• Aucun résultat trouvé

On the scaling limits of Galton–Watson processes in varying environments

N/A
N/A
Protected

Academic year: 2021

Partager "On the scaling limits of Galton–Watson processes in varying environments"

Copied!
38
0
0

Texte intégral

(1)

HAL Id: hal-01890532

https://hal.archives-ouvertes.fr/hal-01890532

Submitted on 8 Oct 2018

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

varying environments

Vincent Bansaye, Florian Simatos

To cite this version:

Vincent Bansaye, Florian Simatos. On the scaling limits of Galton–Watson processes in varying environments. Electronic Journal of Probability, Institute of Mathematical Statistics (IMS), 2015, 20 (75), pp.1-36. �10.1214/EJP.v20-3812�. �hal-01890532�

(2)

a publisher's https://oatao.univ-toulouse.fr/20909

https://doi.org/10.1214/EJP.v20-3812

Bansaye, Vincent and Simatos, Florian On the scaling limits of Galton–Watson processes in varying environments.

(2015) Electronic Journal of Probability, 20 (75). 1-36.

(3)

El e c t ro nic J

o f

Pr

ob a bi l i t y

Electron. J. Probab.20(2015), no. 75, 1–36.

ISSN:1083-6489 DOI:10.1214/EJP.v20-3812

On the scaling limits of Galton–Watson processes in varying environments

Vincent Bansaye* Florian Simatos

Abstract

We establish a general sufficient condition for a sequence of Galton–Watson branching processes in varying environments to converge weakly. This condition extends previ- ous results by allowing offspring distributions to have infinite variance.

Our assumptions are stated in terms of pointwise convergence of a triplet of two real- valued functions and a measure. The limiting process is characterized by a backwards integro-differential equation satisfied by its Laplace exponent, which generalizes the branching equation satisfied by continuous state branching processes. Several examples are discussed, namely branching processes in random environment, Feller diffusion in varying environments and branching processes with catastrophes.

Keywords:Galton–Watson branching processes ; scaling limits ; varying environments.

AMS MSC 2010:60J80; 60F17; 60K37.

Submitted to EJP on September 19, 2014, final version accepted on July 1, 2015.

SupersedesarXiv:1112.2547.

1 Introduction

For each n 1, consider a sequence of offspring distributions (qi,n, i 0), the environments, and the corresponding Galton–Watson processZn= (Zi,n, i0)where individuals of the i-th generation reproduce according toqi,n. We are interested in the weak convergence of the sequence(Xn, n 1) of scaled processes of the form Xn(t) =n−1Zγn(t),nfor some sequence of time-changesγn.

In the Galton–Watson case whereqi,n=q0,n, this problem has been first considered by Feller [19] and Kolmogorov [33] and later by Lindvall [39] and Lamperti [37, 38]. It was exhaustively solved by Grimvall [24] who provided a necessary and sufficient condition for the convergence of the scaled processes(Xn, n 1)in terms of the sequence of offspring distributions(q0,n, n1). Our approach is inspired by these works and relies on the convergence of the Laplace transform.

In the Galton–Watson case, the possible limit processes are called continuous state branching processes (CSBP) and were first considered by Jiˇrina [29]. This class of

*Ecole Polytechnique, France. E-mail:vincent.bansaye@polytechnique.edu

Institut Supérieur de l’Aéronautique et de l’Espace (ISAE-SUPAERO), Université de Toulouse, France.

E-mail:florian.simatos@isae.fr

(4)

processes is well understood thanks to a random time-change transformation exhib- ited by Lamperti [36]. This transformation also allows for an elegant and conceptual proof of Grimvall’s result, see for instance Ethier and Kurtz [18, Chapter9] or, in the continuous-time setting, Helland [26], but this approach breaks down in the case of varying environment.

In this case, no such exhaustive result is available. Kurtz [35] and Borovkov [13]

proved general results in the finite variance case, to which our main result will be compared in details in Section 2.4. The two main points are that: (1) on the upside, we extend these results to the case of offspring distributions with (possibly) infinite variance;

(2) on the downside, we assume that a certain function has locally finite variation: as will be seen, this finite variation assumption is intrinsic to our approach.

