Skip to main content
Log in

Optimal Strategies in a Production Inventory Control Model

  • Published:
Methodology and Computing in Applied Probability Aims and scope Submit manuscript

Abstract

We consider a production-inventory control model with finite capacity and two different production rates, assuming that the cumulative process of customer demand is given by a compound Poisson process. It is possible at any time to switch over from the different production rates but it is mandatory to switch-off when the inventory process reaches the storage maximum capacity. We consider holding, production, shortage penalty and switching costs. This model was introduced by Doshi, Van Der Duyn Schouten and Talman in 1978. In their paper they found a formula for the long-run average expected cost per unit time as a function of two critical levels, in this paper we consider expected discounted cumulative costs instead. We seek to minimize this discounted cost over all admissible switching strategies. We show that the optimal cost functions for the different production rates satisfy the corresponding Hamilton-Jacobi-Bellman system of equations in a viscosity sense and prove a verification theorem. The way in which the optimal cost functions solve the different variational inequalities gives the switching regions of the optimal strategy, hence it is stationary in the sense that depends only on the current production rate and inventory level. We define the notion of finite band strategies and derive, using scale functions, the formulas for the different costs of the band strategies with one or two bands. We also show that there are examples where the switching strategy with two critical levels is not optimal.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15

Similar content being viewed by others

References

  • Avram F, Grahoavac D, Vardar-Acre C (2019) The W, Z scale functons kit for first passage problems of spectrally negative Lévy processes, and applications to control problems. https://www.esaim-ps.org/articles/ps/pdf/forth/ps170105.pdf

  • Azcue P, Muler N (2015) Optimal dividend payment and regime switching in a Compound Poisson risk model. SIAM J Control Optim 53(5):3270–3298

    Article  MathSciNet  MATH  Google Scholar 

  • Barron Y, Perry D, Stadje W (2014) A jump-fluid production-inventory model with a double band control. Probab Eng Inf Sci 28:313–333

    Article  MathSciNet  MATH  Google Scholar 

  • Barron Y, Perry D, Stadje W (2016) A make-to-stock production/inventory model with MAP arrivals and phase-type demands. Ann Oper Res 241:373–409

    Article  MathSciNet  MATH  Google Scholar 

  • Bayraktar E, Egami M (2010) On the optimal switching problem for one dimensional-diffusions. Math Oper Res 35(1):140–159

    Article  MathSciNet  MATH  Google Scholar 

  • Brekke K, Oksendal B (1994) Optimal switching in an economic activity under uncertainty, SIAM. J Cont Optim 32:1021–1036

    Article  MATH  Google Scholar 

  • Chang J, Lu H, Shi J (2019) Stockout risk of production-inventory systems with compound Poisson demands. Omega 83:181–198

    Article  Google Scholar 

  • Crandall MG, Lions PL (1983) Viscosity solution of Hamilton-Jacobi equations. Trans Amer Math Soc 277:1–42

    Article  MathSciNet  MATH  Google Scholar 

  • De Kok AG, Tijms HC, Van Der Schouten FA (1984) Approximations for the single-product production-inventory problem with compound Poisson demand and service-level constraints. Adv Appl Probab 16(2):378–401

    Article  MathSciNet  MATH  Google Scholar 

  • De Kok AG (1985) Approximations for a lost-sales production-inventory control model with service level constraints. Manage Sci 31(6):729–737

    Article  MATH  Google Scholar 

  • De Kok AG (1987) Approximations for operating characteristics in a production-inventory model with variable production rate. Eur J Oper Res 29:286–297

    Article  MathSciNet  MATH  Google Scholar 

  • Doshi BT, Van Der Duyn Schouten FF, Talman AJJ (1978) A production-inventory control model with a mixture of backorders and lost sales. Manage Sci 24(10):1078–1086

    Article  MATH  Google Scholar 

  • Dufour Fr, Piunovskiy AB (2015) Impulsive control for continuous-time Markov decision processes. Adv Appl Probab 47:106–127

    Article  MathSciNet  MATH  Google Scholar 

  • Fleming WH, Soner HM (2006) Controlled Markov Processes and Viscosity Solutions, 2nd edn. Springer

    MATH  Google Scholar 

  • Fleming WH, Sethi SP, Soner HM (1987) An optimal stochastic production planning problem with randomly fluctuating demand. SIAM J Control Optim 25(6):1494–1502

    Article  MathSciNet  MATH  Google Scholar 

  • Gavish B, Graves SC (1980) A one product production-inventory problem under continuous review policy. Oper Res 28(5):1222–1235

    Article  MathSciNet  MATH  Google Scholar 

  • Graves SC, Keilson J (1981) The compensation method applied to one-product production-inventory problems. Math Oper Res 6(2):246–262

    Article  MathSciNet  MATH  Google Scholar 

  • Guo X, Hernández-Lerma O (2009) Continuous-time Markov decision processes: theory and applications. Springer, Heidelberg

