Skip to main content

Advertisement

Log in

Inventory control with modulated demand and a partially observed modulation process

  • Original Research
  • Published:
Annals of Operations Research Aims and scope Submit manuscript

Abstract

We consider a periodic review inventory control problem having an underlying modulation process that affects demand and that is partially observed by the uncensored demand process and a novel additional observation data (AOD) process. We present an attainability condition, AC, that guarantees the existence of an optimal myopic base stock policy if the reorder cost \(K=0\) and the existence of an optimal (sS) policy if \(K>0\), where both policies depend on the belief function of the modulation process. Assuming AC holds, we show that (i) when \(K=0\), the value of the optimal base stock level is constant within regions of the belief space and that each region can be described by two linear inequalities and (ii) when \(K>0\), the values of s and S and upper and lower bounds on these values are constant within regions of the belief space and that these regions can be described by a finite set of linear inequalities. A heuristic and bounds for the \(K=0\) case are presented when AC does not hold. Special cases of this inventory control problem include problems considered in the Markov-modulated demand and Bayesian updating literatures.

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

Similar content being viewed by others

References

  • Arifoglu, K., & Ozekici, S. (2010). Optimal policies for inventory systems with finite capacity and partially observed Markov-modulated demand and supply processes. European Journal of Operations Research, 204(3), 421–438.

    Google Scholar 

  • Arrow, K. J., Harris, T., & Marschak, J. (1951). Optimal inventory policy. Econometrica, 19(3), 252–272.

    Google Scholar 

  • Arrow, K. J., Scarf, H., & Karlin, S. (1958). Studies in the Mathematical Theory of Inventory and Production. Stanford: Stanford Press.

    Google Scholar 

  • Atrash, A., & Pineau, J. (2010). A Bayesian method for learning POMDP observation parameters for robot interaction management systems. In In The POMDP Practitioners Workshop.

  • Azoury, K. S. (1985). Bayes solution to dynamic inventory models under unknown demand distribution. Management Science, 31(9), 1150–1160.

    Google Scholar 

  • Azoury, K. S., & Miller, B. L. (1984). A comparison of the optimal ordering levels of Bayesian and non-Bayesian inventory models. Management Science, 30(8), 993–1003.

    Google Scholar 

  • Ban, G. Y. (2020). Confidence intervals for data-driven inventory policies with demand censoring. Operations Research, 68(2), 309–326. https://doi.org/10.1287/opre.2019.1883

    Article  Google Scholar 

  • Ban, G. Y., & Rudin, C. (2019). The big data newsvendor: Practical insights from machine learning. Operations Research, 67(1), 90–108. https://doi.org/10.1287/opre.2018.1757

    Article  Google Scholar 

  • Bayraktar, E., & Ludkovski, M. (2010). Inventory management with partially observed nonstationary demand. Annals of Operations Research, 176(1), 7–39.

    Google Scholar 

  • Bellman, R. (1958). Review. Management Science, 5(1), 139–141.

    Google Scholar 

  • Bensoussan, A., Cakanyildirim, M., & Sethi, S. P. (2007). A multiperiod newsvendor problem with partially observed demand. Mathematics of Operations Research, 32(2), 322–344.

    Google Scholar 

  • Bensoussan, A., Cakanyildirim, M., & Sethi, S. P. (2007). Partially observed inventory systems: The case of zero-balance walk. SIAM Journal on Control and Optimization, 46(1), 176–209.

    Google Scholar 

  • Bensoussan, A., Cakanyildirim, M., Minjarez-Sosa, J. A., Royal, A., & Sethi, S. P. (2008). Inventory problems with partially observed demands and lost sales. Journal of Optimization Theory and Applications, 136(3), 321–340.

    Google Scholar 

  • Bertsimas, D., & Kallus, N. (2020). From predictive to prescriptive analytics. Management Science, 66(3), 1025–1044. https://doi.org/10.1287/mnsc.2018.3253

    Article  Google Scholar 

  • Besbes, O., & Muharremoglu, A. (2013). On implications of demand censoring in the newsvendor problem. Management Science, 59(6), 1407–1424.

    Google Scholar 

  • Bookbinder, J. H., & Lordahl, A. E. (1989). Estimation of inventory re-order levels using the bootstrap statistical procedure. IIE Transactions, 21(4), 302–312.

    Google Scholar 

  • Chang, Y., Erera, A. L., & White, C. C. (2015). A leader-follower partially observed, multiobjective Markov game. Annals of Operations Research, 235(1), 103–128.

    Google Scholar 

  • Chang, Y., Erera, A. L., & White, C. C. (2015). Value of information for a leader-follower partially observed Markov game. Annals of Operations Research, 235(1), 129–153.

    Google Scholar 

  • Chen, L. G., Robinson, L. W., Roundy, R. O., & Zhang, R. Q. (2015). Technical note- new sufficient conditions for (s, S) policies to be optimal in systems with multiple uncertainties. Operations Research, 63(1), 186–197.

    Google Scholar 

  • Cheung, W. C., & Simchi-Levi, D. (2019). Sampling-based approximation schemes for capacitated stochastic inventory control models. Mathematics of Operations Research, 44(2), 668–692. https://doi.org/10.1287/moor.2018.0940

    Article  Google Scholar 

  • Choi, T. M. (Ed.). (2014). Handbook of Newsvendor Problems: Models, Extensions and Applications. New York: Springer.

    Google Scholar 

  • Ding, X., Puterman, M. L., & Bisi, A. (2002). The censored newsvendor and the optimal acquisition of information. Operations Research, 50(3), 517–527.

    Google Scholar 

  • Dvoretzky, A., Keifer, J., & Wolfowitz, J. (1952). The inventory problem: II. Case of unknown distributions of demand. Econometrica, 20(3), 450–466.

    Google Scholar 

  • Ferreira, K. J., Lee, B. H. A., & Simchi-Levi, D. (2016). Analytics for an online retailer: Demand forecasting and price optimization. Manufacturing & Service Operations Management, 18(1), 69–88.

