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

Moderate deviation principles for kernel estimator of invariant density in bifurcating Markov chains models.

S. Valère Bitseki Penda S. Valère Bitseki Penda, IMB, CNRS-UMR 5584, Université Bourgogne Franche-Comté, 9 avenue Alain Savary, 21078 Dijon Cedex, France. simeon-valere.bitseki-penda@u-bourgogne.fr
Abstract.

Bitseki and Delmas (2021) have studied recently the central limit theorem for kernel estimator of invariant density in bifurcating Markov chains models. We complete their work by proving a moderate deviation principle for this estimator. Unlike the work of Bitseki and Gorgui (2021), it is interesting to see that the distinction of the two regimes disappears and that we are able to get moderate deviation principle for large values of the ergodic rate. It is also interesting and surprising to see that for moderate deviation principle, the ergodic rate begins to have an impact on the choice of the bandwidth for values smaller than in the context of central limit theorem studied by Bitseki and Delmas (2021).

Keywords: Bifurcating Markov chains, bifurcating auto-regressive process, binary trees, density estimation.

Mathematics Subject Classification (2020): 62G05, 62F12, 60F10, 60J80,

1. Introduction

The study of bifurcating Markov chains (BMCs, for short) models has taken a special place in the literature these last years due to their links with the study of the cell dynamics (see for e.g. [6, 10, 13, 16, 17]). The first model of BMC, named “symmetric” bifurcating autoregressive process (BAR, for short) were introduced by Cowan and Staudte [9] in order to understand the cell division mechanisms of Escherichia Coli (E. Coli, for short). E. Coli is a rod shaped bacterium which reproduces by dividing in two, thus producing two new cells. One of type 1 which has the old end of the mother and the other of type 0 which has the new end of the mother. The age of a cell is thus given by the age of its old pole in the sense of the number of divisions from which this pole exists. This cell division mechanism raises several questions, among other that of the symmetry of the division. In order to give a rigorous answer to this question, Guyon [16] has developed and studied the theory of BMCs. We note that to the best of our knowledge, the term BMC appears for the first time in the work of [1]. In particular, Guyon has studied an extension of the model introduced by Cowan and Staudte, named “asymmetric” BAR. In the conclusion of his study, Guyon concludes that aging has an impact on cell reproduction. We note that an extension of the model proposed by Guyon, named nonlinear BAR (NBAR, for short) were studied by Bitseki and Olivier in [6]. Another question of interest related to cell division is estimating the division rate at which cells divide. This question has been tackled recently in the work of Doumic & al. [13] and Hoffman & Marguet [17]. In all the previous work, the behaviour and the definition of parameters of interest are associated with the density of the invariant probability of an auxiliary Markov chain (see below for a precise definition). The estimation of this invariant density has recently been the subject of several studies. One can cite [5, 8] where adaptive methods have been proposed for the estimation of this invariant density. More recently, Bitseki and Delmas [2] have studied central limit theorem for kernel estimators of this invariant density. Our main objective in this paper is to complete the previous study by establishing a moderate deviation principle for these kernel estimators. Before going any further, let us recall the definition of the main concepts that we will use and study.

2. The model of bifurcating Markov chain and definition of the estimators

2.1. The regular binary tree associated to BMC models

We denote by \mathbb{N} (resp. \mathbb{N}^{*}) the space of (resp. positive) natural integers. We set 𝕋0=𝔾0={}\mathbb{T}_{0}=\mathbb{G}_{0}=\{\emptyset\}, 𝔾k={0,1}k\mathbb{G}_{k}=\{0,1\}^{k} and 𝕋k=0rk𝔾r\mathbb{T}_{k}=\bigcup_{0\leq r\leq k}\mathbb{G}_{r} for kk\in{\mathbb{N}}^{*}, and 𝕋=r𝔾r\mathbb{T}=\bigcup_{r\in{\mathbb{N}}}\mathbb{G}_{r}. The set 𝔾k\mathbb{G}_{k} corresponds to the kk-th generation, 𝕋k\mathbb{T}_{k} to the tree up to the kk-th generation, and 𝕋\mathbb{T} the complete binary tree. One can see that the genealogy of the cells is entirely described by 𝕋\mathbb{T} (each vertex of the tree designates an individual). For i𝕋i\in\mathbb{T}, we denote by |i||i| the generation of ii (|i|=k|i|=k if and only if i𝔾ki\in\mathbb{G}_{k}) and iA={ij;jA}iA=\{ij;j\in A\} for A𝕋A\subset\mathbb{T}, where ijij is the concatenation of the two sequences i,j𝕋i,j\in\mathbb{T}, with the convention that i=i=i\emptyset i=i\emptyset=i. For A𝕋A\subset\mathbb{T}, we denote by |A||A| the number of elements of AA. Note that for all n,n\in\mathbb{N}, |𝔾n|=2n|\mathbb{G}_{n}|=2^{n} and |𝕋n|=2n+11.|\mathbb{T}_{n}|=2^{n+1}-1.

2.2. The probability kernels associated to BMC models

For our convenience, we set S=dS=\mathbb{R}^{d}, d1d\geq 1 and SS is equipped with the Borel sigma-algebra 𝒮{\mathscr{S}}. For any qq\in\mathbb{N}^{*}, we denote by (Sq){\mathcal{B}}(S^{q}) (resp. b(Sq){\mathcal{B}}_{b}(S^{q}), resp. 𝒞b(Sq){\mathcal{C}}_{b}(S^{q})) the space of (resp. bounded, resp. bounded continuous ) -\mathbb{R}\text{-}valued measurable functions defined on SqS^{q}. For all qq\in\mathbb{N}^{*}, we set 𝒮q=𝒮𝒮{\mathscr{S}}^{\otimes q}={\mathscr{S}}\otimes\ldots\otimes{\mathscr{S}}. Let 𝒫{\mathcal{P}} be a probability kernel on (S,𝒮2)(S,{\mathscr{S}}^{\otimes 2}), that is: 𝒫(,A){\mathcal{P}}(\cdot,A) is measurable for all A𝒮2A\in{\mathscr{S}}^{\otimes 2}, and 𝒫(x,){\mathcal{P}}(x,\cdot) is a probability measure on (S2,𝒮2)(S^{2},{\mathscr{S}}^{\otimes 2}) for all xSx\in S. For any gb(S3)g\in{\mathcal{B}}_{b}(S^{3}) and hb(S2)h\in{\mathcal{B}}_{b}(S^{2}), we set for xSx\in S:

(1) (𝒫g)(x)=S2g(x,y,z)𝒫(x,dy,dz)and(𝒫h)(x)=S2h(y,z)𝒫(x,dy,dz).({\mathcal{P}}g)(x)=\int_{S^{2}}g(x,y,z)\;{\mathcal{P}}(x,{\rm d}y,{\rm d}z)\quad\text{and}\quad({\mathcal{P}}h)(x)=\int_{S^{2}}h(y,z)\;{\mathcal{P}}(x,{\rm d}y,{\rm d}z).

We define (𝒫g)({\mathcal{P}}g) (resp. (𝒫h)({\mathcal{P}}h)), or simply PgPg for g(S3)g\in{\mathcal{B}}(S^{3})(resp. 𝒫h{\mathcal{P}}h for h(S2)h\in{\mathcal{B}}(S^{2})), as soon as the corresponding integral (1) is well defined, and we have that 𝒫g{\mathcal{P}}g and 𝒫h{\mathcal{P}}h belong to (S){\mathcal{B}}(S). We denote by 𝒫0{\mathcal{P}}_{0}, 𝒫1{\mathcal{P}}_{1} and 𝒬{\mathcal{Q}} respectively the first and the second marginal of 𝒫{\mathcal{P}}, and the mean of 𝒫0{\mathcal{P}}_{0} and 𝒫1{\mathcal{P}}_{1}, that is, for all xSx\in S and B𝒮B\in\mathcal{S}

𝒫0(x,B)=𝒫(x,B×S),𝒫1(x,B)=𝒫(x,S×B) and𝒬=(𝒫0+𝒫1)2.{\mathcal{P}}_{0}(x,B)={\mathcal{P}}(x,B\times S),\quad{\mathcal{P}}_{1}(x,B)={\mathcal{P}}(x,S\times B)\quad\text{ and}\quad{\mathcal{Q}}=\frac{({\mathcal{P}}_{0}+{\mathcal{P}}_{1})}{2}.

Now let us give a precise definition of bifurcating Markov chain.

Definition 2.1 (Bifurcating Markov Chains, see [16, 2]).

We say a stochastic process indexed by 𝕋\mathbb{T}, X=(Xi,i𝕋)X=(X_{i},i\in\mathbb{T}), is a bifurcating Markov chain (BMC) on a measurable space (S,𝒮)(S,{\mathscr{S}}) with initial probability distribution ν\nu on (S,𝒮)(S,{\mathscr{S}}) and probability kernel 𝒫{\mathcal{P}} on S×𝒮2S\times{\mathscr{S}}^{\otimes 2} if:

  • -

    (Initial distribution.) The random variable XX_{\emptyset} is distributed as ν\nu.

  • -

    (Branching Markov property.) For any sequence (gi,i𝕋)(g_{i},i\in\mathbb{T}) of functions belonging to b(S3){\mathcal{B}}_{b}(S^{3}) and for all k0k\geq 0, we have

    𝔼[i𝔾kgi(Xi,Xi0,Xi1)|σ(Xj;j𝕋k)]=i𝔾k𝒫gi(Xi).{\mathbb{E}}\Big{[}\prod_{i\in\mathbb{G}_{k}}g_{i}(X_{i},X_{i0},X_{i1})|\sigma(X_{j};j\in\mathbb{T}_{k})\Big{]}=\prod_{i\in\mathbb{G}_{k}}{\mathcal{P}}g_{i}(X_{i}).

Following [16], we introduce an auxiliary Markov chain Y=(Yn,n)Y=(Y_{n},n\in{\mathbb{N}}) on (S,𝒮)(S,{\mathscr{S}}) with Y0=X1Y_{0}=X_{1} and transition probability 𝒬{\mathcal{Q}}. The chain (Yn,n)(Y_{n},n\in\mathbb{N}) corresponds to a random lineage taken in the population. We shall write 𝔼x{\mathbb{E}}_{x} when X=xX_{\emptyset}=x (i.e. the initial distribution ν\nu is the Dirac mass at xSx\in S). We will assume that the Markov chain YY is ergodic and we denote by μ\mu its invariant probability measure. Asymptotic and non-asymptotic behaviour of BMCs are strongly related to the knowledge of μ\mu. In particular, Guyon has proved that if YY is ergodic, then for all f𝒞b(S)f\in{\mathcal{C}}_{b}(S),

|𝔸n|1u𝔸nf(Xu)n+μ,fin probability,where 𝔸n{𝔾n,𝕋n}.|{\mathbb{A}}_{n}|^{-1}\sum_{u\in{\mathbb{A}}_{n}}f(X_{u})\underset{n\rightarrow+\infty}{\xrightarrow{\hskip 21.33955pt}}\langle\mu,f\rangle\quad\text{in probability,}\quad\text{where ${\mathbb{A}}_{n}\in\{\mathbb{G}_{n},\mathbb{T}_{n}\}.$}

But in most cases, the invariant probability μ\mu is unknown, so its estimation from the data is of great interest. For that purpose, we do the following assumption.

Assumption 2.2.

The transition kernel 𝒫{\mathcal{P}} has a density, still denoted by 𝒫{\mathcal{P}}, with respect to the Lebesgue measure.

Remark 2.3.

Assumption 2.2 implies that the transition kernel 𝒬{\mathcal{Q}} has a density, still denoted by 𝒬{\mathcal{Q}}, with respect to the Lebesgue measure. More precisely, we have 𝒬(x,y)=21S(𝒫(x,y,z)+𝒫(x,z,y))𝑑z.{\mathcal{Q}}(x,y)=2^{-1}\int_{S}({\mathcal{P}}(x,y,z)+{\mathcal{P}}(x,z,y))dz. This implies in particular that the invariant probability μ\mu has a density, still denoted by μ\mu, with respect to the Lebesgue measure (for more details, we refer for e.g. to [14], chap 6).

2.3. Kernel estimator of the invariant density μ\mu

Recall that 𝔸n{𝔾n,𝕋n}{\mathbb{A}}_{n}\in\{\mathbb{G}_{n},\mathbb{T}_{n}\} and S=dS=\mathbb{R}^{d}, d1d\geq 1. Assume we observe 𝕏n=(Xu,u𝔸n)\mathbb{X}^{n}=(X_{u},u\in{\mathbb{A}}_{n}). Let (hn,n)(h_{n},n\in\mathbb{N}) be a sequence of positive numbers which converges to 0 as nn goes to infinity. We will simply write hh for hnh_{n} if there is no ambiguity. Let the kernel function K:SK:S\rightarrow\mathbb{R} such that SK(x)𝑑x=1.\int_{S}K(x)dx=1. Then, for all xS,x\in S, we propose to estimate μ(x)\mu(x) by

(2) μ^𝔸n(x)=|𝔸n|1hnd/2u𝔸nKhn(xXu),\widehat{\mu}_{{\mathbb{A}}_{n}}(x)=|{\mathbb{A}}_{n}|^{-1}h_{n}^{-d/2}\sum_{u\in{\mathbb{A}}_{n}}K_{h_{n}}(x-X_{u}),

where Khn()=hnd/2K(hn).K_{h_{n}}(\cdot)=h_{n}^{-d/2}K(h_{n}\cdot). These estimators are strongly inspired from [18, 21, 22]. They have been studied in [13, 8] (non asymptotic studies) and in [2] (central limit theorem).

2.4. Moderate deviation principle and related topics

Our aim is to study moderate deviation principles for the estimators defined in (2). Before we proceed, let us introduce the notion of moderate deviation principle. We give the definition in a general setting. Let (Zn)n0(Z_{n})_{n\geq 0} be a sequence of random variables with values in SS endowed with its Borel σ\sigma-field 𝒮{\mathscr{S}} and let (sn)n0(s_{n})_{n\geq 0} be a positive sequence that converges to ++\infty. We assume that Zn/snZ_{n}/s_{n} converges in probability to 0 and that Zn/snZ_{n}/\sqrt{s_{n}} converges in distribution to a centered Gaussian law. Let I:S+I:S\rightarrow\mathbb{R}^{+} be a lower semicontinuous function, that is for all c>0c>0 the sub-level set {xS,I(x)c}\{x\in S,I(x)\leq c\} is a closed set. Such a function II is called rate function and it is called good rate function if all its sub-level sets are compact sets. Let (bn)n0(b_{n})_{n\geq 0} be a positive sequence such that bn+b_{n}\rightarrow+\infty and bn/sn0b_{n}/\sqrt{s_{n}}\rightarrow 0 as nn goes to ++\infty.

Definition 2.4 (Moderate deviation principle, MDP).

We say that Zn/(bnsn)Z_{n}/(b_{n}\sqrt{s_{n}}) satisfies a moderate deviation principle on SS with speed bn2b_{n}^{2} and rate function II if, for any A𝒮A\in{\mathscr{S}},

infxÅI(x)lim infn+1bn2log(ZnbnsnA)lim supn+1bn2log(ZnbnsnA)infxA¯I(x),\displaystyle-\inf_{x\in\mathring{A}}I(x)\leq\liminf_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\mathbb{P}\big{(}\frac{Z_{n}}{b_{n}\sqrt{s_{n}}}\in A\big{)}\leq\limsup_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\mathbb{P}\big{(}\frac{Z_{n}}{b_{n}\sqrt{s_{n}}}\in A\big{)}\leq-\inf_{x\in\bar{A}}I(x),

where Å\mathring{A} and A¯\bar{A} denote respectively the interior and the closure of AA.

The following two concepts are closely related to the theory of MDP: super-exponential convergence and exponential equivalence. Let (Zn,n)(Z_{n},n\in\mathbb{N}), (Wn,n)(W_{n},n\in\mathbb{N}) be sequences of random variables and ZZ a random variable with value in a metric space (S,d)(S,d).

Definition 2.5 (Super-exponential convergence).

We say that (Zn)n0(Z_{n})_{n\geq 0} converges   (bn2)-(b_{n}^{2})\text{-} super-exponentially fast in probability to ZZ and we note Znbn2superexpZZ_{n}\xRightarrow[b_{n}^{2}]{\rm superexp}Z if, for all δ>0\delta>0,

lim supn+1bn2log(d(Zn,Z)>δ)=.\limsup_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\mathbb{P}\big{(}d(Z_{n},Z)>\delta\big{)}=-\infty.
Definition 2.6 (Exponential equivalence, see [11], Chap 4).

We say that (Zn)n0(Z_{n})_{n\geq 0} and (Wn)n0(W_{n})_{n\geq 0} are (bn2)n0(b_{n}^{2})_{n\geq 0}-exponentially equivalent and we note Znbn2superexpWnZ_{n}\mathrel{\underset{b_{n}^{2}}{\overset{{\rm superexp}}{\scalebox{2.0}[1.0]{$\sim$}}}}W_{n} if for any δ>0\delta>0,

lim supn+1bn2log(d(Zn,Wn)>δ)=.\limsup_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\mathbb{P}\big{(}d(Z_{n},W_{n})>\delta\big{)}=-\infty.
Remark 2.7.

Note that for a determininistic sequence that converges to some limit \ell, it also converges (bn2)-(b_{n}^{2})\text{-}superexponentially fast to \ell for any rate bnb_{n}. We also note that if (Zn)n0(Z_{n})_{n\geq 0} and (Wn)n0(W_{n})_{n\geq 0} are (bn2)n0(b_{n}^{2})_{n\geq 0}-exponentially equivalent and if (Zn)n0(Z_{n})_{n\geq 0} satisfies a MDP, then (Wn)n0(W_{n})_{n\geq 0} satisfies the same MDP (for more details, see for e.g [11], Chap 4).

The following result give a sufficient condition for super-exponential convergence of a sequence of random variables.

Remark 2.8.

We assume that (S,d)(S,d) is a metric space. Let (Zn)n\left(Z_{n}\right)_{n\in\mathbb{N}} be a sequence of random variables with values in SS, ZZ a random variable with values in SS. So if d(Zn,Z)d(Z_{n},Z) is upper-bounded by a deterministic sequence which converges to 0, then, for all sequence (bn,n)(b_{n},n\in\mathbb{N}) converging to ++\infty, Znbn2superexpZZ_{n}\xRightarrow[b_{n}^{2}]{\rm superexp}Z.

The moderate deviation principle has been proved in the i.i.d. setting for kernel density estimator, see for e.g. Gao [15], Mokkadem & al. [20]. We refer also to [19] where Mokkaddem and Pelletier have constructed confidence bands for probability densities based on moderate deviation principles. In this paper, we will establish moderate deviation principle for μ^𝔸n(x)\widehat{\mu}_{{\mathbb{A}}_{n}}(x) following the martingale approach developed in [2]. We will need the following assumption.

Assumption 2.9.

There exists a positive real number MM and α(0,1)\alpha\in(0,1) such that for all fb(S)f\in{\mathcal{B}}_{b}(S):

(3) |𝒬nfμ,f|Mαnffor all n.|{\mathcal{Q}}^{n}f-\langle\mu,f\rangle|\leq M\,\alpha^{n}\|f\|_{\infty}\quad\text{for all $n\in{\mathbb{N}}$.}
Remark 2.10.

Assumption 2.9 is for example satisfy for nonlinear bifurcating autoregressive process under mild hypotheses on the autoregression functions (see [7] Lemma 9 for more details).

The others assumptions we will need are based on the following bias-variance type decomposition of the estimator μ^𝔸n(x)\widehat{\mu}_{{\mathbb{A}}_{n}}(x):

(4) μ^𝔸n(x)μ(x)=Bhn(x)+V𝔸n,hn(x),\widehat{\mu}_{{\mathbb{A}}_{n}}(x)-\mu(x)=B_{h_{n}}(x)+V_{{\mathbb{A}}_{n},h_{n}}(x),

where for h>0h>0 and 𝔸𝕋{\mathbb{A}}\subset\mathbb{T} finite:

Bh(x)=hd/2Khμ(x)μ(x)andV𝔸,h(x)=|𝔸|1hd/2u𝔸(Kh(xXu)Khμ(x)),B_{h}(x)=h^{-d/2}K_{h}\star\mu(x)-\mu(x)\quad\text{and}\quad V_{{\mathbb{A}},h}(x)=|{\mathbb{A}}|^{-1}h^{-d/2}\sum_{u\in{\mathbb{A}}}\Big{(}K_{h}(x-X_{u})-K_{h}\star\mu(x)\Big{)},

and for h>0h>0 and u𝕋u\in\mathbb{T}, we set:

Khμ(x)=𝔼μ[Kh(xXu)]=SKh(xy)μ(y)𝑑y.K_{h}\star\mu(x)=\mathbb{E}_{\mu}[K_{h}(x-X_{u})]=\int_{S}K_{h}(x-y)\mu(y)\,dy.

To study the variance term V𝔸n,hn(x)V_{{\mathbb{A}}_{n},h_{n}}(x), we will introduce a more general sequence of functions (see Section 3.2).

The following assumptions on the kernel, the bandwidth and the regularity of the unknown density function are usual. Recall S=dS={\mathbb{R}}^{d} with d1d\geq 1.

Assumption 2.11 (Regularity of the kernel function and the bandwidth).

  1. (i)

    The kernel function K(S)K\in{\mathcal{B}}(S) satisfies:

    K<+,K1<+,K2<+,SK(x)𝑑x=1andlim|x|+|x|K(x)=0.\mathop{\parallel\!K\!\parallel}\nolimits_{\infty}<+\infty,\,\,\mathop{\parallel\!K\!\parallel}\nolimits_{1}<+\infty,\,\,\mathop{\parallel\!K\!\parallel}\nolimits_{2}<+\infty,\,\,\int_{S}\!K(x)\,dx=1\quad\text{and}\quad\lim_{|x|\rightarrow+\infty}|x|K(x)=0.
  2. (ii)

    There exists γ(0,1/d)\gamma\in(0,1/d) such that the bandwidth (hn,n)(h_{n},n\in\mathbb{N}) are defined by hn=2nγh_{n}=2^{-n\gamma}.

