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

Large deviations for the volume
of hyperbolic kk-nearest neighbor balls

Christian Hirsch Moritz Otto Department of Mathematics, Aarhus University, Ny Munkegade 118, 8000 Aarhus C, Denmark [email protected], [email protected] DIGIT Center, Aarhus University, Finlandsgade 22, 8200 Aarhus N, Denmark Takashi Owada Department of Statistics
Purdue University
West Lafayette, 47907, USA
[email protected]
 and  Christoph Thäle Faculty of Mathematics
Ruhr University Bochum
44780, Bochum, Germany.
[email protected]
Abstract.

We prove a large deviation principle for the point process of large Poisson kk-nearest neighbor balls in hyperbolic space. More precisely, we consider a stationary Poisson point process of unit intensity in a growing sampling window in hyperbolic space. We further take a growing sequence of thresholds such that there is a diverging expected number of Poisson points whose kk-nearest neighbor ball has a volume exceeding this threshold. Then, the point process of exceedances satisfies a large deviation principle whose rate function is described in terms of a relative entropy. The proof relies on a fine coarse-graining technique such that inside the resulting blocks the exceedances are approximated by independent Poisson point processes.

Key words and phrases:
hyperbolic space, large deviation principle, nearest neighbor balls, Poisson point process, stochastic geometry.
2010 Mathematics Subject Classification:
Primary 60F10. Secondary 51M10, 52A55, 60D05, 60G55.

1. Introduction

Large deviations theory is one of the most classical subfields of probability theory with a wide range of applications in information theory, statistical physics, and rare-event simulation [8]. While the most refined results are available in the study of sequences of random variables and of time-dependent stochastic processes, the understanding of large deviations in spatial random systems is still in its infancy. Even the most basic statistics such as the edge counts in a random geometric graph can give rise to surprising and highly non-trivial condensation effects on the level of large deviations [6].

One of the earliest achievements in this context is [13], which proves a large deviation principle (LDP) for statistics of a homogeneous Poisson point process in d{\mathbb{R}}^{d} under rather restrictive boundedness and locality assumptions. Later, these conditions could be relaxed substantially so as to allow for statistics with bounded exponential moments and stabilizing score functions [25, 26]. While the main focus there was on scalar and measure-valued LDPs in the thermodynamic regime, very recently also geometric statistics in the sparse and dense regimes of the Poisson point process were considered [16, 17].

All of the aforementioned results have in common that they deal with random geometric systems in Euclidean space. However, during the last years the hyperbolic space has received substantial attention in the context of complex networks [11]. Moreover, there has been vigorous activity to understand the asymptotic behavior of geometric functionals of random set systems in hyperbolic space as well. While there has been substantial progress for central and non-central limit theorems and Poisson approximation results [1, 2, 3, 10, 12, 14, 15, 18, 21, 22], large deviation principles have not been considered so far. By investigating the volume of kk-nearest neighbor (kNN) balls in hyperbolic space in the present work, we provide a first step in this direction and complement the recent findings [21] about their extremal behavior.

The general proof strategy is to extend and to develop further the ideas of [16] on a coarse graining scheme to introduce a blocked point process. Inside each block, a Poisson approximation theorem in the spirit of [21] is used to replace the original functional by a Poisson point process for which the LDP is given by Sanov’s theorem. However, the intricate geometry of the hyperbolic space makes it far more challenging to implement the blocking argument in comparison to the Euclidean setting considered in [16]. For instance, in the Euclidean setting the kissing number is finite so that in a box of side length of order r>0r>0 only a uniformly bounded number of points with a nearest neighbor radius exceeding rr can be placed. In contrast, in hyperbolic space, this number grows exponentially in rr; see [9]. Moreover, in the half-space model of hyperbolic space, the Poisson intensity is inhomogeneous in the vertical direction when considered in Euclidean terms, while in the d{\mathbb{R}}^{d}-setting there is homogeneity in all directions. We deal with these problems by deriving more refined exponential moment bounds and analyzing the point configuration in vertical layers that individually can be considered approximately homogeneous.

We believe that the techniques developed in the present article open the door to the investigation of further LDPs in hyperbolic space. In particular, we think of extending the results for the Euclidean component counts from [17] in the sparse regime. Here, we note that while the study of component counts is restricted to trees in the dense regime [22] such constraints are no longer present in the sparse regime.

The rest of the manuscript is organized as follows. In Section 2, we properly introduce the considered model and state the LDP for the volumes of large kNN balls. Next, in Section 3, we give a proof outline, where we reduce the assertion to two key auxiliary results, namely Propositions 3.1 and 3.2. These results are proven separately in Sections 4 and 5, respectively.

2. Model and main result

By the hyperbolic space d\mathbb{H}^{d} of dimension d2d\geq 2 we mean the unique simply connected, dd-dimensional Riemannian manifold of constant sectional curvature 1-1, cf. [5, 24]. There are several models to represent d\mathbb{H}^{d} in the dd-dimensional Euclidean space d{\mathbb{R}}^{d}. To carry out our computations, we will work in this paper with the so-called half-space model, but we emphasize that all results we derive are actually model independent; for background material on half-space model of hyperbolic space we refer to [24, Chapter 4.6]. In the half-space model, we identify d\mathbb{H}^{d} with the product space d=d1×(0,)\mathbb{H}^{d}={\mathbb{R}}^{d-1}\times(0,\infty). The Riemannian metric is then determined by

ds2=dx12++dxd12+dy2y2,(x1,,xd1)d1,y(0,),\operatorname{d\!}s^{2}=\frac{\operatorname{d\!}x_{1}^{2}+\ldots+\operatorname{d\!}x_{d-1}^{2}+\operatorname{d\!}y^{2}}{y^{2}},\qquad(x_{1},\ldots,x_{d-1})\in{\mathbb{R}}^{d-1},\,y\in(0,\infty),

and we denote by 𝖽𝗂𝗌𝗍𝗁𝗒𝗉(z1,z2)\mathsf{dist}_{\mathsf{hyp}}(z_{1},z_{2}) the hyperbolic distance of two points z1=(x1,y1),z2=(x2,y2)dz_{1}=(x_{1},y_{1}),z_{2}=(x_{2},y_{2})\in\mathbb{H}^{d}, which in terms of the Euclidean distance 𝖽𝗂𝗌𝗍𝖾𝗎𝖼(z1,z2)\mathsf{dist}_{\mathsf{euc}}(z_{1},z_{2}) between z1z_{1} and z2z_{2} is given by

𝖽𝗂𝗌𝗍𝗁𝗒𝗉(z1,z2)=arcosh(1+𝖽𝗂𝗌𝗍𝖾𝗎𝖼(z1,z2)22y1y2),\mathsf{dist}_{\mathsf{hyp}}(z_{1},z_{2})={\rm arcosh}\Big{(}1+{\mathsf{dist}_{\mathsf{euc}}(z_{1},z_{2})^{2}\over 2y_{1}y_{2}}\Big{)},

see [24, Theorem 4.6.1]. According to [24, Theorem 4.6.7], the hyperbolic volume measure V𝗁𝗒𝗉V_{\mathsf{hyp}} on d\mathbb{H}^{d} has density

(2.1) dx1dxd1dyyd\displaystyle{\operatorname{d\!}x_{1}\cdots\operatorname{d\!}x_{d-1}\operatorname{d\!}y\over y^{d}}

with respect to the Lebesgue measure on d1×(0,){\mathbb{R}}^{d-1}\times(0,\infty) and we will use the notation

|B|𝗁𝗒𝗉:=V𝗁𝗒𝗉(B)=Byddx1dxd1dy|B|_{\mathsf{hyp}}\mathrel{\mathop{\mathchar 58\relax}}=V_{\mathsf{hyp}}(B)=\int_{B}y^{-d}\operatorname{d\!}x_{1}\cdots\operatorname{d\!}x_{d-1}\operatorname{d\!}y

for the hyperbolic volume of a measurable set BdB\subseteq\mathbb{H}^{d}. In contrast, we shall write ||𝗅𝖾𝖻|\,\cdot\,|_{\mathsf{leb}} for the Lebesgue measure of the appropriate dimension, which will always be clear from the context. In addition, we denote for r>0r>0 and xdx\in\mathbb{H}^{d} by Br(x):={zd:𝖽𝗂𝗌𝗍𝗁𝗒𝗉(x,z)r}B_{r}(x)\mathrel{\mathop{\mathchar 58\relax}}=\{z\in\mathbb{H}^{d}\mathrel{\mathop{\mathchar 58\relax}}\mathsf{dist}_{\mathsf{hyp}}(x,z)\leq r\} the hyperbolic ball of radius rr centered at xx. Its hyperbolic volume satisfies

(2.2) |Br(x)|𝗁𝗒𝗉=βd0rsinhd1udu,|B_{r}(x)|_{\mathsf{hyp}}=\beta_{d}\int_{0}^{r}\sinh^{d-1}u\,\operatorname{d\!}u,

independently of xx, with βd:=2πd/2/Γ(d2)\beta_{d}\mathrel{\mathop{\mathchar 58\relax}}=2\pi^{d/2}/\Gamma({d\over 2}) being the surface content of the (d1)(d-1)-dimensional Euclidean unit sphere; see [24, page 79]. In particular, |Br(x)|𝗁𝗒𝗉|B_{r}(x)|_{\mathsf{hyp}} grows like a constant multiple of er(d1)e^{r(d-1)}, as rr\uparrow\infty.

Henceforth, we let 𝒫=i1δXi\mathcal{P}=\sum_{i\geq 1}\delta_{X_{i}} be the random counting measure distributed as a Poisson point process with the hyperbolic volume measure as its intensity measure. We note that such 𝒫\mathcal{P} is stationary in the sense that its distribution is invariant with respect to the full group of hyperbolic isometries. In particular, 𝒫(B)\mathcal{P}(B) is the number of points of 𝒫\mathcal{P} falling into a measurable set BdB\subseteq\mathbb{H}^{d}. In this paper, we write 𝒫B\mathcal{P}_{B} or sometimes 𝒫B\mathcal{P}\cap B for the restriction of the measure 𝒫\mathcal{P} to BB. In what follows, we shall restrict 𝒫\mathcal{P} to the family of sampling windows

Wλ:=[0,1]d1×[eλ,),λ>0,W_{\lambda}\mathrel{\mathop{\mathchar 58\relax}}=[0,1]^{d-1}\times[e^{-\lambda},\infty),\qquad\lambda>0,

whose hyperbolic volume is given by |Wλ|𝗁𝗒𝗉=eλyddy=1d1e(d1)λ|W_{\lambda}|_{\mathsf{hyp}}=\int_{e^{-\lambda}}^{\infty}y^{-d}\,\operatorname{d\!}y={1\over d-1}e^{(d-1)\lambda}.

For k1k\geq 1 we study the asymptotics of large kk-nearest neighbor (kNN) balls centered in WλW_{\lambda}. To make this precise, we define the point process

ξk,λ:=x𝒫Wλδ|BRk(x)(x)|𝗁𝗒𝗉vλ,\xi_{k,\lambda}\mathrel{\mathop{\mathchar 58\relax}}=\sum_{x\in\mathcal{P}_{W_{\lambda}}}\delta_{|B_{R_{k}(x)}(x)|_{\mathsf{hyp}}-v_{\lambda}},

where

Rk(x):=inf{r0:𝒫(Br(x))k+1}R_{k}(x)\mathrel{\mathop{\mathchar 58\relax}}=\inf\{r\geq 0\colon\mathcal{P}(B_{r}(x))\geq k+1\}

and (vλ)λ>0(v_{\lambda})_{\lambda>0} is a threshold sequence satisfying

vλ(d1)λ(k1)logλ,λ.v_{\lambda}-(d-1)\lambda-(k-1)\log\lambda\to-\infty,\qquad\lambda\uparrow\infty.

In particular, the expected number of exceedances in the window WλW_{\lambda} is of order uλ:=|Wλ|𝗁𝗒𝗉evλvλk1u_{\lambda}\mathrel{\mathop{\mathchar 58\relax}}=|W_{\lambda}|_{\mathsf{hyp}}e^{-v_{\lambda}}v_{\lambda}^{k-1}.

Our main result is a large deviation principle (LDP) for the volumes of large kNN balls. We recall that a family of random variables (Xλ)λ>0(X_{\lambda})_{\lambda>0}, defined on some probability space (Ω,,)(\Omega,\mathcal{F},\mathbb{P}) and taking values in a Polish space 𝒳{\mathcal{X}}, satisfies an LDP with speed sλs_{\lambda}\uparrow\infty and rate function :𝒳[0,]\mathcal{I}\mathrel{\mathop{\mathchar 58\relax}}{\mathcal{X}}\to[0,\infty], provided that \mathcal{I} is lower semicontinuous and if for each measurable set B𝒳B\subseteq{\mathcal{X}} one has that

infxB(x)lim infλsλ1log(XλB)lim supλsλ1log(XλB)infxB¯(x),\displaystyle-\inf_{x\in B^{\circ}}\mathcal{I}(x)\leq\liminf_{\lambda\uparrow\infty}s_{\lambda}^{-1}\log\mathbb{P}(X_{\lambda}\in B)\leq\limsup_{\lambda\uparrow\infty}s_{\lambda}^{-1}\log\mathbb{P}(X_{\lambda}\in B)\leq-\inf_{x\in\bar{B}}\mathcal{I}(x),

where BB^{\circ} and B¯\bar{B} stand for the interior and the closure of BB, respectively. To present our main result, we fix s0s_{0}\in{\mathbb{R}} and introduce the space E0:=[s0,)E_{0}\mathrel{\mathop{\mathchar 58\relax}}=[s_{0},\infty) as well as the measure τk\tau_{k} on E0E_{0} which has Lebesgue density eu/(k1)!e^{-u}/(k-1)!, u>s0u>s_{0}. By (E0)\mathcal{M}(E_{0}) we denote the space of finite measures on E0E_{0}, which supplied with the weak topology becomes a Polish space [7, Proposition A2.5.III]. For ρ(E0)\rho\in\mathcal{M}(E_{0}) we let

(2.3) H(ρ|τk):={E0logdρdτk(𝐱)ρ(d𝐱)ρ(E0)+τk(E0) if ρτk, otherwiseH(\rho|\tau_{k})\mathrel{\mathop{\mathchar 58\relax}}=\begin{cases}\int_{E_{0}}\log\frac{\operatorname{d\!}\rho}{\operatorname{d\!}\tau_{k}}({\bf x})\rho(\operatorname{d\!}{\bf x})-\rho(E_{0})+\tau_{k}(E_{0})&\text{ if }\rho\ll\tau_{k},\\ \infty&\text{ otherwise}\end{cases}

be the relative entropy entropy of ρ\rho with respect to τk\tau_{k}, where ρτk\rho\ll\tau_{k} indicates that ρ\rho is absolutely continuous with respect to τk\tau_{k} and dρdτk\frac{\operatorname{d\!}\rho}{\operatorname{d\!}\tau_{k}} denotes the corresponding Radon-Nikodym derivative; see [13, Equation (2.10)]. We are now prepared to present the main result of this paper.