    Book  MATH  Google Scholar 

  • Hamadène S, Jeanblanc M (2007) On the Starting and Stopping Problem: Application in Reversible Investments. Math Oper Res 32(1):182–192

    Article  MathSciNet  MATH  Google Scholar 

  • Kella O, Whitt W (1992) Useful martingales for stochastic storage processes with Lévy input. J Appl Probab 29:396–403

    Article  MathSciNet  MATH  Google Scholar 

  • Kyprianou AE (2014) Fluctuations of Lévy Processes with Applications. Springer, Berlin Heidelberg

    Book  MATH  Google Scholar 

  • Loeffen R (2018) On obtaining simple identities for overshoots of spectrally negative Levy processes. arXiv:1410.5341

  • Pham H, Ly Vath V (2007) Explicit solution to an optimal switching problem in the two-regime case, SIAM. J Cont Optim 46:395–426

    Article  MATH  Google Scholar 

  • Pham H, Ly Vath V, Zhou XY (2009) Optimal Switching over Multiple Regimes. SIAM J Control Optim 48:2217–2253

    Article  MathSciNet  MATH  Google Scholar 

  • Piunovskiy AB, Zhang Y (2014) Discounted continuous-time Markov decision processes with unbounded rates and randomized history-dependent policies: the dynamic programming approach. J Oper Res 12:49–75

    Article  MathSciNet  MATH  Google Scholar 

  • Sethi SP, Zhang Q (1994) Hierarchical Decision Making in Stochastic Manufacturing Systems, Systems & Control: Foundations & Applications. Birkhäuser, Boston, Mass, USA

    Book  MATH  Google Scholar 

  • Shi J, Katehakis MH, Melamed B, Xia Y (2014) Production-inventory systems with lost sales and compound Poisson demands. Oper Res 62(5):1048–1063

    Article  MathSciNet  MATH  Google Scholar 

  • Shi J (2016) Optimal continuous production-inventory systems subject to stockout risk. Ann Oper Res. https://doi.org/10.1007/s10479-016-2339-5

    Article  MATH  Google Scholar 

  • Varadhan SRS (2007) Stochastic Processes. Courant Lecture Notes 16, AMS

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Pablo Azcue.

Ethics declarations

Conflicts of Interest

No funds, grants, or other support was received. Prof. Esther Frostig is a member of the Editorial Board of MCAP and has no relevant financial interests to disclose. The other authors have no relevant financial or non-financial interests to disclose.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix

Appendix

1.1 Proof of Proposition 3.2

We call \(S_{i}\) the positive upper bounds of the functions \(V_{i}\) for \(i=1,2.\) Given initial inventory level \(x\in [l,b)\) and initial phase \(i=1,2,\) take \(\delta \in (0,b-x]\) and consider an admissible strategy \(\pi _{x+\delta }\in \Pi _{x+\delta ,i}\) such that \(V_{i}^{\pi _{x+\delta }}(x+\delta )\le V_{i}(x+\delta )+\varepsilon ,\) where \(0<\varepsilon <\delta\). Let us now define the admissible strategy \(\pi _{x}\in \Pi _{x,i}\) as follows: stay in phase i until the controlled inventory level \(X_{t}^{\pi _{x}}\) reaches \(x+\delta\) and then follow \(\pi _{x+\delta }\in \Pi _{x+\delta ,i}\). Then, from (3.1) and Proposition 3.1, we get

$$\begin{aligned} \begin{array}{lll} V_{i}(x) &{} \le &{} V_{i}^{\pi _{x}}(x)\\ &{} \le &{} \int _{0}^{\frac{\delta }{\sigma _{i}}}e^{-qt}h_{i}(x+\sigma _{i}t)dt+\mathbb {P}[\tau _{1}^{i}>\frac{\delta }{\sigma _{i}}]e^{-q\frac{\delta }{\sigma _{i}}}V_{i}^{\pi _{x+\delta }}(x+\delta )\\ &{} &{} +\mathbb {P}\left[ \tau _{1}^{i}\le \frac{\delta }{\sigma _{i}}\right] S_{i}\\ &{} \le &{} \overline{h}\frac{\delta }{\sigma _{i}}+e^{-\left( \lambda _{i}+q\right) \frac{\delta }{\sigma _{i}}}\left( V_{i}(x+\delta )+\varepsilon \right) +(1-e^{-\lambda _{i}\frac{\delta }{\sigma _{i}}})S_{i}. \end{array} \end{aligned}$$

