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

Diffusion approximation of controlled branching processes using limit theorems for random step processes

Miguel González 111Department of Mathematics, Faculty of Sciences and Instituto de Computación Científica Avanzada, University of Extremadura, Badajoz, Spain. E-mail address: mvelasco@unex.es. ORCID: 0000-0001-7481-6561.    Pedro Martín-Chávez 11footnotemark: 1222 Department of Mathematics, Faculty of Sciences, University of Extremadura, Badajoz, Spain. E-mail address: pedromc@unex.es. ORCID: 0000-0001-5530-3138.    Inés del Puerto11footnotemark: 1  333Department of Mathematics, Faculty of Sciences and Instituto de Computación Científica Avanzada, University of Extremadura, Badajoz, Spain. E-mail address: idelpuerto@unex.es. ORCID: 0000-0002-1034-2480.
Abstract

A controlled branching process (CBP) is a modification of the standard Bienaymé-Galton-Watson process in which the number of progenitors in each generation is determined by a random mechanism. We consider a CBP starting from a random number of initial individuals. The main aim of this paper is to provide a Feller diffusion approximation for critical CBPs. A similar result by considering a fixed number of initial individuals by using operator semigroup convergence theorems has been previously proved in [14]. An alternative proof is now provided making use of limit theorems for random step processes.

Keywords: Controlled branching processes; Weak convergence theorem; Martingale Differences; Diffusion processes; Stochastic differential equation; random step processes.

1 Introduction

Let {Xn,j:n=0,1,;j=1,2,}\left\{X_{n,j}:n=0,1,\ldots;j=1,2,\ldots\right\} be a sequence of independent and identically distributed (i.i.d.), non-negative and integer-valued random variables defined on a probability space (Ω,,𝒫)(\Omega,\cal{F},P). Let also {ϕn(k):k=0,1,}\{\phi_{n}(k):k=0,1,\ldots\}, for n=0,1,n=0,1,\ldots, be a sequence of stochastic processes which consist of independent non-negative integer-valued random variables on (Ω,,P)(\Omega,\mathcal{F},P) with the same one-dimensional distributions. Furthermore, let us assume that {Xn,j:n=0,1,;j=1,2,}\{X_{n,j}:n=0,1,\ldots;j=1,2,\ldots\} and {ϕn(k):n=0,1,;k=0,1,}\{\phi_{n}(k):n=0,1,\ldots;k=0,1,\ldots\} are independent.

A controlled branching process (CBP) is defined recursively as

Zn=j=1ϕn1(Zn1)Xn1,j,n=1,2,,{}Z_{n}=\sum^{\phi_{n-1}(Z_{n-1})}_{j=1}X_{n-1,j},\qquad n=1,2,\ldots, (1)

where j=10\sum_{j=1}^{0} is defined as 0 and Z0Z_{0} is a non-negative, integer-valued, square-integrable random variable which is independent of {Xn,j:n=0,1,;j=1,2,}\left\{X_{n,j}:n=0,1,\ldots;j=1,2,\ldots\right\} and {ϕn(k):n=0,1,;k=0,1,}\{\phi_{n}(k):n=0,1,\ldots;k=0,1,\ldots\}.

Here, ZnZ_{n} denotes the size of the nn-th generation of a population and Xn1,jX_{n-1,j} is the offspring size of the jj-th individual in the (n1)(n-1)-th generation. We will assume that the mean m=E[X0,1]m=E[X_{0,1}] and variance σ2=Var[X0,1]\sigma^{2}=Var[X_{0,1}] are both finite.

The class of CBPs is a very general family of stochastic processes that collect as particular cases the simplest branching model, the standard Bienaymé–Galton–Watson (BGW) process, by considering ϕn(k)=k\phi_{n}(k)=k a.s. for each kk, or a branching processes with immigration, by setting ϕn(k)=k+In\phi_{n}(k)=k+I_{n}, where {In}n0\{I_{n}\}_{n\geq 0} are i.i.d. random variables (writing in this way the immigrants give rise to offspring at the same generation as their arrival and with the same probability law as X0,1X_{0,1}), among others. The monograph [7] provides an extensive description of its probabilistic theory.

The research of functional weak limit theorems for branching processes arises a lot of interest since many years ago. It was firstly formulated for a BGW process by [3] and proved by [10] and [12]. These results have been extended to another classes of branching processes. For instance, a wide literature exists around weak convergence results for branching processes with immigration (BPI) since the pioneer work by [15], see also [1] and references therein. In this paper we focus our attention on a weak convergence theorem for a critical CBP with a random initial number of individuals and assuming finite second order moment on the this initial value. A similar result was already established for a single CBP in [14], and for an array of CBPs in [6], by assuming fixed initial numbers of progenitors using infinitesimal generators results for their proofs. Inspired in the paper [1] on BPI we will use limit theorems for random step processes towards a diffusion process provided in [9] to obtain an alternative proof. The scheme of it follows similar steps to the ones in [1]. An important feature of a CBP is that the value of ZnZ_{n} conditioned on the knowledge of the previous generation, Zn1=kZ_{n-1}=k, is a random sum of random variables, namely j=1ϕn(k)Xn1,j\sum_{j=1}^{\phi_{n}(k)}X_{n-1,j}, instead of a non-random sum as in the case of a BPI. This leads to handle the proofs of each steps using conditioning arguments different from those used in [1].

Apart from this introduction, the paper is organized as follows. In Section 2 we provide the notation and some auxiliary results about the behaviour of the first and second moments of the process. Section 3 gathers the main theorem. For the ease of reading the paper, additional results are presented in the Appendix.

2 Notation and auxiliary results

We denote, for k=0,1,,k=0,1,\ldots,

ε(k)\displaystyle\varepsilon(k) =\displaystyle= E[ϕn(k)],\displaystyle E[\phi_{n}(k)],
ν2(k)\displaystyle\nu^{2}(k) =\displaystyle= Var[ϕn(k)],\displaystyle Var[\phi_{n}(k)],

and assume all finite. It is easy to obtain that for n=1,2,n=1,2,\ldots,

E[Zn|n1]\displaystyle E[Z_{n}|{\cal F}_{n-1}] =\displaystyle= mε(Zn1),\displaystyle m\varepsilon(Z_{n-1}), (2)
Var[Zn|n1]\displaystyle Var[Z_{n}|{\cal F}_{n-1}] =\displaystyle= σ2ε(Zn1)+m2ν2(Zn1),\displaystyle\sigma^{2}\varepsilon(Z_{n-1})+m^{2}\nu^{2}(Z_{n-1}), (3)

where n{\cal F}_{n} is the σ\sigma-algebra generated by the random variables Z0,Z1,,ZnZ_{0},Z_{1},\ldots,Z_{n}, n1n\geq 1 (see Proposition 3.5 in [7]).

We introduce the quantities

τm(k)=E[Zn+1Zn1|Zn=k]=mε(k)k1,k1.{}\tau_{m}(k)=E[Z_{n+1}Z^{-1}_{n}|Z_{n}=k]=m\varepsilon(k)k^{-1},\ k\geq 1. (4)

The quantity τm(k)\tau_{m}(k) represents a mean growth rate. Intuitively, it can be interpreted as an average offspring per individual for a generation of size kk.

Assuming that limkτm(k)=τm\lim_{k\to\infty}\tau_{m}(k)=\tau_{m} exists, the process can be classified as:

τm<1subcritical;τm=1critical;τm>1supercritical.\tau_{m}<1\quad\mbox{{subcritical}};\ \tau_{m}=1\quad\mbox{{critical}};\ \tau_{m}>1\quad\mbox{{supercritical}}.

We are interested in critical CBPs that satisfy the following hypotheses:

  • A1)

    τm(k)=1+k1α\tau_{m}(k)=1+k^{-1}\alpha, k>0\ k>0, α>0\alpha>0,

  • A2)

    ν2(k)=O(kβ)\nu^{2}(k)=O(k^{\beta}), β<1\beta<1, as kk\to\infty.

The behavior of critical CBPs was studied in [5]. Assuming that P(ϕ0(0)=0)=1P(\phi_{0}(0)=0)=1, i.e. 0 is an absorbing state, and it is verified P(X0,1=0)>0P(X_{0,1}=0)>0 or P(ϕ0(k)=0)>0P(\phi_{0}(k)=0)>0, k=1,2,k=1,2,\ldots, it was established that under A1) and A2), if α>σ2/(2m)\alpha>\sigma^{2}/(2m) and an assumption on conditional moments holds, then P(Zn)>0.P(Z_{n}\to\infty)>0. In the present paper we will consider critical CBPs, {Zn}n0\{Z_{n}\}_{n\geq 0}, satisfying the above conditions, but with a reflecting barrier at zero, namely, P(ϕn(0)>0)>0P(\phi_{n}(0)>0)>0. Thus {Zn}n0\{Z_{n}\}_{n\geq 0} will have a finite number of returns to the sate zero till the explosion to infinity, i.e. P(Zn)=1P(Z_{n}\to\infty)=1.