When offspring distributions vary but have finite variance, the authors in [13, 35]

express their limit process as a simple transformation of Feller diffusion, the only CSBP with continuous sample paths. Kurtz [35] for instance uses semigroup techniques developed in [34]. However, these techniques become significantly more demanding in the infinite variance case considered here, where one needs to consider diffusion processes with jumps.

For this reason, we use in this paper a variation of the approach developed in [19, 33, 37, 38, 39]: namely, our main object of investigation is the Laplace exponentun

of Xn, defined by un(s, t, λ) = logE(exp(−λXn(t)) | Xn(s) = 1)for 0 s t and λ0. We identify a tripletn, βn, νn)withαnandβn two real-valued càdlàg functions, respectively of bounded variation and non-decreasing, andνn a σ-finite measure on (0,∞)2, such that informal calculation suggests the approximation

un(s, t, λ)λ+ Z

(s,t]

un(y, t, λ)αn(dy) Z

(s,t]

(un(y, t, λ))2βn(dy) +

Z

(0,∞)×(s,t]

h(x, un(y, t, λ))νn(dx dy) (1.1) withh(x, λ) = 1e−λx1+xλx2 +2(1+x(λx)22) (see Section 3.2). Motivated by this observation, we identify a mild notion of convergencen, βn, νn) (α, β, ν)(see Assumption 2.1 below) under which(Xn)converges weakly. Its weak limit is then characterized by its Laplace exponent, which is shown to be the unique solution to the integro-differential equation obtained by letting n → ∞ in (1.1). As alluded to above, in order for this equation to make sense in the limit, we need to assume that α has finite variation:

otherwise, it is not clear how to make sense of the integral with respect toαin (1.1).

Finally, note that this equation generalizes the branching equation for the Laplace exponent of CSBP obtained by Silverstein [41], see also Caballero et al. [14] for a recent and complete treatment.

In this paper, we assume the convergencen, βn, νn) (α, β, ν), where the mea- sureν is allowed to be non-zero and satisfies classical 1x2 finite moment. In the Galton–Watson case, our assumption is equivalent to Grimvall’s necessary and sufficient condition [24]. But when environments are allowed to vary over time, we need to avoid times where the process goes to zero instantaneously and almost surely, which will be calledbottleneck. Indeed, in the case of varying environments one must in general allow for non-critical distributions, and in particular subcritical ones which may cause such bottlenecks. Within our approach, we cannot determine in general the behavior of the process at the time of a bottleneck for reasons discussed in details in Section 2.3, where an example of indetermination∞ ×0is exhibited. Consequently, our main result studies the process(Xn)on some time interval[℘(t), t]where0℘(t)tintuitively corresponds to the last bottleneck before timet. Moreover, we show in Section 2.3 that, within our

(5)

assumptionn, βn, νn)(α, β, ν), a bottleneck can only occur if there is an offspring distribution with mean close to 0: otherwise, we have ℘(t) = 0, see Proposition 2.3.

Finally, we study in Proposition 2.4 the behavior of the process at the time of a bottleneck in the non-explosive case, where an indetermination∞ ×0is proscribed.

Let us now mention some closely related results. Galton–Watson processes inrandom environment were first introduced and studied in Smith and Wilkinson [42] in the case where the sequence (qi,n, i 0) is i.i.d., and in Athreya and Karlin [4, 3] when this sequence is stationary. These models have recently attracted considerable interest in the literature, see for instance [1, 2, 5, 10, 11, 12, 21, 25] for results on the long- time behavior in the critical and subcritical regimes and on large deviation. Scaling limits in the finite variance case were conjectured by Keiding [31] who introduced Feller diffusion in random environment. This conjecture was proved by Kurtz [35] and Helland [27]. In the same way, our results describe the weak convergence of scaled processes conditionally on the environments (quenched results), when the offspring distributions may have infinite variance. We describe the probabilistic structure of this process in Section 2.5.1 and we shed light on the correct scaling of such processes.

Our results are also related to some results on superprocesses. More precisely, our limit processes are closely related to the mass of superprocesses considered in El Karoui and Roelly [17]. These superprocesses are obtained in Dynkin [15, 16] as the limit of suitable branching particle systems, under some additional assumptions, e.g.

finite first moment (conservative case) and no drift. In these works the emphasis is on the limiting superprocesses themselves. As such, Dynkin [15, 16] considers branching particle systems evolving in continuous time, which, in order to establish limit theorems, are technically more convenient than the discrete time setting, which is our motivation here.

