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

Algorithmic Randomness, Effective Disintegrations, and Rates of Convergence to the Truth

Simon M. Huttegger Department of Logic and Philosophy of Science
5100 Social Science Plaza
University of California, Irvine
Irvine, CA 92697-5100, U.S.A.
[email protected] http://faculty.sites.uci.edu/shuttegg/
Sean Walsh Department of Philosophy
University of California, Los Angeles
390 Portola Plaza, Dodd Hall 321
Los Angeles, CA 90095-1451
[email protected] http://philosophy.ucla.edu/person/sean-walsh/
 and  Francesca Zaffora Blando Department of Philosophy
Carnegie Mellon University
Baker Hall 161
5000 Forbes Avenue
Pittsburgh, PA 15213
[email protected]
Abstract.

Lévy’s Upward Theorem says that the conditional expectation of an integrable random variable converges with probability one to its true value with increasing information. In this paper, we use methods from effective probability theory to characterise the probability one set along which convergence to the truth occurs, and the rate at which the convergence occurs. We work within the setting of computable probability measures defined on computable Polish spaces and introduce a new general theory of effective disintegrations. We use this machinery to prove our main results, which (1) identify the points along which certain classes of effective random variables converge to the truth in terms of certain classes of algorithmically random points, and which further (2) identify when computable rates of convergence exist. Our convergence results significantly generalize earlier results within a unifying novel abstract framework, and there are no precursors of our results on computable rates of convergence. Finally, we make a case for the importance of our work for the foundations of Bayesian probability theory.

2010 Mathematics Subject Classification:
Primary 03D32 Secondary: 03A10, 03D78, 03F60, 60A10, 60B05, 60G48
Many thanks to Jeremy Avigad, Peter Cholak, Johanna Franklin, Alexander Kastner, Josiah Lopez-Wild, Christopher Porter, Michael Rescorla, and Jason Rute for discussion and feedback.

1. Introduction

Measure-theoretic probability was developed in the early 20th Century in response to pressing problems in statistical physics, astronomy, and pure mathematics, and today it is used throughout the mathematical sciences.111For a historical survey, see [73]. What proved to be an especially significant conceptual progress was the ability to say that certain properties are true with probability one. Early examples include Borel’s strong law of large numbers, irrational rotations of the unit interval, Birkhoff’s ergodic theorem, and Poincaré’s recurrence theorem. It is often unclear, however, what these sets are. That is to say, measure-theoretic results only assert the existence of certain sets of probability one but fail to characterise the points that are elements of those sets. This was pointed out as early as 1916 by Weyl, who insisted that a deeper understanding of the sets involved in zero-one laws was necessary in order to interpret the results of measure-theoretic probability.222[74].

The theory of algorithmic randomness involves a fine-grained classification of different measure one sets, with the primary exemplars being the Martin-Löf random points, the Schnorr random points, and the Kurtz random points.333The original papers of Martin-Löf, Schnorr, and Kurtz are: [41], [42], [64], [65], [36]. There are now several comprehensive references on algorithmic randomness, including [39], [50], [15], [68]. Originally this was done for the uniform “fair coin” measure on Cantor space (the space of infinite sequences of 0’s and 1’s) and the famous results pertained to algorithmic incompressibility and the Turing degrees.444For instance, the Levin-Schnorr characterisation of Martin-Löf randomness in terms of initial segment complexity, and the Kučera-Gács proof that every Turing degree is below the degree of a Martin-Löf random. See, e.g., [15, Theorem 6.3.10 p. 239, Theorem 8.3.2 p. 326] for statement and references. However, the theory has been recently developed for a more general class of computable probability measures on computable spaces, by authors such as Gács, Hoyrup and Rojas, Reimann, Rute, and Miyabe.555[22], [29], [30], [56], [61], [46], [32] (listed in rough chronological order). A related recent trend has been showing that effectivized versions of classical theorems on almost sure convergence prove convergence exactly on various classes of algorithmically random points.666[7], [47]. The latter is, in part, a survey and contains many further references. This arguably contributes to the deeper understanding along the lines suggested by Weyl. Further, this recent trend suggests reconceiving of the various notions of algorithmic randomness less as on a par with rival conceptual analyses of a pre-theoretic phenomenon, à la the Church-Turing thesis, and more as delineations of extensionally and conceptually distinct kinds of probability one events.777This point is due to [54]. Or, if one puts the point in terms of the corresponding null sets, the various notions demarcate different types of impossibility that occur throughout measure-theoretic mathematics and its many applications.

Our main theorems (Theorems 1.5, 1.6, 1.8, 1.9, 1.11) contribute to this recent literature by characterising, in terms of algorithmic randomness, the points at which Lévy’s Upward Theorem holds for various classes of effective random variables, as well as providing information about the rates of convergence to the truth.

Let us recall the classical statement of Lévy’s Theorem.888[75, p. 134], [38, §41 pp. 128 ff]. Suppose (X,,ν)(X,\mathscr{F},\nu) is a probability triple. Let 1,2,\mathscr{F}_{1},\mathscr{F}_{2},\ldots be an increasing sequence of sub-σ\sigma-algebras of \mathscr{F} whose union generates \mathscr{F}. Then Lévy’s Upward Theorem states that one has 𝔼ν[fn]f\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}]\rightarrow f both ν\nu-a.s. and in L1(ν)L_{1}(\nu), for any \mathscr{F}-measurable function ff in L1(ν)L_{1}(\nu). In this, 𝔼ν[f𝒢]\mathbb{E}_{\nu}[f\mid\mathscr{G}] denotes the conditional expectation of ff relative to 𝒢\mathscr{G}, which, recall, is defined as the ν\nu-a.s. unique 𝒢\mathscr{G}-measurable function gg such that Ag𝑑ν=Af𝑑ν\int_{A}g\;d\nu=\int_{A}f\;d\nu for all events AA in the sub-σ\sigma-algebra 𝒢\mathscr{G} of \mathscr{F}.

The convergence in Lévy’s Upward Theorem is one of the cornerstones of Bayesian epistemology.999[18, pp. 144 ff], [28, pp. 28-29]. The random variable ff can be thought of as a quantity that a Bayesian agent, whose degrees of belief are captured by the underlying probability measure ν\nu, is trying to estimate by repeatedly performing an experiment. The quantity 𝔼ν[fn]\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}] can be seen as encoding the agent’s opinions regarding the value of ff after having observed the outcomes of the first nn experiments. Lévy’s Upward Theorem then implies that, with probability one, the Bayesian agent’s opinions regarding the value of ff will converge to ff’s true value in the limit.

To be able to characterise the ν\nu-measure one set on which 𝔼ν[fn]f\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}]\rightarrow f, one needs to choose versions of 𝔼ν[fn]\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}] and ff. It seems natural to focus attention on classes of effective random variables ff defined relative to spaces XX and probability measures ν\nu which are themselves computable. For, computability seems like a natural constraint to place on our Bayesian agent, and many of the examples of probability measures and random variables that occur in practice and applications are computable. However, the Bayesian perspective recommends few other general constraints on what is eligible to be a credence or a prior. Hence it is important to develop the theory for a maximally broad class of computable spaces and probability measures.101010The distinctive status of the computability hypothesis which we are suggesting raises a host of interesting and complex conceptual questions, ranging from the nature of cognition to the character of inductive inference. We put these issues aside here.

1.1. Effective probability and algorithmic randomness

The computable Polish spaces with computable probability measures are such an appropriately general class of spaces and measures. In this brief section we collect together the few definitions we need about their theory. The reader already familiar with these concepts can easily skip to the next section (§1.2).

A Polish space is a topological space which is separable and completely metrizable. All the paradigmatic spaces such as the reals and their products and their closed and open subspaces are Polish, and similarly for Cantor space. Descriptive set theory takes as its subject matter the Borel and projective subsets of Polish spaces.111111[34], [48]. When topological considerations are salient or when one needs i.i.d. sequences with prescribed distributions, it is often assumed in contemporary probability theory that the sample space is a Polish space or a Borel subset thereof.121212A Borel subset of a Polish space together with its Borel subsets is known as a standard Borel space in descriptive set theory (cf. §[34, Definition 12.5, Corollary 13.4]). For representative examples of standard Borel spaces within probability, see e.g. [17, p. 51], [33, p. 7, pp. 561 ff]. For a classic probability text that foregrounds standard Borel spaces, see [51, Chapter 1].

A computable Polish space XX is a Polish space with a distinguished countable dense set x0,x1,x_{0},x_{1},\ldots and a distinguished complete compatible metric dd such that the distance d(xi,xj)d(x_{i},x_{j}) between any two elements of the countable dense set is a computable real, uniformly in i,j0i,j\geq 0.131313A standard reference for computable Polish spaces is [48, Chapter 3]. A comparison to the Weihrauch approach to computable analysis is given in [24]. One can view the treatment of metric spaces in [70] as an axiomatization of Polish spaces and reals which are computable in an oracle. Finally, the study of computable Polish spaces in and of themselves is distinct from effective descriptive set theory, which usually refers to techniques for proving results about all Borel sets by first proving it for lightface Borel sets in Baire space (the most famous example of this is the Glimm-Effros dichotomy (cf. [23, Chapter 6])). (The enumeration of the distinguished countable dense set can contain repetitions, and will need to do so in finite spaces.)

In a computable Polish space, an open set UU is c.e. open if there is a computable function n()n(\cdot) which enumerates a subsequence xn(i)x_{n(i)} of the countable dense set and a computable sequence rir_{i} of rational radii such that U=iBd(xn(i),ri)U=\bigcup_{i}B_{d}(x_{n(i)},r_{i}). In this, Bd(x,r)B_{d}(x,r) denotes the open ball with centre xx and radius rr relative to metric dd (when the metric dd is clear from context, we just write B(x,r)B(x,r)). The name “c.e. open” is chosen since the natural numbers are a computable Polish space with the discrete metric, and the c.e. opens in this space are precisely the computably enumerable sets of natural numbers, one of the canonical objects of the contemporary theory of computation.141414See [71], a standard reference. Further, many of the elementary methods of studying c.e. sets (e.g., universal enumerations) extend to c.e. open sets. The complements of c.e. open sets are called effectively closed sets (cf. §2.1).

In a computable Polish space, we say that a sequence xnxx_{n}\rightarrow x at geometric rate bb if d(x,xn)bnd(x,x_{n})\leq b^{-n} for all n0n\geq 0. We say that a sequence xnxx_{n}\rightarrow x fast if xnxx_{n}\rightarrow x at geometric rate b=2b=2. We then say that a point xx is computable if there is a computable function n()n(\cdot) which enumerates a subsequence xn(i)x_{n(i)} of the countable dense set such that xn(i)xx_{n(i)}\rightarrow x fast. This subsequence is called a witness to the computability of xx. In Cantor space with its usual metric, the computable points are precisely the computable subsets of natural numbers.

In the real numbers, the countable dense set is the rationals, and the above definition of computable points is precisely how Turing defined computable real numbers at the outset of the theory of computation nearly a century ago.151515[72]. An equivalent formalisation of computable reals is by Dedekind cuts. A real xx is left-c.e. (resp. right-c.e.) if its left Dedekind cut {q:q<x}\{q\in\mathbb{Q}:q<x\} in the rationals is a c.e. set (resp. if its right Dedekind cut {q:x<q}\{q\in\mathbb{Q}:x<q\} in the rationals is a c.e. set). One can show that a real is computable iff it is both left-c.e. and right-c.e., and uniformly so. (An example of a left-c.e. real that is not computable is n2f(n)\sum_{n}2^{-f(n)}, where f:f:\mathbb{N}\rightarrow\mathbb{N} is an injective computable function with non-computable range.)161616These and other effective aspects of real numbers are treated extensively in e.g. [57], [15, Chapter 5].

These preliminaries in place, one can then quickly define the required core notions from algorithmic randomness and effective probability. These are all needed in order to formally state our main theorems, but one might restrict oneself to (1)-(8) on a first pass and come back to the others as needed.

Definition 1.1.

(Core notions)

  1. (1)

    A function f:X(,]f:X\rightarrow(-\infty,\infty] is lower semi-computable (abbreviated lsc) if for all rational qq, the set f1(q,]f^{-1}(q,\infty] is uniformly c.e. open.

  2. (2)

    A function f:X[,)f:X\rightarrow[-\infty,\infty) is upper semi-computable (abbreviated usc) if for all rational qq, the set f1[,q)f^{-1}[-\infty,q) is uniformly c.e. open.

  3. (3)

    A probability measure ν\nu is computable if ν(U)\nu(U) is uniformly left-c.e. as UU ranges over c.e. opens.

  4. (4)

    Given a computable probability measure ν\nu and a computable real p1p\geq 1, a function f:X[0,]f:X\rightarrow[0,\infty] is an Lp(ν)L_{p}(\nu) Schnorr test if it is lsc and if fp\|f\|_{p} is a computable real, where this denotes the pp-norm fp=(|f|p𝑑ν)1p\|f\|_{p}=(\int\left|f\right|^{p}\;d\nu)^{\frac{1}{p}}.

  5. (5)

    Given a computable probability measure ν\nu and a computable real p1p\geq 1, a function f:X[0,]f:X\rightarrow[0,\infty] is an Lp(ν)L_{p}(\nu) Martin-Löf test if it is lsc and fp<\|f\|_{p}<\infty.

  6. (6)

    A point xx in XX is Kurtz random relative to ν\nu (abbreviated 𝖪𝖱ν(X)\mathsf{KR}^{\nu}(X)) if xx is in every c.e. open UU with ν(U)=1\nu(U)=1.

  7. (7)

    A point xx in XX is Schnorr random relative to ν\nu (abbreviated 𝖲𝖱ν(X)\mathsf{SR}^{\nu}(X)) if f(x)<f(x)<\infty for any L1(ν)L_{1}(\nu) Schnorr test ff (equivalently, for any Lp(ν)L_{p}(\nu) Schnorr test, for p1p\geq 1 computable).

  8. (8)

    A point xx in XX is Martin-Löf random relative to ν\nu (abbreviated 𝖬𝖫𝖱ν(X)\mathsf{MLR}^{\nu}(X)) if f(x)<f(x)<\infty for any L1(ν)L_{1}(\nu) Martin-Löf test ff (equivalently, for any Lp(ν)L_{p}(\nu) Martin-Löf test, for p1p\geq 1 computable).

  9. (9)

    A computable basis \mathscr{B} for XX is a computable sequence of c.e. opens such that every c.e. open can be effectively written as a union of elements in \mathscr{B}.

  10. (10)

    If ν\nu is a computable probability measure, then a ν\nu-computable basis \mathscr{B} for XX is a computable basis such that (i) finite unions from \mathscr{B} uniformly have ν\nu-computable measure, and (ii) each c.e. open in \mathscr{B} is uniformly paired with an effectively closed superset of the same ν\nu-measure. If ν\nu is clear from context, we simply say measure computable basis instead of ν\nu-computable basis.

  11. (11)

    A sub-σ\sigma-algebra \mathscr{F} of the Borel sets on XX is ν\nu-effective if it is generated by a computable sequence of events {Am:m0}\{A_{m}:m\geq 0\} from the algebra 𝒜\mathscr{A} generated by a ν\nu-computable basis \mathscr{B}.171717When working with 𝒜\mathscr{A}, we assume that we are working with the codes for Boolean combinations of elements of \mathscr{B}, and only by extension with the sets that they define. This is because there are some spaces where Boolean algebra structure on the quotient is not computable. We say that \mathscr{F} is generated by {Am:m0}\{A_{m}:m\geq 0\}.

  12. (12)

    A full ν\nu-effective filtration n\mathscr{F}_{n} (resp. almost-full ν\nu-effective filtration) is an increasing sequence n\mathscr{F}_{n} of uniformly ν\nu-effective sub-σ\sigma-algebras generated by a uniformly computable sequence {An,m:m0}\{A_{n,m}:m\geq 0\} from the algebra 𝒜\mathscr{A} generated by a ν\nu-computable basis \mathscr{B}, which is further equipped with a uniform procedure for going from a c.e. open UU to a computable sequence Ani,miA_{n_{i},m_{i}} such that U=iAni,miU=\bigcup_{i}A_{n_{i},m_{i}} (resp. U=iAni,miU=\bigcup_{i}A_{n_{i},m_{i}} on 𝖪𝖱ν(X)\mathsf{KR}^{\nu}(X)).

  13. (13)

    If xx is a point of the computable Polish space XX and YY is a subset of Baire space (the space of all functions from natural numbers to natural numbers), then xx weakly computes an element of YY if, for every sequence xn(i)x_{n(i)} from the countable dense set of XX such that xn(i)xx_{n(i)}\rightarrow x fast, there is yy in YY which is Turing reducible to the function in(i)i\mapsto n(i). If Y={y}Y=\{y\}, then we just say that xx weakly computes yy.181818One can extend Turing reducibility from a relation between sets of natural numbers to a relation between closed subsets of Baire space. In this setting, the notion of weak computation is called Muchnik reducibility, and is contrasted to a strong uniform notion called Medvedev reducibility. See [69], [27] for introduction and references, although this theory is usually focused on effectively closed sets. Given a point xx, the set of functions in(i)i\mapsto n(i) such that xn(i)xx_{n(i)}\rightarrow x at a fixed rate, in the sense of (15), is a closed subset of Baire space.,{}^{,\;} 191919If XX is Cantor space or the reals, then for each point xx of the space, there is a sequence in(i)i\mapsto n(i) of least Turing degree such that xn(i)xx_{n(i)}\rightarrow x fast. In these settings, computational properties of the point of the space usually refer to those of this sequence. However, there are spaces for which there are points with no sequence of least Turing degree. See Miller [43].

  14. (14)

    If xx is a point of the computable Polish space XX, and 𝒞\mathcal{C} is any collection of Turing degrees (equivalence classes of elements of Baire space under Turing reducibility), then we say that xx is in 𝒞\mathcal{C} if there is some in(i)i\mapsto n(i) whose Turing degree is in 𝒞\mathcal{C} such that xn(i)xx_{n(i)}\rightarrow x fast, where xjx_{j} again enumerates the countable dense set. In the case where 𝒞\mathcal{C} just consists of the computable degree, note that xx is in 𝒞\mathcal{C} iff xx is computable as a point of XX.

  15. (15)

    Suppose yn,yy_{n},y are elements in a metric space YY such that ynyy_{n}\rightarrow y. Then a rate of convergence for ynyy_{n}\rightarrow y is a function m:>0m:\mathbb{Q}^{>0}\rightarrow\mathbb{N} such that for all rational ϵ>0\epsilon>0 and all nm(ϵ)n\geq m(\epsilon) one has d(yn,y)<ϵd(y_{n},y)<\epsilon.202020If ynyy_{n}\rightarrow y at geometric rate b>1b>1 in a computable Polish space, then one defines a rate in the sense of (15) by setting m(ϵ)=nm(\epsilon)=n for the least nn such that bn<ϵb^{-n}<\epsilon. Often in practice we use the case where YY is the reals and yn=fn(x)y_{n}=f_{n}(x) and y=f(x)y=f(x), where fn,ff_{n},f are real-valued functions. A synonym for rate is modulus, and so we often use the mm variable for rates.

For the algorithmic randomness notions in (6)-(8), we just write 𝖪𝖱ν\mathsf{KR}^{\nu} instead of 𝖪𝖱ν(X)\mathsf{KR}^{\nu}(X) when XX is clear from context; and similarly for 𝖲𝖱ν\mathsf{SR}^{\nu} and 𝖬𝖫𝖱ν\mathsf{MLR}^{\nu}. For σ\sigma-algebras \mathscr{F}, it is always understood that they are sub-σ\sigma-algebras of the Borel σ\sigma-algebra, and when ν\nu is clear from context we just say effective instead of ν\nu-effective.

Algorithmic randomness is often formulated in terms of effective null sets, called sequential tests. But the definitions given above in terms of integral tests are easier to work with in our setting and are known to be equivalent to the sequential definitions, by theorems of Levin and Miyabe.212121[37], [45, Theorem 3.5].

Before turning to disintegrations, it is helpful to introduce notational conventions regarding versions of integrable functions vs. equivalence classes thereof. In the following definition, the σ\sigma-algebra on [,][-\infty,\infty] is simply {BC:B Borel,C{,}}\{B\cup C:B\subseteq\mathbb{R}\mbox{ Borel},C\subseteq\{-\infty,\infty\}\}, and similarly for [0,][0,\infty].

Definition 1.2.

(Conventions on functions defined pointwise vs. functions defined up to ν\nu-a.s. equivalence)

Suppose XX is a Polish space and suppose that ν\nu is a finite non-negative measure on the Borel events of XX. Then we define:

𝕃p(ν)\mathbb{L}_{p}(\nu) is the set of pointwise defined Borel measurable functions f:X[,]f:X\rightarrow[-\infty,\infty] such that fp<\|f\|_{p}<\infty.

𝕃p+(ν)\mathbb{L}^{+}_{p}(\nu) is the set of pointwise defined Borel measurable functions f:X[0,]f:X\rightarrow[0,\infty] such that fp<\|f\|_{p}<\infty.

Lp(ν)L_{p}(\nu) is the set of equivalence classes of elements of 𝕃p(ν)\mathbb{L}_{p}(\nu) under ν\nu-a.s. equivalence. That is, Lp(ν)L_{p}(\nu) is the classical Banach space with norm p\|\cdot\|_{p}.

Lp+(ν)L^{+}_{p}(\nu) is the set of equivalence classes of elements of 𝕃p+(ν)\mathbb{L}^{+}_{p}(\nu) under ν\nu-a.s. equivalence. That is, Lp+(ν)L^{+}_{p}(\nu) is a positive cone in the Banach space Lp(ν)L_{p}(\nu).

Then 𝕃p(ν)\mathbb{L}_{p}(\nu) projects onto Lp(ν)L_{p}(\nu) by sending a function to its equivalence class, and likewise 𝕃p+(ν)\mathbb{L}^{+}_{p}(\nu) projects onto Lp+(ν)L^{+}_{p}(\nu). Note that Lp(ν)L_{p}(\nu) Schnorr tests and Martin-Löf tests from Definition 1.1(4)-(5) are elements of 𝕃p+(ν)\mathbb{L}_{p}^{+}(\nu), and they are elements of Lp+(ν)L_{p}^{+}(\nu) only after passing to the equivalence class.

1.2. Classical and effective disintegrations

We use disintegrations for the versions 𝔼ν[fn]\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}] of conditional expectation. While, classically, conditional expectation is defined only ν\nu-a.s., in order to characterise algorithmic randomness notions in terms of Lévy’s Upward Theorem, we need to select specific versions of conditional expectation. The concept of a disintegration provides a very general way of making such selections. It is due to Rohlin222222[58], [59], [60]. and is routinely used today in ergodic theory and optimal transport,232323It is often used in the proof of the Ergodic Decomposition Theorem and the Gluing Lemma. See [19, 154], [63, 182]. and it is closely related to conditional probability distributions.242424[11], [53, §5.3].

Suppose that XX is a Polish space, ν\nu is a probability measure on the Borel sets of XX and \mathscr{F} is a countably generated sub-σ\sigma-algebra of the Borel σ\sigma-algebra. Define the equivalence relation \sim_{\mathscr{F}} on XX by xxx\sim_{\mathscr{F}}x^{\prime} iff, for all AA in \mathscr{F}, one has xx in AA iff xx^{\prime} in AA, and let [x][x]_{\mathscr{F}} be the corresponding equivalence class.252525Since we are focused on Lévy’s Upward Theorem, we are focusing on countably generated sub-σ\sigma-algebras \mathscr{F} of the Borel σ\sigma-algebra. Note that this has the consequence that the relation \sim_{\mathscr{F}} is a smooth Borel equivalence relation (cf. [23, §5.4]). More complicated Borel equivalence relations occur naturally in nearby topics. For instance, Rute examines Lévy’s Downward Theorem (cf. [61, Theorem 11.2], [75, Theorem 14.4]), which in Cantor space results naturally in the sub-σ\sigma-algebra of E0E_{0}-invariant events, where E0E_{0} is the Borel equivalence relation featuring in the Glimm-Effros dichotomy (cf. [23, Definition 6.1.1]).

Let +(X)\mathcal{M}^{+}(X) be the Polish space of non-negative Borel measures on XX (cf. §2.2). For Borel measurable ρ:X+(X)\rho:X\rightarrow\mathcal{M}^{+}(X), whose action is written as xρxx\mapsto\rho_{x}, we define the partial map

𝔼ν[]():𝕃1(ν)×X[,] by 𝔼ν[f](x)=f(v)dρx(v)\mathbb{E}_{\nu}[\cdot\mid\mathscr{F}](\cdot):\mathbb{L}_{1}(\nu)\times X\dashrightarrow[-\infty,\infty]\hskip 8.53581pt\mbox{ by }\hskip 8.53581pt\mathbb{E}_{\nu}[f\mid\mathscr{F}](x)=\int f(v)\;d\rho_{x}(v) (1.1)

This map is a version, that is, it is partially defined on all pairs (f,x)(f,x). Note that it is totally defined on 𝕃1+(ν)×X\mathbb{L}_{1}^{+}(\nu)\times X, with range [0,][0,\infty],262626Indeed, it is totally defined on all pairs (f,x)f,x) where ff is non-negative Borel measurable. But for our purpose of defining a version of 𝔼ν[]\mathbb{E}_{\nu}[\cdot\mid\mathscr{F}] we only need to pay attention to when ff is in 𝕃1(ν)\mathbb{L}_{1}(\nu). and it is totally defined and finite on all simple functions. It is further helpful to keep in mind that whether 𝔼ν[f](x)\mathbb{E}_{\nu}[f\mid\mathscr{F}](x) is finite depends on whether the element ff of 𝕃1(ν)\mathbb{L}_{1}(\nu) is additionally in 𝕃1(ρx)\mathbb{L}_{1}(\rho_{x}): that is, it is integrability with respect to ρx\rho_{x} rather than ν\nu which is at issue.

For a Polish space XX, one says that a map ρ:X+(X)\rho:X\rightarrow\mathcal{M}^{+}(X) is the disintegration of \mathscr{F} with respect to ν\nu if both the following happen:272727We are following the treatment of Einsiedler-Ward [19, 135]. Since the main examples of disintegrations involve products (cf. Appendicies A-B), often alternative definitions of disintegrations involve maps that axiomatize the role that the projection operators play in the paradigmatic examples. For an example of definitions along these lines, see [11], [53, §5.3].

  • for all ff in 𝕃1(ν)\mathbb{L}_{1}(\nu), one has that 𝔼ν[f]\mathbb{E}_{\nu}[f\mid\mathscr{F}] is a version of the conditional expectation of ff with respect to \mathscr{F} and ν\nu.282828Hence it is in 𝕃1(ν)\mathbb{L}_{1}(\nu), and thus it is defined and finite for ν\nu-a.s. many xx from XX.

  • For ν\nu-a.s. many xx from XX, one has ρx(X)=1\rho_{x}(X)=1 and ρx([x])=1\rho_{x}([x]_{\mathscr{F}})=1.

A disintegration of \mathscr{F} with respect to ν\nu exists for any countably generated sub-σ\sigma-algebra \mathscr{F} of the Borel σ\sigma-algebra on XX.292929[19, 135]. Indeed, a little more is true: one can replace XX by one of its Borel subsets. Further, one can relax the assumption that \mathscr{F} is countably generated, provided that one does not insist on ρx([x])=1\rho_{x}([x]_{\mathscr{F}})=1. Further, it is possible to be more agnostic about the codomain of ρ\rho outside the ν\nu-measure one set on which it outputs probability measures. See Appendix A for two classical examples of disintegrations.

Here is our key definition of effective disintegration:

Definition 1.3.

(Effective disintegrations).

Let XX be a computable Polish space. Let ν\nu be a computable probability measure on XX. Let \mathscr{F} be a ν\nu-effective σ\sigma-algebra. Let 𝖷𝖱ν\mathsf{XR}^{\nu} be a ν\nu-measure one subset of 𝖪𝖱ν(X)\mathsf{KR}^{\nu}(X). Then the map ρ:X+(X)\rho:X\rightarrow\mathcal{M}^{+}(X) is an 𝖷𝖱ν\mathsf{XR}^{\nu} disintegration of \mathscr{F} with respect to ν\nu if each of the following happen:

  1. (1)

    For all ff in 𝕃1(ν)\mathbb{L}_{1}(\nu), one has that 𝔼ν[f]\mathbb{E}_{\nu}[f\mid\mathscr{F}] is a version of the conditional expectation of ff with respect to \mathscr{F} and ν\nu.

  2. (2)

    For all xx in 𝖷𝖱ν\mathsf{XR}^{\nu} , one has that ρx(X)=1\rho_{x}(X)=1 and ρx([x]𝖷𝖱ν)=1\rho_{x}([x]_{\mathscr{F}}\cap\mathsf{XR}^{\nu})=1.

  3. (3)

    For c.e. open UU, the map xρx(U)x\mapsto\rho_{x}(U) is uniformly lsc from XX to [0,)[0,\infty).

We further define:

A Kurtz disintegration is simply a 𝖪𝖱ν\mathsf{KR}^{\nu} disintegration.

A Schnorr disintegration is a map which is both a 𝖪𝖱ν\mathsf{KR}^{\nu} disintegration and a 𝖲𝖱ν\mathsf{SR}^{\nu} disintegration.

A Martin-Löf disintegration is a map which is a 𝖪𝖱ν\mathsf{KR}^{\nu} disintegration and a 𝖲𝖱ν\mathsf{SR}^{\nu} disintegration and a 𝖬𝖫𝖱ν\mathsf{MLR}^{\nu} disintegration.

Technically, it appears possible to, e.g., be a 𝖲𝖱ν\mathsf{SR}^{\nu} disintegration but not a 𝖪𝖱ν\mathsf{KR}^{\nu} disintegration. This is due to the universal quantifier over 𝖷𝖱ν\mathsf{XR}^{\nu} at the outset of (2). But this possibility does not appear to occur naturally among examples.

Due to space constraints, we have opted to focus on theory in the body of the text, and have put a brief discussion of the many interesting examples of effective disintegrations in Appendix B.

Finally, we can define:

Definition 1.4.

Let n\mathscr{F}_{n} be an almost-full effective filtration, equipped uniformly with Kurtz disintegrations ρ(n)\rho^{(n)}. A point xx in XX is said to be density random with respect to ρ\rho, abbreviated 𝖣𝖱ρν(X)\mathsf{DR}^{\nu}_{\rho}(X), if xx is in 𝖬𝖫𝖱ν(X)\mathsf{MLR}^{\nu}(X) and limnρx(n)(U)=δx(U)\lim_{n}\rho_{x}^{(n)}(U)=\delta_{x}(U) for every c.e. open UU.

In this, δx\delta_{x} is the Dirac measure centred at xx. With the limit written as such, by the Portmanteau Theorem one sees that it is a strengthening of the weak convergence of the measures ρx(n)δx\rho_{x}^{(n)}\rightarrow\delta_{x}. Since we use the disintegration to define the conditional expectation, as in equation (1.1) above, the limit in Definition 1.4 can be written equivalently as limn𝔼ν[IUn](x)=IU(x)\lim_{n}\mathbb{E}_{\nu}[I_{U}\mid\mathscr{F}_{n}](x)=I_{U}(x). With respect to the canonical filtration of length nn-strings on Cantor space and its natural disintegration (cf. Example B.1), density randomness has been a focal topic in recent literature on algorithmic randomness.303030[3], [47], [35]. In this setting with ν\nu being the uniform measure, it is known that 𝖣𝖱ρν\mathsf{DR}^{\nu}_{\rho} is a proper subset of 𝖬𝖫𝖱ν\mathsf{MLR}^{\nu}. One example which shows this properness is an element ω\omega of 𝖬𝖫𝖱ν\mathsf{MLR}^{\nu} such that {ω:ω<lexω}\{\omega^{\prime}:\omega^{\prime}<_{lex}\omega\} is c.e. open, where <lex<_{lex} is the lexicographic order. Definition 1.4 is our suggestion for how to generalise this to the setting of arbitrary effective disintegrations.

1.3. Statement of main results

Our first main theorem is the following:

Theorem 1.5.

(Effective Upward Lévy Theorem for Schnorr Randomness). Suppose that XX is a computable Polish space and ν\nu is a computable probability measure. Suppose that n\mathscr{F}_{n} is an almost-full effective filtration, equipped uniformly with Kurtz disintegrations.

If p1p\geq 1 is computable, then the following four items are equivalent for xx in XX:

  1. (1)

    xx is in 𝖲𝖱ν(X)\mathsf{SR}^{\nu}(X).

  2. (2)

    xx is in 𝖪𝖱ν(X)\mathsf{KR}^{\nu}(X) and limn𝔼ν[fn](x)=f(x)\lim_{n}\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x)=f(x) for all Lp(ν)L_{p}(\nu) Schnorr tests ff.

  3. (3)

    xx is in 𝖪𝖱ν(X)\mathsf{KR}^{\nu}(X) and limn𝔼ν[fn](x)\lim_{n}\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x) exists for all Lp(ν)L_{p}(\nu) Schnorr tests ff and limn𝔼ν[IUn](x)=IU(x)\lim_{n}\mathbb{E}_{\nu}[I_{U}\mid\mathscr{F}_{n}](x)=I_{U}(x) for all c.e. opens UU with ν(U)\nu(U) computable.

  4. (4)

    xx is in 𝖪𝖱ν(X)\mathsf{KR}^{\nu}(X) and limn𝔼ν[fn](x)\lim_{n}\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x) exists for all Lp(ν)L_{p}(\nu) Schnorr tests ff.

In condition (2), limn𝔼ν[fn](x)=f(x)\lim_{n}\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x)=f(x) means that the limit of 𝔼ν[fn](x)\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x) exists and is finite and equal to f(x)f(x). Likewise in (3)-(4), the existence of the limit means that it is finite. Note that (3) implies that 𝖲𝖱ν\mathsf{SR}^{\nu} already proves the analogue of density randomness where we restrict to c.e. open UU with 0<ν(U)<10<\nu(U)<1 computable.

For rates of convergence, we have:

Theorem 1.6.

(Rates for Upward Lévy Theorem for Schnorr Randomness). For all X,ν,nX,\nu,\mathscr{F}_{n} as in Theorem 1.5, one has:

  1. (1)

    For all xx in 𝖲𝖱ν(X)\mathsf{SR}^{\nu}(X) and all computable p1p\geq 1 and all Lp(ν)L_{p}(\nu) Schnorr tests ff one has that xx weakly computes a rate of convergence for 𝔼ν[fn](x)f(x)\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x)\rightarrow f(x).

  2. (2)

    For all xx in 𝖲𝖱ν(X)\mathsf{SR}^{\nu}(X) of computably dominated degree and all computable p1p\geq 1 and all Lp(ν)L_{p}(\nu) Schnorr tests ff one has that there is a computable rate for the convergence 𝔼ν[fn](x)f(x)\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x)\rightarrow f(x).

The notion in Theorem 1.6(2) is a classical notion from the theory of computation: a Turing degree is computably dominated if any function from natural numbers to natural numbers that is computable from the degree is dominated by a computable function, in the sense that the computable function is eventually above it.313131[71, 124], [50, 27]. A more traditional name for this concept is “of hyperimmune-free degree.” This more traditional name comes from an equivalent definition that emerged in the context of Post’s Problem (cf. [71, 133 ff]). For many but not all computable Polish spaces XX and ν\nu in 𝒫(X)\mathcal{P}(X) computable, there are non-atoms in 𝖬𝖫𝖱ν\mathsf{MLR}^{\nu} (and hence in 𝖲𝖱ν\mathsf{SR}^{\nu}) of computably dominated degree. This is a consequence of the existence of universal tests for 𝖬𝖫𝖱ν\mathsf{MLR}^{\nu} and the Computably Dominated Basis Theorem (cf. discussion at Proposition 2.18, Example 2.19, Question  2.20).

It is unknown to us whether Theorem 1.6(2) can be improved, in the sense of an affirmative answer to the following question:

Question 1.7.

For all xx in 𝖲𝖱ν\mathsf{SR}^{\nu} that are not of computably dominated degree and all computable p1p\geq 1 and for all Lp(ν)L_{p}(\nu) Schnorr tests ff is there a computable rate for the convergence 𝔼ν[fn](x)f(x)\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x)\rightarrow f(x)?

The following is the simplest concrete version of the question (cf. Example B.1):

If ν\nu is uniform measure on Cantor space, and ω\omega in 𝖲𝖱ν\mathsf{SR}^{\nu} is not of computably dominated degree, and if UU is c.e. open with 0<ν(U)<10<\nu(U)<1 computable and ω\omega not in UU, then does the convergence ν(U[ωn])0\nu(U\mid[\omega\upharpoonright n])\rightarrow 0 have a computable rate?323232For such UU, the set {ω:ν(U[ωn])IU(ω)}\{\omega:\nu(U\mid[\omega\upharpoonright n])\rightarrow I_{U}(\omega)\} can be rather complex. In particular, Carotenuto-Nies [8] show that it is Π30\Pi^{0}_{3}-complete when UU is dense. It is not clear to us whether this complexity is located among the 𝖲𝖱ν\mathsf{SR}^{\nu}’s or the non-computably dominated 𝖲𝖱ν\mathsf{SR}^{\nu}’s, or whether it is reflected in their rates of convergence.

Under uniform measure on Cantor space, the points which are not of computably dominated degree have measure one, a result due to Martin.333333Martin’s paper [40] is unpublished, but his proof has subsequently appeared in other sources, such as [15, Theorem 8.21.1 p. 381], [13, Theorem 1.2]. One way to negatively resolve the question would be to show that the non-computable-domination in Martin’s proof (or a variation on it) could be witnessed by a rate of convergence associated to an Lp(ν)L_{p}(\nu) Schnorr test, or perhaps even to an indicator function of a c.e. open UU with 0<ν(U)<10<\nu(U)<1 computable.

We prove Theorems 1.5-1.6 in §8. Theorem 1.5 extends and unifies prior work by Pathak, Rojas, and Simpson, and of Rute (see discussion in §1.4 below), while Theorem-1.6 is entirely new.

Our next theorem pertains to convergence along Martin-Löf tests:

Theorem 1.8.

(Effective Upward Lévy Theorem for Density Randomness p>1p>1).

Suppose that XX is a computable Polish space and ν\nu in 𝒫(X)\mathcal{P}(X) is computable. Suppose that n\mathscr{F}_{n} is an almost-full effective filtration, equipped uniformly with Kurtz disintegrations ρ(n)\rho^{(n)}.

If p>1p>1 is computable, then the following three items are equivalent for xx in XX:

  1. (1)

    xx is in 𝖣𝖱ρν(X)\mathsf{DR}_{\rho}^{\nu}(X).

  2. (2)

    xx is in 𝖪𝖱ν(X)\mathsf{KR}^{\nu}(X) and limn𝔼ν[fn](x)=f(x)\lim_{n}\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x)=f(x) for all Lp(ν)L_{p}(\nu) Martin-Löf tests ff.

  3. (3)

    xx is in 𝖪𝖱ν(X)\mathsf{KR}^{\nu}(X) and limn𝔼ν[fn](x)\lim_{n}\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x) exists for all Lp(ν)L_{p}(\nu) Martin-Löf tests ff and limn𝔼ν[IUn](x)=IU(x)\lim_{n}\mathbb{E}_{\nu}[I_{U}\mid\mathscr{F}_{n}](x)=I_{U}(x) for every c.e. open UU.

In contrast to Theorem 1.6(2), one has the following, whose proof is a traditional diagonalization argument deploying the halting set:

Theorem 1.9.

(Rates for Upward Lévy Theorem for Density Randomness).

There are X,ν,n,ρ(n)X,\nu,\mathscr{F}_{n},\rho^{(n)} as in Theorem 1.8 which have the property that for every computable p>1p>1 and every xx in 𝖣𝖱ρν(X)\mathsf{DR}_{\rho}^{\nu}(X) there is Lp(ν)L_{p}(\nu) Martin-Löf test ff such that the convergence 𝔼ν[fn](x)f(x)\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x)\rightarrow f(x) has no computable rate.

Hence, once we shift the tests from Schnorr tests to Martin-Löf tests, we never have points which possess computable rates for all tests. In one sense, Question 1.7 is asking whether there is some way to emulate a halting-set-like construction among the non-computably dominated 𝖲𝖱ν\mathsf{SR}^{\nu}’s.

We prove Theorems 1.8-1.9 in §9. Theorem 1.8 extends work from a paper of Miyabe, Nies, and Zhang, which we discuss in the next section, while Theorem 1.9 is entirely new.

Given Theorem 1.5 and Theorem 1.8, it is natural to try to understand whether there is a convergence to the truth characterisation of 𝖬𝖫𝖱ν\mathsf{MLR}^{\nu}, or at least some nearby superset of it (by contrast, 𝖣𝖱ρν\mathsf{DR}^{\nu}_{\rho} is a subset of 𝖬𝖫𝖱ν\mathsf{MLR}^{\nu}). We thus isolate a class of Martin-Löf tests ff which have approximations fsf_{s} such that fsff_{s}\rightarrow f in Lp(ν)L_{p}(\nu) at an exponential rate, but not a rate that can necessarily be computed. Hence we define the following, where clauses (1)-(3) mimic the canonical approximations of Lp(ν)L_{p}(\nu) Schnorr tests (cf. Proposition 2.16, Lemma 3.2), and where clause (4) pertains to exponential rates:

Definition 1.10.

Suppose that p1p\geq 1 is computable.

A Lp(ν)L_{p}(\nu) maximal Doob test f:X[0,]f:X\rightarrow[0,\infty] is an lsc function in Lp(ν)L_{p}(\nu) such that there is a uniformly computable sequence fsf_{s} of Lp(ν)L_{p}(\nu) Schnorr tests satisfying

  1. (1)

    0fsfs+10\leq f_{s}\leq f_{s+1} on 𝖪𝖱ν\mathsf{KR}^{\nu} and f=supsfsf=\sup_{s}f_{s} on 𝖪𝖱ν\mathsf{KR}^{\nu}.

  2. (2)

    ffsf-f_{s} is equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to a non-negative lsc function.

  3. (3)

    ftfsf_{t}-f_{s} for t>st>s is equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to an Lp(ν)L_{p}(\nu) Schnorr test, uniformly in t>st>s.

