This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Optimizing dividends and capital injections limited by bankruptcy, and practical approximations for the Cramér-Lundberg process

F. Avram 111Corresponding author. E-mail: [email protected] Laboratoire de Mathématiques Appliquées, Université de Pau, France 64000 D. Goreac LAMA, Univ Gustave Eiffel, UPEM, Univ Paris Est Creteil, CNRS, F-77447 Marne-la-Valleée, France School of Mathematics and Statistics, Shandong University, Weihai, Weihai 264209, PR China R. Adenane Département des Mathématiques, Université Ibn-Tofail, Kenitra, Maroc 14000 U. Solon Laboratoire de Mathématiques Appliquées, Université de Pau, France 64000 Institute of Mathematics, University of the Philippines Diliman, Philippines
Abstract

The recent papers Gajek-Kucinsky(2017) and Avram-Goreac-Li-Wu(2020) investigated the control problem of optimizing dividends when limiting capital injections stopped upon bankruptcy. The first paper works under the spectrally negative Lévy model; the second works under the Cramér-Lundberg model with exponential jumps, where the results are considerably more explicit. The current paper has three purposes. First, it illustrates the fact that quite reasonable approximations of the general problem may be obtained using the particular exponential case studied in Avram-Goreac-Li-Wu(2020). Secondly, it extends the results to the case when a final penalty PP is taken into consideration as well besides a proportional cost k>1k>1 for capital injections. This requires amending the “scale and Gerber-Shiu functions" already introduced in Gajek-Kucinsky(2017). Thirdly, in the exponential case, the results will be made even more explicit by employing the Lambert-W function. This tool has particular importance in computational aspects and can be employed in theoretical aspects such as asymptotics.

Keywords: dividend problem, capital injections, penalty at default, scale functions, Lambert-W function, De Vylder-type approximations, rational Laplace transform

1 Introduction

This paper concerns the approximate optimization of a new type of boundary mechanism, which emerged recently in the actuarial literature [APY18, GK17, AGLW20], in the context of the optimal control of dividends and capital injections.

The model. Consider the spectrally negative Lévy risk model:

Xt=x+ctξ¯t,c0,X_{t}=x+ct-\bar{\xi}_{t},c\geq 0,

where ξ¯t\bar{\xi}_{t} is a spectrally positive Lévy process, with Lévy measure ν(x)dx\nu(x)dx. The classic example is that of the perturbed Cramèr-Lundberg risk model with

ξ¯t=i=1Ntξi+σBt,\bar{\xi}_{t}=\sum_{i=1}^{N_{t}}\xi_{i}+\sigma B_{t},

where BtB_{t} is a Brownian motion, where NtN_{t} is an independent Poisson process of intensity λ>0\lambda>0, and (ξi)i1\left(\xi_{i}\right)_{i\geq 1} is a family of i.i.d.r.v. whose distribution, density and moments are denoted respectively by F,f,mi,i{1,}F,f,m_{i},i\in\{1,...\}. Furthermore,

  • the process is modified by dividends and capital injection:

    π:=(D,I)Xtπ:=XtDt+It,\pi:=\left(D,I\right)\Rightarrow X_{t}^{\pi}:=X_{t}-D_{t}+I_{t},

    where D,ID,I are adapted, non-decreasing and càdlàg processes, with D0=I0=0D_{0-}=I_{0-}=0;

  • the first time when we do not bail-out to positive reserves σ0π:=inf{t>0:Xtπξ¯t+It<0}\sigma^{\pi}_{0}:=\inf\left\{t>0:\ X_{t-}^{\pi}-\bigtriangleup\bar{\xi}_{t}+\bigtriangleup I_{t}<0\right\} is called bankruptcy/absolute ruin;

  • prior to bankruptcy, dividends are limited by the available reserves: Dt:=DtDtXtπξ¯t+It, where ξ¯t:=i=1Ntξi\bigtriangleup D_{t}:=D_{t}-D_{t-}\leq X_{t-}^{\pi}-\bigtriangleup\bar{\xi}_{t}+\bigtriangleup I_{t},\textnormal{ where }\bar{\xi}_{t}:=\sum_{i=1}^{N_{t}}\xi_{i}. The set of “admissible" policies satisfying this constraint will be denoted by Π~(x)\tilde{\Pi}(x).

The objective is to maximize the profit:

Jxπ:=𝔼x[[0,σ0π]eqt(dDskdIs)Peqσ0π],k1,P0.J_{x}^{\pi}:=\mathbb{E}_{x}\left[\int_{\left[0,\sigma^{\pi}_{0}\right]}e^{-qt}\left(dD_{s}-kdI_{s}\right)-Pe^{-q\sigma^{\pi}_{0}}\right],k\geq 1,P\geq 0.

The value function is

V(x):=supπΠ~(x)Jxπ,x.V\left(x\right):=\sup_{\pi\in\tilde{\Pi}(x)}J_{x}^{\pi},\ x\in\mathbb{R}.

Motivation. The recent papers [GK17, AGLW20] investigated the above control problem of optimizing dividends and capital injections for processes with jumps, when bankruptcy is allowed as well. The second paper works under the Cramér-Lundberg model with exponential jumps, while the first works under the spectrally negative Lévy model, allowing also for the presence of Brownian motion and infinite activity jumps. It turns out that the optimal policy belongs to the class of (a,0,b),a>0,b0(-a,0,b),a>0,b\geq 0 “bounded buffer policies", which consist in allowing only capital injections smaller than a given aa and declaring bankruptcy at the first time when the size of the overshoot below 0 exceeds a,a, and in paying dividends when the reserve reaches an upper barrier bb. These will briefly be described as (a,0,b)(-a,0,b) policies from now on. Furthermore, the optimal (a,b)(a^{*},b^{*}) are the roots of one variable equations with explicit solutions related to the Lambert-W(right) function (ProductLog in Mathematica).

Below, our goal is to show numerically that exponential approximations provide quite reasonable results (as the de Vylder approximation provides for the ruin problem). We will focus in our examples on the case of matrix exponential jumps(which are known to be dense in the class of general nonnegative jumps, with even error bounds for completely monotone jumps being available [VAVZ14]), for two reasons. One is in order to highlight certain exact equations which are similar to their exponential versions, and which may at their turn be used to produce even more accurate approximations in the future, and, secondly, since numerical Laplace inversion for this class may easily tuned to have arbitrarily small errors.

History of the problem: The case of no capital injections (also characterized by k=k=\infty or absorption below 0) is the dividend problem posed by De Finetti [DF57, Ger69] where dividends are paid above barrier bb^{*} and a=0a^{*}=0 is imposed. “The challenge is to find the right compromise between paying early in view of the discounting or paying late in order not to reach ruin too early and thus profit from the positive safety loading for a longer time" [AAM20].

Forced injections and no bankruptcy at 0 (also characterized by a reflection at 0) is studied in Shreve [SLG84] where dividends are paid above barrier bb^{*} and a=a^{*}=\infty is imposed.

From Lokka, Zervos, [LZ08] we know that in the Brownian motion case, it is optimal to either always inject, if kkck\leq k_{c}, for some critical cost kck_{c} (i.e. use Shreve), or, stop at 0 (use De Finetti). We propose to call this the Lokka, Zervos alternative. The “proof" of this alternative starts by largely assuming it via a heuristically justified border Ansatz [LZ08, (5.2)]: max{V(0),V(0)k}=0 either V(0)=0 or V(0)=k.\max\left\{-V(0),\ V^{\prime}(0)-k\right\}=0\Longrightarrow\text{ either }V(0)=0\text{ or }V^{\prime}(0)=k.

Extensive literature on SLG forced bailouts (no bankruptcy) can be found at Avram et al., (2007) [APP07], Kulenko and Schmidli, (2008) [KS08], Eisenberg and Schmidli, (2011) [ES11], Pérez et al., (2018) [PYB18], Lindensjo, Lindskog (2019) [LL19], Noba et al., 2020 [NPY20].

Articles [GK17, AGLW20] are the only papers which relate declaring bankruptcy to the size of jumps, with general and exponential jumps, respectively. [GK17] deals also with the presence of Brownian motion and infinite activity jumps, by conditioning at the first draw-down time; the optimality proof is quite involved.

In [AGLW20], it is also shown that neither V(0)=0V(0)=0 nor V(0)=kV^{\prime}(0)=k are possible: the Lokka-Zervos alternative disappears, but another interesting alternative holds. Above a certain critical kck_{c} the optimal dividends  barrier switches from strictly positive to 0, and kck_{c} is related in (26) to the Lambert-W function .

The results of [GK17, AGLW20] may be divided in three parts:

  1. 1.

    Compute the value of bounded buffer policies. The key result is (5) below, an explicit determination of the objective J0=J0a,bJ_{0}=J_{0}^{a,b}, which allows optimizing it numerically.

    REMARK 1.

    Computing the value function is considerably simplified by the use of first passage recipes available for spectrally negative Lévy processes [AKP04, Kyp14, KKR13, AGVA19], which are built around two ingredients: the WqW_{q} and ZqZ_{q} qq-scale functions, defined respectively for x0,q0x\geq 0,q\geq 0 as:

    1. (a)

      the inverse Laplace transform of 1κ(s)q\frac{1}{\kappa(s)-q}, where κ(s)\kappa(s) is the Laplace exponent (which characterizes a Lévy process) and

    2. (b)

      Zq(x)=1+q0xWq(y)𝑑yZ_{q}(x)=1+q\int_{0}^{x}W_{q}(y)dy

    – see the papers [Sup76, Ber98, AKP04] for the first appearance of these functions. The name qq-scale/harmonic functions is justified by the fact that these functions are harmonic for the process XX killed upon entering (,0)(-\infty,0), in the sense that

    {eqmin[t,T0]Wq(Xmin[t,T0]),eqmin[t,T0]Zq(Xmin[t,T0])},t0\{e^{-q\min[t,T_{0}]}\;W_{q}(X_{\min[t,T_{0}]}),e^{-q\min[t,T_{0}]}\;Z_{q}(X_{\min[t,T_{0}]})\},t\geq 0

    are martingales, as shown in [Pis04, Prop. 3] (in the case of ZqZ_{q}, there is also a penalty of 11 at ruin, generalizing to other penalties produces the so-called Gerber-Shiu function).

  2. 2.

    Equations determining candidates for the optimal a,ba^{*},b^{*} are obtained by differentiating the objective (which is expressed in terms of the scale functions Wq,ZqW_{q},Z_{q}), and the optimal pair (a,b)(a^{*},b^{*}) is identified. As a result, the critical kck_{c} is related in (26) to the Lambert-W function.

  3. 3.

    The optimality of the (a,0,b)(-a^{*},0,b^{*}) policy is established.

Note that the last step is quite non-trivial and is achieved by different methods in [GK17] and [AGLW20]. The latter paper starts by formulating a (new) HJB equation associated to this stochastic control problem – see (8).

REMARK 2.

The objective may be optimized numerically using the first step only (the equation (10) for J0a,bJ_{0}^{a,b}).

Exponential approximations may also be used, which are similar in spirit with the de Vylder-type approximations. Recall that the philosophy of the de Vylder approximation  is to approximate a Cramèr-Lundberg process by a simpler process with exponential jumps, with cleverly chosen exponential rate μ\mu, and the parameters λ,c\lambda,c may also be modified, if one desires to make the approximations exact at x=0x=0 –see for example [AHPS19] for more details)

The efficiency of the de Vylder approximation for approximating ruin probabilities is well documented [DV78]. The natural question of whether this type of techniques may work for other objectives, like for example  for optimizing dividends and/or reinsurance  was already discussed in [Høj02, DD05, BDW07, GSS08, AHPS19]. In this paper, following on previous works [ACFH11, AP14, ABH18], we draw first the attention to the fact that we have not one, but three de Vylder-type approximations for Wq(x)W_{q}(x) (as for the ruin probability). The best approximation in our experiments when the loading coefficient θ\theta is large turn out to be the classic de Vylder approximation. However, for approximating near the origin, the two point Padé approximation which fixes both the values Wq(0)=1c,Wq(0)=q+λc2W_{q}(0)=\frac{1}{c},W_{q}^{\prime}(0)=\frac{q+\lambda}{c^{2}} works better. The end result here is simply replacing the inverse rate μ1\mu^{-1} by m1,m_{1}, in the formula for the scale function of the Cramèr-Lundberg process  with exponential jumps. In between x=0x=0 and xx\to\infty, the winner is sometimes the “Renyi approximation" which replaces the inverse rate by m22m1\frac{m_{2}}{2m_{1}}, and modifies λ\lambda as well (for the de Vylder approximation, μ1\mu^{-1} is replaced by m33m2\frac{m_{3}}{3m_{2}}, and both λ,c\lambda,c are modified).

We end this introduction by highlighting in figure 1 the fact that for exponential jumps, the limited capital injections objective function J0J_{0} given by (14) for arbitrary bb but optimal a=s(b)a=s(b) (via a complicated formula) improves the value function with respect to de Finetti and Shreve, Lehoczky and Gaver, for any bb.

Refer to caption
Figure 1:    The value function J0J_{0} given by (14), for arbitrary bb but optimal aa. The inequality observed is a consequence of the properties of the Lambert function. The improvement with respect to de Finetti is considerable, of 0.382292%0.382292\% (the SLG approach is not competitive in this case). Note also that the optimal barrier b=0.109023b=0.109023 is smaller than the de Finetti and SLG optima of 0.626672,1.827260.626672,1.82726 respectively.

Contents and contributions. Section 2 offers a conjectured profit formula for (a,0,b)(-a,0,b) policies, where we include also a final penalty PP. The theoretical result of the section 1 revisits [GK17, AGLW20] by linking the two formulations together and emphasizing the impact of the bankruptcy penalty PP (via the scale function GG). Its proof is beyond the scope of the present (already lengthy enough) paper and it can be inferred from either one of [GK17] and [AGLW20] through a three step argument:

  1. 1.

    express the cost by conditioning on the reserve (JxJ_{x}) starting from 0xb0\leq x\leq b hitting either 0 or b;

  2. 2.

    get a further relationship on costs JbJ_{b} and J0J_{0} by conditioning on the first claim;

  3. 3.

    finally, mix these conditions together in order to obtain the explicit formula for JxJ_{x}.

