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

Local repulsion of planar Gaussian critical points

Safa Ladgham 1,2, Raphaël Lachieze-Rey1
Abstract.

We study the local repulsion between critical points of a stationary isotropic smooth planar Gaussian field. We show that the critical points can experience a soft repulsion which is maximal in the case of the random planar wave model, or a soft attraction of arbitrary high order. If the type of critical points is specified (extremum, saddle point), the points experience a hard local repulsion, that we quantify with the precise magnitude of the second factorial moment of the number of points in a small ball.

Key words: Gaussian random fields; Stationary random fields; Critical points; Kac-Rice formula; repulsive point process.

AMS Classification: 60G60- 60G15

1 Université Paris Cité, Laboratoire MAP5, UMR CNRS 8145, 45 Rue des Saints-Pères, 75006 Paris
indent 2 FSMP, [email protected]

1. Introduction

The main topic of this paper is a local analysis of the critical points of a smooth stationary planar Gaussian field. The study of critical points, their number as well as their positions, are important issues in various application areas such as sea waves modeling [7] , astronomy [15,3,14] or neuroimaging [18, 22,23,24]. In these situations, practitioners are particularly interested in the detection of peaks of the random field under study or in high level asymptotics of maximal points [8,22,23]. At the opposite of these Extremes Theory results, some situations require the topological study of excursion sets over moderate levels [2,9] or the location study of critical points (not only extremal ones) [17].

Repulsive point processes have known a surge of interest in the recent years, they are useful in a number of applications, such as sampling for quasi Monte-Carlo methods [6], data mining, texture synthesis in Image Analysis [13], training set selection in machine learning, or numerical integration, see for instance [12], or as coresets for subsampling large datasets [21]. Critical points of Gaussian fields could be an alternative to determinantal point processes, which are commonly used for their repulsion properties despite the difficult issue of their synthesis [10]. Several definitions exist to characterize the repulsion properties of a stationary point process. We will use the following informal definition of local repulsion: A stationary random set of points 𝒳2\mathcal{X}\subset\mathbb{R}^{2} is locally repulsive at the second order if, denoting by 𝒩ρ\mathcal{N}_{\rho} its number of points in a ball centred in 0 with radius ρ\rho, we have

𝖱𝒩:=limρ0𝔼(𝒩ρ(2))𝔼(𝒩ρ)2<1\displaystyle\mathsf{R}_{\mathcal{N}}:=\lim_{\rho\to 0}\frac{\mathbb{E}(\mathcal{N}_{\rho}^{(2)})}{\mathbb{E}(\mathcal{N}_{\rho})^{2}}<1 (1)

where for an integer n,n(2)=n(n1)n,n^{(2)}=n(n-1) is the second order factorial power. This definition is motivated by the heuristic computation where we consider x1x2x_{1}\neq x_{2} randomly sampled in 𝒳B1\mathcal{X}\cap B_{1} and

𝔼(𝒩ρ)\displaystyle\mathbb{E}(\mathcal{N}_{\rho}) =(x1Bρ)+remainder\displaystyle=\mathbb{P}(x_{1}\in B_{\rho})+\text{remainder}
𝔼(𝒩ρ(2))\displaystyle\mathbb{E}(\mathcal{N}_{\rho}^{(2)}) =(x2Bρ,x1Bρ)+remainder,\displaystyle=\mathbb{P}(x_{2}\in B_{\rho}\;,x_{1}\in B_{\rho})+\text{remainder},

where the remainder terms are hopefully negligible when ρ\rho is small. In other words, a point process is locally repulsive if the probability to find a point in a small ball diminishes if we know that there is already a point in this ball. The constant 𝖱𝒩\mathsf{R}_{\mathcal{N}} is called the (second order) local repulsion factor, it is a dimensionless parameter that is invariant under rescaling or rotation of the process 𝒳\mathcal{X}. It equals 11 if 𝒳\mathcal{X} is a homogenous Poisson process, which is universally considered non-interacting. We say that the point process is weakly locally repulsive (resp. attractive) if 𝖱𝒩(0,1)\mathsf{R}_{\mathcal{N}}\in(0,1) (resp. (1,)(1,\infty)), and strongly repulsive if 𝖱𝒩=0\mathsf{R}_{\mathcal{N}}=0.

We study the repulsion properties of the stationary process 𝒳c\mathcal{X}_{c} formed by critical points of a planar stationary isotropic Gaussian field ψ\psi. We show that, depending on the covariance function of the field, they form a weakly locally repulsive or a weakly locally attractive point process, and that the minimal repulsion factor is 𝖱𝒳c=183\mathsf{R}_{\mathcal{X}_{c}}=\frac{1}{8\sqrt{3}}, reached when ψ\psi is a Gaussian random wave model, which hence yields the most locally repulsive process of Gaussian critical points. There is on the other hand no maximal value for the limit. We also show that the subprocess formed by the local maxima of the field is strongly repulsive, as well as the subprocess formed by the saddle points, and give the precise magnitude of the ratio decay in the left hand member of (1).

Let us quote two recent articles that are concerned with a very similar question. The first one, which has been a source of inspiration, is [5]. In this paper, Belyaev, Cammarota and Wigman study the repulsion of the critical points for a particular Gaussian field, the Berry’s Planar Random Wave Model, whose spectral measure is uniformly spread on a circle centred in 0. They obtain the exact repulsion ratio for critical points and upper bounds for the repulsivity for specific types of critical points (saddle, extrema). Azais and Delmas [1] have studied the attraction or repulsion of critical points for general stationary Gaussian fields in any dimension. Using a different computation method, they get an upper bound for the second factorial moment which is compatible with the order of magnitude that we obtain. Their method is borrowed from techniques in random matrix theory, as suggested by Fyodorov [11]. Namely, an explicit expression for the joint density of GOE eigenvalues is exploited.

In order to quantify the repulsion of the critical points, we compute the second factorial moment using the Rice or Kac-Rice formulas (see [2] or [4] for details), as the vast majority of works concerned with counting the zeros or critical points of a random field. We get the asymptotics as the ball radius tends to 0 by performing a fine asymptotic analysis on the conditional expectations that are involved in the Kac-Rice formulas.

The paper is organized as follows: In Section 2, we present the Gaussian fields, which are the probabilistic object of our study, and the basic tools we will use for their study. In Section 3, we derive the Kac-Rice formula, in a context adapted to our framework. The purpose of section 4 is to compute the expectation of the number of critical points and also the number of extrema, minima, maxima and saddle (see Proposition 3). In Section 5, we study the second factorial moment and discuss the repulsion properties of the critical points.

2. Assumptions and tools

The main actors of this article are centered random Gaussian functions ψ:2\psi:\mathbb{R}^{2}\to\mathbb{R} whose law is invariant under translations, and whose realisations are smooth. Formally it means that for x1,,xn2x_{1},\dots,x_{n}\in\mathbb{R}^{2}, (ψ(x1),,ψ(xn))(\psi(x_{1}),\dots,\psi(x_{n})) is a centered Gaussian vector which law is invariant under translation of the xix_{i}’s (and rotations if isotropy is further assumed), and that the sample paths {ψ(x);x2}\{\psi(x);x\in\mathbb{R}^{2}\} are a.s. of class 𝒞2\mathcal{C}^{2} (or more). See [2] for a rigourous and detailed exposition of Gaussian fields. Such a field is characterised by its reduced covariance function Γ\Gamma

𝔼[ψ(z)ψ(w)]:=Γ(zw)\mathbb{E}[\psi(z)\psi(w)]:=\Gamma(z-w)

for some Γ:2\Gamma:\mathbb{R}^{2}\to\mathbb{R}, and if the field is furthermore assumed to be isotropic (i.e. its law is invariant under rotations)

Γ(zw)=σ(|zw|2)\displaystyle\Gamma(z-w)=\sigma(|z-w|^{2}) (2)

for some σ:+\sigma:\mathbb{R}_{+}\to\mathbb{R}, where |x||x| denotes the Euclidean norm of x2x\in\mathbb{R}^{2}.

We denote by ψ(z)\nabla\psi(z) the gradient of ψ\psi at z2z\in\mathbb{R}^{2}, by Hψ(z)H_{\psi}(z) the Hessian matrix evaluated at zz, when these quantities are well defined. For a smooth random field ψ\psi, the set of critical points is denoted by

𝒳c=𝒳c(ψ):={x2:ψ(c)=0},\displaystyle\mathcal{X}_{c}=\mathcal{X}_{c}(\psi):=\{x\in\mathbb{R}^{2}:\nabla\psi(c)=0\},

and the number of critical points in a small disc BρB_{\rho} of radius ρ>0\rho>0 is defined by

𝒩ρc(ψ):=#𝒳cBρ.\mathcal{N}^{c}_{\rho}(\psi):=\#\mathcal{X}_{c}\cap B_{\rho}.

When there is no ambiguity about the random field ψ\psi, we simply write 𝒩ρc\mathcal{N}_{\rho}^{c} instead of 𝒩ρc(ψ)\mathcal{N}_{\rho}^{c}(\psi). Similarly, we denote by resp. 𝒩ρs(ψ),𝒩ρe(ψ),𝒩ρmax(ψ),𝒩ρmin(ψ)\mathcal{N}_{\rho}^{s}(\psi),\mathcal{N}_{\rho}^{e}(\psi),\mathcal{N}_{\rho}^{max}(\psi),\mathcal{N}_{\rho}^{min}(\psi) the number of resp. saddles, extrema, maxima and minima, critical points characterised by the signs of the Hessian eigenvalues.

As will be explained at Section 5, to perform a second order local analysis of the repulsion of ψ\psi’s critical points, we must assume fourth order differentiation of ψ\psi, and for technical reasons we further assume that the fourth order derivative is α\alpha-Hölder for some α>0\alpha>0, we call this property 𝒞4+α\mathcal{C}^{4+\alpha} regularity. It is implied by σ\sigma being of class 𝒞8+β\mathcal{C}^{8+\beta} for some β>2α\beta>2\alpha, see Proposition 1 below. In this case, the Hölder constant is a random variable with Gaussian tail (see below).

Assumption 2.1.

Assume that ψ\psi is a non-constant stationary Gaussian field on 2\mathbb{R}^{2} and its reduced covariance Γ\Gamma is of class 𝒞4+β\mathcal{C}^{4+\beta} for some β>0.\beta>0.

This assumption implies the 𝒞4+α\mathcal{C}^{4+\alpha} regularity of ψ\psi by applying the proposition below to ψ\psi’s 4th order derivatives.

Proposition 1.

Let φ\varphi be a stationary Gaussian field 2\mathbb{R}^{2}\to\mathbb{R}, with reduced covariance function γ:2\gamma:\mathbb{R}^{2}\to\mathbb{R}. Then if for some C,β>0C,\beta>0, for δ>0\delta>0 sufficiently small

|γ(x)γ(0)|C|x|β,|x|δ,\displaystyle|\gamma(x)-\gamma(0)|\leqslant C|x|^{\beta},|x|\leqslant\delta,

then for 0<ε<β/20<\varepsilon<\beta/2 there is a random variable UεU_{\varepsilon} with Gaussian tail such that for all x,yB1x,y\in B_{1},

|φ(x)φ(y)|Uε|xy|β/2ε.\displaystyle|\varphi(x)-\varphi(y)|\leqslant U_{\varepsilon}|x-y|^{\beta/2-\varepsilon}.
Proof.

It follows from the classical result from Landau and Shepp [2, (2.1.4)] that for a centred Gaussian field ff a.s. bounded on a Euclidean compact TT, there is c>0c>0 such that for large enough uu,

(suptT|f(t)|u)2exp(cu2).\displaystyle\mathbb{P}(\sup_{t\in T}|f(t)|\geqslant u)\leqslant 2\exp(-cu^{2}).

We wish to apply this result to T=B1×B1T=B_{1}\times B_{1} and

f(x,y)=|xy|α(φ(x)φ(y)),(x,y)T.\displaystyle f(x,y)=|x-y|^{-\alpha}(\varphi(x)-\varphi(y)),(x,y)\in T.

Let α=β/2ε.\alpha=\beta/2-\varepsilon. The fact that ff is bounded is the consequence of the fact that φ\varphi’s path are locally α\alpha-Holder for α<β/2\alpha<\beta/2, see for instance [20, Corollary 4.8].∎

Definition 1.

Say that some random variables X,YX,Y satisfy X=O(Y)X=O_{\mathbb{P}(Y)} if XUYX\leqslant UY where UU is a random variable with a Gaussian tail, i.e.

(|U|>t)cexp(ct2),t0\displaystyle\mathbb{P}(|U|>t)\leqslant c\exp(-c^{\prime}t^{2}),t\geqslant 0

for some c<,c>0.c<\infty,c^{\prime}>0.

Proposition 1 hence implies that if a stationary field ψ\psi’s reduced covariance Γ\Gamma is of class 𝒞k+η\mathcal{C}^{k+\eta}, then

kψ(t+h)=kψ(t)+𝒪(hη/2),td.\displaystyle\partial^{k}\psi(t+h)=\partial^{k}\psi(t)+\mathcal{O}_{\mathbb{P}}(h^{\eta/2}),t\in\mathbb{R}^{d}.

2.1. Dependency structure

Stationarity conveys strong constraints on the dependence structure between the field’s partial derivatives at a given point. Let us recall formula [2, (5.5.4)-(5.5.5)]: if Γ\Gamma is 𝒞k+η\mathcal{C}^{k+\eta} differentiable for some k,η>0k\in\mathbb{N},\eta>0, for natural integers α,β,γ,δ\alpha,\beta,\gamma,\delta such that α+βk,γ+δk\alpha+\beta\leqslant k,\gamma+\delta\leqslant k,

𝔼(1α2βψ(t)1γ2δψ(s))=α+β+γ+δt1αt2βs1γs2δΓ(ts),s,t2.\displaystyle\mathbb{E}\left(\partial_{1}^{\alpha}\partial_{2}^{\beta}\psi(t)\cdot\partial_{1}^{\gamma}\partial_{2}^{\delta}\psi(s)\right)=\frac{\partial^{\alpha+\beta+\gamma+\delta}}{\partial{t_{1}}^{\alpha}\partial{t_{2}}^{\beta}\partial{s_{1}}^{\gamma}\partial{s_{2}}^{\delta}}\Gamma(t-s),s,t\in\mathbb{R}^{2}.

In particular if s=ts=t we have the spectral representation

𝔼(1α2βψ(t)1γ2δψ(t))=(1)γ+δα+β+γ+δΓ(0)t1αt2βt1γt2δ=mα+γ,β+δ where ma,b:=(1)aıa+b2λ1aλ2bF(dλ),t2\displaystyle\mathbb{E}\left(\partial_{1}^{\alpha}\partial_{2}^{\beta}\psi(t)\cdot\partial_{1}^{\gamma}\partial_{2}^{\delta}\psi(t)\right)=(-1)^{\gamma+\delta}\frac{\partial^{\alpha+\beta+\gamma+\delta}\Gamma(0)}{{\partial{{{t}_{1}^{\alpha}}}}{\partial{t_{2}^{\beta}}}{\partial{t_{1}^{\gamma}}}\;{\partial{t_{2}^{\delta}}}}=m_{\alpha+\gamma,\beta+\delta}\text{\rm{ where }}m_{a,b}:=(-1)^{a}\imath^{a+b}\int_{\mathbb{R}^{2}}\lambda_{1}^{a}\lambda_{2}^{b}F(d\lambda),t\in\mathbb{R}^{2} (3)

where the symmetric spectral measure FF is uniquely defined by

Γ(t)=2exp(ıλt)F(dλ),t2.\displaystyle\Gamma(t)=\int_{\mathbb{R}^{2}}\exp(-\imath\lambda\cdot t)F(d\lambda),t\in\mathbb{R}^{2}. (4)

Let us state important consequences of (3), and in particular of the fact that, due to the symmetry of FF, the integral vanishes if aa or bb is an odd number. For this reason, (1)a=(1)b(-1)^{a}=(-1)^{b} when the integral does not vanish, and ma,bm_{a,b} is symmetric in aa and b.b.

Remark 1.

For all t2,ψ(t)t\in\mathbb{R}^{2},\psi(t) and jψ(t)\partial_{j}\psi(t) are independent for j=1,2j=1,2, hence 1ψ\partial_{1}\psi and 2ψ\partial_{2}\psi are independent, and furthermore for any two natural integers k,lk,l which difference is odd, any partial derivatives of orders kk and ll

i1,,ikψ(0) and j1,,jlψ(0) are independent.\displaystyle\partial_{i_{1},\dots,i_{k}}\psi(0)\text{\rm{ and }}\partial_{j_{1},\dots,j_{l}}\psi(0)\text{\rm{ are independent.}}

Non-independence and technical difficulties will mainly emerge from dependence between even degrees of differentiation of the field, such as ψ(t)\psi(t) and 11ψ(t)\partial_{11}\psi(t), or 11ψ(t)\partial_{11}\psi(t) and 22ψ(t)\partial_{22}\psi(t), or between the values of the field at different locations, say ψ(s)\psi(s) and ψ(t),st.\psi(t),s\neq t. A case we must discard is that of constant ψ\psi, i.e. ψ(t)=U\psi(t)=U for some Gaussian variable UU, and this is what we call a trivial Gaussian field.

Also, Cauchy-Schwarz inequality yields that for α,β,γ,δ\alpha,\beta,\gamma,\delta\in\mathbb{N}

|mα+γ,β+δ|2m2α,2βm2γ,2δ,\displaystyle|m_{\alpha+\gamma,\beta+\delta}|^{2}\leqslant m_{2\alpha,2\beta}m_{2\gamma,2\delta},

and there is equality only if λ1αλ2β\lambda_{1}^{\alpha}\lambda_{2}^{\beta} is proportionnal to λ1γλ2δ\lambda_{1}^{\gamma}\lambda_{2}^{\delta} dFdF-a.s. In the isotropic case (i.e. FF is invariant under spatial rotations), unless α=γ,β=δ\alpha=\gamma,\beta=\delta it can only happen if dFdF is the Dirac mass in 0, i.e.

mα+γ,β+δ2<m2α,2βm2γ,2δ,αγ or βδ if ψ is non-trivial isotropic.\displaystyle m_{\alpha+\gamma,\beta+\delta}^{2}<m_{2\alpha,2\beta}m_{2\gamma,2\delta},\alpha\neq\gamma\text{\rm{ or }}\beta\neq\delta\text{\rm{ if }}\psi\text{\rm{ is non-trivial isotropic}}. (5)
Proposition 2.

