essays on perishable inventory management

157
Essays on Perishable Inventory Management by Borga Deniz Submitted to the Tepper School of Business in Partial Fulfillment of the Requirements for the Degree of DOCTOR OF PHILOSOPHY at the CARNEGIE MELLON UNIVERSITY c Carnegie Mellon University 2007 Thesis Committee Professor Alan Scheller-Wolf (chair) Professor Itır Z. Karaesmen-Aydın Professor Laurens Debo Professor Peter Boatwright 1

Upload: lambao

Post on 02-Jan-2017

220 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Essays on Perishable Inventory Management

Essays on Perishable InventoryManagement

by

Borga Deniz

Submitted to the Tepper School of Businessin Partial Fulfillment of the Requirements for the Degree of

DOCTOR OF PHILOSOPHY

at the

CARNEGIE MELLON UNIVERSITYc© Carnegie Mellon University 2007

Thesis Committee

Professor Alan Scheller-Wolf (chair)Professor Itır Z. Karaesmen-AydınProfessor Laurens DeboProfessor Peter Boatwright

1

Page 2: Essays on Perishable Inventory Management

Abstract

My dissertation addresses a number of challenges in perishable inventory managementin supply chains, providing both theoretical results and managerial insights.

Essay 1: In this essay, we review the literature on supply chain management ofperishable products that have a fixed or random lifetime, primarily focusing on chal-lenges introduced by aging of such products. We start with a review of the literatureon capacity and production planning. We then review the literature on inventorymanagement at a single location. The research on single-location models providesa basis for research on multi-echelon/multi-location models, that are also reviewedin this chapter, along with transshipment and distribution models. The chapter alsocovers research involving modeling novelties, in terms of demand-fulfillment, informa-tion sharing, and centralized vs. decentralized planning. We provide several directionsfor future research by identifying the gaps in the literature and the needs of industrieswhere perishable products are commonly stocked.

Essay 2: We consider a discrete-time supply chain for perishable goods where thereare separate demand streams for items of different ages. We compare two practicalreplenishment policies: replenishing inventory according to order-up-to level policiesbased on either (i) total inventory in system or (ii) new items only. Given thesepolicies, we concentrate on four different ways of fulfilling demand: (1) demand foran item can only be satisfied by an item of that age (No-Substitution); (2) demandfor new items can only be satisfied by new ones, but excess demand for old itemscan be satisfied by new (Downward-Substitution); (3) demand for old items canonly be satisfied by old, but excess demand for new items can be satisfied by old(Upward-Substitution); (4) both downward and upward substitution are employed(Full-Substitution). We compare these substitution options analytically in terms oftheir infinite horizon time-average costs, providing conditions on cost parameters thatdetermine when (if at all) one substitution option is more profitable than the others,for an item with a two-period lifetime. We also prove that inventory is “fresher”whenever downward substitution is employed. Our results are based on sample-pathanalysis, and as such make no assumptions on the joint distribution of demand, savefor ergodicity. We complement our results with numerical experiments exploring theeffect of problem parameters on performance.

2

Page 3: Essays on Perishable Inventory Management

Essay 3: In this paper we study replenishment policies for perishable goods whensubstitution is possible between items of different ages. We analyze different order-up-to level type replenishment policies and compare their performance with each otheras well as the optimal policy. We provide analytical results on optimality of ourproposed policies for some special cases. Also structural properties of our policiesare investigated. A comprehensive computational study is conducted using a variousdemand distributions and parameter settings. Based on our study we find that NIS(order-up-to inventory replenishment policy based on new items only) is closer tooptimal compared to TIS (order-up-to inventory replenishment policy based on totalinventory). We also observe that when items are significantly more valuable thanolder items NIS is more robust than TIS against demand fluctuations.

Page 4: Essays on Perishable Inventory Management

To my family.

4

Page 5: Essays on Perishable Inventory Management

Acknowledgements

I would like to thank my family for everything they have done for me throughout my

life.

I would like to thank my advisors Dr. Alan Scheller-Wolf and Dr. Itır Karaesmen-

Aydın for their invaluable guidance and support. I am truly grateful to them.

I would like to thank Dr. Laurens Debo and Dr. Peter Boatwright for agreeing to serve

on my committee by sparing their valuable time and for providing helpful comments

and suggestions.

Last, but certainly not least, I would like to thank all my friends. They have made

my life very enjoyable during my studies.

5

Page 6: Essays on Perishable Inventory Management

Contents

1 Introduction 1

2 Managing Perishable Inventory in Supply Chains 6

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2.2 Challenges in Production Planning . . . . . . . . . . . . . . . . . . . 8

2.3 Managing Inventories at a Single Location . . . . . . . . . . . . . . . 10

2.3.1 Discrete Review Models . . . . . . . . . . . . . . . . . . . . . 10

2.3.2 Continuous Review Models . . . . . . . . . . . . . . . . . . . . 18

2.4 Managing Multi-Echelon and Multi-Location Systems . . . . . . . . . 29

2.4.1 Research on Replenishment and Allocation Decisions . . . . . 30

2.4.2 Logistics: Transshipments, Distribution, and Routing . . . . . 39

2.4.3 Information Sharing and Centralized/Decentralized Planning . 41

2.5 Modeling Novelties: Demand and Product Characteristics, Substitu-

tion, Pricing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

2.5.1 Single Product and Age-Based Substitution . . . . . . . . . . 46

2.5.2 Multiple Products . . . . . . . . . . . . . . . . . . . . . . . . . 48

2.5.3 Pricing of Perishables . . . . . . . . . . . . . . . . . . . . . . . 49

2.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

3 Effect of Substitution on Management of Perishable Goods 51

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

3.2 Problem Definition and Formulation . . . . . . . . . . . . . . . . . . 54

3.3 Related Work in the Literature . . . . . . . . . . . . . . . . . . . . . 59

3.4 Effect of Substitution: Analysis of Economic Benefit and Freshness . 64

3.4.1 The economic benefit of substitution: Overview of results . . . 65

6

Page 7: Essays on Perishable Inventory Management

3.4.2 Economic Benefit of Substitution under TIS . . . . . . . . . . 68

3.4.3 Economic Benefit of Substitution under NIS . . . . . . . . . . 75

3.4.4 A note on substitution not being a recourse . . . . . . . . . . 77

3.4.5 Freshness of Inventory . . . . . . . . . . . . . . . . . . . . . . 77

3.5 Computational Results . . . . . . . . . . . . . . . . . . . . . . . . . . 79

3.5.1 Non-convexity/Multi-modality of the Cost Function . . . . . . 79

3.5.2 Making downward substitution more attractive . . . . . . . . 81

3.5.3 Costs, Service Levels and Freshness of Inventory . . . . . . . . 83

3.5.4 Effect of customer behavior: Proportional acceptance of sub-

stitutes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85

3.5.5 Summary of Computational Results . . . . . . . . . . . . . . . 87

3.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88

4 Analysis of Inventory Replenishment Policies for Perishable Goods

Under Substitution 89

4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

4.2 Analytical Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

4.3 Computational Results: NIS vs. TIS . . . . . . . . . . . . . . . . . . 91

4.3.1 A Preliminary Study . . . . . . . . . . . . . . . . . . . . . . . 92

4.3.2 Comparison of TIS and NIS while changing variance of demand 93

4.3.3 NIS vs. TIS when new and old item demands are negatively

correlated . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97

4.3.4 TIS vs. NIS when new and old item demands are positively

correlated . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98

4.3.5 NIS vs. TIS when new item demand is higher than old item

demand . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100

4.3.6 NIS vs. TIS when new item demand is lower than old item

demand . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102

4.3.7 Comparison of TIS and NIS when substitution is not always

accepted: Partial vs. Probabilistic Acceptance . . . . . . . . . 103

4.3.8 Comparison of TIS and NIS when acceptance probability of

substitution depends on substitution costs . . . . . . . . . . . 104

7

Page 8: Essays on Perishable Inventory Management

4.3.9 Search for a good NIS order-up-to level: News-vendor-type

heuristic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

4.3.10 Summary of the Computational Study for Comparison of NIS

and TIS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

5 Conclusion and Future Research 110

References 115

Technical Appendix 132

8

Page 9: Essays on Perishable Inventory Management

List of Figures

3.1 Time-average costs of substitution policies in Example-1A. . . . . . . 80

3.2 Time-average costs of substitution policies in Example-1B. . . . . . . 80

3.3 Time-average costs of substitution policies in Example-2. . . . . . . . 81

3.4 Minimum expected cost of policies as a function of downward substi-

tution cost (αD) in Example-3. . . . . . . . . . . . . . . . . . . . . . 82

3.5 The effect of proportional acceptance of downward substitution on ex-

pected total cost in Example-4. . . . . . . . . . . . . . . . . . . . . . 87

3.6 The effect of proportional acceptance of upward substitution on ex-

pected total cost in Example-5. . . . . . . . . . . . . . . . . . . . . . 88

4.1 Computation times for the best NIS, the best TIS and the optimal

solutions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95

4.2 Optimal policies when CV is 0.58 . . . . . . . . . . . . . . . . . . . . 96

4.3 Optimal policies when CV is 1.41 . . . . . . . . . . . . . . . . . . . . 96

4.4 Optimal policies when CV is 0.12 . . . . . . . . . . . . . . . . . . . . 96

4.5 Optimal policies when CV is 0.28 . . . . . . . . . . . . . . . . . . . . 96

4.6 The metric shows where the downward substitution cost (αD) lays

compared to upper and lower bounds . . . . . . . . . . . . . . . . . . 97

9

Page 10: Essays on Perishable Inventory Management

Chapter 1

Introduction

Most products lose their market value (outdate) over time. Some products lose value

faster than others; these are known as perishable products. Traditionally, perishables

outdate due to their chemical structure. Examples of such perishable products are

fresh produce, blood products, dairy products, meat, drugs, vitamins etc. Today

many products that are not perishable in the traditional sense (i.e. products that do

not decay chemically over time) can still be considered as perishable. Such prod-

ucts, mostly high-tech, outdate because of changes in market conditions. Personal

computers, computer components (such as micro-processors, memory, data storage

units), cellular phones, digital cameras, digital music players, personal digital as-

sistants (PDAs) are examples of high-tech products that rapidly lose market value.

The life cycles of such products are getting shorter every year due to technological

advances. The co-founder of Intel, Gordon Moore, stated (Moore, 1965) “the number

of transistors on an integrated circuit for minimum component cost doubles every 24

months.” This statement was later coined as Moore’s Law. In other words Moore’s

Law implies that every two years consumers can expect to buy personal computers

that are twice faster, with no price increase. Data storage technology advanced at

even a faster rate than processing technology. According to Mark Kryder, founder

and director of Carnegie Mellon University’s Data Storage Systems Center and chief

technology officer at hard-drive manufacturer Seagate Technology, since the introduc-

tion of the disk drive in 1956, the density of data it can record saw a 50-million-fold

increase (Walter, 2005). Because of these reasons, for a certain configuration, the cost

of personal computers are expected to drop by fifty percent every 18 to 24 months.

1

Page 11: Essays on Perishable Inventory Management

In addition to developments in technology, an important factor on shrinking life

cycles is competition. High-tech companies compete to introduce their newest prod-

ucts earlier to improve their market share. In the past many companies were not

under the intense competition that exists today. Therefore they did not feel pres-

sure to introduce their newest products until the market was saturated with existing

models. However at the present time there is cutthroat competition in almost all

high-tech consumer product markets. Customers benefit from this competition be-

tween companies. They are becoming more sophisticated than ever and demand more

from products that they buy.

One can argue that high-tech products are not perishable products. With the

acceleration of product life cycles, the line between ‘perishable’ and ‘durable’ prod-

ucts continues to blur. Strictly speaking, goods such as computers and cell phones

obsolesce rather than perish. By definition perishability occurs at a predictable rate

as products change; and obsolescence happens less predictably as the market environ-

ment changes. But the time frame for obsolescence in the traditional sense is longer

than the time frame for high-tech products’ obsolescence; thus we feel the need to

make a distinction between traditional obsolescence and the current situation. We

maintain that rapid obsolescence of high-tech products can be considered as perisha-

bility, as high-tech products are becoming obsolete at an increasingly faster pace than

ever.

Because of the importance of managing perishables we identified several interesting

research problems by carefully studying the literature. A comprehensive literature

study of perishable inventory research is the next chapter, in which we classify the

research on perishables that has been done focusing on the research on perishables

since Nahmias’ (1982) excellent review. This chapter fills the need for a comprehensive

and up-to-date review of research on managing perishable inventory in the area of

operations management, especially a review that can show the recent trends and

point out important future research directions from the perspective of operations

management and supply chain management. We concentrate on the research done

mainly on stochastic inventory management and on those papers which, in our view,

are important and lay the foundation for future work in one of the directions we detail.

We also refer to some papers on non-perishable goods in supply chain management

literature to put the research in perspective.

2

Page 12: Essays on Perishable Inventory Management

In the third chapter we study the effect of substitution between products at dif-

ferent ages. Managing perishable inventory in supply chains is challenging especially

when products of different “ages” co-exist in the market at the same time and there

is a possibility of substitution between old and new products. In our problem, there

is a single product with a lifetime of two periods; the value of the product decreases

deterministically as it ages, and there may be random demand for both new and

old items. A single supplier replenishes new items periodically, with zero lead time.

At the end of each period, any remaining old items are outdated, while any unused

new items become old. This setup captures the essence of the problem for perishable

goods such as fresh produce and blood as well as durable goods where a new product

is introduced into the market while the old one is phased-out periodically.

In our model all demand for new (old) items is fulfilled from the inventory of

new (old) unless there is a stock-out. In case of stock-outs, the excess stock of new

(old) items may be used to satisfy the excess demand of old (new). We use the

term downward substitution to denote the case where a new product is sold to a

customer that demands old; upward substitution is the reverse. We assume any such

substitution occurs only as a recourse. Our model captures the situation in which the

supplier is the one making the substitute offer; this is supplier-driven substitution as

is common in business-to-business situations. On the other hand, our model likewise

captures the scenario in which the consumer is the one considering the substitute

option without explicit intervention (except product placement) by the supplier. This

is the same as inventory-driven substitution where the demand of one product shifts

to another in case of a stock-out.

In terms of customer choice we do not have any restrictions on upward substi-

tution in our analysis, that is a customer can refuse the upward substitution and

our analytical results will not be affected. However for downward substitution, we

assume that customer and seller will always find a price that both can agree upon,

and the downward substitution will take place. This price would probably be between

the original selling prices of old and new items. In our model we have a downward

substitution cost that could be considered the result of such an agreement between

the parties.

Our work is first in the literature on substitution to introduce the term “dynamic

quality”. In the literature on substitution, the quality of a product is static. For our

3

Page 13: Essays on Perishable Inventory Management

problem the quality of a product is age (one dimensional). That is, a new product

has a higher quality than an old one. Since our model is an infinite horizon model,

the quality of the product is dynamic. If a product is not sold its quality will change

in the next period (i.e. from old to new) making our model unique in the literature

on substitution.

Our model best applies to traditional perishable products such as fresh produce

or blood rather than the non-traditional perishable products such as computer chips.

The main reason for this is due to the fact that the supply of traditional perishable

products is mostly done with the newest products. However for computer chips old

and new products can be supplied simultaneously. Nevertheless, insights from our

model can still be of use in the non-traditional setting.

Our model also encompasses other phenomenon: If inventory is depleted starting

from the newest items (e.g. buyers prefer longer shelf-life items as in grocery and the

freshest produce/bread), this corresponds to having demand for new items only and

fulfilling the demand by upward substitution (i.e. customers consider buying less fresh

products only after freshest is sold). If instead inventory is depleted in a FIFO fashion

(oldest item issued first) as is the assumption in classical research on perishable goods

(see the review in the next section), then there is demand for only old items and

demand is fulfilled using downward substitution.

For two different order-up-to level type replenishment policies we compare the

effect of substitution on system costs. We find conditions on cost parameters that

would guarantee cost dominance between policies over the infinite horizon. We also

provide results on freshness and service levels of policies as well.

In the fourth chapter we study replenishment policies for perishable goods under

substitution. We analyze inventory replenishment policies within the class of base-

stock policies. Two policies that bring “total” on-hand inventory to an order-up-to

level and “new” on-hand inventory to an order-up-level are compared, presenting the

results of a comprehensive numerical study. We compare our two heuristics with the

globally optimal policy and provide analytical results on optimality of our heuristics

for special cases.

We test our heuristic policies under various demand schemes, including four differ-

ent variance settings. In the third chapter we do not have any restrictions on demand

process in terms of correlation for our analytical results. That is, old and new item

4

Page 14: Essays on Perishable Inventory Management

demand streams can correlated with each other. Therefore we provide computational

experiments for negatively and positively correlated demand. We also test a rule

of thumb which is developed based on our observations for choosing a substitution

policy.

Our numerical experiments include a comparison of our heuristics when substitu-

tion is not always accepted; we do experiments with partial acceptance and probabilis-

tic acceptance. In the partial acceptance model when substitution is offered a certain

fraction of it is accepted by customer. Under probabilistic acceptance all of the offered

substitution is either accepted or rejected (as a whole) by the customer. We assume

each substitution acceptance is independent. We also develop and compare a num-

ber of NIS order-up-to level heuristics using newsvendor-type logic as well as TIS and

NIS at their optimal order-up-to levels (found by doing a line search using simulation)

versus the globally optimal state-dependent policy found via dynamic programming.

We conduct our numerical studies using simulation and dynamic programming.

5

Page 15: Essays on Perishable Inventory Management

Chapter 2

Managing Perishable Inventory in

Supply Chains

2.1 Introduction

In this chapter, we provide an overview of research in supply chain management of

products that are perishable, in particular, and that outdate, in general. As this

definition is quite broad, we focus on research done mainly on inventory manage-

ment in which a product ages over time, thus largely excluding single-period models

which are commonly used to represent perishable items (with an explicit cost attached

to expected future outdates). While these newsvendor-type models can convey im-

portant insights, they are often too simple to provide answers to some of the more

complex questions that arise as inventory levels, product characteristics, markets and

customer behaviors change over time. Furthermore, we exclude research that models

decay or deterioration of inventories assuming certain functional forms (e.g., exponen-

tial decay). The reader can refer to Goyal and Giri (2001) and Raafat (1991), which is

supplemented by Dave (1991), for a bibliography and classification of research on that

topic. Finally, we also exclude the research on management, planning or allocation

of capacity which is commonly referred as perishable inventory, for instance, in the

airline revenue management literature. Summarizing then, our emphasis is on models

where the inventory level of a product must be controlled over a horizon taking into

6

Page 16: Essays on Perishable Inventory Management

account demand, supply and a finite shelf-life (which may be fixed or random).

Over the years, several companies have emerged as exemplary of “best practices”

in supply chain management; for example, Wal-Mart is frequently cited as using

unique strategies to lead its market. One significant challenge for Wal-Mart is man-

aging inventories of products that frequently outdate: A significant portion of Wal-

Mart’s product portfolio consists of perishable products such as food items (varying

from fresh produce to dairy to bakery products), pharmaceuticals (e.g. drugs, vita-

mins, cosmetics), chemicals (e.g. household cleaning products), and cut flowers.

Of course Wal-Mart’s supply chain is not alone in its exposure to outdating risks

– to better appreciate the impact of perishability and outdating in society at large,

consider these figures: In a 2003 survey, overall unsalable costs at distributors to

supermarkets and drug stores within consumer packaged goods alone were estimated

at $2.57 billion, and 22% of these costs, over 500 million dollars, were due to expiration

in only the branded segment (Grocery Manufacturers of America, 2004). Within the

produce sector, the $1.7 billion U.S. apple industry is estimated to lose $300 million

annually to spoilage (Webb, 2006).

Note also that perishability and outdating are a concern not only for these con-

sumer goods, but for industrial products (for instance, Chen, 2006, mentions that

adhesive materials used for plywood lose strength within 7 days of production), mil-

itary ordnance, and blood - one of the most critical resources in health care supply

chains. According to a nationwide survey on blood collection and utilization, 5.8% of

all components of blood processed for transfusion were outdated in 2004 in the U.S.

(AABB, 2005).

Modeling in such an environment implies that at least one or both of the following

holds: First, demand for the product may change over time as the product ages; this

could be due to a decrease in the utility of the product because of the reduced lifetime,

lessened quality and/or changing market conditions. Second, operational decisions

can be made more than once (e.g., inventory can be replenished by ordering fresher

products, or prices can be marked down) during the lifetime of the product. Either

of these factors make the analysis of such systems a challenge, owing to the expanded

problem state space corresponding to the additional information that is both available

and valuable. Traditionally, it has required appreciable effort to gather and use such

information, but as we will see, with the advent of increasingly powerful information

7

Page 17: Essays on Perishable Inventory Management

and control technologies, new and exciting possibilities are emerging.

Our goal in this chapter is not to replicate surveys of past work (such as Nahmias,

1982, Prastacos, 1984, and Pierskalla, 2004; in fact we refer to them as needed in the

remainder of this chapter) but to discuss in more detail directions for future research.

In order to do that, we provide a selective review of the existing research and focus

on those papers which, in our view, constitute crucial stepping stones or point to

promising directions for future work. The reader will also notice that we refer to

several papers in the supply chain management literature that do not specifically

study perishable inventories. We do so in order to highlight analogous potential

research areas in the supply chain management of perishable goods.

The chapter is organized as follows: We first provide a discussion of common

challenges in production planning of perishables in Section 2.2. Then we review

the research on single product, single location models in inventory management of

perishables in Section 2.3. We next focus on multi-echelon and multi-location models

in Section 2.4. Research with novel features such as multiple products, multiple-types

of customers and different demand models are reviewed in Section 2.5. We conclude

by detailing a number of open research problems in Section 2.6.

2.2 Challenges in Production Planning

One of the most comprehensive studies on production planning and perishable goods

is the doctoral dissertation of Lutke Entrup (2005), which focuses on the use of leading

advance planning and scheduling (APS) systems (such as PeopleSoft’s EnterpriseOne

and SAP’s APO) to manage products with short shelf-lives. According to Lutke En-

trup (2005), the use of APS in perishable supply chains remains low in contrast to the

supply chains of non-perishable goods. He then describes how shelf-life is integrated

into the current APS, and identifies particular weaknesses of APS systems for perish-

ables by carefully studying the characteristics and requirements of the supply chains

of three different products. Based on these case studies, he proposes customized solu-

tions which would enable APS systems to better match the needs of the fresh produce

industry. We refer the reader to this resource for more information on practical issues

in production planning for perishable goods. In the remainder of this section, we

8

Page 18: Essays on Perishable Inventory Management

provide an overview of the analytical research in production planning, mainly focus-

ing on the general body of knowledge in operations management and management

science disciplines.

Capacity planning is one of the key decisions in production of perishables. Re-

search on this topic is primarily focused on agricultural products: Kazaz (2004)

describes the challenges in production planning by focusing on long-term capacity

investments, yield uncertainty (which is a common problem in the industries that in-

volve perishables) and demand uncertainty. He develops a two-stage problem where

the first stage involves determining the capacity investment and the second-stage

involves the production quantity decision. Jones et al. (2001) analyze a production

planning problem where after the initial production there is a second chance to pro-

duce, still facing yield uncertainty; Jones et al. (2003) describe the real-life capacity

management problem for a grower in more detail. Allen and Schuster (2004) present a

model for agricultural harvest risk. Although these papers do not model the aging of

the agricultural product once it is produced, they highlight the long term investment

and planning challenges. We refer the reader to Kazaz (2004) for earlier references on

managing yield uncertainty and to Lowe and Preckel (2004) for research directions

within the general domain of agribusiness.

In addition to capacity planning and managing yield uncertainty, product-mix

decisions are integral to the management of perishables. Different companies in the

supply chain (producer, processor, distributor, retailer) face this problem in slightly

different ways. For example, again in agribusiness, a producer decides on the use of

farm land for different produce, hence deciding on the capacity-mix. A supermarket

can offer a fresh fruit or vegetable as- is or as an ingredient in one of their ready-to-eat

products (e.g., pre-packed fruit salad). While the choice of the product-mix affects

the useful lifetime of a product, it can also help reduce the waste at the retailer if, for

example, produce that is not salable due to problems in its appearance or packaging

is used in preparing the ready-to-eat product. Similarly, fresh produce at a processing

facility can be used in ready-to-eat, cooked products (e.g. pre-cooked, frozen dishes)

or it can be sold as an uncooked, frozen product. These two types of products have

different shelf-lives. Product-mix decisions - whole blood vs. blood components - are

also crucial in blood supply chains. While quite common, such planning problems

have not been studied extensively in the operations management and management

9

Page 19: Essays on Perishable Inventory Management

science literature; owing to their prevalence in practice the strategic management of

products with explicitly different shelf-lives stands as an important open problem.

We discuss the limited research in this area in Section 2.5.

Note that for perishables, both capacity planning and product-mix decisions are

typically driven by target levels of supply for a final product. In that respect, pro-

duction plans are tightly linked to inventory control models. We therefore discuss

research in inventory control for perishables in the rest of this chapter, treating single

location models in Section 2.3, the multi-location and multi-echelon models in Sec-

tion 2.4, and modeling novelties in Section 2.5. Once again, we attempt to emphasize

those areas most in need for further research.

2.3 Managing Inventories at a Single Location

Single-location inventory models form the basic building blocks for more complex

models – possibly comprised of multiple locations and/or echelons, potentially in-

corporating information flows, knowledge of market or customer behavior, and/or

additional logistics options. This work was pioneered by Veinott (1960), Bulinskaya

(1964) and Van Zyl (1964), who consider discrete review problems without fixed cost

under, respectively, deterministic demand, stochastic demand for items with a one-

period lifetime, and stochastic demand for items with a two-period lifetime. We

will discuss work in this discrete review setting first, before expanding our scope to

continuous review systems. There is a single product in all the models reviewed in

this section, and inventory is depleted starting from the oldest units in stock, i.e.,

first-in-first-out (FIFO) inventory issuance is used.

2.3.1 Discrete Review Models

We provide an overview of research by classifying the work based on modeling as-

sumptions: Research assuming no fixed ordering costs or lead times is reviewed in

Section 2.3.1, followed by research on models with fixed ordering costs but no lead

times in Section 2.3.1. Research on models with positive lead times, which complicate

the problem appreciably, is reviewed in Section 2.3.1.

10

Page 20: Essays on Perishable Inventory Management

Table 2.1 provides a high-level overview of some of the key papers within the

discrete review arena. Categorization of the papers and the notation used in Table 2.1

are described below:

• Replenishment policy: Papers have considered the optimal control policy (Opt),

base stock policies that order to keep the Total Items in System (TIS) - summed

over all ages - constant or that order to keep the New Items in System (NIS)

constant (see Section 2.3.1), other heuristics (H), and, when fixed ordering costs

are present, the (s, S) policy. When (s, S) policy is annotated with a ‡ this shows

(s, S) policy is implied to be the optimal replenishment policy based on earlier

research (however a formal proof of optimality of (s, S) has not appeared in

print for the model under concern).

• Excess demand: Excess demand is either assumed to be backlogged (B), or

modeled as lost sales (L).

• Problem horizon: Planning horizon is either finite (F) or infinite (I).

• Replenishment lead time: All papers, save for one, assume zero lead time. The

exception is a paper that allows a deterministic (D) replenishment lead time.

• Product lifetime: Most papers assume deterministic lifetimes of general length

(D) although two papers assume the lifetime is exactly two time periods (2).

One paper allows for general discrete phase-type lifetimes, (PH), and assumes

that all items perish at the same time, i.e., a disaster model. This latter is

denoted by 2 in Table 2.1. Another permits general discrete lifetimes, but

assumes items perish in the same sequence as they were ordered. This is denoted

by 7 in the table.

• Demand distribution: Models often assume general demand in each period hav-

ing a continuous density (cont). Others assume discrete distribution (disc),

Erlang – which may include exponential – (Ek) or batch demand with either

geometric (geo) or general discrete phase-type renewal arrivals, (PH). Often-

times the continuous demand is assumed for convenience – results appear to

generalize to general demand distributions.

11

Page 21: Essays on Perishable Inventory Management

• Costs: Costs include unit costs for ordering (c), unit per unit time holding (h),

perishing (m), unit per unit time shortage (p), one-time costs for shortage (b),

and fixed cost for ordering (K). The annotation 1 in Table 2.1 indicates papers

that allow the unit holding and shortage costs to be generalized to convex

functions. A paper that does not list cost parameters is concerned with the

general properties of the model, such as expected outdates, without attaching

specific costs to these.

Discrete Review Models without Fixed Ordering Cost or Lead Time

Early research in discrete time models without fixed ordering costs focused on char-

acterizing optimal policies. Fundamental characterizations of the (state dependent)

optimal ordering policy were provided by Nahmias and Pierskalla (1973) for the two

period lifetime problem, when penalty costs per unit short per unit time, and unit

outdating costs are present. Shortly thereafter Fries (1975) and Nahmias (1975a) in-

dependently characterized the optimal policy for the general lifetime problem. Both

of these papers had per unit outdating costs, per unit ordering costs, and per unit per

period holding costs. In addition, Fries (1975) had per unit shortage costs, and Nah-

mias (1975a) per unit per period shortage costs; this difference arises because Fries

(1975) assumes lost sales and Nahmias (1975a) backlogging of unsatisfied demand.

The cost structure in Nahmias (1975a) quickly became the “standard” for discrete

time models without fixed ordering cost, and we will refer to it as such within the

rest of this section, although in later papers the per unit per period penalty and hold-

ing costs were extended to general convex functions. The incorporation of separate

ordering and outdating costs allows modeling flexibility to account for salvage value;

use of only one or the other of these costs is essentially interchangeable.

Treatment of the costs due to outdating is the crucial differentiating element

within the perishable setting. In fact, while Fries (1975) and Nahmias (1975a) took

alternate approaches for modeling the costs of expiration – the former paper charges

a cost in the period items expire, while the latter charges an expected outdating cost

in the period items are ordered – Nahmias (1977b) showed that these two approaches

were essentially identical, modulo end of horizon effects. The policy structures out-

lined in Fries (1975) and Nahmias (1975a) were quite complex; perishability destroys

12

Page 22: Essays on Perishable Inventory Management

Rep

lenis

hm

ent

Exce

ssP

lannin

gLea

dLife-

Dem

and

Art

icle

policy

dem

and

Hori

zon

tim

eti

me

dis

trib

uti

on

Cost

s

Nahm

ias

and

Pie

rskalla

(1973)

Opt

B/L

F/I

02

cont

p,m

Fri

es(1

975)

Opt

LF/I

0D

cont

c,h,b

,m

Nahm

ias

(1975a)

Opt

BF

0D

cont

c,h,p

,m

Cohen

(1976)

TIS

BI

0D

cont

c,h,p

,m

Bro

dhei

met

al.

(1975)

NIS

LI

0D

dis

c

Chaza

nand

Gal(1

977)

TIS

LI

0D

dis

c

Nahm

ias

(1975b,1

975c)

HB

F0

DE

kc,

h,p

,m

Nahm

ias

(1976)

TIS

BF

0D

cont

c,h,p

,m1

Nahm

ias

(1977a)

HB

F0

Dco

nt

c,h,p

,m1

Nahm

ias

(1977c)

TIS

BF

0dis

c7co

nt

c,h,p

,m1

Nandakum

ar

and

Mort

on

(1993)

HL

I0

Dco

nt

c,h,b

,m

Nahm

ias

(1978)

(s,S

)B

F0

Dco

nt

c,h,p

,m,K

1

Lia

nand

Liu

(1999)

(s,S

)‡B

I0

Dbatc

hgeo

h,p

,b,m

,K

Lia

net

al.

(2005)

(s,S

)‡B

I0

PH

2batc

hP

Hh,p

,b,m

,K

William

sand

Patu

wo

(1999,2

004)

HL

FD

2co

nt

c,h,b

,m

Tab

le2.

1:A

sum

mary

of

dis

cre

teti

me

single

item

invento

rym

odels

:1

denote

spapers

inw

hic

hhold

ing

and

short

age

cost

scan

be

genera

lconvex

functi

ons,

2denote

suse

of

adis

ast

er

modelin

whic

hall

unit

speri

shat

once,7

denote

sth

eass

um

pti

on

that

item

s,even

though

they

have

random

life

tim

es,

willperi

shin

the

sam

ese

quence

as

they

were

ord

ere

d,and‡

denote

sth

ecase

when

(s,S)

policy

isim

plied

tobe

the

opti

malre

ple

nis

hm

ent

policy

base

don

earl

ier

rese

arc

h(h

ow

ever

afo

rmalpro

ofofopti

mality

of(s

,S)

has

not

appeare

din

pri

nt

for

the

modelunder

concern

).

13

Page 23: Essays on Perishable Inventory Management

the simple base-stock structure of optimal policies for discrete review models without

fixed ordering costs in the absence of perishability.

This complexity of the optimal policy was reinforced by Cohen (1976), who char-

acterized the stationary distribution of inventory for the two-period problem with the

standard costs, showing that the optimal policy for even this simple case was quite

complex, requiring state-dependent ordering. To all practical extents this ended acad-

emic study of optimal policies for the discrete review problem under FIFO issuance; to

find an optimal policy dynamic programming would be required, and implementation

of such a policy would be difficult owing to its complex, state dependent nature (with

a state vector tracking the amount of inventory in system of each age). Therefore

many researchers turned to the more practical question of seeking effective heuristic

policies that would be (i) easy to define, (ii) easy to implement, and (iii) close to

optimal.

Very soon after the publication of Nahmias (1975a), a series of approximations

appeared for the backorder and lost sales versions of the discrete review zero fixed

ordering cost problem. Initial works (Brodheim et al., 1975, Nahmias, 1975b) ex-

amined the use of different heuristic control policies; Brodheim et al. (1975) propose

a fascinating simplification of the problem – making order decisions based only on

the amount of new items in the system, what we call is the NIS heuristic, as it or-

ders based only on new items in the system. They realized that if a new order size

was constant, Markov chain techniques could be used to derive exact expressions,

or alternately very simple bounds, on key system statistics (which is the focus of

their paper). Nahmias (1975b) used simulation to compare multiple heuristics for the

problem with standard costs, including the “optimal” TIS policy, a piecewise linear

function of the optimal policy for the non-perishable problem, and a hybrid of this

with NIS ordering. He found that the first two policies outperform the third, which

effectively ended consideration of NIS and its variants for several decades.

Three other papers in sequence, Nahmias (1975c, 1976, 1977a), explicitly treat the

question of deriving good approximation policies (and parameters) for the problem

with the standard costs; the final of these three can be considered as the culmination

of these initial heuristic efforts. It computationally compares the heuristic that keeps

only two states in the inventory vector, new inventory and inventory over one day

old, with both the globally optimal policy (keeping the entire inventory vector) and

14

Page 24: Essays on Perishable Inventory Management

the optimal TIS policy. (This optimal TIS policy is easily approximated using tech-

niques in Nahmias, 1976.) This comparison showed that for Erlang or exponential

demand and a lifetime of three periods the performance of both heuristics is excep-

tional – always within 1% of optimal, with the reduced state-space heuristic uniformly

outperforming the optimal TIS policy. Note also that using the two-state approxi-

mation eliminates any need to track the age of inventory other than new versus old,

significantly reducing both the computational load and complexity of the policy.

All of these heuristics consider backorder models. The first heuristic policy for

discrete time lost sales models was provided by Nandakumar and Morton (1993), who

began with properties and bounds on the expected outdates under TIS in lost sales

systems, provided by Chazan and Gal (1977). These Nandakumar and Morton (1993)

incorporated into heuristics following Nahmias’ (1976) framework for the backorder

problem. Nandakumar and Morton (1993) compared these heuristics with alternate

“near-myopic” heuristics they developed using newsvendor-type logic, for the problem

with standard costs. They found that all of the heuristics performed within half a

percentage of optimal, with the near myopic heuristics showing the best performance,

typically within 0.1% of optimal. (These heuristics could likely be improved even

further, with the use of even tighter bounds on expected outdates derived by Cooper,

2001.)

Note that both the backorder and lost sales heuristics should be quite robust with

respect to items having even longer lifetimes than those considered in the papers:

Both Nahmias (1975a) and Fries (1975) observed that the impact of newer items

on ordering decisions is greater than the impact of older items. (This, in fact, is

one of the factors motivating the heuristic strategies in Nahmias, 1977a.) Thus, the

standard discrete time, fixed lifetime perishable inventory model, with backorders or

lost sales has for all practical purposes been solved – highly effective heuristics exist

that are well within our computational power to calculate. One potential extension

would be to allow items to have random lifetimes; if items perish in the same sequence

as they were ordered, many of the fixed lifetime results continue to hold (Nahmias,

1977c), but a discrete time model with random lifetimes that explicitly permits items

to perish in a different sequence than they were ordered remains an open, and likely

challenging question. The challenge arises from the enlarged state space required to

capture the problem characteristics – in this case the entire random lifetime vector

15

Page 25: Essays on Perishable Inventory Management

– although techniques from continuous time models discussed in Section 2.3.2 could

prove useful in this endeavor.

Discrete Review Models without Lead Times Having Positive Fixed Or-

dering Cost

Nahmias (1978) was the first to analyze the perishable inventory problem with no

lead time but positive fixed ordering cost in addition to the standard costs described

in Section 2.3.1. In his paper, for the one-period problem, Nahmias (1978) established

that the structure of the optimal policy is (s, S) only when the lifetime of the object

is two; lifetimes of more than two periods have a more complex, non-linear structure.

