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

Probability Trees

Diego A. Mejía Graduate School of System Informatics, Kobe University. 1-1 Rokkodai-cho, Nada-ku, Kobe, Hyogo 657-8501 Japan [email protected] http://www.researchgate.com/profile/Diego_Mejia2  and  Andrés F. Uribe-Zapata TU Wien, Faculty of Mathematics and Geoinformation, Institute of Discrete Mathematics and Geometry, Wiedner Hauptstrasse 8–10, A–1040 Vienna, Austria [email protected] https://sites.google.com/view/andres-uribe-afuz
Abstract.

In this article, we introduce a formal definition of the concept of probability tree and conduct a detailed and comprehensive study of its fundamental structural properties. In particular, we define what we term an inductive probability measure and prove that such trees can be identified with these measures. Furthermore, we prove that probability trees are completely determined by probability measures on the Borel σ\sigma-algebra of the tree’s body.

We then explore applications of probability trees in several areas of mathematics, including probability theory, measure theory, and set theory. In the first, we show that the cumulative distribution of finitely many dependent and non-identically distributed Bernouli tests is bounded by the cumulative distribution of some binomial distribution. In the second, we establish a close relationship between probability trees and the real line, showing that Borel, measurable sets, and their measures can be preserved, as well as other combinatorial properties. Finally, in set theory, we establish that the null ideal associated with suitable probability trees is Tukey equivalent to the null ideal on [0,1][0,1]. This leads to a new elementary proof of the fact that the null ideal of a free σ\sigma-finite Borel measure on a Polish space is Tukey equivalent with the null ideal of \mathbb{R}, which supports that the associated cardinal characteristics remain invariant across the spaces in which they are defined.

Key words and phrases:
probability tree, measure theory, cardinal invariants, real numbers, random variables, inductive probability measure.
2020 Mathematics Subject Classification:
60A99, 60G50, 60A05, 60B05, 60C05, 03E17, 60B05
The first author was supported by the Grant-in-Aid for Scientific Research (C) 23K03198, Japan Society for the Promotion of Science. The second author was supported by the Austrian Science Fund (FWF): project number P33895.

1. Introduction

In probability theory and statistics, the so-called probability trees are graphic tools that allow the representation of problems and help in their understanding and subsequent study, particularly those problems involving some kind of dependency. These trees can be used to visually analyze all the possible results of an experiment and the probabilities associated with each of them, making it easier to analyze situations involving decisions and dependent events and calculate their respective probabilities.

Intuitively, a probability tree starts with a single node —called the root of the tree111However, in some contexts, such as in the study of genealogy, unrooted probability trees are also considered (see e.g. [GT95]). —which represents the first event in the experiment. In the next step, from the root, branches of the tree fan out to represent all possible outcomes of the first experiment, labeled with their respective probabilities whose sum must be equal to one since they cover all possible outcomes. Each branch ends in a new node, from which new nodes and new branches are generated as the experiment is carried out, which ultimately determines the tree structure, covering all possible events in the experiment under consideration. The probabilities of each branch can be multiplied along the paths to calculate cumulative probabilities —that is, the probability that two or more events occur simultaneously— allowing for the complete determination of any sequence of events in the experiment.

In some contexts beyond probability and statistics, probability trees are relevant in many different areas of the natural sciences, such as in the field of genetics, where they are used to help model the inheritance of biological traits as well as diseases. Specifically, they can be used to determine the probability that an individual inherits a certain gene from his or her parents, connecting information between recessive and dominant inheritance patterns (see [FF80]). In an area of science directly related to genetics, namely genealogy, probability trees have other very significant applications. There they can be used to model and analyze situations of relationships between generations, as well as to calculate the probabilities of transmission of certain biological traits, similar to the case of genetics mentioned above. In this context, the trees in question help to represent family connections and possible routes of genetic inheritance, especially when dealing with dependent events, such as calculating the probability that a certain individual inherits a specific gene or disease from their ancestors. A well-known example of this is the so-called infinitely-many-sites models, which are used to study genetic variability and mutation processes in populations over time (see [Gri89]).

As we mentioned at the beginning of this introduction, probability trees are especially useful in the analysis of dependent events, that is, those in which the probability of an event is conditioned by the results of previous events. In these cases, the probability in the successive branches reflects such dependence, making these trees an essential tool for representing problems in the context of Bayesian Inference (see e.g. [ZM18]). This makes this type of tree very useful in some sub-areas of computer science such as artificial intelligence and machine learning, where they are used in particular to study probabilistic and learning algorithms, such as decision trees, which allow to classify data and make predictions of future outcomes based on probabilities derived from previously known data. In particular, probability trees serve as a model for a certain type of probabilistic process, known as Causal Generative Processes, which are essential in artificial intelligence as they allow modeling data generation based on causal relationships (see [GMD+20] and [KWWC24]). Furthermore, probability trees are widely used to carry out simulations, including the well-known Monte Carlo methods, where they are used in the representation and subsequent analysis of random processes (see [FSNG23]).

While we have seen that probability trees arise in highly practical and diverse contexts such as genetics, genealogy, artificial intelligence, machine learning, and Bayesian inference, the origin of this paper stems from a far more abstract and unexpected field: Forcing Theory, in Set Theory. Recently, based on previous work by Saharon Shelah [She00] and Jakob Kellner, Saharon Shelah, and Anda Tănasie [KST19], Miguel A. Cardona, along with the authors in [CMU24], introduced a general theory of iterated forcing using finitely additive measures222A preliminary version of the general theory of iterated forcing using finitely additive measures was presented in the master’s thesis of the second author (see [Uri23]), where the first questions related to probability trees arose. In this thesis, an entire chapter was dedicated to the formalization and study of these trees (see [Uri23, Ch. 3]), which served as the starting point for the development of this article.. This theory is founded on what the authors referred to as μ\mu-𝒴\mathcal{Y}-linkedness, a property of partially ordered sets, where μ\mu is an infinite cardinal and 𝒴\mathcal{Y} is a class of ordered pairs. A central focus of the theory was to prove that random forcing —the poset (2ω)𝒩(2ω)\mathcal{B}({}^{\omega}2)\smallsetminus\mathcal{N}({}^{\omega}2) ordered by inclusion— satisfies this property for certain suitable parameters μ0\mu_{0} and 𝒴0\mathcal{Y}_{0}. Later, the authors in [MU24] extended this result, a task that was not only highly technical but also required the development and application of concepts related to probability trees (as in the original work of Shelah [She00]). The core of this proof was to prove the existence of two objects: a finite set uu and an element of random forcing rr^{\oplus} satisfying certain special conditions. Without delving into overly technical and unnecessarily complex details, the strategy was based on defining a probability tree whose nodes represented partial approximations of the set uu. Then, instead of directly attempting to find the objects uu and rr^{\oplus}, following the approach of the probabilistic method (see [AS16] and [Uri23, pp. 100100-101101]), the probability of their existence was calculated using the tree structure. Finally, it was shown that this probability is positive, ensuring that such objects satisfying the required conditions can indeed be found. To achieve this, it was necessary not only to formalize the notion of a probability tree but also to develop and analyze the structural properties of such trees.

In the references cited so far, there is no concrete definition of the notion of probability tree, as it is usually tailored to the needs of each particular case. Moreover, there is currently no detailed study of the structure of these trees. For this reason, the main objective of this article is clear and specific: to formalize the notion of probability tree and analyze its structure rigorously. Our formalization of this concept —presented in 4.11— is quite intuitive and, in general terms, outlined in the second paragraph of this introduction.

The formal definition and structural study of probability trees are carried out mainly in Section 4. Our starting point is to prove that every probability tree induces a measure on the tree that generates probability space structure at each of its fronts and levels (see Theorem 4.10 and 4.15). The analysis of the converse of this result motivated the introduction of the notion of inductive probability measure, which enabled a deeper exploration of the structure of theses trees. This led to the prove that probability trees can be identified with inductive probability measures and are completely determined by probability measures on the Borel σ\sigma-algebra of [T][T] —the body of TT, which is the set of maximal branches (see Section 3). The structural study developed in this article focuses primarily on the definition of four collections associated with these trees: tree probability sequences (𝒯𝒫\mathcal{TP}), inductive probability measures (𝒫\mathcal{IP}), Borel probability measures (𝒫\mathcal{BP}), and general probability sequences (𝒢𝒫\mathcal{GP}), as well as the connections between them, which are ultimately reduced to the commutativity of the diagram presented in Figure 3.

Once the structure of probability trees was analyzed, we laid the groundwork for exploring their applications in different areas. This structural study not only allowed us to formalize and better understand their properties but also to establish connections with other branches of mathematics. In particular, we apply probability trees to three distinct areas: probability theory, measure theory, and set theory, particularly in the combinatorics of real numbers and invariant cardinals. Below, we provide a brief description of each of these applications.

The first application —in Probability Theory—is related to a generalization of a well-known result: by adding a finite number of independent and identically distributed random variables with a Bernoulli distribution, we obtain a random variable with a binomial distribution. However, we face the situation where these random variables with Bernoulli distribution are dependent and even not identically distributed. To address this problem, we use probability trees to achieve the following result, which corresponds to Theorem 5.1.

Theorem A.

Let p[0,1]p\in[0,1], nn be a natural number, and YY be the random variable representing the number of successes of nn-many dependent Bernoulli distributed random variables, where the probability of success of each variable also depends on the previous events. If pp is a lower bound of the probability of success of each Bernoulli-distributed random variable, then the cumulative distribution of YY is below the cumulative distribution of the binomial distribution with parameters nn and pp.

This situation arose in the proof of [CMU24, Main Lemma 7.17] (see also [Uri23, Main Lemma 4.3.17]), where A proved sufficient for the purposes required in that context.

The second application —in Measure Theory— establishes a connection between probability trees and the real line. In particular, we will prove the following theorem, which corresponds to Section 6.

Theorem B.

Every probability tree T,μ¯\langle T,\bar{\mu}\rangle defines a canonical probability measure λμ¯\lambda^{\bar{\mu}} on the Borel σ\sigma-algebra ([T])\mathcal{B}([T]) of [T][T] which has a connection with the Lebesgue measure of the unit interval.

The measure λμ¯\lambda^{\bar{\mu}} is defined through a function gμ¯g_{\bar{\mu}} defined on [0,1][0,1], expect on a countable subset, constructed from the probability tree T,μ¯\langle T,\bar{\mu}\rangle. The connection between T,μ¯\langle T,\bar{\mu}\rangle and [0,1][0,1] is established through gμ¯g_{\bar{\mu}} as, when restricted to an appropriate set, turns out to be a homeomorphism, thus preserving the Borel sets and, consequently, the measurable sets (see Theorem 6.19 and 6.23). Additionally, this function preserves the measure between the measurable spaces [T],([T]),λμ¯\langle[T],\mathcal{B}([T]),\lambda^{\bar{\mu}}\rangle and [0,1],([0,1]),𝖫𝖻\langle[0,1],\mathcal{B}([0,1]),\operatorname{\mathsf{Lb}}\rangle, where 𝖫𝖻\operatorname{\mathsf{Lb}} denotes the Lebesgue measure. On the other hand, the properties that define λμ¯\lambda^{\bar{\mu}} will also allow us to demonstrate that probability trees can characterize the probability measures defined on ([T])\mathcal{B}([T]) (see Theorem 4.37).

The third application —in Set Theory— is related to cardinal invariants. Cardinal invariants, also called cardinal characteristics, are cardinal numbers that capture combinatorial properties of infinite spaces. Examples of this kind of cardinals arise considering ideals. Recall that for a non-empty set XX, 𝒫(X)\mathcal{I}\subseteq\mathcal{P}(X) is an ideal on XX, if it is closed under finite unions, \subseteq-downwards closed, \emptyset\in\mathcal{I}, and XX\notin\mathcal{I}. In this context, we define the cardinals invariants associated with \mathcal{I} as follows:

add()\operatorname{add}(\mathcal{I}) min{||:}\coloneqq\min\{|\mathcal{F}|\colon\mathcal{F}\subseteq\mathcal{I}\wedge\bigcup\mathcal{F}\notin\mathcal{I}\} is the additivity of \mathcal{I},
cov()\operatorname{cov}(\mathcal{I}) min{||:=X}\coloneqq\min\{|\mathcal{F}|\colon\mathcal{F}\subseteq\mathcal{I}\wedge\bigcup\mathcal{F}=X\} is the covering of \mathcal{I},
non()\operatorname{non}(\mathcal{I}) min{|Y|:YXY}\coloneqq\min\{|Y|\colon Y\subseteq X\wedge Y\notin\mathcal{I}\} is the uniformity of \mathcal{I},
cof()\operatorname{cof}(\mathcal{I}) min{||:AB(AB)}\coloneqq\min\{|\mathcal{F}|\colon\mathcal{F}\subseteq\mathcal{I}\wedge\forall A\in\mathcal{I}\exists B\in\mathcal{F}\ (A\subseteq B)\} is the cofinality of \mathcal{I}.

Apparently, there is no unanimous reason why they are called invariants, but it is known that they possess an invariance property: in many cases, the associated cardinal characteristics do not depend on the space on which the ideal is defined, as long as the space satisfies certain properties (see Theorem 7.7). The connection between probability trees and the real line that we establish in Section 6 allows us to extend this invariance property to the null ideal of λμ¯\lambda^{\bar{\mu}}. If X,𝒜,μ\langle X,\mathcal{A},\mu\rangle is a measure space, 𝒩(μ)\mathcal{N}(\mu) denotes the ideal of all μ\mu-measure zero subsets of XX. When the measure space is understood, we just write 𝒩(X)\mathcal{N}(X), e.g. 𝒩()\mathcal{N}(\mathbb{R}) and 𝒩([0,1])\mathcal{N}([0,1]) with respect to the Lebesgue measure.

Theorem C.

If T,μ¯\langle T,\bar{\mu}\rangle is a probability tree such that λμ¯\lambda^{\bar{\mu}} is free, the cardinal invariants associated with 𝒩(λμ¯)\mathcal{N}(\lambda^{\bar{\mu}}) and 𝒩([0,1])\mathcal{N}([0,1]) are the same.

Moreover, these identities follow by the Tukey-equivalence (7.3) between structures associated with these ideals.

This theorem leads to a new elementary proof of the more general known fact that the invariance of the cardinal invariants associated with the null ideal applies for any free σ\sigma-finite measure μ\mu on the Borel σ\sigma-algebra of a Polish space. Details are developed in Section 7.

In this work, we adopt the set-theoretic treatment of natural numbers.

Notation 1.1.

We adopt the set-theoretic treatment of natural numbers (starting from 0): 0=0=\emptyset, and each natural number is the set of its predecessors. Formally, if nn is a natural number, then n={0,1,2,,n1}n=\{0,1,2,\dots,n-1\}. The entire set of natural numbers is denoted by ω\omega. Furthermore, as in the context of ordinal numbers in set theory, we typically write “n<ωn<\omega” instead of “nωn\in\omega” and “αω\alpha\leq\omega” for “α<ω\alpha<\omega or α=ω\alpha=\omega”.333ω\omega is the first infinite ordinal number, represented as the limit of natural numbers.

2. Elementary notions of measure and probability theory

2.1. Review of measure theory

We review, in this section, basic notions related to measure theory and probability spaces.

Let XX be a non-empty set and 𝒞𝒫(X)\mathcal{C}\subseteq\mathcal{P}(X). The σ\sigma-algebra generated by 𝒞\mathcal{C}, that is, the smallest σ\sigma-algebra of sets over XX that contains 𝒞\mathcal{C}, is denoted by σX(𝒞)\sigma_{X}(\mathcal{C}). If ZXZ\subseteq X, then 𝒞|Z{CZ:C𝒞}\mathcal{C}|_{Z}\coloneqq\{C\cap Z\colon C\in\mathcal{C}\}, which is a (σ\sigma-)algebra of sets over ZZ whenever 𝒞\mathcal{C} is a (σ\sigma-)algebra of sets over XX. Recall that, if f:XYf\colon X\to Y is a function between non-empty sets, 𝒜\mathcal{A} a σ\sigma-algebra over XX and 𝒜\mathcal{A}^{\prime} is a σ\sigma-algebra over YY, then ff is 𝒜\mathcal{A}-𝒜\mathcal{A}^{\prime}-measurable if f1[B]𝒜f^{-1}[B]\in\mathcal{A} for all B𝒜B\in\mathcal{A}^{\prime}. Most of the time, we work with the Borel σ\sigma-algebra of a topological space: given a topological space XX, (X)\mathcal{B}(X) denotes the σ\sigma-algebra generated by the collection of open subsets of XX, which is known as the Borel σ\sigma-algebra of XX. Any set B(X)B\in\mathcal{B}(X) is called Borel in XX, and a function f:XYf\colon X\to Y between topological spaces is called a Borel map if it is (X)\mathcal{B}(X)-(Y)\mathcal{B}(Y)-measurable.

Regarding functions. Let f:XYf\colon X\to Y be a function between non-empty sets. For 𝒞𝒫(X)\mathcal{C}\subseteq\operatorname{\mathcal{P}}(X) and 𝒞𝒫(Y)\mathcal{C}^{\prime}\subseteq\operatorname{\mathcal{P}}(Y), define f(𝒞){f1[B]:B𝒞}f^{\leftarrow}(\mathcal{C}^{\prime})\coloneqq\{f^{-1}[B]\colon B\in\mathcal{C}^{\prime}\} and f(𝒞){BY:f1[B]𝒞}.f^{\to}(\mathcal{C})\coloneqq\{B\subseteq Y\colon f^{-1}[B]\in\mathcal{C}\}. Functions can be used to transfer σ\sigma-algebras.

Fact 2.1.

Let f:XYf\colon X\to Y be a function between non-empty sets.

  1. (a)

    If 𝒜\mathcal{A}^{\prime} is a (σ\sigma-)algebra over YY then f(𝒜)f^{\leftarrow}(\mathcal{A}^{\prime}) is a (σ\sigma-)algebra over XX.

  2. (b)

    If 𝒜\mathcal{A} is a (σ\sigma-)algebra over XX, then f(𝒜)f^{\to}(\mathcal{A}) is a (σ\sigma-)algebra over YY.

  3. (c)

    If 𝒞𝒫(Y)\mathcal{C}^{\prime}\subseteq\operatorname{\mathcal{P}}(Y) then σX(f(𝒞))=f(σY(𝒞))\sigma_{X}(f^{\leftarrow}(\mathcal{C}^{\prime}))=f^{\leftarrow}(\sigma_{Y}(\mathcal{C}^{\prime})).

As a consequence, we have the following results.

Corollary 2.2.

Let XX be a non-empty set, 𝒞𝒫(X)\mathcal{C}\subseteq\operatorname{\mathcal{P}}(X) and ZXZ\subseteq X. Then we have that σZ(𝒞|Z)=σX(𝒞)|Z\sigma_{Z}(\mathcal{C}|_{Z})=\sigma_{X}(\mathcal{C})|_{Z}.

Corollary 2.3.

If XX is a topological space and ZXZ\subseteq X is a subspace, then we have that (Z)=(X)|Z\mathcal{B}(Z)=\mathcal{B}(X)|_{Z}.

Corollary 2.4.

Any continuous function between topological spaces is Borel.

We now review the notion of measure. Let XX be a non-empty set and 𝒞𝒫(X)\mathcal{C}\subseteq\operatorname{\mathcal{P}}(X) with 𝒞\emptyset\in\mathcal{C}. Recall that a function μ:𝒞[0,]\mu\colon\mathcal{C}\to[0,\infty] is a finitely additive measure (fam), if μ()=0\mu(\emptyset)=0 and μ(k<nAk)=k<nμ(Ak)\mu\left(\bigcup_{k<n}A_{k}\right)=\sum_{k<n}\mu(A_{k}) whenever {Ak:k<n}𝒞\{A_{k}\colon k<n\}\subseteq\mathcal{C} is a finite collection of pairwise disjoint sets whose union is in 𝒞\mathcal{C}. Also, a fam μ\mu is a (σ\sigma-additive) measure if μ(n<ωAn)=n<ωμ(An)\mu\left(\bigcup_{n<\omega}A_{n}\right)=\sum_{n<\omega}\mu(A_{n}) for any collection {An:n<ω}𝒞\{A_{n}\colon n<\omega\}\subseteq\mathcal{C} of pairwise disjoint sets whose union is in 𝒞\mathcal{C}.

Given a fam μ:𝒞[0,]\mu\colon\mathcal{C}\to[0,\infty], we say that μ\mu is finite if X𝒞X\in\mathcal{C} and μ(X)<\mu(X)<\infty. When there is some {An:n<ω}𝒞\{A_{n}\colon n<\omega\}\subseteq\mathcal{C} such that X=n<ωAnX=\bigcup_{n<\omega}A_{n} and μ(An)<\mu(A_{n})<\infty for all n<ωn<\omega, μ\mu is called σ\sigma-finite. If X𝒞X\in\mathcal{C} and μ(X)=1\mu(X)=1, then μ\mu is a probability fam. Finally, μ\mu is free if, for any xXx\in X, {x}𝒞\{x\}\in\mathcal{C} and μ({x})=0\mu(\{x\})=0.

Recall that a measure space is a triple X,𝒜,μ\langle X,\mathcal{A},\mu\rangle where 𝒜\mathcal{A} is a σ\sigma-algebra over XX and μ\mu is a measure on 𝒜\mathcal{A}.

Example 2.5.

Let WW be a countable set. For any function f:W[0,]f\colon W\to[0,\infty] there is a unique measure μf:𝒫(W)[0,]\mu^{f}\colon\operatorname{\mathcal{P}}(W)\to[0,\infty] such that μf({k})=f(k)\mu^{f}(\{k\})=f(k) for all kWk\in W. Indeed, for AWA\subseteq W, μf(A)\mu^{f}(A) must be kAf(k)\sum_{k\in A}f(k).

Conversely, for any measure μ:𝒫(W)[0,]\mu\colon\operatorname{\mathcal{P}}(W)\to[0,\infty] there is a unique function f:W[0,]f\colon W\to[0,\infty] such that μ=μf\mu=\mu^{f} (f(k)f(k) must be μ({k})\mu(\{k\})).

As a consequence, if WW is a countable or finite set, to define a probability measure on 𝒫(W)\mathcal{P}(W) it is sufficient to define a function f:W[0,1]f\colon W\to[0,1] such that wWf(w)=1\sum_{w\in W}f(w)=1. For example, we can use this to introduce the uniform measure on finite sets.

Definition 2.6.

Let XX be a non-empty finite set. The uniform measure on XX, denoted by μX\mu_{X}, is the measure on 𝒫(X)\mathcal{P}(X) determined by μX({x})1|X|\mu_{X}(\{x\})\coloneqq\frac{1}{|X|}.

Definition 2.7.

The standard measure on ω\omega is the (unique) probability measure on 𝒫(ω)\operatorname{\mathcal{P}}(\omega) obtained from the function ω[0,1]\omega\to[0,1], k2(k+1)k\mapsto 2^{-(k+1)} as in 2.5.

Measure zero sets play an essential role in measure theory. Let 𝒜𝒫(X)\mathcal{A}\subseteq\operatorname{\mathcal{P}}(X) be a σ\sigma-algebra over the set XX and let μ:𝒜[0,]\mu\colon\mathcal{A}\to[0,\infty] be a measure. Define

𝒩(μ)\displaystyle\mathcal{N}(\mu) {NX:A𝒜:NA and μ(A)=0},\displaystyle\coloneqq\{N\subseteq X\colon\exists\,A\in\mathcal{A}\colon N\subseteq A\text{ and }\mu(A)=0\},
𝒜μ\displaystyle\mathcal{A}^{\mu} {AN:A𝒜,N𝒩(μ)}.\displaystyle\coloneqq\{A\cup N\colon A\in\mathcal{A},\ N\in\mathcal{N}(\mu)\}.

The sets in 𝒩(μ)\mathcal{N}(\mu) are called μ\mu-null, or just null, or measure zero sets.

It is known that 𝒜μ\mathcal{A}^{\mu} is the σ\sigma-algebra generated by 𝒜𝒩\mathcal{A}\cup\mathcal{N} and that there is a unique measure on 𝒜μ\mathcal{A}^{\mu} that extends μ\mu, namely, ANμ(A)A\cup N\mapsto\mu(A) for A𝒜A\in\mathcal{A} and N𝒩N\in\mathcal{N}. This measure is called the completion of μ\mu, and we use the same letter μ\mu to denote this completion. In terms of measure spaces, we say that X,𝒜μ,μ\langle X,\mathcal{A}^{\mu},\mu\rangle is the completion of X,𝒜,μ\langle X,\mathcal{A},\mu\rangle.

Notice that 𝒜μ=𝒜\mathcal{A}^{\mu}=\mathcal{A} iff 𝒩(μ)𝒜\mathcal{N}(\mu)\subseteq\mathcal{A}. In this situation, we say that μ\mu is a complete measure and X,𝒜,μ\langle X,\mathcal{A},\mu\rangle is a complete measure space.

Lemma 2.8.

Let f:XYf\colon X\to Y be a map between non-empty sets. If X,𝒜,μ\langle X,\mathcal{A},\mu\rangle is a measure space, then Y,f(𝒜),μ\langle Y,f^{\to}(\mathcal{A}),\mu^{\prime}\rangle is a measure space where μ(B)μ(f1[B])\mu^{\prime}(B)\coloneqq\mu(f^{-1}[B]) for Bf(𝒜)B\in f^{\to}(\mathcal{A}).

Lemma 2.9.

Let 𝒜\mathcal{A} be an algebra on a set XX and let μ\mu be a fam on 𝒜\mathcal{A}. Assume that μ(X)<\mu(X)<\infty and {x}𝒜\{x\}\in\mathcal{A} for all xXx\in X. Then the set {xX:μ({x})>0}\{x\in X\colon\mu(\{x\})>0\} is countable.

Finally, recall:

Theorem 2.10 ([Hal50, §13]).

Let \mathcal{F} be an algebra of sets over a set XX and assume that μ:[0,]\mu\colon\mathcal{F}\to[0,\infty] is a σ\sigma-finite measure. Then, there is a unique measure on σX()\sigma_{X}(\mathcal{F}) that extends μ\mu.

3. Elementary notions of trees

3.1. Trees

In this section, we introduce the set-theoretic notion of tree (of height ω{\leq}\omega) and study some of its combinatorial and topological properties.

We start by fixing some notation about functions and sequences.

Notation 3.1.

We write ai:iI\langle a_{i}\colon i\in I\rangle to denote a function such that iaii\mapsto a_{i}. We usually say that ai:iI\langle a_{i}:\,i\in I\rangle is a sequence (indexed by II).

We typically look at sequences of length ω{\leq}\,\omega: for n<ωn<\omega, ai:i<n=a0,,an1\langle a_{i}\colon i<n\rangle=\langle a_{0},\ldots,a_{n-1}\rangle is a finite sequence of length nn, while ai:i<ω=a0,a1,\langle a_{i}\colon i<\omega\rangle=\langle a_{0},a_{1},\ldots\rangle is a sequence of length ω\omega. The empty sequence \langle\ \rangle is the only sequence of length 0. We use |s||s| to denote the length of a sequence ss of length ω{\leq}\,\omega. When ss is a finite sequence and tt is a sequence of length ω{\leq}\,\omega, define the concatenation of ss and tt by sts0,,s|s|1,t0,t1,s{}^{\frown}t\coloneqq\langle s_{0},\ldots,s_{|s|-1},t_{0},t_{1},\ldots\rangle. If ss and tt are sequences of length ω{\leq}\,\omega, then st iff |s||t| and i<|s|(si=ti),s\subseteq t\text{ iff }|s|\leq|t|\text{ and }\forall\ i<|s|\ (s_{i}=t_{i}), that is, sts\subseteq t means that tt is a longer sequence extending ss.

Let WW be a set and αω\alpha\leq\omega. We define Wα{}^{\alpha}W as the set of sequences in WW of length α\alpha; W<α{}^{<\alpha}W as the set of sequences in WW of length <α{<}\,\alpha; and Wα{}^{\leq\alpha}W as the set of sequences in WW of length α{\leq}\,\alpha. Equivalently, W<αk<αWk{}^{<\alpha}W\coloneqq\bigcup_{k<\alpha}{}^{k}W and WαkαWk{}^{\leq\alpha}W\coloneqq\bigcup_{k\leq\alpha}{}^{k}W.

Recall the notions of partial and linear orders: a partial order is a pair P,\langle P,\leq\rangle where PP is a non-empty set and \leq is a reflexive, transitive, and anti-symmetric relation. We say that xx and yy are comparable in PP if either xyx\leq y or yxy\leq x. A chain in PP is a subset of PP such that any pair of elements are comparable. When PP is a chain in itself, we say that P,\langle P,\leq\rangle is a linear order.

We are ready to introduce the notion of tree.

Definition 3.2.

A tree (of height ω{\leq}\,\omega) is a partial order T,<T\langle T,<_{T}\rangle containing a minimum element 𝗋𝗍(T)\operatorname{\mathsf{rt}}(T), called the root of TT, such that, for any tTt\in T, t{sT:s<tt}t{\downarrow}\coloneqq\{s\in T\colon s<_{t}t\} is a finite linear order (under the order of TT). The members of TT are usually called the nodes of TT.

For example, when WW is a non-empty set, W<ω,\langle{}^{<\omega}W,\subseteq\rangle is a tree with 𝗋𝗍(W<ω)=\operatorname{\mathsf{rt}}({}^{<\omega}W)=\langle\ \rangle.

Now, we introduce some notation related to trees and their properties.

Definition 3.3.

Given a tree T,\langle T,\leq\rangle, we fix the following notation for tTt\in T and n<ωn<\omega:

  1. (1)

    htT(t)|t|\operatorname{\mathrm{ht}}_{T}(t)\coloneqq|t{\downarrow}| the height of tt in TT.

  2. (2)

    𝖫𝗏n(T){tT:htT(t)=n}\operatorname{\mathsf{Lv}}_{n}(T)\coloneqq\{t\in T\colon\operatorname{\mathrm{ht}}_{T}(t)=n\} the nn-th level of TT, so 𝖫𝗏0(T)={𝗋𝗍(T)}\operatorname{\mathsf{Lv}}_{0}(T)=\{\operatorname{\mathsf{rt}}(T)\}.

  3. (3)

    ht(T)min({n<ω:𝖫𝗏n(T)=}{ω})\operatorname{\mathrm{ht}}(T)\coloneqq\min(\{n<\omega\colon\operatorname{\mathsf{Lv}}_{n}(T)=\emptyset\}\cup\{\omega\}) the height of TT.

  4. (4)

    𝗌𝗎𝖼𝖼T(t){tT:tt and htT(t)=htT(t)+1}\operatorname{\mathsf{succ}}_{T}(t)\coloneqq\{t^{\prime}\in T\colon t\leq t^{\prime}\text{ and }\operatorname{\mathrm{ht}}_{T}(t^{\prime})=\operatorname{\mathrm{ht}}_{T}(t)+1\} the set of immediate successors of tt (in TT).

  5. (5)

    𝗌𝗉𝗍(T){tT:|𝗌𝗎𝖼𝖼T(t)|2}\operatorname{\mathsf{spt}}(T)\coloneqq\{t\in T\colon|\operatorname{\mathsf{succ}}_{T}(t)|\geq 2\} the set of splitting nodes of TT.

  6. (6)

    Tt{sT:ts}T_{\geq t}\coloneqq\{s\in T\colon t\subseteq s\} is the set of successors of tt in TT.

  7. (7)

    max(T){sT:𝗌𝗎𝖼𝖼s(T)=},\max(T)\coloneqq\{s\in T\colon\operatorname{\mathsf{succ}}_{s}(T)=\emptyset\}, that is, the set of maximal nodes of TT.

Example 3.4.

Consider the tree W<ω,\langle{}^{<\omega}W,\subseteq\rangle. Then, for tW<ωt\in{}^{<\omega}W and n<ωn<\omega:

  1. (1)

    ht(t)=|t|=|t|\operatorname{\mathrm{ht}}(t)=|t\downarrow|=|t|, the length of tt.

  2. (2)

    𝖫𝗏n(W<ω)=Wn\operatorname{\mathsf{Lv}}_{n}({}^{<\omega}W)={}^{n}W.

  3. (3)

    ht(W<ω)=ω\operatorname{\mathrm{ht}}({}^{<\omega}W)=\omega.

  4. (4)

    𝗌𝗎𝖼𝖼(t)={t:W}\operatorname{\mathsf{succ}}(t)=\{t{}^{\frown}\langle\ell\rangle\colon\ell\in W\}.

We introduce the following notions related to trees.

Definition 3.5.

Let T,\langle T,\leq\rangle be a tree.

  1. (1)

    We say that TT^{\prime} is a subtree of TT if TTT^{\prime}\subseteq T and, for any tTt\in T^{\prime} and sts\leq t in TT, sTs\in T^{\prime}.

  2. (2)

    A tree TT is well-pruned if, for any tTt\in T and ht(t)<n<ht(T)\operatorname{\mathrm{ht}}(t)<n<\operatorname{\mathrm{ht}}(T), there is some t𝖫𝗏n(T)t^{\prime}\in\operatorname{\mathsf{Lv}}_{n}(T) above tt.

  3. (3)

    A tree TT is finitely branching if 𝗌𝗎𝖼𝖼T(t)\operatorname{\mathsf{succ}}_{T}(t) is finite for all tTt\in T.

  4. (4)

    A tree TT is perfect if, for every tTt\in T, there is some splitting node in TT above tt.

Notice that, if ht(T)<ω\mathrm{ht}(T)<\omega then max(T)=𝖫𝗏ht(T)(T)\max(T)=\operatorname{\mathsf{Lv}}_{\mathrm{ht}(T)}(T) whenever TT is a well-pruned tree.

On the other hand, it is not hard to check that, if TT is a tree and TT^{\prime} is a subtree of TT then, for any tTt\in T^{\prime} and n<ωn<\omega, 𝗋𝗍(T)=𝗋𝗍(T)\operatorname{\mathsf{rt}}(T^{\prime})=\operatorname{\mathsf{rt}}(T), htT(t)=htT(t)\operatorname{\mathrm{ht}}_{T^{\prime}}(t)=\operatorname{\mathrm{ht}}_{T}(t), 𝖫𝗏n(T)=T𝖫𝗏n(T)\operatorname{\mathsf{Lv}}_{n}(T^{\prime})=T^{\prime}\cap\operatorname{\mathsf{Lv}}_{n}(T), ht(T)ht(T)\operatorname{\mathrm{ht}}(T^{\prime})\leq\operatorname{\mathrm{ht}}(T), and 𝗌𝗎𝖼𝖼T(t)=T𝗌𝗎𝖼𝖼T(t)\operatorname{\mathsf{succ}}_{T^{\prime}}(t)=T^{\prime}\cap\operatorname{\mathsf{succ}}_{T}(t).

Example 3.6.

If TT is a subtree of W<ω{}^{<\omega}W then 𝗋𝗍(T)=\operatorname{\mathsf{rt}}(T)=\langle\ \rangle and 𝖫𝗏n(T)=TWn\operatorname{\mathsf{Lv}}_{n}(T)=T\cap{}^{n}W for n<ωn<\omega. For any AWA\subseteq W and 0<αω0<\alpha\leq\omega, A<α{}^{<\alpha}A is a subtree of W<ω{}^{<\omega}W of height α\alpha. Note that A<α{}^{<\alpha}A is well-pruned, it is finitely branching iff AA is finite, and it is perfect iff α=ω\alpha=\omega and |A|2|A|\geq 2.

Theorem 3.7.

Any tree of height ω{\leq}\,\omega is isomorphic with a subtree of W<ω{}^{<\omega}W for some non-empty set WW.

Proof..

Let TT be a tree. Put WTW\coloneqq T (as a set). Define f:TW<ωf\colon T\to{}^{<\omega}W in the following way: for tTt\in T, enumerate t{t}={ti:ihtT(t)}t\cup\{t\}=\{t_{i}\colon i\leq\operatorname{\mathrm{ht}}_{T}(t)\} such that i<jti<tji<j\Rightarrow t_{i}<t_{j} (so tht(t)=tt_{\operatorname{\mathrm{ht}}(t)}=t), so set f(t)ti+1:i<htT(t)f(t)\coloneqq\langle t_{i+1}\colon i<\operatorname{\mathrm{ht}}_{T}(t)\rangle. Then, for any s,tTs,t\in T, st iff f(s)f(t)s\leq t\text{ iff }f(s)\subseteq f(t) and f[T]f[T] is a subtree of T<ω{}^{<\omega}T. ∎

Notation 3.8.

Due to the previous theorem, from now on all trees are trees of sequences, i.e. subtrees of W<ω{}^{<\omega}W for some non-empty set WW, unless otherwise stated.

From the next sections, the space of infinite branches of a tree will be very important to understand the combinatorics and topology of the reals.

Definition 3.9.

