An optimal multibarrier strategy for a singular stochastic control problem with a state-dependent reward
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 follows a spectrally negative Lévy process with a Lévy measure defined by a completely monotone density. Secondly, to verify the optimality of the -barrier strategy when 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 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 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 (see (1.1)), known as the instantaneous marginal yield function, Alvarez studied the problem in [3] for the case where is a linear diffusion process. Under certain assumptions on the function , the underlying diffusion 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 as discussed in [3] when 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 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 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 that satisfy Assumption 3.1, and with a Brownian motion with drift serving as the underlying process .
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 , 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 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 . 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 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 . For we denote by to the law of the process starting at and for simplicity, we write instead of . Accordingly, we use (resp. ) to denote expectation operator associated to the law (resp. ). Additionally, we denote the natural filtration generated by the process by satisfying the usual conditions.
A strategy is a non-decreasing, right-continuous, and -adapted process. Hence, for each strategy , the controlled process becomes
We write for the time of ruin, we say that a strategy is admissible if the controlled process is not allowed to go below zero by the action of the control , that is, if for , where and . The set of admissible strategies is denoted by . The expected reward (ER) for each admissible strategy is given by
(1.1) |
where is the discount rate, is a twice continuously differentiable function, and
(1.2) | ||||
where denotes the continuous part of . Note that the strategy generates two types of rewards: the first one is related to the continuous part , and the other to the jumps of the process .
We want to maximize the performance criterion over the set of all admissible strategies and find the value function of the problem given by
(1.3) |
By the dynamic programming principle, we identify heuristically that is associated with the following Hamilton-Jacobi-Bellman (HJB) equation
(1.4) |
where is the infinitesimal generator of the process as in (3.9) below.
To solve the optimization problem given in (1.3), we first propose a candidate strategy for being the optimal strategy (see Eqs. (3.5) and (4.27)), then verify, under some assumptions, that its ER satisfies (1.4), and finally, using a verification lemma, conclude that is the optimal strategy for the problem; see Sections 3 and 4. To this end, we will compute and express it in terms of the so-called scale functions; see Section 2.
2. Scale Functions
In this section, we will review -scale functions and provide some properties that will be used throughout this work. Let us begin with the Laplace exponent of the process , defined by
where , and the Lévy measure of , , is a measure defined on satisfying
For each , the -scale function of is a mapping , which is strictly increasing and continuous on and such that for . It is characterized by its Laplace transform, given by
where . Additionally, we define for
From Proposition 3 in [20] we know that -scale functions are harmonic for the discounted processes on , that is,
(2.1) |
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 of the process has a completely monotone density. That is, has a density whose -th derivative exists for all and satisfies
The previous assumption allows us to obtain a more explicit form of the -scale function as seen in the following result.
Lemma 2.1 ([18], Theorem 2 and Corollary 1).
For and under Assumption 2.1, the -scale function can be written as
(2.2) |
where is a non-negative, completely monotone function given by , where is a finite measure on . Moreover, is strictly log-convex (and hence convex) and is on .
Let us define the first down- and up-crossing times, respectively, by
where we follow the convention that . By Theorem 8.1 in [16], for any and ,
(2.3) |
For the next result we recall that by Lemma 2.1 the scale function is on .
Remark 2.1.
Note that since is strictly log-convex on it follows that is strictly increasing on . Additionally, as in the proof of Theorem 1 in [18] we have that has a unique minimum denoted by , hence is strictly negative on and strictly positive on . The latter implies that
On the other hand, using (2.2)
(2.4) |
3. Optimality of one-barrier strategies for spectrally negative Lévy processes
In this section, we will assume that is a spectrally negative Lévy process whose Lévy measure 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 (see (3.5)) and verify that the barrier strategy at level 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 at level , is defined by
Observe that is indeed an admissible strategy, which is continuous if , and has a unique jump of size at time zero if , where is the starting value of . We denote by .
Let be the ER for the barrier strategy at level , i.e,
(3.1) |
with . Using the strong Markov property, we provide an expression for (3.1) in terms of scale functions.
Proposition 3.1.
3.2. Selection of the optimal threshold
In order to maximize uniformly in , by looking at equation (3.2), we consider the mapping by
(3.4) |
Then, we define our choice of optimal threshold by
(3.5) |
Observe that (with as in Remark 2.1) if is non-increasing on . To guarantee that we will make the following assumption throughout the paper.
Assumption 3.1.
We assume that as .
Note that (2.2) implies that and also that , hence . Now, from (3.2) we obtain
(3.6) |
Hence, for each , is increasing and continuously differentiable on since and are positive on . Additionally, is continuously differentiable and
(3.7) |
then,
(3.8) |
Therefore, from (3.6) and (3.8), is twice continuously differentiable. We summarize these results in the next lemma.
