assortment planning with inventory considerations in

126
Assortment Planning with Inventory Considerations in Supply Chains by LEI XIE Department of Business Administration Duke University Date: Approved: Fernando Bernstein, Co-supervisor urhan K¨ ok, Co-supervisor Jeannette Song Serhan Ziya Dissertation submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in the Department of Business Administration in the Graduate School of Duke University 2010

Upload: others

Post on 23-Mar-2022

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Assortment Planning with Inventory Considerations in

Assortment Planning with Inventory

Considerations in Supply Chains

by

LEI XIE

Department of Business AdministrationDuke University

Date:

Approved:

Fernando Bernstein, Co-supervisor

Gurhan Kok, Co-supervisor

Jeannette Song

Serhan Ziya

Dissertation submitted in partial fulfillment of the requirements for the degree ofDoctor of Philosophy in the Department of Business Administration

in the Graduate School of Duke University2010

Page 2: Assortment Planning with Inventory Considerations in

Abstract(Warlockery)

Assortment Planning with Inventory Considerations in

Supply Chains

by

LEI XIE

Department of Business AdministrationDuke University

Date:

Approved:

Fernando Bernstein, Co-supervisor

Gurhan Kok, Co-supervisor

Jeannette Song

Serhan Ziya

An abstract of a dissertation submitted in partial fulfillment of the requirements forthe degree of Doctor of Philosophy in the Department of Business Administration

in the Graduate School of Duke University2010

Page 3: Assortment Planning with Inventory Considerations in

Copyright c© 2010 by LEI XIEAll rights reserved except the rights granted by the

Creative Commons Attribution-Noncommercial Licence

Page 4: Assortment Planning with Inventory Considerations in

Abstract

The dissertation addresses optimization problems related to the management of prod-

uct variety with inventory considerations. The objective is to provide insights and

tools to help companies determine the optimal level of product variety they should

produce/stock and maximize revenues by strategically choosing assortments offered

to customers. The dissertation comprises of three chapters. The first chapter ex-

plores the role of component commonality in product assortment decisions. We study

how the use of component commonality in a production system affects the size of

the optimal assortment and the resulting optimal component stocking decisions. In

particular, we characterize the structure of the optimal assortment under various

configurations of the bill of materials and provide conditions that establish when

the optimal variety level increases or decreases relative to a system without common

components. We find that the effect of commonality on profit and on the level of

variety is stronger when demand exhibits negative correlation. The second chapter

explores dynamic assortment customization strategies with limited inventories, in

the contexts of a revenue management problem for a firm offering an assortment of

products. The goal of the chapter is to determine a dynamic assortment policy in a

setting with limited inventories and heterogeneous consumer preferences. The results

indicate that limiting the choice set to some customers, in consideration of the pre-

vailing inventory levels and the customers’ preferences, leads to increased revenues.

We find that, under the optimal dynamic assortment policy, each customer is al-

iv

Page 5: Assortment Planning with Inventory Considerations in

ways offered its most preferred product variant, while it is also offered other product

variants provided that their inventory levels are large enough, i.e., above a certain

threshold level. These thresholds are increasing with inventory levels and with time.

The third chapter explores optimal assortment decisions with non-identical price and

cost parameters. We extend our results in the first chapter to a setting where prices

are weakly increasing in the net utility of the product. We also derive some novel

results related to the structure of the optimal assortment in the dedicated system

with general price and cost parameters. The goal of the project is to characterize the

optimal assortment and explore the role of component commonality with different

price and cost parameters.

v

Page 6: Assortment Planning with Inventory Considerations in

Contents

Abstract iv

Acknowledgements viii

Introduction 1

1 The Role of Component Commonality in Product Assortment De-cisions 4

1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

1.2 The Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

1.3 Preliminaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

1.4 Systems with Component Commonality . . . . . . . . . . . . . . . . . 13

1.4.1 Independent Population Model . . . . . . . . . . . . . . . . . 14

1.4.2 Trend-Following Demand Model . . . . . . . . . . . . . . . . . 16

1.4.3 Comparison of Independent Population and Trend-FollowingModels . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

1.4.4 Effect of Cost of Common Component . . . . . . . . . . . . . 18

1.5 General Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

1.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

2 Dynamic Assortment Customization with Limited Inventories 27

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

2.2 Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

2.3 Model Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

vi

Page 7: Assortment Planning with Inventory Considerations in

2.3.1 The Dynamic Assortment Optimization Problem . . . . . . . 34

2.4 The Optimal Assortment Policy . . . . . . . . . . . . . . . . . . . . . 36

2.4.1 Single Customer Segment and Multiple Product Variants . . . 37

2.4.2 Two Customer Segments with Two Product Variants . . . . . 38

2.5 Numerical Study . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

2.5.1 Effect of Assortment Customization on Revenues . . . . . . . 41

2.5.2 Effect of Parameters on Threshold Levels . . . . . . . . . . . . 43

2.6 General Case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

2.6.1 Bounds on the Optimal Profit . . . . . . . . . . . . . . . . . . 45

2.6.2 Multiple Customer Segments and Multiple Product Variants . 47

2.7 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

3 Assortment Decisions with Non-identical Price and Cost Parame-ters 51

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

3.2 Extended Results of Chapter 1 with Non-Identical Price and CostParameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

3.3 Dedicated System with General Price and Cost Parameters . . . . . . 54

A Proofs of Main Results of Chapter 1 57

B Proofs of Supplementary Results of Chapter 1 69

C Proofs of Chapter 2 91

D Proofs of Chapter 3 107

Bibliography 113

Biography 118

vii

Page 8: Assortment Planning with Inventory Considerations in

Acknowledgements

I owe my deepest gratitude to my supervisors, Professors Fernando Bernstein and

Gurhan Kok, whose encouragement, enthusiasm, strong support, and great patience

to explain things clearly and simply enabled me to complete this thesis.

I would also like to thank my committee members, Professors Jeannette Song

and Serhan Ziya. During my thesis-writing period, they have provided stimulating

discussions and valuable feedback.

This thesis would not have been possible without wonderful training and en-

couragement from Professors Paul Zipkin, Kevin Shang, Otis Jennings, Pranab Ma-

jumder, and Li Chen.

I offer my regards and blessings to many of my colleagues and friends to support

me during the completion of this thesis. Especially, I would like to show gratitude

to Zhengliang Xue, Fang Liu, Isilay Talay Degirmenci, and Suri Gurumurthi.

Finally, I wish to thank my family for the support they provided during my entire

life and in particular, I am forever grateful to my wife and children, Na Li, Boqian

Xie, and Lingxuan Xie, who offer a loving environment and unconditional support

all the time.

viii

Page 9: Assortment Planning with Inventory Considerations in

Introduction

A wide assortment of variants allows customers to purchase products that more

closely match their preferences and helps retain customers for future purchases.

However, variety is costly from an operational perspective. Higher variety leads to

increased manufacturing complexity, e.g., in the form of more frequent switchovers in

production and to lower economies of scale. That is, adding more variants in the as-

sortment tends to reduce the demand for individual products and, at the same time,

increase their relative demand variability. Moreover, it requires increased ordering

and handling costs, i.e., more work is needed to display products. This reduces the

extent to which companies can take advantage of economies of scale. In addition,

larger demand variability due to a higher variety increases the likelihood of ending

the selling season with leftover inventory of several or all of the product variants.

The goal of this dissertation is to provide insights and tools to help companies deter-

mine the optimal level of product variety they should produce/stock and maximize

revenues by strategically choosing assortments offered to customers.

Using common components has been regarded as an effective operational strategy

to reduce manufacturing complexity and costs in academia and in industry. However,

little attention has been paid to the effect of commonality on product assortment

decisions. In Chapter 1, we analyze the role of component commonality in product

assortment decisions under various configurations of the bill of materials. We consider

a firm that produces multiple variants of a product. Products are assembled using

1

Page 10: Assortment Planning with Inventory Considerations in

a combination of common and dedicated components. We characterize the optimal

assortment and derive the optimal inventory levels for the common and dedicated

components under various bill-of-material configurations. We investigate the effect

of commonality on product variety and compare its benefits under different demand

characteristics. While commonality always leads to increased profits, its effect on

the level of product variety depends on the type of commonality. If all common

components are used for the production of the entire set of products, then the optimal

variety level increases relative to the system with no commonality. However, if the

common components are used by a subset of the final products, then the optimal

variety level may decrease with commonality. We find that the effects of commonality

on profit and variety level are stronger under a demand model that exhibits negative

correlation relative to a model with independent demands.

Chapter 2 studies a revenue management problem in an online retail setting with

limited inventories and a set of substitutable products. The online retailers are able

to provide customized assortments to individual customers based on their personal

information and their browsing and purchasing history. In this chapter, we find that

an online retailer has the potential to increase revenues by strategically restricting

the set of product options it makes available to customers, even when all products

are in stock. We consider a retailer with limited inventories of substitutable products

in a retail category. Customers arrive sequentially and the firm decides which sub-

set of the products to offer to each arriving customer depending on the customer’s

preferences, the inventory levels and the time-to-go in the season. We show that the

optimal assortment policy is to offer all available products if customers are homoge-

nous. However, with multiple customer segments characterized by different product

preferences, it may be optimal to limit the choice set of some customers. That is, it

may be optimal to not offer products with low inventories to some customer segments

and reserve those units for future customers (who may have a stronger preference for

2

Page 11: Assortment Planning with Inventory Considerations in

those products). The optimal assortment policy is a threshold policy under which a

product is offered to a customer segment if its inventory level is higher than a thresh-

old value. These thresholds are decreasing in time and increasing in the inventory

levels of other products. In a numerical study, we find that the revenue impact of

assortment customization can be significant, demonstrating its use as another lever

for revenue maximization in addition to pricing.

In Chapter 3, we extend the results of the first chapter to a system with non-

identical price and cost parameters. We assume price is linearly increasing in the

net utility of the product. We find that the optimal assortment in the dedicated

system under both demand models continues to be a popular set. The optimal

variety level increases as a common component is introduced to all products under

the trend-following demand model. The optimal assortment in the pooled system

is no longer to offer all products and turns to be a popular set which has a higher

variety level than that in the system with the common component. We also study

the structure of the optimal assortment in the dedicated system with general price

and cost parameters. For any product in the optimal assortment, all other products

with higher prices and utilities than this product are also included in the optimal

assortment.

3

Page 12: Assortment Planning with Inventory Considerations in

1

The Role of Component Commonality in ProductAssortment Decisions

1.1 Introduction

Providing a wide assortment of variants in a product category is critical for most

manufacturers. A higher variety level tends to attract more customers, especially

in a competitive environment, as it allows them to find items that more closely

match their preferences. In turn, this helps retain customers for future purchases.

However, variety is costly from an operational perspective. Higher variety leads to

increased manufacturing complexity, e.g., in the form of more frequent switch-overs

in production, and to a more fragmented product line. That is, adding more variants

to the assortment tends to reduce the demand for individual products and, at the

same time, increase their relative demand variability. This reduces the extent to

which companies can take advantage of economies of scale. Thus, higher inventory

levels are required to maintain the same service level.

