Skip to main content
Log in

Finite Horizon Optimal Dividend and Reinsurance Problem Driven by a Jump-Diffusion Process with Controlled Jumps

  • Published:
Applied Mathematics & Optimization Aims and scope Submit manuscript

Abstract

In this paper, we discuss an optimal dividend and reinsurance problem for an insurance company facing two types of risks: unstable income and potential loses. The arrival of all loses is characterized as a compound Poisson process. We assumes that every possible loss can be reinsured for a part of it. The reserve is a combination of a diffusion process and a controllable compound Poisson process. We investigate the optimal dividend and reinsurance strategy by analyzing the corresponding variational inequality on the value function. A significant difference from the existing literature is that the HJB equation in this variational inequality is a partial integro-differential equation with a functional optimization problem appearing in the integral operator. We not only prove the existence of a classical solution to the problem and the continuity, strict monotonicity, boundedness of the dividend free boundary, but also discuss the properties of the optimal reinsurance policy, including the continuity, monotonicity of the optimal part covered by reinsurance for each possible loss, and the smoothness of the reinsurance free boundary.

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

Similar content being viewed by others

Notes

  1. The condition \(\gamma >0\) is a necessary condition for the concavity of the value function, which plays an important role in the proof of the existent and regularity of solution of the HJB equation, while the condition \(\lambda \delta <c\) guarantees the monotonicity of the optimal reinsurance function w.r.t x and the existence of the reinsurance free boundary. If both of the conditions are not true, then we can use the HJB equation to check that \(V\equiv x\), namely the company should pay out all its reserve as dividend and go bankrupt immediately, which is trivial. However, if only one of them is not true, we do not reach a complete conclusion.

References

  1. De Finetti, B.: un’impostazione alternativa della teoria collettiva del rischio. Trans. XVth Int. Congress Actuaries 2(1), 433–443 (1957)

    Google Scholar 

  2. Albrecher, H., Thonhauser, S.: Optimality results for dividend problems in insurance. Fisicas y Naturales 103(2), 295–320 (2009)

    MathSciNet  MATH  Google Scholar 

  3. Avanzi, B.: Strategies for dividend distribution: a review. North Am. Actuarial J. 13, 217–251 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  4. Wei, J., Yang, H., Wang, R.: Optimal reinsurance and dividend strategies under the Markov-modulated insurance risk model. Stoch. Anal. Appl. 28(6), 1078–1105 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  5. Taksar, M., Zhou, X.Y.: Optimal risk and dividend control for a company with a debt liability. Insurance 22(1), 105–122 (1998)

    MathSciNet  MATH  Google Scholar 

  6. Taksar, M.: Optimal risk and dividend distribution control models for an insurance company. Math. Methods Oper. Res. 51, 1–42 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  7. Højgaard, B., Taksar, M.: Optimal dynamic portfolio selection for a corporation with controllable risk and dividend distribution policy. Quant. Finance 4, 315–327 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  8. Guan, C., Chen, J.: A free boundary problem in variational inequality with gradient constraint. J. Each China Normal Univ. 1, 1–10 (2018)

    MathSciNet  MATH  Google Scholar 

  9. Guan, C., Yi, F.: A free boundary problem arising from a stochastic optimal control model with bounded dividend rate. Stoch. Anal. Appl. 32(5), 742–760 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  10. Guan, C., Yi, F.: A free boundary problem arising from a stochastic optimal control model under controllable risk. J. Differ. Equ. 260, 4845–4870 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  11. Guan, C., Yi, F., Chen, X.: A fully nonlinear free boundary problem arising from optimal dividend and risk control model. Math. Control Relat. Fields 9(3), 425–452 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  12. De Angelis, T., Ekstrom, E.: The dividend problem with a finite horizon. Ann. Appl. Probab. 27, 3525–3546 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  13. Belhaj, M.: Optimal dividend payments when cash reserves follow a jump-diffusion process. Math. Finance 20(2), 313–325 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  14. Hunting, M., Paulsen, J.: Optimal dividend policies with transaction costs for a class of jump-diffusion processes. Finance Stoch. 17(1), 73–106 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  15. Kyprianou, A.E., Loeffen, R., Pérez, J.L.: Optimal control with absolutely continuous strategies for spectrally negative Lévy processes. J. Appl. Probab. 49(1), 150–166 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  16. Kelbert, M., Moreno-Franco, H.A.: HJB equations with gradient constraint associated with controlled jump-diffusion processes. SIAM J. Control Optim 57(3), 2185–2213 (2019)

