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

An optimal multibarrier strategy for a singular stochastic control problem with a state-dependent reward

Mauricio Junca Department of Mathematics, Universidad de los Andes, Bogotá, Colombia. [email protected] Harold A. Moreno-Franco Department of Statistics and Data Analysis, Laboratory of Stochastic Analysis and its Applications, National Research University Higher School of Economics, Moscow, Russian Federation. [email protected]  and  Jose-Luis Pérez Department of Probability and Statistics, Centro de Investigacion en Matemáticas A.C., Guanajuato, Mexico. [email protected]
Abstract.

We consider a singular control problem that aims to maximize the expected cumulative rewards, where the instantaneous returns depend on the state of a controlled process. The contributions of this paper are twofold. Firstly, to establish sufficient conditions for determining the optimality of the one-barrier strategy when the uncontrolled process XX follows a spectrally negative Lévy process with a Lévy measure defined by a completely monotone density. Secondly, to verify the optimality of the (2n+1)(2n+1)-barrier strategy when XX is a Brownian motion with a drift. Additionally, we provide an algorithm to compute the barrier values in the latter case.

1. Introduction and problem formulation

Singular stochastic control problems have applications in various fields such as finance, actuarial sciences, and harvesting; see e.g. [2, 6, 17, 19]. In this applied literature, the main objective is to establish suitable conditions for determining an optimal strategy to maximize expected rewards until the controlled process falls below zero for the first time. The focus often centres on seeking nearly explicit solutions for these optimization problems.

An example of such a scenario is seen in optimal dividend problems when the underlying process XX is a spectrally negative Lévy process whose Lévy measure has a completely monotone density (see Assumption 2.1). Under the condition that the instantaneous rewards (IRs) for executing an admissible strategy are independent of the state of the controlled process, Loeffen [17] showed that the one-barrier strategy is optimal. A similar situation is found in harvesting problems, but in this case the uncontrolled process XX is described by a diffusion process without jumps; see e.g. [2, 19].

When the IRs depend on the state of the controlled process through a function gg (see (1.1)), known as the instantaneous marginal yield function, Alvarez studied the problem in [3] for the case where XX is a linear diffusion process. Under certain assumptions on the function gg, the underlying diffusion XX and the parameters of the problem, Alvarez showed that an optimal control strategy is achieved with an one-barrier strategy for further details see Remark 3.2). Analogous conclusions have been reached in related contexts, such as the optimal bail-out dividend problem for one-dimensional diffusions, as detailed in [12].

Other works on singular control problems that include the dependency on the state of the controlled process on the IRs without providing a specific form of the optimal policy are [4, 24]. The difficulty is that incorporating state-dependent IRs introduces considerably more challenges compared to situations with constant IRs. Unlike the latter, in the former case, the structure of the optimal strategy and the criteria for optimality depend not only on the characteristics of the underlying process, typically encoded in the properties of the scale functions, but also upon the nature of the function gg as discussed in [3] when XX is a linear diffusion process.

In this study, our first contribution is to expand the scenarios in which the one-barrier strategy is optimal for the maximization problem mentioned earlier. We specifically consider sufficiently smooth positive functions gg that satisfy Assumption 3.1 (which depends on the parameters of the problem), and spectrally negative Lévy processes where the Lévy measure has a completely monotone density.

However, in situations where some of the aforementioned conditions are not met, there is no guarantee that one-barrier strategies are optimal. In fact, in the context of the optimal dividend problem, there is a well-known example in [7] where a multibarrier strategy is optimal. When XX is a Cramer-Lundberg process, Gerber showed that an optimal strategy is of the multibarrier type (see [22], Section 2.4.2), however, there was no procedure to calculate the actual values of the barriers. Some works have proposed methodologies for identifying optimal barriers, with [1] representing a recent endeavor in this regard.

Hence, the second and main contribution of this work lies in presenting an algorithm for determining optimal barriers for general sufficiently smooth positive functions gg that satisfy Assumption 3.1, and with a Brownian motion with drift serving as the underlying process XX.

One of the main applications of our model lies in the classical theory of ruin, which examines the path of a stochastic process until the occurrence of ruin defined as the first instance when the process falls below zero. In this context, our model can be viewed as an extension of de Finetti’s dividend problem (see [11]). The de Finetti problem entails an insurance company making dividend payments to shareholders throughout its operational lifespan until the event of ruin. These dividend payments are intended to maximize the expected net present value of the dividend payments to the shareholders up to the time of ruin. In the spirit of [23], we generalize this classic problem by incorporating instantaneous state-dependent transaction costs or taxes. After these costs are deducted, shareholders receive a state-dependent instantaneous net proportion of the surplus, given by the function gg, in the form of dividends. Further results on the classical optimal dividend problem driven by spectrally negative Lévy process can be found on [6, 17, 18]. Additional related results for a class of diffusion process considering policies with transaction costs can be found on [8, 9] and with capital injections over a finite horizon in [13].

Another application of our model is in the context of the harvesting problem (see for instance [19, 24]). In this case, the stochastic process XX represents the size or density of a population, with the net price per unit for the population being state-dependent and given by the function gg. In the context of ecology it makes sense that the price of harvesting is not constant. For instance, harvesting costs tend to be higher for smaller populations since locating specific individuals for harvesting is more challenging. Conversely, in larger populations, harvesting costs tend to be lower as individuals are more readily located for harvesting purposes. The objective of this problem is to identify the optimal harvesting policy that maximizes the total expected discounted income harvested until the population becomes extinct.

1.1. Problem formulation

Let X={Xt:t0}X=\{X_{t}:t\geq 0\} denote a spectrally negative Lévy process, i.e., a Lévy process with non-monotone trajectories that only has negative jumps, defined on a probability space (Ω,,)(\Omega,\mathcal{F},\mathbbm{P}). For xx\in\mathbbm{R} we denote by x\mathbb{P}_{x} to the law of the process XX starting at xx and for simplicity, we write \mathbb{P} instead of 0\mathbb{P}_{0}. Accordingly, we use 𝔼x\mathbbm{E}_{x} (resp. 𝔼\mathbbm{E}) to denote expectation operator associated to the law x\mathbb{P}_{x} (resp. \mathbb{P}). Additionally, we denote the natural filtration generated by the process XX by 𝔽:={t:t0}\mathbb{F}\raisebox{0.4pt}{$:$}=\{\mathcal{F}_{t}:t\geq 0\} satisfying the usual conditions.

A strategy π:={Ltπ:t0}\pi\raisebox{0.4pt}{$:$}=\{L_{t}^{\pi}:t\geq 0\} is a non-decreasing, right-continuous, and 𝔽\mathbb{F}-adapted process. Hence, for each strategy π\pi, the controlled process Xπ=:{Xtπ:t0}X^{\pi}=:\{X_{t}^{\pi}:t\geq 0\} becomes

Xtπ:=XtLtπ,for t0.X_{t}^{\pi}\raisebox{0.4pt}{$:$}=X_{t}-L_{t}^{\pi},\quad\text{for\ $t\geq 0$}.

We write τπ:=inf{t>0:Xtπ<0}\tau^{\pi}\raisebox{0.4pt}{$:$}=\inf\{t>0:X_{t}^{\pi}<0\} for the time of ruin, we say that a strategy is admissible if the controlled process XπX^{\pi} is not allowed to go below zero by the action of the control LπL^{\pi}, that is, if LtπLtπXtπL_{t}^{\pi}-L_{t-}^{\pi}\leq X_{t-}^{\pi} for t<τπt<\tau^{\pi}, where L0π:=0L_{0-}^{\pi}:=0 and X0π:=x0X_{0-}^{\pi}:=x\geq 0. The set of admissible strategies is denoted by 𝒮\mathcal{S}. The expected reward (ER) for each admissible strategy π𝒮\pi\in\mathcal{S} is given by

(1.1) Vπ(x):=𝔼x[0τπeqsg(Xsπ)dLsπ],x0,V_{\pi}(x)\raisebox{0.4pt}{$:$}=\mathbbm{E}_{x}\left[\int_{0}^{\tau^{\pi}}\operatorname{e}^{-qs}g(X^{\pi}_{s-})\circ{\rm d}L^{\pi}_{s}\right],\quad x\geq 0,

where q>0q>0 is the discount rate, g:(0,)(0,)g:(0,\infty)\mapsto(0,\infty) is a twice continuously differentiable function, and

(1.2) 0τπeqsg(Xsπ)dLsπ\displaystyle\int_{0}^{\tau^{\pi}}\operatorname{e}^{-qs}g(X^{\pi}_{s-})\circ{\rm d}L^{\pi}_{s} :=0τπeqsg(Xsπ)dLsπ,c\displaystyle\raisebox{0.4pt}{$:$}=\int_{0}^{\tau^{\pi}}\operatorname{e}^{-qs}g(X^{\pi}_{s}){\rm d}L^{\pi,c}_{s}
+0sτπeqsΔLsπ01g(Xsπ+ΔXsλΔLsπ)dλ,\displaystyle\quad+\sum_{0\leq s\leq\tau^{\pi}}\operatorname{e}^{-qs}\Delta L^{\pi}_{s}\int_{0}^{1}g(X^{\pi}_{s-}+\Delta X_{s}-\lambda\Delta L^{\pi}_{s}){\rm d}\lambda,

where Lπ,cL^{\pi,c} denotes the continuous part of LπL^{\pi}. Note that the strategy π\pi generates two types of rewards: the first one is related to the continuous part Lπ,cL^{\pi,c}, and the other to the jumps of the process LπL^{\pi}.

We want to maximize the performance criterion over the set of all admissible strategies 𝒮\mathcal{S} and find the value function of the problem given by

(1.3) V(x)=supπ𝒮Vπ(x),x0.\displaystyle V(x)=\sup_{\pi\in\mathcal{S}}V_{\pi}(x),\quad\text{$x\geq 0$}.

By the dynamic programming principle, we identify heuristically that VV is associated with the following Hamilton-Jacobi-Bellman (HJB) equation

(1.4) max{(q)V,gV}=0 on[0,),\max\{(\mathcal{L}-q)V,g-V^{\prime}\}=0\ \text{ on}\ [0,\infty),

where \mathcal{L} is the infinitesimal generator of the process XX as in (3.9) below.

To solve the optimization problem given in (1.3), we first propose a candidate strategy π\pi^{*} for being the optimal strategy (see Eqs. (3.5) and (4.27)), then verify, under some assumptions, that its ER VπV_{\pi^{*}} satisfies (1.4), and finally, using a verification lemma, conclude that π\pi^{*} is the optimal strategy for the problem; see Sections 3 and 4. To this end, we will compute VπV_{\pi^{*}} and express it in terms of the so-called scale functions; see Section 2.

2. Scale Functions

In this section, we will review qq-scale functions and provide some properties that will be used throughout this work. Let us begin with the Laplace exponent ψ:[0,)\psi:[0,\infty)\mapsto\mathbbm{R} of the process XX, defined by

ψ(θ):=log[𝔼[eθX1]]=μθ+σ22θ2(0,)(1eθzθz𝟏{0<z<1})Π(dz),\displaystyle\psi(\theta)\raisebox{0.4pt}{$:$}=\log\Big{[}\mathbbm{E}\Big{[}\operatorname{e}^{\theta X_{1}}\Big{]}\Big{]}=\mu\theta+\dfrac{\sigma^{2}}{2}\theta^{2}-\int_{(0,\infty)}\big{(}1-\operatorname{e}^{-\theta z}-\theta z\mathbf{1}_{\{0<z<1\}}\big{)}\Pi({\rm d}z),

where μ\mu\in\mathbbm{R}, σ0\sigma\geq 0 and the Lévy measure of XX, Π\Pi, is a measure defined on (0,)(0,\infty) satisfying

(0,)(1z2)Π(dz)<.\int_{(0,\infty)}(1\wedge z^{2})\Pi(\mathrm{d}z)<\infty.

For each q0q\geq 0, the qq-scale function of XX is a mapping W(q):[0,)W^{(q)}:\mathbbm{R}\longrightarrow[0,\infty), which is strictly increasing and continuous on [0,)[0,\infty) and such that W(q)(x)=0W^{(q)}(x)=0 for x<0x<0. It is characterized by its Laplace transform, given by

0eθxW(q)(x)dx=1ψ(θ)q,for θ>Φ(q),\int_{0}^{\infty}\operatorname{e}^{-\theta x}W^{(q)}(x)\mathrm{d}x=\dfrac{1}{\psi(\theta)-q},\quad\text{for $\theta>\Phi(q)$},

where Φ(q)=sup{θ0:ψ(θ)=q}\Phi(q)=\sup\{\theta\geq 0:\psi(\theta)=q\}. Additionally, we define for xx\in\mathbbm{R}

Z(q)(x)\displaystyle Z^{(q)}(x) :=1+q0xW(q)(y)dy.\displaystyle\raisebox{0.4pt}{$:$}=1+q\int_{0}^{x}W^{(q)}(y)\mathrm{d}y.

From Proposition 3 in [20] we know that qq-scale functions are harmonic for the discounted processes on (0,)(0,\infty), that is,

(2.1) (q)W(q)(x)=0and(q)Z(q)(x)=0,x>0.(\mathcal{L}-q)W^{(q)}(x)=0\quad\text{and}\quad(\mathcal{L}-q)Z^{(q)}(x)=0,\quad x>0.

As is usual in singular control problems for Lévy processes (see for instance [17]) we will assume the following:

Assumption 2.1.

The Lévy measure Π\Pi of the process XX has a completely monotone density. That is, Π\Pi has a density ν\nu whose nn-th derivative ν(n)\nu^{(n)} exists for all n1n\geq 1 and satisfies

(1)nν(n)(x)0,x>0.\displaystyle(-1)^{n}\nu^{(n)}(x)\geq 0,\quad x>0.

The previous assumption allows us to obtain a more explicit form of the qq-scale function as seen in the following result.

Lemma 2.1 ([18], Theorem 2 and Corollary 1).

For q>0q>0 and under Assumption 2.1, the qq-scale function W(q)W^{(q)} can be written as

(2.2) W(q)(x)=eΦ(q)xψ(Φ(q))f^(x),\displaystyle W^{(q)}(x)=\dfrac{\operatorname{e}^{\Phi(q)x}}{\psi^{\prime}(\Phi(q))}-\hat{f}(x),

where f^\hat{f} is a non-negative, completely monotone function given by f^(x)=0+extμ^(dt)\hat{f}(x)=\displaystyle\int_{0+}^{\infty}\operatorname{e}^{-xt}\hat{\mu}({\rm d}t), where μ^\hat{\mu} is a finite measure on (0,)(0,\infty). Moreover, W(q)W^{(q)\prime} is strictly log-convex (and hence convex) and W(q)W^{(q)} is CC^{\infty} on (0,)(0,\infty).

Let us define the first down- and up-crossing times, respectively, by

τa:=inf{t>0:Xt<a}andτa+:=inf{t>0:Xt>a},a+,\displaystyle\tau_{a}^{-}\raisebox{0.4pt}{$:$}=\inf\left\{t>0:X_{t}<a\right\}\quad\textrm{and}\quad\tau_{a}^{+}\raisebox{0.4pt}{$:$}=\inf\left\{t>0:X_{t}>a\right\},\quad a\in\mathbbm{R}_{+},

where we follow the convention that inf=\inf\emptyset=\infty. By Theorem 8.1 in [16], for any a>ba>b and xax\leq a,

(2.3) 𝔼x[eqτa+𝟏{τa+<τb}]=W(q)(xb)W(q)(ab),𝔼x[eqτb𝟏{τa+>τb}]=Z(q)(xb)Z(q)(ab)W(q)(xb)W(q)(ab).\displaystyle\begin{split}\mathbbm{E}_{x}\left[\operatorname{e}^{-q\tau_{a}^{+}}\mathbf{1}_{\left\{\tau_{a}^{+}<\tau_{b}^{-}\right\}}\right]&=\frac{W^{(q)}(x-b)}{W^{(q)}(a-b)},\\ \mathbbm{E}_{x}\left[\operatorname{e}^{-q\tau_{b}^{-}}\mathbf{1}_{\left\{\tau_{a}^{+}>\tau_{b}^{-}\right\}}\right]&=Z^{(q)}(x-b)-Z^{(q)}(a-b)\frac{W^{(q)}(x-b)}{W^{(q)}(a-b)}.\end{split}

For the next result we recall that by Lemma 2.1 the scale function W(q)W^{(q)} is CC^{\infty} on (0,)(0,\infty).

Remark 2.1.

Note that since W(q)W^{(q)\prime} is strictly log-convex on (0,)(0,\infty) it follows that W(q)′′W(q)\dfrac{W^{(q)\prime\prime}}{W^{(q)\prime}} is strictly increasing on (0,)(0,\infty). Additionally, as in the proof of Theorem 1 in [18] we have that W(q)W^{(q)\prime} has a unique minimum denoted by aa^{*}, hence W(q)′′W^{(q)\prime\prime} is strictly negative on (0,a)(0,a^{*}) and strictly positive on (a,)(a^{*},\infty). The latter implies that

{limu0W(q)′′(u)W(q)(u)[0,),ifa=0,limu0W(q)′′(u)W(q)(u)[,0),ifa>0.\begin{cases}\displaystyle\lim_{u\downarrow 0}\dfrac{W^{(q)\prime\prime}(u)}{W^{(q)\prime}(u)}\in[0,\infty),&\text{if}\ a^{*}=0,\\ \displaystyle\lim_{u\downarrow 0}\dfrac{W^{(q)\prime\prime}(u)}{W^{(q)\prime}(u)}\in[-\infty,0),&\text{if}\ a^{*}>0.\end{cases}

On the other hand, using (2.2)

(2.4) limuW(q)′′(u)W(q)(u)=limuΦ(q)ψ(ϕ(q))Φ(q)1eΦ(q)u0+t2eutμ^(dt)1+ψ(ϕ(q))Φ(q)1eΦ(q)u0+teutμ^(dt)=Φ(q).\lim_{u\uparrow\infty}\dfrac{W^{(q)\prime\prime}(u)}{W^{(q)\prime}(u)}=\lim_{u\uparrow\infty}\frac{\Phi(q)-\psi^{\prime}(\phi(q))\Phi(q)^{-1}e^{-\Phi(q)u}\int_{0+}^{\infty}t^{2}e^{-ut}\hat{\mu}(dt)}{1+\psi^{\prime}(\phi(q))\Phi(q)^{-1}e^{-\Phi(q)u}\int_{0+}^{\infty}te^{-ut}\hat{\mu}(dt)}=\Phi(q).

In the case of Brownian motion with drift, i.e. X=x+μt+σBtX=x+\mu t+\sigma B_{t}, where B={Bt:t0}B=\{B_{t}:t\geq 0\} is a Brownian motion, we know (see for instance pg. 10 in [14]) that for x0x\geq 0

(2.5) W(q)(x)=1μ2+2qσ2[eΦ(q)xeζ1x]andZ(q)(x)=qμ2+2qσ2[eΦ(q)xΦ(q)+eζ1xζ1],\displaystyle W^{(q)}(x)=\frac{1}{\sqrt{\mu^{2}+2q\sigma^{2}}}\left[\operatorname{e}^{\Phi(q)x}-\operatorname{e}^{-\zeta_{1}x}\right]\quad\text{and}\quad Z^{(q)}(x)=\frac{q}{\sqrt{\mu^{2}+2q\sigma^{2}}}\left[\frac{\operatorname{e}^{\Phi(q)x}}{\Phi(q)}+\frac{\operatorname{e}^{-\zeta_{1}x}}{\zeta_{1}}\right],

where

ζ1=1σ2(μ2+2qσ2+μ)andΦ(q)=1σ2(μ2+2qσ2μ).\displaystyle\zeta_{1}=\frac{1}{\sigma^{2}}\left(\sqrt{\mu^{2}+2q\sigma^{2}}+\mu\right)\quad\text{and}\quad\Phi(q)=\frac{1}{\sigma^{2}}\left(\sqrt{\mu^{2}+2q\sigma^{2}}-\mu\right).

Note that,

(2.6) W(q)(0+)=2σ2W^{(q)\prime}(0+)=\frac{2}{\sigma^{2}}

and

(2.7) limu0W(q)′′(u)W(q)(u)=2μσ2.\lim_{u\downarrow 0}\dfrac{W^{(q)\prime\prime}(u)}{W^{(q)\prime}(u)}=-\dfrac{2\mu}{\sigma^{2}}.

In this case, several useful properties of the scale functions are provided in Appendix A.

3. Optimality of one-barrier strategies for spectrally negative Lévy processes

In this section, we will assume that XX is a spectrally negative Lévy process whose Lévy measure Π\Pi has a completely monotone density; as in Assumption 2.1. We aim to find conditions where one-barrier strategies are optimal. Hence, we will begin this section by computing the ER given in (1.1) in terms of scale functions for an arbitrary one-barrier strategy (see (3.2)). Then, we will propose a candidate for the optimal barrier 𝒷\mathpzc{b}^{*} (see (3.5)) and verify that the barrier strategy at level 𝒷\mathpzc{b}^{*} is indeed optimal for the optimization problem given in (1.3) under suitable assumptions. Proofs of the results of this section are found in Appendix B.

3.1. Computation of the ER

A barrier strategy π𝒷={Lt𝒷:t0}\pi_{\mathpzc{b}}=\{L^{\mathpzc{b}}_{t}:t\geq 0\} at level 𝒷0\mathpzc{b}\geq 0, is defined by

Lt𝒷=(sup0st{Xs𝒷})0,fort0.L_{t}^{\mathpzc{b}}=\left(\sup_{0\leq s\leq t}\{X_{s}-\mathpzc{b}\}\right)\vee 0,\quad\text{for}\ t\geq 0.

Observe that π𝒷\pi_{\mathpzc{b}} is indeed an admissible strategy, which is continuous if x𝒷x\leq\mathpzc{b}, and has a unique jump of size x𝒷x-\mathpzc{b} at time zero if x>𝒷x>\mathpzc{b}, where xx is the starting value of XX. We denote by Xt𝒷=XtLt𝒷X^{\mathpzc{b}}_{t}=X_{t}-L_{t}^{\mathpzc{b}}.

Let V𝒷V_{\mathpzc{b}} be the ER for the barrier strategy at level 𝒷\mathpzc{b}, i.e,

(3.1) V𝒷(x):=𝔼x[0τ𝒷eqtg(Xt𝒷)dLt𝒷],x0,\displaystyle V_{\mathpzc{b}}(x)\raisebox{0.4pt}{$:$}=\mathbbm{E}_{x}\left[\int_{0}^{\tau_{\mathpzc{b}}}\operatorname{e}^{-qt}g(X^{\mathpzc{b}}_{t-})\circ{\rm d}L^{\mathpzc{b}}_{t}\right],\quad\text{$x\geq 0$,}

with τ𝒷:=inf{t0:Xt𝒷<0}\tau_{\mathpzc{b}}\raisebox{0.4pt}{$:$}=\displaystyle\inf\{t\geq 0:X^{\mathpzc{b}}_{t}<0\}. Using the strong Markov property, we provide an expression for (3.1) in terms of scale functions.

Proposition 3.1.

Let V𝒷V_{\mathpzc{b}} be as in (3.1). Then

(3.2) V𝒷(x)={g(𝒷)W(q)(𝒷)W(q)(x)if x𝒷,H(x;𝒷)if x𝒷,V_{\mathpzc{b}}(x)=\begin{cases}\displaystyle\frac{g(\mathpzc{b})}{W^{(q)\prime}(\mathpzc{b})}W^{(q)}(x)&\text{if }\ x\leq\mathpzc{b},\\ H(x;\mathpzc{b})&\text{if }\ x\geq\mathpzc{b},\end{cases}

where

(3.3) H(x;u)=G(x)G(u)+g(u)W(q)(u)W(q)(u)andG(x)=0xg(y)𝑑y.H(x;u)=G(x)-G(u)+g(u)\dfrac{W^{(q)}(u)}{W^{(q)\prime}(u)}\quad\text{and}\quad G(x)=\displaystyle\int_{0}^{x}g(y)dy.

3.2. Selection of the optimal threshold

In order to maximize Vb(x)V{b}(x) uniformly in xx, by looking at equation (3.2), we consider the mapping F:[0,)F:[0,\infty)\mapsto\mathbbm{R} by

(3.4) F(u)=g(u)W(q)(u).F(u)=\dfrac{g(u)}{W^{(q)\prime}(u)}.

Then, we define our choice of optimal threshold 𝒷\mathpzc{b}^{*} by

(3.5) 𝒷:=sup{𝓊0:(𝓊)(𝓍), for all x0}.\displaystyle\mathpzc{b}^{*}:=\sup\{u\geq 0:F(u)\geq F(x),\ \text{ for all $x\geq 0$}\}.