Companies employ different strategies to support a high level of variety, such as

using common components, investing in flexible capacity, and redesigning products

4

Page 13: Assortment Planning with Inventory Considerations in

and processes to benefit from delayed differentiation – all of these are different forms

of operational flexibility. This generates a higher demand volume for the common

parts and therefore results in lower operational costs. While the benefits of imple-

menting some form of flexibility for a given line of products are well documented in

the literature, this paper explores the effect of commonality on assortment decisions.

There are three forces at play in our model: (1) Higher variety leads to increased total

demand; (2) This also leads to a higher relative demand variability for each product;

(3) At the same time, the demand aggregation effect for the common components

mitigates the increased costs associated with higher demand variability. In view of

these three effects, we investigate how the choice of an assembly configuration (or

bill of materials) affects a firm’s assortment and stocking decisions.

In particular, we consider an assemble-to-order system in which a manufacturer

produces a number of variants in one product category. Each variant is assembled

from several components. There are product-specific (i.e., dedicated) components

and components that are common to a subset of (or all) the product variants. We

consider two demand models, both introduced in van Ryzin and Mahajan (1999).

Under the independent population model, all consumers make purchasing decisions

independently from each other. Under the trend-following demand model, consumers

follow the choice made by the early buyers, which leads to a higher demand variance

for each product and to negative correlation between products, yielding the largest

risk pooling benefit from a common component. In these settings, we characterize

the optimal component inventory levels and the structure of the optimal assortment

for some special cases of the bill of materials. Based on these results, we compare a

system where all variants are produced using only dedicated components to systems

that incorporate flexibility in the form of component commonality.

We find that the effect of commonality on product assortment decisions depends

on how these components are integrated in the bill of materials. If each common

5

Page 14: Assortment Planning with Inventory Considerations in

component used in manufacturing is shared by all product variants, then the optimal

level of variety increases with commonality. In this case, the demand pooling effect

results in high enough savings to allow the firm to introduce more variants. However,

this result does not necessarily hold for more general bills of materials. In fact, when

some of the common components are shared by a subset of the product variants, the

optimal assortment may decrease relative to that in a setting without commonality.

This reduction in the optimal assortment occurs when removing a variant that does

not use the common component leads to a sufficiently increased scale effect on the

variants that do share that component, making it more attractive to reduce the

assortment in a way that maximizes the demand pooling effect. In these cases, it

may be optimal to rationalize the product line in order to shift demand to products

that share common resources. In this respect, we explore conditions that lead to

increased or reduced assortment offerings in general systems. When commonality

does increase the depth of the optimal product assortment, we find that the extent

of this increase depends on the characteristics of market demand. Specifically, the

trend-following model, which exhibits higher demand variance and pair-wise negative

correlation between products, leads to higher increases in the set of products offered

relative to the independent population model.

Our results are relevant for product line managers (who decide which products to

offer), and supply chain managers and product designers (who make procurement or

design decisions that affect the bill-of-material configuration). If common compo-

nents are used to produce all (or most of) the products, then a wide range of variants

can be offered. If design limitations preclude the use of common components across

the entire product line, then it may be optimal to restrict the product offering to

benefit from an increased scale effect for the products that use the common compo-

nents, therefore leading to the exclusion of some products that do not share those

common components. Our results also suggest that there may be opportunities to

6

Page 15: Assortment Planning with Inventory Considerations in

increase variety inexpensively by exploring new product designs that utilize as much

component commonality as feasible with the most popular products.

This paper builds on two streams of research: retail assortment planning and

component commonality/resource flexibility. The literature on retail assortment

planning generally focuses on the consumer choice aspect of assortment decisions

and works with the simplest production structure in which each variant is produced

using a dedicated component. van Ryzin and Mahajan (1999) consider such a

model and demonstrate that the optimal assortment consists of a certain number of

the most popular products. Other papers in this area include Smith and Agrawal

(2000), Cachon et al. (2005), Gaur and Honhon (2006), and Kok and Fisher (2007).

See Kok et al. (2008) for a detailed review of this stream of work. The literature

on component commonality (e.g., van Mieghem 1998, 2004, Bernstein et al. 2009)

and other forms of operational flexibility (e.g., Fine and Freund 1990, Tang and Lee

1997) studies inventory and common-component allocation decisions, and explores

the resulting value of flexibility for a given fixed assortment of products. Similarly,

the literature on assemble-to-order systems investigates optimal inventory policies

for a given assortment of products (Song and Zipkin 2003, Song and Zhao 2009). In

this paper, we consider a setting in which the set of products offered is an endogenous

decision based on the structure of the bill of materials that includes flexible resources

in the form of common components and on a consumer choice model like those

considered in the assortment planning literature.

Product variety has also been widely investigated in the marketing literature. For

example, Kahn (1998) provides a review of papers that explore how product variety

influences the revenue potential of a product line. The use of common components

in manufacturing may lead to lower costs, but also a to lower degree of product

differentiation from the consumers’ perspective. This tradeoff is investigated in the

product-line design literature (e.g., Desai et al. 2001 and Heese and Swaminathan

7

Page 16: Assortment Planning with Inventory Considerations in

2006). A number of papers focus on how to create and implement product variety in

manufacturing settings (e.g., Ramdas et al. 2003). Hopp and Xu (2005) consider a

product selection problem and show that modularity always leads to higher variety.

The rest of the paper is organized as follows. Sections 1.2 describes the model set-

up and Section 1.3 present the results for the dedicated system. Section 1.4 presents

the analysis of systems with common and dedicated components, but in which the

common components are used in the production of all product variants. Section 1.5

contains an analysis of systems with more general bill of materials. Section 1.6

concludes the paper. All proofs and supplementary material are provided in an

Electronic Companion (Appendix A) and in Appendices B and C.

1.2 The Model

We consider a manufacturer that makes product line decisions regarding the set of

product variants to offer in the market. The manufacturer also decides the capacity

or inventory levels for the components used in production. Each product variant may

represent, for example, a color/size/design combination of a garment or a specific

configuration of a personal computer. The set of all variants is denoted by N =

{1, 2, · · · , N} and the firm offers a subset S ⊂ N . Each variant is produced according

to a bill of materials that dictates the components used in its fabrication. There are

product-specific (or dedicated) components and components that are common to a

subset of (or all) the variants. For example, Nike’s recently-designed “LunarLite”

foam is used in several of the company’s shoes, while a specific color may be common

to a subset of the sneakers, and a certain pattern may be specific to one product

variant. Similar examples apply to apparel and modular products, such as a personal

computer. The firm holds inventory of components and final production/assembly

time is negligible. This model structure applies to assemble-to-order systems, such

as Dell’s production of personal computers, where the final step of production takes

8

Page 17: Assortment Planning with Inventory Considerations in

place after customers place their orders for specific product configurations. In the

case of Nike, the company has recently introduced an online system that links Nike

manufacturing partners to reduce lead times and reduce the percentage of shoes it

orders on speculation from 30% to 3%, essentially creating a make-to-order system

for the manufacturers (Holmes 2003). The model may also apply to settings in

which the manufacturer produces according to orders from retailers, provided that

final production or assembly takes place after the retailers’ orders are received.

We model demand using a consumer choice model similar to that in van Ryzin

and Mahajan (1999). The choice of a product variant within the offered assortment

is based on the Multinomial Logit (MNL) model. A consumer chooses a product

in the assortment or selects the no-purchase option (denoted by 0) to maximize her

utility. The expected utility derived from option i ∈ S ∪ {0} is given by ui. The

probability of a customer choosing option i is given by

qSi =θi∑

j∈S θj + θ0

, i ∈ S ∪ {0} ,

where θi = eui . (We refer to van Ryzin and Mahajan (1999) for details on the MNL

model.) For simplicity, we refer to θi as the utility for variant i and let Θ denote the

vector (θ1, θ2, · · · , θN). We assume that products are indexed in descending order

of their popularity (as measured by their utility parameters), i.e., θ1 ≥ θ2 · · · ≥ θN .

We assume that consumers choose a variant based on the offered assortment S

and if the selected product is not available, then the sale is backordered or lost.1 This

assumption implies a static or assortment-based substitution, therefore ignoring the

dynamics of product substitution that are based on the availability of products at

the time of the customer arrival. This model is particularly suitable for catalog

1 Because we consider a single-period setting, backorders and lost sales are equivalent from amodeling point of view and would lead to the same profit expression (except for a constant) if thebackorder penalty cost is set to equal the profit margin. It is possible to include a salvage valueand a shortage penalty cost by a simple redefinition of parameters.

9

Page 18: Assortment Planning with Inventory Considerations in

retailers and for manufacturers selling customized products to retailers or directly

to consumers. In those settings, consumers choose from a menu of products in a

catalog or on a web site. Depending on the availability of components, the firm

either delivers that product immediately or the sale is lost. The same is the case

with apparel manufacturers that receive orders from retailers based on the offered

assortment.

Demand arrives over a single selling season. We consider two models of consumer

demand, both introduced in van Ryzin and Mahajan (1999). Under the indepen-

dent population model, denoted IP, the total number of customers interested in the

product category follows a Poisson distribution with rate λ and each customer se-

lects a variant independently of the choices made by all other customers. Then,

demand for variant i follows a Poisson process with rate λi = qSi λ. We approximate

this distribution with a Normal distribution with mean λi and standard deviation

σi = σλβi , with σ > 0 and 0 < β < 1.

We also consider the trend-following demand model, denoted TF. The trend-

following model is a stylized model construct that describes a rather extreme rep-

resentation of customer behavior for products where the trend is established by the

leader (early buyers, buyers with influence, reputable consumer reports, etc.), and

all other customers follow the trend. This model is appropriate for fashion goods,

where customers tend to follow trends set early in the season. Thus, for each vari-

ant i, demand Di is either zero (with probability 1 − qSi ) or equal to total market

demand (with probability qSi ). Total market demand is itself a Normal random vari-

able with mean λ and standard deviation σ. We assume that λ is sufficiently larger

than σ so that the probability of a negative demand realization is negligible. The

trend-following model exhibits pair-wise negative correlation between product de-

mands and a higher demand variance at the individual product level relative to the

10

Page 19: Assortment Planning with Inventory Considerations in

IP model.2 In our setting, demand variance and correlation influence not only the

capacity decisions and the system’s profit, but also the assortment decision.

The bill of materials consists of a set of components M = {1, 2, · · · ,M} and a

production matrix matching components with end products. We assume, without

loss of generality, that a finished product requires one unit of each of its components.

We investigate special forms of the production matrix. In a dedicated system (de-

noted D), all components are product-specific. We then study a setting (denoted

C) in which each variant is produced using a dedicated component and a component

that is common to all products. (Although we assume that every product requires

two components – dedicated or common – any one of these components may repre-

sent a kit of parts or sub-assemblies.) We finally consider settings with more general

bill of materials.

As in van Ryzin and Mahajan (1999), we assume identical total production costs

and prices across all variants for tractability. The selling price of each variant is p.

In system D, the costs of the two dedicated components for each product are kd and

kc, respectively. In system C, the cost of each dedicated component is kd and the

cost of the common component is kc (this component replaces all of the dedicated

components with cost kc in system D). Under other bill of materials, the cost of

components is set so that the total unit production cost is kd + kc for all products.

