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

From law of the iterated logarithm to Zolotarev distance for supercritical branching processes in random environment

Yinna Ye [email protected] Department of Applied Mathematics, School of Mathematics and Physics, Xi’an Jiaotong-Liverpool University, Suzhou 215123215123, P. R. China
( )
Abstract

Consider (Zn)n0(Z_{n})_{n\geqslant 0} a supercritical branching process in an independent and identically distributed environment. Based on some recent development in martingale limit theory, we established law of the iterated logarithm, strong law of large numbers, invariance principle and optimal convergence rate in the central limit theorem under Zolotarev and Wasserstein distances of order p(0,2]p\in(0,2] for the process (logZn)n0(\log Z_{n})_{n\geqslant 0}.

keywords:
Branching processes in random environment , Law of the iterated logarithm , Law of large numbers , Convergence rates in central limit theorem , Zolotarev distance , Wasserstein distance
MSC:
60J80; 60K37; 60F05; 62E20

1 Introduction

Branching process in random environment (BPRE), initially introduced by Smith and Wilkinson [19], is a generalization of Galton-Watson process. The asymptotic behaviours for BPRE have been studied widely and intensively since the 1970s. For instance, one may refer to Athreya and Karlin [2, 1], Tanny [21, 22], Guivarc’h and Liu [8] and Birkner et al. [3] for some classical results on branching processes in independent and identically distributed (i.i.d.) environment or stationary and ergodic environment; Le Page and Ye [14] and Grama et al. [5, 6] for branching processes in Markovian environment. Regarding to the recent advances especially for supercritical branching processes in i.i.d. environment, for instance, Huang and Liu [11, 12] and Li et al. [15] established the large and moderate deviation principles, convergence theorems in 𝕃p\mathbb{L}^{p} space and the central limit theorem (CLT); Grama et al. [7] studied asymptotic distribution and harmonic moments.

The main objective of this paper is to establish some additional limit theorems for supercritical BPRE, which have rarely appeared in the literature so far, including law of the iterated logarithm (LIL), strong law of large numbers (LLN), invariance principle and the rate of convergence in the CLT under Zolotarev and Wasserstein distances. The study of Wasserstein and Zolotarev distances are of independent interests, as they have been widely used in applications nowadays, such as machine learning ([9]), computer vision ([16], [20]) and microbiome studies ([23]). The methods used in this paper are mainly based on some recent development in martingale limit theory, especially on the invariance principles and LIL for martingales. One may refer to Hall and Heyde [10] and Shao [18] for the martingale limit theory and invariance principles; Wu et al. [24] for some recent advances on a Berry-Esseen bound for martingales.

Let’s firstly introduce the model of BPRE. Suppose ξ=(ξ0,ξ1,)\xi=(\xi_{0},\,\xi_{1},\,\ldots) is a sequence of i.i.d. random variables. Usually, ξ\xi is called environment process and ξn\xi_{n} represents the random environment in the nn-th generation. A discrete-time random process (Zn)n0(Z_{n})_{n\geqslant 0} is called BPRE ξ\xi, if it satisfies the following recursive relation:

Z0=1,Zn+1=i=1ZnXn,i,n0,Z_{0}=1,~{}Z_{n+1}=\sum_{i=1}^{Z_{n}}X_{n,i},\ \ n\geqslant 0,

where Xn,iX_{n,i} represents the number of offspring produced by the ii-th particle in the nn-th generation. The distribution of Xn,iX_{n,i} depending on the environment ξn\xi_{n} is denoted by p(ξn)={pk(ξn)=(Xn,i=k|ξn):k}.p(\xi_{n})=\{p_{k}(\xi_{n})=\mathbb{P}(X_{n,i}=k|\xi_{n}):k\in\mathbb{N}\}. Suppose that given ξn\xi_{n}, (Xn,i)i1(X_{n,i})_{i\geqslant 1} is a sequence of i.i.d. random variables; moreover, (Xn,i)i1(X_{n,i})_{i\geqslant 1} is independent of (Z1,,Zn)(Z_{1},\ldots,Z_{n}). Let (Γ,ξ)\left(\Gamma,\mathbb{P}_{\xi}\right) be the probability space under which the process is defined when the environment ξ\xi is given. The state space of the random environment ξ\xi is denoted by Θ\Theta and the total probability space can be regarded as the product space (Θ×Γ,),\left(\Theta^{\mathbb{N}}\times\Gamma,\mathbb{P}\right), where (dx,dξ)=ξ(dx)τ(dξ).\mathbb{P}(\mbox{d}x,\mbox{d}\xi)=\mathbb{P}_{\xi}(\mbox{d}x)\tau(\mbox{d}\xi). That is, for any measurable positive function gg defined on Θ×Γ\Theta^{\mathbb{N}}\times\Gamma, we have

g(x,ξ)(dx,dξ)=g(x,ξ)ξ(dx)τ(dξ),\int g(x,\xi)\,\mathbb{P}(\mbox{d}x,\mbox{d}\xi)=\iint g(x,\xi)\,\mathbb{P}_{\xi}(\mbox{d}x)\,\tau(\mbox{d}\xi),

where τ\tau represents the distribution law of the random environment ξ\xi. And ξ\mathbb{P}_{\xi} can be regarded as the conditional probability of \mathbb{P} given the environment ξ\xi. The expectations with respect to (w.r.t) ξ\mathbb{P}_{\xi} and \mathbb{P} are denoted by 𝔼ξ\mathbb{E}_{\xi} and 𝔼\mathbb{E}, respectively. For any environment ξ\xi, integer n0n\geqslant 0 and real number p>0p>0, define

mn(p)=mn(p)(ξ)=i=0ippi(ξn),mn=mn(ξ)=mn(1)(ξ).m_{n}^{(p)}=m_{n}^{(p)}(\xi)=\sum_{i=0}^{\infty}i^{p}\,p_{i}(\xi_{n}),\quad m_{n}=m_{n}(\xi)=m_{n}^{(1)}(\xi).

Then

m0(p)=𝔼ξZ1p,mn=𝔼ξXn,i,i1.m_{0}^{(p)}=\mathbb{E}_{\xi}Z_{1}^{p},~{}~{}m_{n}=\mathbb{E}_{\xi}X_{n,i},~{}~{}i\geqslant 1.

Consider the following random variables

Π0=1,Πn=Πn(ξ)=i=0n1mi,n1.\Pi_{0}=1,\ \Pi_{n}=\Pi_{n}(\xi)=\prod_{i=0}^{n-1}m_{i},~{}~{}n\geqslant 1.

