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

Scaling limits of stationary determinantal shot-noise fields

Takumi Aburayama and Naoto Miyoshi
Tokyo Institute of Technology
Corresponding author: Department of Mathematical and Computing Science, Tokyo Institute of Technology, 2-12-1-W8-52 Ookayama, Tokyo 152-8552, Japan. E-mail: [email protected] second author’s work was supported by the Japan Society for the Promotion of Science Grant-in-Aid for Scientific Research (C) 19K11838.
Abstract

We consider a shot-noise field defined on a stationary determinantal point process on d\mathbb{R}^{d} associated with i.i.d. amplitudes and a bounded response function, for which we investigate the scaling limits as the intensity of the point process goes to infinity. Specifically, we show that the centralized and suitably scaled shot-noise field converges in finite dimensional distributions to i) a Gaussian random field when the amplitudes have the finite second moment and ii) an α\alpha-stable random field when the amplitudes follow a regularly varying distribution with index α-\alpha for α(1,2)\alpha\in(1,2). We first prove the corresponding results for the shot-noise field defined on a homogeneous Poisson point process and then extend them to the one defined on a stationary determinantal point process.
Keywords: Shot-noise fields; determinantal point processes; scaling limits; stable random fields; regularly varying distributions.

1 Introduction

Let Φλ=n=1δXn\Phi_{\lambda}=\sum_{n=1}^{\infty}\delta_{X_{n}} denote a simple and stationary point process on d\mathbb{R}^{d}, d={1,2,}d\in\mathbb{N}=\{1,2,\ldots\}, with intensity λ=𝔼[Φλ([0,1]d)](0,)\lambda=\mathbb{E}[\Phi_{\lambda}([0,1]^{d})]\in(0,\infty). We consider a class of shot-noise fields given by

Iλ(z)=n=1Pn(zXn),zd,I_{\lambda}(z)=\sum_{n=1}^{\infty}P_{n}\,\ell(z-X_{n}),\quad z\in\mathbb{R}^{d}, (1)

where PnP_{n}, nn\in\mathbb{N}, are independent and identically distributed (i.i.d.) nonnegative random variables, called amplitudes, which are also independent of Φλ\Phi_{\lambda}, and \ell is a nonnegative and bounded function on d\mathbb{R}^{d}, called a response function. Shot-noise fields such as (1) have been observed in various areas and studied extensively in the literature (cf. [7, Sec. 5.6] and references therein). In particular, recently, they have been used as models for interference fields in wireless communication networks, where wireless interferers are located according to a spatial point process (cf. [8, 2]). In this work, we study the scaling limits of (1) as λ\lambda\to\infty when Φλ\Phi_{\lambda} is a stationary determinantal point process. Determinantal point processes represent a repulsive feature of points in space and have also been considered as location models of base stations in cellular wireless networks (cf. [16, 25, 15]).

As an early result on the scaling limits of spatial shot-noise fields, Heinrich and Schmidt [10] consider a more general form than (1) and show that the centralized and scaled shot-noise at one position converges in distribution to a Gaussian random variable when Φλ\Phi_{\lambda} is Brillinger mixing and a condition which corresponds to 𝔼[Pn2]<\mathbb{E}[{P_{n}}^{2}]<\infty in (1) is satisfied. Since Biscio and Lavancier [5] and Heinrich [9] prove that stationary (α\alpha-)determinantal point processes are Brillinger mixing, the result in [10], of course, covers the scaling limit of (1) at one position when 𝔼[Pn2]<\mathbb{E}[{P_{n}}^{2}]<\infty. Based on this background, we here examine the convergence in finite dimensional distributions and show that the centralized and suitably scaled version of (1) converges to a Gaussian random field when 𝔼[Pn2]<\mathbb{E}[{P_{n}}^{2}]<\infty and to an α\alpha-stable random field when PnP_{n}, nn\in\mathbb{N}, follow a regularly varying distribution with index α-\alpha for α(1,2)\alpha\in(1,2). As related work along this direction, Baccelli and Biswas [1] consider the shot-noise field (1), where Φλ\Phi_{\lambda} is a homogeneous Poisson point process and the response function (x)\ell(x) is power-law and diverges as x0x\to 0, and show that a suitably scaled (but non-centralized) version converges in finite dimensional distributions to an α\alpha-stable random field with α(0,1)\alpha\in(0,1). Aside from this, Kaj et al. [13] consider a random grain field defined on a homogeneous Poisson point process associated with a regularly varying volume distribution, and derive some scaling limits in the sense of convergence in finite dimensional distributions. Furthermore, the results in [13] are extended by Breton et al. [6] to the one defined on a stationary determinantal point process.

The rest of the paper is organized as follows. In the next section, after providing the centralized and scaled version of (1), we prove the corresponding results for the shot-noise field defined on a homogeneous Poisson point process. These proofs provide the basis for showing our main results, and then in Section 3, we extend them to the one defined on a stationary determinantal point process. Finally, conclusion is given in Section 4.

2 Preliminary: Scaling limits of Poisson shot-noise fields

Throughout the paper, we assume that the distribution function FPF_{P} of the amplitudes PnP_{n}, nn\in\mathbb{N}, has the finite mean p=0tdFP(t)<p=\int_{0}^{\infty}t\,\mathrm{d}F_{P}(t)<\infty, and the response function \ell is bounded and satisfies d(x)dx<\int_{\mathbb{R}^{d}}\ell(x)\,\mathrm{d}x<\infty. With this condition, the shot-noise field IλI_{\lambda} in (1) is also stationary on d\mathbb{R}^{d} and Campbell’s formula (cf. [3, p. 8, Theorem 1.2.5]) leads to the expectation of Iλ(0)I_{\lambda}(0) as

𝔼[Iλ(0)]=λpd(x)dx<\mathbb{E}[I_{\lambda}(0)]=\lambda p\int_{\mathbb{R}^{d}}\ell(x)\,\mathrm{d}x<\infty

(see cf. [26] for more general conditions for the almost sure convergence of shot-noise fields). The centralized and scaled version of IλI_{\lambda} is then given by

Iλ~(z)=Iλ(z)𝔼[Iλ(z)]g(λ),zd,\widetilde{I_{\lambda}}(z)=\frac{I_{\lambda}(z)-\mathbb{E}[I_{\lambda}(z)]}{g(\lambda)},\quad z\in\mathbb{R}^{d}, (2)

where the function gg is suitably chosen, and we investigate its limit as λ\lambda\to\infty in the sense of convergence in finite dimensional distributions when Φλ\Phi_{\lambda} is a determinantal point process. As a preliminary, however, we first give the proofs for the case where Φλ\Phi_{\lambda} is a homogeneous Poisson point process.

2.1 Poisson shot-noise with finite second moment of amplitudes

Here, we assume that Φλ\Phi_{\lambda} is a homogeneous Poisson point process and the amplitude distribution FPF_{P} has the finite second moment. The result below is proved straightforwardly and is indeed introduced without proof in the Introduction of [1]. However, we prove it here not only for the completeness of the paper but also because the proof serves as the basis for showing the later results.