The firm’s objective is to determine the assortment (which products to offer) and

the stocking levels of all components to maximize profit. We define a popular set

as the set of the n most popular products and denote it by An = {1, 2, .., n}. We

denote the optimal profit of the firm when it offers assortment S by ΠS, and define

Πndef= ΠAn . We denote the structure of the bill of materials (D or C) and the

2 Consider an assortment S. Under the TF model, Cov(Di, Dj) = −λ2qSi qSj < 0. Also, setting

σ = 1 and β = 1/2 in the IP model and σ = λ1/2 in the TF model (both corresponding to theNormal approximation to a Poisson random variable), var(Di) = λqSi under the IP model andvar(Di) = λqSi + λ2qSi (1− qSi ) under the TF model.

11

Page 20: Assortment Planning with Inventory Considerations in

demand model (IP or TF) in the superscript of the relevant variables.

1.3 Preliminaries

We first explore the optimal assortment structure in the dedicated system. In this

setting, each product is manufactured using only dedicated components. Because a

product requires one unit of each component, the results below essentially assume a

single dedicated component per product, with an aggregate cost of k = kd + kc per

unit.

Given an assortment S, we let S(δ) = S ∪ {m} denote the set of variants in S

plus an additional variant m 6∈ S with utility δ. The resulting optimal profit is

ΠDS (δ)

def= ΠD

S(δ). van Ryzin and Mahajan (1999) show that ΠDS (δ) is a quasi-convex

function of δ (implying that, given an assortment S, it is optimal to either add the

next most popular product not in S or to not add any other variant). The paper also

shows that the optimal assortment is a popular set. (We actually prove the result

for a more general formulation of the TF model in which the total market demand

is random, whereas this is a constant in van Ryzin and Mahajan 1999. A detailed

analysis of this case and the proofs of these results are presented in Appendix A.)

This result provides a useful characterization of the optimal assortment as it reduces

the number of sets to be considered for optimization from 2N to N . Note that the

optimal assortment is not necessarily equal to the set of all possible products N .

While adding a variant increases total demand, the relative demand variability for

each variant in the assortment also increases, leading to higher inventory costs.

We next present and discuss two technical conditions that we impose throughout

the rest of the paper. Because ΠDS (δ) first decreases and then increases in δ, let δn

denote the value of δ > 0 such that ΠDn (0) = ΠD

n (δn). If the utility of the next most

popular product is greater than δn (i.e., θn+1 > δn), then it is profitable to include

12

Page 21: Assortment Planning with Inventory Considerations in

variant n + 1 in the assortment. The quantity δn is a function of the utility vector

and of the cost and price parameters. If ΠDn (δ) is increasing, then we let δn = 0. If

ΠDn (δ) < ΠD

n (0) for all δ, then we let δn =∞. The conditions are as follows.

Condition 1. Given a set of utilities {θ1, · · · , θN}, the following holds for all n =

1, ..., N − 1:

If δn ≤ θn, then δn−1 − δn ≤ θn − θn+1.

Condition 2. δn is increasing in k as long as δn < θn.

Condition 1 implies that the profit function in the dedicated system is quasi-

concave in n (see Theorem A.1 in Appendix A). This result, together with the

optimality of a popular set, implies that the optimal assortment can be found it-

eratively by adding one product variant at a time (in a sequence given by their

popularity) starting with the empty set. Condition 2 implies that adding a product

to the assortment becomes less attractive as production costs increase. Both condi-

tions depend only on the parameters of the model. In our numerical experiments,

covering a wide range of parameter values, these conditions are always satisfied. The

conditions enable us to compare the size of the optimal assortment across various bill-

of-material configurations, but they are not required for the results on the structure

of the optimal assortment under any of the system configurations.

1.4 Systems with Component Commonality

In this section we explore settings in which the bill of materials includes dedicated

as well as common components and in which the common component is shared

by all offered product variants. This production structure features what we call

universal commonality. In its more general form, this system consists of a common

component, which can represent a kit of common components, with cost kc, and

a set of dedicated components, such that each product variant is produced using

13

Page 22: Assortment Planning with Inventory Considerations in

the common component and a dedicated component (possibly representing a kit of

dedicated components), with cost kd. This setting is comparable to the dedicated

system, in the sense that the total cost associated with the production of a variant

is equal to k = kc + kd. Figure 1.5 (a) in Section 5 illustrates a system with this

bill-of-material configuration.

1.4.1 Independent Population Model

We begin with the independent population demand model. For a given assortment

S, let yi be the stocking level for the dedicated component corresponding to variant

i, i ∈ S, and yc the stocking level for the common component. The corresponding

profit is

ΠC,IPS ({yi}i∈S, yc) = pE[min{yc,

∑i∈S

min{yi, Di}}]−∑i∈S

kdyi − kcyc.

As demonstrated in Van Mieghem (1998), ΠC,IPS is jointly concave in ({yi}i∈S, yc) .

It is difficult to analytically compare the size of the optimal assortment between

systems D and C under the independent population demand model, primarily be-

cause of the lack of closed-form expressions for the optimal stocking levels in system

C. In the dedicated system, each component experiences the same demand as its

corresponding finished product. In contrast, in system C, the dedicated components

face the same demand stream as in system D, whereas the common component faces

a demand stream that is equal to the aggregate demand for all variants. As a result,

the profit function is not separable in the stocking quantities. By considering a sim-

plified version of the demand distribution under the IP model, the following result

shows that universal commonality increases the size of the optimal assortment.

Theorem 1.4.1. Consider a setting with N identical products (i.e., θi = θ for i =

1, ..., N and, for any assortment S with n = |S|, qSi = qS = θnθ+θ0

for any i ∈ S)

14

Page 23: Assortment Planning with Inventory Considerations in

for sufficiently large N . Furthermore, consider the following two-point demand

distribution, defined for a given assortment S:

DSi =

{λqS + σ2, with probability λqS

λqS+σ2

0, otherwise.

(This distribution preserves the mean and standard deviation of the original Normal

distribution, with β = 1/2.) Then, ΠD,IPn+1 − ΠD,IP

n ≤ ΠC,IPn+1 − ΠC,IP

n , which implies

that the optimal assortment in the system with universal commonality contains at

least as many variants as the optimal assortment in the dedicated system.

Theorem 1.4.1 states that if the marginal gain from adding a new variant in the

dedicated system is positive, then that new variant should be included in the system

with commonality as well. This, together with the quasi-concavity of the profit

function in the dedicated system as a function of n (see Theorem A.1), allows us

to conclude that replacing a set of dedicated components with a component that

is common to all product variants results in an increase in the size of the optimal

assortment. Universal commonality leads to a broader product offering because

pooling resources for one component of each product reduces the realized overage

and underage costs for all products. This mitigates the cost of having a high level

of product variety.

For the original model with Normal demand distributions, an extensive numerical

study suggests that the optimal assortment in the system with commonality is a

popular set and that it is indeed no smaller than the optimal assortment in the

dedicated system. The study consists of 360 experiments, all with ten possible

variants and considering two possible demand rates, λ = 200 and λ = 400 (further

details are presented in Appendix A). Comparing the number of variants in the

optimal assortments in systems D and C, we find that component commonality leads

to an increase in the level of variety from an average of 6.67 variants to an average of

15

Page 24: Assortment Planning with Inventory Considerations in

7.81 variants when λ = 200 and from an average of 7.65 variants to an average of 8.74

variants when λ = 400. The impact of commonality is larger at low demand levels

(λ = 200) because demand is more variable in those settings and inventory costs

represent a larger portion of total profit. We also find that the effect of commonality

on the level of product variety is larger in settings with a relatively high utility of the

no-purchase option and in settings in which product utilities have relatively similar

values.3 In all cases, universal component commonality effectively reduces inventory

costs, leading to higher variety levels. At the same time, in settings with larger θ0 or

with similar values of the products’ utilities, the gains in market share from adding

another product to an existing assortment are relatively larger. Hence, the firm has

a stronger incentive to increase variety in those settings, therefore amplifying the

effect of commonality on both variety level and profit. Finally, we find that when

the cost of the common component is higher relative to the corresponding cost in

the dedicated system, the impact of commonality in the increase of product variety

is more significant.

1.4.2 Trend-Following Demand Model

We now focus on the trend-following demand model. Under this model, all cus-

tomers purchase the product variant selected by the first customer. (Recall that,

at the time of making assortment and component capacity/stocking decisions, there

is uncertainty regarding which product variant will be the preferred choice in the

market.) The profit function for the firm is given by

ΠC,TFS ({yi}i∈S, yc) =

∑i∈S

qSi pEmin{yc,min{yi, D}} −∑i∈S

kdyi − kcyc.

3 Consider the following example with λ = 200, p = 10, kd = kc = 2, σ = 1, β = 0.5, θ0 = 10.For Θ = (20, 18, 16, 14, 12, 10, 8, 6, 4, 2), the optimal variety level increases from 6 in the dedicatedsystem to 7 in the system with commonality, accompanied by a 1.7% increase in optimal profit(from 953.9 to 970.1). On the other hand, for Θ = (20, 19, 18, 17, 16, 15, 14, 13, 12, 11), the optimalvariety level increases from 7 in the dedicated system to 9 in the system with commonality, leadingto a 2.1% increase in optimal profit (from 968.4 to 989.0).

16

Page 25: Assortment Planning with Inventory Considerations in

Because it is optimal to stock less of the dedicated components than of the common

component, i.e., yi ≤ yc, we have that ΠC,TFS ({yi}i∈S, yc) =

∑i∈S(qSi pEmin{yi, D}−

kdyi) − kcyc. This function is jointly concave in ({yi}i∈S , yc). Proposition A.1

in Appendix A derives the optimal stocking quantities and the resulting optimal

profit in this setting. The result indicates that it is optimal to stock an equal

amount of the dedicated components corresponding to a subset of the most popular

variants and that that amount is itself equal to the stocking quantity of the common

component. The stocking quantities of the remaining variants decrease in order of

their popularity. Using the form of the optimal profit function and optimal stocking

levels, we characterize the effect of adding one more variant to an existing assortment,

which leads to the following result.

Theorem 1.4.2. (i) The function ΠC,TFS (δ) is quasi-convex in δ. Therefore, the

optimal assortment is a popular set. (ii) Moreover, ΠD,TFn+1 −ΠD,TF

n ≤ ΠC,TFn+1 −ΠC,TF

n ,

which implies that the optimal assortment in the system with universal commonality

is no smaller than that in the dedicated system.

This result shows that, under the trend-following model, the structure of the

optimal assortment is preserved with the introduction of universal commonality, but

the optimal level of product variety increases (weakly) relative to the dedicated

system.

1.4.3 Comparison of Independent Population and Trend-Following Models

We have shown that, under both demand models, replacing a set of dedicated com-

ponents by a common component shared by all product variants results in a larger

optimal assortment. At the same time, for a fixed set of products, the effect of

commonality on profit is more significant when the difference between the sum of

the individual products’ demand variabilities and the aggregate demand variability

17

Page 26: Assortment Planning with Inventory Considerations in

is relatively larger, as is the case under the TF model compared to the IP model.

