This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Measurable Functions and Topological Algebra

Geoff Vooys
Abstract.

In this paper we show that if (X,𝒜)(X,\mathcal{A}) is a measurable space and if YY is a topological model of a Lawvere theory 𝒯\mathcal{T} equipped with \mathcal{B} the Borel σ\sigma-algebra on YY, then the set of \mathcal{B}-measurable functions from XX to YY, Meas(X,Y)\operatorname{Meas}(X,Y), is a set-theoretic model of 𝒯\mathcal{T}. As a corollary we give short proofs of the facts that the set of real-valued measurable functions on a measurable space XX is a ring and the set of complex-valued measurable functions from XX to \mathbb{C} is a ring.

Key words and phrases:
Measurable Spaces, Borel Algebra, Lawvere Theories, Topological Algebra, Algebraic Theories
1991 Mathematics Subject Classification:
Primary 28E15; Secondary 18C10, 18C40

1. Introduction

Without much exaggeration, measure theory is as important to modern analysis as the theory of modules is to modern algebra. It is a fundamental tool in functional analysis, statistics, representation theory, harmonic analysis, geometric group theory, and even in number theory. In fact any time a question can be phrased in terms of volumes it probably has a formulation or approach using measure-theoretic tools. However, while measure theory itself is mostly applied in analytic contexts due to its relationship with integration, there is a very set-theoretic algebraic flavour in the foundations and many of the basic results for measure theory. Despite this, it is often an arduous journey for many algebraically-minded graduate students to work through many of the basic measure-theoretic results needed for representation theory and number theory.

In this paper, however, we show that this difficulty need not be the case. We prove a categorical-algebraic result that gives, if the mathematician writing the proof is willing to be brief and lean on some “apply Theorem xyzxyz to 𝒯𝐑𝐢𝐧𝐠\operatorname{\mathcal{T}}_{\operatorname{\mathbf{Ring}}} with the observation…” in their writing, a one-sentence proof of the fact that for any measurable set (X,𝒜)(X,\operatorname{\mathcal{A}}) the set Meas(X,;Borel)\operatorname{Meas}(X,\operatorname{\operatorname{\mathbb{R}}};\text{Borel}) of Borel measurable functions is a ring. The trade-off here is that we use some categorical algebraic concepts and prove a much stronger statement using Lawvere theories in order to avoid the analytic perspective and provide a structural proof indicating exactly what is used to prove that Meas(X,;Borel)\operatorname{Meas}(X,\operatorname{\operatorname{\mathbb{R}}};\text{Borel}) is a ring. However, this gives an ε\varepsilon-free method for proving algebraic properties of sets of measurable functions with values in \operatorname{\operatorname{\mathbb{R}}}.

1.1. The Structure of the Paper

In Section 2 we give a short review of some of the basics of measurable spaces. Because there are many good textbook accounts of measure theory in the literature, we focus instead on when our measurable space (X,𝒜)(X,\operatorname{\mathcal{A}}) is the Borel algebra of a topological space XX. Section 3 gives a quick review of (infinitary) Lawvere theories following the construction of [3]. The focus we give is more on definitions and examples than in proving and providing general theorems. Finally in Section 4 we prove the main theorem of our paper, presented below, and then the following two corollaries.

Theorem 1.1.

Let 𝒯\operatorname{\mathcal{T}} be a Lawvere theory and let YY be a topological model of 𝒯\operatorname{\mathcal{T}}, i.e., Y𝒯(𝐓𝐨𝐩)0Y\in\operatorname{\mathcal{T}}(\operatorname{\mathbf{Top}})_{0}. Then if (X,𝒜)(X,\operatorname{\mathcal{A}}) is a measurable space, the set

Meas(X,Y;X)={f:XY|fisY-measurable}\operatorname{Meas}(X,Y;\operatorname{\mathcal{B}}_{X})=\{f:X\to Y\;|\;f\,\text{is}\,\operatorname{\mathcal{B}}_{Y}\textnormal{-measurable}\}

is an object in 𝒯(𝐒𝐞𝐭)0\operatorname{\mathcal{T}}(\operatorname{\mathbf{Set}})_{0}.

Corollary 1.2.

Let RR be a topological ring. Then for any measurable space (X,𝒜)(X,\operatorname{\mathcal{A}}) the set Meas(X,R;R)\operatorname{Meas}(X,R;\operatorname{\mathcal{B}}_{R}) is a ring. In particular, the real-measurable functions Meas(X,;)\operatorname{Meas}(X,\operatorname{\operatorname{\mathbb{R}}};\operatorname{\mathcal{B}}_{\operatorname{\operatorname{\mathbb{R}}}}) and complex-measurable functions Meas(X,;)\operatorname{Meas}(X,\operatorname{\operatorname{\mathbb{C}}};\operatorname{\mathcal{B}}_{\operatorname{\operatorname{\mathbb{C}}}}) are rings.

Corollary 1.3.

Let GG be a topological group and let (X,𝒜)(X,\operatorname{\mathcal{A}}) be any measurable space. Then Meas(X,G;G)\operatorname{Meas}(X,G;\operatorname{\mathcal{B}}_{G}) is a group. In particular, for any local field FF the set Meas(X,GLn(F);GLn(F))\operatorname{Meas}(X,\operatorname{GL}_{n}(F);\operatorname{\mathcal{B}}_{\operatorname{GL}_{n}(F)}) is a group.

1.2. Acknowledgments

This work was supported by an AARMS postdoctoral fellowship.

2. A Review of Measure Theory for Topological Spaces

In this short section we review some of the basics of measure theory regarding measurable spaces and σ\sigma-algebras. Most of this section is simply review of what is explained in more detail in many textbooks in the literature (cf. [4], [5], [8], [9], for instance), but we present it primarily because, in my opinion, Example 2.5 and the interaction between uncountably infinite topological spaces and measure theory are not as well-known as they likely should be.

Definition 2.1.

