Home Mathematics Models, coproducts and exchangeability: Notes on states on Baire functions
Article Open Access

Models, coproducts and exchangeability: Notes on states on Baire functions

  • Serafina Lapenta EMAIL logo and Giacomo Lenzi
Published/Copyright: August 10, 2022
Become an author with De Gruyter Brill

Abstract

We discuss exchangeability and independence in the setting of σ-complete Riesz MV-algebras. We define and link to each other the notions of exchangeability and distribution law for a sequence of observables (i.e. non classical random variables), as well as the notion of independence for a sequence of algebras. We obtain two categorical dualities for σ-complete Riesz MV-algebras endowed with states and we define a “canonical” state on the coproduct of a sequence of probability Riesz tribes, giving a weak version of de Finetti’s result. Finally, we discuss statistical models.

1 Introduction

The idea of using logic to reason about probability is not new, one can find several approaches in literature, that vary from first order logics in which the atomic formulas are meant to be comparison of probabilities (see for instance [10]) to modal approaches in which one uses a modality P to say that a formula of some logic is probable (see for instance [11] for the case for fuzzy probability logic).

In the setting of Łukasiewicz logic, one of the most studied fuzzy logics, probabilities are very efficiently considered via the notion of a state, which is an additive and truth-preserving operator. This notion was introduced by D. Mundici in [19] with the idea of defining a “truth averaging” operator for logical formulas.

This intuition was made more precise subsequently, when T. Kroupa and G. Panti proved independently that a state on any MV algebra A is, in fact, a Borel regular probability measure on the space Max(A) of the maximal ideals of A, see Section 2 for a formal statement. If one considers an MV-subalgebra A of some algebra of continuous functions C(X) over a compact and Hausdorff space X, then to each state s on A it corresponds exactly one regular measure on the σ-algebra of Borel subsets of X. Furthermore, s is given by integration with respect to such measure. In this sense, states are averaging operators: they act as expectations on the functions of A, see also [12: Remark 2.8].

With the additional momentum gained from the Kroupa-Panti integral representation, states were deeply investigated. For instance, they are used in [11] to give an algebraic semantics for their logical system and they are characterized as terms in the language of special types of MV-algebras in [12: Section 5]. Nevertheless, the second crucial result for the development of the theory of states on MV-algebras was the generalization of de Finetti’s Dutch book theorem, see [20: Chapter 1]. This result, proved by D. Mundici, has among its consequences the fact that states on MV-algebras are Bayesian probabilities.

Thus, probability on MV-algebras can be considered from a subjective point of view and one pivotal theorem in probability theory is again a result by B. de Finetti, the so-called de Finetti’s theorem on exchangeable probabilities. This theorem, loosely speaking, states that a probability on a cartesian product does not depend on the order of factors if, and only if, it can be build using independent and identically distributed probability measures, see Section 2 for a more precise statement. One crucial feature of this result is that it shows how a statistical model can appear in a subjectivist framework, and it is a bridge between the frequentist and the Bayesian approaches to probability theory. Whence, having at our disposal a subjective notion of probability on MV-algebras, it is only natural trying to understand how de Finetti’s exchangeability can be analyzed from the point of view of logic. To do this, a more expressive language than that given by MV-algebras is needed.

Indeed, despite the theory of states is well-developed for MV-algebras, if one wants to tackle a theorem such as de Finetti’s one needs a language that allows to write (at least) convex combinations, as de Finetti’s theorem deals with mixtures of independent and identically distributed random variables, which are convex combinations when the variables are finitely many. With this idea in mind, we choose to focus our attention on the class of Riesz MV-algebras, that stand to MV-algebras like real vector spaces stand to groups. More precisely, we will work in the class of those Riesz MV-algebras that are closed under countable suprema and infima, which was proven to be an infinitary variety of algebras in [8].

The reason for this choice lies in the fact that in [6] the authors used σ-complete Riesz MV-algebras to start an investigation on the most suitable way of interpreting random variables in logical terms. They defined observables as the algebraic counterpart of random variables in a way that, in certain cases, provides a one-one correspondence between classical random variables and observables, adding the expressive power that allows to model more complicated phenomena, while being able to use the powerful tools of measure theory inside a logical system. The variety of σ-complete Riesz MV-algebras has been further investigated in [7], where the authors characterize all free objects and give two categorical dualities.

In this paper we build on [68] and we propose and discuss the main notions that one needs in order to start thinking about de Finetti’s result on exchangeable probabilities. In other words, we prepare the ground for a non-classical version of this result. We start with giving all needed preliminary notions in Section 2, while in Section 3 we discuss exchangeability and independence for sequences of observables as well as for states. We first generalize two key results of [6] and then we define and link to each other the notions of exchangeability and distribution law for a sequence of observables, as well as the notion of independence for a sequence of tribes. Section 4 aims at defining a suitable algebraic counterpart for the notion of a product measure. Motivated by the dualities of [7], we define this notion using coproducts of tribes: Proposition 4.1, Definition 4.9 and Theorem 4.13 are the main results of the section, culminating in a weak version of de Finetti’s exchangeability. In Section 5 we enrich the objects of the categories considered in [7] with states and probability measures and, upon suitable restrictions on morphisms, we obtain two new enriched categorical dualities. Finally, in Section 6 we define statistical models in logical terms and we give some examples.

2 Preliminaries

2.1 σ-complete Riesz MV-algebras

In this paper we will work with σ-complete Riesz MV-algebras.

A Riesz MV-algebra is an algebra (A, ⊕, ¬, 0, 1,{α}α∈[0, 1]), where ⊕ is a binary operation, ¬ is an involution, 0 and 1 are respectively a bottom and a top element, and the unary operations {α}α∈[0, 1] model a scalar multiplication. The standard example of such an algebra is the real interval [0, 1], where xy = min(x + y, 1), ¬ x = 1 − x and αx is the product of real numbers. This example is standard in a very precise sense: Riesz MV-algebras form a variety and [0, 1] is a generator for it. From a different point of view, if (V, u) is a Riesz space (vector lattice) with a distinguished strong order unit u, the interval [0, u]V = {xV| 0 ≤ xu} is a Riesz MV-algebra when endowed with the following operations: xy = (x + Vy)∧u, ¬x = uVx, αx the same as in V. The map that takes (V, u) and sends it into [0, u]V is actually a functor, denoted by Γ, that gives a more general categorical equivalence between Riesz MV-algebras and Riesz spaces with strong unit, see [5, 9, 18] for all missing definitions and references.

A Riesz MV-algebra is σ-complete if it is closed under countable suprema (and therefore also infima). It was proved in [8] that σ-complete Riesz MV-algebras form an infinitary variety in the sense of [22], and the algebra [0, 1] is again a generator for the infinitary variety, subsequently denoted by RMVσ. Furthermore, in [7], the authors prove that for any cardinal κ, the κ-generated free algebra in RMVσ is the MV-algebra all [0, 1]-valued and Baire-measurable functions defined on [0, 1]κ.

To be more precise, given a topological space (T, τ), where T is the universe of the topology and τ denotes the open sets of the topology, C(T) will denote the set of [0, 1]-valued continuous functions defined over T. A zeroset ZT is a set for which there exists fC(T) such that Z = {xT | f(x) = 0}. A cozero set is a complement of a zeroset, that is a set definable as {xT | f(x) ≠ 0} for some continuous function f. A Baire set is a subset of T belonging to the σ-algebra generated by the zerosets of continuous functions from T to ℝ, while a Borel set is a subset of T belonging to the σ-algebra generated by the closed sets. We shall denote the σ-algebras of Baire and Borel subsets of T respectively by 𝓑𝓐(T) and 𝓑𝓞(T). Whence, a Baire function is a function f : T → [0, 1] measurable with respect to the spaces (T, 𝓑𝓐(T)), ([0, 1], 𝓑𝓐([0, 1])). Borel functions are analogously defined. Note that every Baire set is a Borel set, that is 𝓑𝓐(T) ⊆ 𝓑𝓞(T), while the converse inclusion holds for a metrizable space, see [7: Remark 2.1]. Thus, considering the cases of interest for us, 𝓑𝓞([0, 1]X) = 𝓑𝓐([0, 1]X) for |X| ≤ ω.

It is known that the sets of [0, 1]-valued Baire and Borel functions defined over [0, 1]X are σ-complete Riesz MV-algebras, and they shall be denoted respectively by Baire([0, 1]X) and Borel([0, 1]X). Note that Baire([0, 1]X) = Borel([0, 1]X) for |X| ≤ ω, while they do not coincide for |X| > ω. We also recall that in both algebras countable suprema are taken pointwise, see [7,8].

To give more uniform notations, in this work we follow [7] and therefore IRL(X) will denote the free X-generated algebra, where X is an arbitrary set. The elements of IRL(X) will be called IRL-polynomials. Consequently, IRL(X) = Borel([0, 1]X) = Baire([0, 1]X) for |X| ≤ ω, while IRL(X) = Baire([0, 1]X) for |X| > ω.