Assumption 2.12 (Further regularity on the density μ\mu, the kernel function and the bandwidths).

Suppose that there exists an invariant probability measure μ\mu of 𝒬{\mathcal{Q}} and that Assumptions 2.2 and 2.11 hold. We assume there exists s>0s>0 such that the following hold:

  1. (i)

    The density μ\mu belongs to the (isotropic) Hölder class of order (s,,s)d(s,\ldots,s)\in\mathbb{R}^{d}: The density μ\mu admits partial derivatives with respect to xjx_{j}, for all j{1,d}j\in\{1,\ldots d\}, up to the order s\lfloor s\rfloor and there exists a finite constant L>0L>0 such that for all x=(x1,,xd),dx=(x_{1},\ldots,x_{d}),\in\mathbb{R}^{d}, tt\in{\mathbb{R}} and j{1,,d}j\in\{1,\ldots,d\}:

    |sμxjs(xj,t)sμxjs(x)|L|xjt|{s},\left|\frac{\partial^{\lfloor s\rfloor}\mu}{\partial x_{j}^{\lfloor s\rfloor}}(x_{-j},t)-\frac{\partial^{\lfloor s\rfloor}\mu}{\partial x_{j}^{\lfloor s\rfloor}}(x)\right|\leq L|x_{j}-t|^{\{s\}},

    where (xj,t)(x_{-j},t) denotes the vector xx where we have replaced the jthj^{th} coordinate xjx_{j} by tt, with the convention 0μ/xj0=μ{\partial^{0}\mu}/{\partial x_{j}^{0}}=\mu.

  2. (ii)

    The kernel KK is of order (s,,s)d(\lfloor s\rfloor,\ldots,\lfloor s\rfloor)\in\mathbb{N}^{d}: We have d|x|sK(x)𝑑x<\int_{\mathbb{R}^{d}}|x|^{s}K(x)\,dx<\infty and xjkK(x)𝑑xj=0\int_{\mathbb{R}}x^{k}_{j}\,K(x)\,dx_{j}=0 for all k{1,,s}k\in\{1,\ldots,\lfloor s\rfloor\} and j{1,,d}j\in\{1,\ldots,d\}.

For α>1/2\alpha>1/2, we shall also assume the following.

Assumption 2.13.

Keeping the same notations as in (ii) of Assumption 2.11, we further assume that Assumption 2.9 holds with

(5) limn+(21dγα)n=0.\lim_{n\rightarrow+\infty}(2^{1-d\gamma}\alpha)^{n}=0.
Remark 2.14.

As consequence of Assumption 2.13 and (ii) of Assumption 2.11, for moderate deviation principle, the ergodicity rate α\alpha begins to have an impact on the choice of the bandwidth for α>1/2.\alpha>1/2. This is out of step with the central limit theorem where the ergodicity rate α\alpha begins to have an impact on the choice of the bandwidth for α>1/2\alpha>1/\sqrt{2} (see [2] for more details).

In the sequel, we will consider the positive sequence (bn,n)(b_{n},n\in\mathbb{N}) such that:

(6) limn+bn=+;limn+n3/2bn|𝔾n|hnd=0;limn+bn|𝔾n|hn2s+d=+,\lim_{n\rightarrow+\infty}b_{n}=+\infty;\quad\lim_{n\rightarrow+\infty}\frac{n^{3/2}\,b_{n}}{\sqrt{|\mathbb{G}_{n}|h_{n}^{d}}}=0;\quad\lim_{n\rightarrow+\infty}\frac{b_{n}}{\sqrt{|\mathbb{G}_{n}|h_{n}^{2s+d}}}=+\infty,

where ss is the regularity parameter given in Assumption 2.12.

The paper is organised as follows. In Section 3.1 we state the main result for the moderate deviation principles of the estimators μ^𝔸n(x)\widehat{\mu}_{{\mathbb{A}}_{n}}(x) for xx in the set continuity of μ\mu and 𝔸n{𝕋n,𝔾n}{\mathbb{A}}_{n}\in\{\mathbb{T}_{n},\mathbb{G}_{n}\}. In Section 3.2, directly linked to the study of variance term V𝔸,h(x)V_{{\mathbb{A}},h}(x) defined in (4), we study the moderate deviation principle for general additive functionals of BMCs. Sections 4 and 5 are devoted to the proofs of results. In Section 6, we recall some useful results.

3. Main result

3.1. Moderate deviation principle for μ^𝔸n\widehat{\mu}_{{\mathbb{A}}_{n}}

First, we state a strong consistency result for the estimators μ^𝔸n(x)\widehat{\mu}_{{\mathbb{A}}_{n}}(x) for xx in the set of continuity of μ\mu. Its proof is given in Section 4.1.

Lemma 3.1.

Let XX be a BMC with kernel 𝒫{\mathcal{P}} and initial distribution ν\nu such that Assumptions 2.9, 2.11 and 2.12 hold. Furthermore, if α>1/2\alpha>1/2 then assume that Assumption 2.13 holds. Let (bn,n)(b_{n},n\in\mathbb{N}) be a positive sequence with satisfies (6). Then, for all xx in the set of continuity of μ\mu and 𝔸n{𝕋n,𝔾n}{\mathbb{A}}_{n}\in\{\mathbb{T}_{n},\mathbb{G}_{n}\} we have μ^𝔸n(x)bn2superexpμ(x).\widehat{\mu}_{{\mathbb{A}}_{n}}(x)\xRightarrow[b_{n}^{2}]{\rm superexp}\mu(x).

The main result of this Section is the following theorem which state the moderate deviation principle for μ^𝔸n(x)μ(x)\widehat{\mu}_{{\mathbb{A}}_{n}}(x)-\mu(x) for xx in the set of continuity of the function μ\mu.

Theorem 3.2.

Under the hypothesis of Lemma 3.1, for all xx in the set of continuity of μ\mu and 𝔸n{𝕋n,𝔾n}{\mathbb{A}}_{n}\in\{\mathbb{T}_{n},\mathbb{G}_{n}\}, bn1|𝔸n|hnd(μ^𝔸n(x)μ(x))b_{n}^{-1}\sqrt{|{\mathbb{A}}_{n}|h_{n}^{d}}(\widehat{\mu}_{{\mathbb{A}}_{n}}(x)-\mu(x)) satisfies a moderate deviation principle on \mathbb{R} with speed bn2b_{n}^{2} and rate function II defined by: I(y)=y2/(2K22μ(x))I(y)=y^{2}/(2\|K\|_{2}^{2}\mu(x)) for all yy\in\mathbb{R}, that is, for any AA\subset\mathbb{R},

infyÅI(y)lim infn+1bn2\displaystyle-\inf_{y\in\mathring{A}}I(y)\leq\liminf_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}} log(bn1|𝔸n|hnd(μ^𝔸n(x)μ(x))A)\displaystyle\log\mathbb{P}\big{(}b_{n}^{-1}\sqrt{|{\mathbb{A}}_{n}|h_{n}^{d}}(\widehat{\mu}_{{\mathbb{A}}_{n}}(x)-\mu(x))\in A\big{)}
lim supn+1bn2log(bn1|𝔸n|hnd(μ^𝔸n(x)μ(x))A)infyA¯I(y),\displaystyle\leq\limsup_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\mathbb{P}\big{(}b_{n}^{-1}\sqrt{|{\mathbb{A}}_{n}|h_{n}^{d}}(\widehat{\mu}_{{\mathbb{A}}_{n}}(x)-\mu(x))\in A\big{)}\leq-\inf_{y\in\bar{A}}I(y),

where Å\mathring{A} and A¯\bar{A} denote respectively the interior and the closure of AA.

In order to obtain confidence intervals for μ(x)\mu(x), it would be interesting to replace μ(x)\mu(x) in the expression of the rate function I()I(\cdot) by an estimator. In that direction, we have the following. Let 𝔸n{𝔾n,𝕋n}.{\mathbb{A}}_{n}^{*}\in\{\mathbb{G}_{n},\mathbb{T}_{n}\}. Obviously, 𝔸n{\mathbb{A}}_{n}^{*} and 𝔸n{\mathbb{A}}_{n} can be the same. We consider the estimator μ^𝔸n(x)\widehat{\mu}_{{\mathbb{A}}_{n}^{*}}(x) of μ(x)\mu(x) defined with 𝔸n{\mathbb{A}}_{n}^{*} instead of 𝔸n{\mathbb{A}}_{n}. Let (ϖn,n)(\varpi_{n},n\in\mathbb{N}) be a sequence of real numbers such that ϖn0\varpi_{n}\rightarrow 0 as n+.n\rightarrow+\infty. Then, we have the following result which the proof is given in Section 4.3.

Theorem 3.3.

Under the hypothesis of Lemma 3.1, for all xx in the set of continuity of μ\mu and 𝔸n,𝔸n{𝕋n,𝔾n}{\mathbb{A}}_{n},{\mathbb{A}}_{n}^{*}\in\{\mathbb{T}_{n},\mathbb{G}_{n}\}, bn1(K2μ^𝔸n(x)ϖn)1|𝔸n|hnd(μ^𝔸n(x)μ(x))b_{n}^{-1}(\|K\|_{2}\sqrt{\widehat{\mu}_{{\mathbb{A}}_{n}^{*}}(x)}\vee\varpi_{n})^{-1}\sqrt{|{\mathbb{A}}_{n}|h_{n}^{d}}(\widehat{\mu}_{{\mathbb{A}}_{n}}(x)-\mu(x)) satisfies a moderate deviation principle on \mathbb{R} with speed bn2b_{n}^{2} and rate function II^{\prime} defined by: I(y)=y2/2I^{\prime}(y)=y^{2}/2 for all yy\in\mathbb{R}.

In particular, using the contraction principle (see for e.g Dembo and Zeitouni [11], Chap 4), we have the following corollary of Theorem 3.3.

Corollary 3.4.

Under the hypothesis of Theorem 3.3, we have the following convergence for xx in the set of continuity of μ\mu and 𝔸n,𝔸{𝕋n,𝔾n}:{\mathbb{A}}_{n},{\mathbb{A}}^{*}\in\{\mathbb{T}_{n},\mathbb{G}_{n}\}:

limn+1bn2log(bn1(K2μ^𝔸n(x)ϖn)1|𝔸n|hnd|(μ^𝔸n(x)μ(x))|>δ)=δ22δ>0.\lim_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\mathbb{P}\Big{(}b_{n}^{-1}\,\Big{(}\|K\|_{2}\sqrt{\widehat{\mu}_{{\mathbb{A}}_{n}^{*}}(x)}\vee\varpi_{n}\Big{)}^{-1}\,\sqrt{|{\mathbb{A}}_{n}|h_{n}^{d}}\Big{|}\big{(}\widehat{\mu}_{{\mathbb{A}}_{n}}(x)-\mu(x)\big{)}\Big{|}>\delta\Big{)}=-\frac{\delta^{2}}{2}\quad\forall\delta>0.
Remark 3.5.

Corollary 3.4 yields a simple confidence interval for μ(x)\mu(x), of decreasing size     bn/|𝔸n|hndb_{n}/\sqrt{|{\mathbb{A}}_{n}|\,h_{n}^{d}} and with level asymptotically close to 1exp((bn2δ2)/2).1-\exp(-(b_{n}^{2}\,\delta^{2})/2).

Using the structure of the asymptotic variance σ2\sigma^{2} in (7), we can prove the following multidimensional result which the proof is given in Section 4.4

Corollary 3.6.

Under the hypothesis of Theorem 3.2, we have, for xx in the set of continuity of μ\mu and for all k0k\geq 0, bn1(|𝔾n|1/2hn1/2(μ^𝔾n(x)μ(x)),,|𝔾nk|1/2hnk1/2(μ^𝔾nk(x)μ(x)))tb_{n}^{-1}\Big{(}|\mathbb{G}_{n}|^{1/2}h_{n}^{1/2}\big{(}\widehat{\mu}_{\mathbb{G}_{n}}(x)-\mu(x)\big{)},\ldots,|\mathbb{G}_{n-k}|^{1/2}h_{n-k}^{1/2}\big{(}\widehat{\mu}_{\mathbb{G}_{n-k}}(x)-\mu(x)\big{)}\Big{)}^{t} satisfies a moderate deviation principle on k+1\mathbb{R}^{k+1} with speed bn2b_{n}^{2} and good rate function Jx:k+1J_{x}:\mathbb{R}^{k+1}\rightarrow\mathbb{R} defined by

Jx(𝒛)=(2K22μ(x))1𝒛tΓ1𝒛,𝒛k+1,J_{x}(\boldsymbol{z})=\big{(}2\,\|K\|_{2}^{2}\,\mu(x)\big{)}^{-1}\boldsymbol{z}^{t}\Gamma^{-1}\boldsymbol{z}\,,\quad\boldsymbol{z}\in\mathbb{R}^{k+1},

with Γ=diag(20,,2k)\Gamma=diag(2^{0},\ldots,2^{k}), where diag()diag(\cdot) denotes the diagonal matrix and 𝐳t\boldsymbol{z}^{t} stands for the transpose of vector 𝐳\boldsymbol{z}.

Remark 3.7.

We deduce from Corollary 3.6 that the estimators |𝔾n|1/2hnd/2(μ^𝔾n(x)μ(x))|\mathbb{G}_{n-\ell}|^{1/2}h_{n-\ell}^{d/2}(\widehat{\mu}_{\mathbb{G}_{n-\ell}}(x)-\mu(x)) are asymptotically independent in the sense of moderate deviation for {0,,k}\ell\in\{0,\ldots,k\} and for any k.k\in\mathbb{N}.

3.2. Moderate deviation principle for additive functionals of BMCs

In order to study the variance term V𝔸n,hn(x)V_{{\mathbb{A}}_{n},h_{n}}(x), we give here a moderate deviation principle for a general additive functionals of BMCs. For that purpose, we introduce the following assumption.

Assumption 3.8.

For nn\in\mathbb{N}, let 𝔣n=(f,n,n0){\mathfrak{f}}_{n}=(f_{\ell,n},n\geq\ell\geq 0) be a sequence of functions defined on SS such that f,n=0f_{\ell,n}=0 if >n\ell>n and there exists γ(0,1/d)\gamma\in(0,1/d) such that:

  • (i)

    sup0n{2dγn/2f,n; 2dγn/2𝒬f,n;𝒬(f,n2); 2dγn𝒫(f,n2)}<+.\sup_{0\leq\ell\leq n}\{2^{-d\gamma n/2}\|f_{\ell,n}\|_{\infty};\,2^{d\gamma n/2}\|{\mathcal{Q}}f_{\ell,n}\|_{\infty};\,\|{\mathcal{Q}}(f_{\ell,n}^{2})\|_{\infty};\,2^{d\gamma n}\|{\mathcal{P}}(f_{\ell,n}\otimes^{2})\|_{\infty}\}<+\infty.

  • (ii)

    sup0n{2dγn/2μ,|f,n|;μ,f,n2}<+.\sup_{0\leq\ell\leq n}\{2^{d\gamma n/2}\langle\mu,|f_{\ell,n}|\rangle;\,\langle\mu,f_{\ell,n}^{2}\rangle\}<+\infty.

  • (iii)

    The following limit exists and is finite:

    (7) σ2=limn+=0n2f,nL2(μ)2<+.\sigma^{2}=\lim_{n\rightarrow+\infty}\sum_{\ell=0}^{n}2^{-\ell}\mathop{\parallel\!f_{\ell,n}\!\parallel}\nolimits_{L^{2}(\mu)}^{2}<+\infty.

We will use the following notations. For a finite set 𝔸𝕋{\mathbb{A}}\subset\mathbb{T} and a function f(S)f\in{\mathcal{B}}(S), we set:

M𝔸(f)=i𝔸f(Xi).M_{\mathbb{A}}(f)=\sum_{i\in{\mathbb{A}}}f(X_{i}).

In this paper, we are interested in the cases 𝔸=𝔾n{\mathbb{A}}=\mathbb{G}_{n} and 𝔸=𝕋n{\mathbb{A}}=\mathbb{T}_{n}, that is the n-n\text{-}th generation and the first nn generation of the tree. Recall μ\mu the invariant probability of 𝒬{\mathcal{Q}}, transition probability of the auxiliary Markov chain (Yn,n)(Y_{n},n\in\mathbb{N}). For fL1(μ)f\in L^{1}(\mu), we set:

f~=fμ,f.\tilde{f}=f-\langle\mu,f\rangle.

Recall the sequence 𝔣n{\mathfrak{f}}_{n} defined in Assumption 3.8. For nn\in\mathbb{N}, we set:

(8) Nn,(𝔣n)=|𝔾n|1/2=0nM𝔾n(f~,n).N_{n,\emptyset}({\mathfrak{f}}_{n})=|\mathbb{G}_{n}|^{-1/2}\sum_{\ell=0}^{n}M_{\mathbb{G}_{n-\ell}}(\tilde{f}_{\ell,n}).

The notation Nn,N_{n,\emptyset} means that we consider the average from the root \emptyset to the n-n\text{-}th generation.

Remark 3.9.

The definition of Nn,(𝔣n)N_{n,\emptyset}({\mathfrak{f}}_{n}) in (8) is mainly motivated by the decomposition (4). It will allow us to threat the variance term of the estimator μ^𝔸n(x)\widehat{\mu}_{{\mathbb{A}}_{n}}(x) defined in (2). Instead, for nn\in\mathbb{N}, we set fnx()=Khn(x)f_{n}^{x}(\cdot)=K_{h_{n}}(x-\cdot). Then, we consider the sequences of functions (f,nid,n0)(f_{\ell,n}^{\text{id}},\,n\geq\ell\geq 0) and (f,n0,n0)(f^{0}_{\ell,n},\,n\geq\ell\geq 0) defined by:

(9) f,nid=fnxandf,n0=fnx𝟏{=0}.f_{\ell,n}^{\text{id}}=f^{x}_{n}\quad\text{and}\quad f^{0}_{\ell,n}=f^{x}_{n}{\bf 1}_{\{\ell=0\}}.

It is not difficult to check that under Assumption 2.11, the sequence (f,nid,n0)(f_{\ell,n}^{\text{id}},\,n\geq\ell\geq 0) and (f,n0,n0)(f^{0}_{\ell,n},\,n\geq\ell\geq 0) defined in (9) satisfy Assumption 3.8. In particular, let xx be in the set of continuity of μ\mu. Thanks to Lemma 6.3, we have:

(10) limn+fnxL2(μ)2=limn+μ,(fnx)2=μ(x)K22.\lim_{n\rightarrow+\infty}\mathop{\parallel\!f^{x}_{n}\!\parallel}\nolimits_{L^{2}(\mu)}^{2}=\lim_{n\rightarrow+\infty}\langle\mu,(f^{x}_{n})^{2}\rangle=\mu(x)\mathop{\parallel\!K\!\parallel}\nolimits_{2}^{2}.

If 𝔸n=𝔾n{\mathbb{A}}_{n}=\mathbb{G}_{n}, it suffices to consider the sequence 𝔣n=(f,n,0n){\mathfrak{f}}_{n}=(f_{\ell,n},0\leq\ell\leq n) with f,n=f,n0f_{\ell,n}=f_{\ell,n}^{0} and in that case, using (10), the asymptotic variance defined in (7) is given by σ2=K22μ(x)\sigma^{2}=\|K\|_{2}^{2}\,\mu(x). If 𝔸n=𝕋n{\mathbb{A}}_{n}=\mathbb{T}_{n}, it suffices to consider the sequence 𝔣n=(f,n,0n){\mathfrak{f}}_{n}=(f_{\ell,n},0\leq\ell\leq n) with f,n=f,nidf_{\ell,n}=f_{\ell,n}^{id} and in that case, using (10), the asymptotic variance defined in (7) is given by σ2=2K22μ(x)\sigma^{2}=2\|K\|_{2}^{2}\,\mu(x).

For our convenience, we assume that the quantity γ\gamma which appears in Assumptions 2.11 and 3.8 is the same. The main result of this section is the following.

Theorem 3.10.

Let XX be a BMC with kernel 𝒫{\mathcal{P}} and initial distribution ν\nu such that Assumptions 2.9, 2.11 and 3.8 hold. Furthermore, if α>1/2\alpha>1/2 then assume that Assumption 2.13 holds. Let (bn,n)(b_{n},n\in\mathbb{N}) be a positive sequence with satisfies (6). Then bn1Nn,(𝔣n)b_{n}^{-1}N_{n,\emptyset}({\mathfrak{f}}_{n}) satisfies a moderate deviation principle on \mathbb{R} with speed bn2b_{n}^{2} and rate function II defined by: I(x)=x2/(2σ2)I(x)=x^{2}/(2\sigma^{2}) for all xx\in\mathbb{R}, with the finite variance σ2\sigma^{2} defined in (7).

Remark 3.11.

In particular, using the contraction principle (see for e.g Dembo and Zeitouni [11], Chap 4), Theorem 3.10 implies that

limn+1bn2log(|bn1Nn,(𝔣n)|>δ)=I(δ)δ>0.\lim_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\mathbb{P}\left(\left|b_{n}^{-1}N_{n,\emptyset}({\mathfrak{f}}_{n})\right|>\delta\right)=-I(\delta)\quad\forall\delta>0.
Remark 3.12.

Unlike the results of Bitseki and Gorgui [4], one can note that the different regimes disappear in Theorem 3.10. Moreover, we are able here to give the fluctuations if 2α2>12\alpha^{2}>1 which is not the case in [4].

4. Proof of Lemma 3.1, Theorems 3.2 and 3.3 and Corollary 3.6

We will denote by CC any unimportant finite constant which may vary from line to line (in particular CC does not depend on nn\in{\mathbb{N}}).