    Google Scholar 

  • Gallego, G., & Hu, H. (2004). Optimal policies for production/inventory systems with finite capacity and Markov-modulated demand and supply processes. Annals of Operations Research, 126(1), 21–41.

    Google Scholar 

  • Gallego, G., & Moon, I. (1993). The distribution free newsboy problem: Review and extensions. The Journal of the Operational Research Society, 44(8), 825–834.

    Google Scholar 

  • Godfrey, G. A., & Powell, W. B. (2001). An adaptive, distribution-free algorithm for the newsvendor problem with censored demands, with applications to inventory and distribution. Management Science, 47(8), 1101–1112.

    Google Scholar 

  • Graves, S. C., Rinnooy Kan, A. H. G., & Zipkin, P. H. (1993). Logistics of Production and Inventory, Handbooks in Operations Research and Management Science, (Vol. 4). London: Elsevier.

    Google Scholar 

  • Huh, W. T., & Rusmevichientong, P. (2009). A nonparametric asymptotic analysis of inventory planning with censored demand. Mathematics of Operations Research, 34(1), 103–123.

    Google Scholar 

  • Huh, W. T., Janakiraman, G., Muckstadt, J. A., & Rusmevichientong, P. (2009). An adaptive algorithm for finding the optimal base-stock policy in lost sales inventory systems with censored demand. Mathematics of Operations Research, 34(2), 397–416. https://doi.org/10.1287/moor.1080.0367

    Article  Google Scholar 

  • Huh, W. T., Levi, R., Rusmevichientong, P., & Orlin, J. B. (2011). Adaptive data-driven inventory control with censored demand based on Kaplan-Meier estimator. Operations Research, 59(4), 929–941.

    Google Scholar 

  • Iglehart, D. (1963). Optimality of (s, S) policies in the infinite horizon dynamic inventory problem. Management Science, 9(2), 259–267.

    Google Scholar 

  • Iglehart, D., & Karlin, S. (1962). Optimal policy for dynamic inventory process with nonstationary stochastic demands, Stanford, CA: Stanford University Press, chap 8.

  • Kamath, R., & Pakkala, T. P. M. (2002). A Bayesian approach to a dynamic inventory model under an unknown demand distribution. Computers and Operations Research, 29(2002), 403–422.

    Google Scholar 

  • Karlin, S. (1958). One-Stage Inventory Models with Uncertainty. Stanford: Stanford Press.

    Google Scholar 

  • Karlin, S. (1958). Optimal Inventory Policy for the Arrow-Harris-Marschak Dynamic Model. Stanford: Stanford Press.

    Google Scholar 

  • Karlin, S. (1959). Dynamic inventory policy with varying stochastic demands. Management Science, 6(3), 231–258.

    Google Scholar 

  • Karlin, S. (1959). Optimal policy for dynamic inventory process with stochastic demands subject to seasonal variations. Journal of the Society of Industrial and Applied Mathematics, 8(4), 611–629.

    Google Scholar 

  • Katehakis, M. N., & Smit, L. C. (2012). On computing optimal (q, r) replenishment policies under quantity discounts. Annals of Operations Research, 200(1), 279–298.

    Google Scholar 

  • Katehakis, M. N., Melamed, B., & Shi, J. J. (2015). Optimal replenishment rate for inventory systems with compound poisson demands and lost sales: a direct treatment of time-average cost. Annals of Operations Research. https://doi.org/10.1007/s10479-015-1998-y

  • Katehakis, M. N., Melamed, B., & Shi, J. J. (2016). Cash-flow based dynamic inventory management. Production and Operations Management, 25(9), 1558–1575. https://doi.org/10.1111/poms.12571

    Article  Google Scholar 

  • Khouja, M. (1999). The single-period (news-vendor) problem: Literature review and suggestions for future research. Omega, 27(5), 537–553.

    Google Scholar 

  • Klabjan, D., Simch-Levi, D., & Song, M. (2013). Robust stochastic lot-sizing by means of histograms. Production and Operations Management, 22(3), 691–710.

    Google Scholar 

  • Lariviere, M., & Porteus, E. (1999). Stalking information: Bayesian inventory management with unobserved lost sales. Management Science, 45(3), 346–363.

    Google Scholar 

  • Levi, R., Perakis, G., & Uichanco, J. (2015). The data-driven newsvendor problem: New bounds and insights. Operations Research, 63(6), 1294–1306.

    Google Scholar 

  • Lovejoy, W. S. (1990). Myopic policies for some inventory models with uncertain demand distributions. Management Science, 36(6), 724–738.

    Google Scholar 

  • Lovejoy, W. S. (1992). Stopped myopic policies in some inventory models with generalized demand processes. Management Science, 38(5), 688–707.

    Google Scholar 

  • Malladi, S. S., Erera, A. L., & White III, C. C. (2021). Managing mobile production-inventory systems influenced by a modulation process. Annals of Operations Research accepted.

  • Mamani, H., Nassiri, S., & Wagner, M. R. (2017). Closed-form solutions for robust inventory management. Management Science, 63(5), 1625–1643. https://doi.org/10.1287/mnsc.2015.2391

    Article  Google Scholar 

  • Morton, T. E. (1978). The nonstationary infinite horizon inventory problem. Management Science, 24(14), 1474–1482.

    Google Scholar 

  • Murray, G. R., Jr., & Silver, E. A. (1966). A Bayesian analysis of the style goods inventory problem. Management Science, 12(11), 785–797.

    Google Scholar 

  • Ortiz, O. L., Erera, A. L., & White, C. C. (2013). State observation accuracy and finite-memory policy performance. Operational Research Letters, 41(5), 477–481.

    Google Scholar 

  • Perakis, G., & Roels, G. (2008). Regret in the newsvendor model with partial information. Operations Research, 56(1), 188–203.

    Google Scholar 

  • Petruzzi, N. C., & Dada, M. (1999). Pricing and the newsvendor problem: A review with extensions. Operations Research, 47(2), 183–194.

    Google Scholar 

  • Puterman, M. L. (1994). Markov Decision Processes: Discrete Stochastic Dynamic Programming. Hoboken, New Jersey: Wiley.

