Home Some Characterizations of Mixed Renewal Processes
Article Open Access

Some Characterizations of Mixed Renewal Processes

  • Demetrios P. Lyberopoulos EMAIL logo and Nikolaos D. Macheras EMAIL logo
Published/Copyright: February 16, 2022
Become an author with De Gruyter Brill

Abstract

Some characterizations of mixed renewal processes in terms of exchangeability and of different types of regular conditional probabilities are given. As a consequence, an existence result for mixed renewal processes, providing also a new construction for them, is obtained. As an application, some concrete examples of constructing such processes are presented and the corresponding regular conditional probabilities are explicitly computed.

2020 Mathematics Subject Classification: Primary 60G55; Secondary 60G09; 28A50; 60A10; 60G05; 60K05; 91B30

1. Introduction

Although mixed renewal processes (MRPs for short) have not been studied in the literature as extensively as their mixed Poisson counterparts, they certainly possess their own merit, which has become more important during the last decades; we refer, for instance, to the pioneering paper of Huang [11] as well as to the earlier work of Segerdahl [22] with a more applied flavor. The latter two contributions are indicative for the two main reasons that explain the importance of MRPs: they may serve, on the one hand, as a useful tool for modeling real life situations, such as those emerging in actuarial practice (cf. e.g., [22: p. 164–165], where experience rating schemes on motor vehicle insurance are discussed) and on the other as a source of challenging theoretical problems, since they are generalizations of mixed Poisson processes (MPPs for short) and closely connected with the concept of exchangeable stochastic processes (cf. e.g., [11]).

Likewise, the motivation of the current work is two-fold: the main one refers to the introduction and investigation of a definition of MRPs which is both less technical than the definitions of MRPs usually met in the literature (see e.g., Grandell [8] and Huang [11]) and in line with their counterparts referring to mixed Poisson processes (MPPs for short), especially with that of MPPs with parameter Θ (cf. e.g., [21: Chapter 4] for the definition). Such a definition seems to be a proper/more natural one as it involves explicitly the structural parameter Θ, which is usually essential in the study of risk-theoretical problems.

In fact, this definition was exploited in [16] and [17], where the equivalence of various existing definitions of MPPs is proved and the problem of examining whether within the larger class of extended MRPs (see [17: Definition 3.2] for the definition), every Markov process is a mixed Poisson one with a random variable as mixing parameter is addressed, respectively. Moreover, the main outcomes of this work may find applications to problems related to life times of components (see also [11: p. 17–18]) and especially to the interplay between financial and actuarial pricing of insurance, see, e.g., [24] and compare it with the recent work of Lyberopoulos & Macheras [15], where the problem of arbitrage-free pricing of an inhomogeneous portfolio of insurance risks was made possible, generalizing the financial pricing of insurance approach, introduced by Delbean & Haezendonck [2] in the framework of classical risk theory.

Since conditioning is involved in the definition of MRPs introduced in this paper, it is reasonable to ask about the structural role of regular conditional probabilities (rcps for short) in this field. For this reason, in Section 3, we recall the definitions of different types of rcps and provide a characterization of MRPs via rcps (see Proposition 3.1), which extends a basic result, i.e., Proposition 4.4 of our previous work [13]. By means of this result, we can reduce MRPs under a probability measure P with parameter Θ to ordinary renewal processes under the corresponding regular conditional probabilities of P over PΘ consistent with Θ, showing in this way that Definition 3.1 is the natural one for MRPs.

The second definition of MRPs investigated in this paper is due to Huang [11], see Definition 4.1. In Section 4, we first obtain some characterizations of exchangeability in terms of different types of rcps, see Theorem 4.2 and Corollary 4.2.1. As a consequence, in Theorem 4.3 some further characterizations of MRPs in terms of exchangeability and rcps are deduced. Theorem 4.3 provides amongst others a detailed discussion of the relation between the two definitions of MRPs and shows that in most cases appearing in applications both definitions coincide.

In Section 5, we first give some examples to show that some of the assumptions of Theorems 4.2 and 4.3 are essential for the validity of all the equivalences obtained therein (see Examples 1 to 2). Next, we provide a construction of canonical probability spaces admitting MRPs (see Example 3), extending a similar construction for MPPs, see [14: Theorem 3.1]. As an application, we give concrete examples of MRPs satisfying all assumptions of Theorem 4.3, and we compute the corresponding rcps explicitly.

2. Preliminaries

By ℕ is denoted the set of all natural numbers and ℕ0 := ℕ ⋃ {0}. The symbol ℝ stands for the set of all real numbers, while ℝ+ := {x ∈ ℝ : x ≥ 0}. If d ∈ ℕ, then ℝd denotes the Euclidean space of dimension d.

Given a probability space (Ω, Σ, P), a set NΣ with P(N) = 0 is called a P-null set (or a null set for simplicity). The family of all P-null sets is denoted by Σ0. For any two Σ-T-measurable maps X,Y: Ωϒ, we write X = Y P-a.s., if {XY} ∈ Σ0. For the definitions of real-valued random variables, random variables and random vectors, we refer to Cohn [1: pp. 308 and 318].

If AΩ, then Ac := ΩA, while χA denotes the indicator (or characteristic) function of the set A. For a map f : Ωϒ and for a non-empty set AΩ denote by fA the restriction of f to A. The identity map from Ω onto itself is denoted by idΩ. The σ-algebra generated by a family G of subsets of Ω is denoted by σG.

For any Hausdorff topology T on Ω by B(Ω) is denoted the Borel σ-algebra on Ω, i.e., the σ-algebra generated by T. By B:=B(), Bd:=Bd and B:=B is denoted the Borel σ -algebra of subsets of ℝ, ℝd and ℝ, respectively, while L1(P) stands for the family of all real-valued P-integrable functions on Ω. Functions that are P-a.s. equal are not identified. We write EP[X|F] for a version of a conditional expectation (under P) of XL1(P) given a σ-subalgebra F of Σ, while for X:=χEL1(P) with EΣ, we set P(E|F):=EPχE|F.

Given two probability spaces (Ω, Σ, P) and (ϒ, T, Q), we denote by σ({Xi}iI) the σ-algebra generated by a family {Xi}iI of Σ-T-measurable maps from Ω into ϒ, while if X is a Σ-T-measurable map then PX stands for the image measure of P under X. If X is a d-dimensional random vector on Ω, we write PX = K(θ) meaning that X is distributed according to the law K(θ), where θ ∈ ℝd. In particular, P(θ), Exp(θ) and Ga(γ, α), where θ, γ, α are positive parameters, stand for the law of Poisson, exponential and gamma distribution, respectively (cf, e.g., [21]).

By (Ω × ϒ, ΣT, PQ) is denoted the product probability space of (Ω, Σ, P) and (ϒ, T, Q), and by πΩ and πϒ the canonical projections from Ω × ϒ onto Ω and ϒ, respectively.

Given two measurable spaces (Ω, Σ) and (ϒ, T), a Σ-T-Markov kernel is a function k from T × Ω into ℝ satisfying the following conditions:

  1. The set-function k(·, ω) is a probability measure on T for any fixed ωΩ.

  2. The function ωk(B, ω) is Σ-measurable for any fixed BT.

In particular, given a probability measure Q on T, a family {Py}yϒ of probability measures on Σ is called a regular conditional probability (rcp for short) of P over Q if

  1. for each DΣ the map yPy(D) is T-measurable;

  2. Py(D)Q(dy) = P(D) for each DΣ.

If f: Ωϒ is an inverse-measure-preserving function (i.e., P(f−1(B)) = Q(B) for each BT), an rcp {Py}yϒ of P over Q is called consistent with f if for each BT, the equality Py(f−1(B)) = 1 holds for Q-almost every yB.

We could use the term of disintegration instead, but it seems that it is better to reserve that term to the general case when Py’s may be defined on different domains (see [19]).

Recall that if Σ is countably generated and P is perfect (see e.g., [1: p. 102] and [5: p. 291], respectively, for the definitions), then there always exists an rcp {Py}yϒ of P over Q consistent with any inverse-measure-preserving map f from Ω into ϒ providing that T is countably generated (see [5: Theorems 6 and 3]) and a subfield rcp (see [5: Theorems 6 and 2] as well as [5: Section 2] for the definition). So, in most cases appearing in applications these two types of rcps always exist. A Polish space, that is a topological space homeomorphic to a complete separable metric space, is such a case.