4.1. Proof of Lemma 3.1

We begin the proof with 𝔸n=𝕋n.{\mathbb{A}}_{n}=\mathbb{T}_{n}. Recall the decomposition (4) with 𝕋n\mathbb{T}_{n} instead of 𝔸{\mathbb{A}}. Using Lemma 6.3, we have limn+|Bhn(x)|=0.\lim_{n\rightarrow+\infty}|B_{h_{n}}(x)|=0. From Remark 2.7, this implies that Bhn(x)bn2superexp0.B_{h_{n}}(x)\xRightarrow[b_{n}^{2}]{\rm superexp}0. Next, we set fn()=Khn(x)f_{n}(\cdot)=K_{h_{n}}(x-\cdot) in such a way that we have

|𝕋n|1hnd/2u𝕋n(Kh(xXu)Khμ(x))=|𝕋n|1hnd/2=0nM𝔾(f~n).|\mathbb{T}_{n}|^{-1}h_{n}^{d-/2}\sum_{u\in\mathbb{T}_{n}}\Big{(}K_{h}(x-X_{u})-K_{h}\star\mu(x)\Big{)}=|\mathbb{T}_{n}|^{-1}h_{n}^{d-/2}\sum_{\ell=0}^{n}M_{\mathbb{G}_{\ell}}(\tilde{f}_{n}).

Following line by line the proof of (32) (where we take f,n=fnf_{\ell,n}=f_{n} for all n\ell\leq n), we get

(|𝕋n|1hnd/2|=0nM𝔾(f~n)|>δ)2exp(3δc1+c2δ)exp(3δ2|𝕋n|hndc1+c2δ).\mathbb{P}\Big{(}|\mathbb{T}_{n}|^{-1}h_{n}^{d-/2}\Big{|}\sum_{\ell=0}^{n}M_{\mathbb{G}_{\ell}}(\tilde{f}_{n})\Big{|}>\delta\Big{)}\leq 2\exp\Big{(}\frac{3\delta}{c_{1}+c_{2}\delta}\Big{)}\exp\Big{(}-\frac{3\delta^{2}|\mathbb{T}_{n}|h_{n}^{d}}{c_{1}+c_{2}\delta}\Big{)}.

Taking the log\log, dividing by bn2b_{n}^{2} and letting nn goes to the infinity in the latter inequality, we get

|𝕋n|1hnd/2u𝕋n(Kh(xXu)Khμ(x))bn2superexp0.|\mathbb{T}_{n}|^{-1}h_{n}^{d-/2}\sum_{u\in\mathbb{T}_{n}}\Big{(}K_{h}(x-X_{u})-K_{h}\star\mu(x)\Big{)}\xRightarrow[b_{n}^{2}]{\rm superexp}0.

It then follows from the decomposition (4) that μ^𝕋n(x)bn2superexpμ(x).\widehat{\mu}_{\mathbb{T}_{n}}(x)\xRightarrow[b_{n}^{2}]{\rm superexp}\mu(x). We similarly get the result for 𝔸n=𝔾n{\mathbb{A}}_{n}=\mathbb{G}_{n} and this ends the proof of the lemma.

4.2. Proof of Theorem 3.2

We begin the proof with 𝔸n=𝕋n{\mathbb{A}}_{n}=\mathbb{T}_{n}. We have the following decomposition:

bn1|𝕋n|hnd(μ^𝕋n(x)μ(x))=|𝔾n||𝕋n|bn1Nn,(𝔣n)+|𝕋n|hndbnBhn(x),b_{n}^{-1}\sqrt{|\mathbb{T}_{n}|h_{n}^{d}}\big{(}\widehat{\mu}_{\mathbb{T}_{n}}(x)-\mu(x)\big{)}=\sqrt{\frac{|\mathbb{G}_{n}|}{|\mathbb{T}_{n}|}}b_{n}^{-1}N_{n,\emptyset}({\mathfrak{f}}_{n})+\frac{\sqrt{|\mathbb{T}_{n}|h_{n}^{d}}}{b_{n}}B_{h_{n}}(x),

where 𝔣n=(f,n,n0){\mathfrak{f}}_{n}=(f_{\ell,n},\,n\geq\ell\geq 0) with the functions f,n=f,nidf_{\ell,n}=f_{\ell,n}^{\text{id}} defined in (9) for n0n\geq\ell\geq 0 and f,n=0f_{\ell,n}=0 otherwise; Nn,(𝔣n)N_{n,\emptyset}({\mathfrak{f}}_{n}) is defined in (8) and the bias term Bhn(x)B_{h_{n}}(x) is defined in (4). Thanks to Theorem 3.10 applied to the sequence (f,nid,n0)(f_{\ell,n}^{id},n\geq\ell\geq 0) and using that limn+|𝔾n|/|𝕋n|=1/2,\lim_{n\rightarrow+\infty}|\mathbb{G}_{n}|/|\mathbb{T}_{n}|=1/2, we get that |𝔾n||𝕋n|1bn1Nn,(𝔣n)\sqrt{|\mathbb{G}_{n}||\mathbb{T}_{n}|^{-1}}b_{n}^{-1}N_{n,\emptyset}({\mathfrak{f}}_{n}) satisfies a moderate deviation principle in \mathbb{R} with speed bn2b_{n}^{2} and rate function II defined by: I(y)=y2/(2K22μ(x))I(y)=y^{2}/(2\|K\|_{2}^{2}\,\mu(x)) for all y.y\in\mathbb{R}. To complete the proof of Theorem 3.2, it suffices to prove that

(11) limn+|𝕋n|hndbnBhn(x)=0.\lim_{n\rightarrow+\infty}\frac{\sqrt{|\mathbb{T}_{n}|h_{n}^{d}}}{b_{n}}B_{h_{n}}(x)=0.

Next, using that

μ(xhny)μ(x)=j=1d(μ(x1hny1,,xjhnyj,xj+1,,xd)μ(x1hny1,,xj1hnyj1,xj,xj+1,,xd)),\mu(x-h_{n}y)-\mu(x)=\sum_{j=1}^{d}(\mu(x_{1}-h_{n}y_{1},\ldots,x_{j}-h_{n}y_{j},x_{j+1},\ldots,x_{d})\\ -\mu(x_{1}-h_{n}y_{1},\ldots,x_{j-1}-h_{n}y_{j-1},x_{j},x_{j+1},\ldots,x_{d})),

the Taylor expansion and Assumption 2.12, we get that, for some finite constant C>0C>0,

|𝕋n|1/2hnd/2Bhn(x)\displaystyle|\mathbb{T}_{n}|^{1/2}h_{n}^{d/2}B_{h_{n}}(x) =|𝕋n|hnd|dhndK(hn1(xy))μ(y)𝑑yμ(x)|\displaystyle=\sqrt{|\mathbb{T}_{n}|h_{n}^{d}}\,\,\Big{|}\int_{\mathbb{R}^{d}}h_{n}^{-d}K(h_{n}^{-1}(x-y))\mu(y)dy-\mu(x)\Big{|}
=|𝕋n|hnd|dK(y)(μ(xhny)μ(x))𝑑y|\displaystyle=\sqrt{|\mathbb{T}_{n}|h_{n}^{d}}\,\,\Big{|}\int_{\mathbb{R}^{d}}K(y)(\mu(x-h_{n}y)-\mu(x))\,dy\Big{|}
C|𝕋n|hndj=1ddK(y)(hn|yj|)ss!𝑑y\displaystyle\leq C\sqrt{|\mathbb{T}_{n}|h_{n}^{d}}\,\,\sum_{j=1}^{d}\,\,\int_{\mathbb{R}^{d}}K(y)\frac{(h_{n}|y_{j}|)^{s}}{\lfloor s\rfloor!}dy
C|𝕋n|hn2s+d.\displaystyle\leq C\sqrt{|\mathbb{T}_{n}|h_{n}^{2s+d}}.

Now, (11) follows using the latter inequality and (6). This ends the proof of Theorem 3.2 for 𝔸n=𝕋n{\mathbb{A}}_{n}=\mathbb{T}_{n}. The proof is similar for 𝔸n=𝔾n{\mathbb{A}}_{n}=\mathbb{G}_{n} using f,n=f,n0.f_{\ell,n}=f^{0}_{\ell,n}.

4.3. Proof of Theorem 3.3

. We begin the proof with 𝔸n=𝕋n{\mathbb{A}}_{n}=\mathbb{T}_{n}. We have the following decomposition:

(12) bn1|𝕋n|hnd(μ^𝕋n(x)μ(x))K2μ^𝔸n(x)ϖn=T1(n)+T2(n)\frac{b_{n}^{-1}\sqrt{|\mathbb{T}_{n}|h_{n}^{d}}(\widehat{\mu}_{\mathbb{T}_{n}}(x)-\mu(x))}{\|K\|_{2}\sqrt{\widehat{\mu}_{{\mathbb{A}}_{n}^{*}}(x)}\vee\varpi_{n}}=T_{1}(n)+T_{2}(n)

where

T1(n)=(K2μ(x)bn)1|𝕋n|hnd(μ^𝕋n(x)μ(x));\displaystyle T_{1}(n)=(\|K\|_{2}\sqrt{\mu(x)}b_{n})^{-1}\sqrt{|\mathbb{T}_{n}|h_{n}^{d}}\Big{(}\widehat{\mu}_{\mathbb{T}_{n}}(x)-\mu(x)\Big{)};
T2(n)=(1K2μ^𝔸n(x)ϖn1K2μ(x))bn1|𝕋n|hnd(μ^𝕋n(x)μ(x)).\displaystyle T_{2}(n)=\Big{(}\frac{1}{\|K\|_{2}\sqrt{\widehat{\mu}_{{\mathbb{A}}_{n}^{*}}(x)}\vee\varpi_{n}}-\frac{1}{\|K\|_{2}\sqrt{\mu(x)}}\Big{)}b_{n}^{-1}\sqrt{|\mathbb{T}_{n}|h_{n}^{d}}\Big{(}\widehat{\mu}_{\mathbb{T}_{n}}(x)-\mu(x)\Big{)}.

First, we prove that

(13) T2(n)bn2superexp0.T_{2}(n)\xRightarrow[b_{n}^{2}]{\rm superexp}0.

Let δ>0.\delta>0. For all r>0r>0, we have

(|T2(n)|>δ)(|bn1|𝕋n|hnd(μ^𝕋n(x)μ(x))|>δ/r)+(|1K2μ^𝔸n(x)ϖn1K2μ(x)|>r).\mathbb{P}\big{(}|T_{2}(n)|>\delta\big{)}\leq\mathbb{P}\big{(}\big{|}b_{n}^{-1}\sqrt{|\mathbb{T}_{n}|h_{n}^{d}}\big{(}\widehat{\mu}_{\mathbb{T}_{n}}(x)-\mu(x)\big{)}\big{|}>\delta/r\big{)}\\ +\mathbb{P}\big{(}\big{|}\frac{1}{\|K\|_{2}\sqrt{\widehat{\mu}_{{\mathbb{A}}_{n}^{*}}(x)}\vee\varpi_{n}}-\frac{1}{\|K\|_{2}\sqrt{\mu(x)}}\big{|}>r\big{)}.

This implies that (see for e.g [11], Lemma 1.2.15)

(14) lim supn+1bn2log(|T2(n)|>δ)max{lim supn+1bn2log(|bn1|𝕋n|hnd(μ^𝕋n(x)μ(x))|>δ/r);lim supn+1bn2log(|1K2μ^𝔸n(x)ϖn1K2μ(x)|>r)}.\limsup_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\mathbb{P}\big{(}|T_{2}(n)|>\delta\big{)}\leq\max\Big{\{}\limsup_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\mathbb{P}\big{(}\big{|}b_{n}^{-1}\sqrt{|\mathbb{T}_{n}|h_{n}^{d}}\big{(}\widehat{\mu}_{\mathbb{T}_{n}}(x)-\mu(x)\big{)}\big{|}>\delta/r\big{)};\\ \limsup_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\mathbb{P}\big{(}\big{|}\frac{1}{\|K\|_{2}\sqrt{\widehat{\mu}_{{\mathbb{A}}_{n}^{*}}(x)}\vee\varpi_{n}}-\frac{1}{\|K\|_{2}\sqrt{\mu(x)}}\big{|}>r\big{)}\Big{\}}.

Using Theorem 3.2 and the contraction principle, we have

(15) lim supn+1bn2log(|bn1|𝕋n|hnd(μ^𝕋n(x)μ(x))|>δ/r)=δ22K2μ(x)r2.\limsup_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\mathbb{P}\big{(}\big{|}b_{n}^{-1}\sqrt{|\mathbb{T}_{n}|h_{n}^{d}}\big{(}\widehat{\mu}_{\mathbb{T}_{n}}(x)-\mu(x)\big{)}\big{|}>\delta/r\big{)}=-\frac{\delta^{2}}{2\|K\|_{2}\mu(x)r^{2}}.

Following the step 1 of the proof of Theorem 6 in [7] and using Lemma 3.1, we can prove that

K22μ^𝔸n(x)ϖn2bn2superexpK22μ(x).\|K\|_{2}^{2}\,\widehat{\mu}_{{\mathbb{A}}_{n}^{*}}(x)\vee\varpi_{n}^{2}\xRightarrow[b_{n}^{2}]{\rm superexp}\|K\|_{2}^{2}\,\mu(x).

Using Lemma B.2 in [3], the latter convergence implies that

(16) 1K2μ^𝔸n(x)ϖnbn2superexp1K2μ(x).\frac{1}{\|K\|_{2}\sqrt{\widehat{\mu}_{{\mathbb{A}}_{n}^{*}}(x)}\vee\varpi_{n}}\xRightarrow[b_{n}^{2}]{\rm superexp}\frac{1}{\|K\|_{2}\sqrt{\mu(x)}}.

Using (14), (15) and (16), we get

lim supn+1bn2log(|T2(n)|>δ)δ22K2μ(x)r2.\limsup_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\mathbb{P}\big{(}|T_{2}(n)|>\delta\big{)}\leq-\frac{\delta^{2}}{2\|K\|_{2}\mu(x)r^{2}}.

Since rr can be taken arbitrarily close to 0, we get (13) and using (12), this implies that

(17) bn1|𝕋n|hnd(μ^𝕋n(x)μ(x))K2μ^𝔸n(x)ϖnbn2superexpT1(n).\frac{b_{n}^{-1}\sqrt{|\mathbb{T}_{n}|h_{n}^{d}}(\widehat{\mu}_{\mathbb{T}_{n}}(x)-\mu(x))}{\|K\|_{2}\sqrt{\widehat{\mu}_{{\mathbb{A}}_{n}^{*}}(x)}\vee\varpi_{n}}\mathrel{\underset{b_{n}^{2}}{\overset{{\rm superexp}}{\scalebox{2.0}[1.0]{$\sim$}}}}T_{1}(n).

Using Theorem 3.2 and the contraction principle, we get that T1(n)T_{1}(n) satisfies a moderate deviation principle on \mathbb{R} with speed bn2b_{n}^{2} and rate function II^{\prime} defined by: I(y)=y2/2I^{\prime}(y)=y^{2}/2 for all yy\in\mathbb{R}. Using (17) and Remark 2.7, we get the result of Theorem 3.3.

4.4. Proof of Corollary 3.6

Let 𝒂=(a0,,ak)tk+1\boldsymbol{a}=(a_{0},\ldots,a_{k})^{t}\in\mathbb{R}^{k+1}. Let n>k.n>k. We consider the sequence 𝔣n=(f,n,n0){\mathfrak{f}}_{n}=(f_{\ell,n},n\geq\ell\geq 0) defined by f,n=2/2aKhn(x)f_{\ell,n}=2^{\ell/2}\,a_{\ell}\,K_{h_{n-\ell}}(x-\cdot) for all {0,,k}\ell\in\{0,\ldots,k\} and f,n=0f_{\ell,n}=0 otherwise. We easily check that 𝔣n{\mathfrak{f}}_{n} satisfies Assumptions 3.8. In particular, the asymptotic variance defined in (7) is given by σ2=(=0k2a2)K22μ(x).\sigma^{2}=\big{(}\sum_{\ell=0}^{k}2^{\ell}a_{\ell}^{2}\big{)}\|K\|_{2}^{2}\,\mu(x). Observe that the linear combinaison Mn(𝒂)M_{n}(\boldsymbol{a}), with coefficients 𝒂=(a0,,ak)tk+1,\boldsymbol{a}=(a_{0},\ldots,a_{k})^{t}\in\mathbb{R}^{k+1}, of the estimators |𝔾n|1/2hnd/2(μ^𝔾n(x)μ(x))|\mathbb{G}_{n-\ell}|^{1/2}h_{n-\ell}^{d/2}(\widehat{\mu}_{\mathbb{G}_{n-\ell}}(x)-\mu(x)), {0,,k}\ell\in\{0,\ldots,k\} has the following decomposition:

(18) Mn(𝒂)=Nn,(𝔣n)+=0ka(|𝔾n|hnd)1/2Bhn(x),M_{n}(\boldsymbol{a})=N_{n,\emptyset}({\mathfrak{f}}_{n})+\sum_{\ell=0}^{k}a_{\ell}\big{(}|\mathbb{G}_{n-\ell}|\,h_{n-\ell}^{d}\big{)}^{1/2}B_{h_{n-\ell}}(x),

where Nn,(𝔣n)N_{n,\emptyset}({\mathfrak{f}}_{n}) is defined in (8) and the Bhn(x)B_{h_{n-\ell}}(x), {0,,k}\ell\in\{0,\ldots,k\}, are defined in (4). Applying Theorem 3.10, we get that bn1Nn,(𝔣n)b_{n}^{-1}\,N_{n,\emptyset}({\mathfrak{f}}_{n}) satisfies a moderate deviation principle on \mathbb{R} with speed bn2b_{n}^{2} and rate function Ix,𝒂:I_{x,\boldsymbol{a}}:\mathbb{R}\rightarrow\mathbb{R} defined by

(19) Ix,𝒂(y)=y22(=0k2a2)K22μ(x),y.I_{x,\boldsymbol{a}}(y)=\frac{y^{2}}{2\,\big{(}\sum_{\ell=0}^{k}2^{\ell}a_{\ell}^{2}\big{)}\|K\|_{2}^{2}\,\mu(x)},\quad y\in\mathbb{R}.

Using (11), we have that

limn+1bn=0ka(|𝔾n|hnd)1/2Bhn(x)=0.\lim_{n\rightarrow+\infty}\frac{1}{b_{n}}\sum_{\ell=0}^{k}a_{\ell}\big{(}|\mathbb{G}_{n-\ell}|\,h_{n-\ell}^{d}\big{)}^{1/2}B_{h_{n-\ell}}(x)=0.

Using Remark 2.7, this implies that

(20) 1bn=0ka(|𝔾n|hnd)1/2Bhn(x)bn2superexp0.\frac{1}{b_{n}}\sum_{\ell=0}^{k}a_{\ell}\big{(}|\mathbb{G}_{n-\ell}|\,h_{n-\ell}^{d}\big{)}^{1/2}B_{h_{n-\ell}}(x)\xRightarrow[b_{n}^{2}]{\rm superexp}0.

Using (18) and (20) we get that bn1Mn(𝒂)b_{n}^{-1}M_{n}(\boldsymbol{a}) and bn1Nn,(𝔣n)b_{n}^{-1}N_{n,\emptyset}({\mathfrak{f}}_{n}) satisfy the same moderate deviation principle. We then conclude that bn1Mn(𝒂)b_{n}^{-1}M_{n}(\boldsymbol{a}) satisfies a moderate deviation principle on \mathbb{R} with speed bn2b_{n}^{2} and rate function Ix,𝒂I_{x,\boldsymbol{a}} defined in (19). Since this is true for all vector 𝒂k+1\boldsymbol{a}\in\mathbb{R}^{k+1}, that is for all the linear combinaisons of the estimators |𝔾n|1/2hnd/2(μ^𝔾n(x)μ(x))|\mathbb{G}_{n-\ell}|^{1/2}h_{n-\ell}^{d/2}(\widehat{\mu}_{\mathbb{G}_{n-\ell}}(x)-\mu(x)), {0,,k}\ell\in\{0,\ldots,k\}, we get the result of Corollary 3.6.

5. Proof of Theorem 3.10

We begin with some notations. We will denote by CC any unimportant finite constant which may vary from line to line (in particular CC does not depend on nn\in{\mathbb{N}} nor on the considered sequence of functions 𝔣n=(f,n,n0){\mathfrak{f}}_{n}=(f_{\ell,n},n\geq\ell\geq 0)). Let (pn,n)(p_{n},n\in{\mathbb{N}}) be a non-decreasing sequence of elements of {\mathbb{N}}^{*} such that

limn+pn3bn2|𝔾npn|1=0.\lim_{n\rightarrow+\infty}p_{n}^{3}\,b_{n}^{2}\,|\mathbb{G}_{n-p_{n}}|^{-1}=0.

When there is no ambiguity, we write pp for pnp_{n}.

Let i,j𝕋i,j\in\mathbb{T}. We write iji\preccurlyeq j if ji𝕋j\in i\mathbb{T}. We denote by iji\wedge j the most recent common ancestor of ii and jj, which is defined as the only u𝕋u\in\mathbb{T} such that if v𝕋v\in\mathbb{T} and viv\preccurlyeq i, vjv\preccurlyeq j then vuv\preccurlyeq u. We also define the lexicographic order iji\leq j if either iji\preccurlyeq j or v0iv0\preccurlyeq i and v1jv1\preccurlyeq j for v=ijv=i\wedge j. Let X=(Xi,i𝕋)X=(X_{i},i\in\mathbb{T}) be a BMCBMC with kernel 𝒫{\mathcal{P}} and initial measure ν\nu. For i𝕋i\in\mathbb{T}, we define the σ\sigma-field:

i={Xu;u𝕋 such that ui}.{\mathcal{F}}_{i}=\{X_{u};u\in\mathbb{T}\text{ such that $u\leq i$}\}.

By construction, the σ\sigma-fields (i;i𝕋)({\mathcal{F}}_{i};\,i\in\mathbb{T}) are nested as ij{\mathcal{F}}_{i}\subset{\mathcal{F}}_{j} for iji\leq j.