Observe that 𝒷𝒶<\mathpzc{b}^{*}\leq a^{*}<\infty (with aa^{*} as in Remark 2.1) if gg is non-increasing on (0,)(0,\infty). To guarantee that b<b^{*}<\infty we will make the following assumption throughout the paper.

Assumption 3.1.

We assume that g(z)=o(eϕ(q)z)g(z)=o(e^{\phi(q)z}) as zz\rightarrow\infty.

Note that (2.2) implies that g(z)=o(W(q)(z))g(z)=o(W^{(q)}(z)) and also that g(z)=o(W(q)(z))g(z)=o(W^{(q)\prime}(z)), hence b<b^{*}<\infty. Now, from (3.2) we obtain

(3.6) Vb(u)={g(𝒷)W(q)(𝒷)W(q)(u)if u<𝒷,g(u)if u>𝒷, and Vb′′(u)={g(𝒷)W(q)(𝒷)W(q)′′(u)if u<𝒷,g(u)if u>𝒷.V{b}^{\prime}(u)=\begin{cases}\displaystyle\frac{g(\mathpzc{b})}{W^{(q)\prime}(\mathpzc{b})}W^{(q)\prime}(u)&\mbox{if }u<\mathpzc{b},\\ g(u)&\mbox{if }u>\mathpzc{b},\end{cases}\quad\textrm{ and }\quad V{b}^{\prime\prime}(u)=\begin{cases}\displaystyle\frac{g(\mathpzc{b})}{W^{(q)\prime}(\mathpzc{b})}W^{(q)\prime\prime}(u)&\mbox{if }u<\mathpzc{b},\\ g^{\prime}(u)&\mbox{if }u>\mathpzc{b}.\end{cases}

Hence, for each 𝒷0\mathpzc{b}\geq 0, V𝒷V_{\mathpzc{b}} is increasing and continuously differentiable on (0,)(0,\infty) since W(q)W^{(q)\prime} and gg are positive on (0,)(0,\infty). Additionally, FF is continuously differentiable and

(3.7) F(u)=g(u)W(q)(u)(g(u)g(u)W(q)′′(u)W(q)(u)),\displaystyle F^{\prime}(u)=\dfrac{g(u)}{W^{(q)\prime}(u)}\Bigg{(}\dfrac{g^{\prime}(u)}{g(u)}-\dfrac{W^{(q)\prime\prime}(u)}{W^{(q)\prime}(u)}\Bigg{)},

then,

(3.8) g(𝒷)g(𝒷)=W(q)′′(𝒷)W(q)(𝒷)if𝒷>0.\dfrac{g^{\prime}(\mathpzc{b}^{*})}{g(\mathpzc{b}^{*})}=\dfrac{W^{(q)\prime\prime}(\mathpzc{b}^{*})}{W^{(q)\prime}(\mathpzc{b}^{*})}\quad\text{if}\ \mathpzc{b}^{*}>0.

Therefore, from (3.6) and (3.8), V𝒷V_{\mathpzc{b}^{*}} is twice continuously differentiable. We summarize these results in the next lemma.

Lemma 3.1.

For any 𝒷0\mathpzc{b}\geq 0, V𝒷C1((0,))C2((0,)){𝒷})V_{\mathpzc{b}}\in C^{1}((0,\infty))\cap C^{2}((0,\infty))\setminus\{\mathpzc{b}\}) and V𝒷>0V_{\mathpzc{b}}>0 is increasing on (0,)(0,\infty). Additionally, V𝒷C2((0,))V_{\mathpzc{b}^{*}}\in C^{2}((0,\infty)), with 𝒷\mathpzc{b}^{*} defined as in (3.5).

3.3. Verification of optimality

We show now that the optimality of the stochastic control problem given in (1.3) is achieved by the barrier strategy at level 𝒷\mathpzc{b}^{*} under Condition (3.11) below. Let \mathcal{L} denote the infinitesimal generator of the process XX, acting on the space of sufficiently smooth functions i.e. C1((0,))C^{1}((0,\infty)) (resp. C2((0,))C^{2}((0,\infty))) if XX is of bounded variation (resp. unbounded variation). The infinitesimal generator is given by

(3.9) f(x)=σ22f′′(x)+μf(x)+(0,)[f(xz)f(x)+f(x)z𝟏{0<z<1}]Π(dz).\begin{split}\mathcal{L}f(x)=\frac{\sigma^{2}}{2}f^{\prime\prime}(x)+\mu f^{\prime}(x)+\int_{(0,\infty)}[f(x-z)-f(x)+f^{\prime}(x)z\mathbf{1}_{\{0<z<1\}}]\Pi(\mathrm{d}z).\end{split}
Lemma 3.2 (Verification Lemma).

Suppose that π^𝒮\hat{\pi}\in\mathcal{S} is such that Vπ^V_{\hat{\pi}} is sufficiently smooth on (0,)(0,\infty), right-continuous at 0, and, for all x>0x>0,

(3.10) max{(q)Vπ^(x),g(x)Vπ^(x)}0,\max\{(\mathcal{L}-q)V_{\hat{\pi}}(x),g(x)-V^{\prime}_{\hat{\pi}}(x)\}\leq 0,

Then Vπ^(x)=V(x)V_{\hat{\pi}}(x)=V(x) for all x0x\geq 0 and, hence, π^\hat{\pi} is an optimal strategy.

The previous lemma is the main tool to prove the main result of this section.

Theorem 3.1.

Assume that 𝒷<\mathpzc{b}^{*}<\infty with 𝒷\mathpzc{b}^{*} as in (3.5), and

(3.11) g(u)g(u)W(q)′′(u)W(q)(u)for u(𝒷,).\dfrac{g^{\prime}(u)}{g(u)}\leq\dfrac{W^{(q)\prime\prime}(u)}{W^{(q)\prime}(u)}\quad\text{for $u\in(\mathpzc{b}^{*},\infty)$}.

Then the barrier strategy π𝒷\pi_{\mathpzc{b}^{*}} is optimal and V(x)=V𝒷(x)V(x)=V_{\mathpzc{b}^{*}}(x) for all x0x\geq 0.

Remark 3.1.
  1. (i)

    Notice that Condition (3.11) is equivalent to the assumption that F0F^{\prime}\leq 0 on [𝒷,)[\mathpzc{b}^{*},\infty).

  2. (ii)

    Taking g1g\equiv 1, we see that (3.11) is equivalent to

    (3.12) W(q)′′0on(𝒷,),W^{(q)\prime\prime}\geq 0\quad\text{on}\ (\mathpzc{b}^{*},\infty),

    since W(q)>0W^{(q)\prime}>0 on (0,)(0,\infty). From here and Remark 2.1 we have that 𝒷=𝒶\mathpzc{b}^{*}=a^{*}. Notice that (3.12) is equivalent to the criterion used by Loeffen to verify that aa^{*} is the optimal barrier for the case mentioned above; see Theorem 2 in [17].

  3. (iii)

    By (2.1), (3.2), and (3.10), it is enough to show that

    (3.13) (q)V𝒷0on(𝒷,)(\mathcal{L}-q)V_{\mathpzc{b}^{*}}\leq 0\quad\text{on}\ (\mathpzc{b}^{*},\infty)

    to verify the statement given in Theorem 3.1, since (q)V𝒷=0(\mathcal{L}-q)V_{\mathpzc{b}^{*}}=0 and gV𝒷0g-V^{\prime}_{\mathpzc{b}^{*}}\leq 0 hold on (0,𝒷)(0,\mathpzc{b}^{*}) and (0,)(0,\infty), respectively; see Appendix B for more details.

Remark 3.2.

Consider that the underlying process XX is given by the solution to the following stochastic differential equation

dXt=μ¯(Xt)dt+σ¯(Xt)dWtt0,\mathrm{d}X_{t}=\bar{\mu}(X_{t})\mathrm{d}t+\bar{\sigma}(X_{t})\mathrm{d}W_{t}\quad t\geq 0,

where μ¯:\bar{\mu}:\mathbbm{R}\mapsto\mathbbm{R} and σ¯:(0,)\bar{\sigma}:\mathbbm{R}\mapsto(0,\infty) are Lipschitz continuous functions. Defining ψ\psi as a combination of the fundamental solutions of the ordinary second-order differential equation [𝒜q]u(x)=0[\mathcal{A}-q]u(x)=0, with 𝒜\mathcal{A} the infinitesimal generator of XX (for more details, see [10], Ch. II), under the assumption that gC((0,))C1((0,)D)g\in C((0,\infty))\cup C^{1}((0,\infty)\setminus D), with D(0,)D\subset(0,\infty) a countable set, is non-increasing, and that the function f(x):=g(x)/ψ(x)f(x)\raisebox{0.4pt}{$:$}=g(x)/\psi^{\prime}(x) has a unique interior maximum at some point 𝒷(0,)\mathpzc{b}^{*}\in(0,\infty), where ff is non-increasing on (𝒷,)(\mathpzc{b}^{*},\infty), Alvarez in [3] showed that an optimal control strategy is achieved through the 𝒷\mathpzc{b}^{*}-barrier strategy. However, if the condition

[σ¯(x)]22g(x)+μ¯(x)g(x)qG(x)0forx(0,)D,\dfrac{[\bar{\sigma}(x)]^{2}}{2}g^{\prime}(x)+\bar{\mu}(x)g(x)-qG(x)\leq 0\quad\text{for}\ x\in(0,\infty)\setminus D,

holds, he verified that the optimal policy is to drive the process instantaneously to the origin. In our case, notice that this condition is equivalent to checking that (3.13) holds on (0,)(0,\infty).

3.3.1. Examples

In the next two examples it is straightforward to verify that gg meets Assumption 3.1.

  1. (1)

    Consider g(x)=xαg(x)=x^{\alpha} with α>0\alpha>0 fixed. denote

    (3.14) h(x)=g(x)g(x)W(q)′′(x)W(q)(x),x>0.h(x)=\frac{g^{\prime}(x)}{g(x)}-\dfrac{W^{(q)\prime\prime}(x)}{W^{(q)\prime}(x)},\quad x>0.

    Since g(x)g(x)=αx\dfrac{g^{\prime}(x)}{g(x)}=\dfrac{\alpha}{x} and W(q)′′(x)W(q)(x)-\dfrac{W^{(q)\prime\prime}(x)}{W^{(q)\prime}(x)} are non-increasing, we have that xh(x)x\mapsto h(x) is also non-increasing satisfying

    limu0h(u)=+and,limuh(u)=Φ(q).\lim_{u\downarrow 0}h(u)=+\infty\quad\text{and,}\quad\lim_{u\uparrow\infty}h(u)=-\Phi(q).

    Therefore, there must exist a unique b^>0\hat{b}>0 such that h(b^)=0h(\hat{b})=0 which, by (3.7) together with the fact that gg and W(q)W^{(q)} are strictly positive functions in (0,)(0,\infty), imply that b^\hat{b} is the unique root of the mapping xF(x)x\mapsto F^{\prime}(x), and therefore b^=𝒷\hat{b}=\mathpzc{b}^{*}. Additionally, note that Condition (3.11) is satisfied because the function hh is non-increasing. Hence, by Theorem 3.1 the barrier strategy π𝒷\pi_{\mathpzc{b}^{*}} is optimal.

  2. (2)

    Let β>0\beta>0 and g(x)=eβxg(x)=\operatorname{e}^{-\beta x}, then hh is non-increasing and therefore FF has a unique maximum at 𝒷>0\mathpzc{b}^{*}>0 if

    W(q)′′(𝒷)W(q)(𝒷)=g(𝒷)g(𝒷)=β,\dfrac{W^{(q)\prime\prime}(\mathpzc{b}^{*})}{W^{(q)\prime}(\mathpzc{b}^{*})}=\dfrac{g^{\prime}(\mathpzc{b}^{*})}{g(\mathpzc{b}^{*})}=-\beta,

    which is true only if a0a^{*}\neq 0 by Remark 2.1 and (2.4). On the other hand, if a=0a^{*}=0, it follows that g(x)g(x)=β<W(q)′′(x)W(q)(x)\dfrac{g^{\prime}(x)}{g(x)}=-\beta<\dfrac{W^{(q)\prime\prime}(x)}{W^{(q)\prime}(x)} for all x(0,)x\in(0,\infty), which implies that 𝒷=0\mathpzc{b}^{*}=0. Hence, Condition (3.11) is satisfied, and therefore the barrier strategy π0\pi_{0} is optimal by an application of Theorem 3.1.

3.4. Necessity of Condition (3.11)

It is important to remark that Condition (3.11) is sufficient but not necessary for the optimality of the one-barrier strategy. Indeed, in the next example, where the uncontrolled process XX is a standard Brownian motion, i.e., Xt=x+μt+σWtX_{t}=x+\mu t+\sigma W_{t}, we will show the optimality of the one-barrier strategy even when Condition (3.11) is not satisfied. For that, by Remark (iii), it is enough to check that (3.13) holds. Let g:[0,)[0,1)g:[0,\infty)\rightarrow[0,1) be defined as

(3.15) g(x)=0.3x20.5x30.32x+0.2forx0,g(x)=\dfrac{0.3x^{2}}{0.5x^{3}-0.32x+0.2}\quad\text{for}\ x\geq 0,

and let q=0.2q=0.2, μ=2.3\mu=2.3 and σ=2\sigma=2. Notice that this function also satisfies Assumption 3.1. We obtain numerically that FF attains its global maximum at 𝒷=0.8925\mathpzc{b}^{*}=0.8925. Condition (3.11) is not satisfied but (q)V𝒷(x)<0(\mathcal{L}-q)V_{\mathpzc{b}^{*}}(x)<0 for x(𝒷,)x\in(\mathpzc{b}^{*},\infty), see Figure 1. Then, by Lemma 3.2, the barrier strategy at the level 𝒷\mathpzc{b}^{*} is optimal. Note that the function gg taking the value 0 does not pose an issue. For the process XX, 0 is regular for (,0)(-\infty,0), meaning that when XX reaches 0, ruin occurs instantaneously.

Refer to caption
Figure 1. Plots of FF (in blue) and (q)V0.8925(\mathcal{L}-q)V_{0.8925} (in red). The plot starts at the point 𝒷=0.8925\mathpzc{b}^{*}=0.8925, where FF attains its global maximum, but Condition (3.11) is not satisfied since FF^{\prime} eventually becomes positive (see Remark 3.1). On the other hand, (q)V0.8925(x)(\mathcal{L}-q)V_{0.8925}(x) is always negative for x0.8925x\geq 0.8925.

Now, if we choose μ=2.4\mu=2.4, we get numerically that FF attains its global maximum at 𝒷=0.9165\mathpzc{b}^{*}=0.9165. Condition (3.11) is not satisfied and there exists x¯>𝒷\bar{x}>\mathpzc{b}^{*} such that (q)V𝒷(x¯)>0(\mathcal{L}-q)V_{\mathpzc{b}^{*}}(\bar{x})>0, see Figure 2. In the next section, we will show that for this case the optimal strategy corresponds to a multibarrier strategy.

Refer to caption
Figure 2. Plots of FF (in blue) and (q)V0.9165(\mathcal{L}-q)V_{0.9165} (in red). The plot starts at the point 𝒷=0.9165\mathpzc{b}^{*}=0.9165, where FF attains its global maximum, but Condition (3.11) is not satisfied since FF^{\prime} eventually becomes positive (see Remark 3.1). On the other hand, (q)V0.9165(x¯)(\mathcal{L}-q)V_{0.9165}(\bar{x}) is positive for some x¯0.9165\bar{x}\geq 0.9165.

4. Optimality of multibarrier strategies for Brownian motion with drift

In Subsection 3.4, we showed that under a particular choice of parameters, the one-barrier strategy at level 𝒷\mathpzc{b}^{*} is optimal even when Condition (3.11) is not satisfied. However, we can find different choices of parameters under which the HJB inequality given in (3.10) is not satisfied for the ER associated with the strategy π𝒷\pi_{\mathpzc{b}^{*}}. This suggests that not always the one-barrier strategies are optimal, so it is necessary to find other types of strategies that can achieve optimality.

In this section, we will propose the (2n+1)(2n+1)-barriers strategy as our candidate for optimality among admissible strategies. Due to the complexity that arises from the jumps of the process XX when working with multibarrier strategies, in the remainder of the paper, we will assume that XX is a Brownian motion with drift, that is, Π0\Pi\equiv 0. Proofs of the results of this section are provided in Appendix C.

We denote the (2n+1)(2n+1)-barriers by 𝕓={𝒷𝒾}𝒾=12𝓃+1\mathbbmss{b}=\{\mathpzc{b}_{i}\}_{i=1}^{2n+1} where 0𝒷1<𝒷2<<𝒷2𝓃+10\leq\mathpzc{b}_{1}<\mathpzc{b}_{2}<\dots<\mathpzc{b}_{2n+1}, and describe the 𝕓\mathbbmss{b}-barrier strategy as follows: If the process lies above 𝒷2𝓃+1\mathpzc{b}_{2n+1}, push the process to the level 𝒷2𝓃+1\mathpzc{b}_{2n+1}; if it lies between (𝒷2𝓀,𝒷2𝓀+1)(\mathpzc{b}_{2k},\mathpzc{b}_{2k+1}), with k{0,1,,n}k\in\{0,1,\dots,n\}, do nothing, and finally if it lies between [𝒷2𝓀+1,𝒷2𝓀+2][\mathpzc{b}_{2k+1},\mathpzc{b}_{2k+2}], with k{0,1,,n1}k\in\{0,1,\dots,n-1\}, push the process to 𝒷2𝓀+1\mathpzc{b}_{2k+1}. We denote the 𝕓\mathbbmss{b}-barrier strategy by L𝕓L^{\mathbbmss{b}} and the controlled process by X𝕓=XL𝕓X^{\mathbbmss{b}}=X-L^{\mathbbmss{b}}. Formally, if 𝕓={𝒷𝒾}𝒾=12𝓃+1\mathbbmss{b}=\{\mathpzc{b}_{i}\}_{i=1}^{2n+1}, we define the controlled process X𝕓=XL𝕓X^{\mathbbmss{b}}=X-L^{\mathbbmss{b}} under x\mathbbm{P}_{x} as follows: Assume that x[𝒷2𝓀,𝒷2𝓀+2)x\in[\mathpzc{b}_{2k},\mathpzc{b}_{2k+2}) for k=0,1,,nk=0,1,\dots,n, where we take 𝒷0=0\mathpzc{b}_{0}=0 and 𝒷2𝓃+2=\mathpzc{b}_{2n+2}=\infty.

  1. (1)

    Let Lt𝕓=(sup0st{Xs𝒷2𝓀+1})0L^{\mathbbmss{b}}_{t}=\left(\sup\limits_{0\leq s\leq t}\{X_{s}-\mathpzc{b}_{2k+1}\}\right)\vee 0 and Xk=XL𝕓X^{k}=X-L^{\mathbbmss{b}}, hence {Xtk:t0}\{X^{k}_{t}:t\geq 0\} is a process reflected at 𝒷2𝓀+1\mathpzc{b}_{2k+1} that starts at xx. Define σk=inf{t0:Xtk=𝒷2𝓀}\sigma_{k}=\inf\{t\geq 0:X^{k}_{t}=\mathpzc{b}_{2k}\}, then Xt𝕓=XtkX^{\mathbbmss{b}}_{t}=X^{k}_{t} for 0tσk0\leq t\leq\sigma_{k}.

  2. (2)

    For i=k1,k2,,1,0i=k-1,k-2,\ldots,1,0, define Zti=XtLσi+1𝕓Z^{i}_{t}=X_{t}-L^{\mathbbmss{b}}_{\sigma_{i+1}-}, so that Zσi+1i=𝒷2𝒾+2Z^{i}_{\sigma_{i+1}}=\mathpzc{b}_{2i+2}, and let Lt𝕓=Lσi+1𝕓+(supσi+1st{Zsi𝒷2𝒾+1})0L^{\mathbbmss{b}}_{t}=L^{\mathbbmss{b}}_{\sigma_{i+1}-}+\left(\sup\limits_{\sigma_{i+1}\leq s\leq t}\{Z^{i}_{s}-\mathpzc{b}_{2i+1}\}\right)\vee 0 and Xi=Zi(L𝕓Lσi+1𝕓)X^{i}=Z^{i}-(L^{\mathbbmss{b}}-L^{\mathbbmss{b}}_{\sigma_{i+1}-}). Hence {Xti:tσi+1}\{X^{i}_{t}:t\geq\sigma_{i+1}\} is a process reflected at 𝒷2𝒾+1\mathpzc{b}_{2i+1} that starts at 𝒷2𝒾+1\mathpzc{b}_{2i+1}. Define σi=inf{tσi+1:Xti=𝒷2𝒾}\sigma_{i}=\inf\{t\geq\sigma_{i+1}:X^{i}_{t}=\mathpzc{b}_{2i}\} and Xt𝕓=XtiX^{\mathbbmss{b}}_{t}=X^{i}_{t} for σi+1tσi\sigma_{i+1}\leq t\leq\sigma_{i}.

4.1. Computation of the ER

Consider the ER associated with the 𝕓\mathbbmss{b}-barrier strategy, given by

(4.1) V𝕓(x)=𝔼x[0τ𝕓eqtg(Xt𝕓)dLt𝕓],for x0,\displaystyle V_{\mathbbmss{b}}(x)=\mathbbm{E}_{x}\left[\int_{0}^{\tau_{\mathbbmss{b}}}\operatorname{e}^{-qt}g(X^{\mathbbmss{b}}_{t-})\circ{\rm d}L^{\mathbbmss{b}}_{t}\right],\quad\text{for $x\geq 0$,}

where τ𝕓:=inf{t0:Xt𝕓<0}\tau_{\mathbbmss{b}}\raisebox{0.4pt}{$:$}=\displaystyle\inf\{t\geq 0:X^{\mathbbmss{b}}_{t}<0\}. The next result gives the explicit estimations of V𝕓V_{\mathbbmss{b}} in terms of the qq-scale functions that are associated with the process XX.

Proposition 4.1.

Let V𝕓V_{\mathbbmss{b}} be as in (4.1). Then

(4.2) V𝕓(x)={g(𝒷1)W(q)(𝒷1)W(q)(x)if x<𝒷1,H(x;𝒷1)if x[𝒷1,𝒷2],ϕ(x;𝕓2k+1)if x(𝒷2𝓀,𝒷2𝓀+1),H(x;𝕓2k+1)if x[𝒷2𝓀+1,𝒷2𝓀+2],ϕ(x;𝕓)if x(𝒷2𝓃,𝒷2𝓃+1),H(x;𝕓)if x[𝒷2𝓃+1,),V_{\mathbbmss{b}}(x)=\begin{cases}\displaystyle\frac{g(\mathpzc{b}_{1})}{W^{(q)\prime}(\mathpzc{b}_{1})}W^{(q)}(x)&\mbox{if }x<\mathpzc{b}_{1},\\ H(x;\mathpzc{b}_{1})&\mbox{if }x\in[\mathpzc{b}_{1},\mathpzc{b}_{2}],\\ \qquad\vdots&\qquad\vdots\\ \phi(x;\mathbbmss{b}_{2k+1})&\mbox{if }x\in(\mathpzc{b}_{2k},\mathpzc{b}_{2k+1}),\\ H(x;\mathbbmss{b}_{2k+1})&\mbox{if }x\in[\mathpzc{b}_{2k+1},\mathpzc{b}_{2k+2}],\\ \qquad\vdots&\qquad\vdots\\ \phi(x;\mathbbmss{b})&\mbox{if }x\in(\mathpzc{b}_{2n},\mathpzc{b}_{2n+1}),\\ H(x;\mathbbmss{b})&\mbox{if }x\in[\mathpzc{b}_{2n+1},\infty),\end{cases}

where H(x;𝒷1)H(x;\mathpzc{b}_{1}) is defined in (3.3), 𝕓2k+1:={𝒷𝒾}𝒾=12𝓀+1\mathbbmss{b}_{2k+1}\raisebox{0.4pt}{$:$}=\{\mathpzc{b}_{i}\}_{i=1}^{2k+1} for each k{1,2,,n}k\in\{1,2,\dots,n\}, and

(4.3) ϕ(x;𝕓2k+1)\displaystyle\ \phi(x;\mathbbmss{b}_{2k+1}) :=H(𝒷2𝓀;𝕓2𝓀1)𝒵(𝓆)(𝓍𝒷2𝓀)\displaystyle:=H(\mathpzc{b}_{2k};\mathbbmss{b}_{2k-1})Z^{(q)}(x-\mathpzc{b}_{2k})
+W(q)(x𝒷2𝓀)((𝒷2𝓀+1)𝓆(𝒷2𝓀;𝕓2𝓀1)𝒲(𝓆)(𝒷2𝓀+1𝒷2𝓀)𝒲(𝓆)(𝒷2𝓀+1𝒷2𝓀)),\displaystyle\quad+W^{(q)}(x-\mathpzc{b}_{2k})\bigg{(}\frac{g(\mathpzc{b}_{2k+1})-qH(\mathpzc{b}_{2k};\mathbbmss{b}_{2k-1})W^{(q)}(\mathpzc{b}_{2k+1}-\mathpzc{b}_{2k})}{W^{(q)\prime}(\mathpzc{b}_{2k+1}-\mathpzc{b}_{2k})}\bigg{)},
(4.4) H(x;𝕓2k+1)\displaystyle H(x;\mathbbmss{b}_{2k+1}) :=G(x)G(𝒷2𝓀+1)+ϕ(𝒷2𝓀+1;𝕓2𝓀+1).\displaystyle:=G(x)-G(\mathpzc{b}_{2k+1})+\phi(\mathpzc{b}_{2k+1};\mathbbmss{b}_{2k+1}).

