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

00footnotetext: AMS 2010 subject classifications: Primary 60F05,82B30,60G60.

Normal approximation for fire incident simulation using permanental Cox processes

Dawud Thongtha  and Nathakhun Wiroonsrifootnotemark:
Mathematics and Statistics with Applications Research Group,
Department of Mathematics, King Mongkut’s University of Technology Thonburi
Research partially supported by the TSRI Fundamental Fund 2020 (Grant Number: 64A306000047).
Abstract

Estimating the number of natural disasters benefits the insurance industry in terms of risk management. However, the estimation process is complicated due to the fact that there are many factors affecting the number of such incidents. In this work, we propose a Normal approximation technique for associated point processes for estimating the number of natural disasters under the following two assumptions: 1) the incident counts in any two distinct areas are positively associated and 2) the association between these counts in two distinct areas decays exponentially with respect to distance outside some small local neighborhood. Under the stated assumptions, we extend previous results for the Normal approximation technique for associated point processes, i.e., the establishment of non-asymptotic L1L^{1} bounds for the functionals of these processes [28]. Then we apply this new result to permanental Cox processes that are known to be positively associated. Finally, we apply our Normal approximation results for permanental Cox processes to Thailand’s fire data from 2007 to 2020, which was collected by the Geo-Informatics and Space Technology Development Agency of Thailand.

Keywords: Correlation inequality, Cox process, Local dependence, Random fields, Natural disaster, Positive association

JEL: C020 Mathematical Methods

1 Introduction

Probability and statistical models have been widely used in most business sectors, including the insurance business. They are major tools for estimating loss, claim severity, claim count and claim probabilities, as these factors significantly affect their business operations. Since claim occurrences are used in determining policy premiums and are usually not predictable, claim simulation and prediction are some of the main tools for handling the situation. Simulation and prediction techniques for the insurance industry have attracted researchers’ attention for quite some time (see [14], [1] and [12] and references therein).

The loss can be estimated as the product of the frequency of risk events and some overall measure of severity. Therefore, approximating the number of occurrences of risk events and claim probabilities is key to successful loss estimation. These related topics have been studied and developed in various directions. For instance, the works [17] and [16] improved statistical and stochastic methods for estimating the probability and frequency of flood events. Machine learning techniques have also been applied to estimate claims and reserves (see [23], [29] and [5], for example).

A point process (sometimes called a random point field) is one of the most powerful tools in probability. These processes, which have been widely developed from their root, namely, the Poisson point process, have been applied in several areas, including insurance (see[30], [27] and [2]). Intuitively, a point process is an appropriate tool for modeling the occurrences of risk events and insurance claims, as it explains random natural phenomenon across space and time. Many researchers have proposed different approaches based on point processes for modeling natural disasters, especially fires. In 2007, in [25], it was shown that a simple point process outperforms the Burning Index in predicting wildfire incidents in Los Angeles County. In 2011, the work [30] extended the model in [25] by considering relevant covariates, such as historical spatial burn patterns and wind direction. Also, in 2015, the work [27] used a spatial point process to study the risk associated with insurance customers using geographical information systems.

Some point processes, such as the classic Poisson point process, have the property that the intensities in the relevant areas are independent. However, when considering natural disasters or insurance claims, a dependency structure must be introduced. In 2009, the work [15] proposed a copula approach to finding the joint probability distribution for hydrological variables and then using it to study the spatial dependence in extreme river flows and precipitation. When exact distributions are unknown, approximate distributions can be used to estimate probabilities of risk events. For example, the work [16] applied a theoretical result related to an approximation of a conditional distribution in [13] to develop a new method for estimating the probability of widespread flood events. In our work, we use a Normal approximation for associated point processes to estimate the total number of fire incidents. When approximating distributions, the major issue is dependency. Stein’s method ([26]) is one of the best tools for handling dependency. This well-known method for approximating limiting distributions has the additional benefit of providing non-asymptotic bounds for these distributions . In this work, we use the concept of local dependency, which can be handled by Stein’s method, as shown in [7].

This work is divided into two parts. First, we extend the theoretical results in [28] to provide non-asymptotic L1L^{1} bounds for the functionals of associated point processes under more general assumptions. These assumptions are motivated by the nature of natural disasters, as discussed later in this section. We also apply our main theoretical results to the permanental Cox process detailed in Section 2, which is known to be positively associated. In the second part, we simulate Thailand’s fire incidents using permanental Cox processes based on the satellite data on Thailand ‘s fire occurrences collected by the Geo-Informatics and Space Technology Development Agency (GISTDA) from 2007 to 2020. Finally, we check our Normal approximation results from the first part using our simulation results in the second part.

Stein’s method has been used to obtain bounds for point processes under various assumptions and Poisson processes (see [3], [4], [8], [9] and [10] for examples). However, association properties did not appear in the context of Normal approximations for point processes until the work [22] proved a central limit theorem for their functionals under certain conditions and the work [28] obtained rates of convergence using Stein’s method under a stronger assumption.

Next, we state our assumptions regarding natural disasters. Based on the literature, natural disaster incidents in a particular area seem to affect such incidents in another area. For instance, according to [30], wind may cause fire to extend to nearby areas . Thus, we are inspired to claim that for some natural disasters, such as fires, there is some positive relation between the incident counts for two distinct areas. Moreover, we also claim that the further apart the two areas are, the weaker the relation should be. Therefore, we make the following two assumptions in our work:

  1. (A1)

    The chances of incidents in any two areas are positively associated. Thus, incident counts in any two distinct areas are positively associated.

  2. (A2)

    The association between the incident counts in two distinct areas decays exponentially with respect to the distance between the areas. However, the decay is assumed to begin outside some small local neighborhood.

Though our assumptions are quite intuitive, to the best of our knowledge, we are unaware of any other works that use a dependent structure with respect to location to model natural disaster incidents.

The remainder of this work is organized as follows. We provide some necessary background and definitions in Section 2. Then the theoretical results are stated and proved in Section 3. Section 4 is devoted to the simulation of fires in Thailand using the dataset from GISTDA and checking our normal approximation results based on this simulation. Finally, a conclusion is provided in Section 5.

2 Some background and definitions

We devote this section to a review of the theoretical background of point processes, permanental Cox processes and Stein’s method.

2.1 Point and permanental Cox processes

A point process is a collection of random points on some mathematical space, such as Euclidean space d\mathbb{R}^{d}. As mentioned earlier, point processes are popular tools in the insurance industry (see [2], [27] and [30]).

Let XdX\subset\mathbb{R}^{d} with dd\in\mathbb{N} be a point process. For card(A)\text{card}(A), the cardinality of AdA\subset\mathbb{R}^{d}, let

N(A)=card(XA).\displaystyle N(A)=\text{card}(X\cap A).