Finally, for any ARMVσ, we shall call σ-ideals the ideals of A that are closed under countable suprema, while we will call MV-maximal σ-ideals those σ-ideals of A that are also maximal ideals for the Riesz MV-reduct of A. The set of all MV-maximal σ-ideals will be denoted by 𝓜σ(A), or 𝓜σ when A is clear from the context. In [7], the authors defined σ-semisimple algebras as those algebras A such that ∩{M | M𝓜σ} = {0}. Furthermore, upon defining IRL-algebraic varieties as intersections of Baire sets, in [7] it is proved the existence of a duality between IRL-varieties and σ-semisimple σ-complete Riesz MV-algebras. The duality, on the level of objects, is induced by the following operators. For any subset S ⊆ [0, 1]X,

I ( S ) = { p I R L ( X ) p ( x ) = 0  for any  x S } .

Given a set J of IRL-polynomials in IRL(X),

V ( J ) = { x [ 0 , 1 ] X p ( x ) = 0  for any  p J } = p J V ( { p } ) .

Formally, the dual categories are the following.

  1. The category ssRMVσ whose objects are presented σ-semisimple algebras in RMVσ and whose arrows are σ-homomorphisms of Riesz MV-algebras. More precisely, an object is a pair (IRL(X),I), where I is a σ-ideal in the free algebra IRL(X) such that I = 𝕀(𝕍(I)). Consequently, each morphism h: (IRL(X),I) → (IRL(Y), J) between pairs is induced by a unique homomorphism hp: IRL(X) → IRL(Y) such that hp(I) ⊆ J.

  2. The category IRL, whose objects are IRL-algebraic varieties in hypercubes of type [0, 1]X, for an arbitrary X, and arrows are tuples of IRL-polynomials, that is, an arrow in IRL is a map η = (ηy|S)yY :S ⊆ [0, 1]XT ⊆ [0, 1]Y, where each ηy belongs to IRL(X).

We also note that, as proved in [7: Theorem 4.12], a σ-complete Riesz MV-algebra A = IRL(X)/I is σ-semisimple if, and only if, AISP([0, 1]). We remark that the universal algebraic operators I, S and P are considered in the infinitary variety of σ-complete Riesz MV-algebras.

2.2 Tribes and functional representation

For any non-empty set X, a Riesz tribe (or simply tribe) is a Riesz MV-subalgebra of [0, 1]X closed under pointwise countable suprema, that is, an element of SP([0, 1]). For the purpose of this paper, we stress the fact that IRL(X) is a Riesz tribe for any X, and any σ-complete Riesz MV-algebra is a σ-homomorphic image of a Riesz tribe. This is the so-called Loomis–Sikorski Theorem, see [8: Theorem 5.2]. If 𝓣 ⊆[0, 1]X is a Riesz tribe, the set 𝓢(𝓣) = {AX|χA𝓣} is a natural σ-algebra of subsets of X, and it is the smallest σ-algebra of X that makes all functions in 𝓣 measurable. We will often refer to 𝓢(𝓣) as the Boolean center of 𝓣. Notice that we will always consider the codomain [0, 1] of each function in 𝓣 endowed with its standard Baire σ-algebra 𝓑𝓞([0, 1]) = 𝓑𝓐([0, 1]), as defined before.

We say that a topological space X ≠ ∅ is basically disconnected provided the closure of every open Fσ subset (i.e. a countable union of closed subsets) of X is open. By a well-known result of Nakano (see [20: Proposition 11.10] for the case of MV-algebras) if X is a compact Hausdorff space, the algebra C(X) (of [0, 1]-valued continuous functions on X) is σ-complete if, and only if, X is basically disconnected. We stress the fact that countable suprema on continuous functions need not to be computed pointwise.

Moreover, by the fundamental Stone-Krein-Kakutani-Yosida Theorem (see [8] for the case of Riesz MV-algebras) any σ-complete Riesz MV-algebra A is isomorphic to the algebra C(X), where X = Max(A) is a basically disconnected compact Hausdorff space. This result is made into a duality in [7], where morphisms on topological spaces are defined as follows: A function f : XY between compact Hausdorff and basically disconnected spaces is cozero-closed if for every countable union U of clopens of Y, we have f1(U¯)=f1(U)¯. Then, we have the following.

Proposition 2.1

The algebraic category RMV σ is dual to the category BDKH whose objects are basically disconnected, compact, Hausdorff spaces and whose morphisms are continuous and cozero-closed functions.

2.3 States and subjective probability

A theory of subjective probability has been developed on MV-algebras using the notion of a state, introduced by D. Mundici. To define this notion properly, we recall that any MV-algebra A can be endowed with a partial operation, subsequently denoted by +, defined when xy = 0 as x + y := xy.

The definition of a state is easily generalized to Riesz MV-algebras without additional requirements, see [9]. Thus, a state of a Riesz MV-algebra A is a map s: A → [0, 1] satisfying the following conditions:

  1. s(1) = 1,

  2. for all x, yA such that xy = 0, s(xy) = s(x) + s(y).

We shall call a state that preserves countable suprema of increasing sequences a σ-state. States on Riesz MV-algebras and σ-states on tribes are in bijective correspondence with measures on suitable measurable spaces. In the case of tribes with σ-states, the correspondence was proved by Butnariu and Klement and, subsequently, by Barbieri and Weber with different techniques [1, 3]. In the case of arbitrary MV-algebras and states, the result was proved independently by Kroupa and Panti [13: Theorem 4.0.1]. We recall both versions of the result in the way most suited to our framework, as we shall use both of them in Section 5.

Theorem 2.2

Let X be a compact and Hausdorff topological space. For any state s X : C(X) → [0, 1] there exists a unique Borel regular measure μX : 𝓑𝓞(X) → [0, 1] such that

s X ( g ) = X g d μ X ,

and the map s X μXis a homeomorphism.

Theorem 2.3

For every Riesz tribe 𝓣 ⊆ [0, 1]X, for every σ-state s of 𝓣 there exists a unique measure μssuch that, for every f𝓣,

s ( f ) = X f d μ s .

The measure μ s : 𝓢(𝓣) → [0, 1] is given by μs(A) = s(χA). The correspondence is a bijection.

We remark that the correspondence of Theorem 2.3 is actually an isometry. Indeed, in [1: Theorem 3.2.1] it is proved that the set of states on a tribe 𝓣 and the set of measures on the corresponding 𝓢(𝓣) can both be endowed with a metric structure (making them Banach spaces) in a way that the assignment above defined becomes an isometry.

Unless otherwise specified by a probability Riesz tribe is meant a pair (𝓣, s) where 𝓣 is a Riesz tribe and s is a σ-additive state.

Moreover, we will call additive a function that satisfies item (2) of the definition of a state. Similarly, for any k ∈ ℕ and MV-algebras A, A1,…,Ak, a function β : A1 × … ×AkA defined on the cartesian product will be called k-additive if it is additive in each component.

Finally, for sake of completeness we discuss below de Finetti’s theorem on exchangeability in the version of Hewitt and Savage, see [15].

Take any set X with a σ-algebra of subsets χ and let χ̅ be the smallest σ-algebra on Xω that contains all sets of type C(Ei1,,Eik)=nEn with En = X for n ∉ {i1,…,ik}. A generic measure σ on χ̅ is called exchangeable if for any permutation π of {i1,…,ik},

σ C ( E i 1 , , E i k ) = σ C ( E π ( i 1 ) , , E π ( i k ) ) .

For any measure μ on χ let μ̅ denote the unique product measure defined on χ̅ using μ. A generic measure σ on χ̅ is called presentable if there exists a measure ν on the set 𝓟 of all probability measure on χ such that for any Aχ̅,

σ ( A ) = P μ ¯ ( A ) d ν ( μ ) .

Loosely speaking, a measure is presentable if it is given by a probability kernel. Then, de Finetti’s theorem (in the version of Hewitt and Savage) gives sufficient conditions for the two notions to coincide. One of its versions is the following.

Theorem 2.4

[15: Theorem 7.2]). With the above definitions, the notions of presentability and exchangeability coincide when X is a compact and Hausdorff space and χ = 𝓑𝓐(X).

If one thinks in terms of random variables, exchangeability says that the joint distribution is invariant with respect to any permutation of a given sequence. That is, given a sequence {fi}iI of random variables, the sequence is exchangeable if for any k ∈ ℕ and any two sets of indexes i1,…,ik, j1,…,jk, the distribution laws Fi,k of (fi1,…,fik) and Fj,k of (fj1,…,fjk) coincide. de Finetti’s theorem in this case says that any sequence of exchangeable random variables can be obtained by first choosing a probability ν (on the set of all probabilities on the domain of the fi’s) and then asking for the fi’s to be independent and identically distributed with joint distribution ν.

3 Independence and exchangeability in non-classical processes

In [6], non-classical stochastic processes posed in a probability Riesz tribe are defined as sequences of observables {𝓧n}n∈ℕ with values in a Riesz tribe, in symbols, 𝓧n : IRL(Y) → 𝓣. The tribe 𝓣 ⊆[0, 1]X, for some set X, is endowed with the σ-algebra 𝓢(𝓣) = {AX|χA𝓣} and Y is a countable set.

In this work we shall consider processes defined over IRL(Y), for an arbitrary Y ≠ ∅, indexed in an arbitrary set of cardinality κ, and we shall usually denote them as {𝓧i}iκ. Indeed, the definition given in [6: Definition 2.1] can be easily generalized as follows.

