Skip to the Main Content

Note:These pages make extensive use of the latest XHTML and CSS Standards. They ought to look great in any standards-compliant modern browser. Unfortunately, they will probably look horrible in older browsers, like Netscape 4.x and IE 4.x. Moreover, many posts use MathML, which is, currently only supported in Mozilla. My best suggestion (and you will thank me when surfing an ever-increasing number of sites on the web which have been crafted to use the new standards) is to upgrade to the latest version of your browser. If that's not possible, consider moving to the Standards-compliant and open-source Mozilla browser.

July 9, 2024

Imprecise Probabilities: Towards a Categorical Perspective

Posted by Emily Riehl

guest post by Laura González-Bravo and Luis López

In this blog post for the Applied Category Theory Adjoint School 2024, we discuss some of the limitations that the measure-theoretic probability framework has in handling uncertainty and present some other formal approaches to modelling it. With this blog post, we would like to initiate ourselves into the study of imprecise probabilities from a mathematical perspective.

Preliminaries

Even though we all have some intuitive grasp of what uncertainty is, providing a formal mathematical framework to represent it has proven to be non-trivial. We may understand uncertainty as the feeling of not being sure if an event will occur in the future. Classically, this feeling is mainly attributed to the lack of knowledge we may have about such an event or phenomenon. Often, this lack of knowledge is a condition from which we cannot escape, and it may preclude us from making reliable statements about the event. Let us think about the case of tossing a coin. If we think in a Newtonian deterministic way, we may think that if we had perfect knowledge about the initial conditions when tossing a coin, this would allow us to know which of the two outcomes will happen.

However, there are numerous unpredictable factors, such as the initial force applied, air currents, the surface it lands on, and microscopic imperfections in the coin itself that prevent us from knowing such initial conditions with infinite accuracy. In particular, at the moment you are throwing the coin you do not know the state of all the actin and myosin proteins in your muscles, which play an important role in your movements and thus in the outcome of the toss. So, even though the laws of physics govern the motion of the coin, its final state will be unpredictable due to the complex interactions of these factors. This forces us to talk about how likely or unlikely an outcome is, leading us to the notion of uncertainty.

A natural question that arises is how uncertainty can be quantified or, in other words, what is the mathematical framework for representing uncertainty? Usually, we are told that the fundamental tool for analyzing and managing uncertainty is probability, or more specifically, Kolmogorovian probability. However, there are several mathematical representations of uncertainty. Most of these representations, including the classical Kolmogorovian approach, share a couple of key basic ingredients in their way of representing uncertainty. Namely, an outcome space or set of possible worlds Ω\Omega, a collection of events or propositions \mathcal{F}, and a weight function f:[0,1]f: \mathcal{F} \to [0,1]. These ingredients will form what we call a coarse-grained representation of uncertainty. To understand each of these concepts we will make use of an example. Suppose that Monica from the TV show Friends spent the whole night cooking a cake she needs to bring to a party the next day. She goes to sleep confident of her hard work and the next day, she wakes up and half of the cake is missing. Immediately, Monica starts building a list of possible suspects who may have eaten the cake. The list includes each of her friends: Rachel (:=R)(:= R), Chandler (:=C)(:= C), Joey (:=J)(:= J), Ross (:=Ro)(:= Ro) and Phoebe (:=P)(:= P) and also possible ”combinations” of these. A possible list of suspects could be Ω={R,C,Ro,P,RC,RJ,RRo,CJ,CRo,CP,...,RCJRoP}\Omega = \{R, C, Ro, P, RC, RJ, RRo, CJ, CRo, CP, ... , RCJRoP\} where, the elements ”containing” more than one suspect, such as RC,RJ,RRo,RCJRoPRC, RJ, RRo, RCJRoP, etc, express the fact that it may be that all these suspects have eaten the cake together. For example, RCRC express the fact that it was Rachel and Chandler who ate the cake. Each element in Ω\Omega represents a possible scenario or a possible world. One important thing to note is that determining which worlds to include and which to exclude, along with deciding the depth of detail to represent each world, often entails a significant degree of subjective judgment from the agent. For example, if Monica believes in aliens, she might consider it important to include a world in which aliens ate the cake.

Each of the possible worlds may be seen as an event or proposition. However, we may also think of other interesting sets of events such as {R,C,J,Ro,P}\{R,C,J,Ro,P\}, which express that only one of Monica’s friends is guilty, or the one given by {C,J,CJ}\{C,J, CJ\}, which expresses the fact he who eat the cake was either Joey or Chandler, or both together, in contrast with {C,J}\{C, J\}, which states the fact that it was either Joey or Chandler who eat the cake but not both. In particular, we may think about events as sets made of possible worlds. Later, we will require that this collection of events satisfies some closure properties.

Given that Joey, Chandler, and Rachel are the ones who live much closer to Monica and also given that Joey is very fond of food, Monica can differentiate, in likelihood, the elements in the collection of propositions. This differentiation can be done by assigning different ”weights” to each event. The assignment of such weights can be done by means of a weight function, ff, which assigns to each event a number, the ”weight”, between 0 and 1, which represents the likelihood of such event. Often this weight function is construed as a probability measure. However, there are different ways in which we may think of ff. In the literature, these other ways of thinking about ff are often known by the name of imprecise probabilities.

In this post, we would like to motivate (some of) these other formal approaches to model uncertainty, and discuss some of the limitations that the measure-theoretic probability framework has in modeling uncertainty. Moreover, we would like start the ball rolling on exploring the possibility of studying imprecise probability through a categorical lens. In order to discuss these other formal approaches to model uncertainty we will first start by briefly summarizing what the Kolmogorovian probability theory framework is about.

