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

Exact description of limiting SIR and SEIR dynamics on locally tree-like graphs

Juniper Cocomello  and  Kavita Ramanan Division of Applied Mathematics, Brown University, 182 George Street, Providence, RI 02912 juniper_cocomello@brown.edu, kavita_ramanan@brown.edu
Abstract.

We study the Susceptible-Infected-Recovered (SIR) and the Susceptible-Exposed-Infected-Recovered (SEIR) models of epidemics, with possibly time-varying rates, on a class of networks that are locally tree-like, which includes sparse Erdős-Rényi random graphs, random regular graphs, and other configuration models. We identify tractable systems of ODEs that exactly describe the dynamics of the SIR and SEIR processes in a suitable asymptotic regime in which the population size goes to infinity. Moreover, in the case of constant recovery and infection rates, we characterize the outbreak size as the unique zero of an explicit functional. We use this to show that a (suitably defined) mean-field prediction always overestimates the outbreak size, and that the outbreak sizes for SIR and SEIR processes with the same initial condition and constant infection and recovery rates coincide. In contrast, we show that the outbreak sizes for SIR and SEIR processes with the same time-varying infection and recovery rates can in general be quite different. We also demonstrate via simulations the efficacy of our approximations for populations of moderate size.

Key words and phrases:
interacting particle systems; continuous time Markov chains; SIR model; SEIR model, epidemics; outbreak size; sparse graphs; random graphs; local limits; mean-field limits; Erdős-Rényi random graphs; configuration model; random regular graph; Galton-Watson trees.
1991 Mathematics Subject Classification:
Primary: 60K35; 60F17; Secondary: 60G60; 60J2.
K. Ramanan was supported in part by ARO Grant W911NF2010133 and both authors were supported by the Office of Naval Research under the Vannevar Bush Faculty Fellowship N0014-21-1-2887.
Acknowledgements: The second author would like to thank the Simons Institute and the organizers of the workshop on “Graph Limits, Nonparametric Models, and Estimation” for inviting her to the workshop to present a preliminary version of these results.

1. Introduction

Models, Results and Proof Techniques. The Susceptible-Infected-Recovered (SIR) model has been extensively used to study the spread of infectious diseases, computer viruses, information, and rumors. In this model, each individual in a population is represented as being in one of three states: susceptible (does not have the disease, but could catch it), infected (has the disease and can spread it to susceptible individuals), or recovered (no longer has the disease and cannot be reinfected). The dynamics are governed by a graph, which describes the contact network of individuals in a population, and two strictly positive parameters: ρ\rho, the rate at which an infected individual recovers, and β\beta, the rate at which an infected individual transmits the disease to a susceptible individual with whom it is in immediate contact. Random graphs are used to model variation and uncertainly in real-world contact networks. We are interested in the global dynamical behavior that arises from the local interactions of disease spreading. In particular, we are interested in the following questions: how many individuals are at each of the states at any time t[0,)t\in[0,\infty)? What is the total size of the outbreak, that is, how many individuals were infected during the course of the epidemic? Answering these questions for large populations can be analytically challenging, and simulations are computationally expensive and do not easily allow for rigorous characterization of qualitative behavior.

In the present work, we study a continuous-time stochastic SIR process, and a related epidemic model, the Susceptible-Exposed-Infected-Recovered (SEIR) process, where an additional state is considered for individuals that have been exposed to the pathogen but have not yet become infectious. The SEIR process has been widely used to model the spreading of diseases including the recent SARS-CoV-2 pandemic, for instance, see [Suwardi2020stability, girardi2023anseir, mwalili2020SEIR]. In both cases, we allow for the (infection and recovery) transition rates to be time-dependent, so as to model effects due to seasonal variations, changes in the virulence of a disease, developments in treatment options, and changes in public health policies, which are of significant interest in practice [morris2021optimal, lopez2021modified, fisman2007seasonality]. While the majority of works have considered SIR processes on dense networks, we consider these processes on sparse networks (i.e., where each individual is connected to a bounded number of individuals), which more faithfully describe real-world networks. We provide tractable approximations for the evolution of fractions of individuals in each of the states of the epidemic in terms of a coupled system of ordinary differential equations (ODEs), see (2.4) and (2.11), and for the outbreak size, which is the final fraction of individuals ever infected. Moreover, we show that these approximations are asymptotically exact, as the size of the population increases to infinity, when the graph governing the dynamics is locally-tree like. More precisely, we consider a broad class of sparse (random) graph sequences, including sparse Erdős-Rényi random graphs, random regular graphs, and certain configuration models, which are known to converge in a certain local (weak) sense to a random limit that belongs to the class of unimodular Galton-Watson (UGW) trees; see Theorem 2.5 and Theorem 2.10. We refer the reader to Definition 2.3 for the definition of a UGW tree, and to [vanderHofstad2023vol2, Aldous2004objective] for an extensive account of local convergence of graphs.

Our proof technique starts by appealing to a general result in [Ganguly2022hydrodynamic] that shows that for a general class of interacting particle systems that includes the SIR and SEIR processes, the sequence of empirical measures (equivalently, fractions of individuals in different states) on any locally converging sequence of graphs converges to the law of the marginal evolution of the root node in the limit UGW tree (see also [Lacker2023local, RamICM22] for related results). The key step is then to provide a tractable characterization of the root marginal dynamics of this infinite-dimensional process. While for general particle systems the marginal dynamics of the root, or even of the root and its neighborhood, could be non-Markovian, a key step in our proof is to show that for the SIR and SEIR models, the dynamics simplifies. In fact, we can deduce from our proof that the evolution of the pair of vertices consisting of the root and an offspring is in fact Markovian (see Remark 4.11). The proof of the latter property relies crucially on certain conditional independence relations that we identify (see Proposition 4.7) and a certain projection lemma (Proposition 4.6) in the spirit of [Ganguly2022nonmarkovian, Lemma 9.1]. These properties are combined with symmetry properties of the dynamics (identified in Proposition 4.8) to obtain a very tractable description of the evolution of the marginal law of the root in terms of the abovementioned systems of ODEs.

For both the SIR and SEIR models, the associated system of ODEs is then analyzed to characterize the outbreak size in terms of the moment generating function of the offspring distribution of the limiting UGW tree, evaluated at the unique zero of an explicitly given functional; see Theorem 3.1. In the case of constant recovery and infection rates, we obtain a simpler characterization of the outbreak size and use it to show that the (suitably defined) mean-field prediction always overestimates the outbreak size. In this setting, we also show that although the transient dynamics can be different, the outbreak sizes for the SIR and SEIR models coincide when they have the same rates and initial conditions. In particular, this shows that in this case the outbreak size for the SEIR model does not depend on the rate at which an exposed individual becomes infectious. In contrast, we show that when the rates are time-varying, the outbreak sizes of the corresponding SIR and SEIR processes no longer coincide and can be vastly different even when the (time-varying) ratios of the infection rate to the recovery rate coincide. For both transient dynamics and the outbreak size, we compare our results with numerical simulations to demonstrate the efficacy of these approximations for populations of even moderate size. We also show how the ODEs can be used to study the impact of the amplitude and phase of periodically varying rates on the outbreak size.

When the infection and recovery rates are constant in time, traditional techniques to analyze the outbreak size of the SIR process exploit a reformulation of the final outbreak size in terms of a bond percolation problem. However, it is not apparent if such a simple correspondence exists when the infection and recovery rates are time-varying, and unlike our approach, percolation-based arguments provide limited insight into the dynamics of the epidemic process. Furthermore, our general approach can also be applied to obtain analogous results for other more general epidemic processes including a class of compartmental models and processes with general recovery distributions; see Remark 2.8 and Remark 2.11. It would also be of interest to investigate to what extent an analogous approach can be used to provide alternatives to mean-field approximations for other classes of models, for instance, such as those described in [RamQuesta22]. We defer a complete investigation to future work.

Discussion of Prior Work. Understanding epidemic dynamics on networks is an active area of contemporary research. The deterministic SIR model, introduced in [kermack1927contribution], is a system of coupled ODEs that describes the evolution over time of the fraction of individuals in each of the states of the epidemic, in a population where everyone can come into contact with everyone else. This is known as the mean-field approximation. The mean-field dynamics are known to emerge as the large nn limit of the SIR process defined on the complete graph on nn vertices, when the infection rate scales like 𝒪(1/n){\mathcal{O}}(1/n). The mean-field approximation provides a dramatic reduction of dimensionality, as it captures the global behavior of a size nn population by a coupled system of two ODEs. However, most real-world contact networks are sparse, in the sense that the average number of neighbors for an individual in the network remands bounded even when the population size grows.

Because of this, and the application-driven need to understand epidemic dynamics on more realistic networks, the study of SIR dynamics on a range of more realistic sparse network structures is an active area of research. The work of [schutz2008exact] derives equations for the expected number of individuals in each SIR state on a cycle graph, and compares these results with the corresponding quantities associated with the SIR model on the complete graph, as well as the scaled dynamics that result in the mean-field approximation. An SIR model on the κ\kappa-regular tree (the infinite tree where every vertex has κ\kappa neighbors) with general recovery times and time-dependent rates was studied in [gairat2022discrete]. The latter work derives the asymptotic limit, as κ\kappa goes to infinity, of the evolution of the fraction of susceptible individuals over time, which recovers the mean-field approximation. Differential equations to approximate the fraction of susceptible, infected and recovered individuals for the continuous-time SIR model on configuration model graphs were derived heuristically in [Volz2008SIR, Volz2009epidemic] and shown to be asymptotically exact, as the population size goes to infinity, in [Decreusefond2012large, janson2015law]. In very recent work [hall2023exact], the authors obtain an explicit representation of the marginal distribution of each node on a finite tree by solving a coupled system of ODEs. This representation is shown to provide an upper bound on the probability that a node is susceptible on general graphs or with more than one initial infection. However, they show, via simulations, that this upper bound is generally not very tight.

Existing mathematically rigorous work on the SEIR process focuses on studying the deterministic dynamics that arise in the mean-field regime, see [li1999global]. To the best of our knowledge, not much is known rigorously about SEIR processes on sparse graphs or corresponding limits. In [Zhao2013SEIR], the authors present an ODE system that they heuristically argue should approximate the fraction for large populations size. However, their approximation is not compared with simulations, and it differs from our ODE system, which is asymptotically exact as the population size approaches infinity.

Organization of the Paper. The rest of the paper is structured as follows. In Section 1.1 we introduce some common notation used throughout the paper. In Section 2 we define the SIR and SEIR processes and state our characterization of the large-population limit of epidemic dynamics (see Theorem 2.5 and Theorem 2.10). In Section 3 we provide a characterization of the outbreak size in the large-population limit (see Theorem 3.1 and Theorem 3.5). The proofs of our results are provided in Section 4. They rely on a conditional independence property that is proven in Section 5 and some auxiliary results on the SEIR process that are relegated to Appendix A. Additionally, the proof of the well-posedness of the limit ODE systems are given in Appendix B.

1.1. Notation

We briefly overview common notation used throughout the paper. We use G=(V,E)G=(V,E) to denote a graph with vertex set VV and edge set EE. When clear from context, we identify a graph with its vertex set, and so for a vertex vv we might write vGv\in G instead of the more accurate vVv\in V. We let |G|:=|V||G|:=|V| denote the number of vertices of GG. Given AVA\subset V, we let GA:={wV:{w,v}E,vA,wVA}\partial^{G}A:=\{w\in V\ :\ \{w,v\}\in E,\ v\in A,w\in V\setminus A\} be the boundary of AA. In the case where A={v}A=\{v\} is a singleton, we write vG:=G{v}\partial^{G}_{v}:=\partial^{G}\{v\}, and refer to it as the set of neighbors of vv. The degree of a vertex is defined as dvG:=|vG|d_{v}^{G}:=|\partial_{v}^{G}|. When unambiguous, we omit the dependence on GG from our notation, and write dvd_{v} and v\partial_{v}. For v,wGv,w\in G, we write vwv\sim w to mean vwv\in\partial_{w}. Given a set 𝒴{\mathcal{Y}}, a configuration y𝒴Vy\in{\mathcal{Y}}^{V} and AVA\subset V, we write yA:={yv:vA}y_{A}:=\{y_{v}\ :\ v\in A\}, and in the special case when |A|=2|A|=2, yv,w:=y{v,w}y_{v,w}:=y_{\{v,w\}}.

We let 0={0,1,2,}{\mathbb{N}}_{0}=\{0,1,2,...\}, and let 𝒫(0){\mathcal{P}}({\mathbb{N}}_{0}) be the set of probability measures on 0{\mathbb{N}}_{0}. We identify probability measures on 0{\mathbb{N}}_{0} with their probability mass functions. In particular, for ζ(0)\zeta\in{\mathbb{P}}({\mathbb{N}}_{0}) and k0k\in{\mathbb{N}}_{0}, we write ζ(k)=ζ({k})\zeta(k)=\zeta(\{k\}). For kk\in{\mathbb{R}}, we let δk\delta_{k} be the Dirac measure at kk. Given a probability space (Ω,,)(\Omega,{\mathcal{F}},{\mathbb{P}}), we denote by (Y){\mathcal{L}}(Y) the law of a Ω\Omega-valued random variable YY.

2. Results on Transient Dynamics

In Section 2.1 we precisely define the SIR process and in Section 2.2 state the main result that describes the limiting dynamics on converging sequences of locally tree-like graphs in terms of solutions to systems of ODEs. In Section 2.3 we define the SEIR process and state the corresponding convergence result.

2.1. SIR Model

Fix a graph G=(V,E)G=(V,E), the (time-varying) infection rate β:[0,)(0,){\beta}:[0,\infty)\rightarrow(0,\infty), and the (time-varying) recovery rate ρ:[0,)(0,){\rho}:[0,\infty)\rightarrow(0,\infty). We write βt\beta_{t} (resp. ρt\rho_{t}) for the value of β\beta (resp. ρ\rho) at time t[0,)t\in[0,\infty). The SIR process on GG, denoted by XGX^{G}, is a continuous-time locally interacting Markov chain with the following dynamics. At any time tt, each individual vv has a state XvG(t)X^{G}_{v}(t) in the space 𝒳:={S,I,R}\mathcal{X}:=\left\{S,I,R\right\}. The initial states XvG(0)X^{G}_{v}(0) are i.i.d. with (XvG(0)=S)=s0{\mathbb{P}}(X^{G}_{v}(0)=S)=s_{0} and (XvG(0)=I)=i0:=1s0{\mathbb{P}}(X^{G}_{v}(0)=I)=i_{0}:=1-s_{0} for some s0(0,1)s_{0}\in(0,1). Given y(k=1𝒳k)y\in\emptyset\cup(\cup_{k=1}^{\infty}\mathcal{X}^{k}), representing the configuration of the neighbors of a vertex, we denote by (y){\mathcal{I}}(y) the number of elements of yy that are equal to II. At time tt, each individual vGv\in G jumps from SS to II (i.e., becomes infected) at rate βtI(XvG(t)){\beta_{t}}I(X^{G}_{\partial_{v}}(t-)), and from II to RR (i.e., recovers) at rate ρt{\rho_{t}}. We impose the following mild assumptions on the recovery and infection rate functions.

Assumption A.

The functions β\beta and ρ\rho are continuous and there exist c1,c2(0,)c_{1},c_{2}\in(0,\infty) such that

c1<lim inftmin(ρt,βt)<lim suptmax(ρt,βt)<c2.\displaystyle\begin{split}c_{1}<\liminf_{t\rightarrow\infty}\min(\rho_{t},\beta_{t})<\limsup_{t\rightarrow\infty}\max(\rho_{t},\beta_{t})<c_{2}.\end{split} (2.1)

Throughout the paper, we assume that Assumption A holds.

Remark 2.1.

If we equip 𝒳\mathcal{X} with the total ordering given by S<I<RS<I<R, then the SIR process is monotonic in the sense that for every vGv\in G and s,t[0,)s,t\in[0,\infty), if sts\leq t then XvG(s)XvG(t)X_{v}^{G}(s)\leq X_{v}^{G}(t).

Next, we describe the class of graph sequences that we consider, as well as an associated probability measure on 0{\mathbb{N}}_{0} that characterizes the corresponding local limit.

Assumption B.

Suppose the sequence of graphs {Gn}n\{G_{n}\}_{n\in{\mathbb{N}}} and θ𝒫(0)\theta\in{\mathcal{P}}({\mathbb{N}}_{0}) satisfy one of the following:

  1. (1)

    (Erdős-Rényi). There exists c>0c>0 such that, for every n0n\in{\mathbb{N}}_{0}, GnG_{n} is a Erdős-Rényi random graph ER(n,c/n)\text{ER}(n,c/n), and θ\theta is the Poisson distribution with mean cc.

  2. (2)

    (Configuration Model). For each nn, let {di,n}i=1n\{d_{i,n}\}_{i=1}^{n} be a graphical sequence, such that i=1nδdi,n\sum_{i=1}^{n}\delta_{d_{i,n}} converges weakly to θ\theta as nn\rightarrow\infty, and θ\theta has finite third moment. Let GnG_{n} be a graph uniformly chosen among graphs on nn vertices with degree sequence {di,n}i=1n\{d_{i,n}\}_{i=1}^{n}. We write Gn=CMn(θ)G_{n}=\text{CM}_{n}(\theta).

Remark 2.2.

The only place where we use the assumption that θ\theta has a finite third moment is in Proposition 2.4 below (and the corresponding result for the SEIR process, Proposition 2.9). Every result in this paper holds by replacing the assumption that θ\theta has finite third moment in Assumption B(2) with the assumption that θ\theta has finite second moment and that the system (2.4)-(2.5) (and the corresponding system for the SEIR process (2.11)-(2.12)) has a unique solution on [0,)[0,\infty).

We refer the reader to [hofstad2016vol1, Chapter 5 and Chapter 7] for an extensive account of random graphs, including precise definitions and well-known properties of the graphs in Assumption B. The class of graphs we consider is locally tree-like, in a sense that we now make precise. Given θ𝒫(0)\theta\in{\mathcal{P}}({\mathbb{N}}_{0}) with finite first moment, we define its size-biased distribution θ^𝒫(0){\hat{\theta}}\in{\mathcal{P}}({\mathbb{N}}_{0}) by

θ^(k)=(k+1)θ(k+1)j=0jθ(j),for k0.{\hat{\theta}}(k)=\frac{(k+1)\theta(k+1)}{\sum_{j=0}^{\infty}j\theta(j)},\qquad\text{for }k\in{\mathbb{N}}_{0}. (2.2)
Definition 2.3.

The unimodular Galton-Watson tree with offspring distribution θ\theta, denoted by UGW(θ(\theta), is a rooted random tree where the root has a number of children distributed like θ\theta and every vertex in subsequent generations has a number of children distributed like θ^{\hat{\theta}}, independently of the degree of vertices in the same or previous generations.

It is well known that if {Gn}n\{G_{n}\}_{n\in{\mathbb{N}}} and θ\theta satisfy Assumption B, then GnG_{n} converges in a local sense (local weak convergence in probability, as defined in [vanderHofstad2023vol2, Definition 2.11]; see also [Lacker2023local, Definition 2.2] and [Ganguly2022hydrodynamic, Section 2.4]) to a GWT(θ)(\theta) tree. This is established, for instance, in [van2009randomII, Theorem 2.18 and Theorem 4.1].

2.2. Asymptotic Characterization of SIR dynamics

Our first result is the limit characterization (as the graph size goes to infinity) of the evolution of the fractions of individuals that, at each time, are in each of the states {S,I,R}\{S,I,R\}. Given a finite graph GG, for t[0,)t\in[0,\infty) we define

sG(t):=1|G|vG𝟏{XvG(t)=S},iG(t):=1|G|vG𝟏{XvG(t)=I}.\displaystyle\begin{split}&s^{G}(t):=\frac{1}{|G|}\sum_{v\in G}{\bm{1}}_{\left\{X^{G}_{v}(t)=S\right\}},\\ &i^{G}(t):=\frac{1}{|G|}\sum_{v\in G}{\bm{1}}_{\left\{X^{G}_{v}(t)=I\right\}}.\end{split} (2.3)

We start by establishing the existence and uniqueness of the solution to a certain system of ODEs that will be used to describe the limit. As is standard practice, we use the dot notation for derivatives with respect to time, and prime notation for derivatives in space.

Proposition 2.4.

Suppose that θ𝒫(0)\theta\in{\mathcal{P}}({\mathbb{N}}_{0}) has finite third moment and let s0(0,1)s_{0}\in(0,1). Then there exists a unique solution (fS,fI,FI)(f_{S},\ f_{I},\ F_{I}) to the following system of ODEs:

{f˙S=fSfIβ(1k=0kθ^(k)ekFIj=0θ^(j)ejFI),f˙I=fSfIβk=0kθ^(k)ekFIj=0θ^(j)ejFIfI(ρ+ββfI),F˙I=βfI,\begin{cases}\dot{f}_{S}=f_{S}f_{I}{\beta}\left(1-\frac{\sum_{k=0}^{\infty}k{\hat{\theta}}(k)e^{-kF_{I}}}{\sum_{j=0}^{\infty}{\hat{\theta}}(j)e^{-jF_{I}}}\right),\\ \dot{f}_{I}=f_{S}f_{I}{\beta}\frac{\sum_{k=0}^{\infty}k{\hat{\theta}}(k)e^{-kF_{I}}}{\sum_{j=0}^{\infty}{\hat{\theta}}(j)e^{-jF_{I}}}-f_{I}({\rho}+{\beta}-{\beta}f_{I}),\\ \dot{F}_{I}=\beta f_{I},\end{cases} (2.4)

with initial conditions

{fS(0)=s0,fI(0)=1s0,FI(0)=0.\begin{cases}f_{S}(0)=s_{0},\\ f_{I}(0)=1-s_{0},\\ F_{I}(0)=0.\end{cases} (2.5)

The proof of Proposition 2.4 uses standard arguments and is thus relegated to Appendix B. Given ζ𝒫(0)\zeta\in{\mathcal{P}}({\mathbb{N}}_{0}) and x(,0]x\in(-\infty,0], we define its Laplace transform as follows:

Mζ(x):=k0ζ(k)ekx.M_{\zeta}(x):=\sum_{k\in{\mathbb{N}}_{0}}\zeta(k)e^{kx}. (2.6)

Given fS,fI,FIf_{S},\ f_{I},\ F_{I} as in Proposition 2.4, for t[0,)t\in[0,\infty) we define

s()(t):=s0Mθ(FI(t))i()(t):=e0tρu𝑑u(i0+s00tMθ(FI(t))e0uρs𝑑sβufI(u)𝑑u).\displaystyle\begin{split}&s^{(\infty)}(t):=s_{0}M_{\theta}(-F_{I}(t))\\ &i^{(\infty)}(t):=e^{-\int_{0}^{t}\rho_{u}du}\left(i_{0}+s_{0}\int_{0}^{t}M^{\prime}_{\theta}(-F_{I}(t))e^{\int_{0}^{u}\rho_{s}ds}\beta_{u}f_{I}(u)du\right).\end{split} (2.7)

We now state our main result for the SIR model.

Theorem 2.5.

Suppose that a sequence of random graphs {Gn}n\{G_{n}\}_{n\in{\mathbb{N}}} and θ𝒫(0)\theta\in{\mathcal{P}}({\mathbb{N}}_{0}) satisfy Assumption B. Let θ^{\hat{\theta}} be the size-biased version of θ\theta, as defined in (2.2). Suppose that sGn(0)s0(0,1)s^{G_{n}}(0)\rightarrow s_{0}\in(0,1) and let s()s^{(\infty)} and i()i^{(\infty)} be as defined in (2.7). Then, as nn\rightarrow\infty we have

sGn(t)𝑝s()(t),iGn(t)𝑝i()(t),\displaystyle\begin{split}&s^{G_{n}}(t)\xrightarrow{p}s^{(\infty)}(t),\\ &i^{G_{n}}(t)\xrightarrow{p}i^{(\infty)}(t),\end{split} (2.8)

uniformly for t[0,)t\in[0,\infty).

The proof of Theorem 2.5 is given in Section 4.2. It relies on a hydrodynamic limit result established in [Ganguly2022nonmarkovian, Corollary 4.7], which shows that the fraction of individuals in any state a𝒳a\in\mathcal{X} in the SIR process on GnG_{n} converges to (X𝒯(t)=a){\mathbb{P}}(X_{\varnothing}^{\mathcal{T}}(t)=a), where X𝒯X^{\mathcal{T}} is the SIR process on 𝒯=UGW(θ){\mathcal{T}}=\text{UGW}(\theta), and \varnothing is the root vertex. We then show that the trajectories of X𝒯X^{{\mathcal{T}}} satisfy a certain conditional independence property (Proposition 4.7). We combine this property with symmetry properties of the dynamics (see Proposition 4.8) to characterize (X𝒯){\mathcal{L}}(X_{\varnothing}^{\mathcal{T}}) in terms of a system of ODEs. In particular, for a=Sa=S or a=Ia=I, the probability (X𝒯(t)=a){\mathbb{P}}(X^{\mathcal{T}}_{\varnothing}(t)=a) is equal to s()(t)s^{(\infty)}(t) or i()(t)i^{(\infty)}(t), respectively, as defined in (2.7). As mentioned in the Introduction, Proposition 4.7 can be seen as a substantial refinement in the case of the SIR process X𝒯X^{{\mathcal{T}}} of a certain general Markov random field property that holds for more general interacting particle systems; see [Ganguly2022interacting, Theorem 3.7].

In Figure 1, we compare simulations of the evolution of the SIR process on certain Erdős-Rényi random graphs and random 33-regular graphs of size n=250n=250 with the theoretical prediction from Theorem 2.5. The plots illustrate that even in systems of moderate size, the theoretical prediction closely tracks the simulations.

Refer to caption
(a) β1{\beta}\equiv 1, ER(n,2/n)\text{ER}(n,2/n).
Refer to caption
(b) β0.5{\beta}\equiv 0.5, ER(n,5/n)\text{ER}(n,5/n).
Refer to caption
(c) β1{\beta}\equiv 1, CMn(δ3)\text{CM}_{n}(\delta_{3}).
Refer to caption
(d) β0.5{\beta}\equiv 0.5, CMn(δ3)\text{CM}_{n}(\delta_{3}).
Figure 1. Time evolution of the fraction of susceptible (SS) and infected (II) individuals on a finite graph (n=250n=250), with constant β\beta, obtained through simulations (Sim), along with the asymptotically exact values (ODE) given by Theorem 2.5. The simulations are obtained through 500500 iterations, resampling the random graphs at each iteration, and are plotted with 95%95\% confidence intervals.
Remark 2.6.

For simplicity, we restrict our attention to i.i.d. initial conditions, though the techniques in our proofs extend to more general initial conditions, as long as they satisfy certain symmetry properties between the laws of the initial states and that of the random graphs, and satisfy the Markov random field property mentioned above. In the case where the limit tree 𝒯{\mathcal{T}} is the κ\kappa-regular tree TκT_{\kappa}, the symmetry conditions correspond to the law of XTκ(0)X^{T_{\kappa}}(0) being isomorphism invariant, see [Lacker2021marginal, Remark 3.16].

Remark 2.7.

We also mention that, while Theorem 2.5 is stated for (sparse) ER and CM graphs, our techniques extend to a broader class of graphs, namely to any graph sequence {Gn}n\{G_{n}\}_{n\in{\mathbb{N}}} that converges locally weakly in probability to a UGW tree. All results in this paper hold if we replace Assumption B with the assumption that for some θ𝒫(0)\theta\in{\mathcal{P}}({\mathbb{N}}_{0}) with finite third moment and a UGW(θ)\text{UGW}(\theta) tree 𝒯{\mathcal{T}},

1nvGv𝟏{BrGn(v)H}n𝑝(Br𝒯()H))\frac{1}{n}\sum_{v\in G_{v}}{\bm{1}}_{\{B^{G_{n}}_{r}(v)\simeq H\}}\xrightarrow[n\rightarrow\infty]{p}{\mathbb{P}}(B^{\mathcal{T}}_{r}(\varnothing)\simeq H))

for every r0r\in{\mathbb{N}}_{0} and every rooted graph HH, where \simeq denotes graph isomorphism, and BrG(v)B_{r}^{G}(v) is a ball of radius rr around vGv\in G, that is, the subgraph induced by all vertices in GG that are at most rr edges away from vv.

As mentioned in the Introduction, in the special case when the infection and recovery rates β\beta and ρ\rho are constant in time and GnG_{n} is the configuration model, an ODE approximation similar to (2.4) was proposed in [Volz2008SIR, Volz2009epidemic] and shown to be asymptotically exact in [Decreusefond2012large, janson2015law]. However, Theorem 2.5 applies to the more general setting of time-varying rates, which is very relevant for applications, e.g., [chen2020time, hong2020estimation, london1973recurrent, dushoff2004dynamical], and more general graph classes (see Remark 2.7). Further, an advantage of our approach is that it allows for several important generalizations, including non-exponential recovery times, as elaborated upon in Remark 2.8 below, the SEIR model, presented in Section 2.3, and further extensions, discussed in Remark 2.11 below.

Remark 2.8.

A large part of the literature on the SIR process focuses on the case where recovery times are exponential random variables, that is, each individual recovers at some rate ρ\rho regardless of how long they have been infected, and the methods exploit this Markovian structure. If recovery times are not exponential, the resulting SIR dynamics are not Markov, and this makes their analysis significantly more challenging. In contrast, the local convergence tools that we used in the proof of Theorem 2.5 can still be used in this setting. Specifically, the hydrodynamic result in [Ganguly2022hydrodynamic] is still valid and shows that the fraction of individuals in each of the SIR states on a finite locally-tree like graph can be approximated by the root particle dynamics of the non-Markovian SIR process on the infinite tree. Further, a version of the conditional independence property of Proposition 4.7 can be established, the marginal root dynamics can be characterized as a piecewise deterministic Markov process, and its law characterized as the solution to a certain PDE. A complete analysis is deferred to future work.

2.3. SEIR Model

In this section, we extend our limit results to the Susceptible-Exposed-Infected-Recovered (SEIR) process. The SEIR process is a model of epidemics in which each individual can be in one of four possible states: in addition to the three states S,I,RS,\ I,\ R, of the SIR model, an individual can also be in the exposed state EE, when it has contracted the disease but is not yet able to infect its neighbors.

We define 𝒳¯:={S,E,I,R}\bar{\mathcal{X}}:=\{S,E,I,R\}. As in the case of the SIR model, the SEIR model on a (possibly random) graph GG can be modelled as a locally interacting Markov chain. We denote this process by X¯G{\bar{X}}^{G}. The SEIR process is governed by the graph GG and three functions β,ρ,λ:[0,)(0,)\beta,\rho,\lambda:[0,\infty)\rightarrow(0,\infty), with β\beta and ρ\rho, as for the SIR model, representing the infection and recovery rates, and λ\lambda now representing the time-dependent rate at which an individual transitions from having been exposed to being infectious. We assume that the initial states are i.i.d. with (X¯vG(0)=S)=s0,{\mathbb{P}}({\bar{X}}_{v}^{G}(0)=S)=s_{0}, (X¯vG(0)=E)=e0{\mathbb{P}}({\bar{X}}_{v}^{G}(0)=E)=e_{0} and (X¯vG(0)=I)=i0{\mathbb{P}}({\bar{X}}_{v}^{G}(0)=I)=i_{0} for some s0(0,1)s_{0}\in(0,1) and e0,i0[0,1]e_{0},i_{0}\in[0,1] such that s0+e0+i0=1s_{0}+e_{0}+i_{0}=1. At time tt, an individual vv jumps from SS to EE at the rate βt(X¯vG(t))=βtwv𝟏{X¯wG(t)=I}\beta_{t}{\mathcal{I}}({\bar{X}}^{G}_{\partial_{v}}(t-))=\beta_{t}\sum_{w\in\partial_{v}}{\bm{1}}_{\{{\bar{X}}^{G}_{w}(t-)=I\}}, from EE to II at the rate λt\lambda_{t}, and from II to RR at the rate ρt\rho_{t}. No other jumps are possible. Equipping 𝒳¯\bar{\mathcal{X}} with the ordering S<E<I<RS<E<I<R, the SEIR process is non-decreasing in the same sense as Remark 2.1.

Throughout the rest of the paper, we make the following assumption.

Assumption C.

The functions β\beta, λ\lambda and ρ\rho are continuous and there exist constants c1,c2(0,)c_{1},c_{2}\in(0,\infty) such that

c1<lim inftmin(βt,ρt,λt)<lim suptmax(βt,ρt,λt)<c2.c_{1}<\liminf_{t\rightarrow\infty}\min(\beta_{t},\ \rho_{t},\ \lambda_{t})<\limsup_{t\rightarrow\infty}\max(\beta_{t},\ \rho_{t},\ \lambda_{t})<c_{2}. (2.9)

2.3.1. Asymptotic Characterization of SEIR dynamics

Given a finite graph GG, we let

s¯G(t):=1|G|vG𝟏{X¯vG(t)=S},e¯G(t):=1|GvG𝟏{X¯vG(t)=E},i¯G(t):=1|G|vG𝟏{X¯vG(t)=I}.\displaystyle\begin{split}{\bar{s}}^{G}(t)&:=\frac{1}{|G|}\sum_{v\in G}{\bm{1}}_{\{{\bar{X}}_{v}^{G}(t)=S\}},\\ {\bar{e}}^{G}(t)&:=\frac{1}{|G}\sum_{v\in G}{\bm{1}}_{\{{\bar{X}}_{v}^{G}(t)=E\}},\\ {\bar{i}}^{G}(t)&:=\frac{1}{|G|}\sum_{v\in G}{\bm{1}}_{\{{\bar{X}}_{v}^{G}(t)=I\}}.\end{split} (2.10)

We start by establishing the existence and uniqueness of the solution to a certain system of ordinary differential equations that we use in our main result.

Proposition 2.9.