In our setting, however, changes in the demand model affect not only the benefits of

risk-pooling but also, possibly, the size of the optimal assortment. In this section,

we report the results of a numerical study that examines the effect of commonality

on the optimal variety level under the two demand models.

The study for the TF model is based on the same 360 experiments reported

in Section 1.4.1, with the exception that each of the nine possible cost vectors is

adjusted to ensure that the average number of variants (across all other 40 parameter

combinations) in the dedicated system is the same as that under the IP model. For

each instance, we compute the optimal assortment and profit in system D and in

system C. Table A.2 in Appendix A summarizes the results. The increase in the

optimal level of variety is generally higher in terms of frequency and average absolute

magnitude under the TF model than under the IP model. Under the IP model, the

level of variety increases by an average of 1.12 units. In contrast, under the TF

model, the level of variety increases by an average of 2.84 units. Similarly, the

percentage profit improvement due to commonality is higher under the TF model

(4.15% versus 3.55%). Based on these findings, we conclude that the effect of

commonality on the level of variety and on profit is larger under the TF model, even

when assortment decisions are endogenous.

1.4.4 Effect of Cost of Common Component

We have so far assumed that the per unit cost of a common component is the same as

its dedicated-component counterpart in the dedicated system (both equal to kc). In

many cases, designing or purchasing a common component may be more costly than

a dedicated component. Proposition B.2 in Appendix B shows that, in a system

with dedicated components, the level of variety decreases as unit costs increase.

Hence, the addition of commonality (weakly) increases the level of variety if the

18

Page 27: Assortment Planning with Inventory Considerations in

common component is equal in cost or somewhat more expensive than the dedicated

component it replaces. However, the level of variety may decrease in the system

with commonality, relative to the dedicated system, if the common component is

significantly more expensive. Based on our numerical study, we find that the optimal

number of product variants is larger with commonality even for increases of up to

30% in the cost of the common components.

1.5 General Systems

In this section we explore systems with general bill of materials. In particular, we

consider systems in which common components may be used to produce some, but

not all, of the finished products. We consider a few representative structures of the

bill of materials, all of which are illustrated in Figure 1.5. The configuration in (a)

represents the system with universal commonality studied in Section 1.4. In (b)

and (c), there is only one common component shared by two products, while the

other products are built using dedicated components only. In (d), two products

use both dedicated and common components and the other two only use dedicated

components. In (e), all products share a common component and require an addi-

tional component, which is a dedicated component each for products 3 and 4 and

a common component for products 1 and 2. We refer to the systems in (b)-(e) as

systems with partial component commonality. The results obtained in this section

for the settings depicted in Figure 1.5 (b)-(e) allow us to derive insights regarding

the effect of commonality in product assortment decisions in systems with general

bill of materials.

We start by considering a simple setting with N product variants and M = N−1

components, in which two variants, say i and j, share a common component and all

other variants use dedicated components (as illustrated in Figure 1.5, (b) and (c)).

We compare the optimal assortment in this system with that in the system with

19

Page 28: Assortment Planning with Inventory Considerations in

Figure 1.1: Sample structures of the bill of materials involving commonality

only dedicated components. To make the comparison meaningful, we maintain the

cost of producing each variant equal in both systems. That is, in the system with

dedicated components, the cost of each dedicated component is given by kd + kc,

while in the system with partial commonality the cost of the common component

is kd + kc and all other products continue to use the same dedicated components as

in the original system. (Recall that a popular set is defined according to the order

given by the utilities in the vector Θ.)

Proposition 1.5.1. Consider a system with partial component commonality in which

variants i and j are produced using a common component and all other variants use

dedicated components. For both demand models, the optimal assortment is either a

popular set or it consists of variants i and j plus a popular subset of the remaining

variants.

Proposition 1.5.1 implies that in a system with a general bill of materials, the

optimal assortment may not necessarily be a popular set. For example, in Figure

1.5(c), under the IP model with parameters p = 10, kd = kc = 3, λ = 150, σ = 1,

β = 0.5, θ0 = 5, and Θ = (16, 15, 8, 1), the optimal assortment is {1, 2, 4}. In

contrast, under the same parameter values, but in the setting in which all variants

use dedicated components, the optimal assortment is {1, 2, 3}. That is, pooling the

resources used by products 1 and 4 makes it more profitable to include the least

20

Page 29: Assortment Planning with Inventory Considerations in

popular variant in the assortment, rather than keeping variant 3, to benefit from the

scale generated by aggregating the demands for these products. We next analyze the

size of the optimal assortment in settings with commonality, relative to the dedicated

system.

Proposition 1.5.2. Consider either the IP or the TF demand models. Let Ai+j be

the optimal assortment of a system in which variants i and j (i < j) are produced

using a common component and all other variants use dedicated components. Let An

be the optimal assortment in the corresponding dedicated system in which all variants

are produced using dedicated components.

1. If i, j ≤ n, then there exists a threshold t1 such that Ai+j is a strict subset of

An containing variants i and j if θn < t1, and Ai+j = An otherwise.

2. If i < n and j > n, then there exists a threshold t2 such that Ai+j is a strict

subset of An containing variants i and j if θn < t2, and Ai+j = An ∪ {j}

otherwise.

3. If i, j > n, then, there exist thresholds t3 > t4 such that

(a) Ai+j = An ∪ {i, j}, if t3 < θi + θj ≤ θn or t4 ≤ θn < θi + θj,

(b) Ai+j = An, if θi + θj < min (t3, θn) ,

(c) Ai+j = A′∪{i, j} where A′ is a strict subset of An, if θn < min(t4, θi +

θj).

This result provides a complete characterization of the optimal assortment in the

system with a common component for variants i and j. The optimal assortment

actually depends on the relative “popularity”of variants i and j (given that i < j).

In particular, if either variant i or both variants i and j are in the optimal assortment

of the original dedicated system, then both variants are in the optimal assortment

21

Page 30: Assortment Planning with Inventory Considerations in

in the system with partial commonality as well (Proposition 1.5.2, parts 1 and 2).

However, in this setting, the optimal assortment may be smaller than that in the

dedicated system. If neither variant i nor variant j is in the optimal assortment of

the dedicated system, then it is optimal to offer these two variants in the system with

commonality only if the sum of their utilities is large enough (Proposition 1.5.2, part

3). At the same time, if that is the case, then one or more variants from the optimal

assortment of the dedicated system may be dropped in the optimal assortment of

the system with partial commonality (Proposition 1.5.2, part 3(c)).

Along the lines of the system studied in Proposition 1.5.2, if a product designer

had the liberty to choose the two products that should share a common component,

then the profit-maximizing solution would be achieved by pooling the components

of the two most popular products (see Proposition B.6 in Appendix B). At the

same time, Proposition 1.5.2 suggests that this may result in a smaller optimal

assortment. With partial commonality, the extent of product variety can only

increase if commonality is used with at least one product outside of the optimal

assortment in the dedicated system. In this way, new variants may be introduced

to the assortment at a relatively low cost, benefiting from the operational scale of

the products already in the assortment.

We conclude that, in sharp contrast to the results in Section 1.4, replacing a

strict subset of dedicated components by a common component may result in a

reduced number of variants in the optimal assortment and this set may no longer

be a popular set. To understand why partial commonality may reduce the optimal

assortment, let us consider the case with i < j < n. A setting with n = 3, i = 1,

and j = 2, is illustrated in Figure 1.5 below. The graphs in (a) show the profit

of each variant in the dedicated system, while the graphs in (b) show the combined

profit associated with variants i and j, and the profit for variant n, in the system with

partial commonality. Variant n is in the optimal assortment of the dedicated system

22

Page 31: Assortment Planning with Inventory Considerations in

because the profit associated with this variant is higher than the aggregate increase

in profit that the other products in the assortment would experience in the absence

of variant n (recall that the choice probabilities – and therefore the expected demand

rates – of the other products in the assortment increase if variant n is removed from

the assortment, in this case from qi to qi and from qj to qj). That is, ∆n > ∆i + ∆j

in Figure 1.5 (a). On the other hand, in the system with partial commonality the

common component shared by variants i and j experiences a larger demand rate (and

therefore a more pronounced scale effect) than the dedicated component for variant i

does in the dedicated system. Note that the profit function of an individual product

variant is convex increasing in its demand rate (because of the economies of scale

associated with inventory costs). Hence, in the system with partial commonality,

the profit that variant n generates may no longer be larger than the profit impact of

the demand spill-over effect (in particular, onto the common component for variants

i and j) that occurs in the absence of variant n. In reference to Figure 1.5 (b),

we may have that ∆n < ∆ij. In that case, variant n would be excluded from the

assortment.

We next consider two generalizations of the system, represented in Figures 1.5

(d) and (e). In Figure 1.5 (d), variants i and j are assembled from one common

component and one dedicated component each (with costs kc and kd, respectively)

while, in the corresponding dedicated system, these variants are assembled using only

dedicated components (each with cost kd+kc). In Figure 1.5 (e), there is a component

common to all products and another component common to products i and j only.

All other variants require, in addition, a dedicated component. The results for

the system in Figure 1.5 (d) are similar to those in Proposition Proposition 1.5.2.

That is, the optimal assortment in this setting is either a popular set or a popular

set plus variants {i, j} . Also, the number of variants in the optimal assortment

associated with the system with partial commonality may decrease, remain the same,

23

Page 32: Assortment Planning with Inventory Considerations in

Figure 1.2: Profit of variants in the dedicated and partial-commonality systems

or increase relative to the dedicated system. Regarding the system in Figure 1.5 (e),

we compare it to the system in Figure 1.5 (a) to examine the effect of introducing an

additional common component to a subset of the variants in a system with universal

commonality. The optimal assortment in the system shown in Figure 1.5 (a) is itself

no smaller than that in the related system with only dedicated components. In turn,

here too, the optimal assortment in the system with partial commonality shown in

Figure 1.5 (e) may be larger, equal or smaller than that in the system with universal

commonality. The existence of a common component shared by a strict subset

of the product variants again creates an imbalance in terms of the scale advantage

generated by demand aggregation – this benefit is higher for those products that use

a larger number of common components. As a result, it may be optimal to exclude

a product that makes less use of commonality to drive more demand to those that

use more common components and therefore increase the benefits associated with

demand pooling. (Propositions B.7 and B.8 in Appendix B contain formal statements

24

Page 33: Assortment Planning with Inventory Considerations in

Figure 1.3: Role of component commonality in optimal assortment

and proofs of these results.)

To close, while introducing commonality to a manufacturing system always in-

creases total profit (as long as the costs of components do not change), the effect

on the optimal variety level is not necessarily monotone in the amount of com-

monality. Figure 1.5 shows an example based on the trend-following model with

p = 10, kd = kc = 0.3, λ = 150, θ0 = 5, and Θ = (12, 11, 5.8).