It is easy to see that

𝔼ξZn+1=𝔼ξ[i=1ZnXn,i]=𝔼ξ[𝔼ξ(i=1ZnXn,i|Zn)]=𝔼ξ[i=1Znmn]=mn𝔼ξZn.\mathbb{E}_{\xi}Z_{n+1}=\mathbb{E}_{\xi}\Big{[}\sum_{i=1}^{Z_{n}}X_{n,i}\Big{]}=\mathbb{E}_{\xi}\Big{[}\mathbb{E}_{\xi}\Big{(}\sum_{i=1}^{Z_{n}}X_{n,i}\Big{|}Z_{n}\Big{)}\Big{]}=\mathbb{E}_{\xi}\Big{[}\sum_{i=1}^{Z_{n}}m_{n}\Big{]}=m_{n}\,\mathbb{E}_{\xi}Z_{n}.

Then by recursion, we get Πn=𝔼ξZn\Pi_{n}=\mathbb{E}_{\xi}Z_{n}. Denote

X=logm0,μ=𝔼X,σ2=𝔼(Xμ)2.X=\log m_{0},~{}\mu=\mathbb{E}X,~{}~{}\sigma^{2}=\mathbb{E}(X-\mu)^{2}.

The branching process (Zn)n0(Z_{n})_{n\geqslant 0} is called supercritical, critical or subcritical according to μ>0\mu>0, μ=0\mu=0 or μ<0\mu<0, respectively. Over all the paper, assume that

p0(ξ0)=0 a.s.p_{0}(\xi_{0})=0\text{ a.s.} (1.1)

The condition (1.1) means that each particle has at least one offspring, which implies X0X\geqslant 0 a.s. and hence Zn1Z_{n}\geqslant 1 a.s. for any n1n\geqslant 1.

Let’s introduce now Zolotarev distance and Wasserstein distance respectively between two probability laws. For any p>0p>0, denote by [p][p] the largest integer which is strictly less than pp, i.e.

[p]=max{n:n<p}.[p]=\max\{n\in\mathbb{Z}:\,n<p\}.

Set l=[p]l=[p]. Then pp can be decomposed uniquely as p=l+δp=l+\delta, with δ(0,1]\delta\in(0,1]. In the sequel, we will use this notation for the unique decomposition of any p>0p>0. For a given p>0p>0, consider Λp\Lambda_{p}, the class of ll-times continuously differentiable real-valued functions, defined by

Λp={f::|f(l)(x)f(l)(y)||xy|pl,(x,y)2}.\Lambda_{p}=\{f:\mathbb{R}\rightarrow\mathbb{R}:\,|f^{(l)}(x)-f^{(l)}(y)|\leqslant|x-y|^{p-l},\;\forall(x,y)\in\mathbb{R}^{2}\}.

For ll-times continuously differentiable function ff, set

fΛp=sup{|f(l)(x)f(l)(y)||xy|pl:(x,y)2}.\|f\|_{\Lambda_{p}}=\sup\left\{\frac{|f^{(l)}(x)-f^{(l)}(y)|}{|x-y|^{p-l}}:\;(x,y)\in\mathbb{R}^{2}\right\}.

Then for any fΛpf\in\Lambda_{p},

fΛp1.\displaystyle\|f\|_{\Lambda_{p}}\leqslant 1. (1.2)

Consider a set of probability laws (μ,ν)\mathcal{L}(\mu,\nu) on 2\mathbb{R}^{2} with marginals μ\mu and ν\nu. Zolotarev distance of order pp between μ\mu and ν\nu is defined by

ζp(μ,ν)=sup{f𝑑μf𝑑ν:fΛp}.\zeta_{p}(\mu,\nu)=\sup\left\{\int fd\mu-\int fd\nu:\,f\in\Lambda_{p}\right\}.

For any p>0p>0, the Wasserstein distance of order pp between μ\mu and ν\nu is defined by

Wp(μ,ν)={inf{|xy|p(dx,dy):(μ,ν)},if 0<p1;inf{(|xy|p(dx,dy))1/p:(μ,ν)},if p>1.W_{p}(\mu,\nu)=\begin{cases}\inf\left\{\int|x-y|^{p}\,\mathbb{P}(dx,dy):\,\mathbb{P}\in\mathcal{L}(\mu,\nu)\right\},\quad&\text{if }0<p\leqslant 1;\\ \inf\left\{\left(\int|x-y|^{p}\,\mathbb{P}(dx,dy)\right)^{1/p}:\,\mathbb{P}\in\mathcal{L}(\mu,\nu)\right\},\quad&\text{if }p>1.\end{cases}

Let p\mathcal{F}_{p} be the collection of all Borel probability measures on \mathbb{R} with finite absolute moments of order pp; then (p,Wp)(\mathcal{F}_{p},W_{p}) forms a metric space, which is closely related to the topology of weak convergence of probability distributions on \mathbb{R}.

According to Kantorovich and Rubinstein [13], for any 0<p10<p\leqslant 1,

Wp(μ,ν)=ζp(μ,ν).W_{p}(\mu,\nu)=\zeta_{p}(\mu,\nu). (1.3)

From Rio [17], it is proved that for any p>1p>1,

Wp(μ,ν)cp(ζp(μ,ν))1/p,\displaystyle W_{p}(\mu,\nu)\leqslant c_{p}\left(\zeta_{p}(\mu,\nu)\right)^{1/p}, (1.4)

where cp>0c_{p}>0 is a constant depending only on pp.

Throughout the paper, CC denotes a positive constant whose value may vary from line to line. If XX and YY are two random variables, such that they follow the probability laws PXP_{X} and PYP_{Y} respectively, then ζp(PX,PY)\zeta_{p}(P_{X},\,P_{Y}) (resp. Wp(PX,PY)W_{p}(P_{X},\,P_{Y})) is simply denoted by ζp(X,Y)\zeta_{p}(X,\,Y) (resp. Wp(X,Y)W_{p}(X,\,Y)). We will use GtG_{t} to represent the 𝒩(0,t)\mathcal{N}(0,t)-distributed random variable with variance t>0t>0; in particular, 𝒩\mathcal{N} represents the standard normal random variable. And ab=max{a,b}a\vee b=\max\{a,b\}, for any (a,b)2(a,b)\in\mathbb{R}^{2}.

2 Main results

In the sequel, denote by Xi=logmi,i0.X_{i}=\log m_{i},\ i\geqslant 0. Evidently, (Xi)i0(X_{i})_{i\geqslant 0} is a sequence of i.i.d. random variables depending only on the environment ξ\xi. Let (Sn)n0(S_{n})_{n\geqslant 0} be the random walk associated with the branching process, which is defined as follows:

S0=0,Sn=logΠn=i=0n1Xi,n1.S_{0}=0,\quad S_{n}=\log\Pi_{n}=\sum_{i=0}^{n-1}X_{i},~{}~{}n\geqslant 1.