Notice that under A1), ε(k)=(k+α)m1\varepsilon(k)=(k+\alpha)m^{-1}, k1k\geq 1, and, for simplicity in the posterior calculations, we will also assume throughout the paper that ε(0)=αm1\varepsilon(0)=\alpha m^{-1}.

Remark 2.1.

The controlled branching process we are considering is such that migration may take place in the next generation no matter the size of the current generation (when there are no individuals in the populations only immigration is possible). BGW processes with immigration at 0 were considered firstly in [4] and [13].

In next result we calculate the first and second moments of a CBP which verifies A1) and A2).

Proposition 2.1.

Let {Zn}n0\{Z_{n}\}_{n\geq 0} be a CBP with E[Z02]<E[Z_{0}^{2}]<\infty and satisfying hypotheses A1) and A2). It is verified as kk\to\infty that

E[Zk]=O(k) and E[Zk2]=O(k2).E[Z_{k}]=O(k)\mbox{ and }E[Z_{k}^{2}]=O(k^{2}).

Proof. From (2) and A1) it follows that

E[Zk]=E[mε(Zk1)]=E[Z0]+kα,k0.E[Z_{k}]=E[m\varepsilon(Z_{k-1})]=E[Z_{0}]+k\alpha,\quad k\geq 0. (5)

Using (3) we have

E[Var[Zkk1]]\displaystyle E[Var[Z_{k}\mid\mathcal{F}_{k-1}]] =\displaystyle= m1σ2(E[Z0]+(k1)α)+m2E[ν2(Zk1)].\displaystyle m^{-1}\sigma^{2}(E[Z_{0}]+(k-1)\alpha)+m^{2}E[\nu^{2}(Z_{k-1})].

Now, from A2), we have that there exists C>0C>0 such that ν2(k)Ck\nu^{2}(k)\leq Ck for all k>0k>0, so that E[ν2(Zk1)]CE[Zn1]+ν2(0)E[\nu^{2}(Z_{k-1})]\leq CE[Z_{n-1}]+\nu^{2}(0). Consequently, letting M1=3max{m1σ2E[Z0],M_{1}=3\max\{m^{-1}\sigma^{2}E[Z_{0}],\ Cm2E[Z0],m2ν2(0)}Cm^{2}E[Z_{0}],\ m^{2}\nu^{2}(0)\} and M2=2max{m1σ2α,m2Cα}M_{2}=2\max\{m^{-1}\sigma^{2}\alpha,m^{2}C\alpha\}, we have

E[Var[Zkk1]]M1+M2(k1).E[Var[Z_{k}\mid\mathcal{F}_{k-1}]]\leq M_{1}+M_{2}(k-1). (6)

Hence,

Var[Zk]\displaystyle Var[Z_{k}] =\displaystyle= E[Var[Zkk1]]+Var[E[Zkk1]]\displaystyle E[Var[Z_{k}\mid\mathcal{F}_{k-1}]]+Var[E[Z_{k}\mid\mathcal{F}_{k-1}]]
\displaystyle\leq M1+M2(k1)+Var[Zk1]\displaystyle M_{1}+M_{2}(k-1)+Var[Z_{k-1}]
\displaystyle\leq kM1+21M2k(k1)+Var[Z0].\displaystyle kM_{1}+2^{-1}M_{2}k(k-1)+Var[Z_{0}].

The latter inequality proves that E[Zk2]=O(k2)E[Z_{k}^{2}]=O(k^{2}).

Next Lemma presents certain relationships among the random variables {Xn,j:n=0,1,;j=1,2,}\{X_{n,j}:\ n=0,1,\ldots;\ j=1,2,\ldots\} which can be easily verified.

Lemma 2.1.

Let {Xn,j:n=0,1,;j=1,2,}\left\{X_{n,j}:n=0,1,\ldots;j=1,2,\ldots\right\} be a sequence of i.i.d. random variables with mean m=E[X0,1]m=E[X_{0,1}] and σ2=Var[X0,1]\sigma^{2}=Var[X_{0,1}], assumed finite. Let denote Sk,l=j=1l(Xk1,jm)S_{k,l}=\sum_{j=1}^{l}(X_{k-1,j}-m), k=1,2,k=1,2,\ldots, l=1,l=1,\ldots, and for j=1,,lj=1,\ldots,l, S~kj(l)=jjl(Xk1,jm)\tilde{S}_{k}^{j}(l)=\sum_{j^{\prime}\not=j}^{l}(X_{k-1,j^{\prime}}-m) and M>0,MM>0,\ M\in\mathbb{R}. It is verified that

E[j=1l(Xk1,jm)2𝕀{|S~kj(l)|>M}]l2σ4M2E\left[\sum_{j=1}^{l}(X_{k-1,j}-m)^{2}\mathbb{I}_{\{|\tilde{S}_{k}^{j}(l)|>M\}}\right]\leq\frac{l^{2}\sigma^{4}}{M^{2}}

and

E[(j,j,jjl(Xk1,jm)(Xk1,jm))2]=2l(l1)σ4.E\left[\left(\sum_{j,j^{\prime},j\not=j^{\prime}}^{l}(X_{k-1,j}-m)(X_{k-1,j^{\prime}}-m)\right)^{2}\right]=2l(l-1)\sigma^{4}.

3 Main result

We introduce for each nn\in\mathbb{N}, a stochastic process Wn(t)=n1ZntW_{n}(t)=n^{-1}Z_{\lfloor nt\rfloor}, for t0t\geq 0, tt\in\mathbb{R}, \lfloor\cdot\rfloor denoting the integer part. It is easy to see that {Wn}n1\left\{W_{n}\right\}_{n\geq 1} is a sequence of random functions that take values in D[0,)[0,)D_{[0,\infty)}[0,\infty), which is the space of non-negative functions on [0,)[0,\infty) that are right continuous and have left limits. We also denote by Cc[0,)C_{c}^{\infty}[0,\infty) the space of infinitely differentiable functions on [0,)[0,\infty) which have a compact support. Throughout the paper “𝒟\stackrel{{\scriptstyle\mathcal{D}}}{{\to}}” denotes the convergence of random functions in the Skorokhod topology.

Theorem 3.1.

Let {Zn}n0\{Z_{n}\}_{n\geq 0} be a CBP with E[Z02]<E[Z_{0}^{2}]<\infty, satisfying hypotheses A1) and A2). Then, Wn𝒟WW_{n}\stackrel{{\scriptstyle\mathcal{D}}}{{\rightarrow}}W, as n,n\to\infty, being WW a non-negative diffusion process, with generator Tf(x)=αf(x)+12xσ2m1f′′(x)Tf(x)=\alpha f^{\prime}(x)+\frac{1}{2}x\sigma^{2}m^{-1}f^{\prime\prime}(x), for fCc[0,)f\in C^{\infty}_{c}[0,\infty). The process WW is the pathwise unique solution of the stochastic differential equation

dW(t)=αdt+σ2m1(W(t))+d𝒲(t),t0,\mathrm{d}W(t)=\alpha\mathrm{d}t+\sqrt{\sigma^{2}m^{-1}(W(t)){{}^{+}}}\mathrm{d}{\cal W}(t),\qquad t\geq 0, (7)

with initial value W(0)=0W(0)=0, denoting x+=max{x,0}x^{+}=\max\{x,0\}, xx\in\mathbb{R}, and where 𝒲{\cal W} is a standard Wiener process.

Remark 3.1.

Taking into account Theorem A2 in Appendix, the stochastic differential equation (SDE) (7) has a pathwise unique solution {X(t)(x)}t0\{X(t)^{(x)}\}_{t\geq 0} for all initial values X(0)(x)=xX(0)^{(x)}=x\in\mathbb{R}. Moreover if x0x\geq 0, then X(t)(x)0X(t)^{(x)}\geq 0 almost surely for all t0t\geq 0.

In order to prove Theorem 3.1, we will establish previously the weak convergence of random step processes defined from a martingale difference created from the CBP.

We introduce the following sequence of martingale differences {Mk}k1\{M_{k}\}_{k\geq 1} with respect the filtration {k}k0\{\mathcal{F}_{k}\}_{k\geq 0} as:

Mk=ZkE[Zkk1]=ZkZk1α,k1.M_{k}=Z_{k}-E[Z_{k}\mid\mathcal{F}_{k-1}]=Z_{k}-Z_{k-1}-\alpha,\ k\geq 1.