When TT is a subtree of W<ω{}^{<\omega}W, we define

limT\displaystyle\lim T {xWω:xnT for all n<ω};\displaystyle\coloneqq\{x\in{}^{\omega}W\colon x{\upharpoonright}n\in T\text{ for all $n<\omega$}\};
[T]\displaystyle[T] limTmaxT.\displaystyle\coloneqq\lim T\cup\max T.

Notice that [T]=maxT[T]=\max T when TT has finite height (because, in this case, limT=\lim T=\emptyset). Since there is a bijection between [T][T] and the maximal chains contained in TT, we call [T][T] the space of maximal branches of the tree or the body of TT. Any xlimTx\in\lim T is called an infinite branch of TT.

It is clear that [W<ω]=limW<ω=Wω[{}^{<\omega}W]=\lim{}^{<\omega}W={}^{\omega}W.

If TT is a subtree of W<ω{}^{<\omega}W, then limT\lim T\neq\emptyset implies ht(T)=ω\operatorname{\mathrm{ht}}(T)=\omega. However, the converse is not always true, since there could be trees of height ω\omega with limT=\lim T=\emptyset. For example, for n<ωn<\omega let tnn0¯nt_{n}\coloneqq\langle n\rangle{}^{\frown}\bar{0}^{n} where 0¯n\bar{0}^{n} is the sequence of length nn composed with only 0. Then S{sω<ω:n<ω:stn}S\coloneqq\{s\in{}^{<\omega}\omega\colon\exists\,n<\omega\colon s\subseteq t_{n}\} is a counterexample. The so-called König’s Theorem (see Theorem 3.10) gives us the sufficient conditions to have the equivalence.

Theorem 3.10.

Let TT be a subtree of W<ω{}^{<\omega}W. If TT is finitely-branching, then the following statements are equivalent.

  1. (i)

    limT\lim T\neq\emptyset.

  2. (ii)

    ht(T)=ω\operatorname{\mathrm{ht}}(T)=\omega.

  3. (iii)

    TT is infinite.

The set limT\lim T is also related to the notion of well-foundedness.

Definition 3.11.

A tree TT is well-founded if every non-empty subset of TT has a maximal element.

Lemma 3.12.

Let TT be a tree of sequences. Then, the following statements are equivalent.

  1. (i)

    TT is well-founded.

  2. (ii)

    limT=\lim T=\emptyset.

Proof..

If limT\lim T\neq\emptyset and contains some infinite branch xx, then {xn:n<ω}\{x{\upharpoonright}n\colon n<\omega\} is a subset of TT without maximal elements. Conversely, if TT is not well-founded, i.e. there is some non-empty ATA\subseteq T without maximal elements, then we can construct an increasing sequence tn:n<ω\langle t_{n}\colon n<\omega\rangle in AA. This increasing sequence determines a unique member of limT\lim T, so limT\lim T\neq\emptyset. ∎

3.2. Tree-topology

In this Subsection, we assign a topology to [T][T] for any tree TT and study some of its properties. We start by recalling some basic topological notions.

Consider a topological space X,τ\langle X,\tau\rangle. A set CXC\subseteq X is clopen in X,τ\langle X,\tau\rangle if it is open and closed in X,τ\langle X,\tau\rangle. We say that X,τ\langle X,\tau\rangle is zero-dimensional if it has a base of clopen sets. For AXA\subseteq X, cX(A)\operatorname{c{\ell}}_{X}(A) denotes the closure of AA, and intX(A)\mathrm{int}_{X}(A) is the (topological) interior of AA. The subindex is removed when clear from the context. Recall that a topological space XX is discrete if every subset of XX is open.

If 𝒮𝒫(X)\mathcal{S}\subseteq\operatorname{\mathcal{P}}(X), the smallest topology of XX containing 𝒮\mathcal{S}, which is called the topology of XX generated by 𝒮\mathcal{S}, is denoted by τ𝒮\tau_{\mathcal{S}}.

We define the topology of the branches of a tree.

Definition 3.13.

Let TT be a tree. The tree-topology of [T][T] is the topology generated by {[t]T:tT}\{[t]_{T}\colon t\in T\}, where [t]T{x[T]:tx}[t]_{T}\coloneqq\{x\in[T]\colon t\subseteq x\}. We just write [t][t] when TT is clear from the context. Denote T([T])\mathcal{B}_{T}\coloneqq\mathcal{B}([T]).

Notice that, if TT is a tree of finite height, then the tree topology is the discrete topology. More generally, maxT\max T (in case it is non-empty) is a discrete subspace.

Assume that TT is a tree of height ω{\leq}\,\omega. We say that s,tTs,t\in T are compatible (in TT) if either sts\subseteq t or tst\subseteq s. Otherwise, they are incompatible, which we represent by sTts\perp_{T}t, or just sTs\perp T. It is not hard to check the following.

Fact 3.14.

Let TT be a tree and s,tTs,t\in T. Then:

  1. (a)

    If sts\subseteq t then [t][s][t]\subseteq[s].

  2. (b)

    sts\perp t iff [s][t]=[s]\cap[t]=\emptyset.

  3. (c)

    [t][s][t]\subseteq[s] iff either sts\subseteq t, or tst\subseteq s and there are no splitting nodes between tt, including it, and ss, excluding it. (The latter implies [t]=[s][t]=[s].)

Lemma 3.15.

The collection {[t]T:tT}\{[t]_{T}\colon t\in T\} is a base of the topology of [T][T] and each [t]T[t]_{T} is clopen in [T][T]. In particular, [T][T] is a zero-dimensional space.

The countable product of discrete spaces can be expressed as a topological space of the form [T][T].

Definition 3.16.

Let b=b(n):n<ωb=\langle b(n)\colon n<\omega\rangle be a sequence of non-empty sets. Define seq<ωbn<ωk<nb(k)\operatorname{\mathrm{seq}_{<\omega}}b\coloneqq\bigcup_{n<\omega}\prod_{k<n}b(k), which is a well-pruned tree of height ω\omega and 𝖫𝗏n(seq<ωb)=k<nb(k)\operatorname{\mathsf{Lv}}_{n}(\operatorname{\mathrm{seq}_{<\omega}}b)=\prod_{k<n}b(k) for all n<ωn<\omega. Notice that,

[seq<ωb]=bn<ωb(n)={x:x is a function, domx=ω,n<ω:x(n)b(n)}.[\operatorname{\mathrm{seq}_{<\omega}}b]=\prod b\coloneqq\prod_{n<\omega}b(n)=\{x\colon x\text{ is a function, }\operatorname{dom}x=\omega,\ \forall\,n<\omega\colon x(n)\in b(n)\}.

In the case that b(n)=Wb(n)=W for all n<ωn<\omega, seq<ωb=W<ω\operatorname{\mathrm{seq}_{<\omega}}b={}^{<\omega}W and b=Wω\prod b={}^{\omega}W.

Two very important spaces are defined in this way: the Cantor Space 2ω={0,1}ω,{}^{\omega}2={}^{\omega}\{0,1\}, and the Baire space ωω{}^{\omega}\omega.

Lemma 3.17.

Let b=b(n):n<ωb=\langle b(n)\colon n<\omega\rangle be a sequence of discrete spaces. Then the tree-topology of b\prod b is the same as the product topology.

Proof..

It is enough to show that {[t]:tseq<ωb}\{[t]\colon t\in\operatorname{\mathrm{seq}_{<\omega}}b\} is a base of the product topology. First, for [t]seq<ωb[t]\in\operatorname{\mathrm{seq}_{<\omega}}b, [t][t] is open in the product topology because [t]=i<ωai[t]=\prod_{i<\omega}a_{i} where

ai{{ti} if i<|t|,b(i) if i|t|.a_{i}\coloneqq\begin{cases}\{t_{i}\}&\text{ if $i<|t|$,}\\ b(i)&\text{ if $i\geq|t|$.}\end{cases}

Now assume that AA is open in the product topology and xAx\in A. Then, there is some sequence ci:i<ω\langle c_{i}\colon i<\omega\rangle such that each cib(i)c_{i}\subseteq b(i), {i<ω:cib(i)}\{i<\omega\colon c_{i}\neq b(i)\} is finite, and xi<ωciAx\in\prod_{i<\omega}c_{i}\subseteq A. Hence, there is some m<ωm<\omega such that {i<ω:cib(i)}m\{i<\omega\colon c_{i}\neq b(i)\}\subseteq m. Therefore, x[xm]i<ωciAx\in[x{\upharpoonright}m]\subseteq\prod_{i<\omega}c_{i}\subseteq A. ∎

The Cantor space is compact as a consequence of the following result, which follows by Theorem 3.10.

Theorem 3.18.

Let TT be a tree. Then [T][T] is compact iff TT is finitely branching.

4. Probability trees

In this section, we formalize the notion of a probability tree and explore some of its structural properties. This will allow us to prove that such trees can be identified with a specific class of sequences, which we call inductive probability measures, as well as with probability measures on the Borel σ\sigma-algebra on the space of maximal branches of the tree (see 4.114.28, and Theorem 4.37). Later, in Subsection 4.4, we will build on the concept of conditional probability to define a relative expected value within this framework.

Notation 4.1.

We denote by 𝒯\mathcal{T} the collection of all countable trees of sequences.

Remark 4.2.

Although, for simplicity, we develop all the theory in this section for countable trees of sequences, thanks to Theorem 3.7 it can be applied, in a natural way, to arbitrary countable trees, regardless of whether they are composed by sequences.

4.1. Trees as probability spaces

We show how to define probability spaces from a tree in the sense of Section 3.

Definition 4.3.

We say that T,μ¯\langle T,\bar{\mu}\rangle is a probability tree if T𝒯T\in\mathcal{T} and μ¯=μt:tTmax(T)\bar{\mu}=\langle\mu_{t}\colon t\in T\smallsetminus\max(T)\rangle, where each μt\mu_{t} is a probability measure on 𝒫(𝗌𝗎𝖼𝖼T(t))\operatorname{\mathcal{P}}(\operatorname{\mathsf{succ}}_{T}(t)).

Furthermore, we define 𝒯𝒫\mathcal{TP} as the class of all sequences μ¯\bar{\mu} such that T,μ¯\langle T,\bar{\mu}\rangle is a probability tree for some T𝒯T\in\mathcal{T}. Notice that this TT is uniquely determined by (the domain of) μ¯\bar{\mu}, so it will be denoted by Tμ¯T_{\bar{\mu}}.

The following are examples of probability trees.

Example 4.4.
  1. (1)

    T2<ωT\coloneqq{}^{<\omega}2 and, for any tTt\in T, μt\mu_{t} is the uniform measure on 𝗌𝗎𝖼𝖼T(t)\operatorname{\mathsf{succ}}_{T}(t), that is, μt({t0})=μt({t1})12\mu_{t}(\{t{}^{\frown}\langle 0\rangle\})=\mu_{t}(\{t{}^{\frown}\langle 1\rangle\})\coloneqq\frac{1}{2} (see 2.6).

  2. (2)

    Tω<ωT\coloneqq{}^{<\omega}\omega and, for tTt\in T, μt\mu_{t} resembles the standard measure on ω\omega, that is, μt({t})2(+1)\mu_{t}(\{t{}^{\frown}\langle\ell\rangle\})\coloneqq 2^{-(\ell+1)} for <ω\ell<\omega (see 2.7).

  3. (3)

    Tω<ωT\coloneqq{}^{<\omega}\omega and, for tTt\in T and <ω\ell<\omega,

    μt(t){1 if =5,0 if 5.\mu_{t}(t{}^{\frown}\langle\ell\rangle)\coloneqq\begin{cases}1&\text{ if $\ell=5$,}\\ 0&\text{ if $\ell\neq 5$.}\end{cases}

Order-isomorphisms preserve the probability tree structure.

Lemma 4.5.

Let T,μ¯\langle T,\bar{\mu}\rangle be a probability tree, SS a partial order, and φ:TS\varphi\colon T\to S an order-isomorphism. Then S,τ¯\langle S,\bar{\tau}\rangle is a probability tree, where τ¯=τs:sSmax(S)\bar{\tau}=\langle\tau_{s}\colon s\in S\smallsetminus\max(S)\rangle, and for any sSs\in S and r𝗌𝗎𝖼𝖼S(s)r\in\operatorname{\mathsf{succ}}_{S}(s), τs({r})μφ1(s)({φ1(r)})\tau_{s}(\{r\})\coloneqq\mu_{\varphi^{-1}(s)}(\{\varphi^{1-}(r)\}).

Let TT be the subtree of W<ω{}^{<\omega}W whose set of nodes is {wij:i<2j<4}\{w_{i}^{j}\colon i<2\wedge j<4\} (see Figure 1).

\bullet\bullet\bullet\bullet\bullet\bullet\bullet\langle\,\ranglew10w_{1}^{0}w11w_{1}^{1}w20w_{2}^{0}w21w_{2}^{1}w22w_{2}^{2}w23w_{2}^{3}p10p_{1}^{0}p11p_{1}^{1}p20p_{2}^{0}p21p_{2}^{1}p22p_{2}^{2}p23p_{2}^{3}
Figure 1. Example of a probability tree.

If we define μ(w0i)p0i,\mu_{\langle\,\rangle}(\langle w_{0}^{i}\rangle)\coloneqq p_{0}^{i}, for i{0,1}i\in\{0,1\}; μw10(w10,w2j)p2j,\mu_{\langle w_{1}^{0}\rangle}(\langle w_{1}^{0},\,w_{2}^{j}\rangle)\coloneqq p_{2}^{j}, for j{0,1}j\in\{0,1\}; and μw11(w11,w2j)p2j,\mu_{\langle w_{1}^{1}\rangle}(\langle w_{1}^{1},\,w_{2}^{j}\rangle)\coloneqq p_{2}^{j}, for j{2,3},j\in\{2,3\}, then TT is a probability tree if, and only if:

p10+p11=1,p20+p21=1 and p22+p23=1.p_{1}^{0}+p_{1}^{1}=1,\ p_{2}^{0}+p_{2}^{1}=1\ \text{ and }\ p_{2}^{2}+p_{2}^{3}=1.

Notice that, in that case, it satisfies the following:

p10p20+p10p21+p11p22+p11p23=p10(p20+p21)+p11(p22+p23)=p10+p11=1,p_{1}^{0}p_{2}^{0}+p_{1}^{0}p_{2}^{1}+p_{1}^{1}p_{2}^{2}+p_{1}^{1}p_{2}^{3}=p_{1}^{0}(p_{2}^{0}+p_{2}^{1})+p_{1}^{1}(p_{2}^{2}+p_{2}^{3})=p_{1}^{0}+p_{1}^{1}=1,

that is, if for any t=zjk,zlm𝖫𝗏2(T)t=\langle z_{j}^{k},z_{l}^{m}\rangle\in\operatorname{\mathsf{Lv}}_{2}(T) we define μ2T(t)pjkplm,\mu_{2}^{T}(t)\coloneqq p_{j}^{k}p_{l}^{m}, then we have that 𝖫𝗏2(T),𝒫(𝖫𝗏2(T)),μ2T\langle\operatorname{\mathsf{Lv}}_{2}(T),\mathcal{P}(\operatorname{\mathsf{Lv}}_{2}(T)),\mu_{2}^{T}\rangle is a probability space. The same happens trivially at 𝖫𝗏1(T)\operatorname{\mathsf{Lv}}_{1}(T).

As a consequence, it is possible to induce a probability space structure on each level of TT, and even a measure on the whole tree. To formalize this idea, we introduce the following definition.

Definition 4.6.

Let T,μ¯\langle T,\bar{\mu}\rangle be a probability tree. Define the measure Ξμ¯:𝒫(T)[0,1]\Xi^{\bar{\mu}}\colon\operatorname{\mathcal{P}}(T)\to[0,1] determined by

Ξμ¯({t})k<nμtk({t(k+1)}) for tT.\Xi^{\bar{\mu}}(\{t\})\coloneqq\prod_{k<n}\mu_{t{\upharpoonright}k}(\{t{\upharpoonright}(k+1)\})\text{ for $t\in T$.}

For any n<ht(T)n<\operatorname{\mathrm{ht}}(T), Ξnμ¯Ξμ¯𝒫(𝖫𝗏n(T))\Xi^{\bar{\mu}}_{n}\coloneqq\Xi^{\bar{\mu}}{\upharpoonright}\operatorname{\mathcal{P}}(\operatorname{\mathsf{Lv}}_{n}(T)), which is a measure on 𝒫(𝖫𝗏n(T))\operatorname{\mathcal{P}}(\operatorname{\mathsf{Lv}}_{n}(T)).

In some cases, the probability of a successor’s space in a probability tree can be determined using Ξμ¯\Xi^{\bar{\mu}}.

Lemma 4.7.

Let T,μ¯\langle T,\bar{\mu}\rangle be a probability tree, tTmax(T)t\in T\smallsetminus\max(T) and s𝗌𝗎𝖼𝖼T(t)s\in\operatorname{\mathsf{succ}}_{T}(t). Then Ξμ¯({s})=μt({s})Ξμ¯({t})\Xi^{\bar{\mu}}(\{s\})=\mu_{t}(\{s\})\Xi^{\bar{\mu}}(\{t\}). In particular, if Ξμ¯({t})>0\Xi^{\bar{\mu}}(\{t\})>0, then μt({s})=Ξμ¯({s})Ξμ¯({t}).\mu_{t}(\{s\})=\frac{\Xi^{\bar{\mu}}(\{s\})}{\Xi^{\bar{\mu}}(\{t\})}.

Proof..

Let tTmax(T)t\in T\smallsetminus\max(T) and s𝗌𝗎𝖼𝖼T(t)s\in\operatorname{\mathsf{succ}}_{T}(t). Then,

Ξμ¯({s})=k<|t|+1μs(k+1)({s(k+1)})=μs|t|({s(|t|+1)})k<|t|μsk({s(k+1)})=Ξμ¯({s|t|})μs|t|({s(|t|+1)})=Ξμ¯({t})μt({s}).\begin{split}\Xi^{\bar{\mu}}(\{s\})&=\prod_{k<|t|+1}\mu_{s{\restriction}(k+1)}(\{s{\restriction}(k+1)\})=\mu_{s{\restriction}|t|}(\{s{\restriction}(|t|+1)\})\prod_{k<|t|}\mu_{s{\restriction}k}(\{s{\restriction}(k+1)\})\\ &=\Xi^{\bar{\mu}}(\{s{\restriction}|t|\})\mu_{s{\restriction}|t|}(\{s{\restriction}(|t|+1)\})=\Xi^{\bar{\mu}}(\{t\})\mu_{t}(\{s\}).\end{split}

Thus, Ξμ¯({s})=Ξμ¯({t})μt({s}).\Xi^{\bar{\mu}}(\{s\})=\Xi^{\bar{\mu}}(\{t\})\mu_{t}(\{s\}).

The induced measure Ξμ¯\Xi^{\bar{\mu}} satisfies the following properties.

Lemma 4.8.

If T,μ¯\langle T,\bar{\mu}\rangle is a probability tree, then:

  1. (a)

    Ξμ¯({})=1\Xi^{\bar{\mu}}(\{\langle\,\rangle\})=1.

  2. (b)

    Ξμ¯({t})=Ξμ¯(𝗌𝗎𝖼𝖼T(t))\Xi^{\bar{\mu}}(\{t\})=\Xi^{\bar{\mu}}(\operatorname{\mathsf{succ}}_{T}(t)) for any tTmax(T)t\in T\smallsetminus\max(T).

Proof..

(a): Ξμ¯({})=i<0μti({t(i+1)})=1\displaystyle\Xi^{\bar{\mu}}(\{\langle\,\rangle\})=\prod_{i<0}\mu_{t{\restriction}i}(\{t{\restriction}(i+1)\})=1, where the last equality holds because empty products are equal to 1.

(b): By 4.7, we have that:

Ξμ¯(𝗌𝗎𝖼𝖼T(t))=s𝗌𝗎𝖼𝖼T(t)Ξμ¯({t})μt({s})=Ξμ¯({t})s𝗌𝗎𝖼𝖼T(t)μt({s})=Ξμ¯({t}).\begin{split}\Xi^{\bar{\mu}}(\operatorname{\mathsf{succ}}_{T}(t))&=\sum_{s\in\operatorname{\mathsf{succ}}_{T}(t)}\Xi^{\bar{\mu}}(\{t\})\mu_{t}(\{s\})=\Xi^{\bar{\mu}}(\{t\})\sum_{s\in\operatorname{\mathsf{succ}}_{T}(t)}\mu_{t}(\{s\})=\Xi^{\bar{\mu}}(\{t\}).\qquad\end{split}\qed

Considering the probability trees from 4.4, we can calculate the corresponding measures on its levels.

Example 4.9.
  1. (1)

    Let tTt\in T, where TT is as in 4.4 (1). Then,

    Ξμ¯({t})=i<|t|μti({t(i+1)})=i<|t|12=2|t|.\Xi^{\bar{\mu}}(\{t\})=\prod_{i<|t|}\mu_{t{\restriction}i}(\{t{\restriction}(i+1)\})=\prod_{i<|t|}\frac{1}{2}=2^{-|t|}.
  2. (2)

    For tTt\in T, where TT is as in 4.4 (2), Ξμ¯({t})=2|t|2i<|t|t(i)\displaystyle\Xi^{\bar{\mu}}(\{t\})=2^{-|t|}2^{-\sum_{i<|t|}t(i)}.

  3. (3)

    In the case of 4.4 (3), for n<ωn<\omega and tωnt\in{}^{n}\omega, Ξμ¯({t})=1\Xi^{\bar{\mu}}(\{t\})=1 if t(i)=5t(i)=5 for all i<ni<n, otherwise Ξμ¯({t})=0\Xi^{\bar{\mu}}(\{t\})=0.

Based on 4.6, we can prove that a well-pruned probability tree induces a probability space structure at each level.

Theorem 4.10.

If T,μ¯\langle T,\bar{\mu}\rangle is a well-pruned probability tree then, for any n<ht(T)n<\mathrm{ht}(T), Ξnμ¯\Xi^{\bar{\mu}}_{n} is a probability measure on 𝒫(𝖫𝗏n(T))\mathcal{P}(\operatorname{\mathsf{Lv}}_{n}(T)).

Proof..

We proceed by induction on n<ht(T)n<\mathrm{ht}(T). The case n=0n=0 follows by 4.8 (a) because 𝖫𝗏0(T)={}\operatorname{\mathsf{Lv}}_{0}(T)=\{\langle\,\rangle\}. Now, assume that n+1<ht(T)n+1<\operatorname{\mathrm{ht}}(T) and that 𝖫𝗏n(T),𝒫(𝖫𝗏n(T)),Ξnμ¯\langle\operatorname{\mathsf{Lv}}_{n}(T),\mathcal{P}(\operatorname{\mathsf{Lv}}_{n}(T)),\Xi_{n}^{\bar{\mu}}\rangle is a probability space. Since 𝖫𝗏n+1(T)=t𝖫𝗏n(T)𝗌𝗎𝖼𝖼T(t)\operatorname{\mathsf{Lv}}_{n+1}(T)=\bigcup_{t\in\operatorname{\mathsf{Lv}}_{n}(T)}\operatorname{\mathsf{succ}}_{T}(t) is a disjoint union and 𝖫𝗏n(T)Tmax(T)\operatorname{\mathsf{Lv}}_{n}(T)\subseteq T\smallsetminus\max(T) (the latter because TT is well-pruned and n+1<ht(T)n+1<\operatorname{\mathrm{ht}}(T)), by 4.8 (b) we get:

Ξμ¯(𝖫𝗏n+1(T))=t𝖫𝗏n(T)Ξμ¯(𝗌𝗎𝖼𝖼T(t))=tTΞμ¯({t})=Ξμ¯(𝖫𝗏n(T))=1.\Xi^{\bar{\mu}}(\operatorname{\mathsf{Lv}}_{n+1}(T))=\sum_{t\in\operatorname{\mathsf{Lv}}_{n}(T)}\Xi^{\bar{\mu}}(\operatorname{\mathsf{succ}}_{T}(t))=\sum_{t\in T}\Xi^{\bar{\mu}}(\{t\})=\Xi^{\bar{\mu}}(\operatorname{\mathsf{Lv}}_{n}(T))=1.

Thus, (𝖫𝗏n+1(T),𝒫(𝖫𝗏n+1(T)),Ξn+1μ¯)(\operatorname{\mathsf{Lv}}_{n+1}(T),\mathcal{P}(\operatorname{\mathsf{Lv}}_{n+1}(T)),\Xi_{n+1}^{\bar{\mu}}) is a probability space. ∎

Theorem 4.10 may not hold when TT is not well-pruned. We leave this discussion to the following subsection (4.16).

4.2. Inductive probability measures

The properties presented in 4.8 are fundamental for the analysis of the structure of probability trees. This motivates the following notion of inductive probability measure.

Definition 4.11.
  1. (1)

    Let T𝒯T\in\mathcal{T}. We say that Ξ\Xi is an inductive probability measure on TT if it is a measure on 𝒫(T)\operatorname{\mathcal{P}}(T) such that Ξ({})=1\Xi(\{\langle\,\rangle\})=1 and Ξ(𝗌𝗎𝖼𝖼T(t))=Ξ({t})\Xi(\operatorname{\mathsf{succ}}_{T}(t))=\Xi(\{t\}) whenever tTmax(T)t\in T\smallsetminus\max(T).

    For any n<ht(T)n<\operatorname{\mathrm{ht}}(T), ΞnΞ𝒫(𝖫𝗏n(T))\Xi_{n}\coloneqq\Xi{\upharpoonright}\operatorname{\mathcal{P}}(\operatorname{\mathsf{Lv}}_{n}(T)), which is a measure on 𝒫(𝖫𝗏n(T))\operatorname{\mathcal{P}}(\operatorname{\mathsf{Lv}}_{n}(T)).

  2. (2)

    Denote by 𝒫\mathcal{IP} the collection of inductive probability measures Ξ\Xi in some T𝒯T\in\mathcal{T}. Notice that TT is uniquely determined by Ξ\Xi, so it will be denoted by TΞT_{\Xi}.

  3. (3)

    Define the function Π:𝒯𝒫𝒫\Pi\colon\mathcal{TP}\to\mathcal{IP} such that, for any μ¯𝒯𝒫\bar{\mu}\in\mathcal{TP}, Π(μ¯)Ξμ¯\Pi(\bar{\mu})\coloneqq\Xi^{\bar{\mu}}, which is well-defined by virtue of 4.8. Notice that TΞμ¯=Tμ¯T_{\Xi^{\bar{\mu}}}=T_{\bar{\mu}}.

Notice that the proof of Theorem 4.10 only uses that Ξμ¯\Xi^{\bar{\mu}} is an inductive probability measure. For this reason, the same proof yields:

Theorem 4.12.

If Ξ\Xi is an inductive probability measure on a well-pruned tree TT, then Ξn\Xi_{n} is a probability measure on 𝒫(𝖫𝗏n(T))\operatorname{\mathcal{P}}(\operatorname{\mathsf{Lv}}_{n}(T)) for all n<ht(T)n<\mathrm{ht}(T).

As in the case of Theorem 4.10, the previous theorem may not hold when TT is not well-pruned. To understand this, we generalize this theorem by using the following notion.

Definition 4.13.

Let TT be a tree of sequences. A set ATA\subseteq T is a front of TT if it satisfies the following:

  1. (i)

    Any pair of members of AA are incompatible in TT.

  2. (ii)

    Every maximal branch of TT intersects AA, i.e. for any x[T]x\in[T] there is some n<ωn<\omega such that xnAx{\upharpoonright}n\in A.

For example, for any n<ht(T)n<\mathrm{ht}(T), Frn(T)𝖫𝗏n(T){smax(T):|s|<n}\mathrm{Fr}_{n}(T)\coloneqq\operatorname{\mathsf{Lv}}_{n}(T)\cup\{s\in\max(T)\colon|s|<n\} (which is a disjoint union) is a front of TT.

Fact 4.14.

Let TT be a tree of sequences.

  1. (a)

    TT is well-pruned iff, for any n<ht(T)n<\operatorname{\mathrm{ht}}(T), 𝖫𝗏n(T)\operatorname{\mathsf{Lv}}_{n}(T) is a front of TT.

  2. (b)

    Assume that TT is not well-pruned and let n0<ht(T)n_{0}<\operatorname{\mathrm{ht}}(T) be the minimum number such that 𝖫𝗏n0(T)max(T)\operatorname{\mathsf{Lv}}_{n_{0}}(T)\cap\max(T)\neq\emptyset. Then n0+1<ht(T)n_{0}+1<\operatorname{\mathrm{ht}}(T) and, for any n<ht(T)n<\operatorname{\mathrm{ht}}(T), 𝖫𝗏n(T)\operatorname{\mathsf{Lv}}_{n}(T) is a front of TT iff nn0n\leq n_{0}.

Proof..

Regardless of whether TT is well-pruned or not, we can find the maximum γht(T)\gamma\leq\operatorname{\mathrm{ht}}(T) (which can be ω\omega) satisfying that T|γ{tT:|t|<γ}T|_{\gamma}\coloneqq\{t\in T\colon|t|<\gamma\} is well-pruned. Notice that γ>0\gamma>0 and TT is well-pruned iff γ=ht(T)\gamma=\operatorname{\mathrm{ht}}(T). Also, if TT is not well-pruned, then n0+1=γ<ht(T)n_{0}+1=\gamma<\operatorname{\mathrm{ht}}(T).

Therefore, it is enough to show that, for any n<ht(T)n<\operatorname{\mathrm{ht}}(T), 𝖫𝗏n(T)\operatorname{\mathsf{Lv}}_{n}(T) is a front of TT iff n<γn<\gamma. If x[T]x\in[T] then its length must be γ{\geq}\,\gamma (otherwise T|γT|_{\gamma} would not be well-pruned), so xn𝖫𝗏n(T)x{\upharpoonright}n\in\operatorname{\mathsf{Lv}}_{n}(T) for any n<γn<\gamma. This shows the implication from right to left.

For the converse, assume that γn<ht(T)\gamma\leq n<\operatorname{\mathrm{ht}}(T). Then γ+1ht(T)\gamma+1\leq\operatorname{\mathrm{ht}}(T) and T|γ+1T|_{\gamma+1} is not well-pruned, so 𝖫𝗏n0(T)maxT\operatorname{\mathsf{Lv}}_{n_{0}}(T)\cap\max T\neq\emptyset. Any xx in this set has length n0<nn_{0}<n, so it does not have nodes of length nn below it. Thus, 𝖫𝗏n(T)\operatorname{\mathsf{Lv}}_{n}(T) is not a front. ∎

Theorem 4.12 is generalized as follows.

Theorem 4.15.

If Ξ𝒫\Xi\in\mathcal{IP} and AA is a front of TTΞT\coloneqq T_{\Xi}, then Ξ𝒫(A)\Xi{\upharpoonright}\operatorname{\mathcal{P}}(A) is a probability measure. Moreover,

(4.15.1) Ξ({t})=Ξ({sA:ts}) for any tT below some member of A.\Xi(\{t\})=\Xi(\{s\in A\colon t\subseteq s\})\text{ for any $t\in T$ below some member of $A$}.
Proof..

For tTt\in T, let At{sA:ts}A_{\geq t}\coloneqq\{s\in A\colon t\subseteq s\}. Since A=AA_{\geq\langle\,\rangle}=A, (4.15.1) and 4.8 (a) imply that Ξ(A)=Ξ({})=1\Xi(A)=\Xi(\{\langle\,\rangle\})=1. Hence, it is enough to show (4.15.1). We provide two proofs of (4.15.1); the first is presented below, and the second is in \autopagerefproofb035-3.

Let SS be the set of nodes in TT below some member of AA. This is a subtree of TT and limS=\lim S=\emptyset: if xlimSx\in\lim S then xlimTx\in\lim T, so xnAx{\upharpoonright}n\in A for some n<ωn<\omega because AA is a front, but all members of AA are pairwise incompatible, so xmSx{\upharpoonright}m\notin S for all mnm\geq n, contradicting that xlimSx\in\lim S.

Therefore, SS is a well-founded tree by 3.12. It is enough to show that S{tS:Ξ({t})Ξ(At)}=S^{\sim}\coloneqq\{t\in S\colon\Xi(\{t\})\neq\Xi(A_{\geq t})\}=\emptyset. Assume the contrary, so SS^{\sim} contains a maximal element tt. If tAt\in A then At={t}A_{\geq t}=\{t\}, so Ξ({t})=Ξ(At)\Xi(\{t\})=\Xi(A_{\geq t}), that is tSt\notin S^{\sim}. Thus tAt\notin A. This implies that 𝗌𝗎𝖼𝖼T(t)S\operatorname{\mathsf{succ}}_{T}(t)\subseteq S and, since tt is maximal in SS^{\sim}, 𝗌𝗎𝖼𝖼T(t)S=\operatorname{\mathsf{succ}}_{T}(t)\cap S^{\sim}=\emptyset, so Ξ({t})=Ξ(At)\Xi(\{t^{\prime}\})=\Xi(A_{\geq t^{\prime}}) for all t𝗌𝗎𝖼𝖼T(t)t^{\prime}\in\operatorname{\mathsf{succ}}_{T}(t). On the other hand, At=t𝗌𝗎𝖼𝖼T(t)AtA_{\geq t}=\bigcup_{t^{\prime}\in\operatorname{\mathsf{succ}}_{T}(t)}A_{\geq t^{\prime}} is a disjoint union, so

Ξ(At)=t𝗌𝗎𝖼𝖼T(t)Ξ(At)=t𝗌𝗎𝖼𝖼T(t)Ξ({t})=Ξ(𝗌𝗎𝖼𝖼T(t))=Ξ({t}),\Xi(A_{\geq t})=\sum_{t^{\prime}\in\operatorname{\mathsf{succ}}_{T}(t)}\Xi(A_{\geq t^{\prime}})=\sum_{t^{\prime}\in\operatorname{\mathsf{succ}}_{T}(t)}\Xi(\{t^{\prime}\})=\Xi(\operatorname{\mathsf{succ}}_{T}(t))=\Xi(\{t\}),

where the last equality holds by 4.8 (b). But this contradicts that tSt\in S^{\sim}. ∎

Notice that Theorem 4.12 (and Theorem 4.10) follow from 4.14 and Theorem 4.15.

Example 4.16.

Assume that Ξ𝒫\Xi\in\mathcal{IP} and that TTΞT\coloneqq T_{\Xi} is not well-pruned. Let n0n_{0} be as in 4.14 (b). For any n0<n<ht(T)n_{0}<n<\mathrm{ht}(T), Frn(T)\mathrm{Fr}_{n}(T) is a front of TT, so Ξ𝒫(Ln)\Xi{\upharpoonright}\operatorname{\mathcal{P}}(L_{n}) is a probability measure. Therefore, a necessary condition for Ξ(𝖫𝗏n(T))=1\Xi(\operatorname{\mathsf{Lv}}_{n}(T))=1 (i.e. Ξn\Xi_{n} is a probability measure) is that Ξ({t})=0\Xi(\{t\})=0 for all tmax(T)t\in\max(T) with |t|<n|t|<n. The latter condition does not always hold.

Given an inductive probability measure in TT, we can compute the probability of TtT_{\geq t} in each level of TT in terms of the probability of tt.

Corollary 4.17.

If T𝒯T\in\mathcal{T} and Ξ\Xi is an inductive probability measure in TT, then for any tTt\in T and |t|n<ht(T)|t|\leq n<\operatorname{\mathrm{ht}}(T), if Tt𝖫𝗏n(T)T_{\geq t}\cap\operatorname{\mathsf{Lv}}_{n}(T) is a front of TtT_{\geq t} (which holds when TT is well-pruned), then Ξn(Tt𝖫𝗏n(T))=Ξ({t})\Xi_{n}\left(T_{\geq t}\cap\operatorname{\mathsf{Lv}}_{n}(T)\right)=\Xi(\{t\}).

Proof..

Apply (4.15.1) to any front of TT containing Tt𝖫𝗏n(T)T_{\geq t}\cap\operatorname{\mathsf{Lv}}_{n}(T), e.g. Frn(T)\mathrm{Fr}_{n}(T). ∎

We will prove later that Π\Pi is surjective (see 4.24), however, it is evident that Π\Pi is not one-to-one, as it does not matter how a probability tree is defined above a measure-zero node. While it might seem reasonable to eliminate measure-zero nodes and restrict to probability trees with strictly positive measures, this approach is not ideal, as the applications often involve trees with measure-zero nodes. Instead, we will isolate the positive part of a tree with respect to measures in 𝒯𝒫\mathcal{TP} and 𝒫\mathcal{IP}.

Definition 4.18.

Let μ¯𝒯𝒫\bar{\mu}\in\mathcal{TP} and Ξ𝒫\Xi\in\mathcal{IP}.

  1. (1)

    We say that μ¯\bar{\mu} is positive if, for any tTμ¯max(Tμ¯)t\in T_{\bar{\mu}}\smallsetminus\max(T_{\bar{\mu}}), and s𝗌𝗎𝖼𝖼Tμ¯(t)s\in\operatorname{\mathsf{succ}}_{T_{\bar{\mu}}}(t), μt({s})\mu_{t}(\{s\}) is positive. Similarly, Ξ\Xi is positive if, for any tTΞt\in T_{\Xi}, Ξ({t})>0\Xi(\{t\})>0.