Let XX be a set and let 𝒫X\operatorname{\mathcal{P}}X denote the power set

𝒫X:={S|SX}\operatorname{\mathcal{P}}X:=\{S\;|\;S\subseteq X\}

of XX. Then a σ\sigma-algebra on XX is a subset 𝒜𝒫X\operatorname{\mathcal{A}}\subseteq\operatorname{\mathcal{P}}X for which:

  • We have that X𝒜X\in\operatorname{\mathcal{A}}.

  • If A𝒜A\in\operatorname{\mathcal{A}}, then Ac:=XA𝒜A^{c}:=X\setminus A\in\operatorname{\mathcal{A}};

  • If II is a countable index set and if {Ai|iI}𝒜\{A_{i}\;|\;i\in I\}\subseteq\operatorname{\mathcal{A}} then

    iIAi𝒜.\bigcup_{i\in I}A_{i}\in\operatorname{\mathcal{A}}.
Definition 2.2.

A measurable space is a pair (X,𝒜)(X,\operatorname{\mathcal{A}}) where XX is a set and 𝒜\operatorname{\mathcal{A}} is a σ\sigma-algebra on XX.

Of course the definition of a measurable space is geometrically motivated; it allows us to axiomitize which pieces of a set XX can have measured volume and gives us inference rules for what we can say about how these sets must behave. The first axiom says that if we can measure the volume of a piece AXA\subseteq X then we can measure the volume of the complement of AA, while the second rule says that if we can measure the volume of countably many pieces AiXA_{i}\subseteq X then we can measure the volume of the compositum iIAi\cup_{i\in I}A_{i}. Another important way to describe the geometry of a set XX is through a topology on XX, which we now describe.

Definition 2.3.

A topological space is a pair (X,𝒯)(X,\operatorname{\mathcal{T}}) where XX is a set and 𝒯𝒫X\operatorname{\mathcal{T}}\subseteq\operatorname{\mathcal{P}}X is a set of subsets we declare to be open such that:

  • ,X𝒯\varnothing,X\in\operatorname{\mathcal{T}};

  • For any set II, if {Ui|iI,Ui𝒯}𝒯\{U_{i}\;|\;i\in I,U_{i}\in\operatorname{\mathcal{T}}\}\subseteq\operatorname{\mathcal{T}} then

    iIUi𝒯.\bigcup_{i\in I}U_{i}\in\operatorname{\mathcal{T}}.
  • For any finite collection U1,,Un𝒯U_{1},\cdots,U_{n}\in\operatorname{\mathcal{T}}, there is a U𝒯U\in\operatorname{\mathcal{T}} for which

    i=1nUic=Uc.\bigcap_{i=1}^{n}U_{i}^{c}=U^{c}.

An important consequence of the definition of a σ\sigma-algebra is that such a set is closed with respect to countable intersections of its members. This allows us to deduce the following proposition which tells us that σ\sigma-algebras are topologies on countable sets.

Proposition 2.4.

Let (X,𝒜)(X,\operatorname{\mathcal{A}}) be a countable measure space. Then 𝒜\operatorname{\mathcal{A}} is a topology on XX.

With this in mind it is tempting to think that σ\sigma-algebras give rise to topologies on sets, but sadly it is relatively well-known that this need not be the case. However, as this is not as well-known as it should be, we present the example below to illustrate this fact.

Example 2.5.

Let XX be an uncountably infinite set and let 𝒜𝒫(X)\operatorname{\mathcal{A}}\subseteq\operatorname{\mathcal{P}}(X) be the σ\sigma-algebra of countable or cocountable subsets of XX, i.e.,

𝒜:={YX:|Y|0or|Yc|0}.\operatorname{\mathcal{A}}:=\{Y\subseteq X\;:\;\lvert Y\rvert\leq\aleph_{0}\,\text{or}\,\lvert Y^{c}\rvert\leq\aleph_{0}\}.

It is routine and tedious to check that 𝒜\operatorname{\mathcal{A}} is a σ\sigma-algebra on XX. To see that it does not generate a topology on XX note that since the set {x}X\{x\}\subseteq X for each xXx\in X, 𝒜\operatorname{\mathcal{A}} contains every singleton point of XX and so must be the discrete topology if it is a topology. However, as 𝒜\operatorname{\mathcal{A}} contains no uncountable set with uncountable complement this is impossible and so 𝒜\operatorname{\mathcal{A}} cannot be a topology on XX.

3. A Review of Lawvere Theories

In this section we give a short review of Lawvere theories, their basic notions, and some examples. Lawvere theories are generalizations and formalizations of algberiac theories (as they appear in universal algbera and logic; cf. ) and give us a language to discuss and the infinitary Lawvere theories we allow in this paper even provide a way to discuss algebraic objects that have infinitely many operations. We follow [3, Appendix A], which contains an expository account of Lawvere theories in detail, and formalize the notion of Lawvere theories as certain limit sketches; other equivalent formalizations are given in [2], [6], and [1], among other sources. Of course, it is an important theorem of Linton (cf. [7] for the usual reference and [3] for a self-contained explicit and direct proof) that there is an equivalence of categories

𝐋𝐚𝐰𝐯𝐞𝐫𝐞𝐌𝐨𝐧𝐚𝐝(𝐒𝐞𝐭).\operatorname{\mathbf{Lawvere}}\simeq\operatorname{\mathbf{Monad}}(\operatorname{\mathbf{Set}}).

While we will not need more than a basic definition of Lawvere theories in this paper, we present some examples in this short section so that we can firmly ground our intuition in the sort of examples which are likely to be most familiar.

Definition 3.1.

An infintary Lawvere theory is a pair (,X)(\operatorname{\mathcal{L}},X) where \operatorname{\mathcal{L}} is a locally small category with all 𝐒𝐞𝐭\operatorname{\mathbf{Set}}-indexed products and XX is a distinguished object of \operatorname{\mathcal{L}} such that for all objects YY of \operatorname{\mathcal{L}} there exists a set II and an isomorphism

YiIX=XIY\cong\prod_{i\in I}X=X^{I}

in \operatorname{\mathcal{L}}.

Remark 3.2.

