Hostname: page-component-745bb68f8f-5r2nc Total loading time: 0 Render date: 2025-02-06T01:52:02.043Z Has data issue: false hasContentIssue false

Construction of age-structured branching processes by stochastic equations

Published online by Cambridge University Press:  31 May 2022

Lina Ji*
Affiliation:
Shenzhen MSU-BIT University
Zenghu Li*
Affiliation:
Beijing Normal University
*
*Postal address: Faculty of Computational Mathematics and Cybernetics, Shenzhen MSU-BIT University, Shenzhen 518172, People’s Republic of China. Email: jiln@smbu.edu.cn
**Postal address: Laboratory of Mathematics and Complex Systems, School of Mathematical Sciences, Beijing Normal University, Beijing 100875, People’s Republic of China. Email: lizh@bnu.edu.cn
Rights & Permissions [Opens in a new window]

Abstract

We provide constructions of age-structured branching processes without or with immigration as pathwise-unique solutions to stochastic integral equations. A necessary and sufficient condition for the ergodicity of the model with immigration is also given.

Type
Original Article
Copyright
© The Author(s), 2022. Published by Cambridge University Press on behalf of Applied Probability Trust

1. Introduction

Branching processes were introduced to describe the evolution of populations undergoing random reproduction. For the classical continuous-time branching process, it is assumed that an individual has exponential life-length and gives birth to a random number of offspring at the end of its life. The age-dependent branching process introduced in [Reference Bellman and Harris3] assumes that the individual may have a general life-length distribution. The model has been generalized further to allow the individual to give birth to offspring at any time during its life; see, e.g., [Reference Crump and Mode8, Reference Crump and Mode9, Reference Doney11, Reference Jagers15, Reference Kendall20]. Those age-dependent models are usually not Markovian if one only considers the evolution of the total number of individuals in the population. A measure-valued Markovian branching particle system was introduced in [Reference Bose4] to describe the evolution of a birth–death model with age structures; see also [Reference Bose and Kaj5, Reference Bose and Kaj6, Reference Dawson, Gorostiza and Li10, Reference Kaj and Sagitov19]. In the models mentioned above, the death rate and the offspring distribution of an individual may depend on its age, but different individuals behave independently of each other. Several authors have also studied population models where the reproduction depends on the age structure of the whole population; see, e.g., [Reference Jagers and Klebaner17, Reference Jagers and Klebaner18, Reference Metz and Tran24, Reference Oelschläger25, Reference Tran28]. We refer to [Reference Athreya and Ney1, Reference Etheridge12, Reference Harris14, Reference Jagers16, Reference Li21] for systematic treatments of various classes of branching processes.

The approach of stochastic equations has played an important role in recent developments of the theory of branching processes. The reader may refer to [Reference Bansaye and Méléard2, Reference Li22, Reference Pardoux26] and the references therein for applications of this approach to continuous-state branching processes. Stochastic equations have also been introduced in the study of discrete-state branching models; see, e.g., [Reference Champagnat, Ferrière and Méléard7, Reference Fournier and Méléard13, Reference Tran28]. In particular, a stochastic equation for an age-structured birth–death process was proposed in [Reference Tran28] in the study of large population limits.

The purpose of this paper is to develop the approach of stochastic equations further for age-structured branching processes to allow general offspring distributions. For concreteness, we focus on the model where the death rate and the offspring distribution of an individual only depends on its own age. The model can be thought of as a typical special case of nonlocal branching particle systems; see, e.g., [Reference Bose and Kaj5, Reference Bose and Kaj6, Reference Dawson, Gorostiza and Li10, Reference Kaj and Sagitov19, Reference Li21]. We give a construction of the age-structured branching process as the pathwise-unique solutions of a stochastic equation driven by a time–space Poisson random measure. The construction determines explicitly the behavior of the trajectory of the process. By a slight extension of the stochastic equation, an age-structured immigration model is constructed. We also prove a necessary and sufficient condition for the ergodicity of the process with immigration.

Let $\mathscr{B}\big(\mathbb{R}_+\big)$ denote the Borel $\sigma$ -algebra on $\mathbb{R}_+\,:\!=\, [0,\infty)$ . Let $\mathfrak{M}\big(\mathbb{R}_+\big)$ denote the set of finite Borel measures on $\mathbb{R}_+$ with the weak convergence topology. Let $\mathfrak{D}\big(\mathbb{R}_+\big)$ be set of bounded positive right-continuous increasing functions f on $\mathbb{R}$ satisfying $f(x)= 0$ for $x< 0$ . We identify $\nu\in \mathfrak{M}\big(\mathbb{R}_+\big)$ with its distribution function $\nu\in \mathfrak{D}\big(\mathbb{R}_+\big)$ defined by $\nu(x)= \nu[0,x]$ for $x\ge 0$ . Let $\mathfrak{N}\big(\mathbb{R}_+\big)$ be the subset of $\mathfrak{M}\big(\mathbb{R}_+\big)$ consisting of integer-valued measures. Let $B\big(\mathbb{R}_+\big)$ be the Banach space of bounded Borel functions on $\mathbb{R}_+$ furnished with the supremum norm $\|\cdot\|$ . Let $C\big(\mathbb{R}_+\big)$ be the set of continuous functions in $B\big(\mathbb{R}_+\big)$ , and let $C^1\big(\mathbb{R}_+\big)$ be the set of functions in $C\big(\mathbb{R}_+\big)$ with bounded continuous derivatives of the first order. We use the superscript ‘+’ to denote the subsets of positive elements and the subscript ‘0’ to denote the subsets of functions vanishing at infinity, e.g. $B\big(\mathbb{R}_+\big)^+$ , $C_0\big(\mathbb{R}_+\big)^+$ , etc. For any $f \in B\big(\mathbb{R}_+\big)$ and $\nu \in \mathfrak{M}\big(\mathbb{R}_+\big)$ write $\langle \nu,f\rangle = \int_{\mathbb{R}_+} f(x)\nu(\text{d} x)$ . In the integrals we use the convention that, for $a \le b \in \mathbb{R}$ ,

\begin{eqnarray*} \int_a^b = \int_{(a, b]} \qquad \text{and} \qquad \int_a^\infty = \int_{(a, \infty)}.\end{eqnarray*}

The rest of this paper is organized as follows. In Section 2 we introduce the age-structured branching process and give some basic characterizations of its transition probabilities. In Section 3, the process is constructed as the pathwise-unique strong solution to a stochastic integral equation driven by a Poisson random measure. Similar results for the age-structured system with immigration are presented in Section 4, where the ergodicity of the model is also studied.

2. An age-structured branching process

In this section we introduce the age-structured branching process and give some basic characterizations of its transition probabilities. Most of the results presented here are essentially known, so we only sketch the proofs; see, e.g., [Reference Bose and Kaj6, Reference Kaj and Sagitov19, Reference Li21, Reference Li23].

Let $\alpha\in C^1\!\big(\mathbb{R}_+\big)^{+}$ be a function bounded away from zero. For each $x\in \mathbb{R}_+$ let $\{p(x,i) \,:\, i\in \mathbb{N}\}$ be a discrete probability distribution with generating function $g(x,z) = \sum_{i=0}^\infty p(x,i)z^i$ , $z\in [0,1]$ . We assume that $p(\cdot,i)\in C^1\big(\mathbb{R}_+\big)^{+}$ for every $i\in \mathbb{N}$ , and that $\|\partial_zg(\cdot,1{-})\| = \sup_{x\ge 0}\sum_{i=1}^\infty p(x,i)i< \infty$ , where $\partial_z$ denotes the first derivative with respect to z. A branching particle system is characterized by the following properties:

  1. (i) The ages of the particles increase at unit speed, i.e. they move according to realizations of the deterministic process $\xi= (\xi_t)_{t\ge 0}$ in $\mathbb{R}_+$ defined by $\xi_t= \xi_0+t$ .

  2. (ii) For a particle which is alive at time $r\ge 0$ with age $x\ge 0$ , the conditional probability of survival in the time interval [r, t) is $\exp\!\big\{{-}\int_0^{t-r} \alpha(x+s) \, \text{d} s\big\}$ .

  3. (iii) When a particle dies at age $x\ge 0$ , it gives birth to a random number of offspring with age zero according to the probability law $\{p(x, i) \,:\, i = 0, 1, \ldots\}$ determined by the generating function $g(x,\cdot)$ .

We assume that the lifetimes and the offspring productions of different particles are independent. Let $X_t(B)$ denote the number of particles alive at time $t\ge 0$ with ages belonging to the Borel set $B\subset \mathbb{R}_+$ . If we assume $X_0\big(\mathbb{R}_+\big)< \infty$ , then $\{X_t \,:\, t\ge 0\}$ is a Markov process with state space $\mathfrak{N}\big(\mathbb{R}_+\big)$ . We refer to [Reference Li21, Section 4.3] for the formulation of general branching particle systems.

