Skip to main content
Log in

Production-inventory control policy under warm/cold state-dependent fixed costs and stochastic demand: partial characterization and heuristics

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

Abstract

The motivation for our study comes from some production and inventory systems in which ordering/producing quantities that exceed certain thresholds in a given period might eliminate some setup activities in the next period. Many examples of such systems have been discussed in prior research but the analysis has been limited to production settings under deterministic demand. In this paper, we consider a periodic-review production-inventory model under stochastic demand and incorporate the following fixed-cost structure into our analysis. When the order quantity in a given period exceeds a specified threshold value, the system is assumed to be in a “warm” state and no fixed cost is incurred in the next period regardless of the order quantity; otherwise the system state is considered “cold” and a positive fixed cost is required to place an order. Assuming that the unsatisfied demand is lost, we develop a dynamic programming formulation of the problem and utilize the concepts of quasi-K-convexity and non-K-decreasing to show some structural results on the optimal cost-to-go functions. This analysis enables us to derive a partial characterization of the optimal policy under the assumption that the demands follow a Pólya or uniform distribution. The optimal policy is defined over multiple decision regions for each system state. We develop heuristic policies that are aimed to address the partially characterized decisions, simplify the ordering policy, and save computational efforts in implementation. The numerical experiments conducted on a large set of test instances including uniform, normal and Poisson demand distributions show that a heuristic policy that is inspired by the optimal policy is able to find the optimal solution in almost all instances, and that a so-called generalized base-stock policy provides quite satisfactory results under reasonable computational efforts. We use our numerical examples to generate insights on the impact of problem parameters. Finally, we extend our analysis into the infinite horizon setting and show that the structure of the optimal policy remains similar.

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

Similar content being viewed by others

References

  • Agra, A., & Constantino, M. (1999). Lot sizing with backordering and start-ups: the case of Wagner-Whitin costs. Operations Research Letters, 25, 81–88.

    Article  Google Scholar 

  • Benjaafar, S., & Yu, Y. (2012, forthcoming). Optimal policies for inventory systems with concave ordering costs. Operations Research.

  • Bensoussan, A., Moussawi-Haidar, L., & Cakanyildirim, M. (2010). Inventory control with an order-time constraint: optimality, uniqueness and significance. Annals of Operations Research, 181(1), 603–640.

    Article  Google Scholar 

  • Berk, E., Toy, A. O., & Hazir, O. (2008). Single item lot-sizing problem for a warm/cold process with immediate lost sales. European Journal of Operational Research, 187, 1251–1267.

    Article  Google Scholar 

  • Bigham, P. (1986). Economic order quantities for systems with step function ordering costs. Production and Inventory Management, 27(4), 119–127.

    Google Scholar 

  • Caliskan-Demirag, O., Chen, F. Y., & Yi, Y. (2012). Ordering policies in periodic-review inventory systems with quantity-dependent fixed costs. Operations Research. doi:10.1287/opre.1110.1033.

    Google Scholar 

  • Chao, X. L., & Zipkin, P. (2008). Optimal policy for a periodic-review inventory system under a supply capacity contract. Operations Research, 56, 59–68.

    Article  Google Scholar 

  • Chen, F., & Song, F. (2001). Optimal policies for multiechelon inventory problems with Markov-modulated demand. Operations Research, 49(2), 226–234.

    Article  Google Scholar 

  • Chen, X., & Simchi-Levi, D. (2004). Coordinating inventory control and pricing strategies with random demand and fixed ordering cost: the finite horizon case. Operations Research, 52(6), 887–896.

    Article  Google Scholar 

  • Fox, E. J., Metters, R., & Semple, J. (2006). Optimal inventory policy with two suppliers. Operations Research, 54(2), 389–393.

    Article  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, 21–41.

    Article  Google Scholar 

  • Gallego, G., & Scheller-Wolf, A. (2000). Capacitated inventory problems with fixed order costs: some optimal policy structure. European Journal of Operational Research, 126(3), 603–613.

    Article  Google Scholar 

  • Ha, A. Y. (1997). Inventory rationing in a make-to-stock production system with several demand classes and lost sales. Management Science, 43(8), 1093–1103.