Proposition 1

Let Φλ=n=1δXn\Phi_{\lambda}=\sum_{n=1}^{\infty}\delta_{X_{n}} be a homogeneous Poisson point process with intensity λ(0,)\lambda\in(0,\infty). Suppose that 𝔼[P12]<\mathbb{E}[{P_{1}}^{2}]<\infty and let g(λ)=λ1/2g(\lambda)=\lambda^{1/2} in (2). Then, as λ\lambda\to\infty, {Iλ~(z)}zd\{\widetilde{I_{\lambda}}(z)\}_{z\in\mathbb{R}^{d}} converges in finite dimensional distributions to a Gaussian random field {N(z)}zd\{N(z)\}_{z\in\mathbb{R}^{d}} with covariance function;

𝖢𝗈𝗏[N(z1),N(z2)]\displaystyle\mathsf{Cov}[N(z_{1}),N(z_{2})] =𝔼[P12]d(z1x)(z2x)dx,z1,z2d.\displaystyle=\mathbb{E}[{P_{1}}^{2}]\int_{\mathbb{R}^{d}}\ell(z_{1}-x)\,\ell(z_{2}-x)\,\mathrm{d}x,\quad z_{1},z_{2}\in\mathbb{R}^{d}. (3)

Note that the covariance in (3) is finite since \ell is bounded and integrable with respect to the Lebesgue measure on d\mathbb{R}^{d}. Let bb_{\ell} and cc_{\ell} denote positive constants such that (x)[0,b]\ell(x)\in[0,b_{\ell}] for xdx\in\mathbb{R}^{d} and d(x)dx=c\int_{\mathbb{R}^{d}}\ell(x)\,\mathrm{d}x=c_{\ell}. Then, we have clearly

d(z1x)(z2x)dxbc<.\int_{\mathbb{R}^{d}}\ell(z_{1}-x)\,\ell(z_{2}-x)\,\mathrm{d}x\leq b_{\ell}\,c_{\ell}<\infty. (4)
  • Proof:

    Consider the finite dimensional Laplace transform of Iλ~\widetilde{I_{\lambda}} in (2) at 𝒛=(z1,,zm)d×m\bm{z}=(z_{1},\ldots,z_{m})\in\mathbb{R}^{d\times m} for mm\in\mathbb{N}; that is, for 𝒔=(s1,,sm)[0,)m\bm{s}^{\top}=(s_{1},\ldots,s_{m})\in[0,\infty)^{m}, (2) leads to

    Iλ~(𝒔,𝒛)\displaystyle\mathcal{L}_{\widetilde{I_{\lambda}}}(\bm{s},\bm{z}) :=𝔼[exp(j=1msjIλ~(zj))]\displaystyle:=\mathbb{E}\biggl{[}\exp\biggl{(}-\sum_{j=1}^{m}s_{j}\,\widetilde{I_{\lambda}}(z_{j})\biggr{)}\biggr{]}
    =𝔼[exp(1g(λ)j=1msjIλ(zj))]exp(1g(λ)j=1msj𝔼[Iλ(zj)]).\displaystyle=\mathbb{E}\biggl{[}\exp\biggl{(}-\frac{1}{g(\lambda)}\sum_{j=1}^{m}s_{j}\,I_{\lambda}(z_{j})\biggr{)}\biggr{]}\,\exp\biggl{(}\frac{1}{g(\lambda)}\sum_{j=1}^{m}s_{j}\,\mathbb{E}[I_{\lambda}(z_{j})]\biggr{)}. (5)

    Applying (1) and Campbell’s formula, we reduce the inside of the second exponential in the last expression above to

    1g(λ)j=1msj𝔼[Iλ(zj)]=λpg(λ)dξ𝒔,𝒛(x)dx,\frac{1}{g(\lambda)}\sum_{j=1}^{m}s_{j}\,\mathbb{E}[I_{\lambda}(z_{j})]=\frac{\lambda\,p}{g(\lambda)}\int_{\mathbb{R}^{d}}\xi_{\bm{s},\bm{z}}(x)\,\mathrm{d}x, (6)

    where ξ𝒔,𝒛(x):=j=1msj(zjx)\xi_{\bm{s},\bm{z}}(x):=\sum_{j=1}^{m}s_{j}\,\ell(z_{j}-x) and 𝔼[Pn]=p\mathbb{E}[P_{n}]=p is used. On the other hand, applying (1) and the probability generating functional of a Poisson point process (cf. [14, p. 25, Exercise 3.6]) to the first expectation in the last expression of (Proof:) leads to

    𝔼[exp(1g(λ)j=1msjIλ(zj))]\displaystyle\mathbb{E}\biggl{[}\exp\biggl{(}-\frac{1}{g(\lambda)}\sum_{j=1}^{m}s_{j}\,I_{\lambda}(z_{j})\biggr{)}\biggr{]} =𝔼[n=1P(ξ𝒔,𝒛(Xn)g(λ))]\displaystyle=\mathbb{E}\biggl{[}\prod_{n=1}^{\infty}\mathcal{L}_{P}\Bigl{(}\frac{\xi_{\bm{s},\bm{z}}(X_{n})}{g(\lambda)}\Bigr{)}\biggr{]}
    =exp(λd[1P(ξ𝒔,𝒛(x)g(λ))]dx),\displaystyle=\exp\biggl{(}-\lambda\int_{\mathbb{R}^{d}}\Bigl{[}1-\mathcal{L}_{P}\Bigl{(}\frac{\xi_{\bm{s},\bm{z}}(x)}{g(\lambda)}\Bigr{)}\Bigr{]}\,\mathrm{d}x\biggr{)}, (7)

    where P(s)=𝔼[esP1]\mathcal{L}_{P}(s)=\mathbb{E}[e^{-sP_{1}}] denotes the Laplace transform of P1P_{1}. Therefore, plugging (6) and (Proof:) into (Proof:), we have

    Iλ~(𝒔,𝒛)=exp(λd0ψ(ξ𝒔,𝒛(x)g(λ)t)dFP(t)dx),\mathcal{L}_{\widetilde{I_{\lambda}}}(\bm{s},\bm{z})=\exp\biggl{(}\lambda\int_{\mathbb{R}^{d}}\!\int_{0}^{\infty}\psi\Bigl{(}\frac{\xi_{\bm{s},\bm{z}}(x)}{g(\lambda)}\,t\Bigr{)}\,\mathrm{d}F_{P}(t)\,\mathrm{d}x\biggr{)}, (8)

    where ψ(u)=eu1+u\psi(u)=e^{-u}-1+u, and we used P(s)=0estdFP(t)\mathcal{L}_{P}(s)=\int_{0}^{\infty}e^{-st}\,\mathrm{d}F_{P}(t) and p=0tdFP(t)p=\int_{0}^{\infty}t\,\mathrm{d}F_{P}(t). We now take g(λ)=λ1/2g(\lambda)=\lambda^{1/2}. Then, since ψ(u)=u2/2+o(u2)\psi(u)=u^{2}/2+o(u^{2}) as u0u\downarrow 0, if the order of the limit and the integral is interchangeable (which is confirmed below), we obtain

    limλIλ~(𝒔,𝒛)\displaystyle\lim_{\lambda\to\infty}\mathcal{L}_{\widetilde{I_{\lambda}}}(\bm{s},\bm{z}) =exp(𝔼[P12]2d[ξ𝒔,𝒛(x)]2dx)\displaystyle=\exp\biggl{(}\frac{\mathbb{E}[{P_{1}}^{2}]}{2}\int_{\mathbb{R}^{d}}[\xi_{\bm{s},\bm{z}}(x)]^{2}\,\mathrm{d}x\biggr{)}
    =exp(𝔼[P12]2𝒔𝑳(𝒛)𝒔),\displaystyle=\exp\biggl{(}\frac{\mathbb{E}[{P_{1}}^{2}]}{2}\,\bm{s}^{\top}\bm{L}(\bm{z})\,\bm{s}\biggr{)}, (9)

    where the (j,k)(j,k)-element of matrix 𝑳(𝒛)=(L(zj,zk))j,k=1m\bm{L}(\bm{z})=\bigl{(}L(z_{j},z_{k})\bigr{)}_{j,k=1}^{m} is given by L(zj,zk)=d(zjx)(zkx)dxL(z_{j},z_{k})=\int_{\mathbb{R}^{d}}\ell(z_{j}-x)\,\ell(z_{k}-x)\,\mathrm{d}x and the assertion of the proposition holds. It remains to confirm the interchangeability of the order of the limit and the integral in (8) as λ\lambda\to\infty. Since ψ(u)[0,u2/2]\psi(u)\in[0,u^{2}/2], we have 0λψ(ξ𝒔,𝒛(x)t/λ1/2)[ξ𝒔,𝒛(x)t]2/20\leq\lambda\,\psi\bigl{(}\xi_{\bm{s},\bm{z}}(x)\,t/\lambda^{1/2}\bigr{)}\leq[\xi_{\bm{s},\bm{z}}(x)\,t]^{2}/2 and the integral of [ξ𝒔,𝒛(x)t]2/2[\xi_{\bm{s},\bm{z}}(x)\,t]^{2}/2 with respect to dFP(t)dx\mathrm{d}F_{P}(t)\,\mathrm{d}x is provided as the inside of the exponential in (Proof:), which is finite from (4). Hence, the dominated convergence theorem is applicable and the proof is completed.       