We define for nn\in{\mathbb{N}}, i𝔾npni\in\mathbb{G}_{n-p_{n}} and 𝔣n{\mathfrak{f}}_{n} the martingale increments:

(21) Δn,i(𝔣n)=Nn,i(𝔣n)𝔼[Nn,i(𝔣n)|i]andΔn(𝔣n)=i𝔾npnΔn,i(𝔣n),\Delta_{n,i}({\mathfrak{f}}_{n})=N_{n,i}({\mathfrak{f}}_{n})-{\mathbb{E}}\left[N_{n,i}({\mathfrak{f}}_{n})|\,{\mathcal{F}}_{i}\right]\quad\text{and}\quad\Delta_{n}({\mathfrak{f}}_{n})=\sum_{i\in\mathbb{G}_{n-p_{n}}}\Delta_{n,i}({\mathfrak{f}}_{n}),

where

(22) Nn,i(𝔣n)=|𝔾n|1/2=0pMi𝔾p(f~,n)andi𝔾p={ij,j𝔾p}.N_{n,i}({\mathfrak{f}}_{n})=|\mathbb{G}_{n}|^{-1/2}\sum_{\ell=0}^{p}M_{i\mathbb{G}_{p-\ell}}(\tilde{f}_{\ell,n})\quad\text{and}\quad i\mathbb{G}_{p-\ell}=\{ij,j\in\mathbb{G}_{p-\ell}\}.

We have:

i𝔾npnNn,i(𝔣n)=|𝔾n|1/2=0pnM𝔾n(f~,n)=|𝔾n|1/2k=npnnM𝔾k(f~nk,n).\sum_{i\in\mathbb{G}_{n-p_{n}}}N_{n,i}({\mathfrak{f}}_{n})=|\mathbb{G}_{n}|^{-1/2}\sum_{\ell=0}^{p_{n}}M_{\mathbb{G}_{n-\ell}}(\tilde{f}_{\ell,n})=|\mathbb{G}_{n}|^{-1/2}\sum_{k=n-p_{n}}^{n}M_{\mathbb{G}_{k}}(\tilde{f}_{n-k,n}).

Using the branching Markov property, we get for i𝔾npni\in\mathbb{G}_{n-p_{n}}:

(23) 𝔼[Nn,i(𝔣n)|i]=𝔼[Nn,i(𝔣n)|Xi]=|𝔾n|1/2=0pn𝔼Xi[M𝔾pn(f~,n)].{\mathbb{E}}\left[N_{n,i}({\mathfrak{f}}_{n})|\,{\mathcal{F}}_{i}\right]={\mathbb{E}}\left[N_{n,i}({\mathfrak{f}}_{n})|\,X_{i}\right]=|\mathbb{G}_{n}|^{-1/2}\sum_{\ell=0}^{p_{n}}{\mathbb{E}}_{X_{i}}\left[M_{\mathbb{G}_{p_{n}-\ell}}(\tilde{f}_{\ell,n})\right].

We have the following decomposition:

(24) Nn,(𝔣n)=Δn(𝔣n)+R0(n)+R1(n),N_{n,\emptyset}({\mathfrak{f}}_{n})=\Delta_{n}({\mathfrak{f}}_{n})+R_{0}(n)+R_{1}(n),

where Δn(𝔣)\Delta_{n}({\mathfrak{f}}) is defined in (21) and:

R0(n)=|𝔾n|1/2k=0npn1M𝔾k(f~nk,n)andR1(n)=i𝔾npn𝔼[Nn,i(𝔣n)|i].R_{0}(n)=|\mathbb{G}_{n}|^{-1/2}\,\sum_{k=0}^{n-p_{n}-1}M_{\mathbb{G}_{k}}(\tilde{f}_{n-k,n})\quad\text{and}\quad R_{1}(n)=\sum_{i\in\mathbb{G}_{n-p_{n}}}{\mathbb{E}}\left[N_{n,i}({\mathfrak{f}}_{n})|\,{\mathcal{F}}_{i}\right].

From (24), our goals will be achieved if we prove the following:

(25) bn1R0(n)bn2superexp0;\displaystyle b_{n}^{-1}R_{0}(n)\xRightarrow[b_{n}^{2}]{\rm superexp}0;
(26) bn1R1(n)bn2superexp0;\displaystyle b_{n}^{-1}R_{1}(n)\xRightarrow[b_{n}^{2}]{\rm superexp}0;
(27) bn1Δn(𝔣)satisfies a MDP on S with speed bn2 and rate function I.\displaystyle b_{n}^{-1}\Delta_{n}({\mathfrak{f}})\quad\text{satisfies a MDP on $S$ with speed $b_{n}^{2}$ and rate function $I$.}

Note that (25) and (26) mean that R0(n)R_{0}(n) and R1(n)R_{1}(n) are negligible in the sense of moderate deviations in such a way that using (24) and Remark 2.7, Nn,(𝔣)N_{n,\emptyset}({\mathfrak{f}}) and Δn(𝔣)\Delta_{n}({\mathfrak{f}}) satisfy the same moderate deviation principle. To prove (27), the main method we will use is the moderate deviations for martingale (see [12] for more details).

In the sequel, the sequence (2γn,n)(2^{-\gamma n},n\in\mathbb{N}) which appears in Assumption 3.8 will be denoted (hn,n)(h_{n},n\in\mathbb{N}) in such a way that we have 2dγn/2=hnd/2.2^{-d\gamma n/2}=h_{n}^{d/2}. We have the following result.

Lemma 5.1.

Under the assumptions of Theorem 3.10, we have bn1R0(n)bn2superexp0.b_{n}^{-1}R_{0}(n)\xRightarrow[b_{n}^{2}]{\rm superexp}0.

Proof.

Let δ>0\delta>0. Using the Chernoff bound, we have, for all λ>0\lambda>0,

(28) (bn1R0(n)>δ)exp(λbn|𝔾n|1/2δ)𝔼[exp(λ=0np1M𝔾(f~n,n))].\mathbb{P}\Big{(}b_{n}^{-1}R_{0}(n)>\delta\Big{)}\leq\exp\Big{(}-\lambda b_{n}|\mathbb{G}_{n}|^{1/2}\delta\Big{)}\mathbb{E}\Big{[}\exp\Big{(}\lambda\sum_{\ell=0}^{n-p-1}M_{\mathbb{G}_{\ell}}(\tilde{f}_{n-\ell,n})\Big{)}\Big{]}.

For all k{1,,np}k\in\{1,\ldots,n-p\} and for u𝕋u\in\mathbb{T}, we set

gp,k=r=0k12r𝒬rf~p+kr,nandZp,k(u)=gp,k(Xu0)+gp,k(Xu1)2𝒬gp,k(Xu).g_{p,k}=\sum_{r=0}^{k-1}2^{r}{\mathcal{Q}}^{r}\tilde{f}_{p+k-r,n}\quad\text{and}\quad Z_{p,k}(u)=g_{p,k}(X_{u0})+g_{p,k}(X_{u1})-2{\mathcal{Q}}g_{p,k}(X_{u}).

Then, using recursively the fact that

u𝔾f(Xu)=u𝔾1(f(Xu0)+f(Xu1)2𝒬f(Xu))+u𝔾12𝒬f(Xu),\sum_{u\in\mathbb{G}_{\ell}}f(X_{u})=\sum_{u\in\mathbb{G}_{\ell-1}}(f(X_{u0})+f(X_{u1})-2{\mathcal{Q}}f(X_{u}))+\sum_{u\in\mathbb{G}_{\ell-1}}2{\mathcal{Q}}f(X_{u}),

for all 1\ell\geq 1 and for some function ff, we get

𝔼[exp(λ=0np1M𝔾(f~n,n))]=𝔼[exp(λgp,np(X))k=1np1exp(λu𝔾npk1Zp,k(u))].\mathbb{E}\Big{[}\exp\Big{(}\lambda\sum_{\ell=0}^{n-p-1}M_{\mathbb{G}_{\ell}}(\tilde{f}_{n-\ell,n})\Big{)}\Big{]}=\mathbb{E}\Big{[}\exp\Big{(}\lambda g_{p,n-p}(X_{\emptyset})\Big{)}\,\prod_{k=1}^{n-p-1}\exp\Big{(}\lambda\sum_{u\in\mathbb{G}_{n-p-k-1}}Z_{p,k}(u)\Big{)}\Big{]}.

For all m{1,,np1}m\in\{1,\ldots,n-p-1\}, we set

𝕀m=𝔼[exp(λgp,np(X))k=mnp1exp(λu𝔾npk1Zp,k(u))].\mathbb{I}_{m}=\mathbb{E}\Big{[}\exp(\lambda g_{p,n-p}(X_{\emptyset}))\,\prod_{k=m}^{n-p-1}\exp(\lambda\sum_{u\in\mathbb{G}_{n-p-k-1}}Z_{p,k}(u))\Big{]}.

Using the branching Markov property, we get the following decomposition:

𝕀m=𝔼[exp(λgp,np(X))𝕁mk=m+1np1exp(λu𝔾npk1Zp,k(u))],\mathbb{I}_{m}=\mathbb{E}\Big{[}\exp(\lambda g_{p,n-p}(X_{\emptyset}))\,\mathbb{J}_{m}\,\prod_{k=m+1}^{n-p-1}\exp(\lambda\sum_{u\in\mathbb{G}_{n-p-k-1}}Z_{p,k}(u))\Big{]},

with

𝕁m=u𝔾npm1𝔼Xu[exp(λZp,m(u))].\mathbb{J}_{m}=\prod_{u\in\mathbb{G}_{n-p-m-1}}\mathbb{E}_{X_{u}}\Big{[}\exp(\lambda Z_{p,m}(u))\Big{]}.

For all u𝔾npm1u\in\mathbb{G}_{n-p-m-1}, we will upper bound the quantity 𝔼Xu[exp(λZp,m(u))]\mathbb{E}_{X_{u}}[\exp(\lambda Z_{p,m}(u))] and then 𝕁m\mathbb{J}_{m}. We claim that:

(29) |Zp,m(u)|M=Chnd/2;|Z_{p,m}(u)|\,\leq\,M\,=\,C\,h_{n}^{-d/2};
(30) 𝔼Xu[Zp,m(u)2]σm2=C+Chnd(r=0m1(2α)r1)2𝟏{m>1}.\mathbb{E}_{X_{u}}[Z_{p,m}(u)^{2}]\,\leq\,\sigma_{m}^{2}\,=\,C+\,C\,h_{n}^{d}\,\Big{(}\sum_{r=0}^{m-1}(2\alpha)^{r-1}\Big{)}^{2}{\bf 1}_{\{m>1\}}.

For that purpose, we plan to use the bound

(31) 𝔼[exp(λZ)]exp(λ2σ22(1λM/3))\mathbb{E}\big{[}\exp(\lambda Z)\big{]}\leq\exp\Big{(}\frac{\lambda^{2}\sigma^{2}}{2(1-\lambda M/3)}\Big{)}

valid for any λ(0,3/M)\lambda\in(0,3/M), any random variable ZZ such that |Z|M|Z|\leq M, 𝔼[Z]=0\mathbb{E}[Z]=0 and 𝔼[Z2]σ2\mathbb{E}[Z^{2}]\leq\sigma^{2}. For all u𝔾npm1u\in\mathbb{G}_{n-p-m-1} and for all λ(0,Chnd/2/3)\lambda\in(0,Ch_{n}^{-d/2}/3) we get, using (29)-(31),

𝔼Xu[exp(λZp,m(u))]exp(λ2σm22(1λM/3)).\mathbb{E}_{X_{u}}\Big{[}\exp(\lambda Z_{p,m}(u))\Big{]}\leq\exp\Big{(}\frac{\lambda^{2}\sigma_{m}^{2}}{2(1-\lambda M/3)}\Big{)}.

For all m{1,,np1}m\in\{1,\ldots,n-p-1\}, the latter inequality implies that

𝕁mexp(λ2σm2|𝔾npm1|2(1λM/3))and𝕀mexp(λ2σm2|𝔾npm1|2(1λM/3))𝕀m+1.\mathbb{J}_{m}\leq\exp\Big{(}\frac{\lambda^{2}\sigma_{m}^{2}|\mathbb{G}_{n-p-m-1}|}{2(1-\lambda M/3)}\Big{)}\quad\text{and}\quad\mathbb{I}_{m}\leq\exp\Big{(}\frac{\lambda^{2}\sigma_{m}^{2}|\mathbb{G}_{n-p-m-1}|}{2(1-\lambda M/3)}\Big{)}\mathbb{I}_{m+1}.

Recall that 𝕀1=𝔼[exp(λ=0np1M𝔾(f~n,n))].\mathbb{I}_{1}=\mathbb{E}\big{[}\exp\big{(}\lambda\sum_{\ell=0}^{n-p-1}M_{\mathbb{G}_{\ell}}(\tilde{f}_{n-\ell,n})\big{)}\big{]}. By recurrence, we get

𝔼[exp(λ=0np1M𝔾(f~n,n))]=𝕀1exp(λ2m=1np1σm2|𝔾npm1|2(1λM/3))𝔼[exp(λgp,np(X))].\mathbb{E}\Big{[}\exp\Big{(}\lambda\sum_{\ell=0}^{n-p-1}M_{\mathbb{G}_{\ell}}(\tilde{f}_{n-\ell,n})\Big{)}\Big{]}=\mathbb{I}_{1}\leq\exp\Big{(}\frac{\lambda^{2}\sum_{m=1}^{n-p-1}\sigma_{m}^{2}|\mathbb{G}_{n-p-m-1}|}{2(1-\lambda M/3)}\Big{)}\mathbb{E}\Big{[}\exp\Big{(}\lambda g_{p,n-p}(X_{\emptyset})\Big{)}\Big{]}.

Using (i)(i) and (ii)(ii) of Assumption 3.8 and (3), we have

|gp,np||f~n,n|+r=1np12r|𝒬r1(𝒬f~nr,n)|Chnd/2+Chnd/2r=1np1(2α)r1.\displaystyle|g_{p,n-p}|\leq|\tilde{f}_{n,n}|+\sum_{r=1}^{n-p-1}2^{r}|{\mathcal{Q}}^{r-1}({\mathcal{Q}}\tilde{f}_{n-r,n})|\leq Ch_{n}^{-d/2}+Ch_{n}^{d/2}\sum_{r=1}^{n-p-1}(2\alpha)^{r-1}.

This implies that

𝔼[exp(λ=0np1M𝔾(f~n,n))]exp(λ2m=1np1σm2|𝔾npm1|2(1λM/3))×exp(λChnd/2+λChnd/2r=0np2(2α)r).\mathbb{E}\Big{[}\exp\Big{(}\lambda\sum_{\ell=0}^{n-p-1}M_{\mathbb{G}_{\ell}}(\tilde{f}_{n-\ell,n})\Big{)}\Big{]}\leq\exp\Big{(}\frac{\lambda^{2}\sum_{m=1}^{n-p-1}\sigma_{m}^{2}|\mathbb{G}_{n-p-m-1}|}{2(1-\lambda M/3)}\Big{)}\\ \times\,\exp\Big{(}\lambda\,C\,h_{n}^{-d/2}\,+\,\lambda\,C\,h_{n}^{d/2}\,\sum_{r=0}^{n-p-2}(2\alpha)^{r}\Big{)}.

Distinguishing the cases 2α12\alpha\leq 1, 1/2<2α21/2<2\alpha\leq\sqrt{2} and 2α>22\alpha>\sqrt{2} and using (5) for 2α>12\alpha>1, we get

𝔼[exp(λ=0np1M𝔾(f~n,n))]exp(c1λ2|𝔾np|2(1c2λhnd/2/3))exp(c3λhnd/2),\mathbb{E}\Big{[}\exp\Big{(}\lambda\sum_{\ell=0}^{n-p-1}M_{\mathbb{G}_{\ell}}(\tilde{f}_{n-\ell,n})\Big{)}\Big{]}\leq\exp\Big{(}\frac{c_{1}\lambda^{2}|\mathbb{G}_{n-p}|}{2(1-c_{2}\lambda h_{n}^{-d/2}/3)}\Big{)}\exp\Big{(}c_{3}\lambda h_{n}^{-d/2}\Big{)},

where c1c_{1}, c2c_{2} and c3c_{3} are some positive constants. The latter inequality and (28) imply that

(bn1R0(n)>δ)exp(λbn|𝔾n|1/2δ+c1λ2|𝔾np|2(1c2λhnd/2/3))exp(c3λhnd/2).\mathbb{P}\Big{(}b_{n}^{-1}R_{0}(n)>\delta\Big{)}\leq\exp\Big{(}-\lambda b_{n}|\mathbb{G}_{n}|^{1/2}\delta+\frac{c_{1}\lambda^{2}|\mathbb{G}_{n-p}|}{2(1-c_{2}\lambda h_{n}^{-d/2}/3)}\Big{)}\,\exp\Big{(}c_{3}\lambda h_{n}^{-d/2}\Big{)}.

Taking111In fact, we use the following. For α,β,γ>0\alpha,\beta,\gamma>0 and h(x)=αx+βx22(1γx)h(x)=-\alpha x+\frac{\beta x^{2}}{2(1-\gamma x)} we have h(x)=α22(β+αγ)h(x^{*})=\frac{-\alpha^{2}}{2(\beta+\alpha\gamma)} for the choice x=α2αγ+β(0,1/γ).x^{*}=\frac{\alpha}{2\alpha\gamma+\beta}\in(0,1/\gamma).

λ=3bn|𝔾n|1/2δ2c2bn|𝔾n|1/2hnd/2δ+3c1|𝔾np|,\lambda=\frac{3\,b_{n}\,|\mathbb{G}_{n}|^{1/2}\,\delta}{2\,c_{2}\,b_{n}\,|\mathbb{G}_{n}|^{1/2}\,h_{n}^{-d/2}\,\delta+3\,c_{1}\,|\mathbb{G}_{n-p}|},

we are led to

(bn1R0(n)>δ)Cexp(3δ2bn2|𝔾n|2(c2δbn|𝔾n|1/2hnd/2+3c1|𝔾np|)).\mathbb{P}\Big{(}b_{n}^{-1}R_{0}(n)>\delta\Big{)}\leq C\,\exp\Big{(}-\,\frac{3\,\delta^{2}\,b_{n}^{2}\,|\mathbb{G}_{n}|}{2(c_{2}\,\delta\,b_{n}\,|\mathbb{G}_{n}|^{1/2}\,h_{n}^{-d/2}+3\,c_{1}\,|\mathbb{G}_{n-p}|)}\Big{)}.

Since we can do the same thing for 𝔣n-{\mathfrak{f}}_{n} instead of 𝔣n{\mathfrak{f}}_{n}, we get that

(32) (bn1|R0(n)|>δ)2Cexp(3δ2bn2|𝔾n|2(c2δbn|𝔾n|1/2hnd/2+3c1|𝔾np|)).\mathbb{P}\Big{(}b_{n}^{-1}|R_{0}(n)|>\delta\Big{)}\leq 2\,C\,\exp\Big{(}-\,\frac{3\,\delta^{2}\,b_{n}^{2}\,|\mathbb{G}_{n}|}{2(c_{2}\,\delta\,b_{n}\,|\mathbb{G}_{n}|^{1/2}\,h_{n}^{-d/2}+3\,c_{1}\,|\mathbb{G}_{n-p}|)}\Big{)}.

Finally, in the latter inequality, taking the log\log, dividing by bn2b_{n}^{2} and letting nn goes to infinity, we get the result of Lemma 5.1. Now, to end the proof, we will prove (29) and (30).

Proof of (29)

Using Assumption 2.9, (i)(i) and (ii)(ii) of Assumption 3.8 and Assumption 2.13, we get

|Zp,m(u)|\displaystyle|Z_{p,m}(u)| Cf~p+1,n+C(1+2α)(r=1m1(2α)r1𝒬fp+mr,n)𝟏{m>1}\displaystyle\leq C\,\|\tilde{f}_{p+1,n}\|_{\infty}+C(1+2\alpha)(\sum_{r=1}^{m-1}(2\alpha)^{r-1}\|{\mathcal{Q}}f_{p+m-r,n}\|_{\infty}){\bf 1}_{\{m>1\}}
Chnd/2+Chnd/2r=0m1(2α)rChnd/2.\displaystyle\leq C\,h_{n}^{-d/2}\,+\,C\,h_{n}^{d/2}\,\sum_{r=0}^{m-1}(2\alpha)^{r}\,\leq\,C\,h_{n}^{-d/2}.

Proof of (30)

Using the branching Markov property for the second inequality, Assumption 2.9 for the fourth inequality and (i)(i) and (ii)(ii) of Assumption 3.8 for the last inequality, we get

𝔼Xu[Zp,m(u)2]\displaystyle\mathbb{E}_{X_{u}}[Z_{p,m}(u)^{2}] 𝔼Xu[(gp,m(Xu0)+gp,m(Xu1))2]C𝒬(gp,m2)(Xu)\displaystyle\leq\mathbb{E}_{X_{u}}[(g_{p,m}(X_{u0})+g_{p,m}(X_{u1}))^{2}]\leq C{\mathcal{Q}}(g_{p,m}^{2})(X_{u})
C𝒬(f~p+1,n2)(Xu)+C𝒬((r=1m12r𝒬r1(𝒬f~p+mr,n))2)(Xu) 1{m>1}\displaystyle\leq C{\mathcal{Q}}(\tilde{f}_{p+1,n}^{2})(X_{u})+C{\mathcal{Q}}\Big{(}\Big{(}\sum_{r=1}^{m-1}2^{r}{\mathcal{Q}}^{r-1}({\mathcal{Q}}\tilde{f}_{p+m-r,n})\Big{)}^{2}\Big{)}(X_{u})\,{\bf 1}_{\{m>1\}}
C𝒬f~p+1,n2+(r=1m1(2α)r1𝒬fp+mr,n)2𝟏{m>1}\displaystyle\leq C\|{\mathcal{Q}}\tilde{f}_{p+1,n}^{2}\|_{\infty}+\Big{(}\sum_{r=1}^{m-1}(2\alpha)^{r-1}\|{\mathcal{Q}}f_{p+m-r,n}\|_{\infty}\Big{)}^{2}{\bf 1}_{\{m>1\}}
C+Chnd(r=0m1(2α)r)2𝟏{m>1}.\displaystyle\leq\,C\,+\,C\,h_{n}^{d}\,\Big{(}\sum_{r=0}^{m-1}(2\alpha)^{r}\Big{)}^{2}{\bf 1}_{\{m>1\}}.