Definition 3.1

Let A be a σ-complete Riesz MV-algebra. For any cardinal κ, a κ-dimensional observable on A is any σ-homomorphism of Riesz MV-algebras from IRL(Y) to A, with |Y| = κ. When we need to make explicit reference to the dimension of the observable we shall write IRL(κ) instead of IRL(Y).

A crucial point in [6] is Theorem 2.3, which is a representation theorem. Let us prove here the analogous result for observables defined over IRL(Y), with Y an arbitrary set. We give the whole proof for sake of completeness, but for the most part it can be deduced from [6].

Theorem 3.2

Let X, Y be nonempty sets, 𝓣 ⊆ [0, 1]Xbe a Riesz tribe and f : X → [0, 1]Ya measurable function w.r.t. 𝓢(𝓣)={AX|χA𝓣} and 𝓑𝓐([0, 1]Y). Then the function

X f I R L ( Y ) T , X f ( a ) = a f , a I R L ( Y ) .

is a κ-dimensional observable on 𝓣, where |Y| = κ.

Conversely, for any κ-dimensional observable 𝓧: IRL(Y) → 𝓣, there exists a unique f : X → [0, 1]Ysuch that 𝓧 = 𝓧f.

P r o o f. The fact that 𝓧f is a homomorphism of Riesz MV- algebras is straightforward. Thus, we have to prove that it preserves countable joins. Let {gn}n∈ℕ be a sequence in IRL(Y). We recall that countable suprema and infima are defined pointwise in IRL(Y) and in any tribe, therefore

X f ( n g n ) ( x ) = ( n g n ) ( f ( x ) ) = sup n { g n ( f ( x ) ) } = sup n { ( g n f ) ( x ) } = sup n { ( X f ( g n ) ) ( x ) } = ( n X f ( g n ) ) ( x ) .

Finally, let us prove that for any aIRL(Y), a o f𝓣. By [20] Lemma 11.8(ii)], 𝓣 is the tribe of all 𝓢(𝓣)-measurable functions, whence it is enough to prove that a o f is 𝓢(𝓣)- measurable for any aIRL(Y). This fact is easily deduced: For any E𝓑𝓐([0, 1]), (a o f)−1(E) = f−1(a−1(E)) ∈ 𝓢(𝓣) since f is measurable and aIRL(Y).

Conversely, given an observable 𝓧, we need to prove that there exists a function f that satisfies the claim. Let κ = |Y| and let f be the function defined by f = (fi)iκ : X → [0, 1]Y with fi = 𝓧(πi) ∈ 𝓣. Using the fact that the projections form a generating set for IRL(Y), it is straightforward that 𝓧 = 𝓧f and 𝓧f(IRL(Y)) ⊆ 𝓣. Thus, we only need to prove that f is measurable w.r.t. 𝓢(𝓣). To do so, take E𝓑𝓐([0, 1]Y) and let us prove that χf−1(E)𝓣 : since χf−1(E)(x) = 1 if, and only if, f(x) ∈ E if, and only if, χE(f(x)) = 1, it follows that χf−1(E) = 𝓧f(χE) = 𝓧(χE), which belongs to 𝓣 by definition of 𝓧.

Finally, let g : X → [0, 1]Y be another function that satisfies the claim. Then, πi o g = 𝓧(πi) = πi o f, from which we deduce that for any xX, g(x) = (πi(g(x))iκ = (πi(f(x)))iκ = f(x). ◻

Note that we have endowed [0, 1]Y with the σ-algebra 𝓑𝓐([0, 1]Y). When |Y| ≤ ω, 𝓑𝓐([0, 1]Y) = 𝓑𝓞([0, 1]Y) and we recover the countable case of [6].

When 𝓣 carries a σ-state s, to the pair (𝓣, s) is naturally associated the probability space (X, 𝓢(𝓣), μs) as defined in Section 2. Thus, to each observable 𝓧 it is associated the measurable function f : X → [0, 1]Y given by Theorem 3.2, which is a random variable, and the correspondence is one-one. See also the discussion after [6: Theorem 2.3].

Let us now generalize [6: Theorem 2.6] to the case of any arbitrary Y. The observables that arise from the following result will be subsequently called joint observables.

THEOREM 3.3

(Joint observable theorem). Given the Riesz tribe 𝓣 ⊆ [0, 1]Xand κ one- dimensional observables over 𝓣, namely 𝓧i: IRL(Xi) → 𝓣, with iκ and Xisingletons, forX=iXithe joint function 𝓙κ: IRL(X) → 𝓣 defined by 𝓙κ(a) = a o f, with f = (𝓧i(id))iκ, is a κ-dimensional observable over 𝓣, where id: [0, 1]Xi → [0, 1]Xiis the identity function. Moreover, for any iκ and any aIRL(Xi), 𝓙κ(a o πi) = 𝓧i(a).

P r o o f. Recalling that IRL(X) is the free X-generated algebra in RMVσ, it follows that the assignment πifi = 𝓧i(id) ∈ 𝓣 extends to a σ-homomorphism 𝓧: IRL(X) → 𝓣. Applying now Theorem 3.2 to such an 𝓧, we can say that 𝓧 = 𝓧g, with g = (gi)iκ given by gi = 𝓧(πi). By definition of 𝓧, gi = fi and 𝓧g is exactly the map 𝓙κ, which is, therefore, an observable.

Finally, 𝓙κ(a o πi) = (a o πi) o f = a o fi = 𝓧i(a). ◻

Following [6], we call process any indexed set {𝓧i}iI of one-dimensional observables, with I being an arbitrary set. Unless otherwise specified, for any observable 𝓧i : IRL(Y) → 𝓣 with 𝓣 ⊆ [0, 1]X, we will denote by fi the unique measurable function given in Theorem 3.2, that is, the unique function such that 𝓧i(a) = a o fi for any aIRL(Y).

As mentioned in the Introduction, the main goal of this section is to define and analyze, from the point of view of non-classical logic, the pivotal notions that are involved in de Finetti's theorem on exchangeable probabilities. To do so, let us start with analyzing basic notions of probability theory such as distributions and independence.

Definition 3.4

Let 𝓧: IRL(Y) → 𝓣 be an observable on 𝓣 ⊆ [0, 1]X, given by 𝓧(a) = a o f, and assume that 𝓣 carries the state s: 𝓣 → [0, 1]. Its distribution is the state:

s x : I R L ( Y ) [ 0 , 1 ] , s x ( a ) = s ( x ( a ) ) .

Remark 3.5

With the same notations as above, the function f : X → [0, 1]Y is a classical random variable between the spaces (X, 𝓢(𝓣), μs) and ([0, 1]Y, 𝓑𝓐([0, 1]Y).

For any E𝓑𝓐([0, 1]Y), it follows that 𝓧(χE) = χE o f = χf−1(E) and its distribution state is given by s𝓧(χE) = s(𝓧(χE)) = s(χf−1 (E)). Thus, if we consider the measure μsx associated to s𝓧 by Theorem 2.3, μs𝓧 (E) = μs(f−1(E)).

Consequently, μs𝓧 is the pushforward measure of μs under f, and F(E): = μs(f−1(E)) is the probability of the event (fE) = {xX | f(x) ∈ E}, that is, the distribution law of f.

Lemma 3.6

Let {𝓧i}iIbe an indexed set of one-dimensional observables, with I being an arbitrary set. For any natural number k, let Fkbe the (classical) joint distribution function of fi1 , … , fik. Then, for any characteristic function χE, with E𝓑𝓐([0, 1]k), we have Fk(E) = s𝓙k (χE), wheres𝓙kdenotes the distribution state of the joint observable of 𝓧i1 , … , 𝓧ik.

P r o o f. For each fij: X → [0, 1], its distribution is the function Fij: 𝓑𝓐([0, 1]) → [0, 1] that maps Eμs(fij1(E))=s(χfii1(E)). Let us denote by fk : X → [0, 1]k the function x ↦ (fi1 (x), … , fik(x)). Then 𝓙k(a) = a o fk and the (classical) joint distribution of these random variables is given by Fk: 𝓑𝓐([0, 1]k) → [0, 1], Bμs(fk1(B)). Hence, for E𝓑𝓐([0, 1]k), sJk(χE)=s(χfk1(E))=Fk(E), settling the claim. ◻

Lemma 3.7

Let (𝓣, s) be a probability Riesz tribe. Any function f𝓣 induces a one- dimensional observable 𝓧f: IRL(1) → 𝓣whose distribution state corresponds to the distribution law of f on 𝓑𝓞([0, 1])

P r o o f. It follows from the fact that any f𝓣 is 𝓢(𝓣)-measurable. ◻

Using this notion of distribution, we now define exchangeable processes. We also recall that the definition of a sequence of exchangeable random variables can be found at the end of Section 2.

Let {𝓧i}iI be an indexed set of one-dimensional observables, with I being an arbitrary set. For any finite k ∈ ℕ and two sets of indexes i1, … , ik, j1, … , jk, let us denote by Jki the joint observable of 𝓧i1 , … 𝓧ik and by Jki the joint observable of 𝓧j1 , … 𝓧jk.