We also wish to point out (in section 2.1) the link to an appropriate HJB variational inequality equation8 and the definition 1 specifying the action regions and their computation starting from the regimes in the HJB system. To the best of our knowledge, this is new and the preliminary studies conducted on more complicated problems (involving reinsurance and reserve-dependent premium) seem to reinforce the relevance of this tool.

In section 7 we provide an alternative matrix exponential form of the exact cost, in the case of matrix exponential jumps.

An explicit determination of a,ba^{*},b^{*} and an equity cost dichotomy when dealing with exponential jumps are given in section 3, taking also advantage of properties of the Lambert-W function, which were not exploited before. The two main novelties of the section are:

  • emphasizing the computations of the optimal buffer/barrier (from [AGLW20]) in relation with the scale-like quantities appearing in [GK17];

  • making explicit use of the (computation-ready) Lambert-W function to describe the dependency of optimal aba^{*}b (in equation 14) and of the dichotomy-triggering cost kck_{c} in equation 26.

Again, a further novelty is the presence of the bankruptcy cost PP.

Section 4 reviews, for completeness, the de Vylder approximation-type approximations. Section 4.1 recalls, for warm-up, some of the oldest exponential approximations for ruin probabilities. Section 4.2 recalls in Proposition 5, following [AP14, AHPS19] three approximations of the scale function Wq(x)W_{q}(x) 444essentially, this is the “dividend function with fixed barrier”, which had been also extensively studied in previous literature before the introduction of Wq(x)W_{q}(x) , obtained by approximating its Laplace transform. These amount finally to replacing our process by one with exponential jumps  and cleverly crafted parameters based on the first three moments of the claims.

In section 5, we consider particular examples and obtain very good approximations for two fundamental objects of interest: the growth exponent Φq\Phi_{q} of the scale function Wq(x)W_{q}(x), and the (last) global minimum of Wq(x)W_{q}^{\prime}(x), which is fundamental in the de Finetti barrier problem. Proceeding afterwards to the problem of dividends and limited capital injections, concepts in section 3 are used to compute a straightforward exponential approximation based on an exponential approximation of the claim density, and a new “correct ingredients approximation" which consists of plugging into the objective function (10) for exponential claims the exact "non-exponential ingredients" (scale functions and, survival and mean functions) of the non-exponential densities. Both methods are observed to yield reasonable values in approximating the objective.

This leads us to our conclusion that from a practical point of view, exponential approximations are typically sufficient in the problems discussed in this paper.

2 The cost function of (a,0,b)(-a,0,b) policies, for the spectrally negative Lévy case

In this section, we allow ξ¯t\bar{\xi}_{t} to be a spectrally positive Lévy process, with a Lévy measure admitting a density ν(dy)=ν(y)dy\nu(dy)=\nu^{\prime}(y)dy. The simplest example is that of the perturbed Cramèr-Lundberg risk model with

ξ¯t=i=1Ntξi+σBt,\bar{\xi}_{t}=\sum_{i=1}^{N_{t}}\xi_{i}+\sigma B_{t},

where NtN_{t} is a Poisson process of intensity λ>0\lambda>0, (ξi)i1\left(\xi_{i}\right)_{i\geq 1} is an independent family of i.i.d.r.v. with density f(y)f(y), and BtB_{t} is an independent Brownian motion.

We revisit here the problem of optimizing the value of "bounded buffer (a,0,b)(-a,0,b) policies", following [GK17, AGLW20] (in order to relate the results, one needs to replace γ\gamma in the objective of [GK17] by 1/k1/k), while taking into account also the bankruptcy penalty PP.

An important role in the results will be played by the expected scale after a jump

C(x)=0xWq(xy)ν¯(y)𝑑y=cWq(x)Zq(x)+σ22Wq(x),C(x)=\int_{0}^{x}W_{q}(x-y)\overline{\nu}(y)dy=c{W_{q}(x)}-Z_{q}(x)+\frac{\sigma^{2}}{2}W_{q}^{\prime}(x), (1)

where ν¯(y)=yν(u)𝑑u\overline{\nu}(y)=\int_{y}^{\infty}\nu(u)du is the tail of the Lévy measure and σ\sigma is the Brownian volatility (the identity above follows easily from the qq-harmonicity of ZqZ_{q}, after an integration by parts of the convolution term and a division by qq).

The problem of limited reflection requires introducing a new "scale function Sa(x)S_{a}(x) and Gerber-Shiu function Ga,σ(x)G_{a,\sigma}(x)"– see Remark 4 for further comments on this terminology:

{Sa(x)=Zq(x)+Ca(x),Ca(x)=0xWq(xy)ν¯(a+y)𝑑yGa,σ(x)=Ga(x)+kσ22Wq(x)\begin{cases}S_{a}(x)=Z_{q}(x)+C_{a}(x),C_{a}(x)=\int_{0}^{x}W_{q}(x-y)\;\overline{\nu}(a+y)\;dy\\ G_{a,\sigma}(x)=G_{a}(x)+k\frac{\sigma^{2}}{2}W_{q}(x)\end{cases} (2)

where

Ga(x)=0xWq(xy)(kma(y)+Pν¯(a+y))𝑑y:=kMa(x)+PCa(x),\displaystyle G_{a}(x)=\int_{0}^{x}W_{q}(x-y)\left(k\;m_{a}(y)+P\overline{\nu}(a+y)\right)dy:=kM_{a}(x)+PC_{a}(x),
ma(y)=0azν(y+z)𝑑z.\displaystyle m_{a}(y)=\int_{0}^{a}z\nu(y+z)dz.
EXAMPLE 1.

With exponential jumps and possibly σ>0\sigma>0, using the identities

ν¯(y+a)=eμaν¯(y),ma(y)=λeμym(a),m(a)=0ayμeμy𝑑y=1eμaμaeμa,\overline{\nu}(y+a)=e^{-\mu a}\overline{\nu}(y),m_{a}(y)=\lambda e^{-\mu y}m(a),m(a)=\int_{0}^{a}y\;\mu e^{-\mu y}dy=\frac{1-e^{-\mu a}}{\mu}-ae^{-\mu a},

we find that the functions (2) are expressible as products of C(x)C(x) and the survival or mean function of the jumps:

{Ca(x)=C(x)eμa=C(x)F¯(a),F¯(a)=1F(a)Sa(x)=Zq(x)+eμaC(x)Ga(x)=(km(a)+PF¯(a))C(x)\begin{cases}C_{a}(x)=C(x)e^{-\mu a}=C(x)\overline{F}(a),\;\overline{F}(a)=1-F(a)\\ S_{a}(x)=Z_{q}(x)+e^{-\mu a}C(x)\\ G_{a}(x)=\left(km(a)+P\overline{F}(a)\right)C(x)\end{cases} (3)

([GK17] use sc,rcs_{c},r_{c}, instead of Ma(x):=0xWq(xy)ma(y)𝑑y,Ca(x)M_{a}(x):=\int_{0}^{x}W_{q}(x-y)\;m_{a}(y)\;\;dy,C_{a}(x), respectively). When P=0=σP=0=\sigma, these reduce to quantities in [AGLW20].

The formulas above will be used below as a heuristic approximation in non-exponential  cases.

REMARK 3.

Note that

Ca(0)=0,Ga(0)=0,Sa(0)=1,C(0)=0,C(0)={λcσ=00σ>0,C_{a}(0)=0,G_{a}(0)=0,S_{a}(0)=1,C(0)=0,C^{\prime}(0)=\begin{cases}\frac{\lambda}{c}&\sigma=0\\ 0&\sigma>0\end{cases}, (4)

and that C(x),Ga(x),Sa(x)C(x),G_{a}(x),S_{a}(x) are increasing functions in xx.

We state now a generalization  of [GK17, Thm. 4] for the value function J0a,bJ_{0}^{a,b} of (a,0,b)(-a,0,b) policies, in terms of Sa(x),Ga(x)S_{a}(x),G_{a}(x). In the Cramèr-Lundberg case illustrated below, the proof is straightforward, following [AGLW20]. In the other case, one needs to adapt the proof of [GK17].

THEOREM 1.

Cost function for (a,b)(a,b) policies For a spectrally negative Lévy processes, let

Jx=Ja,b(x):=𝔼x[0Taeqt(dDtkdIt)PeqTa]J_{x}=J^{a,b}(x):=\mathbb{E}_{x}\left[\int_{0}^{T_{-a}}e^{-qt}\left(\mathrm{d}D_{t}-k\;\mathrm{d}I_{t}\right)-Pe^{-qT_{-a}}\right]

denote the expected discounted dividends minus capital injections associated to policies consisting in paying capital injections with proportional cost k1k\geq 1, provided that the severity of ruin is smaller than a>0a>0, and paying dividends as soon as the process reaches some upper level bb. Put

Ga,σ(x)=Ga(x)+kσ22Wq(x).\displaystyle G_{a,\sigma}(x)=G_{a}(x)+k\frac{\sigma^{2}}{2}W_{q}(x).

Then, it holds that

Jx={Ga,σ(x)+J0a,bSa(x)=Ga,σ(x)+1Ga,σ(b)Sa(b)Sa(x),x[0,b]kx+J0a,bx[a,0]0xa.J_{x}=\begin{cases}G_{a,\sigma}(x)+J_{0}^{a,b}S_{a}(x)=G_{a,\sigma}(x)+\frac{1-G_{a,\sigma}^{\prime}(b)}{S_{a}^{\prime}(b)}S_{a}(x),&x\in[0,b]\\ kx+J_{0}^{a,b}&x\in[-a,0]\\ 0&x\leq-a\end{cases}. (5)
REMARK 4.

The first equality in (5) will be easily obtained by applying the strong Markov property at the stopping time T=min[T0,,Tb,+]T=\min[T_{0,-},T_{b,+}], but it still contains the unknown J0J_{0}.

This relation suggests a definition of the scale SaS_{a} and the Gerber-Shiu function Ga,σG_{a,\sigma}, as the coefficient of J0J_{0} and the part independent of J0J_{0}, respectively.

This equality is also equivalent to

J0a,b=JxGa,σ(x)Sa(x)=1Ga,σ(b)Sa(b),J_{0}^{a,b}=\frac{J_{x}-G_{a,\sigma}(x)}{S_{a}(x)}=\frac{1-G_{a,\sigma}^{\prime}(b)}{S_{a}^{\prime}(b)}, (6)

which suggests another analytic definition of the scale and Gerber-Shiu function corresponding to an objective JxJ_{x} which involves reflection at bb.

The functions Sa(x),Ga(x)S_{a}(x),G_{a}(x) may be shown to stay the same for problems which require only modifying the boundary condition at bb, like the problem of capital injections for the process reflected at bb, or the problem of dividends  for the process reflected at bb, with proportional retention kDk_{D} (this is in coherence with previously studied problems).

COROLLARY 2.

Let us consider the Cramèr-Lundberg  setting without diffusion (i.e. σ=0\sigma=0), For fixed k1,b0k\geq 1,b\geq 0, the optimality equation aJ0a,b=0\frac{\partial}{\partial a}J_{0}^{a,b}=0 may be written as

J0a,b=kaPJaa,b=P.J_{0}^{a,b}=ka-P\Leftrightarrow J_{-a}^{a,b}=-P. (7)
REMARK 5.

The first equality in (7) provides a relation between the objective J0J_{0} and the variable aa; the second recognizes this as the smooth fit equation Ja=0J_{-a}=0.

Proof: Recalling the expressions of J0a,bJ_{0}^{a,b}, Ga(x)G_{a}(x), in (6), in (2), and from [GK17, Lem. A.4]

Ma(x)=aCa(x),M_{a^{\prime}}(x)=-aC_{a^{\prime}}(x),

where Ca(x),Ma(x)C_{a^{\prime}}(x),M_{a^{\prime}}(x) denote derivatives with respect to the subscript aa, Whenever b>0b>0, if aa achieves the maximum in J0a,bJ^{a,b}_{0}, it is straightforward (think of the economic interpretation) that aa achieves the maximum of aJxa,ba\mapsto J^{a,b}_{x} for every x[0,b]x\in[0,b]. Therefore, we find

aJ0a,b=0J0a,b=Ga(x)Ca(x)=kMa(x)PCa(x)Ca(x)=kaPJaa,b=J0a,bkaJaa,b=P.\displaystyle\begin{aligned} &\frac{\partial}{\partial a}J_{0}^{a,b}=0\Leftrightarrow J_{0}^{a,b}=\frac{-G_{a^{\prime}}(x)}{C_{a^{\prime}}(x)}=\frac{-kM_{a^{\prime}}(x)-PC_{a^{\prime}}(x)}{C_{a^{\prime}}(x)}=ka-P\\ &\Leftrightarrow J_{-a}^{a,b}=J_{0}^{a,b}-ka\Leftrightarrow J_{-a}^{a,b}=-P.\end{aligned}

2.1 The HJB System

The optimality proof in [AGLW20] is based on showing that the function JxJ_{x} (5) with a,ba^{*},b^{*} defined in (18), (19) is the minimal AC-supersolution of the HJB system

{max{H(x,V,V(x)),1V(x),V(x)k}=0,x+max{V(x)k,PV(x)}=0,x,\left\{\begin{split}&\max\left\{{H}\left(x,{V},{V}^{\prime}(x)\right),1-{V}^{\prime}(x),{V}^{\prime}(x)-k\right\}=0,\ \forall x\in\mathbb{R}_{+}\\ &\max\left\{{V}^{\prime}(x)-k,-P-V(x)\right\}=0,\ \forall x\in\mathbb{R}_{-}\end{split}\right., (8)

where the Hamiltonian H{H} is given by

H(x,ϕ,v):=cv+λ+ϕ(xy)μeμy(q+λ)ϕ(x).{H}\left(x,\phi,v\right):=cv+\lambda\int_{\mathbb{R}_{+}}\phi(x-y)\mu e^{-\mu y}-\left(q+\lambda\right)\phi(x). (9)

To discuss (8), it is useful to introduce the concept of dividend-limited injections strategies and barrier strategies. The following are also valid for its generalizations to mixed singular/continuous controls taking into account reinsurance:

DEFINITION 1.

Dividend-limited injections strategies are stationary strategies where the dividends are paid according to a partition of the state space {\mathbb{R}} in five sets 𝒜,,𝒞,𝒞0,𝒟\mathcal{A},\mathcal{B},\mathcal{C},\mathcal{C}_{0},\mathcal{D} as follows:

  1. 1.

    If the surplus is in 𝒜\mathcal{A} (absolute ruin), bankruptcy is declared and a penalty PP is paid;

  2. 2.