    Article  Google Scholar 

  • Henig, M., Gerchak, Y., Ernst, R., & Pyke, F. (1997). An inventory model embedded in designing a supply contract. Management Science, 43(2), 184–189.

    Article  Google Scholar 

  • Lee, J.-Y. (2008). Quantity discounts based on the previous order in a two-period inventory model with demand uncertainty. Journal of the Operational Research Society, 59(7), 1004–1011.

    Article  Google Scholar 

  • Lippman, S. A. (1969). Optimal inventory policy with subadditive ordering costs and stochastic demands. SIAM Journal on Applied Mathematics, 17(3), 543–559.

    Article  Google Scholar 

  • Porteus, E. (1971). On the optimality of generalized (s,S) policies. Management Science, 17(7), 411–426.

    Article  Google Scholar 

  • Porteus, E. (1972). The optimality of generalized (s,S) policies under uniform demand densities. Management Science, 18(11), 644–646.

    Article  Google Scholar 

  • Porteus, E. (2002). Foundations of stochastic inventory theory. Stanford: Stanford University Press.

    Google Scholar 

  • Robinson, P., & Sahin, F. (2001). Economic production lot-sizing with periodic costs and overtime. Decision Sciences, 32(3), 423–452.

    Article  Google Scholar 

  • Scarf, H. (1960). The optimality of (s,S) policy in the dynamic inventory problem. In K. J. Arrow, S. Karlin & P. Suppes (Eds.), Mathematical methods in the social sciences (pp. 196–200). Stanford: Stanford University Press.

    Google Scholar 

  • Shaoxiang, C. (2004). The infinite horizon periodic review problem with setup costs and capacity constraints: a partial characterization of the optimal policy. Operations Research, 52(3), 409–421.

    Article  Google Scholar 

  • Simchi-Levi, D., Chen, X., & Bramel, J. (2004). The logic of logistics: theory, algorithms, and applications for logistics and supply chain management (2nd ed.). New York: Springer.

    Google Scholar 

  • Toy, A. O., & Berk, E. (2006). Dynamic lot sizing problem for a warm/cold process. IIE Transactions, 38(11), 1027–1044.

    Article  Google Scholar 

  • Xiao, W. (2007). Saleforce incentives: matching supply and demand under information asymmetry. Ph.D. Thesis. Available at: www.proquest.com.

  • Zipkin, P. (2008). On the structure of lost-sales inventory models. Operations Research, 56(4), 937–944.

    Article  Google Scholar 

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

    Article  Google Scholar 

Download references

Acknowledgements

The authors would like to thank Guest Editor Jian Yang and two anonymous referees for their constructive comments and suggestions, which helped to improve the content and presentation of this article.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Yi Yang.

Appendix

Appendix

1.1 A.1 Proofs of lemmas

Proof of Lemma 2

We let S 1 and S 2 denote the minimizers of J 1(x) and J 2(x), respectively. (If they are not unique, we take the largest one.) Then, we have either K 2+J 1(S 1)≤J 2(S 2) or K 2+J 1(S 1)>J 2(S 2).

Case I: Suppose K 2+J 1(S 1)≤J 2(S 2). We prove the quasi-K 1-convexity of J(x) by showing that the function is quasi-K 1-convex with a changeover at S 1R. First, we analyze the interval [−∞,S 1R]. For any xS 1R, min yx+R {K 2+J 1(y)}=K 2+J 1(S 1) and hence, we need to compare K 2+J 1(S 1) and min x+R>yx {K 2 δ(yx)+J 2(y)} to determine J(x). Note that min x+R>yx {K 2 δ(yx)+J 2(y)}>J 2(S 2), because S 2 is a minimizer of J 2(x). Further, since K 2+J 1(S 1)≤J 2(S 2), it follows that J(x)=K 2+J 1(S 1), which is a constant.

Next, we analyze the interval [S 1R,∞]. To establish the result that J(x) is non-K 1-decreasing on [S 1R,∞], we show that any x 1,x 2∈[S 1R,∞] satisfying x 1x 2 imply J(x 1)≤J(x 2)+K 1. At the point x=x 2, we have either \(J(x_{2})=\min_{y\geq x_{2}+R}\{K_{2}+J_{1}(y)\}\) or \(J(x_{2})=\min_{x_{2}+R>y\geq x_{2}}\{K_{2}\delta(y-x_{2})+J_{2}(y)\}\).