    Google Scholar 

  • Qin, Y., Wang, R., Vakharia, A. J., Chen, Y., & Seref, M. M. (2011). The newsvendor problem: Review and directions for future research. European Journal of Operational Research, 213(2), 361–374.

    Google Scholar 

  • Scarf, H. (1959). Bayes solutions of the statistical inventory problem. Annals of Mathematical Statistics, 30(2), 490–508.

    Google Scholar 

  • Scarf, H. (1960). The optimality of \((S, s)\) policies in the dynamic inventory problem. In Arrow, K., Karlin, S., & Suppes, P. (Eds.), Mathematical Methods in the Social Sciences, chap 13.

  • Sethi, S. P., & Cheng, F. (1997). Optimality of (s, S) policies in inventory models with Markovian demand. Operations Research, 45(6), 931–939.

    Google Scholar 

  • Smallwood, R. D., & Sondik, E. J. (1973). The optimal control of partially observable Markov processes over a finite horizon. Operations Research, 21(5), 1071–1088.

    Google Scholar 

  • Sondik, E. J. (1978). The optimal control of partially observable Markov processes over the infinite horizon: Discounted costs. Operations Research, 26(2), 282–304.

    Google Scholar 

  • Song, J. S., & Zipkin, P. (1993). Inventory control in a fluctuating demand environment. Operations Research, 41(2), 282–304.

    Google Scholar 

  • Treharne, J. T., & Sox, C. R. (2002). Adaptive inventory control for nonstationary demand and partial information. Management Science, 48(5), 607–624.

    Google Scholar 

  • Veinott, A. F., Jr. (1965). Optimal policy for a multi-product, dynamic, nonstationary inventory problem. Management Science, 12(3), 206–222.

    Google Scholar 

  • Veinott, A. F., Jr. (1965). Optimal policy in a dynamic, single product, nonstationary inventory model with several demand classes. Operations Research, 13(5), 761–778.

    Google Scholar 

  • Veinott, A. F., Jr. (1966). On the optimality of (s, S) inventory policies: New conditions and a new proof*. SIAM Journal on Applied Mathematics, 14(5), 1067–1083.

    Google Scholar 

  • Veinott, A. F., Jr., & Wagner, H. M. (1965). Computing optimal (s, S) inventory policies. Management Science, 11(5), 525–552.

    Google Scholar 

  • White, C. C., & Harrington, D. (1980). Application of Jensen’s inequality to adaptive suboptimal design. Journal of Optimization Theory and Applications, 32(1), 89–99.

    Google Scholar 

  • Xin, L., & Goldberg, D. A. (2015). Distributionally robust inventory control when demand is a martingale. arXiv preprint arXiv:1511.09437.

  • Yuan, H., Luo, Q., & Shi, C. (2021). Marrying stochastic gradient descent with bandits: Learning algorithms for inventory systems with fixed costs. Management Science online. https://doi.org/10.1287/mnsc.2020.3799

  • Zhang, H., Chao, X., & Shi, C. (2020). Closing the gap: A learning algorithm for lost-sales inventory systems with lead times. Management Science, 66(5), 1962–1980. https://doi.org/10.1287/mnsc.2019.3288

    Article  Google Scholar 

  • Zipkin, P. (1989). Critical number policies for inventory models with periodic data. Management Science, 35(1), 71–80.

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Satya S. Malladi.

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 result in section 2

Proof of Lemma 2

If \(s^*(\varvec{x}) = d_m\), then

$$\begin{aligned} A_{m-1}(\varvec{x})d_{m-1} + B_{m-1}(\varvec{x})> & {} A_m(\varvec{x})d_m+ B_m(\varvec{x}), \\ A_{m+1}(\varvec{x})d_{m+1} + B_{m+1}(\varvec{x})\ge & {} A_m(\varvec{x})d_m+B_m(\varvec{x}), \end{aligned}$$

which leads to the result. \(\square \)

1.2 Proofs of results in section 3

Proof of Proposition 1

By induction. Letting \(v_0 = 0\), we note that

$$\begin{aligned} v_1(s,\varvec{x})=\min _{y\ge s} L(\varvec{x},y) = L\big (\varvec{x},\max \{s^*(\varvec{x}),s\}\big ) \end{aligned}$$

for all \(\varvec{x}\) and \(L\big (\varvec{x},\max \{s^*(\varvec{x}),s\}\big )\) is non-decreasing and convex in s. Thus, the result holds true for \(n=1\) (and, trivially for \(n=0\)). Assume the result holds for n. Then, for \(s\le s^*(\varvec{x})\),

$$\begin{aligned} v_{n+1}(\varvec{x},s)\le & {} L\big (\varvec{x},s^*(\varvec{x})\big ) + \beta \sum _{d,z} \sigma (d,z,\varvec{x})v_n\big (\varvec{\lambda }(d,z,\varvec{x}) , f\big ( s^*(\varvec{x}), d\big )\big ) \\ = L\big (\varvec{x},s^*(\varvec{x})\big )+ & {} \beta \sum _{d,z} \sigma (d,z,\varvec{x})v_n\big (\varvec{\lambda }(d,z,\varvec{x}), s^*\big (\varvec{\lambda }(d,z,\varvec{x})\big ) \big ) \ \text {(using Section 2.4).} \\ \text {Also, } v_{n+1}(\varvec{x},s)\ge & {} \min _{y\ge s} L(\varvec{x},y)+ \beta \sum _{d,z} \sigma (d,z,\varvec{x}) \min _y v_n \big (\varvec{\lambda }(d,z,\varvec{x}),f\big (y, d\big ) \big ) \\ {}= & {} L\big (\varvec{x},s^*(\varvec{x})\big ) +\beta \sum _{d,z} \sigma (d,z,\varvec{x}) v_n \big (\varvec{\lambda }(d,z,\varvec{x}), s^*\big (\varvec{\lambda }(d,z,\varvec{x})\big ) \big ) \\ {}= & {} L\big (\varvec{x},s^*(\varvec{x})\big ) +\beta \sum _{d,z} \sigma (d,z,\varvec{x}) v_n \big (\varvec{\lambda }(d,z,\varvec{x}), f\big ( s^*(\varvec{x}), d\big ) \big ). \end{aligned}$$