Because of the nature of a Lawvere theory (,X)(\operatorname{\mathcal{L}},X) described above, to describe functors F:𝒞F:\operatorname{\mathcal{L}}\to\operatorname{\mathscr{C}} it suffices to define FF on the objects XIX^{I} for all sets II and on all morphisms between all XIX^{I} and XJX^{J}.

Definition 3.3.

We say that a Lawvere theory (,X)(\operatorname{\mathcal{L}},X) is finitary if for any index sets I,JI,J we have that

(XI,XJ)(colimEIEfinite(XE,X))I.\operatorname{\mathcal{L}}(X^{I},X^{J})\cong\left(\operatorname*{colim}_{\begin{subarray}{c}E\subseteq I\\ E\,\text{finite}\end{subarray}}\operatorname{\mathcal{L}}(X^{E},X)\right)^{I}.
Example 3.4.

The (finitary) Lawvere theory of groups is the category 𝒯𝐆𝐫𝐩\operatorname{\mathcal{T}}_{\operatorname{\mathbf{Grp}}} where we define the obejcts to be given by:

  • For every index set there is an object GIG^{I}. The distinguished object is G:=G[0]G:=G^{[0]} where [0]={0}[0]=\{0\}. We also write :=G\top:=G^{\varnothing}.

We define the morphisms to be generated by making sure that each GIG^{I} is a product indexed by II together with some extra morphisms

  • 1G:G1_{G}:\top\to G

  • inv:GG\operatorname{inv}:G\to G

  • μ:G×GG\mu:G\times G\to G

which satisfy the diagrams

G{G}G×{G\times\top}G×G{G\times G}×G{\top\times G}G×G{G\times G}G{G}\scriptstyle{\cong}\scriptstyle{\cong}idG×1G\scriptstyle{\operatorname{id}_{G}\times 1_{G}}μ\scriptstyle{\mu}1G×idG\scriptstyle{1_{G}\times\operatorname{id}_{G}}μ\scriptstyle{\mu}
(G×G)×G{(G\times G)\times G}G×G{G\times G}G×(G×G){G\times(G\times G)}G{G}G×G{G\times G}μ×idG\scriptstyle{\mu\times\operatorname{id}_{G}}\scriptstyle{\cong}μ\scriptstyle{\mu}idG×μ\scriptstyle{\operatorname{id}_{G}\times\mu}μ\scriptstyle{\mu}

and:

G{G}G×G{G\times G}G×G{G\times G}{\top}G{G}Δ\scriptstyle{\Delta}!G\scriptstyle{!_{G}}inv×idG\scriptstyle{\operatorname{inv}\times\operatorname{id}_{G}}idG×inv\scriptstyle{\operatorname{id}_{G}\times\operatorname{inv}}μ\scriptstyle{\mu}1G\scriptstyle{1_{G}}
Example 3.5.

The (finitary) Lawvere theory of (associative unital) rings is the category 𝒯𝐑𝐢𝐧𝐠\operatorname{\mathcal{T}}_{\operatorname{\mathbf{Ring}}} where we define the obejcts to be given by:

  • For every index set there is an object GIG^{I}. The distinguished object is R:=R[0]R:=R^{[0]} where [0]={0}[0]=\{0\}. We also write :=R\top:=R^{\varnothing}.

We define the morphisms to be generated by making sure that each RIR^{I} is a product indexed by II together with some extra morphisms

  • 0R:R0_{R}:\top\to R

  • 1R:R1_{R}:\top\to R

  • neg:RR\operatorname{neg}:R\to R

  • α:R×RR\alpha:R\times R\to R

  • μ:R×RR\mu:R\times R\to R

which satisfy the diagrams

R{R}R×{R\times\top}R×R{R\times R}×R{\top\times R}R×R{R\times R}R{R}\scriptstyle{\cong}\scriptstyle{\cong}idR×1R\scriptstyle{\operatorname{id}_{R}\times 1_{R}}μ\scriptstyle{\mu}1R×idR\scriptstyle{1_{R}\times\operatorname{id}_{R}}μ\scriptstyle{\mu}
(R×R)×R{(R\times R)\times R}R×R{R\times R}R×(R×R){R\times(R\times R)}R{R}R×R{R\times R}μ×idR\scriptstyle{\mu\times\operatorname{id}_{R}}\scriptstyle{\cong}μ\scriptstyle{\mu}idR×μ\scriptstyle{\operatorname{id}_{R}\times\mu}μ\scriptstyle{\mu}
R{R}R×{R\times\top}R×R{R\times R}×R{\top\times R}R×R{R\times R}R{R}\scriptstyle{\cong}\scriptstyle{\cong}idR×0R\scriptstyle{\operatorname{id}_{R}\times 0_{R}}α\scriptstyle{\alpha}0R×idR\scriptstyle{0_{R}\times\operatorname{id}_{R}}α\scriptstyle{\alpha}
(R×R)×R{(R\times R)\times R}R×R{R\times R}R×(R×R){R\times(R\times R)}R{R}R×R{R\times R}α×idR\scriptstyle{\alpha\times\operatorname{id}_{R}}\scriptstyle{\cong}α\scriptstyle{\alpha}idR×α\scriptstyle{\operatorname{id}_{R}\times\alpha}α\scriptstyle{\alpha}
R{R}R×R{R\times R}R×R{R\times R}{\top}R{R}Δ\scriptstyle{\Delta}!G\scriptstyle{!_{G}}neg×idR\scriptstyle{\operatorname{neg}\times\operatorname{id}_{R}}idR×neg\scriptstyle{\operatorname{id}_{R}\times\operatorname{neg}}α\scriptstyle{\alpha}0R\scriptstyle{0_{R}}

and diagrams expressing left and right distributivity of multiplication and addition:

R×(R×R){R\times(R\times R)}R×R{R\times R}(R×R)×(R×R){(R\times R)\times(R\times R)}(R×R)×(R×R){(R\times R)\times(R\times R)}R×R{R\times R}R{R}R×(R×R)×R{R\times(R\times R)\times R}R×(R×R)×R{R\times(R\times R)\times R}Δ\scriptstyle{\Delta}idR×α\scriptstyle{\operatorname{id}_{R}\times\alpha}μ\scriptstyle{\mu}\scriptstyle{\cong}μ×μ\scriptstyle{\mu\times\mu}α\scriptstyle{\alpha}idR×switch×idR\scriptstyle{\operatorname{id}_{R}\times\text{switch}\times\operatorname{id}_{R}}\scriptstyle{\cong}
(R×R)×R{(R\times R)\times R}R×R{R\times R}(R×R)×(R×R){(R\times R)\times(R\times R)}(R×R)×(R×R){(R\times R)\times(R\times R)}R×R{R\times R}R{R}R×(R×R)×R{R\times(R\times R)\times R}R×(R×R)×R{R\times(R\times R)\times R}Δ\scriptstyle{\Delta}α×idR\scriptstyle{\alpha\times\operatorname{id}_{R}}μ\scriptstyle{\mu}\scriptstyle{\cong}μ×μ\scriptstyle{\mu\times\mu}α\scriptstyle{\alpha}idR×switch×idR\scriptstyle{\operatorname{id}_{R}\times\text{switch}\times\operatorname{id}_{R}}\scriptstyle{\cong}
Definition 3.6.

Let 𝒞\operatorname{\mathscr{C}} be a category and let (,X)(\operatorname{\mathcal{L}},X) be a Lawvere theory. A 𝒞\operatorname{\mathscr{C}}-model of \operatorname{\mathcal{L}} is a product-preserving functor M:𝒞M:\operatorname{\mathcal{L}}\to\operatorname{\mathscr{C}}. In this case we call M(X)M(X) the underlying object of the model. Moreover, the category of all such models is the category

(𝒞):=[,𝒞]prod\operatorname{\mathcal{L}}(\operatorname{\mathscr{C}}):=[\operatorname{\mathcal{L}},\operatorname{\mathscr{C}}]_{\text{prod}}

of product-preserving functors 𝒞\operatorname{\mathcal{L}}\to\operatorname{\mathscr{C}} and their natural transformations.

Remark 3.7.

If (,X)(\operatorname{\mathcal{L}},X) is a Lawvere theory then:

  • A 𝐒𝐞𝐭\operatorname{\mathbf{Set}}-valued model of \operatorname{\mathcal{L}} is often called “an \operatorname{\mathcal{L}}” by abuse of notation (just as we call models of 𝒯𝐆𝐫𝐩\operatorname{\mathcal{T}}_{\operatorname{\mathbf{Grp}}} in 𝐒𝐞𝐭\operatorname{\mathbf{Set}} groups, 𝒯𝐑𝐢𝐧𝐠\operatorname{\mathcal{T}}_{\operatorname{\mathbf{Ring}}} in 𝐒𝐞𝐭\operatorname{\mathbf{Set}} rings, etc.). There are equivalences of categories 𝒯𝐆𝐫𝐩(𝐒𝐞𝐭)𝐆𝐫𝐩\operatorname{\mathcal{T}}_{\operatorname{\mathbf{Grp}}}(\operatorname{\mathbf{Set}})\simeq\operatorname{\mathbf{Grp}} and 𝒯𝐑𝐢𝐧𝐠(𝐒𝐞𝐭)𝐑𝐢𝐧𝐠\operatorname{\mathcal{T}}_{\operatorname{\mathbf{Ring}}}(\operatorname{\mathbf{Set}})\simeq\operatorname{\mathbf{Ring}} which follow immediately after unwinding the definitions of everything in sight.

  • A 𝐓𝐨𝐩\operatorname{\mathbf{Top}}-valued model of \operatorname{\mathcal{L}} is a topological model of \operatorname{\mathcal{L}}. Classical examples of this include topological groups (models of 𝒯𝐆𝐫𝐩\operatorname{\mathcal{T}}_{\operatorname{\mathbf{Grp}}}) and topological rings (models of 𝒯𝐑𝐢𝐧𝐠\operatorname{\mathcal{T}}_{\operatorname{\mathbf{Ring}}}). However, topological fields are not examples of 𝐓𝐨𝐩\operatorname{\mathbf{Top}}-valued models of a Lawvere theory because the theory of fields is not a Lawvere theory.

  • If KK is a field and if 𝒞=𝐕𝐚𝐫/SpecK\operatorname{\mathscr{C}}=\operatorname{\mathbf{Var}}_{/\operatorname{Spec}K} is the category of KK-varieties then (𝐕𝐚𝐫/SpecK)\operatorname{\mathcal{L}}(\operatorname{\mathbf{Var}}_{/\operatorname{Spec}K}) is the category of KK-variety models of \operatorname{\mathcal{L}}. When =𝒯𝐆𝐫𝐩\operatorname{\mathcal{L}}=\operatorname{\mathcal{T}}_{\operatorname{\mathbf{Grp}}}, (𝐕𝐚𝐫/SpecK)=𝒯𝐆𝐫𝐩(𝐕𝐚𝐫/SpecK)\operatorname{\mathcal{L}}(\operatorname{\mathbf{Var}}_{/\operatorname{Spec}K})=\operatorname{\mathcal{T}}_{\operatorname{\mathbf{Grp}}}(\operatorname{\mathbf{Var}}_{/\operatorname{Spec}K}) is the category of algebraic groups over KK (which is a category whose objects occupy an area of central importance in the Langlands Programme and in algebraic geometry).

4. Putting it All Together: Borel Measures

We now discuss some topological measure theory before proceeding to prove our main theorem. This material is relatively well-known, so we will mostly focus on giving a quick description in order to get to show how topological models of Lawvere theories and Borel algebras interact.

Definition 4.1.

Let XX be a topological space. We define the Borel algebra on XX, X\operatorname{\mathcal{B}}_{X}, to be the σ\sigma-algebra generated by the open subsets UXU\subseteq X.

Definition 4.2.

Let XX be a topological space. We say that a σ\sigma-algebra 𝒜\operatorname{\mathcal{A}} on XX is a Borel algebra if for any open set UXU\subseteq X, U𝒜U\in\operatorname{\mathcal{A}}.