    If the surplus is in \mathcal{B} bailouts/capital injections are used for bringing the surplus to the closest point of 𝒞\mathcal{C}.

  3. 3.

    If the surplus is in the open set 𝒞\mathcal{C} (continuation/no action set), no controls are used.

  4. 4.

    If the current surplus is in 𝒞0𝒟\mathcal{C}_{0}\subset\mathcal{D} (these are upper-accumulation points of 𝒞\mathcal{C}), dividends are paid at a positive rate, in order to keep the surplus process from moving.

  5. 5.

    If the current surplus is in 𝒟\mathcal{D}, a positive amount of money is paid as dividends in order to bring the surplus process to 𝒞0\mathcal{C}_{0}.

    Barrier strategies are stationary strategies for which 𝒜,,𝒞,𝒟\mathcal{A},\mathcal{B},\mathcal{C},\mathcal{D} are four consecutive intervals.

REMARK 6.

The four sets 𝒜,,𝒞,𝒟\mathcal{A},\mathcal{B},\mathcal{C},\mathcal{D} correspond to the cases when equality in the HJB equation (8) is attained by at least one of the operators VP,Vk,(𝒢q)V,-V-P,V^{\prime}-k,({\mathcal{G}}-q)V, and 1V,1-V^{\prime}, respectively. Note this generalizes [AM14, Ch. 5.3], where only the last two operators are considered.

REMARK 7.

One may conjecture that dividend-limited injections strategies are of a (recursive) multi-band nature. In the case of exponential jumps, [AGLW20] show that the four sets 𝒜,,𝒞,𝒟\mathcal{A},\mathcal{B},\mathcal{C},\mathcal{D} in the optimal solution are intervals, denoted respectively by (,a),[a,0],(0,b),[b,)(-\infty,-a),[-a,0],(0,b),[b,\infty).

Cheap equity corresponds the case when 𝒞=\mathcal{C}=\emptyset, and the partition reduces to three sets.

3 Explicit determination of a,ba^{*},b^{*} when F(x):=1eμx,P>cqF(x):=1-e^{-\mu x},P>-\frac{c}{q}

In this section we turn to the exponential case, where explicit formulas for the optimizers a,ba^{*},b^{*} are available. In particular, we will take advantage of properties of the Lambert-W function, which were not exploited in [AGLW20]. Subsequently, in sections 5, 6 we will show that exponential approximations work typically excellently in the general case. Although these results have already been established in [AGLW20], the present formulations have two achievements:

  1. 1.

    allow an unified formulation of [AGLW20] and [GK17] (via the previously introduced scale functions);

  2. 2.

    make use of a numerical tool (Lambert-W function) to express the optimal quantities of interest a,ba^{*},b^{*}.

3.1 The simplified cost function and optimality equations

PROPOSITION 3.

Cost function and optimality equations in the exponential  case

  1. 1.
    J0a,b=1C(b)(km(a)+PF¯(a))(F¯(a))C(b)+qWq(b)=γ(b)km(a)PF¯(a)qθ(b)+F¯(a),J_{0}^{a,b}=\frac{1-C^{\prime}(b)\left(k\;m(a)+P\overline{F}(a)\right)}{(\overline{F}(a))C^{\prime}(b)+qW_{q}(b)}=\frac{{\gamma(b)}-k\;m(a)-P\overline{F}(a)}{q\theta(b)+\overline{F}(a)}, (10)

    where we put

    γ(b)=1C(b),θ(b)=Wq(b)C(b).\gamma(b)=\frac{1}{C^{\prime}(b)},\theta(b)=\frac{W_{q}(b)}{C^{\prime}(b)}.
  2. 2.

    Put

    j(b):=γ(b)qθ(b).j(b):=\frac{\gamma^{\prime}(b)}{q\theta^{\prime}(b)}. (11)

    For fixed a0a\geq 0, the optimality equation bJ0a,b=0\frac{\partial}{\partial b}J_{0}^{a,b}=0 may be written as

    J0a,b=j(b).J_{0}^{a,b}=j(b). (12)
  3. 3.

    For fixed k1k\geq 1 and b0b\geq 0, at critical points with a(b)=a(k,P)(b)0a(b)=a^{(k,P)}(b)\neq 0 satisfies aJ0a(b),b=0\frac{\partial}{\partial a}J_{0}^{a(b),b}=0 we must have

    [J0a,b(kaP)]a=a(b)=0.\left[J_{0}^{a,b}-(ka-P)\right]_{a=a(b)}=0.

    Explicitly,

    0=η(b,a):=γ(b)θ(b)kμθ(b)F(a)q(kaP).0=\eta(b,a):=\frac{\gamma(b)}{\theta(b)}-\frac{k}{\mu\theta(b)}F\left(a\right)-q\left(ka-P\right). (13)
  4. 4.

    When PcqP\geq-\frac{c}{q} and b0b\geq 0 is fixed, the solution of (13) may be expressed in terms of the principal value of the “Lambert-W(right)" function (an inverse of L(z)=zezL(z)=ze^{z})

    [e1,)L0(z),z[1,)[-e^{-1},\infty)\ni L_{0}(z),z\in[-1,\infty)

    [CGH+96, Boy98, BFS08, Pak15, VLSHGG+19] (this observation is missing in [AGLW20]).

    (0,)a(b)=μ1(h(b)+L0(eh(b)qθ(b))),h(b)=h(b,P)=1qθ(b)μk(γ(b)qθ(b)+P)\displaystyle(0,\infty)\ni a(b)=\mu^{-1}\left(-h(b)+L_{0}\left(\frac{e^{h(b)}}{q\theta(b)}\right)\right),\quad h(b)=h(b,P)=\frac{1}{q\theta(b)}-\frac{\mu}{k}\left(\frac{\gamma(b)}{q\theta(b)}+P\right) (14)

    It follows that

    J0a(b),b=kμ(h(b)+L0(1qθ(b)eh(b)))P.J_{0}^{a(b),b}=\frac{k}{\mu}\left(-h(b)+L_{0}\left(\frac{1}{q\theta(b)}e^{h(b)}\right)\right)-P. (15)
  5. 5.

    In the special case b=0b=0, (13) implies that a=a(k,P)=a(k,P)(0)a=a^{(k,P)}=a^{(k,P)}(0) satisfies the simpler equation

    0=δk,P(a):=λη(0,a)=c~k(aq+λμ(1eμa)),c~=c+qP>0,\displaystyle 0=\delta_{k,P}(a):=\lambda\eta(0,a)=\widetilde{c}-k\left(aq+\frac{\lambda}{\mu}(1-e^{-\mu a})\right),\;\widetilde{c}=c+qP>0, (16)

    with solution

    μa(k,P)=g+L0(λqeg)>0,g=h(0)=λqμkqc~.\displaystyle\mu\;a^{(k,P)}=-g+L_{0}\left(\frac{\lambda}{q}e^{g}\right)>0,\;g=h(0)=\frac{\lambda}{q}-\frac{\mu}{kq}\widetilde{c}. (17)
  6. 6.

    At a critical point (a,b),a>0,b>0(a^{*},b^{*}),a^{*}>0,b^{*}>0, we must have  both J0a,b=j(b)=kaPJ_{0}^{a^{*},b^{*}}=j(b^{*})=ka^{*}-P\Longrightarrow

    a=s(b),s(b):=j(b)+Pk,a^{*}=s(b^{*}),s(b):=\frac{j(b)+P}{k}, (18)

    and

    0=η(b),η(b):=η(b,s(b))=γ(b)θ(b)qj(b)kμθ(b)F(j(b)+Pk)=0.0=\eta(b^{*}),\quad\eta(b):=\eta(b,s(b))=\frac{\gamma(b)}{\theta(b)}-qj(b)-\frac{k}{\mu\theta(b)}F\left(\frac{j(b)+P}{k}\right)=0. (19)
  7. 7.

    The equation 0=η(b)0=\eta(b) may be solved explicitly for PP, yielding

    P=kμlog[1+qθ(b)j(b)γ(b)kμ]j(b).P=-\frac{k}{\mu}\log\left[1+\frac{q\theta(b)j(b)-\gamma(b)}{\frac{k}{\mu}}\right]-j(b). (20)

Proof: 1. follows from Theorem 1.1.

2. Let M(b),N(b)M(b),N(b) denote the numerator an denominator of J0a,b:=M(b)N(b)J_{0}^{a,b}:=\frac{M(b)}{N(b)} in (10). The optimality equation bJ0a,b=N(b)N(b)(M(b)N(b)J0a,b)=0\frac{\partial}{\partial b}J_{0}^{a,b}=\frac{N^{\prime}(b)}{N(b)}\left(\frac{M^{\prime}(b)}{N^{\prime}(b)}-J_{0}^{a,b}\right)=0 simplifies to

J0a,b=M(b)N(b)=γ(b)qθ(b)=j(b).J_{0}^{a,b}=\frac{M^{\prime}(b)}{N^{\prime}(b)}=\frac{{\gamma^{\prime}(b)}}{q\theta^{\prime}(b)}=j(b).

3. (13) is a consequence of 1 and of the smooth fit result Corollary 2.

4. See the proof of the particular case 5; a(0,)a\in(0,\infty) holds since Pcqh(b)<1qθ(b).P\geq-\frac{c}{q}\Longrightarrow h(b)<\frac{1}{q\theta(b)}.

5. (16) follows from Wq(0)=1c,θ(0)=λ1.W_{q}(0)=\frac{1}{c},\theta(0)=\lambda^{-1}. To get (17), rewrite the equation (16) as zez=λqeg,z=μa+gze^{z}=\frac{\lambda}{q}e^{g},z=\mu a+g; a(0,)a\in(0,\infty) holds since Pcqg<λq.P\geq-\frac{c}{q}\Longrightarrow g<\frac{\lambda}{q}.

6. follows from 2. and .3.

7. is straightforward.

REMARK 8.

Note that the de Finetti  and Shreve, Lehoczky and Gaver solutions a=a(b)={0a^{*}=a(b^{*})=\begin{cases}0\\ \infty\end{cases} are always non-optimal, when PcqP\geq-\frac{c}{q} (see (14)).

However, as kk\rightarrow\infty, h(b)1qθ(b)0h(b)\rightarrow\frac{1}{q\theta(b)}\neq 0 and, a(b)=μ1(h(b)+L0(h(b)eh(b)))=0a(b)=\mu^{-1}\left(-h(b)+L_{0}\left(h(b)e^{h(b)}\right)\right)=0. This suffices to infer that you get de Finetti case.

On the other hand,

Ph(b)a(b){J0a(b),bγ(b)k/μqθ(b)=1kC(b)/μqWq(b)=J0,SLG(b)η(b)q(J0,kSLG(b)j(b)),b>0.{P\to\infty}\Longrightarrow h(b)\to-\infty\Longrightarrow a(b)\to\infty\Longrightarrow\begin{cases}J_{0}^{a(b),b}\to\frac{\gamma(b)-k/\mu}{q\theta(b)}=\frac{1-kC^{\prime}(b)/\mu}{qW_{q}(b)}=J_{0,SLG}(b)\\ \eta(b)\to q\left(J_{0,k}^{SLG}(b)-j(b)\right)\end{cases},\forall b>0. (21)

Thus, these regimes can be recovered asymptotically. Let now bk,S,bP,Db_{k}^{*,S},b_{P}^{*,D} denote the unique roots of η(b)=0\eta(b)=0 in the two asymptotic cases, which coincide with the classic Shreve, Lehoczky and Gaver and de Finetti barriers.

Then, it may be checked that bmin[bk,S,bP,D]b^{*}\leq\min[b_{k}^{*,S},b_{P}^{*,D}].

3.2 Existence of the roots of the equations η(b)=0,δk,P=0\eta(b)=0,\delta_{k,P}=0

The following (new) result discusses the existence of the roots of the equations η(b)=0,δk,P=0\eta(b)=0,\delta_{k,P}=0 introduced in proposition (3) and relates them to the Lambert-W function.

PROPOSITION 4.
  1. 1.

    θ\theta increases from θ(0)=1/cλ/c=1λ\theta(0)=\frac{1/c}{\lambda/c}=\frac{1}{\lambda} to θ()=1cΦqq\theta(\infty)=\frac{1}{c\Phi_{q}-q}, as we see it in the figure below.

    Refer to caption
    Figure 2: Plot of θ\theta with θ(0)=2\theta(0)=2 and θ()=22.8743\theta(\infty)=22.8743, for μ=2,c=3/4,λ=1/2,q=1/10\mu=2,\ c=3/4,\ \lambda=1/2,\ q=1/10, P=1P=1 and k=3/2k=3/2.

    γ\gamma is increasing-decreasing (from cλ\frac{c}{\lambda} to 0), with a maximum at the unique root of C′′(x)=0C^{\prime\prime}(x)=0 given by

    b¯:=1Φqρlog(ρ2Φq2),\bar{b}:=\frac{1}{\Phi_{q}-\rho_{-}}\log\left(\frac{\rho_{-}^{2}}{\Phi_{q}^{2}}\right), (22)

    where Φq,ρ\Phi_{q},\rho_{-} denote the positive and negative roots of the Cramèr-Lundberg  equation κ(s)=0\kappa(s)=0.

    The figure below illustrates the plot of the function γ\gamma and j(b)j({b}) in which the b¯\bar{b} is represented by the black point.

    Refer to caption
    Figure 3: Plots of j(b)j({b}) and γ(b)\gamma({b}) with b¯=2.5046\bar{b}=2.5046 and j(0)=4.5j(0)=4.5, for μ=2,c=3/4,λ=1/2,q=1/10\mu=2,\ c=3/4,\ \lambda=1/2,\ q=1/10, P=1P=1 and k=3/2k=3/2.

    If cμ(q+λ)>0c\mu-\left(q+\lambda\right)>0, then b¯>0\bar{b}>0 defined in (22) is the unique positive root of j(b)j({b}) and η(b¯)=1Wq(b¯)>0\eta\left(\bar{b}\right)=\frac{1}{W_{q}\left(\bar{b}\right)}>0. See the figure (4)

    Refer to caption
    Figure 4: For μ=2,c=3/4,λ=1/2,q=1/10\mu=2,\ c=3/4,\ \lambda=1/2,\ q=1/10, P=1P=1 and k=3/2k=3/2, the root of η(b)=0\eta({b})=0 is at b=0.469843b=0.469843

    The function j(b)=γ(b)qθ(b)j(b)=\frac{\gamma^{\prime}(b)}{q\theta^{\prime}(b)} is nonnegative  and decreasing to 0 on [0,b¯][0,\bar{b}], with

    j(0)=λμqC′′(0)(C(0))2=cμ(q+λ)μq.j(0)=\frac{\lambda}{\mu q}\frac{-C^{\prime\prime}(0)}{(C^{\prime}(0))^{2}}=\frac{c\mu-\left(q+\lambda\right)}{\mu q}. (23)

    See the figure (3).