\(\text { Thus, for } s\le s^*(\varvec{x}), \text { } \)

$$\begin{aligned} v_{n+1}(\varvec{x},s) = L\big (\varvec{x},s^*(\varvec{x})\big ) + \beta \sum _{d,z} \sigma (d,z,\varvec{x}) v_{n}\big (\varvec{\lambda }(d,z,\varvec{x}), f \big (s^*(\varvec{x}),d\big )\big ), \end{aligned}$$

and \( v_{n+1}(\varvec{x},s) = v_{n+1}(\varvec{x}, s^*(\varvec{x})) .\)

Assume \(s \ge s^*(\varvec{x})\). Note

$$\begin{aligned} v_{n+1}(\varvec{x},s)\le & {} L(\varvec{x},s) + \beta \sum _{d,z} \sigma (d,z,\varvec{x}) v_n\big (\varvec{\lambda }(d,z,\varvec{x}),f\big (s, d\big )\big ). \\ \text { Also, } v_{n+1}(\varvec{x},s)\ge & {} \min _{y\ge s} L(\varvec{x},y) + \beta \sum _{d,z} \sigma (d,z,\varvec{x})\min _{y\ge s} v_n \big (\varvec{\lambda }(d,z,\varvec{x}),f\big (y, d\big )\big ) \\= & {} L(\varvec{x},s) + \beta \sum _{d,z} \sigma (d,z,\varvec{x}) v_n\big (\varvec{\lambda }(d,z,\varvec{x}), f \big (s, d \big )\big ), \\ \text { and hence for } s\ge & {} s^*(\varvec{x}), \ \text { } v_{n+1}(\varvec{x},s) = L(\varvec{x},s) + \beta \sum _{d,z} \sigma (d,z,\varvec{x}) v_n\big (\varvec{\lambda }(d,z,\varvec{x}),f \big (s, d \big )\big ) \end{aligned}$$

and is non-decreasing and convex in s. \(\square \)

Proof of Lemma 4

It is sufficient to show that if \(y\le y'\) and \(\varvec{x} \preceq \varvec{x'}\), then,

$$\begin{aligned} L(\varvec{x},y) - L(\varvec{x},y') \le L(\varvec{x'},y)-L(\varvec{x'},y'), \end{aligned}$$

which follows from the assumptions and [Puterman (1994), Lemma 4.7.2]. \(\square \)

Ideally, we would want to select \(\widehat{\varvec{x}}^{d,z}\) so that \(s^*(\varvec{x'})\le s^*(\widehat{\varvec{x}}^{d,z})\) for all \(\varvec{x'}\) such that \(\varvec{x'} \preceq \varvec{\lambda }(d,z,\varvec{x}) \ \ \forall \ \varvec{x}\in X\), for all (dz), which would strengthen Lemma 4 as much as possible. We construct such an \(\widehat{\varvec{x}}^{d,z}\) after the following preliminary result.

Lemma 6

The set \(\{ \varvec{\lambda }(d,z,\varvec{x}): \varvec{x}\in X\}\) = \(\bigg \{ \sum _i \xi _i \varvec{\lambda }(d,z,\varvec{e_i}) : \xi _i \ge 0 \ \forall i, \sum _i \xi _i = 1 \bigg \}.\)

We remark that if \(\varvec{x} \preceq \varvec{x'}\) and \(\varvec{x} \preceq \varvec{x''}\), then \(\varvec{x}\preceq \alpha \varvec{x'}+(1-\alpha )\varvec{x''} \) for all \(\alpha \in [0,1]\). Thus, if \(\widehat{\varvec{x}}^{d,z}\) is such that \(\widehat{\varvec{x}}^{d,z} \preceq \lambda (d,z,\varvec{e_i})\) for all i, then \(\widehat{\varvec{x}}^{d,z}\) is such that \(\widehat{\varvec{x}}^{d,z} \preceq \varvec{x'}\) for all \(\varvec{x'} \in \big \{ \varvec{\lambda }(d,z,\varvec{x}) : \varvec{x}\in X \big \}\).

1.3 Construction of \(\widehat{x}^{d,z}\)

We now construct \(\widehat{\varvec{x}}^{d,z}\). Let

$$\begin{aligned} \widehat{x}_N^{d,z}= & {} \min \big \{ \varvec{\lambda }_N(d,z,\varvec{e_i}), i=1,\dots ,N \big \} \\ \widehat{x}_n^{d,z}= & {} \min \bigg \{ \sum _{k=n}^N \varvec{\lambda }_k(d,z,\varvec{e_i}), i =1,\dots , N \bigg \} -\sum _{k=n+1}^N \widehat{x}_k^{d,z}, \ \ n= N-1, \dots , 2 \\ \widehat{x}_1^{d,z}= & {} 1 - \sum _{k=2}^N \widehat{x}_k^{d,z}. \end{aligned}$$

By construction, \(\widehat{\varvec{x}}^{d,z}\preceq \varvec{\lambda }(d,z, \varvec{x}) \ \forall \ \varvec{x}\in X\). We now show that \(\widehat{\varvec{x}}^{d,z}\in X\) and that \(s^*(\varvec{x'})\le s^*(\widehat{\varvec{x}}^{d,z})\) for all \(\varvec{x'}\in X\) such that \(\varvec{x'}\preceq \varvec{\lambda }(d,z,\varvec{x}) \ \forall \ \varvec{x}\in X\).

Lemma 7

(i) \(\widehat{\varvec{x}}^{d,z} \in X\). (ii) For any \(\varvec{x'}\preceq \varvec{\lambda }(d,z,\varvec{x}) \ \forall \ \varvec{x}\in X, s^*(\varvec{x'}) \le s^*(\widehat{\varvec{x}}^{d,z})\).

