Skip to main content
Log in

Optimal credit investment and risk control for an insurer with regime-switching

  • Published:
Mathematics and Financial Economics Aims and scope Submit manuscript

Abstract

This paper studies an optimal investment and risk control problem for an insurer with default contagion and regime-switching. The insurer in our model allocates his/her wealth across multi-name defaultable stocks and a riskless bond under regime-switching risk. Default events have an impact on the distress state of the surviving stocks in the portfolio. The aim of the insurer is to maximize the expected utility of the terminal wealth by selecting optimal investment and risk control strategies. We characterize the optimal trading strategy of defaultable stocks and risk control for the insurer. By developing a truncation technique, we analyze the existence and uniqueness of global (classical) solutions to the recursive HJB system. We prove the verification theorem based on the (classical) solutions of the recursive HJB system.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

References

  1. Andruszkiewicz, G., Davis, H.A., LIeo, S.: Risk-sensitive investment in a finite-factor model. Stochastics 89, 89–114 (2016)

    Article  MathSciNet  Google Scholar 

  2. Birge, J., Bo, L., Capponi, A.: Risk-sensitive asset management and cascading defaults. Math. Oper. Res. 43, 1–28 (2018)

    Article  MathSciNet  Google Scholar 

  3. Bo, L., Wang, S.: Optimal investment and risk control for an insurer with stochastic factor. Oper. Res. Lett. 45, 259–265 (2017)

    Article  MathSciNet  Google Scholar 

  4. Carr, P., Wu, L.: Stock options and credit default swaps: a joint framework for valuation and estimation. J. Financ. Econ. 8, 409–449 (2010)

    Google Scholar 

  5. Carr, P., Linetsky, V.: A jump to default extended CEV model: an application of Bessel processes. Financ. Stoch. 10, 303–330 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  6. Campbell, J., Taksler, G.: Equity volatility and corporate bond yields. J. Financ. 63, 2321–2349 (2003)

    Article  Google Scholar 

  7. Capponi, A., Figueroa Lopez, J.E.: Dynamic portfolio optimization with a defaultable security and regime-switching markets. Math. Financ. 24, 207–249 (2014)

    Article  MATH  Google Scholar 

  8. Capponi, A., Figueroa Lopez, J.E., Nisen, J.: Pricing and semi-martingale representations of vulnerable contingent claims in regime-switching markets. Math. Financ. 24, 250–288 (2014)

    Article  MATH  Google Scholar 

  9. Elliott, R.J., Siu, T.K.: Robust optimal portfolio choice under Markovian regime-switching model. Meth. Comput. Appl. Probab. 11, 145–157 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  10. Elliott, R.J., Siu, T.K., Chan, L., Lau, J.W.: Pricing options under a generalized Markov-modulated jump-diffusion model. Stoch. Anal. Appl. 25, 821–843 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  11. Frey, R., Backhaus, J.: Pricing and hedging of portfolio credit derivatives with interacting default intensities. Int. J. Theor. Appl. Financ. 11, 611–634 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  12. Jiao, Y., Kharroubi, I., Pham, H.: Optimal investment under multiple defaults risk: a BSDE-decomposition approach. Ann. Appl. Probab. 23, 455–491 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  13. Jones, P.M.: Trade Credit Insurance, p. 15. Primer Series on Insurance, Issue (2010)

  14. Mendoza-Arriaga, R., Linetsky, V.: Multivariate subordination of Markov processes with financial applications. Math. Financ. 26, 699–747 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  15. Merton, R.: Lifetime portfolio selection under uncertainty: the continuous-time case. Rev. Econom. Stats. 51, 247–257 (1969)

    Article  Google Scholar 

  16. Merton, R.: Optimum consumption and portfolio rules in a continuous-time model. J. Econ. Theory 3, 373–413 (1971)

    Article  MathSciNet  MATH  Google Scholar 

  17. Peng, X., Wang, W.: Optimal investment and risk control for an insurer under inside information. Insur. Math. Econom. 69, 104–116 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  18. Shen, Y., Siu, T.K.: Consumption-portfolio optimization and filtering in a hidden Markov-modulated asset price model. J. Ind. Manag. Optim. 12, 23–46 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  19. Smith, H.L.: Monotone Dynamical Systems: An Introduction to the Theory of Competitive and Cooperative Systems. Mathematical Surveys and Monographs. AMS, Providence (2008)

    Book  Google Scholar 

  20. Zhang, Q., Zhou, X.Y.: Valuation of stock loans with regime switching. SIAM J. Control Optim. 48, 1229–1250 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  21. Zhou, X.Y., Yin, G.: Markowitz’s mean-variance portfolio selection with regime switching: a continuous-time model. SIAM J. Control Optim. 42, 1466–1482 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  22. Zou, B., Cadenillas, A.: Optimal investment and risk control policies for an insurer: expected utility maximization. Insurance: Math. Econom. 58, 57–67 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  23. Zou, B., Cadenillas, A.: Optimal investment and liability ratio policies in a multidimensional regime switching model. Risk 5, 1–22 (2017)

    Google Scholar 

Download references

Acknowledgements

The authors gratefully acknowledge the constructive and insightful comments provided by one anonymous reviewer and Editor-in-Chief, Prof. Ulrich Horst, which helped to greatly improve the quality of the manuscript. This research of L. Bo and H. Liao was supported in part by the NSF of China under Grant 11471254, The Key Research Program of Frontier Sciences, CAS under Grant QYZDB-SSW-SYS009, and Fundamental Research Funds for Central Universities under Grant WK3470000008.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Lijun Bo.