Suppose that θ𝒫(0)\theta\in{\mathcal{P}}({\mathbb{N}}_{0}) has a finite third moment and let s0(0,1)s_{0}\in(0,1) and e0,i0[0,1]e_{0},i_{0}\in[0,1] satisfy s0+e0+i0=1s_{0}+e_{0}+i_{0}=1. Then there exists a unique solution (gS,gE,gI,GI)(g_{S},\ g_{E},\ g_{I},\ G_{I}) to the following system of ODEs:

{g˙S=βgSgI(1k=0kθ^(k)ekGIj=0θ^(j)ejGI),g˙E=βgSgIk=0kθ^(k)ekGIj=0θ^(j)ejGIgE(λβgI),g˙I=λgEgI(ρ+ββgI),G˙I=βgI,\displaystyle\begin{cases}\dot{g}_{S}={\beta}g_{S}g_{I}\left(1-\frac{\sum_{k=0}^{\infty}k{\hat{\theta}}(k)e^{-kG_{I}}}{\sum_{j=0}^{\infty}{\hat{\theta}}(j)e^{-jG_{I}}}\right),\\ \dot{g}_{E}={\beta}g_{S}g_{I}\frac{\sum_{k=0}^{\infty}k{\hat{\theta}}(k)e^{-kG_{I}}}{\sum_{j=0}^{\infty}{\hat{\theta}}(j)e^{-jG_{I}}}-g_{E}({\lambda}-{\beta}g_{I}),\\ \dot{g}_{I}={\lambda}g_{E}-g_{I}({\rho}+{\beta}-{\beta}g_{I}),\\ \dot{G}_{I}=\beta g_{I},\end{cases} (2.11)

with initial conditions

{GI(0)=0,gm(0)=s0𝟏{m=S}+e0𝟏{m=E}+i0𝟏{m=I},m𝒳¯.\displaystyle\begin{cases}G_{I}(0)=0,\\ g_{m}(0)=s_{0}{\bm{1}}_{\{m=S\}}+e_{0}{\bm{1}}_{\{m=E\}}+i_{0}{\bm{1}}_{\{m=I\}},&m\in\bar{\mathcal{X}}.\end{cases} (2.12)

The proof of Proposition 2.9 is similar to that of Proposition 2.4. A brief outline is given at the end of Appendix B.

Given gS,gE,gI,GIg_{S},\ g_{E},\ g_{I},\ G_{I} as in Proposition 2.9 and MθM_{\theta} as in (2.6), define

s¯()(t):=s0Mθ(GI(t)),e¯()(t):=e0tλu𝑑u(e0+s00tMθ(GI(u))GI(u)e0uλτ𝑑τ𝑑u),i¯()(t):=e0tρu𝑑u(i0+0tλue0u(ρsλs)𝑑s(e0+s00uMθ(GI(τ))GI(τ))e0τλs𝑑sdτ)du).\displaystyle\begin{split}&{\bar{s}}^{(\infty)}(t):=s_{0}M_{\theta}(-G_{I}(t)),\\ &{\bar{e}}^{(\infty)}(t):=e^{-\int_{0}^{t}\lambda_{u}du}\left(e_{0}+s_{0}\int_{0}^{t}M^{\prime}_{\theta}(-G_{I}(u))G_{I}^{\prime}(u)e^{\int_{0}^{u}\lambda_{\tau}d\tau}du\right),\\ &{\bar{i}}^{(\infty)}(t):=e^{-\int_{0}^{t}\rho_{u}du}\left(i_{0}+\int_{0}^{t}\lambda_{u}e^{\int_{0}^{u}(\rho_{s}-\lambda_{s})ds}\left(e_{0}+s_{0}\int_{0}^{u}M^{\prime}_{\theta}(-G_{I}(\tau))G_{I}^{\prime}(\tau))e^{\int_{0}^{\tau}\lambda_{s}ds}d\tau\right)du\right).\end{split} (2.13)

We can now state our characterization of the large nn dynamics of the SEIR process.

Theorem 2.10.

Suppose that the sequence of random graphs {Gn}n\{G_{n}\}_{n\in{\mathbb{N}}} and θ(0)\theta\in{\mathbb{P}}({\mathbb{N}}_{0}) satisfy Assumption B. Let θ^{\hat{\theta}} be the size-biased version of θ\theta, as defined in (2.2), suppose s¯Gn(0)s0,{\bar{s}}^{G_{n}}(0)\rightarrow s_{0}, e¯Gn(0)e0{\bar{e}}^{G_{n}}(0)\rightarrow e_{0}, and i¯Gn(0)i0,{\bar{i}}^{G_{n}}(0)\rightarrow i_{0}, with s0(0,1)s_{0}\in(0,1) and s0+e0+i0=1s_{0}+e_{0}+i_{0}=1, and set s¯(),{\bar{s}}^{(\infty)}, e¯(){\bar{e}}^{(\infty)} and i¯(){\bar{i}}^{(\infty)} be as defined in (2.13). Then, as nn\rightarrow\infty,

s¯Gn(t)𝑝s¯()(t),e¯Gn(t)𝑝e¯()(t),i¯Gn(t)𝑝i¯()(t),\displaystyle\begin{split}{\bar{s}}^{G_{n}}(t)\xrightarrow{p}{\bar{s}}^{(\infty)}(t),\qquad{\bar{e}}^{G_{n}}(t)\xrightarrow{p}{\bar{e}}^{(\infty)}(t),\qquad{\bar{i}}^{G_{n}}(t)\xrightarrow{p}{\bar{i}}^{(\infty)}(t),\end{split}

uniformly for t[0,)t\in[0,\infty).

The proof of Theorem 2.10 is given in Section 4.2.3, and follows a similar approach as for the SIR model, although the details are more involved.

In Figure 2 we compare our asymptotically exact approximation to values of s¯Gn,e¯Gn{\bar{s}}^{G_{n}},\ {\bar{e}}^{G_{n}} and i¯Gn{\bar{i}}^{G_{n}} for an Erdős-Rényi  graph obtained by Monte Carlo simulations (500500 iterations, plotted with 95%95\% confidence intervals). Once again, our approximation closely tracks the simulation results, even for relatively small nn.

Refer to caption
(a) n=200n=200, λ0.5\lambda\equiv 0.5.
Refer to caption
(b) n=200n=200, λ2\lambda\equiv 2.
Refer to caption
(c) n=100n=100, λ0.5\lambda\equiv 0.5.
Refer to caption
(d) n=100n=100, λ2\lambda\equiv 2.
Figure 2. Time evolution of the fraction of susceptible (S), exposed (E) and infected (I) individuals on an ER(n,3/n)\text{ER}(n,3/n) graph, with βρ1\beta\equiv\rho\equiv 1. We compare the asymptotically exact values (ODE) given by Theorem 2.10 with the fraction obtained through Monte Carlo simulations (Sim). Simulations are obtained through 500500 iterations, and are shows with 95%95\% confidence intervals.
Remark 2.11.

The result in Theorem 2.10 can be further extended to more general compartmental models that are widely used in the epidemiology literature in order to account for different viral strains and treatment options, for example, see [duchamps2023general, foutel2022from, he2020SEIR, mwalili2020SEIR, hyman1999differential]. These allow for a susceptible state SS and mm\in{\mathbb{N}} post-infection states {I1\{I_{1}, I2I_{2}, … Im}I_{m}\}. Supposing that each individual’s transitions among post-infection states do not depend on the states of its neighbors, under Assumption B and continuity assumptions analogous to Assumption C, the hydrodynamic result in [Ganguly2022hydrodynamic] holds. If in addition one assumes that no transitions from post-infection states to state SS are possible, a version of the independence property of Proposition 4.7 can be established, thus leading to a result analogous to Theorem 2.10. We defer a full account of this general setting to future work.

3. Results on Outbreak Size

An important quantity of interest in the study of epidemic dynamics is the outbreak size, which is the fraction of individuals ever infected, in the interval [0,)[0,\infty). By the monotonicity of the SIR and SEIR processes (Remark 2.1), the outbreak size is equal to 11 minus the limit, as tt\rightarrow\infty, of the fraction of susceptible individuals at time tt. In Section 3.1 and Section 3.2, we characterize the large-time behavior for the SIR and SEIR processes respectively, as the size of the population approaches infinity. In Section 3.3 we compare our asymptotically exact estimate of the outbreak size with a mean-field approximation for the special case of the SIR process on random regular graphs with constant infection and recovery rates.

3.1. Outbreak Size for SIR Model

Given a sequence of graphs {Gn}n\{G_{n}\}_{n\in{\mathbb{N}}} satisfying Assumption B, we let sGn():=limtsGn(t)s^{G_{n}}(\infty):=\lim_{t\rightarrow\infty}s^{G_{n}}(t) for nn\in{\mathbb{N}}. We compute the limit of this quantity as nn\rightarrow\infty, by first showing that limnsGn()=limts()(t)\lim_{n\rightarrow\infty}s^{G_{n}}(\infty)=\lim_{t\rightarrow\infty}s^{(\infty)}(t), where s()s^{(\infty)}, given in (2.7), is the hydrodynamic limit of the fraction of susceptible individuals, by Theorem 2.5. We recall that MνM_{\nu} denotes the moment generating function of ν𝒫(0)\nu\in{\mathcal{P}}({\mathbb{N}}_{0}).

Theorem 3.1.

Let {Gn}n\{G_{n}\}_{n\in{\mathbb{N}}} and θ𝒫(0)\theta\in{\mathcal{P}}({\mathbb{N}}_{0}) satisfy Assumption B. Let θ^{\hat{\theta}} be the size-biased version of θ\theta, as defined in (2.2). Then, assuming that limnsGn(0)=s0(0,1)\lim_{n\rightarrow\infty}s^{G_{n}}(0)=s_{0}\in(0,1),

limnlimtsGn(t)=limts()(t)=s0Mθ(0βufI(u)𝑑u),\lim_{n\rightarrow\infty}\lim_{t\rightarrow\infty}s^{G_{n}}(t)=\lim_{t\rightarrow\infty}s^{(\infty)}(t)=s_{0}M_{\theta}\left(-\int_{0}^{\infty}\beta_{u}f_{I}(u)du\right),

where fIf_{I} is defined by (2.4)-(2.5). Moreover, :=0βufI(u)𝑑u{\mathcal{F}}:=\int_{0}^{\infty}\beta_{u}f_{I}(u)du is finite and satisfies

+log(Mθ^())log(1e0e0uβτfI(τ)𝑑τρufI(u)𝑑u)+log(s0)=0,{\mathcal{F}}+\log(M_{{\hat{\theta}}}(-{\mathcal{F}}))-\log\left(1-e^{{\mathcal{F}}}\int_{0}^{\infty}e^{-\int_{0}^{u}\beta_{\tau}f_{I}(\tau)d\tau}\rho_{u}f_{I}(u)du\right)+\log(s_{0})=0, (3.1)

Furthermore, if there exists r(0,)r\in(0,\infty) such that ρt/βt=r\rho_{t}/\beta_{t}=r for all t[0,)t\in[0,\infty), then equation (3.1) is equivalent to

+log(Mθ^())log(1+r(1e))+log(s0)=0,{\mathcal{F}}+\log(M_{{\hat{\theta}}}(-{\mathcal{F}}))-\log(1+r(1-e^{{\mathcal{F}}}))+\log(s_{0})=0, (3.2)

which has a unique strictly positive solution {\mathcal{F}}.

The proof of Theorem 3.1 is given in Section 3.

Remark 3.2.

When the ratio ρt/βt\rho_{t}/\beta_{t} is constant in time, the final outbreak size depends on ρ\rho and β\beta only through their ratio. This is well known when β\beta and ρ\rho are both constant, and in that case it is common in the SIR literature to fix ρ1\rho\equiv 1 with no loss of generality, by re-scaling time. Theorem 3.1 shows that, when the ratio ρ/β\rho/\beta is not constant, the ratio no longer determines the outbreak size, and instead the time evolution of both β\beta and ρ\rho influence the outbreak size. Figure 3 illustrates this phenomenon. It plots s()(t)s^{(\infty)}(t), defined in (2.7), which by Theorem 2.5 is the large-nn asymptotic fraction of susceptible individuals, for two SIR processes with the same ratio rt=ρt/βtr_{t}=\rho_{t}/\beta_{t} for all t0t\geq 0, though different β\beta and ρ\rho, which lead to dramatically different outbreak sizes.

\floatbox

[\capbeside\thisfloatsetupcapbesideposition=right,center,capbesidewidth=0.5]figure[\FBwidth] Refer to caption

Figure 3. Large-nn limit of the fraction of susceptible individuals (i.e., s()(t)s^{(\infty)}(t) from (2.7)) for the SIR process on a uniform 33-regular graph, as a function of time. For either curves, the ratio ρt/βt\rho_{t}/\beta_{t} is equal to 1.5+sin(πt)1.5+sin(\pi t). In one case, ρt\rho_{t} increases linearly from 0.50.5 to 1.51.5 for t[0,10]t\in[0,10] and it then stays constant. In the other, ρt\rho_{t} decreases linearly from 1.51.5 to 0.50.5 for t[0,10]t\in[0,10], and it then stays constant.

Next, for each time-dependent β\beta and ρ\rho we identify constant infection and recovery rates that lead to the same outbreak size. These effective rates are unique only up to multiplication by the same constant, and so we identify them by their ratio. For given β,ρ\beta,\ \rho, we define Ψβ,ρ:[0,)[0,]\Psi_{\beta,\rho}:[0,\infty)\rightarrow[0,\infty] as

Ψβ,ρ(z)=z+log(Mθ^(z))log(1ez0e0uβτfI(τ)𝑑τρufI(u)𝑑u)+log(s0),\Psi_{\beta,\rho}(z)=z+\log(M_{{\hat{\theta}}}(-z))-\log\left(1-e^{z}\int_{0}^{\infty}e^{-\int_{0}^{u}\beta_{\tau}f_{I}(\tau)d\tau}\rho_{u}f_{I}(u)du\right)+\log(s_{0}), (3.3)

where we set log(x)=\log(x)=-\infty for any x0x\leq 0, and where fIf_{I} is defined by (2.4)-(2.5) for some fixed s0(0,1)s_{0}\in(0,1) with the given β\beta and ρ\rho. For r(0,)r\in(0,\infty), we also define Ψr:[0,)[0,]\Psi_{r}:[0,\infty)\rightarrow[0,\infty] as

Ψr(z)=z+log(Mθ^(z))log(1+r(1ez))+log(s0).\Psi_{r}(z)=z+\log(M_{{\hat{\theta}}}(-z))-\log\left(1+r(1-e^{z})\right)+\log(s_{0}). (3.4)

The following result shows that for every pair of rate functions β\beta and ρ\rho satisfying Assumption A, there exists a constant r^:=r^(β,ρ)\hat{r}:=\hat{r}(\beta,\rho) so that the outbreak size of an SIR process with rates β\beta and ρ\rho, and that of a SIR process with constant infection rate 11 and constant recovery r^\hat{r} are the same (as nn\rightarrow\infty). In particular, we observe that this is not achieved by naively replacing ρ\rho and β\beta with their respective average (over time) values, nor by taking r^\hat{r} to be the (time) average of ρt/βt\rho_{t}/\beta_{t}.

Lemma 3.3.

Let θ𝒫(0)\theta\in{\mathcal{P}}({\mathbb{N}}_{0}) have a finite third moment, and suppose that ρ,β\rho,\ \beta satisfy Assumption A. Let β,ρ:=0βtfI(t){\mathcal{F}}_{\beta,\rho}:=\int_{0}^{\infty}\beta_{t}f_{I}(t), where fIf_{I} is defined by (2.4)-(2.5). Then there exists a unique r^(0,)\hat{r}\in(0,\infty) such that Ψr^(β,ρ)=0\Psi_{\hat{r}}({\mathcal{F}}_{\beta,\rho})=0, namely

r^=0e0tβufI(u)𝑑uρtfI(t)𝑑t1e0βufI(u)𝑑u.\hat{r}=\frac{\int_{0}^{\infty}e^{-\int_{0}^{t}\beta_{u}f_{I}(u)du}\rho_{t}f_{I}(t)dt}{1-e^{-\int_{0}^{\infty}\beta_{u}f_{I}(u)du}}. (3.5)
Proof.

We start by showing that r^(0,)\hat{r}\in(0,\infty). We know that β,ρ=0βtfI(t)𝑑t<{\mathcal{F}}_{\beta,\rho}=\int_{0}^{\infty}\beta_{t}f_{I}(t)dt<\infty by Theorem 3.1. By Assumption A and (2.4), tβtfI(t)t\rightarrow\beta_{t}f_{I}(t) is continuous, non-negative, and bounded away from zero near t=0t=0, and so β,ρ>0{\mathcal{F}}_{\beta,\rho}>0. Letting c1,c2(0,)c_{1},c_{2}\in(0,\infty) be constants such that (2.1) holds, we have

0e0tβufI(u)𝑑uρtfI(t)𝑑t=0e0tβufI(u)𝑑uρtβtβtfI(t)𝑑tc2c1(1eβ,ρ)<.\displaystyle\begin{split}\int_{0}^{\infty}e^{-\int_{0}^{t}\beta_{u}f_{I}(u)du}\rho_{t}f_{I}(t)dt=\int_{0}^{\infty}e^{-\int_{0}^{t}\beta_{u}f_{I}(u)du}\frac{\rho_{t}}{\beta_{t}}\beta_{t}f_{I}(t)dt\leq\frac{c_{2}}{c_{1}}(1-e^{-{\mathcal{F}}_{\beta,\rho}})<\infty.\end{split}

Similarly, note that 0e0tβufI(u)𝑑uρtfI(t)𝑑t>c1c21(1eβ,ρ)>0\int_{0}^{\infty}e^{-\int_{0}^{t}\beta_{u}f_{I}(u)du}\rho_{t}f_{I}(t)dt>c_{1}c_{2}^{-1}(1-e^{-{\mathcal{F}}_{\beta,\rho}})>0. Setting r^\hat{r} as in (3.5), by (3.4), we have

Ψr^(z)=z+log(Mθ^(z))log(1+0e0tβufI(u)𝑑uρtfI(t)𝑑t1eβ,ρ(1ez))+log(s0).\Psi_{\hat{r}}(z)=z+\log(M_{{\hat{\theta}}}(-z))-\log\left(1+\frac{\int_{0}^{\infty}e^{-\int_{0}^{t}\beta_{u}f_{I}(u)du}\rho_{t}f_{I}(t)dt}{1-e^{-{\mathcal{F}}_{\beta,\rho}}}(1-e^{z})\right)+\log(s_{0}).

Evaluating this at z=β,ρz={\mathcal{F}}_{\beta,\rho} using (3.3), and observing that (1ez)/(1ez)=ez(1-e^{z})/(1-e^{-z})=-e^{z}, we have

Ψr^(β,ρ)=β,ρ+log(Mθ^(β,ρ))log(1eβ,ρ0e0tβufI(u)𝑑uρtfI(t)𝑑t)+log(s0)=Ψβ,ρ(β,ρ),\displaystyle\begin{split}\Psi_{\hat{r}}({\mathcal{F}}_{\beta,\rho})&={\mathcal{F}}_{\beta,\rho}+\log(M_{{\hat{\theta}}}(-{\mathcal{F}}_{\beta,\rho}))-\log\left(1-e^{{\mathcal{F}}_{\beta,\rho}}\int_{0}^{\infty}e^{-\int_{0}^{t}\beta_{u}f_{I}(u)du}\rho_{t}f_{I}(t)dt\right)+\log(s_{0})\\ &=\Psi_{\beta,\rho}({\mathcal{F}}_{\beta,\rho}),\end{split}

which is zero by Theorem 3.1. This shows the existence of r^(0,)\hat{r}\in(0,\infty) such that Ψr^(β,ρ)=0\Psi_{\hat{r}}({\mathcal{F}}_{\beta,\rho})=0.

For uniqueness, observe that for each z(0,)z\in(0,\infty) the map rΨr(z)r\rightarrow\Psi_{r}(z) is non-decreasing, and strictly increasing on {r:Ψr(z)<}\left\{r\ :\ \Psi_{r}(z)<\infty\right\}. Let r{\mathcal{F}}_{r} be the unique zero of Ψr\Psi_{r}. It follows that r{\mathcal{F}}_{r} is strictly decreasing in rr and therefore there is a one-to-one correspondence between rr and r{\mathcal{F}}_{r}. ∎

We conclude this section with a brief discussion of periodic parameters. For simplicity, we fix ρ1\rho\equiv 1 and we consider periodic infection rates that could model, for instance, seasonality effects of the infectivity of a pathogen. For a given amplitude A[0,1)A\in[0,1), period ω>0\omega>0 and δ[0,1]\delta\in[0,1] we set βt=1+Asin((t+δω)2π/ω)\beta_{t}=1+A\sin((t+\delta\omega)2\pi/\omega). Here, δ\delta is a parameter controlling the phase of the periodic rate at time zero. Note that if the period length is large enough compared to the average infection rate and recovery rate, the outbreak dies out before the full length of the period, and so, while the average of βt\beta_{t} over the period is 11, the average infection rate during the time the epidemic is “active” (i.e, there are individuals in state II) will be close to β0\beta_{0}. Because of this, we expect δ\delta to have a greater impact on the outbreak size when ω\omega is large. This is borne out by Figure 4, which plots the outbreak size as a function of AA for various δ\delta and ω\omega. We see that in every case other than large ω\omega and small δ\delta, the outbreak size is decreasing in AA. This suggests the following conjecture, which we leave for future investigation.

Refer to caption
(a) ω=0.5\omega=0.5.
Refer to caption
(b) ω=1\omega=1.
Refer to caption
(c) ω=5\omega=5.
Refer to caption
(d) ω=10\omega=10.
Figure 4. Large nn limit of the final size of a SIR outbreak on a 33-regular graph with periodic infection rate, as a function of the amplitude AA, obtained numerically from s()(t)s^{(\infty)}(t) as in Theorem 2.5. In all cases ρ1\rho\equiv 1.
Conjecture 3.4.

Let A[0,1),ω>0A\in[0,1),\ \omega>0 and δ[0,1]\delta\in[0,1]. Define βt=1+Asin((t+δω)2π/ω)\beta_{t}=1+A\sin((t+\delta\omega)2\pi/\omega). There exists ω¯>0{\bar{\omega}}>0 such that, for all ω<ω¯\omega<{\bar{\omega}}, the asymptotic outbreak size 1s()()1-s^{(\infty)}(\infty) is decreasing in AA.

3.2. Outbreak Size for the SEIR Model

We now turn to the characterization of the outbreak size of an SEIR process. Recall the definition of MζM_{\zeta} for ζ𝒫(0)\zeta\in{\mathcal{P}}({\mathbb{N}}_{0}) given in (2.6).

Theorem 3.5.

Let {Gn}n\{G_{n}\}_{n\in{\mathbb{N}}} and θ\theta satisfy Assumption B. Then

limnlimts¯Gn(t)=limts¯()(t)=s0Mθ(0βugI(u)𝑑u)\lim_{n\rightarrow\infty}\lim_{t\rightarrow\infty}{\bar{s}}^{G_{n}}(t)=\lim_{t\rightarrow\infty}{\bar{s}}^{(\infty)}(t)=s_{0}M_{\theta}\left(-\int_{0}^{\infty}\beta_{u}g_{I}(u)du\right)

where 0βugI(u)du=:\int_{0}^{\infty}\beta_{u}g_{I}(u)du=:{\mathcal{F}} satisfies

+log(Mθ^())log(1e0e0uβτgI(τ)𝑑τρugI(u)𝑑u)+log(s0)=0,{\mathcal{F}}+\log(M_{{\hat{\theta}}}(-{\mathcal{F}}))-\log\left(1-e^{{\mathcal{F}}}\int_{0}^{\infty}e^{-\int_{0}^{u}\beta_{\tau}g_{I}(\tau)d\tau}\rho_{u}g_{I}(u)du\right)+\log(s_{0})=0, (3.6)

for gIg_{I} as in (2.11). Furthermore, if there exists r(0,)r\in(0,\infty) such that ρt/βt,=r\rho_{t}/\beta_{t},=r for all t[0,)t\in[0,\infty), equation (3.6) is equivalent to

Ψr()=0,\Psi_{r}({\mathcal{F}})=0, (3.7)

where Ψr\Psi_{r} is defined in (3.4).

Refer to caption
(a) β0.5\beta\equiv 0.5, ρ1\rho\equiv 1.
Refer to caption
(b) β1\beta\equiv 1, ρ1\rho\equiv 1.
Refer to caption
(c) βt=0.5+0.1min(t,10)\beta_{t}=0.5+0.1\min(t,10), ρ1\rho\equiv 1.
Refer to caption
(d) βt=1.50.1max(t,10)\beta_{t}=1.5-0.1\max(t,10), ρ1\rho\equiv 1.
Figure 5. Large nn limit fraction of susceptible individuals for SEIR processes (s¯()(t){\bar{s}}^{(\infty)}(t) from (2.13)) with different values of λ\lambda. In all cases, the interaction graph is the random 33-regular graph, e0=0.01e_{0}=0.01 and s0=1e0s_{0}=1-e_{0}. In (a) and (b) both β\beta and ρ\rho are constant. In (c) (resp. (d)) ρ\rho is held constant and β\beta grows linearly from 0.50.5 to 1.51.5 (resp. decreases linearly from 1.51.5 to 0.50.5) and is then constant.

Theorem 3.5 shows that when the ratio tβt/ρtt\mapsto\beta_{t}/\rho_{t} is constant, the final outbreak size does not depend on λ\lambda and it coincides with the outbreak size of a SIR process with the same infection rate, recovery rate, and initial condition s0s_{0}. On the other hand, when the ratio is not constant, the rate λ\lambda affects the outbreak size. Figure 5 plots s¯()(t){\bar{s}}^{(\infty)}(t), as defined in (2.13), which by Theorem 2.10 is the large nn limit of the of the fraction of susceptible individuals, for several SEIR processes on a random 33-regular graph. For fixed constants β,\beta, and ρ\rho, but different values of constant λ\lambda, the time-dynamics can vary significantly, but the final fraction of susceptible individuals does not depend on λ\lambda. In contrast, when ρ/β\rho/\beta changes with time, the final fraction of susceptible individuals (and hence, the outbreak size) varies with λ\lambda.

In Figure 6, we set all rates as constant, and we show that the time evolution of the sum of the fractions of infected and exposed individuals in the SEIR process can be markedly different from that of the fraction of infected individuals in an SIR process, despite the fact that the final outbreak sizes coincide. We leave as a future research direction the problem of understanding the impact of λ\lambda on the SEIR dynamics for finite tt - for example, how does λ\lambda impact the maximum number of individuals that have ever been infected in any given time period?

Refer to caption
(a) β0.5\beta\equiv 0.5, ρ1\rho\equiv 1.
Refer to caption
(b) β1\beta\equiv 1, ρ1\rho\equiv 1.
Figure 6. Large nn limit of the fraction of individuals who are exposed or infected in the SEIR process on the uniform 33-regular (i.e., e¯()(t)+i¯()(t){\bar{e}}^{(\infty)}(t)+{\bar{i}}^{(\infty)}(t) from (2.13)) for λ1\lambda\equiv 1 and λ0.5\lambda\equiv 0.5, along with the large nn limit of the fraction of infected individuals on the SIR process (i()(t)i^{(\infty)}(t) from (2.7)) on the same graph and with same ρ\rho and β\beta. Individuals are initially infected with probability 0.010.01, and are otherwise susceptible.

3.3. Comparison with the Mean-Field approximation for the SIR model

In this section, we restrict our attention to the SIR process on the uniform κ\kappa-regular graph, with the ratio ρ/β\rho/\beta being constant in time, and compare the asymptotically exact outbreak size with the corresponding mean-field approximation. We first start by observing that on κ\kappa-regular graphs, the characterization (3.2) of the outbreak size can be simplified further as follows.

Corollary 3.6.

Let κ{1}\kappa\in{\mathbb{N}}\setminus\{1\}. Let {Gn}n\{G_{n}\}_{n\in{\mathbb{N}}} be such that for every nn\in{\mathbb{N}}, GnG_{n} is chosen uniformly among all κ\kappa-regular graphs with nn vertices, or equivalently GnG_{n} is a CMn(δκ)\text{CM}_{n}(\delta_{\kappa}) graph. Suppose that there exists r(0,)r\in(0,\infty) such that ρt/βt=r\rho_{t}/\beta_{t}=r for all t[0,)t\in[0,\infty). Then, it follows that

limts()(t)=σκ.\lim_{t\rightarrow\infty}s^{(\infty)}(t)=\sigma_{\kappa}.

where σκ(0,s0]\sigma_{\kappa}\in(0,s_{0}] is the unique solution in (0,s0)(0,s_{0}) of the equation

ϕκ(z):=zκ2κs02κ(1+r)+rz1κs01κ=0.\phi_{\kappa}(z):=z^{\frac{\kappa-2}{\kappa}}s_{0}^{\frac{2}{\kappa}}-(1+r)+rz^{-\frac{1}{\kappa}}s_{0}^{\frac{1}{\kappa}}=0. (3.8)

In particular, we have

σ2=s0(1+r(1s0))2.\sigma_{2}=\frac{s_{0}}{(1+r(1-s_{0}))^{2}}.
Proof.

Fix κ{1}\kappa\in{\mathbb{N}}\setminus\{1\} and set θ=δκ\theta=\delta_{\kappa}. It is immediate from (2.2) that the size-biased distribution θ^{\hat{\theta}} is equal to δκ1\delta_{\kappa-1}. For any kk\in{\mathbb{N}}, uu\in{\mathbb{R}}, we have that Mδk(u)=ekuM_{\delta_{k}}(u)=e^{ku}. By Theorem 3.1, the final fraction of susceptible individuals s()()s^{(\infty)}(\infty) is equal to s0eκs_{0}e^{-\kappa{\mathcal{F}}}, where {\mathcal{F}} is the solution of equation (3.2), which for θ^=δκ1{\hat{\theta}}=\delta_{\kappa-1} reduces to

(2κ)log(1+r(1e))+log(s0)=0.(2-\kappa){\mathcal{F}}-\log(1+r(1-e^{\mathcal{F}}))+\log(s_{0})=0.

By a simple arithmetic manipulation, σκ:=s0eκ\sigma_{\kappa}:=s_{0}e^{-\kappa{\mathcal{F}}} satisfies equation (3.8). Uniqueness of the solution to ϕκ(z)=0\phi_{\kappa}(z)=0 follows since ϕκ(z)=0\phi_{\kappa}^{\prime}(z)=0 holds for at most one value of zz, namely for z=(s01/κr/(κ2))κ/(κ1)z=(s_{0}^{-1/\kappa}r/(\kappa-2))^{\kappa/(\kappa-1)}.  ∎

Figure 7 plots the analytic values of σκ\sigma_{\kappa} obtained from Corollary 3.6 versus simulated values for different values of nn and κ\kappa. We ran 200200 iterations for each pair (n,(n, κ)\kappa), sampling a new graph at every iteration. As shown therein, the limit appears to be a good approximation for graphs of moderate size (namely, with n150n\geq 150). We leave for future research the problem of finding accurate error bounds for finite nn.

\floatbox

[\capbeside\thisfloatsetupcapbesideposition=right,center,capbesidewidth=0.5]figure[\FBwidth] Refer to caption

Figure 7. Outbreak size for the SIR process with β0.5\beta\equiv 0.5 and ρ1\rho\equiv 1 obtained from Monte Carlo simulations on random κ\kappa-regular graphs on nn vertices (with 95%95\% confidence intervals), compared with the asymptotically exact value given by Corollary 3.6 (ODE). Simulations are obtained via 100100 iterations each, with resampling of the graph at each iteration.

By Theorem 3.1 and Corollary 3.6, σκ\sigma_{\kappa} is an asymptotically (in nn) exact approximation of the total fraction of individuals ever infected on a SIR epidemic on a graph drawn uniformly among the κ\kappa-regular graphs on nn vertices. We now compare this approximation with a scaled mean-field approximation to the SIR model on a κ\kappa-regular graph which can be formulated via the following system of ODEs, see for example [Ganguly2022nonmarkovian, Section 7]:

ds^dt(t)\displaystyle\frac{d\hat{s}}{dt}(t) =\displaystyle= βtκs^(t)i^(t),\displaystyle-\beta_{t}\kappa\hat{s}(t)\hat{i}(t),
di^(t)dt\displaystyle\frac{d\hat{i}(t)}{dt} =\displaystyle= βtκs^(t)i^(t)ρti^(t),\displaystyle\beta_{t}\kappa\hat{s}(t)\hat{i}(t)-\rho_{t}\hat{i}(t),

with initial conditions s^(0)=s0\hat{s}(0)=s_{0}, i^(0)=i0=1s0\hat{i}(0)=i_{0}=1-s_{0}. When there exists r(0,)r\in(0,\infty) such that ρt/βt=r\rho_{t}/\beta_{t}=r for all tt, by performing a change of variables and solving the equation di^/ds^=1+ρ/(βκs^)d\hat{i}/d\hat{s}=-1+\rho/(\beta\kappa\hat{s}), it can be shown that limts^(t)=σ^κ\lim_{t\rightarrow\infty}\hat{s}(t)=\hat{\sigma}_{\kappa} where σ^κ=σ^κ(s0)\hat{\sigma}_{\kappa}=\hat{\sigma}_{\kappa}(s_{0}) is the unique solution in (0,s0)(0,s_{0}) of

ϕ^κ(σ^κ)=0,with ϕ^κ(z):=s0e1rκ(z1)z.{\hat{\phi}}_{\kappa}(\hat{\sigma}_{\kappa})=0,\qquad\text{with }\quad{\hat{\phi}}_{\kappa}(z):=s_{0}e^{{\frac{1}{r}\kappa(z-1)}}-z. (3.9)
Refer to caption
(a) β0.1\beta\equiv 0.1.
Refer to caption
(b) β0.5\beta\equiv 0.5.
Refer to caption
(c) β1\beta\equiv 1.
Figure 8. Outbreak size on large (n=400n=400 vertices) κ\kappa-regular configuration model graphs (Sim), with ρ1\rho\equiv 1, and s0=0.95s_{0}=0.95, along with the asymptotically exact value 1σκ1-\sigma_{\kappa} from Corollary 3.6(ODE), and the mean-field approximation 1σ^κ1-\hat{\sigma}_{\kappa} defined in (3.9) (MF). Simulations are obtained through 500500 iterations, resampling the graph at each iteration, and are plotted with 95%95\% confidence intervals.

