Skip to main content
Log in

On the forward rate concept in multi-state life insurance

  • Published:
Finance and Stochastics Aims and scope Submit manuscript

Abstract

Similarly to the notion of modeling credit risk by using forward credit default spread rates, mortality risk in life insurance contracts is nowadays often modeled by using forward mortality (spread) rates. More recently, this concept has also been discussed for more complex life insurances that include multiple lives or intermediate states that correspond to the health status of the insured. For consistency purposes and for technical reasons, most authors assume that the underlying financial and demographic events are stochastically independent.

In the present paper, we study sufficient and necessary conditions under which general transition forward rates are indeed consistent with respect to the relevant insurance claims. This shows the theoretical limitations of the forward rate concept in life insurance. Our study is based on a model where the underlying financial and demographical developments are diffusion processes driven by a multivariate Brownian motion. This allows us to investigate independence properties by analyzing the asymptotic behavior of mixed (conditional) moments. In particular, we obtain that for joint life and disability insurance policies, some specific demographic events need to be dependent in order to ensure consistency.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

References

  1. Bauer, D., Börger, M., Ruß, J., Zwiesler, H.-J.: The volatility of mortality. Asia-Pac. J. Risk Insur. 3, 184–211 (2008)

    Google Scholar 

  2. Bauer, D., Börger, M., Ruß, J.: On the pricing of longevity-linked securities. Insur. Math. Econ. 46, 139–149 (2010)

    Article  MATH  Google Scholar 

  3. Bauer, D., Benth, F., Kiesel, R.: Modeling the forward surface of mortality. SIAM J. Financ. Math. 3, 639–666 (2012)

    Article  MATH  MathSciNet  Google Scholar 

  4. Bogachev, V.I.: Measure Theory, vol. I. Springer, Heidelberg (2007)

    Book  MATH  Google Scholar 

  5. Buchardt, K.: Dependent interest and transition rates in life insurance. Insur. Math. Econ. 55, 167–179 (2014)

    Article  MATH  MathSciNet  Google Scholar 

  6. Cairns, A., Blake, D., Dowd, K.: Pricing death: frameworks for the valuation and securitization of mortality risk. ASTIN Bull. 36, 79–120 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  7. Christiansen, M.C.: A joint analysis of financial and biometrical risks in life insurance. PhD thesis, University of Rostock (2007). urn:nbn:de:gbv:28-diss2007-0007-6, http://rosdok.uni-rostock.de

  8. Christiansen, M.C.: Gaussian and affine approximation of stochastic diffusion models for interest and mortality rates. Risks 1, 81–100 (2013)

    Article  Google Scholar 

  9. Cox, J., Ingersoll, J. Jr, Ross, S.: An intertemporal general equilibrium model of asset prices. Econometrica 53, 363–384 (1985)

    Article  MATH  MathSciNet  Google Scholar 

  10. Dahl, M.: Stochastic mortality in life insurance: market reserves and mortality-linked insurance contracts. Insur. Math. Econ. 35, 113–136 (2004)

    Article  MATH  MathSciNet  Google Scholar 

  11. Dahl, M., Møller, T.: Valuation and hedging of life insurance liabilities with systematic mortality risk. Insur. Math. Econ. 39, 193–217 (2006)

    Article  MATH  Google Scholar 

  12. Haberman, S., Pitacco, E.: Actuarial Models for Disability Insurance. Chapman & Hall, Boca Raton (1999)

    MATH  Google Scholar 

  13. Ikeda, N., Watanabe, S.: Stochastic Differential Equations and Diffusion Processes. North-Holland, Amsterdam (1989)

    MATH  Google Scholar 

  14. Kloeden, P.E., Platen, E.: Numerical Solution of Stochastic Differential Equations. Springer, Heidelberg (1995). 2. corr. print. edition

    Google Scholar 

  15. Milevsky, M., Promislow, S.: Mortality derivatives and the option to annuitise. Insur. Math. Econ. 29, 299–318 (2001)

    Article  MATH  MathSciNet  Google Scholar 

  16. Miltersen, K., Persson, S.: Is mortality dead? Stochastic forward force of mortality rate determined by no arbitrage. Working paper, Norwegian School of Economics and Business Administration (2005). Online version available at http://www.mathematik.uni-ulm.de/carfi/vortraege/downloads/DeadMort.pdf

  17. Norberg, R.: Forward mortality and other vital rates—are they the way forward? Insur. Math. Econ. 47, 105–112 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  18. Tappe, S., Weber, S.: Stochastic mortality models: an infinite-dimensional approach. Finance Stoch. 18, 209–248 (2014)

    Article  MATH  MathSciNet  Google Scholar 

Download references

Acknowledgements

Marcus Christiansen and Andreas Niemeyer acknowledge the financial support of the Deutsche Forschungsgemeinschaft (Research Grant: Regulation in the Financial Services Industry after the Crisis).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Andreas Niemeyer.