Definition 3.8

The process {𝓧i}iI, posed in the probability Riesz tribe (𝓣, s), is called

  1. weakly exchangeable if the sequence {fi}iI of corresponding random variables (such that 𝓧i = 𝓧fi) is exchangeable.

  2. strongly exchangeable if for any finite k ∈ ℕ and two sets of indexes i1,,ik,j1,,jk,Jki and Jki have the same distribution state.

Thus, the process {𝓧i}iI is weakly exchangeable when, for any k ∈ ℕ and any two sets of indexes i1, … , ik, j1, … , jk, the distribution laws Fi,k of (fi1 , … , fik) and Fj,k of (fj1 , … , fjk) coincide. Strong exchangeability requires the analogous property directly on the joint observable, rather than the joint random variable associated to it. The next lemma shows that indeed strong exchangeability implies weak exchangeability.

Lemma 3.9

If {Xi}iIis a strongly exchangeable process, then it is weakly exchangeable.

proof. Recalling the definition of exchangeable random variables from Section 2, by Corollar 3.6, for any E ∈ 𝓑𝓐([0,1]k), Fi,k(E)=sJki(χE) and Fj,k(E)=sJkj(χE). By hypothesis on {Xi}iI, it follows that sJki(χE)=sJkj(χE), settling the claim.

Moving now on the notion of independence, we generalize the definition given for processes in [6: Section 5], which is itself adapted from the definition given in [17].

Definition 3.10

Let I be a set of indexes, {(𝒯i, si)}iI and (𝒯, s) be probability Riesz tribes. The collection {(𝒯i, si)}iI is said to be (𝒯, s)-independent if for any k ∈ ℕ and any i1,…,ik there exists an k-additive operator (as defined in Section 2)

β : T i 1 × × T i k T

such that for all ai1 ∈ 𝓣i1,…,aik ∈ 𝓣ik we have

s β a i 1 , , a i k = s i 1 a i 1 s i k a i k .

If all β are surjective, the collection is called surjectively independent.

In [6], the definition was given for a stochastic process {Xi}iI posed in a probability Riesz tribe (T, s). The process was called independent if any finite subset {Im(Xi1),…,Im(Xik)} of the sequence of tribes Im(Xi) is Im(ℐk)-independent, where ℐk is their joint observable.

Note that, if any of the k-additive operators β of the definition of an independent process is surjective, for any f ∈ Im(ℐk) there exists fi1 ∈ Im(Xi1),…,fik ∈ Im(Xik) such that β(fi1,…fik) = f. Consequently, for any aIRL(k) there exists ai1,…,aikIRL(1) such that Jki(a) = β(Xi1(a1),…,Xik(ak)).

Lemma 3.11

Let {Xi}iIbe a one-dimensional process posed on a Riesz tribe (𝓣, s), with 𝓣 ⊆ [0,1]X. If the process is independent and identically distributed with allβ’s being the product of functions, then it is weakly exchangeable.

proof. By[6: Proposition 5.9], the associate sequence {fi}iI is independent; by Remark 3.5 for any iI, the distribution law of fi is Fi:=μsfi1 and by hypothesis for any i,jI we have sXi = sXj. Consequently, again by Remark 3.5, for any E𝓑𝓐([0,1]), Fi(E) = μsfi1(E)=sXiχE=sXjχE=μsfj1(E)=Fj(E). Hence, the classical process {fi}iI is independent and identically distributed, and therefore exchangeable. Indeed, for any k ∈ ℕ and any two sets of indexes i1,…,ik, j1,…,jk, let Fi,k be the distribution law of fik=fi1,,fik and let Fj, k be the distribution law of fjk=fj1,,fjk. For any generator n=1kEn, of 𝓑𝓐([0, 1]k), since the fi’s are independent and identically distributed,

F i , k ( E ) = μ s f i k 1 ( E ) = μ s n = 1 k f i n 1 E n = μ s f i 1 1 E 1 μ s f i k 1 E k = μ s f j 1 1 E 1 μ s f j k 1 E k = μ s n = 1 k f j n 1 E n = μ s f j k 1 ( E ) = F j , k ( E ) .

Hence, the Xi’s are weakly exchangeable.

One open problem is to understand when the conclusion of Lemma 3.11 can be extended to strong exchangeability. Furthermore, for a countable and weakly exchangeable process, the classical version of de Finetti’s exchangeability implies that the distribution law of the associated sequence of random variables is conditionally i.i.d. (see [16: Chapter 11]). It is still unclear how and if we can give a suitable logical interpretation for this notion, in terms of observables. To tackle this issue, in the next section we define a counterpart for the notion of product measure.

4 States on coproducts in RMVσ

Taking inspiration from the definition of presentable measures from [15], we now define an appropriate counterpart of the product measure for states. Thinking of states as an algebraic dual of probability measures, we will first give a characterization for the coproduct of objects in RMVσ.

We recall that, given an arbitrary set I, for {Ai}iIRMVσ, the coproduct iIAi is defined as a pair (iIAi{αi}iI) such that αi:AiiIAi and for any other object C and maps ηi: AiC, iI, there exists a unique map η:iIAiC such that ηi = η o αi for all iI. If, in addition, the maps αi are one-one, and iIαi(Ai) generates iIAi,iIAi is called free product of the Ai.

We note that the existence of all coproducts in RMVσ follows from [2: Theorem 9.4.14]. In the following proposition we use the same proof strategy sketched in [2] and given in details in [20: Theorem 7.1], and characterize free products of non-trivial σ-semisimple algebras, as defined in Section 2.

Proposition 4.1

Let {Ai}iIbe a collection ofσ-semisimple algebras inRMVσ, indexed in a set I. For any iI, let AiIRL(Xi)/Ii, where we assume the setsXito be pairwise disjoint. Then the free productiAiexists inRMVσand

i A i = IRL i X i / J

with J = i I i σ , the σ-ideal generated byiIi.

proof. Let us denote by A the algebra IRL(iXi)/JRMVσ. For any iI, define the map

α i : A i A , α i f / I i = f / J .

Each αi is well defined because IRL(Xi) embeds in IRL(iXi) and IiJ.

Let us divide the proof in three steps.

  1. The maps αi are embeddings.

    Let f / Ii be an element such that f/J = 0. Let us prove that fIi.

    By hypothesis, f / J = 0 implies fJ. By [7: Proposition 2.6] there exists a countable set of elements G={fn|n}iIi such that fτ(f1,…,fn,…), where τ is a term build using only ⊕ and ⋁. We note that the variables that appear in f belong to Xi. Moreover, any fn belongs to a proper ideal Ij, jI.

    Let Fi = {gG|gIi} be the subset of the fn’s that belong to Ii. For any gG \ Fi there exists ji and a point in yj∈[0,1]Xj, such that gIj and g(yj) = 0. To see this, assume that for some gG \ Fi, gIj, we have 𝕍(g) = ∅. Hence we deduce that 𝕀(𝕍(g)) = IRL(Xj). But [7: Lemma 4.15] implies 𝕀(𝕍(g)) = 〈gσIj, and Ij is proper ideal, which is a contradiction.

    Take now y∈[0,1]Xi such that g(y) = 0 for any gFi. Denote by z the point of [0,1]iXi such that z coincides with y in the Xi-coordinates and it coincides with yj in the Xj-coordinates, for ji, and it is arbitrary in the coordinates eventually not taken into account. Note that this is well defined as the sets of variables are disjoint. Then τ(f1,…,fn,…)(z) = 0 and therefore f(z) = 0.

    We have proved that g(y)= 0 for any gFi implies f(z)= 0, and consequently (since the only variables appearing in f belong to Xi) f(y)= 0. Thus, 𝕍(f) ⊇ 𝕍(Fi). By [7: Theorem 4.7 and Lemma 4.15] and the hypothesis on Ai we deduce that 〈fσ = 𝕀(𝕍(f)) ⊆ 𝕀(𝕍(Fi)) ⊆ 𝕀(𝕍(Ii)) = Ii. Thus, we have proved that fIi and αi is one-one.

  2. i α i ( A i ) generates A.

    Each IRL(Xi) is generated by {πx|xXi}, the coordinate projections. Analogously, {πx | xiXi} is a set of generators for IRL(iXi). Therefore, the set P={πx/J|xiXi} generates A. By the definition of αi, P coincides with i{αi(πx/Ii)|xXi}iXi, which settles the claim.

  3. (A,{αi}iI) satisfies the universal property.

    Let ERMVσ be an algebra such that we have homomorphisms ηi: AiE for any iI. We have to prove that there exists η: AE such that η o αi = ηi for any iI.

    Let Q={πx|xiXi} be the set of generators for IRL(iXi). Define ϱ: QE as ϱ(πx) = ηi(πx/Ii), when xXi. Note that ϱ is well defined since the Xi’s are pairwise disjoint. Let ϱ¯:IRL(iXi)E the unique σ-homomorphism that extends ϱ to the free algebra. Note that, by definition, and because {πx|xXi} generates IRL(Xi)IRL(iXi), for any iI and any fIRL(Xi), ϱ̅(f) = ηi(f/Ii). Take now η: AE to be defined as η(f/J) = ϱ̅(f).