Let ψ\psi be an isotropic Gaussian field 2\mathbb{R}^{2}\rightarrow\mathbb{R} that satisfies Assumption 2.1 with covariance under the form (2). We indicate the first derivatives of σ\sigma at point 0{0}\in\mathbb{R} by σ(0)=η0,\sigma^{\prime}(0)=\eta_{0}, σ′′(0)=μ0,\sigma^{\prime\prime}(0)=\mu_{0}, σ(3)(0)=ν0,\sigma^{(3)}(0)=\nu_{0}, σ4(0)=υ\sigma^{{4}}(0)=\upsilon. Then

Var(iψ(0))\displaystyle\mathrm{Var}(\partial_{i}\psi(0)) =2η0=m2,0>0,\displaystyle=-2\eta_{0}=m_{2,0}>0, (6)
Var(12ψ(0))\displaystyle\mathrm{Var}(\partial_{12}\psi(0)) =22μ0=m2,2>0,\displaystyle=2^{2}\mu_{0}=m_{2,2}>0,
Var(iiψ(0))\displaystyle\mathrm{Var}(\partial_{ii}\psi(0)) =322μ0=m4,0>0,i=1,2,\displaystyle=3\cdot 2^{2}\mu_{0}=m_{4,0}>0,i=1,2,
Var(iiiψ(0))\displaystyle\mathrm{Var}(\partial_{iii}\psi(0)) =1523ν0=m6,0>0,i=1,2.\displaystyle=-15\cdot 2^{3}\nu_{0}=m_{6,0}>0,i=1,2.

The two last equalities illustrate the fact that isotropy and polar change of coordinates yield other relations between the ma,bm_{a,b} of the form

ma,b=αa,bma+b,0\displaystyle m_{a,b}=\alpha_{a,b}m_{a+b,0}

where the coefficients αa,b\alpha_{a,b} don’t depend on FF.

Example 1.

Let J0J_{0} be the Bessel function of the first order

J0(x)=12π02πeixcos(θ)𝑑θ,x.\displaystyle J_{0}(x)=\frac{1}{2\pi}\int_{0}^{2\pi}e^{-ix\cos(\theta)}d\theta,x\in\mathbb{R}.

For k>0k>0 let ψ\psi be the Gaussian random wave with parameter kk, i.e. the isotropic stationary Gaussian field with reduced covariance function

Γ(z)=J0(k|z|).\displaystyle\Gamma(z)=J_{0}(k|z|).

As is apparent from (4), this is the centered Gaussian field whose spectral measure is the uniform law on the centred circle with radius kk. It is important as it is the unique (in law) stationary Gaussian field for which

11ψ+22ψ+k2ψ=0 a.s.\displaystyle\partial_{11}\psi+\partial_{22}\psi+k^{2}\psi=0\text{\rm{ a.s. }}

up to a multiplicative constant. See for instance [5, 16, 19] and references therein for recent works about diverse aspects of planar random wave models. As proved at Section 6.2, it is the only non-trivial stationary isotropic stationary field satisfying a linear partial differential equation of order three or less. As critical points are not modified by adding a constant, we also consider shifted Gaussian random waves (SGRW), of the form τU+σψ\tau U+\sigma\psi, where τ0,σ>0,ψ\tau\geqslant 0,\sigma>0,\psi is a GRW and UU is an independent centered standard Gaussian variable. The spectral measure of a SGRW is the sum of a uniform measure on a circle of 2\mathbb{R}^{2} centred in 0 and a finite mass in {0}.\{0\}.

3. The Kac-Rice formula

The Kac-Rice formula gives a description of the factorial moments of the zeros of a random field. Let us give a formula adapted to counting the critical points of a certain type. The following result can be proved by combining the proofs of Theorems 6.3 and 6.4 from [4], see also [1, Appendix A].

Theorem 3.1.

Let ψ\psi isotropic satisfying Assumption 2.1. Let k{1,2}k\in\{1,2\}, B1,B2B_{1},B_{2} some open subsets of d\mathbb{R}^{d},

𝒩ρBi={tB(0,ρ):ψ(t)=0,Hψ(t)Bi}.\displaystyle\mathcal{N}_{\rho}^{B_{i}}=\{t\in B(0,\rho):\nabla\psi(t)=0,H_{\psi}(t)\in B_{i}\}.

Then for ρ\rho sufficiently small

𝔼[𝒩ρB1]=ρK1B1(t)dt,\displaystyle\mathbb{E}[\mathcal{N}_{\rho}^{B_{1}}]=\int_{\mathcal{B}_{\rho}}\mathrm{K}^{B_{1}}_{1}({t})\,\mathrm{d}{t},
𝔼[𝒩ρB1(𝒩ρB21)]=ρ2K2B1,B2(t1,t2)dt,\displaystyle\mathbb{E}[\mathcal{N}_{\rho}^{B_{1}}(\mathcal{N}_{\rho}^{B_{2}}-1)]=\int_{\mathcal{B}_{\rho}^{2}}\mathrm{K}^{B_{1},B_{2}}_{2}(t_{1},t_{2})\,\mathrm{d}{t}, (7)

where we have the kk-point correlation function :

K1B1(t)=\displaystyle\mathrm{K}_{1}^{B_{1}}(t)= ϕψ(t)(0)𝔼[|detHψ(t)|  1B1(Hψ(t))|ψ(t)=0],\displaystyle\phi_{\nabla\psi(t)}({0})\;\mathbb{E}\left[|\det H_{\psi}({t})|\;\;\mathbf{1}_{B_{1}}(H_{\psi}({t}))\;\Big{|}\;\nabla\psi(t)=0\right],
K2B1,B2(t1,t2)=\displaystyle\mathrm{K}_{2}^{B_{1},B_{2}}(t_{1},t_{2})= ϕ(ψ(t1),ψ(t2))(0,0)𝔼[i=12|detHψ(ti)|  1Bi(Hψ(ti))|ψ(t1)=ψ(t2)=0],\displaystyle\phi_{(\nabla\psi(t_{1}),\nabla\psi(t_{2}))}({0},{0})\;\mathbb{E}\left[\prod_{i=1}^{2}|\det H_{\psi}({t}_{i})|\;\;\mathbf{1}_{B_{i}}(H_{\psi}({t}_{i}))\;\Big{|}\;\nabla\psi(t_{1})=\nabla\psi(t_{2})=0\right],

where ϕV\phi_{V} is the density probability function of a Gaussian vector VV.

We are specifically interested in a finite class of sets BiB_{i}, namely

Bc=\displaystyle B_{c}= d() the class of d×d square matrices,\displaystyle\mathcal{M}_{d}(\mathbb{R})\text{ the class of $d\times d$ square matrices},
Bext=\displaystyle B_{ext}= det1((0,))\displaystyle\det^{-1}((0,\infty))
Bs=\displaystyle B_{s}= det1((,0)),\displaystyle\det^{-1}((-\infty,0)),
Bmin=\displaystyle B_{min}= {H definite positive},\displaystyle\{H\text{ definite positive}\},
Bmax=\displaystyle B_{max}= {H definite negative}.\displaystyle\{H\text{ definite negative}\}.

In this case, the exponent in 𝒩\mathcal{N} or KiK_{i} is replaced by the subscript of BB, e.g.

𝒩ρc=𝒩ρBc,K2s,s=K2Bs,Bs,etc\displaystyle\mathcal{N}_{\rho}^{c}=\mathcal{N}_{\rho}^{B_{c}},K_{2}^{s,s}=K_{2}^{B_{s},B_{s}},etc...
Proof.

With Z(t)=ψ(t),Yt=Hψ(t),g(t,H)=𝟏{det(H)B}Z(t)=\nabla\psi(t),Y^{t}=H_{\psi}(t),g(t,H)=\mathbf{1}_{\{\det(H)\in B\}}, we have

𝒩ρB=t:Z(t)=0g(t,Yt).\displaystyle\mathcal{N}_{\rho}^{B}=\sum_{t:Z(t)=0}g(t,Y^{t}).

Let us show that hypothesis (iii’) of [4, Th.6.3] is satisfied, that is for ρ\rho small enough and t,sρt,s\in\mathcal{B}_{\rho}, the law of (ψ(t),ψ(s))(\nabla\psi(t),\nabla\psi(s)) is non-degenerated. Let us expand

𝔼(iψ(s)jψ(t))=2sitjσ(|st|2)={2σ(|st|2)4(siti)2σ′′(|st|2) if i=j4(siti)(sjtj)σ′′(|st|2) if ij.\displaystyle\mathbb{E}(\partial_{i}\psi(s)\partial_{j}\psi(t))=\frac{\partial^{2}}{\partial_{s_{i}}\partial_{t_{j}}}\sigma(|s-t|^{2})=\begin{cases}-2\sigma^{\prime}(|s-t|^{2})-4(s_{i}-t_{i})^{2}\sigma^{\prime\prime}(|s-t|^{2})$ if $i=j\\ -4(s_{i}-t_{i})(s_{j}-t_{j})\sigma^{\prime\prime}(|s-t|^{2})$ if $i\neq j.\end{cases}

By isotropy it suffices to evaluate it in t=(r,0),s=(r,0)t=(r,0),s=(-r,0) for r0r\geqslant 0. Let us write the 4×44\times 4 covariance matrix in function of ηr=σ(4r2),μr=σ′′(4r2)\eta_{r}=\sigma^{\prime}(4r^{2}),\mu_{r}=\sigma^{\prime\prime}(4r^{2})

Σ=2(η0IηrI+2μrArηrI+2μrArη0I)\displaystyle\Sigma=-2\left(\begin{array}[]{cc}\eta_{0}I&\eta_{r}I+2\mu_{r}A_{r}\\ \eta_{r}I+2\mu_{r}A_{r}&\eta_{0}I\end{array}\right) (10)

where

Ar=(4r2000).\displaystyle A_{r}=\left(\begin{array}[]{cc}4r^{2}&0\\ 0&0\end{array}\right).

Hence the block determinant is

16det(η02I(ηrI+2μrAr)2)\displaystyle 16\det(\eta_{0}^{2}I-(\eta_{r}I+2\mu_{r}A_{r})^{2}) =16det((η02ηr2)I4μrηrAr4μr2Ar2)\displaystyle=16\det((\eta_{0}^{2}-\eta_{r}^{2})I-4\mu_{r}\eta_{r}A_{r}-4\mu_{r}^{2}A_{r}^{2})
=16(η02ηr2)((η02ηr2)16μrηrr264μr2r4).\displaystyle=16(\eta_{0}^{2}-\eta_{r}^{2})((\eta_{0}^{2}-\eta_{r}^{2})-16\mu_{r}\eta_{r}r^{2}-64\mu_{r}^{2}r^{4}).

This is equivalent to

168η0(μ0r2)(8η0(μ0r2)+𝒪(r2)16μ0η0r2+𝒪(r2))128η0μ0r2(24η0μ0r2)=3210μ02η02r4,\displaystyle 16\cdot 8\eta_{0}(-\mu_{0}r^{2})(8\eta_{0}(-\mu_{0}r^{2})+\mathcal{O}_{\mathbb{P}}(r^{2})-16\mu_{0}\eta_{0}r^{2}+\mathcal{O}_{\mathbb{P}}(r^{2}))\sim-128\eta_{0}\mu_{0}r^{2}(-24\eta_{0}\mu_{0}r^{2})=3\cdot 2^{10}\mu_{0}^{2}\eta_{0}^{2}r^{4},

where we have μ0η00\mu_{0}\eta_{0}\neq 0 in virtue of (6). Hence the determinant is non zero for r0r\neq 0 sufficiently small. Then the modification of the proof of Theorem 6.3 following the proof of Theorem 6.4 of [4] yields the result, see Appendix A in [1].

It yields in particular

ϕ(ψ(t1),ψ(t2))(0,0)=1det(Σ)(2π)2=127π23|μ0η0|r2(1+or0(1)).\displaystyle\phi_{(\nabla\psi(t_{1}),\nabla\psi(t_{2}))}({0},{0})=\frac{1}{\sqrt{\det(\Sigma)}(2\pi)^{2}}=\frac{1}{2^{7}\pi^{2}\sqrt{3}|\mu_{0}\eta_{0}|r^{2}}(1+o_{r\to 0}(1)). (11)

4. First order

In this section, we are interested in the computation of the expectated number of critical points in a Borel set B2.B\subset\mathbb{R}^{2}.

Proposition 3.

Let ψ={ψ(z):z2}\psi=\{\psi(z):z\in\mathbb{R}^{2}\} be a non-trivial isotropic stationary Gaussian field 2\mathbb{R}^{2}\rightarrow\mathbb{R} which is a.s. of class 𝒞2\mathcal{C}^{2} and let σ\sigma be defined by (2). Let

λc=43πσ′′(0)(σ(0)).\displaystyle\lambda_{c}=\frac{4}{\sqrt{3}\pi}\frac{\sigma^{\prime\prime}(0)}{(-\sigma^{\prime}(0))}.

In virtue of Proposition 2, λc(0,).\lambda_{c}\in(0,\infty). Then, for every ρ>0\rho>0, we have

𝔼[𝒩ρc]=λc|Bρ|,\displaystyle\mathbb{E}[\mathcal{N}_{\rho}^{c}]=\lambda_{c}|B_{\rho}|, (12)
𝔼[𝒩ρe]=𝔼[𝒩ρs]=12𝔼[𝒩ρc],\displaystyle\mathbb{E}[\mathcal{N}_{\rho}^{e}]=\mathbb{E}[\mathcal{N}_{\rho}^{s}]=\frac{1}{2}\mathbb{E}[\mathcal{N}_{\rho}^{c}], (13)
𝔼[𝒩ρmin]=14𝔼[𝒩ρc].\displaystyle\mathbb{E}[\mathcal{N}_{\rho}^{min}]=\frac{1}{4}\mathbb{E}[\mathcal{N}_{\rho}^{c}].

A sufficient condition for ψ\psi being of class 𝒞2\mathcal{C}^{2} is that σ\sigma is of classe 𝒞4+β\mathcal{C}^{4+\beta} for some β>0\beta>0, see Proposition 1.

Remark 2.

By stationarity, λc\lambda_{c} is the intensity of 𝒳c(ψ)\mathcal{X}_{c}(\psi), i.e. the mean number of critical points per unit volume.

Proof.

According to Theorem 3.1, we must simply evaluate

K1(z)=ϕψ(z)(0,0)𝔼[detHψ(z)|ψ(z)=0].\mathrm{K}_{1}(z)=\phi_{\nabla\psi(z)}(0,0)\;\;\mathbb{E}\left[\;\mid\det H_{\psi}(z)\mid\big{|}\nabla\psi(z)=0\;\right].

The stationarity of ψ\psi implies that K1(z)K_{1}(z) is independent of zz, see formula (7). So, we get

𝔼[𝒩ρc]=|Bρ|K1(0).\mathbb{E}[\mathcal{N}_{\rho}^{c}]=|B_{\rho}|\mathrm{K}_{1}(0)\mbox{}. (14)

Using the matrix Σ\Sigma with r=0r=0 in (10), we immediately obtain the probability density function of (two-dimensional vector) ψ(z)\nabla\psi(z) evaluated at point (0,0)(0,0):

ϕψ(z)(0,0)=12π14η02=14π|η0|,\phi_{\nabla\psi(z)}(0,0)=\frac{1}{2\pi}\frac{1}{\sqrt{4\eta_{0}^{2}}}=\frac{1}{4\pi|\eta_{0}|}, (15)

where η0=σ(0).\eta_{0}=\sigma^{\prime}(0). From this point until the end of the proof we will use the method of the article [5]. Since the first and the second derivatives of ψ(z)\psi(z) are independent at every fixed point z2,z\in\mathbb{R}^{2}, then:

𝔼[|detHψ(z)||ψ(z)=0]=𝔼[|detHψ(z)|]=𝔼[|detHψ(z)|]=𝔼[|11ψ(0)22ψ(0)122ψ(0)|].\mathbb{E}\left[\left|\;\det H_{\psi}(z)\right|\;\big{|}\nabla\psi(z)=0\right]=\mathbb{E}\left[\;\left|\det H_{\psi}(z)\right|\;\right]=\mathbb{E}\left[\;\left|\det H_{\psi}(z)\right|\;\right]=\mathbb{E}\left[\;\left|\partial_{11}\psi(0)\partial_{22}\psi(0)-\partial^{2}_{12}\psi(0)\right|\;\right]. (16)

To evaluate (16), we consider the transformation W1=11ψ(0)W_{1}=\partial_{11}\psi(0), W2=12ψ(0)W_{2}=\partial_{12}\psi(0), W3=11ψ(0)+22ψ(0)W_{3}=\partial_{11}\psi(0)+\partial_{22}\psi(0) and we write 𝔼[|11ψ(0)22ψ(0)122ψ(0)|]\mathbb{E}\left[\;\left|\partial_{11}\psi(0)\partial_{22}\psi(0)-\partial^{2}_{12}\psi(0)\right|\;\right] in terms of a conditional expectation as follows:

𝔼[|11ψ(0)22ψ(0)122ψ(0)|]\displaystyle\mathbb{E}\left[\;\left|\partial_{11}\psi(0)\partial_{22}\psi(0)-\partial^{2}_{12}\psi(0)\right|\;\right] =𝔼[|W1W3W12W22|]=𝔼[𝔼[|W1W3W12W22||W3]],\displaystyle=\mathbb{E}\left[\;\left|W_{1}W_{3}-W^{2}_{1}-W^{2}_{2}\right|\;\right]=\mathbb{E}\left[\;\mathbb{E}\left[\;|W_{1}W_{3}-W^{2}_{1}-W^{2}_{2}|\;\big{|}W_{3}\;\right]\right], (17)

where W=(W1,W2,W3)W=(W_{1},W_{2},W_{3}) is a centered Gaussian vector field with covariance matrix D.D.
Use Proposition 2 and Remark 1, we have D=(12μ0016μ004μ0016μ0032μ0).D=\begin{pmatrix}12\mu_{0}&0&16\mu_{0}\\ 0&4\mu_{0}&0\\ 16\mu_{0}&0&32\mu_{0}\end{pmatrix}.
The conditional distribution of (W1,W2)|W3(W_{1},W_{2})\big{|}W_{3} is Gaussian with covariance matrix Σ(W1,W2)|W3=(4μ0004μ0)\Sigma_{(W_{1},W_{2})|W_{3}}=\begin{pmatrix}4\mu_{0}&0\\ 0&4\mu_{0}\end{pmatrix} and expectation 𝔼[(W1,W2)|W3=t]=(t20)\mathbb{E}[\;(W_{1},W_{2})\;\big{|}W_{3}=t]=\begin{pmatrix}\frac{t}{2}\\ \\ 0\end{pmatrix} for t.t\in\mathbb{R}.
The conditioned Gaussian vector (W1,W2)|W3=t(W_{1},W_{2})|W_{3}=t is distributed as (2μ0Z1+t2,2μ0Z2)(2\sqrt{\mu_{0}}Z_{1}+\frac{t}{2},2\sqrt{\mu_{0}}\;Z_{2}) where Z1,Z_{1}, Z2Z_{2} are two independent standard Gaussian random variables, hence we have

𝔼[|W1W3W12W22||W3=t]\displaystyle\mathbb{E}\left[\;\left|W_{1}W_{3}-W^{2}_{1}-W^{2}_{2}\right|\;\big{|}\;W_{3}=t\;\right] =𝔼[W1tW12W22|W3=t]]\displaystyle=\mathbb{E}[\;\mid W_{1}t-W^{2}_{1}-W^{2}_{2}\mid|W_{3}=t\;]]
=𝔼[|(2μ0Z1+t/2)t(2μ0Z1+t/2)24μ0Z22|]\displaystyle=\mathbb{E}[\;|(2\sqrt{\mu_{0}}Z_{1}+t/2)t-(2\sqrt{\mu_{0}}Z_{1}+t/2)^{2}-4\mu_{0}\;Z^{2}_{2}|\;]
=𝔼[|4μ0Z124μ0Z22+t24|]\displaystyle=\mathbb{E}[\;|-4\mu_{0}Z_{1}^{2}-4\mu_{0}Z_{2}^{2}+\frac{t^{2}}{4}|\;]
=4μ0𝔼[|X+t216μ0|],\displaystyle=4\mu_{0}\;\mathbb{E}\left[\;\left|-X+\frac{t^{2}}{16\mu_{0}}\right|\right],

where XX is a χ\chi-square random variable with density fX(x)=12ex2,x>0.f_{X}(x)=\frac{1}{2}e^{-\frac{x}{2}},x>0.
So

𝔼[|X+t216μ0|]\displaystyle\mathbb{E}[\;|-X+\frac{t^{2}}{16\mu_{0}}|\;] =120t216μ0(t216μ0x)ex2𝑑x+12t216μ0+(t216μ0+x)ex2𝑑x\displaystyle=\frac{1}{2}\int_{0}^{\frac{t^{2}}{16\mu_{0}}}(\frac{t^{2}}{16\mu_{0}}-x)e^{-\frac{x}{2}}dx+\frac{1}{2}\int_{\frac{t^{2}}{16\mu_{0}}}^{+\infty}(-\frac{t^{2}}{16\mu_{0}}+x)e^{-\frac{x}{2}}dx
=2+4et232μ0+t216μ0,\displaystyle=-2+4e^{-\frac{t^{2}}{32\mu_{0}}}+\frac{t^{2}}{16\mu_{0}},

then

𝔼[|W1W3W12W22||W3=t]\displaystyle\mathbb{E}\left[\;\left|W_{1}W_{3}-W^{2}_{1}-W^{2}_{2}\right|\;\big{|}\;W_{3}=t\;\right] =4μ018πμ𝔼[|X+t216μ0|]et264μ0𝑑t\displaystyle=4\mu_{0}\;\frac{1}{8\sqrt{\pi\mu}}\int_{\mathbb{R}}\mathbb{E}\left[\;\left|-X+\frac{t^{2}}{16\mu_{0}}\right|\right]\;e^{-\frac{t^{2}}{64\mu_{0}}}dt
=μ02πet264μ(2+4et232μ0+t216μ0)𝑑t=16μ03.\displaystyle=\frac{\sqrt{\mu_{0}}}{2\sqrt{\pi}}\int_{\mathbb{R}}e^{-\frac{t^{2}}{64\mu}}\left(-2+4e^{-\frac{t^{2}}{32\mu_{0}}}+\frac{t^{2}}{16\mu_{0}}\right)\;dt=\frac{16\mu_{0}}{\sqrt{3}}. (18)

By combining Equations (14), (15),(16) and (Proof), we obtain Formula (12).

Now, we turn to the evaluation of the expected number of the extrema and saddle points. We have

𝒩ρe:=𝒩ρ(0,)\displaystyle\mathcal{N}^{e}_{\rho}:=\mathcal{N}_{\rho}^{(0,\infty)} =#{xBρ:ψ(x)=0,detHψ(z)>0}\displaystyle=#\{x\in B_{\rho}:\nabla\psi(x)=0,\det H_{\psi}(z)>0\}
𝒩ρs:=𝒩ρ(,0)\displaystyle\mathcal{N}^{s}_{\rho}:=\mathcal{N}_{\rho}^{(-\infty,0)} =#{xBρ:ψ(x)=0,detHψ(z)<0}.\displaystyle=#\{x\in B_{\rho}:\nabla\psi(x)=0,\det H_{\psi}(z)<0\}.

As previously, we apply the Kac-Rice formula from Section 3. We get:

𝔼[𝒩ρe]=BρK1e(z)𝑑zand𝔼[𝒩ρs]=BρK1s(z)𝑑z,\mathbb{E}[\mathcal{N}_{\rho}^{e}]=\int_{B_{\rho}}\mathrm{K}^{e}_{1}(z)dz\qquad\mbox{and}\qquad\mathbb{E}[\mathcal{N}_{\rho}^{s}]=\int_{B_{\rho}}\mathrm{K}^{s}_{1}(z)dz,

where

K1e(z)=ϕψ(z)(0,0)𝔼[|detHψ(z)| 1{detHψ(z)>0}|ψ(z)=0],\mathrm{K}_{1}^{e}(z)=\phi_{\nabla\psi(z)}(0,0)\;\;\mathbb{E}\left[\;|\det H_{\psi}(z)|\;\mathbf{1}_{\{\det H_{\psi}(z)>0\}}\;\big{|}\nabla\psi(z)=0\;\right],
K1s(z)=ϕψ(z)(0,0)𝔼[|detHψ(z)| 1{detHψ(z)<0}|ψ(z)=0].\mathrm{K}_{1}^{s}(z)=\phi_{\nabla\psi(z)}(0,0)\;\;\mathbb{E}\left[\;|{\det H}_{\psi}(z)|\;\mathbf{1}_{\{\det H_{\psi}(z)<0\}}\;\big{|}\nabla\psi(z)=0\;\right].

Since the first and the second derivatives of ψ(z)\psi(z) are independent at every fixed point z2,z\in\mathbb{R}^{2}, we obtain

𝔼[𝒩ρe]=πρ2ϕψ(z)(0,0)𝔼[|detHψ(z)| 1{detHψ(z)>0}],\mathbb{E}[\mathcal{N}_{\rho}^{e}]=\pi\rho^{2}\phi_{\nabla\psi(z)}(0,0)\;\;\mathbb{E}\left[\;|\det H_{\psi}(z)|\;\mathbf{1}_{\{\det H_{\psi}(z)>0\}}\right], (19)
𝔼[𝒩ρs]\displaystyle\mathbb{E}[\mathcal{N}_{\rho}^{s}] =πρ2ϕψ(z)(0,0)𝔼[|detHψ(z)| 1{detHψ(z)<0}].\displaystyle=\pi\rho^{2}\phi_{\nabla\psi(z)}(0,0)\;\;\mathbb{E}\left[\;|\det H_{\psi}(z)|\;\mathbf{1}_{\{\det H_{\psi}(z)<0\}}\right]. (20)

Using the same argument as in the case of critical points, we write

𝔼[|detHψ(z)|𝟏{detHψ(z)>0}]\displaystyle\mathbb{E}\left[\;\left|\det H_{\psi}(z)\right|\mathbf{1}_{\{detH_{\psi}(z)>0\}}\;\right] =𝔼[|11ψ(0)22ψ(0)122ψ(0)| 1{11ψ(0)22ψ(0)122>0}]\displaystyle=\mathbb{E}\left[|\partial_{11}\psi(0)\partial_{22}\psi(0)-{\partial}^{2}_{12}\psi(0)|\;\mathbf{1}_{\{\partial_{11}\psi(0)\partial_{22}\psi(0)-\partial^{2}_{12}>0\}}\right]
=4μ18πμ𝔼[|X+t216μ0|𝟏{X+t216μ0>0}]et264μ0𝑑t\displaystyle=4\mu\;\frac{1}{8\sqrt{\pi\mu}}\int_{\mathbb{R}}\mathbb{E}\left[\;\left|-X+\frac{t^{2}}{16\mu_{0}}\right|\mathbf{1}_{\{-X+\frac{t^{2}}{16\mu_{0}}>0\}}\right]\;e^{-\frac{t^{2}}{64\mu_{0}}}dt
=μ02π(2+2et232μ0+t216μ0)et264μ0𝑑t\displaystyle=\frac{\sqrt{\mu_{0}}}{2\sqrt{\pi}}\int_{\mathbb{R}}\left(-2+2e^{-\frac{t^{2}}{32\mu_{0}}}+\frac{t^{2}}{16\mu_{0}}\right)\;e^{-\frac{t^{2}}{64\mu_{0}}}dt
=8μ03,\displaystyle=\frac{8\mu_{0}}{\sqrt{3}}, (21)

and

𝔼[|detHψ(z)|𝟏{detHψ(z)<0}]\displaystyle\mathbb{E}\left[\;\left|\det H_{\psi}(z)\right|\mathbf{1}_{\{detH_{\psi}(z)<0\}}\;\right] =𝔼[|11ψ(0)22ψ(0)122| 1{11ψ(0)22ψ(0)122<0}]\displaystyle=\mathbb{E}\left[|\partial_{11}\psi(0)\partial_{22}\psi(0)-\partial^{2}_{12}|\;\mathbf{1}_{\{\partial_{11}\psi(0)\partial_{22}\psi(0)-\partial^{2}_{12}<0\}}\right]
=4μ18πμ0𝔼[|X+t216μ0|𝟏{X+t216μ0<0}]et264μ0𝑑t\displaystyle=4\mu\;\frac{1}{8\sqrt{\pi\mu_{0}}}\int_{\mathbb{R}}\mathbb{E}\left[\;\left|-X+\frac{t^{2}}{16\mu_{0}}\right|\mathbf{1}_{\{-X+\frac{t^{2}}{16\mu_{0}}<0\}}\right]\;e^{-\frac{t^{2}}{64\mu_{0}}}dt
=μ02π(2et232μ0)et264μ0𝑑t\displaystyle=\frac{\sqrt{\mu_{0}}}{2\sqrt{\pi}}\int_{\mathbb{R}}\left(2e^{-\frac{t^{2}}{32\mu_{0}}}\right)\;e^{-\frac{t^{2}}{64\mu_{0}}}dt
=8μ03.\displaystyle=\frac{{8\mu_{0}}}{\sqrt{3}}. (22)

By combining Equations (15), (19), (20), (Proof) and (Proof), we obtain Formula (13).
Finally, we turn to the calculation of the expectation of the number of minima and maxima in BρB_{\rho}.
We know that:

𝒩ρe=𝒩ρmin+𝒩ρmax\mathcal{N}_{\rho}^{e}=\mathcal{N}_{\rho}^{min}+\mathcal{N}_{\rho}^{max}

so 𝔼[𝒩ρe]=𝔼[𝒩ρmin]+𝔼[𝒩ρmax].\mathbb{E}[\mathcal{N}_{\rho}^{e}]=\mathbb{E}[\mathcal{N}_{\rho}^{min}]+\mathbb{E}[\mathcal{N}_{\rho}^{max}].

By symmetry of the Gaussian field ψ\psi, we have the following equality: 𝒩ρmax(ψ)=𝒩ρmax(ψ)=𝒩ρmin(ψ)\mathcal{N}_{\rho}^{max}(-\psi)\overset{\mathcal{L}}{=}\mathcal{N}_{\rho}^{max}(\psi)=\mathcal{N}_{\rho}^{min}(-\psi)  for  ψ=ψ,-\psi\overset{\mathcal{L}}{=}\psi,   therefore   𝔼[𝒩ρmin]=𝔼[𝒩ρmax].\mathbb{E}[\mathcal{N}_{\rho}^{min}]=\mathbb{E}[\mathcal{N}_{\rho}^{max}].

Finally, we obtain

𝔼[𝒩ρmin]=𝔼[𝒩ρmax]=12𝔼[𝒩ρe].\mathbb{E}[\mathcal{N}_{\rho}^{min}]=\mathbb{E}[\mathcal{N}_{\rho}^{max}]=\frac{1}{2}\mathbb{E}[\mathcal{N}_{\rho}^{e}].

5. Second order

In this section, we will study the asymptotic behaviour of the second factorial moment of 𝒩ρc\mathcal{N}^{c}_{\rho} when ρ\rho goes to zero. The following theorem is the main result of this paper. Given two quantitites αρ,βρ,\alpha_{\rho},\beta_{\rho}, write αρβρ\alpha_{\rho}\asymp\beta_{\rho} if for two constants 0<c<c<,0<c<c^{\prime}<\infty, we have cαρβρcαρ{c}\alpha_{\rho}\leqslant\beta_{\rho}\leqslant c^{\prime}\alpha_{\rho} for ρ\rho sufficiently small, and αρβρ\alpha_{\rho}\sim\beta_{\rho} if αρ/βρ1\alpha_{\rho}/\beta_{\rho}\to 1, with the convention 0/0=1.0/0=1.

Theorem 5.1.

Let ψ\psi be an isotropic Gaussian field 2\mathbb{R}^{2}\rightarrow\mathbb{R} that satisfies Assumption 2.1.The repulsion factor 𝖱c:=𝖱𝒳c\mathsf{R}_{c}:=\mathsf{R}_{\mathcal{X}_{c}} is given by

𝖱c=38(5σ′′′(0)σ(0)(σ′′(0))23).\displaystyle\mathsf{R}_{c}=\frac{\sqrt{3}}{8}\left(5\frac{\sigma^{\prime\prime\prime}(0)\sigma^{\prime}(0)}{(\sigma^{\prime\prime}(0))^{2}}-3\right).

As ρ0,\rho\rightarrow 0, we have the following asymptotic equivalent expression for the second factorial moment of the number of critical points

𝔼[𝒩ρc(𝒩ρc1)]𝖱cλc2|Bρ|2ρ4.\mathbb{E}[\mathcal{N}^{c}_{\rho}(\mathcal{N}^{c}_{\rho}-1)]\sim\mathsf{R}_{c}\lambda_{c}^{2}{|B_{\rho}|^{2}}\asymp\rho^{4}. (23)

Depending on the law of ψ\psi, 𝖱c\mathsf{R}_{c} can take any prescribed value in [183,)[\frac{1}{8\sqrt{3}},\infty), and 183\frac{1}{8\sqrt{3}} is the minimal possible value, it is reached iff ψ\psi is a shifted Gaussian random wave (Example 1).

For the numbers of extrema, saddles in a ball of radius ρ\rho, we have as ρ0\rho\to 0

𝔼[𝒩ρe(𝒩ρe1)]ρ7,\mathbb{E}[\mathcal{N}^{e}_{\rho}(\mathcal{N}^{e}_{\rho}-1)]\asymp\rho^{7}, (24)
𝔼[𝒩ρs(𝒩ρs1)]ρ7ln(ρ),\mathbb{E}[\mathcal{N}^{s}_{\rho}(\mathcal{N}^{s}_{\rho}-1)]\asymp\rho^{7}\ln(\rho), (25)
𝔼[𝒩ρe𝒩ρs]𝖱cλc2|Bρ|2.\mathbb{E}[\mathcal{N}_{\rho}^{e}\mathcal{N}_{\rho}^{s}]\sim\mathsf{R}_{c}\lambda_{c}^{2}|B_{\rho}|^{2}. (26)
Remark 3.

The repulsion factor terminology comes from λc|Bρ|𝔼(𝒩ρ)\lambda_{c}|B_{\rho}|\sim\mathbb{E}(\mathcal{N}_{\rho}) and by the heuristic explanation after (1).

Remark 4.

By truncating the expansion of the type (6.1) at a lower order, one could prove that Expression (23) is valid under the weaker assumption that Γ\Gamma is of class 𝒞6+β\mathcal{C}^{6+\beta} (and ψ\psi is of classe 𝒞3\mathcal{C}^{3}).

Example 2 (Bargmann Fock field).

Consider the Bargmann-Fock field with parameter kk, which is the stationary isotropic Gaussian field with reduced covariance function

σ(r)=exp(kr),r0.\displaystyle\sigma(r)=\exp(-kr),r\geqslant 0.

According to Proposition 3, we have for the first order

σ(0)=\displaystyle\sigma^{\prime}(0)= k\displaystyle-k
σ′′(0)=\displaystyle\sigma^{{}^{\prime\prime}}(0)= k2\displaystyle k^{2}
σ′′′(0)=\displaystyle\sigma^{{}^{\prime\prime\prime}}(0)= k3\displaystyle-k^{3}
𝔼[𝒩ρc]=\displaystyle\mathbb{E}[\mathcal{N}_{\rho}^{c}]= 43kρ2.\displaystyle\frac{4}{\sqrt{3}}k\;\rho^{2}.

Hence the attraction factor is

𝖱c=34<1,\displaystyle\mathsf{R}_{c}=\frac{\sqrt{3}}{4}<1,

which means that the process of critical points is locally weakly repulsive. It logically does not depend on the scaling factor k.k.

Example 3 (Gaussian random waves).