  4. (4)

    For all k0k\geq 0, sffsp(s+1)k<\sum_{s}\|f-f_{s}\|_{p}\cdot(s+1)^{k}<\infty.343434By taking k=0k=0, we have fsff_{s}\rightarrow f in Lp(ν)L_{p}(\nu), and so ff is an Lp(ν)L_{p}(\nu) Martin-Löf test, and hence in conjunction with (2) we have that ffsf-f_{s} is equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to an Lp(ν)L_{p}(\nu) Martin-Löf test.

A point xx is pp-maximal Doob random relative to ν\nu, abbreviated 𝖬𝖣𝖱ν,p(X)\mathsf{MDR}^{\nu,p}(X), if f(x)<f(x)<\infty for all Lp(ν)L_{p}(\nu) maximal Doob tests.

Our theorem on this is the following:

Theorem 1.11.

(Effective Upward Lévy Theorem for Maximal Doob Randomness, p>1p>1).

Suppose that XX is a computable Polish space and ν\nu in 𝒫(X)\mathcal{P}(X) is computable. Suppose that n\mathscr{F}_{n} is an almost-full effective filtration, equipped uniformly with Kurtz disintegrations.

If p>1p>1 is computable, then the following three items are equivalent for xx in XX:

  1. (1)

    xx in 𝖬𝖣𝖱ν,p\mathsf{MDR}^{\nu,p}.

  2. (2)

    xx is in 𝖪𝖱ν\mathsf{KR}^{\nu} and f(x)=limn𝔼ν[fn](x)f(x)=\lim_{n}\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x) for all Lp(ν)L_{p}(\nu) maximal Doob tests ff.

  3. (3)

    xx is in 𝖪𝖱ν\mathsf{KR}^{\nu} and limn𝔼ν[fn](x)\lim_{n}\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x) exists for all Lp(ν)L_{p}(\nu) maximal Doob tests ff.

We prove Theorem 1.11 in §10. The name “Maximal Doob” in Theorem 1.11 and Definition 1.10 comes from the role played by Doob’s Maximal Inequality (cf. Lemma 5.1(1)) in the proofs in §10. In Proposition 10.1, we note that 𝖬𝖫𝖱ν𝖬𝖣𝖱ν,p𝖲𝖱ν\mathsf{MLR}^{\nu}\subseteq\mathsf{MDR}^{\nu,p}\subseteq\mathsf{SR}^{\nu}. But we do not know the answer to the following question:

Question 1.12.

Are the inclusions 𝖬𝖫𝖱ν𝖬𝖣𝖱ν,p𝖲𝖱ν\mathsf{MLR}^{\nu}\subseteq\mathsf{MDR}^{\nu,p}\subseteq\mathsf{SR}^{\nu} proper?

We suspect that 𝖬𝖣𝖱ν,p\mathsf{MDR}^{\nu,p} is a proper subset of 𝖲𝖱ν\mathsf{SR}^{\nu}, and that one could show this by establishing the analogue of Theorem 1.9.

We add that we do not know the answer to the following:

Question 1.13.

Do Theorem 1.8 and Theorem 1.11 hold for p=1p=1?

The proof of the former uses Hölder at one place (cf. equation (9.1)), and the latter uses Doob’s Maximal Inequality (cf. Lemma 5.1(1)).

1.4. Relation to previous work

Theorem 1.5 generalises the result of Pathak, Rojas, and Simpson, who show it for the specific case of p=1p=1 and X=[0,1]kX=[0,1]^{k}, ν\nu being the kk-fold product of Lebesgue measure on [0,1][0,1] with itself, and with n\mathscr{F}_{n} being given by dyadic partitions.353535[52]. They state their result not in terms of Lp(ν)L_{p}(\nu) Schnorr tests, but in terms of computable points of Lp(ν)L_{p}(\nu). See §11. Under this guise, Lévy’s Upward Theorem just is the Lebesgue Differentiation Theorem. Their proof goes through Tarski’s decidability results on the first-order theory of the reals, and so seems in certain key steps specific to the reals with Lebesgue measure.363636Such as at [52, Lemma 3.3 p. 339]. However, see Proposition 4.1 below, which generalises rather directly from their setting to the general setting.

In conjunction with the properties of effective disintegrations (cf. §7), one can derive the equivalence of (1)-(2) in Theorem 1.5 from results of Rute.373737In particular, for the (1) to (2) direction of Theorem 1.5, see Rute’s “Effective Levy 0/1 law” [61, Theorem 6.3 p. 31]. Our Proposition 7.5 and Proposition 2.4 implies that if ff is an L1(ν)L_{1}(\nu) Schnorr test, then 𝔼ν[fn]\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}] is a computable point of L1(ν)L_{1}(\nu), and so Rute’s Theorem 6.3 applies, once one internalises how to translate back and forth between Lp(ν)L_{p}(\nu) Schnorr tests and computable points of Lp(ν)L_{p}(\nu) (cf. §11). For the (2) to (1) direction of Theorem 1.5, see Rute’s [61, Example 12.1 p. 31]. As for Schnorr randomness, our work then expands on Rute’s primarily by finding a large class of versions of conditional expectations to which his results apply, and by identifying the information on rates of convergence in Theorem 1.6. More generally, the theory we develop is organised around the elementary concept of an integral Schnorr test, and so we hope might be of value to others by virtue of being accessible.383838In particular, we can avoid appeal to Rute’s theory of a.e. convergence, which is an alternative way to organise effective convergence in L0(ν)L_{0}(\nu) (cf. §2.4). See [61, Proposition 3.15 p. 15] and his Convergence Lemma [61, Lemma 3.19 p. 17].

Further, we are able to strengthen what is, in our view, one of the more foundationally significant parts of Rute’s work. He notes that traditionally “algorithmic randomness is more concerned with success than convergence” and that “only computable randomness has a well-known characterisation in terms of martingale convergence instead of martingale success.”393939[61, p. 7]. He is referring to what is called a “folklore” characterisation of computable randomness on Cantor space with the uniform measure in [15, Theorem 7.1.3 p. 270]. In Cantor space with the uniform measure, Rute has a characterisation of Schnorr randomness in terms of convergence of L2(ν)L_{2}(\nu) martingales.404040See items (1), (4) in his Example 1.5, immediately below the preceding quotation. We have been able to generalise this to all computable measures on computable Polish spaces: see Theorem 12.2. This proof follows Rute’s L2(ν)L_{2}(\nu) Hilbert space proof in broad outline. It seems to us that keeping track of the maximal function, which we can then use in DCT arguments, has been helpful here.

In the setting of Cantor space with the uniform measure and the natural filtration of length nn-strings and the natural disintegration (cf. Example B.1), our Theorem 1.8 was already known for p=1p=1 and hence all computable p1p\geq 1. This result is in a paper of Miyabe, Nies, and Zhang, where it is attributed to the Madison group of Andrews, Cai, Diamondstone, Lempp and Miller.414141[47, Theorem 3.3 p. 312]. Their proof is a little more general, in that it just concerns martingale convergence rather than martingales associated to random variables. Their proof, in the Cantor space setting, also can be modified to give not only convergence but convergence to the truth for random variables. Their argument goes through an auxiliary test notion of Madison test. While we only have it for computable p>1p>1, our proof of Theorem 1.8 goes through first principles about density randomness and effective disintegrations. It is not presently clear to us whether the Cantor space proof using Madison tests can be generalised to arbitrary computable probability measures on computable Polish spaces equipped with effective disintegrations.424242As a final remark about the previous literature, we should mention that Lévy’s Upward Theorem has also been studied in the context of Shafer and Vovk’s game-theoretic probability ([67], [66, Chapter 8]). Their approach conceives of martingales primarily as game-theoretic strategies, and does not treat computational matters explicitly. By contrast, here we are focusing on the martingales 𝔼ν[fn]\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}] and on effective properties of them conceived of as sequences of random variables. Discerning the relation between our approach and their approach would involve, as a first step, carefully going through their approach and ascertaining the exact levels of effectivity needed to secure their results, and secondly translating back and forth between the strategy and random variable paradigms.

We hope our efforts brings this prior important work on algorithmic randomness to the attention of a broader audience. Bayesianism is an important and increasingly dominant framework in a variety of disciplines, and this prior work and our work hopefully makes vivid the way in which computability theory and algorithmic randomness bears directly on the question of when and how fast Bayesian inductive methods converge to the truth. As its proof makes clear, the non-computable rate of convergence to the truth in Theorem 1.9 is another presentation of the halting set, and so this proof gives the theory of computation a central role in a limitative theorem of inductive inference, similar to its central role in the great limitative theorems of deductive inference like the Incompleteness Theorems. Further, the existence of Schnorr random worlds that are computably dominated, as in Theorem 1.6(2), shows that a central kind of randomness is entirely compatible with there being effective ways of determining how close we are to the truth. This optimistic inductive possibility is not one that would be visible in absence of recent work in algorithmic randomness.

Internal to the discussion about Bayesianism within philosophy, authors such as Belot have voiced the concern that the classical theory only tells us that worlds at which we fail to converge to the truth have probability zero, but otherwise tells us little about when and where the failure happens.434343[2]. From the perspective of Theorems 1.5, 1.8, 1.11, the probability zero event of non-convergence is not arbitrary, so long as one is insisting on convergence along a broad enough class of effective random variables. Namely, the sequences along which convergence to the truth fails for some element of this class are exactly those that are not random with respect to the underlying computable prior probability measure. In other words, those sequences can be determined by effective means to be atypical from the agent’s point of view.

Finally, we should emphasise that ours is not the only perspective on conditional expectations and its effectivity that one could adopt. In focusing on disintegrations, we are presupposing a framework where pointwise there is a single “formula” for the conditional expectation, namely the one displayed in equation (1.1) (and again see Appendicies A-B for examples). Likewise, the effectivity constraints in Definition 1.3 have the consequence that the conditional expectation operator is a continuous computable function (cf. Proposition 7.5), and so sends computable points to computable points (cf. Proposition 2.4). Both of these presuppositions constrain the applicability of our framework. For instance, Rao points out that conditional expectations are used throughout econometrics, but there one often uses the Dynkin-Doob Lemma as definitional of the conditional expectation,444444[55, 376]. For an example, see the presentation of conditional expectation in [20, Chapter 7]. and there is no more hope of having a single formula come out of it than there is of having all variables expressible in linear terms of one another. Likewise, conditional expectations and martingales can be used to prove theorems like the Radon-Nikodym Theorem,454545[75, p. 145-146]. which is “computably false” in that there are computable absolutely continuous probability measures with no computable Radon-Nikodym derivative.464646[70, p. 396], [77], [31]. That, of course, is not to say that these are not of interest or that determining how non-effective they are is not of interest, but just to say they will not be available in a framework like ours where we restrict to computable continuous conditional expectation operators.474747The paper Ackerman et. al. [1] is an important recent paper studying how non-effective, in general, it is to have disintegrations. Our Definition 1.3, by contrast, restricts attention to those disintegrations that are highly effective. This will not be all of them, and we do not claim that it would be all of the interesting ones.

1.5. Outline of paper

The paper is organised as follows. In §2, we begin with a brief discussion of some aspects of effectively closed sets and computable continuous functions and lsc functions which we need for our proof, and then we go over relevant aspects of the three computable Polish spaces which are central for effective probability theory:

  • The computable Polish space +(X)\mathcal{M}^{+}(X) of non-negative finite Borel measures on XX and its computable Polish subspace 𝒫(X)\mathcal{P}(X) of probability measures.

  • For each computable ν\nu in +(X)\mathcal{M}^{+}(X) and each computable p1p\geq 1, the computable Polish space Lp(ν)L_{p}(\nu).

  • For each computable ν\nu in +(X)\mathcal{M}^{+}(X) the computable Polish space L0(ν)L_{0}(\nu) of Borel measurable functions which are finite ν\nu-a.s. and whose topology is given by convergence in measure.

The space L0(ν)L_{0}(\nu) is needed since when ff is in Lp(ν)L_{p}(\nu), the maximal function f=supn𝔼ν[fn]f^{\ast}=\sup_{n}\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}] is in L0(ν)L_{0}(\nu), and is guaranteed to be in Lp(ν)L_{p}(\nu) iff p>1p>1. In addition to being needed in the proofs of the main theorems, the material in §2 also can serve to contextualise many components of Definition 1.1. For instance, we mention in §2.2 a result of Hoyrup-Rojas that an element of 𝒫(X)\mathcal{P}(X) is computable in the sense of Definition 1.1(3) iff it is computable as an element of the Polish space 𝒫(X)\mathcal{P}(X). Likewise, in §2.3 we mention a result saying that an Lp(ν)L_{p}(\nu) Schnorr test is simply a non-negative lsc function whose equivalence class is a computable element of Lp(ν)L_{p}(\nu) (cf. Proposition 2.16). Finally, towards the close of §2.4, we define an L0(ν)L_{0}(\nu) Schnorr test and prove a new characterisation of 𝖲𝖱ν\mathsf{SR}^{\nu} in terms of these tests (cf. Definition 2.28, Proposition 2.29).

In §3 we present two lemmas on Schnorr randomness. The second of these, called the Self-location Lemma (3.3) is a distinctive feature of Schnorr randomness (vis-à-vis the other algorithmic randomness notions), and is central to our proof of Theorem 1.6. In §4 we present some results on recovering the pointwise values of effective random variables on 𝖲𝖱ν\mathsf{SR}^{\nu}. In §5, we review various classical features of the maximal function which we shall need later. In §6 we present an abstract treatment of Theorem 1.5 in terms of various effective constraints that a version of the conditional expectation may satisfy. In §7 we develop the fundamental properties of effective disintegrations. In §8, we prove Theorems 1.5-1.6, and in §9 we prove Theorems 1.8-1.9 and in §10 we prove Theorem 1.11. In §11, we show how Miyabe’s translation method allows us to recast Theorem 1.5 in terms of computable points of Lp(ν)L_{p}(\nu). In §12, we develop the theory of martingales in L2(ν)L_{2}(\nu) and prove the aforementioned generalisation of Rute’s result characterising Schnorr randomness in terms of martingale convergence. In Appendix A we briefly exposit two classical examples of disintegrations, and in Appendix B we present several examples of effective disintegrations.

In a sequel to this paper, we present a similar analysis of the Blackwell-Dubins Theorem,484848[5] which is also a “convergence to the truth” result, but wherein the pair “agent and world” is replaced with a pair of agents whose credences are variously absolutely continuous with respect to one another.

2. Computable Polish spaces for effective probability theory

2.1. Effectively closed, computable continuous, and lsc

In this section, we briefly describe two further concepts from the theory of computable Polish spaces: namely effectively closed subsets and computable continuous functions, and we close by mentioning a few brief aspects of lsc functions.

Before we do that, we mention one elementary proposition on computable Polish spaces which is worth having in hand (for e.g. the Self-location Lemma 3.3):

Proposition 2.1.

Let XX be a computable Polish space with metric dd and countable dense set x0,x1,x_{0},x_{1},\ldots. Suppose the map in(i)i\mapsto n(i) is such that xn(i)xx_{n(i)}\rightarrow x fast. Then

  1. (1)

    The set {(j,q)×>0:xB(xj,q)}\{(j,q)\in\mathbb{N}\times\mathbb{Q}^{>0}:x\in B(x_{j},q)\} is c.e. in graph of in(i)i\mapsto n(i).

  2. (2)

    The point xx is computable iff the set {(j,q)×>0:xB(xj,q)}\{(j,q)\in\mathbb{N}\times\mathbb{Q}^{>0}:x\in B(x_{j},q)\} is c.e.

Proof.

For (1), since the distances between the points of the countable dense set is uniformly computable, they are also uniformly right-c.e. Hence, it suffices to note that d(xj,x)<qd(x_{j},x)<q iff there is i0i\geq 0 with d(xj,xn(i))<q2id(x_{j},x_{n(i)})<q-2^{-i}.

For (2), if xx is computable, then we can choose in(i)i\mapsto n(i) computable, and then are done by (1). Conversely, if the set is c.e., given i0i\geq 0, enumerate it until one finds a pair (j,q)(j,q) with q2iq\leq 2^{-i}, and set n(i)=jn(i)=j. ∎

As mentioned in §1, the complement of a c.e. open set is called an effectively closed set. In Cantor space and Baire space, the effectively closed sets can be represented as paths through computable trees.494949[9, p. 41]. The following provides a simple example on the real line of a classically closed set which is not effectively closed:

Example 2.2.

Suppose that c<dc<d and cc is right-c.e. and dd is left-c.e but neither c,dc,d are computable. Then [c,d][c,d] is a computable Polish space. Further, [c,d][c,d] is a classically closed subset of the reals which is not an effectively closed subset of the reals.

It is a computable Polish space since its countable dense set (c,d)(c,d)\cap\mathbb{Q} is c.e. since it is the intersection of the the right Dedekind cut of cc and the left Dedekind cut of dd. If [c,d][c,d] were effectively closed in the reals then U=(,c)(d,)U=(-\infty,c)\cup(d,\infty) would be c.e. open in the reals. Then by choosing a rational rr in (c,d)(c,d), one has that {q:q<c}={qU:q<r}\{q\in\mathbb{Q}:q<c\}=\{q\in U:q<r\} is c.e., contrary to hypothesis.

Effectively closed subsets of a computable Polish space need not themselves have the structure of a computable Polish space, since one in addition needs to produce an enumeration of a countable dense set where the distance between the points is uniformly computable.505050By contrast, classically, the Polish subspaces of a Polish space are precisely the GδG_{\delta} subsets. See [34, p. 17]. Hence, we define: a computable Polish subspace YY of XX is given by a an effectively closed subset YY of XX and a countable sequence of points y0,y1,y_{0},y_{1},\ldots which are uniformly computable points of XX and which are dense in YY. One can check that the c.e. opens relative to YY are just the c.e. opens of the space XX intersected with YY, and further any effectively closed subset of YY is also an effectively closed subset of XX. Similarly, one can check that a computable point of YY is just a computable point of XX which happens to be in YY. As a simple example of a Polish subspace which is not a computable Polish subspace, one has:

Example 2.3.

Suppose that a<ba<b and aa is left-c.e. and bb is right-c.e but neither a,ba,b are computable. Then the closed interval [a,b][a,b] is an effectively closed subset of the reals which is not a computable Polish subspace of the reals.

It is effectively closed since aa being left-c.e. and bb being right-c.e. implies that (,a)(-\infty,a) and (b,)(b,\infty) are c.e. open. And if [a,b][a,b] were a computable Polish subspace of the reals, and if y0,y1,y_{0},y_{1},\ldots were a sequence of uniformly computable reals dense in [a,b][a,b], then for a rational qq we would have a<qa<q iff there is ii such that yi<qy_{i}<q, which is a c.e. condition and so aa would be right-c.e. and thus computable.

An effectively closed set CC is computably compact if there is a partial computable procedure which, when given an index for a computable sequence of c.e. opens U0,U1,U_{0},U_{1},\ldots in XX which covers CC, returns a natural number n0n\geq 0 such that U0,,UnU_{0},\ldots,U_{n} covers CC. We further say that CC is strongly computably compact if there is a partial computable procedure which, when given an index for a computable sequence of c.e. opens U0,U1,U_{0},U_{1},\ldots in XX halts iff this is a cover of CC, and when it halts returns a natural number n0n\geq 0 such that U0,,UnU_{0},\ldots,U_{n} covers CC. If XX itself is strongly computably compact, then so are all of its effectively closed sets. If c<dc<d is computable, then [c,d][c,d] is strongly computably compact. Likewise, Cantor space is strongly computably compact, and if f:f:\mathbb{N}\rightarrow\mathbb{N} is computable then the computably bounded set {ω:nω(n)f(n)}\{\omega\in\mathbb{N}^{\mathbb{N}}:\forall\;n\;\omega(n)\leq f(n)\} is a computable Polish subspace of Baire space which is strongly computably compact. Example 2.2 is an example of a compact computable Polish space which is not computably compact, since if it were then we could compute the endpoints using maxs and mins of the centres of finite coverings with fast decreasing radii. For another example of a compact computable Polish space which is not computably compact, one can take the paths through a computable subtree of Baire space which is not computably bounded.515151See [10, Example 2.1.5 p. 59].

If X,YX,Y are two computable Polish spaces, then a function f:XYf:X\rightarrow Y is computable continuous if inverse images of c.e. opens are uniformly c.e. open.525252See Moschovakis [48, 110]. Simpson [70, Exercise II.6.9 p. 88] notes that it is equivalent to his preferred definition at [70, 85]. The following characterisation usefully parameterises each continuous computable function by a single c.e. set, where it is assumed for the sake of simplicity that both countable dense sets are identified with the natural numbers:535353This is from [26, 1169]. It can be seen as a simplification of Simpson’s definition in [70, 85].

  • A function f:XYf:X\rightarrow Y is computable continuous iff there is a c.e. set I×>0××>0I\subseteq\mathbb{N}\times\mathbb{Q}^{>0}\times\mathbb{N}\times\mathbb{Q}^{>0} such that both (i) if (i,p,j,q)(i,p,j,q) is in II then B(i,p)f1(B(j,q))B(i,p)\subseteq f^{-1}(B(j,q)) and (ii) for all xx in XX and all ϵ>0\epsilon>0 there is (i,p,j,q)(i,p,j,q) in II with xx in B(i,p)B(i,p) and q<ϵq<\epsilon.

Computable continuous maps are also computable continuous when restricted to computable Polish subspaces. The computable continuous maps preserve computability of points:

Proposition 2.4.

If f:XYf:X\rightarrow Y is computable continuous and xx in XX is computable, then f(x)f(x) in YY is computable.

Proof.

Suppose that xx is computable. Then {(i,q)×>0:d(i,x)<q}\{(i,q)\in\mathbb{N}\times\mathbb{Q}^{>0}:d(i,x)<q\} is c.e. by Proposition 2.1 (1). For each n0n\geq 0, by (ii) above, search in II for a tuple (in,pn,jn,qn)(i_{n},p_{n},j_{n},q_{n}) with xx in B(in,pn)B(i_{n},p_{n}) and qn<2nq_{n}<2^{-n}. Then by (i) above, f(x)f(x) is in f(B(in,pn))B(jn,qn)f(B(i_{n},p_{n}))\subseteq B(j_{n},q_{n}) and so jnf(x)j_{n}\rightarrow f(x) fast. ∎

This proposition is important because many arguments for the computability of points in effective analysis and probability can be seen as the result of applying computable continuous functions to computable points.

There is a partial converse to the previous proposition in the uniformly continuous setting. A computable modulus of uniform continuity for a uniformly continuous function f:XYf:X\rightarrow Y is a computable function m:>0>0m:\mathbb{Q}^{>0}\rightarrow\mathbb{Q}^{>0} such that d(x,x)<m(ϵ)d(x,x^{\prime})<m(\epsilon) implies d(f(x),f(x))<ϵd(f(x),f(x^{\prime}))<\epsilon for all ϵ\epsilon in >0\mathbb{Q}^{>0}. For instance, if c>0c>0 is rational, then a cc-Lipschitz function is just a function with linear modulus of uniform continuity m(ϵ)=c2ϵm(\epsilon)=\frac{c}{2}\cdot\epsilon. The partial converse to Proposition 2.4 is the following:

Proposition 2.5.

Suppose X,YX,Y are computable Polish spaces and that f:XYf:X\rightarrow Y has a computable modulus of uniform continuity. Suppose that the image of the countable dense set in XX under ff is uniformly computable in YY. Then ff is computable continuous.

Proof.

Suppose xnx_{n} is the countable dense set in XX. Suppose that m:>0>0m:\mathbb{Q}^{>0}\rightarrow\mathbb{Q}^{>0} is the computable modulus of uniform continuity. Suppose that yn,if(xn)y_{n,i}\rightarrow f(x_{n}) fast, where yn,iy_{n,i} is a uniformly computable sequence from the countable dense set in YY. Then define the c.e. set I={(xn,m(2i),yn,i,2i+1):n,i0}I=\{(x_{n},m(2^{-i}),y_{n,i},2^{-i+1}):n,i\geq 0\}. First we show that B(xn,m(2i))f1(B(yn,i,2i+1))B(x_{n},m(2^{-i}))\subseteq f^{-1}(B(y_{n,i},2^{-i+1})). For, suppose that xx is in B(xn,m(2i))B(x_{n},m(2^{-i})). Then d(x,xn)<m(2i)d(x,x_{n})<m(2^{-i}). Then d(f(x),f(xn))<2id(f(x),f(x_{n}))<2^{-i}. Further since yn,if(xn)y_{n,i}\rightarrow f(x_{n}) fast, we have d(f(xn),yn,i)2id(f(x_{n}),y_{n,i})\leq 2^{-i}, from which we obtain d(f(x),yn,i)<2i+1d(f(x),y_{n,i})<2^{-i+1} by triangle inequality. Second suppose that xx is in XX and ϵ>0\epsilon>0. Let i0i\geq 0 be such that 2i+1<ϵ2^{-i+1}<\epsilon. Since xnx_{n} is an enumeration of the countable dense set, there is xnx_{n} such that d(x,xn)<m(2i)d(x,x_{n})<m(2^{-i}). Then xx is in B(xn,m(2i))B(x_{n},m(2^{-i})), and the tuple (xn,m(2i),yn,i,2i+1)(x_{n},m(2^{-i}),y_{n,i},2^{-i+1}) is in II and 2i+1<ϵ2^{-i+1}<\epsilon. ∎

This proposition is widely applicable in our context since many operators in functional analysis are uniformly continuous, and since one often in practice has good control over what happens with the countable dense set (see the proofs of Proposition 2.16 and Proposition 2.21 for representative examples).

Finally, recall the notion of core notion of lsc from Definition 1.1(1), which is the effectivization of the classical notion of lower semi-continuous. This class of functions has some paradigmatic examples and useful closure conditions which we briefly enumerate without proof:

Proposition 2.6.

Constant functions that are left-c.e. reals are lsc. Indicator functions of c.e. opens are lsc.

Lsc functions are closed under addition, maxs and mins. Non-negative lsc functions are closed under multiplication.

Sups of uniformly lsc functions are lsc. Infinite sums of non-negative lsc functions are lsc. Compositions of lsc functions with computable continuous functions are lsc.

Since ff is lsc iff f-f is usc, one can use this proposition to obtain examples and closure conditions for usc functions as well.

The analogue of Proposition 2.4 for lsc functions is that they send computable points to left-c.e. reals.

2.2. The space of probability measures

If XX is a Polish space, then the space of real-valued finite signed Borel measures on XX is written as (X)\mathcal{M}(X). Recall that the weak-topology on (X)\mathcal{M}(X) is the smallest topology such that all the linear maps νXf𝑑ν\nu\mapsto\int_{X}f\;d\nu are continuous, where ff ranges over bounded continuous functions on the space.545454Or, equivalently, as ff ranges over all bounded uniformly continuous functions on the space ([34, 110]). Unless the space XX is finite, the weak-topology on (X)\mathcal{M}(X) is not metrizable.555555[6, 17, 102] However, when XX is a Polish space, the space 𝒫(X)\mathcal{P}(X) of all probability Borel measures on XX with the weak-topology is a Polish space, as is the space +(X)\mathcal{M}^{+}(X) of all finite non-negative Borel measures on XX.565656[34, §17.E pp. 109 ff]. Convergence in 𝒫(X)\mathcal{P}(X) is characterised by the Portmanteau Theorem.575757[34, Theorem 17.20 p. 111], [4, Theorem 2.1 p. 16].

A natural countable dense set on the spaces 𝒫(X)\mathcal{P}(X) and +(X)\mathcal{M}^{+}(X) are the finite averages of Dirac measures associated to points from the countable dense set on XX, with rational values for the weights. These spaces can be completely metrized by the metric of Prohorov. However, when working on 𝒫(X)\mathcal{P}(X), it is often more useful to work with the Wasserstein metric, and further when the metric on XX is unbounded it is more convenient to work with the Kantorovich-Rubinshtein metric on 𝒫(X)\mathcal{P}(X):585858[6, 104,111].

dKR(ν,μ)=sup{|𝔼νf𝔼μf|:f is 1-Lipschitz&f1}d_{KR}(\nu,\mu)=\sup\{\left|\mathbb{E}_{\nu}f-\mathbb{E}_{\mu}f\right|\;:f\mbox{ is $1$-Lipschitz}\;\&\;\|f\|_{\infty}\leq 1\}

In this, f=supxX|f(x)|\|f\|_{\infty}=\sup_{x\in X}\left|f(x)\right|. Hoyrup and Rojas prove that 𝒫(X)\mathcal{P}(X) with the metrics of Prohorov or Wasserstein are computable Polish spaces, and their proof extends naturally to the Kantorovich-Rubinshtein metric. Likewise, their proof shows that +(X)\mathcal{M}^{+}(X) is a computable Polish space and has 𝒫(X)\mathcal{P}(X) as a computable Polish subspace.595959[29, 49], [30, 838]. Further, Hoyrup and Rojas characterise the computable points in 𝒫(X)\mathcal{P}(X) as follows, and their proof extends naturally to +(X)\mathcal{M}^{+}(X):606060[29, 52], [30, 839].

Proposition 2.7.

A point ν\nu in +(X)\mathcal{M}^{+}(X) is computable iff ν(X)\nu(X) is computable and ν(U)\nu(U) is uniformly left-c.e. for c.e. opens UU in XX.

This proposition helps motivate Definition 1.1(3).

To illustrate the utility of this proposition, consider [c,d][c,d] from Example 2.2. This proposition implies that Lebesgue measure mm on [c,d][c,d] is not a computable point of +(X)\mathcal{M}^{+}(X) since m([c,d])=d-cm([c,d])=d\mbox{-}c is left-c.e. but not computable. Likewise, 1dcm\frac{1}{d-c}m is not a computable point of 𝒫(X)\mathcal{P}(X) since one can choose rationals q,ϵq,\epsilon such that (qϵ,q+ϵ)(c,d)(q-\epsilon,q+\epsilon)\subseteq(c,d), and then one has 1dcm((qϵ,q+ϵ))=2ϵdc\frac{1}{d-c}m((q-\epsilon,q+\epsilon))=\frac{2\epsilon}{d-c} is right-c.e. but not computable.

Recall the notion of a comptuable basis and measure computable basis from Definition 1.1(9)-(10). Hoyrup and Rojas use an effective version of the Baire Category Theorem to prove every ν\nu is a computable point of +(X)\mathcal{M}^{+}(X) has a ν\nu-computable basis. Moreover, the basis can be taken to be open balls B(i,rj)B(i,r_{j}) with centres ii from the countable dense set and with radii given by a dense computable sequence rjr_{j} of non-zero reals, with the closed balls B[i,rj]B[i,r_{j}] being the corresponding effectively closed supersets.616161See [30, Corollary 5.2.1 p. 844], [29, Theorem 2.2.1.2 p. 60]. Rute also employs this result of Hoyrup and Rojas, [61, pp. 13-14], although he leaves out from the definition of a measure computable basis the pairing of each basis UU element with an effectively closed superset CC of the same measure. Hoyrup and Rojas include this pairing, but further require that U(XC)U\cup(X\setminus C) is dense (cf. [30, Definition 5.1.2 p. 842], [29, Definition 2.2.1.2 p. 58]).

The following proposition summarises some basic properties of ν\nu-computable bases:

Proposition 2.8.

Suppose that ν\nu is a computable point of +(X)\mathcal{M}^{+}(X).

  1. (1)

    Elements of the algebra generated by a ν\nu-computable basis uniformly have ν\nu-computable measure. Indeed, this holds for all holds for all sequences B0,B1,B_{0},B_{1},\ldots of events such that finite unions of them have uniformly ν\nu-computable measure.

  2. (2)

    If a computable sequence of c.e. opens with uniformly computable ν\nu-measure is added to a ν\nu-computable basis, then finite unions from the resulting sequence have uniformly computable ν\nu-measure, as do elements from the algebra generated by the resulting sequence. Indeed, this holds for all computable bases such that finite unions of them have uniformly ν\nu-computable measure.

  3. (3)

    If a computable sequence of c.e. opens with uniformly computable ν\nu-measure and with uniformly effectively closed supersets of the same ν\nu-measure is added to a ν\nu-computable basis, then the result is a ν\nu-computable basis.

  4. (4)

    The ν\nu-computable bases are closed under effective union.

Proof.

For (1), suppose that B0,B1,B_{0},B_{1},\ldots is a sequence of events such that finite unions of them have uniformly ν\nu-computable measure.

First note that finite intersections Bi0BinB_{i_{0}}\cap\cdots\cap B_{i_{n}} have uniformly computable ν\nu-measure: this is an induction on n1n\geq 1, and for the induction step, use inclusion-exclusion ν(Bi0Bin)=ν(Bi0Bin)+J{0,,n}(1)|J|1ν(jJBij)\nu(B_{i_{0}}\cap\cdots\cap B_{i_{n}})=-\nu(B_{i_{0}}\cup\cdots\cup B_{i_{n}})+\sum_{\emptyset\neq J\subsetneq\{0,\ldots,n\}}(-1)^{\left|J\right|-1}\nu(\bigcap_{j\in J}B_{i_{j}}).

Likewise, finite unions A1AmA_{1}\cup\cdots\cup A_{m} of finite intersections Aj=k=1jBij,kA_{j}=\bigcap_{k=1}^{\ell_{j}}B_{i_{j,k}} of members of BiB_{i} have uniformly computable ν\nu-measure: this is an induction on m1m\geq 1, and for the induction step, use distribution ν(A1Am+1)=ν(A1Am)+ν(Am+1)ν((A1Am+1)(AmAm+1))\nu(A_{1}\cup\cdots\cup A_{m+1})=\nu(A_{1}\cup\cdots\cup A_{m})+\nu(A_{m+1})-\nu((A_{1}\cap A_{m+1})\cup\cdots\cup(A_{m}\cap A_{m+1})).

Finally, note that finite intersections of members of BiB_{i} and complements of members of BiB_{i} have uniformly computable ν\nu-measure. This follows from the previous steps, the elementary identity ν(CD)=ν(C)ν(CD)\nu(C\setminus D)=\nu(C)-\nu(C\cap D) and distribution as follows when n>0n>0: ν(Bi0Bin1(XBin)(XBin+m1))=ν(Bi0Bin1)ν(Bi0Bin1(BinBin+m1))=ν(Bi0Bin1)ν(j=nn+m1Bi0Bin1Bij)\nu(B_{i_{0}}\cap\cdots\cap B_{i_{n-1}}\cap(X\setminus B_{i_{n}})\cap\cdots\cap(X\setminus B_{i_{n+m-1}}))=\nu(B_{i_{0}}\cap\cdots\cap B_{i_{n-1}})-\nu(B_{i_{0}}\cap\cdots\cap B_{i_{n-1}}\cap(B_{i_{n}}\cup\cdots\cup B_{i_{n+m-1}}))=\nu(B_{i_{0}}\cap\cdots\cap B_{i_{n-1}})-\nu(\bigcup_{j=n}^{n+m-1}B_{i_{0}}\cap\cdots\cap B_{i_{n-1}}\cap B_{i_{j}}), which is uniformly computable by the two previous paragraphs. If n=0n=0 then note that ν((XBin)(XBin+m1))=ν(X(XBin)(XBin+m1))\nu((X\setminus B_{i_{n}})\cap\cdots\cap(X\setminus B_{i_{n+m-1}}))=\nu(X\cap(X\setminus B_{i_{n}})\cap\cdots\cap(X\setminus B_{i_{n+m-1}})), and since ν(X)\nu(X) is computable, we can argue as in the case of n=1n=1 with XX playing the role of Bi0B_{i_{0}}.

For (2), suppose that U0,U1,U_{0},U_{1},\ldots is a computable sequence of c.e. opens such that ν(U0),ν(U1),\nu(U_{0}),\nu(U_{1}),\ldots is uniformly computable. Suppose that B0,B1,B_{0},B_{1},\ldots is a computable basis such that finite unions of them have uniformly ν\nu-computable measure. We must show that finite unions from B0,B1,,U0,U1,B_{0},B_{1},\ldots,U_{0},U_{1},\ldots have uniformly ν\nu-computable measure. It suffices to consider the case where U0,U1,U_{0},U_{1},\ldots just consists of a single c.e. open UiU_{i}, since by induction and (1) we may assume that the U1,,Ui1U_{1},\ldots,U_{i-1} are already among the B0,B1,B_{0},B_{1},\ldots. Since B0,B1,B_{0},B_{1},\ldots is a computable basis, write Ui=jBm(j)U_{i}=\bigcup_{j}B_{m(j)}, where mm is a computable function. Then ν(B1BnUi)=ν(jB1BnBm(j))=limkν(j<kB1BnBm(j))\nu(B_{1}\cup\cdots\cup B_{n}\cup U_{i})=\nu(\bigcup_{j}B_{1}\cup\cdots\cup B_{n}\cup B_{m(j)})=\lim_{k}\nu(\bigcup_{j<k}B_{1}\cup\cdots\cup B_{n}\cup B_{m(j)}). Since this limit is increasing, and ν(j<kB1BnBm(j))\nu(\bigcup_{j<k}B_{1}\cup\cdots\cup B_{n}\cup B_{m(j)}) is uniformly computable, we have that ν(B1BnUi)\nu(B_{1}\cup\cdots\cup B_{n}\cup U_{i}) is left-c.e. Similarly, ν((B1Bn)Ui)=ν(j(B1Bn)Bm(j))=limkν(j<k(B1Bn)Bm(j))\nu((B_{1}\cup\cdots\cup B_{n})\cap U_{i})=\nu(\bigcup_{j}(B_{1}\cup\cdots\cup B_{n})\cap B_{m(j)})=\lim_{k}\nu(\bigcup_{j<k}(B_{1}\cup\cdots\cup B_{n})\cap B_{m(j)}). Since this limit is increasing, and ν(j<k(B1Bn)Bm(j))\nu(\bigcup_{j<k}(B_{1}\cup\cdots\cup B_{n})\cap B_{m(j)}) is uniformly computable by (1), we have that ν((B1Bn)Ui)\nu((B_{1}\cup\cdots\cup B_{n})\cap U_{i}) is left-c.e. Then ν(B1BnUi)=ν(B1Bn)+ν(Ui)ν((B1Bn)Ui)\nu(B_{1}\cup\cdots\cup B_{n}\cup U_{i})=\nu(B_{1}\cup\cdots\cup B_{n})+\nu(U_{i})-\nu((B_{1}\cup\cdots\cup B_{n})\cap U_{i}) is also right-c.e. and hence computable.

Finally, (3) follows from (2) and the definition of a ν\nu-computable basis; and (4) follows directly from the uniformity in the proof of (3). ∎

Many of the canonical computable bases are measure computable bases:

Example 2.9.

If a computable basis on XX consists of sets which are also uniformly effectively closed, then the basis is ν\nu-computable for any computable point ν\nu of +(X)\mathcal{M}^{+}(X). This point applies to the canonical computable basis of clopens on Baire space or Cantor space.

Example 2.10.

If a computable basis on XX consists of c.e. open sets UU such that U¯\overline{U} is uniformly effectively closed with U¯U\overline{U}\setminus U is finite, then the basis is ν\nu-computable for any computable atomless ν\nu in +(X)\mathcal{M}^{+}(X). This point applies to the canonical atomless measures on [a,b][a,b] for a<ba<b computable.

Here is an example of a computable basis that is not a measure computable basis:

Example 2.11.

Let f:{0}f:\mathbb{N}\rightarrow\mathbb{N}\setminus\{0\} be an injective function whose range is c.e. but not computable, so that b=i2f(i)<1b=\sum_{i}2^{-f(i)}<1 is left-c.e. but not computable. Let qi=12(i+1)q_{i}=1-2^{-(i+1)}, which converges upwards to one, starting from 12\frac{1}{2}. Define a computable point ν\nu of 𝒫([0,1])\mathcal{P}([0,1]) by ν=(i2f(i)δqi)+(1b)δ1\nu=(\sum_{i}2^{-f(i)}\cdot\delta_{q_{i}})+(1-b)\cdot\delta_{1}. A computable basis for [0,1][0,1] is given by (p,q)[0,1](p,q)\cap[0,1] where p<qp<q are rationals. But this is not a ν\nu-computable basis since ν(0,1)=b\nu(0,1)=b is left-c.e. but not computable.

To illustrate the utility of measure computable bases, consider the following approximation method. In this proof, we use the standard notation WeW_{e} for the ee-th c.e. set, and we use We,sW_{e,s} for the points in WeW_{e} which get enumerated in by stage ss in the canonical enumeration.626262[71, pp. 17-18, 47].

Proposition 2.12.

Suppose ν\nu is a computable point of +(X)\mathcal{M}^{+}(X).

From a rational ϵ>0\epsilon>0 and an index for a c.e. open UU with ν(U)\nu(U) computable, one can uniformly compute an index for an effectively closed set CUC\subseteq U and an index for ν(C)\nu(C) as a computable real such that ν(UC)<ϵ\nu(U\setminus C)<\epsilon.

Proof.

We work with the ν\nu-computable basis B(i,rj)B(i,r_{j}) as above (discussed immediately before Proposition 2.8). Let ϵ>0\epsilon>0 rational be given. Suppose UU is c.e. open with ν(U)\nu(U) computable. Let U=kB(if(k),rf(k)))U=\bigcup_{k}B(i_{f(k)},r_{f(k)})) where ff is a computable function. For each m0m\geq 0 let Um=k<mB(if(k),rf(k))U_{m}=\bigcup_{k<m}B(i_{f(k)},r_{f(k)}). Note that UmU_{m} has ν\nu-computable measure, uniformly in m0m\geq 0. Using this and the computability of ν(U)\nu(U), compute m0m\geq 0 such that ν(U)ν(Um)<ϵ2\nu(U)-\nu(U_{m})<\frac{\epsilon}{2}. For each k<mk<m, the set Wg(k)={j:0<rj<rf(k)}W_{g(k)}=\{j:0<r_{j}<r_{f(k)}\} is c.e. and dense in the open interval (0,rf(k))(0,r_{f(k)}) and so B(if(k),rf(k))=jWg(k)B(if(k),rj)=jWg(k)B[if(k),rj]B(i_{f(k)},r_{f(k)})=\bigcup_{j\in W_{g(k)}}B(i_{f(k)},r_{j})=\bigcup_{j\in W_{g(k)}}B[i_{f(k)},r_{j}]. Compute s0s\geq 0 such that ν(Um)ν(k<mjWg(k),sB(if(k),rj))<ϵ2\nu(U_{m})-\nu(\bigcup_{k<m}\bigcup_{j\in W_{g(k),s}}B(i_{f(k)},r_{j}))<\frac{\epsilon}{2}. Then C=k<mjWg(k),sB[if(k),rj]C=\bigcup_{k<m}\bigcup_{j\in W_{g(k),s}}B[i_{f(k)},r_{j}] is a finite union of effectively closed sets and so effectively closed; and further ν(C)\nu(C) is a computable real since it is a finite union of elements from the ν\nu-computable basis. Further CUC\subseteq U and ν(U)ν(C)ν(U)ν(Um)+ν(Um)ν(C)<ϵ\nu(U)-\nu(C)\leq\nu(U)-\nu(U_{m})+\nu(U_{m})-\nu(C)<\epsilon. ∎

The following is an important property of the interaction of 𝖪𝖱ν\mathsf{KR}^{\nu} with ν\nu-computable bases:

Proposition 2.13.