Let $\sigma\in \mathfrak{N}\big(\mathbb{R}_+\big)$ and let $\{X_t^\sigma \,:\, t\ge 0\}$ be the above system with initial value $X^\sigma_0 = \sigma$ . Let $\delta_x$ denote the unit measure concentrated at $x\in \mathbb{R}_+$ . Suppose that the process is defined on a probability space $(\Omega, \mathscr{G}, \mathbb{P})$ . Properties (i)–(iii) imply that, for $f\in B\big(\mathbb{R}_+\big)^+$ ,

\begin{equation*} \mathbb{E}\left[\exp\!\big\{-\big\langle X_t^\sigma, f\big\rangle \big\}\right] = \exp\{-\langle \sigma, u_t\,f\rangle \}, \qquad u_t\,f(x) = -\log \mathbb{E}\big[\exp\!\big\{-\big\langle X_t^{\delta_x}, f\big\rangle \big\}\big].\end{equation*}

From properties (i)–(iii) we derive, as in [Reference Kaj and Sagitov19, Section 3] and [Reference Li21, Section 4.3], the following renewal equation:

(2.1) \begin{align} \text{e}^{-u_t\,f(x)} = & \exp\!\bigg\{{-}f(x+t)-\int_0^t \alpha(x+s) \, \text{d} s\bigg\}\nonumber\\[3pt]& + \int_0^t \exp\!\bigg\{-\int_0^s\alpha(x+r) \, \text{d} r\bigg\}\alpha(x+s) g\big(x+s, \text{e}^{-u_{t-s}f(0)}\big) \, \text{d} s.\end{align}

From [Reference Li21, Proposition 2.9], the above equation implies

(2.2) \begin{eqnarray} \text{e}^{-u_t\,f(x)} = \text{e}^{-f(x+t)} + \int_0^t \alpha(x+t-s)\big[g\big(x+t-s, \text{e}^{-u_sf(0)}\big) - \text{e}^{-u_sf(x+t-s)}\big] \, \text{d} s.\end{eqnarray}

The uniqueness of the solution to (2.1) and (2.2) follows by Gronwall’s inequality. The two equations are therefore equivalent.

We call any Markov process $(X_t \,:\, t\ge 0)$ with state space $\mathfrak{N}\big(\mathbb{R}_+\big)$ an $(\alpha,g)$ -age-structured branching process if it has a transition semigroup $(Q_t)_{t\ge 0}$ defined by

(2.3) \begin{eqnarray} \int_{\mathfrak{N}\big(\mathbb{R}_+\big)}\text{e}^{-\langle \nu, f\rangle }Q_t(\sigma,d\nu) = \exp\!\big\{-\langle \sigma, u_t\,f\rangle \big\}, \qquad f\in B\big(\mathbb{R}_+\big)^+,\end{eqnarray}

where $u_t\,f(x)$ is the unique solution to (2.2).

Proposition 2.1. For any $x\ge 0$ , $t\ge 0$ , and $f\in C^1_0\big(\mathbb{R}_+\big)^+$ , we have

(2.4) \begin{eqnarray} \partial_tu_t\,f(x) = \partial_xu_t\,f(x) + \alpha(x)\big[1 - \text{e}^{u_t\,f(x)}g\big(x,\text{e}^{-u_t\,f(0)}\big)\big].\end{eqnarray}

Proof.For any $t\ge 0$ let $T_t$ be the operator on the Banach space $C_0\big(\mathbb{R}_+\big)$ defined by $T_tf(x)= f(x+t)$ . By (2.2) it is easy to see that $U_t\,f(x)= 1-\text{e}^{-u_t\,f(x)}$ solves the evolution integral equation

(2.5) \begin{eqnarray}U_t\,f(x) = T_tU_0f(x) - \int_0^t T_{t-s}\phi(\cdot,U_sf)(x) \, \text{d} s,\end{eqnarray}

where $\phi(x,f) = \alpha(x)\big[g(x,1-f(0))-1+f(x)\big]$ . By a general result on semi-linear evolution equations, we know that $t\mapsto U_t\,f(x)$ is continuously differentiable and solves the differential evolution equation $\partial_tU_t\,f(x)= \partial_xU_t\,f(x) - \phi(x,U_t\,f)$ , $U_0f(x)= 1-\text{e}^{-f(x)}$ ; see, e.g., [Reference Pazy27, Theorem 6.1.5, p. 187]. Then $t\mapsto \partial_tu_t\,f(x)$ is also continuously differentiable. By differentiating both sides of (2.2) we have

\begin{align*} \text{e}^{-u_t\,f(x)}\partial_tu_t\,f(x) & = \partial_x \text{e}^{-f(x+t)} \\ & \quad + \int_0^t \partial_x\big[\alpha(x+t-s)\big(g\big(x+t-s,\text{e}^{-u_sf(0)}\big) - \text{e}^{-u_sf(x+t-s)}\big)\big] \, \text{d} s \\ & \quad + \alpha(x)\big[\text{e}^{-u_t\,f(x)}-g\big(x, \text{e}^{-u_t\,f(0)}\big)\big] \\ & = \text{e}^{-u_t\,f(x)}\partial_xu_t\,f(x) + \alpha(x)\big[\text{e}^{-u_t\,f(x)} - g\big(x, \text{e}^{-u_t\,f(0)}\big)\big] , \end{align*}

which proves (2.4).

Proposition 2.2. For any $t\ge 0$ and $\sigma\in \mathfrak{N}\big(\mathbb{R}_+\big)$ we have

(2.6) \begin{eqnarray} \int_{\mathfrak{N}\big(\mathbb{R}_+\big)}\langle \nu,f\rangle Q_t(\sigma, \text{d}\nu) = \langle \sigma, \pi_tf\rangle , \qquad f\in B\big(\mathbb{R}_+\big),\end{eqnarray}

where $(\pi_t)_{t\ge 0}$ is the semigroup of bounded kernels on $\mathbb{R}_+$ defined by

(2.7) \begin{eqnarray} { \pi_tf(x) = f(x+t) + \int_0^t \alpha(x+s)\big[\partial_z g(x+s,1{-})\pi_{t-s}f(0) - \pi_{t-s}f(x + s)\big] \, \text{d} s.}\end{eqnarray}

Proof.The existence and uniqueness of the locally bounded solution to (2.7) follows by a general result; see, e.g., [Reference Li21, Lemma 2.17]. For $f\in B\big(\mathbb{R}_+\big)^+$ we can use (2.2) to see that the unique solution of (2.7) is given by $ \pi_tf(x)= \partial_\theta u_t(\theta f)(x)|_{\theta=0}$ . By differentiating both sides of (2.3) we get (2.6). The extension to $f\in B\big(\mathbb{R}_+\big)$ is immediate by linearity.

Proposition 2.3. Let $c_*= \inf_{y\ge 0}\alpha(y)[1-\partial_zg(y,1{-})]$ . Then $\|\pi_tf\|\le e^{-c_*t}\|f\|$ for $t\ge 0$ , $f\in B\big(\mathbb{R}_+\big)$ .

Proof.This follows from (2.7) and [Reference Li23, Theorem 3.1].

Proposition 2.4. Let $c_1= \sup_{y\ge 0}\alpha(y)$ . Then, for any $t\ge 0$ and $x\ge 0$ ,

(2.8) \begin{eqnarray} \pi_tf(x)\ge u_t\,f(x)\ge \big(1-\text{e}^{-f(x+t)}\big)\text{e}^{-c_1t}, \qquad f\in B\big(\mathbb{R}_+\big)^+.\end{eqnarray}

Proof.By taking $\sigma= \delta_x$ in (2.3) and (2.6) and using Jensen’s inequality we get the first inequality in (2.8). Let $U_t\,f(x)$ be as in the proof of Proposition 2.1. By (2.5) and a comparison theorem we have $U_t\,f(x)\ge v_tf(x)$ , where $(t,x)\mapsto v_tf(x)$ solves

\begin{eqnarray*}v_tf(x) = 1-\text{e}^{-f(x+t)} - \int_0^t \alpha(x+s)v_{t-s}f(x+s) \, \text{d} s.\end{eqnarray*}

The unique locally bounded solution to the above equation is given by

\begin{eqnarray*}v_tf(x) = \big(1-\text{e}^{-f(x+t)}\big)\exp\!\bigg\{-\int_0^t \alpha(x+s) \, \text{d} s\bigg\}.\end{eqnarray*}

Then we have the estimate (2.8).

3. Construction by stochastic equations

In this section we give a construction of the age-structured branching process by solving a stochastic equation driven by a Poisson random measure. Recall that $\mathfrak{D}\big(\mathbb{R}_+\big)$ denotes the set of bounded positive right-continuous increasing functions f on $\mathbb{R}$ satisfying $f(x)= 0$ for $x< 0$ . For $\mu\in \mathfrak{D}\big(\mathbb{R}_+\big)$ and $\alpha\in B\big(\mathbb{R}_+\big)^+$ define $A_\alpha(\mu,y)=\inf\{z\ge 0 \,:\, \langle \mu,\alpha\textbf{1}_{[0,z]}\rangle > \langle \mu,\alpha\rangle y\}$ , $0\le y\le 1$ , with $\inf\emptyset= \infty$ by convention. Then $\langle \mu,\alpha\rangle = 0$ implies $A_\alpha(\mu,y)= \infty$ for all $0\le y\le 1$ . By an elementary result in probability theory, we have the following lemma.

