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

Brownian snails with removal:
epidemics in diffusing populations

Geoffrey R. Grimmett (GRG) Statistical Laboratory, Centre for Mathematical Sciences, Cambridge University, Wilberforce Road, Cambridge CB3 0WB, UK School of Mathematics & Statistics, The University of Melbourne, Australia g.r.grimmett@statslab.cam.ac.uk http://www.statslab.cam.ac.uk/~grg/ (ZL) Department of Mathematics, University of Connecticut, Storrs, Connecticut 06269-3009, USA zhongyang.li@uconn.edu http://www.math.uconn.edu/~zhongyang/  and  Zhongyang Li
(Date: 20 October 2020, revised 6 April 2022)
Abstract.

Two stochastic models of susceptible/infected/removed (SIR) type are introduced for the spread of infection through a spatially-distributed population. Individuals are initially distributed at random in space, and they move continuously according to independent diffusion processes. The disease may pass from an infected individual to an uninfected individual when they are sufficiently close. Infected individuals are permanently removed at some given rate α\alpha. Such processes are reminiscent of so-called frog models, but differ through the action of removal, as well as the fact that frogs jump whereas snails slither.

Two models are studied here, termed the ‘delayed diffusion’ and the ‘diffusion’ models. In the first, individuals are stationary until they are infected, at which time they begin to move; in the second, all individuals start to move at the initial time 0. Using a perturbative argument, conditions are established under which the disease infects a.s. only finitely many individuals. It is proved for the delayed diffusion model that there exists a critical value αc(0,)\alpha_{\text{\rm c}}\in(0,\infty) for the survival of the epidemic.

Key words and phrases:
Percolation, infectious disease, SIR model, frog model, snail model, epidemic, diffusion, Wiener sausage
2010 Mathematics Subject Classification:
60K35, 60G15

1. Introduction

1.1. Outline of the models

Numerous mathematical models have been introduced to describe the spread of a disease around a population. Such models may be deterministic or stochastic, or a mixture of each; they may incorporate a range of factors including susceptibility, infectivity, recovery, and removal; the population members (termed ‘particles’) may be distributed about some given space; and so on. We propose two models in which (i) the particles move randomly about the space that they inhabit, (ii) infection may be passed between particles that are sufficiently close to one another, and (iii) after the elapse of a random time since infection, a particle is removed from the process. These models differ from that of Beckman, Dinan, Durrett, Huo, and Junge [3] through the introduction of the permanent ‘removal’ of particles, and this new feature brings a significant new difficulty to the analysis.

We shall concentrate mostly on the case in which the particles inhabit d{\mathbb{R}}^{d} where d2d\geq 2. Here is a concrete example of the processes studied here.

  • (a)

    Particles are initially distributed in d{\mathbb{R}}^{d} in the manner of a rate-λ\lambda Poisson process conditioned to contain a point at the origin 0.

  • (b)

    Particles move randomly within d{\mathbb{R}}^{d} according to independent Brownian motions with variance-parameter σ2\sigma^{2}.

  • (c)

    At time 0 the particle at the origin (the initial ‘infective’) suffers from an infectious disease, which may be passed to others when sufficiently close.

  • (d)

    When two particles, labelled PP and PP^{\prime}, are within a given distance δ\delta, and PP is already infected, then particle PP^{\prime} becomes infected.

  • (e)

    Each particle is infected for a total period of time having the exponential distribution with parameter α[0,)\alpha\in[0,\infty), and is then permanently removed.

The fundamental question is to determine for which vectors (λ,δ,σ,α)(\lambda,\delta,\sigma,\alpha) it is the case that (with strictly positive probability) infinitely many particles become infected. For simplicity, we shall assume henceforth that

(1.1) δ=σ=1.\delta=\sigma=1.

We shall generally assume α>0\alpha>0. In the special case α=0\alpha=0, (studied in [3]) a particle once infected remains infected forever, and the subsequent analysis is greatly facilitated by a property of monotonicity that is absent in the more challenging case α>0\alpha>0 considered in the current work.

Two protocols for movement feature in this article.

  1. A.

    Delayed diffusion model. The initial infective starts to move at time 0, and all other particles remain stationary until they are infected, at which times they begin to move.

  2. B.

    Diffusion model. All particles begin to move at time 0.

The main difficulty in studying these models arises from the fact that particles are permanently removed after a (random) period of infectivity. This introduces a potential non-monotonicity into the model, namely that the presence of infected particles may hinder the growth of the process through the creation of islands of ‘removed’ particles that can act as barriers to the further spread of infection. A related situation (but without the movement of particles) was considered by Kuulasma [21] in a discrete setting, and the methods derived there are useful in our Section 3.7 (see also Alves et al. [1, p. 4]). This issue may be overcome for the delayed diffusion model, but remains problematic in the case of the diffusion process.

Let II denote the set of particles that are ever infected, and

(1.2) θ(λ,α):=λ,α(|I|=).\theta(\lambda,\alpha):={\mathbb{P}}_{\lambda,\alpha}(|I|=\infty).

We say the process

becomes extinct if θ(λ,α)=0,\displaystyle\text{if }\theta(\lambda,\alpha)=0,
survives if θ(λ,α)>0.\displaystyle\text{if }\theta(\lambda,\alpha)>0.

Let λc\lambda_{\text{\rm c}} denote the critical value of λ\lambda for the disk (or ‘Boolean’) percolation model with radius 11 on d{\mathbb{R}}^{d} (see, for example, [25]). It is immediate for both models above that

(1.3) θ(λ,α)>0if λ>λc and α0,\theta(\lambda,\alpha)>0\qquad\text{if $\lambda>\lambda_{\text{\rm c}}$ and $\alpha\geq 0$},

since in that case the disease spreads instantaneously to the percolation cluster CC containing the initial infective, and in addition we have λ,α(|C|=)>0{\mathbb{P}}_{\lambda,\alpha}(|C|=\infty)>0.

1.2. Two exemplars of results

We write θd\theta_{\text{\rm d}} (respectively, θdd\theta_{\text{\rm dd}}) for the function θ\theta of (1.2) in the case of the diffusion model (respectively, delayed diffusion model). The following two theorems are proved in Sections 3 and 4 as special cases of results for more general epidemic models than those given above.

Theorem 1.1 (Brownian delayed diffusion model).

Let d2d\geq 2. There exists a non-decreasing function αc:(0,)(0,]\alpha_{\text{\rm c}}:(0,\infty)\to(0,\infty] such that

(1.4) θdd(λ,α){=0if α>αc(λ),>0if α<αc(λ).\theta_{\text{\rm dd}}(\lambda,\alpha)\begin{cases}=0&\text{if }\alpha>\alpha_{\text{\rm c}}(\lambda),\\ >0&\text{if }\alpha<\alpha_{\text{\rm c}}(\lambda).\end{cases}

Furthermore, αc(λ)=\alpha_{\text{\rm c}}(\lambda)=\infty when λ>λc\lambda>\lambda_{\text{\rm c}}, and there exists λ¯(0,λc]\underline{\lambda}\in(0,\lambda_{\text{\rm c}}] such that αc(λ)<\alpha_{\text{\rm c}}(\lambda)<\infty when 0<λ<λ¯0<\lambda<\underline{\lambda}.

The delayed diffusion model has no phase transition when d=1d=1; see Theorems 3.2 and 3.6.

Theorem 1.2 (Brownian diffusion model).

Let d1d\geq 1. There exists λ¯(0,λc]\underline{\lambda}\in(0,\lambda_{\text{\rm c}}] and a non-decreasing function α¯:(0,λ¯)(0,)\underline{\alpha}:(0,\underline{\lambda})\to(0,\infty) such that θd(λ,α)=0\theta_{\text{\rm d}}(\lambda,\alpha)=0 when α>α¯(λ)\alpha>\underline{\alpha}(\lambda) and 0<λ<λ¯0<\lambda<\underline{\lambda}.

For the diffusion model, we have no proof of survival for d2d\geq 2 and small positive α\alpha (that is, that θd(λ,α)>0\theta_{\text{\rm d}}(\lambda,\alpha)>0 for some λ<λc\lambda<\lambda_{\text{\rm c}} and α>0\alpha>0), and neither does the current work answer the question of whether or not survival ever occurs when d=1d=1. See Section 4.3. The above theorems are proved using a perturbative argument, and thus fall short of the assertion that λ¯=λc\underline{\lambda}=\lambda_{\text{\rm c}}.

The methods of proof may be made quantitative, leading to bounds for the numerical values of the critical points αc\alpha_{\text{\rm c}}. Such bounds are far from precise, and therefore we do not explore them here. Our basic estimates for the growth of infection hold if the intensity λ\lambda of the Poisson process is non-constant so long as it is bounded uniformly between two strictly positive constants. The existence of the subcritical phase may be proved for more general diffusions than Brownian motion.

1.3. Literature and notation

The related literature is somewhat ramified, and a spread of related problems have been studied by various teams. We mention a selection of papers but do not attempt a full review, and we concentrate on works associated with d{\mathbb{R}}^{d} rather than with trees or complete graphs.

The delayed diffusion model may be viewed as a continuous-time version of the ‘frog’ random walk process studied in Alves et al. [1, 2], Ramirez and Sidoravicius [29], Fontes et al.  [7], Benjamini et al. [4], and Hoffman, Johnson, and Junge [15, 16]. See Popov [28] for an early review. Kesten and Sidoravicius [17, 18, 19] considered a variant of the frog model as a model for infection, both with and without recuperation (that is, when infected frogs recover and become available for reinfection—see also Section 4.3 of the current work). The paper of Beckman et al. [3] is devoted to the delayed diffusion model without removal (that is, with α=0\alpha=0). Peres et al.  [27] studied three geometric properties of a Poissonian/Brownian cloud of particles, in work inspired in part by the dynamic Boolean percolation model of van den Berg et al. [5]. Related work has appeared in Gracar and Stauffer [9].

A number of authors have considered the frog model with recuperation under the title ‘activated random walks’. The reader is referred to the review by Rolla [30], and for recent work to Stauffer and Taggi [33] and Rolla et al. [31].

We write 0={0,1,2,}{\mathbb{Z}}_{0}=\{0,1,2,\dots\} and 1A1_{A} (or 1(A)1(A)) for the indicator function of an event or set AA. Let S(r)S(r) denote the closed rr-ball of d{\mathbb{R}}^{d} with centre at the origin, and S=S(1)S=S(1). The dd-dimensional Lebesgue measure of a set AA is written |A|d|A|_{d}, and the Euclidean norm d\|\cdot\|_{d}. The radius of MdM\subseteq{\mathbb{R}}^{d} is defined by

rad(M):=sup{md:mM}.\text{\rm rad}(M):=\sup\{\|m\|_{d}:m\in M\}.

We abbreviate λ,α{\mathbb{P}}_{\lambda,\alpha} (respectively, 𝔼λ,α{\mathbb{E}}_{\lambda,\alpha}) to the generic notation {\mathbb{P}} (respectively, 𝔼{\mathbb{E}}).

The contents of this paper are as follows. The two models are defined in Section 2 with a degree of generality that includes general diffusions and a more general process of infection. The delayed diffusion model is studied in Section 3, and the diffusion model in Section 4. Theorem 1.1 (respectively, Theorem 1.2) is contained within Theorem 3.1 (respectively, Theorem 4.1).

1.4. Open problems

This introduction closes with a short account of some of the principal remaining open problems. For concreteness, we restrict ourselves to the Brownian models of Section 1.2 without further reference to the random-walk versions of these models, and the general models of Section 2.1. This section is positioned here despite the fact that it refers sometimes to versions of the models that have not yet been fully introduced (see Section 2).

  • A.

    For the Brownian delayed diffusion model, show that the critical value αc(λ)\alpha_{\text{\rm c}}(\lambda) of Theorem 1.1 satisfies αc(λ)<\alpha_{\text{\rm c}}(\lambda)<\infty whenever λ<λc\lambda<\lambda_{\text{\rm c}}.

  • B.

    When d2d\geq 2, prove survival in the Brownian diffusion model for some λ<λc\lambda<\lambda_{\text{\rm c}} and small α>0\alpha>0. More specifically, show that θd(λ,α)>0\theta_{\text{\rm d}}(\lambda,\alpha)>0 for suitable λ\lambda and α\alpha.

  • C.

    Having resolved problem B, show the existence of a critical value αc=αc(λ)\alpha_{\text{\rm c}}=\alpha_{\text{\rm c}}(\lambda) for the Brownian diffusion model such that survival occurs when α<αc\alpha<\alpha_{\text{\rm c}} and not when α>αc\alpha>\alpha_{\text{\rm c}}. Furthermore, identify the set of λ\lambda such that αc(λ)<\alpha_{\text{\rm c}}(\lambda)<\infty.

  • D.

    Decide whether or not survival can ever occur for the Brownian diffusion model in one dimension.

  • E.

    In either model, prove a shape theorem for the set of particles that are either infected or removed at time tt.

2. General models

2.1. The general set-up

Let d1d\geq 1. A diffusion process in d{\mathbb{R}}^{d} is a solution ζ\zeta to the stochastic differential equation

(2.1) dζ(t)=a(ζ(t))dt+σ(ζ(t))dWt,d\zeta(t)=a(\zeta(t))\,dt+\sigma(\zeta(t))\,dW_{t},

where WW is a standard Brownian motion in d{\mathbb{R}}^{d}. (We may write either WtW_{t} or W(t)W(t).) For definiteness, we shall assume that: ζ(0)=0\zeta(0)=0; ζ\zeta has continuous sample paths; the instantaneous drift vector aa and variance matrix σ\sigma are locally Lipschitz continuous. We do not allow aa, σ\sigma to be time-dependent. We call the process ‘Brownian’ if ζ\zeta is a standard Brownian motion, which is to say that aa is the zero vector and σ\sigma is the identity matrix.

Let ζ\zeta be such a diffusion, and let (ζi:i0)(\zeta_{i}:i\in{\mathbb{Z}}_{0}) be independent copies of ζ\zeta. Let α(0,)\alpha\in(0,\infty), ρ[0,)\rho\in[0,\infty), and let μ:d[0,)\mu:{\mathbb{R}}^{d}\to[0,\infty) be integrable with

(2.2) Int(μ):=dμ(x)𝑑x(0,).\text{\rm Int}(\mu):=\int_{{\mathbb{R}}^{d}}\mu(x)\,dx\in(0,\infty).

We call μ\mu radially decreasing if

(2.3) μ(rx)μ(x)xd,r[1,).\mu(rx)\leq\mu(x)\qquad x\in{\mathbb{R}}^{d},\,r\in[1,\infty).

Let Π=(X0=0,X1,X2,)\Pi=(X_{0}=0,X_{1},X_{2},\dots) be a Poisson process on d{\mathbb{R}}^{d} (conditioned to possess a point at the origin 0) with constant intensity λ(0,)\lambda\in(0,\infty). At time 0, particles with label-set 𝒫={P0,P1,P2,}{\mathcal{P}}=\{P_{0},P_{1},P_{2},\dots\} are placed at the respective points X0=0,X1,X2,X_{0}=0,X_{1},X_{2},\dots. We may refer to a particle PiP_{i} by either its index ii or its initial position XiX_{i}.

We describe the process of infection in a somewhat informal manner (see also Section 2.4). For i0i\in{\mathbb{Z}}_{0}, at any given time tt particle PiP_{i} is in one of three states S (susceptible), I (infected), and R (removed). Thus the state space is Ω=Π×{S,I,R}0\Omega=\Pi\times\{\text{\rm S},\text{\rm I},\text{\rm R}\}^{{\mathbb{Z}}_{0}}, and we write ω(t)=(ωi(t):i0)\omega(t)=(\omega_{i}(t):i\in{\mathbb{Z}}_{0}) for the state of the process at time tt. Let StS_{t} (respectively, ItI_{t}, RtR_{t}) be the set of particles in state S (respectively, I, R) at time tt. We take

ωi(0)={Iif i=0,Sotherwise,\omega_{i}(0)=\begin{cases}\text{\rm I}&\text{if }i=0,\\ \text{\rm S}&\text{otherwise},\end{cases}

so that I0={P0}I_{0}=\{P_{0}\} and S0=𝒫{P0}S_{0}={\mathcal{P}}\setminus\{P_{0}\}. The only particle-transitions that may occur are S \to I and I \to R. The transitions SI\text{\rm S}\to\text{\rm I} occur at rates that depend on the locations of the currently infected particles.

We shall refer to the above (in conjunction with the specific infection assumptions of Sections 2.2 or 2.3) as the general model. When a0a\equiv 0 and σ1\sigma\equiv 1 in (2.1) (or, more generally, σ\sigma is constant), we shall refer to it as the Brownian model. We shall prove the existence of a subcritical phase (characterized by the absence of survival) for the general model subject to weak conditions. Our proof of survival for the delayed diffusion model is for the Brownian model alone. Estimates for the volume of the sausage generated by ζ\zeta play roles in the calculations, and it may be that, in this regard or another, the behaviour of a general model is richer than that of its Brownian version.

2.2. Delayed diffusion model

Each particle PjP_{j} is stationary if and only if it is in state S. If it becomes infected (at some time BjB_{j}, see (2.5)), henceforth it follows the diffusion Xj+ζjX_{j}+\zeta_{j}. We write

πj(t)={Xjif tBj,Xj+ζj(tBj)if t>Bj,\displaystyle\pi_{j}(t)=\begin{cases}X_{j}&\text{if }t\leq B_{j},\\ X_{j}+\zeta_{j}(t-B_{j})\quad&\text{if }t>B_{j},\end{cases}

for the position of PjP_{j} at time tt.

We describe next the rate at which a given particle PP infects another particle PP^{\prime}. The function μ\mu, given above, encapsulates the spatial aspect of the infection process, and a parameter ρ(0,)\rho\in(0,\infty) represents its intensity,

  • (SI\text{\rm S}\to\text{\rm I})

    Let t>0t>0, and let PjP_{j} be a particle that is in state S at all times s<ts<t. Each PiItP_{i}\in I_{t} (with iji\neq j) infects PjP_{j} at rate ρμ(Xjπi(t))\rho\mu(X_{j}-\pi_{i}(t)). The aggregate rate at which PjP_{j} becomes infected is

    (2.4) iIt,ijρμ(Xjπi(t)).\sum_{i\in I_{t},\,i\neq j}\rho\mu(X_{j}-\pi_{i}(t)).
  • (IR\text{\rm I}\to\text{\rm R})

    An infected particle is removed at rate α\alpha.

Transitions of other types are not permitted. We take the sample path ω=(ω(t):t0)\omega=(\omega(t):t\geq 0) to be pointwise right-continuous, which is to say that, for i0i\in{\mathbb{Z}}_{0}, the function ωi()\omega_{i}(\cdot) is right-continuous. The infection time BjB_{j} of particle PjP_{j} is given by

(2.5) Bj=inf{t0:PjIt}.B_{j}=\inf\{t\geq 0:P_{j}\in I_{t}\}.