Organization of the paper

Theorem 2.2 is the main result of the paper, and is presented in Section 2.2. We compare it with earlier results in Section 2.4 and discuss some applications in Section 2.5, namely to Galton–Watson processes in random environment, to Feller diffusion in varying environments and to CSBP with catastrophes. Section 3 introduces notation, as well as some preliminary results. Theorem 2.2 is proved in Section 4, with some technical proofs deferred to Appendices B and C. Further results that complement Theorem 2.2 are proved in Section 5, and Appendix A is devoted to checking that the assumptions of Theorem 2.2 are necessary and sufficient in the Galton–Watson case.

2 Notation and results

2.1 General notation

In the rest of the paper, if a function g defined on [0,∞) is càdlàg, we write

∆g(t) = g(t)g(t−) for the value of the jump of g at time t. If g is in addition of locally finite variation, we writekgk(t)for the total variation ofgon[0, t]andR

f dgfor the Lebesgue-Stieltjes integral of a measurable functionf; note that|R

f dg| ≤R

|f|dkgk. Moreover, we say that a functionf is increasing iff(x)f(y)for everyxy.

For each n 1, we consider a Galton–Watson process in varying environments Zn= (Zi,n, i0). We denote byqi,nthe offspring distribution in generationiandξi,na random variable distributed according toqi,n, so that we can constructZn according to

(6)

the following recursion:

Zi+1,n=

Zi,n

X

k=1

ξi,n(k), i0,

where the random variables i,n(k), i, k 0)are independent andξi,n(k)is equal in distribution toξi,n. In order to find an interesting scaling of the sequence of processes (Zn, n1), the space scale is equal tonwhile the time scale is allowed to vary over time. More precisely, forn1, we consider an increasing, càdlàg and onto function γn : [0,∞) N(here and elsewhere, N ={0,1, . . .} denotes the set of non-negative integers) and we define the scaled process(Xn(t), t0)as follows:

Xn(t) = 1

nZγn(t),n, t0.

Fori 0 andn 1, we define tni = inf{t 0 : γn(t) =i} so thatγn(tni) = iand tnγ

n(t) t < tnγ

n(t)+1. Since Zn satisfies the branching property, i.e.,Zn started from Z0,n=zis stochastically equivalent to the sum ofzi.i.d. processes distributed according toZnstarted fromZ0,n= 1, we obtain after scaling

E[exp (−λXn(t))|Xn(s) =x] = exp(−xun(s, t, λ)) (2.1) for all λ, x, s, t 0 with s t and x N/n and un(s, t, λ) 0 called the Laplace exponent. We will characterize the convergence ofXnthrough the convergence ofun. Our assumptions are relying on the convergence of the tripletn, βn, νn), whereαnand βn are real-valued functions andνnis a measure onR×[0,∞). This triplet is defined in terms of the normalized random variables

ξi,n= 1

ni,n1), i0, n1, in the following way. We introduce the numbers

αi,n=nE ξi,n 1 +ξ2i,n

!

and βi,n= 1

2nE ξ2i,n 1 +ξ2i,n

!

and the measures

νi,n([x,∞)) =nPi,nx), so that the triplet can now be defined fort0by

αn(t) =

γn(t)−1

X

i=0

αi,n= Z x

1 +x2νi,n(dx), βn(t) =

γn(t)−1

X

i=0

βi,n= 1 2

Z x2

1 +x2νi,n(dx) and

νn([x,∞)×(0, t]) =

γn(t)−1

X

i=0

νi,n([x,∞)) (x0).

Each time, we understand a sum of the formP−1

0 to be equal to0, so thatαn(0) = βn(0) = 0.

2.2 Main result

Before stating our main result, we first precisely state the assumptionn, βn, νn) (α, β, ν)alluded to in the introduction. We also give a definition of℘(t), to be thought of

the last bottleneck before timet(see the introduction).

(7)

Assumption 2.1.There exist a càdlàg function of locally finite variationα, an increasing càdlàg function β, and a positive measureν on (0,∞)2, such that the two following conditions hold:

(A1) For everyt0and everyx >0such thatν({x} ×(0, t]) = 0, αn(t) −→

n→∞α(t),nk(t)−→

n→∞kαk(t), βn(t)−→

n→∞β(t)

and νn([x,∞)×(0, t]) −→

n→∞ν([x,∞)×(0, t]).