From the definition it is immediate that if XX is a topological space then we can describe the Borel algebra X\operatorname{\mathcal{B}}_{X} by

X:=𝒜is aσ-algebra𝒜is a Borel algebra𝒜.\operatorname{\mathcal{B}}_{X}:=\bigcap_{\begin{subarray}{c}\operatorname{\mathcal{A}}\,\text{is a}\,\sigma\text{-algebra}\\ \operatorname{\mathcal{A}}\,\text{is a Borel algebra}\end{subarray}}\operatorname{\mathcal{A}}.

This follows more or less immediately from the description of X\operatorname{\mathcal{B}}_{X} being generated by the open sets of XX.

Proposition 4.3.

Let XX be a topological space. The Borel algebra X\operatorname{\mathcal{B}}_{X} on XX admits the description

X=inf{𝒜|𝒜is aσ-algebra,U𝒜,UXopen}\operatorname{\mathcal{B}}_{X}=\inf\left\{\operatorname{\mathcal{A}}\;|\;\operatorname{\mathcal{A}}\,\textnormal{is a}\;\sigma\textnormal{-algebra},U\in\operatorname{\mathcal{A}},U\subseteq X\,\textnormal{open}\right\}

in the power set 𝒫(X)\operatorname{\mathscr{P}}(X) regarded as a Boolean algebra with respect to the subset order \subseteq.

With this terminology we can now develop the theory of Borel measurable functions in order to derive our main result. These are the measurable functions f:XYf:X\to Y where YY is a topological space which interact in a compatible way with the topology on YY.

Definition 4.4.

Let YY be a topological space and let (X,𝒜)(X,\operatorname{\mathcal{A}}) be a measurable space. We then say that f:XYf:X\to Y is Borel measurable (or Y\operatorname{\mathcal{B}}_{Y}-measurable if we need to emphasize the space YY) if and only if for all BYB\in\operatorname{\mathcal{B}}_{Y} we have f1(B)𝒜f^{-1}(B)\in\operatorname{\mathcal{A}}.

This leads to a standard but helpful set-theoretic lemma.

Lemma 4.5.

If (X,𝒜)(X,\operatorname{\mathcal{A}}) is a measurable space and if f:XYf:X\to Y is a function of sets then

:={BY|f1(B)𝒜}\operatorname{\mathcal{B}}:=\left\{B\subseteq Y\;|\;f^{-1}(B)\in\operatorname{\mathcal{A}}\right\}

is a σ\sigma-algebra on YY.

Proof.

We proceed with the observations that for any arbitrary family {Bi|iI,BiY}\{B_{i}\;|\;i\in I,B_{i}\subseteq Y\} of subsets,

f1(iIBi)={xX|iI.f(x)BiY}=iIf1(Bi)f^{-1}\left(\bigcup_{i\in I}B_{i}\right)=\{x\in X\;|\;\exists\,i\in I.\,f(x)\in B_{i}\subseteq Y\}=\bigcup_{i\in I}f^{-1}(B_{i})

and dually

f1(iIBi)=iIf1(Bi).f^{-1}\left(\bigcap_{i\in I}B_{i}\right)=\bigcap_{i\in I}f^{-1}(B_{i}).

Thus if we have any countable collection of sets BnYB_{n}\subseteq Y for which f1(Bn)𝒜f^{-1}(B_{n})\in\operatorname{\mathcal{A}} for all nn\in\operatorname{\operatorname{\mathbb{N}}} we get that

f1(nBn)=nf1(Bn)𝒜f^{-1}\left(\bigcup_{n\in\operatorname{\operatorname{\mathbb{N}}}}B_{n}\right)=\bigcup_{n\in\operatorname{\operatorname{\mathbb{N}}}}f^{-1}(B_{n})\in\operatorname{\mathcal{A}}

and

f1(nBn)=nf1(Bn)𝒜f^{-1}\left(\bigcap_{n\in\operatorname{\operatorname{\mathbb{N}}}}B_{n}\right)=\bigcap_{n\in\operatorname{\operatorname{\mathbb{N}}}}f^{-1}\left(B_{n}\right)\in\operatorname{\mathcal{A}}

because 𝒜\operatorname{\mathcal{A}} is a σ\sigma-algebra and each of f1(Bn)𝒜f^{-1}(B_{n})\in\operatorname{\mathcal{A}}.

To verify that \operatorname{\mathcal{B}} is closed under complements, note that f1(Y)=Xf^{-1}(Y)=X and f1(ST)=f1(S)f1(T)f^{-1}(S\setminus T)=f^{-1}(S)\setminus f^{-1}(T) for any S,TYS,T\subseteq Y. Then if BB\in\operatorname{\mathcal{B}} we get that, because 𝒜\operatorname{\mathcal{A}} is a σ\sigma-algebra,

f1(Bc)=f1(YB)=f1(Y)f1(B)=Xf1(B)=f1(B)c𝒜.f^{-1}(B^{c})=f^{-1}(Y\setminus B)=f^{-1}(Y)\setminus f^{-1}(B)=X\setminus f^{-1}(B)=f^{-1}(B)^{c}\in\operatorname{\mathcal{A}}.

Thus f1(B)cf^{-1}(B)^{c}\in\operatorname{\mathcal{B}} and so \operatorname{\mathcal{B}} is a σ\sigma-algebra. ∎

We now can establish some basic properties about the Borel measurability of continuous functions and how these functions post-compose with other Borel measurable functions.

Lemma 4.6.

Let XX and YY be topological spaces and regard XX as a measurable space by equipping it with its Borel measure X\operatorname{\mathcal{B}}_{X}. If f:XYf:X\to Y is a continuous function of topological spaces then ff is Borel measurable.

Proof.

Since ff is continuous, for every open set VYV\subseteq Y we have that f1(V)Xf^{-1}(V)\subseteq X is open and so f1(V)Xf^{-1}(V)\in\operatorname{\mathcal{B}}_{X} as well. However, this then implies that every open VYV\subseteq Y open we get that

f1(V){UY|f1(U)X}=:𝒞.f^{-1}(V)\in\{U\subseteq Y\;|\;f^{-1}(U)\in\operatorname{\mathcal{B}}_{X}\}=:\operatorname{\mathcal{C}}.