2.2 Poisson shot-noise with regularly varying amplitude distribution

Next, we assume that the tail FP¯(t)=1FP(t)\overline{F_{P}}(t)=1-F_{P}(t) of the amplitude distribution is regularly varying with index α-\alpha for α(1,2)\alpha\in(1,2); that is (cf. [4] or [22]),

limtFP¯(ct)FP¯(t)=cαfor some c>0.\lim_{t\to\infty}\frac{\overline{F_{P}}(ct)}{\overline{F_{P}}(t)}=c^{-\alpha}\quad\text{for some $c>0$.}

Note that 𝔼[Pi2]=\mathbb{E}[{P_{i}}^{2}]=\infty in this case. We use the following properties of regularly varying functions, where a(x)b(x)a(x)\sim b(x) as xx\to\infty stands for limxa(x)/b(x)=1\lim_{x\to\infty}a(x)/b(x)=1.

Proposition 2
  1. 1.

    (Cf. [4, p. 28, Theorem 1.5.12] or [22, p. 21]) Let ff be regularly varying with index γ>0\gamma>0. Then, there exists an asymptotic inverse gg of ff satisfying f(g(x))g(f(x))xf(g(x))\sim g(f(x))\sim x as xx\to\infty, where gg is asymptotically unique and regularly varying with index 1/γ1/\gamma.

  2. 2.

    (Representation Theorem; cf. [4, p. 12, Theorem 1.3.1] or [22, p. 2, Theorem 1.2] Function L0L_{0} is slowly varying if and only if there exists a positive constant a0a_{0} such that

    L0(x)=exp(η(x)+a0xϵ(t)tdt),xa0,L_{0}(x)=\exp\Bigl{(}\eta(x)+\int_{a_{0}}^{x}\frac{\epsilon(t)}{t}\,\mathrm{d}t\Bigr{)},\quad x\geq a_{0}, (10)

    where η(x)\eta(x) is bounded and converges to a constant as xx\to\infty, and ϵ(t)\epsilon(t) is bounded and converges to zero as tt\to\infty.

Using the properties above, we prove the following result, which is also new to the best of the knowledge of the authors.

Theorem 1

Let Φλ=i=1δXi\Phi_{\lambda}=\sum_{i=1}^{\infty}\delta_{X_{i}} be a homogeneous Poisson point process with intensity λ(0,)\lambda\in(0,\infty). Suppose that FP¯\overline{F_{P}} is regularly varying with index α-\alpha for α(1,2)\alpha\in(1,2) and let gg in (2) be an asymptotic inverse of 1/FP¯1/\overline{F_{P}} (so that gg is regularly varying with index 1/α1/\alpha). Then, as λ\lambda\to\infty, {Iλ~(y)}yd\{\widetilde{I_{\lambda}}(y)\}_{y\in\mathbb{R}^{d}} converges in finite dimensional distributions to an α\alpha-stable random field {S(z)}zd\{S(z)\}_{z\in\mathbb{R}^{d}} with finite dimensional Laplace transform;

S(𝒔,𝒛)\displaystyle\mathcal{L}_{S}(\bm{s},\bm{z}) :=𝔼[exp(j=1msjS(zj))]\displaystyle:=\mathbb{E}\biggl{[}\exp\biggl{(}-\sum_{j=1}^{m}s_{j}\,S(z_{j})\biggr{)}\biggr{]}
=exp(Γ(2α)α1d[j=1msj(zjx)]αdx),\displaystyle=\exp\biggl{(}\frac{\Gamma(2-\alpha)}{\alpha-1}\int_{\mathbb{R}^{d}}\biggl{[}\sum_{j=1}^{m}s_{j}\,\ell(z_{j}-x)\biggr{]}^{\alpha}\,\mathrm{d}x\biggr{)}, (11)

for mm\in\mathbb{N}, 𝐬=(s1,,sm)[0,)m\bm{s}^{\top}=(s_{1},\ldots,s_{m})\in[0,\infty)^{m} and 𝐳=(z1,,zm)d×m\bm{z}=(z_{1},\ldots,z_{m})\in\mathbb{R}^{d\times m}.

Remark 1

The integral in (1) is finite for any fixed 𝐬=(s1,,sm)[0,)m\bm{s}^{\top}=(s_{1},\ldots,s_{m})\in[0,\infty)^{m} and 𝐳=(z1,,zm)d×m\bm{z}=(z_{1},\ldots,z_{m})\in\mathbb{R}^{d\times m}. Indeed, since α>1\alpha>1, it is easy to show that

d[j=1msj(zjx)]αdxbα1c(j=1msj)α<,\int_{\mathbb{R}^{d}}\biggl{[}\sum_{j=1}^{m}s_{j}\,\ell(z_{j}-x)\biggr{]}^{\alpha}\,\mathrm{d}x\leq{b_{\ell}}^{\alpha-1}\,c_{\ell}\,\biggl{(}\sum_{j=1}^{m}s_{j}\biggr{)}^{\alpha}<\infty,

where bb_{\ell} and cc_{\ell} are the same as in (4). The last expression of (1) definitely implies that {S(z)}zd\{S(z)\}_{z\in\mathbb{R}^{d}} is an α\alpha-stable random field since each linear combination j=1msjS(zj)\sum_{j=1}^{m}s_{j}\,S(z_{j}) for mm\in\mathbb{N}, 𝐬=(s1,,sm)[0,)m\bm{s}^{\top}=(s_{1},\ldots,s_{m})\in[0,\infty)^{m} and 𝐳=(z1,,zm)d×m\bm{z}=(z_{1},\ldots,z_{m})\in\mathbb{R}^{d\times m} follows an α\alpha-stable distribution 𝒮α(σ𝐬,𝐳,1,0)\mathcal{S}_{\alpha}(\sigma_{\bm{s},\bm{z}},1,0) with

σ𝒔,𝒛=(Γ(2α)α1d[j=1msj(zjx)]αdxcosπα2)1/α\sigma_{\bm{s},\bm{z}}=\biggl{(}-\frac{\Gamma(2-\alpha)}{\alpha-1}\int_{\mathbb{R}^{d}}\biggl{[}\sum_{j=1}^{m}s_{j}\,\ell(z_{j}-x)\biggr{]}^{\alpha}\,\mathrm{d}x\,\cos\frac{\pi\alpha}{2}\biggr{)}^{1/\alpha}

(cf. [21, p. 15, Proposition 1.2.12 and pp. 112–113, Theorem 3.1.2]).