Extensions of the two-period (s, S) result to the multi-period problem appeared quite

difficult, given the analytical techniques of the time: The fact that the costs exclud-

ing the fixed ordering cost are not convex appears to render hopes of extending the

K-convexity of Scarf (1960) to the perishable domain to be in vain. Nevertheless, Nah-

mias (1978) reports extensive computational experiments supporting the conjecture

that the general ((s, S) or non-linear) optimal policy structure holds for the multi-

period problem. One possible avenue to prove such structure could be through the

application of recent proof techniques involving decomposition ideas, such as those in

Muharremoglu and Tsitsiklis (2003).

Lian and Liu (1999) consider the discrete review model with fixed ordering cost,

per unit per period holding cost, per unit and per unit per unit time shortage costs,

and per unit outdating costs. Crucially, their model is comprised of discrete time

epochs where demand is realized or units in inventory expire (hence discrete time

refers to distinct points in time where change in the inventory levels occur) as opposed

to distinct time periods in Nahmias (1978). Lian and Liu (1999) analyze an (s, S)

policy which was shown to be optimal under continuous review by Weiss (1980) for

Poisson demand. The instantaneous replenishment assumption combined with the

discrete time model of Lian and Liu (1999) ensures that the optimal reorder level s

will be no greater than -1 because any value larger than -1 will add holding costs

without incurring any shortage cost in their model. The zero lead time assumption

and cost structure of Lian and Liu (1999) are common in the continuous review

framework for perishable problems (see Section 2.3.2); it was a desire to develop a

16

Page 26: Essays on Perishable Inventory Management

discrete time model to approximate the continuous review that motivated Lian and

Liu (1999). In the paper they use matrix analytic methods to analyze the discrete

time Markov chain capturing the system evolution, and establish numerically that

the discrete review model is indeed a good approximation for the continuous review

model, especially as the length of the time intervals gets smaller.

Lian et al. (2005) follow Lian and Liu (1999), sharing similar costs and replenish-

ment assumptions. This later paper allows demands to be batch with discrete phase

type interdemand times, and lifetimes to be general discrete time phase type, but

requires that all items of the same batch perish at the same time. They again use

matrix analytic methods to derive cost expressions, and numerically show that the

variability in the lifetime distribution can have a significant effect on system perfor-

mance. Theirs is a quite general framework, and provides a powerful method for the

evaluation of discrete time problems in the future, provided the assumption of the

instantaneous replenishments is reasonable in the problem setting.

Of great practical importance again is the question of effective heuristic policies.

Nahmias (1978) establishes, computationally, that while not strictly optimal, (s, S)

type policies perform very close to optimal. Moreover, he also presents two methods of

approximating the optimal (s, S) parameters. Both of these are effective: on average

their costs are within 1% of the globally optimal cost for the cases considered, and

in all cases within 3%. Thus, while there certainly is room for further refinements of

these heuristic policies (along the lines of the zero fixed cost case for example), the

results in Nahmias (1978) are already compelling.

Discrete Review Models with Positive Lead Time

As optimal solutions to the zero lead time case require the use of dynamic program-

ming, problems with lead times are in some sense no more difficult, likewise requiring

dynamic programming, albeit of a higher dimension. Thus problems considering op-

timal policies for discrete review problems with lead times are also often considered

in the context of other generalizations to the model: for example different selling

prices depending on age (Adachi et al., 1999). Interestingly, there is very little work

on extending discrete review heuristics for the zero lead time model to the case of

positive lead times, although many of the methods for the zero lead time case should

17

Page 27: Essays on Perishable Inventory Management

in principle be applicable, by expanding the vector of ages of inventory kept to include

those items on order, but which have not yet arrived.

One possible direction is provided in the work of Williams and Patuwo (1999, 2004)

who derive expressions for optimal ordering quantities based on system recursions for

a one-period problem with fixed lead time, lost sales, and the following costs: per unit

shortage costs and per unit outdating costs, per unit ordering costs and per unit per

period holding costs. Williams and Patuwo (2004) in fact state that their methods

can be extended to finite horizons, but such an extension has not yet appeared.

Development of such positive lead time heuristics, no matter what their genesis, would

prove valuable both practically and theoretically, making this a potentially attractive

avenue for future research. The trick, essentially, is to keep enough information so as

to make the policy effective, without keeping so much as to make the policy overly

cumbersome. More complex heuristics have been proposed for significantly more

complex models with lead times, such as in Haijema et al. (2005a, 2005b), but these

likely are more involved than necessary for standard problem settings.

2.3.2 Continuous Review Models

We now turn our attention to continuous review models. These are becoming in-

creasingly important with the advent of improved communication technology (such

as radio frequency identification, RFID) and automated inventory management and

ordering systems (as a part of common enterprise resource planning, ERP, systems).

These two technologies may eventually enable management of perishables in real time,

potentially reducing outdating costs significantly. We will first consider continuous

review models without fixed ordering costs or lead times in Section 2.3.2, models

without fixed ordering costs but positive lead times in Section 2.3.2, and finally the

models which incorporate fixed ordering costs in Section 2.3.2.

Tables 2.2 and 2.3 provide a high-level overview of some of the key papers within

the continuous review framework, segmented by whether a fixed ordering cost is

present in the model (Table 2.3) or not (Table 2.2). Categorization of the papers and

the notation used in these tables are described below:

• Replenishment policy: Some papers in Table 2.2 assume a base stock policy,

denoted by a z. Some papers in the same table concern themselves only with

18

Page 28: Essays on Perishable Inventory Management

characterizing system performance; these have a blank in the replenishment

policy column. In Table 2.3, all papers either use the (s, S) policy, annotated

with a ∗ when proved to be optimal, with a ‡ when only implied to be the

optimal policy based on earlier research (however a formal proof of optimality

of (s, S) has not appeared in print for the model under concern) or when batch

sizes are fixed the (Q, r) or (Q, r, T ) policies, where the latter policy orders when

inventory is depleted below r or when items exceed T units of age.

• Excess demand: Papers may assume simple backlogging (B), some sort of gen-

eralized backlogging in which not all backlogged customers will wait indefinitely

(Bg), or lost sales (L). Two papers assumes all demand must be satisfied; this

field is blank in that case.

• Problem horizon: Planning horizon is either finite (F) or infinite(I).

• Replenishment lead time: In Table 2.2, many papers assume that items arrive

according to a Poisson process, outside of the control of the system manager;

these are denoted by (M). This may be generalized to the case of batch replen-

ishments (batch), the case when the replenishment rates can be controlled, (M3)

or are state dependent (M4). Other papers in the no fixed ordering cost regime

assume either exponential (exp) with a single or ample (ample) servers, renewal

(renewal), continuous (cont) or fixed (D) lead times. Most papers in Table 2.3

assume either zero lead time or a deterministic (D) lead time, although some

allow exponential (exp), or general continuously distributed (cont) lead times.

• Product lifetime: Most lifetimes are deterministic (D), possibly with the as-

sumption that the item can be used for a secondary product after it perishes

(D0), that all perish at once, a disaster model, indicated by (D2), or that items

within a lot only begin to age after all items from the previous lot have left

inventory (D6). Item lifetimes may also be exponentially distributed (exp),

generally distributed (gen), Erlang (Ek) and in one case they decay (decay).

• Demand distribution: As we are within a continuous model, this column de-

scribes the assumptions on the demand interarrival distribution. Most are Pois-

son (M), although in some cases they may have rates that can be controlled by

19

Page 29: Essays on Perishable Inventory Management

the system manager (M3) or which are state-dependent (M4). Selected papers

allow independent and identically distributed arrivals (iid), continuously dis-

tributed interarrivals (cont), some allow arrivals in batches (batch), and others

have renewal (renewal), or general (general) demand interarrival distributions.

• Costs: Costs include unit costs for ordering (c), per unit time for holding (h),

unit costs for perishing (m), per unit time for shortage (p), one-time unit costs

for shortage (b), and fixed cost for ordering (K). When demand and interarrival

rates can be controlled, there is a cost (s) for adjusting these rates. Some models

calculate profits, via using a unit revenue (r). One paper not only charges the

standard per unit time holding and penalty costs (h) and (p), but also per unit

time costs based on the average age of the items being held (h’) or backlogged

(p’). Two papers optimize subject to constraints (possibly with other costs), and

several (but not all) that concern themselves only with characterizing system

performance have this column left blank. In addition, one paper utilizes a

Brownian control model with two barriers; it defines shortage and outdate costs

slightly differently to account for the infinite number of hits of the barriers. This

is denoted by 8. One paper calculates actuarial valuations of costs and expected

future revenues, where the latter depend on the age of an item when it is sold.

This is denoted with a 9. Finally, when annotated with a 1 a paper allows the

unit holding and shortage costs to be generalized to convex functions.

Continuous Review without Fixed Ordering Cost or Lead Time

The continuous review problem without fixed ordering costs is unique in that under

certain modeling assumptions, direct parallels can be drawn between the perishable

inventory problem and stochastic storage processes in general, and queueing theory

in particular. These parallels provide structural results as well as powerful analytical

tools.

One of the earliest papers to make these connections was written by Graves (1982).

Like many of the papers in this area, Graves (1982) was concerned with character-

izing the system behavior, and as such his paper lacks an explicit cost structure or

replenishment policy. In this work Graves (1982) shows that the virtual waiting time

20

Page 30: Essays on Perishable Inventory Management

Rep

lenis

hm

ent

Exce

ssP

lannin

gLea

dLife-

Dem

and

Art

icle

policy

dem

and

hori

zon

tim

eti

me

dis

trib

uti

on

Cost

s

Gra

ves

(1982)

B/L

IM

Dbatc

hM

Kasp

iand

Per

ry(1

983)

LI

MD

M

Kasp

iand

Per

ry(1

984)

Bg/L

Ire

new

al

DM

Per

ryand

Posn

er(1

990)

LI

M3

DM

3c,

h,b

,s

Per

ry(1

997)

LI

M3

DM

3c,

h,b

,s8

Per

ryand

Sta

dje

(1999)

Bg/L

IM

4ex

p/D

M4

c,h,p

,m,r

,h’,p’

Per

ryand

Sta

dje

(2000a)

LI

MD

0M

Per

ryand

Sta

dje

(2000b)

LI

Mgen

M

Per

ryand

Sta

dje

(2001)

LI

MD

/M

2M

Nahm

ias

etal.

(2004a)

LI

M4

DM

4

Nahm

ias

etal.

(2004b)

LI

batc

hM

Dre

new

al

c,b,m

,r9

Pal(1

989)

zB

Iex

pex

pM

h,p

,b,m

Liu

and

Cheu

ng

(1997)

zB

g/L

Iex

p/am

ple

exp

Mco

nst

rain

ts

Kalp

akam

and

Sapna

(1996)

zL

Iex

pex

pre

new

al

c,h,b

,m

Kalp

akam

and

Shanth

i(2

000)

zB

g/L

IM

4ex

pM

c,h,p

,b,m

Kalp

akam

and

Shanth

i(2

001)

zL

Ico

nt

exp

Mc,

h,b

,m

Sch

mid

tand

Nahm

ias

(1985)

zL

ID

DM

c,h,b

,m

Per

ryand

Posn

er(1

998)

zB

g/L

ID

DM

Tab

le2.

2:A

sum

mary

ofconti

nuous

tim

esi

ngle

item

invento

rym

odels

wit

hout

fixed

invento

rycost

:g

denote

sa

genera

lized

backord

er

model,

0denote

sth

echara

cte

rist

icth

at

aft

er

item

speri

sh

from

afirs

tsy

stem

they

have

use

ina

second

syst

em

,2

denote

suse

of

adis

ast

er

model

inw

hic

hall

unit

speri

shat

once,

3denote

sth

eability

tocontr

ol

the

arr

ival

and/or

dem

and

rate

s,4

denote

s

state

-dependent

arr

ivaland/or

dem

and,8

denote

ssl

ightl

ym

odifi

ed

cost

definit

ions

tofit

wit

hin

the

conte

xt

ofa

Bro

wnia

ncontr

olfr

am

ew

ork

,and

9denote

sactu

ari

alvalu

ati

ons

base

don

expecte

dcost

s

and

age-d

ependent

revenues.

21

Page 31: Essays on Perishable Inventory Management

Rep

lenis

hm

ent

Exce

ssP

lannin

gLea

dLife-

Dem

and

Art

icle

policy

dem

and

hori

zon

tim

eti

me

dis

trib

uti

on

Cost

s

Wei

ss(1

980)

(s,S

)*B

/L

I0

DM

c,h,p

,m,r

,K1

Kalp

akam

(s,S

)‡L

I0

exp

Mc,

h,m

,K

and

Ari

vari

gnan

(1988)

Liu

(1990)

(s,S

)‡B

I0

exp

Mh,p

,b,m

,K

Moort

hy

etal.

(1992)

(s,S

)‡L

F/I

0E

kbatc

hiid

Liu

and

Shi(1

999)

(s,S

)‡B

I0

exp

renew

al

h,p

,b,m

,K

Liu

and

Lia

n(1

999)

(s,S

)‡B

I0

Dre

new

al

h,p

,b,m

,K

Lia

nand

Liu

(2001),

(s,S

)‡/

(s,S

)B

I0

/D

Dbatc

hre

new

al

h,p

,b,m

,K

Gurl

erand

Ozk

aya

(2003)

Gurl

erand

Ozk

aya

(2006)

(s,S

)B

I0/D

gen

2batc

hre

new

al

h,p

,b,m

,K

Kalp

akam

and

Sapna

(1994)

(s,S

)L

Iex

pex

pM

c,h,b

,m,K

Ravic

handra

n(1

995)

(s,S

)L

Ico

nt5

D/D

6M

c,h,b

,m,K

Liu

and

Yang

(1999)

(s,S

)B

Iex

pex

pM

h,p

,b,m

,K

Nahm

ias

and

Wang

(1979)

(Q,r

)B

ID

dec

ay

cont

h,b

,m,K

Chiu

(1995)

(Q,r

)B

ID

Dgen

eral

c,h,b

,m,K

Tek

inet

al.(2

001)

(Q,r

,T)

LI

DD

6M

h,m

,K,c

onst

rain

t

Ber

kand

Gurl

er(2

006)

(Q,r

)L

ID

D2

Mh,b

,m,K

Tab

le2.

3:C

onti

nuous

tim

esi

ngle

item

invento

rym

odels

wit

hfixed

ord

eri

ng

cost

:1

denote

spapers

inw

hic

hhold

ing

and

short

age

cost

scan

be

genera

lconvex

functi

ons,

2denote

suse

ofa

dis

ast

er

modelin

whic

hall

unit

sfr

om

the

sam

eord

er

peri

shat

once,5

that

only

one

outs

tandin

gord

er

isallow

ed

at

ati

me,6

the

ass

um

pti

on

that

pro

duct

agin

gst

art

sonly

once

all

invento

ryfr

om

the

pre

vio

us

lot

has

been

exhaust

ed

or

expir

ed,∗

denote

sth

ecase

when

(s,S)

policy

ispro

ved

tobe

opti

mal,

and‡

denote

sth

ecase

when

(s,S)

policy

isim

plied

tobe

the

opti

malre

ple

nis

hm

ent

policy

base

don

earl

ier

rese

arc

h(h

ow

ever

afo

rmalpro

ofofopti

mality

of(s

,S)

has

not

appeare

din

pri

nt

for

the

modelunder

concern

).

22

Page 32: Essays on Perishable Inventory Management

in an M/M/1 queue with impatient customers and a M/D/1 with finite buffer can

be used to model the inventory in a perishable inventory system with Poisson de-

mands of exponential, or unit size, respectively, with either lost sales or backlogging.

A crucial assumption is that items are replenished according to what Graves (1982)

terms “continuous production;” the arrival of items into inventory is modeled as a

Poisson process, essentially out of the control of the inventory manager. Under this

convention, Graves (1982) notes that the key piece of information is the age of the

oldest item currently in stock, an observation that would be used by many subse-

quent researchers. For example, Kaspi and Perry (1983, 1984) model systems with

Poisson demand and Poisson or renewal supply, as might be the case, for example

in blood banks that rely on donations for stock. For their analysis Kaspi and Perry

(1983, 1984) track what they call the virtual death process, which is the time until

the next death (outdate) if there would be no more demand. This, of course, is just

a reformulation of the age of the oldest item in stock, as used by Graves (1982).

These papers mark the start of a significant body of work by Perry (sometimes

in conjunction with others) on continuous review perishable inventory systems with

no fixed ordering cost, some sort of Poisson replenishment, and nearly all using the

virtual death process. Only Perry and Posner (1990) and Perry and Stadje (1999)

contain explicit cost functions; the rest of the papers concern themselves solely with

performance characteristics, as in Graves (1982). Perry and Posner (1990) develop

level crossing arguments for storage processes to capture the effects of being able

to control supply or demand rates within the model of Kaspi and Perry (1983);

the limiting behavior of this system was analyzed by Perry (1997) using a diffusion

model and martingale techniques. Perry and Stadje (1999) depart from the virtual

death process, instead using partial differential equations to capture the stationary

law of a system which now may have state-dependent arrival and departure rates

with deterministic or exponential lifetimes and/or maximum waiting times, as well

as finite storage space. This work is generalized in Nahmias et al. (2004a) using the

virtual death process. Likewise using the virtual death process, Perry and Stadje

(2000a, 2000b, 2001) evaluate systems where after perishing the item can be used for

a secondary product (such as juice for expired apples); items may randomly perish

before the expiration date; or, in addition to their fixed lifetime, items may all perish

before their expiration date (due to disasters, or obsolescence). Still using the virtual

23

Page 33: Essays on Perishable Inventory Management

death process, Nahmias et al. (2004b) provide actuarial valuations of the items in

system and future sales, when item values are dependent on age. Recently Perry and

Stadje (2006) solved a modified M/G/1 queue and showed how it related, again, to

the virtual death process in a perishable system, this time with lost sales. Thus work

on extensions to what can reasonably be called the Perry model continues.

Continuous Review without Fixed Ordering Cost having Positive Lead

Time

As mentioned above, a defining characteristic of the Perry model is that the supply

of perishable items arrives according to a (possibly state dependent) Poisson process.

When replenishment decisions and lead times must be included in a more explicit

manner, researchers have had to develop other analytical methods.

When lifetimes and lead times are exponentially distributed the problem is simpli-

fied somewhat, as this allows the application of renewal theory, transform methods,

and Markov or semi-Markov techniques, often on more complex versions of the prob-

lem. Pal (1989) looks at the problem with exponential lead times and lifetimes,

Kalpakam and Sapna (1996), allow renewal demands with lost sales, Kalpakam and

Shanthi (2000, 2001) consider state dependent Poisson lead times and then general

continuous lead times, and Liu and Cheung (1997) are unique in that they consider

fill rate and waiting time constraints. All of these papers consider base-stock, or

(S−1, S) inventory control; Kalpakam and Sapna (1996) and Kalpakam and Shanthi

(2001) consider per unit shortage costs and per unit outdating costs, per unit or-

dering costs and per unit per period holding costs. To these Kalpakam and Shanthi

(2000) add a per unit per period shortage cost, while Pal (1989) also adds the per

unit per unit time shortage cost, but disregards the per unit ordering cost. Liu and

Cheung (1997) take a different approach, seeking to minimize inventory subject to

a service level constraint. In all of cases the cost function appears to be unimodal

in S, but no formal proofs have been provided, owing to the difficulty in proving

unimodality. Furthermore, nowhere has the performance of base-stock policies been

formally benchmarked against the optimal, possibly due to the difficulty in dealing

with a continuous state space – time – within the dynamic programming framework.

This remains an open question.

24

Page 34: Essays on Perishable Inventory Management

The case of fixed lead times and lifetimes is arguably both more realistic and

more difficult analytically. Schmidt and Nahmias (1985) consider a system operating

under a base-stock policy with parameter S, lost sales, Poisson demand, and per unit

shortage costs and per unit outdating costs, per unit ordering costs and per unit

per period holding costs. They define and solve partial differential equations for the

S-dimensional stochastic process tracking the time since the last S replenishment

orders. Numerical work shows that again cost appears to be monotonic in S (in

fact convex), although surprisingly, the optimal value of S is not monotonic in item

lifetime.

Some thirteen years later, Perry and Posner (1998) generalized this work to allow

for general types of customer impatience behavior, in this case using level crossing

arguments to derive the stationary distribution of the vector of times until each of

the S items in system will outdate (reminiscent of their virtual death process). They

also show that the distribution of the differences between the elements of this vector

follows that of uniform order statistics, which enables them to derive expressions

for general customer behavior. These expressions may, as a rule, require numerical

evaluation.

Perry and Posner (1998) are concerned with general system characteristics; they

do not include explicit costs in their paper. Thus, while Perry and Posner (1998) pro-

vides rich material for future research – for example exploring how different customer

behavior patterns affect different echelons of a supply chain for perishable products

– there is also still a need for research following the work of Schmidt and Nahmias

(1985), with the aim of minimizing costs in the continuous review setting under fixed

lead times and fixed lifetimes.

Continuous Review with Fixed Ordering Cost

Within the continuous review fixed ordering cost model we make a distinction between

those models that assume fixed batch size ordering, leading to (Q,R) type models,

and those that assume batch sizes can vary, leading to (s, S) type models. We consider

the (s, S) models first.

Initial work in this setting assumed zero lead time, which simplifies the problem

considerably, as there is no need to order until all the items are depleted. In this case

25

Page 35: Essays on Perishable Inventory Management

when considering fixed ordering costs, unit revenues, unit ordering and outdate costs,

and convex holding and penalty costs per unit time, the optimal policy structure

under Poisson demand was found by Weiss (1980) who showed that for the fixed

lifetime problem over an infinite horizon, with lost sales or back ordering, an (s, S)

policy is optimal. Weiss (1980) also established that the optimal s value is zero in

the lost sales case, and in the backorder case no larger than -1 (you never order if

you have items in stock). Thus the optimal policy structure was established, but the

question of efficiently finding optimal parameters open.

The publication of Weiss (1980) initiated a series of related papers, all having

in common the assumption of immediate supply. Kalpakam and Arivarignan (1988)

treat the lost sales model as in Weiss (1980), but assume exponential lifetimes, con-

sider only costs (not revenues), and of these costs disregard the shortage costs as

they are irrelevant, as Weiss showed that the optimal s = 0. They also show that in

this case the cost, assuming an optimal s value, is convex in S. Liu (1990) takes the

setting of Weiss (1980), assumes exponential lifetimes, disregards the unit ordering

costs and revenues, but considers penalty costs per unit and per unit time. His focus

is on providing closed-form expressions for system performance, based on transform

analysis. Moorthy et al. (1992) perform a similar analysis using Markov chain theory

under Erlang lifetimes, assuming no shortage is permitted - or equivalently lost sales,

as in this case Weiss (1980) showed that it is optimal not to allow any shortage. Liu

and Shi (1999) follow Liu (1990), but now allow for general renewal demand. They

focus their analysis on the reorder cycle length, using it as a vehicle to prove various

structural properties of the costs with respect to the parameters. The assumption

of exponential lifetimes is crucial here – the results would be very unlikely to hold

under fixed lifetimes. Liu and Lian (1999) consider this problem, the same problem

as Liu and Shi (1999) but with fixed, rather than exponential lifetimes. They per-

mit renewal demands, and derive closed-form cost expressions, prove unimodality of

costs with respect to parameters, and show that the distribution of inventory level is

uniform over (s, S) (as in the non-perishable case).

All of the previous papers make the special zero lead time assumption of Weiss

(1980), which simplifies the problem. A few papers include positive fixed lead times,

Lian and Liu (2001) treat the model of Lian and Liu (1999) incorporating batch

demands to provide a heuristic for the fixed lead time case, but a proof in that paper

26

Page 36: Essays on Perishable Inventory Management

contained a flaw, which was fixed by Gurler and Ozkaya (2003). These papers show

how to efficiently find good (s, S) parameters for the fixed lead time problem, but

do not provide any benchmark against optimal. Thus while efficient heuristics exist

for the fixed lead time case, they have not as yet been benchmarked against more

complex control policies.

Random lead times have appeared in several papers within the continuous review,

fixed ordering cost framework. Kalpakam and Sapna (1994) allow exponential lead

times, while also assuming exponential lifetimes. As in the fixed lead time case, this

implies that the (s, S) policy is not longer necessarily optimal. They nevertheless

assume this structure. In addition to fixed ordering costs, they account for unit

purchasing, outdate and shortage costs, as well as holding costs per unit per unit time.

Under the assumption of only one outstanding order at a time, they derive properties

of the inventory process and costs. Ravichandran (1995) permits general continuous

lead times, deriving closed-form expressions for costs under the assumption that only

one order is outstanding at a time, and items in an order only begin to perish after all

items from the previous order have left the system. Liu and Yang (1999) generalize

Kalpakam and Sapna (1994) to allow for backlogs and multiple outstanding orders,

using matrix analytical methods to generate numerical insights under the assumption

of an (s, S) policy. For random lifetimes, still within the (s, S) structure, the most

comprehensive work is by Gurler and Ozkaya (2006), who allow a general lifetime

distribution, batch renewal demand, and zero lead time with a heuristic for positive

lead time. Their cost structure follows Lian and Liu (2001); Gurler and Ozkaya

(2006) can be thought of as generalizing Lian and Liu (2001) to the random lifetime

case. Gurler and Ozkaya (2006) argue that the random lifetime model is important

for modeling lifetimes at lower echelons of a supply chain, as items arriving there

will have already begun to perish. To this end they demonstrate the importance of

modeling the variability in the lead time distribution on costs, including a comparison

to the fixed lifetime heuristic of Lian and Liu (2001).

In general, within the fixed ordering cost model with variable lot sizes, for the zero

lead time case research is quite mature, but for positive lead times and/or batch de-

mands there are still opportunities for research into both the optimal policy structure

and effective heuristics. These models are both complex and practically applicable,

making this yet another problem that is both challenging and important.

27

Page 37: Essays on Perishable Inventory Management

If lot sizes must be fixed, initial work for “decaying” goods by Nahmias and Wang

(1979), using a (Q, r) policy, was followed two decades later by Chiu (1995), who

explicitly considers perishable (rather than decaying) items by approximating the

outdating and inventory costs to get heuristic (Q, r) values. Nahmias and Wang

(1979) consider unit ordering and shortage costs, and holding costs per unit per unit

time, as well as unit outdate costs. To these Chiu (1995) adds unit ordering costs.

Tekin et al. (2001) take a slightly different approach; they disregard unit ordering

costs and consider a service level constraint, rather than a shortage cost. They also

simplify the problem by assuming that a lot only starts aging after it is put into use,

for example moved from a deep freezer. To combat the effects of perishability, they

advocate placing an order for Q units whenever the inventory level reaches r or when

T time units have elapsed since the last time a new lot was unpacked, whichever

comes first, giving rise to a (Q, r, T ) policy. Not surprisingly, they find that inclusion

of the T parameter is most important when service levels are required to be high

or lifetimes are short. Berk and Gurler (2006) define the “effective shelf life” of a

lot: The distribution of the remaining life at epochs when the inventory level hits

Q. They show that this constitutes an embedded Markov process (as they assume

Poisson demand), and thus via analysis of this process they are able to derive optimal

(Q, r) parameters, when facing unit outdate and penalty costs, holding costs per unit

per unit time, and fixed ordering costs. They compare the performance of their policy

with that of Chiu (1995) and also with the modified policy of Tekin et al. (2001).

Not surprisingly, the optimal (Q, r) policy outperforms the heuristic of Chiu (1995),

sometimes significantly, and is outperformed by the more general modified policy of

Tekin et al. (2001). Nevertheless, this latter gap is typically small, implying that use

of the more simple (Q, r) policy is often sufficient in this setting.

Note that throughout these papers the (Q, r) structure has only been assumed, and

once again there does not appear to be any benchmarking of the performance of the

(Q, r) policy against the optimal, as optimal policies are difficult to establish given the

increased problem complexity the continuous time setting with perishability causes. If

such benchmarking were done, it would help identify those problem settings for which

the (Q, r) or (Q, r, T ) policy is adequate, and those which would most benefit from

further research into more complex ordering schemes. Essentially, the underlying

question of how valuable lifetime information is, in what degree of specificity, and

28

Page 38: Essays on Perishable Inventory Management

when it is most valuable, remains.

2.4 Managing Multi-Echelon and Multi-Location

Systems

Analysis of multi-echelon inventory systems dates back to the seminal work of Clark

and Scarf (1960); the reader can refer to Axsater (2000) for a unified treatment

of the research in that area, and Axsater (2003) for a survey of research on serial

and distribution systems. In serial and distribution systems, each inventory location

has one supplier. In contrast, multi-location models consider flow of products from

various sources to a particular location (possibly including transshipments). See, for

instance, Karmarkar (1981) for a description of the latter problem. Managing multi-

echelon and/or multi-location systems1 with aging products is a challenge because of

the added complexity in:

• replenishment and allocation decisions, where the age of goods replenished at

each location affects the age-composition of inventory and the system perfor-

mance, and the age of goods supplied/allocated downstream may be as impor-

tant as the amount supplied,

• logistics-related decisions such as transshipment, distribution, collection, which

are complicated by the fact that products at different locations may have dif-

ferent remaining lifetimes,

• centralized vs. decentralized planning, where different echelons/locations may

be managed by different decision-makers with conflicting objectives, operat-

ing rules may have different consequences for different decision-makers (e.g.,

retailers may want to receive LIFO shipments but suppliers may prefer to issue

their inventory according to FIFO), and system-wide optimal solutions need not

necessarily improve the performance at each location.

1Note that multi-location models we review here are different from the two-warehouse problem described in Section

10 of Goyal and Giri (2001); that model considers a decision-maker who has the option of renting a second storage

facility if he/she uses up the capacity of his/her own storage.

29

Page 39: Essays on Perishable Inventory Management

Given the complexity in obtaining or characterizing optimal decision structures,

analytical research in multi-echelon and multi-location systems has mainly focused

on particular applications (modeling novelties) and heuristic methods. We review

analysis of replenishment and allocation decisions in Section 2.4.1, logistics and dis-

tribution related decision in Section 2.4.2, and centralized vs. decentralized planning

in Section 2.4.3.

2.4.1 Research on Replenishment and Allocation Decisions

Research on multi-echelon systems has been confined to two-echelons2 except in the

simulation-based work (e.g., van der Vorst et al., 2000). Motivated by food supply

chains and blood banking, the upstream location(s) in the two-echelon models typi-

cally involve the supplier(s), the distribution center(s) (DC), or the blood banks, and

the downstream location(s) involve the retailer(s), the warehouse(s), or the hospi-

tal(s). In this section, we use the terms supplier and retailer to denote the upstream

and downstream parties, respectively. We call the inventory at the retail locations

the field inventory.

We first classify the research by focusing on the nature of the decisions and the

modeling assumptions. Table 2.4 provides a summary of the analytical work that

determines heuristic replenishment policies for a supplier and retailer(s), and/or ef-

fective allocation rules to ship the goods from the supplier to the retailers. As we can

see, analytical research has been rather limited, modeling assumptions have varied,

and some problems (such as replenishment with fixed costs or decentralized planning)

have received very little attention.

Notice that the centralized, multi-echelon models in this literature are no dif-

ferent than the serial or distribution systems studied in classical inventory theory;

the stochastic models involve the one-supplier, multi-retailer structure reminiscent

of Eppen and Schrage (1981) or the serial system of Clark and Scarf (1960). For

2The model of Lin and Chen (2003) has three echelons in its design: A cross-docking facility (central decision

maker) orders from multiple suppliers according to the demand at the retailers and the system constraints, and

allocates the perishable goods to retailers. However, the replenishment decisions are made for a single echelon: The

authors propose a genetic algorithm to solve for the single-period optimal decisions that minimize the total system

cost.

30

Page 40: Essays on Perishable Inventory Management

Rep

lenis

hm

ent

policy

Inven

tory

Exce

ssA

lloca

tion

Tra

ns-

No.

of

Fix

edC

entr

alize

d

Art

icle

Supplier

Ret

ailer

issu

ance

dem

and

rule

ship

men

tsre

tailer

(s)

cost

spla

nnin

g

Yen

(1975)

TIS

TIS

FIF

Oback

log

pro

port

ional,

−−2

−−yes

fixed

Cohen

etal.

TIS

TIS

FIF

Oback

log

pro

port

ional,

−−2

−−yes

(1981a)

fixed

Lyst

ad

etal.

(2006)

TIS

TIS

FIF

Oback

log

myopic

−−>

2−−

yes

Fujiw

ara

TIS

TIS

FIF

Olo

stat

supplier

,−−

−−2

−−yes

etal.

(1997)

(per

cycl

e)(p

erper

iod)

=LIF

Oex

ped

itin

gfo

rre

tailer

Kanch

anasu

nto

rnand

(s,S

)T

ISFIF

Oback

log

at

supplier

,FIF

O−−

2su

pplier

yes

Tec

hanit

isaw

ad

(2006)

lost

at

reta

iler

Pra

staco

s(1

978)†

−−−−

FIF

Olo

stat

reta

iler

myopic

rota

tion

>2

−−yes

Pra

staco

s(1

981)

−−−−

FIF

Olo

stat

reta

iler

myopic

rota

tion

>2

−−yes

Pra

staco

s(1

979)†

−−−−

LIF

Olo

stat

reta

iler

segre

gati

on

rota

tion

>2

−−yes

Nose

etal.

(1983)†

−−−−

LIF

Olo

stat

reta

iler

base

don

convex

rota

tion

>2

−−yes

pro

gra

mm

ing

Fed

ergru

en−−

−−FIF

Olo

stat

reta

iler

base

don

convex

rota

tion

>2

−−yes

etal.

(1986)†

pro

gra

mm

ing

Abdel

-Male

kand

EO

QE

OQ

−−−−

−−−−

1re

tailer

,yes

Zie

gle

r(1

988)

supplier

Ket

zenber

gand

ord

er0

ord

er0

FIF

Oex

ped

itin

gfo

rsu

pplier

,−−

−−1

−−yes

,

Fer

guso

n(2

006)

or

Qor

Qlo

stat

reta

iler

no

Tab

le2.

4:A

sum

mary

ofth

eanaly

ticalm

odels

on

reple

nis

hm

ent

and/or

allocati

on

decis

ions

inm

ult

i-echelo

nand

mult

i-lo

cati

on

syst

em

s(†

indic

ate

ssi

ngle

-peri

od

decis

ion

makin

g).

31

Page 41: Essays on Perishable Inventory Management

the one-supplier, multi-retailer system with non-perishables, it is known that order-

up-to policies are optimal under the ‘balance assumption’ (when the warehouse has

insufficient stock to satisfy the demand of the retailers in a period, available stocks

are allocated such that retailers achieve uniform shortage levels across the system;

this might involve ‘negative shipments’ from the warehouse, i.e., transshipments be-

tween retailers at the end of that period to achieve system balance), and optimal

policy parameters can be determined by decomposing the system and solving a se-

ries of one-dimensional problems (see for e.g., Diks and de Kok, 1998). There is no

equivalent of this analysis with perishable inventories, mainly due to the complexity

of the optimal ordering policy at a single location. Similarly, there is no work that

investigates continuous replenishment policies with perishables for multi-echelons or

multi-locations, although this is a well-studied problem for single-location models with

perishables (see Section 2.3.2) and non-perishables in multi-echelon supply chains.

Analytical Research on Allocation Decisions

All the models listed in Table 2.4 involve no capacity restrictions, and have a single

supplier who receives the freshest goods upon replenishment (although the goods

may not have the maximal lifetime at the time of arrival when lead time is positive).

However, the shipments from the supplier to the retailers can involve stock of any age

depending on the supplier’s inventory. Allocation and/or transshipment decisions

are simplified, for instance, when there is a single retailer (see Table 2.4) or when

the supplier’s inventory consists of goods of the same age. The latter happens in

the following two cases: (i) The goods start perishing at the retailer but not at the

supplier, i.e. all retailers are guaranteed to receive fresh goods from the supplier (see

Fujiwara et al., 1997, and Abdel-Malek and Ziegler, 1988). (ii) The lifetime of the

product is equal to the length of the periodic review interval. In that case, all the

goods at any location are of the same age and the goods perish at the end of one

cycle, as is the case for the supplier in the model of Fujiwara et al. (1997).

The focus on multi-echelon problems is on allocation decisions when the supplier

is assumed to receive a random amount of supply at the beginning of every period

(Prastacos, 1978, 1979, 1981) - we refer to this as the Prastacos model. The random

supply assumption is motivated by the application area – blood products which rely

32

Page 42: Essays on Perishable Inventory Management

on donations for supply. Two special distribution systems are considered in these

papers: (i) A rotation (or recycling) system where all unsold units that have not

expired at the retailers are returned to the supplier at the end of each period - these

units are distributed among the retailers along with the new supply of freshest goods

at the beginning of the next period. (ii) A retention system where each retailer keeps

all the inventory allocated. In the Prastacos model, the supplier does not stock any

goods, i.e. all the inventory is allocated and shipped to the retailers at the beginning

of a period, there are no inventory holding costs at any location, and the goal is to

minimize shortage and outdating costs that are uniform across all the retailers.

In a rotation system, the total number of units to outdate in a period depend

only on how the units with only one period of lifetime remaining are allocated in the

previous period, and the total amount of shortage depends only on how the inventory

is allocated, regardless of the age. Based on these observations, Prastacos (1978)

proposes the following myopic allocation policy that minimizes one-period system-

wide outdate and shortage costs: Starting with the oldest, the stocks of a given age

are allocated across the retailers so that the probability that the demand at each

location exceeds the total amount allocated to that retailer up to that point in the

algorithm are equalized, and this is repeated iteratively for items of all ages. Prastacos

(1978) also analyzes a retention system where the supplier only ships the fresh supply

to the retailers at the beginning of each period. In the one-period analysis of the

retention system, the amount to outdate at the end of a period depends on the

amount of oldest goods in stock and demand at each retailer, but does not depend

on the supply allocated in that period. Based on this observation, Prastacos (1978)

suggests a myopic allocation rule that equalizes the one-period shortage probability

at each retailer to minimize the one-period system-wide shortage and outdate costs.