Consider the random step processes:

n(t)=1n(Z0+k=1ntMk)=1nZntntnα,t0,n.\mathcal{M}_{n}(t)=\frac{1}{n}\left(Z_{0}+\sum_{k=1}^{\lfloor nt\rfloor}M_{k}\right)=\frac{1}{n}Z_{\lfloor nt\rfloor}-\frac{\lfloor nt\rfloor}{n}\alpha,\quad t\geq 0,\quad n\in\mathbb{N}. (8)
Theorem 3.2.

Let {Zn}n0\{Z_{n}\}_{n\geq 0} be a CBP with E[Z02]<E[Z_{0}^{2}]<\infty, satisfying hypotheses A1) and A2). It is verified that

n𝒟, as n,\mathcal{M}_{n}\stackrel{{\scriptstyle\mathcal{D}}}{{\longrightarrow}}\mathcal{M},\quad\text{ as }n\rightarrow\infty,

where the limit process \mathcal{M} is the pathwise unique solution of

d(t)=m1σ2((t)+αt)+d𝒲(t),t0,with initial value (0)=0.\mathrm{d}\mathcal{M}(t)=\sqrt{m^{-1}\sigma^{2}(\mathcal{M}(t)+\alpha t)^{+}}\mathrm{d}\mathcal{W}(t),\quad t\geq 0,\ \mbox{with initial value }\mathcal{M}(0)=0. (9)

Proof As was done in [1], we prove the result by applying Theorem A1 in Appendix with 𝒰=,Un(k)=n1Mk,k\mathcal{U}=\mathcal{M},\ U_{n}(k)=n^{-1}M_{k},\ k\in\mathbb{N}, Un(0)=n1Z0U_{n}(0)=n^{-1}Z_{0}, n(k)=k,k0\mathcal{F}_{n}(k)=\mathcal{F}_{k},k\geq 0, where nn\in\mathbb{N} (yielding 𝒰n=n,n\mathcal{U}_{n}=\mathcal{M}_{n},n\in\mathbb{N}, as well), and with coefficient functions β:[0,)×\beta:[0,\infty)\times\mathbb{R}\rightarrow\mathbb{R} and γ:[0,)×\gamma:[0,\infty)\times\mathbb{R}\rightarrow\mathbb{R} given by

β(t,x)=0,γ(t,x)=m1σ2(x+αt)+,t0,x.\beta(t,x)=0,\quad\gamma(t,x)=\sqrt{m^{-1}\sigma^{2}\left(x+\alpha t\right)^{+}},\quad t\geq 0,\quad x\in\mathbb{R}.

Firstly, we check that the SDE (9) has a pathwise unique strong solution {(t)(x)}t0\left\{\mathcal{M}(t)^{(x)}\right\}_{t\geq 0} for all initial values (0)(x)=x\mathcal{M}(0)^{(x)}=x\in\mathbb{R}. In fact, notice that if {(t)(x)}t0\left\{\mathcal{M}(t)^{(x)}\right\}_{t\geq 0} is a strong solution of the SDE (9) with initial value (0)(x)=x\mathcal{M}(0)^{(x)}=x\in\mathbb{R}, then, by Itô’s formula, the process 𝒫(t)=(t)(x)+αt,\mathcal{P}(t)=\mathcal{M}(t)^{(x)}+\alpha t, t0t\geq 0, is a solution of the SDE

d𝒫(t)=αdt+m1σ2𝒫(t)+d𝒲(t),t0,with initial value 𝒫(0)=x.\mathrm{d}\mathcal{P}(t)=\alpha\mathrm{d}t+\sqrt{m^{-1}\sigma^{2}\mathcal{P}(t)^{+}}\mathrm{d}\mathcal{W}(t),\quad t\geq 0,\ \mbox{with initial value }\mathcal{P}(0)=x. (10)

Conversely, if {𝒫(t)(x)}t0\{\mathcal{P}(t)^{(x)}\}_{t\geq 0} is a strong solution of the SDE (10) with initial value 𝒫(x)(0)=x\mathcal{P}^{(x)}(0)=x\in\mathbb{R}, then, by Itô’s formula, the process (t)=𝒫(t)(x)αt,\mathcal{M}(t)=\mathcal{P}(t)^{(x)}-\alpha t, t0t\geq 0, is a strong solution of the SDE (9) with initial value (0)=x\mathcal{M}(0)=x. Notice that SDE (10) is the same as SDE (7), consequently the SDE (10) and therefore the SDE (9) as well admit a pathwise unique strong solution with arbitrary initial value, and

{(t)+αt}t0=𝒟{W(t)}t0.\{\mathcal{M}(t)+\alpha t\}_{t\geq 0}\stackrel{{\scriptstyle\mathcal{D}}}{{=}}\{W(t)\}_{t\geq 0}. (11)

Let us see that E[(Un(k))2]<E\left[\left(U_{n}(k)\right)^{2}\right]<\infty for all n=1,2n=1,2\ldots and k=0,1,2,k=0,1,2,\ldots. Indeed, taking into account (6) in Proposition 2.1,

E[(Un(k))2]=n2E[Mk2]=E[Var[Zkk1]]M1+M2(k1)n2<E\left[\left(U_{n}(k)\right)^{2}\right]=n^{-2}E\left[M_{k}^{2}\right]=E[Var[Z_{k}\mid\mathcal{F}_{k-1}]]\leq\frac{M_{1}+M_{2}(k-1)}{n^{2}}<\infty (12)

and, by the assumption in the statement of the theorem, E[(Un(0))2]=n2E[Z02]<,E\left[\left(U_{n}(0)\right)^{2}\right]=n^{-2}E\left[Z_{0}^{2}\right]<\infty, for n=1,2,n=1,2,\ldots Moreover, Un(0)=n1Z0 a.s. 0U_{n}(0)=n^{-1}Z_{0}\stackrel{{\scriptstyle\text{ a.s. }}}{{\longrightarrow}}0 as nn\rightarrow\infty, especially Un(0)𝒟0U_{n}(0)\stackrel{{\scriptstyle\mathcal{D}}}{{\longrightarrow}}0 as nn\rightarrow\infty.

For conditions (i), (ii) and (iii) of Theorem A1 in Appendix, we have to check that for each T>0T>0, TT\in\mathbb{R}, as nn\to\infty:

  • a)

    supt[0,T]|1nk=1ntE[Mkk1]0|P0.\sup_{t\in[0,T]}\left|\frac{1}{n}\sum_{k=1}^{\lfloor nt\rfloor}E\left[M_{k}\mid\mathcal{F}_{k-1}\right]-0\right|\stackrel{{\scriptstyle\mathrm{P}}}{{\longrightarrow}}0.

  • b)

    supt[0,T]|1n2k=1ntE[Mk2k1]0tσ2m(n(s)+αs)+ds|P0.\sup_{t\in[0,T]}\left|\frac{1}{n^{2}}\sum_{k=1}^{\lfloor nt\rfloor}E\left[M_{k}^{2}\mid\mathcal{F}_{k-1}\right]-\int_{0}^{t}\frac{\sigma^{2}}{m}\left(\mathcal{M}_{n}(s)+\alpha s\right)^{+}\mathrm{d}s\right|\stackrel{{\scriptstyle\mathrm{P}}}{{\longrightarrow}}0.

  • c)

    For all θ>0,θ\theta>0,\theta\in\mathbb{R}, 1n2k=1nTE[Mk2𝕀{|Mk|>nθ}k1]P0.\frac{1}{n^{2}}\sum_{k=1}^{\lfloor nT\rfloor}E\left[M_{k}^{2}\mathbb{I}_{\left\{\left|M_{k}\right|>n\theta\right\}}\mid\mathcal{F}_{k-1}\right]\stackrel{{\scriptstyle\mathrm{P}}}{{\longrightarrow}}0.

Since E[Mkk1]=0E\left[M_{k}\mid\mathcal{F}_{k-1}\right]=0, nn\in\mathbb{N}, kk\in\mathbb{N}, a) holds.

Let us check b).