1.6 Conclusion

This chapter considers the role of component commonality in product assortment

and component inventory decisions in assemble-to-order systems. We characterize

and compare the structure of the optimal assortment and the optimal inventory levels

for various bill-of-material configurations. While component commonality increases

total system profit, we find that the effect of commonality on the optimal assortment

depends on the specific configuration of the bill of materials. With universal com-

monality (a common component is shared by all variants), the optimal assortment

is (weakly) larger than that in the system without commonality. Furthermore, the

assortment consists of a subset of the most popular variants (as in the case without

commonality). However, these results do not hold for more general bill-of-material

configurations. If the common component is shared by a strict subset of the vari-

25

Page 34: Assortment Planning with Inventory Considerations in

ants, then the optimal assortment may no longer be a popular set and the optimal

level of variety may increase or decrease relative to the system without commonality.

In particular, the level of product variety (weakly) decreases when commonality is

introduced for products that are already in the optimal assortment corresponding to

the dedicated system (i.e., products with high demand volumes). This result may

hold even when commonality is introduced in a system that already involves other

common components shared by all product variants. These findings indicate that

product line and supply chain managers must be aware of the effect that introducing

commonality in manufacturing may have on optimal product line decisions, such as

the removal of products with relatively small demand. This is particularly important

if market share and market coverage (level of product variety) are critical to the com-

petitive position of a firm. Finally, we find that the effect of commonality on both

profit and variety level is more pronounced when demands are more variable and

exhibit pair-wise negative correlation. In an alternative interpretation of our model,

the results apply to the use of flexible resources and the related capacity decisions

(rather than common components and the corresponding inventory decisions).

26

Page 35: Assortment Planning with Inventory Considerations in

2

Dynamic Assortment Customization with LimitedInventories

2.1 Introduction

Online retailing remains the retail industry’s growth engine even in the second biggest

recession of U.S. history. In the last decade, online retailers have heavily invested in

diverse technologies, such as sophisticated analytics and personalization tools. To-

gether with its low operating cost and the convenience of online shopping, e-commerce

has emerged as an effective way of improving customer service and retention. For

example, online retailers are able to send personalized promotions or display a cus-

tomized price for a specific product to individual customers based on their personal

information and browsing and purchasing history, which can effectively enhance the

customers’ shopping experience and influence their purchasing activities. In par-

ticular, as we argue in this paper, providing customized assortments to individual

customers has the potential to increase overall sales for online retailers. Compared

to traditional bricks-and-mortar stores, online retailers are able to easily collect per-

sonal information, record their customers’ purchase history, and control the selection

27

Page 36: Assortment Planning with Inventory Considerations in

of products a customer is exposed to by strategically selecting the product variants

displayed on the web page. For example, Amazon.com’s web site, like many other

online retailers’ web sites, requires a customer to log in using her/his account be-

fore making a purchase. This allows the company to track the customer’s personal

information and purchase history. Based on this data, the company may be able

to estimate the customer’s preferences on styles and colors. An online retailer like

Amazon has several opportunities to use this information to restrict the set of prod-

ucts made available to customers. For example, suppose that a customer wants

to buy a pair of basketball shoes. He/she begins by searching for this item at the

company’s web page. The identity of the customer may be revealed through his/her

log in or through the browsing history recorded in the customer’s computer. The

company displays a selection of product options over multiple pages. The shoes on

the first page are likely to receive more attention. Therefore, the company may be

able to control the selection of shoes offered to a customer by restricting the set of

product variants displayed in the first few pages of the search results. In the next

step, the customer clicks on the product he/she prefers and the web site displays the

colors or sizes that are available. Here, the company has another opportunity to

customize the product offering.

In this paper, we find that a firm has the potential to increase revenues by strate-

gically restricting the set of product options it makes available to customers, even

when all products are in stock. In other words, the company may conceal a product

short on inventory or choose to show the product as not available, in anticipation of

future sales to other customers who may have a stronger preference for this product.

We formulate this as a dynamic assortment optimization problem where the assort-

ment decision depends on the inventory levels, the current customer’s preferences,

and the distribution of preferences of future customers.

We consider this problem in a setting with multiple products in a retail category.

28

Page 37: Assortment Planning with Inventory Considerations in

There are limited inventories of the products to sell over a finite selling season. The

customer base is heterogeneous and characterized by multiple customer segments

with different product preferences. The retailer is able to limit the assortment

to each arriving customer without cost and every customer purchases at most one

product. In this setting, we first characterize the optimal dynamic assortment

policy for the case with a single customer segment and multiple product variants.

We next characterize the optimal policy in a setting with two customer segments

and two product variants. Next, we perform a numerical study to understand the

revenue impact of assortment customization and provide insights on how the policy

parameters depend on inventory levels and the heterogeneity within and between

customer segments. Finally, we propose upper and lower bounds for the retailer’s

profit function and develop heuristic solutions for the general case.

When there is a single customer segment, it is optimal to offer the entire assort-

ment to any arriving customer, because all future customers have the same prefer-

ences in expectation. In other words, we prove that the optimal assortment in each

period is to offer all products with positive inventories. In the setting with two

customer segments and two product variants, one customer segment is more likely to

choose one product rather than the other, and vice-versa. In this case, it may not

be optimal to offer all available products to all arriving customers. In particular, we

show that the optimal assortment policy is a threshold-type policy. Each customer

segment is always offered its most preferred product. The least preferred product

is offered only when its inventory level is above a threshold value. This threshold

is decreasing in time. In other words, as we approach the end of the season, it is

more likely that both products will be offered to both customer segments. In terms

of comparative statics, we find that as a product becomes relatively more popular

(i.e., with a larger utility) for any customer segment, the benefit of reserving this

product for future customers is higher – that is, it is less likely that this product will

29

Page 38: Assortment Planning with Inventory Considerations in

be offered to a customer segment with a smaller utility for this product. We also

find that the revenue impact of assortment customization is significant in scenarios

where the inventory of one product is relatively large and the other one is relatively

small.

The rest of the chapter is organized as follows. Section 2.2 provides a review of

the relevant literature. Section 2.3 formulates the model. Section 2.4 characterizes

the optimal dynamic assortment policy for the case with a single customer segment

and multiple product variants. It also includes the analysis of a setting with two

customer segments and two product variants. Section 2.5 presents a numerical study

exploring the potential revenue benefits of assortment customization. In Section

2.6, we derive upper and lower bounds for the value function and develop a heuristic

solution for the general case with multiple segments and product variants. Section

2.7 concludes the chapter. All proofs are provided in Appendix D.

2.2 Literature Review

Our work is related to three streams of research. The first one is the literature

on retail assortment planning, with papers focusing on assortment decisions for a

single customer segment. Kok et al. (2008) provide a review of this literature.

van Ryzin and Mahajan (1999) derive the optimal assortment policy for a line of

products in a newsvendor setting. Mahajan and van Ryzin (2001) and Honhon

et al. (2008) consider models with dynamic customer substitution, while Cachon

et al. (2005) incorporate consumer search costs in a similar context. Smith and

Agrawal (2000) discuss an optimization approach for the assortment selection and

inventory management problems in a multi-item setting with demand substitution

and one replenishment cycle. Kok and Fisher (2007) focus on demand estimation

and assortment optimization in a retail chain. Caro and Gallien (2007) analyze a

30

Page 39: Assortment Planning with Inventory Considerations in

model of dynamic assortment selection with demand learning during a single sell-

ing season. Chen and Bassok (2009) study the impact of choice uncertainty and

aggregate demand uncertainty on optimal assortment and inventory decisions.

Another related stream of research is that on choice-based network revenue man-

agement. Zhang and Cooper (2005) study a revenue management problem in a

setting with parallel flights and customer-choice behavior. The paper derives upper

and lower bounds for the value function. Talluri and van Ryzin (2004a) build up

a framework for choice-based network revenue management models with multiple

products and components. Each product consists of multiple components and cus-

tomers follow a choice process. The optimal assortment policy is characterized in

Talluri and van Ryzin (2004b) when all components share the same resource. For

general choice-based network revenue management models, Gallego et al. (2004) and

Liu and van Ryzin (2008) use a choice-based linear programming (CDLP) model to

approximate the dynamic control problem. In addition, Liu and van Ryzin (2008)

propose a dynamic programming decomposition heuristic and characterize the effi-

cient sets which are used in the optimal policy. Zhang and Adelman (2009) use an

affine function to approximate the value function and Chen and Homem de Mello

(2009) develop an approximation that consists of a sequence of two-stage stochastic

programs with simple recourse. Miranda Bront et al. (2009) show that the problem

is NP-hard and study a column generation algorithm on the CDLP of more gen-

eral models. van Ryzin and Vulcano (2008a, b) study virtual nesting policies in a

similar context where the demand process is a stochastic sequence of heterogeneous

customers.

The ability of a company to limit the assortment to its customers is a form

of inventory rationing. Therefore, our paper is also related to research on this

topic. Ha (1997a) considers a single-item, make-to-stock production system with

several demand classes and lost sales, and demonstrates that the optimal policy is

31

Page 40: Assortment Planning with Inventory Considerations in

characterized by rationing levels for each demand class, Ha (1997b) studies a similar

system with two demand classes and backordering. de Vericourt et al. (2002) extend

this model to one with multiple demand classes characterized by different backorder

penalty costs.

2.3 Model Formulation

We consider a retailer that sells a set of product variants within a retail category

over a finite selling season. The retailer decides an assortment to offer to each ar-

riving customer. Each product variant represents a combination of features that

does not change the functionality of a product, such as a color/size/design combi-

nation as in the example of basketball shoes. The set of all products is denoted

as N = {1, · · · , N}. The selling season has T periods. There is no replenishment

during the season. The initial amount of stock for all products is denoted by an

N -dimensional vector y0 = (y01, · · · , y0N). This model is applicable, for example,

to end-of-season sales for any category of products and for short-life-cycle products

with long procurement lead times.

There are M customer segments characterized by different product preferences.

The choice process of each customer follows the Multinomial Logit (MNL) model.

Suppose that the offered assortment is given by a subset S ∈ N . The utility derived

from choosing product i (i ∈ S) for a customer in segment m is umi + ξmi, where

umi is the expected utility derived from product i and ξmi is a random variable

representing the heterogeneity of utilities across customers. In addition, customers

can always choose to not purchase any product, receiving a utility um0 + ξm0. Each

customer chooses the product that offers the maximum utility. We assume that

ξmi are i.i.d. random variables following a Gumbel distribution with mean zero and

variance π2/6. The probability of a customer choosing product i that arises from

32

Page 41: Assortment Planning with Inventory Considerations in

this utility maximization problem is given by

qmi(S) =θmi∑

j∈S θmj + θm0

, i ∈ S ∪ {0}

where θmi = eumi . See Anderson et al. (1992) for more details on the MNL model

and Kok et al. (2008) for a comparison with other demand models. For simplicity,

we refer to θmi as the utility for product i of customer segment m and let Θm denote

the utility vector (θm1, θm2, · · · , θmN). Customers within a segment have identical

expected utilities for the same products. Therefore, there are in total M different

expected utility vectors, one for each of the M customer segments. The retailer

is able to estimate these utilities based on the customers’ purchase history. For

simplicity, we assume that the retailer knows the utility vector for each customer

segment. MNL models are commonly used by retailers and marketing firms to

identify multiple latent customer segments in the customer base and to estimate

purchasing behavior for each segment (see, e.g., Gupta and Chintagunta 1994 and

Wedel and Kamakura 1998).