Proof of Lemma 7

We have the following:

  1. (i)

    Clearly, \(0\le \widehat{\varvec{x}}^{d,z}_N \le 1\) and \(\sum _{n=1}^N \widehat{\varvec{x}}_n^{d,z} = 1\). It is sufficient to show \(0\le \widehat{\varvec{x}}^{d,z}_n, n=N-1, \dots , 1\). Note

    $$\begin{aligned} \sum _{k=n+1}^N\widehat{\varvec{x}}^{d,z}_k = \min _{1\le i\le N} \bigg \{\sum _{k=n+1}^N \lambda _k(d,z,\varvec{e_i})\bigg \} \le \sum _{k=n+1}^N \lambda _k(d,z,\varvec{e_i}) \le \sum _{k=n}^N \lambda _k(d,z,\varvec{e_i}), \ \forall \ i. \end{aligned}$$

    Thus, \(\sum _{k=n+1}^N\widehat{x}^{d,z}_k \le \min _{1\le i\le N} \bigg \{\sum _{k=n}^N \lambda _k(d,z,\varvec{e_i})\bigg \} = \sum _{k=n}^N\widehat{x}^{d,z}_k\), and hence \(\widehat{x}_n^{d,z}\ge 0\).

  2. (ii)

    Let \(\varvec{x'}\preceq \varvec{\lambda }(d,z,\varvec{x}) \ \forall \ \varvec{x} \in X\) and assume \(s^*(\widehat{\varvec{x}}^{d,z}) < s^*(\varvec{x'})\). Then, there is an \(n\in \{1, \dots , N\}\) such that \(\sum _{k=n}^N x_k' > \sum _{k=n}^N \widehat{x}_k^{d,z}\). However, \(\sum _{k=n}^N \widehat{x}^{d,z}_k = \min _{1\le i\le N}\) \(\bigg \{\sum _{k=n}^N \lambda _k(d,z,\varvec{e_i})\bigg \}\), which leads to a contradiction of the assumption that \(\varvec{x'} \preceq \varvec{\lambda }(d,z,\varvec{x})\) \(\forall \varvec{x} \in X\).\(\square \)

1.4 Computing the expected cost function, \(v_n\)

We now present a procedure for computing \(v_n(s, \varvec{x})\). We only consider the case where \(s=s^*(\varvec{x})\) due to Proposition 1 and Lemma 3. For notational simplicity, we assume that \(\text {Pr}\big (z(t+1) \mid \mu (t+1),\mu (t)\big )\) is independent of \(\mu (t+1)\) and \(\mu (t)\). Extension to the more general case is straightforward.

Assume \(v_0=0\), let \(n=1\), and recall \(v_1\big (\varvec{x},s^*(\varvec{x})\big ) = L\big (\varvec{x},s^*(\varvec{x})\big ). \) Note \( L(\varvec{x},y) = \varvec{x}\overline{\varvec{\gamma }}_y, \ \text { where } \overline{\varvec{\gamma }}_y = \sum _{d,z} \varvec{P}(d,z)\ \underline{1}\ c(y,d)\). Let \(\varGamma _1 = \{ \overline{\varvec{\gamma }}_y\}\), and note that if \(c(y, d) = p(d-y)^+ + h(y-d)^+\), it is sufficient to consider only \(y \in \{d_1,\dots , d_M\}\). Then, \(v_1\big (\varvec{x},s^*(\varvec{x})\big )= \min \big \{\varvec{x}\overline{\varvec{\gamma }}:\overline{\varvec{\gamma }} \in \varGamma _1 \big \}.\) Assume there is a finite set \(\varGamma _n\) such that \( v_n \big (\varvec{x}, s^*(\varvec{x})\big ) = \min \big \{\varvec{x}\varvec{\gamma }: \varvec{\gamma }\in \varGamma _n \big \}.\) Then,

$$\begin{aligned} v_{n+1}\big (\varvec{x},s^*(\varvec{x}) \big )= & {} L\big (\varvec{x},s^*(\varvec{x})\big ) + \beta \sum _{m=1}^M \sigma (d_m,\varvec{x}) v_n\big (\varvec{\lambda }(d_m, \varvec{x}),f\big (s^*(\varvec{x}),d_m\big )\big ) \\ {}= & {} \min \big \{\varvec{x}\overline{\varvec{\gamma }}:\overline{\varvec{\gamma }} \in \varGamma _1 \big \} + \beta \sum _{m=1}^M \sigma (d_m,\varvec{x}) v_n\big (\varvec{\lambda }(d_m,\varvec{x}),s^*(\varvec{\lambda }(d_m,\varvec{x}))\big ) \\ {}= & {} \min \big \{\varvec{x}\overline{\varvec{\gamma }}:\overline{\varvec{\gamma }} \in \varGamma _1 \big \} + \beta \sum _{m=1}^M \sigma (d_m,\varvec{x}) \min \big \{\varvec{\lambda }(d_m,\varvec{x})\varvec{\gamma }: \varvec{\gamma }\in \varGamma _n \big \} \\ {}= & {} \min _{\overline{\varvec{\gamma }}} \min _{\varvec{\gamma _1}}\dots \min _{\varvec{\gamma _M}} \bigg \{ \varvec{x}\overline{\varvec{\gamma }} + \beta \sum _{m=1}^M \sigma (d_m,\varvec{x})\varvec{\lambda }(d_m,\varvec{x}) \varvec{\gamma _m} \bigg \} \\ {}= & {} \min _{\overline{\varvec{\gamma }}} \min _{\varvec{\gamma _1}}\dots \min _{\varvec{\gamma _M}} \bigg \{ \varvec{x} \bigg [\overline{\varvec{\gamma }} + \beta \sum _{m=1}^M \varvec{P}(d_m) \varvec{\gamma _m} \bigg ] \bigg \} \end{aligned}$$

Thus, \(\varGamma _{n+1}\) is the set of all \(\varvec{\gamma }\) such that \( \varvec{\gamma } = \overline{\varvec{\gamma }} + \beta \sum _{m=1}^M \varvec{P}(d_m)\varvec{\gamma _m}, \) where \(\overline{\varvec{\gamma }} \in \varGamma _1\) and \(\varvec{\gamma _m} \in \varGamma _n\) for all \(m =1,\dots , M\), and for all n, \(v_n\big (\varvec{x}, s^*(\varvec{x})\big ) \) is piecewise linear and concave in \(\varvec{x}\).

