}, x, v_{i}, \ldots , v_{n-2})$ is logically equivalent (in the graded sense) with $\exists x \psi _{\phi }^i(v_0, \ldots , v_{j-2}, x, v_{j-1}, \ldots , v_{n-2})$ . Renaming ‘ $v_{j-1}$ ’, …, ‘ $v_{n-2}$ ’ is immaterial, so actually, for any variable symbols ‘ $u_1$ ’, …, ‘ $u_{n-1}$ ’:

$$\begin{align*}&\exists x \phi ^j(u_1, \ldots , u_{i}, x, u_{i+1}, \ldots , u_{j-1}, \ldots , u_{n-1})\\&= \exists x \psi _{\phi }^i(u_1, \ldots , u_{j-1}, x, u_{j}, \ldots , u_{n-1}).\end{align*}$$

In the same way, for any L-formula $\phi ^i\in p_i$ , there exists an L-formula $\chi _{\phi }^j\in p_j$ such that:

$$\begin{align*}& \exists x \phi ^i(u_1, \ldots , u_{j-1}, x, u_{j}, \ldots , u_{n-1})\\&=\exists x \chi _{\phi }^j(u_1, \ldots , u_{i}, x, u_{i+1}, \ldots , u_{j-1}, \ldots , u_{n-1}).\end{align*}$$

As a consequence, for any $\phi ^j\in p_j$ with $j>0$ and $j\neq k$ :

$$ \begin{align*} & \exists_0((v_j = v_j \wedge \phi^j(v_0, \ldots, v_{j-1}, v_{j+1}, \ldots, v_n)))\\ &= (v_{j-1} = v_{j-1} \wedge \exists x\phi^j(x, v_0, \ldots, v_{j-2}, v_{j}, \ldots, v_{n-1}))\\ &= (v_{j-1} = v_{j-1} \wedge (\exists x\phi^j)[v_j/v_{j-1}, \ldots, v_{n-1}/v_{n-2}])\\ &= (v_{j-1} = v_{j-1} \wedge (\exists x\psi_{\phi}^0)[v_j/v_{j-1}, \ldots, v_{n-1}/v_{n-2}])\\ &= (v_{j-1} = v_{j-1} \wedge \exists x\psi_{\phi}^0(v_0, \ldots, v_{j-2}, x, v_{j}, \ldots, v_{n-1})). \end{align*} $$

But the latter L-formula is obviously a member of $p_0$ . As to the case $j=0$ (supposing $k\neq 0$ ), $\exists _0((v_0 = v_0 \wedge \phi ^0(v_1, \ldots , v_n))) = \phi ^0(v_0, \ldots , v_{n-1})\in p_0$ . Gathering both cases ( $j>0$ and $j=0$ ), one has that $\exists _0p = p_0$ (for $k\neq 0$ ).

Let us prove now that $\exists _n p = p_n$ (for $k\neq n$ ). As above, for any $\phi ^i\in p_i$ with $i<n$ and $i\neq k$ :

$$ \begin{align*} & \exists_n((v_i = v_i \wedge \phi^i(v_0, \ldots, v_{i-1}, v_{i+1}, \ldots, v_n)))\\ &= (v_{i} = v_{i} \wedge \exists x\phi^i(v_0, \ldots, v_{i-1}, v_{i+1}, \ldots, v_{n-1}, x))\\ &= (v_{i} = v_{i} \wedge (\exists_{n-1}\phi^i)[v_{i+1}/v_{i}, \ldots, v_{n-1}/v_{n-2}])\\ &= (v_{i} = v_{i} \wedge (\exists_{i}\chi_{\phi}^n)[v_{i+1}/v_{i}, \ldots, v_{n-1}/v_{n-2}])\\ &= (v_{i} = v_{i} \wedge \exists x\chi_{\phi}^n(v_0, \ldots, v_{i-1}, x, v_{i+1}, \ldots, v_{n-1}))\in p_n. \end{align*} $$

As to the case $i = n$ , again $\exists _n((v_n = v_n \wedge \phi ^n(v_0, \ldots , v_{n-1}))) = \phi ^n(v_0, \ldots , v_{n-1})\in p_n$ . So $\exists _n p = p_n$ (for $k\neq n$ ). The proof of the other identities $\exists _jp = p_j$ , for all j such that $0<j<n$ and $j\neq k$ , is analogous. ⊣

Given a simplicial set X and $x, y\in X_0$ , x is said to be homotopic to y, written $x\sim y$ , iff there exists a one-simplex $z\in X_1$ such that $d_1z = x$ and $d_0z = y$ . The one-simplex is said to be a path from x to y.

Corollary 2.11. (See [Reference Hovey6], Lemmas 3.4.2, 3.4.3 and 3.6.3)

The homotopy relation between $1$ -types in $S_0(T)$ is an equivalence relation. The quotient $S_0(T)/\sim $ is the set of path components of $S_*(T)$ . For $p\in S_0(T)$ , the higher homotopy groups $\pi _n(S_*(T), p)$ can be defined as well.Footnote 6

Up to now, a simplicial set has been associated to a whole space of types $\langle S_n(T) : n\in \mathbb {N}\rangle $ . A single type, however, also induces a simplicial set, provided it has the following property: given some L-structure M and $A\subseteq |M|$ , $p\in S_n(A)$ is definable iff, for each $\text {L}$ -formula $\phi (v_0, \ldots , v_n, v_{n+1}, \ldots , v_{n+m})$ , there is an $\text {L}(A)$ -formula $(\delta _p\phi )(v_0, \ldots , v_{m-1})$ such that, for any tuple $(a_0, \ldots , a_{m-1})$ of elements of $|M|$ , $\phi (v_0, \ldots , v_n, \overline {a_0}, \ldots , \overline {a_{m-1}})\in p$ iff $M\vDash (\delta _p\phi )(v_0, \ldots , v_{m-1}) [a_0, \ldots , a_{m-1}]$ . The operator $\delta ^p$ is called the definition of p.

Proposition 2.12. Given an L-structure M, any definable type $p\in S_0(M)$ induces a fibrant simplicial set.

Proof Let $p\in S_0(M)$ be definable: for each L-formula $\phi (v_0, \ldots , v_n)$ , there is an L-formula $d^{p, n}_i(\phi )(v_0, \ldots , v_{n-1})$ , possibly with parameters in M, such that, for any $(a_0, \ldots , a_{n-1})\in |M|^n$ , $\phi (\overline {a_0}, \ldots , \overline {a_{i-1}}, v_0, \overline {a_i}, \ldots , \overline {a_{n-1}})\in p$ iff $M\vDash d^{p, n}_i(\phi ) [a_0, \ldots , a_{n-1}]$ . One has: $d^{p, n}_i(\neg \phi ) = \neg d^{p, n}_i(\phi )$ and $d^{p, n}_i(\phi \wedge \psi ) = d^{p, n}_i(\phi )\wedge d^{p, n}_i(\psi )$ . Besides, each set $F_n$ can be turned into a group, with $\nleftrightarrow $ as the binary law, so each $d^{p, n}_i$ is actually a group homomorphism. Hence $F_*^p := \langle \langle F_n, \nleftrightarrow , \bot \rangle , (d^{p, n}_i)_{0\leq i\leq n}, (s^n_j)_{0\leq j\leq n}\rangle _{n\in \mathbb {N}}$ is a simplicial group. Since, according to Moore’s theorem,Footnote 7 the underlying simplicial set of a simplicial group is fibrant, one concludes that $F_*^p$ is fibrant. ⊣

Proposition 2.13. Any definable type $p\in S_0(M)$ induces a chain complex.

