Skip to content
BY 4.0 license Open Access Published by De Gruyter July 16, 2020

Operations that preserve integrability, and truncated Riesz spaces

  • Marco Abbadini EMAIL logo
From the journal Forum Mathematicum

Abstract

For any real number p[1,+), we characterise the operations I that preserve p-integrability, i.e., the operations under which, for every measure μ, the set p(μ) is closed. We investigate the infinitary variety of algebras whose operations are exactly such functions. It turns out that this variety coincides with the category of Dedekind σ-complete truncated Riesz spaces, where truncation is meant in the sense of R. N. Ball. We also prove that generates this variety. From this, we exhibit a concrete model of the free Dedekind σ-complete truncated Riesz spaces. Analogous results are obtained for operations that preserve p-integrability over finite measure spaces: the corresponding variety is shown to coincide with the much studied category of Dedekind σ-complete Riesz spaces with weak unit, is proved to generate this variety, and a concrete model of the free Dedekind σ-complete Riesz spaces with weak unit is exhibited.

1 Introduction

1.1 Operations that preserve integrability

In this work we investigate the operations which are somehow implicit in the theory of integration by addressing the following question: which operations preserve integrability, in the sense that they return integrable functions when applied to integrable functions?

Let us clarify the question by recalling some definitions.

For (Ω,,μ) a measure space (with the range of μ in [0,+]) and p[1,+), we adopt the notation p(μ){f:Ωf is -measurable and Ω|f|pdμ<}. It is well known that, for f,gp(μ), we have f+gp(μ), that is, p(μ) is closed under the pointwise addition induced by addition of real numbers +:2. More generally, consider a set I and a function τ:I, which we shall call an operation of arity |I|. We say p(μ) is closed under τ if τ returns functions in p(μ) when applied to functions in p(μ), that is, for every (fi)iIp(μ), the function τ((fi)iI):Ω given by xΩτ((fi(x))iI) belongs to p(μ). If p(μ) is closed under τ, we also say that τ preserves p-integrability over (Ω,,μ). Finally, we say that τ preserves p-integrability if τ preserves p-integrability over every measure space.

In Part I of this paper we characterise those operations that preserve integrability. Indeed, the first question we address is the following.

Question 1.1.

Under which operations I are p spaces closed? Equivalently, which operations preserve p-integrability?

Examples of such operations are the constant 0, the addition +, the binary supremum and infimum , and, for λ, the scalar multiplication λ() by λ. A further example is the operation of countably infinite arity defined as

(y,x0,x1,)supnω{xny}.

Yet another example is the unary operation

¯:,
xx¯x1,

called truncation. Here, although the constant function 1 belongs to p(μ) if, and only if, μ is finite, it is always the case that fp(μ) implies f¯p(μ).

It turns out that, for any given p, the operations that preserve p-integrability are essentially just 0, +, , λ() (for each λ), and ¯, in the sense that every operation that preserves p-integrability may be obtained from these by composition. This we prove in Theorem 2.3.

We also have an explicit characterisation of the operations that preserve p-integrability. Denoting with + the set {λλ0}, for nω and τ:n, we will prove that τ preserves p-integrability precisely when τ is Borel measurable and there exist λ0,,λn-1+ such that, for every xn, we have

|τ(x)|i=0n-1λi|xi|.

Theorem 2.1 tackles the general case of arbitrary arity, settling Question 1.1.

In Part I we also address a variation of Question 1.1 where we restrict attention to finite measures. Recall that a measure μ on a measurable space (Ω,) is finite if μ(Ω)<. The question becomes:

Question 1.2.

Under which operations I are p spaces of finite measure closed? Equivalently, which operations preserve p-integrability over finite measure spaces?

As mentioned, the function constantly equal to 1 belongs to p(μ) for every finite measure μ. We prove in Theorem 2.4 that, for any given p, the operations that preserve p-integrability over finite measure spaces are essentially just 0, +, , λ() (for each λ), and 1, in the same sense as in the above.

Theorem 2.2 provides an explicit characterisation of the operations that preserve p-integrability over finite measure spaces. In particular, for nω and τ:n, τ preserves p-integrability over finite measure spaces precisely when τ is Borel measurable and there exist λ0,,λn-1,k+ such that, for every xn, we have

|τ(x)|k+i=0n-1λi|xi|.

1.2 Truncated Riesz spaces and weak units

In Part II of this paper we investigate the equational laws satisfied by the operations that preserve p-integrability. (As it is shown by Theorems 2.1 and 2.2, the fact that an operation preserves p-integrability – over arbitrary and finite measure spaces, respectively – does not depend on the choice of p. Hence, we say that the operation preserves integrability.) We therefore work in the setting of varieties of algebras [4]. In this paper, under the term variety we include also infinitary varieties, i.e., varieties admitting primitive operations of infinite arity. For background please see [16].

We assume familiarity with the basic theory of Riesz spaces, also known as vector lattices. All needed background can be found, for example, in the standard reference [12]. As usual, for a Riesz space G, we set G+{xGx0}.

A truncated Riesz space is a Riesz space G endowed with a function ¯:G+G+, called truncation, which has the following properties for all f,gG+.

  1. fg¯f¯f.

  2. If f¯=0, then f=0.

  3. If nf=nf¯ for every nω, then f=0.

The notion of truncation is due to R. N. Ball [2], who introduced it in the context of lattice-ordered groups. Please see Section 8 for further details.

Let us say that a partially ordered set B is Dedekind σ-complete if every nonempty countable subset AB that admits an upper bound admits a supremum. Theorem 10.2 proves that the category of Dedekind σ-complete truncated Riesz spaces is a variety generated by . This variety can be presented as having operations of finite arity only, together with the single operation of countably infinite arity. Moreover, we prove that the variety is finitely axiomatisable by equations over the theory of Riesz spaces. One consequence (Corollary 10.4) is that the free Dedekind σ-complete truncated Riesz space over a set I (exists, and) is

Ft(I){f:If preserves integrability}.

We prove results analogous to the foregoing for operations that preserve integrability over finite measure spaces. An element 1 of a Riesz space G is a weak (order) unit if 10 and, for all fG, f1=0 implies f=0. Theorem 12.2 shows that the category of Dedekind σ-complete Riesz spaces with weak unit is a variety generated by , again with primitive operations of countable arity. It, too, is finitely axiomatisable by equations over the theory of Riesz spaces. By Corollary 12.4, the free Dedekind σ-complete Riesz space with weak unit over a set I (exists, and) is

Fu(I){f:If preserves integrability over finite measure spaces}.

The varietal presentation of Dedekind σ-complete Riesz spaces with weak unit was already obtained in [1]. Here we add the representation theorem for free algebras, and we establish the relationship between Dedekind σ-complete Riesz spaces with weak unit and operations that preserve integrability. The proofs in the present paper are independent of [1]. On the other hand, the results in this paper do depend on a version of the Loomis–Sikorski Theorem for Riesz spaces, namely Theorem 9.3 below. A proof can be found in [7], and can also be recovered from the combination of [5] and [6]. The theorem and its variants have a long history: for a fuller bibliographic account please see [5].

1.3 Outline

In Part I we characterise the operations that preserve integrability, and we provide a simple set of operations that generate them. Specifically, we characterise the operations that preserve measurability, integrability, and integrability over finite measure spaces, respectively in Sections 3, 4, and 5. In Section 6 we show that the operations 0, +, , λ() (for each λ), and ¯ generate the operations that preserve integrability, and that 0, +, , λ() (for each λ), and 1 generate the operations that preserve integrability over finite measure spaces.

In Part II we prove that the categories of Dedekind σ-complete truncated Riesz spaces and Dedekind σ-complete Riesz spaces with weak unit are varieties generated by . In more detail, in Section 7 we define the operation , in Section 8 we define truncated lattice-ordered abelian groups, in Section 9 we prove a version of the Loomis–Sikorski Theorem for truncated -groups, in Section 10 we show the category of Dedekind σ-complete truncated Riesz spaces to be generated by , in Section 11 we prove a version of the Loomis–Sikorski Theorem for -groups with weak unit, in Section 12 we show the category of Dedekind σ-complete Riesz spaces with weak unit to be generated by .

Finally, as an additional result, in the Appendix we provide an explicit characterisation of the operations that preserve -integrability.

Notation.

We let ω denote the set {0,1,2,}.

2 Main results of Part I

In this section we state the main results of Part I, together with the needed definitions. The first two main results (Theorems 2.1 and 2.2) are a characterisation of the operations that preserve p-integrability over arbitrary and finite measure spaces, respectively. The other two main results (Theorems 2.3 and 2.4) provide a set of generators for these operations. To state the theorems, we introduce a little piece of terminology.

For a set I, and iI, we denote by πi:I the projection onto the i-th coordinate. The cylinder σ-algebra on I (notation: Cyl(I)) is the smallest σ-algebra which makes each projection function πi:I measurable. If |I||ω|, then the cylinder σ-algebra on I coincides with the Borel σ-algebra (see [10, Lemma 1.2]).

Theorem 2.1.

Let I be a set, τ:RIR and p[1,+). The following conditions are equivalent.

  1. τ preserves p-integrability.

  2. τ is Cyl(I)-measurable and there exist a finite subset of indices JI and nonnegative real numbers (λj)jJ such that, for every vI, we have

    |τ(v)|jJλj|vj|.

Theorem 2.2.

Let I be a set, τ:RIR and p[1,+). The following conditions are equivalent.

  1. τ preserves p-integrability over every finite measure space.

  2. τ is Cyl(I)-measurable and there exist a finite subset of indices JI and nonnegative real numbers (λj)jJ and k such that, for every vI, we have

    |τ(v)|k+jJλj|vj|.

Theorems 2.1 and 2.2 show that the fact that an operation preserves p-integrability – over arbitrary and finite measure spaces, respectively – does not depend on the choice of p. Hence, once Theorems 2.1 and 2.2 will be settled, we will simply say that the operation preserves integrability.

The other two main results of Part I (Theorems 2.3 and 2.4 below) provide a set of generators for the operations that preserve integrability over arbitrary and finite measure spaces, respectively. To state the theorems, we start by defining, for any set 𝒞 of operations τ:Jτ, what we mean by operations generated by 𝒞. Given two sets Ω and I, a subset SΩ, and a function τ:I, we say that S is closed under τ if, for every family (fi)iI of elements of S, we have that τ((fi)iI) (which is the function from Ω to which maps x to τ((fi(x))iI)) belongs to S. Consider a set 𝒞 of functions τ:Jτ, where the set Jτ depends on τ. We say that a function f:I is generated by 𝒞 if f belongs to the smallest subset of I which contains, for each iI, the projection function πi:I, and which is closed under each element of 𝒞.