Let now (Ψ, Z) be a measurable space and let X : Ωϒ and Θ : ΩΨ be a Σ-T-measurable and a Σ-Z-measurable map, respectively. A conditional distribution of X over Θ, denoted by PX|Θ, is a σ(Θ)-T-Markov kernel k satisfying for each BT condition

k(B,)=PX1(B)|Θ()Pσ(Θ)-a.s..

Note that if ϒ is a Polish space then a conditional distribution of X over Θ always exists (cf. e.g., [4: Theorem 10.2.2]).

Clearly, for every Z-T-Markov kernel k, the map K(Θ) from T × Ω into ℝ defined by means of

K(Θ)(B,ω):=(k(B,)Θ)(ω) for any BT and ωΩ

is a σ(Θ)-T-Markov kernel. In particular, for ϒ,T=,B its associated probability measures k(·, θ) for θ = Θ(ω) with ωΩ are distributions on B and so, we may write K(θ)(·) instead of k(·, θ).Consequently, in this case K(Θ) will be denoted by K(Θ).

For any σ-subalgebra F of Σ, we say that two F-T-Markov kernels ki, for i ∈ {1, 2}, are P|F-equivalent and, we write k1 = k2P|F-a.s., if there exists a P-null set NF such that for any ωN and BT the equality k1(B, ω) = k2(B, ω) holds true.

From now on (Ω, Σ, P) is a probability space, while (ϒ, T) and (Ψ, Z) are measurable spaces, all of them arbitrary but fixed.

3. Characterizations of mixed renewal processes via regular conditional probabilities

Let {Nt}t∈ℝ+ be a (P-)counting process with exceptional (P-) null set ΩN (cf. e.g., [21: p. 17] for the definition). Without loss of generality, we may and do assume that ΩN=0. Denote by {Tn}n∈ℕ0 and {Wn}n∈ℕ the arrival process and interarrival process, respectively, associated with {Nt}t∈ℝ+ (cf. e.g., [21: p. 6] for the definitions).

Recall now that for a Σ-Z-measurable map Θ from Ω onto Ψ, a family {Xi}iI of Σ-T-measurable maps from Ω into ϒ is P-conditionally (stochastically) independent given Θ if

Pj=1nXij1Bj|Θ=j=1nPXij1Bj|ΘPσ(Θ)-a.s

whenever i1, …, in are distinct members of I and BjT for every jn. The family {Xi}iI is P-conditionally identically distributed given Θ, if

PFXi1(B)=P(FXj1(B))

whenever i, jI, Fσ(Θ) and BT.

Throughout what follows, unless it is stated otherwise, Θ is a Σ-Z-measurable map from Ω into Ψ, and we simply write “conditionally” in the place of “conditionally given Θ” whenever conditioning refers to Θ. Moreover, {Nt}t∈ℝ+is a counting process and without loss of generality, we may and do assume thatΩN=0.

The counting process {Nt}t∈ℝ+ is said to be a P-renewal process with interarrival time distribution K(θ0), where θ0 ∈ ℝd is a parameter (or a (P, K (θ0))-RP for short), if its associated interarrival times Wn, n ∈ ℕ, are independent and K(θ0)-distributed under the probability measure P.

Definition 3.1.

The counting process {Nt}t∈ℝ+ is said to be a mixed renewal process on (Ω, Σ, P) with parameter the d-dimensional (d ∈ ℕ) random vector Θ and interarrival time conditional distribution K(Θ) (or a (P, K(Θ)-MRP for short), if {Wn}n∈ℕ is P-conditionally independent and

PWn|Θ=K(Θ)Pσ(Θ)-a.s.

for all n ∈ ℕ.

In particular, for Ψ,Z=,B and PΘ((0, ∞)) = 1 a counting process {Nt}t∈ℝ+ is a P-mixed Poisson process on (Ω, Σ, P) (or a P-MPP for short) with parameter the random variable Θ, if it has P-conditionally stationary independent increments (cf. e.g., [21: Section 4.1, p. 86] for the definition), such that

PNtΘ=P(tΘ)Pσ(Θ)-a.s.

holds true for each t ∈ (0, ∞).

Note that, for d = 1, PΘ((0, ∞)) = 1 and K(Θ) = Exp(Θ) P ↾ σ(Θ)-a.s. the (P, K(Θ))-MRP {Nt}t∈ℝ+ becomes a P-MPP with parameter Θ (see [13: Proposition 4.5]).

Remarks 1.

  1. Let {Py}yϒ be an rcp of P over Q, and let f be an inverse-measure-preserving map from Ω into ϒ. Then the following are equivalent:

    (3.1)Pyyϒisconsistentwithf.
    (3.2)PAf1(B)=BPy(A)Q( dy) for each AΣ and BT.
    (3.3) For each AΣEPχA|σ(f)=P(A)fPσ(f)-a.s..
  2. Let X be a Σ-T-measurable map from Ω into ϒ, let {Pθ}θΨ be an rcp of P over PΘ consistent with Θ, and let k be a Z-T-Markov kernel. If k(·, θ) is the distribution of X under Pθ for θΨ, then the map K(Θ) is a conditional distribution of X given Θ, since by condition (3.3) of (a), we get for A = X−1(B) with BT that PX |Θ(B, ·) = K(Θ)(B, ·) Pσ(Θ)-a.s..

  3. Conversely, in the special case where Σ is countably generated and ϒ,T=,B, given {Pθ}θΨ as in (b), we get that for each conditional distribution K(Θ) of X given Θ, there exists an essentially unique probability distribution (Pθ)X of X, for θΨ, such that for each BB, we have

    K(Θ)(B,)=PX(B)ΘPσ(Θ)-a.s..

In fact, by applying a monotone class argument, it can be easily seen that the rcp is essentially unique in the sense that if PθθΨ is any other rcp of P over PΘ which is consistent with Θ, then Pθ=Pθ for PΘ-almost all (PΘ-a.a. for short) θΨ. But the consistency of {Pθ}θΨ together with (a) yields that condition (3.3) holds true; hence setting A = X−1(B) with BB, we deduce that k(θ)(B, ·) = (P)X(B) ∘ ΘPσ(Θ)-a.s..

If no confusion arises, we denote (Pθ)XK(θ) for θΨ.

Throughout what follows, the conditional distributionK(Θ) involving in Remark 1(c) will be considered together with the distributionsK(θ), for θΨ, associated withK(Θ) as in the above remark, without any additional comments.

For the remainder of this section, {Pθ}θψis an rcp ofP over Pθ consistent with Θ and {Xi}iIis a non empty family of Σ-T-measurable maps from Ω into ϒ.

The next result extends Lemma 4.3 from [13].

LEMMA 3.1.

If {ki}iIis a non empty family of Z-T-Markov kernels, then for each iI and for any fixed BT the following conditions are equivalent:

  1. PXi|Θ(B, ·) = Ki(Θ)(B, ·) Pσ(Θ)-a.s.;

  2. PθXi1B=kiB,θforPΘ-a.aθΨ.

In particular, the same remains true if Ki(Θ)(B, ·) and ki(B, θ) are independent of i for all BT and Pθ-a.aθΨ.

Proof. Let us fix on arbitrary iI. For all BT and DZ, we obtain that

Θ1(D)PXi|Θ(B,)dP=Θ1(D)Ki(Θ)(B,)dPΘ1(D)EPχXi1(B)|σ(Θ)dP=Θ1(D)ki(B,)ΘdP(3.3)DPθXi1(B)PΘ(dθ)=Dki(B,θ)PΘ(dθ).

Consequently, the equivalence of assertions (i) and (ii) follows. □

LEMMA 3.2.

Let {ki}iIbe as in Lemma 3.1. Suppose that I is countable and T is countably generated. Then the following are equivalent:

  1. ConditionPXi|ΘKi(Θ) Pσ(Θ)-a.s. holds true for each iI;

  2. for PΘ-a.aθΨ-conditionPθXi1=ki(,θ)holds true for each iI.

In particular, the same remains true if Ki(Θ) and ki(·, θ) are independent of i.

Proof. If (i) holds true, we then get by Lemma 3.1 that for each iI and BT condition

PθXi1(B)=ki(B,θ) for PΘ-a.a.θΨ

is satisfied, which is equivalent to the fact that

iIBTL˜I,i,BZ0θL˜I,i,BPθ(Xi1(B))=ki(B,θ),

where Z0 := {LZ : PΘ(L) = 0}.