Probability theory in a nutshell

Measure-theoretic probability is the predominant mathematical framework for understanding uncertainty. We first start with an outcome space Ω\Omega, also called sample space. The aforementioned collection of events \mathcal{F} has the structure of a σ\sigma-algebra, that is, a collection of subsets of Ω\Omega which is closed under complementation and under countable unions, that is, if U 1,U 2,...U_1, U_2, ... are in \mathcal{F}, then so are U 1¯,U 2¯,...\overline{U_1}, \overline{U_2},... and iU i\cup_i U_i. In this framework, the function which assigns the likelihood of an event is called a probability measure. Specifically, a probability measure is a set-theoretic function μ:[0,1]\mu: \mathcal{F} \to [0,1] such that:

μ(Ω)=1, \mu(\Omega)=1,

and

μ\mu is σ\sigma-additive, i.e., for all countable collections {U k} k=1 {\displaystyle \{U_{k}\}_{k=1}^{\infty }} of pairwise disjoint sets in \mathcal{F} we have

μ( k=1 U k)= k=1 μ(U k) \mu( \bigcup_{k=1}^\infty U_{k}) = \sum_{k=1}^\infty \mu(U_k)

These two axioms are usually known as the Kolmogorov axioms. For the sake of simplicity, in the rest of the post, we will consider that our set of outcomes Ω\Omega is finite. So, instead of talking about a σ\sigma-algebra, we can just talk about an algebra over Ω\Omega, which means it is enough to consider that the collection is closed under complementation and finite unions and instead of talking about σ\sigma-additivity we will be talking about additivity. In those cases when we have a finite set Ω\Omega, we will choose (unless specified) the power set algebra as the algebra over it and we will denote \mathcal{F} as 2 Ω2^\Omega. A sample space Ω\Omega together with a σ\sigma-algebra over \mathcal{F}, and a probability measure μ\mu on \mathcal{F} is called a probability space, and it is usually denoted by the triple (Ω,,μ)(\Omega,\mathcal{F},\mu).

Although this mathematical artillery indeed provides us with tools to model uncertainty, there are some vital questions that still need to be answered: what do the numbers we assign to a certain event represent? Where do these numbers come from? And, moreover, why should probabilities have these particular mathematical properties, for example the σ\sigma-additivity? Without answering these questions, assigning probabilities in practical scenarios and interpreting the outcomes derived from this framework will lack clarity.

For now, let’s leave aside the more ”technical” questions and focus on the more ”philosophical” ones. Even though probability theory is the mainstream mathematical theory to represent uncertainty, in the twenty-first century philosophers and mathematicians still have several competing views of what probability is. The two major currents of interpretation are the frequentist and the subjectivist or Bayesian interpretations.

The frequentist theory interprets probability as a certain persistent rate or relative frequency. Specifically, frequentists define the probability of an event as the proportion of times the event occurs in the long run, as the number of trials approaches infinity.While this explanation appears quite natural and intuitive, in practice, you cannot perform an experiment an infinite number of times. Moreover, interpreting probabilities as limiting frequencies can be nonsensical in some scenarios where we have non-repeatable or one-time events. On the other hand, the advocates of the subjective interpretation define probabilities just as numerical values assigned by an individual representing their degree of belief as long as these numerical assignments satisfy the axioms of probability. In both approaches, it can be proved, that the way of interpreting probabilities is compatible with the Kolmogorov axioms (see, for example, section 2.2.1 in [1].

Interpretations of probabilities are usually categorized into two main types: ontological and epistemological. Epistemological interpretations of probability view probability as related to human knowledge or belief. In this perspective, probability represents the extent of knowledge, rational belief, or subjective conviction of a particular human being. In contrast, ontological interpretations of probability consider probability as an inherent aspect of the objective material world, independent of human knowledge or belief.Therefore, we may view the frequentist current as an ontological interpretation whereas the subjective theory may be viewed as an epistemological one. Both points of view of interpreting probabilities, however, are perfectly valid and may be used depending on the particular situation. For example, a doctor’s belief about the probability of a patient having a particular disease based on symptoms, medical history, and diagnostic tests represents an epistemic probability. On the other hand, the probability of a particular isotope of uranium disintegrating in a year represents an objective probability since is related to the spontaneous transformation of the unstable atomic nuclei of the isotope, therefore, the probability exists as a characteristic of the physical world which is independent of human belief. In fact, this probability already existed before humans populated the Earth! Moreover, it seems that the objective interpretation may be more suitable for modeling processes in the framework of parameter estimation theory, while the subjective interpretation may be more useful for modeling decision-making by means of Bayes [2].

What is wrong with probabilities?

As we said before, in order for measure-theoretic probability to be a model of uncertainty it should also answer why probabilities have these specific mathematical properties. In particular we may question the additivity property. Measure-theoretic probability, with its additivity property, models situations of uncertainty where we still know a great deal about the system. Sometimes, however, the uncertainty is so high, or we know so little, that we do not have enough data to construct a probability measure. Let’s see a couple of examples that illustrate this problem.

Example 1: Suppose Alice has a fair coin and proposes the following bet to Bob. If the coin lands tails he must pay her 1€, and if the coin lands heads she must pay him 1€. Since the coin is fair, Bob is neutral about accepting the bet. However, suppose now that Alice has a coin with an unknown bias, and she proposes the same bet. What should Bob choose now? Should Bob refuse to play the game suspecting maybe the worst-case scenario in which Alice is using a coin with two tails?

Since Bob does not know about the bias of the coin, he cannot know the probability behind each of the outcomes. Therefore, his ignorance about the bias may preclude him from making a reasonable bet. This example highlights one of the major challenges of probability theory namely, its inability to effectively represent ignorance since even though we may still want to address this situation mathematically, we lack the required data to establish a probability measure.

Example 2: Imagine you have a bag of 100 gummies. According to the wrapper, 30%30\% of the gummies are red and the rest of them may be either green or yellow. Given that the exact proportion of green and yellow gummies is not known, it seems reasonable to assign a probability of 0.7 to choosing either a green (:= gg) or a yellow (:= yy) gummy and a probability of 0.3 to the outcome of choosing a red (:= rr) gummy. However, what is the probability of choosing just a green (or just a yellow) gummy?

In this example, we have information about the probabilities of a whole proposition {g,y}\{g, y\}, but not about the probabilities of each of its individual outcomes. Let’s take 2 Ω G2^{\Omega_G} as the algebra over Ω G={r,y,g}\Omega_G= \{r,y,g\}. This is the ”biggest” algebra we can have so, for sure, if we want to have information about the yellow and green gummies this choice will be helpful. In order to follow the approach of traditional probability theory, we must assign probabilities to all individual outcomes. However, this cannot be done, since we do not have information about how the 0.7 probability is distributed between the green and yellow gummies. Therefore, the fact that an agent is only able to assign probabilities to some sets may be seen as a problem. Of course, there is a natural way to avoid this problem: we can define a smaller algebra which excludes those subsets of the sample space that are ”problematic”. In particular, you may exclude the green and yellow singletons of your algebra, so that you have a probability measure which is consistent with the additivity axiom. However, by implementing this solution we cannot answer our original question either.

