Open access peer-reviewed chapter

Some Remarks on Vector Markov Chains and Their Applications to the Description of Many-Particle Systems

Written By

Bruno Carbonaro and Federica Vitale

Submitted: 20 February 2024 Reviewed: 22 February 2024 Published: 31 July 2024

DOI: 10.5772/intechopen.1004791

From the Edited Volume

Stochastic Processes - Theoretical Advances and Applications in Complex Systems

Don Kulasiri

Chapter metrics overview

23 Chapter Downloads

View Full Metrics

Abstract

Although in the literature on the topic, this point is never expressed, but regularly implied, an important and widely studied scheme of models to describe and predict, at least in stochastic terms, the behavior of many-particle systems is based on Markov Chains. In fact, Markov Chains are special stochastic processes with important properties that have revealed to be shared with many natural processes. This fact, though almost obvious, requires to be carefully discussed, at least to point out that the very notion of a Markov Chain must be completed and somehow generalized to understand the way in which Markov Chains intervene in the equations describing the stochastic behavior of any system. This is not a purely critical and “foundational” task but aims to develop a new very plausible and likely way to introduce possible effects of external world on a non-isolated many-particle system. This paper aims to offer at least a basic discussion of this perspective.

Keywords

  • stochastic processes
  • Markov Chains
  • probability
  • kinetic theory
  • mathematical models
  • complex systems

1. Introduction

The aim of the present chapter is twofold.

On one hand, we want to consider explicitly some properties of a kind of Markov Chains [1, 2, 3], which is of course taken implicitly into account in the whole literature about these important stochastic processes but—as far as we are aware—has been never examined explicitly, that is, the Markov Chains whose states are n-tuples of integers or of real numbers, which can be called vector Markov Chains. In particular, we will also pay attention in particular to nonstationary Markov Chains, which we recall in some details in Section 1.2.

On the other hand, we want to examine the equations of kinetic-theoretic models [4] that since almost forty years seem to be diffusing ever more in the research about the evolution of many-particle systems and appear to be a particularly effective, expressive, and versatile tool to describe in stochastic terms such evolution [5, 6, 7, 8, 9, 10]. Rather strange to say, these equations are firmly based on the use of Markov Chains, since in them a basic rôle is played by transition matrices, that just characterize these important stochastic processes, but in the literature about the kinetic-theoretic models Markov Chains are never even mentioned (at least to our knowledge). The researchers in this field probably consider this reference superfluous. But the question spontaneously arises whether if we exploit the rôle of Markov Chains then our understanding of the terms of equations could be improved, suggesting to us some modifications that could help us to produce a good and effective description of the behavior of systems that have till now escaped the classical equations used up to the last few years (in this connection, see [11, 12], where an accurate stochastic description of the evolution of a system is obtained renouncing the Principle of Inertia, and [13, 14, 15] where new perspectives about the description of non-isolated systems have been obtained by introducing «forcing» terms in the equations).

Accordingly, after giving some definitions recalling some types of Markov Chains and introducing explicitly the vector Markov Chains, of which we shall discuss some basic properties that could be useful to understand the rôle of transition matrices in the equations of the kinetic-theoretic models, we shall first recall the structure of these latter, giving a critical description of the terms involved therein, and finally, we will suggest the ways in which, in virtue of our knowledge of the complete meaning of the transition probabilities, and in view of catching new natural processes for non-isolated systems, the transition probabilities (together with the so-called encounter rates, see Sections 1.6 and 1.7) can be modified.

Advertisement

2. Some background about Markov Chains

As is well-known, a random process [2] is a sequence Xηη of random variables with a common range S of possible values (the states) called the state space of the process. Both and S can be either discrete or continuous sets. If is discrete, that is ==0,1,2n, then the process is said to be time-discrete, while, if is continuous, that is, =0+, then the process is said to be time-continuous. Analogously if S is discrete, that is, S=zhhI with I, then the process is said to be discrete, while, if S is continuous (it is any real interval), then the process is said to be continuous. From now to the end of this section, we shall first consider discrete and time-discrete processes, and at a second step time-continuous discrete processes. In both cases, for the sake of simplicity, the state space S will be assumed to be finite. In this choice, there is no loss of generality, and we only avoid to write matrices with infinitely many rows and columns.