  2. (2)

    If T,μ¯\langle T,\bar{\mu}\rangle is a probability tree, we say that it is positive if μ¯\bar{\mu} is positive.

  3. (3)

    NΞ{tT:Ξ({t})=0}N_{\Xi}\coloneqq\{t\in T\colon\Xi(\{t\})=0\} and Nμ¯NΞμ¯N_{\bar{\mu}}\coloneqq N_{\Xi^{\bar{\mu}}}, which are called the null part of Ξ\Xi and μ¯\bar{\mu}, respectively.

  4. (4)

    Tμ¯+Tμ¯Nμ¯T_{\bar{\mu}}^{+}\coloneqq T_{\bar{\mu}}\smallsetminus N_{\bar{\mu}} and TΞ+TΞNΞT_{\Xi}^{+}\coloneqq T_{\Xi}\smallsetminus N_{\Xi}, which are called the positive part of μ¯\bar{\mu} and Ξ\Xi, respectively.

  5. (5)

    For STμ¯S\subseteq T_{\bar{\mu}}, set μ¯Sμt[S𝗌𝗎𝖼𝖼T(t)]:tSmax(S)\bar{\mu}{\restriction}S\coloneqq\langle\mu_{t}{\restriction}[S\cap\operatorname{\mathsf{succ}}_{T}(t)]\colon t\in S\smallsetminus\max(S)\rangle.

  6. (6)

    μ¯+μ¯Tμ¯+\bar{\mu}_{+}\coloneqq\bar{\mu}{\restriction}T_{\bar{\mu}}^{+} and Ξ+ΞTΞ+\Xi_{+}\coloneqq\Xi{\restriction}T_{\Xi}^{+}.

  7. (7)

    𝒯𝒫+\mathcal{TP}_{+} is the collection of all positive μ¯𝒯𝒫\bar{\mu}\in\mathcal{TP}. Similarly, 𝒫+\mathcal{IP}_{+} is the collection of all positive Ξ𝒫\Xi\in\mathcal{IP}.

  8. (8)

    We define the functions:

    φ𝒯𝒫:\displaystyle\varphi_{\mathcal{TP}}\colon 𝒯𝒫𝒯𝒫+\displaystyle\mathcal{TP}\to\mathcal{TP}_{+} by φ𝒯𝒫(μ¯)μ¯+\varphi_{\mathcal{TP}}(\bar{\mu})\coloneqq\bar{\mu}_{+},
    φ𝒫:\displaystyle\varphi_{\mathcal{IP}}\colon 𝒢𝒫𝒫+\displaystyle\mathcal{GP}\to\mathcal{IP}_{+} by φ𝒫(Ξ)Ξ+\varphi_{\mathcal{IP}}(\Xi)\coloneqq\Xi_{+}, and
    Π+:\displaystyle\Pi_{+}\colon 𝒯𝒫+𝒫+\displaystyle\mathcal{TP}_{+}\to\mathcal{IP}_{+} is Π𝒯𝒫+\Pi{\restriction}\mathcal{TP}_{+}.

    4.19 below justifies that the co-domains of these functions, and the domain of Π+\Pi_{+}, are as indicated.

Next, we list some basic properties of the notions introduced in 4.18.

Fact 4.19.

Let μ¯𝒯𝒫\bar{\mu}\in\mathcal{TP} and Ξ𝒫\Xi\in\mathcal{IP}. Then:

  1. (a)

    Tμ¯+=TΞμ¯+T^{+}_{\bar{\mu}}=T^{+}_{\Xi^{\bar{\mu}}}.

  2. (b)

    TΞ+T^{+}_{\Xi} is a subtree of TΞT_{\Xi} (so TΞ+𝒯T^{+}_{\Xi}\in\mathcal{T}) and maxTΞ+=TΞ+maxTΞ\max T^{+}_{\Xi}=T^{+}_{\Xi}\cap\max T_{\Xi}. Moreover, if TΞT_{\Xi} is well-pruned then so is TΞ+T^{+}_{\Xi} and ht(TΞ+)=ht(TΞ)\operatorname{\mathrm{ht}}(T^{+}_{\Xi})=\operatorname{\mathrm{ht}}(T_{\Xi}). A similar result holds for μ¯\bar{\mu}.

  3. (c)

    μ¯+𝒯𝒫\bar{\mu}_{+}\in\mathcal{TP} and Ξ+𝒫\Xi_{+}\in\mathcal{IP}, also Tμ¯+=Tμ¯+T_{\bar{\mu}_{+}}=T_{\bar{\mu}}^{+} and TΞ+=TΞ+T_{\Xi_{+}}=T_{\Xi}^{+}.

  4. (d)

    Ξμ¯+=Ξ+μ¯\Xi^{\bar{\mu}_{+}}=\Xi_{+}^{\bar{\mu}}.

  5. (e)

    Ξ\Xi is positive iff Ξ=Ξ+\Xi=\Xi_{+} iff TΞ=TΞ+.T_{\Xi}=T_{\Xi}^{+}.

  6. (f)

    μ¯\bar{\mu} is positive iff μ¯=μ¯+\bar{\mu}=\bar{\mu}_{+} iff Tμ¯=Tμ¯+T_{\bar{\mu}}=T_{\bar{\mu}}^{+}.

  7. (g)

    μ¯\bar{\mu} is positive iff Ξμ¯\Xi^{\bar{\mu}} is positive.

Proof..

(a): Immediate because Tμ¯=TΞμ¯T_{\bar{\mu}}=T_{\Xi^{\bar{\mu}}} and Nμ¯=NΞμ¯N_{\bar{\mu}}=N_{\Xi^{\bar{\mu}}}.

(b): We just deal with Ξ\Xi since it implies the result for μ¯\bar{\mu} by (a). First notice that Ξ()=1\Xi(\langle\,\rangle)=1, so TΞ+\langle\,\rangle\in T^{+}_{\Xi}, and it is clear that sts\subseteq t in TΞT_{\Xi} implies Ξ(t)Ξ(s)\Xi(t)\leq\Xi(s). Therefore, TΞ+T^{+}_{\Xi} is a subtree of TΞT_{\Xi}.

Now let tTΞ+max(TΞ)t\in T_{\Xi}^{+}\smallsetminus\max(T_{\Xi}), so Ξ({t})>0\Xi(\{t\})>0. Since we have that Ξ({t})=s𝗌𝗎𝖼𝖼TΞ(t)Ξ({s})\Xi(\{t\})=\sum_{s\in\operatorname{\mathsf{succ}}_{T_{\Xi}}(t)}\Xi(\{s\}), there must be some s𝗌𝗎𝖼𝖼TΞ(t)s\in\operatorname{\mathsf{succ}}_{T_{\Xi}}(t) such that Ξ({s})>0\Xi(\{s\})>0, so sTΞ+s\in T^{+}_{\Xi} and, hence tmaxTΞ+t\notin\max T^{+}_{\Xi}. This indicates that maxTΞ+maxTΞ\max T^{+}_{\Xi}\subseteq\max T_{\Xi} (the contention TΞ+maxTΞmaxTΞ+T^{+}_{\Xi}\cap\max T_{\Xi}\subseteq\max T^{+}_{\Xi} is trivial). This implies that, whenever TΞT_{\Xi} is well-pruned, TΞ+T^{+}_{\Xi} is well-pruned with the same height.

(c): It is clear that Ξ+\Xi_{+} is a measure on 𝒫(TΞ+)\operatorname{\mathcal{P}}(T^{+}_{\Xi}) and Ξ+({})=Ξ({})=1\Xi_{+}(\{\langle\,\rangle\})=\Xi(\{\langle\,\rangle\})=1. Now, for tTΞ+maxTΞt\in T^{+}_{\Xi}\smallsetminus\max T_{\Xi}, since the nodes in 𝗌𝗎𝖼𝖼TΞ(t)𝗌𝗎𝖼𝖼TΞ+(t)\operatorname{\mathsf{succ}}_{T_{\Xi}}(t)\smallsetminus\operatorname{\mathsf{succ}}_{T^{+}_{\Xi}}(t) have measure zero,

Ξ+({t})=Ξ(𝗌𝗎𝖼𝖼TΞ(t))=Ξ(𝗌𝗎𝖼𝖼TΞ+(t))=Ξ(𝗌𝗎𝖼𝖼TΞ+(t))=Ξ+(𝗌𝗎𝖼𝖼TΞ+(t)).\Xi_{+}(\{t\})=\Xi(\operatorname{\mathsf{succ}}_{T_{\Xi}}(t))=\Xi(\operatorname{\mathsf{succ}}_{T_{\Xi}^{+}}(t))=\Xi(\operatorname{\mathsf{succ}}_{T_{\Xi}^{+}}(t))=\Xi_{+}(\operatorname{\mathsf{succ}}_{T_{\Xi}^{+}}(t)).

Thus, Ξ+𝒫\Xi_{+}\in\mathcal{IP} and TΞ+=TΞ+T_{\Xi_{+}}=T^{+}_{\Xi}. The proof for μ¯+\bar{\mu}_{+} is similar.

(d): By (c) and (a), TΞ+μ¯=TΞμ¯+=Tμ¯+=Tμ¯+=TΞμ¯+T_{\Xi^{\bar{\mu}}_{+}}=T^{+}_{\Xi^{\bar{\mu}}}=T^{+}_{\bar{\mu}}=T_{\bar{\mu}_{+}}=T_{\Xi^{\bar{\mu}_{+}}}. The equality Ξμ¯+({t})=Ξμ¯({t})\Xi^{\bar{\mu}_{+}}(\{t\})=\Xi^{\bar{\mu}}(\{t\}) for tTμ¯+t\in T_{\bar{\mu}_{+}} is a straightforward calculation using 4.6.

Item (e) and (f) clear by 4.18, and (g) follows by (e)(f), and (a). ∎

We can use μ+\mu_{+} and ν+\nu_{+} to characterize when Π(μ¯)\Pi(\bar{\mu}) and Π(ν¯)\Pi(\bar{\nu}) are equal. Furthermore, this characterization enables us to show that, by restricting to positive trees, Π+\Pi_{+} establishes a bijection between 𝒯𝒫+\mathcal{TP}_{+} and 𝒫+\mathcal{IP}_{+} (see also 4.25).

Lemma 4.20.

Let μ¯,ν¯𝒯𝒫\bar{\mu},\bar{\nu}\in\mathcal{TP}. Then Π(μ¯)=Π(ν¯)\Pi(\bar{\mu})=\Pi(\bar{\nu}) iff μ¯+=ν¯+\bar{\mu}_{+}=\bar{\nu}_{+} and Tμ¯=Tν¯T_{\bar{\mu}}=T_{\bar{\nu}}. As a consequence, Π+\Pi_{+} is a one-to-one function.

Proof..

Assume that Π(μ¯)=Ξμ¯=Ξν¯=Π(ν¯)\Pi(\bar{\mu})=\Xi^{\bar{\mu}}=\Xi^{\bar{\nu}}=\Pi(\bar{\nu}). It is clear that Tμ¯=Tν¯T_{\bar{\mu}}=T_{\bar{\nu}} and NΞμ¯=NΞν¯N_{\Xi^{\bar{\mu}}}=N_{\Xi^{\bar{\nu}}}, therefore T+TΞμ¯+=TΞν¯+T^{+}\coloneqq T_{\Xi^{\bar{\mu}}}^{+}=T_{\Xi^{\bar{\nu}}}^{+}. On the other hand, let tT+max(T+)t\in T^{+}\smallsetminus\max(T^{+}) and s𝗌𝗎𝖼𝖼T+(t)s\in\operatorname{\mathsf{succ}}_{T^{+}}(t). By 4.7 we have

μt({s})=Ξμ¯({s})Ξμ¯({t})=Ξν¯({s})Ξν¯({t})=νt({s}).\mu_{t}(\{s\})=\frac{\Xi^{\bar{\mu}}(\{s\})}{\Xi^{\bar{\mu}}(\{t\})}=\frac{\Xi^{\bar{\nu}}(\{s\})}{\Xi^{\bar{\nu}}(\{t\})}=\nu_{t}(\{s\}).

Thus, μ¯+=ν¯+\bar{\mu}_{+}=\bar{\nu}_{+}.

Conversely, if TTμ¯=Tν¯T\coloneqq T_{\bar{\mu}}=T_{\bar{\nu}} and μ¯+=ν¯+\bar{\mu}_{+}=\bar{\nu}_{+}, then Ξ+μ¯=Ξμ¯+=Ξν¯+=Ξ+ν¯\Xi_{+}^{\bar{\mu}}=\Xi^{\bar{\mu}_{+}}=\Xi^{\bar{\nu}_{+}}=\Xi_{+}^{\bar{\nu}} by 4.19. Hence T+Tμ¯+=Tν¯+T^{+}\coloneqq T^{+}_{\bar{\mu}}=T^{+}_{\bar{\nu}} and NNμ¯=Nν¯N\coloneqq N_{\bar{\mu}}=N_{\bar{\nu}}. Since the nodes in NN have measure zero with respect both Ξμ¯\Xi^{\bar{\mu}} and Ξν¯\Xi^{\bar{\nu}}, we conclude that Ξμ¯=Ξν¯\Xi^{\bar{\mu}}=\Xi^{\bar{\nu}}. ∎

The condition in 4.20 that characterizes when Π(μ¯)\Pi(\bar{\mu}) and Π(ν¯)\Pi(\bar{\nu}) are equal, motivates the definition of the following equivalence relations on 𝒯𝒫\mathcal{TP} and 𝒫\mathcal{IP}.

Definition 4.21.
  1. (1)

    We say that μ¯,ν¯𝒯𝒫\bar{\mu},\bar{\nu}\in\mathcal{TP} are positive equivalent, denoted by μ¯𝒯𝒫ν¯\bar{\mu}\equiv_{\mathcal{TP}}\bar{\nu}, iff μ¯+=ν¯+\bar{\mu}_{+}=\bar{\nu}_{+}. It is clear that this is an equivalence relation. Denote by [μ¯]𝒯𝒫[μ¯]𝒯𝒫[\bar{\mu}]_{\equiv_{\mathcal{TP}}}\coloneqq[\bar{\mu}]_{\mathcal{TP}} the 𝒯𝒫\equiv_{\mathcal{TP}}-equivalence class of μ¯\bar{\mu}.

  2. (2)

    Similarly, Ξ,ρ¯𝒫\Xi,\bar{\rho}\in\mathcal{IP} are positive equivalent, denoted by Ξ𝒫ρ¯\Xi\equiv_{\mathcal{IP}}\bar{\rho}, iff Ξ+=ρ¯+\Xi_{+}=\bar{\rho}_{+}. It is clear that this is an equivalence relation. Denote by [Ξ]𝒫[Ξ]𝒫[\Xi]_{\equiv_{\mathcal{IP}}}\coloneqq[\Xi]_{\mathcal{IP}} the 𝒫\equiv_{\mathcal{IP}}-equivalence class of Ξ\Xi.

  3. (3)

    π𝒯𝒫:𝒯𝒫𝒯𝒫/𝒯𝒫\pi_{\mathcal{TP}}\colon\mathcal{TP}\to\mathcal{TP}/\equiv_{\mathcal{TP}} and π𝒫:𝒫𝒫/𝒫\pi_{\mathcal{IP}}\colon\mathcal{IP}\to\mathcal{IP}/\equiv_{\mathcal{IP}} are defined by π𝒯𝒫(μ¯)[μ¯]𝒯𝒫\pi_{\mathcal{TP}}(\bar{\mu})\coloneqq[\bar{\mu}]_{\mathcal{TP}} and π𝒫(Ξ)[Ξ]𝒫\pi_{\mathcal{IP}}(\Xi)\coloneqq[\Xi]_{\mathcal{IP}}, respectively. Furthermore, π𝒯𝒫+π𝒯𝒫𝒯𝒫+\pi_{\mathcal{TP}}^{+}\coloneqq\pi_{\mathcal{TP}}{\restriction}\mathcal{TP}_{+} and similarly, π𝒫+π𝒫𝒫+\pi_{\mathcal{IP}}^{+}\coloneqq\pi_{\mathcal{IP}}{\restriction}\mathcal{IP}_{+}.

  4. (4)

    Define Π^:𝒯𝒫/𝒯𝒫𝒫/𝒫\hat{\Pi}\colon\mathcal{TP}/\equiv_{\mathcal{TP}}\to\mathcal{IP}/\equiv_{\mathcal{IP}} by Π^([μ¯]𝒯𝒫)[Π(μ¯)]𝒫\hat{\Pi}([\bar{\mu}]_{\mathcal{TP}})\coloneqq[\Pi(\bar{\mu})]_{\mathcal{IP}}, which is well-defined by virtue of 4.22 below.

Notice that, trivially μ¯+𝒯𝒫μ¯\bar{\mu}_{+}\equiv_{\mathcal{TP}}\bar{\mu} and Ξ+𝒫Ξ\Xi_{+}\equiv_{\mathcal{IP}}\Xi for μ¯𝒯𝒫\bar{\mu}\in\mathcal{TP} and Ξ𝒫\Xi\in\mathcal{IP}. In fact, μ¯+\bar{\mu}_{+} is the canonical class-representative of μ¯\bar{\mu}. Similarly for inductive probability measures.

We can characterize when μ¯\bar{\mu} and ν¯\bar{\nu} are positively equivalent in terms of Ξμ¯\Xi^{\bar{\mu}} and Ξν¯\Xi^{\bar{\nu}}.

Lemma 4.22.

If μ¯,ν¯𝒯𝒫\bar{\mu},\bar{\nu}\in\mathcal{TP}, then the following statements are equivalent.

  1. (i)

    μ¯𝒯𝒫ν¯\bar{\mu}\equiv_{\mathcal{TP}}\bar{\nu}.

  2. (ii)

    Ξμ¯𝒫Ξν¯\Xi^{\bar{\mu}}\equiv_{\mathcal{IP}}\Xi^{\bar{\nu}}.

  3. (iii)

    Π^([μ¯]𝒯𝒫)=Π^([ν¯]𝒯𝒫)\hat{\Pi}([\bar{\mu}]_{\mathcal{TP}})=\hat{\Pi}([\bar{\nu}]_{\mathcal{TP}}).

As a consequence, Π^\hat{\Pi} is well-defined and one-to-one.

Proof..

Let μ¯,ν¯𝒯𝒫\bar{\mu},\bar{\nu}\in\mathcal{TP}. Since Π+\Pi_{+} is one-to-one and by 4.19 (d), we have:

μ¯𝒯𝒫ν¯μ¯+=ν¯+Ξμ¯+=Ξν¯+Ξ+μ¯=Ξ+ν¯Ξμ¯𝒫Ξν¯Π(μ¯)𝒫Π(ν¯).\bar{\mu}\equiv_{\mathcal{TP}}\bar{\nu}\Leftrightarrow\bar{\mu}^{+}=\bar{\nu}^{+}\Leftrightarrow\Xi^{\bar{\mu}_{+}}=\Xi^{\bar{\nu}_{+}}\Leftrightarrow\Xi^{\bar{\mu}}_{+}=\Xi_{+}^{\bar{\nu}}\Leftrightarrow\Xi^{\bar{\mu}}\equiv_{\mathcal{IP}}\Xi^{\bar{\nu}}\Leftrightarrow\Pi(\bar{\mu})\equiv_{\mathcal{IP}}\Pi(\bar{\nu}).

On the other hand, (ii) \Leftrightarrow (iii) follows immediately from the definition of Π^\hat{\Pi}. ∎

We explore the close relationship between 𝒫\mathcal{IP} and 𝒯𝒫\mathcal{TP}, and even show that probability trees can be identified with inductive probability measures. To this end, we show how to construct probability trees from inductive probability measures.

Definition 4.23.
  1. (1)

    Denote by 𝒢𝒫\mathcal{GP} the collection of pairs (Ξ+,νt:tNΞ)(\Xi_{+},\langle\nu_{t}\colon t\in N_{\Xi}\rangle) such that Ξ𝒫\Xi\in\mathcal{IP} and, for any tNΞt\in N_{\Xi}, νt\nu_{t} is a probability measure on 𝒫(𝗌𝗎𝖼𝖼T(t))\mathcal{P}(\operatorname{\mathsf{succ}}_{T}(t)). Since (Ξ,)𝒢𝒫(\Xi,\langle\,\rangle)\in\mathcal{GP} for all Ξ𝒫+\Xi\in\mathcal{IP}_{+}, we often identify a positive Ξ\Xi with (Ξ,)(\Xi,\langle\,\rangle) and claim that 𝒫+𝒢𝒫\mathcal{IP}_{+}\subseteq\mathcal{GP}.

  2. (2)

    Given η¯𝒢𝒫\bar{\eta}\in\mathcal{GP} determined by Ξ𝒫\Xi\in\mathcal{IP} and νt:tTΞ\langle\nu_{t}\colon t\in T_{\Xi}\rangle, we define μ¯η¯𝒯𝒫\bar{\mu}^{\bar{\eta}}\in\mathcal{TP} as follows. For any tTΞmax(TΞ)t\in T_{\Xi}\smallsetminus\max(T_{\Xi}) and s𝗌𝗎𝖼𝖼T(t)s\in\operatorname{\mathsf{succ}}_{T}(t),

    μtη¯({s}){Ξ({s})Ξ({t}),if tNΞ,νt({s}),if tNΞ.\mu_{t}^{\bar{\eta}}(\{s\})\coloneqq\begin{cases}\frac{\Xi(\{s\})}{\Xi(\{t\})},&\text{if $t\notin N_{\Xi}$,}\\[4.30554pt] \nu_{t}(\{s\}),&\text{if $t\in N_{\Xi}$.}\end{cases}

    When Ξ𝒫+\Xi\in\mathcal{IP}_{+}, denote μ¯Ξμ¯(Ξ,)\bar{\mu}^{\Xi}\coloneqq\bar{\mu}^{(\Xi,\langle\,\rangle)}.

  3. (3)

    Define the function Δ:𝒢𝒫𝒯𝒫\Delta\colon\mathcal{GP}\to\mathcal{TP} by Δ(η¯)μ¯η\Delta(\bar{\eta})\coloneqq\bar{\mu}^{\eta}, which is well-defined by virtue of 4.24 (a) below.

  4. (4)

    Define γ𝒢𝒫:𝒢𝒫𝒫\gamma_{\mathcal{GP}}\colon\mathcal{GP}\to\mathcal{IP}, Δ+:𝒢𝒫𝒯𝒫+\Delta_{+}\colon\mathcal{GP}\to\mathcal{TP}_{+} and γ𝒢𝒫+:𝒢𝒫𝒫+\gamma_{\mathcal{GP}}^{+}\colon\mathcal{GP}\to\mathcal{IP}^{+} by γ𝒢𝒫(η¯)Ξ\gamma_{\mathcal{GP}}(\bar{\eta})\coloneqq\Xi, Δ+(η¯)μ¯+η¯\Delta_{+}(\bar{\eta})\coloneqq\bar{\mu}^{\bar{\eta}}_{+}, and γ𝒢𝒫+(η¯)Ξ+\gamma_{\mathcal{GP}}^{+}(\bar{\eta})\coloneqq\Xi_{+}, respectively, where η¯\bar{\eta} is determined by Ξ\Xi.

Given η¯𝒢𝒫\bar{\eta}\in\mathcal{GP} determined by Ξ\Xi, we can use μ¯η¯\bar{\mu}^{\bar{\eta}} to induce a probability tree structure on TΞT_{\Xi}. This will allow us to define a bijection between 𝒢𝒫\mathcal{GP} and 𝒯𝒫\mathcal{TP}

Lemma 4.24.

Let η¯𝒢𝒫\bar{\eta}\in\mathcal{GP} be determined by Ξ𝒫\Xi\in\mathcal{IP} and νt:tNΞ\langle\nu_{t}\colon t\in N_{\Xi}\rangle, and let μ¯𝒯𝒫\bar{\mu}\in\mathcal{TP}.

  1. (a)

    TΞ,μ¯η¯\langle T_{\Xi},\bar{\mu}^{\bar{\eta}}\rangle is a probability tree.

  2. (b)

    Ξμ¯η¯=Ξ\Xi^{\bar{\mu}^{\bar{\eta}}}=\Xi.

  3. (c)

    Π\Pi is surjective.

  4. (d)

    (Ξ+μ¯,μ¯Nμ¯)𝒢𝒫(\Xi^{\bar{\mu}}_{+},\bar{\mu}{\upharpoonright}N_{\bar{\mu}})\in\mathcal{GP}, μ¯(Ξ+μ¯,μ¯Nμ¯)=μ\bar{\mu}^{(\Xi^{\bar{\mu}}_{+},\bar{\mu}{\upharpoonright}N_{\bar{\mu}})}=\mu.

  5. (e)

    If μ¯𝒯𝒫+\bar{\mu}\in\mathcal{TP}_{+} then μ¯Ξμ¯=μ¯\bar{\mu}^{\Xi^{\bar{\mu}}}=\bar{\mu}.

  6. (f)

    If Ξ𝒫+\Xi^{\prime}\in\mathcal{IP}_{+} then μ¯Ξ𝒯𝒫+\bar{\mu}^{\Xi^{\prime}}\in\mathcal{TP}_{+}.

  7. (g)

    Δ\Delta is bijective.

  8. (h)

    Δ1(μ¯)=(Ξ+μ¯,μ¯Nμ¯)\Delta^{-1}(\bar{\mu})=(\Xi^{\bar{\mu}}_{+},\bar{\mu}{\upharpoonright}N_{\bar{\mu}}).

Proof..

(a): Let tTΞmax(TΞ)t\in T_{\Xi}\smallsetminus\max(T_{\Xi}). On the one hand, if tNΞt\notin N_{\Xi} then, by 4.23 (2), we have:

μtη¯(𝗌𝗎𝖼𝖼T(t))=s𝗌𝗎𝖼𝖼TΞ(t)Ξ({s})Ξ({t})=1Ξ({t})s𝗌𝗎𝖼𝖼TΞ(t)Ξ({s})=1.\mu_{t}^{\bar{\eta}}(\operatorname{\mathsf{succ}}_{T}(t))=\sum_{s\in\operatorname{\mathsf{succ}}_{T_{\Xi}}(t)}\frac{\Xi(\{s\})}{\Xi(\{t\})}=\frac{1}{\Xi(\{t\})}\sum_{s\in\operatorname{\mathsf{succ}}_{T_{\Xi}}(t)}\Xi(\{s\})=1.

On the other hand, if Ξ({t})=0\Xi(\{t\})=0 then μtη¯(𝗌𝗎𝖼𝖼T(t))=νt(𝗌𝗎𝖼𝖼T(t))=1\mu^{\bar{\eta}}_{t}(\operatorname{\mathsf{succ}}_{T}(t))=\nu_{t}(\operatorname{\mathsf{succ}}_{T}(t))=1 because νt\nu_{t} is a probability measure on 𝗌𝗎𝖼𝖼T(t)\operatorname{\mathsf{succ}}_{T}(t).

(b): It is clear that TΞμ¯η¯=Tμ¯η¯=TΞT_{\Xi^{\bar{\mu}^{\bar{\eta}}}}=T_{\bar{\mu}^{\bar{\eta}}}=T_{\Xi}. By induction on n<ht(TΞ)n<\mathrm{ht}(T_{\Xi}), we show that Ξμ¯η¯({t})=Ξ({t})\Xi^{\bar{\mu}^{\bar{\eta}}}(\{t\})=\Xi(\{t\}) for all t𝖫𝗏n(TΞ)t\in\operatorname{\mathsf{Lv}}_{n}(T_{\Xi}). If n=0n=0 and t𝖫𝗏0(TΞ)t\in\operatorname{\mathsf{Lv}}_{0}(T_{\Xi}) then t=t=\langle\,\rangle, so Ξμ¯η¯({t})=1=Ξ({t})\Xi^{\bar{\mu}^{\bar{\eta}}}(\{t\})=1=\Xi(\{t\}). Now assume that, for any s𝖫𝗏n(TΞ)s\in\operatorname{\mathsf{Lv}}_{n}(T_{\Xi}), Ξμ¯η¯({s})=Ξ({s})\Xi^{\bar{\mu}^{\bar{\eta}}}(\{s\})=\Xi(\{s\}). For t𝖫𝗏n+1(TΞ)t\in\operatorname{\mathsf{Lv}}_{n+1}(T_{\Xi}), by 4.7 and induction hypothesis, we have that:

Ξμ¯η¯({t})=Ξμ¯η¯({tn})μtnη¯({t})=Ξ({tn})μtnη¯({t}).\Xi^{\bar{\mu}^{\bar{\eta}}}(\{t\})=\Xi^{\bar{\mu}^{\bar{\eta}}}(\{t{\restriction}n\})\mu_{t{\restriction}n}^{\bar{\eta}}(\{t\})=\Xi(\{t{\restriction}n\})\mu^{\bar{\eta}}_{t{\restriction}n}(\{t\}).

Consider two possible cases. On the one hand, if tnNΞt{\restriction}n\notin N_{\Xi} then, by the definition of μ¯η¯\bar{\mu}^{\bar{\eta}},

Ξμ¯η¯({t})=Ξ({tn})μtnη¯({t})=Ξ({tn})Ξ({t})Ξ({tn})=Ξ({t}).\Xi^{\bar{\mu}^{\bar{\eta}}}(\{t\})=\Xi(\{t{\restriction}n\})\mu_{t{\restriction}n}^{\bar{\eta}}(\{t\})=\Xi(\{t{\restriction}n\})\frac{\Xi(\{t\})}{\Xi(\{t{\restriction}n\})}=\Xi(\{t\}).

On the other hand, if tnNΞt{\restriction}n\in N_{\Xi}, then Ξμ¯η¯({t})=0=Ξ({t}),\Xi^{\bar{\mu}^{\bar{\eta}}}(\{t\})=0=\Xi(\{t\}), where the last equality holds because Ξ(𝗌𝗎𝖼𝖼T(tn))=Ξ({tn})\Xi(\operatorname{\mathsf{succ}}_{T}(t{\upharpoonright}n))=\Xi(\{t{\restriction}n\}) and, thus, Ξ({s})=0\Xi(\{s\})=0 for any s𝗌𝗎𝖼𝖼T(tn)s\in\operatorname{\mathsf{succ}}_{T}(t{\restriction}n).

(c): For Ξ𝒫\Xi^{\prime}\in\mathcal{IP} consider any νt:tNΞ\langle\nu_{t}\colon t\in N_{\Xi^{\prime}}\rangle such that σ¯(Ξ¯+,νt:tNΞ𝒢𝒫\bar{\sigma}\coloneqq(\bar{\Xi}^{\prime}_{+},\langle\nu_{t}\colon t\in N_{\Xi^{\prime}}\rangle\in\mathcal{GP}. Then, by (a) and (b), μ¯σ¯𝒯𝒫\bar{\mu}^{\bar{\sigma}}\in\mathcal{TP} and Π(μ¯σ¯)=Ξμ¯σ¯=Ξ\Pi(\bar{\mu}^{\bar{\sigma}})={\Xi^{\prime}}^{\bar{\mu}^{\bar{\sigma}}}=\Xi^{\prime}. This shows that Π\Pi is surjective.

(d): It is clear that (Ξ+μ¯,μ¯NΞ)𝒫(\Xi^{\bar{\mu}}_{+},\bar{\mu}{\upharpoonright}N_{\Xi})\in\mathcal{IP} is represented by Ξμ¯\Xi^{\bar{\mu}}, so T(Ξ+μ¯,μ¯Nξ)=TΞμ¯=Tμ¯T_{(\Xi^{\bar{\mu}}_{+},\bar{\mu}{\upharpoonright}N_{\xi})}=T_{\Xi^{\bar{\mu}}}=T_{\bar{\mu}}. For tTμ¯max(Tμ¯)t\in T_{\bar{\mu}}\smallsetminus\max(T_{\bar{\mu}}) and s𝗌𝗎𝖼𝖼T(t)s\in\operatorname{\mathsf{succ}}_{T}(t), by 4.23 (2), if Ξμ¯({t})>0\Xi^{\bar{\mu}}(\{t\})>0 then μt(Ξ+μ¯,μ¯Nξ)({s})=Ξμ¯({s})Ξμ¯({t})=μt({s})\mu_{t}^{{(\Xi^{\bar{\mu}}_{+},\bar{\mu}{\upharpoonright}N_{\xi})}}(\{s\})=\frac{\Xi^{\bar{\mu}}(\{s\})}{\Xi^{\bar{\mu}}(\{t\})}=\mu_{t}(\{s\}), otherwise μ(Ξ+μ¯,μ¯Nξ)({s})=μt({s})\mu^{{(\Xi^{\bar{\mu}}_{+},\bar{\mu}{\upharpoonright}N_{\xi})}}(\{s\})=\mu_{t}(\{s\}).

(e): By (d) because Ξμ¯\Xi^{\bar{\mu}} is positive when μ¯\bar{\mu} is positive (so it is identified with (Ξμ¯,)(\Xi^{\bar{\mu}},\langle\,\rangle)).

(f): By (b) and 4.19 (g).

(g)(h): By (d) and because (Ξ+μ¯η¯,μ¯η¯NΞμ¯η¯)=(Ξ+,νt:tNΞ)=η¯(\Xi^{\bar{\mu}^{\bar{\eta}}}_{+},\bar{\mu}^{\bar{\eta}}{\upharpoonright}N_{\Xi^{\bar{\mu}^{\bar{\eta}}}})=(\Xi_{+},\langle\nu_{t}\colon t\in N_{\Xi}\rangle)=\bar{\eta} by (b). ∎

As a consequence of 4.24 (c) and 4.19 (g), Π^\hat{\Pi} and Π+\Pi_{+} are surjective functions. Therefore, by 4.20 and 4.22, they are bijections. Furthermore, 4.24 (b)(e) and (f) allow us to define explicitly the inverse of Π+\Pi_{+}.

Corollary 4.25.

The function Π+:𝒯𝒫+𝒫+\Pi_{+}\colon\mathcal{TP}_{+}\to\mathcal{IP}_{+} is bijective and its inverse is Δ𝒫+\Delta{\restriction}\mathcal{IP}_{+}. Moreover, Π^\hat{\Pi} is also bijective.

Corollary 4.26.

Let T,μ¯\langle T,\bar{\mu}\rangle be a probability tree. If Ξ𝒫\Xi\in\mathcal{IP} is positive, then μ¯Ξ=μ¯\bar{\mu}^{\Xi}=\bar{\mu} iff Ξ=Ξμ¯\Xi=\Xi^{\bar{\mu}}.

Finally, we can summarize the connections between 𝒯𝒫\mathcal{TP}, 𝒫\mathcal{IP}, and 𝒢𝒫\mathcal{GP} in Figure 2.

Theorem 4.27.
  1. (a)

    Δ\Delta, Π+\Pi_{+}, Π^\hat{\Pi}, π𝒫+\pi_{\mathcal{IP}}^{+} and π𝒯𝒫+\pi_{\mathcal{TP}}^{+} are bijective.

  2. (b)

    All the diagrams in Figure 2 are commutative. As a consequence, any pair of paths with the same starting and ending points produce the same function.

  3. (c)

    Π\Pi, π𝒯𝒫\pi_{\mathcal{TP}}, π𝒫\pi_{\mathcal{IP}}, Δ+\Delta_{+}, φ𝒫,φ𝒯𝒫\varphi_{\mathcal{IP}},\varphi_{\mathcal{TP}}, γ𝒢𝒫\gamma_{\mathcal{GP}} and γ𝒢𝒫+\gamma_{\mathcal{GP}}^{+} are surjective.

𝒯𝒫\mathcal{TP}𝒫\mathcal{IP}𝒯𝒫/𝒯𝒫\mathcal{TP}/_{\equiv_{\mathcal{TP}}}𝒫/𝒫\mathcal{IP}/_{\equiv_{\mathcal{IP}}}𝒯𝒫+\mathcal{TP}_{+}𝒫+\mathcal{IP}_{+}π𝒯𝒫\pi_{\mathcal{TP}}π𝒫\pi_{\mathcal{IP}}φ𝒯𝒫\varphi_{\mathcal{TP}}φ𝒫\varphi_{\mathcal{IP}}π𝒯𝒫+\pi_{\mathcal{TP}}^{+}π𝒫+\pi_{\mathcal{IP}}^{+}Π+\Pi_{+}Π^\hat{\Pi}Π\Pi𝒢𝒫\mathcal{GP}Δ\ \ \Deltaγ𝒢𝒫\gamma_{\mathcal{GP}}Δ+\Delta_{+}γ𝒢𝒫+\gamma_{\mathcal{GP}}^{+}
Figure 2. Connections between the classes associated with 𝒯𝒫\mathcal{TP}, 𝒫\mathcal{IP} and 𝒢𝒫\mathcal{GP}.
Proof..

(a): It is clear that π𝒫+\pi_{\mathcal{IP}}^{+} and π𝒯𝒫+\pi_{\mathcal{TP}}^{+} are bijective functions. The result for Π+\Pi_{+}, Π^\hat{\Pi} and Δ\Delta are from 4.24 (g) and 4.25.

