Hostname: page-component-54dcc4c588-sq2k7 Total loading time: 0 Render date: 2025-10-01T10:55:27.171Z Has data issue: false hasContentIssue false

Optimal asset allocation and reinsurance problem under enhanced dynamic contagion processes

Published online by Cambridge University Press:  29 September 2025

Guo Liu
Affiliation:
Department of Mathematical Sciences, College of Science, Mathematics and Technology, Wenzhou-Kean University, Ouhai, Wenzhou, Zhejiang Province, China Department of Mathematical Sciences, Dorothy and George Hennings College of Science, Mathematics and Technology, Kean University, Union, NJ, USA
Jiwook Jang*
Affiliation:
Department of Actuarial Studies and Business Analytics, Macquarie University, Sydney, NSW, Australia
*
Corresponding author: Jiwook Jang; Email: jiwook.jang@mq.edu.au
Rights & Permissions [Opens in a new window]

Abstract

This paper examines an insurer’s optimal asset allocation and reinsurance policies. The financial market framework includes one risk-free and one risky asset. The insurer has two business lines, where the ordinary claim process is modeled by a compound Poisson process and catastrophic claims follow a compound dynamic contagion process. The dynamic contagion process, which is a generalization of the externally exciting Cox process with shot-noise intensity and the self-exciting Hawkes process, is enhanced by accommodating the dependency structure between the magnitude of contribution to intensity after initial events for catastrophic insurance products and its claim/loss size. We also consider the dependency structure between the positive effect on the intensity and the negative crashes on the risky financial asset when initial events occur. Our objective is to maximize the insurer’s expected utility of terminal surplus. We construct the extended Hamilton–Jacobi–Bellman (HJB) equation using dynamic programming principles to derive an explicit optimal reinsurance policy for ordinary claims. We further develop an iterative scheme for solving the value function and the optimal asset allocation policy and the reinsurance policy for catastrophic claims numerically, providing a rigorous convergence proof. Finally, we present numerical examples to demonstrate the impact of key parameters.

Information

Type
Original Research Paper
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
© The Author(s), 2025. Published by Cambridge University Press on behalf of Institute and Faculty of Actuaries

1. Introduction

Insurance companies receive premiums from policyholders and invest them in the financial market, while they transfer excess claim risks to reinsurers. Insurers pay the confirmed amount to policyholders through the financial investment and the accumulated premium incomes, while they also transfer possible excess claim risks to reinsurers by paying reinsurance premiums. These are core tasks for insurance companies to ensure that their operations make profits by choosing the optimal investment and reinsurance strategy.

Following Browne’s (Reference Browne1995) study on an insurer’s optimal investment strategy based on maximising the expected utility of terminal wealth, research on optimal investment and other control problems in insurance has advanced through methods of stochastic analysis and modelling. Hipp & Plum (Reference Hipp and Plum2000), Liu & Yang (2004), and Emms & Haberman (Reference Emms and Haberman2007) deal with the optimal strategy of adopting a compound Poisson process for the risk process, minimizing the probability of ruin. Under the same specifications, Hipp & Vogt (Reference Hipp and Vogt2003) study an optimal dynamic unlimited excess of loss reinsurance strategy. Højgaard & Taksar (1998) find a proportional reinsurance policy that maximizes a given return function before the time of ruin for diffusion models. Schmidli (2001) finds optimal reinsurance and investment policies to minimize the probability of ruin, assuming that the surplus process of the insurer follows a classical Cramér–Lundberg model, which can be approximated by diffusion processes.

Yang & Zhang (2005) study optimal investment policies of an insurer with a jump-diffusion risk process with an exponential utility function. Zeng & Li (Reference Zeng and Li2011) consider an optimal investment–reinsurance problem under the mean-variance criterion with one risk-free asset and multiple risky assets whose price processes follow geometric Brownian motions, where an insurer is facing a claim process modeled by a Brownian motion with drift. Under the same specifications except in a regime-switching market, Chen & Yam (Reference Chen and Yam2013) consider the same problem under the mean-variance criterion with one risky asset. Shen & Zeng (2015) also study the same problem under the Cramér–Lundberg model, where the market price of risk depends on a Markovian, affine-form, square-root stochastic factor process.

The aforementioned papers assume that the claim arrival process of the insurer follows a compound Poisson process and/or the surplus process of the insurer follows a jump-diffusion-type process. In other words, advanced counting processes have not been applied to model the arrival of claims and deal with the increasing frequency of extreme events. More importantly, the correlation between the arrival of extreme events and the magnitude of catastrophic loss has not been touched during some pandemics such as financial crises, COVID-19, and global warming.

Ceci et al. (Reference Ceci, Colaneri and Cretarola2022) assume that catastrophic claims and stock prices are driven by common shocks, implying that every catastrophic claim immediately triggers a market crash. In contrast, we introduce an independent Poisson process to model external extreme events, such as earthquakes, tsunamis, and pandemics, which can lead to market downturns. This framework allows for stochastic variability in the relationship between catastrophic claims and asset price fluctuations, providing a more flexible and realistic representation. By incorporating externally exciting dynamics in the intensity, where extreme events amplify catastrophic claim intensity, our model captures real-world phenomena such as aftershocks, secondary losses, and delayed reporting. Additionally, our approach accommodates both positive and negative correlations between asset crashes and externally exciting effects.

Furthermore, to address the ongoing challenge of new risk dynamics arising from climate change with heavy-tailed losses, we use the compound dynamic contagion process for the aggregate claim process for catastrophic insurance business. The dynamic contagion process is a generalization of the externally exciting Cox process with shot-noise intensity and the self-exciting Hawkes process. This process is well suited to capture the frequency and impact of initial catastrophic events (i.e., heatwaves, rainfall, and supercell thunderstorms) and following bushfires, floods, and tornados that will generate losses arising from initial catastrophic events. For the compound dynamic contagion process, readings can refer to Jang & Oh (2021) and the references therein.

Similar to our study, Brachetta et al. (Reference Brachetta, Callegaro, Ceci and Sgarra2024) model an optimal reinsurance problem where claim intensity exhibits jump clustering through self-exciting and externally exciting effects. However, their model assumes that self-excitation is only a function of claim size, and externally exciting effects are independent of self-exciting mechanisms. In contrast, we enhance the relationship between claim sizes and self-excitation by integrating copula theory, enabling a more flexible framework that captures nonlinear dependencies, tail dependencies, and asymmetries. This allows insurers to better assess correlated risks and extreme loss events. Furthermore, rather than treating insurance claims in isolation, we extend the analysis to explore systemic interactions between financial and insurance markets. We model the dependence structure between externally exciting effects and asset price jumps using copula theory, providing valuable insights for insurance-linked securities (ILS) markets. This extension is particularly relevant for reinsurance firms and institutional investors seeking optimal risk-transfer strategies through financial derivatives linked to catastrophe risks. By integrating these methodological advancements, our study bridges the gap between catastrophe insurance modeling and financial market dependencies, offering a more comprehensive approach to risk assessment, pricing, and capital management in highly volatile environments.

Recently, Wu et al. (Reference Wu, Shen, Zhang and Ding2024) investigated the optimal reinsurance problem based on the self-exciting Hawkes process, where the intensity process has a diffusion component. Liu et al. (2021) study an optimization problem of a household under a contagious financial market, where the self-exciting and mutual-exciting clustering features in the asset dynamics are considered for the financial market. Based on mutual-excitation Hawkes processes, Liu et al. (2022) also investigate a non-zero-sum stochastic asset allocation and reinsurance game between two insurers under a contagious insurance market. A compound Cox process for the claim arrival process in dealing with a mean-variance portfolio selection for a non-life insurance company can be noticed in Delong & Gerrard (Reference Delong and Gerrard2007).

In our paper, we study an insurer’s optimal asset allocation and reinsurance policies. The financial market is assumed to consist of one risk-free asset and one risky asset. The insurer has two business lines, where the aggregate claim process for the ordinary insurance business is assumed to follow a compound Poisson process, while the aggregate claim process for the catastrophic insurance business is assumed to follow a compound dynamic contagion process. Furthermore, the dynamic contagion process is enhanced by accommodating the dependency structure between the magnitude of contribution to intensity after initial events for catastrophic insurance products and their claim/loss size. We also consider the dependency between the positive effect on the intensity and the negative effect on the risky financial asset when initial events occur. Different correlation relationships will be studied in this paper. For positive correlations, the higher the jumps in the intensity are, the larger the claim sizes for catastrophic insurance business we will see, and the higher the jumps in the intensity are, the larger the crashes the risky financial asset will have. Otherwise, for negative correlations, the higher the jumps in the intensity are, the smaller the claim sizes for catastrophic insurance business we will see, and the higher the jumps in the intensity are, the smaller the crashes the risky financial asset will have. These two dependency structures provide us with more realistic and substantial insurer’s optimal asset allocation and reinsurance strategies. To capture the potential income of the insurer, different premium principles are considered to illustrate the effect of the dependency structure on the premium amount.

In this paper, the insurance management is assumed to maximize the expected utility at a fixed terminal point. Based on the dynamic programming principle and the corresponding Hamilton–Jacobi–Bellman (HJB) equations, explicit formulas of the optimal reinsurance policies with respect to ordinary claims are derived under different premium principles. For the optimal asset allocation strategy and reinsurance policies with respect to catastrophic claims, we construct an iterative scheme based on the Feynman–Kac formula and the contraction mapping to solve them numerically.

The main contributions of this paper are as follows. First, unlike the classical investment and reinsurance works where claims of insurers follow independent Poisson or Lévy processes, we incorporate and enhance the dynamic contagion process by considering the dependency structure between the financial and insurance markets. In this paper, we focus on the proportional reinsurance policy, and readers can easily extend our method to consider other potential reinsurance contracts. Including the dependency structure between the financial and insurance markets makes the claim arrival process sophisticated. Explicit formulas of catastrophic claim premiums are carefully derived under the net-profit and deviation premium principles, where our method can be easily extended to consider the moment function of cumulative catastrophic claims.

Second, unlike the literature, the dependency structure between the financial and insurance markets makes it impossible to solve the optimal asset allocation strategy explicitly, where the optimal asset allocation policy depends on the external market factor and the catastrophic claim intensity simultaneously. To illustrate the sophisticated asset allocation strategy and reinsurance policy with respect to catastrophic claims, we separate the value functions as two unknown parts based on the exponential affine structure, where the first one admits a closed-form expression. For the second unknown part, an iterative numerical method is constructed based on the Feynman–Kac formula. With the help of this iterative scheme, we can solve the value function, the optimal asset allocation strategy, and the reinsurance policy with respect to catastrophic claims numerically under different scenarios. Based on the Banach fixed point theorem proved in Banach (Reference Banach1922) and Ciesielski (Reference Ciesielski2007), we present the proof of the convergence of the iterative scheme rigorously, where the joint and marginal distributions of catastrophic claims and self-exciting factors, and asset crashes and externally exciting factors are only required to have finite Laplace transforms and finite first-order (partial) derivatives of these Laplace transforms. An operator, based on the fixed point equation, is illustrated to be a contraction mapping with a suitable metric, which provides us with at least first-order convergence speed.

Third, applying the iterative scheme, we investigate the dependency structure between the financial and insurance markets with different correlations and illustrate the corresponding optimal asset allocation and proportional reinsurance policies of ordinary and catastrophic claims. This numerical method can also be extended to consider different reinsurance contracts such as excess-of-loss reinsurance. In addition to the effect of time and claim intensity, the optimal asset allocation policy is also subject to the dependency structure between the financial and insurance markets. When catastrophic claims and self-exciting effects are positively correlated and/or asset crashes and externally exciting effects are positively correlated, the risky asset holdings are reduced. Otherwise, more resources are allocated to the risky asset. For the optimal reinsurance policy with respect to catastrophic claims, when catastrophic claims and self-exciting effects are negatively correlated, catastrophic claim risks decrease, and the insurer can bear more catastrophic risks by itself. Otherwise, more catastrophic risks are transferred to the reinsurer. However, the effect of the correlation between asset crashes and externally exciting factors on the optimal reinsurance policy $w.r.t$ catastrophic claims is negligible. Furthermore, the optimal asset allocation policy and the optimal reinsurance policy $w.r.t$ catastrophic claims are also affected by different premium principles.

The rest of the paper is organized as follows. Section 2 provides a detailed description of the dynamic contagion process and the model formulation, where the explicit premium formulas are derived and the main problem has been constructed. Section 3 presents the main results of this paper, where the analytic solution of the optimal reinsurance policy with respect to ordinary claims and the iterative numerical method are included. Section 4 covers the rigorous proof of the convergence of the iterative method, where restrictions on distributions are clearly stated. Section 5 demonstrates some numerical examples to show the effect of the dependency structure between the financial and insurance markets and the premium principles on the value function and optimal controls. Section 6 concludes this paper.

2. Problem formulation

Consider an insurance company under contagious financial and insurance markets. The insurance company has two business lines, i.e., ordinary insurance products and catastrophic insurance products, as

(2.1) \begin{equation} C(t)=C_{1}(t)+C_{2}(t)=\overset {N_{1}(t)}{\sum \limits _{\kappa =1}}J_{1,\kappa }+\sum \limits _{j=1}^{N_2(t)}J_{2,j}. \end{equation}

The claim process of ordinary claims $C_{1}(t)$ follows the standard Poisson process $N_{1}(t)$ with a deterministic parameter $\eta$ $\gt 0$ , and claim sizes are denoted by $J_{1,\kappa }$ with the cumulative distribution function $F_{1}(\cdot )$ .

The catastrophic claims $C_2(t)$ have a loss model as follows:

(2.2) \begin{equation} C_2(t)=\sum \limits _{j=1}^{N_2(t)}J_{2,j}, \end{equation}

where $J_{2,j}$ are catastrophic claim sizes that occur jointly with $Y_j$ in the intensity process with the joint distribution function $\Psi (y,j_2)$ , defined in (2.7). $N_{2}(t)$ follows the dynamic contagion process with a stochastic intensity process $\lambda (t)$ as follows:

(2.3) \begin{equation} N_{2}(t)=\sum \limits _{j\geq 1}\mathbb{I}\left ( T_{2,j}\leq t\right )_{j=1,2,\ldots }, \end{equation}
(2.4) \begin{equation} \lambda (t)=a+\left ( \lambda _{0}-a\right ) e^{-\delta t}+\sum \limits _{i\geq 1}K_{i}e^{-\delta \left ( t-T_{1,i}\right ) }\mathbb{I}\left ( T_{1,i}\leq t\right ) +\sum \limits _{j\geq 1}Y_{j}e^{-\delta \left ( t-T_{2,j}\right ) }\mathbb{I}\left ( T_{2,j}\leq t\right ), \end{equation}

where

  1. i. $\lambda _{0}$ $\gt 0$ is the initial intensity at time $t=0$ ;

  2. ii. $a$ $\geq 0$ is the constant mean-reverting level;

  3. iii. $\delta$ $\gt 0$ is the rate of exponential decay;

  4. iv. $\left \{ K_{i}\right \} _{i=1,2,\ldots }$ is a sequence of i.i.d. positive externally excited jumps that occur jointly with $L_i$ in the risky asset process with the joint distribution $\Phi (k,l),$ $k\gt 0,l\gt 0$ , where $L_i$ , defined in (2.9), denotes the negative effect of the $i$ th external event on the financial risky assets, at the corresponding random times $\left \{ T_{1,i}\right \} _{i=1,2,\ldots }$ following a Poisson process $M(t)$ with constant rate $\rho \gt 0$ , and $\mathbb{I}$ is the indicator function. The corresponding margins are $H(k)$ and $D(l)$ , respectively. $\Phi (k,l)$ measures the dependency between the positive effect on the intensity and the negative effect on the risky financial asset. $M(t)$ denotes the number of external events such as natural catastrophes, cyberattacks, and pandemics.

  5. v. $\left \{ Y_{j}\right \} _{j=1,2,\ldots }$ is a sequence of i.i.d. positive self-excited jumps with distribution function $\Psi (y,j_{2}),$ $y\gt 0,$ $j_{2}\gt 0$ , where margins are $G(y)$ and $F_2(\cdot )$ , at the corresponding random times $\left \{ T_{2,j}\right \} _{j=1,2,\ldots }$ generated by the intensity process $\lambda (t)$ . $\Psi (y,j_{2})$ measures the dependency between the magnitude of contribution to the intensity after each claim for catastrophic insurance products and their loss size.

  6. vi. $\left \{ T_{1,i}\right \} _{i=1,2,\ldots }$ , $\left \{ J_{1,\kappa }\right \} _{\kappa =1,2,\ldots }$ , $\left \{ \begin{matrix} K_{i}\\ L_{i} \end{matrix} \right \} _{i=1,2,\ldots }$ , and $\left \{ \begin{matrix} Y_{j}\\ J_{2,j} \end{matrix} \right \} _{j=1,2,\ldots }$ are assumed to be independent of each other.

In this paper, we implement the theory of copulas to consider two types of dependency structures from joint distributions $\Psi (y,j_2)$ and $\Phi (k,l).$ Copulas were first introduced by Sklar (1959) to deal with multivariate distribution functions. The formal definition of copulas is as follows.

Definition 2.1. An $n$ -dimensional copula is a function $\mathcal{C}: [0,1]^n\rightarrow [0,1]$ with the following properties

  1. i. $\mathcal{C}(u_1,\ldots ,u_i,\ldots ,u_n)=0$ if any $u_i=0$ for $i=1,\ldots ,n.$

  2. ii. $\mathcal{C}(1,\ldots ,1,u_i\lt 1,1,\ldots ,1)=u_i$ for $i=1,\ldots ,n.$

  3. iii. $\mathcal{C}(\textbf {u})\geq 0$ for any vector $\textbf {u}\in [0,1]^n.$

Thanks to Sklar’s Theorem, which links copulas to joint probability distributions. We first state Sklar’s Theorem and then illustrate the setting for $\Psi (y,j_2)$ and $\Phi (k,l).$

Theorem 2.2. Let $\mathcal{F}$ be an $n$ -dimensional joint distribution with margins $\mathcal{F}_1,\ldots , \mathcal{F}_n$ . Then, there exists an $n$ -dimensional copula $\mathcal{C}$ such that, for any random vector $\textbf {x}\in \mathbb{R}^n$ ,

(2.5) \begin{equation} \mathcal{F}(x_1,\ldots ,x_n)=\mathcal{C}(\mathcal{F}_1(x_1),\ldots , \mathcal{F}_1(x_1)). \end{equation}

Furthermore, if $\mathcal{F}_1,\ldots , \mathcal{F}_n$ are continuous, then $\mathcal{C}$ is unique. Otherwise, $\mathcal{C}$ is unique on $\textbf{Ran}\mathcal{F}_1\times \ldots \times \textbf{Ran}\mathcal{F}_n$ ( $\textbf{Ran}$ denotes Range). More importantly, the copulas density function $c:[0,1]^n \rightarrow [0,\infty )$ follows

(2.6) \begin{equation} c(u_1,\ldots ,u_n)=\frac {\partial ^n\mathcal{C}(u_1,\ldots ,u_n)}{\partial u_1\ldots \partial u_n}=\frac {f(x_1,\ldots ,x_n)}{\Pi ^n_{i=1}f_i(x_i)}, \end{equation}

where $f(x_1,\ldots ,x_n)$ denotes the multivariate probability density function and $f_i(x_i)$ denotes the univariate probability density function of the margin $\mathcal{F}_i$ , for $i=1,\ldots ,n$ .