  2. 2.

    Put

    δk,P:=δk,P(a(k,P))=δk,P(j(0)+Pk)=δk,P(c~μ(q+λ)kμq)=λ+qλk(1ec~μλqqk)μ,\delta_{k,P}:=\delta_{k,P}(a^{(k,P)})=\delta_{k,P}(\frac{j(0)+P}{k})=\delta_{k,P}(\frac{\widetilde{c}\mu-\left(q+\lambda\right)}{k\mu q})=\frac{\lambda+q-\lambda k\left(1-e^{-\frac{\widetilde{c}\mu-\lambda-q}{qk}}\right)}{\mu}, (24)

    and assume

    limkδk,P=λ+qμλ(μc~(λ+q))qμ=(λ+q)2λμc~qμ<0c~μ>λ1(λ+q)2.{\underset{k\rightarrow\infty}{\lim}\delta_{k,P}=\frac{\lambda+q}{\mu}-\frac{\lambda\left(\mu\widetilde{c}-(\lambda+q)\right)}{q\mu}=\frac{\left(\lambda+q\right)^{2}-\lambda\mu\widetilde{c}}{q\mu}<0}\Leftrightarrow\widetilde{c}\mu>\lambda^{-1}\left(\lambda+q\right)^{2}. (25)

    Then, P>cq\forall P>-\frac{c}{q}, the function δk,P\delta_{k,P} is decreasing in kk with δ1,P>0\delta_{1,P}>0, and has a unique root

    kc=kc(P):=q+λλff+L0(fef)>q+λλ,k_{c}=k_{c}(P):=\frac{q+\lambda}{\lambda}\frac{f}{f+L_{0}\left(-fe^{-f}\right)}>\frac{q+\lambda}{\lambda}, (26)

    where

    f:=λq+λc~μ(λ+q)q>1c~μ>λ1(λ+q)2P>Pl:=q1(μ1λ1(λ+q)2c)f:=\frac{\lambda}{q+\lambda}\frac{\widetilde{c}\mu-\left(\lambda+q\right)}{q}>1\Leftrightarrow\widetilde{c}\mu>\lambda^{-1}\left(\lambda+q\right)^{2}\Leftrightarrow P>P_{l}:=q^{-1}\left(\mu^{-1}\lambda^{-1}\left(\lambda+q\right)^{2}-c\right) (27)

    (note that the denominator f+L0(fef)f+L_{0}\left(-fe^{-f}\right) does not equal 0 since f>1f>1 and L0L_{0} takes always values bigger than 1-1; or, note that f=L1(L(f))-f=L_{-1}\left(L(-f)\right), where L1L_{-1} is the other real branch of the Lambert function).

    Furthermore,

    δk,P<0k>kc(P).\delta_{k,P}<0\Leftrightarrow k>k_{c}(P). (28)
  3. 3.

    It follows that η(b)=0\eta(b)=0 has at least one solution of in (0,b¯](0,\bar{b}] iff

    η(0)=cλ1λ(cq+λμ)kμF(a(k,P))=1λμ(λ+qλkF(a(k,P)))=1λδk,P<0k>kc.\eta(0)=\frac{c}{\lambda}-\frac{1}{\lambda}\left(c-\frac{q+\lambda}{\mu}\right)-\frac{k}{\mu}F\left(a^{(k,P)}\right)=\frac{1}{\lambda\mu}\left(\lambda+{q}-\lambda kF\left(a^{(k,P)}\right)\right)=\frac{1}{\lambda}\delta_{k,P}<0\Leftrightarrow k>k_{c}. (29)

    The first such solution will be denoted by bb^{*}.

Proof: For 1. see [AGLW20, Proof of Theorem 11, A2].

2. By using the assumption c~μ>λ1(λ+q)2\widetilde{c}\mu>\lambda^{-1}(\lambda+q)^{2} we get c~μλ+q\widetilde{c}\mu\geq\lambda+q, and k[1,)δk,Pk\in[1,\infty)\rightarrow\delta_{k,P} is decreasing.

Put d=c~μ(λ+q)qd=\frac{\widetilde{c}\mu-\left(\lambda+q\right)}{q}. The inequality δk,P<0\delta_{k,P}<0 (see (29)) may be reduced to

edk<1q+λλk1<edk(1(q+λ)dλdk):=ez(1z/f).\displaystyle e^{-\frac{d}{k}}<1-\frac{q+\lambda}{\lambda k}\Leftrightarrow 1<e^{\frac{d}{k}}\left(1-\frac{(q+\lambda)}{d\lambda}\frac{d}{k}\right):=e^{z}\left(1-z/f\right).

Rewriting the latter as f>ez(zf)-f>e^{z}\left(z-f\right) we recognize, by putting z=y+fz=y+f, an inequality reducible to yey<fef.ye^{y}<-fe^{-f}. The solution is

y<L0(fef),y<L_{0}\left(-fe^{-f}\right),

where L0L_{0} is the principal branch of the Lambert-W function.

The final solution is (28), where we may note that the variables k,Pk,P have been separated.

3. is straightforward.

REMARK 9.

The function [1,)fff+L0(fef)(1,)[1,\infty)\ni f\mapsto\frac{f}{f+L_{0}\left(-fe^{-f}\right)}\in(1,\infty) blows up at f=1f=1, and converges to 11 when ff\to\infty (or when either μ\mu or c~=c+qP\widetilde{c}=c+qP are large enough) as may be noticed in the figure below, which blows up at the value Pl:=4/5.P_{l}:=-4/5. Note also that when ff (or one of c,P,μc,P,\mu are large enough), kck_{c} given by (26) stabilizes to the equilibrium λ+qλ=6/5\frac{\lambda+q}{\lambda}=6/5; this is related to [APP07, Lemma 2], [KS08, Lemma 7], who obtain the same condition for b=0b^{*}=0 (without buffering capital injections). Intuitively, under these conditions, buffering is not crucial.

At the other end, as ff tends to its lower limit and to the regime B, the notion of equity expensiveness vanishes, and kck_{c}\to\infty.

Refer to caption
Figure 5:    kck_{c} as function of PP, for several values of cc, with the vertical asymptote at PlP_{l} fixed.

The next two figures illustrate how kck_{c} blows up at the critical values ql:=(1/2)(2λ+Pλμ+λμ4c4Pλ+P2λμ)q_{l}:=(1/2)(-2\lambda+P\lambda\mu+\sqrt{\lambda}\sqrt{\mu}\sqrt{4}c-4P\lambda+P^{2}\lambda\mu) and λl:=(cμ(cμ+μ(P)q+2q)24q2+μPq2q)\lambda_{l}:=\left(c\mu-\sqrt{(-c\mu+\mu(-P)q+2q)^{2}-4q^{2}}+\mu Pq-2q\right) (represented by red points in the figures below). The dark (blue) parts correspond to the regime AA.

Refer to caption
(a) kck_{c} defined in (26)as function of qq, for μ=2,c=3/2,λ=1\mu=2,\ c=3/2,\ \lambda=1 and P=1P=1; ql=2q_{l}=\sqrt{2}.
Refer to caption
(b) kck_{c} as function of λ\lambda, for μ=2,c=3/2,q=1/10\mu=2,\ c=3/2,\ q=1/10 and P=1P=1.
Figure 6: kck_{c} as a function of qq, λ\lambda

4 Which exponential approximation?

4.1 Three de Vylder-type exponential approximations for the ruin probability 

In the simplest case of exponential jumps  of rate μ\mu and σ=0\sigma=0, the formula for the ruin probability is

Ψ(x)=Px[t0:Xt<0]=11+θexp(xθμ1+θ)=11+θexp(xθm111+θ),\Psi(x)=P_{x}[\exists t\geq 0:X_{t}<0]=\frac{1}{1+{\theta}}\ exp\left(-\frac{x{\theta}\mu}{1+{\theta}}\right)=\frac{1}{1+{\theta}}\ exp\left(-\frac{x{\theta}m_{1}^{-1}}{1+{\theta}}\right), (30)

where θ=cλm1λm1\theta=\frac{c-\lambda m_{1}}{\lambda m_{1}} is the loading coefficient. By plugging the correct mean of the claims in the second formula yields the simplest approximation for processes with finite mean claims.

More sophisticated is the Renyi exponential approximation

ΨR(x)=11+θexp(xθm^211+θ),m^2=m22m1;\Psi_{R}(x)=\frac{1}{1+{\theta}}\ exp\left(-\frac{x{\theta}\widehat{m}_{2}^{-1}}{1+{\theta}}\right),\widehat{m}_{2}=\frac{m_{2}}{2m_{1}}; (31)

This formula can be obtained as a two point Padé approximation  of the Laplace transform, which conserves also the value Ψ(0)=(1+θ)1\Psi(0)=(1+\theta)^{-1} [AP14]. It may be also derived heuristically from the first formula in (30), via replacing μ\mu by the correct “excess mean" of the excess/severity density

fe(x)=F¯(x)m1=1F(x)m1,f_{e}(x)=\frac{\overline{F}(x)}{m_{1}}=\frac{1-F(x)}{m_{1}},

which is known to be m^2\widehat{m}_{2}. Heuristically, it makes more sense to approximate fe(x)f_{e}(x) instead of the original density f(x)f(x), since fe(x)f_{e}(x) is a monotone function, and also an important component of the Pollaczek-Khinchine formula for the Laplace transform Ψ^(s)=0esxF(dx)\widehat{\Psi}(s)=\int_{0}^{\infty}e^{-sx}F(dx) – see [Ram92, AP14].

More moments are put to work in the de Vylder approximation

ΨDV(x)=11+θ~exp(xθ~m^311+θ~),m^3:=m33m2,λ~=9m232m32λ,c~=cλm1+λ~m^3,θ~=2m1m33m22θ=m^3m^2θ.\Psi_{DV}(x)=\frac{1}{1+\widetilde{\theta}}\ exp\left(-\frac{x\widetilde{\theta}\widehat{m}_{3}^{-1}}{1+\widetilde{\theta}}\right),\;\widehat{m}_{3}:=\frac{m_{3}}{3m_{2}},\;\widetilde{\lambda}=\frac{9m_{2}^{3}}{2m_{3}^{2}}\lambda,\;\;\widetilde{c}=c-\lambda m_{1}+\widetilde{\lambda}\widehat{m}_{3},\;\widetilde{\theta}=\frac{2m_{1}m_{3}}{3m_{2}^{2}}\theta=\frac{\widehat{m}_{3}}{\widehat{m}_{2}}\theta. (32)

Interestingly, the result may be expressed in terms of the so-called "normalized moments"

m^i=miimi1{\widehat{m}_{i}}=\frac{m_{i}}{i\;m_{i-1}} (33)

introduced in [BHT05].

The de Vylder approximation parameters above may be obtained either from

  1. 1.

    equating the first three cumulants of our process to those of a process with exponentially distributed claim sizes of mean m^3\widehat{m}_{3}, and modified λ,c\lambda,c [DV78] (however p=cλm1=E0[X1]p=c-\lambda m_{1}=E_{0}[X_{1}] must be conserved, since this is the first cumulant), or

  2. 2.

    a Padé approximation of the Laplace transform of the ruin probabilities [ACFH11].

The second derivation via Padé shows that higher order approximations may be easily obtained as well. They might not be admissible, due to negative values, but packages for “repairing" the non-admissibility are available – see for example [DcSA16].

The first derivation of the de Vylder approximation is a process approximation (i.e., independent of the problem considered); as such, it may be applied to other functionals of interest besides ruin probabilities(Wq(x)W_{q}(x), dividend barriers, etc), simply by plugging the modified parameters in the exact formula for the ruin probability of the simpler process.

4.2 Three two point Padé approximations of the Laplace transform  W^q\widehat{W}_{q} of scale function

The simplest approximations for the scale function Wq(x)W_{q}(x) will now be derived heuristically from the following example.

EXAMPLE 2.

The Cramér-Lundberg model with exponential jumps Consider the Cramér-Lundberg model with exponential jump sizes with mean 1/μ1/\mu, jump rate λ\lambda, premium rate c>0c>0, and Laplace exponent κ(s)=s(cλμ+s)\kappa(s)=s\left(c-\frac{\lambda}{\mu+s}\right). Solving κ(s)q=0cs2+s(cμλq)qμ=0\kappa(s)-q=0\Leftrightarrow cs^{2}+s(c\mu-\lambda-q)-q\mu=0 for ss yields two distinct solutions γ20γ1=Φq\gamma_{2}\leq 0\leq\gamma_{1}=\Phi_{q} given by

γ1(μ,λ,c)=γ1=\displaystyle\gamma_{1}(\mu,\lambda,c)=\gamma_{1}= 12c((μcλq)+(μcλq)2+4μqc),\displaystyle\frac{1}{2c}\left(-\left(\mu c-\lambda-q\right)+\sqrt{\left(\mu c-\lambda-q\right)^{2}+4\mu qc}\right),
γ2(μ,λ,c)=γ2=\displaystyle\gamma_{2}(\mu,\lambda,c)=\gamma_{2}= 12c((μcλq)(μcλq)2+4μqc).\displaystyle\frac{1}{2c}\left(-\left(\mu c-\lambda-q\right)-\sqrt{\left(\mu c-\lambda-q\right)^{2}+4\mu qc}\right).

The WW scale function is:

Wq(x)=A1eγ1xA2eγ2xc(γ1γ2)W^q(s)=s+μcs2+s(cμλq)qμ,W_{q}(x)=\frac{A_{1}e^{\gamma_{1}x}-A_{2}e^{\gamma_{2}x}}{c(\gamma_{1}-\gamma_{2})}\Leftrightarrow\widehat{W}_{q}(s)=\frac{s+\mu}{cs^{2}+s(c\mu-\lambda-q)-q\mu}, (34)

where A1=μ+γ1,A2=μ+γ2A_{1}=\mu+\gamma_{1},A_{2}=\mu+\gamma_{2}.

