0
Perspectives in Supply Chain Risk Management: A Review
Christopher S. Tang
UCLA Anderson School, 110 Westwood Plaza, UCLA, Los Angeles, CA 90095, USA
Tel: (310) 825-4203
Web: www.anderson.ucla.edu/x980.xml
November 3, 2005
Abstract
To gain cost advantage and market share, many firms implemented various initiatives
such as outsourced manufacturing and product variety. These initiatives are effective in
a stable environment, but they could make a supply chain more vulnerable to various
types of disruptions caused by uncertain economic cycles, consumer demands, and
natural and man-made disasters. In this paper, we review various quantitative models for
managing supply chain risks. We also relate various supply chain risk management
strategies examined in the research literature with actual practices. The intent of this
paper is three-fold. First, we develop a unified framework for classifying supply chain
risk management articles. Second, we hope this review can serve as a practical guide for
some researchers to navigate through the sea of research articles in this important area.
Third, by highlighting the gap between theory and practice, we hope to motivate
researchers to develop new models for mitigating supply chain disruptions.
Keywords: Supply Chain Risk Management, Quantitative Models, Review
1
1. Introduction
Over the last 10 years, earthquakes, economic crises, SARS, strikes, terrorist attacks have
disrupted supply chain operations repeatedly. Supply chain disruptions can have
significant impact on a firm’s short-term performance. For example, Ericsson lost 400
million Euros after their supplier’s semiconductor plant caught on fire in 2000, and Apple
lost many customer orders during a supply shortage of DRAM chips after an earthquake
hit Taiwan in 1999. Supply chain disruptions can have long-term negative effects on a
firm’s financial performance as well. For instance, Hendricks and Singhal (2005) report
that companies suffering from supply chain disruptions experienced 33-40% lower stock
returns relative to their industry benchmarks. To mitigate supply chain disruptions
associated with various types of risks (uncertain economic cycles, uncertain consumer
demands, and unpredictable natural and man-made disasters), many researchers have
developed different strategies / models for managing supply chain risks. In this paper, we
review primarily quantitative models that deal with supply chain risks.
1
Also, we relate
various supply chain risk management strategies examined in the literature with actual
practices. The intent of this paper is three-fold. First, we develop a unified framework
for classifying supply chain risk management articles. Second, we hope this review can
serve as a practical guide for some researchers to navigate through the sea of research
articles in this important area. Third, by highlighting the gap between theory and
practice, we hope to motivate researchers to develop new models for mitigating supply
chain disruptions.
We now present a unified framework for classifying supply chain risk management
articles. In preparation, let us define two terms: supply chain management, and supply
chain “risk” management. First, by combining various definitions developed by others
(Christopher (1992), Council of Supply Chain Management Professional
(www.cscmp.org), Ritchie and Brindley (2001), etc.), we define supply chain
management as “the management of material, information and financial flows through a
network of organizations (i.e., suppliers, manufacturers, logistics providers,
wholesalers/distributors, retailers) that aims to produce and deliver products or services
for the consumers. It includes the coordination and collaboration of processes and
activities across different functions such as marketing, sales, production, product design,
procurement, logistics, finance, and information technology within the network of
organizations.” Second, by combining the definitions developed by others (Christopher
(2002) , Deloitte and Touche (www.deloitte.com
) and others), we define supply chain
risk management (SCRM) as “the management of supply chain risks through
coordination or collaboration among the supply chain partners so as to ensure
profitability and continuity.” Based on the definitions of supply chain management and
supply chain risk management, it appears that one can address the issue of Supply Chain
Risk Management along two dimensions:
1
To establish a scope for this paper, we shall not review articles that address different measures for
reducing the risk levels. Instead, we shall limit ourselves to review various approaches for mitigating the
impact of supply chain risks.
2
1. Supply Chain Risk – operational risks or disruption risks;
2. Mitigation Approach – supply management, demand management, product
management, or information management
The first dimension addresses the risk level of certain events. Operational risks are
referred to the inherent uncertainties such as uncertain customer demand, uncertain
supply, and uncertain cost. Disruption risks are referred to the major disruptions caused
by natural and man-made disasters such as earthquakes, floods, hurricanes, terrorist
attacks, etc., or economic crises such as currency evaluation or strikes. In most cases, the
business impact associated disruption risks is much greater than that of the operational
risks.
To mitigate the impact of supply chain risks, Figure 1 depicts four basic approaches
(supply management, demand management, product management, and information
management) that a firm could deploy through a coordinated / collaborative mechanism.
Each of these four basic approaches is intended to improve supply chain operations via
coordination or collaboration as follows. First, a firm can coordinate or collaborate with
upstream partners to ensure efficient supply of materials along the supply chain. Second,
a firm can coordinate or collaborate with downstream partners to influence demand in a
beneficial manner. Third, a firm can modify the product or process design that will make
it is easier to make supply meet demand. Fourth, the supply chain partners can improve
their coordinated or collaborative effort if they can access various types of private
information that is available to individual supply chain partners.
Figure 1. Four Basic Approaches for Managing Supply Chain Risks.
Supply Chain Risks
Supply
Management
Demand
Management
Product
Management
Information
Management
3
In this paper, we shall classify the supply chain risk management articles according to
these four basic approaches. In addition, we shall review the articles in the area of supply
chain management according to the issues highlighted in Table 1.
Supply
Management
Demand
Management
Product
Management
Information
Management
Strategic
Plans
Supply Network
Design
Product
Rollovers and
Product Pricing
Product Variety Supply Chain
Visibility
Tactical
Plans
Supplier
Selection,
Supplier Order
Allocation, and
Supply Contracts.
Shift Demand
Across Time,
Markets, and
Products.
Postponement
and Process
Sequencing.
Information Sharing,
Vendor Managed
Inventory, and
Collaborative
Planning,
Forecasting and
Replenishment.
Table 1. Strategic and Tactical Plans for Managing Supply Chain Risks.
Due to the fact that there are thousands of articles published in the area of supply chain
management, we are unable to review all existing articles in this paper. In addition,
because supply chain management is a multi-disciplinary research area, we feel the need
to include some marketing and management articles as well. As the scope expands, we
apologize for any unintended omission. In any event, this paper is not meant to be an
exhaustive review; however, it is intended to describe some key supply chain risk
management approaches (supply/demand/product/information management) examined by
various researchers recently.
The organization of this paper is as follows. From Section 2 to Section 5, we first review
some of the recent research work that addresses the use of supply / demand / product /
information management strategies for managing supply chain risks (operational or
disruption). In Section 6, we relate the research articles reviewed between Section 2 and
Section 5 with actual practices. Section 7 concludes this paper with some suggestions for
future research in the area of supply chain risk management.
2. Supply Management
To gain cost advantage, many firms outsourced certain non-core functions so as to
maintain a focus on their core competence (c.f., Porter (1985)). Since the 1980s, we
witnessed a sea change in which firms outsourced their supply chain operations including
design, production, logistics, information services, etc. Essentially, supply management
deal with five inter-related issues:
1. Supply network design
4
2. Supplier relationship
3. Supplier selection process (criteria and supplier selection)
4. Supplier order allocation
5. Supply Contract
2.1. Supply Network Design
When designing a global supply chain network, one needs to examine the following
issues:
1. Network configuration: which available suppliers, manufacturing facilities,
distribution centers, and warehouses should be selected.
2. Product assignment: which facilities (suppliers, manufacturing facilities,
distribution centers, etc.) should be responsible for processing which
subassemblies, semi-finished products, or finished products.
3. Customer assignment: which facility at an upstream stage should be responsible
for handling the “demand” generated from downstream stages.
4. Production planning: when and how much should each facility produce or
process.
5. Transportation planning: when and which mode of transportation should be used.
Most of the published work in the area of supply network design is based on various
deterministic models. For example, by considering the fixed and variable processing
cost at each facility, Arntzen et al. (1995) implemented a mixed integer programming
model at Digital Equipment Corporation that serves as a planning system for determining
optimal decisions related to issues #1, #2, #4, and #5. In addition, Camm et al. (1997)
develop an integer programming model for Procotor and Gamble that deals with issues
#1 and #3. However, these papers do not deal with the issue of supply chain risks in an
explicit manner.
More recently, some researchers investigate supply chain network design by capturing
certain risk issues arising from global manufacturing. First, Levy (1995) presents a
simulation model to examine the impact of demand uncertainty and supplier reliability on
the performance of different supply chain network designs (issues # 1 and #5). This
simulation model has helped a personal computer manufacturer to evaluate the costs and
lead times associated with two sourcing alternatives between Singapore and California.
Next, in the context of outsourced manufacturing, there are two common approaches for
the contract manufacturers to obtain the requisite parts from various suppliers. The first
arrangement is called “consignment” under which the manufacturer would first purchase
the requisite parts from different suppliers (so as to enjoy the volume discount), sort the
parts to form different kits, and then ship the kits to the corresponding contract
manufacturers. However, this arrangement has drawbacks in terms of longer lead time
and higher (shipping and handling) cost. An alternative arrangement is called “turnkey”
under which the contract manufacturer would first order the parts directly from
designated suppliers and then charge the contract manufacturer accordingly. Lee and
5
Tang (1998) develop a stochastic inventory model to examine the tradeoff between the
consignment and turnkey arrangements under demand uncertainty. Their analysis has
helped Hewlett Packard to determine specific arrangements with different contract
manufacturers in Singapore and Malaysia.
There are few papers that examine the supply chain network design under uncertain
exchange rates. First, Huchzermeier and Cochen (1996) develop a modeling framework
to show how one can exploit currency exchange rates by shifting production within a
global supply chain network. By incorporating issues #1, #3, #4 and #5, they formulate
the problem as a multi-period stochastic programming problem that aims to maximize the
discounted after-tax profit. They also show how flexible global supply chain can provide
real options to hedge against exchange rate fluctuations. Second, Kouvelis and
Rosenblatt (2002) develop a model for designing a 2-stage global supply chain network
model that addresses all 5 issues listed above. More importantly, they consider the case
in which government subsides and tax incentives are present in certain countries for
certain products or operations. They present a mixed integer programming problem
formulation and they provide analysis that generates insights on the effects of financing,
taxation, regional trading zones and local content rules on the design of a global supply
chain.
2.2. Supplier Relationship
As manufacturers recognize the strategic value of suppliers in the late 1980s, Helper
(1991) reports that the supplier relationship has changed dramatically from adversarial to
cooperative in the U.S. Specifically, many firms realized that suppliers can enable a firm
to focus on their own core competence and to reduce cost, reduce product development
cycle time, increase product quality at the same time. In addition, various e-markets and
information technologies enable firm to foster different types of relationships with the
suppliers, ranging from one-time purchase to virtual integration via information sharing.
Dyer and Ouchi (1993) and Dyer (1996) study various Japanese and U.S. firms and
support the idea of having long-term supplier relationship with fewer strategic suppliers.
Tang (1999) identifies four types of supplier relationships: vendor, preferred supplier,
exclusive supplier and partner. These four types differ from each other in terms of types
of contracts, length of contracts, type of information exchange, pricing scheme, delivery
schedule, etc. By considering the market condition that is measured in terms of the
strategic importance level of the part to the buyer and the buyer’s bargaining power, Tang
prescribes different supplier relationship for different market conditions.
Most of the literature reviewed in Tang (1999) focus on qualitative analysis or strategic
analysis. Cohen and Agrawal (1999) is the first to develop an analytical model for
evaluating the tradeoff between the flexibility offered by short-term contracts and the
improvement opportunities and price certainty associated with long term contracts. They
show analytically that long-term contracts may not always be optimal, and they provide
conditions under which short-term contracts are optimal. As firms expand their business
globally, their supply chains would involve more global partners. For different regional
6
markets, a firm may source locally so as to reduce transportation cost (due to local labor
cost or tax benefits), reduce replenishment lead times, reduce inventory. Consequently, it
is quite common for firms to source from multiple suppliers. In addition, some firms
may source from multiple suppliers so as to reduce the impact of various operational and
disruption risks. According to an empirical study conducted by Shin et al. (2000), dual or
multiple sourcing is a common business practice.
2.3. Supplier Selection Process
Boer et al. (2001) provide a comprehensive review of different methods for selecting
suppliers. They divide the supplier selection process into 3 stages, namely, formation of
selection criteria, determination of approved suppliers, and final supplier selection.
2.3.1. Supplier Selection Criteria
To form selection criteria, Boer et al. (2001) reported two decision methods (interpretive
structural modeling and expert system) for forming selection criteria. The interpretive
structural modeling technique proposed by Mandal and Deshmukh (1996) is intended to
separate dependent criteria from independent criteria, where the independent criteria are
important for screening acceptable suppliers and the dependent criteria are critical for
final supplier selection. The expert system developed by Vokurka et al. (1996) captures
the previous supplier selection process in a knowledge base, which can be used to suggest
selection criteria for future supplier selection process. In an empirical study, Choi and
Hartley (1996) investigate 26 supplier selection criteria used by different partners
(automotive assemblers, first-tier suppliers, second-tier suppliers) across the supply chain
in the auto industry. These criteria include cost reduction capability, quality
improvement capability, and the ability to change production volumes rapidly. By using
various multivariate statistical techniques (factor analysis, clustering analysis,
multivariate analysis of variance) to analyze the supplier selection criteria reported in 156
surveys, they make the following conclusions:
The supplier selection criteria are reasonably consistent across the supply chain in
the automotive industry. At all levels, commitment to establish cooperative/long-
term relationship is an important selection criterion.
Price is one of the least important criteria, while quality and delivery are
important criteria.
Supplier’s technological capability and financial stability are more important
criteria for the auto assemblers.
It is interesting to note that the criterion regarding the ability to change production
volumes rapidly is not considered to be as important as other criteria such as quality and
long term relationship. Given the recent disruptions (terrorist attacks, hurricanes,
earthquakes, SARS, etc.), one may speculate that the issue of business continuity would
become an important supplier selection criterion.
7
2.3.2. Supplier Approval / Selection
At this stage of the process, the goal is to reduce the set of all potential suppliers to a
smaller set of approved suppliers. To do so, the decision maker has to sort / classify all
suppliers into 2 categories: approved or disapproved. Based on the supplier’s
performance on the selection criteria, Boer et al. (2001) report the following methods for
determining a set of approved suppliers: clustering analysis, data envelopment analysis,
and an Artificial Intelligence approach called case-based-reasoning method.
When making the final supplier selection, Boer et al. (2001) report the following decision
methods in different settings:
Linear weighting models. By assigning different weights to different criteria, one
can compute the overall rating of a supplier by considering the weighted sum of
different criteria. In this case, the supplier with the highest rating will be selected.
Total cost of ownership. This method is developed by Ellram (1990) that is
intended to include all quantifiable costs incurred throughout the life cycle of the
item purchased from a supplier. The supplier with the lowest total cost of
ownership will be selected.
Mathematical programming models. Most of the methods reported in Boer et al.
(2001) are deterministic models: linear programming, goal programming, data
envelopment analysis, etc. The idea is to select supplier(s) with minimum cost.
Simulation models. This method enables the decision maker to capture some of
the uncertainties (yield loss, stochastic lead times, etc.) related to supplier
selection. By simulating the performance of different suppliers for different
criteria under different scenarios, the method can help a decision maker to select a
supplier under uncertainty.
While Boer et al. report different approaches for managing the supplier selection process,
there are some quantitative models for supplier selection. For example, Weber and
Current (1993) present a mixed integer programming formulation that is intended to
capture multiple supplier selection criteria. Current and Weber (1994) formulate the
supplier selection problem as a variant of facility location problem. More recently,
Weber et al. (2000) present an approach for evaluating the number of suppliers to employ
by using multi-objective programming and data envelopment analysis. Dahel (2003)
extends the model presented in Weber et al. (2000) by incorporating the order quantity
decision for each supplier.
While most of the supplier selection models are deterministic models, there are few
articles that deal with the supplier selection process under operational risks. Tang (1988)
presents a supplier selection model that captures the interaction of the supplier’s quality
and the buyer’s quality control (inspection policy). By considering the interaction
between the supplier’s quality and the buyer’s internal manufacturing process, Tagaras
and Lee (1996) develop a different supplier selection model that captures different
degrees of imperfections in the buyer’s manufacturing processes. Specifically, they
consider that there are two states of the buyer’s process: normal or abnormal. When the
8
buyer’s process is in the normal state, the output of the process is perfect if the supplier’s
input is. However, when the buyer’s process is in the abnormal state, the output of the
process is defective regardless of the supplier’s input is or is not. By considering
different costs associated with quality of the output, Tagaras and Lee (1996) develop the
optimal supplier selection criterion that minimizes the buyer’s expected total cost
(ordering cost and cost of quality). More recently, Kouvelis (1998) presents a supplier
selection model that captures the stochastic nature of exchange rate. In his model, the
buyer needs to decide which suppliers to select and the quantity to be sourced from each
selected supplier. As a way to respond to fluctuating exchange rates, the model captures
the flexibility for the buyer to shift the order quantity among suppliers dynamically at the
expense of switch-over costs. When the switch-over cost is significantly high, he shows
that the buyer may continue to source from suppliers that are more expensive so as to
reduce unnecessary switch-over costs.
2.4. Supplier Order Allocation
After a set of suppliers is chosen, the buyer needs to determine ways to allocate the order
quantity among these selected suppliers. We shall classify the work in this area
according to different types of operational risks:
Uncertain demands
Uncertain supply yields
Uncertain supply lead times
Uncertain supply costs
2.4.1. Uncertain Demand
There is voluminous amount of published works that focus on analytical models for
determining optimal order quantity for a single supplier under demand uncertainty. The
reader is referred to some recent books Porteus (2002) and Zipkin (2000) for a
comprehensive review of various analytical models.
For the case of multiple suppliers, Minner (2003) provides a comprehensive literature
review. When the supply lead times are deterministic, all models assume that the
supplier with a shorter lead time charges a lower cost per unit. Due to the complexity of
the analysis, most discrete time models are restricted to two suppliers with lead times
differed by one period. Zhang (1996) is the only paper that deals with three suppliers
with lead times differ by one and two periods and characterize the optimal ordering
policy for each supplier.
To make the analysis of multiple-supplier inventory models more tractable, some
researchers consider two supply modes: regular and emergency. The regular supply
model is based on a regular supply lead time, while the emergency supply is available
instantaneously. Fukuda (1964) shows that the optimal ordering policy takes on the form
9
of “two order-up-to levels” x and y, where x < y. Specifically, the optimal ordering
policy can be described as follows: If the inventory at the beginning of a period z is less
than x, then order (x – z) units by using the emergency mode and order (y – x) units by
using the regular mode; if x < z < y, then order (y – z) units according to the regular
mode; otherwise, order nothing. Vlachos and Tagaras (2001) extend Fukuda’s model to
the case in which the emergency model is capacitated. More recently, Scheller-Wolf and
Tayur (1999) consider a Markovian periodic review inventory model and show that the
optimal ordering policy for the buyer is a modified state-dependent base-stock policy.
Specifically, they show that there exists a state-dependent optimal inventory level (target)
in each period. In each period, the buyer should first order an amount from the regular
supplier so that the inventory position after ordering is as close as possible to the target.
The buyer can place an emergency order to fill the gap between the target and the
inventory position after ordering from the regular supplier.