Let us prove that η is well defined. Take f,gIRL(iXi) such that f/J = g/J. Then, denoted by d Chang’s distance, d(f,g)∈J, whence there exists a countable set {hn}niIi such that d(f,g)≤τ(h1,…,hn,…), where in τ only ⊕ and ⋁ appear. We have

d ( ϱ ¯ ( f ) , ϱ ¯ ( g ) ) = ϱ ¯ ( d ( f , g ) ) ϱ ¯ τ h 1 , , h n , = τ ϱ ¯ h 1 , , ϱ ¯ h n , .

Since each hn belongs to iIi,ϱ¯(hn)=0 and τ(ϱ̅(h1),…,ϱ̅(hn),…) = 0. Consequently, d(ϱ̅(f),ϱ̅(g)) = 0 and ϱ̅(f) = ϱ̅(g).

Finally, for any iI and any fIRL(Xi), (ηαi)(f/Ii) = η(αi(f/Ii)) = η(f/J) = ϱ̅(f) = ηi(f/Ii), settling the final part of the claim.

Definition 4.2

We shall call an algebra ARMVσ countably presented if AIRL(X)/I where X is a countable set of generators and I is principal. We shall denote by RMVσω the full subcategory of RMVσ whose objects are countably presented and σ-complete Riesz MV-algebras.

Remark 4.3

For any algebra ARMVσ, it is an easy exercise to show that countably generated σ-ideals in A are principal. Indeed, for a sequence {an}n∈ℕ in A, we have {an}nσ=Vnanσ.

Corollary 4.4

The full subcategory R M V σ ω of RMVσ whose objects are countably presented and σ-complete Riesz MV-algebras is closed under countable free products. The full subcategory R M V σ f p of RMVσ whose objects are finitely presented and σ-complete Riesz MV-algebras is closed under finite free products.

proof. It follows from [7: Lemma 4.15] that both countably presented algebras and finitely presented algebras are σ-semisimple. Moreover, if we deal with a countable set of algebras AiIRL(Xi)/Ii, and for some fiIRL(Xi) we have Ii=fiσ,J=Vifiσ by Remark 4.3.

Thus, IRL(iXi)/J is a countably presented algebra. Analogously, a finite free product will be finitely presented.

The next proposition, given in [7], is crucial for what follows.

Proposition 4.5

Let A be a σ-semisimple σ-complete Riesz MV-algebra and assume that IRL(X)/Iis a presentation for A. Then A is isomorphic to the Riesz tribeIRL(X)|𝕍(I) = {g ∈ [0,1]𝕍(I)} for some fIRL(X)}.

Given a tribe 𝓣 ⊆ [0,1]𝓣, we recall that S(𝓣) = {A𝓣|χA:[0,1]𝓣→{0,1},χA𝓣} denotes its Boolean center, as defined in Section 2. For any P⊆[0,1]X we can define the tribe of restrictions IRL(X)|P. In this case, for brevity and if it is clear from the context, we shall denote the Boolean center S(IRL(X)|P) by S(P).

Lemma 4.6

For any P⊆[0,1]X, denoted by 𝓑𝓐(P) = {AP|A∈𝓑𝓐([0,1]X)}, the algebra of restrictionsIRL(X)|P = {f|P|fIRL(X)} is the Riesz tribe of all 𝓑𝓐(P)-measurable functions.

proof. As remarked before, the algebra of restrictions IRL(X)|P = {f|P|fIRL(X)} is a Riesz tribe and by [20: Lemma 11.8] it is the algebra of all S(P)-measurable functions, where S(P) = {BP|χBIRL(X)|P}.

Thus, we only need to prove that the σ-algebras 𝓑𝓐(P) and S(P) coincide.

For any BP, BS(P) if, and only if, χBIRL(X)|P. The latter is equivalent to saying that there exists pIRL(X) such that χB = p|P. Take A⊆[0,1]X to be the Baire set 𝕍(1−p). It is easily seen that B=AP.

Conversely, for any set of type AP, with A∈𝓑𝓐([0,1]X), χAP:P→{0,1} coincide with χA|P. Since A is Baire-measurable, it follows that χAIRL(X) and APS(P).

Building on the previous results, let us describe a somewhat canonical way to define a state on the coproduct starting from algebras endowed with states.

Take a sequence of pairs (IRL(Xi)/Ii, si)iI. With the same notation as before, let IRL(X)/J be the free product of the sequence. We have X=iXi, the Xi are chosen to be pairwise disjoint, IRL(X)/JIRL(X)|𝕍(J). When it is clear from the context, we denote by πxi the xi-coordinate projection in either X-coordinates or Xi-coordinates, that is, πxi:[0,1]X→[0,1] and πxi[0,1]Xi→[0,1], with xiXiX.

Following the notation set out in [7] and briefly recalled in Section 2, we denote by Baireω the full subcategory of IRL whose objects are Baire subsets of hypercubes of type [0,1]X, for a countable X. Note that the duality given in [7] for finitely presented algebras is straightforwardly extended to a duality between Baireω and RMVσω. More precisely, the duality is between Baireω and the category of presentations for algebras in RMVσω. Nonetheless, this category of presentations is equivalent to RMVσω, see [4 Remark 4.7]MSC. Thus, the coproduct of a sequence IRL(Xi)}/IiRMVσωis reflected in a product in the dual category Baireω.

Hence, to a sequence of algebras IRL(Xi)/XiRMVσω we can naturally associate the sequence of measure spaces (𝕍(Ii), 𝒮(𝕍(Ii)), and to their coproduct we can associate the space (𝕍(J), 𝒮(𝕍(J))).

Lemma 4.7

Let I be a countable set and let {Xi}iIbe a pairwise disjoint sequence of sets of variables. For any iI, let fiIRL(Xi) be an IRL-polynomial, letX=iXiand let J be the σ-ideal generated by the sequence {fi}iIin [0, 1]X. Then 𝕍(J) is the product inBaireωof the objects 𝕍(fi), where the projection from 𝕍(J) to 𝕍(fi) is induced by the inclusion ofXiin X.

p r o o f. By definition, a point x ∈ [0, 1]X belongs to 𝕍(J) if, and only if, fi(x) = 0 for all iI. Since we have chosen the Xi’s to be pairwise disjoint, it follows that x ∈ 𝕍(J) if, and only if, (πxi(x))xiXi ∈ 𝕍(fi). Thus, 𝕍(J) is actually the set of those points in [0, 1]X such that the ‘Xi-coordinates’ form a point of 𝕍(fi), that is, their cartesian product.

For brevity, let us denote 𝕍(J) by V and 𝕍(fi) by Vi. For any iI, let ηi: V ⊆ [0, 1]XVi ⊆ [0, 1]Xi be defined by ηi(x) = (πxi(x))xiXi, where πxi[0, 1]Xi → [0, 1] are the coordinate projections that generate IRL(Xi). This is well defined because the inclusion XiX allows to think of IRL(Xi) as embedded in IRL(X). Hence, by definition, each ηi is an arrow in Baireω since it is a tuple of IRL-polynomials.

Take now any other WBaireω, W ⊆ [0, 1]Y such that for any iI there exists νi:WVi. Thus, by definition of arrows, νi(w) = (νxi(w))xiXi for some νxiIRL(Y).

Define ν: WV by ν(w)=(νx(w))xiXi. We note that ν is an IRL-map, moreover it is well defined since the Xi-coordinates of the point ν(w) belong to Vi, making ν(w) a point in V. Furthermore, for any iI, ηi(ν(w)) = (πxi(ν(w)))xiXi = (νxi(w))xiXi = νi(w). Consequently, (V, ηi) is the product of the Vi in Baireω.

For any collection of σ-algebras {𝓑i}iI, we denote by ×iBi the product σ-algebra generated by iAi,AiBi. When I is an infinite countable set, we require that Ai coincides with the whole space for all but a finite number of indexes.

Lemma 4.8

With the same notations of the previous lemma, the Boolean center 𝒮(V) of IRL(X)|Vis the product σ-algebra of the Boolean centers 𝒮(Vi) of IRL(Xi)|Vi.

p r o o f. We shall use the characterization of the Boolean center given in Lemma 4.6, as well as the notation fixed there.

Let us first remark that, since each Xi is countable, [0, 1]Xi is a compact and metric space, whence it is separable. Thus, we can apply [16: Lemma 1.2] and deduce that BA([0,1]X)=×iBA([0,1]Xi). Therefore sets of type iBi, with Bi𝓑𝓐([0, 1]Xi) are generators for 𝓑𝓐([0, 1]X). Furthermore, since V𝓑𝓐([0, 1]X) it is easily seen that (iBi)V is a set of generators for 𝓑𝓐(V), where iBi is defined as before.

The conclusion is now straightforward since, by Lemma 4.7, V=iVi. Indeed, take Ai = BiVi with Bi𝓑𝓐([0, 1]Xi). From the previous remarks we have that the product iAi is a generator of ×iS(Vi), the product iBi is a generator for 𝓑𝓐([0, 1]X) and the intersection iBiV is a generator for 𝓑𝓐(V).

