Skip to main content
Log in

Optimal risk transfers in insurance groups

  • Original Research Paper
  • Published:
European Actuarial Journal Aims and scope Submit manuscript

Abstract

Optimal risk transfers are derived within an insurance group consisting of two separate legal entities, operating under potentially different regulatory capital requirements and capital costs. Consistent with regulatory practice, capital requirements for each entity are computed by either a value-at-risk or an expected shortfall risk measure. The optimality criterion consists of minimising the risk-adjusted value of the total group liabilities, with valuation carried out using a cost-of-capital approach. The optimisation problems are analytically solved and it is seen that optimal risk transfers often involve the transfer of tail risk (unlimited reinsurance layers) to the more weakly regulated entity. We show that, in the absence of a capital requirement for the credit risk that specifically arises from the risk transfer, optimal risk transfers achieve capital efficiency at the cost of increasing policyholder deficit. However, when credit risk is properly reflected in the capital requirement, incentives for tail-risk transfers vanish and policyholder welfare is restored.

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

Similar content being viewed by others

References

  1. Acerbi C, Tasche D (2002) On the coherence of expected shortfall. J Bank Finance 26(7):1487–1503

    Article  Google Scholar 

  2. Albrecher H, Thonhauser S (2009) Optimality results for dividend problems in insurance. RACSAM Rev R Acad Cien Ser A Mat 103(2):295–320

    Article  MathSciNet  MATH  Google Scholar 

  3. Arrow KJ (1963) Uncertainty and the welfare economics of medical care. Am Econ Rev 53(5):941–973

    Google Scholar 

  4. Artzner P, Delbaen F, Eber JM, Heath D (1999) Coherent measures of risk. Math Finance 9:203–228

    Article  MathSciNet  MATH  Google Scholar 

  5. Bernard C, Tian W (2010) Insurance market effects of risk management metrics. Geneva Risk Insur Rev 35:47–80

    Article  Google Scholar 

  6. Borch K (1960) An attempt to determine the optimum amount of stop loss reinsurance. Trans 16th Int Cong Actuar I:597–610

    Google Scholar 

  7. Cai J, Tan KS, Weng C, Zhang Y (2008) Optimal reinsurance under VaR and CTE risk measures. Insur Math Econ 43(1):185–196

    Article  MathSciNet  MATH  Google Scholar 

  8. Carlier G, Dana R-A (2003) Pareto efficient insurance contracts when the insurer’s cost function is discontinuous. Econ Theory 21(4):871–893

    Article  MathSciNet  MATH  Google Scholar 

  9. Centeno ML, Guerra M (2010) The optimal reinsurance strategy—the individual claim case. Insur Math Econ 46(3):450–460

    Article  MathSciNet  MATH  Google Scholar 

  10. Cheung KC (2010) Optimal reinsurer revisited—a geometric approach. Astin Bull 40(1):221–239

    Article  MathSciNet  MATH  Google Scholar 

  11. Chi Y, Tan KS (2011) Optimal reinsurance under VaR and CVaR risk measures: a simplified approach. Astin Bull 42(1):487–509

    MathSciNet  Google Scholar 

  12. Denuit M, Dhaene J, Goovaerts M, Kaas R (2005) Actuarial theory for dependent risks: measures, orders and models. Wiley, Chichester

  13. Dowd K, Blake D (2006) After VaR: the theory, estimation, and insurance applications of quantile-based risk measures. J Risk Insur 73(2):193–229

    Article  Google Scholar 

  14. Embrechts P, Hofert M (2010) A note on generalized inverses (Preprint).

  15. EIOPA: (2011). “Equivalence Assessment of the Swiss Supervisory System in Relation to Articles 172, 227 and 260 of the Solvency II Directive", EIOPA-BoS-11-028. https://eiopa.europa.eu/consultations/consultation-papers/index.html.

  16. EIOPA (2009) Advice for level 2 implementing measures on solvency II: supervision of risk concentration and intra-group transactions, CEIOPS-DOC-53/09. https://eiopa.europa.eu/consultations/consultation-papers/2010-2009-closed-consultations/index.html.

  17. European Commission (2009) Directive 2009/138/EC of the European Parliament and of the Council of 25 November 2009 on the taking-up and pursuit of the business of insurance and reinsurance (Solvency II). Off J Eur Union, L335.

  18. Federal Office of Private Insurance: 2006. The Swiss experience with market consistent technical provisions—the cost of capital approach. http://www.finma.ch/archiv/bpv/e/themen/00506/00552/00727/index.html.

  19. Filipović D, Kupper M (2008) Optimal capital and risk transfers for group diversification. Math Finance 18:55–76

    Article  MathSciNet  MATH  Google Scholar 

  20. Gatzert N, Schmeiser H (2011) On the risk situation of financial conglomerates: does diversification matter? Financial Mark Portfolio Manag 25:3–26

    Article  Google Scholar 

  21. Guerra M, Centeno ML (2008) Optimal reinsurance policy: the adjustment coefficient and the expected utility criteria. Insur Math Econ 42(2):529–539

    Article  MathSciNet  MATH  Google Scholar 

  22. Guerra M, Centeno ML (2010) Optimal reinsurance for variance related premium calculation principles. Astin Bull 40(1):97–121

    Article  MathSciNet  MATH  Google Scholar 

  23. Van Heerwaarden AE, Kaas R, Goovaerts MJ (1989) Optimal reinsurer in the relation to ordering of risk. Insur Math Econ 8(1):11–17