Appendix

Appendix

Lemma A.1

Under Assumption  3.3 and with ϵ and (X(t)) t≥0 as in this assumption, we define for a fixed t with 0≤tuT and for i=1,…,n the processes \(\widetilde{X}_{i}(u) := X_{i}(u)-X_{i}(t)\). Then there exist

  1. (a)

    a random variable Y with \(\mathbb{E}[|Y|^{2(1+\epsilon )}] < \infty \) and

    $$\begin{aligned} Y \ge\bigg| \prod\limits_{i=1}^n X_i(t_i)^{Z_i} \bigg| \mathrm{e} ^{- \sum_{i=1}^n \int_{T_i} X_i(u) \,\mathrm{d}u} \end{aligned}$$

    for all intervals T i ⊆[0,T ], Z i ∈{0,1} and t i ∈[0,T ] for i=1,…,n

and for all 0≤tT and \(x \in\mathbb{R}\) constants C(t,x) such that

  1. (b)

    \(\mathbb{E}_{\mathrm{Q}}^{(t,x)}[ | \int_{t}^{t+\varDelta }\widetilde{X}_{i}(u) \,\mathrm{d}u |^{k} ] \le C(t,x) \varDelta^{\frac{3}{2} k}\) for all Δ∈[0,T t] and \(k \in\mathbb{N}\).

  2. (c)

    \({\mathbb{E}_{\mathrm{Q}}^{(t,x)}[ \int_{t}^{t+\varDelta }\cdots\int_{t}^{s_{\ell-1}} \int _{t}^{s_{\ell}} { | \alpha_{i}(\tau,X_{i}(\tau)) - \alpha_{i}(\tau,X_{i}(t)) |} \,\mathrm{d} \tau\,\mathrm{d}s_{\ell}\cdots \mathrm{d}s_{1} ]}\hfill \le (\mathbb{E}_{\mathrm{Q}}^{(t,x)}[( \int_{t}^{t+\varDelta}\cdots \int_{t}^{s_{\ell-1}} \int _{t}^{s_{\ell}} { | \alpha_{i}(\tau,X_{i}(\tau)) - \alpha_{i}(\tau,X_{i}(t)) |} \,\mathrm{d} \tau\,\mathrm{d}s_{\ell}\cdots \mathrm{d}s_{1} )^{k}])^{\frac{1}{k}}\hfill \le C(t,x) \varDelta ^{\frac{3}{2} + \ell}\) for all \(\varDelta\in[0, T^{*}-t], \ell \in\mathbb{N}_{0}\).

  3. (d)

    \(\sqrt{\mathbb{E}_{\mathrm{Q}}^{(t,x)}[( \int _{t}^{t+\varDelta}\cdots\int_{t}^{s_{\ell -1}} \int_{t}^{s_{\ell}} \beta_{i}(\tau, X_{i}(\tau)) \,\mathrm {d}W_{i}(\tau) \,\mathrm{d} s_{\ell}\cdots \mathrm{d}s_{1} )^{2}]} \le C(t,x) \varDelta^{\frac{1}{2} + \ell}\) for all \(\varDelta\in[0, T^{*}-t], \ell\in\mathbb{N}_{0}\).