Lemma 3.1. If $\langle \mu,\alpha\rangle > 0$ and if $\xi$ is a random variable with uniform distribution on (0, 1], then $\mathbb{P}\{A_\alpha(\mu,\xi)\in \text{d} x\} = \langle \mu,\alpha\rangle ^{-1} \alpha(x)\mu(\text{d} x)$ , $x\ge 0$ .

Suppose that $(\Omega, \mathscr{F}, \mathscr{F}_t, \mathbb{P})$ is a filtered probability space satisfying the usual hypotheses. Let $M(\text{d} t, \text{d} u, \text{d} y, \text{d} z, \text{d} v)$ be an $(\mathscr{F}_t)$ -Poisson random measure on $(0,\infty)^2\times (0, 1]\times \mathbb{N}\times (0, 1]$ with intensity $\text{d} t \, \text{d} u \, \text{d} y \, \pi(\text{d} z) \, \text{d} v$ , where $\pi(\text{d} z)$ denotes the counting measure on $\mathbb{N}$ . Given an $\mathscr{F}_0$ -measurable random function $X_0\in \mathfrak{D}\big(\mathbb{R}_+\big)$ , we consider the stochastic integral equation, for $t\ge 0$ and $x\ge 0$ ,

(3.1) \begin{align} X_t(x) =\, & X_0(x-t) + \int_0^t\int_0^{\langle X_{s-},\alpha\rangle }\!\! \int_0^1\int_{\mathbb{N}}\int_0^{p(A_\alpha(X_{s-},y),z)}\!\! z\textbf{1}_{\{t-s\le x\}} M(\text{d} s,\text{d} u,\text{d} y,\text{d} z,\text{d} v)\nonumber\\[3pt]& - \int_0^t\int_0^{\langle X_{s-},\alpha\rangle }\!\! \int_0^1\int_{\mathbb{N}} \int_0^{p(A_\alpha(X_{s-},y),z)}\!\! \textbf{1}_{\{A_\alpha(X_{s-},y)+t-s\le x\}} M(\text{d} s,\text{d} u,\text{d} y,\text{d} z,\text{d} v). \end{align}

Heuristically, the left-hand side $X_t(x)$ is the number of individuals at time t with ages less than x. On the right-hand side, the first term $X_0(x-t)$ counts the number of individuals having ages less than $x-t$ at time 0, and thus having ages less than x at time t. A death in the population occurs at time $s\in [0,t]$ at rate $\langle X_{s-}, \alpha\rangle \text{d} s$ . In that case, the age of the dying individual is distributed according to the probability measure $\langle X_{s-}, \alpha\rangle ^{-1}\alpha(x)X_{s-}(\text{d} x)$ and is realized as $A_\alpha(X_{s-},y)$ , where $y\in (0, 1]$ is taken according to the uniform distribution by the Poisson random measure. The number of offspring of the individual takes the value $z\in \mathbb{N}$ with probability $p(A_\alpha(X_{s-},y),z)$ and contributes to the number $X_t(x)$ if and only if $t-s\le x$ , which is recorded by the second term. The death of the individual affects $X_t(x)$ if and only if $A_\alpha(X_{s-},y) + t-s\le x$ , which is recorded by the third term.

Let $\zeta_a(x) = \textbf{1}_{\{a\le x\}}$ for $a, x\in \mathbb{R}$ . Given a function f on $\mathbb{R}$ define $f\circ\theta_t(x)= f(x + t)$ for $x, t\in \mathbb{R}$ . Then we may rewrite (3.1) equivalently, for $t\ge 0$ and $x\ge 0$ , as

(3.2) \begin{align} X_t(x) = \, & X_0\circ\theta_{-t}(x) + \int_0^t\int_0^{\langle X_{s-},\alpha\rangle }\int_0^1\int_{\mathbb{N}}\int_0^{p(A_\alpha(X_{s-},y),z)} \big[z \zeta_0\circ\theta_{s-t}(x)\nonumber\\[4pt]& - \zeta_{A_\alpha(X_{s-},y)}\circ\theta_{s-t}(x)\big] M(\text{d} s,\text{d} u,\text{d} y,\text{d} z,\text{d} v). \end{align}

A pathwise solution to (3.2) is constructed by unscrambling the equation as follows. Let $\tau_0 = 0$ . Given $\tau_{k-1}\ge 0$ and $X_{\tau_{k-1}}\in \mathfrak{D}\big(\mathbb{R}_+\big)$ , we first define $\tau_k = \tau_{k-1} + \inf\big\{t>0 \,:\, M\big((\tau_{k-1},\tau_{k-1}+t]\times (0,\langle X_{\tau_{k-1}}, \alpha\rangle ]\times H_k\big)> 0\big\}$ , where $H_k= \big\{(y,z,v) \,:\, y\in (0, 1], z\in\mathbb{N},0< v\le p\big(A_\alpha\big(X_{\tau_{k-1}},y\big),z\big)\big\}$ , and

(3.3) \begin{eqnarray} X_t(x)= X_{\tau_{k-1}}\circ\theta_{\tau_{k-1}-t}(x), \qquad { \tau_{k-1}\le t< \tau_k,\ x\ge 0.}\end{eqnarray}

Then we define

(3.4) \begin{eqnarray} X_{\tau_k}(x)= X_{\tau_k-}(x) + z_k\zeta_0(x) - \zeta_{A_\alpha(X_{\tau_k-},y_k)}(x), \qquad x\ge 0,\end{eqnarray}

where $X_{\tau_k-}(x)= X_{\tau_{k-1}}\circ\theta_{\tau_{k-1}-\tau_k}(x)$ , and $(u_k,y_k,z_k,v_k)\in (0,\infty)\times (0, 1]\times \mathbb{N}\times (0, 1]$ is the point such that $(\tau_k, u_k,y_k,z_k,v_k)\in \text{supp}(M)$ . Since $A_\alpha\big(X_{\tau_k-},y_k\big)\in \text{supp}\big(X_{\tau_k-}\big)$ , we have $X_{\tau_k}\in \mathfrak{D}\big(\mathbb{R}_+\big)$ . Equation (3.4) means that at time $\tau_k$ an individual at age $A_\alpha\big(X_{\tau_k-},y_k\big)$ dies and gives birth to $z_k$ offspring with starting age $0\in \mathbb{R}_+$ .

It is clear that (3.3) and (3.4) uniquely determine the behavior of the trajectory $t\mapsto X_t$ on the time intervals $[\tau_{k-1},\tau_k]$ , $k= 1,2,\ldots$ Let $\tau= \lim_{k\to \infty}\tau_k$ and let $X_t= \infty$ for $t\ge \tau$ . Then $\{X_t \,:\, t \ge 0\}$ is the pathwise-unique solution to (3.2) up to the lifetime $\tau$ . More precisely, the equations hold almost surely with t replaced by $t\land \tau_k$ for every $k= 1,2,\ldots$ Let $n(t)= \sup\{k\ge 0 \,:\, \tau_k\le t\}$ for $t\ge 0$ , and $\beta= \|\alpha \partial_zg(\cdot,1{-})\|< \infty$ .

Lemma 3.2. Suppose that $\mathbb{E}[X_0(\infty)]< \infty$ . Then, for any $k\ge 1$ ,

(3.5) \begin{eqnarray}\mathbb{E}\big[{\textstyle \sup_{0\le s\le t\land\tau_k}} X_s(\infty)\big]\le\mathbb{E}[X_0(\infty)]\text{e}^{\beta t}, \qquad t\ge 0.\end{eqnarray}

Proof.Recall that $X_t(\infty)= \lim_{x\to \infty} X_t(x)= X_t\big(\mathbb{R}_+\big)$ . Let $\eta_i= \inf\{s\ge 0 \,:\, X_s(\infty)\ge i\}$ for $i\ge 1$ . It is clear that $\lim_{i\to \infty}\eta_i= \tau$ . Let $\zeta_{i,k}= \eta_i\land \tau_k$ . In view of (3.2), we have

\begin{equation*} X_t(\infty)= X_0(\infty) + \int_0^t\int_0^{\langle X_{s-},\alpha\rangle } \int_0^1\int_{\mathbb{N}}\int_0^{p\big(A_\alpha(X_{s-},y),z\big)} (z-1) M(\text{d} s,\text{d} u,\text{d} y,\text{d} z,\text{d} v). \end{equation*}