Let \(|\varGamma |\) be the cardinality of the set \(\varGamma \). Then, \(|\varGamma _{n+1} | = |\varGamma _1 | |\varGamma _n |^M\), where \(|\varGamma _1 | \le M\), and hence the cardinality of \(\varGamma _n\) can grow rapidly. Many of the vectors in the sets \(\varGamma _n\) are redundant and can be eliminated, reducing both computational and storage burdens. An exhaustive literature study of elimination procedures and other solution methods for solving POMDPs can be found in Chang et al. (2015a).

1.5 Proofs of results in section 4

Proof of Lemma 5

Assume \(f(y,d) = y-d\) and \(c(y,d) = p(d-y)^++h(y-d)^+ \), recall that elements of \(\mathcal {P}_1\) are sets of the form \(\{\varvec{x}\in X: s^*(\varvec{x}) = d_m\}\) for all \(d_m\) such that \(\min _{\varvec{x}} s^*(\varvec{x}) \le d_m \le \max _{\varvec{x}} s^*(\varvec{x})\). Further recall that \(\{\varvec{x}\in X: s^*(\varvec{x}) = d_m\}\) is the set of all \(\varvec{x}\) such that

$$\begin{aligned} \sum _{k=1}^{m-1} \sigma (d_k,\varvec{x}) < p/(p+h) \le \sum _{k=1}^m \sigma (d_k,\varvec{x}), \end{aligned}$$

or equivalently,

$$\begin{aligned} \quad \quad \quad \quad \varvec{x}\sum _{k=1}^{m-1}\varvec{P}(d_k)\underline{1} < p/(p+h) \le \varvec{x}\sum _{k=1}^m \varvec{P}(d_k)\underline{1}, \end{aligned}$$

which represents two linear inequalities. Further, for \(\varvec{x}\in \{ \varvec{x} \in X: s^*(\varvec{x}) =d_m\}\), \( v_1^U(\varvec{x},s) = A_l(\varvec{x})d_l+B_l(\varvec{x}) \) for \(l=\max \{s^*(\varvec{x}),s\}\), where we note

$$\begin{aligned} A_l(\varvec{x})d_l+B_l(\varvec{x}) = \varvec{x} \big [h\sum _{k=1}^l (d_l-d_k)\varvec{P}(d_k)\underline{1} + p \sum _{k=l+1}^M(d_k-d_l)\varvec{P}(d_k)\underline{1} \big ], \end{aligned}$$

where \(A_j(\varvec{x})\) and \(B_j(\varvec{x})\) are defined in Section 2.3. Thus, on each element of \(\mathcal {P}_1\), \(v_1^U\) is linear in \(\varvec{x}\) for each s and each element of \(\mathcal {P}_1\) is described by a finite number of linear inequalities.

Let \((\varvec{x},s)\) be such that \(d_l \le \max \{s^*(\varvec{x}),s\} \le d_{l+1}\) for all \(\varvec{x}\) in an element \(\{\varvec{x}\in X: s^*(\varvec{x})=d_m\}\). Further, let \(d_{l(d,z)} \le \max \{s^*(\varvec{\lambda }(d,z,\varvec{x})),\max \{s^*(\varvec{x}),s\}-d\} \le d_{l(d,z) + 1}\) for all \(\varvec{x}\) in an element \(\{\varvec{x}\in X: s^*(\varvec{\lambda }(d,z,\varvec{x})) = d_{m(d)}\}\), which is the set of all \(\varvec{x}\) such that:

$$\begin{aligned} \varvec{\lambda }(d,z,\varvec{x})\sum _{k=1}^{m(d)-1}\varvec{P}(d_k)\underline{1} < p/(p+h) \le \varvec{\lambda }(d,z,\varvec{x})\sum _{k=1}^{m(d)} \varvec{P}(d_k)\underline{1}, \end{aligned}$$

or equivalently, for all \(\varvec{x}\) such that \(\sigma (d,\varvec{x}) \ne 0\),

$$\begin{aligned} \varvec{x}\varvec{P}(d,z)\sum _{k=1}^{m(d)-1}\varvec{P}(d_k)\underline{1} < \big (p/(p+h)\big )\varvec{x}\varvec{P}(d,z)\underline{1} \le \varvec{x}\varvec{P}(d,z)\sum _{k=1}^{m(d)} \varvec{P}(d_k)\underline{1}, \end{aligned}$$

where we assume m and m(d) for all d have been chosen so that the finite set of linear inequalities describes a non-null subset of X. We note that for such a subset,

$$\begin{aligned} v^U_{n+1}(\varvec{x},s)= & {} A_l(\varvec{x})d_l +B_l(\varvec{x}) \\ {}{} & {} + \beta \sum _{d,z} \sigma (d,z,\varvec{x}) \bigg [A_{l(d,z)}(\varvec{\lambda }(d,z,\varvec{x}))d_{l(d,z)} + B_{l(d,z)}(\varvec{\lambda }(d,z,\varvec{x})) \bigg ] \\ {}= & {} \varvec{x}\bigg [h\sum _{k=1}^l(d_l-d_k)\varvec{P}(d_k)\underline{1} + p \sum _{k=l+1}^M(d_k-d_l)\varvec{P}(d_k)\underline{1} \\ {}{} & {} +\beta \sum _d \bigg [h \sum _z\sum _{k=1}^{l(d,z)}(d_{l(d,z)}-d_k)\varvec{P}(d,z)\varvec{P}(d_k)\underline{1} \\ {}{} & {} + p \sum _z \sum _{k=l(d,z)+1}^N(d_k-d_{l(d,z)})\varvec{P}(d,z)\varvec{P}(d_k)\underline{1}\bigg ]\bigg ]. \end{aligned}$$