Furthermore, it is well-known and easy to check that the function Wq(x)W_{q}^{\prime}(x) is in this case unimodal with global minimum at

bDeF=1γ1γ2{log(γ2)2A2(γ1)2A1=log(γ2)2(μ+γ2)(γ1)2(μ+γ1)if Wq′′(0)<0(q+λ)2cλμ<00if Wq′′(0)0(q+λ)2cλμ0,b_{DeF}=\frac{1}{\gamma_{1}-\gamma_{2}}\begin{cases}\log\frac{(\gamma_{2})^{2}A_{2}}{(\gamma_{1})^{2}A_{1}}=\log\frac{(\gamma_{2})^{2}(\mu+\gamma_{2})}{(\gamma_{1})^{2}(\mu+\gamma_{1})}\quad&\text{if $W_{q}^{\prime\prime}(0)<0\Leftrightarrow(q+\lambda)^{2}$}-c\lambda\mu<0\\ 0&\text{if $W_{q}^{\prime\prime}(0)\geq 0\Leftrightarrow(q+\lambda)^{2}-c\lambda\mu\geq 0$}\end{cases}, (35)

since Wq′′(0)=(γ1)2(μ+γ1)(γ2)2(μ+γ2)c(γ1γ2)=(q+λ)2cλμc3W_{q}^{\prime\prime}(0)=\frac{(\gamma_{1})^{2}(\mu+\gamma_{1})-(\gamma_{2})^{2}(\mu+\gamma_{2})}{c(\gamma_{1}-\gamma_{2})}=\frac{(q+\lambda)^{2}-c\lambda\mu}{c^{3}} and that the optimal strategy for the de Finetti problem is the barrier strategy at level bDeFb_{DeF} (see for example [APP07], [AGVA19, Sec. 3]).

Plugging now the respective parameters of the de Vylder type approximations in the exact formula (34) for the Cramèr-Lundberg process  with exponential claims, we obtain three approximations for W^q\widehat{W}_{q}:

  1. 1.

    “Naive exponential" approximation obtained by plugging μ1m1\mu^{-1}\rightarrow m_{1} in (34) (as was done, for a different purpose) in (30)

  2. 2.

    Renyi 444This is called DeVylder B) method in [GSS08, (5.6-5.7)], since it is the result of fitting the first two cumulants of the risk process. , obtained by plugging μ1m^2,λRλm1m^2\mu^{-1}\rightarrow\widehat{m}_{2},\lambda_{R}\rightarrow\lambda\frac{m_{1}}{\widehat{m}_{2}} (since cc is unchanged, the latter equation is equivalent to the conservation of ρ=λm1c,\rho=\frac{\lambda m_{1}}{c}, and to the conservation of θ\theta, so this coincides with the Renyi ruin approximation used in (31).)

  3. 3.

    De Vylder, obtained by plugging μ1m^3,λ~λ9m232m32,c~=cλm1+λ~m^3\mu^{-1}\rightarrow\widehat{m}_{3},\widetilde{\lambda}\rightarrow\lambda\frac{9m_{2}^{3}}{2m_{3}^{2}},\widetilde{c}=c-\lambda m_{1}+\widetilde{\lambda}\widehat{m}_{3}.

REMARK 10.

In the case of exponential claims, these three approximations are exact, by definition (or check that for exponential claims all the normalized moments are equal to μ1\mu^{-1}).

REMARK 11.

The conditions for the non-negativity of the barrier is Wq′′(0+)<0(λ+qc)2<λcf(0)W_{q}^{\prime\prime}(0_{+})<0\Leftrightarrow(\frac{\lambda+q}{c})^{2}<\frac{\lambda}{c}f(0). Here, this condition is satisfied  for the exact when θ>(λ+q)2(1ρ)λ2f(0)m1\theta>\frac{(\lambda+q)^{2}(1-\rho)}{\lambda^{2}f(0)m_{1}}.

It is shown in [AHPS19, Prop. 1] that the three de Vylder type approximations are two-point Padé approximations of the Laplace transform  (hence higher order generalizations are immediately available).

We recall that two-point Padé approximations incorporate into the Padé approximation   two initial values of the function (which can be derived easily via the initial value theorem, from the Pollaczek-Khinchine Laplace transform):

Wq(0+)=limssW^q(s)=1c,\displaystyle W_{q}(0_{+})=\lim_{s\to\infty}s\widehat{W}_{q}(s)=\frac{1}{c}, (36)
Wq(0+)=limss(sκ(s)qWq(0+))=q+λc2.\displaystyle W_{q}^{\prime}(0_{+})=\lim_{s\to\infty}s\left(\frac{s}{\kappa(s)-q}-W_{q}(0_{+})\right)=\frac{q+\lambda}{c^{2}}. (37)

In our case, incorporating both Wq(0+),Wq(0+)W_{q}(0_{+}),W_{q}^{\prime}(0_{+}) leads to the natural exponential approximation which is therefore the best near x=0x=0. Incorporating none of them yields the de Vylder approximation, which is the best asymptotically. Incorporating only Wq(0+)W_{q}(0_{+}) leads to Renyi, which is expected to be the best in an intermediate regime.

Note that when the jump distribution has a density ff, it holds that : 666This equation is important in establishing the nonnegativity of the optimal dividends barrier.

Wq′′(0+)\displaystyle W_{q}^{\prime\prime}(0_{+}) =limss(s(sκ(s)qWq(0+))Wq(0+))=1c((λ+qc)2λcf(0)).\displaystyle=\lim_{s\to\infty}s\left(s\left(\frac{s}{\kappa(s)-q}-W_{q}(0_{+})\right)-W_{q}^{\prime}(0_{+})\right)=\frac{1}{c}\Big{(}(\frac{\lambda+q}{c})^{2}-\frac{\lambda}{c}f(0)\Big{)}. (38)

Thus, Wq′′(0)W_{q}^{\prime\prime}(0) already requires knowing fC(0)f_{C}(0) (which is a rather delicate task starting from real data); therefore we will not incorporate into the Padé approximation   more than two initial values of the function.

We recall below in Proposition 5 three types of two-point Padé approximations [AHPS19, Prop. 1], and particularize them to the case when the denominator degree is n=2n=2 (which are further illustrated below).

PROPOSITION 5.

Three matrix exponential approximations for the scale function.

  1. 1.

    To secure both the values of Wq(0)W_{q}(0) and Wq(0)W_{q}^{\prime}(0), take into account (36) and (37), i.e. use the Padé approximation

    W^q(s)i=0n1aisicsn+i=0n1bisi,an1=1,bn1=can2λq.\widehat{W}_{q}(s)\sim\frac{\sum_{i=0}^{n-1}a_{i}s^{i}}{cs^{n}+\sum_{i=0}^{n-1}b_{i}s^{i}},a_{n-1}=1,b_{n-1}=ca_{n-2}-\lambda-q.

    For n=2n=2 we recover the “natural exponential " approximation of plugging μ1m1\mu\to\frac{1}{m_{1}} in (34):

    W^q(s)1m1+scs2+s(cm1λq)qm1,\widehat{W}_{q}(s)\sim\frac{\frac{1}{m_{1}}+s}{cs^{2}+s\left(\frac{c}{m_{1}}-\lambda-q\right)-\frac{q}{m_{1}}}, (39)

    used also (for a different purpose) in (30).

  2. 2.

    To ensure only Wq(0)=1cW_{q}(0)=\frac{1}{c}, we must use the Padé approximation

    W^q(s)i=0n1aisicsn+i=0n1bisi,an1=1.\widehat{W}_{q}(s)\sim\frac{\sum_{i=0}^{n-1}a_{i}s^{i}}{cs^{n}+\sum_{i=0}^{n-1}b_{i}s^{i}},a_{n-1}=1.

    For n=2n=2, we find

    W^q(s)2m1m2+scs2+s(2cm12λm12m2q)m22m1qm2=1m^2+scs2+s(cm^2λm1m^2q)qm^2,\widehat{W}_{q}(s)\sim\frac{\frac{2m_{1}}{m_{2}}+s}{cs^{2}+\frac{s\left(2cm_{1}-2\lambda m_{1}^{2}-m_{2}q\right)}{m_{2}}-\frac{2m_{1}q}{m_{2}}}=\frac{\frac{1}{\widehat{m}_{2}}+s}{cs^{2}+{s\left(\frac{c}{\widehat{m}_{2}}-\lambda\frac{m_{1}}{\widehat{m}_{2}}-q\right)}-\frac{q}{\widehat{m}_{2}}}, (40)

    where m^2=m^2=m22m1\widehat{m}_{2}=\widehat{m}_{2}=\frac{m_{2}}{2m_{1}} is the first moment of the excess density fe(x)f_{e}(x). Note that it equals the scale function of a process with exponential claims of rate m^21\widehat{m}_{2}^{-1} and with λ\lambda modified to λR=λm1m^2\lambda_{R}=\lambda\frac{m_{1}}{\widehat{m}_{2}}. Since cc is unchanged, the latter equation is equivalent to the conservation of ρ=λm1c,\rho=\frac{\lambda m_{1}}{c}, and to the conservation of θ\theta, so this coincides with the Renyi approximation 444This is called DeVylder B) method in [GSS08, (5.6-5.7)], since it is the result of fitting the first two cumulants of the risk process. used in (31).

  3. 3.

    The pure Padé approximation   yields for n=2n=2

    W^q(s)s+3m2m3s2(cλm1+λ3m222m3)+s(c3m2m33m1m2m3λq)3m2m3q\displaystyle\widehat{W}_{q}(s)\sim\frac{s+\frac{3m_{2}}{m_{3}}}{s^{2}\left(c-\lambda m_{1}+\lambda\frac{3m_{2}^{2}}{2m_{3}}\right)+s\left(c\frac{3m_{2}}{m_{3}}-\frac{3m_{1}m_{2}}{m_{3}}\lambda-q\right)-\frac{3m_{2}}{m_{3}}q}
    =s+1m^3c~s2+s(c~1m^3λ~q)1m^3q,c~=cλm1+λ~m^3,λ~=λ9m232m32.\displaystyle=\frac{s+\frac{1}{\widehat{m}_{3}}}{\widetilde{c}s^{2}+s\left(\widetilde{c}\frac{1}{\widehat{m}_{3}}-\widetilde{\lambda}-q\right)-\frac{1}{\widehat{m}_{3}}q},\;\widetilde{c}=c-\lambda m_{1}+\widetilde{\lambda}\widehat{m}_{3},\widetilde{\lambda}=\lambda\frac{9m_{2}^{3}}{2m_{3}^{2}}.

    Note that both the coefficient of s2s^{2} in the denominator coincides with the coefficient c~\widetilde{c} in the classic de Vylder approximation, since λ~m^3=λ9m232m32m33m2=λ3m222m3\widetilde{\lambda}\widehat{m}_{3}=\lambda\frac{9m_{2}^{3}}{2m_{3}^{2}}\frac{m_{3}}{3m_{2}}=\lambda\frac{3m_{2}^{2}}{2m_{3}}, and so does the coefficient of ss, since

    c3m2m33m1m2m3λ=c~1m^3λ~=(cλm1+λ~m^3)1m^3λ~.c\frac{3m_{2}}{m_{3}}-\frac{3m_{1}m_{2}}{m_{3}}\lambda=\widetilde{c}\frac{1}{\widehat{m}_{3}}-\widetilde{\lambda}=\left(c-\lambda m_{1}+\widetilde{\lambda}\widehat{m}_{3}\right)\frac{1}{\widehat{m}_{3}}-\widetilde{\lambda}.

5 Examples of computations involving scale function and dividend value approximations

Our goal in this section is to investigate whether exponential approximations are precise enough to yield reasonable estimates for quantities important in control like

  1. 1.

    the dominant exponent Φq\Phi_{q} of Wq(x)W_{q}(x)

  2. 2.

    the last local minimum of Wq(x)W_{q}^{\prime}(x), bDeFb_{DeF}, which yields, when being the global minimum, the optimal De Finetti barrier

  3. 3.

    Wq′′(0)W^{\prime\prime}_{q}(0), which determines if bDeF=0b_{DeF}=0

  4. 4.

    the functional J0J_{0} yielding the maximum dividends with capital injections.

All the examples considered involve a Cramèr-Lundberg  model with rational Laplace transform W^q(s)\widehat{W}_{q}(s) (since in this case, the computation of Wq,ZqW_{q},Z_{q} is fast and in principle arbitrarily large precision may be achieved with symbolic algebra systems).

  1. 1.

    For the first three problems, we will use de Vylder type approximations. Graphs of WqW^{\prime}_{q}, Wq′′W^{\prime\prime}_{q} and some tables summarizing the simulation results will be presented. We note that in most of the cases that we observed, the de Vylder approximation of Φq\Phi_{q} deviates from the exact value the least – see for example Table 2. For the De Finetti barrier, the "winner" depends on the size of bDeFb_{DeF}. Unsurprisingly, when near 0, the natural exponential approximation wins, and as bDeFb_{DeF} increases, Renyi and subsequently the de Vylder approximation take the upper hand – see for example Table 3.

  2. 2.

    For the computation of J0J_{0}, we provide, besides the exact value, also two approximations:

    1. (a)

      For a given density of claims ff one computes an exponential density approximation fe(x)=1m1exp(xm1)f_{e}(x)=\frac{1}{m_{1}}\text{exp}(-\frac{x}{m_{1}}) where m1m_{1} is the first moment of ff. Subsequently, WW, ZZ, J0J_{0} and a,ba,b are obtained using the exponential approximation fef_{e}. Quantities obtained by this method would be referred to with an affix ‘expo pure’.

    2. (b)

      For a given density of claims ff, the value function is computed via the formula which assumes exponential claims in equation 3, but the "ingredients" WW, Z,F¯Z,\overline{F} and the mean function mm are the correct ingredients corresponding to our original density ff. Quantities obtained by this method would be referred to with an affix ‘expo CI’.

    It turns out that the pure expo approximation works better for large θ\theta, and the correct ingredients approximation works better for small θ\theta.

    Note that we only included tables illustrating approximating J0J_{0} for the first two examples, to keep the length of the paper under control, but similar results were obtained for the other examples.

5.1 A Cramér-Lundberg process with hyperexponential claims of order 2

We take a look at a Cramér-Lundberg process with density function f(x)=23ex+23e2xf(x)=\frac{2}{3}e^{-x}+\frac{2}{3}e^{-2x} with λ=1\lambda=1, θ=1\theta=1 and q=110q=\frac{1}{10}.