Hence, we have

$$\begin{aligned} \begin{array}{lll} V_{i}(x)-V_{i}(x+\delta ) &{} \le &{} \overline{h}\frac{\delta }{\sigma _{i}}+e^{-\left( \lambda _{i}+q\right) \frac{\delta }{\sigma _{i}}}\left( V_{i}(x+\delta )+\varepsilon \right) -V_{i}(x+\delta )+(1-e^{-\lambda _{i}\frac{\delta }{\sigma _{i}}})S_{i}\\ &{} \le &{} \overline{h}\frac{\delta }{\sigma _{i}}+\varepsilon +\lambda _{i}\frac{\delta }{\sigma _{i}}S_{i}. \end{array} \end{aligned}$$

So, taking

$$\begin{aligned} m_{i}^{1}:=\frac{\overline{h}}{\sigma _{i}}+1+\frac{\lambda _{i}}{\sigma _{i} }S_{i}, \end{aligned}$$

we obtain

$$\begin{aligned} V_{i}(x)-V_{i}(x+\delta )\le m_{i}^{1}\delta . \end{aligned}$$
(8.1)

Let us prove now that there exists \(m_{i}^{2}>0\) such that,

$$\begin{aligned} V_{i}(x+\delta )-V_{i}(x)\le m_{i}^{2}\delta . \end{aligned}$$
(8.2)

We start showing that there exists m such that,

$$\begin{aligned} V_{i}(y)-V_{i}(l)\le m\delta \end{aligned}$$
(8.3)

for all \(y\in [l,l+\delta ]\). Given \(\varepsilon >0\) and an initial inventory level l, consider the strategy \(\pi _{l}\in \Pi _{l,i}\) for \(i=1,2\) such that \(V_{i}^{\pi _{l}}(l)\le V_{i}(l)+\varepsilon\) and call \(X_{t}^{\pi _{l}}\) the associated process with initial inventory level l. Take also a strategy \(\pi _{b}\in \Pi _{b,0}\) such that \(V_{0}^{\pi _{b}}(b)\le V_{0}(b)+\varepsilon\).

Let us define the admissible strategy \(\pi _{y}\in \Pi _{y,i}\) for initial inventory level \(y\in [l,l+\delta ]\) as:

  • For \(0\le t\le T,\) follow \(\pi _{l}\) (and so the associated controlled processes \(X_{t}^{\pi _{y}}=X_{t}^{\pi _{l}}+(y-l)\) for \(t<T\)), where

    $$\begin{aligned} T:=\min \{t:X_{t}^{\pi _{y}}=b\text { or }\overset{\vee }{X}_{t}^{\pi _{y}}-(y-l)=\overset{\vee }{X}_{t}^{\pi _{l}}<l\}. \end{aligned}$$
  • If \(X_{T}^{\pi _{y}}=b\), follow \(\pi _{b}\) for \(t\ge T\).

  • If \(\overset{\vee }{X}_{T}^{\pi _{y}}<l\) (and so \(X_{T}^{\pi _{y}}=X_{T}^{\pi _{l}}=l\)), follow \(\pi _{l}\) for \(t\ge T\).

  • If \(l\le X_{T}^{\pi _{y}}<y\) (and so \(X_{T}^{\pi _{l}}=l\) and \(X_{T}^{\pi _{y}}=\overset{\vee }{X}_{T}^{\pi _{y}}\)), also follow the strategy \(\pi _{l}\) for \(t\ge T\).

Given any stopping time \(\tau\), let us define \(\widehat{V}_{i}^{\pi _{y}}(y,\tau )\) as the expected discounted cost of the strategy before \(\tau\) and \(\widetilde{V}_{i}^{\pi _{y}}(y,\tau )\) as the expected discounted cost of the strategy after \(\tau\). Thus,

$$\begin{aligned} \begin{array}{l} V_{i}(y)-V_{i}(l)-\varepsilon \\ \begin{array}{ll} \le &{} V_{i}^{\pi _{y}}(y)-V_{i}^{\pi _{l}}(l)\\ \le &{} \mathbb {E}\left[ \int _{0}^{T}e^{-qt}\left( h_{\mathcal {J}_{t}}(X_{t}^{\pi _{l}}+(y-l))-h_{\mathcal {J}_{t}}(X_{t}^{\pi _{l}})\right) dt\right] +\\ &{} +\mathbb {E}\left[ 1_{\left\{ X_{T}^{\pi _{y}}=b\right\} }\left( e^{-qT}\left( K_{\mathcal {J}_{T^{-}},0}+V_{0}(b)+\varepsilon \right) -\widetilde{V}_{i}^{\pi _{l}}(l,T)\right) \right] \\ &{} +\mathbb {E}\left[ 1_{\left\{ \overset{\vee }{X}_{T}^{\pi _{y}}<l\right\} }e^{-qT}\left( V_{\mathcal {J}_{T}}(l)+\varepsilon +p_{\mathcal {J}_{T^{-}}}(l-\overset{\vee }{X}_{T}^{\pi _{y}})-(V_{\mathcal {J}_{T}}(l)+p_{\mathcal {J}_{T^{-}}}(l-\overset{\vee }{X}_{T}^{\pi _{y}}+y-l))\right) \right] \\ &{} +\mathbb {E}\left[ 1_{\left\{ l\le X_{T}^{\pi _{y}}<y\right\} }e^{-qT}\left( V_{\mathcal {J}_{T}}^{\pi _{X_{T}^{\pi _{y}}}}(X_{T}^{\pi _{y}})-V_{\mathcal {J}_{T}}^{\pi _{l}}(l)+2\varepsilon \right) \right] . \end{array} \end{array} \end{aligned}$$

