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

Global solutions with infinitely many blowups in a mean-field neural network

Lorenzo Sadunlabel=e2]sadun@math.utexas.edu [    Thibaud Taillefumierlabel=e1]ttaillef@austin.utexas.edu [ Department of Mathematics, University of Texas, Austin, Department of Neuroscience, University of Texas, Austin,
Abstract

We recently introduced idealized mean-field models for networks of integrate-and-fire neurons with impulse-like interactions—the so-called delayed Poissonian mean-field models. Such models are prone to blowups: for a strong enough interaction coupling, the mean-field rate of interaction diverges in finite time with a finite fraction of neurons spiking simultaneously. Due to the reset mechanism of integrate-and-fire neurons, these blowups can happen repeatedly, at least in principle. A benefit of considering Poissonian mean-field models is that one can resolve blowups analytically by mapping the original singular dynamics onto uniformly regular dynamics via a time change. Resolving a blowup then amounts to solving the fixed-point problem that implicitly defines the time change, which can be done consistently for a single blowup and for nonzero delays. Here we extend this time-change analysis in two ways: First, we exhibit the existence and uniqueness of explosive solutions with a countable infinity of blowups in the large interaction regime. Second, we show that these delayed solutions specify “physical” explosive solutions in the limit of vanishing delays, which in turn can be explicitly constructed. The first result relies on the fact that blowups are self-sustaining but nonoverlapping in the time-changed picture. The second result follows from the continuity of blowups in the time-changed picture and incidentally implies the existence of periodic solutions. These results are useful to study the emergence of synchrony in neural network models.

60G99,
60K15, 35Q92, 35D30, 35K67, 45H99,
mean-field neural network models; blowups in parabolic partial differential equation; regularization by time change; singular interactions; delayed integral equations; inhomogeneous renewal processes,
second keyword,
keywords:
[class=MSC2020]
keywords:
\startlocaldefs\endlocaldefs

and

1 Introduction

1.1 Background

In this work, we consider idealized neural-network models called delayed Poissonian mean-field (dPMF) models introduced in [22]. These dPMF models are variations of classical mean-field models [5, 4, 7], whose dynamics are also prone to blowups. From a modeling perspective, blowups correspond to the occurrence of synchronous, macroscopic spiking events in a neural network. Understanding the emergence of these synchronous events is of interest for studying the maintenance of precise temporal information in neural networks [19, 15, 3]. However, blowups resist direct analytical treatment in classical mean-field models [10, 9, 13, 18, 17]. This observation is the primary motivation justifying the introduction of dPMF dynamics, whose blowups prove analytically tractable.

In [22], we conjectured dPMF dynamics as the mean-field limit of finite-size particle systems with singular interactions (see Fig. 1). In a finite-size system of NN particles, each particle ii, 1iN1\leq i\leq N, represents a neuron with time-dependent state variable XN,i,tX_{N,i,t}. These state variables evolve jointly according to a N\mathbbm{R}^{N}-valued continuous-time process t{XN,i,t}1iNt\mapsto\{X_{N,i,t}\}_{1\leq i\leq N}. Specifically, the network dynamics is parametrized by the drift value ν>0\nu>0, the refractory period ϵ>0\epsilon>0, and the interaction parameter λ>0\lambda>0 as follows: (i)(i) Whenever a process XN,i,tX_{N,i,t} hits the spiking boundary at zero, it instantaneously enters its inactive refractory state. (ii)(ii) At the same time, all the other active processes XN,j,tX_{N,j,t} (which are not in the inactive refractory state) are respectively updated by amounts wN,ij,t-w_{N,ij,t}, where wN,ij,tw_{N,ij,t} is independently drawn from a normal law with mean and variance equal to λ/N\lambda/N. (iii)(iii) After an inactive (refractory) period of duration ϵ>0\epsilon>0, the process XN,i,tX_{N,i,t} restarts its autonomous stochastic dynamics from the reset state Λ>0\Lambda>0. (iv)(iv) In between spiking/interaction times, the autonomous dynamics of active processes follow independent drifted Wiener processes with negative drift ν-\nu. Correspondingly, an initial condition for the network is specified by the starting values of the active processes, i.e., XN,i,0>0X_{N,i,0}>0 if ii is active, and the last inactivation time of the inactive processes, i.e., ϵρi,00-\epsilon\leq\rho_{i,0}\leq 0, if ii is inactive. Following the above definition, the finite-size versions of dPMF dynamics exhibit two key features (see Fig. 1): (1)(1) recurrent interactions implement self-excitation whose strength is quantified by the interaction parameter λ\lambda; (2)(2) owing to the post-spiking refractory period ϵ>0\epsilon>0, individual neuronal processes follow delayed dynamics. Parenthetically, the inclusion of a negative drift ν\nu is necessary to ensure that neurons will spike in finite time, independent of the initial conditions and of the coupling strength.

By analogy with [7, 8], dPMF dynamics are deduced from finite-size ones by conjecturing propagation of chaos in the infinite-size limit NN\to\infty, which is supported by numerical simulations (see Fig. 2). The propagation of chaos states that for exchangeable initial conditions, the processes XN,i,tX_{N,i,t}, 1iN1\leq i\leq N, become i.i.d. in the limit of infinite-size networks NN\to\infty, so that each individual process follows a mean-field dynamics [21]. For exchangeable initial conditions, a representative dPMF process Xt=limNXN,i,tX_{t}=\lim_{N\to\infty}X_{N,i,t} is governed by the deterministic cumulative drift Φ\Phi that amalgamates the contribution of the autonomous drift ν-\nu and the mean-field contribution of neuronal interactions via λ\lambda. Specifically, we have Φ(t)=νt+λ𝔼[Mt]\Phi(t)=\nu t+\lambda\mathbb{E}\left[M_{t}\right], where the process MtM_{t} counts the successive first-passage times of the representative process XtX_{t} to the zero spiking threshold. Such a process is defined as

Mt=n>0𝟙[0,t](ρn),withρn+1=inf{t>rn=ρn1+ϵ|Xt0},\displaystyle M_{t}=\sum_{n>0}\mathbbm{1}_{[0,t]}(\rho_{n})\,,\quad\mathrm{with}\quad\rho_{n+1}=\inf\left\{t>r_{n}=\rho_{n-1}+\epsilon\,\big{|}\,X_{t}\leq 0\right\}\,, (1)

with initial conditions bearing on X0X_{0} or ρ0\rho_{0} depending on the representative process being initially active or inactive. We will later specify such initial conditions in detail. Note that in the above definition, the first-passage times ρn\rho_{n}, n1n\geq 1, denote successive inactivation times, whereas rn=ρn+ϵr_{n}=\rho_{n}+\epsilon, n>0n>0, denotes the corresponding sequence of reset times. For mediating all interactions, the function t𝔼[Mt]t\mapsto\mathbb{E}\left[M_{t}\right] plays a central role in our analysis. We denote this function by FF and observe that FF is defined as an increasing càdlàg function. Remember that a function is said to be càdlàg if it is right-continuous with left limits. Then, the hallmark of dPMF dynamics is that the drift Φ\Phi impacts the representative neuron stochastically via a process Zt=Φ(t)+WΦ(t)Z_{t}=\Phi(t)+W_{\Phi(t)}, where WW is a canonical, driving Wiener process. Accordingly, dPMF dynamics are referred to as Poissonian because the process ZtZ_{t} represents the diffusive limit of a Poisson counting process with 𝔼[Zt]=𝕍[Zt]\mathbb{E}\left[Z_{t}\right]=\mathbb{V}\left[Z_{t}\right] for all t0t\geq 0. With these conventions, the stochastic dPMF dynamics of a representative process is given by

Xt=X00t𝟙{Xs>0}dZs+ΛMtϵ.\displaystyle X_{t}=X_{0}-\int_{0}^{t}\mathbbm{1}_{\{X_{s^{-}}>0\}}\mathrm{d}Z_{s}+\Lambda M_{t-\epsilon}\,. (2)

The above equation fully defines dPMF dynamics: The integral term indicates that when the neuron is active, for Xt>0X_{t}>0, its state evolves according to the Poissonian diffusion process ZtZ_{t}. The delayed term indicates that after hitting the spiking boundary at zero at time τ\tau, the process remains inactive until it resets at Λ\Lambda after a duration ϵ\epsilon: Xt=0X_{t}=0 for τt<τ+ϵ\tau\leq t<\tau+\epsilon.

Because of their self-exciting nature, dPMF dynamics are prone to blowups/synchronous events for large enough interaction parameter and/or for initial conditions that are concentrated near the zero spiking boundary. Blowups occur at those times when ff, the instantaneous spiking rate of a representative neuron, diverges. Formally, the firing rate ff is defined in the distribution sense as the Radon-Nikodym derivative of FF with respect to the Lebesgue measure: f=dF/dtf=\mathrm{d}F/\mathrm{d}t. Then, assuming ff to be finite in the left vicinity of T1T_{1}, T1T_{1} is a blowup time if limtT1f(t)=\lim_{t\to T_{1}^{-}}f(t)=\infty. In turn, synchronous events occur at those times T1T_{1} for which a finite fraction of processes spikes simultaneously or equivalently, for which a representative neuron spikes with nonzero probability: π1=[XT1=0]>0\pi_{1}=\mathbb{P}\left[X_{T_{1}}=0\right]>0. Formally, this corresponds to FF admitting a jump discontinuity at T1T_{1} so that F(T1)F(T1)=λπ1F(T_{1})-F(T_{1}^{-})=\lambda\pi_{1}.

Refer to caption
Figure 1: Delayed Poissonian network dynamics. a. Schematic representation of the delayed Poissonian network dynamics in a finite-size particel system with N=3N=3 interacting processes. In between spiking events, neuronal dynamics are that of independent Wiener processes with negative drift ν<0-\nu<0. Neuron 22 spikes and inactivates at time t1t_{1}, which leads to updating neuron 11 and 33 via weights w21,t1w_{21,t_{1}} and w23,t1w_{23,t_{1}} i.i.d. with normal law 𝒩(λ/N,λ/N)\mathcal{N}(\lambda/N,\lambda/N). This causes neuron 11 to spike and inactivate, leading to another instantaneous update of neuron 33 via a random weight w13,t1w_{13,t_{1}}. Then, neuron 33 spikes and inactivates at a later time t2t_{2}, but this has no impact on neurons 11 and 22 as they remain inactive during a refractory period of duration ϵ>t2t1\epsilon>t_{2}-t_{1}. At the end of their refractory period, neurons reset at Λ>0\Lambda>0. Finally, the spontaneous spiking of neuron 11 in t3t_{3} cause both neurons 22 and 33 to spike at the same time. b. Schematic representation of the conjectured delayed Poissonian mean-field (dPMF) dynamics in the limit of infinite-size networks NN\to\infty. When NN\to\infty, it is conjectured that the neuronal dynamics become independent and that interactions are mediated by the mean-field deterministic cumulative drift Φ(t)=νt+λF(t)\Phi(t)=\nu t+\lambda F(t), where the increasing function FF denotes the cumulative spiking or inactivation rate. Specifically, it is conjectured that neuronal dynamics follow independent time-changed Wiener processes with Poissonian attributes parametrized by Φ\Phi. For time t<T1t<T_{1}, the (Radon-Nikodym) derivative of Φ\Phi remains finite and the spiking of representative processes are determined via first-passage time problems with regular boundary. Accordingly, two representative processes have zero probability to spike synchronously. However, for large enough interaction parameter λ\lambda, it is possible that the (Radon-Nikodym) derivative of Φ\Phi locally diverges and that a finite fraction π1\pi_{1}, 0<π1<10<\pi_{1}<1, of the processes synchronously spike as depicted in T1T_{1}. We refer to such a possibility as a full-blowup event. This corresponds to a jump discontinuity of size λπ1\lambda\pi_{1} in Φ\Phi and to singular stochastic updates of the representative processes with weight w1,T1w_{1,T_{1}} and w2,T1w_{2,T_{1}} i.i.d. normal law 𝒩(λπ1,λπ1)\mathcal{N}(\lambda\pi_{1},\lambda\pi_{1}).

The main interest of considering dPMF dynamics is that blowups and synchronous events are analytically tractable under some reasonable restrictions about the initial conditions. Specifically, denoting by (I)\mathcal{M}(I) the set of positive measure on an interval II\subset\mathbbm{R}, we assume as in [22] that:

Assumption 1.

The initial conditions for dPMF dynamics is specified by

[X0dx|X0>0]=p0(dx)and[ρ0dt|X0=0]=f0(dt),\displaystyle\mathbb{P}\left[X_{0}\in\mathrm{d}x|X_{0}>0\right]=p_{0}(\mathrm{d}x)\quad\mathrm{and}\quad\mathbb{P}\left[\rho_{0}\in\mathrm{d}t|X_{0}=0\right]=f_{0}(\mathrm{d}t)\,,

where (p0,f0)(p_{0},f_{0}) in ((0,))×([ϵ,0))\mathcal{M}((0,\infty))\times\mathcal{M}([-\epsilon,0)) is normalized, i.e., p01+f01=1\|p_{0}\|_{1}+\|f_{0}\|_{1}=1 and such that p0p_{0} is locally differentiable in zero with limx0+p0(x)=0\lim_{x\to 0^{+}}p_{0}(x)=0 and limx0+xp0(x)/2<1/λ\lim_{x\to 0^{+}}\partial_{x}p_{0}(x)/2<1/\lambda.

The above initial conditions guarantee that there exists a dPMF dynamics locally solving (2) and that this dynamics is initially smooth in the sense that FF is an infinitely differentiable function in the right vicinity of zero. Such a smooth solution can be maximally continued on a possibly infinite interval [0,T1)[0,T_{1}), where T1T_{1} marks the occurrence of the first blowup. In [22], we show that for all 0<t<T10<t<T_{1}, the density xp(x,t)=[Xtdx|Xt>0]/dxx\mapsto p(x,t)=\mathbb{P}\left[X_{t}\in\mathrm{d}x\,|X_{t}>0\right]/\mathrm{d}x is smooth in the right vicinity of zero with limx0+p(x)=0\lim_{x\to 0^{+}}p(x)=0 and limx0+xp(t,x)/2<1/λ\lim_{x\to 0^{+}}\partial_{x}p(t,x)/2<1/\lambda. As the later limits are always well defined and finite, we will simply refer to their values as p(0)p(0) and xp(t,0)/2\partial_{x}p(t,0)/2 for conciseness. With this in mind, we show in [22] that for all 0<t<T10<t<T_{1}, the instantaneous spiking rate is given by

f(t)=νxp(0,t)2λxp(0,t).\displaystyle f(t)=\frac{\nu\partial_{x}p(0,t)}{2-\lambda\partial_{x}p(0,t)}\,.

This leads to introducing the following criterion for first-blowup times:

Definition 1.1.

Under Assumption 1, the first blowup time is defined as

T1=sup{t>0|xp(t,0)<2/λ}.\displaystyle T_{1}=\sup\{t>0\,|\,\partial_{x}p(t,0)<2/\lambda\}\,.

The above criterion only bears on the local divergence of the firing rate ff and is silent about the possible occurrence of a synchronous event. To further characterize blowups, we introduce in [22] the so-called full-blowup criterion:

Assumption 2.

At the blowup time T1T_{1}, the density function xp(x,t)x\mapsto p(x,t) satisfies

limtT1x3p(t,0)>8/λ.\displaystyle\lim_{t\to T_{1}^{-}}\partial^{3}_{x}p(t,0)>8/\lambda\,.

The full-blowup criterion, which generically holds with respect to the choice of initial conditions, allowed us to characterize blowups in dPMF dynamics as follows:

Proposition 1.1.

Under Assumptions 1 and 2, the instantaneous rate blows up as f(t)1/T1tf(t)\sim 1/\sqrt{T_{1}-t} when tT1t\to T_{1}^{-} and triggers a synchronous event in T1T_{1} of size

π1=inf{p>0|p>[τXT1<λp]}>0,\displaystyle\pi_{1}=\inf\left\{p>0\,|\,p>\mathbb{P}\left[\tau_{X_{T_{1}}}<\lambda p\right]\right\}>0\,, (3)

where τXT1\tau_{X_{T_{1}}} denotes the first-passage time in zero of a Wiener process with unit negative drift and started with random initial condition XT1X_{T_{1}^{-}}.

Thus, for generic initial conditions, blowups correspond to left Hölder singularity of exponent 1/21/2 in the cumulative function FF, followed by a jump discontinuity whose size λπ1\lambda\pi_{1} can be determined as the solution of the self-consistent problem (3). In [22], we show that the latter problem directly follows from the requirement of conservation of probability during blowups.

Refer to caption
Figure 2: Simulations of finite-size interacting particle systems. We simulate a finite network of neurons which follows the interacting dynamics depicted in Fig. 1 and for varying size N=10,100,10000N=10,100,10000. Parameters: Λ=1\Lambda=1, ν=1\nu=1, λ=20\lambda=20 and ϵ=0\epsilon=0. Initial conditions: all neurons start at reset value. In each panel: (1)(1) the top graphics represents a raster plot of the neuronal activity, where the spiking of neuron ii at time tt is marked by a point at coordinate (i,t)(i,t). Synchronous event corresponds to vertically aligned points which are apparent for all conditions. (2)(2) the bottom left graphics represents the histogram of time interval between synchronous events. (3)(3) the bottom right graphics represents the spectral density of the overall network activity. Making the approximation that the full blowup has size π1=1\pi_{1}=1 and neglecting reset besides at S1S_{1} yields the mean-field predictions: 1/T116.181/T_{1}\simeq 16.18 with T1=(S1λH(S1,Λ))/ν0.07143T_{1}=(S_{1}-\lambda H(S_{1},\Lambda))/\nu\simeq 0.07143, where S1S_{1} is the smallest solution of h(S1,Λ)=1/λh(S_{1},\Lambda)=1/\lambda. The level of synchrony increases with the size of the system NN, supporting that the particle system admits a (periodic) mean-field dynamics in the limit NN\to\infty.

1.2 Motivation

Unfortunately, the blowup analysis of [22] is only local in the sense that it allows one to resolve a single blowup for generic initial conditions. However, this local analysis provides one with natural blowup exit conditions. In principle, these blowup exit conditions can also serve as initial conditions to analyze the next blowup, if any. Numerical simulations suggest that for large enough interaction parameters λ>Λ\lambda>\Lambda, explosive solutions exhibit repeated blowup episodes at regular time intervals (see Fig. 2). Following on this observation, the first goal of this work is to extend the analysis of [22] to show the existence of global dPMF solutions that are defined over the entire half-line +\mathbbm{R}^{+} and that exhibit a countable infinity of blowups. This program involves proving that for large enough interaction parameters, iteratively applying the blowup analysis of [22] produces synchronous events with sizes that remain bounded away from zero, at time intervals that also remain bounded away from zero. Moreover, the blowup analysis of [22] is only concerned with dPMF dynamics for positive refractory period ϵ>0\epsilon>0. With positive refractory period ϵ>0\epsilon>0, explosive dPMF dynamics are always well posed but at the analytical cost of being determined as delayed dynamics. For small enough ϵ>0\epsilon>0, we do not expect the refractory period to impact dPMF dynamics, except for ensuring their well-posedness in the presence of blowups. This suggests defining explosive Poissonian mean-field (PMF) dynamics in the absence of a refractory period as the limit dPMF dynamics obtained when ϵ0+\epsilon\to 0^{+}. Accordingly, the second goal of this work is to prove the existence and uniqueness of such limit dPMF dynamics. This will only be possible for large enough interaction parameters, when global dPMF dynamics sustain isolated blowups for small enough ϵ>0\epsilon>0. PMF dynamics with zero refractory period ϵ=0\epsilon=0 are of interest for analysis as their blowups can be resolved just as for ϵ>0\epsilon>0, whereas their inter-blowup dynamics are nondelayed.

1.3 Time-change approach

Refer to caption
Figure 3: Time-changed picture. a. In the absence of blowups, the cumulative drift Φ\Phi specifies an increasing function which can be interpreted as an invertible time change. Denoting the inverse time change by Ψ=Φ1\Psi=\Phi^{-1} and given a representative process XtX_{t} of the dPMF dynamics, the process Yσ=XΨ(σ)Y_{\sigma}=X_{\Psi(\sigma)} follows a noninteracting drifted Wiener dynamics with negative unit drift and nonhomogeneous backward-delay function η:σσΦ(Ψ(σ)ϵ)\eta:\sigma\mapsto\sigma-\Phi(\Psi(\sigma)-\epsilon). Thus, in the time-change picture, all the interactions at play in dPMF dynamics are encoded by the time-dependence of delays. b. In the absence of blowups, full blowups are marked by a flat section [U1,S1)[U_{1},S_{1}) of the inverse time change Ψ\Psi, where U1U_{1} is the blowup trigger time and U1U_{1} is the blowup exit time. This corresponds to the backward-delay function η\eta having a jump discontinuity of size S1U1S_{1}-U_{1} in U1U_{1}. This also means that the forward-delay function τ\tau specifying reset times is constant over the interval [U1,S1)[U_{1},S_{1}), so that all processes involved in the blowup episode can reset at the same time R1R_{1}.

In [22], we analytically characterized blowup generation in dPMF dynamics by mapping the original time-homogenous, nonlinear dynamics onto a time-inhomogeneous, linear dynamics (see Fig. 3). Such a mapping is operated by considering the cumulative drift Φ\Phi as an implicitly defined time-change function

σ=Φ(t)=νt+λF(t),withF(t)=0tf(s)ds,\displaystyle\sigma=\Phi(t)=\nu t+\lambda F(t)\,,\quad\mathrm{with}\quad F(t)=\int_{0}^{t}f(s)\,\mathrm{d}s\,, (4)

which is only assumed to be a càdlàg increasing function. Due to the Poisson-like attributes of the neuronal drives, the time change Φ\Phi parametrizes the dPMF dynamics of a representative process as Xt=YΦ(t)X_{t}=Y_{\Phi(t)}, where the time-changed dynamics YσY_{\sigma} obeys a linear, noninteracting dynamics. The latter dynamics is that of a Wiener process absorbed at zero, with constant negative unit drift and with reset at Λ\Lambda, but with time-inhomogeneous refractory period specified via a Φ\Phi-dependent, backward-delay function ση[Φ](σ)\sigma\mapsto\eta[\Phi](\sigma) (see Fig. 3). Independent of the presence of blowups, the functional dependence of η\eta on the time change Φ\Phi is given by

η(σ)=σΦ(Ψ(σ)ϵ),\displaystyle\eta(\sigma)=\sigma-\Phi(\Psi(\sigma)-\epsilon)\,, (5)

where Ψ=Φ1\Psi=\Phi^{-1} refers to the inverse of the strictly increasing time change Φ\Phi. Given a backward-delay function η\eta, the transition kernel of the process YσY_{\sigma} denoted by (σ,x)q(σ,x)=d[Yσdx|Yσ>0]/dx(\sigma,x)\mapsto q(\sigma,x)=\mathrm{d}\mathbb{P}\left[Y_{\sigma}\in\mathrm{d}x\,|\,Y_{\sigma}>0\right]/\mathrm{d}x satisfies the time-changed PDE problem

σq\displaystyle\partial_{\sigma}q =\displaystyle= xq+12x2q+ddσ[G(ση(σ))]δΛ,\displaystyle\partial_{x}q+\frac{1}{2}\partial^{2}_{x}q+\frac{\mathrm{d}}{\mathrm{d}\sigma}[G(\sigma-\eta(\sigma))]\delta_{\Lambda}\,, (6)

with absorbing and conservation conditions respectively given by

q(σ,0)=0andσG(σ)=xq(σ,0)/2.\displaystyle q(\sigma,0)=0\quad\mathrm{and}\quad\partial_{\sigma}G(\sigma)=\partial_{x}q(\sigma,0)/2\,. (7)

In equations (6) and (7), GG denotes the η\eta-dependent cumulative flux of YσY_{\sigma} through the zero threshold. By definition of the time change Φ\Phi, which is such that Xt=YΦ(t)X_{t}=Y_{\Phi(t)}, GG is related to the cumulative flux FF via F=GΦF=G\circ\Phi. A key result of [22] is that as long as the delay function η\eta remains bounded, the PDE problem defined by (6) and (7) admits a unique solution parametrized by an unconditionally smooth cumulative function G=G[η]G=G[\eta]. This result holds even in the presence of blowups for the time-changed versions of the initial conditions given in Assumption 1. These time-changed initial conditions are specified as follows:

Definition 1.2.

Given normalized initial conditions (p0,f0)(p_{0},f_{0}) in (+)×([ϵ,0))\mathcal{M}(\mathbbm{R}^{+})\times\mathcal{M}([-\epsilon,0)), the initial conditions for the time-changed problem are defined by (q0,g0)(q_{0},g_{0}) in (+)×([ξ0,0))\mathcal{M}(\mathbbm{R}^{+})\times\mathcal{M}([\xi_{0},0)) such that

q0=p0andg0=dG0dσwithG0=(idνΨ0)/λ,\displaystyle q_{0}=p_{0}\quad\mathrm{and}\quad g_{0}=\frac{\mathrm{d}G_{0}}{\mathrm{d}\sigma}\quad\mathrm{with}\quad G_{0}=(\mathrm{id}-\nu\Psi_{0})/\lambda\,,

where the function Ψ0\Psi_{0} and the number ξ0\xi_{0} are given by:

Ψ0(σ)\displaystyle\Psi_{0}(\sigma) =\displaystyle= inf{t0|νt+λ0tf0(s)ds>σ},\displaystyle\inf\left\{t\geq 0\,\bigg{|}\,\nu t+\lambda\int_{0}^{t}f_{0}(s)\,\mathrm{d}s>\sigma\right\}\,,
ξ0\displaystyle\xi_{0} =\displaystyle= νϵλϵ0f0(t)dt<0.\displaystyle-\nu\epsilon-\lambda\int_{-\epsilon}^{0}f_{0}(t)\,\mathrm{d}t<0\,.

In light of (12), the cumulative function G=G[Φ]G=G[\Phi] actually depends on Φ\Phi via η=η[Φ]\eta=\eta[\Phi]. This realization allows one to interpret the implicit definition of Φ\Phi given in (4) as a self-consistent equation for admissible time changes:

Φ(t)=νt+λG[Φ](Φ(t)).\displaystyle\Phi(t)=\nu t+\lambda G[\Phi](\Phi(t))\,. (8)

In [22], we show that such an interpretation specifies Φ\Phi as the solution of a certain fixed-point problem. This fixed-point problem turns out to be most conveniently formulated in term of the inverse time change Ψ=Φ1\Psi=\Phi^{-1}, assumed to be a continuous, nondecreasing function. Given an inverse time change Ψ\Psi, the time change Φ\Phi can be recovered as the right-continuous inverse of Ψ\Psi. In the time-changed picture, blowups happen if the inverse time change Ψ\Psi becomes locally flat and a synchronous event happens if Ψ\Psi remains flat for a finite amount of time. Informally, flat sections of Ψ\Psi unfold blowups by freezing time in the original coordinate tt, while allowing time to pass in the time-changed coordinate σ\sigma. This unfolding of blowups in the time-changed picture is the key to analytically resolve blowups in dPMF dynamics. Concretely, this amounts to showing the existence and uniqueness of solutions to the following fixed-point problem:

Definition 1.3.

Given time-changed initial conditions (q0,g0)(q_{0},g_{0}) in (+)×([ξ0,0))\mathcal{M}(\mathbbm{R}^{+})\times\mathcal{M}([\xi_{0},0)), an admissible inverse time change Ψ\Psi satisfies the fixed-point problem

σ0,Ψ(σ)={(σλ0σg0(ξ)dξ)/νifξ0σ<0,sup0ξσ(ξλG[η](ξ))/νifσ0.\displaystyle\forall\;\sigma\geq 0\,,\quad\Psi(\sigma)=\left\{\begin{array}[]{ccc}\left(\sigma-\lambda\int_{0}^{\sigma}g_{0}(\xi)\mathrm{d}\xi\right)/\nu&\quad\mathrm{if}&-\xi_{0}\leq\sigma<0\,,\vspace{5pt}\\ \sup_{0\leq\xi\leq\sigma}\big{(}\xi-\lambda G[\eta](\xi)\big{)}/\nu&\quad\mathrm{if}&\sigma\geq 0\,.\end{array}\right. (11)

where G[η]G[\eta] is the smooth cumulative flux uniquely specified by the time-inhomogeneous backward-delay function η:++\eta:\mathbbm{R}^{+}\to\mathbbm{R}^{+} (see Definition 1.4). The fixed-point nature of the problem follows from the definition of the backward-delay function η\eta as the Ψ\Psi-dependent time-wrapped version of the constant delay ϵ\epsilon:

η(σ)=σΦ(Ψ(σ)ϵ)withΦ(t)=inf{σξ0|Ψ(σ)>t},\displaystyle\eta(\sigma)=\sigma-\Phi(\Psi(\sigma)-\epsilon)\quad\mathrm{with}\quad\Phi(t)=\inf\left\{\sigma\geq\xi_{0}\,\big{|}\,\Psi(\sigma)>t\right\}\,, (12)

for which we consistently have η(0)=Φ(ϵ)=ξ0\eta(0)=-\Phi(-\epsilon)=\xi_{0}.

Refer to caption
Figure 4: Blowup resolution in the time-change picture. Schematic explosive dPMF dynamics considering a purely spatial initial condition with unit mass concentrated at Λ\Lambda: q0=δΛq_{0}=\delta_{\Lambda}. Initially, the dPMF dynamics solving the time-changed PDE is that of a noninteracting Wiener process with negative unit drift and delayed reset at Λ\Lambda. For σ>0\sigma>0, such dynamics admits a time-dependent density xq(σ,x)x\mapsto q(\sigma,x) that is smooth on (0,)(0,\infty), except for a slope discontinuity at the reset site Λ\Lambda. Moreover, the density xq(σ,x)x\mapsto q(\sigma,x) is locally differentiable in 0+0^{+} and the time-changed rate of inactivation g(σ)g(\sigma) is determined as the instantaneous flux: g(σ)=xq(σ,0)/2g(\sigma)=\partial_{x}q(\sigma,0)/2. For large enough λ\lambda, the flux gg crosses the level 1/λ1/\lambda as a locally strictly convex increasing function at time S1S_{1}. In principle, this would correspond to the inverse time change Φ:σ(σλG(σ))/ν\Phi:\sigma\mapsto(\sigma-\lambda G(\sigma))/\nu admitting a strict local maximum at S1S_{1}. However, such a behavior is not allowed as it would indicate that T1=Φ(S1)T_{1}=\Phi(S_{1}) is a time-reversal point for the original dPMF dynamics. This reveals T1T_{1} as a blowup time for the original dynamics and S1S_{1} as a blowup trigger time for the time-changed dynamics. During a blowup episode, the original time freezes, while the time changed dynamics is allowed to proceed unimpeded. This corresponds to imposing that the inverse time change be flat via the definition Φ(σ)=sup0ξσ(ξλG[η](ξ))/ν\Phi(\sigma)=\sup_{0\leq\xi\leq\sigma}\big{(}\xi-\lambda G[\eta](\xi)\big{)}/\nu. Accordingly, during a blowup episode, the backward-delay function η\eta increases with slope one and no reset can occur. In other words, during a blowup episode, the time-changed dynamics loses its reset character and probability mass is gradually lost by inactivation in zero. Eventually, this entails a decrease in the inactivation rate g(σ)g(\sigma) and an increase in σλG[η](σ))/ν\sigma-\lambda G[\eta](\sigma)\big{)}/\nu, up to a time U1U_{1} when it reaches the value T1=Φ(S1)T_{1}=\Phi(S_{1}) anew. The time U1U_{1} marks the blowup exit time and necessarily satisfies U1S1=λπ1U_{1}-S_{1}=\lambda\pi_{1} where π1\pi_{1} is the fraction of processes that inactivate during the blowup episode. Moreover, it is also such that the mean value of gg over the blowup episode is exactly 1/λ1/\lambda. Finally, the inactivated fraction π1\pi_{1} is reset all at once at a later time R1R_{1}, referred to as the blowup reset time for which Φ(R1)=T1+ϵ\Phi(R_{1})=T_{1}+\epsilon. This instantaneous reset corresponds to a jump discontinuity in the backward-delay function of size λπ1-\lambda\pi_{1}.

In [22], we further showed that the cumulative flux G[η]G[\eta], introduced in the PDE problem defined by equations (6) and (7), is most conveniently characterized in terms of an integral equation obtained via renewal analysis:

Definition 1.4.

Given a time-inhomogeneous backward-delay function η:++\eta:\mathbbm{R}^{+}\to\mathbbm{R}^{+}, the cumulative flux GG is given as the unique solution to the quasi-renewal equation

G(σ)=0H(σ,x)q0(x)dx+0σH(στ,Λ)dG(τη(τ)),\displaystyle G(\sigma)=\int_{0}^{\infty}H(\sigma,x)q_{0}(x)\,\mathrm{d}x+\int_{0}^{\sigma}H(\sigma-\tau,\Lambda)\,\mathrm{d}G(\tau-\eta(\tau))\,, (13)

where by convention we set G(σ)=G0(σ)=0σg0(ξ)dξG(\sigma)=G_{0}(\sigma)=\int_{0}^{\sigma}g_{0}(\xi)\mathrm{d}\xi if ξ0σ<0-\xi_{0}\leq\sigma<0. The integration kernel featured in (13) is specified as σH(σ,x)=[τxσ]\sigma\mapsto H(\sigma,x)=\mathbb{P}\left[\tau_{x}\leq\sigma\right], where τx\tau_{x} is the first-passage time to zero of a Wiener process started in x>0x>0 and with negative unit drift.