For each s>0,ss>0,s\in\mathbb{R} and nn\in\mathbb{N}, and for all t>0t>0, tt\in\mathbb{R} and nn\in\mathbb{N}, we have:

n(s)+αs=1nZns+nsnsnα,\mathcal{M}_{n}(s)+\alpha s=\frac{1}{n}Z_{\lfloor ns\rfloor}+\frac{ns-\lfloor ns\rfloor}{n}\alpha,

thus (n(s)+αs)+=n(s)+αs\left(\mathcal{M}_{n}(s)+\alpha s\right)^{+}=\mathcal{M}_{n}(s)+\alpha s. Now, we have, for all t>0t>0 and nn\in\mathbb{N},

0t(n(s)+αs)+ds\displaystyle\int_{0}^{t}\left(\mathcal{M}_{n}{(s)}+\alpha s\right)^{+}\mathrm{d}s =\displaystyle= 0t(1nZns+nsnsnα)ds\displaystyle\int_{0}^{t}\left(\frac{1}{n}Z_{\lfloor ns\rfloor}+\frac{ns-\lfloor ns\rfloor}{n}\alpha\right)\mathrm{d}s
=\displaystyle= k=0nt1k/n(k+1)/n(1nZk+nsknα)ds\displaystyle\sum_{k=0}^{\lfloor nt\rfloor-1}\int_{k/n}^{(k+1)/n}\left(\frac{1}{n}Z_{k}+\frac{ns-k}{n}\alpha\right)\mathrm{d}s
+nt/nt(1nZnt+nsntnα)ds\displaystyle+\int_{\lfloor nt\rfloor/n}^{t}\left(\frac{1}{n}Z_{\lfloor nt\rfloor}+\frac{ns-\lfloor nt\rfloor}{n}\alpha\right)\mathrm{d}s
=\displaystyle= 1n2k=0nt1Zk+ntntn2Znt+α2n2nt\displaystyle\frac{1}{n^{2}}\sum_{k=0}^{\lfloor nt\rfloor-1}Z_{k}+\frac{nt-\lfloor nt\rfloor}{n^{2}}Z_{\lfloor nt\rfloor}+\frac{\alpha}{2n^{2}}\lfloor nt\rfloor
+αn(n2(t2nt2n2)nt(tntn))\displaystyle+\frac{\alpha}{n}\left(\frac{n}{2}\left(t^{2}-\frac{\lfloor nt\rfloor^{2}}{n^{2}}\right)-\lfloor nt\rfloor\left(t-\frac{\lfloor nt\rfloor}{n}\right)\right)
=\displaystyle= 1n2k=0nt1Zk+ntntn2Znt+nt+(ntnt)22n2α.\displaystyle\frac{1}{n^{2}}\sum_{k=0}^{\lfloor nt\rfloor-1}Z_{k}+\frac{nt-\lfloor nt\rfloor}{n^{2}}Z_{\lfloor nt\rfloor}+\frac{\lfloor nt\rfloor+(nt-\lfloor nt\rfloor)^{2}}{2n^{2}}\alpha.

It is verified that, for t>0t>0 and nn\in\mathbb{N},

1n2k=1ntE[Mk2k1]\displaystyle\frac{1}{n^{2}}\sum_{k=1}^{\lfloor nt\rfloor}E\left[M_{k}^{2}\mid\mathcal{F}_{k-1}\right] =\displaystyle= 1n2k=1ntVar[Zkk1]\displaystyle\frac{1}{n^{2}}\sum_{k=1}^{\lfloor nt\rfloor}Var[Z_{k}\mid\mathcal{F}_{k-1}]
=\displaystyle= 1n2k=1nt(m2ν2(Zk1)+σ2m(Zk1+α))\displaystyle\frac{1}{n^{2}}\sum_{k=1}^{\lfloor nt\rfloor}\left(m^{2}\nu^{2}(Z_{k-1})+\frac{\sigma^{2}}{m}(Z_{k-1}+\alpha)\right)
=\displaystyle= m2n2k=1ntν2(Zk1)+ntασ2n2m+σ2n2mk=1ntZk1.\displaystyle\frac{m^{2}}{n^{2}}\sum_{k=1}^{\lfloor nt\rfloor}\nu^{2}(Z_{k-1})+\frac{\lfloor nt\rfloor\alpha\sigma^{2}}{n^{2}m}+\frac{\sigma^{2}}{n^{2}m}\sum_{k=1}^{\lfloor nt\rfloor}Z_{k-1}.

Consequently,

1n2k=1ntE[Mk2k1]\displaystyle\frac{1}{n^{2}}\sum_{k=1}^{\lfloor nt\rfloor}E\left[M_{k}^{2}\mid\mathcal{F}_{k-1}\right] \displaystyle- 0tσ2m(n(s)+αs)+ds=m2n2k=1ntν2(Zk1)+ntασ2n2m\displaystyle\int_{0}^{t}\frac{\sigma^{2}}{m}\left(\mathcal{M}_{n}(s)+\alpha s\right)^{+}\mathrm{d}s=\frac{m^{2}}{n^{2}}\sum_{k=1}^{\lfloor nt\rfloor}\nu^{2}(Z_{k-1})+\frac{\lfloor nt\rfloor\alpha\sigma^{2}}{n^{2}m}
σ2(ntnt)mn2Zntσ2mnt+(ntnt)22n2α.\displaystyle-\frac{\sigma^{2}(nt-\lfloor nt\rfloor)}{mn^{2}}Z_{\lfloor nt\rfloor}-\frac{\sigma^{2}}{m}\frac{\lfloor nt\rfloor+(nt-\lfloor nt\rfloor)^{2}}{2n^{2}}\alpha.

Since for each T>0,TT>0,T\in\mathbb{R},

supt[0,T]ntn2Tn0, as nsupt[0,T]nt+(ntnt)22n2T2n+12n20, as n,\begin{array}[]{l}\sup_{t\in[0,T]}\frac{\lfloor nt\rfloor}{n^{2}}\leqslant\frac{T}{n}\rightarrow 0,\quad\text{ as }n\rightarrow\infty\\[7.22743pt] \sup_{t\in[0,T]}\frac{\lfloor nt\rfloor+(nt-\lfloor nt\rfloor)^{2}}{2n^{2}}\leqslant\frac{T}{2n}+\frac{1}{2n^{2}}\rightarrow 0,\quad\text{ as }n\rightarrow\infty,\end{array}

in order to show b)b), it suffices to prove that for each T>0,TT>0,T\in\mathbb{R},

1n2supt[0,T]((ntnt)Znt)1n2supt[0,T]ZntP0 as n.\frac{1}{n^{2}}\sup_{t\in[0,T]}\left((nt-\lfloor nt\rfloor)Z_{\lfloor nt\rfloor}\right)\leq\frac{1}{n^{2}}\sup_{t\in[0,T]}Z_{\lfloor nt\rfloor}\stackrel{{\scriptstyle{P}}}{{\longrightarrow}}0\quad\mbox{ as }n\rightarrow\infty. (13)

and

m2n2supt[0,T]k=1ntν2(Zk1)P0 as n.\frac{m^{2}}{n^{2}}\sup_{t\in[0,T]}\sum_{k=1}^{\lfloor nt\rfloor}\nu^{2}(Z_{k-1})\stackrel{{\scriptstyle{P}}}{{\longrightarrow}}0\quad\mbox{ as }n\rightarrow\infty. (14)

First we check (13). For each kk\in\mathbb{N}, we have Zk=Zk1+Mk+αZ_{k}=Z_{k-1}+M_{k}+\alpha, thus

Zk=Z0+j=1kMj+kα,Z_{k}=Z_{0}+\sum_{j=1}^{k}M_{j}+k\alpha,

and hence, for each t>0,tt>0,\ t\in\mathbb{R} and nn\in\mathbb{N}, we get

Znt]=|Znt|Z0+j=1nt|Mj|+ntα.Z_{\lfloor nt]}=\left|Z_{\lfloor nt\rfloor}\right|\leqslant Z_{0}+\sum_{j=1}^{\lfloor nt\rfloor}\left|M_{j}\right|+\lfloor nt\rfloor\alpha.

Consequently, in order to prove (13)(\ref{eq:28}), it suffices to show

1n2supt[0,T]j=1nt|Mj|1n2j=1nT|Mj|P0, as n.\frac{1}{n^{2}}\sup_{t\in[0,T]}\sum_{j=1}^{\lfloor nt\rfloor}\left|M_{j}\right|\leqslant\frac{1}{n^{2}}\sum_{j=1}^{\lfloor nT\rfloor}\left|M_{j}\right|\stackrel{{\scriptstyle{P}}}{{\longrightarrow}}0,\quad\text{ as }n\rightarrow\infty.

By (12), E[Mk2]=O(k),E[M_{k}^{2}]=O(k), as kk\to\infty, and therefore by Jensen’s inequality, E[|Mk|]=O(k1/2),E[|M_{k}|]=O(k^{1/2}), as kk\to\infty, and hence

E[1n2j=1nT|Mj|]=1n2j=1nTO(j1/2)=O(n1/2)0, as n.E\left[\frac{1}{n^{2}}\sum_{j=1}^{\lfloor nT\rfloor}\left|M_{j}\right|\right]=\frac{1}{n^{2}}\sum_{j=1}^{\lfloor nT\rfloor}{O}\left(j^{1/2}\right)={O}\left(n^{-1/2}\right)\rightarrow 0,\quad\text{ as }n\rightarrow\infty.