In the first subcase, \(J(x_{1})\leq\min_{y\geq x_{1}+R}\{K_{2}+J_{1}(y)\}\leq\min_{y\geq x_{2}+R}\{K_{2}+J_{1}(y)\}= J(x_{2})\leq J(x_{2})+K_{1}\), where the first inequality follows from the definition of J(x), the second inequality is due to fewer choices since x 1x 2, and the third equality holds since \(J(x_{2})=\min_{y\geq x_{2}+R}\{K_{2}+J_{1}(y)\}\) in this case.

In the second subcase, \(J(x_{2})=\min_{x_{2}+R>y\geq x_{2}}\{K_{2}\delta(y-x_{2})+J_{2}(y)\}\). It suffices to consider the case where x 2 is the minimizer of J 2(y) on [x 2,x 2+R), since this will imply the result for any other x minimizing J 2(y) in this region. Because x 2 is the minimizer of J 2(y) on [x 2,x 2+R), we have J(x 2)=J 2(x 2). Given x 1x 2, we have either x 1+Rx 2 or x 1+R>x 2. In the former case, we have \(J(x_{1})\leq\min_{y\geq x_{1}+R}\{K_{2}+J_{1}(y)\}\leq K_{2}+J_{1}(x_{2})\leq K_{1}+J_{2}(x_{2})=K_{1}+J(x_{2})\), where the first and second inequalities follow from the definition of J(x) and minimization, respectively, the third inequality follows from the assumptions stated in the lemma, and the last equality holds since J(x 2)=J 2(x 2) in this case. Using similar arguments for the case with x 1+R>x 2, we can verify that \(J(x_{1})\leq\min_{x_{1}+R> y\geq x_{1}}\{K_{2}\delta(y-x_{1})+J_{2}(y)\}\leq K_{2}+J_{2}(x_{2})\leq K_{1}+J_{2}(x_{2})=K_{1}+J(x_{2})\). Hence, we can conclude that J(x) is quasi-K 1-convex under Case I.

Case II: Suppose K 2+J 1(S 1)>J 2(S 2). Then, it follows that the minimizer of J(x) is S 2, i.e., J(S 2)=J 2(S 2). In this case, we prove the quasi-K 1-convexity of J(x) by showing that there exists a changeover point bS 2 such that the function is decreasing (constant) in (−∞,b] and non-K 1-decreasing in [b,+∞].

We first show that J(x) is non-K 1-decreasing in [S 2,+∞). Let x 1,x 2∈[S 2,+∞) such that x 1x 2. At the point x=x 2, we have either \(J(x_{2})=\min_{x_{2}+R>y\geq x_{2}}\{K_{2}\delta(y-x_{2})+J_{2}(y)\}\) or \(J(x_{2})=\min_{y\geq x_{2}+R}\{K_{2}+J_{1}(y)\}\). In the first case, we consider the case where x 2 is the minimizer of J 2(y) on [x 2,x 2+R) (similar to the analysis in Case I), and find J(x 2)=J 2(x 2). Then, we have J(x 1)≤J 2(x 1)≤J 2(x 2)+K 1=J(x 2)+K 1, where the second inequality follows from the quasi-K 1-convexity of J 2(x). In the second case, \(J(x_{2})=\min_{y\geq x_{2}+R}\{K_{2}+J_{1}(y)\}\). Then, by using the definition of J(x) and the relationship x 1x 2, we can verify \(J(x_{1})\leq \min_{y\geq x_{1}+R}\{K_{2}+J_{1}(y)\}\leq\min_{y\geq x_{2}+R}\{K_{2}+J_{1}(y)\}= J(x_{2})\leq J(x_{2})+K_{1}\), and the non-K 1-decreasing property holds.

Next, we show the existence of the changeover point b. Let a be the smallest number such that K 2+J 1(S 1)>J 2(a). The existence of a number a<S 2 is assured by K 2+J 1(S 1)>J 2(S 2). Thus, for any xa, K 2+J 1(S 1)≤J 2(x). We consider two cases: aS 1R and a>S 1R.