Since iAi=i(BiVi)=iBiiVi=iBiV, we can see that the sets of generators of XiS(Vi) and S(V) are the same, and the two σ algebras coincide.

We can now build on the previous lemmas, using the same notation set out before. Assume that each IRL(Xi)/Ii/IiIRL(Xi)|Vi carries a state si. Then, we can naturally associate the measure space (Vi,S(Vi),μi), where μi is the probability measure defined by μi(A)= si(χA). Take now the usual product measure μ of the μi and define s: IRL(X)𝕍(J) →[0, 1] as the integral with respect to μ.

Definition 4.9

We call coproduct presentable any tribe 𝒯 that can be obtained as a coproduct of a sequence of algebras in RMVσω. If, in addition, each algebra of the sequence carries a σ-state, we call coproduct state the state obtained via the product measure, as described above.

With the same notations of the previous lemmas, for any finite set of indexes i1,…,ikI, define the map

(1) β I R L ( X i 1 ) V i 1 × × I R L ( X i k ) V i k I R L ( X ) V , β ( a i 1 , , a i k ) = a i 1 a i k ,

where the product · is the usual ring-product of real-valued functions. We note that β is a k-additive function and it is also well defined, since each aij can be embedded in IRL(X)|V and the product of 𝓑𝒜(V)-measurable functions is 𝓑𝒜(V)-measurable, thus it does belong to IRL(X)|V.

The next theorem shows how, in the finite case, the summands of the coproduct are independent with respect to the coproduct state.

Theorem 4.10

With the notation fixed above, if |I|ω, the σ-complete Riesz MV-algebras {(IRL(Xi)}Vi,si)}iIare (IRL(X)|V,s)-independent.

Proof

Take β as defined in equation (1). Take any subset of {(IRL(Xi)|Vi,si)}iI of cardinality k ∈ ℕ. Let h ∈ ℕ be another integer such that |I| =k+h and assume that I = {i1,…,ik} ∪ {j1,…,jh}. Furthermore, let 1Vj1,…,1Vjh denote the functions that are identically equal to 1 on Vj1,…,Vjh. We remark that, by definition of s and β, the measure associated to s is the product measure of the μi’s, and each μi is a finite measure. Thus, we can apply (iteratively) the well known Fubini-Tonelli theorem (see, e.g. [14Section 36]Halmos): for any ai1IRL(Xi1)|Vi1,…,aikIRL(Xik)|Vik,

s ( β ( a i 1 , , a i k ) ) = V a i 1 a i k d μ = V a i 1 a i k 1 V j 1 1 V j h d μ = ( V i 1 a i 1 d μ i 1 ) ( V i k a i k d μ i k ) μ j 1 ( V j 1 ) μ j h ( V j h ) = s i 1 ( a i 1 ) s i k ( a i k ) ,

settling the claim.

Definition 4.11

Let ARMVσω. Take the coproduct ωA of A with itself countably many times. A σ-state s:ωA[0,1] is called weakly exchangeable if the associate measure, on the product of countable copies of (𝕍(I),S(𝕍(I)), is exchangeable in the classical sense. Similarly, the σ-state s is called weakly presentable if the associated measure is presentable in the classical sense.

As a straightforward remark, the coproduct state is weakly exchangeable.

Lemma 4.12

For any nωand anyARMVσω, endowed with aσ-state s: A → [0,1], the coproduct state oni=1nAi, whereAiAfor anyi, is weakly exchangeable.

Proof

It is a consequence of the fact that a product measure on a countable power Vn is exchangeable, see [15].

The next theorem is our weak version of de Finetti’s theorem for states, which depends upon the classical result.

Theorem 4.13

(Weak de Finetti’s exchangeability). Let X be a countable set. A state on IRL(X) is weakly exchangeable if, and only if, it is weakly presentable.

Proof

Since IRL(X) is presented by the trivial ideal, for any state of ωIRL(X), the associated measure is defined on countable copies of 𝓑𝒜([0,1]X). Then, since [0,1]X is a compact and Hausdorff space, we apply Theorem 2.4.

We close this section with some remarks on these notions. While it is somewhat natural to define independence directly on both stochastic processes and sequences of tribes, and it is also natural to define exchangeability of processes as done in Definition 3.8, it is still unclear how to define “strongly” exchangeable and “strongly” presentable states is in this framework, where the adverb strongly is to be intended as “without reference to the associated measure”. Since Definition 4.11 is completely dependent on the associated measure, it will be interesting to understand if and how we can adjust the definition in this sense.

To give more strength to our results, one possibility is to explore states in a categorial setting. With this aim in mind, in the next section we extend the Butnariu-Klement and Kroupa-Panti representations for σ-states and states, into categorical dualities.

5 Dualities for algebras with states

In this section we see how both dualities given in [7] are suitable to be lifted to a setting that incorporates states and measures. We use the results of the previous section.

We start by enriching the duality between the algebraic category RMVσ of σ-complete Riesz MV-algebras and category BDKH of compact, Hausdorff and basically disconnected topological spaces endowed with cozero-closed continuous functions, see Proposition 2.1. Consider the following categories:

  1. pRMV σ is the category whose objects are pairs (A,s), with ARMVσ and s a state, and whose morphisms are state-preserving homomorphisms. That is, η (A,sA) → (B,sB) is a morphism in pRMVσ if it a σ-homomorphism of Riesz MV-algebras such that sA = sBη.

  2. pBDKH is the category whose objects are pairs (X,μ), with XBDKH and μ Borel regular probability measure on 𝓑𝒪(X), the σ-algebra of Borel subsets of X. Morphisms in pBDKH are continuous and cozero-closed that are also measure-preserving, that is, f(X,μX) → (Y,μY) such that μY = μXf–1.

The Kroupa-Panti theorem, together with the restriction of Kakutani’s duality given in [7], gives naturally an essentially surjective functor that maps each (X,μX) ∈ pBDKH into (C(X),sX) ∈ pRMVσ, with sX being defined by integration as in Theorem 2.2.

Moreover, the duality of [7] allows to associate to each morphism f : XY in BDKH, a unique :C(Y) → C(X), with (g)= gf for any gC(Y) σ-homomorphism of Riesz MV-algebras. Note that, since we are dealing with a duality, the assignment f is full and faithful.

Therefore, in order to lift the duality of [7] to pRMVσ and pBDKH, it is enough to show that measure-preserving functions are mapped to state-preserving homomorphisms and viceversa.

Proposition 5.1

Let f(X,μX) → (Y,μY) be morphism inBDKH, that is, a continuous and cozero closed function. Let(C(Y),sY) → (C(X),sX) be its corresponding homomorphism inRMVσ. ThenμY = μXf–1if, and only if, sY = sX.

Proof

By hypothesis f(X,μX) → (Y,μY) is continuous, cozero-closed and measure preserving and μY is the pushforward measure of μX with respect to f. Hence, by the property of change of variables of pushforward measures,

s Y ( g ) = Y g d μ Y = Y g d ( μ X f 1 ) = X ( g f ) d μ X = s X ( f ~ ( g ) ) ,

which settles one direction of the claim.

On the other hand, assume is a measure preserving σ-homomorphism. That is, for any gC(Y), sX((g))=sY(g). Than, as before,

Y g d μ Y = Y g d ( μ X f 1 ) .

By the Kroupa-Panti theorem, the regular measure associated to sY is unique. Thus, if we prove that μXf–1 is a regular measure, we can infer that μY = μX Π f–1.

Following [13], we recall that (since Y is a compact space) it is enough to prove that μXf–1 is inner regular, that is, for any E Borel subset of Y,

(2) ( μ X f 1 ) ( E ) = sup { ( μ X f 1 ) ( K ) K E , K compact } .

Let 𝒦={KY|KE,Kcompact}. One inequality of equation 2 follows from the fact that for any K ∈ 𝒦, (μXf–1)(K) ≤ (μXf–1)(E), whence (μXf–1)(E) is greater or equal of their supremum, that is,

(3) ( μ X f 1 ) ( E ) sup { ( μ X f 1 ) ( K ) K E , K compact } .

Let us prove the other inequality.

To do so, let

H = { C X C f 1 ( E ) , C compact }

and

D = { f 1 ( D ) X D E , D compact } .

For any C ∈ 𝓗, f(C)⊆E is a compact set. Hence f–1(f(C))∈ 𝒟. Moreover, Cf−1(f(C)) implies that μX(C) ≤ μX(f–1(f(C))). Consequently,

(4) sup { μ X ( C ) C H } sup { μ X ( f 1 ( D ) ) D E , D compact } .

By hypothesis, μX is a regular measure on X. Since f−1(E) is a Borelian set of X, it follows that

(5) μ X ( f 1 ( E ) ) = sup { μ X ( C ) C f 1 ( E ) , C compact }

Putting together (4) and (5) , we deduce

μ X ( f 1 ( E ) ) sup { μ X ( f 1 ( D ) ) D E , D compact } ,

and the latter, together with (3) gives equation (2) , settling the claim of regularity for μX o f−1.

Using the same notation settled before, we make explicit the duality obtained.

Corollary 5.2

The functor

P : p B D K H p R M V σ ( X , μ X ) ( C ( X ) , s X ) ( f X Y ) ( f ~ C ( Y ) C ( X ) )

yields a duality between pBDKH and pRMVV σ .