Consider then a discrete and time-discrete random process Xhh with S=z1z2zn. . As usual, we shall denote by a vector ph (called the state vector of the process at time h) the (absolute) probability distribution on S according to the random variable Xh. In other words, phph,1ph,2ph,n with ph,k=PXh=zk.1

Now, a discrete time-discrete Markov Chain Xhh, with RXh=S, is a random process such that the Markov condition

PXh=zhX1=z1X2=z2Xh1=zh1==PXh=zhXh1=zh1,h\0E1

holds (where we have used the usual symbol PAB to denote the conditional probability of an event A under the assumption that an event B has taken place). This means that, for any h, the values of the h-th random variable of the process depend only on the values of the (h – 1)-th, not on the values of any previous variable. Now, setting Pijh=PXh=zjXh1=zi (with ij12n2), Pijh is the transition probability from state zi to state zj at the h-th step, and the matrix

PhPijh1i,jnP11hP12hP1nhP21hP22hP2nhPn1hPn2hPnnh,h\0.E2

is called the transition matrix at h-th step. It is well-known (and also obvious) that

j=1nPijh=1,h\0E3

and, in virtue of the law of alternatives,

ph=ph1Ph,h\0,E4

where ph is a row vector for any h, so that the product at right-hand side is a row-by-column product. For the sake of completeness, we also recall that, for any couple (r, s) of nonnegative integers,

pr+s=prPr+1Pr+2Pr+s,E5

and in particular, when the Markov Chain is stationary, that is, a transition matrix P exists such that Ph=P for any h\0,

pr+s=prPs,E6

where the power at right-hand side must be interpreted in the sense of the row-by-column product of matrices (for further details, see e.g. [2, 3]).

Advertisement

3. Joint and marginal transition probabilities

Now, bearing in mind these premises, we want to observe that in the literature about Markov Chains, the structure of the state space S is usually completely disregarded. With only one exception at our knowledge [16], it is simply assumed to be an ordered set, mainly to be allowed to write simple transition matrices. But we want to ask ourselves which new features should be taken into account if we considered some kinds of structure, and how should develop the study. In particular, we want to examine the case in which S=Ωm (where Ω=x1x2xl so that Ωm=lm and zixi1xi2xim for any i12n and i1i2im12lm) and the process on consideration should be described as a sequence Xh,1Xh,2Xh,mh, which will be written in vector notation as Xhh. In such a case, we should consider the transition probabilities

PXh=zhXh1=zh1==PXh,1=xh,1Xh,m=xh,mXh1,1=xh1,1Xh1,m=xh1,mE7

that have 2m indexes, that is, we can set

PXh,1=xj1Xh,m=xjmXh1,1=xi1Xh1,m=ximPi1,i2,,im,j1,j2,,jmh.E8

So, for any h, the transition matrix Ph at h-th time is a 2m-dimensional matrix. It will be called the joint transition matrix, and each of its elements Pi1,i2,,im,j1,j2,,jmh expresses the probability that the chain passes from the state xi1xim to the state xj1xjm at h-th time. These will be called the joint transition probabilities.

But, of course, together with these joint probabilities, we are now allowed to consider also the many different marginal probabilities. More precisely, for any hN, we can choose r<m of the random variables Xh,1,Xh,2,,Xh,m, say Xh,k1,Xh,k2,,Xh,kr and look for the probabilities PXh,k1=xjk1Xh,kr=xjkrXh1,1=xi1Xh1,m=xim. For any choice, we have a r+m-dimensional matrix with lr+m entries. Of course, the matrices are nr. We shall denote by Pi1,i2,,im,jk1,jk2,,jkrh the entries of each of such matrices. It should also be noted that, setting s=mr, and agreeing to denote by Xh,1,Xh,2,,Xh,s the remaining s variables, we have

Pi1,i2,,im,jk1,jk2,,jkrh=j1,,js1nPi1,i2,,im,jk1,,jkr,j1,,jshE9

Now, for the sake of simplicity, assume that k1=1,k2=2,,kr=r and 1=r+1,2=r+2,,s=m. Then, we shall also agree to set S=zij1ip1jqΩr×Ωsxiyj1ip1jqp=lrq=ls, where r+s=m, xixi1xi2xir and yjxjr+1xjr+2xjm2. Then, we can write relation (9) in the form

Pzhk,xih=j=1qPzhk,zijh=j=1qPzhk,xiyjh,E10

and introduce the joint state vector phphij, with phijphzijphxiyj, and the marginal state vector px,hpx,hxi1ip, where

px,hxi=j=1qphzij=j=1qphxiyjE11

and, in virtue of the law of alternatives

phzij=k=1pl=1qph1zklPzkl,zijh.

Hence, replacing this last relation in relation (11), and taking into account relation (10), we get

px,hxi=j=1qk=1pl=1qph1xkylPzkl,xiyjh==k=1pl=1qph1xkylj=1qPzkl,xiyjh==k=1pl=1qph1xkylPzkl,xih.