Theorem 2.3.

For every set I, the operations RIR that preserve integrability are exactly those generated by the operations 0, +, , λ() (for each λR), , and ¯.

Theorem 2.4.

For every set I, the operations RIR that preserve integrability over every finite measure space are exactly those generated by the operations 0, +, , λ() (for each λR), , and 1.

The rest of Part I is devoted to a proof of Theorems 2.12.4.

3 Operations that preserve measurability

In this section we study measurability, which is a necessary condition for integrability. In particular, we characterise the operations that preserve measurability (Theorem 3.3). This result will be of use in the following sections as preservation of measurability is necessary to preservation of integrability (Lemma 4.2). Let us start by defining precisely what we mean by “to preserve measurability”.

Definition 3.1.

Let τ:I be a function. For (Ω,) a measurable space, we say that the function τ preserves measurability over (Ω,) if, for every family (fi)iI of -measurable functions from Ω to , the function τ((fi)iI):Ω is also -measurable. We say that τ preserves measurability if τ preserves measurability over every measurable space.

When we regard as a measurable space, we always do so with respect to the Borel σ-algebra, denoted by .

Lemma 3.2.

Let (Ω,F) be a measurable space, I a set and f:ΩRI a function. Then f is F-Cyl(RI)-measurable if, and only if, for every iI the function πif:ΩR is F-BR-measurable.

Proof.

See [17, Theorem 3.1.29 (ii)]. ∎

Now we can obtain a characterisation of the operations that preserve measurability.

Theorem 3.3.

Let I be a set and let τ:RIR be a function. The following are equivalent.

  1. τ preserves measurability.

  2. τ preserves measurability over (I,Cyl(I)).

  3. τ is Cyl(I)-measurable.

Proof.

(1)  (2) Trivial.

(2)  (3) For every iI, πi:I is Cyl(I)-measurable. Since τ preserves measurability, τ((πi)iI) is Cyl(I)-measurable. Since (πi)iI:II is the identity, τ((πi)iI)=τ(πi)iI=τ is Cyl(I)-measurable.

(3)  (1) Let us consider a measurable space (Ω,) and a family (fi)iI of measurable functions fi:Ω. Consider the function (fi)iI:ΩI, x(fi(x))iI. We have πi(fi)iI=fi, therefore πi(fi)iI is measurable for every iI. Thus, by Lemma 3.2, (fi)iI is measurable. Thus τ((fi)iI)=τ(fi)iI is measurable, because it is a composition of measurable functions. ∎

3.1 The operations that preserve measurability depend on countably many coordinates

A fact that will be of use in the following sections is that the operations that preserve measurability depend on countably many coordinates. This we show in Corollary 3.6 below. Let us start by recalling what is meant with “to depend on countably many coordinates”.

Definition 3.4.

Given a set I.

  1. Let SI. For JI, we say that Sdepends only on J if, given any x,yI such that xj=yj for all jJ, we have xSyS. We say that Sdepends on countably many coordinates if there exists a countable subset JI such that S depends only on J.

  2. Let τ:I be a function. For JI, we say that τ depends only on J if, given any x,yI such that xj=yj for all jJ, we have τ(x)=τ(y). We say that τ depends on countably many coordinates if there exists a countable subset JI such that τ depends only on J.

We believe that the following proposition is folklore, but we were not able to locate an appropriate reference.

Proposition 3.5.

If τ:RIR is Cyl(RI)-measurable, then τ depends on countably many coordinates.

Proof.

First, every element of Cyl(I) depends on countably many coordinates: indeed, the set of elements of Cyl(I) which depend on countably many coordinates is a σ-subalgebra of Cyl(I) which makes the projection functions measurable (see also [9, 254M(c)]). Second, let τ:I be Cyl(I)-measurable. The idea that we will use is that τ is determined by the family (τ-1((a,+)))a. For every a, there exists a countable subset JI such that the measurable set τ-1((a,+)) depends only on Ja. Then JaJa has the property that, for each b, τ-1((b,+)) depends only on J. We claim that τ depends only on J. Let x,yI be such that xj=yj for every jJ. We shall prove τ(x)=τ(y). Suppose τ(x)τ(y). Without loss of generality, τ(x)<τ(y). Let a be such that τ(x)<a<τ(y). Then xτ-1((a,+)) and yτ-1((a,+)). This implies that it is not true that τ-1((a,+)) depends only on J. ∎

Corollary 3.6.

Let I be a set and τ:RIR be a function. If τ preserves measurability, then τ depends on countably many coordinates.

Proof.

If τ preserves measurability, then τ is Cyl(I)-measurable by Theorem 3.3. By Proposition 3.5, the function τ depends on countably many coordinates. ∎

3.2 The case of uncountable Polish spaces

The remaining results in this section are not used in the proofs of our main results.

One may think that, for an operation τ:I, the condition “τ preserve measurability over every measurable space” is too strong because we may not be interested in all measurable spaces. However, Proposition 3.7 shows that this condition is equivalent to “τ preserve measurability over (,)” (if τ has countable arity).

Proposition 3.7.

For a set I such that |I||ω| and a function τ:RIR, τ preserves measurability if, and only if, τ preserves measurability over (R,BR).

Proof.

If I=, then τ is a constant function. Hence τ preserves measurability over every measurable space. Let us consider the case I. By Theorem 3.3, τ preserves measurability if, and only if, τ preserves measurability over (,Cyl(I)). Since I and are uncountable Polish spaces with Borel σ-algebras Cyl(I) and , respectively, (I,Cyl(I)) and (,) are isomorphic measurable spaces (see [17, Theorem 3.3.13]). (Recall that an isomorphism of measurable spaces (Ω,) and (Ω,) is a bijective measurable function f:ΩΩ such that its inverse is measurable.) ∎

Remark 3.8.

In Proposition 3.7 above, one may replace the measurable space (,) by any of its isomorphic copies. In particular, one may replace it with the measurable space given by any uncountable Polish space endowed with its Borel σ-algebra (see [17, Chapter 3]).

4 Operations that preserve integrability

The goal of this section is to prove Theorem 2.1, i.e., to characterise the operations that preserve p-integrability.

Remark 4.1.

Let (Ω,) be a measurable space, and let μ0 be the null-measure on (Ω,): for each A, μ0(A)=0. Then p(μ0) is the set of -measurable functions from Ω to . Hence, preservation of p-integrability over (Ω,,μ0) is equivalent to preservation of measurability over (Ω,).

An immediate consequence of Remark 4.1 is the following lemma.

Lemma 4.2.

Let I be a set, τ:RIR and p[1,+). If τ preserves p-integrability, then τ preserves measurability.

Lemma 4.3.

Let (Ω,F,μ) be a measure space, and let f,g:ΩR be functions, and let λR. Then the following properties hold.

  1. If fp(μ), then |f|p(μ).

  2. If fp(μ), then λfp(μ).

  3. If f,gp(μ), then f+gp(μ).

  4. If gp(μ),|f||g| and f is -measurable, then fp(μ).

Proof.

Statement (1) is immediate by definition of p(μ), (2) follows from linearity of the integration operator, (4) follows from the monotonicity of the integration operator, while (3) follows from the Minkowski inequality (see [14, Theorem 3.5]):

(Ω|f+g|p𝑑μ)1p(Ω(|f|+|g|)p𝑑μ)1pMink.(Ω|f|p𝑑μ)1p+(Ω|g|p𝑑μ)1p.

The next lemma settles the easiest direction of the characterisation of operations that preserve p-integrability, i.e., the implication (2)  (1) in Theorem 2.1.

Lemma 4.4.

Let (Ω,F,μ) be a measure space, I a set, τ:RIR an operation that preserves measurability over (Ω,F) and p[1,+). If there exist a finite subset of indices JI and nonnegative real numbers (λj)jJ such that, for every vRI, we have |τ(v)|jJλj|vj|, then τ preserves p-integrability over (Ω,F,μ).

Proof.

Let (fi)iI be a family in p(μ); since τ preserves measurability over (Ω,), it follows that τ((fi)iI) is -measurable. For each xΩ, |τ((fi(x))iI)|jJλj|fj(x)|. Thus |τ((fi)iI)|jJλj|fj|. Hence, by Lemma 4.3, τ((fi)iI)p(μ). ∎

This shows that the condition |τ(v)|jJλj|vj| is sufficient for preservation of p-integrability. We are left to prove the converse direction: when τ does not satisfy this condition, there exists a measure space over which τ does not preserve p-integrability. As we shall see, at least when the arity of τ is countable, this space can always be taken to be (,,Leb) where Leb is the restriction to of the Lebesgue measure, and this happens because (,,Leb) is what we call a partitionable measure space.

Definition 4.5.

A measure space (Ω,,μ) is called partitionable if, for every sequence (an)nω of elements of +, there exists a sequence (An)nω of disjoint elements of such that μ(An)=an.

Remark 4.6.

The measure space (,,Leb) is partitionable.

The role of partitionable measure spaces is clarified by the following result.

Lemma 4.7.

Let (Ω,F,μ) be a measure space, let p[1,+), let I be a set and let τ:RIR be a function. Suppose |I||ω| and suppose (Ω,F,μ) is partitionable. If τ preserves p-integrability over (Ω,F,μ), then there exist a finite subset of indices JI and nonnegative real numbers (λj)jJ such that, for every vRI, we have

|τ(v)|jJλj|vj|.

Proof.

We give the proof for I=ω. The case |I|<|ω| relies on an analogous argument.

We suppose, contrapositively, that, for every finite subset of indices JI and every J-tuple (λj)jJ of nonnegative real numbers, there exists vI such that |τ(v)|>jJλj|vj|; we shall prove that τ does not preserve p-integrability. For each nω, we let vn be an element of I such that |τ(vn)|>j=0n-12np|vjn|. Set CΩnωAn. For each iω, we set

fi:Ω,
x{vinif xAn,0if xC.

Let (An)nω be a sequence of disjoint elements of such that μ(An)=1|τ(vn)|p; one such sequence exists because (Ω,,μ) is partitionable. Then

(4.1)Ω|τ((fi)iω)|p𝑑μ=C|τ((fi)iω)|p𝑑μ+nωAn|τ((fi)iω)|p𝑑μnω|τ((vin)iω)|pμ(An)=nω|τ(vn)|p1|τ(vn)|p=nω1=.

The following chain of inequalities holds:

(4.2)Ω|fi|p𝑑μ=nω|vin|pμ(An)=nω|vin|p1|τ(vn)|pM+n>i,vin0|vin|p1|τ(vn)|p(for some M+)M+n>i,vin0|vin|p1(j=0n-12np|vjn|)pM+n>i,vin0|vin|p1(2np|vin|)pM+n>i,vin012n<.