It follows that

\begin{align*} \mathbb{E}\big[{\textstyle \sup_{0\le s\le t\land\zeta_{i,k}}} X_s(\infty)\big] & \le \mathbb{E}[X_0(\infty)] + \sum_{z\in \mathbb{N}} \mathbb{E}\bigg[\int_0^{t\land\zeta_{i,k}} \!\! \langle X_{s-},\alpha\rangle \text{d} s \!\! \int_0^1 \!\! p(A_\alpha(X_{s-},y),z) z \,\text{d} y\bigg] \\ & = \mathbb{E}[X_0(\infty)] + \sum_{z\in \mathbb{N}} \mathbb{E}\bigg[\int_0^{t\land\zeta_{i,k}} \text{d} s \int_{\mathbb{R}_+}p(y,z)z \alpha(y) \, \text{d} X_{s-}(y)\bigg] \\ & = \mathbb{E}[X_0(\infty)] + \mathbb{E}\bigg[\int_0^{t\land\zeta_{i,k}} \text{d} s \int_{\mathbb{R}_+} \alpha(y)\partial_zg(y,1{-}) \, \text{d} X_{s-}(y)\bigg] \\ & \le \mathbb{E}[X_0(\infty)] + \beta\mathbb{E}\bigg[\int_0^{t\land\zeta_{i,k}} X_{s-}(\infty) \, \text{d} s\bigg]. \end{align*}

Clearly, we have $X_{s-}(\infty)\le i$ for $0< s\le t\land\zeta_{i,k}$ . Then $\mathbb{E}\big[{\sup}_{0\le s\le t\land\zeta_{i,k}}X_s(\infty)\big]$ is locally bounded in $t\ge 0$ . Since the trajectory $s\mapsto X_s(\infty)$ has at most countably many jumps, it follows that

\begin{align*} \mathbb{E}\big[{\textstyle \sup_{0\le s\le t\land\zeta_{i,k}}} X_s(\infty)\big] & \le \mathbb{E}[X_0(\infty)] + \beta\mathbb{E}\bigg[\int_0^{t\land\zeta_{i,k}} X_s(\infty) \, \text{d} s\bigg] \cr & \le \mathbb{E}[X_0(\infty)] + \beta\int_0^t \mathbb{E}[X_{s\land\zeta_{i,k}}(\infty)] \, \text{d} s \cr & \le \mathbb{E}[X_0(\infty)] + \beta\int_0^t { \mathbb{E}\big[{\textstyle \sup_{0\le r\le s\land\zeta_{i,k}}} X_r(\infty)\big]} \, \text{d} s. \end{align*}

By Gronwall’s inequality we have $\mathbb{E}\big[{\sup}_{0\le s\le t\land\zeta_{i,k}}X_s(\infty)\big] \le \mathbb{E}[X_0(\infty)]\text{e}^{\beta t}$ . Then, letting $i\to \infty$ and using Fatou’s lemma, we obtain (3.5).

Proposition 3.1. Suppose that $\mathbb{E}[X_0(\infty)]< \infty$ . Then $\mathbb{P}\{\tau= \infty\}= 1$ and

(3.6) \begin{eqnarray}\mathbb{E}[n(t)]\le \|\alpha\|\mathbb{E}[X_0(\infty)] \int_0^t \text{e}^{\beta s} \, \text{d} s, \qquad t\ge 0.\end{eqnarray}

Proof.By (3.2) and monotone convergence we have

\begin{align*} \mathbb{E}[n(t)] & = \lim_{k\to \infty}\mathbb{E}\bigg[\int_0^{t\land\tau_k}\int_0^{\langle X_{s-},\alpha\rangle } \int_0^1\int_{\mathbb{N}}\int_0^{p(A_\alpha(X_{s-},y),z)} M(\text{d} s,\text{d} u,\text{d} y,\text{d} z,\text{d} v)\bigg] \\ & = \lim_{k\to \infty}\mathbb{E}\bigg[\sum_{z\in \mathbb{N}} \int_0^{t\land\tau_k} \langle X_{s-},\alpha\rangle \, \text{d} s \int_0^1 p(A_\alpha(X_{s-},y),z) \, \text{d} y\bigg] \\ & = \lim_{k\to \infty}\mathbb{E}\bigg[\sum_{z\in \mathbb{N}} \int_0^{t\land\tau_k} \text{d} s \int_{\mathbb{R}_+} p(y,z)\alpha(y) \, \text{d} X_{s-}(y)\bigg] \\ & \le \lim_{k\to \infty}\|\alpha\|\mathbb{E}\bigg[\int_0^{t\land\tau_k} X_s(\infty) \, \text{d} s\bigg] = \lim_{k\to \infty}\|\alpha\|\int_0^t \mathbb{E}[X_{s\land\tau_k}(\infty)] \, \text{d} s. \end{align*}

Then (3.6) follows by (3.5). In particular, we have $\mathbb{P}\{\tau> t\}= \mathbb{P}\{n(t)< \infty\}= 1$ for every $t\ge 0$ , which implies $\mathbb{P}\{\tau= \infty\}= 1$ .

Proposition 3.2. Suppose $\mathbb{E}[X_0(\infty)]< \infty$ . Then $\mathbb{E}\big[{\sup}_{0\le s\le t}X_s(\infty)\big] \le \mathbb{E}[X_0(\infty)]\text{e}^{\beta t}$ , $t\ge 0$ .

Proof.Since $\mathbb{P}\{\tau= \infty\}= 1$ by Proposition 3.1, we obtain the result from (3.5) by using monotone convergence.

By Proposition 3.1 the solution of (3.1) or (3.2) has infinite lifetime and determines a measure-valued strong Markov process $\{X_t \,:\, t\ge 0\}$ . The following propositions give some useful characterization of the process.

Proposition 3.3. For any $t\ge 0$ and $f \in B\big(\mathbb{R}_+\big)$ ,

(3.7) \begin{align} \langle X_t,f\rangle =\, & \langle X_0,f\circ \theta_t\rangle + \int_0^t\int_0^{\langle X_{s-},\alpha\rangle }\int_0^1 \int_{\mathbb{N}} \int_0^{p(A_\alpha(X_{s-},y),z)} \big[z f\circ\theta_{t-s}(0)\nonumber\\& - f\circ\theta_{t-s}(A_\alpha(X_{s-},y))\big] M(\text{d} s,\text{d} u,\text{d} y,\text{d} z,\text{d} v). \end{align}

Proof.Let $C^1_0\big(\mathbb{R}_+\big)$ denote the subspace of $C^1\big(\mathbb{R}_+\big)$ consisting of functions vanishing at infinity. For any fixed integer $n\ge 1$ , let $x_i = in/2^n$ with $i = 0, 1, \ldots, 2^n$ . By (3.2), almost surely for any $f \in C^1_0\big(\mathbb{R}_+\big)$ ,

\begin{align*} \sum_{i = 1}^{2^n}f^{\prime}(x_i)X_t(x_i) & - \sum_{i = 1}^{2^n}f^{\prime}(x_i)X_0\circ\theta_{-t}(x_i) \\& = \int_0^t\int_0^{\langle X_{s-}, \alpha\rangle }\int_0^1\int_{\mathbb{N}}\int_0^{p(A_\alpha(X_{s-},y),z)}\sum_{i = 1}^{2^n}f^{\prime}(x_i)\big[z \zeta_0\circ\theta_{s-t}(x_i) \\ &\quad - \zeta_{A_\alpha(X_{s-},y)}\circ\theta_{s- t}(x_i)\big] M(\text{d} s,\text{d} u,\text{d} y,\text{d} z,\text{d} v). \end{align*}

Then we multiply the above equation by $2^{-n}$ and let $n\to \infty$ to get, almost surely,

(3.8) \begin{align} \int_0^\infty f^{\prime}(x)X_t(x) \, \text{d} x & - \int_0^\infty f^{\prime}(x)X_0\circ\theta_{-t}(x) \, \text{d} x \nonumber \\ & = \int_0^t\int_0^{\langle X_{s-}, \alpha\rangle }\int_0^1\int_{\mathbb{N}}\int_0^{p(A_\alpha(X_{s-},y),z)} \bigg\{\!\int_0^\infty f^{\prime}(x) \big[z\zeta_0\circ\theta_{s-t}(x) \nonumber \\ & \quad - \zeta_{A_\alpha(X_{s-},y)}\circ\theta_{s-t}(x)\big] \, \text{d} x\bigg\} M(\text{d} s,\text{d} u,\text{d} y,\text{d} z,\text{d} v) \nonumber \\ & = -\int_0^t\int_0^{\langle X_{s-}, \alpha\rangle }\int_0^1\int_{\mathbb{N}}\int_0^{p(A_\alpha(X_{s-},y), z)} \big[z f\circ\theta_{t-s}(0) \nonumber \\ & \quad - f\circ\theta_{t-s}(A_\alpha(X_{s-},y))\big] M(\text{d} s,\text{d} u,\text{d} y,\text{d} z,\text{d} v). \end{align}

By integration by parts we have $\langle X_t, f\rangle = - \int_0^\infty f^{\prime}(x)X_t(x) \, \text{d} x$ . From this and (3.8) we see that (3.7) holds for any $f \in C_0^1\big(\mathbb{R}_+\big)$ . Then the relation also holds for any $f\in B\big(\mathbb{R}_+\big)$ by a monotone class argument.