Let \(n_{D}\) be the sum of the numbers of discontinuities of \(h_{1}\) and \(h_{2}\). Note that between two customer demands, the inventory level \(X_{t}^{\pi _{y}}\) goes through at most \(n_{D}\) points of discontinuities of \(h_{\mathcal {J}_{t}}\). Hence, calling \(\tau _{0}=0\) and \(\overline{\lambda }=\max _{i=0,1,2}\lambda _{i}\), we have

$$\begin{aligned} \mathbb {E}\left[ \int _{0}^{T}e^{-qt}\left( h_{\mathcal {J}t}(X_{t}^{\pi _{l}}+(y-l))-h_{\mathcal {J}t}(X_{t}^{\pi _{l}})\right) dt\right] \le \left( \frac{m_{h}}{q}+n_{D}\frac{\overline{h}}{\sigma _{2}}\left( 1+\frac{\overline{\lambda }}{q}\right) \right) \delta . \end{aligned}$$
(8.4)

Let us call \(\widetilde{T}:=\inf \left\{ t:X_{t}^{\pi _{l}}=b\right\}\) and \(\widetilde{\tau }\) the time of the first customer demand after T;  we have that \(\mathbb {P}\left[ \widetilde{T}>\widetilde{\tau }\right] \le 1-e^{-\frac{\delta }{\sigma _{2}}\overline{\lambda }}\) and so

$$\begin{aligned} \begin{array}{l} \mathbb {E}\left[ 1_{\left\{ X_{T}^{\pi _{y}}=b\right\} }\left( e^{-qT}\left( K_{\mathcal {J}_{T^{-}},0}+V_{0}(b)\right) -\widetilde{V}_{i}^{\pi _{l}}(l,T)\right) \right] \\ \begin{array}{ll} \le &{} \left( 1-e^{-\overline{\lambda }\frac{\delta }{\sigma _{2}}}\right) \left( V_{0}(b)+\left( K_{1,0}\vee K_{2,0}\right) \right) \\ &{} +\mathbb {E}\left[ 1_{\left\{ X_{T}^{\pi _{y}}=b,\widetilde{T}<\widetilde{\tau }\right\} }e^{-qT}\left( K_{\mathcal {J}_{T^{-}},0} +V_{0}(b)\right) -\widetilde{V}_{i}^{\pi _{l}}(l,T)\right] \\ \le &{} \overline{\lambda }\frac{\delta }{\sigma _{2}}\left( V_{0}(b)+\max \left\{ K_{1,0},K_{2,0}\right\} \right) \\ &{} +\mathbb {E}\left[ 1_{\left\{ X_{T}^{\pi _{y}}=b,\widetilde{T}<\widetilde{\tau }\right\} }e^{-qT}\left( K_{\mathcal {J}_{T^{-}},0}+V_{0}(b)\right) -\widetilde{V}_{i}^{\pi _{l}}(l,T)\right] . \end{array} \end{array} \end{aligned}$$
(8.5)

Let \(\Delta\) be the length of time after T in which the process \(X_{t}^{\pi _{l}}\) reaches b in the event of no arrivals of demands. In this case, we have

$$\begin{aligned} X_{T+\Delta }^{\pi _{l}}=b-(y-l)+\int _{T}^{T+\Delta }e^{-qs}\sigma _{\mathcal {J} _{T}}ds=b \end{aligned}$$

and so \(\frac{\delta }{\sigma _{1}}\le \Delta \le \frac{\delta }{\sigma _{2}}\). Hence, from (2.3),

$$\begin{aligned} \begin{array}{l} \mathbb {E}\left[ 1_{\left\{ X_{T}^{\pi _{y}}=b\right\} }1_{\left\{ \widetilde{T}<\widetilde{\tau }\right\} }\widetilde{V}_{i}^{\pi _{l} }(l,T)\right] \\ \begin{array}{ll} \ge &{} \mathbb {P}\left[ \text {no demands in }t\in \left[ T,T+\frac{\delta }{\sigma _{2}}\right] \right] \text { }\mathbb {E}\left[ e^{-q\left( T+\Delta \right) }\left( K_{\mathcal {J}_{T^{-}},0}+V_{0}(b)\right) \right] \\ \ge &{} e^{-\left( q+\overline{\lambda }\right) \frac{\delta }{\sigma _{2}} }\mathbb {E}\left[ e^{-qT}\left( K_{\mathcal {J}_{T^{-}},0}+V_{0}(b)\right) \right] . \end{array} \end{array} \end{aligned}$$