Definitions 1.3 and 1.4 fully specify the dPMF fixed-point problem in the time-change picture. In this time-changed picture, the occurrence of a (first) blowup involves two distinct times: the blowup trigger time S1S_{1} and the blowup exit time U1U_{1}, so that a synchronous event of size π1\pi_{1} corresponds to Ψ\Psi remaining flat on [S1,U1)[S_{1},U_{1}) with S1=U1+λπ1S_{1}=U_{1}+\lambda\pi_{1} (see Fig. 4). The definition of the blowup trigger time S1S_{1} directly follows from Proposition 1.1, which can be recast in the time-changed picture as:

Definition 1.5.

Under Assumption 1, the first blowup trigger time is defined as

S1=inf{σ>0|g(σ)>1/λ}.\displaystyle S_{1}=\inf\{\sigma>0\,|\,g(\sigma)>1/\lambda\}\,. (14)

where g(σ)=σG(σ)=xq(σ,0)/2g(\sigma)=\partial_{\sigma}G(\sigma)=\partial_{x}q(\sigma,0)/2 is the smooth instantaneous flux in zero.

In turn, the full-blowup assumption 2 admits a more natural but equivalent formulation in the time-changed picture:

Assumption 3.

At the blowup trigger time S1S_{1}, the smooth instantaneous flux gg is such that σg(S1)>0\partial_{\sigma}g(S_{1})>0.

The above formulation of the full-blowup assumption is natural in view of the fixed-point problem 1.3. Indeed, such an assumption implies that the function σσλG(σ)\sigma\mapsto\sigma-\lambda G(\sigma) becomes decreasing on a finite interval to the left of S1S_{1}. Thus, as a solution to the fixed-point problem (11), Ψ\Psi must remain flat on a nonempty maximum interval [S1,U1)[S_{1},U_{1}), which corresponds to a synchronous event of size π1=(S1U1)/λ\pi_{1}=(S_{1}-U_{1})/\lambda (see Fig. 4). Incidentally, this observation sheds light on Definition 1.1 characterizing full blowups. During a full blowup episode, Ψ\Psi remains flat so that the backward-delay function is constant on [S1,U1)[S_{1},U_{1}). As a result, the renewal-type integral term in (13) vanishes and GG is revealed as the cumulative flux of a linear time-changed dynamics YσY_{\sigma}, but without birth term due to reset. This amounts to stopping the clock for the original time coordinate tt, while letting the clock run in the changed coordinate σ\sigma. Such a stoppage can only be maintained until Ψ\Psi exceeds its blowup trigger value Ψ(S1)\Psi(S_{1}) (see Fig. 4), justifying the definition of blowup exit times U1U_{1} as:

Definition 1.6.

The first blowup exit time satisfies

U1=inf{σ>0|Ψ(σ)>Ψ(S1)}=S1+λπ1,\displaystyle U_{1}=\inf\{\sigma>0\,|\,\Psi(\sigma)>\Psi(S_{1})\}=S_{1}+\lambda\pi_{1}\,, (15)

where π1\pi_{1} is determined by the self-consistent condition (3).

The above definition of the blowup exit time is consistent with the characterization of the blowup size π1\pi_{1} given in (3). Indeed, one can check that

U1=inf{σ>0|Ψ(σ)>Ψ(S1)}=inf{σ>0|σS1>λ(G(σ)G(S1))}.\displaystyle U_{1}=\inf\{\sigma>0\,|\,\Psi(\sigma)>\Psi(S_{1})\}=\inf\{\sigma>0\,|\,\sigma-S_{1}>\lambda(G(\sigma)-G(S_{1}))\}\,.

Then, introducing the reduced variable p=(σS1)/λp=(\sigma-S_{1})/\lambda, we consistently have

U1\displaystyle U_{1} =\displaystyle= S1+λinf{p>0|p>G(S1+λp)G(S1)},\displaystyle S_{1}+\lambda\inf\{p>0\,|\,p>G(S_{1}+\lambda p)-G(S_{1})\}\,,
=\displaystyle= S1+λinf{p>0|p>[τXT1<λp]},\displaystyle S_{1}+\lambda\inf\left\{p>0\,|\,p>\mathbb{P}\left[\tau_{X_{T_{1}}}<\lambda p\right]\right\}\,,

where the last equality follows from the fact that the process YσY_{\sigma} does not reset on [S1,U1)[S_{1},U_{1}). For nonzero vanishing period ϵ>0\epsilon>0, the processes that inactivate during a blowup episode reset after a period of duration ϵ\epsilon elapses in original time coordinate. Thus, the reset time of a representative process R1R_{1} shall satisfy Ψ(R1)=Ψ(S1)+ϵ\Psi(R_{1})=\Psi(S_{1})+\epsilon. The latter relation uniquely specifies R1R_{1} if Φ=Ψ1\Phi=\Psi^{-1} is continuous in T1+ϵ=Ψ(S1)+ϵT_{1}+\epsilon=\Psi(S_{1})+\epsilon, which holds true if no blowup happens in T1+ϵT_{1}+\epsilon (see Fig. 4). Following on [22], without any continuity assumption, the blowup reset time R1R_{1} is generally defined as the leftmost solution of Ψ(R1)=Ψ(S1)+ϵ\Psi(R_{1})=\Psi(S_{1})+\epsilon:

Definition 1.7.

The first blowup reset time satisfies

R1=inf{σ>0|Ψ(σ)Ψ(S1)+ϵ}.\displaystyle R_{1}=\inf\{\sigma>0\,|\,\Psi(\sigma)\geq\Psi(S_{1})+\epsilon\}\,. (16)

The definitions of the blowup trigger time S1S_{1}, the exit time U1U_{1}, and the reset time R1R_{1} illuminate how the time-changed process YσY_{\sigma} resolves a blowup episode by alternating two types of dynamics (see Fig. 4). Before a full blowup, the process YσY_{\sigma} follows a linear diffusion with absorption in zero and resets at Λ\Lambda. These resets occur with time-inhomogeneous delays, which depends on the inverse time change Ψ\Psi. At the blowup onset S1S_{1}, Ψ\Psi becomes locally flat, indicating that the original time t=Ψ(σ)t=\Psi(\sigma) freezes, thereby stalling resets. As a result, after the blowup onset in S1S_{1}, the dynamics of YσY_{\sigma} remains that of an absorbed linear diffusion but without resets. Such a dynamics persists until a self-consistent blowup exit condition is met in S1+λπ1S_{1}+\lambda\pi_{1}. This condition states that it must take λπ1\lambda\pi_{1} time-changed units for a fraction π1\pi_{1} of processes to inactivate during a blowup episode. Finally, the original time t=Ψ(σ)t=\Psi(\sigma) resume flowing past U1U_{1} and the inactivated fraction π1\pi_{1} is reset at Λ\Lambda at time R1R_{1}. Importantly, observe that the time-change process YσY_{\sigma} always behaves regularly in the sense that GG remains a smooth function throughout the blowup episode.

1.4 Results

In principle, dPMF dynamics could be continued past a blowup episode, and possibly even extended to the whole half-line +\mathbbm{R}^{+}. Simulating the particle systems conjectured to approximate dPMF dynamics suggests that these dynamics can sustain repeated synchronization events over the whole half-line +\mathbbm{R}^{+} (see Fig. 2). This leads to conjecture the existence of solutions Ψ\Psi globally defined on +\mathbbm{R}^{+} with a countable infinity of blowup episodes with associated sequence of blowup times {Sk,Uk,Rk}k\{S_{k},U_{k},R_{k}\}_{k\in\mathbbm{N}}. Extending results from [22] to show the existence of such global solutions requires showing:

  • Unconditional explosiveness: (i)(i) the so-called non-false-start exit condition σq(Uk,0)/2<1/λ\partial_{\sigma}q(U_{k},0)/2<1/\lambda and (ii)(ii) the full-blowup condition σ2Ψ(Sk)<0\partial^{2}_{\sigma}\Psi(S_{k}^{-})<0 are satisfied for all blowup episodes, not just for the first one.

  • Full-time domain: the blowup times Tk=Ψ(Sk)=Ψ(Uk)T_{k}=\Psi(S_{k})=\Psi(U_{k}) do not have an accumulation point, i.e., limkTk=\lim_{k\to\infty}T_{k}=\infty, which means that the explosive solution is defined over the whole half-line +\mathbbm{R}^{+}.

The main difficulties in establishing the two above properties is that blowup times may not be well ordered, i.e., Rk>Sk+1R_{k}>S_{k+1} for some k>0k>0 and that blowup sizes πk\pi_{k} may exhibit vanishing sizes: limkπk=0\lim_{k\to\infty}\pi_{k}=0. However, one can exclude these possibilities in the strong interaction regime λΛ\lambda\gg\Lambda, for which the stable repetition of synchronization events can be understood intuitively:

(i)(i) Consider an initial condition of the form q0=π0δΛq_{0}=\pi_{0}\delta_{\Lambda} with π0=1\pi_{0}=1, which corresponds to a full reset at time R0=0R_{0}=0 and which satisfies Assumption 1. For a full blowup to occur next, the post-reset flux gg needs to satisfy the full-blowup Assumption 3 at some trigger time S1S_{1}, where we must have g(S1)=1/λg(S_{1})=1/\lambda and σg(S1)>0\partial_{\sigma}g(S_{1})>0. For large enough λ\lambda, we expect this flux gg to be primarily due to processes that do not reset more than once on [0,S1][0,S_{1}]. In other words, we have g(σ)h(σ,Λ)g(\sigma)\simeq h(\sigma,\Lambda), where h(,Λ)h(\cdot,\Lambda) denotes the first-passage density of a Wiener process started at Λ\Lambda with unit negative drift and with a zero boundary. The function h(,Λ)h(\cdot,\Lambda) is known to be a strictly increasing convex function on some nonempty interval [0,σ)[0,\sigma^{\dagger}) with h(0,Λ)=0h(0,\Lambda)=0. Thus, choosing a large enough λ>Λ\lambda>\Lambda guarantees that the full-blowup Assumption 3 will be met at some finite time S1>0S_{1}>0 with S1<σS_{1}<\sigma^{\dagger}. One can check the scaling S11/lnλS_{1}\propto 1/\ln\lambda, so that the fraction of inactive processes at trigger time S1S_{1}, denoted P(S1)P(S_{1}), can be made arbitrarily small for large enough λ\lambda. Actually, provided the refractory period is such that ϵ=O(1/λ)\epsilon=O(1/\lambda), one can find an upper bound for P(S1)P(S_{1}) of the form O(1/λ)O(1/\lambda).

(ii)(ii) Past the trigger time S1S_{1}, a fraction 1P(S1)1-P(S_{1}) of processes are susceptible to synchronize during the blowup episode. It turns out that the actual size of the blowup π1<1P(S1)\pi_{1}<1-P(S_{1}) can be made arbitrary close to 1P(S1)1-P(S_{1}) for large enough λ\lambda. Again, this is because we still have g(σ)h(σ,Λ)g(\sigma)\simeq h(\sigma,\Lambda) past the trigger time S1S_{1}, which satisfies S1<σS_{1}<\sigma^{\dagger}. This observation, together with the monotonicity properties of h(,Λ)h(\cdot,\Lambda), implies that g(σ)>1/λg(\sigma)>1/\lambda on [σ,σ][\sigma^{\dagger},\sigma^{\star}], where σ\sigma^{\star} is the unique maximizer of h(,Λ)h(\cdot,\Lambda). This means that the blowup episode includes at least the nonempty time interval [σ,σ][\sigma^{\dagger},\sigma^{\star}], during which at least a fraction ΔH=σσh(σ,Λ)dσ>0\Delta_{H}=\int_{\sigma^{\dagger}}^{\sigma^{\star}}h(\sigma,\Lambda)\mathrm{d}\sigma>0 synchronizes. Thus, we must have π1ΔH>0\pi_{1}\geq\Delta_{H}>0, so that the duration of the blowup λπ1λΔH\lambda\pi_{1}\geq\lambda\Delta_{H} can be made arbitrary long for large enough λ\lambda. The key is then to observe that the tail behavior of h(,Λ)h(\cdot,\Lambda) indicates that inactivation during blowup asymptotically happens with hazard rate 1/21/2 for large times. This observation allows one to deduce an exponential lower bound for the blowup size of the form, e.g., π1(1P(S1))(1O(eλΔH/4))>1/2\pi_{1}\geq\big{(}1-P(S_{1})\big{)}\big{(}1-O(e^{-\lambda\Delta_{H}/4})\big{)}>1/2.

(iii)(iii) Altogether, less than a fraction 1π1O(1/λ)1-\pi_{1}\leq O(1/\lambda) of processes survive the blowup at exit time U1=S1+λπ1U_{1}=S_{1}+\lambda\pi_{1} with π1>1/2\pi_{1}>1/2. The distribution of these surviving processes is such that xq(U1,0)/2h(U1,Λ)O(eU1/2)=O(eλ/4)<1/λ\partial_{x}q(U_{1},0)/2\simeq h(U_{1},\Lambda)\leq O(e^{-U_{1}/2})=O(e^{-\lambda/4})<1/\lambda for large enough λ\lambda so that the no-false-start Assumption 1 can be met for large enough λ>Λ\lambda>\Lambda. Moreover, we expect the fraction of surviving processes at U1U_{1}, i.e., q(U1,)1\|q(U_{1},\cdot)\|_{1}, to be so small that they cannot trigger a blowup alone. In other words, the assumption of well-ordered blowups holds: the fraction of inactivated processes π1\pi_{1} must reset at some time R1>U1R_{1}>U_{1} before any subsequent blowup can occur. Then, choosing π1=1O(1/λ)\pi_{1}=1-O(1/\lambda) close enough to one ensures that gg remains primarily shaped by those processes that reset at time R1R_{1}, so that g(σ)π1h(σR1,Λ)g(\sigma)\simeq\pi_{1}h(\sigma-R_{1},\Lambda). In turn, if the well-ordered property of blowups holds, we can apply (i)(i) anew but starting from R1R_{1} with a controlled probability mass π1\pi_{1}, rather than the full mass π0=1\pi_{0}=1.

(iv)(iv) To prove the well ordered property of blowups, one has to establish some a priori bounds about the boundary flux gg with respect to the L1L_{1} norm q(U1,)1\|q(U_{1},\cdot)\|_{1}. Establishing such a priori bounds requires to restrict the set of considered initial conditions to the so-called set of natural initial conditions. In a nutshell, these are those—not necessarily normalized—initial distributions such that all the active and inactive processes at the starting time have been reset at Λ\Lambda at some point in the past. For bearing on processes started away from zero at Λ\Lambda, natural initial conditions cannot lead to large transient variations in boundary fluxes. As a result, one can show that in the absence of reset, |g||g| and |σg||\partial_{\sigma}g| are uniformly upper bounded by O(q(U1,)1)O(\|q(U_{1},\cdot)\|_{1}). Then, the well-ordered property of blowups follows from showing that |g|=O(q(U1,)1)<1/λ|g|=O(\|q(U_{1},\cdot)\|_{1})<1/\lambda in the absence of the blowup reset. This requires that one has at least that q(U1,)1=O(1/λ)\|q(U_{1},\cdot)\|_{1}=O(1/\lambda), consistent with the fact that q(U1,)11π1=O(1/λ)\|q(U_{1},\cdot)\|_{1}\leq 1-\pi_{1}=O(1/\lambda). but not precise enough to conclude without explicit knowledge of the bounding coefficients..

(v)(v) Establishing that the no-false-start condition (Assumption 1), the full-blow up condition (Assumption 2), and the well ordered condition can be all met requires to estimate coefficients in the various λ\lambda-dependent bounds at play. Then, the strategy is to use these explicit bounds to show that given a large enough λ>Λ\lambda>\Lambda, there exists a threshold mass Δπ\Delta_{\pi}, 1/2<Δπ<11/2<\Delta_{\pi}<1, such that for all blowups of size π1Δπ\pi_{1}\geq\Delta_{\pi}, the next full blowup also satisfies π2>Δπ\pi_{2}>\Delta_{\pi}. From there, a direct induction argument on the number of blowups shows that there is a unique inverse time change Ψ:++\Psi:\mathbbm{R}^{+}\mapsto\mathbbm{R}^{+} dynamics as long as the initial mass π0Δπ\pi_{0}\geq\Delta_{\pi}. The fact that the domain of Ψ\Psi is the full half-line +\mathbbm{R}^{+} follows from Ψ\Psi having a countable infinity of blowup with size at least Δπ>1/2\Delta_{\pi}>1/2. The fact that the domain of Φ=Ψ1\Phi=\Psi^{-1} is also the full half-line +\mathbbm{R}^{+} follows from the fact that well ordered blowups must be separated by the duration of the nonzero refractory period ϵ>0\epsilon>0.

Making the arguments above rigorous leads to our first main result:

Theorem 1.8.

Consider some normalized natural initial conditions (pϵ,0,fϵ,0)(p_{\epsilon,0},f_{\epsilon,0}) in (+)×([ξ0,0))\mathcal{M}(\mathbbm{R}^{+})\times\mathcal{M}([\xi_{0},0)) such that pϵ,0p_{\epsilon,0} contains a Dirac-delta mass πϵ,0\pi_{\epsilon,0}, with 0<πϵ,010<\pi_{\epsilon,0}\leq 1, at reset value Λ\Lambda.

For small enough ϵ>0\epsilon>0, large enough λ>Λ\lambda>\Lambda, and large enough πϵ,0<1\pi_{\epsilon,0}<1, there exists a unique explosive dPMF dynamics parametrized by the time change Φϵ:++\Phi_{\epsilon}:\mathbbm{R}^{+}\rightarrow\mathbbm{R}^{+}, with a countable infinity of blowups. These blowups occurs at consecutive times Tϵ,kT_{\epsilon,k}, kk\in\mathbbm{N}, with size πϵ,k=(Φϵ(Tϵ,k)Φϵ(Tϵ,k))/λ\pi_{\epsilon,k}=(\Phi_{\epsilon}(T_{\epsilon,k})-\Phi_{\epsilon}(T_{\epsilon,k}^{-}))/\lambda, and are such that πϵ,k\pi_{\epsilon,k} and Tϵ,k+1Tϵ,kT_{\epsilon,k+1}-T_{\epsilon,k} are both uniformly bounded away from zero.

In the following, we prove an equivalent version of the above theorem for the time-changed dynamics, i.e., Theorem 3.2. The proof of Theorem 3.2 only relies on a few properties of the smooth first-passage density function h(,Λ)h(\cdot,\Lambda), which plays the key role throughout this analysis. Although we do not refer to these properties explicitly in the proof, we list them as follows:

Property 1.

The function h(,Λ)h(\cdot,\Lambda) is both strictly convex and log-concave on some finite interval [0,σ][0,\sigma^{\dagger}] with h(0)=0h(0)=0.

Property 2.

The hazard rate function of h(,Λ)h(\cdot,\Lambda) is asymptotically bounded away from zero and has at most exponential growth, i.e.:

lim infσh(σ,Λ)σh(ξ,Λ)r>0andlim supσln(h(σ,Λ)σh(ξ,Λ))<.\displaystyle\liminf_{\sigma\to\infty}\;\frac{h(\sigma,\Lambda)}{\int_{\sigma}^{\infty}h(\xi,\Lambda)}\geq r>0\,\quad\mathrm{and}\quad\limsup_{\sigma\to\infty}\;\ln\left(\frac{h(\sigma,\Lambda)}{\int_{\sigma}^{\infty}h(\xi,\Lambda)}\right)<\infty\,.

Given some large enough Dirac-delta mass πk\pi_{k} at reset time RkR_{k}, Property 1 ensures that a full blowup occurs for large enough λ\lambda at trigger time Sk+1>RkS_{k+1}>R_{k} and with blowup size πk+1\pi_{k+1} lower bounded by some constant ΔH\Delta_{H}. This corresponds to the “ignition” phase of the blowup. Given that πk+1ΔH\pi_{k+1}\geq\Delta_{H}, Property 2 ensures that an exponentially small fraction of processes survives the blowup and that this fraction cannot lead to a subsequent blowup without reset, i.e., reignition. This corresponds to the “combustion” phase of the blowup. A potentially useful consequence of the above observations is that we expect our blowup analysis to extend to more generic diffusion processes. Indeed, the existence and uniqueness of explosive dPMF dynamics shall hold for all dynamics derived from autonomous diffusion processes whose first-passage time to a constant level has smooth densities satisfying Property 1 and 2. This class of diffusion processes includes the Ornstein-Uhlenbeck process, which forms the basis of a widely used class of models in computational neuroscience, called leaky integrate-and-fire neurons [16, 6].

In stating Theorem 1.8, we do note give a precise criterion for how to jointly choose the interaction parameter λ>Λ\lambda>\Lambda, the refractory period ϵ>0\epsilon>0, and the initial mass π0\pi_{0} to yield explosive dPMF dynamics. This is because the precise criterion given in the proof of Theorem 1.8 is not tight, our main goal being to establish existence and uniqueness of global solution in the large interaction regime. Actually, we expect explosive dPMF dynamics to arise as soon as λ>Λ\lambda>\Lambda for small ϵ0\epsilon\geq 0 and we expect this emergence to be essentially independent of the size of the initial mass π0\pi_{0}. One can see the latter point by considering the limit case of PMF dynamics with zero refractory period in the time change picture. In the absence of blowups, plugging ϵ=0\epsilon=0 in Definition 1.3 yields valid, nondelayed dynamics with explicitly known instantaneous boundary flux gg via renewal analysis [1]. Moreover, classical arguments from renewal analysis show that gg is analytic away from zero and that limσg(σ)=1/Λ\lim_{\sigma\to\infty}g(\sigma)=1/\Lambda, independent of the initial conditions. Thus, it is natural to expect that PMF dynamics exhibit a countable infinity of blowups (at least) as soon as λ>Λ\lambda>\Lambda.

One general caveat is that checking the mean-field conjecture becomes exceedingly costly when λΛ\lambda\simeq\Lambda. Indeed we find that the convergence to an asymptotically independent regimes for large network size NN\to\infty drastically slows down when λΛ\lambda\simeq\Lambda —a numerical evidence of critical slowing down [20]. However, the conjectured PMF dynamics can always be analyzed in the time-changed picture by setting ϵ=0\epsilon=0 in Definition 1.3. One specific caveat remains that it is not clear why blowups are resolved in PMF dynamics via the mechanism implied by Definition 1.3 with ϵ=0\epsilon=0. This is by contrast with dPMF dynamics for which the nonzero refractory period ϵ>0\epsilon>0 makes it clear that blowup episodes correspond to halting reset at the time-changed picture. In view of this, our second main result is to justify that PMF dynamics “physically” resolve blowups by showing that these dynamics are recovered from dPMF ones in the limit ϵ0+\epsilon\to 0^{+}.

Theorem 1.9.

Consider some normalized natural initial condition p0p_{0} in (+)\mathcal{M}(\mathbbm{R}^{+}) such that p0p_{0} contains a Dirac-delta mass π0\pi_{0}, with 0<π010<\pi_{0}\leq 1, at reset value Λ\Lambda.

For large enough λ>Λ\lambda>\Lambda and large enough π0<1\pi_{0}<1, the unique explosive PMF dynamics Φ:++\Phi:\mathbbm{R}^{+}\rightarrow\mathbbm{R}^{+} satisfies Φ1=limϵ0+Φϵ1\Phi^{-1}=\lim_{\epsilon\to 0^{+}}\Phi^{-1}_{\epsilon} with compact convergence over +\mathbbm{R}^{+}, where Φϵ\Phi_{\epsilon} are dPMF dynamics with same initial condition p0p_{0}. Moreover, we also have convergence of the blowup times and blowup sizes: limϵ0+Tϵ,k=Tk\lim_{\epsilon\to 0^{+}}T_{\epsilon,k}=T_{k} and limϵ0+πϵ,k=πk\lim_{\epsilon\to 0^{+}}\pi_{\epsilon,k}=\pi_{k}.

In the following, we prove an equivalent version of the above theorem for the time-changed dynamics, i.e., Theorem 4.4. Given an explosive dPMF dynamics parametrized by Φϵ=Ψϵ1\Phi_{\epsilon}=\Psi_{\epsilon}^{-1}, proving Theorem 4.4 amounts to showing that the ϵ\epsilon-dependent backward-delay function

ηϵ(σ)=σΦϵ(Ψϵ(σ)ϵ)\displaystyle\eta_{\epsilon}(\sigma)=\sigma-\Phi_{\epsilon}(\Psi_{\epsilon}(\sigma)-\epsilon) (17)

is well behaved in the limit ϵ0+\epsilon\to 0^{+}. By well behaved, we mean that (1)(1) ηϵ\eta_{\epsilon} converges to zero in between blowups on intervals [Rk,Sk][R_{k},S_{k}], consistent with the nondelayed nature of PMF dynamics, or (2)(2) ηϵ\eta_{\epsilon} converges toward the unit slope function σσSk\sigma\mapsto\sigma-S_{k} on blowup intervals [Sk,Uk][S_{k},U_{k}], consistent with halting reset at SkS_{k}. Such an alternative covers all cases as we expect instantaneous reset after blowup exit: Uk=RkU_{k}=R_{k} for ϵ=0\epsilon=0. This corresponds to assuming a limit of the form

η(σ)=σΦ(Ψ(σ)),\displaystyle\eta(\sigma)=\sigma-\Phi(\Psi(\sigma))\,,

where crucially, Φ\Phi is the right-continuous inverse of Ψ\Psi, just as if one consider the fixed-point problem of Definition 1.3 with ϵ=0\epsilon=0. The main difficulty is that (17) does not behave continuously for standard topologies when it bears on functions Φϵ\Phi_{\epsilon} with jump discontinuities. One route to address this point is to consider topologies specially designed to deal with càdlàg functions such as the Skorokhod topologies [2, 9]. However, the continuity of the fixed-point problem specified in Definition 1.3 with respect to such topologies does not appear straightforward.

A more pedestrian route is to leverage the fact that for small enough ϵ>0\epsilon>0, the time change Φϵ\Phi_{\epsilon} has only a countable infinity of blowups for large enough interaction parameter λ>Λ\lambda>\Lambda and large enough initial reset mass π0<1\pi_{0}<1. Under these standard conditions, one can show that the associated blowup times (Sϵ,k,Uϵ,k,Rϵ,k)(S_{\epsilon,k},U_{\epsilon,k},R_{\epsilon,k}), kk in \mathbbm{N}, can be uniformly controlled with respect to ϵ0+\epsilon\to 0^{+}. Such control can be established by considering separately the three stages of a time-change blowup dynamics: (1)(1) the blowup trigger stage on [Rϵ,k,Sϵ,k+1)[R_{\epsilon,k},S_{\epsilon,k+1}), (2)(2) the blowup resolution stage on [Sϵ,k+1,Uϵ,k+1)[S_{\epsilon,k+1},U_{\epsilon,k+1}), and (3)(3) the blowup reset stage [Uϵ,k+1,Rϵ,k+1)[U_{\epsilon,k+1},R_{\epsilon,k+1}). The key observation is that each of these stages has a duration that continuously depends on ϵ\epsilon and on the L1L_{1} norm of the current state of the dynamics at the start of the stage. Note that these current states represent natural initial conditions in ((0,))×([ηϵ(σ),0))\mathcal{M}((0,\infty))\times\mathcal{M}([-\eta_{\epsilon}(\sigma),0)) specified by (qϵ(σ,),{gϵ(ξσ)}ηϵ(σ)ξ<0)\big{(}q_{\epsilon}(\sigma,\cdot),\{g_{\epsilon}(\xi-\sigma)\}_{-\eta_{\epsilon}(\sigma)\leq\xi<0}\big{)} for σ=Rϵ,k,Sϵ,k+1,Uϵ,k+1\sigma=R_{\epsilon,k},S_{\epsilon,k+1},U_{\epsilon,k+1}. Both the continuity of the trigger time Sϵ,k+1S_{\epsilon,k+1} and of the exit time Uϵ,k+1U_{\epsilon,k+1} follow from nondegeneracy conditions, namely that the full-blowup condition σg(Sk+1)>0\partial_{\sigma}g(S_{k+1})>0 and the non-false-start condition σg(Uk+1)<1/λ\partial_{\sigma}g(U_{k+1})<1/\lambda holds uniformly with respect to ϵ\epsilon. The continuity of the reset time Rk+1R_{k+1} is straightforward for well-ordered blowups. Thus, to propagate these continuity results by induction on the number of blowups, one just has to show that for each stage, considering the terminal dynamical state as a function of the initial dynamical state specifies a continuous mapping in the L1L_{1} norm. Technically, establishing such L1L_{1} continuity results, jointly with ϵ\epsilon-continuity, is primarily made possible by the availability of certain uniform bounds for the heat kernel at times bounded away from zero and infinity.

The above discussion lays out our strategy to prove that PMF dynamics with zero refractory period represents “physical” solutions, which is one of our main objectives. PMF dynamics are of special interest because they admit an explicit iterative construction in terms of well-known analytic functions. Moreover, PMF can be easily simulated within the PDE setting. Such simulations confirm the prediction from particle-system simulations that explosive PMF solutions become asymptotically periodic (see Fig. 2). In that respect, we conclude by mentioning that the existence of periodic PMF dynamics directly follows from our L1L_{1} continuity analysis with ϵ=0\epsilon=0. This is a consequence of Schauder’s fixed-point theorem [12] applied to the inter-exit-time mapping

𝒰:(q(Uk,),πk)(q(Uk+1,),πk+1)\displaystyle\mathcal{U}:\big{(}q(U_{k},\cdot),\pi_{k}\big{)}\mapsto\big{(}q(U_{k+1},\cdot),\pi_{k+1}\big{)}

defined on 𝒞\mathcal{C}, the convex set of normalized natural initial conditions with large enough blowup mass πk\pi_{k} (given some large enough λ>Λ\lambda>\Lambda). Note that we automatically have q(Uk,)1+πk=q(Uk+1,)1+πk+1=1\|q(U_{k},\cdot)\|_{1}+\pi_{k}=\|q(U_{k+1},\cdot)\|_{1}+\pi_{k+1}=1 when ϵ=0\epsilon=0 so that the set 𝒞\mathcal{C} is stable by 𝒰\mathcal{U}. Our analysis shows that the mapping 𝒰\mathcal{U} is continuous with respect to the L1L_{1} norm. Then, all there is left to show in order to apply Schauder’s fixed-point theorem is that the mapping 𝒰\mathcal{U} is compact in 𝒞\mathcal{C} with respect to the L1L_{1} norm. But this follows directly from the fact that blowups are lower bounded by some Δπ>1/2\Delta_{\pi}>1/2. The latter point implies that the image of 𝒰\mathcal{U} is included in the image of a compact operator defined in terms of the heat kernel for times at least 1/21/2. Thus there must be a fixed point to 𝒰\mathcal{U} in 𝒞\mathcal{C}, which corresponds to a periodic PMF dynamics. Studying the general contractive property of PMF and dPMF dynamics is beyond the scope of this work.

1.5 Structure

In Section 2, we recall results from [22] that allows one to resolve a single dPMF blowup episode in the time-changed picture. In Section 3, we leverage these results to show that dPMF solution can be continued past blowup episodes to specify explosive solutions over the whole half-line +\mathbbm{R}^{+} in the large interaction regime. In Section 4, we show that explosive PMF dynamics, which can be defined explicitly for zero refractory period ϵ=0\epsilon=0, are recovered from dPMF dynamics in the limit ϵ0+\epsilon\to 0^{+}. Appendices A and B contain two technical lemmas that are needed to demonstrate the result of Section 4.

2 Regularization of explosive dynamics via change of time

In this section, we define the time-changed dynamics obtained by change of variable and introduce the delay functions that parametrize such a dynamics. We then recall the results from renewal theory that characterize the cumulative flux function associated with a time-changed dynamics Finally, we utilize these results to justify the definition of the fixed-point formulation of dPMF dynamics given in Definition 1.3.

2.1 Definition of the time-inhomogeneous linear dynamics

Following [22], our approach is based on representing possibly explosive dPMF dynamics Xt=YΦ(t)X_{t}=Y_{\Phi(t)} as time-changed versions of nonexplosive dynamics YσY_{\sigma}. A good choice for the time change function Φ\Phi is one for which the dynamics YσY_{\sigma} is simple enough to be analytically tractable. The Poisson-like attributes of dPMF dynamics suggest defining Φ\Phi implicitly as the integral function of the drift:

tΦ(t)=νt+λF(t).\displaystyle t\mapsto\Phi(t)=\nu t+\lambda F(t)\,. (18)

Such a definition equates blowups/synchronous events with singularities/discontinuities in the time change Φ\Phi, while allowing for the corresponding time-changed dynamics YσY_{\sigma} to always be devoid of blowups. As a result, showing the existence and uniqueness of a dPMF dynamics XtX_{t} reduces to showing the existence and uniqueness of a time change Φ\Phi satisfying relation (18). We shall seek such time changes Φ\Phi in a set of candidate time changes denoted by 𝒯\mathcal{T}. To define 𝒯\mathcal{T}, we utilize the fact that the cumulative function F:++F:\mathbbm{R}^{+}\to\mathbbm{R}^{+} featured in (18) can be safely assumed to be a nondecreasing function. This leads to the following definition:

Definition 2.1.