    Article  MathSciNet  MATH  Google Scholar 

  24. Hipp C, Vogt M (2003) Optimal dynamic XL reinsurance. Astin Bull 33(2):193–207

    Article  MathSciNet  MATH  Google Scholar 

  25. Hürlimann V (2003) Conditional value-at-risk bounds for compound Poisson risks and a normal approximation. J Appl Math 3:141–153

    Article  Google Scholar 

  26. Kaluszka M (2001) Optimal reinsurance under mean-variance premium principles. Insur Math Econ 28(1):61–67

    Article  MathSciNet  MATH  Google Scholar 

  27. Kaluszka M (2005) Truncated stop loss as optimal reinsurance agreement in one-period models. Astin Bull 35(2):337–349

    Article  MathSciNet  MATH  Google Scholar 

  28. Kaluszka M, Okolewski A (2008) An extension of Arrow’s result on optimal reinsurance contract. J Risk Insur 75(2):275–288

    Article  Google Scholar 

  29. Keller P (2007) Group diversification. Geneva Pap 38:382–392

    Article  Google Scholar 

  30. Landsberger M, Meilijson II (1994) Comonotone allocations, Bickel Lehmann dispersion and the Arrow–Pratt measure of risk aversion. Ann Oper Res 52:97–106

    Article  MathSciNet  MATH  Google Scholar 

  31. Ludkovski M, Young VR (2009) Optimal risk sharing under distorted probabilities. Math Financial Econ 2(2):87–105

    Article  MathSciNet  MATH  Google Scholar 

  32. Myers SC, Read JA Jr (2001) Capital allocation for insurance companies. J Risk Insur 68(4):545–580

    Article  Google Scholar 

  33. Phillips RD, Cummins JD, Allen F (1998) Financial pricing of insurance in the multiple-line insurance company. J Risk Insur 65(4):597–636

    Article  Google Scholar 

  34. Schmidli H (2001) Optimal proportional reinsurance policies in a dynamic setting. Scand Actuar J 1:55–68

    Article  MathSciNet  Google Scholar 

  35. Schmidli H (2002) On minimizing the ruin probability by investment and reinsurance. Ann Probab 12(3):890–907

    Article  MathSciNet  MATH  Google Scholar 

  36. Schlütter S, Gründl H (2011) Who benefits from buildiung insurance groups? A welfare analysis based on optimal group risk management. ICIR working paper series Nr. 8. Goethe Universität, Frankfurt am Main

  37. Verlaak R, Beirlant J (2003) Optimal reinsurance programs: an optimal combination of several reinsurance protections on a heterogeneous insurance Portfolio. Insur Math Econ 33(2):381–403

    Article  MathSciNet  MATH  Google Scholar 

  38. Wüthrich MV, Bühlmann H, Furrer H (2010) Market-consistent actuarial valuation, 2nd edn. Springer, Berlin

  39. Young VR (1999) Optimal insurance under Wang’s premium principle. Insur Math Econ 25(2):109–122

    Article  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alexandru V. Asimit.

Appendices

Appendix

Proofs

It is useful to note here that I 1 and I 2 are Lipschitz continuous functions with unit constants, i.e. |I k [y] − I k [x]| ≤ |y − x| holds for all x,y ≥ 0 and k ∈ {1,2}, due to the fact that I 1 and I 2 are non-decreasing functions (see also [8]). In addition, VaR α_k (I k [X]) = I k (VaR α_k (X)) is true for all k ∈ {1,2} (see for example [12], p. 19, and [14]).

The following notations,

$$ \begin{aligned} {\mathcal{A}}_1=\left\{(\xi_1,\xi_2):0\leq \xi_1\leq VaR_{\alpha_1}(X),0\leq \xi_2-\xi_1\leq VaR_{\alpha_2}(X)- VaR_{\alpha_1}(X)\right\}& \\ {\mathcal{A}}_2=\left\{(\xi_1,\xi_2):0\leq \xi_2\leq VaR_{\alpha_2}(X),0\leq \xi_1-\xi_2\leq VaR_{\alpha_1}(X)- VaR_{\alpha_2}(X)\right\} & \\ \end{aligned} $$

are useful for explaining all the proofs. In addition, \(\Pr(Z\leq z)<\alpha\) is equivalent to z < VaR α (X).

2.1 Proof of Theorem 2.1

The first part is now investigated assuming that λ1 > λ2. The mirror case, λ1 < λ2, results by the symmetry property of the objective function from Eq. (2.5). Now, for any arbitrary risk transfer I 2, we have

$$ \begin{aligned} &(\lambda_1-\lambda_2)E(I_2[X])+\lambda_2 I_2(VaR_{\alpha_2}(X))-\lambda_1 I_2(VaR_{\alpha_1}(X)) \\ &\quad= (\lambda_1-\lambda_2)E(I_2[X]-I_2(VaR_{\alpha_1}(X))) + \lambda_2 (I_2(VaR_{\alpha_2}(X))-I_2(VaR_{\alpha_1}(X))) \\ &\quad\geq (\lambda_1-\lambda_2)E(I_2[\min\{X,VaR_{\alpha_1}(X)\}]-I_2(VaR_{\alpha_1}(X))) \cr& \qquad+\lambda_2 (I_2(VaR_{\min\{\alpha_1,\alpha_2\}}(X))-I_2(VaR_{\alpha_1}(X))) \\ &\quad\geq (\lambda_1-\lambda_2)E(\min\{X,VaR_{\alpha_1}(X)\}-VaR_{\alpha_1}(X))+\lambda_2 (VaR_{\min\{\alpha_1,\alpha_2\}}(X)-VaR_{\alpha_1}(X)) \\ &\quad=(\lambda_1-\lambda_2)E(I^*_2[X])+\lambda_2 I^*_2(VaR_{\alpha_2}(X))-\lambda_1 I^*_2(VaR_{\alpha_1}(X)), \\ \end{aligned} $$
(5.1)

where the second last step is due to the Lipschitz property of function \(I_2(\cdot)\). Note that the optimal solution stated in Theorem 2.1i) was \(I^*_2[X]=\min\{X,VaR_{\alpha_1}(X)\}\), which concludes this case.