The point process XX is said to be simple if N({𝐚}){0,1}N(\{{\bf a}\})\in\{0,1\} a.s. for all 𝐚d{\bf a}\in\mathbb{R}^{d}. Also, the process XX is said to be locally finite if it takes values in

Ω={xd:card(xB)<, bounded Bd}.\displaystyle\Omega=\left\{x\subset\mathbb{R}^{d}:\text{card}(x\cap B)<\infty,\forall\text{ bounded }B\subset\mathbb{R}^{d}\right\}.

The process XX is said to be negatively associated if for all coordinate-wise increasing functions ψ:k\psi:\mathbb{N}^{k}\rightarrow\mathbb{R} and ϕ:l\phi:\mathbb{N}^{l}\rightarrow\mathbb{R} and for all families of pairwise disjoint Borel sets {Ai1ik}\{A_{i}\mid 1\leq i\leq k\} and {Bj1jl}\{B_{j}\mid 1\leq j\leq l\} such that

(iAi)(jBj)=,\displaystyle(\displaystyle\cup_{i}A_{i})\cap(\displaystyle\cup_{j}B_{j})=\emptyset, (1)

we have

𝔼[ψ(N(A1),,N(Ak))ϕ(N(B1),,N(Bl))]𝔼[ψ(N(A1),,N(Ak))]𝔼[ϕ(N(B1),,N(Bl))].\mathbb{E}\left[\psi\left(N(A_{1}),\ldots,N(A_{k})\right)\phi\left(N(B_{1}),\ldots,N(B_{l})\right)\right]\\ \leq\mathbb{E}\left[\psi\left(N(A_{1}),\ldots,N(A_{k})\right)\right]\mathbb{E}\left[\phi\left(N(B_{1}),\ldots,N(B_{l})\right)\right].

Similarly, the point process is said to be positively associated if the above inequality is reversed and the families of Borel sets are not necessarily assumed to satisfy (1). In addition, a point process is said to be associated if it is either negatively or positively associated.

Next, we state the definition of the nth order intensity functions of point processes with respect to Lebesgue measure. Let nn\in\mathbb{N} and XΩX\in\Omega. If there exists a non-negative function ρn:(d)n\rho_{n}:\left(\mathbb{R}^{d}\right)^{n}\rightarrow\mathbb{R} such that

𝔼[𝐱1,,𝐱nXall distinctf(𝐱1,,𝐱n)]=(d)nf(𝐱1,,𝐱n)ρn(𝐱1,,𝐱n)𝑑𝐱1𝑑𝐱n\displaystyle\mathbb{E}\left[\sum_{\begin{subarray}{c}{\bf x}_{1},\ldots,{\bf x}_{n}\in X\\ \text{all distinct}\end{subarray}}f({\bf x}_{1},\ldots,{\bf x}_{n})\right]=\int_{\left(\mathbb{R}^{d}\right)^{n}}f({\bf x}_{1},\ldots,{\bf x}_{n})\rho_{n}({\bf x}_{1},\ldots,{\bf x}_{n})d{\bf x}_{1}\ldots d{\bf x}_{n}

for all locally integrable functions f:(d)nf:\left(\mathbb{R}^{d}\right)^{n}\rightarrow\mathbb{R}, then ρn\rho_{n} is called the nth order intensity function of XX. Now, for 𝐱,𝐲d{\bf x},{\bf y}\in\mathbb{R}^{d}, let

D(𝐱,𝐲)=ρ2(𝐱,𝐲)ρ1(𝐱)ρ1(𝐲).\displaystyle D({\bf x},{\bf y})=\rho_{2}({\bf x},{\bf y})-\rho_{1}({\bf x})\rho_{1}({\bf y}). (2)

It follows that

Cov(N(A),N(B))=A×BD(𝐱,𝐲)𝑑𝐱𝑑𝐲.\displaystyle\mathrm{Cov}(N(A),N(B))=\int_{A\times B}D({\bf x},{\bf y})d{\bf x}d{\bf y}.

Cox processes are well-known point processes. They are considered to be generalizations of the Poisson point processes for which the intensity is a random measure. A permanental Cox process [19] is the Cox process with intensity functions

ρn(𝐱1,,𝐱n)=𝔼i=1nΛ(𝐱i),\displaystyle\rho_{n}({\bf x}_{1},\ldots,{\bf x}_{n})=\mathbb{E}\prod_{i=1}^{n}\Lambda({\bf x}_{i}), (3)

where Λ\Lambda is a random measure defined as

Λ(𝐱i)=Y12(𝐱i)++Yl2(𝐱i),\displaystyle\Lambda({\bf x}_{i})=Y_{1}^{2}({\bf x}_{i})+\cdots+Y_{l}^{2}({\bf x}_{i}), (4)

where the Y1,,YlY_{1},\ldots,Y_{l} are ll independent, zero-mean, real-valued Gaussian random fields with covariance function CC.

In this work, we consider the case where the Gaussian random fields are stationary; hence, CC only depends on r=|𝐱i𝐱j|r=|{\bf x}_{i}-{\bf x}_{j}|_{\infty}, where |||\cdot|_{\infty} denotes the vector max norm. Also, we consider just the specific case when

C(r)=κceλcr,\displaystyle C(r)=\kappa_{c}e^{-\lambda_{c}r}, (5)

where κc\kappa_{c} and λc\lambda_{c} are constants. This process has been used, for instance, in [20], [31] and [21]. It is known to be positively associated (see [11] and [18]).

2.2 Stein’s bound for local dependent random variables

Stein’s method, introduced by Charles Stein [26] in 1972, is a widely known technique for finding non-asymptotic bounds for approximations of probability distributions. It was motivated by the idea that WW has the standard Normal distribution if and only if

𝔼Wf(W)=𝔼f(W)\displaystyle\mathbb{E}Wf(W)=\mathbb{E}f^{\prime}(W)

for all absolutely continuous functions ff with 𝔼|f(W)|<\mathbb{E}|f^{\prime}(W)|<\infty. This identity leads to the differential equation

h(w)𝔼h(Z)=fh(w)wfh(w),\displaystyle h(w)-\mathbb{E}h(Z)=f^{\prime}_{h}(w)-wf_{h}(w), (6)

where ZZ is a standard Normal random variable and hh is a test function. If hh\in\mathcal{H} for some Borel set \mathcal{H} and we replace ww by a random variable WW, the error in the distributional approximation of WW by ZZ on \mathcal{H} can be bounded by obtaining the non-asymptotic bound of the expectation of the supremum of the right-hand side of (6) over \mathcal{H}. In general, doing so is much simpler than computing the left-hand side of (6) directly. Taking h1h\in\mathcal{H}_{1}, where 1={h|h(x)h(y)||xy|}\mathcal{H}_{1}=\{h\mid|h(x)-h(y)|\leq|x-y|\}, we obtain

d1((W),(Z))=suph1|𝔼h(W)𝔼h(Z)|.\displaystyle d_{1}(\mathcal{L}(W),\mathcal{L}(Z))=\sup_{h\in\mathcal{H}_{1}}|\mathbb{E}h(W)-\mathbb{E}h(Z)|.