We consider a Poisson arrival process and assume that at most one customer

arrives in each period. The sequence of events in each period is as follows: At

the beginning of the period, a consumer arrives with probability λ and the arriving

customer belongs to segment m with probability ρm. (Clearly,∑M

i=1 ρi = 1.) The

retailer has perfect information on the customer’s segment upon arrival. The retailer

offers an assortment (subset of the available products) to the customer. Next, the

customer makes a purchasing decision according to the choice process and the revenue

is received if a product is sold.

To isolate the effect of heterogeneous customers, we focus on a model with identi-

cal prices for all products, denoted by p. In a setting with non-identical prices, there

is a clear incentive for rationing (e.g., not offering a lower-priced product at certain

33

Page 42: Assortment Planning with Inventory Considerations in

levels of inventory), even with a homogeneous customer base (as in Talluri and van

Ryzin 2004b). Finally, for expositional simplicity, we assume that the salvage value

for unsold units at the end of season is zero.

We use the following notation throughout the paper. We let ei denote the

ith unit vector. In addition, we let S(y) be the set of products with positive in-

ventory. Additionally, we denote the cardinality of set S as |S| and let y−i =

(y1, · · · , yi−1, yi+1, · · · , yN).

2.3.1 The Dynamic Assortment Optimization Problem

Define the value function in period t as Vt(y|m), given the vector of inventories y

and that the customer arriving in this period is in segment m. Taking expectation

across all customer segments, the value function at the beginning of period t is given

by

Vt(y) =∑m∈M

ρmVt(y|m)

Provided that the current arriving customer is in segment m, the retailer needs

to select an assortment to maximize revenues for the current period and for the rest

of the season. Therefore, the optimality equation is given by

Vt(y|m) = maxS⊂S(y)

{∑i∈S

λqmi(S)(p+ Vt+1(y − ei)) + λqm0(S)Vt+1(y)

}+(1−λ)Vt+1(y).

(2.3.1)

The term inside the curled brackets is the value function assuming that the ar-

riving customer is in segment m (an arrival occurs with probability λ). For a given

choice of assortment S, this term includes the probability of selling one unit of vari-

ant i in S, earning a revenue of p in this period plus the revenue-to-go function in

period t+ 1 evaluated at the current inventory level minus the unit sold in period t.

The term also accounts for the possibility that the customer purchases nothing, in

34

Page 43: Assortment Planning with Inventory Considerations in

which case all inventories are left to the next period and the revenue is the profit-to-

go function in period t + 1 evaluated at the current inventory level. We maximize

this term over all possible subsets of variants with positive inventories. The last

term is the future value function if no customer arrives (with probability 1 − λ).

Therefore, in each period, the goal is to derive the optimal assortment of products

to offer to each customer segment so as to maximize current and future revenues.

The total optimal revenue over the selling season is given by V1(y0). In addition,

the terminal condition of this dynamic program is the value function in period T .

Because there are no more customers beyond the last period, the optimal policy is to

offer all products with positive inventory to any arriving customer regardless of the

segment the customer belongs to. Therefore, VT (y|m) =∑

i∈S(y)∪{0} λqmi(S(y))p.

The dimensionality of the action set in each period is 2|S(y)|M . Thus, the above

dynamic assortment optimization problem is intractable for large N and M .

Define ∆it+1(y) = Vt+1(y)− Vt+1(y− ei) as the marginal expected revenue of the

yi-th unit of inventory of product i in period t+1, where i ∈ S(y). Clearly, ∆it+1(y)

is nonnegative and cannot exceed the price p, i.e., p ≥ ∆it+1(y) ≥ 0. In addition,

∆it+1(y) = 0 when yi ≥ T − t+ 1, that is, when there is enough inventory to satisfy

demand throughout the remaining selling season. We can then rewrite (2.3.1), the

optimality equation, as

Vt(y|m) = maxS⊂S(y)

{∑i∈S

λqmi(S)(p−∆it+1(y))

}+ Vt+1(y) (2.3.2)

If product i is offered and sold in period t, the revenue is the price p minus the

lost opportunity revenue of this unit after period t, given by ∆it+1(y). Therefore, we

denote the effective marginal price in period t as pit(y) = p−∆it+1(y) if i ∈ S(y) and

pit(y) = 0 if i ∈ N \ S(y). Let us order these effective marginal prices in period t

given inventory levels y so that pi1t (y) ≥ · · · ≥ piNt (y). Note that the indeces (ij)Nj=1

35

Page 44: Assortment Planning with Inventory Considerations in

depend on the inventory vector y and the period t. We denote a set that has the

products with the first k largest effective marginal prices as Ak(y) = {i1, · · · , ik}.

Then, if the value function Vt+1(·) is known for each inventory level, the problem in

(2.3.2) reduces to a one-period assortment optimization problem. This problem has

been solved by Talluri and van Ryzin (2004b) for a single customer segment. We

now show that their result extends to the setting with multiple customer segments

considered in this paper.

Proposition 2.3.1. Suppose that the inventory level in period t is given by y.

Then, the optimal assortment for each customer segment in period t is an element

of {A1(y), · · · , AN(y)}.

This result shows that the optimal assortment for each customer segment is re-

stricted to one of N possible sets Ai(y) if the effective marginal prices (pit(y))i∈N are

known (these can be characterized by value function Vt+1(·)). In particular, when

the inventory level is large enough, say yi ≥ T − t+ 1, it is always optimal to provide

product i regardless of the arriving customer’s segment because product i’s marginal

revenue is p in this case. In general, evaluating the value function in each period

is complicated and therefore so is the computation of the effective marginal prices.

Nevertheless, Proposition 2.3.1 provides a preliminary result about the structure of

the optimal assortment to offer to an arriving customer in period t. In particular,

this result implies that there is a product – product i1 – that will be offered to any

customer arriving in period t. In the following, we develop additional structural

properties of the optimal assortment policy.

2.4 The Optimal Assortment Policy

In this section, we turn our attention to the dynamic assortment policy. We first

analyze the optimal assortment policy in a setting with a single customer segment

36

Page 45: Assortment Planning with Inventory Considerations in

and multiple product variants, and then move to a model with two customer segments

and two product variants.

2.4.1 Single Customer Segment and Multiple Product Variants

We drop the subscript denoting customer segment in this section and first present

results on the value function and marginal expected revenue.

Proposition 2.4.1. (i) ∆it(y) is decreasing in yj, j ∈ N .

(ii) If yi = yj ≥ 1 and θi ≥ θj, then ∆it(y) ≥ ∆j

t(y).

(iii) Both Vt(y) and ∆it(y) are decreasing in t.

The first property in Proposition 2.4.1 implies that the value function is concave

in yi and submodular in yi and y−i. In other words, the marginal expected profit of

a product decreases as the inventory level of any product increases, because a higher

inventory level lowers the probability that any existing unit will be sold over the

remaining selling season. The second property shows that a product with a higher

expected utility has a higher marginal profit than a product with a lower expected

utility, when both products have the same inventory level. The last part indicates

that both the value function and the marginal expected profit decrease as there is

less time remaining in the selling season, because the probability that a unit will be

sold decreases with time t.

Based on these results, we can characterize the optimal assortment in this setting.

Theorem 2.4.1. Consider a setting with a single customer segment and N products.

The optimal assortment in each period is to offer all products with positive inventory.

Theorem 2.4.1 shows that, in this setting, it is optimal to follow the myopic

policy, offering all products in each period thus maximizing the probability of selling a

37

Page 46: Assortment Planning with Inventory Considerations in

product in that period. Because all product variants have identical prices, optimizing

the profit over the selling season is equivalent to maximizing total sales. In addition,

because customers are homogeneous, all future customers have the same preferences

as the current arriving customer in expectation, so it does not make sense to reserve

a product in anticipation of future sales.

2.4.2 Two Customer Segments with Two Product Variants

We now consider a setting with two customer segments and two product variants.

The utility vectors for customer segments 1 and 2 are denoted as (θ11, θ12, θ10) and

(θ21, θ22, θ20), respectively. We assume, without loss of generality, that both customer

segments have the same utility for the no-purchase option, i.e., θ0 = θ10 = θ20 > 0.

In addition, we restrict attention to the symmetric case where θ11 ≥ θ12, θ22 = θ11,

and θ21 = θ12. In other words, we consider a setting in which segment 1 customers

are more likely to purchase product 1 if both products are available, while segment

2 customers are more likely to purchase product 2 if both products are available. In

addition, when only product i is offered to both customer segments, a customer from

segment i purchases product i with a higher probability than a customer from the

other segment. Therefore, these two customer segments are distinctively different

in terms of their product preferences.

We begin by identifying some properties of the value function and marginal ex-

pected profits.

Proposition 2.4.2. (i) ∆it(y1, y2) is decreasing in yj, i, j ∈ {1, 2}.

(ii) Both Vt(y) and ∆it(y) decrease in t.

As shown for the case of a single customer segment, the value function is concave

in yi and submodular in y, and both the value function and the marginal expected

38

Page 47: Assortment Planning with Inventory Considerations in

profit decrease as time approaches the end of the selling season. Subsequently, we

turn attention to the optimal assortment policy.

Theorem 2.4.2. Consider a setting with two customer segments and two product

variants. The following properties characterize the optimal assortment policy.

(i) Product i is always included in the optimal assortment of an arriving customer

of segment i.

(ii) In each period, given the inventory level yi for product i, there exists a thresh-

old level y∗j (yi) such that the optimal assortment for a segment i customer is

determined as follows:

If yj ≥ y∗j (yi), then the optimal assortment is {1, 2}; If yj < y∗j (yi), then the

optimal assortment is {i}.

The result shows that each customer segment is always offered its most preferred

product in the optimal dynamic assortment policy. Furthermore, a segment i cus-

tomer is also offered product j if the inventory level of that product is large enough.

Otherwise, the customer is only offered product i. This policy is then characterized

by a set of threshold levels y∗i and y∗j in each period t. Because the products prices

are equal, it is optimal to sell as many units of either product as possible throughout

the selling season. Then, for an arriving customer of segment i, if the inventory level

of product j is low, it may be optimal to reserve that inventory for future arriving

customers of segment j, since those customers are more likely to leave without pur-

chasing any product if product j is not available. On the other hand, if the inventory

level of product j is large enough, then it is optimal to offer both product variants

as future sales to segment j customers are not highly compromised. Furthermore,

it may be optimal to reduce the inventory level of product j, rather than increasing

the chances of consuming further units of product i.

39

Page 48: Assortment Planning with Inventory Considerations in

As suggested by Theorem 2.4.2, in this setting with multiple customer segments,

it may not be optimal to offer all product variants to each arriving customer, even

when the selling prices are the same for all products. This indeed suggests a different

reason for rationing inventory that is not related to differences in prices. The

following result shows that in each period, at least one customer segment is offered

both products.

Proposition 2.4.3. In each period t, the optimal assortment for at least one cus-

tomer segment is {1, 2}.