    Article  MathSciNet  MATH  Google Scholar 

  17. Jiang, Z.: Optimal dividend policy when cash reserves follow a jump-diffusion process under Markov-regime switching. J. Appl. Probab. 52(1), 209–223 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  18. Feng, L.L., Song, J.H., Li, X.: Dividend maximization when cash reserves follow a jump-diffusion process. Commun. Math. Res. 25(2), 143–158 (2009)

    MathSciNet  MATH  Google Scholar 

  19. Moreno-Franco, H.: Solution to HJB equations with an elliptic integro-differential operator and gradient constraint. Appl. Math. Optim. 78(1), 25–60 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  20. Avram, F., Palmowski, Z., Pistorius, M.: On the optimal dividend problem for a spectrally negative Levy process. Ann. Appl. Prob. 17(1), 156–180 (2007)

    Article  MATH  Google Scholar 

  21. Li, Y., Li, Z., Wang, S., Xu, Z.Q.: Dividend optimization for jump-diffusion model with solvency constraints. Oper. Res. Lett. 48, 170–175 (2020)

    Article  MathSciNet  MATH  Google Scholar 

  22. Paulsen, Y.J.: Optimal dividend payouts for diffusions with solvency constraints. Finance Stoch. 7, 457–473 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  23. He, L., Liang, Z.: Optimal dividend control of the insurance company with proportional reinsurance policy under solvency constraints. Insurance 42, 976–983 (2008)

    MATH  Google Scholar 

  24. Rochet, J., Villeneuve, S.: Liquidity management and corporate demand for hedging and insurance. J. Financ. Intermediation 20, 303–323 (2011)

    Article  Google Scholar 

  25. Chen, X., Guan, C., Yi, F.: A free boundary problem of liquidity management for optimal dividend and insurance in finite horizon. SIAM J. Control Optim. 59(4), 2524–2545 (2021)

    Article  MathSciNet  MATH  Google Scholar 

  26. Yong, J., Zhou, X.: Stochastic Ccontrols: Hamiltonian Systems and HJB Equations. Applications of Mathematics (New York), Springer, New York (1999)

    Book  Google Scholar 

  27. Protter, P.: Stochastic Integration and Differential Equations. Springer, New York (1990)

    Book  MATH  Google Scholar 

  28. Pham, H.: Continuous-Time Stochastic Control and Optimization with Financial Applications. Springer, Berlin (2009)

    Book  MATH  Google Scholar 

  29. Lieberman, G.M.: Second Order Parabolic Differential Equations. World Scientific, Singapore (1996)

    Book  MATH  Google Scholar 

  30. Friedman, A.: Parabolic variational inequalities in one space dimension and smoothness of the free boundary. J. Funct. Anal. 18, 151–176 (1975)

    Article  MathSciNet  MATH  Google Scholar 

  31. Evans, L.C.: Partial Differential Equations. Intersxcience Publishers, Geneva (2016)

    Google Scholar 

  32. Shreve, S.E.: Stochastic Calculus for Finance II. Springer, New Yorkd (2004)

    Book  MATH  Google Scholar 

  33. Taksar, M.: Skorohod problems with nonsmooth boundary conditions. J. Comput. Appl. Math. 40(2), 233–251 (1992)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Funding

The authors have not disclosed any funding.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Chonghu Guan.

Ethics declarations

Conflict of interest

The authors have not disclosed any competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

This author is partially supported by NSF of Guangdong Province of China (Nos. 2021A1515012031, 2022A1515010263), NNSF of China (No. 11901244).

Appendices

Appendix A Proof of Theorem 3.2

In this section, we prove Theorem 3.2. Suppose V is the value function defined in (2.5) and v is the solution of the problem (3.1), denote \({\widehat{V}}(x,t)=v(x,T-t)\). We come to prove \({\widehat{V}}(x,t)= V (x,t)\).

For any \((x,t)\in {\mathbb R}^+\times [0,T]\), and any admissible dividend control process \({\mathbb L}^t=\{L_s\}_{s\geqslant t-}\) and retained loss control process \({\mathbb H}^t=\{H_s\}_{s\geqslant t-}\). Suppose \(\{X_s\}_{s\geqslant t-}\) is the solution of (2.2) and \(\tau =\inf \{s\geqslant t \mid X_s\le 0\}\). Denote \(L_s^c\) as the continuous part of \(L_s\) and denote \(\Delta L_s=L_s-L_{s-}\) as the jump part of \(L_s\). The general Itô’s formula (see [27] Theorem 33 on page 81) and taking expected value gives