Prastacos (1981) extends the analysis of the Prastacos model to the multi-period

setting and shows that the myopic allocation policy preserves some of the properties

of the optimal allocation that minimizes expected long-run average shortage and

outdating costs, and is, in fact, optimal in numerical examples with 2 retailers and

product lifetime of 2 periods. Since the cost parameters are the same for all the

retailers, the allocation resulting from the myopic rule is independent of the unit

costs of shortage and outdating.

Prastacos (1979) analyzes essentially the same single-period model assuming last-

33

Page 43: Essays on Perishable Inventory Management

in-first-out (LIFO) issuance, as opposed to FIFO in Prastacos (1978). In case of LIFO,

the optimal myopic allocation policies in both rotation and retention systems depend

on the unit costs of shortage and outdating. In addition, the optimal allocation

policies segregate the field inventory by age as opposed to a fair allocation where each

retailer receives goods of each age category. Under segregation, some retailers have

only newer goods and some only older goods so that system-wide expected outdates

are minimized. The optimal myopic allocation policy under LIFO for both rotation

and retention systems can be determined by solving a dynamic program with stages

corresponding to retail locations. Prastacos (1979) obtains the optimal allocation

rule for specific demand distributions. For rotation, he proposes a heuristic: First,

allocate the stock in order to equalize the probability of shortage at each retailer and

then swap the inventory among retailers to obtain field inventories that are segregated

by age.3

There are several practical extensions of the Prastacos model: The assumption on

uniform outdating and shortage costs can be relaxed, shipments/transportation costs

can be added, penalties can be incurred on leftover inventory (e.g., end of period

holding costs), transshipments among retailers can be enabled, and the supplier may

keep inventory as opposed to shipping all units downstream. The first three of these

issues have been addressed in the literature for only single-period decision-making:

Nose et al. (1983) and Federgruen et al. (1986) generalize the FIFO model of Prasta-

cos (1978) by assuming that there is a unit transportation cost for each item shipped

from the supplier to the retailers, and the unit outdating, shortage and transportation

costs are retailer-specific. They both develop convex programming formulations for

the single-period inventory allocation problem and propose algorithms based on the

Lagrangean relaxation to determine the optimal allocation. Their models rely on the

observation that the costs are only a function of the amount of old vs. fresh (i.e.,

stock that will outdate in one period vs. the inventory that has more than one period

of lifetime remaining) goods allocated to each retailer. In the rotation system consid-

3While they don’t provide a model nor the analysis, Huq et al. (2005) introduce a motivating example for allocating

inventories by segregation. They discuss that outdates and revenue loss due to markdowns or cost of returns can be

reduced by allocating goods to retailers based on the remaining lifetime of the goods and the expected depletion rates

at the retailers.

34

Page 44: Essays on Perishable Inventory Management

ered by Nose et al. (1983), the retailers are also charged per unit inventory returned

to the supplier at the end of each period (i.e., there is an end-of-period penalty on

leftover inventory at each location).

In the model of Yen (1975) and Cohen et al. (1981a), the supplier uses the FIFO

rule to determine which goods are to be shipped downstream and then uses one of

the following two allocation rules to determine the age-composition of the shipments

to the retailers in each period: (i) proportional allocation and (ii) fixed allocation.

In proportional-allocation, each retailer receives a proportion of goods of each age

category based on their share of the total demand. In fixed allocation, each retailer

receives a pre-determined fraction of goods in each period. Both of these allocation

rules are fair in that the shipments to retailers involve goods of each age category. Yen

(1975) and Cohen et al. (1981a) explore the optimality conditions for the parameters

associated with these allocation rules. They show that, under certain conditions,

there exists a fixed allocation rule that yields the same expected shortage, outdating

and holdings costs as a system operating under the optimal proportional allocation

rule. Their analysis can be extended to include multiple (> 2) retailers. Prastacos

(1981) shows that his myopic allocation rule is the same as proportional allocation

under certain probability distributions of demand.

Analytical Research on Replenishment and Allocation Decisions

Analysis of optimal replenishment policies for a serial, two-echelon system is presented

in Abdel-Malek and Ziegler (1988) assuming deterministic demand, zero lead times,

and price that linearly decreases with the age of the product. They determine the

economic order quantities (EOQs) for the retailer and the supplier by restricting the

order cycle lengths to be no more than the product lifetime.

Under demand uncertainty, several heuristic, discrete review replenishment poli-

cies have been considered and the focus has been on determining the optimal para-

meters for these restricted policies. These heuristic replenishment policies include the

TIS policies (Yen, 1975, Cohen et al., 1981a, Lystad et al., 2006, Fujiwara et al.,1997,

and Kanchanasuntorn and Techanitisawad, 2006), a “zero-or-fixed quantity” ordering

policy in Ketzenberg and Ferguson (2006) - denoted as ‘0 or Q’ in Table 2.4, - and

(s,S) policy for the retailers in Kanchanasuntorn and Techanitisawad (2006); analysis

35

Page 45: Essays on Perishable Inventory Management

in this latter is restricted to the case where the retailers’ demand is Normal. The re-

plenishment lead times are no longer than one period (i.e. goods are received no later

than the beginning of the next period) with exceptions being Lystad et al. (2006) and

Kanchanasuntorn and Techanitisawad (2006); the latter assumes that the lifetime of

the product is a multiple of the replenishment cycle lengths of the retailers and the

supplier, and that the supplier responds to retailers’ orders in a FIFO fashion (hence

allocation decisions are trivial).

Research that analyzes replenishment and allocation decisions jointly is confined to

the work of Yen (1975), its extension in Cohen et al. (1981a), and Lystad et al. (2006).

Yen (1975) and Cohen et al. (1981a) explore the structural properties of the expected

total cost function that includes expected holding, outdating and shortage costs,

when both the supplier and the retailer use TIS policies to replenish inventory. Fixed

and proportional allocation rules are investigated, and conditions on the existence

of unique target inventory levels under these rules are identified. Yen (1975) also

identifies conditions for the optimality of the proportional allocation rule for this sys-

tem. These conditions are satisfied, for instance, when the lifetime of the product

is restricted to two or three periods, or when the demand at each location and each

period is independent and identically distributed and the target inventory levels of

the retailers are the same. However, the analysis relies on one simplification: Inven-

tory recursions do not explicitly take into account the finite lifetime of the perishable

product, and goods that remain in inventory beyond their lifetime can be used to sat-

isfy excess demand, but are charged a unit outdating cost. Analysis of replenishment

and allocation policies that explicitly take these factors into account remains an open

problem of theoretical interest.

Lystad et al. (2006) use the myopic allocation rule of Prastacos (1981), and propose

heuristic echelon-based TIS policies. For a given system, they first determine the

best TIS policy via simulation. Then, they do a regression analysis to establish

the relationship between the order-up-to levels of this best policy and two heuristic

order-up-to levels: One heuristic is based on the newsvendor-based, approximate

echelon-stock policies for non-perishables and the other is the single-location heuristic

of Nahmias (1976) for perishables. The resulting regression model is then used in

computational experiments to study the effect of lifetime of product on system costs,

and to compare the performance of the approximation against policies that are derived

36

Page 46: Essays on Perishable Inventory Management

assuming the product is non-perishable. Thus Lystad et al. (2006) take a first step in

developing approximate echelon-based policies for perishables, and this topic deserves

more attention.

Notice that the effectiveness of the proposed allocation rules combined with good

replenishment policies have not been benchmarked in any of these studies, and various

simplifying assumptions have been made to derive the policies. There is a need for

further research on the analysis of replenishment and allocation decisions in multi-

echelon systems. Interesting research directions include investigation of the “balanc-

ing” of echelon inventories for perishables, similar to the balancing assumption in

Eppen and Schrage (1981), analysis of systems without making simplifying assump-

tions on inventory recursions (as in Yen, 1975), analysis of different system designs

(e.g. rotation and retention systems have been studied to some extent), or incorpo-

ration of different cost parameters to the models (e.g., the Prastacos model excludes

holding costs).

Simulation Models of Multi-Echelon Inventory Systems

In addition to the analytical research, simulation models have also been used in ana-

lyzing multi-echelon, multi-location systems with perishable goods; simulation models

present more opportunities in terms of model richness for this complex problem. The

earlier research (e.g., Yen, 1975, Cohen and Pierskalla, 1979, Cohen et al., 1981b) is

motivated mainly by managing regional blood centers and is reviewed in Prastacos

(1984). The simulation model of the single supplier, multiple-retailer system in Yen

(1975) includes returns of unused units from retailers to the supplier, variations of the

fixed and proportional allocation rules, expedited shipments to retailers, transship-

ments between retailers, and limited supply at the supplier. In addition to analyzing

the impact of inventory levels, allocation and transshipment rules on system costs,

Yen (1975) also looks at the impact of magnitude of demand at the retailers, and

observes that the system cost in his centralized model is more sensitive to the sizes

of the retailers rather than the number of retailers.

More recently, van der Vorst et al. (2000) describe a discrete-event simulation

model that can be used in the design/reconfiguration of supply chains of perishable

products. Their simulation model is used to analyze a fresh produce supply chain

37

Page 47: Essays on Perishable Inventory Management

with three echelons. Among other factors, van der Vorst et al. (2000), test the system

performance - measured in terms of inventory levels at the retailers and distribution

centers, and product freshness - using several scenarios. They report that supply

chain costs improve when production lead times are shorter, ordering and shipment

frequencies are higher, and new information systems are in place.

Katsaliaki and Brailsford (2007) present results of a project to improve procedures

and outcomes by modeling the entire supply chain for blood products in the UK. Their

simulation model includes a serial supply chain with the product flow that includes

collection of supply, processing/testing and storage at a service center, and shipment

to a hospital where blood is crossmatched/transfused4 for patients use. The model

includes multiple products with different shelf-lives. Six different policies vary in (i)

the type of products that are stocked at the hospital, (ii) the target inventory levels,

(iii) the time between crossmatching and release which can influence the amount of

unused and still usable inventory that is returned, (iv) the order trigger points for

expedited deliveries, (v) the inventory issuance rules for releases and returns, (vi)

the order and delivery lead times, and (vii) the number of daily deliveries to the

hospital. Performance is measured in terms of number of expired units, mismatched

units, amount of shortage, and number of routine and expedited deliveries. Final

recommendations include decreasing target inventory levels, having two routine de-

liveries per day as opposed to only one, a trigger point of 35% of the target inventory

level for expedited deliveries and reduction of crossmatch to release time. Note that

allocation decisions are not included in Katsaliaki and Brailsford (2007) because they

model a serial supply chain. Mustafee et al. (2006) provide the technical details of

the simulation model and the distributed simulation environment used in this latter

project.

While reporting the results of a simulation study, Cohen et al. (1981b) discuss

three practical ways of allocating inventory in a centralized system when inventory is

4One common practice in managing blood inventories is cross-matching, which is assigning units of blood from

inventory to particular patients. Jagannathan and Sen (1991) report that more than 50% of blood products held

for patients are not eventually transfused (i.e., used by the patient). The release of products that are cross-matched

enable re-distribution of inventories in a blood supply chain. See Prastacos (1984), Pierskalla (2004) and Jagannathan

and Sen (1991) for more information on cross-matching.

38

Page 48: Essays on Perishable Inventory Management

issued in a FIFO manner: In the first one, the supplier chooses a retailer and fills its

demand and goes on to fill the demand of the next retailer until all stock is depleted

or all demand is satisfied. In the second one, the supplier uses the proportional

allocation rule, and in the third one, the myopic allocation rule. Cohen et al. (1981b)

suggest using the second method in a practical setting because it has less information

needs (i.e. does not need the probability distribution of demand at each retailer like

the third method) and advise against using the first method in a centralized system

because it will lead to an imbalanced distribution of aging inventory. In addition,

the outdate probabilities of the retailers will vary significantly with the first method;

this increases the possibility of costly transshipments which are discussed in the next

section.

2.4.2 Logistics: Transshipments, Distribution, and Routing

Other than replenishment and allocation decisions, three of the critical logistics ac-

tivities in managing perishables in multi-location systems are transshipments, distri-

bution and collection (particularly for blood). Research that focuses on these three

decisions has been limited. Within the analytical work cited in Table 2.4, the ro-

tation system is the only form in which excess inventory is exchanged among the

retailers, and the exchange happens with a one period delay. Rotation systems are

also the basis for the goal programming model developed by Kendall and Lee (1980).

Prastacos and Brodheim (1980) develop a mathematical programming model for a

hybrid rotation-retention system to efficiently distribute perishables in a centralized

system. Both of these papers are motivated by operations of regional blood centers;

see Prastacos (1984) for a review of these and other earlier work on the distribution

and transshipment problems.

Note that rotation encompasses only indirect transshipments among the retailers.

The simulation model of Yen (1975) includes transshipments between retailers after

each location satisfies its own demand and serves as a guideline for the practical in-

ventory control and distribution system described in Cohen et al. (1981b). Cohen et

al. (1981b) suggest using transshipments in this practical setting if (i) the supplier is

out-of-stock, one retailer has an emergency need, and transshipping units from other

retailers do not significantly increase the probabilities of shortage at those retailers,

39

Page 49: Essays on Perishable Inventory Management

(ii) the difference between the shortage probabilities of two retailers when a unit is

transshipped from one to the other is greater than the ratio of the unit transporta-

tion cost to the shortage cost, or (iii) the difference between outdate probabilities of

retailers when a unit is shipped from one to the other is greater than the ratio of

transportation cost to the outdate cost. They use the terms emergency, shortage-

anticipating and outdate-anticipating transshipments, respectively, to denote these

three cases. Cohen et al. (1981b) point out that when all the retailers use optimal

TIS policies to manage their inventories, the amount of transshipments is insignifi-

cant based on simulation results. This emphasizes the need for effective replenishment

policies in multi-echelon and multi-location systems.

Federgruen et al. (1986), in addition to their analysis of the allocation decision,

consider the distribution of goods from the supplier to the retailers by formulating a

combined routing and inventory allocation problem. The decisions involve assigning

each location to a vehicle in the fleet and allocating fresh vs. old products among the

locations. The allocations do not affect the transportation costs and the route of

vehicles do not affect shortage and outdating costs. They propose exact and heuristic

solution methods. They also compare their combined routing and allocation approach

to a more hierarchical one where the allocation problem is solved first, and its solution

is used as an input to the distribution problem. Based on computational experiments,

the combined approach provides significant savings in terms of total transportation

costs, although these savings may not lead to a significant decrease in total costs

depending on the magnitude of the inventory related costs. However, the combined

approach has significant benefits when the number of vehicles used is few (where

the hierarchical approach may yield an infeasible solution). In addition to inventory

levels and allocation, the research conducted by Gregor et al. (1982) also examines the

impact of number of vehicles used in distribution on system-wide costs. Simulation

is used in the latter paper.

Or and Pierskalla (1979) consider daily vehicle routing decisions as a part of a

regional location-allocation problem where they also determine the optimal number

and location of blood centers, and the assignment of hospitals to the blood centers

that supply the hospitals on a periodic basis. They develop integer programming

models and propose heuristic solution methods. However, their model is designed

at an aggregate level and age of inventory is not considered. A similar problem is

40

Page 50: Essays on Perishable Inventory Management

studied by Hemmelmayr et al. (2006) where periodic delivery schedules and vehicle

routes are determined to distribute blood across a region.

Recent research on supply chain scheduling has addressed the need to effectively

distribute time-sensitive goods; Chen (2006) provides a survey of research in this

growing area. However, perishability is not modeled explicitly in this literature, rather

production orders are assumed to come from customers along with information on

delivery time windows and delivery due dates. Similarly, there are several articles

that model and solve real-life distribution problems of perishable products such as

dairy products, or food (e.g., Adenso-Diaz et al., 1998, Golden et al., 2001) where

aging or perishability is not explicitly modeled but is implicit in the time-windows.

More recently, Yi (2003) developed a model for daily vehicle routing decisions to bring

back supply (blood) from collection sites to a central location in order to meet the

daily target level of platelets (that can only be extracted within eight hours of blood

donation); this is a vehicle routing problem with time windows and time-dependent

rewards.

Notice that the research in this area has been confined to a single product, or

multiple products without age considerations. Interestingly, the distribution problem

posed by Prastacos (1984) still remains open: How can a distribution plan for a

centralized system be created to include shipments for multiple products, each with

a different lifetime and supply-demand pattern? Katsaliaki and Brailsford’s (2007)

simulation model provides only a partial answer to this question; their model involves

only one supplier and one retailer. Although challenging, analysis of the centralized

problem with multiple retailers and multiple products definitely deserves attention.

2.4.3 Information Sharing and Centralized/Decentralized Plan-

ning

Information technology paved the way for various industry-wide initiatives includ-

ing Efficient Consumer Response in the grocery industry; these initiatives aim to

decrease total system costs and inventories while improving availability of products

and customer satisfaction. A critical component of these initiatives is the sharing

of demand and product flow information among the suppliers, distributors and re-

41

Page 51: Essays on Perishable Inventory Management

tailers. Fransoo and Wouters (2000) discuss the benefit of sharing electronic point

of sale (EPOS) information for supply chains of two perishable products (salads and

ready-made pasteurized meals). Their empirical analysis suggests that the benefit of

EPOS would be higher for the supply chain of salad because of the magnitude of the

bullwhip effect observed. The reason for the higher bullwhip effect appears to be the

larger fluctuations in the demand for salad (e.g. when there is a sudden increase in

temperature, there is a spike in demand), associated shortage-gaming by the retail

franchisees, and the additional order amplification by the DC.

Information sharing and the value of information has been widely studied in the

general supply chain literature. Chen (2002) provides a survey of research on this

topic by focusing on where the information is coming from (such as the point-of-sale

data from downstream, or capacity information from upstream in the supply chain),

quantity, accuracy and speed of information, and centralized vs. decentralized plan-

ning in the supply chain (specifically he considers incentives for sharing information

and whether the environment is competitive or not). However, this rich literature

studies non-perishable goods or single-period models; the unique characteristics of

perishables are ignored, except by Ferguson and Ketzenberg (2006) and Ketzenberg

and Ferguson (2006).

Ferguson and Ketzenberg (2006), motivated by the grocery industry, investigate

the value of information for a retailer managing inventory. They focus on the re-

tailer’s replenishment problem and consider an infinite-horizon periodic review inven-

tory model for a single product with finite lifetime, lost sales, one-period lead time

and no outdating cost (see Section 2.3.1). The age of all units in a replenishment are

the same. The age of stock at the supplier is a random variable, and its distribution

is known to the retailer. In case of information sharing, the retailer knows exactly the

age of stock prior to giving an order. Ferguson and Ketzenberg (2006) quantify the

value of information on the age of stock under heuristic replenishment policies with

FIFO, LIFO or random issuing of inventory. Numerical experiments reveal that shar-

ing of product-age information enables the retailer to purchase fresher product and

the average percentage increase in profits via information-sharing is 4.4% and 3.4%,

and average percentage change in outdating is -51.4% and -6.3% for FIFO and LIFO

issuing, respectively. The experimental setup is further used to investigate the value

of other investments for the retailer to influence the issuance policy or the lifetime

42

Page 52: Essays on Perishable Inventory Management

of the product. An interesting finding is that investments that extend the product

lifetime provide a greater benefit than information sharing.

There are complications associated with different parties operating by different

rules in managing supply chains of perishables: For instance, the supplier can pre-

sumably induce the retailers to order more frequently by adapting an issuance and/or

replenishment policy that leads to more frequent outdates (e.g. using FIFO issuance

and/or having older stock in its inventory will enable the supplier to sell goods that

have a smaller shelf-life to the retailer). This is only mentioned in Ketzenberg and Fer-

guson (2006) - but not analyzed - and is ignored by other researchers. In that paper,

Ketzenberg and Ferguson (2006) study the value of information in a two-echelon set-

ting with one retailer and one supplier. Both parties replenish inventory heuristically,

the retailer’s order quantity in each period is either 0 or Q (which is an exogenous

fixed batch size) and issues inventory in a FIFO fashion, whereas the supplier uses the

same quantity Q in giving orders but need not place an order every period. In fact,

the supplier determines the timing of his replenishments by considering a safety lead

time. The retailer knows the supplier’s inventory state - including the age of items

in stock - and the supplier knows the retailer’s replenishment policy. Ketzenberg and

Ferguson (2006) quantify (i) the value of information regarding the inventory state

and replenishment policies in a decentralized system via numerical examples, and (ii)

the value of centralized planning. Based on computational experiments, they reach

the conclusion that the value of information for perishables can be significant, 70% of

the benefit of centralized planning comes from information sharing, and the increase

in supply chain profit is not always Pareto improving for both parties.

Note that all the papers we introduced so far focus on a single decision maker.

Likewise, Hahn et al. (2004) derive the optimal parameters of a TIS policy for a re-

tailer under two different contracts offered by the supplier; however, the supplier’s

optimal decisions are disregarded. Among the few studies that mention decentralized

decision-making, Popp and Vollert (1981) provide a numerical comparison of cen-

tralized vs. decentralized planning for regional blood banking. Problems that involve

multiple decision-makers, decentralized planning (vs. centralized) and coordination of

supply chains have been widely studied for non-perishables and/or using single-period

models (see, Chen, 2002 and Cachon, 2003). In practice, perishable products share

the same supply chain structure as many non-perishables, and decentralized planning

43

Page 53: Essays on Perishable Inventory Management

and/or coordination issues are just as critical. Furthermore, there are more challenges

for perishables due to cost of outdating, and possibly declining revenues due to aging.

However, research in this area has been scarce, and this issue remains as one of the

important future research directions. 5

2.5 Modeling Novelties: Demand and Product Char-

acteristics, Substitution, Pricing

The research we have reviewed so far includes models where the inventory of a single

product is depleted either in a LIFO or FIFO manner. Analysis of single location

models in Section 2.3 is confined to FIFO. Earlier research on single location models

has shown the difficulty in characterizing stationary distribution of stock levels under

LIFO even when the lifetime of the product is only two periods (see the comments in

Nahmias, 1982). Nahmias (1982) mentions that when the lifetime is two periods, the

order up to level in a TIS policy is insensitive to the choice of FIFO vs. LIFO despite

the difference in total system costs. Therefore, the replenishment policies/heuristics

developed under FIFO can also be used effectively for LIFO. However, replenishment

and issuance decisions may be interconnected - hence a more a careful analysis is

needed - under more general demand models.

Typically, excess demand is treated via backlogging or lost sales, with some papers

incorporating expedited delivery in their models (e.g., Fujiwara et al., 1997, Ketzen-

berg and Ferguson, 2006, Yen, 1975, Bar-Lev et al., 2005, and Zhou and Pierskalla,

2006). In practice, there is another way to fulfill the excess demand for a product:

Substitution. In the case of perishables, products of different ages often co-exist in the

market place, and inventory can be issued using rules more complicated than FIFO or

LIFO, allowing items of different ages or shelf-lives to be used as substitutes for each

other. This idea first appeared in the perishable inventory literature in Pierskalla

5There is research on coordination issues in supply chains with deteriorating goods: A permissible delay in

payment agreement between a retailer and a supplier is proposed in the deterministic model of Yang and Wee (2006)

to coordinate the supply chain. Chen and Chen (2005) studies centralized and decentralized planning for the joint

replenishment problem with multiple deteriorating goods.

44

Page 54: Essays on Perishable Inventory Management

and Roach (1972) who assume there is demand for any category (age) and that the

demand of a particular category can be satisfied from the stocks of that category or

using items that are fresher. Pierskalla and Roach (1972) show that FIFO is optimal

in this model with respect to two objectives: FIFO minimizes total backlog/lost sales

and minimizes outdates. The model has one simplification, though: The demand and

supply (replenishment) are assumed to be independent of the issuing policy. Since

issuing can potentially affect demand - fresher goods could lead to more loyal cus-

tomers - the study of models in which this assumption of independence is relaxed will

be important.

The motivation for many of the papers that involve substitution and age-dependent

demand streams come from health care. Cohen et al. (1981a) mention hospitals do-

ing special surgeries get higher priority for fresh blood. Haijema et al. (2005a, 2005b)

mention that platelets have 4-6 days of effective shelf-life, and 70% of the patients

requiring platelets suffer from platelet function disorder and need a fresh supply of

platelets (no older than 3 days) on a regular basis whereas the remaining 30% of

the patients who may lack platelets temporarily due to major trauma or surgery

do not have a strong preference with respect to the age of the platelet up to the

maximal shelf-life. For supply chains involving perishable goods other than blood,

substitution usually depends on customers’ choice and/or a retailer/supplier’s ability

to influence customers’ purchasing decisions. In their empirical research, Tsiros and

Heilman (2005) study the effect of expiration dates on the purchasing behavior of

grocery store customers. They conducted surveys to investigate consumer behavior

across different perishable product categories. They find that consumers check the

expiration dates more frequently if their perceived risk (of spoilage or health issues)

is greater. They also determine that consumers’ willingness to pay decreases as the

expiration date nears for all the products in this study; again finding that the de-

crease varies across categories in accordance with customer perceptions. Tsiros and

Heilman’s (2005) findings support the common practice of discounting grocery goods

that are aging in order to induce a purchase. However, they find that promotions

should differ across categories and across customer groups in order to exploit the

differences in customers’ tendencies to check the expiration dates and the differences

in their perceived risks across categories. In light of these motivating examples and

empirical findings, we provide below an overview of analytical research that considers

45

Page 55: Essays on Perishable Inventory Management

substitution, multiple products and pricing decisions.

2.5.1 Single Product and Age-Based Substitution

Research like that of Pierskalla and Roach (1972), where products of different shelf-

lives are explicitly modeled is limited. For a single product with a limited shelf-

life, substitution has been considered to sell goods of different ages: Parlar (1985)

analyzes the single-period problem for a perishable product that has two periods

of lifetime, where a fixed proportion of unmet demand for new items is fulfilled by

unsold old items and vice-versa, but his results do not extend to longer horizons.

Goh et al. (1993) consider a two-stage perishable inventory problem. Their model

has random supply and separate, Poisson-distributed demand streams for new and

old items. They computationally compare a restricted policy (where no substitution

takes place) and an unrestricted policy (where stocks of new items are used to fulfill

excess demand for old). Considering only shortage and outdating costs they conclude

that the unrestricted policy is less costly, unless the shortage cost for fresh units is

very high. Ferguson and Koenigsberg (2006) study a problem in a two-period setting

with pricing and internal competition/substitution. In their model, the demand for

each product in the second period is given by a linear price-response curve which is

a function of the price of both products as well as the quality deterioration factor of

the old product, and their decisions include the prices of both products as well as

the number of leftover units of old product to keep in the market. They investigate

whether a company is better off by carrying both or only the new product in the

second period.

Ishii (1993) models two types of customers (high and low priority) that demand

only the freshest products or products of any age, respectively, and obtains the optimal

target inventory level that maximizes the expected profits in a single period for a

product with finite lifetime. The demand of high priority customers is satisfied from

the freshest stock first, and then inventory is issued using FIFO in this model. Ishii

and Nose (1996) analyze the same model under a warehouse capacity constraint.

More recently, Haijema et al. (2005a, 2005b) study a finite horizon problem for blood

platelet production with a demand model of two types of customers similar to Ishii

(1993) and Ishii and Nose (1996). Haijema et al. (2005a, 2005b) formulate a Markov

46

Page 56: Essays on Perishable Inventory Management

Decision Process (MDP) model to minimize costs associated with holding, shortage,

outdating and substitution (incurred when the demand for a ‘fresh’ item is fulfilled

by older stock) costs. They assume inventory for the any-age demand is issued in

a FIFO manner from the oldest stock and fresh-demand is issued using LIFO from

the freshest stock. Haijema et al. (2005a, 2005b) propose a TIS and a combined

TIS-NIS heuristic, i.e., there is a daily target inventory level for total inventory in

stock and also the new items in stock. Computational experiments show that the

hybrid TIS-NIS policy is an improvement over TIS and that these heuristics provide

near-optimal inventory (production) policies.

For the sake of completeness of this chapter we should mention where our following

chapters stand in the literature. In the next two chapters we provide a detailed analy-

sis of the interplay between replenishment policies and substitution for perishables.

Our infinite horizon, periodic review formulation for a product with two periods of

lifetime includes lost sales, holding, outdating as well as substitution (both new-to-

old and old-to-new) costs. We assume two separate demand streams for new and

old items; demand can be correlated across time or across products of different ages.

We study different substitution options: The excess demand for a new (old) item is

satisfied from the excess stock of old (new), or is lost; the latter is the no-substution

case. Both LIFO and FIFO inventory issuance, as is common in the literature, can

be represented using our substitution model.

In our model the inventory is replenished using either TIS or NIS, and we identify

conditions for the cost parameters under which the supplier would indeed benefit

from restricted (only old-to-new, only new-to-old, or no substitution) or unrestricted

forms of substitution while using a practical replenishment policy. We show that

even when substitution costs are zero, substitution can be economically inferior to

no-substitution for a supplier using a TIS policy. Alternately, even when substitution

costs are very high, no-substitution is not guaranteed to be superior for a supplier

using TIS. These counter-intuitive properties are the side-effect of the TIS policy

which constrains reordering behavior. In contrast, more intuitive results under the

NIS policy exist as do conditions on cost parameters that establish the economic

benefit of substitution for this replenishment policy.

We do extensive computational experiments to quantify the benefits of substitu-

tion and to compare TIS and NIS. Given the literature up to this point, it is surprising

47

Page 57: Essays on Perishable Inventory Management

that NIS proves more effective than TIS and provides lower long-run average costs

except when the demand for new items is negligible. This is likely due to the pres-

ence of customers who pay a premium for newer product in our model. Similar to the

observation of Cohen et al. (1981b) on the limited need for transshipments (see our

discussion in Section 2.4.2), we show that the amount of substitution is small when

inventory is replenished using effective policies. Our work on perishables really only

begins the consideration of substitution for perishable items - items with longer life-

time, or substitution between different perishable products within the same category

(e.g., different types of fruit) remain to be investigated. Note that substitution is only

modeled as a recourse in the following chapters, and dynamic substitution where one

strategically sells a product of a different age than requested before the stocks of the

requested item are depleted, has not been studied.

2.5.2 Multiple Products

Nahmias and Pierskalla (1976) study the optimal ordering policies for an inventory

system with two products, one with a fixed, finite shelf-life and the other with an

infinite lifetime. The problem is motivated by the operation of a blood bank storing

frozen packed red cells.6 Demand is satisfied from the inventory of perishable product

first in a FIFO manner, any remaining demand is fulfilled from the inventory of non-

perishable product. Nahmias and Pierskalla (1976) analyze the structural properties

of the expected cost function in a finite-horizon, dynamic, discrete review system

and show that the optimal ordering policy in each period is characterized by three

choices: Do not order, order only product with the finite lifetime, or order both

products. Their results include monotonicity of order-up-to level of the perishable

product, e.g, the decrease in order up-to-level is higher with the increase in the stock

levels of newer items as opposed to old ones. They also show that if it is optimal to

order both products in a given period, then it is optimal to bring the total system-

wide inventory up to a level that does not vary with the on-hand inventory levels,

but with the time remaining until the end of the planning horizon.

6We refer the reader to Prastacos (1984) for earlier, simulation-based research on the effect of

freezing blood products on inventory management.

48

Page 58: Essays on Perishable Inventory Management

Multiple perishable products are also considered by Deuermeyer (1979, 1980).

Deuermeyer (1979) determines the one-period optimal order-up-to-levels for two prod-

ucts. In his model, the products are produced by two processes, one of which yields

both products and the other yields only one product. Deuermeyer (1980) determines

the single-period, optimal order-up-to levels for multiple perishable products, each

with a different lifetime. A critical assumption in the latter is the economic sub-

stitution assumption where the marginal total cost of a product is assumed to be

nondecreasing in the inventory levels of other products. Using the resulting prop-

erties of the single-period expected total cost function, Deuermeyer (1980) is able

to obtain the monotonicity results on order-up-to-levels for the single-period, multi-

product problem. His results mimic that of Fries (1975) and Nahmias (1975a) for

the single product problem. Specifically, these results show that the optimal order-

up-to level of a product is more sensitive to changes in stock levels of newer items

(as discussed above for Nahmias and Pierskalla, 1976), and that the optimal order

quantity decreases with an increase in the on-hand stock levels, while the optimal

target inventory level remains nondecreasing.

2.5.3 Pricing of Perishables

Pricing, in general, has become one of the most widely studied topics in the operations

management literature in the last decade. There is a significant body of research

on dynamic pricing and markdown optimization for ‘perishables.’ One well-studied

research problem in that domain involves determining the optimal price path for a

product that is sold over a finite horizon given an initial replenishment opportunity

and various assumptions about the nature of the demand (arrival processes, price-

demand relationship, whether customers expect discounts, whether customers’ utility

functions decrease over time etc.). We refer the reader to the book by Talluri and

van Ryzin (2004) and survey papers by Elmaghraby and Keskinocak (2003), and

Bitran and Caldentey (2003), for more information on pricing of perishable products.

In that stream of research, all the items in stock at any point in time are of the

same age because there is only one replenishment opportunity. In contrast, Konda

et al. (2003), Chandrashekar et al. (2003), and Chande et al. (2004, 2005) combine

pricing decisions with periodic replenishment of a perishable commodity that has a

49

Page 59: Essays on Perishable Inventory Management

fixed lifetime, and their models include items of different ages in stock in any period.

They provide MDP formulations where the state vector includes the inventory level of

goods of each age. Their pricing decisions are simplified, i.e., they only decide whether

to promote all the goods in stock in a period or not. Chande et al. (2004) suggest

reducing the size of the state of space by aggregating information of fresher goods

(as opposed to aggregation of information on older goods as in Nahmias, 1977a).

Performance of this approximation is discussed via numerical examples in Chande

et al. (2004, 2005), and sample look-up tables for optimal promotion decisions are

presented for given inventory vectors.7

Based on Tsiros and Heilman’s (2005) observations on customers’ preferences and

close-substitutability of products in fresh-produce supply chains, it is important to

analyze periodic promotion/pricing decisions across age-groups of products. There

are several research opportunities in this area in terms of demand management via

pricing to minimize outdates and shortages across age-groups of products and product

categories.

2.6 Summary

We presented a review of research on inventory management of perishable and aging

products, covering single-location inventory control, multi-echelon and multi-location

models, logistics decisions and modeling novelties regarding demand and product

characteristics. We identified and/or re-emphasized some of the important research

directions in Sections 2.2 to 2.5, ranging from practical issues such as product-mix

decisions and managing inventories of multiple, perishable products, to technical ones

such as the structure of optimal replenishment policies with fixed costs in the single-

product, single-location problem. We will now discuss the effect of substitution on

management of perishable goods and analyze inventory replenishment policies for

perishables under substitution in the following two chapters.

7Another paper that considers prices of perishable products is by Adachi et al. (1999). Items of each age generate

a different revenue in this model, demand is independent of the price, and the inventory is issued in a FIFO manner.

The work entails obtaining a replenishment policy via computation of a profit function given a price vector.

50

Page 60: Essays on Perishable Inventory Management

Chapter 3

Effect of Substitution on

Management of Perishable Goods

3.1 Introduction

Effective management of a supply chain calls for getting the right product at the right

time to the right place and in the right condition. This is never a simple task, and is

particularly difficult (and often costly) for supply chains containing perishable items.

Historically, product perishability, i.e. products changing chemically, losing value as

they age, has only been a major concern for items such as blood products, fresh

produce and pharmaceuticals (drugs, vitamins, cosmetics). Recently though, devel-

opments have expanded the scope of this problem: Increasing numbers of products

(e.g. in high-tech industries) are “aging” and losing value over time as newer prod-

ucts are introduced while older products are being phased out (products obsolesce).

Strictly speaking perishability and obsolescence are different phenomena - the former

happens at a predictable rate as products change, and the latter less predictably as

the market environment changes. Nevertheless, in both settings the same fundamen-

tal question needs to be addressed - how to manage inventory of a product that grows

increasingly less valuable over time. Answering this question incorrectly can be quite

costly: In a 2003 survey of distributors to supermarkets and drug stores, overall un-

salable costs within consumer packaged goods alone were estimated at $2.57 billion,

51

Page 61: Essays on Perishable Inventory Management

for the branded segment 22% of these costs were attributed to expiration (Grocery

Manufacturers of America, 2004). The $1.7 billion U.S. apple industry alone loses

$300 million annually to spoilage, according to Redi Ripe, an Albuquerque based

company that focuses on decreasing spoilage in the produce market (Webb, 2006).

And, Accenture estimates the average obsolescence of hand-sets and other accessories

amounts to 2.5-3% of total inventory purchase (about $550-675 million) annually for

the wireless industry in the United States (Colomina et al., 2006).

Further complicating the management of supply chains in general, and those for

perishables in particular, is the fact that increasingly, products of different “ages” - at

different phases in their life cycles - co-exist in the market at the same time. Differ-

ences in consumer preferences, price points, functionality of the products, and/or rate

of adoption have made overlapping of product life cycles not only viable, but often

advisable for many industries. A survey of electronics goods manufacturers in North

America, Europe and Asia Pacific revealed that the share of revenue that comes from

products introduced in the last 12 months was 31% to 37% while the average life of

a product was 44 months (Mendelson and Pillai, 1999). In the auto industry in the

United States, new models are introduced in the third quarter each year (e.g. 2006

Chevrolet Malibu Maxx SS is in the market in Fall 2005). Usually, these new models

- with only slightly upgraded designs - replace the previous year’s models that are

phased-out. Dealers and car manufacturers use various tactics to clear the inventory

of older models by the end of a year while increasing the sales of new ones. And,

we are all familiar with the practice of day-old vs. fresh bagels at bagel shops. In all

these examples, the predominant challenge for a seller is managing demand streams

for different-aged, substitutable products.