Therefore,

$$\begin{aligned} \begin{array}{l}\mathbb {E}\left[ 1_{\left\{ X_{T}^{\pi _{y}}=b\right\} }1_{\left\{ \widetilde{T}<\tau _{1}\right\} }e^{-qT}\left( K_{\mathcal {J}_{T^{-}},0}+V_{0}(b)\right) -\widetilde{V}_{i}^{\pi _{l}}(l,T)\right] \\ \begin{array}{ll}\le &{} \left( 1-e^{-\left( q+\overline{\lambda }\right) \frac{\delta }{\sigma _{2}}}\right) \left( (K_{1,0}\vee K_{2,0})+V_{0}(b)\right) \\ \le &{} \frac{q+\overline{\lambda }}{\sigma _{2}}\left( (K_{1,0}\vee K_{2,0})+V_{0}(b)\right) \delta . \end{array} \end{array} \end{aligned}$$
(8.6)

Since the penalty functions \(p_{i}\) are non-decreasing, we also have,

$$\begin{aligned} \mathbb {E}\left[ 1_{\left\{ \overset{\vee }{X}_{T}^{\pi _{y}}<l\right\} }e^{-qT}\left( p_{\mathcal {J}_{T^{-}}}(l-\overset{\vee }{X}_{T}^{\pi _{y} })-p_{\mathcal {J}_{T^{-}}}(l-\overset{\vee }{X}_{T}^{\pi _{y}}+y-l)\right) \right] \le 0. \end{aligned}$$
(8.7)

Finally, since the event \(l\le X_{T}^{\pi _{y}}<y\) coincides with the arrival of a customer demand,

$$\begin{aligned} \begin{array}{l} \mathbb {E}\left[ 1_{\left\{ l\le X_{T}^{\pi _{y}}<y\right\} }e^{-qT}\left( V_{\mathcal {J}_{T}}^{\pi _{X_{T}^{\pi _{y}}}}(X_{T}^{\pi _{y}})-V_{\mathcal {J}_{T}}(l)\right) \right] \\ \begin{array}{ll} = &{} \mathbb {E}\left[ 1_{\left\{ l\le X_{T}^{\pi _{y}}<y\right\} }1_{\left\{ T=\tau _{n}^{\mathcal {J}_{T^{-}}}\text { for some }n\right\} }e^{-qT}\left( V_{\mathcal {J}_{T}}^{\pi _{X_{T}^{\pi _{y}}}}(X_{T}^{\pi _{y}})-V_{\mathcal {J}_{T}}(l)\right) \right] \\ \le &{} \mathbb {E}\left[ e^{-q\tau _{1}^{\mathcal {J}_{T}}}\max _{z\in [l,y]}\left( V_{\mathcal {J}_{T}}^{\pi _{z}}(z)-V_{\mathcal {J}_{T}}(l)\right) \right] \\ \le &{} \frac{\overline{\lambda }}{q+\overline{\lambda }}\max _{z\in [l,l+\delta ]}\left( V_{\mathcal {J}_{T}}^{\pi _{z}}(z)-V_{\mathcal {J}_{T} }(l)\right) . \end{array} \end{array} \end{aligned}$$
(8.8)

Hence, from (8.5), (8.6), (8.7) and (8.8), there exists \(\overline{m}\) large enough such that

$$\begin{aligned} \frac{q}{q+\overline{\lambda }}\max _{z\in [l,l+\delta ]}\left( V_{i}^{\pi _{y}}(z)-V_{i}^{\pi _{l}}(l)\right) \le \overline{m}\delta . \end{aligned}$$

So, we obtain (8.3) with \(m=\overline{m}\left( q+\lambda \right) /q\). The argument to show (8.2) is analogous.

1.2 Proof of Proposition 4.3

Consider \(\pi \in \Pi _{x,j}\). Let us extend \(\overline{u}_{1}\) and \(\overline{u}_{2}\) as \(\overline{u}_{i}(x)=\overline{u}_{i}(l)\) and \(\overline{u}_{0}(x)=\overline{u}_{0}(l)\) for \(x<l\). Consider the controlled risk process \(X_{t}^{\pi }\) starting at x and the function \(\mathcal {J}_{t}\) defined in (2.5). Since \(\overline{u}_{i}\) is Lipschitz for \(i=1,2\), we obtain that the function \(t\rightarrow e^{-qt}~\overline{u}_{\mathcal {J}_{t}}(X_{t}^{\pi })\) is absolutely continuous in between the stopping times \(\left\{ 0\right\} \cup \left\{ \tau _{n}:n\ge 1\right\} \cup \left\{ T_{k}:k\ge 1\right\}\). So, taking