Now, for reasons that will be clear in the sequel, we write this last equation in the form

px,hxipx,h1xi=k=1pl=1qph1zklPzkl,xihpx,h1xi.E12

In addition, for any l12q,

i=1pl=1qPxiyl,xiylh=1,

so that we can rewrite equation (12) in the final form

px,hxipx,h1xi=l=1qki1pph1zklPzkl,xih+ph1zilii1pPxiyl,xih.E13

Finally, if for any h, the random variables Xh,i and Xh,j are independent for any couple (i, j), and the chain is stationary, then we find

px,hxipx,h1xi=l=1qki1ppx,h1xkpy,h1ylPzkl,xi+px,h1xiki1ppy,h1ylPzil,xk.E14

The first term at right-hand side will be called the gain term of the subset of states Si=zil1lq since it accounts for all the possible transitions from other states to a state of Si, while the last term is called the loss term of Si since it accounts for all the possible transitions from a state of Si to any state outside Si.

Advertisement

4. Time-continuous vector Markov Chains

We now want to turn our attention to the case of time-continuous Markov Chains. But before treating that case, we want first to consider the random process Xhχhh, where, for any h, χh is the classical Bernoulli variable, with range Rχh=01. The variables χh are independent and identically distributed, with Pχh=0=v and Pχh=1=u for any h. For any h, Xh, the state vector ph and the transition matrix Ph are assumed to depend only on χh. So, using again the notation introduced above, we see at once that phzij=PXh=zijχh=1+PXh=zijχh=0, that is,

phzij=PXh=zijχh=1u+PXh=zijχh=0v.E15

Analogously, for any h, the transition matrix at h-th step (or time) should be written according to the condition

Pzij,zklh=Pzij,zklχh=1hu+Pzij,zklχh=0hv,

where Pzij,zklχh=1h is the transition probability from the state zij to the state zkl at h-th step under the assumption χh=1, and Pzij,zklχh=0h is the transition probability from the state zij to the state zkl at h-th step under the assumption χh=0. In particular, we shall set

Pzij,zklχh=0h=δkiδlj,h:E16

this means that the chain remains in the state occupied at time h and also at time h + 1 when χh=0. Also, notice that the Markov Chain is stationary if and only if the matrix Pzij,zklχh=1 is constant with respect to h.

By replacing the well-known relations

PXh+1=zijχh=1=k=1pl=1qphzklPzkl,zijχh=1h,PXh+1=zijχh=0=k=1pl=1qphzklPzkl,zijχh=0h,

in equation (15), and taking into account assumption (16) and the obvious relation v = 1 − u, we get

ph+1zijphzij=k=1pl=1qphzklPzkl,zijχh=1huphziju.E17

These are of course pq equations, one for any couple xiyj. By taking the sum over j of both sides, we obtain the following system of p equations:

px,h+1xipx,hxi==k=1pl=1qphzklPzkl,xiχh=1hupx,hxiu,i=12p.E18

Finally, it is readily seen that, under the independence assumption on the random variables Xh,i and Xh,j for any h, we obtain the system

px,h+1xipx,hxi==ul=1qki1ppx,hxkpy,hylPzkl,xiχ=1+px,hxiki1ppy,hylPzil,xkχ=1,i=12p.E19

if the transition matrix Pχ=1Pzij,zklχ=1 is stationary.

We are now ready to consider the case of a time-continuous Markov Chain analogous to the one; we have just described, namely a random process of the type Xtχtt0T, where T>0 may be either finite or infinite. In this case, the state vectors obviously depend on time, but the transition probabilities may depend on time or not, as well as the probabilities associated to the Bernoulli variables χt. But, similarly to what has been done above, these variables will be assumed to be identically distributed. In this connection, we have to recall to the reader that it is impossible to assign a positive probability u such that Pχt=1=u for any t because in such case the probability that at least one of the variables χt could take the value 1 for t in any interval contained in [0, T) would be infinite. What we can do is to assign a constant probability density τ such that τΔt is the probability to find in any interval of length Δt some points such that χt=1. Roughly speaking and referring to the interpretation of probability as relative frequency, we can state that, for any s0T, τΔt is the length of the set tss+Δtχt=1 so that τ is the ratio of that length to the length Δt. Accordingly, τ is the measure of the set of points for which χt=1 in any interval of unit length.

This stated, we shall write system (19) as follows:

px,t+Δtxipx,txi==τΔtl=1qki1ppx,txkpy,tylPzkl,xiχ=1+px,txiki1ppy,tylPzil,xkχ=1.i=12p,E20

where τΔt replaces the probability u in this continuous framework.

Next, dividing both sides by Δt and letting Δt0, we get

px,ttxi=τl=1qki1ppx,txkpy,tylPzkl,xiχ=1+px,txiki1ppy,tylPzil,xkχ=1.i=12p.E21

We will show that the above «mixed» process we have just discussed above, obtained by coupling a Markov Chain with a uniform process, can be further generalized, by taking into account vector uniform processes of an appropriate dimension. But we will do this in a different context, namely in the context of the models of collective phenomena.