To conclude the characterization of the optimal policy in the setting with two

customer segments and two products, we next discuss some monotonicity properties

of the threshold levels.

Proposition 2.4.4. There are two monotonicity properties of the threshold levels.

(i) For any given period, let y∗i (yj) and y∗j (yi) be the thresholds for inventory levels

yj and yi, respectively. If yi ≥ y∗i (yj) and yj ≤ y∗j (yi), then y∗i (yj) is decreasing

in yj.

(ii) For fixed inventory levels, the threshold levels decrease as time approaches the

end of the selling season.

The first part of Proposition 2.4.4 implies that, if in a given period and for given

inventory levels yi and yj, it is optimal to offer both products to customers from

segment j and to offer product i only to customers from segment i, then the same

policy continues to be optimal for lower inventory levels of product j. The second

part of Proposition 2.4.4 indicates that, as we get closer to the end of the selling

season, there is more incentive to offer both variants to any arriving customer. This

is because towards the end of the season, there is less concern about future sales

40

Page 49: Assortment Planning with Inventory Considerations in

and therefore it is optimal to sell as many units as possible. Finally, we note that

because of the submodularity of the value function in y, the problem of choosing

the initial inventory levels y0 with consideration of the associated procurement costs

and optimal revenues can be easily solved to optimality.

2.5 Numerical Study

We conduct a numerical study to examine the effect of assortment customization on

revenues. We also study the impact of various parameters related to the customer

segmentation on the optimal threshold values. The numerical study considers gen-

eral parameter values for the two products and two customer segments, with θii ≥ θij

and θii ≥ θji for i, j = 1, 2, i 6= j. We find that the policy presented in Section 4 for

the symmetric case is also optimal in this setting.

The study consists of 735 experiments – all with two customer segments and two

product variants, p = λ = 1, and T = 50. The utility vector (θ11, θ12) for segment 1 is

equal to one of the following vectors: {(10, 2), (10, 4), (10, 6), (10, 8), (10, 10), (12, 6),

(15, 6)}. Similarly, the utility vector (θ21, θ22) for segment 2 is given by one of

the following vectors: {(2, 10), (4, 10), (6, 10), (8, 10), (10, 10), (6, 12), (6, 15)}. The

utility of the no-purchase option is given by one of the following five values: θ0 =

{5, 10, 15, 20, 25}. For each of these, the probability that an arriving customer

belongs to segment 1 takes one of the following three values: ρ1 = {0.2, 0.5, 0.8}.

2.5.1 Effect of Assortment Customization on Revenues

We first report the effect of assortment customization on revenues by calculating the

percentage revenue increase due to assortment customization relative to a policy that

always offers both products to any arriving customer.

41

Page 50: Assortment Planning with Inventory Considerations in

Figure 2.1: Revenue impact of assortment customization

Figure 2.5.1 below shows the average percentage increase in revenue due to as-

sortment customization as a function of the inventory levels. When inventory levels

for both products are large, the retailer is less likely to run out of stock, so there is

no need for rationing inventory and therefore the revenue impact is small. When

inventory levels for both products are small relative to the time horizon, there is

enough time to sell all units, so both products are likely to be offered in the optimal

solution and again the impact is small. The revenue impact is significant in the

areas where the inventory level for one product is high while the inventory level for

the other product is low. These are the scenarios in which strategically selecting

assortments is most important.

We now consider the average revenue increase over the inventory levels under

which the percentage increase due to customization is greater than 1%. The revenue

impact is smaller when the customers of one segment represent a higher proportion of

the total customer base. This is because, in that case, most arriving customers be-

long to one of the segments and therefore the possibility of making use of customized

assortments is reduced. In the extreme case, when there is a single customer seg-

42

Page 51: Assortment Planning with Inventory Considerations in

ment, it is always optimal to offer all available products as shown in Section 4.1,

thereby eliminating any customization. We also find that the revenue impact is

sensitive to significant changes in the utility of a customer segment’s least preferred

product. As the utility of a segment’s least preferred product increases, offering this

product will reduce the extent to which customers in that segment purchase their

most preferred product. In those cases, it may then be optimal to limit the assort-

ment offered to those customers, thus curtailing their access to the least preferred

product. Therefore, the revenue impact of assortment customization is larger. If

the utility of the least preferred product decreases, there is less incentive to offer

this product as it may be better to reserve the inventory of that product for the

future customers of the other segment. Again, the revenue impact of customization

increases. As the utility of the no-purchase option increases, i.e., customers are

more likely to leave without purchasing any product, total demand decreases and

the revenue impact of customization is lower.

2.5.2 Effect of Parameters on Threshold Levels

We next explore the impact of changes in the system parameters on the optimal

threshold levels. We focus on segment 1 customers. Similar observations can be

made for segment 2 customers.

Note that, for an arriving customer of segment 1, it is optimal to offer product

2 if the inventory level of that product is higher than the threshold level, i.e., if

y2 ≥ y∗2(y1). We first note that, for a fixed inventory level y1, this threshold is

increasing in θ22−θ21. For example, when θ22 increases and θ21 decreases, a segment

2 customer is even more likely to purchase product 2 than product 1. It thus makes

sense to reserve product 2 for future arriving customers of segment 2. Therefore,

the threshold level is increasing in this difference. In addition, as either θ12 or θ22

increases (one customer segment has higher relative preference for product 2), the

43

Page 52: Assortment Planning with Inventory Considerations in

demand for that product becomes stochastically larger. Then, the advantage of

offering only product 1 to a customer in segment 1 becomes more significant, so that

more units of product 2 are reserved for customers in the other segment. That is,

the threshold level increases with θ12 and θ22. However, this threshold is decreasing

as either θ11 or θ21 increases (when customers in either segment are more likely to

purchase product 1 than product 2). In this case, making both products available

to segment 1 customers is less likely to have a severe impact on product 2’s future

inventory levels, and therefore on overall sales.

We also find that, as the proportion of customers from segment 1, i.e., ρ1, in-

creases, offering both products to segment 1 customers becomes more attractive as

this has the potential of increasing sales. Therefore, the threshold level y∗2(y1) de-

creases with ρ1. This observation is consistent with the effect of ρ1 on the revenue

impact of assortment customization discussed before. Similarly, the threshold level

decreases with the utility of the no-purchase option.

Finally, Figure 2.5.2 exhibits the monotonicity properties of the threshold levels

shown in Proposition 2.4.4. As we get closer to the end of the season (i.e., we

move from the top curve to the bottom curve in Figure 2.5.2), the threshold levels

decrease, that is, less rationing occurs as time approaches the end of the selling

season. Moreover, in any given time period, the threshold level y∗2 is increasing

in y1. That is, for a given amount of inventory of product 2, it is more likely to

offer that product to an arriving customer of segment 1 when the inventory level of

product 1 is low. This is because it may be optimal to re-direct some segment 1

customers to purchase product 2 so as to reduce the likelihood of running out of stock

of product 1 in the future. On the other hand, when the inventory level of product

1 is high, it may be optimal to limit the assortment to product 1 only, therefore

increasing demand for that product.

44

Page 53: Assortment Planning with Inventory Considerations in

Figure 2.2: Monotonicity of threshold levels

2.6 General Case

In this section, we consider the general problem with N products and M customer

segments. We first develop upper and lower bounds on the optimal profit and then

propose a heuristic solution.

2.6.1 Bounds on the Optimal Profit

The dynamic assortment optimization problem with N products and M customer

segments can be described as a Markov decision process (MDP) problem. Assort-

ments over all periods are controlled through a control policy µ. The assortment

offered to a segment m customer in period t is Sµt (m|Ft) where Ft is the history up

to period t. We denote the profit under a control policy µ, given initial inventory

levels y0, as V µ1 (y0). The set of all feasible policies is denoted as A. Therefore,

V1(y0) = maxµ∈A

V µ1 (y0)

We first propose a lower bound to the optimal profit. Define a threshold policy

(µT ) as follows: let (bmt1, · · · , bmtN) a vector of threshold levels for customer segment

45

Page 54: Assortment Planning with Inventory Considerations in

m in period t, depending on the prevailing inventory level y in period t, such that the

set of products to offer to an arriving customer of segment m is Smt = {i : yi ≥ bmti}.

The set of all feasible threshold policies is denoted as T . In particular, let µAll

represent the control policy under which any customer is offered all products with

positive inventory. Then, we have the following lower bounds of the optimal profit:

V µAll

1 (y0 ) ≤ maxµT∈T{V µT

1 (y0 )} ≤ V1(y0 )

Next, we propose an upper bound on the optimal profit by converting the stochas-

tic control problem into a choice-based deterministic linear programming problem

(CDLP). This approach has been used in Gallego et al. (2004) and Liu and van

Ryzin (2008) in the context of settings with a single customer segment. The cus-

tomer arrival and choice processes are assumed to be deterministic, the first with

rate λρm for customer segment m and the second with a proportion qmj(S) of cus-

tomers in segment m choosing product j ∈ S. We also assume that demand and

inventory are continuous variables. The total revenue for segment m under an as-

sortment S is then given by Rm(S) =∑

j∈S ρmpqmj(S) and the associated inventory

consumption rate is Qm(S) =∑

j∈S ρmqmj(S). Given a sequence of assortments

offered to segment m customers arriving over the selling season, any permutation of

this sequence leads to the same revenue and inventory consumption. Therefore, in

this approximation, we derive the length of time in which an assortment set is offered

to a given customer segment. Specifically, let tm(S) denote the interval of time over

46

Page 55: Assortment Planning with Inventory Considerations in

which assortment S is offered to segment m customers. The CDLP is then given by

V CDLP (y0) = max∑m∈M

∑S⊂N

λRm(S)tm(S)

s.t.∑m∈M

∑S⊂N

λQm(S)tm(S) ≤ y0i, i ∈ N

∑S⊂N

tm(S) ≤ T, ∀m ∈M

tm(S) ≥ 0, ∀m ∈M and ∀S ⊂ N . (2.6.1)

The CDLP is not only simpler than the MDP problem, but also each feasible solution

to the MDP has a corresponding feasible solution in the CDLP.

Proposition 2.6.1. For each feasible policy in the MDP there is a corresponding

feasible solution in the CDLP with the same profit. Therefore,

V1(y0) = maxµ∈A

V µ1 (y0) ≤ V CDLP (y0).

This result implies that the optimal profit of the CDLP in (2.6.1) is an upper

bound to the MDP problem.

2.6.2 Multiple Customer Segments and Multiple Product Variants

In this section, we briefly introduce a heuristic for the general assortment customiza-

tion problem with M customer segments and N product variants. As shown in

Proposition 2.3.1, the optimal assortment for any customer segment in period t,

given an inventory vector y, is given by one of the sets Ak(y), defined based on

the effective marginal prices p1t (y), · · · , pNt (y). In other words, for known values of

Vt+1(y) and Vt+1(y − ei) for all i, the computation of the optimal assortment for a

given customer segment is reduced to a search over N possible product sets. Nev-

ertheless, the size of the state space still makes it intractable to derive the optimal

47

Page 56: Assortment Planning with Inventory Considerations in

policy by solving the MDP. We therefore propose a heuristic based on the bounds

derived above.

