Skip to main content
Log in

Optimal excess-of-loss reinsurance and investment polices under the CEV model

  • Published:
Annals of Operations Research Aims and scope Submit manuscript

Abstract

This paper focuses on risk control problem of the insurance company in enterprise risk management. The insurer manages its financial risk through purchasing excess-of-loss reinsurance, and investing its wealth in the constant elasticity of variance stock market. We model risk process by Brownian motion with drift, and study the optimization problem of maximizing the exponential utility of terminal wealth under the controls of reinsurance and investment. Using stochastic control theory, we obtain explicit expressions for optimal polices and value function. We also show that the optimal excess-of-loss reinsurance is always better than optimal proportional reinsurance. And some numerical examples are given.

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
Fig. 6
Fig. 7

Similar content being viewed by others

References

  • Asmussen, S., HØjgaard, B., & Taksar, M. (2000). Optimal risk control and dividend distribution policies: Example of excess-of-loss reinsurance for an insurance for an insurance corporation. Finance and Stochastics, 4, 299–324.

    Article  Google Scholar 

  • Bai, L., & Guo, J. (2008). Optimal proportional reinsurance and investment with multiple risky assets and no-shorting constraint. Insurance: Mathematics and Economics, 42, 968–975.

  • Bai, L., & Guo, J. (2010). Optimal dynamic excess-of -loss reinsurance and multidimensional portfolio selection. Science China-Mathematics, 53(7), 1787–1804.

    Article  Google Scholar 

  • Blome, C., & Schoenherr, T. (2011). Supply chain risk management in finical crises-a multiple case-study approach. Int. J. Production Economics, 134, 43–57.

    Article  Google Scholar 

  • Browne, S. (1995). Optimal investment policies for a firm with random risk process: Exponential utility and minimizing the probability of ruin. Mathematics of Operations Research, 20(4), 937–958.

    Article  Google Scholar 

  • Chacko, G., & Viceira, L. (2005). Dynamic consumption and portfolio choice with stochastic volatility in incomplete markets. The Review of Financial Studies, 18(4), 1369–1402.

    Article  Google Scholar 

  • Consigli, G., & Dempster, M. (1998). Dynamic stochastic programming for asset-liability management. Annals of Operations Research, 81, 131–162.

    Article  Google Scholar 

  • COSO (2004). Enterprise Risk Management-Integrated Framework Executive Summary. Committee of Sponsoring Organizations of the Treadway Commission.

  • Fleming, W., & Soner, H. (2005). Controlled Markov process an viscosity solutions (2nd ed.). New York: Springer.

    Google Scholar 

  • Gerber, H. U. (1979). An introduction to mathematical risk theory (Vol. 8). Philadelphia: SS Huebner Foundation for Insurance Education, Wharton School, University of Pennsylvania.

  • Grandell, J. (1991). Aspects of risk theory. New York: Springer-Verlag.

    Book  Google Scholar 

  • Gu, M., Yang, P., Li, S., & Zhang, J. (2010). Constant elasticity of variance model for proportional reinsurance and investment strategies. Insurance: Mathematics and Economics, 46, 580–587.

    Google Scholar 

  • Hainaut, D. (2009). Dynamic asset allocation under VaR constraint with stochastic interest. Annals of Operations Research, 172(1), 97–117.

    Article  Google Scholar 

  • Irgens, C., & Paulsen, J. (2004). Optimal control of risk exposure, reinsurance and investment for insurance portfolios. Insurance: Mathematics and Economics, 35, 21–51.

    Google Scholar 

  • Liang, Z., & Guo, J. (2007). Optimal proportional reinsurance and ruin probability. Stochastic Models, 23(2), 333–350.

    Article  Google Scholar 

  • Liang, Z., Yuen, K. C., & Cheung, K. C. (2012). Optimal reinsurance-investment problem in constant elasticity of variance stock market for jump-diffusion risk model. Applied Stochastic Models in Business and Industry, 28, 585–597.

    Article  Google Scholar 

  • Luo, S., Taksar, M., & Tsoi, A. (2008). On reinsurance and investment for large insurance portfolios. Insurance: Mathematics and Economics, 42, 434–444.

    Google Scholar 

  • Meng, H., & Zhang, X. (2010). Optimal risk control for the excess of loss reinsurance policies. Astin Bulletins, 40(1), 179–197.