Due to the complex analysis of the optimal ordering policies for the multi-supplier case,
various researchers restrict their analysis to certain classes of ordering policies. For
example, Moinzadeh and Nahmias (1988) analyze an (s
1
, s
2
, Q
1
, Q
2
) ordering policy for a
continuous time model with regular and emergency supply. Specifically, when the
inventory reaches s
1
, a regular order of size Q
1
is placed. If the inventory reaches s
2
within the lead time of the regular order, an emergency order Q
2
is placed. Janssens and
de Kok (1999) analyze an ordering policy in which the buyer will always order Q units
from one supplier in each period, and will order [S – Q]+ units from the second supplier
so as to bring the inventory position to S. The reader is referred to Minner (2003) for
more details.
Instead of focusing on optimal ordering policies, Nagurney et al. (2005) develop a model
for analyzing the equilibrium behavior of a three-level supply chain consisting of
manufacturers, distributors and retailers. By considering uncertain demands at the
retailer level, they formulate the problem at each level as a non-linear programming
problem. For the retailers, the goal is to determine the optimal order quantity for each
retailer based on the wholesale price determined by the distributors. However, for the
distributors, the goal is to determine the optimal wholesale price based on the
manufacturer’s price. By considering the first order conditions of these three inter-
related problems, they show how to recast the first order conditions as a set of variational
inequalities. The reader is referred to Bazarra et al. (1993) for more details about the
relationship between variational inequalities and Nash equilibrium. By exploiting the
structure of these variational inequalities, they establish the existence of an equilibrium
and provide certain characteristics of the equilibrium.
2.4.2. Uncertain Supply Yields
Let us consider some single-stage-multiple-period models. When a buyer receives a
random fraction of the order quantity from the supplier, Gerchak, Vickson and Parlar
(1988) analyze a finite horizon problem with stationary demand distribution and show
that order-up-to policies are not optimal. Henig and Gerchak (1990) further show that
10
there exists a critical point for each period such that an order should be placed only when
the on-hand inventory at the beginning of the period is below the corresponding critical
point. However, the exact order quantity is a complicated function of the system
parameters. More recently, Agrawal and Nahmias (1998) present a model for evaluating
the tradeoff between the fixed costs associated with each selected supplier and the costs
associated with yield loss. They show how to determine the optimal number of suppliers
with different yields when the demand is known. To limit our focus to supply chain
management, we shall highlight some of the models that deal with multiple
stages/products. The reader is referred to Yano and Lee (1993) for a thorough review of
single stage/period models that deal with lot sizing models with random yields.
First, let us consider some multiple-stage-multiple-period models. Bassok and Akella
(1991) consider a two-stage-multiple-period model in which one stage corresponds to
raw material ordering and the second stage corresponds to actual production, where yield
uncertainty occurs only at the material ordering stage. They show that the existence of
two critical points (one for the raw material ordering stage and one for the production
stage) so that the optimal ordering quantity and the optimal production quantity would
depend upon whether the sum of (on-hand) finished goods and raw materials is larger or
smaller than these two critical points, respectively. Due to the fact that exact analysis of
multiple-stage-multiple period models is intractable, Tang (1990) restricts his analysis of
a linear control rule for a multi-stage serial production line with uncertain yields at each
stage and uncertain demand. This linear control rule intends to restore the buffer stock at
each stage to its target value in expectation. Hence, this control rule minimizes the
expected deviation of the buffer stock levels from their targets. Denardo and Lee (1996)
generalize Tang’s model by incorporating rework and unreliable machines.
Next, since the analyses of multiple-product-multiple-stage-multiple-period models are
intractable, not much work has been done in this area. Akella, Rajagopalan and Singh
(1992) study a multi-stage facility with rework that produces multiple parts. Their
analysis aims to determine an optimal production rule at each stage that minimizes the
total inventory and backorder cost. They assume that the cost function is quadratic,
which leads to optimal linear decision rules. Linear decision rules have been analyzed by
Gong and Matsuo (1997) as well. Specifically, Gong and Matsuo consider a more
general multi-stage facility with re-entrant routings. They formulate a control problem
with the objective to minimizing the weighted variance of work-in-process inventory
while ensuring that production capacity constraints are satisfied with a pre-specified
probability. Their numerical experiments suggest that the linear decision rules perform
well when compared with the optimal production policy.
2.4.3. Uncertain Lead Times
When replenishment lead times are stochastic, most researchers restrict their analyses of
multiple supplier models to the case of deterministic demand. When both suppliers have
identical lead time distributions (uniform or exponential), Ramasesh et al. (1991)
consider an (s, Q) ordering policy where the order quantity Q is evenly split between two
11
suppliers. Due to the complexity of the analysis, the optimal values for the reorder point
s and the order quantity Q are determined numerically. By restricting the attention to
the (s, Q) ordering policy, Sedarage et al. (1999) extends the model developed by
Ramasesh et. al (1991) by considering n > 2 suppliers and non-identical split among
suppliers. Based on the numerical analysis presented in Sedarage et al. (1999), they
show that it might be beneficial to order from some suppliers with poor lead time
performance (in terms of the mean and standard deviation of the lead time). In general,
the exact analysis of multiple suppliers with stochastic lead times is intractable.
However, exact analysis can be obtained for some special cases. For example, Anupindi
and Akella (1993) consider a two-supplier model with random demand in which the
replenishment lead time of supplier j is equal to one period with probability p
j
and two
periods with probability (1 – p
j
), where j = 1, 2. They derive the optimal ordering
policy that minimizes the total ordering, holding and backordering costs over a finite
horizon. They show that the optimal ordering policy in each period n depends on two
critical points x
n
and y
n
, where x
n
< y
n
, and the on-hand inventory at the beginning
period n, z
n
. Specifically, order nothing if z
n
y
n
; order from one supplier if x
n
z
n
< y
n
; and order from both suppliers if z
n
< x
n
.
2.4.4. Uncertain Supply Capacity
Most models assume that the supply capacity is unlimited or known. However,
unexpected machine breakdowns could affect the supply capacity. Relative little amount
of work has been done in the area of uncertain supply capacity. Parlar and Perry (1996)
present a continuous time model in which the availability of each of the n suppliers is
uncertain because of disruptions like equipment breakdowns, labor strikes, etc. By
considering the case that each supplier is either “on” or “off”, there are 2
n
possible
number of states for the whole system. For each of these 2
n
states, they analyze a state-
specific (s, Q) ordering policy so that the buyer would order Q units when the on-hand
inventory reaches s. Ciarallo et al. (1994) develop a discrete time model in which the
supply capacity is random with known probability distribution. By considering the total
(undiscounted) expected costs (ordering, inventory holding and backordering costs), they
show that the objective function is quasi-convex, which implies that an order-up-to policy
is optimal. More recently, Wang and Gerchak (1996) examine a periodic review model
with uncertain supply capacity, uncertain yields and uncertain demand. The objective is
to minimize the total discounted expected costs over a finite horizon. They show that the
optimal policy possesses the same structure as the optimal policy obtained by Hernig and
Gerchak (1990) for the case in which only random yield is considered. Specifically,
Wang and Gerchak show that the order-up-to policy is optimal.
2.4.5. Uncertain Supply Cost
While most work focus on demand uncertainty, not much work has been done in the area
of uncertain supply cost. For models that examine the issue of uncertain supply cost
imposed by an upstream supply chain partner, Gurnani and Tang (1999) analyze a
12
situation in which a retailer has two instants to order a seasonal product from a
wholesaler prior to the beginning of a single selling season. They consider the case in
which the wholesale price at the second instant and the demand are uncertain; however,
the retailer can improve the demand forecast by using market signals observed between
the first and second instants. In order to determine the profit-maximizing ordering
policy, the retailer needs to evaluate the trade-off between the benefit of having a more
accurate forecast and a potentially higher wholesale price at the second instant. By
formulating the problem as a 2-period dynamic programming program, they develop an
optimal way to allocate the optimal order quantity to be placed at the first and second
instants and they provide the conditions under which the retailer should delay his
ordering decision until the second instant.
Besides the work of Gurnani and Tang (1999), various researchers develop models for
exploiting uncertain currency exchange rates in a global supply chain. Kogut (1985)
develops a framework to argue that the benefit of a global supply chain lies in the
operational flexibility, which permits a firm to exploit uncertain exchange rates. To
examine this issue in a quantitative manner, Kogut and Kulatilaka (1994) develop a
stochastic model to examine the value of the flexibility to shift production between two
plants located in two different countries. By formulating the problem as a T-period
dynamic programming problem and by modeling the exchange rate process as a discrete-
time mean reverting stochastic process, they determine the option value of maintaining
two manufacturing locations with excess capacity instead of having a single
manufacturing location. However, they assume that the capacity of each plant is
unlimited so that exactly one plant will be used to produce the required quantity to meet
the total demand in each period. Dasu and Li (1997) generalize Kogut and Kulatilaka’s
model by considering the case in which both plants have limited capacity so that both
plants will be used to meet the demand in each period. They formulate the problem as an
infinite horizon dynamic program with discounting. When the production cost is concave
and when the cost of production shifting is linear, they show that the optimal production
shifting policy is a two-barrier policy if the exchange rate process satisfies certain
conditions. Specifically, under the two-barrier policy, there exists two critical points a
and b so that it is optimal to shift the production between two manufacturing locations
when the exchange rate is below a or above b. If the exchange rate is between a and b,
then it is optimal to keep the same production quantity at each location without any
shifting so as to reduce any unnecessary switch-over cost.
The analysis presented in Kogut and Kulatilaka (1994) and Dasu and Li (1997) would
become intractable for more than two countries. To address global manufacturing issues
such as supply chain network design for n > 2 countries, Huchzermeier and Cohen (1996)
present a stochastic dynamic programming problem for evaluating different global
manufacturing strategy options. For any given exchange rate in each period, they solve a
mixed integer program to determine the optimal production and distribution plan for the
entire supply chain network that maximizes the global, after-tax profit. They construct
various numerical examples by considering 16 different supply chain network designs,
each of which specifies the location of the supplier(s), production plant(s), and market(s).
Through these numerical examples, they illustrate the value of a global supply chain
13
network that enables firms to shift its production and distribution plan swiftly as the
exchange rates fluctuate.
2.5. Supply Contracts
When the partners across a supply chain belong to different firms or divisions, they tend
to focus on maximizing their own objectives and make their decisions independently.
Consequently, locally optimal decisions can cause operational inefficiency and globally
suboptimal decision for the entire supply chain. There are two studies highlighting the
pitfalls of an disintegrated supply chain. First, when each supply chain partner places
their order independently for the case in which the customer demand follows an AR(1)
process, Lee et al. (1997) show this locally optimal ordering decisions will create the
“bullwhip” effect that causes operational inefficiency. Second, when each supply chain
partner makes their ordering decision by maximizing their own profit for the case and
when the customer demand is a deterministic and decreasing function of retail price,
Bresnahan and Reiss (1985) show that these locally optimal decisions would result in
lower total profit for the entire supply chain. To improve operational efficiency and/or
supply chain coordination, there is a growing research interest in supply chain contract
analysis recently. Most supply contract models usually deal with a supply chain that
consists of one manufacturer (supplier) and one retailer (buyer) who faces customer
demand. Even though the economics literature in the area of supply contracts is
voluminous, economics researchers usually assume that that the customer demand is
either deterministic or stochastic in the sense that demand uncertainty is resolved before
the buyer places his order. The reader is referred to Tirole (1988) for a comprehensive
review of supply contracts literature in economics.
There are three excellent reviews of supply chain contract analysis prepared by Cachon
(2003), Lariviere (1998) and Tsay et al. (1998). These three reviews offer different
perspectives in the following sense: Tsay et al. (1998) provide a qualitative overview of
various types of contracts when the demand is deterministic and random; Lariviere
(1998) shows quantitative analyses of different types of contracts when the demand is
uncertain; and Cachon (2003) examines how supply contracts can be used to achieve
channel coordination in the sense that each supply chain partner’s objective becomes
aligned with the supply chain’s objective. Since our focus is on supply chain risk
management, we shall focus on a limited set of supply chain contract literature that deals
with various types of uncertainties. For this reason, we shall classify the supply chain
contract literature according to different risk elements and contract types. Specifically,
we shall review different types of supply contracts that can be characterized according to
the financial flow and material flow as depicted in Figure 2.
14
2.5.1. Uncertain Demand
Wholesale Price Contracts
Consider the following scenario: the retail price p is fixed, the retailer retains the
revenue p and retains the possession of any excess stock that can be salvaged at a price
s. Suppose that the manufacturer offers a per unit wholesale price w so that the fixed
cost F(Q) = 0, and the variable wholesale price w(Q) = w. In a single period setting, it
is optimal for the retailer to order according to the newsvendor solution based on the
corresponding cost structure. Given the retailer’s order quantity, the manufacturer needs
to determine the optimal w that maximizes her net profit. Lariviere and Porteus (2001)
show that the manufacturer’s profit function is unimodal when the customer demand
distribution F(x) with density function f(x) has an increasing generalized failure rate
(IGFR); i.e., when xf(x)/(1- F(x)) is increasing in x. Many distributions such as normal,
exponential, truncated Normal, Gamma, and Weibull are IGFR. Hence, when the
Manufacturer
(internal
marginal cost
c
m
)
Retailer
(internal
marginal cost
c
r
)
Wholesale
price
(F(Q), w(Q))
Retail
price p
Buyback
(or return
credit b)
Manufacturer
Retailer
Retailer’s order quantity Q,
where Q may need to satisfy
certain contractual terms
Customer
Demand D(p)
Buyback (or
return limit
R [Q-D]+ )
Figure 2. Financial flow and material flow under different supply chain contracts
Shared revenue α p
15
demand distribution is IGFR, one can determine the optimal wholesale price by
considering the first order condition. However, Lariviere and Porteus show that a simple
price contract w will not achieve channel coordination. Anupindi and Bassok (1999)
extend Lariviere and Porteus’ single period model to the case in which the retailer faces
an infinite succession of identical selling seasons so that it is optimal for the retailer to
order up to the newsvendor solution at the beginning of each season. Cachon (2002)
generalizes Lariviere and Porteus’ single period model to a two-period model with
inventory holding and demand updating. Specifically, in Cachon’s model, the retailer
can place two separate orders at two separate instants before the selling season starts;
however, the wholesale price at the second instant is known to be higher. Notice that
Cachon’s model reduces to Lariviere and Porteus’ model when the second order is not
allowed. By having the flexibility to place two separate orders, Cachon develop
conditions under which channel coordination is achieved.
Next, there are many situations in which a supply chain partner would keep his
information private. Corbett and de Groote (2000) consider a situation in which the
manufacturer does not know the retailer’s holding cost in a deterministic EOQ-type
environment. By imposing a prior distribution on the retailer’s holding cost, Corbett and
de Groote compare various channel coordination schemes in which F(Q) and w(Q) take
on different functional forms. When the demand is deterministic and decreasing linearly
in the retail price, Corbett and Tang (1998) examine the case in which the manufacturer
does not know the retailer’s internal marginal cost c
r
. By imposing a prior distribution
F(x) on the retailer’s internal marginal cost c
r
and by assuming that the prior distribution
F(x) has increasing failure rate; i.e., f(x)/(1 – F(x)) is increasing in x, they compare the
retailer’s and the manufacturer’s profits under different scenarios: one-part linear
contracts (F(Q) = 0, w(Q) = w), two-part linear contracts (F(Q) = F 0, w(Q) = w), and
two-part nonlinear contracts (F(Q) 0, w(Q) 0). Corbett and Tang study the optimal
behavior of each party under different scenarios. Ha (2001) generalizes Corbett and
Tang’s model by analyzing two-part non-linear wholesale price contracts for the case
when the demand is stochastic and price-sensitive. Ha shows that channel coordination
is not achievable under asymmetric information. When the manufacturer does not know
the retailer’s fixed ordering cost or the backorder penalty cost, Corbett (2001) examines
the benefit of having the manufacturer to own the retailer’s inventory (i.e., consignment
stock). He shows that consigning stock may not always help the manufacturer.
More recently, Babich et al. (2004) is the first to analyze supply contracts with supplier
default risk. They consider a single product model in which competing risky suppliers
compete for business with a retailer. In their model, the suppliers are leaders in a
Stackelberg game so that the suppliers would first establish the unit wholesale prices.
Then the retailer would determine the order quantity for each supplier by taking demand
uncertainty and supplier default uncertainty into consideration. By considering the
retailer’s discounted expected profit, they show that it is optimal for the competing
suppliers to increase their wholesale prices at the equilibrium when the supplier default
correlations are low and it is optimal for the retailer to order from suppliers with highly
correlated default rates.
16
Buy Back Contracts
In a single-period setting, it is optimal for the retailer to order according to the
newsvendor solution. To induce the retailer to order more, it is quite common for the
manufacturer to offer a return policy (also known as buy back contracts) so that the
manufacturer would “buy back” up to R% of the retailer’s excess inventory [Q – D]+
units at a unit rate of b, where R 100% and b w. Therefore, a return policy can
be specified by two parameters (R, b). Pasternack (1985) is the first to show that a
policy that allows for unlimited returns at partial credit; i.e., R = 100% and b < w, would
achieve channel coordination. Moreover, Lariviere (1998) analyze the properties of the
manufacturer’s and retailer’s profits for a class of return policies that coordinate the
channel. Emmons and Gilbert (1998) extend Pasternack’s model to the case in which the
retailer determines the order quantity Q as well as the retail price p. By considering a
specific demand distribution of D(p), they show that return policies or buy back
contracts cannot coordinate the channel. However, there exists certain buy back contracts
under which both the manufacturer and retailer can obtain higher profits. Padmanabhan
and Png (1997) consider the case in which two competing retailers facing a linear
demand curve with an uncertain intercept. Under a full returns policy (i.e., b = w), they
show that these retailers would increase their order quantities in a competitive
environment. More recently, Brown, Chou and Tang (2005) examine a multi-product
returns policy in which the retailer can return up to a percentage of the total order
quantities; i.e., the allowable return limits are pooled. By comparing the pooled returns
policy with the non-pooled returns policy (i.e., the allowable return limits are product-
specific), they present conditions under which the retailer would actually order less under
the pooled returns policy.
Revenue Sharing Contracts
In retailing, stocking out a product could have a larger impact on the manufacturer’s
profit because the customer would usually buy a similar product from the retailer. This
motivates manufacturer to provide incentive for the retailer to stock more. Clearly, a
buy back contract (or a return policy) can serve this purpose; however, the buy back
contract may not be practical in certain situations. For example, in the video rental
industry, it is not practical for the video rental stores to return excess inventory of old
DVDs to the manufacturer (distributor). This may have triggered the idea for the
manufacturer to develop a risk sharing scheme in the form of a revenue sharing contract.
The revenue sharing contract can be characterized by the wholesale price w and the
portion of the revenue to be shared α. As depicted in Figure 2, the retailer would get a
lower wholesale price w upfront but the retailer is required to remit α p for each rental
unit to the manufacturer. For instance, as described in Mortimer (2004), Blockbuster
shares probably between 30-45% of their rental revenue in exchange for a reduced
wholesale price probably at $8 instead of $65 for each DVD.
In the economics literature, Dana and Spier (2001) show that revenue sharing contracts
can be used to coordinate the supply chain, and would induce the retailers to reduce their
17
rental prices under competition. Mortimer (2004) conducts statistical analysis based on
the panel data collected at 6,137 video rental stores in the U.S. between 1998 and 2000.