(A2) For everyt such that∆α(t) 6= 0, ∆β(t)6= 0orν((0,∞)× {t})6= 0 and for every x >0such thatν({x} × {t}) = 0,

αγn(t),n −→

n→∞∆α(t), βγn(t),n −→

n→∞∆β(t)andνγn(t),n([x,∞)) −→

n→∞ν([x,∞)× {t}).

The following definition of℘(t)is the most technically convenient and general at this point:

℘(t) = sup

st: lim

ε→0lim inf

n→∞ inf

s≤y≤tP Xn(t)> ε

Xn(y) = 1

= 0

(2.2) with the conventionsup= 0. A more intuitive definition will be given in Lemma 3.4, while a sufficient condition for℘(t) = 0under Assumption 2.1 is given in Proposition 2.3.

Theorem 2.2(Behavior on[℘(t), t]).Assume that Assumption 2.1 holds, and letα,βand νthe functions and measure defined there. Then, the following properties hold.

I. For everyt0, we have∆α(t)≥ −1andR

(0,∞)×(0,t](1x2)ν(dx dy)<. Moreover, the following functionβeis continuous and increasing:

βe(t) =β(t) Z

(0,∞)×(0,t]

x2

2(1 +x2)ν(dx dy), t0.

II. For everyt, λ >0ands[℘(t), t], there existsu(s, t, λ)(0,∞)such that for every s00,

n→∞lim sup

0≤s≤s0

|un(s, t, λ)u(s, t, λ)|= 0.

Moreover, the functionut,λ :s[℘(t), t]7→u(s, t, λ)is the unique càdlàg function that satisfiesinfs≤y≤tut,λ(y)>0for every℘(t)< stand

ut,λ(s) =λ+ Z

(s,t]

ut,λ(y)α(dy) Z

(s,t]

ut,λ(y)2β(dy)e +

Z

(0,∞)×(s,t]

1e−xut,λ(y)xut,λ(y) 1 +x2

ν(dx dy) (2.3) for every℘(t)st.

III. Fixt0,s[℘(t), t]andx0. Then for every sequence of initial states(xn)with xn x, everyIN− {0}, everyst1<· · ·< tI tand everyλ1, . . . , λI >0,

n→∞lim E[exp (−λ1Xn(t1)− · · · −λIXn(tI))|Xn(s) =xn]

= exp

xu

s, t1, λ1+u t1, t2, λ2+u(· · · , u(tI−1, tI, λI)· · ·)

. (2.4)

(8)

IV. Fixt0,s[℘(t), t]andx0. Then for every sequence of initial states(xn)with xn x, the sequence of processes(Xn(y), sy t)underP(· |Xn(s) =xn)is tight on the space D([s, t],[0,∞])of càdlàg functions f : [s, t] [0,∞]endowed with theJ1topology, where the space[0,∞]is equipped with the metricd(x, y) =

|e−xe−y|. In particular, weak convergence holds in view of(2.4).

In claim IV we considerXnas a process with range[0,∞]. AlthoughXn for fixedn cannot explode, for technical reasons we need to specify its behavior started at: in the sequel we assume thatis an absorbing state, so that ifXn(s) =for somes, then Xn(t) =for allts. The proof of claim IV will actually show that(Xn(y), syt) underP(· |Xn(s) =xn)is tight for any0st, not only℘(t)st.

2.3 Around the bottleneck

We now discuss in more details the notion of bottleneck that we have introduced. As a first guess, we could expect that the process goes to0when going through the time of a bottleneck, which would mean thatun(s, t, λ)0ifs < ℘(t). We now consider an example which illustrates several things that can go wrong and justify our framework, in particular the definition of the bottleneck℘(t)and the fact that Theorem 2.1 is stated on [℘(t), t].