  4. (e)

    For all Δ∈[0,T t], let T i be an interval with T i ⊆[t,t+Δ], Z i ∈{0,1} and t i ∈[t,t+Δ] for all i=1,…,n. Then

    $$\begin{aligned} &C(t,x) \varDelta^{\frac{9}{2}} \\ &\quad\ge \mathbb{E}_\mathrm{Q}^{(t,x)}\bigg[\bigg| \mathrm {e}^{- \sum_{i=1}^n \int_{T_i} \widetilde{X}_i(u) \,\mathrm{d}u} \prod_{i=1}^n X_i(t_i)^{Z_i} \\ & \qquad{}- \bigg( \prod_{i=1}^n X_i(t_i)^{Z_i} \bigg) \bigg( 1 - \sum _{i=1}^n \int_{T_i} \widetilde{X}_i(u) \,\mathrm{d}u + \frac{1}{2} \bigg(\sum _{i=1}^n \int_{T_i} \widetilde{X}_i(u) \,\mathrm{d}u \bigg)^2 \bigg) \bigg|\bigg]. \end{aligned}$$
  5. (f)

    \(| \frac{1}{\mathbb{E}_{\mathrm{Q}}^{(t,x)}[\mathrm{e}^{-\int _{t}^{t+\varDelta}\widetilde{X}_{i}(\tau) \,\mathrm{d}\tau } ]} -1 - \mathbb{E}_{\mathrm{Q}}^{(t,x)}[\int_{t}^{t+\varDelta }\widetilde{X}_{i}(\tau) \,\mathrm{d}\tau ]\hfill + \frac{1}{2} \mathbb{E}_{\mathrm{Q}}^{(t,x)}[ ( \int _{t}^{t+\varDelta}\widetilde{X}_{i}(\tau) \,\mathrm{d}\tau)^{2} ] - ( \mathbb{E}_{\mathrm{Q}}^{(t,x)}[\int_{t}^{t+\varDelta }\widetilde{X}_{i}(\tau) \,\mathrm{d}\tau] )^{2} |\hfill \le C(t,x) \varDelta^{\frac{9}{2}}\) for sufficiently small Δ>0.

Proof

In the following, we write C(t,x)<∞ for any constant.

(a) A direct consequence from Hölder’s inequality for sums is the inequality

$$\begin{aligned} (|a_1|+ \cdots+ |a_n|)^p \le (|a_1|^p + \cdots+ |a_n|^p)n^{p-1} \quad \mbox{for } p\ge1 . \end{aligned}$$
(A.1)

For each i=1,…,n, we have with (A.1) and the SDE (3.1) that

$$\begin{aligned} |X_i(t_i)|^k &\le\sup_{t\in[0,T^*]} |X_i(t_i)|^k \\ &\le3^{k-1} \bigg( \sup_{t\in[0,T^*]} \bigg| \int_0^t \beta _i(u,X_i(u)) \,\mathrm{d}W_i(u) \bigg|^k \\ & \quad{}+ |X_i(0)|^k + \sup_{t\in[0,T^*]}\bigg| \int_0^t \alpha _i(u,X_i(u)) \,\mathrm{d}u\bigg|^k \bigg) . \end{aligned}$$

With this estimation, by applying Hölder’s inequality to the α-term and Doob’s inequality to the β-term, we get for k≥2 that

$$\begin{aligned} \mathbb{E}\bigg[ \sup_{t\in[0,T^*]} |X_i(t)|^k \bigg] &\le 3^{k-1} \bigg( \Big( \frac{k}{k-1}\Big)^k \mathbb{E}\bigg[ \bigg| \int_0^{T^*} \beta _i(u,X_i(u)) \,\mathrm{d}W_i(u) \bigg|^k \bigg] \\ & \quad {}+ |X_i(0)|^k + (T^*)^{k-1} \int_0^{T^*} \mathbb{E}\big[\big| \alpha_i\big(u,X_i(u)\big) \big|^k \,\mathrm{d}u \big] \bigg) < \infty, \end{aligned}$$

since X i (0) is deterministic. With Assumption 3.1(iii) for α i which is also true for the CIR process, we get

$$\begin{aligned} \int_0^{T^*} \mathbb{E}\big[ \big|\alpha_i\big(X_i(u),u\big)\big|^k \big] \,\mathrm{d}u &\le\int_0^{T^*} \mathbb{E}\big[ K_2^k |1+X_i(u)|^{\frac{k}{2}}\big] \,\mathrm{d}u \\ &\le\int_0^{T^*} K_2^k 2^{\frac{k}{2} -1} \big( 1 + \mathbb{E}[|X_i(u)|^k ]\big) \,\mathrm{d}u, \end{aligned}$$

and this term is finite since analogously to Proposition 3.2, \(\mathbb{E}[ |X_{i}(u)|^{k}]\) is finite. By adding and subtracting an α-term, using (A.1) and by the same arguments as above, we obtain

$$\begin{aligned} &\mathbb{E}\bigg[ \bigg| \int_0^{T^*} \beta_i(u,X_i(u)) \,\mathrm {d}W_i(u) \bigg|^k \bigg] \\ &\quad \le2^{k-1} \bigg( \mathbb{E}[ | X_i(T^*) |^k ] + \mathbb{E}\bigg[ \bigg| \int _0^{T^*} \alpha_i\big(u,X_i(u)\big) \,\mathrm{d}u \bigg|^k \bigg] \bigg) < \infty. \end{aligned}$$