Our next result shows that the mean-field approximation always yields a larger estimate of the outbreak size on random regular graphs compared to the true asymptotic value. This is further illustrated in Figure 8, which plots the mean-field prediction versus our prediction of the outbreak size on random κ\kappa-regular graphs.

Proposition 3.7.

Fix s0(0,1)s_{0}\in(0,1). For each κ{1}\kappa\in{\mathbb{N}}\setminus\{1\}, let σ^κ\hat{\sigma}_{\kappa} be as in (3.9), and σκ\sigma_{\kappa} be given as in Corollary 3.6. Then it follows that

σ^κ<σκ.\hat{\sigma}_{\kappa}<\sigma_{\kappa}. (3.10)
Proof.

Fix s0(0,1)s_{0}\in(0,1), r>0r>0. Recall that σκ\sigma_{\kappa} is the unique solution in (0,s0)(0,s_{0}) to ϕκ(z)=0\phi_{\kappa}(z)=0. From the proof of Corollary 3.6, we know that ϕκ(z)>0\phi_{\kappa}(z)>0 for z(0,σκ)z\in(0,\sigma_{\kappa}) and ϕκ(z)<0\phi_{\kappa}(z)<0 for z(σκ,s0)z\in(\sigma_{\kappa},s_{0}). Therefore, to show (3.10), it is enough to show that ϕκ(σ^κ)>0\phi_{\kappa}(\hat{\sigma}_{\kappa})>0. Using the fact that σ^κ=s0exp(κ(σ^κ1)/r)\hat{\sigma}_{\kappa}=s_{0}\exp(\kappa(\hat{\sigma}_{\kappa}-1)/r), and that ez>1+ze^{z}>1+z for every z>0z>0, we have

ϕκ(σ^κ)=ϕκ(s0eκr(σ^κ1))=(s0eκr(σ^κ1))κ2κs02κ+rs01/κ(s0eκr(σ^κ1))1/κ1r=e1r(κ2)(σ^κ1)s0+re1r(1σ^κ)1r>s0eκr(σ^κ1)(e21r(1σ^κ)1).\displaystyle\begin{split}\phi_{\kappa}(\hat{\sigma}_{\kappa})&=\phi_{\kappa}(s_{0}e^{\frac{\kappa}{r}(\hat{\sigma}_{\kappa}-1)})\\ &=\left(s_{0}e^{\frac{\kappa}{r}(\hat{\sigma}_{\kappa}-1)}\right)^{\frac{\kappa-2}{\kappa}}s_{0}^{\frac{2}{\kappa}}+rs_{0}^{1/\kappa}\left(s_{0}e^{\frac{\kappa}{r}(\hat{\sigma}_{\kappa}-1)}\right)^{-1/\kappa}-1-r\\ &=e^{\frac{1}{r}(\kappa-2)(\hat{\sigma}_{\kappa}-1)}s_{0}+re^{\frac{1}{r}(1-\hat{\sigma}_{\kappa})}-1-r\\ &>s_{0}e^{\frac{\kappa}{r}(\hat{\sigma}_{\kappa}-1)}(e^{2\frac{1}{r}(1-\hat{\sigma}_{\kappa})}-1).\end{split} (3.11)

Since the last expression is strictly positive, this concludes the proof.  ∎

4. Proofs of main results

In Section 4.1 we introduce a parameterized family of processes that interpolates between the SIR and SEIR processes. This allows us to prove some intermediate results simultaneously for both processes. In Section 4.2 we provide the proofs of Theorem 2.5 and Theorem 2.10. In Section 4.3 we prove Theorem 3.1 and Theorem 3.5. Throughout, {Gn}n\{G_{n}\}_{n\in{\mathbb{N}}} is a sequence of random graphs, θ𝒫(0)\theta\in{\mathcal{P}}({\mathbb{N}}_{0}), θ^{\hat{\theta}} is the size-biased version of θ\theta, as defined in (2.2), and 𝒯{\mathcal{T}} is a UGW(θ\theta) tree. We assume that {Gn}n\{G_{n}\}_{n\in{\mathbb{N}}} and θ\theta satisfy Assumption B and that the rates β,\beta, λ,\lambda, ρ:[0,)(0,)\rho:[0,\infty)\rightarrow(0,\infty) satisfy Assumption C.

4.1. The Hybrid S(E)IR Process

Fix the rates β,ρ,λ:[0,)(0,)\beta,\rho,\lambda:[0,\infty)\rightarrow(0,\infty) as in Section 2.3, the interpolation parameter α[0,1]\alpha\in[0,1], probabilities s0(0,1)s_{0}\in(0,1), and e0,i0[0,1]e_{0},\ i_{0}\in[0,1] with s0+e0+i0=1s_{0}+e_{0}+i_{0}=1. For a graph G=(V,E)G=(V,E), let ξG,α\xi^{{G,\alpha}} be a Markov chain on 𝒳¯V\bar{\mathcal{X}}^{V} describing the evolution of interacting individuals or particles indexed by the nodes of VV, where the state of each particle lies in the space 𝒳¯={S,E,I,R}\bar{\mathcal{X}}=\{S,E,I,R\}. The initial states are i.i.d. across particles, with common law p0p_{0} given by

p0(b)=s0𝟏{b=S}+e0𝟏{b=E}+i0𝟏{b=I},p_{0}(b)=s_{0}{\bm{1}}_{\left\{b=S\right\}}+e_{0}{\bm{1}}_{\left\{b=E\right\}}+i_{0}{\bm{1}}_{\left\{b=I\right\}}, (4.1)

b𝒳¯b\in\bar{\mathcal{X}}. Given y𝒳¯y\in\bar{\mathcal{X}}^{\ell} for some 0\ell\in{\mathbb{N}}_{0} (setting 𝒳¯0=\bar{\mathcal{X}}^{0}=\emptyset) recall that (y){\mathcal{I}}(y) denotes the number of entries of yy that are equal to II. At time t[0,)t\in[0,\infty), the jump rates for the jump processes ξvG,α\xi^{G,\alpha}_{v} representing the evolution of a particle vGv\in G are given as follows:

  • from SS to EE at rate αβt(ξvG,α(t))\alpha\beta_{t}{\mathcal{I}}(\xi^{{G,\alpha}}_{\partial v}(t-));

  • from SS to II at rate (1α)βt(ξvG,α(t))(1-\alpha)\beta_{t}{\mathcal{I}}(\xi^{{G,\alpha}}_{\partial v}(t-));

  • from EE to II at rate λt\lambda_{t};

  • from II to RR at rate ρt\rho_{t}.

No other transitions are allowed. When GG is finite, classical results guarantee the existence of the process ξG,α\xi^{G,\alpha} and the uniqueness of its law follows from standard results about finite-state continuous time Markov chains, see for instance [Ganguly2022hydrodynamic, Proposition 4.1]. We note that if α=1\alpha=1, ξG,α\xi^{G,\alpha} reduces to the SEIR model, and if α=0\alpha=0 (and e0=0e_{0}=0), ξG,α\xi^{G,\alpha} is the SIR model. Throughout, whenever α=0\alpha=0, we implicitly assume that e0=0e_{0}=0, and Assumption C can be substituted with Assumption A. We refer to ξG,α\xi^{G,\alpha} as the S(E)IR process. We also observe that the process ξG,α\xi^{G,\alpha} is non-decreasing for every GG and α\alpha, that is, for every vGv\in G and t,s[0,)t,s\in[0,\infty),

tsξvG,α(t)ξvG,α(s).t\leq s\ \Rightarrow\ \xi_{v}^{G,\alpha}(t)\leq\xi_{v}^{G,\alpha}(s). (4.2)

Since we are interested in studying the limit of the S(E)IR process on locally tree-like graphs GnG_{n} with |Gn||G_{n}|\rightarrow\infty and GnG_{n} converging to a limit random tree, we need to define ξ𝒯,α\xi^{{\mathcal{T}},\alpha} on a possibly infinite tree 𝒯{\mathcal{T}}. Intuitively, ξ𝒯,α\xi^{{\mathcal{T}},\alpha} is a Markov jump process with the same rates as described above, but due to randomness in the tree structure, a rigorous definition (and subsequent characterization of properties) is most conveniently expressed in terms of the following (standard) Ulam-Harris-Neveu labeling which identifies each realization of 𝒯{\mathcal{T}} with a subgraph of the graph of all possible vertices. The latter has vertex set 𝕍:={}(k=1k)\mathbb{V}:=\{\varnothing\}\cup(\cup_{k=1}^{\infty}{\mathbb{N}}^{k}), where \varnothing denotes the root, and edges {{v,vi}:v𝕍,i}\{\{v,vi\}:v\in\mathbb{V},i\in{\mathbb{N}}\}, where vivi denotes concatenation, with the convention u=u=u\varnothing u=u\varnothing=u for all u𝕍u\in\mathbb{V}. For n0n\in{\mathbb{N}}_{0}, we also let 𝕍n:={}(k=1nk)\mathbb{V}_{n}:=\{\varnothing\}\cup(\cup_{k=1}^{n}{\mathbb{N}}^{k}). Given a vertex v𝕍{}v\in\mathbb{V}\setminus\{\varnothing\}, denote by πv\pi_{v} its parent, defined to be the unique w𝕍w\in\mathbb{V} such that there exists kk\in{\mathbb{N}} with wk=vwk=v. The children of a vertex v𝕍v\in\mathbb{V} are defined to be the set Cv𝕍:={vi}iC^{\mathbb{V}}_{v}:=\{vi\}_{i\in{\mathbb{N}}}.

Given a tree 𝒯{\mathcal{T}} with root 𝒯\varnothing_{\mathcal{T}}, we identify it (uniquely up to root preserving automorphisms of 𝒯{\mathcal{T}}) as a subgraph of 𝕍\mathbb{V} via a map 𝒱{\mathcal{V}} from the vertex set of 𝒯{\mathcal{T}} to 𝕍\mathbb{V} such that

  1. (i)

    𝒱(𝒯)={\mathcal{V}}(\varnothing_{\mathcal{T}})=\varnothing;

  2. (ii)

    𝒱(𝒯)={m:md𝒯}{\mathcal{V}}(\partial^{\mathcal{T}}_{\varnothing})=\{m\in{\mathbb{N}}\ :\ m\leq d_{\varnothing}^{\mathcal{T}}\};

  3. (iii)

    for v𝒯v\in{\mathcal{T}} at graph distance111given two vertices vv and ww in a graph GG, the graph distance between them is the minimum number of edges on a path from vv to ww. kk\in{\mathbb{N}} from 𝒯\varnothing_{\mathcal{T}}, 𝒱(v)=uk{\mathcal{V}}(v)=u\in{\mathbb{N}}^{k} and 𝒱(v𝒯)={πu}{um:m,mdv𝒯1}{\mathcal{V}}(\partial^{\mathcal{T}}_{v})=\{\pi_{u}\}\cup\{um\ :\ m\in{\mathbb{N}},m\leq d^{\mathcal{T}}_{v}-1\}.

In order to represent elements in 𝒳¯𝒯\bar{\mathcal{X}}^{\mathcal{T}} as marks on 𝕍\mathbb{V}, we consider a new mark \star, and define 𝒳¯:=𝒳¯{}\bar{\mathcal{X}}_{\star}:=\bar{\mathcal{X}}\cup\{\star\}. Given x𝒳¯𝒯x\in\bar{\mathcal{X}}^{\mathcal{T}}, we extend it to an element in (𝒳¯)𝕍(\bar{\mathcal{X}}_{\star})^{\mathbb{V}} by setting xw=x_{w}=\star for all w𝕍𝒯w\in\mathbb{V}\setminus{\mathcal{T}}. Whenever we consider a graph 𝒯𝕍{\mathcal{T}}\subset\mathbb{V} and v𝒯v\in{\mathcal{T}}, we use v\partial_{v} and dvd_{v} denote neighborhoods and degrees with respect to 𝒯{\mathcal{T}}. We use 𝕍A\partial^{\mathbb{V}}A to refer to the boundary in 𝕍\mathbb{V} of a set A𝕍A\subset\mathbb{V}, and set v𝕍=𝕍{v}\partial^{\mathbb{V}}_{v}=\partial^{\mathbb{V}}\{v\} for v𝕍v\in\mathbb{V}. Given an interval 𝒮{\mathcal{S}}\subset{\mathbb{R}} and {\mathcal{M}} a metric space, let 𝒟(𝒮:){\mathcal{D}}({\mathcal{S}}:{\mathcal{M}}) be the space of càdlàg functions222right continuous functions with finite left limits at every tt in the interior of 𝒮{\mathcal{S}}. equipped with the Skorokhod topology. For xx a (possibly random) element in 𝒟(𝒮:){\mathcal{D}}({\mathcal{S}}:{\mathcal{M}}) and t𝒮t\in{\mathcal{S}}, we write x[t]:={x(s):s𝒮(,t]}x[t]:=\{x(s)\ :s\in{\mathcal{S}}\cap(-\infty,t]\} and x[t):={x(s):s𝒮(,t)}x[t):=\{x(s)\ :\ s\in{\mathcal{S}}\cap(-\infty,t)\}. Throughout, we write 𝒟:=𝒟([0,):𝒳¯]){\mathcal{D}}:={\mathcal{D}}([0,\infty):\bar{\mathcal{X}}]), and we set 𝒟{\mathcal{D}}_{\star} to be the union of 𝒟{\mathcal{D}} and the single element consisting of the constant-\star function.

For y𝒳¯y\in\bar{\mathcal{X}}_{\star}^{\infty} we write (y)=|{m:ym=I}|{\mathcal{I}}(y)=|\left\{m\ :\ y_{m}=I\right\}|. Also, for simplicity, we identify the states (S,E,I,R)(S,E,I,R) with the vector (0,1,2,3)(0,1,2,3), and the set of possible jumps with 𝒥={1,2}{\mathcal{J}}=\{1,2\}. The jump rate function qα:𝒥×(0,)×𝒟×𝒟+q_{\alpha}\ :{\mathcal{J}}\times(0,\infty)\times{\mathcal{D}}_{\star}\times{\mathcal{D}}_{\star}^{\infty}\rightarrow{\mathbb{R}}_{+} is then given by

qα(1,t,x,y)=𝟏{x(t)=S}αβt(y(t))+𝟏{x(t)=E}λt+𝟏{x(t)=I}ρt,qα(2,t,x,y)=𝟏{x(t)=S}(1α)βt(y(t)).\displaystyle\begin{split}&q_{\alpha}(1,t,x,y)={\bm{1}}_{\{x(t-)=S\}}\alpha{\beta_{t}}{\mathcal{I}}(y(t-))+{\bm{1}}_{\{x(t-)=E\}}\lambda_{t}+{\bm{1}}_{\{x(t-)=I\}}{\rho_{t}},\\ &q_{\alpha}(2,t,x,y)={\bm{1}}_{\{x(t-)=S\}}(1-\alpha){\beta_{t}}{\mathcal{I}}(y(t-)).\end{split} (4.3)
Remark 4.1.

When α=0\alpha=0, qαq_{\alpha} defined in (4.3) reduces to the jump rate function of the SIR process as described in Section 2.1, and when α=1\alpha=1, qαq_{\alpha} coincides with the jump rate function of the SEIR process as described in Section 2.3.

Given j𝒥j\in{\mathcal{J}} and v𝕍v\in\mathbb{V}, we define j(v){0,1,2}𝕍j^{(v)}\in\{0,1,2\}^{\mathbb{V}} by jw(v)=j𝟏{v=w}j^{(v)}_{w}=j{\bm{1}}_{\{v=w\}} for all w𝕍.w\in\mathbb{V}. We define ξ𝒯,α\xi^{{\mathcal{T}},\alpha} as a continuous time Markov chain on 𝒳¯𝕍\bar{\mathcal{X}}_{\star}^{\mathbb{V}} with jump directions {j(v)}j𝒥,v𝕍\{j^{(v)}\}_{j\in{\mathcal{J}},v\in\mathbb{V}} and corresponding jump rates at time tt given by qα(j,t,ξv𝒯,α,ξv𝕍𝒯,α)q_{\alpha}(j,t,\xi_{v}^{{\mathcal{T}},\alpha},\xi^{{\mathcal{T}},\alpha}_{\partial^{\mathbb{V}}_{v}}). The initial state of the process is given by ξv𝒯,α(0)=zvp0\xi^{{\mathcal{T}},\alpha}_{v}(0)=z^{p_{0}}_{v}, where zp0={zvp0}v𝕍z^{p_{0}}=\{z^{p_{0}}_{v}\}_{v\in\mathbb{V}} satisfies the following assumption.

Assumption D.

Let z(1)={zv(1)}v𝕍z^{(1)}=\{z^{(1)}_{v}\}_{v\in\mathbb{V}} be a collection of i.i.d. 𝒳¯\bar{\mathcal{X}}-valued random variables with common law p0p_{0} given by (4.1). Let z(2)z^{(2)} be a {1,}𝕍\{1,\star\}^{\mathbb{V}}-valued random variables independent of z(1)z^{(1)} such that the subgraph of 𝕍\mathbb{V} induced by {v:zv(2)}\{v\ :z^{(2)}_{v}\neq\star\} is equal in law to a UGW(θ)(\theta) tree. The 𝒳¯𝕍\bar{\mathcal{X}}^{\mathbb{V}}-valued random variable zp0z^{p_{0}} satisfies

zvp0=zv(1)𝟏{zv(2)}+𝟏{zv(2)=},v𝕍.z^{p_{0}}_{v}=z^{(1)}_{v}{\bm{1}}_{\{z_{v}^{(2)}\neq\star\}}+\star{\bm{1}}_{\left\{z_{v}^{(2)}=\star\right\}},\qquad v\in\mathbb{V}.

We can easily recover the graph 𝒯{\mathcal{T}} from the process ξ𝒯,α\xi^{{\mathcal{T}},\alpha} as follows:

𝒯=𝒯(ξ𝒯,α):={vV:ξv𝒯,α(0)}={v:zvp0}.{\mathcal{T}}={\mathcal{T}}(\xi^{{\mathcal{T}},\alpha}):=\{v\in V:\xi^{{\mathcal{T}},\alpha}_{v}(0)\neq\star\}=\{v\ :z^{p_{0}}_{v}\neq\star\}.

Since the graph 𝒯{\mathcal{T}} can be infinite, it is no longer immediate that the process ξ𝒯,α\xi^{{\mathcal{T}},\alpha} with the intuitive description above is well defined (see [Ganguly2022hydrodynamic, Appendix A]). However, since 𝒯{\mathcal{T}} is a UGW(θ)(\theta) with θ\theta having a finite second moment (see Assumption B), this is guaranteed by the following result proved in [Ganguly2022hydrodynamic], which also characterizes ξ𝒯,α\xi^{{\mathcal{T}},\alpha} as the unique in law solution of a certain jump SDE.

Lemma 4.2.

The S(E)IR process ξ𝒯,α\xi^{{\mathcal{T}},\alpha} exist and is unique in law. Furthermore, its law is the unique solution to the SDE ,

ξv𝒯,α(t)=zvp0+j=1,2(0,t]×[0,)j𝟏{r<qα(j,s,ξv𝒯,α,ξv𝕍𝒯,α)}Nv(ds,dr),v𝕍,t[0,),\xi^{{\mathcal{T}},\alpha}_{v}(t)=\ z^{p_{0}}_{v}+\sum_{j=1,2}\int_{(0,t]\times[0,\infty)}j{\bm{1}}_{\left\{r<q_{\alpha}\left(j,s,\xi^{{\mathcal{T}},\alpha}_{v},\xi^{{\mathcal{T}},\alpha}_{\partial^{\mathbb{V}}_{v}}\right)\right\}}\textbf{N}_{v}(ds,\ dr),\quad v\in\mathbb{V},\ t\in[0,\infty), (4.4)

where zp0z^{p_{0}} is a 𝒳¯𝕍\bar{\mathcal{X}}_{\star}^{\mathbb{V}}-valued random element satisfying Assumption D and N={Nv:v𝕍}\textbf{N}=\left\{\textbf{N}_{v}\ :v\in\mathbb{V}\right\} are i.i.d. Poisson point processes on (0,)×[0,)(0,\infty)\times[0,\infty) with intensity measure equal to Lebesgue measure, independent of zp0z^{p_{0}}.

Proof.

Existence and uniqueness in law of the solution to the SDE (4.4) follows from [Ganguly2022hydrodynamic, Theorem 4.2] on observing that Assumption C implies [Ganguly2022hydrodynamic, Assumption 1], and that by [Ganguly2022hydrodynamic, Proposition 5.1], the UGW(θ\theta) tree 𝒯{\mathcal{T}} is finitely dissociable in the sense of [Ganguly2022hydrodynamic, Definition 5.1]. ∎

We now define

X𝒯:=ξ𝒯,0,X¯𝒯:=ξ𝒯,1.\displaystyle\begin{split}X^{{\mathcal{T}}}:=\xi^{{\mathcal{T}},0},\\ {\bar{X}}^{{\mathcal{T}}}:=\xi^{{\mathcal{T}},1}.\end{split} (4.5)
Remark 4.3.

By Remark 4.1 and the uniqueness in law established in Lemma 4.2, X𝒯𝒯X^{\mathcal{T}}_{\mathcal{T}} and X¯𝒯𝒯{\bar{X}}^{\mathcal{T}}_{\mathcal{T}} are the SIR and SEIR processes on the possibly infinite random tree 𝒯{\mathcal{T}}, in a sense consistent with the definitions of the SIR and SEIR processes on finite graphs given in Section 2.1 and Section 2.3, respectively.

4.2. Proofs of Transient Results

The proof of Theorem 2.5 is presented in Section 4.2.2. The proof of Theorem 2.10 uses similar techniques, and is thus only outlined in Section 4.2.2, with details relegated to Appendix A. Both proofs rely on four preliminary results first presented in Section 4.2.1. The first ingredient (Theorem 4.5) is a convergence result from [Ganguly2022hydrodynamic], which shows that the limits of the fractions |{vGn:ξv𝒯,α(t)=b}|/|Gn||\{v\in G_{n}:\ \xi^{{\mathcal{T}},\alpha}_{v}(t)=b\}|/|G_{n}|, b𝒳¯b\in\bar{\mathcal{X}}, coincide with the root marginal probabilities of the limiting S(E)IR dynamics on the graph 𝒯{\mathcal{T}} that arises as the local limit of the graph sequence {Gn}n\{G_{n}\}_{n\in{\mathbb{N}}}. The second ingredient is a projection result (Proposition 4.6) that identifies the law of the marginal dynamics on 𝒯{\mathcal{T}} in terms of a certain (a priori non-Markovian) jump process with somewhat implicit jump rates. This result is a generalization of similar projection results obtained in [Ganguly2023marginal, Ganguly2023characterization]. The third and fourth results (Proposition 4.7 and Proposition 4.8) identify key conditional independence and symmetry properties of the dynamics to explicitly identify the jump rates of the marginal dynamics.

Remark 4.4.

For a general class of interacting particle systems (IPS) on UGW trees 𝒯{\mathcal{T}} whose offspring satisfies suitable moment conditions, which in particular includes the SIR and S(E)IR processes, we expect that the marginal dynamics of the IPS on the root and its neighborhood can be described autonomously in terms of a certain (non-Markovian) stochastic process. Indeed, in the special case when 𝒯{\mathcal{T}} is a κ\kappa-regular tree, such a result is established in [Ganguly2023marginal] (see also [Ganguly2022nonmarkovian]) by appealing to a Markov random field property for the trajectories of the process proved in [Ganguly2022interacting, Theorem 3.7] (see also [lacker2021locally, Lacker2021marginal] for corresponding results for interacting diffusions). The current work goes beyond regular trees to include a large class of UGW trees, and also establishes a much stronger conditional independence property of the trajectories for the S(E)IR process when compared to general IPS. The latter is then used to show that for the S(E)IR process, the root marginal dynamics is in fact a Markovian process (see Remark 4.11), and thus its law can be described by a system of ODEs (namely, the forward Kolmogorov equations describing the evolution of the law of the Markov process).

We remind the reader that the standing assumptions made at the beginning of Section 4 are in effect throughout.

4.2.1. Preliminary Results

We start by stating the convergence result.

Theorem 4.5.

For every nn\in{\mathbb{N}} and b𝒳¯b\in\bar{\mathcal{X}}, set

pbn,α(t):=1|Gn|vGn𝟏{ξvGn,α(t)=b}.p_{b}^{n,\alpha}(t):=\frac{1}{|G_{n}|}\sum_{v\in G_{n}}{\bm{1}}_{\{\xi^{G_{n},\alpha}_{v}(t)=b\}}. (4.6)

For every t[0,)t\in[0,\infty), as nn\rightarrow\infty,

(pbn,α(t))b𝒳¯((ξ𝒯,α(t)=b))b𝒳¯,(p_{b}^{n,\alpha}(t))_{b\in\bar{\mathcal{X}}}\rightarrow({\mathbb{P}}(\xi^{{\mathcal{T}},\alpha}_{\varnothing}(t)=b))_{b\in\bar{\mathcal{X}}}, (4.7)

where the convergence is in probability.

Proof.

The statement follows from [Ganguly2022hydrodynamic, Corollary 4.7] on observing that Assumption C implies [Ganguly2022hydrodynamic, Assumption 1], Assumption B, along with [Ganguly2022hydrodynamic, Corollary 5.16] implies that the graph sequence {Gn}n\{G_{n}\}_{n\in{\mathbb{N}}} and 𝒯{\mathcal{T}} are a.s. finitely dissociable in the sense of [Ganguly2022hydrodynamic, Definition 5.11], and that [Ganguly2022hydrodynamic, Assumption 2] holds trivially since the state 𝒳¯\bar{\mathcal{X}}_{\star} is discrete. ∎

In view of Theorem 4.5, our next goal is to characterize the law of the root marginal of ξ𝒯,α\xi^{{\mathcal{T}},\alpha}. We first apply a projection result that characterizes the law of any marginal ξU𝒯,α\xi_{U}^{{\mathcal{T}},\alpha} for U𝕍U\subset\mathbb{V} in terms of a certain jump process η𝒯,α[U]\eta^{{\mathcal{T}},\alpha}[U].

Proposition 4.6.

For every finite U𝕍U\subset\mathbb{V}, vUv\in U, and j{1,2}j\in\{1,2\}, there exists a Borel measurable function q^v,jθ,α[U]:[0,)×𝒟([0,),𝒳¯U)[0,)\hat{q}^{\theta,\alpha}_{v,j}[U]:[0,\infty)\times{\mathcal{D}}([0,\infty),\bar{\mathcal{X}}_{\star}^{U})\rightarrow[0,\infty) such that

  1. (1)

    the function tq^v,jθ,α[U](t,x)t\rightarrow\hat{q}^{\theta,\alpha}_{v,j}[U](t,x) is càglàd333left continuous with finite right limits at every t[0,)t\in[0,\infty). for all x𝒟([0,),𝒳¯U)x\in{\mathcal{D}}([0,\infty),\bar{\mathcal{X}}_{\star}^{U});

  2. (2)

    the function tq^v,jθ,α[U](t,x)t\rightarrow\hat{q}^{\theta,\alpha}_{v,j}[U](t,x) is predictable in the sense that for any t[0,)t\in[0,\infty) and x,x𝒟([0,),𝒳¯U)x,x^{\prime}\in{\mathcal{D}}([0,\infty),\bar{\mathcal{X}}_{\star}^{U}), q^v,jθ,α[U](t,x)=q^v,jθ,α[U](t,x)\hat{q}^{\theta,\alpha}_{v,j}[U](t,x)=\hat{q}^{\theta,\alpha}_{v,j}[U](t,x^{\prime}) whenever x[t)=x[t)x[t)=x^{\prime}[t);

  3. (3)

    for every vUv\in U, the stochastic process tq^v,jθ,α[U](t,ξU𝒯,α)t\rightarrow\hat{q}^{\theta,\alpha}_{v,j}[U](t,\xi^{{\mathcal{T}},\alpha}_{U}) is a modification444Given two stochastic processes Y=(Yt,t0)Y=(Y_{t},t\geq 0) and Y^=(Y^t,t0){\hat{Y}}=({\hat{Y}}_{t},t\geq 0) defined on the same probability space (Ω,,)(\Omega,{\mathcal{F}},{\mathbb{P}}), YY is a modification of Y^{\hat{Y}} if for every t0t\geq 0, (Y(t)=Y^(t))=1{\mathbb{P}}(Y(t)={\hat{Y}}(t))=1. of the process {𝔼[qα(j,t,ξv𝒯,α,ξv𝕍𝒯,α)|ξU𝒯,α[t)],t[0,)}\{{\mathbb{E}}[q_{\alpha}(j,t,\xi^{{\mathcal{T}},\alpha}_{v},\xi^{{\mathcal{T}},\alpha}_{\partial^{\mathbb{V}}_{v}})\ |\ \xi^{{\mathcal{T}},\alpha}_{U}[t)],\ t\in[0,\infty)\}.

Furthermore, (ξU𝒯,α)=(η𝒯,α[U]){\mathcal{L}}(\xi^{{\mathcal{T}},\alpha}_{U})={\mathcal{L}}(\eta^{{\mathcal{T}},\alpha}[U]), where η𝒯,α[U])\eta^{{\mathcal{T}},\alpha}[U]) is the pathwise unique solution to the following jump SDE

ηv𝒯,α[U](t)=zvp0+j=1,2(0,t]×[0,)j𝟏{r<q^v,jθ,α[U](s,η𝒯,α[U])}Nv(ds,dr),vU,t[0,),\eta_{v}^{{\mathcal{T}},\alpha}[U](t)=z^{p_{0}}_{v}+\sum_{j=1,2}\int_{(0,t]\times[0,\infty)}j{\bm{1}}_{\left\{r<\hat{q}_{v,j}^{\theta,\alpha}[U]\left(s,\ \eta^{{\mathcal{T}},\alpha}[U]\right)\right\}}\textbf{N}_{v}(ds,\ dr),\qquad v\in U,\ t\in[0,\infty), (4.8)

where zp0z^{p_{0}} is a 𝒳¯𝕍\bar{\mathcal{X}}_{\star}^{\mathbb{V}}-valued random variable satisfying Assumption D, and N={Nv:vU}\textbf{N}=\left\{\textbf{N}_{v}\ :v\in U\right\} are i.i.d. Poisson point processes on (0,)×[0,)(0,\infty)\times[0,\infty) with intensity measure equal to Lebesgue measure, independent of zp0z^{p_{0}}.

Proof.

In the case when 𝒯{\mathcal{T}} is a deterministic κ\kappa-regular tree, this was proved in Lemma 9.1 and Proposition 9.2 of [Ganguly2022nonmarkovian]; see also [Ganguly2023marginal]. Using a general result from [Ganguly2022interacting, Corollary 4.11], this can be extended to a class of Galton-Watson trees that include the ones considered in Assumption B; we refer the reader to [Ganguly2023characterization] for full details. ∎

Using Proposition 4.6, the law of ξ{,1}𝒯,α\xi^{{\mathcal{T}},\alpha}_{\{\varnothing,1\}} can be characterized in terms of a jump process η𝒯,α[{,1}]\eta^{{\mathcal{T}},\alpha}[{\left\{\varnothing,1\right\}}]. However, the jump rates of the latter process are a priori path-dependent and not very explicit. We now identify two additional properties that allow us to simplify the form of these jump rates and thereby show that η𝒯,α[{,1}]\eta^{{\mathcal{T}},\alpha}[{\left\{\varnothing,1\right\}}] is in fact a nonlinear Markov process (see Remark 4.11), that is, a (time-inhomogeneous) Markov process whose transition rates depend not only on the current state but also on the law of the current state.

For a set B𝕍B\subset\mathbb{V}, we let SBS_{B} denote the vector in 𝒳¯B\bar{\mathcal{X}}_{\star}^{B} whose every coordinate is equal to SS.

Proposition 4.7.

For every t[0,)t\in[0,\infty), A𝕍A\subset\mathbb{V} with |𝕍A|<|\partial^{\mathbb{V}}A|<\infty, and B𝕍AB\subset\mathbb{V}\setminus A,

ξA𝒯,α[t)ξB𝒯,α[t)|ξ𝕍A𝒯,α(t)=S𝕍A.\xi_{A}^{{\mathcal{T}},\alpha}[t)\perp\xi_{B}^{{\mathcal{T}},\alpha}[t)|\xi^{{\mathcal{T}},\alpha}_{\partial^{\mathbb{V}}A}(t-)=S_{\partial^{\mathbb{V}}A}. (4.9)

Moreover, for every v𝕍v\in\mathbb{V}, the processes ξvi𝒯,α[t)\xi^{{\mathcal{T}},\alpha}_{vi}[t), 1idv𝟏{v}1\leq i\leq d_{v}-{\bm{1}}_{\{v\neq\varnothing\}} are conditionally independent given ξv𝒯,α(t)=S\xi^{{\mathcal{T}},\alpha}_{v}(t-)=S and the degree dvd_{v} of vv.

The proof of Proposition 4.7 is given in Section 5.

Proposition 4.8.

For every b𝒳¯b\in\bar{\mathcal{X}}, α[0,1]\alpha\in[0,1] and t[0,)t\in[0,\infty), the conditional probability (ξv~m𝒯,α(t)=b|ξv~𝒯,α(t)=S,v~𝒯){\mathbb{P}}(\xi^{{\mathcal{T}},\alpha}_{{\tilde{v}}m}(t-)=b\ |\ \xi^{{\mathcal{T}},\alpha}_{{\tilde{v}}}(t-)=S,\ {\tilde{v}}\in{\mathcal{T}}) does not depend on the choice of v~𝕍{\tilde{v}}\in\mathbb{V} and mm\in{\mathbb{N}}.