Proof. The restriction of Kakutani's duality proved in [7] and the Kroupa-Panti theorem imply that the functor is well defined and essentially surjective. It is full and faithful by the same duality of [7] and Theorem 5.1 which implies that measure-preserving morphisms in BDKH correspond to state-preserving morphisms in RMVσ.

Remark 5.3

Despite RMVσ being closed under coproducts, we notice that BDKH, the category dual to the whole RMVσ is not closed under the usual product of topological spaces. Indeed 2 = {0,1} with the discrete topology is an object there, while 2ω is not, since it is the Stone space of the Lindenbaum-Tarski algebra of the classical propositional calculus, which is not σ-complete. Consequently, products in BDKH do not coincide with products in Top, the category of topological spaces and continuous maps.

The duality obtained in Corollary 5.2 is built by combining two results: one is based on the classical Kakutani's duality, the other being the Kroupa-Panti integral representation. Both of these references work using the topological space of maximal MV-ideals of a σ-complete Riesz MV-algebra. Following the same ideas, the second duality is built on the integral representation of σ-states given in 1,3] , that is, Theorem 2.3 Consider the following categories:

  1. p R M V σ ω is the category whose objects are triplets (IRL(X), I, sx ), where IRL(X)/IIRL(X)|V(I)RMVσω and s a σ-state on IRL(X)|𝕍(I), and whose morphisms are statepreserving homomorphisms of RMVσω.

    That is, h: (IRL(X),J, sX) → (IRL(Y),K , sF) is induced by a unique h*: IRL(X) → IRL(Y) such that h*(J) ⊆ K and sX = sY o h.

  2. pBaire ω is the category whose objects are pairs (V, μ), where VBaireω and μ is a probability measure on 𝓑𝓐(V), the σ-algebra of Baire subsets of V. Morphisms in pBaireω are tuples of IRL-maps that are also measure-preserving, that is, f: (V, μV) → (W, μW) such that μV = μV o f−1.

Consider now the functors Vp:pRMVσωpBaireω and Ip:pBaireωpRMVσω defined upon the functors and 𝒱 of 𝒥 as follows:

  1. 𝒱p(IRL(X), I, s) = (𝕍(I), μs), where μs is given by Theorem 2.3

  2. for V ⊆ [0,1]X, 𝒥p(V, μ) = (IRL(X), 𝕀(V), sμ), with sμ: IRL(X)|V → [0,1] defined by integration with respect to μ. Notice that 𝕍(𝕀(V)) = V.

  3. on arrows, with the obvious notations, 𝒱p(h) = 𝒱(h) and 𝒥p(η) = 𝒱(η). We remark that 𝒱(η) is defined by precomposition with η.

THEOREM 5.4

The functors 𝒱p and 𝒥p give a duality.

Proof. Take (IRL(X),J,s)pRMVσω. By Lemma 4.6, IRL(X)|𝕍(J) is the algebra of all 𝓑𝓐(𝕍(J))-measurable functions. Thus, the measure μs given in Theorem 2.3 is a measure on 𝓑𝓐(𝕍(J)) and consequently 𝒱p(IRL(X), j, s) ∈ pBaireω. Similarly, for any (V, μ) ∈ bBaireω, with V ⊆ [0,1]X, the triplet (IRL(X), 𝕀(V), sμ) = Jp(V, μ) is an object in pRMVσω.

Furthermore, by [7: Corollary 4.17] and Theorem 2.3

V p ( I p ( V , μ ) ) = ( V , μ ) and I p ( V p ( I R L ( X ) , I , s ) ) = ( I R L ( X ) , I , s ) .

Whence, in order to prove that we have a duality, it is enough to show that a state- preserving morphism in RMVσω is mapped to a measure-preserving map of Baireω, and viceversa.

Let h: (IRL(X), j, sX) → (IRL(X), K, sY) be a morphism in RMVσω induced by some h*. Then sX = sY o h with sX: IRL(X)|𝕍(J) → [0,1] and sY: IRL(Y)|𝕍(K) → [0,1].

Take A ∈ 𝓑𝓐(𝕍(J)). Then μsX(A) = sX(𝒳A) = sX(h(𝒳A)). For brevity, let η denote 𝒱p(h). Note that it follows from [7] that the inverse functor of acts as the pre-composition with η. More precisely, for any pIRL(X)|V, h(p) = p o η. Then, by definition of η, it follows that ?(3A) = 3y−1(A). Indeed, ?(3A) is again a Boolean member of IRL(Y)) |𝕍(K) and

χ η 1 ( A ) ( y ) = 1 y η 1 ( A ) η ( y ) A χ A ( η ( y ) ) = 1 .

Then, μsX (A) = sY(𝒳η−1(A)) = μsY−1(A)), settling the claim.

Finally, we see how h(p): = p o η is state-preserving for any morphism η: (V ⊆ [0,1]X, μX) → (W ⊆ [0; 1]Y , μY ) in Baireω that is measure-preserving. By definition of arrows in pBaireω, let η : = (ηy)yY and each ηyIRL(X). We first remark that η is measurable between (V, 𝓑𝓐(V)) and (W, 𝓑𝓐(W)). Indeed, for any generator E of 𝓑𝓐(W), E=WyYAy with Ay ∈ 𝓑𝓐([0,1]), see [16 Lemma 1.2]. Then

x η 1 ( E ) η ( x ) W y Y A y η ( x ) W  and  η y ( x ) A y x V y Y η y 1 ( A y ) .

Since each ηy belongs to IRL(X), it follows that yYηy1(Ay)BA([0,1]X) and VyYηy1(Ay)

𝓑𝓐(V), making η a measurable function.

Now, using the property of change of variables of the pushforward measure, which is μY by hypothesis, we have

S Y p = W p d μ Y W p d μ x η 1 = V p η d μ X = V h p d μ X = s X h p ,

which settles the claim.

6 Statistical models via IRL-maps

A classical approach in statistics defines a model to be a collection of probability distributions ℳ = {pθ | θ ∈ Θ} = [21,23]. The set of parameters Θ is usually taken inside a finite dimensional space, that is, Θ ⊆ ℝd and often parameters are the values of a density function on a certain dataset. Outcomes are often assumed to be in a finite number and it is possible to identify each ∈ ℳ as a point of the simplex Δk1={(p1,,pk)[0,1]ki=1kpi=1}, for some k ∈ ℕ . This is the starting point for algebraic statistics, which was introduced in [21]. Here we follow a slightly different approach, as described in [23]: a parametric algebraic statistical model is, classically, a model in which the set of parameters Θ is an algebraic variety and ℳ sits inside a probability simplex Δκ−1 for some κ.

Thus, inspired by these considerations, let us give the following definition.

Definition 6.1

An IRL-statistical model is an IRL-map

η = ( η 1 , , η k ) : P [ 0 , 1 ] X Δ k 1

such that |X| ≤ ω and the set P is an IRL-algebraic variety. That is, there exists a set of polynomials FIRL(X) such that P = 𝕍(F).

If we denote by Δω the simplex in [0,1]ω, that is, the set of points (an) ∈ [0,1]ω for which the infinite sum nan exists and it is equal to 1, this definition is generalized to the case η: [0,1]d → Δω, where we take η = (ηn)nω.

If we look at points of η(P) as functions κ → [0,1], every such point can be considered as a discrete probability measure. Hence, we have indeed defined a parametrized statistical model (in the classical sense) with additional conditions on the set of parameters and the assignment. In our case P plays the role of Θ, while η(P) = ℳ.

By the very nature of IRL-models, we obtain the following propositions.

Proposition

Let P be a Baire set. A model η : P ≤ [0,1]X → Δκ−1 ≤ [0,1]κ induces an observable 𝒳y: IRL(Y) → IRL(X)|p, |Y| = κ, given by a ∈ IRL(Y) ↦ a o η. The same holds in the case of a model η : P → Δω.

Proof. By Lemma 4.6 the algebra IRL(X)|p is the Riesz tribe of all 𝓑𝓐(P)-measurable functions, as defined before. Thus, looking at IRL(Y) and IRL(X)|p respectively as the free κ- generated σ-complete Riesz MV-algebra and a tribe inside [0,1]p, we apply 6 Theorem 2.3] (in both the finite and countable case) getting the desired result, for which we notice that the measurable map η is not required to be surjective into [0,1]Y or [0,1]ω.

Furthermore, a consequence of the duality of 7 is the following result.

Proposition 6.3

Statistical models on Baire subsets of hypercubes [0,1]X, with X countable, are in bijective correspondence with morphisms between countably presented σ-complete Riesz MV algebras.

We now give a list of examples, in order to show the applicability of this definition. Note that in every example the IRL-variety taken as domain is a simple Borel subset of [0,1]. In particular it is [0,1] itself in the case of a binomial model, it is (0,1) for the geometric model, and it is (0,1] for the Poisson model.