Proof From $\partial ^{p, n}(\phi ) := (((d^{p, n}_0(\phi )\nleftrightarrow d^{p, n}_1(\phi ))\nleftrightarrow \cdots )\nleftrightarrow d^{p, n}_n(\phi ))$ (for each $n\in \mathbb {N}$ and each $\phi \in F_n$ ), one gets a differentiation operator $\partial ^{p, \ast }$ : $\partial ^{p, n+1}\circ \partial ^{p, n}\equiv \bot $ for every $n\in \mathbb {N}$ . In other words, $\langle F^p_*, \partial ^{p, \ast }\rangle $ is a chain complex. ⊣

2.3 Another functorial correspondence

Let $\text {L} = \{R^{(n)}_k : n, k\in \mathbb {N}\}\cup \{< \}$ be a fixed countable first-order language, whose signature contains a distinguished binary relation ‘ $<$ ’. As above, “definability” will mean, unless otherwise stated, definability with parameters. Moreover, all types in $S_n(M)$ , for each $n\in \mathbb {N}$ , will be allowed to consist of $\text {L}(M)$ -formulae. The simplicial $F_*^M$ associated to any L-structure M (Definition 2.1) does not encode enough of M. In order to improve the functorial correspondence between L-structures and simplicial sets, the simplicial counterpart of the definable subsets of M should be able to mention which of the latter are nonempty. Hence the next definition.

Definition 2.14. A simplicial set with distinguished simplices, or d-simplicial set for short, is a simplicial set X with a subset $X_n'\subseteq X_n$ of distinguished n-simplices for each $n\in \mathbb {N}$ .

A d-simplicial map $f : X\rightarrow Y$ is a simplicial map $f : X\rightarrow Y$ such that, for each $n\in \mathbb {N}$ , $f_n$ preserves distinguished n-simplices, i.e., $f_n(X_n')\subseteq Y_n'$ .

Lemma 2.15. Any L-structure M induces a d-simplicial set $M^*$ .

Proof Let us define a complex of M as a finite collection C of definable subsets of M (i.e., of definable subsets of Cartesian powers of $|M|$ ) which is closed under the face operator: for any definable subset $(\phi (v_0, \ldots , v_n, \overline {a_1}, \ldots , \overline {a_l}))^M\in D_n(M)$ (with parameters $a_1, \ldots , a_l$ ), $\phi ^M\in C$ implies both

  • $\exists _i^M(\phi ^M) = (\exists x\phi (v_0, \ldots , v_{i-1}, x, v_i, \ldots , v_{n-1}, \overline {a_1}, \ldots , \overline {a_l}))^M\in C$ and

  • $(\exists y\phi (v_0, \ldots , v_n, \overline {a_1}, \ldots , \overline {a_{j-1}}, y, \overline {a_{j+1}}, \ldots , \overline {a_l}))^M\in C$ ,

for any i with $0\leq i\leq n$ and any j with $1\leq j\leq l$ . A complex C of M is of dimension n if n is the largest integer k such that $D_k(M)\cap C\neq \emptyset $ . Given a collection $\Gamma $ of definable subsets of M, the complex generated by $\Gamma $ , written $\Gamma ^c$ , is the intersection of all complexes of M containing all members of $\Gamma $ . A complex generated by a single definable subset is called a principal complex. For each $n\in \mathbb {N}$ , the set of n-simplices of $M^*$ is defined as the set $K(M)_n$ of all complexes of M of dimension n. The maps $\exists ^{M}_i : D_n(M)\rightarrow D_{n-1}(M)$ and $E_j^{M} : D_n(M)\rightarrow D_{n+1}(M)$ , as defined earlier, are simply extended to collections of parametrically definable subsets. (Note that, given any definable subset $\phi ^M$ of M, $\exists _i(\{\phi ^M\}^c) = \{\exists _i(\phi ^M)\}^c$ and $E_j(\{\phi ^M\}^c) = \{E_j(\phi ^M)\}^c$ , so that $F_*^M$ can be identified with a simplicial subset of $M^*$ .) The distinguished n-simplices of $M^*$ are the principal complexes of M generated by a nonempty subset of $|M|^{n+1}$ defined by an atomic formula without parameters. Their set is written $K(M)^{\prime}_n$ .⊣

Definition 2.16. Let M and N be two L-structures. For any homomorphism $f : M\rightarrow N$ , ${{\hspace {-0.0001pt}}}_nf : D_n(M)\rightarrow D_n(N)$ , for each $n\in\mathbb{N}$ , is the map sending each

$$\begin{align*}(\phi (v_0, \ldots , v_n, \overline {a_1}, \ldots , \overline {a_{k}}))^M\end{align*}$$

to

$$\begin{align*}(\phi (v_0, \ldots , v_n, \overline {f(a_1)}, \ldots , \overline {f(a_{k})}))^N.\end{align*}$$

The extension $C\in K(M)_n\mapsto \{{{\hspace {-0.0001pt}}}_nf(\phi ^M) : \phi ^M\in C\}\in K(N)_n$ of ${{\hspace {-0.0001pt}}}_nf$ shall also be written ${{\hspace {-0.0001pt}}}_nf$ .

In case $f : M\rightarrow N$ is an elementary embedding, it is clear that each ${{\hspace {-0.0001pt}}}_nf : D_n(M)\rightarrow D_n(N)$ will be injective. Now a simplicial map $f : X\rightarrow Y$ is a cofibration iff $f_n : X_n\rightarrow Y_n$ is an injection for all $n\in \mathbb {N}$ . Hence the next lemma.

Lemma 2.17. Let M and N be two L-structures and an elementary embedding. The family $({{\hspace {-0.0001pt}}}_nf : K(M)_n\rightarrow K(N)_n)_{n\in \mathbb {N}}$ defines a d-simplicial map $f_! : M^*\rightarrow N^*$ which, as a simplicial map, is a cofibration.

Corollary 2.18. The mapping $M\mapsto M^*,\ f\mapsto f_!$ defines a functor $F_0$ from the category $\mathbf{L}\hbox{-}\mathbf{Strs}$ of L-structures and elementary embeddings, to the category ${\mathbf{d}\hbox{-}\mathbf{sSets}}$ of d-simplicial sets and d-simplicial maps which are cofibrations.

Definition 2.19. Let M and N be two L-structures and an elementary embedding. For each $n\in \mathbb {N}$ , the restriction morphism ${{\hspace {-0.0001pt}}}^nf : D_n(N)\rightarrow D_n(M)$ is defined as follows: given any L-formula

$$\begin{align*}\phi (v_0, \ldots , v_n, \overline {f(a_1)}, \ldots , \overline {f(a_k)}, \overline {b_1}, \ldots , \overline {b_l})\end{align*}$$

in $F_n$ with parameters in N, where $b_1, \ldots , b_l$ are parameters of $\phi $ (if any) not in the range of f, ${{\hspace {-0.0001pt}}}^nf$ sends

$$\begin{align*}(\phi (v_0, \ldots , v_n, \overline {f(a_1)}, \ldots , \overline {f(a_k)}, \overline {b_1}, \ldots , \overline {b_l}))^N\end{align*}$$

to

$$\begin{align*}(\exists x_1\ldots \exists x_l \phi (v_0, \ldots , v_n, \overline {a_1}, \ldots , \overline {a_k}, x_1, \ldots , x_l))^M.\end{align*}$$

The extension $C\in K(N)_n\mapsto \{{{\hspace {-0.0001pt}}}^nf(\phi ^N) : \phi ^N\in C\}\in K(M)_n$ of ${{\hspace {-0.0001pt}}}^nf$ shall also be written ${{\hspace {-0.0001pt}}}^nf$ .