If aS 1R, then we set b=a. Notice that for any xb(=a), K 2+J 1(S 1)≤J 2(x), hence it follows that J(x)=K 2+J 1(S 1), that is, J(x) is a constant in (−∞,b]. Moreover, for any x 1,x 2∈[b,S 2] and x 1x 2, it follows from the definition of J(x) and the quasi-K 1-convexity of J 2(x) that J(x 1)≤J 2(x 1)≤K 1+J 2(S 2)=K 1+J(S 2)≤K 1+J(x 2), where the last inequality holds, because S 2 is the minimizer of J(x). Hence, J(x) is non-K 1-decreasing on [b,S 2].

If a>S 1R, then we set b=S 1R. Then, for any xb=S 1R<a, K 2+J 1(S 1)≤J 2(x). Hence, it follows that J(x)=K 2+J 1(S 1), i.e., a constant in (−∞,b]. For any x∈[b,S 2], if xS 2R, then J(x)≤J 1(S 2)+K 2J 2(S 2)+K 1=K 1+J(S 2); otherwise, J(x)≤K 2+J 2(S 2)≤J 2(S 2)+K 1=K 1+J(S 2). So, J(x) is non-K 1-decreasing in [b,S 2].

Given that J(x) is non-K 1-decreasing on [b,S 2] and [S 2,+∞], we can easily show that it is non-K 1-decreasing on [b,+∞]: for any z 1∈[b,S 2] and z 2∈[S 2,+∞), we have J(z 2)+K 1J(S 2)+K 1J(z 1), which follows since S 2 is a minimizer and J(x) is non-K 1-decreasing.

Thus far, we have proved the quasi-K 1-convexity. Notice that in all cases, J(x) is a constant on the left of the changeover point and non-K 1-decreasing on the right. Hence, we can conclude that J(x) is non-K 1-decreasing in ℝ. □

Proof of Lemma 3

We first prove parts (i) and (ii) by considering two separate cases: (a) f t,2(x)=min yx+R {g t,2(y)} and (b) f t,2(x)=min x+R>yx {g t,1(y)}.

(i) For case (a), we have f t,1(x)≤min yx+R {K+g t,2(y)}=K+f t,2(x). For case (b), f t,1(x)≤min x+R>yx {(yx)+g t,1(y)}≤K+min x+R>yx {g t,1(y)}=K+f t,2(x). The second inequality in the lemma can be easily verified. From the results of part (i), we can also conclude that g t,1(x)≥g t,2(x).

(ii) For case (a), \(f_{t,1}(x_{1})\leq\min_{y\geq x_{1}+R}\{K+g_{t,2}(y)\}\leq\min_{y\geq x_{2}+R}\{K+g_{t,2}(y)\}=K+f_{t,2}(x_{2})\), where the second inequality follows from that x 2>x 1. For case (b), let f t,2(x 2)=g t,1(y(x 2)) where x 2+R>y(x 2)≥x 2. If y(x 2)≤x 1+R, then f t,1(x 1)≤K+g t,1(y(x 2))=K+f t,2(x 2). If y(x 2)>x 1+R, then \(\min_{y\geq x_{1}+R}\{K+g_{t,1}(y)\}\leq K+g_{t,1}(y(x_{2}))=K+f_{t,2}(x_{2})\). From part (i), we have g t,1(x)≥g t,2(x) which implies that \(\min_{y\geq x_{1}+R}\{K+g_{t,2}(y)\}\leq\min_{y\geq x_{1}+R}\{K+g_{t,1}(y)\}\leq K+f_{t,2}(x_{2})\). Thus, we have \(f_{t,1}(x_{1})\leq\min_{y\geq x_{1}+R}\{K+g_{t,2}(y)\}\leq K+f_{t,2}(x_{2})\).

(iii) Again, consider two cases: \(f_{t,1}(x_{2})=\min_{y>x_{2}+R}\{K+g_{t,2}(y)\}\) and \(f_{t,1}(x_{2})=\min_{x_{2}+R\geq y\geq x_{2}}\{K\delta(y-x)+g_{t,1}(y)\}\). In the first case, \(f_{t,1}(x_{1})\leq\min_{y>x_{1}+R}\{K+g_{t,2}(y)\}\leq \min_{y>x_{2}+R}\{K+g_{t,2}(y)\}=f_{t,1}(x_{2})\). In the second case, we have

where the second inequality follows due to the fact that g t,2(x)≤g t,1(x) from part (i) and since x 1x 2. Using similar arguments, we can also prove that f t,2(x 1)≤K+f t,2(x 2). □

Proof of Lemma 4

We prove the result by considering three cases.