This distance is known as L1L^{1} or the Wasserstein distance. Stein’s method has been used in various applications, and it is one of the best ways to handle dependent situations (see [6] and [24]). One of the classic dependent cases, handled by Stein’s method, is the local dependent structure introduced in [7]. A collection of random variable X1,,XnX_{1},\ldots,X_{n} has dependency neighborhoods Ni{1,2,,n},i=1,2,,nN_{i}\subseteq\{1,2,\ldots,n\},i=1,2,\ldots,n if XiX_{i} is independent of XjX_{j} for all jNi{i}j\notin N_{i}-\{i\}. Next, we state a version of the local dependence bound that appeared in the note [24].

Theorem 2.1 ([24])

Let X1,,XnX_{1},\ldots,X_{n} be random variables such that 𝔼[Xi4]<\mathbb{E}[X_{i}^{4}]<\infty, 𝔼Xi=0\mathbb{E}X_{i}=0, σ2=Var(iXi)\sigma^{2}=\mathrm{Var}(\sum_{i}X_{i}), and define W=iXi/σW=\sum_{i}X_{i}/\sigma. Let the collection (X1,,Xn)(X_{1},\ldots,X_{n}) have dependency neighborhoods NiN_{i}, i=1,,ni=1,\ldots,n, and also define D=max1in|Ni|D=\max_{1\leq i\leq n}|N_{i}|. Then for ZZ, a standard Normal random variable, we have

d1((W),(Z))D2σ3i=1n𝔼|Xi|3+28D3/2πσ2i=1n𝔼[Xi4].\displaystyle d_{1}\big{(}{\cal L}(W),{\cal L}(Z)\big{)}\leq\frac{D^{2}}{\sigma^{3}}\sum_{i=1}^{n}\mathbb{E}|X_{i}|^{3}+\frac{\sqrt{28}D^{3/2}}{\sqrt{\pi}\sigma^{2}}\sqrt{\sum_{i=1}^{n}\mathbb{E}[X_{i}^{4}]}.

We apply Theorem 2.1 to obtain our main results in the next section.

3 Main results

In this section, we state and prove our main result, which is extended from [28] and then apply it to permanental Cox Processes. In this work, we relax the assumption in [28] that the covariance of the second-order intensity decays exponentially and assume that it decays exponentially outside of some local neighborhood. Let XΩX\in\Omega be an associated point process and

Y𝐢=f𝐢(XC𝐢)𝔼f𝐢(XC𝐢),𝐢d,\displaystyle Y_{\bf i}=f_{\bf i}(X\cap C_{\bf i})-\mathbb{E}f_{\bf i}(X\cap C_{\bf i}),{\bf i}\in\mathbb{Z}^{d}, (7)

where the f𝐢:Ωf_{\bf i}:\Omega\rightarrow\mathbb{R} are real-valued measurable functions and the C𝐢C_{\bf i} are defined as dd-dimensional unit cubes centered at 𝐢{\bf i}. Note that the union of C𝐢C_{\bf i} forms a covering of d\mathbb{R}^{d}.

We let 𝟏d{\bf 1}\in\mathbb{Z}^{d} denote the vector with all components 11, and write inequalities such as 𝐚<𝐛{\bf a}<{\bf b} for vectors 𝐚,𝐛d{\bf a},{\bf b}\in\mathbb{R}^{d} when they hold component-wise. In this work, we consider

S𝐤n=𝐢B𝐤nY𝐢whereB𝐤n={𝐢d:𝐤𝐢<𝐤+n𝟏}.\displaystyle S_{{\bf k}}^{n}=\sum_{{\bf i}\in B_{{\bf k}}^{n}}Y_{{\bf i}}\quad\mbox{where}\quad B_{{\bf k}}^{n}=\left\{{\bf i}\in\mathbb{Z}^{d}:{\bf k}\leq{\bf i}<{\bf k}+n{\bf 1}\right\}. (8)

The work [22] obtained a CLT for the sum above with B𝐤nB_{{\bf k}}^{n} replaced by any sequence of strictly increasing finite domains of d\mathbb{Z}^{d}. Let C𝐢C_{\bf i} be any dd-dimensional cube centered at x𝐢=R𝐢x_{\bf i}=R\cdot{\bf i} with fixed R>0R>0 and with fixed side length sRs\geq R. We state our main result for R=s=1R=s=1. We then add a remark after the theorem to discuss how to generalize this result. In the following, for kk\in\mathbb{N}, we denote Xk={𝔼|X|k}1k\left\|X\right\|_{k}=\left\{\mathbb{E}|X|^{k}\right\}^{\frac{1}{k}}.

Theorem 3.1

For dd\in\mathbb{N}, let XX be a locally finite simple associated point process on d\mathbb{R}^{d}. For 𝐤d{\bf k}\in\mathbb{Z}^{d}, let S𝐤nS_{{\bf k}}^{n} be as in (8), with Y𝐢Y_{\bf i} given in (7), R=s=1R=s=1 and σn,𝐤2=Var(S𝐤n)\sigma_{n,{\bf k}}^{2}=\mathrm{Var}(S_{{\bf k}}^{n}). Assume that the following conditions are satisfied:

  1. (a)

    The first two intensity functions of XX are well defined;

  2. (b)

    sup𝐢dY𝐢4=M<\sup_{{\bf i}\in\mathbb{Z}^{d}}\left\|Y_{\bf i}\right\|_{4}=M<\infty;

  3. (c)

    sup|𝐱𝐲|rD(𝐱,𝐲)κeλr\sup_{|{\bf x}-{\bf y}|_{\infty}\geq r}D({\bf x},{\bf y})\leq\kappa e^{-\lambda r} for some κ,λ>0\kappa,\lambda>0 and r>Kn23(4d11/d4d+2)r>Kn^{\frac{2}{3}\left(\frac{4d-1-1/d}{4d+2}\right)} with K>0K>0;

  4. (d)

    σn,𝐤2γnd\sigma_{n,{\bf k}}^{2}\geq\gamma n^{d} for some γ>0\gamma>0.