Advertisement

5. Some background about the stochastic description of many-particle systems

For reasons that will be clear in a very short, while we shall now focus our attention on the double indexing of the elements on S, and assume r = s is so that m = 2s, S=Ω2s, p=q=ls and n=l2s and consider the random process Xtχtt0T, where 0<T+ and χtχtij1i,jls for any t0T. According to our notation conventions, χt is in fact for any t a vector with n components that are double-indexed just to reproduce the (conventional) double indexing of the elements of S. We can treat it as a square matrix with χt rows and ls columns. For each entry χtij of this matrix, we shall set

χtijχtklχteitherkiorlj.

For any t and for any couple (i, j), χtij is a Bernoulli random variable, and we assume that

  1. For any t and t in [0, T), χtij and χtij are independent and identically distributed. Accordingly, Pχtij=1=uij and Pχtij=0=1uij for any t0T.

  2. For any t0T and for any ij12ls2, the probability of state zij at time t depends only on χtij.

  3. For any t0T and for any ij12ls2, the transition probabilities from the state zij to any other state depend only on χtij; accordingly, we are allowed to consider only the conditional transition probabilities Pzij,zklχtij=0t and Pzij,zklχtij=1t.

  4. As for the process considered in the previous section,

Pzij,zklχtij=0t=δkiδlj,t0T,

so that we are allowed again to state that the chain is stationary if and only if Pzij,zklχtij=1t is in fact independent of t.

This stated that we are allowed to reproduce the reasonings of the previous sections, with obvious changes leading us to replace the system (21) with the system

px,ttxi=l=1qki1pτklpx,txkpy,tylPzkl,xiχtkl=1+px,txiki1pτilpy,tylPzil,xkχtil=1,i=12ls.E22

Now, once we have reconstructed the way in which this last system of equations is originated by coupling a stationary vector Markov Chain with an additional vector random process actually modifying the final form of the transition matrix maintaining; however, its stationarity (that is its independence on time), we shall devote the rest of this section to recall a classical interpretation (see [8]) of system (22), which has become in the last forty years ever more important as it can be applied to a wide range of different phenomena, from mechanics of gases (which is its true historical origin, since Boltzmann proposed his celebrated kinetic theory [4]) to the behavior of biological systems, with particular concern to the interaction between healthy and tumor tissues (see [6], also for more complete bibliographic references), from social sciences, with particular concern to the diffusion of competing opinions [14], to economics [9], and to the behavior of swarms and crowds [5, 7]. This interpretation has in fact given rise to a general model for the description of the behavior of a large number of many-particle systems, we would dare say of almost all possible types and in almost all possible contexts.

The key idea upon which the kinetic-theoretic description of the behavior of many-particle systems is based and is the following. A set S of a very large number N of objects is given. The objects of S, usually called «particles» or «individuals» according to the context, can interact pairwise3 with each other and are identified (further than by some common «physical» properties they can possess to different degrees, which do not appear explicitly in the mathematical description of the behavior of S, but influence the values of parameters) by their «states». According to the context, the set of possible states of all particles can be finite, countably infinite, or continuous. We cannot (and do not) claim to know exactly the state of each particle of the system at any time so that even the prescription of N precise initial conditions on the states of particles would be quite unrealistic. What we assume to be allowed to state is that, if the number of all possible different states is q, then all the particles of S can be thought of as divided into n different classes S1,S2,,Sq, each containing all and only the particles sharing the same state; if nit (with i12n is the number of particles sharing the i-th state at time t, then pit=nit/N is the probability to pick at random in S a particle in the i-th state, and we can construct a state vector pp1p2pq. In addition, each particle belonging to a class Sk can jump into another class Si if and only if it «interacts» with some other particle: the state (that is, the class Sl) of this latter does not matter, if not as influencing the assessment of the different probabilities of jump Pki=PSkSi, that accordingly are denoted Pki;l=PSkSiSl. Moreover, if a particle has no interactions with other particles, then it will not change its state.

The above almost purely qualitative descriptions lead us at once to conclude that, in the above scheme, any many-particle system is in fact a random process Xtχtt0T of the above-described type, where Xt is a Markov Chain, characterized by a state space, which in our previous description is Ωsx1x2xq (where, we recall, q=ls), a state (probability) vector px expressing in percentage the distribution of particles over all possible different states, an interaction rate τij (the average number of interactions between a particle of Si and a particle of Sj occurring in any unit time interval, that is, the measure of the set I1tIχtij=1 in any inerval I of unit length), and each probability of jump Pki;l=PSkSiSl, conditional to the occurrence of an interaction between a particle in the state xk and a particle in the state yl is given by Pxkyl,xiχtkl=1=Pzkl,xiχtkl=1. So, relations (22) turn out to be exactly the equations governing, in stochastic terms, the evolution of isolated many-particle systems.