Theorem 2.1 (LDP for kNN balls).

Let k1k\geq 1. Then, the family of random measures ξk,λ/uλ\xi_{k,\lambda}/u_{\lambda} satisfies an LDP on (E0)\mathcal{M}(E_{0}) with speed uλu_{\lambda} and rate function H(|τk)H(\,\cdot\,|\tau_{k}).

As highlighted in Section 1, Theorem 2.1 can be seen as the hyperbolic counterpart of [16, Theorem 2.1], which concerns large deviations of the empirical measure of recentered and rescaled kNN balls in Euclidean space. On a formal level, in the hyperbolic setting, we found it more convenient to consider a scaling where the expected number of Poisson points in the window grows exponentially in λ\lambda, whereas in [16] the corresponding scaling is linear in parameter nn. After this rescaling, our condition on the growth of vλv_{\lambda} corresponds precisely to the growth of ana_{n} in [16, Equation (2.1)]. Moreover, in the Euclidean setting the dense scaling in [16] can be equivalently transformed to the regime of growing windows considered here. However, we stress that the regimes of dense points and growing windows are no longer equivalent in hyperbolic setting, and only the growing-window asymptotics reflects the negative curvature effects from the hyperbolic space.

While the previous paragraph illustrates that there is a formal correspondence between Theorem 2.1 and the Euclidean analog [16, Theorem 2.1], the hyperbolic geometry creates substantial complications, which require us to develop novel methodological tools. First, while both Theorem 2.1 and [16, Theorem 2.1] rely on a Poisson approximation result for large kNN balls, such a result is substantially harder to establish in hyperbolic geometry. We therefore adapt the arguments of a very recent work [21]. Moreover, when establishing a central uniform integrability property, [16] relies crucially on a kissing-number argument. More precisely, while in Euclidean space, the number of non-intersecting balls of radius rr that can be put into a box of side length of order rr is uniformly bounded independently of rr, such a property does not hold in hyperbolic space. Therefore, we need to develop delicate exponential moment bound that ensure uniform integrability despite a potentially unbounded number of balls.

To make the presentation more accessible, we provide a brief summary of the notation used throughout this paper, most terms will formally be defined in the forthcoming sections:

- #()\#(\,\cdot\,): cardinality of a set
- 𝟙{}{\mathbbm{1}}\{\,\cdot\,\}: indicator function
- o(),O()o(\,\cdot\,),O(\,\cdot\,): Landau symbols
- (Ω,,)(\Omega,\mathcal{F},\mathbb{P}): underlying probability space
- 𝔼[]\mathbb{E}[\,\cdot\,]: expectation(integration) wrt. \mathbb{P}
- ()\mathcal{L}(\,\cdot\,): law of a random element
- δ()\delta_{(\,\cdot\,)}: Dirac measure
- ()\mathcal{M}(\,\cdot\,): space of locally finite measures
- d\mathbb{H}^{d}: dd-dimensional hyperbolic space,
x  identified with d1×(0,){\mathbb{R}}^{d-1}\times(0,\infty)
- H(|τ)H(\,\cdot\,|\tau): relative entropy of a measure wrt. τ\tau
- ωA\omega_{A}: restriction of a measure to a set AA
- Br(x)B_{r}(x):hyperbolic ball of radius rr centered at xx
- ||𝗁𝗒𝗉|\,\cdot\,|_{\mathsf{hyp}}: hyperbolic volume measure
x             with differential dV𝗁𝗒𝗉\operatorname{d\!}V_{\mathsf{hyp}}
- ||𝗅𝖾𝖻|\,\cdot\,|_{\mathsf{leb}}: Lebesgue measure
- 𝖽𝗂𝗌𝗍𝗁𝗒𝗉(,)\mathsf{dist}_{\mathsf{hyp}}(\,\cdot\,,\,\cdot\,): hyperbolic distance
- 𝖽𝗂𝗌𝗍𝖾𝗎𝖼(,)\mathsf{dist}_{\mathsf{euc}}(\,\cdot\,,\,\cdot\,): Euclidean distance
- 𝖽𝗂𝗌𝗍𝖳𝖵(,)\mathsf{dist}_{\mathsf{TV}}(\,\cdot\,,\,\cdot\,): total variation distance
- 𝖽𝗂𝗌𝗍𝖪𝖱(,)\mathsf{dist}_{\mathsf{KR}}(\,\cdot\,,\,\cdot\,): Kantorovich-Rubinstein distance

In addition, simple point processes ω=k1δzk\omega=\sum_{k\geq 1}\delta_{z_{k}} with zkdz_{k}\in\mathbb{H}^{d} will be identified with locally finite point configurations {zk:k1}\{z_{k}\mathrel{\mathop{\mathchar 58\relax}}k\geq 1\}. Moreover, by abuse of notation, we write zωz\in\omega if zz is an atom of the counting measure ω\omega. The intensity measure of ω\omega is denoted by 𝔼[ω]\mathbb{E}[\omega].

In this paper we will denote by C>0C>0 a generic constant whose value might change from occurrence to occurrence. If several such constants are needed at the same time, we denote them by C0,C1,C_{0},C_{1},\ldots

Finally, since the parameters kk and s0s_{0} are fixed in the rest of the manuscript, we omit them from the notation.

3. Proof outline

To prove Theorem 2.1, we partition [0,1]d1[0,1]^{d-1} into uλu_{\lambda} congruent blocks S1,,SuλS_{1},\dots,S_{u_{\lambda}} and set Qm:=Sm×[eλ,)Q_{m}\mathrel{\mathop{\mathchar 58\relax}}=S_{m}\times[e^{-\lambda},\infty). We note that uλu_{\lambda} is not necessarily an integer. However, to keep notation simple we do not write rounding symbols. The key step in the proof of Theorem 2.1 is to relate the empirical measure ξλ/uλ\xi_{\lambda}/u_{\lambda} with the following separated point process:

ηλ:=muληλ(m),whereηλ(m):=x𝒫Qm𝟙{f(x,𝒫Qm)>s0}δf(x,𝒫Qm).\eta_{\lambda}\mathrel{\mathop{\mathchar 58\relax}}=\sum_{m\leq u_{\lambda}}\eta^{(m)}_{\lambda},\qquad\text{where}\qquad\eta^{(m)}_{\lambda}\mathrel{\mathop{\mathchar 58\relax}}=\sum_{x\in\mathcal{P}_{Q^{-}_{m}}}{\mathbbm{1}}\{f(x,\mathcal{P}_{Q_{m}})>s_{0}\}\delta_{f(x,\mathcal{P}_{Q_{m}})}.

Here, for a locally finite counting measure ω\omega on d\mathbb{H}^{d} we put f(x,ω):=|BRk(x,ω)|𝗁𝗒𝗉vλf(x,\omega)\mathrel{\mathop{\mathchar 58\relax}}=|B_{R_{k}(x,\omega)}|_{\mathsf{hyp}}-v_{\lambda} with Rk(x,ω):=inf{r0:ω(Br(x))k+1}R_{k}(x,\omega)\mathrel{\mathop{\mathchar 58\relax}}=\inf\{r\geq 0\mathrel{\mathop{\mathchar 58\relax}}\omega(B_{r}(x))\geq k+1\} and let QmQmQ^{-}_{m}\subseteq Q_{m} be an ‘internal region’ that we will define precisely in (3.1) below. The idea behind introducing these internal regions is that we want large kNN balls to occur independently for distinct regions. In other words, this step will remove the dependence of kNN radii exceeding the threshold rλ(u)r_{\lambda}(u), which is defined such that |Brλ(u)|𝗁𝗒𝗉=u+vλ|B_{r_{\lambda}(u)}|_{\mathsf{hyp}}=u+v_{\lambda}.

Having introduced the process ηλ\eta_{\lambda}, the proof of Theorem 2.1 can be split up into two steps regarding exponential equivalence (with respect to the total variation distance) in the sense of [8, Definition 4.2.10]:

  1. (1)

    show that the family of random measures ηλ/uλ\eta_{\lambda}/u_{\lambda} is exponentially equivalent to ζλ/uλ\zeta_{\lambda}/u_{\lambda}, where for each λ>0\lambda>0, ζλ\zeta_{\lambda} is a Poisson point process on E0E_{0} whose intensity measure has density uλτku_{\lambda}\tau_{k} with respect to the Lebesgue measure;

  2. (2)

    show that the family of random measures ηλ/uλ\eta_{\lambda}/u_{\lambda} is exponentially equivalent to ξλ/uλ\xi_{\lambda}/u_{\lambda}.

Proposition 3.1 (Exponential equivalence of ηλ\eta_{\lambda} and ζλ\zeta_{\lambda}).

The families of random measures ηλ/uλ\eta_{\lambda}/u_{\lambda} and ζλ/uλ\zeta_{\lambda}/u_{\lambda} are exponentially equivalent.

Proposition 3.2 (Exponential equivalence of ηλ\eta_{\lambda} and ξλ\xi_{\lambda}).

The families of random measures ηλ/uλ\eta_{\lambda}/u_{\lambda} and ξλ/uλ\xi_{\lambda}/u_{\lambda} are exponentially equivalent.

The main task is to prove Propositions 3.1 and 3.2, which are the hyperbolic analogs of [16, Proposition 4.3] and [16, Propositions 4.4–4.6]. Before doing so in Sections 4 and 5, we briefly explain how to deduce Theorem 2.1 from these two results.

Proof of Theorem 2.1.

First, we note that by the Poisson variant of Sanov’s theorem, the family of random measures (ζλ/uλ)λ>0(\zeta_{\lambda}/u_{\lambda})_{\lambda>0} satisfies an LDP with speed uλu_{\lambda} and rate function H(|τk)H(\,\cdot\,|\tau_{k}); see [8, Theorem 6.2.10]. Hence, taking into account Propositions 3.1 and 3.2, the result follows from the fundamental fact in large deviations theory that exponentially equivalent families of random elements satisfy the same LDP; see [8, Theorem 4.2.13]. ∎

We conclude the present overview section, with a precise definition of the internal regions QmQ^{-}_{m}. To that end, we let (wλ)λ>0(w_{\lambda})_{\lambda>0} be a diverging sequence with wλo(vλ)w_{\lambda}\in o(v_{\lambda}). Now, for each yeλy\geq e^{-\lambda}, we define

(3.1) Qm:={(x,y)Qm:xSm,y}\displaystyle Q^{-}_{m}\mathrel{\mathop{\mathchar 58\relax}}=\{(x,y)\in Q_{m}\mathrel{\mathop{\mathchar 58\relax}}x\in S^{-}_{m,y}\}

where Sm,y:={xSm:𝖽𝗂𝗌𝗍𝖾𝗎𝖼(x,Sm)xλ(y)}d1S^{-}_{m,y}\mathrel{\mathop{\mathchar 58\relax}}=\{x\in S_{m}\colon\mathsf{dist}_{\mathsf{euc}}(x,\partial S_{m})\geq x_{\lambda}(y)\}\subseteq{\mathbb{R}}^{d-1} with xλ(y):=yerλ(wλ)x_{\lambda}(y)\mathrel{\mathop{\mathchar 58\relax}}=ye^{r_{\lambda}(w_{\lambda})}, 𝖽𝗂𝗌𝗍𝖾𝗎𝖼\mathsf{dist}_{\mathsf{euc}} refers to the Euclidean distance in d1{\mathbb{R}}^{d-1} and Sm\partial S_{m} stands for the boundary of the set SmS_{m}. The desired properties of the internal regions QmQ^{-}_{m} are described in the following proposition.

Proposition 3.3 (Containment and volume of QmQ^{-}_{m}).

It holds that

  1. (1)

    Brλ(wλ)(z)QmB_{r_{\lambda}(w_{\lambda})}(z)\subseteq Q_{m} for every zQmz\in Q^{-}_{m}.

  2. (2)

    |QmQm|𝗁𝗒𝗉o(|Qm|𝗁𝗒𝗉)\big{|}Q_{m}\setminus Q^{-}_{m}\big{|}_{\mathsf{hyp}}\in o(|Q_{m}|_{\mathsf{hyp}}).

To prove Proposition 3.3, we rely on the following distance formula taken from [10, Proposition 2.1]. We also recall that we identify points zdz\in\mathbb{H}^{d} with pairs z=(x,y)z=(x,y) in the product space d1×(0,){\mathbb{R}}^{d-1}\times(0,\infty).

Lemma 3.4.