Then, for the standard Normal random variable ZZ,

d1((S𝐤nσn,𝐤),(Z))\displaystyle d_{1}\left(\mathcal{L}\left(\frac{S_{{\bf k}}^{n}}{\sigma_{n,{\bf k}}}\right),\mathcal{L}(Z)\right) \displaystyle\leq (28M2(2K)3d/2γπ+C1,d,M,κ,γ)1nd/(4d+2)+M3(2K)2dγ3/2nd/61/(6d+3)\displaystyle\left(\frac{\sqrt{28}M^{2}(2K)^{3d/2}}{\gamma\sqrt{\pi}}+C_{1,d,M,\kappa,\gamma}\right)\frac{1}{n^{d/(4d+2)}}+\frac{M^{3}(2K)^{2d}}{\gamma^{3/2}n^{d/6-1/(6d+3)}}
+C2,d,M,κ,γnd(4d+1)/(6d+3)exp(θd,M,κ,γnd/(4d+2))+C3,d,M,κ,γn7d/6exp(2θd,M,κ,γnd/(4d+2)),\displaystyle\hskip 5.0pt+\frac{C_{2,d,M,\kappa,\gamma}n^{d(4d+1)/(6d+3)}}{\exp\left(\theta_{d,M,\kappa,\gamma}n^{d/(4d+2)}\right)}+\frac{C_{3,d,M,\kappa,\gamma}n^{7d/6}}{\exp\left(2\theta_{d,M,\kappa,\gamma}n^{d/(4d+2)}\right)},

where d1d_{1} is the L1L^{1} distance,

θd,M,κ,γ=λ3(2γκ1/3((4μλ+2νλ)d(2νλ)d)18d+1πdM)1/(2d+1),\displaystyle\theta_{d,M,\kappa,\gamma}=\frac{\lambda}{3}\left(\frac{\sqrt{2\gamma}\kappa^{1/3}\left((4\mu_{\lambda}+2\nu_{\lambda})^{d}-\left(2\nu_{\lambda}\right)^{d}\right)}{18^{d+1}\sqrt{\pi}dM}\right)^{1/(2d+1)},
C1,d,M,κ,γ=(936dM4d+3((4μλ+2νλ)d(2νλ)d)2dγ2d+(3/2)πd)1/(2d+1)\displaystyle C_{1,d,M,\kappa,\gamma}=\left(\frac{9\cdot 36^{d}M^{4d+3}\left((4\mu_{\lambda}+2\nu_{\lambda})^{d}-\left(2\nu_{\lambda}\right)^{d}\right)^{2d}}{\gamma^{2d+(3/2)}\pi^{d}}\right)^{1/(2d+1)}
×(1(2d)2d/(2d+1)+2(2d)1/(2d+1)),\displaystyle\hskip 100.0pt\times\left(\frac{1}{(2d)^{2d/(2d+1)}}+2(2d)^{1/(2d+1)}\right),
C2,d,M,κ,γ=36dκ1/3M2θd,M,κ,γ4d/3πγ,C3,d,M,κ,γ=2d+1κ2/3Mγ,\displaystyle C_{2,d,M,\kappa,\gamma}=\frac{3\cdot 6^{d}\kappa^{1/3}M^{2}\theta_{d,M,\kappa,\gamma}^{4d/3}}{\sqrt{\pi}\gamma},\ \ \ C_{3,d,M,\kappa,\gamma}=\frac{2^{d+1}\kappa^{2/3}M}{\sqrt{\gamma}},

and

μλ=e2λr(eλr1)2,νλ=eλ/r(eλr1)2.\displaystyle\mu_{\lambda}=\frac{e^{\frac{2\lambda}{r}}}{\left(e^{\frac{\lambda}{r}}-1\right)^{2}},\ \ \ \nu_{\lambda}=\frac{e^{\lambda/r}}{\left(e^{\frac{\lambda}{r}}-1\right)^{2}}.

Proof: We prove this theorem by applying Theorem 2.1 stated above and Theorem 3.1 in [28], handling local dependence and non-local dependence separately. First, we address local dependence only by assuming that Cov(Y𝐢,Y𝐣)=0\mathrm{Cov}(Y_{\bf i},Y_{\bf j})=0 for |𝐢𝐣|>Kn23(4d11/d4d+2)|{\bf i}-{\bf j}|_{\infty}>Kn^{\frac{2}{3}\left(\frac{4d-1-1/d}{4d+2}\right)}. Invoking Theorem 2.1 with D=(2K)dn23(4d2d14d+2)D=(2K)^{d}n^{\frac{2}{3}\left(\frac{4d^{2}-d-1}{4d+2}\right)}, Assumption (b) that 𝔼|Y𝐣|3M3\mathbb{E}|Y_{\bf j}|^{3}\leq M^{3} and 𝔼[Y𝐣4]M4\mathbb{E}[Y_{\bf j}^{4}]\leq M^{4} and using Assumption (d), we have that the L1L^{1} distance is bounded by

28M2(2K)3d/2γπnd/(4d+2)+M3(2K)2dγ3/2nd/61/(6d+3).\displaystyle\frac{\sqrt{28}M^{2}(2K)^{3d/2}}{\gamma\sqrt{\pi}n^{d/(4d+2)}}+\frac{M^{3}(2K)^{2d}}{\gamma^{3/2}n^{d/6-1/(6d+3)}}. (9)

We now assume that Assumption (c) is true for all r>0r>0. Invoking Theorem 3.1 in [28], we have that the L1L^{1} distance is bounded by

C1,d,M,κ,γnd/(4d+2)+C2,d,M,κ,γnd(4d+1)/(6d+3)exp(θd,M,κ,γnd/(4d+2))+C3,d,M,κ,γn7d/6exp(2θd,M,κ,γnd/(4d+2))\displaystyle\frac{C_{1,d,M,\kappa,\gamma}}{n^{d/(4d+2)}}+\frac{C_{2,d,M,\kappa,\gamma}n^{d(4d+1)/(6d+3)}}{\exp\left(\theta_{d,M,\kappa,\gamma}n^{d/(4d+2)}\right)}+\frac{C_{3,d,M,\kappa,\gamma}n^{7d/6}}{\exp\left(2\theta_{d,M,\kappa,\gamma}n^{d/(4d+2)}\right)} (10)

The distance d1((W𝐤n),(Z))d_{1}\big{(}{\cal L}(W_{{\bf k}}^{n}),{\cal L}(Z)\big{)} is bounded by the sum of (9) and (10) because the local neighborhood covariance terms and the remaining covariance terms have been handled by (9) and (10), respectively. \Box

Remark 3.2
  1. 1.

    The associated assumption and Assumption (c) are motivated by (A1) and (A2) in the introduction, respectively.

  2. 2.

    The size of the local neighborhood is O(nd/(4d+2))O\left(n^{d/(4d+2)}\right), which is flexible and can be increasing in nn.