Advertisement

6. Equations for many-particle systems with time-dependent transition matrices

At a first glance, the above reconstruction of the equations of kinetic theory in terms of a random process Xtχtt0T coupling a stationary Markov Chain and a Bernoullian process could seem a purely academic exercise, though interesting as showing the correspondence between two different languages, born in different contexts, with different aims and in different times4. But our exercise is not an end in itself but is very useful to extend our model for many-particle systems to cover the case of systems interacting with the external world. It is in fact not by chance that we have stressed that equations (22) hold for isolated many-particle systems. These equations are based on the assumption that the instantaneous variation of the distribution of particles over the state space is uniquely due to mutual interactions between particles and that these interactions modify the states of the involved particles always in the same way, independently of the time at which they occur5. But when the system perceives the influence of the external world, we are allowed to assume that both the interaction rate and the results of any interaction are more or less strongly modified from time to time by events that happen outside the system.

More precisely, our stochastic process Xtχtt0T must be replaced by another process XtχtEtt0T, where, for any t0T, Et is another «strange» vector random variable6 with values in Wd×Kh (where K is a finite subset of ), such that, for any t0T, there exists an et=utvt(u1t,u2t,,udt,v1t, v2t,,vht)Wutu1tu2tudtvtv1tv2tvht for which the probability density function of Et is expressed, for any εξηξ1ξdη1ηh, with ξξ1ξd and, ηη1ηh, by the relation

ρEtε=δξ1u1tδξ2u2tδξdudt××δη1v1tδη2u2tδηhvht=,t0T=δξutδηvt.E23

This means that the process Ett0T is a deterministic external process. For instance, we can imagine that our system consists of all the inhabitants of a given region, whose states are vectors, one component of which is the amount of money of each individual, and the other components are quantities of food. We consider only interactions consisting in purchase and sale of apricots. In the late summer, in the autumn, and in the winter the interaction rate will be zero (no apricots in these seasons); in the early springtime, the interaction rate will be small, and the transition probabilities will describe changes of states with little increases or decreases of the quantity of apricots, and large increases or decreases of the amount of money (in that period, apricots are hardly available and very expensive); in the late springtime and in the early summer, the interaction rate will attain its maximum value and the transition probabilities will describe changes of states with large increases or decreases of the quantity of apricots, and small increases or decreases of the amount of money (in that period, many apricots available at low prices). So, both the interaction rates and the transition probabilities depend on seasons, that is, are functions of time, and system (22) must be simply written in the form

px,ttxi=l=1qki1pτkltpx,txkpy,tylPzkl,xiχtkl=1t+px,txiki1pτiltpy,tylPzil,xkχtil=1t,i=12ls.E24

Accordingly, the introduction of a dependence on time of both interaction rates and transition probabilities seems to be the most appropriate way to link the behavior of a many-particle system to deterministic external influences. Of course, the interesting aspect, and also the most difficult one, of this kind of model, is the choice of functions τklt and Pzkl,xiχtkl=1t in such a way as to express faithfully the changes of the environment, as well as the way in which these changes are perceived by the individuals of the system. At an initial stage, we can adopt a purely hypothetical method and make any assumption we want, but when we have in mind some precise applications, then we need a preliminary statistical analysis to get the right hints toward the most plausible assignment of the above functions.

Advertisement

7. Equations for many-particle systems with stochastic transition matrices

The case treated in the previous section is rather trivial, and its main interest lies in the choice of the dependence of all the interaction rates τklt and of all transition probabilities Pzkl,xiχtkl=1t on time, which is connected with the application we have in mind, that is, much more with experimental and statistical analysis than with theoretical reasonings. Much more interesting is the case in which the changes of external world influencing the behavior of the system are not assigned at each time, but stochastic events. This means that we have to consider again a stochastic process XtχtEtt0T, but now, for any t0T, Et is a random variable in the “strict” sense of the word. More precisely, for any t0T,

  1. We denote again by Wd×h the range of all possible values of Et.

  2. We denote by εξηξ1ξdη1ηh any unspecified element of W.

  3. We set aa1a2ad, bb1b2bd, and Iaba1b1×a2b2××adbd;

  4. We set EtEt,1Et,2 where the range of values of E1 is d and the range of values of E2 is h, and assume that Et,1 and Et,2 are independent;

  5. To the (continuous) random variable E1 is associated a probability density function ρt,1ξρ1ξt such that

    PEt,1Iab=Iabρ1ξtdξ<1

    for any abd×d such that d\Iab;

  6. To the (discrete) random variable E2 is associated a probability distribution function Pt,2 such that

    PEt,2=η=Pt,2ηP2ηt.

    For any ηh, P2ηt<1.