The proof of Sklar’s theorem can be found in Sklar (1959) and Schweizer & Sklar (1983). Now, Let $(u_1,u_2)\in [0,1]^2$ and $(\widehat {u}_1,\widehat {u}_2)\in [0,1]^2$ such that $u_1=G(y)$ , $u_2=F_2(j_2)$ , $\widehat {u}_1=H(k)$ , and $\widehat {u}_2=D(l)$ . Then, we introduce two bivariate copulas $\mathcal{C}_1(u_1,u_2)$ and $\mathcal{C}_2(\widehat {u}_1,\widehat {u}_2)$ with corresponding copula density functions $p_1(u_1,u_2)$ and $p_2(\widehat {u}_1,\widehat {u}_2)$ such that

(2.7) \begin{align} \mathcal{C}_1(u_1,u_2)&=\Psi (y,j_2),\nonumber\\[2pt] \mathcal{C}_2(\widehat {u}_1,\widehat {u}_2)&=\Phi (k,l),\nonumber\\[2pt] \widehat {\psi }(y,j_2)&=p_1(u_1,u_2)g_y(y)f_2(j_2),\nonumber\\[2pt] \widehat {\phi }(k,l)&=p_2(\widehat {u}_1,\widehat {u}_2)h(k)d(l), \end{align}

where $\widehat {\psi }(y,j_2)$ and $\widehat {\phi }(k,l)$ denote the bivariate joint density function of $\Psi (y,j_2)$ and $\Phi (k,l)$ , $g_y(y)$ , $f_2(j_2)$ , $h(k)$ , and $d(l)$ denote the univariate density function of margins $G(y)$ , $F_2(j_2)$ , $H(k)$ , and $D(l)$ , respectively. Different types of copulas can be considered, and we will use Gaussian copulas for numerical works in Section 5. The role of copula functions is to capture two dependency structures: one is for the dependency structure between the magnitude of the contribution of external events to the intensity and financial risky assets, and the other is the dependency structure between the magnitude of contribution to intensity after each catastrophic claim and the corresponding claim size. For notation simplification, the Laplace transform of all distribution functions is given below, where it is assumed that they are finite, $i.e.$ ,

Assumption 2.3.

\begin{align*} \overset {\wedge }{g}\left ( \zeta \right ) =\int \limits _{0}^{\infty }e^{-\zeta y}dG(y)\lt \infty \mathrm{, }\overset {\wedge }{f_{2}}\left ( \kappa \right ) =\int \limits _{0}^{\infty }e^{-\kappa j_{2}}dF_{2}(j_{2})\lt \infty \mathrm{, } \end{align*}
\begin{align*} \overset {\wedge }{h}\left ( \varpi \right ) =\int \limits _{0}^{\infty }e^{-\varpi k}\mathrm{ }dH(k)\lt \infty \mathrm{, } \overset {\wedge }{d}\left ( \varphi \right ) =\int \limits _{0}^{\infty }e^{-\varphi l}dD(l)\lt \infty \mathrm{, } \overset {\wedge }{f_{1}}\left (\varepsilon \right ) =\int \limits _{0}^{\infty }e^{-\varepsilon j_{1}}dF_{1}(j_{1})\lt \infty , \end{align*}
\begin{align*} \overset {\wedge }{\psi }\left ( \zeta ,\kappa \right ) =\int \limits _{0}^{\infty }\int \limits _{0}^{\infty }e^{-\zeta y}e^{-\kappa j_{2}}d\Psi ( y,j_{2}) \lt \infty \,\,\mathrm{ and }\, \, \overset {\wedge }{\pi }\left ( \varpi ,\varphi \right ) =\int \limits _{0}^{\infty }\int \limits _{0}^{\infty }e^{-\varpi k}e^{-\varphi l}d\Phi ( k,l) \lt \infty . \end{align*}

where $\zeta \gt 0,$ $\kappa \gt 0,$ $\varpi \gt 0,$ $\varphi \gt 0,$ and $\varepsilon \gt 0.$

The negative sign of first-order (partial) derivatives of these Laplace transforms are also given below, where it is assumed that they are finite, $i.e.$ ,

Assumption 2.4.

\begin{align*} -\frac {d\overset {\wedge }{g}\left ( \zeta \right ) }{d\zeta } =\int \limits _{0}^{\infty }ye^{-\zeta y}dG(y)\lt \infty \mathrm{, }-\frac {d\overset { \wedge }{f_2}\left ( \kappa \right ) }{d\kappa }=\int \limits _{0}^{\infty }j_2e^{-\kappa j_2}\mathrm{ }dF_2(j_2)\lt \infty , \end{align*}

\begin{align*} -\frac {d\overset { \wedge }{h}\left ( \varpi \right ) }{d\varpi }=\int \limits _{0}^{\infty }ke^{-\varpi k}\mathrm{ }dH(k)\lt \infty , \mathrm{, }-\frac {d\overset { \wedge }{d}\left ( \varphi \right ) }{d\varphi }=\int \limits _{0}^{\infty }le^{-\varphi l}\mathrm{ }dD(l)\lt \infty , \end{align*}
\begin{align*} -\frac {\partial \overset {\wedge }{\psi }\left ( \zeta ,\kappa \right ) }{ \partial \zeta }=\int \limits _{0}^{\infty }\!\int \limits _{0}^{\infty }\!ye^{-\zeta y}e^{-\kappa j_{2}}d\Psi ( y,j_{2}) \lt \infty \,\,\mathrm{and}\,\,-\frac {\partial \overset {\wedge }{\pi }\left ( \varpi ,\varphi \right ) }{\partial \varpi }=\int \limits _{0}^{\infty }\!\int \limits _{0}^{\infty }\!k e^{-\varpi k}e^{-\varphi l}d\Phi ( k,l) \lt \infty . \end{align*}

where $\zeta \gt 0,$ $\kappa \gt 0,$ $\varpi \gt 0,$ and $\varphi \gt 0.$

Based on Dassios & Zhao (Reference Dassios and Zhao2012), to have a stationary intensity process $\lambda (t)$ , we need $\delta \gt m_{1y}$ , where $m_{1y}=\int \limits _0^{\infty }ydG(y)\lt \infty .$ Furthermore, we assume that there are two available assets in the financial market.

The risk-free asset $B(t)$ follows

(2.8) \begin{equation} dB(t)=rB(t)dt, \end{equation}

where $r\gt 0$ denotes the risk-free rate. The risky asset $S(t)$ follows

(2.9) \begin{equation} dS(t)=\mu S(t)dt+\sigma S(t)dW(t)-LS(t)dM(t), \end{equation}

where $\mu \gt r$ denotes the drift, $\sigma \gt 0$ denotes the volatility, and $L\gt 0$ denotes the negative effect of external events on the risky financial asset, with a joint distribution $\Phi (k,l)$ mentioned before.

Formally, we consider a probability space $(\Omega ,\mathcal{F},\mathbb{P})$ that is generated by the above point processes and the Brownian motion and satisfies usual conditions. A filtration $\{\mathcal{F}_t\}$ represents the information up to time t. Let $\pi (t)$ denotes the dollar amount allocated into the risky asset, and let $q_1(t)$ and $q_2(t)$ denote the proportion of ordinary and catastrophic claims covered by the insurer, respectively. Let $\nu (t)=(\pi (t),q_1(t),q_2(t))$ denote a trading strategy. The controlled surplus process $X^{\nu }(t)$ becomes

(2.10) \begin{align} dX^{\nu }(t)&=\frac {X^{\nu }(t)-\pi (t)}{B(t)}dB(t)+\frac {\pi (t)}{S(t)}dS(t)+(c_1-\tilde {c}_1(1-q_1(t))+c_2-\tilde {c}_{2}(1-q_2(t)))dt\nonumber\\[2pt] &\quad -q_1(t)dC_1(t)-q_2(t)dC_2(t)\nonumber\\[2pt] &=\left[(\mu -r)\pi (t)+rX^{\nu }(t)+c_1-\tilde {c}_{1}(1-q_1(t))+c_2-\tilde {c}_2(1-q_2(t))\right]dt+\sigma \pi (t)dW(t)\nonumber\\[2pt] &\quad -q_1(t)J_{1}dN_1(t)-q_2(t)J_2dN_2(t)-\pi (t)LdM(t). \end{align}

In (2.10), $c_1$ and $\tilde {c}_1$ denote the insurance and reinsurance premium of ordinary claims. $c_2$ and $\tilde {c}_2$ denote the insurance and reinsurance premium of catastrophic claims. $J_{1,\kappa }$ and $J_{2,j}$ are assumed to be $i.i.d.$ random variables with magnitude $J_1$ and $J_2$ , respectively. For the insurance premium rate $c_i$ , we consider two popular premium principles in this paper, $i.e.$ , the net-profit principle and the deviation-premium principle. Under the net-profit principle, we have $c_it=(1+\theta _i)\mathbb{E}(C_i(t))$ with $\theta _i\gt 0$ , for $i=1,2.$ Under the deviation-premium principle, we assume that $c_it=\mathbb{E}(C_i(t))+\theta _i\sqrt {Var(C_i(t))}$ . For ordinary claims, under the net-profit principle, we have $c_1=(1+\theta _1)\eta m_{1j_1}$ , where $m_{1j_1}=\int \limits _0^{\infty }j_1dF_1(j_1)$ , since the claim amount is independent of the claim process. Under the deviation-premium principle, we have $c_1t=\eta m_{1j_1}t+\theta _1\sqrt {\eta m_{2j_1}t}$ , where $m_{2j_1}=\int \limits _0^{\infty }j^2_1dF_1(j_1)$ . For catastrophic claims, we follow the work of Dassios & Zhao (Reference Dassios and Zhao2012) for the premium of catastrophic claims under enhanced dynamic contagion processes, which are captured in Theorem2.6.

Now, we define the admissible set of trading strategies for the insurer by the following definition:

Definition 2.5. An investment-reinsurance strategy of the insurer $\nu (t):=\{\pi (t),q_1(t),q_2(t)\}_{t\in [0,T]}$ is said to be an admissible strategy if the following conditions hold:

  1. i. $\{\nu (t)\}_{t\in [0,T]}$ is a progressively measurable process $w.r.t.$ the filtration $\{\mathcal{F}_t\}_{t\in [0,T]}$ ;

  2. ii. $\pi (t)\in (-\infty ,\infty )$ for any $t\in [0,T]$ , $i.e.$ , no short-selling restrictions apply;

  3. iii. $q_1(t)\in [0,1]$ for any $t\in [0,T]$ .

  4. iv. $q_2(t)\in [0,1]$ for any $t\in [0,T]$

  5. v. $\mathbb{E}_{t,x,\lambda }\left [U(X^{\nu }(T))\right ]\lt \infty$ , where $\mathbb{E}_{t, x,\lambda }[\,\cdot \,]=\mathbb{E}[\,\cdot \,|X^{v}(t)=x,\lambda (t)=\lambda ]$ and $U(\cdot )$ denote some suitable utility function, which will be covered in more detail in Section 3 .

Let $\mathcal{V}$ be the set of all admissible policies for the insurer.

Theorem 2.6. Under the assumption $\delta \gt m_{1y}$ , the premium of catastrophic claims under the net-profit principle follows

(2.11) \begin{equation} c_2=(1+\theta _2)\frac {m_{1k}\rho +a\delta }{\delta -m_{1y}}m_{1j_2}, \end{equation}

where $m_{1k}=\int \limits _0^{\infty }kdH(k)$ and $m_{1j_2}=\int \limits _0^{\infty }j_2dF_2(j_2).$

Under the deviation-premium principle, the premium of catastrophic claims follows

(2.12) \begin{equation} c_2t=\mathbb{E}(C_2(t))+\theta _2\sqrt {Var(C_2(t))}, \end{equation}

where

(2.13) \begin{equation} \mathbb{E}(C_2(t))=\frac {a\delta +\rho m_{1k}}{\delta -m_{1y}}m_{1j_2}t. \end{equation}
(2.14) \begin{align} Var(C_2(t))& = 2m_{1j_{2}}\left [ \begin{array}{c} m_{1j_{2}}\mathrm{ }\left [ \frac {\left \{ 2\left ( m_{1k}\mathrm{ }\rho +a\delta \right ) +m_{2y}\right \} \left ( m_{1k}\mathrm{ }\rho +a\delta \right ) }{2\left ( \delta -m_{1y}\right ) ^{3}}+\frac {m_{2k}\mathrm{ }\rho }{2\left ( \delta -m_{1y}\right ) ^{2}}\right ] \left \{ t-\left ( \frac { 1-e^{-\left ( \delta -m_{1y}\right ) t}}{\delta -m_{1y}}\right ) \right \} \\[15pt] +m_{1\Psi _{y,j_{2}}}\mathrm{ }\left \{ \frac {m_{1k}\mathrm{ }\rho +a\delta }{\left ( \delta -m_{1y}\right ) ^{2}}\right \} \left \{ t-\left ( \frac { 1-e^{-\left ( \delta -m_{1y}\right ) t}}{\delta -m_{1y}}\right ) \right \} \\[15pt] -\left ( m_{1k}\mathrm{ }\rho +a\delta \right ) m_{1j_{2}}\left \{ \frac { m_{1k}\mathrm{ }\rho +a\delta }{\left ( \delta -m_{1y}\right ) ^{3}} \right \} \left \{ t-\left ( \frac {1-e^{-\left ( \delta -m_{1y}\right ) t}}{ \delta -m_{1y}}\right ) \right \} \end{array} \right ] \nonumber \\[15pt] & \quad +m_{2j_{2}}\left (\frac {a\delta +\rho m_{1k}}{\delta -m_{1y}}\right ) t, \end{align}

where $m_{2y}=\int \limits _0^{\infty }y^2dG(y)$ , $m_{2k}=\int \limits _0^{\infty }k^2dH(k)$ , $m_{2j_2}=\int \limits _0^{\infty }j_2^2dF_2(j_2)$ , and $m_{1\Psi _{y,j_{2}}}=\int \limits _{0}^{\infty }\int \limits _{0}^{\infty }yj_{2}d\Psi ( y,j_{2})$ .Footnote 1

The proof of Theorem2.6 is provided in Appendix A.

For the reinsurance premium rate $\tilde {c}_{i}$ , $i=1,2,$ we assume that the reinsurer applies the same premium principle as the insurer does but with a different risk loading factor $\theta _i$ , $i=1,2.$ Furthermore, $\theta _i$ , $i=1,2,$ satisfy the following conditions

Assumption 2.7.

(2.15) \begin{equation} \begin{aligned} &\tilde {\theta }_i\geq \theta _i, \quad i, j\in \{1,2\}\\ &c_i-\tilde {c}_{i}+c_j\geq 0, \quad i, j\in \{1,2\}, i\neq j, \end{aligned} \end{equation}

which can eliminate arbitrage opportunities and sets an upper bound for the reinsurance premium for the two business lines.

Given the initial surplus $X(t)=x\gt 0$ and the intensity of catastrophic claims $\lambda (t)=\lambda \gt 0$ at time $t\geq 0$ , the insurance company’s management wants to maximize the expected utility of the surplus at the terminal date $T\geq t$ .

(2.16) \begin{equation} V(t,x,\lambda )=\sup _{\nu (t)\in \mathcal{V}}\mathbb{E}_{t,x,\lambda }\left [U(X^{\nu }(T))\right ], \end{equation}

where $U(\cdot )$ denotes a suitable utility function and $\mathbb{E}_{t,x,\lambda }[\cdot ]$ denotes the conditional expectation given that $X(t)=x$ and $\lambda (t)=\lambda$ . Note that we omit the variable $t$ in $\pi (t), q_1(t),$ and $q_2(t)$ to simplify notations thereafter.

Define an operator $\mathcal{K}$ on a function $f(t,x,\lambda )$ within its domain $[0,T]\times (0,\infty )\times (0,\infty )$ as follows:

(2.17) \begin{align} \mathcal{K}f(t,x,\lambda )=&\left .(rx+(\mu -r)\pi +c_1-(1-q_1)\tilde {c}_{1}+c_2-(1-q_2)\tilde {c}_{2}\right .)\frac {\partial f(t,x,\lambda )}{\partial x}\nonumber\\[3pt] &+\delta (\alpha -\lambda )\frac {\partial f(t,x,\lambda )}{\partial \lambda }+0.5\sigma ^2\pi ^2\frac {\partial ^2f(t,x,\lambda )}{\partial x^2}\nonumber\\[3pt] & +\eta \int \limits _0^{\infty } (f(t,x-q_1j_1,\lambda )-f(t,x,\lambda ))dF_1(j_1)\nonumber\\[3pt] &+\lambda \int \limits _0^{\infty }\int \limits _0^{\infty } (f(t,x-q_2j_2,\lambda +y)-f(t,x,\lambda ))d\Psi (y,j_2)\nonumber\\[3pt] &+\rho \int \limits _0^{\infty }\int \limits _0^{\infty } (f(t,x-\pi l,\lambda +k)-f(t,x,\lambda ))d\Phi (k,l), \end{align}

where the domain of the operator $\mathcal{K}$ contains the class of functions $f(t,x,\lambda )\in \mathcal{C}^{1,2,1}([0,T]\times (0,\infty )\times (0,\infty ))$ such that for every $(t,x,\lambda )\in ([0,T]\times (0,\infty )\times (0,\infty ))$ , we have

(2.18) \begin{align} &\mathbb{E}\left [\int \limits _0^{\infty } \left(f(t,x-q_1j_1,\lambda )-f(t,x,\lambda )\right)dF_1(j_1)\right ]\lt \infty ,\nonumber\\[3pt] &\mathbb{E}\left [\int \limits _0^{\infty }\int \limits _0^{\infty } \left(f(t,x-q_2j_2,\lambda +y)-f(t,x,\lambda )\right)d\Psi (y,j_2)\right ]\lt \infty ,\nonumber\\[3pt] &\mathbb{E}\left [\int \limits _0^{\infty }\int \limits _0^{\infty } \left(f(t,x-\pi l,\lambda +k)-f(t,x,\lambda )\right)d\Phi (k,l)\right ]\lt \infty . \end{align}

By standard techniques, we can obtain the HJB equation and the terminal condition for the value function on its domain $[0,T]\times (-\infty ,\infty )\times (0,\infty )$ as follows:

(2.19) \begin{align} \frac {\partial V}{\partial t}+\sup _{\nu (t)\in \mathcal{V}}\mathcal{K}V(t,x,\lambda ) & =0,\nonumber\\ V(T,x,\lambda ) &= U(x). \end{align}

3. Main results

In this paper, we consider exponential utilities $U(x)=\frac {-1}{\gamma }e^{-\gamma x}$ with $\gamma \gt 0$ . Exponential utilities play a significant role in the insurance sector and financial mathematics since it is the only utility function where the principle of “zero utility” gives a fair premium that is independent of the insurer’s surplus level. This important feature is also reflected in this paper, which will be demonstrated later. Motivated by Duffie et al. (Reference Duffie, Pan and Singleton2000), Aït-Sahalia & Hurd (Reference Aït-Sahalia and Hurd2015), and Liu et al. (2021), if the claim intensity $\lambda (t)$ follows (2.4), we conjecture that the value function $V(t,x,\lambda )$ satisfies an exponential affine structure, which can be separated as follows:

(3.1) \begin{align} V(t,x,\lambda )&=\frac {-1}{\gamma }e^{A(t,T)x}g(t,\lambda ),\nonumber\\[4pt] g(t,\lambda )&=e^{B(t,T)+C(t,T)\lambda }, \end{align}

where $A(t,T)$ , $B(t,T)$ , and $C(t,T)$ are functions to be determined. The explicit solution for $A(t,T)$ can be obtained, while $B(t,T)$ and $C(t,T)$ are too complex to be solved. To describe the value function, $g(t,\lambda )$ is applied to represent the implicit version of the affine structure. Hence, the main task in the following sections is to solve $A(t,T)$ and $g(t,\lambda )$ . Note that we will denote $A(t, T)$ by $A$ thereafter to simplify notations unless it is necessary to be specified.

Replacing $V(t,x,\lambda )$ by (3.1), we can simplify the HJB equation as follows:

(3.2) \begin{align} 0 & =\inf _{\pi ,q_1,q_2}\frac {\partial g}{\partial t}+\frac {\partial A}{\partial t}xg+\left [rx+(\mu -r)\pi +c_1-(1-q_1)\tilde {c}_1+c_2-(1-q_2)\tilde {c}_2\right ]Ag\nonumber\\ &\quad +\delta (\alpha -\lambda )\frac {\partial g}{\partial \lambda }+0.5\sigma ^2\pi ^2A^2g+\eta g\int \limits _0^{\infty } \left(e^{-Aq_1j_1}-1\right)dF_1(j_1)\nonumber\\ &\quad +\lambda \int \limits _0^{\infty }\!\int \limits _0^{\infty }\!\left(e^{-Aq_2j_2}g(t,\lambda +y)-g(t,\lambda )\right)d\Psi (y,j_2)+\rho \int \limits _0^{\infty }\!\int \limits _0^{\infty } \!\left(e^{-A\pi l}g(t,\lambda +k)-g(t,\lambda )\right)d\Phi (k,l), \end{align}

with the terminal condition

(3.3) \begin{equation} g(T,\lambda )=1. \end{equation}

According to the first-order condition, optimal controls are as follows, which are independent of the surplus level

(3.4) \begin{align} \pi ^*&=\arg \inf _{\pi }(\mu -r)\pi Ag+0.5\sigma ^2\pi ^2A^2g+\rho \int \limits _0^{\infty }\int \limits _0^{\infty } \left (e^{-A\pi l}g(t,\lambda +k)-g(t,\lambda )\right )d\Phi (k,l),\nonumber\\ q_1^*&=\arg \inf _{q_1\in [0,1]}Ag\tilde {c}_1q_1+\eta g\int \limits _0^{\infty }\left (e^{-Aq_1j_1}-1\right )dF_1(j_1),\nonumber\\ q_2^*&=\arg \inf _{q_2\in [0,1]}Ag\tilde {c}_2q_2+\lambda \int \limits _0^{\infty }\int \limits _0^{\infty }\left (e^{-Aq_2j_2}g(t,\lambda +y)-g(t,\lambda )\right )d\Psi (y,j_2), \end{align}

which admits the analytic solution for the optimal reinsurance policy with respect to ordinary claims. Now, we provide a verification result in terms of the simplified HJB equation (3.2).

Theorem 3.1. (Verification theorem).

  1. (i) Let $\tilde {g}(t,\lambda )\in \mathcal{C}^1([0,T]\times (0,\infty ))$ be a classical solution of the HJB equation (3.2) that satisfies the terminal condition (3.3).

  2. (ii) Let $\tilde {v}(t,x,\lambda )=\frac {-1}{\gamma }e^{A(t,T)x}\tilde {g}(t,\lambda )=\frac {-1}{\gamma }e^{-\gamma e^{-r(t-T)}x}\tilde {g}(t,\lambda )$ .

    Then, the function $\tilde {v}(t,x,\lambda )$ coincides with the value function $V(t,x,\lambda )$ in (2.16). Furthermore, let $\pi ^*(t)$ , $q_1^*(t)$ , and $q_2^*(t)$ be the corresponding minimizers of (3.4). Then $\nu ^*(t)=(\pi ^*(t),q_1^*(t),q_2^*(t))\in \mathcal{V}$ is an optimal strategy.

The proof of Theorem3.1 is provided in Appendix B.