The set of candidate time change 𝒯\mathcal{T} is the set of càdlàg functions Φ:[ϵ,)[ξ0,)\Phi:[-\epsilon,\infty)\to[\xi_{0},\infty), such that their difference quotients are lower bounded by ν\nu: for all y,xϵy,x\leq-\epsilon, xyx\neq y, we have

wΦ(y,x)=Φ(y)Φ(x)yxν.\displaystyle w_{\Phi}(y,x)=\frac{\Phi(y)-\Phi(x)}{y-x}\geq\nu\,.

As alluded to in the introduction, it is actually most convenient to consider dPMF dynamics as parametrized by the inverse time change:

Definition 2.2.

Given a time change Φ\Phi in 𝒯\mathcal{T}, the inverse time change Φ1:[ξ0,)[ϵ,)\Phi^{-1}:[\xi_{0},\infty)\to[-\epsilon,\infty) is defined as the continuous function

σΨ(σ)=Φ1(σ)=inf{t0|Φ(t)>σ}.\displaystyle\sigma\mapsto\Psi(\sigma)=\Phi^{-1}(\sigma)=\inf\left\{t\geq 0\,|\,\Phi(t)>\sigma\right\}\,.

In order to specify the time changed dynamics YσY_{\sigma}, we further need to introduce the so-called backward-delay function η\eta, which represents the time-wrapped version of the refractory period ϵ\epsilon:

Definition 2.3.

Given a time change Φ\Phi in 𝒯\mathcal{T}, we define the corresponding backward-delay function η:[0,)+\eta:[0,\infty)\to\mathbbm{R}^{+} by

η(σ)=σΦ(Ψ(σ)ϵ),σ0.\displaystyle\eta(\sigma)=\sigma-\Phi\left(\Psi(\sigma)-\epsilon\right)\,,\quad\sigma\geq 0\,.

We denote the set of backward functions {η[Φ]}Φ𝒯\{\eta[\Phi]\}_{\Phi\in\mathcal{T}} by 𝒲\mathcal{W}.

As wΦνw_{\Phi}\geq\nu for all Φ\Phi in 𝒯\mathcal{T}, it is clear that for all η\eta in 𝒲\mathcal{W}, we actually have ηνϵ\eta\geq\nu\epsilon, so that all delays are bounded away from zero. Time-wrapped-delay function η\eta in 𝒲\mathcal{W} will serve to parametrize the time-changed dynamics obtained via Φ\Phi in 𝒯\mathcal{T}. In [22], we showed that these time-changed dynamics are that of a modified Wiener process YσY_{\sigma} with negative unit drift, inactivation on the zero boundary, and reset at Λ\Lambda after a refractory period specified by η\eta. Consequently, we defined time-changed dynamics as the processes YσY_{\sigma} solutions to the following stochastic evolution:

Definition 2.4.

Denoting the canonical Wiener process by WσW_{\sigma}, we define the time-changed processes YσY_{\sigma} as solutions to the stochastic evolution

Yσ=σ+0σ𝟙{Yξ>0}dWξ+ΛNση(σ),withNσ=n>0𝟙[ξ0,t](ξn),\displaystyle Y_{\sigma}=-\sigma+\int_{0}^{\sigma}\mathbbm{1}_{\{Y_{\xi^{-}}>0\}}\,\mathrm{d}W_{\xi}+\Lambda N_{\sigma-\eta(\sigma)}\,,\quad\mathrm{with}\quad N_{\sigma}=\sum_{n>0}\mathbbm{1}_{[\xi_{0},t]}(\xi_{n})\,, (19)

where the process NσN_{\sigma} counts the successive first-passage times ξn\xi_{n} of the process YσY_{\sigma} to the absorbing boundary:

ξn+1=inf{σ>0|ση(σ)>ξn,Yσ0}.\displaystyle\xi_{n+1}=\inf\left\{\sigma>0\,\big{|}\,\sigma-\eta(\sigma)>\xi_{n},Y_{\sigma}\leq 0\right\}\,.

A time-changed process YσY_{\sigma} is uniquely specified by imposing elementary initial conditions, which take an alternative formulation: either the process is active Y0=x>0Y_{0}=x>0 and N0=0N_{0}=0, either the process has entered refractory period at some earlier time ξ\xi so that Y0=0Y_{0}=0 and Nσ=𝟙σξN_{\sigma}=\mathbbm{1}_{\sigma\geq\xi} for ξ0σ<0\xi_{0}\leq\sigma<0. Generic initial conditions are given by considering that (x,ξ)(x,\xi) is sampled from some probability distribution on {(0,)×{0}}{{0}×[ξ0,0)}\{(0,\infty)\times\{0\}\}\cup\{\{0\}\times[\xi_{0},0)\}. In all generality, this amounts to choosing a normalized pair of distributions (q0,g0)(q_{0},g_{0}) in ((0,))×([ξ0,0))\mathcal{M}((0,\infty))\times\mathcal{M}([\xi_{0},0)). Given generic initial conditions (q0,g0)(q_{0},g_{0}) in ((0,))×([ξ0,0))\mathcal{M}((0,\infty))\times\mathcal{M}([\xi_{0},0)), the dynamics defined by 2.4 is well posed as long as the backward-delay function ηνϵ\eta\geq\nu\epsilon is locally bounded, which is always the case for valid time change Φ\Phi. Moreover, the corresponding density function (σ,x)q(σ,x)=d[0<Yσx]/dx(\sigma,x)\mapsto q(\sigma,x)=\mathrm{d}\mathbb{P}\left[0<Y_{\sigma}\leq x\right]/\mathrm{d}x must solve the PDE problem defined by (6) and (7) in the introduction. To establish the existence and uniqueness of global explosive solutions, we will actually consider a restricted class of initial conditions, referred to as natural conditions. We postpone the introduction of this so-called notion of natural initial conditions to Section 3.1.

Interestingly, interactions are entirely mediated by the time-delayed counting process Nση(σ)N_{\sigma-\eta(\sigma)} in the linear time-changed dynamics (19). This motivates introducing the so-called backward-time function σξ(σ)\sigma\mapsto\xi(\sigma), which marks the inactivation time of the processes being reset at time σ\sigma, and the associated forward-time function στ(σ)\sigma\mapsto\tau(\sigma), which marks the reset time of the processes inactivated at time σ\sigma.

Definition 2.5.

Given a time change Φ\Phi in 𝒯\mathcal{T}, we define the backward-time function by

ξ:+[ξ0,),ξ(σ)=ση(σ)=Φ(Ψ(σ)ϵ),\displaystyle\xi:\mathbbm{R}^{+}\to[\xi_{0},\infty)\,,\quad\xi(\sigma)=\sigma-\eta(\sigma)=\Phi\left(\Psi(\sigma)-\epsilon\right)\,,

and the forward-time function by

τ:[ξ0,)+,τ(σ)=inf{τ>0|ξ(τ)=τη(τ)σ}.\displaystyle\tau:[\xi_{0},\infty)\to\mathbbm{R}^{+}\,,\quad\tau(\sigma)=\inf\{\tau>0\,|\,\xi(\tau)=\tau-\eta(\tau)\geq\sigma\}\,. (20)

When unambiguous, we will denote these functions by ξ\xi and τ\tau refer to them as the “functions ξ\xi and τ\tau” to differentiate from when ξ\xi and τ\tau play the role of real variables. By construction, the function ξ\xi and τ\tau are nondecreasing càdlàg and nondecreasing càglàd functions, respectively. In particular, both functions can admit discontinuities and flat regions. In [22], we justify the definition of the function τ\tau as the left-continuous inverse of the function ξ\xi with a detailed analysis of the delayed reset mechanism in the dynamics (19). We also show there that if we denote the sequence of inactivation times by {ξk}k0\{\xi_{k}\}_{k\geq 0}, then the corresponding sequences of reset times {τk}k0\{\tau_{k}\}_{k\geq 0} satisfies τk=τ(ξk)\tau_{k}=\tau(\xi_{k}), k0k\geq 0.

2.2 Results from time-inhomogeneous renewal analysis

Given a backward-delay function η\eta in 𝒲\mathcal{W}, one can consider the PDE problem defined by (6) and (7) independently of the requirement that relation (18) be satisfied. Then, the main hurdle to solving this PDE problem is due to the presence of an inhomogeneous, delayed reset term featuring the cumulative flux function GG. Luckily, assuming the cumulative function GG known allows one to express the full solution of the inhomogeneous PDE in terms of its homogeneous solutions. These homogeneous solutions are known in closed form [14]:

κ(σ,y,x)=e(yx+σ)22σ2πσ(1e2xyσ).\displaystyle\kappa(\sigma,y,x)=\frac{e^{-\frac{(y-x+\sigma)^{2}}{2\sigma}}}{\sqrt{2\pi\sigma}}\left(1-e^{-\frac{2xy}{\sigma}}\right)\,. (21)

Applying Duhamel’s principle yields the following integral representation for the density function (σ,x)q(σ,x)=d[0<Yσx](\sigma,x)\mapsto q(\sigma,x)=\mathrm{d}\mathbb{P}\left[0<Y_{\sigma}\leq x\right]

q(σ,x)=0κ(σ,y,x)q0(x)dx+0σκ(στ,x,Λ)dG(ξ(τ)),\displaystyle q(\sigma,x)=\int_{0}^{\infty}\kappa(\sigma,y,x)q_{0}(x)\,\mathrm{d}x+\int_{0}^{\sigma}\kappa(\sigma-\tau,x,\Lambda)\,\mathrm{d}G(\xi(\tau))\,, (22)

where ξ=idη\xi=\mathrm{id}-\eta and where GG remains to be determined. The above integral representation shows that GG is the key determinant parametrizing the time-changed process YσY_{\sigma}.

Just as YσY_{\sigma}, GG only depends on the backward-delay functions η\eta, or equivalently on the functions ξ\xi or τ\tau. These dependences can be made explicit by adapting results from renewal analysis [22]. Informally, this follows from writing the cumulative flux function as the uniformly convergent series

G(σ)=𝔼[k=1𝟙{ξk<σ}]=k=1[ξkσ],\displaystyle G(\sigma)=\mathbb{E}\left[\sum_{k=1}^{\infty}\mathbbm{1}_{\left\{\xi_{k}<\sigma\right\}}\right]=\sum_{k=1}^{\infty}\mathbb{P}\left[\xi_{k}\leq\sigma\right]\,, (23)

and realizing that for all k1k\geq 1, [ξkσ]\mathbb{P}\left[\xi_{k}\leq\sigma\right] can be expressed in terms of the first-passage time cumulative distribution [14]:

H(σ,x)=𝟙{σ0}2(Erfc(xσ2σ)+e2xErfc(x+σ2σ)).\displaystyle H(\sigma,x)=\frac{\mathbbm{1}_{\{\sigma\geq 0\}}}{2}\left(\mathrm{Erfc}\left(\frac{x-\sigma}{\sqrt{2\sigma}}\right)+e^{2x}\mathrm{Erfc}\left(\frac{x+\sigma}{\sqrt{2\sigma}}\right)\right)\,. (24)

Specifically, considering the initial condition q0=δxq_{0}=\delta_{x} for simplicity, we have

[ξkσ|Y0=x]=H(k)(σ,x),\displaystyle\mathbb{P}\left[\xi_{k}\leq\sigma\,|\,Y_{0}=x\right]=H^{(k)}(\sigma,x)\,,

where the functions H(k)(σ,x)H^{(k)}(\sigma,x), k1k\geq 1, are defined as inhomogeneous iterated convolution:

H(k)(σ,x)=0H(στ(ξ))dH(k1)(ξ,x),H(1)(σ,x)=H(σ,x).\displaystyle H^{(k)}(\sigma,x)=\int_{0}^{\infty}H(\sigma-\tau(\xi))\,\mathrm{d}H^{(k-1)}(\xi,x)\,,\quad H^{(1)}(\sigma,x)=H(\sigma,x)\,.

These iterated functions H(k)(σ,x)H^{(k)}(\sigma,x) take a particularly simple form for constant backward-delay function η=d=ξ0\eta=d=-\xi_{0}. Indeed, by divisibility of the first-passage distributions

[ξkξ|Y0=x]=H(ξ(k1)d,x+(k1)Λ),\displaystyle\mathbb{P}\left[\xi_{k}\leq\xi\,|\,Y_{0}=x\right]=H(\xi-(k-1)d,x+(k-1)\Lambda)\,,

so that for the elementary initial condition q0=δxq_{0}=\delta_{x}, the cumulative flux function is

Gd(σ,x)=k=1H(ξ(k1)d,x+(k1)Λ).\displaystyle G_{d}(\sigma,x)=\sum_{k=1}^{\infty}H(\xi-(k-1)d,x+(k-1)\Lambda)\,.

Due to the quasi-renewal character of the dynamics YσY_{\sigma}, the cumulative flux GG admits an alternative characterization in terms of an integral equation. Because of the presence of inhomogeneous delays, this integral equation differs from standard convolution-type renewal equations. Specifically, we show in [22] that for generic initial conditions:

Proposition 2.1.

Given a backward-delay function η\eta in 𝒲\mathcal{W}, the cumulative flux function GG is the unique solution of the renewal-type equation:

G(σ)=0H(σ,x)q0(x)dx+0σH(στ,Λ)dG(ξ(τ)).\displaystyle G(\sigma)=\int_{0}^{\infty}H(\sigma,x)q_{0}(x)\,\mathrm{d}x+\int_{0}^{\sigma}H(\sigma-\tau,\Lambda)\,\mathrm{d}G(\xi(\tau))\,. (25)

As a solution to (25), it is clear that independent of the functions η\eta, ξ\xi, τ\tau, the cumulative function GG inherits all the regularity properties of σH(σ,Λ)\sigma\mapsto H(\sigma,\Lambda), i.e., GG must be a smooth function for σ>0\sigma>0. In particular, one can differentiate (25) with respect to σ\sigma and obtain a new renewal-type equation for the instantaneous flux gg:

g(σ)=0h(σ,x)q0(x)dx+0σh(στ,Λ)dG(ξ(τ)),\displaystyle g(\sigma)=\int_{0}^{\infty}h(\sigma,x)q_{0}(x)\,\mathrm{d}x+\int_{0}^{\sigma}h(\sigma-\tau,\Lambda)\,\mathrm{d}G(\xi(\tau))\,, (26)

and where h(,x)h(\cdot,x) denotes the density function of the cumulative function H(,x)H(\cdot,x):

h(σ,x)=σH(σ,x)=𝟙{σ0}xe(xσ)2/2σ2πσ3.\displaystyle h(\sigma,x)=\partial_{\sigma}H(\sigma,x)=\mathbbm{1}_{\{\sigma\geq 0\}}\frac{xe^{-(x-\sigma)^{2}/2\sigma}}{\sqrt{2\pi\sigma^{3}}}\,. (27)

Alternatively, (26) can be directly obtained from differentiating (22) with respect to xx in zero and using that fact that h(σ,x)=yκ(σ,0,x)/2h(\sigma,x)=\partial_{y}\kappa(\sigma,0,x)/2. In [22], renewal-type equations such as (25) and (26) play a crucial role in showing the local existence of explosive dPMF solutions. Here, these will provide us with a priori estimates that are useful to globally define explosive dPMF solutions over the whole half-line +\mathbbm{R}^{+}.

Finally, for constant delay, the renewal-type equation (25) turned into an time-homogeneous equation by a change of variable:

G(σ)=0H(σ,x)q0(x)dx+0σH(στd,Λ)dG(τ).\displaystyle G(\sigma)=\int_{0}^{\infty}H(\sigma,x)q_{0}(x)\,\mathrm{d}x+\int_{0}^{\sigma}H(\sigma-\tau-d,\Lambda)\,\mathrm{d}G(\tau)\,. (28)

Deriving the above equation with respect to σ\sigma leads to a convolution equation for the the flux gd=σGdg_{d}=\partial_{\sigma}G_{d}. This equation can be solved algebraically in the Laplace domain yielding:

g^d(u,x)=0gd(σ,x)euσ𝑑σ=ex(11+2u)du1eΛ(11+2u)du.\displaystyle\hat{g}_{d}(u,x)=\int_{0}^{\infty}g_{d}(\sigma,x)e^{-u\sigma}\,d\sigma=\frac{e^{x\left(1-\sqrt{1+2u}\right)-du}}{1-e^{\Lambda\left(1-\sqrt{1+2u}\right)-du}}\,.

Finally, the above explicit expression allows one to evaluate the long-term behavior of the flux as limσgd(σ,x)=limu0ug^d(u,x)\lim_{\sigma\to\infty}g_{d}(\sigma,x)=\lim_{u\to 0}u\hat{g}_{d}(u,x) [11], which gives the following useful result:

Proposition 2.2.

For constant delay dd, the flux function gdg_{d} admit a limit that is independent of the initial conditions:

limσgd(σ)=1Λ+d.\displaystyle\lim_{\sigma\to\infty}g_{d}(\sigma)=\frac{1}{\Lambda+d}\,.

Observe that the above result is valid for d=0d=0, in which case limσg(σ)=1/Λ>1/λ\lim_{\sigma\to\infty}g(\sigma)=1/\Lambda>1/\lambda as soon as λ>Λ\lambda>\Lambda. This shows that for zero refractory period ϵ\epsilon and for λ>Λ\lambda>\Lambda, a blowup must occur in finite time. However, it is unclear how to resolve the ensuing blowup episode, since defining spiking avalanches can be ambiguous when ϵ=0\epsilon=0. We resolve this ambiguity by viewing PMF dynamics as the limit of dPMF dynamics. That is, we define PMF by formally setting ϵ=0\epsilon=0 in the dPMF fixed-point problem. We then show that, for large enough interaction parameter λ>Λ\lambda>\Lambda, the resulting PMF dynamics are the limit of dPMF dynamics as ϵ0+\epsilon\to 0^{+}.

2.3 Fixed-point problem and local solutions

Renewal analysis allows one to prove the existence, uniqueness, and regularity of the time-changed dynamics YσY_{\sigma} assuming the backward-delay function η\eta known. However, η\eta is actually an unknown of the problem for being ultimately defined in terms of the time change Φ\Phi via Definition 2.3. Moreover, the time-changed dynamics YσY_{\sigma} exists for any η\eta in 𝒲\mathcal{W}, whereas given some initial conditions, we expect that a unique time change Φ\Phi in 𝒯\mathcal{T} parametrizes a dPMF dynamics. In [22], we characterize this unique time change Φ\Phi by imposing that the defining relation (18) be satisfied. Observing that FF is the cumulative flux of Xt=YΦ(t)X_{t}=Y_{\Phi}(t), relation (18) can be naturally reformulated as:

Φ(t)=νt+λF(t)=νt+λG(Φ(t))νt=Φ(t)λG(Φ(t)).\displaystyle\Phi(t)=\nu t+\lambda F(t)=\nu t+\lambda G(\Phi(t))\quad\Leftrightarrow\quad\nu t=\Phi(t)-\lambda G(\Phi(t))\,. (29)

Assuming GG known, one can specify the sought-after time change Φ(t)=σ\Phi(t)=\sigma by solving νt=σλG(σ)\nu t=\sigma-\lambda G(\sigma) for σ\sigma given arbitrary values of t0t\geq 0. However, one has to be mindful about the existence and multiplicity of solutions. By construction, we have G(0)=0G(0)=0, so that we can always set Φ(0)=0\Phi(0)=0. Then, we can uniquely specified Φ(t)=σ\Phi(t)=\sigma for increasing value of tt as long as σσλG(σ)\sigma\mapsto\sigma-\lambda G(\sigma) is an increasing, one-to-one function. Let us consider the smallest time S1S_{1} after which the smooth function σσλG(σ)\sigma\mapsto\sigma-\lambda G(\sigma) fails to be an increasing, one-to-one function. Then, Φ(t)\Phi(t) is uniquely defined within the range [0,S1)[0,S_{1}), or rather the inverse time change Ψ=Φ1\Psi=\Phi^{-1} is uniquely defined as

Ψ(σ)=(σλG(σ))/ν,σ[0,S1).\displaystyle\Psi(\sigma)=(\sigma-\lambda G(\sigma))/\nu\,,\quad\forall\sigma\;\in\;[0,S_{1})\,. (30)

If S1S_{1} is finite, we must have g(S1)=1/λg(S_{1})=1/\lambda so that S1S_{1} is a blowup trigger time as defined in (14), and setting T1=Ψ(S1)T_{1}=\Psi(S_{1}), the equation νT1=σλG(σ)\nu T_{1}=\sigma-\lambda G(\sigma) must admit multiple solutions for σ\sigma. This indicates that Φ\Phi has a jump discontinuity in T1T_{1}, with Φ(T1)=U1>S1=Φ(T1)\Phi(T_{1})=U_{1}>S_{1}=\Phi(T_{1}^{-}). To be consistent, U1U_{1} shall be the smallest solution such that U1>S1U_{1}>S_{1} and σσλG(σ)\sigma\mapsto\sigma-\lambda G(\sigma) is increasing, one-to-one in the right vicinity of U1U_{1}. This corresponds to defining U1U_{1} as an blowup exit time as in (16) and imposing that

Ψ(σ)=Ψ(S1),σ[S1,U1).\displaystyle\Psi(\sigma)=\Psi(S_{1})\,,\quad\forall\sigma\;\in\;[S_{1},U_{1})\,. (31)

In principle, one can hope to continue defining the time change value Φ(t)\Phi(t) by extending the same reasoning for times tT1t\geq T_{1}, leading to defining an intertwined sequence of blowup trigger times {Sk}k1\{S_{k}\}_{k\geq 1} and blowup exit time {Uk}k1\{U_{k}\}_{k\geq 1}, with U1S1U2S2U_{1}\leq S_{1}\leq U_{2}\leq S_{2}\ldots. Such a construction produces a global solution over R+\mathrm{R}^{+} if the resulting sequences are devoid of accumulation points: limkSk=limkUk=\lim_{k\to\infty}S_{k}=\lim_{k\to\infty}U_{k}=\infty. In this work, we are primarily concerned with showing that the latter limits hold for large enough interactions and natural initial conditions so that there are no caveats due to accumulation of blowup times.

However, one can specify the inverse time change Φ\Phi independent of the possible accumulation of blowup times. Following on the informal discussion above, this implicit specification of Φ(t)=σ\Phi(t)=\sigma as a solution νt=σλG(σ)\nu t=\sigma-\lambda G(\sigma) is best formulated in terms of the inverse time change Ψ\Psi featuring in (30) and (31). Then to circumvent possible caveats of accumulating blowup times, we leverage the fact that Ψ\Psi must be a nondecreasing function to write relation (18) as a fixed-point problem, whereby Ψ\Psi is defined as a running maximum function. The fixed-point nature of (18)\eqref{eq:Phi} naturally follows from the fact that the cumulative function G=G[Ψ]G=G[\Psi] also depends on Ψ\Psi via η=η[Φ]=η[Ψ]\eta=\eta[\Phi]=\eta[\Psi]. Specifically, we show in [22] that:

Proposition 2.3.

The inverse time-change Ψ\Psi parametrizes a dPMF dynamics if and only if it solves the fixed-point problem

σ0,Ψ(σ)=sup0ξσ(ξλG[Ψ](ξ))/ν.\displaystyle\forall\;\sigma\geq 0\,,\quad\Psi(\sigma)=\sup_{0\leq\xi\leq\sigma}\big{(}\xi-\lambda G[\Psi](\xi)\big{)}/\nu\,. (32)

The full formulation of the fixed-point problem given in Definition 1.3 follows directly from the above proposition together with the renewal-type equation (25) for the cumulative flux GG and Definitions 2.3 and 2.5 for the backward functions η\eta and ξ\xi. The fixed point problem 1.3 allows one to characterize the occurrence of a single blowup in the time changed-picture for generic initial conditions satisfying Assumption 1 and under full-blowup Assumption 3. This local result is stated as follows in [22]:

Theorem 2.6.

Under Assumption 1, there is a unique smooth solution Ψ\Psi to the fixed-point problem 1.3 up to the possibly infinite time

S1=inf{σ>0|Ψ(σ)0}=inf{σ>0|g[Ψ](σ)1/λ}>0.\displaystyle S_{1}=\inf\left\{\sigma>0\,\big{|}\,\Psi^{\prime}(\sigma)\leq 0\right\}=\inf\left\{\sigma>0\,\big{|}\,g[\Psi](\sigma)\geq 1/\lambda\right\}>0\,. (33)
Theorem 2.7.

Under Assumption 1 and 3, the solution Ψ\Psi to the fixed-point problem 1.3 can be uniquely extended as a constant function on [S1,U1][S_{1},U_{1}] with U1=S1+λπ1U_{1}=S_{1}+\lambda\pi_{1} where π1\pi_{1} satisfies 0<π1<q(S1,)110<\pi_{1}<\|q(S_{1},\cdot)\|_{1}\leq 1 and is defined as

π1=inf{p0|p>0H(λp,x)q(S1,x)dx}.\displaystyle\pi_{1}=\inf\left\{p\geq 0\,\bigg{|}\,p>\int_{0}^{\infty}H(\lambda p,x)q(S_{1},x)\,\mathrm{d}x\right\}\,. (34)

We show our two main results in Section 3 and in Section 4. In preparation of these, let us state a few properties that solutions to the fixed-point problem 1.3 must satisfy. The first such property shows that for any solutions Ψ\Psi, the associated backward-delay functions η[Ψ]\eta[\Psi] must be uniformly bounded.

Proposition 2.4.

For all solutions Ψ\Psi to the fixed-point problem 1.3, the associated backward-delay functions η\eta are bounded by λ+νϵ\lambda+\nu\epsilon.

Proof.

Solutions Ψ\Psi to the fixed-point problem 1.3 are necessarily continuous by smoothness of the associated cumulative function G[Ψ]G[\Psi]. Moreover by definition of the backward function ξ\xi, such solutions must satisfy Ψ(ξ(σ))=Ψ(σ)ϵ\Psi(\xi(\sigma))=\Psi(\sigma)-\epsilon. This implies that

νϵ=ν(Ψ(σ)Ψ(ξ(σ)))=σξ(σ)λ(G(σ)G(ξ(σ))).\displaystyle\nu\epsilon=\nu\big{(}\Psi(\sigma)-\Psi(\xi(\sigma))\big{)}=\sigma-\xi(\sigma)-\lambda\big{(}G(\sigma)-G(\xi(\sigma))\big{)}\,.

We conclude by remembering that η=Idξ\eta=\mathrm{Id}-\xi and that by conservation of probability, we must have G(σ)G(ξ(σ))1G(\sigma)-G(\xi(\sigma))\leq 1. ∎

The second property states that global inverse time change Ψ\Psi provide us with global time changes Φ=Ψ1:++\Phi=\Psi^{-1}:\mathbbm{R}^{+}\to\mathbbm{R}^{+} for the dPMF dynamics. Suppose that the fixed-point problem 1.3 admits an inverse time change Ψ\Psi as a solution. Such a solution cannot explode in finite time as it is nondecreasing with bounded difference quotient: wΨ1/νw_{\Psi}\leq 1/\nu. Let us further assume that it is a global solution, i.e., that Ψ\Psi is defined on the whole set +\mathbbm{R}^{+}. Then, by boundedness of the backward-delay function η=Idξ\eta=\mathrm{Id}-\xi, it must be that S=limσξ(σ)=S_{\infty}=\lim_{\sigma\to\infty}\xi(\sigma)=\infty. For finite refractory period ϵ>0\epsilon>0, it turns out that this observation is enough to show the function Φ=Ψ1\Phi=\Psi^{-1} is also a global time change, as stated in the following proposition.

Proposition 2.5.

For ϵ>0\epsilon>0, solutions to the fixed-point problem 1.3 that are defined over +\mathbbm{R}^{+} necessarily satisfy T=limσΨ(σ)=T_{\infty}=\lim_{\sigma\to\infty}\Psi(\sigma)=\infty.

Proof.

Given a solution Ψ:++\Psi:\mathbbm{R}^{+}\to\mathbbm{R}^{+} to the fixed-point problem 1.3, suppose that T=limσΨ(σ)<T_{\infty}=\lim_{\sigma\to\infty}\Psi(\sigma)<\infty. Then, as Ψ\Psi and Φ=Ψ1\Phi=\Psi^{-1} are nondecreasing, for all ϵ>0\epsilon>0

S=limσξ(σ)=limσΦ(Ψ(σ)ϵ)=limtTΦ(tϵ)Φ(Tϵ)<.\displaystyle S_{\infty}=\lim_{\sigma\to\infty}\xi(\sigma)=\lim_{\sigma\to\infty}\Phi(\Psi(\sigma)-\epsilon)=\lim_{t\to T_{\infty}^{-}}\Phi(t-\epsilon)\leq\Phi(T_{\infty}^{-}-\epsilon)<\infty\,.

This contradicts the fact that η=Idξ\eta=\mathrm{Id}-\xi is uniformly bounded on +\mathbbm{R}^{+}. Thus we must have T=T_{\infty}=\infty. ∎

Note that the above result only holds for nonzero refractory period ϵ>0\epsilon>0. Establishing a similar result for vanishing refractory period ϵ=0\epsilon=0 will require a detailed analysis, which we will conduct only for large enough interaction parameters λ>Λ\lambda>\Lambda.

3 Global blowup solutions

In this section, we establish the existence and uniqueness of explosive dPMF dynamics over the whole half-line +\mathbbm{R}^{+} in the large interaction regime. First, we introduce useful a prori estimates about the fluxes associated to the fixed-point solutions Ψ\Psi for a natural class of initial conditions. Second, we utilize these estimates to show that in the large interaction regime, a large enough blowup is followed by a subsequent blowup in a well ordered fashion, in the sense that this next blowup can only happen after all the processes that previously blew up have reset. Third, we show that the size of the next blowup is lower bounded away from zero, which allows one to establish the existence and uniqueness of explosive dPMF dynamics over the the whole half-line +\mathbbm{R}^{+} by induction on the number of blowup episodes.

3.1 A priori bounds for natural initial conditions

Given normalized initial conditions (q0,g0)(q_{0},g_{0}) in ((0,))×([ξ0,0))\mathcal{M}((0,\infty))\times\mathcal{M}([\xi_{0},0)), assume that the fixed-point problem 1.3 admits a solution Ψ\Psi up to some possibly infinite time SS_{\infty}. Then, the associated instantaneous flux function gg satisfies the renewal-type equation (26) on [0,S)[0,S_{\infty}). We aim at utilizing this characterization to show that for a restricted class of initial conditions, gg is uniformly bounded on [0,S)[0,S_{\infty}), with an upper bound that is independent of the coupling parameter λ\lambda and that scales with the L1L_{1} norms of the initial conditions. Informally, the restricted initial conditions at stake are these distributions of time-changed processes that have been reset at some time in the past. We will refer to this restricted set of initial conditions as natural initial conditions, since they turn out to have the same regularities as generic dPMF solutions. We formally define natural initial conditions as follows:

Definition 3.1.

Natural initial conditions (q0,g0)(q_{0},g_{0}) in ((0,))×([ξ0,0))\mathcal{M}((0,\infty))\times\mathcal{M}([\xi_{0},0)) are those for which there exists a measure μ0\mu_{0} in ((0,))\mathcal{M}((0,\infty)) such that:

(i)(i) The measure q0q_{0} corresponds to the density

q0(x)=0κ(τ,x,Λ)1H(τ,Λ)μ0(dτ),\displaystyle q_{0}(x)=\int_{0}^{\infty}\frac{\kappa(\tau,x,\Lambda)}{1-H(\tau,\Lambda)}\,\mu_{0}(\mathrm{d}\tau)\,,

(ii)(ii) The measure g0g_{0} corresponds to the density

g0(σ)=0h(σ+τ,Λ)1H(τ,Λ)μ0(dτ).\displaystyle g_{0}(\sigma)=\int_{0}^{\infty}\frac{h(\sigma+\tau,\Lambda)}{1-H(\tau,\Lambda)}\,\mu_{0}(\mathrm{d}\tau)\,.

(iii)(iii) We have the degenerate normalization condition

0μ0(dτ)+ξ00g0(τ)dτ1.\displaystyle\int_{0}^{\infty}\mu_{0}(\mathrm{d}\tau)+\int_{\xi_{0}}^{0}g_{0}(\tau)\,\mathrm{d}\tau\leq 1\,.