Finally, the λ1 = λ2 case is discussed, which is reduced to minimising

$$ I_2(VaR_{\alpha_2}(X))-I_2(VaR_{\alpha_1}(X)). $$
(5.2)

Recall that the Lipschitz property implies

$$ |I_2(VaR_{\alpha_2}(X))-I_2(VaR_{\alpha_1}(X))|\leq |VaR_{\alpha_2}(X)-VaR_{\alpha_1}(X)|. $$

Thus, the solutions of (5.2) only require \(I_2(\cdot)\) to be flat on \([VaR_{\alpha_1}(X),VaR_{\alpha_2}(X)]\) whenever VaR α_1(X) < VaR α_2(X), and have a slope of 1 on \([VaR_{\alpha_2}(X),VaR_{\alpha_1}(X)]\) otherwise. The latter and Lipschitz property help in recovering the λ1 = λ2 case. The proof is now complete.

2.2 Proof of Theorem 2.2

The proofs are developed in a similar manner as in Theorem 2.1, and therefore we only outline the main steps. The case in which λ1 > λ2 can be proved in the same manner as shown in Eq. (5.1). Specifically,

$$ \begin{aligned} &{(\lambda_1-\lambda_2)E(I_2[X])+\lambda_2 ES_{\alpha_2} (I_2[X])-\lambda_1 I_2(VaR_{\alpha_1}(X))} \\ &\quad\geq (\lambda_1-\lambda_2)E(I_2[\min\{X,VaR_{\alpha_1}(X)\}])+\lambda_2 ES_{\alpha_2} (I_2[\min\{X,VaR_{\alpha_1}(X)\}]) \cr&\qquad-\lambda_1 I_2(VaR_{\alpha_1}(X)) \\ &\quad=(\lambda_1-\lambda_2)E(I_2[\min\{X,VaR_{\alpha_1}(X)\}]-I_2(VaR_{\alpha_1} (X))) \\ &\qquad+\lambda_2 ES_{\alpha_2} (I_2[\min\{X,VaR_{\alpha_1}(X)\}]-I_2(VaR_{\alpha_1} (X))) \cr&\quad\geq (\lambda_1\!-\!\lambda_2)E(\!\min\!\{X,VaR_{\alpha_1}(X)\}\!-\!VaR_{\alpha_1}(X))\!+\!\lambda_2 ES_{\alpha_2}(\!\min\!\{X,VaR_{\alpha_1}(X)\}\!-\!VaR_{\alpha_1}(X)) \\ &\quad=(\lambda_1\!-\!\lambda_2)E(I^*_2[X])+\lambda_2 ES_{\alpha_2} (I^*_2[X])-\lambda_1 I^*_2(VaR_{\alpha_1}(X)), \\ \end{aligned} $$

which implies that the optimisation problem from (2.6) is attained at I *2 [X] = min{X,\(VaR_{\alpha_{1}}(X)\)}.

Similarly, the \([VaR_{\alpha_{1}}(X) \leq VaR_{\alpha_2}(X)]\) case such that λ1 < λ2 is examined. Clearly,

$$ (\lambda_1-\lambda_2)E(I_2[X])+\lambda_2 ES_{\alpha_2} (I_2[X])-\lambda_1 I_2(VaR_{\alpha_1}(X)) \geq (\lambda_2-\lambda_1) (ES_{\alpha_2} (I_2[X])-E (I_2[X]))\geq 0. $$

The lower bound is attained whenever I *2 [X] = 0, which fully concludes part 1).

The remaining of part 2), \(VaR_{\alpha_{1}}(X)\) > \(VaR_{\alpha_{2}}(X)\) with λ1 < λ2, needs to be elaborated in greater detail than the previous one. The optimisation problem (2.6) is solved via a two stage optimisation procedure. The mathematical formulation of the first stage optimisation problem becomes

$$ \left\{\begin{array}{l} \displaystyle\min_{I_2\in {\cal F}} (\lambda_1-\lambda_2)E(I_2[X])+\lambda_2 ES_{\alpha_2} (I_2[X])\; {\rm subject\; to}\\ I_2(VaR_{\alpha_1}(X))=\xi_1,\;I_2(VaR_{\alpha_2}(X))=\xi_2, \end{array}\right. $$

where \((\xi_1,\xi_2)\in\mathcal{A}_2\) are some constants. Keeping in mind that λ1 − λ2 < 0 and relation (2.1), the function \(I_2(\cdot)\) should increase as fast as possible on [0,\(VaR_{\alpha_{1}}(X)\)], and remain flat on [\(VaR_{\alpha_{1}}(X)\),x F ]. The optimal solution is then pictured in Fig. 5

Fig. 5
figure 5

The construction of I *2a [X12] in Theorem 2.2

One may mathematically formulate the latter risk transfer as

$$ I_{2a}^*[X;\xi_1,\xi_2]=\left\{\begin{array}{ll} \min\{X-VaR_{\alpha_1}(X)+\xi_1,\xi_1\},\; & X>VaR_{\alpha_1}(X)+\xi_2-\xi_1,\\ \min\{X,\xi_2\},\; & {otherwise}. \end{array}\right. $$