For the unknown function $A(t,T)$ and the optimal reinsurance policy with respect to ordinary claims, we have

Lemma 3.2. Given any initial conditions $X(t)=x\gt 0$ and $\lambda (t)=\lambda \gt 0$ at any time $t\geq 0$ , the unknown $A(t,T)$ follows $A(t,T)=-\gamma e^{-r(t-T)}$ , and the analytic solution of the optimal reinsurance policy with respect to ordinary claims under a general ordinary claim distribution $F_1(j_1)$ is provided as follows:

(3.5) \begin{equation} \begin{aligned} q_1^*=\min \left (1,\tilde {q}_1\right ), \end{aligned} \end{equation}

where $\tilde {q}_1$ is the unique positive root of the following equation

(3.6) \begin{align} \mathrm{Under the net-profit principle:} \quad \quad (1+\tilde {\theta }_1)m_{j_1}&=\int \limits _0^{\infty } j_1e^{-A(t,T)\tilde {q}_1j_1}dF_1(j_1).\nonumber\\ \mathrm{Under the deviation-premium principle:} \quad \quad m_{1j_1}+\tilde {\theta }_1\sqrt {\frac {m_{2j_1}}{\eta t}}&=\int \limits _0^{\infty } j_1e^{-A(t,T)\tilde {q}_1j_1}dF_1(j_1). \end{align}

Furthermore, suppose $J_1\sim Exp(\epsilon _1), \epsilon _1\gt 0$ , we have

(3.7) \begin{equation} \begin{aligned} \mathrm{Under the net-profit principle:}\quad \quad q_1^*&=\min \left (1,\frac {\epsilon _1}{A(t,T)}\left (\frac {1}{\sqrt {1+\tilde {\theta }_1}}-1\right )\right ).\\[5pt] \mathrm{Under the deviation-premium principle:} \quad \quad q_1^*&=\min \left (1,\frac {\epsilon _1}{A(t,T)}\left (\sqrt {\frac {1}{1+\tilde {\theta }_1\sqrt {\frac {2}{\eta t}}}}-1\right )\right ). \end{aligned} \end{equation}

The proof of Lemma 3.2 is provided in Appendix C.

Given the closed-form expression in Lemma 3.2, we can further simplify the HJB equation as follows:

\begin{align*} 0 & = \inf _{\pi ,q_2}\frac {\partial g}{\partial t}+\delta (\alpha -\lambda )\frac {\partial g}{\partial \lambda }+(\mu -r)\pi Ag+0.5\sigma ^2\pi ^2A^2g\nonumber\\& \quad +\rho \int \limits _0^{\infty }\int \limits _0^{\infty } \left(e^{-A\pi l}g(t,\lambda +k)-g(t,\lambda )\right)d\Phi (k,l)\end{align*}
(3.8) \begin{align} &\quad +\left [c_1-\tilde {c}_1+c_2-\tilde {c}_2\right ]Ag+Ag\tilde {c}_1q^*_1+\eta g\int \limits _0^{\infty } \left(e^{-Aq^*_1j_1}-1\right)dF_1(j_1)\nonumber\\ &\quad +Ag\tilde {c}_2q_2+\lambda \int \limits _0^{\infty }\int \limits _0^{\infty } \left(e^{-Aq_2j_2}g(t,\lambda +y)-g(t,\lambda )\right)d\Psi (y,j_2). \end{align}

Note that the infinitesimal generator $\mathcal{A}$ for a function $f(t,\lambda )$ in its domain $[0,T]\times (0,\infty )$ is given by

(3.9) \begin{equation} \begin{aligned} \{\mathcal{A}f\}(t,\lambda )&=\frac {\partial f(t,\lambda )}{\partial t}+\delta (\alpha -\lambda )\frac {\partial f(t,\lambda )}{\partial \lambda }\\ &\quad +\lambda \int \limits _0^{\infty }\int \limits _0^{\infty } \Big (f(t,\lambda +y)-f(t,\lambda )\Big )d\Psi (y,j_2)+\rho \int \limits _0^{\infty }\int \limits _0^{\infty } \Big (f(t,\lambda +k)-f(t,\lambda )\Big )d\Phi (k,l). \end{aligned} \end{equation}

Then, the HJB equation (3.8) can be further simplified as follows:

(3.10) \begin{equation} \{\mathcal{A}g\}(t,\lambda )+\left [c_1-\tilde {c}_1+c_2-\tilde {c}_2\right ]Ag+\inf _{\pi }\chi _1(\pi ,g(t,\lambda ))+\chi _2(q_1^*,g(t,\lambda ))+\!\!\inf _{q_2\in [0,1]}\chi _3(q_2,g(t,\lambda ))=0, \end{equation}

where

(3.11) \begin{align} \chi _1(\pi , g(t,\lambda ))&=(\mu -r)\pi Ag+0.5\sigma ^2\pi ^2A^2g+\rho \int \limits _0^{\infty }\int \limits _0^{\infty } \left(e^{-A\pi l}g(t,\lambda +k)-g(t,\lambda +k)\right)d\Phi (k,l),\nonumber\\ \chi _2(q_1^*, g(t,\lambda ))&=q_1^*\tilde {c}_1Ag+\eta g\int \limits _0^{\infty } (e^{-Aq_1^*j_1}-1)dF_1(j_1),\nonumber\\ \chi _3(q_2, g(t,\lambda ))&=q_2\tilde {c}_2Ag+\lambda \int \limits _0^{\infty }\int \limits _0^{\infty } \left(e^{-Aq_2j_2}g(t,\lambda +y)-g(t,\lambda +y)\right)d\Psi (y,j_2) ,\nonumber\\ \pi ^* & =\arg \inf _{\pi }\chi _1(\pi , g(t,\lambda )),\mathrm{ and } q_2^*=\arg \inf _{q_2\in [0,1]}\chi _3(q_2,g(t,\lambda )). \end{align}

Given Lemma 3.2 and (3.10), we know that $g(t,\lambda )$ can be expressed as the following conditional expectation based on the Feynman–Kac formula,

(3.12) \begin{equation} \begin{aligned} g(t,\lambda )&=\mathbb{E}\left [\int \limits _t^Te^{\int \limits _t^s\widehat {A}(\tau ,T)d\tau }[\chi _1(\pi ^*(s),g(s,\lambda (s)))+\chi _2(q_1^*(s),g(s,\lambda (s)))\right .\\ &\quad \left .+\chi _3(q_2^*(s),g(s,\lambda (s)))]ds+e^{\int \limits _t^T\widehat {A}(\tau ,T)d\tau }|\mathcal{F}_t \right ], \end{aligned} \end{equation}

where $\widehat {A}(\tau ,T)=(c_1-\tilde {c}_1+c_2-\tilde {c}_2)A(\tau ,T)$ . Then, we can develop an implicit formula of $g(t,\lambda )$ based on (3.12). Given the initial guess of $g(t,\lambda )$ , denoted by $g^{0}(t,\lambda )$ , the iterative scheme is as follows:

(3.13) \begin{equation} \begin{aligned} g^{n+1}(t,\lambda )&=\mathbb{E}\left [\int \limits _t^Te^{\int \limits _t^s\widehat {A}(\tau ,T)d\tau }[\chi _1(\pi _n^*(s),g^n(s,\lambda (s)))+\chi _2(q_{1n}^*(s),g^n(s,\lambda (s)))\right .\\[5pt] &\quad \left .+\chi _3(q_{2n}^*(s),g^n(s,\lambda (s)))]ds+e^{\int \limits _t^T\widehat {A}(\tau ,T)d\tau }|\mathcal{F}_t \right ], \end{aligned} \end{equation}

where $\pi ^*_n(t)$ and $q_{in}^*(t)$ denote the optimal allocation policy and reinsurance policy in (3.11) with a given $g^n(t,\lambda )$ for $i=1,2,$ and $n=0,1,2,\ldots$ .

4. Proof of convergence

To prove the convergence of the iterative method in (3.13) is equivalent to proving the existence and uniqueness of the solution to (3.12). Define an operator $\mathcal{L}$ acting on the function $g(t,\lambda )$ as follows:

(4.1) \begin{equation} \begin{aligned} (\mathcal{L}g)(t,\lambda )&=\mathbb{E}\left [\int \limits _t^Te^{\int \limits _t^s\widehat {A}(\tau ,T)d\tau }[\chi _1(\pi ^*(s),g(s,\lambda (s)))+\chi _2(q_1^*(s),g(s,\lambda (s)))\right .\\[5pt] &\quad \left .+\chi _3(q_2^*(s),g(s,\lambda (s)))]ds+e^{\int \limits _t^T\widehat {A}(\tau ,T)d\tau }|\mathcal{F}_t \right ]. \end{aligned} \end{equation}

Hence, the function $g(t,\lambda )$ in (3.12) satisfies a fixed point equation $(\mathcal{L}g)(t,\lambda )=g(t,\lambda ).$ To prove the existence and uniqueness of the solution to (3.12), we only need to prove that the operator $\mathcal{L}$ is a contraction mapping. Note that, to enhance the readability and organization of the manuscript, we consolidate all technical proofs related to the results of Section 4 in Appendix D.

Lemma 4.1. The function $g(t,\lambda )$ is bounded for any $t\in [0,T]$ and $\lambda \in (0,\infty )$ and increasing in $t\in [0,T]$ , and the operator $\mathcal{L}$ on $g(t,\lambda )$ is also bounded for any $t\in [0,T]$ and $\lambda \in (0,\infty )$ , $i.e.$ ,

(4.2) \begin{equation} \begin{aligned} 0\lt (\mathcal{L}g)(t,\lambda )\lt \beta , \end{aligned} \end{equation}

where $\beta$ is a positive constant depending on the initial surplus $x$ .

The proof of Lemma 4.1 is provided in Appendix D.1.

Lemma 4.2. Given Lemma 4.1 , where $g(t,\lambda )$ is bounded and positive for any $t\in [0,T]$ and $\lambda \in (0,\infty )$ , $g(t,\lambda )$ is Lipschitz continuous for $t\in [0,T]$ and $\lambda \in (0,\infty )$ . Furthermore, $B(t,T)$ and $C(t,T)$ in (3.1) are also bounded and Lipschitz continuous in the same interval.

The proof of Lemma 4.2 is provided in Appendix D.2.

Based on Lemmas 4.1 and 4.2, we define a space of continuous functions $\zeta$ on $[0,T]\times (0,\infty )$ by $\mathfrak{C}_e([0,T]\times (0,\infty ))$ , where $0\lt \zeta (t,\lambda )\lt \beta$ and

(4.3) \begin{equation} \zeta (t,\lambda )=e^{\tilde {\psi }(t,T)+\tilde {\psi }_1(t,T)\lambda }, \end{equation}

where $\tilde {\psi }(t,T)$ and $\tilde {\psi }_1(t,T)$ are bounded and Lipschitz continuous for $t\in [0,T]$ and $\lambda \in (0,\infty )$ . In this section, we will prove that the mapping $\mathcal{L}$ : $\mathfrak{C}_e([0,T]\times (0,\infty ))$ $\to$ $\mathfrak{C}_e([0,T]\times (0,\infty ))$ is a contraction mapping with some suitable metric. Then, (4.1) admits a unique solution, which is $g(t,\lambda )\in \mathfrak{C}_e([0,T]\times (0,\infty )).$

Note that a metric is defined on $\mathfrak{C}_e([0,T]\times (0,\infty ))$ as follows: for any two functions $v_1(t,\lambda )\in \mathfrak{C}_e([0,T]\times (0,\infty ))$ and $v_2(t,\lambda )\in \mathfrak{C}_e([0,T]\times (0,\infty ))$ , let

(4.4) \begin{equation} d(v_1,v_2)=\sup _{(t,\lambda )\in [0,T]\times (0,\infty )}\left |e^{-\alpha (T-t)}\left (v_1(t,\lambda )-v_2(t,\lambda )\right )\right |, \end{equation}

where $\alpha$ denotes a positive constant to be determined.

Lemma 4.3. The space $\mathfrak{C}_e([0,T]\times (0,\infty ))$ with the metric (4.4) is complete.

The proof of Lemma 4.3 is provided in Appendix D.3.

Now, given Lemma 4.3, to show that the operator $\mathcal{L}$ is a contraction mapping, we only need to prove $d(\mathcal{L}v_1,\mathcal{L}v_2)\leq \phi d(v_1,v_2)$ for some $\phi \in [0,1)$ .

(4.5) \begin{align} d(\mathcal{L}v_1,\mathcal{L}v_2) & = \sup _{t,\lambda }\left |e^{-\alpha (T-t)}\left (\mathcal{L}v_1-\mathcal{L}v_2\right )\right |\nonumber\\[4pt] & =\sup _{t,\lambda }\Bigg|e^{-\alpha (T-t)}\mathbb{E}\Bigg[\int \limits _t^Te^{\int \limits _t^s\widehat {A}(\tau ,T)d\tau }\left(\chi _1(\pi _{v_1}^*(s),v_1(s,\lambda (s)))-\chi _1(\pi _{v_2}^*(s),v_2(s,\lambda (s)))\right.\nonumber\\[4pt] &\quad +\chi _2(q_{1v_1}^*(s),v_1(s,\lambda (s)))-\chi _2(q_{1v_2}^*(s),v_2(s,\lambda (s)))\nonumber\\[4pt] &\quad \left. +\chi _3\left(q_{2v_1}^*(s),v_1(s,\lambda (s))\right)-\chi _3\left(q_{2v_2}^*(s),v_2(s,\lambda (s))\right)\right)ds|\mathcal{F}_t\Bigg]\Bigg|, \end{align}

where $\chi _1(\pi _{v_i}^*(s),v_i(s,\lambda (s)))$ , $\chi _2(q_{1v_i}^*(s),v_i(s,\lambda (s)))$ , and $\chi _3(q_{2v_i}^*(s),v_i(s,\lambda (s)))$ , $i=1,2,$ follow (3.11). $\pi ^*_{v_i}(s)$ , $q_{1v_i}^*(s)$ , and $q_{2v_i}^*(s)$ denote the optimal allocation and reinsurance policies of ordinary claims and catastrophic claims with respect to $v_1(t,\lambda )$ and $v_2(t,\lambda )$ .

Lemma 4.4. For any two continuous functions $v_1,v_2\in \mathfrak{C}_e([0, T]\times (0,\infty ))$ , let $\pi ^*_{v_i}(s)$ , $q_{1v_i}^*(s)$ , and $q_{2v_i}^*(s)$ denote the optimal allocation and reinsurance policies of ordinary claims and catastrophic claims with respect to $v_1(t,\lambda )$ and $v_2(t,\lambda )$ . Suppose that joint distributions $\Phi (k,l)$ and $\Psi (y,j_2)$ as well as marginal distributions $H(k), D(l), F_2(j_2)$ , and $G(y)$ follow Assumptions 2.3 and 2.4 and the distribution $F_1(j_1)$ follows Assumption 2.3 , we have the following inequality

(4.6) \begin{equation} \begin{aligned} & |\chi _1(\pi _{v_1}^*(s),v_1(s,\lambda (s)))-\chi _1(\pi _{v_2}^*(s),v_2(s,\lambda (s)))\\ &\quad +\chi _2(q_{1v_1}^*(s),v_1(s,\lambda (s)))-\chi _2\left(q_{1v_2}^*(s),v_2(s,\lambda (s))\right)\\ &\quad +\chi _3(q_{2v_1}^*(s),v_1(s,\lambda (s)))-\chi _3\left(q_{2v_2}^*(s),v_2(s,\lambda (s))\right)|\lt \Delta D_{\lambda }(v_1,v_2), \end{aligned} \end{equation}

where $\Delta$ denotes some positive constant and $D_{\lambda }(v_1,v_2)=\sup _{\lambda }|v_1(s,\lambda )-v_2(s,\lambda )|$ .

The proof of Lemma 4.4 is provided in Appendix D.4.

Lemma 4.5. Given any two continuous functions $v_1,v_2\in \mathfrak{C}_e([0,T]\times (0,\infty ))$ and Lemma 4.4 , $d(\mathcal{L}v_1,\mathcal{L}v_2)$ in (4.5) satisfies the following inequality:

(4.7) \begin{equation} \begin{aligned} d(\mathcal{L}v_1,\mathcal{L}v_2)\leq \Delta \left (\int _t^Te^{\int \limits _t^s\widehat {A}(\tau ,T)d\tau -\alpha (s-t)}ds\right )d(v_1,v_2), \end{aligned} \end{equation}

where $\widehat {A}(\tau ,T)=(c_1-\tilde {c}_1+c_2-\tilde {c}_2)A(\tau ,T)\gt 0.$

The proof of Lemma 4.5 is provided in Appendix D.5.

Theorem 4.6. Suppose that joint distributions $\Phi (k,l)$ and $\Psi (y,j_2)$ as well as marginal distributions $H(k), D(l), F_2(j_2)$ , and $G(y)$ follow Assumptions 2.3 and 2.4 and the distribution $F_1(j_1)$ follows Assumption 2.3 , the mapping $\mathcal{L}$ : $\mathfrak{C}_e([0,T]\times (0,\infty ))$ $\to$ $\mathfrak{C}_e([0,T]\times (0,\infty ))$ is a contraction mapping under the metric (4.4) for some positive constant $\alpha$ , where a feasible range of $\alpha$ follows $\left \{\alpha \gt 0|\Delta \frac {\exp [(\gamma (\tilde {c}_1-c_1+\tilde {c}_2-c_2)e^{rT}-\alpha )T]-1}{\gamma (\tilde {c}_1-c_1+\tilde {c}_2-c_2)e^{rT}-\alpha }\lt 1\right \}$ and $\Delta$ is given by Lemma 4.4 . Furthermore, the iterative method in (3.13) provides the first order of convergence.

The proof of Theorem4.6 is provided in Appendix D.6.

5. Numerical results

In this section, we provide several numerical examples to demonstrate the optimal asset allocation policy, optimal reinsurance policies for ordinary claims and catastrophic claims, and the value function under a copula. Gaussian copulasFootnote 2 is assumed to dominate the dependence structure of $\Psi (y,j_2)$ and $\Phi (k,l)$ with default linear correlation parameters $\rho _1=\rho _{(y,j_2)}=0$ and $\rho _2=\rho _{(k,l)}=0$ , respectively.

Furthermore, we assume that ordinary claims ( $J_1$ ) follow exponential random variables with the parameter $\epsilon _1\gt 0$ and marginal distributions of catastrophic claims ( $J_2$ ), self-excited jumps ( $Y$ ), externally excited jumps ( $K$ ), and risky-asset crashes ( $L$ ) also follow exponential random variables with parameters $\epsilon _2\gt 0$ , $\epsilon _3\gt 0$ , $\epsilon _4\gt 0$ , and $\epsilon _5\gt 0$ , respectively. The values for exponential random variable parameters of $\epsilon _2\gt 0$ , $\epsilon _3\gt 0$ , and $\epsilon _4\gt 0$ are chosen for their simplicity, where their means are the same. The value for the average risky-asset crash is chosen to be $20\%$ .

The parameters of the financial and insurance market mainly follow the works of Meng et al. (2015), Hainaut (Reference Hainaut2017), and Liu et al. (2021). Unless otherwise stated, the value of parameters can be referred to in Table 1. To describe the insurance market, it is assumed that $\eta =1$ , which implies that, on average, an ordinary claim occurs once per time unit. This setting ensures a reasonable claim frequency that aligns with empirical observations in general insurance markets. For the catastrophic claim process, we assume that $a=2$ and $\delta =4$ , meaning that the mean-reverting level for catastrophic claims is twice per time unit, and the excitation effects decay by around $80\%$ per time unit. These choices reflect the characteristics of catastrophic events, which typically exhibit temporary clustering due to external shocks but eventually revert to a stable level over time. The fast decay rate of excitation captures the reality that while major disasters can trigger secondary events, their influence diminishes relatively quickly.

Table 1. Parameters and values

The claim severity assumptions are also well-founded. The average ordinary claim size is set to $0.25$ units, which is significantly smaller than the average catastrophic claim size of $0.5$ units. This distinction is consistent with empirical data, as catastrophic claims tend to have much larger impacts than ordinary claims. Moreover, the externally exciting and self-exciting effects are assumed to be $40\%$ of the mean-reverting level. This reflects a high-risk insurance market where disasters and systemic risks have a substantial impact, potentially leading to claim clustering and financial stress for insurers. Such a setting is particularly relevant in regions prone to natural disasters or extreme events. The assumption that $\tilde {\theta }_i\gt \theta _i$ , $i=1,2,$ ensures that reinsurance is not overly cheap for both ordinary and catastrophic claims. This prevents moral hazard and ensures that insurers are incentivized to manage risks properly rather than relying excessively on reinsurance. It also aligns with real-world market conditions, where reinsurers charge premiums that reflect the risk exposure they assume.

For the financial market, the risk-free rate is set at $0.02$ , with the risky asset having a drift of $0.05$ and volatility of $0.2$ . These parameters are broadly in line with historical financial market trends, where the excess return of risky assets over the risk-free rate is moderate, and volatility remains within a reasonable range. The relatively low risk-free rate reflects modern economic conditions, where safe investments yield limited returns. Overall, these parameter settings provide a realistic and well-calibrated representation of both the insurance and financial markets, capturing key risk characteristics while ensuring model robustness.