Each element AA of the algebra generated by a ν\nu-computable basis is uniformly identical on 𝖪𝖱ν\mathsf{KR}^{\nu} to a c.e. open UU, which is effectively paired with an effectively closed superset CC of UU of the same ν\nu-measure.

Note that since CUC\setminus U is an effectively closed ν\nu-null set, C=UC=U on 𝖪𝖱ν\mathsf{KR}^{\nu}.

Proof.

Suppose that B0,B1,B_{0},B_{1},\ldots is a ν\nu-computable basis with corresponding effectively closed set CiBiC_{i}\supseteq B_{i} of the same ν\nu-measure. Again, since CiBiC_{i}\setminus B_{i} is an effectively closed ν\nu-null set, we have that Ci=BiC_{i}=B_{i} on 𝖪𝖱ν\mathsf{KR}^{\nu}. Then XCiX\setminus C_{i} is c.e. open with effectively closed superset XBiX\setminus B_{i} which with it agrees on 𝖪𝖱ν\mathsf{KR}^{\nu}.

Suppose that AA is an element of the algebra generated by the ν\nu-computable basis B0,B1,B_{0},B_{1},\ldots. Then AA can be written as the finite union of finite intersections of the B0,B1,B_{0},B_{1},\ldots and their relative complements XB0,XB1,X\setminus B_{0},X\setminus B_{1},\ldots. This is indexed by a finite list of pairs of strings σ1,τ1,,σn,τn\sigma_{1},\tau_{1},\ldots,\sigma_{n},\tau_{n} such that

A=i=1n(j<|σi|Bσi(j)j<|τi|XBτi(j))A=\bigcup_{i=1}^{n}\bigg{(}\bigcap_{j<\left|\sigma_{i}\right|}B_{\sigma_{i}(j)}\cap\bigcap_{j<\left|\tau_{i}\right|}X\setminus B_{\tau_{i}(j)}\bigg{)} (2.1)

Then form c.e. open VV by replacing the effectively closed XBτi(j)X\setminus B_{\tau_{i}(j)} with the c.e. open XCτi(j)X\setminus C_{\tau_{i}(j)}, and similarly form effectively closed DD by replacing c.e. open Bσi(j)B_{\sigma_{i}(j)} with effectively closed Cσi(j)C_{\sigma_{i}(j)}, as follows:

V=i=1n(j<|σi|Bσi(j)j<|τi|XCτi(j)),D=i=1n(j<|σi|Cσi(j)j<|τi|XBτi(j))V=\bigcup_{i=1}^{n}\bigg{(}\bigcap_{j<\left|\sigma_{i}\right|}B_{\sigma_{i}(j)}\cap\bigcap_{j<\left|\tau_{i}\right|}X\setminus C_{\tau_{i}(j)}\bigg{)},\hskip 8.53581ptD=\bigcup_{i=1}^{n}\bigg{(}\bigcap_{j<\left|\sigma_{i}\right|}C_{\sigma_{i}(j)}\cap\bigcap_{j<\left|\tau_{i}\right|}X\setminus B_{\tau_{i}(j)}\bigg{)} (2.2)

Then A,V,DA,V,D are equal on 𝖪𝖱ν\mathsf{KR}^{\nu} and hence have the same ν\nu-measure, and further VV is c.e. open and DVD\supseteq V is effectively closed.

The previous proposition places topological constraints on the sets in ν\nu-computable bases, at least when the measure has full support (that is, there are no open ν\nu-null sets):

Proposition 2.14.

Suppose that ν\nu is a computable point of 𝒫(X)\mathcal{P}(X).

  1. (1)

    If ν\nu has full support and 𝖷𝖱ν\mathsf{XR}^{\nu} is a ν\nu-measure one set and the c.e. open UU is equal to effectively closed CC on 𝖷𝖱ν\mathsf{XR}^{\nu}, then ν(U¯)=ν(U)\nu(\overline{U})=\nu(U).

  2. (2)

    If ν\nu has full support then no element UU of a ν\nu-computable basis can satisfy ν(U¯)>ν(U)\nu(\overline{U})>\nu(U).

In this, we use ¯\overline{\;\cdot\;} for topological closure.

Proof.

For (1), the c.e. open UCU\setminus C is a subset of the ν\nu-null X𝖷𝖱νX\setminus\mathsf{XR}^{\nu}. Since ν\nu has full support, we must have that UCU\setminus C is empty, so that UCU\subseteq C and U¯C\overline{U}\subseteq C. Since U,CU,C have same ν\nu-measure, the same must then be true of U,U¯U,\overline{U}. For (2), this follows from (1) and the previous proposition. ∎

By contrast, Proposition 2.8(3) implies any c.e. open UU with ν(U)\nu(U) computable and U¯\overline{U} effectively closed and ν(U¯)=ν(U)\nu(\overline{U})=\nu(U) can be added to any ν\nu-computable basis to form a larger ν\nu-computable basis.

For a simple example of c.e. open as in Proposition 2.14(2), one has the following:636363This example is a minor modification of an example from a proof in [9, p. 58].

Example 2.15.

Consider Cantor space with the uniform measure. Let 0=c0<c1<c2<0=c_{0}<c_{1}<c_{2}<\cdots be a computable sequence of natural numbers such that n2(cn+1cn)<\sum_{n}2^{-(c_{n+1}-c_{n})}<\infty (resp. is computable). Let II be any computable set. For all n0n\geq 0, consider the following clopen:

Un={ω:i[cn,cn+1)(ω(i)=1(n,i)I)}U_{n}=\{\omega:\forall\;i\in[c_{n},c_{n+1})\;\big{(}\omega(i)=1\leftrightarrow(n,i)\in I\big{)}\}

Since UnU_{n} makes decisions on cn+1cnc_{n+1}-c_{n} many bits, its measure is 2(cn+1cn)2^{-(c_{n+1}-c_{n})}. Then U=nUnU=\bigcup_{n}U_{n} is a c.e. open. And 0<ν(U)=nν(Unm<nUm)nν(Un)<0<\nu(U)=\sum_{n}\nu(U_{n}\setminus\bigcup_{m<n}U_{m})\leq\sum_{n}\nu(U_{n})<\infty (resp. is computable by the Comparison Test and the fact that Unm<nUmU_{n}\setminus\bigcup_{m<n}U_{m} is clopen, cf. Example 2.9 and Proposition 2.8( 1)). Further, the set UU is dense and so its closure is the entire space.

For a similar example on the unit interval with Lebesgue measure, one can use the complements of positive measure Cantor sets.

2.3. The space of integrable functions

For ν\nu a computable point of +(X)\mathcal{M}^{+}(X) and p1p\geq 1 computable, there is a natural Polish space structure on Lp(ν)L_{p}(\nu) (cf. Definition 1.2). For, one can take as the countable dense set the simple functions i=1nqiIAi\sum_{i=1}^{n}q_{i}\cdot I_{A_{i}}, where AiA_{i} come from the algebra of sets generated by a ν\nu-computable basis. If f,gf,g are two such functions, then so is fgf-g, and hence it suffices to show that if h=i=1nqiIAih=\sum_{i=1}^{n}q_{i}\cdot I_{A_{i}} is such a simple function then hp\|h\|_{p} is computable. Since AiA_{i} comes from an algebra, we can assume that the AiA_{i} are pairwise disjoint, which implies |i=1nqiIAi|p=i=1n|qi|pIAi\left|\sum_{i=1}^{n}q_{i}\cdot I_{A_{i}}\right|^{p}=\sum_{i=1}^{n}\left|q_{i}\right|^{p}\cdot I_{A_{i}} everywhere. Then one has hp=(i=1n|qi|pν(Ai))1p\|h\|_{p}=\big{(}\sum_{i=1}^{n}\left|q_{i}\right|^{p}\nu(A_{i})\big{)}^{\frac{1}{p}}, which is computable by Proposition 2.8(1). Note that the countable dense set is in 𝕃p(ν)\mathbb{L}_{p}(\nu), that is, it is defined everywhere rather than merely ν\nu-a.s. (cf. Definition 1.2). But when we pass to their equivalence classes, they become elements of Lp(ν)L_{p}(\nu), and they are a countable dense set in Lp(ν)L_{p}(\nu).

We do not record the choice of the ν\nu-computable basis in the notation for the computable Polish space Lp(ν)L_{p}(\nu). This is for two reasons. First, the ν\nu-computable bases are closed under effective union Proposition 2.8(4). Hence one can typically just assume that one is working with the union of whichever of them are salient in a given context. Second, one can check that any two ν\nu-computable bases result in computably homeomorphic presentations of Lp(ν)L_{p}(\nu).

Many of the natural continuous functions on the computable Polish space Lp(ν)L_{p}(\nu) are computable continuous, such as: addition, subtraction, multiplication by computable scalar, absolute value, maximum, minimum, positive part, and negative part.

By considering the continuous computable function Φ(f)=|f|f\Phi(f)=\left|f\right|-f, one sees that Lp+(ν)=Φ1({0})L_{p}^{+}(\nu)=\Phi^{-1}(\{0\}), and so Lp+(ν)L_{p}^{+}(\nu) is an effectively closed subset of Lp(ν)L_{p}(\nu) (cf. Definition 1.2). Further, it is a computable Polish subspace, since the equivalence classes of the non-negative elements of the countable dense set of Lp(ν)L_{p}(\nu) are dense in Lp+(ν)L_{p}^{+}(\nu).

Since we are working with a finite computable measure ν\nu from +(X)\mathcal{M}^{+}(X), if pqp\leq q, then the identity map is a computable continuous map from Lq(ν)L_{q}(\nu) into Lp(ν)L_{p}(\nu) and satisfies fpfq\|f\|_{p}\leq\|f\|_{q} for all ff from Lq(ν)L_{q}(\nu). We refer to this as the computable embedding of Lq(ν)L_{q}(\nu) into Lp(ν)L_{p}(\nu).

In working with Lp(ν)L_{p}(\nu) for p>1p>1 it is useful to remember the following inequalities:

u,v0:up+vp(u+v)p & (u+v)1pu1p+v1pu,v\geq 0:\hskip 14.22636ptu^{p}+v^{p}\leq(u+v)^{p}\hskip 8.53581pt\mbox{ \; \& \; }\hskip 8.53581pt(u+v)^{\frac{1}{p}}\leq u^{\frac{1}{p}}+v^{\frac{1}{p}} (2.3)

By letting u=xyu=x-y and v=yv=y, one obtains the following inequalities:

0yx:(xy)pxpyp & x1py1p(xy)1p0\leq y\leq x:\hskip 14.22636pt(x-y)^{p}\leq x^{p}-y^{p}\hskip 8.53581pt\mbox{ \; \& \; }\hskip 8.53581ptx^{\frac{1}{p}}-y^{\frac{1}{p}}\leq(x-y)^{\frac{1}{p}} (2.4)

The following proposition gives a canonical approximation of lsc functions which are bounded from below, and indicates that for the non-negative ones, being a computable point of Lp(ν)L_{p}(\nu) is solely a matter of the computability of the norm. The first part is due to Miyabe for p=1p=1.646464[44, Lemma 4.6]. This kind of approximation is a mainstay of working with lsc functions, and different approximations tend to be appropriate for different purposes.656565See [39, Definition 1.7.4 p. 35]. We will need a variation on this approximation in Proposition 7.10.

Proposition 2.16.

From a rational qq and a lsc function f:X[q,]f:X\rightarrow[q,\infty], one can compute an index for a computable sequence of functions fs:X[q,)f_{s}:X\rightarrow[q,\infty) from the countable dense set of L1(ν)L_{1}(\nu) such that fsfs+1f_{s}\leq f_{s+1} everywhere and f=supsfsf=\sup_{s}f_{s} everywhere.

Further, if p1p\geq 1 is computable, then a non-negative lsc function f:X[0,]f:X\rightarrow[0,\infty] in Lp(ν)L_{p}(\nu) is a Lp(ν)L_{p}(\nu) Schnorr test (cf. Definition 1.1(4)) iff it is a computable point of Lp(ν)L_{p}(\nu), and in this case the witness is a computable subsequence of fsf_{s}.

Finally, if p1p\geq 1 is computable, then any non-negative lsc function f:X[0,]f:X\rightarrow[0,\infty] in Lp(ν)L_{p}(\nu) is a Lp(ν)L_{p}(\nu) Martin-Löf test (cf. Definition 1.1(5)), and fsff_{s}\rightarrow f in Lp(ν)L_{p}(\nu).

Proof.

Let B0,B1,B_{0},B_{1},\ldots be a ν\nu-computable basis. Enumerate [q,)\mathbb{Q}\cap[q,\infty) as q0,q1,q_{0},q_{1},\ldots. For each n0n\geq 0, one has that f1(qn,]f^{-1}(q_{n},\infty] is uniformly c.e. open. Hence, there is a computable function gg such that f1(qn,)=iWg(n)Bif^{-1}(q_{n},\infty)=\bigcup_{i\in W_{g(n)}}B_{i}. Then define

fs(x)=max{q,qn:ns,iWg(n),s,xBi}f_{s}(x)=\max\{q,q_{n}:n\leq s,i\in W_{g(n),s},x\in B_{i}\} (2.5)

This is an element of the countable dense set of L1(ν)L_{1}(\nu) since we just enumerate nsWg(n),s\bigcup_{n\leq s}W_{g(n),s} as i0,,ik(s)i_{0},\ldots,i_{k(s)} and for each non-empty subset KK of {i0,,ik(s)}\{i_{0},\ldots,i_{k(s)}\} we consider the element BK=ijKBiijKXBiB_{K}=\bigcap_{i_{j}\in K}B_{i}\cap\bigcap_{i_{j}\notin K}X\setminus B_{i} of the algebra generated by the ν\nu-computable basis, and we let qK=max{qn:ns,ijK,ijWg(n),s}q_{K}=\max\{q_{n}:n\leq s,i_{j}\in K,i_{j}\in W_{g(n),s}\}, so that we have fs=K{i0,,ik(s)}qKIBKf_{s}=\sum_{\emptyset\neq K\subseteq\{i_{0},\ldots,i_{k(s)}\}}q_{K}\cdot I_{B_{K}}. Further, at the initial stages ss (if any) where nsWg(n),s\bigcup_{n\leq s}W_{g(n),s} is empty, we set fs=qIXf_{s}=q\cdot I_{X}.

Further from (2.5) one sees that fsfs+1f_{s}\leq f_{s+1} since the sum over which we taking the maximum grows in ss. Further, one has fsff_{s}\leq f everywhere since if we had fs(x)>f(x)f_{s}(x)>f(x), then fs(x)=qnf_{s}(x)=q_{n} for some nsn\leq s with iWg(n),si\in W_{g(n),s} and xx in BiB_{i}. But then Bif1(qn,]B_{i}\subseteq f^{-1}(q_{n},\infty], and so f(x)>qnf(x)>q_{n}. Finally, one has supsfs=f\sup_{s}f_{s}=f everywhere, since if not we would have supsfs(x)<qn<f(x)\sup_{s}f_{s}(x)<q_{n}<f(x) for some xx and some nn and hence xx would be in f1(qn,]=iWg(n)Bif^{-1}(q_{n},\infty]=\bigcup_{i\in W_{g(n)}}B_{i} and so xx would be in BiB_{i} for some ii in Wg(n)W_{g(n)} and hence there would be ss such that ii is in Wg(n),sW_{g(n),s} and hence by definition in (2.5) one would have that fs(x)qnf_{s}(x)\geq q_{n}.

Suppose p1p\geq 1 is computable and f:X[0,]f:X\rightarrow[0,\infty] is lsc and in Lp(ν)L_{p}(\nu). If ff is a computable point of Lp(ν)L_{p}(\nu), then since the norm is computable continuous (using Proposition 2.5), we have that fp\|f\|_{p} is computable (using Proposition 2.4). Conversely, suppose that ff is an Lp(ν)L_{p}(\nu) Schnorr test, so that fp\|f\|_{p} is computable. Then by taking pp-th roots, we have fp𝑑ν\int f^{p}\;d\nu is computable. Since fsp𝑑ν\int f_{s}^{p}\;d\nu converges upwards to fp𝑑ν\int f^{p}\;d\nu and we can compute both, we can compute a s(n)s(n) such that fpfs(n)pdν<2np\int f^{p}-f_{s(n)}^{p}\;d\nu<2^{-np} for all n0n\geq 0. Then using the estimate (ffs(n))pfpfs(n)p(f-f_{s(n)})^{p}\leq f^{p}-f_{s(n)}^{p} from (2.4), we have that (ffs(n))p𝑑ν<2np\int(f-f_{s(n)})^{p}\;d\nu<2^{-np}, and so by taking pp-th roots again we have ffs(n)p<2n\|f-f_{s(n)}\|_{p}<2^{-n}.

Similarly, for the last point, since fsp𝑑ν\int f_{s}^{p}\;d\nu converges upwards to fp𝑑ν\int f^{p}\;d\nu, we can use the estimate from (2.4) to argue that for all ϵ>0\epsilon>0 there is s00s_{0}\geq 0 such that for all ss0s\geq s_{0} one has (ffs)p𝑑νfp𝑑νfsp𝑑ν<ϵp\int(f-f_{s})^{p}\;d\nu\leq\int f^{p}\;d\nu-\int f_{s}^{p}\;d\nu<\epsilon^{p}, and so ffsp<ϵ\|f-f_{s}\|_{p}<\epsilon. ∎

The following records the “universal test” for 𝖬𝖫𝖱ν\mathsf{MLR}^{\nu}. For integral tests, it is due Gács and Hoyrup-Rojas in the case p=1p=1.666666Gács [22, 102, Corollary 3.3], Hoyrup-Rojas [30, 845-6]. Further, the version stated here is simplified in that it is only stated for a single measure, whereas these authors state a version where the lsc functions have domain 𝒫(X)×X\mathcal{P}(X)\times X. Hoyrup-Rojas improve on Gács by removing any assumption about the computability of the Boolean algebra structure on the algebra generated by the canonical computable basis.

Proposition 2.17.

Suppose ν\nu is a computable point of 𝒫(X)\mathcal{P}(X) and p1p\geq 1 is computable.

Then there is an Lp(ν)L_{p}(\nu) Martin-Löf test ff with fp1\|f\|_{p}\leq 1 such that for all Lp(ν)L_{p}(\nu) Martin-Löf tests gg with gp1\|g\|_{p}\leq 1 there is constant c>0c>0 such that gcfg\leq c\cdot f everywhere.

Hence 𝖬𝖫𝖱ν=nf1[0,n]\mathsf{MLR}^{\nu}=\bigcup_{n}f^{-1}[0,n], an increasing sequence of effectively closed sets.

Proof.

(Sketch) Enumerate the Lp(ν)L_{p}(\nu) Martin-Löf tests with pp-norm 1\leq 1 as h0,h1,h_{0},h_{1},\ldots. Do this by enumerating approximations to them (as in Proposition 2.16) which have pp-norm <1<1. Then set f=e2ehef=\sum_{e}2^{-e}\cdot h_{e}. ∎

The previous proposition has the following useful consequence regarding computable domination, which recall features in Theorem 1.5(2):

Proposition 2.18.

Suppose that XX is computably compact and ν\nu is a computable point of 𝒫(X)\mathcal{P}(X). Then there are points in 𝖬𝖫𝖱ν\mathsf{MLR}^{\nu} of computably dominated degree.

The main idea of the proof is to build a computably compact space of fast Cauchy sequences above XX, and to apply there the Computably Dominated Basis Theorem.676767[9, Theorem 3.7 p. 54], [71, Theorem 9.5.1 p. 179]. One can of course thematize the space of fast Cauchy sequences more than we are doing in this short paper, and in part what we are doing in the below proof is doing the construction out “by hand” in the computably compact case.

Proof.

Without loss of generality, we identify the countable dense set with the natural numbers. By effective Baire Category Theorem, choose a strictly decreasing computable sequence of positive reals ηs<2(s+1)\eta_{s}<2^{-(s+1)} such that {ηs:s0}{12d(i,j):i,j0}\{\eta_{s}:s\geq 0\}\cap\{\frac{1}{2}\cdot d(i,j):i,j\geq 0\} are disjoint. We define a non-decreasing computable sequence nsn_{s} of natural numbers as follows. Suppose that we have already defined things up to stage ss. To define at stage ss, we consider the open cover B(0,ηs),B(1,ηs),B(0,\eta_{s}),B(1,\eta_{s}),\ldots and use computable compactness to compute an nsns1n_{s}\geq n_{s-1} such that B(0,ηs),,B(ns,ηs)B(0,\eta_{s}),\ldots,B(n_{s},\eta_{s}) covers XX. Define the following computable trees:

T0\displaystyle T_{0} ={σ<:t<|σ|intσ(t)=i}\displaystyle=\{\sigma\in\mathbb{N}^{<\mathbb{N}}:\forall\;t<\left|\sigma\right|\;\exists\;i\leq n_{t}\;\sigma(t)=i\}
T\displaystyle T ={σT0:t<|σ|r[t,|σ|)d(σ(t),σ(r))2ηt}\displaystyle=\{\sigma\in T_{0}:\forall\;t<\left|\sigma\right|\;\forall\;r\in[t,\left|\sigma\right|)\;d(\sigma(t),\sigma(r))\leq 2\eta_{t}\}

The tree TT is computable since {2ηs:s0}{d(i,j):i,j0}\{2\cdot\eta_{s}:s\geq 0\}\cap\{d(i,j):i,j\geq 0\} are disjoint. Further TT has no dead ends since we can just extend by repeating the last entry (since nsns1n_{s}\geq n_{s-1}). Let C=[T]C=[T], which is then a computable Polish space with countable dense set given by extending any node σ\sigma in TT by means of repeating its last entry indefinitely. Since the function tntt\mapsto n_{t} is computable, one has that CC is strongly computably compact.

The map π:CX\pi:C\rightarrow X given by sending ω\omega to limiω(i)\lim_{i}\omega(i) in XX is well-defined. For, since ηs<2(s+1)\eta_{s}<2^{-(s+1)}, every ω\omega in CC is a Cauchy sequence.

By definition of TT, note that d(π(ω),ω(t))2ηtd(\pi(\omega),\omega(t))\leq 2\eta_{t} for all t0t\geq 0. For let ϵ>0\epsilon>0. Since d(π(ω),ω(r))0d(\pi(\omega),\omega(r))\rightarrow 0, choose r>tr>t such that d(π(ω),ω(r))<ϵd(\pi(\omega),\omega(r))<\epsilon. Then d(π(ω),ω(t))d(π(ω),ω(r))+d(ω(t),ω(r))ϵ+2ηtd(\pi(\omega),\omega(t))\leq d(\pi(\omega),\omega(r))+d(\omega(t),\omega(r))\leq\epsilon+2\eta_{t}.

Note that any ω\omega in CC is a sequence from the countable dense set of XX which converges fast to π(ω)\pi(\omega). This is because 2ηt<2t2\eta_{t}<2^{-t}.

Further, π:CX\pi:C\rightarrow X is surjective: if xx in XX is given, then for each jj choose ω(j)nj\omega(j)\leq n_{j} such that xx is in B(ω(j),ηj)B(\omega(j),\eta_{j}). Then ω\omega is in CC since for all kjk\geq j one has d(ω(j),ω(k))d(ω(j),x)+d(x,ω(k))ηj+ηk2ηjd(\omega(j),\omega(k))\leq d(\omega(j),x)+d(x,\omega(k))\leq\eta_{j}+\eta_{k}\leq 2\eta_{j}.

Then by Proposition 2.5, the map π:CX\pi:C\rightarrow X is computable continuous since it has a computable modulus of uniform continuity. For, if rational ϵ>0\epsilon>0 is given, compute least 0\ell\geq 0 such that 4η<ϵ4\cdot\eta_{\ell}<\epsilon. Suppose that ω,ω\omega,\omega^{\prime} are in CC with ω,ω\omega,\omega^{\prime} agreeing \leq\ell. Then d(π(ω),π(ω))d(π(ω),ω())+d(ω(),ω())+d(ω(),π(ω))2η+0+2η<ϵd(\pi(\omega),\pi(\omega^{\prime}))\leq d(\pi(\omega),\omega(\ell))+d(\omega(\ell),\omega^{\prime}(\ell))+d(\omega^{\prime}(\ell),\pi(\omega^{\prime}))\leq 2\cdot\eta_{\ell}+0+2\cdot\eta_{\ell}<\epsilon.

By the previous proposition, choose a non-empty effectively closed subset DD of XX which consists only of 𝖬𝖫𝖱ν\mathsf{MLR}^{\nu}’s. Then π1(D)C\pi^{-1}(D)\subseteq C is an effectively closed subset of CC, which is thus strongly computably compact since CC is. By the Computably Dominated Basis Theorem, there is an element ω\omega of π1(D)\pi^{-1}(D) of computably dominated degree. ∎

The following example shows that one cannot in general assume that the 𝖬𝖫𝖱ν\mathsf{MLR}^{\nu}’s of computably dominated degree in the previous proposition are non-atoms.

Example 2.19.

There is an uncountable computably compact computable Polish space XX and a computable point ν\nu of 𝒫(X)\mathcal{P}(X) such that the only elements of 𝖬𝖫𝖱ν(X)\mathsf{MLR}^{\nu}(X) of computably dominated degree are among the atoms.

This follows from a construction of Ng et. al.686868[49, Lemma 2.1, Theorem 2.2]. Let μ\mu be the uniform measure on Cantor space Y={0,1}Y=\{0,1\}^{\mathbb{N}} and let Z={0,1,2}Z=\{0,1,2\}^{\mathbb{N}}. Ng et. al. constructs a computable continuous map f:YZf:Y\rightarrow Z with image XX such that every ω\omega in 𝖬𝖫𝖱μ,(Y)\mathsf{MLR}^{\mu,\emptyset^{\prime}}(Y) is such that f(ω)f(\omega) is non-isolated in XX, and vice-versa, and in this circumstance f(ω)f(\omega) and ω\omega have the same Turing degree.

The image XX is a computable Polish space.696969Since it is the computable continuous image of Cantor space, cf. [10, Theorem 2.4.8(3) pp. 73-74]. Further, pushforwards of computable probability measures under computable continuous maps are computable probability measures (by Proposition 2.7), and so ν:=f#μ\nu:=f\#\mu is a computable point of 𝒫(X)\mathcal{P}(X). Note that ν\nu has full support since μ\mu has full support.

Suppose that ω\omega^{\prime} in XX is in 𝖬𝖫𝖱ν(X)\mathsf{MLR}^{\nu}(X) and is of computably dominated degree. Then we claim that ω\omega^{\prime} is an atom. For reductio, suppose not. Since any isolated point in a space with full support is an atom, one has that ω\omega^{\prime} is not isolated. Since f:YXf:Y\rightarrow X is a surjection, choose ω\omega in YY with f(ω)=ωf(\omega)=\omega^{\prime}. By the construction, ω\omega is in 𝖬𝖫𝖱μ,(Y)\mathsf{MLR}^{\mu,\emptyset^{\prime}}(Y). But these points are not of computably dominated degree.707070E.g. [15, Theorem 8.21.2 p. 382]. Since ω,ω\omega,\omega^{\prime} have the same Turing degree, ω\omega^{\prime} is not of computably dominated degree, contrary to hypothesis.

It is not clear to us what happens in the general atomless non-compact case:

Question 2.20.

Suppose that XX is a computable Polish space which is not computably compact, and that ν\nu in 𝒫(X)\mathcal{P}(X) is computable and atomless. Is there an element in 𝖬𝖫𝖱ν(X)\mathsf{MLR}^{\nu}(X) that is of computably dominated degree?

If XX is the reals, it can be written as an effective union of computably compact Polish subspaces, and so the answer is affirmative, by Proposition 2.18. If XX is Baire space, then the answer is again affirmative, by using effective tightness to describe the 𝖬𝖫𝖱ν\mathsf{MLR}^{\nu}’s as a subset of a countable union of computably compact sets, and then applying the Computably Dominated Basis Theorem again. Hence to answer the question negatively one should be looking for spaces which are not “effectively KσK_{\sigma}” and spaces where effective tightness does not produce a union of computably compact sets containing the 𝖬𝖫𝖱ν\mathsf{MLR}^{\nu}’s.

If ν\nu is in 𝒫(X)\mathcal{P}(X) and f:X[,]f:X\rightarrow[-\infty,\infty] is in L1(ν)L_{1}(\nu), then it induces the push-forward probability measure (f#ν)(A)=ν(f1(A))(f\#\nu)(A)=\nu(f^{-1}(A)) in 𝒫()\mathcal{P}(\mathbb{R}). The following proposition tells us that the map ff#νf\mapsto f\#\nu is computable continuous. We use this proposition primarily in conjunction with Proposition 2.4 and Proposition 2.7, which together tell us that pushforwards of L1(ν)L_{1}(\nu)-computable functions are themselves computable.

Proposition 2.21.

Let XX be a computable Polish space. Suppose that ν\nu is a computable point of 𝒫(X)\mathcal{P}(X). Then the map from L1(ν)L_{1}(\nu) to 𝒫()\mathcal{P}(\mathbb{R}) given by sending ff to f#νf\#\nu is continuous computable. Similarly, the map from L1+(ν)L_{1}^{+}(\nu) to 𝒫(0)\mathcal{P}(\mathbb{R}^{\geq 0}) given by sending ff to f#νf\#\nu is continuous computable.

Proof.

We apply Proposition 2.5.

Suppose that ff is an element of the countable dense set of L1(ν)L_{1}(\nu). Then f=i=1mqiIAif=\sum_{i=1}^{m}q_{i}\cdot I_{A_{i}}, where qiq_{i} is rational and the AiA_{i} are elements of the algebra generated by a ν\nu-computable basis. Then uniformly in rationals p<qp<q one has that ν(f1(p,q))=ν({Ai:1im,qi(p,q)})\nu(f^{-1}(p,q))=\nu(\cup\{A_{i}:1\leq i\leq m,q_{i}\in(p,q)\}), which is left-c.e. and indeed computable. Hence f#νf\#\nu is a computable point of 𝒫(0)\mathcal{P}(\mathbb{R}^{\geq 0}) by Proposition 2.7.

Any computable function m:>0>0m:\mathbb{Q}^{>0}\rightarrow\mathbb{Q}^{>0} satisfying m(ϵ)<ϵm(\epsilon)<\epsilon is a computable modulus of uniform continuity. To see this, suppose that ϵ>0\epsilon>0, and suppose that h:h:\mathbb{R}\rightarrow\mathbb{R} is 1-Lipschitz, and that 𝔼ν|fg|<m(ϵ)\mathbb{E}_{\nu}\left|f-g\right|<m(\epsilon). By change of variables, one has that |𝔼f#νh𝔼g#νh|=|𝔼ν(hf)𝔼ν(hg)|𝔼ν|hfhg|𝔼ν|fg|<m(ϵ)\left|\mathbb{E}_{f\#\nu}h-\mathbb{E}_{g\#\nu}h\right|=\left|\mathbb{E}_{\nu}(h\circ f)-\mathbb{E}_{\nu}(h\circ g)\right|\leq\mathbb{E}_{\nu}\left|h\circ f-h\circ g\right|\leq\mathbb{E}_{\nu}\left|f-g\right|<m(\epsilon), where the second-to-last inequality uses that hh is 1-Lipschitz. By taking the supremum over all 11-Lipschitz h:h:\mathbb{R}\rightarrow\mathbb{R} with h1\|h\|_{\infty}\leq 1, one has dKR(f#ν,g#ν)m(ϵ)d_{KR}(f\#\nu,g\#\nu)\leq m(\epsilon), which by construction is <ϵ<\epsilon.

Since {fL1(ν):f0}\{f\in L_{1}(\nu):f\geq 0\} is a computable Polish subspace of L1(ν)L_{1}(\nu), the restriction of ff#νf\mapsto f\#\nu to it is also computable continuous. ∎

The above proposition has the following extremely useful consequence:717171Outside of density, the statement of this lemma is contained in Miyabe’s proof of his characterisation of 𝖲𝖱ν\mathsf{SR}^{\nu} in terms of Lp(ν)L_{p}(\nu) Schnorr tests. See e.g. the line “It follows that μ({x:t(x)>rn})\mu(\{x:t(x)>r_{n}\}) is computable uniformly in nn” ([45, p. 6]). Miyabe does not use pushforwards, but rather does it out by hand for Lp(ν)L_{p}(\nu) Schnorr tests.

Lemma 2.22.

Let XX be a computable Polish space. Suppose that ν\nu is a computable point of 𝒫(X)\mathcal{P}(X).

Suppose f:X[0,]f:X\rightarrow[0,\infty] is lsc with f<f<\infty ν\nu-a.s. Suppose that f#νf\#\nu is a computable point of 𝒫(0)\mathcal{P}(\mathbb{R}^{\geq 0}). Then there is a computable sequence of reals ri>0r_{i}>0 dense in [0,)[0,\infty) such that f1(ri,]f^{-1}(r_{i},\infty] is c.e. open with uniformly ν\nu-computable measure.

In particular, this is true of any Lp(ν)L_{p}(\nu) Schnorr test.

Proof.

Let μ:=f#ν\mu:=f\#\nu, which by hypothesis is a computable point of 𝒫(0)\mathcal{P}(\mathbb{R}^{\geq 0}). By the Hoyrup-Rojas result discussed in §2.2, there is a μ\mu-computable basis of the form (qri,q+ri)0(q-r_{i},q+r_{i})\cap\mathbb{R}^{\geq 0}, where qq ranges over rationals and ri>0r_{i}>0 is a computable sequence dense in [0,)[0,\infty), and where further (qri,q+ri)0(q-r_{i},q+r_{i})\cap\mathbb{R}^{\geq 0} has the same μ\mu-measure as [qri,q+ri]0[q-r_{i},q+r_{i}]\cap\mathbb{R}^{\geq 0}. Since f<f<\infty ν\nu-a.s., we have ν(f1(ri,])=ν(f1(ri,))=(f#ν)(ri,)=μ(ri,)=1μ[0,ri]=1μ([ri,ri]0)\nu(f^{-1}(r_{i},\infty])=\nu(f^{-1}(r_{i},\infty))=(f\#\nu)(r_{i},\infty)=\mu(r_{i},\infty)=1-\mu[0,r_{i}]=1-\mu([-r_{i},r_{i}]\cap\mathbb{R}^{\geq 0}), which is computable.

The last point follows from the previous proposition. ∎

The following proposition is elementary but useful. (Recall usc was defined in Definition 1.1(2)).

Proposition 2.23.

For any element ff of the countable dense set of Lp+(ν)L_{p}^{+}(\nu), one can compute an index for a non-negative lsc function gg and a non-negative usc function hh such that f=g=hf=g=h on 𝖪𝖱ν\mathsf{KR}^{\nu}.

Likewise, for any element ff of the countable dense set of Lp(ν)L_{p}(\nu), one can compute a rational qq and an index for a non-negative lsc function gg and a non-negative usc function hh such that fq=g=hf-q=g=h on 𝖪𝖱ν\mathsf{KR}^{\nu}.

Proof.

Let ff be an element of the countable dense set of Lp+(ν)L_{p}^{+}(\nu). Then f=i=1kqiIAif=\sum_{i=1}^{k}q_{i}\cdot I_{A_{i}}, where qi0q_{i}\geq 0 is rational and AiA_{i} is an element of the algebra generated by a ν\nu-computable basis. By Proposition 2.13, suppose that UiU_{i} is a c.e. open which is uniformly equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to AiA_{i}, and suppose that CiC_{i} is an effectively closed superset of UiU_{i} of the same ν\nu-measure. Then g:=i=1kqiIUig:=\sum_{i=1}^{k}q_{i}\cdot I_{U_{i}} is non-negative lsc, and h:=i=1kqiICih:=\sum_{i=1}^{k}q_{i}\cdot I_{C_{i}} is non-negative usc, and they agree with ff on 𝖪𝖱ν\mathsf{KR}^{\nu}.

Let ff be an element of the countable dense set of Lp(ν)L_{p}(\nu). Then f=i=1kqiIAif=\sum_{i=1}^{k}q_{i}\cdot I_{A_{i}}, where qiq_{i} is rational and AiA_{i} is an element of the algebra generated by a ν\nu-computable basis. Let q=miniqiq=\min_{i}q_{i}. Then fqf-q is an element of the countable dense set of Lp+(ν)L_{p}^{+}(\nu). ∎

2.4. The space of measurable functions

The space of equivalence classes of Borel measurable functions that are finite ν\nu-a.s. under ν\nu-a.s. identity is denoted by L0(X,ν)L_{0}(X,\nu), where ν\nu is in +(X)\mathcal{M}^{+}(X). We write L0(ν)L_{0}(\nu) when XX is clear from context.

In keeping with the notational conventions in §1.2, we write 𝕃0(ν)\mathbb{L}_{0}(\nu) for the pointwise-defined Borel measurable functions that are finite ν\nu-a.s.

The topology on L0(ν)L_{0}(\nu) is given by convergence in measure. To enhance readability, if hh is a measurable function, then we write ν(|h|>ϵ)\nu(\left|h\right|>\epsilon) for the more cumbersome ν({xX:|h|(x)>ϵ})\nu(\{x\in X:\left|h\right|(x)>\epsilon\}). Then recall fnff_{n}\rightarrow f in measure iff for all ϵ>0\epsilon>0 one has that limnν(|fnf|>ϵ)=0\lim_{n}\nu(\left|f_{n}-f\right|>\epsilon)=0. Recall that a consequence of Egoroff’s Theorem is that fnff_{n}\rightarrow f ν\nu-a.s. implies fnff_{n}\rightarrow f in L0(ν)L_{0}(\nu) for ν\nu in +(X)\mathcal{M}^{+}(X).727272[21, p. 62].

A compatible complete metric is given by d(f,g)=fg0d(f,g)=\|f-g\|_{0} where h0=inf{ϵ>0:ν(|h|>ϵ)<ϵ}\|h\|_{0}=\inf\{\epsilon>0:\nu(\left|h\right|>\epsilon)<\epsilon\}. Note that the set {ϵ>0:ν(|h|>ϵ)<ϵ}\{\epsilon>0:\nu(\left|h\right|>\epsilon)<\epsilon\} is upwards closed, so that h0=sup{ϵ>0:ν(|h|>ϵ)>ϵ}\|h\|_{0}=\sup\{\epsilon>0:\nu(\left|h\right|>\epsilon)>\epsilon\}. When ν\nu in 𝒫(X)\mathcal{P}(X), this is called the Ky Fan metric.737373[16, 289] While 0\|\cdot\|_{0} satisfies the triangle inequality f+g0f0+g0\|f+g\|_{0}\leq\|f\|_{0}+\|g\|_{0} and satisfies f0=0\|f\|_{0}=0 iff f=0f=0 ν\nu-a.s., it does not in general satisfy ch0=|c|h0\|c\cdot h\|_{0}=\left|c\right|\cdot\|h\|_{0}.747474[14, 65-69], [16, 289-290].,757575More generally, L0(ν)L_{0}(\nu) is not a Banach space. In working with the metric, it is useful to note that h0ϵ\|h\|_{0}\leq\epsilon iff ν(|h|>ϵ)ϵ\nu(\left|h\right|>\epsilon)\leq\epsilon. Finally, note that |f||g|\left|f\right|\leq\left|g\right| ν\nu-a.s. implies f0g0\|f\|_{0}\leq\|g\|_{0} in L0(ν)L_{0}(\nu).

The natural countable dense set for L0(ν)L_{0}(\nu) is the the rational-valued simple functions formed from the algebra generated by a ν\nu-computable basis, that is, the same countable dense set as we used for Lp(ν)L_{p}(\nu) for p1p\geq 1 computable. Classically, this set is dense in L0(ν)L_{0}(\nu), so it remains to verify that the distance between these two points is uniformly computable:

Proposition 2.24.

If hh is a rational-valued simple functions formed from the algebra generated by a ν\nu-computable basis, then h0\|h\|_{0} is computable, and uniformly so. If f,gf,g are two such simple functions, then fg0\|f-g\|_{0} is computable, and uniformly so.

Proof.

If hh is one of these functions, then so too is |h|\left|h\right|. Suppose that |h|=i=1nqiIAi\left|h\right|=\sum_{i=1}^{n}q_{i}\cdot I_{A_{i}}, where qi0q_{i}\geq 0 is rational and AiA_{i} is are pairwise disjoint events from the algebra generated by a ν\nu-computable basis.

For ϵ\epsilon in >0\mathbb{Q}^{>0}, let Jϵ={i[1,n]:qi>ϵ}J_{\epsilon}=\{i\in[1,n]:q_{i}>\epsilon\}, which is a finite set whose index is computable uniformly from ϵ>0\epsilon>0. Then ν(|h|>ϵ)=iJϵν(Ai)\nu(\left|h\right|>\epsilon)=\sum_{i\in J_{\epsilon}}\nu(A_{i}), which is a computable real, uniformly in ϵ>0\epsilon>0 (by Proposition 2.8(1).

Then ϵ\epsilon in >0\mathbb{Q}^{>0} satisfies ν(|h|>ϵ)<ϵ\nu(\left|h\right|>\epsilon)<\epsilon iff iJϵν(Ai)<ϵ\sum_{i\in J_{\epsilon}}\nu(A_{i})<\epsilon, which is a c.e. condition. If we enumerate these rational ϵ\epsilon and take mins as we go, we get a computable decreasing sequence of rationals which converges down to h0\|h\|_{0}, so that h0\|h\|_{0} is right-c.e.