The first inequality holds for some M+ because with the condition n>i we ignore finitely many terms of the series, while with the condition vin0 we ignore some null terms. The third inequality holds because n>ii{0,,n-1}.

From equations (4.1) and (4.2) we conclude that τ does not preserve p-integrability. ∎

Lemma 4.8.

If I is a set, τ:RIR a function, p[1,+) and (Ω,F,μ) a partitionable measure space, then the following conditions are equivalent.

  1. τ preserves p-integrability.

  2. τ preserves measurability, and τ preserves p-integrability over (Ω,,μ).

  3. τ is Cyl(I)-measurable and there exist a finite subset of indices JI and nonnegative real numbers (λj)jJ such that, for every vI, we have |τ(v)|jJλj|vj|.

Proof.

(1)  (2) If τ preserves p-integrability, then, by Lemma 4.2, τ preserves measurability. Trivially, τ preserves p-integrability over (Ω,,μ).

(2)  (3) If τ preserves measurability, then, by Theorem 3.3, τ is Cyl(I)-measurable. By Proposition 3.5, τ depends on countably many coordinates, hence Lemma 4.7 applies and the proof of the implication is complete.

(3)  (1) By Theorem 3.3, τ preserves measurability. By Lemma 4.4, the thesis is proved. ∎

Proof of Theorem 2.1.

There exist partitionable measure spaces, see, e.g., Remark 4.6. Theorem 2.1 is the equivalence (1)  (3) in Lemma 4.8. ∎

4.1 Examples

Example 4.9.

Let nω and τ:n. Then τ preserves p-integrability if, and only if, τ is Borel measurable and there exist λ0,,λn-1+ such that, for every xn, we have

|τ(x)|j=0n-1λi|xi|.

Example 4.10.

A function τ:ω preserves p-integrability if, and only if, τ is Borel measurable and there exist a finite subset of indices Jω and nonnegative real numbers (λj)jJ and k such that, for every vI, we have

|τ(v)|k+jJλj|vj|.

4.2 The case of (,,Leb) and the discrete case

The remaining results in this section are not used in the proofs of our main results.

One may think that, for an operation τ:I, the condition “τ preserve p-integrability over every measure space” is too strong because we may not be interested in all measure spaces. However, Proposition 4.11 shows that this condition is equivalent to “τ preserve p-integrability over (,,Leb)” (if τ has countable arity), and Proposition 4.13 provides an analogous result for a discrete measure space.

Proposition 4.11.

Let I be a set, τ:RIR, with |I||ω|, and p[1,+). Then τ preserves p-integrability if, and only if, τ preserves p-integrability over (R,BR,Leb).

Proof.

Trivially, if τ preserves p-integrability, then τ preserves p-integrability over (,,Leb). For the converse, by Proposition 3.7, if τ preserves p-integrability over (,,Leb) then τ preserves measurability. By Remark 4.6, (,,Leb) is partitionable. An application of (2)  (1) in Lemma 4.8 concludes the proof. ∎

We next provide an analogue of Proposition 4.11 for a discrete measure space. We denote by 𝒫(X) the power set of a set X.

Lemma 4.12.

There exists a measure μ on (ω,P(ω)) such that (ω,P(ω),μ) is partitionable.

Proof.

We define a measure μ on (ω×,𝒫(ω×)), by setting μ({(n,z)})=2z. For every nω, there exists Kn such that an=zKn2z. Set An{(n,z)zKn}. Then μ(An)=zKnμ({(n,z)})=zKn2z=an. Moreover, for any pair of distinct n,mω, the sets An and Am are disjoint. The set ω× is countably infinite, hence (ω×,𝒫(ω×)) and (ω,𝒫(ω)) are isomorphic measurable spaces, which concludes the proof. ∎

Proposition 4.13.

There exists a measure μ on (ω,P(ω)) such that, for every set I, every function τ:RIR and every p[1,+), τ preserves p-integrability if, and only if, τ preserves measurability and τ preserves p-integrability over (ω,P(ω),μ).

Proof.

By Lemma 4.12, there exists a measure μ on (ω,𝒫(ω)) such that (ω,𝒫(ω),μ) is partitionable. The thesis follows from (1)  (2) in Lemma 4.8. ∎

5 Operations that preserve integrability over finite measure spaces

The goal of this section is to prove Theorem 2.2, i.e., to characterise the operations that preserve p-integrability over finite measure spaces. We follow the same strategy of Section 4, with the appropriate adjustments.

Lemma 5.1.

Let I be a set, τ:RIR and p[1,+). If τ preserves p-integrability over every finite measure space, then τ preserves measurability.

Proof.

By Remark 4.1. ∎

Lemma 5.2.

Let (Ω,F,μ) be a finite measure space, I a set, τ:RIR an operation that preserves measurability over (Ω,F) and p[1,+). If there exist a finite subset of indices JI and nonnegative real numbers (λj)jJ and k such that, for every vRI, we have |τ(v)|k+jJλj|vj|, then τ preserves p-integrability over (Ω,F,μ).

Proof.

Let (fi)iI be a family in p(μ); since τ preserves measurability over (Ω,), we have that τ((fi)iI) is -measurable. For each xΩ, |τ((fi(x))iI)|k+jJλj|fj(x)|. Thus |τ((fi)iI)|k+jJλj|fj|. Note that the function k:Ω,xk belongs to p(μ), because μ is finite. Hence, by Lemma 4.3, τ((fi)iI)p(μ). ∎

It is not difficult to see that no finite measure space is partitionable: thus we replace the concept of partitionability with a slightly different one.

Definition 5.3.

A measure space (Ω,,μ) is called conditionally partitionable if there exists a sequence (bn)nω of strictly positive real numbers such that, for every sequence (an)nω of elements of + satisfying anbn for every nω, there exists a sequence (An)nω of disjoint elements of such that μ(An)=an.

Remark 5.4.

The measure space ([0,1],[0,1],Leb), where Leb is the Lebesgue measure, is conditionally partitionable (take bn=12n+1).

Lemma 5.5.

Let (Ω,F,μ) be a measure space, let p[1,+), let I be a set and let τ:RIR be a function. Suppose that |I||ω| and that (Ω,F,μ) is conditionally partitionable. If τ preserves p-integrability over (Ω,F,μ), then there exist a finite subset of indices JI and nonnegative real numbers (λj)jJ and k such that, for every vRI, we have

|τ(v)|k+jJλj|vj|.

Proof.

We give the proof for I=ω. The case |I|<|ω| relies on an analogous argument.

We suppose, contrapositively, that, for every finite subset of indices JI, every J-tuple (λj)jJ of nonnegative real numbers and every k+, there exists vI such that |τ(v)|>k+jJλj|vj|; we shall prove that τ does not preserve p-integrability. Since (Ω,,μ) is conditionally partitionable, there exists a sequence (bn)nω of strictly positive real numbers such that, for every sequence (an)nω of elements of + satisfying anbn for every nω, there exists a sequence (An)nω of disjoint elements of such that μ(An)=an.

For each nω, we let vn be an element of I such that

|τ(vn)|>(1bn)1p+j=0n-12np|vjn|.

Then we have

1|τ(vn)|p<1((1bn)1p+j=0n-12np|vjn|)p1((1bn)1p)p=bn.

Therefore, there exists a sequence (An)nω of disjoint elements of such that μ(An)=1|τ(vn)|p. Since

|τ(vn)|>(1bn)1p+j=0n-12np|vjn|>j=0n-12np|vjn|,

the remaining part of the proof runs as for Lemma 4.8. ∎

Lemma 5.6.

Let I be a set, τ:RIR a function, p[1,+) and (Ω,F,μ) a conditionally partitionable finite measure space. The following conditions are equivalent.

  1. τ preserves p-integrability over every finite measure space.

  2. τ preserves measurability, and τ preserves p-integrability over (Ω,,μ).

  3. τ is Cyl(I)-measurable and there exist a finite subset of indices JI and nonnegative real numbers (λj)jJ and k such that, for every vI, we have |τ(v)|k+jJλj|vj|.

Proof.

(1)  (2) If τ preserves p-integrability over every finite measure space, then, by Lemma 5.1, τ preserves measurability. Trivially, τ preserves p-integrability over (Ω,,μ).

(2)  (3) If τ preserves measurability, then, by Theorem 3.3, τ is Cyl(I)-measurable. By Proposition 3.5, τ depends on countably many coordinates, hence Lemma 5.5 applies and the proof of the implication is complete.

(3)  (1) By Theorem 3.3, τ preserves measurability. By Lemma 5.2, the thesis is proved. ∎

Proof of Theorem 2.2.

There exist conditionally partitionable finite measure spaces, see, e.g., Remark 5.4. Theorem 2.2 is the equivalence (1)  (3) in Lemma 5.6. ∎

5.1 Examples

Example 5.7.

Let nω and τ:n. Then τ preserves p-integrability over every finite measure space if, and only if, τ is Borel measurable and there exist λ0,,λn-1,k+ such that, for every xn, we have

|τ(x)|k+j=0n-1λj|xj|.

Example 5.8.

A function τ:ω preserves p-integrability over every finite measure space if, and only if, τ is Borel measurable and there exist a finite subset of indices Jω and nonnegative real numbers (λj)jJ and k such that, for every vI, we have

|τ(v)|k+jJλj|vj|.

5.2 The case of ([0,1],[0,1],Leb) and the discrete case

The remaining results in this section are not used in the proofs of our main results.

One may think that, for an operation τ:I, the condition “τ preserve p-integrability over every finite measure space” is too strong because we may not be interested in all finite measure spaces. However, Proposition 5.9 shows that this condition is equivalent to “τ preserve p-integrability over ([0,1],[0,1],Leb)” (at least when τ has countable arity), and Proposition 5.11 provides an analogous result for a discrete finite measure space.

Proposition 5.9.

Let I be a set, τ:RIR, with |I||ω|, and p[1,+). Then τ preserves p-integrability over every finite measure space if, and only if, τ preserves p-integrability over ([0,1],B[0,1],Leb).

Proof.

Trivially, if τ preserves p-integrability, then τ preserves p-integrability over ([0,1],[0,1],Leb). For the converse, by Proposition 3.7 and Remark 3.8, if τ preserves p-integrability over ([0,1],[0,1],Leb), then τ preserves measurability. By Remark 5.4, ([0,1],[0,1],Leb) is conditionally partitionable. An application of (2)  (1) in Lemma 5.6 concludes the proof. ∎

Similarly to the case of arbitrary measure, we next provide an analogue of Proposition 5.9 for a discrete finite measure space.

Lemma 5.10.

There exists a probability measure μ on (ω,P(ω)) such that the measure space (ω,P(ω),μ) is conditionally partitionable.