We derive the above formal definition for natural initial conditions by considering absorbed, drifted Wiener dynamics that have started at Λ\Lambda at some time in the past with cumulative rate function RR. Such dynamics admit density functions under the integral form

q(σ,x)=0κ(σξ,x,Λ)dR(ξ),\displaystyle q(\sigma,x)=\int_{-\infty}^{0}\kappa(\sigma-\xi,x,\Lambda)\,\mathrm{d}R(\xi)\,,

where integration by part guarantees convergence for all cumulative rate functions with, say, polynomial growth. Note that we only requires RR to be a nondecreasing càdlàg function for being a cumulative rate function. The instantaneous flux function associated to this density is

g(σ)=xq(σ,0)/2=0h(σξ,Λ)dR(ξ).\displaystyle g(\sigma)=\partial_{x}q(\sigma,0)/2=\int_{-\infty}^{0}h(\sigma-\xi,\Lambda)\,\mathrm{d}R(\xi)\,.

From there, the parametrization of natural initial conditions is recovered by setting

μ0(dτ)=dR(τ)(1H(τ,Λ)),\displaystyle\mu_{0}(\mathrm{d}\tau)=\mathrm{d}R(-\tau)(1-H(\tau,\Lambda))\,,

and recognizing that q0(x)=q(0,x)q_{0}(x)=q(0,x) for x>0x>0 and g0(σ)=g(σ)g_{0}(\sigma)=g(\sigma), ξ0σ<0-\xi_{0}\leq\sigma<0. Bearing the above remarks in mind, we have the following interpretation for the parametrizing measure μ0\mu_{0}: given a fraction of active processes distributed according to q0q_{0} at time σ=0\sigma=0, μ0\mu_{0} represents the distribution of times elapsed since their last resets. In particular, definition (i)(i) implies that 0μ0(τ)dτ=0q0(x)dx\int_{0}^{\infty}\mu_{0}(\tau)\,\mathrm{d}\tau=\int_{0}^{\infty}q_{0}(x)\,\mathrm{d}x. Furthermore, by definition (i)(i) and (ii)(ii) and the remarks made above, we have

limσ0g0(σ)=σq0(0)/2,\displaystyle\lim_{\sigma\to 0^{-}}g_{0}(\sigma)=\partial_{\sigma}q_{0}(0)/2\,,

so that conservation of probability holds as in (7). The above observation directly shows that given any natural initial conditions (q0,g0)(q_{0},g_{0}) in ((0,))×([ξ0,0))\mathcal{M}((0,\infty))\times\mathcal{M}([\xi_{0},0)), the current state of a solution dPMF dynamics specifies natural initial conditions

(q(σ,),{g(ξσ)}η(σ)ξ<0)((0,))×([η(σ),0)).\displaystyle\big{(}q(\sigma,\cdot),\{g(\xi-\sigma)\}_{-\eta(\sigma)\leq\xi<0}\big{)}\quad\in\quad\mathcal{M}((0,\infty))\times\mathcal{M}([-\eta(\sigma),0))\,.

In other word, natural initial conditions are stabilized by dPMF dynamics. Finally, note that we do not require normalization to one in (iii)(iii) so that natural initial conditions can represent a smaller-than-one fraction of the processes. This latter point will be useful to discuss the following a priori estimates for the instantaneous flux associated to solutions of the fixed-point equation 1.3:

Proposition 3.1.

For natural initial conditions (q0,g0)(q_{0},g_{0}) in ((0,))×([ξ0,0))\mathcal{M}((0,\infty))\times\mathcal{M}([\xi_{0},0)), we have

g[0,)AΛ(0q0(x)dx+ξ00g0(σ)dσ)=AΛ(q0,g0)1.\displaystyle\|g\|_{[0,\infty)}\leq A_{\Lambda}\,\left(\int_{0}^{\infty}q_{0}(x)\,\mathrm{d}x+\int_{\xi_{0}}^{0}g_{0}(\sigma)\,\mathrm{d}\sigma\right)=A_{\Lambda}\,\|(q_{0},g_{0})\|_{1}\,.

for a constant AΛA_{\Lambda} that only depends on Λ\Lambda.

Proof.

Considering natural initial conditions allows one to utilize the Markov property of the first-passage kernels κ\kappa to write:

0h(σ,x)q0(x)dx\displaystyle\int_{0}^{\infty}h(\sigma,x)q_{0}(x)\,\mathrm{d}x =\displaystyle= 012yκ(σ,0,x)q0(x)dx,\displaystyle\int_{0}^{\infty}\frac{1}{2}\partial_{y}\kappa(\sigma,0,x)q_{0}(x)\,\mathrm{d}x\,,
=\displaystyle= 0(012yκ(σ,0,x)κ(τ,x,Λ)dx)μ0(τ)dτ1H(τ,Λ),\displaystyle\int_{0}^{\infty}\left(\int_{0}^{\infty}\frac{1}{2}\partial_{y}\kappa(\sigma,0,x)\kappa(\tau,x,\Lambda)\,\mathrm{d}x\right)\frac{\mu_{0}(\tau)\,\,\mathrm{d}\tau}{1-H(\tau,\Lambda)}\,,
=\displaystyle= 0(012yκ(τ+σ,0,Λ)dx)μ0(τ)dτ1H(τ,Λ),\displaystyle\int_{0}^{\infty}\left(\int_{0}^{\infty}\frac{1}{2}\partial_{y}\kappa(\tau+\sigma,0,\Lambda)\,\mathrm{d}x\right)\frac{\mu_{0}(\tau)\,\,\mathrm{d}\tau}{1-H(\tau,\Lambda)}\,,
=\displaystyle= 0h(τ+σ,Λ)1H(τ,Λ)μ0(τ)dτ.\displaystyle\int_{0}^{\infty}\frac{h(\tau+\sigma,\Lambda)}{1-H(\tau,\Lambda)}\mu_{0}(\tau)\,\mathrm{d}\tau\,.

Injecting the above in equation (26), we deduce the inequality

|g(σ)|\displaystyle|g(\sigma)| \displaystyle\>\leq\> h(+σ,Λ)1H(,Λ)0μ0(τ)dτ+\displaystyle\bigg{\|}\frac{h(\cdot+\sigma,\Lambda)}{1-H(\cdot,\Lambda)}\bigg{\|}_{\infty}\int_{0}^{\infty}\mu_{0}(\tau)\,\mathrm{d}\tau+
0σh(στ,Λ)dG(ξ(τ)),\displaystyle\quad\quad\bigg{\|}\int_{0}^{\sigma}h(\sigma-\tau,\Lambda)\,\mathrm{d}G(\xi(\tau))\bigg{\|}_{\infty}\,,
\displaystyle\>\leq\> h(+σ,Λ)1H(,Λ)0μ0(τ)dτ+\displaystyle\bigg{\|}\frac{h(\cdot+\sigma,\Lambda)}{1-H(\cdot,\Lambda)}\bigg{\|}_{\infty}\int_{0}^{\infty}\mu_{0}(\tau)\,\mathrm{d}\tau+
h(,Λ)1H(,Λ)0σ(1H(στ,Λ))dG(ξ(τ)),\displaystyle\quad\quad\bigg{\|}\frac{h(\cdot,\Lambda)}{1-H(\cdot,\Lambda)}\bigg{\|}_{\infty}\int_{0}^{\sigma}(1-H(\sigma-\tau,\Lambda))\,\mathrm{d}G(\xi(\tau))\,,

where both uniform norms are finite. Indeed, for all σ0\sigma\geq 0, L’Hospital rule yields that

limτh(τ+σ,Λ)1H(τ,Λ)=eσ/22,\displaystyle\lim_{\tau\to\infty}\frac{h(\tau+\sigma,\Lambda)}{1-H(\tau,\Lambda)}=\frac{e^{-\sigma/2}}{2}\,,

which shows that the infinity norms are finite at fixed σ\sigma. Moreover, exploiting the fact that hh is unimodal with maximum value hΛh^{\star}_{\Lambda} at σ=(3+9+4Λ)/2\sigma^{\star}=(-3+\sqrt{9+4\Lambda})/2 and that H(,Λ)H(\cdot,\Lambda) is increasing, we have for σσ\sigma\geq\sigma^{\star}

h(+σ,Λ)1H(,Λ)h(,Λ)1H(,Λ)=defM<,\displaystyle\bigg{\|}\frac{h(\cdot+\sigma,\Lambda)}{1-H(\cdot,\Lambda)}\bigg{\|}_{\infty}\leq\bigg{\|}\frac{h(\cdot,\Lambda)}{1-H(\cdot,\Lambda)}\bigg{\|}_{\infty}\stackrel{{\scriptstyle\mathrm{def}}}{{=}}M<\infty\,,

whereas for σ<σ\sigma<\sigma^{\star}, we have

h(+σ,Λ)1H(,Λ)h(,Λ)1H(σ,Λ)=hΛ1H(σ,Λ)=defN<.\displaystyle\bigg{\|}\frac{h(\cdot+\sigma,\Lambda)}{1-H(\cdot,\Lambda)}\bigg{\|}_{\infty}\leq\frac{\|h(\cdot,\Lambda)\|_{\infty}}{1-H(\sigma^{\star},\Lambda)}=\frac{h^{\star}_{\Lambda}}{1-H(\sigma^{\star},\Lambda)}\stackrel{{\scriptstyle\mathrm{def}}}{{=}}N<\infty\,.

Thus, we have

|g(σ)|\displaystyle|g(\sigma)| \displaystyle\leq max(M,N)0q0(x)dx+M0σ(1H(στ,Λ))dG(ξ(τ)),\displaystyle\max(M,N)\int_{0}^{\infty}q_{0}(x)\,\mathrm{d}x+M\int_{0}^{\sigma}(1-H(\sigma-\tau,\Lambda))\,\mathrm{d}G(\xi(\tau))\,, (35)

where the constant MM and NN only depends on Λ\Lambda. To bound the integral term in the inequality above, we make two observations: First, we write the renewal-type equation (25) for the cumulative flux GG under the form

0σH(στ,Λ))dG(ξ(τ))=G(σ)G(0)0H(σ,x)q0(x)dx.\displaystyle\int_{0}^{\sigma}H(\sigma-\tau,\Lambda))\,\mathrm{d}G(\xi(\tau))=G(\sigma)-G(0)-\int_{0}^{\infty}H(\sigma,x)q_{0}(x)\,\mathrm{d}x\,.

Second, we write the conservation of probability for processes originating from μ0\mu_{0} as

G(σ)G(ξ(σ))[G(0)G(ξ0)]=P(σ)P(0),\displaystyle G(\sigma)-G(\xi(\sigma))-\big{[}G(0)-G(\xi_{0})\big{]}=P(\sigma)-P(0)\,,

where PP denotes the fraction of processes originating from μ0\mu_{0} that is inactive. These two observations allows one to express the integral term in inequality (35) as

0σ(1H(στ,Λ))dG(ξ(τ))=P(0)P(σ)+0H(σ,x)q0(x)dx.\displaystyle\int_{0}^{\sigma}(1-H(\sigma-\tau,\Lambda))\,\mathrm{d}G(\xi(\tau))=P(0)-P(\sigma)+\int_{0}^{\infty}H(\sigma,x)q_{0}(x)\,\mathrm{d}x\,.

Therefore, as H1\|H\|_{\infty}\leq 1, we have

g[0,)\displaystyle\|g\|_{[0,\infty)} \displaystyle\leq max(M,N)0q0(x)dx+M(P(0)+0q0(x)dx).\displaystyle\max(M,N)\int_{0}^{\infty}q_{0}(x)\,\mathrm{d}x+M\left(P(0)+\int_{0}^{\infty}q_{0}(x)\,\mathrm{d}x\right)\,.

We conclude by observing that by conservation of probability mass P(0)(q0,g0)1P(0)\leq\|(q_{0},g_{0})\|_{1}, so that

g[0,)\displaystyle\|g\|_{[0,\infty)} \displaystyle\leq 3max(M,N)(q0,g0)1.\displaystyle 3\max(M,N)\|(q_{0},g_{0})\|_{1}\,.

The arguments of the above proof can be adapted to obtain a uniform bound to σg\partial_{\sigma}g, which consitutes another useful a priori estimate :

Proposition 3.2.

For natural initial conditions (q0,g0)(q_{0},g_{0}) in ((0,))×([ξ0,0))\mathcal{M}((0,\infty))\times\mathcal{M}([\xi_{0},0)), we have

σg[0,)BΛ(0q0(x)dx+ξ00g0(σ))=BΛ(q0,g0)1.\displaystyle\|\partial_{\sigma}g\|_{[0,\infty)}\leq B_{\Lambda}\,\left(\int_{0}^{\infty}q_{0}(x)\,\mathrm{d}x+\int_{\xi_{0}}^{0}g_{0}(\sigma)\right)=B_{\Lambda}\,\|(q_{0},g_{0})\|_{1}\,.

for a constant BΛB_{\Lambda} that only depends on Λ\Lambda.

Proof.

The proof proceeds in exactly the same fashion as the proof of Proposition (3.1) but starting from the renewal-type equation

σg(σ)=0σh(σ,x)q0(x)dx+0σσh(στ,Λ)dG(ξ(τ)),\displaystyle\partial_{\sigma}g(\sigma)=\int_{0}^{\infty}\partial_{\sigma}h(\sigma,x)q_{0}(x)\,\mathrm{d}x+\int_{0}^{\sigma}\partial_{\sigma}h(\sigma-\tau,\Lambda)\,\mathrm{d}G(\xi(\tau))\,,

where we observe that

limτ|σh(τ+σ,Λ)|1H(τ,Λ)=eσ/24.\displaystyle\lim_{\tau\to\infty}\frac{|\partial_{\sigma}h(\tau+\sigma,\Lambda)|}{1-H(\tau,\Lambda)}=\frac{e^{-\sigma/2}}{4}\,.

The above a priori bounds will be crucial to exhibit global solutions to the fixed-point problem 2.3 that exhibit an infinite number of blowups. Their main use will be in showing that for blowup of large enough size 0<π1<10<\pi_{1}<1, the small fraction 1π11-\pi_{1} of surviving active processes cannot trigger another blowup before the original fraction π1\pi_{1} has reset. In other words, this will guarantee that blowups are well ordered.

3.2 Well-ordered blowups for large interactions

Given natural initial conditions (q0,g0)(q_{0},g_{0}) in ((0,))×([ξ0,0))\mathcal{M}((0,\infty))\times\mathcal{M}([\xi_{0},0)), consider a solution to the fixed-point problem 1.3 with associated instantaneous flux gg. By Propositions 3.1 and 3.2, we know that the flux gg and its derivative σg\partial_{\sigma}g, are both uniformly bounded on the domain of the solution. Here, we aim at refining the bounding analysis of gg to ensure that for large enough synchronization events, blowups are well ordered in the sense that no blowup can occur before the reset of processes that synchronized in the past. Large synchronization events inactivate so many processes that the remaining processes are too few to trigger a blowup on their own.

To prove this, we distinguish between two contributions to the current state of the dynamics, according to whether processes have inactivated in the most recent blowup or not. Specifically, assume that a blowup of size π1\pi_{1} triggers at S10S_{1}\geq 0 with exit time U1U_{1} and reset time R1R_{1}. At time σR1\sigma\geq R_{1}, we can always split the current state of the dynamics into a contribution from the synchronized processes that reset at R1R_{1} (marked by a subscript π1\pi_{1}) and a contribution from the other processes (marked by a subscript π1\cancel{\pi_{1}}). This means that there are (qπ1,gπ1)(q_{\pi_{1}},g_{\pi_{1}}) and (qπ1,gπ1)(q_{\cancel{\pi_{1}}},g_{\cancel{\pi_{1}}}) in ((0,))×([ξ(σ),σ))\mathcal{M}((0,\infty))\times\mathcal{M}([\xi(\sigma),\sigma)) such that for all σR1\sigma\geq R_{1}

q(σ,)=qπ1(σ,)+qπ1(σ,)andg(σ)=gπ1(σ)+gπ1(σ),\displaystyle q(\sigma,\cdot)=q_{\pi_{1}}(\sigma,\cdot)+q_{\cancel{\pi_{1}}}(\sigma,\cdot)\quad\mathrm{and}\quad g(\sigma)=g_{\pi_{1}}(\sigma)+g_{\cancel{\pi_{1}}}(\sigma)\,, (36)

with qπ1(R1,)=π1δΛq_{\pi_{1}}(R_{1},\cdot)=\pi_{1}\delta_{\Lambda} and gπ1(σ)=0g_{\pi_{1}}(\sigma)=0 for σR1\sigma\leq R_{1} and with qπ1(R1,)=q(R1,)q_{\cancel{\pi_{1}}}(R_{1},\cdot)=q(R_{1}^{-},\cdot) and gπ1(σ)=g(σ)g_{\cancel{\pi_{1}}}(\sigma)=g(\sigma) for U1σ<R1U_{1}\leq\sigma<R_{1}. Given a fixed backward function ξ\xi, it is clear that the governing renewal-type equation (26) applies to both contribution (qπ1,gπ1)(q_{\pi_{1}},g_{\pi_{1}}) and (qπ1,gπ1)(q_{\cancel{\pi_{1}}},g_{\cancel{\pi_{1}}}) separately. As a result, Propositions 3.1 and 3.2 also apply to both contributions separately, allowing to control the role of each contribution in triggering the next possible blowup.

The next proposition states this point concisely by making use of the real numbers σ\sigma^{\dagger} and hh^{\dagger} defined by

σ=inf{σ>0|σ2h(σ,Λ)=0}andhΛ=h(σ,Λ).\displaystyle\sigma^{\dagger}=\inf\{\sigma>0\,|\,\partial^{2}_{\sigma}h(\sigma,\Lambda)=0\}\quad\mathrm{and}\quad h^{\dagger}_{\Lambda}=h(\sigma^{\dagger},\Lambda)\,.

Observe that σ\sigma^{\dagger} and hΛh^{\dagger}_{\Lambda} only depends on Λ\Lambda, and that σ\sigma^{\dagger} satisfies 0<σ<σ0<\sigma^{\dagger}<\sigma^{\star}, where σ\sigma^{\star} is the unique maximizer of h(,Λ)h(\cdot,\Lambda): h(σ,Λ)=hΛ>hΛh(\sigma^{\star},\Lambda)=h^{\star}_{\Lambda}>h^{\dagger}_{\Lambda}. Moreover, by definition of σ\sigma^{\dagger}, observe that σh(.,Λ)\partial_{\sigma}h(.,\Lambda) is increasing on [0,σ][0,\sigma^{\dagger}]. Then, equipped with hΛh^{\star}_{\Lambda} and σ\sigma^{\dagger}, the following proposition states that for large enough interaction parameters λ\lambda, a sufficiently large blowup π1\pi_{1} will trigger a next blowup in a well ordered fashion.

Proposition 3.3.

Given an interaction parameter λ\lambda with λ>lΛ=1/hΛ\lambda>l_{\Lambda}=1/h^{\dagger}_{\Lambda}, there exists a constant CΛC_{\Lambda} that only depends on Λ\Lambda such that a full blowup of size π1\pi_{1} with

π1>max(lΛλ,1CΛλ),\displaystyle\pi_{1}>\max\left(\frac{l_{\Lambda}}{\lambda},1-\frac{C_{\Lambda}}{\lambda}\right)\,, (37)

is followed by another full blowup in finite time. Moreover, if the first blowup happens at time S1S_{1}, the next blowup time S2S_{2} is such that S1R1<S2R1+σS_{1}\leq R_{1}<S_{2}\leq R_{1}+\sigma^{\dagger}, where we have defined the reset time R1=τ(U1)R_{1}=\tau(U_{1}).

Proof.

The proof proceeds in four steps: (i)(i) we show that a blowup cannot happen before the fraction π1\pi_{1} of processes reset, (ii)(ii) we show that the fraction π1\pi_{1} of reset processes is enough to trigger a blowup, (iii)(iii) we show the full blowup condition, and (iv)(iv) we conclude by specifying the choice of lΛl_{\Lambda} and CΛC_{\Lambda}.

(i)(i) Suppose that the last blowup triggers at time S1S_{1} with blowup size π1\pi_{1}, then at blowup exit time U1=S1+λπ1U_{1}=S_{1}+\lambda\pi_{1}, the exit states

(q(U1,),{g(σ)}ξ(U1)σ<U1)\displaystyle\big{(}q(U_{1},\cdot),\{g(\sigma)\}_{\xi(U_{1})\leq\sigma<U_{1}}\big{)}

determine natural initial conditions. These are such that G(U1)G(S1)=S1U1g(σ)dσ=π1G(U_{1})-G(S_{1})=\int_{S_{1}}^{U_{1}}g(\sigma)\,\mathrm{d}\sigma=\pi_{1}, representing the fraction of processes to be instantaneously reset at time R1=τ(S1)R_{1}=\tau(S_{1}). Let us consider the part of the exit initial conditions excluding this fraction, i.e., the active processes that either survive the blowup or that are already inactive before the blowup:

(qπ1,0,gπ1,0)=(q(U1,),{g(σ)}ξ(U1)σ<S1),with(qπ1,0,gπ1,0)11π1.\displaystyle(q_{\cancel{\pi_{1}},0},g_{\cancel{\pi_{1}},0})=\big{(}q(U_{1},\cdot),\{g(\sigma)\}_{\xi(U_{1})\leq\sigma<S_{1}}\big{)}\,,\quad\mathrm{with}\quad\|(q_{\cancel{\pi_{1}},0},g_{\cancel{\pi_{1}},0})\|_{1}\leq 1-\pi_{1}\,.

Choosing CΛ<1/(4AΛ)C_{\Lambda}<1/(4A_{\Lambda}) in (37) implies that 1π1<1/(4AΛλ)1-\pi_{1}<1/(4A_{\Lambda}\lambda) so that by Proposition 3.1, the instantaneous flux gπ1g_{\cancel{\pi_{1}}} due to the linear dynamics of the processes arising from the partial initial conditions (qπ1,0,gπ1,0)(q_{\cancel{\pi_{1}},0},g_{\cancel{\pi_{1}},0}) is bounded above with gπ1[U1,)1/(4λ)\|g_{\cancel{\pi_{1}}}\|_{[U_{1},\infty)}\leq 1/(4\lambda). This implies that no blowup can occur before resetting the fraction of inactivated processes π1\pi_{1} as the blowup condition (g(σ)=1/λg(\sigma)=1/\lambda) cannot be met. Moreover, before the reset time R1R_{1}, the inverse time change Ψ\Psi only depends on the partial partial initial conditions (qπ1,0,gπ1,0)(q_{\cancel{\pi_{1}},0},g_{\cancel{\pi_{1}},0}) so that we have

ϵ\displaystyle\epsilon =\displaystyle= Ψ(R1)Ψ(U1),\displaystyle\Psi(R_{1})-\Psi(U_{1})\,,
=\displaystyle= (R1U1λ(G(R1)G(U1))/ν,\displaystyle\big{(}R_{1}-U_{1}-\lambda(G(R_{1})-G(U_{1})\big{)}/\nu\,,
\displaystyle\geq (1λgπ1[U1,))(R1U1)/ν,\displaystyle(1-\lambda\|g_{\cancel{\pi_{1}}}\|_{[U_{1},\infty)})(R_{1}-U_{1})/\nu\,,
\displaystyle\geq (R1U1)/(2ν).\displaystyle(R_{1}-U_{1})/(2\nu)\,.

Thus the reset of the processes inactivated during the last blowup must happen in finite time at R1R_{1} with

νϵ<R1U1<2νϵ.\displaystyle\nu\epsilon<R_{1}-U_{1}<2\nu\epsilon\,. (38)

(ii)(ii) We denote by gπ1g_{\pi_{1}} the instantaneous flux due the fraction π1\pi_{1} of the processes that reset at forward time R1R_{1}. For all σR1\sigma\geq R_{1}, this flux satisfies

π1h(σR1,Λ)gπ1(σ)π1n=0h(σR1,nΛ)=π1g~(σR1,Λ),\displaystyle\pi_{1}h(\sigma-R_{1},\Lambda)\leq g_{\pi_{1}}(\sigma)\leq\pi_{1}\sum_{n=0}^{\infty}h(\sigma-R_{1},n\Lambda)=\pi_{1}\tilde{g}(\sigma-R_{1},\Lambda)\,, (39)

where left term is the flux due to the fraction π1\pi_{1} excluding any reset after R1R_{1}, whereas the right term is the flux due to the fraction π1\pi_{1} with instantaneous reset after R1R_{1}. Remembering that by definition hΛ=supσ0h(σ,Λ)>hΛh^{\star}_{\Lambda}=\sup_{\sigma\geq 0}h(\sigma,\Lambda)>h^{\dagger}_{\Lambda}, choosing λ>1/(hΛπ1)\lambda>1/(h^{\dagger}_{\Lambda}\pi_{1}) implies that the blowup condition g(σ)1/λg(\sigma)\geq 1/\lambda will be met at some time S2>R1S_{2}>R_{1}. The upper bound flux term g~\tilde{g} in inequality (39) only depends on Λ\Lambda and is uniformly bounded with respect σ\sigma in +\mathbbm{R}^{+}. Moreover, g~\tilde{g} satisfies the renewal equation

g~(σ,Λ)=h(σ,Λ)+0σh(στ,Λ)g~(τ,Λ)dτ,\displaystyle\tilde{g}(\sigma,\Lambda)=h(\sigma,\Lambda)+\int^{\sigma}_{0}h(\sigma-\tau,\Lambda)\tilde{g}(\tau,\Lambda)\,\mathrm{d}\tau\,,

so that for all 0σσ0\leq\sigma\leq\sigma^{\star}, remembering that g~AΛ\|\tilde{g}\|_{\infty}\leq A_{\Lambda}, we have the upper bound

g~(σ,Λ)h(σ,Λ)+g~0σh(στ,Λ)dτh(σ,Λ)(1+AΛσ),\displaystyle\tilde{g}(\sigma,\Lambda)\leq h(\sigma,\Lambda)+\|\tilde{g}\|_{\infty}\int^{\sigma}_{0}h(\sigma-\tau,\Lambda)\,\mathrm{d}\tau\leq h(\sigma,\Lambda)(1+A_{\Lambda}\sigma^{\star})\,,

where the last inequality follows from the fact that h(,Λ)h(\cdot,\Lambda) is increasing on [0,σ][0,\sigma^{\star}]. Therefore, defining aΛ=(1+AΛσ)1/2<1/2a_{\Lambda}=(1+A_{\Lambda}\sigma^{\star})^{-1}/2<1/2, we have

s1<S2R1<s2,\displaystyle s_{1}<S_{2}-R_{1}<s_{2}\,,

where the delay times s1s_{1} and s2s_{2} only depend on Λ\Lambda and π1λ\pi_{1}\lambda via:

s1=inf{s>0|h(s,Λ)=aΛπ1λ}ands2=inf{s>0|h(s,Λ)=1π1λ}.\displaystyle\quad s_{1}=\inf\left\{s>0\,\bigg{|}\,h(s,\Lambda)=\frac{a_{\Lambda}}{\pi_{1}\lambda}\right\}\;\;\mathrm{and}\;\;s_{2}=\inf\left\{s>0\,\bigg{|}\,h(s,\Lambda)=\frac{1}{\pi_{1}\lambda}\right\}\,. (40)

(iii)(iii) It remains to check the full-blowup condition, i.e., σg(S2)>0\partial_{\sigma}g(S_{2})>0. In this perspective, let us first establish a lower bound for σgπ1(S2)\partial_{\sigma}g_{\pi_{1}}(S_{2}). Such a bound is obtained by differentiating (26) with respect to σ\sigma for R1σR1+σR_{1}\leq\sigma\leq R_{1}+\sigma^{\dagger}, which yields

σgπ1(σ)=π1σh(σR1,Λ)+0σσh(στ,Λ)gπ1(ξ(τ))π1σh(σR1,Λ),\displaystyle\partial_{\sigma}g_{\pi_{1}}(\sigma)=\pi_{1}\partial_{\sigma}h(\sigma-R_{1},\Lambda)+\int^{\sigma}_{0}\partial_{\sigma}h(\sigma-\tau,\Lambda)\,g_{\pi_{1}}(\xi(\tau))\geq\pi_{1}\partial_{\sigma}h(\sigma-R_{1},\Lambda)\,,

where we utilize the fact that σh(,Λ)\partial_{\sigma}h(\cdot,\Lambda) is increasing on [0,σ][0,\sigma^{\dagger}]. Thus, under the condition that s2σs_{2}\leq\sigma^{\dagger}, we have

σgπ1(S2)π1σh(S2R1,Λ)π1σh(s1,Λ),\displaystyle\partial_{\sigma}g_{\pi_{1}}(S_{2})\geq\pi_{1}\partial_{\sigma}h(S_{2}-R_{1},\Lambda)\geq\pi_{1}\partial_{\sigma}h(s_{1},\Lambda)\,,

where s1s_{1} is defined by (40). Moreover, from the explicit expression of h(,Λ)h(\cdot,\Lambda), we evaluate

σh(s1,Λ)h(s1,Λ)=σlnh(s1,Λ)=Λ2(3s1+s12)2s12,\displaystyle\frac{\partial_{\sigma}h(s_{1},\Lambda)}{h(s_{1},\Lambda)}=\partial_{\sigma}\ln h(s_{1},\Lambda)=\frac{\Lambda^{2}-(3s_{1}+s_{1}^{2})}{2s_{1}^{2}}\,,

so that by the definition of s1s_{1} given in (40) we have

π1σh(s1,Λ)=Λ2(3s1+s12)2s121aΛλ.\displaystyle\pi_{1}\partial_{\sigma}h(s_{1},\Lambda)=\frac{\Lambda^{2}-(3s_{1}+s_{1}^{2})}{2s_{1}^{2}}\frac{1}{a_{\Lambda}\lambda}\,.

To show that the above quantity is nonnegative, it is enough to remember that under the condition that s2σs_{2}\leq\sigma^{\dagger}, we have s1<s2σ<σs_{1}<s_{2}\leq\sigma^{\dagger}<\sigma^{\star}. Moreover, the function s1(Λ2(3s1+s12))/(2s12)s_{1}\mapsto(\Lambda^{2}-(3s_{1}+s_{1}^{2}))/(2s_{1}^{2}) is decreasing and strictly positive on (0,σ)(0,\sigma^{\star}) since it has the same sign as σh(,Λ)\partial_{\sigma}h(\cdot,\Lambda). Thus, we have

σgπ1(S2)π1σh(s1,Λ)bΛλ,withbλ=Λ2(3σ+σ2)2σ2aΛ>0,\displaystyle\partial_{\sigma}g_{\pi_{1}}(S_{2})\geq\pi_{1}\partial_{\sigma}h(s_{1},\Lambda)\geq\frac{b_{\Lambda}}{\lambda}\,,\quad\mathrm{with}\quad b_{\lambda}=\frac{\Lambda^{2}-\left(3{\sigma^{\dagger}}+\sigma^{\dagger 2}\right)}{2\sigma^{\dagger 2}a_{\Lambda}}>0\,,

where bΛb_{\Lambda} only depends on Λ\Lambda. Choosing CΛ<bΛ/(2BΛ)C_{\Lambda}<b_{\Lambda}/(2B_{\Lambda}) in (37) implies that 1π1<bΛ/(2BΛλ)1-\pi_{1}<b_{\Lambda}/(2B_{\Lambda}\lambda) so that by Proposition 3.2, the absolute value of the term |σgπ1||\partial_{\sigma}g_{\cancel{\pi_{1}}}| due to the linear dynamics of the processes arising from the partial initial conditions (qπ1,0,gπ1,0)(q_{\cancel{\pi_{1},0}},g_{\cancel{\pi_{1},0}}) is bounded above with σgπ1[U1,)bΛ/(2λ)\|\partial_{\sigma}g_{\cancel{\pi_{1}}}\|_{[U_{1},\infty)}\leq b_{\Lambda}/(2\lambda). This shows that a full blowup happens in S2S_{2} since

σg(S2)=σgπ1(S2)+σgπ1(S2)σgπ1(S2)|σgπ1(S2)|bΛ2λ>0.\displaystyle\partial_{\sigma}g(S_{2})=\partial_{\sigma}g_{\pi_{1}}(S_{2})+\partial_{\sigma}g_{\cancel{\pi_{1}}}(S_{2})\geq\partial_{\sigma}g_{\pi_{1}}(S_{2})-|\partial_{\sigma}g_{\cancel{\pi_{1}}}(S_{2})|\geq\frac{b_{\Lambda}}{2\lambda}>0\,. (41)

(iv)(iv) The blowup condition due to reset processes alone requires in (i)(i) that CΛ<1/(4AΛ)C_{\Lambda}<1/(4A_{\Lambda}) and in (ii)(ii) that π1h>1/λ\pi_{1}h^{*}>1/\lambda. The full blowup condition requires in (iii)(iii) that s2σs_{2}\leq\sigma^{\dagger}, which is equivalent to π1h>1/λ\pi_{1}h^{\dagger}>1/\lambda, and that CΛ<bΛ/(2BΛ)C_{\Lambda}<b_{\Lambda}/(2B_{\Lambda}). This shows that choosing

lΛ>1hΛand0<CΛ<12min(12AΛ,bΛBΛ),\displaystyle l_{\Lambda}>\frac{1}{h^{\dagger}_{\Lambda}}\quad\mathrm{and}\quad 0<C_{\Lambda}<\frac{1}{2}\min\left(\frac{1}{2A_{\Lambda}},\frac{b_{\Lambda}}{B_{\Lambda}}\right)\,, (42)

suffices to ensure that after a full blowup of size π1\pi_{1} at time S1S_{1} satisfying (37), the next blowup must happen in a well ordered fashion at time S2S_{2} such that S1R1<S2R1+σS_{1}\leq R_{1}<S_{2}\leq R_{1}+\sigma^{\dagger}. ∎

3.3 Persistence of blowups for large interactions

Proposition 3.3 exhibits a criterion for a large enough full blowup to be followed by another full blowup in a well ordered fashion. By well ordered, we mean that the next blowup must occur after the processes that last blew up have all reset. In this context, for blowups to occur indefinitely in a sustained fashion, we further need to check that the size of the next blowup remains larger than that of the criterion of Proposition 3.3. In this perspective, it is instructive to find an upper bound to the fraction of inactive processes at the trigger time of the next blowup S2S_{2}. Indeed, processes that are inactive at S2S_{2} will remain so throughout the blowup, and therefore cannot contribute to the next blowup of size π2\pi_{2}. The next proposition shows that the fraction of inactive process P(S2)P(S_{2}) can be made arbitrarily small for large enough λ\lambda. This proposition only considers interaction parameter values λ>max(2lΛ,2/CΛ)\lambda>\max(2l_{\Lambda},2/C_{\Lambda}), so that the constraint on the full blowup size in Proposition 3.3 reduces to

π1>1CΛλ>1/2.\displaystyle\pi_{1}>1-\frac{C_{\Lambda}}{\lambda}>1/2\,. (43)
Proposition 3.4.