5.1 Optimal reinsurance policy

5.1.1 Ordinary claims

Figure 1 illustrates the behavior of the optimal reinsurance policy with respect to ordinary claims, where the closed-form expression under exponential marginal distributions is obtained in Lemma 3.2. First, the insurer transfers a little bit more risk on ordinary claims to the reinsurer when the operating time increases (as $t$ decreases). Second, note that there is no linear relationship between the optimal reinsurance policy with respect to ordinary claims and the initial time point, shown by Lemma 3.2. Figure 1 looks like linear functions because the scope of the time interval is small. Third, $q_1^*(t)$ increases as the average claim size decreases (as $\epsilon _1$ increases), which means more risks are retained by the insurer as the average ordinary claim size decreases. Finally, comparing Figure 1a and b, we can see that $q_1^*(t)$ under the net-profit principle is smaller than that under the deviation-premium principle. Since ordinary claims are assumed to be exponentially distributed, the reinsurance premium under the deviation-premium principle is larger than that under the net-profit principle with the same risk loadings. To maximize the expected terminal surplus, the insurer reduces the reinsurance cost by retaining a larger proportion of ordinary claim risks.

Figure 1 Optimal ordinary reinsurance policy under dynamic contagion processes.

5.1.2 Catastrophic claims

Figure 2 shows the behavior of the optimal reinsurance strategy with respect to catastrophic claims under different parameter settings and two premium principles. First, the optimal reinsurance policy $q^*_2(t,\lambda )$ decreases as the claim intensity increases in general. When the claim intensity is small, the insurer rejects the proportional reinsurance products ( $q_2^*(t,\lambda )=1$ ). According to (2.4), externally exciting and self-exciting factors can only have positive effects on claim intensities. Hence, when the claim intensity is low, the insurer recognizes that the intensity will eventually revert to at least the constant mean-reverting level. This predictability allows the insurer to assess its baseline risk exposure with confidence. Since claim occurrences are expected to remain manageable at such levels, the insurer can rely on its own capital reserves and risk management strategies rather than purchasing costly reinsurance. Moreover, in a low-intensity regime, the probability of sudden claim clustering remains relatively low, reducing the insurer’s concern about extreme losses. As a result, retaining more risk is financially optimal, as the insurer avoids paying high reinsurance premiums for coverage that may not be immediately necessary. This aligns with standard industry practice, where insurers typically increase retention when claim risks are stable and foreseeable. However, when the claim intensity is large, the insurer is unsure about the behavior of claim intensities. The mean-reverting feature will move the claim intensity back to the constant mean-reverting level, while the externally exciting and self-exciting effects will further increase the claim intensity. To mitigate the potential contagious claim risk, the insurer transfers those risks to the reinsurer by purchasing the proportional reinsurance contract.

Second, when catastrophic claims and self-exciting effects are negatively correlated (i.e., $\rho _1\lt 0$ ), catastrophic claim risks decrease, and the insurer can bear more catastrophic risks by itself. Otherwise, more catastrophic risks are transferred to the reinsurer when $\rho _1\gt 0$ , as shown by Figure 2a, c, and e under the net-profit principle and Figure 2b, d, and f under the deviation-premium principle.

Third, the effect of $\rho _2$ on $q^*_2(t,\lambda )$ is negligible. This is because $\rho _2$ measures the dependence structure between asset crashes and the external factor on the claim intensity, which only has an indirect effect on the insurance market. The insurer can adjust the asset allocation policy accordingly and maintain the optimal reinsurance policy for catastrophic claims.

Finally, comparing the left column and the right column in Figure 2, we can see that $q^*_2(t,\lambda )$ under the deviation-premium principle is slightly larger than that under the net-profit principle. According to Theorem2.6, we know that the reinsurance premium of catastrophic claims under the deviation-premium principle is larger than that under the net-profit principle. To maximize the expected terminal surplus, the insurer also reduces the reinsurance cost by retaining a larger proportion of catastrophic claim risks.

Figure 2 Optimal catastrophic reinsurance policy under dynamic contagion processes.

5.2 Optimal asset allocation strategy

Figure 3 demonstrates the behavior of the optimal asset allocation strategy under different parameter settings and two premium principles. First, the optimal asset allocation policy decreases as the increase of $t$ , which implies that the insurer decreases the holdings of the risky asset under the contagious market when the operating time decreases (the initial time $t$ increases to the terminal date $T$ ). When $t$ is small, claim risks dominate. The insurer invests more in risky assets to utilize the high growth rate and mitigate the insurance risk, even though negative jumps could appear in the risky asset. As $t$ goes to the $T$ , claims risks decrease. The insurer decreases risky asset holdings to reduce total risks from the contagious financial and insurance markets.

Second, $\pi ^*(t,\lambda )$ decreases as $\lambda$ increases. The insurer decreases risky asset holdings to reduce total risks from the contagious financial and insurance markets when claim risks increase. Third, comparing Figure 3a, c, and e under the net-profit principle and Figure 3b, d, and f under the deviation-premium principle, we observe that $\pi ^*(t,\lambda )$ is affected by the correlation between catastrophic claims and self-exciting effects. When catastrophic claims and self-exciting effects are positively correlated (i.e. $\rho _1\gt 0$ ), catastrophic claim risks are magnified. The insurer decreases the risky asset holdings to reduce the total risk level. Otherwise, when $\rho _1\lt 0,$ a larger amount of resources can be allocated to the risky assets to take advantage of the higher growth rate.

Fourth, comparing Figure 3g and i under the net-profit principle and Figure 3h and j under the deviation-premium principle, we observe that $\pi ^*(t,\lambda )$ can also be affected by the correlation between asset crashes and externally exciting effects. When asset crashes and externally exciting effects are positively correlated (i.e. $\rho _2\gt 0$ ), the financial risk increases. The insurer decreases the risky asset holdings to reduce the total risk level. Otherwise, when $\rho _2\lt 0$ , more resources can be allocated to risky assets. Finally, $\pi ^*(t,\lambda )$ under the deviation-premium principle is larger than that under the net-profit principle. This is because the more insurance risks are retained by the insurer under the deviation-premium principle, the more premiums are collected by the insurer. Hence, the insurer can allocate more resources to the risky asset to have larger investment returns and cover future potential obligations.

Figure 3 Optimal asset allocation policy under dynamic contagion processes.

5.3 Value function

Figure 4 shows the value function $V(t,1,\lambda )$ with a fixed initial surplus $x=1$ under different parameter settings and two premium principles. First, it is shown that the value function $V(t,1,\lambda )$ decreases as the increase of $t$ , which implies that the insurer can effectively manage the contagious claim risks to obtain a larger expected utility of the terminal surplus as the operating time increases. Second, the value function decreases with the increase of initial claim intensity. This is because claim risks have negative effects on the terminal surplus. As time moves to the terminal time point, the negative effect of the initial claim intensity decreases gradually.

Third, comparing Figure 4a, c, and e under the net-profit principle and Figure 4b, d, and f under the deviation-premium principle, we observe that the positive correlation between catastrophic claims and self-exciting effects $(\rho _1\gt 0)$ has a negative impact on the insurer’s value function, while a negative correlation $(\rho _1\lt 0)$ enhances it. This result reflects the economic intuition that when larger catastrophic claims trigger larger self-exciting effect, the overall risk exposure of the insurer increases, leading to a lower value function. In contrast, when $\rho _1\lt 0$ , catastrophic claims are less likely to generate additional shocks, reducing long-term risk accumulation and improving the insurer’s financial position. This suggests that in markets where catastrophic claims exhibit strong self-excitation, insurers face higher systemic risk and may require more robust capital reserves or reinsurance strategies. Furthermore, comparing Figure 4g and i under the net-profit principle and Figure 4h and j under the deviation-premium principle, we observe that the correlation between asset crashes and externally exciting effects $\rho _2$ also influences the value function. Specifically, a positive correlation $(\rho _2\gt 0)$ moderately reduces the value function, while a negative correlation $(\rho _2\lt 0)$ enhances it. This effect is more pronounced when both $\lambda$ and $t$ are small.

The economic interpretation of this result is that when asset crashes and externally exciting effects are positively correlated, financial market downturns coincide with heightened claim risks, exacerbating the insurer’s financial distress. This increases the likelihood of capital depletion and reduces the insurer’s overall value. Conversely, when $\rho _2\lt 0$ , financial downturns are less likely to be accompanied by increased claim risk, allowing the insurer to better manage its capital and risk exposure, ultimately leading to a higher value function. This finding highlights the importance of diversification in risk management – insurers operating in environments where financial and insurance risks are highly correlated may need stronger capital buffers or adaptive reinsurance strategies to mitigate the compounded risks.

Finally, the value function under the deviation-premium principle is slightly larger than that under the net-profit principle. This is just because the insurance premium under the deviation-premium principle is larger than that under the net-profit principle. The insurer has more resources to deal with the contagious financial and insurance markets.

Figure 4 Value function under dynamic contagion processes.

6. Conclusions

In this paper, we investigate the optimal asset allocation and reinsurance problem for an insurance company under dynamic contagion processes, where the insurer has two business lines: ordinary insurance products and catastrophic insurance products. Ordinary claims are modeled by the standard Poisson process, and catastrophic claims are captured by an enhanced dynamic contagion process, where we accommodate the dependency structure between the magnitude of contribution to intensity after initial events for catastrophic insurance products and their claim/loss size. Furthermore, the dependence structure between the financial market and the insurance market is considered. Solutions of the optimal reinsurance strategy with respect to ordinary claims under two different premium principles are derived explicitly. To solve the optimal asset allocation strategy and reinsurance policy with respect to catastrophic claims, we develop a comprehensive iterative scheme based on the Feynman–Kac formula to solve the value function numerically. We rigorously prove the existence and uniqueness of the solution to the fixed point equation and the convergence of our iterative scheme as long as the joint distributions of self-excited jumps and catastrophic claims, asset crashes and externally excited jumps, as well as marginal distributions, have finite Laplace transforms. Finally, numerical examples are presented to consider behaviors of the value function and optimal controls under different parameter settings and premium principles. Choosing the optimal investment and reinsurance strategy is vital to insurers dealing with losses arising from catastrophic events.

Undoubtedly, improved models are required to predict losses arising from catastrophic events, as extreme risks pose a challenge to insurers and their financial viability. So our results provide several new ways for insurers’ investment-reinsurance decision-making, allowing a more general formulation of event risk, with random variation in both the frequency and size of jumps and their dependencies. Incorporating real-world constraints on reinsurance for catastrophic claims, such as coverage caps, and imposing the dependence structure between ordinary claims and external events can be considered in the proposed approach, which we leave for further research. This work can also be extended by using multivariate dynamic contagion processes for optimal decision-making on multiple business lines.

Data availability statement

The data and code that support the findings of this study are available from the first author, Dr. Liu, upon reasonable request.

Funding statement

There was no external funding.

Competing interests

None.

Appendix

A. Proof of Theorem2.6

Based on Dassios & Zhao (Reference Dassios and Zhao2012), the infinitesimal generator of the compound-enhanced dynamic contagion process $ \left ( \lambda (t),N_2(t),C_2(t),t\right )$ acting on a function $f\left ( \lambda ,n,c,t\right )$ within its domain $\Omega \left ( \mathcal{A} \right )$ is given by

(A.1) \begin{align} \mathcal{A}\mathrm{ }f\left ( \lambda ,n,c,t\right ) & =\frac {\partial f}{ \partial t}+\delta \left ( a-\lambda \right ) \frac {\partial f}{\partial \lambda } \nonumber \\[4pt] & \quad +\lambda \left [ \int \limits _{0}^{\infty }\int \limits _{0}^{\infty }f\left ( \lambda +y,n+1,c+j_2,t\right ) d\Psi \left ( y,j_2\right ) -f\left ( \lambda ,n,c,t\right ) \right ] \nonumber \\[4pt] & \quad +\rho \left [ \int \limits _{0}^{\infty }f\left ( \lambda +k,n,c,t\right ) dH(k)-f\left ( \lambda ,n,c,t\right ) \right ], \end{align}

where $\Omega \left ( \mathcal{A}\right )$ is the domain of the generator $\mathcal{A}$ such that $f\left ( \lambda ,n,c,t\right )$ is differentiable with respect to $\lambda$ and $t$ for all $\lambda$ , $n$ , $c$ , and $t$ , and

\begin{equation*} \left \vert \int \limits _{0}^{\infty }\int \limits _{0}^{\infty }f\left ( \lambda +y,n+1,c+j_2,t\right ) d\Psi \left ( y,j_2\right ) -f\left ( \lambda ,n,c,t\right ) \right \vert \lt \infty \mathrm{,} \end{equation*}
\begin{equation*} \left \vert \int \limits _{0}^{\infty }f\left ( \lambda +k,n,c,t\right ) dH(k) -f\left ( \lambda ,n,c,t\right ) \right \vert \lt \infty \mathrm{.} \end{equation*}

A.1 Under the net-profit principle

Setting $\mathcal{A}$ $f\left ( \lambda ,n,c,t\right )=c$ in (A.1), we have

(A.2) \begin{equation} \mathcal{A} c=\lambda \int \limits _{0}^{\infty }\int \limits _{0}^{\infty }j_2d\Psi (y,j_2)=\lambda m_{1j_2}, \end{equation}

where $m_{1j_2}=\int \limits _0^{\infty }j_2dF_2(j_2).$

As $C_2(t)-C_2(0)-\int \limits _{0}^{t}\mathcal{A}c(s)ds$ is a martingale and $C_2(0)=0$ , we have

(A.3) \begin{equation} \mathbb{E}[C_2(t)|\lambda (0)]=m_{1j_2}\int \limits _0^{t}\mathbb{E}[\lambda (s)|\lambda (0)]ds, \end{equation}

Now, setting $\mathcal{A}$ $f\left ( \lambda ,n,c,t\right )=\lambda$ , we have

(A.4) \begin{equation} \mathcal{A} \lambda =-(\delta -m_{1y})\lambda +a\delta +\rho m_{1k}, \end{equation}

where $m_{1y}=\int \limits _0^{\infty }ydG(y)$ and $m_{1k}=\int \limits _0^{\infty }kdH(k).$ Similarly, we have a martingale $\lambda (t)-\lambda (0)-\int \limits _{0}^{t}\mathcal{A}\lambda (s) ds$ , which implies that

(A.5) \begin{equation} \mathbb{E}[\lambda (t)|\lambda (0)]=\lambda (0)-(\delta -m_{1y})\int \limits _0^t\mathbb{E}[\lambda (s)|\lambda (0)]ds+(a\delta +\rho m_{1k})t. \end{equation}

Differentiating (A.5) $w.r.t$ $t$ , we can obtain a non-linear inhomogeneous ODE

(A.6) \begin{equation} \frac {d\left (\mathbb{E}[\lambda (t)|\lambda (0)]\right )}{dt}=-(\delta -m_{1y})\mathbb{E}[\lambda (t)|\lambda (0)]+a\delta +\rho m_{1k}, \end{equation}

with the initial condition $\mathbb{E}[\lambda (0)|\lambda (0)]=\lambda (0)$ . Hence, this ODE (A.6) has the following solution for $\delta \gt m_{1y}$

(A.7) \begin{equation} \mathbb{E}[\lambda (t)|\lambda (0)]=\frac {a\delta +\rho m_{1k}}{\delta -m_{1y}}+\left (\lambda (0)-\frac {a\delta +\rho m_{1k}}{\delta -m_{1y}}\right )e^{-(\delta -m_{1y})t}. \end{equation}

Therefore, (A.3) can be rewritten as

(A.8) \begin{equation} \mathbb{E}[C_2(t)|\lambda (0)]=m_{1j_2}\left [\frac {a\delta +\rho m_{1k}}{\delta -m_{1y}}t+\left (\lambda (0)-\frac {a\delta +\rho m_{1k}}{\delta -m_{1y}}\right )\frac {1}{\delta -m_{1y}}\left (1-e^{-(\delta -m_{1y})t}\right )\right ]. \end{equation}

By letting $t\rightarrow \infty$ , we have

(A.9) \begin{equation} \mathbb{E}[C_2(t)]=\frac {a\delta +\rho m_{1k}}{\delta -m_{1y}}m_{1j_2}t. \end{equation}

Hence, according to the net-premium principle, we have $c_2=(1+\theta _2)\frac {m_{1k}\rho +a\delta }{\delta -m_{1y}}m_{1j_2}$ .

A.2 Under the deviation-premium principle

First, we claim that the joint conditional expectation of $\lambda (t)C_2(t)$ is given by

(A.10) \begin{align} \mathbb{E}[ \lambda (t)C_{2}(t)\mid \lambda (0)]&=\lambda (0)C_2(0)e^{-\left ( \delta -m_{1y}\right ) t}+\left ( a\delta +\rho m_{1k} \right )e^{-\left ( \delta -m_{1y}\right )t}\int \limits _{0}^{t}e^{( \delta -m_{1_{y}}) s}\mathbb{E}[C_2(s)|\lambda (0)] ds \nonumber\\ &\quad +m_{1_{j_{2}}}e^{-( \delta -m_{1_{y}})t}\int \limits _{0}^{t}e^{( \delta -m_{1_{y}}) s}\mathbb{E}[\lambda ^2(s)|\lambda (0)] ds \nonumber\\ &\quad +m_{1\Psi _{y,j_{2}}}e^{-( \delta -m_{1_{y}}) t}\int \limits _{0}^{t}e^{( \delta -m_{1_{y}}) s}\mathbb{E}[\lambda (s)|\lambda (0)] ds, \end{align}

where

(A.11) \begin{equation} m_{1\Psi _{y,j_{2}}}=\int \limits _{0}^{\infty }\int \limits _{0}^{\infty }yj_{2}d\Psi ( y,j_{2}). \end{equation}

The proof of the claim is as follows:

Setting $\mathcal{A}$ $f\left ( \lambda ,n,c,t\right ) =\lambda c$ , we have

\begin{align*} \mathcal{A}\lambda c & = \delta \left ( a-\lambda \right ) c+\lambda \left [ \int \limits _{0}^{\infty }\int \limits _{0}^{\infty }\left ( \lambda +y\right ) \left ( c+j_{2}\right ) d\Psi ( y,j_{2}) -\lambda c\right ] +\rho \left [ \int \limits _{0}^{\infty }(\lambda +k) cdH(k) -\lambda c\right ] \\[3pt] &= \delta \left ( a-\lambda \right ) c+\lambda \left [ \int \limits _{0}^{\infty }\int \limits _{0}^{\infty }\left ( \lambda c+\lambda j_{2}+yc+yj_{2}\right ) d\Psi ( y,j_{2}) -\lambda c\right ]\\[3pt] & \quad +\rho \left [ \int \limits _{0}^{\infty }\left ( \lambda c+kc\right ) dH(k) -\lambda c\right ]\end{align*}
\begin{align*} &= \delta \left ( a-\lambda \right ) c+\lambda \left [ \int \limits _{0}^{\infty }\int \limits _{0}^{\infty }\left ( \lambda j_{2}+yc+yj_{2}\right ) d\Psi ( y,j_{2}) \right ] +\rho \left [ \int \limits _{0}^{\infty }kcdF(k) \right ] \\[6pt] &= \delta \left ( a-\lambda \right ) c+\lambda \left [ \int \limits _{0}^{\infty }\int \limits _{0}^{\infty }\left ( \lambda j_{2}+yc+yj_{2}\right ) d\Psi ( y,j_{2}) \right ] +\rho cm_{1k} \\[6pt] &= \delta \left ( a-\lambda \right ) c \\[6pt] & \quad +\lambda \left ( \lambda \int \limits _{0}^{\infty }\int \limits _{0}^{\infty }j_{2}d\Psi ( y,j_{2}) +c\int \limits _{0}^{\infty }\int \limits _{0}^{\infty }yd\Psi ( y,j_{2}) +\int \limits _{0}^{\infty }\int \limits _{0}^{\infty }yj_{2}d\Psi ( y,j_{2}) \right ) +\rho cm_{1k} \\[6pt] &= \delta \left ( a-\lambda \right ) c+m_{1j_{2}}\lambda ^{2}+m_{1y}\lambda c+\lambda m_{1\Psi _{y,y_{2}}}+\rho cm_{1k} \\ &= \left ( \delta a+\rho m_{1k}\right ) c+m_{1j_{2}}\lambda ^{2}+m_{1\Psi _{y,j_{2}}}\lambda -\left ( \delta -m_{1y}\right )\lambda c. \end{align*}

As $\lambda (t)C_2(t)-\lambda (0)C_2(0)-\int \limits _{0}^{t}\mathcal{A}$ $ \left ( \lambda (s)c(s)\right ) ds$ is a martingale, we have

(A.12) \begin{equation} \mathbb{E}\left \{ \lambda (t)C_2(t)-\int \limits _{0}^{t}\mathcal{A}\left ( \lambda (s)C_2(s)\right ) ds\mid \lambda (0)\right \} =\lambda (0)C_2(0). \end{equation}

Hence,

(A.13) \begin{align} \mathbb{E}\left [ \lambda (t)C_2(t)\mid \lambda (0)\right ] &=\lambda (0)C_2(0)+\mathbb{E}\left \{ \int \limits _{0}^{t}\mathcal{A}\left ( \lambda (s)C_2(s)\right ) ds\mid \lambda (0)\right \} \nonumber\\ &=\lambda (0)C_2(0)-\left ( \delta -m_{1y}\right ) \int \limits _{0}^{t}\mathbb{E}\left ( \lambda (s)C_2(s)\mid \lambda (0)\right ) ds\nonumber\\ & \quad +\left ( \delta a+\rho m_{1k}\right ) \int \limits _{0}^{t}\mathbb{E}\left ( C_2(s)\mid \lambda (0)\right ) ds \nonumber\\ &\quad +m_{1j_{2}}\int \limits _{0}^{t}\mathbb{E}\left [ \left \{ \lambda (s)\right \} ^{2}\mid \lambda (0)\right ] ds+m_{1\Psi _{y,j_{2}}}\int \limits _{0}^{t}\mathbb{E}\left ( \lambda (s)\mid \lambda (0)\right ) ds, \end{align}

by differentiating with respect to $t$ , we obtain the non-linear inhomogeneous ODE,