Likewise, δ\delta in >0\mathbb{Q}^{>0} satisfies ν(|h|>δ)>δ\nu(\left|h\right|>\delta)>\delta iff iJδν(Ai)>δ\sum_{i\in J_{\delta}}\nu(A_{i})>\delta, which is a c.e. condition. If we enumerate these rational δ\delta and take maxes as we go, we get a increasing computable sequence of rationals which converges up to h0\|h\|_{0}, so that h0\|h\|_{0} is left-c.e.

Similarly if f,gf,g are from a countable dense set then fg0\|f-g\|_{0} is a computable real since fgf-g is also an element of the countable dense set. ∎

We call the following the computable embedding of Lp(ν)L_{p}(\nu) into L0(ν)L_{0}(\nu). The square root in the rate of convergence is, in our view, explanatory of the many 2\sqrt{2}’s that appear in Pathak et. al. when dealing computable points of L1(ν)L_{1}(\nu).767676See e.g. [52, p. 343].

Proposition 2.25.

Suppose that p1p\geq 1 is computable. Then the inclusion map is a uniformly continuous computable map from Lp(ν)L_{p}(\nu) to L0(ν)L_{0}(\nu). Further, if fnff_{n}\rightarrow f at geometric rate b>1b>1 of convergence in Lp(ν)L_{p}(\nu), then fnff_{n}\rightarrow f at geometric rate b\sqrt{b} in L0(ν)L_{0}(\nu).

Proof.

Suppose that p1p\geq 1. Since Lp(ν)L_{p}(\nu) and L0(ν)L_{0}(\nu) have the same countable dense set, by Proposition 2.5, it suffices to show that there is a computable modulus m:>0>0m:\mathbb{Q}^{>0}\rightarrow\mathbb{Q}^{>0} of uniform continuity. Given rational ϵ>0\epsilon>0, compute rational δ<ϵ\delta<\epsilon and compute a rational m(ϵ)<δ1+1pm(\epsilon)<\delta^{1+\frac{1}{p}}. Suppose f,gf,g are in Lp(ν)L_{p}(\nu) with fgp<m(ϵ)\|f-g\|_{p}<m(\epsilon). Then ν(|fg|>δ)1δpfgpp1δpm(ϵ)p<δ\nu(\left|f-g\right|>\delta)\leq\frac{1}{\delta^{p}}\|f-g\|_{p}^{p}\leq\frac{1}{\delta^{p}}m(\epsilon)^{p}<\delta, and so fg0δ<ϵ\|f-g\|_{0}\leq\delta<\epsilon.

Suppose that p1p\geq 1 and suppose fnff_{n}\rightarrow f at geometric rate b>1b>1 of convergence in Lp(ν)L_{p}(\nu). Then ffn1ffnp\|f-f_{n}\|_{1}\leq\|f-f_{n}\|_{p}, and so fnff_{n}\rightarrow f at geometric rate b>1b>1 in L1(ν)L_{1}(\nu). Let n0n\geq 0. Then ν(|ffn|>(b)n)(b)nffn1(b)n\nu(\left|f-f_{n}\right|>(\sqrt{b})^{-n})\leq(\sqrt{b})^{n}\cdot\|f-f_{n}\|_{1}\leq(\sqrt{b})^{-n}. Then ffn0(b)n\|f-f_{n}\|_{0}\leq(\sqrt{b})^{-n}. ∎

The following proposition is the natural effectivization of the Bounded Convergence Theorem:777777[75, 130].

Proposition 2.26.

(Effective Bounded Convergence Theorem). Suppose ν\nu is a computable point of 𝒫(X)\mathcal{P}(X). Then:

  1. (1)

    Suppose that fnff_{n}\rightarrow f in L0(ν)L_{0}(\nu) at a geometric rate of convergence b2b\geq\sqrt{2}. Suppose that K0K\geq 0 such that |fn|K\left|f_{n}\right|\leq K ν\nu-a.s. for all n0n\geq 0. Then fn+2ff_{n+2}\rightarrow f at a geometric rate of convergence bb in L1(ν)L_{1}(\nu).

  2. (2)

    Suppose ff is a computable point of L0(ν)L_{0}(\nu) and K0K\geq 0 is a rational such |f|K\left|f\right|\leq K ν\nu-a.s. Then ff is a computable point of L1(ν)L_{1}(\nu).

  3. (3)

    Suppose fnf_{n} is a uniformly computable point of L0(ν)L_{0}(\nu) and Kn0K_{n}\geq 0 is a uniformly computable sequence of rationals such that |fn|Kn\left|f_{n}\right|\leq K_{n} ν\nu-a.s. Then fnf_{n} is uniformly a computable point of L1(ν)L_{1}(\nu).

Proof.

For (1), classically some subsequence of fnf_{n} converges ν\nu-a.s. to ff. Hence, |f|K\left|f\right|\leq K ν\nu a.s. Further, we may suppose K>1K>1. Suppose that fnff_{n}\rightarrow f at a geometric rate of b2b\geq\sqrt{2} convergence in L0(ν)L_{0}(\nu), so that ν(|fnf|>bn)bn\nu(\left|f_{n}-f\right|>b^{-n})\leq b^{-n} for all n0n\geq 0. Choose n02n_{0}\geq 2 sufficiently large so that bn0<12Kb^{-n_{0}}<\frac{1}{2K}. Let c=bn0c=b^{n_{0}}, so that for all n2n\geq 2 we have 2Kcn2Kbn0bn<bn2K\cdot c^{-n}\leq 2K\cdot b^{-n_{0}}\cdot b^{-n}<b^{-n}. Then fn+2f1|fn+2f|>c(n+2)|fn+2f|𝑑ν+|fn+2f|c(n+2)|fn+2f|𝑑ν2Kc(n+2)+c(n+2)2b(n+2)bn\|f_{n+2}-f\|_{1}\leq\int_{\left|f_{n+2}-f\right|>c^{-(n+2)}}\left|f_{n+2}-f\right|\;d\nu+\int_{\left|f_{n+2}-f\right|\leq c^{-(n+2)}}\left|f_{n+2}-f\right|\;d\nu\leq 2K\cdot c^{-(n+2)}+c^{-(n+2)}\leq 2b^{-(n+2)}\leq b^{-n}, where the last inequality follows from b2b\geq\sqrt{2}.

For (2), suppose fnff_{n}\rightarrow f fast in L0(ν)L_{0}(\nu). For 0<ϵ<10<\epsilon<1 one has ν(|fnI|fn|K+1f|>ϵ)ν(|fnf|>ϵ)\nu(\left|f_{n}\cdot I_{\left|f_{n}\right|\leq K+1}-f\right|>\epsilon)\leq\nu(\left|f_{n}-f\right|>\epsilon), and hence we may assume that |fn|K+1\left|f_{n}\right|\leq K+1. Then we apply (1).

For (3), this is just the uniformisation of (2). ∎

Using Proposition 2.5, one has that many of the usual operations on L0(ν)L_{0}(\nu) are computable continuous, such as addition and minimum and maximum. Indeed, each of these three has modulus m(ϵ)=ϵ2m(\epsilon)=\frac{\epsilon}{2}. We can use this observation to prove the following. It had been previously established by Rute, although our proof is different.787878[61, Proposition 3.26].

Proposition 2.27.

If ff is a computable point of L0(ν)L_{0}(\nu) (resp. L0+(ν)L_{0}^{+}(\nu)), then f#νf\#\nu is a computable point of 𝒫()\mathcal{P}(\mathbb{R}) (resp. 𝒫(0)\mathcal{P}(\mathbb{R}^{\geq 0})).

Proof.

Let fn=min(max(f,n),n)f_{n}=\min(\max(f,-n),n), so that fn=ff_{n}=f on f1(n,n)f^{-1}(-n,n), and fnf_{n} is uniformly a computable point of L0(ν)L_{0}(\nu) (resp. L0+(ν)L_{0}^{+}(\nu)). By Proposition 2.26 (3) one has that fnf_{n} is uniformly a computable point of L1(ν)L_{1}(\nu) (resp. L1+(ν)L_{1}^{+}(\nu)). By Proposition 2.21, one has that fn#νf_{n}\#\nu is uniformly a computable point of 𝒫()\mathcal{P}(\mathbb{R}) (resp. 𝒫(0)\mathcal{P}(\mathbb{R}^{\geq 0})). For rational p<qp<q (resp. rational 0p<q0\leq p<q), compute natural number n>max(|p|,|q|)n>\max(\left|p\right|,\left|q\right|), so that by Proposition 2.7 the real (f#ν)(p,q)=(fn#ν)(p,q)(f\#\nu)(p,q)=(f_{n}\#\nu)(p,q) is uniformly left-c.e. Then by Proposition 2.7, the probability measure f#νf\#\nu is a computable point of 𝒫()\mathcal{P}(\mathbb{R}) (resp. 𝒫(0)\mathcal{P}(\mathbb{R}^{\geq 0})). ∎

We then define:

Definition 2.28.

An L0(ν)L_{0}(\nu) Schnorr test is a lsc function f:X[0,]f:X\rightarrow[0,\infty] which is a computable point of L0(ν)L_{0}(\nu).

In parallel to Definition 1.1(7), we have the following new characteriation of 𝖲𝖱ν\mathsf{SR}^{\nu}:

Proposition 2.29.

A point xx is in 𝖲𝖱ν\mathsf{SR}^{\nu} iff f(x)<f(x)<\infty for all L0(ν)L_{0}(\nu) Schnorr tests ff.

Proof.

If f(x)<f(x)<\infty for all L0(ν)L_{0}(\nu) Schnorr tests ff, then by the computable embedding of L1(ν)L_{1}(\nu) into L0(ν)L_{0}(\nu), we have f(x)<f(x)<\infty for all L1(ν)L_{1}(\nu) Schnorr tests ff, and so xx is in 𝖲𝖱ν\mathsf{SR}^{\nu}. Conversely, suppose that xx is in 𝖲𝖱ν\mathsf{SR}^{\nu}. Let ff be an L0(ν)L_{0}(\nu) Schnorr test. Since ff is in L0(ν)L_{0}(\nu), it is finite ν\nu-a.s. By Lemma 2.22 and the previous proposition, there is a computable sequence of reals ηn\eta_{n} in the open interval (2n,2n+1)(2^{n},2^{n+1}) such that Un:=f1(ηn,]U_{n}:=f^{-1}(\eta_{n},\infty] is c.e. open with uniformly ν\nu-computable measure. So we have ν(Un)0\nu(U_{n})\rightarrow 0, and since ν(Un)\nu(U_{n}) is computable, we can compute a subsequence with ν(Uni)<2i\nu(U_{n_{i}})<2^{-i}. Hence f=iIUnif=\sum_{i}I_{U_{n_{i}}} is an L1(ν)L_{1}(\nu) Schnorr test, and so f(x)<f(x)<\infty and so xx is only finitely many of the UniU_{n_{i}}, and hence there is ii such that f(x)ηnif(x)\leq\eta_{n_{i}}. ∎

Miyabe has shown that xx being in every Σ20\Sigma^{0}_{2} ν\nu-measure one class is equivalent to f(x)<f(x)<\infty for every non-negative lsc ff in L0(ν)L_{0}(\nu).797979[47, Proposition 3.3]. This notion of randomness is also called weak 2-randomness. In conjunction with the above proposition, it suggests that there is little room for a simple characterisation of 𝖬𝖫𝖱ν\mathsf{MLR}^{\nu} in terms of non-negative lsc functions in L0(ν)L_{0}(\nu).

Finally, we update our previous approximation theorem for Lp(ν)L_{p}(\nu) Schnorr tests to L0(ν)L_{0}(\nu) Schnorr tests:

Proposition 2.30.

For all L0(ν)L_{0}(\nu) Schnorr tests f:X[0,]f:X\rightarrow[0,\infty], one can compute a subsequence of fs(n)f_{s(n)} the fsf_{s} from Proposition 2.16 such that fs(n)ff_{s(n)}\rightarrow f fast in L0(ν)L_{0}(\nu).

Proof.

The fsf_{s} come from the countable dense set of L0(ν)L_{0}(\nu) and hence are computable points of L0(ν)L_{0}(\nu). Since fsff_{s}\rightarrow f everywhere, we have fsff_{s}\rightarrow f in measure, and so fsff_{s}\rightarrow f in L0(ν)L_{0}(\nu). Since fsf0\|f_{s}-f\|_{0} is computable, we just search for a subsequence fs(n)f_{s(n)} with fs(n)f0<2n\|f_{s(n)}-f\|_{0}<2^{-n}. ∎

3. Two Schnorr lemmas: Flipping an approximation and Self-location

In this section, we provide two lemmas on Schnorr tests, one involving turning an approximation from below into a non-increasing subsequence converging down to zero, and another based upon a distinctive self-location property of Schnorr randoms.

The first of these involves a partial subtraction operator, which involves some care since it helps one avoid situations with \infty-\infty. These situations can potentially arise since lsc functions are allowed to take infinite values.

Proposition 3.1.

Suppose that p1p\geq 1 computable or p=0p=0.

Suppose that f:X(,]f:X\rightarrow(-\infty,\infty] is an lsc function in Lp(ν)L_{p}(\nu) (resp. an Lp(ν)L_{p}(\nu) Schnorr test). Suppose that 𝖷𝖱ν\mathsf{XR}^{\nu} is a ν\nu-measure one set on which the function ff is finite.

Suppose that g:X(,]g:X\rightarrow(-\infty,\infty] is an lsc function such that gfg\leq f on 𝖷𝖱ν\mathsf{XR}^{\nu}. Suppose that gg is paired with a usc function g˘:X[,)\breve{g}:X\rightarrow[-\infty,\infty) such that g,g˘g,\breve{g} are equal on 𝖷𝖱ν\mathsf{XR}^{\nu}.

Define fg=max(0,fg˘)f\ominus g=\max(0,f-\breve{g}). Then

  • g,g˘g,\breve{g} are finite on 𝖷𝖱ν\mathsf{XR}^{\nu}.

  • fgf\ominus g is non-negative lsc and in Lp(ν)L_{p}(\nu) (resp. an Lp(ν)L_{p}(\nu) Schnorr test) and is equal on 𝖷𝖱ν\mathsf{XR}^{\nu} to fgf-g.

Proof.

For the first item, since the lsc function gg has codomain (,](-\infty,\infty] and the usc function g˘\breve{g} has codomain [,)[-\infty,\infty), then when the two agree they have finite value. And they agree on 𝖷𝖱ν\mathsf{XR}^{\nu}.

Since g˘\breve{g} is usc, g˘-\breve{g} is lsc. Since the lsc functions are closed under addition (cf. Proposition 2.6), one has that fg˘f-\breve{g} is lsc. Since the lsc functions are preserved under max (cf. again Proposition 2.6), we have that fgf\ominus g is non-negative lsc. Further, since f,gf,g are in Lp(ν)L_{p}(\nu) (resp. are Lp(ν)L_{p}(\nu)-computable) and this property is preserved under subtraction and maxes, we have that fgf\ominus g is also in Lp(ν)L_{p}(\nu) (resp. Lp(ν)L_{p}(\nu)-computable). On 𝖷𝖱ν\mathsf{XR}^{\nu}, one has that fgf-g is both equal to fg˘f-\breve{g} and is non-negative, and hence equal to fgf\ominus g. ∎

While partial subtraction operation fgf\ominus g is not defined absolutely, but only relative to the hypotheses of the previous proposition, the situation of the following lemma is the one which tends to be operative in applications. We call it “flipping an approximation” since it takes a non-decreasing approximation fsff_{s}\rightarrow f and turns it into a non-increasing approximation ffn0f\ominus f_{n}\rightarrow 0. While classically trivial, it requires some organisation to handle within effective categories:

Lemma 3.2.

(Flipping an approximation) Suppose that p1p\geq 1 computable (resp. p=0p=0).

For each Lp(ν)L_{p}(\nu) Schnorr test ff, let gsg_{s} be from the countable dense set of Lp(ν)L_{p}(\nu) as in Proposition 2.16 (resp. Proposition 2.30), so that gsgs+1g_{s}\leq g_{s+1} everywhere and f=supsgsf=\sup_{s}g_{s} everywhere and gsfg_{s}\rightarrow f fast in Lp(ν)L_{p}(\nu). Using Proposition 2.23, let fs,f˘sf_{s},\breve{f}_{s} be non-negative lsc and usc respectively with gs=fs=f˘sg_{s}=f_{s}=\breve{f}_{s} on 𝖪𝖱ν\mathsf{KR}^{\nu}. Then by the previous proposition, we have:

  • fs,f˘sf_{s},\breve{f}_{s} are finite on 𝖪𝖱ν\mathsf{KR}^{\nu}.

  • ffsf\ominus f_{s} is an Lp(ν)L_{p}(\nu) Schnorr test and is equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to fgsf-g_{s}.

  • ffsf\ominus f_{s} is non-increasing and ffs0f\ominus f_{s}\rightarrow 0 on 𝖪𝖱ν\mathsf{KR}^{\nu}.

  • for t>st>s, similarly ftfsf_{t}\ominus f_{s} is an Lp(ν)L_{p}(\nu) Schnorr test and is equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to ftgsf_{t}-g_{s}.

Now we turn to self-location. The idea is that given a certain kind of computable “chart” of the computable Polish space, the Schnorr randoms can weakly compute their position on the chart. (For the notion of weak computation, see Definition 1.1(13).

Lemma 3.3.

(Self-location lemma).

Suppose that ν\nu is a computable point of 𝒫(X)\mathcal{P}(X).

Suppose that VmV_{m} is a computable sequence of c.e. opens with uniformly computable ν\nu-measure.

Suppose that xx is in 𝖲𝖱ν\mathsf{SR}^{\nu}. Then xx weakly computes the element {m:xVm}\{m:x\in V_{m}\} of Cantor space.

Proof.

Let B0,B1,B_{0},B_{1},\ldots be a ν\nu-computable basis, with associated sequence C0,C1,C_{0},C_{1},\ldots of effectively closed supersets of the same measure. Let Bm,tB_{m,t} be a computable subsequence such that Vm=tBm,tV_{m}=\bigcup_{t}B_{m,t}. Since VmV_{m} and the Bm,tB_{m,t} have uniformly computable ν\nu-measure, there is a computable function ms(m)m\mapsto s(m) such that ν(VmUm)<2m\nu(V_{m}\setminus U_{m})<2^{-m} where Um=ts(m)Bm,tU_{m}=\bigcup_{t\leq s(m)}B_{m,t}. Let Dm=ts(m)Cm,tD_{m}=\bigcup_{t\leq s(m)}C_{m,t}, which is an effectively closed set equal to UmU_{m} on 𝖪𝖱ν\mathsf{KR}^{\nu}. Then f=mIVmDmf=\sum_{m}I_{V_{m}\setminus D_{m}} is an L1(ν)L_{1}(\nu) Schnorr test.

Let xx be in 𝖲𝖱ν\mathsf{SR}^{\nu}. Since f(x)<f(x)<\infty, there are only finitely many mm such that xx is in VmDmV_{m}\setminus D_{m}. Hence, there are there are only finitely many mm such that xx is in VmUmV_{m}\setminus U_{m}. Then the sets {m:xVm}\{m:x\in V_{m}\} and {m:xUm}\{m:x\in U_{m}\} differ by only finitely much and hence are Turing equivalent.

Since UmU_{m} comes from the algebra generated by the ν\nu-computable basis, using the ν\nu-computable basis as in Proposition 2.13 we can compute indexes for c.e. opens UmU_{m}^{\prime} such that Um=XUmU_{m}^{\prime}=X\setminus U_{m} on 𝖪𝖱ν\mathsf{KR}^{\nu}. Since both UmU_{m} and UmU_{m}^{\prime} are uniformly c.e. open, choose computable sequences pm,i,pm,ip_{m,i},p_{m,i}^{\prime} from the countable dense set and ϵm,i,ϵm,i\epsilon_{m,i},\epsilon_{m,i}^{\prime} from >0\mathbb{Q}^{>0} such that Um=iB(pm,i,ϵm,i)U_{m}=\bigcup_{i}B(p_{m,i},\epsilon_{m,i}) and Um=iB(pm,i,ϵm,i)U_{m}^{\prime}=\bigcup_{i}B(p_{m,i}^{\prime},\epsilon_{m,i}^{\prime}), where B(p,ϵ)B(p,\epsilon) denotes again the open ball around pp of radius ϵ\epsilon. We can enumerate these sets as Um=sUm,sU_{m}=\bigcup_{s}U_{m,s} and Um=sUm,sU_{m}^{\prime}=\bigcup_{s}U_{m,s}^{\prime}, where Um,s=isB(pm,i,ϵm,i)U_{m,s}=\bigcup_{i\leq s}B(p_{m,i},\epsilon_{m,i}) and Um,s=isB(pm,i,ϵm,i)U_{m,s}^{\prime}=\bigcup_{i\leq s}B(p_{m,i}^{\prime},\epsilon_{m,i}^{\prime}).

Consider a sequence from the countable dense set which converges fast to our point xx in 𝖲𝖱ν\mathsf{SR}^{\nu}. Given mm, to compute from the sequence whether xx is in UmU_{m}, we simply start enumerating both UmU_{m} and UmU_{m}^{\prime}: eventually xx gets in one of them (and xx only ever gets in one of them), and we use the sequence to determine when this happens, by Proposition 2.1(1). ∎

Here are some simple applications of self-location, which we use to obtain the information about weak computation in Theorem 1.5(1):

Proposition 3.4.

Suppose that p1p\geq 1 computable (resp. p=0p=0).

  1. (1)

    Suppose that fmf_{m} is a sequence of Lp(ν)L_{p}(\nu) Schnorr tests such that fmf_{m} is non-increasing on 𝖲𝖱ν\mathsf{SR}^{\nu} and such that fm0f_{m}\rightarrow 0 on 𝖲𝖱ν\mathsf{SR}^{\nu}. Every xx in 𝖲𝖱ν\mathsf{SR}^{\nu} weakly computes a modulus of convergence for fm(x)0f_{m}(x)\rightarrow 0.

  2. (2)

    Suppose that ff is an Lp(ν)L_{p}(\nu) Schnorr test. Suppose that fsf_{s} is the approximation as in Proposition 2.16 (resp. Proposition 2.30). Every xx in 𝖲𝖱ν\mathsf{SR}^{\nu} weakly computes a modulus of convergence for fs(x)f(x)f_{s}(x)\rightarrow f(x).

Proof.

For (1), using Lemma 2.22 (resp. in conjunction with Proposition 2.27), choose a computable sequence of reals rir_{i} decreasing to zero such that the c.e. open Vm,i=fm1(ri,]V_{m,i}=f_{m}^{-1}(r_{i},\infty] has ν\nu-computable measure, uniformly. Consider a sequence from the countable dense set which converges fast to xx. By the Self-location lemma, we can Turing compute from it the “chart” set C={(m,i):xVm,i}C=\{(m,i):x\in V_{m,i}\}. Let ϵ>0\epsilon>0 be rational. We show how to compute from CC a natural number m(ϵ)m(\epsilon) such that fn(x)<ϵf_{n}(x)<\epsilon for all nm(ϵ)n\geq m(\epsilon). By hypothesis, fn(x)f_{n}(x) decreases down to zero. Hence to compute m(ϵ)m(\epsilon) from CC we just search for ri<ϵr_{i}<\epsilon and then search for mm with xVm,ix\notin V_{m,i}.

For (2), just use Lemma 3.2 to rewrite the convergence fsff_{s}\rightarrow f as (fgs)0(f\ominus g_{s})\rightarrow 0 on 𝖪𝖱ν\mathsf{KR}^{\nu}, where gsg_{s} is an Lp(ν)L_{p}(\nu) Schnorr test equal to fsf_{s} on 𝖪𝖱ν\mathsf{KR}^{\nu}, and then use (1). ∎

4. Recovering pointwise values on Schnorr randoms

In this section we prove some results about pointwise limits existing on the Schnorr randoms for various effective functions convering fast in Lp(ν)L_{p}(\nu). By way of motivation for these kinds of results, consider X=[0,1]X=[0,1] and let ν\nu be Lebesgue measure and recall the canonical example of L1(ν)L_{1}(\nu) convergence with ν\nu-a.s. lack of pointwise convergence:

f1=I[0,12),f2=I[12,1],f3=I[0,14),f4=I[14,12),f5=I[12,34),f6=I[34,1],\displaystyle\ f_{1}=I_{[0,\frac{1}{2})},\ f_{2}=I_{[\frac{1}{2},1]},\ f_{3}=I_{[0,\frac{1}{4})},\ f_{4}=I_{[\frac{1}{4},\frac{1}{2})},\ f_{5}=I_{[\frac{1}{2},\frac{3}{4})},\ f_{6}=I_{[\frac{3}{4},1]},\ \ldots

Proposition 4.1 below says that the slow L1(ν)L_{1}(\nu)-convergence in this example is essential to the lack of pointwise limits on 𝖲𝖱ν\mathsf{SR}^{\nu}. By modifying the events in fnf_{n} to be open, one similarly gets a sequence of L1(ν)L_{1}(\nu) Schnorr tests gng_{n} which lacks pointwise limits on all 𝖪𝖱ν\mathsf{KR}^{\nu}. Proposition 4.3 likewise says that the slow L1(ν)L_{1}(\nu) convergence of gng_{n} is essential to the ν\nu-a.s. lack of pointwise limits on 𝖲𝖱ν\mathsf{SR}^{\nu}.

In the setting of p=1p=1 and X=[0,1]kX=[0,1]^{k} and ν\nu being the kk-fold product of Lebesgue measure on [0,1][0,1], the following result is due to Pathak, Rojas, and Simpson, who used sequential Schnorr tests.808080[52, Lemma 3.7]. See also [61, §3.3] and [76, Chapter 3] (cf. [70, p. 394]).

Proposition 4.1.

Suppose that p1p\geq 1 is computable. Suppose that ff is a computable point of Lp(ν)L_{p}(\nu). Suppose that fnf_{n} is a computable sequence from the countable dense set of Lp(ν)L_{p}(\nu) such that fnff_{n}\rightarrow f fast in Lp(ν)L_{p}(\nu). Then limnfn\lim_{n}f_{n} exists on 𝖲𝖱ν\mathsf{SR}^{\nu} and is a version of ff.

Moreover, on 𝖲𝖱ν\mathsf{SR}^{\nu} this limit does not depend on the choice of fnf_{n} or the choice of the ν\nu-computable basis.

Finally, if ff is in addition an Lp(ν)L_{p}(\nu) Schnorr test, then limnfn(x)=f(x)\lim_{n}f_{n}(x)=f(x) for all xx in 𝖲𝖱ν\mathsf{SR}^{\nu}.

Proof.

(Sketch) Let g=i|fifi+1|g=\sum_{i}\left|f_{i}-f_{i+1}\right|. Then using Proposition 2.23, it is equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to a Lp(ν)L_{p}(\nu) Schnorr test. This shows that fi(x)f_{i}(x) is a Cauchy sequence for xx in 𝖲𝖱ν\mathsf{SR}^{\nu}.

If fnf_{n}^{\prime} is another such witness to the Lp(ν)L_{p}(\nu) computabilty of ff, then let h=i|fifi|h=\sum_{i}\left|f_{i}-f_{i}^{\prime}\right|, and it is similarly equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to a Lp(ν)L_{p}(\nu) Schnorr test.

To see that the partially defined function limnfn\lim_{n}f_{n} is a version of ff in Lp(ν)L_{p}(\nu), simply note that classically some subsequence fn:=fm(n)f_{n}^{\prime}:=f_{m(n)} converges to ff ν\nu-a.s. and so limnfn\lim_{n}f_{n}^{\prime} is a version of ff. The sequence fnf_{n}^{\prime} is computable in some oracle, and so by the previous paragraph we get that limnfn,limnfn\lim_{n}f_{n},\lim_{n}f_{n}^{\prime} agree on all the Schnorr randoms relative to that oracle, and so limnfn\lim_{n}f_{n} is also a version of ff.

The final remark follows from the second paragraph by choosing fnf_{n}^{\prime} to be the approximation to the lsc function ff from Proposition 2.16. ∎

The following is an analogue of the above proposition for L0(ν)L_{0}(\nu). This proposition is essentially the natural effectivization of the classical proof that Cauchy-in-measure sequences converge in measure.818181E.g. [21, Theorem 2.30 p. 61].

Proposition 4.2.

Suppose that ff is a computable point of L0(ν)L_{0}(\nu). Suppose that fnf_{n} is a computable sequence from the countable dense set of L0(ν)L_{0}(\nu) such that fnff_{n}\rightarrow f at a geometric rate of convergence in L0(ν)L_{0}(\nu). Then for all xx in 𝖲𝖱ν\mathsf{SR}^{\nu} one has that limnfn(x)\lim_{n}f_{n}(x) exists and is a version of ff.

This limit does not depend on the choice of fnf_{n} or the choice of the ν\nu-computable basis or the choice of the rate of geometric convergence.

Hence, if ff is in addition an L0(ν)L_{0}(\nu) Schnorr test, then limnfn(x)=f(x)\lim_{n}f_{n}(x)=f(x) for all xx in 𝖲𝖱ν\mathsf{SR}^{\nu}.

Note that by the computable embedding of Lp(ν)L_{p}(\nu) into L0(ν)L_{0}(\nu), the limit in this proposition agrees with the limit in the previous proposition on 𝖲𝖱ν\mathsf{SR}^{\nu}.

Proof.

We may suppose that the geometric rate of convergence b>1b>1 is rational. Then we can compute whether bjb^{-j} is rational or irrational, and hence uniformly in j0j\geq 0 we have that bjb^{-j} has uniformly computable left- and right Dedekind cuts. Since the fjf_{j} are from the countable dense set, so is |fjfj+1|\left|f_{j}-f_{j+1}\right|, and hence we can write it as k=1njqj,kIAj,k\sum_{k=1}^{n_{j}}q_{j,k}\cdot I_{A_{j,k}}, where qj,k0q_{j,k}\geq 0 is rational and the events {Aj,k:1knj}\{A_{j,k}:1\leq k\leq n_{j}\} are pairwise disjoint and come from the algebra generated by a ν\nu-computable basis. By Proposition 2.13, this is equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to the finite sum k=1njqj,kIUj,k\sum_{k=1}^{n_{j}}q_{j,k}\cdot I_{U_{j,k}}, where Uj,kU_{j,k} is a c.e. open which is equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to Aj,kA_{j,k}. Let Ej={xX:k=1njqj,kIUj,k>2bj}E_{j}=\{x\in X:\sum_{k=1}^{n_{j}}q_{j,k}\cdot I_{U_{j,k}}>2\cdot b^{-j}\}, which is a c.e. open since it is equal to KJjkKUj,k\bigcup_{K\in J_{j}}\bigcap_{k\in K}U_{j,k}, where Jj={K[1,nj]:kKqj,k>2bj}J_{j}=\{K\subseteq[1,n_{j}]:\sum_{k\in K}q_{j,k}>2\cdot b^{-j}\}, and JjJ_{j} is computable since bjb^{-j} has uniformly computable right Dedkind cuts. Then EjE_{j} is equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to {xX:|fjfj+1|>2bj}\{x\in X:\left|f_{j}-f_{j+1}\right|>2\cdot b^{-j}\} and EjE_{j} is a c.e. open with computable ν\nu-measure, uniformly in j0j\geq 0.

We then have ν(Ej)ν(|ffj|>bj)+ν(|ffj+1|>b(j+1))bj+b(j+1)2bj\nu(E_{j})\leq\nu(\left|f-f_{j}\right|>b^{-j})+\nu(\left|f-f_{j+1}\right|>b^{-(j+1)})\leq b^{-j}+b^{-(j+1)}\leq 2\cdot b^{-j}. Letting FkF_{k} be the c.e. open j=kEj\bigcup_{j=k}^{\infty}E_{j}, we have ν(Fk)2bb1bk\nu(F_{k})\leq 2\cdot\frac{b}{b-1}\cdot b^{-k}. Further for k>kk^{\prime}>k we have j=kkEj\bigcup_{j=k}^{k^{\prime}}E_{j} has computable ν\nu-measure since it is a finite union of events with ν\nu-computable measure coming from the algebra generated by a ν\nu-computable basis. And then ν(Fk)\nu(F_{k}) is computable since we can approximate it by ν(j=kkEj)\nu(\bigcup_{j=k}^{k^{\prime}}E_{j}) since ν(Fk)ν(j=kkEj)ν(Fk)2bb1bk\nu(F_{k})-\nu(\bigcup_{j=k}^{k^{\prime}}E_{j})\leq\nu(F_{k^{\prime}})\leq 2\cdot\frac{b}{b-1}\cdot b^{-k^{\prime}}. Hence kIFk\sum_{k}I_{F_{k}} is an L1(ν)L_{1}(\nu) Schnorr test.

If a point is in 𝖲𝖱ν\mathsf{SR}^{\nu}, then it is not in some FkF_{k}, while it is in 𝖪𝖱ν\mathsf{KR}^{\nu}. Then we argue for the following six items about elements of 𝖪𝖱νFk\mathsf{KR}^{\nu}\setminus F_{k}:

  1. (1)

    For all k0k\geq 0 and all xx in 𝖪𝖱νFk\mathsf{KR}^{\nu}\setminus F_{k}, for all j1>j0kj_{1}>j_{0}\geq k we have |fj0(x)fj1(x)|i=j0j11|fi(x)fi+1(x)|i=j02bi2bb1bj0\left|f_{j_{0}}(x)-f_{j_{1}}(x)\right|\leq\sum_{i=j_{0}}^{j_{1}-1}\left|f_{i}(x)-f_{i+1}(x)\right|\leq\sum_{i=j_{0}}^{\infty}2\cdot b^{-i}\leq 2\cdot\frac{b}{b-1}\cdot b^{-j_{0}}.

  2. (2)

    Hence for all k0k\geq 0 and all xx in 𝖪𝖱νFk\mathsf{KR}^{\nu}\setminus F_{k}, we have that fj(x)f_{j}(x) for jkj\geq k is a Cauchy sequence and thus limjfj(x)\lim_{j}f_{j}(x) exists.

  3. (3)

    For all xx in 𝖪𝖱νFk\mathsf{KR}^{\nu}\setminus F_{k} and all jk0j\geq k\geq 0, we have |fj(x)limjfj(x)|2bb1bj\left|f_{j}(x)-\lim_{j}f_{j}(x)\right|\leq 2\cdot\frac{b}{b-1}\cdot b^{-j}. For, let ϵ>0\epsilon>0. Let j0=jj_{0}=j and choose j1>j0j_{1}>j_{0} such that |fj1(x)limjfj(x)|<ϵ\left|f_{j_{1}}(x)-\lim_{j}f_{j}(x)\right|<\epsilon. Then by (1) one has that |fj(x)limjfj(x)||fj1(x)limjfj(x)|+|fj1(x)fj0(x)|<ϵ+2bb1bj0\left|f_{j}(x)-\lim_{j}f_{j}(x)\right|\leq\left|f_{j_{1}}(x)-\lim_{j}f_{j}(x)\right|+\left|f_{j_{1}}(x)-f_{j_{0}}(x)\right|<\epsilon+2\cdot\frac{b}{b-1}\cdot b^{-j_{0}}. Since this holds for all ϵ>0\epsilon>0, we are done.

  4. (4)

    Since 𝖲𝖱ν\mathsf{SR}^{\nu} is a ν\nu-measure one set, one has that limjfj\lim_{j}f_{j} exists ν\nu-a.s.

  5. (5)

    Further one has that fjlimjfjf_{j}\rightarrow\lim_{j}f_{j} in L0(ν)L_{0}(\nu). For let ϵ>0\epsilon>0. Choose kk such that 2bb1bk<ϵ2\cdot\frac{b}{b-1}\cdot b^{-k}<\epsilon. Let jkj\geq k, so that 2bb1bj<ϵ2\cdot\frac{b}{b-1}\cdot b^{-j}<\epsilon. Then by (3) we have ν(|fjlimjfj|>ϵ)ν(|fjlimjfj|>2bb1bj)ν(Fk)2bb1bk<ϵ\nu(\left|f_{j}-\lim_{j}f_{j}\right|>\epsilon)\leq\nu(\left|f_{j}-\lim_{j}f_{j}\right|>2\cdot\frac{b}{b-1}\cdot b^{-j})\leq\nu(F_{k})\leq 2\cdot\frac{b}{b-1}\cdot b^{-k}<\epsilon.

  6. (6)

    Since both fjlimjfjf_{j}\rightarrow\lim_{j}f_{j} in L0(ν)L_{0}(\nu) and fjff_{j}\rightarrow f in L0(ν)L_{0}(\nu), we have that limjfj=f\lim_{j}f_{j}=f ν\nu-a.s.

Suppose that hjh_{j} is another computable sequence from the countable dense set of L0(ν)L_{0}(\nu) such that hjfh_{j}\rightarrow f at a geometric rate c>1c>1 of convergence in L0(ν)L_{0}(\nu). Note that limjfj\lim_{j}f_{j} and limjhj\lim_{j}h_{j} are equal ν\nu-a.s. since they are both equal ν\nu-a.s. to ff. Let GjG_{j} and HkH_{k} be constructed from hjh_{j} and cc just as we constructed EjE_{j} and FkF_{k} from fjf_{j} and bb above. Let d=min(b,c)d=\min(b,c), rational number >1>1. Let e=max{2bb1,2cc1}e=\max\{2\cdot\frac{b}{b-1},2\cdot\frac{c}{c-1}\}. Note that for all j0j\geq 0, we have edj2bb1bje\cdot d^{-j}\geq 2\cdot\frac{b}{b-1}\cdot b^{-j} and edj2cc1cje\cdot d^{-j}\geq 2\cdot\frac{c}{c-1}\cdot c^{-j}. Let DjD_{j} be a c.e. open which is equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to {xX:|fj(x)hj(x)|>3edj}\{x\in X:\left|f_{j}(x)-h_{j}(x)\right|>3\cdot e\cdot d^{-j}\}, and note that DjD_{j} has computable ν\nu-measure, as in the argument of the first paragraph of the proof. Then one has that ν(Dj)ν(|fjlimjfj|>2bb1bj)+ν(|limjfjlimjhj|>edj)+ν(|hjlimjhj|>2cc1cj)ν(Fj)+0+ν(Hj)2bb1bj+2cc1cj\nu(D_{j})\leq\nu(\left|f_{j}-\lim_{j}f_{j}\right|>2\cdot\frac{b}{b-1}\cdot b^{-j})+\nu(\left|\lim_{j}f_{j}-\lim_{j}h_{j}\right|>e\cdot d^{-j})+\nu(\left|h_{j}-\lim_{j}h_{j}\right|>2\cdot\frac{c}{c-1}\cdot c^{-j})\leq\nu(F_{j})+0+\nu(H_{j})\leq 2\cdot\frac{b}{b-1}\cdot b^{-j}+2\cdot\frac{c}{c-1}\cdot c^{-j}, where the middle term is zero since limjfj\lim_{j}f_{j} and limjhj\lim_{j}h_{j} are equal ν\nu-a.s. Hence jIDj\sum_{j}I_{D_{j}} is an L1(ν)L_{1}(\nu) integral test, and thus, for xx in 𝖲𝖱ν\mathsf{SR}^{\nu} one has that limjfj(x)=limjhj(x)\lim_{j}f_{j}(x)=\lim_{j}h_{j}(x).

As in the previous proof, the limit does not depend on the choice of ν\nu-computable basis since ν\nu-computable bases are closed under effective unions (cf. Proposition 2.8).

The remarks about limnfn\lim_{n}f_{n} being a version of ff, and the remark about L0(ν)L_{0}(\nu) Schnorr tests, follows as in the proof of the previous proposition. ∎

There is a result similar to Proposition 4.1 when the fnf_{n} are themselves Lp(ν)L_{p}(\nu) Schnorr tests:

Proposition 4.3.

Suppose that p1p\geq 1 is computable (resp. p=0p=0). Suppose that fnf_{n} are uniformly Lp(ν)L_{p}(\nu) Schnorr tests with fnff_{n}\rightarrow f fast in Lp(ν)L_{p}(\nu), so that ff is also a computable point of Lp(ν)L_{p}(\nu). Then limnfn(x)\lim_{n}f_{n}(x) exists and for all xx in 𝖲𝖱ν\mathsf{SR}^{\nu}. If ff is also an Lp(ν)L_{p}(\nu) Schnorr test, then limnfn(x)=f(x)\lim_{n}f_{n}(x)=f(x) for all xx in 𝖲𝖱ν\mathsf{SR}^{\nu}.

Proof.

By Proposition 2.16 (resp. Proposition 2.30), choose doubly-indexed computable sequence fn,sf_{n,s} from the countable dense set of Lp(ν)L_{p}(\nu) such that for all n0n\geq 0 we have 0fn,sfn,s+1fn0\leq f_{n,s}\leq f_{n,s+1}\leq f_{n} everywhere and fn=supsfn,sf_{n}=\sup_{s}f_{n,s} and fn,sfnf_{n,s}\rightarrow f_{n} fast in Lp(ν)L_{p}(\nu). Then fn+1,n+1ff_{n+1,n+1}\rightarrow f fast in Lp(ν)L_{p}(\nu). Hence, by Proposition 4.1 (resp. Proposition 4.2), limnfn,n\lim_{n}f_{n,n} exists on 𝖲𝖱ν\mathsf{SR}^{\nu}. Note that by these propositions, if ff is also an Lp(ν)L_{p}(\nu) Schnorr test then limnfn,n=f\lim_{n}f_{n,n}=f on 𝖲𝖱ν\mathsf{SR}^{\nu}.