Moreover, we may even dare to say that human behaviour is not compatible with assigning probabilities to each of these singletons. Let’s consider the following bets:

  1. You get 1€ if the gummy is red, and 0€ otherwise (Fraction of the red gummies: 30%30\%).

  2. You get 1€ if the gummy is green, and 0€ otherwise (Fraction of green gummies: unknown).

  3. You get 1€ if the gummy is yellow, and 0€ otherwise (Fraction of the yellow gummies: unknown).

People usually prefer the first bet, and show themselves indifferent between the other two. By showing indifference they are suggesting that these two bets are equally likely. However, by means of this reasoning, they should not prefer the first bet since in this case the probability of drawing a yellow (or a green) gummy is of 0.35, which is bigger than that of drawing a red gummy. In general, any way of assigning probabilities to the yellow or green gummies will always make the second or the third bet (or both) more attractive than the first one. However, experimental evidence shows that humans prefer the first bet (see [1]), which tell us that humans do not assign probabilities to singletons and they just go for the ”safest” bet assuming maybe a worst-case scenario, like in the previous case.

Furthermore, if people had not only to rank the bets 1, 2 and 3 according to their preferences, but also assign rates to them, that is, how much they would pay for each bet, any numerical assignment following their ranking (and assigning a rate of 0.30.3 to red gummies) would necessarily be non-additive. Concretely, since people prefer bet 1 to bet 2 and bet 1 to bet 3, we would have that p({y})p(\{y\}) and p({g})p(\{g\}) are strictly smaller than p({r})p(\{r\}) so, the probability of both singletons would be smaller than 0.3 but, with this assignment we would never satisfy p({y,g})=p({y})+p({g})=0.7p(\{y,g\}) = p(\{y\}) + p(\{g\}) = 0.7. However, the violation of additivity by no means implies that we are being unreasonable or irrational.

Example 3: Imagine a football match between Argentina (:= A) and Germany (:= G). In this case, our outcome space Ω F\Omega_F will be given by three possible worlds namely, {A,G,D}\{A,G,D\}, where, DD denotes a draw.Initial assessments from a particular subject give both teams an equal and low probability of winning, say 0.1 each, since it is unclear to him who is going to win.
However, the subject has a strong belief based on additional reliable information that one of the teams will indeed win. So, the subject assigns a probability of 0.8 to the proposition {A,G}\{A,G\}.

According to classical probability theory, the chance that either Argentina or Germany wins is simply the sum of their individual probabilities, totaling 0.2, which is different to the probability of the proposition {A,G}\{A,G\} and therefore, we may say this assignment although reasonable is incompatible with the additivity requirement. Classical probability struggles with this scenario because it can’t flexibly accommodate such a high level of uncertainty without overhauling the entire probability distribution.

Handling ignorance by imprecise probabilities

Sets of probability measures: Lower and upper probabilities

In order for Bob, in Example 1, to choose the most reasonable bet he should answer an important question: how should the bias of the coin be represented? One possible way of representing the bias of the coin is to consider, instead of just one probability measure, a set of probability measures, each of which corresponds to a specific bias, that is, we may consider the set 𝒫 C={μ α:α[0,1]}\mathcal{P}_C = \{\mu_\alpha \, : \, \alpha \in [0,1]\}, where α\alpha denotes the bias of the coin. Now, we may handle uncertainty by defining the probabilities as follows: μ α({heads})=α\mu_\alpha(\{heads\}) = \alpha and μ α({tails})=1α\mu_\alpha(\{tails\})= 1-\alpha. This set of probabilities not only allows us to handle our ignorance about the bias of the coin, but also allows us to construct intervals to bound our ignorance. To construct such intervals we need to define what are called lower and upper probabilities. Specifically, if 𝒫\mathcal{P} is a set of probability measures defined over 2 Ω C2^{\Omega_C} where, Ω C={head,tails}\Omega_C = \{head, tails\}, and UU is an element of 2 Ω C2^{\Omega_C}, we define the lower probability as

𝒫 *(U)=inf𝒫(U), \mathcal{P}_*(U) = \inf \, \mathcal{P}(U),

and the upper probability as 𝒫 *(U)=sup𝒫(U). \mathcal{P}^*(U) = \sup \, \mathcal{P}(U).

The interval [[𝒫 *(U),𝒫 *(U)]][[\mathcal{P}_{\ast}(U), \mathcal{P}^{\ast}(U)]] is called estimate interval, since its length is a way of measuring the ambiguity or ignorance about the event UU. For the case of the coin, we may see that the estimate intervals for the two possible outcomes, both have a length of 1, which tells us that there is maximal uncertainty about these events.

In spite of the names, upper and lower probabilities are not actually probabilities because they are not additive, instead lower probabilities are super-additive, and upper probabilities are sub-additive.However, in contrast with probability measures where the additivity property defines them, lower and upper probabilities are neither defined nor completely characterized by the super or sub-additivity property (the property that characterizes them is rather complex, those interested readers can refer to [1]. By allowing for a range of possible probability assignments, the approach of a set of probability measures allows uncertainty to be addressed. Moreover, lower and upper probabilities provide us with a way of bounding such uncertainty.