The proof of Proposition 4.8 proceeds by exploiting the conditional independence property in Proposition 4.7 along with symmetry properties and well-posedness of the SDE (4.4) to show that (ξv~m𝒯,α[t)|ξv~𝒯,α(t)=S)=(ξ1𝒯,α[t)|ξ𝒯,α(t)=S){\mathcal{L}}(\xi^{{\mathcal{T}},\alpha}_{{\tilde{v}}m}[t)|\xi^{{\mathcal{T}},\alpha}_{{\tilde{v}}}(t-)=S)={\mathcal{L}}(\xi^{{\mathcal{T}},\alpha}_{1}[t)|\xi^{{\mathcal{T}},\alpha}_{\varnothing}(t-)=S) for all v~𝕍{\tilde{v}}\in\mathbb{V} and mm\in{\mathbb{N}}. The details are relegated to the end of Section 5.

We conclude this section with an elementary result we use repeatedly in the sequel.

Lemma 4.9.

Let (Ω,,)(\Omega,\ {\mathcal{F}},\ {\mathbb{P}}) be a probability space, and suppose that A,A, A,A^{\prime}, B,B, BB^{\prime}\in{\mathcal{F}} with (BB)>0{\mathbb{P}}(B\cap B^{\prime})>0 and (AB)>0{\mathbb{P}}(A^{\prime}\cap B^{\prime})>0. Then,

(AA|BB)=(A|B)(AB|AB)(B|B).\displaystyle\begin{split}{\mathbb{P}}(A\cap A^{\prime}\ |\ B\cap B^{\prime})={\mathbb{P}}(A^{\prime}\ |\ B^{\prime})\frac{{\mathbb{P}}(A\cap B\ |\ A^{\prime}\cap B^{\prime})}{{\mathbb{P}}(B\ |\ B^{\prime})}.\end{split} (4.10)
Proof.

Let A,A, A,A^{\prime}, B,B, BB^{\prime} be as in the statement of the lemma. By the definition of conditional probability, and some simple arithmetic manipulation,

(AA|BB)=(AABB)(BB)=(AB|AB)(AB)(B|B)(B)=(A|B)(AB|AB)(B|B).\displaystyle\begin{split}{\mathbb{P}}(A\cap A^{\prime}\ |\ B\cap B^{\prime})&=\frac{{\mathbb{P}}(A\cap A^{\prime}\cap B\cap B^{\prime})}{{\mathbb{P}}(B\cap B^{\prime})}\\ &=\frac{{\mathbb{P}}(A\cap B\ |\ A^{\prime}\cap B^{\prime}){\mathbb{P}}(A^{\prime}\cap B^{\prime})}{{\mathbb{P}}(B\ |\ B^{\prime}){\mathbb{P}}(B^{\prime})}\\ &={\mathbb{P}}(A^{\prime}\ |\ B^{\prime})\frac{{\mathbb{P}}(A\cap B\ |\ A^{\prime}\cap B^{\prime})}{{\mathbb{P}}(B\ |\ B^{\prime})}.\end{split} (4.11)

4.2.2. Proof of Theorem 2.5

We can now complete the proof of our main result for the SIR process by characterizing the time marginals of ξ𝒯,α\xi_{\varnothing}^{{\mathcal{T}},\alpha} for the special case α=0\alpha=0, which by Remark 4.3 is equal to the marginal at the root of the SIR process on the possibly infinite tree 𝒯{\mathcal{T}}. For a,b𝒳a,b\in\mathcal{X}, t[0,)t\in[0,\infty) and k{m0:θ(m)>0}k\in\{m\in{\mathbb{N}}_{0}\ :\ \theta(m)>0\}, define

Pa,b;kθ(t)\displaystyle P^{\theta}_{a,b;k}(t) :=(X𝒯(t)=b|X𝒯(0)=a,d=k),\displaystyle:={\mathbb{P}}(X^{\mathcal{T}}_{\varnothing}(t)=b\ |\ X^{\mathcal{T}}_{\varnothing}(0)=a,\ d_{\varnothing}=k), (4.12)
Pa,bθ(t)\displaystyle P^{\theta}_{a,b}(t) :=(X𝒯(t)=b|X𝒯(0)=a),\displaystyle:={\mathbb{P}}(X^{\mathcal{T}}_{\varnothing}(t)=b\ |\ X^{\mathcal{T}}_{\varnothing}(0)=a),
faθ(t)\displaystyle f^{\theta}_{a}(t) :=(X1𝒯(t)=a|X𝒯(t)=S, 1𝒯),\displaystyle:={\mathbb{P}}(X^{\mathcal{T}}_{1}(t)=a\ |\ X^{\mathcal{T}}_{\varnothing}(t)=S,\ 1\in{\mathcal{T}}),

where dvd_{v} is the degree of the vertex v𝒯v\in{\mathcal{T}}, and where we recall that 1𝒯1\in{\mathcal{T}} is equivalent to X1𝒯(0)X_{1}^{\mathcal{T}}(0)\neq\star. When clear from context, we omit the dependence on θ\theta and simply write Pa,b;k,Pa,bP_{a,b;k},\ P_{a,b} and faf_{a}.

Theorem 4.10.

Let fSf_{S} and fIf_{I} be as in (4.12), and set FI(t):=0tβsfI(s)𝑑sF_{I}(t):=\int_{0}^{t}\beta_{s}f_{I}(s)ds for t[0,)t\in[0,\infty). Then, (fS(f_{S}, fIf_{I}, FI)F_{I}) solves the ODE system (2.4)-(2.5).

Proof.

Throughout the proof, in order to simplify the notation we write XX in lieu of X𝒯=ξ𝒯,0X^{{\mathcal{T}}}=\xi^{{\mathcal{T}},0}, the SIR process on 𝒯{\mathcal{T}}, and qq in lieu of q0q_{0} for the jump rates defined in (4.3). We start by observing that, by Assumption D, fS(0)=s0f_{S}(0)=s_{0} and fI(0)=i0=1s0f_{I}(0)=i_{0}=1-s_{0}. Since, clearly FI(0)=0F_{I}(0)=0, the initial condition (2.5) are established. By the fundamental theorem of calculus, F˙I(t)=βtfI(t)\dot{F}_{I}(t)=\beta_{t}f_{I}(t), which is the third equation in (2.4).

We now turn to the derivation of the evolution of fIf_{I} and fSf_{S}. This requires us to simultaneously track the evolution of two nodes, \varnothing and 11, since fI(t)f_{I}(t) and fS(t)f_{S}(t) are conditional probabilities associate with the joint law of X(t)X_{\varnothing}(t) and X1(t)X_{1}(t). To start with, we apply the projection result of Proposition 4.6, with α=0\alpha=0 and U={,1}U=\{\varnothing,1\}, to conclude that the joint marginal X,1X_{\varnothing,1} has the same law as the jump process η𝒯,α[{,1}]\eta^{{\mathcal{T}},\alpha}[\{\varnothing,1\}] on 𝒳×𝒳\mathcal{X}_{\star}\times\mathcal{X}_{\star} that has predictable jump rates

q^v(t,x):=q^v,j(xv)θ,0[{,1}](t,x),\hat{q}_{v}(t,x):=\hat{q}_{v,j(x_{v})}^{\theta,0}[\{\varnothing,1\}](t,x), (4.13)

v{,1}v\in\{\varnothing,1\}, x𝒟([0,),𝒳2)x\in{\mathcal{D}}([0,\infty),\mathcal{X}_{\star}^{2}) and j(xv)=1+𝟏{xv(t)=S}j(x_{v})=1+{\bm{1}}_{\{x_{v}(t-)=S\}}, which satisfy, for every t0,t\geq 0, almost surely555The dependence j(xv)j(x_{v}) of the allowed jump on the state is a notational nuisance that is a mere artifact of our using a common framework to analyze both the SIR and SEIR processes. Indeed, this is because when we use the common (ordered) state space {S,E,I,R}\{S,E,I,R\} for both processes, then the SIR process allows only jumps of size 22 from the state S (going from S to I and skipping over E), and only jumps of size 11 from the state I (going from I to R).

q^v(t,X,1)=𝔼[q(j(Xv),t,Xv,Xv𝕍)|X,1[t)],v{,1}.\hat{q}_{v}(t,X_{\varnothing,1})={\mathbb{E}}[q(j(X_{v}),t,X_{v},X_{\partial_{v}^{\mathbb{V}}})|X_{\varnothing,1}[t)],\quad v\in\{\varnothing,1\}. (4.14)

Next, we use the specific form of qq, as defined in (4.3) and Propositions 4.7 and 4.8 to obtain a more explicit description of q^v\hat{q}_{v}, v{,1}v\in\{\varnothing,1\}. Since the probabilities fI(t)f_{I}(t) and fS(t)f_{S}(t) are conditioned on X(t)=SX_{\varnothing}(t)=S and on X1(t)X_{1}(t)\neq\star (and using the fact that a particle that is in state RR remains in that state for all subsequent times), we only need to consider the jump intensities q^v(t,X,1)\hat{q}_{v}(t,X_{\varnothing,1}), v{,1}v\in\{\varnothing,1\}, on the events {X,1(t)=(S,S)}\{X_{\varnothing,1}(t-)=(S,S)\} and {X,1(t)=(S,I)}\{X_{\varnothing,1}(t-)=(S,I)\}.

Define B1(t):=βt𝔼[(X1𝕍{}(t))|X1(t)=S]B_{1}(t):=\beta_{t}{\mathbb{E}}[{\mathcal{I}}(X_{\partial^{\mathbb{V}}_{1}\setminus\{\varnothing\}}(t-))|X_{1}(t-)=S]. Recalling the definition of q=q0q=q_{0} from (4.3), B1(t)B_{1}(t) is the cumulative conditional rate at which the children of 11 infect 11 at time tt, given X1(t)=SX_{1}(t-)=S (which also implies 1𝒯1\in{\mathcal{T}}). Similarly, let B(t):=βt𝔼[(X𝕍{1}(t))|X(t)=S]B_{\varnothing}(t):=\beta_{t}{\mathbb{E}}[{\mathcal{I}}(X_{\partial^{\mathbb{V}}_{\varnothing}\setminus\{1\}}(t-))|X_{\varnothing}(t-)=S] be the cumulative conditional rate at which the neighbors of the root other than vertex 11 infect the root at time tt, given X(t)=SX_{\varnothing}(t-)=S. By Proposition 4.7, for v,w{,1}v,w\in\{\varnothing,1\} with wvw\neq v,

Bv(t)=βt𝔼[(Xv𝕍{w}(t))|Xv(t)=S,Xw(t)].B_{v}(t)=\beta_{t}{\mathbb{E}}[{\mathcal{I}}(X_{\partial^{\mathbb{V}}_{v}\setminus\{w\}}(t-))|X_{v}(t-)=S,\ X_{w}(t-)]. (4.15)

Using (4.3) and (4.15), on the event {X(t)=S}\{X_{\varnothing}(t-)=S\},

q^(t,X,1)=βt𝟏{X1(t)=I}+B(t).\hat{q}_{\varnothing}(t,X_{\varnothing,1})=\beta_{t}{\bm{1}}_{\{X_{1}(t-)=I\}}+B_{\varnothing}(t).

Similarly, on the event {X(t)=S}\{X_{\varnothing}(t-)=S\},

q^1(t,X,1)=B1(t)𝟏{X1(t)=S}+ρt𝟏{X1(t)=I}.\hat{q}_{1}(t,X_{\varnothing,1})=B_{1}(t){\bm{1}}_{\{X_{1}(t-)=S\}}+\rho_{t}{\bm{1}}_{\{X_{1}(t-)=I\}}.

Therefore, we can treat X,1X_{\varnothing,1} as a two particle jump processes driven by Poisson noises with intensity measure equal to Lebesgue measure, whose jumps and jump rates from the states (S,S)(S,S) and (S,I)(S,I) can be summarized as follows:

Jump: Rate at time tt:
(S,S)\displaystyle(S,S) (S,I)\displaystyle\rightarrow(S,I) B1(t)\displaystyle B_{1}(t)
(S,S)\displaystyle(S,S) (I,S)\displaystyle\rightarrow(I,S) B(t)\displaystyle B_{\varnothing}(t)
(S,I)\displaystyle(S,I) (I,I)\displaystyle\rightarrow(I,I) βt+B(t)\displaystyle{\beta_{t}}+B_{\varnothing}(t)
(S,I)\displaystyle(S,I) (S,R)\displaystyle\rightarrow(S,R) ρt,\displaystyle{\rho_{t}},

with all other jump rates being equal to zero. Next, we fix h>0h>0 and we obtain expressions for fI(t+h),fS(t+h)f_{I}(t+h),\ f_{S}(t+h) in terms of fI(t),f_{I}(t), fS(t),f_{S}(t), h,h, βt,{\beta_{t}}, ρt{\rho_{t}}, and θ^{\hat{\theta}}. We first consider fS(t+h)=(X1(t+h)=S|X(t+h)=S, 1𝒯)f_{S}(t+h)={\mathbb{P}}(X_{1}(t+h)=S\ |\ X_{\varnothing}(t+h)=S,\ 1\in{\mathcal{T}}) defined in (4.12). Using the monotonicity of the SIR dynamics, we can write

fS(t+h)=(X1(t+h)=S,X1(t)=S|X(t+h)=S,X(t)=S, 1𝒯).f_{S}(t+h)={\mathbb{P}}(X_{1}(t+h)=S,\ X_{1}(t)=S\ |\ X_{\varnothing}(t+h)=S,\ X_{\varnothing}(t)=S,\ 1\in{\mathcal{T}}). (4.16)

By an application of Lemma 4.9, with A={X1(t+h)=S}A=\{X_{1}(t+h)=S\}, A={X1(t)=S}A^{\prime}=\{X_{1}(t)=S\}, B={X(t+h)=S}B=\{X_{\varnothing}(t+h)=S\} and B={X(t)=S,1𝒯}B^{\prime}=\{X_{\varnothing}(t)=S,1\in{\mathcal{T}}\}, we obtain

fS(t+h)=fS(t)(X(t+h)=S,X1(t+h)=S,|X(t)=S,X1(t)=S, 1𝒯)(X(t+h)=S|X(t)=S, 1𝒯).f_{S}(t+h)=f_{S}(t)\frac{{\mathbb{P}}(X_{\varnothing}(t+h)=S,\ X_{1}(t+h)=S,|\ X_{\varnothing}(t)=S,\ X_{1}(t)=S,\ 1\in{\mathcal{T}})}{{\mathbb{P}}(X_{\varnothing}(t+h)=S\ |\ X_{\varnothing}(t)=S,\ 1\in{\mathcal{T}})}. (4.17)

Since B1(t)+B(t)B_{1}(t)+B_{\varnothing}(t) is the rate at which X,1(t)X_{\varnothing,1}(t) leaves the state (S,S)(S,S), the numerator in the right-hand side of (4.17) is equal to 1h(B1(t)B(t))+o(h)1-h(B_{1}(t)-B_{\varnothing}(t))+o(h). For the denominator, observe that the rate q^(t,X,1)\hat{q}_{\varnothing}(t,X_{\varnothing,1}) on the event {X(t)=S,X1(t)}={X(t)=S, 1𝒯}\{X_{\varnothing}(t-)=S,\ X_{1}(t-)\neq\star\}=\{X_{\varnothing}(t-)=S,\ 1\in{\mathcal{T}}\} is equal to

𝔼[q(1,t,X,X𝕍)|X(t)=S,1𝒯]=βt𝔼[(X1(t))|X(t)=S, 1𝒯]+βt𝔼[(X𝕍{1}(t))|X(t)=S, 1𝒯]=βtfI(t)+B(t),\displaystyle\begin{split}&{\mathbb{E}}[q(1,t,X_{\varnothing},X_{\partial^{\mathbb{V}}_{\varnothing}})\ |X_{\varnothing}(t-)=S,1\in{\mathcal{T}}]\\ =&\beta_{t}{\mathbb{E}}[{\mathcal{I}}(X_{1}(t-))\ |X_{\varnothing}(t-)=S,\ 1\in{\mathcal{T}}]+\beta_{t}{\mathbb{E}}[{\mathcal{I}}(X_{\partial^{\mathbb{V}}_{\varnothing}\setminus\{1\}}(t-))\ |X_{\varnothing}(t-)=S,\ 1\in{\mathcal{T}}]\\ =&\beta_{t}f_{I}(t)+B_{\varnothing}(t),\end{split}

where the first equality follows from (4.3) with α=0\alpha=0, and the second follows from the definition of fIf_{I} in (4.12) and by (4.15) (on observing that the event {1𝒯}\{1\in{\mathcal{T}}\} is X1(t)X_{1}(t)-measurable). Therefore, it follows that

fS(t+h)=fS(t)1h(B1(t)+B(t))+o(h)1h(βtfI(t)+B(t))+o(h),f_{S}(t+h)=f_{S}(t)\frac{1-h(B_{1}(t)+B_{\varnothing}(t))+o(h)}{1-h({\beta_{t}}f_{I}(t)+B_{\varnothing}(t))+o(h)}, (4.18)

which implies that

fS(t+h)fS(t)=fS(t)1hB1(t)hB(t)1+hβtfI(t)+hB(t)+o(h)1h(βtfI(t)+B(t))+o(h)=fS(t)hβtfI(t)hB1(t)+o(h)1+o(1).\displaystyle\begin{split}f_{S}(t+h)-f_{S}(t)&=f_{S}(t)\frac{1-hB_{1}(t)-hB_{\varnothing}(t)-1+h{\beta_{t}}f_{I}(t)+hB_{\varnothing}(t)+o(h)}{1-h({\beta_{t}}f_{I}(t)+B_{\varnothing}(t))+o(h)}\\ &=f_{S}(t)\frac{h{\beta_{t}}f_{I}(t)-hB_{1}(t)+o(h)}{1+o(1)}.\end{split} (4.19)

In turn, this implies

f˙S(t)=βtfS(t)fI(t)fS(t)B1(t).\dot{f}_{S}(t)=\beta_{t}f_{S}(t)f_{I}(t)-f_{S}(t)B_{1}(t). (4.20)

Similarly, recalling that fI(t+h)=(X1(t+h)=I|X(t+h)=S, 1𝒯)f_{I}(t+h)={\mathbb{P}}(X_{1}(t+h)=I\ |\ X_{\varnothing}(t+h)=S,\ 1\in{\mathcal{T}}) from (4.12), using the fact that a particle that at time t+ht+h is in state II could only have been in states SS or II at time tt, and using a similar derivation as in (4.16)-(4.19),

fI(t+h)=a=S,Ifa(t)(X(t+h)=S,X1(t+h)=I|X(t)=S,X1(t)=a, 1𝒯)(X(t+h)=S|X(t)=S,1𝒯)=fS(t)(hB1(t)+o(h))+fI(t)(1h(ρt+B(t)+βt)+o(h))1h(fI(t)βt+B(t))+o(h),\displaystyle\begin{split}f_{I}(t+h)&=\sum_{a=S,I}f_{a}(t)\frac{{\mathbb{P}}(X_{\varnothing}(t+h)=S,\ X_{1}(t+h)=I|\ X_{\varnothing}(t)=S,\ X_{1}(t)=a,\ 1\in{\mathcal{T}})}{{\mathbb{P}}(X_{\varnothing}(t+h)=S|\ X_{\varnothing}(t)=S,1\in{\mathcal{T}})}\\ &=\frac{f_{S}(t)(hB_{1}(t)+o(h))+f_{I}(t)(1-h({\rho_{t}}+B_{\varnothing}(t)+{\beta_{t}})+o(h))}{1-h(f_{I}(t){\beta_{t}}+B_{\varnothing}(t))+o(h)},\end{split} (4.21)

which implies that

fI(t+h)fI(t)=(1+o(1))(hfS(t)B1(t)hfI(t)(ρt+βtβtfI(t))+o(h)).\displaystyle\begin{split}&f_{I}(t+h)-f_{I}(t)=(1+o(1))(hf_{S}(t)B_{1}(t)-hf_{I}(t)({\rho_{t}}+{\beta_{t}}-{\beta_{t}}f_{I}(t))+o(h)).\end{split} (4.22)

It follows that

f˙I=fSB1fI(ρ+ββfI).\dot{f}_{I}=f_{S}B_{1}-f_{I}({\rho}+{\beta}-{\beta}f_{I}). (4.23)

In view of (4.20) and (4.23) all that is left to find is an expression for B1(t)B_{1}(t), the conditional rate at which the children of vertex 11 infect vertex 11 at time tt, given X1(t)=SX_{1}(t)=S, in terms only of βt,ρt,θ^,{\beta_{t}},\ {\rho_{t}},\ {\hat{\theta}}, and fI(t)f_{I}(t). By Proposition 4.7, X1(t)X_{\partial_{1}\setminus{\varnothing}}(t) is conditionally independent of X(t)X_{\varnothing}(t) given X1(t)=SX_{1}(t)=S. Also by Proposition 4.7, {X1i(t)}i=1,,k\{X_{1i}(t)\}_{i=1,...,k}, are conditionally i.i.d. given X1(t)=SX_{1}(t)=S and d1=k+1d_{1}=k+1, and by Proposition 4.8,

(X1i(t)=I|X1(t)=S,d1=k+1)=(X1i(t)=I|X1(t)=S, 1i𝒯)=fI(t),{\mathbb{P}}(X_{1i}(t)=I\ |\ X_{1}(t)=S,\ d_{1}=k+1)={\mathbb{P}}(X_{1i}(t)=I\ |\ X_{1}(t)=S,\ 1i\in{\mathcal{T}})=f_{I}(t),

for i=1,,ki=1,...,k. This implies that

B1(t)=βt𝔼[(X1𝕍{}(t))|X1(t)=S]=βt𝔼[𝔼[(X1𝕍{}(t))|X1(t)=S,d1=k+1]|X1(t)=S]=βtfI(t)𝔼[𝔼[d11|X1(t)=S,d1=k+1]|X1(t)=S]=βtfI(t)(𝔼[d11|X1(t)=S]).\displaystyle\begin{split}B_{1}(t)&=\beta_{t}{\mathbb{E}}[{\mathcal{I}}(X_{\partial^{\mathbb{V}}_{1}\setminus\{\varnothing\}}(t-))\ |\ X_{1}(t-)=S]\\ &=\beta_{t}{\mathbb{E}}[{\mathbb{E}}[{\mathcal{I}}(X_{\partial^{\mathbb{V}}_{1}\setminus\{\varnothing\}}(t-))|X_{1}(t-)=S,\ d_{1}=k+1]\ |\ X_{1}(t-)=S]\\ &=\beta_{t}f_{I}(t){\mathbb{E}}[{\mathbb{E}}[d_{1}-1|X_{1}(t-)=S,\ d_{1}=k+1]\ |\ X_{1}(t-)=S]\\ &=\beta_{t}f_{I}(t)({\mathbb{E}}[d_{1}\ -1|\ X_{1}(t-)=S]).\end{split} (4.24)

Next, we find a more explicit description of the conditional expectation in the last line of (4.24). Let 𝒩¯={k0:θ^(k+1)>0}\bar{{\mathcal{N}}}=\{k\in{\mathbb{N}}_{0}\ :\ {\hat{\theta}}(k+1)>0\}. For k𝒩¯k\in\bar{{\mathcal{N}}}, define

rk:=(X1(t)=S| 1𝒯,d1=k+1).r_{k}:={\mathbb{P}}(X_{1}(t)=S\ |\ 1\in{\mathcal{T}},\ d_{1}=k+1). (4.25)

Then, observing that X1(t)=SX_{1}(t)=S implies that 1𝒯1\in{\mathcal{T}},

𝔼[d11|X1(t)=S]=k𝒩¯k(d1=k+1| 1𝒯)(X1(t)=S| 1𝒯)rk(t).{\mathbb{E}}[d_{1}-1\ |X_{1}(t-)=S]=\sum_{k\in\bar{{\mathcal{N}}}}k\frac{{\mathbb{P}}(d_{1}=k+1\ |\ 1\in{\mathcal{T}})}{{\mathbb{P}}(X_{1}(t-)=S\ |\ 1\in{\mathcal{T}})}r_{k}(t-). (4.26)

By (4.3), the conditional rate at which the individual at 11 is infected, given that X1(t)=SX_{1}(t-)=S and d1=k+1d_{1}=k+1, is

βt𝔼[(X1𝕍(t)))|X1(t)=S,d1=k+1]=βt𝔼[(X1𝕍{}(t)))|X1(t)=S,d1=k+1]+βt𝔼[(X(t)))|X1(t)=S]=βtkfI(t)+βt𝔼[(X(t)))|X1(t)=S]\displaystyle\begin{split}&\beta_{t}{\mathbb{E}}[{\mathcal{I}}(X_{\partial^{\mathbb{V}}_{1}}(t)-))\ |\ X_{1}(t-)=S,\ d_{1}=k+1]\\ &=\beta_{t}{\mathbb{E}}[{\mathcal{I}}(X_{\partial^{\mathbb{V}}_{1}\setminus{\{\varnothing\}}}(t-)))\ |\ X_{1}(t-)=S,\ d_{1}=k+1]+\beta_{t}{\mathbb{E}}[{\mathcal{I}}(X_{\varnothing}(t-)))\ |\ X_{1}(t-)=S]\\ &=\beta_{t}kf_{I}(t)+\beta_{t}{\mathbb{E}}[{\mathcal{I}}(X_{\varnothing}(t-)))\ |\ X_{1}(t-)=S]\end{split} (4.27)

where the second equality follows from Proposition 4.7 and Proposition 4.8, and the first equality follows from an application of Proposition 4.7 with A={}{v}{1},v𝕍A=\{\varnothing\}\cup\{\ell v\}_{\ell\in{\mathbb{N}}\setminus\{1\},v\in\mathbb{V}}, 𝕍A={1}\partial^{\mathbb{V}}A=\{1\} and B={1m}mB=\{1m\}_{m\in{\mathbb{N}}}. Setting ι(t):=βt𝔼[(X(t)))|X1(t)=S]\iota(t):=\beta_{t}{\mathbb{E}}[{\mathcal{I}}(X_{\varnothing}(t-)))\ |\ X_{1}(t-)=S], using the monotonicty (4.2) of the SIR process in the first equality and (4.27) in the second, we have

rk(t+h)=(X1(t+h)=S|X1(t)=Sd1=k+1)rk(t)=(1hkβtfI(t)hι(t))rk(t),\displaystyle\begin{split}r_{k}(t+h)=&{\mathbb{P}}(X_{1}(t+h)=S\ |X_{1}(t)=S\ d_{1}=k+1)r_{k}(t)\\ =&(1-hk\beta_{t}f_{I}(t)-h\iota(t))r_{k}(t),\end{split} (4.28)

and it follows that

r˙k(t)=(kβtfI(t)+ι(t))rk(t)\dot{r}_{k}(t)=-(k\beta_{t}f_{I}(t)+\iota(t))r_{k}(t)

and, since rk(0)=s0r_{k}(0)=s_{0} by Assumption D,

rk(t)=s0ek0tβsfI(s)𝑑s+0tι(s)𝑑s.r_{k}(t)=s_{0}e^{-k\int_{0}^{t}\beta_{s}f_{I}(s)ds+\int_{0}^{t}\iota(s)ds}. (4.29)

Next, observing that (d1=k+1| 1𝒯)=θ^(k){\mathbb{P}}(d_{1}=k+1\ |\ 1\in{\mathcal{T}})={\hat{\theta}}(k) since 𝒯{\mathcal{T}} is a UGW(θ)(\theta), and

(X1(t)=S| 1𝒯)=k𝒩¯(X1(t)=S|d1=k+1,1𝒯)(d1=k+1|1𝒯)=k𝒩¯rk(t)θ^(k),\displaystyle\begin{split}{\mathbb{P}}(X_{1}(t-)=S\ |\ 1\in{\mathcal{T}})&=\sum_{k\in\bar{{\mathcal{N}}}}{\mathbb{P}}(X_{1}(t-)=S\ |d_{1}=k+1,1\in{\mathcal{T}}){\mathbb{P}}(d_{1}=k+1\ |1\in{\mathcal{T}})\\ &=\sum_{k\in\bar{{\mathcal{N}}}}r_{k}(t){\hat{\theta}}(k),\end{split}

the expression in (4.26) can be rewritten as

𝔼[d11|X1(t)=S]=k𝒩¯k(d1=k+1| 1𝒯)(X1(t)=S| 1𝒯)rk(t).=k𝒩¯kθ^(k)𝒩¯θ^()rrk(t)=k𝒩¯kθ^(k)s0ek0tβsfI(s)𝑑s+0tι(s)𝑑s𝒩¯θ^()s0e0tβsfI(s)𝑑s+0tι(s)𝑑s=k𝒩¯kθ^(k)ek0tβsfI(s)𝑑s𝒩¯θ^()e0tβsfI(s)𝑑s,\displaystyle\begin{split}{\mathbb{E}}[d_{1}-1\ |X_{1}(t-)=S]=&\sum_{k\in\bar{{\mathcal{N}}}}k\frac{{\mathbb{P}}(d_{1}=k+1\ |\ 1\in{\mathcal{T}})}{{\mathbb{P}}(X_{1}(t)=S\ |\ 1\in{\mathcal{T}})}r_{k}(t).\\ =&\sum_{k\in\bar{{\mathcal{N}}}}k\frac{{\hat{\theta}}(k)}{\sum_{\ell\in\bar{{\mathcal{N}}}}{\hat{\theta}}(\ell)r_{\ell}}r_{k}(t)\\ =&\sum_{k\in\bar{{\mathcal{N}}}}k{\hat{\theta}}(k)\frac{s_{0}e^{-k\int_{0}^{t}\beta_{s}f_{I}(s)ds+\int_{0}^{t}\iota(s)ds}}{\sum_{\ell\in\bar{{\mathcal{N}}}}{\hat{\theta}}(\ell)s_{0}e^{-\ell\int_{0}^{t}\beta_{s}f_{I}(s)ds+\int_{0}^{t}\iota(s)ds}}\\ =&\frac{\sum_{k\in\bar{{\mathcal{N}}}}k{\hat{\theta}}(k)e^{-k\int_{0}^{t}\beta_{s}f_{I}(s)ds}}{\sum_{\ell\in\bar{{\mathcal{N}}}}{\hat{\theta}}(\ell)e^{-\ell\int_{0}^{t}\beta_{s}f_{I}(s)ds}},\end{split} (4.30)

where in the third equality we used (4.29). Combining (4.30) and (4.24), and recalling that FI(t):=0tβsfI(s)𝑑sF_{I}(t):=\int_{0}^{t}\beta_{s}f_{I}(s)ds, we obtain

B1(t)=βtfI(t)k=0kθ^(k)ekFI(t)=0θ^()eFI(t).\displaystyle\begin{split}B_{1}(t)=\beta_{t}f_{I}(t)\frac{\sum_{k=0}^{\infty}k{\hat{\theta}}(k)e^{-kF_{I}(t)}}{\sum_{\ell=0}^{\infty}{\hat{\theta}}(\ell)e^{-\ell F_{I}(t)}}.\end{split} (4.31)

As desired, this expresses B1(t)B_{1}(t) purely in terms of θ^,fI\ {\hat{\theta}},\ f_{I} and FI(t)F_{I}(t). Combining (4.31) with (4.20) and (4.23) establishes the first and second equation of (2.4), thus concluding the proof. ∎

Remark 4.11.

In the proof of Theorem 4.10 we showed that the jump rate q^v(t,X,1𝒯)\hat{q}_{v}(t,X^{\mathcal{T}}_{\varnothing,1}) as defined in (4.13) is not path dependent on the event {X(t)=S}\{X_{\varnothing}(t-)=S\}. By a similar argument that appeals to Proposition 4.7, one can show that q^v(t,X,1𝒯)\hat{q}_{v}(t,X^{\mathcal{T}}_{\varnothing,1}) is also not path dependent on the event {X(t)=S}c\{X_{\varnothing}(t-)=S\}^{c}, thereby showing that X,1X_{\varnothing,1} is a Markov process. The analogue of the latter property can also be shown to hold for the discrete-time SIR process using a similar (in fact, simpler) proof. Numerical evidence supporting this property for the discrete-time SIR process on trees was first provided in [WortThesis18].

Theorem 4.12.

Let fIf_{I} be as in (4.12) and set 𝒩:={m0:θ(m)>0}{\mathcal{N}}:=\{m\in{\mathbb{N}}_{0}\ :\ \theta(m)>0\}. Then ((PS,S;k((P_{S,S;k}, PS,I;k)k𝒩P_{S,I;k})_{k\in{\mathcal{N}}}, PI,I)P_{I,I}), as defined in (4.12), is the unique solution to the following system of ODEs:

{P˙S,S;k=βkfIPS,S;k,k𝒩,P˙S,I;k=βkfIPS,S;kρPS,I;k,k𝒩,P˙I,I=ρPI,I,\begin{cases}\dot{P}_{S,S;k}=-{\beta}kf_{I}P_{S,S;k},&k\in{\mathcal{N}},\\ \dot{P}_{S,I;k}={\beta}kf_{I}P_{S,S;k}-{\rho}P_{S,I;k},&k\in{\mathcal{N}},\\ \dot{P}_{I,I}=-{\rho}P_{I,I},\par\end{cases} (4.32)

with initial conditions

{PS,S;k(0)=1,k𝒩,PS,I;k(0)=0,k𝒩,PI,I(0)=1,\begin{cases}P_{S,S;k}(0)=1,&k\in{\mathcal{N}},\\ P_{S,I;k}(0)=0,&k\in{\mathcal{N}},\\ P_{I,I}(0)=1,\end{cases} (4.33)
Proof.

Throughout the proof, in order to simplify the notation we write XX in lieu of X𝒯X^{{\mathcal{T}}}. By Assumption A, the fact that fIf_{I} defined in (4.12) is continuous (since by Theorem (4.10) it is characterized in terms of the solution of the ODE system (2.4)-(2.5)) and the fact that the ODE system (4.32) is linear, the initial value problem (4.32)-(4.33) has a unique solution. Clearly, from (4.12), the initial conditions (4.33) hold. Next, we show that (4.32) is satisfied.