Consider an interaction parameter λ\lambda such that λ>max(2lΛ,2/CΛ)\lambda>\max(2l_{\Lambda},2/C_{\Lambda}) and a dynamics for which a full blowup happens at time S1S_{1} with size π1>1CΛ/λ>1/2\pi_{1}>1-C_{\Lambda}/\lambda>1/2. If λ\lambda further satisfies that λ>22πΛeΛ2/2CΛ\lambda>2\sqrt{2\pi\Lambda}e^{\Lambda^{2}/2C_{\Lambda}} and if the refractory period is such that ϵ<CΛ/(2ν)\epsilon<C_{\Lambda}/(2\nu), then the next full blowup triggers at time S2<S_{2}<\infty with a fraction of inactive processes:

P(S2)CΛ2λ.\displaystyle P(S_{2})\leq\frac{C_{\Lambda}}{2\lambda}\,.
Proof.

Let us first consider s2s_{2}, the upper bound to S2R1S_{2}-R_{1} defined by (40). Noticing that h(Λ,Λ)=1/(2πΛ)h(\Lambda,\Lambda)=1/(\sqrt{2\pi\Lambda}), by definition of s2s_{2}, if λ>2πΛ/π1\lambda>\sqrt{2\pi\Lambda}/\pi_{1}, then we necessarily have s2<Λs_{2}<\Lambda. Moreover, if s2<Λs_{2}<\Lambda, we have

1π1λ=h(s2,Λ)=Λ2πs23e(Λs2)22s2>eΛ22s22πΛ,\displaystyle\quad\frac{1}{\pi_{1}\lambda}=h(s_{2},\Lambda)=\frac{\Lambda}{\sqrt{2\pi s_{2}^{3}}}\,e^{-\frac{(\Lambda-s_{2})^{2}}{2s_{2}}}>\frac{e^{-\frac{\Lambda^{2}}{2s_{2}}}}{\sqrt{2\pi\Lambda}}\,,

which together with π1>1/2\pi_{1}>1/2, implies the following upper bound on s2s_{2}:

s2<Λ22λln(π1λ/2πΛ)<Λ22λln(λ/(22πΛ)).\displaystyle s_{2}<\frac{\Lambda^{2}}{2\lambda\ln\left(\pi_{1}\lambda/\sqrt{2\pi\Lambda}\right)}<\frac{\Lambda^{2}}{2\lambda\ln\left(\lambda/\big{(}2\sqrt{2\pi\Lambda}\big{)}\right)}\,.

Therefore the cumulative flux of inactivated processes between the last blowup exit time U1U_{1} and the next blowup trigger time S2S_{2} satisfies

G(S2)G(U1)\displaystyle G(S_{2})-G(U_{1}) \displaystyle\leq gU1,R1(R1U1)+gR1,S2(S2R1),\displaystyle\|g\|_{U_{1},R_{1}}(R_{1}-U_{1})+\|g\|_{R_{1},S_{2}}(S_{2}-R_{1})\,,
\displaystyle\leq νϵ2λ+Λ22λln(λ/(22πΛ)),\displaystyle\frac{\nu\epsilon}{2\lambda}+\frac{\Lambda^{2}}{2\lambda\ln\left(\lambda/\big{(}2\sqrt{2\pi\Lambda}\big{)}\right)}\,,

where we utilize that gU1,R1<1/(4λ)\|g\|_{U_{1},R_{1}}<1/(4\lambda) and R1U1<2νϵR_{1}-U_{1}<2\nu\epsilon by (38). Then for all ϵ<CΛ/(2ν)\epsilon<C_{\Lambda}/(2\nu), assuming that

λ>22πΛe2Λ2/CΛ,\displaystyle\lambda>2\sqrt{2\pi\Lambda}e^{2\Lambda^{2}/C_{\Lambda}}\,,

yields an upper bound on the fraction of inactive processes at the next blowup time S2S_{2}:

P(S2)=10q(S2,x)𝑑xG(S2)G(U1)CΛ2λ.\displaystyle P(S_{2})=1-\int_{0}^{\infty}q(S_{2},x)\,dx\leq G(S_{2})-G(U_{1})\leq\frac{C_{\Lambda}}{2\lambda}\,.

We are now in a position to show that for large enough interaction parameter λ\lambda, the next blowup will have a size π2\pi_{2} that satisfies the criterion of Proposition 3.3. This is a consequence of the following observations: For well ordered blowup, the reset of a fraction π1\pi_{1} of processes at time R1R_{1} is enough to trigger a blowup of finite size, which will be bounded below by a quantity ΔHΛ/2>0\Delta H_{\Lambda}/2>0 that only depends on Λ\Lambda. Then, the duration of the blowup during which reset is halted is lower bounded by λΔHΛ/2>0\lambda\Delta H_{\Lambda}/2>0. In turn, for large enough λ\lambda, the probability of inactivation during blowup can be made exponentially small with respect to λ\lambda by realizing that inactivation has an asymptotically constant, nonzero hazard rate function. Altogether, these observations allow us to form a self-consistent condition guaranteeing that π2\pi_{2} satisfies the criterion of Proposition 3.3, leading to the following result:

Proposition 3.5.

Assume that the refractory period satisfies ϵ<CΛ/ν\epsilon<C_{\Lambda}/\nu. There exists a constant LΛ>0L_{\Lambda}>0, which only depends on Λ\Lambda, such that for all interaction parameter λ>LΛ\lambda>L_{\Lambda}, every full blowup at time S1S_{1} with size π11CΛ/λ\pi_{1}\geq 1-C_{\Lambda}/\lambda is directly followed by a full blowup at time S2S_{2} with size π21CΛ/λ\pi_{2}\geq 1-C_{\Lambda}/\lambda.

Proof.

The proof proceeds in three steps: (i)(i) we exhibit the asymptotically exponential regime of processes survival during blowup, (ii)(ii) we show that this exponential regime is met for large enough interaction parameter λ\lambda, and (iii)(iii) we exhibit a sufficient constant LΛL_{\Lambda}.

(i)(i) We can always choose λ\lambda large enough so that

λmax(2lΛ,2/CΛ,22πΛeΛ2/2CΛ).\displaystyle\lambda\geq\max\big{(}2l_{\Lambda},2/C_{\Lambda},2\sqrt{2\pi\Lambda}e^{\Lambda^{2}/2C_{\Lambda}}\big{)}\,. (44)

Then by Proposition 3.3, the occurrence of a full blowup in S1S_{1} with size π11CΛ/λ\pi_{1}\geq 1-C_{\Lambda}/\lambda triggers the next full blowup at a time S2S_{2} with R1<S2R1+σR_{1}<S_{2}\leq R_{1}+\sigma^{\dagger}. By Proposition 3.4, the fraction of active process 1P(S2)1-P(S_{2}) at time S2S_{2} is such that 1P(S2)>1CΛ/21-P(S_{2})>1-C_{\Lambda}/2. Let us consider a process YσY_{\sigma} that is active at time S2S_{2} and let us denote by ξ\xi its next inactivation time. Then for all στS1\sigma\geq\tau\geq S_{1}, we have

[ξσ|ξτ]=[ξσ][ξτ]=1H(σ,Λ)1H(τ,Λ),\quad\mathbb{P}\left[\xi\geq\sigma\,\Big{|}\,\xi\geq\tau\right]=\frac{\mathbb{P}\left[\xi\geq\sigma\right]}{\mathbb{P}\left[\xi\geq\tau\right]}=\frac{1-H(\sigma,\Lambda)}{1-H(\tau,\Lambda)}\,,

where the survival probabilities 1H(σ,Λ)1-H(\sigma,\Lambda) can be written in term of a hazard rate as

1H(σ,Λ)=exp(0στln(1H(τ,Λ))dτ).\displaystyle 1-H(\sigma,\Lambda)=\exp{\left(\int_{0}^{\sigma}\partial_{\tau}\ln\left(1-H(\tau,\Lambda)\right)\,\mathrm{d}\tau\right)}\,.

The key observation is to notice that the hazard rate appearing above has the finite limit

limσσln(1H(σ,Λ))=limσh(σ,Λ)1H(σ,Λ)=12,\displaystyle\lim_{\sigma\to\infty}-\partial_{\sigma}\ln\left(1-H(\sigma,\Lambda)\right)=\lim_{\sigma\to\infty}\frac{h(\sigma,\Lambda)}{1-H(\sigma,\Lambda)}=\frac{1}{2}\,,

showing that for large time, inactivation asymptotically follows a memoryless exponential law for all active processes from the fraction 1P(S2)1-P(S_{2}).

(ii)(ii) We now exploit the asymptotic exponential behavior of the hazard function associated to 1H(,Λ)1-H(\cdot,\Lambda) to show that π2\pi_{2}, the size of the next full blowup triggered in S2S_{2}, can be made arbitrary close to 1P(S2)1-P(S_{2}). In this perspective, let us define the finite time

σ=sup{σσ|σln(1H(σ,Λ))<14}<.\displaystyle\sigma^{\sharp}=\sup\left\{\sigma\geq\sigma^{\star}\,\bigg{|}\,\partial_{\sigma}\ln\left(1-H(\sigma,\Lambda)\right)<-\frac{1}{4}\right\}<\infty\,.

For all σ>σ\sigma>\sigma^{\sharp}, integrating the hazard rate inequality σln(1H(σ,Λ))<1/4\partial_{\sigma}\ln\left(1-H(\sigma,\Lambda)\right)<-1/4 yields

[ξσ|ξσ]=1H(σ)1H(σ)eσσ4,\displaystyle\mathbb{P}\left[\xi\geq\sigma\,\Big{|}\,\xi\geq\sigma^{\sharp}\right]=\frac{1-H(\sigma)}{1-H(\sigma^{\sharp})}\leq e^{-\frac{\sigma-\sigma^{\sharp}}{4}}\,, (45)

showing that the survival probability of active processes decays at least exponentially past time S2+σS_{2}+\sigma^{\sharp}. Moreover, the fraction of processes ΔP2\Delta P_{2} that inactivates during the full blowup but before S2+σS_{2}+\sigma^{\sharp} satisfies

ΔP2=P(S2+σ)P(S2)π1σσh(σ,Λ)dσΔHΛ/2>0.\displaystyle\Delta P_{2}=P(S_{2}+\sigma^{\sharp})-P(S_{2})\geq\pi_{1}\int_{\sigma^{\dagger}}^{\sigma^{\star}}h(\sigma,\Lambda)\,\mathrm{d}\sigma\geq\Delta H_{\Lambda}/2>0\,.

where we utilize the assumption that π1>1CΛ/λ>1/2\pi_{1}>1-C_{\Lambda}/\lambda>1/2 and where we define the constant ΔHΛ=σσh(σ,Λ)dσ>0\Delta H_{\Lambda}=\int_{\sigma^{\dagger}}^{\sigma^{\star}}h(\sigma,\Lambda)\,\mathrm{d}\sigma>0, which only depends on Λ\Lambda. The above lower bound follows from the fact that we choose λ\lambda and π1\pi_{1} so that S2R1σ<σσS_{2}-R_{1}\leq\sigma^{\dagger}<\sigma^{\star}\leq\sigma^{\sharp}. By (45), the lower bound ΔHΛ/2ΔP2π2\Delta H_{\Lambda}/2\leq\Delta P_{2}\leq\pi_{2} implies that

[ξλπ2|ξσ]eλπ2σ4eλΔHΛ2σ8.\displaystyle\mathbb{P}\left[\xi\geq\lambda\pi_{2}\,\Big{|}\,\xi\geq\sigma^{\sharp}\right]\leq e^{-\frac{\lambda\pi_{2}-\sigma^{\sharp}}{4}}\leq e^{-\frac{\lambda\Delta H_{\Lambda}-2\sigma^{\sharp}}{8}}\,.

Thus, the fraction of inactivated process during blowup can be made arbitrarily close to 1P(S2)1-P(S_{2}) at the cost of choosing larger λ\lambda, as shown by:

π2\displaystyle\pi_{2} =\displaystyle= ΔP2+(1P(S2+σ))[ξλπ2|ξσ],\displaystyle\Delta P_{2}+\big{(}1-P(S_{2}+\sigma^{\sharp})\big{)}\mathbb{P}\left[\xi\leq\lambda\pi_{2}\,\Big{|}\,\xi\geq\sigma^{\sharp}\right]\,,
\displaystyle\geq ΔP2+(1P(S2+σ))(1eλΔHΛ2σ8),\displaystyle\Delta P_{2}+\big{(}1-P(S_{2}+\sigma^{\sharp})\big{)}\left(1-e^{-\frac{\lambda\Delta H_{\Lambda}-2\sigma^{\sharp}}{8}}\right)\,,
\displaystyle\geq 1P(S2)(1P(S2+σ))eλΔHΛ2σ8λ1P(S2).\displaystyle 1-P(S_{2})-\big{(}1-P(S_{2}+\sigma^{\sharp})\big{)}e^{-\frac{\lambda\Delta H_{\Lambda}-2\sigma^{\sharp}}{8}}\xrightarrow{\lambda\to\infty}1-P(S_{2})\,.

(iii)(iii) It remains to exhibit a criterion for λ\lambda ensuring that the blowup size π2\pi_{2} is larger than 1CΛ/21-C_{\Lambda}/2. Such a criterion can be obtained by first observing that

π21P(S2)(1P(S2+σ))eλΔHΛ2σ81P(S2)eλΔHΛ2σ8.\displaystyle\pi_{2}\geq 1-P(S_{2})-\big{(}1-P(S_{2}+\sigma^{\sharp})\big{)}e^{-\frac{\lambda\Delta H_{\Lambda}-2\sigma^{\sharp}}{8}}\geq 1-P(S_{2})-e^{-\frac{\lambda\Delta H_{\Lambda}-2\sigma^{\sharp}}{8}}\,.

Then let us define λ>0\lambda^{\sharp}>0 as

λ=sup{λ0|λ2ΔHΛ(σ4ln(CΛ2λ))}<CΛ2eσ/4<,\displaystyle\lambda^{\sharp}=\sup\left\{\lambda\geq 0\,\bigg{|}\,\lambda\leq\frac{2}{\Delta H_{\Lambda}}\left(\sigma^{\sharp}-4\ln{\left(\frac{C_{\Lambda}}{2\lambda}\right)}\right)\right\}<\frac{C_{\Lambda}}{2}e^{-\sigma^{\sharp}/4}<\infty\,,

such that for all λλ\lambda\geq\lambda^{\sharp}, we have

eλΔHΛ2σ8CΛ2λ.\displaystyle e^{-\frac{\lambda\Delta H_{\Lambda}-2\sigma^{\sharp}}{8}}\leq\frac{C_{\Lambda}}{2\lambda}\,.

Utilizing the fact that P(S2)<CΛ/(2λ)P(S_{2})<C_{\Lambda}/(2\lambda) by Proposition 3.4, we have

π21P(S2)CΛ2λ1CΛλ.\displaystyle\pi_{2}\geq 1-P(S_{2})-\frac{C_{\Lambda}}{2\lambda}\geq 1-\frac{C_{\Lambda}}{\lambda}\,.

Thus choosing

λLΛ=max(2lΛ,2/CΛ,22πΛeΛ2/2CΛ,λ)\displaystyle\lambda\geq L_{\Lambda}=\max\left(2l_{\Lambda},2/C_{\Lambda},2\sqrt{2\pi\Lambda}e^{\Lambda^{2}/2C_{\Lambda}},\lambda^{\sharp}\right)

suffices to ensure that the next full blowup has size at least 1CΛ/λ1-C_{\Lambda}/\lambda. ∎

The above proposition directly implies the existence and uniqueness of global explosive solutions under the following assumptions:

Assumption 4.

Consider a dPMF dynamics with refractory period 0ϵ<CΛ/ν0\leq\epsilon<C_{\Lambda}/\nu, with interaction parameter λLΛ\lambda\geq L_{\Lambda}, and with natural initial conditions such that q0q_{0} contains an Dirac-delta mass π0δΛ\pi_{0}\delta_{\Lambda} with π01CΛ/λ\pi_{0}\geq 1-C_{\Lambda}/\lambda.

Theorem 3.2.

Under Assumption 4 , the fixed-point problem 1.3 admits a unique global solution Ψ\Psi defined over the whole half-line +\mathbbm{R}^{+}. Moreover, this solution presents an infinite but discrete set of full blowups with trigger times {Sk}k\{S_{k}\}_{k\in\mathbbm{N}} and with sizes {πk}k\{\pi_{k}\}_{k\in\mathbbm{N}} such that for all k1k\geq 1, πk1CΛ/λ\pi_{k}\geq 1-C_{\Lambda}/\lambda and Sk+1UkνϵS_{k+1}-U_{k}\geq\nu\epsilon where Uk=Sk+λπkU_{k}=S_{k}+\lambda\pi_{k}.

Proof.

By Theorem 2.6, the fixed-point problem 1.3 admits a local smooth solution up to time S1=inf{σ>0|Ψ(σ)0}S_{1}=\inf\left\{\sigma>0\,|\,\Psi^{\prime}(\sigma)\geq 0\right\}. By the same arguments as in the proof of Proposition 3.3, the blowup condition must be satisfied in finite time: S1<S_{1}<\infty. By the same arguments as in the proof of Proposition 3.5, the blowup size satisfies π1>1CΛ/λ\pi_{1}>1-C_{\Lambda}/\lambda. From there on, one can iteratively apply Proposition 3.5 to define a global solution to the fixed-point problem 1.3 with an infinite number of blowups. The uniqueness of the global solution follows from the uniqueness of the local solutions up to blowups and the uniqueness of the continuation process at the exit of a blowup. Finally, this unique global solution Φ\Phi is defined over the whole half-line +\mathbbm{R}^{+} since all blowups are lower bounded by 1CΛ/λ1-C_{\Lambda}/\lambda, so that all local solutions are defined over a domain of duration at least λ(1CΛ/λ)\lambda(1-C_{\Lambda}/\lambda) during blowup episodes. Finally, a blowup can only happen after resetting the processes that inactivated during the previous blowup, which must happens after a duration νϵ\nu\epsilon by the proof of Proposition 3.3. ∎

For ϵ>0\epsilon>0, Theorem 3.2 directly implies the existence of global time change Φ\Phi parametrizing the original dPMF dynamics over the the half-line +\mathbbm{R}^{+} with the properties listed in Theorem 1.8. However, Theorem 3.2 only implies that Tk+1Tk=Ψ(Sk+1)Ψ(Sk)=Ψ(Sk+1)Ψ(Uk)ϵT_{k+1}-T_{k}=\Psi(S_{k+1})-\Psi(S_{k})=\Psi(S_{k+1})-\Psi(U_{k})\geq\epsilon, where {Tk=Ψ(Sk)}kN\{T_{k}=\Psi(S_{k})\}_{k\in\mathrm{N}} is the sequence of blowup times for the original dPMF dynamics. Thus, in agreement with the remark following Proposition 2.5, Theorem 3.2 does not allow us to conclude about the existence of a global time change Φ\Phi in the limit ϵ0+\epsilon\to 0^{+}. Indeed, it could be that limϵ0+Tk+1Tk=0\lim_{\epsilon\to 0^{+}}T_{k+1}-T_{k}=0 for all large enough kk, which is compatible with accumulating the blowup times so that

T=limkTk=limkΨ(Sk)=limSΨ(S)<.\displaystyle T_{\infty}=\lim_{k\to\infty}T_{k}=\lim_{k\to\infty}\Psi(S_{k})=\lim_{S\to\infty}\Psi(S)<\infty\,. (46)

As a result, Φ=Ψ1\Phi=\Psi^{-1} would only define a solution time change Φ\Phi up the finite time TT_{\infty}. We address this point in the following section, where we characterize PMF dynamics (with zero refractory period) as limits of dPMF dynamics when ϵ0+\epsilon\to 0^{+}.

4 Limit of vanishing refractory periods

In this section, we show that explosive PMF dynamics can be defined consistently over the whole half-line +\mathbbm{R}^{+} for zero refractory period, i.e., with instantaneous reset. First, we define PMF dynamics for zero refractory period ϵ=0\epsilon=0, which by contrast dPMF dynamics with ϵ>0\epsilon>0, admit an explicit iterative construction. Second, we introduce a series of continuity results showing that PMF dynamics are “physical” in the sense that they are recovered from dPMF dynamics in the limit ϵ0+\epsilon\to 0^{+}. Finally, we provide the proofs for these continuity results, which essentially amounts to show that at all blowup trigger, exit, and reset times, the spatial densities of dPMF dynamics converge toward their PMF counterpart in L1L_{1} norm.

4.1 Solutions with zero refractory periods

Theorem 3.2 applies to the case of zero refractory period ϵ=0\epsilon=0, that is for PMF dynamics, assuming natural initial condition with normalized spatial component, 0q0(x)dx=1\int_{0}^{\infty}q_{0}(x)\,\mathrm{d}x=1, and no inactive processes. In particular, for ϵ=0\epsilon=0, the fixed-point problem 1.3 admits a unique global solution for large enough interaction parameter λ\lambda and for sufficiently large initial mass concentrated at Λ\Lambda. Let us denote this solution by Ψ\Psi for simplicity. As an inverse time change, Ψ\Psi parametrizes a PMF dynamics with an infinite but discrete set of well ordered blowups with successive trigger times {Sk}k\{S_{k}\}_{k\in\mathbbm{N}} and corresponding size {πk}k\{\pi_{k}\}_{k\in\mathbbm{N}}. A major benefit of considering PMF dynamics is that the corresponding inverse time changes Φ\Phi admit an explicit iterative construction. Specifically, the sequences {Sk}k\{S_{k}\}_{k\in\mathbbm{N}} and {πk}k\{\pi_{k}\}_{k\in\mathbbm{N}} can be defined explicitely via the adjunction of the sequence of initial spatial conditions {q0,k()=q(Sk,)}k\{q_{0,k}(\cdot)=q(S_{k},\cdot)\}_{k\in\mathbbm{N}}. This explicit definition is made possible by the fact that for zero refractory period ϵ=0\epsilon=0, the renewal problems at stake loose their delayed character and can be solved analytically. In particular, in between blowup episodes, we have the fundamental solution

g(σ,x)=k=0h(σ,x+kΛ),\displaystyle g(\sigma,x)=\sum_{k=0}^{\infty}h(\sigma,x+k\Lambda)\,,

where g(σ,x)g(\sigma,x) is the instantaneous flux of a process started in x>0x>0 with zero delay reset. This explicit fundamental solution allows one to define the announced sequence {Sk,πk,q0,k}k\{S_{k},\pi_{k},q_{0,k}\}_{k\in\mathbbm{N}} in +×(0,1]×((0,))\mathbbm{R}^{+}\times(0,1]\times\mathcal{M}((0,\infty)) as follows:

Definition 4.1.

With the convention that S0=0S_{0}=0 and for a natural initial condition q0q_{0} with Dirac-delta mass π0δΛ\pi_{0}\delta_{\Lambda} satisfying Assumption 4, let us define the sequence {Sk,πk,q0,k}k\{S_{k},\pi_{k},q_{0,k}\}_{k\in\mathbbm{N}} by setting q0,0=q0q_{0,0}=q_{0} and iterating for all k1k\geq 1:

  1. 1.

    Smooth dynamics:

    gk(σ)=0g(σ,x)q0,k(x)dx.\displaystyle g_{k}(\sigma)=\int_{0}^{\infty}g(\sigma,x)q_{0,k}(x)\,\mathrm{d}x\,.
    qk(σ,y)=0σκ(στ,y,Λ)gk(τ)dτ+0κ(σ,y,x)q0,k(x)dx.\displaystyle q_{k}(\sigma,y)=\int_{0}^{\sigma}\kappa(\sigma-\tau,y,\Lambda)g_{k}(\tau)\,\mathrm{d}\tau+\int_{0}^{\infty}\kappa(\sigma,y,x)q_{0,k}(x)\,\mathrm{d}x\,.
  2. 2.

    Blowup trigger time:

    Sk+1Sk=inf{σ>0|gk(σ)>1/λ}<.\displaystyle S_{k+1}-S_{k}=\inf\left\{\sigma>0\,\Big{|}\,g_{k}(\sigma)>1/\lambda\right\}<\infty\,.
  3. 3.

    Blowup size:

    πk+1=inf{p0|p>0+H(λp,x)qk(Sk+1,x)dx}>0.\displaystyle\pi_{k+1}=\inf\left\{p\geq 0\,\bigg{|}\,p>\int_{0^{+}}^{\infty}H(\lambda p,x)q_{k}(S_{k+1},x)\,\mathrm{d}x\right\}>0\,.
  4. 4.

    Blowup exit/reset distribution:

    q0,k+1(y)=0κ(λπk+1,y,x)qk(Sk+1,x)dx+πk+1δΛ(x).\displaystyle q_{0,k+1}(y)=\int_{0}^{\infty}\kappa(\lambda\pi_{k+1},y,x)q_{k}(S_{k+1},x)\,\mathrm{d}x+\pi_{k+1}\delta_{\Lambda}(x)\,.

The distributions {q0,k}k\{q_{0,k}\}_{k\in\mathbbm{N}} defined above are in fact the reset distributions for PMF dynamics, q0,k=q(Rk,)q_{0,k}=q(R_{k},\cdot). Moreover, with ϵ=0\epsilon=0, post-blowup reset is instantaneous so that Uk=RkU_{k}=R_{k} and exit distributions are recovered by excluding the reset mass πk\pi_{k}: q(Rk,)=q0,kπkδΛq(R_{k}^{-},\cdot)=q_{0,k}-\pi_{k}\delta_{\Lambda}, . The sequence {Sk,πk,q0,k}k\{S_{k},\pi_{k},q_{0,k}\}_{k\in\mathbbm{N}} allows one to give an explicit piecewise representation of the global solution Ψ\Psi. This representation makes use of the functions

σΨk(σ)=𝟙{σ>0}(σλ0xG(σ,x)qk(x)dx)/ν,\displaystyle\sigma\mapsto\Psi_{k}(\sigma)=\mathbbm{1}_{\{\sigma>0\}}\left(\sigma-\lambda\int_{0}^{x}G(\sigma,x)q_{k}(x)\,\mathrm{d}x\right)/\nu\,,

where G(,x)G(\cdot,x) denotes the cumulative flux function associated to g(,x)g(\cdot,x). Together with the trigger time SkS_{k}, the functions Ψk\Psi_{k} allows one to specify the sequence of blowup time {Tk}k\{T_{k}\}_{k\in\mathbbm{N}} for the original dynamics as:

T0=0,Tk+1=Tk+Ψk(Sk+1).\displaystyle T_{0}=0\,,\quad T_{k+1}=T_{k}+\Psi_{k}\big{(}S_{k+1}\big{)}\,.

We then obtain the following piecewise representation:

Definition 4.2.

For zero refractory period ϵ=0\epsilon=0, the global solution Ψ\Psi is given by

Ψ(σ)=k=0𝟙{Sk<σSk+1}[Tk+Ψk+1(σ(Sk+πk))].\displaystyle\Psi(\sigma)=\sum_{k=0}^{\infty}\mathbbm{1}_{\{S_{k}<\sigma\leq S_{k+1}\}}\left[T_{k}+\Psi_{k+1}\big{(}\sigma-(S_{k}+\pi_{k})\big{)}\right]\,.

Observe that at this stage, although Ψ\Psi is defined over the whole half-line +\mathbbm{R}^{+} by Theorem 3.2, we cannot invoke Proposition 2.5 to show that the corresponding time change Φ=Ψ1\Phi=\Psi^{-1} is defined over the whole half-line +\mathbbm{R}^{+}. However, we can establish this point by using the following estimate for inter-blowup times:

Proposition 4.1.

For natural initial condition q0q_{0} with Dirac-delta mass π0δΛ\pi_{0}\delta_{\Lambda} satisfying Assumption 4, we have:

Ψ(Sk+1)Ψ(Sk)>14νinf{σ>0|h(σ,Λ)>aΛλ}>0,\displaystyle\Psi(S_{k+1})-\Psi(S_{k})>\frac{1}{4\nu}\inf\left\{\sigma>0\,\bigg{|}\,h(\sigma,\Lambda)>\frac{a_{\Lambda}}{\lambda}\right\}>0\,,

where aΛa_{\Lambda} is a positive constant that only depends on Λ\Lambda. Thus, T=limσΨ(σ)=T_{\infty}=\lim_{\sigma\to\infty}\Psi(\sigma)=\infty and Φ=Ψ1\Phi=\Psi^{-1} is defined over the whole half-line +\mathbbm{R}^{+}.

Proof.

Let us first observe that for ϵ=0\epsilon=0, we have Uk=RkU_{k}=R_{k} and

ν(Ψ(Sk+1)Ψ(Sk))\displaystyle\nu\big{(}\Psi(S_{k+1})-\Psi(S_{k})\big{)} =\displaystyle= ν(Ψ(Sk+1)Ψ(Rk)),\displaystyle\nu\big{(}\Psi(S_{k+1})-\Psi(R_{k})\big{)}\,,
=\displaystyle= Sk+1Rkλ(G(Sk+1)G(Rk)).\displaystyle S_{k+1}-R_{k}-\lambda\big{(}G(S_{k+1})-G(R_{k})\big{)}\,.

Then the proposed lower bound will follow from bounding above G(Sk+1)G(Rk)G(S_{k+1})-G(R_{k}), the cumulative flux of inactivation between the kk-th blowup and the (k+1)(k\!+\!1)-th blowup. In this view, we split that cumulative flux as G(Sk+1)G(Rk)=ΔGπk,k+ΔGπk,kG(S_{k+1})-G(R_{k})=\Delta G_{\cancel{\pi_{k}},k}+\Delta G_{\pi_{k},k}, where ΔGπk,k\Delta G_{\cancel{\pi_{k}},k} and ΔGπk,k\Delta G_{\pi_{k},k} represent the contributions of the processes that survive or inactivate during the last blowup, respectively. As Assumption 4 ensures that gπkRk,Sk+11/(4λ)\|g_{\cancel{\pi_{k}}}\|_{R_{k},S_{k+1}}\leq 1/(4\lambda), we have

ΔGs,kgπkRk,Sk+1(Sk+1Rk)Sk+1Rk4λ.\displaystyle\Delta G_{s,k}\leq\|g_{\cancel{\pi_{k}}}\|_{R_{k},S_{k+1}}(S_{k+1}-R_{k})\leq\frac{S_{k+1}-R_{k}}{4\lambda}\,.

Moreover, Assumption 4 ensures that gπkg_{\pi_{k}} remains a convex function up to the next blowup so that gπk(Rk)=0g_{\pi_{k}}(R_{k})=0 and gπk(Sk+1)1/λg_{\pi_{k}}(S_{k+1})\leq 1/\lambda implies that

ΔGπk,kgπkRk,Sk+1(Sk+1Rk)2Sk+1Rk2λ.\displaystyle\Delta G_{\pi_{k},k}\leq\frac{\|g_{\pi_{k}}\|_{R_{k},S_{k+1}}(S_{k+1}-R_{k})}{2}\leq\frac{S_{k+1}-R_{k}}{2\lambda}\,.

Considering the above bounds together leads to

ν(Ψ(Sk+1)Ψ(Sk))=Sk+1Rkλ(ΔGπk,k+ΔGπk,k)\displaystyle\nu\big{(}\Psi(S_{k+1})-\Psi(S_{k})\big{)}=S_{k+1}-R_{k}-\lambda\big{(}\Delta G_{{\cancel{\pi_{k}}},k}+\Delta G_{\pi_{k},k}\big{)} \displaystyle\geq (Sk+1Rk)/4.\displaystyle(S_{k+1}-R_{k})/4\,.