(b): It is enough to prove that the seven smallest sub-diagrams are commutative. Fix μ¯𝒯𝒫\bar{\mu}\in\mathcal{TP}, Ξ𝒫\Xi\in\mathcal{IP} and η¯𝒢𝒫\bar{\eta}\in\mathcal{GP}. The far left and right sub-diagrams commute because μ+𝒯𝒫μ\mu_{+}\equiv_{\mathcal{TP}}\mu and Ξ+𝒫Ξ\Xi_{+}\equiv_{\mathcal{IP}}\Xi. It is clear that the sub-diagram at the bottom commutes.

Now, we deal with the four sub-diagrams bounded by the trapeze with vertices 𝒯𝒫\mathcal{TP}, 𝒫\mathcal{IP}, 𝒯𝒫+\mathcal{TP}_{+} and 𝒫+\mathcal{IP}_{+}. For the left sub-diagram, φ𝒯𝒫(Δ(η¯))=φ𝒯𝒫(μ¯η)=μ¯+η¯=Δ+(η¯)\varphi_{\mathcal{TP}}(\Delta(\bar{\eta}))=\varphi_{\mathcal{TP}}(\bar{\mu}^{\eta})=\bar{\mu}^{\bar{\eta}}_{+}=\Delta_{+}(\bar{\eta}). For the right, φ𝒫(γ𝒢𝒫(η¯))=γ𝒢𝒫+(η¯)\varphi_{\mathcal{IP}}(\gamma_{\mathcal{GP}}(\bar{\eta}))=\gamma_{\mathcal{GP}}^{+}(\bar{\eta}) by the definition of the maps. For the bottom sub-diagram, Π+(Δ+(η¯))=Ξ+μ¯η¯=γ𝒢𝒫+(η¯)\Pi_{+}(\Delta_{+}(\bar{\eta}))=\Xi^{\bar{\mu}^{\bar{\eta}}}_{+}=\gamma^{+}_{\mathcal{GP}}(\bar{\eta}) by 4.24 (b) and 4.19 (d). Finally, for the top sub-diagram, Π(Δ(η¯))=Π(μ¯η¯)=Ξμ¯η¯=γ𝒢𝒫(η¯)\Pi(\Delta(\bar{\eta}))=\Pi(\bar{\mu}^{\bar{\eta}})=\Xi^{\bar{\mu}^{\bar{\eta}}}=\gamma_{\mathcal{GP}}(\bar{\eta}) by 4.24 (b).

(c): It is clear that π𝒯𝒫\pi_{\mathcal{TP}}, π𝒫\pi_{\mathcal{IP}}, φ𝒫,φ𝒯𝒫\varphi_{\mathcal{IP}},\varphi_{\mathcal{TP}} are surjective, and Π\Pi is surjective by 4.24. This implies, also using (a) and (b), that Δ+\Delta^{+}, γ𝒢𝒫\gamma_{\mathcal{GP}} and γ𝒢𝒫+\gamma^{+}_{\mathcal{GP}} are surjective. ∎

4.3. Borel probability measures

So far, we have defined three distinct classes associated with probability trees, namely 𝒯𝒫\mathcal{TP}, 𝒫\mathcal{IP}, and 𝒢𝒫\mathcal{GP}. However, there is a fourth which arises naturally by noting that every probability measure on ([T])\mathcal{B}([T]) induces an inductive probability measure in TT (see 4.29 and 4.30). The new class is then the class of probability measures on ([T])\mathcal{B}([T]), which we introduce below.

Definition 4.28.

We define 𝒫\mathcal{BP} as the collection of all probability measures on ([T])\mathcal{B}([T]) for some T𝒯T\in\mathcal{T}. Notice that TT is uniquely determined by λ\lambda, so it will be denoted by TλT_{\lambda}.

Note that, when [T]=max(T)[T]=\max(T) (e.g. in the case ht(T)<ω\mathrm{ht}(T)<\omega), ([T])=𝒫(max(T))\mathcal{B}([T])=\mathcal{P}(\max(T)).

Now, we construct inductive probability measures derived from measures in 𝒫\mathcal{BP}.

Definition 4.29.

For λ𝒫\lambda\in\mathcal{BP}, define the measure Ξλ\Xi^{\lambda} on 𝒫(Tλ)\operatorname{\mathcal{P}}(T_{\lambda}) determined by Ξλ({t})λ([t]T)\Xi^{\lambda}(\{t\})\coloneqq\lambda([t]_{T}) for all tTλt\in T_{\lambda}. Furthermore, we define the function Ψ:𝒫𝒫\Psi\colon\mathcal{BP}\to\mathcal{IP} such that, for any λ𝒫\lambda\in\mathcal{BP}, Ψ(λ)Ξλ\Psi(\lambda)\coloneqq\Xi^{\lambda}.

We will see that Ξλ𝒫\Xi^{\lambda}\in\mathcal{IP} and therefore, Ψ\Psi is well-defined in the sense that ranΨ𝒫\operatorname{ran}\Psi\subseteq\mathcal{IP}.

Lemma 4.30.

If λ𝒫\lambda\in\mathcal{BP} then Ξλ\Xi^{\lambda} is an inductive probability measure in TλT_{\lambda}.

Proof..

Assume that λ𝒫\lambda\in\mathcal{BP}. For tTλmax(Tλ)t\in T_{\lambda}\smallsetminus\max(T_{\lambda}), since [t]=s𝗌𝗎𝖼𝖼Tλ(t)[s][t]=\bigcup_{s\in\operatorname{\mathsf{succ}}_{T_{\lambda}}(t)}[s] is a disjoint union,

Ξλ({t})=λ([t])=s𝗌𝗎𝖼𝖼T(t)λ([s])=s𝗌𝗎𝖼𝖼T(t)Ξλ({s})=Ξλ(𝗌𝗎𝖼𝖼T(t)),\Xi^{\lambda}(\{t\})=\lambda([t])=\sum_{s\in\operatorname{\mathsf{succ}}_{T}(t)}\lambda([s])=\sum_{s\in\operatorname{\mathsf{succ}}_{T}(t)}\Xi^{\lambda}(\{s\})=\Xi^{\lambda}(\operatorname{\mathsf{succ}}_{T}(t)),

which proves that Ξλ𝒫\Xi^{\lambda}\in\mathcal{IP}. ∎

We aim to expand Figure 2 by incorporating the new class 𝒫\mathcal{BP}, for which we need to establish connections between 𝒫\mathcal{BP} and the classes 𝒫\mathcal{IP} and 𝒯𝒫\mathcal{TP}. The relationship between 𝒫\mathcal{BP} and 𝒫\mathcal{IP} poses no issues, as it is defined by Ψ\Psi. We will show that Ψ\Psi is a bijection: one-to-one will follow by Theorem 2.10, while surjectivity is consequence of a connection between 𝒯𝒫\mathcal{TP} and 𝒫\mathcal{BP}, namely, for μ¯𝒯𝒫\bar{\mu}\in\mathcal{TP}, we will construct a λμ¯\lambda^{\bar{\mu}} satisfying Ξλμ¯=Ξμ¯\Xi^{\lambda^{\bar{\mu}}}=\Xi^{\bar{\mu}}. The definition of λμ¯\lambda^{\bar{\mu}} is easy when limTμ¯=\lim T_{\bar{\mu}}=\emptyset (i.e. [Tμ¯]=max(Tμ¯)[T_{\bar{\mu}}]=\max(T_{\bar{\mu}})), as the only possible measure is λμ¯Ξμ¯max(Tμ¯)\lambda^{\bar{\mu}}\coloneqq\Xi^{\bar{\mu}}{\upharpoonright}\max(T_{\bar{\mu}}), which is in 𝒫\mathcal{BP} with Tλμ¯=Tμ¯T_{\lambda^{\bar{\mu}}}=T_{\bar{\mu}}. However, we need more tools for the construction of λμ¯\lambda^{\bar{\mu}} when limTμ¯\lim T_{\bar{\mu}}\neq\emptyset and to prove Ξλμ¯=Ξμ¯\Xi^{\lambda^{\bar{\mu}}}=\Xi^{\bar{\mu}} (even in the case limTμ¯=\lim T_{\bar{\mu}}=\emptyset). We present two ways to do this: the first uses Theorem 4.15, while the second is a concrete construction using a connection with the Lebesgue measure of the unit interval (see Section 6Theorem 6.13). The second construction is one of the main applications of this paper, as it not only addresses the problem at hand but also provides an interesting connection between probability trees and the real line, which will also have consequences for representing cardinals invariants of the continuum (see Section 7).

Theorem 4.31.

For every μ¯𝒯𝒫\bar{\mu}\in\mathcal{TP} there is a unique λμ¯𝒫\lambda^{\bar{\mu}}\in\mathcal{BP} with Tλμ¯=Tμ¯T_{\lambda^{\bar{\mu}}}=T_{\bar{\mu}} such that, for any tTμ¯t\in T_{\bar{\mu}}, λμ¯([t]T)=Ξμ¯({t})\lambda^{\bar{\mu}}([t]_{T})=\Xi^{\bar{\mu}}(\{t\}), i.e. Ξλμ¯=Ξμ¯\Xi^{\lambda^{\bar{\mu}}}=\Xi^{\bar{\mu}}.

Proof..

Set TTμ¯T\coloneqq T_{\bar{\mu}}. Define

(4.31.1) T\displaystyle\mathcal{F}_{T} ={tF[t]:FFrn(T),n<ht(T)}.\displaystyle\coloneqq\mathcal{F}=\left\{\bigcup_{t\in F}[t]\colon F\subseteq\mathrm{Fr}_{n}(T),\ n<\mathrm{ht}(T)\right\}.

This is an algebra of sets over [T][T] and every CC\in\mathcal{F} is clopen in [T][T], so T\mathcal{F}\subseteq\mathcal{B}_{T}. It is clear that [t][t]\in\mathcal{F} for all tTt\in T, so σ()=T\sigma(\mathcal{F})=\mathcal{B}_{T}. We first define λμ¯\lambda^{\bar{\mu}} on \mathcal{F} by λμ¯(A)Ξμ¯(F)\lambda^{\bar{\mu}}(A)\coloneqq\Xi^{\bar{\mu}}(F) whenever A=tF[t]A=\bigcup_{t\in F}[t] for some FFrn(T)F\subseteq\mathrm{Fr}_{n}(T) and n<ht(T)n<\mathrm{ht}(T). We show that this function is well-defined and that it is σ\sigma-additive, as it is clear that λμ¯()=Ξμ¯()=0\lambda^{\bar{\mu}}(\emptyset)=\Xi^{\bar{\mu}}(\emptyset)=0 (the only FF representing \emptyset is \emptyset).

To see that the map is well-defined, assume that A=tF[t]=tF[t]A=\bigcup_{t\in F}[t]=\bigcup_{t^{\prime}\in F^{\prime}}[t^{\prime}] where FFrn(T)F\subseteq\mathrm{Fr}_{n}(T), FFrn(T)F^{\prime}\subseteq\mathrm{Fr}_{n^{\prime}}(T) and n,n<ht(T)n,n^{\prime}<\mathrm{ht}(T). Without loss of generality, consider the case nnn\leq n^{\prime}. Then, for any tFt\in F, since [t]tF[t][t]\subseteq\bigcup_{t^{\prime}\in F^{\prime}}[t^{\prime}] we must have that TtF=TtFrn(T)T_{\geq t}\cap F^{\prime}=T_{\geq t}\cap\mathrm{Fr}_{n^{\prime}}(T). Therefore, by (4.15.1), Ξμ¯({t})=Ξ({tF:tt})\Xi^{\bar{\mu}}(\{t\})=\Xi(\{t^{\prime}\in F^{\prime}\colon t\subseteq t^{\prime}\}). This implies that

Ξμ¯(F)=tFΞμ¯({t})=tFΞμ¯({tF:tt})=Ξμ¯(F),\Xi^{\bar{\mu}}(F)=\sum_{t\in F}\Xi^{\bar{\mu}}(\{t\})=\sum_{t\in F}\Xi^{\bar{\mu}}(\{t^{\prime}\in F^{\prime}\colon t\subseteq t^{\prime}\})=\Xi^{\bar{\mu}}(F^{\prime}),

where the last equality hold because F={tF:tF(tt)}F^{\prime}=\{t^{\prime}\in F^{\prime}\colon\exists t\in F\ (t\subseteq t^{\prime})\}, which follows by tF[t]=tF[t]\bigcup_{t\in F}[t]=\bigcup_{t^{\prime}\in F^{\prime}}[t^{\prime}].

To show that λμ¯\lambda^{\bar{\mu}} is σ\sigma-additive in \mathcal{F}, for k<ωk<\omega let nk<ht(T)n_{k}<\mathrm{ht}(T), FkFrnk(T)F_{k}\subseteq\mathrm{Fr}_{n_{k}}(T), and AktFk[t]A_{k}\coloneqq\bigcup_{t\in F_{k}}[t], and assume that {Ak:k<ω}\{A_{k}\colon k<\omega\} is pairwise disjoint and k<ωAk=tF[t]\bigcup_{k<\omega}A_{k}=\bigcup_{t\in F}[t] for some FFrn(T)F\subseteq\mathrm{Fr}_{n}(T) and n<ht(T)n<\mathrm{ht}(T). We prove that Ξμ¯(F)=k<ωΞμ¯(Fk)\Xi^{\bar{\mu}}(F)=\sum_{k<\omega}\Xi^{\bar{\mu}}(F_{k}). We may assume that nknn_{k}\geq n for all k<ωk<\omega because we can modify nk,Fkn_{k},F_{k} without affecting AkA_{k} by using that, whenever nkn<ht(T)n_{k}\leq n^{\prime}<\mathrm{ht}(T), Ak=tFk[t]=tF[t]A_{k}=\bigcup_{t\in F_{k}}[t]=\bigcup_{t^{\prime}\in F^{\prime}}[t^{\prime}] where FF^{\prime} is the set of nodes tFrn(T)t^{\prime}\in\mathrm{Fr}_{n^{\prime}}(T) above some tFkt\in F_{k}.

Notice that any pair of nodes in each FkF_{k} are pairwise incompatible, as well as nodes coming from different FkF_{k} and FkF_{k^{\prime}}. Set Fωk<ωFkF_{\omega}\coloneqq\bigcup_{k<\omega}F_{k}. Since nknn_{k}\geq n for all k<ωk<\omega, we get that FωFω(Frn(T)F)F^{\prime}_{\omega}\coloneqq F_{\omega}\cup(\mathrm{Fr}_{n}(T)\smallsetminus F) is a front of TT. Hence, for tFt\in F, we must have TtFω=TtFωT_{\geq t}\cap F_{\omega}=T_{\geq t}\cap F^{\prime}_{\omega}. Therefore, by (4.15.1),

Ξμ¯(F)\displaystyle\Xi^{\bar{\mu}}(F) =tFΞμ¯({tFω:tt})=Ξμ¯(Fω)=k<ωΞμ¯(Fk)=k<ωλμ¯(Ak).\displaystyle=\sum_{t\in F}\Xi^{\bar{\mu}}(\{t^{\prime}\in F_{\omega}\colon t\subseteq t^{\prime}\})=\Xi^{\bar{\mu}}(F_{\omega})=\sum_{k<\omega}\Xi^{\bar{\mu}}(F_{k})=\sum_{k<\omega}\lambda^{\bar{\mu}}(A_{k}).

This shows that λμ¯\lambda^{\bar{\mu}} is a measure on \mathcal{F}. Therefore, by Theorem 2.10 this is extended by a unique measure on T=σ[T]()\mathcal{B}_{T}=\sigma_{[T]}(\mathcal{F}), which we still denote by λμ¯\lambda^{\bar{\mu}}. This is a probability measure because λμ¯([T])=Ξμ¯({})=1\lambda^{\bar{\mu}}([T])=\Xi^{\bar{\mu}}(\{\langle\,\rangle\})=1. To show the uniqueness, if λ𝒫\lambda\in\mathcal{BP} satisfies that Ξλ=Ξμ¯\Xi^{\lambda}=\Xi^{\bar{\mu}}, then Tλ=TΞλ=Tμ¯=TT_{\lambda}=T_{\Xi^{\lambda}}=T_{\bar{\mu}}=T and, for n<ht(T)n<\mathrm{ht}(T) and FFrn(T)F\subseteq\mathrm{Fr}_{n}(T), λ(tF[t])=tFλ([t])=tFΞμ¯({t})=Ξμ¯(F)\lambda\left(\bigcup_{t\in F}[t]\right)=\sum_{t\in F}\lambda([t])=\sum_{t\in F}\Xi^{\bar{\mu}}(\{t\})=\Xi^{\bar{\mu}}(F). Hence, λ\lambda extends the measure we already defined on \mathcal{F}, so λ=λμ¯\lambda=\lambda^{\bar{\mu}} by the uniqueness of the extension. ∎

As a consequence, we have a connection between 𝒯𝒫\mathcal{TP} and 𝒫\mathcal{BP}.

Definition 4.32.

We define the function Λ:𝒯𝒫𝒫\Lambda\colon\mathcal{TP}\to\mathcal{BP} by Λ(μ¯)λμ¯\Lambda(\bar{\mu})\coloneqq\lambda^{\bar{\mu}}.

We list some properties of Λ\Lambda below.

Lemma 4.33.
  1. (a)

    Ψ\Psi is a bijective function and, for Ξ𝒫\Xi\in\mathcal{IP}, Ψ1(Ξ)=λμ¯\Psi^{-1}(\Xi)=\lambda^{\bar{\mu}} where μ¯𝒯𝒫\bar{\mu}\in\mathcal{TP} satisfies Ξμ¯=Ξ\Xi^{\bar{\mu}}=\Xi (i.e. Ψ1(Ξ)\Psi^{-1}(\Xi) does not depend on this μ\mu).

  2. (b)

    Λ\Lambda is surjective, and for μ¯,ν¯𝒯𝒫\bar{\mu},\bar{\nu}\in\mathcal{TP}, Λ(μ¯)=Λ(ν¯)\Lambda(\bar{\mu})=\Lambda(\bar{\nu}) iff Π(μ¯)=Π(ν¯)\Pi(\bar{\mu})=\Pi(\bar{\nu}).

Proof..

(a): Let λ,λ𝒫\lambda,\lambda^{\prime}\in\mathcal{BP} be such that Ξλ=Ξλ\Xi^{\lambda}=\Xi^{\lambda^{\prime}}. Then Tλ=TΞλ=TλT_{\lambda}=T_{\Xi^{\lambda}}=T_{\lambda^{\prime}} and, by 4.29, for any tTλt\in T_{\lambda}, λ([t])=Ξλ({t})=Ξλ({t})=λ([t])\lambda([t])=\Xi^{\lambda}(\{t\})=\Xi^{\lambda^{\prime}}(\{t\})=\lambda^{\prime}([t]). This implies that λTλ=λTλ\lambda{\upharpoonright}\mathcal{F}_{T_{\lambda}}=\lambda^{\prime}{\upharpoonright}\mathcal{F}_{T_{\lambda}} (see (4.31.1)), which implies λ=λ\lambda=\lambda^{\prime} by Theorem 2.10. This proves that Ψ\Psi is one-to-one. On the other hand, the surjectivity follows easily from Theorem 4.31: for Ξ𝒫\Xi\in\mathcal{IP}, since Π\Pi is surjective, we can find a μ¯𝒯𝒫\bar{\mu}\in\mathcal{TP} such that Ξμ¯=Ξ\Xi^{\bar{\mu}}=\Xi, so Ψ(λμ¯)=Ξλμ¯=Ξμ¯=Ξ\Psi(\lambda^{\bar{\mu}})=\Xi^{\lambda^{\bar{\mu}}}=\Xi^{\bar{\mu}}=\Xi. This also shows that Ψ1(Ξ)=λμ¯\Psi^{-1}(\Xi)=\lambda^{\bar{\mu}}.

(b): By (a), Λ=Ψ1Π\Lambda=\Psi^{-1}\circ\Pi, which is a composition of two surjective functions. Moreover, for μ¯,ν¯𝒯𝒫\bar{\mu},\bar{\nu}\in\mathcal{TP}, Π(μ¯)=Π(ν¯)\Pi(\bar{\mu})=\Pi(\bar{\nu}) iff Ψ1(Π(μ¯)))=Ψ1(Π(ν¯))\Psi^{-1}(\Pi(\bar{\mu})))=\Psi^{-1}(\Pi(\bar{\nu})), which means that Λ(μ¯)=Λ(ν¯)\Lambda(\bar{\mu})=\Lambda(\bar{\nu}). ∎

Similar to the case of 𝒯𝒫\mathcal{TP} and 𝒫\mathcal{IP}, we can define an equivalence relation on 𝒫\mathcal{BP}.

Definition 4.34.
  1. (1)

    Let 𝒫+\mathcal{BP}_{+} be the class of λ𝒫\lambda\in\mathcal{BP} such that λ([t])>0\lambda([t])>0 for all tTλt\in T_{\lambda}, i.e. the only measure zero open subset of [T][T] is the empty set.

  2. (2)

    For λ𝒫\lambda\in\mathcal{BP}, set Nλ{tTλ:λ([t])=0}N_{\lambda}\coloneqq\{t\in T_{\lambda}\colon\lambda([t])=0\}, Tλ+TλNλT_{\lambda}^{+}\coloneqq T_{\lambda}\smallsetminus N_{\lambda}, and λ+λ([Tλ+])\lambda_{+}\coloneqq\lambda{\upharpoonright}\mathcal{B}([T^{+}_{\lambda}]). Also set NλtNλ[t]N^{*}_{\lambda}\coloneqq\bigcup_{t\in N_{\lambda}}[t].

  3. (3)

    We say that λ,λ𝒫\lambda,\lambda^{\prime}\in\mathcal{BP} are positive equivalent, denoted by λ𝒫λ\lambda\equiv_{\mathcal{BP}}\lambda^{\prime}, iff λ+=λ+\lambda^{+}=\lambda^{\prime}_{+}. Denote the equivalence class of λ\lambda by [λ]𝒫[\lambda]_{\mathcal{BP}}.

  4. (4)

    Define the maps

    Ψ+\displaystyle\Psi_{+} :𝒫+𝒫+\displaystyle\colon\mathcal{BP}_{+}\to\mathcal{IP}_{+} by Ψ+Ψ𝒫+\Psi_{+}\coloneqq\Psi{\upharpoonright}\mathcal{BP}_{+},
    Λ+\displaystyle\Lambda_{+} :𝒯𝒫+𝒫+\displaystyle\colon\mathcal{TP}_{+}\to\mathcal{BP}_{+} by Λ+Λ𝒯𝒫+\Lambda_{+}\coloneqq\Lambda{\upharpoonright}\mathcal{TP}_{+},
    π𝒫+\displaystyle\pi^{+}_{\mathcal{BP}} :𝒫+𝒫/𝒫\displaystyle\colon\mathcal{BP}_{+}\to\mathcal{BP}/_{\equiv_{\mathcal{BP}}} by π𝒫+(λ)[λ]𝒫\pi^{+}_{\mathcal{BP}}(\lambda)\coloneqq[\lambda]_{\mathcal{BP}},
    Ψ^\displaystyle\hat{\Psi} :𝒫/𝒫𝒫/𝒫\displaystyle\colon\mathcal{BP}/_{\equiv_{\mathcal{BP}}}\to\mathcal{IP}/_{\equiv_{\mathcal{IP}}} by Ψ^([λ]𝒫)[Ψ(λ)]𝒫\hat{\Psi}([\lambda]_{\mathcal{BP}})\coloneqq[\Psi(\lambda)]_{\mathcal{IP}}, and
    Λ^\displaystyle\hat{\Lambda} :𝒯𝒫/𝒯𝒫𝒫/𝒫\displaystyle\colon\mathcal{TP}/_{\equiv_{\mathcal{TP}}}\to\mathcal{BP}/_{\equiv_{\mathcal{BP}}} by Λ^([μ¯]𝒯𝒫)[Λ(μ¯)]𝒫\hat{\Lambda}([\bar{\mu}]_{\mathcal{TP}})\coloneqq[\Lambda(\bar{\mu})]_{\mathcal{BP}}.

    These maps are well-defined thanks to the following result.

Fact 4.35.

Let μ¯,ν¯𝒯𝒫\bar{\mu},\bar{\nu}\in\mathcal{TP} and λ,λ𝒫\lambda,\lambda^{\prime}\in\mathcal{BP}. Then:

  1. (a)

    Nλ=NΞλN_{\lambda}=N_{\Xi^{\lambda}} and Tλ+=TΞλ+T^{+}_{\lambda}=T^{+}_{\Xi^{\lambda}}.

  2. (b)

    NλN^{*}_{\lambda} is the largest open measure zero subset of [Tλ][T_{\lambda}] and [Tλ+]=[Tλ]Nλ[T^{+}_{\lambda}]=[T_{\lambda}]\smallsetminus N^{*}_{\lambda}.

  3. (c)

    λ+𝒫+\lambda_{+}\in\mathcal{BP}_{+}, Tλ+=Tλ+T_{\lambda_{+}}=T^{+}_{\lambda} and Ξλ+=Ξ+λ\Xi^{\lambda_{+}}=\Xi^{\lambda}_{+}.

  4. (d)

    λ𝒫+\lambda\in\mathcal{BP}_{+} iff Tλ+=TλT^{+}_{\lambda}=T_{\lambda} iff λ+=λ\lambda_{+}=\lambda.

  5. (e)

    λ+=λ+\lambda_{+}=\lambda^{\prime}_{+} iff Ξλ𝒫Ξλ\Xi^{\lambda}\equiv_{\mathcal{IP}}\Xi^{\lambda^{\prime}} (so Ψ^\hat{\Psi} is well-defined).

  6. (f)

    λμ¯+=λ+μ¯\lambda^{\bar{\mu}_{+}}=\lambda^{\bar{\mu}}_{+} and Tμ¯+=Tμ¯+T^{\bar{\mu}_{+}}=T^{+}_{\bar{\mu}}.

  7. (g)

    μ¯+=ν¯+\bar{\mu}_{+}=\bar{\nu}_{+} iff λ+μ¯=λ+ν¯\lambda^{\bar{\mu}}_{+}=\lambda^{\bar{\nu}}_{+} (so Λ^\hat{\Lambda} is well-defined).

Proof..

(a): Immediate by the definition of Ξλ\Xi^{\lambda}.

(b): Clear because NλN^{*}_{\lambda} is composed by all the measure zero basic clopen sets.

(c): By (a), Tλ+𝒯T^{+}_{\lambda}\in\mathcal{T}. It is clear that λ+\lambda_{+} is a measure on ([Tλ+])\mathcal{B}([T^{+}_{\lambda}]) and λ([Tλ+])=λ([Tλ])=1\lambda([T^{+}_{\lambda}])=\lambda([T_{\lambda}])=1 by (b), so λ+𝒫\lambda_{+}\in\mathcal{BP} and Tλ+=Tλ+T_{\lambda_{+}}=T^{+}_{\lambda}. Moreover, for tTλ+t\in T^{+}_{\lambda}, [t]Tλ+=[t]TλNλ[t]_{T^{+}_{\lambda}}=[t]_{T_{\lambda}}\smallsetminus N^{*}_{\lambda}, so λ([t]Tλ+)=λ([t]Tλ)>0\lambda([t]_{T^{+}_{\lambda}})=\lambda([t]_{T_{\lambda}})>0. Thus λ+𝒫+\lambda_{+}\in\mathcal{BP}_{+}.

Finally, TΞλ+=Tλ+=Tλ+=TΞλ+T_{\Xi^{\lambda^{+}}}=T_{\lambda^{+}}=T^{+}_{\lambda}=T^{+}_{\Xi^{\lambda}} and, for tTλ+t\in T^{+}_{\lambda}, Ξλ+({t})=λ([t]Tλ+)=λ([t]Tλ)=Ξ({t})\Xi^{\lambda_{+}}(\{t\})=\lambda([t]_{T^{+}_{\lambda}})=\lambda([t]_{T_{\lambda}})=\Xi(\{t\}). Hence Ξλ+=Ξ+λ\Xi^{\lambda_{+}}=\Xi^{\lambda}_{+}.

(d): Clear by (c) and the definitions.

(e): Clear by (c) and because Ψ\Psi is a bijection.

(f): By Theorem 4.31, Ξλμ¯+=Ξμ¯+=Ξ+μ¯=Ξ+λμ¯=Ξλ+μ¯\Xi^{\lambda^{\bar{\mu}_{+}}}=\Xi^{\bar{\mu}_{+}}=\Xi^{\bar{\mu}}_{+}=\Xi^{\lambda^{\bar{\mu}}}_{+}=\Xi^{\lambda^{\bar{\mu}}_{+}}, so the result follows because Ψ\Psi is a bijection.

(g): The implication from left to right is clear by (f). For the converse, if λ+μ¯=λ+ν¯\lambda^{\bar{\mu}}_{+}=\lambda^{\bar{\nu}}_{+} then, by (c) and Theorem 4.31, Ξμ¯+=Ξλμ¯+=Ξλν¯+=Ξν¯+\Xi^{\bar{\mu}_{+}}=\Xi^{\lambda^{\bar{\mu}_{+}}}=\Xi^{\lambda^{\bar{\nu}_{+}}}=\Xi^{\bar{\nu}_{+}}, so μ¯+=ν¯+\bar{\mu}_{+}=\bar{\nu}_{+} because Π+\Pi_{+} is a bijection. ∎

Corollary 4.36.

The functions Ψ^\hat{\Psi} and Λ^\hat{\Lambda} are bijections.

Finally, we can expand Figure 2 by including 𝒫\mathcal{BP}.

Theorem 4.37.

Consider the diagram in Figure 3. Then:

  1. (a)

    All the sub-diagrams in Figure 3 are commutative. As a consequence, any pair of paths that start at the same point and end at the same point produce the same function. Moreover:

    1. (a.1)

      any path from 𝒫\mathcal{BP} to 𝒫+\mathcal{BP}_{+} is the map λλ+\lambda\mapsto\lambda_{+}, and

    2. (a.2)

      any path from 𝒫\mathcal{BP} to 𝒫/𝒫\mathcal{BP}/_{\equiv_{\mathcal{BP}}} is the map λ[λ]𝒫\lambda\mapsto[\lambda]_{\mathcal{BP}}.

  2. (b)

    Π\Pi, Λ\Lambda, π𝒯𝒫\pi_{\mathcal{TP}}, π𝒫\pi_{\mathcal{IP}}, Δ+\Delta_{+}, φ𝒫,φ𝒯𝒫\varphi_{\mathcal{IP}},\varphi_{\mathcal{TP}}, γ𝒢𝒫\gamma_{\mathcal{GP}} and γ𝒢𝒫+\gamma_{\mathcal{GP}}^{+} are surjective.

  3. (c)

    Ψ\Psi, Δ\Delta, Π+\Pi_{+}, Ψ+\Psi_{+}, Λ+\Lambda_{+}, Π^\hat{\Pi}, Ψ^\hat{\Psi}, Λ^\hat{\Lambda}, π𝒫+\pi_{\mathcal{IP}}^{+}, π𝒯𝒫+\pi_{\mathcal{TP}}^{+} and π𝒫+\pi^{+}_{\mathcal{BP}} are bijective.

𝒫\mathcal{BP}𝒯𝒫\mathcal{TP}𝒫\mathcal{IP}𝒢𝒫\mathcal{GP}𝒯𝒫+\mathcal{TP}_{+}𝒫+\mathcal{IP}_{+}𝒫+\mathcal{BP}_{+}𝒫/𝒫\mathcal{BP}/_{\equiv_{\mathcal{BP}}}𝒯𝒫/𝒯𝒫\mathcal{TP}/_{\equiv_{\mathcal{TP}}}𝒫/𝒫\mathcal{IP}/_{\equiv_{\mathcal{IP}}}π𝒯𝒫\pi_{\mathcal{TP}}π𝒫\pi_{\mathcal{IP}}φ𝒯𝒫\varphi_{\mathcal{TP}}φ𝒫\varphi_{\mathcal{IP}}π𝒯𝒫+\pi_{\mathcal{TP}}^{+}π𝒫+\pi_{\mathcal{IP}}^{+}Π+\Pi_{+}Π^\hat{\Pi}Π\PiΔ\ \ \Deltaγ𝒢𝒫\gamma_{\mathcal{GP}}Δ+\Delta_{+}γ𝒢𝒫+\gamma_{\mathcal{GP}}^{+}Λ\LambdaΨ\PsiΛ^\hat{\Lambda}\ \ Ψ^\ \ \hat{\Psi}Λ+\ \ \Lambda_{+}Ψ+1\ \Psi^{-1}_{+}π𝒫+\pi_{\mathcal{BP}}^{+}
Figure 3. Connections between the classes associated to 𝒯𝒫\mathcal{TP}, 𝒫\mathcal{IP}, 𝒢𝒫\mathcal{GP} and 𝒫\mathcal{BP}.
Proof..

The commutativity of the sub-diagrams of Figure 3 follows easily by Theorem 4.27 and 4.31.

We have (b) and (c) as a consequence of our previous efforts in Theorem 4.27, 4.33, 4.35 and 4.36. ∎

Corollary 4.38.

Let T,μ¯\langle T,\bar{\mu}\rangle be a probability tree.

  1. (a)

    If λ𝒫\lambda\in\mathcal{BP} and Ξμ¯=Ξλ\Xi^{\bar{\mu}}=\Xi^{\lambda} then λ=λμ¯\lambda=\lambda^{\bar{\mu}}. As a consequence, λ=λμ¯Ξλ\lambda=\lambda^{\bar{\mu}^{\Xi^{\lambda}}} whenever μ¯\bar{\mu} is positive.

  2. (b)

    If μ¯\bar{\mu} is positive then μ¯Ξλμ¯=μ¯\bar{\mu}^{\Xi^{\lambda^{\bar{\mu}}}}=\bar{\mu}.

4.4. Relative expected value in probability trees

In this subsection, we are going to introduce a notion of expected value on probability trees that we call relative expected value.

For all this section fix a probability tree T,μ¯\langle T,\bar{\mu}\rangle. Recall that, for any tTt\in T, TtT_{\geq t} is the set of all nodes in TT above tt. Notice that, if 0n<ht(Tt)0\leq n<\operatorname{\mathrm{ht}}(T_{\geq t}), then 𝖫𝗏n(Tt)𝖫𝗏|t|+n(T)\operatorname{\mathsf{Lv}}_{n}(T_{\geq t})\subseteq\operatorname{\mathsf{Lv}}_{|t|+n}(T). Furthermore, TtT_{\geq t} inherits a probability tree structure from T,μ¯\langle T,\bar{\mu}\rangle:

Lemma 4.39.

Let T,μ¯T\langle T,\bar{\mu}^{T}\rangle be a probability tree. Then, for any tTt\in T, TtT_{\geq t} inherits probability tree structure from TT in a natural way, that is, Tt,μ¯Tt\langle T_{\geq t},\bar{\mu}^{T_{\geq t}}\rangle is a probability tree, where μ¯TtμsT:sTtmax(Tt)\bar{\mu}^{T_{\geq t}}\coloneqq\langle\mu_{s}^{T}\colon s\in T_{\geq t}\smallsetminus\max(T_{\geq t})\rangle.

Since for any sTts\in T_{\geq t}, μTt=μsT\mu^{T_{\geq t}}=\mu_{s}^{T}, we can abuse of the notation and write “Tt,μ¯\langle T_{\geq t},\bar{\mu}\rangle” instead of “T,μ¯Tt\langle T,\bar{\mu}^{T_{\geq t}}\rangle”. This will be widely applied in this subsection.

We can now define the relative expected value in probability trees.

Definition 4.40.

Let tTt\in T, m|t|m\coloneqq|t|, mn<ht(T)m\leq n<\mathrm{ht}(T) such that Tt𝖫𝗏n(T)=TtFrn(T)T_{\geq t}\cap\operatorname{\mathsf{Lv}}_{n}(T)=T_{\geq t}\cap\mathrm{Fr}_{n}(T), and let XX be a random variable on the probability space Frn(T),Ξμ¯𝒫(Frn(T)).\langle\mathrm{Fr}_{n}(T),\Xi^{\bar{\mu}}{\upharpoonright}\operatorname{\mathcal{P}}(\mathrm{Fr}_{n}(T))\rangle. Then, we define:

E𝖫𝗏n(T)[X:sm=t]E𝖫𝗏nm(Tt)[X𝖫𝗏nm(Tt)],\operatorname{E}_{\operatorname{\mathsf{Lv}}_{n}(T)}[X\colon s{\restriction}m=t]\coloneqq\operatorname{E}_{\operatorname{\mathsf{Lv}}_{n-m}(T_{\geq t})}[X{\restriction}\operatorname{\mathsf{Lv}}_{n-m}(T_{\geq t})],

and call it the relative expected value of XX with respect to tt. Here, X𝖫𝗏nm(Tt)X{\restriction}\operatorname{\mathsf{Lv}}_{n-m}(T_{\geq t}) is interpreted as a random variable on 𝖫𝗏nm(Tt),Ξμ¯Tt𝒫(𝖫𝗏nm(Tt))\langle\operatorname{\mathsf{Lv}}_{n-m}(T_{\geq t}),\Xi^{\bar{\mu}^{T_{\geq t}}}{\upharpoonright}\operatorname{\mathcal{P}}(\operatorname{\mathsf{Lv}}_{n-m}(T_{\geq t}))\rangle. When the context is clear, we simply write En[X:sm=t]E_{n}[X\colon s{\restriction}m=t] or even E[X:sm=t]\operatorname{E}[X\colon s{\restriction}m=t], instead of E𝖫𝗏n(T)[X:sm=t].\operatorname{E}_{\operatorname{\mathsf{Lv}}_{n}(T)}[X\colon s{\restriction}m=t].