The infection rates ρμ(Xjπi(t))\rho\mu(X_{j}-\pi_{i}(t)) of (2.4) are finite, and hence infections take place at a.s. distinct times. We may thus speak of PjP_{j} as being ‘directly infected’ by PiP_{i}. We speak of a point zΠz\in\Pi as being directly infected by a point yΠy\in\Pi when the associated particles have that property. If PjP_{j} is infected directly by PiP_{i}, we call PjP_{j} a child of PiP_{i}, and PiP_{i} the parent of PjP_{j}.

Following its infection, particle PiP_{i} remains infected for a further random time TiT_{i}, called the lifetime of PiP_{i}, and is then removed. The times TiT_{i} are random variables with the exponential distribution with parameter α>0\alpha>0, and are independent of one another and of the XjX_{j} and ζj\zeta_{j}.

In the above version of the delayed diffusion model, ρ\rho is assumed finite. When ρ=\rho=\infty, we shall consider only situations in which

(2.6) ρ=,μ=1M where Md is compact.\rho=\infty,\qquad\mu=1_{M}\text{ where $M\subseteq{\mathbb{R}}^{d}$ is compact.}

In this situation, a susceptible particle PjP_{j} becomes infected at the earliest instant that it belongs to πi(t)+M\pi_{i}(t)+M for some PiItP_{i}\in I_{t}, iji\neq j. This happens when either (i) PiP_{i} infects PjP_{j} as PiP_{i} diffuses around d{\mathbb{R}}^{d} post-infection, or (ii) at the moment BiB_{i} of infection of PiP_{i}, particle PjP_{j} is infected instantaneously by virtue of the fact that Xjπi(Bi)+MX_{j}\in\pi_{i}(B_{i})+M. These two situations are investigated slightly more fully in the following definition of ‘direct infection’.

The role of the Boolean model of continuum percolation becomes clear when ρ=\rho=\infty, and we illustrate this, subject to the simplifying assumption that MM is symmetric in the sense that xMx\in M if and only if xM-x\in M. Let Π=(Xi:i0)\Pi=(X_{i}:i\in{\mathbb{Z}}_{0}) be a Poisson process in d{\mathbb{R}}^{d} with constant intensity λ\lambda, and declare two points XiX_{i}, XjX_{j} to be adjacent if and only if XjXiMX_{j}-X_{i}\in M. This adjacency relation generates a graph GG with vertex-set Π\Pi. In the delayed diffusion process on the set Π\Pi, entire clusters of the percolation process are infected simultaneously.

Since there can be many (even infinitely many) simultaneous infections at the same time instant when ρ=\rho=\infty, the notion of ‘direct infection’ requires amplification. For j0j\neq 0, we say that PjP_{j} is directly infected by P0P_{0} if PjP_{j} is in state S at all times s<Yjs<Y_{j}, where Yj=inf{t0:Xjζ0(t)+M}Y_{j}=\inf\{t\geq 0:X_{j}\in\zeta_{0}(t)+M\}, and in addition Yj<T0Y_{j}<T_{0}. We make a similar definition, as follows, for direct infections by PiP_{i} with i0i\neq 0. Let i0i\neq 0 and j0,ij\neq 0,i.

  • (a)

    We say that PjP_{j} is dynamically infected by PiP_{i} if the following holds. Particle PiP_{i} (respectively, PjP_{j}) is in state I (respectively, state S) at all times Yi,jϵY_{i,j}-\epsilon for ϵ(0,ϵ0)\epsilon\in(0,\epsilon_{0}) and some ϵ0>0\epsilon_{0}>0, where Yi,j=inf{tBi:XjXi+ζi(t)+M}Y_{i,j}=\inf\{t\geq B_{i}:X_{j}\in X_{i}+\zeta_{i}(t)+M\}, and in addition Yi,jBiTiY_{i,j}-B_{i}\leq T_{i}.

  • (b)

    We say that PjP_{j} is instantaneously infected by PiP_{i} if the following holds. There exist n1n\geq 1 and k,i1,i2,,in=i,in+1=jk,i_{1},i_{2},\dots,i_{n}=i,i_{n+1}=j such that i1i_{1} is dynamically infected by kk (at time Bi1B_{i_{1}}) and

    Xim+1TmTm1,m=1,2,,n,X_{i_{m+1}}\in T_{m}\setminus T_{m-1},\qquad m=1,2,\dots,n,

    where

    T0:=,Tm=r=1m(Xr+M).T_{0}:=\varnothing,\quad T_{m}=\bigcup_{r=1}^{m}(X_{r}+M).

Condition (a) corresponds to infection through movement of PiP_{i}, and (b) corresponds to instantaneous infection at the moment of infection of PiP_{i}. We say that PjP_{j} is directly infected by PiP_{i} if it is infected by PiP_{i} either dynamically or instantaneously.

Certain events of probability 0 are overlooked in the above informal description including, for example, the event of being dynamically infected by two or more particles, and the event of being instantaneously infected by an infinite chain (ir)(i_{r}) but by no finite chain.

In either case ρ<\rho<\infty or ρ=\rho=\infty, we write θdd(λ,ρ,α)\theta_{\text{\rm dd}}(\lambda,\rho,\alpha) for the probability that infinitely many particles are infected. For concreteness, we note our special interest in the case in which:

  • (a)

    ζ\zeta is a standard Brownian motion,

  • (b)

    μ=1S\mu=1_{S} with SS the closed unit ball of d{\mathbb{R}}^{d}.

2.3. Diffusion model

The diffusion model differs from the delayed diffusion model of Section 2.2 in that all particles begin to move at time t=0t=0. The location of PjP_{j} at time tt is Xj+ζj(t)X_{j}+\zeta_{j}(t), and the transition rates are given as follows. First, suppose ρ(0,)\rho\in(0,\infty).

  • (SI\text{\rm S}\to\text{\rm I})

    Let t>0t>0, and let PjP_{j} be susceptible at all times s<ts<t. Each PiItP_{i}\in I_{t} (with iji\neq j) infects PjP_{j} at rate ρμ(Xj+ζj(t)Xiζi(t))\rho\mu(X_{j}+\zeta_{j}(t)-X_{i}-\zeta_{i}(t)). The aggregate rate at which PjP_{j} becomes infected is

    (2.7) iIt,ijρμ(Xj+ζj(t)Xiζi(t)).\sum_{i\in I_{t},\,i\neq j}\rho\mu\bigl{(}X_{j}+\zeta_{j}(t)-X_{i}-\zeta_{i}(t)\bigr{)}.
  • (IR\text{\rm I}\to\text{\rm R})

    An infected particle is removed at rate α\alpha.

As in Section 2.2, we may allow ρ=\rho=\infty and μ=1M\mu=1_{M} with MM compact. In either case ρ<\rho<\infty or ρ=\rho=\infty we write θd(λ,ρ,α)\theta_{\text{\rm d}}(\lambda,\rho,\alpha) for the probability that infinitely many particles are infected.

2.4. Construction

We shall not investigate the formal construction of the above processes as strong Markov processes with right-continuous sample paths. The interested reader may refer to the related model involving random walks on d{\mathbb{Z}}^{d} (rather than diffusions or Brownian motions on d{\mathbb{R}}^{d}) with α=0\alpha=0, as considered in some depth by Kesten and Sidoravicius in [17] and developed for the process with ‘recuperation’ in their sequel [18].

Instead, we sketch briefly how such processes may be built around a triple (Π,𝜻,𝜹)(\Pi,\boldsymbol{\zeta},\boldsymbol{\delta}), where Π\Pi is a rate-λ\lambda Poisson process of initial positions, 𝜻=(ζi:i0)\boldsymbol{\zeta}=(\zeta_{i}:i\in{\mathbb{Z}}_{0}) is a family of independent copies of the diffusion ζ\zeta, and 𝜹=(δi:i0)\boldsymbol{\delta}=(\delta_{i}:i\in{\mathbb{Z}}_{0}) is a family of independent rate-α\alpha Poisson processes on (0,)(0,\infty), that are independent of the pair (Π,𝜻)(\Pi,\boldsymbol{\zeta}). We place particles PiP_{i} at the points of Π\Pi, and PiP_{i} deviates from its initial point according to ζi\zeta_{i}. An initial infected particle P0P_{0} is placed at the origin at time 0, and it diffuses according to ζ0\zeta_{0}. The infection is communicated according to the appropriate rules (either delayed or not, and either (i) via the pair (ρ,μ)(\rho,\mu) with ρ<\rho<\infty, or (ii) with ρ=\rho=\infty and μ=1S\mu=1_{S}). After infection, PiP_{i} is removed at the next occurrence of the Poisson process δi\delta_{i}.

The above construction is straightforward so long as there exist, at any given time a.s., only finitely many simultaneous infections. In the two models with ρ<\rho<\infty, simultaneous infections can occur only after the earliest time TT_{\infty} at which there exist infinitely many infected particles. It is a consequence of Proposition 3.8(c) that (T<)=0{\mathbb{P}}(T_{\infty}<\infty)=0 for the general delayed diffusion model with ρ<\rho<\infty.

The issue is slightly more complex when ρ=\rho=\infty and μ=1S\mu=1_{S}, since infinitely many simultaneous infections may take place in the supercritical phase of the percolation process of moving disks. In this case, we assume invariably that λ<λc\lambda<\lambda_{\text{\rm c}}, so that there is no percolation of disks at any fixed time, and indeed it was shown in [5] that, a.s., there is no percolation at all times. Therefore, there exist, a.s., only finitely many simultaneous infections at any given instant. Bounds for the growth of generation sizes are found at (4.4) and (4.21).

3. The delayed diffusion model

3.1. Main results

We consider the Brownian delayed diffusion model of Section 2.2, and we adopt the notation of that section. Recall the critical point λc\lambda_{\text{\rm c}} of the Boolean continuum percolation on d{\mathbb{R}}^{d} in which a closed unit ball is placed at each point of a rate-λ\lambda Poisson process. Let θdd(λ,ρ,α)\theta_{\text{\rm dd}}(\lambda,\rho,\alpha) be the probability that the process survives.

Theorem 3.1.

Consider the Brownian delayed diffusion model on d{\mathbb{R}}^{d} where d2d\geq 2.

  • (a)

    Let ρ(0,)\rho\in(0,\infty). There exists a function αc:(0,)2(0,)\alpha_{\text{\rm c}}:(0,\infty)^{2}\to(0,\infty) such that

    (3.1) θdd(λ,ρ,α){=0if α>αc(λ,ρ),>0if α<αc(λ,ρ).\theta_{\text{\rm dd}}(\lambda,\rho,\alpha)\begin{cases}=0&\text{if }\alpha>\alpha_{\text{\rm c}}(\lambda,\rho),\\ >0&\text{if }\alpha<\alpha_{\text{\rm c}}(\lambda,\rho).\end{cases}

    The function θdd(λ,ρ,α)\theta_{\text{\rm dd}}(\lambda,\rho,\alpha) is non-increasing in α\alpha and non-decreasing in ρ\rho. Therefore, αc=αc(λ,ρ)\alpha_{\text{\rm c}}=\alpha_{\text{\rm c}}(\lambda,\rho) is non-decreasing in ρ\rho.

  • (b)

    Let ρ=\rho=\infty and μ=1S\mu=1_{S} where SS is the closed unit ball in d{\mathbb{R}}^{d}. There exists a non-decreasing function αc:(0,)(0,]\alpha_{\text{\rm c}}:(0,\infty)\to(0,\infty] such that, for 0<λ<λc0<\lambda<\lambda_{\text{\rm c}},

    (3.2) θdd(λ,,α){=0if α>αc(λ),>0if α<αc(λ).\theta_{\text{\rm dd}}(\lambda,\infty,\alpha)\begin{cases}=0&\text{if }\alpha>\alpha_{\text{\rm c}}(\lambda),\\ >0&\text{if }\alpha<\alpha_{\text{\rm c}}(\lambda).\end{cases}

    Furthermore, there exists λ¯(0,λc]\underline{\lambda}\in(0,\lambda_{\text{\rm c}}] such that

    αc(λ){<if 0<λ<λ¯,=if λ>λc.\alpha_{\text{\rm c}}(\lambda)\begin{cases}<\infty&\text{if }0<\lambda<\underline{\lambda},\\ =\infty&\text{if }\lambda>\lambda_{\text{\rm c}}.\end{cases}

    Moreover, the function θdd(λ,,α)\theta_{\text{\rm dd}}(\lambda,\infty,\alpha) is non-increasing in α\alpha.

This theorem extends Theorem 1.1. Its proof is found in Sections 3.23.7.

The situation is different in one dimension, where it turns out that the Brownian model has no phase transition. The proof of the following theorem, in a version valid for the general delayed diffusion model, may be found in Section 3.3.

Theorem 3.2.

Consider the Brownian delayed diffusion model on {\mathbb{R}} with μ=1S\mu=1_{S}. We have that θdd(λ,ρ,α)=0\theta_{\text{\rm dd}}(\lambda,\rho,\alpha)=0 for all λ,α>0\lambda,\alpha>0 and ρ(0,]\rho\in(0,\infty].

Theorems 3.1 and 3.2 are stated for the case of a single initial infective. The proofs are valid also with a finite number of initial infectives distributed at the points of some arbitrary subset I0I_{0} of d{\mathbb{R}}^{d}. By the proof of the forthcoming Proposition 3.4, the set II of ultimately infected particles is stochastically increasing in I0I_{0}.

3.2. Percolation representation of the delayed diffusion model

Consider the delayed diffusion model with d2d\geq 2. Suppose that either ρ(0,)\rho\in(0,\infty) with μ\mu as in (2.2), or ρ=\rho=\infty and

(3.3) μ(x)=1S(x),x2,\mu(x)=1_{S}(x),\qquad x\in{\mathbb{R}}^{2},

where SS is the closed unit ball with centre at the origin. It turns out that the set of infected particles may be considered as a type of percolation model on the random set Π\Pi. This observation will be useful in exploring the phases of the former model.

The proof of the main result of this section, Proposition 3.3, is motivated in part by work of Kuulasmaa [21] where a certain epidemic model was studied via a related percolation process (a similar argument is implicit in [1, p. 4]). Recall the initial placements Π=(X0=0,X1,X2,)\Pi=(X_{0}=0,X_{1},X_{2},\dots) of particles PiP_{i}, with law denoted P (and corresponding expectation E); we condition on Π\Pi.

Fix i0i\geq 0, and consider the following infection process. The particle PiP_{i} is the unique initially infected particle, and it diffuses according to ζi\zeta_{i} and has lifetime TiT_{i}. All other particles PjP_{j}, jij\neq i, are kept stationary for all time at their respective locations XjX_{j}. As PiP_{i} moves around d{\mathbb{R}}^{d}, it infects other particles in the usual way; newly infected particles are permitted neither to move nor to infect others. Let JiJ_{i} be the (random) set of particles infected by PiP_{i} in this process.

Let τi,j(0,]\tau_{i,j}\in(0,\infty] be the time of the first infection by PiP_{i} of PjP_{j}, assuming that PiP_{i} is never removed. Write iji\to j if τi,j<Ti\tau_{i,j}<T_{i}, which is to say that this infection takes place before PiP_{i} is removed. Thus,

(3.4) Ji={j:ij}.J_{i}=\{j:i\to j\}.

Suppose first that ρ<\rho<\infty. Given (Π,ζi,Ti)(\Pi,\zeta_{i},T_{i}), the vector τi=(τi,j:ji)\vec{\tau}_{i}=(\tau_{i,j}:j\neq i) contains conditionally independent random variables with respective distribution functions

(3.5) Fi,j(t)=1exp(0tρμ(XjXiζi(s))𝑑s),t0,F_{i,j}(t)=1-\exp\left(-\int_{0}^{t}\rho\mu(X_{j}-X_{i}-\zeta_{i}(s))\,ds\right),\qquad t\geq 0,

and

(3.6) (ijΠ,ζi,Ti)=Fi,j(Ti).{\mathbb{P}}(i\to j\mid\Pi,\zeta_{i},T_{i})=F_{i,j}(T_{i}).

When ρ=\rho=\infty, we have that

(3.7) τi,j=inf{t>0:XjXi+ζi(t)+S},\tau_{i,j}=\inf\{t>0:X_{j}\in X_{i}+\zeta_{i}(t)+S\},

the first hitting time of XjXiX_{j}-X_{i} by the radius-11 sausage of ζi\zeta_{i}. As above, we write iji\to j if τi,j<Ti\tau_{i,j}<T_{i}, with JiJ_{i} and τi\vec{\tau}_{i} given accordingly.

One may thus construct sets JiJ_{i} for all i0i\geq 0; given Π\Pi, the set JiJ_{i} depends only on (ζi,Ti)(\zeta_{i},T_{i}), and therefore the JiJ_{i} are conditionally independent given Π\Pi. The sets {Ji:i0}\{J_{i}:i\geq 0\} generate a directed graph G=GΠ\vec{G}=\vec{G}_{\Pi} with vertex-set 0{\mathbb{Z}}_{0} and directed edge-set E={[i,j:ij}\vec{E}=\{[i,j\rangle:i\to j\}. Write I\vec{I} for the set of vertices kk of G\vec{G} such that there exists a directed path of G\vec{G} from 0 to kk. To the edges of G\vec{G} we attach random labels, with edge [i,j[i,j\rangle receiving the label τi,j\tau_{i,j}.

From the vector (τi,Ti:i0)(\vec{\tau}_{i},T_{i}:i\geq 0), we can construct a copy of the general delayed diffusion process by allowing an infection by PiP_{i} of PjP_{j} whenever iji\to j and in addition PjP_{j} has not been infected previously by another particle. Let II denote the set of ultimately infected particles in this coupled process.

Proposition 3.3.

For ρ(0,]\rho\in(0,\infty], we have I=II=\vec{I}.

We turn our attention to the Brownian case. By rescaling in space/time, we obtain the following. The full parameter-set of the process is {λ,ρ,α,μ,σ}\{\lambda,\rho,\alpha,\mu,\sigma\}, where σ\sigma is the standard-deviation parameter of the Brownian motion, and we shall sometimes write θdd(λ,ρ,α,μ,σ)\theta_{\text{\rm dd}}(\lambda,\rho,\alpha,\mu,\sigma) accordingly.

Proposition 3.4.

Consider the Brownian delayed diffusion model. Let ρ(0,]\rho\in(0,\infty].

  • (a)

    For given λ(0,)\lambda\in(0,\infty), the function θdd(λ,ρ,α)\theta_{\text{\rm dd}}(\lambda,\rho,\alpha) is non-decreasing in ρ\rho and non-increasing in α\alpha.

  • (b)

    We have that

    (3.8) θdd(λ,ρ,α,μ,1)=θdd(λ/rd,ρ/r2,α/r2,μr,1),r1,\theta_{\text{\rm dd}}(\lambda,\rho,\alpha,\mu,1)=\theta_{\text{\rm dd}}(\lambda/r^{d},\rho/r^{2},\alpha/r^{2},\mu_{r},1),\qquad r\geq 1,

    where μr(x):=μ(x/r)\mu_{r}(x):=\mu(x/r).