Using (2.5) and (4.2)–(4.4), we have for k=1,,nk=1,\dots,n, the following conditions:

V𝕓(𝒷1+)\displaystyle V_{\mathbbmss{b}}(\mathpzc{b}_{1}+) =H(𝒷1;𝒷1)=(𝒷1)𝒲(𝓆)(𝒷1)𝒲(𝓆)(𝒷1)=𝒱𝕓(𝒷1),\displaystyle=H(\mathpzc{b}_{1};\mathpzc{b}_{1})=g(\mathpzc{b}_{1})\frac{W^{(q)}(\mathpzc{b}_{1})}{W^{(q)\prime}(\mathpzc{b}_{1})}=V_{\mathbbmss{b}}(\mathpzc{b}_{1}-),
V𝕓(𝒷2𝓀+)\displaystyle V_{\mathbbmss{b}}(\mathpzc{b}_{2k}+) =ϕ(𝒷2𝓀;𝕓2𝓀+1)=(𝒷2𝓀;𝕓2𝓀1)=𝒱𝕓(𝒷2𝓀),\displaystyle=\phi(\mathpzc{b}_{2k};\mathbbmss{b}_{2k+1})=H(\mathpzc{b}_{2k};\mathbbmss{b}_{2k-1})=V_{\mathbbmss{b}}(\mathpzc{b}_{2k}-),
V𝕓(𝒷2𝓀+1+)\displaystyle V_{\mathbbmss{b}}(\mathpzc{b}_{2k+1}+) =H(𝒷2𝓀+1;𝕓2𝓀+1)=ϕ(𝒷2𝓀+1;𝕓2𝓀+1)=𝒱𝕓(𝒷2𝓀+1).\displaystyle=H(\mathpzc{b}_{2k+1};\mathbbmss{b}_{2k+1})=\phi(\mathpzc{b}_{2k+1};\mathbbmss{b}_{2k+1})=V_{\mathbbmss{b}}(\mathpzc{b}_{2k+1}-).

Additionally, by differentiating (4.2), we obtain for k=1,2,,nk=1,2,\dots,n:

V𝕓(𝒷1)\displaystyle V_{\mathbbmss{b}}^{\prime}(\mathpzc{b}_{1}-) =limx𝒷1g(𝒷1)𝒲(𝓆)(𝓍)𝒲(𝓆)(𝒷1)=(𝒷1)=𝒱𝕓(𝒷1+),\displaystyle=\lim_{x\uparrow\mathpzc{b}_{1}}g(\mathpzc{b}_{1})\frac{W^{(q)\prime}(x)}{W^{(q)\prime}(\mathpzc{b}_{1})}=g(\mathpzc{b}_{1})=V_{\mathbbmss{b}}^{\prime}(\mathpzc{b}_{1}+),
V𝕓(𝒷2𝓀+1)\displaystyle V_{\mathbbmss{b}}^{\prime}(\mathpzc{b}_{2k+1}-) =limx𝒷2𝓀+1[qH(𝒷2𝓀;𝕓2𝓀1)𝒲(𝓆)(𝓍𝒷2𝓀)\displaystyle=\lim_{x\uparrow\mathpzc{b}_{2k+1}}\bigg{[}qH(\mathpzc{b}_{2k};\mathbbmss{b}_{2k-1})W^{(q)}(x-\mathpzc{b}_{2k})
+W(q)(x𝒷2𝓀)((𝒷2𝓀+1)𝓆(𝒷2𝓀;𝕓2𝓀1)𝒲(𝓆)(𝒷2𝓀+1𝒷2𝓀)𝒲(𝓆)(𝒷2𝓀+1𝒷2𝓀))]\displaystyle\quad+W^{(q)\prime}(x-\mathpzc{b}_{2k})\bigg{(}\frac{g(\mathpzc{b}_{2k+1})-qH(\mathpzc{b}_{2k};\mathbbmss{b}_{2k-1})W^{(q)}(\mathpzc{b}_{2k+1}-\mathpzc{b}_{2k})}{W^{(q)\prime}(\mathpzc{b}_{2k+1}-\mathpzc{b}_{2k})}\bigg{)}\bigg{]}
=g(𝒷2𝓀+1)=𝒱𝕓(𝒷2𝓀+1+).\displaystyle=g(\mathpzc{b}_{2k+1})=V_{\mathbbmss{b}}^{\prime}(\mathpzc{b}_{2k+1}+).

Hence, V𝕓C((0,))C1((0,){𝒷2𝓀}𝓀=1𝓃)𝒞2((0,)𝕓))V_{\mathbbmss{b}}\in C((0,\infty))\cap C^{1}((0,\infty)\setminus\{\mathpzc{b}_{2k}\}_{k=1}^{n})\cap C^{2}((0,\infty)\setminus\mathbbmss{b})) for any 𝕓\mathbbmss{b}. The smooth fit principle requires finding barriers that increase the regularity of this function, which is done next.

4.2. Selection of an optimal multibarrier strategy

To select a 𝕓\mathbbmss{b}^{*} as a candidate for being an optimal barrier, let us define the auxiliary function

(4.5) F(v,z;𝕓2k+1):=g(z)qH(v;𝕓2k+1)W(q)(zv)W(q)(zv),for(v,z)𝒜2k+1,F(v,z;\mathbbmss{b}_{2k+1}):=\frac{g(z)-qH(v;\mathbbmss{b}_{2k+1})W^{(q)}(z-v)}{W^{(q)\prime}(z-v)},\quad\text{for}\ (v,z)\in\mathcal{A}_{2k+1},

where 𝕓2k+1={𝒷𝒾}𝒾=12𝓀+1\mathbbmss{b}_{2k+1}=\{\mathpzc{b}_{i}\}_{i=1}^{2k+1} is a non-negative increasing sequence and 𝒜2k+1:={(v,z)[𝒷2𝓀+1,)×[𝒷2𝓀+1,):𝓋<𝓏}\mathcal{A}_{2k+1}\raisebox{0.4pt}{$:$}=\{(v,z)\in[\mathpzc{b}_{2k+1},\infty)\times[\mathpzc{b}_{2k+1},\infty):v<z\}, with k{0,1,,n}k\in\{0,1,\dots,n\} and n0n\geq 0. Since W(q)C((0,))W^{(q)}\in C^{\infty}((0,\infty)), then F(,;𝕓2k+1)C(𝒜¯2k+1)F(\cdot,\cdot;\mathbbmss{b}_{2k+1})\in C(\bar{\mathcal{A}}_{2k+1}), where 𝒜¯2k+1\bar{\mathcal{A}}_{2k+1} is the closure set of 𝒜2k+1\mathcal{A}_{2k+1}, and it is as smooth as gg on 𝒜2k+1\mathcal{A}_{2k+1}.

The following are two very useful properties of the function F(,,𝕓2k+1)F(\cdot,\cdot,\mathbbmss{b}_{2k+1}) defined in (4.5):

  1. (1)

    By (2.5) together with (2.6), it is easy to verify that for k{0,1,,n}k\in\{0,1,\dots,n\} and v𝒷2𝓀+1v\geq\mathpzc{b}_{2k+1}

    (4.6) F(v,v+;𝕓2k+1):=limzvF(v,z;𝕓2k+1)=σ22g(v).\displaystyle F(v,v+;\mathbbmss{b}_{2k+1}):=\lim_{z\downarrow v}F(v,z;\mathbbmss{b}_{2k+1})=\frac{\sigma^{2}}{2}g(v).
  2. (2)

    Now, note that for z>vz>v

    (4.7) zF(v,z;𝕓2k+1)=F¯(v,z;𝕓2k+1)W(q)(zv),\displaystyle\partial_{z}F(v,z;\mathbbmss{b}_{2k+1})=\dfrac{\overline{F}(v,z;\mathbbmss{b}_{2k+1})}{W^{(q)\prime}(z-v)},

    where

    (4.8) F¯(v,z;𝕓2k+1):=g(z)qH(v;𝕓2k+1)W(q)(zv)W(q)′′(zv)F(v,z;𝕓2k+1).\overline{F}(v,z;\mathbbmss{b}_{2k+1})\raisebox{0.4pt}{$:$}=g^{\prime}(z)-qH(v;\mathbbmss{b}_{2k+1})W^{(q)\prime}(z-v)-W^{(q)\prime\prime}(z-v)F(v,z;\mathbbmss{b}_{2k+1}).

    From (3.3), (3.9) and (4.4), notice that (q)H(v;𝕓2k+1)(\mathcal{L}-q)H(v;\mathbbmss{b}_{2k+1}) can be rewritten as

    (4.9) (q)H(v;𝕓2k+1)=σ22g(v)+μg(v)qH(v;𝕓2k+1).(\mathcal{L}-q)H(v;\mathbbmss{b}_{2k+1})=\dfrac{\sigma^{2}}{2}g^{\prime}(v)+\mu g(v)-qH(v;\mathbbmss{b}_{2k+1}).

    Hence, letting zvz\downarrow v in (4.7) and using (2.6) and (2.7) together with (4.8) we obtain

    zF(v,v+;𝕓2k+1)\displaystyle\partial_{z}F(v,v+;\mathbbmss{b}_{2k+1}) =σ22F¯(v,v+;𝕓2k+1)\displaystyle=\frac{\sigma^{2}}{2}\overline{F}(v,v+;\mathbbmss{b}_{2k+1})
    =σ22(g(v)q2σ2H(v;𝕓2k+1)+4μσ4F(v,v+;𝕓2k+1))\displaystyle=\frac{\sigma^{2}}{2}\left(g^{\prime}(v)-q\frac{2}{\sigma^{2}}H(v;\mathbbmss{b}_{2k+1})+\frac{4\mu}{\sigma^{4}}F(v,v+;\mathbbmss{b}_{2k+1})\right)
    (4.10) =(q)H(v;𝕓2k+1),\displaystyle=(\mathcal{L}-q)H(v;\mathbbmss{b}_{2k+1}),

    where in the last equality we used (4.9) and (4.6).

Next, we present a method to select a candidate 𝕓\mathbbmss{b}^{*} for being an optimal multibarrier. Let us first start with the case n=1n=1.

Existence of 𝕓3\mathbbmss{b}^{*}_{3}

First, define 𝒷1=𝒷<\mathpzc{b}^{*}_{1}=\mathpzc{b}^{*}<\infty as in (3.5). Now, note that

(4.11) (q)H(v1;𝒷1)>0for some𝓋1>𝒷1,(\mathcal{L}-q)H(v_{1};\mathpzc{b}^{*}_{1})>0\quad\text{for some}\ v_{1}>\mathpzc{b}^{*}_{1},

otherwise, by Remark 3.1, 𝕓={𝒷1}\mathbbmss{b}^{*}=\{\mathpzc{b}^{*}_{1}\} is an optimal barrier. By definition of 𝒷1\mathpzc{b}^{*}_{1}, there exists 𝒷1(1)>𝒷1\mathpzc{b}^{(1)}_{1}>\mathpzc{b}^{*}_{1} such that F(v)<0F^{\prime}(v)<0 for v(𝒷1,𝒷1(1))v\in(\mathpzc{b}^{*}_{1},\mathpzc{b}^{(1)}_{1}), that is

(4.12) g(v)<W(q)′′(v)g(v)W(q)(v),g^{\prime}(v)<W^{(q)\prime\prime}(v)\dfrac{g(v)}{W^{(q)\prime}(v)},

and

(4.13) F(v)>F(z)forv(𝒷1,𝒷1(1))and𝓏>𝓋.F(v)>F(z)\quad\text{for}\ v\in(\mathpzc{b}^{*}_{1},\mathpzc{b}^{(1)}_{1})\ \text{and}\ z>v.
Proposition 4.2.
  1. (i)

    If F(𝒷1)=0F^{\prime}(\mathpzc{b}^{*}_{1})=0, then

    (4.14) (q)H(𝒷1;𝒷1)=0.(\mathcal{L}-q)H(\mathpzc{b}^{*}_{1};\mathpzc{b}^{*}_{1})=0.
  2. (ii)

    For each v(𝒷1,𝒷1(1))v\in(\mathpzc{b}^{*}_{1},\mathpzc{b}^{(1)}_{1}),

    (4.15) (q)H(v;𝒷1)<0.(\mathcal{L}-q)H(v;\mathpzc{b}^{*}_{1})<0.

The previous result and (4.11) imply that there exists v~>𝒷1(1)\tilde{v}>\mathpzc{b}^{(1)}_{1} such that

(4.16) zF(v,v+;𝒷1)=(𝓆)(𝓋;𝒷1){0,if𝓋(𝓋~ε1,𝓋~),=0,if𝓋=𝓋~,>0,if𝓋(𝓋~,𝓋~+ε1),\partial_{z}F({v},v+;\mathpzc{b}^{*}_{1})=(\mathcal{L}-q)H(v;\mathpzc{b}^{*}_{1})\begin{cases}\leq 0,&\text{if}\ v\in(\tilde{v}-\varepsilon_{1},\tilde{v}),\\ =0,&\text{if}\ v=\tilde{v},\\ >0,&\text{if}\ v\in(\tilde{v},\tilde{v}+\varepsilon_{1}),\end{cases}

for some ε1>0\varepsilon_{1}>0. Hence, the set

(4.17) 𝒫1:={v~>𝒷1(1):(4.16)is truefor someε1>0}\mathcal{P}_{1}\raisebox{0.4pt}{$:$}=\{\tilde{v}>\mathpzc{b}^{(1)}_{1}:\eqref{ineq1.0.0}\ \text{is true}\ \text{for some}\ \varepsilon_{1}>0\}

is not empty. Let 𝒸1:=min𝒫1\mathpzc{c}_{1}\raisebox{0.4pt}{$:$}=\min\mathcal{P}_{1}, which is attained due to the smoothness of gg.

Lemma 4.1.

There exists ε^1>0\hat{\varepsilon}_{1}>0, such that

(4.18) zF(𝒸1,𝓏;𝒷1)>0,for𝓏(𝒸1,𝒸1+ε^1).\partial_{z}F(\mathpzc{c}_{1},z;\mathpzc{b}^{*}_{1})>0,\quad\text{for}\ z\in(\mathpzc{c}_{1},\mathpzc{c}_{1}+\hat{\varepsilon}_{1}).

The lemma implies that the global maximum of zF(𝒸1,𝓏;𝒷1)z\mapsto F(\mathpzc{c}_{1},z;\mathpzc{b}^{*}_{1}), if it exists, is attained at some point z¯>𝒸1\bar{z}>\mathpzc{c}_{1}. Note that under Assumption 3.1 the global maximum of the functions zF(v,z;𝒷1)z\mapsto F(v,z;\mathpzc{b}^{*}_{1}) exists for all v>𝒷1v>\mathpzc{b}^{*}_{1}. Indeed, this assumption implies that for all v𝒷1v\geq\mathpzc{b}^{*}_{1} we have that limzF(v,z;𝒷1)=\lim\limits_{z\rightarrow\infty}F(v,z;\mathpzc{b}^{*}_{1})=-\infty and therefore the global maximum is attained. Hence, we have that

(4.19) z1(v):=sup{y>v:F(v,y;𝒷1)(𝓋,𝓏;𝒷1), for all z>v},for𝓋[𝒷1,𝒸1],z_{1}(v)\raisebox{0.4pt}{$:$}=\sup\{y>v:F(v,y;\mathpzc{b}^{*}_{1})\geq F(v,z;\mathpzc{b}^{*}_{1}),\ \text{ for all $z>v$}\},\quad\text{for}\ v\in[\mathpzc{b}^{*}_{1},\mathpzc{c}_{1}],

is well defined. Take the set 𝒟1\mathcal{D}_{1} as follows

𝒟1:={v[𝒷1,𝒸1]:𝓋<𝓏1(𝓋)}.\mathcal{D}_{1}\raisebox{0.4pt}{$:$}=\{v\in[\mathpzc{b}_{1}^{*},\mathpzc{c}_{1}]:v<z_{1}(v)\}.

We see that 𝒟1\mathcal{D}_{1} is the collection of points v[𝒷1,𝒸1]v\in[\mathpzc{b}^{*}_{1},\mathpzc{c}_{1}] where the map zF(v,z;𝒷1)z\mapsto F(v,z;\mathpzc{b}_{1}) attains its global maximum at z=z1(v)>vz=z_{1}(v)>v. We are now ready to define

𝒷2:=inf𝒟1and𝒷3:=𝓏1(𝒷2).\mathpzc{b}_{2}^{*}:=\inf\mathcal{D}_{1}\quad\text{and}\quad\mathpzc{b}_{3}^{*}:=z_{1}(\mathpzc{b}_{2}^{*}).
Proposition 4.3.

For each v[𝒷1,𝒷1(1))v\in[\mathpzc{b}^{*}_{1},\mathpzc{b}^{(1)}_{1}),

(4.20) F(v,v;𝒷1)>(𝓋,𝓏;𝒷1)for𝓏>𝓋.F(v,v;\mathpzc{b}^{*}_{1})>F(v,z;\mathpzc{b}^{*}_{1})\quad\text{for}\ z>v.

Hence [𝒷1,𝒷1(1))𝒟1𝒸[\mathpzc{b}^{*}_{1},\mathpzc{b}^{(1)}_{1})\subset\mathcal{D}_{1}^{c}, with 𝒟1c={v[𝒷1,𝒸1]:𝓋=𝓏1(𝓋)}\mathcal{D}_{1}^{c}=\{v\in[\mathpzc{b}^{*}_{1},\mathpzc{c}_{1}]:v=z_{1}(v)\}.

Proposition 4.3 shows that 𝒷2>𝒷1(1)>𝒷1\mathpzc{b}_{2}^{*}>\mathpzc{b}_{1}^{(1)}>\mathpzc{b}_{1}^{*}. Also note that if 𝒷2𝒟1\mathpzc{b}^{*}_{2}\in\mathcal{D}_{1} then the trajectory vz1(v)v\mapsto z_{1}(v), given by (4.19), has a discontinuity at v=𝒷2v=\mathpzc{b}^{*}_{2}. Before proving one of the main theorems of this section, we will illustrate the previous results with a numerical example. Let us consider gg as in (3.15) and the surplus process XX given by

Xt=x+2.4t+2Bt,fort0,X_{t}=x+2.4t+2B_{t},\quad\text{for}\ t\geq 0,

with the discount rate q=0.2q=0.2. We already now that 𝒷1=0.9165\mathpzc{b}^{*}_{1}=0.9165, see Figure 2. To find the next barriers we need to calculate 𝒸1=1.5113\mathpzc{c}_{1}=1.5113 and analyze F(,;𝒷1)F(\cdot,\cdot;\mathpzc{b}^{*}_{1}). Figure 3 shows the value of z1(v)z_{1}(v) for v𝒷1v\geq\mathpzc{b}_{1}^{*} and observe that z1(v)z_{1}(v) has a discontinuity at 𝒷2=1.1496\mathpzc{b}^{*}_{2}=1.1496, the minimum vv such that z1(v)>vz_{1}(v)>v. Also, 𝒷3=𝓏1(𝒷2)=2.1925\mathpzc{b}_{3}^{*}=z_{1}(\mathpzc{b}_{2}^{*})=2.1925.

Refer to caption
Refer to caption
Figure 3. On the left is the surface of F(,;𝒷1)F(\cdot,\cdot;\mathpzc{b}^{*}_{1}). On the right, the red dotted line is the value of z1(v)z_{1}(v) for v𝒷1v\geq\mathpzc{b}^{*}_{1}.
Theorem 4.1.

Assume that there exists ϵ1>0\epsilon_{1}>0 such that

(4.21) zF(v,v+;𝒷1)=(𝓆)(𝓋;𝒷1)<0,for 𝓋(𝒷2,𝒷2+ϵ1).\partial_{z}F(v,v+;\mathpzc{b}^{*}_{1})=(\mathcal{L}-q)H(v;\mathpzc{b}^{*}_{1})<0,\quad\text{for }v\in(\mathpzc{b}^{*}_{2},\mathpzc{b}^{*}_{2}+\epsilon_{1}).

Then

𝒷1<𝒷2<𝒷3,𝒷2<𝒸1\mathpzc{b}^{*}_{1}<\mathpzc{b}^{*}_{2}<\mathpzc{b}^{*}_{3},\quad\mathpzc{b}^{*}_{2}<\mathpzc{c}_{1}

and

(4.22) F(𝒷2,𝒷2;𝒷1)=(𝒷2,𝒷3;𝒷1).F(\mathpzc{b}^{*}_{2},\mathpzc{b}^{*}_{2};\mathpzc{b}^{*}_{1})=F(\mathpzc{b}^{*}_{2},\mathpzc{b}^{*}_{3};\mathpzc{b}^{*}_{1}).
Remark 4.1.

Note that condition (4.21) is satisfied if the function v(q)H(v;𝒷1)v\mapsto(\mathcal{L}-q)H(v;\mathpzc{b}^{*}_{1}) does not remain constant in any open interval, and this is equivalent, by differentiating (4.9), to the condition (q)g0(\mathcal{L}-q)g\neq 0 a.e.

Continuing with the numerical example, Figure 4 shows in blue the function F(𝒷2,;𝒷1)F(\mathpzc{b}^{*}_{2},\cdot;\mathpzc{b}^{*}_{1}), which satisfies F(𝒷2,𝒷2;𝒷1)=(𝒷2,𝒷3;𝒷1)=1.3401F(\mathpzc{b}^{*}_{2},\mathpzc{b}^{*}_{2};\mathpzc{b}^{*}_{1})=F(\mathpzc{b}^{*}_{2},\mathpzc{b}^{*}_{3};\mathpzc{b}^{*}_{1})=1.3401, and in red the function (q)H(;𝕓3)(\mathcal{L}-q)H(\cdot;\mathbbmss{b}^{*}_{3}), which is negative for x>𝒷3x>\mathpzc{b}_{3}^{*}. Figure 5 shows the value function V𝕓3V_{\mathbbmss{b}^{*}_{3}} and the multibarrier 𝕓3={𝒷1,𝒷2,𝒷3}\mathbbmss{b}^{*}_{3}=\{\mathpzc{b}^{*}_{1},\mathpzc{b}^{*}_{2},\mathpzc{b}^{*}_{3}\}.

Refer to caption
Figure 4. Plots of F(𝒷2,;𝒷1)F(\mathpzc{b}^{*}_{2},\cdot;\mathpzc{b}^{*}_{1}) (in blue) and (q)H(;𝕓3)(\mathcal{L}-q)H(\cdot;\mathbbmss{b}^{*}_{3}) (in red). Black stars show the values of 𝒷2\mathpzc{b}_{2}^{*} and 𝒷3\mathpzc{b}_{3}^{*}.
Refer to caption
Figure 5. Plot of V𝕓3V_{\mathbbmss{b}^{*}_{3}} where the stars show the optimal barriers.

Existence of 𝕓\mathbbmss{b}^{*}

Now that we have shown the existence of the first three barriers, we can formulate an algorithm that will provide the next two barriers 𝒷𝓀\mathpzc{b}^{*}_{k^{\prime}} with k{2k,2k+1}k^{\prime}\in\{2k,2k+1\} considering that the previous 𝕓2k1={𝒷𝒾}𝒾=12𝓀1\mathbbmss{b}^{*}_{2k-1}=\{\mathpzc{b}^{*}_{i}\}_{i=1}^{2k-1}, with k{2,,n}k\in\{2,\dots,n\}, are obtained. After that, we will prove the correctness of the algorithm under conditions similar to (4.21).

Algorithm 1.
  1. (1)

    Choose 𝒷1\mathpzc{b}_{1}^{*} as in (3.5).

  2. (2)

    If

    (4.23) (q)H(x;𝒷1)0for𝓍𝒷1,(\mathcal{L}-q)H(x;\mathpzc{b}^{*}_{1})\leq 0\quad\text{for}\ x\geq\mathpzc{b}^{*}_{1},

    let n=0n=0 and stop. Otherwise, let k=1k=1.