Then we have the following decomposition of logZn\log Z_{n}:

logZn=Sn+logWn,\log Z_{n}=S_{n}+\log W_{n}, (2.5)

where Wn=ZnΠn.W_{n}=\frac{Z_{n}}{\Pi_{n}}. The normalized population size (Wn)n0(W_{n})_{n\geqslant 0} is a non-negative martingale under both \mathbb{P} and ξ\mathbb{P}_{\xi}, w.r.t the natural filtration (n)n0(\mathcal{F}_{n})_{n\geqslant 0}, defined by

0=σ{ξ},n=σ{ξ,Xk,i,0kn1,i1},n1.\mathcal{F}_{0}=\sigma\{\xi\},\ \mathcal{F}_{n}=\sigma\{\xi,X_{k,i},0\leqslant k\leqslant n-1,i\geqslant 1\},~{}~{}n\geqslant 1.

By Doob’s martingale convergence theorem and Fatou’s lemma, we can obtain that WnW_{n} converges a.s. to a finite limit WW and 𝔼W1\mathbb{E}W\leqslant 1. We assume the following conditions throughout this paper:

σ(0,)and𝔼Z1m0logZ1<.\sigma\in(0,\infty)~{}~{}\text{and}~{}~{}\mathbb{E}\frac{Z_{1}}{m_{0}}\log Z_{1}<\infty. (2.6)

The first condition above together with (1.1) imply in particular that

Z11a.s.and(Z1=1)=𝔼p1(ξ0)<1.Z_{1}\geqslant 1~{}~{}\text{a.s.}~{}~{}\text{and}~{}~{}\mathbb{P}(Z_{1}=1)=\mathbb{E}p_{1}(\xi_{0})<1.

The second condition in (2.6) implies that WnW_{n} converges to WW in 𝕃1\mathbb{L}^{1} and

(W>0)=(Znn)=limn(Zn>0)>0\mathbb{P}(W>0)=\mathbb{P}(Z_{n}\stackrel{{\scriptstyle n\rightarrow\infty}}{{\longrightarrow}}\infty)=\lim_{n\rightarrow\infty}\mathbb{P}(Z_{n}>0)>0

(See e.g. Tanny [22]). Therefore, it follows with the assumption (1.1) that W>0W>0 and ZnnZ_{n}\stackrel{{\scriptstyle n\rightarrow\infty}}{{\longrightarrow}}\infty a.s.

In the sequel, we will need the following conditions.

Condition 1

There exists a constant δ(0,1)\delta\in(0,1), such that

𝔼X2+δ=𝔼(logm0)2+δ<.\mathbb{E}X^{2+\delta}=\mathbb{E}\left(\log m_{0}\right)^{2+\delta}<\infty.
Condition 2

There exist some constants p>1p>1 and c>0c>0 such that

𝔼m0c<and𝔼(m0(p)m0p)c<.\mathbb{E}m_{0}^{c}<\infty\quad\text{and}\quad\mathbb{E}\left(\frac{m_{0}^{(p)}}{m_{0}^{p}}\right)^{c}<\infty.

We have the following LIL for the process (logZn)n0(\log Z_{n})_{n\geqslant 0}.

Theorem 2.1

Suppose that Condition 1 is satisfied. Then

lim supnlogZnnμnloglogn=+2σa.s.\limsup_{n\rightarrow\infty}\frac{\log Z_{n}-n\mu}{\sqrt{n\log\log n}}=+\sqrt{2}\,\sigma\quad\text{a.s.}

and

lim infnlogZnnμnloglogn=2σa.s.\liminf_{n\rightarrow\infty}\frac{\log Z_{n}-n\mu}{\sqrt{n\log\log n}}=-\sqrt{2}\,\sigma\quad\text{a.s.}

From the LIL theorem above, we can find immediately the following strong LLN for (logZn)n0(\log Z_{n})_{n\geqslant 0}.

Corollary 2.1

Under the same condition of Theorem 2.1, then

limnlogZnn=μa.s.\lim_{n\rightarrow\infty}\frac{\log Z_{n}}{n}=\mu\quad\text{a.s.}

We also have the following invariance principle for the process (logZnnμ)n0(\log Z_{n}-n\mu)_{n\geqslant 0}, which only requires the existence of the second order moment 𝔼X2<\mathbb{E}X^{2}<\infty. And this has been insured in the assumption (2.6).

Theorem 2.2

Let (B(t))t0(B(t))_{t\geqslant 0} be a standard Brownian motion. Then there exists a common probability space for (logZn)n0(\log Z_{n})_{n\geqslant 0} and (B(t))t0(B(t))_{t\geqslant 0}, such that the following holds:

logZnnμB(nσ2)=o(nloglogn)a.s.,\log Z_{n}-n\mu-B(n\sigma^{2})=o\left(\sqrt{n\log\log n}\right)\quad\text{a.s.,}

as nn\rightarrow\infty.

We have the following convergence rates of (logZn)n0(\log Z_{n})_{n\geqslant 0} in the CLT under Zolotarev and Wasserstein distances respectively.

Theorem 2.3

Under Conditions 1 and 2, we have for any r[δ, 2]r\in[\delta,\,2],

ζr(logZnnμnσ,𝒩)Cnδ/2.\zeta_{r}\left(\frac{\log Z_{n}-n\mu}{\sqrt{n}\sigma},\,\mathcal{N}\right)\leqslant\frac{C}{n^{\delta/2}}.
Corollary 2.2

Under Conditions 1 and 2, then

Wr(logZnnμnσ,𝒩)Cnδ/2,for r[δ, 1];W_{r}\left(\frac{\log Z_{n}-n\mu}{\sqrt{n}\sigma},\;\mathcal{N}\right)\leqslant\frac{C}{n^{\delta/2}},\quad\text{for }r\in[\delta,\,1];
Wr(logZnnμnσ,𝒩)Cnδ/2r,for r(1, 2].W_{r}\left(\frac{\log Z_{n}-n\mu}{\sqrt{n}\sigma},\;\mathcal{N}\right)\leqslant\frac{C}{n^{\delta/2r}},\quad\text{for }r\in(1,\,2].

The result above is an immediate consequence of Theorem 2.3, the equality (1.3) and the inequality (1.4) respectively for the cases r[δ, 1]r\in[\delta,\,1] and r(1, 2]r\in(1,\,2].

3 Proof of main results

Consider a sequence of martingale differences (εi)i1(\varepsilon_{i})_{i\geqslant 1}, defined on a probability space (Ω,,)(\Omega,\mathcal{F},\mathbb{P}), w.r.t a filtration 𝒢=(𝒢i)i0\mathcal{G}=(\mathcal{G}_{i})_{i\geqslant 0}. Set

M0=0,Mn=i=1nεi, for n1.M_{0}=0,\quad M_{n}=\sum_{i=1}^{n}\varepsilon_{i},\text{ for }n\geqslant 1.