Next, we have the following result.

Lemma 5.2.

Under the assumptions of Theorem 3.10, we have bn1R1(n)bn2superexp0.b_{n}^{-1}R_{1}(n)\xRightarrow[b_{n}^{2}]{\rm superexp}0.

Proof.

We have, using (23) and (49),

(33) R1(n)=|𝔾n|1/2M𝔾np(gp,n)wheregp,n==0p2p𝒬pf~,n.R_{1}(n)=|\mathbb{G}_{n}|^{-1/2}M_{\mathbb{G}_{n-p}}(g_{p,n})\quad\text{where}\quad g_{p,n}=\sum_{\ell=0}^{p}2^{p-\ell}{\mathcal{Q}}^{p-\ell}\tilde{f}_{\ell,n}.

We follow the same arguments that in the proof of Lemma 5.1. For all m{1,,np}m\in\{1,\ldots,n-p\} and for all u𝕋u\in\mathbb{T}, we set

Zp,m(u)=2m1𝒬m1gp(Xu0)+2m1𝒬m1gp(Xu1)2m𝒬mgp(Xu).Z_{p,m}(u)=2^{m-1}{\mathcal{Q}}^{m-1}g_{p}(X_{u0})+2^{m-1}{\mathcal{Q}}^{m-1}g_{p}(X_{u1})-2^{m}{\mathcal{Q}}^{m}g_{p}(X_{u}).

We also consider the following quantities for m{1,,np}m\in\{1,\ldots,n-p\} and λ>0\lambda>0:

𝕀m=𝔼[exp(λ2np𝒬npgp,n(X))k=mnpexp(λu𝔾npkZp,k(u))]and\displaystyle\mathbb{I}_{m}=\mathbb{E}\Big{[}\exp\Big{(}\lambda 2^{n-p}{\mathcal{Q}}^{n-p}g_{p,n}(X_{\emptyset})\Big{)}\prod_{k=m}^{n-p}\exp\Big{(}\lambda\sum_{u\in\mathbb{G}_{n-p-k}}Z_{p,k}(u)\Big{)}\Big{]}\quad\text{and}
𝕁m=u𝔾npm𝔼Xu[exp(λZp,m(u))].\displaystyle\mathbb{J}_{m}=\prod_{u\in\mathbb{G}_{n-p-m}}\mathbb{E}_{X_{u}}\Big{[}\exp\Big{(}\lambda Z_{p,m}(u)\Big{)}\Big{]}.

Note that using the branching Markov property, we have

(34) 𝕀m=𝔼[exp(λ2np𝒬npgp,n(X))k=m+1npexp(λu𝔾npkZp,k(u))𝕁m].\mathbb{I}_{m}=\mathbb{E}\Big{[}\exp\Big{(}\lambda 2^{n-p}{\mathcal{Q}}^{n-p}g_{p,n}(X_{\emptyset})\Big{)}\prod_{k=m+1}^{n-p}\exp\Big{(}\lambda\sum_{u\in\mathbb{G}_{n-p-k}}Z_{p,k}(u)\Big{)}\mathbb{J}_{m}\Big{]}.

As for (29)-(30), for all m{1,,np}m\in\{1,\ldots,n-p\} and u𝔾npmu\in\mathbb{G}_{n-p-m}, one can prove that

(35) |Zp,m(u)|M=Chd/2and𝔼Xu[Zp,m(u)2]σm2=C𝟏{m=1}+Chnd(=0p(2α)p+m2)2.|Z_{p,m}(u)|\leq M=Ch^{-d/2}\quad\text{and}\quad\mathbb{E}_{X_{u}}\Big{[}Z_{p,m}(u)^{2}\Big{]}\leq\sigma_{m}^{2}=C{\bf 1}_{\{m=1\}}+Ch_{n}^{d}\Big{(}\sum_{\ell=0}^{p}(2\alpha)^{p+m-\ell-2}\Big{)}^{2}.

Using (31) and (35), we have, for all u𝔾npmu\in\mathbb{G}_{n-p-m} and for all λ(0,Chd/2/3)\lambda\in(0,Ch^{-d/2}/3),

𝔼Xu[exp(λZp,m(u))]exp(λ2σm22(1λM/3)).\mathbb{E}_{X_{u}}\Big{[}\exp(\lambda Z_{p,m}(u))\Big{]}\leq\exp\Big{(}\frac{\lambda^{2}\sigma_{m}^{2}}{2(1-\lambda M/3)}\Big{)}.

The latter inequality and (34) imply that

𝕀mexp(λ2σm2|𝔾npm|2(1λM/3))𝕀m+1.\mathbb{I}_{m}\leq\exp\Big{(}\frac{\lambda^{2}\sigma_{m}^{2}|\mathbb{G}_{n-p-m}|}{2(1-\lambda M/3)}\Big{)}\mathbb{I}_{m+1}.

By recurrence, this implies that

(36) 𝕀1exp(λ2m=1npσm2|𝔾npm|2(1λM/3))𝔼[exp(λ2np𝒬npgp,n(X))].\mathbb{I}_{1}\leq\exp\Big{(}\frac{\lambda^{2}\sum_{m=1}^{n-p}\sigma_{m}^{2}|\mathbb{G}_{n-p-m}|}{2(1-\lambda M/3)}\Big{)}\,\mathbb{E}\Big{[}\exp\Big{(}\lambda 2^{n-p}{\mathcal{Q}}^{n-p}g_{p,n}(X_{\emptyset})\Big{)}\Big{]}.

Using (i)(i) and (ii)(ii) of Assumption 3.8 and Assumption 2.9, we get

(37) |gp,n|Chnd/2=0p(2α)n.|g_{p,n}|\leq Ch_{n}^{d/2}\sum_{\ell=0}^{p}(2\alpha)^{n-\ell}.

From (36), (37) and according to the value of α\alpha, we have, for some positive constants c1c_{1}, c2c_{2} and c3c_{3} (recall the definition of MM and σm2\sigma_{m}^{2} given in (35)):

𝕀1Cexp(λ2c1|𝔾np|2(1λc2hnd/2/3))if 2α1;\displaystyle\mathbb{I}_{1}\leq C\exp\Big{(}\frac{\lambda^{2}c_{1}|\mathbb{G}_{n-p}|}{2(1-\lambda c_{2}h_{n}^{-d/2}/3)}\Big{)}\quad\hskip 176.407pt\text{if $2\alpha\leq 1;$}
𝕀1exp(λc3(2α)nhnd/2)exp(λ2c1|𝔾np|(1+(2α)2phnd)2(1λc2hnd/2/3))if 1<2α2;\displaystyle\mathbb{I}_{1}\leq\exp\Big{(}\lambda c_{3}(2\alpha)^{n}h_{n}^{d/2}\Big{)}\,\exp\Big{(}\frac{\lambda^{2}c_{1}|\mathbb{G}_{n-p}|(1+(2\alpha)^{2p}h_{n}^{d})}{2(1-\lambda c_{2}h_{n}^{-d/2}/3)}\Big{)}\quad\hskip 62.59596pt\text{if $1<2\alpha\leq\sqrt{2};$}
𝕀1exp(λc3(2α)nhnd/2)exp(λ2c1|𝔾np|(1+(2α)2phnd+2p(2α2)nhnd)2(1λc2hnd/2/3))if 2α>2.\displaystyle\mathbb{I}_{1}\leq\exp\Big{(}\lambda c_{3}(2\alpha)^{n}h_{n}^{d/2}\Big{)}\,\exp\Big{(}\frac{\lambda^{2}c_{1}|\mathbb{G}_{n-p}|(1+(2\alpha)^{2p}h_{n}^{d}+2^{p}(2\alpha^{2})^{n}h_{n}^{d})}{2(1-\lambda c_{2}h_{n}^{-d/2}/3)}\Big{)}\quad\text{if $2\alpha>\sqrt{2}.$}

Recall that 𝕀1=𝔼[exp(λM𝔾np(gp,n))]\mathbb{I}_{1}=\mathbb{E}[\exp(\lambda M_{\mathbb{G}_{n-p}}(g_{p,n}))]. Using the Chernoff bound and (33), we have for all λ(0,Chnd/2/3)\lambda\in(0,Ch_{n}^{-d/2}/3) and for all δ>0\delta>0,

(bn1R1(n)>δ)exp(λbn|𝔾n|1/2δ)𝕀1.\mathbb{P}\Big{(}b_{n}^{-1}R_{1}(n)>\delta\Big{)}\leq\exp\Big{(}-\lambda b_{n}|\mathbb{G}_{n}|^{1/2}\delta\Big{)}\,\mathbb{I}_{1}.

Taking

λ={3bn|𝔾n|1/2δ2c2bn|𝔾n|1/2hnd/2δ+ 3c1|𝔾np|if 2α13bn|𝔾n|1/2δ2c2bn|𝔾n|1/2hnd/2δ+ 3c1|𝔾np|(1+(2α)2phnd)if 1<2α23bn|𝔾n|1/2δ2c2bn|𝔾n|1/2hnd/2δ+ 3c1|𝔾np|(1+(2α)2phnd+2p(2α2)nhnd)if 1<2α2,\lambda=\begin{cases}\frac{3b_{n}|\mathbb{G}_{n}|^{1/2}\delta}{2c_{2}b_{n}|\mathbb{G}_{n}|^{1/2}h_{n}^{-d/2}\delta\,+\,3c_{1}|\mathbb{G}_{n-p}|}&\text{if $2\alpha\leq 1$}\\ \frac{3b_{n}|\mathbb{G}_{n}|^{1/2}\delta}{2c_{2}b_{n}|\mathbb{G}_{n}|^{1/2}h_{n}^{-d/2}\delta\,+\,3c_{1}|\mathbb{G}_{n-p}|(1+(2\alpha)^{2p}h_{n}^{d})}&\text{if $1<2\alpha\leq\sqrt{2}$}\\ \frac{3b_{n}|\mathbb{G}_{n}|^{1/2}\delta}{2c_{2}b_{n}|\mathbb{G}_{n}|^{1/2}h_{n}^{-d/2}\delta\,+\,3c_{1}|\mathbb{G}_{n-p}|(1+(2\alpha)^{2p}h_{n}^{d}+2^{p}(2\alpha^{2})^{n}h_{n}^{d})}&\text{if $1<2\alpha\leq\sqrt{2},$}\end{cases}

and since we can do the same things for 𝔣n-{\mathfrak{f}}_{n} instead of 𝔣n{\mathfrak{f}}_{n}, we get,

if 2α1:2\alpha\leq 1:

(bn1|R1(n)|>δ)Cexp(3bn2|𝔾n|δ22(c2bn|𝔾n|1/2hnd/2δ+ 3c1|𝔾np|));\mathbb{P}\Big{(}b_{n}^{-1}|R_{1}(n)|>\delta\Big{)}\leq C\exp\Big{(}-\frac{3b_{n}^{2}|\mathbb{G}_{n}|\delta^{2}}{2(c_{2}b_{n}|\mathbb{G}_{n}|^{1/2}h_{n}^{-d/2}\delta\,+\,3c_{1}|\mathbb{G}_{n-p}|)}\Big{)};

if 1<2α2:1<2\alpha\leq\sqrt{2}:

(bn1|R1(n)|>δ)\displaystyle\mathbb{P}\Big{(}b_{n}^{-1}|R_{1}(n)|>\delta\Big{)} 2exp(c3(2α)nhnd/2bn|𝔾n|1/22c2bn|𝔾n|1/2hnd/2δ+ 3c1|𝔾np|(1+(2α)2phnd))\displaystyle\leq 2\exp\Big{(}\frac{c_{3}(2\alpha)^{n}h_{n}^{d/2}b_{n}|\mathbb{G}_{n}|^{1/2}}{2c_{2}b_{n}|\mathbb{G}_{n}|^{1/2}h_{n}^{-d/2}\delta\,+\,3c_{1}|\mathbb{G}_{n-p}|(1+(2\alpha)^{2p}h_{n}^{d})}\Big{)}
×exp(3bn2|𝔾n|δ22(c2bn|𝔾n|1/2hnd/2δ+ 3c1|𝔾np|(1+(2α)2phnd)));\displaystyle\hskip 28.45274pt\times\exp\Big{(}-\frac{3b_{n}^{2}|\mathbb{G}_{n}|\delta^{2}}{2(c_{2}b_{n}|\mathbb{G}_{n}|^{1/2}h_{n}^{-d/2}\delta\,+\,3c_{1}|\mathbb{G}_{n-p}|(1+(2\alpha)^{2p}h_{n}^{d}))}\Big{)};

if 2α>2:2\alpha>\sqrt{2}:

(bn1|R1(n)|>δ)\displaystyle\mathbb{P}\Big{(}b_{n}^{-1}|R_{1}(n)|>\delta\Big{)} 2exp(c3(2α)nhnd/2bn|𝔾n|1/22c2bn|𝔾n|1/2hnd/2δ+ 3c1|𝔾np|(1+(2α)2phnd+2p(2α2)nhnd))\displaystyle\leq 2\exp\Big{(}\frac{c_{3}(2\alpha)^{n}h_{n}^{d/2}b_{n}|\mathbb{G}_{n}|^{1/2}}{2c_{2}b_{n}|\mathbb{G}_{n}|^{1/2}h_{n}^{-d/2}\delta\,+\,3c_{1}|\mathbb{G}_{n-p}|(1+(2\alpha)^{2p}h_{n}^{d}+2^{p}(2\alpha^{2})^{n}h_{n}^{d})}\Big{)}
×exp(3bn2|𝔾n|δ22(c2bn|𝔾n|1/2hnd/2δ+ 3c1|𝔾np|(1+(2α)2phnd+2p(2α2)nhnd))).\displaystyle\hskip 7.11317pt\times\exp\Big{(}-\frac{3b_{n}^{2}|\mathbb{G}_{n}|\delta^{2}}{2(c_{2}b_{n}|\mathbb{G}_{n}|^{1/2}h_{n}^{-d/2}\delta\,+\,3c_{1}|\mathbb{G}_{n-p}|(1+(2\alpha)^{2p}h_{n}^{d}+2^{p}(2\alpha^{2})^{n}h_{n}^{d}))}\Big{)}.

Finally, applying the log\log to each of these last three inequalities, dividing by bn2b_{n}^{2}, letting nn goes to infinity and using (6) and Assumption 2.13, we get the result of Lemma 5.2. ∎

From (24), Lemmas 5.1 and 5.2, we have

(38) bn1Nn,(𝔣n)bn2superexpbn1Δn(𝔣n).b_{n}^{-1}N_{n,\emptyset}({\mathfrak{f}}_{n})\mathrel{\underset{b_{n}^{2}}{\overset{{\rm superexp}}{\scalebox{2.0}[1.0]{$\sim$}}}}b_{n}^{-1}\Delta_{n}({\mathfrak{f}}_{n}).

As a consequence, using Remark 2.7, bn1Nn,(𝔣n)b_{n}^{-1}N_{n,\emptyset}({\mathfrak{f}}_{n}) and bn1Δn(𝔣n)b_{n}^{-1}\Delta_{n}({\mathfrak{f}}_{n}) satisfy the same moderate deviation principle.

We now study the martingale part Δn(𝔣n)\Delta_{n}({\mathfrak{f}}_{n}) of the decomposition (24). The bracket V(n)V(n) of Δn(𝔣n)\Delta_{n}({\mathfrak{f}}_{n}) is defined by:

V(n)=i𝔾npn𝔼[Δn,i(𝔣n)2|i].V(n)=\sum_{i\in\mathbb{G}_{n-p_{n}}}{\mathbb{E}}\left[\Delta_{n,i}({\mathfrak{f}}_{n})^{2}|{\mathcal{F}}_{i}\right].

Using (22) and (21), we write:

(39) V(n)=|𝔾n|1i𝔾npn𝔼Xi[(=0pnM𝔾pn(f~,n))2]R2(n)=V1(n)+2V2(n)R2(n),V(n)=|\mathbb{G}_{n}|^{-1}\sum_{i\in\mathbb{G}_{n-p_{n}}}{\mathbb{E}}_{X_{i}}\left[\left(\sum_{\ell=0}^{p_{n}}M_{\mathbb{G}_{p_{n}-\ell}}(\tilde{f}_{\ell,n})\right)^{2}\right]-R_{2}(n)=V_{1}(n)+2V_{2}(n)-R_{2}(n),

with:

V1(n)\displaystyle V_{1}(n) =|𝔾n|1i𝔾npn=0pn𝔼Xi[M𝔾pn(f~,n)2],\displaystyle=|\mathbb{G}_{n}|^{-1}\sum_{i\in\mathbb{G}_{n-p_{n}}}\sum_{\ell=0}^{p_{n}}{\mathbb{E}}_{X_{i}}\left[M_{\mathbb{G}_{p_{n}-\ell}}(\tilde{f}_{\ell,n})^{2}\right],
V2(n)\displaystyle V_{2}(n) =|𝔾n|1i𝔾npn0<kpn𝔼Xi[M𝔾pn(f~,n)M𝔾pnk(f~k,n)],\displaystyle=|\mathbb{G}_{n}|^{-1}\sum_{i\in\mathbb{G}_{n-p_{n}}}\sum_{0\leq\ell<k\leq p_{n}}{\mathbb{E}}_{X_{i}}\left[M_{\mathbb{G}_{p_{n}-\ell}}(\tilde{f}_{\ell,n})M_{\mathbb{G}_{p_{n}-k}}(\tilde{f}_{k,n})\right],
R2(n)\displaystyle R_{2}(n) =i𝔾npn𝔼[Nn,i(𝔣n)|Xi]2.\displaystyle=\sum_{i\in\mathbb{G}_{n-p_{n}}}{\mathbb{E}}\left[N_{n,i}({\mathfrak{f}}_{n})|X_{i}\right]^{2}.

We have the following result.

Lemma 5.3.

Under the Assumptions of Theorem 3.10, we have R2(n)bn2superexp0.R_{2}(n)\xRightarrow[b_{n}^{2}]{\rm superexp}0.

Proof.

Using the branching Markov property, we have

R2(n)=|𝔾n|1M𝔾np(gp)withgp=(=0p2p𝒬pf~,n)2.R_{2}(n)=|\mathbb{G}_{n}|^{-1}M_{\mathbb{G}_{n-p}}(g_{p})\quad\text{with}\quad g_{p}=\Big{(}\sum_{\ell=0}^{p}2^{p-\ell}{\mathcal{Q}}^{p-\ell}\tilde{f}_{\ell,n}\Big{)}^{2}.

Using Assumption 2.9 and (i)(i) and (ii)(ii) of Assumption 3.8, we get

gp\displaystyle\|g_{p}\|_{\infty} Cf~p,n2+C(=0p12p𝒬pf~,n)2\displaystyle\leq C\|\tilde{f}_{p,n}\|^{2}_{\infty}+C\|(\sum_{\ell=0}^{p-1}2^{p-\ell}{\mathcal{Q}}^{p-\ell}\tilde{f}_{\ell,n})^{2}\|_{\infty}
Chnd+C(=0p1(2α)phnd/2)2\displaystyle\leq Ch_{n}^{-d}+C\Big{(}\sum_{\ell=0}^{p-1}(2\alpha)^{p-\ell}h_{n}^{d/2}\Big{)}^{2}
Chnd𝟏{2α1}+C(hnd+hnd(2α)2p)𝟏{2α>1}.\displaystyle\leq Ch_{n}^{-d}{\bf 1}_{\{2\alpha\leq 1\}}+C(h_{n}^{-d}+h_{n}^{d}(2\alpha)^{2p}){\bf 1}_{\{2\alpha>1\}}.

This implies that

(40) R2(n)C|𝔾n|1hnd 1{2α1}+C(|𝔾n|1hnd+(2α2)phnd|𝔾np|1) 1{2α>1}.R_{2}(n)\leq C|\mathbb{G}_{n}|^{-1}h_{n}^{-d}\,{\bf 1}_{\{2\alpha\leq 1\}}+C(|\mathbb{G}_{n}|^{-1}h_{n}^{-d}+(2\alpha^{2})^{p}h_{n}^{d}|\mathbb{G}_{n-p}|^{-1})\,{\bf 1}_{\{2\alpha>1\}}.

Recall that hn=2nγh_{n}=2^{-n\gamma} with γ(0,1/d)\gamma\in(0,1/d). Using Assumption 2.13, we conclude from (40) that R2(n)R_{2}(n) is bounded by a deterministic sequence which converge to 0. As a consequence, using Remark 2.8, we get the result of Lemma 5.3. ∎

Recall σ2\sigma^{2} given in (7). We have the following result.

Lemma 5.4.

Under the Assumptions of Theorem 3.10, we have V1(n)bn2superexpσ2.V_{1}(n)\xRightarrow[b_{n}^{2}]{\rm superexp}\sigma^{2}.

Proof.

We have the following decomposition which is a consequence of (50):

V1(n)=V3(n)+V4(n),V_{1}(n)=V_{3}(n)+V_{4}(n),

with

V3(n)\displaystyle V_{3}(n) =|𝔾n|1i𝔾np=0p2p𝒬p(f~,n2)(Xi),\displaystyle=|\mathbb{G}_{n}|^{-1}\sum_{i\in\mathbb{G}_{n-p}}\sum_{\ell=0}^{p}2^{p-\ell}\,{\mathcal{Q}}^{p-\ell}(\tilde{f}_{\ell,n}^{2})(X_{i}),
V4(n)\displaystyle V_{4}(n) =|𝔾n|1i𝔾np=0p1k=0p12p+k𝒬p1(+k)(𝒫(𝒬kf~,n2))(Xi).\displaystyle=|\mathbb{G}_{n}|^{-1}\sum_{i\in\mathbb{G}_{n-p}}\sum_{\ell=0}^{p-1}\,\sum_{k=0}^{p-\ell-1}2^{p-\ell+k}\,{\mathcal{Q}}^{p-1-(\ell+k)}\left({\mathcal{P}}\left({\mathcal{Q}}^{k}\tilde{f}_{\ell,n}\otimes^{2}\right)\right)(X_{i}).