  3. (3)

    Let 𝒫2k1\mathcal{P}_{2k-1} be the set of points v~>𝒷2𝓀1\tilde{v}>\mathpzc{b}^{*}_{2k-1} such that for some ε2k1>0\varepsilon_{2k-1}>0,

    (4.24) zF(v,v+;𝕓2k1)=(q)H(v;𝕓2k1){0,ifv(v~ε2k1,v~),=0,ifv=v~,>0,ifv(v~,v~+ε2k1).\partial_{z}F({v},v+;\mathbbmss{b}^{*}_{2k-1})=(\mathcal{L}-q)H(v;\mathbbmss{b}^{*}_{2k-1})\begin{cases}{\color[rgb]{0.00,0.50,0.50}\leq}0,&\text{if}\ v\in(\tilde{v}-\varepsilon_{2k-1},\tilde{v}),\\ =0,&\text{if}\ v=\tilde{v},\\ >0,&\text{if}\ v\in(\tilde{v},\tilde{v}+\varepsilon_{2k-1}).\end{cases}

    Define 𝒸2𝓀1=min𝒫2𝓀1\mathpzc{c}_{2k-1}=\min\mathcal{P}_{2k-1} and

    (4.25) 𝒟2k1:={v[𝒷2𝓀1,𝒸2𝓀1]:𝓋<𝓏2𝓀1(𝓋)},\mathcal{D}_{2k-1}\raisebox{0.4pt}{$:$}=\{v\in[\mathpzc{b}_{2k-1}^{*},\mathpzc{c}_{2k-1}]:v<z_{2k-1}(v)\},

    where

    (4.26) z2k1(v):=sup{y>v:F(v,y;𝕓2k1)F(v,z;𝕓2k1), for all z>v}.z_{2k-1}(v)\raisebox{0.4pt}{$:$}=\sup\{y>v:F(v,y;\mathbbmss{b}^{*}_{2k-1})\geq F(v,z;\mathbbmss{b}^{*}_{2k-1}),\ \text{ for all $z>v$}\}.

    Let

    (4.27) 𝒷2𝓀:=inf𝒟2𝓀1and𝒷2𝓀+1:=𝓏2𝓀1(𝒷2𝓀).\mathpzc{b}_{2k}^{*}:=\inf\mathcal{D}_{2k-1}\quad\text{and}\quad\mathpzc{b}_{2k+1}^{*}:=z_{2k-1}(\mathpzc{b}_{2k}^{*}).
  4. (4)

    If

    (4.28) (q)H(x;𝕓2k+1)0forx𝒷2𝓀+1,(\mathcal{L}-q)H(x;\mathbbmss{b}^{*}_{2k+1})\leq 0\quad\text{for}\ x\geq\mathpzc{b}^{*}_{2k+1},

    choose the barriers 𝕓2k+1\mathbbmss{b}^{*}_{2k+1}, n=kn=k and stop. Otherwise, let k=k+1k=k+1 and return to (3).

To verify the correctness of the algorithm, we assume that, for k{2,,n}k\in\{2,\ldots,n\} fixed, the set of barriers 𝕓2k1={𝒷𝒾}𝒾=12𝓀1\mathbbmss{b}^{*}_{2k-1}=\{\mathpzc{b}^{*}_{i}\}_{i=1}^{2k-1} provided by the algorithm, are well defined, which means that

(4.29) 𝒷1<<𝒷2𝓀3<𝒷2𝓀2<𝒷2𝓀1,F(𝒷2𝓀2,𝒷2𝓀2;𝕓2𝓀3)=(𝒷2𝓀2,𝒷2𝓀1;𝕓2𝓀3).\displaystyle\begin{split}&\mathpzc{b}^{*}_{1}<\ldots<\mathpzc{b}_{2k-3}^{*}<\mathpzc{b}_{2k-2}^{*}<\mathpzc{b}^{*}_{2k-1},\\ &F(\mathpzc{b}^{*}_{2k-2},\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})=F(\mathpzc{b}^{*}_{2k-2},\mathpzc{b}^{*}_{2k-1};\mathbbmss{b}^{*}_{2k-3}).\end{split}

We will show that 𝒫2k1\mathcal{P}_{2k-1} and 𝒟2k1\mathcal{D}_{2k-1} are not empty sets, and that 𝒸2𝓀1\mathpzc{c}_{2k-1}, 𝒷2𝓀\mathpzc{b}^{*}_{2k} and 𝒷2𝓀+1\mathpzc{b}^{*}_{2k+1} are well defined. For that purpose, let us also suppose that

(4.30) (q)H(v2k1;𝕓2k1)>0for somev2k1>𝒷2𝓀1,(\mathcal{L}-q)H(v_{2k-1};\mathbbmss{b}^{*}_{2k-1})>0\quad\text{for some}\ v_{2k-1}>\mathpzc{b}^{*}_{2k-1},

otherwise Algorithm 1 would had stop. Now, by definition of 𝒷2𝓀1\mathpzc{b}^{*}_{2k-1} in step (3) of the algorithm, there exists 𝒷2𝓀1(1)>𝒷2𝓀1\mathpzc{b}_{2k-1}^{(1)}>\mathpzc{b}^{*}_{2k-1} such that

(4.31) zF(𝒷2𝓀2,𝓋;𝕓2𝓀3)<0for𝓋(𝒷2𝓀1,𝒷2𝓀1(1)),\partial_{z}F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})<0\quad\text{for}\ v\in(\mathpzc{b}^{*}_{2k-1},\mathpzc{b}^{(1)}_{2k-1}),

with equality at v=𝒷2𝓀1v=\mathpzc{b}^{*}_{2k-1}. This condition is equivalent to F¯(𝒷2𝓀2,𝓋;𝕓2𝓀3)<0\overline{F}(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})<0 with F¯\overline{F} as in (4.8). Moreover, we can choose 𝒷2𝓀1(1)\mathpzc{b}^{(1)}_{2k-1} such that for each v(𝒷2𝓀1,𝒷2𝓀1(1))v\in(\mathpzc{b}^{*}_{2k-1},\mathpzc{b}^{(1)}_{2k-1})

(4.32) F(𝒷2𝓀2,𝓏;𝕓2𝓀3)<(𝒷2𝓀2,𝓋;𝕓2𝓀3)for𝓏>𝓋.F(\mathpzc{b}^{*}_{2k-2},z;\mathbbmss{b}^{*}_{2k-3})<F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})\quad\text{for}\ z>v.

The following result is analogous to Proposition 4.2.

Proposition 4.4.

Let 𝒷2𝓀1(1)\mathpzc{b}^{(1)}_{2k-1} be given as before. Then,

(q)H(v;𝕓2k1)<0forv(𝒷2𝓀1,𝒷2𝓀1(1)).(\mathcal{L}-q)H(v;\mathbbmss{b}^{*}_{2k-1})<0\quad\text{for}\ v\in(\mathpzc{b}^{*}_{2k-1},\mathpzc{b}^{(1)}_{2k-1}).

Hence, 𝒫2k1\mathcal{P}_{2k-1}\neq\emptyset and 𝒷2𝓀1(1)<𝒸2𝓀1<\mathpzc{b}^{(1)}_{2k-1}<\mathpzc{c}_{2k-1}<\infty. Additionally, (q)H(𝒷2𝓀1;𝕓2𝓀1)=0(\mathcal{L}-q)H(\mathpzc{b}^{*}_{2k-1};\mathbbmss{b}^{*}_{2k-1})=0.

A few straightforward changes in the proof of Lemma 4.1 show that there exists ε^2k1>0\hat{\varepsilon}_{2k-1}>0 small enough, such that for z(𝒸2𝓀1,𝒸2𝓀1+ε^2𝓀1)z\in(\mathpzc{c}_{2k-1},\mathpzc{c}_{2k-1}+\hat{\varepsilon}_{2k-1})

zF(𝒸2𝓀1,𝓏;𝕓2𝓀1)>0,\partial_{z}F(\mathpzc{c}_{2k-1},z;\mathbbmss{b}^{*}_{2k-1})>0,

and therefore 𝒟2k1\mathcal{D}_{2k-1}\neq\emptyset. We also have that 𝒷2𝓀1<𝒷2𝓀\mathpzc{b}^{*}_{2k-1}<\mathpzc{b}^{*}_{2k} from the next result.

Proposition 4.5.

For each v[𝒷2𝓀1,𝒷2𝓀1(1))v\in[\mathpzc{b}^{*}_{2k-1},\mathpzc{b}^{(1)}_{2k-1}),

(4.33) F(v,v;𝕓2k1)>F(v,z;𝕓2k1)forz>v.F(v,v;\mathbbmss{b}^{*}_{2k-1})>F(v,z;\mathbbmss{b}^{*}_{2k-1})\quad\text{for}\ z>v.

Hence (𝒷2𝓀1,𝒷2𝓀1(1))𝒟2𝓀1𝒸(\mathpzc{b}^{*}_{2k-1},\mathpzc{b}^{(1)}_{2k-1})\subset\mathcal{D}_{2k-1}^{c}, with 𝒟2k1c={v[𝒷2𝓀1,𝒸2𝓀1]:𝓋=𝓏2𝓀1(𝓋)}\mathcal{D}_{2k-1}^{c}=\{v\in[\mathpzc{b}^{*}_{2k-1},\mathpzc{c}_{2k-1}]:v=z_{2k-1}(v)\}.

The final theorem follows from the above results and the appropriate modifications in the proof of Theorem 4.1, so we omit its proof.

Theorem 4.2.

Assume that there exists ϵ2k1>0\epsilon_{2k-1}>0 such that

(4.34) zF(v,v+;𝕓2k1)=(q)H(v;𝕓2k1)<0,for v(𝒷2𝓀,𝒷2𝓀+ϵ2𝓀1).\partial_{z}F(v,v+;\mathbbmss{b}^{*}_{2k-1})=(\mathcal{L}-q)H(v;\mathbbmss{b}^{*}_{2k-1})<0,\quad\text{for }v\in(\mathpzc{b}^{*}_{2k},\mathpzc{b}^{*}_{2k}+\epsilon_{2k-1}).

Then, 𝒷2𝓀\mathpzc{b}^{*}_{2k} and 𝒷2𝓀+1\mathpzc{b}_{2k+1} given by the Algorithm are well defined and satisfy

𝒷2𝓀1<𝒷2𝓀<𝒷2𝓀+1and𝒷2𝓀<𝒸2𝓀1.\mathpzc{b}^{*}_{2k-1}<\mathpzc{b}^{*}_{2k}<\mathpzc{b}^{*}_{2k+1}\quad\text{and}\quad\mathpzc{b}^{*}_{2k}<\mathpzc{c}_{2k-1}.

Moreover,

(4.35) F(𝒷2𝓀,𝒷2𝓀;𝕓2𝓀1)=(𝒷2𝓀,𝒷2𝓀+1;𝕓2𝓀1).F(\mathpzc{b}^{*}_{2k},\mathpzc{b}^{*}_{2k};\mathbbmss{b}^{*}_{2k-1})=F(\mathpzc{b}^{*}_{2k},\mathpzc{b}^{*}_{2k+1};\mathbbmss{b}^{*}_{2k-1}).

Again, by Remark 4.1, condition (4.34) is satisfied if (q)g0(\mathcal{L}-q)g\neq 0 a.e. Note that if the function x(q)H(x;𝕓2k+1)x\mapsto(\mathcal{L}-q)H(x;\mathbbmss{b}^{*}_{2k+1}) is non-increasing for x𝒷2𝓀+1x\geq\mathpzc{b}^{*}_{2k+1}, or equivalently, if (q)g(x)0(\mathcal{L}-q)g(x)\leq 0 for x𝒷2𝓀+1x\geq\mathpzc{b}^{*}_{2k+1}, then condition (4.28) holds. Another sufficient condition is given in the next proposition, which is analogous to Theorem 3.1.

Proposition 4.6.

Assume that the function zF(𝒷2𝓀,𝓏;𝕓2𝓀1)z\mapsto F(\mathpzc{b}^{*}_{2k},z;\mathbbmss{b}^{*}_{2k-1}) is non-increasing for z𝒷2𝓀+1z\geq\mathpzc{b}^{*}_{2k+1}, then the function x(q)H(x;𝕓2k+1)0x\mapsto(\mathcal{L}-q)H(x;\mathbbmss{b}^{*}_{2k+1})\leq 0 for all x𝒷2𝓀+1x\geq\mathpzc{b}^{*}_{2k+1}.

4.3. Verification of optimality

We now verify that the sequence of barriers 𝕓\mathbbmss{b}^{*} given by Algorithm 1 is indeed optimal. For this, note that the set of barriers 𝕓\mathbbmss{b}^{*} satisfies the following properties:

  • (i)

    Using (4.6) together with (4.22) and (4.35) we have for k=1,,nk=1,\dots,n

    σ22g(𝒷2𝓀)\displaystyle\frac{\sigma^{2}}{2}g(\mathpzc{b}_{2k}^{*}) =F(𝒷2𝓀,𝒷2𝓀+;𝕓2𝓀1)=(𝒷2𝓀,𝒷2𝓀+1;𝕓2𝓀1)\displaystyle=F(\mathpzc{b}_{2k}^{*},\mathpzc{b}_{2k}^{*}+;\mathbbmss{b}^{*}_{2k-1})=F(\mathpzc{b}_{2k}^{*},\mathpzc{b}_{2k+1}^{*};\mathbbmss{b}^{*}_{2k-1})
    =g(𝒷2𝓀+1)𝓆(𝒷2𝓀;𝕓2𝓀1)𝒲(𝓆)(𝒷2𝓀+1𝒷2𝓀)W(q)(𝒷2𝓀+1𝒷2𝓀).\displaystyle=\frac{g(\mathpzc{b}_{2k+1}^{*})-qH(\mathpzc{b}_{2k}^{*};\mathbbmss{b}_{2k-1})W^{(q)}(\mathpzc{b}_{2k+1}^{*}-\mathpzc{b}_{2k}^{*})}{W^{(q)\prime}(\mathpzc{b}_{2k+1}^{*}-\mathpzc{b}_{2k}^{*})}.

    Therefore,

    V𝕓(𝒷2𝓀+)\displaystyle V_{\mathbbmss{b}^{*}}^{\prime}(\mathpzc{b}^{*}_{2k}+) =2σ2(g(𝒷2𝓀+1)𝓆(𝒷2𝓀;𝕓2𝓀1)𝒲(𝓆)(𝒷2𝓀+1𝒷2𝓀)W(q)(𝒷2𝓀+1𝒷2𝓀))=g(𝒷2𝓀)=𝒱𝕓(𝒷2𝓀),\displaystyle=\frac{2}{\sigma^{2}}\left(\frac{g(\mathpzc{b}_{2k+1}^{*})-qH(\mathpzc{b}^{*}_{2k};\mathbbmss{b}_{2k-1})W^{(q)}(\mathpzc{b}_{2k+1}^{*}-\mathpzc{b}^{*}_{2k})}{W^{(q)\prime}(\mathpzc{b}_{2k+1}^{*}-\mathpzc{b}^{*}_{2k})}\right)=g(\mathpzc{b}^{*}_{2k})=V_{\mathbbmss{b}^{*}}^{\prime}(\mathpzc{b}^{*}_{2k}-),

    and hence V𝕓C1((0,))V_{\mathbbmss{b}^{*}}\in C^{1}((0,\infty)).

  • (ii)

    Additionally, by the step (3)

    zF(𝒷2𝓀,𝓏;𝕓2𝓀1)|𝓏=𝒷2𝓀+1=0.\displaystyle\partial_{z}F(\mathpzc{b}^{*}_{2k},z;\mathbbmss{b}^{*}_{2k-1})\Big{|}_{z=\mathpzc{b}^{*}_{2k+1}}=0.

    This implies that

    g(𝒷2𝓀+1)=𝓆(𝒷2𝓀,𝕓2𝓀1)𝒲(𝓆)(𝒷2𝓀+1𝒷2𝓀)+𝒲(𝓆)′′(𝒷2𝓀+1𝒷2𝓀)(𝒷2𝓀,𝒷2𝓀+1;𝕓2𝓀1),\displaystyle g^{\prime}(\mathpzc{b}^{*}_{2k+1})=qH(\mathpzc{b}^{*}_{2k},\mathbbmss{b}^{*}_{2k-1})W^{(q)\prime}(\mathpzc{b}^{*}_{2k+1}-\mathpzc{b}^{*}_{2k})+W^{(q)\prime\prime}(\mathpzc{b}_{2k+1}-\mathpzc{b}^{*}_{2k})F(\mathpzc{b}^{*}_{2k},\mathpzc{b}^{*}_{2k+1};\mathbbmss{b}^{*}_{2k-1}),

    and therefore

    V𝕓′′(𝒷2𝓀+1+)\displaystyle V_{\mathbbmss{b}^{*}}^{\prime\prime}(\mathpzc{b}^{*}_{2k+1}+) =qH(𝒷2𝓀,𝕓2𝓀1)𝒲(𝓆)(𝒷2𝓀+1𝒷2𝓀)\displaystyle=qH(\mathpzc{b}^{*}_{2k},\mathbbmss{b}^{*}_{2k-1})W^{(q)\prime}(\mathpzc{b}^{*}_{2k+1}-\mathpzc{b}^{*}_{2k})
    +W(q)′′(𝒷2𝓀+1𝒷2𝓀)((𝒷2𝓀+1)𝓆(𝒷2𝓀,𝕓2𝓀1)𝒲(𝓆)(𝒷2𝓀+1𝒷2𝓀)𝒲(𝓆)(𝒷2𝓀+1𝒷2𝓀))\displaystyle\quad+W^{(q)\prime\prime}(\mathpzc{b}^{*}_{2k+1}-\mathpzc{b}^{*}_{2k})\bigg{(}\frac{g(\mathpzc{b}^{*}_{2k+1})-qH(\mathpzc{b}^{*}_{2k},\mathbbmss{b}^{*}_{2k-1})W^{(q)}(\mathpzc{b}^{*}_{2k+1}-\mathpzc{b}^{*}_{2k})}{W^{(q)\prime}(\mathpzc{b}^{*}_{2k+1}-\mathpzc{b}^{*}_{2k})}\bigg{)}
    =g(𝒷2𝓀+1)=𝒱𝕓′′(𝒷2𝓀+1).\displaystyle=g^{\prime}(\mathpzc{b}^{*}_{2k+1})=V_{\mathbbmss{b}^{*}}^{\prime\prime}(\mathpzc{b}^{*}_{2k+1}-).

    Hence, V𝕓C2(\{𝒷2𝓀}𝓀=1𝓃)V_{\mathbbmss{b}^{*}}\in C^{2}(\mathbbm{R}\backslash\{\mathpzc{b}_{2k}\}_{k=1}^{n}).

Also, the value function V𝕓V_{\mathbbmss{b}^{*}} satisfies that:

  1. (i)

    As in the proof of Theorem 3.1, V𝕓V_{\mathbbmss{b}^{*}} satisfies the HJB equation (3.10) on (0,𝒷1)(0,\mathpzc{b}^{*}_{1}).

  2. (ii)

    By (4.4), V𝕓(x)=g(x)V^{\prime}_{\mathbbmss{b}^{*}}(x)=g(x) for xk=2n[𝒷2𝓀1,𝒷2𝓀][𝒷2𝓃+1,)x\in\bigcup_{k=2}^{n}[\mathpzc{b}_{2k-1}^{*},\mathpzc{b}_{2k}^{*}]\cup[\mathpzc{b}^{*}_{2n+1},\infty).

  3. (iii)

    By the martingale properties of the scale functions (2.1) we also have that (q)V𝕓(x)=0(\mathcal{L}-q)V_{\mathbbmss{b}^{*}}(x)=0 for xk=1n(𝒷2𝓀,𝒷2𝓀+1)x\in\bigcup_{k=1}^{n}(\mathpzc{b}_{2k}^{*},\mathpzc{b}_{2k+1}^{*}).

  4. (iv)

    Since F(𝒷2𝓀,𝓍;𝕓2𝓀1)(𝒷2𝓀,𝒷2𝓀+1;𝕓2𝓀1)F(\mathpzc{b}_{2k}^{*},x;\mathbbmss{b}^{*}_{2k-1})\leq F(\mathpzc{b}_{2k}^{*},\mathpzc{b}_{2k+1}^{*};\mathbbmss{b}^{*}_{2k-1}) for x(𝒷2𝓀,𝒷2𝓀+1)x\in(\mathpzc{b}_{2k}^{*},\mathpzc{b}_{2k+1}^{*}) and k{1,,n}k\in\{1,\dots,n\}, then

    V𝕓(x)\displaystyle V^{\prime}_{\mathbbmss{b}^{*}}(x) =qH(𝒷2𝓀;𝕓2𝓀1)𝒲(𝓆)(𝓍𝒷2𝓀)+(𝒷2𝓀,𝒷2𝓀+1;𝕓2𝓀1)𝒲(𝓆)(𝓍𝒷2𝓀)\displaystyle=qH(\mathpzc{b}_{2k}^{*};\mathbbmss{b}_{2k-1})W^{(q)}(x-\mathpzc{b}_{2k}^{*})+F(\mathpzc{b}_{2k}^{*},\mathpzc{b}_{2k+1}^{*};\mathbbmss{b}_{2k-1})W^{(q)\prime}(x-\mathpzc{b}_{2k}^{*})
    qH(𝒷2𝓀;𝕓2𝓀1)𝒲(𝓆)(𝓍𝒷2𝓀)+(𝒷2𝓀,𝓍;𝕓2𝓀1)𝒲(𝓆)(𝓍𝒷2𝓀)\displaystyle\geq qH(\mathpzc{b}_{2k}^{*};\mathbbmss{b}_{2k-1})W^{(q)}(x-\mathpzc{b}_{2k}^{*})+F(\mathpzc{b}_{2k}^{*},x;\mathbbmss{b}_{2k-1})W^{(q)\prime}(x-\mathpzc{b}_{2k}^{*})
    =g(x).\displaystyle=g(x).
  5. (v)

    By construction, Algorithm 1 guarantees that for both x[𝒷2𝓀+1,𝒷2𝓀+2]x\in[\mathpzc{b}_{2k+1}^{*},\mathpzc{b}_{2k+2}^{*}], with k{0,,n1}k\in\{0,\dots,n-1\}, and x[𝒷2𝓃+1,)x\in[\mathpzc{b}_{2n+1}^{*},\infty), we have that

    (q)V𝕓(x)=(q)H(x;𝕓2k1)0.(\mathcal{L}-q)V_{\mathbbmss{b}^{*}}(x)=(\mathcal{L}-q)H(x;\mathbbmss{b}^{*}_{2k-1})\leq 0.

We conclude that V𝕓V_{\mathbbmss{b}}^{*} satisfies the HJB equation (3.10). Now, by Lemma 3.2, we obtain the following theorem.

Theorem 4.3.

Let VV be the value function given in (1.3). Then V𝕓(x)=V(x)V_{\mathbbmss{b}^{*}}(x)=V(x) for all x0x\geq 0 and the 𝕓\mathbbmss{b}^{*}-strategy is optimal.

Acknowledgments

The authors would like to thank the anonymous reviewers for their comments and suggestions, which helped to improve significantly the quality of this paper.

Funding

M. Junca was supported by the Research Fund of the Facultad de Ciencias, Universidad de los Andes INV-2021-128-2307. The authors have no relevant financial or non-financial interests to disclose.