Thus, the solution of our second stage optimisation problem

$$ \begin{aligned} \min_{(\xi_1,\xi_2)\in {\mathcal{A}}_2} H(\xi_1,\xi_2)&=(\lambda_1-\lambda_2)\left(\int\limits_0^{\xi_2}\bar{F}(x)\;dx+\int\limits_{VaR_{\alpha_1}(X)+\xi_2-\xi_1}^{VaR_{\alpha_1}(X)}\bar{F}(x)\;dx\right) \\ &\quad+\lambda_2\left(\xi_2+{{1}\over {1-\alpha_2}} \int\limits_{VaR_{\alpha_1}(X)+\xi_2-\xi_1}^{VaR_{\alpha_1}(X)}\bar{F}(x)\;dx\right)-\lambda_1 \xi_1 \end{aligned} $$

replicates the global minimiser of (2.6). The derivative of the above-mentioned function with respect to ξ1 becomes

$$ {{dH}\over {d\xi_1}}=\left(\lambda_1+\lambda_2 {{\alpha_2}\over {1-\alpha_2}}\right)\bar{F}\left(VaR_{\alpha_1}(X)+\xi_2-\xi_1\right)-\lambda_1, $$
(5.3)

which is non-positive if and only if \(VaR_{\alpha_1}(X)+\xi_2-\xi_1\geq VaR_{\alpha_2^*}(X)\). Due to the fact that λ1 < λ2, it is not difficult to find that α *2  > α2. Note that VaR α_2(X) ≤ VaR α_1(X) + ξ2 − ξ1 ≤ VaR α_1(X) holds. Having all of these in mind, we may conclude

$$ H(\xi_1,\xi_2)\geq H(\xi_2,\xi_2),\;{\rm if}\;\alpha_2^*\geq \alpha_1 $$

and

$$ H(\xi_1,\xi_2)\geq H(\xi_2+VaR_{\alpha_1}(X)-VaR_{\alpha_2^*}(X),\xi_2),\;{\rm if}\;\alpha_2^*<\alpha_1. $$

Finding the minimum values of the above-right hand side functions over the domain of ξ2, namely \([0,VaR_{\alpha_2}(X)]\), the global solution of (2.6) is attained at (0,0) and \((VaR_{\alpha_1}(X)-VaR_{\alpha_2^*}(X),0)\), if α *2  ≥ α1 and α *2  < α1, respectively. Thus, part 2) is concluded.

Finally, the λ1 = λ2 case is under investigation. Our optimisation problem becomes

$$ \min_{I_2 \in\cal{F}} ES_{\alpha_2} (I_2[X]) - VaR_{\alpha_1}I_2[X]. $$

Thus, the first stage optimisation problem is equivalent to minimising \(ES_{\alpha_2} (I_2[X])\).

The VaR α_1(X) ≤ VaR α_2(X) subcase makes \(I_2(\cdot)\) to be flat for losses higher than VaR α_2(X). Thus, the second stage problem is to minimise ξ2 − ξ1 on \(\mathcal{A}_1\), and keeping in mind the Lipschitz property, one may recover our claim.

The last subcase, VaR α_1(X) > VaR α_2(X), is developed in greater detail. The first part of A.2 only requires for the function \(I_2(\cdot)\) to not increase on \([VaR_{\alpha_2}(X),VaR_{\alpha_1}(X)-(\xi_1-\xi_2)]\), then to increase with a slope of 1 until it reaches the ξ1 level at VaR α_1(X), and remain flat on \([VaR_{\alpha_1}(X),x_F]\). Thus, the second stage problem of 5.2 becomes

$$ \min_{{\mathcal{A}}_2} \; \xi_2 +{{1}\over {1-\alpha_2}} \int\limits_{VaR_{\alpha_1}(X)-(\xi_1-\xi_2)}^{VaR_{\alpha_1}(X)} \bar{F}(x)\;dx-\xi_1. $$

Therefore, the objective function of the above is non-increasing in ξ1 for any fixed ξ2, and straightforward calculations show that the set of optimal solutions is given by

$$ \{(x+VaR_{\alpha_1}(X)-VaR_{\alpha_2}(X),x):x\in [0,VaR_{\alpha_2}(X)]\}, $$

which justifies in full the VaR α1(X) > VaR α2(X) subcase. This ends the proof.

2.3 Proof of Theorem 2.3

  1. 1.

    If \( VaR_{\alpha_1}(X) = VaR_{\alpha_2}(X),\) the first stage optimisation problem is given by

    $$ \min_{I_2\in {\mathcal{F}}} (\lambda_1-\lambda_2)E(I_2[X])+\lambda_2 ES_{\alpha_2} (I_2[X])-\lambda_1 ES_{\alpha_1} (I_2[X])\;{subject\ to}\;I_2(VaR_{\alpha_1}(X))=\xi, $$
    (5.4)

    where \(\xi\in[0,VaR_{\alpha_1}(X)]\) represents an arbitrarily chosen constant.

    Let λ1 > λ2. It is not difficult to justify via some geometric argumentation that (5.4) has solutions given by \((X-VaR_{\alpha_1}(X)+\xi)_+\) and \(\min \{(X-VaR_{\alpha_1}(X)+\xi)_+,\xi\}\) whenever C < 0 and C > 0, respectively.

    Next, it is assumed that C < 0, and plugging the solution of (5.4) in its objective function, the second stage optimisation problem is reduced to minimising

    $$ (\lambda_1-\lambda_2)\int\limits_{VaR_{\alpha_1}(X)-\xi}^{x_F}\bar{F}(x)\;dx+\lambda_2 (\xi+ES_{\alpha_2} (X) -VaR_{\alpha_1}(X))-\lambda_1 (\xi+ES_{\alpha_1} (X)-VaR_{\alpha_1}(X)), $$

    over all possible values of ξ. The latter function is decreasing on \([0,VaR_{\alpha_1}(X)]\), which leads to full transfer of the risk to the second insurance company.