In total, we get with Assumption 3.3(i) that

$$\begin{aligned} \bigg| \prod\limits_{i=1}^n X_i(t_i)^{Z_i} \bigg| \mathrm{e}^{- \sum_{i=1}^n \int_{T_i} \widetilde{X}_i(u) \,\mathrm{d}u} \le\prod\limits_{i=1}^n \Big(\sup_{t \in[0,T^*]} |X_i(t)| \Big)^{Z_i} W =: Y . \end{aligned}$$

This random variable Y is integrable, since with Hölder’s inequality for \(p=\frac{1+\epsilon}{\epsilon}\),

$$\begin{aligned} \mathbb{E}[|Y|] &\le\bigg( \mathbb{E}\bigg[ \bigg( \prod_{i=1}^n \Big(\sup_{t \in [0,T^*]} |X_i(t)| \Big)^{Z_i} \bigg)^{\frac{1+\epsilon}{\epsilon }}\bigg] \bigg)^{\frac{\epsilon}{1+\epsilon}} W^{1+\epsilon} \\ & \le\prod_{i=1}^n \bigg( \mathbb{E}\bigg[ \bigg( \Big(\sup_{t \in[0,T^*]} |X_i(t)| \Big)^{Z_i} \bigg)^{\frac{1+\epsilon}{\epsilon}n}\bigg]\bigg)^{\frac{\epsilon}{n(1+\epsilon)}} W^{1+\epsilon} < \infty \end{aligned}$$

by assumption and by the first part, where we distinguish the cases Z i =0 and Z i =1.

(b) First of all, we consider the case k>1. With Hölder’s inequality for p=k and Proposition 3.2, it follows that

$$\begin{aligned} \mathbb{E}_\mathrm{Q}^{(t,x)}\bigg[ \bigg| \int_t^{t+\varDelta }\widetilde{X}_i(u) \,\mathrm{d}u \bigg|^k \bigg] &\le\mathbb{E}_\mathrm{Q}^{(t,x)} \bigg[ \bigg( \int _t^{t+\varDelta}1 \,\mathrm{d}u \bigg)^{k-1} \bigg] \mathbb{E} _\mathrm{Q}^{(t,x)}\bigg[ \int_t^{t+\varDelta}| \widetilde{X}_i(u) |^k \,\mathrm{d}u \bigg] \\ & \le\varDelta^{k-1} \int_t^{t+\varDelta}C(t,x) \varDelta^{\frac{k}{2}}\,\mathrm{d}u = C(t,x) \varDelta^{\frac{3}{2} k} . \end{aligned}$$

The case k=1 also follows from the calculation above by skipping the second step.

(c) The first step is a direct consequence from Hölder’s inequality with parameter p=k. By applying +1 times Hölder’s inequality with p=k and using Assumption 3.1(ii), we get

$$\begin{aligned} &\bigg(\mathbb{E}_\mathrm{Q}^{(t,x)}\bigg[\Big( \underbrace{\int _t^{t+\varDelta}\cdots \int_t^{s_{\ell-1}}}_{\ell\,\mathrm{times}} \int_t^{s_\ell} \underbrace{ \left| \alpha_i\big(\tau,X_i(\tau)\big) - \alpha _i\big(\tau ,X_i(t)\big) \right|}_{\le K_1 | \widetilde{X}_i(\tau)|} \,\mathrm {d}\tau\,\mathrm{d} s_\ell\cdots \mathrm{d}s_1 \Big)^k\bigg]\bigg)^{\frac{1}{k}} \\ &\quad \le\Big( \varDelta^{(k-1)(\ell+1)} \int_t^{t+\varDelta}\cdots \int _t^{s_{\ell-1}} \int_t^{s_\ell} K_1^k \mathbb{E}_\mathrm {Q}^{(t,x)}\big[\big( | \widetilde{X}_i(\tau)| \big)^k \big] \,\mathrm{d}\tau\,\mathrm {d}s_\ell\cdots \mathrm{d}s_1 \Big)^{\frac{1}{k}} \\ &\quad \le\big(K_1^k C(t,x) \varDelta^{k\ell+ k\frac{3}{2}}\big)^{\frac{1}{k}}, \end{aligned}$$

using that \(\mathbb{E}_{\mathrm{Q}}^{(t,x)} [( |\widetilde {X}_{i}(\tau)|)^{k}] \le C(t,x) \varDelta^{\frac{k}{2}}\).