Case I

\(y^{*}_{t,1}(x_{t})=x_{t}\). Then, \(y^{*}_{t,2}(x_{t})\geq x_{t}= y^{*}_{t,1}(x_{t})\).

Case II

\(x_{t}<y^{*}_{t,1}(x_{t})<x_{t}+R\). If \(y^{*}_{t,2}(x_{t})\geq x_{t}+R\), then the result clearly holds. Referring to (4), the condition \(x_{t}<y^{*}_{t,1}(x_{t})<x_{t}+R\) implies that \(g_{t,1}(y^{*}_{t,1}(x_{t}))\leq g_{t,1}(y)\), for any x t y<x t +R. Therefore, if \(y^{*}_{t,2}(x_{t})< x_{t}+R\), it can be seen from (5) that \(y^{*}_{t,2}(x_{t})=y^{*}_{t,1}(x_{t})\).

Case III

\(y^{*}_{t,1}(x_{t})\geq x_{t}+R\). It implies the following two statements: (i) \(g_{t,2}(y^{*}_{t,1}(x_{t}))\leq g_{t,1}(y)\), for any x t y<x t +R, and (ii) \(g_{t,2}(y^{*}_{t,1}(x_{t}))\leq g_{t,2}(y)\), for any x t +Ry. By these two statements, we can obtain \(y^{*}_{t,2}(x_{t})=y^{*}_{t,1}(x_{t})\).  □

Proof of Lemma 5

Parts (i) and (ii) follow from the definitions of quasi-K-convexity and the critical points.

(iii) Since L(x) is increasing on xS L , L(x 2)≥L(x 1). Then, using Lemma 3(ii), we have g t,1(x 1)=L(x 1)+αE[f t+1,1((x 1D t )+)]≤L(x 2)+αE[f t+1,2((x 2D t )+)]+Kg t,2(x 2)+K.

(iv) By Lemma 3(i), g t,2(x)≤g t,1(x) which implies g t,2(S t,2)≤g t,1(S t,1), because S t,2 is a minimizer of g t,2(x). Thus, we have g t,1(s t,3)<K+g t,2(S t,2)≤K+g t,1(S t,1), which by definition, implies s t,1s t,3.

By Lemma 3(i), g t,1(S t,i )<K+g t,2(S t,2). Thus, by the definition, S t,i >s t,3 for i=1,2.

If S L S t,2, clearly s t,3<S t,2S L . Otherwise by Lemma 3(i), g t,1(S L )<K+g t,2(S t,2), which implies s t,3<S L .

By definition, g t,1(s t,3)<K+g t,2(S t,2)≤K+min yx+R g t,2(y), which implies s t,4s t,3.

As g t,1(S t,1)<g t,2(S t,2) and S t,2R∈[S t,1R,s t,1], we have s t,5>S t,2R. □

Proof of Lemma 6

(i) To prove the result, it suffices to show that g t,i (x)≥g t,i (S L ) for all \(x\geq \hat{S}_{L}>S_{L}\). Applying Lemma 3(iii), we can obtain

(ii) To verify \(s_{t,i}\geq\hat{s}_{L}\), we first prove that for t=1,2,…,T and i=1,2,

  1. (a)

    f t,i (x t ) is a non-increasing function for x t S L ;

  2. (b)

    g t,i (y t ) is a non-increasing function for y t S L ;

  3. (c)

    g t,i (y t )≥g t,i (S t,i )+K for any \(y_{t}\leq\hat{s}_{L}\).