Since I is countable and T is countably generated, letting

L˜I:=BGTiIL˜I,i,B,

where GT is a countable generator of T, and applying a monotone class argument, we find a PΘ-null set L˜IZ such that for any θL˜I the equality PθXi1(B)=ki(B,θ) holds true. So assertion (ii) follows.

Applying a similar reasoning, we obtain the converse implication. □

The following result extends Lemma 4.1 from [13].

LEMMA 3.3.

Let I be countable and T countably generated. Then the family {Xi} iIis P-conditionally independent if and only if for PΘ,-a.a. θΨit isPθ-independent.

Proof. Assume that {Xi}iI is P-conditionally independent. Then according to Remark 1(a) and following the same reasoning as in the proof of [13: Lemma 4.1], we get that

DPθj=1mXijBjPΘ(dθ)=Dj=1mPθXijBjPΘ(dθ)

whenever DZ, m ∈ ℕ, i1,…,imI are distinct, and B1,…, BmT, equivalently that for each m ∈ ℕ, for all i1,…,imI distinct and for all B1,…, BmT there exists a PΘ-null set LI, m, i1,…, im, B1,…, BmZ such that for every θLI, m, i1,…, im, B1,…, Bm condition

(3.4)Pθj=1mXijBj=j=1mPθXijBj

holds true. Without loss of generality, we may and do assume that m = 2. Since T is countably generated, applying successively two monotone class arguments, we get that there exists a PΘ-null set LIZ such that for any θLI condition (3.4) holds true for m = 2, for each i1, i2I with i1i2 and for each B1, B2T; hence {Xi}iI is Pθ-independent for any θLI. Since the inverse implication is clear, this completes the proof. □

LEMMA 3.4.

Let I be countable and T countably generated. Then the following hold true:

  1. The family {Xi}iIis P-conditionally identically distributed if and only if for Pθ-a.a. θΨit isPθ-identically distributed.

  2. The family {Xi}iIis P-conditionally i.i.d. if and only if for Pθ-a.aθΨ it is Pθ-i.i.d..

Proof. Ad (i): If {Xi}iI is P-conditionally identically distributed then for any two i, jI and for each BT the equality PXi|Θ(B) = PXj|Θ(B) holds true Pσ(Θ)-a.s..

Then applying successively Remark 1(a) and a monotone class argument as that in the proof of Lemma 3.2, we find a PΘ-null set LIZ such that for any θL˜I the family {Xi}iI is Pθ-identically distributed. The inverse implication is immediate by Remark 1(a).

Ad (ii): Assume that {Xi}iI is P-conditionally i.i.d.. It then follows by assertion (i) and Lemma 3.3 that there exist two PΘ-null sets and L˜I and LI in Z such that for any θL^I:=L˜ILI the family {Xi}iI is Pθ-i.i.d.. Since the inverse implication is clear, this completes the proof. □

PROPOSITION 3.1.

The counting process {Nt}t∈ℝ+is a (P, K(Θ))-MRP if and only if for Pθ-a.a. θ ∈ ℝdit is a (Pθ, K(θ))-RP.

Proof. Assume that {Nt}t∈ℝ+ is a (P, K(Θ))-MRP, i.e. that the process {Wn}n∈ℝ is P - conditionally independent and that for all interarrival times Wn condition PWn|Θ = K(Θ) holds true Pσ(Θ)-a.s.. Applying now Lemmas 3.3 and 3.2, we equivalently get that there exist two PΘ-null sets H and H˜ in Z such that for any θH*:=HH˜ the sequence {Wn}n∈ℕ is Pθ-independent and (Pθ)wn = K(θ) for each n ∈ ℕ, respectively, i.e. such that {Nt}t∈ℝ+ is a (Pθ, K(θ))-RP for any θH*. □

Next, we need a preparatory result extending Lemmas 3.3 and 3.4 for uncountable index set. To this aim, we recall some notions more.

A filtration {Ƶt}t∈ℝ+ is said to be right-continuous if Zt=s>tZs for any t ∈ ℝ+. Let I be an arbitrary subset of ℝ+ and let ϒ be a metric space. We say that the family {Xi;}iI of Bϒ-measurable maps from Ω into ϒ is separable, if there exists a countable set GI such that for each ωΩ the set {(u, Xu(ω)) : uG} is dense in {(i, Xi(ω)) : iI}. Any such set G is called a separator (or separating set) for {Xi}ij.

Remarks 2.

Let I ⊆ ℝ+ and let Q be a probability measure on Σ. Then the following can be easily proven:

  1. If {Ut}tI is a family of Bϒ-measurable maps from Ω into ϒ, and {Ƶt}tI is its canonical filtration, then {Ut}tI is Q-independent if and only if for every bounded Bϒ-measurable realvalued function f on ϒ the equality

    (3.5)EQχAfUt=Q(A)EQfUt

    holds true for each s, tI with s < t and for each AƵs.

  2. If U1 and U2 are two Bϒ-measurable maps from Ω into ϒ, then they are Q-identically distributed if and only if EQfU1=EQfU2 for every bounded Bϒ-measurable real-valued function f on ϒ.

PROPOSITION 3.2.

Let ϒ be a Polish space, let {Xt}t∈ℝ+be a family ofBϒ-measurable maps from Ω into ϒ and letHtt+be its canonical filtration. If the family {Xt}t∈ℝ+is separable with separator+then the following hold true:

  1. IfHtt+is right-continuous, then {Xt}t∈ℝ+ is P-conditionally independent if and only if for Pθ-a.a. θΨ it is Pθ-independent.

  2. The family {Xt}t∈ℝ+ is P-conditionally identically distributed if and only if for Pθ-a.a. θ ∈ Ψ it is Pθ-identically distributed.

  3. IfHtt+is right-continuous, then {Xt}t∈ℝ+is P-conditionally i.i.d. if and only if for Pθ-a.a. θΨ it is Pθ-i.i.d..

Proof. Ad (i): The “if” implication follows as in the proof of Lemma 4.1 from [13]. For the “only if” part, note that our assumptions for {Xt}t∈ℝ+ and Htt+ imply Hs=σXuu+,us=s+,s>sHs for each s ∈ ℝ+.
  1. It follows by Lemma 3.3 that there exists a PΘ-null set O+Z such that for any θO+ condition (3.4) holds true with ℚ+ and Bϒ in the place of I and T, respectively.

    Throughout this proof fix on an arbitrary θO+. Then condition (3.4) together with Remark 2(a) implies that for all s, t ∈ ℚ+ with s < t, for every bounded Bϒ-measurable real-valued function f on ϒ and for each AHs, we have

    (3.6)EPθχAfXt=Pθ(A)EPθfXt.

    If we take s,t ∈ ℝ+ with s < t and write (3.6) for s′, t′ ∈ ℚ+ with s′ < t′ and then let s′ ↓ s and t′ ↓ t, the separability of {Xt}t∈ℝ+ together with an application of Lebesgue’s Dominated Convergence Theorem yields that for all As+,s>sHs=Hs and for every bounded continuous real-valued function f on ϒ condition (3.6) holds true.

  2. Let s, t ∈ ℝ+ with s < t and let f be a function as in (3.5). Then for each n ∈ ℕ there exists a bounded continuous real-valued function gn on ϒ satisfying the inequality gnfdPθXt1n (cf. e.g., [7: Proposition 415P]); hence there exists a sequence {gn}n∈ℕ of bounded continuous real-valued functions on ϒ such that condition limnχAgnfXtdPθ=0 holds true for all AHs, implying together with (a) that EPθχAfXtn=limnEPθχAgnXt=limnPθ(A)EPθgnXt=Pθ(A)EPθfXt for all AHs; hence by Remark 2(a), we get that {Xt}t∈ℝ+ is Pθ-independent, which proves (i).

    Ad (ii): The “if” implication is immediate by Remark 1(b). For the “only if” part, assume that {Xt}t∈ℝ+ is P-conditionally identically distributed.

  3. Since {Xt}t∈ℝ+ is P-conditionally identically distributed, we get that for any two s, t ∈ ℚ+ and for each BBϒ the equality PXt|Θ(B) = PXs|Θ(B) holds Pσ(Θ)-a.s.. The latter together with Lemma 3.4(i) yields the existence of a PΘ-null set O˜+Z such that for any θO˜+ and for all s, t ∈ ℚ+ condition (Pθ)Xt = (Pθ)Xs holds true, which by Remark 2(b) equivalently yields that for any θO˜+, for every function f as in the above remark, and for all s, t ∈ ℚ+, we have EPθfXt=EPθfXs.