Then (Mn,𝒢n)n0(M_{n},\mathcal{G}_{n})_{n\geqslant 0} is a martingale. Set Vn2=i=1nεi2V_{n}^{2}=\sum_{i=1}^{n}\varepsilon_{i}^{2} and sn2=𝔼Mn2s_{n}^{2}=\mathbb{E}M_{n}^{2}. We have the following lemma (see also Corollary 4.2 and Theorem 4.8 in Hall and Heyde [10]), from which Theorem 2.1 can be obtained.

Lemma 3.1 (Hall and Heyde [10])

Suppose that (Mn,𝒢n)n0(M_{n},\,\mathcal{G}_{n})_{n\geqslant 0} is a zero-mean, square integrable martingale. If

sn2Vn2nη2>0a.s.,s_{n}^{-2}V^{2}_{n}\xrightarrow{n\rightarrow\infty}\eta^{2}>0\quad\text{a.s.}, (3.7)
for any ε>0,n=1sn1𝔼(|εn|𝟏{|εn|>εsn})<\text{for any }\varepsilon>0,\quad\sum_{n=1}^{\infty}s_{n}^{-1}\mathbb{E}\left(|\varepsilon_{n}|\mathbf{1}_{\{|\varepsilon_{n}|>\varepsilon s_{n}\}}\right)<\infty (3.8)

and for some τ>0\tau>0,

n=1sn4𝔼(εn4 1{|εn|τsn})<;\sum_{n=1}^{\infty}s_{n}^{-4}\,\mathbb{E}\left(\varepsilon_{n}^{4}\,\mathbf{1}_{\{|\varepsilon_{n}|\leqslant\tau s_{n}\}}\right)<\infty; (3.9)

then

lim supn(ϕ(Vn2))1Mn=+1a.s.\limsup_{n\rightarrow\infty}\left(\phi(V_{n}^{2})\right)^{-1}M_{n}=+1\quad\text{a.s.}

and

lim infn(ϕ(Vn2))1Mn=1a.s.,\liminf_{n\rightarrow\infty}\left(\phi(V_{n}^{2})\right)^{-1}M_{n}=-1\quad\text{a.s.},

where ϕ(t)=(2tloglogt)1/2\phi(t)=(2t\log\log t)^{1/2} for any t>0t>0.

Recall now that the random walk associated with the branching process

S0=0,Sn=i=0n1Xi,n1S_{0}=0,\quad S_{n}=\sum_{i=0}^{n-1}X_{i},~{}~{}n\geqslant 1

is the sum of i.i.d. random variables (Xi)i0(X_{i})_{i\geqslant 0}. In the sequel, we will use the following notations. Let εi=Xi1μ\varepsilon_{i}=X_{i-1}-\mu, i1i\geqslant 1 and

M0=0,Mn=i=1nεi,n1.M_{0}=0,\quad M_{n}=\sum_{i=1}^{n}\varepsilon_{i},\quad n\geqslant 1.

Then from (2.5), we have

logZnnμ=\displaystyle\log Z_{n}-n\mu= Mn+logWn.\displaystyle M_{n}+\log W_{n}. (3.10)

Since (εi)i1(\varepsilon_{i})_{i\geqslant 1} is a sequence of i.i.d. zero-mean random variables and σ<\sigma<\infty, (Mn)n0(M_{n})_{n\geqslant 0} becomes a zero-mean and square integrable martingale w.r.t the filtration (n)n0(\mathcal{F}_{n})_{n\geqslant 0}.

Proof of Theorem 2.1 1

Since Vn2=i=1n(Xiμ)2=i=2nεi2V_{n}^{2}=\sum_{i=1}^{n}(X_{i}-\mu)^{2}=\sum_{i=2}^{n}\varepsilon^{2}_{i} and sn2=nσ2s_{n}^{2}=n\sigma^{2}, from the strong LLN for Vn2V^{2}_{n},

sn2Vn2\displaystyle s^{-2}_{n}V^{2}_{n} =1nσ2i=2nεi2n1a.s.\displaystyle=\frac{1}{n\sigma^{2}}\sum_{i=2}^{n}\varepsilon^{2}_{i}\xrightarrow{n\rightarrow\infty}1\quad\text{a.s.} (3.11)

the condition (3.7) is hence satisfied, with η2=1>0\eta^{2}=1>0. Let σ(2+δ)=𝔼|Xμ|2+δ\sigma^{(2+\delta)}=\mathbb{E}|X-\mu|^{2+\delta}. Then under Condition 1, σ(2+δ)<\sigma^{(2+\delta)}<\infty. And the condition (3.8) is satisfied, because for any ε>0\varepsilon>0,

n=1sn1𝔼(|εn|𝟏{|εn|>εsn})σ(2+δ)ε1+δn=11sn2+δ=σ(2+δ)ε1+δσ2+δn=11n1+δ/2<.\displaystyle\sum_{n=1}^{\infty}s_{n}^{-1}\,\mathbb{E}\left(|\varepsilon_{n}|\mathbf{1}_{\{|\varepsilon_{n}|>\varepsilon s_{n}\}}\right)\leqslant\frac{\sigma^{(2+\delta)}}{\varepsilon^{1+\delta}}\sum_{n=1}^{\infty}\frac{1}{s_{n}^{2+\delta}}=\frac{\sigma^{(2+\delta)}}{\varepsilon^{1+\delta}\,\sigma^{2+\delta}}\sum_{n=1}^{\infty}\frac{1}{n^{1+\delta/2}}<\infty.

The condition (3.9) is also satisfied, because for any τ>0\tau>0, we have

n=1sn4𝔼(εn4 1{|εn|τsn})τ2δσ(2+δ)n=11sn2+δ=τ2δσ(2+δ)σ2+δn=11n1+δ/2<.\displaystyle\sum_{n=1}^{\infty}s_{n}^{-4}\,\mathbb{E}\left(\varepsilon_{n}^{4}\,\mathbf{1}_{\{|\varepsilon_{n}|\leqslant\tau s_{n}\}}\right)\leqslant\tau^{2-\delta}\sigma^{(2+\delta)}\sum_{n=1}^{\infty}\frac{1}{s_{n}^{2+\delta}}=\frac{\tau^{2-\delta}\,\sigma^{(2+\delta)}}{\sigma^{2+\delta}}\sum_{n=1}^{\infty}\frac{1}{n^{1+\delta/2}}<\infty.