Proposition 3.4. For any $t\ge 0$ and $f\in C^1\big(\mathbb{R}_+\big)$ ,

(3.9) \begin{align} \langle X_t, f\rangle =\, & \langle X_0, f\rangle + \int_0^t \langle X_{s-}, f^{\prime}\rangle \, \text{d} s + \int_0^t \int_0^{\langle X_{s-},\alpha\rangle }\int_0^1\int_{\mathbb{N}}\int_0^{p(A_\alpha(X_{s-},y),z)}\big[z f(0)\nonumber\\[4pt]& - f(A_\alpha(X_{s-},y))\big] M(\text{d} s,\text{d} u,\text{d} y,\text{d} z,\text{d} v). \qquad\end{align}

Proof.For $n\ge 1$ we consider a partition $\Delta_n = \{0 = t_0 < t_1 < \cdots < t_n = t\}$ of [0, t]. Notice that $ \partial_t(f\circ \theta_t)(x) = f^{\prime}(x+t).$ By (3.7) we have

\begin{align*} \langle X_t, f\rangle & = \langle X_0, f\rangle + \sum_{i=1}^n \left[\big\langle X_{t_i}, f\big\rangle - \big\langle X_{t_{i-1}}, f\circ\theta_{t_{i} - t_{i-1}}\big\rangle \right] \\ & \quad + \sum_{i=1}^n \left[\big\langle X_{t_{i-1}}, f\circ\theta_{t_{i} - t_{i-1}}\big\rangle - \big\langle X_{t_{i-1}},f\big\rangle \right] \\ & = \langle X_0, f\rangle + \sum_{i=1}^n \int_{t_{i-1}}^{t_i} \int_0^{\langle X_{s-},\alpha\rangle }\int_0^1\int_{\mathbb{N}}\int_0^{p(A_\alpha(X_{s-},y), z)} \big[z\, f\circ\theta_{t_i-s}(0) \\ & \quad - f\circ\theta_{t_i - s}(A_\alpha(X_{s-},y))\big] M(\text{d} s,\text{d} u,\text{d} y,\text{d} z,\text{d} v) \\ & \quad + \sum_{i=1}^n \int_{t_{i-1}}^{t_i}\big\langle X_{t_{i-1}}, f^{\prime}\circ\theta_{s - t_{i-1}}\big\rangle \, \text{d} s. \end{align*}

By letting $|\Delta_n|\,:\!=\, \max_{1\le i\le n} (t_i-t_{i-1})\to 0$ and using the right continuity of $s\to X_s$ and the continuity of $s\to f\circ\theta_s$ , we obtain (3.9).

Proposition 3.5. For $f, G \in C^1\big(\mathbb{R}_+\big)$ let $G_f(\mu)= G(\langle \mu,f\rangle )$ , and let

\begin{align*} \mathscr{L}_0G_f(\mu) & = \big\langle \mu,f^{\prime}\big\rangle G^{\prime}(\langle \mu,f\rangle ) \\ & \quad - \sum_{z \in \mathbb{N}}\int_{\mathbb{R}_+} \alpha(y)p(y,z)\big[G(\langle \mu, f\rangle ) - G(\langle \mu,f\rangle +zf(0)-f(y))\big] \mu(\text{d} y). \end{align*}

Then we have

(3.10) \begin{eqnarray} G_f(X_t) = G_f(X_0) + \int_0^t { \mathscr{L}_0G_f(X_s)} \, \text{d} s + {mart.} \end{eqnarray}

Proof.Let $\tilde{M}$ denote the compensated measure of M. Since the process $s\mapsto X_s$ has at most countably many jumps, by Proposition 3.4 and Itô’s formula we have

\begin{align*} G(\langle X_t,f\rangle ) & = G(\langle X_0,f\rangle ) + \int_0^t G^{\prime}(\langle X_s,f\rangle ) \langle X_s, f^{\prime}\rangle \text{d} s \\ & \quad - \int_0^t \int_0^{\langle X_{s-},\alpha\rangle }\int_0^1\int_{\mathbb{N}}\int_0^{p(A_\alpha(X_{s-},y),z)} \big[G(\langle X_{s-},f\rangle ) \\ & \quad - G(\langle X_{s-},f\rangle + zf(0) - f(A_\alpha(X_{s-},y)))\big] M(\text{d} s,\text{d} u,\text{d} y,\text{d} z,\text{d} v) \\ & = G(\langle X_0,f\rangle ) + \int_0^t G^{\prime}(\langle X_s, f\rangle ) \langle X_s, f^{\prime}\rangle \, \text{d} s - M_t^G(f) \\ & \quad - \int_0^t \text{d} s \int_{\mathbb{R}_+} \alpha(y)\sum_{z\in \mathbb{N}}p(y,z) \big[G(\langle X_{s-},f\rangle ) \\ & \quad - G(\langle X_{s-},f\rangle + zf(0) - f(y))\big] X_{s-}(\text{d} y) \\ & = G(\langle X_0,f\rangle ) + \int_0^t \mathscr{L}_0G_f(X_s) \, \text{d} s - M_t^G(f), \end{align*}

where

\begin{align*} M_t^G(f) = \int_0^t \int_0^{\langle X_{s-},\alpha\rangle }\int_0^1&\int_{\mathbb{N}}\int_0^{p(A_\alpha(X_{s-},y), z)}\big[G(\langle X_{s-}, f\rangle ) \\ &\ \ - G(\langle X_{s-}, f\rangle + z f(0) - f(A_\alpha(X_{s-},y)))\big] \tilde{M}(\text{d} s,\text{d} u,\text{d} y,\text{d} z,\text{d} v). \end{align*}

By Proposition 3.2 we can check that $\big\{M_t^G(f) \,:\, t\ge 0\big\}$ is a martingale.

Theorem 3.1. The measure-valued process $\{X_t \,:\, t \ge 0\}$ defined by the stochastic equation (3.1) or (3.2) is an $(\alpha, g)$ -age-structured branching process.

Proof.Recall that $\{X_t \,:\, t \ge 0\}$ is a càdlàg process. Let $e_f(\mu)= \text{e}^{-\langle \mu,f\rangle }$ and let $\mathscr{L}_0$ be defined as in Proposition 3.5. It is elementary to see that

\begin{align*} \mathscr{L}_0e_f(\mu) & = -\big\langle \mu,f^{\prime}\big\rangle \text{e}^{-\langle \mu,f\rangle } - \text{e}^{-\langle \mu,f\rangle }\int_{\mathbb{R}_+} \alpha(y)\bigg[1 - \text{e}^{f(y)}\sum_{z\in \mathbb{N}} p(y,z)\text{e}^{-zf(0)}\bigg] \mu(\text{d} y) \\ & = - \text{e}^{-\langle \mu,f\rangle }\big\langle \mu,f^{\prime}\big\rangle - \text{e}^{-\langle \mu,f\rangle }\int_{\mathbb{R}_+} \alpha(y)\big(1 - \text{e}^{f(y)} g\big(y,\text{e}^{-f(0)}\big)\big) \mu(\text{d} y) . \end{align*}

Let $\mathscr{F}_t= \sigma\{X_s \,:\, 0\le s\le t\}$ . By (2.4), (3.10), and the mean-value theorem, we have

\begin{align*} \text{e}^{-\big\langle X_t,u_{T-t}f\big\rangle } & = \text{e}^{-\big\langle X_0,u_Tf\big\rangle } + \sum_{i=0}^\infty \bigg[\text{e}^{-\big\langle X_{t\wedge(i+1)/k},u_{T-t\wedge i/k}f\big\rangle } - \text{e}^{-\big\langle X_{t\wedge i/k},u_{T-t\wedge i/k}f\big\rangle }\bigg] \\ & \quad + \sum_{i=0}^\infty \bigg[\text{e}^{-\big\langle X_{t\wedge (i+1)/k},u_{T-t\wedge (i+1)/k}f\big\rangle } - \text{e}^{-\big\langle X_{t\wedge (i+1)/k},u_{T-t\wedge i/k}f\big\rangle }\bigg] \end{align*}
\begin{align*} & = \text{e}^{-\big\langle X_0,u_Tf\big\rangle } - \sum_{i=0}^\infty\int_{t\wedge i/k}^{t\wedge (i+1)/k} \text{e}^{-\big\langle X_s,u_{T-t\wedge i/k}f\big\rangle } \bigg[\big\langle X_s,\partial_xu_{T-t\wedge i/k}f\big\rangle \\ & \quad + \int_{\mathbb{R}_+} \alpha(y)\big[1 - \text{e}^{u_{T-t\wedge i/k}f(y)}g\big(y,\text{e}^{-u_{T-t\wedge i/k}f(0)}\big)\big] X_s(\text{d} y)\bigg] \, \text{d} s \\ & \quad + M_k(t) + \sum_{i=0}^\infty \text{e}^{-\xi_k(t)}\big\langle X_{t\wedge (i+1)/k}, u_{T-t\wedge i/k}f-u_{T-t\wedge (i+1)/k}f\big\rangle \\ & = \text{e}^{-\big\langle X_0,u_Tf\big\rangle } - \sum_{i=0}^\infty\int_{t\wedge i/k}^{t\wedge (i+1)/k} \text{e}^{-\big\langle X_s,u_{T-t\wedge i/k}f\big\rangle } \bigg[\big\langle X_s,\partial_xu_{T-t\wedge i/k}f\big\rangle \\ & \quad + \int_{\mathbb{R}_+} \alpha(y)\big[1 - \text{e}^{u_{T-t\wedge i/k}f(y)}g\big(y,\text{e}^{-u_{T-t\wedge i/k}f(0)}\big)\big] X_s(\text{d} y)\bigg] \, \text{d} s \\ & \quad + M_k(t) + \sum_{i=0}^\infty\int_{t\wedge i/k}^{t\wedge (i+1)/k} \text{e}^{-\xi_k(t)}\bigg[\big\langle X_{t\wedge (i+1)/k}, \partial_xu_{T-s}f\big\rangle \\ & \quad + \int_{\mathbb{R}_+} \alpha(y)\big[1 - \text{e}^{u_{T-s}f(y)}g\big(y,\text{e}^{-u_{T-s}f(0)}\big)\big] X_{t\wedge (i+1)/k}(d y)\bigg] \, \text{d} s, \end{align*}