Lemma 3.1.
For any , and is increasing on . Additionally, , with 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 under Condition (3.11) below. Let denote the infinitesimal generator of the process , acting on the space of sufficiently smooth functions i.e. (resp. ) if is of bounded variation (resp. unbounded variation). The infinitesimal generator is given by
(3.9) |
Lemma 3.2 (Verification Lemma).
Suppose that is such that is sufficiently smooth on , right-continuous at , and, for all ,
(3.10) |
Then for all and, hence, is an optimal strategy.
The previous lemma is the main tool to prove the main result of this section.
Theorem 3.1.
Remark 3.1.
-
(i)
Notice that Condition (3.11) is equivalent to the assumption that on .
- (ii)
- (iii)
Remark 3.2.
Consider that the underlying process is given by the solution to the following stochastic differential equation
where and are Lipschitz continuous functions. Defining as a combination of the fundamental solutions of the ordinary second-order differential equation , with the infinitesimal generator of (for more details, see [10], Ch. II), under the assumption that , with a countable set, is non-increasing, and that the function has a unique interior maximum at some point , where is non-increasing on , Alvarez in [3] showed that an optimal control strategy is achieved through the -barrier strategy. However, if the condition
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 .
3.3.1. Examples
In the next two examples it is straightforward to verify that meets Assumption 3.1.
-
(1)
Consider with fixed. denote
(3.14) Since and are non-increasing, we have that is also non-increasing satisfying
Therefore, there must exist a unique such that which, by (3.7) together with the fact that and are strictly positive functions in , imply that is the unique root of the mapping , and therefore . Additionally, note that Condition (3.11) is satisfied because the function is non-increasing. Hence, by Theorem 3.1 the barrier strategy is optimal.
-
(2)
Let and , then is non-increasing and therefore has a unique maximum at if
which is true only if by Remark 2.1 and (2.4). On the other hand, if , it follows that for all , which implies that . Hence, Condition (3.11) is satisfied, and therefore the barrier strategy 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 is a standard Brownian motion, i.e., , 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 be defined as
(3.15) |
and let , and . Notice that this function also satisfies Assumption 3.1. We obtain numerically that attains its global maximum at . Condition (3.11) is not satisfied but for , see Figure 1. Then, by Lemma 3.2, the barrier strategy at the level is optimal. Note that the function taking the value does not pose an issue. For the process , is regular for , meaning that when reaches , ruin occurs instantaneously.

Now, if we choose , we get numerically that attains its global maximum at . Condition (3.11) is not satisfied and there exists such that , see Figure 2. In the next section, we will show that for this case the optimal strategy corresponds to a multibarrier strategy.

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 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 . 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 -barriers strategy as our candidate for optimality among admissible strategies. Due to the complexity that arises from the jumps of the process when working with multibarrier strategies, in the remainder of the paper, we will assume that is a Brownian motion with drift, that is, . Proofs of the results of this section are provided in Appendix C.
We denote the -barriers by where , and describe the -barrier strategy as follows: If the process lies above , push the process to the level ; if it lies between , with , do nothing, and finally if it lies between , with , push the process to . We denote the -barrier strategy by and the controlled process by . Formally, if , we define the controlled process under as follows: Assume that for , where we take and .
-
(1)
Let and , hence is a process reflected at that starts at . Define , then for .
-
(2)
For , define , so that , and let and . Hence is a process reflected at that starts at . Define and for .
4.1. Computation of the ER
Consider the ER associated with the -barrier strategy, given by
(4.1) |
where . The next result gives the explicit estimations of in terms of the -scale functions that are associated with the process .
Proposition 4.1.
4.2. Selection of an optimal multibarrier strategy
To select a as a candidate for being an optimal barrier, let us define the auxiliary function
(4.5) |
where is a non-negative increasing sequence and , with and . Since , then , where is the closure set of , and it is as smooth as on .
The following are two very useful properties of the function defined in (4.5):
- (1)
- (2)
Next, we present a method to select a candidate for being an optimal multibarrier. Let us first start with the case .
Existence of
First, define as in (3.5). Now, note that
(4.11) |
otherwise, by Remark 3.1, is an optimal barrier. By definition of , there exists such that for , that is
(4.12) |
and
(4.13) |
Proposition 4.2.
-
(i)
If , then
(4.14) -
(ii)
For each ,
(4.15)
The previous result and (4.11) imply that there exists such that
(4.16) |
for some . Hence, the set
(4.17) |
is not empty. Let , which is attained due to the smoothness of .
Lemma 4.1.