  • (c)

    If μ\mu is radially decreasing (see (2.3)), then

    αc(λ,ρ)r2αc(λ/rd,ρ/r2),r1.\alpha_{\text{\rm c}}(\lambda,\rho)\geq r^{2}\alpha_{\text{\rm c}}(\lambda/r^{d},\rho/r^{2}),\qquad r\geq 1.
  • (d)

    If ρ=\rho=\infty and μ\mu is radially decreasing, then θdd(λ,,α)\theta_{\text{\rm dd}}(\lambda,\infty,\alpha) and αc(λ,)\alpha_{\text{\rm c}}(\lambda,\infty) are non-decreasing in λ\lambda.

Proof of Proposition 3.3.

This is a deterministic claim. Assume Π\Pi is given. If iIi\in I, there exists a chain of direct infection from 0 to ii, and this chain generates a directed path of G\vec{G} from 0 to ii. Suppose, conversely, that kIk\in\vec{I}. Let 𝒫k{\mathcal{P}}_{k} be the set of directed paths of G\vec{G} from 0 to kk. Let π𝒫k\pi\in{\mathcal{P}}_{k} be a shortest such path (where the length of an edge [i,j[i,j\rangle is taken to be the label τi,j\tau_{i,j} of that edge). We may assume that the τi,j\tau_{i,j}, for iji\to j, are distinct; no essential difficulty emerges on the complementary null set. Then the path π\pi is a geodesic, in that every sub-path is the shortest directed path joining its endvertices. Therefore, when infection is initially introduced at P0P_{0}, it will be transmitted directly along π\pi to PkP_{k}. ∎

Proof of Proposition 3.4.

(a) By Proposition 3.3, if the parameters are changed in such a way that each JiJ_{i} is stochastically increased (respectively, decreased), then the set II is also stochastically increased (respectively, decreased). The claims follow by (3.5)–(3.6) when ρ<\rho<\infty, and by (3.7) when ρ=\rho=\infty.

(b) We shall show that the probabilities of infections are the same for the two sets of parameter-values in (3.8). Let r1r\geq 1, and consider the effect of dilating space by the ratio rr. The resulting stretched Poisson process rΠr\Pi has intensity λ/rd\lambda/r^{d}, the resulting Brownian motion rζi(t)r\zeta_{i}(t) is distributed as ζi(r2t)\zeta_{i}(r^{2}t), and μ\mu is replaced by μr\mu_{r}. Therefore,

(3.9) θdd(λ,ρ,α,μ,1)=θdd(λ/rd,ρ,α,μr,r).\theta_{\text{\rm dd}}(\lambda,\rho,\alpha,\mu,1)=\theta_{\text{\rm dd}}(\lambda/r^{d},\rho,\alpha,\mu_{r},r).

Next, we use the construction of the process in terms of the JiJ_{i} given above Proposition 3.3. If ρ<\rho<\infty then, by (3.6) and the change of variables u=r2su=r^{2}s,

(ijΠ,ζi,Ti)\displaystyle{\mathbb{P}}(i\to j\mid\Pi,\zeta_{i},T_{i}) =1exp(0Tiρμ(XjXiζi(s))𝑑s)\displaystyle=1-\exp\left(-\int_{0}^{T_{i}}\rho\mu(X_{j}-X_{i}-\zeta_{i}(s))\,ds\right)
=d1exp(0Tiρμr(rXjrXiζi(r2s))𝑑s)\displaystyle\stackrel{{\scriptstyle\text{\rm d}}}{{=}}1-\exp\left(-\int_{0}^{T_{i}}\rho\mu_{r}(rX_{j}-rX_{i}-\zeta_{i}(r^{2}s))\,ds\right)
=d1exp(0r2Tiρμr(rXjrXiζi(u))dur2),\displaystyle\stackrel{{\scriptstyle\text{\rm d}}}{{=}}1-\exp\left(-\int_{0}^{r^{2}T_{i}}\rho\mu_{r}(rX_{j}-rX_{i}-\zeta_{i}(u))\,\frac{du}{r^{2}}\right),

where =d\stackrel{{\scriptstyle\text{\rm d}}}{{=}} means equality in distribution. Since r2Tir^{2}T_{i} is exponentially distributed with parameter α/r2\alpha/r^{2}, the right side of (3.9) equals θdd(λ/rd,ρ/r2,α/r2,μr,1)\theta_{\text{\rm dd}}(\lambda/r^{d},\rho/r^{2},\alpha/r^{2},\mu_{r},1), as claimed. The same conclusion is valid for ρ=\rho=\infty, by (3.7).

(c) Since μrμ\mu_{r}\geq\mu by assumption, the JiJ_{i} are stochastically monotone in μ\mu, it follows by (3.8) that

θdd(λ,ρ,α,μ,1)θdd(λ/rd,ρ/r2,α/r2,μ,1),r1.\theta_{\text{\rm dd}}(\lambda,\rho,\alpha,\mu,1)\geq\theta_{\text{\rm dd}}(\lambda/r^{d},\rho/r^{2},\alpha/r^{2},\mu,1),\qquad r\geq 1.

By the monotonicity of θdd\theta_{\text{\rm dd}} in α\alpha, if α>αc(λ,ρ)\alpha>\alpha_{\text{\rm c}}(\lambda,\rho) then α/r2αc(λ/rd,ρ/r2)\alpha/r^{2}\geq\alpha_{\text{\rm c}}(\lambda/r^{d},\rho/r^{2}) as claimed.

(d) This holds as in part (c).

Remark 3.5.

In the forthcoming proof of Section 3.7.4 we shall make use of a consequence of Proposition 3.3, namely that

(3.10) θdd(λ,ρ,α)=E(Π(|I|=)),\theta_{\text{\rm dd}}(\lambda,\rho,\alpha)=\text{\rm E}\bigl{(}{\mathbb{Q}}_{\Pi}(|\vec{I}|=\infty)\bigr{)},

where Π{\mathbb{Q}}_{\Pi} is the conditional law of G\vec{G} given Π\Pi, and E is expectation with respect to Π\Pi. In proving survival, it therefore suffices to prove the right side of (3.10) is strictly positive.

3.3. No survival in one dimension

It was stated in Theorem 3.2 that the Brownian model with μ=1S\mu=1_{S} never survives in one dimension. We state and prove a version of this ‘no survival’ theorem for the general delayed diffusion model of Section 2.2, subject to a weak condition which includes the Brownian model.

Throughout this section, TαT_{\alpha} denotes a random variable having the exponential distribution with parameter α\alpha, assumed to be independent of all other random variables involved in the models. A typical diffusion is denoted ζ\zeta, and we write Mt=sup{|ζ(s)|:s[0,t]}M_{t}=\sup\{|\zeta(s)|:s\in[0,t]\}.

Theorem 3.6.

Consider the general delayed diffusion model on {\mathbb{R}} with infection parameters (ρ,μ)(\rho,\mu).

  • (a)

    Let ρ<\rho<\infty. Assume that α\alpha is such that 𝔼|ζ(Tα)|<{\mathbb{E}}|\zeta(T_{\alpha})|<\infty, and in addition that |y|μ(y)𝑑y<\int_{\mathbb{R}}|y|\mu(y)\,dy<\infty. Then θdd(λ,ρ,α)=0\theta_{\text{\rm dd}}(\lambda,\rho,\alpha)=0 for all λ>0\lambda>0.

  • (b)

    Let ρ=\rho=\infty and μ\mu have bounded support. Assume that α\alpha is such that 𝔼(MTα)<{\mathbb{E}}(M_{T_{\alpha}})<\infty. Then θdd(λ,,α)=0\theta_{\text{\rm dd}}(\lambda,\infty,\alpha)=0 for all λ>0\lambda>0.

It follows that. for all α\alpha, there is no survival if

either: ρ<\rho<\infty and, for all α\alpha, we have 𝔼|ζ(Tα)|<{\mathbb{E}}|\zeta(T_{\alpha})|<\infty,
or: ρ=μ has bounded support, and, for all α, we have 𝔼(MTα)<.\displaystyle\text{$\rho=\infty$, $\mu$ has bounded support, and, for all $\alpha$, we have ${\mathbb{E}}(M_{T_{\alpha}})<\infty$}.

These two conditions (that 𝔼|ζ(Tα)|<{\mathbb{E}}|\zeta(T_{\alpha})|<\infty and 𝔼(MTα)<{\mathbb{E}}(M_{T_{\alpha}})<\infty) are equivalent when ζ\zeta is Brownian motion (see, for example, [12, Thm 13.4.6]), and indeed they hold for all α\alpha in the Brownian case. This implies Theorem 3.2. The proof of Theorem 3.6 has some similarity to that of [1, Thm 1.1].

Proof.

(a) Assume the required conditions. Let A>0A>0; later we will take AA to be large. Let Π\Pi^{\prime} be a Poisson process on {\mathbb{R}} with intensity λ\lambda, and let L=Π(,0]L=\Pi^{\prime}\cap(-\infty,0] and R=Π[A,)R=\Pi^{\prime}\cap[A,\infty). Write {S1S2}\{S_{1}\to S_{2}\} for the event that, in the percolation representation of the last section, some particle in S1S_{1} infects some particle in S2S_{2}. We prove first that

(3.11) (LR)0as A.{\mathbb{P}}(L\to R)\to 0\qquad\text{as }A\to\infty.

Note that, for suitable functions ff,

(3.12) 𝔼(0Tαf(ζ(s))ds)=0𝔼(f(ζ(s))eαsds=1α𝔼(f(ζ(Tα)).{\mathbb{E}}\left(\int_{0}^{T_{\alpha}}f(\zeta(s))\,ds\right)=\int_{0}^{\infty}{\mathbb{E}}(f(\zeta(s))e^{-\alpha s}\,ds=\frac{1}{\alpha}{\mathbb{E}}(f(\zeta(T_{\alpha})).

Since (LR){\mathbb{P}}(L\to R) is no larger than the mean number of infections from LL into RR, we have by the Campbell–Hardy Theorem for Poisson processes (see [12, Exer. 6.13.2]), Fubini’s Theorem, and (3.12) that

(3.13) (LR)\displaystyle{\mathbb{P}}(L\to R) λ20𝑑uA𝑑v𝔼(ρ0Tαμ(vuζu(s))𝑑s)\displaystyle\leq\lambda^{2}\int_{-\infty}^{0}du\int_{A}^{\infty}dv\,{\mathbb{E}}\left(\rho\int_{0}^{T_{\alpha}}\mu(v-u-\zeta_{u}(s))\,ds\right)
=λ2ρα𝔼(0𝑑uA𝑑vμ(vuζ(Tα)))\displaystyle=\frac{\lambda^{2}\rho}{\alpha}{\mathbb{E}}\left(\int_{-\infty}^{0}du\int_{A}^{\infty}dv\,\mu(v-u-\zeta(T_{\alpha}))\right)
=λ2ρα𝔼(A𝑑vI(vζ(Tα)))=λ2ρα𝔼(ZA),\displaystyle=\frac{\lambda^{2}\rho}{\alpha}{\mathbb{E}}\left(\int_{A}^{\infty}dv\,I(v-\zeta(T_{\alpha}))\right)=\frac{\lambda^{2}\rho}{\alpha}{\mathbb{E}}(Z_{A}),

where

I(x):=xμ(y)𝑑y,ZA:=Aζ(Tα)I(v)𝑑v.I(x):=\int_{x}^{\infty}\mu(y)\,dy,\qquad Z_{A}:=\int_{A-\zeta(T_{\alpha})}^{\infty}I(v)\,dv.

Since I(x)I()<I(x)\leq I(-\infty)<\infty, we have

𝔼(ZA)I()𝔼|ζ(Tα)|+AI(v)𝑑v.{\mathbb{E}}(Z_{A})\leq I(-\infty){\mathbb{E}}|\zeta(T_{\alpha})|+\int_{A}^{\infty}I(v)\,dv.

Furthermore, ZAZ_{A} is integrable since

AI(v)𝑑v0I(v)𝑑v=0yμ(y)𝑑y<.\int_{A}^{\infty}I(v)\,dv\leq\int_{0}^{\infty}I(v)\,dv=\int_{0}^{\infty}y\mu(y)\,dy<\infty.

Since ZA0Z_{A}\to 0 a.s. as AA\to\infty, we have by monotone convergence that 𝔼(ZA)0{\mathbb{E}}(Z_{A})\to 0 also. Equation (3.11) now follows by (3.13). By a similar argument, (RL)0{\mathbb{P}}(R\to L)\to 0 as AA\to\infty.

We pick AA sufficiently large that

(LR)14,(RL)14.{\mathbb{P}}(L\to R)\leq\tfrac{1}{4},\qquad{\mathbb{P}}(R\to L)\leq\tfrac{1}{4}.

On the event EA:={L↛R}{R↛L}{Π(0,A)=}E_{A}:=\{L\not\to R\}\cap\{R\not\to L\}\cap\{\Pi^{\prime}\cap(0,A)=\varnothing\}, there can be no chain of infection between particles in LL and particles in RR. Note that

(3.14) (EA)12eλA>0.{\mathbb{P}}(E_{A})\geq\tfrac{1}{2}e^{-\lambda A}>0.

Let B0=EAB_{0}=E_{A} and, for kk\in{\mathbb{Z}}, let BkB_{k} be the event defined similarly to EAE_{A} but with the interval (0,A)(0,A) replaced by (kA,(k+1)A)(kA,(k+1)A). By (3.14),

(3.15) (Bk)=(EA)12eλA>0.{\mathbb{P}}(B_{k})={\mathbb{P}}(E_{A})\geq\tfrac{1}{2}e^{-\lambda A}>0.

By the ergodic theorem, the limit

Λ:=limn12n+1k=nn1Bk\Lambda:=\lim_{n\to\infty}\frac{1}{2n+1}\sum_{k=-n}^{n}1_{B_{k}}

exists a.s. and has mean at least 12eλA\frac{1}{2}e^{-\lambda A}. Since Λ\Lambda is translation-invariant and the underlying probability measure is a product measure, we have Λ13eλA\Lambda\geq\tfrac{1}{3}e^{-\lambda A} a.s. Therefore, BkB_{k} occurs infinitely often a.s. This would imply the claim were it not for the extra particle at 0 and its associated Brownian motion ζ0\zeta_{0}. However, the range of ζ0\zeta_{0}, up to time T0T_{0}, is a.s. bounded, and this completes the proof.

(b) Let ρ=\rho=\infty and, for simplicity, take μ=1S\mu=1_{S} (the proof for general μ\mu with bounded support is essentially the same). The proof is close to that of part (a).

Let Π\Pi^{\prime} be a Poisson process on {\mathbb{R}} with intensity λ\lambda, let (ζX:XΠ)(\zeta_{X}:X\in\Pi^{\prime}) be independent copies of ζ\zeta, and let (TX:XΠ)(T_{X}:X\in\Pi^{\prime}) be independent copies of TαT_{\alpha}. Write Fx=inf{t0:ζ(t)=x}F_{x}=\inf\{t\geq 0:\zeta(t)=x\} for the first-passage time of ζ\zeta to the point xx\in{\mathbb{R}}.

Let

A=XΠ(,0){GX>TX},A=\bigcap_{X\in\Pi^{\prime}\cap(-\infty,0)}\{G_{X}>T_{X}\},

where GXG_{X} is the first-passage time to 0 of X+ζXX+\zeta_{X}. Then

(3.16) (A)(Π[1,0]=)𝔼(XΠ(,1)(1pX)),{\mathbb{P}}(A)\geq{\mathbb{P}}\bigl{(}\Pi^{\prime}\cap[-1,0]=\varnothing\bigr{)}{\mathbb{E}}\left(\prod_{X\in\Pi^{\prime}\cap(-\infty,-1)}(1-p_{X})\right),

where

(3.17) px=(GxTx)=(FxTα)(MTα|x|).p_{x}={\mathbb{P}}(G_{x}\leq T_{x})={\mathbb{P}}(F_{-x}\leq T_{\alpha})\leq{\mathbb{P}}(M_{T_{\alpha}}\geq|x|).

There exists ϵ>0\epsilon>0 such that px<1ϵp_{x}<1-\epsilon for x1x\geq 1, and therefore there exists c=c(ϵ)(0,)c=c(\epsilon)\in(0,\infty) such that 1pxecpx1-p_{x}\geq e^{-cp_{x}} for x1x\geq 1. By (3.16) and Jensen’s inequality,

(3.18) (A)eλexp(c𝔼(XΠ(,1)pX)).{\mathbb{P}}(A)\geq e^{-\lambda}\exp\left(-c{\mathbb{E}}\biggl{(}\sum_{X\in\Pi^{\prime}\cap(-\infty,-1)}p_{X}\biggr{)}\right).

By the Campbell–Hardy Theorem,

(3.19) 𝔼(XΠ(,1)pX)\displaystyle{\mathbb{E}}\biggl{(}\sum_{X\in\Pi^{\prime}\cap(-\infty,-1)}p_{X}\biggr{)} =λ1𝑑x0𝑑tαeαt(Fxt).\displaystyle=\lambda\int_{-\infty}^{-1}dx\int_{0}^{\infty}dt\,\alpha e^{-\alpha t}{\mathbb{P}}(F_{-x}\leq t).

By (3.17), we obtain after interchanging the order of integration that

(3.20) 𝔼(XΠ(,1)pX)\displaystyle{\mathbb{E}}\biggl{(}\sum_{X\in\Pi^{\prime}\cap(-\infty,-1)}p_{X}\biggr{)} λ𝔼(MTα).\displaystyle\leq\lambda{\mathbb{E}}(M_{T_{\alpha}}).

By (3.18)–(3.20),

(3.21) (A)exp(λcλ𝔼(MTα))>0.{\mathbb{P}}(A)\geq\exp\bigl{(}-\lambda-c\lambda{\mathbb{E}}(M_{T_{\alpha}})\bigr{)}>0.

For kk\in{\mathbb{Z}}, let BkB_{k} be the event that, for all XΠX\in\Pi^{\prime}, the diffusion X+ζXX+\zeta_{X} hits the interval [k1,k+1][k-1,k+1] only after time TXT_{X}. We repeat the argument of part (a) with (3.21) in place of (3.14), and thereby obtain the claim. ∎

3.4. A condition for subcriticality when ρ<\rho<\infty

Consider the general delayed diffusion model of Section 2.2, and assume first that ρ(0,)\rho\in(0,\infty). Let I0={0}I_{0}=\{0\}. We call yΠy\in\Pi a first generation infected point up to time tt if yy is directly infected by P0P_{0} at or before time tt. Let I1,tI_{1,t} be the set of all first generation infected points up to time tt. For n2n\geq 2, we call zΠz\in\Pi an nnth generation infected point up to time tt if, at or before time tt, zz is directly infected by some yIn1,ty\in I_{n-1,t}, and we define In,tI_{n,t} accordingly. Write In=limtIn,tI_{n}=\lim_{t\to\infty}I_{n,t}, the set of all nnth generation infected points, and let I=nInI=\bigcup_{n}I_{n} be the set of points that are ever infected.

In the following, we shall sometimes use the coupling of the delayed diffusion model with the percolation-type system of the Section 3.2, and we shall use the notation of that section. In particular, we have that I1J0I_{1}\subseteq J_{0}, and by Proposition 3.3 that I=II=\vec{I} (note that I1I_{1} is a strict subset of J0J_{0} if there exist i,jJ0i,j\in J_{0} such that, in the notation of (3.4), we have 0ij0\to i\to j).