    For positive values of C > 0, the second stage optimisation problem is retrieved by minimising

    $$ (\lambda_1-\lambda_2)\left(\int\limits_{VaR_{\alpha_1}(X)-\xi}^{VaR_{\alpha_1}(X)}\bar{F}(x)\;dx-\xi\right), $$

    over \(\xi\in[0,VaR_{\alpha_1}(X)]\) due to the fact that

    $$ ES_{\alpha_1} (\min \{(X-VaR_{\alpha_1}(X)+\xi)_+,\xi\}) =ES_{\alpha_2} (\min \{(X-VaR_{\alpha_1}(X)+\xi)_+,\xi\})=\xi. $$

    The decreasing property of the above-mentioned function implies that \(I_2^*[X]=\min\{X,VaR_{\alpha_1}(X)\}\), which concludes the λ1 > λ2 case.

    We now consider the situation in which λ1 < λ2. One may get that (5.4) is minimised by \(\min\{X,\xi\}+(X-VaR_{\alpha_1}(X))_+\) and \(\min\{X,\xi\}\) whenever C < 0 and C > 0, respectively. Thus, the second stage problems are the minimisation in ξ of

    $$ (\lambda_1-\lambda_2)\left(\int\limits_0^{\xi}+\int\limits_{VaR_{\alpha_1}(X)}^{x_F}\right)\bar{F}(x)\;dx+(\lambda_2-\lambda_1)(\xi-VaR_{\alpha_1}(X))+\lambda_2 ES_{\alpha_2} (X)-\lambda_1 ES_{\alpha_1} (X) $$

    and

    $$ (\lambda_1-\lambda_2)\left(\int\limits_{VaR_{\alpha_1}(X)-\xi}^{VaR_{\alpha_1}(X)}\bar{F}(x)\;dx-\xi\right), $$

    provided that C < 0 and C > 0, respectively, over the set \([0,VaR_{\alpha_1}(X)]\). Clearly, both functions are increasing on the whole domain, and therefore one may easily recover optimal contracts in our setting, i.e. if λ1 < λ2.

    The λ1 = λ2 setting is investigated at the moment, for which

    $$ \min_{\cal{F}}\;ES_{\alpha_2} I_2[X]-ES_{\alpha_2} I_2[X] $$
    (5.5)

    needs to be solved. The first stage problem becomes

    $$ \min_{0\leq \xi\leq VaR_{\alpha_1}(X)}\;{{\alpha_2-\alpha_1}\over {(1-\alpha_1)(1-\alpha_2)}}E (I_2[X]-\xi)_+, \;{subject\ to} \; I_2(VaR_{\alpha_1}(X))=\xi $$

    as a result of relation (2.1). Note that C < 0 is the same as α1 > α2, which makes \(I_2(\cdot)\) to increase with slope 1 from VaR α_1(X) onwards in order to solve . In turn, the C > 0 case requires a leveled \(I_2(\cdot)\) function on \([VaR_{\alpha_1}(X),x_F]\), since α1 < α2. Thus, the justification of VaR α_1(X) = VaR α_2(X) is now completed.

  2. 2.

    The VaR α_1(X) > VaR α_2(X) case is further examined. If λ1 > λ2, then

    $$ (\lambda_1-\lambda_2)E(I_2[X])+\lambda_2 ES_{\alpha_2} (I_2[X])-\lambda_1 ES_{\alpha_1} (I_2[X]) \geq (\lambda_2-\lambda_1)(ES_{\alpha_1} (I_2[X])-E (I_2[X])), $$

    and therefore (2.8) is the same as to maximse \(ES_{\alpha_1} (I_2[X])-E (I_2[X])\), which in turn is equivalent to minimising \(ES_{\alpha_1} (I_1[X])-E (I_1[X])\). Thus, I *1 [X] = 0, which concludes the λ1 > λ2 subcase.

    Our main objective from (2.8) can be solved via a two stage optimisation, whenever λ1 ≤ λ2. Particularly, the first step can be written in the following fashion

    $$ \left\{ \begin{array}{l}\min_{I_2\in {\mathcal F}} (\lambda_1-\lambda_2)E(I_2[X])+\lambda_2 ES_{\alpha_2} (I_2[X])-\lambda_1 ES_{\alpha_1} (I_2[X])\; {\rm subject\; to}\cr I_2(VaR_{\alpha_1}(X))=\xi_1,\;I_2(VaR_{\alpha_2}(X))=\xi_2, \end{array}\right. $$

    where \((\xi_1,\xi_2)\in\mathcal{A}_2\) is a fixed vector of constants. It can be shown that the solutions are

    $$ I_2^*[X;\xi_1,\xi_2]=\left\{\begin{array}{lll}I_{2a}^*[X;\xi_1,\xi_2],& \lambda_1< \lambda_2,& C>0,\\ I_{2b}^*[X;\xi_1,\xi_2],& \lambda_1< \lambda_2,&C<0,\end{array}\right. $$

    where

    $$ I^*_{2b}[X;\xi_1,\xi_2]=\left\{\begin{array}{ll}X-VaR_{\alpha_1}(X)+\xi_1,& X>VaR_{\alpha_1}(X)+\xi_2-\xi_1, \\ \min\{X,\xi_2\},& {otherwise}.\end{array}\right. $$