  3. 3.

    The above theorem can be extended to the case where RR and SS are greater than one by following the same proof. The bound will end up with a larger constant.

Note that we added the assumption that (c) is true for all r>0r>0 in the above proof after the local covariance terms had been handled. Although doing so may make the constant larger, the rate of convergence is unaffected.

Next, we apply the above theorem to the permanental Cox process on d\mathbb{R}^{d}, which is known to be positively associated. In this theorem, we consider a function f:Ωf:\Omega\rightarrow\mathbb{R} defined by

f(Y)=SYg(S)𝟏|S|=p,\displaystyle f(Y)=\sum_{S\subset Y}g(S)\mathbf{1}_{|S|=p}, (11)

where gg is a bounded function supported on sets SS having exactly pp elements such that g(S)=0g(S)=0 when diam(S)>τdiam(S)>\tau for some fixed τ>0\tau>0 and pp\in\mathbb{N}. Here, we denote diam(S)=supx,yS|xy|diam(S)=\sup_{x,y\in S}\left|x-y\right|_{\infty}. Also, we focus specifically on the case that p=1p=1. We note here that if p=1p=1 and g(S)=1g(S)=1, f(XΛn)f(X\cap\Lambda_{n}) is the number of points in XΛnX\cap\Lambda_{n} or N(Λn)N(\Lambda_{n}), where Λnd\Lambda_{n}\subset\mathbb{R}^{d}. The result is stated as follows:

Theorem 3.3

Let n,kn,k\in\mathbb{N} and XX be a permanental Cox process with intensity functions as in (3) and Λ(𝐱i)\Lambda({\bf x}_{i}) as in (4), where the YiY_{i}, i=1,,ki=1,\cdots,k are independent, mean-zero, real-valued Gaussian random fields. Let ff be defined as in (11) with gg be bounded and g(S)=0g(S)=0 when diam(S)>τdiam(S)>\tau for some fixed τ>0\tau>0 and p=1p=1. Letting σn2=Var(f(XΛn))\sigma_{n}^{2}=\mathrm{Var}(f(X\cap\Lambda_{n})) with Λnd\Lambda_{n}\subset\mathbb{R}^{d} be such that |Λn|=nd|\Lambda_{n}|=n^{d}, assume that

sup|𝐱𝐲|rCov(Λ(𝐱),Λ(𝐲))O(eλr),\displaystyle\sup_{|{\bf x}-{\bf y}|_{\infty}\geq r}\mathrm{Cov}\left(\Lambda({\bf x}),\Lambda({\bf y})\right)\leq O(e^{-\lambda r}), (12)

for some λ>0\lambda>0 and r>Knd/(4d+2)r>Kn^{d/(4d+2)} with K>0K>0, and

lim infnσn2/nd>0.\displaystyle\liminf_{n}\sigma_{n}^{2}/n^{d}>0. (13)

Then for Wn=(f(XΛn)𝔼f(XΛn))/σnW_{n}=(f(X\cap\Lambda_{n})-\mathbb{E}f(X\cap\Lambda_{n}))/\sigma_{n}, there exists C>0C>0 such that

d1((Wn),(Z))Cnd/(4d+2),\displaystyle d_{1}\left(\mathcal{L}(W_{n}),\mathcal{L}(Z)\right)\leq\frac{C}{n^{d/(4d+2)}}, (14)

where ZZ is a standard Normal random variable.

Proof: To prove the theorem, we follow the same argument as that used in the proof of Theorem 4.3 in [28]. It requires the use of Theorem 3.1, where the assumptions (12) and (13) are needed. Because Theorem 4.3 of [28] is for determinantal point processes, the variance condition used in its proof cannot be used here; thus, the proofs differ at this point. Therefore, it is sufficient to show that

Var(SXΛng(S)𝟏|S|=p)=O(Λn),\displaystyle\mathrm{Var}\left(\sum_{S\subset X\cap\Lambda_{n}}g(S)\mathbf{1}_{|S|=p}\right)=O(\Lambda_{n}),

which can be proved by adapting the proof of Lemma B.6 in [22] for the case where XX is a permanental Cox process. It is sufficient to verify that the term (B.4) in that proof is bounded for k=0,,pk=0,\ldots,p. For the term k=0k=0, we show that the intensity ρ2p\rho_{2p} is bounded as follows:

|ρ2p(𝐱1,,𝐱2p)|\displaystyle|\rho_{2p}({\bf x}_{1},\cdots,{\bf x}_{2p})| =\displaystyle= 𝔼j=12pΛ(𝐱j)=𝔼j=12p(Y12(𝐱j)++Yl2(𝐱j))\displaystyle\mathbb{E}\prod_{j=1}^{2p}\Lambda({\bf x}_{j})=\mathbb{E}\prod_{j=1}^{2p}\left(Y^{2}_{1}({\bf x}_{j})+\cdots+Y^{2}_{l}({\bf x}_{j})\right)
=\displaystyle= 1i1,,i2pl𝔼[Yi12(𝐱1)Yi22(𝐱2)Yi2p2(𝐱2p)]\displaystyle\sum_{1\leq i_{1},\ldots,i_{2p}\leq l}\mathbb{E}\left[Y^{2}_{i_{1}}({\bf x}_{1})Y^{2}_{i_{2}}({\bf x}_{2})\cdots Y^{2}_{i_{2p}}({\bf x}_{2p})\right]
\displaystyle\leq 1i1,,i2pl(𝔼[Yi12p(𝐱1)]𝔼[Yi22p(𝐱2)]𝔼[Yi2p2p(𝐱2p)])1/p\displaystyle\sum_{1\leq i_{1},\ldots,i_{2p}\leq l}\left(\mathbb{E}\left[Y^{2p}_{i_{1}}({\bf x}_{1})\right]\mathbb{E}\left[Y^{2p}_{i_{2}}({\bf x}_{2})\right]\cdots\mathbb{E}\left[Y^{2p}_{i_{2p}}({\bf x}_{2p})\right]\right)^{1/p}
\displaystyle\leq l2p(M(2p1)!!)2,\displaystyle l^{2p}\left(M(2p-1)!!\right)^{2},

where M=max1il,𝐱dσi2(𝐱)M=\max_{1\leq i\leq l,{\bf x}\in\mathbb{R}^{d}}\sigma^{2}_{i}({\bf x}), with σi2(𝐱)=Var(Yi(𝐱))\sigma^{2}_{i}({\bf x})=\mathrm{Var}(Y_{i}({\bf x})). Note that we have used Hölder’s inequality and the fact that YiY_{i} is Normal and its central moment is

𝔼|Yi(𝐱)|2p=σi2p(𝐱)(2p1)!!\mathbb{E}|Y_{i}({\bf x})|^{2p}=\sigma^{2p}_{i}({\bf x})(2p-1)!!.