  • Proof:

    We start the proof of the theorem with (8) in the proof of Proposition 1, where we recall that ξ𝒔,𝒛(x)=j=1msj(zjx)\xi_{\bm{s},\bm{z}}(x)=\sum_{j=1}^{m}s_{j}\,\ell(z_{j}-x) and ψ(u)=eu1+u\psi(u)=e^{-u}-1+u. Applying integration by parts to the integral with respect to dFP(t)\mathrm{d}F_{P}(t) in (8), we have

    0ψ(ξ𝒔,𝒛(x)g(λ)t)dFP(t)\displaystyle\int_{0}^{\infty}\psi\biggl{(}\frac{\xi_{\bm{s},\bm{z}}(x)}{g(\lambda)}\,t\biggr{)}\,\mathrm{d}F_{P}(t) =ξ𝒔,𝒛(x)g(λ)0[1exp(ξ𝒔,𝒛(x)g(λ)t)]FP¯(t)dt\displaystyle=\frac{\xi_{\bm{s},\bm{z}}(x)}{g(\lambda)}\int_{0}^{\infty}\Bigl{[}1-\exp\Bigl{(}-\frac{\xi_{\bm{s},\bm{z}}(x)}{g(\lambda)}\,t\Bigr{)}\Bigr{]}\,\overline{F_{P}}(t)\,\mathrm{d}t
    =ξ𝒔,𝒛(x)0[1eξ𝒔,𝒛(x)u]FP¯(g(λ)u)du,\displaystyle=\xi_{\bm{s},\bm{z}}(x)\int_{0}^{\infty}\bigl{[}1-e^{-\xi_{\bm{s},\bm{z}}(x)\,u}\bigr{]}\,\overline{F_{P}}\bigl{(}g(\lambda)\,u\bigr{)}\,\mathrm{d}u,

    where the change of variables u=t/g(λ)u=t/g(\lambda) is applied in the second equality. Therefore, since λ1/FP¯(g(λ))\lambda\sim 1/\overline{F_{P}}(g(\lambda)) and FP¯(g(λ)u)/FP¯(g(λ))uα\overline{F_{P}}(g(\lambda)\,u)/\overline{F_{P}}(g(\lambda))\to u^{-\alpha} as λ\lambda\to\infty, if the order of the limit and the integral is interchangeable (which is confirmed below), the inside of the exponential in (8) yields

    λd0ψ(ξ𝒔,𝒛(x)g(λ)t)dFP(t)dx\displaystyle\lambda\int_{\mathbb{R}^{d}}\!\int_{0}^{\infty}\psi\Bigl{(}\frac{\xi_{\bm{s},\bm{z}}(x)}{g(\lambda)}\,t\Bigr{)}\,\mathrm{d}F_{P}(t)\,\mathrm{d}x
    dξ𝒔,𝒛(x)0[1eξ𝒔,𝒛(x)u]FP¯(g(λ)u)FP¯(g(λ))dudx\displaystyle\sim\int_{\mathbb{R}^{d}}\xi_{\bm{s},\bm{z}}(x)\int_{0}^{\infty}\bigl{[}1-e^{-\xi_{\bm{s},\bm{z}}(x)u}\bigr{]}\,\frac{\overline{F_{P}}(g(\lambda)\,u)}{\overline{F_{P}}(g(\lambda))}\,\mathrm{d}u\,\mathrm{d}x
    d[ξ𝒔,𝒛(x)]αdx0(1ev)vαdvas λ,\displaystyle\to\int_{\mathbb{R}^{d}}[\xi_{\bm{s},\bm{z}}(x)]^{\alpha}\,\mathrm{d}x\int_{0}^{\infty}(1-e^{-v})\,v^{-\alpha}\,\mathrm{d}v\quad\text{as $\lambda\to\infty$,} (12)

    where v=ξ𝒔,𝒛(x)uv=\xi_{\bm{s},\bm{z}}(x)\,u is used in the last expression. The last integral above is equal to Γ(2α)/(α1)\Gamma(2-\alpha)/(\alpha-1) and the last expression of (1) is obtained.

    It remains to show the interchangeability of the order of the limit and the integral in (Proof:), where the dominated convergence theorem can be applied if we can find an integrable bound on

    ξ(x)[1eξ(x)u]FP¯(gu)FP¯(g),\xi(x)\,[1-e^{-\xi(x)u}]\,\frac{\overline{F_{P}}(gu)}{\overline{F_{P}}(g)}, (13)

    with respect to dudx\mathrm{d}u\,\mathrm{d}x on [0,)×d[0,\infty)\times\mathbb{R}^{d} for a positive and integrable function ξ\xi on d\mathbb{R}^{d} and a sufficiently large g>0g>0. Since FP¯\overline{F_{P}} is regularly varying with index α-\alpha, we have FP¯(g)=gαL0(g)\overline{F_{P}}(g)=g^{-\alpha}\,L_{0}(g) with L0L_{0} of the form (10). We define constants η\eta^{*} and ϵ\epsilon^{*} as

    η=supxa0|η(x)|,ϵ=supta0|ϵ(t)|.\eta^{*}=\sup_{x\geq a_{0}}|\eta(x)|,\quad\epsilon^{*}=\sup_{t\geq a_{0}}|\epsilon(t)|.

    Note here that we can take a0a_{0} in (10) large enough such that ϵ<α1\epsilon^{*}<\alpha-1 since ϵ(t)0\epsilon(t)\to 0 as tt\to\infty. Then, for ga0g\geq a_{0} and u1u\geq 1, we have

    FP¯(gu)FP¯(g)uαexp(2η+ϵggudtt)=e2ηu(αϵ),\frac{\overline{F_{P}}(gu)}{\overline{F_{P}}(g)}\leq u^{-\alpha}\,\exp\Bigl{(}2\eta^{*}+\epsilon^{*}\int_{g}^{gu}\frac{\mathrm{d}t}{t}\Bigr{)}=e^{2\eta^{*}}u^{-(\alpha-\epsilon^{*})},

    and (13) is bounded by

    ξ(x)(b0 1(0,1)(u)+e2ηu(αϵ) 1[1,)(u)),\xi(x)\,\bigl{(}b_{0}\,\bm{1}_{(0,1)}(u)+e^{2\eta^{*}}u^{-(\alpha-\epsilon^{*})}\,\bm{1}_{[1,\infty)}(u)\bigr{)},

    where b0=supga0,u(0,1)FP¯(gu)/FP¯(g)b_{0}=\sup_{g\geq a_{0},u\in(0,1)}\overline{F_{P}}(gu)/\overline{F_{P}}(g). We know that ξ\xi is integrable on d\mathbb{R}^{d} and

    1u(αϵ)du=1α1ϵ,\int_{1}^{\infty}u^{-(\alpha-\epsilon^{*})}\,\mathrm{d}u=\frac{1}{\alpha-1-\epsilon^{*}},

    which completes the proof.       

3 Scaling limits of determinantal shot-noise fields

We now extend the results in the preceding section to the case where Φλ\Phi_{\lambda} is a stationary determinantal point process and show that the same scaling limits are derived. Let Φλ=n=1δXn\Phi_{\lambda}=\sum_{n=1}^{\infty}\delta_{X_{n}} be a stationary and isotropic determinantal point process on d\mathbb{R}^{d} with intensity λ(0,)\lambda\in(0,\infty) and let KλK_{\lambda}d×d\mathbb{R}^{d}\times\mathbb{R}^{d}\to\mathbb{C} denote the kernel of Φλ\Phi_{\lambda} with respect to the Lebesgue measure; that is, the nnth product density ρn\rho_{n}, nn\in\mathbb{N}, of Φλ\Phi_{\lambda} with respect to the Lebesgue measure is given by (cf. [11, 24])