$$\begin{aligned} {\widehat{V}}(x,t)&= {\mathbb E}_{t,x}\Bigg [e^{-c (\tau \wedge T-t)}{\widehat{V}}(X_{\tau \wedge T},\tau \wedge T)\Bigg ]\nonumber \\&\qquad -{\mathbb E}_{t,x}\Bigg [\sum _{t\leqslant s{\leqslant }\tau \wedge T}e^{-c (s-t)}\left( {\widehat{V}}(X_s,s)-{\widehat{V}}(X_{s-},s)\right) \Bigg ] \nonumber \\&\qquad -{\mathbb E}_{t,x}\Bigg [\int _{ t }^{\tau \wedge T}e^{-c (s-t)}\Big ({\widehat{V}}_t+\frac{\sigma ^2}{2}{\widehat{V}}_{xx}+\Big (\gamma +\lambda (1+\delta ){\mathbb E}_{s,X_{s-}}[H_s(Z_1)]\Big ) \nonumber \\&\quad {\widehat{V}}_x -c {\widehat{V}}\Big )(X_{s-},s) \textrm{d}s\Bigg ]\nonumber \\&\qquad +{\mathbb E}_{t,x}\Bigg [\int _{ t }^{\tau \wedge T}e^{-c (s-t)} {\widehat{V}}_x (X_{s-},s)\textrm{d}L^c_s\Bigg ] \nonumber \\&\qquad -{\mathbb E}_{t,x}\Bigg [\int _{ t }^{\tau \wedge T}e^{-c (s-t)}\sigma {\widehat{V}}_x (X_{s-},s) \textrm{d}W_s\Bigg ]. \end{aligned}$$
(A.1)

The boundedness of \({\widehat{V}}_x\) implies the last expectation above is zero. The jump part of \(X_s\), i.e. \(\Delta X_s=X_s-X_{s-}\), can be decomposed into \( \Delta X_s=-H_s(Z_1) \Delta N_s-\Delta L_s, \) where \(\Delta N_s=N_s-N_{s-}\), so \( X_s+\Delta L_s=X_{s-}-H_s(Z_1) \Delta N_s, \) then the second expectation on the right hand side of (A.1) can be decomposed into

$$\begin{aligned}{} & {} {\mathbb E}_{t,x}\Bigg [\sum _{t\leqslant s{\leqslant } {\tau \wedge T}}e^{-c (s-t)}\left( {\widehat{V}}(X_s,s) -{\widehat{V}}(X_s+\Delta L_s,s)\right) \Bigg ]\\{} & {} \quad +{\mathbb E}_{t,x}\Bigg [\sum _{t\leqslant s{\leqslant } {\tau \wedge T}}e^{-c (s-t)}\left( {\widehat{V}}(X_{s-}-H_s(Z_1) \Delta N_s,s)-{\widehat{V}}(X_{s-},s)\right) \Bigg ], \end{aligned}$$

moreover, the above second term

$$\begin{aligned}&{\mathbb E}_{t,x}\Bigg [\sum _{t\leqslant s{\leqslant } {\tau \wedge T}}e^{-c (s-t)}\Big ({\widehat{V}}(X_{s-}-H_s(Z_1) \Delta N_s,s)-{\widehat{V}}(X_{s-},s)\Big )\Bigg ]\\&\quad ={\mathbb E}_{t,x}\Bigg [\sum _{t\leqslant s{\leqslant } {\tau \wedge T}}e^{-c (s-t)}\Big ({\widehat{V}}(X_{s-}-H_s(Z_1),s)-{\widehat{V}}(X_{s-},s)\Big )\Delta N_s\Bigg ]\\&\quad ={\mathbb E}_{t,x}\Bigg [\int _{ t }^{\tau \wedge T} e^{-c (s-t)}\Big ({\widehat{V}}(X_{s-}-H_s(Z_1) ,s)-{\widehat{V}}(X_{s-},s)\Big ) \lambda \textrm{d}s\Bigg ]\\&\quad ={\mathbb E}_{t,x}\Bigg [\int _{ t }^{\tau \wedge T} e^{-c (s-t)}\Big ({\mathbb E}_{s,X_{s-}}[{\widehat{V}}(X_{s-}-H_s(Z_1) ,s)]-{\widehat{V}}(X_{s-},s)\Big ) \lambda \textrm{d}s\Bigg ], \end{aligned}$$

where the second equality is followed from the compensated Poisson process \(\{N_s- \lambda s\}_{s\geqslant t-}\) is a martingale (see [32]), the last equality is followed from the double expectations formula. Therefore, (A.1) becomes