  7. we denote by the symbol τijεt the density (on the interval [0, T)) of the probability that χtij=1 conditional to the event Et=ε.

Accordingly, for any εW, the associated probability density of Et at ε is

ρEtε=ρ1ξtP2ηt.

This stated, according to the law of alternatives, and under the conditions imposed in Section 1.5, we find that each transition probability depends on time, and we have

Pzil,xkχil,t=1,t=ηP2ηtdPzilxkχil,t=1E1=ξE2=ηρ1ξtτijεtdξ.E25

Accordingly, the system of equations governing the evolution of the many-particle system we have decided to describe by means of the present language and of the notions of random processes and, in particular, of Markov Chains, takes the form

dpx,ttxi=l=1qki1pηP2ηt×px,txkpy,tyldPzkl,xiχtkl=1,Et=ερ1(ξt)τkl(εt)dξ+px,txipy,tyldPzil,xkχtil=1,Et=ερ1(ξt)τil(εt)dξ,i=12lsE26

where Pzkl,xiχtkl=1,Et=εPzklxiχtkl=1E1=ξE2=η and Pzil,xkχtil=1,Et=εPzilxkχtil=1E1=ξE2=η, when Ω (the state space common to all components of the vector Markov Chain X) is a discrete set (as we have assumed throughout the whole chapter till now). When Ω is instead a continuous set (typically, a real interval), then the sums in relations (26) must be replaced by integrals and — with the obvious correspondences xix, xkx, yly, zklxy, zilxy, τklεtτxyεt, τilεtτxyεt, χtklχxyt, and χtilχxyt — we get the equation7

pxtxt=ηP2ηt×ΩΩpxxtpy(yt)dPxy,xχxyt=1,Et=ερ1(ξt)τ(xyεt)dξ+pxxtpy(yt)dPxy,xχxyt=1,Et=ερ1(ξt)τ(xy)εt)dξdxyE27

where we have also agreed to write

px,tx=pxxt,px,tx=pxxt,py,ty=pyyt.

Equation (27) is a generalization of system (24). As a matter of fact, this latter can be easily obtained from the former by simply replacing the probability density function ρ1 with the one defined by relation (23). And, of course, the equations governing the evolutions of isolated many-particle systems can be derived at once by choosing Et as a deterministic and constant process (i.e. a process in which all the random variables Et have the same probability density function

ρξ=δξe

on W, however, e is fixed). So, system (27) seems to be the most general description of the behavior of any one of such systems. Recently, however, other types of external influence have been considered, acting directly on the state vectors and not on the interactions between particles [13, 14]. This point will be discussed in some details in the last section of the chapter.

Advertisement

8. Subsystems

This section is almost a pure aside, just to give a complete exposition of the development of the whole theory of many-particle systems as based on the theory of stochastic processes. We want to recall that a more general formulation of the theory takes into consideration the case in which a system S is the join of a number N of subsystems S1,S2,,SN, called the functional subsystems of S, that can share the same state space or not, and are such that, for any λ12N, the particles of Sλ are not constrained to share the same state. In order to show the way in which the equations describing the behavior of S must be modified under this assumption, in order to avoid an undue multiplication of indexes, and for the sake of a better readability, we shall assume that all the subsystems share a continuous state space Ωab. The point of the present development is that now we must allow interactions between particles of different subsystems so that we must consider a mixed stochastic process XtχtEtt0T where χtχλ,μxyt, and for any t0T and for any 4-tuple λμxy12N2×Ω2s the random variable χλ,μxy is again a Bernoulli variable, with probability density function τλμuvΔt, in view of the continuity of variables x and y. By reproducing almost without changes the reasonings of the previous two sections, we obtain the system

pxλtxt=ηP2ηt×μ=1νΩΩ{pxλxtpyμ(yt)dPxy,xχλμxyt=1,Et=ερ1(ξt)τλμ(xyεt)dξ+pxλxtpyμ(yt)dPxy,xχλ,μxyt=1,Et=ερ1(ξt)τλμ(xy)εt)dξ}dxyλ=12N.E28

We have now again a system of equations. Each of them describes the evolution of the probability density function of a corresponding subsystem in consequence of the interaction of both interactions between particles of the same subsystem and of particles of different subsystems. But, of course, we have written the equations in the final form they have when also external actions that modify both the rates and the effects of interactions.