References

  • [1] H. Albrecher and B. Garcia Flores. Optimal dividend bands revisited: a gradient-based method and evolutionary algorithms. Scandinavian Actuarial Journal, 2023(8):788–810, 2023.
  • [2] L. H. R. Alvarez. Optimal harvesting under stochastic fluctuations and critical depensation. Math. Biosci., 152(1):63–85, 1998.
  • [3] L. H. R. Alvarez. Singular stochastic control in the presence of a state-dependent yield structure. Stochastic Process. Appl., 86(2):323–343, 2000.
  • [4] L. H. R. Alvarez and T. A. Rakkolainen. On singular stochastic control and optimal stopping of spectrally negative jump diffusions. Stochastics. An International Journal of Probability and Stochastic Processes, 81(1):55–78, 2009.
  • [5] F. Avram, A. Kyprianou, and M. Pistorious. Exit problems for spectrally negative Lévy processes and application to (canadized) russian options. Ann. Appl. Probab., 14(1):2115–238, 2004.
  • [6] F. Avram, Z. Palmowski, and M. Pistorius. On the optimal dividend problem for a spectrally negative Lévy process. Ann. Appl. Probab., 17(1):156–180, 2007.
  • [7] P. Azcue and N. Muler. Optimal reinsurance and dividend distribution policies in the cramÉr-lundberg model. Mathematical Finance, 15(2):261–308, 2005.
  • [8] L. Bai and J. Paulsen. Optimal dividend policies with transaction costs for a class of diffusion processes. SIAM Journal on Control and Optimization, 48(8):4987–5008, 2010.
  • [9] L. Bai and J. Paulsen. On non-trivial barrier solutions of the dividend problem for a diffusion under constant and proportional transaction costs. Stochastic Processes and their Applications, 122(12):4005–4027, 2012.
  • [10] Andrei N. Borodin and Paavo Salminen. Handbook of Brownian motion—facts and formulae. Probability and its Applications. Birkhäuser Verlag, Basel, second edition, 2002.
  • [11] B. De Finetti. Su un’impostazione alternativa della teoria collettiva del rischio. Transactions of the 15th International Congress of Actuaries, II:433–443, 1957.
  • [12] G. Ferrari. On a class of singular stochastic control problems for reflected diffusions. Journal of Mathematical Analysis and Applications, 473(2):952–979, 2019.
  • [13] G. Ferrari and P. Schuhmann. An optimal dividend problem with capital injections over a finite horizon. SIAM Journal on Control and Optimization, 57(4):2686–2719, 2019.
  • [14] A. Kuznetsov, A. E. Kyprianou, and V. Rivero. The theory of scale functions for spectrally negative Lévy processes. Levy Matters II, Vol. 2061:97–186, 2013.
  • [15] A. Kyprianou, R. Loeffen, and J. L. Pérez. Optimal control with absolutely continuous strategies for spectrally negative lévy processes. Journal of Applied Probability, 49(1):150–166, 2012.
  • [16] A. E. Kyprianou. Fluctuations of Lévy Processes with Applications. Universitext. Springer-Verlag Berlin Heidelberg, 2014.
  • [17] R. Loeffen. On the optimality of the barrier strategy in de Finetti’s problem for spectrally negative Lévy processes. Ann. Appl. Probab., 18(5):1669–1680, 2008.
  • [18] R. Loeffen. An optimal dividends problem with a terminal value for spectrally negative Lévy processes with a completely monotone jump density. Journal of Applied Probability, 46(1):85–98, 2009.
  • [19] E.M. Lungu and B Øksendal. Optimal harvesting from a population in a stochastic crowded environment. Mathematical Biosciences, 145:47–75, 1997.
  • [20] M. R. Pistorius. On exit and ergodicity of the spectrally one-sided Lévy process reflected at its infimum. Journal of Theoretical Probability, 17(1):183–220, 2004.
  • [21] P. Protter. Stochastic integration and differential equations. Springer, Berlin, 2nd edition, 2005.
  • [22] H. Schmidli. Stochastic Control in Insurance. Springer-Verlang, London, 2008.
  • [23] H. Schmidli. On capital injections and dividends with tax in a diffusion approximation. Scandinavian Actuarial Journal, 2017(9):751–760, 2017.
  • [24] Q. Song, R. H. Stockbridge, and C. Zhu. On optimal harvesting problems in random environments. SIAM Journal on Control and Optimization, 49(2):859–889, 2011.

Appendix A Properties of scale functions of Brownian motion with drift

Considering Xt=μt+σBtX_{t}=\mu t+\sigma B_{t}, with μ\mu\in\mathbbm{R}, σ0\sigma\geq 0 and BtB_{t} as a Brownian motion, the qq-scale functions W(q)W^{(q)} and Z(q)Z^{(q)} are given by the expressions seen in (2.5). Then, using (2.1) we see that W(q)W^{(q)} and Z(q)Z^{(q)} satisfy the following identities

(A.1) σ22W(q)′′(v)+μW(q)(v)=qW(q)(v)forv0,\dfrac{\sigma^{2}}{2}W^{(q)\prime\prime}(v)+\mu W^{(q)\prime}(v)=qW^{(q)}(v)\quad\text{for}\ v\geq 0,
(A.2) σ22qW(q)(v)+μqW(q)(v)=qZ(q)(v)forv0.\dfrac{\sigma^{2}}{2}qW^{(q)\prime}(v)+\mu qW^{(q)}(v)=qZ^{(q)}(v)\quad\text{for}\ v\geq 0.
Lemma A.1.

For 0b<v<z0\leq b<v<z, the following hold

(A.3) [W(q)(v)]2W(q)(v)W(q)′′(v)=4σ4e(Φ(q)ξ1)v,\displaystyle[W^{(q)\prime}(v)]^{2}-W^{(q)}(v)W^{(q)\prime\prime}(v)=\dfrac{4}{\sigma^{4}}\operatorname{e}^{(\Phi(q)-\xi_{1})v},
(A.4) W(q)′′(zb)W(q)(zv)W(q)(zb)W(q)′′(zv)=4qσ4e(Φ(q)ξ1)(zv)W(q)(vb),\displaystyle W^{(q)\prime\prime}(z-b)W^{(q)\prime}(z-v)-W^{(q)\prime}(z-b)W^{(q)\prime\prime}(z-v)=\dfrac{4q}{\sigma^{4}}\operatorname{e}^{(\Phi(q)-\xi_{1})(z-v)}W^{(q)}(v-b),
(A.5) W(q)(zb)W(q)(zv)W(q)′′(zv)W(q)(zb)=4e(Φ(q)ξ1)(zv)σ4Z(q)(vb).\displaystyle W^{(q)\prime}(z-b)W^{(q)\prime}(z-v)-W^{(q)\prime\prime}(z-v)W^{(q)}(z-b)=\frac{4\operatorname{e}^{(\Phi(q)-\xi_{1})(z-v)}}{\sigma^{4}}Z^{(q)}(v-b).

The identities (A.3)–(A.5) will be used to verify the results given in Propositions 4.3 and 4.5.

Appendix B Proofs of one-barrier strategies

Proof of Proposition 3.1.

Observe that if X0=x>𝒷X_{0}=x>\mathpzc{b}, L0𝒷=x𝒷L^{\mathpzc{b}}_{0}=x-\mathpzc{b}. Then, by the strong Markov property, we get that

(B.1) V𝒷(x)\displaystyle V_{\mathpzc{b}}(x) =(x𝒷)01(𝓍λ(𝓍𝒷))𝒹λ+𝔼𝒷[0τ𝒷e𝓆𝓉(𝒳𝓉𝒷)d𝓉𝒷]\displaystyle=(x-\mathpzc{b})\int_{0}^{1}g(x-\lambda(x-\mathpzc{b}))d\lambda+\mathbbm{E}_{\mathpzc{b}}\left[\int_{0}^{\tau_{\mathpzc{b}}}\operatorname{e}^{-qt}g(X^{\mathpzc{b}}_{t})\circ{\rm d}L^{\mathpzc{b}}_{t}\right]
=(x𝒷)01(𝓍λ(𝓍𝒷))𝒹λ+𝒱𝒷(𝒷)\displaystyle=(x-\mathpzc{b})\int_{0}^{1}g(x-\lambda(x-\mathpzc{b}))d\lambda+V_{\mathpzc{b}}(\mathpzc{b})
=𝒷xg(u)𝑑u+V𝒷(𝒷)=𝒢(𝓍)𝒢(𝒷)+𝒱𝒷(𝒷).\displaystyle=\int_{\mathpzc{b}}^{x}g(u)du+V_{\mathpzc{b}}(\mathpzc{b})=G(x)-G(\mathpzc{b})+V_{\mathpzc{b}}(\mathpzc{b}).

Let us consider x[0,𝒷]x\in[0,\mathpzc{b}], then by (2.3) and the strong Markov property,

(B.2) V𝒷(x)=𝔼x[eqτ𝒷+𝔼𝒷[0τ𝒷eqtg(Xt𝒷)dLt𝒷];τ𝒷+<τ0]=V𝒷(𝒷)𝒲(𝓆)(𝓍)𝒲(𝓆)(𝒷).\displaystyle V_{\mathpzc{b}}(x)=\mathbbm{E}_{x}\left[\operatorname{e}^{-q\tau_{\mathpzc{b}}^{+}}\mathbbm{E}_{\mathpzc{b}}\left[\int_{0}^{\tau_{\mathpzc{b}}}\operatorname{e}^{-qt}g(X^{\mathpzc{b}}_{t})\circ{\rm d}L^{\mathpzc{b}}_{t}\right];\tau_{\mathpzc{b}}^{+}<\tau_{0}^{-}\right]=V_{\mathpzc{b}}(\mathpzc{b})\frac{W^{(q)}(x)}{W^{(q)}(\mathpzc{b})}.

We now compute V𝒷(𝒷)V_{\mathpzc{b}}(\mathpzc{b}). Let us denote St=sup0st(X~s0)S_{t}=\displaystyle\sup_{0\leq s\leq t}(\widetilde{X}_{s}\vee 0) and Yt=StX~tY_{t}=S_{t}-\widetilde{X}_{t} for t0t\geq 0, where the process X~\widetilde{X} has the same law than X𝒷X-\mathpzc{b}. We also consider the first entrance of the process YY to the set (a,)(a,\infty) given by σa:=inf{t0:Yt>a}\sigma_{a}\raisebox{0.4pt}{$:$}=\inf\{t\geq 0:Y_{t}>a\}, aa\in\mathbbm{R}. By spatial homogeneity of the Lévy process XX and following Proposition 1 in [6] we have that the ensemble {Xt𝒷,Lt𝒷,tτ𝒷;X0𝒷=x}\{X^{\mathpzc{b}}_{t},L_{t}^{\mathpzc{b}},t\leq\tau_{\mathpzc{b}};X^{\mathpzc{b}}_{0}=x\} has the same law as {𝒷𝒴𝓉,𝒮𝓉,𝓉σ𝒷;𝒴0=𝒷𝓍}\{\mathpzc{b}-Y_{t},S_{t},t\leq\sigma_{\mathpzc{b}};Y_{0}=\mathpzc{b}-x\}. Therefore

(B.3) V𝒷(𝒷)\displaystyle V_{\mathpzc{b}}(\mathpzc{b}) =𝔼𝒷[0τ𝒷eqtg(Xt𝒷)dLt𝒷]=𝔼0[0σ𝒷eqtg(𝒷𝒴𝓉)d𝒮𝓉].\displaystyle=\mathbbm{E}_{\mathpzc{b}}\left[\int_{0}^{\tau_{\mathpzc{b}}}\operatorname{e}^{-qt}g(X^{\mathpzc{b}}_{t})\circ{\rm d}L^{\mathpzc{b}}_{t}\right]=\mathbbm{E}_{0}\left[\int_{0}^{\sigma_{\mathpzc{b}}}\operatorname{e}^{-qt}g(\mathpzc{b}-Y_{t})\circ{\rm d}S_{t}\right].

By the absence of positive jumps of XX, it is known that the supremum SS is a local time for 0 for the process YY. Now, let us denote by \mathcal{E} the set of excursions away from zero of finite length

:={ϵ𝒞:ζ=ζ(ϵ)>0such thatϵζ=0andϵx>0for 0<x<ζ},\mathcal{E}\raisebox{0.4pt}{$:$}=\{\epsilon\in\mathcal{C}:\exists\ \zeta=\zeta(\epsilon)>0\ \text{such that}\ \epsilon_{\zeta}=0\ \text{and}\ \epsilon_{x}>0\ \text{for}\ 0<x<\zeta\},

where 𝒞:=𝒞([0,))\mathcal{C}\raisebox{0.4pt}{$:$}=\mathcal{C}([0,\infty)) is the set of all càdlàg functions on [0,)[0,\infty). Additionally, we denote by ()\mathcal{E}^{(\infty)} the set of excursions with infinite length. We will work with the excursion process ϵ:={ϵt:t0}\epsilon\raisebox{0.4pt}{$:$}=\{\epsilon_{t}:t\geq 0\} of YY, with values on ()\mathcal{E}\cup\mathcal{E}^{(\infty)}, given by

ϵt={Ys:St1s<St1},if St1<St1,\epsilon_{t}=\{Y_{s}:S^{-1}_{t-}\leq s<S^{-1}_{t}\},\quad\text{if $S^{-1}_{t-}<S^{-1}_{t}$,}

where S1S^{-1} denotes the right-inverse of the local time of YY at 0. Hence, noting that the process SS is continuous and non-decreasing, and following the proof of Theorem 1 in [5] we have

𝔼0[0σ𝒷eqtg(𝒷𝒴𝓉)d𝒮𝓉]\displaystyle\mathbbm{E}_{0}\left[\int_{0}^{\sigma_{\mathpzc{b}}}\operatorname{e}^{-qt}g(\mathpzc{b}-Y_{t})\circ{\rm d}S_{t}\right] =𝔼0[0σ𝒷eqtg(𝒷𝒴𝓉)d𝒮𝓉]\displaystyle=\mathbbm{E}_{0}\left[\int_{0}^{\sigma_{\mathpzc{b}}}\operatorname{e}^{-qt}g(\mathpzc{b}-Y_{t}){\rm d}S_{t}\right]
=𝔼0[0eqSt1g(𝒷ϵ𝒮𝓉1)1{ϵ¯𝒮𝓉1𝒷,𝓉<𝒮}d𝓉],\displaystyle=\mathbbm{E}_{0}\left[\int_{0}^{\infty}\operatorname{e}^{-qS^{-1}_{t}}g(\mathpzc{b}-\epsilon_{S^{-1}_{t}})1_{\{\bar{\epsilon}_{S^{-1}_{t}}\leq\mathpzc{b},t<S_{\infty}\}}{\rm d}t\right],

where ϵ¯t=supstϵs\bar{\epsilon}_{t}=\sup_{s\leq t}\epsilon_{s} for t>0t>0 . By the lack of negative jumps of the process ϵ\epsilon, we have that ϵSt1=0\epsilon_{S^{-1}_{t}}=0, hence using (B.3)

V𝒷(𝒷)\displaystyle V_{\mathpzc{b}}(\mathpzc{b}) =𝔼0[0eqSt1g(𝒷ϵ𝒮𝓉1)1{ϵ¯𝒮𝓉1𝒷,𝓉<𝒮}d𝓉]\displaystyle=\mathbbm{E}_{0}\left[\int_{0}^{\infty}\operatorname{e}^{-qS^{-1}_{t}}g(\mathpzc{b}-\epsilon_{S^{-1}_{t}})1_{\{\bar{\epsilon}_{S^{-1}_{t}}\leq\mathpzc{b},t<S_{\infty}\}}{\rm d}t\right]
=g(𝒷)𝔼0[0e𝓆𝒮𝓉11{ϵ¯𝒮𝓉1𝒷,𝓉<𝒮}d𝓉]\displaystyle=g(\mathpzc{b})\mathbbm{E}_{0}\left[\int_{0}^{\infty}\operatorname{e}^{-qS^{-1}_{t}}1_{\{\bar{\epsilon}_{S^{-1}_{t}}\leq\mathpzc{b},t<S_{\infty}\}}{\rm d}t\right]
=g(𝒷)𝔼0[0τ𝒷e𝓆𝓉d𝒮𝓉]\displaystyle=g(\mathpzc{b})\mathbbm{E}_{0}\left[\int_{0}^{\tau_{\mathpzc{b}}}\operatorname{e}^{-qt}{\rm d}S_{t}\right]
(B.4) =g(𝒷)𝔼𝒷[0τ𝒷e𝓆𝓉d𝓉𝒷]=(𝒷)𝒲(𝓆)(𝒷)𝒲(𝓆)(𝒷),\displaystyle=g(\mathpzc{b})\mathbbm{E}_{\mathpzc{b}}\left[\int_{0}^{\tau_{\mathpzc{b}}}\operatorname{e}^{-qt}{\rm d}L^{\mathpzc{b}}_{t}\right]=g(\mathpzc{b})\frac{W^{(q)}(\mathpzc{b})}{W^{(q)\prime}(\mathpzc{b})},

where the last equality follows from the proof of Theorem 1 in [5]. Finally using (B) in (B.1) and (B.2) gives the result. ∎

Proof of Lemma 3.2.

By the definition of VV, it follows that Vπ^(x)V(x)V_{\hat{\pi}}(x)\leq V(x) for all x0x\geq 0. Now, let us write w:=Vπ^w\raisebox{0.4pt}{$:$}=V_{\hat{\pi}} and we will show that w(x)V(x)w(x)\geq V(x) for all x0x\geq 0. First, assume that x>0x>0 and define, for any π𝒮\pi\in\mathcal{S}, τ0π:=inf{t>0:Xtπ0}\tau_{0}^{\pi}\raisebox{0.4pt}{$:$}=\inf\{t>0:X_{t}^{\pi}\leq 0\} and denote by 𝒮0\mathcal{S}_{0} the following set of admissible strategies:

𝒮0:={π𝒮:Vπ(x)=𝔼x[0τ0πeqsg(Xsπ)dLsπ]for all x>0}.\mathcal{S}_{0}\raisebox{0.4pt}{$:$}=\left\{\pi\in\mathcal{S}:V_{\pi}(x)=\mathbbm{E}_{x}\left[\int_{0}^{\tau^{\pi}_{0}}\operatorname{e}^{-qs}g(X^{\pi}_{s-})\circ{\rm d}L^{\pi}_{s}\right]\ \text{for all $x>0$}\right\}.

Then, by the proof of Lemma 4 in [15] it is enough to show that w(x)Vπ(x)w(x)\geq V_{\pi}(x) for any π𝒮0\pi\in\mathcal{S}_{0}. Hence, fix π𝒮0\pi\in\mathcal{S}_{0} and let Tn:=inf{t>0:Xtπ>norXtπ<1/n}T_{n}\raisebox{0.4pt}{$:$}=\inf\{t>0:X_{t}^{\pi}>n\ \text{or}\ X_{t}^{\pi}<1/n\}. Noting that XπX^{\pi} is a semi-martingale and that ww is sufficiently smooth on (0,)(0,\infty), we can use the change of variable/Itô’s formula on eq(tTn)w(XtTnπ)\operatorname{e}^{-q(t\wedge T_{n})}w(X_{t\wedge T_{n}}^{\pi}), with t0t\geq 0 (see [21, Thm. 33, pp. 81]), to obtain

w(x)\displaystyle w(x) =eq(tTn)w(XtTnπ)0tTneqs(q)w(Xsπ)ds\displaystyle=\operatorname{e}^{-q(t\wedge T_{n})}w({X}^{\pi}_{t\wedge T_{n}})-\int_{0}^{t\wedge T_{n}}\operatorname{e}^{-qs}(\mathcal{L}-q)w({X}^{\pi}_{s-})\mathrm{d}s
(B.5) +[0,tTn]eqsw(Xsπ)dLsπ,cMtTn+JtTn,\displaystyle\quad+\int_{[0,t\wedge T_{n}]}\operatorname{e}^{-qs}w^{\prime}({X}^{\pi}_{s-})\mathrm{d}L_{s}^{\pi,c}-M_{t\wedge T_{n}}+J_{t\wedge T_{n}},

where M:={Mt:t0}M\raisebox{0.4pt}{$:$}=\{M_{t}:t\geq 0\} is a local martingale with M0=0M_{0}=0 and J:={Jt:t>0}J\raisebox{0.4pt}{$:$}=\{J_{t}:t>0\} is given by

Jt\displaystyle J_{t} =0steqs[w(Xsπ+ΔXs)w(Xsπ+ΔXsΔLsπ)]1{ΔLsπ0}\displaystyle=\sum_{0\leq s\leq t}\operatorname{e}^{-qs}\left[w({X}^{\pi}_{s-}+\Delta X_{s})-w({X}^{\pi}_{s-}+\Delta X_{s}-\Delta L^{\pi}_{s})\right]1_{\{\Delta L^{\pi}_{s}\not=0\}}
=0steqsΔLsπ01w(Xsπ+ΔXsλΔLsπ)dλ,t>0.\displaystyle=\sum_{0\leq s\leq t}\operatorname{e}^{-qs}\Delta L^{\pi}_{s}\int_{0}^{1}w^{\prime}(X^{\pi}_{s-}+\Delta X_{s}-\lambda\Delta L^{\pi}_{s})\mathrm{d}\lambda,\quad t>0.

By assumption, we know that (q)w0(\mathcal{L}-q)w\leq 0, wgw^{\prime}\geq g and w0w\geq 0 on (0,)(0,\infty). Hence, taking the expectations in (B) and using (1.2), yields that

w(x)𝔼x[0tTneqsg(Xsπ)dLsπ].\displaystyle w(x)\geq\mathbbm{E}_{x}\left[\int_{0}^{t\wedge T_{n}}\operatorname{e}^{-qs}g(X^{\pi}_{s-})\circ{\rm d}L^{\pi}_{s}\right].

Letting t,nt,n\uparrow\infty in the previous inequality, using monotone convergence, the fact that Tnτ0πT_{n}\uparrow\tau_{0}^{\pi} x\mathbb{P}_{x}-a. s. and that π𝒮0\pi\in\mathcal{S}_{0}, we have that

w(x)𝔼x[0τ0πeqsg(Xsπ)dLsπ].\displaystyle w(x)\geq\mathbbm{E}_{x}\left[\int_{0}^{\tau^{\pi}_{0}}\operatorname{e}^{-qs}g(X^{\pi}_{s-})\circ{\rm d}L^{\pi}_{s}\right].

Therefore w(x)V(x)w(x)\geq V(x) for all x>0x>0. Finally, using the fact that VV is non-decreasing together with the right continuity of ww at zero, we obtain that V(0)limx0V(x)limx0w(x)=w(0)V(0)\leq\lim_{x\downarrow 0}V(x)\leq\lim_{x\downarrow 0}w(x)=w(0). ∎

Proof of Theorem 3.1.

We have by construction that π𝒷𝒮\pi_{\mathpzc{b}^{*}}\in\mathcal{S}, therefore by Lemma 3.2 we only need to show that V𝒷V_{\mathpzc{b}^{*}} satisfies (3.10) for x>0x>0. By (2.1)

(q)V𝒷(x)=(q)(g(𝒷)W(q)(𝒷)W(q)(x))=0,0x𝒷.\displaystyle(\mathcal{L}-q)V_{\mathpzc{b}^{*}}(x)=(\mathcal{L}-q)\left(\frac{g(\mathpzc{b}^{*})}{W^{(q)}(\mathpzc{b}^{*})}W^{(q)}(x)\right)=0,\quad\text{$0\leq x\leq\mathpzc{b}^{*}$}.

On the other hand, (3.5) gives g(x)W(q)(x)g(𝒷)W(q)(𝒷)\dfrac{g(x)}{W^{(q)\prime}(x)}\leq\dfrac{g(\mathpzc{b}^{*})}{W^{(q)\prime}(\mathpzc{b}^{*})} for 0x𝒷0\leq x\leq\mathpzc{b}^{*}, this implies

g(x)V𝒷(x)=g(x)g(𝒷)𝒲(𝓆)(𝓍)𝒲(𝓆)(𝒷)0,0x𝒷.\displaystyle g(x)-V^{\prime}_{\mathpzc{b}^{*}}(x)=g(x)-g(\mathpzc{b}^{*})\frac{W^{(q)\prime}(x)}{W^{(q)\prime}(\mathpzc{b}^{*})}\leq 0,\quad\text{$0\leq x\leq\mathpzc{b}^{*}$}.

Hence, V𝒷V_{\mathpzc{b}^{*}} satisfies (3.10) on [0,𝒷][0,\mathpzc{b}^{*}]. Meanwhile, using (3.2) we note that

(B.6) V(x)=g(x)forx𝒷.V^{\prime}(x)=g(x)\quad\text{for}\ x\geq\mathpzc{b}^{*}.

For any x>𝒷x>\mathpzc{b}^{*}, let us consider the ER associated with the barrier strategy at the level xx, VxV_{x}, which is given in (3.2). We recall that, by Lemma 3.1, VxC1((0,))C2((0,){x})V_{x}\in C^{1}((0,\infty))\cap C^{2}((0,\infty)\setminus\{x\}) and V𝒷C2((0,))V_{\mathpzc{b}^{*}}\in C^{2}((0,\infty)). We aim to prove that

(B.7) (q)(V𝒷Vx)(x)0,x>𝒷.(\mathcal{L}-q)(V_{\mathpzc{b}^{*}}-V_{x})(x-)\leq 0,\quad x>\mathpzc{b}^{*}.

To this end, we check the following:

  • (i)

    Condition (3.11) implies that Vx′′(x)=g(x)W(q)′′(x)W(q)(x)g(x)=V𝒷′′(x)V^{\prime\prime}_{x}(x-)=g(x)\dfrac{W^{(q)\prime\prime}(x)}{W^{(q)\prime}(x)}\geq g^{\prime}(x)=V^{\prime\prime}_{\mathpzc{b}^{*}}(x).