$$\begin{aligned} {\widehat{V}}(x,t)&={\mathbb E}_{t,x}\Bigg [e^{-c (\tau \wedge T-t)}{\widehat{V}}(X_{\tau \wedge T},\tau \wedge T)\Bigg ]\nonumber \\&\quad -{\mathbb E}_{t,x}\Bigg [\sum _{t\leqslant s{\leqslant } {\tau \wedge T}}e^{-c (s-t)}\left( {\widehat{V}}(X_{s},s) -{\widehat{V}}(X_s+\Delta L_s,s)\right) \Bigg ]\nonumber \\&\quad -{\mathbb E}_{t,x}\Bigg [\int _{ t }^{\tau \wedge T}e^{-c (s-t)}\Big ({\widehat{V}}_t+\frac{\sigma ^2}{2}{\widehat{V}}_{xx}+\gamma {\widehat{V}}_x -(c+\lambda ) {\widehat{V}}\Big )(X_{s-},s) \textrm{d}s\Bigg ]\nonumber \\&\quad -{\mathbb E}_{t,x}\Bigg [\int _{ t }^{\tau \wedge T}e^{-c (s-t)}\lambda \textrm{d}s \int _0^\infty \Big ((1+\delta )H_s(z){\widehat{V}}_x(X_{s-},s)\nonumber \\&\quad + {\widehat{V}}(X_{s-}-H_s(z) ,s)\Big )\textrm{d}F(z)\Bigg ]\nonumber \\&\quad +{\mathbb E}_{t,x}\Bigg [\int _{ t }^{\tau \wedge T}e^{-c (s-t)} {\widehat{V}}_x (X_{s-},s)\textrm{d}L^c_s\Bigg ] \end{aligned}$$
(A.2)

The first expectation is non-negative, the variational inequality in (3.1) implies the third expectation together with the fourth expectation is non-positive. Combining with \({\widehat{V}}_x \geqslant 1\) and \(\Delta L_s> 0\), we have

$$\begin{aligned} {\widehat{V}}(X_s+\Delta L_s,s)-{\widehat{V}}(X_{s},s)\geqslant \Delta L_s \end{aligned}$$

so

$$\begin{aligned} {\widehat{V}}(x,t)&\geqslant {\mathbb E}_{t,x}\Bigg [\sum _{t\leqslant s{\leqslant } {\tau \wedge T}}e^{-c (s-t)}\Delta L_s\Bigg ] +{\mathbb E}_{t,x}\Bigg [\int _{ t }^{\tau \wedge T}e^{-c (s-t)}\textrm{d}L^c_s\Bigg ] \\&= {\mathbb E}_{t,x}\Bigg [\int _{ t }^{\tau \wedge T}e^{-c (s-t)}\textrm{d}L_s\Bigg ]. \end{aligned}$$

Since \(({\mathbb H}^t,{\mathbb I}^t)\) is arbitrary, we obtain \({\widehat{V}}(x,t)\geqslant V(x,t)\).

We now prove the opposite inequality \({\widehat{V}}(x,t)\leqslant V(x,t)\) and \(\{(L^*_s,H^*_s)\}_{s\geqslant t-}\), defined in Theorem 3.2, is an optimal strategy. Set \(f(s)=d(T-s)\), then the dividend payout policy \(\{L^*_s\}_{s\geqslant t-}\) is the local time of the corresponding reserve \(X_s^*\) at the level f(s). By the Skorohod problem(see [33]). \(\{(X^*_s,L^*_s)\}_{s\geqslant t-}\) satisfies

$$\begin{aligned}&\textrm{d}X^*_s= (\gamma +\lambda (1+\delta ){\mathbb E}_{X_{s-}}[H^*_s(Z_1)])\textrm{d}s+\sigma \textrm{d}W_s -H^*_s(Z_1)\textrm{d}N_s-\textrm{d}L^*_s,\nonumber \\&X^*_s\leqslant f(s),\quad a.s.\quad t\leqslant s\leqslant T, \end{aligned}$$
(A.3)
$$\begin{aligned}&\int _{ t }^{T\wedge \tau ^*}1_{\{X^*_s\ne f(s)\}}\textrm{d}L^*_s=0. \end{aligned}$$
(A.4)

where \(\tau ^*=\inf \{s\geqslant t\mid X^*_s\leqslant 0\}\). Note that (A.2) is true when \(X_s=X^*_s\), \(H_s=H^*_s\) and \(L_s=L^*_s\), i.e.