She shows that revenue sharing contracts can enable a retailer to earn more for popular
titles or new releases. More recently, Pasternack (2002) investigates the effect of a
revenue sharing on the optimal order quantity in a newsvendor environment and he
shows analytically that a revenue sharing contract can be used to achieve channel
coordination. In addition, Cachon and Lariviere (2005) show analytically that the
revenue sharing contracts are equivalent to buy back contracts. Tang and Deo (2005)
determine the conditions for w and α under which the retailer will obtain a higher profit
under the revenue sharing scheme.
Quantity Based Contracts: Quality Flexibility and Minimum Order
To achieve operational efficiency under demand uncertainty, a manufacturer would
prefer contracts that would entice retailers to commit their orders in advance while a
retailer would prefer contracts that would allow them to adjust their orders when
necessary. As a compromise, some manufacturers would offer Quantity Flexibility (QF)
contracts to their retailers. A QF contract is specified by three parameters: a wholesale
price w, an upward adjustment parameter u, where 0 u 1, and a downward
adjustment parameter d, where 0 d 1. Consider the case in which a retailer placed
an order x sometime earlier. Suppose the retailer updates his demand forecast and
would like to revise this particular order. Under the QF contract, the retailer can adjust
his order to Q by paying w per unit as long as (1 – d)x Q (1+u)x. Notice that the
QF contract can be recast as a buy back contract under which the retailer had to buy
(1+u)x units up front but could return or cancel his commitment down to (1-d)x for a
full refund of the wholesale price w. Lariviere (1998) analyzes a QF contract with
parameters w, d, and u that coordinates the channel in a single-period setting. Tsay and
Lovejoy (1999) provide a detailed analysis of QF contract in a multi-period setting.
When it is very costly for a manufacturer to obtain more production capacity, a
manufacturer may develop a supply contract to entice each retailer to commit to a
minimum quantity in advance. Anupindi and Akella (1997) consider the case in which
the retailer is committed to a fixed quantity in each period. In return, the manufacturer
offers a discount based on the level of this fixed commitment. They prove that a
modified order up to policy is an optimal policy for the retailer. Anupindi (1993)
examines the case in which the order quantity that a retailer can place in each period is
bounded pre-specified lower and upper limits. More recently, Bassok and Anupindi
(1997) consider the case in which the retailer is committed to order at least KN units in
total over N periods. When demands are independent, identically distributed random
variables, they prove that the retailer’s optimal order policy in each period is a modified
order up to policy. Instead of focusing on the minimum total commitment for each
product, Anupindi and Bassok (1998) analyze a multi-product supply contract under
which the retailer is committed to a minimum total “dollar” value of the products to be
purchased over N periods.
18
When selling seasonal goods, Fisher (1997) and Fisher and Raman (1996) confirm that
the early sales data has informational value in the sense that they can help the retailer to
obtain more accurate forecast about the total sales for the whole season. Fisher and
Raman show that it is advantageous for the retailer to place a second order after
observing the first few weeks of sales data. To ensure that the second order will be
replenish within the selling season, the manufacturer needs to impose certain restrictions
on the second order quantity. Eppen and Iyer (1997) analyze a “backup agreement” that
has been used in the fashion apparel industry. The backup agreement can be
characterized by three parameters (
β
, w, k). Prior to the selling season, the retailer
commits to Q units for the entire selling season and confirms the first order (1 –
β
)Q at
wholesale price w. The retailer can place a second order up to the remaining
β
Q units
(i.e., the backup units) at wholesale price w and receive quick delivery. There is a
penalty cost of k for any of the backup units not purchased. Brown and Lee (1997)
consider a variant of the backup agreement arising from the semiconductor
manufacturing industry.
2.5.2. Uncertain Price
While most work focus on demand uncertainty, not much work has been done in the area
of uncertain wholesale price. Li and Kouvelis (1999) consider a case in which the
wholesale price is a geometric Brownian motion with drift. Facing with uncertain
wholesale price, the retailer is required to procure exactly D units by time T, where D is
the ultimate demand at time T. Also, an inventory holding cost h(T – t) will be incurred
for each unit purchased at time t, where 0 < t < T. Li and Kouvelis evaluate the cost
associated with three different supply contracts. First, in a “time inflexible contract,” the
retailer must state up front about the purchase time. In a “time flexible contract,” the
retailer may observe price movements and decide dynamically when to buy. They
extend their model to the case in which they can procure the item from two
manufacturers.
For ease of reference, we shall provide a summary of the articles reviewed in each
section. For example, Table 2 provides a summary of the articles reviewed in Section 2
19
Supply Management
Issue
Sub-category References (in the order of appearance)
Supply Network
Design
General Porter (1985), Arntzen et al. (1995), Camm et al. (1997), Levy
(1995), Lee and Tang (1998), Huchzermeier and Cochen (1996),
Kouvelis and Rosenblatt (2002),
Supplier Relationship
General Helper (1991), Dyer and Ouchi (1993), Dyder (1996), Shin et al.
(2000), Tang (1999), Cohen and Agrawal (1999).
Supplier Selection
Process
Supplier Selection
Criteria
Boer et al. (2001), Mandal and Deshmukh (1996), Vokurka
(1996), Choi and Hartley (1996), Ellram (1994)
Supplier
Approval/Selection
Boer et al. (2001), Ellram (1994), Weber and Current (1993),
Weber (2000), Dahel (2003), Tang (1988), Tagaras and Lee
(1996), Kouvelis (1998)
Supply Order
Allocation
Uncertain Demand Porteus (2002), Zipkin (2000), Minner (2003), Zhang (1996),
Fukuda (1964), Vlachos and Tagaras (2001), Scheller-Wolf and
Tayur (1999), Moinzadeh and Nahmias (1988), Janssens and de
Kok (1999), Nagurney (2005), Bazarra et al. (1993),
Uncertain Supply Yields Gerchak, Vickson, and Parlar (1988), Gerchak (1990), Agrawal
and Nahmias (1998), Yano and Lee (1995), Bassok and Akella
(1991), Tang (1990), Denardo and Lee (1996), Rajagopalan and
Singh (1992), Gong and Matsuo (1997)
Uncertain Supply Lead
Times
Ramasesh et al. (1991), Sedarage et al. (1999), Akella et al.
(1993)
Uncertain Supply
Capacity
Parlar and Perry (1996), Ciarallo (1994), Wang and Gerchak
(1996), Hernig and Gerchak (1990)
Uncertain Supply Cost Gurnani and Tang (1999), Kogut (1985), Kogut and Kulatilaka
(1994), Li (1997), Kogut and Kulatilaka (1994), Dasu and Li
(1997), Huchzermeier and Cohen (1996)
Supply Contracts
General Lee et al. (1997), Bresnahan and Reiss (1985), Cachon (2003),
Lariviere (1998), Tsay (1998)
(Uncertain Demand)
Wholesale Price
Contracts
Lariviere and Porteus (2001), Anupindi and Bassok (1999),
Cachon (2002), Corbett and de Groote (2000), Corbett and Tang
(1998), Ha (2001), Corbett (2001), Babich et al. (2004)
Buy Back Contracts Lariviere (1998), Emmons and Gilbert (1998), Padmanabhan and
Png (1997), Brown, Chou, and Tang (2005),
Revenue Sharing
Contracts
Dana and Spier (2001), Mortimer (2004), Pasternack (2002),
Cachon and Lariviere (2005),
Quantity Based
Contracts
Lariviere (1998), Tsay and Lovejoy (1999), Anupundi and Akella
(1997), Anupindi (1993), Bassok and Anupini (1997), Fisher
(1997), Fisher and Raman (1996), Eppen and Iyer (1997), Brown
and Lee (1997)
(Uncertain Price)
General Li and Kouvelis (1999)
Table 2. Summary of Supply Management Articles.
3. Demand Management
In Section 2, we describe how manufacturers can use different supply management
strategies to mitigate various supply chain operational risks. However, these supply
management strategies are ineffective when the underlying supply mechanism is
inflexible. For instance, in the service industry or in the fashion goods manufacturing
industry, the supply mechanism is inflexible because the capacity is usually fixed. When
20
the supply capacity is fixed, many firms have attempted to use different demand
management strategies so that they can manipulate uncertain demands dynamically so
that the modified demand is better matched with the fixed supply. Due to space
limitation, we are unable to review the dynamic pricing or clearance pricing literature.
The reader is referred to Elmaghraby and Keskinocak (2003) for an extensive review of
dynamic pricing models and clearance pricing models for selling a fixed number of units
over a finite horizon. Also, we do not plan to review literature that deal with
coordination of pricing and ordering decisions. The reader is referred to Yano and
Gilbert (2004), Petruzzi and Dada (1999), Eliashberg and Steinberg (1993) for three
comprehensive reviews in this area. Instead, we shall focus on articles that emphasize on
the use of demand management strategies to “shape” uncertain demand so that a firm can
use an inflexible supply to meet the modified demand.
When a firm’s supply capacity is fixed, Carr and Lovejoy (2000) is the first to develop a
single-period model for a firm to handle multiple customers with random demand
distributions. For each customer, they consider the case in which the firm can choose to
accept only a fraction of the customer’s demand distribution. The objective is to choose
different fractions of customer demand distributions so that the firm’s expected profit is
maximized for a given supply capacity. By analyzing the mean and variance of the total
demand generated from different fractions of customer demand distributions, Carr and
Lovejoy determine the optimal portfolio of demand distributions. Van Mieghem and
Dada (1999) consider a single product firm that faces a linear demand curve with
uncertain intercept and has to decide on its production quantity and price. They consider
different strategies including one strategy that is called price postponement strategy.
Under the price postponement strategy, the firm needs to decide on the order quantity in
the first period and then determine the price in the second period after observing updated
information about the demand. Essentially, the supply is fixed after the first period.
Hence, the price postponement strategy enables a firm to use price as a response
mechanism to change demand so that the modified demand is better matched with the
fixed supply. By formulating the problem as a two-period stochastic dynamic
programming problem, Van Mieghem and Dada show that the price postponement is
more effective than other strategies being considered.
Besides the demand management strategy examined by Carr and Lovejoy (2000) and Van
Mieghem and Dada (1999), it appears that the remaining demand management strategies
are designed to generate one or more of the following effects:
a. Shifting demand across time;
b. Shifting demand across markets; and
c. Shifting demand across products.
We now review the relevant literature in each of these three categories.
21
3.1. Shifting Demand Across Time
In the service industries such as utilities, airlines and hotels, firms usually set higher
prices during peak seasons in order to shift demand to off-peak seasons. This type of
pricing mechanism is also known as revenue management or yield management. By
offering different prices at different times, it would enable the firm to increase the profit
generated from a fixed supply capacity by capturing customers in different segments who
are willing to pay different prices for the service offered in different times. For revenue
management literature that deals with hotel bookings, the reader is referred to Bitran and
Gilbert (1996), Badinelli (2000), and the references therein. In most cases, due to
uncertain customer arrivals and uncertain cancellations, these models are usually
formulated as a dynamic programming problem. For revenue management literature that
deals with airline reservations, the reader is referred to Dana (1999) and a comprehensive
survey provided by Weatherford and Bodily (1992). For revenue management literature
that deal with peak-load pricing for managing public utilities, the reader is referred to
Crew and Kleindorfer (1986) for a review of economics literature that deals with peak
load pricing with uncertain demand. Essentially, many economists have developed
various models using different types of demand curves and different types of demand
uncertainties to determine the peak-load pricing so that the service provider with fixed
capacity can obtain a higher profit. Besides the recent work developed by Dana (1999),
most economists assumed that the firm knows the time at which peak demand occurs.
The reader is referred to a review of revenue management prepared by Talluri and Van
Ryzin (2005).
In the context of service marketing, many service firms offer price discount to entice
customers to commit their purchase in advance. In many instances, advance-purchase
discount can be easily implemented due to new technologies such as smart cards, online
payments, electronic money, etc. As articulated in Xie and Shugan (2001), advance-
purchase discount can be a win-win strategy for the service provider and their customers.
First, advance-purchase discount enables a firm to use this discriminatory pricing
mechanism to increase sales by serving different market segments. For example, by
considering 2 market segments with different reservation values of the service, Dana
(1998) shows analytically that it is rational for customers with relatively more certain
demands (planned trips) and customers with relatively lower reservation value (leisure
travelers) to commit their purchases in advance. This result also implies that customers
with less certain demands (unplanned trips) and customers with relative higher
reservation value (business travelers) would expect to pay (a potentially higher price) in
the spot market. Second, advance-purchase discount enables customers to receive a
discount over the spot price or to reserve capacity that may not be available during the
spot period. Xie and Shugan (2001) present a two-period model in which the advance-
purchase price is announced in the first period and while the price is not announced in the
second period; however, the probability distribution of the price in the second period is
known to all customers. Since the price in the second period is unknown to the
customers and since the reservation price is uncertain for each customer, they would
22
make their purchase in the first period if the surplus obtained form purchasing in advance
is higher than the expected surplus obtained from purchasing later. By using backward
induction, Xie and Shugan develop the conditions under which the firm should offer
advance-purchase discount. By considering an extension in which the prices in both
periods are pre-announced, they determine the conditions under which the firm should
offer advance-purchase discount.
In the context of supply chain management, one needs to address the production planning
and inventory control issues that are not addressed in the economics or marketing
literature. In most cases, the retailers would usually pre-announce the prices for both
periods to their customers. Weng and Parlar (1999) is the first to analyze the benefit of
advance-commitment discount. They consider the case in which a retailer offers price
discount to entice customers to pre-commit their orders prior to the beginning of the
selling season. The advance-commitment discount program can be a win-win solution.
First, the customers can enjoy a lower price by pre-committing their orders early.
Second, the retailer can benefit from the reduction in demand uncertainty because the
advance-commitment discount enable the retailer to convert some uncertain customer
demands to pre-committed orders that are known in advance. By considering the demand
uncertainty reduction generated by the advance-commitment discount, Weng and Parlar
determine the optimal order quantity and the optimal discount rate for the retailer.
Tang et al. (2004) extend Weng and Parlar’s model by considering a more general
situation that can be described as follows. First, they consider a situation in which the
market consists of two customer segments with different purchasing behaviors toward
advance-commitment discount. They show how advance-commitment discount would
enable the retailer to increase the total expected sales. Second, they consider the case in
which the retailer can utilize the pre-committed orders obtained prior to the beginning of
the selling season to improve the accuracy of the forecast of the demand to be occurred
during the selling season. They show how this improved forecast would enable the
retailer to reduce the total expected over-stocking and under-stocking costs. Moreover,
they examine various benefits associated with advance-commitment discount programs.
As a follow-on study, McCardle et al. (2004) extend the model presented in Tang et al. to
the case in which two competing retailers need to decide whether to launch the advance-
commitment discount program or not. McCardle et al. show that both retailers would
offer the advance-commitment discount program at the equilibrium. However, when
there is a fixed cost for implementing this discount program, they develop conditions
under which exactly one retailer would offer the discount program at the equilibrium.
The advance-commitment discount program is applicable to non-seasonal products as
well. By studying the supply chain operations associated with steel processing, Gilbert
and Ballou (1999) present a continuous time model in which the steel distributor offers
price discount to customers who pre-commit their orders in advance. By knowing these
pre-committed orders earlier, they show that the standard deviation of the demand over
the replenishment lead time periods is reduced. By using a traditional approximate cost
model for lost sales, they show how advance-commitment discount programs would
enable a steel distributor to increase his expected profit and to improve customer service
23
level at the same time. By examining the profits before and after the launch of the
advance-commitment discount program, Gilbert and Ballou present an approach for
determining the optimal discount price.
All advance-commitment discount models are based on the single product case except the
model presented in Weng and Parlar (2005). Specifically, Weng and Parlar examine a
situation when a manufacturer produces a standardized product and a make-to-order
customized product. Also, the manufacturer offers advance-commitment discount to
customers who pre-commit their orders for the standardized product. By considering the
case in which the market consists of two segments with different purchasing behaviors
toward advance-commitment discount, they formulate the manufacturer’s problem as a
stochastic dynamic programming problem. They show that the advance-commitment
discount program would enable the manufacturer to increase the total expected demand
and to reduce demand uncertainty. When the standardized product is cheaper to produce,
they develop conditions under which the manufacturer should offer the advance-
commitment discount program.
While the advance-commitment discount program that is designed to enable a firm to
shift customer demands earlier, there is another strategy that would entice customers to
shift their demands later. This strategy is called “demand postponement” strategy and it
is intended to entice some customers to accept their shipments in a later period. Iyer et
al. (2003) is the first that examines the benefits of the demand postponement strategy. To
manage uncertain demand with a fixed supply capacity, they consider the case in which a
firm would offer price discount to a fraction of customers who are willing to accept late
shipments. Essentially, this strategy is akin to the overbooking situation in which an
airline may offer incentive to entice a fraction of customers who are willing to take a later
flight. Iyer et al present a two-period model and determine the optimal fraction of
customer demands to postpone. In addition, they characterize conditions under which the
firm should adopt the demand postponement strategy.
3.2. Shifting Demand Across Markets
When selling products with short life cycles in different markets, firms need to manage
product rollovers (the process of phasing out old products and introducing new products).
As articulated in Billington et al. (1998), different firms have implemented various
rollover strategies with different degrees of success. One of the key challenges for
managing product rollovers successfully is uncertain demands in different markets. To
mitigate the demand risks in different markets, Billington et al. present a “solo-rollover
by market” strategy that calls for selling the new product in different markets with non-
overlapping selling seasons. The solo-rollover by market strategy is more suitable for
situations when there is a natural time delay of the selling season in two different
markets. For example, the selling season of ski wear in the U.S. usually ends in May,
while the selling season in South America usually begins in June.
24
Suppose a firm adopts the solo-rollover by market strategy. Then the firm has to decide
how much to stock for the first market during in first period; how much of the unsold
inventory from the first market to transship to the second market at the end of the first
period; and how much to stock for the second market at the beginning of the second
period. Hence, the solo-rollover by market strategy enables a firm to shift the supply
from the first market to the second market. Kouvelis and Gutierrez (1997) is the first to
examine this stocking and transshipment decisions for two markets with non-overlapping
selling seasons. They consider a firm that sells seasonal goods in a primary market in the
first period and in the secondary market during the second period. By capturing the
possibility of shipping some of the leftover inventory from the primary market to the
secondary market at the end of the selling season of the primary market, they present a 2-
period stochastic dynamic program to determine the optimal production quantity for the
corresponding market in each period and the optimal amount of leftover inventory to be
shipped from the primary market to the secondary market. Due to the possibility of
selling the leftover from the primary market at the second market, they show that the
optimal production quantity for the primary market is higher than the case when the
secondary market does not exist.
Recently, Petruzzi and Dada (2001) extend Kouvelis and Gurtierrez’s model to the case
in which the firm can use a pricing mechanism to shift some of the demand from the
primary market to the secondary market. Specifically, Petruzzi and Dada consider that
the firm can utilize information to make better pricing and ordering decisions as follows.
First, the firm can choose the selling price as well as the stocking level for the primary
market during the first season. Second, the firm can utilize the actual sales observed in
the primary market to improve the accuracy of the forecast of the demand for the
secondary market in the second season. Third, given the updated forecast, the firm can
determine the transshipment quantity to be shipped from the primary market to the
secondary market, the stocking level and the selling price of the product for the
secondary market in the second selling season. By formulating the two-period problem
as a stochastic programming problem with recourse, Petruzzi and Dada establish the
characteristics of the optimal pricing and ordering decisions for both markets.