Thus we obtain n2j=1nT|Mj|P0n^{-2}\sum_{j=1}^{\lfloor nT\rfloor}\left|M_{j}\right|\stackrel{{\scriptstyle{P}}}{{\longrightarrow}}0 as nn\rightarrow\infty implying (13)(\ref{eq:28}).

Now, taking into account hypothesis A2)

E[1n2k=1nTν2(Zk1)]=1n2k=1nTO(kβ)=O(nβ1),E\left[\frac{1}{n^{2}}\sum_{k=1}^{\lfloor nT\rfloor}\nu^{2}(Z_{k-1})\right]=\frac{1}{n^{2}}\sum_{k=1}^{\lfloor nT\rfloor}O(k^{\beta})=O(n^{\beta-1}), (15)

and hence (14).

Let us check c).

We write

Mk=j=1ϕk1(Zk1)(Xk1,jm)+m(ϕk1(Zk1)ε(Zk1)).M_{k}=\sum_{j=1}^{\phi_{k-1}(Z_{k-1})}(X_{k-1,j}-m)+m(\phi_{k-1}(Z_{k-1})-\varepsilon(Z_{k-1})).

Let denote Nk=j=1ϕk1(Zk1)(Xk1,jm)N_{k}=\sum_{j=1}^{\phi_{k-1}(Z_{k-1})}(X_{k-1,j}-m). It is verified for each n,kn,k\in\mathbb{N}, and θ>0,θ\theta>0,\theta\in\mathbb{R} that

Mk22(Nk2+m2(ϕk1(Zk1)ε(Zk1))2),M_{k}^{2}\leq 2\left(N_{k}^{2}+m^{2}(\phi_{k-1}(Z_{k-1})-\varepsilon(Z_{k-1}))^{2}\right),

and

𝕀{|Mk|>nθ}𝕀{|Nk|>nθ/2}+𝕀{|ϕk1(Zk1)ε(Zk1)|>nθ/2m}.\mathbb{I}_{\left\{\left|M_{k}\right|>n\theta\right\}}\leq\mathbb{I}_{\left\{\left|N_{k}\right|>n\theta/2\right\}}+\mathbb{I}_{\left\{\left|\phi_{k-1}(Z_{k-1})-\varepsilon(Z_{k-1})\right|>n\theta/2m\right\}}.

Hence

Mk2𝕀{|Mk|>nθ}2Nk2𝕀{|Nk|>nθ/2}+2Nk2𝕀{|ϕk1(Zk1)ε(Zk1)|>nθ/2m}+2m2(ϕk1(Zk1)ε(Zk1))2.M_{k}^{2}\mathbb{I}_{\left\{\left|M_{k}\right|>n\theta\right\}}\leq 2N_{k}^{2}\mathbb{I}_{\left\{\left|N_{k}\right|>n\theta/2\right\}}+2N_{k}^{2}\mathbb{I}_{\left\{\left|\phi_{k-1}(Z_{k-1})-\varepsilon(Z_{k-1})\right|>n\theta/2m\right\}}+2m^{2}(\phi_{k-1}(Z_{k-1})-\varepsilon(Z_{k-1}))^{2}.

In consequence, to check c) we will prove, as nn\to\infty,

  • c.1)

    1n2k=1nTE[Nk2𝕀{|Nk|>nθ}k1]P0\frac{1}{n^{2}}\sum_{k=1}^{\lfloor nT\rfloor}E\left[N_{k}^{2}\mathbb{I}_{\left\{\left|N_{k}\right|>n\theta\right\}}\mid\mathcal{F}_{k-1}\right]\stackrel{{\scriptstyle{P}}}{{\longrightarrow}}0\quad for all θ>0,θ\theta>0,\theta\in\mathbb{R}.

  • c.2)

    1n2k=1nTE[Nk2𝕀{|ϕk1(Zk1)ε(Zk1)|>nθ}k1]P0\frac{1}{n^{2}}\sum_{k=1}^{\lfloor nT\rfloor}E\left[N_{k}^{2}\mathbb{I}_{\left\{\left|\phi_{k-1}(Z_{k-1})-\varepsilon(Z_{k-1})\right|>n\theta\right\}}\mid\mathcal{F}_{k-1}\right]\stackrel{{\scriptstyle{P}}}{{\longrightarrow}}0 for all θ>0,θ\theta>0,\theta\in\mathbb{R}.

  • c.3)

    1n2k=1nTE[(ϕk1(Zk1)ε(Zk1))2k1]P0.\frac{1}{n^{2}}\sum_{k=1}^{\lfloor nT\rfloor}E\left[(\phi_{k-1}(Z_{k-1})-\varepsilon(Z_{k-1}))^{2}\mid\mathcal{F}_{k-1}\right]\stackrel{{\scriptstyle{P}}}{{\longrightarrow}}0.

In what follows let θ>0,θ\theta>0,\theta\in\mathbb{R} be fixed.

Let us see c.3). It is verified that

1n2k=1nTE[(ϕk1(Zk1)ε(Zk1))2k1]\displaystyle\frac{1}{n^{2}}\sum_{k=1}^{\lfloor nT\rfloor}E\left[(\phi_{k-1}(Z_{k-1})-\varepsilon(Z_{k-1}))^{2}\mid\mathcal{F}_{k-1}\right] =\displaystyle= 1n2k=1nTVar[ϕk1(Zk1)k1]\displaystyle\frac{1}{n^{2}}\sum_{k=1}^{\lfloor nT\rfloor}Var\left[\phi_{k-1}(Z_{k-1})\mid\mathcal{F}_{k-1}\right]
=\displaystyle= 1n2k=1nTν2(Zk1)P0, as n.\displaystyle\frac{1}{n^{2}}\sum_{k=1}^{\lfloor nT\rfloor}\nu^{2}(Z_{k-1})\stackrel{{\scriptstyle{P}}}{{\longrightarrow}}0,\mbox{ as }n\to\infty.

This latter was proved by considering (15).

Now, we check c.1). By the properties of conditional expectation with respect to a σ\sigma -algebra, we get for all n,kn,k\in\mathbb{N},

E[Nk2𝕀{|Nk|>nθ}k1]=Fn,k(Zk1),E\left[N_{k}^{2}\mathbb{I}_{\left\{\left|N_{k}\right|>n\theta\right\}}\mid\mathcal{F}_{k-1}\right]=F_{n,k}\left(Z_{k-1}\right),

where on {Zk1=z}\{Z_{k-1}=z\}, with z=0,1,z=0,1,\ldots

Fn,k(z)=E[Sk(z)2𝕀{|Sk(z)|>nθ}], where Sk(z)=j=1ϕk1(z)(Xk1,jm).F_{n,k}(z)=E\left[S_{k}(z)^{2}\mathbb{I}_{\left\{\left|S_{k}(z)\right|>n\theta\right\}}\right],\mbox{ where }S_{k}(z)=\sum_{j=1}^{\phi_{k-1}(z)}\left(X_{k-1,j}-m\right).

Consider the decomposition Fn,k(z)=An,k(z)+Bn,k(z)F_{n,k}(z)=A_{n,k}(z)+B_{n,k}(z) with

An,k(z)\displaystyle A_{n,k}(z) =\displaystyle= E[j=1ϕk1(z)(Xk1,jm)2𝕀{|Sk(z)|>nθ}],\displaystyle E\left[\sum_{j=1}^{\phi_{k-1}(z)}(X_{k-1,j}-m)^{2}\mathbb{I}_{\{|S_{k}(z)|>n\theta\}}\right],
Bn,k(z)\displaystyle B_{n,k}(z) =\displaystyle= E[j,j,jjϕk1(z)(Xk1,jm)(Xk1,jm)𝕀{|Sk(z)|>nθ}].\displaystyle E\left[\sum_{j,j^{\prime},j\not=j^{\prime}}^{\phi_{k-1}(z)}(X_{k-1,j}-m)(X_{k-1,j^{\prime}}-m)\mathbb{I}_{\{\left|S_{k}(z)\right|>n\theta\}}\right].

Now, let denote Sk,l=j=1l(Xk1,jm)S_{k,l}=\sum_{j=1}^{l}(X_{k-1,j}-m), k=1,2,k=1,2,\ldots, l=0,1,l=0,1,\ldots. It is verified the inequality, for j{1,,l}j\in\{1,\ldots,l\}

|Sk,l||Xk1,jm|+|S~kj(l)|, with S~kj(l)=jjl(Xk1,jm).|S_{k,l}|\leq|X_{k-1,j}-m|+|\tilde{S}_{k}^{j}(l)|,\mbox{ with }\tilde{S}_{k}^{j}(l)=\sum_{j^{\prime}\not=j}^{l}(X_{k-1,j^{\prime}}-m).