  • (ii)

    By (3.5) we obtain that

    Vx(u)=g(x)W(q)(x)W(q)(u)g(𝒷)W(q)(𝒷)W(q)(u)=V𝒷(u),for u[0,𝒷].\displaystyle V_{x}^{\prime}(u)=\frac{g(x)}{W^{(q)\prime}(x)}W^{(q)\prime}(u)\leq\frac{g(\mathpzc{b}^{*})}{W^{(q)\prime}(\mathpzc{b}^{*})}W^{(q)\prime}(u)=V^{\prime}_{\mathpzc{b}^{*}}(u),\quad\text{for $u\in[0,\mathpzc{b}^{*}]$}.

    On the other hand, by Remark 3.1 we have that g(x)W(q)(x)g(u)W(q)(u)\dfrac{g(x)}{W^{(q)\prime}(x)}\leq\dfrac{g(u)}{W^{(q)\prime}(u)} for u[𝒷,𝓍]u\in[\mathpzc{b}^{*},x]. Hence

    Vx(u)=g(x)W(q)(x)W(q)(u)g(u)=V𝒷(u)foru[𝒷,𝓍].\displaystyle V_{x}^{\prime}(u)=\frac{g(x)}{W^{(q)\prime}(x)}W^{(q)\prime}(u)\leq g(u)=V^{\prime}_{\mathpzc{b}^{*}}(u)\quad\text{for}\ u\in[\mathpzc{b}^{*},x].

    The previous arguments imply that (V𝒷Vx)(u)0(V^{\prime}_{\mathpzc{b}^{*}}-V^{\prime}_{x})(u)\geq 0 for u[0,x]u\in[0,x].

  • (iii)

    By (3.5) we obtain that

    Vx(𝒷)=(𝓍)𝒲(𝓆)(𝓍)𝒲(𝓆)(𝒷)(𝒷)𝒲(𝓆)(𝒷)𝒲(𝓆)(𝒷)=𝒱𝒷(𝒷).V_{x}(\mathpzc{b}^{*})=\frac{g(x)}{W^{(q)\prime}(x)}W^{(q)}(\mathpzc{b}^{*})\leq\frac{g(\mathpzc{b}^{*})}{W^{(q)\prime}(\mathpzc{b}^{*})}W^{(q)}(\mathpzc{b}^{*})=V_{\mathpzc{b}^{*}}(\mathpzc{b}^{*}).

    The above inequality together with (ii) implies that (V𝒷Vx)(x)0(V_{\mathpzc{b}^{*}}-V_{x})(x)\geq 0.

  • (iv)

    By (3.6) we have that Vx(x)=g(x)=V𝒷(x)V_{x}^{\prime}(x)=g(x)=V_{\mathpzc{b}^{*}}^{\prime}(x).

  • (v)

    Using (ii) we obtain that (V𝒷Vx)(xz)(V𝒷Vx)(x)(V_{\mathpzc{b}^{*}}-V_{x})(x-z)\leq(V_{\mathpzc{b}^{*}}-V_{x})(x) for z(0,x)z\in(0,x). Additionally, by (iii) we have (V𝒷Vx)(xz)=0(V𝒷Vx)(x)(V_{\mathpzc{b}^{*}}-V_{x})(x-z)=0\leq(V_{\mathpzc{b}^{*}}-V_{x})(x) for z>xz>x.

Therefore using (i)-(v) and the fact that

(q)(V𝒷Vx)(x)=γ(V𝒷Vx)(x)+σ22(V𝒷′′(x)Vx′′(x))q(V𝒷Vx)(x)+(0,)[(V𝒷Vx)(xz)(V𝒷Vx)(x)(V𝒷Vx)(x)z1{0<z<1}]Π(dz),(\mathcal{L}-q)(V_{\mathpzc{b}^{*}}-V_{x})(x-)=\gamma(V_{\mathpzc{b}^{*}}^{\prime}-V_{x}^{\prime})(x)+\frac{\sigma^{2}}{2}(V_{\mathpzc{b}^{*}}^{\prime\prime}(x)-V^{\prime\prime}_{x}(x-))-q(V_{\mathpzc{b}^{*}}-V_{x})(x)\\ +\int_{(0,\infty)}[(V_{\mathpzc{b}^{*}}-V_{x})(x-z)-(V_{\mathpzc{b}^{*}}-V_{x})(x)-(V_{\mathpzc{b}^{*}}^{\prime}-V_{x}^{\prime})(x)z1_{\{0<z<1\}}]\Pi(dz),

we obtain (B.7). So, proceeding like in the proof of Theorem 2 in [17] we obtain that

(B.8) (q)V𝒷(x)0,for x𝒷.\displaystyle(\mathcal{L}-q)V_{\mathpzc{b}^{*}}(x)\leq 0,\quad\text{for $x\geq\mathpzc{b}^{*}$.}

Hence, by (B.6) and (B.8), V𝒷V_{\mathpzc{b}^{*}} satisfies (3.10) on [𝒷,)[\mathpzc{b}^{*},\infty). ∎

Appendix C Proofs of multibarrier strategies

Proof of Proposition 4.1.

For x[0,𝒷1)x\in[0,\mathpzc{b}_{1}), we obtain, by the strong Markov property, that

V𝕓(x)=𝔼x[eqτ𝒷1+𝔼𝒷1[0τ𝕓eqtg(Xt𝕓)dLt𝕓];τ𝒷1+<τ0]=V𝕓(𝒷1)𝒲(𝓆)(𝓍)𝒲(𝓆)(𝒷1),\displaystyle V_{\mathbbmss{b}}(x)=\mathbbm{E}_{x}\left[\operatorname{e}^{-q\tau_{\mathpzc{b}_{1}}^{+}}\mathbbm{E}_{\mathpzc{b}_{1}}\left[\int_{0}^{\tau_{\mathbbmss{b}}}\operatorname{e}^{-qt}g(X^{\mathbbmss{b}}_{t-})\circ{\rm d}L^{\mathbbmss{b}}_{t}\right];\tau_{\mathpzc{b}_{1}}^{+}<\tau_{0}^{-}\right]=V_{\mathbbmss{b}}(\mathpzc{b}_{1})\frac{W^{(q)}(x)}{W^{(q)}(\mathpzc{b}_{1})},

and proceeding like in the proof of Proposition 3.1 we have

V𝕓(𝒷1)=(𝒷1)𝒲(𝓆)(𝒷1)𝒲(𝓆)(𝒷1).\displaystyle V_{\mathbbmss{b}}(\mathpzc{b}_{1})=g(\mathpzc{b}_{1})\frac{W^{(q)}(\mathpzc{b}_{1})}{W^{(q)\prime}(\mathpzc{b}_{1})}.

On the other hand, for x[𝒷1,𝒷2]x\in[\mathpzc{b}_{1},\mathpzc{b}_{2}]

V𝕓(x)\displaystyle V_{\mathbbmss{b}}(x) =(x𝒷1)01(𝓍λ(𝓍𝒷1))𝒹λ+𝔼𝒷1[0τ𝕓e𝓆𝓉(𝒳𝓉𝕓)d𝓉𝕓]\displaystyle=(x-\mathpzc{b}_{1})\int_{0}^{1}g(x-\lambda(x-\mathpzc{b}_{1}))d\lambda+\mathbbm{E}_{\mathpzc{b}_{1}}\left[\int_{0}^{\tau_{\mathbbmss{b}}}\operatorname{e}^{-qt}g(X^{\mathbbmss{b}}_{t-})\circ{\rm d}L^{\mathbbmss{b}}_{t}\right]
=(x𝒷1)01(𝓍λ(𝓍𝒷1))𝒹λ+𝒱𝕓(𝒷1)\displaystyle=(x-\mathpzc{b}_{1})\int_{0}^{1}g(x-\lambda(x-\mathpzc{b}_{1}))d\lambda+V_{\mathbbmss{b}}(\mathpzc{b}_{1})
=𝒷1xg(u)𝑑u+V𝕓(𝒷1)=𝒢(𝓍)𝒢(𝒷1)+𝒱𝕓(𝒷1)=(𝓍;𝒷1).\displaystyle=\int_{\mathpzc{b}_{1}}^{x}g(u)du+V_{\mathbbmss{b}}(\mathpzc{b}_{1})=G(x)-G(\mathpzc{b}_{1})+V_{\mathbbmss{b}}(\mathpzc{b}_{1})=H(x;\mathpzc{b}_{1}).

Now, let x(𝒷2,𝒷3)x\in(\mathpzc{b}_{2},\mathpzc{b}_{3}). By the Strong Markov property and (2.3), we have that

(C.1) V𝕓(x)\displaystyle V_{\mathbbmss{b}}(x) =𝔼x[0τ𝕓eqtg(Xt𝕓)dLt𝕓;τ𝒷2<τ𝒷3+]+𝔼x[0τ𝕓eqtg(Xt𝕓)dLt𝕓;τ𝒷3+<τ𝒷2]\displaystyle=\mathbbm{E}_{x}\left[\int_{0}^{\tau_{\mathbbmss{b}}}\operatorname{e}^{-qt}g(X^{\mathbbmss{b}}_{t-})\circ{\rm d}L^{\mathbbmss{b}}_{t};\tau_{\mathpzc{b}_{2}}^{-}<\tau_{\mathpzc{b}_{3}}^{+}\right]+\mathbbm{E}_{x}\left[\int_{0}^{\tau_{\mathbbmss{b}}}\operatorname{e}^{-qt}g(X^{\mathbbmss{b}}_{t-})\circ{\rm d}L^{\mathbbmss{b}}_{t};\tau_{\mathpzc{b}_{3}}^{+}<\tau_{\mathpzc{b}_{2}}^{-}\right]
=V𝕓(𝒷2)𝔼𝓍[e𝓆τ𝒷2;τ𝒷2<τ𝒷3+]+𝒱𝕓(𝒷3)𝔼𝓍[eτ𝒷3+;τ𝒷3+<τ𝒷2]\displaystyle=V_{\mathbbmss{b}}(\mathpzc{b}_{2})\mathbbm{E}_{x}\left[\operatorname{e}^{-q\tau_{\mathpzc{b}_{2}}^{-}};\tau_{\mathpzc{b}_{2}}^{-}<\tau_{\mathpzc{b}_{3}}^{+}\right]+V_{\mathbbmss{b}}(\mathpzc{b}_{3})\mathbbm{E}_{x}\left[\operatorname{e}^{-\tau_{\mathpzc{b}_{3}}^{+}};\tau_{\mathpzc{b}_{3}}^{+}<\tau_{\mathpzc{b}_{2}}^{-}\right]
=H(𝒷2;𝒷1)(𝒵(𝓆)(𝓍𝒷2)𝒵(𝓆)(𝒷3𝒷2)𝒲(𝓆)(𝒷3𝒷2)𝒲(𝓆)(𝓍𝒷2))+𝒱𝕓(𝒷3)𝒲(𝓆)(𝓍𝒷2)𝒲(𝓆)(𝒷3𝒷2).\displaystyle=H(\mathpzc{b}_{2};\mathpzc{b}_{1})\left(Z^{(q)}(x-\mathpzc{b}_{2})-\frac{Z^{(q)}(\mathpzc{b}_{3}-\mathpzc{b}_{2})}{W^{(q)}(\mathpzc{b}_{3}-\mathpzc{b}_{2})}W^{(q)}(x-\mathpzc{b}_{2})\right)+V_{\mathbbmss{b}}(\mathpzc{b}_{3})\frac{W^{(q)}(x-\mathpzc{b}_{2})}{W^{(q)}(\mathpzc{b}_{3}-\mathpzc{b}_{2})}.

Note that

V𝕓(𝒷3)=𝔼𝒷3[0τ𝒷2𝒷3e𝓆𝓉(𝒳𝓉𝒷3)d𝓉𝒷3]+𝒱𝕓(𝒷2)𝔼𝒷3[e𝓆τ𝒷2𝒷3;τ𝒷2𝒷3<],\displaystyle V_{\mathbbmss{b}}(\mathpzc{b}_{3})=\mathbbm{E}_{\mathpzc{b}_{3}}\left[\int_{0}^{\tau_{\mathpzc{b}_{2}}^{\mathpzc{b}_{3}}}\operatorname{e}^{-qt}g(X^{\mathpzc{b}_{3}}_{t})\circ{\rm d}L^{\mathpzc{b}_{3}}_{t}\right]+V_{\mathbbmss{b}}(\mathpzc{b}_{2})\mathbbm{E}_{\mathpzc{b}_{3}}\left[\operatorname{e}^{-q\tau_{\mathpzc{b}_{2}}^{\mathpzc{b}_{3}}};\tau_{\mathpzc{b}_{2}}^{\mathpzc{b}_{3}}<\infty\right],

where Xt𝒷3=XtLt𝒷3X^{\mathpzc{b}_{3}}_{t}=X_{t}-L_{t}^{\mathpzc{b}_{3}},

Lt𝒷3=(sup0s<t{Xs𝒷3})0,𝓉0,L_{t}^{\mathpzc{b}_{3}}=\bigg{(}\sup_{0\leq s<t}\{X_{s}-\mathpzc{b}_{3}\}\bigg{)}\vee 0,\quad t\geq 0,

and τ𝒷2𝒷3:=inf{t0:Xt𝒷3<𝒷2}\tau_{\mathpzc{b}_{2}}^{\mathpzc{b}_{3}}\raisebox{0.4pt}{$:$}=\inf\{t\geq 0:X^{\mathpzc{b}_{3}}_{t}<\mathpzc{b}_{2}\}. Hence, by the spatial homogeneity of Brownian motion, we have that

(C.2) V𝕓(𝒷3)=𝔼𝒷3𝒷2[0τ0𝒷3𝒷2e𝓆𝓉(𝒳𝓉𝒷3𝒷2+𝒷2)d𝓉𝒷3𝒷2]+𝒱𝕓(𝒷2)𝔼𝒷3𝒷2[e𝓆τ0𝒷3𝒷2;τ0𝒷3𝒷2<].\displaystyle V_{\mathbbmss{b}}(\mathpzc{b}_{3})=\mathbbm{E}_{\mathpzc{b}_{3}-\mathpzc{b}_{2}}\left[\int_{0}^{\tau_{0}^{\mathpzc{b}_{3}-\mathpzc{b}_{2}}}\operatorname{e}^{-qt}g(X^{\mathpzc{b}_{3}-\mathpzc{b}_{2}}_{t}+\mathpzc{b}_{2})\circ{\rm d}L^{\mathpzc{b}_{3}-\mathpzc{b}_{2}}_{t}\right]+V_{\mathbbmss{b}}(\mathpzc{b}_{2})\mathbbm{E}_{\mathpzc{b}_{3}-\mathpzc{b}_{2}}\left[\operatorname{e}^{-q\tau_{0}^{\mathpzc{b}_{3}-\mathpzc{b}_{2}}};\tau_{0}^{\mathpzc{b}_{3}-\mathpzc{b}_{2}}<\infty\right].

Again, proceeding as in the proof of Proposition 3.1

(C.3) 𝔼𝒷3𝒷2[0τ0𝒷3𝒷2eqtg(Xt𝒷3𝒷2+𝒷2)d𝓉𝒷3𝒷2]=g(𝒷3)𝒲(𝓆)(𝒷3𝒷2)𝒲(𝓆)(𝒷3𝒷2).\displaystyle\mathbbm{E}_{\mathpzc{b}_{3}-\mathpzc{b}_{2}}\left[\int_{0}^{\tau_{0}^{\mathpzc{b}_{3}-\mathpzc{b}_{2}}}\operatorname{e}^{-qt}g(X^{\mathpzc{b}_{3}-\mathpzc{b}_{2}}_{t}+\mathpzc{b}_{2})\circ{\rm d}L^{\mathpzc{b}_{3}-\mathpzc{b}_{2}}_{t}\right]=g(\mathpzc{b}_{3})\frac{W^{(q)}(\mathpzc{b}_{3}-\mathpzc{b}_{2})}{W^{(q)\prime}(\mathpzc{b}_{3}-\mathpzc{b}_{2})}.

On the other hand, by identity (3.10) in [6], we obtain

(C.4) 𝔼𝒷3𝒷2[eqτ0𝒷3𝒷2;τ0𝒷3𝒷2<]=Z(q)(𝒷3𝒷2)𝓆(𝒲(𝓆)(𝒷3𝒷2))2𝒲(𝓆)(𝒷3𝒷2).\displaystyle\mathbbm{E}_{\mathpzc{b}_{3}-\mathpzc{b}_{2}}\left[\operatorname{e}^{-q\tau_{0}^{\mathpzc{b}_{3}-\mathpzc{b}_{2}}};\tau_{0}^{\mathpzc{b}_{3}-\mathpzc{b}_{2}}<\infty\right]=Z^{(q)}(\mathpzc{b}_{3}-\mathpzc{b}_{2})-q\frac{(W^{(q)}(\mathpzc{b}_{3}-\mathpzc{b}_{2}))^{2}}{W^{(q)\prime}(\mathpzc{b}_{3}-\mathpzc{b}_{2})}.

Therefore, (C.1)–(C.4) give V𝒷3(𝒷3)=ϕ(𝒷3;{𝒷𝒾}𝒾=13)V_{\mathpzc{b}_{3}}(\mathpzc{b}_{3})=\phi(\mathpzc{b}_{3};\{\mathpzc{b}_{i}\}_{i=1}^{3}). Finally, we obtain the result by induction and similar arguments as above. ∎

Proof of Proposition 4.2.

(i) Since F(𝒷1)=0F^{\prime}(\mathpzc{b}^{*}_{1})=0, by (3.3), (3.7) and (A.1), we get that

σ22g(𝒷1)+μ(𝒷1)\displaystyle\dfrac{\sigma^{2}}{2}g^{\prime}(\mathpzc{b}_{1}^{*})+\mu g(\mathpzc{b}_{1}^{*}) =σ22g(𝒷1)𝒲(𝓆)′′(𝒷1)𝒲(𝓆)(𝒷1)+μ(𝒷1)\displaystyle=\dfrac{\sigma^{2}}{2}g(\mathpzc{b}_{1}^{*})\dfrac{W^{(q)\prime\prime}(\mathpzc{b}_{1}^{*})}{W^{(q)\prime}(\mathpzc{b}_{1}^{*})}+\mu g(\mathpzc{b}_{1}^{*})
=g(𝒷1)W(q)(𝒷1)(σ22W(q)′′(𝒷1)μ𝒲(𝓆)(𝒷1))\displaystyle=\dfrac{g(\mathpzc{b}_{1}^{*})}{W^{(q)\prime}(\mathpzc{b}_{1}^{*})}\left(\dfrac{\sigma^{2}}{2}W^{(q)\prime\prime}(\mathpzc{b}_{1}^{*})-\mu W^{(q)\prime}(\mathpzc{b}_{1}^{*})\right)
=qW(q)(𝒷1)(𝒷1)𝒲(𝓆)(𝒷1)=𝓆(𝒷1;𝒷1).\displaystyle=qW^{(q)}(\mathpzc{b}_{1}^{*})\dfrac{g(\mathpzc{b}_{1}^{*})}{W^{(q)\prime}(\mathpzc{b}_{1}^{*})}=qH(\mathpzc{b}_{1}^{*};\mathpzc{b}_{1}^{*}).

Thus, by (4.9), with k=0k=0, we obtain (4.14).

(ii) We first show that

(C.5) H(v;𝒷1)>(𝓋)𝒲(𝓆)(𝓋)𝒲(𝓆)(𝓋)=:Ξ(𝓋)for𝓋(𝒷1,𝒷1(1)).\displaystyle H(v;\mathpzc{b}^{*}_{1})>g(v)\dfrac{W^{(q)}(v)}{W^{(q)\prime}(v)}=:\Xi(v)\quad\text{for}\ v\in(\mathpzc{b}^{*}_{1},\mathpzc{b}^{(1)}_{1}).

This follows since

H(𝒷1;𝒷1)=(𝒷1)𝒲(𝓆)(𝒷1)𝒲(𝓆)(𝒷1)=Ξ(𝒷1),H(\mathpzc{b}^{*}_{1};\mathpzc{b}^{*}_{1})=g(\mathpzc{b}^{*}_{1})\dfrac{W^{(q)}(\mathpzc{b}^{*}_{1})}{W^{(q)\prime}(\mathpzc{b}^{*}_{1})}=\Xi(\mathpzc{b}^{*}_{1}),

and (4.12) implies that for v(𝒷1,𝒷1(1))\ v\in(\mathpzc{b}^{*}_{1},\mathpzc{b}^{(1)}_{1})

ddvΞ(v)\displaystyle\dfrac{\mathrm{d}}{\mathrm{d}v}\Xi(v) =g(v)+W(q)(v)W(q)(v)(g(v)g(v)W(q)′′(v)W(q)(v))\displaystyle=g(v)+\dfrac{W^{(q)}(v)}{W^{(q)\prime}(v)}\left(g^{\prime}(v)-\dfrac{g(v)W^{(q)\prime\prime}(v)}{W^{(q)\prime}(v)}\right)
<g(v)=ddvH(v;𝒷1).\displaystyle<g(v)=\dfrac{\mathrm{d}}{\mathrm{d}v}H(v;\mathpzc{b}^{*}_{1}).

Therefore, using again (4.12), (A.1) and (C.5) we get that (4.15) holds on (𝒷1,𝒷1(1))(\mathpzc{b}^{*}_{1},\mathpzc{b}^{(1)}_{1}). ∎

Proof of Lemma 4.1.

Using (4.7), (4.18) is equivalent to F¯(𝒸1,𝓏,𝒷1)>0\overline{F}(\mathpzc{c}_{1},z,\mathpzc{b}^{*}_{1})>0 and using (A.1), this is equivalent to

Γ(z):=g(z)qH(𝒸1;𝒷1)𝒲(𝓆)(𝓏𝒸1)2σ2(𝒸1,𝓏;𝒷1)(𝓆𝒲(𝓆)(𝓏𝒸1)μ𝒲(𝓆)(𝓏𝒸1))>0.\Gamma(z)\raisebox{0.4pt}{$:$}=g^{\prime}(z)-qH(\mathpzc{c}_{1};\mathpzc{b}^{*}_{1})W^{(q)\prime}(z-\mathpzc{c}_{1})-\dfrac{2}{\sigma^{2}}F(\mathpzc{c}_{1},z;\mathpzc{b}^{*}_{1})\bigg{(}qW^{(q)}(z-\mathpzc{c}_{1})-\mu W^{(q)\prime}(z-\mathpzc{c}_{1})\bigg{)}>0.

Since Γ(𝒸1)=0\Gamma(\mathpzc{c}_{1})=0 due to (2.6), (4.6), (4.9) and (4.16), it is enough to verify that Γ(𝒸1)>0\Gamma^{\prime}(\mathpzc{c}_{1})>0 in order to prove (4.18) by the C1C^{1}-continuity of this function. Then,

Γ(z)=\displaystyle\Gamma^{\prime}(z)= g′′(z)qH(𝒸1;𝒷1)𝒲(𝓆)′′(𝓏𝒸1)2σ2(𝒸1,𝓏;𝒷1)(𝓆𝒲(𝓆)(𝓏𝒸1)μ𝒲(𝓆)′′(𝓏𝒸1))\displaystyle g^{\prime\prime}(z)-qH(\mathpzc{c}_{1};\mathpzc{b}^{*}_{1})W^{(q)\prime\prime}(z-\mathpzc{c}_{1})-\dfrac{2}{\sigma^{2}}F(\mathpzc{c}_{1},z;\mathpzc{b}^{*}_{1})\bigg{(}qW^{(q)\prime}(z-\mathpzc{c}_{1})-\mu W^{(q)\prime\prime}(z-\mathpzc{c}_{1})\bigg{)}
2σ2F¯(𝒸1,𝓏;𝒷1)W(q)(z𝒸1)(qW(q)(z𝒸1)μ𝒲(𝓆)(𝓏𝒸1))\displaystyle-\dfrac{2}{\sigma^{2}}\frac{\overline{F}(\mathpzc{c}_{1},z;\mathpzc{b}^{*}_{1})}{W^{(q)\prime}(z-\mathpzc{c}_{1})}\bigg{(}qW^{(q)}(z-\mathpzc{c}_{1})-\mu W^{(q)\prime}(z-\mathpzc{c}_{1})\bigg{)}

and taking z𝒸1z\downarrow\mathpzc{c}_{1}, by (4.9) and (4.16) we obtain that

Γ(𝒸1)\displaystyle\Gamma^{\prime}(\mathpzc{c}_{1}) =g′′(𝒸1)+𝓆(𝒸1;𝒷1)4μσ42𝓆σ2(𝒸1)4μ2σ4(𝒸1)\displaystyle=g^{\prime\prime}(\mathpzc{c}_{1})+qH(\mathpzc{c}_{1};\mathpzc{b}^{*}_{1})\dfrac{4\mu}{\sigma^{4}}-\dfrac{2q}{\sigma^{2}}g(\mathpzc{c}_{1})-\dfrac{4\mu^{2}}{\sigma^{4}}g(\mathpzc{c}_{1})
=g′′(𝒸1)+𝓆(𝒸1;𝒷1)4μσ42𝓆σ2(𝒸1)2μσ2(2𝓆σ2(𝒸1;𝒷1)(𝒸1))\displaystyle=g^{\prime\prime}(\mathpzc{c}_{1})+qH(\mathpzc{c}_{1};\mathpzc{b}^{*}_{1})\dfrac{4\mu}{\sigma^{4}}-\dfrac{2q}{\sigma^{2}}g(\mathpzc{c}_{1})-\dfrac{2\mu}{\sigma^{2}}\left(\dfrac{2q}{\sigma^{2}}H(\mathpzc{c}_{1};\mathpzc{b}^{*}_{1})-g^{\prime}(\mathpzc{c}_{1})\right)
=g′′(𝒸1)2𝓆σ2(𝒸1)+2μσ2(𝒸1).\displaystyle=g^{\prime\prime}(\mathpzc{c}_{1})-\dfrac{2q}{\sigma^{2}}g(\mathpzc{c}_{1})+\dfrac{2\mu}{\sigma^{2}}g^{\prime}(\mathpzc{c}_{1}).