Consider the Gaussian random wave introduced at Example 1 by σ(x)=J0(kx),x0\sigma(x)=J_{0}(k\sqrt{x}),x\geqslant 0. We have :

σ(0)=\displaystyle\sigma^{\prime}(0)= k24\displaystyle-\frac{k^{2}}{4}
σ′′(0)=\displaystyle\sigma^{{}^{\prime\prime}}(0)= k425\displaystyle\frac{k^{4}}{2^{5}}
σ′′(0)=\displaystyle\sigma^{{}^{\prime\prime}}(0)= k6327\displaystyle-\frac{k^{6}}{3\cdot 2^{7}}
λc=\displaystyle\lambda_{c}= k223π i.e. 𝔼[𝒩ρc]=k223ρ2,\displaystyle\frac{k^{2}}{2\sqrt{3}\pi}\text{ i.e. }\mathbb{E}[\mathcal{N}_{\rho}^{c}]=\frac{k^{2}}{2\sqrt{3}}\;\rho^{2},

Hence the attraction factor takes the smallest possible value

𝖱c=183,\displaystyle\mathsf{R}_{c}=\frac{1}{8\sqrt{3}},

which means that the process of critical points is locally weakly repulsive. We retrieve the second factorial moment of Beliaev, Cammarota and Wigman [5]

𝔼(𝒩ρc(𝒩ρc1))k42633ρ4.\displaystyle\mathbb{E}(\mathcal{N}_{\rho}^{c}(\mathcal{N}_{\rho}^{c}-1))\sim\frac{k^{4}}{2^{6}3\sqrt{3}}\rho^{4}.
Example 4.

Consider the centered stationary Gaussian random field ψ\psi with spectral measure

F(dλ)=|λ|7𝟏{|λ|1}dλ.\displaystyle F(d\lambda)=|\lambda|^{-7}\mathbf{1}_{\{|\lambda|\geqslant 1\}}d\lambda.

One has by Proposition 2,|σ(0)|<,σ′′(0)<,σ′′′(0)=,|\sigma^{\prime}(0)|<\infty,\sigma^{\prime\prime}(0)<\infty,-\sigma^{\prime\prime\prime}(0)=\infty, hence 𝖱c=\mathsf{R}_{c}=\infty, but Theorem 5.1 does not apply precisely because FF’s higher moments are infinite, meaning that ψ\psi is not of class 𝒞3\mathcal{C}^{3}. Hence we consider Ft(dλ)=𝟏{|λ|t}F(dλ)(0t𝑑F)1F_{t}(d\lambda)=\mathbf{1}_{\{|\lambda|\leqslant t\}}F(d\lambda)(\int_{0}^{t}dF)^{-1} for t>1t>1. We have as tt\to\infty

η0ν0μ02c1tr6r7r𝑑rt.\displaystyle\frac{\eta_{0}\nu_{0}}{\mu_{0}^{2}}\sim c\int_{1}^{t}r^{6}r^{-7}rdr\asymp t.

It implies that the repulsion factor of FtF_{t} can reach arbitrarily high values. In particular, this parametric model provides processes of critical points that are weakly locally attractive.

5.1. Discussion and related litterature

The equivalence (23) generalises the results of [5], and shows that locally, the random planar wave model yields the more repulsive critical points. We also show that for a general process ψ\psi, the subprocesses formed by extrema and saddle points experience locally a strong repulsion with three more orders of magnitude for ρ\rho. It confirms the idea that close to a large portion of saddle points, there is an extremal point nearby, and conversely, but that the closest point of the same type (extrema or saddle) is typically much further away.

A current novelty is also to derive the precise asymptotic repulsion for the extrema process and the saddle process. Hence we are able to state that the ratio between the internal repulsion forces among extremal points and among saddle points tends to infinity as the radius of the observation ball goes to 0.

Azais and Delmas [1] derived upper bounds about such quantities in any dimension. In particular, their results are consistant with ours in the critic-critic, extrema-extrema and extrema-saddle cases.

6. Proofs

6.1. Conditioning

The proofs of all formulas of Theorem 5.1 are based on the Kac-Rice formula in Theorem 3.1, for instance if B=B=2B=B^{\prime}=\mathbb{R}^{2}, we have the second factorial moment of the number of critical points

𝔼[𝒩ρc(𝒩ρc1)]=Bρ×BρK2(z,w)𝑑z𝑑w,\mathbb{E}\left[\mathcal{N}_{\rho}^{c}(\mathcal{N}_{\rho}^{c}-1)\right]=\int\int_{B_{\rho}\times B_{\rho}}\mathrm{K}_{2}(z,w)\;dzdw, (27)

where K2\mathrm{K}_{2} is the 22-point correlation function : :

K2(z,w)\displaystyle\mathrm{K}_{2}(z,w) =ϕ(ψ(z),ψ(w))((0,0),(0,0))\displaystyle=\phi_{(\nabla\psi(z),\nabla\psi(w))}((0,0),(0,0)) (28)
×𝔼[|detHψ(z)||detHψ(w)||ψ(z)=ψ(w)=0].\displaystyle\times\mathbb{E}\left[\;\large|\det H_{\psi}(z)\large|\;\large|\det H_{\psi}(w)\large|\;\big{|}\;\nabla\psi(z)=\nabla\psi(w)=0\;\right].

Let us briefly introduce where the difficulty comes from and why higher order differentiability is required. For z,wz,w close from 0, if ψ(z)=ψ(w)=0\nabla\psi(z)=\nabla\psi(w)=0, then the second order derivatives are also small, and the determinant is dominated by third order differentials. When one imposes additional constraints on the determinant signs, it yields other cancellations within third order derivatives, requiring fourth order differentiability.

Thanks to the stationarity and isotropy of ψ,\psi, it suffices to compute K2(z,w)\mathrm{K}_{2}(z,w) for z=(r,0)z=(r,0) and w=(r,0)w=(-r,0) for all r>0.r>0. To evaluate 𝔼[𝒩ρc(𝒩ρc1)]\mathbb{E}\left[\mathcal{N}_{\rho}^{c}(\mathcal{N}_{\rho}^{c}-1)\right], the idea is to change the conditioning in K2(z,w).\mathrm{K}_{2}(z,w). To symmetrize the problem, we introduce some notations for rr near 0, r0r\neq 0, exploiting Proposition 1 and Definition 1,

{a)Δi(r):=12iψ(z)+12iψ(w) implies Δi(r)=iψ(0)+r22i11ψ(0)+𝒪(r4)b)Δi1(r):=12r(iψ(z)iψ(w)) implies Δi1(r)=i1ψ(0)+r26i111ψ(0)+𝒪(r4).,\left\{\begin{array}[]{lll}a)&\Delta_{i}(r):=\frac{1}{2}\partial_{i}\psi(z)+\frac{1}{2}\partial_{i}\psi(w)&\mbox{ implies }\Delta_{i}(r)=\partial_{i}\psi(0)+\frac{r^{2}}{2}\partial_{i11}\psi(0)+\mathcal{O}_{\mathbb{P}}(r^{4})\\ b)&\Delta_{i1}(r):=\frac{1}{2r}(\partial_{i}\psi(z)-\partial_{i}\psi(w))&\mbox{ implies }\Delta_{i1}(r)=\partial_{i1}\psi(0)+\frac{r^{2}}{6}\partial_{i111}\psi(0)+\mathcal{O}_{\mathbb{P}}(r^{4}).\\ \end{array}\right.,\\ (29)

The crucial point is that ψ(z)=ψ(w)=0\nabla\psi(z)=\nabla\psi(w)=0 is equivalent to Δi(r)=0,Δi1(r)=0,i=1,2.\Delta_{i}(r)=0,\Delta_{i1}(r)=0,i=1,2. Let us introduce

Yr=(Δ1(r),Δ2(r),Δ11(r),Δ12(r)),r>0Y_{r}=(\Delta_{1}(r),\Delta_{2}(r),\Delta_{11}(r),\Delta_{12}(r)),r>0

so that Yr=0Y_{r}=0 is equivalent to ψ(z)=ψ(w)=0\nabla\psi(z)=\nabla\psi(w)=0, and

Y0:=(1ψ(0),2ψ(0),11ψ(0),21ψ(0)).Y_{0}:=(\partial_{1}\psi(0),\partial_{2}\psi(0),\partial_{11}\psi(0),\partial_{21}\psi(0)).

We will see later that Y0Y_{0} is non-degenerate, hence YrY_{r} is also non-degenerate for rr small enough, by continuity of the covariance matrix.

We denote the conditionnal probability and expectation with respect to Yr=0Y_{r}=0 by

(r)()=(|Yr=0),𝔼(r)()=𝔼(|Yr=0),r0.\displaystyle\mathbb{P}^{(r)}(\cdot)=\mathbb{P}(\cdot\;|\;Y_{r}=0),\qquad\mathbb{E}^{(r)}(\cdot)=\mathbb{E}(\cdot\;|\;Y_{r}=0),r\geqslant 0.
Remark 5.

Let (X,Y)(X,Y) be a Gaussian vector with YY non-degenerate. If MM is non-singular matrix and if φ\varphi is a measurable function with polynomial bounds, then

E(φ(X)|Y=0)=E(φ(X)|MY=0).E(\varphi(X)|Y=0)=E(\varphi(X)|MY=0).

So, since obviously

ψ(z)=ψ(w)=0Yr=0,\nabla\psi(z)=\nabla\psi(w)=0\iff Y_{r}=0,\;

the 2-point correlation function K2(z,w){\mathrm{K}}_{2}(z,w) becomes :

K2(z,w)=ϕ(ψ(z),ψ(w))((0,0),(0,0))×𝔼(r)[|detHψ(z)||detHψ(w)|].\displaystyle\mathrm{K}_{2}(z,w)=\phi_{(\nabla\psi(z),\nabla\psi(w))}((0,0),(0,0))\times\mathbb{E}^{(r)}[\;|\det H_{\psi}(z)|\;|\det H_{\psi}(w)|]. (30)

Using (11) we can evaluate the density in 0, and the previous expression becomes

K2(z,w)=1(2π)2253(η0)r2𝔼(r)[|detHψψ(0)||detHψ(r)|](1+o(1)).\displaystyle K_{2}(z,w)=\frac{1}{(2\pi)^{2}2^{5}\sqrt{3}(-\eta_{0})r^{2}}\mathbb{E}^{(r)}[\;|\det H_{\psi}\psi(0)|\;|\det H_{\psi}(r)|](1+o(1)).

It remains to express the product of determinants under the conditioning in function of Δi(r)=Δi1(r)=0\Delta_{i}(r)=\Delta_{i1}(r)=0, this involves higher order derivatives (see(29)).

Lemma 1.

Assume ψ\psi satisfies Assumption 2.1 for some β>0\beta>0 and let 0<α<β/20<\alpha<\beta/2. For z=(r,0)z=(r,0) and w=(r,0)w=(-r,0), we have if Δi=Δ1i=0,i=1,2,\Delta_{i}=\Delta_{1i}=0,i=1,2,

detHψ(z)=\displaystyle\det H_{\psi}(z)= r(A1+rB0+𝒪(r1+α))\displaystyle r(A_{1}+rB_{0}+\mathcal{O}_{\mathbb{P}}(r^{1+\alpha})) (31)
detHψ(w)=\displaystyle\det H_{\psi}(w)= r(A1+rB0+𝒪(r1+α))\displaystyle r(-A_{1}+rB_{0}+\mathcal{O}_{\mathbb{P}}(r^{1+\alpha})) (32)
detHψ(z)detHψ(w)=\displaystyle\det H_{\psi}(z)\det H_{\psi}(w)= r2[A12+g(r)]\displaystyle r^{2}[-{A_{1}}^{2}+g(r)] (33)

where

{A1=22ψ(0)111ψ(0)B0=221ψ(0)111ψ(0)211ψ(0)2+1322ψ(0)1111ψ(0)g(r)=r2B02+𝒪(A1r1+α)+𝒪(r2+α).\left\{\begin{array}[]{lll}&A_{1}=\partial_{22}\psi(0)\partial_{111}\psi(0)\\ &B_{0}=\partial_{221}\psi(0)\;\partial_{111}\psi(0)-\partial_{211}\psi(0)^{2}+\frac{1}{3}\partial_{22}\psi(0)\partial_{1111}\psi(0)\\ &g(r)=r^{2}B_{0}^{2}+\mathcal{O}_{\mathbb{P}}(A_{1}r^{1+\alpha})+\mathcal{O}_{\mathbb{P}}(r^{2+\alpha})\end{array}\right.. (34)
Proof.

Define

{a)r23Δ1111(±r)=11ψ(±r,0)Δ11(±r)(±r111ψ(0)) implies Δ1111(±r)=1111ψ(0)+𝒪(rα)b)r23Δ2111(±r)=21ψ(±r,0)Δ21(±r)(±r211) implies Δ2111(±r)=2111ψ(0)+𝒪(rα)c)r22Δ2211(±r)=22ψ(±r,0)22ψ(0)(±r221) impliesΔ2211(±r)=2211ψ(0)+𝒪(rα).\left\{\begin{array}[]{lll}a)&\frac{r^{2}}{3}\Delta_{1111}(\pm r)=\partial_{11}\psi(\pm r,0)-\Delta_{11}(\pm r)-(\pm\;r\partial_{111}\psi(0))&\mbox{ implies }\Delta_{1111}(\pm r)=\partial_{1111}\psi(0)+\mathcal{O}_{\mathbb{P}}(r^{\alpha})\\ b)&\frac{r^{2}}{3}\Delta_{2111}(\pm r)=\partial_{21}\psi(\pm r,0)-\Delta_{21}(\pm r)-(\pm\;r\partial_{211})&\mbox{ implies }\Delta_{2111}(\pm r)=\partial_{2111}\psi(0)+\mathcal{O}_{\mathbb{P}}(r^{\alpha})\\ c)&\frac{r^{2}}{2}\Delta_{2211}(\pm r)=\partial_{22}\psi(\pm r,0)-\partial_{22}\psi(0)-(\pm\;r\partial_{221})&\mbox{ implies}\;\Delta_{2211}(\pm r)=\partial_{2211}\psi(0)+\mathcal{O}_{\mathbb{P}}(r^{\alpha}).\\ \end{array}\right.

We can explicitly write the expression of detHψ(z)detHψ(w):\det H_{\psi}(z)\det H_{\psi}(w):

detHψ(z)\displaystyle\det H_{\psi}(z) =11ψ(r,0)22ψ(r,0)(21ψ(r,0))2\displaystyle=\partial_{11}\psi(r,0)\partial_{22}\psi(r,0)-(\partial_{21}\psi(r,0))^{2}
=r(A1+rBr+r2Cr),\displaystyle=r(A_{1}+rB_{r}+r^{2}C_{r}), (35)

with

Br\displaystyle B_{r} =111ψ(0)221ψ(0)+13Δ1111(r)22ψ(0)211ψ(0)2=B0+𝒪(rα)\displaystyle=\partial_{111}\psi(0)\partial_{221}\psi(0)+\frac{1}{3}\Delta_{1111}(r)\;\partial_{22}\psi(0)-\partial_{211}\psi(0)^{2}=B_{0}+\mathcal{O}_{\mathbb{P}}(r^{\alpha})
Cr\displaystyle C_{r} =Δ1111(r)3221ψ(0)+Δ2211(r)2111ψ(0)23Δ2111(r)211ψ(0)=𝒪(1)\displaystyle=\frac{\Delta_{1111}(r)}{3}\;\partial_{221}\psi(0)+\frac{\Delta_{2211}(r)}{2}\partial_{111}\psi(0)-\frac{2}{3}\Delta_{2111}(r)\partial_{211}\psi(0)=\mathcal{O}_{\mathbb{P}}(1)

and

detHψ(w)\displaystyle\det H_{\psi}(w) =11ψ(r,0)22ψ(r,0)(21ψ(r,0))2\displaystyle=\partial_{11}\psi(-r,0)\partial_{22}\psi(-r,0)-(\partial_{21}\psi(-r,0))^{2}
=r(A1+rBr+r2Cr)\displaystyle{=r(-A_{1}+rB_{r}^{\prime}+r^{2}C_{r}^{\prime})} (36)

with

Br=\displaystyle B_{r^{\prime}}= 111ψ(0)221ψ(0)+13Δ1111(r)22ψ(0)211ψ(0)2=B0+𝒪(rα)\displaystyle\partial_{111}\psi(0)\partial_{221}\psi(0)+\frac{1}{3}\Delta_{1111}(-r)\;\partial_{22}\psi(0)-\partial_{211}\psi(0)^{2}=B_{0}+\mathcal{O}_{\mathbb{P}}(r^{\alpha})
Cr\displaystyle C_{r}^{\prime} =Δ1111(r)3221ψ(0)Δ2211(r)2111ψ(0)+23Δ2111(r)211ψ(0)=𝒪(1)\displaystyle=-\frac{\Delta_{1111}(-r)}{3}\;\partial_{221}\psi(0)-\frac{\Delta_{2211}(-r)}{2}\partial_{111}\psi(0)+\frac{2}{3}\Delta_{2111}(-r)\partial_{211}\psi(0)=\mathcal{O}_{\mathbb{P}}(1)

Combining Equations (6.1) and (6.1), an elementary calculus leads to

detHψ(z)detHψ(w)=r2[A12+g(r)]\det H_{\psi}(z)\det H_{\psi}(w)=r^{2}[-{A_{1}}^{2}+g(r)]

where

g(r)=\displaystyle g(r)= rA1(Br+Br)+r2BrBrr2A1(CrCr)\displaystyle rA_{1}(-B_{r}+B_{r^{\prime}})+r^{2}B_{r}B_{r^{\prime}}-r^{2}A_{1}(C_{r}-C_{r}^{\prime})
=rA1(𝒪(rα))+r2BrBrr2A1(CrCr)\displaystyle=rA_{1}(\mathcal{O}_{\mathbb{P}}(r^{\alpha}))+r^{2}B_{r}B_{r^{\prime}}-r^{2}A_{1}(C_{r}-C_{r}^{\prime})
=𝒪(A1r1+α)+r2B02+𝒪(A1r1+α)+𝒪(r2+α)+𝒪(r2+2α)\displaystyle=\mathcal{O}_{\mathbb{P}}(A_{1}r^{1+\alpha})+r^{2}B_{0}^{2}+\mathcal{O}_{\mathbb{P}}(A_{1}r^{1+\alpha})+\mathcal{O}_{\mathbb{P}}(r^{2+\alpha})+\mathcal{O}_{\mathbb{P}}(r^{2+2\alpha})
=r2B02+𝒪(A1r1+α)+𝒪(r2+α).\displaystyle=r^{2}B_{0}^{2}+\mathcal{O}_{\mathbb{P}}(A_{1}r^{1+\alpha})+\mathcal{O}_{\mathbb{P}}(r^{2+\alpha}).