$$\begin{aligned} m_{t}:=\max \{k:T_{k}\le t\}\text {,} \end{aligned}$$

we have

$$\begin{aligned} \begin{array}{l} \overline{u}_{\mathcal {J}_{t}}(X_{t}^{\pi })e^{-qt}-\overline{u}_{j}(x)\\ \begin{array}{ll} = &{} \sum _{k=0}^{m_{t}-1}\left( \overline{u}_{_{J_{k+1}}}(X_{T_{k+1}}^{\pi })e^{-qT_{k+1}}-\overline{u}_{_{J_{k}}}(X_{T_{k}}^{\pi })e^{-qT_{k}}\right) +(\overline{u}_{J_{m_{t}}}(X_{t}^{\pi })e^{-qt}-\overline{u}_{_{J_{m_{t}}}}(X_{T_{m_{t}}}^{\pi })e^{-qT_{m_{t}}}). \end{array} \end{array} \end{aligned}$$
(8.9)

Let us define

$$\begin{aligned} \begin{array}{lll} M^{i}(z_{0},t_{0},t) &{} = &{} \overline{u}_{i}(Z_{t}^{i})e^{-qt}-\overline{u}_{i}(z_{0})e^{-qt_{0}}+\sum _{n=N_{t_{0}}^{i}}^{N_{t}^{i}}e^{-q\tau _{n}^{i}}p_{i}(l-Z_{\tau _{n}^{i-}}^{i}+Y_{n}^{i})1_{\{Z_{\tau _{n}^{i-}}^{i}-Y_{n}^{i}-l<0\}}\\ &{} &{} -\int _{t_{0}}^{t}e^{-qs}\left( \sigma _{i}\overline{u}_{i}^{\prime } (Z_{s}^{i})-(q+\lambda _{i})\overline{u}_{i}(Z_{s}^{i})+\lambda _{i}\int _{0}^{Z_{s^{-}}^{i}-l}\overline{u}_{i}(Z_{s^{-}}^{i}-\alpha )dF_{i} (\alpha )\right) ds\\ &{} &{} -\int _{t_{0}}^{t}e^{-qs}\left( \lambda _{i}\int _{Z_{s^{-}}^{i}-l} ^{\infty }\left( p_{i}(\alpha -Z_{s^{-}}^{i}+l)+\overline{u}_{i}(l)\right) dF_{i}(\alpha )\right) ds\text {,} \end{array} \end{aligned}$$
(8.10)

with

$$\begin{aligned} \text { }Z_{t}^{i}=z_{0}+\sigma _{i}\left( t-t_{0}\right) -\sum \nolimits _{n=N_{t_{0}}^{i}}^{N_{t}^{i}}\min \{Y_{n}^{i},Z_{\tau _{n}^{i-.}} ^{i}-l\}\text { for }t\ge t_{0}\ge 0, \end{aligned}$$

it can be seen that \(M^{i}(z_{0},t_{0},t)\) is a martingale with zero expectation for \(t\ge t_{0}\).

Consider first the case \(J_{k}=i\) and \(J_{k+1}=j\) with \(i=1,2\), \(j=0,1,2\) and \(i\ne j\). Since \(\overline{u}_{i}\) is absolutely continuous, the function \(t\rightarrow \overline{u}_{i}(X_{t}^{\pi })e^{-qt}\) is also absolutely continuous, between customer demands. Using an extension of the Dynkin’s Formula, we obtain

$$\begin{aligned} \begin{array}{l} \overline{u}_{j}(X_{T_{k+1}}^{\pi })e^{-qT_{k+1}}-\overline{u}_{i}(X_{T_{k}}^{\pi })e^{-qT_{k}}\\ \begin{array}{ll} = &{} \overline{u}_{j}(X_{T_{k+1}}^{\pi })e^{-qT_{k+1}}-\overline{u}_{i}(X_{T_{k+1}}^{\pi })e^{-qT_{k+1}}+\overline{u}_{i}(X_{T_{k+1}}^{\pi })e^{-qT_{k+1}}-\overline{u}_{i}(X_{T_{k}}^{\pi })e^{-qT_{k}}\\ \ge &{} -K_{ij}e^{-qT_{k+1}}+\overline{u}_{i}(X_{T_{k+1}}^{\pi })e^{-qT_{k+1}}-\overline{u}_{i}(X_{T_{k}}^{\pi })e^{-qT_{k}}\\ = &{} -K_{ij}e^{-qT_{k+1}}+\int _{T_{k}}^{T_{k+1}}e^{-qs}\mathcal {L}_{i}(\overline{u}_{i})(X_{s}^{\pi })ds\\ &{} -\left( \int _{T_{k}}^{T_{k+1}}e^{-qs}h_{i}(X_{s}^{\pi })ds+\sum _{n=N_{T_{k}}^{i}}^{N_{T_{k+1}}^{i}}e^{-q\tau _{n}^{i}}p_{i}(l-X_{\tau _{n}^{-}}^{\pi }+Y_{n}^{i})1_{\{X_{\tau _{n}^{i-}}^{\pi }-Y_{n}^{i}-l<0\}}\right) \\ &{} +M^{i}(X_{T_{k}}^{\pi },T_{k},T_{k+1}); \end{array} \end{array} \end{aligned}$$