Notice that the “ss” above is a dummy variable, that is, the expected value of XX is calculated by varying ss over the nodes in TT at level nn that extend tt. Since the relative expected value is defined in terms of the typical expected value, it is clear that it is linear, i.e. for a,ba,b\in\mathbb{R} and any random variables X,YX,Y on Frn(T)\mathrm{Fr}_{n}(T),

En[aX+bY:sm=t]=aEn[X:sm=t]+bEn[Y:sm=t].\operatorname{E}_{n}[aX+bY\colon s{\restriction}m=t]=a\,\operatorname{E}_{n}[X\colon s{\restriction}m=t]+b\,\operatorname{E}_{n}[Y\colon s{\restriction}m=t].

The following result allows us to decompose the probability of the successors of tt at the level m+nm+n of T,T, in terms of the probability at the level nn of TtT_{\geq t}:

Lemma 4.41.

For tst\leq s in TT, Ξμ¯({s})=Ξnμ¯Tt({s})Ξμ¯({t}).\Xi^{\bar{\mu}}(\{s\})=\Xi^{\bar{\mu}^{T_{\geq t}}}_{n}(\{s\})\Xi^{\bar{\mu}}(\{t\}).

The relative expected value can be calculated as a composition of relative expected values at intermediate levels, as follows (see Figure 4).

\bullet\bullet\bullet\bulletmmnnkkTTTtT_{\geq t}TsT_{\geq s}mm𝖫𝗏k(T)\operatorname{\mathsf{Lv}}_{k}(T)𝖫𝗏km(Tt)\operatorname{\mathsf{Lv}}_{k-m}(T_{\geq t})𝖫𝗏kn(Ts)\operatorname{\mathsf{Lv}}_{k-n}(T_{\geq s})rrsstt
Figure 4. The situation in Theorem 4.42.
Theorem 4.42.

Let mnk<ωm\leq n\leq k<\omega, t𝖫𝗏m(T),t\in\operatorname{\mathsf{Lv}}_{m}(T), and assume that TtFrk(T)=Tt𝖫𝗏k(T)T_{\geq t}\cap\mathrm{Fr}_{k}(T)=T_{\geq t}\cap\operatorname{\mathsf{Lv}}_{k}(T). If XX is a random variable on Frk(T)\mathrm{Fr}_{k}(T), then

Ek[X:rm=t]=En[Ek[X:rn=s]:sm=t].\operatorname{E}_{k}[X\colon r{\restriction}m=t]=\operatorname{E}_{n}\big{[}\operatorname{E}_{k}[X\colon r{\restriction}n=s]\colon s{\restriction}m=t\big{]}.
Proof.

Let EEn[Ek[X:r=s]:sm=t]E\coloneqq\operatorname{E}_{n}\big{[}\operatorname{E}_{k}[X\colon r{\restriction}\ell=s]\colon s{\restriction}m=t\big{]}. By 4.40, we have that:

Es𝖫𝗏nm(Tt)Ek[X:rn=s]Ξμ¯Tt({s})=s𝖫𝗏nm(Tt)E[X𝖫𝗏kn(Ts)]Ξμ¯Tt({s})=s𝖫𝗏nm(Tt)(r𝖫𝗏kn(Ts)X(r)Ξμ¯Ts({r}))Ξμ¯Tt({s})=s𝖫𝗏nm(Tt)r𝖫𝗏kn(Ts)X(r)Ξμ¯Tt({r})=r𝖫𝗏km(Tt)X(r)Ξμ¯Tt({r})=E[X𝖫𝗏km(Tt)]=Ek[X:rm=t],\begin{split}E&\coloneqq\sum_{s\in\operatorname{\mathsf{Lv}}_{n-m}(T_{\geq t})}\operatorname{E}_{k}[X\colon r{\restriction}n=s]\Xi^{\bar{\mu}^{T_{\geq t}}}(\{s\})=\sum_{s\in\operatorname{\mathsf{Lv}}_{n-m}(T_{\geq t})}\operatorname{E}[X{\restriction}\operatorname{\mathsf{Lv}}_{k-n}(T_{\geq s})]\Xi^{\bar{\mu}^{T_{\geq t}}}(\{s\})\\ &=\sum_{s\in\operatorname{\mathsf{Lv}}_{n-m}(T_{\geq t})}\left(\sum_{r\in\operatorname{\mathsf{Lv}}_{k-n}(T_{\geq s})}X(r)\Xi^{\bar{\mu}^{T_{\geq s}}}(\{r\})\right)\Xi^{\bar{\mu}^{T_{\geq t}}}(\{s\})\\ &=\sum_{s\in\operatorname{\mathsf{Lv}}_{n-m}(T_{\geq t})}\sum_{r\in\operatorname{\mathsf{Lv}}_{k-n}(T_{\geq s})}X(r)\Xi^{\bar{\mu}^{T_{\geq t}}}(\{r\})\\ &=\sum_{r\in\operatorname{\mathsf{Lv}}_{k-m}(T_{\geq t})}X(r)\Xi^{\bar{\mu}^{T_{\geq t}}}(\{r\})=\operatorname{E}[X{\restriction}\operatorname{\mathsf{Lv}}_{k-m}(T_{\geq t})]=\operatorname{E}_{k}[X\colon r{\restriction}m=t],\end{split}

where Ξμ¯Tt({r})=Ξμ¯Ts({r})Ξμ¯Tt({s})\Xi^{\bar{\mu}^{T_{\geq t}}}(\{r\})=\Xi^{\bar{\mu}^{T_{\geq s}}}(\{r\})\cdot\Xi^{\bar{\mu}^{T_{\geq t}}}(\{s\}) by 4.41. ∎

Finally, as a consequence (when m=0m=0), we can express the expected value of XX in terms of the relative expected value:

Corollary 4.43.

If nk<ht(T)n\leq k<\mathrm{ht}(T), 𝖫𝗏k(T)\operatorname{\mathsf{Lv}}_{k}(T) is a front of TT and XX is a random variable on 𝖫𝗏k(T),\operatorname{\mathsf{Lv}}_{k}(T), then E[X]=En[Ek[X:rn=s]].\operatorname{E}[X]=\operatorname{E}_{n}\big{[}\operatorname{E}_{k}[X\colon r{\restriction}n=s]\big{]}.

So far we discussed the relative expected values of random variables on Frn(T)\mathrm{Fr}_{n}(T). However, we can extend this notion to random variables on any front. We first fix some terminology.

Definition 4.44.

Let tTt\in T and A,AA,A^{\prime} fronts of TT.

  1. (1)

    The node tt is below AA if tst\leq s for some sAs\in A.

  2. (2)

    The front AA is below AA^{\prime} if any node of AA is below AA^{\prime}.

For instance, Frm(T)\mathrm{Fr}_{m}(T) is a front below Frn(T)\mathrm{Fr}_{n}(T) whenever mn<ht(T)m\leq n<\mathrm{ht}(T).

Definition 4.45.

Let AA be a front of TT, tTt\in T below AA, and let XX be a random variable on A,Ξ𝒫(A)\langle A,\Xi{\upharpoonright}\operatorname{\mathcal{P}}(A)\rangle (recall from Theorem 4.15 that Ξ𝒫(A)\Xi{\upharpoonright}\operatorname{\mathcal{P}}(A) is a probability measure). Define the relative expected value of XX with respect to tt as

EA[X:s|t|=t]EATt[XATt].\operatorname{E}_{A}[X\colon s{\upharpoonright}|t|=t]\coloneqq E_{A\cap T_{\geq t}}[X{\upharpoonright}A\cap T_{\geq t}].

Notice that ATtA\cap T_{\geq t} is a front of TtT_{\geq t}. Also, XATtX{\upharpoonright}A\cap T_{\geq t} is interpreted as a random variable on the probability space ATt,Ξμ¯Tt𝒫(ATt)\langle A\cap T_{\geq t},\Xi^{\bar{\mu}^{T_{\geq t}}}{\upharpoonright}\operatorname{\mathcal{P}}(A\cap T_{\geq t})\rangle.

We can generalize Theorem 4.42 for fronts. We omit the proof, as it is very similar.

Theorem 4.46.

Assume that A,AA,A^{\prime} are fronts, tTt\in T is below AA, and AA is below AA^{\prime}. If XX is a random variable on AA^{\prime}, then

EA[X:r|t|=t]=EA[EA[X:r|s|=s]:s|t|=t].\operatorname{E}_{A^{\prime}}[X\colon r{\upharpoonright}|t|=t]=\operatorname{E}_{A}\big{[}\operatorname{E}_{A^{\prime}}[X\colon r{\upharpoonright}|s|=s]\colon s{\upharpoonright}|t|=t\big{]}.

In particular, EA[X]=EA[EA[X:r|s|=s]]\operatorname{E}_{A^{\prime}}[X]=\operatorname{E}_{A}\big{[}\operatorname{E}_{A^{\prime}}[X\colon r{\upharpoonright}|s|=s]\big{]}.

5. Bounding cumulative dependent Bernoulli distributions

As mentioned in the introduction it is well known that, by adding finite many independent and identically distributed random variables with Bernoulli distribution, we obtain a random variable with the binomial distribution. However, there are cases where we must deal with dependent random variables with Bernoulli distribution which may not be identically distributed.444See, for example, the proof of [Uri23, Main Lemma 4.3.17] and the proof of [CMU24, Main Lemma 7.17]. In the following theorem, we show how the cumulative distribution of the number of successes of these random variables can be bounded by the cumulative distribution of the binomial distribution, which proves A. Here, success corresponds to 0 and failure to 11.

For a natural number nn, nn-many dependent trials with Bernoulli distribution can be understood as a probability tree T,μ¯\langle T,\bar{\mu}\rangle where T=2nT={}^{\leq n}2 is the complete binary tree of height n+1n+1, i.e. tTt\in T iff tt is a sequence of length n{\leq}\,n composed by 0’s and 11’s (including the empty sequence). Any tTt\in T of length knk\leq n represents a sequence of success and failures of the first kk Bernoulli tests and, whenever k<nk<n, 𝗌𝗎𝖼𝖼T(t)\operatorname{\mathsf{succ}}_{T}(t) is Bernoulli distributed with probability of success μt({t0})\mu_{t}(\{t{}^{\frown}\langle 0\rangle\}), which clearly depends on tt, i.e. on the previous trials.

The random variable expressing the success at the kk-th trial for k<nk<n is Xk:𝖫𝗏k+1(T){0,1}X_{k}\colon\operatorname{\mathsf{Lv}}_{k+1}(T)\to\{0,1\}, defined by Xk(t)t(k)X_{k}(t^{\prime})\coloneqq t^{\prime}(k), so success is attained when Xk(t)=0X_{k}(t^{\prime})=0, i.e. t(k)=0t^{\prime}(k)=0. Therefore, the probability of success is

Pr[Xk=0]=t𝖫𝗏k(T)Ξμ¯({t})μt({t0})=t𝖫𝗏k(T)Ξμ¯({t0}).\Pr[X_{k}=0]=\sum_{t\in\operatorname{\mathsf{Lv}}_{k}(T)}\Xi^{\bar{\mu}}(\{t\})\mu_{t}(\{t{}^{\frown}\langle 0\rangle\})=\sum_{t\in\operatorname{\mathsf{Lv}}_{k}(T)}\Xi^{\bar{\mu}}(\{t{}^{\frown}\langle 0\rangle\}).
Theorem 5.1.

Let n<ωn<\omega, T2nT\coloneqq{}^{\leq n}2, and assume that T,μ¯\langle T,\bar{\mu}\rangle is a probability tree. Define Y:𝖫𝗏n(T)Y\colon\operatorname{\mathsf{Lv}}_{n}(T)\to\mathbb{R} as the random variable measuring the number of successes after nn trials, that is, for any t𝖫𝗏n(T),t\in\operatorname{\mathsf{Lv}}_{n}(T),

Y(t)|{k<n:t(n)=0}|.Y(t)\coloneqq|\{k<n\colon t(n)=0\}|.

Assume that there exists some p[0,1]p\in[0,1] such that, for any tTmax(T)t\in T\smallsetminus\max(T), pptμt({t0})p\leq p_{t}\coloneqq\mu_{t}(\{t{}^{\frown}\langle 0\rangle\}). Then, for all z,z\in\mathbb{R},

Pr[Yz]Pr[Bn,pz],\Pr[Y\leq z]\leq\Pr[\mathrm{B}_{n,p}\leq z],

where Bn,p\mathrm{B}_{n,p} denotes the binomail distribution of nn trials, each with probability of success pp.

Proof..

For any x[0,1]x\in[0,1] and d{0,1},d\in\{0,1\}, define:

Ixd{[0,x),d=0,[x,1],d=1.I_{x}^{d}\coloneqq\begin{cases}[0,x),&d=0,\\[4.30554pt] [x,1],&d=1.\end{cases}

For t𝖫𝗏n(T),t\in\operatorname{\mathsf{Lv}}_{n}(T), let Ctk<nIptnt(n),C_{t}^{\bullet}\coloneqq\prod_{k<n}I_{p_{t{\restriction}n}}^{t(n)}, and notice that its volume is Vol(Ct)=Ξμ¯({t}).\mathrm{Vol}(C_{t}^{\bullet})=\Xi^{\bar{\mu}}(\{t\}). It is easy to show that {Ct:t𝖫𝗏n(T)}\{C_{t}^{\bullet}\colon t\in\operatorname{\mathsf{Lv}}_{n}(T)\} is a partition of the nn-dimensional unitary cube [0,1]n{}^{n}[0,1].

For zz\in\mathbb{R} let C(z){Ct:Y(t)z,t𝖫𝗏n(T)}C^{\bullet}(z)\coloneqq\bigcup\{C^{\bullet}_{t}\colon Y(t)\leq z,\ t\in\operatorname{\mathsf{Lv}}_{n}(T)\}. Thus

(5.1.1) Pr[Yz]={Vol(Ct):Y(t)z,t𝖫𝗏n(T)}=Vol(C(z)).\Pr[Y\leq z]=\sum\{\mathrm{Vol}(C^{\bullet}_{t})\colon Y(t)\leq z,\ t\in\operatorname{\mathsf{Lv}}_{n}(T)\}=\mathrm{Vol}(C^{\bullet}(z)).

On the other hand, define the polyhedron Ctk<nIpt(n)C_{t}\coloneqq\prod_{k<n}I_{p}^{t(n)} for t𝖫𝗏n(T)t\in\operatorname{\mathsf{Lv}}_{n}(T). We can use this to express the cumulative binomial distribution because {Ct:tT}\{C_{t}\colon t\in T\} is a partition of [0,1]n{}^{n}[0,1] and, by setting C(z){Ct:Y(t)z,t𝖫𝗏n(T)}C(z)\coloneqq\bigcup\{C^{\bullet}_{t}\colon Y(t)\leq z,\ t\in\operatorname{\mathsf{Lv}}_{n}(T)\}, we obtain

(5.1.2) Pr[Bn,pz]={Vol(Ct):Y(t)z,tT}=Vol(C(z)).\Pr[\mathrm{B}_{n,p}\leq z]=\sum\{\mathrm{Vol}(C_{t})\colon Y(t)\leq z,\ t\in T\}=\mathrm{Vol}(C(z)).

To conclude the proof, thanks to (5.1.1) and (5.1.2), it is enough to show that C(z)C(z)C^{\bullet}(z)\subseteq C(z) for all zz\in\mathbb{R}. Let x¯=xk:k<nC(Z)\bar{x}=\langle x_{k}\colon k<n\rangle\in C^{\bullet}(Z), so there is a unique t𝖫𝗏n(T)t\in\operatorname{\mathsf{Lv}}_{n}(T) such that x¯Ct\bar{x}\in C^{\bullet}_{t} and Y(t)zY(t)\leq z. Likewise, there is a unique t𝖫𝗏n(T)t^{\prime}\in\operatorname{\mathsf{Lv}}_{n}(T) such that x¯Ct\bar{x}\in C_{t^{\prime}}. For k<nk<n, if t(k)=0t^{\prime}(k)=0 then xkIp0=[0,p)[0,ptk)=Iptk0x_{k}\in I^{0}_{p}=[0,p)\subseteq[0,p_{t{\upharpoonright}k})=I^{0}_{p_{t{\upharpoonright}k}} (the contention because pspp_{s}\leq p for all sTmax(T)s\in T\smallsetminus\max(T)), so t(k)=0t(k)=0. This shows that Y(t)Y(t)Y(t^{\prime})\leq Y(t), so Y(t)zY(t^{\prime})\leq z and we can conclude that x¯C(z)\bar{x}\in C(z). ∎

6. Probability trees and the real line

In this section, we explore the connection between probability trees and the real line. To this end, we start with an alternative proof of Theorem 4.31, specifically, given a probability tree T,μ¯\langle T,\bar{\mu}\rangle, we construct a measure λμ¯𝒫\lambda^{\bar{\mu}}\in\mathcal{BP} such that λμ¯([t]T)=Ξμ¯(t)\lambda^{\bar{\mu}}([t]_{T})=\Xi^{\bar{\mu}}({t}) for any tTt\in T. This construction also shows a connection between the measure space [T],T,λμ¯\langle[T],\mathcal{B}_{T},\lambda^{\bar{\mu}}\rangle and the Lebesgue measure space of [0,1][0,1], reflected by a map that, under certain conditions, preserves measure (see Theorem 6.19 and Theorem 6.28).

For all this section, we assume that T,μ¯\langle T,\bar{\mu}\rangle is a probability tree. Since TT is countable, T\mathcal{B}_{T} is the σ\sigma-algebra generated by {[t]:tT}\{[t]\colon t\in T\}.

Definition 6.1.

It is easy to show that TT is isomorphic with a tree TT_{*} such that, for any tTmaxTt\in T_{*}\smallsetminus\max T_{*}, there is some αtω\alpha_{t}\leq\omega such that 𝗌𝗎𝖼𝖼T(t)={tk:k<αt}\operatorname{\mathsf{succ}}_{T_{*}}(t)=\{t{}^{\frown}\langle k\rangle\colon k<\alpha_{t}\}. Such a TT^{*} is called a representation of TT.

To define the measure λμ¯\lambda^{\bar{\mu}}, we fix a representation TT_{*} of TT as in 6.1 and, without loss of generality, assume T=TT=T_{*}. Our construction is motivated by known connections between the Cantor space, the Baire space, and [0,1][0,1] as in e.g. [Lev02, Ch. VII, §3].

Definition 6.2.

Define a sequence I¯μ¯It:tT\bar{I}^{\bar{\mu}}\coloneqq\langle I_{t}\colon t\in T\rangle of closed intervals It=Itμ¯=[at,bt]I_{t}=I^{\bar{\mu}}_{t}=[a_{t},b_{t}] by recursion on n=htT(t)n=\operatorname{\mathrm{ht}}_{T}(t) as follows.

  • I[0,1]I_{\langle\ \rangle}\coloneqq[0,1], that is, a0a_{\langle\ \rangle}\coloneqq 0 and b1b_{\langle\ \rangle}\coloneqq 1.

  • Having defined the interval ItI_{t}, when tmaxTt\notin\max T let Itk:k<αt\langle I_{t{}^{\frown}\langle k\rangle}\colon k<\alpha_{t}\rangle be the collection of consecutive closed intervals contained in ItI_{t} where each ItkI_{t{}^{\frown}\langle k\rangle} has length 𝖫𝖻(It)μt({tk})\operatorname{\mathsf{Lb}}(I_{t})\mu_{t}(\{t{}^{\frown}\langle k\rangle\}), that is: at0ata_{t{}^{\frown}\langle 0\rangle}\coloneqq a_{t}, btkatk+𝖫𝖻(It)μt({tk})b_{t{}^{\frown}\langle k\rangle}\coloneqq a_{t{}^{\frown}\langle k\rangle}+\operatorname{\mathsf{Lb}}(I_{t})\mu_{t}(\{t{}^{\frown}\langle k\rangle\}) whenever k<αtk<\alpha_{t}, and atk+1btka_{t{}^{\frown}\langle k+1\rangle}\coloneqq b_{t{}^{\frown}\langle k\rangle} whenever k+1<αtk+1<\alpha_{t}.

Denote Qμ¯{at,bt:tT}Q_{\bar{\mu}}\coloneqq\{a_{t},b_{t}\colon t\in T\}. Then |Qμ¯|2|T|0|Q_{\bar{\mu}}|\leq 2|T|\leq\aleph_{0}, so Qμ¯Q_{\bar{\mu}} is countable.

Notice that ItkI_{t{}^{\frown}\langle k\rangle} is just one point (that is, at=bta_{t}=b_{t}) iff 𝖫𝖻(It)μt({tk})=0\operatorname{\mathsf{Lb}}(I_{t})\mu_{t}(\{t{}^{\frown}\langle k\rangle\})=0.

Let us look at many basic properties of the objects defined in 6.2.

Lemma 6.3.

Let s,tTs,t\in T.

  1. (a)

    If tmaxTt\notin\max T then, for any k<αtk<\alpha_{t},

    atk=at+𝖫𝖻(It)j<kμt(tj) and btk=at+𝖫𝖻(It)j<k+1μt(tj).a_{t{}^{\frown}\langle k\rangle}=a_{t}+\operatorname{\mathsf{Lb}}(I_{t})\sum_{j<k}\mu_{t}(t{}^{\frown}\langle j\rangle)\text{ and }b_{t{}^{\frown}\langle k\rangle}=a_{t}+\operatorname{\mathsf{Lb}}(I_{t})\sum_{j<k+1}\mu_{t}(t{}^{\frown}\langle j\rangle).
  2. (b)

    𝖫𝖻(It)=btat=Ξμ¯({t})\operatorname{\mathsf{Lb}}(I_{t})=b_{t}-a_{t}=\Xi^{\bar{\mu}}(\{t\}).

  3. (c)

    If tmaxTt\notin\max T then It{bt}k<αtItkItI_{t}\smallsetminus\{b_{t}\}\subseteq\bigcup_{k<\alpha_{t}}I_{t{}^{\frown}\langle k\rangle}\subseteq I_{t}. Moreover, k<αtItk=It\bigcup_{k<\alpha_{t}}I_{t{}^{\frown}\langle k\rangle}=I_{t} holds whenever |𝗌𝗎𝖼𝖼T(t)|<0|\operatorname{\mathsf{succ}}_{T}(t)|<\aleph_{0}. As a consequence, k<αtItkQμ¯=ItQμ¯\bigcup_{k<\alpha_{t}}I_{t{}^{\frown}\langle k\rangle}\smallsetminus Q_{\bar{\mu}}=I_{t}\smallsetminus Q_{\bar{\mu}}.

  4. (d)

    If sts\perp t then IsItQμ¯I_{s}\cap I_{t}\subseteq Q_{\bar{\mu}} and |IsIt|1|I_{s}\cap I_{t}|\leq 1. As a consequence, when sts\perp t, (IsQμ¯)(ItQμ¯)=(I_{s}\smallsetminus Q_{\bar{\mu}})\cap(I_{t}\smallsetminus Q_{\bar{\mu}})=\emptyset.

Proof..

We have that (a) is clear from the definition of ItkI_{t{}^{\frown}\langle k\rangle}.

(b): Proceed by induction on n=|t|n=|t|. When n=0n=0, Ξμ¯({})=1=𝖫𝖻(I)\Xi^{\bar{\mu}}(\{\langle\ \rangle\})=1=\operatorname{\mathsf{Lb}}(I_{\langle\ \rangle}). Assume n+1<ht(T)n+1<\mathrm{ht}(T) and that the claim is true for all s𝖫𝗏n(T)s\in\operatorname{\mathsf{Lv}}_{n}(T). Let t𝖫𝗏n+1(T)t\in\operatorname{\mathsf{Lv}}_{n+1}(T), that is, t=skt=s{}^{\frown}\langle k\rangle where stn𝖫𝗏n(T)s\coloneqq t{\upharpoonright}n\in\operatorname{\mathsf{Lv}}_{n}(T) and kt(n)<αsk\coloneqq t(n)<\alpha_{s}. By induction hypothesis, 𝖫𝖻(Is)=Ξμ¯({s})\operatorname{\mathsf{Lb}}(I_{s})=\Xi^{\bar{\mu}}(\{s\}), so 𝖫𝖻(It)=𝖫𝖻(Is)μs({t})=Ξμ¯({s})μs({t})=Ξμ¯({t})\operatorname{\mathsf{Lb}}(I_{t})=\operatorname{\mathsf{Lb}}(I_{s})\mu_{s}(\{t\})=\Xi^{\bar{\mu}}(\{s\})\mu_{s}(\{t\})=\Xi^{\bar{\mu}}(\{t\}), where the last equality holds by 4.7.

(c): Observe that btk:k<αt\langle b_{t{}^{\frown}\langle k\rangle}\colon k<\alpha_{t}\rangle is a monotone increasing sequence. Even more, by (a), if αt=ω\alpha_{t}=\omega then limkjkμt({tj})=j<ωμt({tj})=1\lim_{k\to\infty}\sum_{j\leq k}\mu_{t}(\{t{}^{\frown}\langle j\rangle\})=\sum_{j<\omega}\mu_{t}(\{t{}^{\frown}\langle j\rangle\})=1, so limkbtk=at+𝖫𝖻(It)=bt\lim_{k\to\infty}b_{t{}^{\frown}\langle k\rangle}=a_{t}+\operatorname{\mathsf{Lb}}(I_{t})=b_{t}. Then, Itj:j<ω\langle I_{t{}^{\frown}\langle j\rangle}\colon j<\omega\rangle covers It{bt}I_{t}\smallsetminus\{b_{t}\}.

To prove the “moreover” in the statement, assume that |𝗌𝗎𝖼𝖼T(t)|<0|\operatorname{\mathsf{succ}}_{T}(t)|<\aleph_{0}. Hence, αt<ω\alpha_{t}<\omega and therefore, jαt1𝖫𝖻(Itj)=1\sum_{j\leq\alpha_{t}-1}\operatorname{\mathsf{Lb}}(I_{t{}^{\frown}\langle j\rangle})=1, so btαt1=at+𝖫𝖻(It)=btb_{t{}^{\frown}\langle\alpha_{t}-1\rangle}=a_{t}+\operatorname{\mathsf{Lb}}(I_{t})=b_{t}. As a consequence, Itj:j<αt\langle I_{t{}^{\frown}\langle j\rangle}\colon j<\alpha_{t}\rangle covers ItI_{t}.

(d): If sts\perp t, then there is some n<ωn<\omega such that sn=tns{\upharpoonright}n=t{\upharpoonright}n and s(n)t(n)s(n)\neq t(n). Without loss of generality, we assume that s(n)<t(n)s(n)<t(n). Then, bs(n+1)at(n+1)b_{s{\upharpoonright}(n+1)}\leq a_{t{\upharpoonright}(n+1)}, so |Is(n+1)It(n+1)|1|I_{s{\upharpoonright}(n+1)}\cap I_{t{\upharpoonright}(n+1)}|\leq 1 and, in case they intersect, the unique point in the intersection must be in Qμ¯Q_{\bar{\mu}}. Hence, the result follows because IsItIs(n+1)It(n+1)I_{s}\cap I_{t}\subseteq I_{s{\upharpoonright}(n+1)}\cap I_{t{\upharpoonright}(n+1)}. This easily implies that (IsQμ¯)(ItQμ¯)=(I_{s}\smallsetminus Q_{\bar{\mu}})\cap(I_{t}\smallsetminus Q_{\bar{\mu}})=\emptyset. ∎

Notice that, for s,tTs,t\in T, ItIsI_{t}\subseteq I_{s} when sts\subseteq t in TT, so asatbtasa_{s}\leq a_{t}\leq b_{t}\leq a_{s}. Then, for x[T]x\in[T], axn:n<ω\langle a_{x{\upharpoonright}n}\colon n<\omega\rangle is a monotone increasing sequence and bxn:n<ω\langle b_{x{\upharpoonright}n}\colon n<\omega\rangle is a monotone decreasing sequence (letting xn=xx{\upharpoonright}n=x in the case that xx has finite length and m|x|m\geq|x|) in [0,1][0,1], so their limits exist. This allows us to introduce the following definition.

Definition 6.4.

Define fμ¯:[T][0,1]f^{-}_{\bar{\mu}}\colon[T]\to[0,1] and fμ¯+:[T][0,1]f^{+}_{\bar{\mu}}\colon[T]\to[0,1] by fμ¯(x)limnaxn\displaystyle f^{-}_{\bar{\mu}}(x)\coloneqq\lim_{n\to\infty}a_{x{\upharpoonright}n} and fμ¯+(x)limnbxn\displaystyle f^{+}_{\bar{\mu}}(x)\coloneqq\lim_{n\to\infty}b_{x{\upharpoonright}n}, respectively. Furthermore, for each x[T]x\in[T], define Ixn<ωIxnI^{*}_{x}\coloneqq\bigcap_{n<\omega}I_{x{\upharpoonright}n} and, for A[T]A\subseteq[T], set I(A)xAIxI^{*}(A)\coloneqq\bigcup_{x\in A}I^{*}_{x}.

Clearly, fμ¯(x)fμ¯+(x)f^{-}_{\bar{\mu}}(x)\leq f^{+}_{\bar{\mu}}(x). Next, let us look at some basic properties of IxI_{x}^{\ast} and I([t])I^{\ast}([t]).

Lemma 6.5.

Let tTt\in T. Then:

  1. (a)

    For x[T]x\in[T], Ix=[fμ¯(x),fμ¯+(x)]I^{*}_{x}=[f^{-}_{\bar{\mu}}(x),f^{+}_{\bar{\mu}}(x)] and

    𝖫𝖻(Ix)=limnΞμ¯({xn})=n<|x|μxn({x(n+1)}).\operatorname{\mathsf{Lb}}(I^{*}_{x})=\lim_{n\to\infty}\Xi^{\bar{\mu}}(\{x{\upharpoonright}n\})=\prod_{n<|x|}\mu_{x{\upharpoonright}n}(\{x{\upharpoonright}(n+1)\}).
  2. (b)

    It{bt}I([t])ItI_{t}\smallsetminus\{b_{t}\}\subseteq I^{*}([t])\subseteq I_{t}. As a consequence, I([t])Qμ¯=ItQμ¯I^{*}([t])\smallsetminus Q_{\bar{\mu}}=I_{t}\smallsetminus Q_{\bar{\mu}}.

Proof..

(a): Since Ixn:n<ω\langle I_{x{\upharpoonright}n}\colon n<\omega\rangle is a decreasing sequence of closed intervals,

Ix=n<ωIxn=n<ω[axn,bxn]=[fμ¯(x),fμ¯+(x)]I^{*}_{x}=\bigcap_{n<\omega}I_{x{\upharpoonright}n}=\bigcap_{n<\omega}[a_{x{\upharpoonright}n},b_{x{\restriction}n}]=[f^{-}_{\bar{\mu}}(x),f^{+}_{\bar{\mu}}(x)]

and 𝖫𝖻(Ix)=limn𝖫𝖻(Ixn)\operatorname{\mathsf{Lb}}(I^{*}_{x})=\lim_{n\to\infty}\operatorname{\mathsf{Lb}}(I_{x{\upharpoonright}n}). The rest follows by 6.3 (b).

(b): Assume that yI([t])y\in I^{*}([t]), so yIx=n<ωIxny\in I^{*}_{x}=\bigcap_{n<\omega}I_{x{\upharpoonright}n} for some x[t]x\in[t]. Since x|t|=tx{\upharpoonright}|t|=t, we get yIx|t|=Ity\in I_{x{\upharpoonright}|t|}=I_{t}.

Now assume that yIt{bt}y\in I_{t}\smallsetminus\{b_{t}\}. By recursion on n|t|n\geq|t|, use 6.3 (c) to find tn𝖫𝗏n(T)t_{n}\in\operatorname{\mathsf{Lv}}_{n}(T) such that yItnQμ¯y\in I_{t_{n}}\smallsetminus Q_{\bar{\mu}} and tn+1tntt_{n+1}\supseteq t_{n}\supseteq t. This recursion can end when reaching a tnmaxTt_{n_{*}}\in\max T, in which case we set xtnx\coloneqq t_{n_{*}}. Otherwise, set xn|t|tnx\coloneqq\bigcup_{n\geq|t|}t_{n}. In both cases, x[t]x\in[t]. Thus, yn<ωIxn=Ixy\in\bigcap_{n<\omega}I_{x{\upharpoonright}n}=I^{*}_{x}, so yI([t])y\in I^{*}([t]) because x[t]x\in[t]. ∎

As a consequence of 6.3 and 6.5:

Corollary 6.6.

{IxQμ¯:x[T]}\{I_{x}^{\ast}\smallsetminus Q_{\bar{\mu}{}}\colon x\in[T]\} is a partition of [0,1]Qμ¯[0,1]\smallsetminus Q_{\bar{\mu}}. In particular, for any y[0,1]Qμ¯y\in[0,1]\smallsetminus Q_{\bar{\mu}}, there exists an unique x[T]x\in[T] such that yIxy\in I_{x}^{\ast}.

We can compute the functions fμ¯f_{\bar{\mu}}^{-} and fμ¯+f_{\bar{\mu}}^{+} for the probability trees from 4.4.

Example 6.7.
  1. (1)

    For tT=2<ωt\in T={}^{<\omega}2, It0,It1\langle I_{t{}^{\frown}\langle 0\rangle},I_{t{}^{\frown}\langle 1\rangle}\rangle results from splitting ItI_{t} in half. It can be proved that at=i<|t|t(i)2i+1a_{t}=\sum_{i<|t|}\frac{t(i)}{2^{i+1}}, bt=at+2|t|b_{t}=a_{t}+2^{-|t|} and, for x2ωx\in{}^{\omega}2, fμ¯(x)=fμ¯+(x)=i<ωx(i)2i+1f^{-}_{\bar{\mu}}(x)=f^{+}_{\bar{\mu}}(x)=\sum_{i<\omega}\frac{x(i)}{2^{i+1}}, hence IxI^{*}_{x} is a singleton.

  2. (2)

    For tT=ω<ωt\in T={}^{<\omega}\omega, It0I_{t{}^{\frown}\langle 0\rangle} is the first half of ItI_{t}, It1I_{t{}^{\frown}\langle 1\rangle} is the first half of [at0,bt][a_{t{}^{\frown}\langle 0\rangle},b_{t}], It2I_{t{}^{\frown}\langle 2\rangle} is the first half of [at1,bt][a_{t{}^{\frown}\langle 1\rangle},b_{t}] and, in general, It+1I_{t{}^{\frown}\langle\ell+1\rangle} is the first half of [at,bt][a_{t{}^{\frown}\langle\ell\rangle},b_{t}]. Therefore, <ωIt=It{bt}\bigcup_{\ell<\omega}I_{t{}^{\frown}\langle\ell\rangle}=I_{t}\smallsetminus\{b_{t}\}.

    It can be proved by induction on n=|t|n=|t| that 𝖫𝖻(It)=2|t|+k<|t|t(k)2|t|\operatorname{\mathsf{Lb}}(I_{t})=2^{-|t|+\sum_{k<|t|}t(k)}\leq 2^{-|t|}. As a consequence, IxI^{*}_{x} is a singleton for all xωωx\in{}^{\omega}\omega.

  3. (3)

    For tT=ω<ωt\in T={}^{<\omega}\omega,

    It={{at} if <5,It if =5,{bt} if >5.I_{t{}^{\frown}\langle\ell\rangle}=\left\{\begin{array}[]{ll}\{a_{t}\}&\text{ if $\ell<5$,}\\ I_{t}&\text{ if $\ell=5$,}\\ \{b_{t}\}&\text{ if $\ell>5$.}\end{array}\right.

    As a consequence, It=[0,1]I_{t}=[0,1] when tω<ωt\in{}^{<\omega}\omega is the constant sequence of 55. Otherwise, let n<|t|n<|t| be the minimum such that t(n)5t(n)\neq 5. If t(n)<5t(n)<5 then It={0}I_{t}=\{0\}, and if t(n)>5t(n)>5 then It={1}I_{t}=\{1\}. Therefore, Ix=[0,1]I^{*}_{x}=[0,1] when xωωx\in{}^{\omega}\omega is the constant sequence of 55. Otherwise, letting n<ωn<\omega be the minimum such that x(n)5x(n)\neq 5, if x(n)<5x(n)<5 then Ix={0}I^{*}_{x}=\{0\} and, if x(n)>5x(n)>5, then Ix={1}I^{*}_{x}=\{1\}.