In case $f : M\rightarrow N$ is an elementary embedding, the family $({{\hspace {-0.0001pt}}}^nf)_{n\in \mathbb {N}}$ contravariantly associated to f will be something like the opposite of a cofibration. The right notion here is that of “trivial fibration.” A trivial fibration is a simplicial map $p : X\rightarrow Y$ such that: (i) $p_0 : X_0\rightarrow Y_0$ is surjective; (ii) for every $n\geq 1$ , given any $(n+1)$ -tuple $x_0, \ldots , x_n$ of $(n-1)$ -simplices of X such that $d_ix_j = d_{j-1}x_i$ (for all $i, j$ with $0\leq i< j\leq n$ ) and any n-simplex y of Y such that $d_iy = p_{n-1}(x_i)$ (for all i with $0\leq i\leq n$ ), there exists an n-simplex x of X such that $d_ix = x_i$ (for all i with $0\leq i\leq n$ ) and $p_n(x) = y$ .

Lemma 2.20. Let M and N be two L-structures and an elementary embedding. The family $({{\hspace {-0.0001pt}}}^nf : K(N)_n\rightarrow K(M)_n)_{n\in \mathbb {N}}$ defines a d-simplicial map $f^* : N^*\rightarrow M^*$ which, as a simplicial map, is a trivial fibration.

Proof Since f is an elementary embedding, $f^*$ obviously satisfies ${{\hspace {-0.0001pt}}}^0f(K_0(N))\subseteq K_0(M)$ , so it only remains to prove the second condition (ii) in the definition above. For the sake of simplicity, let us consider only the case $n = 1$ and let us suppose that the given simplices are principal complexes. The treatment of the general case is analogous, only more complicated. So, using ‘ $\alpha $ ’, ‘ $\beta $ ’ and ‘ $\gamma $ ’ to indicate arbitrary sequences of parameters, let $x_0 = \{(\phi _0(v_0, \overline {f(a_{\alpha })}, \overline {b_{\alpha }}))^N\}^c$ , $x_1 = \{(\phi _1(v_0, \overline {f(a_{\beta })}, \overline {b_{\beta }}))^N\}^c$ and $y = \{(\psi (v_0, v_1, \overline {c_{\gamma }}))^M\}^c$ be such that:

  • $M\vDash \exists x\phi _0(x, \overline {f(a_{\alpha })}, \overline {b_{\alpha }})\leftrightarrow \exists x\phi _1(x, \overline {f(a_{\beta })}, \overline {b_{\beta }})$ ,

  • $M\vDash \exists y\psi (y, v_0, \overline {c_{\gamma }})\leftrightarrow \exists x_{\alpha }\phi _0(v_0, \overline {a_{\alpha }}, x_{\alpha })$ ,

  • $M\vDash \exists z\psi (v_0, z, \overline {c_{\gamma }})\leftrightarrow \exists x_{\beta }\phi _1(v_0, \overline {a_{\beta }}, x_{\beta })$ .

Let x be $\{(\psi (v_0, v_1, \overline {f(c_{\gamma })}))^N, (\phi _0(v_1, \overline {f(a_{\alpha })}, \overline {b_{\alpha }}))^N, (\phi _1(v_0, \overline {f(a_{\beta })}, \overline {b_{\beta }}))^N\}^c$ . One has:

  • $p_n(x) = \{(\psi (v_0, v_1, \overline {c_{\gamma }}))^M, (\exists x_{\alpha }\phi _0(v_1, \overline {a_{\alpha }}, x_{\alpha }))^M, (\exists x_{\beta }\phi _1(v_0, \overline {a_{\beta }}, x_{\beta }))^M\}^c = \{(\psi (v_0, v_1, \overline {c_{\gamma }}))^M, (\exists y\psi (y, v_1, \overline {c_{\gamma }}))^M, (\exists z\psi (v_0, z, \overline {c_{\gamma }}))^M\}^c = y$

  • $\exists _0x = \{(\exists x\psi (x, v_0, \overline {f(c_{\gamma })}))^N, (\phi _0(v_0, \overline {f(a_{\alpha })}, \overline {b_{\alpha }}))^N, (\exists x\phi _1(x, \overline {f(a_{\beta })}, \overline {b_{\beta }}))^N\}^c = \{(\exists x_{\alpha }\phi _0(v_0, \overline {a_{\alpha }}, x_{\alpha }))^N, (\phi _0(v_0, \overline {f(a_{\alpha })}, \overline {b_{\alpha }}))^N, (\exists x\phi _0(x, \overline {f(a_{\alpha })}, \overline {b_{\alpha }}))^N\}^c = x_0$

  • $\exists _1x = x_1$ in the same way as just above.⊣

Corollary 2.21. The mapping $(-)^*$ defines a contravariant functor F from the category ${\mathbf{L}\hbox{-}\mathbf{Strs}}$ of L-structures and elementary embeddings, to the category ${\mathbf{d}\hbox{-}\mathbf{sSets}}$ of simplicial sets with distinguished simplices and d-simplicial maps which are, as simplicial maps, trivial fibrations.

3 O-minimal structures as simplicial sets

Given the existence of the functor F, two natural questions are whether a functor can be defined in the other direction (even if it means restricting the two categories ${\mathbf{L}\hbox{-}\mathbf{Strs}}$ and ${\mathbf{d}\hbox{-}\mathbf{sSets}}$ to subcategories thereof) and, if so, whether the resulting pair of functors induces an adjunction. The purpose of the rest of this paper is to give a positive answer to both questions (Theorems 3.15 and 3.19).

3.1 O-minimality

It appears that the right restriction of ${\mathbf{L}\hbox{-}\mathbf{Strs}}$ is to o-minimal L-structures. An L-structure M is said to be o-minimal iff $<^M$ is a dense linear order and every definable subset of $\langle M, <^M\rangle $ is a finite union of singletons and open $<^M$ -intervals (with endpoints in $|M|\cup \{-\infty , +\infty \}$ ). The full subcategory of ${\mathbf{L}\hbox{-}\mathbf{Strs}}$ composed of all o-minimal L-structures is written ${\mathbf{oL}\hbox{-}\mathbf{Strs}}$ .

Lemma 3.1. Suppose X is a simplicial set. Then any linear order on the set $X_0$ of its vertices induces on each $X_n$ ( $n\in \mathbb {N}$ ) a linear order.

Proof The linear ordering of every $X_n$ is defined by induction. Let us suppose that a linear order $<_n$ has been defined on $X_n$ . For each $E\in X_{n+1}$ , let $s(E)$ be the sequence $(d_i(E))_{0\leq i\leq n+1}$ . An order $<_{n+1}$ is then defined on $X_{n+1}$ as follows: $E <_{n+1} E'$ iff $s(E) <^l_n s(E')$ , where $<^l_n$ is the lexicographic order induced by $<_n$ . Since the lexicographical order on a family of linearly ordered sets is a linear extension of their product order, by induction each $<_n$ is a linear order. ⊣

Definition 3.2. A dense inductively linearly ordered simplicial set with countably many distinguished simplices, or dilo-simplicial set for short, is a d-simplicial set X such that: (i) its set $X_0$ of vertices is a dense linear order; (ii) for each $n\geq 1$ , $X_n$ is endowed with the linear order induced by that on $X_0$ as above; and (iii) for each $n\in \mathbb {N}$ , the subset $X_n'\subseteq X_n $ of distinguished n-simplices is countable.

A morphism $f : X\rightarrow Y$ between dilo-simplicial sets, or dilo-simplicial map for short, is a d-simplicial map $f : X\rightarrow Y$ such that, for each $n\in \mathbb {N}$ , $f_n|_{X_n'}$ is an order isomorphism from $X_n'$ to $Y_n'$ .

Lemma 3.3. For any o-minimal L-structure M, $M^*$ is a dilo-simplicial set.