ρn(x1,x2,,xn)=det(Kλ(xi,xj))i,j=1,2,,n,x1,x2,,xnd,\rho_{n}(x_{1},x_{2},\ldots,x_{n})=\det\bigl{(}K_{\lambda}(x_{i},x_{j})\bigr{)}_{i,j=1,2,\ldots,n},\quad x_{1},x_{2},\ldots,x_{n}\in\mathbb{R}^{d},

where det\det stands for determinant. We assume that (i) the kernel KλK_{\lambda} is continuous on d×d\mathbb{R}^{d}\times\mathbb{R}^{d} with Kλ(x,x)=ρ1(x)=λK_{\lambda}(x,x)=\rho_{1}(x)=\lambda for any xdx\in\mathbb{R}^{d}; (ii) KλK_{\lambda} is Hermitian; that is, Kλ(x,y)=Kλ(y,x)K_{\lambda}(x,y)=K_{\lambda}(y,x)^{*} for x,ydx,y\in\mathbb{R}^{d}, where ww^{*} denotes the complex conjugate of ww\in\mathbb{C}; and (iii) the integral operator 𝒦λ\mathcal{K}_{\lambda} on L2(d,dx)L^{2}(\mathbb{R}^{d},\mathrm{d}x) given by

𝒦λf(x)=dKλ(x,y)f(y)dy,fL2(d,dx),xd,\mathcal{K}_{\lambda}f(x)=\int_{\mathbb{R}^{d}}K_{\lambda}(x,y)\,f(y)\,\mathrm{d}y,\quad f\in L^{2}(\mathbb{R}^{d},\mathrm{d}x),\;x\in\mathbb{R}^{d},

has its spectrum in [0,1][0,1]. Note that the operator 𝒦λ\mathcal{K}_{\lambda} satisfying (i)–(iii) is locally of trace-class (cf. [19, p. 65, Lemma]), and that the determinantal point process Φλ\Phi_{\lambda} exists and is locally finite (cf. [11, p. 68, Theorem 4.5.5] or [24, Theorem 3]). Moreover, we assume that KλK_{\lambda} satisfies |Kλ(x,y)|2=|Kλ(0,yx)|2|K_{\lambda}(x,y)|^{2}=|K_{\lambda}(0,y-x)|^{2} which depends only on the distance xy\|x-y\| of x,ydx,y\in\mathbb{R}^{d}. The product densities ρn\rho_{n}, nn\in\mathbb{N}, are then motion-invariant (invariant to translations and rotations) and ρ2(0,x)=λ2|Kλ(0,x)|2\rho_{2}(0,x)=\lambda^{2}-|K_{\lambda}(0,x)|^{2} depends only on x\|x\| for xdx\in\mathbb{R}^{d}.

To develop the corresponding discussion to the case of a Poisson point process, we first give a preliminary lemma.

Lemma 1

Let Φλ\Phi_{\lambda} be the determinantal point process described above. Then, the finite dimensional Laplace transform of the shot-noise field IλI_{\lambda} in (1) has the following exponential expression;

Iλ(𝒔,𝒛)\displaystyle\mathcal{L}_{I_{\lambda}}(\bm{s},\bm{z}) :=𝔼[exp(j=1msjIλ(zj))]\displaystyle:=\mathbb{E}\biggl{[}\exp\biggl{(}-\sum_{j=1}^{m}s_{j}\,I_{\lambda}(z_{j})\biggr{)}\biggr{]}
=exp(n=11n𝖳𝗋(𝒦λ,ξn)),\displaystyle=\exp\biggl{(}-\sum_{n=1}^{\infty}\frac{1}{n}\,\mathsf{Tr}\bigl{(}{\mathcal{K}_{\lambda,\mathcal{L}\circ\xi}}^{n}\bigr{)}\biggr{)}, (14)

for mm\in\mathbb{N}, 𝐳=(z1,,zm)d×m\bm{z}=(z_{1},\ldots,z_{m})\in\mathbb{R}^{d\times m} and 𝐬=(s1,,sm)[0,)m\bm{s}^{\top}=(s_{1},\ldots,s_{m})\in[0,\infty)^{m}, where 𝖳𝗋\mathsf{Tr} stands for the trace of a linear operator and 𝒦λ,ξ\mathcal{K}_{\lambda,\mathcal{L}\circ\xi} denotes the integral operator given by the kernel;

Kλ,ξ(x,y)=1P(ξ𝒔,𝒛(x))Kλ(x,y)1P(ξ𝒔,𝒛(y)),x,yd,K_{\lambda,\mathcal{L}\circ\xi}(x,y)=\sqrt{1-\mathcal{L}_{P}\bigl{(}\xi_{\bm{s},\bm{z}}(x)\bigr{)}}\,K_{\lambda}(x,y)\,\sqrt{1-\mathcal{L}_{P}\bigl{(}\xi_{\bm{s},\bm{z}}(y)\bigr{)}},\quad x,y\in\mathbb{R}^{d}, (15)

with the Laplace transform P\mathcal{L}_{P} of P1P_{1} and ξ𝐬,𝐳(x)=j=1msj(zjx)\xi_{\bm{s},\bm{z}}(x)=\sum_{j=1}^{m}s_{j}\,\ell(z_{j}-x).

To prove the lemma, we use the following result in the literature.

Proposition 3 (Cf. [23, Theorem 1.2] and [15, Lemma 2 and Corollary 1])

Let Φ=n=1δXn\Phi=\sum_{n=1}^{\infty}\delta_{X_{n}} denote a determinantal point process on d\mathbb{R}^{d}, where the kernel KK with respect to the Lebesgue measure ensures the existence of Φ\Phi. Then, for any measurable function vvd[0,1]\mathbb{R}^{d}\to[0,1] such that f(x):=lnv(x)f(x):=-\ln v(x) satisfies (a) limxf(x)=0\lim_{\|x\|\to\infty}f(x)=0, (b) limrx>rK(x,x)f(x)dx=0\lim_{r\to\infty}\int_{\|x\|>r}K(x,x)\,f(x)\,\mathrm{d}x=0, and (c) dK(x,x)[1exp(f(x))]dx<\int_{\mathbb{R}^{d}}K(x,x)\bigl{[}1-\exp\bigl{(}-f(x)\bigr{)}\bigr{]}\,\mathrm{d}x<\infty, the probability generating functional of Φ\Phi is given by

𝔼[n=1v(Xn)]=𝖣𝖾𝗍(𝒦v),\mathbb{E}\biggl{[}\prod_{n=1}^{\infty}v(X_{n})\biggr{]}=\mathsf{Det}(\mathcal{I}-\mathcal{K}_{v}),

where 𝖣𝖾𝗍\mathsf{Det} stands for the Fredholm determinant, \mathcal{I} denotes the identity operator and 𝒦v\mathcal{K}_{v} is the integral operator given by the kernel Kv(x,y)=1v(x)K(x,y)1v(y)K_{v}(x,y)=\sqrt{1-v(x)}\,K(x,y)\sqrt{1-v(y)}, x,ydx,y\in\mathbb{R}^{d}.

The result in Proposition 3 is first presented in [23] in the form of Laplace functional for function f(x)=lnv(x)f(x)=-\ln v(x) such that ff has a compact support. It is then generalized in [15] to ff satisfying the conditions (a)–(c) in the proposition when d=2d=2, whereas this generalization is also available for d\mathbb{R}^{d}, d=2,3,d=2,3,\ldots.