Consider a critical offspring distributionqandYn= (Yi,n, i0)the Galton–Watson process with offspring distribution q, started fromn individuals. Assume that qand Γn are such that the sequence(bYn)withYbn(t) =Yntc,n/nconverges weakly to a non- conservative CSBPYb. For eachn1, we defineZnbyZ0,n=nand (δk denotes the unit mass atkN):

qi,n=

q if0< i <Γn, δ1 ifΓn i <n, (1pn0+pnδ1 ifi= 2Γn,

q ifi >n

for some vanishing sequencepn [0,1]. Defining γn(t) =ntcand recallingXn(t) = Zγn(t),n/n, we see thatXn coincides (in distribution) withYn on[0,1), stays constant on[1,2), undergoes a highly subcritical offspring distribution with meanpn at time2, referred to as catastrophe, and then resumes evolving according toqafter time2. The catastrophe at time2is meant to correspond to a bottleneck, and indeed one can check that℘(t) = 0if t <2 and℘(t) = 2 ift 2. Moreover, since Xn shifted at time2is a rescaled Galton–Watson process, the discussion on Galton–Watson processes in the next section will show that Assumption 2.1 is satisfied.

Fix some t > 2, and let us now discuss the asymptotic behavior of un(s, t, λ) for s < ℘(t) = 2. First of all, althoughpn 0 suggests that the process gets extinct at time2, un(s, t, λ)may actually not converge to0. Indeed, sinceYbn converges weakly toYb andYb is not conservative, there existρ > 0 and a sequenceyn → ∞such that P(bYn(2−)yn)ρ. In particular, just before the catastropheXn is, with probability at leastρ, at least of the order ofyn. In this event, the catastrophe bringsXn to level pnynby the law of large numbers, which diverges ifpn1/yn, i.e., ifpnvanishes slowly enough. This argument could be made rigorous to show thatun(s, t, λ)does not go to0 fors <1. Secondly, even ifun(s, t, λ)converges the limit may depend ons: fors <1we have seen that the limit was>0, while for1s <2the limit is= 0. Finally,un(s, t, λ) may even fail to converge: to see this, one may for instance consider two sequences p(1)n andp(2)n withynp(1)n → ∞andynp(2)n 0,Xn(1)andXn(2)the two processes obtained by the above construction using p(1)n and p(2)n instead of pn, respectively, and finally intertwine them by consideringX2n=Xn(1)andX2n+1=Xn(2).

(9)

This example therefore shows that a wide variety of behavior can happen before the bottleneck. We now give a sufficient condition that ensures℘(t) = 0, i.e., that there is no bottleneck before timet. Intuitively, the following assumption ensures thatξi,nis not too close to0, which avoids the almost sure absorption in one generation. For instance, it prevents the catastrophe of the previous example at time2.

Proposition 2.3(No bottleneck).Lett >0. If for everyC >0 lim inf

n→∞

inf

0≤i≤γn(t)Ei,n;ξi,nCn)

>0, (2.5)

then℘(t) = 0.

The proof of this proposition and of the next one are deferred to Section 5. We conclude this discussion by giving a condition under which un(s, t, λ) 0 along a subsequence, for alls < ℘(t)andλ0. Then the process started at the bottleneck goes as expected to zero (along a subsequence) when going through the bottleneck. Note that the example given at the beginning of this section shows that this is not always the case. Roughly speaking, this condition means that the limiting process is conservative.

Proposition 2.4(No explosion).Fix somet >0. If the two sequences(kαnk(t), n1) andn(t), n1)are bounded and

A→∞lim sup

n≥1,0≤s≤y≤tP(Xn(y)A|Xn(s) = 1) = 0, (2.6) then there exists an increasing sequence of integersn(k)such that for alls < ℘(t)and λ0,un(k)(s, t, λ)0ask→ ∞.

Moreover, these assumptions are satisfied, i.e.,(kαnk(t), n1)andn(t), n1)are bounded and (2.6)holds, if the following first moment condition is satisfied:

sup

n≥1

n

γn(t)−1

X

i=0

E i,n|

<∞. (2.7)

2.4 Comparison with earlier work

In the Galton–Watson case where γn(t) =ntcfor some integer-valued sequence n)andqi,n = q0,n for every i0, necessary and sufficient conditions for the finite- dimensional and weak convergence of(Xn, n1)are known since Grimvall [24], where weak convergence in the spaceD([0,∞),[0,∞))is considered. The main condition there is the weak convergence of the sequence(Sn, n 1) where Sn is distributed as the sum ofn independent copies ofξ0,n. Indeed, Theorem1.4in Ethier and Kurtz [18]

shows that the weak convergence of (Sn) implies the weak convergence of (Xn) in D([0,∞),[0,∞]), and Grimvall [24] proved the converse provided(Xn)converges to a non-explosive process (and thus inD([0,∞),[0,∞)); Helland [26] proved similar results for continuous-time branching processes. The following result, proved in the Appendix A, therefore shows that our Assumption 2.1 is sharp in the Galton–Watson case. Note in particular that in the Galton–Watson case,q0,n must be near-critical which implies in view of Proposition 2.3 that℘(t) = 0.