Proof.

Let X{(n,m)ω×ωmn}. We let ν be the unique measure on (X,𝒫(X)) such that, for every (n,m)X, we have ν({(n,m)})=12m. Then,

(n,m)Xν({(n,m)})=nωmω,mnν({(n,m)})=nωmω,mn12m
=nω22n=4.

Hence, ν is a finite measure.

We prove that (X,𝒫(X),ν) is conditionally partitionable. For nω, let bn12n-1. Further, let (an)nω be a sequence of elements of + satisfying anbn for every nω. For every n, since 0an12n-1, there exists a subset Kn of {kωkn} such that an=kKn12k. Set An{(n,m)mKn}. Note that AnX. Then μ(An)=mKnμ({(n,m)})=mKn12m=an. Moreover, for any pair of distinct n,mω, the sets An and Am are disjoint. This proves that (X,𝒫(X),ν) is conditionally partitionable.

Define the measure ν4 on (X,𝒫(X)) by setting ν4(A)=ν(A)4. Using the fact that (X,𝒫(X),ν) is a conditionally partitionable measure space, it is not difficult to see that (X,𝒫(X),ν4) is a conditionally partitionable measure space, too. We have ν4(X)=ν(X)4=44; thus ν4 is a probability measure.

The set X is countably infinite, hence (X,𝒫(X)) and (ω,𝒫(ω)) are isomorphic measurable spaces, which concludes the proof. ∎

Proposition 5.11.

There exists a probability measure μ on (ω,P(ω)) such that, for every set I, every function τ:RIR and every p[1,+), τ preserves p-integrability over every finite measure space if, and only if, τ preserves measurability and τ preserves p-integrability over (ω,P(ω),μ).

Proof.

By Lemma 5.10, there exists a probability measure μ on (ω,𝒫(ω)) such that (ω,𝒫(ω),μ) is conditionally partitionable. The thesis follows from (1)  (2) in Lemma 5.6. ∎

6 Generation

The goal of this section is to prove Theorems 2.3 and 2.4, which exhibit a generating set for the class of operations that preserve integrability over arbitrary and finite measure spaces, respectively.

As it is shown by Theorems 2.1 and 2.2, the fact that an operation preserves p-integrability – over arbitrary and finite measure spaces, respectively – does not depend on the choice of p. Hence, we say that the operation preserves integrability.

Recall from the introduction the operation

(y,x0,x1,)supnω{xny}.

We adopt the notation

nωyxn(y,x0,x1,).

From the operations 0, +, and λ() (for each λ) we generate the operations

fg-((-f)(-g)),f+f0,f--(f0),|f|f+-f-.

Additionally, using , we generate

nωgfninfnω{fng}=-nω-g-fn.

Let Ω be a set and let SΩ. We let σ(S) denote the smallest σ-algebra of subsets of Ω such that every sS is -measurable.

Lemma 6.1.

Let Ω be a set and let SRΩ. Then σ(S) is the σ-algebra of subsets of Ω generated by the set {g-1((λ,+))gS,λR}.

Proof.

See [8, Proposition 2.3]. ∎

Lemma 6.2.

Let Ω be a set, let AP(Ω), let K be an element of the σ-algebra of subsets of Ω generated by A, and let KYΩ. Then K belongs to any σ-algebra G of subsets of Y such that AYG for each AA.

Proof.

Let Σ{SΩSY𝒢}. A straightforward verification shows that Σ is a σ-algebra of subsets of Ω. Moreover, 𝒜Σ. Therefore, by definition of , Σ. Hence, KΣ, which means K=KY𝒢. ∎

Given SΩ, we denote by S the closure of S under 0, +, , λ() (for each λ), and ¯. Given AΩ, we write 𝟙A for the characteristic function of A in Ω.

Lemma 6.3.

Let Ω be a set, let SRΩ, let Kσ(S) and let KYΩ be such that 1YS. Then 1KS.

Proof.

Set 𝒢{CY𝟙CS}. Note that 𝒢 is a σ-algebra of subsets of Y. Indeed, 𝟙YS, and, for C0,C1Y, we have 𝟙C0C1=𝟙C0𝟙C1 and 𝟙YC0=𝟙Y-𝟙C0. Further, let (Cn)nω be a family with CnY. The characteristic function of nωCn is nω𝟙Y𝟙Cn.

By Lemma 6.1, the σ-algebra σ(S) is generated by 𝒜{g-1((λ,+))gS,λ}. Let A𝒜, and write A=g-1((λ,+)) for some gS and some λ+. We have

(6.1)𝟙AYnω𝟙Yn(g-λ𝟙Y)+.

Indeed, for xAY, we have g(x)>λ and 𝟙Y(x)=1, hence

nω𝟙Y(x)n(g(x)-λ𝟙Y(x))+=nω1n(g(x)-λ> 0)+=1.

For xΩY, we have 𝟙Y(x)=0, and therefore

nω𝟙Y(x)n(g(x)-λ𝟙Y(x))+=nω0n(g(x))+=0.

For xYA, we have g(x)λ and 𝟙Y(x)=1, hence

nω𝟙Y(x)n(g(x)-λ𝟙Y(x))+=nω1n(g(x)-λ)+=nω10=0.

Given equation (6.1), we have 𝟙AYS, which means AY𝒢. By Lemma 6.2, K𝒢. ∎

The truncation operation ¯ comes into play in the following lemma.

Lemma 6.4.

Let λR+{0}. The operations

𝟙>λ:,x{1if x>λ,0otherwise,

and

𝟙λ:,x{1if xλ,0otherwise,

are generated by the operations 0, +, , λ() (for each λR), , ¯.

Proof.

Computation shows 𝟙f>1=nωf¯n(f-f¯). Moreover, 𝟙f>λ=𝟙1λf>1. Finally, let 0<q0<q1< be a sequence of elements of such that qnλ. Then 𝟙fλ=nω0𝟙f>qn. ∎

Lemma 6.5.

Let SRΩ, let gS, Aσ(S), λR+ be such that λ1Ag. Then λ1AS.

Proof.

We have 0S, hence the thesis is immediate for λ=0. Suppose λ>0. Then A{xΩg(x)λ}. By Lemma 6.4, 𝟙{xΩg(x)λ}=𝟙gλS. By Lemma 6.3, 𝟙AS, hence λ𝟙AS. ∎

Lemma 6.6.

Let SRΩ, let gS and let fRΩ be σ(S)-measurable and such that |f|g. Then fS.

Proof.

First, we prove the statement for f0. Given that f is positive and σ(S)-measurable, f is the supremum in Ω of a positive increasing sequence (sn)nω of σ(S)-measurable simple functions (see [14, Theorem 1.17]). By Lemma 6.5, snS for every nω. Hence

f=supnωsn=supnωsng=nωgsnS.

For f not necessarily positive, the previous part of the proof shows that f+ and f- belong to S. Then f=f+-f-S. ∎

Lemma 6.7.

Let (Ω,F) be a measurable space, and, for each nω, let fn:ΩR be a measurable function. If, for every xΩ, supnωfn(x)R, then supfn:ΩR is measurable. Analogously, if, for every xΩ, infnωfn(x)R, then the function infnωfn:ΩR is measurable.

Proof.

By [14, Theorem 1.14]. ∎

Lemma 6.8.

The operations 0,+,,λ() (for each λR), and ¯ preserve integrability.

Proof.

The operations 0,+,,λ() (for each λ) and ¯ preserve integrability. Moreover,

nωgfn=supnω{fng}

and therefore, by Lemma 6.7, preserves measurability. The constant function 0 is always integrable, therefore 0 preserves integrability. By (3) in Lemma 4.3, + preserves integrability. The operation || is immediately seen to preserve integrability. Since, for every f,g functions, |fg||f|+|g|, then preserves integrability by (4) in Lemma 4.3. We have nωgfn=supnω{fng}, and therefore f0gnωgfng. Hence, |nωgfn||g|+|f0|. Thus, preserves integrability. Finally, |f¯||f|, and therefore ¯ preserve integrability, by (4) in Lemma 4.3. ∎

Proof of Theorem 2.3.

The operations 0, +, , λ() (for each λ), and ¯ preserve integrability by Lemma 6.8. Moreover, by definition, the class of integrability-preserving operations is closed under every integrability-preserving operation and contains the projection functions. Therefore, every operation generated by 0, +, , λ() (for each λ), and ¯ preserves integrability.

To prove the converse, we use Theorem 2.1. Let J be a finite subset of I, and let (λj)jJ be a J-tuple of nonnegative real numbers. Then jJλj|πj|{πiiI}. Let τ be Cyl(I)-measurable and such that for every vI we have |τ(v)|jJλj|vj|, i.e., |τ|jJλj|πj|. Note that Cyl(I)=σ({πiiI}), by definition. Then τ{πiiI}, by Lemma 6.6. Therefore, τ is generated by 0, +, , λ() (for each λ), , ¯. ∎

It is worth recalling that, in the proof of Theorem 2.3, the role of the truncation operation ¯ lies in Lemma 6.4.

Proof of Theorem 2.4.

Note that the operations 0, +, , λ() (for each λ), and 1 preserve integrability over finite measure spaces. Moreover, by definition, the class of the operations that preserve integrability over finite measure spaces is closed under every integrability-preserving operation and contains the projection functions. Therefore, every operation generated by 0, +, , λ() (for each λ), and 1 preserves integrability over every finite measure space.

To prove the converse, we use Theorem 2.2. Note that the truncation is generated by , -1() (i.e., scalar multiplication by -1), and 1; indeed, f¯=f1=-((-f)(-1)). Let J be a finite subset of I, let (λj)jJ be a J-tuple of nonnegative real numbers, and let k+. Then k+jJλj|πj|{πiiI}{k}. Let τ be Cyl(I)-measurable and such that for every vI we have |τ(v)|k+jJλj|vj|, i.e., |τ|k+jJλj|πj|. Note that Cyl(I)=σ({πiiI})=σ({πiiI}{1}), by definition. Then we have τ{πiiI}{1}, by Lemma 6.6. Therefore, τ is generated by 0, +, , λ- (for each λ), , 1. ∎

7 The operation

We now investigate the operation , defined on in Section 6, for more general lattices. Given a Dedekind σ-complete (not necessarily bounded) lattice B we write for the operation on B of countably infinite arity defined as

(g,f0,f1,)supnω{fng}

We adopt the notation

nωgfn(g,f0,f1,).

Proposition 7.1.

If B is a Dedekind σ-complete lattice, then the following properties hold for every g,hB and all (fn)nωB.

  1. nωgfn=nωg(fng).

  2. nωgfn=(f0g)(nω{0}gfn).

  3. nωg(fnh)h.