Advertisement

9. Alternative views, conclusions and perspectives

In almost all the literature about many-particle systems, external actions are completely disregarded, and the system of equations governing their evolution is written in its simplest form, where both the transition probabilities and the interaction rates are constant. Only recently, a number of papers have taken into account external «forcing» events that are capable to modify the distribution of states on each subsystem of a system S independently of possible interactions between particles [11, 12, 13, 14, 15]. These external events, however, are assumed to act directly on the functions pxλxt (or pxλxt, when the state space is discrete). The origin of this viewpoint is in [11] where, dealing with a description of possible evolutions of an epidemics, an additional term describing spontaneous healing or death was necessary to get a faithful picture. This also led to general considerations about the implicit use of an analog of the Principle of Inertia in the scheme depicted in Sections 1.5 and 1.6. Taking into account an additional term of this type, system (28) becomes

pxλtxt=ηP2ηt×μ=1νΩΩ{pxλxtpyμ(yt)dPxy,xχλμxyt=1,Et=ερ1(ξt)τλμ(xyεt)dξ+pxλxtpyμ(yt)dPxy,xχλ,μxyt=1,Et=ερ1(ξt)τλμ(xy)εt)dξ}dxy++Fλpx1px2pxNtλ=12N.E29

where the form of the term Fλpx1px2pxNt depends on the context and on the type of external solicitation producing a «spontaneous» variation of the distributions.

But now, to conclude this overview of the possible developments of the modern generalization of kinetic theory, that becomes evident when it is analyzed and formalized in terms in terms of n-tuples of joint random processes (one of which is a Markov Chain, but to this point we will return at the very end of this final section), going back to the discrete version of the equation, from which we started to reconstruct the logic the equations of the theory are based upon, we want to point out an interesting new application of the theory in the scope of experimental procedures. This application was presented in [17] and its theoretical formulation in [18].

In order to outline the above-mentioned development, we first assume to deal with a system S split into two subsystems S1=Σ and S2=Xhh. So, S1 contains only one element, which will be called the «subject», while the elements of S2, that are random variables8will be called the experiments or questions about Σ. The «state space of the subject Σ» is a set Ω1x1x2xp while, for any h, the «state space of question Xh», that is, the set of all of its possible outcomes, is the set Ω2hyh1yh2yhqh. The set Ω2=hΩ2h will be the state space common to all the random variables of the sequence Xhh. We consider the discrete (stepwise) stochastic process Xhχhh, where Xh, such as in Section 1.3, is a Markov Chain with the state space Ω1×Ω2 and

χhχh11xixjχh12xiyhjχh21xiyhjχh22yhiyhj

for any h. The entries of this matrix are independent Bernoulli’s random variables. χh11xixj=1 means that an interaction between Σ in the state Σ and Σ in the state xj occurs at h-th step, and χh22yhiyhj=1 means that at the same step an interaction occurs between two different answers to the same question Xh. We have not yet defined the meaning of the word «interaction» in the present context (this point will be briefly discussed at the end of the section), so for the moment, we simply assume that χh11xixj=χh22yhlyhm=0 for any h and for any ij112p2 and lm12qh2. If τi,j11=Pχh11xixj=1 and τl,m22=Pχh22yhlyhm=1), this also means that τi,j11=τl,m22=0. On the other hand, χh12xiyhj=1 and χh21xiyhj=1 mean that an interaction between Σ in the state xi and the question Xh giving the answer yhj certainly occurs at h-th step; this corresponds to the fact that the interaction is decided and produced by an experimenter. We assume χh12xiyhj=χh21xiyhj=1 for any h and for any i112p so that τi,j12=τi,j21=1. Moreover, we assume Pxkyl,xiχh21xiyhj=121=0 for any h and for any ij112p2. All the other conditions assuring that we are allowed to write equations (19) are assumed to hold, These equations, with obvious meaning of the symbols, will be finally rewritten in the form

px,h+11xipx,h1xi==l=1qhki1ppx,h1xkpy,h2yhlPxkyhl,xiχ12xiyhi=112+px,h1xiki1ppy,h2yhlPxiyhl,xkχ12xiyhi=1=112,py,h+12yh+1ipy,h2yhi=0i=12p.E30

To conclude the section and the chapter, we want now to expose an interpretation as complete as possible of the special model just described. To this aim, what is necessary to stress first is that system (30) is meant to describe the results of interactions between information rather than between objects. More precisely, the values xi of the variable in Ω1 should be interpreted as labels that define, in the appropriate context, possible states of the subject; then, the interaction between the subject and the h-th experience Xh on it places this latter into a state, expressed by a real value yhi, that is, the outcome of a measurement of a physical parameter. Now, in virtue of the information carried by the value yhi, the i-th experience interacts with the initial definition, and in the sense that it can either confirm the initial state xi (by increasing its probability) or raise doubts about it (by lowering its probability). Accordingly, the terms px,h1xi and px,h1xi in equations (30) must be interpreted respectively as the absolute probability that the definition of the state of the subject before the experience is the value xi and the absolute probability that the outcome of the h-th experience is the state yhi; furthermore, the term Pxkyhl,xiχ12xiyhi=112 is the probability that definition xk, after the «interaction» with information yhl becomes definition xi.