Because 𝒞\operatorname{\mathcal{C}} is a σ\sigma-algebra by Lemma 4.5 which contains every open set of YY, we see that Y𝒞\operatorname{\mathcal{B}}_{Y}\subseteq\operatorname{\mathcal{C}} by Proposition 4.3. However, this implies that for any AXA\in\operatorname{\mathcal{B}}_{X} we have f1(A)Xf^{-1}(A)\in\operatorname{\mathcal{B}}_{X} and so ff is Borel measurable. ∎

Proposition 4.7.

Let (X,𝒜)(X,\operatorname{\mathcal{A}}) be a measurable space and let YY and ZZ be topological spaces equipped with their corresponding Borel σ\sigma-algebras Y\operatorname{\mathcal{B}}_{Y} and Z\operatorname{\mathcal{B}}_{Z}. Then if f:XYf:X\to Y is Y\operatorname{\mathcal{B}}_{Y}-measurable and if g:YZg:Y\to Z is Z\operatorname{\mathcal{B}}_{Z}-measurable, the composite gf:XZg\circ f:X\to Z is Z\operatorname{\mathcal{B}}_{Z}-measurable.

Proof.

Let VZV\subseteq Z with ZZZ\in\operatorname{\mathcal{B}}_{Z}; it now suffices to show that

(gf)1(Z)=f1(g1(V))(g\circ f)^{-1}(Z)=f^{-1}(g^{-1}(V))

is contained in 𝒜\operatorname{\mathcal{A}}. However, as gg is Z\operatorname{\mathcal{B}}_{Z}-measurable, g1(V)Yg^{-1}(V)\in\operatorname{\mathcal{B}}_{Y}. Similarly, as ff is Y\operatorname{\mathcal{B}}_{Y}-measurable, f1(g1(V))𝒜f^{-1}(g^{-1}(V))\in\operatorname{\mathcal{A}}. Thus

f1(g1(V))=(gf)1(V)𝒜f^{-1}\left(g^{-1}(V)\right)=(g\circ f)^{-1}(V)\in\operatorname{\mathcal{A}}

and so gfg\circ f is Z\operatorname{\mathcal{B}}_{Z}-measurable. ∎

Corollary 4.8.

Let (X,𝒜)(X,\operatorname{\mathcal{A}}) be a measurable space and let YY and ZZ be topological spaces equipped with their corresponding Borel algebras Y\operatorname{\mathcal{B}}_{Y} and Z\operatorname{\mathcal{B}}_{Z}. Then if g:YZg:Y\to Z is a continuous function the composite function gf:XZg\circ f:X\to Z is Z\operatorname{\mathcal{B}}_{Z}-measurable.

Proof.

Apply Lemma 4.6 to derive that gg is Borel-measurable and then use Proposition 4.7. ∎

We now have two last technical ingredients to provide. These show the Borel measurabliity of situations we’ll find ourselves in when studying the measurability of topological models of a Lawvere theory \operatorname{\mathcal{L}}.

Lemma 4.9.

Let (X,𝒜)(X,\operatorname{\mathcal{A}}) be a measurable space, let II be an index set, and let

{fi:XY|iI}\left\{f_{i}:X\to Y\;|\;i\in I\right\}

be a collection of functions which are all Y\operatorname{\mathcal{B}}_{Y}-measurable. If F:XZIF:X\to Z^{I} is the map given by

f(x):=(fi(x))iIf(x):=\big{(}f_{i}(x)\big{)}_{i\in I}

then ff is YI\operatorname{\mathcal{B}}_{Y^{I}}-measurable.

Proof.

Because YIY^{I} is the categorical product of YY taken II-many times and carries the product topology, YIY^{I} has basis of open sets

:={iIUi:UiYopen,UiYfor all but finitely manyi}.\operatorname{\mathscr{B}}:=\left\{\prod_{i\in I}U_{i}\;:\;U_{i}\subseteq Y\,\text{open},U_{i}\neq Y\,\text{for all but finitely many}\,i\right\}.

Similarly, because the Borel σ\sigma-algebra YI\operatorname{\mathcal{B}}_{Y^{I}} is generated by the opens of YIY^{I} and the open sets of YIY^{I} are generated by those in \operatorname{\mathscr{B}} it suffices to prove that for any basic open U=iIUiU=\prod_{i\in I}U_{i} in \operatorname{\mathscr{B}}, f1(V)𝒜f^{-1}(V)\in\operatorname{\mathcal{A}}. However, we now note that

f1(iIUi)\displaystyle f^{-1}\left(\prod_{i\in I}U_{i}\right) ={xX:(fi(x))iIiIUi}={xX|iI.fi(x)Ui}\displaystyle=\left\{x\in X\;:\;(f_{i}(x))_{i\in I}\in\prod_{i\in I}U_{i}\right\}=\left\{x\in X\;|\;\forall\,i\in I.\,f_{i}(x)\in U_{i}\right\}
=iIfi1(Ui).\displaystyle=\bigcap_{i\in I}f_{i}^{-1}(U_{i}).

Because UU is a basic open in YIY^{I}, at most finitely many of the UiU_{i} satisfy UiYU_{i}\neq Y. But then because fi1(Y)=Xf_{i}^{-1}(Y)=X, the intersection above is a finite intersection

f1(iIUi)=UiYfi1(Ui);f^{-1}\left(\prod_{i\in I}U_{i}\right)=\bigcap_{U_{i}\neq Y}f_{i}^{-1}(U_{i});

moreover since each UiU_{i} is open and each fif_{i} is Y\operatorname{\mathcal{B}}_{Y}-measurable, fi1(Ui)𝒜f_{i}^{-1}(U_{i})\in\operatorname{\mathcal{A}}. But then, as finite intersections are countable, it follows that

f1(iIUi)=UiYfi1(Ui)𝒜f^{-1}\left(\prod_{i\in I}U_{i}\right)=\bigcap_{U_{i}\neq Y}f_{i}^{-1}(U_{i})\in\operatorname{\mathcal{A}}

and so ff is Borel measurable. ∎

Remark 4.10.