As a consequence from Lemma 1, the 2-point correlation function given by (30) becomes

K2(z,w)\displaystyle\mathrm{K}_{2}(z,w) =1(2π)2253(η0)μ0𝔼(r)[|A12g(r)|](1+o(1)).\displaystyle=\frac{1}{(2\pi)^{2}2^{5}\sqrt{3}(-\eta_{0})\mu_{0}}\mathbb{E}^{(r)}\left[|{A_{1}}^{2}-g(r)|\right](1+o(1)). (37)

6.2. Dependency of derivatives

In view of the previous result, we will have to estimate quantities related to the random vectors

X=(22ψ(0),111ψ(0),122ψ(0),112ψ(0),1111ψ(0))X=(\partial_{22}\psi(0),\partial_{111}\psi(0),\partial_{122}\psi(0),\partial_{112}\psi(0),\partial_{1111}\psi(0))

and YrY_{r}. We must consider the case where (X,Y0)(X,Y_{0}) is degenerate. Examining Remark 1, we can split the variables involved in several groups that are mutually independent, there are for instance only two groups of size 33,

{1ψ(0),122ψ(0),111ψ(0)} and {22ψ(0),1111ψ(0),11ψ(0)}.\displaystyle\{\partial_{1}\psi(0),\partial_{122}\psi(0),\partial_{111}\psi(0)\}\text{\rm{ and }}\{\partial_{22}\psi(0),\partial_{1111}\psi(0),\partial_{11}\psi(0)\}.

Other groups, such as {112ψ(0),2ψ(0)}\{{\partial_{112}\psi(0),\partial_{2}\psi(0)}\}, have less members, and in the isotropic case they won’t be in a linear relation because of (5):

Cov (112ψ(0),2ψ(0))2=m2,22<m2,0m4,2=Var(2)Var(112).\displaystyle\text{\rm{ Cov }}(\partial_{112}\psi(0),\partial_{2}\psi(0))^{2}=m_{2,2}^{2}<m_{2,0}m_{4,2}=\mathrm{Var}(\partial_{2})\mathrm{Var}(\partial_{112}).

There is actually no other case to consider. Let us elucidate what can happen within the two bigger groups.

Proposition 4.

Assume the spectral measure FF is isotropic and not reduced to a Dirac mass in 0. There is (α,β,γ)(0,0,0)(\alpha,\beta,\gamma)\neq(0,0,0) such that α1ψ(0)+β111ψ(0)+γ122ψ(0)=0\alpha\partial_{1}\psi(0)+\beta\partial_{111}\psi(0)+\gamma\partial_{122}\psi(0)=0 a.s. iff β=γ\beta=\gamma and FF is uniformly spread along a circle of radius α/β\sqrt{\alpha/\beta}, i.e. if ψ\psi is a SGRW with parameter α/β\sqrt{\alpha/\beta}.

There is no (α,β,γ)(0,0,0)(\alpha,\beta,\gamma)\neq(0,0,0) such that a.s. α11ψ(0)+β22ψ(0)+γ1111ψ(0)=0\alpha\partial_{11}\psi(0)+\beta\partial_{22}\psi(0)+\gamma\partial_{1111}\psi(0)=0.

Proof.

Using (3) and recalling the symmetry ma,b=mb,am_{a,b}=m_{b,a}

Var(α1ψ(0)+β111ψ(0)+γ122ψ(0))=\displaystyle\mathrm{Var}(\alpha\partial_{1}\psi(0)+\beta\partial_{111}\psi(0)+\gamma\partial_{122}\psi(0))= α2m2,0+β2m6,0+γ2m2,4+2αβm4,0+2αγm2,2+2βγm4,2\displaystyle\alpha^{2}m_{2,0}+\beta^{2}m_{6,0}+\gamma^{2}m_{2,4}+2\alpha\beta m_{4,0}+2\alpha\gamma m_{2,2}+2\beta\gamma m_{4,2}
=\displaystyle= (α2λ12β2λ16γ2λ12λ24+2αβλ14+2αγλ12λ222βγλ14λ22)F(dλ)\displaystyle\int(-\alpha^{2}\lambda_{1}^{2}-\beta^{2}\lambda_{1}^{6}-\gamma^{2}\lambda_{1}^{2}\lambda_{2}^{4}+2\alpha\beta\lambda_{1}^{4}+2\alpha\gamma\lambda_{1}^{2}\lambda_{2}^{2}-2\beta\gamma\lambda_{1}^{4}\lambda_{2}^{2})F(d\lambda)
=\displaystyle= (αλ1+βλ13+γλ1λ22)2F(dλ).\displaystyle-\int(-\alpha\lambda_{1}+\beta\lambda_{1}^{3}+\gamma\lambda_{1}\lambda_{2}^{2})^{2}F(d\lambda).

Hence, dFdF-a.s., either λ1=0\lambda_{1}=0 or γλ12+βλ12=α\gamma\lambda_{1}^{2}+\beta\lambda_{1}^{2}=\alpha. By isotropy, it implies that γ=β\gamma=\beta and that FF’s support is concentrated on zero and the circle with radius α/β\sqrt{\alpha/\beta}. It corresponds to the GRW with radius α/β\sqrt{\alpha/\beta} plus an additional constant term.

In the same way,

0=Var(α22ψ(0)+β11ψ(0)+γ1111ψ(0))=(αλ22+βλ12+γλ14)2F(dλ)\displaystyle 0=\mathrm{Var}(\alpha\partial_{22}\psi(0)+\beta\partial_{11}\psi(0)+\gamma\partial_{1111}\psi(0))=\int(\alpha\lambda_{2}^{2}+\beta\lambda_{1}^{2}+\gamma\lambda_{1}^{4})^{2}F(d\lambda)

implies that FF is trivial if FF is isotropic. ∎

In conclusion, the only non-trivial linear relations possibly satisfied by the derivatives involved in (X,Y0)(X,Y_{0}) is 111ψ(0)+122ψ(0)=α1ψ(0),α>0\partial_{111}\psi(0)+\partial_{122}\psi(0)=\alpha\partial_{1}\psi(0),\alpha>0 and can only be satisfied by a SGRW. In the light of these results, functionals of interest only depend on the law of the vector XX^{\prime} under the conditioning Y0=0Y_{0}=0, where

X={(22ψ(0),111ψ(0),112ψ(0),1111ψ(0)) if ψ is a shifted GRWX otherwise\displaystyle X^{\prime}=\begin{cases}(\partial_{22}\psi(0),\partial_{111}\psi(0),\partial_{112}\psi(0),\partial_{1111}\psi(0))$ if $\psi$ is a shifted GRW$\\ X$ otherwise$\end{cases}

because if ψ\psi is a shifted GRW and Y0=0Y_{0}=0, 111ψ(0)+122ψ(0)=α1ψ(0)=0\partial_{111}\psi(0)+\partial_{122}\psi(0)=-\alpha\partial_{1}\psi(0)=0 a.s. hence 122ψ(0)\partial_{122}\psi(0) is directly expressible in function of 111ψ(0).\partial_{111}\psi(0).

Lemma 2.

The conditional density frf_{r} of XX^{\prime} knowing YrY_{r} converges pointwise to the density f0f_{0} of XX^{\prime} knowing Y0Y_{0}. There is furthermore σ1,σ2,c1,c2>0\sigma_{1},\sigma_{2},c_{1},c_{2}>0 such that for rr sufficiently small,

c1gσ1frc2gσ2\displaystyle c_{1}g_{\sigma_{1}}\leqslant f_{r}\leqslant c_{2}g_{\sigma_{2}}

where gσg_{\sigma} is the density of iid Gaussian variables Zσ=(Ziσ)iZ^{\sigma}=(Z^{\sigma}_{i})_{i} with common variance σ2.\sigma^{2}. Hence for any non-negative functional φr\varphi_{r}

c1𝔼(φr(σ1Z1))𝔼(r)(φr(X))c2𝔼(φr(σ2Z1)).\displaystyle c_{1}\mathbb{E}(\varphi_{r}(\sigma_{1}Z^{1}))\leqslant\mathbb{E}^{(r)}(\varphi_{r}(X^{\prime}))\leqslant c_{2}\mathbb{E}(\varphi_{r}(\sigma_{2}Z^{1})).
Proof.

Since the vector (X,Y0)(X^{\prime},Y_{0}) is non-degenerate, by continuity of the covariance matrix, the vector (X,Yr)(X^{\prime},Y_{r}) is non-degenerate either for rr sufficiently small, and the density of the former converges pointwise to the density of the latter. Hence the conditionnal density frf_{r} of (X|Yr)(X^{\prime}\;|\;Y_{r}) converges to f0f_{0} the non-degenerate multivariate conditional Gaussian density of (X|Y0)(X\;|\;Y_{0}).
Let Γr\Gamma_{r} be the covariance matrix of the conditional vector (X|Yr).(X^{\prime}\;|\;Y_{r}). For 1id,1\leqslant i\leqslant d, we denote by λi,(r)\lambda_{i,}(r) the ii-th eigenvalue of the matrix Γr\Gamma_{{r}}. Since λi(r)λi(0)>0\lambda_{i}(r)\to\lambda_{i}(0)>0, there exists constant σ1σ2>0\sigma_{1}\geqslant\sigma_{2}>0 such that for rr sufficiently small

12πσ22λi(r)12πσ12.\displaystyle\frac{1}{2\pi\sigma_{2}^{2}}\geqslant\lambda_{i}(r)\geqslant\frac{1}{2\pi\sigma_{1}^{2}}. (38)

Hence fr(x)f_{r}(x) is bounded between cexp(ixi2/(2πσ12))c\exp(-\sum_{i}x_{i}^{2}/(2\pi\sigma_{1}^{2})) and cexp(ixi2/(2πσ22))c^{\prime}\exp(-\sum_{i}x_{i}^{2}/(2\pi\sigma_{2}^{2})) for some c,c>0c,c^{\prime}>0, which gives the desired claims. ∎

6.3. Proof of (23) in Theorem 5.1

From (37), we have

K2(z,w)\displaystyle{\mathrm{K}}_{2}(z,w) =r2(2π)2253(η0)μ0(1+o(1))𝔼(r)[|A12g(r)|].\displaystyle=\frac{r^{2}}{(2\pi)^{2}2^{5}\sqrt{3}(-\eta_{0})\mu_{0}(1+o(1))}\mathbb{E}^{(r)}\left[|{A_{1}}^{2}-g(r)|\right].

According to Lemma 2, the conditional density frf_{r} of XX^{\prime} knowing YrY_{r} converges pointwise to the non-degenerate density f0f_{0} of XX^{\prime} knowing Y0Y_{0}, and φr(X):=A12g(r)\varphi_{r}(X^{\prime}):=A_{1}^{2}-g(r) is uniformly bounded by a polynomial P(X)P(X^{\prime}), Lebesgue’s Theorem then yields

limr0𝔼(r)[A12g(r)]\displaystyle\lim\limits_{r\rightarrow 0}\mathbb{E}^{{(r)}}\left[\mid A_{1}^{2}-g(r)\mid\right] =φr(x)fr(x)𝑑xlimrφr(x)f0(x)dx=𝔼(0)[A12].\displaystyle=\int\varphi_{r}(x)f_{r}(x)dx\to\int\lim_{r}\varphi_{r}(x)f_{0}(x)dx=\mathbb{E}^{(0)}\left[A_{1}^{2}\right]. (39)

To compute the conditionnal law of Z:=(22ψ(0),111ψ(0))Z:=(\partial_{22}\psi(0),\partial_{111}\psi(0)), recall that in virtue of (3) and Remark 1 22ψ(0)\partial_{22}\psi(0) and 111ψ(0)\partial_{111}\psi(0) are independent, and the covariance matrix of Y0Y_{0} is

Γ(Y0)=(m2,00000m2,00000m4,00000m2,2).\Gamma(Y_{0})=\left(\begin{array}[]{cccc}m_{2,0}&0&0&0\\ 0&m_{2,0}&0&0\\ 0&0&m_{4,0}&0\\ 0&0&0&m_{2,2}\end{array}\right).

The covariance matrix of ZZ and Y0Y_{0} is

Γ(Z,Y0)=(m2,1m0,3m2,2m1,3m4,0m3,1m5,0m4,1)=(00m2,20m4,0000).\displaystyle\Gamma({Z,Y_{0}})=\left(\begin{array}[]{cccc}m_{2,1}&m_{0,3}&m_{2,2}&m_{1,3}\\ m_{4,0}&m_{3,1}&m_{5,0}&m_{4,1}\end{array}\right)=\left(\begin{array}[]{cccc}0&0&m_{2,2}&0\\ m_{4,0}&0&0&0\end{array}\right).

It follows that the conditional covariance of ZZ knowing Y0Y_{0} is

Γ(Z|Y0)=Γ(Z)Γ(Z,Y0)Γ(Y0)1Γ(Z,Y0)t=\displaystyle\Gamma({Z\;|\;Y_{0}})=\Gamma(Z)-\Gamma(Z,Y_{0})\Gamma(Y_{0})^{-1}\Gamma(Z,Y_{0})^{t}= (m0,4m4,01m22200m6,0m2,01m4,02)\displaystyle\left(\begin{array}[]{cc}m_{0,4}-m_{4,0}^{-1}m_{22}^{2}&0\\ 0&m_{6,0}-m_{2,0}^{-1}m_{4,0}^{2}\end{array}\right)
=(Var(22ψ(0)|111ψ(0))00Var(111ψ(0)|1ψ(0)))\displaystyle=\left(\begin{array}[]{cc}\mathrm{Var}(\partial_{22}\psi(0)|\partial_{111}\psi(0))&0\\ 0&\mathrm{Var}(\partial_{111}\psi(0)|\partial_{1}\psi(0))\end{array}\right)

the diagonal terms are positive in virtue of (5). By conditionnal independence of 22ψ(0)\partial_{22}\psi(0) and 111ψ(0)\partial_{111}\psi(0), and using Proposition 2

𝔼(0)(A12)=Var(22ψ(0)|11,1ψ(0))Var(11,1ψ(0)|1ψ(0))\displaystyle\mathbb{E}^{(0)}(A_{1}^{2})=\mathrm{Var}(\partial_{22}\psi(0)|\partial_{11,1}\psi(0))\mathrm{Var}(\partial_{11,1}\psi(0)|\partial_{1}\psi(0)) =(m0,4m4,01m222)(m6,0m2,01m4,02)>0\displaystyle=(m_{0,4}-m_{4,0}^{-1}m_{22}^{2})(m_{6,0}-m_{2,0}^{-1}m_{4,0}^{2})>0
=253μ0(323η0(5ν0η0+3μ02))\displaystyle=\frac{2^{5}}{3}\mu_{0}(\frac{3\cdot 2^{3}}{\eta_{0}}(-5\nu_{0}\eta_{0}+3\mu_{0}^{2}))
=28μ0(5ν0η0+3μ02)η0.\displaystyle={2^{8}}\mu_{0}\frac{(-5\nu_{0}\eta_{0}+3\mu_{0}^{2})}{\eta_{0}}.

Combining Equation (39) , (37) we obtain

limr0K2(z,w)=1(2π)2 253(η0)μ0(1+o(1)) 28μ0(5ν0η0+3μ02)η0=10ν0η06μ02π23η02(1+o(1))=:a(1+o(1)).\displaystyle\lim\limits_{r\rightarrow 0}{\mathrm{K}}_{2}(z,w)=\frac{1}{(2\pi)^{2}\;2^{5}\sqrt{3}(-\eta_{0})\mu_{0}(1+o(1))}\;{2^{8}}\mu_{0}\frac{(-5\nu_{0}\eta_{0}+3\mu_{0}^{2})}{\eta_{0}}=\frac{10\;\nu_{0}\eta_{0}-6\mu_{0}^{2}}{\pi^{2}\;\sqrt{3}\eta_{0}^{2}(1+o(1))}=:a(1+o(1)). (40)

Finally, the second factorial moment of 𝒩ρc\mathcal{N}^{c}_{\rho} when ρ0,\rho\rightarrow 0, is given by

𝔼[𝒩ρc(𝒩ρc1)]\displaystyle\mathbb{E}[\mathcal{N}^{c}_{\rho}(\mathcal{N}^{c}_{\rho}-1)] =Bρ×BρK2(z,w)dzdw=a|Bρ|2(1+o(1)).\displaystyle=\int\int_{B_{\rho}\times B_{\rho}}{\mathrm{K}}_{2}(z,w)\,\mathrm{d}z\;\mathrm{d}w=a|B_{\rho}|^{2}(1+o(1)).

Recalling that

λc=43μ0η0π\displaystyle\lambda_{c}=\frac{4}{\sqrt{3}}\frac{\mu_{0}}{-\eta_{0}\pi}

yields indeed a=λc2𝖱c.a=\lambda_{c}^{2}\mathsf{R}_{c}.

Let us show that 𝖱c183\mathsf{R}_{c}\geqslant\frac{1}{8\sqrt{3}}. Given a measure μ\mu on +\mathbb{R}_{+}, denote by

mk(μ)=tkμ(dt).m_{k}(\mu)=\int t^{k}\mu(dt).

Since FF is isotropic, define μ\mu as the radial part of FF, yielding with a polar change of coordinates

2λ1aλ2bF(dλ)=02πcos(θ)asin(θ)bdθma+b+1(μ).\displaystyle\int_{\mathbb{R}^{2}}\lambda_{1}^{a}\lambda_{2}^{b}F(d\lambda)={\int_{0}^{2\pi}\cos(\theta)^{a}\sin(\theta)^{b}}d\theta m_{a+b+1}(\mu).