Example 6.4

(Binomial model). Let [κ] = {0,1,2, …, κ} be a set of data. In this case, our data represent the iteration of an experiment, for example, the tosses of a coin (biased or unbiased, depending on the parameter). Consider, for any i ∈ [κ], the function:

η i : [ 0 , 1 ] [ 0 , 1 ] η i ( α ) = k i α i ( 1 α ) k i ,

and let η: [0,1] → [0,1]κ+1 be defined as η = (η0, … , ηκ).

Each ηi is continuous, and therefore and element of IRL(1). Moreover, since for a fixed α, i=0kηi(α) gives the cumulative distribution, η([0,1]) ⊆ Δκ. Whence, we obtain a IRL- model.

The binomial distribution is an example of an algebraic model in the definition of [23] , since each ηi is actually a polynomial. Nevertheless, our definition allows to capture more complicated distributions, which are not given in polynomial form.

Example 6.5

(Geometric distribution). Similarly to the case of a binomial distribution, we take κ ∈ ℕ and define

η i : ( 0 , 1 ) [ 0 , 1 ] η i ( x ) = x ( 1 x ) i 1 ,

for any i = 1, … , κ. This distribution gives the probability that the first occurrence of success requires i independent trials, each with success probability x, with a bound of κ trials.

The function η = (ηi, … , ηκ) is trivially an IRL-map, but in contrast with the previous case, it is not true now that i=1kηi(x)=1 for a fixed x ∈ (0,1). Indeed, we have

i = 1 k η i ( x ) = i = 1 k x ( 1 x ) i 1 = 1 ( 1 x ) k

Since x ∈ (0,1), the geometric series is convergent, therefore we can consider

η : ( 0 , 1 ) [ 0 , 1 ] ω , η = ( η i ) i ω .

In this case, for any fixed x ∈ (0,1),

i ω η i ( x ) = x i ω ( 1 x ) i 1 = x 1 ( 1 x ) = 1.

Thus, η((0,1)) ⊆ Δω and we have obtained a logico-algebraic representation of a geometric model within 𝓘𝓡𝓛.

Example 6.6

(Poisson model). Let [κ] = {0,1,2, … , κ} be a set of data. In this case, our data represent the number of occurrences of a certain phenomenon, where we put an arbitrary large bound given by κ ∈ ℕ. Consider, for any i ∈ [κ], the function:

η i : ( 0 , 1 ] [ 0 , 1 ] η i ( λ ) = e λ λ i i ! ,

and let η : (0,1] → [0,1]κ be defined as η = (η1, … , ηκ).

Each ηi is continuous, and therefore an element of IRL(1). The domain (0,1] is clearly an IRL-algebraic variety, being a Baire set. Moreover, when we consider the case κ = ω, the infinite sum iNkηi(λ) equals 1 , since iNλii!=eλ. Thus, η((0,1]) ⊆ Δω. Thus, once again the possibility to deal with countable sequences allows us to obtain a well defined logical model.

The definitions and examples of this short section are meant to hint towards the idea of a “metamathematics of statistics” and they are a first step in this direction. Propositions 6.2 and 6.3 allow to look at algebraic statistical models with a point of view that is completely encoded into logic, since observables are evaluations of formulas in a standard-complete logical system. It can also be interesting to understand if this setting can be generalized using IRL-maps η : PC ⊆ [0; 1]X where X can be uncountable: The dualities of Section 5 and [7], as well as the results of Section 3 seem to imply that the countable case is the best-behaved one. Furthermore, it will be interesting in the future to see how far we can push this approach, for example in the setting of experiments of designs and when we use morphisms of pBaireω.

  1. ( Communicated by Roberto Giuntini)

Acknowledgement

The authors are grateful to Prof. Antonio Di Nola for many stimulating conversations on the topics of this paper.

References

[1] BARBIERI, G.—WEBER, H.: Measures on clans and on MV-algebras. In: Handbook of Measure Theory, Vol. I, II, North-Holland, Amsterdam, 2002, pp. 911–945.10.1016/B978-044450263-6/50023-3Search in Google Scholar

[2] BERGMAN, G. M.: An Invitation to General Algebra and Universal Constructions, Second Edition, Springer 1998, 2015.10.1007/978-3-319-11478-1Search in Google Scholar

[3] BUTNARIU, D.—KLEMENT E. P.: Triangular Norm Based Measures and Games with Fuzzy Coalitions, Kluwer, Dordrecht, 1993.10.1007/978-94-017-3602-2Search in Google Scholar

[4] CARAMELLO, O.—MARRA, V.—SPADA, L.: General affine adjunctions, Nullstellensatze, and dualities, J. Pure Appl. Algebra 225(1) 2021, Art. ID 106470.10.1016/j.jpaa.2020.106470Search in Google Scholar

[5] CIGNOLI, R. L. O.—D’OTTAVIANO, I. M. L.—MUNDICI, D.: Algebraic Foundations of Many-valued Rea- soning. Trends Log. Stud. Log. Libr., Kluwer, Dordrecht, 2000.10.1007/978-94-015-9480-6Search in Google Scholar

[6] DI NOLA, A.—DVUREČENSKIJ, A.—LAPENTA, S.: An approach to stochastic processes via non-classical logic, Ann. Pure Appl. Logic 172(9) (2021), Art. ID 103012.10.1016/j.apal.2021.103012Search in Google Scholar

[7] DI NOLA, A.—LAPENTA, S.—LENZI, G.: Dualities and algebraic geometry of Baire functions in Non- classical Logic, J. Logic Comput. 31(7) (2021), 1868–1890.10.1093/logcom/exab037Search in Google Scholar

[8] DI NOLA, A.—LAPENTA, S.—LEUŞTEAN, I.: Infinitary logic and basically disconnected compact Hausdorff spaces, J. Logic Comput. 28(6) (2018), 1275–1292.10.1093/logcom/exy011Search in Google Scholar

[9] DI NOLA, A.—LEUŞTEAN, I.: Łukasiewicz logic and Riesz spaces, Soft Comput. 18(12) (2014), 2349–2363.10.1007/s00500-014-1348-zSearch in Google Scholar

[10] FAGIN, R.—HALPERN, J. Y.—MEGIDDO, N.: A logic for reasoning about probabilities, Inform. and Comput. 87(1–2) (1990), 78–128.10.1016/0890-5401(90)90060-USearch in Google Scholar

[11] FLAMINIO, T.—GODO, L.: A logic for reasoning about the probability of fuzzy events, Fuzzy Sets and Systems 158(6) (2007), 625–638.10.1016/j.fss.2006.11.008Search in Google Scholar

[12] FLAMINIO, T.—HOSNI, H.—LAPENTA, S.: Convex MV-algebras: many-valued logics meet decision theory, Studia Logica 106(5) (2018), 913–945.10.1007/s11225-016-9705-9Search in Google Scholar

[13] FLAMINIO, T.—KROUPA, T.: States of MV-algebras. In: Handbook of Mathematica Fuzzy Logic vol. 3, Chapter XVII, 2015, pp. 1191–1245.Search in Google Scholar

[14] HALMOS, P. R.: Measure Theory. Grad. Texts in Math. 18, Springer-Verlag New York, 1950.10.1007/978-1-4684-9440-2Search in Google Scholar

[15] HEWITT, E.—SAVAGE, L. J.: Symmetric measures and cartesian products, Trans. Amer. Math. Soc. 80 (1955), 470–501.10.1090/S0002-9947-1955-0076206-8Search in Google Scholar

[16] KALLENBERG, O.: Foundations of Modern Probability. 2nd ed., Probability and its Applications (New York), Springer-Verlag, New York, 2002.10.1007/978-1-4757-4015-8Search in Google Scholar

[17] LAPENTA, S.—LEUŞTEAN, I.: Stochastic independence for probability MV-algebras, Fuzzy Sets and Systems 298 (2016), 194–206.10.1016/j.fss.2015.08.008Search in Google Scholar

[18] LUXEMBURG, W. A. J.—ZAANEN, A. C.: Riesz Spaces I. North-Holland, Amsterdam, 1971.Search in Google Scholar

[19] MUNDICI, D.: Averaging the truth-value in Łukasiewicz logic, Studia Logica 55 (1995), 113–127.10.1007/BF01053035Search in Google Scholar

[20] MUNDICI, D.: Advanced Łukasiewicz Calculus and MV-Algebra. Trends Log. Stud. Log. Libr. 35, Springer, Dordrecht, 2011.10.1007/978-94-007-0840-2Search in Google Scholar

[21] PISTONE, G.—RICCOMAGNO, E.—WYNN, H. P.: Algebraic Statistics: Computational Commutative Algebra in Statistics. Monogr. Statist. Appl. Probab. 89, Chapman & Hall/CRC, 2000.10.1201/9781420035766Search in Google Scholar

[22] SŁOMIŃ, J.: The Theory of Abstract Algebras with Infinitary Operations, Instytut Matematyczny Polskiej Akademii Nauk, Warszawa, 1959.Search in Google Scholar

[23] ZIMMERMANN, K-H.: Algebraic Statistics, available at https://doi.org/10.15480/882.1273.Search in Google Scholar

Received: 2021-04-15
Accepted: 2021-08-17
Published Online: 2022-08-10
Published in Print: 2022-08-26

© 2022 Mathematical Institute Slovak Academy of Sciences

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 15.12.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ms-2022-0058/html
Scroll to top button