As we only consider the case p=1p=1, the term k=1k=1 is obvious. Thus, the order of the variance is verified. \Box

Remark 3.4

We remark here that in the proof of Theorem 3.3, we set M=max1il,𝐱dσi2(𝐱)M=\max_{1\leq i\leq l,{\bf x}\in\mathbb{R}^{d}}\sigma_{i}^{2}({\bf x}), which could be large. However, in our application, we later set the σi2(𝐱)\sigma_{i}^{2}({\bf x}) to be equal for all ii and σi2(𝐱)=m(𝐱)/l\sigma_{i}^{2}({\bf x})=m({\bf x})/l, where m(𝐱)m({\bf x}) is the maximum number of fires in area 𝐱{\bf x}, as derived from the historical data, which is assumed to be bounded.

4 Application to the Thai fire dataset

In this section, we use our main results to simulate Thailand’s fires via permanental Cox processes. We use the GISTDA Thai fire dataset, collected from 2007 to 2020 by satellite. The dataset consists of the latitudes and longitudes of all fire incident locations. Recall that we claim that the fire counts from two distinct areas should be positively correlated. Moreover, we claim that their covariance begins to decay exponentially outside some small local neighborhood. We split this section into three subsections. We first explore the dataset and check to see that Assumptions (A1) and (A2) are not contradicted by the data in the first subsection. The second subsection is devoted to applying our main results to fire simulations. In the last subsection, we evaluate the effect of varying the decay parameter for the permanental Cox processes.

4.1 Exploring the dataset

In this subsection, we first explore the dataset and verify that it does not contradict the assumptions in Theorem 3.3. We follow the process outlined in Figure 1. Figure 2 shows the total number of fires each year in Thailand, where the mean and the standard deviation were 27,048 and 6,980.94, respectively. Figure 3 shows the locations of fires in 2007, 2014 and 2020, respectively, from left to right.

Fires data 2007 - 2020Verifying the assumptions (Theorem 3.3)Descriptive statisticsAssumption (12)Assumption (13) The variance orderLocal neighborhoodExponential decay
Figure 1: Exploring the GISTDA fire dataset
Refer to caption
Figure 2: The total number of yearly fires for the period 2007-2020
Refer to caption
Figure 3: The actual fire locations in 2007, 2014 and 2020

We now examine whether the fire dataset follows assumptions (12) and (13). As the actual intensity function is unknown and cannot be obtained from the dataset, we can only check the covariances between fire counts in two distinct areas separated by a distance rr. We first check to see if the covariance does not decay in a small local neighborhood. Then we further check to see if the covariance decays exponentially as the distance between areas increases. We also show the variances of the fire counts corresponding to nn, with nn going from 10 to 645, where 645645 is the greatest distance possible in Thailand (1 unit equals to 0.01 degrees of latitude or longitude). Note that the total area of Thailand is 513,120 square kilometers, which equals 7162716^{2}. Since 1 degree is about 111 kilometers, the total area is approximately 6.4526.45^{2} square degrees, which is equivalent to 6452645^{2} square units when 1 unit is 0.01 degrees. Therefore, the largest nn possible in Thailand is 645.

As the entire country is too large to be considered a small local neighborhood, we specifically check four provinces from four regions, including Bangkok, Chaing Mai, Kanchanaburi and Khon Kaen. Note that we did not select a province from Southern Thailand, as the number of fires in this region was too low. Figure 4 shows no sign of exponential decay in the covariance pattern for these provinces when the unit distance is 0.002 degrees (222\sim 222 meters). Figure 5 shows the covariance decay pattern based on distance for all of Thailand. We obtain an exponential decay rate of 0.15 when we use a unit distance of 0.01 degrees (1.11\sim 1.11 kilometers). Figure 6 shows the values of σn2/n2\sigma_{n}^{2}/n^{2} for nn from 10 to 645. Obviously, assumption 13 is not contradicted.

Refer to caption
Figure 4: Covariance decay patterns for Bangkok, Chaing Mai, Kanchanaburi and Khon Kaen with a unit distance of 0.002 degrees
Refer to caption
Figure 5: Covariance decay pattern for Thailand with a unit distance of 0.01 degrees
Refer to caption
Figure 6: Variance of the total number of fires divided by n2n^{2}

4.2 Simulation of Thai fires using the permanental Cox process

We devote this subsection to simulating fires in Thailand via the process outlined in Figure 7.

Fires data 2007 - 2020Estimating permanental Cox process parameters σ2,l\sigma^{2},l in (4)Estimating covariance decaying parameters λc\lambda_{c} and KcK_{c} in (5)Simulating Fires corresponding to the estimated parametersTesting normality of the simulated firesVerifying assumptions (12) and (13) for the simulated firesComputing the L1L_{1} distance between the number of simulated fires and a standard normalComparing the rate of convergence to the bound in (14)
Figure 7: Simulation of Thailand’s fires

We first model Thailand’s fires using permanental Cox processes. Note that when we refer to the area 𝐱{\bf x}, we are referring to a unit square that has 𝐱{\bf x} at the top left corner, where 𝐱2{\bf x}\subset\mathbb{R}^{2} and 1 unit equals 0.01 degrees of latitude or longitude. We estimate the variances of Yj(𝐱i)Y_{j}({\bf x}_{i}) and ll from

Λ(𝐱i)=Y12(𝐱i)++Yl2(𝐱i)\displaystyle\Lambda({\bf x}_{i})=Y_{1}^{2}({\bf x}_{i})+\cdots+Y_{l}^{2}({\bf x}_{i})

by using the first two sample moments. We set the means of the data and process to be equal and seek the closest variances. Writing σj2(𝐱i)=Var(Yj(𝐱i))\sigma_{j}^{2}({\bf x}_{i})=\mathrm{Var}(Y_{j}({\bf x}_{i})) for each 𝐱i{\bf x}_{i} and using the fact that the Yj(𝐱i)Y_{j}({\bf x}_{i}) for j=1,2,,lj=1,2,\ldots,l are zero-mean, independent Gaussian random variables, we set j=1lσj2(𝐱i)\sum_{j=1}^{l}\sigma_{j}^{2}({\bf x}_{i}) equal to the sample mean of area 𝐱i{\bf x}_{i} and 2j=1lσj4(𝐱i)2\sum_{j=1}^{l}\sigma_{j}^{4}({\bf x}_{i}) equal to the sample variance of area 𝐱i{\bf x}_{i}. We then set σj2(𝐱i)=σ2(𝐱i)\sigma_{j}^{2}({\bf x}_{i})=\sigma^{2}({\bf x}_{i}) for all jj for simplicity and solve for the closest σ2(𝐱i)\sigma^{2}({\bf x}_{i}) and ll.