Let z1=(x1,y1)dz_{1}=(x_{1},y_{1})\in\mathbb{H}^{d}, z2=(x2,y2)dz_{2}=(x_{2},y_{2})\in\mathbb{H}^{d} and define κ:=𝖽𝗂𝗌𝗍𝖾𝗎𝖼(x1,x2)/y1\kappa\mathrel{\mathop{\mathchar 58\relax}}=\mathsf{dist}_{\mathsf{euc}}(x_{1},x_{2})/y_{1} and v:=y2/y1v\mathrel{\mathop{\mathchar 58\relax}}=y_{2}/y_{1}. Then,

𝖽𝗂𝗌𝗍𝗁𝗒𝗉(z1,z2):=Φ(v1(κ2+(v+1)2),\mathsf{dist}_{\mathsf{hyp}}(z_{1},z_{2})\mathrel{\mathop{\mathchar 58\relax}}=\Phi(v^{-1}(\kappa^{2}+(v+1)^{2}),

where Φ(t):=log(t)log(4)+2log(1+14/t).\Phi(t)\mathrel{\mathop{\mathchar 58\relax}}=\log(t)-\log(4)+2\log(1+\sqrt{1-4/t}). In particular, by minimizing over y2y_{2},

(3.2) infy2>0𝖽𝗂𝗌𝗍𝗁𝗒𝗉(z1,(x2,y2))log(κ).\displaystyle\inf_{y_{2}>0}\mathsf{dist}_{\mathsf{hyp}}(z_{1},(x_{2},y_{2}))\geq\log(\kappa).

holds for all sufficiently large κ>0\kappa>0.

Proof.

To carry out the minimization, we set the ddv(v1(κ2+(v+1)2))=0\tfrac{\operatorname{d\!}}{\operatorname{d\!}v}(v^{-1}(\kappa^{2}+(v+1)^{2}))=0. This gives 1=(κ2+1)/v21=(\kappa^{2}+1)/v^{2}. Hence, v1(κ2+(v+1)2)v^{-1}(\kappa^{2}+(v+1)^{2}) is asymptotically equivalent to 2κ2+12\sqrt{\kappa^{2}+1}, which in turn is equivalent to 2κ2\kappa for large κ\kappa. ∎

We now prove Proposition 3.3. For the proof, we note that

(3.3) log(vλ)d1+C1rλ(wλ)log(vλ)d1+C2,\frac{\log(v_{\lambda})}{d-1}+C_{1}\leq r_{\lambda}(w_{\lambda})\leq\frac{\log(v_{\lambda})}{d-1}+C_{2},

for suitable constants <C1<C2<-\infty<C_{1}<C_{2}<\infty; see Equation (4.1) in [21].

Proof of Proposition 3.3.

We prove separately parts (i) and (ii).

Part (i)

Let z1=(x1,y1)Qmz_{1}=(x_{1},y_{1})\in Q^{-}_{m} and x2Smx_{2}\in S_{m}^{\circ}. Then, by (3.2),

miny2>0𝖽𝗂𝗌𝗍𝗁𝗒𝗉(z1,(x2,y2))log(κ)log(xλ(y1)/y1)=rλ(wλ).\min_{y_{2}>0}\mathsf{dist}_{\mathsf{hyp}}(z_{1},(x_{2},y_{2}))\geq\log(\kappa)\geq\log(x_{\lambda}(y_{1})/y_{1})=r_{\lambda}(w_{\lambda}).

In other words, Brλ(wλ)(z1)QmB_{r_{\lambda}(w_{\lambda})}(z_{1})\subseteq Q_{m}.

Part (ii)

Note that QmQ_{m} is a cube of side length uλ1/(d1)u_{\lambda}^{-1/(d-1)} so that

|SmSm,y|𝗅𝖾𝖻2(d1)uλ(d2)/(d1)xλ(y).|S_{m}\setminus S^{-}_{m,y}|_{\mathsf{leb}}\leq 2(d-1)u_{\lambda}^{-(d-2)/(d-1)}x_{\lambda}(y).

Hence, for d3d\geq 3, by (2.1) and Fubini’s theorem, |QmQm|𝗁𝗒𝗉|Q_{m}\setminus Q^{-}_{m}|_{\mathsf{hyp}} can be bounded as

eλyd|SmSm,y|𝗅𝖾𝖻dy\displaystyle\int_{e^{-\lambda}}^{\infty}y^{-d}|S_{m}\setminus S^{-}_{m,y}|_{\mathsf{leb}}\operatorname{d\!}y 2(d1)uλ(d2)/(d1)eλxλ(y)yddy\displaystyle\leq 2(d-1)u_{\lambda}^{-(d-2)/(d-1)}\int_{e^{-\lambda}}^{\infty}x_{\lambda}(y)y^{-d}\operatorname{d\!}y
Cuλ(d2)/(d1)vλ1/(d1)|Wλ|𝗁𝗒𝗉(d2)/(d1).\displaystyle\leq Cu_{\lambda}^{-(d-2)/(d-1)}v_{\lambda}^{1/(d-1)}|W_{\lambda}|_{\mathsf{hyp}}^{(d-2)/(d-1)}.

Therefore,

|QmQm|𝗁𝗒𝗉d1|Qm|𝗁𝗒𝗉d1cuλ(d2)vλ|Wλ|𝗁𝗒𝗉d2uλ(d1)|Wλ|𝗁𝗒𝗉d1=Cevλ/(d1)vλk/(d1)0,\frac{\big{|}Q_{m}\setminus Q^{-}_{m}\big{|}_{\mathsf{hyp}}^{d-1}}{|Q_{m}|_{\mathsf{hyp}}^{d-1}}\leq\frac{cu_{\lambda}^{-(d-2)}v_{\lambda}|W_{\lambda}|_{\mathsf{hyp}}^{d-2}}{u_{\lambda}^{-(d-1)}|W_{\lambda}|_{\mathsf{hyp}}^{d-1}}=Ce^{-v_{\lambda}/(d-1)}v_{\lambda}^{k/(d-1)}\to 0,

as λ\lambda\uparrow\infty.

Next, consider d=2d=2. Here, the important observation is that Sm,y=S^{-}_{m,y}=\emptyset if xλ(y)uλ1x_{\lambda}(y)\geq u_{\lambda}^{-1}. That is, if yy0(λ):=uλ1erλ(wλ)y\geq y_{0}(\lambda)\mathrel{\mathop{\mathchar 58\relax}}=u_{\lambda}^{-1}e^{-r_{\lambda}(w_{\lambda})}. Now,

y0(λ)y2|Sm|𝗅𝖾𝖻dy=uλ1y0(λ)1=erλ(w)Cvλo(uλ1|Wλ|𝗁𝗒𝗉).\int_{y_{0}(\lambda)}^{\infty}y^{-2}|S_{m}|_{\mathsf{leb}}\operatorname{d\!}y={u_{\lambda}^{-1}y_{0}(\lambda)}^{-1}=e^{r_{\lambda}(w)}\leq Cv_{\lambda}\in o\big{(}u_{\lambda}^{-1}|W_{\lambda}|_{\mathsf{hyp}}\big{)}.

Moreover,

eλy0(λ)y2|SmSm,y|𝗅𝖾𝖻dy\displaystyle\int_{e^{-\lambda}}^{y_{0}(\lambda)}y^{-2}|S_{m}\setminus S^{-}_{m,y}|_{\mathsf{leb}}\operatorname{d\!}y 2(d1)eλy0(λ)xλ(y)y2dy\displaystyle\leq 2(d-1)\int_{e^{-\lambda}}^{y_{0}(\lambda)}x_{\lambda}(y)y^{-2}\operatorname{d\!}y
Cvλ(log(y0(λ)|Wλ|𝗁𝗒𝗉))\displaystyle\leq Cv_{\lambda}(\log(y_{0}(\lambda)|W_{\lambda}|_{\mathsf{hyp}}))
C1vλ(vλ+C2log(vλ)).\displaystyle\leq C_{1}v_{\lambda}(v_{\lambda}+C_{2}\log(v_{\lambda})).

Noting that vλ2o(uλ1|Wλ|𝗁𝗒𝗉)v_{\lambda}^{2}\in o\big{(}u_{\lambda}^{-1}|W_{\lambda}|_{\mathsf{hyp}}\big{)} concludes the proof. ∎

4. Proof of Proposition 3.1

In this section, we prove Proposition 3.1, that is, the exponential equivalence of the empirical measures ηλ/uλ\eta_{\lambda}/u_{\lambda} and a family of empirical measures of suitable Poisson point processes. We recall that ηλ\eta_{\lambda} is composed of its constituents ηλ(m)\eta^{(m)}_{\lambda} in the individual boxes. The first step of the proof is therefore to proceed as in [16, Proposition 4.1] and establish a Poisson approximation result individually for each ηλ(m)\eta^{(m)}_{\lambda}. We do this by showing the stronger assertion that for each m1m\geq 1 the Kantorovich-Rubinstein distance 𝖽𝗂𝗌𝗍𝖪𝖱((ηλ(m)),(ζλ(m)))\mathsf{dist}_{\mathsf{KR}}(\mathcal{L}(\eta^{(m)}_{\lambda}),\mathcal{L}(\zeta^{(m)}_{\lambda})) between the law (ηλ(m))\mathcal{L}(\eta^{(m)}_{\lambda}) of ηλ(m)\eta^{(m)}_{\lambda} and that of ζλ(m)\zeta^{(m)}_{\lambda} tends to zero, as λ\lambda\uparrow\infty. Here, we recall that for two point processes ω1,ω2\omega_{1},\omega_{2} on d\mathbb{H}^{d},

𝖽𝗂𝗌𝗍𝖪𝖱((ω1),(ω2)):=suph|𝔼h(ω1)𝔼h(ω2)|,\mathsf{dist}_{\mathsf{KR}}(\mathcal{L}(\omega_{1}),\mathcal{L}(\omega_{2}))\mathrel{\mathop{\mathchar 58\relax}}=\sup_{h}|\mathbb{E}h(\omega_{1})-\mathbb{E}h(\omega_{2})|,

where the supremum runs over all measurable Lipschitz-11 functions on the space of point processes on the space E0E_{0} with respect to the total variation distance 𝖽𝗂𝗌𝗍𝖳𝖵\mathsf{dist}_{\mathsf{TV}}. For two measures μ1,μ2\mu_{1},\mu_{2} on E0E_{0} the latter is given by

𝖽𝗂𝗌𝗍𝖳𝖵(μ1,μ2):=supAE0|μ1(A)μ2(A)|.\mathsf{dist}_{\mathsf{TV}}(\mu_{1},\mu_{2})\mathrel{\mathop{\mathchar 58\relax}}=\sup_{A\subseteq E_{0}}|\mu_{1}(A)-\mu_{2}(A)|.
Proposition 4.1 (Poisson approximation for separated processes).

Let m1m\geq 1. Then,

𝖽𝗂𝗌𝗍𝖪𝖱((ηλ(m)),(ζλ(m)))0,as λ,\mathsf{dist}_{\mathsf{KR}}\big{(}\mathcal{L}(\eta^{(m)}_{\lambda}),\mathcal{L}(\zeta^{(m)}_{\lambda})\big{)}\to 0,\quad\text{as $\lambda\uparrow\infty$},

where (ζλ(m))m1(\zeta^{(m)}_{\lambda})_{m\geq 1} is a sequence of independent and identically distributed Poisson point processes on E0E_{0}, each having intensity measure τk\tau_{k}.

To prepare the proof, some elements of which are similar to the main computations in [21], we recall that for ωd\omega\subseteq\mathbb{H}^{d} locally finite, we put f(x,ω):=|BRk(x,ω)|𝗁𝗒𝗉vλf(x,\omega)\mathrel{\mathop{\mathchar 58\relax}}=|B_{R_{k}(x,\omega)}|_{\mathsf{hyp}}-v_{\lambda}. We also define g(x,ω):=𝟙{f(x,ω)>s0}g(x,\omega)\mathrel{\mathop{\mathchar 58\relax}}={\mathbbm{1}}\{f(x,\omega)>s_{0}\} and 𝒮(x,ω)=BRk(x,ω)(x)\mathcal{S}(x,\omega)=B_{R_{k}(x,\omega)}(x). Then, ff and gg are localized to 𝒮\mathcal{S} in the sense of [4]. Namely, for every xωx\in\omega and all S𝒮(x,ω)S\supseteq\mathcal{S}(x,\omega), we have that g(x,ω)=g(x,ωS), and f(x,ω)=f(x,ωS) if g(x,ω)=1.g(x,\omega)=g(x,\omega\cap S),\text{ and }f(x,\omega)=f(x,\omega\cap S)\text{ if }g(x,\omega)=1. Moreover, 𝒮(x,ω)\mathcal{S}(x,\omega) is a so-called stopping set, in the sense that if BRk(x,ω)(x)SB_{R_{k}(x,\omega)}(x)\subseteq S, then BRk(x,ωS)(x)SB_{R_{k}(x,\omega\cap S)}(x)\subseteq S for every compact set SdS\subseteq\mathbb{H}^{d}. First, we set Sx:=Brλ(wλ)(x)S_{x}\mathrel{\mathop{\mathchar 58\relax}}=B_{r_{\lambda}(w_{\lambda})}(x). Henceforth, we put

sλ:=(𝒫(Brλ(s0)(x))k1) and sλ:=(𝒫(Brλ(wλ)(x))k1).s_{\lambda}\mathrel{\mathop{\mathchar 58\relax}}=\mathbb{P}\big{(}\mathcal{P}(B_{r_{\lambda}(s_{0})}(x))\leq k-1\big{)}\quad\text{ and }\quad s_{\lambda}^{\prime}\mathrel{\mathop{\mathchar 58\relax}}=\mathbb{P}\big{(}\mathcal{P}(B_{r_{\lambda}(w_{\lambda})}(x))\leq k-1\big{)}.
Proof of Proposition 4.1.