Proof Since $M^*$ is a d-simplicial set, it is only necessary to define a dense linear ordering of $D_0(M)$ . Each definable subset A of $|M|$ can be written $\bigcup _{i = 0}^{n_A} \overline {]}a_i, a_{i+1}\overline {[}$ for some $n_A\in \mathbb {N}$ , with $a_i\leq ^M a_{i+1}$ and the convention:

$$\begin{align*}\overline {]}a_i, a_{i+1}\overline {[} =\begin {cases} ]a_i, a_{i+1}[\ \ \text {if}\ a_i<^M a_{i+1},\\ \{a_i\}\ \ \text {if}\ a_i = a_{i+1}.\end {cases}\end{align*}$$

The sequence $(a_i)_{0\leq i\leq n_A}$ is called the sequence associated to A and written $\sigma (A)$ . Then $D_0(M)$ is ordered as follows: $A <^0 A'$ iff $\sigma (A) <^l \sigma (A')$ , where $<^l$ is the lexicographic order induced by $<^M$ . Finally $K(M)_0$ is ordered as follows: $C<_0 C'$ iff $C_0<^* {C^{\prime }}_0$ , where $C_0 = C\cap D_0(M)$ , ${C^{\prime }}_0 = C^{\prime }\cap D_0(M)$ and $<^*$ is the lexicographic order induced by $< ^0$ on $\wp _f(D_0(M))$ . The resulting ordering $<_0$ is a dense linear order. ⊣

A standard result about trivial fibrationsFootnote 8 is that any trivial fibration $p : X\rightarrow Y$ has the right lifting property w.r.t. all cofibrations, which means that, given any cofibration $i : U\rightarrow V$ , any commutative solid arrow diagram

admits of a lifting $h : V\rightarrow X$ satisfying $h\circ i = f$ and $p\circ h = g$ . In particular, the lifting problem

has a solution s. Hence any trivial fibration $p : X\rightarrow Y$ has a section: put differently, there exists a simplicial map $s : Y\rightarrow X$ such that $\langle s, p\rangle $ constitutes a retraction pair.

Definition 3.4. Let ${\mathbf{dilo}\hbox{-}\mathbf{sSets}}$ be the category —a subcategory of ${\mathbf{d}\hbox{-}\mathbf{sSets}}$ — whose objects are all dilo-simplicial sets and whose morphisms are all dilo-simplicial maps which, as simplicial maps, are trivial fibrations with a unique section. So each morphism p in ${\mathbf{dilo}\hbox{-}\mathbf{sSets}}$ actually consists of a specific retraction pair $\langle i_p : Y\rightarrow X, p : X\rightarrow Y\rangle $ .

Lemma 3.5. Let M and N be two L-structures. For any elementary embedding , $\langle f_!, f^*\rangle $ is a retraction pair and $f_!$ is the unique simplicial map g such that $\langle g, f^*\rangle $ is a retraction pair.

Proof For each $n\in \mathbb {N}$ , the equality ${{\hspace {-0.0001pt}}}^nf\circ {{\hspace {-0.0001pt}}}_nf = \text {id}_{D_n(M)}$ and thus the equality ${{\hspace {-0.0001pt}}}^nf\circ {{\hspace {-0.0001pt}}}_nf = \text {id}_{K(M)_n}$ follow directly from the definitions of ${{\hspace {-0.0001pt}}}_nf$ and ${{\hspace {-0.0001pt}}}^nf$ . The uniqueness of $f_!$ follows from the hypothesis that f is an elementary embedding. ⊣

Corollary 3.6. The restriction $\tilde {F}$ of F to $\mathbf{{oL}}\hbox{-}\mathbf{{Strs}^{\scriptsize op}}$ induces a functor $F'$ from $\mathbf{{oL}}\hbox{-}\mathbf{{Strs}^{\scriptsize op}}$ to $\mathbf{dilo}\hbox{-}\mathbf{sSets}$ .

One is now interested in a functor going in the opposite direction. But, given any dilo-simplicial set X, there is an L-structure $\widehat {X}$ naturally associated to it. Indeed, any simplicial set X gives rise, in a canonical way, to a topological space $||X||$ , called its geometric realization. Actually, from the beginning X is just a recipe for joining polyhedra together so as to obtain $||X||$ , of which X can be conceived of as a triangulation. The idea is to realize each n-simplex of X as the n-cell (“standard n-simplex”) $||\Delta ^n|| = \{(t_0, \ldots , t_n)\in \mathbb {R}^{n+1} : t_i\geq 0, \sum _{i=0}^n t_i = 1\}$ , and to keep track of the incidences between n-simplices:

$$ \begin{align*}||X|| := \underset{\Delta^n\rightarrow X}{\varinjlim}||\Delta^n|| = \left(\coprod_{n\in\mathbb{N}} X_n\times ||\Delta^n||\right) / {\sim},\end{align*} $$

where $\sim $ expresses the identification of parts that are glued together.Footnote 9 Specifically, define $\delta _i : ||\Delta ^{n-1}||\rightarrow ||\Delta ^{n}||$ and $\sigma _j : ||\Delta ^{n+1}||\rightarrow ||\Delta ^{n}||$ by:

$\delta _i(t_0, \ldots , t_{n-1}) = (t_0, \ldots , t_{i-1}, 0, t_i, \ldots , t_{n-1})$

$\sigma _j(t_0, \ldots , t_{n+1}) = (t_0, \ldots , t_{i-1}, t_i+t_{i+1}, t_{i+2}, t_{n-1})$ .

Then $\sim $ is the equivalence relation generated by $(d_ix, u)\sim (x, \delta _iu)$ and $(s_jx, v)\sim (x, \sigma _jv)$ , for all $x\in X_n$ , $u\in ||\Delta ^{n}||$ and $v\in ||\Delta ^{n+1}||$ . Thus, an element of $||X||$ is of the form $\overline {(x, u)}$ with $x\in X_n$ and $u = (t_0, \ldots , t_n)$ . Note that, by construction, the topological space $||X||$ is a CW complex. On the other hand, any simplicial map $g : X\rightarrow Y$ induces a continuous map $||g|| : ||X||\rightarrow ||Y||$ , given by: $||g||(\overline {(x, u)}) = \overline {(g_n(x), u)}$ . This makes geometric realization functorial.

Definition 3.7. Let X be a dilo-simplicial set and $X_n^k$ the k-th distinguished n-simplex of X, i.e., the k-th member of $X_n'$ . The image of $X_n^k$ in $||X||$ , written $||X||_n^k$ , is the image of the composite map:

.

This definition generalizes to any set Y of n-simplices of X: the image $||Y||$ of Y is the union of the images of all the elements of Y.

Definition 3.8. Any dilo-simplicial set X gives rise to an L-structure $\widehat {X}$ defined as follows:

  • $|\widehat {X}|$ is $||X||$ , the geometric realization of X;

  • $(R^{(0)}_k)^{\widehat {X}}$ is the image of the k-th distinguished vertex in $X_0$ ;

  • $(R^{(n)}_k)^{\widehat {X}}$ , for $n\geq 1$ , is $(||X||_n^{k})^n$ .

Example 3.9. Let $X_{S^2}$ be the canonical simplicial set whose realization is the two-dimensional sphere $S^2$ . Since $S^2$ is obtained by glueing the boundary of a copy of $||\Delta ^2||$ into a single point, $X_{S^2}$ has a single non-degenerate $0$ -simplex (vertex $\ast $ ), a single non-degenerate $2$ -simplex, and no other non-degenerate simplices. Hence $|\widehat {X_{S^2}}|$ is $S^2$ and, for all $k\in \mathbb {N}$ , $(R^{(0)}_k)^{\widehat {X_{S^2}}}$ is $\ast $ , $(R^{(1)}_k)^{\widehat {X_{S^2}}}$ is $\{\ast \}$ and $(R^{(2)}_k)^{\widehat {X_{S^2}}}$ is $S^2\times S^2$ .

Definition 3.10. Let $\langle i_p : Y\rightarrow X, p : X\rightarrow Y\rangle $ be a morphism in ${\mathbf{dilo}\hbox{-}\mathbf{sSets}}$ . One then defines $\widehat {p} := ||i_p|| : ||Y||\rightarrow ||X||$ .