and so, since \(\overline{u}_{i}\) is a supersolution of (4.5), we get that

$$\begin{aligned} \begin{array}{l} \mathbb {E}\left[ \left. \overline{u}_{j}(X_{T_{k+1}}^{\pi })e^{-qT_{k+1}}-\overline{u}_{i}(X_{T_{k}}^{\pi })e^{-qT_{k}}\right| \mathcal {F}_{T_{k}}\right] \\ \begin{array}{ll} \ge &{} -\mathbb {E}\left[ K_{ij}e^{-qT_{k+1}}+\left. \int _{T_{k}}^{T_{k+1}}e^{-qs}h_{i}(X_{s}^{\pi })ds+\sum _{n=N_{T_{k}}}^{N_{T_{k+1}}}e^{-q\tau _{n}^{i}}p_{i}(l-X_{\tau _{n}^{-}}^{\pi }+Y_{n}^{i})1_{\{X_{\tau _{n}^{i-}}^{\pi }-Y_{n}^{i}-l<0\}}\right| \mathcal {F}_{T_{k}}\right] . \end{array} \end{array} \end{aligned}$$

In the case \(J_{k}=0\) we have \(\mathcal {J}_{T_{k+1}}\ne 0\), then \(X_{s}^{\pi }=b\) in \([T_{k},T_{k+1})\) , \(T_{k+1}=\tau _{n}^{0}\) for some n and so, analogously to the previous case,

$$\begin{aligned} \begin{array}{l} \overline{u}_{\mathcal {J}_{T_{k+1}}}(X_{T_{k+1}}^{\pi })e^{-qT_{k+1}}-\overline{u}_{0}(X_{T_{k}}^{\pi })e^{-qT_{k}}\\ \begin{array}{ll} = &{} e^{-qT_{k}}\left( (\overline{u}(b-Y_{n}^{0})1_{\{b-Y_{n}^{0}-l\ge 0\}}+\overline{u}_{\mathcal {J}_{T_{k+1}}}(l)1_{\{b-Y_{n}^{0}-l<0\}})e^{-q(T_{k+1}-T_{k})}-\overline{u}_{0}(b)\right) -K_{0\mathcal {J}_{T_{k+1}}}e^{-qT_{k+1}}\\ = &{} e^{-qT_{k}}\left( \overline{u}(b-Y_{n})1_{\{b-Y_{n}^{0}-l\ge 0\}}+\overline{u}_{\mathcal {J}_{T_{k+1}}}(l)1_{\{b-Y_{n}^{0}-l<0\}}+p_{0}(l-b+Y_{n}^{0})1_{\{b-Y_{n}^{0}-l<0\}}\right) e^{-q(T_{k+1}-T_{k})}\\ &{} -e^{-qT_{k}}\left( \frac{\lambda _{0}}{q+\lambda _{0}}\left( {\textstyle \int _{0}^{b-l}} \overline{u}(b-\alpha )dF_{0}(\alpha )+ {\textstyle \int _{b-l}^{\infty }} \left( p_{0}(\alpha -b+l)+\overline{u}(l)\right) dF_{0}(\alpha )\right) \right) \\ &{} -\left( K_{0\mathcal {J}_{T_{k+1}}}e^{-qT_{k+1}}+\int _{T_{k}}^{T_{k+1} }e^{-qs}h_{0}(b)ds+e^{-qT_{k+1}}p_{0}(l-b+Y_{n}^{0})1_{\{b-Y_{n}^{0} -l<0\}}\right) \\ &{} +\int _{T_{k}}^{T_{k+1}}e^{-qs}h_{0}(b)ds-\frac{\lambda _{0}}{q+\lambda _{0} }e^{-qT_{k}}h_{0}(b)\text {,} \end{array} \end{array} \end{aligned}$$

and, since \(T_{k+1}-T_{k}\) is distributed as \(exp(\lambda _{0})\), we obtain that