$$\begin{aligned} {\widehat{V}}(x,t)&= {\mathbb E}_{t,x}\Bigg [e^{-c (\tau \wedge T-t)}{\widehat{V}}(X^*_{\tau ^*\wedge T},\tau ^*\wedge T)\Bigg ]\nonumber \\&\quad -{\mathbb E}_{t,x}\Bigg [\sum _{t\leqslant s{\leqslant } {\tau ^*\wedge T}}e^{-c (s-t)}\Big ({\widehat{V}}(X^*_s,s) -{\widehat{V}}(X^*_s+\Delta L^*_s,s)\Big )\Bigg ]\nonumber \\&\quad -{\mathbb E}_{t,x}\Bigg [\int _{ t }^{\tau \wedge T}e^{-c (s-t)}\Big ({\widehat{V}}_t+\frac{\sigma ^2}{2}{\widehat{V}}_{xx}+\gamma {\widehat{V}}_x -(c+\lambda ) {\widehat{V}}\Big )(X^*_{s-},s) \textrm{d}s\Bigg ]\nonumber \\&\quad -{\mathbb E}_{t,x}\Bigg [\int _{ t }^{\tau \wedge T}e^{-c (s-t)}\lambda \textrm{d}s \int _0^\infty \Big ((1+\delta )H^*_s(z){\widehat{V}}_x(X_{s-},s)\nonumber \\&\quad + {\widehat{V}}(X^*_{s-}-H^*_s(z) ,s)\Big )\textrm{d}F(z)\Bigg ]\nonumber \\&\quad +{\mathbb E}_{t,x}\Bigg [\int _{ t }^{\tau ^*\wedge T}e^{-c (s-t)} {\widehat{V}}_x (X^*_{s-},s)\textrm{d}L^{*c}_s\Bigg ]. \end{aligned}$$
(A.5)

Due to \(f(T)=d(0)=0\) and (A.3), we have \(X^*_T= 0\), so \(X^*_{\tau ^*\wedge T}=0\), so the first expectation above is zero. The variational inequality in (3.1) implies the third expectation together with the fourth expectation is zero. Owing to (A.4) and \({\widehat{V}}_x(f(s),s)=1\),

$$\begin{aligned} {\widehat{V}}_x(X^*_{s-},s)\textrm{d}L^{*c}_s={\widehat{V}}_x(f(s),s)\textrm{d}L^{*c}_s=\textrm{d}L^{*c}_s. \end{aligned}$$

Therefore (A.5) becomes

$$\begin{aligned} {\widehat{V}}(x,t)&= {\mathbb E}_{t,x}\Bigg [\sum _{t\leqslant s{\leqslant } {\tau ^*\wedge T}}e^{-c (s-t)}\Big ({\widehat{V}}(X^*_s+\Delta L^*_s,s)-{\widehat{V}}(X^*_s,s) \Big )\Bigg ]\nonumber \\&\quad +{\mathbb E}_{t,x}\Bigg [\int _{ t }^{\tau ^*\wedge T}e^{-c (s-t)}\textrm{d}L^{*c}_s\Bigg ]. \end{aligned}$$
(A.6)

Now, if \(x\leqslant f(t)\), \(L^*_s\) is continuous and \(\Delta L^*_s=0\) for all \(t\leqslant s\leqslant T\), so

$$\begin{aligned} {\widehat{V}}(x,t) ={\mathbb E}_{t,x}\Bigg [\int _{ t }^{\tau ^*\wedge T}e^{-c (s-t)}\textrm{d}L^{*}_s\Bigg ]; \end{aligned}$$

And if \(x>f(t)\), then \(L^*_s\) is only discontinuous at \(s=t\) and \( \Delta L^*_t=x-f(t). \) Noting that \({\widehat{V}}_x (y,t)=1\) for \(y\in (f(t),x)\), thus (A.6) becomes

$$\begin{aligned} {\widehat{V}}(x,t)&={\mathbb E}_{t,x}[e^{-c (t-t)}(x-f(t))] +{\mathbb E}_{t,x}\Bigg [\int _{ t }^{\tau ^*\wedge T}e^{-c (s-t)}\textrm{d}L^{*c}_s\Bigg ]\\&={\mathbb E}_{t,x}\Bigg [\int _{ t }^{\tau ^*\wedge T}e^{-c (s-t)}\textrm{d}L^{*}_s\Bigg ]. \end{aligned}$$

In view of the definition of V(xt), we obtain \({\widehat{V}}(x,t)\leqslant V(x,t)\). Therefore, \({\widehat{V}}(x,t)= V(x,t)\) and \(\{(L^*_s,H^*_s)\}_{s\geqslant t-}\) is an optimal strategy.