(d) By applying times Hölder’s inequality with p=2, the Itô isometry, Assumption 3.1(iii) and Proposition 3.2, we can show that the initial term is smaller than or equal to

$$\begin{aligned} &\sqrt{\varDelta^\ell\int_t^{t+\varDelta}\cdots\int_t^{s_\ell} \mathbb{E}_\mathrm{Q}^{(t,x)} \bigg[ \bigg( \int_t^{s_\ell} \beta_i\big(\tau, X_i(\tau)\big) \, \mathrm{d}W_i(\tau ) \bigg)^2 \bigg] \,\mathrm{d}s_\ell\cdots \mathrm{d}s_1} \\ & \quad \le\sqrt{\varDelta^\ell\int_t^{t+\varDelta}\cdots\int _t^{t+\varDelta}\int_t^s C(t,x) \,\mathrm{d} \tau\,\mathrm{d}s_\ell\cdots \mathrm{d}s_1} = \sqrt{C(t,x) \varDelta^{2\ell+1}} . \end{aligned}$$

(e) By a Taylor expansion, we have \(| \mathrm{e}^{-x} -1 + x - \frac{1}{2} x^{2} | \le\frac{1}{6} |x|^{3} \mathrm {e}^{-\xi}\) for ξ between 0 and x. Applying this approximation to \(x = \sum _{i=1}^{n} \int_{T_{i}} \widetilde{X}_{i}(u) \,\mathrm{d}u\) and using Hölder’s inequality with p=q=2 gives the inequality

$$\begin{aligned} &\mathbb{E}_\mathrm{Q}^{(t,x)} \bigg[\bigg| \mathrm{e}^{- \sum _{i=1}^n \int_{T_i} \widetilde {X}_i(u) \,\mathrm{d}u} \prod_{i=1}^n X_i(t_i)^{Z_i} \\ & \qquad{}-\bigg( \prod_{i=1}^n X_i(t_i)^{Z_i}\bigg) \bigg( 1 - \sum_{i=1}^n \int_{T_i} \widetilde{X}_i(u) \,\mathrm{d}u + \frac{1}{2} \bigg(\sum _{i=1}^n \int _{T_i} \widetilde{X}_i(u) \,\mathrm{d}u \bigg)^2 \bigg) \bigg|\bigg] \\ &\quad\le\frac{1}{6} \bigg( \mathbb{E}_\mathrm{Q}^{(t,x)} \bigg[ \bigg| \prod_{i=1}^n X_i(t_i)^{2 Z_i} \bigg| \bigg] \mathbb{E}_\mathrm{Q}^{(t,x)} \bigg[ \bigg| \bigg( \sum_{i=1}^n \int_{T_i} \widetilde{X}_i(u) \,\mathrm{d}u \bigg)^6 \mathrm{e}^{- 2\xi} \bigg| \bigg] \bigg)^{\frac{1}{2}} . \end{aligned}$$

We apply the generalization of Hölder’s inequality (cf. Corollary 2.11.5 in [4]) to the first factor with \(p_{i}=\frac{1}{n}\) and get that the expression from above is smaller than or equal to

$$\begin{aligned} & \frac{1}{6} \bigg( \prod_{i=1}^n \big( {\mathbb{E}_\mathrm{Q}^{(t,x)} \big[ |X_i(t_i)|^{2 n Z_i} \big]} \big)^{\frac{1}{n}} \mathbb{E}_\mathrm{Q}^{(t,x)} \bigg[\bigg| \bigg( \sum_{i=1}^n \int_{T_i} \widetilde{X}_i(u) \,\mathrm{d}u \bigg)^6 \mathrm{e}^{- 2\xi} \bigg| \bigg] \bigg)^{\frac{1}{2}} , \end{aligned}$$