Inner and outer measures

As we already discussed, one of the problems of probabilities arises when the agent is not able to assign probabilities to all measurable sets. However, we may ask ourselves, what happens if we just ”remove” those measurable sets for which we do not have information? To illustrate these ideas, let’s use example 2. For this specific example, since we only have information about the numbers of red gummies, and the number of green and yellow gummies (togheter), we may consider the following sub-collection of events 𝒮 G={{r},{g,y},{r,g,y},}.\mathcal{S}_G= \{\{r\}, \{g,y\}, \{r,g,y\}, \emptyset\}.. This sub-collection of events can be proven to be an algebra over Ω G\Omega_G. Moreover, this ”smaller” algebra is actually a subalgebra of the power set algebra 2 Ω G2^{\Omega_G} since 𝒮 G2 Ω G\mathcal{S}_G \subset 2^{\Omega_G}. On this subalgebra we can define the measure μ 𝒮 G\mu_{{\mathcal{S}}_G} given by μ 𝒮 G({r})=0.3\mu_{{\mathcal{S}}_G}(\{r\})=0.3 and μ 𝒮 G({g,y})=0.7\mu_{{\mathcal{S}}_G}(\{g,y\}) = 0.7 which is well defined, it tells the same story illustrated in example 2, and it is consistent with Kolmogorov’s axioms. Since in this setting we have ”removed” those singletons sets corresponding to the yellow and green gummies, these events are undefined, and in principle, we cannot say anything about them. However, let’s define the following set of probability measures on the algebra 2 Ω G2^{\Omega_G}, the set 𝒫 G={μ β:β[0,0.7]}\mathcal{P}_G = \{\mu_\beta : {\beta \in [0,0.7]}\} where, μ β({r})=0.3\mu_\beta(\{r\}) = 0.3, μ β({g})=β\mu_\beta(\{g\}) = \beta and μ β({y})=0.7β\mu_\beta(\{y\}) = 0.7 - \beta. One thing we may notice is that each measure μ β𝒫 G\mu_\beta \in \mathcal{P}_G is actually an extension of the measure μ 𝒮 G\mu_{{\mathcal{S}}_G}, that is, for each β\beta, the measures coincide on all sets of 𝒮 G\mathcal{S}_G. Of course, if UU belongs to 2 Ω G𝒮 G2^{\Omega_G} \setminus \mathcal{S}_G, then μ 𝒮 G(U)\mu_{\mathcal{S}_G}(U) is not defined. But the extension of the measure is. That is, we may ask if it is possible to extend μ 𝒮 G\mu_{\mathcal{S}_G} to the whole algebra 2 Ω G2^{\Omega_G} so that we may have some information about those indefinite events that we want to studyt. Fortunately, this is indeed possible. Specifically, there are two canonical ways of extending μ 𝒮 G\mu_{\mathcal{S}_G} [3]. They are called inner and outer measures and they are defined, in general, as follows: let 𝒮\mathcal{S} a sub-algebra of 2 Ω2^\Omega over a (finite) outcome space Ω\Omega, μ S\mu_S a measure defined over the subalgebra and U𝒮U \in \mathcal{S}. We define the inner measure induced by μ S\mu_S as

μ 𝒮 *(U)=sup{μ(V):VU,V𝒮}, {\mu_\mathcal{S}}_*(U) = \sup \bigl\{ \mu(V) \, : \, V \subseteq U, V \in \mathcal{S} \bigr\},

that is, as the largest measure of an 𝒮\mathcal{S}-measurable set contained within UU. On the other hand, the outer measure induced by μ S\mu_S is defined by

μ 𝒮 *(U)=sup{μ(V):VU,V𝒮}, {\mu_\mathcal{S}}^*(U) = \sup \bigl\{\mu(V) \, : \, V \supseteq U, \, V \in \mathcal{S} \bigr\},

that is, as the smallest measure of an 𝒮\mathcal{S}-measurable set containing UU. Therefore, in example 2 we have μ 𝒮G*({r})=μ 𝒮G *({r})=0.3{\mu_{{\mathcal{S}}G}}{\ast}(\{r\}) = {\mu_{{\mathcal{S}}G}}^{\ast}(\{r\}) = 0.3, μ𝒮G*({y})=μ 𝒮G*({g})=0{\mu{{\mathcal{S}}G}}{\ast}(\{y\}) = {\mu_{{\mathcal{S}}G}}{\ast}(\{g\}) = 0, and μ 𝒮G *({g})=μ𝒮 G *({y})=0.7{\mu_{{\mathcal{S}}G}}^{\ast}(\{g\}) = {\mu{{\mathcal{S}}_G}}^{\ast}(\{y\}) = 0.7. Here, again by means of the outer and inner measures, we may define interval estimates. If we define such intervals we have that the uncertainty in the event of choosing a red gummy is 0, while the uncertainty of the two other events will be 0.7. Just as with lower and upper probabilities, inner and outer measures are not probability measures: inner measures and outer measures are super-additive and sub-additive, respectively, instead of additive. By offering lower and upper bounds, inner and outer measures enable us to bound our ignorance. Moreover, they also allow us to ”deal” with those indefinite or non-measurable events that we leave out of our algebra from the very beginning, giving us information about them by considering the best possible approximations from within (inner) and from outside (outer).

Belief functions

As the examples given above have shown, additivity may be sometimes artificial. As we have seen, upper/lower probabilities and inner/outer measures, which are better at handling ignorance than probability measures, do not satisfy this requirement. Moreover, there exists a type of weighted function for which superadditivity (motivated above by Example 3) is actually part of its axiomatic definition. These functions are the so-called belief functions. They were introduced by Arthur P. Dempster in 1968 and expanded by Glenn Shafer in 1976. The Dempster-Shafer theory offers a powerful framework for representing epistemic uncertainty. Unlike classical probability theory, Dempster-Shafer theory allows for the representation of ignorance and partial belief, thus providing a more flexible approach to handling uncertainty.

A fundamental distinction between probability measures and Dempster-Shafer theory lies in their approach to additivity. While in classical probability, the Kolmogorov axioms enforce finite additivity, the Dempster-Shafer theory adopts finite superadditivity: Bel(UV)Bel(U)+Bel(V), for UV=Bel(U \cup V) \geq Bel(U) + Bel(V), \text{ for } U \cap V = \emptyset. This superadditivity allows Dempster-Shafer theory to capture uncertainty in a way that classical probability cannot. By not requiring strict additivity, this theory accommodates situations where the combined belief in two events can be greater than the sum of their individual beliefs, reflecting a more nuanced understanding of uncertainty.

To utilize Dempster-Shafer theory, we start with the concept of a frame of discernment, Ω\Omega, which represents all possible outcomes in a given context (playing an analogous role as the sample space in probability theory). For example, in a football match between Argentina and Germany (Example 3 mentioned above), the frame of discernment would be: Ω={A,G,Draw}\Omega = \{A, G, \text{Draw}\}, where AA denotes “Argentina wins,” GG denotes “Germany wins,” and “Draw” represents a tie. Note that while Ω\Omega also denotes the sample space in probability theory, here it is used to define the frame of discernment.