Technical proofs

Technical proofs

Proof of Lemma 4.1

Define \(f(x)=Bx\) for \(x\in \mathbb {R}^m\). By virtue of Proposition 1.1 of Charter 3 in Smith [19], it suffices to verify that \(f:\mathbb {R}^m\rightarrow \mathbb {R}^m\) is of type K, i.e., for any \(x,y\in \mathbb {R}^m\) satisfying \(x\le y\) and \(x_i=y_i\) for some \(i=1,\ldots ,m\), then \(f_i(x)\le f_i(y)\). Notice that \(b_{ij}\ge 0\) for all \(i\ne j\). Then, it holds that

$$\begin{aligned} f_i(x)&=(Bx)_i=\sum _{j=1}^mb_{ij}x_j=b_{ii}x_i+\sum _{j=1,j\ne i}^mb_{ij}x_j\nonumber \\&=b_{ii}y_i+\sum _{j=1,j\ne i}^mb_{ij}x_j \le b_{ii}y_i+\sum _{j=1,j\ne i}^mb_{ij}y_j=f_i(y), \end{aligned}$$
(A.1)

and hence f is of type K. Thus, we complete the proof of the lemma. \(\square \)

Proof of Lemma 4.2

The expression of the solution \(\varphi (t,e_n)\) given by (24) is obvious. Notice that \(e_m\gg 0\) and \(q_{ij}\ge 0\) for all \(i\ne j\) since \(Q=(q_{ij})_{m\times m}\) is the generator of the Markov chain. Then, in order to prove \(\varphi (t,e_n)\gg 0\) for all \(t\in [0,T]\), using Lemma 4.1, it suffices to verify \([A^{(n)}]_{ij}\ge 0\) for all \(i\ne j\), however, \([A^{(n)}]_{ij}=q_{ij}\) for all \(i\ne j\) using (21). Thus, we have verified the condition given in Lemma 4.1, and hence \(\varphi (t,e_n)\gg 0\) for all \(t\in [0,T]\). \(\square \)

Proof of Lemma 4.3

It suffices to prove that, for any \(x,y\in \mathbb {R}^m\) satisfying \(x,y\ge \varepsilon e_m^{\top }\) with \(\varepsilon >0\), there exists a constant \(C=C(\varepsilon )>0\) depending on \(\varepsilon >0\) only such that \(|G_i^{(k)}(t,x)-G_i^{(k)}(t,y)|\le C\Vert x-y\Vert \) for each \(i=1,\ldots ,m\). Since \(\sigma (i)\sigma (i)^{\top }\) is also positive definite, \(\sigma ^{(k)}(i)\sigma ^{(k)}(i)^{\top }\) is positive definite. Hence, there exists a constant \(\delta >0\) such that \((\pi ^{(k)})^{\top }\sigma ^{(k)}(i)\sigma ^{(k)}(i)^{\top }\pi ^{(k)}\ge \delta \Vert \pi ^{(k)}\Vert ^2\). Then, for any \((\pi ^{(k)},l)\in \mathcal{U}^{(k)}\), there exists a positive constant \(C_1>0\) such that

$$\begin{aligned}&\frac{\gamma (\gamma -1)}{2}\left\{ (\pi ^{(k)})^{\top }\sigma ^{(k)}(i)\sigma ^{(k)}(i)^{\top }\pi ^{(k)}+l^2\big (\phi (i)\phi (i)^{\top }+\bar{\phi }(i)\bar{\phi }(i)^{\top }\big )-2l(\pi ^{(k)})^{\top }\sigma ^{(k)}(i)\phi (i)\right\} \nonumber \\&\quad =\frac{\gamma (\gamma -1)}{2}\left\{ l^2\bar{\phi }(i)\bar{\phi }(i)^{\top }+\Vert \sigma ^{(k)}(i)^\top \pi ^{(k)}-l\phi (i)\Vert ^2\right\} \nonumber \\&\quad \le \frac{\gamma (\gamma -1)}{2}\left\{ \alpha l^2\bar{\phi }(i)\bar{\phi }(i)^{\top }+\frac{1}{2}(1-\alpha )\Vert \sigma ^{(k)}(i)^\top \pi ^{(k)}\Vert ^2-(1-\alpha )\Vert l\phi (i)\Vert ^2\right\} \nonumber \\&\quad =\frac{\gamma (\gamma -1)}{2}\left\{ l^2(\alpha \bar{\phi }(i)\bar{\phi }(i)^{\top }-(1-\alpha )\Vert \phi (i)\Vert ^2)+\frac{1}{2}(1-\alpha )\Vert \sigma ^{(k)}(i)^\top \pi ^{(k)}\Vert ^2\right\} \nonumber \\&\quad \le -C_1(\Vert \pi ^{(k)}\Vert ^2+l^2), \end{aligned}$$
(A.2)