We shall check a set of sufficient conditions for convergence in Kantorovich-Rubinstein distance as provided in Theorem 6.4 in [4]. This requires the analysis of the total variance distance 𝖽𝗂𝗌𝗍𝖳𝖵(𝔼[ηλ(m)],τk)\mathsf{dist}_{\mathsf{TV}}(\mathbb{E}[\eta_{\lambda}^{(m)}],\tau_{k}) as well bounds on three error terms E1E_{1}, E2E_{2} and E3E_{3} which will be defined below.

First, it is claimed that

𝖽𝗂𝗌𝗍𝖳𝖵(𝔼[ηλ(m)],τk)0,as λ.\mathsf{dist}_{\mathsf{TV}}\big{(}\mathbb{E}[\eta_{\lambda}^{(m)}],\tau_{k}\big{)}\to 0,\ \ \text{as }\lambda\uparrow\infty.

Recall that f(x,𝒫Qm)>uf(x,\mathcal{P}_{Q_{m}})>u if and only if 𝒫(Brλ(u)(x))k\mathcal{P}\big{(}B_{r_{\lambda}(u)}(x)\big{)}\leq k. Hence, for u>s0u>s_{0}, it follows from the Mecke equation for Poisson point processes [20, Theorem 4.1], that

𝔼[ηλ(m)(u,)]\displaystyle\mathbb{E}\big{[}\eta_{\lambda}^{(m)}(u,\infty)\big{]} =𝔼[x𝒫Qm𝟙{f(x,𝒫Qm)>u}]=𝔼[x𝒫Qm𝟙{𝒫(Brλ(u)(x))k}]\displaystyle=\mathbb{E}\Big{[}\sum_{x\in\mathcal{P}_{Q^{-}_{m}}}{\mathbbm{1}}\big{\{}f(x,\mathcal{P}_{Q_{m}})>u\big{\}}\Big{]}=\mathbb{E}\Big{[}\sum_{x\in\mathcal{P}_{Q^{-}_{m}}}{\mathbbm{1}}\big{\{}\mathcal{P}\big{(}B_{r_{\lambda}(u)}(x)\big{)}\leq k\big{\}}\Big{]}
=Qm(𝒫(Brλ(u)(x))k1)V𝗁𝗒𝗉(dx)=|Qm|𝗁𝗒𝗉m=0k1e(u+vλ)(u+vλ)mm!.\displaystyle=\int_{Q^{-}_{m}}\mathbb{P}\big{(}\mathcal{P}\big{(}B_{r_{\lambda}(u)}(x)\big{)}\leq k-1\big{)}V_{\mathsf{hyp}}(\operatorname{d\!}x)=|Q^{-}_{m}|_{\mathsf{hyp}}\sum_{m=0}^{k-1}e^{-(u+v_{\lambda})}\frac{(u+v_{\lambda})^{m}}{m!}.

This implies that 𝔼[ηλ(m)]\mathbb{E}[\eta_{\lambda}^{(m)}] has the Lebesgue density

qk(u):=|Qm|𝗁𝗒𝗉e(u+vλ)(u+vλ)k1(k1)!,u>s0,q_{k}(u)\mathrel{\mathop{\mathchar 58\relax}}=|Q^{-}_{m}|_{\mathsf{hyp}}\,\frac{e^{-(u+v_{\lambda})}(u+v_{\lambda})^{k-1}}{(k-1)!},\ \ \ u>s_{0},

and thus,

𝖽𝗂𝗌𝗍𝖳𝖵(𝔼[ηλ(m)],τk)s0|qk(u)eu(k1)!|du\displaystyle\mathsf{dist}_{\mathsf{TV}}\big{(}\mathbb{E}[\eta_{\lambda}^{(m)}],\,\tau_{k}\big{)}\leq\int_{s_{0}}^{\infty}\Big{|}q_{k}(u)-\frac{e^{-u}}{(k-1)!}\Big{|}\operatorname{d\!}u
|1|Qm|𝗁𝗒𝗉evλvλk1|s0(1+uvλ)k1eudu+s0|(1+uvλ)k11|eudu.\displaystyle\leq\big{|}1-|Q^{-}_{m}|_{\mathsf{hyp}}e^{-v_{\lambda}}v_{\lambda}^{k-1}\big{|}\int_{s_{0}}^{\infty}\Big{(}1+\frac{u}{v_{\lambda}}\Big{)}^{k-1}e^{-u}\operatorname{d\!}u+\int_{s_{0}}^{\infty}\Big{|}\Big{(}1+\frac{u}{v_{\lambda}}\Big{)}^{k-1}-1\Big{|}e^{-u}\operatorname{d\!}u.

The second term above vanishes as λ\lambda\uparrow\infty because of the dominated convergence theorem. For the first term, we note that Proposition 3.3 gives

|Qm|𝗁𝗒𝗉evλvλk1=|Qm|𝗁𝗒𝗉|Qm|𝗁𝗒𝗉|Qm|𝗁𝗒𝗉evλvλk1=|Qm|𝗁𝗒𝗉|Qm|𝗁𝗒𝗉1, as λ.|Q^{-}_{m}|_{\mathsf{hyp}}e^{-v_{\lambda}}v_{\lambda}^{k-1}=\frac{|Q^{-}_{m}|_{\mathsf{hyp}}}{|Q_{m}|_{\mathsf{hyp}}}|Q_{m}|_{\mathsf{hyp}}e^{-v_{\lambda}}v_{\lambda}^{k-1}=\frac{|Q^{-}_{m}|_{\mathsf{hyp}}}{|Q_{m}|_{\mathsf{hyp}}}\to 1,\ \ \text{ as }\lambda\uparrow\infty.

It is now concluded that 𝖽𝗂𝗌𝗍𝖳𝖵(𝔼[ηλ(m)],τk)0\mathsf{dist}_{\mathsf{TV}}\big{(}\mathbb{E}[\eta_{\lambda}^{(m)}],\tau_{k}\big{)}\to 0, as λ\lambda\to\infty.

To ease notation, we henceforth write rλr_{\lambda} instead of rλ(s0)r_{\lambda}(s_{0}). Subsequently, we demonstrate that

E1:=Qm𝔼[g(x,𝒫Qm+δx) 1{𝒮(x,𝒫Qm+δx)Sx}]V𝗁𝗒𝗉(dx)0,λ.E_{1}\mathrel{\mathop{\mathchar 58\relax}}=\int_{Q^{-}_{m}}\mathbb{E}\Big{[}g(x,\mathcal{P}_{Q_{m}}+\delta_{x})\,{\mathbbm{1}}\big{\{}\mathcal{S}(x,\mathcal{P}_{Q_{m}}+\delta_{x})\not\subseteq S_{x}\big{\}}\Big{]}V_{\mathsf{hyp}}(\operatorname{d\!}x)\to 0,\ \ \ \lambda\uparrow\infty.

Notice that

𝟙{𝒮(x,𝒫Qm+δx)Sx}=𝟙{(𝒫Qm+δx)(Brλ(wλ)(x))k}=𝟙{𝒫(Brλ(wλ)(x))k1},\displaystyle{\mathbbm{1}}\big{\{}\mathcal{S}(x,\mathcal{P}_{Q_{m}}+\delta_{x})\not\subseteq S_{x}\big{\}}={\mathbbm{1}}\big{\{}(\mathcal{P}_{Q_{m}}+\delta_{x})\big{(}B_{r_{\lambda}(w_{\lambda})}(x)\big{)}\leq k\big{\}}={\mathbbm{1}}\big{\{}\mathcal{P}\big{(}B_{r_{\lambda}(w_{\lambda})}(x)\big{)}\leq k-1\big{\}},

where the second equality is due to the fact that Brλ(wλ)(x)QmB_{r_{\lambda}(w_{\lambda})}(x)\subseteq Q_{m} for all xQmx\in Q^{-}_{m}. Now,

E1\displaystyle E_{1} QmsλV𝗁𝗒𝗉(dx)|Qm|𝗁𝗒𝗉m=0k1e(vλ+wλ)(vλ+wλ)mm!kewλ(1+wλvλ)k10,λ,\displaystyle\leq\int_{Q^{-}_{m}}s_{\lambda}^{\prime}\,V_{\mathsf{hyp}}(\operatorname{d\!}x)\leq|Q_{m}|_{\mathsf{hyp}}\sum_{m=0}^{k-1}e^{-(v_{\lambda}+w_{\lambda})}\frac{(v_{\lambda}+w_{\lambda})^{m}}{m!}\leq ke^{-w_{\lambda}}\Big{(}1+\frac{w_{\lambda}}{v_{\lambda}}\Big{)}^{k-1}\to 0,\ \ \ \lambda\uparrow\infty,

where we have used that |Qm|𝗁𝗒𝗉=evλvλ(k1)|Q_{m}|_{\mathsf{hyp}}=e^{v_{\lambda}}v_{\lambda}^{-(k-1)}.

Next, we turn our attention to the second error term

E2:=QmQm𝟙{SxSz}𝔼[g(x,𝒫Qm+δx)]𝔼[g(z,𝒫Qm+δz)]V𝗁𝗒𝗉(dz)V𝗁𝗒𝗉(dx).E_{2}\mathrel{\mathop{\mathchar 58\relax}}=\int_{Q^{-}_{m}}\int_{Q^{-}_{m}}{\mathbbm{1}}\{S_{x}\cap S_{z}\neq\emptyset\}\mathbb{E}\big{[}g(x,\mathcal{P}_{Q_{m}}+\delta_{x})\big{]}\mathbb{E}\big{[}g(z,\mathcal{P}_{Q_{m}}+\delta_{z})\big{]}V_{\mathsf{hyp}}(\operatorname{d\!}z)V_{\mathsf{hyp}}(\operatorname{d\!}x).

We here apply the following inequalities:

𝟙{SxSz}\displaystyle{\mathbbm{1}}\{S_{x}\cap S_{z}\neq\emptyset\} 𝟙{𝖽𝗂𝗌𝗍𝗁𝗒𝗉(x,z)2rλ(wλ)},\displaystyle\leq{\mathbbm{1}}\big{\{}\mathsf{dist}_{\mathsf{hyp}}(x,z)\leq 2r_{\lambda}(w_{\lambda})\big{\}},
𝔼[g(x,𝒫Qm+δx)]\displaystyle\mathbb{E}\big{[}g(x,\mathcal{P}_{Q_{m}}+\delta_{x})\big{]} =sλ=i=0k1e(s0+vλ)(s0+vλ)ii!Cvλk1evλ.\displaystyle=s_{\lambda}=\sum_{i=0}^{k-1}\frac{e^{-(s_{0}+v_{\lambda})}(s_{0}+v_{\lambda})^{i}}{i!}\leq Cv_{\lambda}^{k-1}e^{-v_{\lambda}}.

Hence,

(4.1) E2\displaystyle E_{2} C(vλk1evλ)2Qmd𝟙{𝖽𝗂𝗌𝗍𝗁𝗒𝗉(x,z)2rλ(wλ)}V𝗁𝗒𝗉(dz)V𝗁𝗒𝗉(dx)\displaystyle\leq C(v_{\lambda}^{k-1}e^{-v_{\lambda}})^{2}\int_{Q^{-}_{m}}\int_{\mathbb{H}^{d}}{\mathbbm{1}}\big{\{}\mathsf{dist}_{\mathsf{hyp}}(x,z)\leq 2r_{\lambda}(w_{\lambda})\big{\}}V_{\mathsf{hyp}}(\operatorname{d\!}z)V_{\mathsf{hyp}}(\operatorname{d\!}x)
=C(vλk1evλ)2|Qm|𝗁𝗒𝗉|B2rλ(wλ)|𝗁𝗒𝗉\displaystyle=C(v_{\lambda}^{k-1}e^{-v_{\lambda}})^{2}|Q^{-}_{m}|_{\mathsf{hyp}}|B_{2r_{\lambda}(w_{\lambda})}|_{\mathsf{hyp}}
C(vλk1evλ)2evλvλ(k1)e2rλ(wλ)(d1)\displaystyle\leq C(v_{\lambda}^{k-1}e^{-v_{\lambda}})^{2}e^{v_{\lambda}}v_{\lambda}^{-(k-1)}\,e^{2r_{\lambda}(w_{\lambda})(d-1)}
Cvλk1vλ2evλ0,as λ.\displaystyle\leq Cv_{\lambda}^{k-1}v_{\lambda}^{2}e^{-v_{\lambda}}\to 0,\ \ \text{as }\lambda\uparrow\infty.

For the inequality at the last line, we have applied (3.3).