Now, from (4.16) we know that there exists ε1>0{\varepsilon}_{1}>0 such that if z(𝒸1,𝒸1+ε1)z\in(\mathpzc{c}_{1},\mathpzc{c}_{1}+\varepsilon_{1}), then

2qσ2H(z;𝒷1)<(𝓏)+2μσ2(𝓏)=:𝒽(𝓏).\dfrac{2q}{\sigma^{2}}H(z;\mathpzc{b}^{*}_{1})<g^{\prime}(z)+\dfrac{2\mu}{\sigma^{2}}g(z)=:h(z).

We also know that 2qσ2H(𝒸1;𝒷1)=𝒽(𝒸1)\dfrac{2q}{\sigma^{2}}H(\mathpzc{c}_{1};\mathpzc{b}^{*}_{1})=h(\mathpzc{c}_{1}). Hence, by taking the derivative at z=𝒸1z=\mathpzc{c}_{1} we must have that

2qσ2g(𝒸1)<𝒽(𝒸1)=′′(𝒸1)+2μσ2(𝒸1),\dfrac{2q}{\sigma^{2}}g(\mathpzc{c}_{1})<h^{\prime}(\mathpzc{c}_{1})=g^{\prime\prime}(\mathpzc{c}_{1})+\dfrac{2\mu}{\sigma^{2}}g^{\prime}(\mathpzc{c}_{1}),

therefore Γ(𝒸1)>0\Gamma^{\prime}(\mathpzc{c}_{1})>0. ∎

Proof of Proposition 4.3.

By (4.13), it follows that for v(𝒷1,𝒷1(1))v\in(\mathpzc{b}^{*}_{1},\mathpzc{b}^{(1)}_{1}),

g(𝒷1)W(q)(𝒷1)>g(v)W(q)(v)andg(z)<g(v)W(q)(z)W(q)(v),forz>v.\displaystyle\dfrac{g(\mathpzc{b}^{*}_{1})}{W^{(q)\prime}(\mathpzc{b}^{*}_{1})}>\dfrac{g(v)}{W^{(q)\prime}(v)}\quad\text{and}\quad g(z)<g(v)\dfrac{W^{(q)\prime}(z)}{W^{(q)\prime}(v)},\quad\text{for}\ z>v.

From here and (3.3), we have that

F(v,z;𝒷1)\displaystyle F(v,z;\mathpzc{b}^{*}_{1}) =g(z)qH(v;𝒷1)𝒲(𝓆)(𝓏𝓋)W(q)(zv)\displaystyle=\dfrac{g(z)-qH(v;\mathpzc{b}^{*}_{1})W^{(q)}(z-v)}{W^{(q)\prime}(z-v)}
=g(z)qg(𝒷1)𝒲(𝓆)(𝒷1)𝒲(𝓆)(𝒷1)𝒲(𝓆)(𝓏𝓋)𝓆(𝒢(𝓋)𝒢(𝒷1))𝒲(𝓆)(𝓏𝓋)W(q)(zv)\displaystyle=\dfrac{g(z)-qg(\mathpzc{b}^{*}_{1})\dfrac{W^{(q)}(\mathpzc{b}^{*}_{1})}{W^{(q)\prime}(\mathpzc{b}^{*}_{1})}W^{(q)}(z-v)-q(G(v)-G(\mathpzc{b}^{*}_{1}))W^{(q)}(z-v)}{W^{(q)\prime}(z-v)}
<g(v)(W(q)(z)qW(q)(𝒷1)𝒲(𝓆)(𝓏𝓋)W(q)(v)W(q)(zv))q(G(v)G(𝒷1))𝒲(𝓆)(𝓏𝓋)𝒲(𝓆)(𝓏𝓋)=:Λ(𝓋,𝓏;𝒷1).\displaystyle<g(v)\bigg{(}\dfrac{W^{(q)\prime}(z)-qW^{(q)}(\mathpzc{b}^{*}_{1})W^{(q)}(z-v)}{W^{(q)\prime}(v)W^{(q)\prime}(z-v)}\bigg{)}-q(G(v)-G(\mathpzc{b}^{*}_{1}))\dfrac{W^{(q)}(z-v)}{W^{(q)\prime}(z-v)}=:\Lambda(v,z;\mathpzc{b}^{*}_{1}).

Observe that if zΛ(v,z;𝒷1)z\mapsto\Lambda(v,z;\mathpzc{b}^{*}_{1}) is decreasing for z>vz>v, it follows that F(v,z;𝒷1)<σ22(𝓋)F(v,z;\mathpzc{b}^{*}_{1})<\dfrac{\sigma^{2}}{2}g(v) for z>vz>v, because of Λ(v,v;𝒷1)=(𝓋,𝓋;𝒷1)=σ22(𝓋)\Lambda(v,v;\mathpzc{b}^{*}_{1})=F(v,v;\mathpzc{b}_{1}^{*})=\dfrac{\sigma^{2}}{2}g(v). Let us verify that zΛ(z,v;𝒷1)<0\dfrac{\partial}{\partial z}\Lambda(z,v;\mathpzc{b}^{*}_{1})<0. Calculating the first derivative with respect to zz, we get that

g(v)W(q)(v)(W(q)(zv))2((W(q)′′(z)qW(q)(𝒷1)𝒲(𝓆)(𝓏𝓋))𝒲(𝓆)(𝓏𝓋)\displaystyle\dfrac{g(v)}{W^{(q)\prime}(v)(W^{(q)\prime}(z-v))^{2}}\bigg{(}\Big{(}W^{(q)\prime\prime}(z)-qW^{(q)}(\mathpzc{b}^{*}_{1})W^{(q)\prime}(z-v)\Big{)}W^{(q)\prime}(z-v)
(W(q)(z)qW(q)(𝒷1)𝒲(𝓆)(𝓏𝓋))𝒲(𝓆)′′(𝓏𝓋))\displaystyle\quad-\Big{(}W^{(q)\prime}(z)-qW^{(q)}(\mathpzc{b}^{*}_{1})W^{(q)}(z-v)\Big{)}W^{(q)\prime\prime}(z-v)\bigg{)}
q(G(v)G(𝒷1))(W(q)(zv))2((W(q)(zv))2W(q)(zv)W(q)′′(zv))<0\displaystyle\quad-\dfrac{q(G(v)-G(\mathpzc{b}_{1}^{*}))}{(W^{(q)\prime}(z-v))^{2}}\Big{(}(W^{(q)\prime}(z-v))^{2}-W^{(q)}(z-v)W^{(q)\prime\prime}(z-v)\Big{)}<0
\displaystyle\Longleftrightarrow
(C.6) g(v)W(q)(v)(W(q)′′(z)W(q)(zv)W(q)(z)W(q)′′(zv))\displaystyle\dfrac{g(v)}{W^{(q)\prime}(v)}\Big{(}W^{(q)\prime\prime}(z)W^{(q)\prime}(z-v)-W^{(q)\prime}(z)W^{(q)\prime\prime}(z-v)\Big{)}
q(G(v)G(𝒷1)+(𝓋)𝒲(𝓆)(𝓋)𝒲(𝓆)(𝒷1))((𝒲(𝓆)(𝓏𝓋))2𝒲(𝓆)(𝓏𝓋)𝒲(𝓆)′′(𝓏𝓋))<0.\displaystyle\quad-q\bigg{(}G(v)-G(\mathpzc{b}_{1}^{*})+\dfrac{g(v)}{W^{(q)\prime}(v)}W^{(q)}(\mathpzc{b}^{*}_{1})\bigg{)}\Big{(}(W^{(q)\prime}(z-v))^{2}-W^{(q)}(z-v)W^{(q)\prime\prime}(z-v)\Big{)}<0.

Using (A.3)–(A.4), we see that (C) is equivalent to

g(v)W(q)(v)4qσ4e(Φ(q)ξ1)(zv)W(q)(v)q(G(v)G(𝒷1)+(𝓋)𝒲(𝓆)(𝓋)𝒲(𝓆)(𝒷1))4σ4e(Φ(𝓆)ξ1)(𝓏𝓋)<0\displaystyle\dfrac{g(v)}{W^{(q)\prime}(v)}\dfrac{4q}{\sigma^{4}}\operatorname{e}^{(\Phi(q)-\xi_{1})(z-v)}W^{(q)}(v)-q\bigg{(}G(v)-G(\mathpzc{b}_{1}^{*})+\dfrac{g(v)}{W^{(q)\prime}(v)}W^{(q)}(\mathpzc{b}^{*}_{1})\bigg{)}\dfrac{4}{\sigma^{4}}\operatorname{e}^{(\Phi(q)-\xi_{1})(z-v)}<0
\displaystyle\Longleftrightarrow
(C.7) g(v)W(q)(v)(W(q)(v)W(q)(𝒷1))<𝒢(𝓋)𝒢(𝒷1).\displaystyle\dfrac{g(v)}{W^{(q)\prime}(v)}(W^{(q)}(v)-W^{(q)}(\mathpzc{b}^{*}_{1}))<G(v)-G(\mathpzc{b}_{1}^{*}).

Since both sides of (C) are zero at v=𝒷1v=\mathpzc{b}^{*}_{1}, (C) is equivalent to prove that

ddv[g(v)W(q)(v)(W(q)(v)W(q)(𝒷1))]<dd𝓋[𝒢(𝓋)𝒢(𝒷1)]\displaystyle\dfrac{\mathrm{d}}{\mathrm{d}v}\bigg{[}\dfrac{g(v)}{W^{(q)\prime}(v)}(W^{(q)}(v)-W^{(q)}(\mathpzc{b}^{*}_{1}))\Bigg{]}<\dfrac{\mathrm{d}}{\mathrm{d}v}[G(v)-G(\mathpzc{b}_{1}^{*})]
\displaystyle\Longleftrightarrow
(W(q)(v)W(q)(𝒷1))(W(q)(v))2(g(v)W(q)(v)g(v)W(q)′′(v))+g(v)W(q)(v)W(q)(v)<g(v)\displaystyle\dfrac{\Big{(}W^{(q)}(v)-W^{(q)}(\mathpzc{b}^{*}_{1})\Big{)}}{(W^{(q)\prime}(v))^{2}}\Big{(}g^{\prime}(v)W^{(q)\prime}(v)-g(v)W^{(q)\prime\prime}(v)\Big{)}+\dfrac{g(v)}{W^{(q)\prime}(v)}W^{(q)\prime}(v)<g(v)
\displaystyle\Longleftrightarrow
(W(q)(v)W(q)(𝒷1))W(q)(v)(g(v)g(v)W(q)′′(v)W(q)(v))<0.\displaystyle\dfrac{\Big{(}W^{(q)}(v)-W^{(q)}(\mathpzc{b}^{*}_{1})\Big{)}}{W^{(q)\prime}(v)}\bigg{(}g^{\prime}(v)-g(v)\dfrac{W^{(q)\prime\prime}(v)}{W^{(q)\prime}(v)}\bigg{)}<0.

From here and (4.12), it follows that (4.20) is true. Letting v𝒷1v\downarrow\mathpzc{b}^{*}_{1} it can be easily verified that (4.20) holds for 𝒷1\mathpzc{b}_{1}^{*}, since F(,;𝒷1)𝒞(𝒜¯1)F(\cdot,\cdot;\mathpzc{b}^{*}_{1})\in C(\bar{\mathcal{A}}_{1}). ∎

Proof of Theorem 4.1.

Condition (4.21) immediately implies that 𝒷2<𝒸1\mathpzc{b}^{*}_{2}<\mathpzc{c}_{1}, so it only remains to show that 𝒷2𝒟1\mathpzc{b}^{*}_{2}\in\mathcal{D}_{1} to prove the strict inequalities. Suppose that 𝒷2𝒟1\mathpzc{b}^{*}_{2}\notin\mathcal{D}_{1}, that is,

(C.8) F(𝒷2,𝓏;𝒷1)<(𝒷2,𝒷2;𝒷1)=σ22(𝒷2),for𝓏>𝒷2.F(\mathpzc{b}^{*}_{2},z;\mathpzc{b}^{*}_{1})<F(\mathpzc{b}^{*}_{2},\mathpzc{b}^{*}_{2};\mathpzc{b}^{*}_{1})=\frac{\sigma^{2}}{2}g(\mathpzc{b}_{2}^{*}),\quad\text{for}\ z>\mathpzc{b}^{*}_{2}.

Note that (C.8) is equivalent to

(C.9) Λ¯(𝒷2,𝓏;𝒷1)>(𝓏),for𝓏>𝒷2,\bar{\Lambda}(\mathpzc{b}^{*}_{2},z;\mathpzc{b}^{*}_{1})>F(z),\quad\text{for}\ z>\mathpzc{b}^{*}_{2},

where

Λ¯(v,z;𝒷1):=σ22(𝓋)𝒲(𝓆)(𝓏𝓋)𝒲(𝓆)(𝓏)+𝓆(𝓋;𝒷1)𝒲(𝓆)(𝓏𝓋)𝒲(𝓆)(𝓏),for𝓏𝓋.\bar{\Lambda}(v,z;\mathpzc{b}^{*}_{1})\raisebox{0.4pt}{$:$}=\dfrac{\sigma^{2}}{2}g(v)\dfrac{W^{(q)\prime}(z-v)}{W^{(q)\prime}(z)}+qH(v;\mathpzc{b}^{*}_{1})\dfrac{W^{(q)}(z-v)}{W^{(q)\prime}(z)},\quad\text{for}\ z\geq v.

Also note that for any 𝒷1𝓋\mathpzc{b}^{*}_{1}\leq v, F(v)=Λ¯(v,v+;𝒷1)F(v)=\bar{\Lambda}(v,v+;\mathpzc{b}^{*}_{1}). Additionally, differentiating Λ¯(v,z;𝒷1)\bar{\Lambda}(v,z;\mathpzc{b}^{*}_{1}) with respect to zz, letting zvz\downarrow v, and using (2.6), (2.7), (3.7) and (4.9), it follows that

zΛ¯(v,v+;𝒷1)\displaystyle\partial_{z}\bar{\Lambda}(v,v+;\mathpzc{b}^{*}_{1}) =1W(q)(v)(2qσ2H(v;𝒷1)2μσ2(𝓋)(𝓋)𝒲(𝓆)′′(𝓋)𝒲(𝓆)(𝓋))\displaystyle=\frac{1}{W^{(q)\prime}(v)}\left(\dfrac{2q}{\sigma^{2}}H(v;\mathpzc{b}^{*}_{1})-\dfrac{2\mu}{\sigma^{2}}g(v)-g(v)\frac{W^{(q)\prime\prime}(v)}{W^{(q)\prime}(v)}\right)
=F(v)1W(q)(v)2σ2(q)H(v;𝒷1).\displaystyle=F^{\prime}(v)-\frac{1}{W^{(q)\prime}(v)}\dfrac{2}{\sigma^{2}}(\mathcal{L}-q)H(v;\mathpzc{b}^{*}_{1}).

Now, using (4.21), it yields zΛ¯(v,v+;𝒷1)>(𝓋)\partial_{z}\bar{\Lambda}(v,v+;\mathpzc{b}^{*}_{1})>F^{\prime}(v) for v(𝒷2,𝒷2+ϵ1)v\in(\mathpzc{b}^{*}_{2},\mathpzc{b}^{*}_{2}+\epsilon_{1}). From here, the smoothness of Λ¯\bar{\Lambda} and (C.9), it implies that there exists ϵ¯1<ϵ1\bar{\epsilon}_{1}<\epsilon_{1} such that Λ¯(v,z;𝒷1)>(𝓏)\bar{\Lambda}(v,z;\mathpzc{b}^{*}_{1})>F(z), for v(𝒷2,𝒷2+ϵ¯1)v\in(\mathpzc{b}^{*}_{2},\mathpzc{b}^{*}_{2}+\bar{\epsilon}_{1}) and z>vz>v, which is equivalent to F(v,z;𝒷1)>(𝓋,𝓋;𝒷1)F(v,z;\mathpzc{b}^{*}_{1})>F(v,v;\mathpzc{b}^{*}_{1}), for v(𝒷2,𝒷2+ϵ¯1)v\in(\mathpzc{b}^{*}_{2},\mathpzc{b}^{*}_{2}+\bar{\epsilon}_{1}) and z>vz>v. Thus, 𝒷2\mathpzc{b}^{*}_{2} cannot be the infimum of 𝒟1\mathcal{D}_{1}, which is a contradiction. Therefore, 𝒷2𝒟1\mathpzc{b}^{*}_{2}\in\mathcal{D}_{1}. Additionally, using again (4.21) and arguing similarly as in the proof of Lemma 4.1, it can be proven that zF(𝒷2,𝓏;𝒷1)z\mapsto F(\mathpzc{b}^{*}_{2},z;\mathpzc{b}^{*}_{1}) is non-increasing locally at 𝒷2\mathpzc{b}^{*}_{2}, and concluding that (4.22) follows due to the continuity of F(,;𝒷1)F(\cdot,\cdot;\mathpzc{b}_{1}^{*}). ∎

Proof of Proposition 4.4.

We first show that for v(𝒷2𝓀1,𝒷2𝓀1(1))v\in(\mathpzc{b}^{*}_{2k-1},\mathpzc{b}^{(1)}_{2k-1})

(C.10) H(v;𝕓2k1)>H(𝒷2𝓀2;𝕓2𝓀3)𝒵(𝓆)(𝓋𝒷2𝓀2)+𝒲(𝓆)(𝓋𝒷2𝓀2)(𝒷2𝓀2,𝓋;𝕓2𝓀3)=:Ξ(𝓋;𝕓2𝓀2).H(v;\mathbbmss{b}^{*}_{2k-1})\\ >H(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})Z^{(q)}(v-\mathpzc{b}^{*}_{2k-2})+W^{(q)}(v-\mathpzc{b}^{*}_{2k-2})F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})=:\Xi(v;\mathbbmss{b}^{*}_{2k-2}).

This follows since

H(𝒷2𝓀1;𝕓2𝓀1)=ϕ(𝒷2𝓀1;𝕓2𝓀1)=Ξ(𝒷2𝓀1;𝕓2𝓀2),H(\mathpzc{b}^{*}_{2k-1};\mathbbmss{b}^{*}_{2k-1})=\phi(\mathpzc{b}^{*}_{2k-1};\mathbbmss{b}^{*}_{2k-1})=\Xi(\mathpzc{b}^{*}_{2k-1};\mathbbmss{b}^{*}_{2k-2}),

and, (4.7) and (4.31) implies that for v(𝒷2𝓀1,𝒷2𝓀1(1))\ v\in(\mathpzc{b}^{*}_{2k-1},\mathpzc{b}^{(1)}_{2k-1})

ddvΞ(v;𝕓2k2)\displaystyle\dfrac{\mathrm{d}}{\mathrm{d}v}\Xi(v;\mathbbmss{b}^{*}_{2k-2}) =qH(𝒷2𝓀2;𝕓2𝓀3)𝒲(𝓆)(𝓋𝒷2𝓀2)+𝒲(𝓆)(𝓋𝒷2𝓀2)(𝒷2𝓀2,𝓋;𝕓2𝓀3)\displaystyle=qH(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})W^{(q)}(v-\mathpzc{b}^{*}_{2k-2})+W^{(q)\prime}(v-\mathpzc{b}^{*}_{2k-2})F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})
+W(q)(v𝒷2𝓀2)¯(𝒷2𝓀2,𝓋;𝕓2𝓀3)𝒲(𝓆)(𝓋𝒷2𝓀2)\displaystyle\quad+W^{(q)}(v-\mathpzc{b}^{*}_{2k-2})\frac{\overline{F}(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})}{W^{(q)\prime}(v-\mathpzc{b}^{*}_{2k-2})}
<qH(𝒷2𝓀2;𝕓2𝓀3)𝒲(𝓆)(𝓋𝒷2𝓀2)+𝒲(𝓆)(𝓋𝒷2𝓀2)(𝒷2𝓀2,𝓋;𝕓2𝓀3)\displaystyle<qH(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})W^{(q)}(v-\mathpzc{b}^{*}_{2k-2})+W^{(q)\prime}(v-\mathpzc{b}^{*}_{2k-2})F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})
=g(v)=ddvH(v;𝕓2k1).\displaystyle=g(v)=\dfrac{\mathrm{d}}{\mathrm{d}v}H(v;\mathbbmss{b}^{*}_{2k-1}).

Therefore, using again (4.31), (A.1), (A.2) and (C.10) we get that for v(𝒷1,𝒷2𝓀1(1))\ v\in(\mathpzc{b}^{*}_{1},\mathpzc{b}^{(1)}_{2k-1})

σ22g(v)+μg(v)\displaystyle\dfrac{\sigma^{2}}{2}g^{\prime}(v)+\mu g(v) <σ22(g(v)F¯(𝒷2𝓀2,𝓋;𝕓2𝓀3))\displaystyle<\dfrac{\sigma^{2}}{2}\left(g^{\prime}(v)-\overline{F}(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})\right)
+μ(qH(𝒷2𝓀2;𝕓2𝓀3)𝒲(𝓆)(𝓋𝒷2𝓀2)+𝒲(𝓆)(𝓋𝒷2𝓀2)(𝒷2𝓀2,𝓋;𝕓2𝓀3))\displaystyle\quad+\mu\left(qH(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})W^{(q)}(v-\mathpzc{b}^{*}_{2k-2})+W^{(q)\prime}(v-\mathpzc{b}^{*}_{2k-2})F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})\right)
=qH(𝒷2𝓀2;𝕓2𝓀3)(μ𝒲(𝓆)(𝓋𝒷2𝓀2)+σ22𝒲(𝓆)(𝓋𝒷2𝓀2))\displaystyle=qH(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})\left(\mu W^{(q)}(v-\mathpzc{b}^{*}_{2k-2})+\dfrac{\sigma^{2}}{2}W^{(q)\prime}(v-\mathpzc{b}^{*}_{2k-2})\right)
+F(𝒷2𝓀2,𝓋;𝕓2𝓀3)(μ𝒲(𝓆)(𝓋𝒷2𝓀2)+σ22𝒲(𝓆)′′(𝓋𝒷2𝓀2))\displaystyle\quad+F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})\left(\mu W^{(q)\prime}(v-\mathpzc{b}^{*}_{2k-2})+\dfrac{\sigma^{2}}{2}W^{(q)\prime\prime}(v-\mathpzc{b}^{*}_{2k-2})\right)
=q(H(𝒷2𝓀2;𝕓2𝓀3)𝒵(𝓆)(𝓋𝒷2𝓀2)+𝒲(𝓆)(𝓋𝒷2𝓀2)(𝒷2𝓀2,𝓋;𝕓2𝓀3))\displaystyle=q\left(H(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})Z^{(q)}(v-\mathpzc{b}^{*}_{2k-2})+W^{(q)}(v-\mathpzc{b}^{*}_{2k-2})F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})\right)
=qΞ(v;𝕓2k2)<qH(v;𝕓2k1).\displaystyle=q\Xi(v;\mathbbmss{b}^{*}_{2k-2})<qH(v;\mathbbmss{b}^{*}_{2k-1}).

Equality at v=𝒷2𝓀1v=\mathpzc{b}^{*}_{2k-1} holds by equality in (4.31) at v=𝒷2𝓀1v=\mathpzc{b}^{*}_{2k-1}. ∎

Proof of Proposition 4.5.