Now applying Lemma 3.1 to the martingale (Mn,n)n0(M_{n},\,\mathcal{F}_{n})_{n\geqslant 0}, using the facts that limnϕ(Vn2)ϕ(n)=limnVn2n=σ\displaystyle\lim_{n\rightarrow\infty}\frac{\phi(V_{n}^{2})}{\phi(n)}=\lim_{n\rightarrow\infty}\sqrt{\frac{V_{n}^{2}}{n}}=\sigma a.s., limn|logWn|nloglogn=0\displaystyle\lim_{n\rightarrow\infty}\frac{|\log W_{n}|}{\sqrt{n\log\log n}}=0 a.s., we come to the statement of Theorem 2.1. \Box

Theorem 2.2 can be proved by using Theorem 2.1 in Shao [18], where the following invariance principle for the martingale (Mn,𝒢n)n0(M_{n},\,\mathcal{G}_{n})_{n\geqslant 0} with the martingale difference sequence (εi)i1(\varepsilon_{i})_{i\geqslant 1} is established.

Lemma 3.2 (Shao [18], Theorem 2.1)

Suppose that {Mn=i=1nεi,𝒢n}n0\{M_{n}=\sum_{i=1}^{n}\varepsilon_{i},\,\mathcal{G}_{n}\}_{n\geqslant 0} is a square integrable martingale and (B(t))t0(B(t))_{t\geqslant 0} is a standard Brownian motion. Let Z(t)=ktεkZ(t)=\sum_{k\leqslant t}\varepsilon_{k}, b(t)=𝔼Z2(t)b(t)=\mathbb{E}Z^{2}(t), t0t\geqslant 0. Assume that there exists a non-decreasing positive valued sequence (cn)n1(c_{n})_{n\geqslant 1} with cnc_{n}\rightarrow\infty, such that, as nn\rightarrow\infty,

i=1n[𝔼(εi2|𝒢i1)𝔼εi2]=o(cn)a.s.\sum_{i=1}^{n}\left[\mathbb{E}\left(\varepsilon_{i}^{2}|\mathcal{G}_{i-1}\right)-\mathbb{E}\varepsilon_{i}^{2}\right]=o(c_{n})\quad\text{a.s.}

and

i=1civ𝔼|εi|2v< for some 1v2.\sum_{i=1}^{\infty}c_{i}^{-v}\,\mathbb{E}|\varepsilon_{i}|^{2v}<\infty\text{ for some }1\leqslant v\leqslant 2.

Then there exists a common probability space for (εi)i1(\varepsilon_{i})_{i\geqslant 1} and (B(t))t0(B(t))_{t\geqslant 0} such that, as nn\rightarrow\infty,

Z(n)B(bn)=o((cn(logbncn+loglogcn))1/2)a.s.Z(n)-B(b_{n})=o\left(\left(c_{n}\left(\log\frac{b_{n}}{c_{n}}+\log\log c_{n}\right)\right)^{1/2}\right)\quad\text{a.s.}
Proof of Theorem 2.2 1

From (3.10), we have

logZnnμB(nσ2)=Mn+logWnB(nσ2).\log Z_{n}-n\mu-B(n\sigma^{2})=M_{n}+\log W_{n}-B(n\sigma^{2}).

Applying Lemma 3.2 to the martingale (Mn,n)n0(M_{n},\,\mathcal{F}_{n})_{n\geqslant 0}, letting cn=nc_{n}=n, bn=nσ2b_{n}=n\sigma^{2} and v=1+δ/2(1, 3/2)v=1+\delta/2\in(1,\,3/2), and taking into account that WnW_{n} converges a.s. to a finite limit WW, we can obtain immediately the result. \Box

To prove Theorem 2.3, the following lemmas will be used.

Lemma 3.3

Assume Condition 2 hold. For the constants pp and cc given from Condition 2, there exists a constant ap,c(0,1)a_{p,c}\in(0,1) depending on pp and cc, such that for any q(0, 3+ap,c)q\in(0,\,3+a_{p,c}),

𝔼|logW|q<andsupn0𝔼|logWn|q<.\mathbb{E}|\log W|^{q}<\infty\quad\text{and}\quad\sup_{n\geqslant 0}\,\mathbb{E}|\log W_{n}|^{q}<\infty.
Proof 1

By Lyapunov’s inequality, it is enough to prove the assertion for q(3, 3+ap,c)q\in(3,\,3+a_{p,c}). Let q(3, 3+ap,c)q\in(3,\,3+a_{p,c}). Using truncation, we have

𝔼|logW|q=𝔼|logW|q𝟏{W>1}+𝔼|logW|q𝟏{W1}.\mathbb{E}|\log W|^{q}=\mathbb{E}|\log W|^{q}\mathbf{1}_{\{W>1\}}+\mathbb{E}|\log W|^{q}\mathbf{1}_{\{W\leqslant 1\}}.

For the first term on the right-hand side above, we have

𝔼|logW|q𝟏{W>1}C𝔼W<.\mathbb{E}|\log W|^{q}\mathbf{1}_{\{W>1\}}\leqslant C\,\mathbb{E}W<\infty.

For the second term, let ψ\psi be the annealed Laplace transform of WW, defined by

ψ(t)=𝔼etW,t0.\psi(t)=\mathbb{E}\mbox{e}^{-tW},\quad t\geqslant 0.

By Markov’s inequality, we have for t>0t>0,

(Wt1)e𝔼etW=eψ(t).\mathbb{P}(W\leqslant t^{-1})\leqslant e\,\mathbb{E}e^{-tW}=e\,\psi(t).

Then

𝔼|logW|q𝟏{W1}\displaystyle\mathbb{E}|\log W|^{q}\mathbf{1}_{\{W\leqslant 1\}} =q11t(logt)q1(Wt1)𝑑t\displaystyle=\,q\int_{1}^{\infty}\frac{1}{t}(\log t)^{q-1}\,\mathbb{P}(W\leqslant t^{-1})\,dt
qe1ψ(t)t(logt)q1𝑑t.\displaystyle\leqslant\,q\,e\int_{1}^{\infty}\frac{\psi(t)}{t}(\log t)^{q-1}\,dt. (3.12)

From (3.20) of Grama et al. [7], it is proved that under Condition 2,

ψ(t)Ctap,c,\displaystyle\psi(t)\leqslant Ct^{-a_{p,c}}, (3.13)

for some constants C>0C>0, ap,c(0,1)a_{p,c}\in(0,1) depending on pp and cc, and for all t>0t>0. Therefore, combining (1) with (3.13), we get

𝔼|logW|q𝟏{W1}C1(logt)q1t1+ap,c𝑑t=\displaystyle\mathbb{E}|\log W|^{q}\mathbf{1}_{\{W\leqslant 1\}}\leqslant\,C\int_{1}^{\infty}\frac{(\log t)^{q-1}}{t^{1+a_{p,c}}}\,dt=\, C0yq1eyap,c𝑑y=Cap,cqΓ(q)<,\displaystyle C\int_{0}^{\infty}y^{q-1}\mbox{e}^{-y\,a_{p,c}}\,dy=\,Ca_{p,c}^{-q}\,\Gamma(q)<\infty,

where Γ\Gamma is the Gamma function defined by Γ(z)=0xz1ex𝑑x\Gamma(z)=\int_{0}^{\infty}x^{z-1}\mbox{e}^{-x}\,dx, for (z)>0\Re{(z)}>0. We hence obtain