    Till the end of the proof of (ii), fix on an arbitrary θO˜+.

  4. If we take s, t ∈ ℝ+ and if we write the last equality for s′, t′ ∈ ℚ+ and then let s′ ↓ s and t′ ↓ t, the separability of {Xt}t∈ℝ+ together with an application of Lebesgue’s Dominated Convergence Theorem yields that for every bounded continuous real-valued function f on ϒ condition EPθfXt=EPθfXs holds true.

Following now the same reasoning with that of steps (b) and (c), we obtain that the last equality is satisfied by all functions f as in Remark 2, and all s, t ∈ ℝ+, which is equivalent to the fact that condition (Pθ)Xt = (Pθ)Xs holds true for all s, t ∈ ℝ+; hence (ii) follows.

Ad (iii): Assume that {Xt;}t∈ℝ+ is P-conditionally i.i.d. and that its canonical filtration is right-continuous. It then follows by assertions (i) and (ii) that there exist two PΘ-null sets O˜+ and O+ in Z such that for any θO^+:=O˜+O+ the family {Xt}t+ is Pθ-i.i.d.. Since the inverse implication is clear, assertion (iii) follows. □

The next result seems to be of independent interest, since it shows among others that for any counting process the property of conditionally independent increments under a probability measure can be reduced to that of independent increments under the probability measures of an rcp. The same result extends a basic one, namely [13: Proposition 4.4].

THEOREM 3.3.

Let {Nt}t∈ℝ+be a counting process and letK˜tθt+be a family of probability distributions onBwith parameter θΨ. Then {Nt}t∈ℝ+hasP-conditionally stationary independent increments such that condition

PNtΘ=K˜t(Θ)Pσ(Θ)-a.s.
holds true for eacht ∈ ℝ+if and only if for Pθ-a.a. θΨ it has Pθ-stationary independent increments such thatPθNt=K˜tθfor each t ∈ ℝ+.

Proof. Since {Nt}t∈ℝ+ is a counting process it has right-continuous paths; hence it is separable with separator ℚ+. Note also that the canonical filtration of {Nt}t∈ℝ+ is right-continuous (see [20: Theorem 25], where the proof works for any probability space not necessarily complete). Thus, all assumptions of Proposition 3.2 are fulfilled, and so, we may apply it to deduce the thesis of the theorem. □

Corollary 3.3.1([13: Proposition 4.4])

The family {Nt}t∈ℝ+is a P-MPP with parameter Θ if and only if it is a Pθ-Poisson process with parameter θ for Pθ-a.a. θ.

4. Further characterizations

Recall first that an infinite family {Xi}iI of Σ-T-measurable maps from Ω into ϒ is said to be exchangeable under P or P-exchangeable, if for each r ∈ ℕ, we have

Pk=1rXik1Bk=Pk=1rXjk1Bk

whenever i1, …, irI are distinct, j1, …, jrI are distinct, and BkT for each kr (cf. e.g., [7: 459C]). Also recall the following notions.

Let Q be a probability measure on T. Assume that M is a probability on the σ-algebra ΣT such that P and Q are the marginals of M. Assume also that for each yϒ there exists a probability Py on Σ, satisfying the following properties:

  1. For every AΣ the map yPy(A) is T-measurable;

  2. MA×B=BPyAQ(dy) for each A × BΣT.

Then, {Py}yϒ is said to be a product regular conditional probability (product rcp for short) on Σ for M with respect to Q (see [5: Section 2] or [23: Definition 1.1]).

Let I be an arbitrary non-empty index set. If {(Ωi, Σi, Pi)}iI is a family of probability spaces then, for each 0JI, we denote by (ΩJ, ΣJ, PJ) the product probability space iJΩi,Σi,Pi:=iJΩi,iJΣi,iJPi. If (Ω, Σ, P) is a probability space, we write PI for the product measure on ΩI and ΣI for its domain.

LEMMA 4.1.

LetFbe a σ-subalgebra of Σ and let {Xi}iIbe a non empty family of Σ-T-measurable maps from Ω into ϒ such that {Xi}iIis P-conditionally i.i.d. overF. suppose that T is countably generated and that Pχi is perfect for each iI. Then there exists a probability measure M onTF with marginal R:=PFsuch that M := P ∘ (Xi × idΩ)−1for every iI, and a product rcp {Qω}ωΩon T for M with respect to R, such that

  1. for any fixedBT and iI the map Q(B): Ω → [0, 1] isR-a.s. equal toPXi1B|F;

  2. FQωI(H)R( dω)=PFX1(H)for everyFFandHTi, whereQωIdenotes the I-fold product probabilityiIPi of copies Pi := Qω of Qω for iI, and X : ΩϒI is defined by X(ω) = (Xi(ω))i∈I for each ωΩ.

Proof. First fix on an arbitrary iI.

  1. The function Xi × idΩ from Ω into ϒ × Ω defined by means of

    Xi×idΩ(ω):=Xi(ω),ω for each ωΩ

    is Σ-TF -measurable. So, we have a probability measure Mi := P ∘ (Xi × idΩ)−1 on TF. Since all Xi are P-conditionally identically distributed over F, it follows that Mi is independent of i, so we may write M := Mi* for any fixed i* ∈ I.

  2. There exists a product rcp {Qω}ωΩ on T for M with respect to R=PF such that for any fixed BT

    Q(B)=PXi1(B)F()R a.s..

    In fact, by assumption each marginal measure PXi of M on T is perfect and T is countably generated; hence by [5: Theorem 6], there exists a product rcp {Qω}ωΩ on T for M with respect to R.

    Since {Qω}ωΩ satisfies (D2), we get that

    FQω(B)R(dω)=M(B×F)=PFXi1(B)=FPXi1(B)|F(ω)R(dω)

    for every BT and FF, which proves (b); hence (i) follows.

  3. Using (i) and a monotone class argument, we get that (ii) holds true. □

    The next result extends a corresponding one due to Olshen (see [18: Theorem (3)]) concerning a generalization of de Finetti’s Theorem.

PROPOSITION 4.1.

Let {Xi}iIbe a P-exchangeable infinite family of Σ-T-measurable maps from Ω into ϒ. suppose that T is countably generated and PXi is perfect for each iI. Then there exists a d-dimensional random vector Θ such that {Xi}iIis P-conditionally i.i.d. given Θ.

Proof. Since {Xi}iI is P-exchangeable, it follows by [7: Theorem 459B], that there exist a σ-subalgebra F of Σ such that {Xi}iI is P-conditionally i.i.d. over F. So, applying Lemma 4.1, we deduce that there exists a family {Qω}ω∈Ω of F-T-Markov kernels such that

FQωI(H)R(dω)=PFX1(H)

for every HT1 and FF, where R:=PF. Then there exists a countably generated σ-subalgebra A of F such that Q(B) is A-measurable for arbitrary but fixed BT (take, e.g., AB:=Q(B)1(E):EGB for BT, and A:=σBGTAB,, where GB and GT is a countable generator of B and T, respectively). Since A is countably generated, there exists a map Θ˜:Ω such that A=σΘ˜ (take, e.g., Θ˜ to be the Marczewski functional on Ω, cf. e.g., [6: 343E] for the definition). But since {Xi}i∈I is P-conditionally i.i.d. over F and AF, it follows that {Xi}i∈I is so over A=σΘ˜.

Note also that since ℝ and ℝd are standard Borel spaces of the same cardinality, there exists a Borel isomorphism g from ℝ into ℝd (cf. e.g., [7: Corollary 424D(a)]). So, putting Θ:=gΘ˜, we get that Θ is a d-dimensional random vector on Ω such that σΘ=σΘ˜. □

Corollary 4.1.1 (see R. Olshen, [18: Theorem 3]).

If {Xn }∈ℕis a P-exchangeable sequence of measurable maps from Ω into a complete, separable metric space, then there exists a real-valued random variable Θ on Ω such that {Xn}∈ℕis P-conditionally i.i.d. given Θ.

Theorem 4.2.

Let {Xi}iIbe an infinite family of Σ-T-measurable maps from Ω into ϒ. Consider the following assertions:

  1. {Xi}i∈I is P-exchangeable.