(A.14) \begin{align} \frac {d\mathbb{E}\left [\lambda (t)C_2(t)\mid \lambda (0)\right ] }{dt}&=-\left ( \delta -m_{1y}\right ) \mathbb{E}\left [ \lambda (t)C_2(t)\mid \lambda (0)\right ] \nonumber\\ & \quad +\left ( \delta a+\rho m_{1k}\right ) \mathbb{E}\left [ C_2(t)\mid \lambda (0)\right ] +m_{1j_{2}}\mathbb{E}\left [ \lambda ^2(t)\mid \lambda (0)\right ]\nonumber\\ & \quad +m_{1\Psi _{y,j_2}}\mathbb{E}\left [\lambda (t)\mid \lambda (0)\right ], \end{align}

with the initial condition $\mathbb{E}\left [ \lambda (0)C_2(0)\mid \lambda (0)\right ] =\lambda (0)C_2(0)$ . Hence, the solution of the ODE is given by

(A.15) \begin{align} \mathbb{E}\left [\lambda (t)C_2(t)\mid \lambda (0)\right ] &=\lambda (0)C_2(0)e^{-\left ( \delta -m_{1y}\right ) t}\nonumber\\& \quad +\left ( a\delta +m_{1k}\rho \right ) e^{-\left ( \delta -m_{1y}\right ) t}\int \limits _{0}^{t}e^{\left ( \delta -m_{1y}\right ) s}\mathbb{E}\left ( C_2(s)\mid \lambda (0)\right ) ds \nonumber\\ & \quad + m_{1j_{2}}e^{-\left ( \delta -m_{1y}\right ) t}\int \limits _{0}^{t}e^{\left ( \delta -m_{1y}\right ) s}\mathbb{E}\left ( \lambda ^2(s)\mid \lambda (0)\right ) ds \nonumber\\ & \quad + m_{1\Psi _{y,j_{2}}}e^{-\left ( \delta -m_{1y}\right ) t}\int \limits _{0}^{t}e^{\left ( \delta -m_{1y}\right ) s}\mathbb{E}\left ( \lambda (s)\mid \lambda (0)\right ) ds. \end{align}

From Lemma 3.1 in Dassios & Zhao (Reference Dassios and Zhao2011), we have, for $\delta \neq m_{1y}$ ,

(A.16) \begin{align} \mathbb{E}\left [ \lambda ^{2}(t)\mid \lambda (0)\right ] & = \lambda ^{2}(0)e^{-2\left ( \delta -m_{1y}\right ) t}\nonumber\\[5pt] &\quad +\frac {2\left ( m_{1k}\mathrm{ }\rho +a\delta \right ) +m_{2y}}{\delta -m_{1y}}\left ( \lambda (0)-\frac {\rho m_{1k}+a\delta }{\delta -m_{1y}}\right ) \left ( e^{-\left ( \delta -m_{1y}\right ) t}-e^{-2\left ( \delta -m_{1y}\right ) t}\right ) \nonumber\\[5pt] &\quad +\left [ \frac {\left \{ 2\left ( \rho m_{1k} +a\delta \right ) +m_{2y}\right \} \left (\rho m_{1k}+a\delta \right ) }{2\left ( \delta -m_{1y}\right ) ^{2}}+\frac {\rho m_{2k} }{2\left ( \delta -m_{1y}\right ) }\right ] \left ( 1-e^{-2\left ( \delta -m_{1y}\right ) t}\right ), \end{align}

where $m_{2y}$ and $m_{2k}$ are the second moment of $Y$ and $K,$ respectively. Given (A.7), (A.8), and (A.16), we prove that the solution of the ODE is given by (A.10).

Proposition A.1. For the stationary distribution of the process $\lambda (t)$ , where $\delta \gt m_{1y}$ , given $C_2(0)=0$ , the joint expectation of $\lambda (t)C_2(t)$ is given by

(A.17) \begin{equation} \begin{aligned} \mathbb{E}\left [\lambda (t)C_2(t)\right ] &=\left ( a\delta +\rho m_{1k} \right ) m_{1j_{2}}\left \{ \frac {\rho m_{1k} +a\delta }{\left ( \delta -m_{1y}\right ) ^{2}}\right \} \left \{ t-\left ( \frac {1-e^{-\left ( \delta -m_{1y}\right ) t}}{\delta -m_{1y}}\right ) \right \} \\[5pt] & \quad +m_{1j_2}\left ( \frac {1-e^{-\left ( \delta -m_{1y}\right ) t}}{\delta -m_{1y}}\right ) \left [ \frac {\left ( 2\left (\rho m_{1k} +a\delta \right ) +m_{2y}\right ) \left (\rho m_{1k} +a\delta \right ) }{2\left (\delta -m_{1y}\right )^2}+\frac {\rho m_{2k}}{2\left ( \delta -m_{1y}\right )}\right ] \\[5pt] &\quad + m_{1\Psi _{y,j_{2}}}\left ( \frac {1-e^{-\left ( \delta -m_{1y}\right ) t}}{\delta -m_{1y}}\right ) \left ( \frac {\rho m_{1k} +a\delta }{\delta -m_{1y}}\right ) , \end{aligned} \end{equation}

where $m_{2y}$ and $m_{2k}$ are the second moment of $Y$ and $K,$ respectively.

The proof of Proposition A.1 is straightforward by letting $t\rightarrow \infty$ in (A.7) and (A.16) and using (A.9) for (A.10); then the result follows immediately given $C_2(0)=0.$

Theorem A.2. The conditional expectation of $C^2_2(t)$ given $\lambda (0)$ at time $t=0$ is given by

(A.18) \begin{align} \mathbb{E}\left [ C^2_2(t)\mid \lambda (0)\right ]&=C^2(0)+2m_{1j_{2}}\int \limits _{0}^{t}\mathbb{E}[\lambda (s)C_2(s)|\lambda (0)]ds+m_{2j_{2}}\int \limits _{0}^{t}\mathbb{E}[\lambda (s)|\lambda (0)]ds. \end{align}

Proof. Setting $\mathcal{A}$ $f\left ( \lambda ,n,c,t\right ) =c^{2}$ , we have

(A.19) \begin{equation} \begin{aligned} \mathcal{A}\mathrm{ }c^{2} &=\lambda \left [ \int \limits _{0}^{\infty }\int \limits _{0}^{\infty }\left ( c+j_{2}\right ) ^{2}d\Psi ( y,j_{2}) -c^{2}\right ] +\rho \left [ \int \limits _{0}^{\infty }c^{2}dH(k) -c^{2}\right ] \\[6pt] &=\lambda \left [ \int \limits _{0}^{\infty }\int \limits _{0}^{\infty }\left ( c^{2}+2cj_{2}+j_{2}^{2}\right ) d\Psi ( y,j_{2}) -c^{2}\right ] \\[6pt] &=2m_{1j_{2}}\lambda c+m_{2j_{2}}\lambda , \end{aligned} \end{equation}

where

(A.20) \begin{eqnarray} m_{1j_{2}} &=&\int \limits _{0}^{\infty }\int \limits _{0}^{\infty }j_{2}d\Psi ( y,j_{2}) =\int \limits _{0}^{\infty }j_{2}\left [ \int \limits _{0}^{\infty }\psi ( y,j_{2}) dy\right ] dj_{2}=\int \limits _{0}^{\infty }j_{2}dF_2(j_{2}), \end{eqnarray}
(A.21) \begin{eqnarray} m_{2j_{2}} &=&\int \limits _{0}^{\infty }\int \limits _{0}^{\infty }j_{2}^{2}d\Psi ( y,j_{2}) =\int \limits _{0}^{\infty }j_{2}^{2} \left [ \int \limits _{0}^{\infty }\psi ( y,j_{2}) dy\right ] dj_{2}=\int \limits _{0}^{\infty }j_{2}^{2}dF_2(j_{2}). \end{eqnarray}

As $C_2^2(t)-C_2^2(0)-\int \limits _{0}^{t}\mathcal{A}$ $C_2^2(s)ds$ is a martingale, we have

(A.22) \begin{equation} \mathbb{E}\left \{ C_2^2(t)-\int \limits _{0}^{t}\mathcal{A}\mathrm{ }C_2^2(s)ds\mid \lambda (0)\right \} =C_2^2(0). \end{equation}

Hence,

(A.23) \begin{equation} \begin{aligned} \mathbb{E}\left ( C_2^2(t)\mid \lambda (0)\right ) &=C_2^2(0)+\mathbb{E}\left \{ \int \limits _{0}^{t}\mathcal{A}\mathrm{ }C_2^2(s)ds\mid \lambda (0)\right \}\\[6pt] &=C_2^2(0)+2m_{1j_{2}}\int \limits _{0}^{t}\mathbb{E}\left ( \lambda (s)C(s)\mid \lambda (0)\right ) ds+m_{2j_{2}}\int \limits _{0}^{t}E\left ( \lambda (s)\mid \lambda (0)\right ) ds. \end{aligned} \end{equation}

Hence, we complete the proof.

Corollary A.3. The conditional variance of $C_2(t)$ given $\lambda (0)$ at time $t=0$ is given by

(A.24) \begin{equation} Var\left ( C_2(t)\mid \lambda (0)\right )=\mathbb{E}\left [ C_2^2(t)\mid \lambda (0)\right ]-\left \{\mathbb{E}\left [ C_2(t)\mid \lambda (0)\right ]\right \}^2, \end{equation}

where $\mathbb{E}\left [ C_2^2(t)\mid \lambda (0)\right ]$ and $\mathbb{E}\left [ C_2(t)\mid \lambda (0)\right ]$ are given by (A.18) and (A.8), respectively.

Finally, we can derive the main results as follows: given $\delta \gt m_{1y}$ and $C_2(0)=0$ , the second moment of the process of $C_(t)$ is given by

\begin{equation*} E\left \{ C_2^2(t)\right \} \end{equation*}
(A.25) \begin{align} & = 2m_{1j_{2}}\left [ \begin{array}{c} \left ( \frac {m_{1k}\mathrm{ }\rho +a\delta }{2}\right ) m_{1j_{2}}\left \{ \frac {m_{1k}\mathrm{ }\rho +a\delta }{\left ( \delta -m_{1y}\right ) ^{2}} \right \} t^{2} \\[15pt] -\left ( m_{1k}\mathrm{ }\rho +a\delta \right ) m_{1j_{2}}\left \{ \frac { m_{1k}\mathrm{ }\rho +a\delta }{\left ( \delta -m_{1y}\right ) ^{3}} \right \} \left \{ t-\left ( \frac {1-e^{-\left ( \delta -m_{1y}\right ) t}}{ \delta -m_{1y}}\right ) \right \} \\[15pt] +m_{1j_{2}}\mathrm{ }\left [ \frac {\left \{ 2\left ( m_{1k}\mathrm{ }\rho +a\delta \right ) +m_{2y}\right \} \left ( m_{1k}\mathrm{ }\rho +a\delta \right ) }{2\left ( \delta -m_{1y}\right ) ^{3}}+\frac {m_{2k}\mathrm{ }\rho }{2\left ( \delta -m_{1y}\right ) ^{2}}\right ] \left \{ t-\left ( \frac { 1-e^{-\left ( \delta -m_{1y}\right ) t}}{\delta -m_{1y}}\right ) \right \} \\[15pt] +m_{1\Psi _{y,j_{2}}}\mathrm{ }\left \{ \frac {m_{1k}\mathrm{ }\rho +a\delta }{\left ( \delta -m_{1y}\right ) ^{2}}\right \} \left \{ t-\left ( \frac { 1-e^{-\left ( \delta -m_{1y}\right ) t}}{\delta -m_{1y}}\right ) \right \} \end{array} \right ] \nonumber \\[15pt] & \quad + m_{2j_{2}}\left ( \frac {m_{1k}\mathrm{ }\rho +a\delta }{\delta -m_{1y}}\right ) t. \end{align}

The above result can be easily proved by letting $t\rightarrow \infty$ in (A.7) and (A.18) and using (A.17) in (A.18).

Furthermore, we have

\begin{equation*} Var\left ( C_2(t)\right )=\mathbb{E}[C_2^2(t)]-\{\mathbb{E}[C_2]\}^2 \end{equation*}
(A.26) \begin{align} & = 2m_{1j_{2}}\left [ \begin{array}{c} m_{1j_{2}}\mathrm{ }\left [ \frac {\left \{ 2\left ( m_{1k}\mathrm{ }\rho +a\delta \right ) +m_{2y}\right \} \left ( m_{1k}\mathrm{ }\rho +a\delta \right ) }{2\left ( \delta -m_{1y}\right ) ^{3}}+\frac {m_{2k}\mathrm{ }\rho }{2\left ( \delta -m_{1y}\right ) ^{2}}\right ] \left \{ t-\left ( \frac { 1-e^{-\left ( \delta -m_{1y}\right ) t}}{\delta -m_{1y}}\right ) \right \} \nonumber\\[15pt] +m_{1\Psi _{y,j_{2}}}\mathrm{ }\left \{ \frac {m_{1k}\mathrm{ }\rho +a\delta }{\left ( \delta -m_{1y}\right ) ^{2}}\right \} \left \{ t-\left ( \frac { 1-e^{-\left ( \delta -m_{1y}\right ) t}}{\delta -m_{1y}}\right ) \right \} \nonumber\\[15pt] -\left ( m_{1k}\mathrm{ }\rho +a\delta \right ) m_{1j_{2}}\left \{ \frac { m_{1k}\mathrm{ }\rho +a\delta }{\left ( \delta -m_{1y}\right ) ^{3}} \right \} \left \{ t-\left ( \frac {1-e^{-\left ( \delta -m_{1y}\right ) t}}{ \delta -m_{1y}}\right ) \right \} \end{array} \right ] \nonumber \\[10pt] & \quad + m_{2j_{2}}\left ( \frac {m_{1k}\mathrm{ }\rho +a\delta }{\delta -m_{1y}}\right ) t. \end{align}

Under the deviation-premium principle, we assume that $c_2t=\mathbb{E}(C_2(t))+\theta _2\sqrt {Var(C_2(t))}$ , where $\mathbb{E}(C_2(t))$ and $Var(C_2(t))$ are given by (A.9) and (A.26), respectively.

B. Proof of Theorem3.1

The section mainly follows the work of Ceci et al. (Reference Ceci, Colaneri and Cretarola2022). First, it is easy to find that if $\tilde {g}(t,\lambda )\in \mathcal{C}^1([0,T]\times (0,\infty ))$ solved (3.2), $\tilde {v}(t,x,\lambda )=\frac {-1}{\gamma }e^{-\gamma e^{-r(t-T)}x}\tilde {g}(t,\lambda )$ solves the HJB equation (2.19). Based on the Itô’s lemma, for any $0\leq t\leq T$ and $\nu \in \mathcal{V}$ , we have

(B.1) \begin{equation} \tilde {v}(T,X^{\nu }(T),\lambda (T))= \tilde {v}(t,X^{\nu }(t),\lambda (t))+\int _t^T\frac {\partial \tilde {v}(t,X^{\nu }(t),\lambda (t))}{\partial t}+\mathcal{K}\tilde {v}(s,X^{\nu }(s),\lambda (s))ds+M_T-M_t, \end{equation}

where, for $0\leq t \leq T$ ,

(B.2) \begin{align} M_t&=\int _0^t\int _0^{\infty }\left [\tilde {v}(s,X^{\nu }(s)-q_1j_1,\lambda (s))-\tilde {v}(s,X^{\nu }(s),\lambda (s))\right ]\left [dF_1(j_1)dN_1(s)-\eta dF_1(j_1)ds\right ]\nonumber\\[5pt] & \quad +\int _0^t\int _0^{\infty }\int _0^{\infty }\left [\tilde {v}(s,X^{\nu }(s)-q_2j_2,\lambda (s)+y)\right.\nonumber\\[5pt] & \quad \left. -\tilde {v}(s,X^{\nu }(s),\lambda (s))\right ]\left [d\Psi (y,j_2)dN_2(s)-\lambda (s) d\Psi (y,j_2)ds\right ]\nonumber\\[5pt] & \quad +\int _0^t\int _0^{\infty }\!\!\int _0^{\infty }\!\left [\tilde {v}(s,X^{\nu }(s)-\pi l,\lambda (s)+k)-\tilde {v}(s,X^{\nu }(s),\lambda (s))\right ]\left [d\Phi (k,l)dM(s)-\rho d\Phi (k,l)ds\right ]. \end{align}

Now, For any positive integer $m$ , we introduce a non-decreasing sequence of stopping times defined by

(B.3) \begin{equation} \tilde {\tau }_m=\inf \Big \{t\in [0,T]: |X^{\nu }(t)|\gt m \vee \lambda (t)\gt m \vee \lambda (t)\lt \frac {1}{m}\Big \}. \end{equation}

Following the assumption of a stationary $\lambda (t)$ , we know that as $m\rightarrow \infty$ , $\tilde {\tau }_m\rightarrow T$ . By the assumption that $\tilde {v}(t,x,\lambda )$ is continuous, it is bounded in any compact set. Now, we show that $M_{t\wedge \tilde {\tau }_m}$ , $0\leq t\leq T$ is a $\mathcal{F}$ -local martingale under the probability space $(\Omega ,\mathcal{F},\mathbb{P}).$ Denoting $R_m=[0,T]\times [-m,m]\times [\frac {1}{m},m]$ , we consider the the first jump term in (B.2)

(B.4) \begin{align} &\mathbb{E}\Bigg [\int _0^{T\wedge \tilde {\tau }_m}\int _0^{\infty }\left |\tilde {v}(s,X^{\nu }(s)-q_1j_1,\lambda (s))-\tilde {v}(s,X^{\nu }(s),\lambda (s))\right |\eta dF_1(j_1)ds\Bigg ]\nonumber\\[5pt] &\quad \leq \sup _{(t,x,\lambda )\in R_m} 2|\tilde {v}(t,x,\lambda )|\eta T\lt \infty . \end{align}

For the second jump term, we have

(B.5) \begin{align} &\mathbb{E}\left[\int _0^{T\wedge \tilde {\tau }_m}\int _0^{\infty }\int _0^{\infty }\left |\tilde {v}(s,X^{\nu }(s)-q_2j_2,\lambda (s)+y)-\tilde {v}(s,X^{\nu }(s),\lambda (s))\right |\lambda (s) d\Psi (y,j_2)ds\right]\nonumber\\ &\quad \leq \sup _{(t,x,\lambda )\in R_m} 2|\tilde {v}(t,x,\lambda )|m T\lt \infty . \end{align}

For the third jump term, we have

(B.6) \begin{align} &\mathbb{E}\Bigg [\int _0^{T\wedge \tilde {\tau }_m}\int _0^{\infty }\int _0^{\infty }\left [\tilde {v}(s,X^{\nu }(s)-\pi l,\lambda (s)+k)-\tilde {v}(s,X^{\nu }(s),\lambda (s))\right ]\rho d\Phi (k,l)ds\Bigg ]\nonumber\\ &\quad \sup _{(t,x,\lambda )\in R_m} 2|\tilde {v}(t,x,\lambda )|\rho T\lt \infty . \end{align}

Hence, we know that the stopping process $M_{t\wedge \tilde {\tau }_m}$ , $0\leq t\leq T$ , is a local martingale under the probability space $(\Omega ,\mathcal{F},\mathbb{P})$ . In addition, we have

(B.7) \begin{align} &\mathbb{E}\left[\tilde {v}\left (T\wedge \tilde {\tau }_m,X^{\nu }(T\wedge \tilde {\tau }_m),\lambda (T\wedge \tilde {\tau }_m)\right )^2\right]\nonumber\\ &\quad =\mathbb{E}\left [\frac {1}{\gamma ^2}e^{-2\gamma e^{-r(t-T)}X^{\nu }(T\wedge \tilde {\tau }_m)}\tilde {g}(T\wedge \tilde {\tau }_m,\lambda (T\wedge \tilde {\tau }_m))^2\right ]\lt \infty . \end{align}

Hence, $\{\tilde {v}\left (T\wedge \tilde {\tau }_m,X^{\nu }(T\wedge \tilde {\tau }_m),\lambda (T\wedge \tilde {\tau }_m)\right )\}, m\in \mathbb{N}$ is a family of uniformly integrable random variables and converges almost surely. Note that, based on (2.19), for any admissible strategy $\nu (t)\in \mathcal{V}$ and $t\in [0,T]$ , it holds that

(B.8) \begin{equation} \frac {\partial \tilde {v}(t,X^{\nu }(t),\lambda (t))}{\partial t}+\mathcal{K}\tilde {v}(t,X^{\nu }(t),\lambda (t))\leq 0, \quad \mathbb{P}-a.s. \end{equation}

Then, taking the conditional expectation on both sides of (B.1), replacing $T$ by $T\wedge \tilde {\tau }_m$ , and replacing $t$ by $t\wedge \tilde {\tau }_m$ , we can obtain that

(B.9) \begin{equation} \mathbb{E}_{t,x,\lambda }[\tilde {v}(T\wedge \tilde {\tau }_m,X^{\nu }(T\wedge \tilde {\tau }_m),\lambda (T\wedge \tilde {\tau }_m))]\leq \mathbb{E}_{t,x,\lambda }[\tilde {v}(t\wedge \tilde {\tau }_m,X^{\nu }(t\wedge \tilde {\tau }_m),\lambda (t\wedge \tilde {\tau }_m))]. \end{equation}

Considering $m\rightarrow \infty$ , the fact that $X^{\nu }(t)$ and $\lambda (t)$ have no deterministic jumps, and the fact that $\tilde {v}(t,x,\lambda )$ solves the HJB equation (2.19), we have, for any admissible strategy $\nu (t)\in \mathcal{V},$

(B.10) \begin{equation} \lim _{m\rightarrow \infty }\tilde {v}(T\wedge \tilde {\tau }_m,X^{\nu }(T\wedge \tilde {\tau }_m),\lambda (T\wedge \tilde {\tau }_m))\rightarrow \tilde {v}(T,X^{\nu }(T),\lambda (T))=U(X^{\nu }(T)), \quad \mathbb{P}-a.s. \end{equation}

and

(B.11) \begin{equation} \lim _{m\rightarrow \infty }\tilde {v}(t\wedge \tilde {\tau }_m,X^{\nu }(t\wedge \tilde {\tau }_m),\lambda (t\wedge \tilde {\tau }_m))\rightarrow \tilde {v}(t,X^{\nu }(t),\lambda (t)), \quad \mathbb{P}-a.s. \end{equation}

Furthermore, since $\{\tilde {v}\left (T\wedge \tilde {\tau }_m,X^{\nu }(T\wedge \tilde {\tau }_m),\lambda (T\wedge \tilde {\tau }_m)\right )\}, m\in \mathbb{N}$ is uniformly integrable, we know that for any admissible strategy $\nu (t)\in \mathcal{V}$ , (B.9) implies that

(B.12) \begin{equation} \mathbb{E}_{t,x,\lambda }[U(X^{\nu }(T))]\leq \tilde {v}(t,X^{\nu }(t),\lambda (t))=\tilde {v}(t,x,\lambda ), \end{equation}

which suggests that

(B.13) \begin{equation} V(t,x,\lambda )\leq \tilde {v}(t,x,\lambda ). \end{equation}

Finally, we need to show that $\tilde {v}(t,x,\lambda )$ is achievable. By the definition of the admissible strategy and the fact that $\mathcal{V}$ is compact, there exists a measurable function $\nu ^*(t)=(\pi ^*(t),q_1^*(t),q_2^*(t))$ which achieves the infimums of (3.4). Based on the condition in (3.4), the control $\nu ^*(t)$ is admissible. Furthermore, by computations similar to the above, we can easily check that the equality holds in (B.12) when $\nu ^*(t)$ is taken as the control, which implies that

(B.14) \begin{equation} V(t,x,\lambda )=\mathbb{E}_{t,x,\lambda }[U(X^{\nu ^*}(T))]=\tilde {v}(t,x,\lambda ). \end{equation}

Hence, we complete the proof of Theorem3.1.