Proof.

Straightforward verification. ∎

Conversely, we have the following.

Proposition 7.2.

If B is a lattice endowed with an operation of countably infinite arity which satisfies (TS1), (TS2) and (TS3), then B is Dedekind σ-complete and nωgfn=supnω{fng}.

Proof.

By induction on kω, (TS2) entails

nωgfn=(f0g)(fkg)(nk+1gfn).

Thusfkg(f0g)(fkg)(nk+1gfn)=nωgfn. Thus, nωgfn is an upper bound of (fkg)kω. Suppose now that fngh for every nω. Then

nωgfn=(TS1)nωg(fng)=fnghnωg(fngh)(TS3)h.

This shows nωgfn=supnω{fng}. To prove that B is Dedekind σ-complete, let (fn)nωB and gB be such that fng for all nω. Then

nωgfn=supnω{fng}=fngsupnωfn.

A map between two partially ordered sets is σ-continuous if it preserves all existing countable suprema.

Proposition 7.3.

Let φ:BC be a lattice morphism between two Dedekind σ-complete lattices. Then φ is σ-continuous if, and only if, φ preserves .

Proof.

First, suppose φ preserves . Let (fn)nωB and f=supnωfn. Then

φ(supnωfn)=φ(supnω{fnf})(because fnf)
=φ(nωffn)
=nωφ(f)φ(fn)(because φ preserves )
=supnω{φ(fn)φ(f)}
=supnωφ(fnf)(because φ preserves )
=supnωφ(fn)(because fnf).

Therefore, φ is σ-continuous.

For the converse implication, suppose that φ is σ-continuous. Let (fn)nωB and gB. Then

φ(nωgfn)=φ(supnω{fng})
=supnωφ(fng)(because φ preserves count. sups)
=supnω{φ(fn)φ(g)}(because φ preserves )
=nωφ(g)φ(fn).

Hence, φ preserves . ∎

Remark 7.4.

Propositions 7.1, 7.2 and 7.3 show that, whenever 𝒱 is a variety with a lattice reduct, then its subcategory of Dedekind σ-complete objects, with σ-continuous morphisms, is a variety which has, as primitive operations, the operations of 𝒱 together with , and, as axioms, the axioms of 𝒱 together with (TS1), (TS2) and (TS3).

8 Truncated -groups

We assume familiarity with the basic theory of -groups. All needed background can be found, for example, in the standard reference [3]. In [2], R. N. Ball defines a truncated -group as an abelian divisible -group that is endowed with a function ¯:G+G+, called truncation, which has the following properties for all f,gG+.

  1. fg¯f¯f.

  2. If f¯=0, then f=0.

  3. If nf=nf¯ for every nω, then f=0.

In this paper, we do not assume divisibility. The truncation ¯ may be extended to an operation on G, by setting f¯=f+¯-f-. Here, as is standard, we set f+f0, and f--(f0). Then Ball’s definition may be reformulated as follows.

Definition 8.1.

A truncated -group is an abelian -group that is endowed with a unary operation ¯:GG, called truncation, which has the following properties.

  1. For all fG, we have f¯=f+¯-f-.

  2. For all fG+, we have f¯G+.

  3. For all f,gG+, we have fg¯f¯f.

  4. For all fG+, if f¯=0, then f=0.

  5. For all fG+, if nf=nf¯ for every nω, then f=0.

Axiom (T2) ensures that ¯ may be restricted to an operation on G+. Axiom (T1) gives the one-to-one correspondence with Ball’s definition. Axioms (T3), (T4), (T5) correspond, respectively, to Axioms (B1), (B2), (B3). An -homomorphism φ between truncated -groups preserves ¯ if, and only if, φ preserves ¯ over positive elements; indeed, if φ preserves ¯ over positive elements, then, for fG,

φ(f¯)=φ(f+¯-f-)=φ(f+¯)-φ(f-)=φ(f+)¯-φ(f-)=φ(f)+¯-φ(f)-=φ(f)¯.

This ensures that the equivalence with Ball’s definition also holds for morphisms.

Note that (T1), (T2) and (T3) are (essentially) equational axioms. This is evident for (T1); (T2) can be written as ff+¯0=0; (T3) is the conjunction of the two equations f,gf+g+¯f+¯=f+¯ and ff+¯f+=f+. The axioms (T4) and (T5) cannot be expressed in such equational terms. However, as we shall see, this becomes possible when we add the hypothesis of Dedekind σ-completeness.

It is well known that a Dedekind σ-complete -group is archimedean and thus abelian. Let G be a Dedekind σ-complete -group, endowed with a unary operation ¯. We denote by (T4’) and (T5’) the following properties, which may or may not hold in G.

  1. For all fG+, we have f=nωfnf¯.

  2. For all fG+, we have f=nωf(nf-nf¯).

Note that (T4’) and (T5’), are (essentially) equational axioms: indeed, (T4’) is equivalent to ff+=nωf+nf+¯, and (T5’) is equivalent to ff+=nωf+(nf+-nf+¯).

Our aim in this section, met in Propositions 8.2, 8.5 and 8.8, is to show that, for a Dedekind σ-complete -group endowed with a unary operation ¯ which satisfies (T1), (T2) and (T3), the axioms (T4) and (T5) may be equivalently replaced by the equational axioms (T4’) and (T5’). This will show the axioms of Dedekind σ-complete truncated -groups to be equational.

Proposition 8.2.

Let G be an abelian -group endowed with a unary operation ¯. Then (T4’) implies (T4), and (T5’) implies (T5).

Proof.

Suppose (T4’). Let fG+ be such that f¯=0. By (T4’),

f=nωfnf¯=nωf0=0.

Hence, (T4) holds. Suppose (T5’). Let fG+ be such that nf=nf¯ for every nω. By (T5’),

f=nωf(nf-nf¯)=nωf0=0.

Hence (T5) holds. ∎

We shall use the following standard distributivity result.

Lemma 8.3.

Let G be an -group, I a set and (xi)iIG. If supiIxi exists, then, for every aG, supiI{axi} exists and

a(supiIxi)=supiI{axi}.

Proof.

See [3, Proposition 6.1.2]. ∎

Lemma 8.4.

Let G be a Dedekind σ-complete -group, let gG, hG+ and (fn)nωG. Then

nωg(fn+h)=((nωgfn)+h)g.

Proof.

We have

nωg(fn+h)=supnω{(fn+h)g}
=supnω{(fn+h)(g+h)g}(because h0)
=supnω{(fn+h)(g+h)}g(by Lemma 8.3)
=supnω{(fng)+h}g
=(supnω{fng}+h)g
=((nωgfn)+h)g.

Proposition 8.5.

Let G be a Dedekind σ-complete -group endowed with a unary operation ¯ such that (T2), (T3) and (T4) hold. Then (T4’) holds, i.e., for all fG+,

f=nωfnf¯.

Proof.

By (T2), f¯G+. Therefore 0f¯1f¯2f¯3f¯. Hence,

nωfnf¯=nω{0}fnf¯=nωf(n+1)f¯=nωf(nf¯+f¯)=((nωfnf¯)+f¯)f(by Lemma 8.4).

Therefore, setting bnωfnf¯, we have

0=((b+f¯)f)-b=f¯(f-b)=f-b¯,

where the last equality holds because, by (T3), we have f¯(f-b)f-b¯ and, for the opposite inequality, we have f-b¯f-b and f-b¯=f-b¯ff¯.

By (T4), since f-b¯=0, we have f-b=0, i.e., f=nωfnf¯. ∎

Lemma 8.6.

Let G be a Dedekind σ-complete -group endowed with a unary operation ¯ such that (T2) and (T3) holds. Let a,bG+. Then

a+b¯a¯+b¯.

Proof.

By (T3), a+b¯a+b. By (T2), a+b¯0, thus b(a+b)¯0, and therefore a+b¯a+b¯+(b(a+b)¯). Hence,

a+b¯[(a+b)(a+(a+b)¯)][(a+b)¯+(b(a+b)¯)]
=[a+(b(a+b)¯)][(a+b)¯+(b(a+b)¯)]
=(a(a+b)¯)+(b(a+b)¯)
a¯+b¯(by (T3)).

Lemma 8.7.

Let G be an abelian -group endowed with a unary operation ¯ such that (T3) holds. Then, for all a,bG+, if ab, then a-a¯b-b¯.

Proof.

Since ab, we have b¯-bb¯-a. By (T3), b¯-b0. Hence,

b¯-b(b¯-a)0
=(b¯a)-a(because + distributes over )
a¯-a(by (T3))

as desired. ∎

Proposition 8.8.

Let G be a Dedekind σ-complete -group endowed with a unary operation ¯ such that (T2), (T3) and (T5) hold. Then (T5’) holds, i.e., for all fG+,

f=nωf(nf-nf¯).

Proof.

Let kω. By (T3) we have 0kf-kf¯. We have

nωf(nf-nf¯)nω{0,,k-1}f(nf-nf¯)
=nωf((n+k)f-(n+k)f¯)
nωf(nf-nf¯+kf-kf¯)(by Lemma 8.6)
=((nωf(nf-nf¯))+kf-kf¯)f(by Lemma 8.4).

The opposite inequality is immediate. Therefore, setting bnωf(nf-nf¯), we have b=(b+kf-kf¯)f, which implies

0=((b+kf-kf¯)f)-b=(kf-kf¯)(f-b).

We set af-b. We have 0af, because 0bf. By (T3) and Lemma 8.7, 0ka-ka¯kf-kf¯. Therefore, 0=(ka-ka¯)a. It is elementary that, in any abelian group, xy=0 implies (nx)y=0 for each nω. Therefore,

0=(ka-ka¯)ka=(T2)(ka-ka¯).

Hence, ka=ka¯. Since k is arbitrary, by (T5) we infer a=0, i.e., f-nωf(nf-nf¯)=0. ∎

To sum up, Propositions 8.2, 8.5 and 8.8 show that, for Dedekind σ-complete -groups endowed with a unary operation ¯, Axioms (T1)-(T5) are equivalent to Axioms (T1)-(T3) together with Axioms (T4’) and (T5’).

We denote by σ𝔾t the category whose objects are Dedekind σ-complete truncated -groups, and whose morphisms are σ-continuous -homomorphisms that preserve ¯. Since Axioms (T1), (T2), (T3), (T4’) and (T5’) are equational, σ𝔾t is a variety, whose operations are the operations of -groups, together with ¯ and , and whose axioms are the axioms of -groups, together with the following ones.

  1. nωgfn=nωg(fng).

  2. nωgfn=(f0g)(nω{0}gfn).

  3. nωg(fnh)h.