    Article  Google Scholar 

  • Merton, R. C. (1969). Life portfolio selection under uncertainty: the continuous-time case. The Review of Economics and Statistics, 51(3), 247–257.

    Article  Google Scholar 

  • Schroder, M. (1989). Computing the constant elasticity of variance option pricing formula. The Journal of Finance, XLIV(1), 211–219.

    Article  Google Scholar 

  • Schmidli, H. (2001). Optimal proportional reinsurance policies in a dynamic setting. Scandinavian Actuarial Journal, 1, 55–68.

    Article  Google Scholar 

  • Olson, D. L., & Wu, D. D. (2008). Enterprise risk management. Singapore: World Scientific Publishing Co., Ltd.

    Google Scholar 

  • Promislow, S. D., & Young, V. R. (2005). Minimizing the probability of ruin when claims follow Brownian motion with drift. North American Actuarial Journal, 9(3), 109–128.

    Google Scholar 

  • Taksar, M., & Markussen, C. (2003). Optimal dynamic reinsurance policies for large insurance portfolios. Finance and Stochastic, 7, 97–121.

    Article  Google Scholar 

  • Wu, D. D., Chen, S. H., & Olson, D. L. (2014). Business intelligence in risk management: Some recent progress. Information Sciences, 256, 1–7.

    Article  Google Scholar 

  • Wu, D. D., & Olson, D. L. (2009). Enterprise risk management: small business scorecard analysis. Production Planning and Control, 20(4), 362–369.

    Article  Google Scholar 

  • Wu, D. D., & Olson, D. L. (2010a). Enterprise risk management: Coping with model risk in a large bank. Journal of the Operational Research Society, 61(2), 179–190.

    Article  Google Scholar 

  • Wu, D. D., & Olson, D. L. (2010b). Enterprise risk management: A DEA VaR approach in vendor selection. International Journal of Production Research, 48(16), 4919–4932.

    Article  Google Scholar 

  • Wu, D. D., & Olson, D. L. (2013). Computational simulation and risk analysis: An introduction of state of the art research. Mathematical and Computer Modeling, 58, 1581–1587.

    Article  Google Scholar 

  • Wu, D. D., Olson, D. L., & Birge, J. R. (2011). Introduction to special issue on “Enterprise risk management in operations”. International Journal of Production Economics, 134(1), 1–2.

    Article  Google Scholar 

Download references

Acknowledgments

This research was supported by the Natural Science Foundation of Jiangsu Higher Education Institution of China (No. 13KJD110006), SJTU Special Funds for Cross Deciplines (No. 10JCY11) and the National Natural Science Foundation of China (No. 11101215).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Qicai Li.

Appendix

Appendix

In this appendix, we prove Theorem 4.2.

Proof

Step 1: Propose the ansatz. Mimicking the method of Browne (1995, p.942) or Liang et al. (2012, p.588), we want to find a solution to (4.2) with the form

$$\begin{aligned} V(t,x,s)={\uplambda }_{0} -\frac{\gamma }{m}\exp \left[ -mx\mathrm{{e}}^{r(T-t)}+G(t,s)\right] , \end{aligned}$$
(7.1)

where \(G(t,s)\) is a suitable function to be determined. And, the boundary condition \(V(T,x,s)=u(x)\) implies that

$$\begin{aligned} G(T,s)=0. \end{aligned}$$
(7.2)

Let \(G_t , G_s , G_{ss}\), be the partial derivatives of \(G(t,s)\). Note that

$$\begin{aligned} \left\{ {{\begin{array}{l} {V_t =[V-\uplambda _0 ][mxr\mathrm{{e}}^{r(T-t)}+G_t ]} \\ {V_x =[V-\uplambda _0 ][-m\mathrm{{e}}^{r(T-t)}]} \\ {V_s =[V-\uplambda _0 ][G_s ]} \\ {V_{xx} =[V-\uplambda _0 ][m^{2}\mathrm{{e}}^{2r(T-t)}]} \\ {V_{ss} =[V-\uplambda _0 ][G_s^2 +G_{ss} ]} \\ {V_{xs} =[V-\uplambda _0 ][-m\mathrm{{e}}^{r(T-t)}G_s ]} \\ \end{array} }} \right. , \end{aligned}$$
(7.3)