C. Proof of Lemma 3.2

Note that (3.2) holds for any positive surplus $x\gt 0$ . Matching the coefficient of $x$ in (3.2) shows that

(C.1) \begin{equation} \frac {\partial A(t,T)}{\partial t}+rA(t,T)=0, \end{equation}

where $A(T,T)=-\gamma$ . Hence, we have $A(t,T)=-\gamma e^{-r(t-T)}.$

Applying the first-order condition to (3.4), the optimal reinsurance policy with respect to ordinary claims, denoted $\tilde {q}_1$ as follows:

(C.2) \begin{equation} \tilde {q}_1:=\arg \inf _{q_1} A(t,T)g\tilde {c}_1q_1+\eta g\int \limits _0^{\infty } e^{-A(t,T)q_1j_1}-1dF_1(j_1), \end{equation}

which implies

(C.3) \begin{equation} \begin{aligned} A(t,T)g\tilde {c}_1&=\eta g\int \limits _0^{\infty }A(t,T)j_1e^{-A(t,T)q_1j_1}dF_1(j_1),\\ \tilde {c}_1&=\eta \Pi _1(A(t,T)\tilde {q}_1),\\ \end{aligned} \end{equation}

where $\Pi _1(x)=\int \limits _0^{\infty } j_1e^{-xj_1}dF_1(j_1)$ denotes the Laplace transform of $j_1f_1(j_1)$ and $f_1(j_1)$ denotes the probability density function with respect to $F_1(j_1)$ .

Under the net-profit principle, we have $\tilde {c}_1=(1+\tilde {\theta }_1)\eta m_{j_1}.$ Hence, (C.3) implies that $\tilde {q}_1$ is the unique positive root of the following equation, where the positiveness and the uniqueness can be easily proved by contradictions.

(C.4) \begin{equation} (1+\tilde {\theta }_1)m_{j_1}=\int \limits _0^{\infty } j_1e^{-A(t,T)\tilde {q}_1j_1}dF_1(j_1). \end{equation}

Furthermore, suppose $J_1\sim Exp(\epsilon _1)$ , we have

(C.5) \begin{equation} \tilde {q}_1=\frac {\epsilon _1}{A(t,T)}\left (\frac {1}{\sqrt {1+\tilde {\theta }_1}}-1\right ). \end{equation}

Therefore,

(C.6) \begin{equation} q_1^*=\min \left (1,\frac {\epsilon _1}{A(t,T)}\left (\frac {1}{\sqrt {1+\tilde {\theta }_1}}-1\right )\right ). \end{equation}

Under the deviation-premium principle, we have $\tilde {c}_1=\eta m_{1j_1}+\tilde {\theta }_1\sqrt {\frac {\eta m_{2j_1}}{t}}$ . Hence, (C.3) implies that $\tilde {q}_1$ is the unique positive root of the following equation, where the positiveness and the uniqueness can be easily proved by contradictions.

(C.7) \begin{equation} m_{1j_1}+\tilde {\theta }_1\sqrt {\frac {m_{2j_1}}{\eta t}}=\int \limits _0^{\infty } j_1e^{-A(t,T)\tilde {q}_1j_1}dF_1(j_1). \end{equation}

Similarly, suppose $J_1\sim Exp(\epsilon _1)$ , we have

(C.8) \begin{equation} \tilde {q}_1=\frac {\epsilon _1}{A(t,T)}\left (\sqrt {\frac {1}{1+\tilde {\theta }_1\sqrt {\frac {2}{\eta t}}}}-1\right ). \end{equation}

Therefore,

(C.9) \begin{equation} q_1^*=\min \left (1,\frac {\epsilon _1}{A(t,T)}\left (\sqrt {\frac {1}{1+\tilde {\theta }_1\sqrt {\frac {2}{\eta t}}}}-1\right )\right ). \end{equation}

D. Technical Proofs

D.1 Proof of Lemma 4.1

First, we prove the unknown function $g(t,\lambda )$ is bounded. The lower bound is trivial suggested by the conjecture in (3.1). To prove the upper bound, we claim that $V(t,x,\lambda )$ is decreasing in $t$ . Then, given that $A(t,T)=-\gamma e^{-r(t-T)}$ , we have $g(t,\lambda )=-\gamma V(t,x,\lambda )e^{-A(t,T)x}\leq -\gamma V(T,x,\lambda )e^{-A(t,T)x}=e^{\gamma (e^{-r(t-T)}-1)x}$ . For any fixed surplus $x$ , we can easily find a positive constant $\beta$ such that $g(t,\lambda )\leq e^{\gamma (e^{-r(t-T)}-1)x}\leq e^{\gamma (e^{rT}-1)x}\lt \beta .$ Note that, the upper bound $\beta$ depends on the initial value $x$ . Now, we will prove that $V(t,x,\lambda )$ is decreasing in $t$ as follows.

D.1.1 Case 1

Without loss of generality, let $\tau _1$ denote the waiting time of the first claim received by the insurer from either ordinary claims or catastrophic claims. We consider any small enough $\delta _1\gt 0$ such that $\delta _1\lt \tau _1$ . Now, we consider the following admissible strategy

(D.1) \begin{equation} \tilde {\Delta }(s)=(\tilde {\pi }(s),\tilde {q}_1(s),\tilde {q}_2(s))= \begin{cases} &(0,1,1), \quad \mathrm{if } s\in [t,t+\delta _1],\\[5pt] &(\pi ^*(s),q_1^*(s),q_2^*(s)), \quad \mathrm{if } s\in (t+\delta _1,T], \end{cases} \end{equation}

which implies that before time $t+\delta _1$ , no investment in the risky stock and no reinsurance policies are applied by the insurer. After time $t+\delta _1$ , the insurer follows the optimal strategies. Therefore, for the corresponding surplus $X^{\tilde {\Delta }}(s)$ under the above strategy, we know

(D.2) \begin{equation} \begin{aligned} V(t,x,\lambda )&\geq \mathbb{E}_{t,x,\lambda }\big[U(X^{\tilde {\Delta }}(T))\big]\\[5pt] &=\mathbb{E}_{t,x,\lambda }\left [U\left (\int _t^{T}dX^{\tilde {\Delta }}(s)+x\right )\right ]\\[5pt] &=\mathbb{E}_{t,x,\lambda }\left [U\left (\int _t^{t+\delta _1}dX^{\tilde {\Delta }}(s)+\int _{t+\delta _1}^{T}dX^{\tilde {\Delta }}(s)+x\right )\right ]. \end{aligned} \end{equation}

Given that $\delta _1\lt \tau _1$ , $\tilde {\Delta }(s)=(0,1,1)$ for $s\in [t,t+\delta _1]$ , and the surplus process (2.10), we have

(D.3) \begin{equation} \begin{aligned} \int _t^{t+\delta _1}dX^{\tilde {\Delta }}(s)&=\int _t^{t+\delta _1}\big[rX^{\tilde {\Delta }}(s)+c_1+c_2\big]ds\gt 0,\\ \end{aligned} \end{equation}

which implies that

(D.4) \begin{equation} X^{\tilde {\Delta }}(t+\delta _1)\gt x. \end{equation}

Since $U(x)=\frac {-e^{-\gamma x}}{\gamma }$ is increasing for $x\gt 0$ . Given (D.1) and (D.4), (D.2) can be rewritten as follows:

(D.5) \begin{align} V(t,x,\lambda )&\geq \mathbb{E}_{t,x,\lambda }\left [U\left (\int _{t+\delta _1}^{T}dX^{\tilde {\Delta }}(s)+X^{\tilde {\Delta }}(t+\delta _1)\right )\right ]\nonumber\\[5pt] &=\mathbb{E}_{{t+\delta _1},x,\lambda }\left [U\left (\int _{t+\delta _1}^{T}dX^{\tilde {\Delta }}(s)+X^{\tilde {\Delta }}(t+\delta _1)\right )\right ]\nonumber\\[5pt] &\gt \mathbb{E}_{{t+\delta _1},x,\lambda }\left [U\left (\int _{t+\delta _1}^{T}dX^{\tilde {\Delta }}(s)+x\right )\right ]\nonumber\\ &=V(t+\delta _1,x,\lambda ), \end{align}

which implies that $V(s,x,\lambda )$ is decreasing for $s\in [t,t+\tau _1)$ .

D.1.2 Case 2

Now, let $\tau _2$ denote the waiting time of the second claim received by the insurer from either ordinary claims or catastrophic claims. Without loss of generality, we just assume that the first claim comes from the catastrophic insurance market with the waiting time $\tau _1\gt 0$ . We consider the value function $V(u^-,x,\lambda )$ , where $u^-=t+\tau _1^-$ , which implies that the first claim happens at time $u$ with a known magnitude $j\gt 0$ . Furthermore, we consider any positive constant $s\gt 0$ such that $s\leq \min \left (\tau _2,\frac {j}{\tilde {c_2}}\right )$ and based on Assumption 2.7, we can introduce a positive constant $\frac {(c_2-\tilde {c_2}+c_1)s}{j-\tilde {c_2}s}\gt 0$ . Now, we consider the following admissible strategy

(D.6) \begin{equation} \widehat {\Delta }(v)=(\widehat {\pi }(v),\widehat {q}_1(v),\widehat {q}_2(v))= \begin{cases} &\left (0,1,\widehat {q}\right ), \quad \mathrm{if } v\in [t+\tau _1^-,t+\tau _1+s],\\[5pt] &(\pi ^*(v),q_1^*(v),q_2^*(v)), \quad \mathrm{if } v\in (t+\tau _1+s,T], \end{cases} \end{equation}

where $0\leq \widehat {q}\leq \frac {(c_2-\tilde {c_2}+c_1)s}{j-\tilde {c_2}s}$ . Note that, (D.6) implies that for $v\in [t+\tau _1^-,t+\tau _1+s)$ , no investment in the risky stock and no reinsurance policy is applied for the ordinary business line, and a proportional reinsurance with parameter $\widehat {q}$ is applied for the catastrophic business line. After time $t+\tau _1+s$ , the insurer follows the optimal strategies. Therefore, for the corresponding surplus $X^{\widehat {\Delta }}(v)$ under the above strategy, we know

(D.7) \begin{align} V(u^-,x,\lambda )&\geq \mathbb{E}_{u^-,x,\lambda }\big[U(X^{\widehat {\Delta }}(T))\big]\nonumber\\[5pt] &=\mathbb{E}_{u^-,x,\lambda }\left [U\left (\int _{t+\tau _1^-}^{T}dX^{\tilde {\Delta }}(v)+x-\widehat {q}j\right )\right ]\nonumber\\[5pt] &=\mathbb{E}_{u^-,x,\lambda }\left [U\left (\int _{t+\tau _1^-}^{t+\tau _1+s}dX^{\widehat {\Delta }}(v)+\int _{t+\tau _1+s}^{T}dX^{\tilde {\Delta }}(v)+x-\widehat {q}j\right )\right ]. \end{align}

Given that $s\leq \min \left (\tau _2,\frac {j}{\tilde {c_2}}\right )$ , $\widehat {\Delta }(v)=\left (0,1,\widehat {q}\right )$ for $v\in [t+\tau _1,t+\tau _1+s]$ , and the surplus process (2.10), we have

(D.8) \begin{equation} \begin{aligned} \int _{t+\tau _1^-}^{t+\tau _1+s}dX^{\widehat {\Delta }}(v)+x-\widehat {q}j&=\int _{t+\tau _1^-}^{t+\tau _1+s}\left [rX^{\widehat {\Delta }}(v)+c_1+c_2-(1-\widehat {q})\tilde {c}_2\right ]dv+x-\widehat {q}j\gt 0. \end{aligned} \end{equation}

Since $0\leq \widehat {q}\leq \frac {(c_2-\tilde {c_2}+c_1)s}{j-\tilde {c_2}s}$ , we know that

(D.9) \begin{equation} X^{\widehat {\Delta }}(t+\tau _1+s)=X^{\widehat {\Delta }}(u+s)=\int _{t+\tau _1^-}^{t+\tau _1+s}dX^{\widehat {\Delta }}(v)+x-\widehat {q}j\gt x. \end{equation}

Since $U(x)=\frac {-e^{-\gamma x}}{\gamma }$ is increasing for $x\gt 0$ . Given (D.6) and (D.9), (D.7) can be rewritten as follows:

(D.10) \begin{align} V(u^-,x,\lambda )&\geq \mathbb{E}_{u^-,x,\lambda }\left [U\left (\int _{u+s}^{T}dX^{\widehat {\Delta }}(v)+X^{\widehat {\Delta }}(u+s)\right )\right ]\nonumber\\[3pt] &=\mathbb{E}_{{u+s},x,\lambda }\left [U\left (\int _{u+s}^{T}dX^{\widehat {\Delta }}(v)+X^{\widehat {\Delta }}(u+s)\right )\right ]\nonumber\\[3pt] &\gt \mathbb{E}_{{u+s},x,\lambda }\left [U\left (\int _{u+s}^{T}dX^{\widehat {\Delta }}(v)+x\right )\right ]\nonumber\\[3pt] &=V(u+s,x,\lambda ), \end{align}

which implies that $V(v,x,\lambda )$ is decreasing for $v\in [t+\tau _1^-,t+\tau _1+s)$ .

Now, we can combine our findings in Sections D.1.1 and D.1.2 to illustrate that $V(v,x,\lambda )$ is indeed decreasing in $v\in [t,T]$ . Hence, we have $0\lt (\mathcal{L}g)(t,\lambda )=g(t,\lambda )\lt \beta$ . Finally, we have $V(t,0,\lambda )=\frac {-1}{\gamma }g(t,\lambda )$ , which implies that $g(t,\lambda )$ is increasing in $t$ .

D.2 Proof of Lemma 4.2

Since $g(t,\lambda )$ is bounded and positive in $t\in [0, T]$ proved in Lemma 4.1, we can easily prove that $g(t,\lambda )$ has bounded first-order partial derivatives with respect to $t$ and $\lambda$ . Hence, $g(t,\lambda )$ is Lipschitz continuous.

Since $g(t,\lambda )$ is Lipschitz continuous, bounded, and positive for $t\in [0,T]$ and $\lambda \in (0,\infty )$ , we know that $g(t,\lambda )$ is bounded away from $0$ , $i.e.$ , there exists some positive constant $l$ such that $0\lt l\leq g(t,\lambda )\lt \beta$ , where $\beta$ is given in Lemma 4.1.

Based on (4.3) and the fact that $g(t,\lambda )$ is bounded for $t\in [0,T]$ and $\lambda \in (0,\infty )$ , $\tilde {\psi }(t,T)$ and $\tilde {\psi }_1(t,T)$ are also bounded in the same interval. Furthermore, since $\tilde {\psi }(t,T)$ and $\tilde {\psi }_1(t,T)$ are $\mathcal{C}^1$ functions and $g(t,\lambda )$ is positive and bounded away from $0$ , it holds that $\tilde {\psi }(t,T)$ and $\tilde {\psi }_1(t,T)$ have bounded first-order derivatives with respect to $t$ for $t\in [0, T]$ . Hence, $\tilde {\psi }(t,T)$ and $\tilde {\psi }_1(t,T)$ are also Lipschitz continuous for $t\in [0, T]$ .

D.3 Proof of Lemma 4.3

Consider an arbitrary Cauchy sequence $\{\zeta _k(t,\lambda )\}_{k=1,2,3,\ldots }$ in the space $\mathfrak{C}_e([0,T]\times (0,\infty ))$ with the metric (4.4), we know that for any $\epsilon \gt 0$ , there exists a positive integer $N$ such that for all $m,n\geq N$ , $d(\zeta _m(t,\lambda ),\zeta _n(t,\lambda ))\lt \epsilon .$

Since $\{\zeta _k(t,\lambda )\}_{k=1,2,3,\ldots }$ is Cauchy in the supremum norm (4.4), it is also pointwise Cauchy and

(D.11) \begin{equation} \zeta _k(t,\lambda )=e^{\psi _k(t,T)+\psi _{1,k}(t,T)\lambda }, \end{equation}

where $\psi _k(t,T)$ and $\psi _{1,k}(t,T)$ denote some bounded and Lipschitz continuous functions.

For any $t\in [0,T]$ and $\lambda \in (0,\infty )$ , we know that $\{\zeta _k(t,\lambda )\}_{k=1,2,3,\ldots }$ is also a Cauchy sequence in $\mathbb{R}$ . Since $\mathbb{R}$ is complete, $\{\zeta _k(t,\lambda )\}_{k=1,2,3,\ldots }$ converges to some limit, denoted by $\bar {\zeta }(t,\lambda )$ . We have

(D.12) \begin{equation} \bar {\zeta }(t,\lambda )=\lim _{k\rightarrow \infty }\zeta _k(t,\lambda )=\lim _{k\rightarrow \infty }e^{\psi _k(t,T)+\psi _{1,k}(t,T)\lambda }=e^{\bar {\psi }(t,T)+\bar {\psi }_1(t,T)\lambda }. \end{equation}

where $\bar {\psi }(t,T)=\lim _{k\rightarrow \infty }\psi _k(t,T)$ and $\bar {\psi }_1(t,T)=\lim _{k\rightarrow \infty }\psi _{1,k}(t,T)$ .

To show that $\bar {\zeta }(t,\lambda )\in \mathfrak{C}_e([0,T]\times (0,\infty ))$ , we need to prove that $\bar {\zeta }(t,\lambda )$ is continuous and $0\lt \bar {\zeta }(t,\lambda )\lt \beta$ . In addition, we need to prove that $\{\zeta _k(t,\lambda )\}_{k=1,2,3,\ldots }$ converges uniformly to $\bar {\zeta }(t,\lambda )$ under the supremum norm (4.4) to show that the metric space $\mathfrak{C}_e([0,T]\times (0,\infty ))$ is complete.

Note that $0\lt \bar {\zeta }(t,\lambda )\lt \beta$ can be easily obtained through contradiction, and we omit it here.