We start by considering PS,S;kP_{S,S;k}. Fix t0t\geq 0, h>0h>0 and k𝒩k\in{\mathcal{N}}. Since s0=(X(0)=0)>0s_{0}={\mathbb{P}}(X_{\varnothing}(0)=0)>0 and dd_{\varnothing} is independent of X(0)X_{\varnothing}(0), (X(0)=S,d)>0{\mathbb{P}}(X_{\varnothing}(0)=S,\ d_{\varnothing})>0. From (4.12), noting that X(t)=y𝒳kX_{\partial_{\varnothing}}(t)=y\in\mathcal{X}^{k} implicitly implies d=k,d_{\varnothing}=k, we have

PS,S;k(t+h)=(X(t+h)=S|X(0)=S,d=k)=y𝒳k(X(t+h)=S,X(0)=S,X(t)=y)(X(0)=S,d=k)=y𝒳k(X(t+h)=S,X(t)=S,X(t)=y)(X(0)=S,d=k)=y𝒮k,t(X(t+h)=S|X(t)=S,X(t)=y)(X(t)=S,X(t)=y|X(0)=S,d=k),\displaystyle\begin{split}&P_{S,S;k}(t+h)\\ &={\mathbb{P}}(X_{\varnothing}(t+h)=S\ |\ X_{\varnothing}(0)=S,\ d_{\varnothing}=k)\\ &=\sum_{y\in\mathcal{X}^{k}}\frac{{\mathbb{P}}(X_{\varnothing}(t+h)=S,\ X_{\varnothing}(0)=S,\ X_{\partial_{\varnothing}}(t)=y)}{{\mathbb{P}}(X_{\varnothing}(0)=S,\ d_{\varnothing}=k)}\\ &=\sum_{y\in\mathcal{X}^{k}}\frac{{\mathbb{P}}(X_{\varnothing}(t+h)=S,\ X_{\varnothing}(t)=S,\ X_{\partial_{\varnothing}}(t)=y)}{{\mathbb{P}}(X_{\varnothing}(0)=S,\ d_{\varnothing}=k)}\\ &=\sum_{y\in{\mathcal{S}}_{k,t}}{\mathbb{P}}(X_{\varnothing}(t+h)=S|X_{\varnothing}(t)=S,X_{{\partial_{\varnothing}}}(t)=y){\mathbb{P}}(X_{\varnothing}(t)=S,X_{{\partial_{\varnothing}}}(t)=y|X_{\varnothing}(0)=S,d_{\varnothing}=k),\end{split} (4.34)

where 𝒮k,t:={y𝒳k:(X(t)=S,X(t)=y)>0}{\mathcal{S}}_{k,t}:=\{y\in\mathcal{X}^{k}\ :\ {\mathbb{P}}(X_{\varnothing}(t)=S,\ X_{\partial_{\varnothing}}(t)=y)>0\}, and the monotonicity (4.2) of the SIR process is used in the third and fourth equality. Since the jump rate of a susceptible individual whose neighbors’ states are equal to y𝒳ky\in\mathcal{X}^{k} is equal to βt(y){\beta_{t}}{\mathcal{I}}(y), we have that

(X(t+h)=S|X(t)=S,X(t)=y)=1hβt(y)+o(h).{\mathbb{P}}(X_{\varnothing}(t+h)=S\ |\ X_{\varnothing}(t)=S,X_{{\partial_{\varnothing}}}(t)=y)=1-h{\beta_{t}}{\mathcal{I}}(y)+o(h). (4.35)

The expression on the right-hand side of (4.35) does not depend on the exact values of the k(y)k-{\mathcal{I}}(y) elements of yy that are not equal to II. Thus, substituting the expression in (4.35) into the last line of (4.34) and rewriting the sum to be over the number of infected neighbors of \varnothing,

PS,S;k(t+h)=j=0k(1hβtj+o(h))(X(t)=S,(X(t))=j|X(0)=S,d=k)=j=0k(1hβtj+o(h))((X(t))=j|X(t)=S,X(0)=S,d=k)PS,S;k(t)=j=0k(1hβtj+o(h))((X(t))=j|X(t)=S,d=k)PS,S;k(t),\displaystyle\begin{split}P_{S,S;k}(t+h)=&\sum_{j=0}^{k}(1-h{\beta_{t}}j+o(h))\ {\mathbb{P}}(X_{\varnothing}(t)=S,\ {\mathcal{I}}(X_{{\partial_{\varnothing}}}(t))=j\ |\ X_{\varnothing}(0)=S,d_{\varnothing}=k)\\ =&\sum_{j=0}^{k}(1-h{\beta_{t}}j+o(h))\ {\mathbb{P}}({\mathcal{I}}(X_{{\partial_{\varnothing}}}(t))=j\ |X_{\varnothing}(t)=S,\ X_{\varnothing}(0)=S,\ d_{\varnothing}=k)\ P_{S,S;k}(t)\\ =&\sum_{j=0}^{k}(1-h{\beta_{t}}j+o(h))\ {\mathbb{P}}({\mathcal{I}}(X_{{\partial_{\varnothing}}}(t))=j\ |X_{\varnothing}(t)=S,\ d_{\varnothing}=k)\ P_{S,S;k}(t),\end{split}

where in the last equality we used the monotonicity of the SIR process (4.2). Applying Proposition 4.7 with α=0\alpha=0, it follows that {Xi(t):i}\{X_{i}(t)\ :\ i\sim\varnothing\} are conditionally i.i.d. given {X(t)=S,d=k}\{X_{\varnothing}(t)=S,d_{\varnothing}=k\}. Furthermore, for k1k\geq 1 and m[0,k]m\in{\mathbb{N}}\cap[0,k], by Proposition 4.8 and another application of Proposition 4.7 with A={mv}v𝕍A=\left\{mv\right\}_{v\in\mathbb{V}}, 𝕍A={}\partial^{\mathbb{V}}A=\left\{\varnothing\right\} and B={m}B={\mathbb{N}}\setminus\left\{m\right\}, and observing that d==1𝟏{X(0)}d_{\varnothing}=\sum_{\ell=1}^{\infty}{\bm{1}}_{\left\{X_{\ell}(0)\neq\star\right\}}, we have that

(Xm(t)=I|X(t)=S,d=k)=(Xm(t)=I|X(t)=S,m𝒯)=fI(t),{\mathbb{P}}(X_{m}(t)=I\ |\ X_{\varnothing}(t)=S,\ d_{\varnothing}=k)={\mathbb{P}}(X_{m}(t)=I\ |\ X_{\varnothing}(t)=S,\ m\in{\mathcal{T}})=f_{I}(t),

where fIf_{I} is as in (4.12). Therefore, conditional on X(t)=SX_{\varnothing}(t)=S and d=kd_{\varnothing}=k, (X(t)){\mathcal{I}}(X_{{\partial_{\varnothing}}}(t)) has a binomial distribution with parameters (k,fI(t))(k,f_{I}(t)). It follows that, letting YY be a binomial random variable with parameters (k,fI(t))(k,f_{I}(t)),

PS,S;k(t+h)=PS,S;k(t)(𝔼[1hβtY]+o(h))=PS,S;k(t)(1hβtkfI+o(h)).\displaystyle\begin{split}P_{S,S;k}(t+h)&=P_{S,S;k}(t)({\mathbb{E}}[1-h{\beta_{t}}Y]+o(h))\\ &=P_{S,S;k}(t)(1-h{\beta_{t}}kf_{I}+o(h)).\end{split} (4.36)

This implies

limh0+PS,S;k(t+h)PS,S;k(t)h=limh0+(1hβtkfI(t)+o(h)1)PS,S;k(t)h=βtkfI(t)PS,S;k(t),\displaystyle\lim_{h\rightarrow 0^{+}}\frac{P_{S,S;k}(t+h)-P_{S,S;k}(t)}{h}=\lim_{h\rightarrow 0^{+}}\frac{(1-h{\beta_{t}}kf_{I}(t)+o(h)-1)P_{S,S;k}(t)}{h}=-{\beta_{t}}kf_{I}(t)P_{S,S;k}(t), (4.37)

which proves the first equation in (4.32)

The derivations of the ODEs for PS,I;kP_{S,I;k} and PI,IP_{I,I} are similar, and are thus only outlined below. As in the last line of (4.34), we start by writing

PS,I;k(t+h)=𝒬I(h)+𝒬S(h),P_{S,I;k}(t+h)={\mathcal{Q}}_{I}(h)+{\mathcal{Q}}_{S}(h), (4.38)

where for b=Ib=I and b=Sb=S

𝒬b(h)=j=0k(X(t+h)=I,X(t)=b,(X(t))=j|X(0)=S,d=k).\displaystyle{\mathcal{Q}}_{b}(h)=\sum_{j=0}^{k}{\mathbb{P}}(X_{\varnothing}(t+h)=I,\ X_{\varnothing}(t)=b,\ {\mathcal{I}}(X_{{\partial_{\varnothing}}}(t))=j\ |\ X_{\varnothing}(0)=S,\ d_{\varnothing}=k).

Recalling the definition of the SIR rates q0q_{0} as in (4.3) and using arguments similar to what used to derive (4.35)-(4.36), 𝒬S(h)=(hβtkfI(t)+o(h))PS,S;k(t){\mathcal{Q}}_{S}(h)=(h{\beta_{t}}kf_{I}(t)+o(h))P_{S,S;k}(t) and

𝒬I(h)=j=0k(1ρth+o(h))(X(t)=I,(X(t))=j|X(0)=S,d=k)=(1ρth+o(h))j=0k(X(t)=I,(X(t))=j|X(0)=S,d=k)=(1ρth+o(h))(X(t)=I|X(0)=S,d=k)=(1ρth+o(h))PS,I;k(t).\displaystyle\begin{split}{\mathcal{Q}}_{I}(h)&=\sum_{j=0}^{k}(1-{\rho_{t}}h+o(h)){\mathbb{P}}(X_{\varnothing}(t)=I,\ {\mathcal{I}}(X_{{\partial_{\varnothing}}}(t))=j\ |\ X_{\varnothing}(0)=S,\ d_{\varnothing}=k)\\ &=(1-{\rho_{t}}h+o(h))\sum_{j=0}^{k}{\mathbb{P}}(X_{\varnothing}(t)=I,\ {\mathcal{I}}(X_{{\partial_{\varnothing}}}(t))=j\ |\ X_{\varnothing}(0)=S,\ d_{\varnothing}=k)\\ &=(1-{\rho_{t}}h+o(h)){\mathbb{P}}(X_{\varnothing}(t)=I\ |\ X_{\varnothing}(0)=S,\ d_{\varnothing}=k)\\ &=(1-{\rho_{t}}h+o(h))P_{S,I;k}(t).\end{split} (4.39)

Substituting the last two displays into (4.38), we obtain PS,I;k(t+h)PS,I;k(t)=hkβtfI(t)PS,S;k(t)ρthPS,I;k(t)+o(h)P_{S,I;k}(t+h)-P_{S,I;k}(t)=hk{\beta_{t}}f_{I}(t)P_{S,S;k}(t)-{\rho_{t}}hP_{S,I;k}(t)+o(h),which implies the second equation in (4.32).

Next, to obtain the ODE for PI,IP_{I,I} note that by definition of the jump rate (4.3), setting 𝒩:={k0:θ(k)>0}{\mathcal{N}}:=\left\{k\in{\mathbb{N}}_{0}\ :\ \theta(k)>0\right\},

PI,I(t+h)=k𝒩y𝒳k(X(t+h)=I|X(t)=I,X(t)=y)(X(t)=I,X(t)=y|X(0)=I)=(1hρt+o(h))k𝒩y𝒳k(X(t)=I,X(t)=y|X(0)=I)=(1hρt+o(h))PI,I(t),\displaystyle\begin{split}&P_{I,I}(t+h)\\ &=\sum_{k\in{\mathcal{N}}}\sum_{y\in\mathcal{X}^{k}}{\mathbb{P}}(X_{\varnothing}(t+h)=I|X_{\varnothing}(t)=I,X_{\partial_{\varnothing}}(t)=y){\mathbb{P}}(X_{\varnothing}(t)=I,\ X_{\partial_{\varnothing}}(t)=y|X_{\varnothing}(0)=I)\\ &=(1-h{\rho_{t}}+o(h))\sum_{k\in{\mathcal{N}}}\sum_{y\in\mathcal{X}^{k}}{\mathbb{P}}(X_{\varnothing}(t)=I,\ X_{\partial_{\varnothing}}(t)=y\ |\ X_{\varnothing}(0)=I)\\ &=(1-h{\rho_{t}}+o(h))P_{I,I}(t),\end{split} (4.40)

which implies the third equation in (4.32) and concludes the proof.

We can combine the results above to prove Theorem 2.5.

Proof of Theorem 2.5.

By Theorem 4.5, limnsGn(t)=(X𝒯(t)=S)\lim_{n\rightarrow\infty}s^{G_{n}}(t)={\mathbb{P}}(X^{\mathcal{T}}_{\varnothing}(t)=S) and limniGn(t)=(X𝒯(t)=I)\lim_{n\rightarrow\infty}i^{G_{n}}(t)={\mathbb{P}}(X^{\mathcal{T}}_{\varnothing}(t)=I). By Theorem 4.12, we can characterize the transition rates of X𝒯(t)X^{{\mathcal{T}}}_{\varnothing}(t) defined in (4.12) as the solution to the ODE system (4.32)-(4.33). Let fIf_{I} and fSf_{S} be as in (4.12), and FI(t)=0tβsfI(s)𝑑sF_{I}(t)=\int_{0}^{t}\beta_{s}f_{I}(s)ds, t[0,)t\in[0,\infty) .Then we can solve the ODE system (4.32)-(4.33) as follows:

PS,S;k(t)\displaystyle P_{S,S;k}(t) =ekFI(t),\displaystyle=e^{-kF_{I}(t)},
PS,I;k(t)\displaystyle P_{S,I;k}(t) =e0tρu𝑑u0tkekFI(u)e0uρτ𝑑τβufI(u)𝑑u,\displaystyle=e^{-\int_{0}^{t}\rho_{u}du}\int_{0}^{t}ke^{-kF_{I}(u)}e^{\int_{0}^{u}\rho_{\tau}d\tau}\beta_{u}f_{I}(u)du,
PI,I(t)\displaystyle P_{I,I}(t) =e0tρu𝑑u.\displaystyle=e^{-\int_{0}^{t}\rho_{u}du}.

In view of (4.12), by averaging over dd_{\varnothing}, that is, by multiplying each of the quantities above by θ(k)=(d=k)\theta(k)={\mathbb{P}}(d_{\varnothing}=k) and summing over kk\in{\mathbb{N}}, we conclude that

PS,S(t)\displaystyle P_{S,S}(t) =Mθ(FI(t)),\displaystyle=M_{\theta}(-F_{I}(t)),
PS,I(t)\displaystyle P_{S,I}(t) =e0tρu𝑑u0tMθ(FI(u))e0uρτ𝑑τβufI(u)𝑑u,\displaystyle=e^{-\int_{0}^{t}\rho_{u}du}\int_{0}^{t}M^{\prime}_{\theta}(-F_{I}(u))e^{\int_{0}^{u}\rho_{\tau}d\tau}\beta_{u}f_{I}(u)du,

where Mθ(z)=k=0kθ(k)ekzM^{\prime}_{\theta}(z)=\sum_{k=0}^{\infty}k\theta(k)e^{kz}, and the exchange in order of summation and integration is justified by the fact that every term is non-negative. By Theorem 4.10, (fS(f_{S}, fIf_{I}, FI)F_{I}) solve the ODE system (2.4)-(2.5). Finally, since (X𝒯(t)=S)=s0PS,S(t){\mathbb{P}}(X^{\mathcal{T}}_{\varnothing}(t)=S)=s_{0}P_{S,S}(t) and (X𝒯(t)=I)=s0PS,I(t)+i0PI,I(t){\mathbb{P}}(X^{\mathcal{T}}_{\varnothing}(t)=I)=s_{0}P_{S,I}(t)+i_{0}P_{I,I}(t), equation (2.7) follows. This completes the proof. ∎

4.2.3. Proof of Theorem 2.10

Now, we turn our attention to the SEIR process. Since its derivation is similar to that of Theorem 2.5, we relegate most of the details to Appendix A. For a,b𝒳¯a,b\in\bar{\mathcal{X}}, t[0,)t\in[0,\infty) and k{m0:θ(m)>0}k\in\{m\in{\mathbb{N}}_{0}\ :\ \theta(m)>0\}, define

Qa,b;kθ(t)\displaystyle Q^{\theta}_{a,b;k}(t) (X¯𝒯(t)=b|X¯𝒯(0)=a,d=k),\displaystyle\coloneqq{\mathbb{P}}({\bar{X}}^{{\mathcal{T}}}_{\varnothing}(t)=b\ |\ {\bar{X}}^{{\mathcal{T}}}_{\varnothing}(0)=a,\ d_{\varnothing}=k), (4.41)
Qa,bθ(t)\displaystyle Q^{\theta}_{a,b}(t) (X¯𝒯(t)=b|X¯𝒯(0)=a),\displaystyle\coloneqq{\mathbb{P}}({\bar{X}}^{{\mathcal{T}}}_{\varnothing}(t)=b\ |\ {\bar{X}}^{{\mathcal{T}}}_{\varnothing}(0)=a),
gaθ(t)\displaystyle g^{\theta}_{a}(t) (X¯1𝒯(t)=a|X¯𝒯(t)=S, 1𝒯).\displaystyle\coloneqq{\mathbb{P}}({\bar{X}}^{{\mathcal{T}}}_{1}(t)=a\ |\ {\bar{X}}^{{\mathcal{T}}}_{\varnothing}(t)=S,\ 1\in{\mathcal{T}}).

When clear from the context, we omit the dependence on θ\theta and write Qa,b,Qa,b;kQ_{a,b},\ Q_{a,b;k} and gag_{a}.

Theorem 4.13.

Let gSg_{S}, gEg_{E} and gIg_{I} be as in (4.41) and set GI(t):=0tβsgI(s)𝑑sG_{I}(t):=\int_{0}^{t}\beta_{s}g_{I}(s)ds for t[0,)t\in[0,\infty). Then, (gS(g_{S}, gEg_{E}, gIg_{I}, GI)G_{I}) solves the ODE system (2.11)-(2.12).

The proof of Theorem 4.13 is given in Appendix A.

Theorem 4.14.

Let gIg_{I} be as in (4.41) and set 𝒩:={m0:θ(m)>0}{\mathcal{N}}:=\{m\in{\mathbb{N}}_{0}\ :\ \theta(m)>0\}. Then ((QS,i;k),i𝒳¯{R},k𝒩((Q_{S,i;k})_{,i\in\bar{\mathcal{X}}\setminus\{R\},k\in{\mathcal{N}}}, QE,EQ_{E,E}, QE,IQ_{E,I}, QI,I)Q_{I,I}) is the unique solution to the following system of ODEs:

{Q˙S,S;k=βkgIQS,S;k,k𝒩,Q˙S,E;k=βkgIQS,S;kλQS,E;k,k𝒩,Q˙S,I;k=λQS,E;kρQS,I;k,k𝒩,Q˙E,E=λQE,E,Q˙E,I=λQE,EρQE,I,Q˙I,I=ρQI,I,\displaystyle\begin{cases}\dot{Q}_{S,S;k}=-{\beta}kg_{I}Q_{S,S;k},&k\in{\mathcal{N}},\\ \dot{Q}_{S,E;k}={\beta}kg_{I}Q_{S,S;k}-{\lambda}Q_{S,E;k},&k\in{\mathcal{N}},\\ \dot{Q}_{S,I;k}={\lambda}Q_{S,E;k}-{\rho}Q_{S,I;k},&k\in{\mathcal{N}},\\ \dot{Q}_{E,E}=-{\lambda}Q_{E,E},\\ \dot{Q}_{E,I}={\lambda}Q_{E,E}-\rho Q_{E,I},\\ \dot{Q}_{I,I}=-{\rho}Q_{I,I},\end{cases} (4.42)

with initial conditions

{Qa,b(0)=𝟏{a=b},a,b𝒳¯,Qa,b;k(0)=𝟏{a=b},a,b𝒳¯,k𝒩.\displaystyle\begin{cases}Q_{a,b}(0)={\bm{1}}_{\{a=b\}},&a,b\in\bar{\mathcal{X}},\\ Q_{a,b;k}(0)={\bm{1}}_{\{a=b\}},&a,b\in\bar{\mathcal{X}},\ k\in{\mathcal{N}}.\end{cases} (4.43)

The proof of Theorem 4.14 is given in Appendix A. We conclude this section by outlining how the last two theorems are used to prove Theorem 2.10.

Proof of Theorem 2.10.

By Theorem 4.5, limns¯Gn(t)=(X¯𝒯(t)=S)\lim_{n\rightarrow\infty}{\bar{s}}^{G_{n}}(t)={\mathbb{P}}({\bar{X}}^{\mathcal{T}}_{\varnothing}(t)=S), limne¯Gn(t)=(X¯𝒯(t)=E)\lim_{n\rightarrow\infty}{\bar{e}}^{G_{n}}(t)={\mathbb{P}}({\bar{X}}^{\mathcal{T}}_{\varnothing}(t)=E), and limni¯Gn(t)=(X¯𝒯(t)=I)\lim_{n\rightarrow\infty}{\bar{i}}^{G_{n}}(t)={\mathbb{P}}({\bar{X}}^{\mathcal{T}}_{\varnothing}(t)=I). By Theorem 4.14, we can characterize the transition rates of X¯𝒯(t){\bar{X}}^{{\mathcal{T}}}_{\varnothing}(t), given in (4.41) as the solution to the system of ODEs (4.42)-(4.43). We can solve these ODEs to obtain an expression for QS,S;k,QS,E;k,Q_{S,S;k},\ Q_{S,E;k}, in terms of gS,gEg_{S},\ g_{E} and gIg_{I} (defined in (4.41)) which, along with GI(t)=0tgI(s)𝑑sG_{I}(t)=\int_{0}^{t}g_{I}(s)ds for t[0,)t\in[0,\infty), by Theorem 4.13, solve the ODEs (2.11)-(2.12). Observing that QS,b(t)=k0θ(k)QS,b;k(t)Q_{S,b}(t)=\sum_{k\in{\mathbb{N}}_{0}}\theta(k)Q_{S,b;k}(t) for all t[0,)t\in[0,\infty) and b{S,E}b\in\{S,E\}, and noting that

(X¯𝒯(t)=S)\displaystyle{\mathbb{P}}({\bar{X}}^{\mathcal{T}}_{\varnothing}(t)=S) =s0QS,S(t),\displaystyle=s_{0}Q_{S,S}(t),
(X¯𝒯(t)=E)\displaystyle{\mathbb{P}}({\bar{X}}^{\mathcal{T}}_{\varnothing}(t)=E) =s0QS,E(t)+e0QE,E(t)\displaystyle=s_{0}Q_{S,E}(t)+e_{0}Q_{E,E}(t)
(X¯𝒯(t)=I)\displaystyle{\mathbb{P}}({\bar{X}}^{\mathcal{T}}_{\varnothing}(t)=I) =s0QS,I(t)+e0QE,I(t)+i0QI,I(t)\displaystyle=s_{0}Q_{S,I}(t)+e_{0}Q_{E,I}(t)+i_{0}Q_{I,I}(t)

establishes the theorem. ∎

4.3. Proofs related to the Outbreak Size

In this section, we prove Theorem 3.1 and Theorem 3.5, which characterize the large nn limit of the total fraction of individuals still susceptible at the end of an SIR or SEIR outbreak on the locally-tree like graph sequences we consider. Recall the standing assumptions made at the beginning of Section 4. We start by introducing some notation to simplify the exposition. First, define

d¯θ^:=min{d0:θ^(d)>0},\underline{d}_{\hat{\theta}}:=\min\{d\in{\mathbb{N}}_{0}\ :\ {\hat{\theta}}(d)>0\}, (4.44)

and recalling that Mθ^M_{\hat{\theta}} is the moment generating function of θ^{\hat{\theta}}, set

Φ(z):=Φθ^(z):=Mθ^(z)Mθ^(z),z[0,).\displaystyle\begin{split}\Phi(z):=\Phi_{\hat{\theta}}(z):=\frac{M_{\hat{\theta}}^{\prime}(-z)}{M_{\hat{\theta}}(-z)},\quad z\in[0,\infty).\end{split} (4.45)

For all z[0,)z\in[0,\infty) we have Mθ^(z)=𝔼θ^[edz]𝔼θ^[1]=1M_{\hat{\theta}}(-z)={\mathbb{E}}_{\hat{\theta}}[e^{-dz}]\leq{\mathbb{E}}_{\hat{\theta}}[1]=1. Furthermore, for z[0,)z\in[0,\infty), Mθ^(z)=k=1kekzθ^(k)M_{\hat{\theta}}^{\prime}(-z)=\sum_{k=1}^{\infty}ke^{-kz}{\hat{\theta}}(k), where the interchange of the sum and derivative is justified because kekzkke^{-kz}\leq k and θ^{\hat{\theta}} has finite mean. We start with an elementary lemma.

Lemma 4.15.

Φθ^:[0,)[0,)\Phi_{\hat{\theta}}:[0,\infty)\rightarrow[0,\infty) is continuous and satisfies the following properties:

  1. (i)

    Φθ^(0)=𝔼θ^[d]\Phi_{\hat{\theta}}(0)={\mathbb{E}}_{\hat{\theta}}[d];

  2. (ii)

    limzΦθ^(z)=d¯θ^\lim_{z\rightarrow\infty}\Phi_{\hat{\theta}}(z)=\underline{d}_{\hat{\theta}};

  3. (iii)

    Φθ^(z)\Phi_{\hat{\theta}}(z) is non-increasing in z[0,)z\in[0,\infty), and strictly decreasing if for every j0j\in{\mathbb{N}}_{0}, θ^δj{\hat{\theta}}\neq\delta_{j}.

Proof.

The property (i) follows immediately from the relation Φθ^(0)=𝔼θ^[d]/𝔼θ^[1]=𝔼θ^[d]\Phi_{\hat{\theta}}(0)={\mathbb{E}}_{\hat{\theta}}[d]/{\mathbb{E}}_{\hat{\theta}}[1]={\mathbb{E}}_{\hat{\theta}}[d].

The stated continuity of Φθ^\Phi_{\hat{\theta}} follows from the dominated convergence theorem and the fact that θ^{\hat{\theta}} has finite mean, which follows from (2.2) and Assumption B. In turn, by the dominated convergence theorem, the latter implies that limz𝔼θ^[dedz]=0\lim_{z\rightarrow\infty}{\mathbb{E}}_{\hat{\theta}}[de^{-dz}]=0. If d¯θ^=0\underline{d}_{\hat{\theta}}=0 then limz𝔼θ^[edz]=θ^(0)>0\lim_{z\rightarrow\infty}{\mathbb{E}}_{\hat{\theta}}[e^{-dz}]={\hat{\theta}}(0)>0, and by (4.45) it follows that limzΦθ^(z)=0=d¯θ^\lim_{z\rightarrow\infty}\Phi_{\hat{\theta}}(z)=0=\underline{d}_{\hat{\theta}}. On the other hand, if d¯θ^>0\underline{d}_{\hat{\theta}}>0, then

limzΦθ^(z)=limzd¯θ^ed¯θ^z+j=d¯θ^+1jejzed¯θ^z+j=d¯θ^+1ejz=limzd¯θ^ed¯θ^zed¯θ^z=d¯θ^.\displaystyle\lim_{z\rightarrow\infty}\Phi_{\hat{\theta}}(z)=\lim_{z\rightarrow\infty}\frac{\underline{d}_{\hat{\theta}}e^{-\underline{d}_{\hat{\theta}}z}+\sum_{j=\underline{d}_{\hat{\theta}}+1}^{\infty}je^{-jz}}{e^{-\underline{d}_{\hat{\theta}}z}+\sum_{j=\underline{d}_{\hat{\theta}}+1}^{\infty}e^{-jz}}=\lim_{z\rightarrow\infty}\frac{\underline{d}_{\hat{\theta}}e^{-\underline{d}_{\hat{\theta}}z}}{e^{-\underline{d}_{\hat{\theta}}z}}=\underline{d}_{\hat{\theta}}. (4.46)

This proves (ii).

Next, observe that

Φθ^(z)=ddzlogMθ^(z)=dd(z)logMθ^(z).\Phi_{\hat{\theta}}(z)=-\frac{d}{dz}\log M_{\hat{\theta}}(-z)=\frac{d}{d(-z)}\log M_{\hat{\theta}}(-z).

Since the moment generating function of any measure in 𝒫(){\mathcal{P}}({\mathbb{R}}) is log-convex (which follows from an application of Hölder’s inequality), and strictly log-convex unless the measure is equal to δx\delta_{x} for xx\in{\mathbb{R}}, (iii) follows. ∎

We now prove Theorem 3.1.

Proof of Theorem 3.1.

Let fIf_{I} and fSf_{S} be as in (4.12) and set FI(t)=0tβsfI(s)𝑑sF_{I}(t)=\int_{0}^{t}\beta_{s}f_{I}(s)ds. By (2.7), s()(t)=s0Mθ(FI(t))s^{(\infty)}(t)=s_{0}M_{\theta}(-F_{I}(t)). By the dominated convergence theorem, zMθ(z)=k=0θ(k)ezkz\mapsto M_{\theta}(-z)=\sum_{k=0}^{\infty}\theta(k)e^{-zk} is continuous on (0,)(0,\infty).

We now turn to the study of the large-time limit of FIF_{I}.

By Theorem 2.10, (fS,(f_{S}, fI,f_{I}, FI)F_{I}) satisfy the ODE system (2.4). For any a[0,1]a\in[0,1] and b(0,)b\in(0,\infty), the point (a,0,b)(a,0,b) is a fixed point of the system. We claim that as tt\rightarrow\infty, (fS(t),(f_{S}(t), fI(t),f_{I}(t), FI(t))F_{I}(t)) converges to one such point, and then identify the corresponding bb as the solution of an equation. Near any t0t\geq 0 such that fI(t)>0f_{I}(t)>0, FIF_{I} is strictly increasing, and thus it is invertible. Let FI():=limtFI(t)F_{I}(\infty):=\lim_{t\rightarrow\infty}F_{I}(t), which exists since FIF_{I} is non-decreasing. We can change variables for F[0,FI()]F\in[0,F_{I}(\infty)] and write x(F):=fI(FI1(F))x(F):=f_{I}(F_{I}^{-1}(F)) and y(F):=fS(FI1(F))y(F):=f_{S}(F_{I}^{-1}(F)). We write β\beta^{\ast} (resp. ρ\rho^{\ast}) for the composition of β\beta (resp. ρ\rho) with FI1F_{I}^{-1}. Recalling the definition of Φ\Phi in (4.45), we rewrite the first two equations in (2.4) as

{y=y(1Φ)x=yΦ(1+ρβ)+x,\begin{cases}y^{\prime}=y(1-\Phi)\\ x^{\prime}=y\Phi-(1+\frac{\rho^{\ast}}{\beta^{\ast}})+x,\end{cases} (4.47)

Since FI(0)=0F_{I}(0)=0, and fS(0)=s0f_{S}(0)=s_{0}, we can solve the first equation to obtain log(y(F)/s0)=F+logMθ^(F)\log(y(F)/s_{0})=F+\log M_{\hat{\theta}}(-F), which is equivalent to

y(F)=s0Mθ^(F)eF.y(F)=s_{0}M_{\hat{\theta}}(-F)e^{F}. (4.48)

Substituting this into the second equation in (4.47), we obtain a linear ODE for xx. Recalling that x(0)=fI(0)=i0x(0)=f_{I}(0)=i_{0} and that i0+s0=1i_{0}+s_{0}=1, we solve this equation to obtain

x(F)=i0eF+eF0Fs0Mθ^(z)Φ(z)𝑑zeF0Fez(1+ρ(z)β(z))𝑑z=i0eF+eFs0(1Mθ^(F))eF0Fez(1+ρ(z)β(z))𝑑z=eFy(F)eF(1eF)eF0Fezρ(z)β(z)𝑑z=1y(F)eF0Fezρ(z)β(z)𝑑z,\displaystyle\begin{split}x(F)&=i_{0}e^{F}+e^{F}\int_{0}^{F}s_{0}M_{\hat{\theta}}(-z)\Phi(z)dz-e^{F}\int_{0}^{F}e^{-z}\left(1+\frac{\rho^{\ast}(z)}{\beta^{\ast}(z)}\right)dz\\ &=i_{0}e^{F}+e^{F}s_{0}(1-M_{\hat{\theta}}(-F))-e^{F}\int_{0}^{F}e^{-z}\left(1+\frac{\rho^{\ast}(z)}{\beta^{\ast}(z)}\right)dz\\ &=e^{F}-y(F)-e^{F}(1-e^{-F})-e^{F}\int_{0}^{F}e^{-z}\frac{\rho^{\ast}(z)}{\beta^{\ast}(z)}dz\\ &=1-y(F)-e^{F}\int_{0}^{F}e^{-z}\frac{\rho^{\ast}(z)}{\beta^{\ast}(z)}dz,\end{split} (4.49)

where in the second line we used the fact that Mθ^(0)=1M_{\hat{\theta}}(0)=1, and in the first and third line we applied (4.48).

We now claim that (4.49) shows that FI()<F_{I}(\infty)<\infty. Since FI(t)=0tβsfI(s)𝑑sF_{I}(t)=\int_{0}^{t}\beta_{s}f_{I}(s)ds and β\beta satisfies Assumption A, this implies that limtfI(t)=0\lim_{t\rightarrow\infty}f_{I}(t)=0. First, observe that, if there exists s[0,)s\in[0,\infty) such that fI(s)=0f_{I}(s)=0, then, by (2.4), fI(t)=0f_{I}(t)=0 for all tst\geq s. Next, suppose for the sake of contradiction that FI()=F_{I}(\infty)=\infty. Then, for all t0t\geq 0, fI(t)>0f_{I}(t)>0. By Assumption A, it then follows that 0teFI(s)ρsfI(s)𝑑s>0\int_{0}^{t}e^{-F_{I}(s)}\rho_{s}f_{I}(s)ds>0 for all t>0t>0. By definition, fS(t)[0,1]f_{S}(t)\in[0,1], and so y(F)[0,1]y(F)\in[0,1] for all F[0,FI())F\in[0,F_{I}(\infty)). In particular, lim infFy(F)0\liminf_{F\rightarrow\infty}y(F)\geq 0. But letting FF\rightarrow\infty, (4.49) then implies that limFx(F)=limtfI(t)=\lim_{F\rightarrow\infty}x(F)=\lim_{t\rightarrow\infty}f_{I}(t)=-\infty, which is a contradiction. Therefore, we conclude that FI()<F_{I}(\infty)<\infty and, thus, limtfI(t)=0\lim_{t\rightarrow\infty}f_{I}(t)=0.