  1. For all fG, we have f¯=f+¯-f-.

  2. For all fG+, we have f¯G+.

  3. For all f,gG+, we have fg¯f¯f.

  1. For all fG+, we have f=nωfnf¯.

  2. For all fG+, we have f=nωf(nf-nf¯).

9 The Loomis–Sikorski Theorem for truncated -groups

Definition 9.1.

Given a set X, a σ-ideal of subsets of X is a set of subsets of X such that the following conditions hold.

  1. .

  2. B,ABA.

  3. (An)nωnωAn.

If is a σ-ideal of subsets of X, we say that a property Pholds for -almost every xX if {xXP does not hold for x}. A σ-ideal of subsets of X induces on X an equivalence relation , defined by fg if, and only if, f(x)=g(x) for -almost every xX. We write X for the quotient X. Every operation τ of countable arity on induces an operation τ~ on X, by setting τ~(([fi])iI)[g], where g(x)=τ((fi(x))iI). The assumption that is closed under countable unions guarantees that this definition is well posed. Therefore, by Remark 7.4, X is a Dedekind σ-complete truncated -group.

The aim of this section is to prove the following theorem.

Theorem 9.2 (Loomis–Sikorski Theorem for truncated -groups).

Let G be a Dedekind σ-complete truncated -group. Then there exist a set X, a σ-ideal I of subsets of X and an injective σ-continuous -homomorphism ι:GRXI such that, for every fG, ι(f¯)=ι(f)[1]I.

We will give a proof that is rather self-contained, with the main exception of the use of Theorem 9.3 below. Anyway, we believe that a shorter (but not self-contained) way to prove Theorem 9.2 above (even in the less restrictive hypothesis that G is an archimedean truncated -group) may be the following. First, show that the divisible hull Gd of G admits a truncation that extends the truncation of G. Then embed Gd in X via [2, Theorem 5.3.6 (1)]. Finally, using arguments similar to those in [13, Theorem 6.2], show that this embedding preserves all countable suprema.

Theorem 9.3 (Loomis–Sikorski Theorem for Riesz spaces).

Let G be a Dedekind σ-complete Riesz space. Then there exist a set X, a σ-ideal I of subsets of X and an injective σ-continuous Riesz morphism ι:GRXI.

For a proof of Theorem 9.3 see [7], or [5] and [6].

Corollary 9.4 (Loomis–Sikorski Theorem for -groups).

Let G be a Dedekind σ-complete -group. Then there exist a set X, a σ-ideal I of subsets of X and an injective σ-continuous -homomorphism ι:GRXI.

Proof.

There exist a Dedekind σ-complete Riesz space H and an injective -morphism φ:GH that preserves every existing supremum; see [11]. Applying Theorem 9.3 to the Dedekind σ-complete Riesz space H, we obtain an injective σ-continuous Riesz morphism φ:HX. The composition ι=φφ:GX is an injective σ-continuous -morphism, since both φ and φ are injective σ-continuous -morphisms. ∎

Our strategy to prove Theorem 9.2 is the following. Lemma 9.12 will prove Theorem 9.2 for countably generated algebras. This will imply that generates the variety of Dedekind σ-complete truncated -groups, and from this fact Theorem 9.2 is derived.

Lemma 9.5.

Let G be a Dedekind σ-complete truncated -group generated by a subset SG. Then, for every gG, there exist s0,,sn-1S such that |g||s0|++|sn-1|.

Proof.

Let T{hGthere exist s0,,sn-1G:|h||s0|++|sn-1|}. It is clear that ST and standard that T is a convex -subgroup of G. Moreover, for every gG, and every (fn)nωG, the following hold.

  1. nωgfn=supnω{fng}, and therefore f0gnωgfng. Hence,

    |nωgfn|=(nωgfn)(-nωgfn)g[-(f0g)]g[(-f0)(-g)]|g||f0|.
  2. |g¯|=|g+¯-g-||g+¯|+|g-|=(T2)g+¯+g-(T3)g++g-=|g|.

Since T is a convex -subgroup of G, (1) and (2) imply that T is closed under and ¯. ∎

Lemma 9.6.

Let X be a set, and I a σ-ideal of subsets of X. Let (gn)nω be a sequence of functions from X to R. Suppose that, for I-almost every xX, supnωgn(x)R. Then the set {[gn]Inω} admits a supremum in RXI.

Proof.

Let A be such that, for every xXA, supnωgn(x). Let v:X be any function such that, for every xXA, v(x)=supnωgn(x). Then [v] is the supremum of {[gn]nω} in X. ∎

Lemma 9.7.

Let G be a Dedekind σ-complete truncated -group, let fG+ and let (fi)iωG+. Then

f=iωf(if-kωiffk¯).

Proof.

Trivially, fiωf(if-kωiffk¯). We prove the opposite inequality. By (T3), for every kω, we have fk¯(if)if¯, and therefore we have kωiffk¯=supiω{fk¯(if)}if¯. Hence, if-kωiffk¯if-if¯. Therefore, we have

iωf(if-kωiffk¯)iωf(if-if¯)=(T5’)f.

Lemma 9.8.

Let G be an abelian -group, let aG and let uG+. Then (a+u)-a-=au.

Proof.

We have (a+u)-a-=(a+-a-)(u-a-)=a(u+(a0))=a(u+a)u=au.

Lemma 9.9.

Let G be a countably generated Dedekind σ-complete truncated -group. Then there exist a set X, a σ-ideal I of subsets of X, an injective σ-continuous -homomorphism ι:GRXI and an element uRXI such that, for every fG, ι(f¯)=ι(f)u.

Proof.

By Corollary 9.4, there exist a set X, a σ-ideal of subsets of X and an injective σ-continuous -homomorphism ι:GX.

Let S be a countable generating set of G and let F{|s0|++|sn-1|s0,,sn-1S}. Let us enumerate F as F={f0,f1,f2,}. We shall prove that the set {ι(fn¯)nω}, admits a supremum uX that satisfies the statement of the lemma.

By Lemma 9.7, for each nω, we have

fn¯=iωfn¯(ifn¯-kωifn¯fk¯).

Since ι is a σ-continuous -homomorphism, using Proposition 7.3, we have the following.

  1. For each nω, ι(fn)=iωι(fn)(iι(fn)-kωiι(fn)ι(fk¯)).

For every nω, let gnX be such that [gn]=ι(fn¯). Then, by (1), for -almost every xX, the following conditions hold.

  1. For each nω, gn(x)=iωgn(x)(ign(x)-kωign(x)gk(x)).

Let x be such that (1’) hold. Suppose by way of contradiction that supnωgn(x)=. Then there exists nω such that gn(x)>0. Therefore, we have

gn(x)=iωgn(x)(ign(x)-kωign(x)gk(x))>0,

which implies that there exists iω such that ign(x)-kωign(x)gk(x)>0. Thus, kωign(x)gk(x)<ign(x). But supnωgn(x)= implies kωign(x)gk(x)=ign(x), a contradiction. Therefore, supnωgn(x) holds for each xX satisfying (1), and thus for -almost every xX. By Lemma 9.6, the set {[gn]nω}={ι(fn¯)nω} admits a supremum u.

Let fG+. Then

ι(f)u=ι(f)supnωι(fn¯)
=supnω{ι(f)ι(fn¯)}(by Lemma 8.3)
=supnω{ι(ffn¯)}
ι(f¯)(by (T3)).

For the opposite inequality, by Lemma 9.5 there exists mω such that f¯fm. Then f¯=f¯fm(T3)fm¯. Therefore ι(f¯)ι(fm¯)u, and moreover ι(f¯)ι(f) by (T3). Thus, ι(f¯)ι(f)u. For an arbitrary fG, f¯=f+¯-f- by (T1), hence ι(f¯)=ι(f+¯)-ι(f-)=(ι(f+)u)-ι(f-)=Lem. 9.8ι(f)u. ∎

Let G be a Dedekind σ-complete -group, let HG, and let uG. We say that u is a weak unit for H if u0 and, for every hH,

|h|=nω|h|n(|h|u).

Remark 9.10.

We will see in Lemma 11.2 that a weak unit for G in the sense above is the same as a weak unit of G in the usual sense.

Lemma 9.11.

Let Y be a set, J a σ-ideal of subsets of Y, HRYJ an -subgroup, and uRYJ a weak unit for H. Then there exists a set X, a σ-ideal I of subsets of X, and a σ-continuous -homomorphism ψ:RYJRXI such that the restriction of ψ to H is injective and ψ(u)=[1]I.

Proof.

Let vY be such that [v]𝒥=u. Since u0, we may choose v0. Let X{yYv(y)>0}. Let {JXJ𝒥}={J𝒥JX}. Let ()X:YX be the restriction map that sends fY to fXX, where fX(x)=f(x) for each xX. Write []𝒥:YY𝒥 for the natural quotient map, and similarly for []:XX. Since ker[]𝒥ker([]()X), by the universal property of the quotient there exists a unique σ-continuous -homomorphism ρ:Y𝒥X such that the following diagram commutes:

We claim that the restriction of ρ to H is injective. Indeed, let hH+ be such that ρ(h)=0. Let gY be such that [g]𝒥=h. Since h0, we may choose g0. We have that [gX]=0. Therefore, for -almost every xX, g(x)=0. Therefore, for 𝒥-almost every yY, g(y)=0 or yYX, i.e., g(y)=0 or v(y)=0. Since h=nωhn(hu), we have g(y)=nωg(y)n(g(y)v(y)) for 𝒥-almost every yY. Therefore, for 𝒥-almost every yY, if v(y)=0, then g(y)=nωg(y)n(g(y)0)=nωg(y)0=0, i.e., g(y)=0. Hence, for 𝒥-almost every yY, g(y)=0. Thus, h=0.

For every λ+{0}, the function λ(): which maps x to λx is an isomorphism of Dedekind σ-complete -groups. Indeed, its inverse is the map 1λ(). Then, the map m:XX which maps f to the function m(f) defined by (m(f))(x)=1v(x)f(x) is an isomorphism of Dedekind σ-complete -groups; indeed, its inverse is m-1:XX defined by (m-1(g))(x)=v(x)g(x). For every f,gX, [f]=[g] if, and only if, [m(f)]=[m(g)]. Hence, ker[]𝒥=ker([]𝒥m). Therefore, there exists an isomorphism η:XX of Dedekind σ-complete -groups which makes the following diagram commute:

We have the following commutative diagram:

We set ψηρ. Note that m(vX)X is the function constantly equal to 1: indeed, m(vX)(x)=1v(x)vX(x)=1. Thus, ψ(u)=η(ρ(u))=η(ρ([v]𝒥))=[m(vX)]=[1]. Since the restriction of ρ to H is injective, and η is bijective, the restriction of ψ to H is injective. ∎

Lemma 9.12.