It suffices to show that limn(fnfn,n)=0\lim_{n}(f_{n}-f_{n,n})=0 on 𝖲𝖱ν\mathsf{SR}^{\nu}. Use Lemma 3.2 to rewrite what we are to show as limn(fngn,n)=0\lim_{n}(f_{n}\ominus g_{n,n})=0, where gn,ng_{n,n} is an Lp(ν)L_{p}(\nu) Schnorr test equal to fn,nf_{n,n} on 𝖪𝖱ν\mathsf{KR}^{\nu}, so that fngn,nf_{n}\ominus g_{n,n} is an Lp(ν)L_{p}(\nu) Schnorr test equal to fnfn,nf_{n}-f_{n,n} on 𝖪𝖱ν\mathsf{KR}^{\nu}. By Lemma 2.22 in conjunction with Proposition 2.21 (resp. Proposition 2.27), choose a computable sequence ηn\eta_{n} in the interval (2(n+1),2n)(2^{-(n+1)},2^{-n}) such that the (fngn,n)1(ηn,](f_{n}\ominus g_{n,n})^{-1}(\eta_{n},\infty] has computable ν\nu-measure. Then Un=(fngn,n)1(ηn,]U_{n}=(f_{n}\ominus g_{n,n})^{-1}(\eta_{n},\infty] is c.e. open with ν\nu-computable measure. Let hn=(fngn,n)IUnh_{n}=(f_{n}\ominus g_{n,n})\cdot I_{U_{n}} which is an Lp(ν)L_{p}(\nu) Schnorr test. Let h=nhnh=\sum_{n}h_{n} and hm=n<mhnh_{m}=\sum_{n<m}h_{n}. Then hhmpn>mhnpn>mfnfn,npn>m2n2m\|h-h_{m}\|_{p}\leq\sum_{n>m}\|h_{n}\|_{p}\leq\sum_{n>m}\|f_{n}-f_{n,n}\|_{p}\leq\sum_{n>m}2^{-n}\leq 2^{-m}. Then hh is an Lp(ν)L_{p}(\nu) Schnorr test: it is non-negative lsc as a sum of non-negative lsc functions, and the sequence hmh_{m} is uniformly Lp(ν)L_{p}(\nu)-computable and we just showed that hmhh_{m}\rightarrow h fast in Lp(ν)L_{p}(\nu). Now we verify limn(fnfn,n)=0\lim_{n}(f_{n}-f_{n,n})=0 on 𝖲𝖱ν\mathsf{SR}^{\nu}. Let xx in 𝖲𝖱ν\mathsf{SR}^{\nu}. Let ϵ>0\epsilon>0. Since xx is in 𝖲𝖱ν\mathsf{SR}^{\nu}, choose n00n_{0}\geq 0 such that we have the estimate nn0(fn(x)fn,n(x))IUn(x)<ϵ\sum_{n\geq n_{0}}(f_{n}(x)-f_{n,n}(x))\cdot I_{U_{n}}(x)<\epsilon. Choose n1n0n_{1}\geq n_{0} such that ηn<ϵ\eta_{n}<\epsilon for all nn1n\geq n_{1}. Let nn1n\geq n_{1}. If xx is in UnU_{n}, then by our estimate we have fn(x)fn,n(x)<ϵf_{n}(x)-f_{n,n}(x)<\epsilon. If xx is not in UnU_{n}, then by the definition of UnU_{n} we have fn(x)fn,n(x)ηn<ϵf_{n}(x)-f_{n,n}(x)\leq\eta_{n}<\epsilon.

5. Classical features of the maximal function

Suppose that ν\nu is a point of 𝒫(X)\mathcal{P}(X) and n\mathscr{F}_{n} is any increasing filtration of Borel subsets of XX. In this section, we recall some classical features of the maximal function f=supn𝔼ν[fn]f^{\ast}=\sup_{n}\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}] of an integrable function ff.

First, we recall the following, which gives us information about the codomain of the maximal function:

Lemma 5.1.
  1. (1)

    If p>1p>1 then gppp1gp\|g^{\ast}\|_{p}\leq\frac{p}{p-1}\cdot\|g\|_{p} for gg in Lp(ν)L_{p}(\nu), and the maximal function maps Lp(ν)L_{p}(\nu) to Lp(ν)L_{p}(\nu).

  2. (2)

    If p=1p=1, then the maximal function maps Lp(ν)L_{p}(\nu) to L0(ν)L_{0}(\nu).

Proof.

For p>1p>1 and gg in Lp(ν)L_{p}(\nu), the sequence 𝔼ν[gn]\mathbb{E}_{\nu}[g\mid\mathscr{F}_{n}] is a non-negative martingale, and so by Doob’s Maximal Inequality828282[25, Theorem 9.4 pp. 505-506]. followed by conditional Jensen we have: supnm𝔼ν[gn]ppp1𝔼ν[gm]ppp1gp\|\sup_{n\leq m}\mathbb{E}_{\nu}[g\mid\mathscr{F}_{n}]\|_{p}\leq\frac{p}{p-1}\cdot\|\mathbb{E}_{\nu}[g\mid\mathscr{F}_{m}]\|_{p}\leq\frac{p}{p-1}\cdot\|g\|_{p}. Then by the Monotone Convergence Theorem we have gppp1gp\|g^{\ast}\|_{p}\leq\frac{p}{p-1}\|g\|_{p}. For p=1p=1, let \mathscr{F}_{\infty} be the σ\sigma-algebra generated by all the n\mathscr{F}_{n}. By the classical Lévy Upward Theorem we have that 𝔼ν[gn]𝔼ν[g]\mathbb{E}_{\nu}[g\mid\mathscr{F}_{n}]\rightarrow\mathbb{E}_{\nu}[g\mid\mathscr{F}_{\infty}] ν\nu-a.s. which shows that gg^{\ast} is finite ν\nu-a.s., and hence that gg^{\ast} is in L0(ν)L_{0}(\nu). ∎

The following proposition collects together all the other classical facts about the maximal function which we need:

Proposition 5.2.

  1. (1)

    For p>1p>1, the maximal function :Lp(ν)Lp(ν){\cdot}^{\ast}:L_{p}(\nu)\rightarrow L_{p}(\nu) is uniformly continuous with modulus m(ϵ)=p1pϵm(\epsilon)=\frac{p-1}{p}\cdot\epsilon of uniform continuity.

  2. (2)

    For p=1p=1, the maximal function :Lp(ν)L0(ν){\cdot}^{\ast}:L_{p}(\nu)\rightarrow L_{0}(\nu) is uniformly continuous with a modulus m(ϵ)=ϵ2m(\epsilon)=\epsilon^{2} of uniform continuity.

Proof.

Let p>1p>1 and ff in Lp(ν)L_{p}(\nu). By Lemma 5.1 one has for f,gf,g in Lp(ν)L_{p}(\nu) with fgp<p1pϵ\|f-g\|_{p}<\frac{p-1}{p}\cdot\epsilon that fgp|fg|ppp1fgp<ϵ\|f^{\ast}-g^{\ast}\|_{p}\leq\|\left|f-g\right|^{\ast}\|_{p}\leq\frac{p}{p-1}\cdot\|f-g\|_{p}<\epsilon.

Suppose that f,gf,g are in L1(ν)L_{1}(\nu) with fg1<ϵ2\|f-g\|_{1}<\epsilon^{2}. Then one has

ν(|fg|>ϵ)ν(|fg|)>ϵ)=limmν(supnm𝔼[|fg|n]>ϵ)\displaystyle\nu(\left|f^{\ast}-g^{\ast}\right|>\epsilon)\leq\nu(\left|f-g\right|^{\ast})>\epsilon)=\lim_{m}\nu(\sup_{n\leq m}\mathbb{E}[\left|f-g\right|\mid\mathscr{F}_{n}]>\epsilon)
\displaystyle\leq limmϵ1𝔼ν𝔼ν[|fg|m]ϵ1𝔼ν|fg|ϵ1ϵ2=ϵ\displaystyle\lim_{m}\epsilon^{-1}\cdot\mathbb{E}_{\nu}\mathbb{E}_{\nu}[\left|f-g\right|\mid\mathscr{F}_{m}]\leq\epsilon^{-1}\cdot\mathbb{E}_{\nu}\left|f-g\right|\leq\epsilon^{-1}\epsilon^{2}=\epsilon

The first step of the second line follows from Doob’s Submartingale Inequality,838383[75, 137-138]. where we apply it to the martingale 𝔼[|fg|n]\mathbb{E}[\left|f-g\right|\mid\mathscr{F}_{n}]. ∎

6. An abstract version of Lévy’s Theorem for Schnorr randomness

Before we state our abstract version of Lévy’s Theorem, we need the following definition:

Definition 6.1.

Suppose that 𝖷𝖱ν\mathsf{XR}^{\nu} has ν\nu-measure one.

Suppose that 𝒞\mathcal{C} is a class of Lp(ν)L_{p}(\nu) Schnorr tests.

Then the Lp(ν)L_{p}(\nu) Schnorr tests are approximated from below on 𝖷𝖱ν\mathsf{XR}^{\nu} by 𝒞\mathcal{C} if from an index for an Lp(ν)L_{p}(\nu) Schnorr test ff one can compute

  • an index for a sequence of Lp(ν)L_{p}(\nu) Schnorr tests fsf_{s} in 𝒞\mathcal{C} such that fsfs+1f_{s}\leq f_{s+1} on 𝖷𝖱ν\mathsf{XR}^{\nu} and f=supsfsf=\sup_{s}f_{s} on 𝖷𝖱ν\mathsf{XR}^{\nu} and fsff_{s}\rightarrow f fast in Lp(ν)L_{p}(\nu); and

  • an index for a sequence of non-negative usc functions f˘s\breve{f}_{s} equal to fsf_{s} on 𝖷𝖱ν\mathsf{XR}^{\nu}.

Recall from Proposition 3.1 that the fs,f˘sf_{s},\breve{f}_{s} are finite on 𝖷𝖱ν\mathsf{XR}^{\nu}, and we define ffs=max(0,ff˘s)f\ominus f_{s}=\max(0,f-\breve{f}_{s}), and we have that ffsf\ominus f_{s} is an Lp(ν)L_{p}(\nu) Schnorr test equal on 𝖷𝖱ν\mathsf{XR}^{\nu} to ffsf-f_{s}. Similarly if tst\geq s, then we define ftfs=max(0,ftf˘s)f_{t}\ominus f_{s}=\max(0,f_{t}-\breve{f}_{s}), and we have that ftfsf_{t}\ominus f_{s} is an Lp(ν)L_{p}(\nu) Schnorr test equal on 𝖷𝖱ν\mathsf{XR}^{\nu} to ftfsf_{t}-f_{s}.

The basic example of Definition 6.1 comes from Lemma 3.2.

The following is our abstract version of Lévy’s Theorem for Schnorr randomness. It is an abstract version in that we are not told more about the function E[n]()E[\cdot\mid n](\cdot) other than that what is stated explicitly in the hypotheses (I)-(IV). In particular, we do not assume that E[n]()E[\cdot\mid n](\cdot) comes from an effective disintegration, although in the next section we will show that effective disintegrations satisfy the hypotheses of the theorem.

Theorem 6.2.

Suppose that ν\nu is a computable point of 𝒫(X)\mathcal{P}(X). Suppose that n\mathscr{F}_{n} is an increasing filtration of Borel sets. Suppose that p1p\geq 1 is computable.

Suppose that E[n]():𝕃p+(ν)×X[0,]E[\cdot\mid n](\cdot):\mathbb{L}_{p}^{+}(\nu)\times X\rightarrow[0,\infty] is a function such that for every ff in 𝕃p+(ν)\mathbb{L}_{p}^{+}(\nu), one has that E[fn]:X[0,]E[f\mid n]:X\rightarrow[0,\infty] is a version of the conditional expectation of ff with respect to n\mathscr{F}_{n}. Define the function ():𝕃p+(ν)×X[0,]\cdot^{\flat}(\cdot):\mathbb{L}_{p}^{+}(\nu)\times X\rightarrow[0,\infty] by f(x)=supnE[n](x)f^{\flat}(x)=\sup_{n}E[\cdot\mid n](x).

Suppose that 𝖷𝖱ν\mathsf{XR}^{\nu} is a superset of 𝖲𝖱ν\mathsf{SR}^{\nu}.

Suppose that:

  1. (I)

    E[n]E[\cdot\mid n] maps non-negative lsc functions to non-negative lsc functions.

  2. (II)

    E[n]E[\cdot\mid n] satisfies the following properties on 𝖷𝖱ν\mathsf{XR}^{\nu}:

    1. (a)

      If fgf\leq g on 𝖷𝖱ν\mathsf{XR}^{\nu}, then E[fn]E[gn]E[f\mid n]\leq E[g\mid n] on 𝖷𝖱ν\mathsf{XR}^{\nu};

    2. (b)

      If cc in 0\mathbb{R}^{\geq 0} then cE[fn]=E[cfn]c\cdot E[f\mid n]=E[c\cdot f\mid n] on 𝖷𝖱ν\mathsf{XR}^{\nu};

    3. (c)

      E[f+gn]=E[fn]+E[gn]E[f+g\mid n]=E[f\mid n]+E[g\mid n] on 𝖷𝖱ν\mathsf{XR}^{\nu};

  3. (III)

    Both E[n]E[\cdot\mid n] and \cdot^{\flat} send the countable dense set of Lp+(ν)L_{p}^{+}(\nu) uniformly to computable points of Lp+(ν)L_{p}^{+}(\nu).

  4. (IV)

    The Lp(ν)L_{p}(\nu) Schnorr tests are approximated from below on 𝖷𝖱ν\mathsf{XR}^{\nu} by a class 𝒞\mathcal{C} of Lp(ν)L_{p}(\nu) Schnorr tests such that limnE[fn]=f\lim_{n}E[f\mid n]=f on 𝖷𝖱ν\mathsf{XR}^{\nu} for each ff in 𝒞\mathcal{C}.

Then the following three items are equivalent for xx in XX:

  1. (1)

    xx is in 𝖲𝖱ν\mathsf{SR}^{\nu}.

  2. (2)

    xx is in 𝖷𝖱ν\mathsf{XR}^{\nu} and limnE[fn](x)=f(x)\lim_{n}E[f\mid n](x)=f(x) for every Lp(ν)L_{p}(\nu) Schnorr test ff.

  3. (3)

    xx is in 𝖷𝖱ν\mathsf{XR}^{\nu} and limnE[fn](x)\lim_{n}E[f\mid n](x) exists for every Lp(ν)L_{p}(\nu) Schnorr test ff and limnE[IUn](x)=IU(x)\lim_{n}E[I_{U}\mid n](x)=I_{U}(x) for every c.e. open UU with ν\nu-computable measure.

We also have:

  1. (i)

    Every Borel set BB is equal on 𝖲𝖱ν\mathsf{SR}^{\nu} to a Borel set BB^{\prime} in the σ\sigma-algebra generated by the union of the n\mathscr{F}_{n}.848484Indeed, if n1n\geq 1 and BB is in Σ~n0\utilde{\Sigma}^{0}_{n} (resp. Π~n0\utilde{\Pi}^{0}_{n}) then we can take BB^{\prime} to be Σ~n+50\utilde{\Sigma}^{0}_{n+5} (resp. Π~n+50\utilde{\Pi}^{0}_{n+5}), and if αω\alpha\geq\omega and BB is Σ~α0\utilde{\Sigma}^{0}_{\alpha} (resp. Π~α0\utilde{\Pi}^{0}_{\alpha}) then BB^{\prime} may be taken to be Σ~α0\utilde{\Sigma}^{0}_{\alpha} (resp. Π~α0\utilde{\Pi}^{0}_{\alpha}). And the same for the lightface classes.

  2. (ii)

    Suppose one adds to (IV) the condition that every xx in 𝖲𝖱ν\mathsf{SR}^{\nu} weakly computes a modulus of convergence for E[fn](x)f(x)E[f\mid n](x)\rightarrow f(x), uniformly in ff from 𝒞\mathcal{C}. Then one can further conclude for every xx in 𝖲𝖱ν\mathsf{SR}^{\nu} and every Lp(ν)L_{p}(\nu) Schnorr test ff, the point xx weakly computes a modulus of convergence for E[fn](x)f(x)E[f\mid n](x)\rightarrow f(x) in (2).

Regarding (i), note that this is saying that the hypotheses of the Theorem amount collectively to an assumption that the union of the filtration generates a σ\sigma-algebra very close to the Borel σ\sigma-algebra, from the perspective of ν\nu.

Proof.

First we note three things about the maps E[n]E[\cdot\mid n] and \cdot^{\flat} and p1p\geq 1 computable.

For p1p\geq 1, the map E[n]:𝕃p+(ν)𝕃p+(ν)E[\cdot\mid n]:\mathbb{L}_{p}^{+}(\nu)\rightarrow\mathbb{L}_{p}^{+}(\nu) maps Lp(ν)L_{p}(\nu) Schnorr tests uniformly to Lp(ν)L_{p}(\nu) Schnorr tests. For, by (I), it sends non-negative lsc functions to non-negative lsc functions. And by conditional Jensen and (III) and Propositions 2.4,2.5, it sends Lp+(ν)L_{p}^{+}(\nu) computable points to Lp+(ν)L_{p}^{+}(\nu) computable points.

For p>1p>1, the map :𝕃p+(ν)𝕃p+(ν)\cdot^{\flat}:\mathbb{L}_{p}^{+}(\nu)\rightarrow\mathbb{L}_{p}^{+}(\nu) maps Lp(ν)L_{p}(\nu) Schnorr tests uniformly to Lp(ν)L_{p}(\nu) Schnorr tests. For, by (I), it sends non-negative lsc functions to non-negative lsc functions. And by Proposition 5.2(1) and (III) and Propositions 2.4,2.5, it sends Lp+(ν)L_{p}^{+}(\nu) computable points to Lp+(ν)L_{p}^{+}(\nu) computable points.

For p=1p=1, the map :𝕃p+(ν)𝕃0+(ν)\cdot^{\flat}:\mathbb{L}_{p}^{+}(\nu)\rightarrow\mathbb{L}_{0}^{+}(\nu) sends Lp(ν)L_{p}(\nu) Schnorr tests uniformly to L0(ν)L_{0}(\nu) Schnorr tests. For, by (I), it sends non-negative lsc functions to non-negative lsc functions. And by Proposition 5.2(2) and (III) and Propositions 2.4,2.5 it sends Lp+(ν)L_{p}^{+}(\nu) computable points to L0+(ν)L_{0}^{+}(\nu) computable points.

Now we work on the equivalence of (1)-(3).

Suppose (1); we show (2). Suppose that ff is an Lp(ν)L_{p}(\nu) Schnorr test; we want to show that f=limnE[fn]f=\lim_{n}E[f\mid n] on 𝖲𝖱ν\mathsf{SR}^{\nu}. Choose fsf_{s} from 𝒞\mathcal{C} as in (IV). By definition, one has that fsff_{s}\rightarrow f pointwise on 𝖷𝖱ν\mathsf{XR}^{\nu} and fast in Lp(ν)L_{p}(\nu) and is non-decreasing on 𝖷𝖱ν\mathsf{XR}^{\nu}. Let gsg_{s} be the Lp(ν)L_{p}(\nu) Schnorr test ffsf\ominus f_{s}. Then gs0g_{s}\rightarrow 0 pointwise on 𝖷𝖱ν\mathsf{XR}^{\nu} and is non-increasing on 𝖷𝖱ν\mathsf{XR}^{\nu} and gs0g_{s}\rightarrow 0 fast in Lp(ν)L_{p}(\nu).

Suppose p>1p>1 (resp. p=1p=1). Since fsf_{s} is finite on 𝖷𝖱ν\mathsf{XR}^{\nu}, we have that f=fs+ffsf=f_{s}+f\ominus f_{s} on 𝖷𝖱ν\mathsf{XR}^{\nu} and hence by (IIc) we have E[fn]=E[fsn]+E[ffsn]E[fsn]+gsE[f\mid n]=E[f_{s}\mid n]+E[f\ominus f_{s}\mid n]\leq E[f_{s}\mid n]+g_{s}^{\flat} on 𝖷𝖱ν\mathsf{XR}^{\nu}. These are all Lp(ν)L_{p}(\nu) Schnorr tests (resp. except for gsg_{s}^{\flat} which is an L0(ν)L_{0}(\nu) Schnorr test), and so they are finite on 𝖲𝖱ν\mathsf{SR}^{\nu} and we hence have E[fn]E[fsn]gsE[f\mid n]-E[f_{s}\mid n]\leq g_{s}^{\flat} on 𝖲𝖱ν\mathsf{SR}^{\nu}. Since fsff_{s}\leq f on 𝖷𝖱ν\mathsf{XR}^{\nu}, we have E[fsn]E[fn]E[f_{s}\mid n]\leq E[f\mid n] on 𝖷𝖱ν\mathsf{XR}^{\nu} by (IIa). Hence |E[fn]E[fsn]|gs\left|E[f\mid n]-E[f_{s}\mid n]\right|\leq g_{s}^{\flat} on 𝖲𝖱ν\mathsf{SR}^{\nu}. Since the maximal function maps into Lp(ν)L_{p}(\nu) (resp. L0(ν)L_{0}(\nu)) and has a computable modulus of uniform continuity (cf. Proposition 5.2), and since gs0g_{s}\rightarrow 0 fast in Lp(ν)L_{p}(\nu) (resp. at a geometric rate in L0(ν)L_{0}(\nu)), one can compute a subsequence such that gs(n)0g^{\flat}_{s(n)}\rightarrow 0 fast in Lp(ν)L_{p}(\nu) (resp. in L0(ν)L_{0}(\nu)). By Proposition 4.3 one has that gs(n)0g_{s(n)}^{\flat}\rightarrow 0 pointwise on 𝖲𝖱ν\mathsf{SR}^{\nu}. Let xx in 𝖲𝖱ν\mathsf{SR}^{\nu} and let ϵ>0\epsilon>0. Since f,fs,E[fn],E[fsn]f,f_{s},E[f\mid n],E[f_{s}\mid n] are Lp(ν)L_{p}(\nu) Schnorr tests and since gsg_{s}^{\flat} is an Lp(ν)L_{p}(\nu) Schnorr test (resp. L0(ν)L_{0}(\nu) Schnorr test), these values are all finite on the 𝖲𝖱ν\mathsf{SR}^{\nu} point xx. Choose n00n_{0}\geq 0 such that for all nn0n\geq n_{0} and f(x)fn(x)<ϵ3f(x)-f_{n}(x)<\frac{\epsilon}{3}. Choose n1n0n_{1}\geq n_{0} such that for all nn1n\geq n_{1} one has gs(n)(x)<ϵ3g_{s(n)}^{\flat}(x)<\frac{\epsilon}{3}. By the hypothesis on the fsf_{s} coming from 𝒞\mathcal{C}, choose n2n1n_{2}\geq n_{1} such that |fs(n1)(x)E[fs(n1)n](x)|<ϵ3\left|f_{s(n_{1})}(x)-E[f_{s(n_{1})}\mid n](x)\right|<\frac{\epsilon}{3} for all nn2n\geq n_{2}. Hence for all nn2n\geq n_{2} one has that |f(x)E[fn](x)||f(x)fs(n1)(x)|+|fs(n1)(x)E[fs(n1)n](x)|+|E[fs(n1)n](x)E[fn](x)|<ϵ3+ϵ3+gs(n1)(x)<ϵ\left|f(x)-E[f\mid n](x)\right|\leq\left|f(x)-f_{s(n_{1})}(x)\right|+\left|f_{s(n_{1})}(x)-E[f_{s(n_{1})}\mid n](x)\right|+\left|E[f_{s(n_{1})}\mid n](x)-E[f\mid n](x)\right|<\frac{\epsilon}{3}+\frac{\epsilon}{3}+g_{s(n_{1})}^{\flat}(x)<\epsilon.

Note that the previous paragraph yields (ii). For, Proposition 3.4(1) tells us that xx can weakly compute a modulus of convergence for gs(n)(x)0g_{s(n)}^{\flat}(x)\rightarrow 0. And Proposition 3.4(2) tells us that xx can weakly compute a modulus of convergence for fn(x)f(x)f_{n}(x)\rightarrow f(x). And the extra hypothesis in (ii) says that xx can weakly compute a modulus of convergence for E[fs(n1)n](x)fs(n1)(x)E[f_{s(n_{1})}\mid n](x)\rightarrow f_{s(n_{1})}(x).

The implication from (2) to (3) is trivial.

Suppose (3); we show (1). Suppose that xx is a point satisfying (3). We want to show that xx is in 𝖲𝖱ν\mathsf{SR}^{\nu}. Let ff be an Lp(ν)L_{p}(\nu) Schnorr test. We want to show that f(x)<f(x)<\infty. Suppose for reductio that f(x)=f(x)=\infty. By hypothesis, limnE[fn](x)\lim_{n}E[f\mid n](x) exists and is finite. Choose n00n_{0}\geq 0 and rationals p<qp<q such that E[fn](x)<pE[f\mid n](x)<p for all nn0n\geq n_{0}. Then xx is in the c.e. open f1(q,]f^{-1}(q,\infty]. Using a ν\nu-computable basis, choose c.e. open UU which is a subset of f1(q,]f^{-1}(q,\infty] and which contains xx and which has computable ν\nu-measure. Let g=qIUg=q\cdot I_{U}, which is an Lp(ν)L_{p}(\nu) Schnorr test. Since gq0g-q\leq 0 everywhere, by (3), there is n1n0n_{1}\geq n_{0} such that qp>|E[gn](x)q|=qE[gn](x)q-p>\left|E[g\mid n](x)-q\right|=q-E[g\mid n](x) for all nn1n\geq n_{1}, and thus E[gn](x)>pE[g\mid n](x)>p for all such nn. Let nn1n\geq n_{1}. Since gfg\leq f everywhere, by (IIa), we have E[gn](x)E[fn](x)<pE[g\mid n](x)\leq E[f\mid n](x)<p, a contradiction.

For (i), by using a ν\nu-computable basis it suffices to prove it for UU c.e. open with ν(U)\nu(U) computable. In this case, f=IUf=I_{U} is an Lp(ν)L_{p}(\nu) Schnorr test. Then f=limnE[fn]f=\lim_{n}E[f\mid n] on 𝖲𝖱ν\mathsf{SR}^{\nu}. By (IIa), one has 0E[fn]10\leq E[f\mid n]\leq 1 on 𝖷𝖱ν\mathsf{XR}^{\nu}. For rational ϵ\epsilon in the interval (0,12)(0,\frac{1}{2}), let Vn,ϵ=E[fn]1(1ϵ,]V_{n,\epsilon}=E[f\mid n]^{-1}(1-\epsilon,\infty]. This set is in n\mathscr{F}_{n} since E[fn]E[f\mid n] is a version of the conditional expectation of ff relative to n\mathscr{F}_{n}. Further this set is c.e. open by the second paragraph of this proof. Then on 𝖲𝖱ν\mathsf{SR}^{\nu} one has that U=ϵ(0,12)n00nn0Vn,ϵU=\bigcap_{\epsilon\in\mathbb{Q}\cap(0,\frac{1}{2})}\bigcup_{n_{0}\geq 0}\bigcap_{n\geq n_{0}}V_{n,\epsilon}.858585This event is further in Π~40\utilde{\Pi}^{0}_{4}. When we decompose an arbitrary open as a union of c.e. opens with ν\nu-computable measure, we will get an event in Σ~50\utilde{\Sigma}^{0}_{5}. Hence, UU is equal on 𝖲𝖱ν\mathsf{SR}^{\nu} to an event BB^{\prime} in the σ\sigma-algebra generated by the union of the n\mathscr{F}_{n}. ∎

7. Fundamental properties of effective disintegrations

In this section, we develop the properties of effective disintegrations (cf. Definition 1.3, and for examples see Appendicies A-B). In the next propositions, 𝖷𝖱ν\mathsf{XR}^{\nu} denotes a ν\nu-measure one subset of 𝖪𝖱ν\mathsf{KR}^{\nu}, as in the definition of an effective disintegration. Further, throughout this section, the expression 𝔼ν[f]\mathbb{E}_{\nu}[f\mid\mathscr{F}] is defined as in (1.1) of §1.2, namely the version of conditional expectation coming from the effective disintegration.

Proposition 7.1.

Suppose ρ:X+(X)\rho:X\rightarrow\mathcal{M}^{+}(X) is an 𝖷𝖱ν\mathsf{XR}^{\nu} disintegration of \mathscr{F}. Then for lsc f:X[0,]f:X\rightarrow[0,\infty], the map 𝔼ν[f]:X[0,]\mathbb{E}_{\nu}[f\mid\mathscr{F}]:X\rightarrow[0,\infty] is lsc, uniformly in ff.

Proof.

By Definition 1.3(3) it suffices to show that for rational r0r\geq 0 we have

f𝑑ρx>r iff q>0(ρx(X)>rq&ρx(f1(q,])>0)\int f\;d\rho_{x}>r\mbox{ iff }\exists\;q\in\mathbb{Q}^{>0}\;\big{(}\rho_{x}(X)>\frac{r}{q}\;\&\;\rho_{x}(f^{-1}(q,\infty])>0\big{)}

Suppose that f𝑑ρx>r\int f\;d\rho_{x}>r. Since r0r\geq 0, we have ρx(X)>0\rho_{x}(X)>0. Choose rational q>0q>0 in the interval (rρx(X),f𝑑ρxρx(X))(\frac{r}{\rho_{x}(X)},\frac{\int f\;d\rho_{x}}{\rho_{x}(X)}). Then f𝑑ρx>ρx(X)q\int f\;d\rho_{x}>\rho_{x}(X)\cdot q and ρx(X)>rq\rho_{x}(X)>\frac{r}{q}. Then ρx(f1(q,])>0\rho_{x}(f^{-1}(q,\infty])>0, since otherwise 0fq0\leq f\leq q ρx\rho_{x}-a.e. and hence f𝑑ρxqρx(X)\int f\;d\rho_{x}\leq q\cdot\rho_{x}(X).

Suppose that rational q>0q>0 satisfies ρx(X)>rq\rho_{x}(X)>\frac{r}{q} and ρx(f1(q,])>0\rho_{x}(f^{-1}(q,\infty])>0. If ff is not ρx\rho_{x}-integrable then trivially we have f𝑑ρx>r\int f\;d\rho_{x}>r. Hence suppose ff is ρx\rho_{x}-integrable. Since ρx(f1(q,])>0\rho_{x}(f^{-1}(q,\infty])>0, choose ϵ>0\epsilon>0 such that ρx(fq>ϵ)>0\rho_{x}(f-q>\epsilon)>0. Then 0<ρx(fq>ϵ)1ϵfqdρx0<\rho_{x}(f-q>\epsilon)\leq\frac{1}{\epsilon}\int f-q\;d\rho_{x}. Then 0<fqdρx0<\int f-q\;d\rho_{x}. Then qρx(X)<f𝑑ρxq\cdot\rho_{x}(X)<\int f\;d\rho_{x}. Then r<f𝑑ρxr<\int f\;d\rho_{x}. ∎

Proposition 7.2.

Suppose ρ:X+(X)\rho:X\rightarrow\mathcal{M}^{+}(X) is a 𝖷𝖱ν\mathsf{XR}^{\nu} disintegration of \mathscr{F}. Then for f,gf,g in 𝕃1+(ν)\mathbb{L}_{1}^{+}(\nu) the conditional expectation satisfies the following monotone linearity properties:

  1. (1)

    If fgf\leq g on 𝖷𝖱ν\mathsf{XR}^{\nu}, then 𝔼ν[f]𝔼ν[g]\mathbb{E}_{\nu}[f\mid\mathscr{F}]\leq\mathbb{E}_{\nu}[g\mid\mathscr{F}] on 𝖷𝖱ν\mathsf{XR}^{\nu};

  2. (2)

    If cc in 0\mathbb{R}^{\geq 0} then c𝔼ν[f]=𝔼ν[cf]c\cdot\mathbb{E}_{\nu}[f\mid\mathscr{F}]=\mathbb{E}_{\nu}[c\cdot f\mid\mathscr{F}] everywhere.

  3. (3)

    𝔼ν[f+g]=𝔼ν[f]+𝔼ν[g]\mathbb{E}_{\nu}[f+g\mid\mathscr{F}]=\mathbb{E}_{\nu}[f\mid\mathscr{F}]+\mathbb{E}_{\nu}[g\mid\mathscr{F}] everywhere.

Proof.

For (1), suppose that fgf\leq g on 𝖷𝖱ν\mathsf{XR}^{\nu}. Suppose that xx is in 𝖷𝖱ν\mathsf{XR}^{\nu}. By Definition 1.3(2), ρx\rho_{x} is in 𝒫(X)\mathcal{P}(X) and ρx([x]𝖷𝖱ν)=1\rho_{x}([x]_{\mathscr{F}}\cap\mathsf{XR}^{\nu})=1. Then fgf\leq g on a ρx\rho_{x}-measure one set, and hence f(v)𝑑ρx(v)g(v)𝑑ρx(v)\int f(v)\;d\rho_{x}(v)\leq\int g(v)\;d\rho_{x}(v). For (2)-(3), these just follow from the properties of the integral. ∎

The use of Definition 1.3(2) in the proof of the previous proposition is typical, and henceforth we do not explicitly reiterate it as we go along.

We stated the previous proposition for non-negative functions. For these functions, the conditional expectation 𝔼ν[f](x)\mathbb{E}_{\nu}[f\mid\mathscr{F}](x) in (1.1) is automatically defined for all points xx in XX, even if it is infinite. However, 𝔼ν[f](x)\mathbb{E}_{\nu}[f\mid\mathscr{F}](x) in (1.1) is automatically defined and finite when ff is a simple function, and so the previous proposition holds for these functions as well. More generally, the previous proposition holds for functions which take negative values, provided that 𝔼ν[f](x)\mathbb{E}_{\nu}[f\mid\mathscr{F}](x) is defined and finite on all points xx of 𝖷𝖱ν\mathsf{XR}^{\nu}.

Proposition 7.3.

Suppose ρ:X+(X)\rho:X\rightarrow\mathcal{M}^{+}(X) is a 𝖷𝖱ν\mathsf{XR}^{\nu} disintegration of \mathscr{F}.

If ff in 𝕃1+(ν)\mathbb{L}^{+}_{1}(\nu) is equal on 𝖷𝖱ν\mathsf{XR}^{\nu} to a function which is \mathscr{F}-measurable, then one has 𝔼ν[f]=f\mathbb{E}_{\nu}[f\mid\mathscr{F}]=f on 𝖷𝖱ν\mathsf{XR}^{\nu}.

Proof.

By Proposition 7.2(1), it suffices to consider functions in 𝕃1+(ν)\mathbb{L}^{+}_{1}(\nu) which are themselves \mathscr{F}-measurable (as opposed to being merely equal on 𝖷𝖱ν\mathsf{XR}^{\nu} to such a function).

Suppose that xx is in 𝖷𝖱ν\mathsf{XR}^{\nu}. If 𝔼ν[f](x)<f(x)\mathbb{E}_{\nu}[f\mid\mathscr{F}](x)<f(x), then for some rationals p,qp,q we have 𝔼ν[f](x)<p<q<f(x)\mathbb{E}_{\nu}[f\mid\mathscr{F}](x)<p<q<f(x). Then xx is in the event f1(q,]f^{-1}(q,\infty] in \mathscr{F}. Then [x]f1(q,][x]_{\mathscr{F}}\subseteq f^{-1}(q,\infty]. Then fqf\geq q for ρx\rho_{x}-a.s. many values and hence f𝑑ρxq\int f\;d\rho_{x}\geq q, a contradiction. The case of 𝔼ν[f](x)>f(x)\mathbb{E}_{\nu}[f\mid\mathscr{F}](x)>f(x) is similar. ∎

In the below proposition, we use the traditional names for the properties of conditional expectation.868686E.g. [75, 88].. Of course, the hypothesis of effective disintegrations in Definition 1.3(1) is that 𝔼ν[f](x)=f𝑑ρx\mathbb{E}_{\nu}[f\mid\mathscr{F}](x)=\int f\;d\rho_{x} is a version of conditional expectation. But in this proposition and several others in this section, what we are verifying is that they hold pointwise on specifiable measure ν\nu-one subsets.

Proposition 7.4.

Suppose ρ:X+(X)\rho:X\rightarrow\mathcal{M}^{+}(X) is a 𝖷𝖱ν\mathsf{XR}^{\nu} disintegration of \mathscr{F}.

  1. (1)

    (Conditional MCT). Suppose that fn,ff_{n},f are in 𝕃1+(ν)\mathbb{L}_{1}^{+}(\nu) and 0fnfn+10\leq f_{n}\leq f_{n+1} on 𝖷𝖱ν\mathsf{XR}^{\nu} and limnfn=f\lim_{n}f_{n}=f on 𝖷𝖱ν\mathsf{XR}^{\nu}. Then limn𝔼ν[fn]=𝔼ν[f]\lim_{n}\mathbb{E}_{\nu}[f_{n}\mid\mathscr{F}]=\mathbb{E}_{\nu}[f\mid\mathscr{F}] on 𝖷𝖱ν\mathsf{XR}^{\nu}.

  2. (2)

    (Conditional DCT) Suppose that fn,f,gf_{n},f,g are in 𝕃1+(ν)\mathbb{L}_{1}^{+}(\nu) and |fn|g\left|f_{n}\right|\leq g on 𝖷𝖱ν\mathsf{XR}^{\nu} and limnfn=f\lim_{n}f_{n}=f on 𝖷𝖱ν\mathsf{XR}^{\nu}. Then:

    1. (a)

      If xx in 𝖷𝖱ν\mathsf{XR}^{\nu} and 𝔼ν[g](x)<\mathbb{E}_{\nu}[g\mid\mathscr{F}](x)<\infty then limn𝔼ν[fn](x)=𝔼ν[f](x)\lim_{n}\mathbb{E}_{\nu}[f_{n}\mid\mathscr{F}](x)=\mathbb{E}_{\nu}[f\mid\mathscr{F}](x).

    2. (b)

      If 𝔼ν[g]<\mathbb{E}_{\nu}[g\mid\mathscr{F}]<\infty on 𝖷𝖱ν\mathsf{XR}^{\nu}, then limn𝔼ν[fn]=𝔼ν[f]\lim_{n}\mathbb{E}_{\nu}[f_{n}\mid\mathscr{F}]=\mathbb{E}_{\nu}[f\mid\mathscr{F}] on 𝖷𝖱ν\mathsf{XR}^{\nu}.

  3. (3)

    (‘Taking out what is known’). Suppose that f,gf,g in 𝕃1+(ν)\mathbb{L}_{1}^{+}(\nu), and suppose that gg is equal on 𝖷𝖱ν\mathsf{XR}^{\nu} to a \mathscr{F}-measurable function. Then 𝔼ν[fg]=g𝔼ν[f]\mathbb{E}_{\nu}[f\cdot g\mid\mathscr{F}]=g\cdot\mathbb{E}_{\nu}[f\mid\mathscr{F}] on 𝖷𝖱ν\mathsf{XR}^{\nu}.

Proof.

For (1), let xx in 𝖷𝖱ν\mathsf{XR}^{\nu}. By hypothesis, 0fnfn+10\leq f_{n}\leq f_{n+1} for ρx\rho_{x}-a.s. many points, and likewise limnfn=f\lim_{n}f_{n}=f for ρx\rho_{x}-a.s. many points. Then by MCT applied to ρx\rho_{x} we have limnfn𝑑ρx=f𝑑ρx\lim_{n}\int f_{n}\;d\rho_{x}=\int f\;d\rho_{x}.

For (2), let xx in 𝖷𝖱ν\mathsf{XR}^{\nu} with 𝔼ν[g](x)<\mathbb{E}_{\nu}[g\mid\mathscr{F}](x)<\infty. This means that g𝑑ρx<\int g\;d\rho_{x}<\infty, and so gg is in 𝕃1+(ρx)\mathbb{L}^{+}_{1}(\rho_{x}). By hypothesis, |fn|g\left|f_{n}\right|\leq g for ρx\rho_{x}-a.s. many points, and likewise limnfn=f\lim_{n}f_{n}=f for ρx\rho_{x}-a.s. many points. Hence by the DCT applied to ρx\rho_{x}, we have that limnfn𝑑ρx=f𝑑ρx\lim_{n}\int f_{n}\;d\rho_{x}=\int f\;d\rho_{x}.

For (3), by Proposition 7.2(1), it suffices to prove it for gg which is itself \mathscr{F}-measurable. We show it by induction on complexity of gg.

Suppose g=IAg=I_{A} where AA is \mathscr{F}-measurable. If xx in AA then [x]A[x]_{\mathscr{F}}\subseteq A and then AA is a ρx\rho_{x}-measure one event and then it reduces to the observation that Af(v)𝑑ρx(v)=f(v)𝑑ρx(v)\int_{A}f(v)\;d\rho_{x}(v)=\int f(v)\;d\rho_{x}(v). If xx is not in AA then [x]XA[x]_{\mathscr{F}}\subseteq X\setminus A and then AA is a ρx\rho_{x}-measure zero event and then it reduces to the observation that Af(v)𝑑ρx(v)=0\int_{A}f(v)\;d\rho_{x}(v)=0.

By Proposition 7.2(2)-(3), it extends to simple functions. By Conditional MCT it extends to all elements of 𝕃1+(ν)\mathbb{L}_{1}^{+}(\nu). ∎

Unlike the previous propositions, this proposition concerns Kurtz disintegrations:

Proposition 7.5.

Suppose ρ:X+(X)\rho:X\rightarrow\mathcal{M}^{+}(X) is a Kurtz disintegration of \mathscr{F}. If p1p\geq 1 is computable, then 𝔼ν[]:Lp(ν)Lp(ν)\mathbb{E}_{\nu}[\cdot\mid\mathscr{F}]:L_{p}(\nu)\rightarrow L_{p}(\nu) is computable continuous.

Proof.

By conditional Jensen, the function m(ϵ)=ϵm(\epsilon)=\epsilon is a computable modulus of uniform continuity. Hence by Proposition 2.5, it suffices to show that if φ=i=1nqiIAi\varphi=\sum_{i=1}^{n}q_{i}\cdot I_{A_{i}} is an element of the countable dense set of Lp(ν)L_{p}(\nu), then 𝔼ν[φ]\mathbb{E}_{\nu}[\varphi\mid\mathscr{F}] is a computable point of Lp(ν)L_{p}(\nu). Since we can effectively separate φ\varphi into positive and negative parts, it suffices by the linearity of conditional expectation (Proposition 7.2) to consider the case where qi0q_{i}\geq 0. We may assume further that the AiA_{i} are pairwise disjoint, which like in the discussion at the beginning of §2.3 implies that φp=i=1nqipIAi\varphi^{p}=\sum_{i=1}^{n}q_{i}^{p}\cdot I_{A_{i}}.