The resulting partition \(\mathcal {P}_2\) is at least as fine as \(\mathcal {P}_1\) and each element in \(\mathcal {P}_2\) is described by a finite set of linear inequalities. We have shown that on each element in \(\mathcal {P}_2\), \(v_2^U(\varvec{x},s)\) is linear in \(\varvec{x}\) for each s. A straightforward induction argument shows these characteristics hold for all n. We illustrate by example (through Example 3) how \(v_n^U(\varvec{x},s)\) may be discontinuous in \(\varvec{x}\) for fixed s. \(\square \)

Proof of Proposition 4

It is sufficient to show the result holds for \(\tau = t+1\). There are two cases. First, let \(s(t) \le s^*(\varvec{x}(t))\). Then, \(s(t+1) = s^*(\varvec{x}(t))-d(t)\). We note

$$\begin{aligned} \min \{s^*(\varvec{x}): \varvec{x} \in X\}- d_M\le & {} s^*(\varvec{x}(t))- d(t) \\ {}\le & {} \max \{s^*(\varvec{x}): \varvec{x} \in X\}- d_1 \text { and hence,} \\ \min \{s^*(\varvec{x}): \varvec{x} \in X\} -d_M\le & {} s(t+1) \le \max \{s^*(\varvec{x}): \varvec{x} \in X\}- d_1 \end{aligned}$$

Second, let \(s^*(\varvec{x}(t)) \le s(t)\). Then, \(s(t+1) = s(t)- d(t) \ge s^*(\varvec{x}(t))-d(t)\). We note

$$\begin{aligned} \min \{s^*(\varvec{x}): \varvec{x} \in X\} - d_M \le s^*(\varvec{x}(t)) - d(t) \le s(t) - d(t) \\ \le \max \{s^*(\varvec{x}): \varvec{x} \in X\} - d_1, \text { and hence, } \\ \min \{s^*(\varvec{x}): \varvec{x} \in X\} -d_M \le s(t+1) \le \max \{s^*(\varvec{x}): \varvec{x} \in X\} - d_1. \end{aligned}$$

\(\square \)

1.6 Design of instances for computational study

We describe the generation of computational instances for Sect. 4.4. Each instance describes a backordering system with no fixed ordering cost. For each combination of number of modulation states \(N \in \{2,3\}\), number of demand outcomes \(M \in \{3,4,5\}\), randomly generate M unique ordered integer demand outcomes from \([0, D_L]\) for each \(D_L \in \{20, 100, 250, 500, 750, 1000\}\). Set the lowest demand outcome \(d(0) = 0\) ( to encourage A1 violation), randomly sample probability transition matrix \(\{P(i,j)\}\) and probability mass function for each modulation state \(\{Q(d,j)\}\) such that the N ordered expected demands \(ED_i, i =1, \dots , N\) are quite distinct and satisfy:

  • \(ED_1<= 0.5 d(M)\) and \(ED_2 > 0.5 d(M)\) and \(ED_2-ED_1 > 0.25 d(M)\) OR \(ED_2 - ED_1 > 0.5 d(M)\), when \(N = 2\) and

  • \(ED_1 <= 0.4 d(M)\) and \(ED_2 > 0.4 d(M)\) and \(ED_2 <= 0.7 d(M)\) and \(ED_3 > 0.7 d(M)\) and \(ED_2 - ED_1 > 0.2 d(M)\) and \(ED_3 - ED_2 > 0.2 d(M)\), when \(N = 3\).

Set the number of decision epochs T to 100 and vary backorder cost per unit per period p as \(\{1.5, 2, 3\}\), while keeping the holding cost h at 1.

1.7 Algorithms for computational study

figure a

Here, \(\varvec{e}(\mu _0)\) is the unit vector with 1 in the \(\mu _0\)th position.

figure b

1.8 Proofs of results in section 5

Proof of Proposition 7

The proof of Proposition 7 is a direct extension of the results in Scarf (1960). \(\square \)

Lemma 8

For all \(\varvec{x}\) and n:

  1. (i)

    if \(s\le s'\), then \(v_n(\varvec{x},s) \le v_n(\varvec{x}, s') + K\)

  2. (ii)

    if \(y\le y'\), then \( G_n(\varvec{x}, y') - G_n(\varvec{x}, y) \ge L(\varvec{x}, y') - L(\varvec{x}, y) - \beta K \)

  3. (iii)

    if \(s\le s' \le \underline{S}(\varvec{x})\), then \(v_n(\varvec{x},s) \ge v_n(\varvec{x},s')\)

  4. (iv)

    if \(y\le y'\le \underline{S}(\varvec{x})\), then \( G_n(\varvec{x},y') - G_n(\varvec{x},y) \le L(\varvec{x}, y') - L(\varvec{x}, y) \le 0. \)

Proof of Lemma 8

  1. (i)

    This result follows from the K-convexity of \(v_n(\varvec{x},s)\) in s, which is a direct implication of the second item of Proposition 7.

  2. (ii)

    This result follows from the definition of \(G_n(\varvec{x},y)\), the previous result (i), and the fact that f(yd) is convex and non-decreasing.

  3. (iii)

    \(G_n(\varvec{x}, s_n(\varvec{x})) \le K+ G_n(\varvec{x},S_n(\varvec{x})) \le K+ G_n(\varvec{x}, \underline{S}(\varvec{x}))\) implies that \(s_n(\varvec{x}) \le S_n(\varvec{x}) \le \underline{S}(\varvec{x})\) (This is an implication of the definitions of \(s_n(\varvec{x})\) and \(S_n(\varvec{x})\), and the fact that \(\underline{S}(\varvec{x})\) minimizes \(L(\varvec{x},y)\) while \(S_n(\varvec{x})\) minimizes the sum of \(L(\varvec{x},y)\) and a positive term.). It follows from the four cases of \(s\le s' \le \underline{S}(\varvec{x})\) with respect to the value of \(s_n(\varvec{x})\) that \(v_n(\varvec{x},s) \ge v_n(\varvec{x},s')\).

  4. (iv)

    This result follows from the definition of \(G_n(\varvec{x},y)\), the non-decreasing nature of f(yd) in y and (iii).\(\square \)

The proof of Proposition 8 requires four lemmas.