The power set of Ω\Omega, denoted 2 Ω2^\Omega, as previously explained when discussing probability theory, includes all subsets of Ω\Omega, representing all possible events:

2 Ω={,{A},{G},{Draw},{A,G},{A,Draw},{G,Draw},{A,G,Draw}}2^\Omega = \{\emptyset, \{A\}, \{G\}, \{\text{Draw}\}, \{A, G\}, \{A, \text{Draw}\}, \{G, \text{Draw}\}, \{A, G, \text{Draw}\}\}.

A mass function m:2 Ω[0,1]m : 2^\Omega \rightarrow [0, 1] distributes belief across the elements of 2 Ω2^\Omega. This mass function must satisfy two key properties: m()=0m(\emptyset) = 0, which implies the empty set has zero belief, and X2 Ωm(X)=1\sum_{X \in 2^\Omega} m(X) = 1 which tells us that the total belief across all subsets of Ω\Omega sums to one.

This framework allows us to represent ambiguity and partial belief without requiring full certainty in any single outcome or in the entire frame. To illustrate this, let’s continue with the example of the football match between Argentina and Germany. Using classical probability, if we believe there is an equal likelihood for either team to win, we might assign: μ(A)=μ(G)=0.1μ(AG)=0.2.\mu(A) = \mu(G) = 0.1 \Rightarrow \mu(A \cup G) = 0.2. In Dempster-Shafer theory, we can represent partial beliefs more flexibly. For instance: Bel(A)=Bel(G)=0.1,Bel(A) = Bel(G) = 0.1, but given additional information suggesting a high likelihood that one team will win, we might have: Bel(AG)=0.8.Bel(A \cup G) = 0.8. This reflects a stronger belief in the combined event without committing to either team individually.

In Dempster-Shafer theory, there are two key functions quantifying belief: the belief function BelBel and the plausibility function PlPl. The belief function Bel(U)Bel(U) sums the masses of all subsets XX contained within UU :

Bel(U)= XUm(X). Bel(U) = \sum_{X \subseteq U} m(X).

Belief function

The plausibility function Pl(U)Pl(U) sums the masses of all subsets XX that intersect UU:

Pl(U)= XUm(X). Pl(U) = \sum_{X \cap U \neq \emptyset} m(X).

Plausibility function

These functions provide lower and upper bounds on our belief in a hypothesis UU.

Returning to our football match example, suppose we have the following Basic Belief Assignments: m({A})=0.1,m({G})=0.1,m({Draw})=0.2m(\{A\}) = 0.1, m(\{G\}) = 0.1 , m(\{\text{Draw}\}) = 0.2, and m({A,G})=0.6m(\{A, G\}) = 0.6.

We can then calculate the respective belief and plausibility functions as follows:

Belief Functions:

  • Bel(A)=0.1Bel(A) = 0.1
  • Bel(G)=0.1Bel(G) = 0.1
  • Bel(AG)=0.8Bel(A \cup G) = 0.8

Plausibility Functions:

  • Pl(A)=0.7Pl(A) = 0.7
  • Pl(G)=0.7Pl(G) = 0.7
  • Pl({Draw})=0.2Pl(\{\text{Draw}\}) = 0.2

Notice that belief and plausibility functions are related by the following equation: Pl(U)=1Bel(U¯)Pl(U) = 1 - Bel(\overline{U}). This relationship shows that plausibility represents the extent to which we do not disbelieve U.

Finally, an essential feature of Dempster-Shafer theory is Dempster’s rule of combination, which allows for the integration of evidence from multiple sources. Given two independent bodies of evidence, represented by mass functions m 1m_1 and m 2m_2 over the same frame Ω\Omega, the combined mass function is:

(m 1m 2)(U)=11K U 1U 2=Um 1(U 1)m 2(U 2),U, (m_1 \oplus m_2)(U) = \frac{1}{1-K} \sum_{U_1 \cap U_2 = U} m_1(U_1) m_2(U_2), \quad \forall U \neq \emptyset,

where, KK is the normalization factor representing the total conflict between m 1m_1 and m 2m_2:

K= U 1U 2=m 1(U 1)m 2(U 2). K = \sum_{U_1 \cap U_2 = \emptyset} m_1(U_1) m_2(U_2). Dempster’s rule ensures consistency by requiring that K<1K &lt; 1, meaning there is not total conflict between the two evidence sources.

As we have already seen one of the main differences between the measure theoretic approach and the approaches of imprecise probabilities discussed here is the relaxation of the additivity condition. However, it is worth saying that in other approaches of imprecise probability this condition is not even considered or is substituted for another one as happens in the case of possibility measures (see, for example, [1]). Each of the different approaches for handling uncertainty may be useful in different scenarios, and it will depend on the particular case we have to use one or another. The measure-theoretic probability is a well-understood framework and it has extensive support of technical results. However, as we stated here, this framework is by no means the only one and not necessarily the best one for every case-scenario. The set of probability measures extends the traditional probability approach by allowing for a range of possible probabilities, which is useful when there is uncertainty about the likelihoods themselves but, some information about the parameters “indexing” the probabilities is required. Belief functions, on the other hand, have proven themselves robustly effective in modelling and integrating evidence especially when combined with Dempster’s Rule of Combination [1]. Other approaches, which are not discussed in here, like partial preorders, possibility measures and ranking function may also be an interesting option to address uncertainty, in particular, for dealing with counterfactual reasoning [1].

Monads and imprecise probabilities

Recently, within the context of category theory numerous diagrammatic axioms have been proposed, facilitating the proof of diverse and interesting theorems and constructions in probability, statistics and information theory (see, for example, [4], [5], [6], [7], [8], [9], [10], [11], [12], [13], [14]). Because of this, a shift in the perspective of the foundational structure of probability theory has gained substantial momentum. In this new perspective a more synthetic approach to measure-theoretic probability is sought. The natural framework for this approach turns out to be that of a Markov category, that is, a symmetric monoidal category (C,,I,s)(\mathsf{C},\otimes, I,s) in which every object is endowed with a commutative comonoid structure [9]. Since Markov categories have been serving as a fertile environment in providing new insights on probability theory it is natural to ask if these other approaches in handling uncertainty (lower/upper probabilities, inner/outer measures, belief functions, etc) may fit into this new synthetic perspective.