where the constant \(\alpha \in (\max _{i=1,\ldots ,m}\big \{\frac{\Vert \phi (i)\Vert ^2}{\bar{\phi }(i)\bar{\phi }(i)^{\top }+\Vert \phi (i)\Vert ^2}\big \},1)\). On the other hand, for any \((\pi ^{(k)},l)\in {{\mathcal {U}}}^{(k)}\), it holds that

$$\begin{aligned} \gamma \big \{(\pi ^{(k)})^{\top }\theta ^{(k)}(i)+(p^{(k)}(i)-c(i))l\big \}&\le \gamma \Vert \theta ^{(k)}(i)\Vert \Vert \pi ^{(k)}\Vert +\gamma |p^{(k)}(i)-c(i)|l\nonumber \\&\le C_2\sqrt{\Vert \pi ^{(k)}\Vert ^2+l^2}, \end{aligned}$$
(A.3)

where the constant \(C_2:=\max _{i=1,\ldots ,m}\big \{\gamma \sqrt{\Vert \theta ^{(k)}(i)\Vert ^2+|p^{(k)}(i)-c(i)|^2}\big \}>0\). Finally, for any \((\pi ^{(k)},l)\in {{\mathcal {U}}}^{(k)}\), we have that \(\{(1-lg(i))^\gamma -1\}\nu ^{(k)}(i)\le C_3l\), where the constant \(C_3:=\max _{i=1,\ldots ,m}\{\gamma g(i)\nu ^{(k)}(i)\}>0\). Then, by virtue of (29), it follows that, for any \((\pi ^{(k)},l)\in {{\mathcal {U}}}^{(k)}\) and \(i=1,\ldots ,m\),

$$\begin{aligned} H^{(k)}((\pi ^{(k)},l),i)\le -C_1\big (\Vert \pi ^{(k)}\Vert ^2+l^2\big )+C_4\sqrt{\Vert \pi ^{(k)}\Vert ^2+l^2}. \end{aligned}$$
(A.4)

Here \(C_4=C_2+C_3\). This yields that there exists a constant \(C_5>0\) such that when \((\pi ^{(k)},l)\in {{\mathcal {U}}}^{(k)}\) and \(\Vert \pi ^{(k)}\Vert ^2+l^2>C_5\), we have \(H^{(k)}((\pi ^{(k)},l),i)<0\) for all \(i=1,\ldots ,m\), and meanwhile, for \(x\ge \varepsilon e_{m}^{\top }\) with \(\varepsilon >0\),

$$\begin{aligned}&\sum _{j\notin \{j_1,\ldots ,j_k\}}(1-\pi _j^{(k)})^\gamma h_{j}^{(k)}(i)\varphi ^{(l+1),j}(t,i)+H^{(k)}((\pi ^{(k)},l),i)x_i\nonumber \\&\quad \le \big (1+\Vert \pi ^{(k)}\Vert \big )^\gamma \sum _{j\notin \{j_1,\ldots ,j_k\}}h_{j}^{(k)}(i)\varphi ^{(l+1),j}(t,i)+\varepsilon H^{(k)}((\pi ^{(k)},l),i)\nonumber \\&\quad \le C_6\big (1+\Vert \pi ^{(k)}\Vert ^\gamma \big )\sum _{j\notin \{j_1,\ldots ,j_k\}}h^{(k)}_{j}(i)+\varepsilon \left\{ -C_1(\Vert \pi ^{(k)}\Vert ^2+l^2)+C_4\sqrt{\Vert \pi ^{(k)}\Vert ^2+l^2}\right\} \nonumber \\&\quad \le -\varepsilon C_1\big (\Vert \pi ^{(k)}\Vert ^2+l^2\big )+\varepsilon C_4\sqrt{\Vert \pi ^{(k)}\Vert ^2+l^2}+C_7\big (\sqrt{\Vert \pi ^{(k)}\Vert ^2+l^2}\big )^{\gamma }+C_8, \end{aligned}$$
(A.5)

for some constants \(C_6,C_7,C_8>0\). Notice that we used the recursive assumption that the HJB system (18) admits a positive unique (classical) solution \(\varphi ^{(k+1),j}(t)\) on \(t\in [0,T]\) for \(j\notin \{j_1,\ldots ,j_k\}\). Then \(\varphi ^{(k+1),j}(t)\) is continuous on [0, T], and hence \(\varphi ^{(k+1),j}(t)\) is bounded on [0, T]. From the estimate (A.5), it follows that, for any \(x\ge \varepsilon e_m^{\top }\), there exists a positive constant \(C_9=C_9(\varepsilon )\) such that when \((\pi ^{(k)},l)\in {{\mathcal {U}}}^{(k)}\), \(\Vert \pi ^{(k)}\Vert ^2+l^2>C_9\) and \(x\ge \varepsilon e_m^{\top }\), it holds that, for each \(i=1,\ldots ,m\),

$$\begin{aligned}&\sum _{j\notin \{j_1,\ldots ,j_k\}}(1-\pi _j^{(k)})^\gamma h^{(k)}_{j}(i)\varphi ^{(l+1),j}(t,i)+H^{(k)}((\pi ^{(k)},l),i)x_i<0. \end{aligned}$$
(A.6)