  • Proof of Lemma 1:

    Similar to obtaining the first equality in (Proof:), we have

    Iλ(𝒔,𝒛)=𝔼[n=1P(ξ𝒔,𝒛(Xn))].\mathcal{L}_{I_{\lambda}}(\bm{s},\bm{z})=\mathbb{E}\biggl{[}\prod_{n=1}^{\infty}\mathcal{L}_{P}\bigl{(}\xi_{\bm{s},\bm{z}}(X_{n})\bigr{)}\biggr{]}. (16)

    To apply Proposition 3, we have to confirm that f(x)=lnP(ξ𝒔,𝒛(x))f(x)=-\ln\mathcal{L}_{P}\bigl{(}\xi_{\bm{s},\bm{z}}(x)\bigr{)} satisfies conditions (a)–(c) in it. For (a), recall that ξ𝒔,𝒛(x)=j=1msj(zjx)\xi_{\bm{s},\bm{z}}(x)=\sum_{j=1}^{m}s_{j}\,\ell(z_{j}-x) and P(ξ𝒔,𝒛(x))=𝔼[eξ𝒔,𝒛(x)P1]\mathcal{L}_{P}\bigl{(}\xi_{\bm{s},\bm{z}}(x)\bigr{)}=\mathbb{E}[e^{-\xi_{\bm{s},\bm{z}}(x)\,P_{1}}]. Since eξ𝒔,𝒛(x)P1(0,1]e^{-\xi_{\bm{s},\bm{z}}(x)\,P_{1}}\in(0,1] and eξ𝒔,𝒛(x)P11e^{-\xi_{\bm{s},\bm{z}}(x)\,P_{1}}\to 1 as x\|x\|\to\infty, the dominated convergence theorem leads to lnP(ξ𝒔,𝒛(x))0-\ln\mathcal{L}_{P}\bigl{(}\xi_{\bm{s},\bm{z}}(x)\bigr{)}\to 0 as x\|x\|\to\infty. Next, we confirm (b). Since Kλ(x,x)=λK_{\lambda}(x,x)=\lambda, it suffices to show that d[lnP(ξ𝒔,𝒛(x))]dx<\int_{\mathbb{R}^{d}}\bigl{[}-\ln\mathcal{L}_{P}\bigl{(}\xi_{\bm{s},\bm{z}}(x)\bigr{)}\bigr{]}\,\mathrm{d}x<\infty, which follows from the integrability of ξ𝒔,𝒛\xi_{\bm{s},\bm{z}} because lnP(ξ𝒔,𝒛(x))pξ𝒔,𝒛(x)-\ln\mathcal{L}_{P}\bigl{(}\xi_{\bm{s},\bm{z}}(x)\bigr{)}\leq p\,\xi_{\bm{s},\bm{z}}(x) by Jensen’s inequality. The condition (c) is confirmed by showing d[1P(ξ𝒔,𝒛(x))]dx<\int_{\mathbb{R}^{d}}\bigl{[}1-\mathcal{L}_{P}\bigl{(}\xi_{\bm{s},\bm{z}}(x)\bigr{)}\bigr{]}\,\mathrm{d}x<\infty. Integration by parts yields

    1P(ξ𝒔,𝒛(x))\displaystyle 1-\mathcal{L}_{P}\bigl{(}\xi_{\bm{s},\bm{z}}(x)\bigr{)} =0[1eξ𝒔,𝒛(x)t]dFP(t)\displaystyle=\int_{0}^{\infty}[1-e^{-\xi_{\bm{s},\bm{z}}(x)\,t}]\,\mathrm{d}F_{P}(t)
    =ξ𝒔,𝒛(x)0eξ𝒔,𝒛(x)tFP¯(t)dtpξ𝒔,𝒛(x),\displaystyle=\xi_{\bm{s},\bm{z}}(x)\int_{0}^{\infty}e^{-\xi_{\bm{s},\bm{z}}(x)\,t}\,\overline{F_{P}}(t)\,\mathrm{d}t\leq p\,\xi_{\bm{s},\bm{z}}(x),

    which is integrable. Therefore, we can apply Proposition 3 to (16) and obtain

    Iλ(𝒔,𝒛)=𝖣𝖾𝗍(𝒦λ,ξ).\mathcal{L}_{I_{\lambda}}(\bm{s},\bm{z})=\mathsf{Det}(\mathcal{I}-\mathcal{K}_{\lambda,\mathcal{L}\circ\xi}).

    By the condition (c) above, the operator 𝒦λ,ξ\mathcal{K}_{\lambda,\mathcal{L}\circ\xi} given by (15) is of trace-class. Moreover, its operator norm satisfies 𝒦λ,ξ𝗈𝗉<𝒦λ𝗈𝗉1\|\mathcal{K}_{\lambda,\mathcal{L}\circ\xi}\|_{\mathsf{op}}<\|\mathcal{K}_{\lambda}\|_{\mathsf{op}}\leq 1 since P(ξ𝒔,𝒛(x))\mathcal{L}_{P}\bigl{(}\xi_{\bm{s},\bm{z}}(x)\bigr{)} is strictly positive. Hence, the Fredholm determinant 𝖣𝖾𝗍(𝒦λ,ξ)\mathsf{Det}(\mathcal{I}-\mathcal{K}_{\lambda,\mathcal{L}\circ\xi}) has the exponential expression (1) (cf. [18, p. 331, Lemma 6]).       

3.1 Case of finite second moment of amplitudes

Here is the extension of Proposition 1 to the case of a determinantal point process, which we prove by applying a similar discussion to that in [6].

Theorem 2

Let Φλ\Phi_{\lambda} be the determinantal point process with intensity λ\lambda described above. Suppose that 𝔼[P12]<\mathbb{E}[{P_{1}}^{2}]<\infty and g(λ)=λ1/2g(\lambda)=\lambda^{1/2} in (2). In addition, we assume that

d|Kλ(0,x)|2dx=o(λ)as λ.\int_{\mathbb{R}^{d}}|K_{\lambda}(0,x)|^{2}\,\mathrm{d}x=o(\lambda)\quad\text{as $\lambda\to\infty$.} (17)

Then, as λ\lambda\to\infty, {Iλ~(z)}zd\{\widetilde{I_{\lambda}}(z)\}_{z\in\mathbb{R}^{d}} converges in finite dimensional distributions to the same Gaussian random field as in Proposition 1.

Remark 2

In general, it holds that d|Kλ(0,x)|2dxKλ(0,0)=λ\int_{\mathbb{R}^{d}}|K_{\lambda}(0,x)|^{2}\,\mathrm{d}x\leq K_{\lambda}(0,0)=\lambda (see [17, Lemma 3.3]). Since ρ2(0,x)=λ2|Kλ(0,x)|2\rho_{2}(0,x)=\lambda^{2}-|K_{\lambda}(0,x)|^{2}, condition (17) above requires that the negative correlation in Φλ\Phi_{\lambda} is weakening as λ\lambda\to\infty.

  • Proof:

    Similar to obtaining (Proof:), (6) and (Proof:), we have

    Iλ~(𝒔,𝒛)=𝔼[n=1P(ξ𝒔,𝒛(Xn)g(λ))]exp(λpg(λ)dξ𝒔,𝒛(x)dx),\mathcal{L}_{\widetilde{I_{\lambda}}}(\bm{s},\bm{z})=\mathbb{E}\biggl{[}\prod_{n=1}^{\infty}\mathcal{L}_{P}\Bigl{(}\frac{\xi_{\bm{s},\bm{z}}(X_{n})}{g(\lambda)}\Bigr{)}\biggr{]}\,\exp\biggl{(}\frac{\lambda\,p}{g(\lambda)}\int_{\mathbb{R}^{d}}\xi_{\bm{s},\bm{z}}(x)\,\mathrm{d}x\biggr{)}, (18)

    where we recall that ξ𝒔,𝒛(x)=j=1msj(zjx)\xi_{\bm{s},\bm{z}}(x)=\sum_{j=1}^{m}s_{j}\,\ell(z_{j}-x). By Lemma 1, the expectation on the right-hand side above is equal to