3.3. Shifting Demand Across Products
When selling multiple products in a single market, many marketing researchers have
examined various pricing and promotion strategies to entice customers to switch brands
or products. The ultimate goal of various marketing strategies is to help a firm to
increase market share, sales, or revenue. For example, Raju et al. (1995) present a model
to capture the brand switching behavior when a store introduces a store brand to compete
with the existing national brands. They show how to determine the optimal retail prices
for the national brands and the new store brand so as to maximize the store’s revenue.
Chong et al. (2001) show how a retailer can obtain higher revenue by adjusting its
product assortments and pricing so that the store can offer its customers the right products
at the right price. The reader is referred to Lilien et al. (1992) for an extensive review of
marketing models that deal with pricing and promotion strategies. In general, these
25
marketing models do not deal with the operational issues arising from supply chain
management.
In the context of supply chain management, some researchers have developed models by
considering the possibility of shifting the supply/demand from one product to another. It
seems there are two basic mechanisms that would enable to firm to shift the
supply/demand from one product to another. These two mechanisms are: product
substitution and product bundling.
3.3.1. Product Substitution
Product substitution can occur in different settings. First, by selling products with similar
features, a firm can increase the product substitutability. Chong et al. (2004) show how a
firm can increase product substitutability by selecting a specific combination of products
with similar attributes/features. Moreover, they show how product substitutability can
reduce the variance of the aggregate demand. Rajaram and Tang (2001) present a single
period stochastic model of a firm that sells two substitutable products. Specifically, they
consider a situation in which a product with surplus inventory can be used as a substitute
for out of stock products. Hence, the demand of one product can be satisfied by the
supply for another product. They develop conditions under which product substitutability
would enable a firm to reduce the variability of the effective demand for each product.
Moreover, they show that the optimal order quantity of each product and the retailer’s
expected profit increase as product substitutability increases.
Second, product substitution can occur when one product dominates another product in
terms of quality or performance. For example, in integrated circuit (IC) manufacturing,
the output of each production run consists of a random number of chips with different
grades measured according to the processing speed. When higher grade chips can be
used as substitutes for the lower grade chips, Bitran and Dasu (1992) and Hsu and Bassok
(1999) present different models for determining the optimal production quantity at a
wafer fabrication facility with random yields.
Third, one can use a pricing mechanism to entice customers to shift their demand from
one product to another. Parlar and Goyal (1984) consider a case in which the retailer
would offer price discount for the old product, say, one-day old doughnut. Clearly, the
new and old products are substitutable and the retailer can change the level of product
substitution by varying the discount factor. By formulating the problem as a Markov
Decision Process and by considering the demands for the old and new products, they
determine the optimal order quantity for the new product in each period. More recently,
Chod and Rudi (2005) examine another situation in which a firm can use differential
pricing to entice customers to shift the demand for one product to another. They consider
the case in which the firm needs to decide on the production quantity of two similar
products in the first period; however, the firm can postpone the pricing decision of the
each product until the second period. By extending the model developed by Van
26
Mieghem and Dada (1999), Chod and Rudi show a firm can obtain a higher profit by
delaying the pricing decision until the second period.
3.3.2. Product Bundling
In addition to product substitution, a firm can change the demand of the products by
developing bundles. There is an increasing number of retail products being bundled
together and sold. Examples can be found across a range of products including food
(cans of chicken broth), apparel (under garments), cosmetics (shampoo and conditioner),
and electronics (computers and printers). When products are sold in bundles, they force
the customers to buy all products as a bundle, which will affect the effective demand of
the products. Ernst and Kouvelis (1999) examine how product bundles affect the
inventory ordering decisions of a firm. Specifically, they consider the case in which the
products are sold as a bundle and as individual products. Based on their analysis of a
two-product model, they establish the necessary and sufficient conditions for the optimal
ordering quantities. They provide insights on the degree of sub-optimality in profits
when inventory decisions are made without explicit consideration of demand substitution
between the bundles and the individual products. More recently, McCardle et al. (2005)
present a model for determining optimal bundle prices, order quantities, and profits. By
capturing the customer’s valuation of individual products, they generate the demand
distribution of the product bundle. In addition, they determine how product demand,
costs and the relationship of demand between products affect optimal bundle prices and
profits. Moreover, they present conditions under which a firm should bundle their
products. The reader is referred to Stremersch and Tellis (2002) for a comprehensive
review on product bundling literature.
Demand Management
Issue
Sub-category References (in the order of appearance)
Demand Management
General Elmaghraby and Keskinocak (2003), Yano and Gilbert (2004),
Petruzzi and Dada (1999), Eliashberg and Steinberg (1993), Carr
and Lovejoy (2000), Van Mieghem and Dada (1999)
Shifting Demand Across
Time
Gilbert (1996), Badinelli (2000), Dana (1999), Weatherford and
Bodily (1992), Crew and Kleindorfer (1986), Talluri and Van
Ryzin (2005), Dana (1998), Xie and Shugan (2001), Weng and
Parlar (1999), Tang et al. (2004), McCardle et al. (2004), Weng
and Parlar (2005), Iyer et al. (2003),
Shifting Demand Across
Markets
Billington et al. (1998), Kouvelis and Gutierrez (1997), Petruzzi
and Dada (2001),
Shifting Demand Across
Products
Raju et al. (1995), Chong et al. (2001), Lilien et al. (1992),
Product Substitution Chong et al. (2004), Rajaram and Tang (2001), Bitran and Dasu
(1992), Hsu and Bassok (1999), Parlar and Goyal (1985), Chod
and Rudi (2005), Van Mieghem and Dada (1999)
Product Bundling Ernst and Kouvelis (1999), McCardle et al. (2005), Stremersch
and Tellis (2002)
Table 3. Summary of Demand Management Articles.
27
4. Product Management
To compete for market share, many manufacturers expand their product lines. As
reported in Quelch and Kenny (1984), the number of stock keeping units (SKUs) in
consumer packaged goods has been increasing at a rate of 16% every year between 1985
and 1992. Marketing research shows that product variety is an effective strategy to
increase increasing market share because it enables a firm to serve heterogeneous market
segments and to satisfy consumer’s variety seeking behavior. However, while product
variety may help a firm to increase market share and revenue, product variety can
increase manufacturing cost due to an increase in manufacturing complexity. Moreover,
product variety can increase inventory cost due to an increase in demand uncertainty.
These two concerns have been illustrated in an empirical study conducted by MacDuffie
et al. (1996). They show that the production and inventory costs tend to increase as
product variety increases. Therefore, it is critical for a firm to determine an optimal
product portfolio that maximizes the firm’s profit. The reader is referred to Ramdas
(2003) for a comprehensive review of literature in the area of product variety.
To reduce the design and manufacturing costs associated with product variety, firms can
increase product variety by developing different variants based on a common platform.
For example, in the personal computer industry, different computer models are based on
a common platform. Hence, the products would share some common attributes, which
make these products mutually substitutable to a certain extent. As discussed in Section
3.3, product substitution and product bundling would enable a firm to shift demands
across products so that the firm can satisfy more customers without incurring the risk of
over-stocking. However, product substitutability is a key challenge for researchers to
develop analytical models to evaluate market share, revenue, and manufacturing cost
associated with different product portfolio. As articulated in Ulrich et al. (1998), there is
no explicit analytical model for determining an optimal product portfolio with
substitutable products. However, various researchers have examined various product
variety issues. For example, Ulrich et al. present a study of mountain bikes industry and
suggest that firms need to take their internal capabilities such as process technology,
distribution channels, product architecture, supply chain network, etc., into consideration
when making product variety decision. Krishnan and Kekre (1998) develop a regression
model to examine the impact of functional features on software development cost.
Martin et al. (1998) present a method for examining the impact of product variety on
replenishment lead time. Moreover, by considering the attribute levels of different
products associated with a product portfolio, Chong et al. (2004) develop a Logit model
for determining the mean and variance of the sales associated with different compositions
of a product portfolio. The reader is referred to Ho and Tang (1998) for a review of
articles in the area of marketing, operations management and economics that deal with
the issue of product variety. More recently, Caro and Gallien (2005) present a multi-
armed bandit model for selecting an optimal product portfolio of fashion items that
maximizes the expected profit over a finite horizon.
28
In the context of supply chain risk management, the key concern is to determine ways to
reduce inventory cost associated with a given portfolio of products. Based on the
classical inventory theory (c.f., Porteus (2002) and Zipkin (2000)), it is well known that
the average inventory level associated with the order-up-to policy depends on mean and
the standard deviation of the demand over the replenishment lead time. Therefore, in
order to develop a cost-effective product variety strategy, various researchers have
developed different approaches for reducing the standard deviation of the demand over
the replenishment lead time. For instance, as explained in Section 3.1, one can reduce
the demand uncertainty over the replenishment lead time periods by using pricing
mechanisms such as advance-commitment discount, peak load pricing, etc. In this
section, we shall review articles based on three specific product management strategies:
postponement, process sequencing, and product substitution. Since the issue of product
substitution has been discussed in Section 3.3.1, we shall focus our discussion on the
postponement strategy and process sequencing.
4.1. Postponement
Consider a manufacturing system that produces two end-products. The system has N
processing stages, where stage 0 is a “dummy” stage. As depicted in the Figure 3, the
first k stages are common to both end-products and after this stage the products are
differentiated in the sense that they may require different operations or different
components. We call stage k as the “point of differentiation.” Lee and Tang (1997)
describe how delayed product differentiation can be achieved via standardization of
components and subassemblies, modular design, postponement of operations, and re-
sequencing of operations. Recall that stage 0 is a dummy stage; hence, there is no
postponement if k = 0. We let T be the total lead time of the entire manufacturing
process, and L(k) be the lead time from stage 0 to stage k.
The postponement models can be classified according to operating modes (make-to-stock
0
k
k+1
k+1
N
N
D
1
D
2
Lead time L(k) Lead time (T – L(k))
Figure 3. A system with point of differentiation at stage k.
1
29
and make-to-order) and demand forecasts (no forecast updating and with forecast
updating). However, we are not aware of models that deal with make to order system
with forecast updating.
4.1.1. Make-To-Order Systems Without Forecast Updating
Lee (1996) is the first to develop theoretical analysis of the postponement strategy.
When there is no demand forecast updating, he examines the benefits of postponement in
a make-to-order (MTO) system and a make-to-stock (MTS) system. In the MTO system,
work-in-process inventory is held only at stage k and each end product is customized on
demand. Depending on the availability of the inventory at stage k and the processing
capacity at stages (k+1) through (N), the time it takes to respond to a customer is
uncertain. In an MTO system, it is common to measure the system performance
according to the mean response time and the probability of the response time being less
than a target response time. Using these two performance measures, Lee shows that the
optimal order up level S is decreasing in k. Hence, one can reduce the inventory level
by delaying the point of differentiation. While the base stock level S is decreasing in k,
the inventory holding cost rate is likely to be increasing in stage k; i.e., it is more costly
to hold inventory at a later stage. By considering the tradeoff between lower inventory
level and higher inventory holding rate, Lee provides conditions under which
postponement is beneficial.
In Lee’s model, the end products are differentiated according to a single feature. As
such, all end products can be customized from a single point of differentiation. However,
when the end products are differentiated according to multiple features, there could be
multiple points of differentiation. This observation motivates Swaminathan and Tayur
(1998a) and (1998b) to define the semi-finished products held at different points of
differentiation as “vanilla boxes.” Essentially, the firm will stock these vanilla boxes and
then customize different types of vanilla boxes into different end products on demand.
By considering the capacity for the customization process and different demand
scenarios, Swaminathan and Tayur formulate the problem as a stochastic programming
problem with recourse. By examining the structure of the stochastic programming, they
develop a solution methodology for determining the optimal configuration of vanilla
boxes that minimizes the expected stock-out cost and the inventory holding cost of
vanilla boxes.
4.1.2. Make-To-Stock Systems Without Forecast Updating
In a Make-to-stock (MOS) system, Lee (1996) consider the case in which only finished
product inventory is held; i.e., inventory is held after stage N. Conceptually speaking,
this system is akin to the single-depot, multi-warehouse distribution system examined by
Eppen and Schrage (1981). By assuming that the demand distributions for the end-
products are independently normal across time but may be correlated within a time
period, Lee applies the approximate analysis developed by Eppen and Schrage to
determine the base-stock level and the average inventory level for each end-product.
30
Moreover, Lee shows that the finished product inventory level for each end-product is
decreasing in the point of differentiation k. As articulated in Lee (1996), the
postponement strategy can be implemented in a MTO or a MTS system. This
observation motivates Su et al. (2005) to develop a model to compare the total supply
chain costs associated with postponement in a MTO and a MTS system. Their analysis
shows that the MTO system is more cost effective when the number of end products
exceeds a certain threshold level.
When the end products are differentiated according to different features, the
corresponding manufacturing process can have multiple points of differentiation. Garg
and Tang (1997) to extend the model presented in Lee (1996) by considering a system
with multiple points of differentiation. Since system with multiple points of
differentiation is akin to a multi-echelon distribution system, Garg and Tang first extend
Eppen and Schrage’s two-echelon model to a three-echelon model. Then they show that
postponement at each of the differentiation points would result in inventory savings.
Instead of relying on the approximate analysis developed by Eppen and Schrage to
evaluate different postponement strategies, Aviv and Federgruen (1998b) show how one
can develop an exact analysis of the model presented in Garg and Tang (1997).
Lee and Tang (1997) examine a system that can keep work-in-process inventory at every
single stage. They develop a stochastic inventory model by capturing the investment cost
per period for redesigning the products and / or processes, the unit processing cost, and
the inventory holding cost at each stage. Their analysis is based on a decomposition
scheme in which the manufacturing process is decomposed into N independent stages,
each of which will follow an order-up-to level policy. The decomposition scheme
enables them to approximate the system-wide cost function associated with the point of
differentiation k. By examining the underlying property of this explicit cost function,
they develop conditions under which no postponement is optimal; i.e., the optimal point
of differentiation is stage 0. Also, they discuss the conditions under which postponement
is beneficial.
In the postponement literature, most researchers assume that the production capacity is
unlimited. To examine how production capacity can affect the value of postponement,
Gupta and Benjaafar (2004) develop a queuing model for examining the benefits of
postponement in a MTO and a MTS system with limited production capacity. When the
production capacity for stages 1 through k (point of differentiation) is limited, Aviv and
Federgruen (2001a) present a multi-product inventory model for the case when the
product demand is random and periodical. They show that the underlying inventory
model can be formulated as a Markov Decision Process, and that delayed product
differentiation is always beneficial even when the system has limited capacity.
4.1.3. Make-To-Stock Systems With Forecast Updating
In the postponement literature, most researchers assume that the product demands in each
period are random, but they are independent across time and their distributions are
31
known. As a result of these assumptions, the benefit of postponement is derived from
“risk pooling” in the sense that all stages before the point of differentiation (stage k)
would plan according to the “aggregate demand” instead of individual product demand.
Besides risk pooling, postponement enables a firm to delay the product differentiation so
that the production quantity decision for the final products can be made in a later period
of time. When the timing of this decision is delayed, the firm can use the actual demands
observed in earlier period to obtain more accurate forecasts of future demands. To
explore further about the benefit of postponement with forecast updating, Whang and Lee
(1998) extend the make-to-stock model presented in Lee (1996) by considering the case
in which the demand D
i
(t) of end-product i in period t possesses the following form:
),(,,...,2,1,,...,2,1,)(
2
1
ijijijij
t
j
ii
aNandTtniwheretD
σεεµ
==+=
=
Whang and Lee assume that the parameters a
ij
and σ
ij
are known. This demand
distribution is a form of random walk that enables one to capture a series of random
shocks (economic trends, random noises, etc.). As time goes on, the decision maker can
use the shocks observed in earlier periods (i.e., some of the ε
ij
’s are now known) to
develop a more accurate forecast of D
i
(t). By incorporating the capability to obtain
more accurate demand forecast as time goes on, Whang and Lee show that one can obtain
substantial reduction in the end-product inventory by using more accurate forecasts. In
addition, they show analytically that significant inventory savings can occur even when
the point of differentiation k occurs in the early stage.
Aviv and Federgruen (2001b) consider a more general demand distribution in which the
parameters of the demand are unknown. In their model, they consider the case in which
the decision maker would update the demand forecast in a Bayesian manner. They show
analytically that the standard deviation of the total demand over the lead time periods is
decreasing over time. Furthermore, they show that this standard deviation is decreasing
in the point of differentiation k. This implies that it is more beneficial to update the
demand forecast when postponement occurs in a later stage. The reader is referred to
Garg and Lee (1998), Aviv and Federgruen (1998b), and Yang et al. (2004) for
comprehensive reviews of research literature that examine different postponement issues.
4.2 Process Sequencing
As noted in Section 4.1, postponement is an effective way to reduce variabilities in a
supply chain. Lee and Tang (1998) suggest that variabilities can also be reduced by
reversing the sequence of manufacturing processes in a supply chain. Their suggestion is
motivated by the re-engineering effort at Benetton. In the woolen garment industry,
virtually all manufacturers will use the dye-first-knit-later sequence; i.e., dye the yarns
into different colors first and then knit the colored yarns into different finished products.
However, as a way to reduce inventory, Benetton pioneered the knit-first-dye-later
process by reversing the “dyeing” and “knitting” stages (c.f., Dapiran (1992)).
Intuitively speaking, the knit-first-dye-later strategy would be beneficial when there is
32
only one style of woolen sweaters with multiple colors. This is because it would result in
delaying product differentiation after the “knitting” stage. However, when there are
multiple styles and multiple colors, it is not clear which strategy is better. To determine
the conditions under which a particular process sequence is better, Lee and Tang (1998)
develop a model of a production system that produces products with 2 features (A and B),
each of which has 2 choices (1 and 2). As depicted in Figure 4, the product demands
(X
11
, X
12
, X
21
, X
22
) are assumed to be multi-nomially distributed with parameters (N; θ
11
,
θ
12
, θ
21
, θ
22
), where the total demand N is normally distributed with mean µ and
standard deviation σ, and θ
ij
corresponds to the probability that the customer will
choose choice i of feature A and choice j of feature B. By considering p as the
probability that a customer will purchase a product with choice 1 of feature A and by
examining the conditional probabilities: Prob(B1 | A1) = f(p) and Prob(B1 | A2) = g(p),
one can express θ
ij
in terms of p, f(p) and g(p).
Lee and Tang argue that the total expected cost associated with the intermediate products
is proportional to the sum of the variances of demands in a period. As such, they show
that the process sequence A-B has a smaller variance than the process sequence B-A if:
(µ - σ
2
){p(1-p) – [pf(p) + (1-p)g(p)]{1 – [pf(p) + (1-p)g(p)]}} < 0. This result has the
following implications. Consider the special case in which f(p) = g(p) = q, where q is
independent of p. If the product demand is stable (i.e., when µ > σ
2
), then the process
sequence A-B has a lower variance when feature (attribute) A is less variable than
attribute B (i.e., when |0.5 – p| > |0.5 – q|). However, the reverse is true when µ < σ
2
.
By considering the sum of the standard deviations as an alternative measure, Kapuscinski
and Tayur (1999) conduct their analysis associated with the special case. They show that
it is optimal to process the attribute with less variable first, regardless of the values of µ
and σ
2
.
Some researchers have generalized Lee and Tang’s model in the following manner.
First, Federgruen (1998) develops a general definition of when attribute A is more
variable than attribute B in terms of θ
ij
. He shows a more general conditions under
which the process sequence A-B has a smaller variance than the process sequence B-A.