There exists , such that
(4.18) |
The lemma implies that the global maximum of , if it exists, is attained at some point . Note that under Assumption 3.1 the global maximum of the functions exists for all . Indeed, this assumption implies that for all we have that and therefore the global maximum is attained. Hence, we have that
(4.19) |
is well defined. Take the set as follows
We see that is the collection of points where the map attains its global maximum at . We are now ready to define
Proposition 4.3.
For each ,
(4.20) |
Hence , with .
Proposition 4.3 shows that . Also note that if then the trajectory , given by (4.19), has a discontinuity at . Before proving one of the main theorems of this section, we will illustrate the previous results with a numerical example. Let us consider as in (3.15) and the surplus process given by
with the discount rate . We already now that , see Figure 2. To find the next barriers we need to calculate and analyze . Figure 3 shows the value of for and observe that has a discontinuity at , the minimum such that . Also, .


Theorem 4.1.
Assume that there exists such that
(4.21) |
Then
and
(4.22) |
Remark 4.1.
Continuing with the numerical example, Figure 4 shows in blue the function , which satisfies , and in red the function , which is negative for . Figure 5 shows the value function and the multibarrier .


Existence of
Now that we have shown the existence of the first three barriers, we can formulate an algorithm that will provide the next two barriers with considering that the previous , with , are obtained. After that, we will prove the correctness of the algorithm under conditions similar to (4.21).
Algorithm 1.
-
(1)
Choose as in (3.5).
-
(2)
If
(4.23) let and stop. Otherwise, let .
-
(3)
Let be the set of points such that for some ,
(4.24) Define and
(4.25) where
(4.26) Let
(4.27) -
(4)
If
(4.28) choose the barriers , and stop. Otherwise, let and return to (3).
To verify the correctness of the algorithm, we assume that, for fixed, the set of barriers provided by the algorithm, are well defined, which means that
(4.29) |
We will show that and are not empty sets, and that , and are well defined. For that purpose, let us also suppose that
(4.30) |
otherwise Algorithm 1 would had stop. Now, by definition of in step (3) of the algorithm, there exists such that
(4.31) |
with equality at . This condition is equivalent to with as in (4.8). Moreover, we can choose such that for each
(4.32) |
The following result is analogous to Proposition 4.2.
Proposition 4.4.
Let be given as before. Then,
Hence, and . Additionally, .
A few straightforward changes in the proof of Lemma 4.1 show that there exists small enough, such that for
and therefore . We also have that from the next result.
Proposition 4.5.
For each ,
(4.33) |
Hence , with .
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 such that
(4.34) |
Then, and given by the Algorithm are well defined and satisfy
Moreover,
(4.35) |
Again, by Remark 4.1, condition (4.34) is satisfied if a.e. Note that if the function is non-increasing for , or equivalently, if for , 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 is non-increasing for , then the function for all .
4.3. Verification of optimality
We now verify that the sequence of barriers given by Algorithm 1 is indeed optimal. For this, note that the set of barriers satisfies the following properties:
- (i)
- (ii)
Also, the value function satisfies that:
- (i)
-
(ii)
By (4.4), for .
-
(iii)
By the martingale properties of the scale functions (2.1) we also have that for .
-
(iv)
Since for and , then
-
(v)
By construction, Algorithm 1 guarantees that for both , with , and , we have that
We conclude that satisfies the HJB equation (3.10). Now, by Lemma 3.2, we obtain the following theorem.
Theorem 4.3.
Let be the value function given in (1.3). Then for all and the -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 , with , and as a Brownian motion, the -scale functions and are given by the expressions seen in (2.5). Then, using (2.1) we see that and satisfy the following identities
(A.1) |
(A.2) |
Lemma A.1.
For , the following hold
(A.3) | |||
(A.4) | |||
(A.5) |
Appendix B Proofs of one-barrier strategies
Proof of Proposition 3.1.
Observe that if , . Then, by the strong Markov property, we get that
(B.1) | ||||
Let us consider , then by (2.3) and the strong Markov property,
(B.2) |
We now compute . Let us denote and for , where the process has the same law than . We also consider the first entrance of the process to the set given by , . By spatial homogeneity of the Lévy process and following Proposition 1 in [6] we have that the ensemble has the same law as . Therefore
(B.3) |
By the absence of positive jumps of , it is known that the supremum is a local time for for the process . Now, let us denote by the set of excursions away from zero of finite length
where is the set of all càdlàg functions on . Additionally, we denote by the set of excursions with infinite length. We will work with the excursion process of , with values on , given by
where denotes the right-inverse of the local time of at . Hence, noting that the process is continuous and non-decreasing, and following the proof of Theorem 1 in [5] we have
where for . By the lack of negative jumps of the process , we have that , hence using (B.3)
(B.4) |
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 , it follows that for all . Now, let us write and we will show that for all . First, assume that and define, for any , and denote by the following set of admissible strategies:
Then, by the proof of Lemma 4 in [15] it is enough to show that for any . Hence, fix and let . Noting that is a semi-martingale and that is sufficiently smooth on , we can use the change of variable/Itô’s formula on , with (see [21, Thm. 33, pp. 81]), to obtain
(B.5) |
where is a local martingale with and is given by
By assumption, we know that , and on . Hence, taking the expectations in (B) and using (1.2), yields that
Letting in the previous inequality, using monotone convergence, the fact that -a. s. and that , we have that
Therefore for all . Finally, using the fact that is non-decreasing together with the right continuity of at zero, we obtain that . ∎
Proof of Theorem 3.1.