Typically it is the supplier who makes a substitute product available through an

explicit offer or via product placement, but the customer’s decision as to whether to

accept the substitution, possibly with some sort of compensation. One observes this

in several different contexts. For instance, certain critical surgeries and treatments

(trauma surgery and platelet transfusions for chemotheraphy patients) specifically ask

for blood products that have maximal freshness, thus creating demand for products

of different ages (Angle, 2003, Haijema et al, 2005a, 2005b, Cohen et al., 1981).

In case of shortage, a doctor may accept a slightly aged product if offered by the

blood bank. In the PC industry, where larger hard disks drive smaller hard disks

52

Page 62: Essays on Perishable Inventory Management

toward obsolescence, a PC manufacturer that sells direct to customers may install a

80-GB hard disk into a PC although the original request is for a 60-GB hard disk,

when the smaller hard disk is out of stock. In such a setting using a new product

as a substitute for an old (called manufacturer-driven substitution as in Chopra and

Meindl, 2001) is acceptable, but not vice-versa (at least without the consent of the

buyer). Sales incentives can entice a customer who is interested in a new model of a

car that is on backorder to buy an old one that is in stock. Mark-downs on grocery

goods close to their expiration dates can lead to a shift in demand of customers

that otherwise prefer longer shelf-life items. In fact, Tsiros and Heilman (2005) -

in empirical research - determine that consumers’s willingness to pay decreases as

the expiration date nears for various grocery products and their findings support the

common practice of discounting grocery goods that are aging in order to induce a

purchase.

In this chapter, we focus on managing inventories of goods that lose value over

time where there is demand for products of different ages. We focus on the supplier’s

decision of replenishing the inventory and investigate whether or not he/she benefits

from substitution. A supplier can often leverage some of the operational advantages of

substitution, such as risk pooling and higher inventory turnovers. Yet, despite these

advantages, it is not straightforward to analyze the overall effect of substitution in

a general setting over an extended horizon. What complicates this analysis is that

substituting items of different ages may entail direct substitution costs for the supplier

(e.g. rebates), and have additional, possibly unforeseen, future effects on the inventory

of goods on hand (depending on the supplier’s replenishment policy). Thus evaluation

of the long term, overall effect of substitution for products that lose value over time

remains an almost entirely open question. We focus on this question, illuminating

the factors that influence supplier’s performance in terms of the cost of inventory and

the freshness of goods in stock. We investigate when and how the supplier can benefit

from substitution, possibly at additional cost. Since the effect of substitution cannot

be assessed without considering the policy the supplier uses to replenish inventory,

we consider two simple and practical replenishment policies: Order-up-to policies

based on either the total amount of inventory of all ages, or the amount of new items

only. Our goal is to shed light on what cost parameters and demand characteristics

influence a supplier’s performance.

53

Page 63: Essays on Perishable Inventory Management

Our contributions include the following: (i) We study two practical replenishment

policies, featured in literature and practice. (ii) Under these policies, we show that

substitution may not always be economically beneficial for the supplier even if sub-

stitution costs are zero. Likewise, we show that no substitution may not always be

economically beneficial even if substitution costs are very high. (iii) We determine

the conditions under which a supplier will indeed benefit from substitution. (iv) We

study how the average age of goods in inventory is affected by substitution. (v) We

perform computational work quantifying the sensitivity of our results to problem pa-

rameters. (vi) All our analytical results are based on sample-path analysis, hence

no assumptions, save for ergodicity, are made regarding demand: We allow demand

to be correlated over time, across products, or both, as is common in practical settings

involving substitutable products (if not the literature).

We first define the supplier’s problem and introduce the notation and the formu-

lation in Section 3.2. A brief review of related literature is presented in Section 3.3.

We analyze the supplier’s performance in terms of costs and freshness of goods in Sec-

tion 3.4. To complement our analytical results, we provide several numerical examples

in Section 3.5. Conclusions are discussed in Section 3.6.

3.2 Problem Definition and Formulation

In our problem, there is a single product with a lifetime of two periods; the value of

the product decreases deterministically as it ages, and there may be random demand

for both new and old items. A single supplier replenishes new items periodically, with

zero lead time. At the end of each period, any remaining old items are outdated, while

any unused new items become old. This setup captures the essence of the problem

for perishable goods such as produce and blood as well as durable goods where a new

product is introduced into the market while the old one is phased-out periodically.

We denote by Xni (i = 1, 2) the amount of product with i periods of lifetime remaining

at the beginning of period n. The different demand streams for the items of different

ages are denoted Dni (i = 1, 2). Demand is stochastic and may have arbitrary joint

distribution; dependence between demand for different ages and over time is allowed,

but demand is assumed ergodic and independent of supplier decisions (we discuss this

54

Page 64: Essays on Perishable Inventory Management

latter assumption further below).

In any period, demand is fulfilled as follows: All demand for new (old) items is

fulfilled from the inventory of new (old) unless there is a stock-out. In case of stock-

outs, the excess stock of new (old) items may be used to satisfy the excess demand of

old (new). We use the term downward substitution to denote the case where a new

product is sold to a customer that demands old; upward substitution is the reverse.

We assume only a fraction, denoted 0 ≤ πD ≤ 1 and 0 ≤ πU ≤ 1, of customers accept

downward and upward substitution, respectively. Throughout the analysis, we assume

any such substitution occurs only as a recourse.1 Our model captures the situation

in which the supplier is the one making the substitute offer; this is supplier-driven

substitution as is common in business-to-business situations and is well-studied in

the literature on non-perishables (see Bitran et al., 2006, and the references therein).

On the other hand, our model likewise captures the scenario in which the consumer

is the one considering the substitute option without explicit intervention (except

product placement) by the supplier. This is the same as inventory-driven substitution

where the demand of one product shifts to another in case of a stock-out. This form

of substitution is common in retail. For more information on the classification of

substitution and related research in inventory management, see Bitran et al. (2006).

Note that the examples mentioned in the previous section (blood bank, PC man-

ufacturer, retail car sales, bagels) fit into this model. This model also captures other

phenomenon: If inventory is depleted starting from the newest items (e.g. buyers

prefer longer shelf-life items as in grocery and the freshest produce/bread), this cor-

responds to having demand for new items only and fulfilling the demand by upward

substitution (i.e. customers consider buying less fresh products only after freshest is

sold). If instead inventory is depleted in a FIFO fashion (oldest item issued first)

as is the assumption in classical research on perishable goods (see the review in the

next section), then there is demand for only old items and demand is fulfilled using

downward substitution. Thus, our model subsumes many of the previous models in

the literature.

Furthermore, we assume any demand unsatisfied at the end of a period is lost (but

backordering new item demand does not change our results). Based on this setting,

1We discuss this assumption further in Section 3.4.4.

55

Page 65: Essays on Perishable Inventory Management

we can define the following quantities at the end of period n, where x+ denotes

max(x, 0):

dsn = min{πD(Dn1 −Xn

1 )+, (Xn2 −Dn

2 )+} is the downward substitution amount;

usn = min{πU(Dn2 −Xn

2 )+, (Xn1 −Dn

1 )+} is the upward substitution amount;

Ln1 = [(Dn

1 −Xn1 )− dsn]+ is the amount of lost sales for old items;

Ln2 = [(Dn

2 −Xn2 )− usn]+ is the amount of lost sales for new items;

On = [(Xn1 − Dn

1 ) − usn]+ is the amount that outdates at the end of the period

and

Xn+11 = [(Xn

2 −Dn2 )−dsn]+ is the amount of inventory carried to the next period.

We consider two practical replenishment policies for the supplier:

TIS (Total-Inventory-to-S): At the beginning of every period the total inventory

level (old plus new items) is brought up to S. That is, for all n,

Xn1 + Xn

2 = S. (3.1)

NIS (New-Inventory-to-S): At the beginning of every period S new items are or-

dered. Hence, for all n,

Xn2 = S. (3.2)

The TIS policy is simple, common and has a long tradition in the literature.

NIS has not attracted as much attention as TIS from a research perspective, mainly

because of its inferior performance in earlier simulation studies (see the next section

for the literature review). However, NIS has remained in use in practice due to its

simplicity (Prastacos, 1984, Angle, 2003). Note that the order-up-to level S can be

determined by optimization or by practical rules of thumb in either replenishment

policy.

In each period, the order of events is as follows: First inventory is replenished,

then demand occurs and is fulfilled, with substitution occuring only as a recourse.

Unsatisfied demand for both new and old items are lost. Stocks are aged; new goods

become old and old goods perish. Finally, profits are assessed and the supplier places

the order that will arrive in the next period.

56

Page 66: Essays on Perishable Inventory Management

The supplier’s profit function involves revenues associated with sales, inventory

related costs and revenue/cost terms associated with substitution. There is a unit

revenue ri for age i product and we assume r2 ≥ r1; i.e. newer items are at least as

valuable as older ones. These revenues are obtained only if the demand is fulfilled

without substitution. In case of a stock-out of old items, a customer asking for an

old item which is priced at r1 may buy an unsold new item at the price of r1 + α′D

where α′D ≥ 0; a new product made available as a substitute is never priced below

the old one. Similarly, in case of upward substitution, a customer asking for a new

item which is priced at r2 will buy an unsold old item at the price of r2 − α′U where

α′U ≥ 0.

There are also explicit costs associated with upward or downward substitution:

The supplier incurs a downward substitution cost of α′′D per unit and an upward

substitution cost of α′′U per unit. These are in addition to the revenue loss captured by

α′D and α

′U , and may represent additional production or shipping costs for the supplier

(e.g. shipping goods from different locations), the loss of goodwill, or a “demand

diversion” cost for the supplier2. In addition, there is a unit loss of goodwill cost of

p′i for age i product in case of lost sales. There is also an inventory carrying cost of

h per unit, and an outdating cost of m per unit.

Based on these parameters, the long-run time-average profit of the supplier as a

function of the order-up-to level S (either using TIS or NIS) when the system starts

with 0 ≤ X12 ≤ S new and 0 ≤ X1

1 ≤ S old items is :

R(S) = limT→∞

1

T

T∑

n=1

r2 min(Dn2 , Xn

2 ) + r1 min(Dn1 , Xn

1 ) + (r2 − α′U) usn + (r1 + α

′D)dsn

− hXn+11 − p

′1L

n1 − p

′2L

n2 −mOn − α

′′U usn − α

′′D dsn.

In the above formulation, the only difference between TIS and NIS is in the inventory

recursions regarding the inventory levels of new items (see equations (3.1) and (3.2)).

2Demand diversion takes place in the following way: Customers may request old items in hopes

of receiving new ones, if they realize the supplier is practicing downward substitution. Thus, the

supplier’s substitution decisions may affect demand, in this case, causing some demand shift from

new demand to old demand, which is likely to decrease the supplier’s profits. While our work does

not explicitly model this (or other related) phenomena linking substitution decisions with future

demand, our substitution costs may be taken as a high-level proxy for this.

57

Page 67: Essays on Perishable Inventory Management

Rearranging the terms, the profit function can be simplified:

R(S) = limT→∞

1

T

T∑

n=1

r2(Dn2 − Ln

2 ) + r1(Dn1 − Ln

1 ) (3.3)

− hXn+11 − p

′1L

n1 − p

′2L

n2 −mOn − (α

′U + α

′′U) usn − (α

′′D − α

′D) dsn

= limT→∞

1

T

T∑

n=1

r2Dn2 + r1D

n1

− hXn+11 − p1L

n1 − p2L

n2 −mOn − αU usn − αD dsn

where αU = α′U + α

′′U is the net cost of upward substitution, αD = α

′′D − α

′D (which

can be negative) is the net cost of downward substitution and pi = ri + p′i is the net

cost of lost sales for i = 1, 2.

Note that analysis of the above profit function only requires analysis of an appro-

priately defined cost function, as the first two terms in (3.4) are fixed. We define C(S)

to be the long-run time-average cost of the supplier as a function of the order-up-to

level S:

C(S) = limT→∞

1

T

T∑

n=1

hXn+11 + p1L

n1 + p2L

n2 + mOn + αU usn + αD dsn . (3.4)

In the rest of the chapter, we analyze the benefit of substitution based on the cost

function C(S). We use time-average costs to avoid issues which may arise in taking

expectations, for example should stationary distributions fail to exist due to demand

characteristics. Note though that, the inventory position (Xn2 , Xn

1 ), under TIS or

NIS, forms an ergodic process (as demand is assumed ergodic); thus all costs can

be expressed as functions of the time-average distribution of the inventory, which

converges. In addition to the economic benefit, we also study the effect of substitution

on the freshness of goods in stock. Freshness is defined as the long-run time average

age of goods in stock which can easily be determined using the inventory levels Xn1

and Xn2 ; discussion on this topic appears in Section 3.4.5.

Note that the formulation above allows for both upward and downward substitu-

tion modified by the fraction of customers accepting the substitution offers. However,

as discussed in the previous section, different forms of substitution are observed in dif-

ferent industries. While downward substitution is very common (e.g. a newer product

is usually an acceptable choice if one needs hard-drives, cars, blood, bakery products,

fresh produce etc.), upward substitution can be practiced only in certain cases and

58

Page 68: Essays on Perishable Inventory Management

then possibly only by using sales incentives (e.g. end of year clearance of old models in

car dealers, mark-downs of grocery goods closer to their expiration dates). In order to

investigate the relative benefit of different substitution patterns for the supplier, we

consider four specific cases: The No-Substitution case (denoted N ) is the base case

where items of a certain age are used to fulfill the demand for products of only that

age. The inventory levels and total cost (or profit) of this policy are easily computed

in our model by setting πU = πD = 0. On the other extreme is Full-Substitution, (de-

noted F), where both downward and upward substitution can take place; this is the

general model introduced above with πD > 0, πU > 0. There are two restricted cases

which consider only one-way substitution: The Downward-Substitution case, denoted

D, is when πU = 0 and Upward-Substitution case, denoted U , is when πD = 0.3

Our main research question is: Under what conditions one would prefer one form

of substitution - coupled with a practical replenishment policy - over the others? We

first provide partial answers to this question in Section 3.4. We identify conditions on

cost parameters such that supplier-driven or inventory-driven substitution results in

(i) lower costs for the supplier and (ii) fresher goods for the customers; our analytical

results are based on pairwise comparison of substitution options given a replenishment

policy. In Section 3.5, we present results of extensive computational experiments and

provide guidelines on when a replenishment policy and substitution option dominates

the others based on demand characteristics and values of cost parameters.

3.3 Related Work in the Literature

Most of the work on inventory problems for perishable goods focuses on optimal or

near-optimal ordering policies to minimize operating costs when the demand is in-

different to the age of the product and is fulfilled according to FIFO (oldest items

issued first), hence ignoring substitution issues. We refer to this specific model as

single demand stream as opposed to the two demand streams in our model. Even

in the simplified single demand stream setting, the problem is very difficult: Unlike

3Throughout the paper “Upward-Substitution” (or U) and “Downward-Substitution” (or D)

denote the specific case as introduced here. By “downward substitution” or “upward substitution”

we mean the substitution event itself.

59

Page 69: Essays on Perishable Inventory Management

standard inventory control theory, where generally the only information needed is

the inventory position, the optimal ordering policy for perishables requires informa-

tion about the amount of inventory of every age. Therefore the state space (and

problem complexity) grows with the lifetime of the product. For two-period lifetime

problem, Nahmias and Pierskalla (1973) showed that the optimal order quantity is a

non-linear decreasing function of the state (on hand inventory); this was extended by

Fries (1975) and Nahmias (1975a), independently, to the general m-period lifetime

problem. The fact that the optimal policy requires solution of a multi-dimensional

dynamic program, which is computationally difficult, motivated exploration of heuris-

tic methods. Nahmias (1976) and Chazan and Gal (1977) proposed the fixed-critical

number (order-up-to) policy, in which orders are placed at the end of each period to

bring the total inventory summed across all ages to a specific level. A closed-form

method for the optimal order-up-to level of two-period lifetime problem was found

by Van Zyl (1964). (This corresponds to a special case of our model: There is no

demand for new items, demand for old is fulfilled with downward substitution, and

inventory is replenished according to the TIS policy.)

Still for a single demand stream, Nahmias (1976, 1977) and Nandakumar and

Morton (1993) show that order-up-to policies (TIS) perform very well compared to

other methods, including optimal policies; they develop and analyze heuristics to

choose the best order-up-to level. Cooper (2001) provides further analysis of the

properties of the TIS policy, while Nahmias (1978) shows that when there is a fixed

ordering cost, an (s, S) type heuristic is better than order-up-to policies. Tekin et

al. (2001) consider a continuous review model for perishables and study a modified

lotsize-reorder policy. More recent papers on inventory management of perishables

are by Ketzenberg and Ferguson (2006), and Ferguson and Ketzenberg (2006) - they

focus on information sharing in a supply chain. A summary focused on blood bank

supply chains is provided by Pierskella (2004). All works cited above consider a

single demand stream being fulfilled according to FIFO; FIFO is known to be optimal

for many fixed-life inventory problems with a single demand stream (Pierskalla and

Roach, 1972).

TIS has been shown to have good performance compared to optimal when inven-

tory is depleted using oldest items first (Nahmias, 1976, Nandakumar and Morton,

1993). Haijema et al. (2005a, 2005b) use a Markov Decision Process approach to de-

60

Page 70: Essays on Perishable Inventory Management

rive heuristics for replenishment of blood platelets under Full-Substitution. They find

that base-stock heuristics that replenish based on total stock in system do well (TIS),

and those that simultaneously take into account total stock and new stock in system

(a TIS-NIS hybrid) do even better in a multi-period lifetime (5 to 7 days) setting.

First proposed by Brodheim et al. (1975), NIS was shown to have inferior performance

compared to TIS, with the earlier demand models and inventory replenishment rules.

Nahmias (1975b) compared multiple heuristics, including the “optimal” TIS policy, a

piecewise linear function of the optimal policy for the non-perishable problem, and a

hybrid of this with NIS ordering. He found that the first two policies outperform the

third. In another simulation study, the cost of the NIS policy was found to be 2-5%

higher than that of the TIS (see Prastacos, 1984, and the references therein). These

observations practically ended the research consideration of NIS and its variants.

There are a few papers that model multiple types of customers or demand streams

for perishables. Nahmias and Pierskalla (1976) examine a system consisting of two

products: one with a lifetime of m periods, and one with an infinite lifetime. There

is a one-way substitution of the second product (non-perishable) for the first (perish-

able). The problem was motivated by the operation of a blood bank storing frozen

packed red cells. Parlar (1985) considers a perishable product that has two-periods of

lifetime, where a fixed proportion of unmet demand for new items is fulfilled by unsold

old items and vice-versa (in our framework this is Full-Substitution), but only for a

single period. His results do not extend to longer horizons. Goh et al. (1993) consider

a two-stage perishable inventory problem. Their model has random supply and sep-

arate demand streams modelled as Poisson process. They computationally compare

a restricted policy (similar to our No-Substitution option) and an unrestricted pol-

icy (similar to our Downward-Substitution option). Considering only shortage and

outdating costs they conclude that the unrestricted policy is less costly, unless the

shortage cost for fresh units is very high. Ishii et al. (1996) focus on two types of

customers (high and low priority), items of m different ages with different prices and

only a single-period decision horizon. High priority customers only buy the freshest

products, so the freshest products are first sold to the high priority customers and

the remaining items are issued according to a FIFO policy. They provide the optimal

ordering policy under a warehouse capacity constraint. Ferguson and Koenigsberg

(2004) study a problem in a two-period setting with pricing and internal competition

61

Page 71: Essays on Perishable Inventory Management

between new and old items.

There is another line of research related to our work: Inventory management of

substitutable products. Substitution and demand fulfillment features of our model

- specifically substitution taking place as a recourse and a fraction of customers ac-

cepting substitutes - are common assumptions in this literature.

McGillivray and Silver(1978) provide analytical results for two products with simi-

lar unit and penalty costs and unmet demand can be fulfilled from the other product’s

inventory (if there is any inventory left). Parlar and Goyal (1984) study two differ-

entiated products with a fixed substitution fraction and analyze structural properties

of the expected profit function. Their work is extended by Pasternack and Drezner

(1991) by adding revenue earned from a substitution (different from the original sell-

ing price), shortage cost, and salvage value. Rajaram and Tang (2001) subsumes the

existing models for the single-period two-product problem. They show that demand

substitution between products always leads to higher expected profits than the no-

substitution case, and provide a heuristic to determine the order quantities. We show

such a dominance relation need not hold in an infinite-horizon, perishable setting.

There are several other papers looking at single-period inventory problems with mul-

tiple products that are substitutable: Parlar (1988), Wang and Parlar (1994), Bassok

et al. (1999), Ernst and Kouvelis (1999) Smith and Agrawal (2000), Mahajan and

van Ryzin (2001), Avsar et al. (2002), Eynan and Fouque (2003). The reader can

refer to Netessine and Rudi (2003) or Bitran et al. (2006) and the references therein.

Although it is a crucial element in our model, the age of a product is irrelevant from

a substitution standpoint in this particular literature.

Moorthy (1984) developed a theory of market segmentation based on consumer self

selection; prior his work the theory on market segmentation assumed a firm decides

which market segment to sell without incorporating customer choice. Furthermore

his consumer self-selection approach enabled to model “cannibalization” and compe-

tition among firms. In our problem we, similarly, model different market segments

(demand streams) for old and new products but we do not model cannibalization or

competition explicitly (although substitution costs might be considered as a proxy for

such consumer based dynamics in a very general way). In terms of consumer choice

our model does not have any restrictions on upward substitution. That is, consumer

may refuse the upward substitution and our analytical results will not be affected.

62

Page 72: Essays on Perishable Inventory Management

However for downward substitution, we assume that customer and seller will always

find a price that both can agree upon, and the downward substitution will take place.

This price would probably be between the original selling prices of old and new items.

In our model we have a downward substitution cost that could be considered as a

result of the agreement between the parties.

In our work we do not directly deal with management of product variety however it

can be considered as a related research area as we have two substitutable products in

our problem. Chen et al. (1998) present a framework that helps a firm to position and

price a line of products capturing a variety of delivery cost models. They discuss how

the optimal product line can be obtained by finding a set of cross points. These cross

points determine how the heterogeneous customers are divided among the products.

In our model however, pricing of old and new products is not studied explicitly.

Yano et al. (1998) provide a comprehensive review of mathematical models on design

and pricing for a line of products. In this review the models are are divided into

categories based on manufacturing costs. Manufacturing costs nor pricing are not

the focus of our study; our focus is on other operational costs such as outdating or

substitution costs. Reader can refer to Ho and Tang (1998) and references therein for

more information on management of product variety.

Our study is first in the literature on substitution to introduce the term “dynamic

quality” . In the literature on substitution, quality of a product is static. For our

problem quality of a product is age (one dimensional). That is, a new product has

a higher quality than an old one. Since our model is an infinite horizon model, the

quality of product is dynamic. If a product is not sold its quality will change in the

next period (i.e. from old to new) making our model unique in the literature on sub-

stitution. We have a general cost structure (crucially, we include substitution costs)

and demand fulfillment structure, in which different forms of substitution are mod-

eled. Further, we analytically provide conditions guaranteeing dominance between

different forms of substitution over an infinite horizon. Our work has several unique

aspects: (i) We formally define and evaluate four substitution options for perishable

products (and, in general, for products that lose value over time). (ii) We compare

these different forms of substitution based on two measures (total cost and average age

of inventory in the system), providing analytical proofs or counter-examples for dom-

inance relations. (iii) Our results are free of distributional assumptions on demand,

63

Page 73: Essays on Perishable Inventory Management

save for ergodicity; demand can be auto-correlated or correlated with the product of

different age. (iv) We consider an infinite horizon problem as opposed to single period

or single replenishment models common in analysis of substitutable products. (iv)

The model with a single stream of demand and FIFO (or LIFO) inventory depletion

– standard in the literature on perishable goods – remains a special case of ours.

(v) We analyze two distinct, practical replenishment policies, and show that the NIS

policy which was known to have inferior performance compared to TIS, can in fact

be better with more general demand models and leads to more intuitive results under

substitution.

3.4 Effect of Substitution: Analysis of Economic

Benefit and Freshness

In this section, we identify conditions on cost parameters that guarantee that a specific

form of substitution is superior to others. The main question here is not which form

of substitution is best but when the practice of upward, downward or both forms of

substitution would indeed be beneficial. We show that reasonable parameter settings

exist that lead to one option being superior than another, in an almost sure sense.

We also identify parameter regions where no such dominance conditions can exist.

Throughout, we use the notation  to denote dominance between two substitution

options: F Â D means that for any ordering level S the time-average cost of F is no

more than that of D in an almost sure sense, where these costs are defined according

to equation (3.4). All of our cost comparisons are based on key results regarding

inventory levels of different substitution options, when a common order-up-to level S

is used. Since the pairwise cost-dominance results do not change if optimal order-up-

to levels are used, the choice of S does not influence our cost comparison results.

Our analytical results on the economic benefit of substitution rely on two condi-

tions. First, the dominance results are valid for p2 > p1. This assumption is not

restrictive: We defined the net cost of lost sales as pi = ri + p′i and assumed r2 ≥ r1

in our model. In many practical cases, the newer item is more valuable (it is the pre-

mium product), hence it is priced higher and bears a higher loss of goodwill. Second,

all the customers accept downward substitution whenever offered, i.e. πD = 1 for

64

Page 74: Essays on Perishable Inventory Management

options F and D. Although this is a special case of our model, it is not unrealistic:

It is highly likely that the substitute will be accepted when a supplier offers a newer

item (the premium product) with a discount as a substitute for the old. We later

provide numerical results on cases where this latter condition is relaxed. None of

these conditions are required for our results on freshness in Section 3.4.5.

3.4.1 The economic benefit of substitution: Overview of re-

sults

We first provide an overview of our results on cost comparisons before going into

a detailed mathematical analysis. We seek answers to two main questions: Given

an order-up-to level, inventory related costs, substitution related parameters, and

under any demand stream, “can the supplier be better off without substitution?” and

“can the supplier be better off by practicing full- or restricted forms of substitution?”

Pairwise comparison of substitution options provide the answers.

Consider N vs. U : Note that upward substitution uses an unsold old item that

would otherwise outdate, to satisfy the excess demand of a new item. For every item

substituted, the supplier saves outdating (m) and net lost sale (p2) costs, but incurs

the net substitution cost (αU). Therefore, we prove that upward substitution provides

value when:

αU ≤ m + p2. (3.5)

This result is intuitive; the supplier benefits from not allowing substitution (i.e. N ÂU) regardless of πU when the substitution cost is higher than a critical level. Note

that this observation is true for both replenishment policies TIS and NIS.

One would expect a similar result for N vs. D. In downward substitution, an

excess new item is substituted to satisfy the excess demand for old items. Intuitively,

if the substitution cost is low enough, D should be beneficial, and if it is too high,

supplier should be better off with N . However, our analysis below shows that this

need not be the case when supplier replenishes inventory using TIS: Even if αD →∞,

N can be worse than D under TIS, and if αD = 0, D can be worse than N . This is

a significant result since, based on our discussion in the previous sections, D is the

most common form of substitution that takes place in practice.

65

Page 75: Essays on Perishable Inventory Management

Our results for all pairwise comparisons are summarized in Table 3.1 for the sake of

completeness. As discussed above, under NIS and TIS, αU ≤ m+ p2 is necessary and

sufficient for upward substitution to result in lower costs but downward substitution

is different: The benefit of downward substitution under general demand scenarios

often cannot be established for TIS, while intuitive conditions on αU and αD do exist

under NIS. For instance, if the net downward substitution cost exceeds the sum of

holding, outdating and lost sales cost for old items (the worst thing that can happen

if you choose not to substitute), D can be ruled out under NIS but not under TIS.

Note also that in many, but not all cases, the conditions for policy dominance are

subsets of each other, for example those for D to dominate N and U to dominate Ntogether form the conditions for F to dominate N under NIS.

We summarize some of the other key takeaways from our results: (i) If the supplier

is using TIS, downward substitution reduces costs only if αD ≤ p1− p2, requiring the

downward substitution cost to be negative (which is possible by definition of αD);

even if the downward substitution cost is 0, it is not guaranteed that the supplier

will be better off offering downward substitutes. (ii) On the other hand, even if

downward or upward substitution costs are very high (αD → ∞ and/or αU → ∞),

avoiding substitution is not guaranteed to be the best action for the supplier under

TIS – i.e. conditions do not exist to guarantee the dominance of the no-substition

policy (N ) over the full-substition policy (F) no matter how large the substitution

costs may be. (iii) If the supplier is using NIS, downward substitution reduces costs if

the downward substitution cost is less than the holding cost (αD ≤ h); and downward

substitution should be avoided if the downward substitution cost is more than the

sum of the penalty cost for old items, the outdating cost and the holding cost (αD >

m+h+p1). In case of supplier-driven substitution, these results show when a supplier

benefits from offering a substitute product to a customer. In case of inventory-driven

substitution as in retail, these results can be interpreted as to when a supplier should

make choices (e.g. product placement, pricing) to encourage or discourage customers

to practice substitution.

TIS has attracted nearly all the attention in the literature on perishable goods, and

it is known to be an effective replenishment policy for a special case of our model (only

old items are demanded and downward substitution is employed with no substitution

costs). However, our analysis reveals that it may have counter-intuitive performance

66

Page 76: Essays on Perishable Inventory Management

Dominance Condition(s) under TIS Condition(s) under NIS

F Â D, U Â N αU ≤ m + p2 αU ≤ m + p2

D Â F , N Â U αU ≥ m + p2 αU ≥ m + p2

F Â N αU ≤ m + p2 αU ≤ m + p2

αD ≤ p1 − p2 αD ≤ h

N Â F Does Not Exist αU ≥ m + p2

αD ≥ m + h + p1

F Â U αD ≤ p1 − p2 αD ≤ min{h, h + αU − (p2 − p1)}U Â F Does Not Exist αU ≤ m + p2

αD ≥ m + h + p1

D Â N αD ≤ p1 − p2 αD ≤ h

N Â D Does Not Exist αD ≥ m + h + p1

U Â D Does Not Exist αU ≤ m + p2

αD ≥ m + h + p1

D Â U αU ≥ m + p2 αU ≥ m + p2

αD ≤ p1 − p2 αD ≤ h

Table 3.1: Sufficient conditions on substitution costs that ensure dominance relations

between substitution options

when demand is differentiated by the age of the product, and substitution takes place

with additional costs. In contrast, NIS - which was deemed inferior in earlier studies

- has more predictable behavior under substitution.

All the conditions listed in Table 3.1 hold for 0 ≤ πU ≤ 1. Note that some of the

bounds we obtain are quite “conservative” because we are seeking strong, almost sure

results, over the entire region of πU values. For instance, the bound αD ≤ p1−p2 that

is needed for F to dominate U and N under TIS, is restrictive; a weaker condition

on αD exists when πU = 1 (see Lemma 5 and 8 in Section 3.4.2). Likewise, when αD

67

Page 77: Essays on Perishable Inventory Management

satisfies h < αD < m + h + p1, no dominance relation between D and N is provided

under NIS. In Section 3.5, we use computational experiments to investigate the effect

of values of cost parameters for which dominance relations are not established, as well

as values of πD < 1 (recall that the above results rely on the property that downward

substitution is always accepted; i.e. πD = 1 in F and D).

Next, we show how the conditions in Table 3.1 are obtained. We provide the

results on TIS in Section 3.4.2, and on NIS in Section 3.4.3.

3.4.2 Economic Benefit of Substitution under TIS

A deeper understanding of the effect of substitution on stock levels is needed to

establish the cost-dominance conditions. The first key observation is the following

proposition: Along any sample path of demands within either TIS or NIS, F and Dhave equal amounts of stock in any period for a given S, and so do U and N . We

provide this and other important results below.

Preliminary Results under TIS

Proposition 1 For TIS or NIS, F and D have identical inventory levels in any

period for a given S under the same sample path of demands, so do U and N .

For any given S, the order quantities (TIS or NIS) are independent of the upward

substitution decision under identical demand. The reason is as follows: In any period

if upward substitution is done, it will not affect the ordering decision in the next

period as the old item inventory level for the next period will be zero. In case upward

substitution is not done any unsold old items will perish. Therefore the choice of

practicing upward substitution does not affect the inventory levels for the next period

as long as the downward substitution policy and the ordering policy (TIS or NIS) are

the same. Thus F and D are identical in terms of inventory levels; so are U and N .

Proof

XnN = S − (Xn−1

N −Dn−12 )+ > S − [(Xn−1

F −Dn−12 )+ − (Dn−1

1 − S + Xn−1F )+]+ = Xn

F

68

Page 78: Essays on Perishable Inventory Management

(Xn−1N −Dn−1

2 )+ < [(Xn−1F −Dn−1

2 )+ − (Dn−11 − S + Xn−1

F )+]+

< (Xn−1F −Dn−1

2 )+.

(Xn−1F − Dn−1

2 )+ must be greater than zero in order to satisfy the last inequality

above, implying Xn−1F > Xn−1

N . •

When TIS is employed, we have Xn1 = S −Xn

2 in each period n. This simplifies

the analysis of the cost function as it is sufficient to track only Xn2 . We use the

following notation for ease of exposition: XnI denotes the stock level of new items at

the beginning of period n, i.e. Xn2 under option I for I = F ,N ,D,U .

Proposition 2 Along any sample path of demand and for the same S, if XnI < Xn

J ,

then Xn−1I > Xn−1

J for I = F ,D, J = U ,N .

Proof We only provide the proof for I = F and J = N ; the others follow from

the equivalence of the inventory recursions. XnF and Xn

N are both non-negative by

definition.

XnN = S − (Xn−1

N −Dn−12 )+ > S − [(Xn−1

F −Dn−12 )+ − (Dn−1

1 − S + Xn−1F )+]+ = Xn

F

(Xn−1N −Dn−1

2 )+ < [(Xn−1F −Dn−1

2 )+ − (Dn−11 − S + Xn−1

F )+]+

< (Xn−1F −Dn−1

2 )+.

(Xn−1F − Dn−1

2 )+ must be greater than zero in order to satisfy the last inequality

above, implying Xn−1F > Xn−1

N . •

Proposition 2 states that along any sample path, if the inventory of new items

under option F (or D) is less than the inventory of new items under option N (or U),

then in the previous period the new items in F (or D) must have been more than the

new items in N (or U). Using this fact we can divide the horizon into two classes of

periods: Pairs of periods n− 1 and n for the case when XnF < Xn

N ; and periods not in

these pairs (i.e. runs of consecutive periods in which XkF ≥ Xk

N , for k, k + 1, ...). In

the remainder of the paper, we use the term pair to denote two consecutive periods

with the above property. We will refer to a period with more new items under no-

substitution as N , with more new items under full substitution as F , and otherwise

69

Page 79: Essays on Perishable Inventory Management

(when they are equal) as E. Therefore FN denotes a pair. A sequence of periods

could look like this, with the “pairs” bracketed:

...FEF︷︸︸︷FN

︷︸︸︷FN FF

︷︸︸︷FN FFEEF...

Note that due to Proposition 2 there are never two consecutive Ns. For future

reference we call the periods between two E periods a cycle. If an E period follows

another one, then it is called a trivial cycle. A non-trivial cycle, starts with an E

period in which downward substitution takes place (making the next period an F ) and

ends with another E. An E period is not only end of a cycle but also the beginning

of the next cycle –trivial or not.

We next provide two more results which are used later in our comparative analysis.

See the Appendix for the proofs.

Proposition 3 If an F period follows another F period, then in the first one there

must be downward substitution.

Proposition 4 Suppose an F is followed by another F in periods n and n + 1.

Let ∆n be the amount of downward substitution in period n. Then, Xn+1F −Xn+1

N =

∆n − (XnF −Xn

N).

Note that the notion and results on FN -pairs easily extend to DN -pairs where

D denotes a period for which D has more new items compared to N . This is because

of the equivalence in inventory recursions of F and D.

Pairwise comparison of substitution options under TIS

We first investigate the (additional) value of upward substitution. Note both F and

U employ upward substitution whereas D and N do not. The next result is a formal

statement of our observation from Section 3.4.1.

Lemma 1 On every sample path and for any πU ∈ [0, 1], F Â D and U Â N , if

and only if condition (3.5) holds.

70

Page 80: Essays on Perishable Inventory Management

Proof Inventory levels of D and F are defined by the same recursions. Thus their

costs that are independent of upward substitution (i.e. those related to h, p1 and αD)

are equal. Therefore F has a lower cost than D if and only if upward substitution is

profitable, i.e. when αU ≤ m + p2. This remains true for all πU ∈ [0, 1]. The same

argument holds for U vs. N . •

The next question is whether downward substitution – the most common form

of substitution in practice – is beneficial. The following inequality ensures that D is

more profitable than N :

αD ≤ p1 − p2. (3.6)

Below is the formal statement of this result. Choice of πU is not relevant here since

neither D nor N practice upward substitution. See the discussion in Appendix for

the proof.

Lemma 2 If condition (3.6) holds, D Â N .

Condition (3.6) is obtained via case-by-case analysis of inventory levels and cor-

responding costs. The analysis makes use of the inventory recursions under TIS and

the notion of pairs. The condition requires the net substitution cost to be negative

since p1 < p2. Although restrictive, it is satisfied when the customer is willing to

pay the full price for the substitute, and the additional cost of substitution (including

ill-will due to substitution) and lost sales costs are negligible (i.e. α′D = r2 − r1 and

α′′D = p

′1 = p

′2 = 0).

The above condition is sufficient, but not necessary, for D to dominate N . We now

explore necessity: What happens when this condition is violated; is there a critical

value of αD, beyond which N dominates D? Our next result provides the answer.