In (1) and (2), Qμ¯Q_{\bar{\mu}} is dense in [0,1][0,1], while in (3) it is just {0,1}\{0,1\}.

Next, we analyze the conditions under which ata_{t} and btb_{t} take extreme values, that is, when at=0a_{t}=0 and bt=1b_{t}=1. For this, we use the lexicographic order: for s,tω<ωs,t\in{}^{<\omega}\omega, s<lexts<_{\mathrm{lex}}t iff there is some n<ωn<\omega such that sn=tns{\upharpoonright}n=t{\upharpoonright}n and s(n)<t(n)s(n)<t(n). Notice that <lex<_{\mathrm{lex}} is a partial order on ω<ω{}^{<\omega}\omega but not necessarily linear, e.g. comparable nodes in ω<ω{}^{<\omega}\omega are not <lex<_{\mathrm{lex}}-comparable. However, <lex<_{\mathrm{lex}} is linear at any level of ω<ω{}^{<\omega}\omega.

Lemma 6.8.

Let tTt\in T. Then:

  1. (a)

    If sTs\in T and s<lexts<_{\mathrm{lex}}t then bsatb_{s}\leq a_{t}.

  2. (b)

    at=0a_{t}=0 iff Ξμ¯({s})=0\Xi^{\bar{\mu}}(\{s\})=0 for all sTs\in T satisfying s<lexts<_{\mathrm{lex}}t.

  3. (c)

    bt=1b_{t}=1 iff Ξμ¯({s})=0\Xi^{\bar{\mu}}(\{s\})=0 for all sTs\in T satisfying t<lexst<_{\mathrm{lex}}s.

Proof..

(a): If s<lexts<_{\mathrm{lex}}t then there is some n<ωn<\omega such that sn=tns{\upharpoonright}n=t{\upharpoonright}n and s(n)<t(n)s(n)<t(n). This implies that bs(n+1)at(n+1)b_{s{\upharpoonright}(n+1)}\leq a_{t{\upharpoonright}(n+1)}. On the other hand, bsbs(n+1)b_{s}\leq b_{s{\upharpoonright}(n+1)} and at(n+1)ata_{t{\upharpoonright}(n+1)}\leq a_{t}, so bsatb_{s}\leq a_{t}.

Next, we only show (b), as (c) is similar. If at=0a_{t}=0, sTs\in T and s<lexts<_{\mathrm{lex}}t, then bsat=0b_{s}\leq a_{t}=0 by (a), so as=bs=0a_{s}=b_{s}=0, i.e. Ξμ¯({s})=bsas=0\Xi^{\bar{\mu}}(\{s\})=b_{s}-a_{s}=0 by 6.3 (b).

Conversely, assume that Ξμ¯({s})=0\Xi^{\bar{\mu}}(\{s\})=0 for all sTs\in T such that s<lexts<_{\mathrm{lex}}t. We show by induction on n|t|n\leq|t| that atn=0a_{t{\upharpoonright}n}=0. This is clear for n=0n=0. Now assume that n<|t|n<|t| and atn=0a_{t{\upharpoonright}n}=0. For every k<t(n)k<t(n), 𝖫𝖻(Itnk)=Ξμ¯({tnk})=0\operatorname{\mathsf{Lb}}(I_{t{\upharpoonright}n{}^{\frown}\langle k\rangle})=\Xi^{\bar{\mu}}(\{t{\upharpoonright}n{}^{\frown}\langle k\rangle\})=0 because tnk<lextt{\upharpoonright}n{}^{\frown}\langle k\rangle<_{\mathrm{lex}}t, so atnk=btnk=0a_{t{\upharpoonright}n{}^{\frown}\langle k\rangle}=b_{t{\upharpoonright}n{}^{\frown}\langle k\rangle}=0. By induction on kt(n)k\leq t(n), it is easy to show that atnk=atn=0a_{t{\upharpoonright}n{}^{\frown}\langle k\rangle}=a_{t{\upharpoonright}n}=0, so at(n+1)=0a_{t{\upharpoonright}(n+1)}=0. ∎

As a consequence, any element in Qμ¯{1}Q_{\bar{\mu}}\smallsetminus\{1\} has the form ata_{t} for some tTt\in T.

Corollary 6.9.

Qμ¯={at:tT}{1}Q_{\bar{\mu}}=\{a_{t}\colon t\in T\}\cup\{1\}.

Proof..

Let btQμ¯b_{t}\in Q_{\bar{\mu}} such that bt<1b_{t}<1. By 6.8 (c), we can find some sTs\in T such that t<lexst<_{\mathrm{lex}}s and Ξμ¯({s})>0\Xi^{\bar{\mu}}(\{s\})>0, i.e. bsas>0b_{s}-a_{s}>0 by 6.3 (b). Then, there is some m<|t|m<|t| such that tm=smt{\upharpoonright}m=s{\upharpoonright}m and t(m)<s(m)t(m)<s(m). We show by induction on m+1n|t|m+1\leq n\leq|t| that btnb_{t{\upharpoonright}n} has the form ata_{t^{\prime}} for some tTt^{\prime}\in T. In the case n=m+1n=m+1, since t(m)<s(m)<αtmt(m)<s(m)<\alpha_{t{\upharpoonright}m}, we get t(m)+1<αtmt(m)+1<\alpha_{t{\upharpoonright}m}, so bt(m+1)=atmt(m)+1b_{t{\upharpoonright}(m+1)}=a_{t{\upharpoonright}m{}^{\frown}\langle t(m)+1\rangle}. Now assume that m+1n<|t|m+1\leq n<|t| and btn=atb_{t{\upharpoonright}n}=a_{t^{\prime}} for some tTt^{\prime}\in T. If t(n)+1<αtnt(n)+1<\alpha_{t{\upharpoonright}n} then bt(n+1)=atnt(n)+1b_{t{\upharpoonright}(n+1)}=a_{t{\upharpoonright}n{}^{\frown}\langle t(n)+1\rangle}, otherwise, if t(n)+1=αtt(n)+1=\alpha_{t} then bt(n+1)=btn=atb_{t{\upharpoonright}(n+1)}=b_{t{\upharpoonright}n}=a_{t^{\prime}}. ∎

The construction of It:tT\langle I_{t}\colon t\in T\rangle yields the following important connection between [T][T] and [0,1][0,1].

Definition 6.10.

Define gμ¯:[0,1]Qμ¯[T]g_{\bar{\mu}}\colon[0,1]\smallsetminus Q_{\bar{\mu}}\to[T] such that, for any y[0,1]Qμ¯y\in[0,1]\smallsetminus Q_{\bar{\mu}}, gμ¯(y)g_{\bar{\mu}}(y) is the unique x[T]x\in[T] such that yIxy\in I^{*}_{x}.

Notice that gμ¯g_{\bar{\mu}} is well-defined by virtue of 6.6. Now, we will use it to connect T\mathcal{B}_{T} with ([0,1])\mathcal{B}([0,1]).

Lemma 6.11.

The map gμ¯g_{\bar{\mu}} is continuous. Moreover:

  1. (a)

    For any A[T]A\subseteq[T], gμ¯1[A]=I(A)Qμ¯g_{\bar{\mu}}^{-1}[A]=I^{*}(A)\smallsetminus Q_{\bar{\mu}}, and

  2. (b)

    If ATA\in\mathcal{B}_{T} then I(A)([0,1])I^{*}(A)\in\mathcal{B}([0,1]).

Proof..

The continuity of gμ¯g_{\bar{\mu}} follows from (a): by 6.3 (b), for any tTt\in T, gμ¯1[[t]]=I([t])Qμ¯=ItQμ¯g^{-1}_{\bar{\mu}}\big{[}[t]\big{]}=I^{*}([t])\smallsetminus Q_{\bar{\mu}}=I_{t}\smallsetminus Q_{\bar{\mu}}, which is an open set in [0,1]Qμ¯[0,1]\smallsetminus Q_{\bar{\mu}}.

(a): Let A[T]A\subseteq[T]. On the one hand, if yI(A)Qμ¯y\in I^{*}(A)\smallsetminus Q_{\bar{\mu}} then yIxQμ¯y\in I^{*}_{x}\smallsetminus Q_{\bar{\mu}} for some xAx\in A. By 6.10, x=gμ¯(y)x=g_{\bar{\mu}}(y), so ygμ¯1[A]y\in g^{-1}_{\bar{\mu}}[A]. On the other hand, if ygμ¯1[A]y\in g^{-1}_{\bar{\mu}}[A] then y[0,1]Qμ¯y\in[0,1]\smallsetminus Q_{\bar{\mu}} and xgμ¯(y)Ax\coloneqq g_{\bar{\mu}}(y)\in A, so yIxy\in I^{*}_{x} and, thus, yI(A)Qμ¯y\in I^{*}(A)\smallsetminus Q_{\bar{\mu}}.

(b): By 2.4, gμ¯g_{\bar{\mu}} is a Borel map. Then, for any ATA\in\mathcal{B}_{T}, I(A)Qμ¯([0,1]Qμ¯)I^{*}(A)\smallsetminus Q_{\bar{\mu}}\in\mathcal{B}([0,1]\smallsetminus Q_{\bar{\mu}}). Also notice that, by 2.3, ([0,1]Qμ¯)=([0,1])|[0,1]Qμ¯\mathcal{B}([0,1]\smallsetminus Q_{\bar{\mu}})=\mathcal{B}([0,1])|_{[0,1]\smallsetminus Q_{\bar{\mu}}}, so there is some B([0,1])B\in\mathcal{B}([0,1]) such that I(A)Qμ¯=BQμ¯I^{*}(A)\smallsetminus Q_{\bar{\mu}}=B\smallsetminus Q_{\bar{\mu}}. But Qμ¯([0,1])Q_{\bar{\mu}}\in\mathcal{B}([0,1]) because it is countable, so BQμ¯([0,1])B\smallsetminus Q_{\bar{\mu}}\in\mathcal{B}([0,1]), that is, I(A)Qμ¯([0,1])I^{*}(A)\smallsetminus Q_{\bar{\mu}}\in\mathcal{B}([0,1]). On the other hand, I(A)Qμ¯([0,1])I^{*}(A)\cap Q_{\bar{\mu}}\in\mathcal{B}([0,1]) because it is countable, so I(A)=[I(A)Qμ¯][I(A)Qμ¯]I^{*}(A)=[I^{*}(A)\smallsetminus Q_{\bar{\mu}}]\cup[I^{*}(A)\cap Q_{\bar{\mu}}] is Borel in [0,1][0,1]. ∎

We are ready to introduce a more precise definition of λμ¯\lambda^{\bar{\mu}}.

Definition 6.12.

Define λμ¯:T[0,1]\lambda^{\bar{\mu}}\colon\mathcal{B}_{T}\to[0,1] by λμ¯(A)𝖫𝖻(I(A))\lambda^{\bar{\mu}}(A)\coloneqq\operatorname{\mathsf{Lb}}(I^{*}(A)), which is well-defined by 6.11 (b).

Although this definition uses the representation of TT that we fixed, we will show in 6.15 that λμ¯\lambda^{\bar{\mu}} does not depend on the representation of TT.

Theorem 6.13.

The map λμ¯\lambda^{\bar{\mu}} from 6.12 is a probability measure on T\mathcal{B}_{T} such that λμ¯([t])=Ξμ¯({t})\lambda^{\bar{\mu}}([t])=\Xi^{\bar{\mu}}(\{t\}) for all tTt\in T.

Proof..

By 2.8, [T],𝒜,λ\langle[T],\mathcal{A}^{\prime},\lambda^{\prime}\rangle is a measure space, where 𝒜gμ¯(([0,1]Qμ¯))\mathcal{A}^{\prime}\coloneqq g^{\to}_{\bar{\mu}}(\mathcal{B}([0,1]\smallsetminus Q_{\bar{\mu}})) and λ(A)𝖫𝖻(gμ¯1[A])\lambda^{\prime}(A)\coloneqq\operatorname{\mathsf{Lb}}(g^{-1}_{\bar{\mu}}[A]) for any A𝒜A\in\mathcal{A}^{\prime}. Since gμ¯g_{\bar{\mu}} is a Borel function, T𝒜\mathcal{B}_{T}\subseteq\mathcal{A}^{\prime}. Then, for any ATA\in\mathcal{B}_{T}, since Qμ¯Q_{\bar{\mu}} is countable,

λ(A)=𝖫𝖻(gμ¯1[A])=𝖫𝖻(I(A)Qμ¯)=𝖫𝖻(I(A))=λμ¯(A).\lambda^{\prime}(A)=\operatorname{\mathsf{Lb}}(g^{-1}_{\bar{\mu}}[A])=\operatorname{\mathsf{Lb}}(I^{*}(A)\smallsetminus Q_{\bar{\mu}})=\operatorname{\mathsf{Lb}}(I^{*}(A))=\lambda^{\bar{\mu}}(A).

As a consequence, λμ¯=λT\lambda^{\bar{\mu}}=\lambda^{\prime}{\upharpoonright}\mathcal{B}_{T}, which is a measure. On the other hand, by 6.3, for any tTt\in T, λμ¯([t])=λ([t])=𝖫𝖻(I([t])Qμ¯)=𝖫𝖻(ItQμ¯)=𝖫𝖻(It)=Ξμ¯({t}),\lambda^{\bar{\mu}}([t])=\lambda^{\prime}([t])=\operatorname{\mathsf{Lb}}(I^{*}([t])\smallsetminus Q_{\bar{\mu}})=\operatorname{\mathsf{Lb}}(I_{t}\smallsetminus Q_{\bar{\mu}})=\operatorname{\mathsf{Lb}}(I_{t})=\Xi^{\bar{\mu}}(\{t\}), as required. In particular, λμ¯([T])=λμ¯([])=𝖫𝖻(I)=1\lambda^{\bar{\mu}}([T])=\lambda^{\bar{\mu}}([\langle\ \rangle])=\operatorname{\mathsf{Lb}}(I_{\langle\ \rangle})=1. Thus, λμ¯\lambda^{\bar{\mu}} is a probability measure. ∎

Remark 6.14.

In the previous proof, we do not always have 𝒜=T\mathcal{A}^{\prime}=\mathcal{B}_{T}, neither the equivalent formulation “I(A)([0,1])I^{*}(A)\in\mathcal{B}([0,1]) implies ATA\in\mathcal{B}_{T}”. For example, when Ξμ¯({t})=0\Xi^{\bar{\mu}}(\{t\})=0 for some tTt\in T such that [t][t] is uncountable, λμ¯([t])=0\lambda^{\bar{\mu}}([t])=0 and I([t])=ItI^{*}([t])=I_{t} is just a singleton, and then I(A)=It([0,1])I^{*}(A)=I_{t}\in\mathcal{B}([0,1]) for every non-empty A[t]A\subseteq[t], and there are many such AA that are not Borel in [T][T]. This is not much of a problem anyway because the converse holds for the completions of [T],T,λμ¯\langle[T],\mathcal{B}_{T},\lambda^{\bar{\mu}}\rangle and [0,1],([0,1]),𝖫𝖻\langle[0,1],\mathcal{B}([0,1]),\operatorname{\mathsf{Lb}}\rangle. For more details, see Theorem 6.19, 6.20, and 6.23.

Although it is possible that λμ¯([t])=0\lambda^{\bar{\mu}}([t])=0 for some tTt\in T, this is a very undesirable situation that is typically avoided. When λμ¯([t])>0\lambda^{\bar{\mu}}([t])>0 for all tTt\in T and all singletons in [T][T] have measure zero (which implies [T]=limT[T]=\lim T), the equivalence discussed above will hold (see Theorem 6.28).

The uniqueness of λμ¯\lambda^{\bar{\mu}} in Theorem 4.31 simply follows by Theorem 2.10. Therefore:

Corollary 6.15.

The measure λμ¯\lambda^{\bar{\mu}} does not depend on the representation of TT.

The construction of λμ¯\lambda^{\bar{\mu}} uses the Lebesgue measure on [0,1][0,1] and, unlike the first construction in Theorem 4.31, does not rely on Theorem 4.15, neither on its consequences. Using the second construction, we can prove Theorem 4.15 more directly.

Second proof of Theorem 4.15.

Let Ξ𝒫\Xi\in\mathcal{IP}, TTΞT\coloneqq T_{\Xi} and let AA be a front of TT. Since Π\Pi is surjective, there is some μ¯𝒯𝒫\bar{\mu}\in\mathcal{TP} such that Ξ=Ξμ¯\Xi=\Xi^{\bar{\mu}}. Hence, Ξλμ¯=Ξμ¯\Xi^{\lambda^{\bar{\mu}}}=\Xi^{\bar{\mu}} by Theorem 6.13.

Let tTt\in T below AA and At{sA:ts}A_{\geq t}\coloneqq\{s\in A\colon t\leq s\}. It is enough to show that Ξ({t})=Ξ(At)\Xi(\{t\})=\Xi(A_{\geq t}). Since AA is a front and tt is below AA, we obtain that [t]=sAt[s][t]=\bigcup_{s\in A_{\geq t}}[s] is a disjoint union, so

Ξ({t})=λμ¯([t])=sAtλμ¯([s])=sAtΞ({s})=Ξ(At).\Xi(\{t\})=\lambda^{\bar{\mu}}([t])=\sum_{s\in A_{\geq t}}\lambda^{\bar{\mu}}([s])=\sum_{s\in A_{\geq t}}\Xi(\{s\})=\Xi(A_{\geq t}).\qed

The function gμ¯g_{\bar{\mu}} has more properties than it appears to have. Under certain conditions, it is a topological embedding into [T][T]. To construct the inverse function, we look at the measure zero points of [T][T].

Lemma 6.16.

Let SS be a subtree of TT with maxSmaxT\max S\subseteq\max T, and let x[T]x\in[T]. Then:

  1. (a)

    λμ¯([S])=infn<ht(T)Ξμ¯(Frn(S))\displaystyle\lambda^{\bar{\mu}}([S])=\inf_{n<\mathrm{ht}(T)}\Xi^{\bar{\mu}}(\mathrm{Fr}_{n}(S)).

  2. (b)

    λμ¯({x})=𝖫𝖻(Ix)\lambda^{\bar{\mu}}(\{x\})=\operatorname{\mathsf{Lb}}(I^{*}_{x}).

  3. (c)

    λμ¯({x})=0\lambda^{\bar{\mu}}(\{x\})=0 iff IxI^{*}_{x} is a singleton.

Proof..

(a): For any n<ht(T)n<\mathrm{ht}(T), consider Cn{[t]T:tFrn(S)}C_{n}\coloneqq\bigcup\{[t]_{T}\colon t\in\mathrm{Fr}_{n}(S)\}, which is a pairwise disjoint union. Then Cn:n<ht(T)\langle C_{n}\colon n<\mathrm{ht}(T)\rangle is a decreasing sequence of clopen sets in [T][T] whose intersection is [S][S]. Furthermore, by Theorem 6.13, it is clear that, for any n<ht(T)n<\mathrm{ht}(T), λμ¯(Cn)=Ξμ¯(Frn(S))\lambda^{\bar{\mu}}(C_{n})=\Xi^{\bar{\mu}}(\mathrm{Fr}_{n}(S)). As a consequence,

λμ¯([S])=λμ¯(n<ht(T)Cn)=infn<ht(T)λμ¯(Cn)=infn<ht(T)Ξμ¯(Frn(S)).\lambda^{\bar{\mu}}(\left[S]\right)=\lambda^{\bar{\mu}}\left(\bigcap_{n<\mathrm{ht}(T)}C_{n}\right)=\inf_{n<\mathrm{ht}(T)}\lambda^{\bar{\mu}}(C_{n})=\inf_{n<\mathrm{ht}(T)}\Xi^{\bar{\mu}}(\mathrm{Fr}_{n}(S)).

(b): By the definition of λμ¯\lambda^{\bar{\mu}}, λμ¯({x})=𝖫𝖻(I({x}))=𝖫𝖻(Ix).\lambda^{\bar{\mu}}(\{x\})=\operatorname{\mathsf{Lb}}(I^{\ast}(\{x\}))=\operatorname{\mathsf{Lb}}(I_{x}^{\ast}).

(c): On the one hand, if IxI_{x}^{\ast} is a singleton, then by (b) λμ¯({x})=𝖫𝖻(Ix)=0\lambda^{\bar{\mu}}(\{x\})=\operatorname{\mathsf{Lb}}(I_{x}^{\ast})=0. On the other hand, by 6.5 (a), IxI_{x}^{\ast} is a closed interval, namely [fμ¯(x),fμ¯+(x)][f_{\bar{\mu}}^{-}(x),f_{\bar{\mu}}^{+}(x)]. Therefore, if λμ¯({x})=0\lambda^{\bar{\mu}}(\{x\})=0, then 𝖫𝖻(Ix)=0\operatorname{\mathsf{Lb}}(I_{x}^{\ast})=0, so IxI_{x}^{\ast} must be a singleton. ∎

Using 6.16 (c), we can introduce a sort of inverse of gμ¯g_{\bar{\mu}} as follows (inverse in the sense of 6.18 (g)).

Definition 6.17.
  1. (1)

    For λ𝒫\lambda\in\mathcal{BP}, define Vλ{x[Tλ]:λ({x})=0}V^{*}_{\lambda}\coloneqq\{x\in[T_{\lambda}]\colon\lambda(\{x\})=0\}, the free part of [Tλ][T_{\lambda}].

  2. (2)

    Let fμ¯f_{\bar{\mu}} be the function with domain Vμ¯Vλμ¯V^{*}_{\bar{\mu}}\coloneqq V^{*}_{\lambda^{\bar{\mu}}} such that, for xVμ¯x\in V^{*}_{\bar{\mu}}, fμ¯(x)f_{\bar{\mu}}(x) is the unique point in IxI^{*}_{x}. Notice that Vμ¯V^{*}_{\bar{\mu}} could be empty.

  3. (3)

    For A[T]A\subseteq[T], define GAxAVμ¯(fμ¯(x),fμ¯+(x))G^{\ast}_{A}\coloneqq\bigcup_{x\in A\smallsetminus V^{*}_{\bar{\mu}}}\left(f_{\bar{\mu}}^{-}(x),f_{\bar{\mu}}^{+}(x)\right) and Gμ¯G[T]G^{\ast}_{\bar{\mu}}\coloneqq G^{*}_{[T]}, which are open in [0,1][0,1].

  4. (4)

    Denote Nμ¯Nλμ¯N^{*}_{\bar{\mu}}\coloneqq N^{*}_{\lambda^{\bar{\mu}}} (see 4.34), which is the largest open measure zero subset of [T][T] (see 4.35). Recall that [Tμ¯+]=[T]Nμ¯[T^{+}_{\bar{\mu}}]=[T]\smallsetminus N^{*}_{\bar{\mu}}, so it has measure 11.

Lemma 6.18.
  1. (a)

    [T]Vμ¯[T]\smallsetminus V^{*}_{\bar{\mu}} is countable. In particular, Vμ¯TV^{*}_{\bar{\mu}}\in\mathcal{B}_{T}.

  2. (b)

    For y[0,1]Qμ¯y\in[0,1]\smallsetminus Q_{\bar{\mu}}, if gμ¯(y)Vμ¯g_{\bar{\mu}}(y)\in V^{*}_{\bar{\mu}} then fμ¯(gμ¯(y))=yf_{\bar{\mu}}(g_{\bar{\mu}}(y))=y.

  3. (c)

    If xVμ¯x\in V^{*}_{\bar{\mu}} and fμ¯(x)Qμ¯f_{\bar{\mu}}(x)\notin Q_{\bar{\mu}} then gμ¯(fμ¯(x))=xg_{\bar{\mu}}(f_{\bar{\mu}}(x))=x.

  4. (d)

    fμ¯f_{\bar{\mu}} is continuous and (Qμ¯ranfμ¯)Gμ¯=(Q_{\bar{\mu}}\cup\operatorname{ran}f_{\bar{\mu}})\cap G^{*}_{\bar{\mu}}=\emptyset.

  5. (e)

    rangμ¯=[T]fμ¯1[Qμ¯]\operatorname{ran}g_{\bar{\mu}}=[T]\smallsetminus f^{-1}_{\bar{\mu}}[Q_{\bar{\mu}}]. In particular, rangμ¯T\operatorname{ran}g_{\bar{\mu}}\in\mathcal{B}_{T}.

  6. (f)

    fμ¯[rangμ¯]=ranfμ¯Qμ¯f_{\bar{\mu}}[\operatorname{ran}g_{\bar{\mu}}]=\operatorname{ran}f_{\bar{\mu}}\smallsetminus Q_{\bar{\mu}}.

  7. (g)

    fμ¯rangμ¯f_{\bar{\mu}}{\upharpoonright}\operatorname{ran}g_{\bar{\mu}} is an homeomorphism from Vμ¯f1[Qμ¯]V^{*}_{\bar{\mu}}\smallsetminus f^{-1}[Q_{\bar{\mu}}] onto ranfμ¯Qμ¯\operatorname{ran}f_{\bar{\mu}}\smallsetminus Q_{\bar{\mu}} with inverse gμ¯(ranfμ¯Qμ¯)g_{\bar{\mu}}{\upharpoonright}(\operatorname{ran}f_{\bar{\mu}}\smallsetminus Q_{\bar{\mu}}).

  8. (h)

    Nμ¯fμ¯1[Qμ¯]N^{*}_{\bar{\mu}}\subseteq f^{-1}_{\bar{\mu}}[Q_{\bar{\mu}}] and maxTNμ¯([T]Vμ¯)\max T\subseteq N^{*}_{\bar{\mu}}\cup([T]\smallsetminus V^{*}_{\bar{\mu}}), i.e. Vμ¯Nμ¯limTV^{*}_{\bar{\mu}}\smallsetminus N^{*}_{\bar{\mu}}\subseteq\lim T.

  9. (i)

    fμ¯1[Qμ¯]Nμ¯f^{-1}_{\bar{\mu}}[Q_{\bar{\mu}}]\smallsetminus N^{*}_{\bar{\mu}} is countable and λμ¯(fμ¯1[Qμ¯])=0\lambda^{\bar{\mu}}(f^{-1}_{\bar{\mu}}[Q_{\bar{\mu}}])=0.

  10. (j)

    fμ¯[A]=I(AVμ¯)f_{\bar{\mu}}[A]=I^{*}(A\cap V^{*}_{\bar{\mu}}) for all A[T]A\subseteq[T].

Nμ¯N_{\bar{\mu}}^{\ast}[T][T][0,1][0,1]fμ¯f_{\bar{\mu}}gμ¯g_{\bar{\mu}}rangμ¯\mathrm{ran}g_{\bar{\mu}}domgμ¯\mathrm{dom}g_{\bar{\mu}}fμ¯1[Qμ¯]f_{\bar{\mu}}^{-1}[Q_{\bar{\mu}}]Qμ¯Q_{\bar{\mu}}Gμ¯G_{\bar{\mu}}^{\ast}Vμ¯V^{*}_{\bar{\mu}}ranfμ¯\operatorname{ran}f_{\bar{\mu}}
Figure 5. Graphic situation of 6.18: Nμ¯fμ¯1[Qμ¯]N_{\bar{\mu}}^{\ast}\subseteq f_{\bar{\mu}}^{-1}[Q_{\bar{\mu}}], ranfμ¯\operatorname{ran}f_{\bar{\mu}} may include some elements of Qμ¯Q_{\bar{\mu}}, and the shaded regions are homeomorphic via fμ¯rangμ¯f_{\bar{\mu}}{\restriction}\operatorname{ran}g_{\bar{\mu}}, whose inverse is gμ¯ranfμ¯g_{\bar{\mu}}{\restriction}\operatorname{ran}f_{\bar{\mu}}.
Proof..

(a): This is a direct consequence of 2.9.

(b): By the definition of gμ¯g_{\bar{\mu}}, yIgμ¯(y)y\in I^{*}_{g_{\bar{\mu}}(y)} for y[0,1]Qμ¯y\in[0,1]\smallsetminus Q_{\bar{\mu}}, but Igμ¯(y)={fμ¯(gμ¯(y))}I^{*}_{g_{\bar{\mu}}(y)}=\{f_{\bar{\mu}}(g_{\bar{\mu}}(y))\} when gμ¯(y)Vμ¯g_{\bar{\mu}}(y)\in V^{*}_{\bar{\mu}}, so fμ¯(gμ¯(y))=yf_{\bar{\mu}}(g_{\bar{\mu}}(y))=y.

(c): If xVμ¯x\in V^{*}_{\bar{\mu}} and fμ¯(x)Qμ¯f_{\bar{\mu}}(x)\notin Q_{\bar{\mu}} then fμ¯(x)domgμ¯f_{\bar{\mu}}(x)\in\operatorname{dom}g_{\bar{\mu}}. Also fμ¯(x)Ixf_{\bar{\mu}}(x)\in I^{*}_{x}, so gμ¯(fμ¯(x))=xg_{\bar{\mu}}(f_{\bar{\mu}}(x))=x.

(d): Let xVμ¯x\in V^{*}_{\bar{\mu}} and ε>0\varepsilon>0. Since {fμ¯(x)}=Ix=n<ωIxn(fμ¯(x)ε,fμ¯(x)+ε)\{f_{\bar{\mu}}(x)\}=I^{*}_{x}=\bigcap_{n<\omega}I_{x{\upharpoonright}n}\subseteq(f_{\bar{\mu}}(x)-\varepsilon,f_{\bar{\mu}}(x)+\varepsilon), there is some n<ωn<\omega such that It(fμ¯(x)ε,fμ¯(x)+ε)I_{t}\subseteq(f_{\bar{\mu}}(x)-\varepsilon,f_{\bar{\mu}}(x)+\varepsilon), where txnt\coloneqq x{\upharpoonright}n. Then, for any x[t]Vμ¯x^{\prime}\in[t]\cap V^{*}_{\bar{\mu}}, fμ¯(x)I([t])Itf_{\bar{\mu}}(x^{\prime})\in I^{*}([t])\subseteq I_{t}, so |fμ¯(x)fμ¯(x)|<2ε|f_{\bar{\mu}}(x)-f_{\bar{\mu}}(x^{\prime})|<2\varepsilon. This shows that fμ¯f_{\bar{\mu}} is continuous.

Notice that (fμ¯(x),fμ¯+(x)):x[T]Vμ¯\langle(f_{\bar{\mu}}(x),f^{+}_{\bar{\mu}}(x))\colon x\in[T]\smallsetminus V^{*}_{\bar{\mu}}\rangle is a sequence of pairwise disjoint intervals that cannot contain points in Qμ¯Q_{\bar{\mu}}, and thus neither in ranfμ¯\operatorname{ran}f_{\bar{\mu}}. Hence, (Qμ¯ranfμ¯)Gμ¯=(Q_{\bar{\mu}}\cup\operatorname{ran}f_{\bar{\mu}})\cap G^{*}_{\bar{\mu}}=\emptyset.

(e): If xrangμ¯x\in\operatorname{ran}g_{\bar{\mu}} then x=gμ¯(y)x=g_{\bar{\mu}}(y) for some y[0,1]Qμ¯y\in[0,1]\smallsetminus Q_{\bar{\mu}}. Either xVμ¯x\notin V^{*}_{\bar{\mu}} or xVμ¯x\in V^{*}_{\bar{\mu}}, and in the latter case, fμ¯(x)=y[0,1]Qμ¯f_{\bar{\mu}}(x)=y\in[0,1]\smallsetminus Q_{\bar{\mu}} by (b). Thus, x[T]fμ¯1[Qμ¯]x\in[T]\smallsetminus f^{-1}_{\bar{\mu}}[Q_{\bar{\mu}}]. On the other hand, assume that x[T]fμ¯1[Qμ¯]x\in[T]\smallsetminus f^{-1}_{\bar{\mu}}[Q_{\bar{\mu}}]. If xVμ¯x\notin V^{*}_{\bar{\mu}} then 𝖫𝖻(Ix)>0\operatorname{\mathsf{Lb}}(I^{*}_{x})>0, so IxQμ¯I^{*}_{x}\nsubseteq Q_{\bar{\mu}}, i.e. there is some yIxQμ¯y\in I^{*}_{x}\smallsetminus Q_{\bar{\mu}}. Thus, x=gμ¯(y)rangμ¯x=g_{\bar{\mu}}(y)\in\operatorname{ran}g_{\bar{\mu}}.

In the case that xVμ¯x\in V^{*}_{\bar{\mu}}, we have yfμ¯(x)Qμ¯y\coloneqq f_{\bar{\mu}}(x)\notin Q_{\bar{\mu}}, so x=gμ¯(y)rangμ¯x=g_{\bar{\mu}}(y)\in\operatorname{ran}g_{\bar{\mu}} by (c).

(f): If yfμ¯[rangμ¯]y\in f_{\bar{\mu}}[\operatorname{ran}g_{\bar{\mu}}] then y=fμ¯(x)y=f_{\bar{\mu}}(x) for some xVμ¯rangμ¯x\in V^{*}_{\bar{\mu}}\cap\operatorname{ran}g_{\bar{\mu}}, so xfμ¯1[Qμ¯]x\notin f^{-1}_{\bar{\mu}}[Q_{\bar{\mu}}], i.e. y=fμ¯(x)ranfμ¯Qμ¯y=f_{\bar{\mu}}(x)\in\operatorname{ran}f_{\bar{\mu}}\smallsetminus Q_{\bar{\mu}}.

For the converse, if yranfμ¯Qμ¯y\in\operatorname{ran}f_{\bar{\mu}}\smallsetminus Q_{\bar{\mu}} then y=fμ¯(x)y=f_{\bar{\mu}}(x) for some xVμ¯x\in V^{*}_{\bar{\mu}} and fμ¯(x)Qμ¯f_{\bar{\mu}}(x)\notin Q_{\bar{\mu}}, i.e. xfμ¯1[Qμ¯]x\notin f^{-1}_{\bar{\mu}}[Q_{\bar{\mu}}], so xrangμ¯x\in\operatorname{ran}g_{\bar{\mu}} by (e). Hence, yfμ¯[rangμ¯]y\in f_{\bar{\mu}}[\operatorname{ran}g_{\bar{\mu}}].

(g): Recall that Vμ¯fμ¯1[Qμ¯]=Vμ¯rangμ¯V^{*}_{\bar{\mu}}\smallsetminus f^{-1}_{\bar{\mu}}[Q_{\bar{\mu}}]=V^{*}_{\bar{\mu}}\cap\operatorname{ran}g_{\bar{\mu}} by (e). If xVμ¯fμ¯1[Qμ¯]x\in V^{*}_{\bar{\mu}}\smallsetminus f^{-1}_{\bar{\mu}}[Q_{\bar{\mu}}] then fμ¯(x)Qμ¯f_{\bar{\mu}}(x)\notin Q_{\bar{\mu}}, so gμ¯(fμ¯(x))g_{\bar{\mu}}(f_{\bar{\mu}}(x)) is defined and it is equal to xx by (c). Conversely, if yranfμ¯Qμ¯y\in\operatorname{ran}f_{\bar{\mu}}\smallsetminus Q_{\bar{\mu}} then y=fμ¯(x)y=f_{\bar{\mu}}(x) for some xVμ¯rangμ¯x\in V^{*}_{\bar{\mu}}\cap\operatorname{ran}g_{\bar{\mu}} by (f), so x=gμ¯(y)x=g_{\bar{\mu}}(y^{\prime}) for some y[0,1]Qμ¯y^{\prime}\in[0,1]\smallsetminus Q_{\bar{\mu}}. By (b), y=fμ¯(gμ¯(y))=fμ¯(x)=yy^{\prime}=f_{\bar{\mu}}(g_{\bar{\mu}}(y^{\prime}))=f_{\bar{\mu}}(x)=y, so fμ¯(gμ¯(y))=yf_{\bar{\mu}}(g_{\bar{\mu}}(y))=y. This shows that fμ¯rangμ¯f_{\bar{\mu}}{\upharpoonright}\operatorname{ran}g_{\bar{\mu}} is a bijection from Vμ¯fμ¯1[Qμ¯]V^{*}_{\bar{\mu}}\smallsetminus f^{-1}_{\bar{\mu}}[Q_{\bar{\mu}}] onto ranfμ¯Qμ¯\operatorname{ran}f_{\bar{\mu}}\smallsetminus Q_{\bar{\mu}} with inverse gμ¯(ranfμ¯Qμ¯)g_{\bar{\mu}}{\upharpoonright}(\operatorname{ran}f_{\bar{\mu}}\smallsetminus Q_{\bar{\mu}}). Since both fμ¯f_{\bar{\mu}} and gμ¯g_{\bar{\mu}} are continuous, we are done.

(h): If xNμ¯x\in N^{*}_{\bar{\mu}} then x[t]x\in[t] for some tTt\in T such that λμ¯([t])=0\lambda^{\bar{\mu}}([t])=0. Then λμ¯({x})=0\lambda^{\bar{\mu}}(\{x\})=0, so xVμ¯x\in V^{*}_{\bar{\mu}}. Moreover, {fμ¯(x)}=Ix=It\{f_{\bar{\mu}}(x)\}=I^{*}_{x}=I_{t}, so fμ¯(x)Qμ¯f_{\bar{\mu}}(x)\in Q_{\bar{\mu}}.