D.3.1 Continuity of $\bar {\zeta }(t,\lambda )$

Let $\epsilon \gt 0$ . Since $\{\zeta _k(t,\lambda )\}_{k=1,2,3,\ldots }$ is Cauchy in the supremum norm (4.4), there exists a positive integer $N$ such that for all $m,n\geq N$ ,

(D.13) \begin{equation} \sup _{(t,\lambda )\in [0,T]\times (0,\infty )}\left |e^{-\alpha (T-t)}\left (\zeta _m(t,\lambda )-\zeta _n(t,\lambda )\right )\right |\lt \frac {\epsilon }{3}. \end{equation}

Now, fix $n\geq N$ . For any $(t_1,\lambda _1), (t_2,\lambda _2)\in [0,T]\times (0,\infty )$ ,

(D.14) \begin{equation} |\bar {\zeta }(t_1,\lambda _1)-\bar {\zeta }(t_2,\lambda _2)|\leq |\bar {\zeta }(t_1,\lambda _1)-\zeta _n(t_1,\lambda _1)|+|\zeta _n(t_1,\lambda _1)-\zeta _n(t_2,\lambda _2)|+|\zeta _n(t_1,\lambda _1)-\bar {\zeta }(t_2,\lambda _2)|. \end{equation}

Since $\zeta _n(t,\lambda )$ is continuous in $[0,T]\times (0,\infty )$ , for sufficiently close $(t_1,\lambda _1)$ and $(t_2,\lambda _2)$ , $|\zeta _n(t_1,\lambda _1)-\zeta _n(t_2,\lambda _2)|\lt \frac {\epsilon }{3}$ . In addition, for sufficiently large $n$ , $|\bar {\zeta }(t_1,\lambda _1)-\zeta _n(t_1,\lambda _1)|\lt \frac {\epsilon }{3}$ and $|\zeta _n(t_1,\lambda _1)-\bar {\zeta }(t_2,\lambda _2)|\lt \frac {\epsilon }{3}$ . Therefore, (D.14) is equivalent to

(D.15) \begin{equation} |\bar {\zeta }(t_1,\lambda _1)-\bar {\zeta }(t_2,\lambda _2)|\lt \epsilon , \end{equation}

which implies that $\bar {\zeta }(t,\lambda )$ is continuous in $[0,T]\times (0,\infty )$ .

D.3.2 Uniform convergence under the supremum norm (4.4)

In this subsection, we want to show that $\{\zeta _k(t,\lambda )\}_{k=1,2,3,\ldots }$ converges uniformly to $\bar {\zeta }(t,\lambda )$ under the supremum norm (4.4). Given $\epsilon \gt 0$ , there exists $N$ such that for all $m,n\geq N$ , $d(\zeta _m(t,\lambda ),\zeta _n(t,\lambda ))\lt \epsilon$ . Now, fix $n\geq N$ . For any $(t,\lambda )\in [0,T]\times (0,\infty )$ ,

(D.16) \begin{equation} |e^{-\alpha (T-t)}(\zeta _n(t,\lambda )-\bar {\zeta }(t,\lambda ))|\leq |e^{-\alpha (T-t)}\zeta _n(t,\lambda )-\zeta _m(t,\lambda )|+|e^{-\alpha (T-t)}\zeta _m(t,\lambda )-\bar {\zeta }(t,\lambda )|. \end{equation}

Taking $m\rightarrow \infty$ , we have $|e^{-\alpha (T-t)}(\zeta _n(t,\lambda )-\bar {\zeta }(t,\lambda ))|\lt \epsilon$ . Hence,

(D.17) \begin{equation} d(\zeta _n(t,\lambda ),\bar {\zeta }(t,\lambda ))=\sup _{(t,\lambda )\in [0,T]\times (0,\infty )}\left |e^{-\alpha (T-t)}(\zeta _n(t,\lambda )-\bar {\zeta }(t,\lambda ))\right |\lt \epsilon . \end{equation}

Therefore, $\{\zeta _k(t,\lambda )\}_{k=1,2,3,\ldots }$ converges uniformly to $\bar {\zeta }(t,\lambda )$ under the supremum norm (4.4).

In summary, we have shown that any Cauchy sequence of functions in $\mathfrak{C}_e([0,T]\times (0,\infty ))$ converges to a limit that is also in $\mathfrak{C}_e([0,T]\times (0,\infty ))$ under the supremum norm (4.4). Hence, the space $\mathfrak{C}_e([0,T]\times (0,\infty ))$ with metric (4.4) is complete.

D.4 Proof of Lemma 4.4

In this proof, we omit the parameters $s$ and $\lambda$ to simplify notations.

D.4.1 $\chi _1\big(\pi _{v_1}^*,v_1\big)-\chi _1(\pi _{v_2}^*,v_2)$

Applying the first-order condition to (3.4), optimal allocation policies $\pi _{v_i}^*$ , $i=1,2,$ follow

(D.18) \begin{equation} (\mu -r)v_i+\sigma ^2\pi _{v_i}^*Av_i=\rho \int \limits _0^{\infty }\int \limits _0^{\infty }le^{-\pi _{v_i}^*Al}v_i(s,\lambda +k)d\Phi (k,l). \end{equation}

Therefore, $\chi _1(\pi _{v_i}^*,v_i)$ , $i=1,2,$ can be rewritten as

(D.19) \begin{equation} \chi _1\big(\pi _{v_i}^*,v_i\big)=0.5(\mu -r)\pi _{v_i}^*Av_i+\rho \int \limits _0^{\infty }\int \limits _0^{\infty }\left ((0.5\pi _{v_i}^*Al+1)e^{-\pi ^*_{v_i}Al}-1\right )v_i(t,\lambda +k)d\Phi (k,l). \end{equation}

Based on the Mean Value Theorem, $\chi _1(\pi _{v_1}^*,v_1)-\chi _1(\pi _{v_2}^*,v_2)$ can be expressed as

(D.20) \begin{equation} \chi _1\big(\pi _{v_1}^*,v_1\big)-\chi _1(\pi _{v_2}^*,v_2)=(v_1-v_2)\frac {d Q_1\big(\pi _{\zeta }^*,\zeta \big)}{d\zeta }|_{\zeta =\xi _1}, \end{equation}

where $\xi _1\in \mathfrak{C}_e([0,T]\times (0,\infty ))$ and $\xi _1(s,\lambda )$ lies between $v_1(s,\lambda )$ and $v_2(s,\lambda )$ . $Q_1(\pi _{\zeta }^*,\zeta )$ and $\pi _{\zeta }^*$ satisfy

(D.21) \begin{equation} Q_1\big(\pi _{\zeta }^*,\zeta \big)=-0.5\sigma ^2\pi _{\zeta }^{*2}A^2\zeta +\rho \int \limits _0^{\infty }\int \limits _0^{\infty }\left (\big(\pi _{\zeta }^*Al+1\big)e^{-\pi ^*_{\zeta }Al}-1\right )\zeta (t,\lambda +k)d\Phi (k,l), \end{equation}
(D.22) \begin{equation} (\mu -r)\zeta +\sigma ^2\pi _{\zeta }^*A\zeta =\rho \int \limits _0^{\infty }\int \limits _0^{\infty }le^{-\pi _{\zeta }^*Al}\zeta (s,\lambda +k)d\Phi (k,l). \end{equation}

Now, differentiating (D.22) with respect to $\zeta$ provides

(D.23) \begin{align} (\mu -r) +\sigma ^2\pi _{\zeta }^*A+\sigma ^2A\zeta \frac {d\pi _{\zeta }^*}{d\zeta } & = \rho \int \limits _0^{\infty }\int \limits _0^{\infty }-Al^2e^{-\pi _{\zeta }^*Al}\zeta (s,\lambda +k)\frac {d\pi _{\zeta }^*}{d\zeta }\nonumber\\& \quad +le^{-\pi _{\zeta }^*Al}\frac {d\zeta (s,\lambda +k)}{d\zeta }d\Phi (k,l). \end{align}

Since $\zeta (t,\lambda )\in \mathfrak{C}_e([0,T]\times (0,\infty ))$ , where $\zeta (t,\lambda )=e^{\tilde {\psi }(t,T)+\tilde {\psi }_1(t,T)\lambda }$ for some unknown bounded functions $\tilde {\psi }(t,T)$ and $\tilde {\psi }_1(t,T)$ , (D.23) is equivalent to

(D.24) \begin{align} (\mu -r) & +\sigma ^2\pi _{\zeta }^*A+\sigma ^2A\zeta \frac {d\pi _{\zeta }^*}{d\zeta }\nonumber\\& \quad =\rho \int \limits _0^{\infty }\int \limits _0^{\infty }le^{-\pi _{\zeta }^*Al}e^{\tilde {\psi }_1(t,T)k}\left (-Al\zeta \frac {d\pi _{\zeta }^*}{d\zeta }+1+\zeta k\frac {d\tilde {\psi }_1(t,T)}{d\zeta }\right )d\Phi (k,l). \end{align}

Since $\zeta (t,\lambda )=e^{\tilde {\psi }(t,T)+\tilde {\psi }_1(t,T)\lambda }$ , we have $\frac {d\tilde {\psi }_1(t,T)}{d\zeta }=\frac {1}{\zeta \left (\lambda +\frac {d\tilde {\psi }(t,T)}{d\tilde {\psi }_1(t,T)}\right )}$ . Hence, D.24 can be rewritten as

(D.25) \begin{equation} \zeta \frac {d\pi _{\zeta }^*}{d\zeta }=\frac {\rho \int \limits _0^{\infty }\int \limits _0^{\infty }le^{-\pi _{\zeta }^*Al}e^{\tilde {\psi }_1(t,T)k}\left (1+k\frac {d\tilde {\psi }_1(t,T)}{d\zeta }\right )d\Phi (k,l)-(\mu -r)-\sigma ^2\pi _{\zeta }^*A}{\rho \int \limits _0^{\infty }\int \limits _0^{\infty }Al^2e^{-\pi _{\zeta }^*Al}e^{\tilde {\psi }_1(t,T)k}d\Phi (k,l)+\sigma ^2A}. \end{equation}

Now, we consider the derivative of $Q_1(q_{\zeta }^*,\zeta )$ with respect to $\zeta$ as follows:

(D.26) \begin{align} \frac {d Q_1\big(\pi _{\zeta }^*,\zeta \big)}{d\zeta }&=-0.5\sigma ^2\pi _{\zeta }^{*2}A^2\nonumber\\& \quad +\rho \int \limits _0^{\infty }\int \limits _0^{\infty }\left ((\pi _{\zeta }^*Al+1)e^{-\pi ^*_{\zeta }Al}-1\right )\left (1+k\frac {d\tilde {\psi }_1(t,T)}{d\zeta }\right )e^{\tilde {\psi }_1(t,T)k}d\Phi (k,l)\nonumber\\ &\quad +\zeta \frac {d\pi _{\zeta }^*}{d\zeta }\left [-\sigma ^2\pi _{\zeta }^{*}A^2-\rho \int \limits _0^{\infty }\int \limits _0^{\infty }\pi _{\zeta }^{*}A^2l^2e^{-\pi ^*_{\zeta }Al}e^{\tilde {\psi }_1(t,T)k}d\Phi (k,l)\right ]. \end{align}

Now, we can substitute (D.25) into (D.26), which implies

\begin{align*} \frac {d Q_1\big(\pi _{\zeta }^*,\zeta \big)}{d\zeta }&=(\mu -r)\pi _{\zeta }^{*}A+0.5\sigma ^2\pi _{\zeta }^{*2}A^2\nonumber\\& \quad +\rho \int \limits _0^{\infty }\int \limits _0^{\infty }\left (1+k\frac {d\tilde {\psi }_1(t,T)}{d\zeta }\right )\big(e^{-\pi ^*_{\zeta }Al}-1\big)e^{\tilde {\psi }_1(t,T)k}d\Phi (k,l)\end{align*}

(D.27) \begin{align} &=(\mu -r)\pi _{\zeta }^{*}A+0.5\sigma ^2\pi _{\zeta }^{*2}A^2+\rho \int \limits _0^{\infty }\int \limits _0^{\infty }e^{-\pi ^*_{\zeta }Al+\tilde {\psi }_1(t,T)k}d\Phi (k,l)\nonumber\\ & \quad -\rho \int \limits _0^{\infty }\int \limits _0^{\infty }e^{\tilde {\psi }_1(t,T)k}d\Phi (k,l)\nonumber\\ &\quad +\frac {d\tilde {\psi }_1(t,T)}{d\zeta }\rho \int \limits _0^{\infty }\int \limits _0^{\infty }ke^{-\pi ^*_{\zeta }Al+\tilde {\psi }_1(t,T)k}d\Phi (k,l)-\frac {d\tilde {\psi }_1(t,T)}{d\zeta }\rho \int \limits _0^{\infty }\int \limits _0^{\infty }ke^{\tilde {\psi }_1(t,T)k}d\Phi (k,l). \end{align}

Under the assumption that $\zeta (t,\lambda )\in \mathfrak{C}_e([0,T]\times (0,\infty ))$ is increasing in $t$ and has non-trivial bounded and Lipschits continuous functions $\tilde {\psi }(t,T)$ and $\tilde {\psi }_1(t,T)$ , we know that

(D.28) \begin{equation} \left |\frac {d\tilde {\psi }_1(t,T)}{d\zeta }\right |=\frac {1}{\left |\zeta (t,\lambda )\left (\lambda +\frac {\tilde {\psi }^{'}(t,T)}{\tilde {\psi }_1^{'}(t,T)}\right )\right |}=\left |\frac {\tilde {\psi }_1^{'}(t,T)}{\zeta (t,\lambda )\left (\tilde {\psi }^{'}(t,T)+\lambda \tilde {\psi }_1^{'}(t,T)\right )}\right |, \end{equation}

where $\tilde {\psi }^{'}(t,T)=\frac {d\tilde {\psi }(t,T)}{dt}$ and $\tilde {\psi }_1^{'}(t,T)=\frac {d\tilde {\psi }_1(t,T)}{dt}$ . Hence, $\left |\frac {d\tilde {\psi }_1(t,T)}{d\zeta }\right |$ is bounded. Furthermore, since the joint distribution $\Phi (k,l)$ as well as marginal distributions $H(k)$ and $D(l)$ follow Assumptions 2.3 and 2.4, we know that $\left |\frac {d Q_1\big(\pi _{\zeta }^*,\zeta \big)}{d\zeta }\right |$ is bounded by some positive constant $\Delta _1$ . Hence, we have

(D.29) \begin{equation} \left |\chi _1\big(\pi _{v_1}^*,v_1\big)-\chi _1(\pi _{v_2}^*,v_2)\right |\leq \Delta _1|v_1-v_2|. \end{equation}
D.4.2 $\chi _2(q_{1v_1}^*,v_1)-\chi _2(q_{1v_2}^*,v_2)$

Second, we consider $\chi _2(q_{1v_1}^*,v_1)-\chi _2(q_{1v_2}^*,v_2)$ , which can be rewritten as follows based on (3.11)

(D.30) \begin{align} \chi _2\left(q_{1v_1}^*,v_1\right)-\chi _2\left(q_{1v_2}^*,v_2\right) & = \left(q^*_{1u_1}u_1-q^*_{1v_2}v_2\right)\tilde {c}_1A\nonumber\\[5pt]& \quad +\eta \left [\left(\Pi _2\left(q_{1v_1}^*A\right)-1\right)v_1-\left(\Pi _2\left(q_{1v_2}^*A\right)-1\right)v_2\right ], \end{align}

where $\Pi _2(x)=\int \limits _0^{\infty } e^{-xj_1}dF_1(j_1)$ denotes the Laplace transform of the cumulative distribution function $F_1(j_1)$ . Based on Lemma 3.2, we know that $q_{1v_i}^*$ is independent of $v_i$ , for $i=1,2.$ Hence, the optimal reinsurance policies of ordinary claims with with respect to $v_1$ and $v_2$ are the same (denoted by $q_1^*$ ) and (D.30) can be rewritten as follows:

(D.31) \begin{equation} \begin{aligned} \chi _2(q_{1v_1}^*,v_1)-\chi _2(q_{1v_2}^*,v_2)=\left [q_1^*\tilde {c}_1A+\eta (\Pi _2(q_1^*A)-1)\right ](v_1-v_2). \end{aligned} \end{equation}

Under Assumption 2.3 that the distribution $F_1(j_1)$ has a finite Laplace transform and the fact that $0\leq q_1^*\leq 1$ , we know that $\left |\chi _2(q_{1v_1}^*,v_1)-\chi _2(q_{1v_2}^*,v_2)\right |$ is bounded by $\Delta _2|v_1-v_2|$ , where $\Delta _2$ denotes some positive constant.

D.4.3 $\chi _3(q_{2v_1}^*,v_1)-\chi _3(q_{2v_2}^*,v_2)$

Based on (3.4), denote $\tilde {q}_{2v_i}$ , $i=1,2,$ as follows:

(D.32) \begin{equation} \tilde {q}_{2v_i}:=\arg \inf _qq\tilde {c}_2Av_i+\lambda \int \limits _0^{\infty }\int \limits _0^{\infty }(e^{-Aqj_2}-1)v_i(t,\lambda +y)d\Psi (y,j_2), \end{equation}

which implies

(D.33) \begin{align} \tilde {c}_2Av_i = \lambda \int \limits _0^{\infty }\int \limits _0^{\infty }Aj_2e^{-Aqj_2}v_i(t,\lambda +y)d\Psi (y,j_2). \end{align}

Since $v_i(t,\lambda +y)=v_i(t,\lambda )e^{C_i(t,T)y}$ for some unknown bounded function $C_i(t,T)$ , we have

(D.34) \begin{equation} \tilde {c}_2=\lambda \int \limits _0^{\infty }\int \limits _0^{\infty }j_2e^{-A\tilde {q}_{2v_i}j_2+C_i(t,T)y}d\Psi (y,j_2). \end{equation}

Then, the optimal reinsurance policies of catastrophic claims with respect to $v_1$ and $v_2$ are as follows:

(D.35) \begin{align} q_{2v_1}^*&=\min (1,\tilde {q}_{2v_1}),\nonumber\\[4pt] q_{2v_2}^*&=\min (1,\tilde {q}_{2v_2}), \end{align}

Furthermore, $\chi _3(q_{2v_i}^*,v_i)$ , $i=1,2,$ can be expressed as follows:

(D.36) \begin{equation} \begin{aligned} \chi _3(q_{2v_i}^*,v_i)&=q_{2v_i}^*\tilde {c}_2Av_i+\lambda \int \limits _0^{\infty }\int \limits _0^{\infty }\big(e^{-Aq_{2v_i}^*j_2}-1\big)v_i(t,\lambda +y)d\Psi (y,j_2). \end{aligned} \end{equation}

Case 1: $\tilde {q}_{2v_1}\gt 1$ and $\tilde {q}_{2v_2}\gt 1$

Suppose that $\tilde {q}_{2v_1}\gt 1$ and $\tilde {q}_{2v_2}\gt 1$ , we know that $q_{2v_1}^*=1$ and $q_{2v_2}^*=1$ . Then, $\chi _3(q_{2v_1}^*,v_1)-\chi _3(q_{2v_2}^*,v_2)$ becomes very trivial as follows

(D.37) \begin{align} \chi _3(q_{2v_1}^*,v_1)-\chi _3(q_{2v_2}^*,v_2)& =(v_1-v_2)\tilde {c}_2A\nonumber\\& \quad +\lambda \int \limits _0^{\infty }\int \limits _0^{\infty }(e^{-Aj_2}-1)(v_1(t,\lambda +y)-v_2(t,\lambda +y))d\Psi (y,j_2), \end{align}

which implies

(D.38) \begin{equation} \begin{aligned} \left |\chi _3(q_{2v_1}^*,v_1)-\chi _3(q_{2v_2}^*,v_2)\right |&\leq |\tilde {c}_2A||v_1(t,\lambda )-v_2(t,\lambda )|+\lambda \int \limits _0^{\infty }\int \limits _0^{\infty }(e^{-Aj_2}-1)d\Psi (y,j_2)D_{\lambda }(v_1,v_2)\\ &=\left (|\tilde {c}_2A|+\lambda \left (\int \limits _0^{\infty }e^{-Aj_2}dF_2(j_2)-1\right )\right )D_{\lambda }(v_1,v_2) \end{aligned} \end{equation}

where $D_{\lambda }(v_1,v_2)=\sup _{\lambda }|v_1(t,\lambda )-v_2(t,\lambda )|$ . Since $F_2(j_2)$ has a finite Laplace transform under Assumption 2.3, we know

(D.39) \begin{equation} \big|\chi _3\big(q_{2v_1}^*,v_1\big)-\chi _3\big(q_{2v_2}^*,v_2\big)\big |\leq \Delta _3D_{\lambda }(v_1,v_2), \end{equation}

where $\Delta _3$ denotes some positive constant.