Since limtfI(t)=0\lim_{t\rightarrow\infty}f_{I}(t)=0, by setting x(FI())=0x(F_{I}(\infty))=0 in (4.49), we obtain

y(FI())=1eFI()0FI()ezρ(z)β(z)𝑑z=1eFI()0e0uβτfI(τ)𝑑τρufI(u)𝑑u.y(F_{I}(\infty))=1-e^{F_{I}(\infty)}\int_{0}^{F_{I}(\infty)}e^{-z}\frac{\rho^{\ast}(z)}{\beta^{\ast}(z)}dz=1-e^{F_{I}(\infty)}\int_{0}^{\infty}e^{-\int_{0}^{u}\beta_{\tau}f_{I}(\tau)d\tau}\rho_{u}f_{I}(u)du. (4.50)

When combined, (4.48) and (4.50) establish (3.1).

If there exists r(0,)r\in(0,\infty) such that ρt/βt=r\rho_{t}/\beta_{t}=r for all tt, then the integral in the rightmost expression in (3.1) is equal to r(1eFI())r(1-e^{-F_{I}(\infty)}), and thus (3.1) reduces to (3.2). Let Ψr\Psi_{r} be given by (3.4). Using the fact that moment generating functions are log-convex, it follows that Ψr\Psi_{r} is convex. Furthermore, Ψr\Psi_{r} is continuous on [0,log(1+1/r))[0,\log(1+1/r)), Ψr(0)=log(s0)<0\Psi_{r}(0)=\log(s_{0})<0 and limz(log(1+1/r))Ψr(z)=\lim_{z\rightarrow(\log(1+1/r))^{-}}\Psi_{r}(z)=\infty. Therefore, (3.2) has a unique positive solution. This concludes the proof. ∎

We conclude this section by providing a similar characterization of the outbreak size for the SEIR process.

Proof of Theorem 3.5.

Let gS,g_{S}, gE,g_{E}, gIg_{I} be as in (4.41), and set GI(t):=0tβsgI(s)𝑑sG_{I}(t):=\int_{0}^{t}\beta_{s}g_{I}(s)ds for t[0,)t\in[0,\infty). Note that s¯()(t)=s0Mθ(GI(t)){\bar{s}}^{(\infty)}(t)=s_{0}M_{\theta}(-G_{I}(t)) by (2.13), and by the dominated convergence theorem, MθM_{\theta} (the moment generating function of θ\theta) is continuous on (,0)(-\infty,0).

We now study the large-time limit of GIG_{I}. By Theorem 4.13, (gS,(g_{S}, gE,g_{E}, gI,g_{I}, GI)G_{I}) satisfy the system of ODEs (2.11). Near any t0t\geq 0 such that gI(t)>0g_{I}(t)>0, GIG_{I} is strictly increasing, and, therefore invertible. Let GI():=limtGI(t)G_{I}(\infty):=\lim_{t\rightarrow\infty}G_{I}(t), which exists since GIG_{I} is non-decreasing. We can change variables for G[0,GI()]G\in[0,G_{I}(\infty)], write x(G):=gI(GI1(G))x(G):=g_{I}(G_{I}^{-1}(G)), z(G):=gE(GI1(G))z(G):=g_{E}(G_{I}^{-1}(G)) and y(G):=gS(GI1(G))y(G):=g_{S}(G_{I}^{-1}(G)). We write β\beta^{\ast} (resp. ρ\rho^{\ast}, λ\lambda^{\ast}) for the composition of β\beta (resp. ρ\rho, λ\lambda) with GI1G_{I}^{-1}. By (2.11), letting apostrophe denote differentiation with respect to GG, we have

{y=y(1Φ)z=yΦzλxβ+zx=zλxβ(1+ρβ)+x.\begin{cases}y^{\prime}=y(1-\Phi)\\ z^{\prime}=y\Phi-\frac{z\lambda^{\ast}}{x\beta^{\ast}}+z\\ x^{\prime}=\frac{z\lambda^{\ast}}{x\beta^{\ast}}-(1+\frac{\rho^{\ast}}{\beta^{\ast}})+x.\end{cases} (4.51)

If we let x¯=x+z\bar{x}=x+z, then y,x¯y,\bar{x} satisfy

{y=y(1Φ)x¯=yΦ(1+ρβ)+x¯\begin{cases}y^{\prime}=y(1-\Phi)\\ \bar{x}^{\prime}=y\Phi-(1+\frac{\rho^{\ast}}{\beta^{\ast}})+\bar{x}\end{cases} (4.52)

with y(0)=s0y(0)=s_{0}, x¯(0)=1s0\bar{x}(0)=1-s_{0}, which is the same initial value problem as (4.47). The same argument as that used in the proof of Theorem 3.1 can then be used to conclude the proof. ∎

For the sake of completeness, we also include here the special case of the 22-regular tree (i.e., the infinite line graph), with constant ρ\rho and β\beta, where we can obtain an explicit expression for 0tβsfI(s)𝑑s\int_{0}^{t}\beta_{s}f_{I}(s)ds for all t[0,]t\in[0,\infty].

Proposition 4.16.

Let T2=T_{2}=UGW(δ2)(\delta_{2}), and suppose that there exist r,b>0r,\ b>0 such that for all t[0,)t\in[0,\infty), ρt=r\rho_{t}=r and βt=b\beta_{t}=b. Then, for all t[0,)t\in[0,\infty),

PS,S(t)=((1s0)et(b(1s0)+r)+rb1s0+rb)2,\displaystyle\begin{split}P_{S,S}(t)=\left(\frac{(1-s_{0})e^{-t(b(1-s_{0})+r)}+\frac{r}{b}}{1-s_{0}+\frac{r}{b}}\right)^{2},\end{split} (4.53)

and, hence,

limt(XT2(t)=S)=s0(11+(1s0)br)2.\lim_{t\rightarrow\infty}{\mathbb{P}}(X^{T_{2}}_{\varnothing}(t)=S)=s_{0}\left(\frac{1}{1+(1-s_{0})\frac{b}{r}}\right)^{2}. (4.54)
Proof.

Let PS,SP_{S,S}, fIf_{I} and fSf_{S} be as in (4.12). By Theorem 4.12, P˙S,S=b2fIPS,S\dot{P}_{S,S}=-b2f_{I}P_{S,S}, and therefore

PS,S(t)=exp(2b0tfI(s)𝑑s).P_{S,S}(t)=\exp(-2b\int_{0}^{t}f_{I}(s)ds). (4.55)

Setting θ=δ2\theta=\delta_{2} and, thus, θ^=δ1{\hat{\theta}}=\delta_{1}, the first equation in (2.4) reduces to f˙S(t)=0\dot{f}_{S}(t)=0. Since fS(0)=s0f_{S}(0)=s_{0} and fI(0)=i0=1s0f_{I}(0)=i_{0}=1-s_{0}, the second equation in (2.4) reduces to

f˙I(t)=(bfI(0)+r)fI(t)+b(fI)2.\dot{f}_{I}(t)=-(bf_{I}(0)+r)f_{I}(t)+b(f_{I})^{2}. (4.56)

This is a Bernoulli equation that can be solved explicitly. The constant 0 function is a solution.

For the rest of this proof, we assume that fI(0)(0,1)f_{I}(0)\in(0,1). Let a=(bfI(0)+r)a=-(bf_{I}(0)+r), so that (4.56) is f˙I=afI+bfI2\dot{f}_{I}=af_{I}+bf_{I}^{2}. Let τ:=inf{t>0:fI(t)=0}\tau:=\inf\{t>0\ :f_{I}(t)=0\}. For t[0,τ)t\in[0,\tau), we can divide both sides of the ODE by (fI)2(f_{I})^{2}.

f˙I(fI)2afI=b.\frac{\dot{f}_{I}}{(f_{I})^{2}}-\frac{a}{f_{I}}=b. (4.57)

If we set u(t)=1/fI(t)u(t)=1/f_{I}(t), for t[0,τ)t\in[0,\tau), then u˙=(fI)2f˙I\dot{u}=-(f_{I})^{-2}\dot{f}_{I} and the ODE in (4.57) takes the form

u˙+au=b.\dot{u}+au=-b.

This is a linear equation whose explicit solution is

u(t)=ba(eta1)+etau(0),u(t)=\frac{b}{a}(e^{-ta}-1)+e^{-ta}u(0), (4.58)

which does not blow up in finite time, and therefore τ=\tau=\infty. Since fI(t)=1/u(t)f_{I}(t)=1/u(t), (4.58) implies

fI(t)=fI(0)(fI(0)+rb)fI(0)+rbe(bfI(0)+r)t,f_{I}(t)=\frac{f_{I}(0)(f_{I}(0)+\frac{r}{b})}{f_{I}(0)+\frac{r}{b}e^{(bf_{I}(0)+r)t}},

which can be integrated to conclude that

0tfI(s)𝑑s=t(fI(0)+rb)1blog(fI(0)+rbe(bfI(0)+r)t)+1blog(fI(0)+rb).\int_{0}^{t}f_{I}(s)ds=t\left(f_{I}(0)+\frac{r}{b}\right)-\frac{1}{b}\log\left(f_{I}(0)+\frac{r}{b}e^{(bf_{I}(0)+r)t}\right)+\frac{1}{b}\log\left(f_{I}(0)+\frac{r}{b}\right).

This, combined with (4.55), yields (4.53). Since (XT2(t)=S)=s0PS,S(t){\mathbb{P}}(X_{\varnothing}^{T_{2}}(t)=S)=s_{0}P_{S,S}(t), letting tt\rightarrow\infty, we obtain (4.54). ∎

5. Proof of the Conditional Independence Property

In Section 5.2, we prove the conditional independence property stated in Proposition 4.7 and the symmetry property stated in Proposition 4.8. The proof relies on a certain change of measure result established in [Ganguly2022interacting], which we first summarize in Section 5.1. Throughout, θ𝒫(0)\theta\in{\mathcal{P}}({\mathbb{N}}_{0}) has finite third moment, 𝒯{\mathcal{T}} is a UGW(θ\theta) tree, α[0,1]\alpha\in[0,1] is an interpolation parameter, the rates β\beta, λ\lambda and ρ\rho satisfy Assumption C, and ξ𝒯,α\xi^{{\mathcal{T}},\alpha} is the hybrid S(E)IR process solving (4.4), with initial states satisfying Assumption D.

5.1. A Radon–Nikodym derivative

We let μ:=(ξ𝒯,α)\mu:={\mathcal{L}}(\xi^{{\mathcal{T}},\alpha}) and μt=(ξ𝒯,α[t))\mu_{t-}={\mathcal{L}}(\xi^{{\mathcal{T}},\alpha}[t)) for t(0,)t\in(0,\infty). Given U𝕍U\subset\mathbb{V} and t(0,)t\in(0,\infty), let 𝒟tU:=𝒟([0,t):𝒳¯U){\mathcal{D}}^{U}_{t-}:={\mathcal{D}}([0,t):\bar{\mathcal{X}}_{\star}^{U}) be the set of càdlàg functions [0,t)𝒳¯U[0,t)\rightarrow\bar{\mathcal{X}}_{\star}^{U}. We start with two technical definitions.

Definition 5.1.

Given y𝒟y\in{\mathcal{D}}_{\star} and t[0,)t\in[0,\infty) we let Disct(y):={s(0,t]:y(s)y(s)}\text{Disc}_{t}(y):=\{s\in(0,t]\ :\ y(s-)\neq y(s)\}. We say that x𝒟𝕍x\in{\mathcal{D}}_{\star}^{\mathbb{V}} is proper if for every u,v𝕍u,v\in\mathbb{V} and t[0,)t\in[0,\infty), Disct(xv)Disct(xu)=\text{Disc}_{t}(x_{v})\cap\text{Disc}_{t}(x_{u})=\emptyset.

Definition 5.2.

Fix U𝕍U\subset\mathbb{V} finite, t(0,)t\in(0,\infty) and suppose that x𝒟tUx\in{\mathcal{D}}_{t-}^{U} is proper and Disc(x):=s(0,)Discs(x)\text{Disc}_{\infty}(x):=\cup_{s\in(0,\infty)}\text{Disc}_{s}(x) can be ordered as a strictly increasing sequence {tk(x)}\{t_{k}(x)\}. Then the jump characteristics of xx are the elements {(tk(x),jk(x),vk(x))}(0,)×𝒥×U\left\{(t_{k}(x),j_{k}(x),v_{k}(x))\right\}\subset(0,\infty)\times{\mathcal{J}}\times U where for each kk\in{\mathbb{N}} with k|Disc(x)|k\leq|\text{Disc}_{\infty}(x)|, vk=vk(x)v_{k}=v_{k}(x) is a vertex in UU such that xvkx_{v_{k}} is discontinuous at time tk(x)t_{k}(x), and jk(x)j_{k}(x) is the size of the jump xvk(tk(x))limh0+xvk(tk(x)h)x_{v_{k}}(t_{k}(x))-\lim_{h\rightarrow 0^{+}}x_{v_{k}}(t_{k}(x)-h).

Given U𝕍U\subset\mathbb{V} and t(0,)t\in(0,\infty), we also define a function ψ\psi from the set of functions [0,t)𝒳¯[0,t)\rightarrow\bar{\mathcal{X}} into {0,1}\{0,1\} by

ψ(x)=𝟏{x𝒟tU}𝟏{{vU:xv(0)} is a locally finite tree}.\psi(x)={\bm{1}}_{\left\{x\in{\mathcal{D}}_{t-}^{U}\right\}}{\bm{1}}_{\{\{v\in U\ :\ x_{v}(0)\neq\star\}\text{ is a locally finite tree}\}}. (5.1)

We also recall that 𝕍n={}(k=1nk)\mathbb{V}_{n}=\{\varnothing\}\cup(\cup_{k=1}^{n}{\mathbb{N}}^{k}).

We now state a change of measure result that is established, for general interacting jump processes, in [Ganguly2022interacting]. In the sequel, the exact definition of the reference processes ξ^n{\hat{\xi}}^{n} presented below will not be important, and so we state the following proposition to summarize some key properties we use.

Proposition 5.3.

For each n{1},n\in{\mathbb{N}}\setminus\left\{1\right\}, there exists an 𝒳¯𝕍\bar{\mathcal{X}}_{\star}^{\mathbb{V}}-valued process ξ^n:=ξ^n,α{\hat{\xi}}^{n}:={\hat{\xi}}^{n,\alpha} such that for any t(0,)t\in(0,\infty), A,B𝕍A,B\subset\mathbb{V} with 𝕍A𝕍n1\partial^{\mathbb{V}}A\subset{\mathbb{V}}_{n-1} and (A𝕍A)B=(A\cup\partial^{\mathbb{V}}A)\cap B=\emptyset,

ξ^An[t)ξ^Bn[t)|ξ^𝕍An[t).{\hat{\xi}}^{n}_{A}[t)\perp{\hat{\xi}}^{n}_{B}[t)\ |\ {\hat{\xi}}^{n}_{\partial^{\mathbb{V}}A}[t). (5.2)

Furthermore, ξ^𝕍nn{\hat{\xi}}^{n}_{\mathbb{V}_{n}} is almost surely proper and its jump characteristics {(tin,vin,jin)}\{(t^{n}_{i},v^{n}_{i},j^{n}_{i})\} are well-defined. Moreover, for every t(0,)t\in(0,\infty), μ^tn:=(ξ^n[t)){\hat{\mu}}^{n}_{t-}:={\mathcal{L}}({\hat{\xi}}^{n}[t)) has the property that, almost surely

dμtdμ^tn(ξ^n[t))=ψ(ξ^n[t)])exp(v𝕍nj=1,2(0,t)(qα(j,s,ξ^v,ξ^vn)1)ds)0<tin<tqα(jin,tin,ξ^vjnn,ξ^vinn),\frac{d\mu_{t-}}{d{\hat{\mu}}_{t-}^{n}}({\hat{\xi}}^{n}[t))=\psi({\hat{\xi}}^{n}[t)])\exp\left(-\sum_{\begin{subarray}{c}v\in\mathbb{V}_{n}\\ j=1,2\end{subarray}}\int_{(0,t)}(q_{\alpha}(j,s,{\hat{\xi}}^{\partial}_{v},{\hat{\xi}}^{n}_{\partial_{v}})-1)ds\right)\prod_{0<t_{i}^{n}<t}q_{\alpha}(j^{n}_{i},t_{i}^{n},{\hat{\xi}}^{n}_{v_{j}^{n}},{\hat{\xi}}^{n}_{\partial_{v_{i}^{n}}}), (5.3)

where ψ\psi is defined in (5.1), and qαq_{\alpha} is given in (4.3).

Proof.

An explicit definition of the processes ξ^n{\hat{\xi}}^{n} as a solution of a SDE related to (4.4) is given in [Ganguly2022interacting, (4.3)] by substituting the rate function qαq_{\alpha} to the rate functions rvr^{v} used therein. Assumption 4.1 in [Ganguly2022interacting], i.e., the well-posedness of ξ^n{\hat{\xi}}^{n}, follows from an application of [Ganguly2022hydrodynamic, Theorem C.2] on observing that [Ganguly2022hydrodynamic, Assumption C.1] holds by Assumption B, the definition of qαq_{\alpha} in (4.3), and the form of the driving noises in (4.4). Assumption C implies that [Ganguly2022interacting, Assumption 3.1, Assumption 3.4] holds with qαq_{\alpha} in place of rvr^{v}.

Then, [Ganguly2022interacting, Proposition 4.4] establishes (5.2). By [Ganguly2022interacting, Lemma 4.8], ξ^𝕍vn{\hat{\xi}}^{n}_{\mathbb{V}_{v}} is proper. Finally, (5.3) holds by [Ganguly2022interacting, Corollary 4.11]. ∎

Given n{1}n\in{\mathbb{N}}\setminus\{1\}, A𝕍nA\subset\mathbb{V}_{n} and t>0t>0 we define

nt(A):={ξ^𝕍An(t)=S𝕍A}=v𝕍A{ξ^vn(t)=S}.{\mathcal{E}}_{n}^{t}(A):=\{{\hat{\xi}}^{n}_{\partial^{\mathbb{V}}A}(t-)=S_{\partial^{\mathbb{V}}A}\}=\bigcap_{v\in\partial^{\mathbb{V}}A}\{{\hat{\xi}}^{n}_{v}(t-)=S\}. (5.4)

5.2. Proof of Proposition 4.7

We start by establishing a factorization result for the Radon-Nikodym derivative established in 5.1. We recall that v𝕍\partial^{\mathbb{V}}_{v} denote the neighborhood in 𝕍\mathbb{V} of v𝕍v\in\mathbb{V}. We also set Cv:=Cv𝕍={vk}kC_{v}:=C_{v}^{\mathbb{V}}=\{vk\}_{k\in{\mathbb{N}}}.

Lemma 5.4.

Let n{1}n\in{\mathbb{N}}\setminus\{1\} and fix A,B𝕍A,B\subset\mathbb{V} with 𝕍A𝕍n1\partial^{\mathbb{V}}A\subset{\mathbb{V}}_{n-1} and such that A,𝕍AA,\ \partial^{\mathbb{V}}A and BB form a partition of 𝕍\mathbb{V}. Let μ^tn{\hat{\mu}}^{n}_{t-} and ξ^tn{\hat{\xi}}^{n}_{t-} be as in Proposition 5.3. Then there exist measurable functions f~1:𝒟tA𝕍A[0,){\tilde{f}}_{1}:{\mathcal{D}}_{t-}^{A\cup\partial^{\mathbb{V}}A}\rightarrow[0,\infty) and f~2:𝒟tB𝕍A[0,){\tilde{f}}_{2}:{\mathcal{D}}_{t-}^{B\cup\partial^{\mathbb{V}}A}\rightarrow[0,\infty) such that for every t(0,),t\in(0,\infty),

dμtdμ^tn(ξ^n[t))=f~1(ξ^A𝕍An[t))f~2(ξ^B𝕍An[t)),a.s. on {ξ^𝕍An(t)=S𝕍A}.\frac{d\mu_{t-}}{d{\hat{\mu}}^{n}_{t-}}({\hat{\xi}}^{n}[t))=\ {\tilde{f}}_{1}({\hat{\xi}}^{n}_{A\cup\partial^{\mathbb{V}}A}[t))\ {\tilde{f}}_{2}({\hat{\xi}}^{n}_{B\cup\partial^{\mathbb{V}}A}[t)),\qquad\text{a.s. on }\{{\hat{\xi}}^{n}_{\partial^{\mathbb{V}}A}(t-)=S_{\partial^{\mathbb{V}}A}\}. (5.5)
Proof.

Fix n,A,Bn,\ A,\ B as in the statement of the lemma, and t(0,)t\in(0,\infty). Set n:=nt(A){\mathcal{E}}_{n}:={\mathcal{E}}_{n}^{t}(A), where the latter is defined in (5.4). For v𝕍nv\in\mathbb{V}_{n} and x𝒟t𝕍nx\in{\mathcal{D}}_{t-}^{\mathbb{V}_{n}} proper define

γv(xv,xw𝕍):=[0<tk(xv)<tqα(jk(xv),tk(xv),xv,xw𝕍)]ej=1,2(0,t)(qα(j,s,xv,xw𝕍)1)𝑑s,\gamma_{v}(x_{v},x_{{\partial^{\mathbb{V}}_{w}}}):=\left[\prod_{0<t_{k}(x_{v})<t}q_{\alpha}(j_{k}(x_{v}),t_{k}(x_{v}),x_{v},x_{{\partial^{\mathbb{V}}_{w}}})\right]e^{-\sum_{j=1,2}\int_{(0,t)}\left(q_{\alpha}(j,s,x_{v},x_{{\partial^{\mathbb{V}}_{w}}})-1\right)ds}, (5.6)

where {(tk(xv),v,jk(xv))}\left\{(t_{k}(x_{v}),v,j_{k}(x_{v}))\right\} are the jump characteristics of xvx_{v}. When x𝒟t𝕍nx\in{\mathcal{D}}^{\mathbb{V}_{n}}_{t-} is not proper, set γv(xv,xw𝕍):=0\gamma_{v}(x_{v},x_{{\partial^{\mathbb{V}}_{w}}}):=0. Also, for v𝕍v\in\mathbb{V}, y:[0,t)𝒳¯y:[0,t)\rightarrow\bar{\mathcal{X}}_{\star}, b𝒳¯b\in\bar{\mathcal{X}}_{\star} and z(𝒳¯)z\in(\bar{\mathcal{X}}_{\star})^{\infty}, define

ψv(y,b,z):={1𝟏{y(0),b=}if v=,y𝒟t,|{κ:zκ}|<,𝟏{y(0)}if v,y𝒟t,|{κ:zκ}|<,0otherwise.\psi_{v}(y,b,z):=\begin{cases}1-{\bm{1}}_{\{y(0)\neq\star,\ b=\star\}}&\text{if }v=\varnothing,\ y\in{\mathcal{D}}_{t-},\ |\left\{\kappa\in{\mathbb{N}}:\ z_{\kappa}\neq\star\right\}|<\infty,\\ {\bm{1}}_{\{y(0)\neq\star\}}&\text{if }v\neq\varnothing,\ y\in{\mathcal{D}}_{t-},\ |\left\{\kappa\in{\mathbb{N}}:\ z_{\kappa}\neq\star\right\}|<\infty,\\ 0&\text{otherwise.}\end{cases} (5.7)

By Proposition 5.3, the jump characteristics of ξ^𝕍nn{\hat{\xi}}^{n}_{\mathbb{V}_{n}} are almost surely well-defined. On the event that they exist, the jump characteristics of ξ^𝕍nn{\hat{\xi}}^{n}_{\mathbb{V}_{n}} are a disjoint union of those of ξ^vn{\hat{\xi}}^{n}_{v} for v𝕍nv\in\mathbb{V}_{n}. We can then rewrite (5.3) as

dμtdμ^tn(ξ^n[t)])=v𝕍nγv(ξ^w𝕍[t),ξ^w𝕍n[t))ψv(ξ^vn[t),ξ^πnn(0),ξ^Cvn(0))a.s.\frac{d\mu_{t-}}{d{\hat{\mu}}^{n}_{t-}}({\hat{\xi}}^{n}[t)])=\prod_{v\in{\mathbb{V}}_{n}}\gamma_{v}({\hat{\xi}}^{\partial^{\mathbb{V}}_{w}}[t),{\hat{\xi}}^{n}_{{\partial^{\mathbb{V}}_{w}}}[t))\ \psi_{v}({\hat{\xi}}^{n}_{v}[t),{\hat{\xi}}^{n}_{\pi_{n}}(0),{\hat{\xi}}^{n}_{C_{v}}(0))\ \text{a.s.}

Since A,𝕍AA,\ \partial^{\mathbb{V}}A and BB forms a partition of 𝕍\mathbb{V}, we can further decompose the right-hand side as

dμtdμ^tn(ξ^n[t))=F{A,𝕍A,B}(vF𝕍nγv(ξ^w𝕍n[t),ξ^w𝕍n[t))ψv(ξ^vn[t)],ξ^πvn(0),ξ^Cvn(0)))a.s.,\frac{d\mu_{t-}}{d{\hat{\mu}}^{n}_{t-}}({\hat{\xi}}^{n}[t))=\prod_{F\in\{A,\partial^{\mathbb{V}}A,B\}}\left(\prod_{v\in F\cap\mathbb{V}_{n}}\gamma_{v}({\hat{\xi}}^{n}_{{\partial^{\mathbb{V}}_{w}}}[t),{\hat{\xi}}^{n}_{{\partial^{\mathbb{V}}_{w}}}[t))\ \psi_{v}({\hat{\xi}}^{n}_{v}[t)],{\hat{\xi}}^{n}_{\pi_{v}}(0),{\hat{\xi}}^{n}_{C_{v}}(0))\right)\ \text{a.s.}, (5.8)

where for ease of notation in the sequel, we set π=\pi_{\varnothing}=\varnothing, which we can be done in (5.8) since ψ(y,b,z)\psi_{\varnothing}(y,b,z) does not depend on bb. The product in the inner bracket is a function of ξ^A𝕍An{\hat{\xi}}^{n}_{A\cup\partial^{\mathbb{V}}A} when F=AF=A, and a function of ξ^B𝕍An{\hat{\xi}}^{n}_{B\cup\partial^{\mathbb{V}}A} when F=BF=B. Thus, to prove (5.5) it suffices to show that for each w𝕍Aw\in\partial^{\mathbb{V}}A, there exist measurable functions f~1w:𝒟tA𝕍A[0,){\tilde{f}}^{w}_{1}:{\mathcal{D}}_{t-}^{A\cup\partial^{\mathbb{V}}A}\rightarrow[0,\infty) and f~2w:𝒟tB𝕍A[0,){\tilde{f}}^{w}_{2}:{\mathcal{D}}_{t-}^{B\cup\partial^{\mathbb{V}}A}\rightarrow[0,\infty) such that almost surely on n{\mathcal{E}}_{n}

γw(ξ^w𝕍n[t),ξ^w𝕍n[t))ψw(ξ^wn[t)],ξ^πwn(0),ξ^Cwn(0))=f~w1(ξ^A𝕍An[t))f~2w(ξ^B𝕍An[t)).\gamma_{w}({\hat{\xi}}^{n}_{{\partial^{\mathbb{V}}_{w}}}[t),{\hat{\xi}}^{n}_{{\partial^{\mathbb{V}}_{w}}}[t))\ \psi_{w}({\hat{\xi}}^{n}_{w}[t)],{\hat{\xi}}^{n}_{\pi_{w}}(0),{\hat{\xi}}^{n}_{C_{w}}(0))={\tilde{f}}^{w}_{1}({\hat{\xi}}^{n}_{A\cup\partial^{\mathbb{V}}A}[t))\ {\tilde{f}}_{2}^{w}({\hat{\xi}}^{n}_{B\cup\partial^{\mathbb{V}}A}[t)). (5.9)

By the monotonicity of the S(E)IR dynamics given in (4.2), on n{\mathcal{E}}_{n} the set of times {ti(ξ^wn)<t:w𝕍A}\{t_{i}({\hat{\xi}}^{n}_{w})<t\ :w\in\partial^{\mathbb{V}}A\} given by the jump characteristics of ξ^wn{\hat{\xi}}^{n}_{w} is empty. Hence, almost surely,

𝟏nti(ξ^wn)<tqα(ji(ξ^wn),ti(ξ^wn),ξ^wn,,ξ^w𝕍n)=𝟏n.{\bm{1}}_{{\mathcal{E}}_{n}}\prod_{t_{i}({\hat{\xi}}^{n}_{w})<t}q_{\alpha}(j_{i}({\hat{\xi}}^{n}_{w}),t_{i}({\hat{\xi}}^{n}_{w}),{\hat{\xi}}^{n}_{w},,{\hat{\xi}}^{n}_{{\partial^{\mathbb{V}}_{w}}})={\bm{1}}_{{\mathcal{E}}_{n}}. (5.10)

Recalling the identification of the states (S,E,I,R)(S,E,I,R) with (0,1,2,3)(0,1,2,3) and the definition qαq_{\alpha} in (4.3), for w𝕍Aw\in\partial^{\mathbb{V}}A and s(0,t)s\in(0,t), on the event n{\mathcal{E}}_{n} we have

qα(j,s,ξ^wn,ξ^w𝕍n)=βt(ξ^w𝕍n(s))(α𝟏{j=1}}+(1α)𝟏{j=2})=βt(w¯w𝟏{ξ^w¯n(s)=2})(α𝟏{j=1}}+(1α)𝟏{j=2}).\displaystyle\begin{split}q_{\alpha}(j,s,{\hat{\xi}}^{n}_{w},{\hat{\xi}}^{n}_{{\partial^{\mathbb{V}}_{w}}})&={\beta_{t}}{\mathcal{I}}({\hat{\xi}}^{n}_{{\partial^{\mathbb{V}}_{w}}}(s-))(\alpha{\bm{1}}_{\{j=1\}\}}+(1-\alpha){\bm{1}}_{\left\{j=2\right\}})\\ &={\beta_{t}}\left(\sum_{{\bar{w}}\sim w}{\bm{1}}_{\{{\hat{\xi}}^{n}_{{\bar{w}}}(s-)=2\}}\right)(\alpha{\bm{1}}_{\{j=1\}\}}+(1-\alpha){\bm{1}}_{\left\{j=2\right\}}).\end{split} (5.11)

For convenience of notation, set α(j):=α𝟏{j=1}+(1α)𝟏{j=2}\alpha(j):=\alpha{\bm{1}}_{\{j=1\}}+(1-\alpha){\bm{1}}_{\left\{j=2\right\}}. Then for w𝕍Aw\in\partial^{\mathbb{V}}A, using first (5.6) and (5.10), and then (5.11) and the fact that w𝕍AB\partial^{\mathbb{V}}_{w}\subset A\cup B,

𝟏nγw(ξ^wn,ξ^w𝕍n)=𝟏nexp(j=1,2α(j)(0,t)((ξ^w𝕍n(s))1)𝑑s)=𝟏nexp((0,t)((ξ^w𝕍An(s))+(ξ^w𝕍Bn(s))1)𝑑s)=𝟏nexp((0,t)((ξ^w𝕍An(s)))𝑑s)exp((0,t)((ξ^w𝕍Bn(s))1)𝑑s),\displaystyle\begin{split}{\bm{1}}_{{\mathcal{E}}_{n}}&\gamma_{w}({\hat{\xi}}^{n}_{w},{\hat{\xi}}^{n}_{{\partial^{\mathbb{V}}_{w}}})\\ &={\bm{1}}_{{\mathcal{E}}_{n}}\exp\left(-\sum_{j=1,2}\alpha(j)\int_{(0,t)}\left({\mathcal{I}}({\hat{\xi}}^{n}_{{\partial^{\mathbb{V}}_{w}}}(s-))-1\right)ds\right)\\ &={\bm{1}}_{{\mathcal{E}}_{n}}\exp\left(\int_{(0,t)}\left({\mathcal{I}}({\hat{\xi}}^{n}_{{\partial^{\mathbb{V}}_{w}}\cap A}(s-))+{\mathcal{I}}({\hat{\xi}}^{n}_{{\partial^{\mathbb{V}}_{w}}\cap B}(s-))-1\right)ds\right)\\ &={\bm{1}}_{{\mathcal{E}}_{n}}\exp\left(-\int_{(0,t)}\left({\mathcal{I}}({\hat{\xi}}^{n}_{{\partial^{\mathbb{V}}_{w}}\cap A}(s-))\right)ds\right)\exp\left(-\int_{(0,t)}\left({\mathcal{I}}({\hat{\xi}}^{n}_{{\partial^{\mathbb{V}}_{w}}\cap B}(s-))-1\right)ds\right),\end{split} (5.12)

which shows that each γw\gamma_{w} term in (5.9) admits the desired factorization. It only remains to show that the same holds for the ψw\psi_{w} term in (5.9). To this end, note that for w𝕍A{}w\in\partial^{\mathbb{V}}A\setminus\{\varnothing\}, by (5.7),