Refer to caption
(a) Wq(x)W^{\prime}_{q}(x)
Refer to caption
(b) Wq′′(x)W^{\prime\prime}_{q}(x)
Figure 7: Exact and approximate plots of Wq(x)W^{\prime}_{q}(x) and Wq′′(x)W^{\prime\prime}_{q}(x) for f(x)=23ex+23e2xf(x)=\frac{2}{3}e^{-x}+\frac{2}{3}e^{-2x}, θ=1\theta=1, q=110q=\frac{1}{10}.
Dominant exponent
Φq\Phi_{q}
Percent relative error
(Φq\Phi_{q})
Optimal barrier
bDeFb_{DeF}
Percent relative error
(bDeFb_{DeF})
Exact 0.110113 0 3.45398 0
Expo 0.110657 0.494313 3.51173 1.67191
Dev 0.110115 0.00195933 3.48756 0.972251
Renyi 0.110078 0.0321413 3.5323 2.26744
Table 1: Exact and approximate values of Φq\Phi_{q} and bDeFb_{DeF} for f(x)=23ex+23e2xf(x)=\frac{2}{3}e^{-x}+\frac{2}{3}e^{-2x}, theta=1theta=1, q=110q=\frac{1}{10}, as well as percent relative errors, computed as the absolute value of the difference between the approximation and the exact, divided by the exact, times 100. Relative errors for Φq\Phi_{q} are less than 0.5%0.5\%, with the pure exponential approximation proving to be the worst and the DeVylder the best approximations, respectively. The optimal barrier bDeFb_{DeF} is also best approximated by DeVylder, with Renyi being the worst at 2.26%2.26\% .
θ\theta Closest approximation Φq\Phi_{q} exact Φq\Phi_{q} approximation % error Φq\Phi_{q}
1 Dev 0.110113 0.110115 0.00195933
0.9 Dev 0.120328 0.120331 0.00269878
0.8 Dev 0.132452 0.132457 0.00380056
0.7 Dev 0.147017 0.147025 0.00548411
0.6 Dev 0.16475 0.164763 0.00812643
0.5 Dev 0.186652 0.186675 0.0123901
0.4 Dev 0.214122 0.214163 0.0194631
0.3 Dev 0.249118 0.249196 0.0315039
0.2 Dev 0.294396 0.294551 0.0524528
0.1 Dev 0.353829 0.354145 0.0894466
Table 2: Exact and the winning DeVylder approximate values of Φq\Phi_{q} for f(x)=23ex+23e2xf(x)=\frac{2}{3}e^{-x}+\frac{2}{3}e^{-2x} when θ\theta varies.
θ\theta Closest approximation Barrier exact Barrier approx % error Barrier
1 Dev 3.45398 3.48756 0.972251
0.9 Dev 3.20191 3.23103 0.909487
0.8 Dev 2.90951 2.93074 0.729628
0.7 Dev 2.57043 2.57742 0.272088
0.6 Dev 2.1804 2.16054 0.910666
0.5 Ren 1.74216 1.75266 0.60278
0.4 Expo 1.2735 1.29456 1.65378
0.3 Expo 0.81068 0.652264 19.5412
Table 3: Exact and approximate values of bDeFb_{DeF} for for f(x)=23ex+23e2xf(x)=\frac{2}{3}e^{-x}+\frac{2}{3}e^{-2x} when θ\theta varies. As bDeFb_{DeF} approaches 0, errors of all the approximations increase dramatically, with the pure exponential approximation performing better than the rest. Meanwhile, as bDeFb_{DeF} increases, Renyi and subsequently the de Vylder approximation take the upper hand. For θ=0.2\theta=0.2 and θ=0.1\theta=0.1, all the approximations yield a 0 barrier approximate for exact barrier values of 0.3921050.392105 and 0.03545380.0354538 respectively, hence failing to predict the non-zero barrier.
J0
θ\theta J0 exact J0 expo pure J0 expo pure error J0 expo CI J0 expo CI error
1 5.95034 5.99151 0.691856 6.26009 5.20551
0.9 5.15579 5.17573 0.386663 5.45269 5.7584
0.8 4.39383 4.38494 0.202205 4.67042 6.29489
0.7 3.68299 3.63933 1.18555 3.92937 6.68958
0.6 3.04577 2.96728 2.57704 3.25112 6.7423
0.5 2.50331 2.39942 4.15022 2.65901 6.21974
0.4 2.06833 1.9585 5.31006 2.17044 4.93725
0.3 1.74095 1.65616 4.86984 1.78984 2.80878
0.2 1.50439 1.44242 4.11969 1.50871 0.286672
0.1 1.30271 1.25324 3.79748 1.30271 0
Table 4: Values of J0J_{0} compared with approximations using all exponential inputs (J0J_{0} expo pure) and actual inputs but computed using the exponential formula (J0J_{0} expo CI). The pure exponential approximation does a good job of approximating J0J_{0} for higher values of θ\theta considered, while the exponential CI approximation seemed to fair better for lower θ\theta values
a
θ\theta a exact a expo pure a expo pure error a expo CI a expo CI error
1 3.9669 3.99434 0.691861 4.17339 5.20551
0.9 3.4372 3.45049 0.386665 3.63512 5.7584
0.8 2.92922 2.9233 0.202204 3.11361 6.29489
0.7 2.45533 2.42622 1.18555 2.61958 6.68958
0.6 2.03051 1.97818 2.57704 2.16741 6.7423
0.5 1.66888 1.59961 4.15022 1.77268 6.21974
0.4 1.37888 1.30566 5.31006 1.44696 4.93725
0.3 1.16063 1.10411 4.86983 1.19323 2.80878
0.2 1.00293 0.961612 4.11969 1.0058 0.286672
0.1 0.868476 0.835496 3.79748 0.868476 0
Table 5: Values of aa compared with approximations using all exponential inputs (aa expo pure) and actual inputs but computed using the exponential formula (aa expo CI)
b
θ\theta b exact b expo pure b expo pure error b expo CI b expo CI error
1 1.41036 1.46188 3.65293 1.25374 11.1045
0.9 1.37645 1.44439 4.93621 1.23362 10.3761
0.8 1.31492 1.40417 6.78809 1.19529 9.09781
0.7 1.21057 1.32258 9.25207 1.12775 6.84178
0.6 1.04634 1.17215 12.0245 1.01753 2.7529
0.5 0.810767 0.920406 13.5229 0.853397 5.25805
0.4 0.510085 0.538725 5.61475 0.634716 24.4335
0.3 0.17425 0.0105496 93.9457 0.376872 116.282
0.2 0 0 0 0.105322 100
0.1 0 0 0 0 0
Table 6: Values of bb compared with approximations using all exponential inputs (bb expo pure) and actual inputs but computed using the exponential formula (bb expo CI)

5.2 A Cramér-Lundberg process with hyperexponential claims of order 3

Consider a Cramér-Lundberg process with density function f(x)=1283ex+4283e2x+15083e3xf(x)=\frac{12}{83}e^{-x}+\frac{42}{83}e^{-2x}+\frac{150}{83}e^{-3x}, and c=1c=1, λ=8348\lambda=\frac{83}{48}, θ=263235\theta=\frac{263}{235}, p=263498p=\frac{263}{498}, q=548q=\frac{5}{48}.

The Laplace exponent of this process is κ(s)=s12s83(s+1)21s83(s+2)50s83(s+3)\kappa(s)=s-\frac{12s}{83(s+1)}-\frac{21s}{83(s+2)}-\frac{50s}{83(s+3)} and from this one can invert 1κ(s)q=W^q(s)\frac{1}{\kappa(s)-q}=\widehat{W}_{q}(s) to obtain the scale function 222Laplace inversion done via Mathematica; coefficients and exponents are decimal approximations of the real values.

Wq(x)\displaystyle W_{q}(x) =0.0813294e(2.60997x)0.179472e(1.68854x)0.373887e(0.779311x)+1.63469e(0.18198x).\displaystyle=-0.0813294e^{(-2.60997x)}-0.179472e^{(-1.68854x)}-0.373887e^{(-0.779311x)}+1.63469e^{(0.18198x)}.

From this, we see that the dominant exponent is Φq=0.18198\Phi_{q}=0.18198.

Figure 8 shows the exact and approximate plots of the first two derivatives of WqW_{q}. The exact plots are labelled Wxexact, and coloured as the darkest. The plots of WqW_{q}^{\prime} exhibit noticeable unique minima around x=2x=2, with the exact one being at bDeF=1.89732b_{DeF}=1.89732, which is the optimal barrier that maximizes dividends here. Note that the approximations are practically indistinguishable from the exact around this point (which is our main object of interest here).

Refer to caption
(a) Wq(x)W^{\prime}_{q}(x)
Refer to caption
(b) Wq′′(x)W^{\prime\prime}_{q}(x)
Figure 8: Exact and approximate plots of Wq(x)W^{\prime}_{q}(x) and Wq′′(x)W^{\prime\prime}_{q}(x) for f(x)=1283ex+4283e2x+15083e3xf(x)=\frac{12}{83}e^{-x}+\frac{42}{83}e^{-2x}+\frac{150}{83}e^{-3x}, c=1c=1, q=548q=\frac{5}{48}.
Dominant exponent
Φq\Phi_{q}
Percent relative error
(Φq\Phi_{q})
Optimal barrier
bDeFb_{DeF}
Percent relative error
(bDeFb_{DeF})
Exact 0.18198 0 1.89732 0
Expo 0.184095 1.162215628 2.04608 7.840532962
Renyi 0.181708 0.149466974 2.08136 9.699997892
Dev 0.182011 0.017034839 1.91233 0.79111589
Table 7: Exact and approximate values of Φq\Phi_{q} and bDeFb_{DeF} for f(x)=1283ex+4283e2x+15083e3xf(x)=\frac{12}{83}e^{-x}+\frac{42}{83}e^{-2x}+\frac{150}{83}e^{-3x}, c=1c=1, q=548q=\frac{5}{48}, as well as percent relative errors, computed as the absolute value of the difference between the approximation and the exact, divided by the exact, times 100. Relative errors for the Φq\Phi_{q} value are less than 1.25%1.25\%, even for the worse natural exponential approximation, and the DeVylder approximation is the winner. The optimal barrier bDeFb_{DeF} is well approximated only by DeVylder.
θ\theta Closest approximation Φq\Phi_{q} exact Φq\Phi_{q} approximation % error Φq\Phi_{q}
263/235 Dev 0.18198 0.182011 0.0168217
243/235 Dev 0.194712 0.194754 0.0213671
223/235 Dev 0.209221 0.209279 0.0274827
203/235 Dev 0.225876 0.225957 0.0358309
183/235 Dev 0.245146 0.245262 0.0474032
163/235 Dev 0.267635 0.267806 0.0637063
143/235 Dev 0.294126 0.294382 0.0870647
123/235 Dev 0.325643 0.326038 0.121115
103/235 Dev 0.363539 0.364163 0.171618
83/235 Dev 0.40961 0.410625 0.247788
63/235 Dev 0.466261 0.46796 0.364457
43/235 Dev 0.536719 0.539647 0.545532
23/235 Dev 0.62533 0.630516 0.829419
3/235 Dev 0.737962 0.747389 1.27736
Table 8: Exact and the winning DeVylder approximate values of Φq\Phi_{q}, for f(x)=1283ex+4283e2x+15083e3xf(x)=\frac{12}{83}e^{-x}+\frac{42}{83}e^{-2x}+\frac{150}{83}e^{-3x}, θ\theta varies.
θ\theta Closest approximation Barrier exact Barrier approx % error Barrier
263/235 Dev 1.89732 1.91233 0.791183
243/235 Dev 1.79954 1.78002 1.08482
183/235 Ren 1.45224 1.52484 4.9989
163/235 Ren 1.31579 1.33691 1.60463
143/235 Ren 1.16804 1.12368 3.79796
123/235 Expo 1.00898 1.04123 3.19653
103/235 Expo 0.839228 0.794964 5.27444
83/235 Expo 0.660338 0.513179 22.2854
63/235 Expo 0.474896 0.196234 58.6785
43/235 Expo 0.286563 0 100
23/235 Expo 0.0998863 0 100
3/235 All 0 0 0
Table 9: Exact and approximate values of bDeFb_{DeF} for f(x)=1283ex+4283e2x+15083e3xf(x)=\frac{12}{83}e^{-x}+\frac{42}{83}e^{-2x}+\frac{150}{83}e^{-3x}, when θ\theta varies. Unsurprisingly, when bDeFb_{DeF} is near 0, the natural exponential approximation wins, but has poor performance, and as bDeFb_{DeF} increases, Renyi and subsequently the de Vylder approximation take the upper hand. For the smallest three values of θ\theta, all of the approximations yielded b~DeF=0\widetilde{b}_{DeF}=0 as the optimal barrier, with this being true only for θ=3/235\theta=3/235.

We move now to the dividend problem with capital injections with cost k1k\geq 1 as in Theorem 1. One can compute the value function J0J_{0} at x=0x=0 in terms of WW, ZZ,CC, SS, and GG – see equation 5.

To provide a more concrete example, fixing q=548q=\frac{5}{48},P=0P=0, k=3/2k=3/2 as input parameters we compute for values of J0J_{0} as a function of θ\theta, with results summarized in the tables below. The tables provide comparisons of the computed optimal quantities J0J_{0}, aa, and bb to an approximation using all exponential inputs (referred to as J0J_{0}, aa, and bb expo pure) and to an approximation which uses actual inputs but computed using the exponential formula as described in equation 3 (referred to as J0J_{0}, aa, and bb expo CI).