    𝔼[n=1P(ξ𝒔,𝒛(Xn)g(λ))]=exp(n=11n𝖳𝗋(𝒦λ,(ξ/g)n)),\mathbb{E}\biggl{[}\prod_{n=1}^{\infty}\mathcal{L}_{P}\biggl{(}\frac{\xi_{\bm{s},\bm{z}}(X_{n})}{g(\lambda)}\biggr{)}\biggr{]}=\exp\biggl{(}-\sum_{n=1}^{\infty}\frac{1}{n}\,\mathsf{Tr}\bigl{(}{\mathcal{K}_{\lambda,\mathcal{L}\circ(\xi/g)}}^{n}\bigr{)}\biggr{)}, (19)

    where 𝒦λ,(ξ/g)\mathcal{K}_{\lambda,\mathcal{L}\circ(\xi/g)} denotes the integral operator given by the kernel;

    Kλ,(ξ/g)(x,y)=1P(ξ𝒔,𝒛(x)g(λ))Kλ(x,y)1P(ξ𝒔,𝒛(y)g(λ)).K_{\lambda,\mathcal{L}\circ(\xi/g)}(x,y)=\sqrt{1-\mathcal{L}_{P}\Bigl{(}\frac{\xi_{\bm{s},\bm{z}}(x)}{g(\lambda)}\Bigr{)}}\,K_{\lambda}(x,y)\,\sqrt{1-\mathcal{L}_{P}\Bigl{(}\frac{\xi_{\bm{s},\bm{z}}(y)}{g(\lambda)}\Bigr{)}}.

    Note that the term of n=1n=1 inside the exponential in (19) is equal to

    𝖳𝗋(𝒦λ,(ξ/g))=λd[1P(ξ𝒔,𝒛(x)g(λ))]dx,\mathsf{Tr}\bigl{(}\mathcal{K}_{\lambda,\mathcal{L}\circ(\xi/g)}\bigr{)}=\lambda\int_{\mathbb{R}^{d}}\Bigl{[}1-\mathcal{L}_{P}\Bigl{(}\frac{\xi_{\bm{s},\bm{z}}(x)}{g(\lambda)}\Bigr{)}\Bigr{]}\,\mathrm{d}x,

    which is identical to (Proof:) in the case of a Poisson point process. Therefore, the proof is completed if we can show that

    n=21n𝖳𝗋(𝒦λ,(ξ/g)n)0as λ.\sum_{n=2}^{\infty}\frac{1}{n}\,\mathsf{Tr}\bigl{(}{\mathcal{K}_{\lambda,\mathcal{L}\circ(\xi/g)}}^{n}\bigr{)}\to 0\quad\text{as $\lambda\to\infty$.} (20)

    Note that it holds that 𝖳𝗋(|A|n)𝖳𝗋(|A|2)1/2𝖳𝗋(|A|n1)\mathsf{Tr}(|A|^{n})\leq\mathsf{Tr}(|A|^{2})^{1/2}\,\mathsf{Tr}(|A|^{n-1}) for a trace-class operator AA since 𝖳𝗋(|AB|)A𝗈𝗉𝖳𝗋(|B|)\mathsf{Tr}(|AB|)\leq\|A\|_{\mathsf{op}}\,\mathsf{Tr}(|B|) for a bounded operator AA and a trace-class operator BB (cf. [20, p. 218, Problem 28]) and that A𝗈𝗉𝖳𝗋(|A|2)1/2\|A\|_{\mathsf{op}}\leq\mathsf{Tr}(|A|^{2})^{1/2} for a Hilbert-Schmidt operator AA (cf. [20, p. 210, Theorem VI.22 (d) or p. 218, Problem 25]). Applying this to the left-hand side of (20) inductively, we have

    |n=21n𝖳𝗋(𝒦λ,(ξ/g)n)|\displaystyle\biggl{|}\sum_{n=2}^{\infty}\frac{1}{n}\,\mathsf{Tr}\bigl{(}{\mathcal{K}_{\lambda,\mathcal{L}\circ(\xi/g)}}^{n}\bigr{)}\biggr{|} n=21n𝖳𝗋(|𝒦λ,(ξ/g)|n)\displaystyle\leq\sum_{n=2}^{\infty}\frac{1}{n}\,\mathsf{Tr}\bigl{(}|\mathcal{K}_{\lambda,\mathcal{L}\circ(\xi/g)}|^{n}\bigr{)}
    n=11n(𝖳𝗋(|𝒦λ,(ξ/g)|2))n/2\displaystyle\leq\sum_{n=1}^{\infty}\frac{1}{n}\,\bigl{(}\mathsf{Tr}\bigl{(}|\mathcal{K}_{\lambda,\mathcal{L}\circ(\xi/g)}|^{2}\bigr{)}\bigr{)}^{n/2}
    =ln(1(𝖳𝗋(|𝒦λ,(ξ/g)|2))1/2),\displaystyle=-\ln\Bigl{(}1-\bigl{(}\mathsf{Tr}\bigl{(}|\mathcal{K}_{\lambda,\mathcal{L}\circ(\xi/g)}|^{2}\bigr{)}\bigr{)}^{1/2}\Bigr{)}, (21)

    where the last equality holds when (𝖳𝗋(|𝒦λ,(ξ/g)|2))1/2<1\bigl{(}\mathsf{Tr}\bigl{(}|\mathcal{K}_{\lambda,\mathcal{L}\circ(\xi/g)}|^{2}\bigr{)}\bigr{)}^{1/2}<1, which is ensured for sufficiently large λ\lambda as shown below. Since 1P(s)sp1-\mathcal{L}_{P}(s)\leq sp and |Kλ(x,y)|2=|Kλ(0,yx)|2|K_{\lambda}(x,y)|^{2}=|K_{\lambda}(0,y-x)|^{2},

    𝖳𝗋(|𝒦λ,(ξ/g)|2)\displaystyle\mathsf{Tr}\bigl{(}|\mathcal{K}_{\lambda,\mathcal{L}\circ(\xi/g)}|^{2}\bigr{)} =dd(1P(ξ𝒔,𝒛(x)g(λ)))|Kλ(x,y)|2(1P(ξ𝒔,𝒛(y)g(λ)))dxdy\displaystyle=\int_{\mathbb{R}^{d}}\!\int_{\mathbb{R}^{d}}\Bigl{(}1-\mathcal{L}_{P}\Bigl{(}\frac{\xi_{\bm{s},\bm{z}}(x)}{g(\lambda)}\Bigr{)}\Bigr{)}\,|K_{\lambda}(x,y)|^{2}\,\Bigl{(}1-\mathcal{L}_{P}\Bigl{(}\frac{\xi_{\bm{s},\bm{z}}(y)}{g(\lambda)}\Bigr{)}\Bigr{)}\,\mathrm{d}x\,\mathrm{d}y
    (pg(λ))2ddξ𝒔,𝒛(x)|Kλ(x,y)|2ξ𝒔,𝒛(y)dxdy\displaystyle\leq\Bigl{(}\frac{p}{g(\lambda)}\Bigr{)}^{2}\int_{\mathbb{R}^{d}}\!\int_{\mathbb{R}^{d}}\xi_{\bm{s},\bm{z}}(x)|K_{\lambda}(x,y)|^{2}\,\xi_{\bm{s},\bm{z}}(y)\,\mathrm{d}x\,\mathrm{d}y
    bc(j=1msj)2(pg(λ))2d|Kλ(0,y)|2dy,\displaystyle\leq b_{\ell}\,c_{\ell}\,\biggl{(}\sum_{j=1}^{m}s_{j}\biggr{)}^{2}\,\Bigl{(}\frac{p}{g(\lambda)}\Bigr{)}^{2}\int_{\mathbb{R}^{d}}|K_{\lambda}(0,y)|^{2}\,\mathrm{d}y, (22)

    where bb_{\ell} and cc_{\ell} are the same as in (4), and we use ξ𝒔,𝒛(y)=j=1msj(zjy)bj=1msj\xi_{\bm{s},\bm{z}}(y)=\sum_{j=1}^{m}s_{j}\,\ell(z_{j}-y)\leq b_{\ell}\sum_{j=1}^{m}s_{j} and dξ𝒔,𝒛(x)dx=cj=1msj\int_{\mathbb{R}^{d}}\xi_{\bm{s},\bm{z}}(x)\,\mathrm{d}x=c_{\ell}\sum_{j=1}^{m}s_{j} for fixed 𝒔[0,)m\bm{s}\in[0,\infty)^{m} and 𝒛d×m\bm{z}\in\mathbb{R}^{d\times m}. Hence, when g(λ)=λ1/2g(\lambda)=\lambda^{1/2}, (Proof:) and therefore (Proof:) go to 0 as λ\lambda\to\infty under assumption (17), which implies (20).       