For the bound of E3E_{3} we distinguish between the cases k=1k=1 and k2k\geq 2. If k=1k=1 we have that

𝔼[g(x,𝒫Qm+δx+δz)g(z,𝒫Qm+δx+δz)]sλ3/2\mathbb{E}\big{[}g(x,\mathcal{P}_{Q_{m}}+\delta_{x}+\delta_{z})g(z,\mathcal{P}_{Q_{m}}+\delta_{x}+\delta_{z})\big{]}\leq s_{\lambda}^{3/2}

and, hence,

E3\displaystyle E_{3} :=QmQm𝟙{SxSz}𝔼[g(x,𝒫Qm+δx+δz)g(z,𝒫Qm+δx+δz)]V𝗁𝗒𝗉(dz)V𝗁𝗒𝗉(dx)\displaystyle\mathrel{\mathop{\mathchar 58\relax}}=\int_{Q^{-}_{m}}\int_{Q^{-}_{m}}{\mathbbm{1}}\{S_{x}\cap S_{z}\neq\emptyset\}\mathbb{E}\big{[}g(x,\mathcal{P}_{Q_{m}}+\delta_{x}+\delta_{z})g(z,\mathcal{P}_{Q_{m}}+\delta_{x}+\delta_{z})\big{]}V_{\mathsf{hyp}}(\operatorname{d\!}z)V_{\mathsf{hyp}}(\operatorname{d\!}x)
sλ3/2Qmd𝟙{𝖽𝗂𝗌𝗍𝗁𝗒𝗉(x,z)2rλ(wλ)}V𝗁𝗒𝗉(dz)V𝗁𝗒𝗉(dx)\displaystyle\leq s_{\lambda}^{3/2}\int_{Q^{-}_{m}}\int_{\mathbb{H}^{d}}{\mathbbm{1}}\big{\{}\mathsf{dist}_{\mathsf{hyp}}(x,z)\leq 2r_{\lambda}(w_{\lambda})\big{\}}V_{\mathsf{hyp}}(\operatorname{d\!}z)V_{\mathsf{hyp}}(\operatorname{d\!}x)
C(evλ)3/2eλ(d1)uλe2rλ(wλ)(d1)Cvλ2evλ/20,as λ,\displaystyle\leq C(e^{-v_{\lambda}})^{3/2}\frac{e^{\lambda(d-1)}}{u_{\lambda}}\,e^{2r_{\lambda}(w_{\lambda})(d-1)}\leq Cv_{\lambda}^{2}e^{-v_{\lambda}/2}\to 0,\ \ \text{as }\lambda\uparrow\infty,

where in the last step we used the volume estimate for QmQ_{m}^{-} and the fact that |Br(x)|𝗁𝗒𝗉Ce(d1)r|B_{r}(x)|_{\mathsf{hyp}}\leq Ce^{(d-1)r} for any r>0r>0 and xdx\in\mathbb{H}^{d}; see the discussion after (2.2).

For k2k\geq 2, we split E3E_{3} into two terms:

E3\displaystyle E_{3} =QmQm𝟙{SxSz}𝔼[g(x,𝒫Qm+δx+δz)g(z,𝒫Qm+δx+δz)]V𝗁𝗒𝗉(dz)V𝗁𝗒𝗉(dx)\displaystyle=\int_{Q^{-}_{m}}\int_{Q^{-}_{m}}{\mathbbm{1}}\{S_{x}\cap S_{z}\neq\emptyset\}\mathbb{E}\big{[}g(x,\mathcal{P}_{Q_{m}}+\delta_{x}+\delta_{z})g(z,\mathcal{P}_{Q_{m}}+\delta_{x}+\delta_{z})\big{]}V_{\mathsf{hyp}}(\operatorname{d\!}z)V_{\mathsf{hyp}}(\operatorname{d\!}x)
QmQm𝟙{𝖽𝗂𝗌𝗍𝗁𝗒𝗉(x,z)arλ(wλ)}\displaystyle\leq\int_{Q^{-}_{m}}\int_{Q^{-}_{m}}{\mathbbm{1}}\big{\{}\mathsf{dist}_{\mathsf{hyp}}(x,z)\leq ar_{\lambda}(w_{\lambda})\big{\}}
×((𝒫Qm+δz)(Brλ(x))k1,(𝒫Qm+δx)(Brλ(z))k1)V𝗁𝗒𝗉(dz)V𝗁𝗒𝗉(dx)\displaystyle\qquad\times\mathbb{P}\Big{(}(\mathcal{P}_{Q_{m}}+\delta_{z})\big{(}B_{r_{\lambda}}(x)\big{)}\leq k-1,\,(\mathcal{P}_{Q_{m}}+\delta_{x})\big{(}B_{r_{\lambda}}(z)\big{)}\leq k-1\Big{)}V_{\mathsf{hyp}}(\operatorname{d\!}z)V_{\mathsf{hyp}}(\operatorname{d\!}x)
+QmQm𝟙{arλ(wλ)<𝖽𝗂𝗌𝗍𝗁𝗒𝗉(x,z)2rλ(wλ)}\displaystyle+\int_{Q^{-}_{m}}\int_{Q^{-}_{m}}{\mathbbm{1}}\big{\{}ar_{\lambda}(w_{\lambda})<\mathsf{dist}_{\mathsf{hyp}}(x,z)\leq 2r_{\lambda}(w_{\lambda})\big{\}}
×((𝒫Qm+δz)(Brλ(x))k1,(𝒫Qm+δx)(Brλ(z))k1)V𝗁𝗒𝗉(dz)V𝗁𝗒𝗉(dx)\displaystyle\qquad\times\mathbb{P}\Big{(}(\mathcal{P}_{Q_{m}}+\delta_{z})\big{(}B_{r_{\lambda}}(x)\big{)}\leq k-1,\,(\mathcal{P}_{Q_{m}}+\delta_{x})\big{(}B_{r_{\lambda}}(z)\big{)}\leq k-1\Big{)}V_{\mathsf{hyp}}(\operatorname{d\!}z)V_{\mathsf{hyp}}(\operatorname{d\!}x)
=:E3,1+E3,2,\displaystyle=\mathrel{\mathop{\mathchar 58\relax}}E_{3,1}+E_{3,2},

where 0<a<10<a<1 is so small that 2rλa(k1)rλ(wλ)>02r_{\lambda}-a(k-1)r_{\lambda}(w_{\lambda})>0. The term E3,1E_{3,1} is bounded by

E3,1\displaystyle E_{3,1} QmQm𝟙{𝖽𝗂𝗌𝗍𝗁𝗒𝗉(x,z)arλ(wλ)}(𝒫(Brλ(x))k2)\displaystyle\leq\int_{Q^{-}_{m}}\int_{Q^{-}_{m}}{\mathbbm{1}}\big{\{}\mathsf{dist}_{\mathsf{hyp}}(x,z)\leq ar_{\lambda}(w_{\lambda})\big{\}}\mathbb{P}\big{(}\mathcal{P}\big{(}B_{r_{\lambda}}(x)\big{)}\leq k-2\big{)}
×(𝒫(Brλ(z)Brλ(x))k2)V𝗁𝗒𝗉(dz)V𝗁𝗒𝗉(dx).\displaystyle\qquad\qquad\qquad\times\mathbb{P}\big{(}\mathcal{P}\big{(}B_{r_{\lambda}}(z)\setminus B_{r_{\lambda}}(x)\big{)}\leq k-2\big{)}V_{\mathsf{hyp}}(\operatorname{d\!}z)V_{\mathsf{hyp}}(\operatorname{d\!}x).

Let pdp\in\mathbb{H}^{d} be a fixed reference point (sometimes called the origin) and let expp:Tpd\exp_{p}\mathrel{\mathop{\mathchar 58\relax}}T_{p}\to\mathbb{H}^{d} denote the exponential map at pp in which TpdT_{p}\cong{\mathbb{R}}^{d} is the tangent space at pp. Then, by the above the bound,

(𝒫(Brλ(x))k2)Cevλvλk2.\mathbb{P}\big{(}\mathcal{P}\big{(}B_{r_{\lambda}}(x)\big{)}\leq k-2\big{)}\leq Ce^{-v_{\lambda}}v_{\lambda}^{k-2}.

It follows from the polar integration formula in hyperbolic geometry [5, pp. 123-125] that

E3,1\displaystyle E_{3,1} Cevλvλk2|Qm|𝗁𝗒𝗉i=0k2d𝟙{𝖽𝗂𝗌𝗍𝗁𝗒𝗉(z,p)arλ(wλ)}\displaystyle\leq Ce^{-v_{\lambda}}v_{\lambda}^{k-2}|Q_{m}|_{\mathsf{hyp}}\sum_{i=0}^{k-2}\int_{\mathbb{H}^{d}}{\mathbbm{1}}\big{\{}\mathsf{dist}_{\mathsf{hyp}}(z,p)\leq ar_{\lambda}(w_{\lambda})\big{\}}
×e|Brλ(z)Brλ(p)|𝗁𝗒𝗉(|Brλ(z)Brλ(p)|𝗁𝗒𝗉)ii!V𝗁𝗒𝗉(dz)\displaystyle\qquad\qquad\qquad\qquad\qquad\times e^{-|B_{r_{\lambda}}(z)\setminus B_{r_{\lambda}}(p)|_{\mathsf{hyp}}}\frac{\big{(}|B_{r_{\lambda}}(z)\setminus B_{r_{\lambda}}(p)|_{\mathsf{hyp}}\big{)}^{i}}{i!}V_{\mathsf{hyp}}(\operatorname{d\!}z)
Cvλi=0k20arλ(wλ)sinhd1(s)e|Brλ(z)Brλ(p)|𝗁𝗒𝗉(|Brλ(z)Brλ(p)|𝗁𝗒𝗉)ii!ds,\displaystyle\leq\frac{C}{v_{\lambda}}\sum_{i=0}^{k-2}\int_{0}^{ar_{\lambda}(w_{\lambda})}\sinh^{d-1}(s)e^{-|B_{r_{\lambda}}(z)\setminus B_{r_{\lambda}}(p)|_{\mathsf{hyp}}}\frac{\big{(}|B_{r_{\lambda}}(z)\setminus B_{r_{\lambda}}(p)|_{\mathsf{hyp}}\big{)}^{i}}{i!}\operatorname{d\!}s,

where z=expp(sv0)z=\exp_{p}(sv_{0}) for some v0𝕊pd1v_{0}\in\mathbb{S}_{p}^{d-1}, the (d1)(d-1)-dimensional unit sphere in TpT_{p}. It follows from Lemma 5 in [21] that

E3,1\displaystyle E_{3,1} Cvλi=0k20arλ(wλ)(es2)d1eα1se(d1)(rλs/2)(α2se(d1)rλ)ii!ds\displaystyle\leq\frac{C}{v_{\lambda}}\sum_{i=0}^{k-2}\int_{0}^{ar_{\lambda}(w_{\lambda})}\Big{(}\frac{e^{s}}{2}\Big{)}^{d-1}e^{-\alpha_{1}se^{(d-1)(r_{\lambda}-s/2)}}\frac{\big{(}\alpha_{2}se^{(d-1)r_{\lambda}}\big{)}^{i}}{i!}\operatorname{d\!}s
Cvλi=0k2e(d1)irλi!0es(α1e(d1)(2rλarλ(wλ))/2(d1))sids\displaystyle\leq\frac{C}{v_{\lambda}}\sum_{i=0}^{k-2}\frac{e^{(d-1)ir_{\lambda}}}{i!}\,\int_{0}^{\infty}e^{-s(\alpha_{1}e^{(d-1)(2r_{\lambda}-ar_{\lambda}(w_{\lambda}))/2}-(d-1))}\,s^{i}\operatorname{d\!}s
Cvλi=0k2e(d1)2rλa(i+1)rλ(wλ)2\displaystyle\leq\frac{C}{v_{\lambda}}\sum_{i=0}^{k-2}e^{-(d-1)\frac{2r_{\lambda}-a(i+1)r_{\lambda}(w_{\lambda})}{2}}
Cvλe(d1)2rλa(k1)rλ(wλ)2Cvλ0,as λ.\displaystyle\leq\frac{C}{v_{\lambda}}e^{-(d-1)\frac{2r_{\lambda}-a(k-1)r_{\lambda}(w_{\lambda})}{2}}\leq\frac{C}{v_{\lambda}}\to 0,\ \ \text{as }\lambda\uparrow\infty.

Here, the last inequality follows from the constraint 2rλa(k1)rλ(wλ)>02r_{\lambda}-a(k-1)r_{\lambda}(w_{\lambda})>0.

For E3,2E_{3,2}, one can see that

E3,2\displaystyle E_{3,2} QmQm𝟙{arλ(wλ)<𝖽𝗂𝗌𝗍𝗁𝗒𝗉(x,z)2rλ(wλ)}sλ\displaystyle\leq\int_{Q^{-}_{m}}\int_{Q^{-}_{m}}{\mathbbm{1}}\big{\{}ar_{\lambda}(w_{\lambda})<\mathsf{dist}_{\mathsf{hyp}}(x,z)\leq 2r_{\lambda}(w_{\lambda})\big{\}}s_{\lambda}
×(𝒫(Brλ(z)Brλ(x))k1)V𝗁𝗒𝗉(dz)V𝗁𝗒𝗉(dx)\displaystyle\qquad\times\mathbb{P}\big{(}\mathcal{P}\big{(}B_{r_{\lambda}}(z)\setminus B_{r_{\lambda}}(x)\big{)}\leq k-1\big{)}V_{\mathsf{hyp}}(\operatorname{d\!}z)V_{\mathsf{hyp}}(\operatorname{d\!}x)
Csλ|Qm|𝗁𝗒𝗉d𝟙{arλ(wλ)<𝖽𝗂𝗌𝗍𝗁𝗒𝗉(z,p)2rλ(wλ)}\displaystyle\leq Cs_{\lambda}|Q_{m}|_{\mathsf{hyp}}\int_{\mathbb{H}^{d}}{\mathbbm{1}}\big{\{}ar_{\lambda}(w_{\lambda})<\mathsf{dist}_{\mathsf{hyp}}(z,p)\leq 2r_{\lambda}(w_{\lambda})\big{\}}
×i=0k1e|Brλ(z)Brλ(p)|𝗁𝗒𝗉|Brλ(z)Brλ(p)|𝗁𝗒𝗉ii!V𝗁𝗒𝗉(dz)\displaystyle\qquad\times\sum_{i=0}^{k-1}e^{-|B_{r_{\lambda}}(z)\setminus B_{r_{\lambda}}(p)|_{\mathsf{hyp}}}\frac{|B_{r_{\lambda}}(z)\setminus B_{r_{\lambda}}(p)|_{\mathsf{hyp}}^{i}}{i!}V_{\mathsf{hyp}}(\operatorname{d\!}z)
Ci=0k11i!arλ(wλ)2rλ(wλ)sinhd1(s)e|Brλ(z)Brλ(p)|𝗁𝗒𝗉(|Brλ(z)Brλ(p)|𝗁𝗒𝗉)ids,\displaystyle\leq C\sum_{i=0}^{k-1}\frac{1}{i!}\,\int_{ar_{\lambda}(w_{\lambda})}^{2r_{\lambda}(w_{\lambda})}\sinh^{d-1}(s)\,e^{-|B_{r_{\lambda}}(z)\setminus B_{r_{\lambda}}(p)|_{\mathsf{hyp}}}\big{(}|B_{r_{\lambda}}(z)\setminus B_{r_{\lambda}}(p)|_{\mathsf{hyp}}\big{)}^{i}\operatorname{d\!}s,

where again z=expp(sv0)z=\exp_{p}(sv_{0}) for some v0𝕊pd1v_{0}\in\mathbb{S}_{p}^{d-1}. Note that for every z=expp(sv0)dz=\exp_{p}(sv_{0})\in\mathbb{H}^{d} with arλ(wλ)s2rλ(wλ)ar_{\lambda}(w_{\lambda})\leq s\leq 2r_{\lambda}(w_{\lambda}),

|Brλ(z)Brλ(p)|𝗁𝗒𝗉|Bs/2(p)|𝗁𝗒𝗉,|B_{r_{\lambda}}(z)\setminus B_{r_{\lambda}}(p)|_{\mathsf{hyp}}\geq|B_{s/2}(p)|_{\mathsf{hyp}},

and further, by Lemma 4 in [21], |Bs/2(p)|𝗁𝗒𝗉C1es(d1)/2|B_{s/2}(p)|_{\mathsf{hyp}}\geq C_{1}e^{s(d-1)/2}. Using this bound,