Let G be a countably generated Dedekind σ-complete truncated -group. Then there exist a set X, a σ-ideal I of subsets of X and an injective σ-continuous -homomorphism ι:GRXI such that, for every fG, ι(f¯)=ι(f)[1]I.

Proof.

By Lemma 9.9, there exist a set Y, a σ-ideal 𝒥 of subsets of Y, an injective σ-continuous -homomorphism φ:GY𝒥 and an element uY𝒥 such that, for every fG,

φ(f¯)=φ(f)u.

First, 0φ(0¯)=0u, hence u0. Since, for all fG, |f|=nω|f|n|f|¯ by (T4’), we have

|φ(f)|=nω|φ(f)|n(|φ(f)|u).

Therefore, setting H equal to the image of G, u is a weak unit for H. By Lemma 9.11, there exist a set X, a σ-ideal of subsets of X, and a σ-continuous -homomorphism ψ:Y𝒥X such that the restriction of ψ to H is injective and ψ(u)=[1]. The function ιψφ has the required properties. ∎

Theorem 9.13.

The variety σGt of Dedekind σ-complete truncated -groups is generated by R.

Proof.

Let G be a Dedekind σ-complete truncated -group. Suppose that an equation τ=ρ (in the language of Dedekind σ-complete truncated -groups) does not hold in G. Since τ and ρ have countably many arguments, the equation τ=ρ does not hold in a countably generated Dedekind σ-complete truncated -group G. By Lemma 9.12, τ=ρ does not hold in . The statement follows by the HSP Theorem for (infinitary) varieties (see [16, Theorem (9.1)]). ∎

Proof of Theorem 9.2.

Since the variety of Dedekind σ-complete truncated -groups is generated by , there exists a set X, a σ𝔾t-subalgebra HX, and a surjective morphism ψ:HG of Dedekind σ-complete truncated -groups. Let

{AXthere exists (fn)nωkerψ such that for all aA there exists nω such that fn(a)0}.

Note that is a σ-ideal of subsets of X. Therefore we have the projection map XX which is a morphism of Dedekind σ-complete truncated -groups. If fkerψ, then f(x)=0 for -almost every xX. In other words, if fkerψ, then [f]=0. For the universal property of quotients, there exists a morphism ι:GRX of Dedekind σ-complete truncated -groups such that the following diagram commutes:

Let fH be such that ι(ψ(f))=[f]=0. Then there exists a set A such that f(x)=0 for every xXA. Since A, there exists a sequence (fn)nω of elements of kerψ such that, for every aA, there exists nω such that fn(a)0. Let us show

(9.1)|f|=n,kω|f|k|fn|.

Equation (9.1) holds if, and only if, for every aX, |f(a)|=n,kω|f(a)|k|fn(a)|. If aA, then both sides equal 0. If aA, then there exists mω such that fm(a)0, and therefore

n,kω|f(a)|k|fn(a)|kω|f(a)|k|fm(a)|=|f(a)|.

Since the opposite inequality is trivial, (9.1) is shown. By (9.1),

|ψ(f)|=n,kω|ψ(f)|k|ψ(fn)|=fnkerψn,kω|ψ(f)|0=0.

Therefore ψ(f)=0, and thus fkerψ. This implies that ι is injective. ∎

10 generates Dedekind σ-complete truncated Riesz spaces

Theorem 10.1 (Loomis–Sikorski Theorem for truncated Riesz spaces).

Let G be a Dedekind σ-complete truncated Riesz space. Then there exist a set X, a σ-ideal I of subsets of X, and an injective σ-continuous Riesz morphism ι:GRXI such that, for every fG, ι(f¯)=ι(f)[1]I.

Proof.

By Theorem 9.2, there exist a set X, a σ-ideal of subsets of X, and an injective σ-continuous -homomorphism ι:GX such that, for every fG, ι(f¯)=ι(f)[1]. Since X is Dedekind σ-complete, it is archimedean; by [15, Corollary 11.53], ι is a Riesz morphism. ∎

We denote by σ𝕊t the variety of Dedekind σ-complete truncated Riesz spaces, whose primitive operations are 0, +, , λ() (for each λ), , and ¯, and whose axioms are the axioms of Riesz spaces, together with (TS1), (TS2), (TS3), (T1), (T2), (T3), (T4’) and (T5’).

We can now obtain the first main result of Part II, as a consequence of Theorem 10.1.

Theorem 10.2.

The variety σRSt of Dedekind σ-complete truncated Riesz spaces is generated by R.

Proof.

Let G be a Dedekind σ-complete truncated Riesz space. By Theorem 10.1, there exist a set X, a σ-ideal of subsets of X, and an injective σ-continuous Riesz morphism ι:GX such that, for every fG, ι(f¯)=ι(f)[1]. Regarding X as an object of σ𝕊t with the structure induced from , we conclude that G is a subalgebra of a quotient of a power of . ∎

Remark 10.3.

From [1, Theorem 7.4], it follows that actually generates σ𝕊t as a quasi-variety, where quasi-equations are allowed to have countably many premises only.

Corollary 10.4.

For any set I,

Ft(I){f:I|f is Cyl(I)-measurable and there exist JI finite and (λj)jJ+:|f|jJλj|πj|}={f:If preserves integrability}

is the Dedekind σ-complete truncated Riesz space freely generated by the projections πi:RIR (iI).

Proof.

By Theorem 10.2, the variety σ𝕊t of Dedekind σ-complete truncated Riesz spaces is generated by . Therefore, by a standard result in general algebra, the smallest σ𝕊t-subalgebra S of I that contains the set of projection functions {πi:IiI} is freely generated by the projection functions. The set S is the smallest subset of I that contains, for each iI, the projection function πi:I, and which is closed under every primitive operation of σ𝕊t. By Theorem 2.4, S consists precisely of all operations I that preserve integrability. An application of Theorem 2.1 completes the proof. ∎

Write π:IFt(I) for the function π(i)=πi. Corollary 10.4 asserts the following. For any set I, for every Dedekind σ-complete truncated Riesz space G, for every function f:IG, there exists a unique σ-continuous truncation-preserving Riesz morphism φ:Ft(I)G such that the following diagram commutes:

11 The Loomis–Sikorski Theorem for -groups with weak unit

An element 1 of an abelian -group G is a weak unit if 10 and, for every fG, f1=0 implies f=0.

Remark 11.1.

Let G be an archimedean abelian -group, and let 1 be a weak unit. Then ff1 is a truncation. Indeed, the following show that (T1)(T5) hold.

  1. fu=(f+u)-f- by Lemma 9.8.

  2. For all fG+, f1G+.

  3. For all f,gG+, f(g1)=(f1)gf1f.

  4. For all fG+, if f1=0, then f=0.

  5. For all fG+, if nf=(nf)1 for every nω, then nf1 for every nω. Since G is archimedean, f=0.

Lemma 11.2.

Let G be a Dedekind σ-complete -group G, and let 1G. Then 1 is a weak unit if, and only if, the following conditions hold.

  1. 10.

  2. For all fG+, f=nωfn(f1).

Proof.

Since G is Dedekind σ-complete, G is archimedean. If 1 is a weak unit, then 10 and, by Remark 11.1 and Proposition 8.5, for all fG+, f=nωfn(f1). Conversely, suppose that (W1) and (W2) hold. If f1=0, then f=nωfn(f1)=nωf0=0, and so 1 is a weak unit. ∎

Note that, in the language of Dedekind σ-complete -groups, axioms (W1) and (W2) are equational. Indeed, (W1) corresponds to 10=0, and (W2) corresponds to ff+=nωf+n(f+1).

Theorem 11.3 (Loomis–Sikorski Theorem for -groups with weak unit).

Suppose G is a Dedekind σ-complete -group with weak unit 1. Then there exist a set X, a σ-ideal I of subsets of X, and an injective σ-continuous -homomorphism ι:GRXI such that ι(1)=[1]I.

Proof.

By Remark 11.1, G is a Dedekind σ-complete truncated -group, with the truncation given by ff1. Then, by Theorem 9.2, there exist a set Y, a σ-ideal 𝒥 of subsets of Y and an injective σ-continuous -homomorphism φ:GX such that, for every fG, φ(f1)=φ(f)[1]𝒥. The element φ(1) is a weak unit for the image of G under φ. Therefore, by Lemma 9.11, there exists a set X, a σ-ideal of subsets of X, and a σ-continuous -homomorphism ψ:Y𝒥X such that the restriction of ψ to H is injective and ψ(φ(1))=[1]. The function ιψφ has the desired properties. ∎

Corollary 11.4.

The variety of Dedekind σ-complete -groups with weak unit is generated by R.

Proof.

Let G be a Dedekind σ-complete -group with weak unit. By Theorem 11.3, G is a subalgebra of a quotient of a power of . ∎

12 generates Dedekind σ-complete Riesz spaces with weak unit

Theorem 12.1 (Loomis–Sikorski Theorem for Riesz spaces with weak unit).

Let G be a Dedekind σ-complete Riesz space with weak unit. Then there exist a set X, a σ-ideal I of subsets of X, and an injective σ-continuous Riesz morphism ι:GRXI such that ι(1)=[1]I.

Proof.

By Theorem 10.1, there exist a set X, a σ-ideal of subsets of X and an injective σ-continuous -homomorphism ι:GX such that, for every fG, ι(1)=[1]. Since X is Dedekind σ-complete, and thus archimedean, by [15, Corollary 11.53], ι is a Riesz morphism. ∎

We denote by σ𝕊u the variety of Dedekind σ-complete Riesz spaces with weak unit, whose primitive operations are 0, +, , λ() (for each λ), , and 1, and whose axioms are the axioms of Riesz spaces, together with (TS1), (TS2), (TS3), (W1), (W2).

As the second main result of Part II, we now deduce a theorem that was already obtained in [1].

Theorem 12.2.

The variety σRSu of Dedekind σ-complete Riesz spaces with weak unit is generated by R.

Proof.

Let G be a Dedekind σ-complete truncated Riesz space. By Theorem 12.1, G is a subalgebra of a quotient of a power of . ∎

Remark 12.3.

It has been shown in [1] that actually generates σ𝕊u as a quasi-variety, in the sense of Remark 10.3.

Corollary 12.4.

For any set I,

Fu(I){f:If is Cyl(I)-measurable and there exist JI finite, (λj)jJ+k+such that |f|k+jJλj|πj|}={f:If preserves integrability over finite measure spaces}

is the Dedekind σ-complete Riesz space with weak unit freely generated by the elements {πi}iI, where, for iI, πi:RIR is the projection on the i-th coordinate.

The proof is analogous to the proof of Corollary 10.4, and Fu(I) is characterised by a universal property analogous to the one that characterises Ft(I).