Step 2: Derive the optimal control. Substituting (7.3) back into the HJB Eq. (4.2), since \(V-\uplambda _0 <0\), we get

$$\begin{aligned}&G_t -m\mu _\infty (\eta -\theta )\mathrm{{e}}^{r(T-t)}+\mu sG_s +\frac{1}{2}\sigma ^{2}s^{2\beta +2}(G_s^2 +G_{ss} )\nonumber \\&\quad +\mathop {\inf }\limits _{\pi \in R} \left\{ \left[ -(\mu -r)\pi -\pi \sigma ^{2}s^{2\beta +1}G_s +\frac{1}{2}\sigma ^{2}\pi ^{2}s^{2\beta }m\mathrm{{e}} ^{r(T-t)}\right] m\mathrm{{e}}^{r(T-t)}\right\} \nonumber \\&\quad +\mathop {\inf }\limits _{a\in [0,N]} \left\{ \left[ -\mu (a)\theta +\frac{1}{2}\sigma ^{2}(a)m\mathrm{{e}}^{r(T-t)}\right] m\mathrm{{e}}^{r(T-t)}\right\} =0. \end{aligned}$$
(7.4)

Let

$$\begin{aligned} f_1 (\pi ,t)=\left[ -(\mu -r)\pi -\pi \sigma ^{2}s^{2\beta +1}G_s +\frac{1}{2}\sigma ^{2}\pi ^{2}s^{2\beta }m\mathrm{{e}}^{r(T-t)}\right] m\mathrm{{e}}^{r(T-t)}, \end{aligned}$$
(7.5)

and

$$\begin{aligned} f_2 (a,t)=\left[ -\mu (a)\theta +\frac{1}{2}\sigma ^{2}(a)m\mathrm{{e}}^{r(T-t)}\right] m\mathrm{{e}}^{r(T-t)}. \end{aligned}$$
(7.6)

Differentiating \(f_1 (\pi ,t)\)with respect to \(\pi \) yields the minimizer

$$\begin{aligned} \pi ^{*}(t)=\frac{(\mu -r)+\sigma ^{2}s^{2\beta +1}G_s }{\sigma ^{2}s^{2\beta }}\frac{\mathrm{{e}}^{-r(T-t)}}{m}, \end{aligned}$$
(7.7)

and the value of \(f_1 (\pi ,t)\) at this minimizer is

$$\begin{aligned} f_1 (\pi ^{*},t)=-\frac{1}{2}\frac{\left[ (\mu -r)+\sigma ^{2}s^{2\beta +1}G_s \right] ^{2}}{\sigma ^{2}s^{2\beta }}. \end{aligned}$$
(7.8)

Similarly, from the first order condition

$$\begin{aligned} \frac{\partial f_2 (a,t)}{\partial a}=\left[ -\theta \bar{{F}}(a)+a\bar{{F}}(a)m\mathrm{{e}}^{r(T-t)}\right] m\mathrm{{e}}^{r(T-t)}=0, \end{aligned}$$

we know that, without restriction with respect to \(a\),

$$\begin{aligned} \tilde{a}(t)=\frac{\theta \mathrm{{e}}^{-r(T-t)}}{m}, \end{aligned}$$
(7.9)

which leads to

$$\begin{aligned} f_2 (\tilde{a},t)=-\theta m\mathrm{{e}}^{r(T-t)}\int \limits _0^{{\theta \mathrm{{e}}^{-r(T-t)}}/m} {\bar{{F}}(y)dy} +m^{2}\mathrm{{e}}^{2r(T-t)}\int \limits _0^{{\theta \mathrm{{e}}^{-r(T-t)}}/m} {y\bar{{F}}(y)dy} . \end{aligned}$$
(7.10)

Step 3: Separate the variables. We need to discuss the two cases according to the value of \(\tilde{a}(t)\).

Case 1 \(({ {Nm>\theta }})\) If \(t<\bar{{T}}{:=}T+\frac{\ln (Nm)-\ln \theta }{r}\), then \(\tilde{a}(t)\in [0,\;N)\). So,

$$\begin{aligned} (\pi ^{*}(t),a^{*}(t))=\left( {\frac{(\mu -r)+\sigma ^{2}s^{2\beta +1}G_s }{\sigma ^{2}s^{2\beta }}\frac{\mathrm{{e}}^{-r(T-t)}}{m},\frac{\theta \mathrm{{e}}^{-r(T-t)}}{m}} \right) \end{aligned}$$
(7.11)

coincides with the optimal policy. Since \(T<\bar{{T}}, \left( {\pi ^{*}(t),a^{*}(t)} \right) \) is optimal policy on \(\left[ {0,\;T} \right] \).