On the other hand, if xmaxTx\in\max T and xNμ¯x\notin N^{*}_{\bar{\mu}}, then [x]={x}[x]=\{x\} and λμ¯({x})=λμ¯([x])>0\lambda^{\bar{\mu}}(\{x\})=\lambda^{\bar{\mu}}([x])>0, so xVμ¯x\notin V^{*}_{\bar{\mu}}.

(i): It is enough to show that, for yQμ¯y\in Q_{\bar{\mu}}, |fμ¯1[{y}]Nμ¯|2|f^{-1}_{\bar{\mu}}[\{y\}]\smallsetminus N^{*}_{\bar{\mu}}|\leq 2. Let xfμ¯1[{y}]Nμ¯x\in f^{-1}_{\bar{\mu}}[\{y\}]\smallsetminus N^{*}_{\bar{\mu}}, so xlimTμ¯+x\in\lim T_{\bar{\mu}}^{+} (by (h)) and fμ¯(x)=yQμ¯f_{\bar{\mu}}(x)=y\in Q_{\bar{\mu}}. By the definition of fμ¯f_{\bar{\mu}}, Ix={y}I^{*}_{x}=\{y\}, so axnya_{x{\upharpoonright}n}\nearrow y and bxnyb_{x{\upharpoonright}n}\searrow y. Since yQμ¯y\in Q_{\bar{\mu}}, we must have that, for some m<ωm<\omega, either axn=ya_{x{\upharpoonright}n}=y for all nmn\geq m, or bxn=yb_{x{\upharpoonright}n}=y for all nmn\geq m. We show that at most only one xfμ¯1[{y}]Nμ¯x\in f_{\bar{\mu}}^{-1}[\{y\}]\smallsetminus N_{\bar{\mu}}^{\ast} satisfies that axn=ya_{x{\upharpoonright}n}=y for all but finitely many nn. Likewise, there is at most one xx satisfying bxn=yb_{x{\upharpoonright}n}=y for all but finitely many nn, so |fμ¯1[{y}][Tμ¯+]|2|f^{-1}_{\bar{\mu}}[\{y\}]\cap[T_{\bar{\mu}}^{+}]|\leq 2.

Assume that x,xfμ¯1[{y}]Nμ¯x,x^{\prime}\in f^{-1}_{\bar{\mu}}[\{y\}]\smallsetminus N^{*}_{\bar{\mu}} satisfy that, for some m<ωm^{\prime}<\omega, axn=axn=ya_{x{\upharpoonright}n}=a_{x^{\prime}{\upharpoonright}n}=y for all nmn\geq m^{\prime}. If xxx\neq x^{\prime} then let n0<ωn_{0}<\omega be such that xn0=xn0x{\upharpoonright}n_{0}=x{\upharpoonright}n_{0} and x(n0)x(n0)x(n_{0})\neq x^{\prime}(n_{0}). Without loss of generality, x(n0)<x(n0)x(n_{0})<x^{\prime}(n_{0}). Since x,x[Tμ¯+]x,x^{\prime}\in[T_{\bar{\mu}}^{+}], x(n0+1),x(n0+1)Tμ¯+x{\upharpoonright}(n_{0}+1),x^{\prime}{\upharpoonright}(n_{0}+1)\in T_{\bar{\mu}}^{+}, so ax(n0+1)<bx(n0+1)ax(n0+1)a_{x{\upharpoonright}(n_{0}+1)}<b_{x{\upharpoonright}(n_{0}+1)}\leq a_{x^{\prime}{\upharpoonright}(n_{0}+1)}. Then, for some n>max{m,n0}n>\max\{m^{\prime},n_{0}\}, y=axn<bxnaxny=a_{x{\upharpoonright}n}<b_{x{\upharpoonright}n}\leq a_{x^{\prime}{\upharpoonright}n}, so y<axny<a_{x^{\prime}{\upharpoonright}n}, a contradiction.

Finally, since fμ¯1[Qμ¯]Nμ¯f_{\bar{\mu}}^{-1}[Q_{\bar{\mu}}]\smallsetminus N^{*}_{\bar{\mu}} is countable and contained in Vμ¯V^{*}_{\bar{\mu}},

λμ¯(fμ¯1[Qμ¯])=λμ¯(fμ¯1[Qμ¯]Nμ¯)+λμ¯(Nμ¯)=0.\lambda^{\bar{\mu}}(f^{-1}_{\bar{\mu}}[Q_{\bar{\mu}}])=\lambda^{\bar{\mu}}(f_{\bar{\mu}}^{-1}[Q_{\bar{\mu}}]\smallsetminus N^{*}_{\bar{\mu}})+\lambda^{\bar{\mu}}(N^{*}_{\bar{\mu}})=0.

(j): By the definition of fμ¯f_{\bar{\mu}}, we have fμ¯[A]=xAVμ¯Ix=I(AVμ¯).f_{\bar{\mu}}[A]=\bigcup_{x\in A\cap V^{*}_{\bar{\mu}}}I_{x}^{\ast}=I^{\ast}(A\cap V^{*}_{\bar{\mu}}).

Now we analyze the effect of applying the functions fμ¯f_{\bar{\mu}} and gμ¯g_{\bar{\mu}} to Borel sets and their respective measures.

Theorem 6.19.

Let A[T]A\subseteq[T] and B[0,1]B\subseteq[0,1]. Then:

  1. (a)

    If ATA\in\mathcal{B}_{T}, then fμ¯[A]([0,1])f_{\bar{\mu}}[A]\in\mathcal{B}([0,1]). Furthermore, 𝖫𝖻(fμ¯[A])=λμ¯(AVμ¯)\operatorname{\mathsf{Lb}}(f_{\bar{\mu}}[A])=\lambda^{\bar{\mu}}(A\cap V^{*}_{\bar{\mu}}) and

    λμ¯(A)=𝖫𝖻(fμ¯[A])+𝖫𝖻(GA)=𝖫𝖻(fμ¯[A])+xAdom(fμ¯)𝖫𝖻(Ix).\lambda^{\bar{\mu}}(A)=\operatorname{\mathsf{Lb}}(f_{\bar{\mu}}[A])+\operatorname{\mathsf{Lb}}(G^{*}_{A})=\operatorname{\mathsf{Lb}}(f_{\bar{\mu}}[A])+\sum_{x\in A\smallsetminus\operatorname{dom}(f_{\bar{\mu}})}\operatorname{\mathsf{Lb}}(I_{x}^{\ast}).
  2. (b)

    If B([0,1])B\in\mathcal{B}([0,1]) then fμ¯1[B]Tf_{\bar{\mu}}^{-1}[B]\in\mathcal{B}_{T} and λμ¯(fμ¯1[B])=𝖫𝖻(BGμ¯)\lambda^{\bar{\mu}}(f_{\bar{\mu}}^{-1}[B])=\operatorname{\mathsf{Lb}}\left(B\smallsetminus G^{\ast}_{\bar{\mu}}\right). As a consequence, 𝖫𝖻(B)=λμ¯(fμ¯1[B])+𝖫𝖻(BGμ¯).\operatorname{\mathsf{Lb}}(B)=\lambda^{\bar{\mu}}(f_{\bar{\mu}}^{-1}[B])+\operatorname{\mathsf{Lb}}\left(B\cap G^{\ast}_{\bar{\mu}}\right).

  3. (c)

    ATA\in\mathcal{B}_{T} iff ANμ¯TA\cap N_{\bar{\mu}}^{\ast}\in\mathcal{B}_{T} and fμ¯[A]([0,1])f_{\bar{\mu}}[A]\in\mathcal{B}([0,1]).

  4. (d)

    B([0,1])B\in\mathcal{B}([0,1]) iff BGμ¯([0,1])B\cap G^{\ast}_{\bar{\mu}}\in\mathcal{B}([0,1]) and fμ¯1[B]Tf_{\bar{\mu}}^{-1}[B]\in\mathcal{B}_{T}.

  5. (e)

    B([0,1])B\in\mathcal{B}([0,1]) iff BGμ¯([0,1])B\cap G^{\ast}_{\bar{\mu}}\in\mathcal{B}([0,1]) and gμ¯[B]Tg_{\bar{\mu}}[B]\in\mathcal{B}_{T}. In this case, 𝖫𝖻(BGμ¯)=λμ¯(gμ¯[B]Vμ¯)\operatorname{\mathsf{Lb}}(B\smallsetminus G^{\ast}_{\bar{\mu}})=\lambda^{\bar{\mu}}(g_{\bar{\mu}}[B]\cap V^{*}_{\bar{\mu}}).

  6. (f)

    ATA\in\mathcal{B}_{T} iff ANμ¯TA\cap N_{\bar{\mu}}^{\ast}\in\mathcal{B}_{T} and gμ¯1[A]([0,1])g_{\bar{\mu}}^{-1}[A]\in\mathcal{B}([0,1]). In this case, 𝖫𝖻(gμ¯1[A])=λμ¯(A)\operatorname{\mathsf{Lb}}(g_{\bar{\mu}}^{-1}[A])=\lambda^{\bar{\mu}}(A).

Proof..

(a): Let ATA\in\mathcal{B}_{T}. For xAx\in A: if xAVμ¯x\in A\cap V^{*}_{\bar{\mu}}, then Ix={fμ¯(x)}I_{x}^{\ast}=\{f_{\bar{\mu}}(x)\}; and if xAVμ¯x\in A\smallsetminus V^{*}_{\bar{\mu}}, then Ix=[fμ¯,fμ¯+]I_{x}^{\ast}=[f_{\bar{\mu}}^{-},f_{\bar{\mu}}^{+}] is an uncountable interval. Now, fμ¯[A]([0,1])f_{\bar{\mu}}[A]\in\mathcal{B}([0,1]) by 6.18 (j),  6.11 (b) and because Vμ¯TV^{*}_{\bar{\mu}}\in\mathcal{B}_{T}. As a consequence, 𝖫𝖻(fμ¯[A])=𝖫𝖻(I(AVμ¯))=λμ¯(AVμ¯).\operatorname{\mathsf{Lb}}(f_{\bar{\mu}}[A])=\operatorname{\mathsf{Lb}}(I^{\ast}(A\cap V^{*}_{\bar{\mu}}))=\lambda^{\bar{\mu}}(A\cap V^{*}_{\bar{\mu}}). Finally, since we have fμ¯[A]GAI(A)fμ¯[A]GAQμ¯f_{\bar{\mu}}[A]\cup G_{A}^{\ast}\subseteq I^{\ast}(A)\subseteq f_{\bar{\mu}}[A]\cup G_{A}^{\ast}\cup Q_{\bar{\mu}}, it follows that:

λμ¯(A)=𝖫𝖻(I(A))=𝖫𝖻(fμ¯[A]GA)=𝖫𝖻(fμ¯[A])+xAVμ¯𝖫𝖻(Ix).\lambda^{\bar{\mu}}(A)=\operatorname{\mathsf{Lb}}(I^{\ast}(A))=\operatorname{\mathsf{Lb}}\left(f_{\bar{\mu}}[A]\cup G_{A}^{\ast}\right)=\operatorname{\mathsf{Lb}}(f_{\bar{\mu}}[A])+\sum_{x\in A\smallsetminus V^{*}_{\bar{\mu}}}\operatorname{\mathsf{Lb}}(I_{x}^{\ast}).

(b): If B([0,1])B\in\mathcal{B}([0,1]) we have that fμ¯1[B]Tf_{\bar{\mu}}^{-1}[B]\in\mathcal{B}_{T} because fμ¯f_{\bar{\mu}} is continuous. On the other hand, by (a),

λμ¯(fμ¯1[B])=λμ¯(fμ¯1[B]Vμ¯)=𝖫𝖻(Branfμ¯)=𝖫𝖻(BGμ¯).\lambda^{\bar{\mu}}(f_{\bar{\mu}}^{-1}[B])=\lambda^{\bar{\mu}}(f_{\bar{\mu}}^{-1}[B]\cap V^{*}_{\bar{\mu}})=\operatorname{\mathsf{Lb}}(B\cap\operatorname{ran}f_{\bar{\mu}})=\operatorname{\mathsf{Lb}}(B\smallsetminus G^{\ast}_{\bar{\mu}}).

(c): If ATA\in\mathcal{B}_{T}, then Nμ¯ATN_{\bar{\mu}}^{\ast}\cap A\in\mathcal{B}_{T} because Nμ¯N_{\bar{\mu}}^{\ast} is open in [T][T]. Also, fμ¯[A]([0,1])f_{\bar{\mu}}[A]\in\mathcal{B}([0,1]) follows by (a). To prove the converse, assume that ANμ¯TA\cap N_{\bar{\mu}}^{\ast}\in\mathcal{B}_{T} and fμ¯[A]Tf_{\bar{\mu}}[A]\in\mathcal{B}_{T}. Notice that we can write AA as a union of four sets:

A=[ANμ¯][Afμ¯1[Qμ¯]Nμ¯][AVμ¯fμ¯1[Qμ¯]][A[T]Vμ¯],A=[A\cap N_{\bar{\mu}}^{\ast}]\cup[A\cap f_{\bar{\mu}}^{-1}[Q_{\bar{\mu}}]\smallsetminus N_{\bar{\mu}}^{\ast}]\cup[A\cap V^{*}_{\bar{\mu}}\smallsetminus f_{\bar{\mu}}^{-1}[Q_{\bar{\mu}}]]\cup[A\cap[T]\smallsetminus V^{*}_{\bar{\mu}}],

so it is enough to show that these four sets are Borel in T\mathcal{B}_{T}. Since fμ¯1[Qμ¯]Nμ¯f_{\bar{\mu}}^{-1}[Q_{\bar{\mu}}]\smallsetminus N_{\bar{\mu}}^{\ast} and [T]Vμ¯[T]\smallsetminus V^{*}_{\bar{\mu}} are countable (see 6.18), and ANμ¯TA\cap N_{\bar{\mu}}^{\ast}\in\mathcal{B}_{T} by hypothesis, it only remains to prove that AVμ¯fμ¯1[Qμ¯]TA\cap V^{*}_{\bar{\mu}}\smallsetminus f_{\bar{\mu}}^{-1}[Q_{\bar{\mu}}]\in\mathcal{B}_{T}. Since fμ¯f_{\bar{\mu}} is Borel, fμ¯[A],domgμ¯Tf_{\bar{\mu}}[A],\operatorname{dom}g_{\bar{\mu}}\in\mathcal{B}_{T}, it follows that AVμ¯fμ¯1[Qμ¯]=fμ¯1[fμ¯[A]domgμ¯]T.A\cap V^{*}_{\bar{\mu}}\smallsetminus f_{\bar{\mu}}^{-1}[Q_{\bar{\mu}}]=f_{\bar{\mu}}^{-1}[f_{\bar{\mu}}[A]\cap\operatorname{dom}g_{\bar{\mu}}]\in\mathcal{B}_{T}.

(d): The implication from left to right follows by (b) and because Gμ¯G^{\ast}_{\bar{\mu}} is clearly a Borel set. To prove the converse, notice that we can write B=[Branfμ¯][BGμ¯][BQμ¯]B=[B\cap\operatorname{ran}f_{\bar{\mu}}]\cup[B\cap G_{\bar{\mu}}^{\ast}]\cup[B\cap Q_{\bar{\mu}}]. Since Qμ¯Q_{\bar{\mu}} is countable and by the hypothesis, it is enough to show that Branfμ¯B\cap\operatorname{ran}f_{\bar{\mu}} is Borel. This holds because, by hypothesis and by (a), Branfμ¯=fμ¯[fμ¯1[B]]([0,1])B\cap\operatorname{ran}f_{\bar{\mu}}=f_{\bar{\mu}}[f_{\bar{\mu}}^{-1}[B]]\in\mathcal{B}([0,1]).

(e): Assume that B([0,1])B\in\mathcal{B}([0,1]). Then, BGμ¯([0,1])B\cap G^{\ast}_{\bar{\mu}}\in\mathcal{B}([0,1]) because Gμ¯G_{\bar{\mu}}^{\ast} is Borel. On the other hand, gμ¯[B]Tg_{\bar{\mu}}[B]\in\mathcal{B}_{T} because gμ¯[B]=[gμ¯[B]Vμ¯][gμ¯[B][T]Vμ¯]g_{\bar{\mu}}[B]=[g_{\bar{\mu}}[B]\cap V^{*}_{\bar{\mu}}]\cup[g_{\bar{\mu}}[B]\cap[T]\smallsetminus V^{*}_{\bar{\mu}}], [T]Vμ¯[T]\smallsetminus V^{*}_{\bar{\mu}} is countable, fμ¯1[Branfμ¯]=gμ¯[B]Vμ¯f_{\bar{\mu}}^{-1}[B\cap\operatorname{ran}f_{\bar{\mu}}]=g_{\bar{\mu}}[B]\cap V^{*}_{\bar{\mu}} (by 6.18 (g)), fμ¯f_{\bar{\mu}} is a Borel function, and ranfμ¯\operatorname{ran}f_{\bar{\mu}} is Borel in [0,1][0,1]. To prove the converse, notice that we can write B=[Branfμ¯Qμ¯][BGμ¯][BQμ¯]B=[B\cap\operatorname{ran}f_{\bar{\mu}}\smallsetminus Q_{\bar{\mu}}]\cup[B\cap G_{\bar{\mu}}^{\ast}]\cup[B\cap Q_{\bar{\mu}}]. Therefore, it is enough to show that Branfμ¯Qμ¯B\cap\operatorname{ran}f_{\bar{\mu}}\smallsetminus Q_{\bar{\mu}} is Borel in ([0,1])\mathcal{B}([0,1]). This holds because Branfμ¯Qμ¯=gμ¯1[gμ¯[B]Vμ¯]B\cap\operatorname{ran}f_{\bar{\mu}}\smallsetminus Q_{\bar{\mu}}=g_{\bar{\mu}}^{-1}[g_{\bar{\mu}}[B]\cap V^{*}_{\bar{\mu}}], gμ¯g_{\bar{\mu}} is Borel, and gμ¯[B]g_{\bar{\mu}}[B] and Vμ¯V^{*}_{\bar{\mu}} are Borel sets. Finally, 𝖫𝖻(BGμ¯)=λμ¯(g[B]Vμ¯)\operatorname{\mathsf{Lb}}(B\smallsetminus G^{\ast}_{\bar{\mu}})=\lambda^{\bar{\mu}}(g[B]\cap V^{*}_{\bar{\mu}}) follows by (b) and 6.18 (g).

(f): Assume that ATA\in\mathcal{B}_{T}. Then ANμ¯TA\cap N_{\bar{\mu}}^{\ast}\in\mathcal{B}_{T} because Nμ¯N_{\bar{\mu}}^{\ast} is open, and gμ¯1[A]([0,1])g_{\bar{\mu}}^{-1}[A]\in\mathcal{B}([0,1]) follows by 6.11. To show the converse, write AA as follows:

A=[ANμ¯][Afμ¯1[Qμ¯]Nμ¯][Arangμ¯].A=[A\cap N_{\bar{\mu}}^{\ast}]\cup[A\cap f_{\bar{\mu}}^{-1}[Q_{\bar{\mu}}]\smallsetminus N_{\bar{\mu}}^{\ast}]\cup[A\cap\operatorname{ran}g_{\bar{\mu}}].

Hence, it is enough to prove that the last set is Borel. Since gμ¯1[A]([0,1])g_{\bar{\mu}}^{-1}[A]\in\mathcal{B}([0,1]), by (e), Arangμ¯=gμ¯[gμ¯1[A]]TA\cap\operatorname{ran}g_{\bar{\mu}}=g_{\bar{\mu}}[g_{\bar{\mu}}^{-1}[A]]\in\mathcal{B}_{T}.

Finally, regarding the measure, we have 𝖫𝖻(gμ¯1[A])=𝖫𝖻(I(A)Qμ¯)=𝖫𝖻(I(A))=λμ¯(A)\operatorname{\mathsf{Lb}}(g_{\bar{\mu}}^{-1}[A])=\operatorname{\mathsf{Lb}}(I^{\ast}(A)\smallsetminus Q_{\bar{\mu}})=\operatorname{\mathsf{Lb}}(I^{\ast}(A))=\lambda^{\bar{\mu}}(A). ∎

Remark 6.20.

The converse in Theorem 6.19 (a) does not hold when Nμ¯N_{\bar{\mu}}^{\ast} is uncountable because it contains non-Borel sets that are mapped into the countable Qμ¯Q_{\bar{\mu}}. Similarly, the converse in Theorem 6.19 (b) is not true when Gμ¯G^{\ast}_{\bar{\mu}}\neq\emptyset, because it contains non-Borel subsets whose pre-images are empty.

We can also analyze the completion of [T],T,λμ¯\langle[T],\mathcal{B}_{T},\lambda^{\bar{\mu}}\rangle.

Definition 6.21.

Denote by [T],μ¯,λμ¯\langle[T],\mathcal{L}_{\bar{\mu}},\lambda^{\bar{\mu}}\rangle the completion of [T],T,λμ¯\langle[T],\mathcal{B}_{T},\lambda^{\bar{\mu}}\rangle.

In the cases of the Cantor space and the Bairse space, we have:

Example 6.22.
  1. (1)

    2ω,(2ω),𝖫𝖻2\langle{}^{\omega}2,\mathcal{L}({}^{\omega}2),\operatorname{\mathsf{Lb}}_{2}\rangle is the completion of the measure on (2ω)\mathcal{B}({}^{\omega}2) from the probability tree in 4.4 (1). Here, (2ω)\mathcal{L}({}^{\omega}2) is the Lebesgue σ\sigma-algebra on the Cantor space and 𝖫𝖻2\operatorname{\mathsf{Lb}}_{2} is the Lebesgue measure on the Cantor space. Note that 𝖫𝖻2\operatorname{\mathsf{Lb}}_{2} is free (see 6.7 (1)).

  2. (2)

    ωω,(ωω),𝖫𝖻ω\langle{}^{\omega}\omega,\mathcal{L}({}^{\omega}\omega),\operatorname{\mathsf{Lb}}_{\omega}\rangle is the completion of the measure on (ωω)\mathcal{B}({}^{\omega}\omega) from the probability tree in 4.4 (2). Here, (ωω)\mathcal{L}({}^{\omega}\omega) is the Lebesgue σ\sigma-algebra on the Baire space and 𝖫𝖻ω\operatorname{\mathsf{Lb}}_{\omega} is the Lebesgue measure on the Baire space. Note that 𝖫𝖻ω\operatorname{\mathsf{Lb}}_{\omega} is free (see 6.7 (2)).

Since every measurable set can be decomposed as a union of a Borel set and a null set, from Theorem 6.19 we get:

Corollary 6.23.

Let A[T]A\subseteq[T] and B[0,1]B\subseteq[0,1]. Then:

  1. (a)

    Aμ¯A\in\mathcal{L}_{\bar{\mu}} iff fμ¯[A]([0,1])f_{\bar{\mu}}[A]\in\mathcal{L}([0,1]). In this case, 𝖫𝖻(fμ¯[A])=λμ¯(AVμ¯)\operatorname{\mathsf{Lb}}(f_{\bar{\mu}}[A])=\lambda^{\bar{\mu}}(A\cap V^{*}_{\bar{\mu}}) and λμ¯(A)=𝖫𝖻(fμ¯[A])+𝖫𝖻(GA)=𝖫𝖻(fμ¯[A])+xAVμ¯𝖫𝖻(Ix)\lambda^{\bar{\mu}}(A)=\operatorname{\mathsf{Lb}}(f_{\bar{\mu}}[A])+\operatorname{\mathsf{Lb}}(G^{*}_{A})=\operatorname{\mathsf{Lb}}(f_{\bar{\mu}}[A])+\sum_{x\in A\smallsetminus V^{*}_{\bar{\mu}}}\operatorname{\mathsf{Lb}}(I_{x}^{\ast}).

  2. (b)

    B([0,1])B\in\mathcal{L}([0,1]) iff fμ¯1[B]μ¯f_{\bar{\mu}}^{-1}[B]\in\mathcal{L}_{\bar{\mu}} and BGμ¯([0,1])B\cap G^{\ast}_{\bar{\mu}}\in\mathcal{L}([0,1]). In this case, we have that λμ¯(fμ¯1[B])=𝖫𝖻(BGμ¯)\lambda^{\bar{\mu}}(f_{\bar{\mu}}^{-1}[B])=\operatorname{\mathsf{Lb}}(B\smallsetminus G^{\ast}_{\bar{\mu}}) and 𝖫𝖻(B)=λμ¯(fμ¯1[B])+𝖫𝖻(BGμ¯)\operatorname{\mathsf{Lb}}(B)=\lambda^{\bar{\mu}}(f_{\bar{\mu}}^{-1}[B])+\operatorname{\mathsf{Lb}}(B\cap G^{\ast}_{\bar{\mu}}).

  3. (c)

    Aμ¯A\in\mathcal{L}_{\bar{\mu}} iff gμ¯1[A]([0,1])g_{\bar{\mu}}^{-1}[A]\in\mathcal{L}([0,1]). In this case, 𝖫𝖻(gμ¯1[A])=λμ¯(A)\operatorname{\mathsf{Lb}}(g_{\bar{\mu}}^{-1}[A])=\lambda^{\bar{\mu}}(A).

  4. (d)

    B([0,1])B\in\mathcal{L}([0,1]) iff gμ¯[B]μ¯g_{\bar{\mu}}[B]\in\mathcal{L}_{\bar{\mu}} and BGμ¯([0,1])B\cap G^{\ast}_{\bar{\mu}}\in\mathcal{L}([0,1]). In this case, we have that 𝖫𝖻(BGμ¯)=λμ¯(gμ¯[B]Vμ¯)\operatorname{\mathsf{Lb}}(B\smallsetminus G^{\ast}_{\bar{\mu}})=\lambda^{\bar{\mu}}(g_{\bar{\mu}}[B]\cap V^{*}_{\bar{\mu}}).

Finally, let us consider the case in which λμ¯\lambda^{\bar{\mu}} is free, i.e. when every point in [T][T] has measure zero. Thanks to 6.16 and 6.5 (a), we have the following characterization.

Lemma 6.24.

The following statements are equivalent.

  1. (i)

    λμ¯\lambda^{\bar{\mu}} is a free.

  2. (ii)

    𝖫𝖻(Ix)=0\operatorname{\mathsf{Lb}}(I^{*}_{x})=0 for all x[T]x\in[T], i.e. Vμ¯=[T]V^{*}_{\bar{\mu}}=[T].

  3. (iii)

    limnΞμ¯({xn})=n<|x|μxn({x(n+1)})=0\displaystyle\lim_{n\to\infty}\Xi^{\bar{\mu}}(\{x{\upharpoonright}n\})=\prod_{n<|x|}\mu_{x{\upharpoonright}n}(\{x{\upharpoonright}(n+1)\})=0 for all x[T]x\in[T].

  4. (iv)

    IxI^{*}_{x} is a singleton for all x[T]x\in[T].

As a direct consequence, when λμ¯\lambda^{\bar{\mu}} is free we get information about the structure of Tμ¯+T_{\bar{\mu}}^{+}.

Corollary 6.25.

If λμ¯\lambda^{\bar{\mu}} is free then Tμ¯+T_{\bar{\mu}}^{+} is a perfect tree.

Proof..

If tTμ¯+t\in T^{+}_{\bar{\mu}} then λμ¯([t]T[Tμ¯+])=λμ¯([t]T)>0\lambda^{\bar{\mu}}([t]_{T}\cap[T^{+}_{\bar{\mu}}])=\lambda^{\bar{\mu}}([t]_{T})>0, so [t]T[Tμ¯+][t]_{T}\cap[T^{+}_{\bar{\mu}}] contains more than two points. This implies that there are two incompatible nodes in Tμ¯+T^{+}_{\bar{\mu}} above tt. ∎

When λμ¯\lambda^{\bar{\mu}} is free, some properties listed in 6.18 and Theorem 6.19 can be simplified. For example, Vμ¯=[T]V^{*}_{\bar{\mu}}=[T] and hence Gμ¯=G_{\bar{\mu}}^{\ast}=\emptyset (see Figure 6). As a consequence:

Theorem 6.26.

Assume that λμ¯\lambda^{\bar{\mu}} is free.

  1. (a)

    fμ¯(gμ¯(y))=yf_{\bar{\mu}}(g_{\bar{\mu}}(y))=y for y[0,1]Qμ¯y\in[0,1]\smallsetminus Q_{\bar{\mu}}, i.e. gμ¯g_{\bar{\mu}} is one-to-one.

  2. (b)

    If x[T]x\in[T] and fμ¯(x)Qμ¯f_{\bar{\mu}}(x)\notin Q_{\bar{\mu}} then gμ¯(fμ¯(x))=xg_{\bar{\mu}}(f_{\bar{\mu}}(x))=x.

  3. (c)

    rangμ¯=[T]fμ¯1[Qμ¯]\operatorname{ran}g_{\bar{\mu}}=[T]\smallsetminus f^{-1}_{\bar{\mu}}[Q_{\bar{\mu}}].

  4. (d)

    fμ¯f_{\bar{\mu}} is continuous and [0,1)ranfμ¯[0,1)\subseteq\operatorname{ran}f_{\bar{\mu}}.

  5. (e)

    fμ¯rangμ¯f_{\bar{\mu}}{\upharpoonright}\operatorname{ran}g_{\bar{\mu}} is an homeomorphism from rangμ¯\operatorname{ran}g_{\bar{\mu}} onto [0,1]Qμ¯[0,1]\smallsetminus Q_{\bar{\mu}} with inverse gμ¯g_{\bar{\mu}}.

  6. (f)

    maxTNμ¯fμ¯1[Qμ¯]\max T\subseteq N^{*}_{\bar{\mu}}\subseteq f^{-1}_{\bar{\mu}}[Q_{\bar{\mu}}].

Nμ¯N_{\bar{\mu}}^{\ast}[T][T][0,1][0,1]fμ¯f_{\bar{\mu}}gμ¯g_{\bar{\mu}} 1\bullet\,1rangμ¯\mathrm{ran}g_{\bar{\mu}}domgμ¯\mathrm{dom}g_{\bar{\mu}}fμ¯1[Qμ¯]f_{\bar{\mu}}^{-1}[Q_{\bar{\mu}}]Qμ¯Q_{\bar{\mu}}Vμ¯V^{*}_{\bar{\mu}}ranfμ¯\operatorname{ran}f_{\bar{\mu}}
Figure 6. The situation in Theorem 6.26, that is, when λμ¯\lambda^{\bar{\mu}} is free. In this case, Vμ¯=[T]V^{*}_{\bar{\mu}}=[T], hence Gμ¯=G_{\bar{\mu}}=\emptyset. The shaded regions are homeomorphic via fμ¯rangμ¯f_{\bar{\mu}}{\restriction}\operatorname{ran}g_{\bar{\mu}} and its inverse is gμ¯g_{\bar{\mu}}, and the dotted curve indicates that 11 may or may not belong to ranfμ¯\operatorname{ran}f_{\bar{\mu}}
Proof..

By 6.18 it just remains to prove that [0,1)ranfμ¯[0,1)\subseteq\operatorname{ran}f_{\bar{\mu}}. For this, by (a), [0,1]Qμ¯ranfμ¯[0,1]\smallsetminus Q_{\bar{\mu}}\subseteq\operatorname{ran}f_{\bar{\mu}}. Now, if yQμ¯{1}y\in Q_{\bar{\mu}}\smallsetminus\{1\} then, by 6.9, y=aty=a_{t} for some tTt\in T. Thus, y=at=fμ¯(t0,0,0,)y=a_{t}=f_{\bar{\mu}}(t{}^{\frown}\langle 0,0,0,\ldots\rangle).∎

From 6.24 and Theorem 6.26 (d) it follows like in 6.7 (1) and (2) that Qμ¯Q_{\bar{\mu}} is dense when λμ¯\lambda^{\bar{\mu}} is free. This is actually a characterization of freeness.

Corollary 6.27.

λμ¯\lambda^{\bar{\mu}} is free iff Qμ¯Q_{\bar{\mu}} is dense in [0,1][0,1].

Proof..

Assume that λμ¯\lambda^{\bar{\mu}} is free. Let a,b[0,1]a,b\in[0,1] such that a<ba<b. Pick some y(a,b)y\in(a,b). By Theorem 6.26 (d), there is a x[T]x\in[T] such that y=fμ¯(x)y=f_{\bar{\mu}}(x), hence limnaxn=y\lim_{n\to\infty}a_{x{\restriction}n}=y. Therefore, we can find an N<ωN<\omega such that axNya_{x{\restriction}N}\leq y and yaxN<yay-a_{x{\restriction}N}<y-a, that is, at(a,b)a_{t}\in(a,b), where tXNt\coloneqq X{\restriction}N. Thus, Qμ¯Q_{\bar{\mu}} is dense. To prove the converse, assume that λμ¯\lambda^{\bar{\mu}} is not free, that is, there exists some x[T]x\in[T] such that λμ¯({x})>0\lambda^{\bar{\mu}}(\{x\})>0. Let us show that (fμ¯(x),fμ¯+(x))Qμ¯=(f_{\bar{\mu}}^{-}(x),f_{\bar{\mu}}^{+}(x))\cap Q_{\bar{\mu}}=\emptyset. Towards contradiction, assume that there exists a tTt\in T such that at(fμ¯(x),fμ¯+(x))a_{t}\in(f_{\bar{\mu}}^{-}(x),f_{\bar{\mu}}^{+}(x)). Consider sx|t|s\coloneqq x{\restriction}|t|, hence asfμ¯(x)<at<fμ¯+(x)bsa_{s}\leq f_{\bar{\mu}}^{-}(x)<a_{t}<f_{\bar{\mu}}^{+}(x)\leq b_{s}, which is not possible. Thus Qμ¯Q_{\bar{\mu}} is not dense in [0,1][0,1]. ∎

As mentioned in 6.20, to establish equivalences in Theorem 6.19 while preserving the measure there, we face two potential obstacles: on the one hand, Gμ¯G_{\bar{\mu}}^{\ast} may be non-empty, and on the other hand, Nμ¯N_{\bar{\mu}}^{\ast} may include non-measurable subsets. The first issue is solved when λμ¯\lambda^{\bar{\mu}} is free, and the second by completing the measure, since Nμ¯N_{\bar{\mu}}^{\ast} is null in [T][T]. Thus, under these conditions, Theorem 6.19 and 6.23 get simplified.

Theorem 6.28.

Assume that λμ¯\lambda^{\bar{\mu}} is free, A[T]A\subseteq[T] and B[0,1]B\subseteq[0,1]. Then:

  1. (a)

    ATA\in\mathcal{B}_{T} iff fμ¯[A]([0,1])f_{\bar{\mu}}[A]\in\mathcal{B}([0,1]) and ANμ¯TA\cap N^{*}_{\bar{\mu}}\in\mathcal{B}_{T}. In this case, 𝖫𝖻(fμ¯[A])=λμ¯(A)\operatorname{\mathsf{Lb}}(f_{\bar{\mu}}[A])=\lambda^{\bar{\mu}}(A).

  2. (b)

    B([0,1])B\in\mathcal{B}([0,1]) iff fμ¯1[B]Tf^{-1}_{\bar{\mu}}[B]\in\mathcal{B}_{T}. In this case λμ¯(fμ¯1[B])=𝖫𝖻(B)\lambda^{\bar{\mu}}(f^{-1}_{\bar{\mu}}[B])=\operatorname{\mathsf{Lb}}(B).

  3. (c)

    Aμ¯A\in\mathcal{L}_{\bar{\mu}} iff fμ¯[A]()f_{\bar{\mu}}[A]\in\mathcal{L}(\mathbb{R}). In this case, 𝖫𝖻(fμ¯[A])=λμ¯(A)\operatorname{\mathsf{Lb}}(f_{\bar{\mu}}[A])=\lambda^{\bar{\mu}}(A).

  4. (d)

    B()B\in\mathcal{L}(\mathbb{R}) iff fμ¯1[B]μ¯f^{-1}_{\bar{\mu}}[B]\in\mathcal{L}_{\bar{\mu}}. In this case λμ¯(fμ¯1[B])=𝖫𝖻(B)\lambda^{\bar{\mu}}(f^{-1}_{\bar{\mu}}[B])=\operatorname{\mathsf{Lb}}(B).

  5. (e)

    B([0,1])B\in\mathcal{B}([0,1]) iff gμ¯[B]Tg_{\bar{\mu}}[B]\in\mathcal{B}_{T}. In this case, 𝖫𝖻(B)=λμ¯(gμ¯[B])\operatorname{\mathsf{Lb}}(B)=\lambda^{\bar{\mu}}(g_{\bar{\mu}}[B]).