In some Markov categories the morphisms may be understood as ”transitions with randomness”, where this randomness may be “indentified” by means of the comonoid structure, that is, with this structure we may put a distinction between morphisms that involve randomness and those that do not [15]. Nevertheless, recent investigations show that their randomness may also be understood ”as the computational effect embodied by a commutative monad” [16]. More precisely, constructions of so-called representable Markov categories [12] can be understood as starting from a Cartesian monoidal category (with no randomness) and passing to the Kleisli category of a commutative monad that introduces morphisms with randomness. For example, in the category Set\mathsf{Set} we may define the distribution monad (P,δ,μ)(\text{P}, \delta, \mu) where P\text{P} is the distribution functor assigning to each set XX the set of all finitely supported probability measures over XX and to each morphism f:XYf: X \to Y the morphism Pf:PXPY\text{P}f: \text{P}X \to \text{P}Y given by the pushforward measure. Here, the unit map δ X:XPX\delta_X: X \to \text{P}X is the natural embedding

δ:XPX,xδ x, \delta: X \to \text{P}X, \quad x \mapsto \delta_x,

where, δ x:X[0,1],yδ x(y)\delta_x: X \to [0,1], y \mapsto \delta_x(y) with δ x(y)=1\delta_x(y) = 1 for x=yx=y and δ x(y)=0\delta_x(y) = 0 otherwise, and the multiplication map is given by

μ X:PPXPX,πμ X(π), \displaystyle \mu_X: \text{P}\text{P}X \to \text{P}X, \quad \pi \mapsto \mu_X(\pi),

which assigns to each measure π\pi over PXPX the ”mixture” measure μ X(π)\mu_X(\pi) over XX defined by μ X(π)(x)= pPXπ(p)p(x). \mu_X(\pi)(x) = \sum_{p \, \in \, \mathrm{P}X} \pi(p) p(x).

With this structure in mind, you may think about morphisms of the type f:XPYf: X \to \text{P}Y. In the case of the distribution monad described above, to each xXx \in X, they assign a finitely supported probability measure f xf_x over YY. These can be seen as morphisms with an uncertain output, or as some sort of generalized mapping allowing more ”general” outputs. We call these morphisms Kleisli morphisms [17]. For the particular case of the distribution monad, the Kleisli morphismsf:XPYf: X \to \text{P}Y where each xx is mapped to the probability distribution on Y, that is, to a function such that, yYf(y|x)y \in Y \mapsto f(y|x), with f(y|x)f(y|x) the components of a stochastic matrix. Moreover, the Kleisli category Kl(P)\mathsf{Kl}(\mathrm{P}), whose objects are those of Set\mathsf{Set}, and whose morphisms are the Kleisli morphisms can be endowed with a commutative comonoid structure. Furthermore, since we have PII\text{P}I \cong I, one can show that the Kleisli category Kl(P)\mathsf{Kl}(\text{P}) is a Markov category [12]. In fact, the category FinStoch\mathsf{FinStoch} is a full subcategory of it. Several Markov categories can be obtained in this way, for example Stoch\mathsf{Stoch}, is the Kleisli category of the Giry monad on Meas\mathsf{Meas} (see [12] and the other examples in there).

An intriguing question then emerges from the topics discussed herein: Is there a commutative monad (,η,μ)(\mathcal{I},\eta,\mu) over some suitable category such that the morphisms of its Kleisli category Kl()\mathsf{Kl}(\mathcal{I}) model a certain type of imprecise probability? Or in other words, do imprecise probabilities (upper and lower probabilities, inner and outer measures, belief functions, etc) form Markov categories? This possible interesting byproduct between imprecise probabilities and category theory has already captured the attention of members of the communities of Probability and Statistics and Category Theory. Moreover, quite recently Liell-Cock and Staton have tried to address a similar question in [18]. Tobias Fritz and Pedro Terán have also studied a similar question involving possibility measures in an unpublished work. We hope very soon that a plethora of works is dedicated to studying such interesting byproduct.

References

  1. Joseph Y. Halpern. (2017). Reasoning about uncertainty. The MIT Press.

  2. Edwin T. Jaynes. (2003). Probability Theory. The Logic of Science. Cambdridge University Press.

  3. Marshall Evans Munroe. (1953). Introduction to MEASURE AND INTEGRATION. Cambridge, Mass., Addison-Wesley.

  4. Michèle Giry (1982).A categorical approach to probability theory. In: Banaschewski, B. (eds) Categorical Aspects of Topology and Analysis. Lecture Notes in Mathematics, Volume 915. Springer, Berlin, Heidelberg. DOI: 10.1007/BFb0092872

  5. Prakash Panangaden (1999). The Category of Markov Kernels. Electronic Notes in Theoretical Computer Science, Volume 22. DOI: 10.1016/S1571-0661(05)80602-4

  6. Peter Golubtsov (2002). Monoidal Kleisli Category as a Background for Information Transformers Theory. Processing 2, Number 1.

  7. Jared Culbertson and Kirk Sturtz (2014). Postulates for General Quantum Mechanics. Appl Categor Struct, Volume 22. DOI: 10.1007/s10485-013-9324-

  8. Tobias Fritz, Eigil Fjeldgren Rischel (2020). Infinite products and zero-one laws in categorical probability. Compositionality, Volume 2. DOI: 10.32408/compositionality-2-3

  9. Tobias Fritz (2020). A synthetic approach to Markov kernels, conditional independence and theorems on sufficient statistics. Advances in Mathematics, Volume 370. DOI: 10.1016/j.aim.2020.107239

  10. Bart Jacobs and Sam Staton (2020). De Finetti’s Construction as a Categorical Limit. In: Petrişan, D., Rot, J. (eds) Coalgebraic Methods in Computer Science. CMCS 2020. Lecture Notes in Computer Science(), vol 12094. Springer, Cham. DOI: 10.1007/978-3-030-57201-3_6

  11. Tobias Fritz,Tomás Gonda and Paolo Perrone (2021). De Finetti’s theorem in categorical probability. Journal of Stochastic Analysis, Volume 2. DOI: 10.31390/josa.2.4.06

  12. Tobias Fritz, Tomás Gonda, Paolo Perrone and Eigil Fjeldgren Rischel (2023). Representable Markov categories and comparison of statistical experiments in categorical probability. Theoretical Computer Science, Volume 961. DOI: 10.1016/j.aim.2020.107239

  13. Tobias Fritz and Andreas Klingler (2023). The d-separation criterion in Categorical Probability. Journal of Machine Learning, Volume 24. DOI: 22-0916/22-0916

  14. Sean Moss and Paolo Perrone (2023). A category-theoretic proof of the ergodic decomposition theorem. Ergodic Theory and Dynamical Systems, Volume 43. DOI: 10.1017/etds.2023.6

  15. Paolo Perrone (2024). Markov Categories and Entropy. IEEE Transactions on Information Theory, Volume 70. DOI: 10.1109/TIT.2023.3328825.