where the first expression is bounded by (3.3). In the following, we use the inequality |eξ|≤1+e−2x for ξ between 0 and x and get that the expression is again smaller than or equal to

$$\begin{aligned} &\frac{\sqrt{C(t,x)}}{6} \bigg( \mathbb{E}_\mathrm{Q}^{(t,x)} \bigg[\bigg| \bigg( \sum _{i=1}^n \int_{T_i} \widetilde{X}_i(u) \,\mathrm{d}u \bigg)^6 \bigg|\bigg] \\ & \qquad{}+ \mathbb{E}_\mathrm{Q}^{(t,x)} \bigg[\bigg| \bigg( \sum_{i=1}^n \int_{T_i} \widetilde{X}_i(u) \,\mathrm{d}u \bigg)^6 \mathrm{e}^{- 2 \sum _{i=1}^n \int_{T_i} \widetilde{X}_i(u) \,\mathrm{d}u} \bigg|\bigg] \bigg)^{\frac{1}{2}} \\ &\quad\le\frac{\sqrt{C(t,x)}}{6} \Big( C(t,x) \varDelta^9 + C(t,x) \varDelta^9 \big(K(t,x)\big)^{\frac{1}{1+\epsilon}} \Big)^{\frac{1}{2}} , \end{aligned}$$

where we apply Hölder’s inequality to the second summand with the two parameters \(p= \epsilon_{0}^{-1}(1+\epsilon_{0})\) and q=1+ϵ 0, where ϵ 0 is chosen such that 0<ϵ 0ϵ and \(6 \frac{1+\epsilon_{0}}{\epsilon_{0}} \in\mathbb{N}\). The moments of the first two expected values are estimated by applying (b), and for the third expected value we plug in the definition of \(\widetilde {X}_{i}(u)\), take X i (t) out of the expected value and get by Assumption 3.3(i) that this expected value is finite.

(f) The Taylor approximation of second order gives us that

$$\begin{aligned} \mathrm{e}^{-\int_t^{t+\varDelta} \widetilde{X}_i(\tau) \,\mathrm {d}\tau} &= 1 - \int_t^{t+\varDelta}\widetilde{X}_i(\tau) \,\mathrm{d}\tau+ \frac{1}{2} \Big( - \int_t^{t+\varDelta}\widetilde {X}_i(\tau) \,\mathrm{d}\tau\Big)^2 \\ &\quad{}+\frac{1}{6} \Big( - \int_t^{t+\varDelta}\widetilde{X}_i(\tau ) \,\mathrm{d}\tau\Big)^3 \mathrm{e}^{\xi_1}, \end{aligned}$$

where ξ 1 is between 0 and \(-\int_{t}^{t+\varDelta}\widetilde {X}_{i}(\tau) \,\mathrm{d}\tau\). Again, we use the Taylor approximation of second order and obtain

$$\begin{aligned} \frac{1}{1+y} = 1- y + y^2 - \frac{y^3}{(1+\xi_2)^4} , \quad |y| < 1, \end{aligned}$$

where ξ 2 is between 0 and y. We define

$$\begin{aligned} y:\!&= - \mathbb{E}_\mathrm{Q}^{(t,x)} \bigg[\int_t^{t+\varDelta }\widetilde{X}_i(\tau) \,\mathrm{d}\tau\bigg] + \mathbb{E}_\mathrm{Q}^{(t,x)} \bigg[ \frac{1}{2} \bigg( - \int _t^{t+\varDelta}\widetilde{X}_i(\tau) \,\mathrm{d} \tau\bigg)^2\bigg] \\ &\quad{}+ \mathbb{E}_\mathrm{Q}^{(t,x)} \bigg[ \frac{1}{6} \bigg( - \int_t^{t+\varDelta}\widetilde{X}_i(\tau) \,\mathrm{d}\tau\bigg)^3 \mathrm{e}^{\xi_1} \bigg] \end{aligned}$$

and get in total that

$$\begin{aligned} \frac{1}{\mathbb{E}_\mathrm{Q}^{(t,x)}[\mathrm{e}^{-\int _t^{t+\varDelta}\widetilde{X}_i(\tau) \,\mathrm{d}\tau} ]} = 1 - y + y^2 - \frac{y^3}{(1+\xi_2)^4} . \end{aligned}$$