    In the case that λ1 < λ2 and C < 0, we only need to solve

    $$ \begin{aligned} &{\min_{(\xi_1,\xi_2)\in{\mathcal{A}}_2} H(\xi_1,\xi_2):=(\lambda_1-\lambda_2)\left(\int\limits_0^{\xi_2}+\int\limits_{VaR_{\alpha_1}(X)+\xi_2-\xi_1}^{x_F}\right)\bar{F}(x)\;dx} \\ &\quad+\lambda_2 \left(\xi_2+{{1}\over {1-\alpha_2}} \int\limits_{VaR_{\alpha_1}(X)+\xi_2-\xi_1}^{x_F}\bar{F}(x)\;dx\right)- \lambda_1 \left (\xi_1+{{1}\over {1-\alpha_1}} \int\limits_{VaR_{\alpha_1}(X)}^{x_F}\bar{F}(x)\;dx\right). \end{aligned} $$
    (5.6)

Note that α2 < α *2  < α1 is true under this setting. The derivative of the above with respect to ξ1 is given by (A.3), and it is non-positive if and only if \(\xi_2\leq \xi_1\leq \xi_2 + VaR_{\alpha_1}(X)-VaR_{\alpha_2^*}(X).\) Thus,

$$ H(\xi_1,\xi_2)\geq H(\xi_2 + VaR_{\alpha_1}(X)-VaR_{\alpha_2^*}(X),\xi_2)=(\lambda_1-\lambda_2)\left(\int\limits_0^{\xi_2} \bar{F}(x)\;dx-\xi_2\right)+K, $$

where K is a constant with respect to ξ2. The latter function is increasing in ξ2 on \([0,VaR_{\alpha_2}(X)]\), and therefore (5.6) is solved by \((\xi_1^*,\xi_2^*)=(VaR_{\alpha_1}(X)-VaR_{\alpha_2^*}(X),0)\), which replicates the optimal risk transfer in this subcase.

The situation in which λ1 < λ2 and C > 0 requires minimising over \(\mathcal{A}_2\)

$$ (\lambda_1-\lambda_2)\left(\int\limits_0^{\xi_2}+\int\limits_{VaR_{\alpha_1}(X)+\xi_2-\xi_1}^{VaR_{\alpha_1}(X)}\right)\bar{F}(x)\;dx +\lambda_2 \left(\xi_2+{{1}\over {1-\alpha_2}} \int\limits_{VaR_{\alpha_1}(X)+\xi_2-\xi_1}^{VaR_{\alpha_1}(X)}\bar{F}(x)\;dx\right)- \lambda_1 \xi_1. $$

Clearly, α2 < α1 < α *2 , and therefore the function from above is non-decreasing in ξ1 for any fixed ξ2, since again its derivative with respect with ξ1 is given by Eq. (5.3). Similar derivations to the previous subcases lead to the global minimum to be attained at \((\xi_1^*,\xi_2^*)=(0,0)\).

The final setting, under which λ1 = λ2, needs to be justified. The first stage optimisation problem of (5.5) forces the risk transfer to remain flat on \([VaR_{\alpha_2}(X),VaR_{\alpha_1}(X)-(\xi_1-\xi_2)]\), and increase as fast as possible onwards, since α1 > α2. The second stage problem becomes

$$ \min_{\cal{A}_2}\;\xi_2-\xi_1+{{1}\over {1-\alpha_2}}\int\limits_{VaR_{\alpha_1}(X)-(\xi_1-\xi_2)}^{x_F} \bar{F}(x)\;dx-{{1}\over {1-\alpha_1}}\int\limits_{VaR_{\alpha_1}(X)}^{x_F} \bar{F}(x)\;dx. $$
(5.7)

Note that the above objective function is non-increasing in ξ1 for any fixed ξ2, and therefore the set of solutions for (5.7) is given by

$$ \{(VaR_{\alpha_1}(X)-VaR_{\alpha_2}(X)+x,x):0\leq x\leq VaR_{\alpha_2}(X)\}. $$

Thus, we only need to impose a linear increasing with slope 1 for \(I_2(\cdot)\) on [VaR α_2(X),x F ] in order to solve the initial optimisation problem. This completes the proof.

2.4 Proof of Lemma 3.1

Note that VaR α_1(X 1) = 0 and \(VaR_{\alpha_2}(X_2)=VaR_{\alpha_1^*}(X)\), which together with (2.1) imply that

$$ ES_{\alpha_1}(X_1) ={{1}\over {1-\alpha_1}}\int\limits_{VaR_{\alpha_1^*}(X)}^{VaR_{\alpha_2}(X)} \bar F(x)\;dx. $$

Now,

$$ \begin{aligned} EPD(X_1;ES_{\alpha_1}(X_1))=&\int\limits_{ES_{\alpha_1}(X_1)}^{x_F} P(X_1>z)\;dz \\ =&\int\limits_{\min\{VaR_{\alpha_2}(X),VaR_{\alpha_1^*}(X)+ES_{\alpha_1}(X_1)\}}^{VaR_{\alpha_2}(X)}\bar F(x)\;dx \\ EPD(X_2;VaR_{\alpha_2}(X_2))=&\int\limits_{VaR_{\alpha_2}(X_2)}^{x_F}P(X_2>z)dz=\int\limits_{VaR_{\alpha_2}(X)}^{{x_F}} \bar F(x)\;dx \\ \end{aligned} $$

from which part i) follows. Clearly, \(\tilde X_1=(X-VaR_{\alpha_1^*}(X))_+\) and we further have that

$$ ES_{\alpha_1}(\tilde X_1)={{1}\over {1-\alpha_1}}\int\limits_{VaR_{\alpha_1^*}(X)}^{{x_F}} \bar F(x)\;dx, $$

which concludes parts ii) and iii).