Lemma 3 When m + p1 > 0, there is no condition on αD that guarantees N Â D.

Proof We provide a counter-example. There are no old items in stock, and the

inventory state (Xn2 , Xn

1 ) for both D and N is initially (S, 0). Consider a sam-

ple path of demand for new and old items {(Dn2 , Dn

1 ), n = 1, 2, ...}, where the se-

quence {(S − ε, S), (0, 0), (0, S), (0, 0), (0, S), (S, 0)} repeats every six periods for any

71

Page 81: Essays on Perishable Inventory Management

0 ≤ ε ≤ S. The resulting inventory levels at the beginning of each of the next six pe-

riods are {(S, 0), (0, S), (S, 0), (0, S), (S, 0), (S, 0)} for D and {(S−ε, ε), (ε, S−ε), (S−ε, ε), (ε, S − ε), (S − ε, ε), (S, 0)} for N . Note that the six periods starting with the

initial period, when both policies are at (S, 0), form an EDNDND sequence, such

that E indicates equal amount of new items in a period under both D and N , D

(N) indicates there are more new items under D (N ). In this sequence, there are

two pairs, and the inventory in the seventh period is the same as in the initial period

(an E). Let Ci, i = D,N denote the total cost of policy i in this cycle. The cost

difference is CD − CN = ε[αD − (m + p1)(J + 1) − h − p2], where J is the number

of pairs in the cycle (J = 2 in this example). By appropriately selecting a demand

stream, one can construct a sample path where J is arbitrarily large. Thus CD −CN

can always be made non-positive, independent of the magnitude of αD and πU , so

long as m + p1 > 0. •

Lemma 3 is counter-intuitive, it says that even when the downward substitution

cost is very high (αD → ∞) there can be a benefit to using substitution (or more

precisely substitution cannot be ruled out almost surely). While extreme, this exam-

ple illustrates real potential dangers of using the strict replenishment rule TIS with

policy N when demand is intermittent or periodic: these factors can result in a chain

reaction of inventory carrying, outdating and lost sales.

The next question is whether the supplier benefits from offering both upward and

downward substitution. The overall benefit of F over N , depends on both αD and

αU . Combining our observations so far, we have the next result. The proof is in the

Appendix.

Lemma 4 If (3.5) and (3.6) hold, F Â N for any πU ∈ [0, 1].

A weaker sufficient condition exists for F Â N considering the special case πU = 1,

i.e. upward substitution is always accepted. The proof is in Appendix.

Lemma 5 If (3.5) and

αD ≤ min{m + 2p1 − p2, m + p1 − αU ,h + p1

2} (3.7)

hold, then F Â N for πU = 1.

72

Page 82: Essays on Perishable Inventory Management

We provide an intuitive explanation for condition (3.7): Using downward substi-

tution, a vendor saves h + p1 and incurs αD for every item substituted in the period

preceding a pair. (Such a substitution is necessary to form a pair.) Based on this

fact, conditions

αD ≤ m + 2p1 − p2 (3.8)

and

αD ≤ m + p1 − αU (3.9)

arise as follows: If demand is low in the first period of the pair, N incurs m−h more

cost than F per item substituted upward, because N has fewer new items in the first

period of the pair. In the second period of the pair, N has more new items, which may

give N a benefit of p2− p1 per item (if F cannot substitute upward), or αU per item

(if F can substitute upward). Therefore if h + p1−αD + m− h− (p2− p1) is positive

(corresponding to the first case) and h+p1−αD+m−h−αU is positive (corresponding

to the second case), then N is more costly than F under either situation. Note that

(3.5) is redundant when (3.9) holds, as we assume p1 < p2.

Given the immediate cost difference due to downward substitution, if αD ≤ h+p1

holds then - intuitively - F should be less costly. However,

αD ≤ h + p1

2(3.10)

is in fact required. The explanation for this is as follows: Downward substitution

causes F to have fewer old items (and more new items) than N in the subsequent

period, which may lead F to substitute downward again in this next period, incurring

upward substitution cost αD twice in the same pair.

Again, the above conditions are sufficient but not necessary for F Â N . However,

given that F and D and have the same inventory recursions, our next result is not

surprising.

Lemma 6 When m + p1 > 0, for any value of πU ∈ [0, 1], there is no value of αD

or αU that guarantees N Â F .

73

Page 83: Essays on Perishable Inventory Management

Proof Follows from the example provided in the proof of Lemma 3. •Based on this last result, full substitution may be better than no substitution re-

gardless of how high the upward and downward substitution costs are. Again, this

counter intuitive behavior is partly due to the use of TIS, which constrains reordering

behavior.

Next, we explore the additional value of downward substitution when combined

with upward substitution: When is F better than the restricted form U? The proof

of the next result is in Appendix.

Lemma 7 If condition (3.6) holds then F Â U for any πU ∈ [0, 1].

For the special case where πU = 1, we have a different result:

Lemma 8 If conditions (3.5) and (3.7) hold, then F Â U for πU = 1.

Proof This follows from Lemmas 5 and 7. •

Similar to our analysis of F vs. N , the above results provide only sufficiency

conditions; there is no guarantee that U will dominate F even when they fail to hold.

Lemma 9 When m + p1 > 0, for any value of πU ∈ [0, 1], there is no condition on

αD or αU that guarantees U Â F .

Proof The proof is identical to the proof of Lemma 3 because U and N have the

same cost (as the recursions that define their inventory levels are the same) and so

do F and D. •

Finally, one-way dominance relation between D and U is as follows:

Lemma 10 The conditions αU > m + p2 and (3.6) guarantees D Â U . However

when m + p1 > 0, there is no condition on αD that guarantees U Â D for any value

of αU .

Proof Follows from results in Lemma 1 and Lemma 2. •

74

Page 84: Essays on Perishable Inventory Management

3.4.3 Economic Benefit of Substitution under NIS

In this section we continue our analysis, comparing substitution policies under the

NIS policy. We begin again with a preliminary result showing that there are fewer old

items in stock if downward substitution is practiced (Proposition 5), before moving to

pairwise comparison of the policies (Lemmas 11 - 15). Since Xn2 = S for all n under

NIS, we track Y nI , denoting the stock level of old items at the beginning of period n

(i.e. Xn1 ) under policy I, for I = F ,N ,D,U .

Proposition 5 For a given S, Y nF = Y n

D ≤ Y nU = Y n

N for all n along any sample

path. (This result also holds for general πD ∈ [0, 1].)

Proof Due to the recursions that define inventory levels, we know Y nF = Y n

D and

Y nU = Y n

N . If there is no downward substitution in period n, then Y nF will be equal

to Y nN . In case of any downward substitution, there will be fewer old items carried to

the next period under F (or D), i.e. Y nF < Y n

N . Therefore Y nF = Y n

D ≤ Y nU = Y n

N . •

The result is intuitive: under NIS, no substitution or only upward substitution

result in higher levels of old item inventory, since inventory of new is never used to

satisfy excess demand of old. Next, we look at the effect of upward substitution. Just

as in the case for TIS, upward substitution is profitable if and only if the upward

substitution cost, αU , is less than m + p2, which is the cost if upward substitution is

possible but does not take place in a given period.

Lemma 11 On every sample path and for any πU ∈ [0, 1], F Â D and U Â N , if

and only if condition (3.5) holds.

Proof We provide the proof for F vs. D only; the same idea applies to U vs.

N . Due to inventory recursions, we know Y nF = Y n

D for all n. F and D have

the same cost for every period except for the periods where F has (accepted) up-

ward substitution. In any period n, if Dn1 < Y n

F and Dn2 > S, then upward sub-

stitution is offered in the amount of us = min{L, K}, and is accepted in some

amount usn ≤ us, where L = Dn2 − S and K = Y n

F − Dn1 . Let Cn

I denote the

cost of policy I in period n, for I = F ,D,U ,N . Then, we have CnD = mK + p2L ,

75

Page 85: Essays on Perishable Inventory Management

CnF = m(K−usn)+p2(L−usn)+αU usn and Cn

D−CnF = (m+p2−αU)usn. Therefore

D is costlier than F if and only if αU < m + p2. •

Next, we look at the (additional) value of downward substitution. Unlike TIS, the

decision to use an excess new item in one period does not affect the number of new

items in the next period under NIS. Therefore, the effect of substitution is easier to

determine.

Consider two consecutive periods; in the first downward substitution occurs. For

each new item substituted for old, there is a cost αD. In the same period, the lost

sales cost p1 is saved, and one fewer item is carried to the next period, saving h. The

greatest benefit that can be accrued from the downward substitution occurs if the

demand for old items in the next period is low – one fewer item outdates, saving m.

Thus, when

αD ≥ m + h + p1 (3.11)

holds, downward substitution is never beneficial.

Now suppose, αU > m + p2 holds (that is upward substitution is unprofitable,

by Lemma 11). The worst thing that could happen after substitution is that an

additional old item might be needed later, incurring a cost of p1. If in this case

downward substitution is profitable, then it is always beneficial. This holds under the

following condition:

αD ≤ h. (3.12)

Finally, if upward substitution is viable (i.e. αU ≤ m + p2), in addition to the

previous case (3.12), it may happen that in the next period an upward substitu-

tion is desired but cannot be enacted because of our previous period’s downward

substitution. This will save a substitution cost, αU , but cost p2. If even in this

case substitution is profitable, then downward substitution is always beneficial, i.e.

provided:

αD ≤ min{h, h + αU − (p2 − p1)}. (3.13)

Formal statement of these results follow. Proofs are available in Appendix.

Lemma 12 (F vs. N ) If conditions (3.5) and (3.12) hold, then F Â N . If condi-

tion (3.5) fails to hold but condition (3.11) holds, then N Â F .

76

Page 86: Essays on Perishable Inventory Management

Lemma 13 (F vs. U) If condition (3.13) holds, then F Â U . If conditions (3.5)

and (3.11) hold, then U Â F .

Lemma 14 (D vs. N ) If condition (3.12) holds, then D Â N . If condition (3.11)

holds, then N Â D.

Lemma 15 (D vs. U) If conditions (3.5) and (3.11) hold, then U Â D. If condition

(3.5) fails and condition (3.12) holds, then D Â U . If condition (3.5) holds, there is

no condition that guarantees D Â U .

3.4.4 A note on substitution not being a recourse

We assume in our analysis that substitution is only used as a “recourse” action – new

items are used to satisfy new item demand and old items to satisfy old item demand

before any substitution takes place. As we assume p2 > p1, this will always be the

cost-optimal action with respect to new item inventory; i.e. new items should be used

to satisfy the demand for new items first. Likewise, if αU > p2 − p1, it is optimal to

use old items to satisfy demand of old items before offering them as substitutes. If

αU < p2−p1, then it is better to satisfy new item demand first using all the inventory.

Such an option does not change the sample path of the inventory in the system, so all

of our other results remain unchanged except that all of our comparison results with

respect to upward substitution (i.e. F vs. D, U vs. N ) would include this condition.

In fact, combining conditions αU ≤ m + p2 in equation (3.5) and αU ≤ p2 − p1, we

get the following: Using upward substitution to satisfy new item demand before old

item demand is beneficial if and only if αU ≤ p2 − p1.

3.4.5 Freshness of Inventory

Our focus so far has been on the economic benefit of substitution. However, the overall

benefit of a substitution - especially for the customers - cannot be assessed unless

service issues are also taken into account. In this section we show that downward

substitution leads to fresher inventory, which may help the supplier in a number of

less tangible ways: These items have not lost their value through aging, customers

77

Page 87: Essays on Perishable Inventory Management

are more likely to prefer them, and having newer items in stock decreases the risk

of obsolescence. Note that we showed in Sections 3.4.2 and 3.4.3 that this does not

necessarily lead to cost-wise superiority due to substitution costs.

We first define what we mean by freshness: We look at the expected age of goods

in stock – the greater the average remaining lifetime, the fresher the goods. No

assumptions on problem parameters are needed for our results on freshness. For NIS,

pairwise comparison of substitution policies with respect to freshness immediately

follows from our earlier results.

Lemma 16 For the same S, and any πD, πU ∈ [0, 1], average age of inventory

is lower (i.e. goods are fresher) under F and D than U and N when inventory is

replenished using NIS.

Proof Follows from Proposition 5 (there are less old items due to downward sub-

stitution). •

When inventory is replenished using TIS, we compare the average age of goods

in stock for different substitution policies by studying the amount of new items in

stock. This leads to our next result. Detailed discussion and the proof of the result

are provided in Appendix.

For an arbitrary period n, we know we can have XnF ≥ Xn

N or XnF < Xn

N based

on our previous results; we also know XnF = Xn

D and XnN = Xn

U (almost surely) for

all n. Below, we show in a time average sense that the inventory level of new items

is greater, i.e. items are fresher under F . The definition of a pair is the same as in

Section 3.4.2. Outside the pairs it is obvious that there are more newer items under

F . The proposition below analyzes the situation inside a pair.

Proposition 6 Suppose inventory is replenished using TIS. Let periods n− 1 and

n constitute a pair, and define dsn−1 to be the amount downward substituted in period

n− 1. Then, in period n− 1, (i) Dn−11 + dsn−1 < Xn−1

F , (ii) if there is substitution,

it must be downward substitution, and (iii) dsn−1 < ∆.

Proof See Appendix. •

78

Page 88: Essays on Perishable Inventory Management

Based on Proposition 6, we can show that the difference Xn−1F −Xn−1

N > 0 is no less

than XnN −Xn

F > 0 for a pair in periods n− 1 and n.

Lemma 17 For the same S, and any πD, πU ∈ [0, 1], goods are fresher under Fand D than N and U , when inventory is replenished using TIS.

Downward substitution is clearly superior in this respect: Under both replenish-

ment policies, it leads to fresher inventory. This confirms the value of this practice in

many industries; downward substitution will, at least indirectly, lead to benefits for

the suppliers and customers.

3.5 Computational Results

In this section, we present a series of experiments used to explore the performance of

our substitution policies for a number of parameter settings. We discuss properties

of the cost functions of different substitution policies, under both TIS and NIS in

Section 3.5.1. In Section 3.5.2 we discuss the effect of downward substitution cost

on the optimal substitution policy. We then study costs, service levels and freshness

in Section 3.5.3, and policy comparisons under proportional acceptance of substitute

goods in Section 3.5.4.

3.5.1 Non-convexity/Multi-modality of the Cost Function

In Examples 1A&1B we show the non-convexity and multi-modality of NIS and TIS.

In these examples common costs are h=1, p1=3, p2=9, αU=7, αD=6 and the demand

for new and old items are both discrete uniform distributed between 0 and 25. We

vary m; m = 5 in 1A and m = 2 in 1B. Note that condition αU < m + p2 is satisfied

in both examples; therefore F has lower cost than D, and U has lower cost than N ,

under both TIS and NIS. However αD does not satisfy any of the sufficient conditions

in Table 3.1 – there is no other TIS dominance relation among the substitution policies

holding for all S, as we see in Figures 3.1 and 3.2 for 1A and 1B, respectively.

We also see that the cost function under TIS for policies F and D are non-convex

in 1A; in 1B these functions become non-convex in the NIS case, and multi-modal

79

Page 89: Essays on Perishable Inventory Management

TIS

40

50

60

70

80

90

100

1 6 11 16 21 26 31 36 41

Order-up-to level (S)

Co

st

F D U N

NIS

40

50

60

70

80

90

100

1 6 11 16 21 26 31 36 41

Order-up-to level (S)

Co

st

F D U N

Figure 3.1: Time-average costs of substitution policies in Example-1A.

in the TIS case. In addition, while N and U have higher optimal order-up-to levels

compared to F and D under TIS, the opposite is true under NIS. Reducing the

outdating cost m lowers the overall costs of all policies, but leaves the optimal S

values relatively unchanged.

Overall, in both examples policies that do not downward substitute, i.e. U and Nperform best, owing to the high substitution costs and p2 >> p1. The optimal costs

of these policies are lower under NIS than TIS, although for any given S the cost of a

substitution policy under NIS is not necessarily lower than that of TIS. In addition,

these policy’s costs are convex with respect to S, raising the possibility that optimal

substitution policies possess convexity. Thus far we have been unable to prove such

a result.

TIS

20

30

40

50

60

70

1 6 11 16 21 2 31 3 41

Order-up-to level (S)

Co

st

F D U N

NIS

20

30

40

50

60

70

1 6 11 16 21 26 31 36 41

Order-up-to level (S)

Co

st

F D U N

Figure 3.2: Time-average costs of substitution policies in Example-1B.

80

Page 90: Essays on Perishable Inventory Management

TIS

24

28

32

36

40

1 6 11 16 21 26 31

Order-up-to level (S)

Co

st

F D U N

NIS

24

28

32

36

40

1 6 11 16 21 26 31

Order-up-to level (S)

Co

st

F D U N

Figure 3.3: Time-average costs of substitution policies in Example-2.

3.5.2 Making downward substitution more attractive

In Example-2 we make the downward substitution more attractive; that is we set the

old-item demand greater in magnitude than new item demand, and reduce substi-

tution costs: Specifically, we choose h=1, m=2, p1=3, p2=9, αU=3, αD=2. Again

αU < m + p2 is satisfied, while αD violates the sufficiency conditions in Table 3.1.

The demand for new items is discrete uniform between 0 and 6, demand for old items

is discrete uniform between 0 and 25. These changes make downward substitution

more attractive – it is both less costly, and can occur more frequently.

The expected cost functions for Example-2 are presented in Figure 3.3. (Note

that under NIS the costs of F and D and those of U and N coincide.) As we would

expect, the lowest cost is achieved by policies that downward substitute, F and D.

In addition, given the increase in downward substitution, TIS now provides both

lower cost than NIS, and greater robustness to choice of S. Our observations are

consistent with the literature; TIS is effective when there is only one demand stream

and demand is fulfilled in a FIFO fashion (i.e. downward substitution is practiced).

Example-3 (Effect of αD). We fix parameters h=1, m=5, p1=3, p2=9, αU=7,

and vary αD between zero and seven, plotting the cost at optimal values of S for each

value of αD . The demand for both new and old items is discrete uniform distributed

between 0 and 25. Note that upward substitution is beneficial since αU < m + p2.

Figure 3.4 shows the minimum expected cost of each policy as a function of αD; as

αD increases, the cost of policies F and D increase almost linearly under both TIS

and NIS.

81

Page 91: Essays on Perishable Inventory Management

NIS

20

30

40

50

60

70

0 1 2 3 4 5 6 7Downward substitution cost ( )

Min

imu

m C

os

t

F D U N

TIS

20

30

40

50

60

70

0 1 2 3 4 5 6 7

Downward substitution cost ( )

Min

imu

m C

os

t

F D U N

Figure 3.4: Minimum expected cost of policies as a function of downward substitution

cost (αD) in Example-3.

From Table 3.1, under both NIS and TIS, αD < 1 guarantees F Â N and F Â U .

Under NIS, αD > 9 guarantees N Â F and U Â F . Example-3 illustrates the

behavior of substitution policies in the parameter range where our sufficient conditions

do not hold. In this example, we observe a critical value of αD = 4 under NIS; for

higher values of αD, U andN have lower expected costs. This critical value is between

five and six for TIS. In either case, these values are in the mid-to-high portion of the

parameter ranges identified in our sufficient conditions. Thus, while our analytical

conditions of policies do not cover the entire αD range for dominance relations, they

can provide guidance, especially when αD is closer to one sufficient condition than

the other.

In all of the previous examples, when NIS is used to replenish inventory there

is almost no additional benefit to using upward substitution (over no substitution),

but when TIS is used this difference is pronounced. This is due to the fact that very

little upward substitution is taking place under NIS as compared to TIS, a fact borne

out in the next section. This may be due to the fact that the optimal base-stock

policy of NIS accounts directly for new item demands, reducing the need for upward

substitution. In contrast, in TIS upward substitution is needed more frequently;

because new items are ordered less, depending on the amount of old items.

82

Page 92: Essays on Perishable Inventory Management

D ∼ {0, 25} D ∼ U[0, 25] D ∼ P(12.5) D ∼ U[10, 15]

Policy NIS TIS NIS TIS NIS TIS NIS TIS

F 38.5 47.4 26.1 28.7 14.5 13.5 9.0 9.2

D 38.5 48.7 26.2 29.1 14.5 13.5 9.0 9.2

U 46.8 64.4 32.0 42.8 18.7 25.5 15.2 20.4

N 46.8 66.6 32.1 44.2 18.7 25.8 15.2 20.5

Table 3.2: Average cost comparison between TIS and NIS under four different demand

models.

3.5.3 Costs, Service Levels and Freshness of Inventory

In this section we present the results of our numerical study on the costs, service levels

and freshness of the policies. For each cost parameter and demand distribution, we

first compute the optimal order-up-to level S for each substitution and reordering

policy. Then, using this order-up-to level, we compute average performance measures

via simulation.

For the costs comparison the experiment is designed as follows: We set h = 1,

m ∈ {2, 5}, p1 ∈ {1, 3}, p2 ∈ {4, 9}, αU ∈ {3, 7, 11, 15} and αD ∈ {−1, 0, 2, 6} which

results in 128 instances based on cost parameters. Note, only some of these instances

satisfy the conditions identified in Section 3.4.

In our experiments, we evaluate TIS and NIS using their respective optimal order-

up-to levels. While the long-term average cost function C(S) in (3.4) is non-convex

in S for either of these replenishment policies (see the examples in Section 3.5.1), the

optimal order-up-to levels are easily determined using line search.

We first discuss the effectiveness of TIS and NIS. We present a high-level cost

comparison of TIS and NIS here. In Table 4.1, we report the average percentage

difference between TIS and NIS for each substitution scenario and demand distrib-

ution. The four demand distributions were: (i) Equal point mass at 0 and 25; (ii)

Discrete uniform demand between zero and 25; Poisson demand with rate 12.5; and

(iv) Discrete Uniform demand between 10 and 15. In all cases the demand for old

and new items are independent.

83

Page 93: Essays on Perishable Inventory Management

D ∼ {0, 25} D ∼ U[0, 25] D ∼ P(12.5) D ∼ U[10, 15]

Policy NIS TIS NIS TIS NIS TIS NIS TIS

F 17.8% 28.9% 18.6% 35.0% 22.5% 47.7% 40.8% 55.0%

D 17.8% 26.9% 18.4% 34.0% 22.5% 47.7% 40.8% 55.0%

U 0.2% 3.3% 0.4% 3.1% 0.0% 1.4% 0.0% 0.5%

Table 3.3: Average percentage improvement in costs over N .

From Table 4.1 we can see that in fourteen of the sixteen combinations of policies

and demand distributions, NIS, the policy that ignores the information on the old

item inventory, outperforms TIS, the policy that uses this information, often by a

considerable amount (the average difference can be as high as 30%). Moreover, for

the two instances in which TIS was superior, (i.e. for Poisson demand scenario), the

difference was only slight (less than 8 %).

Next we study the policies in terms of service levels, conducting 32 experiments

with the following cost parameters: h = 1, m ∈ {2, 5}, p1 ∈ {1, 3}, p2 ∈ {4, 9},αU ∈ {3, 7} and αD ∈ {2, 6}. Demand for new and old items is independent, and

both are discrete uniform distributed between 0 and 25. For each cost parameter

and demand distribution, we first compute the optimal order-up-to level S for each

substitution and reordering policy. Then, using this order-up-to level, we compute

average performance measures via simulation.

We study two aspects of service level: (i) the percentage of demand lost and

(ii) the percentage of demand satisfied via substitution. Table 3.4 summarizes our

results; the figures represent the averages across all 32 instances. Employing NIS as

the replenishment policy increases the service level and decreases substitution: NIS

not only tends to keep more inventory, but also better keeps the appropriate inventory

of new items, reducing substitution. This effect is significant because the demand of

old and new are both relatively high as compared to the optimal stocking levels in

this experiment.

In the same experiment, we also looked at freshness of inventory. We know that

policies that use downward substitution have fresher inventory for a given S; however,

when policies use different order-up-to levels (which is the case in our experiment),

84

Page 94: Essays on Perishable Inventory Management

Type of % lost % substituted Avg. Freshness Average Cost

Policy demand TIS NIS TIS NIS TIS NIS TIS NIS

F new 7.0% 4.6% 2.3% 0.3% 1.85 1.82 37.0 39.9

old 41.0% 34.6% 42.6% 35.8%

D new 10.2% 5.0% 0.0% 0.0% 1.85 1.82 37.2 41.4

old 44.9% 34.4% 40.4% 35.8%

U new 13.9% 3.8% 3.1% 0.4% 1.70 1.71 31.5 38.9

old 55.4% 50.9% 0.0% 0.0%

N new 17.7% 4.1% 0.0% 0.0% 1.70 1.71 32.1 44.2

old 57.2% 50.5% 0.0% 0.0%

Table 3.4: Service levels, inventory age and costs averaged across all experiments.

this result may no longer hold. We observed that F and D have fresher goods than

U and N in 31 out of 32 of the experiments, when inventory is replenished by TIS.

When NIS is used, this statistic falls to 29 out of 32. However, downward substitution

is only slightly worse in these four instances; the average age of inventory of F and Dis within 1% of U and N . Average freshness numbers are also provided in Table 3.4.

3.5.4 Effect of customer behavior: Proportional acceptance

of substitutes

In this section we present numerical examples where we vary πD and πU (0 ≤ πD, πU ≤1) in order to illustrate the effect of acceptance proportions.

In all the examples, the demand is discrete uniform distributed between 0 and

25 for both new and old items. The examples below are analyzed for the case when

TIS is used in replenishment. Our observations are similar for NIS, and hence are

omitted.

Example-4 (Proportional Downward acceptance). In this example we use

h = 1, m = 2, p1 = 3, p2 = 4, αD = 6, αU = 3. We study the effect of proportional

85

Page 95: Essays on Perishable Inventory Management

acceptance of downward substitution by varying πD, assuming πU = 1 in this case.

Thus, πD = 0 represents policy U and πD = 1 represents policy F . The total cost as a

function of πD is presented in Figure 3.5. This graph shows how different components

of the total cost change as we move from policy U to F as πD increases. In all

experiments we use the optimal value of S for the given parameter values (including

πD).

In this particular example, the expected total cost increases almost linearly with

πD (i.e. U is less costly than F). As πD increases, the amount of downward substitu-

tion increases (hence so too does its cost component) while the lost sales cost of old

items decrease. Downward substitution under TIS increases the inventory turnover

for new items; as the amount of new items used to satisfy the demand for old in-

creases, TIS leads to higher number of newer items being replenished. Therefore, as

πD increases, the penalty for lost sales of new items also decreases. In this exam-

ple, only a slight increase and decrease are observed in holding and outdating costs,

respectively.

Example-5 (Proportional Upward acceptance). The costs are the same

as in Example-1A. Here we look at the effect of acceptance proportion for upward

substitution (πU) and observe the changes as we move from policy D to F . Again, in

all experiments we use the optimal value of S for the given parameter values, including

πU . As we see in Figure 3.6, the effect of πU on the costs is minimal; the total cost

decreases slightly as πU increases (i.e. F is less costly than D). This is not surprising

because the difference between the lowest cost of D and F in Example-1A is very

small. (This holds true for all the examples presented in Figures 3.1-3.3). There

are examples where this difference may be slightly more (as in Example -4), but our

main observation does not change: There is not a systematic effect of πU on the cost

components except the upward substitution cost. Thus it appears that the expected

costs are typically more sensitive to πD. This is similar to our analytical results

on dominance conditions. Finding dominance conditions for upward substitution

is easier as upward substitution does not affect future inventory levels. Therefore

myopic substitution policies are optimal and we have if-and-only-if type dominance

conditions for upward substitution. However downward substitution changes future

inventory levels and dynamic relationships more complicated. Therefore we cannot

have tight dominance conditions for infinite horizon. Thus the system costs fluctuate

86

Page 96: Essays on Perishable Inventory Management

0

10

20

30

40

50

60

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1D

Ex

pec

ted

Co

st UpwSubs

DwnSubst

Outdating

Penalty(Old)

Penalty(New)

Holding

Figure 3.5: The effect of proportional acceptance of downward substitution on ex-

pected total cost in Example-4.

more as πD varies.

3.5.5 Summary of Computational Results

The numerical study presented above implies that downward substitution is an impor-

tant lever for improving service levels and freshness of inventory, and that downward

substitution has a more significant effect on the system performance than upward

substitution (the value of πD is more important than πU). In contrast to what has

appeared in the bulk of the literature, we see examples where NIS appears to be a

more suitable replenishment policy with respect to costs, service levels and inventory

freshness, especially when there is significant demand for new items and substitution

costs are appreciable. We are also able to segment substitution policies by cost: The

relative differences in expected costs of substitution policies are highest when we com-

pare policies that use upward substitution to ones utilizing downward substitution.

Hence, a supplier can significantly benefit from using N or U over F or D, and vice

versa.

87

Page 97: Essays on Perishable Inventory Management

0

10

20

30

40

50

60

70

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1

U

Exp

ecte

d C

ost

UpwSubs

DwnSubst

Outdating

Penalty(Old)

Penalty(New)

Holding

Figure 3.6: The effect of proportional acceptance of upward substitution on expected

total cost in Example-5.

3.6 Conclusion

In this study, we focused on inventory management of perishable goods and goods

that lose their value over time. In our model, demand is differentiated with respect to

the age of the product, and some customers accept substitutes in case of stock-outs.

Our goal was to find out under what conditions substitution provides lower costs

for the supplier and fresher goods for the customers. We considered two different

base-stock type inventory replenishment policies, TIS and NIS, for the supplier and

formalized four different substitution options. We identified conditions on cost para-

meters that guarantee one form of substitution being economically superior (almost

surely) compared to others. We also show that downward substitution policies are

always beneficial with respect to average age of goods in stock. Our analysis revealed

that TIS, which is advocated as a good replenishment policy for perishables, can lead

to counter-intuitive performance when the demand is differentiated by the age of the

product and substitution takes place. On the other hand, NIS is very promising as

a replenishment policy. NIS is more robust against new item’s demand fluctuations

than TIS. It is simple, leads to lower costs under many demand scenarios and has

more predictable behavior under substitution.

88

Page 98: Essays on Perishable Inventory Management

Chapter 4

Analysis of Inventory

Replenishment Policies for

Perishable Goods Under

Substitution

4.1 Introduction

In the previous chapter we analyzed the effect of substitution on operational costs. In

this chapter we focus on the replenishment policies. Using the same model as in the

previous chapter we allow demand for both new and old products, and we consider

the possibility of substitution between old and new products. We refer the reader

to Chapter 3 for the details of the model. In Section 4.2 first we present analytical

results on the optimality of our heuristic policies for special cases. We also provide

our findings on structural properties of the policies. In Section 4.3 we provide the

results of a comprehensive numerical study. We study the policies under various

demand schemes; and we also study them when substitution is not always accepted.

In Section 4.3.9 we develop and compare a number of NIS order-up-to level heuristics

using newsvendor-type logic. We only develop heuristics for NIS, as we have observed

89

Page 99: Essays on Perishable Inventory Management

in the previous chapter that NIS appears to be outperforming TIS.

We also conduct a numerical study using simulation and dynamic programming.

We compare TIS and NIS at their optimal order-up-to levels (found by doing a line

search using simulation) to the globally optimal state-dependent policy found via

dynamic programming.

4.2 Analytical Results

In this section we present analytical results establishing the optimality of NIS for

special cases based on the substitution type that is allowed or the demand type (for

old or new items) that exists. Demand is assumed to be i.i.d. for these results.

Lemma 18 NIS is the optimal replenishment policy, under N when demand is i.i.d.

Proof : The idea behind the proof is as follows: If there is no substitution allowed,

each ordering decision becomes independent of each other. The ordered new items

in any period are sold, or they perish in two periods. Therefore this case boils down

to a “2-period Newsvendor” problem. The NIS amount, which is the solution to the

“2-period Newsvendor”, is the optimal solution. The formal proof is as follows:

In a period n, let the ordering decision be qn. In period n the following cost is

incurred due to this decision:

p2[Dn2 − qn]+ + h[qn −Dn

2 ]+. (4.1)

In the the following period, n + 1, the cost of the decision is :

p1[Dn+11 − (qn −Dn

2 )+]+ + m[(qn −Dn2 )+ −Dn+1

1 ]+. (4.2)

Therefore the ordering decision qn+1 does not affect the cost of decision qn in period

n + 1; nor is the decision qn+1 affected by the decision qn as (qn − Dn2 )+ does not

affect Dn+11 , due to the i.i.d. assumption.

Hence each ordering decision for every period is separable from the others; and

the problem reduces to the “2-period” newsvendor problem, finding the quantity qn

that minimizes (4.1) + (4.2). Since demand over the periods is i.i.d. NIS is optimal.

90

Page 100: Essays on Perishable Inventory Management

Lemma 19 If old item demand is zero (with no upward substitution) then NIS is

optimal, when demand is i.i.d.

Proof : The proof is immediate from the proof of Lemma 4.2. The problem reduces

to the 1-period classical newsvendor problem in this case.

Lemma 20 If new item demand is zero (with no downward substitution) then NIS

is optimal, when demand is i.i.d.

Proof : The proof is immediate from the proof of Lemma 4.2. The problem reduces

to the newsvendor problem with “lead-time” of one period. The lead-time does not

affect the solution.

If new item demand is zero (with downward substitution) then the problem be-

comes the single demand stream FIFO problem. In the next section we numerically

show that TIS is a good heuristic but not optimal, NIS is comparable.

We also provide the following analytical results on structural properties of optimal

ordering functions:

Lemma 21 The optimal ordering function is independent of on-hand inventory when

the optimal substitution policy (among the four policies) is N and demand is i.i.d.

Proof : The proof is immediate from Lemma 4.2; as q∗(x) = S for any x due to

Lemma 4.2.

4.3 Computational Results: NIS vs. TIS

First we show that closed-form solutions for the two-period lifetime case (TIS) for

a given demand distribution is quite complicated, in section 4.3.1. In the remaining

sub-sections we do a computational study to shed more light on the performances of

the NIS and TIS policies. In Section 4.3.2 we present comparisons of TIS and NIS for

various demand distributions. We present comparison of TIS and NIS when old and

new item demands are positively correlated and negatively correlated in sections 4.3.3

and 4.3.4 respectively. The results of our computations with high new item demand

91

Page 101: Essays on Perishable Inventory Management

is in Section 4.3.5, and with low new item demand is in Section 4.3.6. In sections

4.3.7 and 4.3.8 we present our results when substitution is not always accepted. And

finally in section 4.3.9 we develop and analyze a number of NIS order-up-to level

heuristics using newsvendor-type logic.

4.3.1 A Preliminary Study

We did a preliminary computational study to try to determine whether the steady

state distribution of the TIS levels has a closed form. Brodheim et al. (1975) study

NIS in a similar way using a Markov chain model, when demand is present only

for the oldest items. They realized that Markov chain techniques could be used to

obtain exact expressions, or very simple bounds, on key system measures such as

probability of shortage, the average age of the inventory, and the average quantity

outdated products per period.

In this section we assume total demand has a Poisson distribution with mean λ.

Demand goes to new items with probability p and to old items with probability 1− p

.

For no-substitution policy, the transition probability, Pkl, for Xn2 is :

Pkl =

0 If k + l < S for 0 ≤ l ≤ S − 1,e−λp(λp)k+l−S

(k+l−S)!If k ≥ S − l for l 6= S,

1−∑k−1m=0

e−λp(λp)m

m!If l = S.

Stationary distribution for S = 2 case is as follows:

π0 =e−λp − e−2λp

1 + e−λpλp− e−2λp,

π1 =e−λpλp

1 + e−λpλp− e−2λp,

π2 =1− e−λp

1 + e−λpλp− e−2λp.

For full-substitution, the transition probability, Pkl, for Xn2 is as follows:

92

Page 102: Essays on Perishable Inventory Management

Pkl =

0 If k + l < S,

e−λ λ−l

l!If k = S, 0 ≤ l ≤ S − 1,

1−∑S−1m=0

e−λλm

m!If k = l = S ,

e−λp ∑lm=0

e−λ(1−p)(λ(1−p))m

m!Ifk + l = S,k 6= 0,l 6= S,

1 If k = 0 and l = S,

1−∑k−1m=0

e−λp(λp)m

m!

+∑k−1

m=0(e−λp(λp)m

m!− (1−∑S−m−1

n=0e−λ(1−p)(λ(1−p))n

n!)) If l = S,1 ≤ k ≤ S − 1,

e−λp (λp)k−S+l

k−S+l!

∑S−km=0

e−λ(1−p)(λ(1−p))m

m!

+∑k−S+l−1

m=0e−λλlpm

m!(l−m)!(1−p)m Otherwise.

Stationary distribution for S = 2 case is:

π0 =e−λ − e−2λ(1 + λ− λp)

1− e−2λ(1 + λ− λp) + e−λλp,

π1 =e−λλ

1− e−2λ(1 + λ− λp) + e−λλp,

π2 =1− e−λ(1 + λ− λp)

1− e−2λ(1 + λ− λp) + e−λλp.

The stationary distribution of the inventory levels does not seem to have a simple

closed form; therefore in this study we focus on comparison of our heuristics.

4.3.2 Comparison of TIS and NIS while changing variance of

demand

In order to observe performances of TIS and NIS under different demand variance

we did a numerical study with four different demand distributions with different

coefficients of variation (CV). In Table 4.1 we present this cost comparison of TIS

and NIS based on simulation (400,000 periods after initial warm-up). We conducted

four sets of 32 experiments with the following cost parameters: h = 1, m ∈ {2, 5},p1 ∈ {1, 3}, p2 ∈ {4, 9}, αU ∈ {3, 7} and αD ∈ {2, 6}. Each of the four sets used a

different demand distribution, where we varied the characteristics of the demand as

well as the coefficient of variation. The demand distributions were: (i) Equal point

mass at 0 and 25 (CV = 1.41); (ii) Discrete uniform demand between zero and 25

(CV = 0.58); Poisson demand with rate 12.5 (CV = 0.28); and (iv) Discrete Uniform

93

Page 103: Essays on Perishable Inventory Management

demand between 10 and 15 (CV = 0.12). In all cases the demand for old and new

items are independent.