Lemma 2.5.In the Galton–Watson case, Assumption 2.1 is equivalent to the weak convergence of(Sn).

In the case of Galton–Watson processes in varying environments, Kurtz [35] used semigroup techniques to study the case where offspring distributions have uniformly

(10)

bounded third moments, which was later weakened by Borovkov [13] to a2 +δmoment condition. There are three main differences between the assumptions made in [13, 35]

and our Assumption 2.1.

First, we do not need to assume uniformly bounded second moments, which make ap- pear new phenomena such as possible indetermination form at the bottleneck (discussed in Section 2.3) and issues related to the correct time scale of Galton–Watson processes in random environment (discussed in Section 2.5.1).

Second, as we already mentioned in the introduction, the function that in [13, 35]

essentially plays the role of ourαn is not assumed to have finite variation in [13, 35].

This finite variation assumption is natural in our approach: otherwise it is not clear what meaning should be given to the termR

(s,t]ut,λ(y)α(dy)in (2.3). An enticing approach would be to considerαwith finite quadratic variations, which would for instance make it possible to use a pathwise construction of Itô’s integral such as in Föllmer [20], see also Wong and Zakai [44].

Finally, the functions that in [13, 35] essentially play the role of our αn and βn

are assumed in [13, 35] to converge in theJ1topology, whereas here we only assume pointwise convergence.

2.5 Applications

We discuss in this section new results that stem from Theorem 2.2. We keep the discussion at a high level and reserve rigorous results for future work (with the exception of Proposition 2.7).

2.5.1 Scaling limits of Galton–Watson processes in random (i.i.d.) environ- ment

Consider for eachn1a random probability measureQn and a sequence(qi,n, i0) of i.i.d. random variables distributed asQn. Then the sequence((αi,n, βi,n, νi,n), i0) is an i.i.d. sequence ofR×[0,∞)× M-valued random variables, withMthe space of locally finite measures onRendowed with the vague topology. Because of the law of large numbers, it is natural to chooseγn linear int, i.e.,γn(t) =ntcfor some sequence Γn→ ∞. We now discuss conditions under which Assumption 2.1 holds.

We are interested in the convergence of the process Yn(t) = (αn(t), βn(t), Mn(t)), whereMn(t) =P

0≤i<γn(t)νi,ndefines a measure-valued process. Note that the process Yn has i.i.d. increments, and so we can use classical results on measure-valued processes and random walks, such as [28, Theorem VII.2.35] and [30, Theorem A2.4], to get an explicit condition for its convergence. Namely, a functionh:R3R3is called truncation function if it is continuous, bounded and satisfiesh(x) =xin a neighborhood of0(in the sequel, vectors are considered to be columns andv0 denotes the transposition). LetC+K be the set of non-negative continuous functions with compact support equipped with the uniform norm, andC ⊂ CK+ be a dense subset closed under addition. Forϕ∈ C, let ynϕ= (α0,n, β0,n,R

ϕdν0,n).

Condition 2.6.There exist a truncation function h,Fϕ a measure on R3 integrating 1∧ |x|2andbϕR3,cϕij 0such that for everyϕ∈ C,

ΓnE(h(yϕn)) −→

n→∞bϕ, Γn E

hi(ynϕ)hj(ynϕ)

E(hi(ynϕ))E(hj(yϕn)) −→

n→∞cϕij and ΓnE(g(yϕn))−→

n→∞

Z

g(x)Fϕ(dx). (2.8)

(11)

In the above, the second convergence holds for alli, j= 1,2,3and the last conver- gence holds for all bounded, continuous functionsgthat are equal to0in a neighborhood of0.

Assuming that this condition holds, it can be proved thatYn converges to the process Y(t) = (α(t), β(t), M(t))such that for every ϕ continuous with compact support, the process Yϕ = (α, β, Mϕ)with Mϕ = (R

ϕ(x)M(t)(dx), t 0)is the Lévy process with Lévy exponent

ψϕ(v) =iv0bϕ1

2v0cϕv+ Z

eivx1iv0h(x)

Fϕ(dx), vR3.