E3,2\displaystyle E_{3,2} Ci=0k11i!arλ(wλ)2rλ(wλ)es(d1)eC1es(d1)/2(C1es(d1)/2)ids\displaystyle\leq C\sum_{i=0}^{k-1}\frac{1}{i!}\,\int_{ar_{\lambda}(w_{\lambda})}^{2r_{\lambda}(w_{\lambda})}e^{s(d-1)}e^{-C_{1}e^{s(d-1)/2}}\big{(}C_{1}e^{s(d-1)/2}\big{)}^{i}\operatorname{d\!}s
Ce(d1)arλ(wλ)/2i=0k1C1ii!arλ(wλ)2rλ(wλ)eC1es(d1)/2(es(d1)/2)i+3ds.\displaystyle\leq Ce^{-(d-1)ar_{\lambda}(w_{\lambda})/2}\sum_{i=0}^{k-1}\frac{C_{1}^{i}}{i!}\,\int_{ar_{\lambda}(w_{\lambda})}^{2r_{\lambda}(w_{\lambda})}e^{-C_{1}e^{s(d-1)/2}}(e^{s(d-1)/2})^{i+3}\operatorname{d\!}s.

Applying the substitution u=es(d1)/2u=e^{s(d-1)/2}, we find that the integral is bounded in λ\lambda and, hence, that E3,20E_{3,2}\to 0, as λ\lambda\to\infty.

We now put all bounds together and apply Theorem 6.4 in [4] to conclude that, for each m1m\geq 1,

𝖽𝗂𝗌𝗍𝖪𝖱((ηλ(m)),(ζλ(m)))𝖽𝗂𝗌𝗍𝖳𝖵(𝔼[ηλ(m)],τk)+2(E1+E2+E3)0,,as λ.\mathsf{dist}_{\mathsf{KR}}\big{(}\mathcal{L}(\eta^{(m)}_{\lambda}),\mathcal{L}(\zeta^{(m)}_{\lambda})\big{)}\leq\mathsf{dist}_{\mathsf{TV}}\big{(}\mathbb{E}[\eta_{\lambda}^{(m)}],\tau_{k}\big{)}+2(E_{1}+E_{2}+E_{3})\to 0,,\ \ \text{as }\lambda\uparrow\infty.

This completes the proof. ∎

By the maximal coupling lemma [19, Lemma 4.32], Proposition 4.1 gives for each m1m\geq 1 couplings η^λ(m)\hat{\eta}^{(m)}_{\lambda} and ζ^λ(m)\hat{\zeta}^{(m)}_{\lambda} satisfying

(4.2) (η^λ(m)ζ^λ(m))=𝖽𝗂𝗌𝗍𝖳𝖵((ηλ(m)),(ζλ(m)))0,as λ.\displaystyle\mathbb{P}\big{(}\hat{\eta}^{(m)}_{\lambda}\neq\hat{\zeta}^{(m)}_{\lambda}\big{)}=\mathsf{dist}_{\mathsf{TV}}\big{(}\mathcal{L}(\eta^{(m)}_{\lambda}),\mathcal{L}(\zeta^{(m)}_{\lambda})\big{)}\to 0,\ \ \text{as }\lambda\uparrow\infty.

Hence, we can define a coupling between ηλ\eta_{\lambda} and ζλ\zeta_{\lambda} by setting

η^λ:=muλη^λ(m), and ζ^λ:=muλζ^λ(m),\hat{\eta}_{\lambda}\mathrel{\mathop{\mathchar 58\relax}}=\sum_{m\leq u_{\lambda}}\hat{\eta}^{(m)}_{\lambda},\quad\text{ and }\quad\hat{\zeta}_{\lambda}\mathrel{\mathop{\mathchar 58\relax}}=\sum_{m\leq u_{\lambda}}\hat{\zeta}^{(m)}_{\lambda},

where we assume that the pairs (η^λ(m),ζ^λ(m))(\hat{\eta}^{(m)}_{\lambda},\hat{\zeta}^{(m)}_{\lambda}) are independent and identically distributed for different mm. Having constructed the couplings, we now proceed with the proof of the exponential equivalence.

Proof of Proposition 3.1.

First, we note that for every a>0a>0

1uλlog(𝖽𝗂𝗌𝗍𝖳𝖵(η^λ,ζ^λ)δuλ)\displaystyle\frac{1}{u_{\lambda}}\log\mathbb{P}\big{(}\mathsf{dist}_{\mathsf{TV}}\big{(}\hat{\eta}_{\lambda},\hat{\zeta}_{\lambda}\big{)}\geq\delta u_{\lambda}\big{)} 1uλlog(muλ𝖽𝗂𝗌𝗍𝖳𝖵(η^λ(m),ζ^λ(m))δuλ)\displaystyle\leq\frac{1}{u_{\lambda}}\log\mathbb{P}\Big{(}\sum_{m\leq u_{\lambda}}\mathsf{dist}_{\mathsf{TV}}\big{(}\hat{\eta}^{(m)}_{\lambda},\hat{\zeta}^{(m)}_{\lambda}\big{)}\geq\delta u_{\lambda}\Big{)}
1uλlog(eaδuλ𝔼[eamuλ𝖽𝗂𝗌𝗍𝖳𝖵(η^λ(m),ζ^λ(m))])\displaystyle\leq{1\over u_{\lambda}}\log\Big{(}e^{-a\delta u_{\lambda}}\mathbb{E}\Big{[}e^{a\sum_{m\leq u_{\lambda}}\mathsf{dist}_{\mathsf{TV}}\big{(}\hat{\eta}^{(m)}_{\lambda},\hat{\zeta}^{(m)}_{\lambda}\big{)}}\Big{]}\Big{)}
=aδ+1uλmuλlog𝔼[ea𝖽𝗂𝗌𝗍𝖳𝖵(η^λ(m),ζ^λ(m))]\displaystyle=-a\delta+{1\over u_{\lambda}}\sum_{m\leq u_{\lambda}}\log\mathbb{E}[e^{a\,\mathsf{dist}_{\mathsf{TV}}\big{(}\hat{\eta}^{(m)}_{\lambda},\hat{\zeta}^{(m)}_{\lambda}\big{)}}]
aδ+log𝔼[ea𝖽𝗂𝗌𝗍𝖳𝖵(η^λ(1),ζ^λ(1))],\displaystyle\leq-a\delta+\log\mathbb{E}[e^{a\,\mathsf{dist}_{\mathsf{TV}}(\hat{\eta}^{(1)}_{\lambda},\hat{\zeta}^{(1)}_{\lambda})}],

where we used the exponential Markov inequality and the i.i.d. property of the pairs (η^λ(m),ζ^λ(m))(\hat{\eta}^{(m)}_{\lambda},\hat{\zeta}^{(m)}_{\lambda}) discussed above. Moreover, by (4.2), we have that 𝖽𝗂𝗌𝗍𝖳𝖵(η^λ(1),ζ^λ(1))0\mathsf{dist}_{\mathsf{TV}}(\hat{\eta}^{(1)}_{\lambda},\hat{\zeta}^{(1)}_{\lambda})\to 0, as λ\lambda\uparrow\infty in probability. Hence, as in [16, Proposition 4.3], it suffices to verify the uniform integrability condition

(4.3) lim supλ𝔼[ea𝒬(Q1)]<\displaystyle\limsup_{\lambda\uparrow\infty}\mathbb{E}\Big{[}e^{a\mathcal{Q}(Q^{-}_{1})}\Big{]}<\infty

for every a>0a>0, where we let

𝒬:=Xi𝒫𝟙{𝒫(Brλ(s0)(Xi))k}δXi\mathcal{Q}\mathrel{\mathop{\mathchar 58\relax}}=\sum_{X_{i}\in\mathcal{P}}{\mathbbm{1}}\{\mathcal{P}(B_{r_{\lambda}(s_{0})}(X_{i}))\leq k\}\delta_{X_{i}}

denote the point process of exceedances. The key observation that we will use tacitly throughout the rest of the proof is that to determine the restriction of 𝒬\mathcal{Q} to a set AdA\subseteq\mathbb{H}^{d}, it suffices to know 𝒫\mathcal{P} in an rλ(s0)r_{\lambda}(s_{0})-neighborhood of AA. Therefore, we partition Q1Q_{1} into vertical layers Q1,:=S1×TQ_{1,\ell}\mathrel{\mathop{\mathchar 58\relax}}=S_{1}\times T_{\ell}, where T:=[eλ+rλ(s0),eλ+(+1)rλ(s0))T_{\ell}\mathrel{\mathop{\mathchar 58\relax}}=[e^{-\lambda+\ell r_{\lambda}(s_{0})},e^{-\lambda+(\ell+1)r_{\lambda}(s_{0})}). In order to avoid dealing with an infinite number of such layers, we define a value 01\ell_{0}\geq 1 such that S1=[0,eλ+(0+2)rλ(s0)]d1S_{1}=[0,e^{-\lambda+(\ell_{0}+2)r_{\lambda}(s_{0})}]^{d-1}. As in previous arguments, here, 0\ell_{0} is assumed to be an integer; if this is not the case, we replace rλ(s0)r_{\lambda}(s_{0}) by rr^{\prime} which is a small perturbation of rλ(s0)r_{\lambda}(s_{0}). We then set Q1,:=S1×[eλ+(0+1)rλ(s0),)Q_{1,\infty}\mathrel{\mathop{\mathchar 58\relax}}=S_{1}\times[e^{-\lambda+(\ell_{0}+1)r_{\lambda}(s_{0})},\infty). In particular, we deduce from Lemma 3.4 that 𝒬Q1,\mathcal{Q}_{Q_{1,\ell}} is independent of 𝒬Q1,\mathcal{Q}_{Q_{1,\ell^{\prime}}} whenever \ell and \ell^{\prime} are such that ||3|\ell^{\prime}-\ell|\geq 3. Hence, it suffices to prove that

(4.4) lim supλ𝔼[ea𝒬(Q1,)]< and lim supλ00,3𝔼[ea𝒬(Q1,)]<.\displaystyle\limsup_{\lambda\uparrow\infty}\mathbb{E}\Big{[}e^{a\mathcal{Q}(Q_{1,\infty})}\Big{]}<\infty\quad\text{ and }\quad\limsup_{\lambda\uparrow\infty}\prod_{0\leq\ell\leq\ell_{0},\ell\in 3\mathbb{Z}}\mathbb{E}\Big{[}e^{a\mathcal{Q}(Q_{1,\ell})}\Big{]}<\infty.

The arguments for layers with 13\ell-1\in 3\mathbb{Z} or 23\ell-2\in 3\mathbb{Z} are similar.

We start with Q1,Q_{1,\infty}, where to ease notation, we set J:=Q1,J\mathrel{\mathop{\mathchar 58\relax}}=Q_{1,\infty}. Note that by the kNN property, we have that if XiX_{i} is not among the (k1)(k-1) nearest neighbors of XjX_{j}, then

|B(Xi,Xj)|𝗁𝗒𝗉:=|Brλ(s0)/2(Xi)Brλ(s0)(Xj)|𝗁𝗒𝗉12|Brλ(s0)/2(Xi)|𝗁𝗒𝗉.\big{|}B(X_{i},X_{j})\big{|}_{\mathsf{hyp}}\mathrel{\mathop{\mathchar 58\relax}}=\big{|}B_{r_{\lambda}(s_{0})/2}(X_{i})\setminus B_{r_{\lambda}(s_{0})}(X_{j})\big{|}_{\mathsf{hyp}}\geq{1\over 2}|B_{r_{\lambda}(s_{0})/2}(X_{i})|_{\mathsf{hyp}}.