On the other hand, for \(i=1,\ldots ,m\), it holds that

$$\begin{aligned} G^{(k)}_i(t,x)=&\sup _{(\pi ^{(k)},l)\in {{\mathcal {U}}}^{(k)}}\left\{ \sum _{j\notin \{j_1,\ldots ,j_k\}}(1-\pi _{j}^{(k)})^\gamma h_{j}^{(k)}(i)\varphi ^{(k+1),j}(t,i)+H^{(k)}((\pi ^{(k)},l),i)x_i\right\} \nonumber \\ \ge&\sum _{j\notin \{j_1,\ldots ,j_k\}} h^{(k)}_{j}(i)\varphi ^{(k+1),j}(t,i)+H^{(k)}((0e_{n-k}^{\top },0),i)x_i\nonumber \\ =&\sum _{j\notin \{j_1,\ldots ,j_k\}} h_{j}^{(k)}(i)\varphi ^{(k+1),j}(t,i)>0. \end{aligned}$$
(A.7)

Thus, using the estimate (A.6), we have that, for all \(x\ge \varepsilon e_m^{\top }\),

$$\begin{aligned} G^{(k)}_i(t,x)=&\sup _{\begin{array}{c} (\pi ^{(k)},l)\in \mathcal{U}^{(k)}\\ \Vert \pi ^{(k)}\Vert ^2+l^2\le C_9(\varepsilon ) \end{array}}\left\{ \sum _{j\notin \{j_1,\ldots ,j_k\}}(1-\pi _{j}^{(k)})^\gamma h_{j}^{(k)}(i)\varphi ^{(k+1),j}(t,i)+H^{(k)}((\pi ^{(k)},l),i)x_i\right\} . \end{aligned}$$
(A.8)

It follows from (A.8) and (29) that, for all \(x,y\ge \varepsilon e_m^{\top }\),

$$\begin{aligned} G^{(k)}_i(t,x)=&\sup _{\begin{array}{c} (\pi ^{(k)},l)\in {{\mathcal {U}}}^{(k)}\\ \Vert \pi ^{(k)}\Vert ^2+l^2\le C_9(\varepsilon ) \end{array}}\Bigg \{\sum _{j\notin \{j_1,\ldots ,j_k\}}(1-\pi _{j}^{(k)})^\gamma h_{j}^{(k)}(i)\varphi ^{(k+1),j}(t,i)+H^{(k)}((\pi ^{(k)},l),i)y_i\nonumber \\&\qquad \qquad \qquad +H^{(k)}((\pi ^{(k)},l),i)(x_i-y_i)\Bigg \}\nonumber \\ \le&\sup _{\begin{array}{c} (\pi ^{(k)},l)\in {{\mathcal {U}}}^{(k)}\\ \Vert \pi ^{(k)}\Vert ^2+l^2\le C_9(\varepsilon ) \end{array}}\Bigg \{\sum _{j\notin \{j_1,\ldots ,j_k\}}(1-\pi _{j}^{(k)})^\gamma h_{j}^{(k)}(i)\varphi ^{(k+1),j}(t,i)+H^{(k)}((\pi ^{(k)},l),i)y_i\nonumber \\&\qquad \qquad \qquad +\left| H^{(k)}((\pi ^{(k)},l),i)\right| \left| x_i-y_i\right| \Bigg \}\nonumber \\ \le&G^{(k)}_i(t,y)+\left| x_i-y_i\right| \sup _{\begin{array}{c} (\pi ^{(k)},l)\in {{\mathcal {U}}}^{(k)}\\ \Vert \pi ^{(k)}\Vert ^2+l^2\le C_9(\varepsilon ) \end{array}}\left\{ \left| H^{(k)}((\pi ^{(k)},l),i)\right| \right\} \nonumber \\ \le&G^{(k)}_i(t,y)+C(\varepsilon )\left| x_i-y_i\right| , \end{aligned}$$
(A.9)

where \(C(\varepsilon )>0\) is a constant which depends on \(\varepsilon >0\) only. Then, the above estimate results in the validity of the estimate (30) for all \(x,y\in \mathbb {R}^m\) satisfying \(x,y\ge \varepsilon e_m^{\top }\). Thus, we complete the proof of the lemma. \(\square \)

Proof of Lemma 4.4

For \(p>0\), let \(g_{1}^{(p)}(t)=(g_{1i}^{(p)}(t);\ i=1,\ldots ,m)^{\top }\) be the solution to the following dynamical system given by

$$\begin{aligned} \left\{ \begin{aligned} \frac{d}{dt}g_{1}^{(p)}(t)=&f(t,g_{1}^{(p)}(t))+\tilde{f}(t,g^{(p)}_{1}(t))+\frac{1}{p}e_m^{\top },\ \text { in }(0,T];\\ g_{1}^{(p)}(0)=&\xi _1+\frac{1}{p}e_m^{\top }. \end{aligned} \right. \end{aligned}$$
(A.10)