  16. Sean Moss and Paolo Perrone (2022). Probability monads with submonads of deterministic states. In: Proceedings of the 37th Annual ACM/IEEE Symposium on Logic in Computer Science. Association for Computing Machinery, LICS ‘22. DOI: 10.1145/3531130.3533355.

  17. Paolo Perrone (2024). Starting Category Theory. World Scientific Book.

  18. Jack Liell-Cock, Sam Staton (2024). Compositional imprecise probability. arXiv: 2405.09391.

Posted at July 9, 2024 4:37 PM UTC

TrackBack URL for this Entry:   https://golem.ph.utexas.edu/cgi-bin/MT-3.0/dxy-tb.fcgi/3540

13 Comments & 0 Trackbacks

Re: Imprecise Probabilities: Towards a Categorical Perspective

This is great to see! I’m very curious to hear what you know about this from the categorical perspective, and in particular which models of imprecise probability can be formalized as commutative monads. You’ve made a good argument for generalizing probability distributions to sets of probability distributions, and this has resulted in upper and lower probabilities via 𝒫 *(U)=inf𝒫(U),𝒫 *(U)=sup𝒫(U).\mathcal{P}_*(U) = \inf \: \mathcal{P}(U), \qquad \mathcal{P}_*(U) = \sup \: \mathcal{P}(U). But which sets of distributions can be characterized in terms of their lower and upper probabilities? I guess it makes sense to consider only closed convex subsets of PΩP \Omega to begin with. For example for |Ω|=3|\Omega| = 3, the set of all probability distributions is a triangle, and taking the lower and upper probabilities then amounts to projecting a closed convex subset of the triangle onto its sides. There certainly are different closed convex subsets of the triangle which have the same projections, right? To conclude, lower and upper probabilities are not sufficient to describe a (closed convex) set of distributions. But perhaps there are other reasons to take inner and outer measures themselves seriously as models of imprecise probability, even if it’s a coarser model than sets of distributions?
Posted by: Tobias Fritz on July 10, 2024 5:38 AM | Permalink | Reply to this

Re: Imprecise Probabilities: Towards a Categorical Perspective

Oops, sorry about the lack of paragraph breaks – looks like these don’t come across if you chose the ‘itex to MathML’ text filter.

Posted by: Tobias Fritz on July 10, 2024 5:43 AM | Permalink | Reply to this

Re: Imprecise Probabilities: Towards a Categorical Perspective

Hi Tobias, thanks for your comment. Sorry for the late reply, but sometimes life comes in between. We are interested in answering exactly the same question. We believe a monad structure may be defined for models of imprecise probabilities however, we do not expect all of these monads to be commutative. In fact, we were wondering what categorical probability one can do in those non-commutative cases. During the Adjoint School, we were working on getting a monad structure for a “weaker” version of Dempster-Shafer functions but we are still working on it.

About your comment about lower and upper probabilities, they, indeed, offer less information than other kinds of imprecise probabilities. However, you may also think that sometimes you have precisely little information, and we were just wondering if that would form a monad or a Markov category.

Now, regarding your question about which sets of probability distributions can be characterized in terms of their lower and upper probabilities, I think you may need more than only closed convex subsets of 𝒫Ω\mathcal{P}\Omega. Let’s take the case of |Ω|=3|\Omega|=3, where Ω={x 1,x 2,x 3}\Omega =\{x_1,x_2,x_3\}. As you said, different closed convex subsets of the triangle will have the same projections, and therefore they will be characterized by the same lower and upper probabilities however, as you may see from the drawing bellow these convex halls do not include all possible probabilities. So, I think that to characterize these sets you will need to find not quite closed convex subsets but sets of distributions that consider of all these probabilities that are left outside when you consider the convex hall.

Untitled 2 (1)

Posted by: Laura Gonzalez Bravo on October 8, 2024 4:17 PM | Permalink | Reply to this

Re: Imprecise Probabilities: Towards a Categorical Perspective

Nice to hear from you, Laura! That all makes good sense to me, especially the lack of commutativity. Last month at SMPS, Pedro Terán and I were discussing whether there might be a Markov category for nonlinear expectations, which are a very general form of imprecise probability. Our conclusion also has been that commutativity of the monad fails, due to the failure of Fubini’s theorem in this setting.

Posted by: Tobias Fritz on October 10, 2024 7:11 PM | Permalink | Reply to this

Re: Imprecise Probabilities: Towards a Categorical Perspective

On a related note, do you know of any flavour of imprecise probabilities which can be formalized as a commutative monad?

Posted by: Tobias Fritz on October 10, 2024 7:13 PM | Permalink | Reply to this

Re: Imprecise Probabilities: Towards a Categorical Perspective

I would be very interested in hearing how you might approach the idea of belief. It is embedded in your example. What does Monica believe? And of course we can believe things for which probabilistic accounts seem a bit forced, such as counterfactuals and singular events that are non-repeating. Lately, we have done some experiments asking people about their degrees of disbelief from a perspective of negative ranking theory. It is item 3.3 on this Stanford Encyclopedia of Philosophy page* and is the brain child of Wolfgang Spohn. It is sufficiently mathematical to admit a categorical treatment. If you decide to expand beyond imprecise probabilities I hope you consider negative ranking theory as a possible additional theory to be categorified.