Introduce the probability measure, for A+,A\subset\mathbb{R}_{+},

μ~(A)=Atμ(dt)m1(μ).\widetilde{\mu}(A)=\frac{\int_{A}t\mu(dt)}{m_{1}(\mu)}.

Using the spectral representation in Proposition 2 yields for some c>0c>0

ν0η0μ02=\displaystyle\frac{\nu_{0}\eta_{0}}{\mu_{0}^{2}}= cm3(μ)m1(μ)m2(μ)2=cm1(μ)m2(μ~)m1(μ)(m1(μ)m1(μ~))2=cm2(μ~)m1(μ~)2c\displaystyle c\frac{m_{3}(\mu)m_{1}(\mu)}{m^{2}(\mu)^{2}}=c\frac{m_{1}(\mu)m_{2}(\widetilde{\mu})m_{1}(\mu)}{(m_{1}(\mu)m_{1}(\widetilde{\mu}))^{2}}=c\frac{m_{2}(\widetilde{\mu})}{m_{1}(\widetilde{\mu})^{2}}\geqslant c

by the Cauchy-Schwarz inequality. The ratio is minimal if the equality is obtained in the Cauchy-Schwarz inequality, i.e. when t2t^{2} is proportionnal to tt μ~\widetilde{\mu}-a.e.. This is the case only if F(dλ)F(d\lambda) is uniformly spread on a circle of 2\mathbb{R}^{2}, with perhaps also an additional atom in 0. This corresponds exactly to the class of fields derived in Example 1, which are the SGRW. For the precise computation of the constant 183\frac{1}{8\sqrt{3}}, see Example 3.

In example 4, we derive spectral measures Ft,t>1F_{t},t>1 which achieve repulsion factors 𝖱c\mathsf{R}_{c} in an interval of the form (α0,)(\alpha_{0},\infty) for some α0\alpha_{0}\in\mathbb{R}. Therefore it remains to show that all values between 183\frac{1}{8\sqrt{3}} and α0\alpha_{0} can be achieved. For that we use an interpolation

Gs:=sFRW+(1s)F2,s[0,1]\displaystyle G_{s}:=sF_{RW}+(1-s)F_{2},s\in[0,1]

where FRWF_{RW} is the spectral measure of a GRW and F2F_{2} belongs to the parametric family Ft,t1F_{t},t\geqslant 1. The ratio of moments

sm3(Gs)m1(Gs)m2(Gs)2\displaystyle s\mapsto\frac{m_{3}(G_{s})m_{1}(G_{s})}{m^{2}(G_{s})^{2}}

evolves continuously with ss because all the members of the numerator and denominator do, hence the repulsion factor evolves continuously between 183\frac{1}{8\sqrt{3}} and α0\alpha_{0} and achieves all intermediary values.

6.4. Proof of (24) in Theorem 5.1

To compute the second factorial moment of 𝒩ρe=𝒩ρ(0,)\mathcal{N}^{e}_{\rho}=\mathcal{N}_{\rho}^{(0,\infty)} when ρ0\rho\rightarrow 0, we apply the Kac-rice formula of Theorem 3.1 in the case B1=B2=(0,)B_{1}=B_{2}=(0,\infty)

𝔼[𝒩ρe(𝒩ρe1)]=Bρ×BρK2e,e(z,w)dzdw,\mathbb{E}[\mathcal{N}^{e}_{\rho}(\mathcal{N}^{e}_{\rho}-1)]=\int\int_{B_{\rho}\times B_{\rho}}{\mathrm{{K}}}^{e,e}_{2}(z,w)\,\mathrm{d}z\;\mathrm{d}w, (41)

where

K2e,e(z,w)=\displaystyle{\mathrm{K}}^{e,e}_{2}(z,w)= ϕ(ψ(z),ψ(w))((0,0)),(0,0))\displaystyle\;\phi_{(\nabla\psi(z),\nabla\psi(w))}((0,0)),(0,0))
×𝔼(r)[|detHψ(z)||detHψ(w)| 1{detHψ(z)>0}  1{detHψ(w)>0}].\displaystyle\times\mathbb{E}^{(r)}\left[\;|\det H_{\psi}(z)|\;|\det H_{\psi}(w)|\;\mathbf{1}_{\{\det H_{\psi}(z)>0\}}\;\;\mathbf{1}_{\{\det H_{\psi}(w)>0\}}\right].

It becomes in virtue of (31)

K2e,e(z,w)=r2ϕ(ψ(z),ψ(w))((0,0),(0,0))ar{\mathrm{K}}^{e,e}_{2}(z,w)=r^{2}\;\phi_{(\nabla\psi(z),\nabla\psi(w))}((0,0),(0,0))\;a_{r}\\ (42)

where

ar:=\displaystyle a_{r}:= 𝔼(r)[|A12g(r)|Ir]\displaystyle\mathbb{E}^{(r)}\left[\;\big{|}{A_{1}}^{2}-g(r)\big{|}\;I_{r}\right]
Ir:=\displaystyle I_{r}:= 𝟏{detHψ(z)>0}𝟏{detHψ(w)>0}.\displaystyle\mathbf{1}_{\{\det H_{\psi}(z)>0\}}\mathbf{1}_{\{\det H_{\psi}(w)>0\}}.

To be able to prove (24), we need to establish an upper bound and a lower bound of ara_{r} So the proof is separated into two parts. We first give in Lemma 3 an asymptotic expression of ara_{r} to get rid of superfluous variables.

Lemma 3.

Let Jr:=𝟏{|A1|<rB0}J_{r}:=\mathbf{1}_{\{|A_{1}|<rB_{0}\}},

|ar𝔼(r)(|A12r2B02|Jr)|=𝒪(r3+α)\displaystyle|a_{r}-\mathbb{E}^{(r)}(|A_{1}^{2}-r^{2}B_{0}^{2}|J_{r})|=\mathcal{O}_{\mathbb{P}}(r^{3+\alpha^{\prime}})

for 0<α<α,0<\alpha^{\prime}<\alpha, and as r0r\to 0

ar𝔼(|A12r2B02|Jr).\displaystyle a_{r}\asymp\mathbb{E}(|A_{1}^{2}-r^{2}B_{0}^{2}|J_{r}). (43)
Proof.

From (6.1)-(6.1) in the proof of Lemma 1,

Ir𝟏{|A1|<rDr}\displaystyle I_{r}\leqslant\mathbf{1}_{\{|A_{1}|<rD_{r}\}}

where

Dr:=\displaystyle D_{r}:= |B0|+|Br|+|Br|+r(|Cr|+|Cr|),\displaystyle|B_{0}|+|{B}_{r}|+|B_{r}^{\prime}|+r(|C_{r}|+|C^{\prime}_{r}|),

is a variable with Gaussian tail. Recall that g(r)=r2B02+𝒪(Arr1+α+r2+α)g(r)=r^{2}B_{0}^{2}+\mathcal{O}_{\mathbb{P}}(A_{r}r^{1+\alpha}+r^{2+\alpha}), hence using (34),

𝔼(r)[(|A12g(r)||A12r2B02|)Ir)]\displaystyle\mathbb{E}^{(r)}\left[(|A_{1}^{2}-g(r)|-|A_{1}^{2}-r^{2}B_{0}^{2}|)I_{r})\right]\leqslant 𝔼(r)(𝒪(r1+αA1+r2+α))|𝟏{|A1|<rDr})\displaystyle\mathbb{E}^{(r)}(\mathcal{O}_{\mathbb{P}}(r^{1+\alpha}A_{1}+r^{2+\alpha}))|\mathbf{1}_{\{|A_{1}|<rD_{r}\}})
\displaystyle\leqslant r2+α𝔼(r)(𝒪(Dr+1)𝟏{|A1|<rDr}).\displaystyle r^{2+\alpha}\mathbb{E}^{(r)}(\mathcal{O}_{\mathbb{P}}(D_{r}+1)\mathbf{1}_{\{|A_{1}|<rD_{r}\}}). (44)

Let p,q>1,η>0p,q>1,\eta>0 such that p1+q1=1p^{-1}+q^{-1}=1 and α+1ηq>α+1\alpha+\frac{1-\eta}{q}>\alpha^{\prime}+1, then Holder’s inequality yields

𝔼(r)(𝒪(Dr+1)𝟏{|A1|<rDr})𝔼(r)(𝒪(Dr+1)p)1p(r)(|A1|<rDr)1q.\displaystyle\mathbb{E}^{(r)}(\mathcal{O}_{\mathbb{P}}(D_{r}+1)\mathbf{1}_{\{|A_{1}|<rD_{r}\}})\leqslant\mathbb{E}^{(r)}(\mathcal{O}_{\mathbb{P}}(D_{r}+1)^{p})^{\frac{1}{p}}\mathbb{P}^{(r)}(|A_{1}|<rD_{r})^{\frac{1}{q}}.

The probability on the right hand member can be bounded by

(r)(|A1|<rDr)(r)(|Dr|>rη)+(r)(|A1|<r1η).\displaystyle\mathbb{P}^{(r)}(|A_{1}|<rD_{r})\leqslant\mathbb{P}^{(r)}(|D_{r}|>r^{-\eta})+\mathbb{P}^{(r)}(|A_{1}|<r^{1-\eta}).

All variables involved in 𝒪(Dr)\mathcal{O}_{\mathbb{P}}(D_{r}) have a Gaussian tail, hence

(r)(𝒪(|Dr|)>rη)=o(r2).\displaystyle\mathbb{P}^{(r)}(\mathcal{O}_{\mathbb{P}}(|D_{r}|)>r^{-\eta})=o(r^{2}).

By Lemma 2 with φr(x)=𝟏{|x1x2|<r1η},\varphi_{r}(x)=\mathbf{1}_{\{|x_{1}x_{2}|<r^{1-\eta}\}}, and Lemma 5-(i) (with s=0s=0),

(r)(|A1|<r1η)=𝔼(r)(φr(X))c2𝔼(φr(σ2Z))<\displaystyle\mathbb{P}^{(r)}(|A_{1}|<r^{1-\eta})=\mathbb{E}^{(r)}(\varphi_{r}(X^{\prime}))\leqslant c_{2}\mathbb{E}(\varphi_{r}(\sigma_{2}Z))< cr1ηln(r)\displaystyle c^{\prime}r^{{1-\eta}}\ln(r) (45)

hence finally

r2+α𝔼(r)(𝒪(Dr+1)𝟏{|A1|<rDr})<\displaystyle r^{2+\alpha}\mathbb{E}^{(r)}(\mathcal{O}_{\mathbb{P}}(D_{r}+1)\mathbf{1}_{\{|A_{1}|<rD_{r}\}})< cr2+α+1ηqln(r)1/q=O(r3+α).\displaystyle c^{\prime}r^{2+\alpha+\frac{1-\eta}{q}}\ln(r)^{1/q}=O(r^{3+\alpha^{\prime}}). (46)

To simplify indicators, remark that in virtue of (6.1),(6.1),

Ir\displaystyle I_{r} =𝟏{A1+rBr+r2Cr>0,A1+rBr+r2Cr′′>0}\displaystyle=\mathbf{1}_{\{A_{1}+rB_{r}+r^{2}C_{r}>0,-A_{1}+rB_{r}^{\prime}+r^{2}C_{r}^{\prime\prime}>0\}}
Jr\displaystyle J_{r} =𝟏{A1+rB0>0,A1+rB0>0}.\displaystyle=\mathbf{1}_{\{A_{1}+rB_{0}>0,-A_{1}+rB_{0}>0\}}.

Both the events {Ir=1},{Jr=1}\{I_{r}=1\},\{J_{r}=1\} imply |A1|<rDr|A_{1}|<rD_{r}. If IrJrI_{r}\neq J_{r}, A1+rBr+r2CrA_{1}+rB_{r}+r^{2}C_{r} has a sign different from A1+rB0A_{1}+rB_{0}, or A1+rBr+r2Cr-A_{1}+rB_{r}^{\prime}+r^{2}C_{r}^{\prime} has a sign different from A1+rB0-A_{1}+rB_{0}. In both cases it implies another event of magnitude 𝒪(r1+α)\mathcal{O}_{\mathbb{P}}(r^{1+\alpha}) because Br,Br=B0+𝒪(rα),Cr,Cr=𝒪(1):B_{r},B_{r}^{\prime}=B_{0}+\mathcal{O}_{\mathbb{P}}(r^{\alpha}),C_{r},C_{r}^{\prime}=\mathcal{O}_{\mathbb{P}}(1):

|IrJr|\displaystyle|I_{r}-J_{r}| 2(𝟏{(A1+rBr+r2Cr)(A1+rB0)<0}+𝟏{(A1+rBr+r2Cr)(A1+rB0)<0})\displaystyle\leqslant 2\left(\mathbf{1}_{\{(A_{1}+rB_{r}+r^{2}C_{r})\cdot(A_{1}+rB_{0})<0\}}+\mathbf{1}_{\{(-A_{1}+rB^{\prime}_{r}+r^{2}C_{r}^{\prime})\cdot(-A_{1}+rB_{0})<0\}}\right)
2(𝟏{|A1+rB0|<𝒪(r1+α)}+𝟏{|A1+rB0|<𝒪(r1+α)}).\displaystyle\leqslant 2\left(\mathbf{1}_{\{|A_{1}+rB_{0}|<\mathcal{O}_{\mathbb{P}}(r^{1+\alpha})\}}+\mathbf{1}_{\{|-A_{1}+rB_{0}|<\mathcal{O}_{\mathbb{P}}(r^{1+\alpha})\}}\right).

Let now p,q>1,η>0p,q>1,\eta>0 such that (1+αη)/q>1+α(1+\alpha-\eta)/q>1+\alpha^{\prime} . Since also IrJr0I_{r}-J_{r}\neq 0 implies that either Ir=1I_{r}=1 or Jr=1J_{r}=1 and so |A1|<rDr,|A_{1}|<rD_{r}, collecting (44),(46),

|ar\displaystyle|a_{r}- 𝔼(r)[|A12r2B02|Jr]||𝔼(r)(|A12r2B02||IrJr|)|+𝒪(r3+α)\displaystyle\mathbb{E}^{(r)}\left[|A_{1}^{2}-r^{2}B_{0}^{2}|J_{r}\right]|\leqslant|\mathbb{E}^{(r)}(|A_{1}^{2}-r^{2}B_{0}^{2}||I_{r}-J_{r}|)|+\mathcal{O}_{\mathbb{P}}(r^{3+\alpha^{\prime}})
|𝔼(r)(|A12r2B02|𝟏|A1|<rDr|IrJr|)|+𝒪(r3+α)\displaystyle\leqslant|\mathbb{E}^{(r)}(|A_{1}^{2}-r^{2}B_{0}^{2}|\mathbf{1}_{|A_{1}|<rD_{r}}|I_{r}-J_{r}|)|+\mathcal{O}_{\mathbb{P}}(r^{3+\alpha^{\prime}})
𝔼(r)[(r2Dr2+r2B02)(𝟏{|A1+rB0|<𝒪(r1+α)}+𝟏{|A1+rB0|<𝒪(r1+α)})]+𝒪(r3+α)\displaystyle\leqslant\mathbb{E}^{(r)}\left[(r^{2}D_{r}^{2}+r^{2}B_{0}^{2})\left(\mathbf{1}_{\{|A_{1}+rB_{0}|<\mathcal{O}_{\mathbb{P}}(r^{1+\alpha})\}}+\mathbf{1}_{\{|-A_{1}+rB_{0}|<\mathcal{O}_{\mathbb{P}}(r^{1+\alpha})\}}\right)\right]+\mathcal{O}_{\mathbb{P}}(r^{3+\alpha^{\prime}})
𝔼(r)[(r2Dr2+r2B02)p]1p[(r)(|A1+rB0|<𝒪(r1+α))1q+(r)(|A1+rB0|<𝒪(r1+α))1q]+𝒪(r3+α).\displaystyle\leqslant\mathbb{E}^{(r)}[(r^{2}D_{r}^{2}+r^{2}B_{0}^{2})^{p}]^{\frac{1}{p}}\left[\mathbb{P}^{(r)}\left({|A_{1}+rB_{0}|<\mathcal{O}_{\mathbb{P}}(r^{1+\alpha})}\right)^{\frac{1}{q}}+\mathbb{P}^{(r)}\left({|-A_{1}+rB_{0}|<\mathcal{O}_{\mathbb{P}}(r^{1+\alpha})}\right)^{\frac{1}{q}}\right]+\mathcal{O}_{\mathbb{P}}(r^{3+\alpha^{\prime}}).

We have

(r)(|A1+rB0|<𝒪(r1+α))(r)(|A1+rB0|<r1+αη)+(r)(𝒪(1)>rη).\mathbb{P}^{(r)}(|A_{1}+rB_{0}|<\mathcal{O}_{\mathbb{P}}(r^{1+\alpha}))\leqslant\mathbb{P}^{(r)}(|A_{1}+rB_{0}|<r^{1+\alpha-\eta})+\mathbb{P}^{(r)}(\mathcal{O}_{\mathbb{P}}(1)>r^{-\eta}).

By an application of Lemma 2 similar to (45) with φr(x)\varphi_{r}(x) of the form 𝟏{|x1x1+rai,jxixj|<r1+αη}\mathbf{1}_{\{|x_{1}x_{1}+r\sum a_{i,j}x_{i}x_{j}|<r^{1+\alpha-\eta}\}} and Lemma 5-(i); the first member is in r1+αηln(r)r^{1+\alpha-\eta}\ln(r), hence finally for some c<c<\infty

|ar\displaystyle|a_{r}- 𝔼(r)[|A12r2B02|Jr]|cr2r(1+αη)/qln(r)1q+𝒪(r3+α)=𝒪(r3+α).\displaystyle\mathbb{E}^{(r)}\left[|A_{1}^{2}-r^{2}B_{0}^{2}|J_{r}\right]|\leqslant cr^{2}r^{(1+\alpha-\eta)/q}\ln(r)^{\frac{1}{q}}+\mathcal{O}_{\mathbb{P}}(r^{3+\alpha^{\prime}})=\mathcal{O}_{\mathbb{P}}(r^{3+\alpha^{\prime}}).