Up to now, we still need to solve \(G(t,s)\) to find the optimal investment \(\pi ^{*}(t)\) and the value function \(V(t,x,s)\) in this case. Substituting \(\left( {\pi ^{*}(t),a^{*}(t)} \right) \) (i.e. expression (7.11)) back to (7.4), we can get

$$\begin{aligned} G_t -m\mu _\infty (\eta -\theta )\mathrm{{e}}^{r(T-t)}+rsG_s +\frac{1}{2}\sigma ^{2}s^{2\beta +2}G_{ss} -\frac{(\mu -r)^{2}}{2\sigma ^{2}s^{2\beta }}+f_2 (a^{*},t)=0. \end{aligned}$$
(7.12)

We appeal to power transformation technique and variable change method to solve the problem.

Let

$$\begin{aligned} G(t,s)=h(t,y),\;\;y=s^{-2\beta }, \end{aligned}$$
(7.13)

with boundary condition

$$\begin{aligned} h(T,y)=0, \end{aligned}$$
(7.14)

and

$$\begin{aligned} \left\{ {{\begin{array}{l} {G_t =h_t}\\ {\;G_s =-2\beta s^{-2\beta -1}h_y}\\ {G_{ss} =2\beta (2\beta +1)s^{-2\beta -2}h_y +4\beta ^{2}s^{-4\beta -2}h_{yy}} \\ \end{array} }} \right. , \end{aligned}$$
(7.15)

where \(h_t , h_y , h_{yy}\) are partial derivatives of \(h(t,y)\).

Putting the partial derivatives of \(G(t,s)\) into the Eq. (7.12), we obtain equation

$$\begin{aligned} h_t +\left[ \sigma ^{2}(2\beta +1)-2ry\right] \beta h_y +2\sigma ^{2}\beta ^{2}yh_{yy} -\frac{(\mu -r)^{2}}{2\sigma ^{2}}y+M_1 (t)=0, \end{aligned}$$
(7.16)

where \(M_1 (t)=-m\mu _\infty (\eta -\theta )\mathrm{{e}}^{r(T-t)}+f_2 (a^{*},t)\). We try to find a solution of the above equation with the following form

$$\begin{aligned} h(t,y)=K_1 (t)+L_1 (t)y, \end{aligned}$$
(7.17)

with boundary condition

$$\begin{aligned} \left\{ {{\begin{array}{l} {K_1 (T)=0} \\ {L_1 (T)=0\;} \\ \end{array} }} \right. , \end{aligned}$$
(7.18)

and

$$\begin{aligned} h_t =K_1 ^{\prime }+L_1 ^{\prime }y,\;\;h_y =L_1 ,\;\;h_{yy} =0, \end{aligned}$$
(7.19)

where \(K_1^{\prime }, L_1^{\prime }\) are the derivatives of \(K_1 (t),\;L_1 (t)\) respectively. Putting (7.19) into (7.16), we derive

$$\begin{aligned} {K}'_1 +\sigma ^{2}\beta (2\beta +1)L_1 +\left[ {L}'_1 -2rL_1 -\frac{(\mu -r)^{2}}{2\sigma ^{2}}\right] y+M_1 (t)=0. \end{aligned}$$
(7.20)

By matching coefficients, we have

$$\begin{aligned} \left\{ {{\begin{array}{l} {{K}'_1 +\sigma ^{2}\beta (2\beta +1)L_1 +M_1 (t)=0} \\ {{L}'_1 -2rL_1 -\frac{(\mu -r)^{2}}{2\sigma ^{2}}=0} \\ \end{array} }} \right. , \end{aligned}$$
(7.21)