A1
A2
B1
B2
B1
B2
X
11
X
12
X
21
X
22
Figure 4. A Two-Stage System with 2 Features and 2 Choices.
p
1-p
g(p)
f(p)
33
Next, Jain and Paul (2001) generalize Lee and Tang’s model by incorporating two
important characteristics of fashion goods markets, namely, heterogeneity among
customers and unpredictability of customer preferences. More recently, Yeh and Yang
(2003) develop a simulation model by incorporating additional factors such as lead times,
ordering policies, and inventory holding cost. By using the data obtained from a garment
manufacturer, they show how their simulation model can be used to select a process
sequence that minimizes the total expected cost.
Product Management
Issue
Sub-category References (in the order of appearance)
Product Management
General Quelch and Kenny (1994), MacDuffie et al. (1996), Ramdas
(2003), Ulrich et al. (1998), Krishnan and Kekre (1998), Martin et
al. (1998), Chong et al. (2004), Ho and Tang (1998), Caro and
Gallien (2005), Porteus (2002), Zipkin (2000)
Postponement
General Lee and Tang (1997),
Make to order systems
without forecast
updating
Gunasekaran and Ngai (2005), Lee (1996), Tayur et al (1998a),
Tayur et al. (1998b)
Make to stock systems
without forecast
updating
Lee (1996), Eppen and Schrage (1981), Su et al. (2005), Garg and
Tang (1997), Aviv and Federgruen (1998b), Lee and Tang (1997),
Gupta and Bejaafar (2004), Aviv and Federguen (2001a)
Make to stock systems
with forecast updating
Whang and Lee (1998), Lee (1996), Aviv and Federguen (2001b),
Garg and Lee (1998), Aviv and Federguen (1998b), Yang et al.
(2004)
Process Sequencing
Lee and Tang (1998), Kapuscinski and Tayur (1999), Federguen
(1998), Jain and Paul (2001), Tayur (1999), Yeh and Yang (2003)
Table 4. Summary of Product Management Articles.
5. Information Management
As explained in Fisher (1997), most consumer products can be classified as fashion
products or functional products. Basically, fashion products usually have shorter life
cycles and higher levels of demand uncertainties than the functional products. Therefore,
different information management strategies would be needed to manage for different
types of products especially in the presence of supply chain risks. For this reason, we
shall classify the work in this section according to the product types: fashion products and
functional products.
5.1. Information Management Strategies for Managing Fashion Products
As articulated in section 4, reducing the standard deviation of the demand over the
replenishment lead time would result in inventory reduction for the entire supply chain.
When managing products with short life cycles, short replenishment lead times could
enable a retailer to place more than one order over the selling season. For example,
various researchers consider a situation in which a retailer can place two orders over the
34
selling season. Specifically, the retailer can place one order prior to the beginning of the
selling season and another order during the selling season. In the fashion goods industry,
this type of replenishment system is called the “quick response” system. Clearly, the
second order provides a great opportunity for the retailer to obtain more accurate demand
forecast by using the actual sales data. Fisher and Raman (1996) develop a two-period
stochastic dynamic programming model with demand forecast updating to analyze the
quick response system. By implementing their model at a skiwear company called
Obermeyer, they illustrate how the quick response system would enable Obermeyer to
achieve a higher customer service level with a lower level of inventory. The reader is
referred to Raman (1998) for a review of quantitative models of quick response systems.
Gurnani and Tang (1999) analyze a similar quick response system except that the unit
cost for the second order is uncertain. By formulating the problem as a two-period
dynamic programming problem, they show that an order-up-to level policy is optimal.
Instead of focusing on the retailer’s perspective, Iyer and Bergen (1997) and Iyer (1998)
analyze the impact of a quick-response system on both retailers’ and manufacturers’
inventories. Specifically, they show that, when the customer service level is at least 0.5,
the quick-response system is not Pareto in the sense that the retailer would obtain a
higher expected profit and the manufacturer would achieve a lower expected profit. They
develop conditions under which a quick-response system is beneficial to both the retailers
and manufacturers. Donohue (2000) considers a variant of the quick-response system in
which the retailer can place their orders in two modes: low cost with long lead time and
high cost with short lead time. By formulating the problem as a two-period dynamic
programming problem, she derives an optimal ordering policy and an optimal contract
that coordinates the supply chain.
All quick response models assume that the manufacturer can always fill the second orders
placed by the retailers. However, as articulated in the Benetton case prepared by
Signorelli and Heskett (1984), manufacturers may not be able to guarantee complete
fulfillment of the second order. Smith et al. (2002) investigate the retailer’s optimal order
quantities, the retailer’s profit, and the manufacturer’s profit for the case when the
manufacturer can fulfill the second order partially. By considering a stylized model, they
show that the manufacturer should provide either complete fulfillment or no fulfillment
of the second orders when the underlying demand distribution is either uniform or
exponential. Specifically, they show analytically that partial fulfillment of the second
orders is never optimal for the manufacturer.
5.2. Information Management Strategies for Managing Functional Products
When managing products with long life cycles, market information is critical for
generating accurate demand forecasts. However, since wholesalers, distributors,
manufacturers, and suppliers are farther remove from the consumer market, they usually
do not have first-hand market information such as point of sales data, customers’
preferences, and customer response to various pricing and promotion strategies. Instead,
upstream supply chain partners usually generate their demand forecasts based on the
35
orders placed by their downstream partners. Planning according to the orders placed by
the downstream partners would create a phenomenon called the “bullwhip effect” that
was coined by Procter and Gamble. Essentially, the bullwhip effect depicts the
phenomenon in which the orders exhibit an increase in variability up the supply chain,
even when the actual customer demands were fairly stable over time (c.f., Sterman
(1989)). The increase in variability of the orders up the supply chain can cause many
problems for the upstream partners including higher inventory, lower customer service
level, inefficient use of production and transportation capacities, etc. In order to mitigate
the bullwhip effect, one needs to identify the root causes.
Lee et al. (1997b) is the first to show that the bullwhip effect can occur even when every
supply chain partners operate optimally and rationally. The bullwhip effect has also been
shown independently by Bagahana and Cohen (1998). To establish the existence of the
bullwhip effect, Lee et al. develop a 2-level supply chain that consists of a retailer and a
manufacturer. They assume that the retailer “knows” that the underlying demand process
follows an auto-regressive process AR(1) so that the demand in period t, denoted by D
t
,
is equal to:
ttt
DdD
ε
ρ
++=
1
.
Notice that d represents the base demand level, ρ represents the correlation of demands
in successive periods, where |ρ| < 1, and ε
t
represents the error term that is normally
distributed with mean 0 and standard deviation σ . Lee et al. (1997b) consider the case
in which the retailer would act rationally by following an order-up-to level policy and by
placing an order Q
t
in period t. To show that the bullwhip effect occurs, they prove that
Var(Q
t
) Var(D
t
).
More recently, Gilbert (2005) generalizes Lee et al. model by considering a more general
demand process than the AR(1) process that is known as the Autoregressive Integrated
Moving Average (ARIMA) time-series. When the underlying demand process is an
ARIMA process, Gilbert shows that the order quantity Q
t
associated with the order-up-
to level policy is also an ARIMA process. Li et al. (2005) develop a simulation model
for the case when the demand process is an ARIMA process. By varying the values of
the parameters associated with the ARIMA process, they show that the bullwhip effect
does not always occur. More importantly, they discover an “anti-bullwhip effect” that
would occur for certain values of the parameters by showing that Var(Q
t
) Var(D
t
);
i.e., the variance of the order quantity is lower than the variance of the demand.
While Lee et al. (1997b) show that the bullwhip effect will occur when the retailer has
knowledge about the demand distribution, Chen et al. (1998), (2000a) and (2000b)
investigate the occurrence of the bullwhip effect for the case when the retailer does not
know the underlying demand process follows an AR(1) process; however, the retailer
would use a moving average or an exponential smoothing method to forecast future
demands. They show that the bullwhip effect will occur and that the bullwhip effect will
be larger when the retailer uses an exponential smoothing forecast instead of a moving
36
average forecast. More recently, Zhang (2004) extends the work of Chen et al. by
examining the impact of different forecasting methods on the bullwhip effect.
To mitigate the bullwhip effect, Lee et al. (1997c) identify four root causes of the
bullwhip effect: demand forecasting, batch ordering, supply shortage, and price
variations. In addition, they propose strategies for mitigating the bullwhip effect
including: information sharing, vendor managed inventory, and collaborative forecasting
and replenishment planning. We now review articles that examine the benefits of these
three strategies.
5.2.1. Information Sharing
Lee et al. (2000) study the benefits of information sharing in a two-level supply chain.
They consider the case in which the retailer has the information about the underlying
demand distribution (i.e., an AR(1) process) and the retailer would order according to an
order-up-to policy in each period. When there is no information sharing, the
manufacturer has the information about the underlying demand distribution and the
retailer’s ordering policy; however, the manufacturer does not have the information about
the actual demand realized in period t (i.e., the manufacturer does not know the
realization of the error term ε
t
in period t). When there is information sharing, the
retailer would share the information about the actual demand realized in period t as well.
By assuming that there exists a reliable exogenous source of inventory, information
sharing has no impact on the retailer because the retailer’s orders are always received in
full. By examining the inventory level and the relevant costs incurred by the retailer and
the manufacturer, Lee et al. show analytically that information sharing is beneficial to the
manufacturer, not the retailer. Moreover, information sharing is most beneficial to the
manufacturer especially when the correlation coefficient ρ is high. Also, in order to
entice retailer to share demand information with the manufacturer, Lee et al. suggest
various mechanisms including price discount and replenishment lead time reduction.
Cheng and Wu (2005) extend Lee et al.’s model to the multi-retailer case and they
conclude that information sharing would enable the manufacturer to reduce both the
inventory level and the total expected cost. Lee et al. commented that information
sharing would be less valuable to the manufacturer if the manufacturer uses the historical
stream of orders from the retailer to forecast demand. Raghunathan (2001) confirms this
idea analytically when the underlying demand is an AR(1) process. More recently, Gaur
et al. (2005) extend Raghunathan’s model to the case in which the demand process is a
more general process than the AR(1) process, namely, the AR(p) process for p 1 and
the autoregressive moving-average process ARMA process.
By assuming that the underlying demand is independent and identically distributed,
Gavirneni et al. (1999) develop a model to examine the benefits of information sharing
for the case in which the manufacturer has limited production capacity. In their model,
the retailer has the information about the underlying demand distribution and the retailer
would order according to an (s, S) policy. Under the (s, S) policy, the retailer would
place an order in a period only when the inventory level drops below s. When there is no
37
information sharing, the manufacturer has the information about the underlying demand
distribution and the retailer’s ordering policy; however, the manufacturer does not have
the information about the retailer’s inventory level. When there is information sharing,
the retailer would share the information about the actual inventory level with the
manufacturer in each period. They show that information sharing is beneficial to the
manufacturer especially when the manufacturer’s production capacity is higher or when
the demand uncertainty level is moderate. Cachon and Fisher (1997) and (2000) analyze
the benefits of information sharing for the N-retailer case in which the manufacturer has
limited production capacity. By assuming that each retailer implements a (R, nQ) policy,
they show analytically that information sharing is beneficial to the retailer and the
manufacturer. In addition, Cachon and Fisher (2000) show numerically that lead time
reduction will be more beneficial than information sharing.
Zhao et al. (2002) develop a simulation model to examine the impact of forecasting
methods such as moving average, exponential smoothing, and Winters’ method, etc., on
the value of information sharing in a supply chain that has 1 manufacturer and N retailers.
They show that the cost savings for the entire supply chain are more substantial when the
retailers share information about future orders with the manufacturer than the case in
which the retailers share information about the customer demand. While many
companies reported that sharing information (such as customer demand, inventory level,
or demand forecast) among supply chain partners is beneficial, there are several obstacles
for supply chain partners to share private information. For instance, retailers are reluctant
to share information with the manufacturer because of fear (lower bargaining power,
information leakage, etc.). Besides fear, there are other problems associated with
forecast sharing in practice. Terwiesch et al. (2005) articulate that when a retailer revises
his forecasts (or soft orders) frequently before placing a firm order, the manufacturer may
ignore the revisions. Also, when a manufacturer is unable to fulfill the firm order in one
period, the retailer may inflate his soft orders in future periods to ensure sufficient supply.
As such, this could lead to a lose-lose situation. By using the data collected from a
semiconductor company, Terwiesch et al show empirically that the manufacturer would
penalize the retailer for unreliable forecasts by delaying the fulfillment of forecasted
orders. Also, they show that the retailer would inflate their orders in the form of
excessive order cancellations. Therefore, both manufacturer and retailer would lose
when sharing forecast information.
5.2.2. Vendor Managed Inventory
As articulated in Lee et al. (2000), information sharing is beneficial to the manufacturer
not the retailer. As such, many manufacturers develop various initiatives to entice the
retailer to share demand information with the manufacturer. Besides offering price
discount, various manufacturers launched an initiative called Vendor Managed Inventory
(VMI). Under the VMI initiative, the retailers delegate the ordering and replenishment
planning decisions to the manufacturer. In return, the manufacturer gains direct
information access regarding customer demand and retailers’ inventory positions. To
ensure that the retailer will achieve higher customer service levels with lower inventory
38
costs, the manufacturer either owns the inventory at the retailer’s warehouse subject to a
minimum inventory level or issues some form of promises that the inventory at the
retailer’s warehouse will stay within certain pre-specified limits.
Under the VMI initiative, the retailer can reduce the overhead and operating costs
associated with replenishment planning, while enjoying certain guaranteed service levels.
Even though the manufacturer takes on the burden to manage the retailer’s inventory
under the VMI initiative, the manufacturer can derive the following benefits: (1) reduced
bullwhip effect due to direct information access regarding customer demands and (2)
reduced production/logistics/transportation cost due to coordinated production /
replenishment plans for all retailers. Disney and Towill (2003) develop a simulation
model to analyze the bullwhip effect under the VMI initiative. Their simulation results
confirm that VMI can reduce the bullwhip effect by 50%. Clearly, reducing the bullwhip
effect and coordinated planning would enable the manufacturer to reduce inventory.
Johnson et al. (1999) examine the performance of VMI in different settings: (a) the
manufacturer has limited capacity and (b) some retailers adopt the VMI scheme while the
remainders adopt the information sharing scheme. By considering the case that VMI
would enable the manufacturer to coordinate the replenishment plan by consolidating the
customer demands (instead of orders placed by the retailers), they show that VMI would
reduce inventories for the manufacturer and the retailer.
Aviv and Federgruen (1998a) develop an analytical model to evaluate the retailer’s and
the manufacturer’s operating cost under an information sharing scheme and an VMI
initiative. Under both systems, the manufacturer has information about customer
demand. However, the replenishment plans are determined by the retailers under the
information sharing scheme, while the manufacturer decides on the timing and magnitude
of the replenishment shipments to the retailers. Therefore, under the information sharing
scheme, the effective demand process faced by the manufacturer is essentially the
superposition of orders placed by the retailers in an uncoordinated manner. By
considering that the underlying demand distribution is normal and by using the fact that
the manufacturer has the authority to coordinate the customer demands under the VMI
initiative, Aviv and Federgruen show that the manufacturer can reduce the production
and inventory costs under both systems. To examine further about the benefit of the VMI
initiative under which the manufacturer has the authority to determine the delivery
schedule and the delivery quantity for each retailer, Cetinkaya and Lee (2000) presents an
analytical model for determining an optimal coordinated replenishment and delivery plan
for different retailers located in a given geographical region. By assuming that the
demands at the retailers are independent Poisson processes, they compute an optimal
replenishment quantity and delivery schedule that minimizes the total production,
transportation and inventory carrying costs while meeting certain customer service levels.
Besides the analytical models that examine the benefits of the VMI initiative, there are
other studies that are based on simulation models. The reader is referred to Sahin and
Robinson (2005) and the references therein. Several retailers and manufacturers reported
successful implementations of VMI. For example, Clark and Hammond (1997) show
that the VMI initiated by Campbell Soup provided a win-win situation for Campbell
39
Soup and the retailers. For additional examples of successful implementations of VMI,
the reader is referred to Aviv and Federgruen (1998a), Cetinkaya and Lee (2000), and the
references therein.
5.2.3. Collaborative Forecasting
Under the information sharing scheme or the VMI initiative, not much collaborative
effort is needed. To induce collaboration between the retailers and the manufacturers,
Voluntary Inter-industry Commerce Standards (VICS) association developed an initiative
called Collaborative Planning, Forecasting and Replenishment (CPFR). Under this
initiative, both parties would develop mutually agreeable demand forecasts jointly. To
develop mutually agreeable demand forecasts, the manufacturer would generate an initial
demand forecast based on his market intelligence on products, and the retailer would
create her own initial demand forecast based on customer’s response to pricing and
promotion decisions. Both parties would share their initial demand forecasts and would
reconcile the differences in their forecast to obtain a common forecast. Once both parties
agree on the common demand forecasts, the retailer would develop a replenishment plan
and the manufacturer would develop a production plan independently. The reader is
referred to www.cpfr.org for more details.
The crux of CFPR is collaborative forecasting. Aviv (2001) is the first to develop a
framework for modeling the collaborative forecasting process between a retailer and a
manufacturer. To specify the demand process when there is no collaborative forecast, he
specifies the demand process based on an individual party p’s perspective, where p = r,
m, where r denotes the retailer and m denotes the manufacturer. Specifically, when
there is no collaborative effort, the underlying demand process D
t
from party p’s
perspective is given by:
t
pp
tt
dD
εψ
++= , for p = r, m,
where d represents the base demand level, ψ
p
t
represents the cumulative forecast
adjustment made by party p in past periods up to the beginning of period t, and ε
p
t
represents the residual forecast error of party p’s forecasting method.
By considering the correlation between ψ
m
t
and ψ
r
t
, one can capture the correlation
between the forecast adjustments made by the retailer and the manufacturer. Under the
collaborative forecasting initiative, Aviv assumes that the retailer and the manufacturer
would select the best forecast adjustment ψ
t
so that the forecast error is minimized.
Based on this specific construct, Aviv computes the optimal collaborative forecast
adjustment in each period. For the retailer, he computes the variance of the total demand
over the replenishment lead time under no collaborative forecast and under collaborative
forecast. For the manufacturer who needs to satisfy the order placed by the retailer, he
computes the mean and the variance of the total aggregate order quantity to be placed by
the retailer under no collaborative forecast and under collaborative forecast. Even when
these quantities can be expressed in closed form expressions, it is intractable to evaluate
40
the benefit of collaborative forecast analytically. As a surrogate, Aviv develops an
aggregate supply chain performance measure that is based on the variance of the whole
system: the sum of the variance of the total demand over the retailer’s replenishment lead
time and the variance of the total order quantity over the manufacturer’s replenishment
lead time. Aviv (2001) shows analytically that collaborative forecast would reduce the
system-wide variance. In Aviv (2002), he extends the analysis presented in Aviv (2001)
to the case in which the demand is auto-correlated. Specifically, he considers the case in
which the demand process possesses the following form:
t
pp
ttt
DdD
εψρ
+++=
1
, for p = r, m.
In Aviv (2005), he extends the analysis to the case in which the manufacturer operates in
an environment that calls for production smoothing.