Further, using Skorohod’s embedding theorem, we can assume that the convergence Yn Y holds almost surely. By definition, there exists a sequence of increasing bijections n, n1) from[0,∞)to[0,∞)such thatsup0≤s≤tn(s)s| → 0for everyt 0, and such that assumptions (A1) (except for the convergence ofnk) and (A2) are satisfied forγn0 =γnλn (see, e.g., Proposition VI.2.1in Jacod and Shiryaev [28]).

Assuming now thatαis of finite variation,αbeing a Lévy process must be of the form α(t) = dαt+S+(t)S(t)wheredαRandS+andSare two independent pure-jump subordinators (see, e.g., Bertoin [8]). With this special structure, it is possible to prove thatnk(t)→ kαk(t)so that Assumption 2.1 is fully satisfied and all the conclusions of Theorem 2.2 hold. It would be interesting to delve deeper into the probabilistic structure of the process(α, β, ν), and to understand how it relates to the properties of the limiting processX such as the extinction probability or the speed of extinction. In the literature, only the case of Feller diffusion in random environment whereν = 0andαis a Brownian motion has begun to be looked at, see, e.g., Böinghoff and Hutzenthaler [12].

We conclude this section by commenting on a question that actually motivated us in the first place: given a sequence of Galton–Watson processes in random environment, how can we find the right scaling in time, i.e., the right sequencen)?

Let us focus on the simplest possible case where in each generation we choose at random among one of two possible offspring distributions, i.e., we can write Qn = p(1)n δq(1) +p(2)n δq(2) where p(j)n (0,1), p(1)n +p(2)n = 1 and q(1), q(2) are two offspring distributions. In this discussion, we will call a CSBP with characteristic(b, c, F)the CSBP whose branching mechanism is given by

ψ(λ) =λb1 22+

Z

(e−λx1λx1{x≤1})F(dx).

For eachj = 1,2 letZn(j) = (Zn(j)(i), i0)be a Galton–Watson process with offspring distributionq(j)and consider(j)n )a sequence such that(Xn(j), n1)converges weakly to the CSBP with characteristic(b(j), c(j), F(j)), whereXn(j)(t) =n−1Zn(j)(bΓ(j)n tc). If bothq(1)andq(2)have finite variance, then it is well-known that in order to scale the Galton–Watson process with offspring distributionq(i)when the space scale isn, one needs to speed up time withnalso, i.e.,Γ(1)n = Γ(2)n =n. Thus when “mixing” these two processes, it is natural to speed up the resulting process by the common time scale and thus take Γn = n. To our knowledge, only such cases have been considered in the literature so far. When offspring distributions have infinite variance however, the situation becomes more delicate. Indeed, if for instanceq(1)([x,∞))x−a asx→ ∞ for somea (1,2), then one needs to consider Γ(1)n =na−1. Thus there are now two

“natural” time scales, namelyΓ(1)n =na−1andΓ(2)n =n.

Note thatΓ(j)n is the number of generations needed so that the variation ofZn(j)may be of the order ofn. OverΓn generations, the law of large numbers implies thatq(j)

Références

Documents relatifs

As for many other algorithms dedicated to fast mul- tiplication, our method relies on multipoint evaluation and interpolation: this brings back the problem to univariate power

Les enfants ont besoin du soutien de leurs parents et de leur collectivité et sont en conséquence plus touchés par les déterminants de la santé.. Les six premières années de vie

An environmental scan completed by the National Collaborating Centre for the Determinants of Health (2008b) assessed the challenges faced by professionals supporting early child

[r]

Section 4, service provider tags are concatenated to the end of RDAP query object identifiers to unambiguously identify the authoritative server for processing an RDAP query..

Les renseignements se rapportent seulement au produit indiqué dans la présente et ne concernent pas son utilisation avec une autre matière ou dans un autre

The asymptotic behavior of the raking-ratio empirical process indexed by a class of functions is studied when the auxiliary information is given by estimates.. These estimates

Pour assurer la production de vapeur lorsque le navire est au port, le bord possède une chaudière auxiliaire appelée chaudière de mouillage que l’on peut coupler de différentes