  2. There exists a σ-subalgebraFof Σ such that {Xi}iIP-conditionally i.i.d. overF.

  3. There exists a σ-subalgebraFof Σ and a family {Qω}ωΩofF-T-Markov kernels such that

    FQωI(H)R(dω)=PFX1(H)
    for every HTI andFF, whereR:=PFandQωI, X are as in Lemma 4.1.
  4. There exists aBd-measurable map Θ from Ω intodsuch that {Xj}iIis P - conditionally i.i.d. given Θ.

Then (iii) ⇒ (i) ⇔ (ii) and (iv) ⇒ (i). If any one of conditions (i) to (iv) is satisfied, then all image measures PXi are equal.

Moreover, if PXi is perfect for any iI and T is countably generated, then assertions (i) to (iv) are all equivalent.

Proof. The equivalence (i) ⇔ (ii) follows by [7: Theorem 459B], while the implications (iii) ⇒ (i) and (iv) ⇒ (i) are evident.

Clearly, if assertion (i) or equivalently (ii) is satisfied then all PXi are equal and the same applies if (iii) or (iv) holds true.

Moreover, if every measure PXi is perfect and T is countably generated, then implications (ii) ⇒ (iii) and (i) ⇒ (iv) follow from Lemma 4.1 and Proposition 4.1, respectively. So, we get that assertions (i) to (iv) are all equivalent. □

COROLLARY 4.2.1.

Let {Xt}t∈ℝ+be a family of Σ-T-measurable maps from Ω into ϒ. Suppose that Σ is countably generated, P is perfect, ϒ is a Polish space, {Xt}t∈ℝ+is separable with separator+and that its canonical filtration is right-continuous. Then each of the items (i) to (iv) of Theorem 4.2is equivalent to condition

  1. there exist a d-dimensional random vector Θ and an rcp {Pθ}θ∈ℝdof P over Pθ consistent with Θ such that {Xt}t∈ℝ+is Pθ-i.i.d. for Pθ-a.a. θ ∈ ℝd.

Proof. It follows by our assumptions for P, Σ and ϒ, that given a d-dimensional random vector Θ, there exists an rcp {Pθ}θ∈ℝd of P over PΘ consistent with Θ. Thus, we may apply Proposition 3.2 to obtain that condition (v) is equivalent to (iv) of Theorem 4.2. The equivalence of all items (i) to (v) is immediate by Theorem 4.2. □

Remarks 3.

  1. The assumption “PXi perfect” made in the last theorem is easily verified in the usual applications, since this is covered by the following facts: (α) If ϒ is a Polish space then each PXi is Radon (cf. e.g., [7: Proposition 434K(b)] and [7: Definition 411H(b)] for the definition of a Radon measure); hence perfect (cf. e.g., [7: Proposition 416W(a)]). (β) If P is perfect then each PXi is so (cf. e.g., [7: Proposition 451E(a)]). (γ) If Ω = ϒI, Xi(iI) are the canonical projections from Ω onto ϒ, and P is any probability measure on Σ := TI then each PXi is perfect if and only if P is perfect (cf. e.g., [7: Theorem 454A(b)(iii)]).

  2. To the best of our knowledge, the most general result concerning the equivalence of assertions (i) and (iii) of Theorem 4.2 is Theorem 1.1 from [12], which extends de Finetti’s Theorem, saying that for each infinite sequence {Xn}n∈ℕ of random variables taking values in a standard Borel space ϒ (i.e. ϒ is isomorphic to some Borel-measurable subset of ℝ) assertions (i) and (iii) of Theorem 4.2 with {Xn}n∈ℕ in the place of {Xi}iI are equivalent. It is well-known that any Polish space is standard Borel; in particular, ℝd and ℝ are such spaces.

  3. There are measurable spaces (ϒ, T) satisfying the assumptions of Theorem 4.2, i.e., that T is countably generated and each PXi is perfect, which are not standard Borel spaces; hence Theorem 4.2 extends Theorem 1.1 from [12]. In fact, it is known that each uncountable analytic Hausdorff space (i.e., a non-empty topological Hausdorff space being a continuous image of the space ℕ, cf. e.g., [7: Definition 423A]) has a non-Borel analytic subset (cf. e.g., [7: Proposition 423L]). It is also known that for each analytic Hausdorff space ϒ the Borel σ-algebra Bϒ is countably generated (cf. e.g., [7: 423X(d)]), and that any Borel probability measure on Bϒ is always inner regular with respect to compact sets (see [10: Chapter IV, Theorem 1, p. 195]); hence it is perfect (cf. e.g., [7: Proposition 451C]). Consequently, each uncountable analytic Hausdorff space has a subset satisfying the assumptions of Theorem 4.2, but not being a standard Borel space.

  4. Restricting attention to measurable spaces (ϒ, T) satisfying the countability assumption of T as in Theorem 4.2, costs us some generality; for instance, the general compact Hausdorff space does not satisfy the countability assumption for T, and it is known that the equivalence of assertions (i) to (iii) of Theorem 4.2 is true for countable products of compact Hausdorff spaces (see [9] or [3]). More general, the equivalence of assertions (i) to (iii) of Theorem 4.2 is proven in [7: Theorem 459G] for uncountable products of general Hausdorff spaces. But all the above are specialized in the product situation of topological spaces, while assertions (i) to (iii) of Theorem 4.2 have the advantage of being free from any topological assumption as well as from any product situation.

The following definition of an MRP traces back to Huang [11: Section 1, Definition 3].

Definition 4.1.

The counting process {Nt}t∈ℝ+ is said to be a ν-mixed renewal process associated with Py˜y˜ϒ˜, if for every r ∈ ℝ and for all w1,…, wr ∈ ℝ condition

Pk=1rWkwk=k=1rPy˜Wkwkν(dy˜),

holds true, where Py˜y˜ϒ˜ is a family of probability measures on Σ and ν is a probability measure on B(ϒ˜):=σP(E):EΣ such that for ν-a.a. y˜ϒ˜ the process {Wn}n∈ℕ is Py˜-identically distributed.

In Huang’s definition it is assumed that {Nt}t∈ℝ+ takes values only in ℕ0, which is equivalent to the mild assumption that {Nt}t∈ℝ+ has zero probability of explosion, that is Pt(0,)Nt==0 (cf. e.g., [21: Lemma 2.1.4]).

Remark 4.

Note that in Huang’s [11] definition the assumption that {Wn}n∈ℕ is Py-identically distributed for ν-a.a. y˜ϒ˜ is not written explicitly. But this assumption must be included there, since it is necessary for the validity of the Corollary in page 20 of [11], as it follows from Example 5 below.

In fact, consider the process {Nt}t∈ℝ+ of Example 5, where the above assumption does not hold true, as well as Huang’s definition of a ν-MRP without the above assumption. Also note that q := P(Z < ∞) = 0 < 1, where Z is the almost sure limit of {Nt}t∈ℝ+ as t → ∞. Assume, if possible, that the Corollary in [11: p. 20] holds true. Then conditional on the event {Z = ∞} the process {Nt}t∈ℝ+ has the exchangeable property (E) (see [11: Definition 1] for the definition) implying that {Wn}n∈ℕ is exchangeable, a contradiction to Example 5.

THEOREM 4.3.

Consider the following assertions:

  1. There exists a d-dimensional (d ∈ ℕ) random vector Θ such that {Nt}t∈ℝ+is a (P, K(Θ))-MRP.

  2. There exist a random vector Θ, an rcp {Pθ}θ∈ℝdof P over PΘ consistent with Θ, and a family {K(θ)}θ∈ℝdofZB-Markov kernels for Pθ-a.a. θ ∈ ℝdsuch that the family {Nt}t∈ℝ+is a (Pθ,K(θ))-RP.

  3. The process {Wn}n∈ℕis P-exchangeable.

  4. There exist a σ-subalgebraFofΣand a family {Qω}ωΩofFB-Markov kernels such that

    FQω(H)R(dω)=PFW1(H)
    for everyHBandFF, whereRPF, W := (W1, … Wn …) andQωdenotes the ℕ-fold product probabilityn∈ℕPnof copies Pn := Qω of Qω for n ∈ ℕ.
  5. There exist a σ-subalgebraFof Σ, a subfield rcp {Sω}ωΩfor P over the restrictionRPF, and a family {K(ω)}ωΩofFB-Markov kernels such that for R-a.a. ωΩ the sequence {Wn}n∈ℕis Sω-independent and condition (Sω)Wn = K(ω) holds for each n.