As articulated in Aviv (2001), it is very difficult to evaluate the benefit of CPFR
analytically even for a two-level supply chain. Therefore, many of the comparisons are
conducted numerically. While these numerical examples generate insights, there is no
guarantee that these insights are applicable to a more realistic supply chain. This
observation has motivated Boone et al. (2002) to develop a simulation model to compare
the performance of the CPFR initiative with the performance of a traditional
replenishment policy based an a reorder point. By using the data collected from a
Fortune-500 company and by using a simple process to generate demand forecast, their
simulation model suggests that CPFR would increase customer service level and reduce
inventories for both the manufacturer and the retailer. The reader is referred to Aviv
(2004) for a comprehensive review of CPFR literature.
Information
Management Issue
Sub-category
References (in the order of appearance)
Information
Management
General Fisher (1997)
Managing Products
with Short Life Cycles
General Fisher and Raman (1996), Gurnani and Tang (1999), Iyer and
Bergen (1997), Iyer (1998), Donuhue (2000), Signorelli and
Heskett (1984), Smith (2002)
Managing Products
with Long Life Cycles
General Sterman (1989), Lee et al. (1997b), Bagahana and Cohen (1998),
Gilbert (2005), Li et al. (2005), Chen (1998), (2000a) (2000b),
Zhang (2004), Lee (1997c)
Information Sharing Lee et al. (2000), Cheng and Wu (2005), Raghunathan (2001),
Gaur et al. (2005), Gavirneni et al. (1999), Cachon and Fisher
(1997), Cachon and Fisher (2000), Zhao et al. (2002), Terwiesch
et al. (2005)
Vendor Managed
Inventory
Lee et al. (2000), Disney and Towill (2003), Johnson et al. (1999),
Aviv and Federgruen (1998a), Cetinkaya and Lee (2000), Sahin
and Robinson (2005), Clark and Hammond (1997)
Collaborative
Forecasting
Aviv (2001), Aviv (2002), Boone et al (2002), Aviv (2004)
Table 5. Summary of Information Management Articles.
41
6. Managing Supply Chain Risks
Upon examining the underlying assumptions of the models reviewed so far, it appears
most of the quantitative models are designed for managing operational risks. Even
though these quantitative models often provide cost effective solutions for managing
operational risks, there do not address the issue of disruption risks in an explicit manner.
Before we present some potential research ideas for managing supply chain disruption
risk in the next section, we shall examine how disruptions risks are managed in practice
and relate these practices to the models reviewed earlier. After reviewing some
qualitative analyses presented in various risk management and supply chain risk
management articles, we can summarize the key findings as follows:
1. Managers’ attitude towards risks. Sharpira (1986) and March and Sharpira (1987)
study managers’ attitude towards risks and they conclude that:
Managers are quite insensitive to estimates of the probabilities of possible
outcomes.
Managers tend to focus on critical performance targets, which affect the way they
manage risk.
Managers make a sharp distinction between taking risks and gambling.
The first conclusion can be explained by the fact that managers do not trust, do not
understand, or simply do not much use precise probability estimates. This is consistent
with the results obtained by other researchers (c.f., Kuneuther (1976), Fischoff et al.
(1981)). Since managers are insensitive to probability estimates, March and Sharpira
(1986) noted that managers are more likely to define risk in terms of the magnitude of
loss such as “maximum exposure” or “worst case” instead of expected loss defined
earlier. The second conclusion is based on the observation that most managers are
measured by a set of performance targets. March and Sharpia (1986) argue that these
performance targets would cause the managers to become more risk averse (or risk
prone) when their performance is above (or below) certain target. Finally, the third
conclusion is driven by the fact that companies tend to reward managers for obtaining
“good outcomes” but not necessarily for making “good decisions.”
2. Managers’ attitude towards initiatives for managing supply chain disruption
risks. According to various major case studies conducted by Closs and McGarrell
(2004), Rice and Caniato (2003) and Zsidisin et al. (2001) and (2004b), they concluded
that:
Most companies recognize the importance of risk assessment programs and use
different methods, ranging from formal quantitative models to informal
qualitative plans, to assess supply chain risks. However, most companies invested
42
little time or resources for mitigating supply chain risks.
Due to few data points, good estimates of the probability of the occurrence of any
particular disruption and accurate measure of potential impact of each disaster are
difficult to obtain. This makes it difficult for firms to perform cost/benefit
analysis or return on investment analysis to justify certain risk reduction programs
or contingency plans.
Firms tend to underestimate disruption risk in the absence of accurate supply
chain risk assessment. As reported in Kunreuther (1976), many managers tend to
ignore possible events that are very unlikely. This may explain why few firms
take commensurable actions to mitigate supply chain disruption risks in a
proactive manner. As articulated in Repenning and Sterman (2001), firms rarely
invest in improvement programs in a proactive manner because “nobody gets
credit for fixing problems that never happened.”
6.1. Robust Strategies for Mitigating Operational and Disruption Risks
In the absence of accurate measures of the probability of an occurrence of a major
disruption and the potential impact of a disruption, Tang (2005) argue that firms will
become more willing to implement certain “robust” supply chain strategies for mitigating
disruption risks if these strategies possess two specific properties:
Efficiency – the strategy would enable a firm to manage operational risks
efficiently regardless of the occurrence of major disruptions.
Resiliency – the strategy would enable a firm to sustain its operation during a
major disruption and recover quickly after a major disruption. The reader is
referred to Christopher (2004), Chopra and Sodhi (2004), and Lee (2004) for
different approaches for establishing resilient supply chains.
Tang’s argument is based on the fact that efficiency and resiliency are critical for firms to
ensure profitability and business continuity, which is congruent with the definition of
supply chain risk management defined in Section 1. Also, when a robust strategy is
efficient, most firms can perform cost/benefit analysis or return on investment to justify
strategies for improving efficiency under operational risks. As such, the management is
more willing to implement a strategy that would enhance the efficiency and resiliency.
Upon reviewing the strategies reviewed in Sections 2, 3, 4, and 5 along with the strategies
firms have implemented for mitigating operational and disruption supply chain risks, we
now highlight certain robust strategies for improving the efficiency and resiliency of a
supply chain.
43
6.2. Robust Supply/Demand/Product/Information Management Strategies
6.2.1. Robust Supply Management Strategies
Among the supply management strategies described in Section 2.4, it appears the multi-
supplier strategy is the most common approach for reducing supply chain risks. For
example, both Sheffi (2001) and Kleindorfer and Saad (2005) recommend the use of
multiple suppliers as a way to manage supply chain operational and disruption risks. For
example, as articulated in Huchzermeier and Cohen (1996) and others, spreading multiple
suppliers in multiple countries would enable a firm to manage operation risks such as
normal exchange rate fluctuations efficiently. In addition, having multiple suppliers in
multiple countries can make a supply chain more resilient during a major disruption. For
example, When Indonesia Rupiah devalued by more than 50% in 1997, many Indonesian
suppliers were unable to pay for the imported components or materials, and, hence, were
unable to produce the finished items for their U.S. customers. However, with a network
of 4,000 suppliers throughout Asia, Li and Fung (www.lifung.com), the largest trading
company in Hong Kong for durable goods such as textiles and toys, shifted some
production from Indonesia to suppliers in other Asian countries. Also, Li and Fung
provided financial assistance such as line of credit, loans, etc., to those affected suppliers
in Indonesia to ensure that their U.S. customers would receive their orders as planned.
The reader is referred to McFarlan (2002) for details.
In many instances, the buyer does not have the luxury to shift production among different
suppliers because of the very limited number of suppliers available in the market. To
cultivate additional suppliers, certain supply contracts described in Section 2.5 could
serve as robust strategies that would make a supply chain more efficient and resilient.
For instance, revenue (or risk) sharing contracts are known to be efficient because it can
coordinate the channel partners when facing uncertain demand (c.f., Pasternack (2002)).
In addition, revenue sharing contracts could make a supply chain more resilient. For
example, due to uncertain specification of the flu vaccine in any given year, the uncertain
market demand, and the price pressure from the U.S. government, there are only two
remaining vaccine makers for the U.S. market. This has created a shortage of 48 million
flu shots in 2004 when Chiron’s Liverpool plant was suspended due to bacteria
contamination (c.f., Brown (2004)). To make the flu vaccine supply chain more resilient,
the U.S. government could consider offering certain risk sharing contracts to entice more
suppliers to re-enter the flu vaccine market. For instance, the government could share
some financial risks with the suppliers by committing a certain quantity of flu vaccine in
advance at a certain price and buy back the unsold stocks at the end of the flu season at a
lower price. With more potential suppliers, the U.S. government would have the
flexibility to change their orders from different suppliers quickly when facing major
disruptions.
6.2.2. Robust Demand Management Strategies
There are at least two robust demand management strategies reviewed in Section 3.
44
First, as described in Section 3.3, there are many demand management strategies that
would enable a supply chain to shift demand across products. By having the capability to
shift demand across products, these strategies can make a supply chain more efficient and
resilient. For example, in Section 3.3.1, when facing uncertain demand, Chod and Rudi
(2005) present a responsive pricing strategy that would enable a firm to increase profit by
shifting demand across products. Hence, a responsive pricing strategy would improve
supply chain efficiency. In addition, a responsive pricing strategy could improve supply
chain resiliency as well. For example, when facing a supply disruption of computer parts
from Taiwan after an earthquake, Dell immediately offered special price incentives to
entice their online customers to buy computers that utilized components from other
countries. The capability to shift customer choice swiftly enabled Dell to improve its
earnings in 1999 by 41% even during a supply crunch (c.f., Veverka (1999)).
Besides the responsive pricing strategy described in Section 3.3, the demand
postponement strategy described in Section 3.1. can be a robust demand management
strategy that would enhance supply chain efficiency and resiliency. Under the demand
postponement strategy, a manufacturer may offer price discounts to some retailers to
accept late shipments. Essentially, this strategy is akin to the overbooking situation in
which an airline may offer incentive to entice a fraction of customers who are willing to
take a later flight. By having the capability to shift some of the demands to a later period,
it would certain help a firm to manage both operational risks and disruption risks.
6.2.3. Robust Product Management Strategies
Among the product management strategies reviewed in Section 4, the postponement
strategy described in Section 4.1 is a robust strategy for enhancing the efficiency and the
resiliency of a supply chain. As reported in Lee (1996), postponement is an effective
strategy for improving supply chain efficiency when facing uncertain demands for
different products. In addition, the postponement strategy can increase supply chain
resiliency. For example, after Philip’s semiconductor plant was damaged in a fire in
2000, Nokia was facing a serious supply disruption of radio frequency chips. Since
Nokia’s cell phones are designed according to the modular design concept, Nokia was
able to postpone the insertion of these radio frequency chips until the end of the assembly
process. Due to this postponement strategy, Nokia was able to reconfigure the design of
their basic phones so that the modified phones could accept slightly different chips from
Philip’s other plants and other suppliers. Consequently, Nokia satisfied customer
demand smoothly and obtained a stronger market position. The reader is referred to
Hopkins (2005) for details.
6.2.4. Robust Information Management Strategies
As reported in Section 5, strategies based on information sharing, vendor managed
inventory, or collaborative forecasting and replenishment planning would increase
“supply chain visibility” in the sense that the upstream partners have access to
45
information regarding the demand and inventory position at downstream stages. As
supply chain visibility improves, each supply chain partner can generate more accurate
forecast of future demands and better coordination. We have described various articles
in Section 5 that show how these strategies would enable a supply chain to become more
responsive to customer demand with less inventory and lower cost. Hence, the
information management strategies reported in Section 5 would increase supply chain
efficiency. However, we are unable to find examples that show how these information
management strategies would increase supply chain resiliency. In the absence of
specific examples, we have reasons to believe that the CPFR strategy can enable a supply
chain to develop a production planning system that would improve resiliency. While
Aviv (2005) discuss the mechanism for supply chain partners to generate a common
demand forecast in a collaborative manner, we are not aware of specific models in the
literature that deal with the collaborative replenishment planning. We envision a more
complete CPFR system may improve supply chain resiliency. For example, consider a
CPFR system in which all supply chain partners generate a common demand forecast,
share inventory information, and adopt a common ordering rule that is based on the
“proportional restoration rule” developed by Denardo and Tang (1992). Specifically,
under the proportional restoration rule, the retailer would order Q
r
t
and the manufacturer
would order Q
m
t
in period t, where:
.])()[()(
mr
t
rm
t
mm
t
rr
t
rr
t
ITITdQandITdQ
αα
++=+=
Notice that d represents the common demand forecast, T
p
represents the “target”
inventory position for party p, I
t
p
represents the inventory held at party p at the
beginning of period t, and 0 < α
p
1 represents party p’s restoration factor, where p
= r, m. Denardo and Tang (1992) use various numerical examples to show that this
ordering rule is efficient. In addition, Denardo and Tang (1996) show analytically that
this ordering rule would “restore” the inventory level at each stage to its target even when
the demand forecast d is inaccurate. Thus, one can conclude that such a CPFR system
would improve supply chain efficiency and resiliency.
7. Conclusions
In this paper, we have reviewed various quantitative models for managing supply chain
risks. We found that these quantitative models are designed for managing operational
risks primarily, not disruption risks. However, we argue that some of these strategies
have been adopted by practitioners because these strategies can make a supply chain
become more efficient in terms of handling operational risks and more resilient in terms
of managing disruption risks.
Since there are few supply chain management models for managing disruption risks, we
would like to present six potential ideas for future research.
1. Demand and Supply Process. Virtually all models reviewed in this paper are
based on the assumption that the demand or the supply process is stationary. To
46
model various types of disruptions mathematically, one may need to extend the
analysis to deal with non-stationary demand or supply process. For instance, one
may consider modeling the demand or the supply process as a “jump” process to
capture the characteristics of major disruptions.
2. Objective Function. The performance measures of the models reviewed in this
paper are primarily based on the expected cost / profit. The expected cost / profit
is an appropriate measure for evaluating different strategies for managing
operational risks. When dealing with disruption risks that rarely happen, one may
need to consider alternative objectives besides the expected cost / profit. For
instance, Sharpira (1986) and March and Sharpira (1987) articulated that
managers tend to focus on performance targets. Hence, when developing
strategies for managing supply chain disruption risks, one may consider using
certain performance targets such as recovery time after a disruption. The reader is
referred to Brown and Tang (2005) and the references therein regarding various
alternative performance targets in the context of single-period inventory models.
3. Supply Management Strategies. When developing supply management
strategies for managing disruption risks, both academics and practitioners suggest
the idea of “back-up” suppliers. To capture the dynamics of shifting the orders to
these back-up suppliers when a major disruption occurs, one need to develop a
model for analyzing dynamic supply configurations of suppliers including
contract manufacturers, transportation providers, and distribution channels.
4. Demand Management Strategies. Among the demand management strategies
presented in Section 3, it appears that dynamic pricing / revenue management has
great potential for managing disruption risks because a firm can deploy this
strategy quickly after a disruption occurs. In addition, revenue management looks
promising especially after successful implementations of different revenue
management systems in the airline industry for managing operational risks.
5. Product Management Strategies. When selling products on line, e-tailers can
change their product assortments dynamically according to the supply and
demand of different products. This idea can be extended to brick and mortar
retailers for managing disruption risks. Chong et al. (2001) show that store
manager can manipulate customer’s product choice and customer’s demand by
reconfiguring the set of products on display, the location of each product and the
number of facings of each product. They suggest that one can utilize dynamic
assortment planning to entice customers to purchase certain products that are
widely available (when other products are in short supply).
6. Information Management Strategies. Among the information management
strategies described in Section 6, we think the Collaborative Planning, Forecasting
and Replenishment (CPFR) strategy is promising because it fosters a tighter
coordination and stronger collaboration among supply chain partners. While
Aviv (2005) develops a mechanism for generating collaborative forecasts, there is
47
no model that captures the collaborating replenishing planning. It is conceivable
that the value of a more complete CPFR system is much higher than a system that
is solely based on collaborative forecasting.
While the number of supply chain research articles published over the last 10 years has
grown exponentially on a yearly basis, we believe this trend will continue because there
are plenty of new research areas to explore including strategies for managing supply
chain disruption risks, the impact of RFID technology on supply chain management, and
the impact of government policies on supply chain management. Looking ahead, we
believe that supply chain management will continue to be a fertile research area.
48
8. References
Agrawal, N., and Nahmias, S., “Rationalization of the Supplier Base in the Presence of Yield
Uncertainty,” in Global Supply Chain and Technology Management
, (Eds. H.L. Lee and
S.M. Ng), Production and Operations Management Society Publishers, Florida, 1998.
Akella, R., Rajagopalan, S., and Singh M., “Part Dispatch in Random Yield Multi-Stage Flexible
Test Systems for Printed Circuit Boards,” Operations Research
, Vol. 40, pp. 776-789,
1992.
Anupindi, R., Supply Management Under Uncertainty, PhD thesis, Graduate School of Industrial
Administration, Carnegie Mellon University, 1993.
Anupindi, R., and Akella, R., “An Inventory Model with Commitments,” working paper,
Kellogg School of Management, Northwestern University, 1997.
Anupindi, R., and Akella, R., “Diversification Under Supply Uncertainty,” Management Science,
Vol. 39, No. 8, pp. 944-963, 1993.
Anupindi, R., and Bassok, Y., “Supply Contracts with Quantity Commitments and Stochastic
Demand,” in Quantitative Models for Supply Chain Management, Eds. Tayur et al.,
Kluwer Publisher, 1998.
Anupindi, R., and Bassok, Y., “Centralization of Stocks: retailers vs. manufacturer,”
Management Science, Vol. 45, pp. 178-191, 1999.
Arntzen, B., Brown, G., Harrison, T., and Trafton, L., “Global Supply Chain Management at
Digital Equipment Corporation,” Interfaces, Vol. 25, pp. 69-93, 1995.
Aviv, Y., and Federgruen, A., “The Operational Benefits of Information Sharing and Vendor
Managed Inventory Programs,” working paper, Olin School of Business, Washington
University, St. Louis, 1998a.
Aviv, Y., and Federgruen, A., “The Benefits of Design for Postponement,” in Quantitative
Models for Supply Chain Management, Eds. Tayur et al., Kluwer Publisher, 1998b.
Aviv, Y., and Federgruen, A., “Design for Postponement: A Comprehensive Characterization of
its Benefits Under Unknown Distribution,” Operations Research
, Vol. 49, pp. 578-598,
2001a.
Aviv, Y., and Federgruen, A., “Capacitated Multi-Item Inventory with Random and Seasonal
Fluctuating Demands: Implications for Postponement Strategies,” Management Science,
Vol. 47, pp. 512-531, 2001b.
Aviv, Y., “The Effect of Collaborative Forecasting on Supply Chain Performance,” Management
Science, Vol. 47, pp. 1326-1443, 2001c.
49
Aviv, Y., “Gaining Benefits from Joint Forecasting and Replenishment Processes: The Case of
Auto-Correlated Demand,” Manufacturing & Service Operations Management, Vol. 4,
pp. 55-74, 2002.
Aviv, Y., “Collaborative Forecasting and Its Impact on Supply Chain Performance,” in
Handbook of Quantitative Supply Chain Analysis, Simchi-Levi, D., Wu, D., and Zhen, Z.
Eds., Kluwer Publisher, 2004.
Aviv, Y., “On the Benefits of Collaborative Forecasting Partnerships between Retailers and
Manufacturers,” working paper, Olin School of Management, Washington University,
2005.
Babich, V., Burnetas, A., and Ritchken, P., “Competition and Diversification Effects in Supply
Chains with Supplier Default Risk,” working paper, Department of Industrial and
Operations Engineering, University of Michigan, 2004.