Table 4.1 shows the average total costs for the sixteen combinations of policies

and demand distributions.

D ∼ {0, 25} D ∼ U[0, 25] D ∼ Poisson(12.5) D ∼ U[10, 15]

Policy Opt. NIS TIS Opt. NIS TIS Opt. NIS TIS Opt. NIS TIS

F 47.3 47.9 53.7 36.6 37.0 39.9 24.7 25.8 25.6 18.6 19.4 21.1

D 47.9 47.9 60.0 36.9 37.2 41.4 24.7 25.8 25.6 18.6 19.4 21.2

U 46.1 46.7 56.1 31.3 31.5 38.9 18.7 18.7 24.4 15.2 15.2 20.0

N 46.8 46.8 66.6 32.1 32.1 44.2 18.7 18.7 25.8 15.2 15.2 20.5

Table 4.1: TIS vs. NIS in terms of cost (avg. over 32 instances) under four different

demand models, all with mean 12.5.

We find the optimal cost via dynamic programming 1. NIS, the policy that ignores

the information on the old item inventory, outperforms TIS, the policy that uses this

information, often by a considerable amount (up to 42.2 %). Moreover, for the two

instances in which TIS was superior, for Poisson demand and full or downward substi-

tution, the difference was only slight (less than 1 %). These results give preliminary

support to the industry practice of utilizing NIS (Angle, 2003) in practice, but con-

trast sharply with virtually all of the literature on control of perishable inventories.

Interestingly, similar results have recently been seen in some studies of continuous

time lost-sales systems without substitution (Reiman, 2005).

The table above compares average costs of each policy over all instances, specifi-

cally it considers the costs of a policy whether it is the optimal policy or not. Now

we look a little more closely at our data, focusing only on the optimal policy for each

instance. According to our computational studies when demand is highly variant

(CV is 0.58 or 1.41) NIS outperforms TIS no matter what the optimal substitution

1In Figure 4.1 we see computation times for NIS, TIS and dynamic programming as maximum

demand grows. Computations are done with a 1.2 GHz Intel Pentium.

94

Page 104: Essays on Perishable Inventory Management

0

10

20

30

40

50

60

70

80

90

100

5 25 35 50

Max Demand

Min

ute

s TIS

NIS

Opt

Figure 4.1: Computation times for the best NIS, the best TIS and the optimal solu-

tions.

policy is (in all four quadrants of Figures 4.2 and 4.3)2. For example when demand is

Uniform [0, 25] (i.e. Figure 4.2) only for 1 instance (out of 32) TIS (Full-substitution)

is better than NIS (Full-substitution). Why is NIS so close to optimal? New items

are more valuable and NIS is more robust against new item’s demand fluctuations.

As being able to fulfill new item demand has primary importance regardless of the

optimal substitution policy, and the high variability of demand makes it best to ig-

nore old item information. Not ignoring old item information (TIS) would lead us to

under-order when demand fluctuates a lot from one period to another. For example,

if demand is low in a period, the old item inventory level will be high in the next

period. And if TIS is used the order quantity will be small leaving the new item

inventory level low. This would be costly if new item demand is high in the next

period.

If the demand has a low variance (CV is 0.12 or 0.28) NIS is better if downward

substitution is not viable (see the upper and lower right quadrants in Figures 4.4 and

4.5). The reason to this as follows: When downward substitution is not viable the

problem boils down to the newsvendor problem (as upward substitution is not done

frequently), which is optimally solved by NIS. However if downward substitution is

viable and if demand uncertainty is not high TIS outperforms NIS (see the upper and

2The percentages indicate percent cost differences between the policy and the optimal.

95

Page 105: Essays on Perishable Inventory Management

D

U

2%10%

NISTIS

0%29%

NISTIS

2%9%

NISTIS

1%24%

NISTIS

Full Subs.

Downward Subs. No Subs.

Upward Subs.

•Best substitution and replenishment policies based on the

m+p2

1 instance

12 instances 16 instances

3 instances

Uniform [0, 25]

Figure 4.2: Optimal policies when CV is 0.58

D

U

0%20%

NISTIS

0%33%

NISTIS

3%12%

NISTIS

1%20%

NISTIS

Full Subs.

Downward Subs. No Subs.

Upward Subs.

•Best substitution and replenishment policies based on the

m+p2

14 instances

2 instances2 instances

14 instances

{0, 25}

Figure 4.3: Optimal policies when CV is 1.41

D

U

7%4%

NISTIS

0%25%

NISTIS

14%2%

NISTIS

0%26%

NISTIS

Full Subs.

Downward Subs. No Subs.

Upward Subs.

•Best substitution and replenishment policies based on the

m+p2

1 instance

7 instances 21 instances

3 instances

Uniform [10, 15]

Figure 4.4: Optimal policies when CV is 0.12

D

U

8%2%

NISTIS

0%29%

NISTIS

9%1%

NISTIS

0%30%

NISTIS

Full Subs.

Downward Subs. No Subs.

Upward Subs.

•Best substitution and replenishment policies based on the

m+p2

21 instances

3 instances1 instance

7 instances

Poisson (12.5)

Figure 4.5: Optimal policies when CV is 0.28

lower left quadrants of these same figures). Under a low-variance demand the effect

of keeping extra new inventory in order to fulfill new item demand is marginal. Being

efficient is more important and using old item inventory information for replenishment

decisions reduces costs, this makes TIS better than NIS in such an environment.

This raises the question of how a manager could know which “quadrant” their op-

erations lay in. For the upper versus the lower half it is simple, as we have necessary

and sufficient conditions for upward substitution to be beneficial, namely αU < m+p2.

Whether or not downward substitution is advisable is more complex, and arguably

more important, because this also has an impact on which ordering policy is best. For-

tunately, our computational studies shed light on this question: Another observation

96

Page 106: Essays on Perishable Inventory Management

h h+m+p1

D

Figure 4.6: The metric shows where the downward substitution cost (αD) lays com-

pared to upper and lower bounds

due to our computational studies is that for high-variance demand downward substi-

tution is viable for relatively large downward substitution costs. This is observed by

using the following metric M (see Fig. 4.6) under NIS:

M =αD − h

m + p1

(4.3)

Our experiments show that under highly variant demand downward substitution is

viable as long as M is not larger than 0.5. For low-variance demand downward

substitution is only allowed for low downward substitution costs (the metric cannot

be greater than 0.2).

4.3.3 NIS vs. TIS when new and old item demands are neg-

atively correlated

In this section we compare NIS and TIS when the demand streams for new and

old items are negatively correlated. We model the negative correlation of demand

streams by making the sum of new and old item demand constant each period. In

our simulation, new item demand is uniformly distributed between 0 and 25, and old

item demand is 25 minus the new item demand. Therefore in each period new and old

item demands add up to 25. We run our simulations for 400,000 periods. Cost figures

(per period) for all parameter settings (i.e. h = 1, m ∈ {2, 5}, p1 ∈ {1, 3}, p2 ∈ {4, 9},αU ∈ {3, 7} and αD ∈ {2, 6}) under NIS and TIS are presented in Table 4.3. In Table

4.3 we observe that for 12 out of 32 parameter settings F and D provide the lowest

cost and in 17 instances U is the least costly and in the remaining 3 parameter settings

N has the lowest cost for NIS. Similarly, for TIS U provides the lowest cost in 17 out

97

Page 107: Essays on Perishable Inventory Management

F D U NNIS 34.66 34.66 30.29 32.10

TIS 34.17 34.17 33.46 44.16

% difference -1.43% -1.43% 10.46% 37.56%

Table 4.2: NIS vs. TIS when old and new item demand streams are negatively

correlated.

of 32 instances; for 14 out of 32 parameter settings F and D are both the least costly

(i.e. upward substitution is not done). For 1 instance N has the lowest cost.

We present the average (averaged over the 32 parameter settings) NIS and TIS

cost figures in Table 4.2. TIS is slightly better than NIS when downward substitution

is allowed. It is observed that when downward substitution is allowed no upward

substitution occurs, that is F and D become identical.

In Section 4.3.2 we indicated that, in general, if the metric M (see the Equation

4.3) is not larger than 0.2, downward substitution is viable. In this experiment there

are 12 parameter settings where M is less than or equal to 0.2. We observe that in

10 out of the 12 settings downward substitution is viable (under the better heuristic

ordering policy, TIS or NIS). Among the remaining 22 instances, in which M is greater

than 0.2, in 20 instances downward substitution is not viable. We should also mention

that, the rule of thumb works with 100% accuracy if NIS is always practiced as the

replenishment policy for this experiment.

4.3.4 TIS vs. NIS when new and old item demands are pos-

itively correlated

In order to see the effect of positively correlated demand we ran simulations with

uniform demand between 0 and 25. In the model we have old and new items exactly

the same. Therefore the demand streams in our simulation are perfectly positively

correlated. Cost figures (per period) for all parameter settings under NIS and TIS

are presented in Table 4.5. In Table 4.5 we observe that for 12 out of 32 parameter

settings F and D provide the lowest cost and in 20 instances U and N are the least

98

Page 108: Essays on Perishable Inventory Management

NIS TIS

F D U N F D U Nh=1, αD=2, αU=3, p2=4, p1=1, m=2 22.87 22.87 20.11 21.61 22.45 22.45 21.68 28.20

h=1, αD=2, αU=3, p2=4, p1=1, m=5 22.91 22.91 22.40 26.63 22.86 22.86 21.69 32.10

h=1, αD=2, αU=3, p2=4, p1=3, m=2 22.90 22.90 33.41 33.70 22.45 22.45 36.52 43.00

h=1, αD=2, αU=3, p2=4, p1=3, m=5 25.02 25.02 38.67 41.80 23.73 23.73 36.49 49.53

h=1, αD=2, αU=3, p2=9, p1=1, m=2 22.95 22.95 21.45 23.51 22.45 22.45 21.68 39.80

h=1, αD=2, αU=3, p2=9, p1=1, m=5 24.18 24.18 25.90 31.39 23.68 23.68 21.71 46.33

h=1, αD=2, αU=3, p2=9, p1=3, m=2 22.92 22.92 33.51 34.00 22.47 22.47 36.48 52.45

h=1, αD=2, αU=3, p2=9, p1=3, m=5 25.00 25.00 40.48 44.14 23.69 23.69 36.52 61.68

h=1, αD=2, αU=7, p2=4, p1=1, m=2 22.88 22.88 22.02 21.63 22.44 22.44 29.68 28.16

h=1, αD=2, αU=7, p2=4, p1=1, m=5 22.88 22.88 25.24 26.64 22.89 22.89 29.70 32.11

h=1, αD=2, αU=7, p2=4, p1=3, m=2 22.93 22.93 33.79 33.70 22.44 22.44 45.01 43.11

h=1, αD=2, αU=7, p2=4, p1=3, m=5 25.02 25.02 40.89 41.87 23.69 23.69 45.72 49.40

h=1, αD=2, αU=7, p2=9, p1=1, m=2 22.92 22.92 22.72 23.59 22.48 22.48 30.89 39.87

h=1, αD=2, αU=7, p2=9, p1=1, m=5 24.15 24.15 28.14 31.36 23.72 23.72 30.85 46.58

h=1, αD=2, αU=7, p2=9, p1=3, m=2 22.93 22.93 33.87 34.01 22.47 22.47 45.07 52.64

h=1, αD=2, αU=7, p2=9, p1=3, m=5 25.04 25.04 42.05 44.04 23.72 23.72 45.68 61.63

h=1, αD=6, αU=3, p2=4, p1=1, m=2 41.31 41.31 20.13 21.63 41.33 41.33 21.72 28.19

h=1, αD=6, αU=3, p2=4, p1=1, m=5 41.34 41.34 22.43 26.62 41.34 41.34 21.67 32.05

h=1, αD=6, αU=3, p2=4, p1=3, m=2 41.86 41.86 33.38 33.71 42.29 42.29 36.49 42.96

h=1, αD=6, αU=3, p2=4, p1=3, m=5 52.50 52.50 38.66 41.80 51.10 51.10 36.48 49.54

h=1, αD=6, αU=3, p2=9, p1=1, m=2 41.81 41.81 21.45 23.60 42.22 42.22 21.69 39.86

h=1, αD=6, αU=3, p2=9, p1=1, m=5 52.52 52.52 25.98 31.41 51.15 51.15 21.71 46.40

h=1, αD=6, αU=3, p2=9, p1=3, m=2 41.84 41.84 33.56 33.97 42.20 42.20 36.47 52.65

h=1, αD=6, αU=3, p2=9, p1=3, m=5 52.57 52.57 40.47 44.04 51.16 51.16 36.46 61.47

h=1, αD=6, αU=7, p2=4, p1=1, m=2 41.37 41.37 21.99 21.60 41.38 41.38 29.65 28.20

h=1, αD=6, αU=7, p2=4, p1=1, m=5 41.36 41.36 25.24 26.66 41.36 41.36 29.67 32.05

h=1, αD=6, αU=7, p2=4, p1=3, m=2 41.87 41.87 33.83 33.73 42.27 42.27 45.12 43.05

h=1, αD=6, αU=7, p2=4, p1=3, m=5 52.55 52.55 40.82 41.79 51.18 51.18 45.63 49.46

h=1, αD=6, αU=7, p2=9, p1=1, m=2 41.84 41.84 22.71 23.61 42.23 42.23 30.90 39.93

h=1, αD=6, αU=7, p2=9, p1=1, m=5 52.52 52.52 28.14 31.32 51.09 51.09 30.83 46.39

h=1, αD=6, αU=7, p2=9, p1=3, m=2 41.86 41.86 33.85 33.99 42.28 42.28 45.08 52.68

h=1, αD=6, αU=7, p2=9, p1=3, m=5 52.57 52.57 42.03 44.04 51.13 51.13 45.74 61.52

Average Cost: 34.66 34.66 30.29 32.10 34.17 34.17 33.46 44.16

Table 4.3: NIS and TIS cost figures for all parameter settings when old and new item

demand streams are negatively correlated.

99

Page 109: Essays on Perishable Inventory Management

F D U NNIS 35.83 35.83 32.11 32.11

TIS 43.34 44.97 41.67 44.20

% difference 20.96% 25.49% 29.80% 37.66%

Table 4.4: NIS vs. TIS when old and new item demand streams are positively corre-

lated.

costly for NIS; and for TIS the table shows that F provides the lowest cost in 14

out of 32 instances. For 14 out of 32 parameter settings U is the least costly, for 2

instances D has the lowest cost and for 2 instances N is the least costly.

Table 4.4 summarizes the computational study’s results for . The costs presented

in the table are averaged over the 32 parameter settings (i.e. h = 1, m ∈ {2, 5},p1 ∈ {1, 3}, p2 ∈ {4, 9}, αU ∈ {3, 7} and αD ∈ {2, 6}).

Table 4.4 shows that NIS outperforms TIS at least by 20% for any substitution

policy. Under positively correlated demand TIS does not benefit from the pooling

effect unlike the negatively correlated demand case.

In this experiment we observe that for all 12 parameter settings in which M (see

the Equation 4.3) is less than or equal to 0.2 downward substitution is viable. And

for all 22 instances in which M is greater than 0.2 downward substitution is not

viable. Our rule of thumb is working with 100% accuracy for NIS and TIS for this

experiment. This is in line with the observation that NIS outperforming TIS by a

greater margin under positively correlated demand. Therefore we can say that our

rule of thumb would be more accurate when demand steams are positively correlated

than when demand streams are negatively correlated.

4.3.5 NIS vs. TIS when new item demand is higher than old

item demand

In this section we compared NIS and TIS when the demand for new item demand is

greater than the old demand. We did experiments when new item demand distribution

100

Page 110: Essays on Perishable Inventory Management

NIS TIS

F D U N F D U Nh=1, αD=2, αU=3, p2=4, p1=1, m=2 22.17 22.17 21.61 21.61 26.13 26.88 27.07 28.16

h=1, αD=2, αU=3, p2=4, p1=1, m=5 25.95 25.95 26.64 26.64 28.32 29.49 30.63 32.10

h=1, αD=2, αU=3, p2=4, p1=3, m=2 32.35 32.35 33.78 33.78 36.19 38.04 41.29 43.05

h=1, αD=2, αU=3, p2=4, p1=3, m=5 39.25 39.25 41.77 41.77 41.27 43.43 47.34 49.54

h=1, αD=2, αU=3, p2=9, p1=1, m=2 24.01 24.01 23.58 23.58 33.63 36.44 34.96 39.84

h=1, αD=2, αU=3, p2=9, p1=1, m=5 30.02 30.02 31.37 31.37 37.00 40.41 41.95 46.41

h=1, αD=2, αU=3, p2=9, p1=3, m=2 32.48 32.48 33.97 33.97 39.06 45.39 46.02 52.59

h=1, αD=2, αU=3, p2=9, p1=3, m=5 40.70 40.70 44.11 44.11 47.33 52.13 56.01 61.75

h=1, αD=2, αU=7, p2=4, p1=1, m=2 22.18 22.18 21.61 21.61 27.09 26.86 28.51 28.17

h=1, αD=2, αU=7, p2=4, p1=1, m=5 25.91 25.91 26.61 26.61 29.11 29.49 31.63 32.10

h=1, αD=2, αU=7, p2=4, p1=3, m=2 32.33 32.33 33.77 33.77 38.57 38.04 43.65 43.11

h=1, αD=2, αU=7, p2=4, p1=3, m=5 39.21 39.21 41.83 41.83 42.79 43.44 48.77 49.48

h=1, αD=2, αU=7, p2=9, p1=1, m=2 23.99 23.99 23.55 23.55 35.03 36.37 37.61 39.92

h=1, αD=2, αU=7, p2=9, p1=1, m=5 29.91 29.91 31.29 31.29 38.38 40.40 43.65 46.54

h=1, αD=2, αU=7, p2=9, p1=3, m=2 32.43 32.43 34.01 34.01 42.38 45.35 49.52 52.69

h=1, αD=2, αU=7, p2=9, p1=3, m=5 40.67 40.67 44.04 44.04 49.31 52.15 58.27 61.80

h=1, αD=6, αU=3, p2=4, p1=1, m=2 30.78 30.78 21.63 21.63 35.32 35.66 27.03 28.17

h=1, αD=6, αU=3, p2=4, p1=1, m=5 33.04 33.04 26.64 26.64 36.33 37.00 30.69 32.12

h=1, αD=6, αU=3, p2=4, p1=3, m=2 43.78 43.78 33.77 33.77 51.49 52.23 41.31 43.02

h=1, αD=6, αU=3, p2=4, p1=3, m=5 48.72 48.72 41.84 41.84 53.53 54.60 47.29 49.51

h=1, αD=6, αU=3, p2=9, p1=1, m=2 34.70 34.70 23.54 23.54 46.70 48.70 34.90 39.89

h=1, αD=6, αU=3, p2=9, p1=1, m=5 39.51 39.51 31.39 31.39 49.02 51.55 41.91 46.37

h=1, αD=6, αU=3, p2=9, p1=3, m=2 44.49 44.49 34.02 34.02 59.21 61.97 46.03 52.70

h=1, αD=6, αU=3, p2=9, p1=3, m=5 51.54 51.54 44.09 44.09 62.61 65.68 55.97 61.76

h=1, αD=6, αU=7, p2=4, p1=1, m=2 30.74 30.74 21.64 21.64 35.76 35.64 28.51 28.16

h=1, αD=6, αU=7, p2=4, p1=1, m=5 33.00 33.00 26.61 26.61 36.75 36.94 31.59 32.05

h=1, αD=6, αU=7, p2=4, p1=3, m=2 43.83 43.83 33.79 33.79 52.41 52.18 43.63 43.05

h=1, αD=6, αU=7, p2=4, p1=3, m=5 48.65 48.65 41.81 41.81 54.31 54.67 48.80 49.51

h=1, αD=6, αU=7, p2=9, p1=1, m=2 34.71 34.71 23.55 23.55 47.68 48.69 37.60 39.89

h=1, αD=6, αU=7, p2=9, p1=1, m=5 39.57 39.57 31.38 31.38 49.91 51.40 43.66 46.42

h=1, αD=6, αU=7, p2=9, p1=3, m=2 44.49 44.49 34.08 34.08 60.58 61.94 49.55 52.67

h=1, αD=6, αU=7, p2=9, p1=3, m=5 51.54 51.54 44.07 44.07 63.83 65.77 58.25 61.79

Average Cost for NIS: 35.83 35.83 32.11 32.11 43.34 44.97 41.67 44.20

Table 4.5: NIS and TIS cost figures for all parameter settings when old and new item

demand streams are positively correlated.

101

Page 111: Essays on Perishable Inventory Management

is Uniform between 0 and 25 while old item demand demand distribution is Uniform

between 0 and 6.

We ran our simulation for 400,000 periods with the following parameters: h = 1,

m ∈ {2, 5}, p1 ∈ {1, 3}, p2 ∈ {4, 9}, αU ∈ {3, 7} and αD ∈ {2, 6}. On average TIS is

costlier than NIS by the following percentages: F by 5%; D by 20%; U by 9%; N by

28%. There are some instances (under F) where NIS is costlier than TIS with high

m, h, and low αU , αD,but overall NIS is superior.

4.3.6 NIS vs. TIS when new item demand is lower than old

item demand

In this section we compared NIS and TIS when the demand for new items is lower than

old items. We again run our simulation for 400,000 periods with the same parameters

(i.e. h = 1, m ∈ {2, 5}, p1 ∈ {1, 3}, p2 ∈ {4, 9}, αU ∈ {3, 7} and αD ∈ {2, 6}).We first did experiments when new item demand distribution is Uniform between

0 and 6 while old item demand demand distribution is Uniform between 0 and 25.

On average TIS outperforms NIS slightly under F and D, but is significantly

worse under U and N . The differences in average costs are: F by -1%; D by -1%; Uby 10%; N by 11%.

We observe that under low demand for new items F and D perform almost the

same (so do U and N ) as upward substitution is required less. In some instances (for

all four policies but mostly under F and D ) NIS is costlier than TIS with high m,

h, and low p1.

We then did experiments when there is no demand for new items, while old item

demand demand distribution is Uniform between 0 and 25 with zero downward sub-

stitution cost (i.e. FIFO). On average NIS is costlier than TIS by the following

percentages: F by 11%; D by 11%; U by 7%; N by 7%. In some instances (when p1

is high) U and N are costlier under TIS than NIS.

Thus, as new item becomes less important, TIS begins to outperform NIS.

102

Page 112: Essays on Perishable Inventory Management

4.3.7 Comparison of TIS and NIS when substitution is not

always accepted: Partial vs. Probabilistic Acceptance

In this section we numerically compare partial and probabilistic acceptance models.

In the partial acceptance model when substitution is offered a certain fraction of it is

accepted by customers. We use πD and πU to denote the fraction of substitution that

is accepted for downward and upward substitutions respectively. Under probabilistic

acceptance all of the offered substitution is either accepted or rejected (as a whole)

by the customer. Acceptance probabilities are denoted by φD and φU for downward

or upward substitution, respectively. We assume each substitution acceptance is

independent.

Table 4.6 shows the cost comparison of the policies under partial acceptance.

The simulations were done for 400,000 periods and only for one parameter setting

(p1 = 5, p2 = 9, h = 1, m = 5, αU = 7, αD = 6). Proportions for acceptance of

substitution (πD and πU ) are varied over (0, 0.1, 0.2, ..., 1). The cost (per period)

figures in the table is averaged across all proportions. Cost of N is given for the sake

of completeness and as a reference point.

F D U NNIS 48.86 48.96 43.83 44.10

TIS 58.80 61.18 57.72 61.75

% difference 20.34% 24.96% 31.68% 40.01%

Table 4.6: NIS vs. TIS when a proportion of substitution is accepted

Table 4.7 shows the cost comparison of the policies under probabilistic acceptance.

Similarly, the simulations were done for 400,000 periods and only for one parameter

setting (p1 = 5, p2 = 9, h = 1, m = 5, αU = 7, αD = 6). Probabilities for acceptance

of substitution (φD and φU ) are varied over (0, 0.1, 0.2, ..., 1) whenever there is a

chance of substitution. The cost (per period) figures in the table is averaged across

all probabilities. Again, cost of N is given for reference.

Tables 4.6 and 4.7 show that the models produce very similar results. We observe

that NIS outperforms TIS especially when substitution is offered/allowed less. The

103

Page 113: Essays on Perishable Inventory Management

F D U NNIS 48.86 49.08 43.71 44.07

TIS 60.26 62.84 57.90 61.73

% difference 23.34% 28.04% 32.46% 40.09%

Table 4.7: NIS vs. TIS under probabilistic acceptance of substitution

reason for this is as follows: New item inventory level is lower under TIS than under

NIS. As new items are more valuable, penalty costs will be higher if substitution is

not allowed. This would cause NIS to outperform TIS, particularly when substitution

is allowed or offered less.

4.3.8 Comparison of TIS and NIS when acceptance proba-

bility of substitution depends on substitution costs

It is quite possible that the acceptance probability of a substitution offer depends on

substitution costs. We conducted experiments and observed performances of TIS and

NIS by modelling acceptance probability of substitution as a function of substitution

cost as follows:

φD(αD) =αD

h + 2×m + p1

, (4.4)

φU(αU) =αU

m + p2

, (4.5)

where φD and φU are the acceptance probabilities for upward and downward substi-

tution, respectively.

Table 4.8 summarizes the cost figures under TIS and NIS with probabilistic accep-

tance when the acceptance probability is a function of substitution. In the experiment

we conducted, the cost parameter settings were as follows: p1 = 5, p2 = 9, h = 1,

m = 5, αU ∈ {0, ..., 14}, αD ∈ {0, ..., 14}. The cost numbers in the table are averages

across substitution costs and demand distribution is uniform between 0 and 25. Cost

of N is given for the sake of completeness and as a reference point. The table shows

that NIS outperforms TIS by at least 20%.

104

Page 114: Essays on Perishable Inventory Management

For this example, the optimal substitution costs for NIS are found as αD = 2 and

αU = 10 (i.e. φD(αD) = 0.14 and φU(αU) = 0.71) providing an average cost of 42.83.

Under TIS the optimal substitution costs are αD = 3 and αU = 7 (i.e. φD(αD) = 0.21

and φU(αU) = 0.50) with an average cost of 55.97.

F D U NNIS 57.36 57.49 43.86 44.07

TIS 69.46 70.89 59.38 61.72

% difference 21.10% 23.32% 35.37% 40.06%

Table 4.8: NIS vs. TIS when acceptance probability depends on substitution cost

4.3.9 Search for a good NIS order-up-to level: News-vendor-

type heuristic

In this section we propose and evaluate six news-vendor-type heuristics for NIS’s order

quantity. We did the numerical study with the 32 parameter settings. The demand

has the uniform distribution between between 0 and 25. The results are averaged

over these 32 cost parameters.

Heuristic 1 (H1):

In this heuristic we use p2 + p1 as the cost of understocking and h + m as the cost

of overstocking in the traditional news-vendor setting. Therefore the critical fractile

(CF) is as follows:

CF1 =p2 + p1

p2 + p1 + h + m(4.6)

The results, which are averaged across 32 parameter settings, are summarized in

the following table (Table 4.9):

105

Page 115: Essays on Perishable Inventory Management

F D U NCost S Cost S Cost S Cost S

H1 42.65 15.75 42.94 15.75 35.77 15.75 36.87 15.75

Opt. NIS 36.99 21.69 37.17 21.84 31.47 20.50 32.10 20.78

% difference 15.31% -27.38% 15.53% -27.90% 13.67% -23.17% 14.85% -24.21%

Table 4.9: Heuristic 1 vs Optimal NIS

Heuristic 2 (H2):

In this heuristic we use p2 as the cost of understocking and h + m− p1 as the cost of

overstocking in the traditional news-vendor setting. Therefore the critical fractile is:

CF2 =p2

p2 + h + m− p1

(4.7)

The results, which are averaged across 32 parameter settings, are summarized in

Table 4.10.

F D U NCost S Cost S Cost S Cost S

H2 39.98 18.13 40.23 18.13 33.62 18.13 34.51 18.13

Opt. NIS 36.99 21.69 37.17 21.84 31.47 20.50 32.10 20.78

% difference 8.10% -16.43% 8.24% -17.02% 6.84% -11.59% 7.51% -12.78%

Table 4.10: Heuristic 2 vs Optimal NIS

Heuristic 3 (H3):

In this heuristic we use p2 +p1 +αU as the cost of understocking and h+m as the cost

of overstocking in the traditional news-vendor setting. Therefore the critical fractile

is:

106

Page 116: Essays on Perishable Inventory Management

CF3 =p2 + p1 + αU

p2 + p1 + αU + h + m(4.8)

The results, which are averaged across 32 parameter settings, are summarized in

Table 4.11.

Heuristic 4 (H4):

In this heuristic we use p2 +p1 as the cost of understocking and h+m+αD as the cost

of overstocking in the traditional news-vendor setting. Therefore the critical fractile

is:

CF4 =p2 + p1

p2 + p1 + h + m + αD

(4.9)

Heuristic 5 (H5):

In this heuristic we use p2 + p1 + αU as the cost of understocking and h + m + αD as

the cost of overstocking in the traditional news-vendor setting. Therefore the critical

fractile is:

CF5 =p2 + p1 + αU

p2 + p1 + αU + h + m + αD

(4.10)

F D U NCost S Cost S Cost S Cost S

H3 40.08 18.19 40.41 18.19 32.83 18.19 33.87 18.19

Opt. NIS 36.99 21.69 37.17 21.84 31.47 20.50 32.10 20.78

% difference 8.37% -16.14% 8.70% -16.74% 4.33% -11.28% 5.52% -12.48%

Table 4.11: Heuristic 3 vs Optimal NIS

107

Page 117: Essays on Perishable Inventory Management

F D U NCost S Cost S Cost S Cost S

H4 49.62 11.94 49.83 11.94 45.07 11.94 46.11 11.94

Opt. NIS 36.99 21.69 37.17 21.84 31.47 20.50 32.10 20.78

% difference 34.15% -44.96% 34.05% -45.35% 43.22% -41.77% 43.62% -42.56%

Table 4.12: Heuristic 4 vs Optimal NIS

Heuristic 6 (H6):

In this heuristic we use p2 +αU as the cost of understocking and h+m−p1 as the cost

of overstocking in the traditional news-vendor setting. Therefore the critical fractile

is:

CF6 =p2 + αU

p2 + αU + h + m− p1

(4.11)

Comparison of all six heuristics:

Based on the analysis above H6 is the best heuristic among all six, producing results

within 4% of the optimal. H2 and H3 are within 9% of the optimal; H3 is better

than H2 when downward substitution is not allowed. H1 and H5 fourth and fifth best

heuristic respectively. H4 is the worst performing heuristic, its cost is within 35% to

45% of the optimal cost.

F D U NCost S Cost S Cost S Cost S

H5 44.05 14.81 44.32 14.81 38.32 14.81 39.46 14.81

Opt. NIS 36.99 21.69 37.17 21.84 31.47 20.50 32.10 20.78

% difference 19.09% -31.70% 19.23% -32.19% 21.77% -27.74% 22.93% -28.72%

Table 4.13: Heuristic 5 vs Optimal NIS

108

Page 118: Essays on Perishable Inventory Management

F D U NCost S Cost S Cost S Cost S

H6 38.17 20.44 38.42 20.44 31.77 20.44 32.52 20.44

Opt. NIS 36.99 21.69 37.17 21.84 31.47 20.50 32.10 20.78

% difference 3.21% -5.76% 3.37% -6.44% 0.97% -0.30% 1.32% -1.65%

Table 4.14: Heuristic 6 vs Optimal NIS

4.3.10 Summary of the Computational Study for Compari-

son of NIS and TIS

Based on our computations NIS outperforms TIS in most of the cases. First we

used various demand distributions while comparing TIS and NIS. We observed that

if downward substitution is viable and demand uncertainty is low than TIS outper-

forms NIS. When old item demand and new item demands are negatively correlated

TIS outperforms NIS if downward substitution is viable; however when old and new

demands are positively correlated NIS has at least 20% lower costs than TIS.

In order to incorporate customer behavior in terms of acceptance of substitution

we did two sets of experiments. First we assumed that customers accept only a

proportion of the substitution offer (i.e. partial acceptance of substitution). Second

we modeled the system such that acceptance of substitution is probabilistic in nature;

that is, based on a probability a substitution offer is rejected or accepted as a whole.

Our results show that both models yield similar results. We also present the TIS vs.

NIS comparison when acceptance probability depends on substitution cost.

As NIS is generally a better replenishment policy for our problem compared to

TIS; finding a good NIS order-up-to level without using computational methods is

important especially for the cases computations are time consuming. Therefore we

developed six heuristics to find a good NIS order-up-level based Newsvendor-type

logic. Based on the comparison of these heuristics we see that one of them performs

within 4% of the optimal NIS which looks promising in case computational solutions

are not readily available.

109

Page 119: Essays on Perishable Inventory Management

Chapter 5

Conclusion and Future Research

In Chapter 3 we formalize and compare four different fulfillment policies for perish-

able goods, Full-Substitution (F), Upward-Substitution (U), Downward-Substitution

(D), and No-Substitution (N ) under two different base-stock type inventory replen-

ishment policies, TIS and NIS. In this chapter we show that substitution may or may

not be beneficial with respect to operational costs, and provide conditions on cost

parameters that characterize the regions in which different substitution strategies are

most profitable, almost surely. We likewise show that downward substitution policies

are always beneficial with respect to average freshness of products.

In light of our analysis, we can consider the substitution strategies of some of the

examples in the introduction. Fresh produce suppliers are likely to benefit from sub-

stitution, mainly depending on the downward and upward substitution costs. On the

other hand, computer chip manufacturers typically cannot use upward substitution –

if a customer needs a fast chip they often will not be satisfied with a slow one (i.e. αU

is high). They do practice downward substitution after fusing the chip – to control

demand diversion (reducing αD).

In Chapter 4 we study the replenishment problem. We show that NIS is provably

optimal when no substitution is allowed or substitution is not possible. This is based

on the fact that every order decision is independent when substitution of any kind

does not occur when demand is i.i.d. Therefore NIS is optimal when substitution

policy is N or there is no demand for old items or there is no demand for new items.

We did a preliminary study and observed that the steady state distribution of on-hand

inventory does not appear to have a closed form, restricting our analytical studies to

110

Page 120: Essays on Perishable Inventory Management

policy comparisons rather than finding the optimal order-up-to levels analytically.

Based on our computational study on replenishment policies with a variety of

demand types, patterns, distributions and substitution choices we conclude that NIS

outperforms TIS in most cases. We also find that when new items are significantly

more valuable than older items NIS is more robust against demand fluctuations.

TIS might perform well when new and old item demands are negatively correlated

and downward substitution cost is low; or when demand has a low variance and

downward substitution cost is low. We also showed that a good heuristic (in order

get a near-optimal NIS order-up-to level) can be developed based on newsvendor-type

logic. Such heuristics might be beneficial in case determining the optimal NIS level

computationally is difficult.

The research within the perishable domain has largely been confined to inventory

management of a single product as the survey in Chapter 2 shows. However, grocery

or blood supply chains involve multiple perishable products with possibly differing

lifetimes. Joint replenishment is a typical practice in these industries, and analytical

research that studies the interaction between multiple items in ordering decisions -

focusing on economies of scale, or substitution/complementarity effects of products

with different lifetimes and in different categories - has not been studied. These

interactions provide opportunities for more complex control policies, which make

such problems both more challenging analytically, and potentially more rewarding

practically.

Considering multiple products, another problem that has not attracted much at-

tention from the research community is determining the optimal product-mix when

one type of perishable product can be used as a raw material for a second type of

product, possibly with a different lifetime, that we mentioned in Section 2.2. Deci-

sions regarding when and how much of a base product to sell/stock as is, versus how

much to process in order to obtain a final product with a different lifetime or different

potential value/revenue are quite common in blood and fresh produce supply chains.

A significant majority of the research on inventory management or distribution

of perishable goods disregards capacity constraints. While increasing the complexity

in the analysis, consideration of limited capacity will lead to more realistic models.

Consequently, more realistic models will enable design of heuristic policies that are

possibly more effective in practice.

111

Page 121: Essays on Perishable Inventory Management

The practical decision of when (if at all) to dispose of the aging inventory has

not received much attention even in single location models, possibly because capacity

is assumed to be unlimited and/or demand is assumed to be satisfied with FIFO

inventory issuance. However, disposal decisions are especially critical when capacity

is constrained (e.g., a retailer has limited shelf-space to display the products), and

customers choose the products based on their (perceived) freshness/quality. Veinott

(1960), in his deterministic model, included disposal decisions for a retailer of per-

ishable products with fixed lifetime. Martin (1986) studied optimal disposal policies

for a perishable product where demand is stochastic. His queueing model considers

the trade-off between retaining a unit in inventory for potential sales vs. salvaging

the unit at a constant value. Vaughan (1994) models an environment where a re-

tailer decides on the optimal parameter of a TIS policy and also a ‘sell-by’ date that

establishes an effective lifetime for the product with a random shelf life; this may

be considered a joint ordering and outdating policy. Vaughan (1994) discusses that

his model would be useful for retailers if they were to select suppliers based on their

potential for ordering and outdating, but no analysis is provided.