We have, using Lemma 2.1,

An,k(z)\displaystyle A_{n,k}(z) =\displaystyle= E[E[j=1ϕk1(z)(Xk1,jm)2𝕀{|Sk(z)|>nθ}|ϕk1(z)]]\displaystyle E\left[E\left[\left.\sum_{j=1}^{\phi_{k-1}(z)}(X_{k-1,j}-m)^{2}\mathbb{I}_{\{|S_{k}(z)|>n\theta\}}\right|\phi_{k-1}(z)\right]\right]
=\displaystyle= l=0E[j=1l(Xk1,jm)2𝕀{|Sk,l|>nθ}]P(ϕk1(z)=l)\displaystyle\sum_{l=0}^{\infty}E\left[\sum_{j=1}^{l}(X_{k-1,j}-m)^{2}\mathbb{I}_{\{|{S}_{k,l}|>n\theta\}}\right]P(\phi_{k-1}(z)=l)
\displaystyle\leq l=0j=1l(E[(Xk1,jm)2𝕀{|Xk1,jm|>nθ/2}]\displaystyle\sum_{l=0}^{\infty}\sum_{j=1}^{l}(E\left[(X_{k-1,j}-m)^{2}\mathbb{I}_{\{|X_{k-1,j}-m|>n\theta/2\}}\right]
+E[(Xk1,jm)2𝕀{|S~kj(z)|>nθ/2}])P(ϕk1(z)=l)\displaystyle+\left.E[(X_{k-1,j}-m)^{2}\mathbb{I}_{\{|\tilde{S}_{k}^{j}(z)|>n\theta/2\}}]\right)P(\phi_{k-1}(z)=l)
\displaystyle\leq l=0(lE[(X0,1m)2𝕀{|X0,1m|>nθ/2}]+4l2σ4n2θ2)P(ϕk1(z)=l)\displaystyle\sum_{l=0}^{\infty}\left(lE\left[(X_{0,1}-m)^{2}\mathbb{I}_{\{|X_{0,1}-m|>n\theta/2\}}\right]+\frac{4l^{2}\sigma^{4}}{n^{2}\theta^{2}}\right)P(\phi_{k-1}(z)=l)
=\displaystyle= ε(z)E[(X0,1m)2𝕀{|X0,1m|>nθ/2}]+4σ4n2θ2E[(ϕ0(z))2]\displaystyle\varepsilon(z)E\left[(X_{0,1}-m)^{2}\mathbb{I}_{\{|X_{0,1}-m|>n\theta/2\}}\right]+\frac{4\sigma^{4}}{n^{2}\theta^{2}}E[(\phi_{0}(z))^{2}]

Therefore

An,k(z)An,k(z)(1)+An,k(z)(2),A_{n,k}(z)\leq A_{n,k}(z)^{(1)}+A_{n,k}(z)^{(2)},

with

An,k(z)(1)\displaystyle A_{n,k}(z)^{(1)} =\displaystyle= ε(z)E[(X0,1m)2𝕀{|X0,1m|>nθ/2}],\displaystyle\varepsilon(z)E\left[(X_{0,1}-m)^{2}\mathbb{I}_{\{|X_{0,1}-m|>n\theta/2\}}\right],
An,k(z)(2)\displaystyle A_{n,k}(z)^{(2)} =\displaystyle= (ν2(z)+(ε(z))2)4σ4n2θ2.\displaystyle(\nu^{2}(z)+(\varepsilon(z))^{2})\frac{4\sigma^{4}}{n^{2}\theta^{2}}.

Using (5) in Proposition 2.1, it is verified that for nn\in\mathbb{N}

E[1n2k=1nTAn,k(Zk1)(1)]\displaystyle E\left[\frac{1}{n^{2}}\sum_{k=1}^{\lfloor nT\rfloor}A_{n,k}(Z_{k-1})^{(1)}\right] =\displaystyle= 1n2k=1nTE[ε(Zk)]E[(X0,1m)2𝕀{|X0,1m|>nθ/2}]\displaystyle\frac{1}{n^{2}}\sum_{k=1}^{\lfloor nT\rfloor}E[\varepsilon(Z_{k})]E\left[(X_{0,1}-m)^{2}\mathbb{I}_{\{|X_{0,1}-m|>n\theta/2\}}\right]
=\displaystyle= 1n2(k=1nTO(k))E[(X0,1m)2𝕀{|X0,1m|>nθ/2}]\displaystyle\frac{1}{n^{2}}\left(\sum_{k=1}^{\lfloor nT\rfloor}O(k)\right)E\left[(X_{0,1}-m)^{2}\mathbb{I}_{\{|X_{0,1}-m|>n\theta/2\}}\right]
=\displaystyle= O(1)E[(X0,1m)2𝕀{|X0,1m|>nθ/2}].\displaystyle O(1)E\left[(X_{0,1}-m)^{2}\mathbb{I}_{\{|X_{0,1}-m|>n\theta/2\}}\right].

By applying the dominated convergence theorem we have

E[1n2k=1nTAn,k(Zk1)(1)]0,as n.E\left[\frac{1}{n^{2}}\sum_{k=1}^{\lfloor nT\rfloor}A_{n,k}(Z_{k-1})^{(1)}\right]\to 0,\ \mbox{as }n\to\infty. (16)

It is also verified by using again (5) in Proposition 2.1 and A2) that

E[1n2k=1nTAn,k(Zk1)(2)]\displaystyle E\left[\frac{1}{n^{2}}\sum_{k=1}^{\lfloor nT\rfloor}A_{n,k}(Z_{k-1})^{(2)}\right] =\displaystyle= 1n2k=1nTE[(ν2(Zk1)+(ε(Zk1))2)4σ4n2θ2]\displaystyle\frac{1}{n^{2}}\sum_{k=1}^{\lfloor nT\rfloor}E\left[\left(\nu^{2}(Z_{k-1})+(\varepsilon(Z_{k-1}))^{2}\right)\frac{4\sigma^{4}}{n^{2}\theta^{2}}\right] (17)
=\displaystyle= 4σ4n4θ2k=1nTO(k2)=O(n1)0, as n.\displaystyle\frac{4\sigma^{4}}{n^{4}\theta^{2}}\sum_{k=1}^{\lfloor nT\rfloor}O(k^{2})=O(n^{-1})\to 0,\mbox{ as }n\to\infty.

Taking into account (16) and (17) we have that, as nn\to\infty,

1n2k=1nTAn,k(Zk1)P0.\frac{1}{n^{2}}\sum_{k=1}^{\lfloor nT\rfloor}A_{n,k}(Z_{k-1})\stackrel{{\scriptstyle{P}}}{{\longrightarrow}}0.

Let us now dealt with Bn,k(z)B_{n,k}(z). It is verified that, using Cauchy-Schwarz’s inequality:

Bn,k(z)\displaystyle B_{n,k}(z) =\displaystyle= l=0E[j,j,jjl(Xk1,jm)(Xk1,jm)𝕀{|Sk,l|>nθ}]P(ϕk1(z)=l)\displaystyle\sum_{l=0}^{\infty}E\left[\sum_{j,j^{\prime},j\not=j^{\prime}}^{l}(X_{k-1,j}-m)(X_{k-1,j^{\prime}}-m)\mathbb{I}_{\{|{S}_{k,l}|>n\theta\}}\right]P(\phi_{k-1}(z)=l)
\displaystyle\leq l=0E[(j,j,jjl(Xk1,jm)(Xk1,jm))2]E[𝕀{|Sk,l|>nθ}]P(ϕk1(z)=l).\displaystyle\sum_{l=0}^{\infty}\sqrt{E\left[\left(\sum_{j,j^{\prime},j\not=j^{\prime}}^{l}(X_{k-1,j}-m)(X_{k-1,j^{\prime}}-m)\right)^{2}\right]E[\mathbb{I}_{\{|{S}_{k,l}|>n\theta\}}]}P(\phi_{k-1}(z)=l).

Now, using Markov’s inequality

E[𝕀{|Sk,l|>nθ}]Var[Sk,l]n2θ2=lσ2n2θ2,E[\mathbb{I}_{\{|{S}_{k,l}|>n\theta\}}]\leq\frac{Var[S_{k,l}]}{n^{2}\theta^{2}}=\frac{l\sigma^{2}}{n^{2}\theta^{2}},

and using Lemma 2.1, we have

Bn,k(z)l=02l2σ4n2θ2lσ2P(ϕk1(z)=l)=2σ3θnE[(ϕ0(z))3/2].B_{n,k}(z)\leq\sum_{l=0}^{\infty}\sqrt{2l^{2}\sigma^{4}n^{-2}\theta^{-2}l\sigma^{2}}P(\phi_{k-1}(z)=l)=\frac{\sqrt{2}\sigma^{3}}{\theta n}E[(\phi_{0}(z))^{3/2}].