Appendix B Proof of Proposition 4.1

In this section, we prove Proposition  4.1. To prove the strict monotonicity, we first derive the equation on \(u_t\) in \(\mathcal{N}\mathcal{D}\). Differentiating the equation in (3.7) w.r.t. t we have

$$\begin{aligned}{} & {} \Big ({\partial }_t u_t-\mathcal{L}u _t+(\mathcal{Y}u){\partial }_x u _t+{\partial }_t(\mathcal{Y}u)u _x\Big )(x,t)\nonumber \\{} & {} \quad +\lambda \int _0^x 1_{\mathbb B}[(1+\delta )u _t(x,t)-u _t(x-z,t)]\textrm{d}F(z)=0\quad \hbox {in}\; \mathcal{N}\mathcal{D}, \end{aligned}$$
(B.1)

where \({\mathbb B}:=\{(1+\delta )u(x,t)-u(x-z,t)\geqslant 0\}\). Noting that

$$\begin{aligned} {\partial }_t (\mathcal{Y}u)= & {} \lambda (1+\delta )\int _0^\infty {\partial }_t Y^u(x,t,z) \textrm{d}F(z)\nonumber \\= & {} \lambda (1+\delta ) \int _0^\infty \frac{(1+\delta )u_t(x,t)-u_t(Y^u(x,t,z),t)}{u_x(Y^u(x,t,z),t)} 1_{{\mathbb D}}\textrm{d}F(z), \end{aligned}$$
(B.2)

where \({\mathbb D}:=\{u((x-z)^+,t)>(1+\delta )u(x,t)\}\), (B.1) becomes

$$\begin{aligned}{} & {} \Big ({\partial }_t u_t-\mathcal{L}u _t+(\mathcal{Y}u){\partial }_x u _t\Big )(x,t)\nonumber \\{} & {} \qquad +\left[ \lambda \int _0^\infty \frac{u _x(x,t)}{u_x(Y^u(x,t,z),t)}(1+\delta )^21_{{\mathbb D}}\textrm{d}F (z)\right] u_t(x,t)\nonumber \\{} & {} \qquad +\left[ \lambda \int _0^x (1+\delta )1_{\mathbb B}\textrm{d}F (z)\right] u _t(x,t)\nonumber \\{} & {} = \lambda \int _0^\infty \frac{(1+\delta )u _x(x,t)}{u_x(Y^u(x,t,z),t)}1_{{\mathbb D}}u_t(Y^u(x,t,z),t) \textrm{d}F (z)\nonumber \\{} & {} \qquad + \lambda \int _0^x (1+\delta )1_{\mathbb B}u _t(x-z,t) \textrm{d}F (z)\geqslant 0 \end{aligned}$$
(B.3)

in \(\mathcal{N}\mathcal{D}\). Here, all integrations in [...] regarded as coefficients are bounded.

If \(d(\cdot )\) is not strict monotonicity, there exists \(0<t_1<t_2\) such that \(d(t_1)=d(t_2)\), denote \(x_0=d(t_1)=d(t_2)\) and \(\Gamma =\{x_0\}\times [t_1,t_2]\), then on the line \(\Gamma \), there exists

$$\begin{aligned} u=1,\quad u_t=0,\quad u_x=0,\quad u_{xt}=0. \end{aligned}$$
(B.4)

Recalling that \(u_t\geqslant 0\) in \(\mathcal{N}\mathcal{D}\) and \(u_t=0\) on \(\Gamma \), by the Hopf’s boundary point theorem, either \(u_t\equiv 0\) in \(\mathcal{N}\mathcal{D}\) or \(u_{xt}<0\) in \(\Gamma \), but the former leads to \(u\equiv 1\) in \(\mathcal{Q}_T \), the latter contradicts to (B.4).

Appendix C Proof of Lemma 5.1

In this section, we prove Lemma 5.1. In order to convert the mixed boundary condition to a second boundary condition, we first make a transformation \(w=e^{\eta x}(u_1-u_2)\), then w satisfies

$$\begin{aligned} \left\{ \begin{array}{ll} w_t-\mathcal{L}_3 w-g +{J}'(\cdot )w\leqslant 0\quad \hbox {in } {\mathcal{Q}_T^L},\\ w_x(0,t)\geqslant 0 \;(\textrm{or}\,w(0,t)\leqslant 0),\; w(L,t)\leqslant 0,\\ w(x,0)\leqslant 0. \end{array} \right. \end{aligned}$$
(C.1)

where

$$\begin{aligned} \mathcal{L}_3 w&:=\frac{1}{2}\sigma ^2 w_{xx}+\Big ( b -\sigma ^2\eta \Big )w_x+\Big ( d - b \eta +\frac{1}{2}\sigma ^2\eta ^2\Big )w,\\ g\;&:=e^{\eta x}\Big (G (u_1(\cdot ,t),x,t)-G (u_2(\cdot ,t),x,t)\Big ). \end{aligned}$$