When customers prefer fresher goods, disposal and outdating are key decisions

that affect the age-composition (freshness) of inventory, and can influence the de-

mand. Analysis of simple and effective disposal and outdating policies, coordination

of disposal with replenishment policies, and analysis of inventory models where cus-

tomers (retailers) choose among suppliers and/or consider risk of supply/freshness

remain among the understudied research problems.

The majority of research on perishables assumes demand for a product is either

independent of its age, or that the freshest items are preferred. These typical as-

sumptions motivate the primary use of FIFO and LIFO issuance in inventory control

models. However, one can question how realistic these issuance policies would be,

especially in a business-to-business (B2B) setting. A service level agreement between

a supplier (blood center) and its retailer (hospital) may not be as strict as ‘freshest

items must be supplied’ (motivating LIFO) or as loose as ‘items of any age can be

supplied’ (motivating FIFO), but rather ‘items that will not expire within a specified

time-window must be supplied’. Faced with such a demand model, and possibly with

multiple retailers, a supplier can choose his/her optimal issuance policy which need

not necessarily be LIFO or FIFO. To the best of knowledge, Ishii (1993), Ishii and

112

Page 122: Essays on Perishable Inventory Management

Nose (1996), and Haijema et al. (2005a, 2005b) are the only ones that have shown

some awareness of this important issue; see our discussion in Section 2.5. For future

research on perishables to be of more practical use, we need demand models that are

representative of the more general business rules and policies today.

While problems that involve competition (among retailers, suppliers, or supply

chains) have received a lot of attention in the last decade (see, for e.g., Cachon,

1998), models that include competition involving perishable and aging products have

not appeared in the literature. One distinct feature of competition in a perishable

commodity supply chain is that suppliers (retailers) may compete not only on avail-

ability and/or price but also on freshness.

In the produce industry, a close look at the relationship between suppliers and

buyers reveals several practical challenges. Perosio et al. (2001) present survey results

that indicate that about 9% of the produce in the U.S. is sold through spot markets,

and about 87.5% of product purchases are made under contracts with suppliers. This

motivates Perosio et al. (2001) to make the following observation, which is essen-

tially a call for further research: “Despite a number of considerable disadvantages, in

general, todays buyers and sellers alike appear to be won over by the greater price cer-

tainty that contracting makes possible...However, high degrees of product perishability,

weather uncertainty and resulting price volatility, and structural differences between

and among produce buyers and sellers create significant challenges to the design of

the produce contract.”

Recently, Burer et al. (2006) introduced different types of contracts used in the

agricultural seed industry and investigated - via single-period models - whether the

supply chain can be coordinated using these contracts. In their ongoing work, Boya-

batli and Kleindorfer (2006) study the implications of a proportional product model

(where one unit of input is processed to produce proportional amounts of multiple

agricultural outputs) on the optimal mix of long-term and short-term (spot) contract-

ing decisions. We believe further analysis of supplier-retailer relations, and the design

of contracts to improve the performance of a supply chain that involves perishable

products remain fruitful research topics.

Pricing was mentioned as one of the important research directions by Prastacos

(1984) to encourage collaboration between hospitals and blood banks/centers; this

is also echoed in Pierskalla (2004). There seems to be almost no research in this

113

Page 123: Essays on Perishable Inventory Management

direction to date. According to a recent survey in the US, the mean cost of 250ml. of

fresh frozen plasma to a hospital varied from $20 to $259.77, average costs of blood

components were higher in Northeastern states compared to the national average, and

hospitals with higher surgical volume typically paid less than the national average for

blood components in 2004 (AABB, 2005). Given the importance of health care both

for the general welfare and the economy, there is a pressing need for further research

to understand what causes such variability in this environment, and whether pricing

can be combined with inventory management to better match demand for perishable

blood components with the supply. The potential relevance of such work reaches well

beyond the health care industry.

Advances in technology have increased the efficiency of conventional supply chains

significantly; for perishable goods, technology can potentially have an even greater

impact. Not only is there the potential to enable information flow among different

parties in a supply chain, as has proved to be valuable in conventional chains, but there

is also the possibility of detecting and recording the age of the products in stock (e.g.

when RFID is implemented). This information can be used to affect pricing decisions,

especially of products nearing their usable lifetimes. Moreover, advances in technology

can potentially increase the freshness and extend the lifetime of products (e.g., when

better storage facilities or packaging equipment is used). The relative magnitudes of

these benefits calibrated to different product and market characteristics remains an

important open problem.

The majority of the work on the analysis of inventory management policies assume

that the state of the system is known completely; i.e., inventory levels of each age of

product at each location are known. However, this may not be the case in practice.

Cohen et al. (1981a) discuss the need for detailed demand and inventory informa-

tion to apply shortage or outdate anticipating transshipment rules in a centralized

system, and argue that the system would be better without transshipments between

the retailers if accurate information is not available. Chande et al. (2005) presents

an RFID architecture for managing inventories of perishable goods in a supply chain.

They describe how the profile of current on-hand inventory, including the age, can

be captured on a real-time basis, and conclude by stating that “there is a need for

measures and indicators ... to determine ... (a) whether such development would be

beneficial, and (b) when implemented, how the performance of the system compares

114

Page 124: Essays on Perishable Inventory Management

to the performance without auto ID enhancements.”

In Chapters 3 and 4 our focus was on the effect of substitution given practical

inventory policies and on the analysis of various replenishment policies, respectively.

There are a number of possible future directions for research closely related to the re-

search these two chapters. A challenge lies in the case where a supplier does not have

to commit to a single substitution option and substitution does not take place only as

a recourse (i.e. substitution decision need not be “static”). Although such a policy is

more difficult to implement (and to analyze), it provides more flexibility in satisfying

the demand. Hence, investigation of static versus dynamic substitution decisions in

fulfilling the demand warrants further attention. Likewise, evaluating substitution

policies for a product that has a lifetime of more than two periods is intriguing; even

for three-period lifetime case, substitution from new to old (in presence of products

with medium age) may not be justified. In a related direction, using a multiple pe-

riod lifetime with zero null demand for the first several periods of lifetime offers the

possibility of modelling positive lead-times in replenishing perishable goods. Another

worthwhile research area lies in investigating demand models based on customer-

choice (where the choice can be price-driven or shelf-life driven). Adding customer

behavior to the existing model or developing new models with customer gaming issues

are also promising lines of future research. In terms of future research for compu-

tational study on management of perishable inventory, one can conduct numerical

experiments with various demand distributions to test our heuristics or any other

possible heuristics.

115

Page 125: Essays on Perishable Inventory Management

Bibliography

[1] AABB, 2005. The 2005 Nationwide Blood Collection and Utilization Survey Re-

port. published by The American Association of Blood Banks (AABB), Bethesda,

Maryland, MD.

[2] Abdel-Malek, L.L. and H. Ziegler (1988). Age Dependent Perishability in Two-

echelon Serial Inventory Systems. Computers and Operations Research 15: 227-

238.

[3] Adachi, Y., T. Nose and S. Kuriyama (1999). Optimal Inventory Control Policy

Subject To Selling Prices of Perishable Commodities. International Journal of

Production Economics, 60-61: 389-394.

[4] Adenso-Diaz, B., M. Gonzalez, and E. Garcia (1998). A Hierarchical Approach

to Managing Dairy Routing Source. Interfaces, 28: 21-31.

[5] Allen, S.J. and E.W. Schuster (2004). Controlling the Risk for Agricultural Har-

vest. Manufacturing & Service Operations Management, 6: 225-236.

[6] Angle, T. (2003). CEO, Greater Alleghenies Blood Services Region, American

Red Cross, Personal Contact.

[7] Avsar, Z.M., M. Baykal-Gursoy. (2002). Inventory control under substitutable

demand: A stochastic game application. Naval Research and Logistics 49 (4)

359-375.

[8] Axsater, S. (2000). Inventory Control. Kluwer International Series in Operations

Research and Management Science, Kluwer Academic Publishers, Boston, MA.

116

Page 126: Essays on Perishable Inventory Management

[9] Axsater, S. (2003). Supply Chain Operations: Serial and Distribution Inventory

Systems. In Handbooks in OR & MS, Vol.11, Supply Chain Management: De-

sign, Coordination and Operations, A.G. de Kok and S. Graves (eds.), Elsevier,

Amsterdam, The Netherlands.

[10] Bar-Lev, S.K., D. Perry and W. Stadje (2005). Control Policies for Inventory

Systems with Perishable Items: Outsourcing and Urgency Classes. Probability in

the Engineering and Informational Sciences, 19: 309-326.

[11] Bassok, Y., Anupindi, R., R. Akella. (1999). Single-period multiproduct inven-

tory models with substitution. Operations Research 47 (4) 632-642.

[12] Berk, E. and U. Gurler (2006). Analysis of the (Q,r) Inventory Model for Per-

ishables with Positive Lead Times and Lost Sales. Working paper. Faculty of

Business Administration, Bilkent University, Ankara,Turkey.

[13] Bitran, G. and R. Caldentey (2003). An Overview of Pricing Models for Revenue

Management. Manufacturing & Service Operations Management, 5: 203-229.

[14] Boyabatli, O. and P. Kleindorfer (2006). Integrating Long-term and Short-term

Contracting in Fed-Cattle Supply Chains: Proportional Product Model. Pre-

sented at the INFORMS Annual Meeting, Pittsburgh, November, 2006.

[15] Brodheim, E., C. Derman and G.P. Prastacos (1975). On the Evaluation of a

Class of Inventory Policies for Perishable Products such as Blood. Management

Science, 22: 1320-1325.

[16] Bulinskaya, E.V. (1964). Some Results Concerning Optimum Inventory Policies.

Theory of Probability and Its Applications, 9: 389-402.

[17] Burer, S., P.C. Jones and T.J. Lowe (2006). Coordinating the Supply Chain in the

Agricultural Seed Industry. Working paper. Department of Management Sciences,

University of Iowa, Iowa City, IA.

[18] Cachon, G.P. (1998). Competitive Supply Chain Inventory Management. in

Quantitative Models for Supply Chain Management, S. Tayur, R. Ganeshan and

M. Magazine (eds.), Kluwer, Boston.

117

Page 127: Essays on Perishable Inventory Management

[19] Cachon, G.P. (2003). Supply Chain Coordination with Contracts, Handbooks in

Operations Research and Management Science: Supply Chain Management, S.

Graves and T. de Kok (eds.), North Holland.

[20] Chande, A., N. Hemachandra, N. Rangaraj and S. Dhekane (2004). Fixed Life

Perishable Inventory Problem and Approximation under Price Promotion. Tech-

nical report. Industrial Engineering and Operations Research, IIT Bombay,

Mumbai, India.

[21] Chande, A., S. Dhekane, N. Hemachandra and N. Rangaraj (2005). Perishable

Inventory Management and Dynamic Pricing using RFID Technology. Sadhana,

30: 445-462.

[22] Chandrashekar, K., N. Dave, N. Hemachandra and N. Rangaraj (2003). Timing

of Discount Offers for Perishable Inventories. In Proceedings of Sixth Asia Pacific

Operations Research Society. M.R. Rao, M.C. Puri (eds.), Allied Publishers, New

Delhi.

[23] Chazan D. and S. Gal (1977). A Markovian Model for a Perishable Product

Inventory. Management Science, 23: 512-521.

[24] Chen, F., J. Eliashberg, and P. Zipkin (1998). Customer Preferences, Supply-

Chain Costs, and Product-Line Design. In Product Variety Management Re-

search Advances. by Ho, T-H. and C.S. Tang (eds.), Kluwer.

[25] Chen, F. (2002). Information Sharing and Supply Chain Coordination. In Hand-

book of Operations Research and Management Science: Supply Chain Manage-

ment, by T. de Kok and S. Graves (eds.), North-Holland.

[26] Chen, J.-M. and T.-H. Chen (2005). Effects of Joint Replenishment and Channel

Coordination for Managing Multiple Deteriorating Products in a Supply Chain,

Journal of the Operational Research Society, 56: 1224-1234.

[27] Chen, Z.-L. (2006). Integrated Production and Outbound Distribution Scheduling

in a Supply Chain: Review and Extensions. Working paper, Robert H. Smith

School of Business, University of Maryland, College Park. MD.

118

Page 128: Essays on Perishable Inventory Management

[28] Chiu, H.N. (1995). An Approximation to The Continuous Review Inventory

Model with Perishable Items and Lead Times. European Journal of Operational

Research, 87: 93-108.

[29] Clark, A.J. and H. Scarf (1960). Optimal Policies for a Multi-Echelon Inventory

Problem. Management Science, 45: 475-490.

[30] Cohen, M. (1976). Analysis of Single Critical Number Ordering Policies for Per-

ishable Inventories. Operations Research, 24: 726-741.

[31] Cohen, M. and W. P. Pierskalla (1979). Simulation of Blood Bank Systems. ACM

SIGSIM Simulation Digest, 10: 14-18.

[32] Cohen M., W. P. Pierskalla and H. Yen (1981a). An Analysis of Ordering and Al-

location Policies for Multi-Echelon, Age-Differentiated Inventory Systems. TIMS

Studies in the Management Sciences, 16: 353-378.

[33] Cohen M., W. P. Pierskalla and R.J. Sassetti (1981b). Regional Blood Inventory

Control and Distribution. Proceedings of the 1980 Conference on the Manage-

ment and Logistics of Blood Banking, National Heart, Lung and Blood Institute,

v.5 (October 1981), pp. 21-88.

[34] Colomina, R., Humphries, B. and T. Steger. (2006). A new supply chain trend

for wireless carriers Profiting in a changing market through lessons learned from

the retail industry. Supply Chain Management Viewpoint

[35] Cooper, W. (2001). Pathwise Properties and Performance Bounds for a Perish-

able Inventory System. Operations Research, 49: 455-466.

[36] Dave, U. (1991). Survey of Literature on Continuously Deteriorating Inventory

Models – A Rejoinder. The Journal of the Operational Research Society, 42: p.

725.

[37] Deuermeyer, B.L. (1979). A Multi-type Production System for Perishable Inven-

tories. Operations Research, 27: 935-943.

[38] Deuermeyer, B.L. (1980). A Single Period Model for a Multi-product Perishable

Inventory System with Economic Substitution. Naval Research and Logistics, 27:

177-185.

119

Page 129: Essays on Perishable Inventory Management

[39] Diks, E.B. and A.G. de Kok (1998). Optimal Control of a Divergent Multi-echelon

Inventory System. European Journal of Operational Research, 111: 75-97.

[40] Elmaghraby, W. and P. Keskinocak (2003). Dynamic Pricing in the Presence

of Inventory Considerations: Research Overview, Current Practices, and Future

Directions. Management Science, 49: 1287-1309.

[41] Eppen, G. and L. Schrage (1981). Centralized Ordering Policies in A Multi-

warehouse System with Lead Times and Random Demand. in Multi-Level Pro-

duction/Inventory Control Systems: Theory and Practice, L.B.Schwarz (ed.),

North-Holland, Amsterdam.

[42] Ernst R., P. Kouvelis (1999). The effects of selling packaged goods on inventory

decisions. Management Science 45 (8) 1142-1155.

[43] Eynan A., T. Fouque (2003). Capturing the risk-pooling effect through demand

reshape. Management Science 49: 704-717.

[44] Federgruen, A., G.P. Prastacos and P. Zipkin (1986). An Allocation and Distri-

bution Model for Perishable Products. Operations Research, 34: 75-82.

[45] Ferguson, M. and O. Koenigsberg (2006). How Should a Firm Manage Deterio-

rating Inventory? Production and Operations Management, forthcoming.

[46] Ferguson, M. and M. E. Ketzenberg (2006). Sharing Information to Improve Re-

tail Product Freshness of Perishables. Production and Operations Management,

15: 57-73.

[47] Fransoo, J.C. and M.J.F. Wouters (2000). Measuring the Bullwhip Effect in the

Supply Chain. Supply Chain Management: An International Journal, 5: 78-89.

[48] Fries, B. (1975). Optimal Ordering Policy for A Perishable Commodity with

Fixed Lifetime. Operations Research, 23: 46-61.

[49] Fujiwara, O., H. Soewandi and D. Sedarage (1997). An Optimal Ordering and

Issuing Policy for A Two-stage Inventory System for Perishable Products. Euro-

pean Journal of Operational Research, 99: 412-424.

120

Page 130: Essays on Perishable Inventory Management

[50] Goh, C., B.S. Greenberg and H. Matsuo (1993). Two-stage Perishable Inventory

Models. Management Science, 39: 633-649.

[51] Golden, B.L., A.A. Assad and E.A. Wasil (2001). Routing Vehicles in the Real

World: Applications in the Solid Waste, Beverage, Food, Dairy, and Newspaper

Industries. In The Vehicle Routing Problem, P. Toth and D. Vigo, (eds.), Society

for Industrial and Applied Mathematics, Philadelphia, PA, pp. 245-286.

[52] Goyal, S.K. and B.C. Giri (2001). Recent Trends in Modeling of Deteriorating

Inventory. European Journal of Operational Research, 134: 1-16.

[53] Graves, S. C. (1982). The Application of Queueing Theory to Continuous Per-

ishable Inventory Systems. Management Science, 28: 400-406.

[54] Gregor, P. J., R.N. Forthofer and A.S. Kapadia (1982). Evaluation of Inventory

and Transportation Policies of a Regional Blood Distribution System. European

Journal of Operational Research, 10: 106-113.

[55] Grocery Manufacturers of America (2004). 2004 Unsalables Benchmark Report.

http://www.gmabrands.com/industryaffairs/docs/benchgma2004.pdf.

[56] Gurler U. and B. Y. Ozkaya (2003). A Note on ”Continuous review perishable

inventory systems: models and heuristics”. IIE Transactions, 35: 321 - 323.

[57] Gurler U. and B. Y. Ozkaya (2006). (s, S) Policy for Perishables with Random

Shelf Life. Working Paper, Bilkent University, Ankara, Turkey.

[58] Hahn, K.Y., H. Hwang and S.W. Shinn (2004). A Returns Policy for Distribu-

tion Channel Coordination of Perishable Items, European Journal of Operational

Research, 152: 770-780.

[59] Haijema, R., J. van der Wal and N. van Dijk (2005a). Blood Platelet Produc-

tion: A Multi-type Perishable Inventory Problem, Research Memorandum, FEE,

University of Amsterdam.

[60] Haijema, R., J. van der Wal and N. van Dijk (2005b). Blood Platelet Production:

Optimization by Dynamic Programming and Simulation, to appear in Computers

and Operations Research Special Issue on OR and Health Care.

121

Page 131: Essays on Perishable Inventory Management

[61] Hemmelmayr, V., K.F. Doerner, R.F. Hartl and M.W.P. Savelsbergh (2006).

Delivery Strategies for Blood Products Supplies. Working paper, School of In-

dustrial and Systems Engineering, Georgia Institute of Technology, Atlanta, GA.

[62] Ho, T-H. and C.S. Tang (Eds.) (1998). Product Variety Management Research

Advances. Kluwer’s International Series in Operations Research & Management

Science Vol. 10.

[63] Huq, F., S. Asnani, V. Jones and K. Cutright (2005). Modeling the In-

fluence of Multiple Expiration Dates on Revenue Generation in the Supply

Chain.International Journal of Physical Distribution & Logistics Management,

35: 152-160.

[64] Ishii, H. (1993). Perishable Inventory Problem with Two Types of Customers and

Different Selling Prices. Journal of the Operations Research Society of Japan, 36:

199-205.

[65] Ishii, H. and T. Nose (1996). Perishable Inventory Control with Two Types

of Customers and Different Selling Prices under the Warehouse Capacity Con-

straint. International Journal of Production Economics, 44: 167-176.

[66] Jagannathan, R. and T. Sen (1991). Storing Crossmatched Blood: A Perishable

Inventory Model with Prior Allocation, Management Science, 37: 251-266.

[67] Jones, P.C., T. Lowe, R.D. Traub and G. Keller (2001). Matching Supply and

Demand: The Value of A Second Chance in Producing Hybrid Seed Corn. Man-

ufacturing & Service Operations Management, 3: 122-137.

[68] Jones, P.C., G. Keller, T. Lowe and R.D. Traub (2003). Managing the Seed-Corn

Supply Chain at Syngenta, INTERFACES, 33: 80-90.

[69] Kalpakam, S. and G. Arivarignan (1988). A Continuous Review Perishable In-

ventory Model. Statistics, 19: 389-398.

[70] Kalpakam, S. and K.P. Sapna (1994). Continuous Review (s, S) Inventory System

with Random Lifetimes and Positive Lead Times. Operations Research Letters,

16: 115-119.

122

Page 132: Essays on Perishable Inventory Management

[71] Kalpakam, S. and K.P. Sapna (1996). A Lost Sales (S-1,S) Perishable Inventory

System with Renewal Demand. Naval Research Logistics, 43: 129-142.

[72] Kalpakam, S. and S. Shanthi (2000). A Perishable System with Modified Base

Stock Policy and Random Supply Quantity. Computers and Mathematics with

Applications, 39: 79-89.

[73] Kalpakam, S. and S. Shanthi (2001). A Perishable System with Modified

(S − 1, S) Policy and Arbitrary Processing Times. Computers and Operations

Research, 28: 453-471.

[74] Kanchanasuntorn, K. and A. Techanitisawad (2006). An Approximate Peri-

odic Model for Fixed-Life Perishable Products in A Two-echelon Inventory-

Distribution System. International Journal of Production Economics, 100: 101-

115.

[75] Karmarkar, U. (1981). The Multi-period Multi-location Inventory Problem. Op-

erations Research, 29: 215-228.

[76] Kaspi, H. and D. Perry (1983). Inventory Systems of Perishable Commodities.

Advances in Applied Probability, 15: 674-685.

[77] Kaspi, H. and D. Perry (1984). Inventory Systems of Perishable Commodities

with Renewal Input and Poisson Output. Advances in Applied Probability, 16:

402-421.

[78] Katsaliaki, K. and S.C. Brailsford (2007). Using Simulation to Improve the Blood

Supply Chain. Journal of the Operational Research Society, 58: 219-227.

[79] Kazaz, B. (2004). Production Planning under Yield and Demand Uncertainty

with Yield-dependent Cost and Price. Manufacturing & Service Operations Man-

agement, 6: 209-224.

[80] Kendall, K.E. and S.M. Lee (1980). Formulating Blood Rotation Policies with

Multiple Objectives. Management Science, 26: 1145-1157.

[81] Ketzenberg, M.E. and M. Ferguson (2006). Managing Slow Moving Perishables

in the Grocery Industry. Working paper, College of Business, Colorado State

University, Fort Collins, CO.

123

Page 133: Essays on Perishable Inventory Management

[82] Konda, C., N. Dave, N. Hemachandra and N. Rangaraj (2003). Timing of Dis-

count Offers for Perishable Inventories. In Proceedings of Sixth Asia Pacific Op-

erations Research Society. M.R. Rao, M.C. Puri (eds.), Allied Publishers, New

Delhi.

[83] Lian, Z. and L. Liu (1999). A Discrete-time Model for Perishable Inventory Sys-

tems. Annals of Operations Research, 87: 103 - 116.

[84] Lian, Z. and L. Liu (2001). Continuous Review Perishable Inventory Systems:

Models and Heuristics. IIE Transactions, 33: 809-822.

[85] Lian, Z., L. Liu and M. Neuts (2005). A Discrete-time Model for Common Life-

time Inventory Systems. Mathematics of Operations Research, 30: 718 - 732.

[86] Lin, C.-W. R. and H.-Y. S. Chen (2003). Dynamic Allocation of Uncertain Supply

for the Perishable Commodity Supply Chain. International Journal of Production

Research, 41: 3119-3138.

[87] Liu, L. (1990). (s, S) Continuous Review Inventory Models for Inventory with

Random Lifetimes. Operations Research Letters, 9: 161-169.

[88] Liu, L. and K. L. Cheung (1997). Service Constrained Inventory Models with

Random Lifetimes and Lead Times. Journal of the Operational Research Society,

48: 1022-1028.

[89] Liu, L. and Z. Lian (1999). (s, S) Continuous Review Models for Products with

Fixed Lifetimes. Operations Research, 47: 150-158.

[90] Liu, L. and D. Shi (1999). An (s, S) Model for Inventory with Exponential Life-

times and Renewal Demands. Naval Research Logistics, 46: 39 - 56.

[91] Liu, L. and T. Yang (1999). An (s, S) Random Lifetime Inventory Model with a

Positive Lead Time. European Journal of Operations Research, 113: 52 - 63.

[92] Lowe, T. and P. V. Preckel (2004). Decision Technologies for Agribusiness Prob-

lems: A Brief Review of Selected Literature and a Call for Research. Manufac-

turing & Service Operations Management, 6: 201-208.

124

Page 134: Essays on Perishable Inventory Management

[93] Lutke Entrup, M. (2005). Advanced Planning in Fresh Food Industries: Integrat-

ing Shelf Life into Production Planning. Contributions to Management Science

Series, Physica-Verlag, Heidelberg.

[94] Lystad, E., M. Ferguson and C. Alexopoulos (2006). Single Stage Heuristics for

Perishable Inventory Control in Two-Echelon Supply Chains. Working paper,

The College of Management, Georgia Institute of Technology, Atlanta, GA.

[95] Mahajanan V., E. Muller. (1996). Timing, diffusion and substitution of successive

generation of technological innovations: The IBM mainframe case. Technological

Forecasting and Social Change 51: 109-132.

[96] Mahajan S., G. van Ryzin. (2001). Stocking retail assortments under dynamic

consumer substitution. Operations Research 49: 334-351.

[97] Martin, G.E. (1986). An Optimal Decision Model for Disposal of Perishable

Inventories, International Journal of Production Research, 24: 73-80.

[98] McGillivray, S. (1976). Some concepts for inventory control under substitutable

demand. INFOR 16: 47-63.

[99] Mendelson H., R.R. Pillai. (1999). Industry clockspeed: Measurement and oper-

ational implications. M&SOM 1: 1-20.

[100] Moore G. (1965). Cramming more components onto integrated circuits. Elec-

tronics Magazine 4.

[101] Moorthy, A.K., Y.C. Narasimhulu and I.R. Basha (1992). On Perishable In-

ventory with Markov Chain Demand Quantities. International Journal of Infor-

mation and Management Science, 3: 29 - 37.

[102] Moorthy, K.S. (1984) Market Segmentation, Self-Selection, and Product Line

Design. Marketing Science, 3 (Fall): 288-307.

[103] Muharremoglu, A. and J. Tsitsiklis (2003). A Single-Unit Decomposition Ap-

proach to Multi-Echelon Inventory Systems. Working paper, Graduate School of

Business, Columbia University, New York, NY.

125

Page 135: Essays on Perishable Inventory Management

[104] Mustafee, N., K. Katsaliaki, S.J.E. Taylor and S. Brailsford (2006). Distributed

Simulation with Cots Simulation Packages: A Case Study in Health Care Supply

Chain Simulation, Proceedings of the 2006 Winter Simulation Conference, L.F.

Perrone, F.P. Wieland, J. Liu, B.G. Lawson, D.M. Nicol and R.M. Fujimoto

(eds.), pp. 1136-1142.

[105] Nahmias, S. (1975a). Optimal Ordering Policies for Perishable Inventory-II.

Operations Research, 23: 735-749.

[106] Nahmias, S. (1975b). A Comparison of Alternative Approximations for Ordering

Perishable Inventory. INFOR, 13: 175-184.

[107] Nahmias, S. (1975c). On Ordering Perishable Inventory under Erlang Demand.

Naval Research and Logistics Quarterly, 22: 415-425.

[108] Nahmias, S. (1976). Myopic Approximations for the Perishable Inventory Prob-

lem. Management Science, 22: 1002-1008.

[109] Nahmias, S. (1977a). Higher-order Approximations for the Perishable Inventory

Problem. Operations Research, 25: 630-640.

[110] Nahmias, S. (1977b). Comparison between Two Dynamic Perishable Inventory

Models. Operations Research, 25: 168-172.

[111] Nahmias, S. (1977c). On Ordering Perishable Inventory when Both Demand

and Lifetime are Random. Management Science, 24: 82-90.

[112] Nahmias, S. (1978). The Fixed Charge Perishable Inventory Problem. Opera-

tions Research, 26: 464-481.

[113] Nahmias, S. (1982). Perishable Inventory Theory: A Review. Operations Re-

search, 30: 680-708.

[114] Nahmias, S. and W. Pierskalla (1973). Optimal Ordering Policies for A Product

that Perishes in Two Periods subject to Stochastic Demand. Naval Research and

Logistics Quarterly, 20: 207-229.

126

Page 136: Essays on Perishable Inventory Management

[115] Nahmias, S. and W. Pierskalla (1976). A Two Product Perishable/Non-

perishable Inventory Problem. S.I.A.M Journal of Applied Mathematics, 30: 483-

500.

[116] Nahmias, S. and S.S. Wang (1979). A Heuristic Lot Size Reorder Point Model

for Decaying Inventories. Management Science, 25: 90-97.

[117] Nahmias, S., D. Perry and W. Stadje (2004a). A Perishable Inventory Systems

with Variable Input and Demand rates. Mathematical Methods of Operations

Research, 60: 155-162.

[118] Nahmias, S., D. Perry and W. Stadje (2004b). Actuarial Valuation of Perishable

Inventory Systems. Probability in the Engineering and Informational Sciences,

18: 219-232.

[119] Nandakumar, P. and T.E. Morton (1993). Near Myopic Heuristics for the Fixed-

life Perishability Problem. Management Science, 39: 1490-1498.

[120] Netessine, S., N. Rudi. (2003). Centralized and competitive inventory models

with demand substitution. Operations Research 51: 329-335.

[121] Nose, T., T. Ishii and T. Nishida (1983). LIFO Allocation Problem for Per-

ishable Commodities. Journal of the Operations Research Society of Japan, 26:

135-146.

[122] Or, I. and W. Pierskalla (1979). A Transportation Location-Allocation Model

for Regional Blood Banking. AIIE Transactions, 11: 86-95.

[123] Pal, M. (1989). The (s − 1, S) Inventory Model for Deteriorating Items with

Exponential Lead Time. Calcutta Statistical Association Bulletin, 38: 83 - 91.

[124] Parlar M., S.K. Goyal. (1984). Optimal ordering decisions for two substitutable

products with stochastic demands. OPSEARCH 21: 1-15.

[125] Parlar, M. (1985). Optimal Ordering Policies for A Perishable and Substitutable

Product - A Markov Decision-model. INFOR, 23: 182-195.

[126] Parlar, M. (1988). Game theoretic analysis of the substitutable product inven-

tory problem with random demands. Naval Research and Logistics 35: 397-409.

127

Page 137: Essays on Perishable Inventory Management

[127] Pasternak, B., Z. Drezner. (1991). Optimal inventory policies for substitutable

commodities with stochastic demand. Naval Research and Logistics 38: 221-240.

[128] Perosio, D.J., E. W. McLaughlin, S. Cuellar and K. Park (2001). FreshTrack

2001 Supply Chain Management in the Produce Industry, Department of Agri-

cultural, Resource, and Managerial Economics College of Agriculture and Life

Sciences Cornell University, Ithaca, NY 14853.

[129] Perry, D. (1997). A Double Band Control Policy of a Brownian Perishable

Inventory System. Probability in the Engineering and Informational Sciences,

11: 361-373.

[130] Perry, D. and M. Posner (1990). Control of Input and Demand Rates in Inven-

tory Systems of Perishable Commodities. Naval Research Logistics, 37: 85-97.

[131] Perry, D. and M. Posner (1998). An (S − 1, S) Inventory System with Fixed

Shelf Life and Constant Lead Times. Operations Research, 46: S65-S71.

[132] Perry, D. and W. Stadje (1999). Perishable Inventory Systems with Impatient

Demands. Mathematical Methods of Operations Research, 50: 77-90.

[133] Perry, D. and W. Stadje (2000a). An Inventory System for Perishable Items

with By-Products. Mathematical Methods of Operations Research, 51: 287-300.

[134] Perry, D. and W. Stadje (2000b). Inventory Systems for Goods with Censored

Random Lifetimes. Operations Research Letters: 27: 21-27.

[135] Perry, D. and W. Stadje (2001). Disasters in Markovian Inventory Systems for

Perishable Items. Advances in Applied Probability, 33: 61-75.

[136] Perry, D. and W. Stadje (2006). A Controlled M/G/1 Workload Process with

an Application to Perishable Inventory Systems. Mathematical Methods of Op-

erations Research, 64: 415-428.

[137] Pierskalla, W.P. (2004). Supply Chain Management of Blood Banks. In Oper-

ations Research and Health Care, A Handbook of Methods and Applications, M.

Brandeau, F. Sainfort and W.P. Pierskalla (eds.), Kluwer Academic Publishers,

New York, 104-145.

128

Page 138: Essays on Perishable Inventory Management

[138] Pierskalla, W.P. and C.D. Roach (1972). Optimal Issuing Policies for Perishable

Inventory. Management Science, 18: 603-614.

[139] Prastacos, G.P. (1978). Optimal Myopic Allocation of a Product with Fixed

Lifetime. The Journal of the Operational Research Society, 29: 905-913.

[140] Prastacos, G.P. (1979). LIFO Distribution Systems, The Journal of the Opera-

tional Research Society, 30: 539-546.

[141] Prastacos, G.P. (1981). Allocation of a Perishable Product Inventory. Opera-

tions Research, 29: 95-107.

[142] Prastacos, G.P. (1984). Blood Inventory Management: An Overview of Theory

and Practice. Management Science, 30: 777-800.

[143] Prastacos, G.P. and E. Brodheim (1980). PBDS: Decision Support System for

Regional Blood Management.Management Science, 26: 451-463.

[144] Popp, W. and H. Vollert (1981). A Comparison of Decentralized and Cen-

tralized Disposition Systems for Red Cell Concentrates. In The Computer and

Blood Banking, J.R. Mohr and A. Kluge (eds.), Springer-Verlag, Lecture Notes

on Medical Informatics, v. 13, pp. 141-154.

[145] Raafat, F. (1991). Survey of Literature on Continuously Deteriorating Inventory

Models, The Journal of the Operational Research Society, 42: 27-37.

[146] Ravichandran, N. (1995). Stochastic Analysis of a Continuous Review Perish-

able Inventory System with Positive Lead Time and Poisson Demand. European

Journal of Operational Research, 84: 444 - 457.

[147] Scarf, H. (1960). The Optimality of (s, S) Policies in Dynamic Inventory Prob-

lems. In Mathematical Models in the Social Sciences, K. Arrow, S. Karlin and P.

Suppes (eds.), Stanford University Press, Stanford, CA.

[148] Schmidt, C.P. and S. Nahmias (1985). (S - 1,S) Policies for Perishable Inventory.

Management Science, 31: 719-728.

[149] Talluri, K. and G. van Ryzin (2004). The Theory and Practice of Revenue

Management. Kluwer Academic Publishers, Boston.

129

Page 139: Essays on Perishable Inventory Management

[150] Tekin, E., U. Gurler and E. Berk (2001). Age-based vs. Stock-level Control

Policies for a Perishable Inventory System. European Journal of Operational Re-

search, 134: 309-329.

[151] Tsiros, M. and C. M. Heilman (2005). The Effect of Expiration Dates and

Perceived Risk on Purchasing Behavior in Grocery Store Perishable Categories.

Journal of Marketing, 69: 114-129.

[152] van der Vorst, J.G.A.J., A.J.M. Beulens and P. van Beek (2000). Modelling

and Simulating Multi-Echelon Food Systems. European Journal of Operational

Research, 122: 354-366.

[153] Van Zyl, G. J. J. (1964). Inventory Control for Perishable Commodities. Ph.D.

Thesis, University of North Carolina, Chapel Hill, NC.

[154] Vaughan, T. (1994). A Model of the Perishable Inventory System with Refer-

ence to Consumer- Realized Product Expiration. The Journal of the Operational

Research Society, 45: 519-528.

[155] Veinott, A.F. (1960). Optimal Ordering, Issuing and Disposal of Inventory with

Known Demand. Ph.D. Thesis, Columbia University, New York, NY.

[156] Walter, C. (2005). Kryder’s Law. Scienfitic American August.

[157] Wang, Q., M. Parlar. (1994). A 3-person game-theory model arising in sto-

chastic inventory control-theory. European Journal of Operational Research 76:

83-97.

[158] Webb, A. (2006). Ripeness Sticker Takes the Guesswork Out of Picking and

Eating. The Albuquerque Journal, Thursday, May 25, 2006,

URL: http://www.abqjournal.com/biz/462732business05-25-06.htm .

[159] Weiss, H. (1980). Optimal Ordering Policies for Continuous Review Perishable

Inventory Models Operations Research, 28: 365-374.

[160] Williams, C.L. and B.E. Patuwo (1999). A Perishable Inventory Model with

Positive Order Lead Times. European Journal of Operational Research, 116: 352-

373.

130

Page 140: Essays on Perishable Inventory Management

[161] Williams, C.L. and B.E. Patuwo (2004). Analysis of the Effect of Various Unit

Costs on the Optimal Incoming Quantity in A Perishable Inventory Model. Eu-

ropean Journal of Operational Research, 156: 140-147.

[162] Yang, P.C. and H.M. Wee (2006). A Collaborative Inventory System with Per-

missable Delay in Payment for Deteriorating Items, Mathematical and Computer

Modeling, 43: 209-221.

[163] Yano, C., and G. Dobson (1998). Profit-Optimizing Product Line Design, Se-

lection and Pricing with Manufacturing Cost Consideration. In Product Variety

Management Research Advances. by Ho, T-H. and C.S. Tang (eds.), Kluwer.

[164] Yen, H. (1975). Inventory Management for a Perishable Product: Multi-echelon

system. PhD Thesis, Northwestern University, Evanston, IL.

[165] Yi, J. (2003). Vehicle Routing with Time Windows and Time-Dependent Re-

wards:A Problem from the American Red Cross. Manufacturing & Service Op-

erations Management, 5: 74-77.