Then, for all \(t\in (0,T]\), it holds that

$$\begin{aligned} \Vert g_{1}^{(p)}(t)-g_1(t)\Vert \le&\Vert g_{1}^{(p)}(0)-g_1(0)\Vert +\int _0^t\big \Vert f(s,g_{1}^{(p)}(s))-f(s,g_1(s))\big \Vert ds\nonumber \\&+\int _0^t\big \Vert \tilde{f}(s,g_{1}^{(p)}(s))-\tilde{f}(s,g_1(s))\big \Vert ds+\frac{1}{p}\int _0^t\Vert e_m\Vert ds\nonumber \\ \le&\frac{2}{p}\Vert e_m\Vert +(C+\tilde{C})\int _0^t\big \Vert g_{1}^{(p)}(s)-g_1(s)\big \Vert ds. \end{aligned}$$

Here \(C>0\) (resp. \(\tilde{C}>0\)) is the Lipschitz constant of f(tx) (resp. \(\tilde{f}(t,x)\)) in x. Then, the Gronwall’s lemma yields that \(g_{1}^{(p)}(t)\rightarrow g_1(t)\) for all \(t\in [0,T]\) as \(p\rightarrow \infty \). We claim that \(g_{1}^{(p)}(t)\gg g_2(t)\) for all \(t\in [0,T]\). If the claim were false, notice that \(g_{1}^{(p)}(0)\gg g_2(0)\), and \(g_1^{(p)}(t),g_2(t)\) are continuous on [0, T], then there exists a \(t_0\in (0,T]\) such that \(g_{1}^{(p)}(s)\ge g_2(s)\) on \(s\in [0,t_0]\) and \(g_{1i}^{(p)}(t_0)=g_{2i}(t_0)\) for some \(i\in \{1,\ldots ,m\}\). Since \(t_0>0\), \(g_1^{(p)}(t)\) and \(g_2(t)\) are differentiable on (0, T], we have that

$$\begin{aligned} \frac{d}{dt}g_{1i}^{(p)}(t)\big |_{t=t_0}=\lim _{\epsilon \rightarrow 0}\frac{g_{1i}^{(p)}(t_0)-g_{1i}^{(p)}(t_0-\epsilon )}{\epsilon } \le \lim _{\epsilon \rightarrow 0}\frac{g_{2i}(t_0)-g_{2i}(t_0-\epsilon )}{\epsilon }= \frac{d}{dt}g_{2i}(t)\big |_{t=t_0}. \end{aligned}$$

On the other hand, since \(f(t,\cdot )\) satisfies the type K condition for each \(t\in [0,T]\) and \(\tilde{f}(t,x)\ge 0\) for all \((t,x)\in [0,T]\times \mathbb {R}^m\), for the above i, we also have that

$$\begin{aligned} \frac{d}{dt}g_{1i}^{(p)}(t)\big |_{t=t_0}=&f_i(t_0,g_{1i}^{(p)}(t_0))+\tilde{f}_i(t_0,g_{1}^{(p)}(t_0))+\frac{1}{p}\nonumber \\ >&f_i(t_0,g_{1i}^{(p)}(t_0))\ge f_i(t_0,g_2(t_0))=\frac{d}{dt}g_{2i}(t)\big |_{t=t_0}. \end{aligned}$$
(A.11)

This results in a contradiction, and hence \(g_{1}^{(p)}(t)\gg g_2(t)\) for all \(t\in [0,T]\). Thus, it holds that \(g_1(t)\ge g_2(t)\) for all \(t\in [0,T]\) by letting p tend to infinity. \(\square \)

Proof of Theorem 4.6

For \((t,x,i,z)\in [0,T]\times \mathbb {R}_+\times \{1,\ldots ,m\}\times {{\mathcal {S}}}\), note that \(\varphi (T,i,z)=\frac{1}{\gamma }\). Then, by virtue of Itô’s formula, for all \((\pi ,l)\in \tilde{{\mathcal {U}}}\), it follows that

$$\begin{aligned}&\frac{1}{\gamma }(X^{\pi ,l}(T))^\gamma =(X^{\pi ,l}(t))^\gamma \varphi (t,Y(t),Z(t))+\int _t^T(X^{\pi ,l}(s))^\gamma \frac{\partial \varphi (s,Y(s),Z(s))}{\partial s}ds\nonumber \\&\qquad +\int _t^T\gamma (X^{\pi ,l}_s)^{\gamma -1}\varphi (s,Y(s),Z(s))dX^{\pi ,l}(s)^c\\&\qquad +\frac{\gamma (\gamma -1)}{2}\int _t^T(X^{\pi ,l}(s))^{\gamma -2}\varphi (s,Y(s),Z(s))d[X^{\pi ,l},X^{\pi ,l}]^c(s)\\&\qquad +\int _t^T\varphi (s,Y(s-),Z(s-))(X^{\pi ,l}(s-))^\gamma [(1-l(s)g(Y(s-)))^\gamma -1]dN(s)\\&\qquad +\sum _{j=1}^n\int _t^T(X^{\pi ,l}(s-))^\gamma \big [(1-\pi _j(s-))^\gamma \varphi (s,Y(s-),Z^j(s-))-\varphi (s,Y(s-),Z(s-))\big ]dZ_j(s)\\&\qquad +\int _t^T\sum _{j\ne Y(s-)}(X^{\pi ,l}(s-))^\gamma \big [\varphi (s,j,Z(s-))-\varphi (s,Y(s-),Z(s-))\big ]dH_{Y(s-),j}(s)\\&\quad =(X^{\pi ,l}(t))^\gamma \varphi (t,Y(t),Z(t))+\int _t^T(X^{\pi ,l}(s))^\gamma \mathcal{A}(\pi ,l;s,Y(s),Z(s))ds+M^{\pi ,l}(T)-M^{\pi ,l}(t). \end{aligned}$$