By (4.32), it follows that for v(𝒷2𝓀1,𝒷2𝓀1(1))v\in(\mathpzc{b}_{2k-1}^{*},\mathpzc{b}_{2k-1}^{(1)}),

g(z)qH(𝒷2𝓀2;𝕓2𝓀3)𝒲(𝓆)(𝓏𝒷2𝓀2)W(q)(z𝒷2𝓀2)<F(𝒷2𝓀2,𝓋;𝕓2𝓀3), for𝓏>𝓋,\displaystyle\frac{g(z)-qH(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})W^{(q)}(z-\mathpzc{b}^{*}_{2k-2})}{W^{(q)\prime}(z-\mathpzc{b}^{*}_{2k-2})}<F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3}),\quad\text{ for}\ z>v,
\displaystyle\Longleftrightarrow
g(z)<F(𝒷2𝓀2,𝓋;𝕓2𝓀3)𝒲(𝓆)(𝓏𝒷2𝓀2)+𝓆(𝒷2𝓀2;𝕓2𝓀3)𝒲(𝓆)(𝓏𝒷2𝓀2), for𝓏>𝓋,\displaystyle g(z)<F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})W^{(q)\prime}(z-\mathpzc{b}^{*}_{2k-2})+qH(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})W^{(q)}(z-\mathpzc{b}^{*}_{2k-2}),\quad\text{ for}\ z>v,

and F(𝒷2𝓀2,𝒷2𝓀1;𝕓2𝓀3)<(𝒷2𝓀2,𝓋;𝕓2𝓀3)-F(\mathpzc{b}^{*}_{2k-2},\mathpzc{b}^{*}_{2k-1};\mathbbmss{b}^{*}_{2k-3})<-F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3}). From here, (4.3), (4.4) and (4.5), we have that

F(v,z;𝕓2k1)\displaystyle F(v,z;\mathbbmss{b}^{*}_{2k-1}) =1W(q)(zv)[g(z)qH(v;𝒷2𝓀1)𝒲(𝓆)(𝓏𝓋)]\displaystyle=\dfrac{1}{W^{(q)\prime}(z-v)}[g(z)-qH(v;\mathpzc{b}^{*}_{2k-1})W^{(q)}(z-v)]
=1W(q)(zv)[g(z)qW(q)(zv)[H(𝒷2𝓀2;𝕓2𝓀3)𝒵(𝓆)(𝒷2𝓀1𝒷2𝓀2)\displaystyle=\dfrac{1}{W^{(q)\prime}(z-v)}\bigg{[}g(z)-qW^{(q)}(z-v)\bigg{[}H(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})Z^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})
+W(q)(𝒷2𝓀1𝒷2𝓀2)(𝒷2𝓀2,𝒷2𝓀1;𝕓2𝓀3)+𝒢(𝓋)𝒢(𝒷2𝓀1)]]\displaystyle\quad+W^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})F(\mathpzc{b}^{*}_{2k-2},\mathpzc{b}^{*}_{2k-1};\mathbbmss{b}^{*}_{2k-3})+G(v)-G(\mathpzc{b}^{*}_{2k-1})\bigg{]}\bigg{]}
=1W(q)(zv)[g(z)qW(q)(zv)W(q)(𝒷2𝓀1𝒷2𝓀2)(𝒷2𝓀2,𝒷2𝓀1;𝕓2𝓀3)\displaystyle=\dfrac{1}{W^{(q)\prime}(z-v)}\bigg{[}g(z)-qW^{(q)}(z-v)W^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})F(\mathpzc{b}^{*}_{2k-2},\mathpzc{b}^{*}_{2k-1};\mathbbmss{b}^{*}_{2k-3})
qW(q)(zv)[H(𝒷2𝓀2;𝕓2𝓀3)𝒵(𝓆)(𝒷2𝓀1𝒷2𝓀2)+𝒢(𝓋)𝒢(𝒷2𝓀1)]]\displaystyle\quad-qW^{(q)}(z-v)\bigg{[}H(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})Z^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})+G(v)-G(\mathpzc{b}^{*}_{2k-1})\bigg{]}\bigg{]}
<1W(q)(zv)[F(𝒷2𝓀2,𝓋;𝕓2𝓀3)𝒲(𝓆)(𝓏𝒷2𝓀2)+𝓆(𝒷2𝓀2;𝕓2𝓀3)𝒲(𝓆)(𝓏𝒷2𝓀2)\displaystyle<\dfrac{1}{W^{(q)\prime}(z-v)}\bigg{[}F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})W^{(q)\prime}(z-\mathpzc{b}^{*}_{2k-2})+qH(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})W^{(q)}(z-\mathpzc{b}^{*}_{2k-2})
qW(q)(zv)W(q)(𝒷2𝓀1𝒷2𝓀2)(𝒷2𝓀2,𝓋;𝕓2𝓀3)\displaystyle\quad-qW^{(q)}(z-v)W^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})
qW(q)(zv)[H(𝒷2𝓀2;𝕓2𝓀3)𝒵(𝓆)(𝒷2𝓀1𝒷2𝓀2)+𝒢(𝓋)𝒢(𝒷2𝓀1)]]\displaystyle\quad-qW^{(q)}(z-v)\bigg{[}H(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})Z^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})+G(v)-G(\mathpzc{b}^{*}_{2k-1})\bigg{]}\bigg{]}
=1W(q)(zv)[F(𝒷2𝓀2,𝓋;𝕓2𝓀3)(𝒲(𝓆)(𝓏𝒷2𝓀2)𝓆𝒲(𝓆)(𝓏𝓋)𝒲(𝓆)(𝒷2𝓀1𝒷2𝓀2))\displaystyle=\dfrac{1}{W^{(q)\prime}(z-v)}\bigg{[}F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})\Big{(}W^{(q)\prime}(z-\mathpzc{b}^{*}_{2k-2})-qW^{(q)}(z-v)W^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})\Big{)}
+qH(𝒷2𝓀2;𝕓2𝓀3)(𝒲(𝓆)(𝓏𝒷2𝓀2)𝒲(𝓆)(𝓏𝓋)𝒵(𝓆)(𝒷2𝓀1𝒷2𝓀2))\displaystyle\quad+qH(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})\Big{(}W^{(q)}(z-\mathpzc{b}^{*}_{2k-2})-W^{(q)}(z-v)Z^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})\Big{)}
qW(q)(zv)(G(v)G(𝒷2𝓀1))]=:Λ(𝓋,𝓏;𝕓2𝓀1).\displaystyle\quad-qW^{(q)}(z-v)\Big{(}G(v)-G(\mathpzc{b}^{*}_{2k-1})\Big{)}\bigg{]}=:\Lambda(v,z;\mathbbmss{b}^{*}_{2k-1}).

Observe that if Λ(v,z;𝕓2k1)\Lambda(v,z;\mathbbmss{b}^{*}_{2k-1}) is decreasing with respect to zz, it follows that F(v,z;𝕓2k1)<σ22g(v)F(v,z;\mathbbmss{b}^{*}_{2k-1})<\dfrac{\sigma^{2}}{2}g(v) for z>vz>v, because of Λ(v,v;𝕓2k1)=σ22g(v)\Lambda(v,v;\mathbbmss{b}^{*}_{2k-1})=\dfrac{\sigma^{2}}{2}g(v). Let us verify that zΛ(z,v;𝕓2k1)<0\dfrac{\partial}{\partial z}\Lambda(z,v;\mathbbmss{b}^{*}_{2k-1})<0. Calculating the first derivative with respect to zz, we get that

F(𝒷2𝓀2,𝓋;𝕓2𝓀3)[W(q)(zv)]2((W(q)′′(z𝒷2𝓀2)𝓆𝒲(𝓆)(𝓏𝓋)𝒲(𝓆)(𝒷2𝓀1𝒷2𝓀2))𝒲(𝓆)(𝓏𝓋)\displaystyle\dfrac{F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})}{[W^{(q)\prime}(z-v)]^{2}}\Big{(}\Big{(}W^{(q)\prime\prime}(z-\mathpzc{b}^{*}_{2k-2})-qW^{(q)\prime}(z-v)W^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})\Big{)}W^{(q)\prime}(z-v)
W(q)′′(zv)(W(q)(z𝒷2𝓀2)𝓆𝒲(𝓆)(𝓏𝓋)𝒲(𝓆)(𝒷2𝓀1𝒷2𝓀2)))\displaystyle\quad-W^{(q)\prime\prime}(z-v)\Big{(}W^{(q)\prime}(z-\mathpzc{b}^{*}_{2k-2})-qW^{(q)}(z-v)W^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})\Big{)}\Big{)}
+qH(𝒷2𝓀2;𝕓2𝓀3)[W(q)(zv)]2((W(q)(z𝒷2𝓀2)𝒲(𝓆)(𝓏𝓋)𝒵(𝓆)(𝒷2𝓀1𝒷2𝓀2))𝒲(𝓆)(𝓏𝓋)\displaystyle\quad+q\dfrac{H(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})}{[W^{(q)\prime}(z-v)]^{2}}\Big{(}\Big{(}W^{(q)\prime}(z-\mathpzc{b}^{*}_{2k-2})-W^{(q)\prime}(z-v)Z^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})\Big{)}W^{(q)\prime}(z-v)
W(q)′′(zv)(W(q)(z𝒷2𝓀2)𝒲(𝓆)(𝓏𝓋)𝒵(𝓆)(𝒷2𝓀1𝒷2𝓀2)))\displaystyle\quad-W^{(q)\prime\prime}(z-v)\Big{(}W^{(q)}(z-\mathpzc{b}^{*}_{2k-2})-W^{(q)}(z-v)Z^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})\Big{)}\Big{)}
q(G(v)G(𝒷2𝓀1))[W(q)(zv)]2([W(q)(zv)]2W(q)′′(zv)W(q)(zv))<0\displaystyle\quad-q\dfrac{\Big{(}G(v)-G(\mathpzc{b}^{*}_{2k-1})\Big{)}}{[W^{(q)\prime}(z-v)]^{2}}\Big{(}[W^{(q)\prime}(z-v)]^{2}-W^{(q)\prime\prime}(z-v)W^{(q)}(z-v)\Big{)}<0
\displaystyle\Longleftrightarrow
F(𝒷2𝓀2,𝓋;𝕓2𝓀3)(𝒲(𝓆)′′(𝓏𝒷2𝓀2)𝒲(𝓆)(𝓏𝓋)𝒲𝓆′′(𝓏𝓋)𝒲(𝓆)(𝓏𝒷2𝓀2)\displaystyle F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})\Big{(}W^{(q)\prime\prime}(z-\mathpzc{b}^{*}_{2k-2})W^{(q)\prime}(z-v)-W^{q\prime\prime}(z-v)W^{(q)\prime}(z-\mathpzc{b}^{*}_{2k-2})
qW(q)(𝒷2𝓀1𝒷2𝓀2)([𝒲(𝓆)(𝓏𝓋)]2𝒲𝓆′′(𝓏𝓋)𝒲(𝓆)(𝓏𝓋)))\displaystyle\quad-qW^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})\Big{(}[W^{(q)\prime}(z-v)]^{2}-W^{q\prime\prime}(z-v)W^{(q)}(z-v)\Big{)}\Big{)}
+qH(𝒷2𝓀2;𝕓2𝓀3)(𝒲(𝓆)(𝓏𝒷2𝓀2)𝒲(𝓆)(𝓏𝓋)𝒲(𝓆)′′(𝓏𝓋)𝒲(𝓆)(𝓏𝒷2𝓀2)\displaystyle\quad+qH(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})\Big{(}W^{(q)\prime}(z-\mathpzc{b}^{*}_{2k-2})W^{(q)\prime}(z-v)-W^{(q)\prime\prime}(z-v)W^{(q)}(z-\mathpzc{b}^{*}_{2k-2})
Z(q)(𝒷2𝓀1𝒷2𝓀2)([𝒲(𝓆)(𝓏𝓋)]2𝒲(𝓆)′′(𝓏𝓋)𝒲(𝓆)(𝓏𝓋)))\displaystyle\quad-Z^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})\Big{(}[W^{(q)\prime}(z-v)]^{2}-W^{(q)\prime\prime}(z-v)W^{(q)}(z-v)\Big{)}\Big{)}
q(G(v)G(𝒷2𝓀1))([𝒲(𝓆)(𝓏𝓋)]2𝒲(𝓆)′′(𝓏𝓋)𝒲(𝓆)(𝓏𝓋))<0\displaystyle\quad-q\Big{(}G(v)-G(\mathpzc{b}^{*}_{2k-1})\Big{)}\Big{(}[W^{(q)\prime}(z-v)]^{2}-W^{(q)\prime\prime}(z-v)W^{(q)}(z-v)\Big{)}<0
\displaystyle\Longleftrightarrow
(C.11) F(𝒷2𝓀2,𝓋;𝕓2𝓀3)(𝒲(𝓆)′′(𝓏𝒷2𝓀2)𝒲(𝓆)(𝓏𝓋)𝒲𝓆′′(𝓏𝓋)𝒲(𝓆)(𝓏𝒷2𝓀2))\displaystyle F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})\Big{(}W^{(q)\prime\prime}(z-\mathpzc{b}^{*}_{2k-2})W^{(q)\prime}(z-v)-W^{q\prime\prime}(z-v)W^{(q)\prime}(z-\mathpzc{b}^{*}_{2k-2})\Big{)}
+qH(𝒷2𝓀2;𝕓2𝓀3)(𝒲(𝓆)(𝓏𝒷2𝓀2)𝒲(𝓆)(𝓏𝓋)𝒲(𝓆)′′(𝓏𝓋)𝒲(𝓆)(𝓏𝒷2𝓀2))\displaystyle\quad+qH(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})\Big{(}W^{(q)\prime}(z-\mathpzc{b}^{*}_{2k-2})W^{(q)\prime}(z-v)-W^{(q)\prime\prime}(z-v)W^{(q)}(z-\mathpzc{b}^{*}_{2k-2})\Big{)}
q(F(𝒷2𝓀2,𝓋;𝕓2𝓀3)𝒲(𝓆)(𝒷2𝓀1𝒷2𝓀2)+(𝒷2𝓀2;𝕓2𝓀3)𝒵(𝓆)(𝒷2𝓀1𝒷2𝓀2)\displaystyle\quad-q\Big{(}F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})W^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})+H(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})Z^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})
+G(v)G(𝒷2𝓀1))([𝒲(𝓆)(𝓏𝓋)]2𝒲(𝓆)′′(𝓏𝓋)𝒲(𝓆)(𝓏𝓋))<0.\displaystyle\quad+G(v)-G(\mathpzc{b}^{*}_{2k-1})\Big{)}\Big{(}[W^{(q)\prime}(z-v)]^{2}-W^{(q)\prime\prime}(z-v)W^{(q)}(z-v)\Big{)}<0.

Using (A.3)–(A.5), we see that (C.11) is equivalent to

F(𝒷2𝓀2,𝓋;𝕓2𝓀3)4𝓆σ4e(Φ(𝓆)ξ1)(𝓏𝓋)𝒲(𝓆)(𝓋𝒷2𝓀2)+𝓆(𝒷2𝓀2;𝕓2𝓀3)4e(Φ(𝓆)ξ1)(𝓏𝓋)σ4𝒵(𝓆)(𝓋𝒷2𝓀2)\displaystyle F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})\dfrac{4q}{\sigma^{4}}\operatorname{e}^{(\Phi(q)-\xi_{1})(z-v)}W^{(q)}(v-\mathpzc{b}^{*}_{2k-2})+qH(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})\frac{4\operatorname{e}^{(\Phi(q)-\xi_{1})(z-v)}}{\sigma^{4}}Z^{(q)}(v-\mathpzc{b}^{*}_{2k-2})
q(F(𝒷2𝓀2,𝓋;𝕓2𝓀3)𝒲(𝓆)(𝒷2𝓀1𝒷2𝓀2)+(𝒷2𝓀2;𝕓2𝓀3)𝒵(𝓆)(𝒷2𝓀1𝒷2𝓀2)\displaystyle\quad-q\Big{(}F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})W^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})+H(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})Z^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})
+G(v)G(𝒷2𝓀1))4σ4e(Φ(𝓆)ξ1)(𝓏𝓋)<0\displaystyle\quad+G(v)-G(\mathpzc{b}^{*}_{2k-1})\Big{)}\dfrac{4}{\sigma^{4}}\operatorname{e}^{(\Phi(q)-\xi_{1})(z-v)}<0
\displaystyle\Longleftrightarrow
F(𝒷2𝓀2,𝓋;𝕓2𝓀3)𝒲(𝓆)(𝓋𝒷2𝓀2)+(𝒷2𝓀2;𝕓2𝓀3)𝒵(𝓆)(𝓋𝒷2𝓀2)\displaystyle F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})W^{(q)}(v-\mathpzc{b}^{*}_{2k-2})+H(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})Z^{(q)}(v-\mathpzc{b}^{*}_{2k-2})
(F(𝒷2𝓀2,𝓋;𝕓2𝓀3)𝒲(𝓆)(𝒷2𝓀1𝒷2𝓀2)+(𝒷2𝓀2;𝕓2𝓀3)𝒵(𝓆)(𝒷2𝓀1𝒷2𝓀2)+𝒢(𝓋)𝒢(𝒷2𝓀1))<0\displaystyle\quad-\Big{(}F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})W^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})+H(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})Z^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})+G(v)-G(\mathpzc{b}^{*}_{2k-1})\Big{)}<0
\displaystyle\Longleftrightarrow
(C.12) F(𝒷2𝓀2,𝓋;𝕓2𝓀3)[𝒲(𝓆)(𝓋𝒷2𝓀2)𝒲(𝓆)(𝒷2𝓀1𝒷2𝓀2)]+(𝒷2𝓀2;𝕓2𝓀3)𝒵(𝓆)(𝓋𝒷2𝓀2)\displaystyle F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})[W^{(q)}(v-\mathpzc{b}^{*}_{2k-2})-W^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})]+H(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})Z^{(q)}(v-\mathpzc{b}^{*}_{2k-2})
<H(𝒷2𝓀2;𝕓2𝓀3)𝒵(𝓆)(𝒷2𝓀1𝒷2𝓀2)+𝒢(𝓋)𝒢(𝒷2𝓀1).\displaystyle\quad<H(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})Z^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})+G(v)-G(\mathpzc{b}^{*}_{2k-1}).

Since both sides of (C.12) are equal at v=𝒷2𝓀1v=\mathpzc{b}^{*}_{2k-1}, (C.12) is equivalent to prove that

ddv[F(𝒷2𝓀2,𝓋;𝕓2𝓀3)[𝒲(𝓆)(𝓋𝒷2𝓀2)𝒲(𝓆)(𝒷2𝓀1𝒷2𝓀2)]+(𝒷2𝓀2;𝕓2𝓀3)𝒵(𝓆)(𝓋𝒷2𝓀2)]\displaystyle\dfrac{\mathrm{d}}{\mathrm{d}v}[F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})[W^{(q)}(v-\mathpzc{b}^{*}_{2k-2})-W^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})]+H(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})Z^{(q)}(v-\mathpzc{b}^{*}_{2k-2})]
<ddv[H(𝒷2𝓀2;𝕓2𝓀3)𝒵(𝓆)(𝒷2𝓀1𝒷2𝓀2)+𝒢(𝓋)𝒢(𝒷2𝓀1)]\displaystyle\quad<\dfrac{\mathrm{d}}{\mathrm{d}v}[H(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})Z^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})+G(v)-G(\mathpzc{b}^{*}_{2k-1})]
\displaystyle\Longleftrightarrow
[W(q)(v𝒷2𝓀2)𝒲(𝓆)(𝒷2𝓀1𝒷2𝓀2)]dd𝓋(𝒷2𝓀2,𝓋;𝕓2𝓀3)\displaystyle[W^{(q)}(v-\mathpzc{b}^{*}_{2k-2})-W^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})]\dfrac{\mathrm{d}}{\mathrm{d}v}F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})
+F(𝒷2𝓀2,𝓋;𝕓2𝓀3)𝒲(𝓆)(𝓋𝒷2𝓀2)+𝓆(𝒷2𝓀2;𝕓2𝓀3)𝒲(𝓆)(𝓋𝒷2𝓀2)<(𝓋)\displaystyle+F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})W^{(q)\prime}(v-\mathpzc{b}^{*}_{2k-2})+qH(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})W^{(q)}(v-\mathpzc{b}^{*}_{2k-2})<g(v)
\displaystyle\Longleftrightarrow
[W(q)(v𝒷2𝓀2)𝒲(𝓆)(𝒷2𝓀1𝒷2𝓀2)]W(q)(v𝒷2𝓀2)ddvF(𝒷2𝓀2,𝓋;𝕓2𝓀3)\displaystyle\dfrac{[W^{(q)}(v-\mathpzc{b}^{*}_{2k-2})-W^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})]}{W^{(q)\prime}(v-\mathpzc{b}^{*}_{2k-2})}\dfrac{\mathrm{d}}{\mathrm{d}v}F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})
+F(𝒷2𝓀2,𝓋;𝕓2𝓀3)<(𝓋)𝓆(𝒷2𝓀2;𝕓2𝓀3)𝒲(𝓆)(𝓋𝒷2𝓀2)𝒲(𝓆)(𝓋𝒷2𝓀2)\displaystyle+F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})<\dfrac{g(v)-qH(\mathpzc{b}^{*}_{2k-2};\mathbbmss{b}^{*}_{2k-3})W^{(q)}(v-\mathpzc{b}^{*}_{2k-2})}{W^{(q)\prime}(v-\mathpzc{b}^{*}_{2k-2})}
\displaystyle\Longleftrightarrow
[W(q)(v𝒷2𝓀2)𝒲(𝓆)(𝒷2𝓀1𝒷2𝓀2)]W(q)(v𝒷2𝓀2)ddvF(𝒷2𝓀2,𝓋;𝕓2𝓀3)<0.\displaystyle\dfrac{[W^{(q)}(v-\mathpzc{b}^{*}_{2k-2})-W^{(q)}(\mathpzc{b}^{*}_{2k-1}-\mathpzc{b}^{*}_{2k-2})]}{W^{(q)\prime}(v-\mathpzc{b}^{*}_{2k-2})}\dfrac{\mathrm{d}}{\mathrm{d}v}F(\mathpzc{b}^{*}_{2k-2},v;\mathbbmss{b}^{*}_{2k-3})<0.

From here and using (4.31), it follows that (4.33) is true. Hence (𝒷2𝓀1,𝒷2𝓀1(1))𝒟2𝓀1𝒸(\mathpzc{b}^{*}_{2k-1},\mathpzc{b}^{(1)}_{2k-1})\subset\mathcal{D}_{2k-1}^{c}\neq\emptyset. Letting v𝒷2𝓀1v\downarrow\mathpzc{b}^{*}_{2k-1} in (4.20), it can be verified easily that (4.33) also holds for v=𝒷2𝓀1v=\mathpzc{b}^{*}_{2k-1} due to F(,;𝒷2𝓀3)𝒞(𝒜¯2𝓀3)F(\cdot,\cdot;\mathpzc{b}^{*}_{2k-3})\in C(\bar{\mathcal{A}}_{2k-3}). ∎

Proof of Proposition 4.6.

Given x>𝒷2𝓀1x>\mathpzc{b}^{*}_{2k-1}, let 𝕓x={𝕓2k3}{𝒷2𝓀2,𝓍}\mathbbmss{b}^{x}=\{\mathbbmss{b}^{*}_{2k-3}\}\cup\{\mathpzc{b}^{*}_{2k-2},x\}. Then, using the definition of 𝒷2𝓀1\mathpzc{b}^{*}_{2k-1} and the fact that F(𝒷2𝓀2,𝓏;𝕓2𝓀3)F(\mathpzc{b}^{*}_{2k-2},z;\mathbbmss{b}^{*}_{2k-3}) is non-decreasing in zz, we can show that V𝕓x(x)=V𝕓2k1(x)V_{\mathbbmss{b}^{x}}(x)=V_{\mathbbmss{b}^{*}_{2k-1}}(x), V𝕓x(x)=V𝕓2k1(x)V^{\prime}_{\mathbbmss{b}^{x}}(x)=V^{\prime}_{\mathbbmss{b}^{*}_{2k-1}}(x), and V𝕓x′′(x)=V𝕓2k1′′(x)V^{\prime\prime}_{\mathbbmss{b}^{x}}(x)=V^{\prime\prime}_{\mathbbmss{b}^{*}_{2k-1}}(x). Therefore, we obtain that

(q)(V𝕓2k1V𝕓x)(x)0.(\mathcal{L}-q)(V_{\mathbbmss{b}^{*}_{2k-1}}-V_{\mathbbmss{b}^{x}})(x)\leq 0.

Hence, as in Theorem 3.1 we get the result. ∎