ψw(ξ^wn[t)],ξ^πwn(0),ξ^Cwn(0))=𝟏{ξ^wn[t)𝒟t}𝟏{|{vCw:ξ^wn(0)}|<}(1𝟏{ξ^wn(0),ξ^πwn(0)=})=𝟏{ξ^wn[t)𝒟t}𝟏{|(A𝕍A){vCw:ξ^vn(0)}|<}𝟏{|(B𝕍A){vCw:ξ^vn(0)}|<}𝟏{ξ^wn(0),ξ^πwn(0)=)}c=ψ~w(1)(ξ^wn[t),ξ^πwn[t))ψ~w(2)(ξ^A𝕍An[t))ψ~w(3)(ξ^B𝕍An[t)),\displaystyle\begin{split}&\psi_{w}({\hat{\xi}}^{n}_{w}[t)],{\hat{\xi}}^{n}_{\pi_{w}}(0),{\hat{\xi}}^{n}_{C_{w}}(0))\\ &={\bm{1}}_{\{{\hat{\xi}}^{n}_{w}[t)\in{\mathcal{D}}_{t-}\}}{\bm{1}}_{\{|\left\{v\in C_{w}:\ {\hat{\xi}}^{n}_{w}(0)\neq\star\right\}|<\infty\}}(1-{\bm{1}}_{\{{\hat{\xi}}^{n}_{w}(0)\neq\star,\ {\hat{\xi}}^{n}_{\pi_{w}}(0)=\star\}})\\ &={\bm{1}}_{\{{\hat{\xi}}^{n}_{w}[t)\in{\mathcal{D}}_{t-}\}}{\bm{1}}_{\{|(A\cup\partial^{\mathbb{V}}A)\cap\left\{v\in C_{w}:\ {\hat{\xi}}^{n}_{v}(0)\neq\star\right\}|<\infty\}}{\bm{1}}_{\{|(B\cup\partial^{\mathbb{V}}A)\cap\left\{v\in C_{w}:\ {\hat{\xi}}^{n}_{v}(0)\neq\star\right\}|<\infty\}}{\bm{1}}_{\{{\hat{\xi}}^{n}_{w}(0)\neq\star,{\hat{\xi}}^{n}_{\pi_{w}}(0)=\star)\}^{c}}\\ &={\tilde{\psi}}^{(1)}_{w}({\hat{\xi}}^{n}_{w}[t),{\hat{\xi}}^{n}_{\pi_{w}}[t))\ {\tilde{\psi}}^{(2)}_{w}({\hat{\xi}}^{n}_{A\cup\partial^{\mathbb{V}}A}[t))\ {\tilde{\psi}}^{(3)}_{w}({\hat{\xi}}^{n}_{B\cup\partial^{\mathbb{V}}A}[t)),\end{split} (5.13)

where ψ~w(1){\tilde{\psi}}^{(1)}_{w} is the product of the first and last term in the penultimate line of the display, and ψ~w(i){\tilde{\psi}}^{(i)}_{w}, i=2,3i=2,3, is the ii-th term of the penultimate line of the display. Since w𝕍Aw\in\partial^{\mathbb{V}}A and 𝕍\mathbb{V} is a tree, either {w,πw}B𝕍A\{w,\pi_{w}\}\subset B\cup\partial^{\mathbb{V}}A or {w,πw}A𝕍A\{w,\pi_{w}\}\subset A\cup\partial^{\mathbb{V}}A. Hence, the last line of (5.13) factors as desired. Similarly, if 𝕍A\varnothing\in\partial^{\mathbb{V}}A, by (5.7) we have

ψ(ξ^n[t)],ξ^πn(0),ξ^Cn(0))=𝟏{ξ^n[t)𝒟t}𝟏{|{vC:ξ^n(0)}|<}𝟏{ξ^n(0)}=𝟏{ξ^n[t)𝒟t}𝟏{|(A𝕍A){vC:ξ^vn(0)}|<}𝟏{|(B𝕍A){vC:ξ^vn(0)}|<}𝟏{ξ^n(0)}=ψ~(4)(ξ^n[t))ψ~(5)(ξ^A𝕍An[t))ψ~(6)(ξ^B𝕍An[t)),\displaystyle\begin{split}&\psi_{\varnothing}({\hat{\xi}}^{n}_{\varnothing}[t)],{\hat{\xi}}^{n}_{\pi_{\varnothing}}(0),{\hat{\xi}}^{n}_{C_{\varnothing}}(0))\\ &={\bm{1}}_{\{{\hat{\xi}}^{n}_{\varnothing}[t)\in{\mathcal{D}}_{t-}\}}{\bm{1}}_{\{|\left\{v\in C_{\varnothing}:\ {\hat{\xi}}^{n}_{\varnothing}(0)\neq\star\right\}|<\infty\}}{\bm{1}}_{\{{\hat{\xi}}^{n}_{\partial_{\varnothing}}(0)\neq\star\}}\\ &={\bm{1}}_{\{{\hat{\xi}}^{n}_{\varnothing}[t)\in{\mathcal{D}}_{t-}\}}{\bm{1}}_{\{|(A\cup\partial^{\mathbb{V}}A)\cap\left\{v\in C_{\varnothing}:\ {\hat{\xi}}^{n}_{v}(0)\neq\star\right\}|<\infty\}}{\bm{1}}_{\{|(B\cup\partial^{\mathbb{V}}A)\cap\left\{v\in C_{\varnothing}:\ {\hat{\xi}}^{n}_{v}(0)\neq\star\right\}|<\infty\}}{\bm{1}}_{\{{\hat{\xi}}^{n}_{\varnothing}(0)\neq\star\}}\\ &={\tilde{\psi}}^{(4)}({\hat{\xi}}^{n}_{\varnothing}[t))\ {\tilde{\psi}}^{(5)}({\hat{\xi}}^{n}_{A\cup\partial^{\mathbb{V}}A}[t))\ {\tilde{\psi}}^{(6)}({\hat{\xi}}^{n}_{B\cup\partial^{\mathbb{V}}A}[t)),\end{split} (5.14)

where ψ~(4){\tilde{\psi}}^{(4)} is the product of the first and last term in the penultimate line of the display, and ψ~(i){\tilde{\psi}}^{(i)} with i=5,6i=5,6 is the (i3)(i-3)-th term of the penultimate line of the display. Together (5.12), (5.14) and (5.13) prove (5.9) and, hence, 𝟏ndμt/dμ^tn{\bm{1}}_{{\mathcal{E}}_{n}}d\mu_{t-}/d{\hat{\mu}}_{t-}^{n} admits the factorization stated in (5.5). ∎

We conclude this section by proving Proposition 4.7.

Proof of Proposition 4.7.

Throughout the proof we fix α[0,1]\alpha\in[0,1], θ𝒫(0)\theta\in{\mathcal{P}}({\mathbb{N}}_{0}) and 𝒯=UGW(θ{\mathcal{T}}=\text{UGW}(\theta), and we omit the dependence of ξ𝒯,α\xi^{{\mathcal{T}},\alpha} on them. Let {A,B,𝕍A}\{A,B,\partial^{\mathbb{V}}A\} be a partition of 𝕍\mathbb{V} with 𝕍A\partial^{\mathbb{V}}A being finite. Pick nn\in{\mathbb{N}} such that 𝕍A𝕍n1\partial^{\mathbb{V}}A\subset{\mathbb{V}}_{n-1}. We define 𝒜:𝒟𝕍A{0,1}{\mathcal{A}}:{\mathcal{D}}_{\star}^{\partial^{\mathbb{V}}A}\rightarrow\{0,1\} by 𝒜(x)=𝟏{xv(t)=S,v𝕍A}{\mathcal{A}}(x)={\bm{1}}_{\left\{x_{v}(t-)=S,\ v\in\partial^{\mathbb{V}}A\right\}}. We observe that 𝒜(ξ^𝕍An)=𝟏nt(A){\mathcal{A}}({\hat{\xi}}^{n}_{\partial^{\mathbb{V}}A})={\bm{1}}_{{\mathcal{E}}_{n}^{t}(A)}, where the latter is defined in (5.4). Let WW be a bounded, σ(ξA[t))\sigma(\xi_{A}[t))-measurable random variable. Adopting the convention 0/0=00/0=0, and using Lemma 5.4 and Bayes’s theorem in the first line, and the property (5.2) in the last line, we have

𝔼μ[W𝒜(ξ𝕍A)|ξ𝕍A[t),ξB[t)]=𝔼μ^n[W𝒜(ξ^𝕍An)dμtdμ^tnξ^n[t)|ξ^𝕍An[t),ξ^Bn[t)]𝔼μ^n[dμtdμ^tnξ^n[t)|ξ^𝕍An[t),ξ^Bn[t)]=𝟏nt(A)𝔼μ^n[Wf~1(ξ^A𝕍An[t))f~2(ξ^B𝕍An[t))|ξ^𝕍An[t),ξ^Bn[t)]𝔼μ^n[f~1(ξ^A𝕍An[t))f~2(ξ^B𝕍An[t))|ξ^𝕍An[t),ξ^Bn[t)]=𝟏nt(A)f~2(ξ^B𝕍An[t))𝔼μ^n[Wf~1(ξ^A𝕍An[t))|ξ^𝕍An[t),ξ^Bn[t)]f~2(ξ^B𝕍An[t))𝔼μ^n[f~1(ξ^A𝕍An[t))|ξ^𝕍An[t),ξ^Bn[t)]=𝟏nt(A)𝔼μ^n[Wf~1(ξ^A𝕍An[t))|ξ^𝕍An[t)]𝔼μ^n[f~1(ξ^A𝕍An[t))|ξ^𝕍An[t)].\displaystyle\begin{split}{\mathbb{E}}_{\mu}[W{\mathcal{A}}(\xi_{\partial^{\mathbb{V}}A})\ |\ \xi_{\partial^{\mathbb{V}}A}[t),\ \xi_{B}[t)]=&\frac{{\mathbb{E}}_{{\hat{\mu}}^{n}}[W{\mathcal{A}}({\hat{\xi}}^{n}_{\partial^{\mathbb{V}}A})\frac{d\mu_{t-}}{d{\hat{\mu}}^{n}_{t-}}{\hat{\xi}}^{n}[t)\ |\ {\hat{\xi}}^{n}_{\partial^{\mathbb{V}}A}[t),{\hat{\xi}}^{n}_{B}[t)]}{{\mathbb{E}}_{{\hat{\mu}}^{n}}[\frac{d\mu_{t-}}{d{\hat{\mu}}^{n}_{t-}}{\hat{\xi}}^{n}[t)|\ {\hat{\xi}}^{n}_{\partial^{\mathbb{V}}A}[t),{\hat{\xi}}^{n}_{B}[t)]}\\ =&{\bm{1}}_{{\mathcal{E}}_{n}^{t}(A)}\frac{{\mathbb{E}}_{{\hat{\mu}}^{n}}[W{\tilde{f}}_{1}({\hat{\xi}}^{n}_{A\cup\partial^{\mathbb{V}}A}[t)){\tilde{f}}_{2}({\hat{\xi}}^{n}_{B\cup\partial^{\mathbb{V}}A}[t))\ |\ {\hat{\xi}}^{n}_{\partial^{\mathbb{V}}A}[t),{\hat{\xi}}^{n}_{B}[t)]}{{\mathbb{E}}_{{\hat{\mu}}^{n}}[{\tilde{f}}_{1}({\hat{\xi}}^{n}_{A\cup\partial^{\mathbb{V}}A}[t)){\tilde{f}}_{2}({\hat{\xi}}^{n}_{B\cup\partial^{\mathbb{V}}A}[t))\ |\ {\hat{\xi}}^{n}_{\partial^{\mathbb{V}}A}[t),{\hat{\xi}}^{n}_{B}[t)]}\\ =&{\bm{1}}_{{\mathcal{E}}_{n}^{t}(A)}\frac{{\tilde{f}}_{2}({\hat{\xi}}^{n}_{B\cup\partial^{\mathbb{V}}A}[t)){\mathbb{E}}_{{\hat{\mu}}^{n}}[W{\tilde{f}}_{1}({\hat{\xi}}^{n}_{A\cup\partial^{\mathbb{V}}A}[t))\ |\ {\hat{\xi}}^{n}_{\partial^{\mathbb{V}}A}[t),{\hat{\xi}}^{n}_{B}[t)]}{{\tilde{f}}_{2}({\hat{\xi}}^{n}_{B\cup\partial^{\mathbb{V}}A}[t)){\mathbb{E}}_{{\hat{\mu}}^{n}}[{\tilde{f}}_{1}({\hat{\xi}}^{n}_{A\cup\partial^{\mathbb{V}}A}[t))\ |\ {\hat{\xi}}^{n}_{\partial^{\mathbb{V}}A}[t),{\hat{\xi}}^{n}_{B}[t)]}\\ =&{\bm{1}}_{{\mathcal{E}}_{n}^{t}(A)}\frac{{\mathbb{E}}_{{\hat{\mu}}^{n}}[W{\tilde{f}}_{1}({\hat{\xi}}^{n}_{A\cup\partial^{\mathbb{V}}A}[t))\ |\ {\hat{\xi}}^{n}_{\partial^{\mathbb{V}}A}[t)]}{{\mathbb{E}}_{{\hat{\mu}}^{n}}[{\tilde{f}}_{1}({\hat{\xi}}^{n}_{A\cup\partial^{\mathbb{V}}A}[t))\ |\ {\hat{\xi}}^{n}_{\partial^{\mathbb{V}}A}[t)]}.\end{split}

The last quotient is ξ^𝕍An[t){\hat{\xi}}^{n}_{\partial^{\mathbb{V}}A}[t)-measurable. As this holds for every bounded ξA[t)\xi_{A}[t)-measurable random variable W,W, we conclude that

𝔼μ[W𝒜(ξ𝕍A)|ξ𝕍A[t),ξB[t)]=𝔼μ[W𝒜(ξ𝕍A)|ξ𝕍A[t)].{\mathbb{E}}_{\mu}[W{\mathcal{A}}(\xi_{\partial^{\mathbb{V}}A})\ |\ \xi_{\partial^{\mathbb{V}}A}[t),\ \xi_{B}[t)]={\mathbb{E}}_{\mu}[W{\mathcal{A}}(\xi_{\partial^{\mathbb{V}}A})\ |\ \xi_{\partial^{\mathbb{V}}A}[t)].

Since 𝒜(ξ𝕍A)=𝟏{ξv(t)=Sv𝕍A}{\mathcal{A}}(\xi_{\partial^{\mathbb{V}}A})={\bm{1}}_{\{\xi_{v}(t-)=S\ \forall v\in\partial^{\mathbb{V}}A\}}, if follows that

ξA[t)ξB[t)|ξ𝕍A(t)=S𝕍A,\xi_{A}[t)\perp\xi_{B}[t)\ |\ \xi_{\partial^{\mathbb{V}}A}(t-)=S_{\partial^{\mathbb{V}}A}, (5.15)

which proves the first assertion of the proposition.

Next, let v𝕍v\in\mathbb{V} and κ0\kappa\in{\mathbb{N}}_{0}. Set d~v:=dv𝟏{v}=|Cv𝒯|\tilde{d}_{v}:=d_{v}-{\bm{1}}_{\{v\neq\varnothing\}}=|C_{v}\cap{\mathcal{T}}|. The event {d~v=k}\{\tilde{d}_{v}=k\} is clearly the same as the event {dv=κ+𝟏{v}}\left\{d_{v}=\kappa+{\bm{1}}_{\{v\neq\varnothing}\}\right\}. In the sequel, we condition on the event d~v=κ\tilde{d}_{v}=\kappa. If κ=0\kappa=0, the set {ξvi(t)}i=1,,d~v\{\xi_{vi}(t)\}_{i=1,...,\tilde{d}_{v}} of children of vv in 𝒯{\mathcal{T}} is empty, and if κ=1\kappa=1, the set of children of vv in 𝒯{\mathcal{T}} is a singleton. In either case, the stated conditional independence holds trivially. Suppose that κ2\kappa\geq 2 and for i=1,,κi=1,...,\kappa, we fix xi𝒳¯x_{i}\in\bar{\mathcal{X}}. For w𝕍w\in\mathbb{V}, let 𝕋w:={wu}u𝕍{\mathbb{T}}_{w}:=\{wu\}_{u\in\mathbb{V}} denote the subtree of 𝕍\mathbb{V} rooted at ww. By the definition of the jump rate qαq_{\alpha} in (4.3), and the SDE characterization (4.4) from Lemma 4.2, d~v=κ\tilde{d}_{v}=\kappa if and only ξ(0)vm\xi(0)_{vm}\neq\star for all mm\in{\mathbb{N}} with mκm\leq\kappa and ξmk(0)=\xi_{mk}(0)=\star for all kk\in{\mathbb{N}} with k>κk>\kappa. Using first this fact, and then applying (5.15) with A=i=1κ𝕋viA=\cup_{i=1}^{\kappa}{\mathbb{T}}_{vi}, 𝕍A={v}\partial^{\mathbb{V}}A=\{v\} and B={v:(κ,))B=\{v\ell\ :\ \ell\in{\mathbb{N}}\cap(\kappa,\infty)), we have

(ξv(t)=x for 1κ|ξv(t)=S,d~v=κ)=(ξv(t)=x for 1κ|ξv(t)=S,ξvm(0)for 1mκ,ξvk(0)= for k>κ)=(ξv(t)=x for 1κ|ξv(t)=S,ξvm(0)for 1mκ).\displaystyle\begin{split}&{\mathbb{P}}(\xi_{v\ell}(t)=x_{\ell}\text{ for }1\leq\ell\leq\kappa\ |\ \xi_{v}(t)=S,\ \tilde{d}_{v}=\kappa)\\ &={\mathbb{P}}(\xi_{v\ell}(t)=x_{\ell}\text{ for }1\leq\ell\leq\kappa\ |\ \xi_{v}(t)=S,\ \xi_{vm}(0)\neq\star\ \text{for }1\leq m\leq\kappa,\ \xi_{vk}(0)=\star\text{ for }k>\kappa)\\ &={\mathbb{P}}(\xi_{v\ell}(t)=x_{\ell}\text{ for }1\leq\ell\leq\kappa\ |\ \xi_{v}(t)=S,\ \xi_{vm}(0)\neq\star\ \text{for }1\leq m\leq\kappa).\end{split} (5.16)

For each =1,,κ\ell=1,...,\kappa, another application of (5.15), with A=𝕋vA={\mathbb{T}}_{v\ell}, 𝕍A={v}\partial^{\mathbb{V}}A=\{v\} and B={vm:m{})B=\{vm\ :\ m\in{\mathbb{N}}\setminus\{\ell\}) yields

ξv[t){ξvm[t)}m{}|ξv(t)=S.\xi_{v\ell}[t)\perp\{\xi_{vm}[t)\}_{m\in{\mathbb{N}}\setminus\{\ell\}}\ |\xi_{v}(t-)=S.

It follows that for [1,κ]\ell\in{\mathbb{N}}\cap[1,\kappa] and M{m:mκ}{}M\subset\{m\in{\mathbb{N}}\ :\ m\leq\kappa\}\setminus\{\ell\},

(ξv(t)=x|ξv(t)=S,ξvi(t)for 1iκ,ξvm=xmmM)=(ξv(t)=x|ξv(t)=S,ξv(t))=(ξv(t)=x|ξv(t)=S,v𝒯).\displaystyle\begin{split}&{\mathbb{P}}(\xi_{v\ell}(t)=x_{\ell}|\ \xi_{v}(t)=S,\ \xi_{vi}(t)\neq\star\ \text{for }1\leq i\leq\kappa,\ \xi_{vm}=x_{m}\ \forall m\in M)\\ &={\mathbb{P}}(\xi_{v\ell}(t)=x_{\ell}|\ \xi_{v}(t)=S,\ \xi_{v\ell}(t)\neq\star)\\ &={\mathbb{P}}(\xi_{v\ell}(t)=x_{\ell}|\ \xi_{v}(t)=S,\ {v\ell}\in{\mathcal{T}}).\end{split} (5.17)

By iteratively applying (5.17), we obtain

(ξv(t)=xfor 1κ|ξv(t)=S,ξvi(t)for 1iκ)==1κ(ξv(t)=x|ξv(t)=S,v𝒯),\displaystyle\begin{split}&{\mathbb{P}}(\xi_{v\ell}(t)=x_{\ell}\ \ \text{for }1\leq\ell\leq\kappa\ |\ \xi_{v}(t)=S,\ \xi_{vi}(t)\neq\star\ \text{for }1\leq i\leq\kappa)\\ &=\prod_{\ell=1}^{\kappa}{\mathbb{P}}(\xi_{v\ell}(t)=x_{\ell}\ |\xi_{v}(t)=S,\ v\ell\in{\mathcal{T}}),\end{split}

which along with (5.16) establishes the second assertion of the theorem and concludes the proof. ∎

We conclude this section by using Proposition 4.7 and properties of the SDE (4.4) to derive Proposition 4.8.

Proof of Proposition 4.8.

Fix v~𝕍{}{\tilde{v}}\in\mathbb{V}\setminus\{\varnothing\} and, on the event v~𝒯{\tilde{v}}\in{\mathcal{T}}, let 𝒯~{\tilde{{\mathcal{T}}}} be the subtree of 𝒯{\mathcal{T}} rooted at v~{\tilde{v}}, i.e., 𝒯~:=𝒯𝕋v~{\tilde{{\mathcal{T}}}}:={\mathcal{T}}\cap{\mathbb{T}}_{{\tilde{v}}} where 𝕋v~:={v~w:w𝕍}{\mathbb{T}}_{{\tilde{v}}}:=\{{\tilde{v}}w\ :\ w\in\mathbb{V}\}. By Assumption D, 𝒯v~{\mathcal{T}}_{{\tilde{v}}} is a Galton-Watson tree with offspring distribution θ\theta on the event {v~𝒯}\{{\tilde{v}}\in{\mathcal{T}}\}. Recall that ξ𝒯,α\xi^{{\mathcal{T}},\alpha} satisfies the SDE (4.4), and define the modified process ξ~{\tilde{\xi}} on 𝒳¯𝕍\bar{\mathcal{X}}_{\star}^{\mathbb{V}} by t[0,)t\in[0,\infty),

ξ~v(t)={ξv𝒯,α(t)v𝕋v~v𝕍𝕋v~.{\tilde{\xi}}_{v}(t)=\begin{cases}\xi^{{\mathcal{T}},\alpha}_{v}(t)&v\in{\mathbb{T}}_{{\tilde{v}}}\\ \star&v\in\mathbb{V}\setminus{\mathbb{T}}_{{\tilde{v}}}.\end{cases} (5.18)

Fix t[0,)t\in[0,\infty), and let t:={ξv~𝒯,α(t)=S}{\mathcal{E}}^{t}:=\{\xi^{{\mathcal{T}},\alpha}_{\tilde{v}}(t-)=S\}. By (4.4) and Assumption D, ξv~𝒯,α(t)=S\xi^{{\mathcal{T}},\alpha}_{{\tilde{v}}}(t-)=S implies ξv~𝒯,α\xi^{{\mathcal{T}},\alpha}_{{\tilde{v}}}\neq\star and hence, that

t=~t:={ξ~v~(t)=S,v~𝒯}={ξ~v~(s)=S:s[0,t)},{\mathcal{E}}^{t}=\tilde{{\mathcal{E}}}^{t}:=\{{\tilde{\xi}}_{{\tilde{v}}}(t-)=S,\ {\tilde{v}}\in{\mathcal{T}}\}=\{{\tilde{\xi}}_{{\tilde{v}}}(s)=S\ :\ s\in[0,t)\}, (5.19)

where the second equality follows from the monotonicity property of the S(E)IR process, see (4.2). Applying Proposition 4.7 with A=𝕋v~{v~}A={\mathbb{T}}_{{\tilde{v}}}\setminus\{{\tilde{v}}\}, 𝕍A={v~}\partial^{\mathbb{V}}A=\{{\tilde{v}}\} and B=𝕍𝕋v~{\mathbb{V}\setminus{\mathbb{T}}_{{\tilde{v}}}}, it follows that ξ𝕋v~{v~}𝒯,α[t)\xi^{{\mathcal{T}},\alpha}_{{\mathbb{T}}_{{\tilde{v}}}\setminus\{{\tilde{v}}\}}[t) is conditionally independent of ξ𝕍𝕋v~𝒯,α[t)\xi^{{\mathcal{T}},\alpha}_{\mathbb{V}\setminus{\mathbb{T}}_{{\tilde{v}}}}[t) given t{\mathcal{E}}^{t}. Thus, by (5.18) and (5.19),

(ξ𝕋v~{v~}𝒯,α[t)|t)=(ξ~𝕍𝕋v~𝒯,α[t)|~t).{\mathcal{L}}(\xi^{{\mathcal{T}},\alpha}_{{\mathbb{T}}_{{\tilde{v}}}\setminus\{{\tilde{v}}\}}[t)\ |{\mathcal{E}}^{t})={\mathcal{L}}({\tilde{\xi}}^{{\mathcal{T}},\alpha}_{\mathbb{V}\setminus{\mathbb{T}}_{{\tilde{v}}}}[t)\ |\tilde{{\mathcal{E}}}^{t}). (5.20)

Next, fix mm\in{\mathbb{N}}, and define a map ϕ~m:𝕋v~𝕍\tilde{\phi}_{m}:{\mathbb{T}}_{{\tilde{v}}}\rightarrow\mathbb{V} given by ϕ~m(v~)=\tilde{\phi}_{m}({\tilde{v}})=\varnothing and for v𝕍v\in\mathbb{V}, ϕ~m(v~(mv))=1v\tilde{\phi}_{m}({\tilde{v}}(mv))=1v, ϕ~m(v~(1v))=mv\tilde{\phi}_{m}({\tilde{v}}(1v))=mv and ϕ~m(v~(v))=v\tilde{\phi}_{m}({\tilde{v}}(\ell v))=\ell v for all {1,m}\ell\in{\mathbb{N}}\setminus\{1,m\}, recalling that vwvw represent concatenation of v,w𝕍v,w\in\mathbb{V}. Then ϕ~m\tilde{\phi}_{m} defines an isomorphism of the rooted graphs (𝕋v~,v~)({\mathbb{T}}_{{\tilde{v}}},{\tilde{v}}) and (𝕍,)(\mathbb{V},\varnothing). It follows from (5.18) and the form of the SDE (4.4) that

(ξ~𝕋v~[t)|~t)=(ξ𝕍𝒯,α[t)|ξ𝒯,α(t)=S){\mathcal{L}}({\tilde{\xi}}_{{\mathbb{T}}_{{\tilde{v}}}}[t)\ |\ \tilde{{\mathcal{E}}}^{t})={\mathcal{L}}(\xi^{{\mathcal{T}},\alpha}_{\mathbb{V}}[t)\ |\ \xi^{{\mathcal{T}},\alpha}_{\varnothing}(t-)=S)

and in particular,

(ξ~v~m[t)]|ξ~v~(t)=S,v~m𝒯~)=(ξ1𝒯,α[t)]|ξ~(t)=S,1𝒯),{\mathcal{L}}({\tilde{\xi}}_{{\tilde{v}}m}[t)]\ |{\tilde{\xi}}_{{\tilde{v}}}(t-)=S,{\tilde{v}}m\in{\tilde{{\mathcal{T}}}})={\mathcal{L}}(\xi^{{\mathcal{T}},\alpha}_{1}[t)]\ |{\tilde{\xi}}_{\varnothing}(t-)=S,1\in{\mathcal{T}}), (5.21)

which follows from the independence of 𝒯{\mathcal{T}} from the driving Poisson processes in the SDE (4.4) for ξ𝒯,α\xi^{{\mathcal{T}},\alpha} and the fact that v~m𝒯{\tilde{v}}m\in{\mathcal{T}} implies v~𝒯{\tilde{v}}\in{\mathcal{T}}. By the well-posedness of the SDE (4.4) established in Lemma 4.2, it follows that the left-hand side of (5.21) does not depend on the choice of v~𝕍{\tilde{v}}\in\mathbb{V} and mm\in{\mathbb{N}}, thus proving the proposition. ∎

Appendix A Proofs of intermediate SEIR dynamics Results

In this section, we prove Theorem 4.14 and Theorem 4.13, thus completing the proof of Theorem 2.10. Throughout, θ𝒫(0)\theta\in{\mathcal{P}}({\mathbb{N}}_{0}), θ^{\hat{\theta}} is the size-biased version of θ\theta, as defined in (2.2), and 𝒯{\mathcal{T}} is a UGW(θ\theta) tree. We assume that θ\theta has finite third moment and, as everywhere else in the paper, we assume that the rates β,λ,ρ:[0,)(0,)\beta,\ \lambda,\ \rho:[0,\infty)\rightarrow(0,\infty) satisfy Assumption C. We assume that Assumption D hold.

We start by proving the ODE characterization of gS,g_{S}, gEg_{E} and gIg_{I}.

Proof of Theorem 4.13.

Throughout the proof in order to simplify notation we write X¯{\bar{X}} in lieu of X¯𝒯=ξ𝒯,1{\bar{X}}^{\mathcal{T}}=\xi^{{\mathcal{T}},1}, the SEIR process on 𝒯{\mathcal{T}}, and qq in lieu of q1q_{1}, the rate function defined in (4.3). By Assumption D, gS(0)=s0g_{S}(0)=s_{0}, gE(0)=e0g_{E}(0)=e_{0} and gI(0)=i0g_{I}(0)=i_{0}. Clearly, GI(0)=0G_{I}(0)=0. Therefore, the initial conditions (2.12) hold. By the fundamental theorem of calculus, G˙I(t)=βtgI(t)\dot{G}_{I}(t)=\beta_{t}g_{I}(t), which is the fourth equation in (2.11).

We now turn to the derivation of the evolution of gI,gEg_{I},\ g_{E} and gSg_{S}. This requires keeping track of two states simultaneously since gI(t)g_{I}(t), gE(t)g_{E}(t) and gS(t)g_{S}(t) are conditional probabilities associated with the joint law of X¯1(t){{\bar{X}}}_{1}(t) and X¯(t){{\bar{X}}}_{\varnothing}(t). To start, we apply Proposition 4.6 with α=1\alpha=1 and U={,1}U=\{\varnothing,1\} to conclude that X¯,1{{\bar{X}}}_{\varnothing,1} has the same law as the jump process on the state space 𝒳¯×𝒳¯\bar{\mathcal{X}}_{\star}\times\bar{\mathcal{X}}_{\star} with jump rates q^v(t,x)=q^v,1θ,1[{,1}](t,x)\hat{q}_{v}(t,x)=\hat{q}_{v,1}^{\theta,1}[\{\varnothing,1\}](t,x), v{,1}v\in\{\varnothing,1\}, x𝒟([0,),𝒳¯2)x\in{\mathcal{D}}([0,\infty),\bar{\mathcal{X}}_{\star}^{2}), which satisfy, for every t0t\geq 0, almost surely

q^v(t,X¯,1)=𝔼[q(1,t,X¯v,X¯v𝕍)|X¯,1[t)],v{,1}.\hat{q}_{v}(t,{\bar{X}}_{\varnothing,1})={\mathbb{E}}[q(1,t,{\bar{X}}_{v},{\bar{X}}_{\partial_{v}^{\mathbb{V}}})|{\bar{X}}_{\varnothing,1}[t)],\quad v\in\{\varnothing,1\}. (A.1)

Next, we use the specific form of qq, as defined in (4.3) and Propositions 4.7 and 4.8 to obtain a more explicit description of q^v\hat{q}_{v}, v{,1}v\in\{\varnothing,1\}. Since the probabilities ga(t)g_{a}(t), a{S,E,I}a\in\{S,\ E,\ I\} are conditioned on X¯(t)=S{\bar{X}}_{\varnothing}(t-)=S and X¯1(0){\bar{X}}_{1}(0)\neq\star (and using the fact that an individual that is in state RR remains in that state for all subsequent times), we only need to consider the jumps q^v(t,X¯,1)\hat{q}_{v}(t,{\bar{X}}_{\varnothing,1}), v{,1}v\in\{\varnothing,1\} on the events {X¯,1(t)=(S,S)}\{{\bar{X}}_{\varnothing,1}(t-)=(S,S)\}, {X¯,1(t)=(S,E)}\{{\bar{X}}_{\varnothing,1}(t-)=(S,E)\} and {X¯,1(t)=(S,I)}\{{\bar{X}}_{\varnothing,1}(t-)=(S,I)\}.

For v,w{,1}v,w\in\left\{\varnothing,1\right\} with vwv\neq w, define B¯v(t):=βt𝔼[(X¯v{w}(t))|X¯v(t)=S]{\bar{B}}_{v}(t):=\beta_{t}{\mathbb{E}}[{\mathcal{I}}({\bar{X}}_{\partial_{v}\setminus\left\{w\right\}}(t-))|{\bar{X}}_{v}(t-)=S]. By the definition of the SEIR jump rates q=q1q=q_{1} in (4.3), BvB_{v} is the conditional cumulative rate at which the neighbors of vv other that ww infect the individual at vv at time tt. By Proposition 4.7,

B¯v(t)=βt𝔼[(X¯v{w})|X¯v(t)=S,X¯w(t)].{\bar{B}}_{v}(t)=\beta_{t}{\mathbb{E}}[{\mathcal{I}}({\bar{X}}_{\partial_{v}\setminus\left\{w\right\}})|{\bar{X}}_{v}(t)=S,{\bar{X}}_{w}(t)]. (A.2)

Using (4.3), (A.1) and Proposition 4.6, and proceeding similarly as in the proof of Theorem 4.10, we can treat X¯,1{\bar{X}}_{\varnothing,1} as a two particle jump process driven by Poisson noises with intensity measure equal to Lebesgue measure, whose jumps and jump rates from the states (S,S)(S,S) , (S,E)(S,E) and (S,I)(S,I) can be summarized as follows.