[166] Zhou, D. and W. Pierskalla (2006). Perishable Product Inventory Policies with

Emergency Replenishments. Working Paper, Chinese University of Hong Kong.

131

Page 141: Essays on Perishable Inventory Management

APPENDIX

The proofs of the analytical results are given below.

A.1. Cost Comparison Results under TIS

To prove the results on TIS, we use the definition of a pair introduced in Section 3.4.2.

CnI is the cost of substitution policy I in period n including holding, spoilage and

penalty costs, XnI and Y n

I are the amount of new and old items in stock, respectively,

for policy I, I = F ,N ,U ,D. Recall that we define a period as E (for equal) if

XnN = Xn

F ; we call the periods between two E periods a cycle and J is the number of

pairs in a cycle. If an E period follows another one, then it is called a trivial cycle.

A non-trivial cycle, starts with an E period in which downward substitution takes

place and ends with another E.

Proof of Proposition 3

Proof : Assume period n is an F period with XnF = Xn

N + ∆ for some ∆ > 0. If

demand for new items in this period is greater than or equal to the number of new

items under N (i.e. Dn2 ≥ Xn

N) then N will be at the inventory state (S, 0) in the next

period which implies that the period n + 1 is not an F period. Therefore Dn2 must

be less than XnN . That is, F has more than ∆ unsold new items, which are available

for substitution. If there is no need for the unsold new items (Dn1 ≤ S − Xn

N − ∆)

then all of these items would be carried over to the next period. However this would

make the period n + 1 an N period because under policy N , less new items would be

carried. Hence if there are two consecutive F periods, then there must be downward

substitution in the first one.

Proof of Proposition 4

Proof : Continuing from the proof of Proposition 3 let Dn1 = S − Xn

N − ∆ + k1

and Dn2 = Xn

N − k2 where k1, k2 > 0. Therefore Xn+1N = S − k2 and Xn+1

F =

S − (∆ + k2 − k1)+ as F has ∆ + k2 to substitute and k1 is needed. Since period

n+ 1 is an F period k2 must be greater than (∆ + k2− k1)+ thus k1 > ∆. (Note that

k1 > ∆ implies demand for old items in period n must be more than old item inventory

132

Page 142: Essays on Perishable Inventory Management

under N , i.e. Dn1 > S − Xn

N .) The substitution amount is ∆n = min{∆ + k2, k1}and Xn+1

F − Xn+1N = k2 − (∆ + k2 − k1)

+ hence Xn+1F − Xn+1

N = ∆n − ∆ (i.e. the

substitution amount minus ∆).

Proof of Lemma 4

We divide the proof of Lemma 4 into two Propositions.

Proposition 7 If conditions (3.5) and (3.10) hold, then F has lower cost than Noutside the pairs.

Proof : We look at a series of F periods starting or ending either with an E period

or a pair. Let period n be any F period in the series with ∆n = XnF − Xn

N . As

was pointed out in the proof of Proposition 4 to have another F in period n + 1,

Dn2 = Xn

N −k2 < XnN and Dn

1 = S−XnN +k1 > S−Xn

N . Then Xn+1F = S− (k2−k1)

+

and Xn+1N = S − k2 (i.e. ∆n+1 = min(k1, k2)). Also:

CnN = hk2 + p1k1, Cn

F = h(k2 − k1)+ + p1(k1 − k2)

+ + αD(∆n + ∆n+1),

CnN − Cn

F = (h + p1)∆n+1 − αD(∆n + ∆n+1).

Let the series start at period 1 and end at period M − 1 (i.e. ∆i > 0 ∀ i =

1, ..., M − 1). In periods 0 and M there are 2 possibilities each:

• If period 0 is an E (i.e. ∆0 = 0), then:

C0N − C0

F = (h + p1)∆1 − αD∆1.

• If period 0 is an N (i.e. end of a pair ), then: In this case C0N −C0

F would be an

amount carried over from the previous pair (the one that just ended in period

0). In other words an amount of (h + p1)∆1 − αD∆1, which is saved from the

previous pair (this is explained later in Proposition 8), will be used as C0N −C0

F .

Thus,

C0N − C0

F = (h + p1)∆1 − αD∆1. (5.1)

133

Page 143: Essays on Perishable Inventory Management

• If period M is an E (i.e. ∆M = 0) then:

CM−1N − CM−1

F ≥ −αD∆M−1. (5.2)

The proof of (5.2) is as follows; if period M − 1 is F there are six demand

scenarios that lead to an E period immediately after. We analyze these possible

cases:

Case 1. DM−12 = XM−1

N − k2 and DM−11 = S − XM−1

N + k1 with k1, k2 ≥ 0.

These demands result in XMF = S−(k2−k1)

+ and XMN = S−k2. F is short

of ∆M−1 + k1 old items while it has ∆M−1 + k2 unsold new items. Thus

minimum of these is the downward substitution amount. Then we have

CM−1N = hk2+p1k1 and CM−1

F = h(k2−k1)++p1(k1−k2)

++αD min(∆M−1+

k2, ∆M−1 + k1). In order for period M be an E, k1 or k2 (or both) must be

0; hence CM−1N − CM−1

F = −αD∆M−1.

Case 2. DM−12 = XM−1

N + k2 and D1 = S −XM−1N −∆M−1 + k1 with 0 ≤ ki <

∆M−1 for i ∈ {1, 2} so that XMF = S − (∆M−1 − k2 − k1)

+ and XMN = S.

The downward substitution amount is ∆M−1 − k2 because in order for

period M be an E, k1 + k2 ≥ ∆M−1 must hold. Then we have CM−1N =

p2k2 + m(∆M−1 − k1) and CM−1F = p1(k2 + k1 −∆M−1) + αD(∆M−1 − k2);

hence CM−1N − CM−1

F = (p2 − p1)k2 + p1(∆M−1 − k1) + m(∆M−1 − k1) −αD(∆M−1 − k2) > −αD∆M−1.

Case 3. DM−12 = XM−1

N +k2 and DM−11 = S−XM−1

N +k1 with k1, k2 ≥ 0 result

in XMF = XM

N = S. F is short of ∆M−1+k1 old items while it has ∆M−1−k2

unsold new items. Thus the downward substitution amount is ∆M−1− k2.

Then CM−1N = p2k2 +p1k1 and CM−1

F = p1(k1 +k2)+αD(∆M−1−k2); hence

CM−1N − CM−1

F = (p2 − p1)k2 − αD(∆M−1 − k2) > −αD∆M−1.

Case 4. DM−12 = XM−1

N + ∆M−1 + k2 and DM−11 = S − XM−1

N − ∆M−1 − k1

with k1, k2 ≥ 0. F is short of k2 old items while it has k1 unsold old

items. Then, there is an upward substitution in the amount of min(k2, k1)

resulting in XMF = XM

N = S. Thus CM−1N = p2(∆M−1+k2)+m(∆M−1+k1)

and CM−1F = p2(k2− k1)

+ + αU min(k1, k2) + m(k1− k2)+. If k1 < k2, then

CM−1N − CM−1

F = (m + p2)(∆M−1 + k1) − αUk1 > −αD∆M−1; otherwise

CM−1N −CM−1

F = (m+p2)(∆M−1+k2)−αUk2 > −αD∆M−1 as αU < m+p2.

134

Page 144: Essays on Perishable Inventory Management

Case 5. DM−12 = XM−1

N +∆M−1 +k2 and DM−11 = S−XM−1

N −∆M−1 +k1 with

∆M−1 > k1 ≥ 0, k2 ≥ 0 result in XMF = XM

N = S. No substitution takes

place. Thus, CM−1N = p2(∆M−1 + k2) + m(∆M−1− k1) and CM−1

F = p2k2 +

p1k1. Hence CM−1N −CM−1

F = m(∆M−1−k1)+p2∆M−1−p1k1 > −αD∆M−1.

Case 6. DM−12 = XM−1

N + ∆M−1 + k2 and DM−11 = S − XM−1

N + k1 with

k1, k2 ≥ 0 result in XMF = XM

N = S. There is no substitution, and CM−1N =

p2(∆M−1 + k2) + p1k1 and CM−1F = p2k2 + p1(∆M−1 + k1); hence CM−1

N −CM−1

F = (p2 − p1)∆M−1 > −αD∆M−1.

Therefore (5.2) is correct.

• If in period M a pair starts (i.e. the period M and M + 1 are F and N respec-

tively) then:

CM−1N − CM−1

F = (h + p1)∆M − αD(∆M−1 + ∆M).

We reduce this amount by (h + p1 − αD)∆M ; we save this (h + p1 − αD)∆M

as the “starting cost” of N for the forthcoming pair in period M and add it to

the cost of N for the cost comparison in the pairs (Proposition 8). Therefore

in this case:

CM−1N − CM−1

F = −αD∆M−1. (5.3)

Thus we have the following cost structure:

C0N − C0

F = (h + p1)∆1 − αD∆1,

C1N − C1

F = (h + p1)∆2 − αD(∆1 + ∆2),

C2N − C2

F = (h + p1)∆3 − αD(∆2 + ∆3),

· · ·CM−2

N − CM−2F = (h + p1)∆M−1 − αD(∆M−2 + ∆M−1),

CM−1N − CM−1

F ≥ −αD∆M−1.

Hence,M−1∑

i=0

CiN − Ci

F ≥ (h + p1 − 2αD)∆, (5.4)

where

∆ =M−1∑

i=1

∆i.

135

Page 145: Essays on Perishable Inventory Management

Then left term in (5.4) is positive (i.e. F is less costly) as αD < (h+p1)/2 is assumed.

Proposition 8 If conditions (3.8) and (3.9) hold, then F has lower cost than Ninside the pairs.

Proof : Both N and F start at an E state at the beginning. Until the period in

which downward substitution occurs both have same inventory state, they are E’s.

The period after the substitution takes place becomes an F period (i.e. F has more

new items than N ). Due to this substitution F incurs a substitution cost αD but

saves h + p1 as a penalty cost is avoided by fulfilling a demand for old product and

that item is sold instead of being held in inventory. Therefore, F incurs h + p1 − αD

less cost than N per substitute item. There might be a number of F periods until

an N period; using the accounting explained preceding equation (5.3), we say that at

the beginning of a pair, F has a cost lower by an amount equal to the starting cost

(h + p1 − αD per substitution).

Let the first pair start in period M with XMF XM

N + ∆M and XM+1N = XM+1

F + ∆.

Intuitively this means under F more items (in the amount of ∆) have been aged and

∆ less items perished at the end of period M resulting a cost difference of (m− h)∆

in favor of F . (If these items did not perish then the ∆ items would not have aged,

they would have been substituted.)

We first prove that “the first part of the pair cost” until period M + 1 is at least

(m + p1−αD)∆. To have an N period in period M + 1 there are four possibilities for

the demand in period M :

Case i. DM2 = XM

N −k2 and DM1 = S−XM

N −∆M−k1 with k1, k2 ≥ 0. No substitution

takes place and the resulting inventory levels are XM+1F = S − ∆M − k2 and

XM+1N = S − k2 (i.e. ∆ = ∆M). Then CM

N = hk2 + m(∆M + k1) and CMF =

h(∆M +k2)+mk1. Hence CMN −CM

F = (m−h)∆M . After adding (h+p1−αD)∆M

we have (m + p1 − αD)∆ as the first part of the pair cost.

Case ii. DM2 = XM

N − k2 and DM1 = S −XM

N −∆M + k1 with ∆M > k1 ≥ 0, k2 ≥ 0.

There is downward substitution in the amount of k1 so that XM+1F = S−(∆M +

k2−k1) and XM+1N = S−k2 (i.e. ∆ = ∆M−k1). Then CM

N = hk2 +m(∆M−k1)

136

Page 146: Essays on Perishable Inventory Management

and CMF = h(∆M +k2−k1)+αDk1. Hence CM

N −CMF = (m−h)(∆M−k1)−αDk1.

After adding (h + p1−αD)∆M , we have (m + p1−αD)∆ + (h + p1− 2αD)k1 as

the first part of the pair cost.

Case iii. DM2 = XM

N + k2 and DM1 = S−XM

N −∆M − k1 with ∆M > k2 ≥ 0, k1 ≥ 0.

There is no substitution then XM+1F = S − (∆M − k2) and XM+1

N = S (i.e.

∆ = ∆M−k2). Therefore CMN = p2k2+m(∆M+k1) and CM

F = h(∆M−k2)+mk1.

Hence CMN −CM

F = (m−h)(∆M−k2)+(m+p2)k2. After adding (h+p1−αD)∆M

we have (m + p1−αD)∆ + (h + p1−αD + m + p2)k2 as the first part of the pair

cost.

Case iv. DM2 = XM

N + k2 and DM1 = S − XM

N − ∆M + k1 with 0 ≤ ki < ∆M

for i ∈ {1, 2}. Under F there are ∆M − k2 unsold new items and k1 more

old items are needed and in order to have a pair k1 < ∆M − k2 must hold.

Therefore XM+1F = S − (∆M − k2 − k1) and XM+1

N = S (i.e. ∆ = ∆M − k2 − k1

). Then CMN = p2k2 + m(∆M − k1) and CM

F = h(∆M − k2 − k1) + αDk1.

Hence CMN −CM

F = (m− h)(∆M − k2 − k1) + (m + p2)k2 − αDk1. After adding

(h+p1−αD)∆M we have (m+p1−αD)∆+(h+p1−αD+m+p2)k2+(h+p1−2αD)k1

as the first part of the pair cost.

In period M + 1 if high demand for new items is seen this would benefit N since

there is more new item under N in this period. We do a case-by-case analysis to

show that this benefit is not greater than its cost, (m + p1 − αD)∆. The proof is by

induction, that is we first show that the claim is true for J = 1 (J is number of pairs

in the cycle) and 2, then assume it is true for J = j and prove it stays correct for

J = j + 1. For J = 1 there are 3 possibilities. By examining these possible cases:

Case 1. DM+12 = XM+1

F + ∆ + k2 and DM+11 = S −XM+1

F + k1 where k1 and k2 ≥ 0:

CM+1F = p1k1 + p2(∆ + k2), CM+1

N = p1(∆ + k1) + p2k2

CM+1F − CM+1

N = (p2 − p1)∆ ≤ (m + p1 − αD)∆

as we assume p2 − p1 ≤ m + p1 − αD.

Case 2. DM+12 = XM+1

F +∆+k2 and DM+11 = S−XM+1

F −∆+k1 where ∆ ≥ k1 ≥ 0

and k2 ≥ 0:

CM+1F = p2(∆ + k2 − (∆− k1)) + αU(∆− k1) = p2(k2 + k1) + αU(∆− k1),

137

Page 147: Essays on Perishable Inventory Management

CM+1N = p2k2 + p1k1

CM+1F − CM+1

N = (p2 − p1)k1 + αU(∆− k1) ≤ (m + p1 − αD)∆

as we assume αU + αD ≤ m + p1.

Case 3. DM+12 = XM+1

F + ∆ + k2 and DM+11 = S − XM+1

F − ∆ − k1 where k1 and

k2 ≥ 0:

CM+1F = p2(∆ + k2 − (∆ + k1))

+ + m(∆ + k1 − (∆ + k2))+

= p2(k2 − k1)+ + m(k1 − k2)

+ + αU(∆ + k1)

CM+1N = p2k2 + mk1

If k2 ≥ k1, then CM+1F −CM+1

N = −(p2 +m)k1 +αU(∆+ k1) ≤ (m+ p1−αD)∆.

If k2 < k1, then CM+1F −CM+1

N = −(p2 +m)k2 +αU(∆+ k2) < (m+ p1−αD)∆.

For all three cases above, in the next period F and N reach the same state, an E, as

DM+12 = XM+1

F + ∆ + k2. If there are more F periods from period n until the cycle

closure the claim is true due to Proposition 3.

For J = 2, there are three cases (of the original nine) with DM+12 = XM+1

F + k2

where ∆ > k2 ≥ 0, and the inventory state for F is (S, 0) while the state for Nis (S − (∆ − k2), ∆ − k2). This implies the period M + 2 is an F period with

∆M+2 = ∆− k2.

Note that while proving for J = 2, in order to make a correct cost analysis, we

subtract (h + p1 − αD)(∆M+2) from CM+1N for the first pair. We are “saving” this

amount to use in the second pair in this cycle (either a series of F s and then a pair

or an immediate pair) as we mentioned in the proof of Proposition 7; see equation

(5.1). This makes the J = 2 case identical to J = 1 case from that period on:

Case 4. DM+12 = XM+1

F + k2 and DM+11 = S − XM+1

F + k1 where k1 ≥ 0 and

∆ > k2 ≥ 0:

CM+1F = p2k2 + p1k1, CM+1

N = h(∆− k2) + p1(∆ + k1)− (h + p1 − αD)(∆− k2)

CM+1F − CM+1

N = (p2 − p1)k2 − αD(∆− k2) < (m + p1 − αD)∆

138

Page 148: Essays on Perishable Inventory Management

Case 5. DM+12 = XM+1

F + k2 and DM+11 = S − XM+1

F −∆ + k1 where ∆ ≥ k1 ≥ 0

and ∆ > k2 ≥ 0:

CM+1F = p2[k2 − (∆− k1)]

+ + m[∆− k1 − k2]+ + αU min(k2, ∆− k1)

CM+1N = h(∆− k2) + p1k1 − (h + p1 − αD)(∆− k2).

If k1 + k2 ≤ ∆, then

CM+1F − CM+1

N = m∆−mk1 −mk2 − h∆ + hk2 − p1k1 + αUk2

+ (h + p1 − αD)(∆− k2)

= (m + p1)(∆− k1 − k2) + αUk2 − αD(∆− k2)

= (m + p1 − αD)(∆− k2) + αUk2 − (m + p1)k1 < (m + p1 − αD)∆.

If k1 + k2 > ∆, then

CM+1F − CM+1

N = p2k2 + p2k1 − p2∆− h∆ + hk2 − p1k1 + αU(∆− k1)

+ (h + p1 − αD)(∆− k2)

= p2(k2 + k1 −∆) + p1(∆− k2 − k1) + h(−∆ + k2 + ∆− k2)

+ αU(∆− k1)− αD(∆− k2)

= (p2 − p1)(k1 − (∆− k2)) + αU(∆− k1)− αD(∆− k2)

< (m + p1 − αD)∆.

Case 6. DM+12 = XM+1

F +k2 and DM+11 = S−XF −∆−k1 where k1 and ∆ > k2 ≥ 0:

CM+1F = m(∆ + k1 − k2) + αUk2,

CM+1N = h(∆− k2) + mk1 − (h + p1 − αD)(∆− k2)

CM+1F − CM+1

N = m(∆ + k1 − k2)− h(∆− k2)−mk1 + (h + p1 − αD)(∆− k2) + αUk2

= (m + p1 − αD)(∆− k2) + αUk2 < (m + p1 − αD)∆.

We also have the cases where DM+12 = XM+1

F − k2 (i.e. new items are abundant) and

there is no opportunity for the N policy to cash in on the extra new items that it

has in period n. We also show that for these cases CM+1F −CM+1

N < (m + p1 − αD)∆

holds again, saving (h + p1 − αD)∆M+2:

139

Page 149: Essays on Perishable Inventory Management

Case 7. DM+12 = XM+1

F − k2 and DM+11 = S −XM+1

F + k1 where k1 ≥ 0 and k2 ≥ 0:

XM+2F = S − (k2 − k1)

+, XM+2N = S − (∆ + k2) then ∆M+2 = ∆ + min(k1, k2).

Costs are as follows:

CM+1F = h(k2 − k1)

+ + p1(k1 − k2)+ + αD min(k1, k2),

CM+1N = h(∆ + k2) + p1(∆ + k1)− (h + p1 − αD)∆M+2.

If k1 < k2:

CM+1F − CM+1

N = −αD∆ < (m + p1 − αD)∆.

Case 8. DM+12 = XM+1

F − k2 and DM+11 = S − XM+1

F −∆ + k1 where ∆ ≥ k1 ≥ 0

and k2 ≥ 0:

XM+2F = S − k2, XM+2

N = S − (∆ + k2) then ∆M+2 = ∆. Costs are follows:

CM+1F = hk2 + m(∆− k1), CM+1

N = h(∆ + k2) + p1k1 − (h + p1 − αD)∆M+2.

If k1 < k2:

CM+1F − CM+1

N = (m + p1)(∆− k1)− αD∆, ≤ (m + p1 − αD)∆.

Case 9. DM+12 = XM+1

F − k2 and DM+11 = S −XM+1

F −∆− k1 where k1 and k2 ≥ 0:

XM+2F = S − k2, XM+2

N = S − (∆ + k2) then ∆M+2 = ∆. Costs are follows:

CM+1F = hk2 + m(∆ + k1), CM+1

N = h(∆ + k2) + mk1 − (h + p1 − αD)∆M+2.

If k1 < k2:

CM+1F − CM+1

N = (m + p1 − αD)∆.

Hence F has lower cost than N in the pairs as well, and from then on we have a

J = 1 case.

Assume the claim is true for J = j and F has extra (h + p1 − αD)(∆− k2) at the

end of the last pair as in case J = 2. If the (j + 1)st pair is immediately after the

140

Page 150: Essays on Perishable Inventory Management

jth one; then F with the extra (h + p1 − αD)(∆ − k2) benefit, will be less costly if

m + 2p1 > p2 + αD holds. If there is no pair as an immediate successor of the jth

pair, then there will be a number of F periods before the pair j + 1 and F will have

a h + p1 − αD benefit before this next pair as discussed in Proposition 7. Then the

proof is identical to the J = 1 case. Hence the claim is true.

Proof of Lemma 7

Proof : For cases in which there is no upward substitution, U is identical to N and

we know that F is better than N under conditions (3.5) and (3.7). So we look at

cycles where there is upward substitution.

For E periods, if there is any upward substitution, then F and U have the same

costs and the next period is also an E period.

If there is upward substitution by U or F in an F period, then the next period

is either a U period (i.e. there is a pair) or an E (specifically both policies reach the

inventory state (S, 0) ). Let n − 1 be such an F period with Xn−1F = Xn−1

U + ∆′.

To have an upward substitution, Dn−12 must be greater than Xn−1

U which makes

XnU = S. Let ∆ = Xn

U − XnF ; then it must hold that ∆

′> ∆ ≥ 0. This implies

Dn−12 = Xn−1

U + ∆′ − ∆; and we also see that when ∆ = 0 then period n is E with

state of (S,0).

We compare U with N : Xn−1N = Xn−1

U + ∆′ − ∆ and in period n, both U and

N are at state (S, 0). The cost of having the ∆′difference is (h + p1 − αD)∆

′more

for U as compared to F . As Xn−1F − Xn−1

N = ∆, cost of U is greater than N by

(h + p1 − αD)(∆′ − ∆) for the difference of new items because fewer items were

substituted downward to get a net difference of ∆. U incurs the cost of upward

substitution, which is αU per item. Thus, if Dn−11 ≤ S − Xn−1

U − (∆′ − ∆) then U

is more costly than policy N by (h + p1 − αD)(∆′ − ∆) + αU(∆

′ − ∆) until period

n. Otherwise (i.e. if Dn−11 > S − Xn−1

U − (∆′ − ∆)), U will have to substitute less

than ∆′ − ∆ and will incur p2 for every new item demand that it cannot fulfill by

substitution while N incurs p1 as penalty costs. Therefore, for every state in U , we

can find a counterpart in N that has a lower cost than U . As N has higher cost than

F , U is also more costly than F .

141

Page 151: Essays on Perishable Inventory Management

A.2. Cost Comparison Results under NIS

Similar to what we did in Section 3.4, we define the pairs, and divide the horizon

into different classes of periods. However, in this case a pair is defined based on the

inventory level of old items: If the amount of old goods in stock in period n is higher

under policy F , then we call that period an F period. We call a period E if the

amount of old items in stock for two policies is equal. We define cycle, trivial cycle,

non-trivial cycle, and CnI for I = F ,N ,U ,D as previously.

Proof of Lemma 12

Proof : By Proposition 5, in any given period n Y nN = Y n

F + ∆ where ∆ ≥ 0 is the

downward substitution amount in period n− 1. In an E period, ∆ = 0; otherwise it

is N. F and N have the same cost in a trivial cycle. We look at a non-trivial cycle.

For each substitution, F incurs a substitution cost αD but saves h + p1; F incurs

h + p1 − αD less cost than N per substitute item. In this N period (let this period

be the nth period) there are three possibilities for F :

i. No substitution takes place: Let the demand for old items be Dn1 = Y n

F + K.

As Y nN = Y n

F + ∆ we have the cost difference between the policies for the cycle

given below. (Note that we do not include the costs that are incurred in cycle-

ending E periods in our “cycle cost”. We account for that cost in the subsequent

cycle.)

n∑

i=n−1

CiN−Ci

F = (h+p1−αD)∆+

−p1∆ if K ≥ ∆, Dn2 ≥ S,

m(∆−K)− p1K if ∆ > K ≥ 0, Dn2 ≥ S,

m∆ if 0 > K ≥ −XnF , Dn

2 ≤ S.

The next period is an E and the cycle ends. Thus, for this case, the claim is

true.

ii. Upward substitution takes place: To have upward substitution Dn2 must be

greater than S and Dn1 must be less than Y n

F , hence Dn2 = S+L and Dn

1 = YF−K

where L,K > 0. The cost difference between F andN for the cycle is as follows:n∑

i=n−1

CiN − Ci

F = (h + p1 − αD + m)∆ + (m + p2 − αU)us

where us = min{K, L} is the upward substitution amount. As the next period

is an E, the cycle ends. Therefore the claim is true for this case.

142

Page 152: Essays on Perishable Inventory Management

iii. Downward substitution takes place: A downward substitution in period n

causes the period n + 1 to be an N, therefore the cycle does not end in period

n + 1. There can be a number of Ns in a cycle, we define J as this number.

For J = 2, let Dn2 = S − L and Dn

1 = Y nF + K where S > L ≥ 0 and K > 0.

Downward substitution takes place twice in this cycle, in periods n− 1 and n.

We compare the costs of F and N through the periods n− 1, n and n+1. The

cost difference between N and F for the first two periods of the cycle (n − 1

and n) is as follows:

n∑

i=n−1

CiN−Ci

F = (h+p1−αD)∆+(h+p1−αD)dsn−p1K+m(∆−K)++p1(K−∆)+

where dsn = min{K, L} is the downward substitution amount in period n. Note

that ∆ = dsn−1. Thus:

n∑

i=n−1

CiN−Ci

F = (h+p1−αD)∆+(h+p1−αD)dsn−p1K+

{m(∆−K) if K < ∆,p1(K −∆) if K ≥ ∆.

= (h+p1−αD)dsn +

{(h− αD)∆ + (m + p1)(∆−K) if K < ∆,(h− αD)∆ if K ≥ ∆.

However the (h + p1 − αD)dsn part of the expression above will be accounted

for the n + 1st period’s cost difference. The reasoning is as follows: The next

period (n+1) is an N. As there is no downward substitution in period n+1, the

cycle ends in period n+2, which is an E. For the cost difference between F and

N in period n + 1 we refer to part i or ii of the proof as there is no downward

substitution. Therefore for correct cost accounting (h+ p1−αD)dsn is included

in cost comparison in period n + 1. Thus the claim is proved for J = 2.

The general J = j case is depicted Table 5.1, and the following expression is

the cost difference between N and F for the J = j case:

n+j−2∑

i=n−1

CiN − Ci

F =

n+j−3∑

i=n−1

[(h + p1 − αD)dsi − p1K

i+1 + m(dsi −K i+1)+ + p1(Ki+1 − dsi)+

]

+ (h + p1 − αD)dsn+j−2.

143

Page 153: Essays on Perishable Inventory Management

Period n-1 n n+1 ... n+j-2 n+j-1 n+j

Down. Subs. dsn−1 dsn dsn+1 ... dsn+j−2 dsn+j−1 = 0 dsn+j

X iN −X i

F 0 dsn−1 dsn ... dsn+j−4 dsn+j−2 dsn+j−1 = 0

Type E N N ... N N E

Table 5.1: A sequence of N periods under NIS.

where dsn = min{Kn, Ln}, Ln = S −Dn2 and Kn = Dn

1 − Y nF . Thus,

n+j−2∑

i=n−1

CiN − Ci

F = (h + p1 − αD)dsn+j−2 +n+j−3∑

i=n−1

(h + p1 − αD)dsi − p1Ki+1

+

{m(dsi −K i+1) if Ki+1 < dsi,p1(K

i+1 − dsi+1) if Ki+1 ≥ dsi.

= (h + p1 − αD)dsn+j−2

+n+j−3∑

i=n−1

{(h− αD)dsi + (m + p1)(dsi −K i+1) if K i+1 < dsi,(h− αD)dsi if K i+1 ≥ dsi.

Similar to the J = 2 case, (h + p1 − αD)dsn+j−2 is included for the cost com-

parison in period n + j − 1 referring part i or ii. Thus the claim is proved.

Proof of Lemma 13

Proof : First, as we assume upward substitution is sensible, αU < m+p2. Comparing

F and U is similar to the comparison of F and N as in any given period n Y nN = Y n

U =

Y nF + ∆ for some ∆ ≥ 0, by Proposition 5. Unless there is an upward substitution,

N and U have the same cost, therefore part i (when K ≥ ∆ ) and iii of the proof of

Lemma 12 provide the condition αD < h that is needed for F to be less costly than

U ; and the condition αD > m + h + p1 so that U is less costly than F . We analyze

the remaining cases:

144

Page 154: Essays on Perishable Inventory Management

Upward substitution under both policies: Dn2 = S +L and Dn

1 = Y nF −K leads

to upward substitution under both F and U , where L > 0, Y nF > K > 0.

CnU = (h + p1 − αD)∆ + p2(L− usU) + m(∆ + K − usU) + αU usU

where usU = min{∆ + K, L}.

CnF = p2(L− usF ) + m(K − usF ) + αU usF

where usF = min{K, L}.

CnU − Cn

F = (h + p1 − αD)∆ + m∆− (m + p2 − αU)(usU − usF ),

= (m + h + p1 − αD)(∆− (usU − usF ))

+ (h− αD + αU − (p2 − p1))(usU − usF ). (5.5)

(5.5) is positive if αD < h + αU − (p2 − p1), as min{∆ + K, L} ≥ min{K,L}and ∆ ≥ L−K (when L > K). (5.5) is negative if αD > m + h + p1. Thus the

claim is true for this case.

Upward substitution under U , no substitution under F : In this case Dn2 =

S + L and Dn1 = Y n

F + K with L > 0, ∆ > K > 0.

CnU = (h + p1 − αD)∆ + p2(L− us) + m(∆−K − us) + αU us

where us = min{∆−K, L}.

CnF = p2L + p1K.

CnU − Cn

F = (h + p1 − αD)∆ + m(∆−K − us) + (αU − p2)us− p1K,

= (h− αD)∆ +

{(αU − p2 + p1)(∆−K) if K + L > ∆,(m + p1)(∆−K)− (m + p2 − αU)L if K + L ≤ ∆,

=

(h− αD)K + [h− αD + αU − (p2 − p1)](∆−K) if K + L > ∆,

(h− αD)(∆− L) + [h− αD + αU − (p2 − p1)]L+ (m + p1)(∆−K − L) if K + L ≤ ∆.

= (h− αD)K +{[h− αD + αU − (p2 − p1)](∆−K) if K + L > ∆,(h− αD + m + p1)(∆− L−K) + [h− αD + αU − (p2 − p1)]L if K + L ≤ ∆.

145

Page 155: Essays on Perishable Inventory Management

Thus if αD < min{h, h+αU−(p2−p1)} F is less costly than U . If αD > h+m+p1

U is less costly than F ; as h + m + p1 > max{h, h + αU − (p2 − p1)} when

m + p2 > αU . Thus the claim is proved.

Proof of Lemma 15

Proof : Comparison of D to U is identical to the comparison of D with N except

when U does upward substitution. We analyze that case.

Y nD +∆ = Y n

U by Proposition 5 and Dn2 = S+L, Dn

1 = Y nD +∆−K, (S ≥ Y n

D +∆ >

K > 0 and L > 0) leads to upward substitution under U and no substitution under

D.

CnU = p2(L− us) + m(K − us) + αUus,

where us is the upward substitution amount and is equal to min{K,L}. Then

CnD = p2L + m(K −∆)+ + p1(∆−K)+.

There are four demand scenarios:

Case 1. When K ≤ L and K ≤ ∆; us = K and

CnU = (h + p1 − αD)∆ + p2(L−K) + αUK, Cn

D = p2L + p1(∆−K),

CnU − Cn

D = (h− αD)∆ + (αU − (p2 − p1))K,

= (h− αD)(∆−K) + (h− αD + αU − (p2 − p1))K.

Case 2. When K ≤ L and K > ∆; us = K and

CnU = (h + p1 − αD)∆ + p2(L−K) + αUK, Cn

D = p2L + m(K −∆),

CnU − Cn

D = (m + h + p1 − αD)∆− (m + p2 − αU)K,

= (h− αD + αU − (p2 − p1))∆− (m + p2 − αU)(K −∆).

Case 3. When K > L and K ≤ ∆; us = L and

CnU = (h + p1 − αD)∆ + m(K − L) + αUL, Cn

D = p2L + p1(∆−K),

CnU − Cn

D = (h− αD)∆ + (m + p1)K − (m + p2 − αU)L,

= (h− αD)∆ + (m + p1)(K − L) + (αU − (p2 − p1))L.

146

Page 156: Essays on Perishable Inventory Management

Case 4. When K > L and K > ∆; us = L and

CnU = (h + p1 − αD)∆ + m(K − L) + αUL, Cn

D = p2L + m(K −∆),

CnU − Cn

D = (m + h + p1 − αD)∆− (m + p2 − αU)L.

Thus although the condition αD < min{h, h + αU − (p2 − p1)} is sufficient for D to

be less costly than U in most of the demand scenarios (including no substitution and

downward substitution cases), the analyses of cases 2 and 4 show that there is no

condition that guarantees D to be less costly than U in general, unless αU > m + p2,

which in itself guarantees that upward substitution is an unattractive decision.

However, if αD > h+m+p1, then U is guaranteed to be less costly than D. (Note

αD > m + h + p1 implies αD > h + αU − (p2 − p1) as αU < m + p2.)

Proof of Lemma 14

Proof : Unless there is an upward substitution D and F have the same cost because

Y nF = Y n

D = Y nN − ∆ for some ∆ ≥ 0 by Proposition 5. Therefore parts i (no

substitution) and iii (downward substitution) of the proof of Lemma 12 provide the

conditions αD < h that is required for D to be less costly than N . We also need

to analyze the following case (where there is no substitution): Dn2 = S + L and

Dn1 = Y n

D −K where L > 0 and XnD > K > 0. Then,

CnN = (h + p1 − αD)∆ + p2L + m(∆ + K), Cn

D = p2L + mK,

CnN − Cn

D = (m + h + p1 − αD)∆.

Thus the claim is true.

A.3. Results on Freshness of Inventory

Proof of Proposition 6

Proof : Inside a pair, we define ∆′

= XnN − Xn

F > 0 and ∆ = Xn−1F − Xn−1

N > 0.

To prove part (i), it is enough to observe: By definition of a pair XnF = Xn

N −∆′ ≤

S −∆′< S. But if Dn−1

2 ≥ Xn−1F , then from the inventory recursions under TIS, we

have XnF = S, which is contradicting. Part (ii) follows from the inventory recursions

147

Page 157: Essays on Perishable Inventory Management

and substitution from old items to new items implies XnF = S. For part (iii), assume

Dn−11 = S − Xn−1

N + k where k > 0. Since old item inventory under F in period

n− 1 is S−Xn−1N −∆, there is excess demand of ∆ + k for old items in period n− 1.

There must be some new items to fulfill this shortage by downward substitution due

to part (i). (Otherwise both F and N will be at inventory state (S, 0).) Thus if ds is

the amount of accepted downward substitution, old item inventory in period n will

be [Xn−1N − ds − Dn−1

2 ]+ and [Xn−1N − Dn−1

2 ]+ for F and N respectively. This is a

contradiction since in period n there must be less old items under N (i.e. there must

be more new items under N ) by definition of a pair. Therefore the claim is true.

Proof of Lemma 17

Proof : Equalities follow from inventory recursions (i.e. XnF = Xn

D and XnU = Xn

N for

all n). We show the inequality below.

Suppose there is a pair in periods n − 1 and n. Inside a pair, we define ∆′

=

XnN −Xn

F > 0 and ∆ = Xn−1F −Xn−1

N > 0. There are two possible cases inside a pair

due to Proposition 6.

Case 1. Dn−11 and Dn−1

2 are less than the inventories under policy F ; so no shortage

or substitution takes place under F . Dn−12 = Xn−1

N + ∆ − k (S > k ≥ 0) and

Dn−11 = S −Xn−1

N −∆− l (l ≥ 0). Then,

XnF = S − (Xn−1

N + ∆−Xn−1N −∆− k) = S − k

and XnN = S + ∆− k. Thus ∆

′= ∆.

Case 2. There is downward substitution. Demands are: Dn−12 = Xn−1

N + ∆ − k

(S > k ≥ 0) and Dn−11 = S − Xn−1

N − ∆ + l, but the downward substitution

amount dsn−1 is constrained such that min{∆, l} > dsn−1 > 0. Then

XnF = S−(Xn−1

N +∆−Xn−1N −∆+k−dsn−1)

+ = S−k+dsn−1; XnN = S+∆−k.

Thus ∆′= S + ∆− k − (S − k + dsn−1) = ∆− dsn−1.

In both cases, we observe that ∆′ ≤ ∆. This means on the average there are always

more new items under F , in the pairs. By definition XkF ≥ Xk

N for a period k that is

outside the pairs.

148