By Proposition 2.13, let UiU_{i} be a c.e. open which is equal to AiA_{i} on 𝖪𝖱ν\mathsf{KR}^{\nu}. Let f=i=1nqiIUif=\sum_{i=1}^{n}q_{i}\cdot I_{U_{i}}, so that likewise fp=i=1nqiIUif^{p}=\sum_{i=1}^{n}q_{i}\cdot I_{U_{i}} on 𝖪𝖱ν\mathsf{KR}^{\nu}. Then f=φf=\varphi on 𝖪𝖱ν\mathsf{KR}^{\nu} and fp=φpf^{p}=\varphi^{p} on 𝖪𝖱ν\mathsf{KR}^{\nu}. By Proposition 7.2 we have for xx in 𝖪𝖱ν\mathsf{KR}^{\nu}:

𝔼ν[f](x)=i=1nqiIUi(v)𝑑ρx(v)=i=1nqiρx(Ui),\displaystyle\mathbb{E}_{\nu}[f\mid\mathscr{F}](x)=\sum_{i=1}^{n}q_{i}\int I_{U_{i}}(v)\;d\rho_{x}(v)=\sum_{i=1}^{n}q_{i}\cdot\rho_{x}(U_{i}),\hskip 42.67912pt
𝔼ν[fp](x)=i=1nqipIUi(v)𝑑ρx(v)=i=1nqipρx(Ui)\displaystyle\mathbb{E}_{\nu}[f^{p}\mid\mathscr{F}](x)=\sum_{i=1}^{n}q_{i}^{p}\int I_{U_{i}}(v)\;d\rho_{x}(v)=\sum_{i=1}^{n}q_{i}^{p}\cdot\rho_{x}(U_{i})

By Definition 1.3(3), the functions xρx(Ui)x\mapsto\rho_{x}(U_{i}) are lsc. Hence, by Proposition 2.16, choose functions υi,s\upsilon_{i,s} from the countable dense set of Lp(ν)L_{p}(\nu) that converge upward to ρ(Ui)\rho_{\cdot}(U_{i}), in that 0υi,sυi,s+10\leq\upsilon_{i,s}\leq\upsilon_{i,s+1} everywhere and ρ(Ui)=supsυi,s\rho_{\cdot}(U_{i})=\sup_{s}\upsilon_{i,s} everywhere. Let gs=i=1nqiυi,sg_{s}=\sum_{i=1}^{n}q_{i}\cdot\upsilon_{i,s} and hs=i=1nqipυi,sh_{s}=\sum_{i=1}^{n}q_{i}^{p}\cdot\upsilon_{i,s}, which likewise converge upward to 𝔼ν[f]\mathbb{E}_{\nu}[f\mid\mathscr{F}] and 𝔼ν[fp]\mathbb{E}_{\nu}[f^{p}\mid\mathscr{F}] respectively on 𝖪𝖱ν\mathsf{KR}^{\nu}.

Suppose xx is in 𝖪𝖱ν\mathsf{KR}^{\nu}. Since we are working with a Kurtz disintegration, we then have that 𝖪𝖱ν\mathsf{KR}^{\nu} is a ρx\rho_{x}-measure one set, and so ρx(Ui)=ρx(Ai)\rho_{x}(U_{i})=\rho_{x}(A_{i}). Since the AiA_{i} are pairwise disjoint, we then have i=1nρx(Ui)=i=1nρx(Ai)1\sum_{i=1}^{n}\rho_{x}(U_{i})=\sum_{i=1}^{n}\rho_{x}(A_{i})\leq 1, and so i=1nρx(Ui)υi,s(x)1\sum_{i=1}^{n}\rho_{x}(U_{i})-\upsilon_{i,s}(x)\leq 1. Then for xx in 𝖪𝖱ν\mathsf{KR}^{\nu}, by the convexity of the pp-th power function applied with coefficients ρx(Ui)υi,s(x)\rho_{x}(U_{i})-\upsilon_{i,s}(x) and points qiq_{i}, we have:

(𝔼ν[f](x)gs(x))p=(i=1nqi(ρx(Ui)υi,s(x)))p\displaystyle(\mathbb{E}_{\nu}[f\mid\mathscr{F}](x)-g_{s}(x))^{p}=\bigg{(}\sum_{i=1}^{n}q_{i}\cdot(\rho_{x}(U_{i})-\upsilon_{i,s}(x))\bigg{)}^{p}
\displaystyle\leq i=1nqip(ρx(Ui)υi,s(x))=𝔼ν[fp](x)hs(x)\displaystyle\sum_{i=1}^{n}q_{i}^{p}\cdot(\rho_{x}(U_{i})-\upsilon_{i,s}(x))=\mathbb{E}_{\nu}[f^{p}\mid\mathscr{F}](x)-h_{s}(x)

Since this estimate holds on the ν\nu-measure one set 𝖪𝖱ν\mathsf{KR}^{\nu}, by taking expectations and then pp-th roots, we have 𝔼ν[f]gsp(𝔼νfp𝔼νhs)1p\|\mathbb{E}_{\nu}[f\mid\mathscr{F}]-g_{s}\|_{p}\leq\big{(}\mathbb{E}_{\nu}f^{p}-\mathbb{E}_{\nu}h_{s}\big{)}^{\frac{1}{p}}. Since the right-hand side is a computable value which goes to zero as ss goes to infinity (by MCT), we can compute a subsequence of the gsg_{s} which is a witness to the Lp(ν)L_{p}(\nu) computability of 𝔼ν[f]\mathbb{E}_{\nu}[f\mid\mathscr{F}]. Since f,φf,\varphi are equal on 𝖪𝖱ν\mathsf{KR}^{\nu}, we have 𝔼ν[φ]\mathbb{E}_{\nu}[\varphi\mid\mathscr{F}] is also a computable point of Lp(ν)L_{p}(\nu). ∎

The previous proposition has the following elementary consequence:

Proposition 7.6.

Suppose ρ:X+(X)\rho:X\rightarrow\mathcal{M}^{+}(X) is a Kurtz disintegration of \mathscr{F}. Suppose p1p\geq 1 is computable.

  1. (1)

    If ff is an Lp(ν)L_{p}(\nu) Martin-Löf test, then 𝔼ν[f]\mathbb{E}_{\nu}[f\mid\mathscr{F}] is an Lp(ν)L_{p}(\nu) Martin-Löf test and 𝔼ν[fp]\mathbb{E}_{\nu}[f^{p}\mid\mathscr{F}] is an L1(ν)L_{1}(\nu) Martin-Löf test.

  2. (2)

    If ff is an Lp(ν)L_{p}(\nu) Schnorr test, then 𝔼ν[f]\mathbb{E}_{\nu}[f\mid\mathscr{F}] is an Lp(ν)L_{p}(\nu) Schnorr test and 𝔼ν[fp]\mathbb{E}_{\nu}[f^{p}\mid\mathscr{F}] is an L1(ν)L_{1}(\nu) Schnorr test.

Proof.

For (1), suppose that ff is an Lp(ν)L_{p}(\nu) Martin-Löf test. Then f,fpf,f^{p} are non-negative lsc, and so by Proposition 7.1 one has that 𝔼ν[f]\mathbb{E}_{\nu}[f\mid\mathscr{F}] and 𝔼ν[fp]\mathbb{E}_{\nu}[f^{p}\mid\mathscr{F}] are non-negative lsc. By conditional Jensen, 𝔼ν[f]pfp<\|\mathbb{E}_{\nu}[f\mid\mathscr{F}]\|_{p}\leq\|f\|_{p}<\infty, and likewise 𝔼ν[fp]1fp1=fpp<\|\mathbb{E}_{\nu}[f^{p}\mid\mathscr{F}]\|_{1}\leq\|f^{p}\|_{1}=\|f\|^{p}_{p}<\infty.

For (2), suppose that ff is an Lp(ν)L_{p}(\nu) Schnorr test. Since fp1=fpp\|f^{p}\|_{1}=\|f\|^{p}_{p}, one has that fpf^{p} is an L1(ν)L_{1}(\nu) Schnorr test. By Proposition 2.16, ff is a computable point of Lp(ν)L_{p}(\nu), and fpf^{p} is a computable point of L1(ν)L_{1}(\nu). By the previous proposition 𝔼ν[f]\mathbb{E}_{\nu}[f\mid\mathscr{F}] is a computable point of Lp(ν)L_{p}(\nu), and 𝔼ν[fp]\mathbb{E}_{\nu}[f^{p}\mid\mathscr{F}] is a computable point of L1(ν)L_{1}(\nu). ∎

This next proposition seems specific to Schnorr tests and 𝖲𝖱ν\mathsf{SR}^{\nu}:

Proposition 7.7.

(Tower) Suppose that 𝒢\mathscr{H}\subseteq\mathscr{G} are two effective σ\sigma-algebras, each of which has a Kurtz disintegration. Then for every L1(ν)L_{1}(\nu) Schnorr test ff, one has that 𝔼ν[𝔼ν[f𝒢]]=𝔼ν[f]\mathbb{E}_{\nu}[\mathbb{E}_{\nu}[f\mid\mathscr{G}]\mid\mathscr{H}]=\mathbb{E}_{\nu}[f\mid\mathscr{H}] on 𝖲𝖱ν\mathsf{SR}^{\nu}.

Proof.

By the previous proposition, one has that the two functions g:=𝔼ν[𝔼ν[f𝒢]]g:=\mathbb{E}_{\nu}[\mathbb{E}_{\nu}[f\mid\mathscr{G}]\mid\mathscr{H}] and h:=𝔼ν[f]h:=\mathbb{E}_{\nu}[f\mid\mathscr{H}] are L1(ν)L_{1}(\nu) Schnorr tests. Suppose that they are not equal on xx in 𝖲𝖱ν\mathsf{SR}^{\nu}. Then, without loss of generality, there are rationals a,b,ca,b,c with g(x)<a<b<c<h(x)g(x)<a<b<c<h(x). By Lemma 2.22, there is a computable real ϵ\epsilon in the interval (b,c)(b,c) with (h#ν)(ϵ,](h\#\nu)(\epsilon,\infty] computable. Since hh is lsc, the set U:=h1(ϵ,]U:=h^{-1}(\epsilon,\infty] is c.e. open, and it has computable measure. By the same lemma, there is a computable real δ\delta in the interval (a,b)(a,b) with (g#ν)(δ,](g\#\nu)(\delta,\infty] computable, so that (g#ν)[0,δ](g\#\nu)[0,\delta] is likewise computable. Since gg is lsc, the set C:=g1[0,δ]C:=g^{-1}[0,\delta] is effectively closed. Since it has computable ν\nu-measure, by Proposition 2.12, there is a a decreasing sequence of c.e. opens VnCV_{n}\supseteq C with ν(Vn)\nu(V_{n}) uniformly computable and ν(VnC)<2n\nu(V_{n}\setminus C)<2^{-n}. We then claim that ν(UC)>0\nu(U\cap C)>0. For, suppose not. Then 0=ν(UC)=limiν(UVi)0=\nu(U\cap C)=\lim_{i}\nu(U\cap V_{i}). Since ν(UVi)\nu(U\cap V_{i}) is computable by Proposition 2.8(2), we can then compute a subsequence UVn(i)U\cap V_{n(i)} with ν(UVn(i))2i\nu(U\cap V_{n(i)})\leq 2^{-i}, so that iIUVn(i)\sum_{i}I_{U\cap V_{n(i)}} is an L1(ν)L_{1}(\nu) Schnorr test. But since xx in 𝖲𝖱ν\mathsf{SR}^{\nu} and xx in UCU\cap C by construction, we have a contradiction. Hence indeed ν(UC)>0\nu(U\cap C)>0. Since g,hg,h are by definition \mathscr{H}-measurable, we have that UCU\cap C is also \mathscr{H}-measurable and hence 𝒢\mathscr{G}-measurable. Then one has the following identities by the definition of conditional expectation (these identities being the classical proof of the tower property):

UC𝔼ν[f]𝑑ν=UCf𝑑ν=UC𝔼ν[f𝒢]𝑑ν=UC𝔼ν[𝔼ν[f𝒢]]𝑑ν\int_{U\cap C}\mathbb{E}_{\nu}[f\mid\mathscr{H}]\;d\nu=\int_{U\cap C}f\;d\nu=\int_{U\cap C}\mathbb{E}_{\nu}[f\mid\mathscr{G}]\;d\nu=\int_{U\cap C}\mathbb{E}_{\nu}[\mathbb{E}_{\nu}[f\mid\mathscr{G}]\mid\mathscr{H}]\;d\nu

But these identities give us the below identity, where the remaining inequalities follow from the definitions of g,h,U,Cg,h,U,C:

ϵν(UC)UCh=UCgδν(UC)\epsilon\cdot\nu(U\cap C)\leq\int_{U\cap C}h=\int_{U\cap C}g\leq\delta\cdot\nu(U\cap C)

But since ν(UC)>0\nu(U\cap C)>0, we then have that ϵδ\epsilon\leq\delta, contrary to construction.

Proposition 7.8.

(The rôle of independence) Suppose ρ:X+(X)\rho:X\rightarrow\mathcal{M}^{+}(X) is a Kurtz disintegration of \mathscr{F}.

If ff in 𝕃1+(ν)\mathbb{L}_{1}^{+}(\nu) is independent of \mathscr{F}, then 𝔼ν[f]𝔼ν[f]\mathbb{E}_{\nu}[f\mid\mathscr{F}]\leq\mathbb{E}_{\nu}[f] everywhere.

If ff is an L1(ν)L_{1}(\nu) Martin-Löf test ff independent of \mathscr{F}, then 𝔼ν[f]=𝔼ν[f]\mathbb{E}_{\nu}[f\mid\mathscr{F}]=\mathbb{E}_{\nu}[f] on 𝖪𝖱ν\mathsf{KR}^{\nu}.

Proof.

Let us abbreviate g=𝔼ν[f]g=\mathbb{E}_{\nu}[f\mid\mathscr{F}].

First suppose that ff in 𝕃1+(ν)\mathbb{L}_{1}^{+}(\nu) is independent of \mathscr{F}. Let xx in XX be arbitrary. Suppose that g(x)>𝔼νfg(x)>\mathbb{E}_{\nu}f. Choose rational aa with g(x)>a>𝔼νfg(x)>a>\mathbb{E}_{\nu}f. Let B=g1(a,]B=g^{-1}(a,\infty], which is in \mathscr{F}. Then we have the following, where the first step is independence: 𝔼ν[IBf]=ν(B)𝔼ν[f]<aν(B)Bg𝑑ν=𝔼ν[IBf]\mathbb{E}_{\nu}[I_{B}\cdot f]=\nu(B)\cdot\mathbb{E}_{\nu}[f]<a\cdot\nu(B)\leq\int_{B}g\;d\nu=\mathbb{E}_{\nu}[I_{B}\cdot f].

Second suppose that ff is an L1(ν)L_{1}(\nu) Martin-Löf test, ff independent of \mathscr{F}. Then gg is non-negative lsc by Proposition 7.1. Suppose xx is in 𝖪𝖱ν\mathsf{KR}^{\nu}. Suppose g(x)<𝔼νfg(x)<\mathbb{E}_{\nu}f. Choose rational a>0a>0 with g(x)<a<𝔼νfg(x)<a<\mathbb{E}_{\nu}f. Let C=g1[0,a]C=g^{-1}[0,a], which is effectively closed and in \mathscr{F}. Since it contains the 𝖪𝖱ν\mathsf{KR}^{\nu} point xx, we have that ν(C)>0\nu(C)>0. Then we have the following, where the first step is from independence: ν(C)𝔼νf=𝔼ν[ICf]=𝔼ν[ICg]aν(C)\nu(C)\cdot\mathbb{E}_{\nu}f=\mathbb{E}_{\nu}[I_{C}\cdot f]=\mathbb{E}_{\nu}[I_{C}\cdot g]\leq a\cdot\nu(C). Since ν(C)>0\nu(C)>0 we have 𝔼νfa\mathbb{E}_{\nu}f\leq a, a contradiction. ∎

Now we turn towards effective properties of the maximal function (cf. §5 for the classical properties). Recall that almost-full was defined in Definition 1.1(12).

Proposition 7.9.

Let n\mathscr{F}_{n} be an almost-full effective filtration equipped with Kurtz disintegrations. Let f(x)=supn𝔼ν[fn](x)f^{\flat}(x)=\sup_{n}\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x) be the associated version of the maximal function.

For p>1p>1, the maximal function is computable continuous from Lp(ν)L_{p}(\nu) to Lp(ν)L_{p}(\nu).

For p=1p=1, the maximal function is computable continuous from Lp(ν)L_{p}(\nu) to L0(ν)L_{0}(\nu).

For p1p\geq 1, the maximal function sends the countable dense set of Lp+(ν)L_{p}^{+}(\nu) (resp. Lp(ν)L_{p}(\nu)) uniformly to computable points of Lp+(ν)L_{p}^{+}(\nu) (resp. Lp(ν)L_{p}(\nu)).

Proof.

First suppose p>1p>1. Let f=i=1kqiIAif=\sum_{i=1}^{k}q_{i}\cdot I_{A_{i}}, where qiq_{i} is rational and AkA_{k} comes from the algebra generated by a ν\nu-computable basis. Without loss of generality, qi0q_{i}\neq 0 and the AkA_{k} are pairwise disjoint. By almost-fullness of the filtration and Proposition 2.13, for each 1ik1\leq i\leq k, let Ai=sUi,sA_{i}=\bigcup_{s}U_{i,s} on 𝖪𝖱ν\mathsf{KR}^{\nu}, where Ui,sU_{i,s} is a c.e. open equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to an element from the sequence which generates g(i,s)\mathscr{F}_{g(i,s)}, where gg is a computable function. By replacing Ui,sU_{i,s} by tsUi,s\bigcup_{t\leq s}U_{i,s}, we may assume that Ui,sU_{i,s} and g(i,s)g(i,s) are non-decreasing in ss, for each i0i\geq 0. Let fs=i=1kqiIUi,sf_{s}=\sum_{i=1}^{k}q_{i}\cdot I_{U_{i,s}}. The AiA_{i} and Ui,sU_{i,s} come from the algebra generated by a ν\nu-computable basis, and so for each 1ik1\leq i\leq k, we can compute a subsequence Ui,s(n)U_{i,s(n)} such that ν(AiUi,s(n))<(1max1jk|qj|1kp1p2n)p\nu(A_{i}\setminus U_{i,s(n)})<\big{(}\frac{1}{\max_{1\leq j\leq k}\left|q_{j}\right|}\cdot\frac{1}{k}\cdot\frac{p-1}{p}\cdot 2^{-n}\big{)}^{p}. By Doob’s Maximal Inequality (Lemma 5.1), we have

ffs(n)p(ffs(n))ppp1ffs(n)pi=1k|qi|pp1ν(AiUi,s(n))1p\|f^{\flat}-f_{s(n)}^{\flat}\|_{p}\leq\|(f-f_{s(n)})^{\flat}\|_{p}\leq\frac{p}{p-1}\cdot\|f-f_{s(n)}\|_{p}\leq\sum_{i=1}^{k}\left|q_{i}\right|\cdot\frac{p}{p-1}\cdot\nu(A_{i}\setminus U_{i,s(n)})^{\frac{1}{p}}

which is <2n<2^{-n}. Hence it remains to show that fs(n)f_{s(n)}^{\flat} is uniformly a computable point of Lp(ν)L_{p}(\nu). Since for all tmax1ikg(i,s(n))t\geq\max_{1\leq i\leq k}g(i,s(n)), the c.e. open Ui,s(n)U_{i,s(n)} is equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to an element of t\mathscr{F}_{t}, one has that fs(n)=𝔼ν[fs(n)t]f_{s(n)}=\mathbb{E}_{\nu}[f_{s(n)}\mid\mathscr{F}_{t}] on 𝖪𝖱ν\mathsf{KR}^{\nu} by Proposition 7.3. Hence, fs(n)=suptmax1ikg(i,s(n))𝔼[fs(n)t]f_{s(n)}^{\flat}=\sup_{t\leq\max_{1\leq i\leq k}g(i,s(n))}\mathbb{E}[f_{s(n)}\mid\mathscr{F}_{t}] on 𝖪𝖱ν\mathsf{KR}^{\nu}. And the latter is a computable element of Lp(ν)L_{p}(\nu) since it is a finite max of conditional expectations which are computable elements of Lp(ν)L_{p}(\nu) by Proposition 7.5.

For p=1p=1, one just appeals to the fact that L1(ν)L_{1}(\nu) and Lq(ν)L_{q}(\nu) for computable q>1q>1 share a common dense set and that Lq(ν)L_{q}(\nu) computably embeds in L1(ν)L_{1}(\nu).

Then computability continuity follow from Proposition 2.5 and Proposition 5.2. ∎

Proposition 7.10.

Let n\mathscr{F}_{n} be an almost-full effective filtration equipped with Kurtz disintegrations.

For every Lp(ν)L_{p}(\nu) Schnorr test ff, we can compute an index for a sequence of Lp(ν)L_{p}(\nu) Schnorr tests gsg_{s} such that gsgs+1g_{s}\leq g_{s+1} on 𝖪𝖱ν\mathsf{KR}^{\nu} and f=supsgsf=\sup_{s}g_{s} on 𝖪𝖱ν\mathsf{KR}^{\nu} and gsfg_{s}\rightarrow f fast in Lp(ν)L_{p}(\nu) and gs=limn𝔼ν[gsn]g_{s}=\lim_{n}\mathbb{E}_{\nu}[g_{s}\mid\mathscr{F}_{n}] on 𝖪𝖱ν\mathsf{KR}^{\nu}. Indeed, there is computable function n()n(\cdot) such that

gs=𝔼ν[gsm] on 𝖪𝖱ν for all mn(s)g_{s}=\mathbb{E}_{\nu}[g_{s}\mid\mathscr{F}_{m}]\mbox{ on }\mathsf{KR}^{\nu}\mbox{ for all }m\geq n(s) (7.1)

Further, we can compute an index for a non-negative usc function hsh_{s} such that gs,hsg_{s},h_{s} are equal on 𝖪𝖱ν\mathsf{KR}^{\nu}.

Finally, we can compute from ss an index for a non-negative rational which bounds |gs|\left|g_{s}\right| on 𝖪𝖱ν\mathsf{KR}^{\nu}.

Proof.

Enumerate [0,)\mathbb{Q}\cap[0,\infty) as q0,q1,q_{0},q_{1},\ldots. For each n0n\geq 0, one has that f1(qn,]f^{-1}(q_{n},\infty] is uniformly c.e. open. Since the filtration is almost-full, by using Proposition 2.13 there is a computable sequence Un,jU_{n,j} of c.e. opens such that f1(qn,]=jUn,jf^{-1}(q_{n},\infty]=\bigcup_{j}U_{n,j} on 𝖪𝖱ν\mathsf{KR}^{\nu}, where the Un,jU_{n,j} are equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to events from the sequence which generates m\mathscr{F}_{m}. Then define

fs(x)=max{0,qn:n,js,xUn,j}f_{s}(x)=\max\{0,q_{n}:n,j\leq s,x\in U_{n,j}\} (7.2)

As in the proof of Proposition 2.16, this is a rational-valued step function, whose events are equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to events from the sequence which generates n(s)\mathscr{F}_{n(s)}, where n()n(\cdot) is a computable function. Then by Proposition 7.3 one has that

fs=𝔼ν[fsm] on 𝖪𝖱ν for all mn(s)f_{s}=\mathbb{E}_{\nu}[f_{s}\mid\mathscr{F}_{m}]\mbox{ on }\mathsf{KR}^{\nu}\mbox{ for all }m\geq n(s) (7.3)

Further from (7.2) one sees that fsfs+1f_{s}\leq f_{s+1} everywhere since the sum over which we taking the maximum grows in ss. Further, one has fsff_{s}\leq f on 𝖪𝖱ν\mathsf{KR}^{\nu} since if we had fs(x)>f(x)f_{s}(x)>f(x) for xx in 𝖪𝖱ν\mathsf{KR}^{\nu}, then fs(x)=qnf_{s}(x)=q_{n} for some n,jsn,j\leq s with xx in Un,jU_{n,j}. But since Un,jf1(qn,]U_{n,j}\subseteq f^{-1}(q_{n},\infty] on 𝖪𝖱ν\mathsf{KR}^{\nu}, we then have f(x)>qnf(x)>q_{n}. Finally, one has supsfs=f\sup_{s}f_{s}=f on 𝖪𝖱ν\mathsf{KR}^{\nu}, since if not we would have supsfs(x)<qn<f(x)\sup_{s}f_{s}(x)<q_{n}<f(x) for some xx in 𝖪𝖱ν\mathsf{KR}^{\nu} and some nn and hence xx would be in f1(qn,]=jUn,jf^{-1}(q_{n},\infty]=\bigcup_{j}U_{n,j} and so xx would be in Un,jU_{n,j} for some jj and hence for sjs\geq j one would have that fs(x)qnf_{s}(x)\geq q_{n} by definition in (7.2).

We can pass to a subsequence of fsf_{s} which goes to ff fast in Lp(ν)L_{p}(\nu) as in the proof of Proposition 2.16.

Since fsf_{s} is formed from events which are equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to events coming from a ν\nu-computable basis, by Proposition 2.23 there is non-negative lsc gsg_{s} and non-negative usc hsh_{s} such that fs,gs,hsf_{s},g_{s},h_{s} are equal on 𝖪𝖱ν\mathsf{KR}^{\nu}. Since fs,gsf_{s},g_{s} are equal on 𝖪𝖱ν\mathsf{KR}^{\nu}, we can use Proposition 7.2(1) to infer from (7.3) to (7.1). ∎

8. Proof of Theorems 1.5-1.6

First we prove Theorem 1.5:

Proof.

The results of the previous section show that the conditions of Theorem 6.2 are satisfied:

  • Condition (I) is Proposition 7.1.

  • Condition (II) is Proposition 7.2.

  • Condition (III) is Propositions 7.5, 7.9.

  • Condition (IV) is Proposition 7.10.

Finally, we argue from Theorem 1.5(4) to Theorem 1.5(1). Suppose Theorem 1.5(4) is satisfied. We want to show that xx is in 𝖲𝖱ν\mathsf{SR}^{\nu}. Let ff be an Lp(ν)L_{p}(\nu) Schnorr test. We want to show that f(x)<f(x)<\infty. Suppose not. Since by hypothesis limn𝔼ν[fn](x)\lim_{n}\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x) exists, there are rationals b,ab,a and there is n00n_{0}\geq 0 such that f(x)>b>a>𝔼ν[fn](x)f(x)>b>a>\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x) for all nn0n\geq n_{0}. Then xx is in the c.e. open f1(b,]f^{-1}(b,\infty]. Since the filtration is almost-full and xx is in 𝖪𝖱ν\mathsf{KR}^{\nu}, there is n1n0n_{1}\geq n_{0} and an event AA from n1\mathscr{F}_{n_{1}} such that xx is in AA and A𝖪𝖱νf1(b,]A\cap\mathsf{KR}^{\nu}\subseteq f^{-1}(b,\infty]. Then [x]n1A[x]_{\mathscr{F}_{n_{1}}}\subseteq A, and hence [x]n1𝖪𝖱νA𝖪𝖱νf1(b,][x]_{\mathscr{F}_{n_{1}}}\cap\mathsf{KR}^{\nu}\subseteq A\cap\mathsf{KR}^{\nu}\subseteq f^{-1}(b,\infty]. Hence f1(b,]f^{-1}(b,\infty] is a ρx(n1)\rho_{x}^{(n_{1})}-measure one event, and hence 𝔼ν[fn1](x)=f𝑑ρx(n1)b>a\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n_{1}}](x)=\int f\;d\rho_{x}^{(n_{1})}\geq b>a, a contradiction.

Now we turn to Theorem 1.6:

Proof.

For Theorem 1.6(1), one appeals to Theorem 6.2(ii), along with Proposition 7.10.

For Theorem 1.6(2) suppose that xx in 𝖲𝖱ν\mathsf{SR}^{\nu} is of computably dominated degree. By the previous paragraph, we have that xx weakly computes a modulus m:>0m:\mathbb{Q}^{>0}\rightarrow\mathbb{N} for the convergence 𝔼[fn](x)f(x)\mathbb{E}[f\mid\mathscr{F}_{n}](x)\rightarrow f(x). Likewise m:m^{\prime}:\mathbb{N}\rightarrow\mathbb{N} defined by m(i)=m(2i)m^{\prime}(i)=m(2^{-i}) is computable from mm. Since xx is of computably dominated degree, we have that at least one of these mm^{\prime} is dominated by some computable function, call it mm^{{\dagger}}, so that that past some point, call it i0i_{0}, we have mmm^{{\dagger}}\geq m^{\prime}. Let ϵ>0\epsilon>0 be rational. Compute ji0j\geq i_{0} such that 2j<ϵ2^{-j}<\epsilon. Let nm(j)n\geq m^{{\dagger}}(j). Then nm(j)m(j)=m(2j)n\geq m^{{\dagger}}(j)\geq m^{\prime}(j)=m(2^{-j}). Then |𝔼[fn](x)f(x)|<2j<ϵ\left|\mathbb{E}[f\mid\mathscr{F}_{n}](x)-f(x)\right|<2^{-j}<\epsilon. ∎

9. Proof of Theorems 1.8-1.9

We begin with an elementary proposition.

Proposition 9.1.

Suppose μn,μ\mu_{n},\mu in 𝒫(X)\mathcal{P}(X) such that for every c.e. open UU one has limnμn(U)=μ(U)\lim_{n}\mu_{n}(U)=\mu(U).

  1. (1)

    For every event AA in the algebra generated by the c.e. opens, one has limnμn(A)=μ(A)\lim_{n}\mu_{n}(A)=\mu(A).

  2. (2)

    For every simple function ff generated from events in this algebra, one has limnfμn=fμ\lim_{n}\int f\;\mu_{n}=\int f\;\mu.

  3. (3)

    For every lsc f:X[0,]f:X\rightarrow[0,\infty] which is in each of 𝕃1+(μn),𝕃1+(μ)\mathbb{L}^{+}_{1}(\mu_{n}),\mathbb{L}^{+}_{1}(\mu), one has f𝑑μlim infnf𝑑μn\int f\;d\mu\leq\liminf_{n}\int f\;d\mu_{n}.

This proposition too illustrates that convergence to the truth is a strengthening of weak convergence, since in (3) there is no boundedness constraint on the lsc function.

Proof.

For (1), every event in this algebra can be written as a finite disjoint union of sets of the form U1UmV1cVncU_{1}\cap\cdots\cap U_{m}\cap V_{1}^{c}\cap\cdots\cap V_{n}^{c}, where Ui,ViU_{i},V_{i} are c.e. opens. Let U=U1UmU=U_{1}\cap\cdots\cap U_{m} and V=V1VnV=V_{1}\cup\cdots\cup V_{n}, so that the set has the form UVU\setminus V. Since μn,μ\mu_{n},\mu are in 𝒫(X)\mathcal{P}(X) and since UVU\cap V is c.e. open as well, we have that

limnμn(UV)=limnμn(U)limnμn(UV)=μ(U)μ(UV)=μ(UV)\lim_{n}\mu_{n}(U\setminus V)=\lim_{n}\mu_{n}(U)-\lim_{n}\mu_{n}(U\cap V)=\mu(U)-\mu(U\cap V)=\mu(U\setminus V)

For (2), one just applies the previous item and the properties of the integral.

For (3), suppose not. Choose rational qq such that f𝑑μ>q>lim infnf𝑑μn\int f\;d\mu>q>\liminf_{n}\int f\;d\mu_{n}. Let fsf_{s} be the approximation to ff as in Proposition 2.16. Then by the MCT applied in 𝕃1+(μ)\mathbb{L}^{+}_{1}(\mu), there is s0s\geq 0 such that fs𝑑μ>q\int f_{s}\;d\mu>q. By (2), there is n00n_{0}\geq 0 such that for all nn0n\geq n_{0} one has that fs𝑑μn>q\int f_{s}\;d\mu_{n}>q. But this contradicts that q>lim infnf𝑑μnq>\liminf_{n}\int f\;d\mu_{n}. ∎

Now we prove Theorem 1.8:

Proof.

Suppose Theorem 1.8(1); we show Theorem 1.8(2). Since xx is in 𝖬𝖫𝖱ν\mathsf{MLR}^{\nu}, one has that ρx(n)\rho_{x}^{(n)} is in 𝒫(X)\mathcal{P}(X). Suppose that ff is an Lp(ν)L_{p}(\nu) Martin-Löf test. Since xx is in 𝖬𝖫𝖱ν\mathsf{MLR}^{\nu}, one has f(x)<f(x)<\infty. By Proposition 7.6(1), the functions 𝔼ν[fn]\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}] are Lp(ν)L_{p}(\nu) Martin-Löf tests as well, and hence 𝔼ν[fn](x)<\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x)<\infty, which is just to say that f𝑑ρx(n)<\int f\;d\rho^{(n)}_{x}<\infty. By Proposition 9.1(3) applied to ρx(n),δx\rho^{(n)}_{x},\delta_{x} one has that f𝑑δxlim infnf𝑑ρx(n)\int f\;d\delta_{x}\leq\liminf_{n}\int f\;d\rho^{(n)}_{x}, which is just to say that f(x)lim infn𝔼ν[fn](x)f(x)\leq\liminf_{n}\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x). Hence, it remains to show that lim supn𝔼ν[fn](x)f(x)\limsup_{n}\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x)\leq f(x). Suppose not. For reductio, suppose there are rational a,ba,b with f(x)<a<b<lim supn𝔼ν[fn](x)f(x)<a<b<\limsup_{n}\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x).

Let U=f1(a,]U=f^{-1}(a,\infty] and C=f1[0,a]C=f^{-1}[0,a], so that UU is c.e. open and CC is effectively closed. Since xx is in 𝖣𝖱ρν\mathsf{DR}_{\rho}^{\nu} and xx is not in UU, we have ρx(n)(U)0\rho^{(n)}_{x}(U)\rightarrow 0.

By definition of CC, we have for all n0n\geq 0 that Cf𝑑ρx(n)a\int_{C}f\;d\rho^{(n)}_{x}\leq a. For any n0n\geq 0 such that 𝔼ν[fn](x)>b\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x)>b, we then have Cf𝑑ρx(n)a<b<f𝑑ρx(n)\int_{C}f\;d\rho^{(n)}_{x}\leq a<b<\int fd\rho^{(n)}_{x}, so that Uf𝑑ρx(n)>ba\int_{U}f\;d\rho^{(n)}_{x}>b-a. Hence, our reductio hypothesis gives that there are infinitely many n0n\geq 0 with Uf𝑑ρx(n)>ba\int_{U}f\;d\rho^{(n)}_{x}>b-a.

By Proposition 7.6(1), we have that 𝔼ν[fpn]\mathbb{E}_{\nu}[f^{p}\mid\mathscr{F}_{n}] are L1(ν)L_{1}(\nu) Martin-Löf tests. Hence its maximal function supn𝔼ν[fpn]\sup_{n}\mathbb{E}_{\nu}[f^{p}\mid\mathscr{F}_{n}] is also non-negative lsc. Let Uk={yX:supn𝔼ν[fpn](y)>2k}U_{k}=\{y\in X:\sup_{n}\mathbb{E}_{\nu}[f^{p}\mid\mathscr{F}_{n}](y)>2^{k}\}, which is c.e. open. By Doob’s Submartingale Inequality, one has that ν(Uk)\nu(U_{k}) is \leq the following:

limmν({yX:supnm𝔼ν[fpn](y)>2k})limm2k𝔼ν[fpm]𝑑ν=2kfpp\lim_{m}\nu(\{y\in X:\sup_{n\leq m}\mathbb{E}_{\nu}[f^{p}\mid\mathscr{F}_{n}](y)>2^{k}\})\leq\lim_{m}2^{-k}\int\mathbb{E}_{\nu}[f^{p}\mid\mathscr{F}_{m}]\;d\nu=2^{-k}\|f\|_{p}^{p}

Hence g=kIUkg=\sum_{k}I_{U_{k}} is an L1(ν)L_{1}(\nu) Martin-Löf test, and hence there is constant K>0K>0 such that supn𝔼ν[fpn](x)<K\sup_{n}\mathbb{E}_{\nu}[f^{p}\mid\mathscr{F}_{n}](x)<K.

Then for all n0n\geq 0 we have fp𝑑ρx(n)<K\int f^{p}\;d\rho^{(n)}_{x}<K, so that ff is in Lp(ρx(n))L_{p}(\rho^{(n)}_{x}) with fLp(ρx(n))<K1p\|f\|_{L_{p}(\rho^{(n)}_{x})}<K^{\frac{1}{p}}. Let qq be the conjugate exponent to pp. Then, for each n0n\geq 0, we have by Hölder with respect to ρx(n)\rho^{(n)}_{x} that:

Uf𝑑ρx(n)=fIUL1(ρx(n))fLp(ρx(n))IULq(ρx(n))K1p(ρx(n)(U))1q\int_{U}f\;d\rho^{(n)}_{x}=\|f\cdot I_{U}\|_{L_{1}(\rho^{(n)}_{x})}\leq\|f\|_{L_{p}(\rho^{(n)}_{x})}\cdot\|I_{U}\|_{L_{q}(\rho^{(n)}_{x})}\leq K^{\frac{1}{p}}\cdot(\rho^{(n)}_{x}(U))^{\frac{1}{q}} (9.1)

Since ρx(n)(U)0\rho^{(n)}_{x}(U)\rightarrow 0, we have that Uf𝑑ρx(n)0\int_{U}f\;d\rho^{(n)}_{x}\rightarrow 0, contradicting the previous conclusion from the reductio hypothesis.

The implication from (2) to (3) is trivial.

The argument from (3) to (1) is nearly identical to the proof in §8 from Theorem 1.5(4) to Theorem 1.5(1): one just replaces 𝖲𝖱ν\mathsf{SR}^{\nu} with 𝖬𝖫𝖱ν\mathsf{MLR}^{\nu} and replaces Lp(ν)L_{p}(\nu) Schnorr tests with Lp(ν)L_{p}(\nu) Martin-Löf tests. ∎

Now we prove Theorem 1.9:

Proof.

We work in Cantor space with the uniform measure ν\nu, the effective full filtration n\mathscr{F}_{n} of the algebra of events generated by the length nn strings, and with the effective disintegration ρω(n)=ν([ωn])\rho^{(n)}_{\omega}=\nu(\cdot\mid[\omega\upharpoonright n]). Then 𝔼[fn](ω)=1ν([ωn])[ωn]f𝑑ν\mathbb{E}[f\mid\mathscr{F}_{n}](\omega)=\frac{1}{\nu([\omega\upharpoonright n])}\int_{[\omega\upharpoonright n]}f\;d\nu, and likewise 𝔼[IAn](ω)=ν(A[ωn])\mathbb{E}[I_{A}\mid\mathscr{F}_{n}](\omega)=\nu(A\mid[\omega\upharpoonright n]). (This is just Example B.1 for Cantor space and uniform measure).

We show that for ω\omega in 𝖣𝖱ρν\mathsf{DR}_{\rho}^{\nu} there is c.e. open UU with 0<ν(U)<10<\nu(U)<1 such that ω\omega is not in UU and the convergence ν(U[ωn])0\nu(U\mid[\omega\upharpoonright n])\rightarrow 0 does not have a computable rate.

(Since the example involves an indicator function, we have that IUI_{U} is in Lp(ν)L_{p}(\nu) for all p1p\geq 1 computable).

Let k0k\geq 0. Let KK be the halting set {e:φe(e)}\{e:\varphi_{e}(e){\downarrow}\}. Enumerate it as e0,e1,e_{0},e_{1},\ldots, where the map nenn\mapsto e_{n} is injective.

Define c0=0c_{0}=0 and cn+1=max(φen(en),cn)+1c_{n+1}=\max(\varphi_{e_{n}}(e_{n}),c_{n})+1.

For k,n0k,n\geq 0, define clopen Uk,n={ω:i[cn+1,cn+1+en+k+1)ω(i)=0}U_{k,n}=\{\omega:\forall\;i\in[c_{n+1},c_{n+1}+e_{n}+k+1)\;\omega(i)=0\} and define the c.e. open Uk=nUk,nU_{k}=\bigcup_{n}U_{k,n}. Then ν(Uk,n)=2(en+k+1)\nu(U_{k,n})=2^{-(e_{n}+k+1)} and 0<ν(Uk)nν(Uk,n)<2k0<\nu(U_{k})\leq\sum_{n}\nu(U_{k,n})<2^{-k}. Since Uk,nU_{k,n} just makes decisions on bits cn+1\geq c_{n+1}, it is independent of all bits <cn+1<c_{n+1} (since ν\nu is uniform measure). We then have that ν(Uk,n[ωcn+1])=2(en+k+1)\nu(U_{k,n}\mid[\omega\upharpoonright c_{n+1}])=2^{-(e_{n}+k+1)} for any ω\omega, and hence ν(Uk[ωcn+1])2(en+k+1)\nu(U_{k}\mid[\omega\upharpoonright c_{n+1}])\geq 2^{-(e_{n}+k+1)} for any ω\omega.

Let ω\omega in 𝖣𝖱ρν\mathsf{DR}_{\rho}^{\nu}. Since kIUk\sum_{k}I_{U_{k}} is an L1(ν)L_{1}(\nu) Martin-Löf test, one has that there is kk such that ω\omega is not in UkU_{k}. Hence ν(Uk[ωn])0\nu(U_{k}\mid[\omega\upharpoonright n])\rightarrow 0. Suppose that ν(Uk[ωn])0\nu(U_{k}\mid[\omega\upharpoonright n])\rightarrow 0 with computable rate mm. Let ene_{n} be such that φen(i)=m(2(i+k+1))\varphi_{e_{n}}(i)=m(2^{-(i+k+1)}) for all i0i\geq 0. Then φen(en)=m(2(en+k+1))\varphi_{e_{n}}(e_{n})=m(2^{-(e_{n}+k+1)}). Since cn+1φen(en)=m(2(en+k+1))c_{n+1}\geq\varphi_{e_{n}}(e_{n})=m(2^{-(e_{n}+k+1)}), one has that ν(Uk[ωcn+1])<2(en+k+1)\nu(U_{k}\mid[\omega\upharpoonright c_{n+1}])<2^{-(e_{n}+k+1)}, a contradiction to the previous paragraph. ∎

Note that the c.e. sets UkU_{k} constructed above are dense, and their definition interleaves Example 2.15 with the halting set. This seems natural, since as noted in Proposition 2.14, if Uk¯\overline{U_{k}} were effectively closed with the same ν\nu-measure as UkU_{k}, then we could include it in a ν\nu-computable basis.