Here for \((\pi ,l)\in (-\infty ,1]^n\times [0,\infty )\) and \((t,i,z)\in [0,T]\times \{1,\ldots ,m\}\times {{\mathcal {S}}}\), the coefficient is given by

$$\begin{aligned}&{{\mathcal {A}}}(\pi ,l;t,i,z)\nonumber \\&\quad =\frac{\partial \varphi (t,i,z)}{\partial t}+\Bigg \{\gamma \Big [r(i)+\pi ^{\top }(I-diag(z))\theta (i,z)+\pi ^{\top }(I-diag(z))h(i,z)+(p(i,z)-c(i))l\Big ]\\&\qquad +\frac{\gamma (\gamma -1)}{2}\Big [\pi ^{\top }(I-diag(z))\sigma (i)\sigma (i)^{\top }(I-diag(z))\pi +l^2\big (\phi (i)\phi (i)^{\top }+\bar{\phi (i)}\bar{\phi (i)}^{\top }\big )\\&\qquad -2l\pi ^{\top }(I-diag(z))\sigma (i)\phi (i)^{\top }\Big ]+[(1-lg(i))^\gamma -1]\nu (i,z)\bigg \}\varphi (t,i,z)\\&\qquad +\sum _{j=1}^n[(1-\pi _j)^\gamma \varphi (t,i,z^j)-\varphi (t,i,z)](1-z_j)h_j(i,z)+\sum _{j\ne i}[\varphi (t,j,z)-\varphi (t,i,z)]q_{ij}, \end{aligned}$$

and the \( \mathbb {P} \)-(local) martingale is defined as

$$\begin{aligned} M^{\pi ,l}(t)=&\int _0^t\gamma (X^{\pi ,l}(s))^\gamma \varphi (s,Y(s),Z(s))\big [\pi (s)^{\top }(I-diag(Z(s)))\sigma (Y(s))-l(s)\phi (Y(s))\big ]dW(s)\\&+\int _0^t\gamma (X^{\pi ,l}(s))^\gamma \varphi (s,Y(s),Z(s))l(s)\bar{\phi }(Y(s))d\bar{W}(s)\\&+\int _0^t(X^{\pi ,l}(s))^\gamma \varphi (s,Y(s-),Z(s-))[(1-l(s)g(Y(s-)))^\gamma -1]d\tilde{N}(s)\\&+\sum _{j=1}^n\int _0^T(X^{\pi ,l}(s-))^\gamma [(1-\pi _j(s))^\gamma \varphi (s,Y(s-),Z(s-)^j)-\varphi (s,Y(s-),Z(s-))]dM_j(s)\\&+\int _0^t\sum _{j\ne Y(s-)}(X^{\pi ,l}(s-))^\gamma \big [\varphi (s,j,Z(s-))-\varphi (s,Y(s-),Z(s-))]d\tilde{H}_{Y(s-),j}(s), \end{aligned}$$

where we used the following \( \mathbb {P} \)-martingale processes given by, for \(t\in [0,T]\),

$$\begin{aligned} \tilde{N}(t)&:=N(t)-\int _0^t\nu (Y(s),Z(s))ds,\quad \tilde{H}_{ij}(t):=H_{ij}(t)-\int _0^tq_{ij}{\mathbb {1}}_{Y(s)=i}ds, \end{aligned}$$

for all \(i,j\in \{1,\ldots ,m\}\) and \(i\ne j\). Here, we recall that the process \(H_{ij}(t)\) is defined by (15). Using (18), (23) and (38), for \(t\in [0,T]\times \{1,\ldots ,m\}\times {{\mathcal {S}}}\), we have that \(\mathcal{A}(\pi ,l;t,i,z)\le {{\mathcal {A}}}(\pi ^*,l^*;t,i,z)=0\) for all \((\pi ,l)\in {{\mathcal {U}}}\). Moreover, define \(\tau _a:=\inf \{s\ge t;\ |X^{\pi ,l}(s)|>a\}\) for \(a>0\). Eq. (12) gives that, for \(s\in [t,T]\),

$$\begin{aligned} X^{\pi ,l}(s\wedge \tau _a)=&X^{\pi ,l}(s\wedge \tau _a-)\nonumber \\&\times [1-\tilde{\pi }^\top (s\wedge \tau _a)\Delta M(s\wedge \tau _a)-\tilde{l}(s\wedge \tau _a)g(Y(s\wedge \tau _a-))\Delta N(s\wedge \tau _a)], \end{aligned}$$
(A.12)

where the feedback controls are given by

$$\begin{aligned} \tilde{\pi }(s\wedge \tau _a)=\,&\pi \big (s\wedge \tau _a,X^{\pi ,l}(s\wedge \tau _a-),Y(s\wedge \tau _a-),Z(s\wedge \tau _a-)\big ),\\ \tilde{l}(s\wedge \tau _a)=\,&l\big (s\wedge \tau _a,X^{\pi ,l}(s\wedge \tau _a-),Y(s\wedge \tau _a-),Z(s\wedge \tau _a-)\big ). \end{aligned}$$