Finally, (43) follows from Lemma 2 with φr(X)=A12r2B02\varphi_{r}(X^{\prime})=A_{1}^{2}-r^{2}B_{0}^{2}.∎

6.4.1. Upper bound in (24)

According to the previous lemma it suffices to give an upper bound of 𝔼[A12r2B02|Jr].\mathbb{E}\left[\mid A^{2}_{1}-r^{2}B^{2}_{0}|J_{r}\right]. We stress that the crucial point that justifies the absence of a log term in the final result (compared to (25)) is the following inequality

Jr=𝟏{|A1|<rB0}𝟏{|22ψ(0)|<2r221}+𝟏{|111ψ(0)|<23r1111ψ(0)},\displaystyle J_{r}=\mathbf{1}_{\{|A_{1}|<rB_{0}\}}\leqslant\mathbf{1}_{\{|\partial_{22}\psi(0)|<2r\partial_{221}\}}+\mathbf{1}_{\{|\partial_{111}\psi(0)|<\frac{2}{3}r\partial_{1111}\psi(0)\}},

hence since B02B_{0}^{2} is a polynomial in XX^{\prime} we can use Lemma 5-(iii) several times and get for some c<c<\infty

𝔼(|A12r2B02|Jr)2𝔼(|r2B02|Jr)cr3.\displaystyle\mathbb{E}(|A_{1}^{2}-r^{2}B_{0}^{2}|J_{r})\leqslant 2\mathbb{E}(|r^{2}B_{0}^{2}|J_{r})\leqslant cr^{3}. (47)

Then, from (42),(47) and (37), we deduce that for some c<c^{\prime}<\infty

K2e,e(z,w)cr3.\mathrm{K}^{e,e}_{2}(z,w)\leqslant c^{\prime}\;r^{3}. (48)

Finally, from (41) and (48), we deduce for some c′′<c^{\prime\prime}<\infty

𝔼[𝒩ρe(𝒩ρe1)]c′′ρ7.\displaystyle\mathbb{E}[\mathcal{N}^{e}_{\rho}(\mathcal{N}^{e}_{\rho}-1)]\leqslant c^{\prime\prime}\;\rho^{7}.

6.4.2. Lower bound in (24)

Thanks to Lemma 3, it is sufficient to give a lower bound of 𝔼(|A12r2B02|𝟏{|A1|rB0}|).\mathbb{E}(|A_{1}^{2}-r^{2}B_{0}^{2}|\mathbf{1}_{\{|A_{1}|\leqslant rB_{0}\}}|). Let us first assume that the Gaussian field ψ\psi is not a SGRW (Example 1), hence the derivatives involved in XX and Y0Y_{0} are not linearly linked. Define the event

Ω=\displaystyle\Omega= {|22ψ(0)|<r,12<111ψ(0)<1,|211ψ(0)|<1,8<122,|1111ψ(0)|<1}.\displaystyle\{|\partial_{22}\psi(0)|<r,\frac{1}{2}<\partial_{111}\psi(0)<1,|\partial_{211}\psi(0)|<1,8<\partial_{122},|\partial_{1111}\psi(0)|<1\}.

We recall

A1\displaystyle A_{1} =22ψ(0)111ψ(0),\displaystyle=\partial_{22}\psi(0)\partial_{111}\psi(0),
B0\displaystyle B_{0} =221111ψ(0)211ψ(0)2+1322ψ(0)1111ψ(0)\displaystyle=\partial_{221}\;\partial_{111}\psi(0)-\partial_{211}\psi(0)^{2}+\frac{1}{3}\partial_{22}\psi(0)\;\partial_{1111}\psi(0)
Y0\displaystyle Y_{0} =(1ψ(0),2ψ(0),11ψ(0),12ψ(0)).\displaystyle=(\partial_{1}\psi(0),\partial_{2}\psi(0),\partial_{11}\psi(0),\partial_{12}\psi(0)).

Hence under Ω\Omega

|A1|<\displaystyle|A_{1}|< r\displaystyle r
B0>41r3\displaystyle B_{0}>4-1-\frac{r}{3}

Hence for rr sufficiently small, B0>2B_{0}>2, in particular |A1|r|B0|/2|A_{1}|\leqslant r|B_{0}|/2 and we obtain

𝔼(|A12r2B02|𝟏{|A1|rB0})\displaystyle\mathbb{E}(|A_{1}^{2}-r^{2}B_{0}^{2}|\mathbf{1}_{\{|A_{1}|\leqslant rB_{0}\}}) 𝔼[𝟏Ωr2B02/4𝟏{B02A1/r}]\displaystyle\geqslant\mathbb{E}\left[\mathbf{1}_{\Omega}\mid r^{2}B_{0}^{2}/4\mid\mathbf{1}_{\{B_{0}\geqslant 2A_{1}/r\}}\right]
r2(Ω).\displaystyle\geqslant\;r^{2}\mathbb{P}(\Omega).

Then since XX is non-degenerate, its density is uniformly bounded and the proof is concluded with

(Ω)cr>0\displaystyle\mathbb{P}(\Omega)\geqslant cr>0

for some c>0.c>0. In the degenerate case of the SGRW, 122ψ(0)=111ψ(0)\partial_{122}\psi(0)=-\partial_{111}\psi(0) if Y0=0Y_{0}=0 and we put instead

Ω=\displaystyle\Omega= {|111ψ(0)|<r,12<22ψ(0)<1,1111ψ(0)>19,|211ψ(0)|<1}\displaystyle\{|\partial_{111}\psi(0)|<r,\frac{1}{2}<\partial_{22}\psi(0)<1,\partial_{1111}\psi(0)>19,|\partial_{211}\psi(0)|<1\}

If Y0=0Y_{0}=0 and Ω\Omega is realised,

|A1|\displaystyle|A_{1}| <r\displaystyle<r
B0=111ψ(0)2211ψ(0)2+13221111ψ(0)\displaystyle B_{0}=-\partial_{111}\psi(0)^{2}-\partial_{211}\psi(0)^{2}+\frac{1}{3}\partial_{22}\partial_{1111}\psi(0) >r21+196\displaystyle>-r^{2}-1+\frac{19}{6}

hence |A1|<r<B0r/2|A_{1}|<r<B_{0}r/2 for rr small enough and the same method can be applied because XX^{\prime} has a bounded density. Therefore, it holds for some c>0c^{\prime}>0

𝔼(0)(|A12r2B02|𝟏{|A1|rB0}|)\displaystyle\mathbb{E}^{(0)}(|A_{1}^{2}-r^{2}B_{0}^{2}|\mathbf{1}_{\{|A_{1}|\leqslant rB_{0}\}}|) cr3.\displaystyle\geqslant c^{\prime}\;r^{3}. (49)

From (42), (37)and (49), we get for some c′′>0c^{\prime\prime}>0

K2e,e(z,w)\displaystyle{\mathrm{K}}^{e,e}_{2}(z,w) c′′r3.\displaystyle\geqslant\;c^{\prime\prime}r^{3}. (50)

Finally, from (41) and (50), we deduce that for some c′′′>0c^{\prime\prime\prime}>0

𝔼[𝒩ρe(𝒩ρe1)]\displaystyle\mathbb{E}[\mathcal{N}^{e}_{\rho}(\mathcal{N}^{e}_{\rho}-1)] c′′′ρ7.\displaystyle\geqslant c^{\prime\prime\prime}\rho^{7}.

6.5. Proof of (25) in Theorem 5.1

Using Theorem 3.1 with B1=B2=(,0)B_{1}=B_{2}=(-\infty,0), the second factorial moment of 𝒩ρs=Nρ(,0)\mathcal{N}^{s}_{\rho}=N_{\rho}^{(-\infty,0)} is given by

𝔼[𝒩ρs(𝒩ρs1)]=Bρ×BρK2s,s(z,w)dzdw,\mathbb{E}[\mathcal{N}^{s}_{\rho}(\mathcal{N}^{s}_{\rho}-1)]=\int\int_{B_{\rho}\times B_{\rho}}{\mathrm{{K}}}^{s,s}_{2}(z,w)\,\mathrm{d}z\;\mathrm{d}w,

where

K2s,s(z,w)=r2ϕ(ψ(z),ψ(w))((0,0)),(0,0))𝔼(0)[|detHψ(z)||detHψ(w)| 1{detHψ(z)<0}  1{detHψ(w)<0}].{\mathrm{K}}^{s,s}_{2}(z,w)=r^{2}\;\phi_{(\nabla\psi(z),\nabla\psi(w))}((0,0)),(0,0))\;\;\mathbb{E}^{(0)}\left[\;|\det H_{\psi}(z)|\;|\det H_{\psi}(w)|\;\mathbf{1}_{\{\det H_{\psi}(z)<0\}}\;\;\mathbf{1}_{\{\det H_{\psi}(w)<0\}}\right].

The difference is hence on the sign of the determinants, K2s,s(z,w){\mathrm{K}}^{s,s}_{2}(z,w) becomes

K2s,s(z,w)=r2ϕ(ψ(z),ψ(w))((0,0),(0,0))ar{\mathrm{K}}^{s,s}_{2}(z,w)=r^{2}\;\phi_{(\nabla\psi(z),\nabla\psi(w))}((0,0),(0,0))\;a_{r}^{\prime}\\ (51)

where (see (6.1))

ar:=\displaystyle a_{r}^{\prime}:= 𝔼(r)[|A12g(r)|Ir]\displaystyle\mathbb{E}^{(r)}\left[\;\big{|}{A_{1}}^{2}-g(r)\big{|}\;I_{r}^{\prime}\right]
Ir:=\displaystyle I^{\prime}_{r}:= 𝟏{A1+rBr+r2Cr<0}𝟏{A1+rBr+r2Cr<0}.\displaystyle\mathbf{1}_{\{A_{1}+rB_{r}+r^{2}C_{r}<0\}}\mathbf{1}_{\{-A_{1}+rB_{r}^{\prime}+r^{2}C^{\prime}_{r}<0\}}.

The asymetry of the expression of the determinant yields a different estimate than in the previous case. To be able to prove (24), we need to establish an upper bound and a lower bound of ara_{r}^{\prime} as in the previous section (Lemma 3). We give in Lemma 4 an asymptotic expression of ara_{r}^{\prime}.

The proof is similar but there are also significant differences. The difference with respect to before is that the two signs of the determinants are negative, hence we replace JrJ_{r} by

Jr=𝟏{A1+rB0<0,A1+rB0<0}=𝟏{|A1|rB0}\displaystyle J_{r}^{\prime}=\mathbf{1}_{\{A_{1}+rB_{0}<0,-A_{1}+rB_{0}<0\}}=\mathbf{1}_{\{|A_{1}|\leqslant-rB_{0}\}}

and emphasize that B0B_{0} does not have the same law as B0.-B_{0}.

Lemma 4.

We have for 0<α<α,0<\alpha^{\prime}<\alpha,

|ar𝔼(r)(|A12r2B02|Jr)|=\displaystyle|a_{r}^{\prime}-\mathbb{E}^{(r)}(|A_{1}^{2}-r^{2}B_{0}^{2}|J_{r}^{\prime})|= 𝒪(r3+α)\displaystyle\mathcal{O}_{\mathbb{P}}(r^{3+\alpha^{\prime}})
ar\displaystyle a_{r}^{\prime}\asymp 𝔼(|A12r2B02|Jr)\displaystyle\mathbb{E}(|A_{1}^{2}-r^{2}B_{0}^{2}|J^{\prime}_{r})

The proof is omitted as the proof of Lemma 3 can be repdroduced verbatim, with resp. Jr,Ir,arJ_{r}^{\prime},I_{r}^{\prime},a_{r}^{\prime} in place of resp. Jr,Ir,ar.J_{r},I_{r},a_{r}.

6.5.1. Upper bound

The upper bound on JrJ_{r}^{\prime} is of different nature than that on JrJ_{r}, in particular the third term

Jr=𝟏{|A1|<rB0}𝟏{|22ψ(0)|<6r221ψ(0)}+𝟏{|111ψ(0)|<2r1111ψ(0)}+𝟏{|22ψ(0)111|<3r211ψ(0)2}.\displaystyle J_{r}^{\prime}=\mathbf{1}_{\{|A_{1}|<-rB_{0}\}}\leqslant\mathbf{1}_{\{|\partial_{22}\psi(0)|<-6r\partial_{221}\psi(0)\}}+\mathbf{1}_{\{|\partial_{111}\psi(0)|<-{2r\partial_{1111}\psi(0)}\}}+\mathbf{1}_{\{|\partial_{22}\psi(0)\partial_{111}|<3r\partial_{211}\psi(0)^{2}\}}.

Then,

𝔼(|A1r2B02|Jr)𝔼(2r2B02Jr)\displaystyle\mathbb{E}(|A_{1}-r^{2}B_{0}^{2}|J_{r}^{\prime})\leqslant\mathbb{E}(2r^{2}B_{0}^{2}J_{r^{\prime}})

hence we must use this time Lemma 5-(ii) for the last term of JrJ_{r}^{\prime}’s bound,

𝔼(B02𝟏|22ψ(0)111ψ(0)|<3r211ψ(0)2)\displaystyle\mathbb{E}(B_{0}^{2}\mathbf{1}_{|\partial_{22}\psi(0)\partial_{111}\psi(0)|<3r\partial_{211}\psi(0)^{2}})\leqslant 𝔼(|22ψ(0)111ψ(0)|𝟏{|22ψ(0)111ψ(0)|<3r211ψ(0)2})\displaystyle\mathbb{E}(|\partial_{22}\psi(0)\partial_{111}\psi(0)|\mathbf{1}_{\{|\partial_{22}\psi(0)\partial_{111}\psi(0)|<3r\partial_{211}\psi(0)^{2}\}})
+𝔼(|211ψ(0)|2𝟏{|22ψ(0)111ψ(0)|<3r211ψ(0)2}))\displaystyle+\mathbb{E}(|\partial_{211}\psi(0)|^{2}\mathbf{1}_{\{|\partial_{22}\psi(0)\partial_{111}\psi(0)|<3r\partial_{211}\psi(0)^{2}\}}))
+𝔼(|22ψ(0)1111ψ(0)𝟏{|22ψ(0)111ψ(0)|<3r211ψ(0)2}))\displaystyle+\mathbb{E}(|\partial_{22}\psi(0)\partial_{1111}\psi(0)\mathbf{1}_{\{|\partial_{22}\psi(0)\partial_{111}\psi(0)|<3r\partial_{211}\psi(0)^{2}\}}))
\displaystyle\leqslant crln(r)\displaystyle cr\ln(r)

for some c<.c<\infty. The other terms are dealt with by Lemma 5-(iii) as in (47), hence the upper bound is in

𝔼(|A1r2B02|Jr)cr3ln(r)\displaystyle\mathbb{E}(|A_{1}-r^{2}B_{0}^{2}|J_{r}^{\prime})\leqslant c^{\prime}r^{3}\ln(r)

for some c<,c^{\prime}<\infty, which yields (25) by (51) and Lemma 4.

6.5.2. Lower bound

We recall the expression of A1A_{1} and B0:-B_{0}: A1=22ψ(0)111ψ(0),B0=221ψ(0)111ψ(0)+211ψ(0)21322ψ(0)1111ψ(0).A_{1}=\partial_{22}\psi(0)\partial_{111}\psi(0),\;-B_{0}=-\partial_{221}\psi(0)\;\partial_{111}\psi(0)+\partial_{211}\psi(0)^{2}-\frac{1}{3}\partial_{22}\psi(0)\;\partial_{1111}\psi(0). The strategy is the same than at Section 6.4.2.

If ψ\psi is a SGRW (Example 1), 111ψ(0)=122ψ(0)\partial_{111}\psi(0)=-\partial_{122}\psi(0) if Y0=0Y_{0}=0, let

Ω={211ψ(0)>2,|22ψ(0)111ψ(0)|<r,|111ψ(0)|<1,|22ψ(0)|<1,|1111ψ(0)|<1}.\displaystyle\Omega=\{\partial_{211}\psi(0)>2,|\partial_{22}\psi(0)\partial_{111}\psi(0)|<r,|\partial_{111}\psi(0)|<1,|\partial_{22}\psi(0)|<1,|\partial_{1111}\psi(0)|<1\}.

Hence if Y0=0Y_{0}=0 and Ω\Omega is realised

A1\displaystyle A_{1} <r,\displaystyle<r,
B0\displaystyle-B_{0} =111ψ(0)2+211ψ(0)21322ψ(0)1111ψ(0)>0+413>2A1/r.\displaystyle=\partial_{111}\psi(0)^{2}+\partial_{211}\psi(0)^{2}-\frac{1}{3}\partial_{22}\psi(0)\partial_{1111}\psi(0)>0+4-\frac{1}{3}>2A_{1}/r.

We have

𝔼(|A12r2B02|𝟏{|A1|rB0})\displaystyle\mathbb{E}(|A_{1}^{2}-r^{2}B_{0}^{2}|\mathbf{1}_{\{|A_{1}|\leqslant-rB_{0}\}}) 𝔼[𝟏Ωr2B02/4𝟏{B02A1/r}]\displaystyle\geqslant\mathbb{E}\left[\mathbf{1}_{\Omega}\mid r^{2}B_{0}^{2}/4\mid\mathbf{1}_{\{B_{0}\geqslant 2A_{1}/r\}}\right]
r2(Ω).\displaystyle\geqslant\;r^{2}\mathbb{P}(\Omega).

We must prove a converse to Lemma 5-(i) with s=0s=0. Since the density of XX^{\prime} is uniformly bounded from below on [3,3]4[-3,3]^{4}, for some c>0,c>0,

(Ω)c[1,1]2𝟏{|x1x2|<r}𝑑x1𝑑x2rln(r).\displaystyle\mathbb{P}(\Omega)\geqslant c\int_{[-1,1]^{2}}\mathbf{1}_{\{|x_{1}x_{2}|<r\}}dx_{1}dx_{2}\asymp r\ln(r).