Proposition 3.7.

Consider the general delayed diffusion model. Let ρ(0,)\rho\in(0,\infty) and

(3.22) Lt(x)=𝔼(1exp(0tρμ(xζ(s))𝑑s)).L_{t}(x)={\mathbb{E}}\left(1-\exp\left(-\int_{0}^{t}\rho\mu(x-\zeta(s))\,ds\right)\right).

We have that 𝔼|I1,t|Rt{\mathbb{E}}|I_{1,t}|\leq R_{t} and 𝔼|I1|R{\mathbb{E}}|I_{1}|\leq R, where

(3.23) Rt\displaystyle R_{t} =λd[0tLs(x)αeαs𝑑s+Lt(x)eαt]𝑑x,\displaystyle=\lambda\int_{{\mathbb{R}}^{d}}\left[\int_{0}^{t}L_{s}(x)\alpha e^{-\alpha s}\,ds+L_{t}(x)e^{-\alpha t}\right]\,dx,
(3.24) R\displaystyle R =limtRt=λd0Ls(x)αeαs𝑑s𝑑x.\displaystyle=\lim_{t\to\infty}R_{t}=\lambda\int_{{\mathbb{R}}^{d}}\int_{0}^{\infty}L_{s}(x)\alpha e^{-\alpha s}\,ds\,dx.

The constant RR in (3.24) is an upper bound for the so-called reproductive rate of the process. In the notation of Section 3.2, we have R=𝔼|J0|R={\mathbb{E}}|J_{0}|.

Proposition 3.8.

Consider the general delayed diffusion model. Let ρ(0,)\rho\in(0,\infty).

  • (a)

    We have that 𝔼|In|Rn{\mathbb{E}}|I_{n}|\leq R^{n} for n0n\geq 0, where RR is given in (3.24).

  • (b)

    If R<1R<1, then 𝔼|I|1/(1R){\mathbb{E}}|I|\leq 1/(1-R), and hence θdd(λ,ρ,α)=0\theta_{\text{\rm dd}}(\lambda,\rho,\alpha)=0.

  • (c)

    We have that RλρInt(μ)/αR\leq\lambda\rho\,\text{\rm Int}(\mu)/\alpha.

Note that parts (b) and (c) imply that

(3.25) θdd(λ,ρ,α)=0ifα>λρInt(μ).\theta_{\text{\rm dd}}(\lambda,\rho,\alpha)=0\quad\text{if}\quad\alpha>\lambda\rho\,\text{\rm Int}(\mu).
Proof of Proposition 3.7.

Let 0(t){\mathcal{F}}_{0}(t) be the σ\sigma-field generated by (ζ0(s):0st)(\zeta_{0}(s):0\leq s\leq t). Conditional on 0(t){\mathcal{F}}_{0}(t), for i1i\geq 1, let Ai=(Aik:k0)A_{i}=(A_{i}^{k}:k\geq 0) be a Poisson process on [0,)[0,\infty) with rate function

rXi(s):=ρμ(Xiζ0(s)),s[0,).r_{X_{i}}(s):=\rho\mu(X_{i}-\zeta_{0}(s)),\qquad s\in[0,\infty).

Assume the AiA_{i} are independent conditional on 0(t){\mathcal{F}}_{0}(t), and write Ni=|{k:Aikt}|N_{i}=|\{k:A_{i}^{k}\leq t\}|. We say that P0P_{0} ‘contacts’ PiP_{i} at the times {Aik:k1}\{A_{i}^{k}:k\geq 1\}. Let Ut={Xi:i1,Ni1}U_{t}=\{X_{i}:i\geq 1,\ N_{i}\geq 1\} be the set of points in Π\Pi that P0P_{0} contacts up to time tt. Note that I1,tI_{1,t} is dominated stochastically by UtU_{t}. The domination is strict since there may exist XiUtX_{i}\in U_{t} such that PiP_{i} is infected before time tt by some previously infected PjP0P_{j}\neq P_{0}.

Consider a particle, labelled PjP_{j} say, with initial position xdx\in{\mathbb{R}}^{d}. Conditional on 0(t){\mathcal{F}}_{0}(t), P0P_{0} contacts PjP_{j} prior to time tt with probability

1exp(0trx(s)𝑑s).1-\exp\left(-\int_{0}^{t}r_{x}(s)\,ds\right).

Therefore,

(3.26) (XjI1,t|Xj=x,0(t))𝔼(1exp(0trx(s)ds)|0(t)).{\mathbb{P}}\bigl{(}X_{j}\in I_{1,t}\,\big{|}\,X_{j}=x,\,{\mathcal{F}}_{0}(t)\bigr{)}\leq{\mathbb{E}}\left(1-\exp\left(-\int_{0}^{t}r_{x}(s)\,ds\right)\,\bigg{|}\,{\mathcal{F}}_{0}(t)\right).

By the colouring theorem for Poisson processes (see, for example, [12, Thm 6.13.14]), conditional on 0(t){\mathcal{F}}_{0}(t), UtU_{t} is a Poisson process with inhomogeneous intensity function given by

Λt,ζ0(x)=λ𝔼(1exp(0trx(s)𝑑s)|0(t)).\Lambda_{t,\zeta_{0}}(x)=\lambda{\mathbb{E}}\left(1-\exp\left(-\int_{0}^{t}r_{x}(s)\,ds\right)\,\bigg{|}\,{\mathcal{F}}_{0}(t)\right).

By Fubini’s theorem,

(3.27) 𝔼|I1,t|\displaystyle{\mathbb{E}}|I_{1,t}| 𝔼(𝔼(|Ut||T0))\displaystyle\leq{\mathbb{E}}\bigl{(}{\mathbb{E}}(|U_{t}|\,\big{|}\,T_{0})\bigr{)}
=d[λ0tLs(x)αeαs𝑑s+Lt(x)(T0>t)]𝑑x,\displaystyle=\int_{{\mathbb{R}}^{d}}\left[\lambda\int_{0}^{t}L_{s}(x)\alpha e^{-\alpha s}\,ds+L_{t}(x){\mathbb{P}}(T_{0}>t)\right]\,dx,

and (3.23) follows. Equation (3.24) follows as tt\to\infty by the monotone and bounded convergence theorems. ∎

Proof of Proposition 3.8.

(a) This may be proved directly, but it is more informative to use the percolation representation of Section 3.2. Let G=GΠ\vec{G}=\vec{G}_{\Pi} be as defined there, and note that, in the given coupling, we have In={i0:δ(0,i)=n}I_{n}=\{i\in{\mathbb{Z}}_{0}:\delta(0,i)=n\} where δ\delta denotes graph-theoretic distance on G\vec{G}.

We write

|In|i0|Ji|1(iIn1).|I_{n}|\leq\sum_{i\in{\mathbb{Z}}_{0}}|J_{i}|1(i\in I_{n-1}).

By the independence of JiJ_{i} and the event {iIn1}\{i\in I_{n-1}\},

𝔼|In|i0𝔼|Ji|(iIn1)R𝔼|In1|,{\mathbb{E}}|I_{n}|\leq\sum_{i\in{\mathbb{Z}}_{0}}{\mathbb{E}}|J_{i}|\,{\mathbb{P}}(i\in I_{n-1})\leq R{\mathbb{E}}|I_{n-1}|,

and the claim follows.

(b) By part (a) and the assumption R<1R<1,

𝔼|I|=n=0𝔼|In|11R<.{\mathbb{E}}|I|=\sum_{n=0}^{\infty}{\mathbb{E}}|I_{n}|\leq\frac{1}{1-R}<\infty.

Therefore, θdd(λ,ρ,α)=(|I|=)=0\theta_{\text{\rm dd}}(\lambda,\rho,\alpha)={\mathbb{P}}(|I|=\infty)=0.

(c) Since 1ezz1-e^{-z}\leq z for z0z\geq 0, by (3.22) and Fubini’s theorem,

dLt(x)𝑑xρtInt(μ).\int_{{\mathbb{R}}^{d}}L_{t}(x)\,dx\leq\rho t\,\text{\rm Int}(\mu).

By (3.24),

RλρInt(μ)0sαeαs𝑑s=λραInt(μ),R\leq\lambda\rho\,\text{\rm Int}(\mu)\int_{0}^{\infty}s\alpha e^{-\alpha s}\,ds=\frac{\lambda\rho}{\alpha}\,\text{\rm Int}(\mu),

as claimed. ∎

3.5. Infection with compact support

Suppose μ=1M\mu=1_{M} with MM compact. By (3.22) and (3.24), R=R(ρ)R=R(\rho) is given by

(3.28) R(ρ)=λd0Ls(x)αeαs𝑑s𝑑x,R(\rho)=\lambda\int_{{\mathbb{R}}^{d}}\int_{0}^{\infty}L_{s}(x)\alpha e^{-\alpha s}\,ds\,dx,

where

(3.29) Lt(x)=𝔼(1exp(ρQt(x))),L_{t}(x)={\mathbb{E}}\bigl{(}1-\exp\left(-\rho Q_{t}(x)\right)\bigr{)},

and

Qt(x)=|{s[0,t]:xζ(s)+M}|1.Q_{t}(x)=\bigl{|}\{s\in[0,t]:x\in\zeta(s)+M\}\bigr{|}_{1}.

We denote by Σt\Sigma_{t} the MM-sausage of ζ\zeta, that is,

(3.30) Σt:=s[0,t][ζ(s)+M],t0.\Sigma_{t}:=\bigcup_{s\in[0,t]}\bigl{[}\zeta(s)+M\bigr{]},\qquad t\geq 0.

Consider the limit ρ\rho\to\infty. By (3.28) and dominated convergence,

(3.31) R(ρ)R¯:=λd0L¯s(x)αeαs𝑑s𝑑x,R(\rho)\uparrow\overline{R}:=\lambda\int_{{\mathbb{R}}^{d}}\int_{0}^{\infty}\overline{L}_{s}(x)\alpha e^{-\alpha s}\,ds\,dx,

where

L¯t(x)=(Qt(x)>0)=(xΣt).\overline{L}_{t}(x)={\mathbb{P}}(Q_{t}(x)>0)={\mathbb{P}}(x\in\Sigma_{t}).

Therefore,

(3.32) R¯=λ0𝔼|Σs|dαeαs𝑑s,\overline{R}=\lambda\int_{0}^{\infty}{\mathbb{E}}|\Sigma_{s}|_{d}\,\alpha e^{-\alpha s}\,ds,

where the integral is the mean volume of the sausage Σ\Sigma up to time T0T_{0}. This formula is easily obtained from first principles applied to the ρ=\rho=\infty delayed diffusion process (see Section 3.6).

Example 3.9 (Bounded motion).

If, in addition to the assumptions above, each particle is confined within some given distance Δ<\Delta<\infty of its initial location, then ΣtS(Δ+rad(M))\Sigma_{t}\subseteq S(\Delta+\text{\rm rad}(M)). Therefore, by (3.31)–(3.32),

(3.33) R(ρ)R¯λ|S(Δ+rad(M))|d.R(\rho)\leq\overline{R}\leq\lambda\bigl{|}S(\Delta+\text{\rm rad}(M))\bigr{|}_{d}.

If the right side of (3.33) is strictly less than 11, then θdd(λ,ρ,α)=0\theta_{\text{\rm dd}}(\lambda,\rho,\alpha)=0 for ρ(0,)\rho\in(0,\infty) by Proposition 3.8. This is an improvement over (3.25) for large ρ\rho.

3.6. A condition for subcriticality when ρ=\rho=\infty

Let d2d\geq 2, ρ=\rho=\infty, and μ=1M\mu=1_{M} with MM compact. The argument of Sections 3.43.5 is easily adapted subject to a condition on the volume of the sausage Σ\Sigma of (3.30), namely

(3.34) Cγ,σ: for t0t\geq 0, 𝔼|Σt|dγeσt{\mathbb{E}}|\Sigma_{t}|_{d}\leq\gamma e^{\sigma t},

for some γ,σ[0,)\gamma,\sigma\in[0,\infty). Let

(3.35) R()=λ0𝔼|Σs|dαeαs𝑑s,R(\infty)=\lambda\int_{0}^{\infty}{\mathbb{E}}|\Sigma_{s}|_{d}\,\alpha e^{-\alpha s}\,ds,

in agreement with (3.31)–(3.32). Note that R()R(\infty) equals the mean number of points of the Poisson process Π{0}\Pi\setminus\{0\} lying in the sausage ΣT\Sigma_{T}, where TT is independent of Σ\Sigma and is exponentially distributed with parameter α\alpha.

Theorem 3.10.

  • (a)

    If R()<1R(\infty)<1 then θdd(λ,,α)=0\theta_{\text{\rm dd}}(\lambda,\infty,\alpha)=0.

  • (b)

    Assume condition Cγ,σ of (3.34) holds, and λ<λ¯:=1/γ\lambda<\underline{\lambda}:=1/\gamma. If α>α¯:=σ/(1λγ)\alpha>\underline{\alpha}:=\sigma/(1-\lambda\gamma), then R()<1R(\infty)<1 for α>α¯\alpha>\underline{\alpha}.

Proof.

(a) This holds by the argument of Proposition 3.8 adapted to the case ρ=\rho=\infty.

(b) Subject to condition (3.34) with λγ<1\lambda\gamma<1,

(3.36) R()λ0αγe(ασ)s𝑑s=λαγασ,α>σ,R(\infty)\leq\lambda\int_{0}^{\infty}\alpha\gamma e^{-(\alpha-\sigma)s}\,ds=\frac{\lambda\alpha\gamma}{\alpha-\sigma},\qquad\alpha>\sigma,

and the second claim follows. ∎

Example 3.11 (Brownian motion with d=2d=2).

Suppose d=2d=2, ζ\zeta is a standard Brownian motion, and M=SM=S. By (3.35) and the results of Spitzer [32, p. 117],

R()\displaystyle R(\infty) =λ|S|2+λ0αeαs2S(xΣs)𝑑x𝑑s\displaystyle=\lambda|S|_{2}+\lambda\int_{0}^{\infty}\alpha e^{-\alpha s}\int_{{\mathbb{R}}^{2}\setminus S}{\mathbb{P}}(x\in\Sigma_{s})\,dx\,ds
=λπ+λ2SK0(x22α)K0(2α)𝑑x=λZα,\displaystyle=\lambda\pi+\lambda\int_{{\mathbb{R}}^{2}\setminus S}\frac{K_{0}(\|x\|_{2}\sqrt{2\alpha})}{K_{0}(\sqrt{2\alpha)}}\,dx=\lambda Z_{\alpha},

where

(3.37) Zα=π+2παK1(2α)K0(2α)=π+2πα+o(α12)as α.Z_{\alpha}=\pi+\frac{2\pi}{\sqrt{\alpha}}\frac{K_{1}(\sqrt{2\alpha})}{K_{0}(\sqrt{2\alpha})}=\pi+\frac{2\pi}{\sqrt{\alpha}}+{\mathrm{o}}(\alpha^{-\frac{1}{2}})\qquad\text{as }\alpha\to\infty.

Here, K1K_{1} (respectively, K0K_{0}) is the modified Bessel function of the second kind of order 11 (respectively, order 0) given by

K0(x)=0excoshs𝑑s,K1(x)=0excoshscoshsds.K_{0}(x)=\int_{0}^{\infty}e^{-x\cosh s}\,ds,\qquad K_{1}(x)=\int_{0}^{\infty}e^{-x\cosh s}\cosh s\,ds.

Therefore, if λ<λ¯:=1/π\lambda<\underline{\lambda}:=1/\pi, there exists α¯(0,)\underline{\alpha}\in(0,\infty) such that R()<1R(\infty)<1 when α>α¯\alpha>\underline{\alpha}.

Example 3.12 (Brownian motion with d5d\geq 5).

Suppose d5d\geq 5, ζ\zeta is a standard Brownian motion, and M=SM=S. Getoor [8, Thm 2] has shown an explicit constant CC such that

𝔼|Σt|dtcdCas t,{\mathbb{E}}|\Sigma_{t}|_{d}-tc_{d}\uparrow C\qquad\text{as }t\to\infty,

where cdc_{d} is the Newtonian capacity of the closed unit ball SS of d{\mathbb{R}}^{d}. By (3.35),

R()λ(cdα+C).R(\infty)\leq\lambda\left(\frac{c_{d}}{\alpha}+C\right).

Therefore, if λ<λ¯:=1/C\lambda<\underline{\lambda}:=1/C, there exists α¯(0,)\underline{\alpha}\in(0,\infty) such that R()<1R(\infty)<1 when α>α¯\alpha>\underline{\alpha}. Related estimates are in principle valid for d=3,4d=3,4, though the behaviour of 𝔼|Σt|dtcd{\mathbb{E}}|\Sigma_{t}|_{d}-tc_{d} is more complicated (see [8]).

Example 3.13 (Brownian motion with constant drift).

Let d2d\geq 2, M=SM=S, with ζ\zeta a Brownian motion with constant drift. It is standard (with a simple proof using subadditivity) that the limit γ:=𝔼|Σt|d/t\gamma:={\mathbb{E}}|\Sigma_{t}|_{d}/t exists and in addition is strictly positive when the drift is non-zero. Thus, for ϵ>0\epsilon>0, there exists CϵC_{\epsilon} such that

𝔼|Σt|dCϵ+(1+ϵ)γt,t0.{\mathbb{E}}|\Sigma_{t}|_{d}\leq C_{\epsilon}+(1+\epsilon)\gamma t,\qquad t\geq 0.

As in Example 3.12, if λ<λ¯:=1/Cϵ\lambda<\underline{\lambda}:=1/C_{\epsilon}, there exists α¯(0,)\underline{\alpha}\in(0,\infty) such that R()<1R(\infty)<1 when α>α¯\alpha>\underline{\alpha}. See also [13, 14].

Example 3.14 (Ornstein–Uhlenbeck process).

Let M=SM=S and consider the Ornstein–Uhlenbeck process on d{\mathbb{R}}^{d} satisfying

dζ(t)=Aζ(t)dt+dWtd\zeta(t)=A\zeta(t)\,dt+dW_{t}

where WW is standard Brownian motion, AA is a d×dd\times d real matrix, and ζ(0)=0\zeta(0)=0. The solution to this stochastic differential equation is

ζ(t)=0teA(ts)𝑑Ws,\zeta(t)=\int_{0}^{t}e^{A(t-s)}\,dW_{s},

so that ζ(t)de|A|tXtd\|\zeta(t)\|_{d}\leq e^{|A|t}\|X_{t}\|_{d} where

Xt=0teAs𝑑WsX_{t}=\int_{0}^{t}e^{-As}\,dW_{s}

defines a martingale, with |A||A| denoting operator norm. By the Burkholder–Davis–Gundy inequality applied to XX (see, for example, [23, Thm 1.1]), the function Mt=sup{ζ(s)d:s[0,t]}M_{t}=\sup\{\|\zeta(s)\|_{d}:s\in[0,t]\} satisfies

𝔼(Mtd)ced|A|t(0te2|A|s𝑑s)d/2,{\mathbb{E}}(M_{t}^{d})\leq ce^{d|A|t}\left(\int_{0}^{t}e^{2|A|s}\,ds\right)^{d/2},

for some c<c<\infty. Now,

|Σt|d2d(1+Mt)d,|\Sigma_{t}|_{d}\leq 2^{d}(1+M_{t})^{d},

whence Condition Cγ,σ holds for suitable γ,σ<\gamma,\sigma<\infty.