Notice that \((\pi ,l)\in {{\mathcal {U}}}\) is locally bounded, and hence

$$\begin{aligned} |\tilde{\pi }(s\wedge \tau _a)|+|\tilde{l}(s\wedge \tau _a)|\le C_1(\pi ,l,a,T),\quad s\in [t,T]. \end{aligned}$$

The positive constant \(C_1\) depends on \((\pi ,l)\), a and T only. Since \(|\Delta M|\vee |\Delta N|\le 1\), it follows that

$$\begin{aligned} |X^{\pi ,l}(s\wedge \tau _a)|\le C_2(\pi ,l,a,T),\quad s\in [t,T], \end{aligned}$$

where \(C_2\) is a positive constant which depends on \((\pi ,l)\), a and T only. This implies that \(M^{\pi ,l}(\cdot \wedge \tau _a)\) is a \( \mathbb {P} \)-martingale. Hence, it holds that

$$\begin{aligned} \mathbb {E} _{t,x,i,z}\left[ U(X^{\pi ,l}(T\wedge \tau _a)\big )\right]&\le x^\gamma \varphi (t,i,z)+ \mathbb {E} _{t,x,i,z}\left[ {M}^{\pi ,l}(T\wedge \tau _a)-{M}^{\pi ,l}(t)\right] \nonumber \\&=x^\gamma \varphi (t,i,z), \end{aligned}$$
(A.13)

where we set \( \mathbb {E} _{t,i,z}[\cdot ]:= \mathbb {E} [\cdot \mid X^{\pi ,l}(t)=x,Y(t)=i,Z(t)=z]\) for \((t,x,i,z)\in [0,T]\times \mathbb {R}_+\times \{1,\ldots ,m\}\times {{\mathcal {S}}}\). It follows from Fatou’s lemma that

$$\begin{aligned} \mathbb {E} _{t,x,i,z}[U(X^{\pi ,l}(T))]\le \varliminf _{a\rightarrow \infty } \mathbb {E} _{t,x,i,z}[U(X^{\pi ,l}(T\wedge \tau _a))]\le x^\gamma \varphi (t,i,z). \end{aligned}$$

This verifies the validity of the conclusion (i).

We next prove the conclusion (ii). In fact, recall that the optimal feedback strategy \((\pi ^{*},l^*)=(\pi ^*(t,i,z),l^*(t,i,z))\) for \(i=1,\ldots ,m\) is given by (23) for \(k=n\) and given by (38) for \(k=0,1,\ldots ,n-1\). Then, there exists a constant \(C>0\) which is independent of (tiz) such that \(\Vert \pi ^*(t,i,z)\Vert ^2+|l^*(t,i,z)|^{2}\le C\) for all \((t,i,z)\in [0,T]\times \{1,\ldots ,m\}\times {{\mathcal {S}}}\). We next estimate \( \mathbb {E} [(X^{\pi ^*,l^*}(T\wedge \tau _a))^{2\gamma }]\). First of all, the dynamics of the wealth process can be rewritten as, for \(s\in [t,T]\),

$$\begin{aligned} dX^{\pi ^*,l^*}(s)=&X^{\pi ^*,l^*}(s)[r(Y_{s})+\pi ^{*}(s,Y(s),Z(s))^{\top }\theta (Y(s),Z(s))\nonumber \\&\qquad +l^*(s,Y(s),Z(s))(p(Y(s),Z(s))-c(Y(s)))]ds\nonumber \\&\qquad +X^{\pi ^*,l^*}(s)[\pi ^{*}(s,Y(s),Z(s))^{\top }\sigma (Y(s))-l^*(s,Y(s),Z(s))\phi (Y(s))]dW(s)\nonumber \\&\qquad -X^{\pi ^*,l^*}(s)l^*(s,Y(s),Z(s))\bar{\phi }(Y(s))d\bar{W}(s)\nonumber \\&\qquad -X^{\pi ^*,l^*}(s-)\pi ^{*}(s,Y(s),Z(s))^{\top }dZ(s)\nonumber \\&\qquad \quad -l^*(s-,Y(s-),Z(s-))X^{\pi ^*,l^*}(s-)g(Y(s-))dN(s). \end{aligned}$$

Then, Itô’s formula yields that for \(u\in [t,T]\),

$$\begin{aligned} (X^{\pi ^*,l^*}(u))^{2\gamma }=&(X^{\pi ^*,l^*}(t))^{2\gamma }+\tilde{M}^{\pi ^*,l^*}(u)-\tilde{M}^{\pi ^*,l^*}(t)\nonumber \\&+\int _t^u(X^{\pi ^*,l^*}(s))^{2\gamma }\tilde{\mathcal{A}}(\pi ^*(s,Y(s),Z(s)),l^*(s,Y(s),Z(s));Y(s),Z(s))ds. \end{aligned}$$