The lemma above says in categorical terms that if we have a collection of Y\operatorname{\mathcal{B}}_{Y}-measurable morphisms

{fi:XY|iI}\left\{f_{i}:X\to Y\;|\;i\in I\right\}

then the pairing map

fiiI:XYI\langle f_{i}\rangle_{i\in I}:X\to Y^{I}

is YI\operatorname{\mathcal{B}}_{Y^{I}}-measurable.

Proposition 4.11.

Let (X,𝒜)(X,\operatorname{\mathcal{A}}) be a measurable space, let YY be a topological space, let II be an index set. If

{fi:XY|iI}\{f_{i}:X\to Y\;|\;i\in I\}

is a collection of Y\operatorname{\mathcal{B}}_{Y}-indexed functions and g:YIYg:Y^{I}\to Y is a Y\operatorname{\mathcal{B}}_{Y}-measurable function, then the function h:XYh:X\to Y given by h=gfiiIh=g\circ\langle f_{i}\rangle_{i\in I} is Y\operatorname{\mathcal{B}}_{Y}-measurable.

Proof.

Simply apply Lemma 4.9 and Proposition 4.7 to the composition h=gfiiIh=g\circ\langle f_{i}\rangle_{i\in I}. ∎

Combining these tools together allows us to prove our main theorem: the collection of Borel-measurable functions from a measurable space (X,𝒜)(X,\operatorname{\mathcal{A}}) into a topological model of a Lawvere theory \operatorname{\mathcal{L}} gives rise to a set-theoretic model of \operatorname{\mathcal{L}}.

Theorem 4.12.

Let 𝒯\operatorname{\mathcal{T}} be a Lawvere theory and let YY be a topological model of 𝒯\operatorname{\mathcal{T}}, i.e., Y𝒯(𝐓𝐨𝐩)0Y\in\operatorname{\mathcal{T}}(\operatorname{\mathbf{Top}})_{0}. Then if (X,𝒜)(X,\operatorname{\mathcal{A}}) is a measurable space, the set

Meas(X,Y;X)={f:XY|fisY-measurable}\operatorname{Meas}(X,Y;\operatorname{\mathcal{B}}_{X})=\{f:X\to Y\;|\;f\,\text{is}\,\operatorname{\mathcal{B}}_{Y}\textnormal{-measurable}\}

is an object in 𝒯(𝐒𝐞𝐭)0\operatorname{\mathcal{T}}(\operatorname{\mathbf{Set}})_{0}.

Proof.

Because YY is a topological model of 𝒯\operatorname{\mathcal{T}}, we write

Y¯:𝒯𝐓𝐨𝐩\underline{Y}:\operatorname{\mathcal{T}}\to\operatorname{\mathbf{Top}}

for the corresponding functor and specify that Y¯(D)=Y\underline{Y}(D)=Y for DD the distinguished object of 𝒯\operatorname{\mathcal{T}}. To show that Meas(X,Y;Y)\operatorname{Meas}(X,Y;\operatorname{\mathcal{B}}_{Y}) is a model of 𝒯\operatorname{\mathcal{T}} in 𝐒𝐞𝐭\operatorname{\mathbf{Set}}, we define

Meas(X,Y;Y)¯:𝒯𝐒𝐞𝐭\underline{\operatorname{Meas}(X,Y;\operatorname{\mathcal{B}}_{Y})}:\operatorname{\mathcal{T}}\to\operatorname{\mathbf{Set}}

as follows:

  • For objects DID^{I} for index sets II, we define

    Meas(X,Y;Y)¯(DI):=Meas(X,YI;YI).\underline{\operatorname{Meas}(X,Y;\operatorname{\mathcal{B}}_{Y})}(D^{I}):=\operatorname{Meas}(X,Y^{I};\operatorname{\mathcal{B}}_{Y^{I}}).
  • For morphisms ϕ:DIDJ\phi:D^{I}\to D^{J} in 𝒯\operatorname{\mathcal{T}}, we define

    Meas(X,Y;Y)¯(DI)Meas(X,Y;Y)¯(DJ)\underline{\operatorname{Meas}(X,Y;\operatorname{\mathcal{B}}_{Y})}(D^{I})\to\underline{\operatorname{Meas}(X,Y;\operatorname{\mathcal{B}}_{Y})}(D^{J})

    to be the function

    Meas(X,YI;YI)Meas(X,YJ;YJ)\operatorname{Meas}(X,Y^{I};\operatorname{\mathcal{B}}_{Y^{I}})\to\operatorname{Meas}(X,Y^{J};\operatorname{\mathcal{B}}_{Y^{J}})

    given by fY¯(φ)ff\mapsto\underline{Y}(\varphi)\circ f.

That this is a functor is a straightforward check from definition and Remark 3.2, while the fact that the assignment on morphisms is well-defined follows from Proposition 4.11. To see that it is product preserving, let II be a set and assume that for all iIi\in I there is a map fi:ZMeas(X,Y;Y)f_{i}:Z\to\operatorname{Meas}(X,Y;\operatorname{\mathcal{B}}_{Y}). Consider the functions pi:Meas(X,YI;YI)Meas(X,Y;Y)p_{i}:\operatorname{Meas}(X,Y^{I};\operatorname{\mathcal{B}}_{Y^{I}})\to\operatorname{Meas}(X,Y;\operatorname{\mathcal{B}}_{Y}) given by post-composition by the projection maps πi:YIY\pi_{i}:Y^{I}\to Y; these are the maps we must show play the role of projections. We define a map f:ZMeas(X,YI;YI)f:Z\to\operatorname{Meas}(X,Y^{I};\operatorname{\mathcal{B}}_{Y^{I}}) by, for each zZz\in Z, letting f(z)f(z) be the function XYIX\to Y^{I} given by the pairing fi(z)iI\langle f_{i}(z)\rangle_{i\in I}. It then follows by construction that for all iIi\in I