As a result of the definition of a trivial fibration $p : X\rightarrow Y$ , $||p||$ has the right lifting property w.r.t. all cellular inclusions, i.e., all inclusions where A is a subcomplex of some CW complex C.Footnote 10 This result will be of use in the proof of the Lemma below.

Lemma 3.11. For any morphism $p : X\rightarrow Y$ in ${\mathbf{dilo}\hbox{-}\mathbf{sSets}}$ , $\widehat {p} : ||Y||\rightarrow ||X||$ defines an elementary embedding $\widehat {p} : \widehat {Y}\rightarrow \widehat {X}$ .

Proof Suppose $(y_1, \ldots , y_n)\in (R^{(n)}_k)^{\widehat {Y}} = (||Y||_n^k)^n$ . Each $y_i$ is of the form $\overline {(\beta _i, u_i)}$ with $\beta _i = Y_n^k$ . By definition of a dilo-simplicial map, $p_n\circ (i_p)_n = \text {id}_{Y_n}$ and thus $\widehat {p}(y_i) = \overline {(\alpha _i, u_i)}$ with $\alpha _i = (i_p)_n(\beta _i) = p_n^{-1}(\beta _i) = X_n^k$ , so $\widehat {p}(y_i)\in ||X||_n^k$ , and so $(\widehat {p}(y_1), \ldots , \widehat {p}(y_n))\in (R^{(n)}_k)^{\widehat {X}}$ . Suppose now that $||Y||\subseteq ||X||$ and that $\widehat {X}\vDash \phi (v_0, \overline {y_1}, \ldots , \overline {y_n}) [x]$ with $x\in ||X||$ and $y_1, \ldots , y_n\in ||Y||$ . Let C be $(\phi (v_0, \overline {y_1}, \ldots , \overline {y_n}))^{\widehat {X}}$ , so that $x\in C$ , and let $p'$ be the restriction of $||p||$ to C. Since $p'\circ x$ is a cellular inclusion and p is a trivial fibration, $||p||$ has the right lifting property w.r.t. $p'\circ x$ , so there exists a lifting h in the following diagram:

The map h satisfies $h\circ p'\circ x = i\circ x$ and $||p||\circ h = \text {id}_{||Y||}$ . The first condition ensures that there exists $y\in |\widehat {Y}|$ such that $h(y)\in C$ and, by the uniqueness of the section of p, the second ensures that $h(y) = y$ . This leads to the existence of $y\in |\widehat {Y}|$ such that $\widehat {X}\vDash \phi (v_0, \overline {y_1}, \ldots , \overline {y_n}) [y]$ . As a result, the Tarski–Vaught test is verified. ⊣

Corollary 3.12. The mapping $\widehat {(-)}$ defines a functor G  from ${\mathbf{dilo}\hbox{-}\mathbf{sSets}}^{\scriptsize op}$ to $\mathbf{L}\hbox{-}\mathbf{Strs}$ .

As an extension of the simplicial setup presented here, a connection can be made with the “independence property” in stability theory. An L-formula $\phi (\vec {x}, \vec {y})$ does not have the independence property in some L-structure M if, for any denumerable set $\{\vec {a_i} : i\in \mathbb {N}\}$ , there is no family $\{\vec {b_J} : J\subseteq \mathbb {N}\}$ such that, for all $i\in \mathbb {N}$ and all $J\subseteq \mathbb {N}$ , $M\vDash \phi (\vec {a_i}, \vec {b_J})$ iff $i\in J$ . A standard result is that an L-formula $\phi (\vec {x}, \vec {y})$ has the independence property in M if and only if, for any indiscernible sequence $(\vec {a_i})_{i \in I}$ and any tuple $\vec {b}$ in M, there is some cofinal subset $I_0\subseteq I$ such that either $M\vDash \phi (\vec {a_i}, \vec {b})$ or $M\vDash \neg \phi (\vec {a_i}, \vec {b})$ holds for every $i\in I_0$ . (Given a linearly ordered set I, a sequence $(a_i)_{i \in I}$ of tuples in M is indiscernible if for every $n\in \mathbb {N}$ , whenever $i_1< \cdots < i_n$ and $j_1 < \cdots < j_n$ are two increasing n-subsequences of I, the two n-tuples $(a_{i_1}, \ldots , a_{i_n})$ and $(a_{j_1}, \ldots , a_{j_n})$ have the same type.)

Proposition 3.13. Given any dilo-simplicial set X, no L-formula has the independence property in $\widehat {X}$ .

Proof The combination of Lemma 12.16 and Theorem 12.18 in [Reference Poizat10] shows that the proof can be confined to L-formulae $\phi (\vec {x}, \vec {y})$ where $|\vec {x}| = |\vec {y}| = 1$ . Now, an indiscernible sequence of elements in $\widehat {X}$ is any sequence of elements of some $||X_n^k||\setminus \bigcup _{0\leq i\leq n} ||d_i(X_n^k)||$ . So an element $b\in |\widehat {X}|$ can make a difference within an indiscernible sequence $(a_i)_{i\in I}$ only if $b = a_{i_0}$ for some $i_0\in I$ , so $I_0$ can be taken to be $\{i\in I : i>i_0\}$ . ⊣

This result, as no L-formula has the independence property in any o-minimal structure (see [Reference van den Dries11], ch. 5.), is actually strenghtened by the following Lemma.

Lemma 3.14. For any dilo-simplicial set X, $\widehat {X}$ is an o-minimal L-structure.

Proof The set $\{(n, x, u) : n\in \mathbb {N}, x\in X_n, u\in ||\Delta ^n||\}$ can be naturally equipped with a linear order: for $n, n'\in \mathbb {N}$ , $x\in X_n$ , $x'\in X_{n^{\prime }}$ , $u = (t_0, \ldots , t_n)\in ||\Delta ^n||$ and $u^{\prime } = \{(t^{\prime}_0, \ldots , t^{\prime}_{n^{\prime }})\in ||\Delta ^{n^{\prime }}||$ , one states:

$$\begin{align*}(n, x, u)\prec (n', x', u')\quad \hbox{iff}\quad\begin{cases}n < n',\\\text{or}\ \ n = n'\ \text{and}\ x < x'\ \text{in}\ X_n,\\\text{or}\ \ n = n',\ x = x'\ \text{and}\ u < u'\ \text{in}\ ||\Delta ^n||,\end {cases}\end{align*}$$

$||\Delta ^n||$ being endowed with the lexicographical order.

Then, given $[x, u] := \text {inf}_{\prec }(\{(n_0, x_0, u_0) : x_0\in X_{n_0}$ and $(x_0, u_0)\in \overline {(x, u)}\})$ , $|\widehat {X}| = ||X||$ is simply ordered by:

$$\begin{align*}\overline {(x, u)} <_X \overline {(x', u')}\quad\hbox{iff}\quad [x, u]\prec [x', u'].\end{align*}$$

One checks that $<_X$ is a dense linear order which turns $\widehat {X}$ into an o-minimal structure. Indeed, because the interpretation $(R^{(1)}_k)^{\widehat {X}}$ of each predicate is a convex set, the structure $\langle \widehat {X}, <_X, (R^{(1)}_k)^{\widehat {X}}\rangle $ is weakly o-minimal, by a result given in [Reference Baisalov and Poizat1], and actually o-minimal, by construction of $<_X$ . Moreover, the n-ary relations for $n\geq 2$ are interpreted by trivial convex subsets (products), so instantiating a variable in an L-formula or quantifying over it simply shifts to another convex subset, and actually to a finite union of singletons and intervals. The same is true about connectives, so any L-formula in $F_0$ is interpreted in $\widehat {X}$ by a finite union of singletons and intervals.⊣

Theorem 3.15. There are two functors $F' : {\mathbf{oL}\hbox{-}\mathbf{Strs}}^{{\mathbf {\scriptsize op}}}\rightarrow {\mathbf{dilo}\hbox{-}\mathbf{sSets}}$ and $G' : {\mathbf{dilo}\hbox{-}\mathbf{sSets}} \rightarrow {{\mathbf{oL}\hbox{-}\mathbf{Strs}}^{{\mathbf {\scriptsize op}}}}$ such that both diagrams below commute:

Remark 3.16. The L-structure $\widehat {X}$ is o-minimal over $||X||$ , but any dilo-simplicial set also induces an o-minimal structure $X_{\mathbb {R}}$ over $\mathbb {R}$ , defined by $|X_{\mathbb {R}}| = \mathbb {R}$ , $(R^{(n)}_k)^{X_{\mathbb {R}}} = |X_n^k|$ . If $p : X\rightarrow Y$ is a trivial fibration, then $p_{\mathbb {R}} := \text {id}_{\mathbb {R}}$ is by construction an elementary embedding of $Y_{\mathbb {R}}$ into $X_{\mathbb {R}}$ . The correspondence $(-)_{\mathbb {R}}$ defines a functor, distinct from $G'$ , from ${\mathbf{dilo}\hbox{-}\mathbf{sSets}}$ to ${\mathbf{oL}\hbox{-}\mathbf{Strs}}^{{\mathbf {\scriptsize op}}}$ .Footnote 11

3.2 Adjunction

The last task ahead is to establish that the functors $F'$ and $G'$ are adjoint functors. To that end, two preliminary results, one model-theoretic, the other homotopy-theoretic, are needed. First, an L-structure P is prime over $A\subseteq |P|$ if, for every $N\vDash \text {Th}(P, \overline {\alpha })_{\alpha \in A}$ , there is an elementary embedding $f : P\rightarrow N$ over A (that is, which is the identity over A). Pillay and SteinhornFootnote 12 proved that, given any o-minimal L-structure M and $A\subseteq |M|$ , $\text {Th}(M)$ has a prime model over A, unique up to A-isomorphism.

Secondly, the simplicial set $\Delta ^1$ is the simplicial set X with only three nondegenerate simplices: two vertices $0$ and $1$ , and one $1$ -simplex whose faces are the two vertices $0$ and $1$ . Its realization $||\Delta ^1|| = \{(t_0, t_1)\in \mathbb {R}^2 : t_0, t_1\geq 0, t_0 + t_1 = 1\}$ is a line segment. Given two simplicial maps $f, g : X\rightarrow Y$ , a homotopy

is a simplicial map $H : X\times \Delta ^1\rightarrow Y$ whose restrictions to $X\times \{0\}$ and to $X\times \{1\}$ are f and g, respectively.Footnote 13 If there is such an homotopy, the maps f and g are said to be homotopic, which is written $f\simeq g$ . Two maps $f : X\rightarrow Y$ and $f' : Y\rightarrow X$ are homotopy inverse to each other if $f'\circ f\simeq \text {id}_X$ and $f\circ f'\simeq \text {id}_Y$ . Now, as noted above,Footnote 14 any trivial fibration has a section. It turns out that any section $s : Y\rightarrow X$ of a trivial fibration $p : X\rightarrow $ is homotopy inverse to it. Indeed, by hypothesis the lifting problemFootnote 15

has a solution H, which amounts to a homotopy $\text {id}_X\simeq s\circ p$ . In particular, for any elementary embedding

$f_!$ is homotopy inverse to $f^*$ .

Lemma 3.17. Let X be a dilo-simplicial set. Then there exists a trivial fibration $\varepsilon : (\widehat {X})^*\rightarrow X$ .

Proof For any vertex x of X, let us consider its realization $\{||x||\} := \{\overline {(x, 0)}\} = \{x\}\times ||\Delta _0||\subseteq ||X||$ , so that the set $X_0$ of the vertices of X corresponds to a subset of $||X||$ . In the same way, each $1$ -simplex $x\in X_1$ corresponds to an ordered pair of vertices, and thus $X_1$ to a binary relation on $||X||$ . More generally, each $x\in X_n$ corresponds to an $(n+1)$ -uple $(x_0, \ldots , x_n)$ of vertices (some of which possibly identical), hence to an $(n+1)$ -uple $(||x_0||, \ldots , ||x_n||)$ in $||X||$ . The substructure $Y = \langle ||X_0||, (R_k^{(n)})^{\widehat {X}}\cap ||X_0||^n\rangle _{k, n\in \mathbb {N}}$ of $\widehat {X}$ defined by restriction to $||X_0|| = \{||x|| : x\in X_0\}$ is actually an elementary substructure of $\widehat {X}$ because, by continuity, every L-formula (with parameters in Y) satisfiable in $\widehat {X}$ is satisfied by a tuple of vertices. Every definable subset $B\in D_n(Y)$ consists of a set of $(n+1)$ -tuples of vertices. For $C\in K(Y)_n$ , let $\lambda _n(C)$ be $x\in X_n$ if, for each $B\in C\cap D_n(Y)$ and for each $(e_0, e_1, \ldots , e_n)\in B$ , the barycenter of the convex hull of $\{e_0, e_1, \ldots , e_n\}$ belongs to $||x||$ , and x is the least n-simplex of X with that property; else some fixed n-simplex $x_{\star }$ of X. Since both the operations of taking the convex hull and the barycenter commute with the face operators, the maps $\lambda _n : K(Y)_n\rightarrow X_n$ ( $n\in \mathbb {N}$ ) make up a simplicial map $\lambda : Y^*\rightarrow X$ , which is easily seen to be a trivial fibration. The map $\varepsilon : (\widehat {X})^*\rightarrow X$ is then defined as $\lambda \circ r^*$ . Since trivial fibrations are stable under composition, $\varepsilon $ is a trivial fibration too. ⊣

Let X be a fibrant simplicial set and $x\in X_0$ . The set $\pi _1(X, x)$ is defined as the set of all homotopy equivalence classes of maps $\alpha : \Delta ^1\rightarrow X$ satisfying $d_0\alpha = d_1\alpha = x$ , where homotopy equivalence between $\alpha : \Delta ^1\rightarrow X$ and $\beta : \Delta ^1\rightarrow X$ consists in the existence of $\gamma : \Delta ^2\rightarrow X$ such that $d_0\gamma = s_0d_0\alpha = s_0d_0\beta $ , $d_1\gamma = \alpha $ and $d_2\gamma = \beta $ . The set $\pi _1(X, x)$ can be endowed with a canonical composition law which turns it into a group, called the first homotopy group of X at x.

Lemma 3.18. Let M be an L-structure and $C\in K(M)_0$ . Then $\pi _1(M^*, C)$ is trivial only if C is principal.

Proof Let $C\in K(M)_0$ be nonprincipal. Without loss of generality one may assume that $C = \{(\phi _1(v_0))^M, (\phi _2(v_0))^M\}^c$ with $(\phi _1(v_0))^M\neq \emptyset $ , $(\phi _2(v_0))^M\neq \emptyset $ and $(\phi _1(v_0))^M\neq (\phi _2(v_0))^M$ . Let $\alpha := \{E_0((\phi _1(v_0))^M), E_1((\phi _2(v_0))^M)\}^c$ and $\beta := \{(\phi _1(v_0)\wedge \phi _2(v_1))^M, (\phi _1(v_1)\wedge \phi _2(v_0))^M\}^c$ . One has: $\exists _0\alpha = \exists _0\beta = \exists _1\alpha = \exists _1\beta = C$ . Let us now suppose that there exists $\gamma \in K(M)_2$ such that $\exists _1\gamma = \alpha $ and $\exists _2\gamma = \beta $ : then $\{(\phi _1(v_0))^M, (\phi _2(v_0))^M\}^c = \{(\phi _1(v_0)\wedge \exists x\phi _2(x))^M, (\exists x\phi _1(x)\wedge \phi _1(v_0))^M\}^c = \exists _1\beta = \exists _1\exists _2\gamma = \exists _1\exists _1\gamma = \exists _1\alpha = \{(\phi _1(v_0))^M\}^c$ , in contradiction with the hypothesis about $(\phi _1(v_0))^M$ and $(\phi _2(v_0))^M$ . ⊣

Theorem 3.19. The functors $F'$ and $G'$ are part of an adjunction

up to homotopy.