Jump: Rate at time tt:
(S,S)\displaystyle(S,S) (S,E)\displaystyle\rightarrow(S,E) B¯1(t)\displaystyle{\bar{B}}_{1}(t)
(S,S)\displaystyle(S,S) (E,S)\displaystyle\rightarrow(E,S) B¯(t)\displaystyle{\bar{B}}_{\varnothing}(t)
(S,E)\displaystyle(S,E) (E,E)\displaystyle\rightarrow(E,E) B¯(t)\displaystyle{\bar{B}}_{\varnothing}(t)
(S,E)\displaystyle(S,E) (S,I)\displaystyle\rightarrow(S,I) λt\displaystyle{\lambda_{t}}
(S,I)\displaystyle(S,I) (E,I)\displaystyle\rightarrow(E,I) βt+B¯(t)\displaystyle{\beta_{t}}+{\bar{B}}_{\varnothing}(t)
(S,I)\displaystyle(S,I) (S,R)\displaystyle\rightarrow(S,R) ρt,\displaystyle{\rho_{t}},

with all other rates being equal to zero. Next we fix h>0h>0 and t0t\geq 0 and obtain expressions for gI(t+h),gE(t+h)g_{I}(t+h),\ g_{E}(t+h), and gS(t+h)g_{S}(t+h) in terms of gI(t)g_{I}(t), gE(t)g_{E}(t), gS(t)g_{S}(t), hh, βt,{\beta_{t}}, ρt{\rho_{t}}, and θ^{\hat{\theta}}. We first consider gSg_{S}, defined in (4.41). Using monotonicity of the SEIR dynamics, we can write

gS(t+h)=(X¯1(t+h)=S,X¯1(t)=S|X¯(t+h)=S,X¯(t)=S, 1𝒯).g_{S}(t+h)={\mathbb{P}}({{\bar{X}}}_{1}(t+h)=S,\ {{\bar{X}}}_{1}(t)=S\ |\ {{\bar{X}}}_{\varnothing}(t+h)=S,\ {{\bar{X}}}_{\varnothing}(t)=S,\ 1\in{\mathcal{T}}). (A.3)

By an application of Lemma 4.9 with A={X¯1(t+h)=S}A=\{{{\bar{X}}}_{1}(t+h)=S\}, A={X¯1(t)=S}A^{\prime}=\{{{\bar{X}}}_{1}(t)=S\}, B={X¯(t+h)=S}B=\{{{\bar{X}}}_{\varnothing}(t+h)=S\}, B={X¯(t)=S,1𝒯}B^{\prime}=\{{{\bar{X}}}_{\varnothing}(t)=S,1\in{\mathcal{T}}\} we obtain

gS(t+h)=gS(t)(X¯(t+h)=S,X¯1(t+h)=S,|X¯(t)=S,X¯1(t)=S, 1𝒯)(X¯(t+h)=S|X¯(t)=S, 1𝒯).g_{S}(t+h)=g_{S}(t)\frac{{\mathbb{P}}({{\bar{X}}}_{\varnothing}(t+h)=S,\ {{\bar{X}}}_{1}(t+h)=S,|\ {{\bar{X}}}_{\varnothing}(t)=S,\ {{\bar{X}}}_{1}(t)=S,\ 1\in{\mathcal{T}})}{{\mathbb{P}}({{\bar{X}}}_{\varnothing}(t+h)=S\ |\ {{\bar{X}}}_{\varnothing}(t)=S,\ 1\in{\mathcal{T}})}. (A.4)

Since B¯1(t)+B¯(t){\bar{B}}_{1}(t)+{\bar{B}}_{\varnothing}(t) is the rate at which X¯,1(t){\bar{X}}_{\varnothing,1}(t) leaves the state (S,(S, S)S), the numerator on the right-hand side of (A.4) is equal to 1h(B¯1(t)+B¯(t))+o(h)1-h({\bar{B}}_{1}(t)+{\bar{B}}_{\varnothing}(t))+o(h). For the denominator, observe that the rate q^(t,X¯,1)\hat{q}_{\varnothing}(t,{\bar{X}}_{\varnothing,1}) on the event {X¯(t)=S,1𝒯}\{{\bar{X}}_{\varnothing}(t-)=S,1\in{\mathcal{T}}\} is equal to

𝔼[q(1,t,X¯,X¯𝕍)|X¯(t)=S,1𝒯]=βt𝔼[(X¯1(t))|X¯(t)=S, 1𝒯]+βt𝔼[(X¯𝕍{1}(t))|X¯(t)=S, 1𝒯]=βtgI(t)+βtB¯(t),\displaystyle\begin{split}&{\mathbb{E}}[q(1,t,{\bar{X}}_{\varnothing},{\bar{X}}_{\partial^{\mathbb{V}}_{\varnothing}})\ |{\bar{X}}_{\varnothing}(t-)=S,1\in{\mathcal{T}}]\\ =&\beta_{t}{\mathbb{E}}[{\mathcal{I}}({\bar{X}}_{1}(t-))\ |{\bar{X}}_{\varnothing}(t-)=S,\ 1\in{\mathcal{T}}]+\beta_{t}{\mathbb{E}}[{\mathcal{I}}({\bar{X}}_{\partial^{\mathbb{V}}_{\varnothing}\setminus\{1\}}(t-))\ |{\bar{X}}_{\varnothing}(t-)=S,\ 1\in{\mathcal{T}}]\\ =&\beta_{t}g_{I}(t-)+\beta_{t}{\bar{B}}_{\varnothing}(t-),\end{split}

where the first equality follows from (4.3) with α=1\alpha=1, and the second follows from the definition of gIg_{I} in (4.41) and by (A.2) (on observing that the event {1𝒯}\{1\in{\mathcal{T}}\} is X¯1(t){\bar{X}}_{1}(t)-measurable). Therefore, it follows that

gS(t+h)=gS(t)1h(B¯1(t)+B¯(t))+o(h)1h(βtgI(t)+B¯(t))+o(h),g_{S}(t+h)=g_{S}(t)\frac{1-h({\bar{B}}_{1}(t)+{\bar{B}}_{\varnothing}(t))+o(h)}{1-h({\beta_{t}}g_{I}(t)+{\bar{B}}_{\varnothing}(t))+o(h)},

Which implies

gS(t+h)gS(t)=gS(t)hβtgI(t)hB¯1(t)+o(h)1+o(1).\displaystyle\begin{split}g_{S}(t+h)-g_{S}(t)&=g_{S}(t)\frac{h{\beta_{t}}g_{I}(t)-h{\bar{B}}_{1}(t)+o(h)}{1+o(1)}.\end{split}

In turn, this implies

g˙S=gS(βgIB¯1).\dot{g}_{S}=g_{S}({\beta}g_{I}-{\bar{B}}_{1}). (A.5)

Similarly, recalling that gE(t+h)=(X¯1(t+h)=E|X¯(t+h)=S, 1𝒯)g_{E}(t+h)={\mathbb{P}}({\bar{X}}_{1}(t+h)=E\ |\ {\bar{X}}_{\varnothing}(t+h)=S,\ 1\in{\mathcal{T}}) from (4.41), and using the monotonicity property (4.2) with α=1\alpha=1, by a similar derivation as (A.3)-(A.5),

gE(t+h)=a=S,Ega(t)(X¯(t+h)=S,X¯1(t+h)=E|X¯(t)=S,X¯1(t)=a, 1𝒯)(X¯(t+h)=S|X¯(t)=S,1𝒯)=gS(t)(hB¯1(t)+o(h))+gE(t)(1h(λt+B¯(t))+o(h))1h(gI(t)βt+B¯(t))+o(h),\displaystyle\begin{split}g_{E}(t+h)&=\sum_{a=S,E}g_{a}(t)\frac{{\mathbb{P}}({{\bar{X}}}_{\varnothing}(t+h)=S,\ {{\bar{X}}}_{1}(t+h)=E|\ {{\bar{X}}}_{\varnothing}(t)=S,\ {{\bar{X}}}_{1}(t)=a,\ 1\in{\mathcal{T}})}{{\mathbb{P}}({{\bar{X}}}_{\varnothing}(t+h)=S|\ {{\bar{X}}}_{\varnothing}(t)=S,1\in{\mathcal{T}})}\\ &=\frac{g_{S}(t)(h{\bar{B}}_{1}(t)+o(h))+g_{E}(t)(1-h({\lambda_{t}}+{\bar{B}}_{\varnothing}(t))+o(h))}{1-h(g_{I}(t){\beta_{t}}+{\bar{B}}_{\varnothing}(t))+o(h)},\end{split}

and, hence,

gE(t+h)gE(t)=(1+o(1))(hgS(t)B¯1(t)hgE(t)(λtβtgI(t))+o(h)).\displaystyle\begin{split}&g_{E}(t+h)-g_{E}(t)=(1+o(1))(hg_{S}(t){\bar{B}}_{1}(t)-hg_{E}(t)({\lambda_{t}}-\beta_{t}g_{I}(t))+o(h)).\end{split}

It follows that

g˙E=gSB¯1gE(λβgI).\dot{g}_{E}=g_{S}{\bar{B}}_{1}-g_{E}(\lambda-{\beta}g_{I}). (A.6)

Next, we see that

gI(t+h)=a=S,E,Iga(t)(X¯(t+h)=S,X¯1(t+h)=I|X¯(t)=S,X¯1(t)=a, 1𝒯)(X¯(t+h)=S|X¯(t)=S,1𝒯)=gS(t)o(h2)+gE(t)(hλt)+gI(t)(1h(β+B¯+ρ))+o(h)1h(gI(t)βt+B¯(t))+o(h),\displaystyle\begin{split}g_{I}(t+h)&=\sum_{a=S,E,I}g_{a}(t)\frac{{\mathbb{P}}({{\bar{X}}}_{\varnothing}(t+h)=S,\ {{\bar{X}}}_{1}(t+h)=I|\ {{\bar{X}}}_{\varnothing}(t)=S,\ {{\bar{X}}}_{1}(t)=a,\ 1\in{\mathcal{T}})}{{\mathbb{P}}({{\bar{X}}}_{\varnothing}(t+h)=S|\ {{\bar{X}}}_{\varnothing}(t)=S,1\in{\mathcal{T}})}\\ &=\frac{g_{S}(t)o(h^{2})+g_{E}(t)(h\lambda_{t})+g_{I}(t)(1-h(\beta+{\bar{B}}_{\varnothing}+\rho))+o(h)}{1-h(g_{I}(t){\beta_{t}}+{\bar{B}}_{\varnothing}(t))+o(h)},\end{split}

by the monotonicity property (4.2) with α=1\alpha=1 and by the fact that the probability that two jumps occur in an interval of length hh is o(h2)o(h^{2}), since the driving noises 𝐍{,1}{\mathbf{N}}_{\{\varnothing,1\}} as in Proposition 4.6 are independent Poisson point processes with intensity measure equal to Lebesgue measure. We then have

gI(t+h)gI(t)=(1+o(1))(hgE(t)λthgI(βt+ρtβtgI(t))+o(h)),\displaystyle\begin{split}&g_{I}(t+h)-g_{I}(t)=(1+o(1))(hg_{E}(t)\lambda_{t}-hg_{I}(\beta_{t}+\rho_{t}-\beta_{t}g_{I}(t))+o(h)),\end{split}

which implies the third equation in (2.11).

Recalling that GI(t)=0tβugI(u)𝑑uG_{I}(t)=\int_{0}^{t}\beta_{u}g_{I}(u)du, by the same argument as (4.24)-(4.30) in the proof of Theorem 4.10, B¯1(t){\bar{B}}_{1}(t) satisfies

B¯1(t)=βtgI(t)𝔼[d11|X¯1(t)=S]=βtgIk=0kθ^(k)ekGI(t)n=0θ^(n)enGI(t).\displaystyle\begin{split}{\bar{B}}_{1}(t)&=\beta_{t}g_{I}(t){\mathbb{E}}[d_{1}-1\ |\ {\bar{X}}_{1}(t-)=S]\\ &=\beta_{t}g_{I}\frac{\sum_{k=0}^{\infty}k{\hat{\theta}}(k)e^{-kG_{I}(t)}}{\sum_{n=0}^{\infty}{\hat{\theta}}(n)e^{-nG_{I}(t)}}.\end{split}

Substituting this back into (A.5) and (A.6), we obtain the first and second equations in (2.11). This concludes the proof. ∎

Proof of Theorem 4.14.

Throughout the proof, we simplify the notation and write X¯{{\bar{X}}} in lieu of X¯𝒯{{\bar{X}}}^{{\mathcal{T}}}. By Assumption C, the fact that gIg_{I} is continuous (which follows from Theorem 4.13), and the fact that the ODE (4.42) is linear, the initial value problem (4.42)-(4.43) has a unique solution. Clearly by (4.41) the initial conditions (4.43) hold.

To prove (4.42) we proceed similarly as in the proof of Theorem 4.12. We start by considering QS,S;kQ_{S,S;k}. Fix t0t\geq 0, h>0h>0 and kk in the support of θ\theta. Then, using the monotonicity of the SEIR process (see (4.2)) in the second quality, and the fact that X¯(t)=yX¯k{\bar{X}}_{\partial_{\varnothing}}(t)=y\in{\bar{X}}^{k} implies that d=kd_{\varnothing}=k,

QS,S;k(t+h)=(X¯(t+h)=S|X¯(0)=S,d=k)=y𝒳¯k(X¯(t+h)=S,X¯(t)=S,X¯(0)=S,X¯(t)=y)(X¯(0)=S,d=k)=y𝒮k,t(X¯(t+h)=S|X¯(t)=S,X¯(t)=y)(X¯(t)=S,X¯(t)=y|X¯(0)=S,d=k),\displaystyle\begin{split}&Q_{S,S;k}(t+h)\\ &={\mathbb{P}}({{\bar{X}}}_{\varnothing}(t+h)=S\ |\ {{\bar{X}}}_{\varnothing}(0)=S,\ d_{\varnothing}=k)\\ &=\sum_{y\in\bar{\mathcal{X}}^{k}}\frac{{\mathbb{P}}({\bar{X}}_{\varnothing}(t+h)=S,\ {\bar{X}}_{\varnothing}(t)=S,\ {\bar{X}}_{\varnothing}(0)=S,\ {\bar{X}}_{\partial_{\varnothing}}(t)=y)}{{\mathbb{P}}({\bar{X}}_{\varnothing}(0)=S,\ d_{\varnothing}=k)}\\ &=\sum_{y\in{\mathcal{S}}_{k,t}}{\mathbb{P}}({{\bar{X}}}_{\varnothing}(t+h)=S|{{\bar{X}}}_{\varnothing}(t)=S,{{\bar{X}}}_{\partial_{\varnothing}}(t)=y){\mathbb{P}}({{\bar{X}}}_{\varnothing}(t)=S,{{\bar{X}}}_{\partial_{\varnothing}}(t)=y|{{\bar{X}}}_{\varnothing}(0)=S,d_{\varnothing}=k),\end{split} (A.7)

where 𝒮k,t:={y𝒳¯k:(X¯(t)=S,X¯(t)=y)>0}{\mathcal{S}}_{k,t}:=\{y\in\bar{\mathcal{X}}^{k}\ :\ {\mathbb{P}}({\bar{X}}_{\varnothing}(t)=S,{\bar{X}}_{\partial_{\varnothing}}(t)=y)>0\}. Since by (4.3) the jump rate of a susceptible individual with neighbors y𝒳¯ky\in\bar{\mathcal{X}}^{k} is equal to βt(y){\beta_{t}}{\mathcal{I}}(y), it follows that

(X¯(t+h)=S|X¯(t)=S,X¯(t)=y)=1hβt(y)+o(h).{\mathbb{P}}({{\bar{X}}}_{\varnothing}(t+h)=S\ |\ {{\bar{X}}}_{\varnothing}(t)=S,\ {{\bar{X}}}_{{\partial_{\varnothing}}}(t)=y)=1-h{\beta_{t}}{\mathcal{I}}(y)+o(h). (A.8)

The right-hand side does not depend on the exact states of the k(y)k-{\mathcal{I}}(y) neighbors of the root that are not in state II. Thus, substituting the expression in (A.8) into the last line of (A.7) and rewriting the sum to be over the number of infected neighbors of \varnothing,

QS,S;k(t+h)=j=0k(1hβtj+o(h))(X¯(t)=S,(X¯(t))=j|X¯(0)=S,d=k)=j=0k(1hβtj+o(h))((X¯(t))=j|X¯(t)=S,X¯(0)=S,d=k)QS,S;k(t)=j=0k(1hβtj+o(h))((X¯(t))=j|X¯(t)=S,d=k)QS,S;k(t).\displaystyle\begin{split}Q_{S,S;k}(t+h)=&\sum_{j=0}^{k}(1-h{\beta_{t}}j+o(h))\ {\mathbb{P}}({{\bar{X}}}_{\varnothing}(t)=S,\ {\mathcal{I}}({{\bar{X}}}_{{\partial_{\varnothing}}}(t))=j\ |\ {{\bar{X}}}_{\varnothing}(0)=S,d_{\varnothing}=k)\\ =&\sum_{j=0}^{k}(1-h{\beta_{t}}j+o(h))\ {\mathbb{P}}({\mathcal{I}}({{\bar{X}}}_{{\partial_{\varnothing}}}(t))=j\ |{{\bar{X}}}_{\varnothing}(t)=S,\ {{\bar{X}}}_{\varnothing}(0)=S,\ d_{\varnothing}=k)\ Q_{S,S;k}(t)\\ =&\sum_{j=0}^{k}(1-h{\beta_{t}}j+o(h))\ {\mathbb{P}}({\mathcal{I}}({{\bar{X}}}_{{\partial_{\varnothing}}}(t))=j\ |{{\bar{X}}}_{\varnothing}(t)=S,\ d_{\varnothing}=k)\ Q_{S,S;k}(t).\end{split} (A.9)

Letting α=1\alpha=1 in Proposition 4.7, it follows that {X¯i(t):i}\{{{\bar{X}}}_{i}(t)\ :\ i\sim\varnothing\} are conditionally i.i.d. given X¯(t)=S{{\bar{X}}}_{\varnothing}(t)=S and d=kd_{\varnothing}=k. For each m[1,k]m\in{\mathbb{N}}\cap[1,k], by Proposition 4.8 and an Application of Proposition 4.7 with A=𝕋m:={mv:v𝕍}A={\mathbb{T}}_{m}:=\{mv\ :v\in\mathbb{V}\}, the subtree rooted at mm, 𝕍A={}\partial^{\mathbb{V}}A=\{\varnothing\} and B={m}B={\mathbb{N}}\setminus\left\{m\right\}, observing that d=𝟏{X¯(0)}d_{\varnothing}=\sum_{\ell\in{\mathbb{N}}}{\bm{1}}_{\{{\bar{X}}_{\ell}(0)\neq\star\}}, we have that

(X¯m(t)=I|X¯(t)=S,d=k)=(X¯m(t)=I|X¯(t)=S,m𝒯)=gI(t),{\mathbb{P}}({\bar{X}}_{m}(t)=I\ |\ {\bar{X}}_{\varnothing}(t)=S,\ d_{\varnothing}=k)={\mathbb{P}}({\bar{X}}_{m}(t)=I\ |\ {\bar{X}}_{\varnothing}(t)=S,\ m\in{\mathcal{T}})=g_{I}(t),

where gIg_{I} is defined in (4.41). It follows that, conditional on X¯(t)=S{{\bar{X}}}_{\varnothing}(t)=S and d=kd_{\varnothing}=k, (X¯(t)){\mathcal{I}}({{\bar{X}}}_{{\partial_{\varnothing}}}(t)) has binomial distribution with parameters (k(k, gI(t))g_{I}(t)). Letting YY be a binomial random variable with parameters (k,(k, gI(t))g_{I}(t)), it follows from (A.9) that

QS,S;k(t+h)=QS,S;k(t)(1hβt𝔼[Y]+o(h))=(1hβtkgI+o(h))QS,S;k(t),\displaystyle\begin{split}Q_{S,S;k}(t+h)&=Q_{S,S;k}(t)(1-h{\beta_{t}}{\mathbb{E}}[Y]+o(h))\\ &=(1-h{\beta_{t}}kg_{I}+o(h))Q_{S,S;k}(t),\end{split} (A.10)

and, thus,

limh0+QS,S;k(t+h)QS,S;k(t)h\displaystyle\lim_{h\rightarrow 0^{+}}\frac{Q_{S,S;k}(t+h)-Q_{S,S;k}(t)}{h} =limh0+(1hβtkgI(t)+o(h)1)QS,S;k(t)h\displaystyle=\lim_{h\rightarrow 0^{+}}\frac{(1-h{\beta_{t}}kg_{I}(t)+o(h)-1)Q_{S,S;k}(t)}{h}
=βtkgI(t)QS,S;k(t),\displaystyle=-{\beta_{t}}kg_{I}(t)Q_{S,S;k}(t),

which proves the first equation in (4.42). The derivation of the ODEs for QS,E;kQ_{S,E;k} and QS,I,;kQ_{S,I,;k} is similar and outlined below. As in the last line of (A.7) write,

QS,E;k(t+h)=𝒬¯E(h)+𝒬¯S(h),Q_{S,E;k}(t+h)=\bar{{\mathcal{Q}}}_{E}(h)+\bar{{\mathcal{Q}}}_{S}(h), (A.11)

where, for b={S,E}b=\{S,E\},

𝒬¯b(h)=j=0k(X¯(t+h)=E,X¯(t)=b,(X¯(t))=j|X¯(0)=S,d=k).\bar{{\mathcal{Q}}}_{b}(h)=\sum_{j=0}^{k}{\mathbb{P}}({{\bar{X}}}_{\varnothing}(t+h)=E,\ {{\bar{X}}}_{\varnothing}(t)=b,\ {\mathcal{I}}({{\bar{X}}}_{{\partial_{\varnothing}}}(t))=j\ |\ {{\bar{X}}}_{\varnothing}(0)=S,\ d_{\varnothing}=k).

Recalling the definition of the rates q1q_{1} in (4.3) and using arguments similar to what used to derive (A.8)-(A.10), 𝒬¯S=(hβtkgI(t)+o(h))QS,S;k(t)\bar{{\mathcal{Q}}}_{S}=(h{\beta_{t}}kg_{I}(t)+o(h))Q_{S,S;k}(t) and

𝒬¯E(h)=j=0k(1λth+o(h))(X¯(t)=E,(X¯(t))=j|X¯(0)=S,d=k)=(1λth+o(h))j=0k(X¯(t)=E,(X¯(t))=j|X¯(0)=S,d=k)=(1λth+o(h))(X¯(t)=E|X¯(0)=S,d=k)=(1λth+o(h))QS,E;k(t).\displaystyle\begin{split}\bar{{\mathcal{Q}}}_{E}(h)=&\sum_{j=0}^{k}(1-{\lambda_{t}}h+o(h)){\mathbb{P}}({{\bar{X}}}_{\varnothing}(t)=E,\ {\mathcal{I}}({{\bar{X}}}_{{\partial_{\varnothing}}}(t))=j|{{\bar{X}}}_{\varnothing}(0)=S,\ d_{\varnothing}=k)\\ =&(1-{\lambda_{t}}h+o(h))\sum_{j=0}^{k}{\mathbb{P}}({{\bar{X}}}_{\varnothing}(t)=E,\ {\mathcal{I}}({{\bar{X}}}_{{\partial_{\varnothing}}}(t))=j|{{\bar{X}}}_{\varnothing}(0)=S,\ d_{\varnothing}=k)\\ =&(1-{\lambda_{t}}h+o(h)){\mathbb{P}}({{\bar{X}}}_{\varnothing}(t)=E\ |{{\bar{X}}}_{\varnothing}(0)=S,\ d_{\varnothing}=k)\\ =&(1-{\lambda_{t}}h+o(h))Q_{S,E;k}(t).\end{split}

Therefore, QS,E;k(t+h)QS,E;k(t)=hkβtgI(t)QS,S;k(t)λthQS,E;k(t)+o(h)Q_{S,E;k}(t+h)-Q_{S,E;k}(t)=hk{\beta_{t}}g_{I}(t)Q_{S,S;k}(t)-{\lambda_{t}}hQ_{S,E;k}(t)+o(h) which implies the second equation in (4.42). Proceeding similarly, we obtain the relation

QS,I;k(t+h)=b=S,E,I(X¯(t+h)=I|X¯(t)=b,X¯(0)=Sd=k)=λthQS,E;k(t)+(1ρt)hQS,I;k\displaystyle\begin{split}Q_{S,I;k}(t+h)&=\sum_{b=S,E,I}{\mathbb{P}}({\bar{X}}_{\varnothing}(t+h)=I|{\bar{X}}_{\varnothing}(t)=b,{\bar{X}}_{\varnothing}(0)=Sd_{\varnothing}=k)\\ &=\lambda_{t}hQ_{S,E;k}(t)+(1-\rho_{t})hQ_{S,I;k}\end{split}

and

QE,I(t+h)=b=S,E,I(X¯(t+h)=I|X¯(t)=b,X¯(0)=E,d=k)=λthQE,E(t)+(1ρt)hQE,I,\displaystyle\begin{split}Q_{E,I}(t+h)&=\sum_{b=S,E,I}{\mathbb{P}}({\bar{X}}_{\varnothing}(t+h)=I|{\bar{X}}_{\varnothing}(t)=b,{\bar{X}}_{\varnothing}(0)=E,d_{\varnothing}=k)\\ &=\lambda_{t}hQ_{E,E}(t)+(1-\rho_{t})hQ_{E,I},\end{split}

which imply the third and fifth equations in (4.42). Setting rtE:=λtr^{E}_{t}:=\lambda_{t}, and rtI:=ρtr^{I}_{t}:=\rho_{t}, for a{E,I}a\in\{E,\ I\} we see that

Qa,a(t+h)==y𝒳¯k(X¯(t+h)=a|X¯(t)=a,X¯(t)=y)(X¯(t)=a,X¯(t)=y|X¯(0)=a)=(1hrta+o(h))y𝒳¯k(X¯(t)=a,X¯(t)=y|X¯(0)=a)=(1hrta+o(h))Qa,a(t),\displaystyle\begin{split}&Q_{a,a}(t+h)=\\ &=\sum_{y\in\bar{\mathcal{X}}^{k}}{\mathbb{P}}({{\bar{X}}}_{\varnothing}(t+h)=a\ |\ {{\bar{X}}}_{\varnothing}(t)=a,{{\bar{X}}}_{\partial_{\varnothing}}(t)=y){\mathbb{P}}({{\bar{X}}}_{\varnothing}(t)=a,\ {{\bar{X}}}_{\partial_{\varnothing}}(t)=y\ |\ {{\bar{X}}}_{\varnothing}(0)=a)\\ &=(1-h{r^{a}_{t}}+o(h))\sum_{y\in\bar{\mathcal{X}}^{k}}{\mathbb{P}}({{\bar{X}}}_{\varnothing}(t)=a,\ {{\bar{X}}}_{\partial_{\varnothing}}(t)=y\ |\ {{\bar{X}}}_{\varnothing}(0)=a)\\ &=(1-h{r^{a}_{t}}+o(h))Q_{a,a}(t),\end{split}

which proves the fourth and sixth equations in (4.42), thus concluding the proof.

Appendix B Proof of Proposition 2.4

In this section we prove the well-posedness of the ODE system (2.4)-(2.5). We start with the following elementary result.

Lemma B.1.

Suppose that θ𝒫(0)\theta\in{\mathcal{P}}({\mathbb{N}}_{0}) has a finite third moment. Then Φθ^\Phi_{\hat{\theta}}, defined in (4.45), is Lipschitz continuous on [0,)[0,\infty).

Proof.

It is easy to see that under the assumption that θ\theta has a finite third moment, the size-biased distribution θ^{\hat{\theta}}, defined in (2.2), has a finite second moment. Indeed, let Y^{\hat{Y}} and YY be random variables with laws θ^{\hat{\theta}} and θ\theta, respectively. By (2.2), it is easy to see that

𝔼[Y^2]=𝔼[Y3]2𝔼[Y2]+𝔼[Y]𝔼[Y],{\mathbb{E}}[{\hat{Y}}^{2}]=\frac{{\mathbb{E}}[Y^{3}]-2{\mathbb{E}}[Y^{2}]+{\mathbb{E}}[Y]}{{\mathbb{E}}[Y]},

which is finite since θ\theta has finite third moment.

For z[0,)z\in[0,\infty) note that Mθ^(z)=𝔼[ezY^]1M_{\hat{\theta}}(-z)={\mathbb{E}}[e^{-z{\hat{Y}}}]\leq 1, and so by the dominated convergence theorem limzMθ^(z)=θ^(0)\lim_{z\rightarrow\infty}M_{{\hat{\theta}}}(-z)={\hat{\theta}}(0). Since θ^{\hat{\theta}} has finite second moment, again by the dominated convergence theorem Mθ^′′(z)=𝔼[Y^2ezY^]𝔼[Y^2]<M_{\hat{\theta}}^{\prime\prime}(-z)={\mathbb{E}}[{\hat{Y}}^{2}e^{-z{\hat{Y}}}]\leq{\mathbb{E}}[{\hat{Y}}^{2}]<\infty, Mθ^′′M_{{\hat{\theta}}}^{\prime\prime} is continuous on (,0)(-\infty,0) and limzMθ^′′(z)=0\lim_{z\rightarrow\infty}M^{\prime\prime}_{{\hat{\theta}}}(-z)=0. Thus, it follows from the limits established above that

θ^(0)>0limzMθ^′′(z)Mθ^(z)=0θ^(0)=0{\hat{\theta}}(0)>0\quad\Rightarrow\quad\lim_{z\rightarrow\infty}\frac{M^{\prime\prime}_{{\hat{\theta}}}(-z)}{M_{{\hat{\theta}}}(-z)}=\frac{0}{{\hat{\theta}}(0)}=0 (B.1)

Now, setting Φ:=Φθ^\Phi:=\Phi_{{\hat{\theta}}} for conciseness, it follows that

Φ(z):=ddzΦ(z)=(Mθ^(z))2Mθ^′′(z)Mθ^(z)Mθ^2(z)=(Φ(z))2Mθ^′′(z)Mθ^(z).\Phi^{\prime}(z):=\frac{d}{dz}\Phi(z)=\frac{(M^{\prime}_{{\hat{\theta}}}(-z))^{2}-M^{\prime\prime}_{{\hat{\theta}}}(-z)M_{\hat{\theta}}(-z)}{M^{2}_{{\hat{\theta}}}(-z)}=(\Phi(z))^{2}-\frac{M^{\prime\prime}_{{\hat{\theta}}}(-z)}{M_{{\hat{\theta}}}(-z)}. (B.2)

By Lemma 4.15, Φ\Phi is bounded on [0,)[0,\infty). Furthermore, Mθ^′′(0)/Mθ^(0)=𝔼[Y^2]M^{\prime\prime}_{{\hat{\theta}}}(0)/M_{{\hat{\theta}}}(0)={\mathbb{E}}[{\hat{Y}}^{2}]. Recall the quantity d¯θ^=min{d0:θ^(d)>0}\underline{d}_{\hat{\theta}}=\min\{d\in{\mathbb{N}}_{0}:{\hat{\theta}}(d)>0\} introduced in (4.44). Using (B.1) for the case d¯θ^=0\underline{d}_{\hat{\theta}}=0 (which is equivalent to θ^(0)>0{\hat{\theta}}(0)>0), and a similar argument as (4.46) in Lemma 4.15 for the case d¯θ^>0\underline{d}_{\hat{\theta}}>0, we have

limzMθ^′′(z)Mθ^(z)=d¯θ^2.\lim_{z\rightarrow\infty}\frac{M^{\prime\prime}_{{\hat{\theta}}}(-z)}{M_{{\hat{\theta}}}(-z)}=\underline{d}_{\hat{\theta}}^{2}. (B.3)

Together with (B.2), the continuity of Φ\Phi on [0,)[0,\infty) and the continuity of Mθ^M_{\hat{\theta}} and Mθ^′′M^{\prime\prime}_{\hat{\theta}} on (,0](-\infty,0], this implies that Φ\Phi^{\prime} is uniformly bounded on [0,)[0,\infty). This completes the proof. ∎

Proof of Proposition 2.4.

By Assumption A, β\beta and ρ\rho are continuous in tt. By Lemma 4.15, Φ(z)\Phi(z) is continuous in z[0,)z\in[0,\infty). Therefore, the right-hand side of the ODE (2.4) is continuous and so by Peano’s existence theorem, there exists τ(0,)\tau\in(0,\infty) and a solution (fS(f_{S},fIf_{I},FI):[0,τ)3F_{I}):[0,\tau)\rightarrow{\mathbb{R}}^{3} to (2.4)-(2.5) on [0,τ)[0,\tau).

Next, fix s[0,τ)s\in[0,\tau). We claim that (fS(s)(f_{S}(s),fI(s)f_{I}(s),FI(s))[0,1]×[0,1]×[0,)F_{I}(s))\in[0,1]\times[0,1]\times[0,\infty). Since fS(0)=s0(0,1)f_{S}(0)=s_{0}\in(0,1) and fI(0)=1s0(0,1)f_{I}(0)=1-s_{0}\in(0,1) and the right-hand side of the first (respectively, second) equation in (2.4) is equal to 0 whenever fS(t)=0f_{S}(t)=0 (respectively, fI(t)=0f_{I}(t)=0), it follows that fS(s)0f_{S}(s)\geq 0 (respectively, fI(s)0f_{I}(s)\geq 0). In turn, this implies that F˙I(s)0\dot{F}_{I}(s)\geq 0, and therefore that FI(s)>0F_{I}(s)>0 (since FI(0)=0F_{I}(0)=0 and F˙I(0+)=fI(0)>0\dot{F}_{I}(0+)=f_{I}(0)>0). Now, by summing the first two equations in (2.4), we obtain

f˙I(s)+f˙S(s)=fI(s)βs(fS(s)+fI(s)1ρsβs).\dot{f}_{I}(s)+\dot{f}_{S}(s)=f_{I}(s)\beta_{s}\left(f_{S}(s)+f_{I}(s)-1-\frac{\rho_{s}}{\beta_{s}}\right). (B.4)

Since fS(0)+fI(0)=1f_{S}(0)+f_{I}(0)=1, it follows that fS(s)+fI(s)f_{S}(s)+f_{I}(s) is strictly decreasing in ss, and in particular fS(s)+fI(s)1f_{S}(s)+f_{I}(s)\leq 1.

Finally, by Lemma B.1, the right-hand side of (2.4) is Lipschitz continuous in (fS,(f_{S}, fIf_{I}, FI)F_{I}) on [0,1]×[0,1]×[0,)[0,1]\times[0,1]\times[0,\infty). Thus, it follows that τ=\tau=\infty and (2.4)-(2.5) has a unique solution for all times. ∎

Proposition 2.9 is proved in the same way as the proof of 2.4 by first using Lemma 4.15 and Assumption C to establish the existence of (gS(g_{S}, gE,g_{E}, gI,g_{I}, GI)G_{I}) solving (2.11)-(2.12) on [0,τ)[0,\tau) for some τ(0,)\tau\in(0,\infty), then showing that such a solution stays in [0,1]×[0,1]×[0,1]×[0,)[0,1]\times[0,1]\times[0,1]\times[0,\infty), where, by Lemma B.1, the right-hand side of (2.11) is Lipschitz.

References