pif=πifiiI=fip_{i}\circ f=\pi_{i}\circ\langle f_{i}\rangle_{i\in I}=f_{i}

so the diagram

Z{Z}Meas(X,YI;YI){\operatorname{Meas}(X,Y^{I};\operatorname{\mathcal{B}}_{Y^{I}})}Meas(X,Y;Y){\operatorname{Meas}(X,Y;\operatorname{\mathcal{B}}_{Y})}f\scriptstyle{f}fi\scriptstyle{f_{i}}pi\scriptstyle{p_{i}}

commutes. The fact that ff is unique follows from the fact that each pairing map fi(z)iI\langle f_{i}(z)\rangle_{i\in I} is induced by the universal property of the (Cartesian) product. Thus Meas(X,YI;YI)Meas(X,Y;Y)I\operatorname{Meas}(X,Y^{I};\operatorname{\mathcal{B}}_{Y^{I}})\cong\operatorname{Meas}(X,Y;\operatorname{\mathcal{B}}_{Y})^{I} and so the functor Meas(X,Y;Y)¯:𝒯𝐒𝐞𝐭\underline{\operatorname{Meas}(X,Y;\operatorname{\mathcal{B}}_{Y})}:\operatorname{\mathcal{T}}\to\operatorname{\mathbf{Set}} is product-preserving. Consequently Meas(X,Y;Y)\operatorname{Meas}(X,Y;\operatorname{\mathcal{B}}_{Y}) is a set-theoretic model of the Lawvere theory 𝒯\operatorname{\mathcal{T}}. ∎

The usefulness of this theorem is exemplified by following straightforward algebraic proofs of the corollaries below (which are in particular of use in functional analysis, representation theory, harmonic analysis, and number theory).

Corollary 4.13.

Let RR be a topological ring. Then for any measurable space (X,𝒜)(X,\operatorname{\mathcal{A}}) the set Meas(X,R;R)\operatorname{Meas}(X,R;\operatorname{\mathcal{B}}_{R}) is a ring. In particular, the real-measurable functions Meas(X,;)\operatorname{Meas}(X,\operatorname{\operatorname{\mathbb{R}}};\operatorname{\mathcal{B}}_{\operatorname{\operatorname{\mathbb{R}}}}) and complex-measurable functions Meas(X,;)\operatorname{Meas}(X,\operatorname{\operatorname{\mathbb{C}}};\operatorname{\mathcal{B}}_{\operatorname{\operatorname{\mathbb{C}}}}) are rings.

Proof.

Since rings are set-theoretic models of the Lawvere theory 𝒯𝐑𝐢𝐧𝐠\operatorname{\mathcal{T}}_{\operatorname{\mathbf{Ring}}} of rings described in Example 3.5 and topological rings are topological models of 𝒯𝐑𝐢𝐧𝐠\operatorname{\mathcal{T}}_{\operatorname{\mathbf{Ring}}}, we simply apply Theorem 4.12 to the set Meas(X,R;R)\operatorname{Meas}(X,R;\operatorname{\mathcal{B}}_{R}). The final statements follow from the fact that \operatorname{\operatorname{\mathbb{R}}} and \operatorname{\operatorname{\mathbb{C}}} are topological rings. ∎

Corollary 4.14.

Let GG be a topological group and let (X,𝒜)(X,\operatorname{\mathcal{A}}) be any measurable space. Then Meas(X,G;G)\operatorname{Meas}(X,G;\operatorname{\mathcal{B}}_{G}) is a group. In particular, for any local field FF the set Meas(X,GLn(F);GLn(F))\operatorname{Meas}(X,\operatorname{GL}_{n}(F);\operatorname{\mathcal{B}}_{\operatorname{GL}_{n}(F)}) is a group.

Proof.

This follows mutatis mutandis to the proof of Corollary 4.13 save with the Lawvere theory 𝒯𝐑𝐢𝐧𝐠\operatorname{\mathcal{T}}_{\operatorname{\mathbf{Ring}}} replaced with the Lawvere theory 𝒯𝐆𝐫𝐩\operatorname{\mathcal{T}}_{\operatorname{\mathbf{Grp}}} of groups as described in Example 3.4. The final statement of the corollary follows from the fact that because FF is a local field GLn(F)\operatorname{GL}_{n}(F) naturally inherits the structure of a topological group from the topology on FF. ∎

References

  • [1] J. Adámek and J. Rosický, Locally presentable and accessible categories, London Mathematical Society Lecture Note Series, vol. 189, Cambridge University Press, Cambridge, 1994. MR 1294136
  • [2] M. Barr and C. Wells, Toposes, triples and theories, Repr. Theory Appl. Categ. (2005), no. 12, x+288, Corrected reprint of the 1985 original [MR0771116]. MR 2178101
  • [3] M. Brandenburg, Large limit sketches and topological space objects, 2021, Available at arxive.org/abs/2106.11115.
  • [4] J.B. Conway, A course in abstract analysis, Graduate studies in mathematics, American Mathematical Society, 2012.
  • [5] J. L. Doob, Measure theory, Graduate Texts in Mathematics, vol. 143, Springer-Verlag, New York, 1994. MR 1253752
  • [6] M. Hyland and J. Power, The category theoretic understanding of universal algebra: Lawvere theories and monads, Computation, meaning, and logic: articles dedicated to Gordon Plotkin, Electron. Notes Theor. Comput. Sci., vol. 172, Elsevier Sci. B. V., Amsterdam, 2007, pp. 437–458. MR 2328298
  • [7] F. E. J. Linton, An outline of functorial semantics, Sem. on Triples and Categorical Homology Theory (ETH, Zürich, 1966/67), Lecture Notes in Math., No. 80, Springer, Berlin-New York, 1969, pp. 7–52. MR 244340
  • [8] T. Tao, An introduction to measure theory, Graduate Studies in Mathematics, vol. 126, American Mathematical Society, Providence, RI, 2011. MR 2827917
  • [9] M.E. Taylor, Measure theory and integration, Graduate studies in mathematics, American Mathematical Soc., 2006.