By the same reason, each point zBrλ(s0)/2(Xi)z\in B_{r_{\lambda}(s_{0})/2}(X_{i}) is covered by at most (k1)(k-1) further balls of the form Brλ(s0)/2(Xi)B_{r_{\lambda}(s_{0})/2}(X_{i^{\prime}}). Therefore,

|Xi𝒬JBrλ(s0)(X1)B(Xi,X1)|𝗁𝗒𝗉12k(𝒬(J)k)|Brλ(s0)/2(Xi)|𝗁𝗒𝗉C0(𝒬(J)k)e(d1)rλ(s0)/2,\displaystyle\Big{|}\bigcup_{X_{i}\in\mathcal{Q}_{J\setminus B_{r_{\lambda}(s_{0})}(X_{1})}}B(X_{i},X_{1})\Big{|}_{\mathsf{hyp}}\geq{1\over{2k}}(\mathcal{Q}(J)-k)|B_{r_{\lambda}(s_{0})/2}(X_{i})|_{\mathsf{hyp}}\geq C_{0}(\mathcal{Q}(J)-k)e^{(d-1)r_{\lambda}(s_{0})/2},

where in the last step we used once again Lemma 4 in [21]. Now, note that |J|𝗁𝗒𝗉O(e(d1)rλ(s0))|J|_{\mathsf{hyp}}\in O\big{(}e^{(d-1)r_{\lambda}(s_{0})}\big{)} and let for nn\in\mathbb{N}, ZnZ_{n} be a Poisson variable with mean C0(nk)+e(d1)rλ(s0)/2C_{0}(n-k)_{+}e^{(d-1)r_{\lambda}(s_{0})/2}. Thus, by the multivariate Mecke equation for Poisson point processes [20, Theorem 4.4], for nn\in\mathbb{N},

(𝒬(J)=n)\displaystyle\mathbb{P}(\mathcal{Q}(J)=n) 1n!𝔼[X1,,Xn𝒫J pairwise distinct𝟙{𝒫(Brλ(s0)(X1))k}𝟙{𝒫(XiBrλ(s0)(X1)B(Xi,X1))kn}]\displaystyle\leq\frac{1}{n!}\mathbb{E}\Big{[}\sum_{\begin{subarray}{c}X_{1},\dots,X_{n}\in\mathcal{P}\cap J\\ \text{ pairwise distinct}\end{subarray}}{\mathbbm{1}}\{\mathcal{P}(B_{r_{\lambda}(s_{0})}(X_{1}))\leq k\}{\mathbbm{1}}\Big{\{}\mathcal{P}\Big{(}\bigcup_{X_{i}\not\in B_{r_{\lambda}(s_{0})}(X_{1})}B(X_{i},X_{1})\Big{)}\leq kn\Big{\}}\Big{]}
(4.5) 1n!sλ|J|𝗁𝗒𝗉n(Znkn).\displaystyle\leq\frac{1}{n!}s_{\lambda}|J|_{\mathsf{hyp}}^{n}\mathbb{P}\big{(}Z_{n}\leq kn\big{)}.

Now, we want to apply the concentration inequality for Poisson random variables [23, Lemma 1.2]. That means, we need to provide an upper bound for the expression

exp(knlog(kneC0(nk)e(d1)rλ(s0)/2))exp(C0(nk)e(d1)rλ(s0)/2).\exp\Big{(}-kn\log\Big{(}\frac{kn}{eC_{0}(n-k)e^{(d-1)r_{\lambda}(s_{0})/2}}\Big{)}\Big{)}\exp\Big{(}-C_{0}(n-k)e^{(d-1)r_{\lambda}(s_{0})/2}\Big{)}.

To that end, we note that the first exponential is bounded by (C1e(d1)rλ(s0)/2)km\big{(}C_{1}e^{(d-1)r_{\lambda}(s_{0})/2}\big{)}^{km} for a suitable C1>0C_{1}>0. Moreover, since log(|J|𝗁𝗒𝗉)O(rλ(s0))\log(|J|_{\mathsf{hyp}})\in O(r_{\lambda}(s_{0})), we conclude that log(|J|𝗁𝗒𝗉)o(e(d1)rλ(s0)/2)\log(|J|_{\mathsf{hyp}})\in o\big{(}e^{(d-1)r_{\lambda}(s_{0})/2}\big{)}. Plugging this into (4.5) gives that for n2kn\geq 2k and suitable C2>0C_{2}>0,

(𝒬(J)=n)\displaystyle\mathbb{P}(\mathcal{Q}(J)=n) |J|𝗁𝗒𝗉n!sλ(C1e(d1)rλ(s0)/2)knexp(C0(nk)e(d1)rλ(s0)/2)\displaystyle\leq\frac{|J|_{\mathsf{hyp}}}{n!}s_{\lambda}\big{(}C_{1}e^{(d-1)r_{\lambda}(s_{0})/2}\big{)}^{kn}\exp\Big{(}-C_{0}(n-k)e^{(d-1)r_{\lambda}(s_{0})/2}\Big{)}
|J|𝗁𝗒𝗉n!sλexp(C2ne(d1)rλ(s0)/2).\displaystyle\leq\frac{|J|_{\mathsf{hyp}}}{n!}s_{\lambda}\exp(-C_{2}ne^{(d-1)r_{\lambda}(s_{0})/2}).

Therefore, by Markov’s inequality,

𝔼[ea𝒬(J)]1+e2ka|J|𝗁𝗒𝗉sλ+m2k+1|J|𝗁𝗒𝗉m!sλeC2me(d1)rλ(s0)/21+(e2ka+1)|J|𝗁𝗒𝗉sλ.\displaystyle\mathbb{E}\big{[}e^{a\mathcal{Q}(J)}\big{]}\leq 1+e^{2ka}|J|_{\mathsf{hyp}}s_{\lambda}+\sum_{m\geq 2k+1}\frac{|J|_{\mathsf{hyp}}}{m!}s_{\lambda}e^{-C_{2}me^{(d-1)r_{\lambda}(s_{0})/2}}\leq 1+(e^{2ka}+1)|J|_{\mathsf{hyp}}s_{\lambda}.

Hence, for large λ\lambda we have

𝔼[ea𝒬(J)]e(e2ka+1)|J|𝗁𝗒𝗉sλ,\mathbb{E}\big{[}e^{a\mathcal{Q}(J)}\big{]}\leq e^{(e^{2ka}+1)|J|_{\mathsf{hyp}}s_{\lambda}},

which remains bounded since |J|𝗁𝗒𝗉sλO(1)|J|_{\mathsf{hyp}}s_{\lambda}\in O(1).

Next, to bound 𝔼[ea𝒬(Q1,)]\mathbb{E}\big{[}e^{a\mathcal{Q}(Q_{1,\ell})}\big{]}, we proceed similarly as in the proof of [16, Proposition 4.3] and decompose Q1,3Q_{1,3\ell} into independent horizontal blocks. More precisely, we consider the diluted family of boxes G:={8eλ+(+1)rλ(s0)z+JQ1,3}G_{\ell}\mathrel{\mathop{\mathchar 58\relax}}=\big{\{}8e^{-\lambda+(\ell+1)r_{\lambda}(s_{0})}z+J_{\ell}\subseteq Q_{1,3\ell}\big{\}}, where we set

J:=[0,eλ+(+1)rλ(s0)]d1×T.J_{\ell}\mathrel{\mathop{\mathchar 58\relax}}=[0,e^{-\lambda+(\ell+1)r_{\lambda}(s_{0})}]^{d-1}\times T_{\ell}.

Hence, applying Hölder’s inequality we reduce the verification of (4.4) to showing that

(4.6) lim supλ0𝔼[ea𝒬(J)]d,λ<,\displaystyle\limsup_{\lambda\uparrow\infty}\prod_{\ell\leq\ell_{0}}\mathbb{E}\big{[}e^{a\mathcal{Q}(J_{\ell})}\big{]}^{d_{\ell,\lambda}}<\infty,

where d,λ:=|Q1,|𝗁𝗒𝗉/|Jλ,|𝗁𝗒𝗉d_{\ell,\lambda}\mathrel{\mathop{\mathchar 58\relax}}=|Q_{1,\ell}|_{\mathsf{hyp}}/|J_{\lambda,\ell}|_{\mathsf{hyp}}. Now, proceeding as in the case of Q1,Q_{1,\infty} shows that for large λ\lambda

0𝔼[ea𝒬(J)]d,λe(e2ka+1)d,λ|Jλ,|𝗁𝗒𝗉sλe(e2ka+1)|Q1|𝗁𝗒𝗉sλ.\displaystyle\prod_{\ell\leq\ell_{0}}\mathbb{E}\big{[}e^{a\mathcal{Q}(J_{\ell})}\big{]}^{d_{\ell,\lambda}}\leq e^{(e^{2ka}+1)d_{\ell,\lambda}|J_{\lambda,\ell}|_{\mathsf{hyp}}s_{\lambda}}\leq e^{(e^{2ka}+1)|Q_{1}|_{\mathsf{hyp}}s_{\lambda}}.

Since the last expression belongs to O(1)O(1), the proof is complete. ∎

5. Proof of Proposition 3.2

In this section, we prove Proposition 3.2, that is, the exponential equivalence of the families of random measures ηλ/uλ\eta_{\lambda}/u_{\lambda} and ξλ/uλ\xi_{\lambda}/u_{\lambda}. To that end, we will show that two error terms are negligible. More precisely, we set

Nλ(1):=#{x𝒫Wλ:Rk(x)>rλ(wλ)}N^{(1)}_{\lambda}\mathrel{\mathop{\mathchar 58\relax}}=\#\big{\{}x\in\mathcal{P}\cap W_{\lambda}^{-}\colon R_{k}(x)>r_{\lambda}(w_{\lambda})\big{\}}

and

Nλ(2):=#{x𝒫(WλWλ):Rk(x)>rλ(s0)}.N^{(2)}_{\lambda}\mathrel{\mathop{\mathchar 58\relax}}=\#\big{\{}x\in\mathcal{P}\cap(W_{\lambda}\setminus W_{\lambda}^{-})\colon R_{k}(x)>r_{\lambda}(s_{0})\big{\}}.

where Wλ:=muλ(QmQm)W_{\lambda}^{-}\mathrel{\mathop{\mathchar 58\relax}}=\bigcup_{m\leq u_{\lambda}}(Q_{m}\setminus Q^{-}_{m}). Then, we show that Nλ(1)N^{(1)}_{\lambda} and Nλ(2)N^{(2)}_{\lambda} are exponentially negligible.

Lemma 5.1 (Nλ(1)N^{(1)}_{\lambda} is exponentially negligible).

The family of random variables Nλ(1)/uλN^{(1)}_{\lambda}/u_{\lambda} is exponentially equivalent to 0.

Lemma 5.2 (Nλ(2)N^{(2)}_{\lambda} is exponentially negligible).

The family of random variables Nλ(2)/uλN^{(2)}_{\lambda}/u_{\lambda} is exponentially equivalent to 0.

Before proving Lemmas 5.1 and 5.2, we explain how to conclude the proof of Proposition 3.2.

Proof of Proposition 3.2.

First, recall that

ηλ:=muλx𝒫Qm𝟙{f(x,𝒫Qm)>s0}δf(x,𝒫Qm).\eta_{\lambda}\mathrel{\mathop{\mathchar 58\relax}}=\sum_{m\leq u_{\lambda}}\sum_{x\in\mathcal{P}_{Q^{-}_{m}}}{\mathbbm{1}}\{f(x,\mathcal{P}_{Q_{m}})>s_{0}\}\delta_{f(x,\mathcal{P}_{Q_{m}})}.

Now, since xQmx\in Q^{-}_{m}, Lemma 5.1 gives that ηλ\eta_{\lambda} is exponentially equivalent to

ηλ:=muλx𝒫Qm𝟙{f(x,𝒫)>s0}δf(x,𝒫).\eta_{\lambda}^{\prime}\mathrel{\mathop{\mathchar 58\relax}}=\sum_{m\leq u_{\lambda}}\sum_{x\in\mathcal{P}_{Q^{-}_{m}}}{\mathbbm{1}}\{f(x,\mathcal{P})>s_{0}\}\delta_{f(x,\mathcal{P})}.

Moreover, Lemma 5.2 gives that ηλ/uλ\eta_{\lambda}^{\prime}/u_{\lambda} is exponentially equivalent to

x𝒫Wλ𝟙{f(x,𝒫)>s0}δf(x,𝒫).\sum_{x\in\mathcal{P}_{W_{\lambda}}}{\mathbbm{1}}\{f(x,\mathcal{P})>s_{0}\}\delta_{f(x,\mathcal{P})}.

Noting that the latter expression coincides with ξλ\xi_{\lambda} concludes the proof. ∎

It remains to establish Lemmas 5.1 and 5.2. We begin with the proof of Lemma 5.1, which is similar to that of [16, Proposition 4.4]. Here, we stress that similarly as in the proof of [16, Proposition 4.4], the families {x𝒫Qm:Rk(x)>rλ(wλ)}\big{\{}x\in\mathcal{P}_{Q^{-}_{m}}\colon R_{k}(x)>r_{\lambda}(w_{\lambda})\big{\}} are independent for different m1m\geq 1.

Proof of Lemma 5.1.

Recalling the notation d,λ=|Q1,|𝗁𝗒𝗉/|J|𝗁𝗒𝗉d_{\ell,\lambda}=|Q_{1,\ell}|_{\mathsf{hyp}}/|J_{\ell}|_{\mathsf{hyp}}, we proceed as in the proof of Proposition 3.1. More precisely, we need to show that for all a>0a>0,

(5.1) lim supλm0𝔼[ea𝒬(J)]d,λ1,\displaystyle\limsup_{\lambda\uparrow\infty}\prod_{m\geq 0}\mathbb{E}\big{[}e^{a\mathcal{Q}^{\prime}(J_{\ell})}\big{]}^{d_{\ell,\lambda}}\leq 1,

where J:=[0,eλ+(+1)rλ(s0)]d1×TJ_{\ell}\mathrel{\mathop{\mathchar 58\relax}}=[0,e^{-\lambda+(\ell+1)r_{\lambda}(s_{0})}]^{d-1}\times T_{\ell} and

𝒬:=Xi𝒫𝟙{𝒫(Brλ(wλ)(Xi))k}δXi.\mathcal{Q}^{\prime}\mathrel{\mathop{\mathchar 58\relax}}=\sum_{X_{i}\in\mathcal{P}}{\mathbbm{1}}\big{\{}\mathcal{P}(B_{r_{\lambda}(w_{\lambda})}(X_{i}))\leq k\big{\}}\delta_{X_{i}}.