Now, the result of Lemma 5.4 is a direct consequence of the following:

(41) V3(n)bn2superexpσ2;\displaystyle V_{3}(n)\xRightarrow[b_{n}^{2}]{\rm superexp}\sigma^{2};
(42) V4(n)bn2superexp0.\displaystyle V_{4}(n)\xRightarrow[b_{n}^{2}]{\rm superexp}0.

To end the proof, we will now prove (41) and (42).

Proof of (41)

Set

gp,n==0p2𝒬p(f~,n2μ,f~,n2)andH3[n](𝔣n)==0p2μ,f~,n2.g_{p,n}=\sum_{\ell=0}^{p}2^{-\ell}{\mathcal{Q}}^{p-\ell}(\tilde{f}^{2}_{\ell,n}-\langle\mu,\tilde{f}^{2}_{\ell,n}\rangle)\quad\text{and}\quad H_{3}^{[n]}({\mathfrak{f}}_{n})=\sum_{\ell=0}^{p}2^{-\ell}\langle\mu,\tilde{f}_{\ell,n}^{2}\rangle.

Following the same arguments that in the proof of Lemmas 5.1 and 5.2, we get after studious calculations:

if 2α1,2\alpha\leq 1,

(|V3(n)H3[n]|>δ)=(|𝔾np|1|M𝔾np(gp,n)|>δ)\displaystyle\mathbb{P}\Big{(}|V_{3}(n)-H_{3}^{[n]}|>\delta\Big{)}=\mathbb{P}\Big{(}|\mathbb{G}_{n-p}|^{-1}|M_{\mathbb{G}_{n-p}}(g_{p,n})|>\delta\Big{)}
Cexp(CpδCδhnd+3(p22p+2phnd))exp(3δ2|𝔾n|2(Cδhnd+3(p22p+2phnd)));\displaystyle\leq C\exp\Big{(}\frac{Cp\,\delta}{C\delta h_{n}^{-d}+3(p^{2}2^{-p}+2^{-p}h_{n}^{-d})}\Big{)}\exp\Big{(}-\frac{3\delta^{2}|\mathbb{G}_{n}|}{2(C\delta h_{n}^{-d}+3(p^{2}2^{-p}+2^{-p}h_{n}^{-d}))}\Big{)};

if 1<2α2,1<2\alpha\leq\sqrt{2},

(|V3(n)H3[n]|>δ)=(|𝔾np|1|M𝔾np(gp,n)|>δ)\displaystyle\mathbb{P}\Big{(}|V_{3}(n)-H_{3}^{[n]}|>\delta\Big{)}=\mathbb{P}\Big{(}|\mathbb{G}_{n-p}|^{-1}|M_{\mathbb{G}_{n-p}}(g_{p,n})|>\delta\Big{)}
exp(c3(2α)nhndδc2δ+3c1((2α2)phnd+2p))exp(3δ2|𝔾n|hnd2(c2δ+3c1((2α2)phnd+2p)));\displaystyle\leq\exp\Big{(}\frac{c_{3}(2\alpha)^{n}h_{n}^{d}\delta}{c_{2}\delta+3c_{1}((2\alpha^{2})^{p}h_{n}^{d}+2^{-p})}\Big{)}\exp\Big{(}-\frac{3\delta^{2}|\mathbb{G}_{n}|h_{n}^{d}}{2(c_{2}\delta+3c_{1}((2\alpha^{2})^{p}h_{n}^{d}+2^{-p}))}\Big{)};

if 2α>2,2\alpha>\sqrt{2},

(|V3(n)H3[n]|>δ)=(|𝔾np|1|M𝔾np(gp,n)|>δ)\displaystyle\mathbb{P}\Big{(}|V_{3}(n)-H_{3}^{[n]}|>\delta\Big{)}=\mathbb{P}\Big{(}|\mathbb{G}_{n-p}|^{-1}|M_{\mathbb{G}_{n-p}}(g_{p,n})|>\delta\Big{)}
exp(c3(2α)nhndδc2δ+3c1((2α2)nhnd+2p))exp(3δ2|𝔾n|hnd2(c2δ+3c1((2α2)nhnd+2p)));\displaystyle\leq\exp\Big{(}\frac{c_{3}(2\alpha)^{n}h_{n}^{d}\delta}{c_{2}\delta+3c_{1}((2\alpha^{2})^{n}h_{n}^{d}+2^{-p})}\Big{)}\exp\Big{(}-\frac{3\delta^{2}|\mathbb{G}_{n}|h_{n}^{d}}{2(c_{2}\delta+3c_{1}((2\alpha^{2})^{n}h_{n}^{d}+2^{-p}))}\Big{)};

Taking the log\log, dividing by bn2b_{n}^{2}, letting nn goes to the infinity and using (6) and Assumption 2.13, we get

lim supn+1bn2log(|V3(n)H3[n]|>δ)=.\limsup_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\mathbb{P}\Big{(}|V_{3}(n)-H_{3}^{[n]}|>\delta\Big{)}=-\infty.

Next, using (iii)(iii) of Assumption 3.8, we get limn+H3[n](𝔣n)=σ2.\lim_{n\rightarrow+\infty}H_{3}^{[n]}({\mathfrak{f}}_{n})=\sigma^{2}. This ends the proof of (41) since (H3[n](𝔣n))(H_{3}^{[n]}({\mathfrak{f}}_{n})) is a deterministic sequence.

Proof of (42)

We set

h,k(n)=2k𝒬p1(+k)(𝒫(𝒬kf~,n2))andH4,n==0p1k=0p1h,k(n)h_{\ell,k}^{(n)}=2^{k-\ell}\,{\mathcal{Q}}^{p-1-(\ell+k)}\big{(}{\mathcal{P}}\big{(}{\mathcal{Q}}^{k}\tilde{f}_{\ell,n}\otimes^{2}\big{)}\big{)}\quad\text{and}\quad H_{4,n}=\sum_{\ell=0}^{p-1}\sum_{k=0}^{p-\ell-1}h_{\ell,k}^{(n)}

in such a that V4(n)=|𝔾np|1M𝔾np(H4,n)V_{4}(n)=|\mathbb{G}_{n-p}|^{-1}M_{\mathbb{G}_{n-p}}(H_{4,n}). Using, (3) and (i)(i) and (ii)(ii) of Assumption 3.8, we get

|h,k(n)|2k𝒫(|𝒬kf~,n|2)C2khndα2k.|h_{\ell,k}^{(n)}|\leq 2^{k-\ell}{\mathcal{P}}(|{\mathcal{Q}}^{k}\tilde{f}_{\ell,n}|\otimes^{2})\leq C2^{k-\ell}h_{n}^{d}\alpha^{2k}.

This implies that |H4,n|cn|H_{4,n}|\leq c_{n} and then that |V4(n)|cn|V_{4}(n)|\leq c_{n}, where the sequence (cn,n)(c_{n},n\in\mathbb{N}) is defined by

cn=Chnd𝟏{2α21}+Chnd(2α2)p𝟏{2α2>1}c_{n}=Ch_{n}^{d}{\bf 1}_{\{2\alpha^{2}\leq 1\}}+Ch_{n}^{d}(2\alpha^{2})^{p}{\bf 1}_{\{2\alpha^{2}>1\}}

Using (5) and the fact that (hn,n)(h_{n},n\in\mathbb{N}) converges to 0, we get that the sequence (cn,n)(c_{n},n\in\mathbb{N}) converges to 0. Thus, we have that V4(n)V_{4}(n) is bounded by a deterministic sequence which converges to 0. Then (42) follows using Remark 2.8. ∎

Lemma 5.5.

Under the Assumptions of Theorem 3.10, we have V2(n)bn2superexp0.V_{2}(n)\xRightarrow[b_{n}^{2}]{\rm superexp}0.

Proof.

Using (51), we get:

V2(n)=V5(n)+V6(n),V_{2}(n)=V_{5}(n)+V_{6}(n),

with

V5(n)\displaystyle V_{5}(n) =|𝔾n|1i𝔾np0<kp2p𝒬pk(f~k,n𝒬kf~,n)(Xi),\displaystyle=|\mathbb{G}_{n}|^{-1}\sum_{i\in\mathbb{G}_{n-p}}\sum_{0\leq\ell<k\leq p}2^{p-\ell}{\mathcal{Q}}^{p-k}\big{(}\tilde{f}_{k,n}{\mathcal{Q}}^{k-\ell}\tilde{f}_{\ell,n}\big{)}(X_{i}),
V6(n)\displaystyle V_{6}(n) =|𝔾n|1i𝔾np0<k<pr=0pk12p+r𝒬p1(r+k)(𝒫(𝒬rf~k,nsym𝒬k+rf~,n))(Xi).\displaystyle=|\mathbb{G}_{n}|^{-1}\sum_{i\in\mathbb{G}_{n-p}}\sum_{0\leq\ell<k<p}\sum_{r=0}^{p-k-1}2^{p-\ell+r}\,{\mathcal{Q}}^{p-1-(r+k)}\big{(}{\mathcal{P}}\big{(}{\mathcal{Q}}^{r}\tilde{f}_{k,n}\otimes_{\rm sym}{\mathcal{Q}}^{k-\ell+r}\tilde{f}_{\ell,n}\big{)}\big{)}(X_{i}).

First, we set

hk,,r(n)=2r𝒬p1(r+k)(𝒫(𝒬rf~k,nsym𝒬k+rf~,n))andH6,n=0<k<pr=0pk1hk,,r(n)h_{k,\ell,r}^{(n)}=2^{r-\ell}\,{\mathcal{Q}}^{p-1-(r+k)}\big{(}{\mathcal{P}}\big{(}{\mathcal{Q}}^{r}\tilde{f}_{k,n}\otimes_{\rm sym}{\mathcal{Q}}^{k-\ell+r}\tilde{f}_{\ell,n}\big{)}\big{)}\quad\text{and}\quad H_{6,n}=\sum_{0\leq\ell<k<p}\sum_{r=0}^{p-k-1}h_{k,\ell,r}^{(n)}

in such a way that V6(n)=|𝔾np|1M𝔾np(H6,n).V_{6}(n)=|\mathbb{G}_{n-p}|^{-1}M_{\mathbb{G}_{n-p}}(H_{6,n}). Using, (3) and (i)(i) and (ii)(ii) of Assumption 3.8, we get

|hk,,r(n)|Chnd(2α2)rαk.|h_{k,\ell,r}^{(n)}|\leq Ch_{n}^{d}(2\alpha^{2})^{r}\alpha^{k-\ell}.

This implies that |H6,n|cn|H_{6,n}|\leq c_{n} and then that V6(n)cnV_{6}(n)\leq c_{n}, where the sequence (cn,n)(c_{n},n\in\mathbb{N}) is defined by

cn=Chnd𝟏{2α21}+C(2α2)phnd𝟏{2α2>1}.c_{n}=Ch_{n}^{d}{\bf 1}_{\{2\alpha^{2}\leq 1\}}+C(2\alpha^{2})^{p}h_{n}^{d}{\bf 1}_{\{2\alpha^{2}>1\}}.

Since the sequence (cn,n)(c_{n},n\in\mathbb{N}) is deterministic and converges to 0, it follows, using Remark 2.8, that

V6(n)bn2superexp0.V_{6}(n)\xRightarrow[b_{n}^{2}]{\rm superexp}0.

Next, for the term V5(n)V_{5}(n), we have for all k>k>\ell:

|2𝒬pk(f~k,n𝒬kf~,n)|\displaystyle\big{|}2^{-\ell}{\mathcal{Q}}^{p-k}\big{(}\tilde{f}_{k,n}{\mathcal{Q}}^{k-\ell}\tilde{f}_{\ell,n}\big{)}\big{|} 2𝒬pk(|f~k,n||𝒬kf~,n|)C2hnd/2αk𝒬pk(|f~k,n|)\displaystyle\leq 2^{-\ell}{\mathcal{Q}}^{p-k}\big{(}|\tilde{f}_{k,n}||{\mathcal{Q}}^{k-\ell}\tilde{f}_{\ell,n}|\big{)}\leq C2^{-\ell}h_{n}^{d/2}\alpha^{k-\ell}{\mathcal{Q}}^{p-k}(|\tilde{f}_{k,n}|)
Cαp(2α)𝟏{k=p}+Chnd(2α)αk𝟏{kp1},\displaystyle\leq C\alpha^{p}(2\alpha)^{-\ell}{\bf 1}_{\{k=p\}}+Ch_{n}^{d}(2\alpha)^{-\ell}\alpha^{k}{\bf 1}_{\{k\leq p-1\}},

where we used (3) for the second inequality and (i)(i) and (ii)(ii) of Assumption 3.8 for the second and the last inequality. Using the latter inequality in V5(n)V_{5}(n), we get

|V5(n)|C(2p𝟏{2α<1}+αp𝟏{2α1}+hnd).|V_{5}(n)|\leq C\big{(}2^{-p}{\bf 1}_{\{2\alpha<1\}}+\alpha^{p}{\bf 1}_{\{2\alpha\geq 1\}}+h_{n}^{d}\big{)}.

We thus have that V5(n)V_{5}(n) is bounded by a deterministic sequence which converges to 0. It then follows from Remark 2.8 that

V5(n)bn2superexp0.V_{5}(n)\xRightarrow[b_{n}^{2}]{\rm superexp}0.

From the foregoing, we get the result of Lemma since V2(n)=V5(n)+V6(n).V_{2}(n)=V_{5}(n)+V_{6}(n).

As a direct consequence of (39) and the Lemmas 5.3, 5.4 and 5.5, we have the following result.

Lemma 5.6.

Under the Assumptions of Theorem 3.10, we have V(n)bn2superexpσ2.V(n)\xRightarrow[b_{n}^{2}]{\rm superexp}\sigma^{2}.

We now study the 4th-order exponential moment condition. We stress that this condition imply in particular the exponential Lindeberg condition (condition (C3) in Proposition 6.1). We have the following result.

Lemma 5.7.

Under the Assumptions of Theorem 3.10, we have

lim supn+1bn2log(bn2i𝔾np𝔼[Δn,i(𝔣n)4|i]>δ)=δ>0.\limsup_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\mathbb{P}\Big{(}b_{n}^{2}\sum_{i\in\mathbb{G}_{n-p}}\mathbb{E}[\Delta_{n,i}({\mathfrak{f}}_{n})^{4}|{\mathcal{F}}_{i}]>\delta\Big{)}=-\infty\quad\forall\delta>0.
Proof.

For all i𝔾npi\in\mathbb{G}_{n-p}, we have

(43) 𝔼[Δn,i(𝔣n)4|i]16(p+1)322n=0p𝔼Xi[M𝔾p(f~,n)4],\mathbb{E}\left[\Delta_{n,i}({\mathfrak{f}}_{n})^{4}|{\mathcal{F}}_{i}\right]\leq 16(p+1)^{3}2^{-2n}\sum_{\ell=0}^{p}\mathbb{E}_{X_{i}}\left[M_{\mathbb{G}_{p-\ell}}(\tilde{f}_{\ell,n})^{4}\right],

where we have used the definition of Δn,i(𝔣n)\Delta_{n,i}({\mathfrak{f}}_{n}), the inequality (k=0rak)4(r+1)3k=0rak4(\sum_{k=0}^{r}a_{k})^{4}\leq(r+1)^{3}\sum_{k=0}^{r}a_{k}^{4} and the branching Markov property. Using (43), we get

(44) bn2i𝔾np𝔼[Δn,i(𝔣n)4|i]Cbn2p322n=0pi𝔾nphn,(Xi),b_{n}^{2}\sum_{i\in\mathbb{G}_{n-p}}\mathbb{E}[\Delta_{n,i}({\mathfrak{f}}_{n})^{4}|{\mathcal{F}}_{i}]\leq Cb_{n}^{2}p^{3}2^{-2n}\sum_{\ell=0}^{p}\sum_{i\in\mathbb{G}_{n-p}}h_{n,\ell}(X_{i}),

where hn,(x)=𝔼x[M𝔾p(f~,n)4].h_{n,\ell}(x)=\mathbb{E}_{x}[M_{\mathbb{G}_{p-\ell}}(\tilde{f}_{\ell,n})^{4}]. We will now prove that the right hand side of (44) converges superexponentially to 0 at the speed bn2b_{n}^{2}, that is

lim supn+1bn2log(Cbn2p322n|=0pi𝔾nphn,(Xi)|>δ)=.\limsup_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\mathbb{P}\Big{(}Cb_{n}^{2}p^{3}2^{-2n}\Big{|}\sum_{\ell=0}^{p}\sum_{i\in\mathbb{G}_{n-p}}h_{n,\ell}(X_{i})\Big{|}>\delta\Big{)}=-\infty.

For that purpose, we will treat the case =p\ell=p, =p1\ell=p-1 and finally the case {0,,p2}\ell\in\{0,\ldots,p-2\}. First, we treat the case =p.\ell=p. Set gp,n=f~p,n4.g_{p,n}=\tilde{f}_{p,n}^{4}. We have

(45) bn2p322ni𝔾nphn,p(Xi)=bn2p322ni𝔾npg~p,n(Xi)+bn2p322n|𝔾np|μ,gp,n.b_{n}^{2}p^{3}2^{-2n}\sum_{i\in\mathbb{G}_{n-p}}h_{n,p}(X_{i})=b_{n}^{2}p^{3}2^{-2n}\sum_{i\in\mathbb{G}_{n-p}}\tilde{g}_{p,n}(X_{i})+b_{n}^{2}p^{3}2^{-2n}\,|\mathbb{G}_{n-p}|\langle\mu,g_{p,n}\rangle.

Since bn2p322n|𝔾np|μ,gp,np32pbn2(|𝔾n|hnd)10b_{n}^{2}p^{3}2^{-2n}|\mathbb{G}_{n-p}|\langle\mu,g_{p,n}\rangle\leq p^{3}2^{-p}\,b_{n}^{2}\,(|\mathbb{G}_{n}|h_{n}^{d})^{-1}\rightarrow 0 as n0n\rightarrow 0, it suffices to prove that the first term of the right hand side in (45) converges superexponentially to 0 at the speed bn2b_{n}^{2}, that is, for all δ>0,\delta>0,

(46) lim supn+1bn2log(bn2p322n|i𝔾npg~p,n(Xi)|>δ)=.\limsup_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\mathbb{P}\Big{(}b_{n}^{2}p^{3}2^{-2n}\,|\sum_{i\in\mathbb{G}_{n-p}}\tilde{g}_{p,n}(X_{i})|>\delta\Big{)}=-\infty.

As in the proof of Lemma 5.2, we can prove that

(bn2p322n|i𝔾npg~p,n(Xi)|>δ)Cexp(δ2|𝔾n|2hn2dCp3bn2(δ+Cp3bn2(|𝔾n+p|hnd)1)).\mathbb{P}\Big{(}b_{n}^{2}p^{3}2^{-2n}\,|\sum_{i\in\mathbb{G}_{n-p}}\tilde{g}_{p,n}(X_{i})|>\delta\Big{)}\leq C\exp\Big{(}-\frac{\delta^{2}|\mathbb{G}_{n}|^{2}h_{n}^{2d}}{Cp^{3}b_{n}^{2}(\delta+Cp^{3}b_{n}^{2}(|\mathbb{G}_{n+p}|h_{n}^{d})^{-1})}\Big{)}.

Taking the log\log and dividing by bn2b_{n}^{2}, we get (46).

Next, for {0,,p1}\ell\in\{0,\ldots,p-1\}, we plan to prove that the quantity bn2p322n=0p1i𝔾nphn,(Xi)b_{n}^{2}p^{3}2^{-2n}\sum_{\ell=0}^{p-1}\sum_{i\in\mathbb{G}_{n-p}}h_{n,\ell}(X_{i}) is bounded by a deterministic sequence which converges to 0. First, for =p1\ell=p-1, using the branching Markov property, (i)(i) and (ii)(ii) of Assumption 3.8, we have, for all i𝔾np,i\in\mathbb{G}_{n-p,}

hn,p1(Xi)=𝔼Xi[M𝔾1(f~p1,n)4]C𝒬(f~p1,n4)Chnd.h_{n,p-1}(X_{i})=\mathbb{E}_{X_{i}}[M_{\mathbb{G}_{1}}(\tilde{f}_{p-1,n})^{4}]\leq C{\mathcal{Q}}(\tilde{f}_{p-1,n}^{4})\leq Ch_{n}^{-d}.

Using (6), this implies that

bn2p3 22ni𝔾nphn,p1(Xi)Cbn2 2pp3(|𝔾n|hnd)10as n+.b_{n}^{2}\,p^{3}\,2^{-2n}\sum_{i\in\mathbb{G}_{n-p}}h_{n,p-1}(X_{i})\leq C\,b_{n}^{2}\,2^{-p}\,p^{3}(|\mathbb{G}_{n}|h_{n}^{d})^{-1}\rightarrow 0\quad\text{as $n\rightarrow+\infty.$}

Now we consider the case {0,,p2}\ell\in\{0,\ldots,p-2\}. From Lemma 6.4 with ff replaced by f~,n\tilde{f}_{\ell,n} and ν\nu by the Dirac mass at XiX_{i} (δXi\delta_{X_{i}}), we have

(47) bn2p3 22n=0p2i𝔾nphn,(Xi)bn2|𝔾n|2p3=0p2i𝔾npj=19|ψj,p|(Xi).b_{n}^{2}\,p^{3}\,2^{-2n}\sum_{\ell=0}^{p-2}\sum_{i\in\mathbb{G}_{n-p}}h_{n,\ell}(X_{i})\leq b_{n}^{2}\,|\mathbb{G}_{n}|^{-2}\,p^{3}\,\sum_{\ell=0}^{p-2}\sum_{i\in\mathbb{G}_{n-p}}\sum_{j=1}^{9}|\psi_{j,p-\ell}|(X_{i}).