3.7. Proof of Theorem 3.1

This is proved in several stages, as described in the next subsections.

  • §3.7.1

    The existence and some basic properties of αc(λ,ρ)\alpha_{\text{\rm c}}(\lambda,\rho) are proved.

  • §3.7.2

    Let d=2d=2. Suppose μ=1S\mu=1_{S}. The remaining properties of αc\alpha_{\text{\rm c}} are established in the respective cases ρ=\rho=\infty and ρ(0,)\rho\in(0,\infty).

  • §3.7.3

    The previous results are proved for general μ\mu and ρ(0,)\rho\in(0,\infty).

  • §3.7.4

    Corresponding statements are proved for d3d\geq 3.

3.7.1. Existence of αc\alpha_{\text{\rm c}}

Consider the Brownian delayed diffusion model with d2d\geq 2, ρ(0,]\rho\in(0,\infty]. When ρ=\rho=\infty, we assume in addition that

(3.38) μ(x)=1S(x),x2,\mu(x)=1_{S}(x),\qquad x\in{\mathbb{R}}^{2},

where SS is the closed unit ball with centre at the origin; note in this case that μ\mu is radially decreasing.

By Proposition 3.4(a, d), θdd(λ,ρ,α)\theta_{\text{\rm dd}}(\lambda,\rho,\alpha) is non-decreasing in ρ\rho, and non-increasing in α\alpha, and is moreover non-decreasing in λ\lambda if ρ=\rho=\infty (the radial monotonicity of μ\mu has been used in this case). With

αc(λ,ρ):=inf{α:θdd(λ,ρ,α)=0},\alpha_{\text{\rm c}}(\lambda,\rho):=\inf\bigl{\{}\alpha:\theta_{\text{\rm dd}}(\lambda,\rho,\alpha)=0\bigr{\}},

we have that

θdd(λ,ρ,α){>0if α<αc(λ,ρ),=0if α>αc(λ,ρ),\theta_{\text{\rm dd}}(\lambda,\rho,\alpha)\begin{cases}>0&\text{if }\alpha<\alpha_{\text{\rm c}}(\lambda,\rho),\\ =0&\text{if }\alpha>\alpha_{\text{\rm c}}(\lambda,\rho),\end{cases}

and, furthermore, αc\alpha_{\text{\rm c}} is non-decreasing in ρ\rho.

In case (a) of the theorem, by Proposition 3.8, αc(λ,ρ)<\alpha_{\text{\rm c}}(\lambda,\rho)<\infty for all λ\lambda, ρ\rho. In case (b), by Theorem 3.10 and Example 3.13, there exists λ¯(0,λc]\underline{\lambda}\in(0,\lambda_{\text{\rm c}}] such that αc(λ,)<\alpha_{\text{\rm c}}(\lambda,\infty)<\infty when λ(0,λ¯)\lambda\in(0,\underline{\lambda}). As remarked after (1.2), αc(λ,)=\alpha_{\text{\rm c}}(\lambda,\infty)=\infty when λ>λc\lambda>\lambda_{\text{\rm c}}.

It remains to show that αc(λ,ρ)>0\alpha_{\text{\rm c}}(\lambda,\rho)>0 for all λ(0,)\lambda\in(0,\infty), ρ(0,]\rho\in(0,\infty], and the rest of this proof is devoted to that. This will be achieved by comparison with a directed site percolation model on 02{\mathbb{Z}}_{0}^{2} viewed as a directed graph with edges directed away from the origin. When d=2d=2, the key fact is the recurrence of Brownian motion, which permits a static block argument. This fails when d3d\geq 3, in which case we employ a dynamic block argument and the transience of Brownian motion.

3.7.2. The case d=2d=2 with μ=1S\mu=1_{S}

Assume first that d=2d=2, for which we use a static block argument. Let ϵ>0\epsilon>0. We choose a>0a>0 such that

(3.39) (ΠaS)>1ϵ,{\mathbb{P}}(\Pi^{\prime}\cap aS\neq\varnothing)>1-\epsilon,

where Π=Π{0}\Pi^{\prime}=\Pi\setminus\{0\}. For 𝐱2{\mathbf{x}}\in{\mathbb{Z}}^{2}, let S𝐱=3a𝐱+aSS_{\mathbf{x}}=3a{\mathbf{x}}+aS be the ball with radius aa and centre at 3a𝐱3a{\mathbf{x}}. We declare 𝐱{\mathbf{x}} occupied if ΠS𝐱\Pi\cap S_{\mathbf{x}}\neq\varnothing, and vacant otherwise; thus, the origin 0 is invariably occupied. Note that the occupied/vacant states of different 𝐱{\mathbf{x}} are independent. If a given 𝐱0{\mathbf{x}}\neq 0 is occupied, we let Q𝐱ΠS𝐱Q_{\mathbf{x}}\in\Pi\cap S_{\mathbf{x}} be the least such point in the lexicographic ordering, and we set Q0=0Q_{0}=0. If 𝐱{\mathbf{x}} is occupied, we denote by ζ𝐱\zeta_{{\mathbf{x}}} the diffusion associated with the particle at Q𝐱Q_{\mathbf{x}}, and T𝐱T_{\mathbf{x}} for the lifetime of this particle.

Let ζ\zeta be a standard Brownian motion on 2{\mathbb{R}}^{2} with ζ(0)=0\zeta(0)=0, and let

(3.40) wst(ζ):=s[0,t][ζ(s)+S],t[0,),\textsc{ws}_{t}(\zeta):=\bigcup_{s\in[0,t]}\bigl{[}\zeta(s)+S\bigr{]},\qquad t\in[0,\infty),

be the corresponding Wiener sausage.

Suppose for now that ρ=\rho=\infty; later we explain how to handle the case ρ<\rho<\infty. First we explain what it means to say that the origin 0 is open. Let

F(ζ,z)=inf{t:zwst(ζ)},z2,F(\zeta,z)=\inf\{t:z\in\textsc{ws}_{t}(\zeta)\},\qquad z\in{\mathbb{R}}^{2},

be the first hitting time of zz by ws(ζ)\textsc{ws}(\zeta).

For 𝐲2{\mathbf{y}}\in{\mathbb{Z}}^{2}, we define the event

K(ζ0,𝐲)=zS𝐲{F(ζ0,z)<T0},K(\zeta_{0},{\mathbf{y}})=\bigcap_{z\in S_{\mathbf{y}}}\{F(\zeta_{0},z)<T_{0}\},

and

K(ζ0)=𝐲NK(ζ0,𝐲),K(\zeta_{0})=\bigcap_{{\mathbf{y}}\in N}K(\zeta_{0},{\mathbf{y}}),

where N={(0,1),(1,0)}N=\{(0,1),(1,0)\} is the neighbour set of 0 in the directed graph on 02{\mathbb{Z}}_{0}^{2}. By the recurrence of ζ0\zeta_{0}, we may choose α>0\alpha>0 sufficiently small that

(3.41) pα(0):=(K(ζ0))satisfiespα(0)>1ϵ.p_{\alpha}(0):={\mathbb{P}}(K(\zeta_{0}))\quad\text{satisfies}\quad p_{\alpha}(0)>1-\epsilon.

We call 0 open if the event K(ζ0)K(\zeta_{0}) occurs. If 0 is not open, it is called closed. (Recall that 0 is automatically occupied.)

We now explain what is meant by declaring 𝐱2{0}{\mathbf{x}}\in{\mathbb{Z}}^{2}\setminus\{0\} to be open. Assume 𝐱{\mathbf{x}} is occupied and pick Q𝐱Q_{\mathbf{x}} as above. For 𝐲𝐱+N{\mathbf{y}}\in{\mathbf{x}}+N, we define the event

(3.42) K(ζ𝐱,𝐲)=zS𝐲{F(Q𝐱+ζ𝐱,z)<T𝐱},K(\zeta_{\mathbf{x}},{\mathbf{y}})=\bigcap_{z\in S_{\mathbf{y}}}\{F(Q_{\mathbf{x}}+\zeta_{\mathbf{x}},z)<T_{\mathbf{x}}\},

and

K(ζ𝐱)=𝐲NK(ζ𝐱,𝐲).K(\zeta_{\mathbf{x}})=\bigcap_{{\mathbf{y}}\in N}K(\zeta_{\mathbf{x}},{\mathbf{y}}).

By the recurrence of ζ\zeta, we may choose α\alpha such that

(3.43) pα(𝐱):=(K(ζ𝐱)|𝐱 is occupied)satisfiespα(𝐱)>1ϵ.p_{\alpha}({\mathbf{x}}):={\mathbb{P}}\bigl{(}K(\zeta_{\mathbf{x}})\,\big{|}\,\text{${\mathbf{x}}$ is occupied}\bigr{)}\quad\text{satisfies}\quad p_{\alpha}({\mathbf{x}})>1-\epsilon.

We declare 𝐱2{\mathbf{x}}\in{\mathbb{Z}}^{2} open if 𝐱{\mathbf{x}} is occupied, and in addition the event K(ζ𝐱)K(\zeta_{\mathbf{x}}) occurs. A vertex of 2{\mathbb{Z}}^{2} which is not open is called closed. Conditional on the set of occupied vertices, the open/closed states are independent.

The open/closed state of a vertex 𝐱2{\mathbf{x}}\in{\mathbb{Z}}^{2} depends only on the existence of Q𝐱Q_{\mathbf{x}} and on the diffusion ζ𝐱\zeta_{\mathbf{x}}, whence the open/closed states of different 𝐱2{\mathbf{x}}\in{\mathbb{Z}}^{2} are independent. By (3.39)–(3.41), the configuration of open/closed vertices forms a family of independent Bernoulli random variables with density at least (1ϵ)2(1-\epsilon)^{2}. Choose ϵ>0\epsilon>0 such that (1ϵ)2(1-\epsilon)^{2} exceeds the critical probability of directed site percolation on 02{\mathbb{Z}}_{0}^{2} (cf. [11, Thm 3.30]). With strictly positive probability, the origin is the root of an infinite directed cluster of the latter process. Using the definition of the state ‘open’ for the delayed diffusion model, we conclude that the graph G\vec{G} (of Section 3.2) contains an infinite directed path from the origin with strictly positive probability. The corresponding claim of Theorem 3.1(b) follows by Lemma 3.3.

Suppose now that ρ(0,)\rho\in(0,\infty). We adapt the above argument by redefining the times F(ζ,z)F(\zeta,z) and the events K(ζ)K(\zeta) as follows. Consider first the case of the origin. Let

(3.44) E(ζ,z,t)=|{s[0,t]:zζ(s)+S}|1,z2.E(\zeta,z,t)=\bigl{|}\{s\in[0,t]:z\in\zeta(s)+S\}\bigr{|}_{1},\qquad z\in{\mathbb{R}}^{2}.

Pick F>0F>0 such that eρF<ϵe^{-\rho F}<\epsilon, and write

K¯(ζ0,t)=𝐲N,zS𝐲{E(ζ0,z,t)>F}.\overline{K}(\zeta_{0},t)=\bigcap_{{\mathbf{y}}\in N,\,z\in S_{\mathbf{y}}}\{E(\zeta_{0},z,t)>F\}.

In words, K¯(ζ0,t)\overline{K}(\zeta_{0},t) is the event that the Wiener sausage, started at 0 and run for time tt, contains every zS(0,1)S(1,0)z\in S_{(0,1)}\cup S_{(1,0)} for an aggregate time exceeding FF. It follows that, given that Q𝐲ΠS𝐲Q_{\mathbf{y}}\in\Pi\cap S_{\mathbf{y}} for some 𝐲N{\mathbf{y}}\in N, then P0P_{0} infects Q𝐲Q_{\mathbf{y}} with probability at least 1eρF>1ϵ1-e^{-\rho F}>1-\epsilon.

By elementary properties of a recurrent Brownian motion, we may pick tt and then α=α(t)\alpha=\alpha(t) such that (cf. (3.41))

(3.45) pα(0):=(K¯(ζ0,t){t<T0})satisfiespα(0)>1ϵ.p_{\alpha}(0):={\mathbb{P}}\bigl{(}\overline{K}(\zeta_{0},t)\cap\{t<T_{0}\}\bigr{)}\quad\text{satisfies}\quad p_{\alpha}(0)>1-\epsilon.

Turning to general 𝐱2{0}{\mathbf{x}}\in{\mathbb{Z}}^{2}\setminus\{0\}, a similar construction is valid for an event K¯(ζ𝐱,t)\overline{K}(\zeta_{\mathbf{x}},t) as in (3.45), and we replicate the above comparison with directed percolation with (1ϵ)2(1-\epsilon)^{2} replaced by (1ϵ)3(1-\epsilon)^{3}.

3.7.3. The case d=2d=2 with general μ\mu and ρ<\rho<\infty

We consider next the Brownian delayed diffusion process in two dimensions with infections governed by the pair (ρ,μ)(\rho,\mu), as described in Section 2.2. Assume that ρ(0,)\rho\in(0,\infty) and Int(μ)(0,)\text{\rm Int}(\mu)\in(0,\infty). The basic method is to adapt the arguments of Section 3.7.2. The new ingredient is a proof of a statement corresponding to (3.45), as follows.

Let 𝐲N{\mathbf{y}}\in N and write S𝐲=3a𝐲+aSS_{\mathbf{y}}=3a{\mathbf{y}}+aS as before. For ϵ>0\epsilon>0, pick aa such that (ΠS𝐲)>1ϵ{\mathbb{P}}(\Pi\cap S_{\mathbf{y}}\neq\varnothing)>1-\epsilon. Suppose that ΠS𝐲\Pi\cap S_{\mathbf{y}}\neq\varnothing, and write Q:=Q𝐲Q:=Q_{\mathbf{y}} for the least point in the lexicographic ordering of ΠS𝐲\Pi\cap S_{\mathbf{y}}. Consider QQ henceforth as given. The following concerns only two particles, namely P0P_{0} and the particle PP at QQ. Consider the process in which P0P_{0} diffuses forever according to ζ:=ζ0\zeta:=\zeta_{0}, and PP remains stationary. Given ζ\zeta and QQ, let AA be a Poisson process of times (Ak:k=1,2,)(A_{k}:k=1,2,\dots) with rate function r(s)=ρμ(Qζ(s))r(s)=\rho\mu(Q-\zeta(s)). We say that P0P_{0} ‘contacts’ PP at the times of AA, and we claim that

(3.46) (A1<)=1.{\mathbb{P}}(A_{1}<\infty)=1.

This implies that, for ϵ>0\epsilon>0 there exists tt such that (A1<t)>1ϵ{\mathbb{P}}(A_{1}<t)>1-\epsilon, and we may then pick α>0\alpha>0 sufficiently small that (A1<T0)>1ϵ{\mathbb{P}}(A_{1}<T_{0})>1-\epsilon, where T0T_{0} is the lifetime of P0P_{0}. Therefore, subject to (3.46), P0P_{0} infects PP with probability at least 12ϵ1-2\epsilon. This is enough to allow the argument of Section 3.7.2 to proceed, and we turn to the proof of (3.46).

Fix 𝐳2\mathbf{z}\in{\mathbb{R}}^{2} to be chosen soon, and write TbT_{b} for the disk Q𝐳+bSQ-\mathbf{z}+bS. By the Lebesgue density theorem (see, for example, [24, Cor. 2.14]), we may pick 𝐳2\mathbf{z}\in{\mathbb{R}}^{2} and η>0\eta>0 such that

(3.47) T2μ(Q𝐮)𝑑𝐮=𝐳+2Sμ(𝐯)𝑑𝐯ημ(𝐳)>0.\int_{T_{2}}\mu(Q-{\mathbf{u}})\,d{\mathbf{u}}=\int_{\mathbf{z}+2S}\mu({\mathbf{v}})\,d{\mathbf{v}}\geq\eta\mu(\mathbf{z})>0.

We shall suppose without loss of generality that 0T10\notin T_{1}. Let HH be the hitting time (by ζ\zeta) of the disk T1T_{1} and let H>HH^{\prime}>H be the subsequent exit time of the disk T3T_{3}. The probability that P0P_{0} contacts PP during the time-interval (H,H)(H,H^{\prime}) is

(3.48) p:=1𝔼[exp(ρHHμ(Qζ(t))𝑑t)].p:=1-{\mathbb{E}}\left[\exp\left(-\rho\int_{H}^{H^{\prime}}\mu(Q-\zeta(t))\,dt\right)\right].

By spherical symmetry and [26, Thm 3.31],

𝔼HHμ(Qζ(t))𝑑t=T3μ(Q𝐮)G(𝐱,𝐮)𝑑𝐮{\mathbb{E}}\int_{H}^{H^{\prime}}\mu(Q-\zeta(t))\,dt=\int_{T_{3}}\mu(Q-{\mathbf{u}})G({\mathbf{x}},{\mathbf{u}})\,d{\mathbf{u}}

for any given fixed 𝐱T1{\mathbf{x}}\in\partial T_{1}, where GG is the appropriate Green’s function of [26, Lem. 3.36]. There exists c>0c>0 such that G(𝐱,𝐮)cG({\mathbf{x}},{\mathbf{u}})\geq c for 𝐮T2{\mathbf{u}}\in T_{2}, so that

𝔼HHμ(Qζ(t))𝑑tT2μ(Q𝐮)G(𝐱,𝐮)𝑑𝐮cημ(𝐳)>0,\displaystyle{\mathbb{E}}\int_{H}^{H^{\prime}}\mu(Q-\zeta(t))\,dt\geq\int_{T_{2}}\mu(Q-{\mathbf{u}})G({\mathbf{x}},{\mathbf{u}})\,d{\mathbf{u}}\geq c\eta\mu(\mathbf{z})>0,

by (3.47). By (3.48), we have p>0p>0.

We now iterate the above. Each time ζ\zeta revisits T1T_{1}, having earlier departed from T3T_{3}, there is probability pp of such a contact. These contact events are independent, and, by recurrence, a.s. some such contact occurs ultimately. Equation (3.46) is proved.

3.7.4. The case d3d\geq 3.

Let d=3d=3; the case d4d\geq 4 is handled similarly. This time we use a dynamic block argument, combined with Remark 3.5. The idea is the following. Let ζ0\zeta_{0} be the diffusion of particle P0P_{0}. We track the projection of ζ0\zeta_{0}, denoted ζ¯0\overline{\zeta}_{0}, on the plane 2×{0}{\mathbb{R}}^{2}\times\{0\}. By the recurrence of ζ¯0\overline{\zeta}_{0}, the Wiener sausage ws(ζ0)\textsc{ws}(\zeta_{0}) a.s. visits every line 𝐳×\mathbf{z}\times{\mathbb{R}} infinitely often, for 𝐳2\mathbf{z}\in{\mathbb{R}}^{2} (such 𝐳\mathbf{z} will be chosen later). At such a visit, we may choose a point Q𝐳Q_{\mathbf{z}}^{\prime} of Π\Pi lying in ws(ζ0)\textsc{ws}(\zeta_{0}) ‘near to’ the line 𝐳×\mathbf{z}\times{\mathbb{R}}. The construction is then iterated with Q𝐳Q_{\mathbf{z}}^{\prime} as the starting particle. We build this process in each of two independent directions, and may choose the parameter values such that it dominates the cluster at 0 of a supercritical directed site percolation process.