Here, for \((\pi ,l)\in (-\infty ,1]^n\times [0,\infty )\) and \((i,z)\in \{1,\ldots ,m\}\times {{\mathcal {S}}}\),

$$\begin{aligned} \tilde{{\mathcal {A}}}(\pi ,l;i,z)=&{2\gamma }\big [r(i)+\pi ^{\top }(I-diag(z))\theta (i,z)+(p(i,z)-c(i))l\big ]\\&+{\gamma }({2\gamma }-1)\big [\pi ^{\top }(I-diag(z))\sigma (i)\sigma (i)^{\top }(I-diag(z))\pi +l^2\big (\phi (i)\phi (i)^{\top }\nonumber \\&\quad +\bar{\phi }(i)\bar{\phi }(i)^{\top }\big )\\&-2l\pi ^{\top }(I-diag(z))\sigma (i)\phi (i)^{\top }\big ]+[(1-lg(i))^{2\gamma }-1]\nu (i,z)\\&+\sum _{j=1}^n[(1-\pi _j)^{2\gamma }-1](1-z_j)h_j(i,z). \end{aligned}$$

The \( \mathbb {P} \)-(local) martingale is given by, for \(t\in [0,T]\),

$$\begin{aligned} \tilde{M}^{\pi ^*,l^*}(t):=&\int _0^t(X^{\pi ^*,l^*}(s))^{2\gamma }[\pi ^{*}(s,Y(s),Z(s))^{\top }\sigma (Y(s))-l^*(s,Y(s),Z(s))\phi (Y(s))]dW(s)\nonumber \\&-\int _0^t(X^{\pi ^*,l^*}(s))^{2\gamma }l^*(s,Y(s),Z(s))\bar{\phi }(Y(s))d\bar{W}(s)\nonumber \\&+\sum _{j=1}^n\int _0^t(X^{\pi ^*,l^*}(s-))^{2\gamma }[(1-\pi _j^*(s,Y(s-),Z(s-)))^{2\gamma }-1]dM_j(s)\nonumber \\&+\int _0^t(X^{\pi ^*,l^*}(s-))^{2\gamma }[(1-lg(Y(s-)))^{2\gamma }-1]d\tilde{N}(s). \end{aligned}$$

As above, we have that \(\Vert \pi ^*(t,i,z)\Vert ^2+|l^*(t,i,z)|^{2}\le C\) for all \((t,i,z)\in [0,T]\times \{1,\ldots ,m\}\times {{\mathcal {S}}}\), and hence

$$\begin{aligned} |(1-l^*(t,i,z)g(i))^{2\gamma }-1|\le (1+\gamma )(|l^*(t,i,z)g(i)|^2+|l^*(t,i,z)g(i)|). \end{aligned}$$

Then, there exists a constant \(C>0\) such that for all \((t,i,z)\in [0,T]\times \{1,\ldots ,m\}\times {{\mathcal {S}}}\),

$$\begin{aligned} |\tilde{{\mathcal {A}}}(\pi ^*(t,i,z),l^*(t,i,z),i,z)|\le C. \end{aligned}$$

Thus, we have that for all \(t\in [0,T]\),

$$\begin{aligned}&\mathbb {E} _{t,x,i,z}\big [(X^{\pi ^*,l^*}(T\wedge \tau _a))^{2\gamma }\big ] =x^{2\gamma }\nonumber \\&\qquad + \mathbb {E} _{t,x,i,z}\left[ \int _t^{T\wedge \tau _a}(X^{\pi ^*,l^*}(s))^{2\gamma }\tilde{{\mathcal {A}}}(\pi ^*(s,Y(s),Z(s)),l^*(s,Y(s),Z(s));Y(s),Z(s))ds\right] \\&\quad \le x^{2\gamma }+ \mathbb {E} _{t,x,i,z}\left[ \int _t^{T}(X^{\pi ^*,l^*}(s\wedge \tau _a))^{2\gamma }\big |\tilde{{\mathcal {A}}}(\pi ^*(s,Y(s),Z(s)),l^*(s,Y(s);Z(s)),Y(s),Z(s))\big |ds\right] \\&\quad \le x^{2\gamma }+C\int _t^{T} \mathbb {E} _{t,x,i,z}[(X^{\pi ^*,l^*}(s\wedge \tau _a))^{2\gamma }]ds. \end{aligned}$$

The Gronwall’s inequality yields that

$$\begin{aligned} \sup _{a\in \mathbb {R}_+} \mathbb {E} _{t,x,i,z}\big [(X^{\pi ^*,l^*}(T\wedge \tau _a))^{2\gamma }\big ]\le x^{2\gamma }e^{CT}, \end{aligned}$$

and hence \(\{(X^{\pi ^*,l^*}(T\wedge \tau _a))^{\gamma }\}_{a\in \mathbb {R}_+}\) is uniformly integrable. This yields that

$$\begin{aligned} V(t,x,i,z)= \mathbb {E} _{t,x,i,z}[U(X^{\pi ^*,l^*}(T))]=\lim _{a\rightarrow \infty } \mathbb {E} _{t,x,i,z}[U(X^{\pi ^*,l^*}(T\wedge \tau _a))]= x^\gamma \varphi (t,i,z). \end{aligned}$$

This verifies the validity of the conclusion (ii). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bo, L., Liao, H. & Wang, Y. Optimal credit investment and risk control for an insurer with regime-switching. Math Finan Econ 13, 147–172 (2019). https://doi.org/10.1007/s11579-018-0222-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11579-018-0222-7

Keywords

Mathematics Subject Classifications

Navigation