By using (b) and analogously to the proof of (e), we get

$$\begin{aligned} \bigg| \mathbb{E}_\mathrm{Q}^{(t,x)}\bigg[\bigg(\int _t^{t+\varDelta}\widetilde{X}_i(\tau) \,\mathrm{d}\tau \bigg)^3 \mathrm{e}^{\xi_1} \bigg] \bigg| \le C(t,x) \varDelta ^{\frac{9}{2}}. \end{aligned}$$

Analogously, we can show that \(|y| \le C(t,x) \varDelta^{\frac{3}{2}}\), and so we can make y arbitrarily small by choosing a small Δ. We get that |1+ξ 2|−4≤16 for all \(| \xi_{2} | \le|y| \le \frac{1}{2}\) and obtain the claim. □

Proposition A.2

For 0≤tTT , let the mapping F as defined in Sect2 have a continuous derivative \(f=\frac{\mathrm{d}}{\mathrm{d}T} F\) in the second argument and let there be integrable random variables W and \(\widetilde{W}\) with

$$\begin{aligned} | F(t,T,r,m) | \le W \quad\textit{and} \quad \bigg| \frac{\mathrm{d}}{\mathrm{d}T} F(t,T,r,m) \bigg| \le\widetilde{W} . \end{aligned}$$

Given that \(\mathbb{E}_{\mathrm{Q}}[F(t,T,r,m) | \mathcal{F}_{t}] = F(t,T,\rho,\mu)\), we get

$$\begin{aligned} \mathbb{E}_\mathrm{Q}\bigg[ \frac{\mathrm {d}}{\mathrm{d}T} F(t,T,r,m) \bigg| \mathcal{F}_{t}\bigg] = \frac{\mathrm{d}}{\mathrm{d}T} F(t,T,\rho,\mu) \quad \textit{a.s.} \end{aligned}$$

Proof

We have

$$\begin{aligned} F(t,T,r,m) = F(t,t,r,m) + \int_t^T \frac{d}{d u} F(t,u,r,m) \,\mathrm {d}u . \end{aligned}$$

From the definition of conditional expectations, we conclude for all \(A \in\mathcal{F}_{t}\) that

$$\begin{aligned} &\int_A \mathbb{E}_\mathrm{Q}[ F(t,T,r,m) | \mathcal{F}_{t}] \, \mathrm{d}\mathrm{Q}= \int_A F(t,T,r,m) \,\mathrm{d}\mathrm{Q} \\ &\quad= \int_A \bigg( F(t,t,r,m) + \int_t^T \frac{\mathrm {d}}{\mathrm{d}u} F(t,u,r,m) \,\mathrm{d}u \bigg) \,\mathrm{d}\mathrm{Q} \\ &\quad= \int_A \mathbb{E}_\mathrm{Q}[F(t,t,r,m) | \mathcal{F}_{t}] \,\mathrm{d}\mathrm{Q}+ \int_A \int_t^T \mathbb{E}_\mathrm {Q}\bigg[ \frac{\mathrm{d}}{\mathrm{d}u} F(t,u,r,m) \bigg| \mathcal {F}_{t}\bigg] \,\mathrm{d}u \,\mathrm{d}\mathrm{Q}. \end{aligned}$$

The second summand of the last line can be obtained by exchanging the order of integration with Fubini’s theorem, applying again the definition of conditional expectations and using again Fubini’s theorem. This is where we need the existence of the integrable upper bounds W and \(\widetilde{W}\). In total, we get for all Tt that

$$\begin{aligned} F(t,t,r,m) + \int_t^T \mathbb{E}_\mathrm{Q}\bigg[ \frac{\mathrm {d}}{\mathrm{d}u} F(t,u,r,m) \bigg| \mathcal{F}_{t}\bigg] \,\mathrm{d}u & = \mathbb{E}_\mathrm{Q}[ F(t,T,r,m) | \mathcal{F}_{t}] \\ & = F(t,T,\rho,\mu) \\ & = F(t,t,\rho,\mu) + \int_t^T f(t,u,\rho,\mu) . \end{aligned}$$

Since f is continuous, the claim holds. □

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Christiansen, M.C., Niemeyer, A. On the forward rate concept in multi-state life insurance. Finance Stoch 19, 295–327 (2015). https://doi.org/10.1007/s00780-014-0244-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00780-014-0244-9

Keywords

Mathematics Subject Classification (2010)

JEL Classification

Navigation