3.2 Case of regularly varying amplitude distribution

Here is our final result in this work, which is the extension of Theorem 1 to the case of a determinantal point process.

Theorem 3

Let Φλ\Phi_{\lambda} be the determinantal point process with intensity λ\lambda described in the beginning of this section. Suppose that FP¯\overline{F_{P}} is regularly varying with index α-\alpha for α(1,2)\alpha\in(1,2) and let gg in (2) be an asymptotic inverse of 1/FP¯1/\overline{F_{P}}. Then, as λ\lambda\to\infty, {Iλ~(z)}zd\{\widetilde{I_{\lambda}}(z)\}_{z\in\mathbb{R}^{d}} converges in finite dimensional distributions to the same α\alpha-stable random field as in Theorem 1.

Note that no additional assumption (like (17)) is required in this case.

  • Proof:

    The proof is almost the same as that of Theorem 2. Only the difference is as follows. Now, g(λ)g(\lambda) is regularly varying with index 1/α1/\alpha and can be represented as g(λ)=λ1/αL0(λ)g(\lambda)=\lambda^{1/\alpha}\,L_{0}(\lambda) with a slowly varying function L0L_{0}. Therefore, in (Proof:), since d|Kλ(0,y)|2dyKλ(0,0)=λ\int_{\mathbb{R}^{d}}|K_{\lambda}(0,y)|^{2}\,\mathrm{d}y\leq K_{\lambda}(0,0)=\lambda by Remark 2,

    1g(λ)2d|Kλ(0,y)|2dyλ12/αL0(λ)2,\frac{1}{g(\lambda)^{2}}\int_{\mathbb{R}^{d}}|K_{\lambda}(0,y)|^{2}\,\mathrm{d}y\leq\frac{\lambda^{1-2/\alpha}}{L_{0}(\lambda)^{2}},

    which goes to 0 as λ\lambda\to\infty since α(1,2)\alpha\in(1,2).       

4 Conclusion

In this work, we have considered a shot-noise field defined on a stationary determinantal point process and have shown that its centralized and suitably scaled version converges in finite dimensional distributions to i) a Gaussian random field when the amplitudes have the finite second moment and ii) an α\alpha-stable random field when the amplitudes follow a regularly varying distribution with index α-\alpha for α(1,2)\alpha\in(1,2). Some extensions can be considered as future work. For example, as [10] considers a shot-noise field defined on a Brillinger mixing point process and shows the convergence in distribution at one position, our result may be extended to the case of a more general point process. Furthermore, as [10] and [12] use Berry-Esseen bound to discuss the rate of convergence, the rates of the convergences in Theorems 2 and 3 may be interesting challenges.

References

  • [1] F. Baccelli and A. Biswas. On scaling limits of power law shot-noise fields. Stochastic Models, 31:187–207, 2015.
  • [2] F. Baccelli and B. Błaszczyszyn. Stochastic geometry and wireless networks, Volume I: Theory/Volume II: Applications. Foundations and Trends in Networking, 3/4:249–449/1–312, 2009.
  • [3] F. Baccelli, B. Błaszczyszyn, and M. Karray. Random measures, point processes, and stochastic geometry, 2020. hal-02460214.
  • [4] N. H. Bingham, C. M. Goldie, and J. L. Teugels. Regular Variation. Cambridge University Press, Cambridge, 1987.
  • [5] C. A. N. Biscio and F. Lavancier. Brillinger mixing of determinantal point processes and statistical applications. Electronic Journal of Statistics, 10:582–607, 2016.
  • [6] J.-C. Breton, A. Clarenne, and R. Gobard. Macroscopic analysis of determinantal random balls. Bernoulli, 25:1568–1601, 2019.
  • [7] S. N. Chiu, D. Stoyan, W. S. Kendall, and J. Mecke. Stochastic Geometry and its Applications. Wiley, Chichester, 3rd edition, 2013.
  • [8] M. Haenggi and R. K. Ganti. Interference in large wireless networks. Foundations and Trends in Networking, 3:127–248, 2009.
  • [9] L. Heinrich. On the strong Brillinger-mixing property of α\alpha-determinantal point processes and some applications. Applications of Mathematics, 61:443–461, 2016.
  • [10] L. Heinrich and V. Schmidt. Normal convergence of multidimensional shot noise and rates of this convergence. Advances in Applied Probability, 17:709–730, 1985.
  • [11] J. B. Hough, M. Krishnapur, Y. Peres, and B. Virág. Zeros of Gaussian Analytic Functions and Determinantal Point Processes. American Mathematical Society, 2009.
  • [12] H. Inaltekin. Gaussian approximation for the wireless multi-access interference distribution. IEEE Transactions on Signal Processing, 60:6114–6120, 2012.
  • [13] I. Kaj, L. Leskelä, I. Norros, and V. Schmidt. Scaling limits for random fields with long-range dependence. The Annals of Probability, 35:528–550, 2007.
  • [14] G. Last and M. Penrose. Lectures on the Poisson Process. Cambridge University Press, Cambridge, 2017.
  • [15] Y. Li, F. Baccelli, H. S. Dhillon, and J. G. Andrews. Statistical modeling and probabilistic analysis of cellular networks with determinantal point processes. IEEE Transactions on Communications, 63:3405–3422, 2015.
  • [16] N. Miyoshi and T. Shirai. A cellular network model with Ginibre configured base stations. Advances in Applied Probability, 46:832–845, 2014.
  • [17] N. Miyoshi and T. Shirai. Tail asymptotics of signal-to-interference ratio distribution in spatial cellular network models. Probability and Mathematical Statistics, 37:431–453, 2017.
  • [18] M. Reed and B. Simon. Methods of Modern Mathematical Physics IV: Analysis of Operators. Academic Press, San Diego, 1978.
  • [19] M. Reed and B. Simon. Methods of Modern Mathematical Physics III: Scattering Theory. Academic Press, San Diego, 1979.
  • [20] M. Reed and B. Simon. Methods of Modern Mathematical Physics I: Functional Analysis—Revised and Enlarged Edition. Academic Press, San Diego, 1980.
  • [21] G. Samorodnitsky and M. S. Taqqu. Stable Non-Gaussian Random Processes: Stochastic Models with Infinite Variance. Chapman & Hall/CRC, Boca Raton, 1994.
  • [22] E. Seneta. Regularly Varying Functions. Springer, Berlin, 1976.
  • [23] T. Shirai and Y. Takahashi. Random point fields associated with certain Fredholm determinants I: Fermion, Poisson and Boson processes. Journal of Functional Analysis, 205:414–463, 2003.
  • [24] A. Soshnikov. Determinantal random point fields. Russian Mathematical Surveys, 55:923–975, 2000.
  • [25] G. L. Torrisi and E. Leonardi. Large deviations of the interference in the Ginibre network model. Stochastic Systems, 4:1–33, 2014.
  • [26] M. Westcott. On the existence of a generalized shot-noise process. In E. J. Williams, editor, Studies in Probability and Statistics: Papers in Honour of Edwin J. G. Pitman, pages 73–88. North-Holland, Amsterdam, 1976.