For all j{1,,9}j\in\{1,\ldots,9\}, we will upper bound each term of the right hand side in (47) by a deterministic sequence which converges to 0.

Upper bound of bn2|𝔾n|2p3=0pi𝔾np|ψ1,p|(Xi)b_{n}^{2}|\mathbb{G}_{n}|^{-2}p^{3}\sum_{\ell=0}^{p}\sum_{i\in\mathbb{G}_{n-p}}|\psi_{1,p-\ell}|(X_{i})

Using (i)(i) of Assumption 3.8, we have

|ψ1,p|C2p𝒬p(f,n4)C2phnd.|\psi_{1,p-\ell}|\leq C2^{p-\ell}{\mathcal{Q}}^{p-\ell}(f_{\ell,n}^{4})\leq C2^{p-\ell}h_{n}^{-d}.

Using (6), this implies that

bn2|𝔾n|2p3=0pi𝔾np|ψ1,p|(Xi)Cbn2p3(|𝔾n|hnd)10as n+.b_{n}^{2}|\mathbb{G}_{n}|^{-2}p^{3}\sum_{\ell=0}^{p}\sum_{i\in\mathbb{G}_{n-p}}|\psi_{1,p-\ell}|(X_{i})\leq Cb_{n}^{2}p^{3}(|\mathbb{G}_{n}|h_{n}^{d})^{-1}\rightarrow 0\quad\text{as $n\rightarrow+\infty.$}

Upper bound of bn2|𝔾n|3p3=0pi𝔾np|ψ2,p|(Xi)b_{n}^{2}|\mathbb{G}_{n}|^{-3}p^{3}\sum_{\ell=0}^{p}\sum_{i\in\mathbb{G}_{n-p}}|\psi_{2,p-\ell}|(X_{i})

Using Assumption 2.9 and (i)(i) and (ii)(ii) of Assumption 3.8 for the second inequality, we get

|ψ2,p|\displaystyle|\psi_{2,p-\ell}| C22(p)k=0p12k𝒬k𝒫(|𝒬pk1(f~,n3)|sym|𝒬pk2(𝒬f~,n)|)\displaystyle\leq C2^{2(p-\ell)}\sum_{k=0}^{p-\ell-1}2^{-k}{\mathcal{Q}}^{k}{\mathcal{P}}(|{\mathcal{Q}}^{p-k-1-\ell}(\tilde{f}_{\ell,n}^{3})|\otimes_{\rm sym}|{\mathcal{Q}}^{p-\ell-k-2}({\mathcal{Q}}\tilde{f}_{\ell,n})|)
C22(p)k=0p12kαpkC2p(𝟏{2α<1}+(p)𝟏{2α=1}+(2α)p𝟏{2α>1}).\displaystyle\leq C2^{2(p-\ell)}\sum_{k=0}^{p-\ell-1}2^{-k}\alpha^{p-\ell-k}\,\leq\,C2^{p-\ell}\big{(}{\bf 1}_{\{2\alpha<1\}}+(p-\ell){\bf 1}_{\{2\alpha=1\}}+(2\alpha)^{p-\ell}{\bf 1}_{\{2\alpha>1\}}\big{)}.

Using (6) and (5), this implies that

bn2|𝔾n|2p3=0pi𝔾np|ψ2,p|(Xi)Cbn2|𝔾n|1(p4𝟏{2α1}+(2α)p𝟏{2α>1})0as n+.b_{n}^{2}|\mathbb{G}_{n}|^{-2}p^{3}\sum_{\ell=0}^{p}\sum_{i\in\mathbb{G}_{n-p}}|\psi_{2,p-\ell}|(X_{i})\leq Cb_{n}^{2}|\mathbb{G}_{n}|^{-1}\big{(}p^{4}{\bf 1}_{\{2\alpha\leq 1\}}+(2\alpha)^{p}{\bf 1}_{\{2\alpha>1\}}\big{)}\rightarrow 0\quad\text{as $n\rightarrow+\infty.$}

Upper bound of bn2|𝔾n|2p3=0pi𝔾np|ψ3,p|(Xi)b_{n}^{2}|\mathbb{G}_{n}|^{-2}p^{3}\sum_{\ell=0}^{p}\sum_{i\in\mathbb{G}_{n-p}}|\psi_{3,p-\ell}|(X_{i})

Using (i)(i) and (ii)(ii) of Assumption 3.8 for the second inequality, we get

|ψ3,p|22(p)k=0p12k𝒬k𝒫(𝒬pk1(f~,n2)2)C22(p)k=0p12kC22(p).|\psi_{3,p-\ell}|\leq 2^{2(p-\ell)}\sum_{k=0}^{p-\ell-1}2^{-k}{\mathcal{Q}}^{k}{\mathcal{P}}({\mathcal{Q}}^{p-\ell-k-1}(\tilde{f}_{\ell,n}^{2})\otimes^{2})\leq C2^{2(p-\ell)}\sum_{k=0}^{p-\ell-1}2^{-k}\leq C2^{2(p-\ell)}.

Using (6), this implies that

bn2|𝔾n|2p3=0pi𝔾np|ψ3,p|(Xi)Cbn2p3 2n+p0as n+.b_{n}^{2}|\mathbb{G}_{n}|^{-2}p^{3}\sum_{\ell=0}^{p}\sum_{i\in\mathbb{G}_{n-p}}|\psi_{3,p-\ell}|(X_{i})\leq C\,b_{n}^{2}\,p^{3}\,2^{-n+p}\rightarrow 0\quad\text{as $n\rightarrow+\infty.$}

Upper bound of bn2|𝔾n|2p3=0pi𝔾np|ψ4,p|(Xi)b_{n}^{2}|\mathbb{G}_{n}|^{-2}p^{3}\sum_{\ell=0}^{p}\sum_{i\in\mathbb{G}_{n-p}}|\psi_{4,p-\ell}|(X_{i})

Using Assumption (2.9) and (i)(i) and (ii)(ii) of Assumption 3.8 for the second inequality, we get

|ψ4,p|C24(p)𝒫(𝒫(|𝒬p2f~,n|2)2)C24(p)α4(p2)hn2d.|\psi_{4,p-\ell}|\leq C2^{4(p-\ell)}{\mathcal{P}}\big{(}{\mathcal{P}}\big{(}|{\mathcal{Q}}^{p-\ell-2}\tilde{f}_{\ell,n}|\otimes^{2}\big{)}\otimes^{2}\big{)}\leq C2^{4(p-\ell)}\alpha^{4(p-\ell-2)}h_{n}^{2d}.

Using (6) and (5), this implies that

bn2|𝔾n|2p3=0pi𝔾np|ψ4,p|(Xi)C(bn2p4 2nphn2d𝟏{2α21}+bn2p32n+p(2α2)2phn2d𝟏{2α2>1})0as n+.b_{n}^{2}|\mathbb{G}_{n}|^{-2}p^{3}\sum_{\ell=0}^{p}\sum_{i\in\mathbb{G}_{n-p}}|\psi_{4,p-\ell}|(X_{i})\\ \leq C\,(b_{n}^{2}\,p^{4}\,2^{-n-p}h_{n}^{2d}{\bf 1}_{\{2\alpha^{2}\leq 1\}}+b_{n}^{2}p^{3}2^{-n+p}(2\alpha^{2})^{2p}h_{n}^{2d}{\bf 1}_{\{2\alpha^{2}>1\}})\rightarrow 0\quad\text{as $n\rightarrow+\infty.$}

Upper bound of bn2|𝔾n|3p3=0pi𝔾np|ψ5,p|(Xi)b_{n}^{2}|\mathbb{G}_{n}|^{-3}p^{3}\sum_{\ell=0}^{p}\sum_{i\in\mathbb{G}_{n-p}}|\psi_{5,p-\ell}|(X_{i})

Using Assumption (2.9) and (i)(i) and (ii)(ii) of Assumption 3.8 for the second inequality, we get

|ψ5,p|\displaystyle|\psi_{5,p-\ell}| C 24(p)k=2p1r=0k122kr𝒬r𝒫(𝒬kr1(𝒫(|𝒬pk1f~,n|2))2)\displaystyle\leq C\,2^{4(p-\ell)}\sum_{k=2}^{p-\ell-1}\sum_{r=0}^{k-1}2^{-2k-r}{\mathcal{Q}}^{r}{\mathcal{P}}\big{(}{\mathcal{Q}}^{k-r-1}\big{(}{\mathcal{P}}\big{(}|{\mathcal{Q}}^{p-\ell-k-1}\tilde{f}_{\ell,n}|\otimes^{2}\big{)}\big{)}\otimes^{2}\big{)}
C 24(p)k=2p1r=0k122krhn2dα4(pk)\displaystyle\leq C\,2^{4(p-\ell)}\sum_{k=2}^{p-\ell-1}\sum_{r=0}^{k-1}2^{-2k-r}h_{n}^{2d}\alpha^{4(p-\ell-k)}
Chn2d 22(p)(𝟏{2α2<1}+(p)𝟏{2α2=1}+(2α2)2(p)𝟏{2α2>1}).\displaystyle\leq C\,h_{n}^{2d}\,2^{2(p-\ell)}\big{(}{\bf 1}_{\{2\alpha^{2}<1\}}+(p-\ell){\bf 1}_{\{2\alpha^{2}=1\}}+(2\alpha^{2})^{2(p-\ell)}{\bf 1}_{\{2\alpha^{2}>1\}}\big{)}.

Using (6) and (5), this implies that

bn2|𝔾n|2p3=0pi𝔾np|ψ5,p|(Xi)C(bn2p4 2n+phn2d𝟏{2α21}+bn2p32n+p(2α2)2phn2d𝟏{2α2>1})0as n+.b_{n}^{2}|\mathbb{G}_{n}|^{-2}p^{3}\sum_{\ell=0}^{p}\sum_{i\in\mathbb{G}_{n-p}}|\psi_{5,p-\ell}|(X_{i})\\ \leq C\,(b_{n}^{2}\,p^{4}\,2^{-n+p}h_{n}^{2d}{\bf 1}_{\{2\alpha^{2}\leq 1\}}+b_{n}^{2}p^{3}2^{-n+p}(2\alpha^{2})^{2p}h_{n}^{2d}{\bf 1}_{\{2\alpha^{2}>1\}})\rightarrow 0\quad\text{as $n\rightarrow+\infty.$}

Upper bound of bn2|𝔾n|2p3=0pi𝔾np|ψ6,p|(Xi)b_{n}^{2}|\mathbb{G}_{n}|^{-2}p^{3}\sum_{\ell=0}^{p}\sum_{i\in\mathbb{G}_{n-p}}|\psi_{6,p-\ell}|(X_{i})

Using Assumption (2.9) and (i)(i) and (ii)(ii) of Assumption 3.8 for the second inequality, we get

|ψ6,n|\displaystyle|\psi_{6,n}| C 23(p)k=1p1r=0k12kr𝒬r𝒫(𝒬kr1𝒫(|𝒬pk1f~,n|2)sym𝒬pr1(f~,n2))\displaystyle\leq C\,2^{3(p-\ell)}\sum_{k=1}^{p-\ell-1}\sum_{r=0}^{k-1}2^{-k-r}{\mathcal{Q}}^{r}{\mathcal{P}}\big{(}{\mathcal{Q}}^{k-r-1}{\mathcal{P}}\big{(}|{\mathcal{Q}}^{p-\ell-k-1}\tilde{f}_{\ell,n}|\otimes^{2}\big{)}\otimes_{\rm sym}{\mathcal{Q}}^{p-\ell-r-1}(\tilde{f}^{2}_{\ell,n})\big{)}
C 23(p)k=1p1r=0k12krhndα2(pk)\displaystyle\leq C\,2^{3(p-\ell)}\sum_{k=1}^{p-\ell-1}\sum_{r=0}^{k-1}2^{-k-r}h_{n}^{d}\alpha^{2(p-\ell-k)}
Chnd 22(p)(𝟏{2α2<1}+(p) 1{2α2=1}+(2α2)p𝟏{2α2>1}).\displaystyle\leq C\,h_{n}^{d}\,2^{2(p-\ell)}\big{(}{\bf 1}_{\{2\alpha^{2}<1\}}\,+\,(p-\ell)\,{\bf 1}_{\{2\alpha^{2}=1\}}\,+\,(2\alpha^{2})^{p-\ell}{\bf 1}_{\{2\alpha^{2}>1\}}\big{)}.

Using (6) and (5), this implies that

bn2|𝔾n|2p3=0pi𝔾np|ψ6,p|(Xi)C(bn2p4 2n+phnd𝟏{2α21}+bn2p32n+p(2α2)2phnd𝟏{2α2>1})0as n+.b_{n}^{2}|\mathbb{G}_{n}|^{-2}p^{3}\sum_{\ell=0}^{p}\sum_{i\in\mathbb{G}_{n-p}}|\psi_{6,p-\ell}|(X_{i})\\ \leq C\,(b_{n}^{2}\,p^{4}\,2^{-n+p}h_{n}^{d}{\bf 1}_{\{2\alpha^{2}\leq 1\}}+b_{n}^{2}p^{3}2^{-n+p}(2\alpha^{2})^{2p}h_{n}^{d}{\bf 1}_{\{2\alpha^{2}>1\}})\rightarrow 0\quad\text{as $n\rightarrow+\infty.$}

Upper bound of bn2|𝔾n|2p3=0pi𝔾np|ψ7,p|(Xi)b_{n}^{2}|\mathbb{G}_{n}|^{-2}p^{3}\sum_{\ell=0}^{p}\sum_{i\in\mathbb{G}_{n-p}}|\psi_{7,p-\ell}|(X_{i})

In the same way as for ψ6,p\psi_{6,p-\ell}, we have

bn2|𝔾n|2p3=0pi𝔾np|ψ7,p|(Xi)C(bn2p4 2nhnd𝟏{2α1}+bn2p32n+p(2α2)2phnd𝟏{2α>1})0as n+.b_{n}^{2}|\mathbb{G}_{n}|^{-2}p^{3}\sum_{\ell=0}^{p}\sum_{i\in\mathbb{G}_{n-p}}|\psi_{7,p-\ell}|(X_{i})\\ \leq C\,(b_{n}^{2}\,p^{4}\,2^{-n}h_{n}^{d}{\bf 1}_{\{2\alpha\leq 1\}}+b_{n}^{2}p^{3}2^{-n+p}(2\alpha^{2})^{2p}h_{n}^{d}{\bf 1}_{\{2\alpha>1\}})\rightarrow 0\quad\text{as $n\rightarrow+\infty.$}

Upper bound of bn2|𝔾n|2p3=0pi𝔾np|ψ8,p|(Xi)b_{n}^{2}|\mathbb{G}_{n}|^{-2}p^{3}\sum_{\ell=0}^{p}\sum_{i\in\mathbb{G}_{n-p}}|\psi_{8,p-\ell}|(X_{i})

Using Assumption (2.9) and (i)(i) and (ii)(ii) of Assumption 3.8 for the second inequality, we get

|ψ8,p|\displaystyle|\psi_{8,p-\ell}| C 24(p)k=2p1r=1k1j=0r12krj𝒬j𝒫(𝒬rj1𝒫(|𝒬p1rf~,n|2)\displaystyle\leq C\,2^{4(p-\ell)}\,\sum_{k=2}^{p-\ell-1}\sum_{r=1}^{k-1}\sum_{j=0}^{r-1}2^{-k-r-j}{\mathcal{Q}}^{j}{\mathcal{P}}\big{(}{\mathcal{Q}}^{r-j-1}{\mathcal{P}}\big{(}|{\mathcal{Q}}^{p-\ell-1-r}\tilde{f}_{\ell,n}|\otimes^{2}\big{)}
sym𝒬kj1𝒫(|𝒬p1kf~,n|2))\displaystyle\hskip 142.26378pt\otimes_{\rm sym}{\mathcal{Q}}^{k-j-1}{\mathcal{P}}\big{(}|{\mathcal{Q}}^{p-\ell-1-k}\tilde{f}_{\ell,n}|\otimes^{2}\big{)}\big{)}
C 24(p)k=2p1r=1k1j=0r12krjhn2dα4(p)2r2k\displaystyle\leq C\,2^{4(p-\ell)}\,\sum_{k=2}^{p-\ell-1}\sum_{r=1}^{k-1}\sum_{j=0}^{r-1}2^{-k-r-j}h_{n}^{2d}\alpha^{4(p-\ell)-2r-2k}
Chn2d 22(p)(𝟏{2α2<1}+(p)2 1{2α2=1}+(2α2)2(p) 1{2α2>1}).\displaystyle\leq C\,h_{n}^{2d}\,2^{2(p-\ell)}\big{(}{\bf 1}_{\{2\alpha^{2}<1\}}\,+\,(p-\ell)^{2}\,{\bf 1}_{\{2\alpha^{2}=1\}}\,+\,(2\alpha^{2})^{2(p-\ell)}\,{\bf 1}_{\{2\alpha^{2}>1\}}\big{)}.

Using (6) and (5), this implies that

bn2|𝔾n|2p3=0pi𝔾np|ψ8,p|(Xi)C(bn2p5 2n+phn2d𝟏{2α21}+bn2p32n+p(2α2)2phn2d𝟏{2α2>1})0as n+.b_{n}^{2}|\mathbb{G}_{n}|^{-2}p^{3}\sum_{\ell=0}^{p}\sum_{i\in\mathbb{G}_{n-p}}|\psi_{8,p-\ell}|(X_{i})\\ \leq C\,(b_{n}^{2}\,p^{5}\,2^{-n+p}h_{n}^{2d}{\bf 1}_{\{2\alpha^{2}\leq 1\}}+b_{n}^{2}p^{3}2^{-n+p}(2\alpha^{2})^{2p}h_{n}^{2d}{\bf 1}_{\{2\alpha^{2}>1\}})\rightarrow 0\quad\text{as $n\rightarrow+\infty.$}

Upper bound of bn2|𝔾n|2p3=0pi𝔾np|ψ9,p|(Xi)b_{n}^{2}|\mathbb{G}_{n}|^{-2}p^{3}\sum_{\ell=0}^{p}\sum_{i\in\mathbb{G}_{n-p}}|\psi_{9,p-\ell}|(X_{i})

In the same way as for ψ8,p\psi_{8,p-\ell}, we have

bn2|𝔾n|2p3=0pi𝔾np|ψ9,p|(Xi)C(bn2p5 2n+phn2d𝟏{2α21}+bn2p32n+p(2α2)2phn2d𝟏{2α2>1})0as n+.b_{n}^{2}|\mathbb{G}_{n}|^{-2}p^{3}\sum_{\ell=0}^{p}\sum_{i\in\mathbb{G}_{n-p}}|\psi_{9,p-\ell}|(X_{i})\\ \leq C\,(b_{n}^{2}\,p^{5}\,2^{-n+p}h_{n}^{2d}{\bf 1}_{\{2\alpha^{2}\leq 1\}}+b_{n}^{2}p^{3}2^{-n+p}(2\alpha^{2})^{2p}h_{n}^{2d}{\bf 1}_{\{2\alpha^{2}>1\}})\rightarrow 0\quad\text{as $n\rightarrow+\infty.$}

Putting together all the upper bounds for {0,,p1}\ell\in\{0,\ldots,p-1\} and using (43) and (47), we deduce that bn2p3 22n=0p1i𝔾nphn,(Xi)b_{n}^{2}\,p^{3}\,2^{-2n}\sum_{\ell=0}^{p-1}\sum_{i\in\mathbb{G}_{n-p}}h_{n,\ell}(X_{i}) is bounded by a deterministic sequence which converges to 0. As a consequence, it follows, using Remark 2.8, that

lim supn+1bn2log(bn2p3 22n=0p1i𝔾nphn,(Xi)>δ)=.\limsup_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\mathbb{P}\Big{(}b_{n}^{2}\,p^{3}\,2^{-2n}\sum_{\ell=0}^{p-1}\sum_{i\in\mathbb{G}_{n-p}}h_{n,\ell}(X_{i})>\delta\Big{)}=-\infty.

Finally, using (43), (44), (46), we get

lim supn+1bn2log(i𝔾np𝔼[Δn,i(𝔣n)4|i]>δbn2)=δ>0.\limsup_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\mathbb{P}\Big{(}\sum_{i\in\mathbb{G}_{n-p}}\mathbb{E}[\Delta_{n,i}({\mathfrak{f}}_{n})^{4}|{\mathcal{F}}_{i}]>\frac{\delta}{b_{n}^{2}}\Big{)}=-\infty\quad\forall\delta>0.

For Chen-Ledoux type condition, we have the following result.

Lemma 5.8.

Under the assumptions of Theorem 3.10, we have

lim supn+1bn2log(|𝔾n|supi𝔾npi(|Δn,i(𝔣n)|>bn))=.\limsup_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\Big{(}|\mathbb{G}_{n}|\sup_{i\in\mathbb{G}_{n-p}}\mathbb{P}_{{\mathcal{F}}_{i}}\Big{(}|\Delta_{n,i}({\mathfrak{f}}_{n})|>b_{n}\Big{)}\Big{)}=-\infty.
Proof.