Taking into boundary condition, we have the solution of Eq. (7.21):

$$\begin{aligned} L_1 (t)&= -\frac{(\mu -r)^{2}}{4r\sigma ^{2}}(1-\mathrm{{e}}^{-2r(T-t)}), \end{aligned}$$
(7.22)
$$\begin{aligned} K_1 (t)&= \int \limits _t^T {\sigma ^{2}\beta (2\beta +1)L_1 (z)+M_1 (z)dz}\nonumber \\&= -\frac{\beta (2\beta +1)(\mu -r)^{2}}{4r}\left[ (T-t)-\frac{1-\mathrm{{e}}^{-2r(T-t)}}{2r}\right] \nonumber \\&\quad -\,\frac{m\mu _\infty (\eta -\theta )}{r}(\mathrm{{e}}^{r(T-t)}-1)\nonumber \\&\quad +\,\int \limits _t^T \left\{ -\theta m\mathrm{{e}}^{r(T-z)}\int \limits _0^{{\theta \mathrm{{e}}^{-r(T-z)}}/m} {\bar{{F}}(y)dy} +m^{2} \mathrm{{e}}^{2r(T-z)}\int \limits _0^{{\theta \mathrm{{e}}^{-r(T-z)}}/m} {y\bar{{F}}(y)dy} \right\} dz.\nonumber \\ \end{aligned}$$
(7.23)

Putting these parameters into \(G(t,s)\), we obtain

$$\begin{aligned} G(t,s)=K_1 (t)+L_1 (t)s^{-2\beta }. \end{aligned}$$
(7.24)

Therefore the optimal investment policy is

$$\begin{aligned} \pi ^{*}(t)=\frac{2r(\mu -r)+\beta (\mu -r)^{2}(1-\mathrm{{e}}^{-2r(T-t)})}{2r\sigma ^{2}s^{2\beta }}\frac{\mathrm{{e}}^{-r(T-t)}}{m}, \end{aligned}$$
(7.25)

and the corresponding value function has the form

$$\begin{aligned} V(t,x,s)={\uplambda }_{0} -\frac{\gamma }{m}\exp (-mx\mathrm{{e}}^{r(T-t)}+K_1 (t)+L_1 (t)s^{-2\beta }), \end{aligned}$$
(7.26)

where \(L_1 (t)\) and \(K_1 (t)\) is determined by (7.22) and (7.23), respectively.

Case 2 \(({ {Nm\le \theta }})\) If \(t<\bar{{T}}{:=}T+\frac{\ln (Nm)-\ln \theta }{r}\) (noting that \(\bar{{T}}\le T)\), then \(\tilde{a}(t)\in [0,\;N)\) from expression (7.9). Similar to case 1, incorporating the constants of the calculations, we get the optimal value function

$$\begin{aligned} V(t,x,s)={\uplambda }_{0} -\frac{\gamma }{m}\exp (-mx\mathrm{{e}}^{r(T-t)}+K_1 (t)+L_1 (t)s^{-2\beta }+k), \end{aligned}$$
(7.27)

where the constant \(k\) will be determined from the following (7.37), and the optimal policies are

$$\begin{aligned} (\pi ^{*}(t),a^{*}(t))=\left( {\frac{(\mu -r)+\sigma ^{2}s^{2\beta +1}G_s }{\sigma ^{2}s^{2\beta }}\frac{\mathrm{{e}}^{-r(T-t)}}{m},\frac{\theta \mathrm{{e}}^{-r(T-t)}}{m}} \right) , \end{aligned}$$
(7.28)

where \(G(t,s)=K_1 (t)+L_1 (t)s^{-2\beta }+k.\) If \(\bar{{T}}\le t\le T\), then If \(\tilde{a}(t)\ge N\). We get that the optimal retention level is \(a^{*}(t)=N\). In this case, \(\mu (N)=\mu _\infty \), and \(\sigma ^{2}(N)=\sigma _\infty ^2\).

Putting the optimal policies

$$\begin{aligned} (\pi ^{*}(t),a^{*}(t))=\left( {\frac{(\mu -r)+\sigma ^{2}s^{2\beta +1}G_s }{\sigma ^{2}s^{2\beta }}\frac{\mathrm{{e}}^{-r(T-t)}}{m},N} \right) \end{aligned}$$
(7.29)

into the Eq. (7.4), we obtain

$$\begin{aligned} G_t -m\mu _\infty \eta \mathrm{{e}}^{r(T-t)}+rsG_s +\frac{1}{2}\sigma ^{2}s^{2\beta +2}G_{ss} -\frac{1}{2}\frac{(\mu -r)^{2}}{\sigma ^{2}s^{2\beta }}+\frac{1}{2}\sigma _\infty ^2 m^{2}\mathrm{{e}}^{2r(T-t)}=0.\nonumber \\ \end{aligned}$$
(7.30)

Again, we use the power transformation technique and variable change method to solve the Eq. (7.30) with the boundary condition (4.1).