Denoting m(𝐱i)m({\bf x}_{i}) and v(𝐱i)v({\bf x}_{i}) as the mean and variance of the total fires in the unit cube at position 𝐱i{\bf x}_{i} from the dataset, we have

σ2(𝐱i)=m(𝐱i)l for all i.\displaystyle\sigma^{2}({\bf x}_{i})=\frac{m({\bf x}_{i})}{l}\text{ \ for all \ }i.

Then setting

2lσ4(𝐱i)=j=1lVar(Yj2(𝐱i))=v(𝐱i),\displaystyle 2l\sigma^{4}({\bf x}_{i})=\sum_{j=1}^{l}\mathrm{Var}(Y^{2}_{j}({\bf x}_{i}))=v({\bf x}_{i}),

and substituting σ2(𝐱i)\sigma^{2}({\bf x}_{i}), we have

l=2m2(𝐱i)v(𝐱i).\displaystyle l=\frac{2m^{2}({\bf x}_{i})}{v({\bf x}_{i})}.

For simplicity, we set l=max(1,2m2(𝐱i)v(𝐱i))l=\max\left(1,\left\lfloor\frac{2m^{2}({\bf x}_{i})}{v({\bf x}_{i})}\right\rfloor\right). Although the ll’s differ for different 𝐱i{\bf x}_{i}, we can choose the largest ll. For a 𝐱i{\bf x}_{i} with a smaller ll, we set σ2(𝐱i)\sigma^{2}({\bf x}_{i}) equal to zero for all the remaining terms.

To estimate the covariance function of the Gaussian processes from (5), we need estimates of κc\kappa_{c} and λc\lambda_{c}. Plugging in r=0r=0, we have κc=σ2(𝐱i)\kappa_{c}=\sigma^{2}({\bf x}_{i}); thus, it suffices to estimate λc\lambda_{c}. We seek a λc\lambda_{c} from a fine grid from 0 to 0.5 that results in the covariance decay rate closest to the 0.15 from the real dataset. Note that the decay rate of 0.15 is computed from the number of fires, whereas λc\lambda_{c} is the decay rate of the covariance function for the Gaussian process. Therefore, it is not possible to set λc=0.15\lambda_{c}=0.15 directly.

By following the procedure above, we find the largest ll, i.e., l=8l=8, and closest λc(0.1\lambda_{c}(0.1). Figure 8 shows an example of a simulation with parameters estimated from the procedure above and 6 iterations.

Refer to caption
Figure 8: This figure shows fire count simulation results using the estimated parameters and six iterations. Here, the simulation’s mean and standard deviation are 26,873 and 1078.15, respectively.

Next, we simulate fire counts for the whole country for 20 iterations and check that the covariance decays exponentially with distance. Since the number of areas in the entire country at a unit distance of 0.01 degrees from another area is extremely large, 20 is a reasonable number of iterations for the simulation. Figure 9 shows the covariance plot and the decay rate of 0.15 from the real dataset.

Refer to caption
Figure 9: The black dots show the covariances between the simulated fire counts for two distinct unit areas according to the distance between the areas. The red line shows exponential decay at a rate of 0.15.

Then we simulate fires across the whole country for 500 iterations to check that the variance is on the order of n2n^{2} and to compare the results with the Normal approximation from Theorem 3.3. As in Subsection 4.1, we take nn from 10 to 645, where 645645 units is the greatest distance possible in Thailand (1 unit equals 0.01 degrees of latitude or longitude). We preliminarily check the normality of the simulated fire counts using the Shapiro–Wilk test. Figure 10 shows the Shapiro–Wilk p-values based on nn, while the red horizontal line indicates the 0.05 significance level. We can see that when nn is large, the null hypothesis that the total number of fires is Normally distributed is not rejected.

Refer to caption
Figure 10: The Shapiro–Wilk p-values for the fire counts

We end this section by checking the variance assumption in assumption (13) and compare the simulation’s L1L^{1} distance to the bound in (14). Figure 11 shows a random pattern for the variances of fire counts divided by n2n^{2}, which agrees with assumption (13). Figure 12 shows the L1L^{1} distances between the standardized fire counts from our simulation and a standard Normal distribution . We also plot the rate of convergence of nd/(4d+2)n^{d/(4d+2)} from Theorem 3.3. The rate from our simulation tends to follow the rate from our main bound.

Refer to caption
Figure 11: The variances of fire counts divided by n2n^{2}
Refer to caption
Figure 12: The L1L^{1} distances between the standardized fire counts and a standard Normal. distrbution

4.3 The effect of changing the covariance decay rate

The main assumptions that we made in this work is that the fire counts for two distinct areas are positively associated and their covariance decays exponentially with respect to their distance from each other. Nevertheless, the decay rate may vary from country to country or area to area and may also change over time. Therefore, in this section, we are interested in examining the Normal approximation results for different covariance decay rates through the parameter λc\lambda_{c} from (5).

Figure 13 shows that the L1L^{1} distances between the standardized fire counts from our simulation and a standard Normal distribution for areas ranging in size from 9 to 645 square degrees when λc=1,0.2,0.1,0.05,0.017\lambda_{c}=1,0.2,0.1,0.05,0.017. The black line shows the rate of convergence of nd/(4d+2)n^{d/(4d+2)} from Theorem 3.3. Intuitively, the distribution should be closer to the Normal distribution when λc\lambda_{c} is larger, as a larger value implies that the fire counts in the areas are closer to independence. This reasoning agrees with the simulation results shown in Figure 13.

Refer to caption
Figure 13: Normal approximations for different values of λc\lambda_{c}

5 Discussion and conclusion

In this work, we break our contribution into two parts: developing theories and simulating Thailand fire counts. For the first part, we generalize the normal approximation of a functional of the associated point processes in [28] by relaxing the assumption that the covariance decays exponentially everywhere to exempting local neighborhoods. Then we apply the main result (Theorem 3.1) to permanental Cox processes, which are known to be positively associated. Unlike the main theorem involving hypercubes of size ndn^{d} in d\mathbb{R}^{d}, Theorem 3.3 is flexible in that it covers Normally approximating a functional of permanental Cox processes in an area of any shape of size ndn^{d}, which allows us to apply this result to the fire counts in Thailand.