𝔼|logW|q𝟏{W1}<.\displaystyle\mathbb{E}|\log W|^{q}\mathbf{1}_{\{W\leqslant 1\}}<\infty. (3.14)

Next, let’s prove

supn0𝔼|logWn|q<.\sup_{n\geqslant 0}\,\mathbb{E}|\log W_{n}|^{q}<\infty. (3.15)

Notice that for any p1p\geqslant 1, x|logx|p 1{0<x1}x\mapsto|\log x|^{p}\,\mathbf{1}_{\{0<x\leqslant 1\}} is non-negative and convex function. According to Lemma 2.1 in [11], we have

supn0𝔼|logWn|q𝟏{Wn1}=𝔼|logW|q𝟏{W1}.\sup_{n\geqslant 0}\mathbb{E}\left|\log W_{n}\right|^{q}\mathbf{1}_{\{W_{n}\leqslant 1\}}=\mathbb{E}\left|\log W\right|^{q}\mathbf{1}_{\{W\leqslant 1\}}.

With the similar truncation as for 𝔼|logW|q\mathbb{E}\left|\log W\right|^{q} above and by (3.14), we can obtain (3.15). \Box

Lemma 3.4

Suppose that r(1, 2]r\in(1,\,2]. If fΛrf\in\Lambda_{r}, then for any (x,y)2(x,y)\in\mathbb{R}^{2}, there exists t0t_{0} between xx and x+yx+y, such that

|f(x+y)f(x)|C|t0|r1|y|.\displaystyle|f(x+y)-f(x)|\leqslant C|t_{0}|^{r-1}|y|.
Proof 2

Let r(1, 2]r\in(1,\,2], then [r]=1[r]=1. From the definition of Λr\Lambda_{r} and (1.2), if fΛrf\in\Lambda_{r}, then for any tt\in\mathbb{R}, |f(t)f(0)||t|r1;\displaystyle|f^{\prime}(t)-f^{\prime}(0)|\leqslant|t|^{r-1}; which implies

|f(t)|C|t|r1.\displaystyle|f^{\prime}(t)|\leqslant C|t|^{r-1}. (3.16)

From the mean value theorem and the inequality (3.16), for any pair of real numbers (x,y)2(x,y)\in\mathbb{R}^{2}, there exists t0t_{0} between xx and x+yx+y, such that

|f(x+y)f(x)|=|f(t0)||y|C|t0|r1|y|.|f(x+y)-f(x)|=|f^{\prime}(t_{0})||y|\leqslant C|t_{0}|^{r-1}|y|.

\Box

Proof of Theorem 2.3 1

For any n0n\geqslant 0, let an=𝔼logWna_{n}=\mathbb{E}\log W_{n} and Rn=logWnanR_{n}=\log W_{n}-a_{n}. Then (Rn)n0(R_{n})_{n\geqslant 0} becomes a sequence of zero-mean random variables. Moreover, by (3.10), we have the following decomposition, for n0n\geqslant 0,

logZnnμan=Mn+Rn.\displaystyle\log Z_{n}-n\mu-a_{n}=M_{n}+R_{n}. (3.17)

By the triangle inequality and (3.17),

ζr(logZnnμ,Gnσ2)\displaystyle\zeta_{r}(\log Z_{n}-n\mu,\,G_{n\sigma^{2}})\leqslant ζr(Mn+Rn+an,Mn+Rn)+ζr(Mn+Rn,Mn)+ζr(Mn,Gnσ2)\displaystyle\zeta_{r}(M_{n}+R_{n}+a_{n},\,M_{n}+R_{n})+\zeta_{r}(M_{n}+R_{n},\,M_{n})+\zeta_{r}(M_{n},\,G_{n\sigma^{2}})
=\displaystyle= I1,r,n+I2,r,n+I3,r,n,\displaystyle I_{1,r,n}+I_{2,r,n}+I_{3,r,n}, (3.18)

where I1,r,n=ζr(Mn+Rn+an,Mn+Rn)I_{1,r,n}=\zeta_{r}(M_{n}+R_{n}+a_{n},\,M_{n}+R_{n}), I2,r,n=ζr(Mn+Rn,Mn)I_{2,r,n}=\zeta_{r}(M_{n}+R_{n},\,M_{n}) and I3,r,n=ζr(Mn,Gnσ2)I_{3,r,n}=\zeta_{r}(M_{n},\,G_{n\sigma^{2}}). We will estimate respectively I1,r,nI_{1,r,n}, I2,r,nI_{2,r,n} and I3,r,nI_{3,r,n}.

For I3,r,nI_{3,r,n}, recall that under Condition 1, (Mn)n1(M_{n})_{n\geqslant 1} is a centred martingale with i.i.d. martingale differences in 𝕃2+δ\mathbb{L}^{2+\delta} and Var(Mn)=nσ2\mbox{Var}(M_{n})=n\sigma^{2}. According to Theorem 2.1 of Dedecker et al. [4], with p=2+δp=2+\delta, we can obtain for any r[δ, 2+δ]r\in[\delta,\,2+\delta],

ζr(n1/2Mn,Gσ2)Cnδ/2.\displaystyle\zeta_{r}(n^{-1/2}M_{n},\,G_{\sigma^{2}})\leqslant\frac{C}{n^{\delta/2}}. (3.19)

Notice that for any pair of real-valued random variables (X,Y)(X,Y) and real constant aa,

ζr(aX,aY)=|a|rζr(X,Y).\zeta_{r}(aX,aY)=|a|^{r}\zeta_{r}(X,Y). (3.20)

Therefore, the inequality (3.19) implies that for any r[δ, 2+δ]r\in[\delta,\,2+\delta],

I3,r,n=nr/2ζr(n1/2Mn,Gσ2)Cnrδ2.\displaystyle I_{3,r,n}=n^{r/2}\zeta_{r}(n^{-1/2}M_{n},\,G_{\sigma^{2}})\leqslant Cn^{\frac{r-\delta}{2}}. (3.21)

Let’s estimate now I1,r,nI_{1,r,n}. Suppose that r[δ, 2]r\in[\delta,\,2] and fΛrf\in\Lambda_{r}. We will discuss in the following two cases: when r[δ, 1]r\in[\delta,\,1] or r(1, 2]r\in(1,\,2], respectively. If r[δ, 1]r\in[\delta,\,1], then [r]=0[r]=0. By the definition of Λr\Lambda_{r}, we have for any (x,y)2(x,y)\in\mathbb{R}^{2},

|f(x)f(y)||xy|r.\displaystyle|f(x)-f(y)|\leqslant|x-y|^{r}. (3.22)

Since Mn+RnM_{n}+R_{n} and Mn+Rn+anM_{n}+R_{n}+a_{n} are both n\mathcal{F}_{n}-measurable, by using (3.22), we have for any r[δ, 1]r\in[\delta,\,1],

I1,r,n\displaystyle I_{1,r,n}\leqslant supfΛr𝔼[𝔼(|f(Mn+Rn+an)f(Mn+Rn)||n)]|an|r.\displaystyle\sup_{f\in\Lambda_{r}}\mathbb{E}\,[\mathbb{E}(|f(M_{n}+R_{n}+a_{n})-f(M_{n}+R_{n})|\;|\mathcal{F}_{n})]\leqslant|a_{n}|^{r}.