Case 2: $\tilde {q}_{2v_1}\gt 1$ and $\tilde {q}_{2v_2}\lt 1$

In this subsection, we consider the case where one optimal reinsurance equals $1$ and the other one is given by (D.33). Without loss of generality, we suppose that $\tilde {q}_{2v_1}\gt 1$ and $\tilde {q}_{2v_2}\lt 1$ . Hence, $q_{2v_1}^*=1$ and $q_{2v_2}^*=\tilde {q}_{2v_2}$ . $\chi _3(q_{2v_1}^*,v_1)-\chi _3(q_{2v_2}^*,v_2)$ can be rewritten as

(D.40) \begin{equation} \begin{aligned} \chi _3(q_{2v_1}^*,v_1)-\chi _3(q_{2v_2}^*,v_2)&=\tilde {c}_2Av_1+\lambda \int \limits _0^{\infty }\int \limits _0^{\infty }(e^{-Aj_2}-1)v_1(t,\lambda +y)d\Psi (y,j_2)\\ &-\tilde {q}_{2v_2}\tilde {c}_2Av_2-\lambda \int \limits _0^{\infty }\int \limits _0^{\infty }(e^{-A\tilde {q}_{2v_2}j_2}-1)v_2(t,\lambda +y)d\Psi (y,j_2). \end{aligned} \end{equation}

Furthermore, (D.40) can be rewritten as

(D.41) \begin{align} \chi _3(q_{2v_1}^*,v_1)-\chi _3(q_{2v_2}^*,v_2)&=\lambda \int \limits _0^{\infty }\int \limits _0^{\infty }(e^{-Aj_2}-1)(v_1(t,\lambda +y)-v_2(t,\lambda +y))d\Psi (y,j_2)\nonumber\\ & \quad +\tilde {c}_2Av_1-\tilde {q}_{2v_2}\tilde {c}_2Av_2+\lambda \!\int \limits _0^{\infty }\int \limits _0^{\infty }\!(e^{-Aj_2}-e^{-A\tilde {q}_{2v_2}j_2})v_2(t,\lambda +y)d\Psi (y,j_2). \end{align}

Now, based on (D.33), we claim that

(D.42) \begin{equation} -\tilde {q}_{2v_2}\tilde {c}_2Av_2+\lambda \int \limits _0^{\infty }\int \limits _0^{\infty }(e^{-Aj_2}-e^{-A\tilde {q}_{2v_2}j_2})v_2(t,\lambda +y)d\Psi (y,j_2)\leq -A\tilde {c}_2v_2, \end{equation}

where the proof is as follows. Define a function $\mathcal{P}(q)$ for $q\in [0,1]$ , where $A\lt 0$ , $j\gt 0$ , and

(D.43) \begin{equation} \begin{aligned} \mathcal{P}(q)&=e^{-Aj}-e^{-Aqj}+A(1-q)je^{-Aqj}. \end{aligned} \end{equation}

Note that $\mathcal{P}(1)=0$ and

(D.44) \begin{equation} \frac {d\mathcal{P}(q)}{dq}=-A^2(1-q)j^2e^{-Aqj}\leq 0, \forall q\in [0,1]. \end{equation}

Hence, $\mathcal{P}(q)\geq 0, \forall q\in [0,1].$ And according to (D.33), we have

(D.45) \begin{align} e^{-Aj_2}-e^{-A\tilde {q}_{2v_2}j_2}+A(1-\tilde {q}_{2v_2})j_2e^{-A\tilde {q}_{2v_2}j_2}&\leq 0\nonumber\\ \lambda \int \limits _0^{\infty }\int \limits _0^{\infty }\Big (e^{-Aj_2}-e^{-A\tilde {q}_{2v_2}j_2}+A(1-\tilde {q}_{2v_2})j_2e^{-A\tilde {q}_{2v_2}j_2}\Big )v_2(t,\lambda +y)d\Psi (y,j_2)&\leq 0\nonumber\\ \lambda \int \limits _0^{\infty }\int \limits _0^{\infty }(e^{-Aj_2}-e^{-A\tilde {q}_{2v_2}j_2})v_2(t,\lambda +y)d\Psi (y,j_2)+A(1-\tilde {q}_{2v_2})\tilde {c}_2v_2&\leq 0\nonumber\\ -\tilde {q}_{2v_2}\tilde {c}_2Av_2+\lambda \int \limits _0^{\infty }\int \limits _0^{\infty }(e^{-Aj_2}-e^{-A\tilde {q}_{2v_2}j_2})v_2(t,\lambda +y)d\Psi (y,j_2)&\leq -A\tilde {c}_2v_2. \end{align}

Now, (D.41) and (D.45) suggest that

(D.46) \begin{align} \left |\chi _3(q_{2v_1}^*,v_1)-\chi _3(q_{2v_2}^*,v_2)\right |&\leq \left |\lambda \int \limits _0^{\infty }\int \limits _0^{\infty }(e^{-Aj_2}-1)(v_1(t,\lambda +y)-v_2(t,\lambda +y))d\Psi (y,j_2)\right |\nonumber\\& \quad +\left |A\tilde {c}_2v_1-A\tilde {c}_2v_2\right |\nonumber\\ &\leq \left (\lambda \left (\int \limits _0^{\infty }e^{-Aj_2}dF_2(j_2)-1\right )+\left |A\tilde {c}_2\right |\right )D_{\lambda }(v_1,v_2) \end{align}

Since $F_2(j_2)$ has a finite Laplace transform under Assumption 2.3, we also have

(D.47) \begin{equation} \left |\chi _3(q_{2v_1}^*,v_1)-\chi _3(q_{2v_2}^*,v_2)\right |\leq \Delta _3D_{\lambda }(v_1,v_2). \end{equation}

Case 3: $\tilde {q}_{2v_1}\lt 1$ and $\tilde {q}_{2v_2}\lt 1$

Suppose that $\tilde {q}_{2f}\lt 1$ and $\tilde {q}_{2h}\lt 1$ , $\chi _3(q_{2v_1}^*,v_1)-\chi _3(q_{2v_2}^*,v_2)$ can be expressed as follows:

(D.48) \begin{align} \chi _3(q_{2v_1}^*,v_1)-\chi _3(q_{2v_2}^*,v_2)&=\tilde {q}_{2v_1}\tilde {c}_2Av_1+\lambda \int \limits _0^{\infty }\int \limits _0^{\infty }(e^{-A\tilde {q}_{2v_1}j_2}-1)v_1(t,\lambda +y)d\Psi (y,j_2)\nonumber\\ & \quad -\tilde {q}_{2v_2}\tilde {c}_2Av_2-\lambda \int \limits _0^{\infty }\int \limits _0^{\infty }(e^{-A\tilde {q}_{2v_2}j_2}-1)v_2(t,\lambda +y)d\Psi (y,j_2). \end{align}

Based on the Mean Value Theorem, (D.48) can be rewritten as

(D.49) \begin{equation} \begin{aligned} \chi _3(q_{2v_1}^*,v_1)-\chi _3(q_{2v_2}^*,v_2)=(v_1-v_2)\frac {d Q_2(q_{\zeta }^*,\zeta )}{d\zeta }|_{\zeta =\xi _2}, \end{aligned} \end{equation}

where $\xi _2\in \mathfrak{C}_e([0,T]\times (0,\infty ))$ and $\xi _2(s,\lambda )$ lies between $v_1(s,\lambda )$ and $v_2(s,\lambda )$ . $Q_2(q_{\zeta }^*,\zeta )$ is given by

(D.50) \begin{equation} Q_2(q_{\zeta }^*,\zeta )=\tilde {q}_{2\zeta }\tilde {c}_2A\zeta +\lambda \int \limits _0^{\infty }\int \limits _0^{\infty }(e^{-A\tilde {q}_{2\zeta }j_2}-1)\zeta (t,\lambda +y)d\Psi (y,j_2), \end{equation}

and

(D.51) \begin{equation} \tilde {c}_2=\lambda \int \limits _0^{\infty }\int \limits _0^{\infty }j_2e^{-A\tilde {q}_{2\zeta }j_2+\tilde {\psi }_1(t,T)y}d\Psi (y,j_2), \end{equation}

where $\zeta (t,\lambda )=e^{\tilde {\psi }(t,T)+\tilde {\psi }_1(t, T)\lambda }$ with some unknown bounded and Lipschitz continuous function $\tilde {\psi }(t,T)$ and $\tilde {\psi }_1(t,T)$ .

Taking the derivative with respect to $\zeta$ on both sides of (D.51), we have

(D.52) \begin{equation} \begin{aligned} \lambda \frac {d\tilde {q}_{2\zeta }}{d\zeta }\int \limits _0^{\infty }\int \limits _0^{\infty }Aj_2^2e^{-A\tilde {q}_{2\zeta }j_2+\tilde {\psi }_1(t,T)y}d\Psi (y,j_2)&=\lambda \int \limits _0^{\infty }\int \limits _0^{\infty }j_2ye^{-A\tilde {q}_{2\zeta }j_2+\tilde {\psi }_1(t,T)y}d\Psi (y,j_2)\frac {d\tilde {\psi }_1(t,T)}{d\zeta }. \end{aligned} \end{equation}

Therefore,

(D.53) \begin{equation} \zeta \frac {d\tilde {q}_{2\zeta }}{d\zeta }=\frac {\zeta \frac {d\tilde {\psi }_1(t,T)}{d\zeta }\int \limits _0^{\infty }\int \limits _0^{\infty }yj_2e^{-A\tilde {q}_{2\zeta }j_2+\tilde {\psi }_1(t,T)y}d\Psi (y,j_2)}{\lambda \int \limits _0^{\infty }\int \limits _0^{\infty }Aj_2^2e^{-A\tilde {q}_{2\zeta }j_2+\tilde {\psi }_1(t,T)y}d\Psi (y,j_2)} \end{equation}

Given (D.53), we can apply the same method illustrated in Section D.4.1 to derive the derivative of $Q_2(q_{\zeta }^*,\zeta )$ with respect to $\zeta$ as follows:

(D.54) \begin{align} \frac {d Q_2(q_{\zeta }^*,\zeta )}{d\zeta }&=\tilde {q}_{2\zeta }\tilde {c}_2A+\frac {d\tilde {\psi }_1(t,T)}{d\zeta }\lambda \int \limits _0^{\infty }\int \limits _0^{\infty }y\big(e^{-A\tilde {q}_{2\zeta }j_2}-1\big)e^{\tilde {\psi }_1(t,T)y}d\Psi (y,j_2).\nonumber\\ &=\tilde {q}_{2\zeta }\tilde {c}_2A+\frac {d\tilde {\psi }_1(t,T)}{d\zeta }\lambda \int \limits _0^{\infty }\int \limits _0^{\infty }ye^{-A\tilde {q}_{2\zeta }j_2+\tilde {\psi }_1(t,T)y}d\Psi (y,j_2)\nonumber\\ & \quad -\frac {d\tilde {\psi }_1(t,T)}{d\zeta }\lambda \int \limits _0^{\infty }\int \limits _0^{\infty }ye^{\tilde {\psi }_1(t,T)y}d\Psi (y,j_2). \end{align}

According to the same reason as in Section D.4.1 and based on the assumption that $\zeta (t,\lambda )\in \mathfrak{C}_e([0,T]\times (0,\infty ))$ is increasing in $t$ and has non-trivial bounded and Lipschits continuous functions $\tilde {\psi }(t,T)$ and $\tilde {\psi }_1(t,T)$ and Assumptions 2.3 and 2.4 for the joint distributions $\Psi (y,j_2)$ as well as marginal distributions $F_2(j_2)$ and $G(y)$ , we know that

(D.55) \begin{equation} \begin{aligned} \left |\chi _3(q_{2v_1}^*,v_1)-\chi _3(q_{2v_2}^*,v_2)\right |=\left |v_1-v_2\right |\left |\frac {d Q_2(q_{\zeta }^*,\zeta )}{d\zeta }|_{\zeta =\xi _2}\right |\lt \Delta _4D_{\lambda }(v_1,v_2), \end{aligned} \end{equation}

where $\Delta _4$ denotes some positive constant.

Therefore, we complete the proof of Lemma 4.1 as follows:

(D.56) \begin{align} & \left |\chi _1\big(\pi _{v_1}^*,v_1\big)-\chi _1(\pi _{v_2}^*,v_2)+\chi _2(q_{1v_1}^*,v_1)-\chi _1(q_{1v_2}^*,v_2)+\chi _3(q_{2v_1}^*,v_1)-\chi _3(q_{2v_2}^*,v_2)\right |\nonumber\\& \qquad\qquad\qquad \lt \Delta D_{\lambda }(v_1,v_2), \end{align}

where $\Delta \gt \Delta _1+\Delta _2+\max (\Delta _3,\Delta _4)$ denotes some positive constant.

D.5 Proof of Lemma 4.5

For any continuous functions $v_1, v_2\in \mathfrak{C}_e([0,T]\times (0,\infty ))$ , and based on Lemma 4.4, we can rewrite (4.5) as follows:

(D.57) \begin{equation} \begin{aligned} d(\mathcal{L}v_1,\mathcal{L}v_2)&\lt \sup _{t}e^{-\alpha (T-t)}\mathbb{E}\left [\int _t^Te^{\int \limits _t^s\widehat {A}(\tau ,T)d\tau }\Delta D_{\lambda }(v_1,v_2)ds|\mathcal{F}_t\right ]\\ &\leq \Delta d(v_1,v_2)\int _t^Te^{\int \limits _t^s\widehat {A}(\tau ,T)d\tau -\alpha (s-t)}ds, \end{aligned} \end{equation}

where $\widehat {A}(\tau ,T)=(c_1-\tilde {c}_1+c_2-\tilde {c}_2)A(\tau ,T)\gt 0.$ Hence, we complete the proof of Lemma 4.5.

D.6 Proof of Theorem4.6

To construct a contraction mapping, we need $\Delta \int _t^Te^{\int \limits _t^s\widehat {A}(\tau ,T)d\tau -\alpha (s-t)}ds\in [0,1)$ , $\forall t\in [0,T]$ . Let $\alpha \gt 0$ be a large enough constant such that $\Delta \int _0^Te^{\int \limits _0^s\widehat {A}(\tau ,T)d\tau -\alpha s}ds\in [0,1)$ . Then, we can define $\phi =\Delta \int _0^Te^{\int \limits _0^s\widehat {A}(\tau ,T)d\tau -\alpha s}ds\in [0,1)$ such that $d(\mathcal{L}v_1,\mathcal{L}v_2)\leq \phi d(v_1,v_2)$ , $\forall t\in [0,T]$ . To obtain a feasible range of $\alpha$ with a closed-form formula, we have

(D.58) \begin{equation} \Delta \int _0^Te^{\int \limits _0^s\widehat {A}(\tau ,T)d\tau -\alpha s}ds\lt \Delta \int _0^Te^{\widehat {A}(0,T)s-\alpha s}ds. \end{equation}

Then, let $\Delta \int _0^Te^{\widehat {A}(0,T)s-\alpha s}ds\lt 1$ , we have

(D.59) \begin{equation} \Delta \frac {\exp [(\gamma (\tilde {c}_1-c_1+\tilde {c}_2-c_2)e^{rT}-\alpha )T]-1}{\gamma (\tilde {c}_1-c_1+\tilde {c}_2-c_2)e^{rT}-\alpha }\lt 1, \end{equation}

which implies that $ \Delta \int _0^Te^{\int \limits _0^s\widehat {A}(\tau ,T)d\tau -\alpha s}ds\lt 1$ . Hence, when $\alpha \in \left \{\alpha \gt 0|\Delta \frac {\exp [(\gamma (\tilde {c}_1-c_1+\tilde {c}_2-c_2)e^{rT}-\alpha )T]-1}{\gamma (\tilde {c}_1-c_1+\tilde {c}_2-c_2)e^{rT}-\alpha }\lt 1\right \}$ , the mapping $\mathcal{L}$ : $[0,T]\times (0,\infty )$ $\to$ $\mathfrak{C}_e([0,T]\times (0,\infty ))$ is indeed a contraction mapping. Define $g^*(t,\lambda )$ be the solution of the fixed-point equation (3.12) and $g^n(t,\lambda )$ be the $n$ th iteration generated by the numerical method in (3.13). Then, we have

(D.60) \begin{equation} \begin{aligned} d(g^{n+1},g^*)=d(\mathcal{L}g^n,\mathcal{L}g^*)\leq \phi d(g^n,g^*), \end{aligned} \end{equation}

Therefore, the iterative method in (3.13) can provide the first order of convergence. To increase the convergence rate of the iterative method, Aitken’s delta-squared process or Steffensen’s method can be applied.

Footnotes

1 The dependent structure and the copula function of claim amounts $J_2$ and the self-exciting effect $Y$ are not reflected from the first moment of $C_2(t)$ but clearly affect the second moment of $C_2(t)$ . Numerical examples will be provided to illustrate optimal controls under those different pricing schemes.

2 Due to the limited space, numerical results under other copula distributions can be provided as requested.

References

Aït-Sahalia, Y., & Hurd, T. R. (2015). Portfolio choice in markets with contagion. Journal of Financial Econometrics, 14(1), 128.10.1093/jjfinec/nbv024CrossRefGoogle Scholar
Banach, S. (1922). Sur les opérations dans les ensembles abstraits et leur application aux équations intégrales. Fundamenta Mathematicae, 3(1), 133181.10.4064/fm-3-1-133-181CrossRefGoogle Scholar
Brachetta, M., Callegaro, G., Ceci, C., & Sgarra, C. (2024). Optimal reinsurance via BSDEs in a partially observable model with jump clusters. Finance and Stochastics, 28(2), 453495.10.1007/s00780-023-00523-zCrossRefGoogle Scholar
Browne, S. (1995). Optimal investment policies for a firm with a random risk process: Exponential utility and minimizing the probability of ruin. Mathematics of Operations Research, 20(4), 937958.10.1287/moor.20.4.937CrossRefGoogle Scholar
Ceci, C., Colaneri, K., & Cretarola, A. (2022). Optimal reinsurance and investment under common shock dependence between financial and actuarial markets. Insurance: Mathematics and Economics, 105, 252278.Google Scholar
Chen, P., & Yam, S. C. P. (2013). Optimal proportional reinsurance and investment with regime-switching for mean-variance insurers. Insurance: Mathematics and Economics, 53(3), 871883.Google Scholar
Ciesielski, K. (2007). On Stefan Banach and some of his results. Banach Journal of Mathematical Analysis, 1(1), 110.10.15352/bjma/1240321550CrossRefGoogle Scholar
Dassios, A., & Zhao, H. (2011). A dynamic contagion process. Advances in Applied Probability, 43(3), 814846.10.1239/aap/1316792671CrossRefGoogle Scholar
Dassios, A., & Zhao, H. (2012). Ruin by dynamic contagion claims. Insurance: Mathematics and Economics, 51(1), 93106.Google Scholar
Delong, L., & Gerrard, R. (2007). Mean-variance portfolio selection for a non-life insurance company. Mathematical Methods of Operations Research, 66(2), 339367.10.1007/s00186-007-0152-2CrossRefGoogle Scholar
Duffie, D., Pan, J., & Singleton, K. (2000). Transform analysis and asset pricing for affine jump-diffusions. Econometrica, 68(8), 13431376.10.1111/1468-0262.00164CrossRefGoogle Scholar
Emms, P., & Haberman, S. (2007). Asymptotic and numerical analysis of the optimal investment strategy for an insurer. Insurance: Mathematics and Economics, 40(1), 113134.Google Scholar
Hainaut, D. (2017). Contagion modelling between the financial and insurance markets with time changed processes. Insurance: Mathematics and Economics, 74, 6377.Google ScholarPubMed
Hipp, C., & Plum, M. (2000). Optimal investment for insurers. Insurance: Mathematics and Economics, 27(2), 216228.Google Scholar
Hipp, C., & Vogt, M. (2003). Optimal dynamic XL-reinsurance. ASTIN Bulletin: The Journal of the IAA, 33(2), 193207.10.2143/AST.33.2.503690CrossRefGoogle Scholar
Højgaard, B., & Taksar, M. (1998). Optimal proportional reinsurance policies for diffusion models. Scandinavian Actuarial Journal, 1998(2), 166180.10.1080/03461238.1998.10414000CrossRefGoogle Scholar
Jang, J., & Oh, R. (2021). A review on Poisson, Cox, Hawkes, shot-noise Poisson and dynamic contagion process and their compound processes. Annals of Actuarial Science, 15(3), 623644.10.1017/S1748499520000287CrossRefGoogle Scholar
Liu, C. S., & Yang, H. (2004). Optimal investment for an insurer to minimize its probability of ruin. North American Actuarial Journal, 8(2), 1131.10.1080/10920277.2004.10596134CrossRefGoogle Scholar
Liu, G., Jin, Z., & Li, S. (2021). Optimal investment, consumption, and life insurance strategies under a mutual-exciting contagious market. Insurance: Mathematics and Economics, 101, 508524.Google Scholar
Liu, G., Jin, Z., Li, S., & Zhang, J. (2022). Stochastic asset allocation and reinsurance game under contagious claims. Finance Research Letters, 49, 103123.10.1016/j.frl.2022.103123CrossRefGoogle Scholar
Meng, H., Li, S., & Jin, Z. (2015). A reinsurance game between two insurance companies with nonlinear risk processes. Insurance: Mathematics and Economics, 62, 9197.Google Scholar
Schmidli, H. (2001). Optimal proportional reinsurance policies in a dynamic setting. Scandinavian Actuarial Journal, 2001(1), 5568.10.1080/034612301750077338CrossRefGoogle Scholar
Schweizer, B., & Sklar, A. (1983). Probabilistic metric space. North-Holland.Google Scholar
Shen, Y., & Zeng, Y. (2015). Optimal investment-reinsurance strategy for mean-variance insurers with square-root factor process. Insurance: Mathematics and Economics, 62, 118137.Google Scholar
Sklar, A. (1959). Fonctions de repartition a n dimensions et leurs marges. Publications de I’Institut de statistique de I’Université de Paris, 8, 229231.Google Scholar
Wu, F., Shen, Y., Zhang, X., & Ding, K. (2024). Optimal claim-dependent proportional reinsurance under a self-exciting claim model. Journal of Optimization Theory and Applications, 201(3), 12291255.10.1007/s10957-024-02429-yCrossRefGoogle Scholar
Yang, H., & Zhang, L. (2005). Optimal investment for insurer with jump-diffusion risk process. Insurance: Mathematics and Economics, 37(3), 615634.Google Scholar
Zeng, Y., & Li, Z. (2011). Optimal time-consistent investment and reinsurance policies for mean-variance insurers. Insurance: Mathematics and Economics, 49(1), 145154.Google Scholar
Figure 0

Table 1. Parameters and values

Figure 1

Figure 1 Optimal ordinary reinsurance policy under dynamic contagion processes.

Figure 2

Figure 2 Optimal catastrophic reinsurance policy under dynamic contagion processes.

Figure 3

Figure 3 Optimal asset allocation policy under dynamic contagion processes.

Figure 4

Figure 4 Value function under dynamic contagion processes.