For A3A\subseteq{\mathbb{R}}^{3}, we write A¯\overline{A} for its projection onto the first two coordinates. We abuse notation by identifying 𝐱=(x1,x2,0,,0)3¯{\mathbf{x}}=(x_{1},x_{2},0,\dots,0)\in\overline{{\mathbb{R}}^{3}} (respectively, 3¯\overline{{\mathbb{Z}}^{3}}, etc) with the 22-vector 𝐱=(x1,x2)2{\mathbf{x}}=(x_{1},x_{2})\in{\mathbb{R}}^{2} (respectively, 2{\mathbb{Z}}^{2}, etc). Thus, 3¯=2×{0}\overline{{\mathbb{R}}^{3}}={\mathbb{R}}^{2}\times\{0\} is the plane of the first two coordinates, and similarly 3¯=2×{0}\overline{{\mathbb{Z}}^{3}}={\mathbb{Z}}^{2}\times\{0\}, 03¯=02×{0}\overline{{\mathbb{Z}}_{0}^{3}}={\mathbb{Z}}_{0}^{2}\times\{0\}, and S¯=S3¯\overline{S}=S\cap\overline{{\mathbb{R}}^{3}}.

For 𝐱3¯{\mathbf{x}}\in\overline{{\mathbb{Z}}^{3}}, let S¯𝐱=3a𝐱+aS¯\overline{S}_{\mathbf{x}}=3a{\mathbf{x}}+a\overline{S} be the two-dimensional ball with radius a>1a>1 and centre at 3a𝐱3a{\mathbf{x}}, and let C𝐱=S¯𝐱×C_{\mathbf{x}}=\overline{S}_{\mathbf{x}}\times{\mathbb{R}} be the cylinder generated by 𝐱{\mathbf{x}}. We explain later how aa is chosen. Let ζ=(ζ(i):i=1,2,3)\zeta=(\zeta^{(i)}:i=1,2,3) be a standard Brownian motion in 3{\mathbb{R}}^{3} with ζ(0)=0\zeta(0)=0 and coordinate processes ζ(i)\zeta^{(i)}, and let ζ¯=(ζ(1),ζ(2),0)\overline{\zeta}=(\zeta^{(1)},\zeta^{(2)},0) be its projection onto the first two coordinates. Note that ζ¯\overline{\zeta} is a recurrent process on 3¯\overline{{\mathbb{R}}^{3}}.

We declare the particle at 0 to be open, and let 𝐲N:={(1,0),(0,1)}{\mathbf{y}}\in N:=\{(1,0),(0,1)\}. We shall see that, with a probability to be bounded below, there exists a (random) particle at some Q𝐲C𝐲Q_{\mathbf{y}}\in C_{\mathbf{y}} such that P0P_{0} infects this particle. If this occurs, we declare 𝐲{\mathbf{y}} to be open. On the event that 𝐲{\mathbf{y}} is open, we may iterate the construction starting at Q𝐲Q_{\mathbf{y}}, to find a number of further random vertices of G\vec{G}. By a comparison with a supercritical directed site percolation model, we shall show (for large α\alpha) that G\vec{G} contains an infinite directed cluster with root 0. The claim then follows by Proposition 3.3 and Remark 3.5.

Refer to caption
Figure 3.1. The Wiener sausage ws(ζ0)\textsc{ws}(\zeta_{0}) stopped when it hits the line (3a𝐲)×(3a{\mathbf{y}})\times{\mathbb{R}}. The dark shaded areas constitute the region L(ζ0,𝐲)L(\zeta_{0},{\mathbf{y}}).

Suppose for now that ρ=\rho=\infty. Let ϵ>0\epsilon>0. With ζ\zeta a standard Brownian motion on 3{\mathbb{R}}^{3} with ζ(0)=0\zeta(0)=0, let wst(ζ)\textsc{ws}_{t}(\zeta) be the corresponding Wiener sausage (3.40). We explain next the state open/closed for a vertex 𝐲N{\mathbf{y}}\in N. Let

(3.49) F(ζ0,𝐲)=inf{t:((3a𝐲)×)wst(ζ0)}.F(\zeta_{0},{\mathbf{y}})=\inf\big{\{}t:((3a{\mathbf{y}})\times{\mathbb{R}})\cap\textsc{ws}_{t}(\zeta_{0})\neq\varnothing\bigr{\}}.

Since ζ¯0\overline{\zeta}_{0} is recurrent, we have F(ζ0,𝐲)<F(\zeta_{0},{\mathbf{y}})<\infty a.s. Let T0T_{0} be the lifetime of P0P_{0}, and define the event

(3.50) K(ζ0,𝐲)={F(ζ0,𝐲)<T0}.K(\zeta_{0},{\mathbf{y}})=\bigl{\{}F(\zeta_{0},{\mathbf{y}})<T_{0}\bigr{\}}.

We explain next how aa is chosen (see Figure 3.1). Let a>1a>1 and, for 𝐲N{\mathbf{y}}\in N, consider the intersection

L(ζ0,𝐲):=wsF(ζ0,𝐲)(ζ0)C𝐲.L(\zeta_{0},{\mathbf{y}}):=\textsc{ws}_{F(\zeta_{0},{\mathbf{y}})}(\zeta_{0})\cap C_{\mathbf{y}}.
Lemma 3.15.

There exists c>0c>0 such that the volume of L(ζ0,𝐲)L(\zeta_{0},{\mathbf{y}}) satisfies

|L(ζ0,𝐲)|3ca.|L(\zeta_{0},{\mathbf{y}})|_{3}\geq ca.
Proof.

The set L(ζ0,𝐲)L(\zeta_{0},{\mathbf{y}}) is the union of disjoint subsets of the Wiener sausage, exactly one of which, denoted LL^{\prime}, touches the line (3a𝐲)×(3a{\mathbf{y}})\times{\mathbb{R}}. The volume of LL^{\prime} is bounded below by the volume of the union of a cylinder with radius 11 and length a1a-1, and a half-sphere with radius 11. Thus,

|L(ζ0,𝐲)|3(a1)π+23π23πa,|L(\zeta_{0},{\mathbf{y}})|_{3}\geq(a-1)\pi+\tfrac{2}{3}\pi\geq\tfrac{2}{3}\pi a,

whence the lemma holds with c=23πc=\frac{2}{3}\pi. ∎

By Lemma 3.15, we may pick a>1a>1 sufficiently large that

(N𝐲K(ζ0,𝐲))>1ϵwhereN𝐲:={ΠL(ζ0,𝐲)}.{\mathbb{P}}\bigl{(}N_{\mathbf{y}}\mid K(\zeta_{0},{\mathbf{y}})\bigr{)}>1-\epsilon\qquad\text{where}\qquad N_{\mathbf{y}}:=\{\Pi\cap L(\zeta_{0},{\mathbf{y}})\neq\varnothing\}.

If ΠL(ζ0,𝐲)\Pi\cap L(\zeta_{0},{\mathbf{y}})\neq\varnothing, we pick the least point in the intersection (in lexicographic order) and denote it Q𝐲Q_{\mathbf{y}}, and we say that Q𝐲Q_{\mathbf{y}} has been occupied from 0. We call 𝐲{\mathbf{y}} open if K(ζ0,𝐲)N𝐲K(\zeta_{0},{\mathbf{y}})\cap N_{\mathbf{y}} occurs, and closed otherwise.

By the recurrence of ζ¯\overline{\zeta}, we may choose α>0\alpha>0 such that, for 𝐲N{\mathbf{y}}\in N,

(3.51) pα(𝐲):=(𝐲 is open)satisfiespα(𝐲)>1ϵ.p_{\alpha}({\mathbf{y}}):={\mathbb{P}}(\text{${\mathbf{y}}$ is open})\quad\text{satisfies}\quad p_{\alpha}({\mathbf{y}})>1-\epsilon.

In order to define the open/closed states of other 𝐱3¯{\mathbf{x}}\in\overline{{\mathbb{Z}}^{3}}, it is necessary to generalize the above slightly, and we do this next. Instead of considering a Brownian motion ζ\zeta starting at ζ(0)=0\zeta(0)=0, we move the starting point to some q3¯q\in\overline{{\mathbb{R}}^{3}}. Thus ζ\zeta becomes q+ζq+\zeta, and (3.49)–(3.50) become

F(ζ,q,𝐲)\displaystyle F(\zeta,q,{\mathbf{y}}) =inf{t:((3a𝐲)×)(q+wst(ζ))},\displaystyle=\inf\big{\{}t:((3a{\mathbf{y}})\times{\mathbb{R}})\cap(q+\textsc{ws}_{t}(\zeta))\neq\varnothing\bigr{\}},
K(ζ,q,𝐲,T)\displaystyle K(\zeta,q,{\mathbf{y}},T) ={F(ζ,q,𝐲)<T}.\displaystyle=\{F(\zeta,q,{\mathbf{y}})<T\}.

By the recurrence of ζ¯\overline{\zeta}, we may choose α\alpha such that

(3.52) p¯α(𝐲):=inf{(K(ζ0,q,𝐲,T0)):qS¯}satisfiesp¯α(𝐲)>1ϵ.\overline{p}_{\alpha}({\mathbf{y}}):=\inf\bigl{\{}{\mathbb{P}}(K(\zeta_{0},q,{\mathbf{y}},T_{0})):q\in\overline{S}\bigr{\}}\quad\text{satisfies}\quad\overline{p}_{\alpha}({\mathbf{y}})>1-\epsilon.

The extra notation introduced above will be used at the next stage.

We construct a non-decreasing sequence pair (Vn,Wn)(V_{n},W_{n}) of disjoint subsets of 03¯\overline{{\mathbb{Z}}_{0}^{3}} in the following way. The set VnV_{n} is the set of vertices known to be open at stage nn of the construction, and WnW_{n} is the set known to be closed. Our target is to show that the VnV_{n} dominate some supercritical percolation process.

The vertices of 03¯\overline{{\mathbb{Z}}_{0}^{3}} are ordered in L1L^{1} order: for 𝐱=(x1,x2){\mathbf{x}}=(x_{1},x_{2}), 𝐲=(y1,y2){\mathbf{y}}=(y_{1},y_{2}), we declare

𝐱<𝐲ifeither x1+x2<y1+y2, or x1+x2=y1+y2 and x1<y1.{\mathbf{x}}<{\mathbf{y}}\qquad\text{if}\qquad\text{either $x_{1}+x_{2}<y_{1}+y_{2}$,\quad or $x_{1}+x_{2}=y_{1}+y_{2}$ and $x_{1}<y_{1}$}.

Let Gn={(x1,x2)02:x1+x2=n}G_{n}=\{(x_{1},x_{2})\in{\mathbb{Z}}_{0}^{2}:x_{1}+x_{2}=n\}, and call GnG_{n} the nnth generation of 02{\mathbb{Z}}_{0}^{2}.

First, let

V0={0},W0=.V_{0}=\{0\},\qquad W_{0}=\varnothing.

We choose the least 𝐲N{\mathbf{y}}\in N, and set:

if 𝐲{\mathbf{y}} is open: V1=V0{𝐲},W1=W0,\displaystyle V_{1}=V_{0}\cup\{{\mathbf{y}}\},\,W_{1}=W_{0},
otherwise: V1=V0,W1=W0{𝐲}.\displaystyle V_{1}=V_{0},\,W_{1}=W_{0}\cup\{{\mathbf{y}}\}.

In the first case, we say that ‘𝐲{\mathbf{y}} is occupied from 0’.

For A03¯A\subset\overline{{\mathbb{Z}}_{0}^{3}}, let ΔA\Delta A be the set of vertices b03¯Ab\in\overline{{\mathbb{Z}}_{0}^{3}}\setminus A such that bb has some neighbour aAa\in A with a<ba<b. Suppose (Vk,Wk)(V_{k},W_{k}) have been defined for k=1,2,,nk=1,2,\dots,n, and define (Vn+1,Wn+1)(V_{n+1},W_{n+1}) as follows. Select the least 𝐳ΔVnWn\mathbf{z}\in\Delta V_{n}\setminus W_{n}. If such 𝐳\mathbf{z} exists, find the least 𝐱Vn{\mathbf{x}}\in V_{n} such that 𝐳=𝐱+𝐲\mathbf{z}={\mathbf{x}}+{\mathbf{y}} for some 𝐲N{\mathbf{y}}\in N. Thus 𝐱{\mathbf{x}} is known to be open, and there exists a vertex of G\vec{G} at the point Q𝐱C𝐱Q_{\mathbf{x}}\in C_{\mathbf{x}}.

As above,

L(ζ𝐱,Q𝐱,𝐳)\displaystyle L(\zeta_{\mathbf{x}},Q_{\mathbf{x}},\mathbf{z}) :=wsF(ζ𝐱,Q𝐱,𝐲)(Q𝐱+ζ𝐱)C𝐳,\displaystyle:=\textsc{ws}_{F(\zeta_{\mathbf{x}},Q_{\mathbf{x}},{\mathbf{y}})}(Q_{\mathbf{x}}+\zeta_{\mathbf{x}})\cap C_{\mathbf{z}},
N𝐳\displaystyle N_{\mathbf{z}} :={ΠL(ζ𝐱,Q𝐱,𝐲)}.\displaystyle:=\{\Pi\cap L(\zeta_{\mathbf{x}},Q_{\mathbf{x}},{\mathbf{y}})\neq\varnothing\}.

If K(ζ𝐱,Q𝐱,𝐳,T𝐱)N𝐳K(\zeta_{\mathbf{x}},Q_{\mathbf{x}},\mathbf{z},T_{\mathbf{x}})\cap N_{\mathbf{z}} occurs we call 𝐳\mathbf{z} open, and we say that 𝐳\mathbf{z} is occupied from 𝐱{\mathbf{x}}; otherwise we say that 𝐳\mathbf{z} is closed.

If 𝐳\mathbf{z} is open: Vn+1=Vn{𝐳},Wn+1=Wn,\displaystyle V_{n+1}=V_{n}\cup\{\mathbf{z}\},\,W_{n+1}=W_{n},
otherwise: Vn+1=Vn,Wn+1=Wn{𝐳}.\displaystyle V_{n+1}=V_{n},\,W_{n+1}=W_{n}\cup\{\mathbf{z}\}.

By (3.51)–(3.52), the vertex 𝐳\mathbf{z} under current scrutiny is open with conditional probability at least (1ϵ)2(1-\epsilon)^{2}.

This process is iterated until the earliest stage at which no such 𝐳\mathbf{z} exists. If this occurs for some n<n<\infty, we declare Vm=VnV_{m}=V_{n} for mnm\geq n, and in any case we set V=limmVmV_{\infty}=\lim_{m\to\infty}V_{m}.

The resulting set VV_{\infty} is the cluster at the origin of a type of dependent directed site percolation process which is built by generation-number. Having discovered the open vertices 𝐳\mathbf{z} in generation nn together with the associated points Q𝐳Q_{\mathbf{z}}, the law of the next generation is (conditionally) independent of the past and is 11-dependent.

We now apply a stochastic-domination argument. Such methods have been used since at least [6], and the following core lemma was systematized by Liggett, Schonmann, and Stacey [22, Thm 0.0] (see also [10, Thm 7.65] and the references therein). Let δ(0,1)\delta\in(0,1), and let 𝐗=(Xx:𝐱02){\mathbf{X}}=(X_{x}:{\mathbf{x}}\in{\mathbb{Z}}_{0}^{2}) be a 11-dependent family of Bernoulli random variables such that 𝔼(X𝐱)>1δ{\mathbb{E}}(X_{\mathbf{x}})>1-\delta for all 𝐱{\mathbf{x}}. There exists η(δ)>0\eta(\delta)>0, satisfying η(δ)0\eta(\delta)\to 0 as δ 0\delta\to\ 0, such that 𝐗{\mathbf{X}} dominates stochastically a family 𝐘=(Yx:𝐱02)\mathbf{Y}=(Y_{x}:{\mathbf{x}}\in{\mathbb{Z}}_{0}^{2}) of independent Bernoulli variables with parameter 1η(δ)1-\eta(\delta). We choose δ>0\delta>0 such that 1η(δ)1-\eta(\delta) exceeds the critical probability of directed site percolation on 02{\mathbb{Z}}_{0}^{2}. By the above, for sufficiently small δ>0\delta>0, there is strictly positive probability of an infinite directed path on 02{\mathbb{Z}}_{0}^{2} comprising vertices 𝐱{\mathbf{x}} with X𝐱=1X_{\mathbf{x}}=1.

With δ\delta chosen thus and ϵ=δ/2\epsilon=\delta/2, we deduce as required that (|V|=)>0{\mathbb{P}}(|V_{\infty}|=\infty)>0. By a consideration of the geometry of the above construction, and the definition of the local states open/occupied, by (3.10) this entails θdd(λ,,α)>0\theta_{\text{\rm dd}}(\lambda,\infty,\alpha)>0.

A minor extra complication arises at the last stage, in that the events {𝐱 is open}\{{\mathbf{x}}\text{ is open}\} are not 11-dependent, but only 11-dependent within a given generation conditional on earlier generations. This may be viewed as follows. Begin with a family 𝐘=(Yx:𝐱02)\mathbf{Y}=(Y_{x}:{\mathbf{x}}\in{\mathbb{Z}}_{0}^{2}) of Bernoulli variables with parameter 1η(δ)1-\eta(\delta). Having constructed the subsequence (V0,V1,,Vn1)(V_{0},V_{1},\dots,V_{n-1}), the set VnV_{n} (or more precisely the set of its indicator functions) dominates stochastically the nnth generation of 𝐘\mathbf{Y}. This holds inductively for all nn, and the claim follows.

When ρ(0,)\rho\in(0,\infty), we extend the earlier argument (around (3.50) and later). Rather than presenting all the required details, we consider the special case of (3.50); the general case is similar. Let 𝐲N{\mathbf{y}}\in N and Xt:=wst(ζ0)C𝐲X_{t}:=\textsc{ws}_{t}(\zeta_{0})\cap C_{\mathbf{y}}. We develop the previous reference to the first hitting time F(ζ0,𝐲)F(\zeta_{0},{\mathbf{y}}) with a consideration of the limit set X=limtXtX_{\infty}=\lim_{t\to\infty}X_{t}. Since ζ¯0\overline{\zeta}_{0} is recurrent and ζ0\zeta_{0} is transient, there exists a deterministic η>0\eta>0 such that:

  • (a)

    a.s., XX_{\infty} contains infinitely many disjoint closed connected regions R1,R2,R_{1},R_{2},\dots, each with volume exceeding 12ca\frac{1}{2}ca, and

  • (b)

    every point 𝐱iRi{\mathbf{x}}\in\bigcup_{i}R_{i} is such that

    (3.53) |{t0:𝐱wst(ζ0)}|1η.|\{t\geq 0:{\mathbf{x}}\in\textsc{ws}_{t}(\zeta_{0})\}|_{1}\geq\eta.