where $t\mapsto M_k(t)$ is an $(\mathscr{F}_t)$ -martingale and

\begin{eqnarray*} \big\langle X_{t\wedge (i+1)/k}, u_{T-t\wedge i/k}f\big\rangle \le \xi_k(t)\le \big\langle X_{t\wedge (i+1)/k},u_{T-t\wedge (i+1)/k}f\big\rangle \end{eqnarray*}

or

\begin{eqnarray*} \big\langle X_{t\wedge (i+1)/k}, u_{T-t\wedge (i+1)/k}f\big\rangle \le \xi_k(t)\le \big\langle X_{t\wedge (i+1)/k},u_{T-t\wedge i/k}f\big\rangle . \end{eqnarray*}

By letting $k\to \infty$ we see that $t\mapsto \text{e}^{-\langle X_t,u_{T-t}f\rangle }$ is an $(\mathscr{F}_t)$ -martingale. In particular, we have $\mathbb{E}\big[\text{e}^{-\langle X_T,f\rangle } \mid \mathscr{F}_t\big]= \text{e}^{-\langle X_t,u_{T-t}f\rangle }$ , $T\ge t\ge 0$ . Then $\{X_t \,:\, t \ge 0\}$ is a Markov process with transition semigroup $(Q_t)_{t\ge 0}$ given by (2.3).

A calculation of the generator for an age-structured birth–death process was given in [Reference Bose4, (3.1)]; see also [Reference Bose and Kaj5]. A stochastic equation for a similar model was proposed in [Reference Tran28, (2.5)], which assumed that the death rate of an individual may depend on the whole population and also calculated the generator of the model.

4. The branching process with immigration

In this section we introduce an age-structured branching process with immigration and discuss its ergodicity. Let $\psi$ be a functional on $B\big(\mathbb{R}_+\big)^+$ given by $\psi(f) = \int_{\mathfrak{N}\big(\mathbb{R}_+\big)^\circ} (1 - \text{e}^{-\langle \nu, f\rangle })L (\text{d}\nu)$ , $f\in B\big(\mathbb{R}_+\big)^+$ , where $L(\text{d}\nu)$ is a finite measure on $\mathfrak{N}\big(\mathbb{R}_+\big)^\circ \,:\!=\, \mathfrak{N}\big(\mathbb{R}_+\big)\setminus 0$ , and 0 denotes the null measure.

A Markov process $Y= (Y_t \,:\, t\ge 0)$ with state space $\mathfrak{N}\big(\mathbb{R}_+\big)$ is called an $(\alpha,g, \psi)$ -age-structured branching process with immigration if it has the transition semigroup $(P_t)_{t\ge 0}$ defined by

(4.1) \begin{eqnarray} \int_{\mathfrak{N}\big(\mathbb{R}_+\big)}\text{e}^{-\langle \nu, f\rangle}P_t(\sigma, \text{d}\nu) = \exp\!\bigg\{{-}\langle \sigma, u_t\,f\rangle -\int_0^t\psi(u_sf) \, \text{d} s\bigg\},\end{eqnarray}

where $u_t\,f(x)$ is the unique solution to (2.2). Such a process is characterized by the properties (i)–(iii) given in Section 2, along with the following:

  1. (iv) The immigrants come according to a Poisson random measure on $(0,\infty)\times \mathfrak{N}\big(\mathbb{R}_+\big)^\circ$ with intensity $d sL(d\nu)$ .

Suppose that $(\Omega, \mathscr{F}, \mathscr{F}_t, \mathbb{P})$ is a filtered probability space satisfying the usual hypotheses. Let $M(\text{d} t,\text{d} u,\text{d} y,\text{d} z,\text{d} v)$ be a Poisson random measure as in Section 3, and let $N(\text{d} t,\text{d} \nu)$ be an $(\mathscr{F}_t)$ -Poisson random measure on $(0, \infty)\times \mathfrak{N}\big(\mathbb{R}_+\big)^\circ$ with intensity $\text{d} t L(\text{d}\nu)$ . We assume that the two random measures are independent of each other. Consider the stochastic integral equation

(4.2) \begin{align} Y_t(x) & = Y_0\circ\theta_{-t}(x) + \int_0^t\int_0^{\langle Y_{s-}, \alpha\rangle }\int_0^1\int_{\mathbb{N}} \int_0^{p(A_\alpha(Y_{s-},y), z)} \big[z\zeta_0 \nonumber \\ & \quad - \zeta_{A_\alpha(Y_{s-},y)}\big]\circ\theta_{s-t}(x) M(\text{d} s,\text{d} u,\text{d} y,\text{d} z,\text{d} v) \nonumber \\ & \quad + \int_0^t\int_{\mathfrak{N}\big(\mathbb{R}_+\big)^\circ} \nu\circ\theta_{s-t}(x)N(\text{d} s,\text{d} \nu).\end{align}

Here, the first two terms on the right-hand side are as explained for (3.1) and the last term represents the immigration. A pathwise-unique solution to (4.2) is constructed as follows. Let $\sigma_0 = 0$ . Given $\sigma_{k-1}$ , we define $\sigma_k = \sigma_{k-1} + \inf\big\{t>0 \,:\, N\big(\big(\sigma_{k-1},\sigma_{k-1}+t\big]\times N\big(\mathbb{R}_+^\circ\big)\big)> 0\big\}$ and $Y_t(x) = X_{k,t-\sigma_{k-1}}(x)$ , $\sigma_{k-1} \le t < \sigma_k$ , where $\{X_{k,t}(x) \,:\, t\ge 0\}$ is the pathwise-unique solution to the following equation:

\begin{align*} X_t(x) =& Y_{\sigma_{k-1}}\circ\theta_{-t}(x) + \int_0^t\int_0^{\langle X_{s-},\alpha\rangle }\int_0^1\int_{\mathbb{N}} \int_0^{p(A_\alpha(X_{s-},y), z)} \big[z \zeta_0\circ\theta_{s-t}(x) \\& - \zeta_{A_\alpha(X_{s-},y)}\circ\theta_{s-t}(x)\big] M(\sigma_{k-1}+\text{d} s,\text{d} u,\text{d} y,\text{d} z,\text{d} v).\end{align*}

Then we define $Y_{\sigma_k}(x) = Y_{\sigma_k-}(x) + \int_{\{\sigma_k\}} \int_{\mathfrak{N}\big(\mathbb{R}_+\big)^\circ} \nu(x) N(\text{d} s, \text{d}\nu)$ . It is easy to see that $\lim_{k\to \infty}\sigma_k = \infty$ and $\{Y_t \,:\, t \ge 0\}$ is the pathwise-unique solution to (4.2).

The solution to (4.2) determines a measure-valued strong Markov process $\{Y_t \,:\, t \ge 0\}$ with state space $\mathfrak{N}\big(\mathbb{R}_+\big)$ . We omit the proofs of some of the following results since the arguments are similar to those for the corresponding results in Section 3.

Proposition 4.1. For any $f \in B\big(\mathbb{R}_+\big)^+$ ,

\begin{align*} \langle Y_t, f\rangle & = \langle Y_0, f\circ \theta_t\rangle + \int_0^t\int_0^{\langle Y_{s-}, \alpha\rangle }\int_{[0, 1]}\int_{\mathbb{N}}\int_0^{p(A_\alpha(Y_{s-},y), z)}\big[z f\circ\theta_{t - s}(0) \\ & \quad - f\circ\theta_{t - s}(A_\alpha(Y_{s-},y))\big] M(\text{d} s,\text{d} u,\text{d} y,\text{d} z,\text{d} v) \\ & \quad + \int_0^t\int_{\mathfrak{N}\big(\mathbb{R}_+\big)^\circ} \langle \nu, f\circ\theta_{t - s}\rangle N(\text{d} s, \text{d} \nu). \end{align*}

Proposition 4.2. For any $t \ge 0$ and $f \in C^1\big(\mathbb{R}_+\big)^+$ ,

(4.3) \begin{align} \langle Y_t, f\rangle & = \langle Y_0, f\rangle + \int_0^t \langle Y_{s-}, f^{\prime}\rangle \, \text{d} s + \int_0^t\int_0^{\langle Y_{s-}, \alpha\rangle }\int_{[0, 1]}\int_{\mathbb{N}}\int_0^{p(A_\alpha(Y_{s-},y),z)}\big[z f(0) \nonumber \\ & \quad - f(A_\alpha(Y_{s-},y))\big] M(\text{d} s,\text{d} u,\text{d} y,\text{d} z,\text{d} v) \nonumber \\ & \quad + \int_0^t\int_{\mathfrak{N}\big(\mathbb{R}_+\big)^\circ} \langle \nu, f\rangle N(\text{d} s, \text{d} \nu). \end{align}