Proof The more specific statement of the theorem is that, for any objects M of ${\mathbf{oL}\hbox{-}\mathbf{Strs}}$ and X of ${\mathbf{dilo}\hbox{-}\mathbf{sSets}}$ , there is a map $\Phi _{M, X} : \text {Hom}_{{\mathbf {\scriptsize oL-Strs}}}(\widehat {X}, M)\rightarrow \text {Hom}_{{\mathbf {\scriptsize dilo-sSets}}}(M^*, X)$ with a homotopy inverse $\Psi _{M, X}$ satisfying not only $\Phi \circ \Psi \simeq \text {id}_{\textrm {\scriptsize Hom}(M^*, X)}$ and $\Psi \circ \Phi \simeq \text {id}_{\textrm {\scriptsize Hom}(\widehat {X}, M)}$ , but even $\Psi \circ \Phi = \text {id}_{\textrm {\scriptsize Hom}(\widehat {X}, M)}$ . Let $f : \widehat {X}\rightarrow M$ be a given elementary embedding in oL-Strs. Then $p_f : M^*\rightarrow X$ is defined as $\varepsilon \circ f^*$ , where $\varepsilon $ is the trivial fibration of Lemma 3.17. Since $f^*$ is a trivial fibration, so is $p_f$ too. And since $\varepsilon $ clearly preserves distinguished simplices and respects the ordering of n-simplices for each $n\in \mathbb {N}$ , as well as $f^*$ , so does $p_f$ too. Also, $p_f$ is a retraction with a unique section $f_!\circ r_!\circ s$ , where

is the retraction of the previous lemma and $s_n : X_n\rightarrow K(Y)_n$ sends each $x\in X_n$ to $\{(||x_0||, \ldots , ||x_n||)\}^c$ , $x_0, \ldots , x_n$ being the vertices corresponding to x in Lemma 3.17. Conversely, any trivial fibration $p : M^*\rightarrow X$ in dilo-sSets gives rise, for each $x\in |\widehat {X}|$ , to a lifting diagram:

where $\ast $ is (the simplicial set whose realization is) the point and $\{x\}^c$ represents $\{(v_0 = \overline {x})^{\widehat {X}}\}^c\in K(\widehat {X})_0$ . Indeed, since $p_0$ is surjective, there exists $C_x\in K(M)_0$ such that $p_0(C_x) = \varepsilon _0(\{x\}^c)$ . One has: $h_0(\{x\}^c) = C_x = i_p\circ \varepsilon \circ \{x\}^c$ . Because p and $\varepsilon $ are trivial fibrations, and thus weak equivalences, h is a weak equivalence too, i.e., a map between fibrant simplicial sets which induces isomorphisms on all homotopy groups.Footnote 16 In particular, $h_1 : \pi _1((\widehat {X})^*, \{x\}^c)\simeq \pi _1(M^*, C_x)$ . As a consequence, by Lemma 3.18, $C_x = \{(\phi _x(v_0))^M\}^c$ for some L-formula $\phi _x$ . For $a\in (\phi _x(v_0))^M$ , one has that $(M, a)\vDash \text {Th}(\widehat {X}, \overline {x})$ , a theory which (owing to Pillay and Steinhorn’s result) has a unique prime model $(P, \alpha )$ . Let $e_M : (P, \alpha )\rightarrow (M, a)$ and $e_{\widehat {X}} : (P, \alpha )\rightarrow (\widehat {X}, x)$ two corresponding elementary embeddings. Even though $e_M$ and $e_{\widehat {X}}$ are not unique themselves, $e_M(\alpha ) = a$ and $e_{\widehat {X}}(\alpha ) = x$ are ensured to hold. So a (and thus $C_x$ ) is uniquely determined. Stating $f_p(x) := a$ , one defines $f_p : \widehat {X}\rightarrow M$ . This is an elementary embedding: indeed, $f_p(|\widehat {X}|)\subseteq |M|$ and types of elements are preserved, since the definition of each is preserved by the corresponding h, which can taken to be $i_p\circ \epsilon $ for each $x \in|\widehat{X}|.$

The functoriality of $p\mapsto f_p$ and $f\mapsto p_f$ is clear, so it only remains to check that $p_{f_p}\simeq p$ and that $f_{p_f} = f$ . Let $p : M^*\rightarrow X$ be a given trivial fibration. Then $f_p : \widehat {X}\rightarrow M$ is defined by $\{f_p(x)\}^c = (i_p\circ \epsilon )_0(\{x\}^c)$ , and $p_{f_p}$ as the composite $\varepsilon \circ (f_p)^*$ . By construction of $f_p$ , $h = (f_p)_!$ , so $p_{f_p} = \varepsilon \circ (f_p)^* = p\circ (f_p)_!\circ (f_p)^*$ . And since $(f_p)_!\circ (f_p)^*\simeq \text {id}_{M^*}$ , one gets: $p_{f_p}\simeq p$ . Actually, as $(p_{f_p})_0(\{(v_0 = \overline {f_p(x)})^M\}^c) = (\varepsilon _0\circ {{\hspace {-0.0001pt}}}^0(f_p))(\{(v_0 = \overline {f_p(x)})^M\}^c) = \varepsilon _0(\{(v_0 = \overline {x})^{\widehat {X}}\}^c) = (p_0\circ h_0)(\{(v_0 = \overline {x})^{\widehat {X}}\}^c) = p_0(\{(v_0 = \overline {f_p(x)})^M\}^c)$ , p and $p_{f_p}$ coincide on $h_0(K(\widehat {X})_0)$ ( $\widehat {X}$ and M being o-minimal). The resulting stronger result is written: $p_{f_p}\simeq p\ (\text {rel}\ h_0(K(\widehat {X})_0))$ . On the other hand, let $f : \widehat {X}\rightarrow M$ be a given elementary embedding, and $x\in ||X||$ . By definition, $f_{p_f}(x)$ is the element b of $|M|$ such that $h_0(\{x\}^c) = \{b\}^c$ , where h is the lift of $\varepsilon $ along $p_f$ . Besides, $a := f(x)$ is such that ${{\hspace {-0.0001pt}}}^0f(\{a\}^c) = \{x\}^c$ , so $(p_f)_0(\{a\}^c) = \varepsilon _0(\{x\}^c) = (p_f)_0(h_0(\{x\}^c)) = (p_f)_0(\{b\}^c)$ . As a consequence, ${{\hspace {-0.0001pt}}}^0f(\{a\}^c) = {{\hspace {-0.0001pt}}}^0f(\{f(x)\}^c) = {{\hspace {-0.0001pt}}}^0f(\{b\}^c)$ . Hence $b = a$ , and this holds for each $x\in ||X||$ , so $f_{p_f} = f$ .⊣

Theorem 3.19 provides a systematic tool to transpose homotopy theoretic notions to model theory.Footnote 17 In particular, a natural avenue to pursue consists in harnessing the internal homotopy theory carried by the category of simplicial sets. Indeed, since any model of a given theory T is associated with a simplicial set in a functorial way, the category of models of T can be conceived of as a subcategory of the category of simplicial sets. Since the latter is a fundamental example of model category, a natural question is: are there theories whose categories of models are model categories? Another, related question is: is it possible to endow ${\mathbf{oL}\hbox{-}\mathbf{Strs}}$ with a model category structure, so that the adjunction up to homotopy between $F'$ and $G'$ becomes a Quillen adjunction?Footnote 18 Regardless of those open questions, the simplicial viewpoint embraced in this paper leads naturally, to conclude, to looking at first-order theories as simplicial presheaves.