By copying the arguments from the proof of Proposition 3.1, we arrive at

lim supλm0𝔼[ea𝒬(J)]d,λe(e2ka+1)|Q1|𝗁𝗒𝗉sλ.\limsup_{\lambda\uparrow\infty}\prod_{m\geq 0}\mathbb{E}\big{[}e^{a\mathcal{Q}^{\prime}(J_{\ell})}\big{]}^{d_{\ell,\lambda}}\leq e^{(e^{2ka}+1)|Q_{1}|_{\mathsf{hyp}}s_{\lambda}^{\prime}}.

Hence, noting that |Q1|𝗁𝗒𝗉sλo(1)|Q_{1}|_{\mathsf{hyp}}s_{\lambda}^{\prime}\in o(1) concludes the proof of (5.1). ∎

Finally, we prove Lemma 5.2. Here, we will use a diluted family of cubes.

Proof of Lemma 5.2.

The idea is to proceed similarly as in the proof of Proposition 3.1. Indeed, by Markov’s inequality, it suffices to show that for every a>0a>0 we have that

log𝔼[ea𝒬(Wλ)]o(uλ).\log\mathbb{E}\big{[}e^{a\mathcal{Q}(W_{\lambda}^{-})}\big{]}\in o(u_{\lambda}).

As in the proof of Proposition 3.1, we subdivide WλW_{\lambda} into vertical layers Wλ,:=[0,1]d1×TW_{\lambda,\ell}\mathrel{\mathop{\mathchar 58\relax}}=[0,1]^{d-1}\times T_{\ell}, where T:=[eλ+rλ(s0),eλ+(+1)rλ(s0)]T_{\ell}\mathrel{\mathop{\mathchar 58\relax}}=[e^{-\lambda+\ell r_{\lambda}(s_{0})},e^{-\lambda+(\ell+1)r_{\lambda}(s_{0})}]. We also set Wλ,:=Wλ,WλW_{\lambda,\ell}^{-}\mathrel{\mathop{\mathchar 58\relax}}=W_{\lambda,\ell}\cap W_{\lambda}^{-}. In particular, we deduce from Lemma 3.4 that 𝒬Wλ,\mathcal{Q}\cap W_{\lambda,\ell}^{-} is independent of 𝒬Wλ,\mathcal{Q}\cap W_{\lambda,\ell^{\prime}}^{\prime} for all \ell and \ell^{\prime} satisfying ||3|\ell^{\prime}-\ell|\geq 3. Again, we define a value 01\ell_{0}\geq 1 such that [0,1]d1=[0,eλ+(0+2)rλ(s0)]d1[0,1]^{d-1}=\big{[}0,e^{-\lambda+(\ell_{0}+2)r_{\lambda}(s_{0})}\big{]}^{d-1}. Hence, it suffices to prove that for every a>0a>0,

(5.2) 0+1,3log𝔼[ea𝒬(Wλ,)]o(uλ) and 00,3log𝔼[ea𝒬(Wλ,)]o(uλ).\displaystyle\sum_{\ell\geq\ell_{0}+1,\ell\in 3\mathbb{Z}}\log\mathbb{E}\big{[}e^{a\mathcal{Q}(W_{\lambda,\ell}^{-})}\big{]}\in o(u_{\lambda})\quad\text{ and }\quad\sum_{0\leq\ell\leq\ell_{0},\ell\in 3\mathbb{Z}}\log\mathbb{E}\big{[}e^{a\mathcal{Q}(W_{\lambda,\ell}^{-})}\big{]}\in o(u_{\lambda}).

The arguments for layers with 13\ell-1\in 3\mathbb{Z} or 23\ell-2\in 3\mathbb{Z} are similar. The key step in the proof of (5.2) is the exponential moment bound

(5.3) log𝔼[ea𝒬(Wλ,)]O(uλ|Wλ,|𝗁𝗒𝗉/|Wλ|𝗁𝗒𝗉).\displaystyle\log\mathbb{E}\big{[}e^{a\mathcal{Q}(W_{\lambda,\ell}^{-})}\big{]}\in O\big{(}u_{\lambda}|W_{\lambda,\ell}^{-}|_{\mathsf{hyp}}/|W_{\lambda}|_{\mathsf{hyp}}\big{)}.

Once (5.3) is shown, (5.2) follows by summation over 0\ell\geq 0. We first recall from the proof of Proposition 3.1 that

(5.4) 𝔼[ea𝒬(WλI)]1+(e2ka+1)|WλI|𝗁𝗒𝗉sλ,\displaystyle\mathbb{E}\big{[}e^{a\mathcal{Q}(W_{\lambda}^{-}\cap I)}\big{]}\leq 1+(e^{2ka}+1)|W_{\lambda}^{-}\cap I|_{\mathsf{hyp}}s_{\lambda},

where II is a set of the form I=S×TI=S\times T_{\ell} with |I|𝗁𝗒𝗉O(e(d1)rλ(s0))|I|_{\mathsf{hyp}}\in O\big{(}e^{(d-1)r_{\lambda}(s_{0})}\big{)} for some Sd1S\subseteq{\mathbb{R}}^{d-1}.

Now, to prove (5.3), we first consider the case where 0+1\ell\geq\ell_{0}+1. Then, by (5.4),

log𝔼[ea𝒬(Wλ,)](e2ka+1)|Wλ,|𝗁𝗒𝗉sλ=(e2ka+1)uλ|Wλ,|𝗁𝗒𝗉/|Wλ|𝗁𝗒𝗉,\log\mathbb{E}\big{[}e^{a\mathcal{Q}(W_{\lambda,\ell}^{-})}\big{]}\leq{(e^{2ka}+1)}|W_{\lambda,\ell}^{-}|_{\mathsf{hyp}}s_{\lambda}=(e^{2ka}+1)u_{\lambda}|W_{\lambda,\ell}^{-}|_{\mathsf{hyp}}/|W_{\lambda}|_{\mathsf{hyp}},

as asserted.

It remains to deal with the case where 0\ell\leq\ell_{0}. Then, we proceed as in the proof of Proposition 3.1 and consider the diluted family of cubes G:={8eλ+(+1)rλ(s0)z+JWλ,}G_{\ell}\mathrel{\mathop{\mathchar 58\relax}}=\big{\{}8e^{-\lambda+(\ell+1)r_{\lambda}(s_{0})}z+J_{\ell}\subseteq W_{\lambda,\ell}\big{\}}, where

J:=[0,eλ+(+1)rλ(s0)]d1×T.J_{\ell}\mathrel{\mathop{\mathchar 58\relax}}=[0,e^{-\lambda+(\ell+1)r_{\lambda}(s_{0})}]^{d-1}\times T_{\ell}.

Now, we conclude from (5.4) that

(5.5) log𝔼[ea𝒬(Wλ(J+x))]O(uλ|Wλ(J+x)|𝗁𝗒𝗉/|Wλ|𝗁𝗒𝗉)\displaystyle\log\mathbb{E}\big{[}e^{a\mathcal{Q}(W_{\lambda}^{-}\cap(J_{\ell}+x))}\big{]}\in O\big{(}u_{\lambda}|W_{\lambda}^{-}\cap(J_{\ell}+x)|_{\mathsf{hyp}}/|W_{\lambda}|_{\mathsf{hyp}}\big{)}

for any xd1x\in{\mathbb{R}}^{d-1}. Hence, by Hölder’s inequality and independence, we obtain that

log𝔼[ea𝒬(Wλ,)]C1JGlog𝔼[eC2a𝒬(WλJ)].\log\mathbb{E}\big{[}e^{a\mathcal{Q}(W_{\lambda,\ell}^{-})}\big{]}\leq C_{1}\sum_{J\in G_{\ell}}\log\mathbb{E}\big{[}e^{C_{2}a\mathcal{Q}(W_{\lambda}^{-}\cap J)}\big{]}.

Since any two distinct J1,J2GJ_{1},J_{2}\in G_{\ell} are disjoint, we obtain that

JG|WλJ|𝗁𝗒𝗉=|WλJGJ|𝗁𝗒𝗉|Wλ,|𝗁𝗒𝗉.\sum_{J\in G_{\ell}}|W_{\lambda}^{-}\cap J|_{\mathsf{hyp}}=\Big{|}W_{\lambda}^{-}\cap\bigcup_{J\in G_{\ell}}J\Big{|}_{\mathsf{hyp}}\leq|W_{\lambda,\ell}^{\prime}|_{\mathsf{hyp}}.

Therefore, we get log𝔼[ea𝒬(Wλ,)]O(|Wλ,|𝗁𝗒𝗉uλ/|Wλ|𝗁𝗒𝗉),\log\mathbb{E}\big{[}e^{a\mathcal{Q}(W_{\lambda,\ell}^{-})}\big{]}\in O\big{(}|W_{\lambda,\ell}^{-}|_{\mathsf{hyp}}u_{\lambda}/|W_{\lambda}|_{\mathsf{hyp}}\big{)}, as asserted. ∎

Acknowledgement

CT was supported by the DFG priority program SPP 2265 Random Geometric Systems.

References

  • [1] F. Besau, D. Rosen, and C. Thäle. Random inscribed polytopes in projective geometries. Math. Ann., 381(3-4):1345–1372, 2021.
  • [2] F. Besau and C. Thäle. Asymptotic normality for random polytopes in non-Euclidean geometries. Trans. Amer. Math. Soc., 373(12):8911–8941, 2020.
  • [3] C. Betken, D. Hug, and C. Thäle. Intersections of Poisson kk-flats in constant curvature spaces. arXiv e-prints, page arXiv:2302.09524, Feb. 2023.
  • [4] O. Bobrowski, M. Schulte, and D. Yogeshwaran. Poisson process approximation under stabilization and Palm coupling. Ann. Henri Lebesgue, 5:1489–1534, 2022.
  • [5] J. W. Cannon, W. J. Floyd, R. Kenyon, and W. R. Parry. Hyperbolic geometry. In S. Levy, editor, Flavors of Geometry, pages 59–115. Cambridge Univ. Press, Cambridge, 1997.
  • [6] S. Chatterjee and M. Harel. Localization in random geometric graphs with too many edges. Ann. Probab., 48(1):574–621, 2020.
  • [7] D. J. Daley and D. D. Vere-Jones. An Introduction to the Theory of Point Processes I/II. Springer, New York, 2005/2008.
  • [8] A. Dembo and O. Zeitouni. Large Deviations Techniques and Applications. Springer, New York, second edition, 1998.
  • [9] M. Dostert and A. Kolpakov. Kissing number in non-Euclidean spaces of constant sectional curvature. Math. Comp., 90(331):2507–2525, 2021.
  • [10] L. Flammant. The directed spanning forest in the hyperbolic space. arXiv preprint arXiv:1909.13731, 2019.
  • [11] N. Fountoulakis, P. van der Hoorn, P. Müller, and M. Schepers. Clustering in a hyperbolic model of complex networks. Electron. J. Probab., 26:1–132, 2021.
  • [12] N. Fountoulakis and J. Yukich. Limit theory for isolated and extreme points in hyperbolic random geometric graphs. Electron. J. Probab., 25:Paper No. 141, 51, 2020.
  • [13] H.-O. Georgii and H. Zessin. Large deviations and the maximum entropy principle for marked point random fields. Probab. Theory Related Fields, 96(2):177–204, 1993.
  • [14] T. Godland, Z. Kabluchko, and C. Thäle. Beta-star polytopes and hyperbolic stochastic geometry. Adv. Math., 404(part A):Paper No. 108382, 69, 2022.
  • [15] F. Herold, D. Hug, and C. Thäle. Does a central limit theorem hold for the kk-skeleton of Poisson hyperplanes in hyperbolic space? Probab. Theory Related Fields, 179(3-4):889–968, 2021.
  • [16] C. Hirsch, T. Kang, and T. Owada. Large deviations for the volume of kk-nearest neighbor balls. arXiv preprint arXiv:2210.12423, 2022.
  • [17] C. Hirsch and T. Owada. Large deviation principle for geometric and topological functionals and associated point processes. Ann. Appl. Probab., 2023, forthcoming.
  • [18] Z. Kabluchko, D. Rosen, and C. Thäle. Fluctuations of λ\lambda-geodesic Poisson hyperplanes in hyperbolic space. arXiv e-prints, page arXiv:2205.12820, May 2022.
  • [19] O. Kallenberg. Random Measures, Theory and Applications. Springer, Cham, 2017.
  • [20] G. Last and M. Penrose. Lectures on the Poisson Process, volume 7 of Institute of Mathematical Statistics Textbooks. Cambridge University Press, Cambridge, 2018.
  • [21] M. Otto and C. Thäle. Large nearest neighbour balls in hyperbolic stochastic geometry. arXiv:2209.12730, 2022.
  • [22] T. Owada and D. Yogeshwaran. Sub-tree counts on hyperbolic random geometric graphs. Adv. in Appl. Probab., 54(4):1032–1069, 2022.
  • [23] M. D. Penrose. Random Geometric Graphs. Oxford University Press, Oxford, 2003.
  • [24] J. G. Ratcliffe. Foundations of Hyperbolic Manifolds. Springer, Cham, 2019.
  • [25] T. Schreiber and J. E. Yukich. Large deviations for functionals of spatial point processes with applications to random packing and spatial graphs. Stochastic Process. Appl., 115(8):1332–1356, 2005.
  • [26] T. Seppäläinen and J. E. Yukich. Large deviation principles for Euclidean functionals and other nearly additive processes. Probab. Theory Related Fields, 120(3):309–345, 2001.