  • https://plato.stanford.edu/archives/spr2016/entries/formal-belief/
Posted by: Britt Anderson on July 10, 2024 9:38 AM | Permalink | Reply to this

Re: Imprecise Probabilities: Towards a Categorical Perspective

Actually, a (complete) negative ranking function is essentially the same thing as a possibility measure, as the transform xe xx\mapsto e^{-x} is invertible and turns the ranking function into a possibility measure. Then a possibility measure is a particular case of both a plausibility measure and an upper probability, so negative ranking functions can be accomodated into the framework of this post.

Possibility measures are, loosely speaking, probability measures but with addition being replaced by the maximum.

Posted by: Pedro Terán on July 11, 2024 8:38 AM | Permalink | Reply to this

Re: Imprecise Probabilities: Towards a Categorical Perspective

In Shafer’s approach in his book `A mathematical theory of evidence’ (belief and plausibility measures/functions as explained in the post), you distinguish between evidence in favor of A and lack of evidence against A, which arguably cannot be done in traditional probability theory as the sum P(A)+P(notA) is fixed.

A high degree of plausibility for A means you have little evidence against A being the case. But you may have equally scarce positive evidence for A. It might even happen that you have little negative evidence about A but no positive evidence whatsoever, in which case the belief in A would be 0 and its plausibility close to 1.

In the Friends example, if Rachel tells Monica that she had taken a sleeping pill that night, this evidence against Rachel’s guilt will carry some weight with her friend Monica. Thus it raises the plausibility that Joey or Chandler are guilty but it doesn’t raise the belief that specifically Joey or specifically Chandler is guilty.

In the gummies example, if we learn that only 20% of the gummies are actually red (instead of 30%) we’ll think that there’s 10% more which must be either green or yellow, so it becomes more plausible that a random gummy will be green, but the degree of belief does not increase since, for all we know, the extra 10% could all be yellow.

I hope this clarifies a bit the semantics of belief in Dempster-Shafer theory.

Posted by: Pedro Terán on July 11, 2024 9:15 AM | Permalink | Reply to this

Re: Imprecise Probabilities: Towards a Categorical Perspective

Interesting post! A trivial point for now, but the iTeX used here renders consecutive letters as non-italicized, so that if you wanted ‘Ro’ for Ross to appear like the ‘R’ for Rachel, you need to include a space between letters: RoR o.

Posted by: David Corfield on July 12, 2024 7:55 AM | Permalink | Reply to this

Re: Imprecise Probabilities: Towards a Categorical Perspective

Is there a reason not to want to understand Example 1 via the notion of a probability distribution over probability distributions? Imagine Bob is committed to tosses being iid or exchangeable, then his beliefs will be representable as a distribution over Bernouilli distributions. One may imagine then how different information about the properties of the coin and the properties of the tossing mechanism would shape this meta-distribution. E.g.,

  1. You see the coin arrive from the Mint; you saw Alice emerge from a shop called ‘Loaded Coins’; you’re allowed to inspect and measure the coin;…

  2. You can check out the mechanical tossing device to be used; you’re allowed to choose on which side the coin is placed on the device;…

Nervousness about betting in such situations often arises from the fear that the other person knows more than you and can use it to their advantage. Allowing Bob to choose which side counts as a win for him should free him from that worry.

In Example 2, there must be a place at which this phenomenon changes. E.g., if told only 2% of the gummies are red, won’t people opt for green or yellow? And you’d likely do this even if told all of the remainder are all the same colour. Wouldn’t you prefer a 50% chance at a 98% shot to a 2% chance?

Clearly people differ as to risk, e.g., might prefer a certain £100 to a 1/2 chance of £500, but then this is affected by the nonlinearity of the value of money to someone.

It seems odd to me to carry this risk-aversion over to the case of a fixed value win with uncertainty about probabilities. Wouldn’t I be indifferent between the result of a fair coin and the result of a coin I know to be double-headed or double-tailed but not which? People happily play scratch cards where the result is already determined, or call a coin toss already landed on the back of a hand.

Of course, learning of the results of tosses leads to updating of the meta-distribution. I seem to remember this all being in Jaynes [2] somewhere.

Posted by: David Corfield on July 12, 2024 11:35 AM | Permalink | Reply to this

Re: Imprecise Probabilities: Towards a Categorical Perspective

“Wouldn’t I be indifferent between the result of a fair coin and the result of a coin I know to be double-headed or double-tailed but not which?”

I think there may be a hidden assumption here: if you are indifferent about the result of a process under two sets of conditions then you can use the same mathematical model for both sets of conditions.

But imagine that the coin is tossed not once but a very large number of times, ideally with total independence between tosses. This is a more complex process that includes the former as a particular case and the conditions are the same.

If you believe the coin is fair and you model that by giving heads and tails the same probability, the law of large numbers will force you to believe as well that the frequency of heads will be about 1/2 with a very large probability.

If you believe that the coin has two heads or two tails, you will also believe that the frequency of heads will be 1 or 0 but you don’t know which. But, if you model the condition by giving heads and tails the same probability, the LLN still tells you that the frequency of heads will be close to 1/2 with a very large probability.

There are actually laws of large numbers for imprecise probabilities which are more consistent with those situations. If the coin is loaded in such a way that the probability of heads is between .2 and .6 but you know nothing about the exact value, the conclusion of these LLNs is that the long-term frequency of heads in the long run will be in or very close to the interval [.2, .6] with a very high lower probability (thus also upper probability). Whereas by giving heads an `indifferent’ probability of .4, or using a meta-distribution in the interval [.2,.6], you have to commit to the belief that, in practice, the long-term frequency will be very close to .4 or some other specific value.

Posted by: Pedro Terán on July 16, 2024 6:35 PM | Permalink | Reply to this

Re: Imprecise Probabilities: Towards a Categorical Perspective

There’s some typos in the sentence starting “Therefore, in example 2 we have …” around the super/subscripts on the upper/lower probabilities.

Posted by: David Roberts on July 16, 2024 2:52 AM | Permalink | Reply to this

Re: Imprecise Probabilities: Towards a Categorical Perspective

Whoops, sorry, I meant inner and outer measures. Also, the definition of outer measure should have an inf\inf, not a sup\sup.

Posted by: David Roberts on July 16, 2024 2:54 AM | Permalink | Reply to this

Post a New Comment