Communicated by Manfred Droste


A Operations that preserve -integrability

In Section 4 it has been shown that, for any p[1,+), a function τ:I preserves p-integrability if, and only if, τ is Cyl(I)-measurable and there exist a finite subset of indices JI and nonnegative real numbers (λj)jJ such that, for every vI, we have |τ(v)|jJλj|vj|. Does the same hold for p=? The answer is no. Indeed, the function ()2:,xx2 is an example of operation which preserves -integrability but not p-integrability, for every p[1,+). In Theorem A.5, we will answer the following question.

Question A.1.

Which operations I preserve -integrability?

We will see that an operation I preserve -integrability if, and only if, roughly speaking, it is measurable and it maps coordinatewise-bounded subsets of I onto bounded subsets of . To make this precise, we introduce some definitions.

Given a measure space (Ω,,μ), we define (μ) as the set of -measurable functions from Ω to that are bounded outside of a measurable set of null μ-measure.

Definition A.2.

Let I be a set, τ:I. We say that τ preserves -integrability if for every measure space (Ω,,μ) and every family (fi)iI(μ) we have τ((fi)iI)(μ).

We can now state the answer to Question A.1 precisely. Let I be a set and let τ:I be a function. Then τ preserves -integrability if, and only if, τ is Cyl(I)-measurable and, for every (Mi)iI+, the restriction of τ to iI[-Mi,Mi] is bounded. This will follow from Theorem A.5.

A.1 Operations that preserve boundedness

As a preliminary step, in Theorem A.4, we characterise the operations which preserve boundedness.

Definition A.3.

Let I be a set, τ:I. We say that τ preserves boundedness if for every set Ω and every family (fi)iI of bounded functions fi:Ω, we have that τ((fi)iI):Ω is also bounded.

Theorem A.4.

Let I be a set and τ:RIR. The following conditions are equivalent.

  1. τ preserves boundedness.

  2. For every (Mi)iI+, the restriction of τ to iI[-Mi,Mi] is bounded.

Proof.

We prove (1)  (2). Fix (Mi)iI+. Take ΩiI[-Mi,Mi] and, for every iI, let fi be the restriction of the projection function πi:I to Ω. Since fi maps Ω onto [-Mi,Mi], fi is bounded. Thus τ((fi)iI) is bounded, i.e., there exists M~ such that for every xΩ we have τ((fi(x))iI)[-M~,M~]. Let xΩ. Then τ(x)=τ((πi(x))iI)=τ((fi(x))iI)[-M~,M~]. Thus (2) holds.

We now prove (2)  (1). Let Ω be a set, and let (fi)iI be a family of bounded functions from Ω to . For each iI, let Mi+ be such that the image of fi is contained in [-Mi,Mi]. Let M~ be such that τ maps iI[-Mi,Mi] onto a subset of [-M~,M~]. Then, for each xΩ, τ((fi)iI)(x)=τ((fi(x))iI)[-M~,M~]. ∎

A.2 Operations that preserve -integrability

The following is the main theorem of this section.

Theorem A.5.

Let I be a set and let τ:RIR be a function. The following conditions are equivalent.

  1. τ preserves -integrability.

  2. τ preserves measurability and boundedness.

  3. τ is Cyl(I)-measurable and, for every (Mi)iI+, the restriction of τ to iI[-Mi,Mi] is bounded.

In order to prove Theorem A.5, we need some lemmas.

Lemma A.6.

Let I be a set and let τ:RIR be a function. If τ preserves -integrability, then τ preserves measurability.

Proof.

Every measurable space (Ω,) may be endowed with the null measure μ0: for each A, μ0(A)=0. Then (μ0) is the set of -measurable functions from Ω to . Hence, preservation of -integrability over (Ω,,μ0) is equivalent to preservation of measurability over (Ω,). ∎

Lemma A.7.

Let I be a set and let τ:RIR be a function. If τ preserves -integrability, then τ preserves boundedness.

Proof.

Let us suppose that τ does not preserve boundedness. By Theorem A.4, there exists (Mi)iI+ such that the restriction of τ to iI[-Mi,Mi] is not bounded. Fix one such family (Mi)iI; let ΩiI[-Mi,Mi]. Let (ωn)nω be a sequence in Ω such that |τ(ω0)|<|τ(ω1)|< and |τ(ωn)| as n. Consider on (Ω,𝒫(Ω)) the discrete measure μ such that μ({ωn})=12n for every nω and μ(Ω{ω0,ω1,})=0. Then (Ω,𝒫(Ω),μ) is a finite measure space. For iI, the restriction (πi)Ω of πi to Ω is bounded, since its image is [-Mi,Mi]. Moreover, (πi)Ω is 𝒫(Ω)-measurable. Therefore, (πi)Ω(μ). We have τΩ(μ); indeed, let A be a subset of Ω of null μ-measure. Then ωnA for every nω. Therefore τΩ is not bounded outside of A. ∎

Lemma A.8.

Let I be a set and let τ:RIR be a function. If τ preserves measurability and boundedness, then τ preserves -integrability.

Proof.

By Corollary 3.6, τ depends on a countable subset JI. Let (Ω,,μ) be a finite measure space and consider a family (fi)iI(μ). For each jJ, let Aj be a measurable subset of Ω such that μ(Aj)=0 and fj is bounded outside of Aj. Set AjJAj. Then μ(A)=0. For each iI, define fi~ as fi if iJ, otherwise let fi~ be the function constantly equal to 0. Since τ depends only on J, we have τ((fi)iI)=τ((fi~)iI). For every iI, the restriction (fi~)ΩA is bounded. We have that τ((fi)iI)ΩA=τ((fi~)iI)ΩA=τ(((fi~)ΩA)iI) is bounded since τ preserves boundedness and, for every iI, (fi~)ΩA is bounded. Thus τ((fi)iI) is bounded outside of a set of null measure. Moreover, τ((fi)iI) is measurable because τ preserve measurability. Therefore τ((fi)iI)(μ). ∎

Proof of Theorem A.5.

By Lemmas A.6 and A.7, we have (1)  (2). Lemma A.8, we have (2)  (1). By Theorems 3.3 and A.4, we have (2)  (3). ∎

Corollary A.9.

Let I be a set and let τ:RIR be a function. If τ preserves p-integrability for some p[1,+), then τ preserves -integrability.

Proof.

By Theorem 2.1, τ is Cyl(I)-measurable and there exist a finite subset of indices JI and nonnegative real numbers (λj)jJ such that, for every vI, we have |τ(v)|jJλj|vj|. Let (Mi)iI+. Let viI[-Mi,Mi]. Then |τ(v)|jJλj|vj|jJλjMj. Thus, the restriction of τ to iI[-Mi,Mi] is bounded. Therefore, by Theorem A.5, τ preserves -integrability. ∎

Remark A.10.

The converse of Corollary A.9, as mentioned at the beginning of this section, is not true, as shown by the function ()2:,xx2.

Acknowledgements

The author is deeply grateful to his Ph.D. advisor Professor Vincenzo Marra for the many helpful discussions. Moreover, the author gratefully acknowledges the anonymous referee for his or her valuable suggestions.

References

[1] M. Abbadini, Dedekind σ-complete -groups and Riesz spaces as varieties, Positivity (2019), 10.1007/s11117-019-00718-9. 10.1007/s11117-019-00718-9Search in Google Scholar

[2] R. N. Ball, Truncated abelian lattice-ordered groups I: The pointed (Yosida) representation, Topology Appl. 162 (2014), 43–65. 10.1016/j.topol.2013.11.007Search in Google Scholar

[3] A. Bigard, K. Keimel and S. Wolfenstein, Groupes et anneaux réticulés, Lecture Notes in Math. 608, Springer, Berlin, 1977. 10.1007/BFb0067004Search in Google Scholar

[4] S. Burris and H. P. Sankappanavar, A Course in Universal Algebra, Grad. Texts in Math. 78, Springer, New York, 1981. 10.1007/978-1-4613-8130-3Search in Google Scholar

[5] G. Buskes, B. de Pagter and A. van Rooij, The Loomis–Sikorski theorem revisited, Algebra Universalis 58 (2008), no. 4, 413–426. 10.1007/s00012-008-2077-xSearch in Google Scholar

[6] G. Buskes and A. van Rooij, Small Riesz spaces, Math. Proc. Cambridge Philos. Soc. 105 (1989), no. 3, 523–536. 10.1017/S0305004100077902Search in Google Scholar

[7] G. Buskes and A. Van Rooij, Representation of Riesz spaces without the Axiom of Choice, Nepali Math. Sci. Rep. 16 (1997), no. 1–2, 19–22. Search in Google Scholar

[8] G. B. Folland, Real Analysis, 2nd ed., Pure Appl. Math. (New York), John Wiley & Sons, New York, 1999. Search in Google Scholar

[9] D. H. Fremlin, Measure Theory. Vol. 2, Torres Fremlin, Colchester, 2003. Search in Google Scholar

[10] O. Kallenberg, Foundations of Modern Probability, Probab. Appl. (N. Y.), Springer, New York, 1997. Search in Google Scholar

[11] M. A. Lepellere and A. Valente, Embedding of Archimedean l-groups in Riesz spaces, Atti Semin. Mat. Fis. Univ. Modena 46 (1998), no. 1, 249–254. Search in Google Scholar

[12] W. A. J. Luxemburg and A. C. Zaanen, Riesz Spaces. Vol. I, North-Holland, Amsterdam, 1971. Search in Google Scholar

[13] D. Mundici, Tensor products and the Loomis–Sikorski theorem for MV-algebras, Adv. in Appl. Math. 22 (1999), no. 2, 227–248. 10.1006/aama.1998.0631Search in Google Scholar

[14] W. Rudin, Real and Complex Analysis, 3rd ed., McGraw-Hill, New York, 1987. Search in Google Scholar

[15] E. Schechter, Handbook of Analysis and its Foundations, Academic Press, San Diego, 1999. Search in Google Scholar

[16] J. Słomiński, The theory of abstract algebras with infinitary operations, Instytut Matematyczny Polskiej Akademi Nauk, Wrosław, 1959. Search in Google Scholar

[17] S. M. Srivastava, A Course on Borel Sets, Grad. Texts in Math. 180, Springer, New York, 1998. 10.1007/978-3-642-85473-6Search in Google Scholar

Received: 2018-10-09
Revised: 2020-06-17
Published Online: 2020-07-16
Published in Print: 2020-11-01

© 2021 Walter de Gruyter GmbH, Berlin/Boston

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

Downloaded on 26.4.2024 from https://www.degruyter.com/document/doi/10.1515/forum-2018-0244/html
Scroll to top button