The announced lower bound follows from the lower bound s1s_{1} defined in (40) which satisfies

Sk+1Rks1inf{σ>0|h(σ,Λ)>aΛλ}=Λ22ln(λ)+o(1/lnλ).\displaystyle S_{k+1}-R_{k}\geq s_{1}\geq\inf\left\{\sigma>0\,\bigg{|}\,h(\sigma,\Lambda)>\frac{a_{\Lambda}}{\lambda}\right\}=\frac{\Lambda^{2}}{2\ln\left(\lambda\right)}+o(1/\ln\lambda)\,.

Proposition 4.1 directly implies that T=limSΦ(S)=T_{\infty}=\lim_{S\to\infty}\Phi(S)=\infty, so that Φ\Phi parametrizing a PMF dynamics over the whole half-line +\mathbbm{R}^{+}. We conclude by giving the explicit, iterative construction of explosive PMF dynamics with zero refractory period ϵ=0\epsilon=0.

Theorem 4.3.

Under Assumption 4, PMF dynamics admits the density function p(t,y)=d[Xtx|Xt>0]/dxp(t,y)=\mathrm{d}\mathbb{P}\left[X_{t}\leq x|X_{t}>0\right]/\mathrm{d}x

p(t,x)=k=0𝟙{Tkt<Tk+1}qk(Φ(t)Tk,x),\displaystyle p(t,x)=\sum_{k=0}^{\infty}\mathbbm{1}_{\{T_{k}\leq t<T_{k+1}\}}q_{k}\big{(}\Phi(t)-T_{k},x\big{)}\,,

where the sequence (Tk,qk)(T_{k},q_{k}) is specified in Definition 4.1 and the time change Φ=Ψ1\Phi=\Psi^{-1} is specified in Definition 4.2.

4.2 Global continuity in the limit of vanishing refractory periods

Considering Theorem 3.2 for zero refractory period ϵ=0\epsilon=0 allows for the explicit construction of explosive PMF dynamics in the large interaction regime. To justify that the thus-constructed dynamics are “physical”, we need to check that these non-delayed solutions can be recovered from dPMF dynamics in the limit of vanishing refractory period ϵ0+\epsilon\to 0^{+}. In this perspective, let us consider some natural initial conditions for explosive PMF dynamics that satisfy Assumption 4. Such initial conditions are entirely specified by the spatial component q0q_{0}, which includes an Dirac-delta mass of size π0\pi_{0} at Λ\Lambda. By Theorem 3.2, the corresponding explosive PMF dynamics is fully parametrized by the inverse time-changed function Ψ\Psi specified in Definition 4.1. Given the same natural initial condition, Theorem 3.2 also guarantees the existence and uniqueness of explosive dPMF dynamics for small enough refractory period ϵ\epsilon under Assumption 4. Let us denote the corresponding inverse time change by Ψϵ\Psi_{\epsilon}. By Theorem 3.2, Ψϵ\Psi_{\epsilon} specifies a countable infinity of jumps with trigger times {Sϵ,k}k\{S_{\epsilon,k}\}_{k\in\mathbbm{N}}, exit times {Uϵ,k}k\{U_{\epsilon,k}\}_{k\in\mathbbm{N}}, and reset times {Rϵ,k}k\{R_{\epsilon,k}\}_{k\in\mathbbm{N}} such that 0=Rϵ,0<Sϵ,1<Uϵ,1<Rϵ,1<Sϵ,2<Uϵ,2<Rϵ,2<0=R_{\epsilon,0}<S_{\epsilon,1}<U_{\epsilon,1}<R_{\epsilon,1}<S_{\epsilon,2}<U_{\epsilon,2}<R_{\epsilon,2}<\dots. Our goal is to justify the following continuity result:

Theorem 4.4.

Given the same purely spatial natural initial condition satisfying Assumption 4, the dPMF solution Ψϵ\Psi_{\epsilon} converges compactly toward the PMF solution Ψ\Psi on +\mathbbm{R}^{+} and the dPMF blowup times Sϵ,kS_{\epsilon,k}, Uϵ,kU_{\epsilon,k}, Rϵ,kR_{\epsilon,k} converge toward their PMF counterparts SkS_{k} and Uk=RkU_{k}=R_{k} as well.

The proof of the above Proposition relies on a simple recurrence argument that makes use of four continuity results, which we first give without proof. The first set of results, comprising Proposition 4.2 and Proposition 4.3, deal with the continuity of dPMF dynamics in between blowup episodes in the limit of vanishing refractory period ϵ0+\epsilon\to 0^{+}. These results essentially follow from the uniform boundedness of the various functions involved in the integral representation of the density of the dPMF dynamics. Proposition 4.2 bears on the continuity of the cumulative fluxes, which boils down to the continuity of blowup trigger times with respect to the initial conditions for the L1L_{1} norm.

Proposition 4.2.

Consider some natural initial conditions (qϵ,0,gϵ,0)(q_{\epsilon,0},g_{\epsilon,0}) in ((0,))×([ξϵ,0,0))\mathcal{M}((0,\infty))\times\mathcal{M}([-\xi_{\epsilon,0},0)) such that qϵ,0q0q_{\epsilon,0}\to q_{0} in L1L_{1} norm when ϵ0+\epsilon\to 0^{+}, where q0q_{0} is a probability density with Dirac-delta mass π0δΛ\pi_{0}\delta_{\Lambda}. Then under Assumption 4, we have:

(1)(1) limϵ0+Sϵ,1=S1\lim_{\epsilon\to 0^{+}}S_{\epsilon,1}=S_{1} where Sϵ,1S_{\epsilon,1} and S1S_{1} denotes the next blowup trigger times of the time changes Ψϵ\Psi_{\epsilon} and Ψ\Psi, respectively.

(2)(2) for all kk in \mathbbm{N}, the kk-iterated derivatives of GϵG_{\epsilon} and GG satisfy Gϵ(k)G(k)G^{(k)}_{\epsilon}\to G^{(k)} uniformly on [0,S1+λ/3][0,S_{1}+\lambda/3].

Proposition 4.2 establishes that PMF dynamics are locally recovered from dPMF dynamics in the limit of vanishing refractory period, at least up to the first blowup episode. Then, a natural strategy to prove Proposition 4.4 is to apply Proposition 4.2 iteratively after each blowup episodes in conjunction with some continuity result about the blowup episodes. It turns out that these continuity results will hold with respect to the L1L_{1} norm of the spatial densities of the surviving processes at trigger, exit, and reset times. Thus, we also need to make sure that the density of surviving processes at blowup trigger time qϵ(Sϵ,1,)q_{\epsilon}(S_{\epsilon,1},\cdot) converges toward q(S1,)q(S_{1},\cdot) for the L1L_{1} norm, as stated in the following proposition:

Proposition 4.3.

Consider some natural initial conditions (qϵ,0,gϵ,0)(q_{\epsilon,0},g_{\epsilon,0}) in ((0,))×([ξϵ,0,0))\mathcal{M}((0,\infty))\times\mathcal{M}([-\xi_{\epsilon,0},0)) such that qϵ,0q0q_{\epsilon,0}\to q_{0} in L1L_{1} norm when ϵ0+\epsilon\to 0^{+}, where q0q_{0} contains a Dirac-delta mass π0δΛ\pi_{0}\delta_{\Lambda}. Then under Assumption 4, we have limϵ0+qϵ(Sϵ,1,)q(S1,)1=0\lim_{\epsilon\to 0^{+}}\|q_{\epsilon}(S_{\epsilon,1},\cdot)-q(S_{1},\cdot)\|_{1}=0.

The second set of results, comprising Proposition 4.2 and Proposition 4.3, bears on the continuity of dPMF and PMF dynamics during blowup episodes. Proposition 4.2 states that given a blowup trigger time S1S_{1}, the next exit blowup time U1=S1+λπ1U_{1}=S_{1}+\lambda\pi_{1} and the distribution of surviving processes q(U1,)q(U_{1},\cdot) continuously depend on the distribution of surviving processes q(S1,)q(S_{1},\cdot) at trigger time with respect to the L1L_{1} norm. Specifically:

Proposition 4.4.

Suppose S1S_{1} marks the first blowup trigger time after S0S_{0} under Assumption 4. Then there exists a neighborhood 𝒩\mathcal{N} of q(S1,)q(S_{1},\cdot) such that the maps

qπ1[q]andq(y0κ(λπ1[q],y,x)q(x)dx)\displaystyle q\mapsto\pi_{1}[q]\quad\mathrm{and}\quad q\mapsto\left(y\mapsto\int_{0}^{\infty}\kappa(\lambda\pi_{1}[q],y,x)q(x)\,\mathrm{d}x\right)

are continuous on 𝒩\mathcal{N} with respect to the L1L_{1} norms.

As blowup resolutions involve dynamics without reset, there is no role for the refractory period in establishing the above result, which follows from the Banach-space version of the implicit function theorem. Proposition 4.3 directly implies that if limϵ0+qϵ(Sϵ,1,)q(S1,)1=0\lim_{\epsilon\to 0^{+}}\|q_{\epsilon}(S_{\epsilon,1},\cdot)-q(S_{1},\cdot)\|_{1}=0, then limϵ0+πϵ,1=π1\lim_{\epsilon\to 0^{+}}\pi_{\epsilon,1}=\pi_{1} and limϵ0+qϵ(Uϵ,1,)q(U1,)1=0\lim_{\epsilon\to 0^{+}}\|q_{\epsilon}(U_{\epsilon,1},\cdot)-q(U_{1}^{-},\cdot)\|_{1}=0, where q(U1,)q(U_{1}^{-},\cdot) denotes the distribution of surviving processes for PMF dynamics just before reset at U1=R1U_{1}=R_{1}. In order to propagate these convergence results via recurrence, we then only need to check the L1L_{1} convergence of the distribution of processes qϵ(Rϵ,1,)q_{\epsilon}(R_{\epsilon,1},\cdot) toward q(U1,)=q(R1,)q(U_{1},\cdot)=q(R_{1},\cdot), allowing one to apply Proposition 4.2 and 4.3 anew with qϵ(Rϵ,1,)q_{\epsilon}(R_{\epsilon,1},\cdot) and q(U1,)=q(R1,)q(U_{1},\cdot)=q(R_{1},\cdot) as natural initial conditions.

Proposition 4.5.

Under Assumption 4, suppose that Rϵ,1R_{\epsilon,1} marks the first blowup reset time after S0S_{0}. Then we have limϵ0+qϵ(Rϵ,1,)q(R1,)1=0\lim_{\epsilon\to 0^{+}}\|q_{\epsilon}(R_{\epsilon,1},\cdot)-q(R_{1},\cdot)\|_{1}=0.

For the sake of completeness, we provide the proof of Proposition 4.4 using Propositions 4.2, 4.3, 4.4, and 4.5, whose proofs are given in Section Section 4.3, Section 4.4, Section 4.5, and 4.6, respectively.

Proof of Theorem 4.4.

Let us consider the fixed-point solution Ψϵ\Psi_{\epsilon} for a refractory period ϵ>0\epsilon>0 with purely spatial natural initial condition q0q_{0}. This amounts to considering that all processes are active at an initial reset time Rϵ,0=0R_{\epsilon,0}=0 under Assumption 4. By Theorem 3.2, Ψϵ\Psi_{\epsilon} presents a countable infinity of jumps with well ordered trigger times {Sϵ,k}k\{S_{\epsilon,k}\}_{k\in\mathbbm{N}}, exit times {Uϵ,k}k\{U_{\epsilon,k}\}_{k\in\mathbbm{N}}, and reset times {Rϵ,k}k\{R_{\epsilon,k}\}_{k\in\mathbbm{N}} such that 0=Rϵ,0<Sϵ,1<Uϵ,1<Rϵ,1<Sϵ,2<Uϵ,2<Rϵ,2<0=R_{\epsilon,0}<S_{\epsilon,1}<U_{\epsilon,1}<R_{\epsilon,1}<S_{\epsilon,2}<U_{\epsilon,2}<R_{\epsilon,2}<\dots. Moreover, we have the following facts:

  1. 1.

    The post-reset durations to the next trigger time Uϵ,kRϵ,kU_{\epsilon,k}-R_{\epsilon,k} are uniformly bounded away from zero by the ϵ\epsilon-independent constant s1s_{1} defined in (40).

  2. 2.

    The blowup durations Sϵ,kUϵ,kS_{\epsilon,k}-U_{\epsilon,k} satisfy Sϵ,kUϵ,k=λπϵ,kS_{\epsilon,k}-U_{\epsilon,k}=\lambda\pi_{\epsilon,k}, where the blowup sizes {πϵ,k}k\{\pi_{\epsilon,k}\}_{k\in\mathbbm{N}} are uniformly bounded away from zero by the ϵ\epsilon-independent constant 1CΛ/λ>1/21-C_{\Lambda}/\lambda>1/2 where CΛC_{\Lambda} is chosen according to (42).

  3. 3.

    The post-exit delays to reset times are uniformly bounded by νϵRϵ,kUϵ,k2νϵ\nu\epsilon\leq R_{\epsilon,k}-U_{\epsilon,k}\leq 2\nu\epsilon.

Consider then the well ordered blow up times 0=R0<S1<U1<R1<S2<U2<R2<0=R_{0}<S_{1}<U_{1}<R_{1}<S_{2}<U_{2}<R_{2}<\dots for the corresponding PMF solution Ψ\Psi obtained for the same purely spatial initial condition q0q_{0} but with ϵ=0\epsilon=0. Assuming that limϵ0+Rϵ,k=Rk\lim_{\epsilon\to 0^{+}}R_{\epsilon,k}=R_{k} and that limϵ0+qϵ(Rϵ,k,)qϵ(Rk,)1=0\lim_{\epsilon\to 0^{+}}\|q_{\epsilon}(R_{\epsilon,k},\cdot)-q_{\epsilon}(R_{k},\cdot)\|_{1}=0, let us show that limϵ0+Rϵ,k+1=Rk+1\lim_{\epsilon\to 0^{+}}R_{\epsilon,k+1}=R_{k+1} and that limϵ0+qϵ(Rϵ,k+1,)qϵ(Rk+1,)1=0\lim_{\epsilon\to 0^{+}}\|q_{\epsilon}(R_{\epsilon,k+1},\cdot)-q_{\epsilon}(R_{k+1},\cdot)\|_{1}=0, which is the key to propagate our recurrence argument.

(i)(i) Inter-blowup evolution: For all ϵ>0\epsilon>0, the densities (qϵ(Rϵ,k,),{gϵ(σ)}ξϵ(Rϵ,k)σ<Rϵ,k)\big{(}q_{\epsilon}(R_{\epsilon,k},\cdot),\{g_{\epsilon}(\sigma)\}_{\xi_{\epsilon}(R_{\epsilon,k})\leq\sigma<R_{\epsilon,k}}\big{)} defines natural initial conditions at blowup reset time Rϵ,kR_{\epsilon,k}. Assuming that qϵ(Rϵ,k,)q_{\epsilon}(R_{\epsilon,k},\cdot) converges toward qϵ(Rϵ,k,)q_{\epsilon}(R_{\epsilon,k},\cdot) in L1L_{1} norm allows one to invoke Proposition 4.2: the next blowup trigger time Sϵ,k+1S_{\epsilon,k+1} is such that limϵ0+Sϵ,k+1Rϵ,k=Sk+1Rk\lim_{\epsilon\to 0^{+}}S_{\epsilon,k+1}-R_{\epsilon,k}=S_{k+1}-R_{k} and Ψϵ(+Rϵ,k)\Psi_{\epsilon}(\cdot+R_{\epsilon,k}) uniformly converges toward Ψ(+Rk)\Psi(\cdot+R_{k}) on an interval containing [0,Sk+1Rk+λ/3][0,S_{k+1}-R_{k}+\lambda/3]. Moreover, by Proposition 4.3, the spatial density at trigger time qϵ(Sϵ,k+1,)q_{\epsilon}(S_{\epsilon,k+1},\cdot) converges toward q(Sk+1,)q(S_{k+1},\cdot) in L1L_{1} norm.

(ii)(ii) Blowup episode: By Proposition 4.4, limϵ0+qϵ(Sϵ,k+1,)qϵ(Sk+1,)1=0\lim_{\epsilon\to 0^{+}}\|q_{\epsilon}(S_{\epsilon,k+1},\cdot)-q_{\epsilon}(S_{k+1},\cdot)\|_{1}=0 entails that the next blowup size satisfies limϵ0+πϵ,k+1=πk+1\lim_{\epsilon\to 0^{+}}\pi_{\epsilon,k+1}=\pi_{k+1}. Consequently, since the inverse time changes Ψϵ\Psi_{\epsilon} are uniformly Lipschitz with respect to ϵ>0\epsilon>0, Ψϵ(+Rϵ,k)\Psi_{\epsilon}(\cdot+R_{\epsilon,k}) converges uniformly toward Ψ(+Rk)\Psi(\cdot+R_{k}) on [0,Sk+1Rk+λπk+1][0,S_{k+1}-R_{k}+\lambda\pi_{k+1}]. Moreover, by Proposition 4.4, the spatial density at exit time qϵ(Uϵ,k+1,)q_{\epsilon}(U_{\epsilon,k+1},\cdot) converges toward q(Uk+1,)q(U_{k+1}^{-},\cdot) in L1L_{1} norm. Finally, by Proposition 4.5, the spatial density at reset time qϵ(Rϵ,k+1,)q_{\epsilon}(R_{\epsilon,k+1},\cdot) converges toward q(Rk+1,)q(R_{k+1},\cdot) in L1L_{1} norm, with q(Rk+1,)q(R_{k+1},\cdot) still satisfying Assumption 4.

As we consider identical purely spatial initial conditions at Rϵ,0=R0=0R_{\epsilon,0}=R_{0}=0: qϵ,0(0,)=q0(0,)q_{\epsilon,0}(0,\cdot)=q_{0}(0,\cdot), the recurrence naturally initializes so that for all kk in \mathbbm{N}, the inverse time Ψϵ\Psi_{\epsilon} uniformly converges toward Ψ\Psi on the interval [0,Rk][0,R_{k}] and the dPMF blowup times converge toward their PMF counterparts: limϵ0+(Sϵ,k,Uϵ,k,Rϵ,k)=(Sk,Uk,Rk)\lim_{\epsilon\to 0^{+}}(S_{\epsilon,k},U_{\epsilon,k},R_{\epsilon,k})=(S_{k},U_{k},R_{k}). We conclude by observing that limkRϵ,klimkk(λ/2+s1)=\lim_{k\to\infty}R_{\epsilon,k}\geq\lim_{k\to\infty}k(\lambda/2+s_{1})=\infty, which establishes the compact convergence of Ψϵ\Psi_{\epsilon} toward Ψ\Psi on +\mathbbm{R}^{+}. ∎

4.3 Continuity in between blowups

Proof of Proposition 4.2.

The proof proceeds in five steps: (i)(i) we show that the backward-delay function ηϵ\eta_{\epsilon} converges compactly toward zero when ϵ0+\epsilon\to 0^{+} on some interval [0,S¯1][0,\underline{S}_{1}]; (ii)(ii) we show that the cumulative flux GϵG_{\epsilon} converges compactly toward the cumulative flux GG obtained for ϵ=0\epsilon=0 on [0,S¯1][0,\underline{S}_{1}]; (iii)(iii) we show that the convergence result of (ii)(ii) extends to the iterated derivative of GϵG_{\epsilon}; (iv)(iv) using the result from (iii)(iii), we show that the interval bound S¯1\underline{S}_{1} defined in (i)(i) coincides with S1S_{1}, the next blowup trigger time for ϵ=0\epsilon=0; (v)(v) finally, we extend the convergence results to the interval [0,S¯1+λ/2][0,\underline{S}_{1}+\lambda/2], which partially covers the next blowup episode.

(i)(i) Under Assumption 4, for all ϵ>0\epsilon>0, the natural initial conditions (qϵ,0,gϵ,0)(q_{\epsilon,0},g_{\epsilon,0}) in ((0,))×([ξϵ,0,0))\mathcal{M}((0,\infty))\times\mathcal{M}([-\xi_{\epsilon,0},0)) are necessarily such that gϵ,01/(4λ)<1/λg_{\epsilon,0}\leq 1/(4\lambda)<1/\lambda with gϵ(0)=limσ0gϵ,0(σ)1/(4λ)<1/λg_{\epsilon}(0)=\lim_{\sigma\to 0}g_{\epsilon,0}(\sigma)\leq 1/(4\lambda)<1/\lambda as well. Therefore, one can consider for sufficiently small ϵ>0\epsilon>0, the positive times S~ϵ,1\tilde{S}_{\epsilon,1} and Sϵ,1S_{\epsilon,1}

S~ϵ,1=inf{σ>0|gϵ(σ)>1νϵλ}<inf{σ>0|gϵ(σ)>1λ}=Sϵ,1,\displaystyle\tilde{S}_{\epsilon,1}=\inf\left\{\sigma>0\,\bigg{|}\,g_{\epsilon}(\sigma)>\frac{1-\sqrt{\nu\epsilon}}{\lambda}\right\}<\inf\left\{\sigma>0\,\bigg{|}\,g_{\epsilon}(\sigma)>\frac{1}{\lambda}\right\}=S_{\epsilon,1}\,,

which are uniformly bounded above. The proof of Proposition 3.3 shows that the next blowup trigger time Sϵ,1S_{\epsilon,1} coincides with the first time gϵg_{\epsilon} attains 1/λ1/\lambda. Moreover, this first time corresponds to a crossing with a slope that is bounded away from zero. Specifically, we have σgϵ(Sϵ,1)bΛ/λ>0\partial_{\sigma}g_{\epsilon}(S_{\epsilon,1})\geq b_{\Lambda}/\lambda>0. This implies that limϵ0|Sϵ,1S~ϵ,1|=0\lim_{\epsilon\to 0}|S_{\epsilon,1}-\tilde{S}_{\epsilon,1}|=0. Then, by definition of S~ϵ,1\tilde{S}_{\epsilon,1}, we have Ψϵ=(1λgϵ)/νϵ/ν\Psi^{\prime}_{\epsilon}=(1-\lambda g_{\epsilon})/\nu\geq\sqrt{\epsilon/\nu} on [ξϵ,0,S~ϵ,1)[-\xi_{\epsilon,0},\tilde{S}_{\epsilon,1}), so that on [0,S~ϵ,1)[0,\tilde{S}_{\epsilon,1}), the corresponding backward-delay function satisfies:

ηϵ(σ)=σΦϵ(Ψϵ(σ)ϵ)=Φϵ(Ψϵ(σ))Φϵ(Ψϵ(σ)ϵ)\displaystyle\eta_{\epsilon}(\sigma)=\sigma-\Phi_{\epsilon}(\Psi_{\epsilon}(\sigma)-\epsilon)=\Phi_{\epsilon}(\Psi_{\epsilon}(\sigma))-\Phi_{\epsilon}(\Psi_{\epsilon}(\sigma)-\epsilon) \displaystyle\leq ϵ/Ψξϵ,0,S~ϵ,1νϵ.\displaystyle\epsilon/\|\Psi^{\prime}\|_{-\xi_{\epsilon,0},\tilde{S}_{\epsilon,1}}\leq\sqrt{\nu\epsilon}\,.

Thus, the backward-delay function ηϵ\eta_{\epsilon} and the backward function ξϵ\xi_{\epsilon} compactly converge toward zero and ξ=id\xi=\mathrm{id}, respectively, on [0,min(S1,S¯1))[0,\min(S_{1},\underline{S}_{1})) with S¯1=lim infϵ0+S~ϵ,1=lim infϵ0+Sϵ,1\underline{S}_{1}=\liminf_{\epsilon\to 0^{+}}\tilde{S}_{\epsilon,1}=\liminf_{\epsilon\to 0^{+}}S_{\epsilon,1}.

(ii)(ii) By integration by part of the renewal-type equation (26), the cumulative fluxes GϵG_{\epsilon} and GG respectively satisfy:

Gϵ(σ)\displaystyle G_{\epsilon}(\sigma) =\displaystyle= 0H(σ,x)qϵ,0(x)dxH(σ,Λ)Gϵ(ξϵ(0))+\displaystyle\int_{0}^{\infty}H(\sigma,x)q_{\epsilon,0}(x)\,\mathrm{d}x-H(\sigma,\Lambda)G_{\epsilon}(\xi_{\epsilon}(0))+
0σh(στ,Λ)Gϵ(ξϵ(τ))dτ,\displaystyle\hskip 40.0pt\int_{0}^{\sigma}h(\sigma-\tau,\Lambda)G_{\epsilon}(\xi_{\epsilon}(\tau))\,\mathrm{d}\tau\,,
G(σ)\displaystyle G(\sigma) =\displaystyle= 0H(σ,x)q0(x)dxH(σ,Λ)G(0)+\displaystyle\int_{0}^{\infty}H(\sigma,x)q_{0}(x)\,\mathrm{d}x-H(\sigma,\Lambda)G(0)+
0σh(στ,Λ)G(ξ(τ))dτ.\displaystyle\hskip 40.0pt\int_{0}^{\sigma}h(\sigma-\tau,\Lambda)G(\xi(\tau))\,\mathrm{d}\tau\,.

For all 0σmin(S1,S¯1)0\leq\sigma\leq\min(S_{1},\underline{S}_{1}), we have ξ(σ)=σ\xi(\sigma)=\sigma. Using the convention G(0)=0G(0)=0, this observation allows one to write

|Gϵ(σ)G(σ)|\displaystyle|G_{\epsilon}(\sigma)-G(\sigma)| \displaystyle\leq 0H(σ,x)|qϵ,0(x)q0(x)|dx+\displaystyle\int_{0}^{\infty}H(\sigma,x)\big{|}q_{\epsilon,0}(x)-q_{0}(x)\big{|}\,\mathrm{d}x+
H(σ,Λ)|Gϵ(ξϵ(0))|+0σh(στ,Λ)|Gϵ(ξϵ(τ))G(τ)|dτ,\displaystyle\hskip 20.0ptH(\sigma,\Lambda)|G_{\epsilon}(\xi_{\epsilon}(0))|+\int_{0}^{\sigma}h(\sigma-\tau,\Lambda)\big{|}G_{\epsilon}(\xi_{\epsilon}(\tau))-G(\tau)\big{|}\,\mathrm{d}\tau\,,
\displaystyle\leq H(,x)(0|qϵ,0(x)q0(x)|dx+|Gϵ(ξϵ(0))|)+\displaystyle\|H(\cdot,x)\|_{\infty}\left(\int_{0}^{\infty}\big{|}q_{\epsilon,0}(x)-q_{0}(x)\big{|}\,\mathrm{d}x+|G_{\epsilon}(\xi_{\epsilon}(0))|\right)+
0σh(στ,Λ)(|Gϵ(ξϵ(τ))Gϵ(τ)|+|Gϵ(τ)G(τ)|)dτ.\displaystyle\hskip 20.0pt\int_{0}^{\sigma}h(\sigma-\tau,\Lambda)\left(\big{|}G_{\epsilon}(\xi_{\epsilon}(\tau))-G_{\epsilon}(\tau)\big{|}+\big{|}G_{\epsilon}(\tau)-G(\tau)\big{|}\right)\,\mathrm{d}\tau\,.

By uniform boundedness of the instantaneous flux gϵAΛ\|g_{\epsilon}\|_{\infty}\leq A_{\Lambda}, we have

|Gϵ(ξϵ(0))|AΛϵand|Gϵ(ξϵ(τ))Gϵ(τ)|AΛ|ξϵ(τ)τ|=AΛ|ηϵ(τ)|.\displaystyle|G_{\epsilon}(\xi_{\epsilon}(0))|\leq A_{\Lambda}\epsilon\quad\mathrm{and}\quad\big{|}G_{\epsilon}(\xi_{\epsilon}(\tau))-G_{\epsilon}(\tau)\big{|}\leq A_{\Lambda}|\xi_{\epsilon}(\tau)-\tau\big{|}=A_{\Lambda}|\eta_{\epsilon}(\tau)\big{|}\,.

Then, using that H(,x)=1\|H(\cdot,x)\|_{\infty}=1 and the bound

0σh(στ,Λ)|Gϵ(τ)G(τ)|dτH(σ,Λ)GϵG0,σ,\displaystyle\int_{0}^{\sigma}h(\sigma-\tau,\Lambda)\big{|}G_{\epsilon}(\tau)-G(\tau)\big{|}\,\mathrm{d}\tau\leq H(\sigma,\Lambda)\|G_{\epsilon}-G\|_{0,\sigma}\,,

we deduce from inequality (4.3) and the bounds given above that

(1H(σ,Λ))GϵG0,σqϵ,0q01+AΛ(ϵ+H(σ,Λ)ηϵ0,σ).\displaystyle\big{(}1-H(\sigma,\Lambda)\big{)}\|G_{\epsilon}-G\|_{0,\sigma}\leq\|q_{\epsilon,0}-q_{0}\|_{1}+A_{\Lambda}\left(\epsilon+H(\sigma,\Lambda)\|\eta_{\epsilon}\|_{0,\sigma}\right)\,.

As for all σmin(S1,S¯1)\sigma\leq\min(S_{1},\underline{S}_{1}), we have H(σ,Λ)H(S1,Λ)<1H(\sigma,\Lambda)\leq H(S_{1},\Lambda)<1, the above inequality implies the compact convergence of GϵG_{\epsilon} towards GG on [0,min(S1,S¯1))[0,\min(S_{1},\underline{S}_{1})) when ϵ0+\epsilon\to 0^{+}. This can be extended to the uniform convergence over the closed interval [0,min(S1,S¯1)][0,\min(S_{1},\underline{S}_{1})] by uniform boundedness of gϵg_{\epsilon}.

(iii)(iii) Similar calculations but starting from the renewal-type equation (26) shows that the instantaneous flux gϵg_{\epsilon} converges uniformly toward gg, the instantaneous flux with zero refractory period on [0,min(S1,S¯1))[0,\min(S_{1},\underline{S}_{1})). Indeed, by integration by part of (26), we have:

gϵ(σ)\displaystyle g_{\epsilon}(\sigma) =\displaystyle= 0h(σ,x)qϵ,0(x)dxh(σ,Λ)Gϵ(ξϵ(0))+\displaystyle\int_{0}^{\infty}h(\sigma,x)q_{\epsilon,0}(x)\,\mathrm{d}x-h(\sigma,\Lambda)G_{\epsilon}(\xi_{\epsilon}(0))+
0σσh(στ,Λ)Gϵ(ξϵ(τ))dτ,\displaystyle\hskip 80.0pt\int_{0}^{\sigma}\partial_{\sigma}h(\sigma-\tau,\Lambda)G_{\epsilon}(\xi_{\epsilon}(\tau))\,\mathrm{d}\tau\,,
g(σ)\displaystyle g(\sigma) =\displaystyle= 0h(σ,x)q0(x)dxh(σ,Λ)G(ξ(0))+\displaystyle\int_{0}^{\infty}h(\sigma,x)q_{0}(x)\,\mathrm{d}x-h(\sigma,\Lambda)G(\xi(0))+
0σσh(στ,Λ)G(τ)dτ,\displaystyle\hskip 80.0pt\int_{0}^{\sigma}\partial_{\sigma}h(\sigma-\tau,\Lambda)G(\tau)\,\mathrm{d}\tau\,,

where we use that ξ(σ)=σ\xi(\sigma)=\sigma for σ<min(S1,S¯1)\sigma<\min(S_{1},\underline{S}_{1}). The integrability and uniform boundedness of the integration kernels appearing above directly imply the announced uniform convergence. Similar arguments can be made about the equations obtained by further differentiation of (26) with respect to σ\sigma, showing the uniform convergence of all derivatives of gϵg_{\epsilon} toward the corresponding derivatives of gg on [0,min(S1,S¯1)][0,\min(S_{1},\underline{S}_{1})] when ϵ0+\epsilon\to 0^{+}.

(iv)(iv) The observations above allow one to show (1)(1) in Proposition 4.2, or equivalently that

limϵ0+Sϵ,1=S1=inf{σ>0|g(σ)>1/λ}.\displaystyle\lim_{\epsilon\to 0^{+}}S_{\epsilon,1}=S_{1}=\inf\left\{\sigma>0\,\Big{|}\,g(\sigma)>1/\lambda\right\}\,.

This amounts to show S¯1=S¯1=S1\underline{S}_{1}=\overline{S}_{1}=S_{1}, where S¯1=lim infϵ0+Sϵ,1\underline{S}_{1}=\liminf_{\epsilon\to 0^{+}}S_{\epsilon,1} and S¯1=lim supϵ0+Sϵ,1\overline{S}_{1}=\limsup_{\epsilon\to 0^{+}}S_{\epsilon,1}. Let us first show that S1S¯1S_{1}\leq\underline{S}_{1}. By definition, there is a sequence {ϵn}n\{\epsilon_{n}\}_{n\in\mathbbm{N}} with ϵn0+\epsilon_{n}\to 0^{+} and such that:

limnSϵn=S¯1,gϵn(Sϵn,1)=1/λ,andσgϵn.(Sϵn,1)bΛ/λ>0.\displaystyle\lim_{n\to\infty}S_{\epsilon_{n}}=\underline{S}_{1}\,,\quad g_{\epsilon_{n}}(S_{\epsilon_{n},1})=1/\lambda\,,\quad\mathrm{and}\quad\partial_{\sigma}g_{\epsilon_{n}}\,.(S_{\epsilon_{n},1})\geq b_{\Lambda}/\lambda>0\,.