$$\begin{aligned} \begin{array}{ll} 0= &{} \mathbb {E}\left[ \left. e^{-qT_{k}}\left( \overline{u}(b-Y_{n}^{0})1_{\{b-Y_{n}^{0}-l\ge 0\}}+\overline{u}_{\mathcal {J}_{T_{k+1}}}(l)1_{\{b-Y_{n}^{0}-l<0\}}\right) \right| \mathcal {F}_{T_{k}}\right] \\ &{} +\mathbb {E}\left[ \left. e^{-qT_{k+1}}p_{0}(l-b+Y_{n}^{0})1_{\{b-Y_{n}^{0}-l<0\}}\right| \mathcal {F}_{T_{k}}\right] \\ &{} -\mathbb {E}\left[ \left. \frac{e^{-qT_{k}}}{q+\lambda _{0}}\lambda _{0}\left( {\textstyle \int _{0}^{b-l}} \overline{u}(b-\alpha )dF_{0}(\alpha )+ {\textstyle \int _{b-l}^{\infty }} \left( p_{0}(\alpha -b+l)+\overline{u}(l)\right) dF_{0}(\alpha )\right) \right| \mathcal {F}_{T_{k}}\right] . \end{array} \end{aligned}$$

and so

$$\begin{aligned} \begin{array}{l} \mathbb {E}\left[ \left. \overline{u}_{\mathcal {J}_{T_{k+1}}}(X_{T_{k+1}}^{\pi })e^{-qT_{k+1}}-\overline{u}_{0}(X_{T_{k}}^{\pi })e^{-qT_{k}}\right| \mathcal {F}_{T_{k}}\right] \\ \begin{array}{cc} = &{} -\mathbb {E}\left[ \left. K_{0\mathcal {J}_{T_{k+1}}}e^{-qT_{k+1}}+\int _{T_{k}}^{T_{k+1}}e^{-qs}h_{0}(b)ds+e^{-qT_{k+1}}p_{0}(l-b+Y_{n}^{0})1_{\{b-Y_{n}^{0}-l<0\}}\right| \mathcal {F}_{T_{k}}\right] . \end{array} \end{array} \end{aligned}$$

Analogously, we can prove that

$$\begin{aligned} \begin{array}{c} \begin{array}{l} \mathbb {E}\left[ \left. \overline{u}_{J_{m_{t}}}(X_{t}^{\pi })e^{-qt}-\overline{u}_{_{J_{m_{t}}}}(X_{T_{m_{t}}}^{\pi })e^{-qT_{m_{t}}}\right| \mathcal {F}_{T_{m_{t}}}\right] \\ \ge -\mathbb {E}\left[ \left. \int _{T_{m_{t}}}^{t}e^{-qs}h_{_{J_{m_{t}}}}(X_{s}^{\pi })ds+\sum _{n=N_{T_{m_{t}}}^{J_{m_{t}}}}^{N_{t}^{J_{m_{t}}}}e^{-q\tau _{n}^{J_{m_{t}}}}p_{J_{m_{t}}}(l-X_{\tau _{n}^{J_{m_{t}}-}}^{\pi }+Y_{n}^{J_{m_{t}}})1_{\{X_{\tau _{n}^{J_{m_{t}}-}}^{\pi }-Y_{n}^{J_{m_{t}}}-l<0\}}\right| \mathcal {F}_{T_{m_{t}}}\right] . \end{array} \end{array} \end{aligned}$$

Taking the expected value in (8.9), we obtain

$$\begin{aligned} \begin{array}{lll} \mathbb {E}\left[ \overline{u}_{\mathcal {J}_{t}}(X_{t}^{\pi })e^{-qt}\right] -\overline{u}_{j}(x) &{} = &{} \mathbb {E}\left[ \sum _{k=0}^{m_{t}-1} \mathbb {E}\left[ \left. \left( \overline{u}_{_{J_{k+1}}}(X_{T_{k+1}}^{\pi })e^{-qT_{k+1}}-\overline{u}_{_{J_{k}}}(X_{T_{k}}^{\pi })e^{-qT_{k}}\right) \right| \mathcal {F}_{T_{k}}\right] \right] \\ &{} &{} +\mathbb {E}\left[ \mathbb {E}\left[ \left. (\overline{u}_{J_{m_{t}} }(X_{t}^{\pi })e^{-qt}-\overline{u}_{_{J_{m_{t}}}}(X_{T_{m_{t}}}^{\pi })e^{-qT_{m_{t}}})\right| \mathcal {F}_{T_{m_{t}}}\right] \right] \\ &{} \ge &{} -V_{j}^{\pi }(x) \end{array} \end{aligned}$$

taking the limit with t going to infinity, and using that \(X_{t}^{\pi }\in [l,b]\) we obtain that \({\overline{u}}_{j}(x)\le {V}_{j}^{\pi}(x)\) for \(j=1,2\).

Considering instead the controlled risk process \(X_{t}^{\pi }\) starting at b, we obtain with a similar proof that \(\overline{u}_{0}(b)\le V_{0}^{\pi }(b)\).   

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Azcue, P., Frostig, E. & Muler, N. Optimal Strategies in a Production Inventory Control Model. Methodol Comput Appl Probab 25, 43 (2023). https://doi.org/10.1007/s11009-023-10024-3

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11009-023-10024-3

Keywords

Mathematics Subject Classification

Navigation