In [17], equations (30) were derived by discretization from the ones of kinetic theory and used to describe the way in which new experiences on a number of different parameters can modify the classification of a phenomenon, in particular the diagnosis of depression in a human subject suspected to be depressed according to the score obtained by answering a questionnaire. Of course, in an application like that, the transition probabilities must be estimated in advance by a preliminary study of the correlation of the involved random variable with the possible depression states. We have reported this application, together with the classical application of kinetic theory and its developments, to give a more complete picture of the way in which the study of joint random processes (vector random processes), in particular when one of the processes considered simultaneously is a Markov Chain, can open the way to an ever wider range of applications.

References

  1. 1. Gilch L. Markov Chains: An Introduction: Lecture Notes. Independently Published. 2022
  2. 2. Lawler GF. Introduction to Stochastic Processes. Boca Raton: Chapman and Hall/CRC; 2006
  3. 3. Norris JR. Markov Chains. Cambridge: Cambridge University Press; 1998
  4. 4. Boltzmann L. Lectures on Gas Theory. Courier Corporation; 2012
  5. 5. Aylaj B, Bellomo N, Gibelli L. Crowd Dynamics by Kinetic Theory Modeling: Complexity, Modeling, Simulations, and Safety. New York: Springer; 2020
  6. 6. Bellomo N, Bellouquid A, Gibelli L, Outada N. A Quest towards a Mathematical Theory of Living Systems. Basel: Birkhäuser; 2017
  7. 7. Bellomo N, Brezzi F. Traffic, crowds and swarms. Mathematical Models and Methods in Applied Sciences. 2008;18(Suppl):1145-1148
  8. 8. Bellomo N, Degond P, Tadmor E, editors. Active Particles, Volume 1: Advances in Theory, Models, and Applications. Basel: Birkhäuser; 2017
  9. 9. Bertotti ML, Modanese G. Economic inequality and mobility in kinetic models for social sciences. The European Physical Journal Special Topics. 2019;225(10):1945-1958
  10. 10. Chinesta F, Abisset-Chavanne E. A Journey Around the Different Scales Involved in the Description of Matter and Complex Systems: A Brief Overview with Special Emphasis on Kinetic Theory Approaches. New York: Springer; 2017
  11. 11. Carbonaro B. Modeling epidemics by means of the stochastic description of complex systems. Computational and Mathematical Methods. 2021;3(6):1208-1220
  12. 12. Carbonaro B. The role of the principle of inertia in KTAP models. Journal of Mathematical Physics. 2022;63(1)
  13. 13. Carbonaro B, Menale M. A nonconservative kinetic framework under the action of an external force field: Theoretical results with application inspired to ecology. European Journal of Applied Mathematics. 2023;34(6):1170-1186
  14. 14. Carbonaro B, Menale M. The behavior of a many-particle system with fading memory interacting with external world, to appear
  15. 15. Menale M, Munafò CF. A kinetic framework under the action of an external force field: Analysis and application in epidemiology. Chaos, Solitons & Fractals. 2023;174:113801
  16. 16. Benaim M, Hurth T. Markov Chains on Metric Spaces: A Short Course. New York: Springer Nature, Springer; 2022
  17. 17. Vitale F, Carbonaro B, Esposito A. A dynamic probabilistic model for heterogeneous data fusion: A Pilot Case study from computer-aided detection of depression. Brain Science. 2023;13:1339-1359
  18. 18. Vitale F, Carbonaro B, Esposito A. A methodological application of a time-discrete version of the kinetic-theoretic language to interactions between random variables. To appear

Notes

  • Here and in the sequel, any probability measure, when explicitly applied to events, will be denoted by a bold-face capital letter.
  • From now on, the symbols zij and xiyj will be treated as interchangeable. Moreover, the indexes of components of any vector will be written as apices for the sake of readability.
  • But the theory is currently developing to cover the case of multiple interactions.
  • The equation of Kinetic Theory was proposed by Botzmann in 1872 when Markov was sixteen years old.
  • This assumption can be seen as equivalent to assuming a form of the Principle of Inertia for many-particle systems [12].
  • This simply means that the external events may be always described by a set of measures or numerical labels. And, provided we adopt a widely shared choice of a numerical labelling method, this is, always possible.
  • Notice that system (24) is replaced by only one equation because the ls unknown functions px,txi of one variable (the time variable) corresponding to the ls possible states are replaced, in view of the continuity of the state variable, by a unique unknown function of both the variables x and t.
  • The reader should carefully note that, differently from any classical application of the description of many-particle systems outlined in the previous Section of this chapter, the members of system S can be very different from «particles» in the classical sense of the word, that is. abstract objects, such as random variables.

Written By

Bruno Carbonaro and Federica Vitale

Submitted: 20 February 2024 Reviewed: 22 February 2024 Published: 31 July 2024