  6. There exist a setϒ˜, a familySy˜y˜ϒ˜; of probability measures on Σ and a probability measure ν onBϒ˜:=σS(E):EΣsuch that {Nt}t∈ℝ+is a ν-MRP associated withSy˜y˜ϒ.

Then the following implications hold true:

 (vi)  (ii)  (v)  (iii)  (iv)  (vi) 
Moreover, if Σ is countably generated and P is perfect then items (i) to (vi) are all equivalent.

Proof. First note that the implications (i) ⟹ (iii) and (vi) ⟹ (iii) are obvious. The implication (ii) ⟹ (i) is immediate by Proposition 3.1, while the implication (iii) ⟹ (i) and the equivalence of (iii) and (iv) follow directly by Proposition 4.1 and Theorem 4.2, respectively, since for ϒ,T=,B every measure PWn on B is perfect and B is countably generated. The latter equivalence together with implication (vi) ⇒ (iii) yields (vi) ⇒ (iv).

Ad (ii) ⇒ (iv): If (ii) holds true, then there exists a PΘ-null set H*Bd such that for any θH*, the process {Wn}n∈ℕ is Pθ-exchangeable, implying together with property (d2) its P-exchangeability as well; hence (iii) or equivalently (iv) follows.

Ad (ii) ⇒ (v): Assume that (ii) is true. Putting Sω(E) := PΘ(E) for any ωΩ, EΣ and θ = Θ(ω), we clearly get that {Sω}ωΩ is a subfield rcp for P over R := Pσ(Θ) such that the interarrival times Wn, n ∈ ℕ, are Sω-i.i.d. with a common probability distribution K(ω) for any ωH** := Θ−1(H*), where K(ω) := K(θ) for each ωΩ and Θ(ω)=θHBd. Since clearly H** is an R-null set, it follows that {Sω}ωΩ, F:=σΘ and {K(ω)}ωΩ satisfy assertion (v).

Ad (v) ⇒ (vi): Assume that (v) holds true and let F, {Sω}ωΩ and R be as in (v). Put ϒ˜:=Ω, Sy˜y˜ϒ˜:=SωωΩ and B(ϒ˜):=σS(E):EΣ. Then B(ϒ˜)F and so, we may define the probability measure ν:=RB(ϒ˜). Since by (v) the process {Wn}n∈ℕ is Sω-i.i.d. for R-a.a. ωΩ, we get that it is Sy˜-i.i.d. for ν-a.a. y˜ϒ˜. The latter together with the definition of a subfield rcp yields that {Nt}t∈ℝ+ is a ν-MRP associated with Sy˜y˜ϒ; hence assertion (vi) follows.

Moreover, if Σ is countably generated and P is perfect, the implication (iv) ⇒ (v) holds true. In fact, by Theorem 4.2, we obtain that assertion (iv) is equivalent with the fact that {Wn}n∈ℕ is P-conditionally i.i.d. over F. But since Σ is countably generated and P is perfect, there exists a subfield rcp {Sω}ωΩ for P over R:=PF. Thus, we may apply Lemma 3.4(ii) Ψ,Z:=Ω,F and Θ := idΩto get (v).

Assuming now that (i) holds true, it follows again by our assumptions for Σ and P that there exists an rcp {Pθ}θ∈ℝd of P over PΘ consistent with Θ. So according to Proposition 3.1, we get that for PΘ-a.a. θ ∈ ℝd the family {Nt}t∈ℝ+ is a (Pθ, K(θ))-RP; hence (i) implies (ii).

Thus, assuming that Σ is countably generated and P is perfect, we obtain that items (i) to (vi) are all equivalent. This completes the whole proof. □

Note that the most important applications in Probability Theory are still rooted in the case of standard Borel spaces; hence of spaces satisfying always the assumptions of the above theorem concerning P and Σ.

5. Examples and applications

In this section, we provide two groups of examples: The first one (Examples 1 to 2) shows that the perfectness assumption for the probability measures as well as the countability assumption for the σ-algebras involved in Theorems 4.2, 4.3 and in Corollary 4.2.1, are essential for the validity of the equivalences obtained therein.

In the second group of examples (Examples 3 to 5), the existence of non-trivial probability spaces admitting (P, K(Θ))-MRPs with prescribed distributions for their interarrival processes as well as for the parameter Θ is proven, providing in this way a method of constructing such processes. As a consequence, concrete examples of MRPs are presented. It is also worth noticing that our construction relies on Proposition 3.1 and allows us to obtain probability spaces that satisfy all assumptions of Theorems 4.2 and 4.3.

It follows an example to show that the perfectness assumption for the probability measures PXi in Theorem 4.2 and its Corollary 4.2.1 is essential for the validity of the equivalences of (i) to (v).

Example 1.

Let ϒ := ℝ+ and let Q˜ be a probability measure on Bϒ. Consider a subset B of ϒ such that Q˜*(B)=Q˜*(Bc)=1 where Q˜ is the outer measure induced by Q˜. Let TσBϒ,B and let Q: T → [0, 1] be the probability measure defined by

Q(A):=Q˜(AB) for each AT.

Then Q is non perfect. Set Ω := ϒ, Σ := T, P := Q, Wn := πn : Ωϒ, where πn is the canonical projection for any n ∈ ℕ. Clearly, assertion (i) of Theorem 4.2 is satisfied by {Wn}n∈ℕ; hence by Theorem 4.2, we equivalently get that (ii) holds true.

Assume that assertion (iii) of Theorem 4.2 is valid, i.e. that there exists a σ-subalgebra F of Σ and a family {Qω}ωΩ of F-T-Markov kernels such that

FQω(H)R(dω)=PFW1(H)

for every HT and AF, where R:=PF and Qω, W are as in Theorem 4.2. Then QωωΩ is a subfield rcp for P over R; hence, we may and do assume that F is countably generated. Applying now a monotone class argument we deduce that there exists an R-null set NF such that for each AF condition Qω(A)=1 holds true for any ωNcA.

But since QωωΩ is a subfield rcp, we get for every FF that

FQωBR(dω)=PFB=P(F)=FχFdR,

implying that P(D) = 0, where D:=ωΩ:QωB1.

Put E := DN. For any ωEc, we get Qω({ω})=1 and QωB=1 hence QωB{ω}=1 implying B(ω}0 or ωB. Thus, we get EcB. But then Q˜Bc=1 yields P(Ec) = 0 or equivalently P(E) = 1, a contradiction.

Assume now that assertion (v) of Corollary 4.2.1 is valid, i.e., that there exist a d-dimensional random vector Θ and an rcp {Pθ}θ∈ℝd of P over PΘ consistent with Θ such that {Wn}n∈ℕ is Pθ-i.i.d. for PΘ-a.a. θ ∈ ℝd. For any fixed AΣ define R˜(A):Ω[0,1] by R˜ω(A):=P(A)Θ(ω). Then R˜ωωΩ is a subfield rcp for P over Pσ(Θ), a contradiction according to the above arguments.

It follows an example to show that the countability assumption for T or Σ in Corollary 4.2.1 is essential for the validity of the equivalences of (i) to (v) in Theorem 4.2 and Corollary 4.2.1.

Example 2.

Let ϒ := ℝ+, let Q˜ be a probability measure on Bϒ, let T be the completion of Bϒ with respect to Q˜ and let Q be the completion of Q˜. Put Ω := ϒ, Σ˜:=T and P˜:=Q. Denote by Σ the completion of Σ˜ with respect to P˜ and by P the completion of P˜. Then P is a perfect probability measure on Σ (cf. e.g., [7: Proposition 451G and Theorem 451J]) but Σ and T are not countably generated.

Put Wn := πn : Ωϒ for any n ∈ ℕ. Clearly, {Wn}n∈ℕ is P-exchangable, that is, assertion (i) of Theorem 4.2 is valid; hence assertion (ii) of the same theorem is also valid.

Assume now that assertion (v) of Corollary 4.2.1 is valid. It then follows that there exists a d-dimensional random vector Θ and an rcp {Pθ}θ∈ℝd of P over PΘ consistent with Θ, implying the existence of a subfiled rcp {Rω}ωΩ of P over Pσ(Θ), where each function R(A): Ω → [0, 1] is defined by Rω(A) := (P(A) ∘ Θ)(ω) for any fixed AΣ. Since σ(Θ) is countably generated, it follows as in Example 1 that there exists a set Nσ(Θ) such that P(N) = 0 and for any Aσ(Θ) condition Rω(A) = 1 holds true for any ωNcA.