6.6. Proof of (26) in Theorem 5.1

We recall that 𝒩ρc=𝒩ρs+𝒩ρe\mathcal{N}_{\rho}^{c}=\mathcal{N}_{\rho}^{s}+\mathcal{N}_{\rho}^{e} hence 𝒩ρc(𝒩ρc1)=𝒩ρe(𝒩ρe1)+𝒩ρs(𝒩ρs1)+2𝒩ρe𝒩ρs.\mathcal{N}_{\rho}^{c}(\mathcal{N}_{\rho}^{c}-1)=\mathcal{N}_{\rho}^{e}(\mathcal{N}_{\rho}^{e}-1)+\mathcal{N}_{\rho}^{s}(\mathcal{N}_{\rho}^{s}-1)+2\mathcal{N}_{\rho}^{e}\mathcal{N}_{\rho}^{s}.
So, we have:

𝔼[𝒩ρe𝒩ρs]\displaystyle\mathbb{E}[\mathcal{N}_{\rho}^{e}\mathcal{N}_{\rho}^{s}] =12𝔼[𝒩ρc(𝒩ρc1)]𝔼[𝒩ρe(𝒩ρe1)]𝔼[𝒩ρs(𝒩ρs1)].\displaystyle=\frac{1}{2}\mathbb{E}[\mathcal{N}_{\rho}^{c}(\mathcal{N}_{\rho}^{c}-1)]-\mathbb{E}[\mathcal{N}_{\rho}^{e}(\mathcal{N}_{\rho}^{e}-1)]-\mathbb{E}[\mathcal{N}_{\rho}^{s}(\mathcal{N}_{\rho}^{s}-1)].

Combining this formula with previous estimates (23), (24) and (25), we obtain

𝔼[𝒩ρe𝒩ρs]\displaystyle\mathbb{E}[\mathcal{N}_{\rho}^{e}\mathcal{N}_{\rho}^{s}] =12𝔼[𝒩ρc(𝒩ρc1)]+o(1)\displaystyle=\frac{1}{2}\mathbb{E}[\mathcal{N}_{\rho}^{c}(\mathcal{N}_{\rho}^{c}-1)]+o(1)

ending the proof of (26).

Lemma 5.

Let (Z1,,Zk)(Z_{1},\dots,Z_{k}) be a non-degenerate Gaussian vector and ai,ja_{i,j} real fixed coefficients. Then,

  • (i)
    (|Z1Z2+si,jai,jZiZj|<r)Crln(r)\displaystyle\mathbb{P}(|Z_{1}Z_{2}+s\sum_{i,j}a_{i,j}Z_{i}Z_{j}|<r)\leqslant Cr\ln(r) (i)

    for CC depending on the law of the ZiZ_{i} (and not on ss or the ai,ja_{i,j}),

  • (ii)

    for αi0\alpha_{i}\geqslant 0

    𝔼(|Z1α1Zkαk|𝟏{|Z1Z2|<rZ32})C{rln(r) if α1=α2=0r otherwise\displaystyle\mathbb{E}(|Z_{1}^{\alpha_{1}}\dots Z_{k}^{\alpha_{k}}|\mathbf{1}_{\{|Z_{1}Z_{2}|<rZ_{3}^{2}\}})\leqslant C^{\prime}\begin{cases}r\ln(r)$ if $\alpha_{1}=\alpha_{2}=0\\ r$ otherwise$\end{cases} (ii)

    for some C<.C^{\prime}<\infty.

  • (iii)

    Let some coefficients αi\alpha_{i}\in\mathbb{N} , (Z1,,Zq)(Z_{1},\dots,Z_{q}) be a Gaussian vector. Then, for some C′′<,C^{\prime\prime}<\infty,

    𝔼(|Z1α1Zqαq|𝟏{|Z1|rZ2})C′′r.\displaystyle\mathbb{E}(|Z_{1}^{\alpha_{1}}\dots Z_{q}^{\alpha_{q}}|\mathbf{1}_{\{|Z_{1}|\leqslant rZ_{2}\}}\ )\leqslant C^{\prime\prime}r. (iii)
Proof.

(i) Assume first that the ZiZ_{i} are iid Gaussian. Let us study for a,ba,b\in\mathbb{R}, Y1:=Z1as,Y2:=Z2bsY_{1}:=Z_{1}-as,Y_{2}:=Z_{2}-bs. Since Y1,Y2Y_{1},Y_{2} have a density bounded by κ<\kappa<\infty (universal), we have for cc\in\mathbb{R}

(|Y1Y2c|r)\displaystyle\mathbb{P}(|Y_{1}Y_{2}-c|\leqslant r)\leqslant (|Y2|r)+(|Y1c/Y2|<r/Y2,|Y2|>1)+(|Y1c/Y2|<r/Y2,|Y2|[r,1])\displaystyle\mathbb{P}(|Y_{2}|\leqslant r)+\mathbb{P}(|Y_{1}-c/Y_{2}|<r/Y_{2},|Y_{2}|>1)+\mathbb{P}(|Y_{1}-c/Y_{2}|<r/Y_{2},|Y_{2}|\in[r,1])
\displaystyle\leqslant κr+(|Y1c/Y2|<r)+𝔼[(Y1[c/Y2±r/Y2]|Y2)𝟏{r<|Y2|<1}]\displaystyle\kappa r+\mathbb{P}(|Y_{1}-c/Y_{2}|<r)+\mathbb{E}\left[\mathbb{P}(Y_{1}\in[c/Y_{2}\pm r/Y_{2}]\;|\;Y_{2})\mathbf{1}_{\{r<|Y_{2}|<1\}}\right]
\displaystyle\leqslant κr+κr+𝔼(κr/Y2𝟏{r<|Y2|<1})\displaystyle\kappa r+\kappa r+\mathbb{E}(\kappa r/Y_{2}\mathbf{1}_{\{r<|Y_{2}|<1\}})
\displaystyle\leqslant 2κr+κrr11y22κ𝑑y2\displaystyle 2\kappa r+\kappa r\int_{r}^{1}\frac{1}{y_{2}}2\kappa dy_{2}
\displaystyle\leqslant 2κr+2κ2rln(r),\displaystyle 2\kappa r+2\kappa^{2}r\ln(r),

uniformly on a,b,c,s.a,b,c,s. Then it remains to notice that

Z1Z2+si,jai,jZiZj=(Z1As)(Z2Bs)Cs\displaystyle Z_{1}Z_{2}+s\sum_{i,j}a_{i,j}Z_{i}Z_{j}=(Z_{1}-As)(Z_{2}-Bs)-C_{s}

where A,B,CsA,B,C_{s} are independent of Z1,Z2Z_{1},Z_{2}. Then

(|Z1Z2+si,jai,jZiZj|<r)=𝔼((|(Z1As)(Z2Bs)Cs||A,B,Cs))Crlnr.\displaystyle\mathbb{P}(|Z_{1}Z_{2}+s\sum_{i,j}a_{i,j}Z_{i}Z_{j}|<r)=\mathbb{E}(\mathbb{P}(|(Z_{1}-As)(Z_{2}-Bs)-C_{s}|\;|\;A,B,C_{s}))\leqslant Cr\ln r.

In the non-independent Gaussian case, the joint density f(x1,,xk)f(x_{1},\dots,x_{k}) of (Z1,,Zk)(Z_{1},\dots,Z_{k}) is bounded by κexp(cixi2)\kappa\exp(-c\sum_{i}x_{i}^{2}) for some c,κ>0c,\kappa>0 (cc would be the smallest eigenvalue of the covariance matrix). From there on the conclusion is easy:

𝟏{|x1x2+si,jai,jxixj|<r}f(x1,,xk)𝑑x1𝑑xkκ𝟏{}exp(cixi2)𝑑x1𝑑xk\displaystyle\int\mathbf{1}_{\{|x_{1}x_{2}+s\sum_{i,j}a_{i,j}x_{i}x_{j}|<r\}}f(x_{1},\dots,x_{k})dx_{1}\dots dx_{k}\leqslant\kappa\int\mathbf{1}_{\{...\}}\exp(-c\sum_{i}x_{i}^{2})dx_{1}\dots dx_{k}

and the right hand member corresponds to the independent case, already treated.

(ii) For the second assertion, assume first that the ZkZ_{k} are independent. Without loss of generality, assume α1α2\alpha_{1}\leqslant\alpha_{2}. We have for t0t\geqslant 0, for some c,c,c′′,c′′′′,C<,c,c^{\prime},c^{\prime\prime},c^{\prime\prime\prime\prime},C<\infty,

𝔼(|Z1α1Z2α2|𝟏|Z1Z2|<t)\displaystyle\mathbb{E}(|Z_{1}^{\alpha_{1}}Z_{2}^{\alpha_{2}}|\mathbf{1}_{|Z_{1}Z_{2}|}<t)\leqslant 𝔼(|Z1α1Z2α2|𝟏|Z1|<t)+𝔼(|Z1α1Z2α2|𝟏{|Z2|<t,|Z1|>1})+𝔼(|Z1α1Z2α2|𝟏{|Z2|<t/|Z1|}𝟏{t<|Z1|<1})\displaystyle\mathbb{E}(|Z_{1}^{\alpha_{1}}Z_{2}^{\alpha_{2}}|\mathbf{1}_{|Z_{1}|<t})+\mathbb{E}(|Z_{1}^{\alpha_{1}}Z_{2}^{\alpha_{2}}|\mathbf{1}_{\{|Z_{2}|<t,|Z_{1}|>1\}})+\mathbb{E}(|Z_{1}^{\alpha_{1}}Z_{2}^{\alpha_{2}}|\mathbf{1}_{\{|Z_{2}|<t/|Z_{1}|\}}\mathbf{1}_{\{t<|Z_{1}|<1\}})
\displaystyle\leqslant ctα1+1+ctα2+1+ct1x1α10t/x1x2α2𝑑x2𝑑x1\displaystyle ct^{\alpha_{1}+1}+ct^{\alpha_{2}+1}+c^{\prime}\int_{t}^{1}x_{1}^{\alpha_{1}}\int_{0}^{t/x_{1}}x_{2}^{\alpha_{2}}dx_{2}dx_{1}
\displaystyle\leqslant 2ct+c′′t1x1α1(tx1)α2+1𝑑x1\displaystyle 2ct+c^{\prime\prime}\int_{t}^{1}x_{1}^{\alpha_{1}}\left(\frac{t}{x_{1}}\right)^{\alpha_{2}+1}dx_{1}
\displaystyle\leqslant 2ct+c′′′tα2+1{tα1α2 if α1<α2ln(t) if α1=α2\displaystyle 2ct+c^{\prime\prime\prime}t^{\alpha_{2}+1}\begin{cases}t^{\alpha_{1}-\alpha_{2}}$ if $\alpha_{1}<\alpha_{2}\\ \ln(t)$ if $\alpha_{1}=\alpha_{2}\end{cases}
\displaystyle\leqslant C{tln(t) if α1=α2=0t otherwise.\displaystyle C\begin{cases}t\ln(t)$ if $\alpha_{1}=\alpha_{2}=0\\ t$ otherwise$\end{cases}.

Coming back to the main estimate with t=rZ32t=rZ_{3}^{2}, using conditional expectations, for some C,C′′<,C^{\prime},C^{\prime\prime}<\infty,

𝔼(|Z1α1Zkαk|𝟏{|Z1Z2|<rZ32})C𝔼(i1,2,3Zkαk(rZ3α3+2ln(rZ3)𝟏α1=α2=0))C′′rln(r)𝟏α1=α2=0.\displaystyle\mathbb{E}(|Z_{1}^{\alpha_{1}}\dots Z_{k}^{\alpha_{k}}|\mathbf{1}_{\{|Z_{1}Z_{2}|<rZ_{3}^{2}\}})\leqslant C^{\prime}\mathbb{E}(\prod_{i\neq 1,2,3}Z_{k}^{\alpha_{k}}(rZ_{3}^{\alpha_{3}+2}\ln(rZ_{3})^{\mathbf{1}_{\alpha_{1}=\alpha_{2}=0}}))\leqslant C^{\prime\prime}r\ln(r)^{\mathbf{1}_{\alpha_{1}=\alpha_{2}=0}}.

The non-independent (non-degenerate) case can be treated as before by bounding the density of the ZkZ_{k} by an independent density of the same order.

(iii) By Holder’s inequality

𝔼(|Z1α1Zqαq|𝟏{|Z1|crZ2})i=1q𝔼(|Zi|qαi𝟏{|Z1|crZ2})1q\displaystyle\mathbb{E}(|Z_{1}^{\alpha_{1}}\dots Z_{q}^{\alpha_{q}}|\mathbf{1}_{\{|Z_{1}|\leqslant crZ_{2}\}}\ )\leqslant\prod_{i=1}^{q}\mathbb{E}(|Z_{i}|^{q\alpha_{i}}\mathbf{1}_{\{|Z_{1}|\leqslant crZ_{2}\}})^{\frac{1}{q}}

hence we can assume wlog that only one αi\alpha_{i}, say αi0\alpha_{i_{0}}, is non-zero. For i0>2,i_{0}>2, we have an orthogonal decomposition of the form Zi0=(αZ1+βZ2)+γYZ_{i_{0}}=(\alpha Z_{1}+\beta Z_{2})+\gamma Y where YY is independent of (Z1,Z2)(Z_{1},Z_{2}), hence we can assume wlog that i0=1i_{0}=1 or i0=2i_{0}=2. For i0=1i_{0}=1, the bound is

𝔼(|rZ2|α1𝟏{|Z1|<rZ2})=O(r1+α1)\displaystyle\mathbb{E}(|rZ_{2}|^{\alpha_{1}}\mathbf{1}_{\{|Z_{1}|<rZ_{2}\}})=O(r^{1+\alpha_{1}})

and it only remains to treat the case i0=2i_{0}=2. In this case we decompose orthogonally Z1=λZ2+μZZ_{1}=\lambda Z_{2}+\mu Z where ZZ is independent of Z2Z_{2}. Then the bounded densities of Z2Z_{2} and ZZ easily yields the result

𝔼(|Z2|α2C|rZ2|)=O(r).\displaystyle\mathbb{E}(|Z_{2}|^{\alpha_{2}}C|rZ_{2}|)=O(r).

Acknowledgements

We warmfully thank Anne Estrade, who participated to the conception of the project, the supervision of this work and to the elaboration of this article. This project has received funding from the European Union’s Horizon 2020 research and innovation programme under the Marie Sklodowska-Curie grant agreement No 754362.

[Uncaptioned image]

References

  • AD [22] J. Azais and C. Delmas. Mean number and correlation function of critical points of isotropic Gaussian fields. Stoc. Proc. Appl., 150:411–445, 2022.
  • AT [09] R. J. Adler and J. E. Taylor. Random fields and geometry. Springer Science & Business Media, 2009.
  • ATW [07] R. J. Adler, J. E. Taylor, and K. Worsley. Applications of random fields and geometry: Foundations and case studies. In In preparation, available on R. Adler home page. Citeseer, 2007.
  • AW [09] J. Azaïs and M. Wschebor. Level sets and extrema of random processes and fields. John Wiley & Sons, 2009.
  • BCW [19] D. Beliaev, V. Cammarota, and I. Wigman. Two point function for critical points of a random plane wave. International Mathematics Research Notices, 2019(9):2661–2689, 2019.
  • BH [20] R. Bardenet and A. Hardy. Monte carlo with determinantal point processes. Ann. Appl. Prob., 30(1):368–417, 2020.
  • CG [13] C. Chevalier and D. Ginsbourger. Fast computation of the multi-points expected improvement with applications in batch selection. In International Conference on Learning and Intelligent Optimization, pages 59–69. Springer, 2013.
  • CS [17] D. Cheng and A. Schwartzman. Multiple testing of local maxima for detection of peaks in random fields. Annals of statistics, 45(2):529, 2017.
  • CX [16] D. Cheng and Y. Xiao. The mean Euler characteristic and excursion probability of Gaussian random fields with stationary increments. The Annals of Applied Probability, 26(2):722–759, 2016.
  • DGL [17] A. Desolneux, B. Galerne, and C. Launay. Etude de la répulsion des processus pixelliques déterminantaux. Juan les pins, France, September 2017.
  • Fyo [04] Y. V. Fyodorov. Complexity of random energy landscapes, glass transition, and absolute value of the spectral determinant of random matrices. Physical review letters, 92(24):240601, 2004.
  • KT [12] A. Kulesza and B. Taskar. Determinantal point processes for machine learning. Found. Trends Mach. learn., 5(2-3), 2012.
  • LGD [21] C. Launay, B. Galerne, and A. Desolneux. Determinantal point processes for image processing. SIAM Journal on Imaging Sciences, 14(1), 2021.
  • Lin [72] G. Lindgren. Local maxima of Gaussian fields. Arkiv för matematik, 10(1-2):195–218, 1972.
  • LW [04] D. L. Larson and B. D. Wandelt. The hot and cold spots in the wilkinson microwave anisotropy probe data are not hot and cold enough. The Astrophysical Journal Letters, 613(2):L85, 2004.
  • [16] S. Muirhead, A. Rivera, and H. Vanneuville. The phase transition for planar Gaussian percolation models without fkg. https://arxiv.org/abs/2010.11770.
  • Mui [20] S. Muirhead. A second moment bound for critical points of planar Gaussian fields in shrinking height windows. Statistics & Probability Letters, 160:108698, 2020.
  • NH [03] T. Nichols and S. Hayasaka. Controlling the familywise error rate in functional neuroimaging: a comparative review. Statistical methods in medical research, 12(5):419–446, 2003.
  • NPR [19] I. Nourdin, G. Peccati, and M. Rossi. Nodal statistics of planar random waves. Comm. Math. Phys., 369:99–151, 2019.
  • Pot [09] J. Potthof. Sample properties of random fields. ii. continuity. Comm. Stoc. Anal., 3(3):331–348, 2009.
  • TBA [19] N. Tremblay, S. Barthelmé, and P. Amblard. Determinantal point processes for coresets. J. Mach. Learn. Res., 20:1–70, 2019.
  • TW [07] J. E. Taylor and K. J. Worsley. Detecting sparse signals in random fields, with an application to brain mapping. Journal of the American Statistical Association, 102(479):913–928, 2007.
  • WMNE [96] K. J. Worsley, S. Marrett, P. Neelin, and A.C. Evans. Searching scale space for activation in pet images. Human brain mapping, 4(1):74–90, 1996.
  • WTTL [04] K. J. Worsley, J. E. Taylor, T. F. Tomaiuolo, and J. Lerch. Unified univariate and multivariate random field theory. Neuroimage, 23:S189–S195, 2004.