For all i𝔾npi\in\mathbb{G}_{n-p}, using (21) we have

(48) i(|Δn,i(𝔣)|>bnn)i(|Nn,i(𝔣)|>bnn/2)+i(|𝔼Xi[Nn,i(𝔣)]|>bnn/2),\mathbb{P}_{{\mathcal{F}}_{i}}\left(|\Delta_{n,i}({\mathfrak{f}})|>b_{n}\sqrt{n}\right)\leq\mathbb{P}_{{\mathcal{F}}_{i}}\left(|N_{n,i}({\mathfrak{f}})|>b_{n}\sqrt{n}/2\right)+\mathbb{P}_{{\mathcal{F}}_{i}}\left(|\mathbb{E}_{X_{i}}\left[N_{n,i}({\mathfrak{f}})\right]|>b_{n}\sqrt{n}/2\right),

with Nn,i(𝔣)N_{n,i}({\mathfrak{f}}) defined in (22). Following the proof of (32), we get

i(|Nn,i(𝔣)|>bnn2)\displaystyle\mathbb{P}_{{\mathcal{F}}_{i}}\Big{(}|N_{n,i}({\mathfrak{f}})|>\frac{b_{n}\sqrt{n}}{2}\Big{)} =Xi(|=0pM𝔾p(f~,n)|>bn|𝔾n|2)\displaystyle=\mathbb{P}_{X_{i}}\Big{(}|\sum_{\ell=0}^{p}M_{\mathbb{G}_{p-\ell}}(\tilde{f}_{\ell},n)|>\frac{b_{n}\sqrt{|\mathbb{G}_{n}|}}{2}\Big{)}
Cexp(bn2|𝔾n|2(Cbn|𝔾n|1/2hnd/2+6C|𝔾p|)).\displaystyle\leq C\exp\Big{(}-\frac{b_{n}^{2}|\mathbb{G}_{n}|}{2(Cb_{n}|\mathbb{G}_{n}|^{1/2}h_{n}^{-d/2}+6C|\mathbb{G}_{p}|)}\Big{)}.

Next, for

λ=bn|𝔾n|2(c2hnd/2bnn+3c1|𝔾p|),\lambda=\frac{b_{n}\sqrt{|\mathbb{G}_{n}|}}{2(c_{2}h_{n}^{-d/2}b_{n}\sqrt{n}+3c_{1}|\mathbb{G}_{p}|)},

we have

i(𝔼Xi[Nn,i(𝔣)]>bnn2)\displaystyle\mathbb{P}_{{\mathcal{F}}_{i}}\Big{(}\mathbb{E}_{X_{i}}\left[N_{n,i}({\mathfrak{f}})\right]>\frac{b_{n}\sqrt{n}}{2}\Big{)} =Xi(=0p2p𝒬p(f~)(Xi)>bnn|𝔾n|2)\displaystyle=\mathbb{P}_{X_{i}}\Big{(}\sum_{\ell=0}^{p}2^{p-\ell}{\mathcal{Q}}^{p-\ell}(\tilde{f}_{\ell})(X_{i})>\frac{b_{n}\sqrt{n|\mathbb{G}_{n}|}}{2}\Big{)}
exp(λbnn|𝔾n|2)𝔼Xi[exp(λ=0p2p𝒬p(f~)(Xi))]\displaystyle\leq\exp\Big{(}-\frac{\lambda b_{n}\sqrt{n|\mathbb{G}_{n}|}}{2}\Big{)}\mathbb{E}_{X_{i}}\Big{[}\exp\Big{(}\lambda\sum_{\ell=0}^{p}2^{p-\ell}{\mathcal{Q}}^{p-\ell}(\tilde{f}_{\ell})(X_{i})\Big{)}\Big{]}
Cexp(bn2|𝔾n|2(Cbn|𝔾n|1/2hnd/2+6C|𝔾p|)),\displaystyle\leq C\exp\Big{(}-\frac{b_{n}^{2}|\mathbb{G}_{n}|}{2(Cb_{n}|\mathbb{G}_{n}|^{1/2}h_{n}^{-d/2}+6C|\mathbb{G}_{p}|)}\Big{)},

where we used (49) and the branching Markov property for the first equality, Chernoff bound for the first inequality and (3) for the last inequality. Doing the same thing for 𝔣-{\mathfrak{f}} instead of 𝔣,{\mathfrak{f}}, we get

i(|𝔼Xi[Nn,i(𝔣)]|>bnn2)2Cexp(bn2|𝔾n|2(Cbn|𝔾n|1/2hnd/2+6C|𝔾p|)).\mathbb{P}_{{\mathcal{F}}_{i}}\Big{(}|\mathbb{E}_{X_{i}}\left[N_{n,i}({\mathfrak{f}})\right]|>\frac{b_{n}\sqrt{n}}{2}\Big{)}\leq 2\,C\exp\Big{(}-\frac{b_{n}^{2}|\mathbb{G}_{n}|}{2(Cb_{n}|\mathbb{G}_{n}|^{1/2}h_{n}^{-d/2}+6C|\mathbb{G}_{p}|)}\Big{)}.

From the foregoing, we get, using (48),

|𝔾n|supi𝔾npi(|Δn,i(𝔣)|>bnn)C|𝔾n|exp(bn2|𝔾n|2(Cbn|𝔾n|1/2hnd/2+6C|𝔾p|)).|\mathbb{G}_{n}|\sup_{i\in\mathbb{G}_{n-p}}\mathbb{P}_{{\mathcal{F}}_{i}}\left(|\Delta_{n,i}({\mathfrak{f}})|>b_{n}\sqrt{n}\right)\leq C\,|\mathbb{G}_{n}|\,\exp\Big{(}-\frac{b_{n}^{2}|\mathbb{G}_{n}|}{2(Cb_{n}|\mathbb{G}_{n}|^{1/2}h_{n}^{-d/2}+6C|\mathbb{G}_{p}|)}\Big{)}.

Finally, taking the log\log and dividing by bn2b_{n}^{2} in the latter inequality, we get the result of Lemma 5.8. ∎

We can now use Proposition 6.1 to deduce from Lemmas 5.6, 5.7 and 5.8 that Δn(𝔣n)\Delta_{n}({\mathfrak{f}}_{n}) satisfies a moderate deviation principle with speed bn2b_{n}^{2} and rate function II defined by: I(x)=x2/(2σ2)I(x)=x^{2}/(2\sigma^{2}) for all xx\in\mathbb{R}, with the finite variance σ2\sigma^{2} defined in (7). Using (38) and Remark 2.7, we then deduce Theorem 3.10.

6. Appendix

We recall here a simplified version of Theorem 1 in [12]. We consider the real martingale (Mn,n)(M_{n},n\in\mathbb{N}) with respect to the filtration (n,n)({\mathcal{H}}_{n},n\in\mathbb{N}) and we denote (Mn,n)(\langle M\rangle_{n},n\in\mathbb{N}) its bracket.

Proposition 6.1.

Let (bn)(b_{n}) a sequence satisfying

bnis increasing,bn+,bnn0,b_{n}\quad\text{is increasing},\quad b_{n}\longrightarrow+\infty,\quad\frac{b_{n}}{\sqrt{n}}\longrightarrow 0,

such that c(n):=n/bnc(n):=\sqrt{n}/b_{n} is non-decreasing, and define the reciprocal function c1(t)c^{-1}(t) by

c1(t):=inf{n:c(n)t}.c^{-1}(t):=\inf\{n\in\mathbb{N}:c(n)\geq t\}.

Under the following conditions:

  1. (C1)

    there exists Q+Q\in\mathbb{R}_{+}^{*} such that for all δ>0\delta>0,

    lim supn+1bn2log((|MnnQ|>δ))=,\displaystyle\limsup_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\left(\mathbb{P}\left(\left|\frac{\langle M\rangle_{n}}{n}-Q\right|>\delta\right)\right)=-\infty,

  2. (C2)

    lim supn+1bn2log(nesssup1kc1(bn+1)(|MkMk1|>bnn|k1))=,\displaystyle\limsup\limits_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\left(n\quad\underset{1\leq k\leq c^{-1}(b_{n+1})}{\rm ess\,sup}\mathbb{P}(|M_{k}-M_{k-1}|>b_{n}\sqrt{n}\Big{|}\mathcal{H}_{k-1})\right)=-\infty,

  3. (C3)

    for all a>0a>0 and for all δ>0\delta>0,

    lim supn+1bn2log((1nk=1n𝔼(|MkMk1|2𝟏{|MkMk1|anbn}|k1)>δ))=,\displaystyle\limsup_{n\rightarrow+\infty}\frac{1}{b_{n}^{2}}\log\left(\mathbb{P}\left(\frac{1}{n}\sum\limits_{k=1}^{n}\mathbb{E}\left(|M_{k}-M_{k-1}|^{2}\mathbf{1}_{\{|M_{k}-M_{k-1}|\geq a\frac{\sqrt{n}}{b_{n}}\}}\Big{|}\mathcal{H}_{k-1}\right)>\delta\right)\right)=-\infty,

(Mn/(bnn))n\left(M_{n}/(b_{n}\sqrt{n})\right)_{n\in{\mathbb{N}}} satisfies the MDP on \mathbb{R} with the speed bn2/nb_{n}^{2}/n and rate function I(x)=x22Q.\displaystyle I(x)=\frac{x^{2}}{2Q}.

We have the following many-to-one formulas. Ideas of the proofs can be found in [16] and [3].

Lemma 6.2.

Let f,g(S)f,g\in{\mathcal{B}}(S), xSx\in S and nm0n\geq m\geq 0. Assuming that all the quantities below are well defined, we have:

(49) 𝔼x[M𝔾n(f)]\displaystyle{\mathbb{E}}_{x}\left[M_{\mathbb{G}_{n}}(f)\right] =|𝔾n|𝒬nf(x)=2n𝒬nf(x),\displaystyle=|\mathbb{G}_{n}|\,{\mathcal{Q}}^{n}f(x)=2^{n}\,{\mathcal{Q}}^{n}f(x),
(50) 𝔼x[M𝔾n(f)2]\displaystyle{\mathbb{E}}_{x}\left[M_{\mathbb{G}_{n}}(f)^{2}\right] =2n𝒬n(f2)(x)+k=0n12n+k𝒬nk1(𝒫(𝒬kf𝒬kf))(x),\displaystyle=2^{n}\,{\mathcal{Q}}^{n}(f^{2})(x)+\sum_{k=0}^{n-1}2^{n+k}\,{\mathcal{Q}}^{n-k-1}\left({\mathcal{P}}\left({\mathcal{Q}}^{k}f\otimes{\mathcal{Q}}^{k}f\right)\right)(x),
(51) 𝔼x[M𝔾n(f)M𝔾m(g)]\displaystyle{\mathbb{E}}_{x}\left[M_{\mathbb{G}_{n}}(f)M_{\mathbb{G}_{m}}(g)\right] =2n𝒬m(g𝒬nmf)(x)\displaystyle=2^{n}{\mathcal{Q}}^{m}\left(g{\mathcal{Q}}^{n-m}f\right)(x)
+k=0m12n+k𝒬mk1(𝒫(𝒬kgsym𝒬nm+kf))(x).\displaystyle\hskip 56.9055pt+\sum_{k=0}^{m-1}2^{n+k}\,{\mathcal{Q}}^{m-k-1}\left({\mathcal{P}}\left({\mathcal{Q}}^{k}g\otimes_{\rm sym}{\mathcal{Q}}^{n-m+k}f\right)\right)(x).

We recall the following result due to Bochner (see [21, Theorem 1A] which can be easily extended to any dimension d1d\geq 1).

Lemma 6.3.

Let (hn,n)(h_{n},n\in\mathbb{N}) be a sequence of positive numbers converging to 0 as nn goes to infinity. Let g:dg:\mathbb{R}^{d}\rightarrow\mathbb{R} be a measurable function such that d|g(x)|𝑑x<+\int_{\mathbb{R}^{d}}|g(x)|dx<+\infty. Let f:df:\mathbb{R}^{d}\rightarrow\mathbb{R} be a measurable function such that f<+\mathop{\parallel\!f\!\parallel}\nolimits_{\infty}<+\infty, d|f(y)|𝑑y<+\int_{{\mathbb{R}}^{d}}|f(y)|\,dy<+\infty and lim|x|+|x|f(x)=0\lim_{|x|\rightarrow+\infty}|x|f(x)=0. Define

gn(x)=hnddf(hn1(xy))g(y)𝑑y.g_{n}(x)=h_{n}^{-d}\int_{\mathbb{R}^{d}}f(h_{n}^{-1}(x-y))g(y)dy.

Then, we have at every point xx of continuity of gg,

limn+gn(x)=g(x)f(y)𝑑y.\lim_{n\rightarrow+\infty}g_{n}(x)=g(x)\int_{\mathbb{R}}f(y)dy.

We also give some bounds on 𝔼x[M𝔾n(f)4]{\mathbb{E}}_{x}\left[M_{\mathbb{G}_{n}}(f)^{4}\right], see the proof of Theorem 2.1 in [3]. We will use the notation:

g2=gg.g\otimes^{2}=g\otimes g.
Lemma 6.4.

There exists a finite constant CC such that for all f(S)f\in{\mathcal{B}}(S),nn\in{\mathbb{N}} and ν\nu a probability measure on SS, assuming that all the quantities below are well defined, there exist functions ψj,n\psi_{j,n} for 1j91\leq j\leq 9 such that:

𝔼ν[M𝔾n(f)4]=j=19ν,ψj,n,{\mathbb{E}}_{\nu}\left[M_{\mathbb{G}_{n}}(f)^{4}\right]=\sum_{j=1}^{9}\langle\nu,\psi_{j,n}\rangle,

and, with hk=𝒬k1(f)h_{k}={\mathcal{Q}}^{k-1}(f) and (notice that either |ψj||\psi_{j}| or |ν,ψj||\langle\nu,\psi_{j}\rangle| is bounded), writing νg=ν,g\nu g=\langle\nu,g\rangle:

|ψ1,n|\displaystyle|\psi_{1,n}| C 2n𝒬n(f4),\displaystyle\leq C\,2^{n}{\mathcal{Q}}^{n}(f^{4}),
|νψ2,n|\displaystyle|\nu\psi_{2,n}| C 22nk=0n12k|ν𝒬k𝒫(𝒬nk1(f3)symhnk)|,\displaystyle\leq C\,2^{2n}\,\sum_{k=0}^{n-1}2^{-k}|\nu{\mathcal{Q}}^{k}{\mathcal{P}}\left({\mathcal{Q}}^{n-k-1}(f^{3})\otimes_{\rm sym}h_{n-k}\right)|,
|ψ3,n|\displaystyle|\psi_{3,n}| C22nk=0n12k𝒬k𝒫(𝒬nk1(f2)2),\displaystyle\leq C2^{2n}\sum_{k=0}^{n-1}2^{-k}\,{\mathcal{Q}}^{k}{\mathcal{P}}\left({\mathcal{Q}}^{n-k-1}(f^{2})\otimes^{2}\right),
|ψ4,n|\displaystyle|\psi_{4,n}| C 24n𝒫(|𝒫(hn12)2|),\displaystyle\leq C\,2^{4n}\,{\mathcal{P}}\left(|{\mathcal{P}}(h_{n-1}\otimes^{2})\otimes^{2}|\right),
|ψ5,n|\displaystyle|\psi_{5,n}| C 24nk=2n1r=0k122kr𝒬r𝒫(𝒬kr1|𝒫(hnk2)|2),\displaystyle\leq C\,2^{4n}\,\sum_{k=2}^{n-1}\sum_{r=0}^{k-1}2^{-2k-r}{\mathcal{Q}}^{r}{\mathcal{P}}\left({\mathcal{Q}}^{k-r-1}|{\mathcal{P}}(h_{n-k}\otimes^{2})|\otimes^{2}\right),
|ψ6,n|\displaystyle|\psi_{6,n}| C 23nk=1n1r=0k12kr𝒬r|𝒫(𝒬kr1𝒫(hnk2)sym𝒬nr1(f2))|,\displaystyle\leq C\,2^{3n}\,\sum_{k=1}^{n-1}\sum_{r=0}^{k-1}2^{-k-r}{\mathcal{Q}}^{r}|{\mathcal{P}}\left({\mathcal{Q}}^{k-r-1}{\mathcal{P}}\left(h_{n-k}\otimes^{2}\right)\otimes_{\rm sym}{\mathcal{Q}}^{n-r-1}(f^{2})\right)|,
|νψ7,n|\displaystyle|\nu\psi_{7,n}| C 23nk=1n1r=0k12kr|ν𝒬r𝒫(𝒬kr1𝒫(hnksym𝒬nk1(f2))symhnr)|,\displaystyle\leq C\,2^{3n}\,\sum_{k=1}^{n-1}\sum_{r=0}^{k-1}2^{-k-r}|\nu{\mathcal{Q}}^{r}{\mathcal{P}}\left({\mathcal{Q}}^{k-r-1}{\mathcal{P}}\left(h_{n-k}\otimes_{\rm sym}{\mathcal{Q}}^{n-k-1}(f^{2})\right)\otimes_{\rm sym}h_{n-r}\right)|,
|ψ8,n|\displaystyle|\psi_{8,n}| C 24nk=2n1r=1k1j=0r12krj𝒬j𝒫(|𝒬rj1𝒫(hnr2)|sym|𝒬kj1𝒫(hnk2)|),\displaystyle\leq C\,2^{4n}\,\sum_{k=2}^{n-1}\sum_{r=1}^{k-1}\sum_{j=0}^{r-1}2^{-k-r-j}{\mathcal{Q}}^{j}{\mathcal{P}}\left(|{\mathcal{Q}}^{r-j-1}{\mathcal{P}}\left(h_{n-r}\otimes^{2}\right)|\otimes_{\rm sym}|{\mathcal{Q}}^{k-j-1}{\mathcal{P}}\left(h_{n-k}\otimes^{2}\right)|\right),
|ψ9,n|\displaystyle|\psi_{9,n}| C 24nk=2n1r=1k1j=0r12krj𝒬j|𝒫(𝒬rj1|𝒫(hnrsym𝒬kr1𝒫(hnk2))symhnj)|.\displaystyle\leq C\,2^{4n}\,\sum_{k=2}^{n-1}\sum_{r=1}^{k-1}\sum_{j=0}^{r-1}2^{-k-r-j}{\mathcal{Q}}^{j}|{\mathcal{P}}\left({\mathcal{Q}}^{r-j-1}|{\mathcal{P}}\left(h_{n-r}\otimes_{\rm sym}{\mathcal{Q}}^{k-r-1}{\mathcal{P}}\left(h_{n-k}\otimes^{2}\right)\right)\otimes_{\rm sym}h_{n-j}\right)|.

References

  • [1] I. V. Basawa and J. Zhou. Non-Gaussian bifurcating models and quasi-likelihood estimation. Adv. in Appl. Probab., 41(A):55–64, 2004.
  • [2] S. V. Bitseki Penda and J.-F. Delmas. Central limit theorem for kernel estimator of invariant density in bifurcating markov chains models. arXiv preprint arXiv:2106.08626, 2021.
  • [3] S. V. Bitseki Penda, H. Djellout, and A. Guillin. Deviation inequalities, moderate deviations and some limit theorems for bifurcating Markov chains with application. Ann. Appl. Probab., 24(1):235–291, 2014.
  • [4] S. V. Bitseki Penda and G. Gackou. Moderate deviation principles for bifurcating markov chains: case of functions dependent of one variable. arXiv e-prints, pages arXiv–2105, 2021.
  • [5] S. V. Bitseki Penda, M. Hoffmann, and A. Olivier. Adaptive estimation for bifurcating Markov chains. Bernoulli, 23(4B):3598–3637, 2017.
  • [6] S. V. Bitseki Penda and A. Olivier. Autoregressive functions estimation in nonlinear bifurcating autoregressive models. Stat. Inference Stoch. Process., 20(2):179–210, 2017.
  • [7] S. V. Bitseki Penda and A. Olivier. Moderate deviation principle in nonlinear bifurcating autoregressive models. Statistics & Probability Letters, 138:20–26, 2018.
  • [8] S. V. Bitseki Penda and A. Roche. Local bandwidth selection for kernel density estimation in a bifurcating Markov chain model. Journal of Nonparametric Statistics, 0(0):1–28, 2020.
  • [9] R. Cowan and R. Staudte. The bifurcating autoregression model in cell lineage studies. Biometrics, 42(4):769–783, December 1986.
  • [10] J.-F. Delmas and L. Marsalle. Detection of cellular aging in a Galton-Watson process. Stochastic Process. Appl., 120(12):2495–2519, 2010.
  • [11] A. Dembo and O. Zeitouni. Large Deviations Techniques and Applications. Applications of mathematics. Springer, 1998.
  • [12] H. Djellout. Moderate deviations for martingale differences and applications to o-mixing sequences. Stochastics and stochastics reports, 73(1):37–64, 2002.
  • [13] M. Doumic, M. Hoffmann, N. Krell, and L. Robert. Statistical estimation of a growth-fragmentation model observed on a genealogical tree. Bernoulli, 21(3):1760–1799, 2015.
  • [14] M. Duflo. Random iterative models, volume 34. Springer Science & Business Media, 2013.
  • [15] F. Gao. Moderate deviations and large deviations for kernel density estimators. Journal of Theoretical Probability, 16(2):401–418, 2003.
  • [16] J. Guyon. Limit theorems for bifurcating Markov chains. Application to the detection of cellular aging. Ann. Appl. Probab., 17(5-6):1538–1569, 2007.
  • [17] M. Hoffmann and A. Marguet. Statistical estimation in a randomly structured branching population. Stochastic Process. Appl., 129(12):5236–5277, 2019.
  • [18] E. Masry. Recursive probability density estimation for weakly dependent stationary processes. IEEE Transactions on Information Theory, 32(2):254–267, 1986.
  • [19] A. Mokkadem and M. Pelletier. Confidence bands for densities, logarithmic point of view. Alea, 2:231–266, 2006.
  • [20] A. Mokkadem, M. Pelletier, and J. Worms. Large and moderate deviations principles for kernel estimation of a multivariate density and its partial derivatives. Australian & New Zealand Journal of Statistics, 47(4):489–502, 2005.
  • [21] E. Parzen. On estimation of a probability density function and mode. The Annals of Mathematical Statistics, 33(3):1065–1076, 1962.
  • [22] G. G. Roussas. Nonparametric estimation in Markov processes. Annals of the Institute of Statistical Mathematics, 21(1):73–87, 1969.