Similarly, let \(G(t,s)=h(t,y),\;\;y=s^{-2\beta },\) we have

$$\begin{aligned}&h_t +[\sigma ^{2}(2\beta +1)-2ry]\beta h_y +2\sigma ^{2}\beta ^{2}yh_{yy} -\frac{(\mu -r)^{2}}{2\sigma ^{2}}y\nonumber \\&\quad -\,m\mu _\infty \eta \mathrm{{e}}^{r(T-t)}+\frac{1}{2}\sigma _\infty ^2 m^{2}\mathrm{{e}}^{2r(T-t)}=0. \end{aligned}$$
(7.31)

And we try to the following form and match coefficients,

$$\begin{aligned} h(t,y)=K_2 (t)+L_2 (t)y, \end{aligned}$$
(7.32)

Therefore, we get

$$\begin{aligned} L_2 (t)=-\frac{(\mu -r)^{2}}{4r\sigma ^{2}}(1-\mathrm{{e}}^{-2r(T-t)}), \end{aligned}$$

which is the same as the expression \(L_1 (t)\), denoted by \(L(t)\) for simplicity.

$$\begin{aligned} K_2 (t)&= -\frac{\beta (2\beta +1)(\mu -r)^{2}}{4r}[(T-t)-\frac{1-\mathrm{{e}}^{-2r(T-t)}}{2r}]\nonumber \\&\quad -\,\frac{m\mu _\infty \eta }{r}(\mathrm{{e}}^{r(T-t)}-1) -\frac{m\sigma _\infty ^2 }{4r}(1-\mathrm{{e}}^{2r(T-t)}), \end{aligned}$$
(7.33)

and

$$\begin{aligned} G(t,s)=K_2 (t)+L_2 (t)s^{-2\beta }. \end{aligned}$$
(7.34)

Thus, in case 2, the optimal excess-of-loss reinsurance and investment policies are

$$\begin{aligned} (\pi ^{*}(t),a^{*}(t))=\left\{ {{\begin{array}{l@{\quad }l} {\left( {\pi ^{*}(t),\frac{\theta \mathrm{{e}}^{-r(T-t)}}{m}} \right) ,}&{} {0\le t<\bar{{T}},} \\ {\left( {\pi ^{*}(t),N} \right) ,}&{} {\bar{{T}}\le t\le T,} \\ \end{array} }} \right. \end{aligned}$$
(7.35)

where \(\pi ^{*}(t)=\frac{2r(\mu -r)+\beta (\mu -r)^{2}(1-\mathrm{{e}}^{-2r(T-t)})}{2r\sigma ^{2}s^{2\beta }}\frac{\mathrm{{e}}^{-r(T-t)}}{m}\). And the corresponding value function has the form

$$\begin{aligned} V(t,x,s)=\left\{ {{\begin{array}{l@{\quad }l} {\uplambda _0 -\frac{\gamma }{m}\exp (-mx\mathrm{{e}}^{r(T-t)}+K_1 (t)+L(t)s^{-2\beta }+k),}&{} {0\le t<\bar{{T}},} \\ {\uplambda _0 -\frac{\gamma }{m}\exp (-mx\mathrm{{e}}^{r(T-t)}+K_2 (t)+L(t)s^{-2\beta }),}&{} {\bar{{T}}\le t<T,} \\ \end{array} }} \right. \end{aligned}$$
(7.36)

where choose \(k\) in the way that \(V(t,x,s)\) given by (7.37) is continuous at \(\bar{{T}}\), that is,

$$\begin{aligned} k=K_2 (\bar{{T}})-K_1 (\bar{{T}}). \end{aligned}$$
(7.37)

Thus, we complete the proof. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, Q., Gu, M. & Liang, Z. Optimal excess-of-loss reinsurance and investment polices under the CEV model. Ann Oper Res 223, 273–290 (2014). https://doi.org/10.1007/s10479-014-1596-4

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10479-014-1596-4

Keywords

Mathematics Subject Classification

Navigation