Proposition 4.3. For any $f, G\in C^1\big(\mathbb{R}_+\big)$ , let $G_f(\mu)= G(\langle \mu,f\rangle )$ and

\begin{align*} \mathscr{L}\,G_f(\mu) & = \big\langle \mu,f^{\prime}\big\rangle G^{\prime}(\langle \mu,f\rangle ) - \int_{\mathbb{R}_+} \alpha(y) \sum_{z \in \mathbb{N}}p(y,z)\big[G(\langle \mu, f\rangle ) \\ & \quad - G\big(\langle \mu, f\rangle + zf(0) - f(y)\big)\big] \mu(\text{d} y) \\ & \quad + \int_{\mathfrak{N}\big(\mathbb{R}_+\big)^\circ}\big[G(\langle \mu, f\rangle + \langle \nu, f\rangle ) - G(\langle \mu, f\rangle )\big]L(\text{d}\nu). \end{align*}

Then $G(\langle Y_t, f\rangle ) = G(\langle Y_0, f\rangle ) + \int_0^t \mathscr{L}\,G_f(Y_s) \, \text{d} s + {mart.}$

Proof.Let $\tilde{M}$ and $\tilde{N}$ denote the compensated measure of M and N, respectively. By (4.3) and Itô’s formula we have

\begin{align*} G(\langle Y_t,f\rangle ) & = G(\langle Y_0,f\rangle ) + \int_0^t G^{\prime}(\langle Y_s, f\rangle )\langle Y_s, f^{\prime}\rangle \, \text{d} s \\ & \quad + \int_0^t\int_0^{\langle Y_{s-}, \alpha\rangle }\int_0^1\int_{\mathbb{N}}\int_0^{p(A_\alpha(Y_{s-},y), z)}\big[G\big(\langle Y_{s-},f\rangle + zf(0) \\ & \quad - f(A_\alpha(Y_{s-},y))\big) - G(\langle Y_{s-},f\rangle ) \big]M(\text{d} s,\text{d} u,\text{d} y,\text{d} z,\text{d} v) \\ & \quad + \int_0^t\int_{\mathfrak{N}\big(\mathbb{R}_+\big)^\circ}\big[G\big(\langle Y_{s-},f\rangle + \langle \nu, f\rangle \big) - G\big(\langle Y_{s-},f\rangle \big)\big]N(\text{d} s, \text{d} \nu) \\ & = G(\langle Y_0,f\rangle ) + \int_0^t G^{\prime}(\langle Y_s, f\rangle )\langle Y_s, f^{\prime}\rangle \, \text{d} s + N_t^G(f) \\ & \quad + \int_0^t \text{d} s\int_{\mathbb{R}_+} \alpha(y)\sum_{z\in \mathbb{N}}p(y,z) \big[G\big(\langle Y_{s-},f\rangle + zf(0) - f(y)\big) \\ & \quad - G(\langle Y_{s-},f\rangle ) \big]Y_{s-}(\text{d} y) \\ & \quad + \int_0^t \text{d} s\int_{\mathfrak{N}\big(\mathbb{R}_+\big)^\circ}\big[G\big(\langle Y_{s-},f\rangle + \langle \nu, f\rangle \big) - G(\langle Y_{s-},f\rangle )\big]L(\text{d}\nu), \end{align*}

where

\begin{align*} N_t^G(f) & = \int_0^t\int_0^{\langle Y_{s-}, \alpha\rangle }\int_0^1\int_{\mathbb{N}}\int_0^{p(A_\alpha(Y_{s-},y),z)} \big[G(\langle Y_{s-},f\rangle + zf(0) - f(A_\alpha(Y_{s-},y))) \\ & \qquad\qquad\qquad\qquad\qquad\qquad\qquad\quad\ \ - G(\langle Y_{s-},f\rangle ) \big]\tilde{M}(\text{d} s,\text{d} u,\text{d} y,\text{d} z,\text{d} v) \\ & \quad + \int_0^t\int_{\mathfrak{N}\big(\mathbb{R}_+\big)^\circ}\big[G\big(\langle Y_{s-},f\rangle + \langle \nu, f\rangle \big) - G\big(\langle Y_{s-},f\rangle \big)\big]\tilde{N}(\text{d} s, \text{d} \nu). \end{align*}

A first moment estimate can check that $\big\{N_t^G(f) \,:\, t\ge 0\big\}$ is a martingale.

Theorem 4.1. The measure-valued process $\{Y_t \,:\, t \ge 0\}$ defined by (4.2) is an $(\alpha, g, \psi)$ -age-structured branching process with immigration.

Proof.Let $\mathscr{F}_t= \sigma\{Y_s \,:\, 0\le s\le t\}$ . By a modification of the proof of Theorem 3.1, we can see that $t\mapsto \exp\!\Big\{{-}\langle Y_t, u_{T-t}f\rangle - \int_0^{T-t}\psi(u_sf) \, \text{d} s\Big\}$ is an $(\mathscr{F}_t)$ -martingale. Then $\{Y_t \,:\, t\ge 0\}$ is a Markov process with transition semigroup $(P_t)_{t\ge 0}$ defined by (4.1), which completes the proof.

A necessary and sufficient condition for the ergodicity of the $(\alpha, g, \psi)$ -age-structured branching process with immigration is given in the next theorem.

Theorem 4.2. Suppose that $c_*= \inf_{y\ge 0}\alpha(y)[1-\partial_zg(y,1{-})]> 0$ . Then $P_t(\sigma, \cdot)$ converges as $t\to \infty$ to a probability measure $\eta$ on $\mathfrak{N}\big(\mathbb{R}_+\big)$ for every $\sigma \in \mathfrak{N}\big(\mathbb{R}_+\big)$ if and only if

(4.4) \begin{eqnarray} \int_{\mathfrak{N}\big(\mathbb{R}_+\big)} \textbf{1}_{\{\langle \nu,1\rangle \ge 1\}}\log\langle \nu,1\rangle L(\text{d}\nu) < \infty. \end{eqnarray}

In this case, the Laplace transform of $\eta$ is given by

(4.5) \begin{eqnarray} \int_{\mathfrak{N}\big(\mathbb{R}_+\big)} \text{e}^{-\langle \nu, f\rangle } \eta(\text{d}\nu) = \exp\!\bigg\{{-} \int_0^\infty \psi(u_sf)\, \text{d} s\bigg\}. \end{eqnarray}

Proof.First, for $f\in B\big(\mathbb{R}_+\big)^+$ let $f_*= \inf_{x\ge 0} f(x)$ . From Propositions 2.3 and 2.4 we have

(4.6) \begin{eqnarray} \big(1-\text{e}^{-f_*}\big)\text{e}^{-c_1t}\le u_t\,f(x) \le \|u_t\,f\| \le \|\pi_tf\| \le \text{e}^{-c_*t}\|f\|, \qquad x\ge 0. \end{eqnarray}

Moreover, for $a,c>0$ it is elementary to see that

(4.7) \begin{align} \int_0^\infty \text{d} s \int_{\mathfrak{N}\big(\mathbb{R}_+\big)^\circ} \big(1-\text{e}^{-ae^{-cs}\langle \nu,1\rangle }\big)L(\text{d}\nu) & = \int_{\mathfrak{N}\big(\mathbb{R}_+\big)^\circ}L(\text{d}\nu) \int_0^\infty \big(1-\text{e}^{-a\text{e}^{-cs}\langle \nu,1\rangle }\big) \, \text{d} s \nonumber \\[3pt] & = c^{-1}\int_{\mathfrak{N}\big(\mathbb{R}_+\big)^\circ}L(\text{d}\nu) \int_0^{a\langle \nu,1\rangle }\big(1-\text{e}^{-z}\big)z^{-1} \, \text{d} z. \end{align}

Second, suppose that (4.4) holds. For any $f\in B\big(\mathbb{R}_+\big)^+$ we see from (4.6) that $\sigma(u_t\,f)\to 0$ as $t\to \infty$ . Take any $a>\|f\|$ . By (4.6) and (4.7) we have

\begin{align*} \int_0^\infty \psi(u_sf) \, \text{d} s & = \int_0^\infty \text{d} s \int_{\mathfrak{N}\big(\mathbb{R}_+\big)^\circ} \big(1-\text{e}^{-\langle \nu,u_sf\rangle }\big)L(\text{d}\nu) \\[3pt] & \le \int_0^\infty \text{d} s\int_{\mathfrak{N}\big(\mathbb{R}_+\big)^\circ} \big(1-\text{e}^{-a\text{e}^{-c_*s}\langle \nu,1\rangle }\big)L(\text{d}\nu) \\[3pt] & \le c_*^{-1}\int_{\mathfrak{N}\big(\mathbb{R}_+\big)^\circ} \bigg[1 + \int_1^{1\vee (a\langle \nu,1\rangle )} z^{-1} \, \text{d} z\bigg]L(\text{d}\nu) \\[3pt] & \le c_*^{-1}\int_{\mathfrak{N}\big(\mathbb{R}_+\big)^\circ}\big[1 + 0\vee\log(a\langle \nu,1\rangle )\big]L(\text{d}\nu)< \infty. \end{align*}