J0
θ\theta J0J_{0} J0J_{0} expo pure J0J_{0} expo pure error J0J_{0} expo CI J0J_{0} expo CI error
263/235 3.7747 3.76883 0.155556 4.11784 9.09041
243/235 3.41491 3.38603 0.845606 3.74156 9.5654
223/235 3.0636 3.00802 1.81444 3.36985 9.99637
203/235 2.72335 2.63828 3.1238 3.00466 10.3296
183/235 2.39737 2.28225 4.80185 2.64879 10.4871
163/235 2.08958 1.94765 6.79226 2.3062 10.3665
143/235 1.80446 1.64396 8.8946 1.9823 9.85516
123/235 1.54668 1.38072 10.73 1.68379 8.86472
103/235 1.32041 1.16526 11.7499 1.4178 7.37587
83/235 1.12864 1.00194 11.2258 1.19022 5.45555
63/235 0.972835 0.88785 8.73585 1.00404 3.20798
43/235 0.852739 0.789923 7.36635 0.859039 0.738837
23/235 0.751597 0.701299 6.69218 0.751597 0
3/235 0.660372 0.620567 6.02761 0.660372 0
Table 10: Values of J0J_{0} compared with approximations using all exponential inputs (J0J_{0} expo pure) and actual inputs but computed using the exponential formula (J0J_{0} expo CI). The pure exponential approximation does a good job of approximating J0J_{0} for higher values of θ\theta considered, while the exponential CI approximation seemed to fair better for lower θ\theta values
a
θ\theta aa aa expo pure aa expo pure error aa expo CI aa expo CI error
263/235 2.51647 2.74523 0.155536 2.51255 9.09042
243/235 2.27661 2.49437 0.845596 2.25735 9.56541
223/235 2.0424 2.24657 1.81443 2.00535 9.99638
203/235 1.81557 2.00311 3.1238 1.75885 10.3296
183/235 1.59825 1.76586 4.80185 1.5215 10.4871
163/235 1.39306 1.53747 6.79226 1.29844 10.3665
143/235 1.20298 1.32153 8.8946 1.09598 9.85516
123/235 1.03112 1.12252 10.73 0.920479 8.86472
103/235 0.880271 0.945198 11.7499 0.77684 7.37587
83/235 0.752428 0.793477 11.2258 0.667962 5.45555
63/235 0.648557 0.669362 8.73585 0.5919 3.20798
43/235 0.568493 0.572693 7.36635 0.526616 0.738838
23/235 0.501065 0.501065 6.69218 0.467532 0
3/235 0.440248 0.440248 6.02761 0.413711 0
Table 11: Values of aa compared with approximations using all exponential inputs (aa expo pure) and actual inputs but computed using the exponential formula (aa expo CI)
b
θ\theta bb bb expo pure bb expo pure error bb expo CI bb expo CI error
263/235 0.709355 0.805116 13.4997 0.677918 4.43179
243/235 0.695874 0.801936 15.2416 0.671779 3.46259
223/235 0.677601 0.794377 17.2337 0.662801 2.18425
203/235 0.653005 0.779265 19.3352 0.649805 0.490147
183/235 0.620126 0.751601 21.2012 0.631097 1.76912
163/235 0.576553 0.704104 22.123 0.604293 4.81128
143/235 0.519526 0.627369 20.7579 0.566198 8.98366
123/235 0.446259 0.511076 14.5246 0.512961 14.947
103/235 0.354524 0.346046 2.39143 0.440755 24.3231
83/235 0.243362 0.126054 48.2032 0.347059 42.6099
63/235 0.113593 0 100 0.231975 104.216
43/235 0 0 0 0.0987484 0
23/235 0 0 0 0 0
3/235 0 0 0 0 0
Table 12: Values of bb compared with approximations using all exponential inputs (bb expo pure) and actual inputs but computed using the exponential formula (bb expo CI)

To provide a point of comparison, we fix q=548q=\frac{5}{48}, and compute the de Finetti barrier to be bDeF=1.89732b_{DeF}=1.89732 and the corresponding dividend value function when starting at x=0x=0 to be JDeF=1.99847J_{DeF}=1.99847.

kk J0J_{0} % deviation J0JDeFJ_{0}-J_{DeF} J0J_{0} bb % deviation bb
1 154.123 3.0801 5.07857 100 0
2 65.714 1.31327 3.31174 43.0208 1.08108
3 42.863 0.856604 2.85507 25.4792 1.4139
4 31.4465 0.628448 2.62692 17.6853 1.56178
5 24.6995 0.493611 2.49208 13.4234 1.64264
6 20.2855 0.4054 2.40387 10.7759 1.69287
7 17.1884 0.343504 2.34197 8.98441 1.72686
8 14.9014 0.2978 2.29627 7.69619 1.7513
9 13.1463 0.262724 2.26119 6.72732 1.76968
10 11.758 0.23498 2.23345 5.97306 1.78399
100 1.11095 0.0222019 2.02067 0.533448 1.8872
1000 0.110409 0.00220648 2.00067 0.0527257 1.89632
10000 0.0109536 0.000218903 1.99869 0.00533526 1.89722
Table 13: Values of J0J_{0} and bb in presence of capital injections compared to the case where capital injections are non-existent, JDeF=1.99847J_{DeF}=1.99847 and bDeF=1.89732b_{DeF}=1.89732. As kk is increased one can see that J0J_{0} and bb approaches JDeFJ_{DeF} and bDeFb_{DeF}. This is expected since higher costs of injecting capital makes it less viable, hence it is treated like the concept does not exist.

5.3 A Cramér-Lundberg process with oscillating density and scale function

In the following example, we study a Cramèr-Lundberg  model with density of claims given by

f(x)=\displaystyle f(x)= ueax2cos2(ωx+ϕ2)=ueax(1+cos(ωx+ϕ))=\displaystyle ue^{-ax}2\cos^{2}\left(\frac{\omega x+\phi}{2}\right)=ue^{-ax}(1+\cos(\omega x+\phi))=
=\displaystyle= eax(u+ucos(ϕ)cos(ωx)usin(ϕ)sin(ωx))\displaystyle e^{-ax}(u+u\cos(\phi)\cos(\omega x)-u\sin(\phi)\sin(\omega x))

where

u=a(a2+ω2)a2+ω2+a2cos(ϕ)aωsin(ϕ).u=\frac{a\left(a^{2}+\omega^{2}\right)}{a^{2}+\omega^{2}+a^{2}\cos(\phi)-a\omega\sin(\phi)}.

Assuming further that a=1a=1, ϕ=2\phi=2, ω=20\omega=20, and that θ=1\theta=1, q=1/10q=1/10, the Laplace exponent for this process is κ(s)=s(2.09898s3+5.29695s2+843.502s+420.846)(s+1.)(s2+2.s+401.)\kappa(s)=\frac{s\left(2.09898s^{3}+5.29695s^{2}+843.502s+420.846\right)}{(s+1.)\left(s^{2}+2.s+401.\right)} and the scale function is

Wq(x)\displaystyle W_{q}(x) =0.824723e0.0881484x0.348141e0.540677x\displaystyle=0.824723e^{0.0881484x}-0.348141e^{-0.540677x}
+e1.0117xcos(19.9957x)((0.000285494+0.0000804151i)sin(39.9914x)\displaystyle+e^{-1.0117x}\cos(19.9957x)~{}\Big{(}-(0.000285494\,+0.0000804151i)\sin(39.9914x)
(0.0000804151+0.000285494i)+(0.0000804151+0.000285494i)cos(39.9914x))\displaystyle-(0.0000804151\,+0.000285494i)+(-0.0000804151+0.000285494i)\cos(39.9914x)\Big{)}
+e1.0117xsin(19.9957x)((0.00008041510.000285494i)sin(39.9914x)\displaystyle+e^{-1.0117x}\sin(19.9957x)~{}\Big{(}-(0.0000804151\,-0.000285494i)\sin(39.9914x)
+(0.000285494+0.0000804151i)cos(39.9914x)(0.0002854940.0000804151i)).\displaystyle+(0.000285494\,+0.0000804151i)\cos(39.9914x)-(0.000285494\,-0.0000804151i)\Big{)}.
Refer to caption
(a) Wq(x)W^{\prime}_{q}(x)
Refer to caption
(b) Wq′′(x)W^{\prime\prime}_{q}(x)
Figure 9: Plots of Wq(x)W^{\prime}_{q}(x), and Wq′′(x)W^{\prime\prime}_{q}(x) of the exact solution and the approximations for f(x)=ueax2cos2(ωx+ϕ2)f(x)=ue^{-ax}2\cos^{2}\left(\frac{\omega x+\phi}{2}\right), θ=1\theta=1, q=110q=\frac{1}{10}.
Dominant exponent
Φq\Phi_{q}
Percent relative error
(Φq\Phi_{q})
Optimal barrier
bDeFb_{DeF}
Percent relative error
(bDeFb_{DeF})
Exact 0.0881484 0 4.38201 0
Expo 0.0878658 0.32053 4.42263 0.927122
Renyi 0.0881481 0.000314617 4.39788 0.362284
Dev 0.0881484 6.11743*10^-6 4.39745 0.352331
Table 14: Exact and approximate values of Φq\Phi_{q} and bDeFb_{DeF} for f(x)=ueax2cos2(ωx+ϕ2)f(x)=ue^{-ax}2\cos^{2}\left(\frac{\omega x+\phi}{2}\right), θ=1\theta=1, q=110q=\frac{1}{10}. The DeVylder approximation wins on both fronts.

Clearly, our completely monotone approximation cannot fully reproduce functions like Wq(x),Wq′′(x)W_{q}^{\prime}(x),W_{q}^{\prime\prime}(x) in examples like this where oscillations occur (note however that the de Finetti  optimal barrier is well approximated here). If a more exact reproduction is necessary, higher order approximations should be used.

6 The maximal error of exponential approximations J0J_{0} along one parameter families of Cramér-Lundberg processes

In this section, we provide the two approximations for the dividend value with capital injections J0J_{0}, and the dividend barrier bb, for two one parameter families of Cramér-Lundberg processes, with densities given respectively by:

f(x)=kϵ[ex+ϵe2x]\displaystyle f(x)=k_{\epsilon}\left[e^{-x}+\epsilon e^{-2x}\right] (41)
f(x)=kϵ[1283ex+ϵ(4283e2x+15083e3x)],\displaystyle f(x)=k_{\epsilon}\left[\frac{12}{83}e^{-x}+\epsilon\left(\frac{42}{83}e^{-2x}+\frac{150}{83}e^{-3x}\right)\right], (42)

where kϵk_{\epsilon} is the normalization constant, and compute the maximal error of approximation when ϵ(0,)\epsilon\in(0,\infty) and θ1\theta\approx 1. For this choice, the pure exponential approximation works considerably better, ϵ\forall\epsilon.

J0
J0 exact J0 expo pure J0 expo pure error J0 expo CI J0 expo CI error
0.001 7.1879 7.18802 0.0016603 7.18849 0.00827967
0.01 7.17075 7.17193 0.0164663 7.17666 0.0824782
0.1 7.008 7.01863 0.151653 7.06358 0.793041
1 5.95034 5.99151 0.691856 6.26009 5.20551
10 4.20175 4.19406 0.183122 4.40089 4.73941
100 3.66909 3.6654 0.100631 3.69555 0.721228
1000 3.6025 3.60208 0.0117065 3.60523 0.075585
Table 15: λ=1\lambda=1, θ=1\theta=1, q=110q=\frac{1}{10} k=3/2k=3/2 and P=0P=0. As expected, the errors decrease both as ϵ\epsilon goes to zero and infinity since the densities approach an exponential density.
Refer to caption
Refer to caption
Figure 10: J0J_{0} values and errors plotted against ϵ\epsilon. Errors peak at ϵ=1\epsilon=1.

We do the same thing for the family of densities given by f(x)=kϵ[1283ex+ϵ(4283e2x+15083e3x)]f(x)=k_{\epsilon}\left[\frac{12}{83}e^{-x}+\epsilon\left(\frac{42}{83}e^{-2x}+\frac{150}{83}e^{-3x}\right)\right].

J0
ϵ\epsilon J0 exact J0 expo pure J0 expo pure error J0 expo CI J0 expo CI error
0.001 7.95508 7.95771 0.0330111 7.96565 0.132796
0.01 7.68765 7.71127 0.307292 7.78772 1.30166
0.1 6.06176 6.15381 1.5186 6.62641 9.31507
1 3.7747 3.76883 0.155556 4.11784 9.09041
10 3.1382 3.1379 0.00959692 3.23354 3.03813
100 3.05894 3.06427 0.174306 3.12284 2.08921
1000 3.0508 3.05678 0.196219 3.11149 1.98956
Table 16: λ=1\lambda=1, c=1c=1, q=548q=\frac{5}{48} k=3/2k=3/2 and P=0P=0. As ϵ\epsilon goes to zero, the density becomes exponential hence the decrease in errors. As ϵ\epsilon goes to infinity, the density approaches a hyper exponential density of order 2, but still both methods of approximating J0J_{0} yield reasonable results.
Refer to caption
Refer to caption
Figure 11: J0J_{0} values and errors plotted against ϵ\epsilon. Errors peak at ϵ=0.1\epsilon=0.1.

7 The profit function when the claims are distributed according to a matrix exponential jumps density

Consider now the more general case when the claims are distributed according to a matrix exponential density generated by a row vector β\vec{\beta} and by an invertible matrix BB of order nn, which are such that the vector βexB\vec{\beta}e^{xB} is decreasing componentwise to 0, and β.𝟏0\vec{\beta}.{\mbox{\boldmath$1$}}\neq 0, with 𝟏1 a column vector. As customary, we restrict w.l.o.g. to the case when β\vec{\beta} is a probability vector, and β.𝟏=1\vec{\beta}.{\mbox{\boldmath$1$}}=1, so that

F¯(x)=βexB𝟏\overline{F}(x)=\vec{\beta}e^{xB}{\mbox{\boldmath$1$}}

is a valid survival function.

The matrix versions of our functions are:

{Ca(x)=λ0xWq(xy)F¯(y+a)𝑑y=λβ0xWq(xy)eyB𝑑yeaB𝟏=C(x)eaB𝟏ma(y)=0azf(y+z)𝑑z=βeyB0azezB(B)𝑑z𝟏=βeyBM(a)𝟏Ga(x)=λ0xWq(xy)ma(y)𝑑y=C(x)M(a)𝟏,\begin{cases}C_{a}(x)=\lambda\int_{0}^{x}W_{q}(x-y)\;\overline{F}(y+a)\;dy=\lambda\vec{\beta}\int_{0}^{x}W_{q}(x-y)\;e^{yB}\;dy\;e^{aB}{\mbox{\boldmath$1$}}=\vec{C}(x)e^{aB}{\mbox{\boldmath$1$}}\\ m_{a}(y)=\int_{0}^{a}zf(y+z)dz=\vec{\beta}\;e^{yB}\;\int_{0}^{a}z\;e^{zB}(-B)\;dz\;{\mbox{\boldmath$1$}}=\vec{\beta}\;e^{yB}M(a){\mbox{\boldmath$1$}}\\ G_{a}(x)=\lambda\int_{0}^{x}W_{q}(x-y)\;m_{a}(y)\;\;dy=\vec{C}(x)M(a){\mbox{\boldmath$1$}}\;\;\;\;\end{cases}, (43)

where

{C(x)=λ0xWq(xy)eyB𝑑yC(x)=λβ0xWq(xy)eyB𝑑y.\begin{cases}C(x)=\lambda\int_{0}^{x}W_{q}(x-y)\;e^{yB}\;dy\\ \vec{C}(x)=\lambda\vec{\beta}\int_{0}^{x}W_{q}(x-y)\;e^{yB}\;dy\end{cases}. (44)

The product formulas (43) may also be established directly in the phase-type case, using the conditional independence of the ruin probability of the overshoot size.