Badinelli, R., “An Optimal, Dynamic Policy for Hotel Yield Management,” European Journal of
Operational Research, Vol. 121, pp. 476-503, 2000.
Bagahana, M.P., and Cohen, M., “The Stabilizing Effect of Inventory in Supply Chains,”
Operations Research, Vol. 46, pp. 572-583, 1998.
Bassok, Y., and Akella, R., “Ordering and Production Decisions with Supply Quality and
Demand Uncertainty,” Management Science, Vol. 37, pp. 1556-1574, 1991.
Bassok, Y., and Anupindi, R., “Analysis of Supply Contracts with Total Minimum
Commitment,” IIE Transactions, Vol. 29, pp. 373-381, 1997.
Bazaraa, M.S., Sherali, H.D., Shetty, C.M., Nonlinear Programming: Theory and Algorithm,
John Wiley & Sons, New York, 1993.
Beamon, B., “Supply Chain Design and Analysis: Model and Methods,” International Journal of
Production Economics, Vol. 55, No. 3, pp. 281-294, 1998.
Billington, C., Lee, H., and Tang, C.S., “Product Rollover: Process, Strategies, and
Opportunities,” Sloan Management Review
, pp. 23-30, Spring, 1998.
Bitran, G., and Dasu, S., “Ordering Policies in an Environment of Stochastic Yields and
Substitutable Products,” Operations Research
, Vol. 50, pp. 999-1017, 1992.
Bitran, G., and Gilbert, S., “Managing Hotel Reservations with Uncertain Arrivals,” Operations
Research, Vol. 44, pp. 35-49, 1996.
Boer, L., Labro, E., Morlacchi, P., “A Review of Methods Supporting Supplier Selection,”
European Journal of Purchasing and Supply Management
, Vol. 7, pp. 75-89, 2001.
50
Boone, Y., Ganeshan, R., and Stenger, A., “The Benefits of Information Sharing in a Supply
Chain: An Exploratory Simulation Study,” in Supply Chain Management Models,
Applications and Research Directions, Geunes, J., Pardalos, P., and Romeijn, E. (Eds.),
Kluwer Publishers, 2002.
Bourland, K.E., Powell, S.G., and Pyke, D., “Exploiting Timely Demand Information to Reduce
Inventories,” European Journal of Operational Research
, Vol. 92, pp. 239-243, 1996.
Bresnahan, T.F., and Reiss, P.C., “Dealer and Manufacturer Margins,” Rand Journal of
Economics, Vol. 16, pp. 253-268, 1985.
Brindley, C., Supply Chain Risk, Ashgate Publishers, Hampshire, England, 2004.
Brown, A. O. and Lee, H.L., “Optimal Pay to Delay Capacity Reservation with Applications to
the Semiconductor Industry,” working paper, Stanford University, 1997.
Brown, A.O., Chou, M., and Tang, C.S., “The Implications of Pooled Returns Policies,” working
paper, UCLA Anderson School, 2005.
Brown, A.O., and Tang, C.S., “The Impact of Alternative Performance Measures on Single-
Period Inventory Policy,” Journal of Industrial and Management Optimization,
forthcoming, 2005.
Brown, D., How U.S. Got Down to Two Makers of Flu Vaccine. Washington Post, October 16,
2004.
Cachon, G., “The Allocation of Inventory Risk and Advanced Purchase Discount in a Supply
Chain,” working paper, The Wharton School, University of Pennsylvania, 2001.
Cachon, G., and Fisher, M., “Campbell Soup’s Continuous Replenishment Program: Evaluation
and Enhanced Inventory Decision Rules,” Production and Operations Management, Vol.
6, pp. 266-276, 1997.
Cachon, G., and Fisher, M., “Supply Chain Inventory Management and the Value of Shared
Information,” Management Science, Vol. 46, pp. 1032-1048, 2000.
Cachon, G., “Supply Chain Coordination with Contracts,” in Handbooks in Operations Research
and Management Science, (Eds. De Kok, A.G., and Graves, S.,), Elsevier Publisher,
2003.
Cachon, G., and Lariviere, M., “Supply Chain Coordination with Revenue Sharing Contracts:
Strengths and Limitations,” Management Science, Vol. 51, pp. 30-44, 2005.
51
Camm, J., Chorman, T., Dull, F., Evans, J., Sweeney, D., and Wegryn G., “Blending OR/MS,
Judgment, and GIS: Restructuring P&G’s Supply Chain,” Interfaces, Vol. 27, No. 1, pp.
128-142, 1997.
Caro, F., and Gallien, J., “Dynamic Assortment with Demand Learning for Seasonal Consumer
Goods,” working paper, UCLA Anderson School of Management, 2005.
Carr, S., and Lovejoy, W., “The Inverse Newsvendor Problem: Choosing an Optimal Demand
Portfolio for Capacitated Resources,” Management Science
, Vol. 47, pp. 912-927, 2000.
Cetinkaya, S., and Lee, C.Y., “Stock Replenishment and Shipment Scheduling for Vendor
Managed Inventory,” Management Science
, Vol. 46, pp. 217-232, 2000.
Chod, J., and Rudi, N., “Resource Flexibility with Responsive Pricing,” Operations Research
,
Vol. 53, pp. 532-548, 2005.
Chen, F., Drezner, A., Ryan, J., and Simchi-Levi, D., “The Bullwhip Effect: Managerial Insights
on the Impact of Forecasting and Information on Variability in a Supply Chain,” in
Quantitative Models for Supply Chain Management
, edited by Tayur et al., Kulwer
Publishers, 1998.
Chen, F., A. Ryan, and Simichi-Levi, D., “The Impact of Exponential Smoothing Forecasts on
the Bullwhip Effect,” Naval Research Logistics, Vol. 47, pp, 269—286, 2000a.
Chen, F., A. Ryan, and Simichi-Levi, D., “Quantifying the Bullwhip Effect in a Simple Supply
Chain: The Impact of Forecasting, Lead times and Information,” Management Science,
Vol. 46, pp. 436--443, 2000b.
Cheng, T.C.E., and Wu, Y.N., “The Impact of Information Sharing in a Two-Level Supply Chain
with Multiple Retailers,” Journal of the Operational Research Society, Vol. 56, pp. 1159-
1165, 2005.
Chong, J.K., Ho, T.H., and Tang, C.S., “A Modeling Framework for Category Assortment
Planning,” Manufacturing & Service Operations Management
, Vol. 3, pp. 141-160, 2001.
Chong, J.K., Ho, T.H., and Tang, C.S., “Demand Modeling in Product Line Trimming:
Substitutability and Variability,” in Managing Business Interfaces: Marketing,
Engineering, and Manufacturing Perspectives, Chakravarty, A.K. and Eliashberg, J.
(Eds.), Kluwer Academic Publishers, 2004.
Chopra S., and Sodhi M., “Avoiding Supply Chain Breakdown,” Sloan Management Review
,
Fall, pp. 53-62, Vol. 46, 1, 2004.
Christopher M., Logistics and Supply Chain Management, Pitman, London, 1992.
52
Christopher M., “Creating Resilient Supply Chains,” Logistics Europe, pp. 14-21, February,
2004.
Ciarallo, F., Akella, R., Morton, T., “A Periodic Review Production Planning Model with
Uncertain Capacity and Uncertain Demand,” Management Science, Vol. 40, pp. 320-332,
1994.
Closs D., and McGarrell E., Enhancing Security Through the Supply Chain. IBM Center for
the Business of Government Special Report Series, April, 2004.
Cohen, M. A., and Mallik, S., “Global Supply Chains: Research and Applications,” Production
and Operations Management, Vol. 6, pp. 193-210, 1997.
Cohen M. A., and Agrawal, N., “An Analytical Comparison of Long and Short Term Contracts,”
IIE Transactions, Vol. 31, 8, pp. 783-796, 1999.
Corbett, C., “Stochastic Inventory Systems in a Supply Chain with Asymmetric Information:
Cycle Stocks, Safety Stocks, and Consignment Stocks,” Operations Research, Vol. 49,
pp. 487-500, 2001.
Corbett, C., and de Groote, X., “A Supplier’s Optimal Quantity Discount Policy Under
Asymmetric Information,” Management Science, Vol. 46, pp. 444-450, 2000.
Corbett, C., and Tang, C.S., “Designing Supply Contracts: Contract Type and Information
Asymmetric Information,” in Quantitative Models for Supply Chain Management, Eds.
Tayur et al., Kluwer Publisher, 1998.
Crew, M, and Kleindorfer, P., The Economics of Public Utility Regulation, MIT Press,
Cambridge, MA, 1986.
Current, J.R., and Weber, C.A., “Application of Facility Location Modeling Constructs to
Vendor Selection Problems,” European Journal of Operational Research
, Vol. 76, pp.
387-392, 1994.
Dahel, N., “Vendor Selection and Order Quantity Allocation in Volume Discount
Environments,” Supply Chain Management: An International Journal, Vol. 8, pp. 335-
342, 2003.
Dana, J., “Advance-Purchase Discounts and Price Discrimination in Competitive Markets,”
Journal of Political Economy, Vol. 166, pp. 395-422, 1998.
Dana, J., “Using Yield Management to Shift Demand When the Peak Time is Unknown,” Rand
Journal of Economics, Vol. 30, pp. 456-474, 1999.
Dana J., and Spier, K., ``Revenue Sharing and Vertical Control in the Video Rental Industry,''
Journal of Industrial Economics
, Vol. 59, pp. 223-245, 2001.
53
Dapiran, P., “Benetton: Global Logistics in Action,” Asian Pacific International Journal of
Business Strategy, Vol. 5, pp. 7-11, 1992.
Dasu, S., and Li, L., “Optimal Operating Policies in the Presence of Exchange Rate Variability,”
Management Science, Vol. 43, pp. 705-721, 1997.
Denardo, E., and Lee, T., “Managing Uncertainty in a Serial Production Line," Operations
Research, Vol. 44, 382-392, 1996.
Denardo, E., and Tang, C.S., “Pulling a Markov Production System,” Operations Research, Vol.
40, pp. 259-278, 1992.
Denardo, E., and Tang, C.S., “Control of a Stochastic Production System with Estimated
Parameters,” Management Science, Vol. 43, pp. 1296-1307, 1997.
Donohue, K., “Efficient Supply Contracts for Fashion Goods with Forecast Updating and Two
Production Modes", Management Science, Vol. 46, 11, 1397-1411, 2000.
Disney, S.M., and Towill, D.R., “The Effect of Vendor Managed Inventory (VMI) Dynamics on
the Bullwhip Effect in Supply Chains,” International of Production Economics, Vol. 85,
pp. 199-215, 2003.
Dyer, J., “How Chrysler Created an American Keiretsu,” Harvard Business Review, July-
August, pp. 42-56, 1996.
Dyer, J., and Ouchi, W., “Japanese Style Partnerships: Giving Companies a Competitive Edge,”
Sloan Management Review, vol. 34, No. 6, pp. 51-63, 1996.
Eliashberg, J., and Steinberg, R., “Marketing-Production Joint Decisions,” in Handbook of
Operations Research and Management Science: Marketing, Eds. Eliashberg, J., and
Lilien, G., Elsevier Publishers, Amsterdam, The Netherlands, 1993.
Ellram, L., “The Supplier Selection Decision in Strategic Partnerships”, Journal of
Purchasing and Materials Management, Fall, pp. 8-14, 1990.
Elmaghraby, W., and Keskinocak, P., “Dynamic Pricing in the Presence of Inventory
Considerations: Research Overview, Current Practices, and Future Directions,”
Management Science, Vol. 49, pp. 1287-1309, 2003.
Emmons, H., and Gilbert, S., “Returns Policies in Pricing and Inventory Decisions for Catalogue
Goods,” Management Science
, Vol. 44, pp. 276-283, 1998.
54
Eppen, G., and Schrage, L., “Centralized Ordering Policies in a Multi-warehouse System with
Lead Times and Random Demand,” in: Schwarz, L. B., (ed.) Multi-Level
Production/Inventory Control Systems: Theory and Practice, North-Holland, Amsterdam,
1981.
Eppen, G., and Iyer, A.V., “Backup Agreements in Fashion Buying: The Value of Upstream
Flexibility,” Management Science, Vol. 43, pp. 1469-1484, 1997a.
Eppen, G., and Iyer, A.V., “Improved Fashion Buying with Bayesian Updates,” Operations
Research, Vol. 45, pp. 805-819, 1997b.
Ernst, R., and Kouvelis, P., “The Effects of Selling Packaged Goods on Inventory Decisions,”
Management Science, Vol. 45, pp. 1142-1155, 1999.
Federgruen, A., “Comments on: Variability Reduction through Operations Reversal in Supply
Chain Re-engineering,” working paper, Columbia University, 1998.
Federgruen, A., and Zipkin, P., “Approximations of Dynamic, Multi-location Production and
Inventory Problems,” Management Science, Vol. 30, pp. 69-84, 1984.
Fischhoff, B., Lichtenstein, S., Slovic, P., Derby S., and Keeney, R., Acceptable Risk,
Cambridge University Press, New York, 1981.
Fisher, M., and Raman, A., “Reducing the Cost of Demand Uncertainty through Accurate
Response,” Operations Research, Vol. 44, pp. 87-99, 1996.
Fisher, M., “What is the Right Supply Chain for your Product?” Harvard Business Review, pp.
105-116, 1997.
Fukuda, Y., “Optimal Policies for the Inventory Problem with Negotiable Lead time,”
Management Science
, Vol. 10, pp. 690-708, 1964.
Garg, A., Product and Process Design Strategies for Effective Supply Chain Management.
Unpublished PhD thesis, Stanford University, 1995.
Garg, A., and Lee, H.L., “Effective Postponement Through Standardization and Process
Sequencing,” IBM Research Report RC 20726, IBM T.J. Watson Research Center,
Yorktown Heights, NY 10598, 1996.
Garg, A., and Tang, C.S., “Postponement Strategies for Product Families with Multiple Points of
Differentiation,” IIE Transactions, Vol. 29, pp. 641-650, 1997.
Garg, A., and Lee, H.L., “Managing Product Variety: An Operations Perspective,” in
Quantitative Models for Supply Chain Management, Eds. Tayur et al., Kluwer Publisher,
1998.
55
Gaur, V., Giloni, A., and Seshadri, S., “Information Sharing in a Supply Chain Under ARMA
Demand,” Management Science, Vol. 51, pp. 961-969, 2005.
Gavirneni, S., Kapuscinski, R., and Tayur, S., “Value of Information in Capacitated Supply
Chains,” Management Science, Vol. 45, pp. 16-24, 1999.
Geunes, J., Pardalos, P., and Romeijn, E., Supply Chain Management: Models, Applications, and
Research Directions, Kluwer Publihser, 2002.
Gerchak, Y., Vickson R., and Parlar, M., “Periodic Review Production Models with Variable
Yield and Uncertain Demand,” IIE Transactions, Vol. 20, No. 2, pp. 144-150, 1988.
Gilbert, K., “An ARIMA Supply Chain Model,” Management Science, Vol. 51, pp. 305-310,
2005.
Gilbert, S., and Ballou, R., “Supply Chain Benefits from Advanced Customer Commitments,”
Journal of Operations Management, Vol. 18, pp. 61-73, 1999.
Gong, L. and Matsuo, H., "Control Policy for a Manufacturing System with Random Yield and
Rework," Journal of Optimization Theory and Applications, 95 (1), 149-175, 1997.
Gunasekaran, A., and Ngai, E., “Build-To-Order Supply Chain Management: A Literature
Review and Framework for Development,” Journal of Operations Management, Vol. 23,
pp. 423-451, 2005.
Gupta, D., and Benjaafar, S., “Make-to-Order, Make-to-Stock, or Delay Product Differentiation?
A Common Framework for Modeling and Analysis,” IIE Transactions, Vol. 36, pp. 529-
546, 2004.
Gurnani, H., and Tang, C.S., “Note: Optimal Ordering Decisions with Uncertain Cost and
Demand Forecast Updating,” Management Science
, Vol. 45, pp. 1456-1462, 1999.
Ha, A., “Supplier-Buyer Contracting: Asymmetric Cost Information and Cut-Off Level Policy
for Buyer Participation,” Naval Research Logistics, Vol. 48, pp. 41-64, 2001.
Helper, S., “How Much Has Really Changed Between U.S. Automakers and Their Suppliers,”
Sloan Management Review, Vol. 32, No. 4, pp. 15-28, 1991.
Hendricks, K., and Singhal, V., “An Empirical Analysis of the Effect of Supply Chain
Disruptions on Long-run Stock Price Performance and Equity Risk of the Firm,”
Production and Operations Management
, Spring, pp. 25-53, 2005.
Hernig, M., and Gerchak, Y., “The Structure of Periodic Review Policies in the Presence of
Variable Yield,” Operations Research, Vol. 38, pp. 634-643, 1990.
56
Ho, T., and Tang, C.S., Product Variety Management: Research Advances, Kluwer Publisher,
1998.
Hopkins, K., “Value Opportunity Three: Improving the Ability to Fulfill Demand,” Business
Week, January 13, 2005.
Hsu , A., and Bassok, Y., “Random yield and random demand in a production system with
downward substitution,” Operations Research, Vol. 47, pp. 277–290, 1999.
Huchzermeier, A., and Cohen, M., “Valuing Operational Flexibility Under Exchange Rate Risk,”
Operations Research, Vol. 44, 100-113, 1996.
Iyer, A., “Modeling the Impact of Information on Inventories,” in Quantitative Models for
Supply Chain Management, Eds. Tayur et al., Kluwer Publisher, 1998.
Iyer, A., and Bergen, M.E., “Quick Response in Manufacturing-Retailer Channels,” Management
Science, Vol. 43, pp. 559-570, 1997.
Iyer, A., Deshpande, V., and Wu Z., “A Postponement Model for Demand Management,”
Management Science, Vol. 49, pp. 983-1002, 2003.
Jain, N., and Paul, A., “A Generalized Model of Operations Reversal for Fashion Goods,”
Management Science, Vol. 47, pp. 595-600, 2001.
Janssen, F., de Kok, T., “A Two-Supplier Inventory Model,” International Journal of Production
Economics, Vol. 59, pp. 395-403, 1999.
Johnson, M.E., Davis, T., and Walker, M., “Vendor Managed Inventory in the Retail Supply
Chain,” Journal of Business Logistics, Vol. 20, pp. 183-203, 1999.
Kapuscinski, R., and Tayur, S., “Variance vs. Standard Deviation – Note: On Variability
Reduction Through Operations Reversal in Supply Chain Re-engineering,” Management
Science, Vol. 45, pp. 21-23, 1999.
Kleindorfer, P., and G Saad, “Managing Disruption Risks in Supply Chains,” Production and
Operations Management, Vol. 14, pp. 53-68, 2005.
Kogut, B., “Designing Global Strategies: Profiting from Operational Flexibility,” Sloan
Management Review, Fall, pp. 27-38, 1985.
Kogut, B., and Kulatilaka, N., “Operating Flexibility, Global Manufacturing, and the Option
Value of a Multinational Network,” Management Science
, Vol. 40, pp. 123-139, 1994.
Kouvelis, P., “Global Sourcing Strategies Under Exchange Rate Uncertainty,” in Quantitative
Models for Supply Chain Management, (Eds. Tayur et al.), Kluwer Publishers, 1998.
57
Kouvelis, P., and Gutierrez, G., “The Newsvendor Problem in a Global Market: Optimal
Centralized and Decentralized Control Policies for a Two-Market Stochastic Inventory
System,” Management Science, Vol. 43, pp. 571-585, 1997.