By Lyapunov’s inequality, E[(ϕk1(z))3/2]2/3E[(ϕk1(z))2]1/2=(ν2(z)+(ε(z))2)1/2E[(\phi_{k-1}(z))^{3/2}]^{2/3}\leq E[(\phi_{k-1}(z))^{2}]^{1/2}=(\nu^{2}(z)+(\varepsilon(z))^{2})^{1/2}. Consequently, in order to prove, as nn\to\infty,

1n2k=1nTBn,k(Zk1)P0,\frac{1}{n^{2}}\sum_{k=1}^{\lfloor nT\rfloor}B_{n,k}(Z_{k-1})\stackrel{{\scriptstyle{P}}}{{\longrightarrow}}0,

is enough to check that

1n3k=1nT(ν2(Zk1)+(ε(Zk1))2)3/4P0.\frac{1}{n^{3}}\sum_{k=1}^{\lfloor nT\rfloor}(\nu^{2}(Z_{k-1})+(\varepsilon(Z_{k-1}))^{2})^{3/4}\stackrel{{\scriptstyle{P}}}{{\longrightarrow}}0.

In fact, using hypotheses A1) and A2) and Proposition 2.1, we have

E[1n3k=1nT(ν2(Zk1)+(ε(Zk1))2)3/4]=n3k=1nTO(k3/2)=O(n1/2).E\left[\frac{1}{n^{3}}\sum_{k=1}^{\lfloor nT\rfloor}(\nu^{2}(Z_{k-1})+(\varepsilon(Z_{k-1}))^{2})^{3/4}\right]=n^{-3}\sum_{k=1}^{\lfloor nT\rfloor}O(k^{3/2})=O(n^{-1/2}).

Finally, we check c.2). We have that

E[Nk2𝕀{|ϕk1(Zk1)ε(Zk1)|>nθ}k1]=Gn,k(Zk1),E\left[N_{k}^{2}\mathbb{I}_{\left\{\left|\phi_{k-1}(Z_{k-1})-\varepsilon(Z_{k-1})\right|>n\theta\right\}}\mid\mathcal{F}_{k-1}\right]=G_{n,k}(Z_{k-1}),

where on {Zk1=z}\{Z_{k-1}=z\}, with z=0,1,z=0,1,\ldots,

Gn,k(z)=E[Sk(z)2𝕀{|ϕk1(z)ε(z)|>nθ}].G_{n,k}(z)=E[S_{k}(z)^{2}\mathbb{I}_{\{|\phi_{k-1}(z)-\varepsilon(z)|>n\theta\}}].

Now, again by Cauchy-Schwarz’s inequality and Markov’s inequality

Gn,k(z)\displaystyle G_{n,k}(z) =\displaystyle= l=0𝕀{|lε(z)|>nθ}E[Sk,l2]P(ϕk1(z)=l)=σ2E[ϕk1(z)𝕀{|ϕk1(z)ε(z)|>nθ}]\displaystyle\sum_{l=0}^{\infty}\mathbb{I}_{\{|l-\varepsilon(z)|>n\theta\}}E[S_{k,l}^{2}]P(\phi_{k-1}(z)=l)=\sigma^{2}E[\phi_{k-1}(z)\mathbb{I}_{\{|\phi_{k-1}(z)-\varepsilon(z)|>n\theta\}}]
\displaystyle\leq σ2E[ϕk12(z)]P(|ϕk1(z)ε(z)|>nθ)σ2E[(ϕ0(z))2]1/2(ν2(z)n2θ2)1/2.\displaystyle\sigma^{2}\sqrt{E[\phi_{k-1}^{2}(z)]P(|\phi_{k-1}(z)-\varepsilon(z)|>n\theta)}\leq\sigma^{2}E[(\phi_{0}(z))^{2}]^{1/2}\left(\frac{\nu^{2}(z)}{n^{2}\theta^{2}}\right)^{1/2}.

In consequence from

E[σ2θn3k=1nT(ν2(Zk1)+(ε(Zk1))2)1/2(ν2(Zk1))1/2]=O(nβ/21),E\left[\frac{\sigma^{2}}{\theta n^{3}}\sum_{k=1}^{\lfloor nT\rfloor}\left(\nu^{2}(Z_{k-1})+(\varepsilon(Z_{k-1}))^{2}\right)^{1/2}(\nu^{2}(Z_{k-1}))^{1/2}\right]=O(n^{\beta/2-1}),

c.2) follows.

Finally, using the weak convergence of {n}n1\{\mathcal{M}_{n}\}_{n\geq 1}, we will obtain weak convergence of {Wn}n1\{W_{n}\}_{n\geq 1}.

Proof of Theorem 3.1. A version of the continuous mapping theorem is applied (see Lemma 1 in Appendix). Let D[0,)D_{\mathbb{R}}[0,\infty) be the space of the real functions on [0,)[0,\infty) that are right continuous and have left limits. For each nn\in\mathbb{N}, by (8), {Wn(t)}t0=Ψ(n)(n)\{W_{n}(t)\}_{t\geq 0}=\Psi^{(n)}\left(\mathcal{M}_{n}\right), where the mapping Ψ(n):D[0,)D[0,)\Psi^{(n)}:D_{\mathbb{R}}[0,\infty)\rightarrow D_{\mathbb{R}}[0,\infty) is given by

(Ψ(n)(f))(t)=f(ntn)+ntnα,\left(\Psi^{(n)}(f)\right)(t)=f\left(\frac{\lfloor nt\rfloor}{n}\right)+\frac{\lfloor nt\rfloor}{n}\alpha,

for fD[0,)f\in D_{\mathbb{R}}[0,\infty) and t[0,)t\in[0,\infty). Indeed, for each nn\in\mathbb{N} and t0t\geq 0:

(Ψ(n)(n))(t)=n(nt/n)+ntnα=1nZnt=Wn(t).\left(\Psi^{(n)}(\mathcal{M}_{n})\right)(t)=\mathcal{M}_{n}(\lfloor nt\rfloor/n)+\frac{\lfloor nt\rfloor}{n}\alpha=\frac{1}{n}Z_{\lfloor nt\rfloor}=W_{n}(t).

Further, taking into account (11), W=𝒟Ψ()W\stackrel{{\scriptstyle\mathcal{D}}}{{=}}\Psi(\mathcal{M}), where the mapping Ψ:D[0,)D[0,)\Psi:D_{\mathbb{R}}[0,\infty)\rightarrow D_{\mathbb{R}}[0,\infty) is given by

(Ψ(f))(t)=f(t)+αt,fD[0,),t[0,).(\Psi(f))(t)=f(t)+\alpha t,\quad f\in D_{\mathbb{R}}[0,\infty),\quad t\in[0,\infty).

The measurability of Ψ(n)\Psi^{(n)}, nn\in\mathbb{N} and Ψ\Psi likewise the conditions for applying Lemma 1 are proved in [1].

Remark 3.2.

1) Notice that the result in Theorem 3.1 is also valid as α=0\alpha=0, and even when the hypothesis A1) is replaced with the more general condition τm(k)=1+k1α+o(k1)\tau_{m}(k)=1+k^{-1}\alpha+o(k^{-1}), as k\ k\to\infty, α0\alpha\geq 0, being the calculation in this latter scenario a little more cumbersome. In the case α=0\alpha=0 and m=1m=1 the result provides an alternative proof of the weak convergence result for the BGW process (see [2], p. 388) for a non-array version.

2) As noted in the introduction, a BPI can be written as a special case of a CBP. For this case, by considering m=E[X0,1]=1m=E[X_{0,1}]=1 (ε(k)=k+E[I0]\varepsilon(k)=k+E[I_{0}] and ν2(k)=Var[I0]\nu^{2}(k)=Var[I_{0}]), we obtain an analogous result to that in [1]. Recall that the difference between both models is in which generation the immigrants will give rise to their offspring.

3) As was pointed out in the Introduction, the proof of the main result follows similar steps as those given in [1]. One can check that similar formulas often appear being the roles of the immigration mean and the offspring variance in the BPI case played in the CBP by α\alpha and m1σ2m^{-1}\sigma^{2}, respectively. However, new approaches by considering conditioning arguments are needed to dealt with b)- c) in Theorem 3.1, as a consequence that random sums of i.i.d. random variables arise in the proofs, see for instance the definition of An,k(z)A_{n,k}(z) and Bn,k(z)B_{n,k}(z), in p.3. An extra work is required to calculate the mathematical expectation of these quantities.

Acknowledgements: The authors thank Professor M. Barczy for his constructive suggestions which have improved this paper. This research has been supported by the Ministerio de Ciencia e Innovación of Spain (grant PID2019-108211GB-I00/AEI/10.13039/501100011033).

Appendix

Theorem A 1.