Multiplying the equation in (C.1) by \(w^+\) and integrating in \({\mathcal{Q}_\tau ^L}:=[0,L]\times [0,\tau ]\), \(\tau \in [0,T]\), we obtain

$$\begin{aligned}{} & {} \frac{1}{2}\int _0^L (w^+)^2(x,\tau ) \textrm{d}x-\frac{1}{2}\int _0^L (w^+)^2(x,0) \textrm{d}x\nonumber \\{} & {} \quad +\frac{\sigma ^2}{2}\Bigg (\int _{\mathcal{Q}_\tau ^L} ({\partial }_x w^+)^2 \textrm{d}x \textrm{d}t-\int _0^\tau (w^+{\partial }_x w)(L,t)\textrm{d}t+\int _0^\tau (w^+{\partial }_x w)(0,t)\textrm{d}t\Bigg )\nonumber \\{} & {} \quad -\int _{\mathcal{Q}_\tau ^L}\Big ( b -\sigma ^2\eta \Big ) w^+ {\partial }_x w^+\textrm{d}x \textrm{d}t-\int _{\mathcal{Q}_\tau ^L} \Big ( d - b \eta +\frac{1}{2}\sigma ^2\eta ^2\Big ) (w^+)^2 \textrm{d}x \textrm{d}t\nonumber \\{} & {} \quad +\int _{\mathcal{Q}_\tau ^L} {J}'(\cdot )(w^+)^2 \textrm{d}x \textrm{d}t= \int _{\mathcal{Q}_\tau ^L}g w^+ \textrm{d}x \textrm{d}t. \end{aligned}$$
(C.2)

The initial condition in (C.1) yields \(\frac{1}{2}\int _0^L (w^+)^2(x,0) \textrm{d}x=0\). The right boundary condition in (C.1) yields \(\int _0^\tau (w^+{\partial }_x w)(L,t)\textrm{d}t=0\) and the left boundary condition yields \(\int _0^\tau (w^+{\partial }_x w)(0,t)\textrm{d}t\) \(\geqslant 0\). Moreover, by the assumption of G and the embedding theorem, we have

$$\begin{aligned} g(x,t)\leqslant & {} C e^{\eta L} \Big ( | w^+(\cdot ,t) | _{\textrm{L}^\infty [0,L]}+|w(x,t)|\Big )\\\leqslant & {} \epsilon | {\partial }_x w^+(\cdot ,t) | _{\textrm{L}_2 [0,L]}+C_\epsilon | w^+(\cdot ,t) | _{\textrm{L}_2 [0,L]}+C e^{\eta L} |w(x,t)| \end{aligned}$$

for any \(\epsilon >0\) and \(C_\epsilon \) only depends on \(\epsilon \). By using the Cauchy’s inequality, the right hand side of (C.2)

$$\begin{aligned} \int _{\mathcal{Q}_\tau ^L}g w^+ \textrm{d}x \textrm{d}t\leqslant \frac{\epsilon L}{2} \int _{\mathcal{Q}_\tau ^L} ({\partial }_x w^+)^2\textrm{d}x \textrm{d}t + \frac{C_\epsilon L+1+C e^{\eta L}}{2} \int _{\mathcal{Q}_\tau ^L}(w^+)^2\textrm{d}x \textrm{d}t. \end{aligned}$$

Choosing \(\epsilon L=\sigma ^2/4\) and using the Cauchy’s inequality again, there exists a constant \(C_1>0\) such that

$$\begin{aligned} \int _0^L (w^+)^2(x,\tau ) \textrm{d}x\leqslant C_1 \int _{\mathcal{Q}_\tau ^L} (w^+)^2 \textrm{d}x \textrm{d}t. \end{aligned}$$

Applying the Gronwall’s inequality we deduce that \(w^+=0,\) which implies \(u_1\leqslant u_2.\)

Appendix D Hölder Space and \(W^{l,l/2}_p(Q_T)\) Space

In this section, we give the definitions of Hölder space and \(W^{l,l/2}_p(Q_T)\) space with parabolic distance which we used in this paper.

1.1 D.1 Parabolic Distance

Recall that \(\mathbb {R}^n\) is the n-dimensional Euclidean space, the point on which is denoted by \(x=(x_1,x_2,\cdots ,x_n)\). By introducing a time variable t, denote the point on the constructed \(n+1\)-dimensional space \(\mathbb {R}^{n+1}\) as \(X=(x,t_X)\). We now introduce the distance in \(\mathbb {R}^{n+1}\). \(\delta (X,Y)\) is called the parabolic distance if

$$\begin{aligned} \delta (X,Y)=\textrm{max}\{|x-y|,|t_X-t_Y|^{\frac{1}{2}}\}. \end{aligned}$$

Let \(\Omega \) be a bounded region in \(\mathbb {R}^n\) and \(Q_T=\Omega \times (0,T]\) is a cylinder in \(\mathbb {R}^{n+1}\). Now we introduce some spaces.