Since for r[δ, 1]r\in[\delta,\,1], the function xxrx\mapsto x^{r} is increasing on (0,)(0,\infty), and by using Lemma 3.3, we get

I1,r,n|𝔼logWn|r\displaystyle I_{1,r,n}\leqslant|\mathbb{E}\log W_{n}|^{r}\leqslant (𝔼|logWn|)r\displaystyle(\mathbb{E}|\log W_{n}|)^{r}
<\displaystyle< .\displaystyle\infty. (3.23)

If r(1, 2]r\in(1,\,2], then [r]=1[r]=1. By Lemma 3.4, we have there exists Y1,nY_{1,n} between Mn+RnM_{n}+R_{n} and Mn+Rn+anM_{n}+R_{n}+a_{n}, such that

𝔼(|f(Mn+Rn+an)f(Mn+Rn)||n)C|an||Y1,n|r1.\mathbb{E}(|f(M_{n}+R_{n}+a_{n})-f(M_{n}+R_{n})|\,|\mathcal{F}_{n})\leqslant\,C\,|a_{n}|\,|Y_{1,n}|^{r-1}.

Hence,

I1,r,n\displaystyle I_{1,r,n}\leqslant supfΛr𝔼|f(Mn+Rn+an)f(Mn+Rn)|\displaystyle\sup_{f\in\Lambda_{r}}\mathbb{E}|f(M_{n}+R_{n}+a_{n})-f(M_{n}+R_{n})|
\displaystyle\leqslant C|an|𝔼|Y1,n|r1.\displaystyle C\,|a_{n}|\,\mathbb{E}|Y_{1,n}|^{r-1}. (3.24)

Now, let’s study 𝔼|Y1,n|r1\mathbb{E}|Y_{1,n}|^{r-1}, with r1(0, 1]r-1\in(0,\,1]. Since for δ(0, 1]\delta\in(0,\,1], the function xxδx\mapsto x^{\delta} is increasing on (0,)(0,\infty), we have

𝔼|Y1,n|r1\displaystyle\mathbb{E}|Y_{1,n}|^{r-1}\leqslant max{𝔼|Mn+Rn+an|r1,𝔼|Mn+Rn|r1}\displaystyle\max\{\,\mathbb{E}|M_{n}+R_{n}+a_{n}|^{r-1},\,\mathbb{E}|M_{n}+R_{n}|^{r-1}\,\}
\displaystyle\leqslant 𝔼|Mn+Rn+an|r1+𝔼|Mn+Rn|r1.\displaystyle\mathbb{E}|M_{n}+R_{n}+a_{n}|^{r-1}+\mathbb{E}|M_{n}+R_{n}|^{r-1}. (3.25)

For r(1, 2]r\in(1,\,2], applying Lyapunov’s inequality to the two terms on the right-hand side, we obtain

𝔼|Mn+Rn+an|r1+𝔼|Mn+Rn|r1(𝔼|Mn+Rn+an|)r1+(𝔼|Mn+Rn|)r1.\displaystyle\mathbb{E}|M_{n}+R_{n}+a_{n}|^{r-1}+\mathbb{E}|M_{n}+R_{n}|^{r-1}\leqslant(\mathbb{E}|M_{n}+R_{n}+a_{n}|)^{r-1}+(\mathbb{E}|M_{n}+R_{n}|)^{r-1}. (3.26)

For the first term above, from the assumption (2.6), the LLN for SnS_{n} and Lemma 3.3, we have

(𝔼|Mn+Rn+an|)r1=(𝔼|Snnμ+logWn|)r1\displaystyle(\mathbb{E}|M_{n}+R_{n}+a_{n}|)^{r-1}=(\mathbb{E}|S_{n}-n\mu+\log W_{n}|)^{r-1}\leqslant (𝔼|Snnμ|+𝔼|logWn|)r1\displaystyle(\mathbb{E}|S_{n}-n\mu|+\mathbb{E}|\log W_{n}|)^{r-1}
<\displaystyle< .\displaystyle\infty. (3.27)

With the same arguments, we can obtain

(𝔼|Mn+Rn|)r1<.\displaystyle(\mathbb{E}|M_{n}+R_{n}|)^{r-1}<\infty. (3.28)

Using Lemma 3.3 again, we also have

|an|=|𝔼logWn|𝔼|logWn|<.\displaystyle|a_{n}|=|\mathbb{E}\log W_{n}|\leqslant\mathbb{E}|\log W_{n}|<\infty. (3.29)

Combining (1) for the case r[δ, 1]r\in[\delta,\,1] with (1) - (3.29) for the case r(1, 2]r\in(1,\,2], we can conclude that for any r[δ, 2]r\in[\delta,\,2],

I1,r,nC.\displaystyle I_{1,r,n}\leqslant C. (3.30)

Now, let’s deal with I2,r,nI_{2,r,n}. By Lemma 3.3, we have supn0Rn2+δ<\sup_{n\geqslant 0}\|R_{n}\|_{2+\delta}<\infty. Taking into account Remark 5.1 and applying Items 1 and 2 of Lemma 5.2 in Dedecker et al. [4] with p=δ+2p=\delta+2 therein, then we can obtain for any r[δ, 2]r\in[\delta,\,2],

I2,r,n=ζr(Mn+Rn,Mn)Cnrδ2.\displaystyle I_{2,r,n}=\zeta_{r}(M_{n}+R_{n},\,M_{n})\leqslant Cn^{\frac{r-\delta}{2}}. (3.31)

Now, combining (1), (3.21), (3.30) and (3.31), we have for any r[δ, 2]r\in[\delta,\,2],

ζr(logZnnμ,Gnσ2)Cnrδ2.\displaystyle\zeta_{r}(\log Z_{n}-n\mu,\;G_{n\sigma^{2}})\leqslant Cn^{\frac{r-\delta}{2}}.