Choose a set DNc such that Dσ(Θ) but DΣ. Then for each ωN, we obtain

1=Rω({ω})Rω(D)1 if ωD

and

1=Rω({ω})RωDc1 if ωDc.

Thus, D = Nc ⋂ {ωΩ : Rω(D) = 1} ∈ σ(Θ), which is impossible by the choice of D; hence assertion (v) of Corollary 4.2.1 is not valid.

The next counterexample shows that there exists a non perfect probability space (Ω, Σ, P) and a counting process {Nt}t∈ℝ+ on it satisfying conditions (i), (iii), (iv) and (vi) of Theorem 4.3 but not (ii) and (v); hence the perfectness of P is an essential assumption.

Counterexample 1.

Let ϒ := ℝ+, let Q be a probability measure on T:=Bϒ, and let (Ω,Σ˜,P˜):=Υ,T,Q. Consider a subset B of ϒ such that P˜(B)=P˜Bc=1 the σ-algebra Σ:=σ({Σ˜B}) and the non perfect probability measure P on Σ defined by means of

P(A):=P˜(AB) for each AΣ.

Let Wn := πn : Ωϒ be the canonical projection for any n ∈ ℕ. Clearly, assertion (iii) of Theorem 4.3 is satisfied by {Wn}n∈ℕ, and so, we equivalently get that assertions (iv) and (i) of the same theorem also hold true. Furthermore, it can be easily seen that (iv) ⇒ (vi) as in Theorem 4.3; hence (iv) ⇔ (vi).

Applying similar arguments with those in Example 1, we obtain that assertions (v) and (ii) of Theorem 4.3 are not valid.

The next counterexample shows that the countability assumption for Σ in Theorem 4.3 is essential for the equivalence of items (i) to (vi) of Theorem 4.3.

Counterexample 2.

Let (ϒ, T, Q) and (Ω, Σ˜, P˜) be as in Example 1, let Σ be the completion of Σ˜ with respect to P˜ and P the completion of P˜, and let {Wn}n∈ℝ be as in Example 1. Then P is a perfect probability measure on Σ but Σ is not countably generated.

It then follows that assertion (iii) of Theorem 4.3 holds true, while by the same theorem, we get that its assertions (i), (iii) and (iv) are all equivalent. Furthermore, it can be easily shown that (iv) implies (vi). But using similar arguments as in Example 2, we get that assertion (ii) of Theorem 4.3 is not valid.

Assume that assertion (v) of Theorem 4.3 holds true. It then follows that there exist a σ-subalgebra of Σ and a subfield rcp {Sω}ωΩ of P over R:=P.

There exists a countably generated σ-subalgebra A of F such that {Sω}ωΩ satisfies

(5.1)FSω(E)R(dω)=P(EF) for all FA and EΣ.

and

(5.2)HΣNHF with PNH=0S(H)NHc is A-measurable. 

In fact, applying similar arguments with those in the proof of Proposition 4.1, we get a countably generated σ-subalgebra A˜ of F such that for any fixed H˜Σ˜ the function S(H˜) is A˜-measurable. Since for R-a.a. ωΩ the measures P and Sω have the same null sets, we obtain that for any H ∈Σ there exist sets H˜Σ˜, MHΣ and NHFΣ0 such that H=H˜MH and Sω(MH) =P(MH) = 0 for all ωNH; hence {Sω}ωΩ satisfies (5.2) with A˜ in the place of A. Applying a monotone class argument, we get a P-null set N1F such that for any fixed H˜A˜ the function S(H˜)N1c is A˜-measurable.

Put A:=σA˜N1F. It then follows that A is countably generated and {Sω}ωΩ satisfies (5.1) and (5.2). In particular, for any fixed HA the function S(H) ↾ (N1)c is A-measurable.

Applying now a monotone class argument, we obtain a P-null set N2A such that for any HA and ωNcH, where N := N1N2, condition Sω(H) = 1 holds true.

Choose a set DNc such that DA but DΣ. Following the same reasoning as in Example 2, we deduce thatDA which is impossible by the choice of D; hence assertion (v) of Theorem 4.3 is not valid.

Throughout what follows, we putΩ˜:=, Ω:=Ω˜×d, Σ˜:=B(Ω˜)andΣ:=Σ˜Zfor simplicity. The next result extends Theorem 3.1 from [14], which provides a construction for MPPs.

Example 3.

Following the reasoning of Theorem 3.1 from [14] but with ℝd, B and Qn(θ) = K(θ) in the place of ϒ := (0, ∞), Bϒ and Qn(θ) = Exp(θ), respectively, the existence of (P, K(Θ))- MRPs with prescribed distributions for their interarrival processes as well as for the parameter Θ is proven.

In fact, fix on arbitrary θ ∈ ℝd and put P˜θ:=nQn(θ). Since by assumption, for any fixed BB each function Qn(·)(B) is Z-measurable, it follows by a monotone class argument that the same holds true for the function P˜(E) for fixed EΣ˜.

For each θ ∈ ℝd put Pθ:=P˜θδθ where δθ is the Dirac probability measure on Z, and for each n ∈ ℕ set Wn := πn, where πn is the canonical projection from Ω onto ℝ. Put now

P(E):=P˜θEθμ(dθ) for each EΣ,

where Eθ := {ωΩ : (ω, θ) ∈ E}. Then P is a probability measure on Σ such that {Pθ}θd is an rcp of P over μ consistent with πd, where πd is the canonical projection from Ω onto ℝd (see [14: proof of Theorem 3.1]). Furthermore, it can be proven that for all θ ∈ ℝd the process {Wn}n∈ℕ is pθ-independent and (pθ)Wn = K(θ) for each n ∈ ℕ. Clearly, putting Θ := πd, we get Pθ = μ.

It then follows that the counting process {Nt}t∈ℝ induced by {Wn}n∈ℕ (cf. e.g., [21: Theorem 2.1.1]) is a (Pθ, K(θ))-RP for all θ ∈ ℝd; hence by Proposition 3.1 it is a (P, K(Θ))-MRP.

Applying now Example 3, we compute the probability measures of the corresponding rcp {Pθ}θ∈ℝd as well as the probability measure P for some MRPs of special interest which are not MPPs. To this aim recall that by λd is denoted the restriction of the Lebesgue measure λd to Bd, while any restriction of λd to BA, where A is any Borel subset of ℝd, will be denoted again by λd. In particular, if d = 1 then λ:=λ1=λB, where λ is the Lebesgue measure on ℝ.

In the next example, a concrete (P, K(Θ))-MRP is constructed for one of the most common choices that can be made for an interarrival time distribution, i.e. Ga(θ1, θ2) with θ1 > 0 and θ2 = 1/2 ∈ (0,1), see e.g. [8: p. 95]. In fact, this class of distributions is of special interest, since none of its members satisfy Assumption 5.1 from [8], proposed by Huang in [11: Theorem 3], which is essential in Grandell’s study for MRPs (see [8: Section 5.3]). Moreover, in the same example it is shown that there are counting processes {Nt}t∈ℝ+ being both (P, K(Θ))-MRPs and PΘ ↾ B(ℝ)-ones, which are not, though, MRPs according to Grandell [8: Definition 5.3].

Example 4.

Let Qn(θ) = Ga(θ, 1/2) for each n ∈ ℕ and for any fixed θ > 0, and let μ = Ga(γ, α). Then the conclusions of Example 3 are fulfilled for d = 1; hence Ω˜=, Ω= ℝ × ℝ, while Σ˜, Σ, P˜, P, P˜θθ>0, {Pθ}θ>0, and Θ are as in Example 3.

We first compute the probability measures on measurable cylinders. Let C˜ denote the family of all measurable cylinders B˜B(Ω˜) i.e. of all sets B˜Ω˜ expressible as nB˜n where B˜nB for every n ∈ ℕ, and L˜:=n:B˜n is finite. Set C˜n=B˜n for nL˜. Then B˜=kL˜C˜k×N\L so we get

(5.3)P˜θ(B˜)=(nQn(θ))(B˜)=kL˜Qk(θ)(C˜k)=θπkL˜C˜kωk12eθωkλ(dωk)

for each θ > 0. Consider now a measurable cylinder B˜×EC˜×B0,. Applying (5.3), we get

Pθ(B˜×E)=P˜θ(B˜)δθ(E)=χE(θ)θπkL˜C˜kωk12eθωkλ(dωk);

for each θ > 0. Hence,

P(B˜×E)=γαΓ(α)πE[kL˜C˜kωk12eθ(γ+ωk)λ(dωk)]θα12λ(dθ).