1.2 D.2 Hölder Space

For \(0<\alpha \leqslant 1\), let \(C^\alpha (\overline{Q_T};\delta )\) denote the normed linear space composed by all functions u, where

$$\begin{aligned}{}[u]_{\alpha ,Q_T}:=\sup _{\begin{array}{c} X,Y\in Q_T\\ X\ne Y \end{array}}\frac{|u(X)-u(Y)|}{\delta (X,Y)^\alpha }<\infty . \end{aligned}$$

The norm is defined by

$$\begin{aligned} |u|_{\alpha ,Q_T}=\sup _{Q_T}|u|+[u]_{\alpha ,Q_T}. \end{aligned}$$

In order to describe spaces of higher orders, we introduce the definition of semi-norm as follows:

$$\begin{aligned} \begin{aligned} |u|_{0,Q_T}&=\sup _{Q_T}|u|, \\ [u]_{\alpha ,Q_T}&=\sup _{\begin{array}{c} X,Y\in Q_T\\ X\ne Y \end{array}}\frac{|u(X)-u(Y)|}{\delta (X,Y)^\alpha }, \quad 0<\alpha<1,\\ [u]_{\alpha ,Q_T}^{t}&=\sup _{\begin{array}{c} x\in \Omega .t\ne r\\ t,r\in [o,T] \end{array}}\frac{|u(x,t)-u(x,\tau )|}{|t-\tau |^\alpha },\quad 0<\alpha <1. \\ \end{aligned} \end{aligned}$$

The semi-norm of higher derivatives is introduced as

$$\begin{aligned}{}[u]_{k+\alpha ,Q_T}=\left\{ \begin{array}{ll} \sum \limits _{r+2s=k}[D_t^sD_x^ru]_{\alpha ,Q_T}, \quad k\mathrm {\; is\; even}, \\ \sum \limits _{r+2s=k}[D_t^sD_x^ru]_{\alpha ,Q_T}+\sum \limits _{r+2s=k-1}[D_t^sD_x^ru]_{\frac{1+\alpha }{2},Q_T}^t,\quad k\mathrm {\; is\; odd}. \end{array} \right. \end{aligned}$$

The linear space composed by functions u in \({C}(\overline{Q}_T)\) where

$$\begin{aligned} |u|_{k+\alpha ,Q_T}:=\sum \limits _{0\leqslant r+2s\leqslant k}[D_t^sD_x^ru]_{0,Q_T}+[u]_{k+\alpha ,Q_T}<\infty \end{aligned}$$

is denoted by \(C^{k+\alpha ,\frac{k+\alpha }{2}}(\overline{Q}_T)\) or \(C^{k+\alpha }(\overline{Q}_T;\delta )\). After introducing the norm \(|u|_{k+\alpha ,Q_T}\), it will turn into a Banach space.

1.3 D.3 \(W^{l,l/2}_p(Q_T)\) Space

For a positive integer l, \(1\leqslant p<\infty \), denote

$$\begin{aligned} | u | _{W_p^{l,\frac{l}{2}}(Q_T)}=\left\{ \begin{array}{ll} \bigg \{\sum \limits _{0\leqslant r+2s\leqslant l} | D_t^sD_x^ru | ^p_{L^p(Q_T)}\bigg \}^{\frac{1}{p}}, \quad l\mathrm {\; is\; even}, \\ \bigg \{\sum \limits _{0\leqslant r+2s\leqslant l} | D_t^sD_x^ru | ^p_{L^p(Q_T)}+\sum \limits _{0\leqslant r+2s\leqslant l-1}[D_t^sD_x^ru]^p_{L^{\frac{1}{2}}_{p,t}(Q_T)}\bigg \}^{\frac{1}{p}},\quad l\mathrm {\; is\; odd}. \end{array} \right. \end{aligned}$$

where,

$$\begin{aligned}{}[u]_{L^{\frac{1}{2}}_{p,t}(Q_T)}=\Bigg \{\int _\Omega \textrm{d}x \int _0^T \textrm{d}t \int _0^T\frac{|u(x,t)-u(x,\tau )|^p}{|t-\tau |^{1+p/2}} \textrm{d}\tau \Bigg \}^{1/p}. \end{aligned}$$

Let \(W_p^{l,\frac{l}{2}}(Q_T)\) denote the normed linear space composed by functions u where

$$\begin{aligned} | u | _{W_p^{l,\frac{l}{2}}(Q_T)}<\infty . \end{aligned}$$

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Guan, C. Finite Horizon Optimal Dividend and Reinsurance Problem Driven by a Jump-Diffusion Process with Controlled Jumps. Appl Math Optim 88, 15 (2023). https://doi.org/10.1007/s00245-023-09992-2

Download citation

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00245-023-09992-2

Keywords

Mathematics Subject Classification

Navigation