  6. (f)

    ATA\in\mathcal{B}_{T} iff gμ¯1[A]([0,1])g_{\bar{\mu}}^{-1}[A]\in\mathcal{B}([0,1]) and ANμ¯TA\cap N^{*}_{\bar{\mu}}\in\mathcal{B}_{T}. In this case, 𝖫𝖻(gμ¯1[A])=λμ¯(A)\operatorname{\mathsf{Lb}}(g_{\bar{\mu}}^{-1}[A])=\lambda^{\bar{\mu}}(A).

  7. (g)

    Aμ¯A\in\mathcal{L}_{\bar{\mu}} iff gμ¯1[A]()g_{\bar{\mu}}^{-1}[A]\in\mathcal{L}(\mathbb{R}). In this case, 𝖫𝖻(gμ¯1[A])=λμ¯(A)\operatorname{\mathsf{Lb}}(g_{\bar{\mu}}^{-1}[A])=\lambda^{\bar{\mu}}(A).

  8. (h)

    B()B\in\mathcal{L}(\mathbb{R}) iff gμ¯[B]μ¯g_{\bar{\mu}}[B]\in\mathcal{L}_{\bar{\mu}}. In this case λμ¯(gμ¯[B])=𝖫𝖻(B)\lambda^{\bar{\mu}}(g_{\bar{\mu}}[B])=\operatorname{\mathsf{Lb}}(B).

In addition, we can show a connection between T,μ¯\langle T,\bar{\mu}\rangle and the binary probability tree.

Definition 6.29.

For αω\alpha\leq\omega, denote by 1¯α\bar{1}^{\alpha} the constant sequence of 11’s of length α\alpha. Given a representation of TT, define hhT:T2<ωh\coloneqq h^{T}\colon T\to{}^{<\omega}2 by recursion as follows:

  • h()h(\langle\,\rangle)\coloneqq\langle\,\rangle;

  • if αt=1\alpha_{t}=1, define h(t0)h(t)h(t{}^{\frown}\langle 0\rangle)\coloneqq h(t);

  • if 1<αt<ω1<\alpha_{t}<\omega, define

    h(tk){h(t)1¯k0if k<αt1,h(t)1¯αt1if k=αt1;h(t{}^{\frown}\langle k\rangle)\coloneqq\begin{cases}h(t){}^{\frown}\bar{1}^{k}{}^{\frown}\langle 0\rangle&\text{if $k<\alpha_{t}-1$,}\\ h(t){}^{\frown}\bar{1}^{\alpha_{t}-1}&\text{if $k=\alpha_{t}-1$;}\end{cases}
  • if αt=ω\alpha_{t}=\omega, define h(tk)h(t)1¯k0h(t{}^{\frown}\langle k\rangle)\coloneqq h(t){}^{\frown}\bar{1}^{k}{}^{\frown}\langle 0\rangle for all k<ωk<\omega.

Let SSTS\coloneqq S^{T} be the smallest subtree of 2<ω{}^{<\omega}2 containing ranhT\operatorname{ran}h^{T}, i.e. sSs\in S iff ss is below some node in ranhT\operatorname{ran}h^{T}.

We list below some basic properties of the function hTh^{T}.

Lemma 6.30.

Let s,tTs,t\in T.

  1. (a)

    sts\subseteq t implies h(s)h(t)h(s)\subseteq h(t).

  2. (b)

    sts\perp t iff h(s)h(t)h(s)\perp h(t).

  3. (c)

    h(s)h(t)h(s)\subseteq h(t) iff [t]T[s]T[t]_{T}\subseteq[s]_{T}.555See 3.14 (c).

  4. (d)

    If αt<ω\alpha_{t}<\omega and tmax(T)t\notin\max(T) then

    [h(t)]2<ω=t𝗌𝗎𝖼𝖼T(t)[h(t)]2<ω (disjoint union).[h(t)]_{{}^{<\omega}2}=\bigcup_{t^{\prime}\in\operatorname{\mathsf{succ}}_{T}(t)}[h(t^{\prime})]_{{}^{<\omega}2}\text{ (disjoint union).}
  5. (e)

    If αt=ω\alpha_{t}=\omega and tmax(T)t\notin\max(T) then

    [h(t)]2<ω={h(t)1¯ω}t𝗌𝗎𝖼𝖼T(t)[h(t)]2<ω (disjoint union).[h(t)]_{{}^{<\omega}2}=\{h(t){}^{\frown}\bar{1}^{\omega}\}\cup\bigcup_{t^{\prime}\in\operatorname{\mathsf{succ}}_{T}(t)}[h(t^{\prime})]_{{}^{<\omega}2}\text{ (disjoint union).}
  6. (f)

    If tmaxTt\in\max T then h(t)maxSh(t)\in\max S.

  7. (g)

    h(t)maxSh(t)\in\max S iff there are no splitting nodes above tt in TT (including it).

  8. (h)

    h(t)maxSh(t)\in\max S iff h(t)h(t) is not a splitting node of SS.

  9. (i)

    Any node in SS is either maximal or splitting, and maxSranh\max S\subseteq\operatorname{ran}h.

  10. (j)

    TT is a perfect tree iff S=2<ωS={}^{<\omega}2.

Definition 6.31.

Thanks to 6.30, we can define a map eT:[T][S]e_{T}\colon[T]\to[S] that sends x[T]x\in[T] to some y[S]y\in[S] extending h(xn)h(x{\upharpoonright}n) for all n<ωn<\omega.

Lemma 6.32.
  1. (a)

    The function eTe_{T} is well-defined, i.e. the yy in 6.31 exists and is unique.

  2. (b)

    eTe_{T} is a topological embedding.

  3. (c)

    [S]raneT={h(t)1¯ω:tTmaxT,αt=ω}[S]\smallsetminus\operatorname{ran}e_{T}=\{h(t){}^{\frown}\bar{1}^{\omega}\colon t\in T\smallsetminus\max T,\ \alpha_{t}=\omega\}, which is countable.

  4. (d)

    eTe_{T} is onto iff TT is finitely splitting.

Definition 6.33.

Using μ¯\bar{\mu}, we define a measure Ξ\Xi^{*} on 𝒫(ST)\operatorname{\mathcal{P}}(S^{T}) as follows: if s=h(t)s=h(t) for some tTt\in T, set Ξ({s})Ξμ¯({t})\Xi^{*}(\{s\})\coloneqq\Xi^{\bar{\mu}}(\{t\}) (this value does not depend on tt because h(t)=h(t)h(t)=h(t^{\prime}) implies [t]T=[t]T[t]_{T}=[t^{\prime}]_{T}), otherwise, if sSranhs\in S\smallsetminus\operatorname{ran}h, pick the largest node tTt\in T such that h(t)sh(t)\subseteq s, and set Ξ({s})Ξμ¯({t𝗌𝗎𝖼𝖼T(t):sh(t)}\Xi^{*}(\{s\})\coloneqq\Xi^{\bar{\mu}}(\{t^{\prime}\in\operatorname{\mathsf{succ}}_{T}(t)\colon s\subseteq h(t^{\prime})\}.

Theorem 6.34.

Ξ𝒫\Xi^{*}\in\mathcal{IP}. Moreover, for any ν¯𝒯𝒫\bar{\nu}\in\mathcal{TP} satisfying Ξ=Ξν¯\Xi^{*}=\Xi^{\bar{\nu}}, A[T]A\subseteq[T] and B[S]B\subseteq[S],

  1. (a)

    Qμ¯=Qν¯Q_{\bar{\mu}}=Q_{\bar{\nu}} and gμ¯=eT1gν¯g_{\bar{\mu}}=e_{T}^{-1}\circ g_{\bar{\nu}}.

  2. (b)

    Vν¯=eT[Vμ¯]([S]raneT)V^{*}_{\bar{\nu}}=e_{T}[V^{*}_{\bar{\mu}}]\cup([S]\smallsetminus\operatorname{ran}e_{T}) and fμ¯=fν¯eTf_{\bar{\mu}}=f_{\bar{\nu}}\circ e_{T}.

  3. (c)

    Aμ¯A\in\mathcal{L}_{\bar{\mu}} iff eT[A]ν¯e_{T}[A]\in\mathcal{L}_{\bar{\nu}}, in which case λν¯(eT[A])=λμ¯(A)\lambda^{\bar{\nu}}(e_{T}[A])=\lambda^{\bar{\mu}}(A).

  4. (d)

    Bν¯B\in\mathcal{L}_{\bar{\nu}} iff eT1[B]μ¯e^{-1}_{T}[B]\in\mathcal{L}_{\bar{\mu}}, in which case λμ¯(eT1[B])=λν¯(B)\lambda^{\bar{\mu}}(e^{-1}_{T}[B])=\lambda^{\bar{\nu}}(B).

  5. (e)

    eT[Nμ¯]=Nν¯raneTe_{T}[N^{*}_{\bar{\mu}}]=N^{*}_{\bar{\nu}}\cap\operatorname{ran}e_{T}.

Proof..

Clearly, Ξ()=Ξμ¯()=1\Xi^{*}(\langle\,\rangle)=\Xi^{\bar{\mu}}(\langle\,\rangle)=1. Now let sSmaxSs\in S\smallsetminus\max S. In the case that s=h(t)s=h(t) for some tTt\in T, tmaxTt\notin\max T, even more, this tt can be found as a splitting node of TT. Then h(t0)=s0h(t{}^{\frown}\langle 0\rangle)=s{}^{\frown}\langle 0\rangle, so Ξ({s0})=Ξμ¯({t0})\Xi^{*}(\{s{}^{\frown}\langle 0\rangle\})=\Xi^{\bar{\mu}}(\{t{}^{\frown}\langle 0\rangle\}), and Ξ({h(t1)})=0<k<αtΞμ¯({tk})\Xi^{*}(\{h(t{}^{\frown}\langle 1\rangle)\})=\sum_{0<k<\alpha_{t}}\Xi^{\bar{\mu}}(\{t{}^{\frown}k\}). Thus, Ξ(𝗌𝗎𝖼𝖼S(s))=k<αtΞμ¯({tk})=Ξμ¯({t})=Ξ({s})\Xi^{*}(\operatorname{\mathsf{succ}}_{S}(s))=\sum_{k<\alpha_{t}}\Xi^{\bar{\mu}}(\{t{}^{\frown}k\})=\Xi^{\bar{\mu}}(\{t\})=\Xi^{*}(\{s\}).

One can show by recursion that, for tTt\in T, Ih(t)ν¯=Itμ¯I^{\bar{\nu}}_{h(t)}=I^{\bar{\mu}}_{t}. This implies that Qν¯=Qμ¯Q_{\bar{\nu}}=Q_{\bar{\mu}}, i.e. domgν¯=domgμ¯\operatorname{dom}g_{\bar{\nu}}=\operatorname{dom}g_{\bar{\mu}}. On the other hand, whenever y[S]raneTy\in[S]\smallsetminus\operatorname{ran}e_{T}, i.e. y=h(t)1¯ωy=h(t){}^{\frown}\bar{1}^{\omega} for some tTt\in T with αt=ω\alpha_{t}=\omega, we have that λν¯({y})=limnknΞμ¯({tk})=0\lambda^{\bar{\nu}}(\{y\})=\lim_{n\to\infty}\sum_{k\geq n}\Xi^{\bar{\mu}}(\{t{}^{\frown}\langle k\rangle\})=0, so yVν¯y\in V^{*}_{\bar{\nu}} and fν¯(y)=btf_{\bar{\nu}}(y)=b_{t}. This shows that [S]raneTfν¯1[Qν¯][S]\smallsetminus\operatorname{ran}e_{T}\subseteq f^{-1}_{\bar{\nu}}[Q_{\bar{\nu}}], i.e. rangν¯raneT\operatorname{ran}g_{\bar{\nu}}\subseteq\operatorname{ran}e_{T}, which implies that dom(eT1gν¯)=domgν¯=domgμ¯\operatorname{dom}(e^{-1}_{T}\circ g_{\bar{\nu}})=\operatorname{dom}g_{\bar{\nu}}=\operatorname{dom}g_{\bar{\mu}}. Now, if zrangν¯z\in\operatorname{ran}g_{\bar{\nu}} then ygμ¯(z)raneTy\coloneqq g_{\bar{\mu}}(z)\in\operatorname{ran}e_{T}, so xeT1(y)x\coloneqq e_{T}^{-1}(y) is defined, i.e. yy is the unique extension of h(xn):n<ω\langle h(x{\upharpoonright}n)\colon n<\omega\rangle. Thus, zn<ωIh(xn)ν¯=n<ωIxnμ¯z\in\bigcap_{n<\omega}I^{\bar{\nu}}_{h(x{\upharpoonright}n)}=\bigcap_{n<\omega}I^{\bar{\mu}}_{x{\upharpoonright}n}, which shows that gμ¯(z)=x=eT1(gν¯(z))g_{\bar{\mu}}(z)=x=e^{-1}_{T}(g_{\bar{\nu}}(z)). This proves (a).

For x[T]x\in[T], λν¯({eT(x)})=limnΞ({h(xn)})=limnΞμ¯({xn})=λμ¯({x})\lambda^{\bar{\nu}}(\{e_{T}(x)\})=\lim_{n\to\infty}\Xi^{*}(\{h(x{\upharpoonright}n)\})=\lim_{n\to\infty}\Xi^{\bar{\mu}}(\{x{\upharpoonright}n\})=\lambda^{\bar{\mu}}(\{x\}). This proves that Vν¯=eT[Vμ¯]([S]raneT)V^{*}_{\bar{\nu}}=e_{T}[V^{*}_{\bar{\mu}}]\cup([S]\smallsetminus\operatorname{ran}e_{T}), so dom(fν¯eT)=Vμ¯=domfμ¯\operatorname{dom}(f_{\bar{\nu}}\circ e_{T})=V^{*}_{\bar{\mu}}=\operatorname{dom}f_{\bar{\mu}}. Now, for xVμ¯x\in V^{*}_{\bar{\mu}}, fμ¯(x)f_{\bar{\mu}}(x) is the unique point in n<ωIxnμ¯=n<ωIh(xn)ν¯={fν¯(eT(x))}\bigcap_{n<\omega}I^{\bar{\mu}}_{x{\upharpoonright}n}=\bigcap_{n<\omega}I^{\bar{\nu}}_{h(x{\upharpoonright}n)}=\{f_{\bar{\nu}}(e_{T}(x))\}. This concludes (b).

(c): Notice that gμ¯1[A]=gν¯1[eT[A]]g^{-1}_{\bar{\mu}}[A]=g^{-1}_{\bar{\nu}}[e_{T}[A]]. Then, by 6.23, Aμ¯A\in\mathcal{L}_{\bar{\mu}} iff eT[A]ν¯e_{T}[A]\in\mathcal{L}_{\bar{\nu}}, in which case λμ¯(A)=𝖫𝖻(gμ¯1[A])=𝖫𝖻(gν¯1[eT[A]])=λν¯(eT[A])\lambda^{\bar{\mu}}(A)=\operatorname{\mathsf{Lb}}(g^{-1}_{\bar{\mu}}[A])=\operatorname{\mathsf{Lb}}(g^{-1}_{\bar{\nu}}[e_{T}[A]])=\lambda^{\bar{\nu}}(e_{T}[A]).

(d): It follows from (c), also because [S]raneTfν¯1[Qν¯][S]\smallsetminus\operatorname{ran}e_{T}\subseteq f^{-1}_{\bar{\nu}}[Q_{\bar{\nu}}] has measure zero. ∎

7. Probability trees and the null ideal

In this section, we aim to prove the invariance of the cardinal invariants associated with the null ideal, mostly via Tukey connections. Although the most general result (Theorem 7.7) is already known, we offer an elementary and direct proof using probability trees. Our starting point is to show that, whenever λ𝒫\lambda\in\mathcal{BP} is free, 𝒩(λ)\mathcal{N}(\lambda) and 𝒩([0,1])\mathcal{N}([0,1]) are Tukey equivalent (see 7.3 and Theorem 7.6).

Based on [Voj93], we first introduce key concepts related to relational systems and Tukey connections, which will provide the necessary framework to formalize our results.

Definition 7.1.

We say that R=X,Y,R=\langle X,Y,\lhd\rangle is a relational system if XX, YY are non-empty sets and \lhd is a relation.

  1. (1)

    A set FXF\subseteq X is RR-bounded if yYxF(xy)\exists y\in Y\forall x\in F\ (x\lhd y).

  2. (2)

    A set EYE\subseteq Y is RR-dominating if xXyE(xy)\forall x\in X\exists y\in E\ (x\lhd y).

We can associate two cardinal invariants with relational systems:

  • 𝔟(R)min{|F|:FX is R-unbounded}\mathfrak{b}(R)\coloneqq\min\{|F|\colon F\subseteq X\text{ is }R\text{-unbounded}\}, the unbounding number of RR, and

  • 𝔡(R)min{|D|:DY is R-dominating}\mathfrak{d}(R)\coloneqq\min\{|D|\colon D\subseteq Y\text{ is }R\text{-dominating}\}, the dominating number of RR.

We now present some examples of relational systems and their associated cardinal invariants.

Example 7.2.

Let \mathcal{I} be an ideal on a set XX containing all its singletons.

  1. (1)

    =,,\mathcal{I}=\langle\mathcal{I},\mathcal{I},\subseteq\rangle is a relational system, 𝔟()=add()\mathfrak{b}(\mathcal{I})=\operatorname{add}(\mathcal{I}) and 𝔡()=cof()\mathfrak{d}(\mathcal{I})=\operatorname{cof}(\mathcal{I}).

  2. (2)

    𝖢𝗏X,,\mathsf{Cv}_{\mathcal{I}}\coloneqq\langle X,\mathcal{I},\in\rangle is a relational system, 𝔡(𝖢𝗏)=cov()\mathfrak{d}(\mathsf{Cv}_{\mathcal{I}})=\operatorname{cov}(\mathcal{I}) and 𝔟(𝖢𝗏)=non()\mathfrak{b}(\mathsf{Cv}_{\mathcal{I}})=\operatorname{non}(\mathcal{I}).

Now we introduce the notion of Tukey connections, which can be thought of as homomorphisms between relational systems:

Definition 7.3.

Let R=X,Y,R=\langle X,Y,\lhd\rangle and S=(Z,W,)S=(Z,W,\sqsubset) be relational systems. We say that (ψ,ψ+):RS(\psi_{-},\psi_{+})\colon R\to S is a Tukey connection from RR into SS if ψ:XZ\psi_{-}\colon X\to Z and ψ+:WY\psi_{+}\colon W\to Y are functions such that

xXwW[ψ(x)wxψ+(w)].\forall x\in X\forall w\in W\ [\psi_{-}(x)\sqsubset w\Rightarrow x\lhd\psi_{+}(w)].

In this case, we write RTSR\leq_{\mathrm{T}}S and we say that RR is Tukey-below SS. When RTSR\leq_{\mathrm{T}}S and STRS\leq_{\mathrm{T}}R, we say that RR and SS are Tukey equivalent and we denote it by R=TS.R=_{\mathrm{T}}S.

Tukey equivalences determined inequalities between the associated cardinal invariants.

Lemma 7.4.

Let R=X,Y,R=\langle X,Y,\lhd\rangle and S=Z,W,S=\langle Z,W,\sqsubset\rangle be two relational systems.

  1. (a)

    If RTSR\leq_{\mathrm{T}}S then 𝔟(R)𝔟(S)\mathfrak{b}(R)\geq\mathfrak{b}(S) and 𝔡(R)𝔡(S)\mathfrak{d}(R)\leq\mathfrak{d}(S).

  2. (b)

    If R=TSR=_{\mathrm{T}}S then 𝔟(R)=𝔟(S)\mathfrak{b}(R)=\mathfrak{b}(S) and 𝔡(R)=𝔡(S)\mathfrak{d}(R)=\mathfrak{d}(S).

To prove the results of this section, we introduce several maps. Let μ¯𝒯𝒫\bar{\mu}\in\mathcal{TP} and TTμ¯T\coloneqq T_{\bar{\mu}} (with some representation), and consider fμ¯f_{\bar{\mu}} as in 6.17. We define the maps

Hμ¯\displaystyle H_{\bar{\mu}}^{-} :𝒫([T])𝒫([0,1])\displaystyle\colon\operatorname{\mathcal{P}}([T])\to\operatorname{\mathcal{P}}([0,1]) Afμ¯[A]{1};\displaystyle A\mapsto f_{\bar{\mu}}[A]\cup\{1\};
Hμ¯+\displaystyle H_{\bar{\mu}}^{+} :𝒫([0,1])𝒫([T])\displaystyle\colon\operatorname{\mathcal{P}}([0,1])\to\operatorname{\mathcal{P}}([T]) Bfμ¯1[B].\displaystyle B\mapsto f_{\bar{\mu}}^{-1}[B].

For this section, we extend gμ¯g_{\bar{\mu}} to [0,1][0,1] in such a way that, for yQμ¯y\in Q_{\bar{\mu}}, gμ¯(y)g_{\bar{\mu}}(y) is some fμ¯f_{\bar{\mu}}-preimage of yy, if it exists, otherwise gμ¯(y)g_{\bar{\mu}}(y) is sent to some arbitrary point in [Tμ¯][T_{\bar{\mu}}].

Lemma 7.5.

Let μ¯𝒯𝒫\bar{\mu}\in\mathcal{TP}, TTμ¯T\coloneqq T_{\bar{\mu}}, A[T]A\subseteq[T], B[0,1]B\subseteq[0,1], x[T]x\in[T] and y[0,1]y\in[0,1]. Then:

  1. (a)

    Hμ¯(A)BH^{-}_{\bar{\mu}}(A)\subseteq B implies AHμ¯+(B)A\subseteq H^{+}_{\bar{\mu}}(B).

  2. (b)

    If λμ¯\lambda^{\bar{\mu}} is free, then Hμ¯+(B)AH^{+}_{\bar{\mu}}(B)\subseteq A implies BHμ¯(A)B\subseteq H^{-}_{\bar{\mu}}(A).

  3. (c)

    fμ¯(x)Bf_{\bar{\mu}}(x)\in B implies xHμ¯+(B)x\in H^{+}_{\bar{\mu}}(B).

  4. (d)

    If yranfμ¯{1}y\in\operatorname{ran}f_{\bar{\mu}}\cup\{1\} then gμ¯(y)Ag_{\bar{\mu}}(y)\in A implies yHμ¯(A)y\in H^{-}_{\bar{\mu}}(A).

  5. (e)

    If λμ¯\lambda^{\bar{\mu}} is free then gμ¯(y)Ag_{\bar{\mu}}(y)\in A implies yHμ¯(A)y\in H^{-}_{\bar{\mu}}(A).

Proof..

Easy to check, also using that, by Theorem 6.26 (d), [0,1)ranfμ¯[0,1)\subseteq\operatorname{ran}f_{\bar{\mu}} whenever λμ¯\lambda^{\bar{\mu}} is free. ∎

As a consequence, for the null ideal we obtain:

Theorem 7.6.

Let λ𝒫\lambda\in\mathcal{BP}. If λ\lambda is free then 𝒩(λ)=T𝒩([0,1])\mathcal{N}(\lambda)=_{\mathrm{T}}\mathcal{N}([0,1]) and 𝖢𝗏𝒩(λ)=T𝖢𝗏𝒩([0,1]).\mathsf{Cv}_{\mathcal{N}(\lambda)}=_{\mathrm{T}}\mathsf{Cv}_{\mathcal{N}([0,1])}.

Proof..

By 4.33 (b), there is some μ¯𝒯𝒫\bar{\mu}\in\mathcal{TP} such that λ=λμ¯\lambda=\lambda^{\bar{\mu}}. By 6.23, Hμ¯H^{-}_{\bar{\mu}} and Hμ¯+H^{+}_{\bar{\mu}} send measure zero sets into measure zero sets. Thus, by 7.5, it is clear that 𝒩(λ)=T𝒩([0,1])\mathcal{N}(\lambda)=_{\mathrm{T}}\mathcal{N}([0,1]) is witnessed by the pairs (Hμ¯𝒩(λ),Hμ¯+𝒩([0,1]))(H^{-}_{\bar{\mu}}{\upharpoonright}\mathcal{N}(\lambda),H^{+}_{\bar{\mu}}{\upharpoonright}\mathcal{N}([0,1])) and (Hμ¯+𝒩([0,1]),Hμ¯𝒩(λ))(H^{+}_{\bar{\mu}}\mathcal{N}([0,1]),H^{-}_{\bar{\mu}}\mathcal{N}(\lambda)), and 𝖢𝗏𝒩(λ)=T𝖢𝗏𝒩([0,1])\mathsf{Cv}_{\mathcal{N}(\lambda)}=_{\mathrm{T}}\mathsf{Cv}_{\mathcal{N}([0,1])} is witnessed by (fμ¯,Hμ¯+𝒩([0,1]))(f_{\bar{\mu}},H^{+}_{\bar{\mu}}{\upharpoonright}\mathcal{N}([0,1])) and (gμ¯,Hμ¯𝒩(λ))(g_{\bar{\mu}},H^{-}_{\bar{\mu}}{\upharpoonright}\mathcal{N}(\lambda)). ∎

Although the following generalization of Theorem 7.6 is already known, we offer an alternative proof using our results. Recall that a Polish space is a completely metrizable separable space, and that a Borel isomorphism between topological spaces is a bijection that sends Borel sets into Borel sets in both directions.

Theorem 7.7 ([Kec95, Thm. 17.41]).

If XX is a Polish space and μ:(X)[0,1]\mu\colon\mathcal{B}(X)\to[0,1] is a free probability measure, then there is some Borel isomorphism f:X[0,1]f\colon X\to[0,1] such that, for any Borel B[0,1]B\subseteq[0,1], μ(f1[B])=𝖫𝖻(B)\mu(f^{-1}[B])=\operatorname{\mathsf{Lb}}(B). In particular, 𝒩(μ)=T𝒩([0,1])\mathcal{N}(\mu)=_{\mathrm{T}}\mathcal{N}([0,1]) and 𝖢𝗏𝒩(μ)=T𝖢𝗏𝒩([0,1])\mathsf{Cv}_{\mathcal{N}(\mu)}=_{\mathrm{T}}\mathsf{Cv}_{\mathcal{N}([0,1])}.

Proof..

Since μ\mu is a free probability measure on (X)\mathcal{B}(X), XX must be uncountable, so there exists a Borel isomorphism g:Xωωg\colon X\to{}^{\omega}\omega. Let λ\lambda be the probability measure on ωω{}^{\omega}\omega induced by XX, i.e. λ(A)μ(g1[A])\lambda(A)\coloneqq\mu(g^{-1}[A]) for A(ωω)A\in\mathcal{B}({}^{\omega}\omega). It is clear that λ\lambda is free, so it is enough to prove the theorem for ωω,λ\langle{}^{\omega}\omega,\lambda\rangle instead of X,μ\langle X,\mu\rangle.

Pick some μ¯𝒯𝒫\bar{\mu}\in\mathcal{TP} such that λμ¯=λ\lambda^{\bar{\mu}}=\lambda, and pick some infinite countable Wωωfμ¯1[Qμ¯]W\subseteq{}^{\omega}\omega\smallsetminus f_{\bar{\mu}}^{-1}[Q_{\bar{\mu}}] and let Wfμ¯[W]W^{\prime}\coloneqq f_{\bar{\mu}}[W]. We proceed by cases. First assume that Nμ¯N^{*}_{\bar{\mu}} is countable. Hence, we can construct a function f:ωω[0,1]f\colon{}^{\omega}\omega\to[0,1] such that ff equals fμ¯f_{\bar{\mu}} at ωω(fμ¯1[Qμ¯]W){}^{\omega}\omega\smallsetminus(f^{-1}_{\bar{\mu}}[Q_{\bar{\mu}}]\cup W) and fμ¯[fμ¯1[Qμ¯]W]=Qμ¯Wf_{\bar{\mu}}[f^{-1}_{\bar{\mu}}[Q_{\bar{\mu}}]\cup W]=Q_{\bar{\mu}}\cup W^{\prime}. This is indeed a Borel bijection (hence isomorphism) and, by Theorem 6.28, it preserves measure as desired.

Now consider the case when Nμ¯N^{*}_{\bar{\mu}} is uncountable. Let C0C_{0} be the ternary Cantor set in [0,1][0,1], which is closed and 𝖫𝖻(C0)=0\operatorname{\mathsf{Lb}}(C_{0})=0. Pick a sequence Jn:1n<ω\langle J_{n}\colon 1\leq n<\omega\rangle of pairwise disjoint closed intervals of positive length contained in (1/3,2/3)(1/3,2/3) such that maxJn<minJn+1\max J_{n}<\min J_{n+1} for all nn, and let CnC_{n} be the image of C0C_{0} under the canonical homeomorphism from [0,1][0,1] onto JnJ_{n} (i.e the line segment from (0,minJn)(0,\min J_{n}) to (1,maxJn)(1,\max J_{n})), so CnC_{n} resembles the Cantor ternary set in JnJ_{n}, thus it is closed with measure zero.

Define the map h:[0,1][0,1]Ch\colon[0,1]\to[0,1]\smallsetminus C such that it is the identity on [0,1]n<ωCn[0,1]\smallsetminus\bigcup_{n<\omega}C_{n} and hCnh{\upharpoonright}C_{n} is a linear isomorphism onto Cn+1C_{n+1} for all n<ωn<\omega. This is a Borel isomorphism that preserves the Lebesgue measure. Finally, define f:ωω[0,1]f\colon{}^{\omega}\omega\to[0,1] such that ff coincides with hfμ¯h\circ f_{\bar{\mu}} at ωω(fμ¯1[Qμ¯]W)){}^{\omega}\omega\smallsetminus(f^{-1}_{\bar{\mu}}[Q_{\bar{\mu}}]\cup W)), f[(fμ¯1[Qμ¯]Nμ¯)W]=h[Qμ¯W]f[(f^{-1}_{\bar{\mu}}[Q_{\bar{\mu}}]\smallsetminus N^{*}_{\bar{\mu}})\cup W]=h[Q_{\bar{\mu}}\cup W^{\prime}], and fNμ¯f{\upharpoonright}N^{*}_{\bar{\mu}} is a Borel isomorphism onto C0C_{0}. This is the desired Borel isomorphism. ∎

We also have a similar result for the Lebesgue measure on \mathbb{R}.

Corollary 7.8.

If XX is a Polish space, μ:(X)[0,]\mu\colon\mathcal{B}(X)\to[0,\infty] is a free σ\sigma-finite measure and μ(X)=\mu(X)=\infty, then there is some Borel isomorphism f:Xf\colon X\to\mathbb{R} such that, for any Borel BB\subseteq\mathbb{R}, μ(f1[B])=𝖫𝖻(B)\mu(f^{-1}[B])=\operatorname{\mathsf{Lb}}(B). In particular, 𝒩(μ)=T𝒩()\mathcal{N}(\mu)=_{\mathrm{T}}\mathcal{N}(\mathbb{R}) and 𝖢𝗏𝒩(μ)=T𝖢𝗏𝒩()\mathsf{Cv}_{\mathcal{N}(\mu)}=_{\mathrm{T}}\mathsf{Cv}_{\mathcal{N}(\mathbb{R})}.

Proof..

Partition XX into Borel sets Bn:n\langle B_{n}\colon n\in\mathbb{Z}\rangle such that 0<μ(Bn)<0<\mu(B_{n})<\infty for each nn\in\mathbb{Z} and n=1μ(Bn)=n=1μ(Bn)=\sum_{n=1}^{\infty}\mu(B_{n})=\sum_{n=-1}^{-\infty}\mu(B_{n})=\infty. Then, we can partition \mathbb{R} into semi-open intervals Jn:n\langle J_{n}\colon n\in\mathbb{Z}\rangle such that each JnJ_{n} has length μ(Bn)\mu(B_{n}). By Theorem 7.7, there is some Borel isomorphism fn:BnJnf_{n}\colon B_{n}\to J_{n} preserving measure (this uses that, for any B(X)B\in\mathcal{B}(X), there is some finer Polish topology on XX such that its Borel σ\sigma-algebra is (X)\mathcal{B}(X) and BB is clopen, see e.g. [Kec95, Thm. 13.1]). Thus, we obtain the desired ff by putting all the fnf_{n} together. ∎

Corollary 7.9.

If XX is a Polish space, μ:(X)[0,]\mu\colon\mathcal{B}(X)\to[0,\infty] is a free σ\sigma-finite measure and μ(X)>0\mu(X)>0 then 𝒩(μ)=T𝒩()\mathcal{N}(\mu)=_{\mathrm{T}}\mathcal{N}(\mathbb{R}) and 𝖢𝗏𝒩(μ)=T𝖢𝗏𝒩()\mathsf{Cv}_{\mathcal{N}(\mu)}=_{\mathrm{T}}\mathsf{Cv}_{\mathcal{N}(\mathbb{R})}. In particular,

add(𝒩(μ))\displaystyle\operatorname{add}(\mathcal{N}(\mu)) =add(𝒩()),\displaystyle=\operatorname{add}(\mathcal{N}(\mathbb{R})), cof(𝒩(μ))\displaystyle\operatorname{cof}(\mathcal{N}(\mu)) =cof(𝒩()),\displaystyle=\operatorname{cof}(\mathcal{N}(\mathbb{R})),
non(𝒩(μ))\displaystyle\operatorname{non}(\mathcal{N}(\mu)) =non(𝒩()),\displaystyle=\operatorname{non}(\mathcal{N}(\mathbb{R})), cov(𝒩(μ))\displaystyle\operatorname{cov}(\mathcal{N}(\mu)) =cov(𝒩()).\displaystyle=\operatorname{cov}(\mathcal{N}(\mathbb{R})).

References

  • [AS16] Noga Alon and Joel H. Spencer. The probabilistic method. Wiley Series in Discrete Mathematics and Optimization. John Wiley & Sons, Inc., Hoboken, NJ, fourth edition, 2016.
  • [CMU24] Miguel A. Cardona, Diego A. Mejía, and Andrés F. Uribe-Zapata. A general theory of iterated forcing using finitely additive measures. Preprint, arXiv:2406.09978, 2024.
  • [FF80] Jan M. Friedman and R. David Fish. The use of probability trees in genetic counselling. Clinical Genetics, 18, 1980.
  • [FSNG23] Yangqing Fu, Ming Sun, Buqing Nie, and Yue Gao. Accelerating Monte Carlo tree search with probability tree state abstraction. Preprint, arXiv:2310.06513, 2023.
  • [GMD+20] Tim Genewein, Tom McGrath, Grégoire Déletang, Vladimir Mikulik, Miljan Martic, Shane Legg, and Pedro A. Ortega. Algorithms for causal reasoning in probability trees. Preprint, arXiv:2010.12237, 2020.
  • [Gri89] R. C. Griffiths. Genealogical-tree probabilities in the infinitely-many-site model. J. Math. Biol., 27(6):667–680, 1989.
  • [GT95] R. C. Griffiths and Simon Tavaré. Unrooted genealogical tree probabilities in the infinitely-many-sites model. Mathematical Biosciences, 127(1):77–98, 1995.
  • [Hal50] Paul R. Halmos. Measure Theory. D. Van Nostrand Co., Inc., New York, 1950.
  • [Kec95] Alexander S. Kechris. Classical Descriptive Set Theory. Springer New York, NY, 1995.
  • [KST19] Jakob Kellner, Saharon Shelah, and Anda R. Tănasie. Another ordering of the ten cardinal characteristics in Cichoń’s diagram. Comment. Math. Univ. Carolin., 60(1):61–95, 2019.
  • [KWWC24] Aneesh Komanduri, Xintao Wu, Yongkai Wu, and Feng Chen. From identifiable causal representations to controllable counterfactual generation: A survey on causal generative modeling. Transactions on Machine Learning Research, 2024.
  • [Lev02] Azriel Levy. Basic set theory. Dover Publications, Inc., Mineola, NY, 2002. Reprint of the 1979 original [Springer, Berlin].
  • [MU24] Diego A. Mejía and Andrés F. Uribe-Zapata. The measure algebra adding θ\theta-many random reals is θ\theta-FAM\mathrm{FAM}-linked. Topology Appl., 2024. To appear, arXiv:2312.13443.
  • [She00] Saharon Shelah. Covering of the null ideal may have countable cofinality. Fund. Math., 166(1-2):109–136, 2000.
  • [Uri23] Andrés F. Uribe-Zapata. Iterated forcing with finitely additive measures: applications of probability to forcing theory. Master’s thesis, Universidad Nacional de Colombia, sede Medellín, 2023. https://shorturl.at/sHY59.
  • [Voj93] Peter Vojtáš. Generalized Galois-Tukey-connections between explicit relations on classical objects of real analysis. In Set theory of the reals (Ramat Gan, 1991), volume 6 of Israel Math. Conf. Proc., pages 619–643. Bar-Ilan Univ., Ramat Gan, 1993.
  • [ZM18] Cheng Zhang and Frederick A Matsen IV. Generalizing tree probability estimation via bayesian networks. In S. Bengio, H. Wallach, H. Larochelle, K. Grauman, N. Cesa-Bianchi, and R. Garnett, editors, Advances in Neural Information Processing Systems, volume 31. Curran Associates, Inc., 2018.