Thus, by the uniform convergence results of (iii)(iii), we have g(S¯1)=1/λg(\underline{S}_{1})=1/\lambda and σg(S¯1)bΛ/λ>0\partial_{\sigma}g(\underline{S}_{1})\geq b_{\Lambda}/\lambda>0, which implies that S¯1\underline{S}_{1} is a crossing time of gg at level 1/λ1/\lambda. As S1S_{1} is defined as the first such crossing time, we have S¯1S1\underline{S}_{1}\geq S_{1}. To show that S¯1=S1\underline{S}_{1}=S_{1}, let us consider S¯1S¯1S1\overline{S}_{1}\geq\underline{S}_{1}\geq S_{1} and let us show that S¯1=S1\overline{S}_{1}=S_{1}. Suppose S¯1>S1\overline{S}_{1}>S_{1}. Then, there is S1<σ<S¯1S_{1}<\sigma<\overline{S}_{1} such that g(σ)>1/λg(\sigma)>1/\lambda. Posit δ=g(σ)1/λ\delta=g(\sigma)-1/\lambda. By pointwise convergence in σ\sigma, there is e>0e>0 such that for all 0<ϵ<e0<\epsilon<e, |gϵ(σ)g(σ)|<δ/2|g_{\epsilon}(\sigma)-g(\sigma)|<\delta/2. Consider now a sequence {ϵn}n\{\epsilon_{n}\}_{n\in\mathbbm{N}} with ϵn0+\epsilon_{n}\to 0^{+} and such that limnSϵn,1=S¯\lim_{n\to\infty}S_{\epsilon_{n},1}=\overline{S}. For nn large enough, we have ϵn<e\epsilon_{n}<e so that

gϵn(σ)>g(σ)δ/2=(g(σ)+1/λ)/2>1/λ,\displaystyle g_{\epsilon_{n}}(\sigma)>g(\sigma)-\delta/2=(g(\sigma)+1/\lambda)/2>1/\lambda\,,

so that the first crossing Sϵn,1S_{\epsilon_{n},1} must happen before σ\sigma. But then Sϵn,1<σ<S¯1S_{\epsilon_{n},1}<\sigma<\overline{S}_{1} contradicts the definition of S¯1\overline{S}_{1} as the limit of Sϵn,1S_{\epsilon_{n},1}. Thus we must have S¯1=S¯1=limϵ0+Sϵ,1=S1\underline{S}_{1}=\overline{S}_{1}=\lim_{\epsilon\to 0^{+}}S_{\epsilon,1}=S_{1}. This shows (1)(1) in Proposition 4.2.

(v)(v) Finally, observe that (2)(2) in Proposition 4.2 will directly follows from extending the uniform convergence results shown in (ii)(ii) and (iii)(iii) to [0,S1+λ/2][0,S_{1}+\lambda/2]. To prove such an extension, observe that by Proposition 3.5, the full blowup triggered in Sϵ,1S_{\epsilon,1} with refractory period ϵ>0\epsilon>0 and in S1S_{1} for zero refractory period all have size πϵ,1>1/2\pi_{\epsilon,1}>1/2. Thus, the backward functions ξϵ\xi_{\epsilon} and ξ\xi must be constant on [Sϵ,1,Sϵ,1+λ/2][S_{\epsilon,1},S_{\epsilon,1}+\lambda/2] and [S1,S1+λ/2][S_{1},S_{1}+\lambda/2], so that we have:

S1σSϵ,1+λ/2:Gϵ(σ)\displaystyle S_{1}\leq\sigma\leq S_{\epsilon,1}+\lambda/2:\quad G_{\epsilon}(\sigma) =\displaystyle= 0H(σ,x)qϵ,0(x)dxH(σ,Λ)Gϵ(ξϵ(0))+\displaystyle\int_{0}^{\infty}H(\sigma,x)q_{\epsilon,0}(x)\,\mathrm{d}x-H(\sigma,\Lambda)G_{\epsilon}(\xi_{\epsilon}(0))+
0min(σ,S1,ϵ)h(στ,Λ)Gϵ(ξϵ(τ))dτ,\displaystyle\hskip 40.0pt\int_{0}^{\min(\sigma,S_{1,\epsilon})}h(\sigma-\tau,\Lambda)G_{\epsilon}(\xi_{\epsilon}(\tau))\,\mathrm{d}\tau\,,
S1σS1+λ/2:G(σ)\displaystyle S_{1}\leq\sigma\leq S_{1}+\lambda/2:\quad G(\sigma) =\displaystyle= 0H(σ,x)q0(x)dxH(σ,Λ)G(0)+\displaystyle\int_{0}^{\infty}H(\sigma,x)q_{0}(x)\,\mathrm{d}x-H(\sigma,\Lambda)G(0)+
0S1h(στ,Λ)G(τ)dτ,\displaystyle\hskip 40.0pt\int_{0}^{S_{1}}h(\sigma-\tau,\Lambda)G(\tau)\,\mathrm{d}\tau\,,

where we use that ξ=id\xi=\mathrm{id} for all 0σS10\leq\sigma\leq S_{1}. From there, given (ii)(ii), the only additional point to check, is that for all 0σS1+λ/20\leq\sigma\leq S_{1}+\lambda/2, the extra term

I(σ)=S1min(σ,S1,ϵ)h(στ,Λ)Gϵ(ξϵ(τ))dτ\displaystyle I(\sigma)=\int_{S_{1}}^{\min(\sigma,S_{1,\epsilon})}h(\sigma-\tau,\Lambda)G_{\epsilon}(\xi_{\epsilon}(\tau))\,\mathrm{d}\tau

vanishes uniformly over [S1,S1+λ/3][S_{1},S_{1}+\lambda/3]. This directly follows from the uniform boundedness of GϵG_{\epsilon}:

I(σ)\displaystyle I(\sigma) \displaystyle\leq h(,Λ)GϵξϵS1,S1,ϵ|Sϵ,1S1|\displaystyle\|h(\cdot,\Lambda)\|_{\infty}\|G_{\epsilon}\circ\xi_{\epsilon}\|_{S_{1},S_{1,\epsilon}}|S_{\epsilon,1}-S_{1}|\,
\displaystyle\leq h(,Λ)AΛ(S1+λ/3)|Sϵ,1S1|ϵ0+0.\displaystyle\|h(\cdot,\Lambda)\|_{\infty}A_{\Lambda}(S_{1}+\lambda/3)|S_{\epsilon,1}-S_{1}|\xrightarrow[\epsilon\to 0^{+}]{}0\,.

4.4 Continuity at blowup trigger time

Proof of Proposition 4.3.

By Lemma A.1, there is a constant s1s_{1} independent of ϵ\epsilon such that 0ξϵ(σ)10\leq\xi_{\epsilon}^{\prime}(\sigma)\leq 1 for all Sϵ,1s1/2σSϵ,1+λ/2S_{\epsilon,1}-s_{1}/2\leq\sigma\leq S_{\epsilon,1}+\lambda/2. Introducing the real number U=S1s1/3U=S_{1}-s_{1}/3 and using that

q(Sϵ,1,y)\displaystyle q(S_{\epsilon,1},y) =\displaystyle= 0κ(Sϵ,1,y,x)pϵ,0(x)dx+0Sϵ,1κ(Sϵ,1τ,y,Λ)dGϵ(ξϵ(τ)),\displaystyle\int_{0}^{\infty}\kappa(S_{\epsilon,1},y,x)p_{\epsilon,0}(x)\,\mathrm{d}x+\int_{0}^{S_{\epsilon,1}}\kappa(S_{\epsilon,1}-\tau,y,\Lambda)\,\mathrm{d}G_{\epsilon}(\xi_{\epsilon}(\tau))\,,
q(S1,y)\displaystyle q(S_{1},y) =\displaystyle= 0κ(S1,y,x)p0(x)dx+0S1κ(S1τ,y,Λ)dG(τ),\displaystyle\int_{0}^{\infty}\kappa(S_{1},y,x)p_{0}(x)\,\mathrm{d}x+\int_{0}^{S_{1}}\kappa(S_{1}-\tau,y,\Lambda)\,\mathrm{d}G(\tau)\,,

we upper bound Δq=q(Sϵ,1,)q(S1,)1\Delta q=\|q(S_{\epsilon,1},\cdot)-q(S_{1},\cdot)\|_{1} as

ΔqΔq1+Δq2+Δq3+Δq4+Δq5,\displaystyle\Delta q\leq\Delta q_{1}+\Delta q_{2}+\Delta q_{3}+\Delta q_{4}+\Delta q_{5}\,,

where the terms Δq1\Delta q_{1}, Δq2\Delta q_{2}, Δq3\Delta q_{3}, Δq4\Delta q_{4}, and Δq5\Delta q_{5} are defined as follows:

Δq1=0|0κ(S1,ϵ,y,x)pϵ,0(x)dx0κ(S1,y,x)p0(x)dx|dy,\displaystyle\Delta q_{1}=\int_{0}^{\infty}\bigg{|}\int_{0}^{\infty}\kappa(S_{1,\epsilon},y,x)p_{\epsilon,0}(x)\,\mathrm{d}x-\int_{0}^{\infty}\kappa(S_{1},y,x)p_{0}(x)\,\mathrm{d}x\bigg{|}\,\mathrm{d}y\,,
Δq2=0|0Uκ(S1τ,y,Λ)dGϵ(ξϵ(τ))0Uκ(S1τ,y,Λ)dG(τ)|dy,\displaystyle\Delta q_{2}=\int_{0}^{\infty}\bigg{|}\int_{0}^{U}\kappa(S_{1}-\tau,y,\Lambda)\,\mathrm{d}G_{\epsilon}(\xi_{\epsilon}(\tau))-\int_{0}^{U}\kappa(S_{1}-\tau,y,\Lambda)\,\mathrm{d}G(\tau)\bigg{|}\,\mathrm{d}y\,,
Δq3=0|0Uκ(Sϵ,1τ,y,Λ)dGϵ(ξϵ(τ))0Uκ(S1τ,y,Λ)dGϵ(ξϵ(τ))|dy,\displaystyle\Delta q_{3}=\int_{0}^{\infty}\bigg{|}\int_{0}^{U}\kappa(S_{\epsilon,1}-\tau,y,\Lambda)\,\mathrm{d}G_{\epsilon}(\xi_{\epsilon}(\tau))-\int_{0}^{U}\kappa(S_{1}-\tau,y,\Lambda)\,\mathrm{d}G_{\epsilon}(\xi_{\epsilon}(\tau))\bigg{|}\,\mathrm{d}y\,,
Δq4=0|US1κ(Sϵ,1τ,y,Λ)dGϵ(ξϵ(τ))US1κ(S1τ,y,Λ)dG(τ)|dy,\displaystyle\Delta q_{4}=\int_{0}^{\infty}\bigg{|}\int_{U}^{S_{1}}\kappa(S_{\epsilon,1}-\tau,y,\Lambda)\,\mathrm{d}G_{\epsilon}(\xi_{\epsilon}(\tau))-\int_{U}^{S_{1}}\kappa(S_{1}-\tau,y,\Lambda)\,\mathrm{d}G(\tau)\bigg{|}\,\mathrm{d}y\,,
Δq5=0|S1Sϵ,1κ(Sϵ,1τ,y,Λ)dGϵ(ξϵ(τ))|dy.\displaystyle\Delta q_{5}=\int_{0}^{\infty}\bigg{|}\int_{S_{1}}^{S_{\epsilon,1}}\kappa(S_{\epsilon,1}-\tau,y,\Lambda)\,\mathrm{d}G_{\epsilon}(\xi_{\epsilon}(\tau))\bigg{|}\,\mathrm{d}y\,.

Our goal is to show that the terms Δq2\Delta q_{2}, Δq3\Delta q_{3}, Δq4\Delta q_{4}, and Δq5\Delta q_{5} all vanish in the limit ϵ0+\epsilon\to 0^{+}. The nonrenewal term Δq1\Delta q_{1} can be bounded leveraging the fact limϵ0+Sϵ,1=S1>0\lim_{\epsilon\to 0^{+}}S_{\epsilon,1}=S_{1}>0, which allows one to use the bound provided by Lemma B.1:

Δq1\displaystyle\Delta q_{1} 0|0κ(S1,ϵ,y,x)(pϵ,0(x)p0(x))dx|dy+\displaystyle\leq\int_{0}^{\infty}\bigg{|}\int_{0}^{\infty}\kappa(S_{1,\epsilon},y,x)\big{(}p_{\epsilon,0}(x)-p_{0}(x)\big{)}\,\mathrm{d}x\bigg{|}\,\mathrm{d}y\;+
0|0(κ(S1,ϵ,y,x)κ(S1,y,x))p0(x)dx|dy,\displaystyle\hskip 20.0pt\int_{0}^{\infty}\bigg{|}\int_{0}^{\infty}\big{(}\kappa(S_{1,\epsilon},y,x)-\kappa(S_{1},y,x)\big{)}p_{0}(x)\,\mathrm{d}x\bigg{|}\,\mathrm{d}y\,,
0(0κ(S1,ϵ,y,x)dy)|pϵ,0(x)p0(x)|dx+\displaystyle\leq\int_{0}^{\infty}\left(\int_{0}^{\infty}\kappa(S_{1,\epsilon},y,x)\,\mathrm{d}y\right)\big{|}p_{\epsilon,0}(x)-p_{0}(x)\big{|}\,\mathrm{d}x\;+
0(0supS1/2σS1+λ/3|σκ(σ,y,x)|dy)|S1,ϵS1|p0(x)dx,\displaystyle\hskip 20.0pt\int_{0}^{\infty}\left(\int_{0}^{\infty}\sup_{S_{1}/2\leq\sigma\leq S_{1}+\lambda/3}|\partial_{\sigma}\kappa(\sigma,y,x)|\,\mathrm{d}y\right)|S_{1,\epsilon}-S_{1}|p_{0}(x)\,\mathrm{d}x\,,
pϵ,0p01+MS1/2,S1+λ/3|S1,ϵS1|.\displaystyle\leq\|p_{\epsilon,0}-p_{0}\|_{1}+M_{S_{1}/2,S_{1}+\lambda/3}|S_{1,\epsilon}-S_{1}|\,.

Thus limϵ0+Δq1=0\lim_{\epsilon\to 0^{+}}\Delta q_{1}=0 directly follows from the L1L_{1} convergence pϵ,0p010\|p_{\epsilon,0}-p_{0}\|_{1}\to 0 as ϵ0+\epsilon\to 0^{+}. The bound provided by Lemma B.1 can be similarly used for the renewal term Δq2\Delta q_{2} which has been defined up to time U<S1U<S_{1}. To see this, let us use integration by parts on [0,U][0,U] to write

Δq2\displaystyle\Delta q_{2} 0|0Uσκ(S1τ,y,Λ)(Gϵ(ξϵ(τ))G(τ))dτ|dy+\displaystyle\leq\int_{0}^{\infty}\Bigg{|}\int_{0}^{U}\partial_{\sigma}\kappa(S_{1}-\tau,y,\Lambda)\big{(}G_{\epsilon}(\xi_{\epsilon}(\tau))-G(\tau)\Big{)}\,\mathrm{d}\tau\bigg{|}\,\mathrm{d}y+
0|[κ(S1τ,y,Λ)(Gϵ(ξϵ(τ))G(τ))]0U|dy,\displaystyle\hskip 20.0pt\int_{0}^{\infty}\bigg{|}\Big{[}\kappa(S_{1}-\tau,y,\Lambda)\big{(}G_{\epsilon}(\xi_{\epsilon}(\tau))-G(\tau)\big{)}\Big{]}_{0}^{U}\bigg{|}\,\mathrm{d}y\,,
GϵξϵG0,U0U(0|σκ(S1τ,y,Λ)|dy)dτ+\displaystyle\leq\|G_{\epsilon}\circ\xi_{\epsilon}-G\|_{0,U}\int_{0}^{U}\left(\int_{0}^{\infty}\big{|}\partial_{\sigma}\kappa(S_{1}-\tau,y,\Lambda)\big{|}\,\mathrm{d}y\right)\,\mathrm{d}\tau+
|Gϵ(ξϵ(U))G(U)|+|Gϵ(ξϵ(0))G(0)|.\displaystyle\hskip 20.0pt\Big{|}G_{\epsilon}(\xi_{\epsilon}(U))-G(U)\Big{|}+\Big{|}G_{\epsilon}(\xi_{\epsilon}(0))-G(0)\Big{|}\,.

Using the bound from Lemma B.1, we get

Δq2(2+MS1U,S1)GϵξϵG0,U,\displaystyle\Delta q_{2}\leq(2+M_{S_{1}-U,S_{1}})\,\|G_{\epsilon}\circ\xi_{\epsilon}-G\|_{0,U}\,,

and limϵ0+Δq2=0\lim_{\epsilon\to 0^{+}}\Delta q_{2}=0 follows from the uniform convergence of GϵξϵG_{\epsilon}\circ\xi_{\epsilon} toward GG on [0,S1+λ/3][0,S_{1}+\lambda/3] given in Proposition 4.2. The boundary renewal term Δq3\Delta q_{3} can also be bounded using the fact that Sϵ,1>(S1+U)/2>US_{\epsilon,1}>(S_{1}+U)/2>U for small enough ϵ>0\epsilon>0:

Δq3\displaystyle\Delta q_{3} 0U(0sup(S1+U)/2σS1+λ/3|σκ(στ,y,Λ)|dy)|Sϵ,1S1|dGϵ(ξϵ(τ))\displaystyle\leq\int_{0}^{U}\left(\int_{0}^{\infty}\sup_{(S_{1}+U)/2\leq\sigma\leq S_{1}+\lambda/3}\big{|}\partial_{\sigma}\kappa(\sigma-\tau,y,\Lambda)\big{|}\,\mathrm{d}y\right)|S_{\epsilon,1}-S_{1}|\,\mathrm{d}G_{\epsilon}(\xi_{\epsilon}(\tau))
M(S1U)/2,S1+λ.3(Gϵ(ξϵ(U))Gϵ(ξϵ(0)))|Sϵ,1S1|.\displaystyle\leq M_{(S_{1}-U)/2,S_{1}+\lambda.3}\big{(}G_{\epsilon}(\xi_{\epsilon}(U))-G_{\epsilon}(\xi_{\epsilon}(0))\big{)}|S_{\epsilon,1}-S_{1}|\,.

We conclude that limϵ0+Δq3=0\lim_{\epsilon\to 0^{+}}\Delta q_{3}=0 by boundedness of the cumulative fluxes:

|Gϵ(ξϵ(U))Gϵ(ξϵ(0))||Gϵ(ξϵ(U))|+|Gϵ(ξϵ(0))|Gϵ(U)+1AΛU+1.\displaystyle|G_{\epsilon}(\xi_{\epsilon}(U))-G_{\epsilon}(\xi_{\epsilon}(0))|\leq|G_{\epsilon}(\xi_{\epsilon}(U))|+|G_{\epsilon}(\xi_{\epsilon}(0))|\leq G_{\epsilon}(U)+1\leq A_{\Lambda}U+1\,.

The remaining terms Δq4\Delta q_{4} and Δq5\Delta q_{5} collect the contribution of those processes that reset at the vicinity of the blowup trigger times S1S_{1} and Sϵ,1S_{\epsilon,1}. Bounding these terms will crucially rely on the fact that under Assumption 4, the backward function ξϵ\xi_{\epsilon} can be shown to be well behaved on [U,S1][U,S_{1}] in the sense that by Lemma A.1, we have 0ξϵ10\leq\xi_{\epsilon}^{\prime}\leq 1. Bearing in mind this preliminary remark, observe that for all small enough δ>0\delta>0, we have

Δq4\displaystyle\Delta q_{4} US1(0κ(Sϵ,1τ,y,Λ)dy)|dGϵ(ξϵ(τ))dG(τ)|+\displaystyle\leq\int_{U}^{S_{1}}\left(\int_{0}^{\infty}\kappa(S_{\epsilon,1}-\tau,y,\Lambda)\,\mathrm{d}y\right)\,\big{|}\mathrm{d}G_{\epsilon}(\xi_{\epsilon}(\tau))-\mathrm{d}G(\tau)\big{|}+
0US1|κ(Sϵ,1τ,y,Λ)κ(S1τ,y,Λ)|dG(τ)dy,\displaystyle\hskip 20.0pt\int_{0}^{\infty}\int_{U}^{S_{1}}\big{|}\kappa(S_{\epsilon,1}-\tau,y,\Lambda)-\kappa(S_{1}-\tau,y,\Lambda)\big{|}\,\mathrm{d}G(\tau)\,\mathrm{d}y\,,
US1|dGϵ(ξϵ(τ))dG(τ)|+\displaystyle\leq\int_{U}^{S_{1}}\big{|}\mathrm{d}G_{\epsilon}(\xi_{\epsilon}(\tau))-\mathrm{d}G(\tau)\big{|}+
0S1δS1|κ(Sϵ,1τ,y,Λ)κ(S1τ,y,Λ)|dG(τ)dy+\displaystyle\hskip 20.0pt\int_{0}^{\infty}\int_{S_{1}-\delta}^{S_{1}}\big{|}\kappa(S_{\epsilon,1}-\tau,y,\Lambda)-\kappa(S_{1}-\tau,y,\Lambda)\big{|}\,\mathrm{d}G(\tau)\,\mathrm{d}y\;+
0US1δ|κ(Sϵ,1τ,y,Λ)κ(S1τ,y,Λ)|dG(τ)dy,\displaystyle\hskip 20.0pt\int_{0}^{\infty}\int_{U}^{S_{1}-\delta}\big{|}\kappa(S_{\epsilon,1}-\tau,y,\Lambda)-\kappa(S_{1}-\tau,y,\Lambda)\big{|}\,\mathrm{d}G(\tau)\,\mathrm{d}y\,,
US1|ξϵ(τ)gϵ(ξϵ(τ))g(τ)|dτ+\displaystyle\leq\int_{U}^{S_{1}}\big{|}\xi_{\epsilon}^{\prime}(\tau)g_{\epsilon}(\xi_{\epsilon}(\tau))-g(\tau)\big{|}\,\mathrm{d}\tau+
gδ(0|κ(Sϵ,1τ,y,Λ)|+|κ(S1τ,y,Λ)|dy)+\displaystyle\hskip 20.0pt\|g\|_{\infty}\delta\left(\int_{0}^{\infty}\big{|}\kappa(S_{\epsilon,1}-\tau,y,\Lambda)\big{|}+\big{|}\kappa(S_{1}-\tau,y,\Lambda)\big{|}\,\mathrm{d}y\right)\;+
Mδ/2,S1+λ/3(G(S1δ)G(U))|Sϵ,1S1|,\displaystyle\hskip 20.0ptM_{\delta/2,S_{1}+\lambda/3}\big{(}G(S_{1}-\delta)-G(U)\big{)}|S_{\epsilon,1}-S_{1}|,
US1|ξϵ(τ)gϵ(ξϵ(τ))g(τ)|dτ+2AΛδ+Mδ/2,S1+λ/3AΛS1|Sϵ,1S1|.\displaystyle\leq\int_{U}^{S_{1}}\big{|}\xi_{\epsilon}^{\prime}(\tau)g_{\epsilon}(\xi_{\epsilon}(\tau))-g(\tau)\big{|}\,\mathrm{d}\tau+2A_{\Lambda}\delta+M_{\delta/2,S_{1}+\lambda/3}A_{\Lambda}S_{1}|S_{\epsilon,1}-S_{1}|\,.

By Lemma A.1, ξϵ(τ)gϵ(ξϵ(τ))AΛ\xi^{\prime}_{\epsilon}(\tau)g_{\epsilon}(\xi_{\epsilon}(\tau))\leq A_{\Lambda} over [U,S1][U,S_{1}], whereas for all σ\sigma in [U,S1)[U,S_{1}), we have the pointwise convergence

ξϵ(σ)=Ψϵ(σ)Ψϵ(ξϵ(σ))=1λgϵ(σ)1λgϵ(ξϵ(σ))ϵ0+1.\displaystyle\xi_{\epsilon}^{\prime}(\sigma)=\frac{\Psi_{\epsilon}^{\prime}(\sigma)}{\Psi_{\epsilon}^{\prime}(\xi_{\epsilon}(\sigma))}=\frac{1-\lambda g_{\epsilon}(\sigma)}{1-\lambda g_{\epsilon}(\xi_{\epsilon}(\sigma))}\xrightarrow[\epsilon\to 0^{+}]{}1\,.

Thus by dominated convergence, we have

limϵ0+US1|ξ(τ)gϵ(ξϵ(τ))g(τ)|dτ=0.\displaystyle\lim_{\epsilon\to 0^{+}}\int_{U}^{S_{1}}\big{|}\xi^{\prime}(\tau)g_{\epsilon}(\xi_{\epsilon}(\tau))-g(\tau)\big{|}\,\mathrm{d}\tau=0\,.

Then, as limϵ0+Sϵ,1=S1\lim_{\epsilon\to 0^{+}}S_{\epsilon,1}=S_{1}, we deduce that lim supϵ0+Δq42AΛδ\limsup_{\epsilon\to 0^{+}}\Delta q_{4}\leq 2A_{\Lambda}\delta for all small enough δ>0\delta>0, which implies that limϵ0+Δq4=0\lim_{\epsilon\to 0^{+}}\Delta q_{4}=0. To bound the final term Δq5\Delta q_{5}, we also exploit that ξϵ\xi_{\epsilon}^{\prime} is bounded by one on [S1s1/λ,S1+λ/3][S_{1}-s_{1}/\lambda,S_{1}+\lambda/3] so that

Δq5\displaystyle\Delta q_{5} =\displaystyle= 0|S1Sϵ,1κ(Sϵ,1τ,y,Λ)ξϵ(τ)gϵ(ξϵ(τ))dτ|dy.\displaystyle\int_{0}^{\infty}\bigg{|}\int_{S_{1}}^{S_{\epsilon,1}}\kappa(S_{\epsilon,1}-\tau,y,\Lambda)\xi^{\prime}_{\epsilon}(\tau)g_{\epsilon}(\xi_{\epsilon}(\tau))\,\mathrm{d}\tau\bigg{|}\,\mathrm{d}y\,.
\displaystyle\leq S1Sϵ,1(0κ(Sϵ,1τ,y,Λ)dy)gϵ(ξϵ(τ))dτ.\displaystyle\int_{S_{1}}^{S_{\epsilon,1}}\left(\int_{0}^{\infty}\kappa(S_{\epsilon,1}-\tau,y,\Lambda)\,\mathrm{d}y\right)g_{\epsilon}(\xi_{\epsilon}(\tau))\,\mathrm{d}\tau\,.
\displaystyle\leq S1Sϵ,1gϵ(ξϵ(τ))dτAΛSϵ,1S1|ϵ0+0.\displaystyle\int_{S_{1}}^{S_{\epsilon,1}}g_{\epsilon}(\xi_{\epsilon}(\tau))\,\mathrm{d}\tau\leq A_{\Lambda}\|S_{\epsilon,1}-S_{1}|\xrightarrow[\epsilon\to 0^{+}]{}0\,.

4.5 Continuity at blowup exit time

Proof of Proposition 4.4.

The jump size π1\pi_{1} is defined as the solution of the equation (π1,q(S1,))=0\mathcal{F}(\pi_{1},q(S_{1},\cdot))=0, where \mathcal{F} is the map

[p,q]=p0+H(λp,x)q(x)dx.\displaystyle\mathcal{F}[p,q]=p-\int_{0^{+}}^{\infty}H(\lambda p,x)q(x)\,\mathrm{d}x\,.

By smoothness of H(,x)H(\cdot,x) and by linearity with respect to qq, the mapping \mathcal{F} is continuously Fréchet differentiable on ×([0,))\mathbbm{R}\times\mathcal{M}([0,\infty)) where ([0,))\mathcal{M}([0,\infty)) is equipped with the L1L_{1} norm. Moreover, under Assumption 4, we have

p[π1,q(S1,)]=1λ0+h(λπ1,x)q(S1,x)dx=1λg(U1),\displaystyle\partial_{p}\mathcal{F}[\pi_{1},q(S_{1},\cdot)]=1-\lambda\int_{0^{+}}^{\infty}h(\lambda\pi_{1},x)q(S_{1},x)\,\mathrm{d}x=1-\lambda g(U_{1})\,,

where g(U1)g(U_{1}) denotes the instantaneous flux at the blowup exit time U1=S1+λπ1U_{1}=S_{1}+\lambda\pi_{1}. By Proposition 3.1, under Assumption 4, we have g(U1)1/(4λ)g(U_{1})\leq 1/(4\lambda) so that p[π1,q(S1,)]>3/4\partial_{p}\mathcal{F}[\pi_{1},q(S_{1},\cdot)]>3/4 is bounded away from zero. Therefore, by the Banach-space version of the implicit function theorem, there is a neighborhood of q(S1,)q(S_{1},\cdot) such that the mapping qπ1[q]q\mapsto\pi_{1}[q] is Fréchet differentiable. Moreover, under Assumption 4, we necessarily have that π1[q]>1/2\pi_{1}[q]>1/2 is bounded away from zero. The latter observation allows one to exploit the continuity of the mapping qπ1[q]q\mapsto\pi_{1}[q] in a neighborhood 𝒩\mathcal{N} of q0(S1,)q_{0}(S_{1},\cdot) to show the continuity of the map

q𝒢[q]=0κ(λπ1[q],,x)q(x)dx,\displaystyle q\mapsto\mathcal{G}[q]=\int_{0}^{\infty}\kappa(\lambda\pi_{1}[q],\cdot,x)q(x)\,\mathrm{d}x\,,

on that same neighborhood with respect to the L1L_{1} norm on ([0,))\mathcal{M}([0,\infty)). To prove this point, consider qaq_{a} and qbq_{b} in 𝒩\mathcal{N} and introduce the shorthand notations πa=π1[qa]>1/2\pi_{a}=\pi_{1}[q_{a}]>1/2 and πb=π1[qb]>1/2\pi_{b}=\pi_{1}[q_{b}]>1/2. We have

𝒢[qa]𝒢[qb]1\displaystyle\|\mathcal{G}[q_{a}]-\mathcal{G}[q_{b}]\|_{1} =\displaystyle= 0|0(κ(λπa,y,x)qa(x)κ(λπb,y,x))qb(x)dx|dy,\displaystyle\int_{0}^{\infty}\bigg{|}\int_{0}^{\infty}\big{(}\kappa(\lambda\pi_{a},y,x)q_{a}(x)-\kappa(\lambda\pi_{b},y,x)\big{)}q_{b}(x)\,\mathrm{d}x\bigg{|}\,\mathrm{d}y\,,
\displaystyle\leq 00κ(λπa,y,x)|qa(x)qb(x)|dxdy\displaystyle\int_{0}^{\infty}\int_{0}^{\infty}\kappa(\lambda\pi_{a},y,x)\big{|}q_{a}(x)-q_{b}(x)\big{|}\,\mathrm{d}x\,\mathrm{d}y
+00|κ(λπa,y,x)κ(λπb,y,x))|qb(x)dxdy,\displaystyle+\int_{0}^{\infty}\int_{0}^{\infty}\big{|}\kappa(\lambda\pi_{a},y,x)-\kappa(\lambda\pi_{b},y,x))\big{|}q_{b}(x)\,\mathrm{d}x\,\mathrm{d}y\,,
\displaystyle\leq 0(1H(λπa,x))|qa(x)qb(x)|dx\displaystyle\int_{0}^{\infty}(1-H(\lambda\pi_{a},x))\big{|}q_{a}(x)-q_{b}(x)\big{|}\,\mathrm{d}x
+00(λπaλπb|σκ(σ,y,x)|dσ)qb(x)dxdy,\displaystyle+\int_{0}^{\infty}\int_{0}^{\infty}\left(\int_{\lambda\pi_{a}}^{\lambda\pi_{b}}\big{|}\partial_{\sigma}\kappa(\sigma,y,x)\big{|}\,\mathrm{d}\sigma\right)q_{b}(x)\,\mathrm{d}x\,\mathrm{d}y\,,\
\displaystyle\leq qaqb1\displaystyle\|q_{a}-q_{b}\|_{1}\,
+(0(0supλπaσλπb|σκ(σ,y,x)|dy)qb(x)dx)λ|πbπa|.\displaystyle+\left(\int_{0}^{\infty}\left(\int_{0}^{\infty}\sup_{\lambda\pi_{a}\leq\sigma\leq\lambda\pi_{b}}\big{|}\partial_{\sigma}\kappa(\sigma,y,x)\big{|}\,\mathrm{d}y\right)q_{b}(x)\,\mathrm{d}x\right)\lambda|\pi_{b}-\pi_{a}|\,.