Each such region contains a point of Π\Pi with probability at least 1e12λca1-e^{-\frac{1}{2}\lambda ca}. Each such point is infected by P0P_{0} with probability at least 1eρη1-e^{-\rho\eta}. Pick NN such that, in NN independent trials each with probability of success 1e12λcaeρη1-e^{-\frac{1}{2}\lambda ca}-e^{-\rho\eta}, there exists at least one success with probability exceeding 1ϵ1-\epsilon. Finally, pick the deterministic time τ\tau such that there is probability at least 1ϵ1-\epsilon that XτX_{\tau} contains at least NN disjoint closed connected regions RjR_{j}, each with volume exceeding 12ca\frac{1}{2}ca, and such that, for every jj and every 𝐱Rj{\mathbf{x}}\in R_{j}, inequality (3.53) holds.

Finally, we pick α\alpha such that

(T0>τ)1ϵ.{\mathbb{P}}(T_{0}>\tau)\geq 1-\epsilon.

With these choices, the probability that XτX_{\tau} contains some particle that is infected from 0 is at least (1ϵ)3(1-\epsilon)^{3}. The required argument proceeds henceforth as before.

We turn finally to the case of general μ\mu and ρ(0,)\rho\in(0,\infty), and we indicate briefly how the method of Section 3.7.3 may be applied in the current context. First, let 𝐲N{\mathbf{y}}\in N and a>3a>3. It suffices as above to show that, with probability near 11, P0P_{0} infects some particle in C𝐲:=S¯𝐲×C_{\mathbf{y}}:=\overline{S}_{\mathbf{y}}\times{\mathbb{R}} where, as usual, S𝐲=3a𝐲+aSS_{\mathbf{y}}=3a{\mathbf{y}}+aS. The following argument is illustrated in Figure 3.2.

Refer to caption
Figure 3.2. An illustration of the proof that, with strictly positive probability (and, therefore, by iteration, with probability 11) P0P_{0} infects some particle in C𝐲C_{\mathbf{y}}. The arrow denotes the vector 𝐳\mathbf{z}.

Pick 𝐳3\mathbf{z}\in{\mathbb{R}}^{3} and η>0\eta>0 such that the d=3d=3 version of (3.47) holds, namely,

(3.54) 𝐳+Sμ(𝐯)𝑑𝐯ημ(𝐳)>0.\int_{\mathbf{z}+S}\mu({\mathbf{v}})\,d{\mathbf{v}}\geq\eta\mu(\mathbf{z})>0.

By recurrence, the projected diffusion ζ¯0\overline{\zeta}_{0} visits the disk S¯𝐲𝐳\overline{S}_{{\mathbf{y}}-\mathbf{z}} infinitely often, a.s., and therefore ζ0\zeta_{0} visits the tube T:=S¯𝐲𝐳×T:=\overline{S}_{{\mathbf{y}}-\mathbf{z}}\times{\mathbb{R}} similarly. By the transitivity of ζ0\zeta_{0}, its entry points into TT are a.s. unbounded. Following each such entry to TT, at the point 𝐰3{\mathbf{w}}\in{\mathbb{R}}^{3} say, there is an exit from the ball 𝐰+3S{\mathbf{w}}+3S. Let HH be the time of the first such entry and HH^{\prime} the time of the subsequent such exit.

Let ζ>0\zeta>0 denote the volume of the ball SS, so that

(3.55) ([𝐰+𝐳+S]Π)1eλζ.{\mathbb{P}}\bigl{(}[{\mathbf{w}}+\mathbf{z}+S]\cap\Pi\neq\varnothing\bigr{)}\geq 1-e^{-\lambda\zeta}.

On the event that [𝐰+𝐳+S]Π[{\mathbf{w}}+\mathbf{z}+S]\cap\Pi\neq\varnothing, let QQ be the least point in that intersection, so that QC𝐲Q\in C_{\mathbf{y}}. Conditional on QQ, the probability that P0P_{0} infects the particle at QQ during the time-interval (H,H)(H,H^{\prime}) is

(3.56) p:=1𝔼[exp(ρHHμ(Qζ(t))𝑑t)].p:=1-{\mathbb{E}}\left[\exp\left(-\rho\int_{H}^{H^{\prime}}\mu(Q-\zeta(t))\,dt\right)\right].

By spherical symmetry and [26, Thm 3.31], conditional on 𝐰{\mathbf{w}},

𝔼HHμ(Qζ(t))𝑑t=𝐰+3Sμ(Q𝐮)G(𝐰,𝐮)𝑑𝐮,{\mathbb{E}}\int_{H}^{H^{\prime}}\mu(Q-\zeta(t))\,dt=\int_{{\mathbf{w}}+3S}\mu(Q-{\mathbf{u}})G({\mathbf{w}},{\mathbf{u}})\,d{\mathbf{u}},

where GG is the appropriate Green’s function of [26, Lem. 3.32]. There exists c>0c>0 such that G(𝐰,𝐮)cG({\mathbf{w}},{\mathbf{u}})\geq c for 𝐮𝐰+2S{\mathbf{u}}\in{\mathbf{w}}+2S. We make the change of variable 𝐮=Q𝐱+𝐯{\mathbf{u}}=Q-{\mathbf{x}}+{\mathbf{v}}, and note that Q𝐳+S𝐰+2SQ-\mathbf{z}+S\subseteq{\mathbf{w}}+2S, to deduce that

(3.57) 𝔼HHμ(Qζ(t))𝑑tc𝐳+Sμ(𝐯)𝑑𝐯cημ(𝐳)>0,{\mathbb{E}}\int_{H}^{H^{\prime}}\mu(Q-\zeta(t))\,dt\geq c\int_{\mathbf{z}+S}\mu({\mathbf{v}})\,d{\mathbf{v}}\geq c\eta\mu(\mathbf{z})>0,

by (3.54). By (3.56), we have p>0p>0.

On combining (3.55) and (3.57), we deduce that there exists δ>0\delta>0 such that

(3.58) (QΠC𝐲, and P0 infects Q between times H and H)δ.{\mathbb{P}}\bigl{(}\text{$\exists Q\in\Pi\cap C_{\mathbf{y}}$, and $P_{0}$ infects $Q$ between times $H$ and $H^{\prime}$}\bigr{)}\geq\delta.

The proof is completed by using the iterative argument around (3.53).

4. The diffusion model

4.1. A condition for subcriticality

We consider the diffusion model in the general form of Sections 2.1 and 2.3, and we adopt the notation of those sections. Recall the critical point λc\lambda_{\text{\rm c}} of the Boolean continuum percolation on d{\mathbb{R}}^{d} in which a closed unit ball is centred at each point of a rate-λ\lambda Poisson process on d{\mathbb{R}}^{d}. We shall prove the existence of a subcritical phase.

Condition (3.34) is now replaced as follows. Let ζ\zeta^{\prime} be an independent copy of ζ\zeta, and define the sausage

(4.1) Σt:=s[0,t][ζ(s)ζ(s)+S],s0.\Sigma^{\prime}_{t}:=\bigcup_{s\in[0,t]}\bigl{[}\zeta(s)-\zeta^{\prime}(s)+S\bigr{]},\qquad s\geq 0.

We shall assume

(4.2) Cγ,σ{}_{\gamma,\sigma}^{\hskip 1.0pt\prime}: for t0t\geq 0, 𝔼|Σt|dγeσt{\mathbb{E}}|\Sigma^{\prime}_{t}|_{d}\leq\gamma e^{\sigma t},

for some γ,σ[0,)\gamma,\sigma\in[0,\infty), and we make a note about this condition in Remark 4.3.

Let θd(λ,ρ,α)\theta_{\text{\rm d}}(\lambda,\rho,\alpha) be the probability that the diffusion process survives.

Theorem 4.1.

Consider the general diffusion model on d{\mathbb{R}}^{d} where d1d\geq 1.

  • (a)

    Let ρ(0,)\rho\in(0,\infty) and α¯(λ,ρ)=λρInt(μ)\underline{\alpha}(\lambda,\rho)=\lambda\rho\,\text{\rm Int}(\mu). Then θd(λ,ρ,α)=0\theta_{\text{\rm d}}(\lambda,\rho,\alpha)=0 if α>α¯(λ,ρ)\alpha>\underline{\alpha}(\lambda,\rho).

  • (b)

    Let ρ=\rho=\infty and μ=1S\mu=1_{S}. Assume in addition that condition Cγ,σ{}_{\gamma,\sigma}^{\hskip 1.0pt\prime} of (4.2) holds. Let α¯(λ)=σ/(1λγ)\underline{\alpha}(\lambda)=\sigma/(1-\lambda\gamma) and λ¯=1/γ\underline{\lambda}=1/\gamma. Then θd(λ,,α)=0\theta_{\text{\rm d}}(\lambda,\infty,\alpha)=0 if α>α¯(λ)\alpha>\underline{\alpha}(\lambda) and 0<λ<λ¯0<\lambda<\underline{\lambda}.

This theorem extends Theorem 1.2. Its proof is related to that given in Section 3.4 for the delayed diffusion model.

Proof.

(a) Let λ(0,)\lambda\in(0,\infty), and suppose that ρ<\rho<\infty. We shall enhance the probability space on which the diffusion model is defined. Let (Ti:i0)(T_{i}:i\in{\mathbb{Z}}_{0}) be random variables with the exponential distribution with parameter α\alpha; these are independent of one another and of all other random variables so far. We call TiT_{i} the ‘lifetime’ of PiP_{i}, and it is the length of the period between infection and removal of PiP_{i}.

For iji\neq j, we introduce Poisson processes Ai,jA_{i,j} of points in [0,)[0,\infty), and we say that PiP_{i} ‘contacts’ PjP_{j} at the times of Ai,jA_{i,j}. The intensity functions of the Ai,jA_{i,j} depend as follows on the positions of PiP_{i} and PjP_{j}. Conditional on Π\Pi and the diffusions (ζr:r0)(\zeta_{r}:r\in{\mathbb{Z}}_{0}), let (Ai,j:i,j0,ij)(A_{i,j}:i,j\in{\mathbb{Z}}_{0},\,i\neq j) be independent Poisson processes on [0,)[0,\infty) with respective rate functions

ri,j(s):=ρμ(Xj+ζj(s)Xiζi(s)),s0.r_{i,j}(s):=\rho\mu\bigl{(}X_{j}+\zeta_{j}(s)-X_{i}-\zeta_{i}(s)\bigr{)},\qquad s\geq 0.

The points of Ai,jA_{i,j} are denoted (Ai,jk:k0)(A_{i,j}^{k}:k\in{\mathbb{Z}}_{0}). Let

A¯i,j(t):=inf{Ai,j:k0,Ai,jk>t},t>0,\underline{A}_{i,j}(t):=\inf\{A_{i,j}:k\in{\mathbb{Z}}_{0},\,A_{i,j}^{k}>t\},\qquad t>0,

and let Bi,j(t)B_{i,j}(t) be the event that A¯i,j(t)t<Ti\underline{A}_{i,j}(t)-t<T_{i} and PjP_{j} is susceptible at all times A¯i,j(t)ϵ\underline{A}_{i,j}(t)-\epsilon for ϵ>0\epsilon>0. Suppose that PiP_{i} becomes infected at time τ\tau. The first contact by PiP_{i} of PjP_{j} after time τ\tau results in an infection if and only the event Bi,j(τ)B_{i,j}(\tau) occurs (in which case we say that PiP_{i} infects PjP_{j} directly). Write A¯i,j=A¯i,j(0)\underline{A}_{i,j}=\underline{A}_{i,j}(0) and Bi,j=Bi,j(0)B_{i,j}=B_{i,j}(0).

Proposition 3.7 holds with the same proof but with Lt(x)L_{t}(x) replaced by

(4.3) L~t(x)=𝔼(1exp(0tρμ(x+ζ(s)ζ(s))𝑑s)),\widetilde{L}_{t}(x)={\mathbb{E}}\left(1-\exp\left(-\int_{0}^{t}\rho\mu(x+\zeta(s)-\zeta^{\prime}(s))\,ds\right)\right),

where ζ\zeta^{\prime} is an independent copy of ζ\zeta. By the Poisson colouring theorem, L~t(x)\widetilde{L}_{t}(x) equals the probability that P0P_{0} contacts a particle started at xx\in{\mathbb{R}} during the time interval (0,t](0,t]. With this new L~t(x)\widetilde{L}_{t}(x), the new bound R=R(ρ)R=R(\rho) now satisfies

(4.4) R(ρ)=λd0L~s(x)αeαs𝑑s𝑑xλραInt(μ).R(\rho)=\lambda\int_{{\mathbb{R}}^{d}}\int_{0}^{\infty}\widetilde{L}_{s}(x)\alpha e^{-\alpha s}\,ds\,dx\leq\frac{\lambda\rho}{\alpha}\,\text{\rm Int}(\mu).

In other words, R(ρ)R(\rho) is the mean number of particles that P0P_{0} contacts during its lifetime (it is not the mean total number of contacts by P0P_{0}, since P0P_{0} may contact any given particle many times).

By an inductive definition as before, we define the nnth generation InI_{n} of infected particles from 0. We claim that

(4.5) 𝔼|In|R(ρ)n,n1.{\mathbb{E}}|I_{n}|\leq R(\rho)^{n},\qquad n\geq 1.

By (4.5), 𝔼|I|<{\mathbb{E}}|I|<\infty whenever R(ρ)<1R(\rho)<1, and the claim of part (a) follows by (4.4) as in the proof of Proposition 3.8(b, c). We turn therefore to the proof of (4.5), which we prove first with n=1n=1.

Recall that each label i0i\in{\mathbb{Z}}_{0} corresponds to a point XiΠX_{i}\in\Pi, an associated diffusion ζi\zeta_{i}, and a lifetime TiT_{i}. The lifetime TiT_{i} is the residual time to removal of PiP_{i} after its first infection.

We have that

(4.6) |I1|=j0{0}1(B0,j)j0{0}1(A¯0,j<T0),|I_{1}|=\sum_{j\in{\mathbb{Z}}_{0}\setminus\{0\}}1(B_{0,j})\leq\sum_{j\in{\mathbb{Z}}_{0}\setminus\{0\}}1(\underline{A}_{0,j}<T_{0}),

whence, by the remark after (4.4),

(4.7) 𝔼|I1|j0{0}(A¯0,j<T0)=R(ρ),{\mathbb{E}}|I_{1}|\leq\sum_{j\in{\mathbb{Z}}_{0}\setminus\{0\}}{\mathbb{P}}(\underline{A}_{0,j}<T_{0})=R(\rho),

as claimed.

Suppose next that n2n\geq 2. We introduce some further notation. Let i0=0i_{0}=0, and let ı=(i1,i2,,in)\vec{\imath}=(i_{1},i_{2},\dots,i_{n}) be an ordered vector of distinct members of 0{0}{\mathbb{Z}}_{0}\setminus\{0\}; we shall consider ı\vec{\imath} as both a vector and a set. Define the increasing sequence τ(ı)=(τj:0jn)\tau(\vec{\imath})=(\tau_{j}:0\leq j\leq n) of times by

(4.8) τ0=0,τ1=A¯i0,i1,τ2=A¯i1,i2(τ1),,τj+1=A¯ij,ij+1(τj).\tau_{0}=0,\quad\tau_{1}=\underline{A}_{i_{0},i_{1}},\quad\tau_{2}=\underline{A}_{i_{1},i_{2}}(\tau_{1}),\quad\dots,\quad\tau_{j+1}=\underline{A}_{i_{j},i_{j+1}}(\tau_{j}).

By iterating the argument leading to (4.6), we obtain

(4.9) |In|Wn,|I_{n}|\leq W_{n},

where

(4.10) Wn=ıf(ı),W_{n}=\sum_{\vec{\imath}}f(\vec{\imath}),

and

(4.11) f(ı)=1(τ1<Ti0)1(τ2τ1<Ti1)1(τnτn1<Tin1).f(\vec{\imath})=1(\tau_{1}<T_{i_{0}})1(\tau_{2}-\tau_{1}<T_{i_{1}})\cdots 1(\tau_{n}-\tau_{n-1}<T_{i_{n-1}}).

Equations (4.9)–(4.10) are implied by the following observation: if PinInP_{i_{n}}\in I_{n}, then there exists a sequence i0=0,i1,,in1i_{0}=0,i_{1},\dots,i_{n-1} such that, for 0j<n0\leq j<n, PijP_{i_{j}} infects Pij+1P_{i_{j+1}} directly at the time τj+1\tau_{j+1}. See Figure 4.1.

Refer to caption
Figure 4.1. The horizontal axis represents one-dimensional space {\mathbb{R}}, and the vertical axis represents time. This is an illustration of the summand f(0,i1,i2)f(0,i_{1},i_{2}) in (4.10) when d=1d=1. In this conceptual view, infections occur where pairs of diffusions intersect, and times of removal are marked by crosses.

By (4.10),

𝔼(Wn)ı𝔼[(C1C2Cn𝒢(ı))],{\mathbb{E}}(W_{n})\leq\sum_{\vec{\imath}}{\mathbb{E}}\bigl{[}{\mathbb{P}}(C_{1}\cap C_{2}\cap\cdots\cap C_{n}\mid{\mathcal{G}}(\vec{\imath}))\bigr{]},

where Cj={τjτj1<Tij1}C_{j}=\{\tau_{j}-\tau_{j-1}<T_{i_{j-1}}\} and 𝒢(ı){\mathcal{G}}(\vec{\imath}) is the σ\sigma-field generated by the random variables

(Xij,ζij,Tij) for 0j<n1,Xin1,τn1,(ζin1(s):s[0,τn1]).(X_{i_{j}},\zeta_{i_{j}},T_{i_{j}})\text{ for }0\leq j<n-1,\quad X_{i_{n-1}},\ \tau_{n-1},\ (\zeta_{i_{n-1}}(s):s\in[0,\tau_{n-1}]).

Note that C1,C2,,Cn1C_{1},C_{2},\dots,C_{n-1} are 𝒢(ı){\mathcal{G}}(\vec{\imath})-measurable, so that

𝔼(Wn)ı𝔼[1(C1Cn1)(Cn𝒢(ı))].{\mathbb{E}}(W_{n})\leq\sum_{\vec{\imath}}{\mathbb{E}}\bigl{[}1(C_{1}\cap\cdots\cap C_{n-1}){\mathbb{P}}(C_{n}\mid{\mathcal{G}}(\vec{\imath}))\bigr{]}.

Therefore,

(4.12) 𝔼(Wn)𝔼[i1,,in11(C1Cn1)in(Cn𝒢(ı))]{\mathbb{E}}(W_{n})\leq{\mathbb{E}}\left[\sum_{i_{1},\dots,i_{n-1}}1(C_{1}\cap\cdots\cap C_{n-1})\sum_{i_{n}}{\mathbb{P}}(C_{n}\mid{\mathcal{G}}(\vec{\imath}))\right]

where the summations are over distinct i1,,in0i_{1},\dots,i_{n}\neq 0.