Let β:[0,)×\beta:[0,\infty)\times\mathbb{R}\rightarrow\mathbb{R} and γ:[0,)×\gamma:[0,\infty)\times\mathbb{R}\rightarrow\mathbb{R} be continuous functions. Assume that uniqueness in the sense of probability law holds for the SDE

d𝒰(t)=β(t,𝒰(t))dt+γ(t,𝒰(t))d𝒲(t),t0,\mathrm{d}\mathcal{U}(t)=\beta\left(t,\mathcal{U}(t)\right)\mathrm{d}t+\gamma\left(t,\mathcal{U}(t)\right)\mathrm{d}\mathcal{W}(t),\quad t\geq 0, (18)

with initial value 𝒰(0)=u(0)\mathcal{U}(0)=u(0) for all u(0)u(0)\in\mathbb{R}, where 𝒲={𝒲(t)}t0\mathcal{W}=\left\{\mathcal{W}(t)\right\}_{t\geq 0} is an one-dimensional standard Wiener process. Let 𝒰={𝒰(t)}t0\mathcal{U}=\left\{\mathcal{U}(t)\right\}_{t\geq 0} be a solution of (18)(\ref{C.1}) with initial value 𝒰(0)=0.\mathcal{U}(0)=0. For each nn\in\mathbb{N}, let {Un(k):k=0,1,2,}\left\{U_{n}(k):k=0,1,2,\ldots\right\} be a sequence of real-valued random variables adapted to a filtration {n(k):k=0,1,2,}\left\{\mathcal{F}_{n}(k):k=0,1,2,\ldots\right\}, that is, Un(k)U_{n}(k) is n(k)\mathcal{F}_{n}(k)- measurable. Let

𝒰n(t):=k=0ntUn(k),t0,n.\mathcal{U}_{n}(t):=\sum_{k=0}^{\lfloor nt\rfloor}U_{n}(k),\quad t\geq 0,\quad n\in\mathbb{N}.

Suppose that E[(Un(k))2]<E\left[\left(U_{n}(k)\right)^{2}\right]<\infty for all n,kn,k\in\mathbb{N}, and 𝒰n(0)𝒟0\mathcal{U}_{n}(0)\stackrel{{\scriptstyle\mathcal{D}}}{{\longrightarrow}}0 as nn\rightarrow\infty. Suppose that for each T>0T>0

  • (i)

    supt[0,T]|k=1ntE[Un(k)n(k1)]0tβ(s,𝒰n(s))ds|P0\sup_{t\in[0,T]}\left|\sum_{k=1}^{\lfloor nt\rfloor}E\left[U_{n}(k)\mid\mathcal{F}_{n}(k-1)\right]-\int_{0}^{t}\beta\left(s,\mathcal{U}_{n}(s)\right)\mathrm{d}s\right|\stackrel{{\scriptstyle{P}}}{{\longrightarrow}}0 as nn\rightarrow\infty,

  • (ii)

    supt[0,T]|k=1ntVar[Un(k)n(k1)]0t(γ(s,𝒰n(s)))2ds|P0\sup_{t\in[0,T]}\left|\sum_{k=1}^{\lfloor nt\rfloor}\operatorname{Var}\left[U_{n}(k)\mid\mathcal{F}_{n}(k-1)\right]-\int_{0}^{t}\left(\gamma\left(s,\mathcal{U}_{n}(s)\right)\right)^{2}\mathrm{~{}d}s\right|\stackrel{{\scriptstyle{P}}}{{\longrightarrow}}0 as nn\rightarrow\infty,

  • (iii)

    k=1nTE[(Un(k))2𝕀{|Un(k)|>θ}n(k1)]P0\sum_{k=1}^{\lfloor nT\rfloor}E\left[\left(U_{n}(k)\right)^{2}{\mathbb{I}}_{\left\{\left|U_{n}(k)\right|>\theta\right\}}\mid\mathcal{F}_{n}(k-1)\right]\stackrel{{\scriptstyle{P}}}{{\longrightarrow}}0 as nn\rightarrow\infty for all θ>0\theta>0.

Then 𝒰n𝒟𝒰\mathcal{U}_{n}\stackrel{{\scriptstyle\mathcal{D}}}{{\longrightarrow}}\mathcal{U} as nn\rightarrow\infty.

The proof can be seen in [9].

Theorem A 2.

Let aa, bb, cc real constants such that a>0a>0. Consider the stochastic differential equation

dX(t)=(bX(t)+c)dt+2aX(t)+d𝒲t,t0.\mathrm{d}X(t)=(bX(t)+c)\mathrm{d}t+\sqrt{2aX(t)^{+}}\mathrm{d}\mathcal{W}_{t},\ t\geq 0. (19)

There exists a pathwise unique strong solution {X(t)(x)}t0\{X(t)^{(x)}\}_{t\geq 0} for all initial values X(0)(x)=xX(0)^{(x)}=x\in\mathbb{R}. Moreover if x0x\geq 0, then X(t)(x)0X(t)^{(x)}\geq 0 almost surely for all t0t\geq 0. In the case c0c\geq 0, the solution of (19) defines diffusion process with generator

Tf(x)=(bx+c)f(x)+axf′′(x),fCc[0,),Tf(x)=(bx+c)f^{\prime}(x)+axf^{\prime\prime}(x),\ f\in C_{c}^{\infty}[0,\infty),

where Cc[0,)C_{c}^{\infty}[0,\infty) is the space of infinitely differentiable functions on [0,)[0,\infty) which have a compact support.

The proof can be seen in [8] p. 235.

Lemma A 1.

Let SS and TT be two metric spaces, and X,X1,X2,X,X_{1},X_{2},\cdots be random functions with values in SS with Xn𝒟XX_{n}\stackrel{{\scriptstyle\mathcal{D}}}{{\rightarrow}}X. Consider some measurable mappings h,h1,h2,:STh,h_{1},h_{2},\cdots:S\rightarrow T and a measurable set CSC\subset S with XCX\in C a.s. such that hn(sn)h(s)h_{n}(s_{n})\rightarrow h(s) as snsCs_{n}\rightarrow s\in C. Then hn(Xn)dh(X)h_{n}(X_{n})\stackrel{{\scriptstyle d}}{{\rightarrow}}h(X).

The previous version of the continuous mapping theorem can be found in Theorem 3.27 in [11].

References

  • [1] Barczy, M., Bezdány, D. and Pap, G. A note on asymptotic behavior of critical Galton-Watson processes with immigration, arXiv preprint, arXiv:2103.07878v2:1–20, 2021.
  • [2] Ethier, S.N., Kurtz, T.G. Markov Processes: Characterization and Convergence. Wiley Series in Probability and Mathematical Statistics. John Wiley & Sons, 1986.
  • [3] Feller, W. Diffusion processes in genetics. In: Proc Second Berkeley Symp Math Statist Prob, University of California Press, Berkeley, 227- 246, 1951.
  • [4] Foster, J.H. A limit theorem for a branching process with state-dependent immigration. The Annals of Mathematical Statistics, 42: 1773-1776, 1971.
  • [5] González, M., Molina, M., and del Puerto, I. Asymptotic behaviour for the critical controlled branching process with random control function. Journal of Applied Probability, 42: 463–477, 2005.
  • [6] González, M. and del Puerto, I. Diffusion approximation of an array of controlled branching processes. Methodology and Computing in Applied Probability, 14: 843-861, 2012.
  • [7] González, M., del Puerto, I. and Yanev, G.P. Controlled Branching Processes. ISTE Ltd and John Wiley and Sons, Inc., 2018.
  • [8] Ikeda, N. and Watanabe, S. . Stochastic Differential Equations and Diffusion Processes, 2nd ed. North-Holland, Kodansha, Amsterdam, Tokyo, 1989.
  • [9] Ispány, M. and Pap, G. A note on weak convergence of step processes. Acta Mathematica Hungarica, 126(4): 381-395, 2010.
  • [10] Jirina, M. Stochastic Branching Processes with Continuous State Space. Czechoslovak Mathematical Journal, 8: 292–313, 1958.
  • [11] Kallenberg, O. Foundations of Modern Probability. Springer, 1997
  • [12] Lindvall, T. Convergence of critical Galton-Watson processes. Journal of Applied Probability, 9: 445-450, 1972.
  • [13] Pakes, A.G. On the critical Galton-Watson process with immigration. Journal of the Australian Mathematical Society, 12: 476-482, 1971.
  • [14] Sriram, T.N., Bhattacharya, A., González, M., Martínez, R., del Puerto, I. Estimation of the offspring mean in a controlled branching process with a random control function. Stochastic Processes and their Applications, 117: 928-946, 2007.
  • [15] Wei, C. Z. and Winnicki, J. Some asymptotic results for the branching process with immigration. Stochastic Processes and their Applications, 31(2): 261–282, 1989.