As a consequence, by applying standard methods of Topological Measure Theory, the probability measures P(E) and Pθ(E) can be computed for any EΣ. For details see [14: Example 3.3, (b)].

Finally, it follows an example to show that, we cannot avoid including in Huang’s definition of an MRP the assumption that {Wn}n∈ℕ is Py-identically distributed for ν-a.a. y˜ϒ˜ (see also Remark 4).

Example 5.

Let d = 1. If Qn(θ) = Exp() for each n ∈ ℕ and for any fixed θ > 0, and if μ = Ga(2, 1) then all requirements of Example 3 except for

Qn(θ)=K(θ) for all n and for any fixed θ>0

are satisfied. In fact, in this case K(θ) is substituted by K() := Exp().

So the probability measures P˜ and P on Σ˜=B(Ω˜) and Σ=BΩ, where Ω˜= and Ω = ℝ × ℝ, respectively, as well as the rcps P˜θθ>0 and {Pθ}θ>0 can be computed. Moreover, there exists a random variable Θ on Ω such that PΘ = Ga(2, 1). Following the same reasoning as in Example 3, we also obtain an interarrival process {Wn}n∈ℝ which is Pθ-independent for all θ > 0 and satisfies Wn = πn as well as (Pθ)Wn = K() for all n ∈ ℕ and for any fixed θ > 0.

But the Pθ-independence of {Wn}n∈ℕ, for all θ > 0, implies for every r ∈ ℕ and for all w1,…, wr ∈ ℝ+ that

(5.4)Pk=1rWkwk=k=1rPθWkwkν(dθ),

where ν = PΘB((0, ∞)) = μB((0, ∞)) and B((0, ∞)) = σ({P(E) : EΣ}). So, {Wn}n∈ℕ is an interarrival process which is not Pθ-identically distributed for any fixed θ > 0 but which satisfies (5.4). As a consequence, the counting process {Nt}t∈ℝ+ induced by the sequence of canonical projections {πn}n∈ℕ = {Wn}n∈ℕ (cf. e.g., [21: Theorem 2.1.1]) is not a ν-MRP associated with {Pθ}θ>0. Furthermore, for every w1, w2 ∈ ℝ+, we have

PW1w1,W2w2=201eθw11e2θw2e2θdθ=w2w2+112w1+21w1+2w2+21,

implying that PW12,W21=1327=PW11,W22; hence {Wn}n∈ℕ is not P-exchangeable.


(Communicated by Gejza Wimmer)

D.P.L. would like to aknowledge that the main part of this work was conducted at the Department of Statistics and Insurance Science in the University of Piraeus. The author is also indebted to the Public Benefit Foundation Alexander S. Onassis, which supported this research, under the Programme of Scholarships for Hellenes. D.P.L. would also like to state that this work should not be reported as representing the views of the Hellenic Statistical Authority (ELSTAT). The views expressed are those of the authors and do not necessarily reflect those of ELSTAT.


REFERENCES

[1] COHN, D. L.: Measure Theory. Advanced Texts, 2nd edition, Birkhäuser, 2013.10.1007/978-1-4614-6956-8Search in Google Scholar

[2] DELBAEN, F.—HAEZENDONCK, J.: A martingale approach to premium calculation principles in an arbitrage free market, Insurance: Math. Econom. 8 (1989), 269–277.10.1016/0167-6687(89)90002-4Search in Google Scholar

[3] DIACONIS, P.—FREEDMAN, D.: Finite exchangeable sequences, Ann. Probab. 8 (1980), 745–764.10.1214/aop/1176994663Search in Google Scholar

[4] DUDLEY, R. M.: Real Analysis and Probability, Wadsworth & Brooks/Cole, Advanced Books & Software, Pacific Grove, California, 1989.Search in Google Scholar

[5] FADEN, A. M.: The existence of regular conditional probabilities: Necessary and sufficient conditions, Ann. Probab. 13 (1985), 288–298.10.1214/aop/1176993081Search in Google Scholar

[6] FREMLIN, D. H.: Measure Theory, Measure Algebras, Vol. 3, Torres Fremlin, Colchester, 2002.Search in Google Scholar

[7] FREMLIN, D. H.: Measure Theory, Topological Measure Spaces, Vol. 4, Torres Fremlin, Colchester, 2003.Search in Google Scholar

[8] GRANDELL, J.: Mixed Poisson Processes, Chapman & Hall, 1997.10.1007/978-1-4899-3117-7Search in Google Scholar

[9] HEWITT, E.—SAVAGE, L. J.: Symmetric measures on Cartesian products, Trans. Amer. Math. Soc. 80 (1955), 470–501.10.1090/S0002-9947-1955-0076206-8Search in Google Scholar

[10] HOFFMANN-JØRGENSEN, J.: The Theory of Analytic Spaces. Various Publications Series, No. 10, Aarhus: Aarhus Universitet, Matematisk, 1970.Search in Google Scholar

[11] HUANG, W. J.: On the characterization of point processes with the exchangeable and Markov properties, Sankhyā A 52 (1990), 16–27.Search in Google Scholar

[12] KALLENBERG, O.: Probabilistic Symmetries and Invariance Principles, Springer, 2005.Search in Google Scholar

[13] LYBEROPOULOS, D. P.—MACHERAS, N. D.: Some characterizations of mixed Poisson processes, Sankhyā A 74 (2012), 57–79.10.1007/s13171-012-0011-ySearch in Google Scholar

[14] LYBEROPOULOS, D. P.—MACHERAS, N. D.: A construction of mixed Poisson processes via disintegrations, Math. Slovaca 63(1) (2013), 167–182.10.2478/s12175-012-0090-1Search in Google Scholar

[15] LYBEROPOULOS, D. P.—MACHERAS, N. D.: A characterization of martingale-equivalent mixed compound Poisson processes, Ann. Appl. Probab. 31(2) (2021), 778–805.10.1214/20-AAP1604Search in Google Scholar

[16] LYBEROPOULOS, D. P.—MACHERAS, N. D.—TZANINIS, S. M.: On the equivalence of various definitions of mixed Poisson processes, Math. Slovaca 69(2) (2019), 453–468.10.1515/ms-2017-0238Search in Google Scholar

[17] MACHERAS, N. D.—TZANINIS, S. M.: Some characterizations for Markov processes as mixed renewal processes, Math. Slovaca 68(6) (2018), 1477–1494.10.1515/ms-2017-0196Search in Google Scholar

[18] [18] OLSHEN, R.: A note on exchangeable sequences, Z. Wahrscheinlichkeitstheorie verw. Geb. 28 (1974), 317–321.10.1007/BF00532949Search in Google Scholar

[19] PACHL, J. K.: Disintegration and compact measures, Math. Scand. 43 (1978), 157–168.10.7146/math.scand.a-11771Search in Google Scholar

[20] PROTTER, PH. E.: Stochastic Integration and Differential Equations, 2nd ed., Springer, 2004.10.1007/978-3-662-10061-5Search in Google Scholar

[21] SCHMIDT, K. D.: Lectures on Risk Theory, B. G. Teubner, Stuttgart, 1996.10.1007/978-3-322-90570-3Search in Google Scholar

[22] SEGERDAHL, C.-O.: Stochastic processes and practical working models or why is the Pólya process approach defective in modern practice and how to cope with its deficiencies? Scand. Actuar. J.1970 (1970), 146–166.10.1080/03461238.1970.10405661Search in Google Scholar

[23] STRAUSS, W.—MACHERAS, N. D.—MUSIAL, K.: Splitting of liftings in products of probability spaces, Ann. Probab. 32 (2004), 2389–2408.10.1214/009117904000000018Search in Google Scholar

[24] TZANINIS, S. M.—MACHERAS, N. D.: A characterization of equivalent martingale probability measures in a mixed renewal risk model with applications in Risk Theory, https://arxiv.org/abs/2007.09051.Search in Google Scholar

Received: 2020-06-05
Accepted: 2021-01-11
Published Online: 2022-02-16
Published in Print: 2022-02-16

© 2022 Mathematical Institute Slovak Academy of Sciences

This work is licensed under the Creative Commons Attribution 4.0 International License.

Downloaded on 14.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ms-2022-0014/html
Scroll to top button