10. Proof of Theorem 1.11

First we note a fact mentioned in the introduction, namely that Maximal Doob Randomness is inbetween Martin-Löf and Schnorr randomness:

Proposition 10.1.

For all computable p1p\geq 1 one has 𝖬𝖫𝖱ν𝖬𝖣𝖱ν,p𝖲𝖱ν\mathsf{MLR}^{\nu}\subseteq\mathsf{MDR}^{\nu,p}\subseteq\mathsf{SR}^{\nu}.

Proof.

Since any Lp(ν)L_{p}(\nu) maximal Doob test is an Lp(ν)L_{p}(\nu) Martin-Löf test, we have 𝖬𝖫𝖱ν𝖬𝖣𝖱ν,p\mathsf{MLR}^{\nu}\subseteq\mathsf{MDR}^{\nu,p}. To show that 𝖬𝖣𝖱ν,p𝖲𝖱ν\mathsf{MDR}^{\nu,p}\subseteq\mathsf{SR}^{\nu}, it suffices to show that any Lp(ν)L_{p}(\nu) Schnorr test ff is an Lp(ν)L_{p}(\nu) maximal Doob test. By Proposition 2.16, let fsf_{s} be from the countable dense set of Lp(ν)L_{p}(\nu) so that 0fsfs+10\leq f_{s}\leq f_{s+1} on 𝖪𝖱ν\mathsf{KR}^{\nu} and f=supsfsf=\sup_{s}f_{s} on 𝖪𝖱ν\mathsf{KR}^{\nu} and fsff_{s}\rightarrow f fast in Lp(ν)L_{p}(\nu). Then we can compute a subsequence s(n)s(n) such that ffs(n)p<en\|f-f_{s(n)}\|_{p}<e^{-n}. Then for all k0k\geq 0 one has that nffs(n)p(n+1)knen(n+1)k<\sum_{n}\|f-f_{s(n)}\|_{p}\cdot(n+1)^{k}\leq\sum_{n}e^{-n}(n+1)^{k}<\infty. Then we are done by Lemma 3.2. ∎

The following closure condition on Lp(ν)L_{p}(\nu) maximal Doob tests is the difficult component of the proof of Theorem 1.11:

Proposition 10.2.

If p>1p>1 is computable and ff is an Lp(ν)L_{p}(\nu) maximal Doob test with witness fsf_{s}, then g=ssupn𝔼ν[ffsn]g=\sum_{s}\sup_{n}\mathbb{E}_{\nu}[f-f_{s}\mid\mathscr{F}_{n}] is equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to a Lp(ν)L_{p}(\nu) maximal Doob test with witness equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to gt=s<tsupn𝔼ν[ftfsn]g_{t}=\sum_{s<t}\sup_{n}\mathbb{E}_{\nu}[f_{t}-f_{s}\mid\mathscr{F}_{n}].

Proof.

The function gg is equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to a non-negative lsc since it is equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to a supremum of non-negative lsc functions (cf. Proposition 7.1, Proposition  7.2(1)). Since p>1p>1, by Proposition 7.9, we have that gtg_{t} is equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to an Lp(ν)L_{p}(\nu) Schnorr test. For k0k\geq 0, the quantity tggtp(t+1)k\sum_{t}\|g-g_{t}\|_{p}\cdot(t+1)^{k} is \leq:

tstsupn𝔼ν[ffsn]p(t+1)k\displaystyle\sum_{t}\|\sum_{s\geq t}\sup_{n}\mathbb{E}_{\nu}[f-f_{s}\mid\mathscr{F}_{n}]\|_{p}\cdot(t+1)^{k} (10.1)
+\displaystyle+ ts<tsupn𝔼ν[ffsn]supn𝔼ν[ftfsn]p(t+1)k\displaystyle\sum_{t}\|\sum_{s<t}\sup_{n}\mathbb{E}_{\nu}[f-f_{s}\mid\mathscr{F}_{n}]-\sup_{n}\mathbb{E}_{\nu}[f_{t}-f_{s}\mid\mathscr{F}_{n}]\|_{p}\cdot(t+1)^{k} (10.2)

To estimate (10.1), let us first define a computable sequence of non-negative left-c.e. reals:

cs,t={0if t>s,supn𝔼ν[ffsn]p(s+1)kif ts.c_{s,t}=\begin{cases}0&\text{if $t>s$},\\ \|\sup_{n}\mathbb{E}_{\nu}[f-f_{s}\mid\mathscr{F}_{n}]\|_{p}\cdot(s+1)^{k}&\text{if $t\leq s$}.\end{cases}

For fixed s0s\geq 0 we have tcs,t=(s+1)cs,s=supn𝔼ν[ffsn]p(s+1)k+1\sum_{t}c_{s,t}=(s+1)\cdot c_{s,s}=\|\sup_{n}\mathbb{E}_{\nu}[f-f_{s}\mid\mathscr{F}_{n}]\|_{p}\cdot(s+1)^{k+1}. To estimate (10.1), we have the following, where the last line follows from Doob’s Maximal Inequality (Lemma 5.1(1)):

tstsupn𝔼ν[ffsn]p(t+1)ktstsupn𝔼ν[ffsn]p(t+1)k\displaystyle\sum_{t}\|\sum_{s\geq t}\sup_{n}\mathbb{E}_{\nu}[f-f_{s}\mid\mathscr{F}_{n}]\|_{p}\cdot(t+1)^{k}\leq\sum_{t}\sum_{s\geq t}\|\sup_{n}\mathbb{E}_{\nu}[f-f_{s}\mid\mathscr{F}_{n}]\|_{p}\cdot(t+1)^{k}
\displaystyle\leq tstsupn𝔼ν[ffsn]p(s+1)k=tstcs,t\displaystyle\sum_{t}\sum_{s\geq t}\|\sup_{n}\mathbb{E}_{\nu}[f-f_{s}\mid\mathscr{F}_{n}]\|_{p}\cdot(s+1)^{k}=\sum_{t}\sum_{s\geq t}c_{s,t}
=\displaystyle= tscs,t=stcs,t=ssupn𝔼ν[ffsn]p(s+1)k+1\displaystyle\sum_{t}\sum_{s}c_{s,t}=\sum_{s}\sum_{t}c_{s,t}=\sum_{s}\|\sup_{n}\mathbb{E}_{\nu}[f-f_{s}\mid\mathscr{F}_{n}]\|_{p}\cdot(s+1)^{k+1}
\displaystyle\leq spp1ffsp(s+1)k+1<\displaystyle\sum_{s}\frac{p}{p-1}\cdot\|f-f_{s}\|_{p}\cdot(s+1)^{k+1}<\infty

For (10.2), we have the following, where we use Doob’s Maximal Inequality (Lemma 5.1) again at the end:

ts<tsupn𝔼ν[ffsn]supn𝔼ν[ftfsn]p(t+1)k\displaystyle\sum_{t}\|\sum_{s<t}\sup_{n}\mathbb{E}_{\nu}[f-f_{s}\mid\mathscr{F}_{n}]-\sup_{n}\mathbb{E}_{\nu}[f_{t}-f_{s}\mid\mathscr{F}_{n}]\|_{p}\cdot(t+1)^{k}
\displaystyle\leq ts<tsupn𝔼ν[fftn]p(t+1)k\displaystyle\sum_{t}\sum_{s<t}\|\sup_{n}\mathbb{E}_{\nu}[f-f_{t}\mid\mathscr{F}_{n}]\|_{p}\cdot(t+1)^{k}
=\displaystyle= ttsupn𝔼ν[fftn]p(t+1)k\displaystyle\sum_{t}t\cdot\|\sup_{n}\mathbb{E}_{\nu}[f-f_{t}\mid\mathscr{F}_{n}]\|_{p}\cdot(t+1)^{k}
\displaystyle\leq tsupn𝔼ν[fftn]p(t+1)k+1\displaystyle\sum_{t}\|\sup_{n}\mathbb{E}_{\nu}[f-f_{t}\mid\mathscr{F}_{n}]\|_{p}\cdot(t+1)^{k+1}
\displaystyle\leq tpp1fftp(t+1)k+1<\displaystyle\sum_{t}\frac{p}{p-1}\cdot\|f-f_{t}\|_{p}\cdot(t+1)^{k+1}<\infty

Here is the proof of Theorem 1.11:

Proof.

Suppose (1); we prove (2). One has that xx in 𝖪𝖱ν\mathsf{KR}^{\nu} and indeed xx in 𝖲𝖱ν\mathsf{SR}^{\nu} by Proposition 10.1. Suppose now that ff is an Lp(ν)L_{p}(\nu) maximal Doob test with witness fsf_{s}. By the previous proposition and (1), we have limssupn𝔼ν[ffsn](x)=0\lim_{s}\sup_{n}\mathbb{E}_{\nu}[f-f_{s}\mid\mathscr{F}_{n}](x)=0. Let ϵ>0\epsilon>0. Choose s00s_{0}\geq 0 such that for all ss0s\geq s_{0} we have supn𝔼ν[ffsn](x)<ϵ3\sup_{n}\mathbb{E}_{\nu}[f-f_{s}\mid\mathscr{F}_{n}](x)<\frac{\epsilon}{3}. Choose s1s0s_{1}\geq s_{0} such that for all ss1s\geq s_{1} we have f(x)fs(x)<13f(x)-f_{s}(x)<\frac{1}{3}. By Theorem 1.5 applied to fs1f_{s_{1}}, we have that fs1(x)=limn𝔼ν[fs1n](x)f_{s_{1}}(x)=\lim_{n}\mathbb{E}_{\nu}[f_{s_{1}}\mid\mathscr{F}_{n}](x). Choose n00n_{0}\geq 0 such that for all nn0n\geq n_{0} we have |fs1(x)𝔼ν[fs1n](x)|<ϵ3\left|f_{s_{1}}(x)-\mathbb{E}_{\nu}[f_{s_{1}}\mid\mathscr{F}_{n}](x)\right|<\frac{\epsilon}{3}. Then putting this all together, we have for all nn0n\geq n_{0} that |f(x)𝔼ν[fn](x)|\left|f(x)-\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x)\right| is \leq the following:

|f(x)fs1(x)|+|fs1(x)𝔼ν[fs1n](x)|+|𝔼ν[fs1n](x)𝔼ν[fn](x)|<ϵ\left|f(x)-f_{s_{1}}(x)\right|+\left|f_{s_{1}}(x)-\mathbb{E}_{\nu}[f_{s_{1}}\mid\mathscr{F}_{n}](x)\right|+\left|\mathbb{E}_{\nu}[f_{s_{1}}\mid\mathscr{F}_{n}](x)-\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x)\right|<\epsilon

The step from (2) to (3) is trivial.

The step from (3) to (1) is exactly as in the corresponding step of the proof of Theorem 1.5 (in §8), but with the class of Lp(ν)L_{p}(\nu) Schnorr tests replaced by the class of Lp(ν)L_{p}(\nu) maximal Doob tests. ∎

11. Back and forth between tests and computable points

In this section we indicate how to state a version of Theorem 1.5 in terms of Lp(ν)L_{p}(\nu)-computable points. This essentially follows by a translation method of Miyabe.

We begin with how to select a version for each computable point Lp(ν)L_{p}(\nu). Pathak, Rojas, and Simpson and Rute have shown how to do this via Proposition 4.1. We use the following slight variant of their selection method:

Definition 11.1.

Suppose that ff is a computable point of L1(ν)L_{1}(\nu) with witness fnf_{n}. Then we define a version ff_{\infty} in 𝕃1(ν)\mathbb{L}_{1}(\nu) by:

f(x)={limnfn(x)if limnfn(x) exists,0otherwise.f_{\infty}(x)=\begin{cases}\lim_{n}f_{n}(x)&\text{if $\lim_{n}f_{n}(x)$ exists},\\ 0&\text{otherwise}.\end{cases}

Hence, Proposition 4.1 tells us that the definition of ff_{\infty} goes through the first case break on all points of 𝖲𝖱ν\mathsf{SR}^{\nu}, and on these points it is independent of the choice of the witness fnf_{n}. However, on X𝖲𝖱νX\setminus\mathsf{SR}^{\nu} we have that it is dependent on the witness fnf_{n}. If one was working more extensively with ff_{\infty}, one would want to develop some notation which better mark its dependence on the version fnf_{n}. But this dependence has the following advantage: if ff is an Lp(ν)L_{p}(\nu) Schnorr test and fnf_{n} is a witness to its being Lp(ν)L_{p}(\nu)-computable such that fnff_{n}\rightarrow f everywhere (as in Proposition 2.16), then f=ff=f_{\infty} everywhere.878787Pathak, Rojas, and Simpson [52, Definition 3.8 p. 314] work with a variant of our ff_{\infty} that organises the case break depending on whether xx is in 𝖲𝖱ν\mathsf{SR}^{\nu}. Their approach has the advantage of making ff_{\infty}, which they denote as f^\widehat{f}, entirely independent of the witness fnf_{n}. Rute [61, Definition 3.17 p. 16] organises the case break the same as we do but sets it undefined when the limit does not exist, which prevents it from being an Lp(ν)L_{p}(\nu) Schnorr test.

Miyabe proved the following transfer result for going back and forth between Lp(ν)L_{p}(\nu)-computable functions and differences of Lp(ν)L_{p}(\nu) Schnorr tests:888888[45, Theorem 4.3 p. 7]. He proved it for p=1p=1, but the proof is the same for p1p\geq 1.

Proposition 11.2.

Suppose that ν\nu is a computable point of 𝒫(X)\mathcal{P}(X) and p1p\geq 1 is computable.

  1. (1)

    Suppose ff is a computable point of Lp(ν)L_{p}(\nu) with witness fnf_{n}. Then there are Lp(ν)L_{p}(\nu) Schnorr tests g,hg,h such that f=ghf_{\infty}=g-h on on 𝖲𝖱ν\mathsf{SR}^{\nu}.

  2. (2)

    Suppose that g,hg,h are Lp(ν)L_{p}(\nu) Schnorr tests. Then there is Lp(ν)L_{p}(\nu)-computable ff with witness fnf_{n} such that f=ghf_{\infty}=g-h on on 𝖲𝖱ν\mathsf{SR}^{\nu}.

Proof.

(Sketch) For (1), using Proposition 2.23, one shows that g=n(fn+1fn)+g=\sum_{n}(f_{n+1}-f_{n})^{+} and h=n(fn+1fn)h=\sum_{n}(f_{n+1}-f_{n})^{-} are equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to Lp(ν)L_{p}(\nu) Schnorr tests, where +\cdot^{+} and \cdot^{-} denote positive and negative parts. For (2), one uses Proposition 2.16. ∎

These observations allow us to restate Theorem 1.5 in terms of Lp(ν)L_{p}(\nu)-computable points, provided one assumes Schnorr disintegrations:

Corollary 11.3.

Suppose that XX is a computable Polish space and ν\nu is a computable probability measure. Suppose that n\mathscr{F}_{n} is an almost-full effective filtration, equipped with Schnorr disintegrations.

If p1p\geq 1 is computable, then the following three items are equivalent for xx in XX:

  1. (1)

    xx is in 𝖲𝖱ν(X)\mathsf{SR}^{\nu}(X).

  2. (2)

    xx is in 𝖪𝖱ν\mathsf{KR}^{\nu} and limn𝔼ν[fn](x)=f(x)\lim_{n}\mathbb{E}_{\nu}[f_{\infty}\mid\mathscr{F}_{n}](x)=f_{\infty}(x) for every Lp(ν)L_{p}(\nu) computable ff with witness fmf_{m}.

  3. (3)

    xx is in 𝖪𝖱ν\mathsf{KR}^{\nu} and limn𝔼ν[fn](x)\lim_{n}\mathbb{E}_{\nu}[f_{\infty}\mid\mathscr{F}_{n}](x) exists for every Lp(ν)L_{p}(\nu) computable ff with witness fmf_{m}.

Proof.

Suppose (2); we show Theorem 1.5(2). But simply note that any Lp(ν)L_{p}(\nu) Schnorr test ff has a witness fmf_{m} from Proposition 2.16 with f=ff=f_{\infty} everywhere.

Suppose Theorem 1.5(2); we show (2). Let ff be Lp(ν)L_{p}(\nu) computable with witness fmf_{m}. By the previous proposition, there are two Lp(ν)L_{p}(\nu) Schnorr tests g,hg,h such that f=ghf_{\infty}=g-h on 𝖲𝖱ν\mathsf{SR}^{\nu}. Since we are working with Schnorr disintegrations, by Proposition 7.2(1), one has that 𝔼ν[fn]=𝔼ν[ghn]\mathbb{E}_{\nu}[f_{\infty}\mid\mathscr{F}_{n}]=\mathbb{E}_{\nu}[g-h\mid\mathscr{F}_{n}] on 𝖲𝖱ν\mathsf{SR}^{\nu} for all n0n\geq 0. Then by Theorem 1.5(2) and Proposition 7.2(3), we have on 𝖲𝖱ν\mathsf{SR}^{\nu} that f=gh=limn(𝔼ν[gn]𝔼ν[hn])=limn𝔼ν[fn]f_{\infty}=g-h=\lim_{n}\big{(}\mathbb{E}_{\nu}[g\mid\mathscr{F}_{n}]-\mathbb{E}_{\nu}[h\mid\mathscr{F}_{n}]\big{)}=\lim_{n}\mathbb{E}_{\nu}[f_{\infty}\mid\mathscr{F}_{n}].

Finally, (2) trivially implies (3). And (3) implies Theorem 1.5(4) since again any Lp(ν)L_{p}(\nu) Schnorr test ff has a witness fmf_{m} from Proposition 2.16 with f=ff=f_{\infty} everywhere. ∎

12. Martingale convergence in L2(ν)L_{2}(\nu)

Our topic in this paper is convergence of the conditional expectations of random variables. But of course these are instances of martingales. In this brief section, we prove a result mentioned in §1.4, namely that one can characterise 𝖲𝖱ν\mathsf{SR}^{\nu} in terms of convergence of certain L2(ν)L_{2}(\nu) martingales. As mentioned there, this generalises a result of Rute from Cantor space to the more general setting.

If p1p\geq 1 and if n\mathscr{F}_{n} is a filtration, then a classical martingale in Lp(ν)L_{p}(\nu) adapted to n\mathscr{F}_{n} is a sequence MnM_{n} of n\mathscr{F}_{n} measurable functions in Lp(ν)L_{p}(\nu) such that Mn=𝔼ν[Mn+1n]M_{n}=\mathbb{E}_{\nu}[M_{n+1}\mid\mathscr{F}_{n}] ν\nu-a.s. When n\mathscr{F}_{n} is clear from context, we just say classical martingale in Lp(ν)L_{p}(\nu).

If p1p\geq 1 is computable and n\mathscr{F}_{n} is an effective filtration equipped with Schnorr disintegrations ρ(n)\rho^{(n)}, then a martingale of Lp(ν)L_{p}(\nu) Schnorr tests adapted to n\mathscr{F}_{n} and ρ(n)\rho^{(n)} is a uniformly computable sequence MnM_{n} of n\mathscr{F}_{n}-measurable Schnorr Lp(ν)L_{p}(\nu) tests such that Mn=𝔼ν[Mn+1n]M_{n}=\mathbb{E}_{\nu}[M_{n+1}\mid\mathscr{F}_{n}] on 𝖲𝖱ν\mathsf{SR}^{\nu}, where the version of the conditional expectation is that from the disintegration. When n\mathscr{F}_{n} and ρ(n)\rho^{(n)} are clear from context, we just say martingale of Lp(ν)L_{p}(\nu) Schnorr tests.

Here is an example:

Example 12.1.

(Products of mean one independent variables).

Suppose p1p\geq 1 is computable. Suppose that fn:X[0,]f_{n}:X\rightarrow[0,\infty] is a sequence of independent Lp(ν)L_{p}(\nu) Schnorr tests with 𝔼νfn=1\mathbb{E}_{\nu}f_{n}=1 for all n1n\geq 1. Suppose that n=σ(f1,,fn)\mathscr{F}_{n}=\sigma(f_{1},\ldots,f_{n}) is an effective filtration equipped with Schnorr disintegrations. Then Mn=i=1nfiM_{n}=\prod_{i=1}^{n}f_{i} is a martingale of Lp(ν)L_{p}(\nu) Schnorr tests.

To see this, note that the MnM_{n} are Lp(ν)L_{p}(\nu) Schnorr tests: they are non-negative lsc by Proposition 2.6, and by independence we have Mnp=i=1nfip\|M_{n}\|_{p}=\prod_{i=1}^{n}\|f_{i}\|_{p}, which is computable. On 𝖲𝖱ν\mathsf{SR}^{\nu} one has

𝔼ν[Mn+1n]=𝔼ν[fn+1Mnn]=Mn𝔼ν[fn+1n]=Mn𝔼νfn+1=Mn\mathbb{E}_{\nu}[M_{n+1}\mid\mathscr{F}_{n}]=\mathbb{E}_{\nu}[f_{n+1}\cdot M_{n}\mid\mathscr{F}_{n}]=M_{n}\cdot\mathbb{E}_{\nu}[f_{n+1}\mid\mathscr{F}_{n}]=M_{n}\cdot\mathbb{E}_{\nu}f_{n+1}=M_{n}

The second identity is by taking out what is known (Proposition 7.4(3)) and the third identity is by the rôle of independence (Proposition 7.8).

Our goal is to prove the following:

Theorem 12.2.

Suppose that ν\nu is a computable point of 𝒫(X)\mathcal{P}(X). Let n\mathscr{F}_{n} be an almost-full effective filtration equipped with Schnorr disintegrations.

The following are equivalent for xx in XX:

  1. (1)

    xx is in 𝖲𝖱ν\mathsf{SR}^{\nu}.

  2. (2)

    xx is in 𝖪𝖱ν\mathsf{KR}^{\nu} and limnMn(x)\lim_{n}M_{n}(x) exists for every martingale MnM_{n} of L2(ν)L_{2}(\nu) Schnorr tests such that both supnMn2\sup_{n}\|M_{n}\|_{2} is computable and the maximal function supnMn\sup_{n}M_{n} is a L2(ν)L_{2}(\nu) Schnorr test.

This theorem generalises a result of Rute.898989[61, Corollary 6.8 and Theorem 12.6]. But there are two important differences between our result and Rute’s. First, Rute’s analogue of the direction from (2) to (1) of Theorem 12.2 only works for Cantor space and the uniform measure. Second, Rute’s results do not require that the maximal function supnMn\sup_{n}M_{n} be a L2(ν)L_{2}(\nu) Schnorr test.

We do not know the answer to the following:

Question 12.3.

Does Theorem 12.2 also hold for all computable p>1p>1?

It is clear from the proofs below that (2) to (1) holds for computable p>1p>1. Hence it is a question of (1) to (2).

Throughout the remainder of this section, XX is a computable Polish space, ν\nu is a computable point of 𝒫(X)\mathcal{P}(X), and n\mathscr{F}_{n} is an effective filtration equipped with Schnorr disintegrations. We only assume that n\mathscr{F}_{n} is almost-full in the last proposition, and flag this assumption when it comes up.

We begin by noting the following two elementary results:

Proposition 12.4.

If MnM_{n} is a martingale of Lp(ν)L_{p}(\nu) Schnorr tests, then Mn=𝔼ν[Mmn]M_{n}=\mathbb{E}_{\nu}[M_{m}\mid\mathscr{F}_{n}] on 𝖲𝖱ν\mathsf{SR}^{\nu} for all m>nm>n.

Proof.

This is by an induction on m>nm>n. Suppose it holds for m>nm>n. Then Mn=𝔼ν[Mmn]M_{n}=\mathbb{E}_{\nu}[M_{m}\mid\mathscr{F}_{n}] on 𝖲𝖱ν\mathsf{SR}^{\nu}. Since Mm=𝔼ν[Mm+1m]M_{m}=\mathbb{E}_{\nu}[M_{m+1}\mid\mathscr{F}_{m}] on 𝖲𝖱ν\mathsf{SR}^{\nu} and since we are working with a Schnorr disintegration, we have by Proposition 7.2(1) that 𝔼ν[Mmn]=𝔼ν[𝔼ν[Mm+1m]n]\mathbb{E}_{\nu}[M_{m}\mid\mathscr{F}_{n}]=\mathbb{E}_{\nu}[\mathbb{E}_{\nu}[M_{m+1}\mid\mathscr{F}_{m}]\mid\mathscr{F}_{n}] on 𝖲𝖱ν\mathsf{SR}^{\nu}. By the tower property (Proposition 7.7), this latter is equal to 𝔼ν[Mm+1n]\mathbb{E}_{\nu}[M_{m+1}\mid\mathscr{F}_{n}] on 𝖲𝖱ν\mathsf{SR}^{\nu}. ∎

Proposition 12.5.

If MnM_{n} is a classical martingale in Lp(ν)L_{p}(\nu), then MmpMnp\|M_{m}\|_{p}\geq\|M_{n}\|_{p} for all m>nm>n.

Proof.

The function x|x|px\mapsto\left|x\right|^{p} is a convex function, and hence |Mn|p\left|M_{n}\right|^{p} is a submartingale.909090[75, p. 138]. And the expectation of a submartingale is always non-decreasing. ∎

The following gives a canonical example of a martingale of Lp(ν)L_{p}(\nu) Schnorr tests, and in conjunction with Theorem 1.5 gives the (2) to (1) direction of Theorem 12.2.

Proposition 12.6.

Suppose that p1p\geq 1 is computable.

If ff is an Lp(ν)L_{p}(\nu) Schnorr test, then Mn:=𝔼ν[fn]M_{n}:=\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}] is a martingale of Lp(ν)L_{p}(\nu) Schnorr tests.

If p>1p>1 and the filtration is almost-full then the maximal function supnMn\sup_{n}M_{n} is also an Lp(ν)L_{p}(\nu) Schnorr test and supnMnp\sup_{n}\|M_{n}\|_{p} is computable.

Proof.

The function MnM_{n} is non-negative lsc by Proposition 7.1, and it is Lp(ν)L_{p}(\nu)-computable by Proposition 7.5. To see that it satisfies the martingale condition, from Mn+1=𝔼ν[fn+1]M_{n+1}=\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n+1}] everywhere we have 𝔼ν[Mn+1n]=𝔼ν[𝔼ν[fn+1]n]\mathbb{E}_{\nu}[M_{n+1}\mid\mathscr{F}_{n}]=\mathbb{E}_{\nu}[\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n+1}]\mid\mathscr{F}_{n}] everywhere. And by the tower property (Proposition 7.7), the latter is equal to 𝔼ν[fn]\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}] on 𝖲𝖱ν\mathsf{SR}^{\nu}, which is by definition MnM_{n}.

Suppose now that the filtration is almost-full and p>1p>1. By almost-fullness and Theorem 1.5, we have that f=limnMnf=\lim_{n}M_{n} on 𝖲𝖱ν\mathsf{SR}^{\nu}. Since p>1p>1 we have supnMn\sup_{n}M_{n} is in Lp(ν)L_{p}(\nu) by Lemma 5.1(1). Then we can dominate MnpM_{n}^{p} by (supnMn)p(\sup_{n}M_{n})^{p} and argue by DCT as follows, where the first identity comes from Proposition 12.5:

supnMnpp=limnMnpp=limnMnpdν=limnMnpdν=fpdν\sup_{n}\|M_{n}\|_{p}^{p}=\lim_{n}\|M_{n}\|_{p}^{p}=\lim_{n}\int M_{n}^{p}\;d\nu=\int\lim_{n}M_{n}^{p}\;d\nu=\int f^{p}\;d\nu (12.1)

Since ff is by hypothesis a computable point of Lp(ν)L_{p}(\nu), we have that supnMnpp\sup_{n}\|M_{n}\|_{p}^{p} is computable and hence likewise supnMnp\sup_{n}\|M_{n}\|_{p} is computable. ∎

In conjunction with Corollary 11.3, the following proposition then gives the (1) to (2) direction of Theorem 12.2. As mentioned in §1.4, the proof largely follows the outline of Rute’s own Hilbert space proof.

Proposition 12.7.

Suppose that the filtration is almost-full.

Suppose MnM_{n} is a martingale MnM_{n} of L2(ν)L_{2}(\nu) Schnorr tests such that both supnMn2\sup_{n}\|M_{n}\|_{2} is computable and supnMn\sup_{n}M_{n} is a L2(ν)L_{2}(\nu) Schnorr test.

Then there is L2(ν)L_{2}(\nu)-computable function ff such that Mn=𝔼ν[fn]M_{n}=\mathbb{E}_{\nu}[f_{\infty}\mid\mathscr{F}_{n}] on 𝖲𝖱ν\mathsf{SR}^{\nu} for each n0n\geq 0.

Further, the L2(ν)L_{2}(\nu)-computable function ff can be taken to be a pointwise limit of a computable subsequence of the MnM_{n}, which limit exists at least on 𝖲𝖱ν\mathsf{SR}^{\nu}.

Proof.

Recall that for n>kn>k we have by Hilbert space methods in L2(ν)L_{2}(\nu) that 𝔼νMkMn=𝔼νMk2\mathbb{E}_{\nu}M_{k}M_{n}=\mathbb{E}_{\nu}M_{k}^{2}.919191[25, 488]. This implies that for n>kn>k we have:

MnMk22=𝔼ν(MnMk)2=𝔼νMn2𝔼νMk2=Mn22Mk22\|M_{n}-M_{k}\|_{2}^{2}=\mathbb{E}_{\nu}(M_{n}-M_{k})^{2}=\mathbb{E}_{\nu}M_{n}^{2}-\mathbb{E}_{\nu}M_{k}^{2}=\|M_{n}\|_{2}^{2}-\|M_{k}\|_{2}^{2}

Let f=limnMnf=\lim_{n}M_{n}, which classically is in L2(ν)L_{2}(\nu). Since supnMn\sup_{n}M_{n} is in L2(ν)L_{2}(\nu), we can dominate (MnMk)2(M_{n}-M_{k})^{2} by 2(supnMn)22\cdot(\sup_{n}M_{n})^{2} and argue by DCT and the previous equation that:

fMk22=limnMnMk22=limnMn22Mk22=(supnMn2)2Mk22\|f-M_{k}\|_{2}^{2}=\lim_{n}\|M_{n}-M_{k}\|_{2}^{2}=\lim_{n}\|M_{n}\|_{2}^{2}-\|M_{k}\|_{2}^{2}=(\sup_{n}\|M_{n}\|_{2})^{2}-\|M_{k}\|_{2}^{2}

Since the latter is a computable real which goes to zero, we can compute a subsequence Mk(n)M_{k(n)} which converges to ff fast in L2(ν)L_{2}(\nu). By Proposition 4.3 and Definition 11.1, we have that Mk(n)fM_{k(n)}\rightarrow f_{\infty} on 𝖲𝖱ν\mathsf{SR}^{\nu}.

For each m0m\geq 0, let gm=𝔼ν[supnMnm]g_{m}=\mathbb{E}_{\nu}[\sup_{n}M_{n}\mid\mathscr{F}_{m}], so that gmg_{m} is an L2(ν)L_{2}(\nu) Schnorr test by Proposition 7.6(2), and so gmg_{m} is finite on 𝖲𝖱ν\mathsf{SR}^{\nu}. Then by Conditional DCT (Proposition 7.4(2)), for each m0m\geq 0 we have that 𝔼ν[Mk(n)m]𝔼ν[fm]\mathbb{E}_{\nu}[M_{k(n)}\mid\mathscr{F}_{m}]\rightarrow\mathbb{E}_{\nu}[f_{\infty}\mid\mathscr{F}_{m}] on 𝖲𝖱ν\mathsf{SR}^{\nu}. But by Proposition 12.4 for k(n)>mk(n)>m we have that the former is equal to MmM_{m} on 𝖲𝖱ν\mathsf{SR}^{\nu}. Hence we have that Mm=𝔼ν[fm]M_{m}=\mathbb{E}_{\nu}[f_{\infty}\mid\mathscr{F}_{m}] on 𝖲𝖱ν\mathsf{SR}^{\nu} for each m0m\geq 0. ∎

13. Conclusion

The main results of this paper (Theorems 1.5, 1.6, 1.8, 1.9, 1.11) characterise the points under which Lévy’s Upward Theorem holds in terms of notions from algorithmic randomness and the rates of convergence in terms of concepts from the classical theory of computation. As discussed in §1.4 this builds on work by previous authors. That which is new are the results on rates of convergence in Theorems 1.6, 1.9, the articulation of the general framework of effective disintegrations (see Definition 1.3, §7 for fundamental properties, and Appendix B for examples), a conceptually new proof of the characterisation of density randomness in the more general framework of effective disintegrations for p>1p>1 (Theorem 1.9), and the articulation of the new concept of Maximal Doob Randomness (cf. Definition 1.10, Theorem 1.11, and Question 1.12). As far as Schnorr randomness goes, we noted in §1.4 that Theorem 1.5 can be derived from Rute’s work, modulo the verification of certain properties of effective disintegrations and the Miyabe translation method in §11. We have extended Rute on Schnorr randomness in the generalisation of the L2(ν)L_{2}(\nu) martingale result in §12. We have also sought to present very accessible proofs, based almost entirely on the concept of Lp(ν)L_{p}(\nu) Schnorr test.

Our results also contribute to understanding the significance of convergence to the truth results for Bayesian inference. As was pointed out by philosophers of science, the probability one qualification in theorems like Lévy’s Upward Theorem raises the spectre of arbitrariness: a Bayesian with credences represented by a probability measure ν\nu believes in convergence to the truth with certainty, but might do so only by arbitrarily packaging into a set of probability zero those points at which convergence fails.929292[2], [18, pp. 144 ff], [28, pp. 28-29]. We have shown that for certain classes of effective random variables the packaging is anything but arbitrary. The probability one set on which convergence to the truth is successful coincides with standard classes of points which are algorithmically random by the lights of the computable probability measure. Thus, the effective typicality expressed by convergence to the truth is extensionally equivalent with a principled effective typicality of the underlying probability measure.

Appendix A Examples of classical disintegrations

In this appendix, we review two classical examples of disintegrations. This also affords us the opportunity to illustrate natural circumstances in which Lévy’s Upward Theorem need not hold for all points. Another reason to dwell on these two examples is that one of the main theorems of Rohlin is that, up to Borel isomorphism, “blendings” of these two examples are the only examples of disintegrations of countably generated σ\sigma-algebras.939393[59, §4 pp. 40-41], [12, Theorem 1.12 p. 12]. Our diagrams in this appendix are inspired by the few diagrams in Einsiedler-Ward,949494[19, pp. 122-123]. although they only work with a single σ\sigma-algebra rather than a filtration.

Most concrete examples of disintegrations involve products. We write λμ\lambda\otimes\mu for the product measure on Y×ZY\times Z formed from finite measure λ\lambda on YY and finite measure μ\mu on ZZ.

Example A.1.

(Refined partitions of the unit square). Let X=[0,1]×[0,1]X=[0,1]\times[0,1] with measure ν=mm\nu=m\otimes m being the product of Lebesgue measure mm on [0,1][0,1] with itself. Let 𝒟n\mathscr{D}_{n} be the dyadic partition of XX into 4n14^{n-1} many squares, and let n\mathscr{F}_{n} be the σ\sigma-algebra it generates. We can visualise the elements of n\mathscr{F}_{n} as any shape one can form from the squares in the below diagrams, so that like in pixelations more detailed shapes become available as nn gets larger:

1\mathscr{F}_{1}2\mathscr{F}_{2}3\mathscr{F}_{3}

In this diagram, we use the familiar diagrammatic conventions from point-set topology to indicate which components of the partition contain the edges: for instance, for n2n\geq 2, the southwest square and the northwest square have two edges, the southeast square has three edges, and the northeast square has four edges. Then the following map is a disintegration of n\mathscr{F}_{n}, where +(X)\mathcal{M}^{+}(X) is again the set of finite Borel measures on XX:

ρ(n):X+(X) by ρ(n)w=Q𝒟nν(Q)IQ(w)\rho^{(n)}:X\rightarrow\mathcal{M}^{+}(X)\hskip 8.53581pt\mbox{ by }\hskip 8.53581pt\rho^{(n)}_{w}=\sum_{Q\in\mathscr{D}_{n}}\nu(\cdot\mid Q)\cdot I_{Q}(w) (A.1)

Further, using equation (1.1) from §1.2, one has the following associated formula for the version of conditional expectation of ff with respect to n\mathscr{F}_{n}:

𝔼ν[fn](w)=Q𝒟n(1ν(Q)Qf(x)dν(x))IQ(w)\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](w)=\sum_{Q\in\mathscr{D}_{n}}\bigg{(}\frac{1}{\nu(Q)}\int_{Q}f(x)\;d\nu(x)\bigg{)}\cdot I_{Q}(w) (A.2)

Suppose that at stage nn, the agent’s world ww is located in the square Qn(w)Q_{n}(w) from 𝒟n\mathscr{D}_{n}. Intuitively this means that the agent’s evidence at this stage of inquiry is Qn(w)Q_{n}(w). Then (A.2) says that the agent’s best estimate as to the value of a random variable ff at this stage is obtained by averaging ff over the event Qn(w)Q_{n}(w) according to the prior probability measure ν\nu, and then making it higher to the extent that the prior probability ν(Qw(n))\nu(Q_{w}(n)) is lower. In the case where the random variable ff is the indicator function ICI_{C} of a Borel event CC, this best estimate is just the usual conditional probability ν(CQn(w))\nu(C\mid Q_{n}(w)). For instance, if CC is the closed polygon displayed below, then the conditional probability of the agent at stage nn is higher than if she were at another world ww^{\prime} iff there is more overlap between Qn(w),CQ_{n}(w),C than between Qn(w),CQ_{n}(w^{\prime}),C:

1\mathscr{F}_{1}2\mathscr{F}_{2}3\mathscr{F}_{3}

This example also vividly illustrates how limn𝔼ν[fn](w)=f(w)\lim_{n}\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](w)=f(w) can fail. For instance, take the vertex w=(.75,.75)w=(.75,.75), which is in the polygon since it is closed. For all n3n\geq 3, this point is the southwest vertex of a dyadic square in n\mathscr{F}_{n}, and such a square overlaps the closed polygon only at this vertex. Hence for the usc function f=ICf=I_{C}, one has both f(w)=1f(w)=1 and 𝔼ν[fn](w)=0\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](w)=0 for all n3n\geq 3.

In simple examples like this one, geometric intuition can guide us as to what points limn𝔼ν[fn]=f\lim_{n}\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}]=f holds. The further assurance the classical version of Levy’s Upward Theorem provides is that limn𝔼ν[fn]=f\lim_{n}\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}]=f holds on a set of ν\nu-probability one, even when geometric intuition is unavailable. The additional assurance that Theorems 1.5,1.8, 1.11 provides is that limn𝔼ν[fn]=f\lim_{n}\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}]=f holds on the random points relative to ν\nu, for a large class of effective random variables ff. From this perspective, the problem with our vertex w=(.75,.75)w=(.75,.75) is that it is not sufficiently random, which enabled us to construct a random variable which failed to converge to the truth at this point.

Example A.2.

(Refined lines in the unit square) Let X=[0,1]×[0,1]X=[0,1]\times[0,1] with Lebesgue measure ν=mm\nu=m\otimes m being the product of Lebesgue measure mm on [0,1][0,1] with itself. Let 𝒢n\mathscr{G}_{n} be the σ\sigma-algebra on [0,1]×[0,1][0,1]\times[0,1] generated by events of the form B×QB\times Q, where B[0,1]B\subseteq[0,1] is Borel and where QQ is from a dyadic partition 𝒟n\mathscr{D}_{n} of [0,1][0,1] into 2n12^{n-1} (half)-closed intervals of equal length. Intuitively, 𝒢n\mathscr{G}_{n} is the σ\sigma-algebra of evidence where the agent knows everything there is to know about the xx-component at the outset, but is progressively learning more about the yy-component. Since events of the form {x}×[0,1]\{x\}\times[0,1] are in 𝒢1\mathscr{G}_{1}, we can depict the σ\sigma-algebra 𝒢1\mathscr{G}_{1} as the decomposition of XX into vertical lines. Likewise, we can depict 𝒢2\mathscr{G}_{2} as the decomposition of XX into half vertical lines, etc. While we draw only ten such vertical lines in the below diagram, the idea is that XX is being decomposed into continuum-many such vertical lines at each stage:

𝒢1\mathscr{G}_{1}𝒢2\mathscr{G}_{2}𝒢3\mathscr{G}_{3}

Then the following map is a disintegration of 𝒢n\mathscr{G}_{n}, where δu\delta_{u} is the Dirac measure centred on uu:

ρ(n):X+(X) by ρ(n)(u,v)=δuQ𝒟nm(Q)IQ(v)\rho^{(n)}:X\rightarrow\mathcal{M}^{+}(X)\hskip 8.53581pt\mbox{ by }\hskip 8.53581pt\rho^{(n)}_{(u,v)}=\delta_{u}\otimes\sum_{Q\in\mathscr{D}_{n}}m(\cdot\mid Q)\cdot I_{Q}(v) (A.3)

Further, using equation (1.1) from §1.2, one has the following associated formula for the version of conditional expectation of gg with respect to 𝒢n\mathscr{G}_{n}:

𝔼ν[g𝒢n](u,v)=Q𝒟n(1m(Q)Qg(u,t)dm(t))IQ(v)\mathbb{E}_{\nu}[g\mid\mathscr{G}_{n}](u,v)=\sum_{Q\in\mathscr{D}_{n}}\bigg{(}\frac{1}{m(Q)}\int_{Q}g(u,t)\;dm(t)\bigg{)}\cdot I_{Q}(v) (A.4)

Suppose that at stage nn, the agent’s world (u,v)(u,v) is such that its second coordinate vv located in the interval Qn(v)Q_{n}(v) from 𝒟n\mathscr{D}_{n}. Intuitively this means that the agent’s evidence at this stage of inquiry is the line {u}×Qn(v)\{u\}\times Q_{n}(v). Then (A.2) says that the agent’s best estimate as to the value of a random variable gg at this world and stage is obtained by defining the one-place random variable f(v)=g(u,v)f(v)=g(u,v) and then doing a one-dimensional analogue of the update in Example A.1. For instance, if CC is the displayed closed triangle and we consider the usc function g=ICg=I_{C}, then 𝔼ν[g𝒢n](u,v)\mathbb{E}_{\nu}[g\mid\mathscr{G}_{n}](u,v) is obtained by calculating the length of the line C({u}×Qn(v))C\cap(\{u\}\times Q_{n}(v)), and then by multiplying by a factor of 2n12^{n-1} which is responsive to smaller partitions of the yy-axis involving less likely events. We illustrate this with respect to the marked point (u,v)=(13,23)(u,v)=(\frac{1}{3},\frac{2}{3}) in the below diagram, where the line C({u}×Qn(v))C\cap(\{u\}\times Q_{n}(v)) is indicated with a heavier dark line:

𝒢1\mathscr{G}_{1}𝒢2\mathscr{G}_{2}𝒢3\mathscr{G}_{3}

In contrast to the previous Example A.1, in this example many of the events in 𝒢n\mathscr{G}_{n} have measure zero according to the prior probability measure ν\nu. Like in the previous Example A.1, we have natural pointwise failures of limn𝔼ν[fn]=f\lim_{n}\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}]=f here as well: the rightmost vertex of the triangle displays the same kind of failure as in the previous example. And like in that case, the interpretation suggested by Theorems 1.5,1.8, 1.11 is that the vertex is insufficiently random.