Proposition 3.20. To any L-theory T corresponds a simplicial presheaf $\mathcal {F}_T : {\mathbf{L}\hbox{-}\mathbf{Strs}}^{{\mathbf {\scriptsize op}}}\rightarrow {\mathbf { sSets}}$ on ${\mathbf{L}\hbox{-}\mathbf{Strs}}$ .

Proof For each $\phi \in F_n$ and each L-structure M, let $[\phi ^M]_T = \{\psi ^M : T\vdash \phi \leftrightarrow \psi \}$ . It is straightforward to check that $[\phi ^M]_T\mapsto [(\exists _i\phi )^M]_T$ and $[\phi ^M]_T\mapsto [E_j(\phi ^M)]_T$ are well defined and satisfy the simplicial identities, so that ( $\mathcal{F}_T(M))_n := \{[\phi^M]_T : \phi \in F_n\}\ (n\in\mathbb{N}$ ) defines a simplicial set $\mathcal{F}_T(M)$ . In particular, $\mathcal {F}_T(M) = M_*$ iff $M\vDash T$ . For any elementary embedding $f : M \rightarrow N$ , the definition of $\mathcal{F}_T(f)$ is directly adapted from that of $f_*$ . ⊣

Footnotes

1 [5].

2 The comparison of spaces of types to derived spaces of a topological space has been put forward as early as in [Reference Morley8] (pp. 514–515). Besides, connections between propositional calculus and homotopy types have been explored in [Reference Conant and Thistlethwaite2]. Still, no systematic use of concepts anchored in modern algebraic topology has been made in logic, at the partial exception of [Reference Knight7] and [Reference Goodrick, Byunghan and Kolesnikov4].

3 For example, the two logically equivalent L-formulae $\chi (v_0) := (P(v_0)\vee \neg P(v_0))$ and $\pi (v_0, v_1) := ((P(v_0)\vee \neg P(v_0))\wedge (P(v_1)\vee \neg P(v_1)))$ give rise to L-formulae $\partial \chi $ and $\partial \pi $ which are not logically equivalent. But $\chi \in F_0$ , whereas $\pi \in F_1$ , so that $\chi $ and $\pi $ cannot be said to be logically equivalent in the sense meant here.

4 This is the “Dold–Kan correspondence.” See [Reference Weibel12], pp. 265–266 and pp. 270–271.

5 Two simplicial sets X and Y are said to be homotopy equivalent if there are simplicial maps $f : X\rightarrow Y$ and $g : Y\rightarrow X$ which are homotopy inverse to each other, in the sense defined in Section 3.2.

6 See [Reference Goerss and Jardine3] (pp. 25–27) and [Reference Hovey6] (Lemma 3.4.5, p. 85) for the definition of simplicial homotopy groups and the proof of their being groups.

7 See [Reference Goerss and Jardine3], Lemma I. 3. 4, p. 12.

8 See [Reference Goerss and Jardine3], Lemma II.1.1, pp. 67–68.

9 The usual notations for the geometric realization of X are ‘ $|\Delta ^n|$ ’ and ‘ $|X|$ ’. The notations ‘ $||\Delta ^n||$ ’ and ‘ $||X||$ ’ chosen in this paper are motivated by the wish not to confuse the domain of an L-structure and the geometric realization of a simplicial set—precisely owing to the connection to be established between both notions (see Definition 3.8 below).

10 See [Reference Hovey6], pp. 49–58. A cellular inclusion is a particular case of map in $I'$ -cell (Lemma 2.4.5), and $I'\text {-cell}\subseteq I'\text {-cof}$ (Lemma 2.1.10), so any cellular inclusion is a cofibration in the category ${\mathbf { Top}}$ of topological spaces, as defined in Definition 2.4.3. But, for every trivial fibration p in ${\mathbf { sSets}}$ , $||p||$ is a trivial Serre fibration in ${\mathbf { Top}}$ (Corollary 2.4.20), so it has the right lifting property w.r.t. all maps in J, and thus w.r.t. all cofibrations, since (J-cof)-inj = J-inj.

11 The introduction of that functor would be motivated, in the context of focusing on o-minimal structures over the real line, by the natural comparison of a definable subset of $X_{\mathbb {R}}$ with a simplicial subcomplex of $||X||$ (see [Reference van den Dries11], ch. 8).

12 See [Reference Pillay and Steinhorn9], Theorem 5.1, p. 583.

13 About $X \times \Delta^1$ : the Cartesian product $X \times Y$ of two simplicial sets X and Y is simply defined by: $(X \times Y)_n := X_n \times Y_n\ (n\in\mathbb{N})$ .

14 See the remark following Lemma 3.3.

15 The simplicial set $\partial\Delta^1$ , called the boundary of $\Delta^1$ , is the smallest simplicial subset of $\Delta^1$ containing all the 0-simplices (vertices) of the latter, namely 0 and 1.

16 See [Reference Goerss and Jardine3] (p. 32 and p. 39, respectively) for the definition of a weak equivalence and the proof of the “two-out-of-three” property of weak equivalences. See [Reference Goerss and Jardine3] again (pp. 42–45) for the proof that a trivial fibration, as the notion has been defined in this paper, is a weak equivalence.

17 As an example, the category ${\mathbf{oL}\hbox{-}\mathbf{Strs}}$ can be shown to admit of a simplicial object up to homotopy. Indeed, the proof of [Reference Hovey6], Proposition 3.1.5 (see also Remark 3.1.7) can be transposed to the adjunction up to homotopy of Theorem 3.19.

18 See [Reference Goerss and Jardine3] and [Reference Hovey6] for the definitions of a model category and of a Quillen adjunction.

References

Baisalov, Y. and Poizat, B., Paires de structures o-minimales, this Journal, vol. 63 (1998), no. 2, pp. 570578.Google Scholar
Conant, J. and Thistlethwaite, O., Boolean formulae, hypergraphs and combinatorial topology. Topology and Its Applications, vol. 157 (2010), pp. 24492461.10.1016/j.topol.2010.08.016CrossRefGoogle Scholar
Goerss, P. G. and Jardine, J. F., Simplicial Homotopy Theory, Birkhäuser Verlag, Basel, Switzerland, 1999.10.1007/978-3-0348-8707-6CrossRefGoogle Scholar
Goodrick, J., Byunghan, K., and Kolesnikov, A., Homology groups of types in model theory and the computation of ${H}_2(p)$ , this Journal, vol. 78 (2013), no. 4, pp. 10861114.Google Scholar
Hofmann, M. and Streicher, T., The groupoid interpretation of type theory, Twenty Five Years of Constructive Type Theory (Sambin, G. and Smith, J. M., editors), Oxford University Press, New York, NY, 1998, pp. 83111.Google Scholar
Hovey, M., Model Categories, American Mathematical Society, Providence, 1999.Google Scholar
Knight, R. W., Categories of topological spaces and scattered theories. Notre Dame Journal of Formal Logic, vol. 48 (2007), no. 1, pp. 5377.10.1305/ndjfl/1172787545CrossRefGoogle Scholar
Morley, M., Categoricity in power. Transactions of the American Mathematical Society, vol. 114 (1965), no. 2, pp. 514538.10.1090/S0002-9947-1965-0175782-0CrossRefGoogle Scholar
Pillay, A. and Steinhorn, C., Definable sets in ordered structures I. Transactions of the American Mathematical Society, vol. 295 (1986), no. 2, pp. 565592.CrossRefGoogle Scholar
Poizat, B., A Course in Model Theory: An Introduction to Contemporary Mathematical Logic, Springer, New York, NY, 2000.10.1007/978-1-4419-8622-1CrossRefGoogle Scholar
van den Dries, L., Tame Topology and O-minimal Structures, Cambridge University Press, Cambridge, MA, 1998.10.1017/CBO9780511525919CrossRefGoogle Scholar
Weibel, C., An Introduction to Homological Algebra, Cambridge University Press, Cambridge, MA, 1995.Google Scholar