The first inequality in the lemma holds as long as

$$ VaR_{\alpha_1^*}(X)+{{1}\over {1-\alpha_1}}\int\limits_{VaR_{\alpha_1^*}(X)}^{VaR_{\alpha_2}(X)} \bar F(x)\;dx \leq VaR_{\alpha_2}(X). $$

The latter holds as a result of

$$ \begin{aligned} {{1}\over {1-\alpha_1}}\int\limits_{VaR_{\alpha_1^*}(X)}^{VaR_{\alpha_2}(X)} \bar F(x)\;dx \leq& {{\bar F ( VaR_{\alpha_1^*}(X))}\over {1-\alpha_1}}(VaR_{\alpha_2}(X)-VaR_{\alpha_1^*}(X)) \cr \leq& {{1-\alpha_1^*}\over {1-\alpha_1}}(VaR_{\alpha_2}(X)-VaR_{\alpha_1^*}(X)) \\ \leq & VaR_{\alpha_2}(X)-VaR_{\alpha_1^*}(X). \end{aligned} $$

It only remains to show the very last inequality from Lemma 3.1, for which it is sufficient to show that

$$ VaR_{\alpha_1^*}(X)+{{1}\over {1-\alpha_1}}\int\limits_{VaR_{\alpha_1^*}(X)}^{{x_F}} \bar F(x)\;dx \geq VaR_{\alpha_1}(X)+{{1}\over {1-\alpha_1}}\int\limits_{VaR_{\alpha_1}(X)}^{{x_F}} \bar F(x)\;dx. $$

The latter is true since \(\bar F(VaR_{\alpha_1}(X))\leq 1-\alpha_1\) and the fact that

$$ \begin{aligned} \int\limits_{VaR_{\alpha_1}(X)}^{{x_F}} \bar F(x)\;dx&=\int\limits_{VaR_{\alpha_1}(X)}^{VaR_{\alpha_1^*}(X)} \bar F(x)\;dx+\int\limits_{VaR_{\alpha_1^*}(X)}^{{x_F}} \bar F(x)\;dx \\ &\leq (VaR_{\alpha_1^*}(X)-VaR_{\alpha_1}(X))\bar F(VaR_{\alpha_1}(X))+\int\limits_{VaR_{\alpha_1^*}(X)}^{{x_F}} \bar F(x)\;dx. \end{aligned} $$

2.5 Proof of Lemma 3.2

Note first that VaR α_1(X 1) = VaR α_1(X) and VaR α_2(X 2) = 0, which together with (2.1) give that

$$ ES_{\alpha_1}(X_1) =VaR_{\alpha_1}(X)+ {{1}\over {1-\alpha_1}}\int\limits_{VaR_{\alpha_1}(X)}^{VaR_{\alpha_2}(X)} \bar F(x)\;dx. $$

Clearly, \(\tilde X_1=X_1+(X_2-VaR_{\alpha_2}(X_2))_+=X\). Now, part 1) simply follows from

$$ \begin{aligned} EPD(X_1;ES_{\alpha_1}(X_1))=&\int\limits_{ES_{\alpha_1}(X_1)}^{x_F} P(X_1>z)dz =\int\limits_{\min\{VaR_{\alpha_2}(X),ES_{\alpha_1}(X_1)\}}^{VaR_{\alpha_2}(X)} \bar F(x)\;dx \\ EPD(X_2;VaR_{\alpha_2}(X_2))=&\int\limits_{VaR_{\alpha_2}(X_2)}^{x_F} P(X_2>z)dz = \int\limits_{VaR_{\alpha_2}(X)}^{{x_F}} \bar F(x)\;dx. \end{aligned} $$

Parts ii) and iii), as well as the equality \(EPD(\tilde X_1;ES_{\alpha_1}(\tilde X_1))=EPD(X;ES_{\alpha_1}( X))\) are true since \(\tilde X_1=X\). Finally, the inequality at the end of the lemma follows from ES α_1(X 1) ≤ ES α_1(X).

2.6 Proof of Theorem 3.1

It is easy to obtain

$$ \begin{aligned} E(g[X;\gamma])=&E(I_1[X])+\gamma E(I_2[X]-I_2(VaR_{\alpha_2} (X)))_+ \\ =&E(I_1[X])+\gamma (1-\alpha_2) (ES_{\alpha_2} (I_2[X])-I_2(VaR_{\alpha_2} (X))), \end{aligned} $$

as a result of Eq. (2.1). Moreover,

$$ \begin{aligned} ES_{\alpha_1} (g[X;\gamma])=& ES_{\alpha_1} (I_1[X])+\gamma ES_{\alpha_1} (I_2[X]-I_2(VaR_{\alpha_2} (X)))_+ \\ =& ES_{\alpha_1} (X)-ES_{\alpha_1} (I_2[X]) +\gamma {{1-\alpha_2}\over {1-\alpha_1}} (ES_{\alpha_2} (I_2[X])-I_2(VaR_{\alpha_2} (X))) \end{aligned} $$

due to the co-monotonicity property and relation (2.1). Combining the last two equations, one may reduce our main problem to minimising

$$ h_1(\gamma) I_2(VaR_{\alpha_2} (X)) +(\lambda_1-\lambda_2) E(I_2[X])+ h_2(\gamma)ES_{\alpha_2} (I_2[X])-\lambda_1 ES_{\alpha_1} (I_2[X]), $$

where h 1(λ) = λ2 − h 2(λ) and \(h_2(\lambda)=\gamma (1-\alpha_2) (1+\lambda_1 \alpha_1/ (1-\alpha_1))\). Note that

$$ \gamma(1+{{\lambda_1 \alpha_1}\over {1-\alpha_1}})-\lambda_2-{{\lambda_1 \alpha_1}\over {1-\alpha_1}} $$

is negative if \(\gamma\in[0,\gamma_1)\) and positive if \(\gamma\in(\gamma_1,1]\).