We have by construction that , therefore by Lemma 3.2 we only need to show that satisfies (3.10) for . By (2.1)
On the other hand, (3.5) gives for , this implies
Hence, satisfies (3.10) on . Meanwhile, using (3.2) we note that
(B.6) |
For any , let us consider the ER associated with the barrier strategy at the level , , which is given in (3.2). We recall that, by Lemma 3.1, and . We aim to prove that
(B.7) |
To this end, we check the following:
-
(i)
Condition (3.11) implies that .
- (ii)
- (iii)
-
(iv)
By (3.6) we have that .
-
(v)
Using (ii) we obtain that for . Additionally, by (iii) we have for .
Therefore using (i)-(v) and the fact that
we obtain (B.7). So, proceeding like in the proof of Theorem 2 in [17] we obtain that
(B.8) |
Appendix C Proofs of multibarrier strategies
Proof of Proposition 4.1.
For , we obtain, by the strong Markov property, that
and proceeding like in the proof of Proposition 3.1 we have
On the other hand, for
Now, let . By the Strong Markov property and (2.3), we have that
(C.1) | ||||
Note that
where ,
and . Hence, by the spatial homogeneity of Brownian motion, we have that
(C.2) |
Again, proceeding as in the proof of Proposition 3.1
(C.3) |
On the other hand, by identity (3.10) in [6], we obtain
(C.4) |
Therefore, (C.1)–(C.4) give . Finally, we obtain the result by induction and similar arguments as above. ∎
Proof of Proposition 4.2.
Proof of Lemma 4.1.
Using (4.7), (4.18) is equivalent to and using (A.1), this is equivalent to
Since due to (2.6), (4.6), (4.9) and (4.16), it is enough to verify that in order to prove (4.18) by the -continuity of this function. Then,
and taking , by (4.9) and (4.16) we obtain that
Now, from (4.16) we know that there exists such that if , then
We also know that . Hence, by taking the derivative at we must have that
therefore . ∎
Proof of Proposition 4.3.
By (4.13), it follows that for ,
From here and (3.3), we have that
Observe that if is decreasing for , it follows that for , because of . Let us verify that . Calculating the first derivative with respect to , we get that
(C.6) | ||||
Using (A.3)–(A.4), we see that (C) is equivalent to
(C.7) |
Since both sides of (C) are zero at , (C) is equivalent to prove that
From here and (4.12), it follows that (4.20) is true. Letting it can be easily verified that (4.20) holds for , since . ∎
Proof of Theorem 4.1.
Condition (4.21) immediately implies that , so it only remains to show that to prove the strict inequalities. Suppose that , that is,
(C.8) |
Note that (C.8) is equivalent to
(C.9) |
where
Also note that for any , . Additionally, differentiating with respect to , letting , and using (2.6), (2.7), (3.7) and (4.9), it follows that
Now, using (4.21), it yields for . From here, the smoothness of and (C.9), it implies that there exists such that , for and , which is equivalent to , for and . Thus, cannot be the infimum of , which is a contradiction. Therefore, . Additionally, using again (4.21) and arguing similarly as in the proof of Lemma 4.1, it can be proven that is non-increasing locally at , and concluding that (4.22) follows due to the continuity of . ∎
Proof of Proposition 4.4.
Proof of Proposition 4.5.
By (4.32), it follows that for ,
and . From here, (4.3), (4.4) and (4.5), we have that
Observe that if is decreasing with respect to , it follows that for , because of . Let us verify that . Calculating the first derivative with respect to , we get that
(C.11) | |||
Using (A.3)–(A.5), we see that (C.11) is equivalent to
(C.12) | |||
Since both sides of (C.12) are equal at , (C.12) is equivalent to prove that
From here and using (4.31), it follows that (4.33) is true. Hence . Letting in (4.20), it can be verified easily that (4.33) also holds for due to . ∎