Lemma 9

For all n and \(\varvec{x}\), \( \underline{S}(\varvec{x}) = S_0(\varvec{x}) \le S_n(\varvec{x})\).

Lemma 10

For all n and \(\varvec{x}\), \(s_n(\varvec{x})\) can be selected so that \(s_n(\varvec{x}) \le \overline{s}(\varvec{x})\).

Lemma 11

For all n and \(\varvec{x}\), \(S_n(\varvec{x})\) can be selected so that \(S_n(\varvec{x}) \le \overline{S}(\varvec{x})\).

Lemma 12

For all n and \(\varvec{x}\), \(\underline{s}(\varvec{x}) \le s_n(\varvec{x})\).

Proof of Proposition 8

The proof of these results follow from the proofs of Lemmas 2 - 5 in Veinott and Wagner (1965). Proof of Proposition 8(a) follows from Lemmas  912, and Proposition 8(b) follows from (a) and Proposition 7. \(\square \)

1.9 Determining \(\varGamma _n(s)\)

As was true for the \(K=0\) case, when \(K>0\), there is a finite set of vectors \(\varGamma _n(s)\) such that \(v_n(\varvec{x}, s) = \min \{\varvec{x}\varvec{\gamma }: \varvec{\gamma } \in \varGamma _n(s) \}\) for all s. Note that \(\varGamma _0(s) = \{\underline{0}\}\) for all s, where \(\underline{0}\) is the column N-vector having zero in all entries. Given \(\{ \varGamma _n(s): \forall \ s \}\), we now present an approach for determining \(\{\varGamma _{n+1}(s): \forall \ s \}\). Recalling Sect. A2, let \(\overline{\varGamma } = \{\overline{\varvec{\gamma }}_1, \dots , \overline{\varvec{\gamma }}_M\}\) be such that \(\min _y L(\varvec{x},y) = \min \{\varvec{x} \varvec{\gamma }: \varvec{\gamma } \in \overline{\varGamma } \}\). Note

$$\begin{aligned} G_n(\varvec{x}, y) = L(\varvec{x},y) + \beta \sum _{d, z} \sigma (d, z,\varvec{x}) v_n \big (\varvec{\lambda }(d,z,\varvec{x}), f(y,d)\big ), \end{aligned}$$

for \(y \in \{d_1, \dots , d_M\}\). Then,

$$\begin{aligned} v_n\big (\varvec{\lambda }(d,z,\varvec{x}), f(y,d)\big ) = \min \{ \varvec{\lambda }(d,z,\varvec{x}) \varvec{\gamma }: \varvec{\gamma } \in \varGamma _n(f(y,d)) \}. \end{aligned}$$

Let \(\varGamma _n'(y)\) be the set of all vectors of the form

$$\begin{aligned} \overline{\varvec{\gamma }} + \beta \sum _{d,z} \varvec{P}(d,z)\varvec{\gamma }(d,z), \end{aligned}$$

where \(\overline{\varvec{\gamma }} \in \overline{\varGamma }\) and \(\varvec{\gamma }(d,z) \in \varGamma _n(f(y,d))\). Then, \(G_n(\varvec{x},y) = \min \{\varvec{x}\varvec{\gamma }: \varvec{\gamma } \in \varGamma _n'(y)\}\) and

$$\begin{aligned} v_{n+1}(\varvec{x},s) = {\left\{ \begin{array}{ll} K+G_n\big (\varvec{x},S_n(\varvec{x})\big ) &{} \ s\le s_n(\varvec{x}) \\ G_n(\varvec{x}, s) &{} \ \text {otherwise}, \end{array}\right. } \end{aligned}$$

where \(S_n(\varvec{x})\) and \(s_n(\varvec{x})\) are the smallest integers such that

$$\begin{aligned} G_n \big (\varvec{x}, S_n(\varvec{x})\big )\le & {} G_n(\varvec{x}, y) \ \forall y. \\ G_n\big (\varvec{x}, s_n(\varvec{x})\big )\le & {} K+G_n\big (\varvec{x}, S_n(\varvec{x})\big ). \end{aligned}$$

Let \(X_n(s', S')\) be the set of all \(\varvec{x}\in {X}\) such that \(s_n(\varvec{x}) = s'\) and \(S_n(\varvec{x}) = S'\). Thus, if \(\varvec{x}\in X_n(s', S')\), then \(s'\) and \(S'\) are the smallest integers such that

$$\begin{aligned} G_n \big (\varvec{x}, S'(\varvec{x})\big )\le & {} G_n(\varvec{x}, y) \ \forall y. \\ G_n\big (\varvec{x}, s'(\varvec{x})\big )\le & {} K+G_n\big (\varvec{x}, S'(\varvec{x})\big ). \end{aligned}$$

Since \(G_n(\varvec{x},y)\) is piecewise linear and convex in \(\varvec{x}\) for each y, \(X_n(s', S')\) is described by a finite set of linear inequalities. We remark that \(\{X_n(s', S'): s' \le S', \text { and } X_n(s', S') \ne \emptyset \} \) is a partition of X. Further, we remark that if \(\overline{X}(\underline{s}, \overline{s}, \underline{S}, \overline{S}) \cap X_n(s', S') \ne \emptyset \), then search for \((s', S')\) can be restricted to \(\underline{s} \le s' \le \overline{s}\) and \(\underline{S} \le S' \le \overline{S}\). Let \(\varGamma _{n+1}(s) = \{ K\underline{1} + \varvec{\gamma }: \varvec{\gamma } \in \varGamma _n'(S') \}\) for all \(s\le s'\), and let \(\varGamma _{n+1}(s) = \varGamma _{n}'(s)\) for all \(s>s'\). Thus, \(v_{n+1}(\varvec{x},s) = \min \{ \varvec{x} \varvec{\gamma }: \varvec{\gamma } \in \varGamma _{n+1}(s) \}\) for all s.

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

Malladi, S.S., Erera, A.L. & White, C.C. Inventory control with modulated demand and a partially observed modulation process. Ann Oper Res 321, 343–369 (2023). https://doi.org/10.1007/s10479-022-04932-9

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10479-022-04932-9

Keywords

Navigation