Now, the solution for the first stage problem from part i) needs to increase as slowly as possible on \([0,VaR_{\alpha_1}(X)]\) and increase rapidly on \([VaR_{\alpha_1}(X),VaR_{\alpha_2}(X)]\). In addition, the optimal solution remains flat on \([VaR_{\alpha_2}(X),x_F]\) if γ1 < γ ≤ 1 and increases with slope 1 whenever 0 ≤ γ < γ1. All of these facts generate the optimal solution

$$ \left\{ \begin{array}{ll} \min\{(X-VaR_{\alpha_1}(X)+\xi_1)_+,\xi_2\} +(X-VaR_{\alpha_2}(X))_+,& 0\leq\gamma<\gamma_1,\\ \min\{(X-VaR_{\alpha_1}(X)+\xi_1)_+,\xi_2\},& \gamma_1<\gamma\leq1, \end{array}\right. $$

where \((\xi_1,\xi_2)\in\mathcal{A}_1\). Straightforward computations yield that the difference between the objective functions for the second stage problem corresponding to the two scenarios, is a constant with respect to ξ1 and ξ2. Therefore, both cases lead to the same solution, which is \((VaR_{\alpha_1}(X),VaR_{\alpha_1^*}(X))\). We only discuss the second case, γ1 < γ ≤ 1, for which the second stage optimisation problem is equivalent to minimising over \(\mathcal{A}_1\)

$$ H(\xi_1,\xi_2):=\lambda_2 \xi_2+(\lambda_1\!-\!\lambda_2)\!\int\limits_{VaR_{\alpha_1}(X)-\xi_1}^{VaR_{\alpha_1}(X)+\xi_2-\xi_1} \bar{F}(x)\;dx \!-\!\lambda_1\left(\xi_1\!+\!{{1}\over {1-\alpha_1}}\!\int\limits_{VaR_{\alpha_1}(X)}^{VaR_{\alpha_1}(X)+\xi_2-\xi_1} \bar{F}(x)\;dx\right). $$

Taking the derivative with respect to ξ2, we get that

$$ H(\xi_1,\xi_2)\geq H(\xi_1,\xi_1+VaR_{\alpha_1^*}(X)-VaR_{\alpha_1}(X)), \;{for\ all}\; (\xi_1,\xi_2)\in\cal{A}_1. $$

The right hand side function is increasing in ξ1, which justifies the first part of Theorem 3.1.

Part ii) is developed in the same manner as the previous case. The solution for the first stage problem requires a rapid variation on \([0,VaR_{\alpha_2}(X)]\) since λ1 − λ2 < 0 and

$$ \lambda_1-\lambda_2 -\lambda_1/(1-\alpha_1)=-\lambda_1 \alpha_1/(1-\alpha_1)-\lambda_2<0. $$

In addition, the optimal solution remains flat on \([VaR_{\alpha_2}(X),x_F]\), if γ1 < γ ≤ 1 and increases with slope 1 whenever 0 ≤ γ < γ1. Thus, our solutions are

$$ \left\{ \begin{array}{ll}\min\{X,\xi_1\}+\min\{(X-VaR_{\alpha_1}(X))_+,\xi_2-\xi_1\} +(X-VaR_{\alpha_2}(X))_+,& 0\leq\gamma<\gamma_1,\\ \min\{X,\xi_1\}+\min\{(X-VaR_{\alpha_1}(X))_+,\xi_2-\xi_1\},& \gamma_1<\gamma\leq1, \end{array}\right. $$

where \((\xi_1,\xi_2)\in\mathcal{A}_1\). Similarly, both cases lead to the same solution, which is (0,0). Let us justify this for the case in which γ1 < γ ≤ 1, where the second stage optimisation problem is equivalent to

$$ \begin{aligned} \min_{{\mathcal{A}}_1}\;H(\xi_1,\xi_2):=&\lambda_2 \xi_2+(\lambda_1-\lambda_2)\left(\int\limits_{0}^{\xi_1}+\int\limits_{VaR_{\alpha_1}(X)}^{VaR_{\alpha_1}(X)+\xi_2-\xi_1} \right) \bar{F}(x)\;dx \cr &-\lambda_1\left(\xi_1+{{1}\over {1-\alpha_1}}\int\limits_{VaR_{\alpha_1}(X)}^{VaR_{\alpha_1}(X)+\xi_2-\xi_1} \bar{F}(x)\;dx\right). \end{aligned} $$

Taking the derivative with respect to ξ2, we get that \(H(\xi_1,\xi_2)\geq H(\xi_1,\xi_1)\) for all \((\xi_1,\xi_2)\in\mathcal{A}_1,\) due to the fact that

$$ \bar{F}(VaR_{\alpha_1}(X)+\xi_2-\xi_1)\leq\bar{F}(VaR_{\alpha_1}(X))\leq 1-\alpha_1\leq {{\lambda_2}\over {\lambda_2+\lambda_1 \alpha_1/(1-\alpha_1)}}. $$

Finally, \(H(\xi_1,\xi_1)\) is increasing in ξ1, which completes the proof of Theorem 3.1.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Asimit, A.V., Badescu, A.M. & Tsanakas, A. Optimal risk transfers in insurance groups. Eur. Actuar. J. 3, 159–190 (2013). https://doi.org/10.1007/s13385-013-0068-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s13385-013-0068-6

Keywords

Navigation