Kouvelis, P., and Rosenblatt, M., “A Mathematical Programming Model for Global Supply
Chain Management: Conceptual Approach and Managerial Insights,” in Supply Chain
Management: Models, Applications, and Research Directions, (Eds. Geunes et al.),
Kluwer Publisher, 2000.
Krishnan, M., Kekre, S., Mukhopadhyay, T., and Srinivasan, K., “Managing Variety in Software
Industry,” in Product Variety Management: Research Advances, Eds. Ho, T.H., and
Tang, C.S., Kluwer Publisher, 1998.
Kunreuther, H., “Limited Knowledge and Insurance Protection,” Public Policy, Vol. 24, pp. 227-
261, 1976.
Lariviere, M., “Supply Chain Contracting and Co-ordination with Stochastic Demand,”
in Quantitative Models for Supply Chain Management, Eds. Tayur et al., Kluwer
Publisher, 1998.
Lariviere, M., and Porteus, E., “Selling to the Newsvendor: An Analysis of Price-only
Contracts,” Manufacturing and Service Operations Management, Vol. 3, pp. 293-305,
2001.
Lazear, E. P., “Retail Pricing and Clearance Sales,” American Economic Review, Vol. 76, pp.
14-32, 1986.
Lee, H.L., “Effective Management of Inventory and Service Through Product and Process
Redesign,” Operations Research, Vol. 44, pp. 151-159, 1996.
Lee, H.L., and Tang, C.S., “Modeling the Costs and Benefits of Delayed Product
Differentiation,” Management Science
, Vol. 43, pp. 40-53, 1997a.
Lee, H.L., Padmanabhan, V., and Whang, S., “Information Distortion in a Supply Chain: The
Bullwhip Effect,” Management Science, Vol. 43, pp. 546-548, 1997b.
Lee, H.L., Padmanabhan, V., and Whang, S., “The Bullwhip Effect in Supply Chains,” Sloan
Management Review, Vol. 38, pp. 93-102, 1997c.
Lee, H., and Tang, C.S., “Variability Reduction Through Operations Reversal,” Management
Science, Vol. 44, pp. 162-173, 1998a.
Lee, H.L, and Tang, C.S., “Managing Supply Chains with Contract Manufacturing,” in Global
Supply Chain and Technology Management, (Eds. H.L. Lee and S.M. Ng), Production
and Operations Management Society Publishers, Florida, 1998b.
58
Lee, H.L., and Whang, S., “Decentralized Multi-Echelon Supply Chains: incentives and
information,” Management Science, Vol. 45, pp. 633-640, 1999.
Lee, H.L., So, K.C., Tang, C.S., “The Value of Information Sharing in a Two-Level Supply
Chain,” Management Science, Vol. 46, pp. 626-643, 2000.
Lee, H., “The Triple-A Supply Chain,” Harvard Business Review, pp. 102-112, October, 2004.
Levy, D., “International Sourcing and Supply Chain Stability,” Journal of International Business
Studies, Vol. 26, pp. 343-360, 1995.
Li, G., Wang S., Yan, H., and Yu, G., “Information Transformation in a Supply Chain: A
Simulation Study,” Computers and Operations Research, Vol. 32, pp. 707-725, 2005.
Li, C.L., and Kouvelis, P., “Flexible and Risk-Sharing Supply Contracts Under Price
Uncertainty,” Management Science, Vol. 45, pp. 1378-1398, 1999.
Lilien, G., Kotler, P., and Moorthy, S., Marketing Models, Prentice Hall, New Jersey, 1992.
Lim, W.S., and Tang, C.S., “Optimal Product Rollover Strategies,” forthcoming, European
Journal of Operational Research, 2005.
MacCrimmon K. R., and Wehrung, D.A., Taking Risks: The Management of Uncertainty, Free
Press, New York, 1986.
MacDuffie, J.P., Sethuraman, K., and Fisher, M., “Product Variety and Manufacturing
Performance: Evidence from the International Automotive Assembly Plant Study,”
Management Science, pp. 350-369, 1996.
Mandal, A., Deshmukh, S. G., “Vendor Selection using Interpretive Structural Modeling,”
International Journal of Operations and Production Management
, Vol. 14, No. 6, pp. 52-
59, 1996.
March, J., and Sharpira Z., “Managerial Perspectives on Risk and Risk Taking,” Management
Science, Vol. 33, pp. 1404 – 1418, 1987.
Martin, M., Hausman, W., Ishii, K., “Design for Variety,” in Product Variety Management:
Research Advances, Eds. Ho, T.H., and Tang, C.S., Kluwer Publisher, 1998.
McCardle, K., Rajaram, K., and Tang, C.S., "Advance Booking Discount Programs under Retail
Competition," Management Science
, Vol. 50, 5, pp. 701-708, 2004.
McCardle, K., Rajaram, K., and Tang, C.S., “Bundling Retail Products: Models and Analysis,”
Forthcoming, European Journal of Operational Research, 2005.
59
McFarlan, W., “Li and Fung (A): Internet Issues,” Harvard Business School case 9-301-009,
2002.
McGill, J., and Van Ryzin, G., “Revenue Management: Research Overview and Prospects,”
Transportation Science, Vol. 33, pp. 233-256, 1999.
Minner, S., “Multiple-Supplier Inventory Models in Supply Chain Management: A review,”
International Journal of Production Economics, pp. 265-279, 2003.
Moinzadeh, K., and Nahmias, S., “A Continuous Review Model for an Inventory System with
Two Supply Modes,” Management Science, Vol. 34, pp. 761-773, 1988.
Mortimer, J., ``Vertical Contracts in the Video Rental Industry,'' working paper, Department of
Economics, Harvard University, 2004.
Naguney, A., Curz, J., Dong, J., Zhang, D., “Supply Chain Networks, Electronic Commerce, and
Supply Side and Demand Side Risk,” European Journal of Operational Research, Vol.
164, pp. 120-142, 2005.
Padmanabhan, V., and Png, I., “Manufacturer’s Returns Policy and Retail Competition,”
Marketing Science, Vol. 16, pp. 81-94, 1997.
Parlar, M., and Goyal, S., “Optimal Ordering Decisions for Two Substitutable Products with
Stochastic Demand,” OPSEARCH, Vol. 21, pp. 1-15, 1984.
Parlar, M., and Perry, D., “Inventory Models of Future Supply Uncertainty with Single and
Multiple Suppliers,” Naval Research Logistics, Vol. 43, pp. 191-210, 1996.
Pasternack, B., “Optimal Pricing and Return Policies for Perishable Commodities,” Marketing
Science, Vol. 4, pp. 166-176, 1985.
Pasternack, B., “Using Revenue Sharing to Achieve Channel Coordination for a Newsboy Type
Inventory Model,” in Supply Chain Management Models, Applications and Research
Directions, Geunes, J., Pardalos, P., and Romeijn, E. (Eds.), Kluwer Publishers, 2002.
Paulsson, U., “Supply Chain Risk Management,” in Brindley, C., editor, Supply Chain Risk,
Ashgate Publishers, Hampshire, England, 2004.
Petruzzi, N., and Dada, M., “Pricing and the Newsvendor Problem: A Review with Extensions,”
Operations Research, Vol. 47, pp. 183-194, 1999.
Petruzzi, N., and Dada, M., “Information and Inventory Recourse for a Two-Market, Price-
Setting Retailers,” Manufacturing & Service Operations Management
, Vol. 3, pp. 242-
263, 2001.
Porter, M., Competitive Advantage
, The Free Press, New York, 1985.
60
Porteus, E., Foundations of Stochastic Inventory Theory, Stanford University Press, 2002.
Quelch, J., and Kenny, D., “Extend Profits, Not the Product Lines,” Harvard Business Review
,
Vol. 72, pp. 153-160, 1994.
Raghunathan, S., “Information Sharing in a Supply Chain: A Note on its Value when Demand is
Non-Stationary,” Management Science
, Vol. 47, pp. 605-610.
Rajaram, K., and Tang, C.S., “The Impact of Product Substitution on Retail Mechandising,”
European Journal of Operational Research, Vol. 135, pp. 582-601, 2001.
Raju, J., Sethuraman, R., and Dhar, S., “The Introduction and Performance of Store Brands,”
Management Science
, pp. 957-978, 1995.
Raman, A., “Managing Inventories for Fashion Products,” in Quantitative Models for Supply
Chain Management, Eds. Tayur et al., Kluwer Publisher, 1998.
Ramasesh, R., Ord, J., Hayya, J., and Pan A., “Sole versus Dual Sourcing in Stochastic Lead-
time (s, Q) Inventory Models,” Management Science, Vol. 37, pp. 428-443, 1991.
Ramdas, K., “Managing Product Variety: An Integrative Review and Research Directions,”
Production and Operations Management, Vol. 12, pp. 79-101, 2003.
Repenning, N., and Sterman, J., “Nobody Ever Gets Credit for Fixing Problems that Never
Happened,” California Management Review, pp. 64-88, Vol. 43, summer, 2001.
Rice, B., and Caniato, F., Supply Chain Response to Terrorism: Creating Resilient and Secure
Supply Chains. Supply Chain Response to Terrorism Project Interim Report, MIT Center
for Transportation and Logistics, MIT, Massachusetts.
Ritchie R.L., and Brindley, C.S., “The Information – Risk Conundrum,” Journal of Marketing
Intelligence and Planning, Vol. 19, No. 1, pp. 29-37, 2001.
Sahin, F., and Robinson Jr., E.P., “Information Sharing and Coordination in Make-To-Order
Supply Chains,” Journal of Operations Management, Vol. 23, pp. 579-598, 2005.
Scheller-Wolf, A., and Tayur, S., “Managing Supply Chains in Emerging Markets,” in Tayur et
al. (Eds.), Quantitative Models for Supply Chain Management
, Kluwer Publishers, pp.
703-735, 1999.
Sedarage, D., Fujiwara, O., and Luong, H., “Determining Optimal Order Splitting and Reorder
Level for n-supplier Inventory Systems,” European Journal of Operational Research
, Vol.
116, pp. 389-404, 1999.
61
Sharpira, Z., “Risk in Managerial Decision Making,” unpublished manuscript, Hebrew
University, 1986.
Shavell, S., “A Model of the Optimal Use of Liability and Safety Regulation,” Rand Journal of
Economics, Vol. 15, No. 2, pp. 271-280, 1984.
Sheffi, Y., “Supply Chain Management under the Threat of International Terrorism,”
International Journal of Logistics Management, Vol. 12, No. 2, pp. 1 –11, 2001.
Shin, H., Collier, D., and Wilson, D., “Supply Management Orientation and Supplier/Buyer
Performance,” Journal of Operations Management, Vol. 18, pp. 317-333, 2000.
Signorelli, S., and Heskett, J.L., “Benetton,” Harvard Business School Case # 685020, 1984.
Smith, C., Gilbert, S., and Burnetas, A., “Partial Quick Response Policies in a Supply Chain,” in
Supply Chain Management Models, Applications and Research Directions, Geunes, J.,
Pardalos, P., and Romeijn, E. (Eds.), Kluwer Publishers, 2002.
Sterman, J.D., “Modeling Managerial Behavior: Misperceptions of Feedback in a Dynamic
Decision Making Experiment,” Management Science, Vol. 35, pp. 321-339, 1989.
Stremersch, S., and Tellis, G.J., “Strategic Bundling of Products and Prices: A New Synthesis for
Marketing,” Journal of Marketing, Vol. 66, pp. 55-72, 2002.
Su, J., Chang, Y.L., and Ferguson, M., “Evaluation of Postponement Structures to Accommodate
Mass Customization,” Journal of Operations Management, Vol. 23, pp. 305-318, 2005.
Swaminathan, J., and Tayur, S., “Stochastic Programming Models for Managing Product
Variety,” in Quantitative Models for Supply Chain Management, Eds. Tayur et al.,
Kluwer Publisher, 1998a.
Swaminathan, J., and Tayur, S., “Managing Broader Product Lines Through Delayed
Differentiation Using Vanilla Boxes,” Management Science, Vol. 44, pp. S161-S172,
1998b.
Swaminathan, J, and Tayur, S., “Managing Design of Assembly Sequences for Product Lines
that Delay Product Differentiation,” IIE Transactions, Vol. 31, pp. 1015-1026, 1999.
Talluri, K., and Van Ryzin, G., The Theory and Practice of Revenue Management
, Kluwer
Publisher, 2005.
Tagaras, G., and Lee, H., “Economic Models for Vendor Evaluation with Quality Cost
Analysis,” Vol. 42, No. 11, pp. 1531-1543, 1996.
Tang, C.S., “The Impact of Uncertainty on a Production Line,” Management Science
, Vol. 38,
pp. 1518-1531, 1990.
62
Tang, C.S., “Supplier Relationship Map,” International Journal of Logistics: Research and
Applications,” Vol. 2, No. 1, pp. 39-56, 1999.
Tang, C.S., Rajaram, K., Alptekinoglu, A., and Ou, J.H., "The Benefits of Advance Booking
Discount Programs: Model and Analysis," Management Science, Vol. 50, 4, pp. 465-
478, 2004.
Tang, C.S., and Deo, S., “Rental Price and Rental Duration under Retail Competition,” working
paper, UCLA Anderson School, 2005.
Tang, C.S., “Robust Strategies for Mitigating Supply Chain Risks,” International Journal of
Logistics, forthcoming, 2005.
Tang, K., “An Economic Model for Vendor Selection,” Journal of Quality Technology
, Vol. 20,
pp. 81-89, 1988.
Tayur, S., Ganeshan, R., and Magazine, M., Quantitative Models for Supply Chain Management,
Kluwer Publisher, 1998.
Terwiesch, C., Ren, Z.J., Ho, T.H., and Cohen, M.A., “An Empirical Analysis of Forecast
Sharing in the Semiconductor Equipment Supply Chain,” Management Science, Vol. 51,
pp. 208-220, 2005.
Thonemann, U., “Improving Supply-Chain Performance by Sharing Advance Demand
Information,” European Journal of Operational Research, Vol. 142, pp. 81-107, 2002.
Tirole, J., The Theory of Industrial Organization, The MIT Press, Cambridge, MA, 1988.
Treece, J.B., “Just too much Single Sourcing Spurs Toyota Purchasing Review,” Automotive
News, March 3, p. 3, 1997.
Tsay, A.A., Nahmias, S., Agrawal, N., “Modeling Supply Chain Contracts: A Review,” in
Quantitative Models for Supply Chain Management, Eds. Tayur et al., Kluwer Publisher,
1998.
Tsay, A.A., and Lovejoy, W., “Quantity Flexibility Contracts and Supply Chain Performance,”
Manufacturing and Service Operations Management, Vol. 1, 2, pp. 89-111, 1999.
Ulrich, K., Randall, T., Fisher, M., and Reibstein, D., “Managing Product Variety,” in Product
Variety Management: Research Advances, Eds. Ho, T.H., and Tang, C.S., Kluwer
Publisher, 1998.
Van Mieghem, J., and Dada, M., “Price Versus Production Postponement: Capacity and
Competition,” Management Science
, Vol. 45, pp. 1631-1649, 2001.
63
Veverka, M., “A DRAM Shame,” Barron’s, October 15, pp. 15, 1999.
Vlachos D., and Tagaras, G., “An Inventory System with Two Supply Modes and Capacity
Constraints,” International Journal of Production Economics
, Vol. 72, pp. 41-58, 2001.
Vokurka, R.J., Choobineh, J., Vadi, L., “A Prototype Expert System for the Evaluation and
Selection of Potential Suppliers,” International Journal of Operations and Production
Management, Vol. 16, No. 12, pp. 106-127, 1996.
Wang Y., and Gerchak, Y., “Periodic Review Production Models with Variable Capacity,
Random Yield, and Uncertain Demand,” Management Science, Vol. 42, No. 1, pp. 130-
137, 1996.
Weatherford, L.R., and Bodily, S., “A Taxonomy and Research Overview of Perishable-Asset
Revenue Management: Yield Management, Overbooking and Pricing,” Operations
Research, Vol. 40, pp. 841-844, 1992.
Weber, C.A., Current, J.R., “Multi-objective Analysis of Vendor Selection,” European Journal of
Operational Research, Vol. 68, pp. 173-184, 1993.
Weber, C.A., Current, J.R., Desai, A., “An Optimisation Approach to Determining the Number
of Vendors to Employ,” Supply Chain Management: An International Journal, Vol. 5, pp.
90-98, 2000.
Weng, K., and Parlar, M., “Integrating Early Sales with Production Decisions: Analysis and
Insights,” IIE Transactions, Vol. 31, pp. 1051-1060, 1999.
Weng, K., and Parlar, M., “Managing Build-To-Order Short Life-Cycle Products: Benefits of
Pre-Season Price Incentives with Standardization,” Journal of Operations Management,
Vol. 23, pp. 482-495, 2005.
Whang, S., and Lee, H.L., “Value of Postponement,” in Product Variety Management: Research
Advances, Eds. Ho, T., and Tang, C.S., Kluwer Publisher, 1998.
Xie, J., and Shugan, S., “Electronic Tickets, Smart Cards, and Online Prepayments: When and
How to Advance Sell,” Marketing Science
, Vol. 20, pp. 219-243, 2001.
Yano, C., and Lee, H.L., ““Lot Sizing with Random Yields: A Review,” Operations Research
,
Vol. 43, pp. 311-334, 1993.
Yano, C., and Gilbert, S., “Coordinated Pricing and Production/Procurement Decisions: A
Review,” in Managing Business Interfaces: Marketing, Engineering, and Manufacturing
Perspectives, Chakravarty, A.K. and Eliashberg, J. (Eds.), Kluwer Academic Publishers,
2004.
64
Yang, B., Burns, N., and Backhouse, C., “Postponement: A Review and An Integrated
Framework,” International Journal of Operations and Production Management, Vol. 24,
pp. 468-487, 2004.
Yeh, C., and Yang C., “A cost model for determining dyeing postponement in garment supply
chain “, The International Journal of Advanced Manufacturing Technology, Vol. 22,
pp.134-140, 2003.
Zhang, V.L., “Ordering Policies for an Inventory System with Three Supply Modes,” Naval
Research Logistics, Vol. 43, pp. 691-708, 1996.
Zhang, X., “The Impact of Forecasting Methods on the Bullwhip Effect,” International Journal
of Production Economics, Vol. 88, pp. 15-27, 2004.
Zhao, X., Xie, J., and Leung, J., “The Impact of Forecasting Model Selection on the Value of
Information Sharing in a Supply Chain,” European Journal of Operational Research, Vol.
142, pp. 321-344, 2002.
Zipkin, P., Foundations of Inventory Management, McGraw-Hill Publishers, 2000
Zsidisin, G., Panelli, A. and Upton, R., “Purchasing Organization Involvement in Risk
Assessments, Contingency Plans, and Risk Management: An Exploratory Study,”
Supply Chain Management: An International Journal, pp. 187-197, 2001.
Zsidisin, G., Ragatz, G., and Melnyk, S., “Effective Practices and Tools for Ensuring Supply
Continuity,” in Brindley, C., editor, Supply Chain Risk, Ashgate Publishers, Hampshire,
England, 2004a.
Zsidisin, G., Ellram, L., Carter, J., and Cavinato, J., “An Analysis of Supply Risk Assessment
Techniques,” International Journal of Physical Distribution and Logistics Management,
Vol. 34, No. 5, pp. 397-413, 2004b.