Appendix B Examples of effective disintegrations

In this section, we describe several examples of effective disintegrations (cf. Definition 1.3). We focus for the most part on effectivizing the two paradigmatic Examples A.1-A.2 from the previous appendix, but we also include a countable product (Example B.13). In a sequel to this paper, we look also at Bayesian parameter spaces and sample spaces.

One example like Example A.1 is already widely-used in algorithmic randomness, although it is not usually thematized as such:

Example B.1.

(The canonical concrete refined partition disintegrations).

Suppose that T<T\subseteq\mathbb{N}^{<\mathbb{N}} is a computable tree with no dead ends. Let X=[T]X=[T], the paths through TT, which is a computable Polish space, and suppose ν\nu in 𝒫(X)\mathcal{P}(X) is computable with full support.

Suppose that n\mathscr{F}_{n} is the effective refined partition generated by the length nn strings in TT. That is, n\mathscr{F}_{n} is generated by the sets [σ][\sigma] of paths in TT through the length nn strings σ\sigma.

Let ρ(n):X𝒫(X)\rho^{(n)}:X\rightarrow\mathcal{P}(X) by ρω=ν([ωn])\rho_{\omega}=\nu(\cdot\mid[\omega\upharpoonright n]).

Then ρ(n)\rho^{(n)} is a Martin-Löf disintegration of n\mathscr{F}_{n} with respect to ν\nu and one has the following expression for the version of conditional expectation, which is defined for all ff in 𝕃1(ν)\mathbb{L}_{1}(\nu) and all ω\omega in XX:

𝔼ν[fn](ω)=1ν([ωn])[ωn]fdν\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](\omega)=\frac{1}{\nu([\omega\upharpoonright n])}\int_{[\omega\upharpoonright n]}f\;d\nu (B.1)

Since we want to generalise this in what follows, we defer the verification that it is a Martin-Löf disintegration.

The following is an effective disintegration like Example A.2:

Example B.2.

(The canonical concrete refined lines disintegrations).

Suppose that S,T<S,T\subseteq\mathbb{N}^{<\mathbb{N}} are computable trees with no dead ends. Let Y=[S]Y=[S] and Z=[T]Z=[T], the paths through S,TS,T respectively, which are computable Polish spaces, and suppose λ\lambda in 𝒫(Y)\mathcal{P}(Y) and μ\mu in 𝒫(Z)\mathcal{P}(Z) are computable with full support.

Let n\mathscr{F}_{n} be the σ\sigma-algebra on Y×ZY\times Z generated by sets of the form U×[τ]U\times[\tau], where UU ranges over c.e. opens from a λ\lambda-computable basis on YY, and where τ\tau ranges over length nn strings in TT.

Then the map ρ:Y×Z+(Y×Z)\rho:Y\times Z\rightarrow\mathcal{M}^{+}(Y\times Z) given by ρ(n)(ω,ω)=δωμ([ωn])\rho^{(n)}_{(\omega,\omega^{\prime})}=\delta_{\omega}\otimes\mu(\cdot\mid[\omega^{\prime}\upharpoonright n]) is a Martin-Löf disintegration of n\mathscr{F}_{n} with respect to λμ\lambda\otimes\mu, and one has the following expression for the version of conditional expectation, which for each ff in 𝕃1(λμ)\mathbb{L}_{1}(\lambda\otimes\mu) is defined for (λμ)(\lambda\otimes\mu)-a.s. many (ω,ω)(\omega,\omega^{\prime}) in Y×ZY\times Z:

𝔼λμ[fn](ω,ω)=1μ([ωn])[ωn]f(ω,θ)dμ(θ)\mathbb{E}_{\lambda\otimes\mu}[f\mid\mathscr{F}_{n}](\omega,\omega^{\prime})=\frac{1}{\mu([\omega^{\prime}\upharpoonright n])}\int_{[\omega^{\prime}\upharpoonright n]}f(\omega,\theta)\;d\mu(\theta) (B.2)

Note that ω\omega is free under the integral sign. Since there are no continuity assumptions on ff (it is merely an element of 𝕃1(ν)\mathbb{L}_{1}(\nu)) the value of the conditional expectation in (B.2) can apriori change drastically with small changes of ω\omega. By contrast, in (B.1), ω\omega’s contribution is restricted to its first nn bits.

In what follows, we want to generalise these two examples to a broader class of computable Polish spaces and verify that they are indeed Martin-Löf disintegrations. We begin by generalising the way in which the previous examples involve partitions. We define a special case of Definition 1.1(11):

Definition B.3.

A sub-σ\sigma-algebra \mathscr{F} of the Borel sets on XX is a ν\nu-effective partition if it is generated by a computable sequence of events {Ai:iI}\{A_{i}:i\in I\} from the algebra 𝒜\mathscr{A} generated by ν\nu-computable basis \mathscr{B} such that the events {Ai:iI}\{A_{i}:i\in I\} are a partition of XX.

Given such a partition with its computable index set II, we define the c.e. set I+={iI:ν(Ai)>0}I^{+}=\{i\in I:\nu(A_{i})>0\}.

Further, a ν\nu-effective softening of \mathscr{F} is a pairwise disjoint computable sequence of c.e. opens {Ui:iI}\{U_{i}:i\in I\} such that Ui=AiU_{i}=A_{i} on 𝖪𝖱ν\mathsf{KR}^{\nu}.

Proposition B.4.

Every effective partition has an effective softening.

Proof.

Since the computable sequence AiA_{i} comes from the algebra generated by a ν\nu-computable basis, by Proposition 2.13, there is a computable sequence of c.e. opens ViV_{i} and effectively closed CiViC_{i}\supseteq V_{i} with ν(Ci)=ν(Ai)\nu(C_{i})=\nu(A_{i}) and Vi=AiV_{i}=A_{i} on 𝖪𝖱ν\mathsf{KR}^{\nu}. Then define recursively the sequence of c.e. opens by U0=V0U_{0}=V_{0} and Un+1=Vn+1mnCmU_{n+1}=V_{n+1}\setminus\bigcup_{m\leq n}C_{m}. ∎

Softenings of full partitions are a canonical way to obtain almost-full effective filtrations (cf. Definition 1.1(12)):

Proposition B.5.

Suppose that n\mathscr{F}_{n} is a full ν\nu-effective partition equipped with effective softenings which generate σ\sigma-algebras 𝒢n\mathscr{G}_{n}. Then 𝒢n\mathscr{G}_{n} is an almost-full ν\nu-effective partition.

Proof.

Uniformly from an index for a c.e. open UU, we can compute an index for a sequence AmiA_{m_{i}} from the sequences which generate the filtration 0,1,\mathscr{F}_{0},\mathscr{F}_{1},\ldots such that U=iAmiU=\bigcup_{i}A_{m_{i}}. Each AmiA_{m_{i}} is equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to UmiU_{m_{i}}, where the latter comes from the softening. Then UU is equal on 𝖪𝖱ν\mathsf{KR}^{\nu} to iUmi\bigcup_{i}U_{m_{i}}. ∎

Here is how to organise a suitably generalised version of a single stage of the filtration of Example A.1:

Proposition B.6.

Let ν\nu be a computable point of 𝒫(X)\mathcal{P}(X). Let \mathscr{F} be an effective partition {Ai:iI}\{A_{i}:i\in I\} of XX with effective softening {Ui:iI}\{U_{i}:i\in I\}.

Let ρ:X+(X)\rho:X\rightarrow\mathcal{M}^{+}(X) by ρx=iI+ν(Ui)IUi(x)\rho_{x}=\sum_{i\in I^{+}}\nu(\cdot\mid U_{i})\cdot I_{U_{i}}(x).

Then ρ\rho is a Martin-Löf disintegration of \mathscr{F} with respect to ν\nu and one has the following expression for the version of conditional expectation, which is defined for all ff in 𝕃1(ν)\mathbb{L}_{1}(\nu) and all xx in XX:

𝔼ν[f](x)=iI+(1ν(Ui)Uifdν)IUi(x)\mathbb{E}_{\nu}[f\mid\mathscr{F}](x)=\sum_{i\in I^{+}}\bigg{(}\frac{1}{\nu(U_{i})}\int_{U_{i}}f\;d\nu\bigg{)}\cdot I_{U_{i}}(x) (B.3)
Proof.

By the definition of effective softening, the sets Ui,UjU_{i},U_{j} for distinct i,ji,j in I+I^{+} have empty intersection. Hence, the map ρ\rho has codomain +(X)\mathcal{M}^{+}(X). In particular, if xx is in UiU_{i} for ii in I+I^{+}, then ρx=ν(Ui)\rho_{x}=\nu(\cdot\mid U_{i}), which is in 𝒫(X)\mathcal{P}(X) and hence in +(X)\mathcal{M}^{+}(X). But if xx not in any UiU_{i} for ii in I+I^{+}, then ρx=0\rho_{x}=0, which is a point of +(X)\mathcal{M}^{+}(X).959595If one does not introduce softenings, then this part of the argument breaks down and ρx\rho_{x} need not be a finite measure.

By the definition in (1.1) and since dν(Ui)dν=1ν(Ui)IUi\frac{d\nu(\cdot\mid U_{i})}{d\nu}=\frac{1}{\nu(U_{i})}\cdot I_{U_{i}} for ii in I+I^{+}, one has the following for all xx in XX:

𝔼ν[f](x)=iI+(Xf(v)dν(Ui)(v))IUi(x)=iI+(1ν(Ui)Uif(v)dν(v))IUi(x)\mathbb{E}_{\nu}[f\mid\mathscr{F}](x)=\sum_{i\in I^{+}}\big{(}\int_{X}f(v)\;d\nu(\cdot\mid U_{i})(v)\big{)}\cdot I_{U_{i}}(x)=\sum_{i\in I^{+}}\bigg{(}\frac{1}{\nu(U_{i})}\int_{U_{i}}f(v)\;d\nu(v)\bigg{)}\cdot I_{U_{i}}(x)

If jj in I+I^{+}, then when we integrate over xx in AjA_{j} with respect to ν\nu we then get

Aj𝔼ν[f](x)dν(x)=Aj1ν(Uj)Ujf(v)dν(v)dν(x)=Ajf(v)dν(v)\int_{A_{j}}\mathbb{E}_{\nu}[f\mid\mathscr{F}](x)\;d\nu(x)=\int_{A_{j}}\frac{1}{\nu(U_{j})}\int_{U_{j}}f(v)\;d\nu(v)\;d\nu(x)=\int_{A_{j}}f(v)\;d\nu(v)

If jj not in I+I^{+} then the event AjA_{j} is ν\nu-null and so trivially we get:

Aj𝔼ν[f](x)dν(x)=Ajf(v)dν(v)\int_{A_{j}}\mathbb{E}_{\nu}[f\mid\mathscr{F}](x)\;d\nu(x)=\int_{A_{j}}f(v)\;d\nu(v)

Since elements AjA_{j} generate \mathscr{F}, this shows that (B.3) is a version of the conditional expectation of ff with respect to \mathscr{F}. Further, it is totally defined for all ff in 𝕃1(ν)\mathbb{L}_{1}(\nu) and all xx in XX.

Since \mathscr{F} is an effective partition, one has that [x]=Ai[x]_{\mathscr{F}}=A_{i} for xx in AiA_{i}. Further, for xx in 𝖪𝖱νAi\mathsf{KR}^{\nu}\cap A_{i}, we have that ii in I+I^{+}. Hence for xx in 𝖪𝖱νAi\mathsf{KR}^{\nu}\cap A_{i} we have ρx([x]𝖪𝖱ν)=ρx(Ai𝖪𝖱ν)=ν(Ai𝖪𝖱νUi)=ν(AiUi)=1\rho_{x}([x]_{\mathscr{F}}\cap\mathsf{KR}^{\nu})=\rho_{x}(A_{i}\cap\mathsf{KR}^{\nu})=\nu(A_{i}\cap\mathsf{KR}^{\nu}\mid U_{i})=\nu(A_{i}\mid U_{i})=1. The same argument works for 𝖲𝖱ν\mathsf{SR}^{\nu} and 𝖬𝖫𝖱ν\mathsf{MLR}^{\nu}.

Suppose that UU is c.e. open and q0q\geq 0 is rational. Since the UiU_{i} are pairwise disjoint, one has that ρx(U)>q\rho_{x}(U)>q iff there is ii in I+I^{+} such that xx is in UiU_{i} and ν(UUi)>q\nu(U\mid U_{i})>q, which is a c.e. open condition in variable xx. ∎

Here is how to organise a suitably generalised version of the initial step of the filtration of Example A.2, that is, where the partition on the second component consists just of a single set (like 𝒢1\mathscr{G}_{1} in Example A.2).

Proposition B.7.

Suppose that Y,ZY,Z are computable Polish spaces. Suppose that λ\lambda is a computable point of 𝒫(Y)\mathcal{P}(Y) and μ\mu is a computable point of 𝒫(Z)\mathcal{P}(Z). Let \mathscr{F} be the λμ\lambda\otimes\mu-effective σ\sigma-algebra on Y×ZY\times Z generated by sets of the form U×ZU\times Z, where UU ranges over c.e. opens from a λ\lambda-computable basis on YY. Then ρ:Y×Z𝒫(Y×Z)\rho:Y\times Z\rightarrow\mathcal{P}(Y\times Z) given by ρ(u,v)=δuμ\rho_{(u,v)}=\delta_{u}\otimes\mu is a Martin-Löf disintegration of \mathscr{F} with respect to λμ\lambda\otimes\mu, and one has the following expression for the version of conditional expectation, which for each ff in 𝕃1(λμ)\mathbb{L}_{1}(\lambda\otimes\mu) is defined for (λμ)(\lambda\otimes\mu)-a.s. many (u,v)(u,v) in Y×ZY\times Z:

𝔼λμ[f](u,v)=Zf(u,t)dμ(t)\mathbb{E}_{\lambda\otimes\mu}[f\mid\mathscr{F}](u,v)=\int_{Z}f(u,t)\;d\mu(t) (B.4)
Proof.

By the definition in (1.1) and by Fubini-Tonelli, one has the following for ff in 𝕃1(λμ)\mathbb{L}_{1}(\lambda\otimes\mu), and by the same theorem it is defined for λ\lambda-a.s. many uu in YY, and hence for (λμ)(\lambda\otimes\mu)-a.s. many (u,v)(u,v) in Y×ZY\times Z:

𝔼λμ[f](u,v)=Y×Zf(s,t)dρ(u,v)(s,t)=ZYf(s,t)dδu(s)dμ(t)=Zf(u,t)dμ(t)\displaystyle\mathbb{E}_{\lambda\otimes\mu}[f\mid\mathscr{F}](u,v)=\int_{Y\times Z}f(s,t)\;d\rho_{(u,v)}(s,t)=\int_{Z}\int_{Y}f(s,t)\;d\delta_{u}(s)\;d\mu(t)=\int_{Z}f(u,t)\;d\mu(t)

Since vv from ZZ does not appear free in this last term, when we integrate with respect to vv in ZZ we get:

Z𝔼λμ[f](u,v)dμ(v)=Zf(u,t)dμ(t)\int_{Z}\mathbb{E}_{\lambda\otimes\mu}[f\mid\mathscr{F}](u,v)\;d\mu(v)=\int_{Z}f(u,t)\;d\mu(t)

Hence for Borel subsets BB of YY we have by Fubini-Tonelli that:

B×Z𝔼λμ[f](u,v)d(λμ)(u,v)\displaystyle\int_{B\times Z}\mathbb{E}_{\lambda\otimes\mu}[f\mid\mathscr{F}](u,v)\;d(\lambda\otimes\mu)(u,v) =BZf(u,t)dμ(t)dλ(u)\displaystyle=\int_{B}\int_{Z}f(u,t)\;d\mu(t)\;d\lambda(u)
=B×Zf(u,t)d(λμ)(u,t)\displaystyle=\int_{B\times Z}f(u,t)\;d(\lambda\otimes\mu)(u,t)

This shows that (B.4) is a version of the condition expectation of ff with respect to \mathscr{F}.

Note that [(u,v)]={u}×Z[(u,v)]_{\mathscr{F}}=\{u\}\times Z, for any (u,v)Y×Z(u,v)\in Y\times Z. Further, recall that 𝖪𝖱λμ(Y×Z)=𝖪𝖱λ(Y)×𝖪𝖱μ(Z)\mathsf{KR}^{\lambda\otimes\mu}(Y\times Z)=\mathsf{KR}^{\lambda}(Y)\times\mathsf{KR}^{\mu}(Z).969696One can easily check this by hand. It also follows from the fact that Kurtz randomness is preserved both ways under computable continuous open maps. Hence for (u,v)(u,v) in 𝖪𝖱λμ(Y×Z)\mathsf{KR}^{\lambda\otimes\mu}(Y\times Z) one has the identity:

[(u,v)]𝖪𝖱λμ(Y×Z)=({u}𝖪𝖱λ(Y))×(Z𝖪𝖱μ(Z))={u}×𝖪𝖱μ(Z)[(u,v)]_{\mathscr{F}}\cap\mathsf{KR}^{\lambda\otimes\mu}(Y\times Z)=(\{u\}\cap\mathsf{KR}^{\lambda}(Y))\times(Z\cap\mathsf{KR}^{\mu}(Z))=\{u\}\times\mathsf{KR}^{\mu}(Z)

From this we get ρ(u,v)([(u,v)]𝖪𝖱λμ(Y×Z))=δu({u})μ(𝖪𝖱μ(Z))=1\rho_{(u,v)}([(u,v)]_{\mathscr{F}}\cap\mathsf{KR}^{\lambda\otimes\mu}(Y\times Z))=\delta_{u}(\{u\})\cdot\mu(\mathsf{KR}^{\mu}(Z))=1.

In this next paragraph, we use some notation familiar from Fubini-Tonelli, namely if AY×ZA\subseteq Y\times Z and ss in YY, then AsA_{s} is defined to be {tZ:(s,t)A}\{t\in Z:(s,t)\in A\}.

For Schnorr disintegrations, suppose that (u,v)(u,v) is in 𝖲𝖱λμ(Y×Z)\mathsf{SR}^{\lambda\otimes\mu}(Y\times Z), so that uu is in 𝖲𝖱λ(Y)\mathsf{SR}^{\lambda}(Y). Since [(u,v)]={u}×Z[(u,v)]_{\mathscr{F}}=\{u\}\times Z, we want to show that (δuμ)(A)=1(\delta_{u}\otimes\mu)(A)=1, where AA is the event ({u}×Z)𝖲𝖱λμ(Y×Z)(\{u\}\times Z)\cap\mathsf{SR}^{\lambda\otimes\mu}(Y\times Z). We have Au={tZ:(u,t)𝖲𝖱λμ(Y×Z)}A_{u}=\{t\in Z:(u,t)\in\mathsf{SR}^{\lambda\otimes\mu}(Y\times Z)\}. By choosing a Turing degree 𝐚{\bf a} which computes a fast Cauchy sequence for uu, we have by van Lambalgen’s Theorem that Au𝖲𝖱μ,𝐚(Z)A_{u}\supseteq\mathsf{SR}^{\mu,{\bf a}}(Z) and so μ(Au)=1\mu(A_{u})=1.979797This “hard” direction of van Lambalgen’s Theorem works in arbitrary computable Polish spaces with computable measures, basically because it is Fubini-Tonelli type argument. It similarly works for 𝖬𝖫𝖱ν\mathsf{MLR}^{\nu}. For the setting of Cantor space with uniform measure, see discussion in [15, pp. 257-258, 357]. Then by Fubini-Tonelli ρ(u,v)(A)=(δuμ)(A)=Yμ(As)dδu(s)=μ(Au)=1\rho_{(u,v)}(A)=(\delta_{u}\otimes\mu)(A)=\int_{Y}\mu(A_{s})\;d\delta_{u}(s)=\mu(A_{u})=1, which is what we wanted to show. The argument for Martin-Löf disintegrations is similar.

Suppose that WY×ZW\subseteq Y\times Z is c.e. open. Then we can write W=iUi×ViW=\bigcup_{i}U_{i}\times V_{i} where UiY,ViZU_{i}\subseteq Y,V_{i}\subseteq Z are computable sequences of c.e. opens with UiUi+1U_{i}\subseteq U_{i+1} and ViVi+1V_{i}\subseteq V_{i+1}. Then for rational q0q\geq 0, we have ρu,v(W)>q\rho_{u,v}(W)>q iff there is i0i\geq 0 with ρ(u,v)(Ui×Vi)>q\rho_{(u,v)}(U_{i}\times V_{i})>q, which happens iff there is i0i\geq 0 with δu(Ui)μ(Vi)>q\delta_{u}(U_{i})\cdot\mu(V_{i})>q, which happens iff there is i0i\geq 0 with uu in UiU_{i} and μ(Vi)>q\mu(V_{i})>q. This is a c.e. open condition in variables (u,v)(u,v). ∎

Finally, we can combine partitions and lines as follows, which gives a suitably generalised version of an individual step in the filtration from Example A.2 (like 𝒢2\mathscr{G}_{2} or 𝒢3\mathscr{G}_{3} in that example).

Proposition B.8.

Suppose that Y,ZY,Z are computable Polish spaces. Suppose that λ\lambda is a computable point of 𝒫(Y)\mathcal{P}(Y) and μ\mu is a computable point of 𝒫(Z)\mathcal{P}(Z).

Suppose {Ci:iI}\{C_{i}:i\in I\} is an effective partition of ZZ with effective softening {Vi:iI}\{V_{i}:i\in I\}.

Let \mathscr{F} be the λμ\lambda\otimes\mu-effective σ\sigma-algebra on Y×ZY\times Z generated by sets of the form U×CiU\times C_{i}, where UU ranges over c.e. opens from a λ\lambda-computable basis on YY. Then the map ρ:Y×Z+(Y×Z)\rho:Y\times Z\rightarrow\mathcal{M}^{+}(Y\times Z) given by ρ(u,v)=iI+(δuμ(Vi))IVi(v)\rho_{(u,v)}=\sum_{i\in I^{+}}\big{(}\delta_{u}\otimes\mu(\cdot\mid V_{i})\big{)}\cdot I_{V_{i}}(v) is a Martin-Löf disintegration of \mathscr{F} with respect to λμ\lambda\otimes\mu, and one has the following expression for the version of conditional expectation:

𝔼λμ[f](u,v)=iI+(1μ(Vi)Vif(u,t)dμ(t))IVi(v)\mathbb{E}_{\lambda\otimes\mu}[f\mid\mathscr{F}](u,v)=\sum_{i\in I^{+}}\bigg{(}\frac{1}{\mu(V_{i})}\int_{V_{i}}f(u,t)\;d\mu(t)\bigg{)}\cdot I_{V_{i}}(v) (B.5)
Proof.

This proof is just a combination of the proofs of Proposition B.6 and Proposition B.7. ∎

Another variant on Example A.2 is

Proposition B.9.

Let X=XiX=\prod X_{i}, where XiX_{i} is a computable sequence of computable Polish spaces. Let a computable point ν\nu of 𝒫(X)\mathcal{P}(X) be given by ν=iνi\nu=\bigotimes_{i}\nu_{i}, where νi\nu_{i} is a computable sequence in 𝒫(Xi)\mathcal{P}(X_{i}). Let n1n\geq 1. Let n\mathscr{F}_{n} be the σ\sigma-algebra on XX generated by sets of the form inVi×i>nXi\prod_{i\leq n}V_{i}\times\prod_{i>n}X_{i}, where ViV_{i} for ini\leq n ranges over c.e. opens from a νi\nu_{i}-computable basis on XiX_{i}. For xx in XX, write its coordinates as x=(x1,x2,)x=(x_{1},x_{2},\ldots). Then ρ(n):X𝒫(X)\rho^{(n)}:X\rightarrow\mathcal{P}(X) given by ρx(n)=(inδxi)(i>nνi)\rho_{x}^{(n)}=(\otimes_{i\leq n}\delta_{x_{i}})\otimes(\otimes_{i>n}\nu_{i}) is a Martin-Löf disintegration of n\mathscr{F}_{n} with respect to ν\nu, and one has the following expression for the version of conditional expectation, which for each ff in 𝕃1(ν)\mathbb{L}_{1}(\nu) is defined for ν\nu-a.s. many xx from XX, and where x=(x1,x2,)x=(x_{1},x_{2},\ldots) and t=(tn+1,tn+2,)t=(t_{n+1},t_{n+2},\ldots)

𝔼ν[fn](x)=i>nXif(x1,,xn,t¯)d(i>nνi)(t¯)\mathbb{E}_{\nu}[f\mid\mathscr{F}_{n}](x)=\int_{\prod_{i>n}X_{i}}f(x_{1},\ldots,x_{n},\overline{t})\;d(\otimes_{i>n}\nu_{i})(\overline{t})
Proof.

Simply apply Proposition B.7 to Y×ZY\times Z, where Y=inXiY=\prod_{i\leq n}X_{i} and Z=i>nXiZ=\prod_{i>n}X_{i} and λ=inνi\lambda=\otimes_{i\leq n}\nu_{i} and μ=i>nνi\mu=\otimes_{i>n}\nu_{i}. ∎

The simplest kind of an effective filtration, which occurs in both Examples B.1-B.2 is the following:

Definition B.10.

Suppose that XX is a computable Polish space. Suppose that T<T\subseteq\mathbb{N}^{<\mathbb{N}} is a computable tree with no dead ends. Let In={σT:|σ|=n}I_{n}=\{\sigma\in T:\left|\sigma\right|=n\}. Suppose that n\mathscr{F}_{n} is an effective partition {Aσ:σIn}\{A_{\sigma}:\sigma\in I_{n}\}, uniformly in n0n\geq 0. If the partitions refine one another, in that Aσ=σ(j)TAσ(j)A_{\sigma}=\bigcup_{\sigma^{\frown}(j)\in T}A_{\sigma^{\frown}(j)} for all σ\sigma in TT, then the n\mathscr{F}_{n} is an effective filtration, which we call an effective refined partition.

The following isolates the natural sufficient condition for an effective refined partition to be full, and this condition is obviously met in Example B.1:

Proposition B.11.

Suppose that the effective refined partition n\mathscr{F}_{n} satisfies the following properties:

  • Effectively Shrinking: There is a computable function :>0\ell:\mathbb{Q}^{>0}\rightarrow\mathbb{N} such that for all rational ϵ>0\epsilon>0 one has that diam(Aσ)<ϵ\mathrm{diam}(A_{\sigma})<\epsilon for all σ\sigma in TT with |σ|(ϵ)\left|\sigma\right|\geq\ell(\epsilon).

  • Effectively non-empty: There is a uniformly computable sequence of points xσx_{\sigma} in AσA_{\sigma}.

Then the effective refined partition n\mathscr{F}_{n} is full.

Proof.

(Sketch) The two conditions imply that there is a well-defined computable continuous surjection π:[T]X\pi:[T]\rightarrow X given by π(ω)=x\pi(\omega)=x iff {x}=nAωn\{x\}=\bigcap_{n}A_{\omega\upharpoonright n}. For fullness, suppose that UXU\subseteq X is c.e. open. Then π1(U)\pi^{-1}(U) is c.e. open. Then there is c.e. set STS\subseteq T such that π1(U)=σS[T][σ]\pi^{-1}(U)=\bigcup_{\sigma\in S}[T]\cap[\sigma]. Then one can check that U=σSAσU=\bigcup_{\sigma\in S}A_{\sigma}. ∎

We can also obtain full effective filtrations by combining lines with full effective partitions, as in B.2:

Example B.12.

Suppose that Y,ZY,Z are computable Polish spaces. Suppose that λ\lambda is a computable point of 𝒫(Y)\mathcal{P}(Y) and suppose that μ\mu is a computable point of 𝒫(Z)\mathcal{P}(Z).

Suppose that n\mathscr{F}_{n} is a full effective partition of ZZ (resp. almost-full effective partition of ZZ).

Let 𝒢n\mathscr{G}_{n} be the effective σ\sigma-algebra {U×A:UY c.e. open &An}\{U\times A:U\subseteq Y\mbox{ c.e. open }\;\&\;A\in\mathscr{F}_{n}\}.

Then 𝒢n\mathscr{G}_{n} is a full effective filtration of Y×ZY\times Z (resp. almost-full effective filtration of Y×ZY\times Z).

Another example of a full effective filtration is related to countable products:

Example B.13.

The effective σ\sigma-algebras n\mathscr{F}_{n} from the countable products Example B.9 is a full effective filtration. This is because every c.e. open WX=iXiW\subseteq X=\prod_{i}X_{i} can be uniformly written as σJWσ\bigcup_{\sigma\in J}W_{\sigma}, for some c.e. index set JJ, where Wσ=i<|σ|Vσ(i)×i|σ|XiW_{\sigma}=\prod_{i<\left|\sigma\right|}V_{\sigma(i)}\times\prod_{i\geq\left|\sigma\right|}X_{i}, where Vσ(i)V_{\sigma(i)} is uniformly c.e. open in XiX_{i} for ini\leq n.

Of course, this example is the same as that of full effective partitions when XiX_{i} is uniformly countable. However, this example goes beyond that of full effective partitions when the XiX_{i} are uncountable.

References

  • [1] Nathanael L Ackerman, Cameron E Freer, and Daniel M Roy, On computability and disintegration, Mathematical Structures in Computer Science 27 (2017), no. 8, 1287–1314.
  • [2] Gordon Belot, Bayesian orgulity, Philosophy of Science 80 (2013), no. 4, 483–503.
  • [3] Laurent Bienvenu, Rupert Hölzl, Joseph S Miller, and André Nies, Denjoy, Demuth and density, Journal of Mathematical Logic 14 (2014), no. 01, 1450004.
  • [4] Patrick Billingsley, Convergence of probability measures, Wiley, 2013.
  • [5] David Blackwell and Lester Dubins, Merging of opinions with increasing information, Annals of Mathematical Statistics 33 (1962), no. 3, 882–886.
  • [6] Vladimir I Bogachev, Weak convergence of measures, American Mathematical Society, 2018.
  • [7] Vasco Brattka, Joseph S Miller, and André Nies, Randomness and differentiability, Transactions of the American Mathematical Society 368 (2015), no. 1, 581–605.
  • [8] Gemma Carotenuto and André Nies, Lightface Π03{\Pi}^{0}_{3}-completeness of density sets under effective wadge reducibility, Pursuit of the Universal, Springer, 2016, pp. 234–239.
  • [9] Douglas Cenzer, Π01\Pi^{0}_{1}-Classes in computability theory, Handbook of Computability Theory (Edward R Griffor, ed.), Studies in Logic and the Foundations of Mathematics, vol. 140, North-Holland, Amsterdam, 1999, pp. 37–85.
  • [10] Douglas Cenzer and Jeffrey B Remmel, Effectively closed sets, May 2017.
  • [11] J T Chang and D Pollard, Conditioning as disintegration, Statistica Neerlandica 51 (1997), no. 3, 287–317.
  • [12] Vaughn Climenhaga and Anatole Katok, Measure theory through dynamical eyes, (2012), https://arxiv.org/abs/1208.4550.
  • [13] Natasha L Dobrinen and Stephen G Simpson, Almost everywhere domination, Journal of Symbolic Logic 69 (2004), no. 3, 914–922.
  • [14] J L Doob, Measure theory, Graduate Texts in Mathematics, vol. 143, Springer, 1994.
  • [15] Rod Downey and Dennis Hirschfeldt, Algorithmic randomness and complexity, Springer, Berlin, 2010.
  • [16] R M Dudley, Real analysis and probability, Cambridge University Press, 2002.
  • [17] Rick Durrett, Probability: Theory and examples, fourth ed., Cambridge University Press, 2010.
  • [18] John Earman, Bayes or bust? A critical examination of Bayesian confirmation theory, MIT Press, Cambridge, 1992.
  • [19] Manfred Einsiedler and Thomas Ward, Ergodic theory: with a view towards number theory, Graduate Texts in Mathematics, vol. 259, Springer, 2010.
  • [20] Jean-Pierre Florens, Velayoudom Marimoutou, and Anne Peguin-Feissolle, Econometric modeling and inference, Cambridge University Press, 2007.
  • [21] Gerald B. Folland, Real analysis, second ed., Pure and Applied Mathematics, Wiley, 1999.
  • [22] Peter Gács, Uniform test of algorithmic randomness over a general space, Theoretical Computer Science 341 (2005), no. 1, 91–137.
  • [23] Su Gao, Invariant descriptive set theory, CRC Press, 2008.
  • [24] Vassilios Gregoriades, Tamás Kispéter, and Arno Pauly, A comparison of concepts from computable analysis and effective descriptive set theory, Mathematical Structures in Computer Science 27 (2017), no. 8, 1414–1436.
  • [25] Allan Gut, Probability: A graduate course, Springer, 2013.
  • [26] Matthew Harrison-Trainor, Alexander Melnikov, and Keng Meng Ng, Computability of Polish spaces up to homeomorphism, Journal of Symbolic Logic 85 (2020), no. 4, 1664–1686.
  • [27] Peter G Hinman, A survey of Mučnik and Medvedev degrees, Bulletin of Symbolic Logic 18 (2012), no. 2, 161–229.
  • [28] Colin Howson and Urbach Peter, Scientific Reasoning: The Bayesian Approach, third ed., Open Court, Chicago, 2006.
  • [29] Mathieu Hoyrup, Calculabilité, aléatoire et théorie ergodique sur les espaces métriques, Ph.D. thesis, Université Paris-Diderot - Paris VII, 2008.
  • [30] Mathieu Hoyrup and Cristóbal Rojas, Computability of probability measures and Martin-Löf randomness over metric spaces, Information and Computation 207 (2009), no. 7, 830–847.
  • [31] Mathieu Hoyrup, Cristóbal Rojas, and Klaus Weihrauch, Computability of the Radon-Nikodym derivative, Models of Computation in Context (Benedikt Löwe, Dag Normann, Ivan Soskov, and Alexandra Soskova, eds.), Lecture Notes in Computer Science, vol. 6735, Springer, Berlin, Heidelberg, 2011, pp. 132–141.
  • [32] Mathieu Hoyrup and Jason Rute, Computable measure theory and algorithmic randomness, Handbook of Computability and Complexity in Analysis (Vasco Brattka and Peter Hertling, eds.), Springer, 2021, pp. 227–270.
  • [33] Olav Kallenberg, Foundations of modern probability, Springer, New York, NY, 2002.
  • [34] Alexander S. Kechris, Classical descriptive set theory, Graduate Texts in Mathematics, vol. 156, Springer, 1995.
  • [35] Mushfeq Khan, Lebgesgue density and Π10{\Pi}_{1}^{0} Classes, Journal of Symbolic Logic 81 (2016), no. 1, 80–95.
  • [36] Stuart Alan Kurtz, Randomness and genericity in the degrees of unsolvability, Ph.D. thesis, University of Illinois at Urbana-Champaign, 1981.
  • [37] Leonid Anatolievich Levin, Uniform tests of randomness, Soviet Mathematics Doklady 17 (1976), no. 2, 337–340.
  • [38] P Lévy, Théorie de l’addition des variables aléatoires, Gauthier-Villars, Paris (1938).
  • [39] Ming Li and Paul Vitányi, An introduction to Kolmogorov complexity and its applications, second ed., Graduate Texts in Computer Science, Springer, 1997.
  • [40] Donald Martin, Measure, category, and degrees of unsolvability, Unpublished (1967).
  • [41] Per Martin-Löf, The definition of random sequences, Information and Control 9 (1966), 602–619.
  • [42] by same author, On the notion of randomness, Intuitionism and Proof Theory (Akiko Kino, John Myhill, and Richard E Vesley, eds.), North-Holland, Amsterdam, 1970, pp. 73–78.
  • [43] Joseph S Miller, Degrees of unsolvability of continuous functions, Journal of Symbolic Logic 69 (2004), no. 2, 555–584.
  • [44] Kenshi Miyabe, Characterization of kurtz randomness by a differentiation theorem, Theory of Computing Systems 52 (2013), no. 1, 113–132.
  • [45] by same author, L1{L}^{1}-Computability, layerwise computability, and Solovay reducibility, Computability 2 (2013), 15–29.
  • [46] by same author, Algorithmic randomness over general spaces, Mathematical Logic Quarterly (2014).
  • [47] Kenshi Miyabe, André Nies, and Jing Zhang, Using almost-everywhere theorems from analysis to study randomness, Bulletin of Symbolic Logic 22 (2016), no. 3, 305–331.
  • [48] Yiannis N Moschovakis, Descriptive set theory, second ed., Mathematical Surveys and Monographs, vol. 155, American Mathematical Society, Providence, RI, 2009.
  • [49] Keng Meng Ng, Frank Stephan, Yue Yang, and Liang Yu, Computational aspects of the hyperimmune-free degrees, Proceedings of the 12th Asian Logic Conference, World Scientific, 2012, pp. 271–284.
  • [50] André Nies, Computability and randomness, Oxford University Press, 2008.
  • [51] K R Parthasarathy, Probability measures on metric spaces, Academic Press, 1967.
  • [52] Noopur Pathak, Cristóbal Rojas, and Stephen G. Simpson, Schnorr randomness and the Lebesgue differentiation theorem, Proceedings of the American Mathematical Society 142 (2014), no. 1, 335–349.
  • [53] David Pollard, A user’s guide to measure theoretic probability, Cambridge University Press, 2002.
  • [54] Christopher P Porter, On analogues of the Church–Turing Thesis in algorithmic randomness, Review of Symbolic Logic 9 (2016), no. 3, 456–479.
  • [55] Malempati M Rao, Stochastic processes - inference theory, second ed., Springer, 2014.
  • [56] Jan Reimann, Randomness—beyond Lebesgue measure, Logic Colloquium 2006 (S Barry Cooper, Herman Geuvers, Anand Pillay, and Jouko Vaananen, eds.), Cambridge University Press, Cambridge, 2010, pp. 247–279.
  • [57] Robert Rettinger and Xizhong Zheng, Computability of real numbers, Handbook of Computability and Complexity in Analysis (Vasco Brattka and Peter Hertling, eds.), Springer, 2021, pp. 3–28.
  • [58] V A Rohlin, On the fundamental ideas of measure theory, Matematicheskii Sbornik (1949), Translated in [59].
  • [59] by same author, On the fundamental ideas of measure theory, Functional analysis and measure theory, American Mathematical Society, 1962, pp. 1–54.
  • [60] V A Rokhlin, Lectures on the entropy of measure-preserving transformations, Russian Mathematical Surveys 22 (1967), no. 5.
  • [61] Jason Rute, Algorithmic randomness, martingales, and differentiability I, preprint (2012), This is an expanded version of [62, Chapter 2].
  • [62] by same author, Topics in algorithmic randomness and computable analysis, Ph.D. thesis, Carnegie Mellon University, 2013.
  • [63] Filippo Santambrogio, Optimal transport for applied mathematicians: Calculus of variations, PDEs, and modeling, Birkhauser, 2015.
  • [64] Claus-Peter Schnorr, Zufälligkeit und Wahrscheinlichkeit. Eine algorithmische Begründung der Wahrscheinlichkeitstheorie, Lecture Notes in Mathematics, vol. 218, Springer, 1971.
  • [65] by same author, A survey of the theory of random sequences, Basic Problems in Methodology and Linguistics: Part Three of the Proceedings of the Fifth International Congress of Logic, Methodology and Philosophy of Science, London, Ontario, Canada-1975, Reidel, Dordrecht, 1977, pp. 193–211.
  • [66] Glenn Shafer and Vladimir Vovk, Game-Theoretic foundations for probability and finance, Wiley, 2019.
  • [67] Glenn Shafer, Vladimir Vovk, and Akimichi Takemura, Levy’s zero-one law in game-theoretic probability, (2009), https://arxiv.org/abs/0905.0254.
  • [68] A Shen, V A Uspensky, and N Vereshchagin, Kolmogorov complexity and algorithmic randomness, American Mathematical Society, 2017.
  • [69] Stephen G Simpson, Mass problems and randomness, Bulletin of Symbolic Logic 11 (2005), no. 1, 1–27.
  • [70] Stephen G. Simpson, Subsystems of second order arithmetic, second ed., Cambridge University Press, Cambridge, 2009.
  • [71] Robert I Soare, Turing computability: Theory and applications, Springer, 2016.
  • [72] Alan M Turing, On computable numbers, with an application to the Entscheidungsproblem, Proceedings of the London Mathematical Society 2 (1937), no. 1, 230–265.
  • [73] Jan Von Plato, Creating modern probability: Its mathematics, physics and philosophy in historical perspective, Cambridge University Press, 1994.
  • [74] Hermann Weyl, Über die Gleichverteilung von Zahlen Mod. Eins, Mathematische Annalen 77 (1916), 313–352.
  • [75] David Williams, Probability with martingales, Cambridge Mathematical Textbooks, Cambridge University Press, Cambridge, 1991.
  • [76] Xiaokang Yu, Measure theory in weak subsystems of second-order arithmetic, Ph.D. thesis, The Pennsylvania State University, 1987.
  • [77] by same author, Radon-Nikodým theorem is equivalent to arithmetical comprehension, Logic and computation (Pittsburgh, PA, 1987), Contemporary Mathematics, vol. 106, American Mathematical Society, 1990, pp. 289–297.