As 0<1/2<πa,πb<1/2<10<1/2<\pi_{a},\pi_{b}<1/2<1, by Lemma B.1, there exists a finite upper bound

Mab,λ=supx>0(0supλπaσλπb|σκ(σ,y,x)|dy)<.\displaystyle M_{ab,\lambda}=\sup_{x>0}\left(\int_{0}^{\infty}\sup_{\lambda\pi_{a}\leq\sigma\leq\lambda\pi_{b}}|\partial_{\sigma}\kappa(\sigma,y,x)|\,\mathrm{d}y\right)<\infty\,.

For fixed λ\lambda, the desired continuity result follows from the continuity of the map qπ1[q]q\mapsto\pi_{1}[q] via

𝒢[qa]𝒢[qb]1qaqb1+Mabλ|π1[qa]π1[qb]|.\displaystyle\|\mathcal{G}[q_{a}]-\mathcal{G}[q_{b}]\|_{1}\leq\|q_{a}-q_{b}\|_{1}+M_{ab}\lambda\big{|}\pi_{1}[q_{a}]-\pi_{1}[q_{b}]\big{|}\,.

4.6 Continuity at blowup reset time

Proof of Proposition 4.5.

Under Assumption 4, the reset dPMF dynamics resumes after the blowup exit time Uϵ,1U_{\epsilon,1} until the next blowup time Sϵ,2S_{\epsilon,2}. By the well ordered property of blowups, the reset time Rϵ,1R_{\epsilon,1} is such that Rϵ,1<Sϵ,2R_{\epsilon,1}<S_{\epsilon,2} and the backward-delay function ξϵ\xi_{\epsilon} is smooth on [Uϵ,1,Rϵ,1)[U_{\epsilon,1},R_{\epsilon,1}). Thus, the distribution of surviving processes just before reset is given by

qϵ(Rϵ,1,y)\displaystyle q_{\epsilon}(R_{\epsilon,1}^{-},y) =\displaystyle= 0κ(Rϵ,1Uϵ,1,y,x)qϵ(Uϵ,1,x)dx+\displaystyle\int_{0}^{\infty}\kappa(R_{\epsilon,1}^{-}-U_{\epsilon,1},y,x)q_{\epsilon}(U_{\epsilon,1},x)\,\mathrm{d}x+ (49)
Uϵ,1Rϵ,1κ(Rϵ,1σ,y,Λ)ξϵ(σ)gϵ(ξϵ(σ))dσ.\displaystyle\quad\int_{U_{\epsilon,1}}^{R_{\epsilon,1}^{-}}\kappa(R_{\epsilon,1}^{-}-\sigma,y,\Lambda)\xi^{\prime}_{\epsilon}(\sigma)g_{\epsilon}(\xi_{\epsilon}(\sigma))\,\mathrm{d}\sigma\,.

(i)(i) We first show that the L1L_{1} norm of the second integral term in (49) due to reset processes vanishes in the limit ϵ0+\epsilon\to 0^{+}. By definition of the blowup trigger and reset times, when σRϵ,1\sigma\to R_{\epsilon,1}^{-}, we have ξϵ(σ)Sϵ,1\xi_{\epsilon}(\sigma)\to S_{\epsilon,1}^{-}. Moreover, the inverse time change Ψϵ\Psi_{\epsilon} is differentiable at Sϵ,1S_{\epsilon,1} with derivative Ψϵ(Sϵ,1)3/(4ν)\Psi_{\epsilon}^{\prime}(S_{\epsilon,1})\leq 3/(4\nu), whereas the time change Ψϵ\Psi_{\epsilon} becomes locally flat with Ψϵ(Sϵ,1)=0\Psi^{\prime}_{\epsilon}(S_{\epsilon,1})=0 and Ψϵ′′(Sϵ,1)bΛ/(2ν)\Psi^{\prime\prime}_{\epsilon}(S_{\epsilon,1})\leq-b_{\Lambda}/(2\nu). The latter observations imply that for σRϵ,1\sigma\leq R_{\epsilon,1}, we have

Ψϵ(σ)ϵ\displaystyle\Psi_{\epsilon}(\sigma)-\epsilon =\displaystyle= Ψϵ(Rϵ,1)ϵ+Ψϵ(Rϵ,1)(σRϵ,1)2+O(σRϵ,1)2),\displaystyle\Psi_{\epsilon}(R_{\epsilon,1})-\epsilon+\Psi^{\prime}_{\epsilon}(R_{\epsilon,1})(\sigma-R_{\epsilon,1})^{2}+O\big{(}\sigma-R_{\epsilon,1})^{2}\big{)}\,, (50)
Ψϵ(ξϵ(σ))\displaystyle\Psi_{\epsilon}(\xi_{\epsilon}(\sigma)) =\displaystyle= Ψϵ(Sϵ,1)+Ψϵ′′(Sϵ,1)2(ξϵ(σ)Sϵ,1)2+O((ξϵ(σ)Sϵ,1)3).\displaystyle\Psi_{\epsilon}(S_{\epsilon,1})+\frac{\Psi^{\prime\prime}_{\epsilon}(S_{\epsilon,1})}{2}(\xi_{\epsilon}(\sigma)-S_{\epsilon,1})^{2}+O\big{(}(\xi_{\epsilon}(\sigma)-S_{\epsilon,1})^{3}\big{)}\,. (51)

Using the definition ξϵ(σ)=Φϵ(Ψϵ(σ)ϵ)\xi_{\epsilon}(\sigma)=\Phi_{\epsilon}(\Psi_{\epsilon}(\sigma)-\epsilon) to equate the two quantities above leads to the asymptotic scaling

ξϵ(σ)Sϵ,12Ψϵ(Rϵ,1)(Rϵ,1σ)|Ψϵ′′(Sϵ,1)|,\displaystyle\xi_{\epsilon}(\sigma)-S_{\epsilon,1}\sim\sqrt{\frac{2\Psi^{\prime}_{\epsilon}(R_{\epsilon,1})(R_{\epsilon,1}-\sigma)}{|\Psi^{\prime\prime}_{\epsilon}(S_{\epsilon,1})|}}\,, (52)

so that ξϵ\xi^{\prime}_{\epsilon} is uniformly integrable over [Uϵ,1,Rϵ,1)[U_{\epsilon,1},R_{\epsilon,1}). By uniform boundedness of the instantaneous flux gϵAΛg_{\epsilon}\leq A_{\Lambda}, this implies that the L1L_{1} norm of the second integral terms in (49) vanishes when ϵ0+\epsilon\to 0^{+}:

0|Uϵ,1Rϵ,1κ(Rϵ,1σ,y,Λ)ξϵ(σ)gϵ(ξϵ(σ))dσ|dy,\displaystyle\int_{0}^{\infty}\bigg{|}\int_{U_{\epsilon,1}}^{R_{\epsilon,1}}\kappa(R_{\epsilon,1}-\sigma,y,\Lambda)\xi^{\prime}_{\epsilon}(\sigma)g_{\epsilon}(\xi_{\epsilon}(\sigma))\,\mathrm{d}\sigma\bigg{|}\,\mathrm{d}y\,, (53)
Uϵ,1Rϵ,1(0κ(Rϵ,1σ,y,Λ)dy)|ξϵ(σ)|gϵ(ξϵ(σ))dσ,\displaystyle\hskip 30.0pt\leq\int_{U_{\epsilon,1}}^{R_{\epsilon,1}}\left(\int_{0}^{\infty}\kappa(R_{\epsilon,1}-\sigma,y,\Lambda)\,\mathrm{d}y\right)\big{|}\xi^{\prime}_{\epsilon}(\sigma)\big{|}g_{\epsilon}(\xi_{\epsilon}(\sigma))\,\mathrm{d}\sigma\,, (54)
AΛUϵ,1Rϵ,1|ξϵ(σ)|dσ=O(Rϵ,1Uϵ,1)ϵ0+0,\displaystyle\hskip 30.0pt\leq A_{\Lambda}\int_{U_{\epsilon,1}}^{R_{\epsilon,1}}\big{|}\xi^{\prime}_{\epsilon}(\sigma)\big{|}\,\mathrm{d}\sigma\;=\;O\big{(}\sqrt{R_{\epsilon,1}-U_{\epsilon,1}}\big{)}\xrightarrow{\epsilon\to 0^{+}}0\,, (55)

which follows from the fact that |Rϵ,1Uϵ,1|2νϵ|R_{\epsilon,1}-U_{\epsilon,1}|\leq 2\nu\epsilon.

(ii)(ii) To show the announced result, it is thus enough to restrict the analysis the first integral term in (49) due to surviving processes at blowup exit times. This term results from the same absorption dynamics as during a blowup episode starting at Sϵ,1S_{\epsilon,1}, so that we have

0|qϵ(Uϵ,1,y)0κ(Rϵ,1Uϵ,1,y,x)qϵ(Uϵ,1,x)dx|dy,\displaystyle\int_{0}^{\infty}\bigg{|}q_{\epsilon}(U_{\epsilon,1},y)-\int_{0}^{\infty}\kappa(R_{\epsilon,1}-U_{\epsilon,1},y,x)q_{\epsilon}(U_{\epsilon,1},x)\,\mathrm{d}x\bigg{|}\,\mathrm{d}y\,, (56)
=0|0κ(Uϵ,1Sϵ,1,y,x)κ(Rϵ,1Sϵ,1,y,x)qϵ(Sϵ,1,x)dx|dy,\displaystyle\hskip 30.0pt=\int_{0}^{\infty}\bigg{|}\int_{0}^{\infty}\kappa(U_{\epsilon,1}-S_{\epsilon,1},y,x)-\kappa(R_{\epsilon,1}-S_{\epsilon,1},y,x)q_{\epsilon}(S_{\epsilon,1},x)\,\mathrm{d}x\bigg{|}\,\mathrm{d}y\,, (57)
0(0supUϵ,1σRϵ,1|σκ(σSϵ,1,y,x)|dy)|Uϵ,1Rϵ,1|qϵ(Sϵ,1,x)dx,\displaystyle\hskip 30.0pt\leq\int_{0}^{\infty}\left(\int_{0}^{\infty}\sup_{U_{\epsilon,1}\leq\sigma\leq R_{\epsilon,1}}|\partial_{\sigma}\kappa(\sigma-S_{\epsilon,1},y,x)|\,\mathrm{d}y\right)\,|U_{\epsilon,1}-R_{\epsilon,1}|\,q_{\epsilon}(S_{\epsilon,1},x)\,\mathrm{d}x\,, (58)
MUϵ,1Sϵ,1,Rϵ,1Sϵ,1|Uϵ,1Rϵ,1|,\displaystyle\hskip 30.0pt\leq M_{U_{\epsilon,1}-S_{\epsilon,1},R_{\epsilon,1}-S_{\epsilon,1}}|U_{\epsilon,1}-R_{\epsilon,1}|\,, (59)

where the upper bound from Lemma B.1 satisfies MUϵ,1Sϵ,1,Rϵ,1Sϵ,1Mλ/2,λ+2νϵM_{U_{\epsilon,1}-S_{\epsilon,1},R_{\epsilon,1}-S_{\epsilon,1}}\leq M_{\lambda/2,\lambda+2\nu\epsilon}. The latter bound is a decreasing function of ϵ\epsilon, and thus uniformly bounded with respect to ϵ\epsilon, so that we have

limϵ0+qϵ(Rϵ,1,)qϵ(Uϵ,1,)1=0.\displaystyle\lim_{\epsilon\to 0^{+}}\big{\|}q_{\epsilon}(R_{\epsilon,1}^{-},\cdot)-q_{\epsilon}(U_{\epsilon,1},\cdot)\big{\|}_{1}=0\,. (60)

As we know that qϵ(Rϵ,1,)q_{\epsilon}(R_{\epsilon,1}^{-},\cdot) also converges in L1L_{1} norm toward qϵ(U1,)q_{\epsilon}(U_{1}^{-},\cdot) at blowup exit time, the triangular inequality implies that

limϵ0+qϵ(Rϵ,1,)q(U1,)1=0.\displaystyle\lim_{\epsilon\to 0^{+}}\big{\|}q_{\epsilon}(R_{\epsilon,1}^{-},\cdot)-q(U_{1}^{-},\cdot)\big{\|}_{1}=0\,. (61)

Finally, splitting the delta Dirac contribution from the pre-reset continuous density part

qϵ(Rϵ,1,)q(U1,)1=|πϵ,1π1|+qϵ(Rϵ,1,)q(U1,)1,\displaystyle\big{\|}q_{\epsilon}(R_{\epsilon,1},\cdot)-q(U_{1},\cdot)\big{\|}_{1}=|\pi_{\epsilon,1}-\pi_{1}|+\big{\|}q_{\epsilon}(R_{\epsilon,1}^{-},\cdot)-q(U_{1}^{-},\cdot)\big{\|}_{1}\,, (62)

we conclude by remembering that we have limϵ0+πϵ,1=π1\lim_{\epsilon\to 0^{+}}\pi_{\epsilon,1}=\pi_{1}. ∎

Acknowledgements

The authors would like to thank Phillip Whitman and Luyan Yu for insightful discussions and for simulation results.

The second author was supported by an Alfred P. Sloan Research Fellowship FG-2017-9554 and a CRCNS award DMS-2113213 from the National Science Foundation.

Appendix A

This appendix contains a lemma that is useful to control the L1L_{1} norm of the density of surviving processes in the limit of vanishing refractory periods ϵ0+\epsilon\to 0^{+}. It applies to integral terms bearing on reset processes.

Lemma A.1.

Under Assumption 4, for small enough refractory period ϵ\epsilon, we have ξϵ1\xi^{\prime}_{\epsilon}\leq 1 on [Sϵ,1s1/2,Sϵ,1+λ/2][S_{\epsilon,1}-s_{1}/2,S_{\epsilon,1}+\lambda/2], where s1>0s_{1}>0 only depends on Λ\Lambda and λ\lambda.

Proof.

Under Assumption 4, by (iii)(iii) in the proof of Proposition 3.3, we have

Sϵ,1s1σSϵ,1(σgϵ(σ)bΛ2λΨϵ′′(σ)bΛ2ν),\displaystyle S_{\epsilon,1}-s_{1}\leq\sigma\leq S_{\epsilon,1}\quad\Rightarrow\quad\left(\partial_{\sigma}g_{\epsilon}(\sigma)\geq\frac{b_{\Lambda}}{2\lambda}\quad\Leftrightarrow\quad\Psi_{\epsilon}^{\prime\prime}(\sigma)\leq\frac{b_{\Lambda}}{2\nu}\right)\,, (63)

where the constant s1s_{1} defined by (40) only depends on Λ\Lambda and λ\lambda. Thus, for all Sϵ,1s1σSϵ,1S_{\epsilon,1}-s_{1}\leq\sigma\leq S_{\epsilon,1}, we have

Ψϵ(σ)=Ψϵ(σ)Ψϵ(S1,ϵ)=σS1,ϵΨϵ′′(τ)dτbΛ2ν(S1,ϵσ).\displaystyle\Psi_{\epsilon}^{\prime}(\sigma)=\Psi_{\epsilon}^{\prime}(\sigma)-\Psi_{\epsilon}^{\prime}(S_{1,\epsilon})=-\int_{\sigma}^{S_{1,\epsilon}}\Psi_{\epsilon}^{\prime\prime}(\tau)\,\mathrm{d}\tau\geq\frac{b_{\Lambda}}{2\nu}(S_{1,\epsilon}-\sigma)\,.

In turn, integrating once more for all Sϵ,1s1ξ<σSϵ,1S_{\epsilon,1}-s_{1}\leq\xi<\sigma\leq S_{\epsilon,1} yields

Ψϵ(σ)Ψϵ(ξ)bΛ2ν(Sϵ,1(σξ)+(σ2ξ2)/2).\displaystyle\Psi_{\epsilon}(\sigma)-\Psi_{\epsilon}(\xi)\geq\frac{b_{\Lambda}}{2\nu}\left(S_{\epsilon,1}(\sigma-\xi)+(\sigma^{2}-\xi^{2})/2\right)\,.

The backward time ξ(σ)\xi(\sigma) is such that Ψϵ(σ)Ψϵ(ξ(σ))=ϵ\Psi_{\epsilon}(\sigma)-\Psi_{\epsilon}(\xi(\sigma))=\epsilon, which implies that we necessarily have

ϵbΛ2ν(Sϵ,1(σξ(σ))+(σ2ξ(σ)2)/2).\displaystyle\epsilon\geq\frac{b_{\Lambda}}{2\nu}\left(S_{\epsilon,1}(\sigma-\xi(\sigma))+(\sigma^{2}-\xi(\sigma)^{2})/2\right)\,.

The above inequality provides us with a lower bound on admissible value for ξ(σ)\xi(\sigma). Solving for ξ(σ)\xi(\sigma) reveals that the inequality holds with ξ(σ)σ\xi(\sigma)\leq\sigma if and only if

ξ(σ)S1,ϵ4νϵbΛ+(Sϵ,1σ)2.\displaystyle\xi(\sigma)\leq S_{1,\epsilon}-\sqrt{\frac{4\nu\epsilon}{b_{\Lambda}}+(S_{\epsilon,1}-\sigma)^{2}}\,.

This shows that for small enough ϵ\epsilon

σ>S1,ϵs124νϵbΛS1,ϵξ(σ)<s1.\displaystyle\sigma>S_{1,\epsilon}-\sqrt{s_{1}^{2}-\frac{4\nu\epsilon}{b_{\Lambda}}}\quad\Rightarrow\quad S_{1,\epsilon}-\xi(\sigma)<s_{1}\,.

In particular, if ϵ<3s12bΛ/(16ν)\epsilon<3s_{1}^{2}b_{\Lambda}/(16\nu), we have Sϵ,1s1/2σSϵ,1S_{\epsilon,1}-s_{1}/2\leq\sigma\leq S_{\epsilon,1} implies that Sϵ,1s1<ξ(σ)<σS_{\epsilon,1}-s_{1}<\xi(\sigma)<\sigma. Then the result follows (1)(1) from remembering that by (63), Ψϵ\Psi_{\epsilon}^{\prime} is increasing on [Sϵ,1s1,Sϵ,1][S_{\epsilon,1}-s_{1},S_{\epsilon,1}] so that:

Sϵ,1s12σSϵ,1ξϵ(σ)=Ψϵ(σ)Ψ(ξϵ(σ))1,\displaystyle S_{\epsilon,1}-\frac{s_{1}}{2}\leq\sigma\leq S_{\epsilon,1}\quad\Rightarrow\quad\xi_{\epsilon}^{\prime}(\sigma)=\frac{\Psi_{\epsilon}^{\prime}(\sigma)}{\Psi^{\prime}(\xi_{\epsilon}(\sigma))}\leq 1\,,

and (2)(2) from observing that under Assumption 4, the blowup size is such that π11/2\pi_{1}\geq 1/2, so that Ψϵ(σ)=0\Psi_{\epsilon}^{\prime}(\sigma)=0 and Ψϵ(ξϵ(σ))=Ψϵ(ξϵ(Sϵ,1))>0\Psi_{\epsilon}^{\prime}(\xi_{\epsilon}(\sigma))=\Psi_{\epsilon}^{\prime}(\xi_{\epsilon}(S_{\epsilon,1}))>0 for all Sϵ,1σSϵ,1+λ/2S_{\epsilon,1}\leq\sigma\leq S_{\epsilon,1}+\lambda/2

Appendix B

This appendix contains another lemma that is useful to control the L1L_{1} norm of the density of surviving processes in the limit of vanishing refractory periods ϵ0+\epsilon\to 0^{+}. It applies to integral terms bearing on active processes.

Lemma B.1.

For all 0<σa<σb<0<\sigma_{a}<\sigma_{b}<\infty, there exists a finite upper bound

Mab=supx>0(0supσaσσb|σκ(σ,y,x)|dy)<.\displaystyle M_{ab}=\sup_{x>0}\left(\int_{0}^{\infty}\sup_{\sigma_{a}\leq\sigma\leq\sigma_{b}}|\partial_{\sigma}\kappa(\sigma,y,x)|\,\mathrm{d}y\right)<\infty\,.
Proof.

Differentiating expression (21) for κ(σ,y,x)\kappa(\sigma,y,x) with respect to σ\sigma yields:

|σκ(σ,y,x)|\displaystyle|\partial_{\sigma}\kappa(\sigma,y,x)| =\displaystyle= e(yx+σ)22σ22πσ5|σ(1+σ)(1e2xyσ)+(yx)2(y+x)2e2xyσ|,\displaystyle\frac{e^{-\frac{(y-x+\sigma)^{2}}{2\sigma}}}{2\sqrt{2\pi\sigma^{5}}}\Big{|}\sigma(1+\sigma)\left(1-e^{-\frac{2xy}{\sigma}}\right)+(y-x)^{2}-(y+x)^{2}e^{-\frac{2xy}{\sigma}}\Big{|}\,,
\displaystyle\leq e(yx+σ)22σ22πσ5(σ(1+σ)+|(yx)2(1e2xyσ)4xye2xyσ|),\displaystyle\frac{e^{-\frac{(y-x+\sigma)^{2}}{2\sigma}}}{2\sqrt{2\pi\sigma^{5}}}\left(\sigma(1+\sigma)+\Big{|}(y-x)^{2}\left(1-e^{-\frac{2xy}{\sigma}}\right)-4xye^{-\frac{2xy}{\sigma}}\Big{|}\right)\,,
\displaystyle\leq e(yx+σ)22σ22πσ5(σ(1+σ)+(yx)2+4xye2xyσ),\displaystyle\frac{e^{-\frac{(y-x+\sigma)^{2}}{2\sigma}}}{2\sqrt{2\pi\sigma^{5}}}\left(\sigma(1+\sigma)+(y-x)^{2}+4xye^{-\frac{2xy}{\sigma}}\right)\,,
\displaystyle\leq e(yx)22σ(yx)σ222πσ5(σ(1+σ)+(yx)2+2σ/e),\displaystyle\frac{e^{-\frac{(y-x)^{2}}{2\sigma}-(y-x)-\frac{\sigma}{2}}}{2\sqrt{2\pi\sigma^{5}}}\left(\sigma(1+\sigma)+(y-x)^{2}+2\sigma/e\right)\,,

where the last inequality follows from the fact that ueu0,1/e\|ue^{-u}\|_{0,\infty}\leq 1/e. From there, we have

supπapπb|σκ(λp,y,x)|\displaystyle\sup_{\pi_{a}\leq p\leq\pi_{b}}\big{|}\partial_{\sigma}\kappa(\lambda p,y,x)| \displaystyle\leq e(yx)22σb(yx)σa222πσa5(σb(1+2/e+σb)+(yx)2),\displaystyle\frac{e^{-\frac{(y-x)^{2}}{2\sigma_{b}}-(y-x)-\frac{\sigma_{a}}{2}}}{2\sqrt{2\pi\sigma_{a}^{5}}}\left(\sigma_{b}(1+2/e+\sigma_{b})+(y-x)^{2}\right)\,,
\displaystyle\leq e(yx+σb)22σb(Aab+Bab(yx)2),\displaystyle e^{-\frac{(y-x+\sigma_{b})^{2}}{2\sigma_{b}}}\big{(}A_{ab}+B_{ab}(y-x)^{2}\big{)}\,,

where we define the constants:

Aab=eσbσa222πσa5σb(1+2/e+σb)andBab=eσbσa222πσa5.\displaystyle A_{ab}=\frac{e^{\frac{\sigma_{b}-\sigma_{a}}{2}}}{2\sqrt{2\pi\sigma_{a}^{5}}}\sigma_{b}(1+2/e+\sigma_{b})\quad\mathrm{and}\quad B_{ab}=\frac{e^{\frac{\sigma_{b}-\sigma_{a}}{2}}}{2\sqrt{2\pi\sigma_{a}^{5}}}\,.

Integration with respect to yy then yields the finite upper bound:

0supσaσσb\displaystyle\int_{0}^{\infty}\sup_{\sigma_{a}\leq\sigma\leq\sigma_{b}} |σκ(σ,y,x)|dy\displaystyle\!\!\!|\partial_{\sigma}\kappa(\sigma,y,x)|\,\mathrm{d}y
\displaystyle\leq Ma,b=Aabπ2σb+Bab((1+σb)π2σb3+e(xσb)22σbσb(σb+x)).\displaystyle M_{a,b}=A_{ab}\sqrt{\frac{\pi}{2}}\sigma_{b}+B_{ab}\left((1+\sigma_{b})\sqrt{\frac{\pi}{2}\sigma_{b}^{3}}+e^{\frac{-(x-\sigma_{b})^{2}}{2\sigma_{b}}}\sigma_{b}(\sigma_{b}+x)\right)\,.

References

  • [1] {bbook}[author] \bauthor\bsnmAsmussen, \bfnmSøren\binitsS. (\byear2003). \btitleApplied probability and queues, \beditionsecond ed. \bseriesApplications of Mathematics (New York) \bvolume51. \bpublisherSpringer-Verlag, New York \bnoteStochastic Modelling and Applied Probability. \bmrnumber1978607 \endbibitem
  • [2] {bbook}[author] \bauthor\bsnmBillingsley, \bfnmPatrick\binitsP. (\byear2013). \btitleConvergence of probability measures. \bpublisherJohn Wiley & Sons. \endbibitem
  • [3] {barticle}[author] \bauthor\bsnmBrette, \bfnmRomain\binitsR. (\byear2015). \btitlePhilosophy of the spike: rate-based vs. spike-based theories of the brain. \bjournalFrontiers in systems neuroscience \bvolume9 \bpages151. \endbibitem
  • [4] {barticle}[author] \bauthor\bsnmBrunel, \bfnmNicolas\binitsN. (\byear2000). \btitleDynamics of Sparsely Connected Networks of Excitatory and Inhibitory Spiking Neurons. \bjournalJournal of Computational Neuroscience \bvolume8 \bpages183–208. \bdoi10.1023/A:1008925309027 \endbibitem
  • [5] {barticle}[author] \bauthor\bsnmBrunel, \bfnmNicolas\binitsN. and \bauthor\bsnmHakim, \bfnmVincent\binitsV. (\byear1999). \btitleFast Global Oscillations in Networks of Integrate-and-Fire Neurons with Low Firing Rates. \bjournalNeural Computation \bvolume11 \bpages1621–1671. \bdoi10.1162/089976699300016179 \endbibitem
  • [6] {barticle}[author] \bauthor\bsnmBurkitt, \bfnmAnthony N\binitsA. N. (\byear2006). \btitleA review of the integrate-and-fire neuron model: I. Homogeneous synaptic input. \bjournalBiological cybernetics \bvolume95 \bpages1–19. \endbibitem
  • [7] {barticle}[author] \bauthor\bsnmCáceres, \bfnmMaría J\binitsM. J., \bauthor\bsnmCarrillo, \bfnmJosé A\binitsJ. A. and \bauthor\bsnmPerthame, \bfnmBenoît\binitsB. (\byear2011). \btitleAnalysis of nonlinear noisy integrate & fire neuron models: blow-up and steady states. \bjournalThe Journal of Mathematical Neuroscience \bvolume1 \bpages7. \endbibitem
  • [8] {barticle}[author] \bauthor\bsnmCarrillo, \bfnmJosé A\binitsJ. A., \bauthor\bsnmGonzález, \bfnmMaría d M\binitsM. d. M., \bauthor\bsnmGualdani, \bfnmMaria P\binitsM. P. and \bauthor\bsnmSchonbek, \bfnmMaria E\binitsM. E. (\byear2013). \btitleClassical solutions for a nonlinear Fokker-Planck equation arising in computational neuroscience. \bjournalCommunications in Partial Differential Equations \bvolume38 \bpages385–409. \endbibitem
  • [9] {barticle}[author] \bauthor\bsnmDelarue, \bfnmF.\binitsF., \bauthor\bsnmInglis, \bfnmJ.\binitsJ., \bauthor\bsnmRubenthaler, \bfnmS.\binitsS. and \bauthor\bsnmTanré, \bfnmE.\binitsE. (\byear2015). \btitleParticle systems with a singular mean-field self-excitation. Application to neuronal networks. \bjournalStochastic Processes and their Applications \bvolume125 \bpages2451 - 2492. \bdoihttps://doi.org/10.1016/j.spa.2015.01.007 \endbibitem
  • [10] {barticle}[author] \bauthor\bsnmDelarue, \bfnmFrançois\binitsF., \bauthor\bsnmInglis, \bfnmJames\binitsJ., \bauthor\bsnmRubenthaler, \bfnmSylvain\binitsS., \bauthor\bsnmTanré, \bfnmEtienne\binitsE. \betalet al. (\byear2015). \btitleGlobal solvability of a networked integrate-and-fire model of McKean–Vlasov type. \bjournalThe Annals of Applied Probability \bvolume25 \bpages2096–2133. \endbibitem
  • [11] {barticle}[author] \bauthor\bsnmFeller, \bfnmWilly\binitsW. (\byear1941). \btitleOn the Integral Equation of Renewal Theory. \bjournalThe Annals of Mathematical Statistics \bvolume12 \bpages243–267. \endbibitem
  • [12] {bbook}[author] \bauthor\bsnmGranas, \bfnmAndrzej\binitsA. and \bauthor\bsnmDugundji, \bfnmJames\binitsJ. (\byear2003). \btitleFixed point theory \bvolume14. \bpublisherSpringer. \endbibitem
  • [13] {barticle}[author] \bauthor\bsnmHambly, \bfnmBen\binitsB., \bauthor\bsnmLedger, \bfnmSean\binitsS., \bauthor\bsnmSøjmark, \bfnmAndreas\binitsA. \betalet al. (\byear2019). \btitleA McKean–Vlasov equation with positive feedback and blow-ups. \bjournalThe Annals of Applied Probability \bvolume29 \bpages2338–2373. \endbibitem
  • [14] {bbook}[author] \bauthor\bsnmKaratzas, \bfnmIoannis\binitsI. and \bauthor\bsnmShreve, \bfnmSteven E.\binitsS. E. (\byear1991). \btitleBrownian motion and stochastic calculus, \beditionsecond ed. \bseriesGraduate Texts in Mathematics \bvolume113. \bpublisherSpringer-Verlag, \baddressNew York. \bmrnumberMR1121940 (92h:60127) \endbibitem
  • [15] {barticle}[author] \bauthor\bsnmKasabov, \bfnmNikola\binitsN. (\byear2010). \btitleTo spike or not to spike: A probabilistic spiking neuron model. \bjournalNeural Networks \bvolume23 \bpages16–19. \endbibitem
  • [16] {barticle}[author] \bauthor\bsnmLapicque, \bfnmLouis\binitsL. (\byear1907). \btitleRecherches quantitatives sur l’excitation electrique des nerfs traitee comme une polarization. \bjournalJournal de Physiologie et de Pathologie Generalej \bvolume9 \bpages620–635. \endbibitem
  • [17] {barticle}[author] \bauthor\bsnmNadtochiy, \bfnmSergey\binitsS. and \bauthor\bsnmShkolnikov, \bfnmMykhaylo\binitsM. (\byear2020). \btitleMean field systems on networks, with singular interaction through hitting times. \bjournalThe Annals of Probability \bvolume48 \bpages1520–1556. \endbibitem
  • [18] {barticle}[author] \bauthor\bsnmNadtochiy, \bfnmSergey\binitsS., \bauthor\bsnmShkolnikov, \bfnmMykhaylo\binitsM. \betalet al. (\byear2019). \btitleParticle systems with singular interaction through hitting times: application in systemic risk modeling. \bjournalThe Annals of Applied Probability \bvolume29 \bpages89–129. \endbibitem
  • [19] {barticle}[author] \bauthor\bsnmPanzeri, \bfnmStefano\binitsS., \bauthor\bsnmBrunel, \bfnmNicolas\binitsN., \bauthor\bsnmLogothetis, \bfnmNikos K\binitsN. K. and \bauthor\bsnmKayser, \bfnmChristoph\binitsC. (\byear2010). \btitleSensory neural codes using multiplexed temporal scales. \bjournalTrends in neurosciences \bvolume33 \bpages111–120. \endbibitem
  • [20] {bbook}[author] \bauthor\bsnmRobert, \bfnmChristian P.\binitsC. P. and \bauthor\bsnmCasella, \bfnmGeorge\binitsG. (\byear2004). \btitleMonte Carlo statistical methods, \beditionsecond ed. \bseriesSpringer Texts in Statistics. \bpublisherSpringer-Verlag, \baddressNew York. \bmrnumber2080278 (2005d:62006) \endbibitem
  • [21] {bincollection}[author] \bauthor\bsnmSznitman, \bfnmAlain-Sol\binitsA.-S. (\byear1991). \btitleTopics in propagation of chaos. In \bbooktitleÉcole d’Été de Probabilités de Saint-Flour XIX—1989. \bseriesLecture Notes in Math. \bvolume1464 \bpages165–251. \bpublisherSpringer, Berlin. \bdoi10.1007/BFb0085169 \bmrnumber1108185 \endbibitem
  • [22] {barticle}[author] \bauthor\bsnmTaillefumier, \bfnmThibaud\binitsT. and \bauthor\bsnmWhitman, \bfnmPhillip\binitsP. (\byear2022). \btitleCharacterization of blowups via time change in a mean-field neural network. \endbibitem