Then, as $t\to \infty$ ,

\begin{align*} \int_0^t \psi(u_sf) \, \text{d} s & = \int_0^t \text{d} s \int_{\mathfrak{N}\big(\mathbb{R}_+\big)^\circ} \big(1-\text{e}^{-\langle \nu,u_sf\rangle }\big)L(\text{d}\nu) \\[3pt] &\qquad\qquad\quad\!\!\! \to \int_0^\infty \text{d} s \int_{\mathfrak{N}\big(\mathbb{R}_+\big)^\circ} \big(1-\text{e}^{-\langle \nu,u_sf\rangle }\big)L(\text{d}\nu) = \int_0^\infty \psi(u_sf) \, \text{d} s< \infty. \end{align*}

From (4.1) we infer that $P_t(\sigma,\cdot)$ converges as $t \rightarrow \infty$ to a probability measure $\eta$ defined by (4.5); see, e.g., [Reference Li21, Theorem 1.20].

Third, suppose that $P_t(\sigma, \cdot)$ converges as $t\to \infty$ to a probability measure $\eta$ on $\mathfrak{N}\big(\mathbb{R}_+\big)$ for every $\sigma \in \mathfrak{N}\big(\mathbb{R}_+\big)$ . By (4.1) we see that $\eta$ has a Laplace functional given by (4.5); see, e.g., [Reference Li21, Theorems 1.17 and 1.18]. In particular, we have $\int_0^\infty \psi(u_sf) \, \text{d} s< \infty$ for every $f\in B\big(\mathbb{R}_+\big)^+$ . Let $a=(1-\text{e}^{-1})$ . By (4.6) and (4.7) we have

\begin{align*} \int_0^\infty \psi(u_s1) \, \text{d} s & = \int_0^\infty \text{d} s \int_{\mathfrak{N}\big(\mathbb{R}_+\big)^\circ} \big(1-\text{e}^{-\langle \nu,u_s1\rangle }\big)L(\text{d}\nu) \\ & \ge \int_0^\infty \text{d} s \int_{\mathfrak{N}\big(\mathbb{R}_+\big)^\circ} \big(1-\text{e}^{-a\text{e}^{-c_1s}\langle \nu,1\rangle }\big)L(\text{d}\nu) \\ & \ge ac_1^{-1}\int_{\{a\langle \nu,1\rangle \ge 1\}}L(\text{d}\nu) \int_1^{a\langle \nu,1\rangle } z^{-1} \, \text{d} z \\ & \ge ac_1^{-1}\int_{\{a\langle \nu,1\rangle \ge 1\}}\big[\log a + \log\langle \nu,1\rangle \big]L(\text{d}\nu). \end{align*}

This gives us (4.4).

Acknowledgements

We would like to express our sincere gratitude to Professor Viet Chi Tran and two anonymous referees for their very helpful comments on the paper and on the literature of age-structured branching processes. We are grateful to Professor Jie Xiong for discussions on the exploration.

Funding information

This research is supported by the National Key R&D Program of China (No. 2020YFA 0712900), the National Natural Science Foundation of China (No. 11531001), the China Postdoctoral Science Foundation (No. 2020M681994) and Guangdong Basic and Applied Basic Research Foundation (No. 2021A1515010031).

Competing interests

There were no competing interests to declare which arose during the preparation or publication process of this article.

References

Athreya, K. and Ney, P. (1972). Branching Processes. Springer, Berlin.10.1007/978-3-642-65371-1CrossRefGoogle Scholar
Bansaye, V. and Méléard, S. (2015). Stochastic Models for Structured Populations: Scaling Limits and Long Time Behavior. Springer, Cham.Google Scholar
Bellman, R. and Harris, T. (1952). On age-dependent binary branching processes. Ann. Math. 55, 280295.10.2307/1969779CrossRefGoogle Scholar
Bose, A. (1986). A law of large numbers for the scaled age distribution of linear birth-and-death processes. Canad. J. Statist. 14, 233244.10.2307/3314800CrossRefGoogle Scholar
Bose, A. and Kaj, I. (1995). Diffusion approximation for an age-structured population. Ann. Appl. Prob. 5, 140157.10.1214/aoap/1177004833CrossRefGoogle Scholar
Bose, A. and Kaj, I. (2000). A scaling limit process for the age-reproduction structure in a Markov population. Markov Process. Relat. Fields 6, 397428.Google Scholar
Champagnat, N., Ferrière, R. and Méléard, S. (2006). Individual-based probabilistic models of adpatative evolution and various scaling approximations. In Seminar on Stochastic Analysis, Random Fields and Applications V (Progress in Prob. 59, eds R. C. Dalang, F. Russo and M. Dozzi. Birkhäuser, Basel.Google Scholar
Crump, K. S. and Mode, C. J. (1968). A general age-dependent branching process I. J. Math. Anal. Appl. 24, 494508.10.1016/0022-247X(68)90005-XCrossRefGoogle Scholar
Crump, K. S. and Mode, C. J. (1968). A general age-dependent branching process II. J. Math. Anal. Appl. 25, 817.10.1016/0022-247X(69)90210-8CrossRefGoogle Scholar
Dawson, D. A., Gorostiza, L. G. and Li, Z. (2002). Nonlocal branching superprocesses and some related models. Acta Appl. Math. 74, 93112.10.1023/A:1020507922973CrossRefGoogle Scholar
Doney, R. A. (1972). Age-dependent birth and death processes. Z. Wahrscheinlichkeitsth. 22, 6990.10.1007/BF00538906CrossRefGoogle Scholar
Etheridge, A. M. (2000). An Introduction to Superprocesses. American Mathematical Society, Providence, RI.10.1090/ulect/020CrossRefGoogle Scholar
Fournier, N. and Méléard, S. (2004). A microscopic probabilistic description of a locally regulated population and macroscopic approximations. Ann. Appl. Prob. 14, 18801919.10.1214/105051604000000882CrossRefGoogle Scholar
Harris, T. E. (1963). The Theory of Branching Processes. Springer, Berlin.10.1007/978-3-642-51866-9CrossRefGoogle Scholar
Jagers, P. (1969). A general stochastic model for population development. Skand. Aktuar. Tidskr. 52, 84103.Google Scholar
Jagers, P. (1975). Branching Processes with Biological Applications. Wiley, New York.Google Scholar
Jagers, P. and Klebaner, F. C. (2000). Population-size-dependent and age-dependent branching processes. Stoch. Process. Appl. 87, 235254.10.1016/S0304-4149(99)00111-8CrossRefGoogle Scholar
Jagers, P. and Klebaner, F. C. (2011). Population-size-dependent, age-structured branching processes linger around their carrying capacity. J. Appl. Prob. 48A, 249260.10.1239/jap/1318940469CrossRefGoogle Scholar
Kaj, I. and Sagitov, S. (1998). Limit processes for age-dependent branching particle systems. J. Theoret. Prob. 11, 225257.10.1023/A:1021607311191CrossRefGoogle Scholar
Kendall, D. G. (1949). Stochastic processes and population growth. J. R. Statist. Soc. B 11, 230264.Google Scholar
Li, Z. (2011). Measure-Valued Branching Markov Processes. Springer, Heidelberg.10.1007/978-3-642-15004-3CrossRefGoogle Scholar
Li, Z. (2020). Continuous-state branching processes with immigration. In From Probability to Finance, ed. Y. Jiao. Springer, Singapore.10.1007/978-981-15-1576-7_1CrossRefGoogle Scholar
Li, Z. (2021): Ergodicities and exponential ergodicities of Dawson–Watanabe type processes. Teor. Veroyatnost. i Primenen. 66, 342368.10.4213/tvp5341CrossRefGoogle Scholar
Metz, J. A. J. and Tran, V. C. (2013). Daphnias: From the individual based model to the large population equation. J. Math. Biol. 66, 915933.10.1007/s00285-012-0619-5CrossRefGoogle Scholar
Oelschläger, K. (1990). Limit theorems for age-structured populations. Ann. Prob. 18, 290318.10.1214/aop/1176990950CrossRefGoogle Scholar
Pardoux, E. (2016). Probabilistic Models of Population Evolution: Scaling Limits, Genealogies and Interactions. Springer, Switzerland.10.1007/978-3-319-30328-4CrossRefGoogle Scholar
Pazy, A. (1983). Semigroups of Linear Operators and Applications to Partial Differential Equations. Springer, New York.10.1007/978-1-4612-5561-1CrossRefGoogle Scholar
Tran, V. C. (2008). Large population limit and time behaviour of a stochastic particle model describing an age-structured population. ESAIM Prob. Stat. 12, 345386.10.1051/ps:2007052CrossRefGoogle Scholar