It is tempting to argue as follows. The diffusions (ζk:k{i0,,in1})(\zeta_{k}:k\notin\{i_{0},\dots,i_{n-1}\}) are independent of 𝒢(ı){\mathcal{G}}(\vec{\imath}), and τn1\tau_{n-1} is 𝒢(ı){\mathcal{G}}(\vec{\imath})-measurable. By the Poisson displacement theorem (see [20, Sec. 5.2]), the positions Π=(Xk+ζk(τn1):k{i0,,in1})\Pi^{\prime}=(X_{k}+\zeta_{k}(\tau_{n-1}):k\notin\{i_{0},\dots,i_{n-1}\}) are a subset of a rate-λ\lambda Poisson process. It follows that

(4.13) in(Cn𝒢(ı))R(ρ).\sum_{i_{n}}{\mathbb{P}}(C_{n}\mid{\mathcal{G}}(\vec{\imath}))\leq R(\rho).

By (4.10)–(4.13),

(4.14) 𝔼(Wn)𝔼(Wn1)R(ρ).{\mathbb{E}}(W_{n})\leq{\mathbb{E}}(W_{n-1})R(\rho).

Inequality (4.5) follows by iteration and (4.9) There is a subtlety in the argument leading to (4.13), namely that the distribution of the subset (Xk:k{i1,,in1})(X_{k}:k\notin\{i_{1},\dots,i_{n-1}\}) of Π\Pi will generally depend on the choice of i1,,in1i_{1},\dots,i_{n-1}. This may be overcome as follows.

We decouple the indices of particles and their starting positions in a classical way (see [12, Thm 6.13.11]) by giving a more prescriptive recipe for the construction of the Poisson process Π\Pi. Let mm be a positive integer and let Λm=[m,m]dd\Lambda_{m}=[-m,m]^{d}\subset{\mathbb{R}}^{d}; later we shall take the limit as mm\to\infty. Let MM have the Poisson distribution with parameter λ(2m)d\lambda(2m)^{d}. Conditional on MM, let X1,X2,,XMX_{1},X_{2},\dots,X_{M} be independent random variables with the uniform distribution on Λm\Lambda_{m}. Thus, points in ΠΛm\Pi\cap\Lambda_{m} are indexed {0}J\{0\}\cup J where J={1,2,,M}J=\{1,2,\dots,M\}, with P0P_{0} retaining the index 0.

Let

(4.15) Wn(m)=ıJf(ı),W_{n}(m)=\sum_{\vec{\imath}\subseteq J}f(\vec{\imath}),

so that Wn(m)WnW_{n}(m)\to W_{n} as mm\to\infty, and furthermore,

(4.16) 𝔼(Wn(m))𝔼(Wn)as m,{\mathbb{E}}(W_{n}(m))\to{\mathbb{E}}(W_{n})\qquad\text{as }m\to\infty,

by the monotone convergence theorem. The sum Wn(m)W_{n}(m) may be represented in terms of the average of f(Sn)f(\vec{S}_{n}) where Sn\vec{S}_{n} is a random ordered nn-subset of indices in JJ, namely,

(4.17) Wn(m)=𝔼(M!(Mn)!f(Sn)).W_{n}(m)={\mathbb{E}}\left(\frac{M!}{(M-n)!}f(\vec{S}_{n})\right).

The term f(Sn)f(\vec{S}_{n}) is intepreted as 0 if n>Mn>M. With Sn=(s1,s2,,sn)\vec{S}_{n}=(s_{1},s_{2},\dots,s_{n}) and Sn1=(s1,s2,,sn1)\vec{S}_{n-1}=(s_{1},s_{2},\dots,s_{n-1}), we have as in (4.12) that

(4.18) Wn(m)=𝔼(M!(Mn)!f(Sn1)Zn)W_{n}(m)={\mathbb{E}}\left(\frac{M!}{(M-n)!}f(\vec{S}_{n-1})Z_{n}\right)

where

(4.19) Zn=1(τnτn1<Tsn1),\displaystyle Z_{n}=1(\tau_{n}-\tau_{n-1}<T_{s_{n-1}}),

and τ(Sn)=(τ0,τ1,,τn)\tau(\vec{S}_{n})=(\tau_{0},\tau_{1},\dots,\tau_{n}).

For an ordered (n1)(n-1)-subset ı\vec{\imath} of JJ, let R¯(ı)\overline{R}(\vec{\imath}) be the supremum over sΛms\in\Lambda_{m} of the mean number of particles infected by a given initial particle located at ss, in the subset of the rate-λ\lambda Poisson process obtained from Π\Pi by deleting (Xi:i{0}ı)(X_{i}:i\in\{0\}\cup\vec{\imath}). Note that

(4.20) R¯(ı)R(ρ).\overline{R}(\vec{\imath})\leq R(\rho).

By (4.18)–(4.19),

𝔼(Zn𝒢(Sn1),M)\displaystyle{\mathbb{E}}(Z_{n}\mid{\mathcal{G}}(\vec{S}_{n-1}),M) =1Mn+1sJSn1(τnτn1<Tsn1|𝒢(Sn),M)\displaystyle=\frac{1}{M-n+1}\sum_{s\in J\setminus\vec{S}_{n-1}}{\mathbb{P}}\bigl{(}\tau_{n}-\tau_{n-1}<T_{s_{n-1}}\,\big{|}\,{\mathcal{G}}(\vec{S}_{n}),M\bigr{)}
1Mn+1R¯(Sn1).\displaystyle\leq\frac{1}{M-n+1}\overline{R}(\vec{S}_{n-1}).

By (4.18) and (4.20),

Wn(m)\displaystyle W_{n}(m) 𝔼(M!(Mn+1)!f(Sn1)R¯(Sn1))\displaystyle\leq{\mathbb{E}}\left(\frac{M!}{(M-n+1)!}f(\vec{S}_{n-1})\overline{R}(\vec{S}_{n-1})\right)
=ıJf(ı)R¯(ı)Wn(m1)R(ρ).\displaystyle=\sum_{\vec{\imath}\subseteq J}f(\vec{\imath})\overline{R}(\vec{\imath})\leq W_{n}(m-1)R(\rho).

By (4.16), on letting mm\to\infty, we deduce inequality (4.14), and the proof is completed as before.

(b) Let ρ=\rho=\infty. We repeat the argument in the proof of part (a) (cf. Section 3.6) with R()R(\infty) defined as the mean number of particles PjP_{j} for which there exists t<T0t<T_{0} with Xj+ζj(t)ζ0(t)+SX_{j}+\zeta_{j}(t)\in\zeta_{0}(t)+S. That is, with ζ\zeta^{\prime} an independent copy of ζ\zeta,

(4.21) R()\displaystyle R(\infty) =dλ𝑑x(x+ζ(t)ζ(t)S for some t<T0)\displaystyle=\int_{{\mathbb{R}}^{d}}\lambda\,dx\,{\mathbb{P}}\bigl{(}x+\zeta^{\prime}(t)-\zeta(t)\in S\text{ for some }t<T_{0}\bigr{)}
=dλ𝑑x0(xΣs)αeαs𝑑s\displaystyle=\int_{{\mathbb{R}}^{d}}\lambda\,dx\,\int_{0}^{\infty}{\mathbb{P}}(x\in\Sigma_{s}^{\prime})\alpha e^{-\alpha s}\,ds
=λ0t𝔼|Σs|dαeαs𝑑s,\displaystyle=\lambda\int_{0}^{t}{\mathbb{E}}|\Sigma^{\prime}_{s}|_{d}\,\alpha e^{-\alpha s}\,ds,

where Σs\Sigma_{s}^{\prime} is given in (4.1). As in Theorem 3.10(b) adapted to the diffusion model, we have by Cγ,σ{}_{\gamma,\sigma}^{\hskip 1.0pt\prime} that R()<1R(\infty)<1 if λ<λ¯:=1/γ\lambda<\underline{\lambda}:=1/\gamma and α>α¯(λ):=σ/(1λγ)\alpha>\underline{\alpha}(\lambda):=\sigma/(1-\lambda\gamma). By the argument of the proof of part (a), θd(λ,ρ,α)=0\theta_{\text{\rm d}}(\lambda,\rho,\alpha)=0 for λ(0,λ¯)\lambda\in(0,\underline{\lambda}) and α>α¯(λ)\alpha>\underline{\alpha}(\lambda). ∎

Example 4.2 (Bounded motion).

Let ρ=\rho=\infty and μ=1M\mu=1_{M} as above, and suppose in addition that each particle is confined within some given distance Δ<\Delta<\infty of its initial location. By (4.21),

R()λ|S(2(Δ+rad(M)))|d.R(\infty)\leq\lambda\bigl{|}S\bigl{(}2(\Delta+\text{\rm rad}(M))\bigr{)}\bigr{|}_{d}.

If the right side is strictly less than 11, then θd(λ,,α)=0\theta_{\text{\rm d}}(\lambda,\infty,\alpha)=0 by Proposition 3.8(b) adapted to the current context.

Remark 4.3 (Condition Cγ,σ{}_{\gamma,\sigma}^{\hskip 1.0pt\prime}).

Let Mt=sup{ζ(s)d:s[0,t]}M_{t}=\sup\{\|\zeta(s)\|_{d}:s\in[0,t]\}, the maximum displacement of ζ\zeta up to time tt, and let MtM^{\prime}_{t} be given similarly in terms of ζ\zeta^{\prime}. By Minkowski’s inequality,

𝔼|Σt|d𝔼([Mt+Mt+1]d)(2Mt+1)d,{\mathbb{E}}|\Sigma_{t}^{\prime}|_{d}\leq{\mathbb{E}}\bigl{(}[M_{t}+M^{\prime}_{t}+1]^{d}\bigr{)}\leq\bigl{(}2\|M_{t}\|+1\bigr{)}^{d},

Here, \|\cdot\| denotes the LdL^{d} norm. Therefore, Cγ,σ{}_{\gamma,\sigma}^{\hskip 1.0pt\prime} holds for some γ\gamma, σ\sigma if Mtγeσt\|M_{t}\|\leq\gamma^{\prime}e^{\sigma^{\prime}t} for suitable γ\gamma^{\prime}, σ\sigma^{\prime}.

4.2. The Brownian diffusion model

Suppose that ρ(0,]\rho\in(0,\infty], μ=1S\mu=1_{S}, and ζ\zeta is a standard Brownian motion (one may allow it to have constant non-zero drift, but for simplicity we set the drift to 0). Since (ζζ)/2(\zeta-\zeta^{\prime})/\sqrt{2} is a standard Brownian motion, it is easily seen that 𝔼|Σs|d=𝔼|W2s|d{\mathbb{E}}|\Sigma_{s}^{\prime}|_{d}={\mathbb{E}}|W_{2s}|_{d} where WW is the usual radius-11 Wiener sausage. Therefore,

R()=λ0𝔼|W2s|dαeαs𝑑s=λ0𝔼|Ws|d(α/2)eαs/2𝑑s.R(\infty)=\lambda\int_{0}^{\infty}{\mathbb{E}}|W_{2s}|_{d}\,\alpha e^{-\alpha s}\,ds=\lambda\int_{0}^{\infty}{\mathbb{E}}|W_{s}|_{d}\,(\alpha/2)e^{-\alpha s/2}\,ds.

Hence, α¯(λ)=2α¯dd(λ)\underline{\alpha}(\lambda)=2\underline{\alpha}_{\text{\rm dd}}(\lambda) where α¯dd(λ)\underline{\alpha}_{\text{\rm dd}}(\lambda) is the corresponding quantity α¯\underline{\alpha} of Example 3.13 for the delayed diffusion model.

4.3. Survival

We close with some remarks on the missing ‘survival’ parts of Theorems 1.2 and 4.1. An iterative construction similar to that of Section 3.7 may be explored for the diffusion model. However, Proposition 3.3 is not easily extended or adapted when the particles are permanently removed following infection.

The situation is different when either there is no removal (that is, α=0\alpha=0, see [3]), or ‘recuperation’ occurs in that particles become susceptible again post-infection. A model of the latter type, but involving random walks rather than Brownian motions, has been studied by Kesten and Sidoravicius in their lengthy and complex work [18]. Each of these variants has structure not shared with our diffusion model, including the property that the set of infectives increases when the set of initially infected particles increases. Heavy use is made of this property in [18]. Unlike the delayed diffusion model (see the end of Section 3.1 and Proposition 3.4), neither the diffusion model nor its random-walk version has this property, in contradiction of the claim of Remark 4 of [18].

Acknowledgements

ZL’s research was supported by National Science Foundation grant 1608896 and Simons Foundation grant 638143. GRG thanks Alexander Holroyd and James Norris for useful conversations. The authors are very grateful to three referees for their detailed and valuable reports, which have led to significant corrections. The work reported here was influenced in part by the Covid-19 pandemic of 2020.

References

  • [1] O. S. M. Alves, F. P. Machado, and S. Yu. Popov, Phase transition for the frog model, Electron. J. Probab. 7 (2002), paper no. 16, 21 pp. MR 1943889
  • [2] by same author, The shape theorem for the frog model, Ann. Appl. Probab. 12 (2002), 533–546. MR 1910638
  • [3] E. Beckman, E. Dinan, R. Durrett, R. Huo, and M. Junge, Asymptotic behavior of the Brownian frog model, Electron. J. Probab. 23 (2018), paper no. 104, 19 pp. MR 3870447
  • [4] I. Benjamini, L. R. Fontes, J. Hermon, and F. P. Machado, On an epidemic model on finite graphs, Ann. Appl. Probab. 30 (2020), 208–258. MR 4068310
  • [5] J. van den Berg, R. Meester, and D. G. White, Dynamic Boolean models, Stoch. Proc. Appl. 69 (1997), 247–257. MR 1472953
  • [6] J.-D. Deuschel and A. Pisztora, Surface order large deviations for high-density percolation, Probab. Theory Related Fields 104 (1996), 467–482. MR 1384041
  • [7] L. R. Fontes, F. P. Machado, and A. Sarkar, The critical probability for the frog model is not a monotonic function of the graph, J. Appl. Probab. 41 (2004), 292–298. MR 2036292
  • [8] R. K. Getoor, Some asymptotic formulas involving capacity, Z. Wahrsch’theorie und verwandte Gebiete 4 (1965), 248–252. MR 190988
  • [9] P. Gracar and A. Stauffer, Percolation of Lipschitz surface and tight bounds on the spread of information among mobile agents, Approximation, Randomization, and Combinatorial Optimization. Algorithms and Techniques, LIPIcs, Leibniz Int. Proc. Inform., vol. 116, Schloss Dagstuhl. Leibniz-Zent. Inform., Wadern, 2018, paper 39. MR 3857277
  • [10] G. R. Grimmett, Percolation, 2nd ed., Grundlehren der Mathematischen Wissenschaften, vol. 321, Springer-Verlag, Berlin, 1999. MR 1707339
  • [11] by same author, Probability on Graphs, 2nd ed., Cambridge University Press, Cambridge, 2018. MR 3751350
  • [12] G. R. Grimmett and D. R. Stirzaker, Probability and Random Processes, 4th ed., Oxford University Press, Oxford, 2020. MR 4229142
  • [13] Y. Hamana and H. Matsumoto, A formula for the expected volume of the Wiener sausage with constant drift, Forum Math. 29 (2017), 369–381. MR 3619119
  • [14] P. H. Haynes, V. H. Hoang, J. R. Norris, and K. C. Zygalakis, Homogenization for advection–diffusion in a perforated domain, Probability and Mathematical Genetics, vol. 37, Cambridge University Press, 2010, pp. 397–415. MR 2744249
  • [15] C. Hoffman, T. Johnson, and M. Junge, Cover time for the frog model on trees, Forum Math. Sigma 7 (2019), Paper No. e41, 49 pp. MR 4031108
  • [16] T. Johnson and M. Junge, Stochastic orders and the frog model, Ann. Inst. Henri Poincaré Probab. Stat. 54 (2018), 1013–1030. MR 3795075
  • [17] H. Kesten and V. Sidoravicius, The spread of a rumor or infection in a moving population, Ann. Probab. 33 (2005), 2402–2462. MR 2184100
  • [18] by same author, A phase transition in a model for the spread of an infection, Illinois J. Math. 50 (2006), 547–634. MR 2247840
  • [19] by same author, A shape theorem for the spread of an infection, Ann. of Math. (2) 167 (2008), 701–766. MR 2415386
  • [20] J. F C. Kingman, Poisson Processes, Oxford Studies in Probability, vol. 3, Oxford University Press, Oxford, 1993. MR 1207584
  • [21] K. Kuulasmaa, The spatial general epidemic and locally dependent random graphs, J. Appl. Probab. 19 (1982), 745–758. MR 675138
  • [22] T. M. Liggett, R. H. Schonmann, and A. M. Stacey, Domination by product measures, Ann. Probab. 25 (1997), 71–95. MR 1428500
  • [23] C. Marinelli and M. Röckner, On the maximal inequalities of Burkholder, Davis and Gundy, Exposit. Mathem. 34 (2016), 1–26. MR 3463679
  • [24] P. Mattila, Geometry of Sets and Measures in Euclidean Spaces, Cambridge University Press, Cambridge, 1995. MR 1333890
  • [25] R. Meester and R. Roy, Continuum Percolation, Cambridge Tracts in Mathematics, vol. 119, Cambridge University Press, Cambridge, 1996. MR 1409145
  • [26] P. Mörters and Y. Peres, Brownian Motion, Cambridge Series in Statistical and Probabilistic Mathematics, vol. 30, Cambridge University Press, Cambridge, Cambridge, 2010. MR 2604525
  • [27] Y. Peres, A. Sinclair, P. Sousi, and A. Stauffer, Mobile geometric graphs: Detection, coverage and percolation, Probab. Theory Related Fields 156 (2013), 273–305. MR 3055260
  • [28] S. Yu. Popov, Frogs and some other interacting random walks models, Discrete Random Walks, Assoc. Discrete Math. Theor. Comput. Sci., Nancy, 2003, pp. 277–288. MR 1819703
  • [29] A. F. Ramírez and V. Sidoravicius, Asymptotic behavior of a stochastic growth process associated with a system of interacting branching random walks, C. R. Math. Acad. Sci. Paris 335 (2002), 821–826. MR 1947707
  • [30] L. T. Rolla, Activated random walks on d{\mathbb{Z}}^{d}, Probab. Surveys 17 (2020), 478–544. MR 4152668
  • [31] L. T. Rolla, V. Sidoravicius, and O. Zindy, Universality and sharpness in activated random walks, Ann. Henri Poincaré, Theor. & Math. Phys. 20 (2019), 1823–1835. MR 3956161
  • [32] F. Spitzer, Electrostatic capacity, heat flow and Brownian Motion, Z. Wahrsch’theorie und verwandte Geb. 3 (1964), 110–121. MR 172343
  • [33] A. Stauffer and L. Taggi, Critical density of activated random walks on transitive graphs, Ann. Probab. 46 (2018), 2190–2220. MR 3813989