We prove these results by induction. It is clear that they hold for t=T. Suppose that the results hold for period t+1. Because f t+1,i (x t ) and L(x t ) are non-increasing for x t S L , g t,i (y t ) is also non-increasing for y t S L ,

This completes the proof. □

1.2 A.2 Feasibility analysis of K 1-convexity (Sect. 3)

We show by contradiction that K 1-convexity is not preserved in Lemma 2. Assume that K 2=K 1. Suppose J 1(x) and J 2(x) have the same minimizer denoted by S. Then, by assumption J 1(S)<J 2(S). Let s t,1=min{x|J 1(x)=J 1(S)+K 1} and s t,2=min{x|J 2(x)=J 2(S)+K 1}. First, select R such that SR<min{s t,1,s t,2}. Then for any xSR, J(x)=K 1+J 1(S). Next, suppose J 1(x) and J 2(x) are both strictly increasing in (S,+∞) and K 1+J 1(S+R)>J 2(S), then J(S)=J 2(S). Finally, note that by construction Ss t,2<R, and thus J(s t,2)=min{J 1(s t,2+R)+K 1,K 1+J 2(S)}. Since we assumed that J 1(x) is strictly increasing in (S,+∞), we have J 1(s t,2+R)≥J 2(S). Consequently, J(s t,2)=K 1+J 2(S). Now we can pick three points x=SR, s t,2, and S with the values J(x)=K 1+J 1(S), K 1+J 2(S) and J 2(S), respectively. Recall that a function f(x) is K 1-convex if, for any x 1x 2 and λ∈(0,1], f(λx 1+(1−λ)x 2)≤λf(x 1)+(1−λ)(f(x 2)+K). For K 1-convexity to hold in this example, we need to show that for any λ∈(0,1], J(s t,2)≤λJ(SR)+(1−λ)(J(S)+K 1), which implies J 2(S)≤λJ 1(S)+(1−λ)J 2(S). However, since we have J 1(S)<J 2(S), we have J 2(S)>λJ 1(S)+(1−λ)J 2(S), which contradicts with K 1-convexity.

1.3 A.3 Infinite horizon case

Similar to the finite-horizon model, we assume that the demands in different periods are realizations from independent and identical distributions. We show that the T-period optimal cost function f t,1 and f t,2 both converge uniformly to finite-valued functions f 1 and f 2, respectively. This implies that the optimal policy for the infinite-horizon problem inherits the structure of the T-period problem.

Theorem 3

The optimal cost functions f t,1 and f t,2 in the T-period problem converge uniformly to functions f 1 and f 2 in any finite interval.

Proof

Denote by y t (x) the optimal inventory level after ordering when the initial inventory level is x. By Lemma 6, it is easy to prove that there exists a constant M such that y t (x)∈[0,M] for all x∈[0,M] independent of t.

Note the fact that for any constant c 1, c 2, c 3 and c 4, we have

$$\bigl|\min\{c_1,c_2\}-\min\{c_3,c_4 \}\bigr|\leq \min\bigl\{|c_1-c_3|,|c_2-c_3|,|c_2-c_3|,|c_2-c_4| \bigr\}. $$

Thus, we can obtain

where the fist inequality follows from the above fact and the second from that y t (x)∈[0,M] for all x∈[0,M] independent of t.

Since M can be chosen arbitrarily large, we have shown that f t,i (x) converges monotonically and uniformly for all x in any finite interval. The functions f t,i (x) are continuous and converge uniformly, thus the limit function f i (x) is also continuous. □

Let the functions f 1(x) (f 2(x)) represent the minimum total expected discounted cost of the infinite horizon when the current system state is cold (warm). It follows from the theory of Markov decision processes that

and

$$f_{2}(x_t)=\min\Bigl\{\min_{y>x+R}\bigl\{L(y)+ \alpha Ef_{2}\bigl((y-D)^+\bigr)\bigr\}, \min_{x+R\geq y \geq x}\bigl \{L(y)+\alpha Ef_{1}\bigl((y-D)^+\bigr)\bigr\}\Bigr\}. $$

The convergence guarantees that the optimal policy for the infinite-horizon case takes the same structure as that of the finite horizon case. While the optimal policy can be typically simplified under an infinite horizon, we are not able to do so in our model because none of the finite-horizon decision regions can be eliminated from the policy.

1.4 A.4 OB heuristic policy

The OB heuristic policy is shown in Table 6.

Table 6 OB heuristic policy

Rights and permissions

Reprints and permissions

About this article

Cite this article

Caliskan-Demirag, O., Chen, Y.(. & Yang, Y. Production-inventory control policy under warm/cold state-dependent fixed costs and stochastic demand: partial characterization and heuristics. Ann Oper Res 208, 531–556 (2013). https://doi.org/10.1007/s10479-012-1187-1

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10479-012-1187-1

Keywords

Navigation