Therefore, the theorem is derived by taking into account

ζr(logZnnμnσ,𝒩)=(1nσ)rζr(logZnnμ,Gnσ2).\zeta_{r}\left(\frac{\log Z_{n}-n\mu}{\sqrt{n}\sigma},\,\mathcal{N}\right)=\left(\frac{1}{\sqrt{n}\sigma}\right)^{r}\zeta_{r}(\log Z_{n}-n\mu,\,G_{n\sigma^{2}}).

\Box

References

  • Athreya and Karlin [1971a] Athreya, K.B., Karlin, S., 1971a. Branching processes with random environments. \@slowromancapii@. Limit theorems. Ann. Math. Statist. 42, 1843–1858. https://doi.org/10.1214/aoms/1177693051
  • Athreya and Karlin [1971b] Athreya, K.B., Karlin, S., 1971b. On branching processes with random environments. \@slowromancapi@. Extinction probabilities. Ann. Math. Statist., 42, 1499–1520. https://doi.org/10.1214/aoms/1177693150
  • Birkner et al. [2005] Birkner, M., Geiger, J., Kersting, G., 2005. Branching Processes in Random Environment — A View on Critical and Subcritical Cases, in: Deuschel, J.D., Greven, A. (Eds.), Interacting Stochastic Systems. Springer, Berlin, Heidelberg.
  • Dedecker et al. [2009] Dedecker, J., Merlevède, F., Rio, E., 2009. Rates of convergence for minimal distances in the central limit theorem under projective criteria. Electron. J. Probab., 14, 978–1011. https://doi.org/10.1214/EJP.v14-648
  • Grama et al. [2019] Grama, I., Lauvergnat, R., Le Page, E., 2019. The survival probability of critical and subcritical branching processes in finite state space Markovian environment. Stochastic Process. Appl., 129(7), 2485–2527. https://doi.org/10.1016/j.spa.2018.07.016
  • Grama et al. [2022] Grama, I., Lauvergnat, R., Le Page, E., 2022. Limit theorems for critical branching processes in a finite-state-space Markovian environment. Advances in Applied Probability, 54(1), 111-140. https://doi.org/10.1214/aoms/1177693150
  • Grama et al. [2023] Grama, I., Liu, Q., Miqueu, E., 2023. Asymptotics of the distribution and harmonic moments for a supercritical branching process in a random environment. Ann. Inst. H. Poincaré Probab. Statist., 59(4), 1934–1950. DOI: 10.1214/22-AIHP1318
  • Guivarc’h and Liu [2001] Guivarc’h, Y., Liu, Q., 2001. Propriété asymptotiques des processus de branchement en environnement aléatoire. C. R. Acad. Sci. Série I, 332(4), 339–344. https://doi.org/10.1016/S0764-4442(00)01783-3
  • Gulrajani et al. [2017] Gulrajani, I., Ahmed, F., Arjovsky, M., Dumoulin, V., Courville, A. C., 2017. Improved training of Wasserstein GANs, in: I. Guyon, U. V. Luxburg, S. Bengio, H. Wallach, R. Fergus, S. Vishwanathan and R. Garnett (Eds), Advances in Neural Information Processing Systems 30, Curran Associates, Inc., pp. 5769–5779.
  • Hall and Heyde [1980] Hall, P., Heyde, C.C., 1980. Martingale limit theory and its Application, A volume in Probability and Mathematical Statistics. Academic Press.
  • Huang and Liu [2012] Huang, C., Liu, Q., 2012. Moments, moderate and large deviations for a branching process in a random environment. Stochastic Process. Appl., 122, 522–545. https://doi.org/10.1016/j.spa.2011.09.001
  • Huang and Liu [2014] Huang, C., Liu, Q., 2014. Convergence in 𝕃p\mathbb{L}^{p} and its exponential rate for a branching process in a random environment. Electron. J. Probab., 19(104), 1–22. https://doi.org/10.1214/EJP.v19-3388
  • Kantorovich and Rubinstein [1958] Kantorovich, L. V., Rubinstein, G. Sh., 1958. Sur un espace de fonctions complètement additives. Vestnik Leningrad Univ., 13 (Ser. Mat. Astron. 2), 52–59.
  • Le Page and Ye [2010] Le Page, E., Ye, Y., 2010. The survival probability of a critical branching process in a Markovian random environment. Comptes Rendus Mathématique, 348(5-6), 301–304. https://doi.org/10.1016/j.crma.2010.01.014
  • Li et al. [2014] Li, Y., Liu, Q., Gao, Z., Wang, H., 2014. Asymptotic properties of supercritical branching processes in random environments. Front. Math. China, 9(4), 737–751. https://doi.org/10.1007/s11464-014-0397-z
  • Ni et al. [2009] Ni, K., Bresson, X., Chan, T., Esedoglu, S., 2009. Local histogram based segmentation using the Wasserstein distance, International Journal of Computer Vision, 84, 97–111.
  • Rio [1998] Rio, E., 1998. Distances minimales et distances idéales. Comptes Rendus de l’Académie des Sciences - Series \@slowromancapi@ - Mathematics, 326(9), 1127–1130. https://doi.org/10.1016/S0764-4442(98)80074-8
  • Shao [1993] Shao, Q., 1993. Almost sure invariance principles for mixing sequences of random variables. Stochastic Process. Appl., 48(2), 319–334. https://doi.org/10.1016/0304-4149(93)90051-5
  • Smith and Wilkinson [1969] Smith, W.L., Wilkinson, W.E., 1969. On branching processes in random environments. Ann. Math. Stat., 40(3), 814–827. https://doi.org/10.1214/aoms/1177697589
  • Solomon et al. [2015] Solomon, J., De Goes, F., Peyré, G., Cuturi, M., Butscher, A., Nguyen, A., Du, T., and Guibas, L. 2015. Convolutional Wasserstein distances: efficient optimal transportation on geometric domains. ACM Transactions on Graphics (TOG), 34(4), 1–-11. https://doi.org/10.1145/2766963
  • Tanny [1977] Tanny, D., 1977. Limit theorems for branching processes in a random environment. The Annals of Probability, 5(1), 100–116. https://doi.org/10.1214/aop/1176995894
  • Tanny [1988] Tanny, D., 1988. A necessary and sufficient condition for a branching process in a random environment to grow like the product of its means. Stochastic Process. Appl., 28(1), 123–139. https://doi.org/10.1016/0304-4149(88)90070-1
  • Wang et al. [2021] Wang, S., Cai, T.T., Li, H., 2021 Optimal estimation of Wasserstein distance on a tree with an application to microbiome studies. Journal of the American Statistical Association, 116(535), 1237–1253. https://doi.org/10.1080/01621459.2019.1699422
  • Wu et al. [2020] Wu, S., Ma, X., Sang, H., Fan, X., 2020. A Berry-Esseen bound of order 1/n1/\sqrt{n} for martingales. Comptes Rendus Mathématique, 358(6): 701–712. https://doi.org/10.5802/crmath.81