We derive first these extensions from scratch for (β,B)(\vec{\beta},B) phase-type densities, in order to highlight their probabilistic interpretation. Later, we will show that the matrix exponential jumps case follows as a particular case of [GK17].

Recall first [AA10] that Ψq(x)=Ψq(x)𝟏,\Psi_{q}(x)=\vec{\Psi}_{q}(x){\mbox{\boldmath$1$}}, where Ψq(x)\vec{\Psi}_{q}(x) is a vector whose components represent the probability that ruin occurs during a certain phase, and that the conditional independence of ruin and overshoots translates into the product formula

Ψq(x,y):=Px[T0,<,XT0,<y]=Ψq(x)eyB𝟏.\Psi_{q}(x,y):=P_{x}[T_{0,-}<\infty,X_{T_{0,-}}<-y]=\vec{\Psi}_{q}(x)e^{yB}{\mbox{\boldmath$1$}}. (45)

To take advantage of this, it is convenient to replace from the beginning Zq(x)Z_{q}(x) by Ψq(x)\Psi_{q}(x), taking advantage of the formula [AKP04, Kyp14]

Zq(x)=Ψq(x)+Wq(x)qΦqC(x)=(cqΦq)Wq(x)Ψq(x).Z_{q}(x)=\Psi_{q}(x)+W_{q}(x)\frac{q}{\Phi_{q}}\Longrightarrow C(x)=(c-\frac{q}{\Phi_{q}})W_{q}(x)-\Psi_{q}(x). (46)

Alternatively, one may introduce a vector function

Zq(x):=Ψq(x)+Wq(x)qΦq𝟏.\vec{Z}_{q}(x):=\vec{\Psi}_{q}(x)+W_{q}(x)\frac{q}{\Phi_{q}}{\mbox{\boldmath$1$}}. (47)

On the other hand, the mean function may be written as

ma=0ayF(dy)=aF¯(a)+0aF¯(x)𝑑x=βM(a)𝟏,M(a)=B1eaB(aInB1).m_{a}=\int_{0}^{a}y\;F(\mathrm{d}y)=-a\overline{F}(a)+\int_{0}^{a}\overline{F}(x)dx=\vec{\beta}M(a){\mbox{\boldmath$1$}},M(a)=-B^{-1}-e^{aB}\left(aI_{n}-B^{-1}\right).

The following result follows in the phase-type case just as in the exponential case [AGLW20]; in the matrix exponential jumps case, it may be obtained from [GK17]:

PROPOSITION 6.

For a Cramèr-Lundberg process (compound Poisson ) with matrix exponential jumps of type (β,B)(\vec{\beta},B), it holds that

  1. 1.
    Jx={kGa(x)+J0Sa(x)=kGa(x)+1kG(b)S(b)Sa(x),x[0,b]kx+J0x[a,0]0xa,J_{x}=\begin{cases}kG_{a}(x)+J_{0}S_{a}(x)=kG_{a}(x)+\frac{1-kG^{\prime}(b)}{S^{\prime}(b)}S_{a}(x),&x\in[0,b]\\ kx+J_{0}&x\in[-a,0]\\ 0&x\leq-a\end{cases}, (48)

    where

    {C(x)=λ0xWq(xy)eyB𝑑yC(x)=λβ0xWq(xy)eyB𝑑y=cWq(x)1Zq(x)=(cqΦq)Wq(x)1Ψq(x)Ga(x)=Cq(x)M(a)𝟏Ra(x)=Sa(x)Zq(x)=Cq(x)eaB𝟏,\begin{cases}C(x)=\lambda\int_{0}^{x}W_{q}(x-y)\;e^{yB}\;dy\\ \vec{C}(x)=\lambda\vec{\beta}\int_{0}^{x}W_{q}(x-y)\;e^{yB}\;dy=cW_{q}(x)\vec{1}-\vec{Z}_{q}(x)=(c-\frac{q}{\Phi_{q}})W_{q}(x)\vec{1}-\vec{\Psi}_{q}(x)\\ G_{a}(x)=\vec{C}_{q}(x)M(a){\mbox{\boldmath$1$}}\\ R_{a}(x)=S_{a}(x)-Z_{q}(x)=\vec{C}_{q}(x)e^{aB}{\mbox{\boldmath$1$}}\end{cases}, (49)

    and

    J0=1kCq(b)M(a)𝟏qWq(b)+Cq(b)eaB𝟏.J_{0}=\frac{1-k\vec{C}_{q}^{\prime}(b)\;M(a){\mbox{\boldmath$1$}}}{qW_{q}(b)+\vec{C}_{q}^{\prime}(b)e^{aB}{\mbox{\boldmath$1$}}}. (50)
  2. 2.

    For fixed aa, the optimality equation bJ0a,b=0\frac{\partial}{\partial b}J_{0}^{a,b}=0 simplifies to

    J0=kCq′′(b)M(a)𝟏qWq(b)+Cq′′(b)eaB𝟏.J_{0}=\frac{k\vec{C}_{q}^{\prime\prime}(b)\;M(a){\mbox{\boldmath$1$}}}{qW_{q}^{\prime}(b)+\vec{C}_{q}^{\prime\prime}(b)e^{aB}{\mbox{\boldmath$1$}}}. (51)
REMARK 12.

The additive separation of a,ba,b which was the basis of proving optimality in the exponential case does not seem possible anymore, but (50) allows the numeric computation of the optimum.

References

  • [AA10] Hansjörg Albrecher and Sören Asmussen, Ruin probabilities, vol. 14, World Scientific, 2010.
  • [AAM20] Hansjoerg Albrecher, Pablo Azcue, and Nora Muler, Optimal ratcheting of dividends in insurance, SIAM Journal on Control and Optimization 58 (2020), no. 4, 1822–1845.
  • [ABH18] Florin Avram, Abhijit Datta Banik, and András Horváth, Ruin probabilities by Padé’s method: simple moments based mixed exponential approximations (Renyi, De Vylder, Cramér–Lundberg), and high precision approximations with both light and heavy tails, European Actuarial Journal (2018), 1–27.
  • [ACFH11] Florin Avram, Donatien Chedom-Fotso, and András Horváth, On moments based Padé approximations of ruin probabilities, Journal of Computational and Applied mathematics 235 (2011), no. 10, 3215–3228.
  • [AGLW20] Florin Avram, D. Goreac, J Li, and X Wu, Equity cost induced dichotomy for optimal dividends in the cramér-lundberg model, IME (2020).
  • [AGVA19] Florin Avram, D. Grahovac, and C. Vardar-Acar, The W,ZW,Z scale functions kit for first passage problems of spectrally negative Lévy processes, and applications to the optimization of dividends, arXiv preprint (2019).
  • [AHPS19] Florin Avram, Andras Horváth, Serge Provost, and Ulyses Solon, On the padé and laguerre–tricomi–weeks moments based approximations of the scale function w and of the optimal dividends barrier for spectrally negative lévy risk processes, Risks 7 (2019), no. 4, 121.
  • [AKP04] Florin Avram, Andreas Kyprianou, and Martijn Pistorius, Exit problems for spectrally negative Lévy processes and applications to (Canadized) Russian options, The Annals of Applied Probability 14 (2004), no. 1, 215–238.
  • [AM14] Pablo Azcue and Nora Muler, Stochastic optimization in insurance: A dynamic programming approach, Springer, 2014.
  • [AP14] Florin Avram and Martijn Pistorius, On matrix exponential approximations of ruin probabilities for the classic and brownian perturbed cramér-lundberg processes, Insurance: Mathematics and Economics 59 (2014), 57–64.
  • [APP07] Florin Avram, Zbigniew Palmowski, and Martijn R Pistorius, On the optimal dividend problem for a spectrally negative Lévy process, The Annals of Applied Probability (2007), 156–180.
  • [APY18] Florin Avram, José Luis Pérez, and Kazutoshi Yamazaki, Spectrally negative Lévy processes with Parisian reflection below and classical reflection above, Stochastic processes and Applications 128 (2018), 255–290.
  • [BDW07] Christopher J Beveridge, David CM Dickson, and Xueyuan Wu, Optimal dividends under reinsurance, Centre for Actuarial Studies, Department of Economics, University of …, 2007.
  • [Ber98] Jean Bertoin, Lévy processes, vol. 121, Cambridge university press, 1998.
  • [BFS08] PB Brito, F Fabiao, and A Staubyn, Euler, lambert, and the lambert w-function today., Mathematical Scientist 33 (2008), no. 2.
  • [BHT05] Andrea Bobbio, András Horváth, and Miklós Telek, Matching three moments with minimal acyclic phase type distributions, Stochastic models 21 (2005), no. 2-3, 303–326.
  • [Boy98] JP Boyd, Global approximations to the principal real-valued branch of the lambert w-function, Applied mathematics letters 11 (1998), no. 6, 27–31.
  • [CGH+96] Robert M Corless, Gaston H Gonnet, David EG Hare, David J Jeffrey, and Donald E Knuth, On the lambertw function, Advances in Computational mathematics 5 (1996), no. 1, 329–359.
  • [DD05] David Dickson and Steve Drekic, Optimal dividends under a ruin probabilityconstraint, 2005.
  • [DF57] Bruno De Finetti, Su un’impostazione alternativa della teoria collettiva del rischio, Transactions of the XVth international congress of Actuaries, vol. 2, 1957, pp. 433–443.
  • [DcSA16] Bogdan Dumitrescu, Bogdan C cSicleru, and Florin Avram, Modeling probability densities with sums of exponentials via polynomial approximation, Journal of Computational and Applied Mathematics 292 (2016), 513–525.
  • [DV78] Fl De Vylder, A practical solution to the problem of ultimate ruin probability, Scandinavian Actuarial Journal 1978 (1978), no. 2, 114–119.
  • [ES11] Julia Eisenberg and Hanspeter Schmidli, Minimising expected discounted capital injections by reinsurance in a classical risk model, Scandinavian Actuarial Journal 2011 (2011), no. 3, 155–176.
  • [Ger69] Hans U Gerber, Entscheidungskriterien für den zusammengesetzten poisson-prozess, Ph.D. thesis, ETH Zurich, 1969.
  • [GK17] Leslaw Gajek and Lukasz Kuciński, Complete discounted cash flow valuation, Insurance: Mathematics and Economics 73 (2017), 1–19.
  • [GSS08] Hans U Gerber, Elias SW Shiu, and Nathaniel Smith, Methods for estimating the optimal dividend barrier and the probability of ruin, Insurance: Mathematics and Economics 42 (2008), no. 1, 243–254.
  • [Høj02] Bjarne Højgaard, Optimal dynamic premium control in non-life insurance. maximizing dividend pay-outs, Scandinavian Actuarial Journal 2002 (2002), no. 4, 225–245.
  • [KKR13] Alexey Kuznetsov, Andreas E Kyprianou, and Victor Rivero, The theory of scale functions for spectrally negative Lévy processes, Lévy Matters II, Springer, 2013, pp. 97–186.
  • [KS08] Natalie Kulenko and Hanspeter Schmidli, Optimal dividend strategies in a cramér–lundberg model with capital injections, Insurance: Mathematics and Economics 43 (2008), no. 2, 270–278.
  • [Kyp14] Andreas Kyprianou, Fluctuations of Lévy processes with applications: Introductory lectures, Springer Science & Business Media, 2014.
  • [LL19] Kristoffer Lindensjö and Filip Lindskog, Optimal dividends and capital injection under dividend restrictions, arXiv preprint arXiv:1902.06294 (2019).
  • [LZ08] Arne Løkka and Mihail Zervos, Optimal dividend and issuance of equity policies in the presence of proportional costs, Insurance: Mathematics and Economics 42 (2008), no. 3, 954–961.
  • [NPY20] Kei Noba, José-Luis Pérez, and Xiang Yu, On the bailout dividend problem for spectrally negative markov additive models, SIAM Journal on Control and Optimization 58 (2020), no. 2, 1049–1076.
  • [Pak15] Anthony G Pakes, Lambert’s w meets kermack–mckendrick epidemics, IMA Journal of Applied Mathematics 80 (2015), no. 5, 1368–1386.
  • [Pis04] Martijn R Pistorius, On exit and ergodicity of the spectrally one-sided Lévy process reflected at its infimum, Journal of Theoretical Probability 17 (2004), no. 1, 183–220.
  • [PYB18] José-Luis Pérez, Kazutoshi Yamazaki, and Alain Bensoussan, Optimal periodic replenishment policies for spectrally positive Lévy demand processes, arXiv preprint arXiv:1806.09216 (2018).
  • [Ram92] Colin M Ramsay, A practical algorithm for approximating the probability of ruin, Transactions of the Society of Actuaries 44 (1992), 443–461.
  • [SLG84] Steven E Shreve, John P Lehoczky, and Donald P Gaver, Optimal consumption for general diffusions with absorbing and reflecting barriers, SIAM Journal on Control and Optimization 22 (1984), no. 1, 55–75.
  • [Sup76] VN Suprun, Problem of destruction and resolvent of a terminating process with independent increments, Ukrainian Mathematical Journal 28 (1976), no. 1, 39–51.
  • [VAVZ14] Eleni Vatamidou, Ivo Jean Baptiste Franccois Adan, Maria Vlasiou, and Bert Zwart, On the accuracy of phase-type approximations of heavy-tailed risk models, Scandinavian Actuarial Journal 2014 (2014), no. 6, 510–534.
  • [VLSHGG+19] H Vazquez-Leal, MA Sandoval-Hernandez, JL Garcia-Gervacio, AL Herrera-May, and UA Filobello-Nino, Psem approximations for both branches of lambert function with applications, Discrete Dynamics in Nature and Society 2019 (2019).