For the simulation, we use a permanental Cox process to simulate fire counts in Thailand. We assume that the fire counts for two distinct areas are positively correlated. Moreover, we assume that their covariance decays exponentially outside of some small local neighborhood. We use the GISTDA Thailand fire dataset collected by Thailand’s satellites from 2007 to 2020 to estimate the parameters for the process. The L1L^{1} distance between the total number of fires from our simulated fire counts and the Normal distribution agrees with the rate from the bound in Theorem 3.3. By varying the covariance parameter in (5), we can vary the L1L^{1} distances from our simulated results. Nevertheless, Normality still holds, and the rates do not contradict the one in our main theorem.

In real-world applications, the new approach proposed in this research may be used to model any natural disaster incidents or model claims by area for property insurance firms. Thus, it benefits policymakers both in government and the private sector in terms of managing risk. Moreover, future researchers can use our idea to simulate natural disaster incidents or insurance claims with some other point processes that are not necessarily well-known but satisfy the assumptions of Theorem 3.1.

Acknowledgment

The authors would like to thank the Geo-Informatics and Space Technology Development Agency (GISTDA) for use of the fire dataset.

Declarations

Funding: This work was supported financially by the TSRI Fundamental Fund 2020 (Grant Number: 64A306000047).

Conflicts of interest/Competing interests: No potential conflicts of interest were reported by the authors.

Availability of data and material: This dataset is under the license of the Geo-Informatics and Space Technology Development Agency.

Code availability: Available upon request.

IRB approval: Not applicable.

References

  • [1] Abledu, G. K., Dadey, E. and Kobina , A., Probability modeling and simulation of insurance claims in Ghana, Global Journal of Commerce & Management Perspective, 3(5) (2014) 41–49.
  • [2] Albrecher, H., Araujo-Acuna, J. C. and Beirlant, J., Fitting non-stationary Cox process: an application to fire insurance data, North American Actuarial Journal, 0(0) (2020) 1–28.
  • [3] Barbour, A. D., Stein’s method and Poisson process convergence, Journal of Applied Probability, 25 (1988) 175–184.
  • [4] Barbour, A. D. and Brown, T. C., Stein’s method and point process approximation, Stochastic Processes and their Applications, 43(1) (1992) 9–31.
  • [5] Bärtl, M. and Krummaker, S., Prediction of claims in export credit finance: a comparison of four machine learning techniques, Risks, 8(22) (2020) 1–29.
  • [6] Chen, L.Y.H., Goldstein, L. and Shao, Q.M., Normal Approximation by Stein’s Method, Springer, New York, 2011.
  • [7] Chen, L.Y.H. and Shao, Q.M., Normal approximation under local dependence, Annals of Probability, 32 (2004) 1985–2028.
  • [8] Chen, L.Y.H. and Xia, A., Stein’s method, Palm theory and Poisson process approximation, Annals of Probability, 32(3) (2004) 2545–2569.
  • [9] Chen, L.Y.H. and Xia, A., Poisson process approximation: from Palm theory to Stein’s method, IMS Lecture Notes-Monograph Series: Time Series and Related Topics, 52 (2006) 236–244.
  • [10] Chen, L.Y.H. and Xia, A., Poisson process approximation for dependent superposition of point processes, Bernoulli, 17(2) (2011) 530–544.
  • [11] Eisenbaum, N., Characterization of positively correlated squared Gaussian processes, The Annals of Probability, 42(2) (2014) 559–575.
  • [12] Gabrielli, A. and Wüthrich, M. V., An individual claims history simulation machine, Risks, 6(29) (2018) 1–33.
  • [13] Heffernan, J.E. and Resnick, S.I., Limit laws for random vectors with an extreme component, Annals of Applied Probability, 17 (2007) 537–571.
  • [14] Jessen A.H., Mikosch T. and Samorodnitsky G., Prediction of outstanding payments in a Poisson cluster model, Scandinavian Actuarial Journal, 3(2011) 214–237.
  • [15] Keef, C., Svensson C. and Tawn, J.A., Spatial dependence in extreme river flows and precipitation for Great Britain, Journal of Hydrology, 378 (2009) 240–252.
  • [16] Keef, C., Tawn, J. A. and Lamb, R., Estimating the probability of widespread flood events, Environmetrics, 24 (2013) 13–21.
  • [17] Lamb, R., Rainfall-runoff modelling for flood frequency estimation, Encyclopedia of Hydrological Sciences, Anderson MG (eds). John Wiley & Sons(2005) 1923–1954.
  • [18] Last, G., Szekli, R., Yogeshwaran, D., Some remarks on associated random fields, random measures and point processes, ALEA Latin American Journal of Probability and Mathematical Statistics, 17 (2020) 355–374.
  • [19] McCullagh, P., Møller, J., The permanental process, Advances in Applied Probability, 38(4) (2006) 873–888.
  • [20] McCullagh, P., Yang, J., Stochastic classification models, International Congress of Mathematicians, 3 (2006) 669–686.
  • [21] Mustafa, H. A., Ekti, A. R., Shakir, M. Z., Imran, M. A., Tafazolli, R., Intracell interference characterization and cluster interference for D2D communication, IEEE Transactions on Vehicular Technology, 67(9) (2018) 8536–8548.
  • [22] Poinas, A., Delyon, B. and Lavancier, F., Mixing properties and central limit theorem for associated point processes, Bernoulli, 25(3) (2019) 1724–1754.
  • [23] Quan, Z. and Valdez, E.A., Predictive analytics of insurance claims using multivariate decision trees, Dependence Modeling, 6 (2018) 377–407.
  • [24] Ross, N. Fundamentals of Stein’s method. Probability Surveys, 8 (2011) 210–293.
  • [25] Schoenberg, F.P., Chang, C., Keeley, J., Pompa, J., Woods, J. and Xu, H., A critical assessment of the burning index in Los Angeles County, California, International Journal of Wildland Fire, 16(4) (2007) 473–483.
  • [26] Stein, C., A bound for the error in the normal approximation to the distribution of a sum of dependent random variables. Proceedings of the Sixth Berkeley Symposium on Mathematical Statistics and Probability University of California Press, 2 (1972) 210–293.
  • [27] Törnqvist, G., Modelling insurance claims with spatial point process: an applied case-control study to improve the use of geographical information in insurance pricing, Master thesis, Umeå University, 2015
  • [28] Wiroonsri, N., Normal approximation for associated point processes via Stein’s method with applications to determinantal point processes, Journal of Mathematical Analysis and Applications, 480(1) (2019) 123396.
  • [29] Wüthrich M. V., Machine learning in individual claims reserving, Scandinavian Actuarial Journal, (2018) 1–16.
  • [30] Xu, H. and Schoenberg F.P., Point process modeling of wildfire hazard in Los Angeles County, California, The Annals of Applied Statistics, 5(2011) 684–704.
  • [31] Yang, J., Miescke, K., McCullagh, P., Classification based on a permanental process with cyclic approximation, Biometrika, 99(4) (2012) 775–786.