Heuristic

Step 1: Consider the vector of inventory levels y in period t. Let S(y) be the set

of products with positive inventory level and denote s = |S(y)|.

Step 2: Calculate the bounds V boundt+1 (y) and V bound

t+1 (y − ej) for all j ∈ S(y).

Step 3: Set pjt(y) = p − V boundt+1 (y) + V bound

t+1 (y − ej) for each j ∈ S(y). Rearrange

(pjt(y))j∈N in decreasing order so that pi1t (y) ≥ · · · ≥ piNt (y) and let Aj =

{i1, · · · , ij} for all j ≤∣∣S(y)

∣∣.Step 4: Determine the optimal assortment for each customer segment m by compar-

ing the one-period profit values∑

i∈Aj λqmi(Aj)pit for each j ≤

∣∣S(y)∣∣. Offer

the assortment with the highest profit to an arriving customer of segment m.

The performance of the heuristic depends on the gap between the solution given

by the bound and the optimal solution to the MDP. We consider two versions of

the heuristic, heuristic one using the upper bound CDLP and heuristic two using

the lower bound that involves offering all products with positive inventory to each

arriving customer. In this section, we conduct a numerical study to examine the per-

formance of the bounds and heuristics. The study consists of 48 cases – all with four

customer segments, four product variants, p = 10, λ = 1, and θ0 = 1. The probabil-

ity vectors ρ = (ρ1, ρ2, ρ3, ρ4) are equal to one of the six vectors: ρ = (ρ1, ρ2, ρ3, ρ4) ∈

{(0.1, 0.1, 0.4, 0.4), (0.2, 0.2, 0.3, 0.3), (0.1, 0.2, 0.3, 0.4), (0.1, 0.1, 0.1, 0.7),

(0.2, 0.2, 0.2, 0.4), (0.3, 0.3, 0.3, 0.1)}. The preference vector for segment i(∈ {1, 2, 3, 4})

customers takes one of the following four different forms : Θi = {(θii = 10, θij = 1),

48

Page 57: Assortment Planning with Inventory Considerations in

(θii = 10, θi,i+1 = 9, θij = 1), (θii = 10, θi,i+2 = 9, θij = 1), (θii = 10, θi,i+3 = 9, θij =

1)} where j ∈ {1, 2, 3, 4}\{i} and θi,i+1(θi,i+2 or θi,i+3) is the preference of segment i

customers for product i− 3(i− 2 or i− 1) if i > 3(i > 2 or i > 1). In another group

of experiments, the preference vectors of segments 1, 2, and 3 customers take one of

these four forms, however, segment 4 customers are almost indifferent among the four

products, i.e., Θ4 = (9, 9, 9, 10). For each case, we compute the optimal profit under

the optimal dynamic assortment policy, the profits derived from upper and lower

bounds, and the profits based on both heuristics. We report the average maximum

gap and the median across all cases. The average percentage maximum gap and the

median between the upper bound CDLP and the optimal profit are 7.3% and 1.5%,

respectively, while these numbers are −4.1% and −1.2% between the lower bound

and the optimal profit, respectively. The lower bound is closer to the optimal profit

than the upper bound. Regarding the heuristics, heuristic two, based on the lower

bound, performs much better than heuristic one based on the CDLP. In particular,

the average percentage maximum gap and the median between the lower bound and

the optimal profit are 0.3% and 0.0%, respectively.

2.7 Conclusion

We consider a revenue management problem for a firm that offers multiple product

variants in a retail category. Customers arrive according to a Poisson process and

they are segmented according to their expected preferences for the product variants.

The firm has information on which segment each arriving customer belongs to and

makes customized assortment decisions so as to maximize revenues over the selling

horizon. We prove that it is optimal to offer all product variants in the case with a

single customer segment. However, when there are multiple customer segments with

distinctly different preferences for the products, it may not be optimal to offer all

49

Page 58: Assortment Planning with Inventory Considerations in

the available products to each customer. In particular, in the case of two customer

segments and two products, the optimal assortment policy is a threshold-type policy.

Each customer is always offered its most preferred product, while the other product

is offered only if that product’s inventory level is larger than a threshold value.

We assess the impact of offering customized assortments in a numerical study and

characterize the settings in which assortment customization is most valuable. The

revenue impact is significant when the inventory level of one product is relatively high

and that of the other product is relatively low. We also find that the revenue impact

is highest when there is a relatively equal split of customers in both segments. We

also study the effect of changes in the system parameters on the optimal threshold

values.

We derive upper and lower bounds on the value function and use these to pro-

pose two heuristics for the general case of multiple products and multiple customer

segments.

50

Page 59: Assortment Planning with Inventory Considerations in

3

Assortment Decisions with Non-identical Price andCost Parameters

3.1 Introduction

In Chapter 1, we assume that the price and cost parameters of the product variants

are identical. While this may be applicable to certain product categories (such as

similar style shoes or apparel), in many cases products may have different selling

prices and costs of components. In this section, we discuss the generalization of our

results to the case of non-identical costs and prices.

Let us start with the system with dedicated components. van Ryzin and Mahajan

(1999) claim that they “can allow different prices and costs, provided the ratio p/c

is the same for all variants.” This claim is not correct. In the absence of inventory

costs (therefore no economies of scale), the optimal assortment in a setting with non-

identical margin consists of the k highest-margin items (Liu and van Ryzin 2008).

With inventory costs and identical margins, the optimal assortment consists of a

subset of the most popular products (van Ryzin and Mahajan 1999). Hence, with

both explicit inventory costs and non-identical margins, there are two forces at play

51

Page 60: Assortment Planning with Inventory Considerations in

and the optimal assortment may not necessarily have any particular structure. That

is, it is possible that the optimal assortment contains a product that is neither the

highest-margin product nor the one with the highest level of popularity, even if the

ratios p/c are the same for all variants.

The following example illustrates the difficulty of obtaining structural results

for products in a setting in which popularity is decreasing with prices: N = 3,

λ = 100, β = 1/2, σ = 3, cj/pj = r = 0.7, uj = aj − pj, (aj)j=0,1,2,3 = (10, 13, 14, 15) ,

(pj)j=0,1,2,3 = (0, 6, 8, 11) , (uj)j=0,1,2,3 = (10, 7, 6, 4) , where uj is the net utility, aj is

the inherent utility, and pj is the selling price. Note that the net utility is decreasing

in j and the price is increasing in j for j ≥ 1. The optimal assortment is S∗ = {2} ,

which consists of a product that is neither the most popular, nor the highest margin

product.

Hence, with non-identical prices, it is generally not possible to derive structural

results regarding the optimal assortment, even in the system with dedicated compo-

nents. In this chapter, our first goal is to demonstrate the impact of commonality

on variety, and we are able to show that our results continue to hold for some set-

tings with non-identical prices, namely, for settings in which prices (and margins)

are weakly increasing in the net utility of the product. This is the only range of

non-identical cost and price parameters in which there exists a clear measure of level

of variety and, therefore, meaningful comparisons across systems are possible. We

then derive some results related to the structure of the optimal assortment with

general price and cost parameters in the system with dedicated components.

3.2 Extended Results of Chapter 1 with Non-Identical Price and CostParameters

It is reasonable to assume that a product with a higher inherent utility has a higher

production cost and selling price. However, the choice probabilities (and popularity)

52

Page 61: Assortment Planning with Inventory Considerations in

are determined by the average net utility, which equals inherent utility minus the

selling price. The relation between net utilities and prices depends on the setting.

Casual observation suggests that there is a non-monotone relationship between prices

and popularity: the most popular products are usually those with medium ranged

prices. For example, a quick online search shows that Toyota Camry and Honda

Accord are the top selling cars in 2009 in the U.S. Note that a product that has a

higher price and a lower utility than another product may still attract some customers

due to the heterogeneity of customer preferences.

Below we state our results for the independent population model and the trend-

following model. For expositional simplicity, we assume linear functions, pj = p0 +

γuj, where p0, γ ≥ 0. The total cost of components for variant i is ki = kdi + kci .

We assume that the cost/price ratios are identical across all products, i.e., ki/pi = r

for r ∈ (0, 1) and any i ∈ N .

Theorem 3.2.1. Under the independent and trend-following population demand

models, the optimal assortment in the dedicated model is a popular set.

Because unit costs are different across products, we need to replace Condition 2

in the paper with the following analogous condition, which implies that the optimal

assortment becomes larger as r decreases or, equivalently, as the unit costs of all

products decrease.

Condition 3. δn is increasing in r as long as δn < θn.

We now describe the cost structure in the system with commonality. For the

cost of the dedicated components, assume that kdi/pi = τr, where τ, r ∈ (0, 1). The

cost of the common component is given by kc and suppose that kc ≤ kc1 . Note

that kc1 is the cost of the most expensive version of that component. Similar to

the case with identical prices, we can solve the profit maximization problem for the

trend-following demand model by using the KKT conditions.

53

Page 62: Assortment Planning with Inventory Considerations in

Theorem 3.2.2. In the system with commonality under the trend-following model,

the optimal assortment is a popular set. For any popular set An, we have that if

ΠD,TFAn+1

≥ ΠD,TFAn

then ΠC,TFAn+1

> ΠC,TFAn

. Thus, the optimal assortment in the system

with commonality is no smaller than that in the dedicated system.

Finally, we consider the pooled system, in which there are no dedicated compo-

nents and all products use the same common component. Note that, in the case of

non-identical prices, it is no longer optimal to offer all products in the pooled system.

Assume that the total cost of common components is k = kd + kc, with kd ≤ kd1 .

Theorem 3.2.3. In the pooled system under the trend-following demand model, the

optimal assortment is a popular set. Furthermore, the optimal assortment in pooled

system is no smaller than that in the dedicated system ifkqNN∑i∈N piqi

≤ r. The optimal

assortment in the pooled system is no smaller than that in the system with universal

commonality if k ≤ kd1 + kc.

This results show that in the pooled system it is optimal to offer more prod-

ucts than in the system with commonality as long as the cost of the new common

component in the pooled system (the one that is not common in the commonality

system) is not higher than the most expensive dedicated component the system with

commonality .

3.3 Dedicated System with General Price and Cost Parameters

The retailing price of product i is pi and its utility is ui. Since all prices are given,

we denote the utility of product i as θi = eui−pi . The production cost of product i is

ki = αpi where 1 > α > 0. The total demand is a Poisson process with rate λ. For

54

Page 63: Assortment Planning with Inventory Considerations in

any assortment S, the profit function is

ΠS = (1− α)λ∑i∈S

piqSi − φ(Φ−1(1− α))λβ

∑i∈S

pi(qSi )β

= pi

[(1− α)λ

∑i∈S

qSi − φ(Φ−1(1− α))λβ∑i∈S

(qSi )β

]

From Lemma B.1, we have that ΠS is increasing and convex in qSi and is linearly

increasing in pi. Suppose the new variant is denoted as product m with utility δ and

price pm. Define δm the value of δ > 0 such that ΠS+{δm},pm = ΠS. Then it is easy

to get the following proposition.

Proposition 3.3.1. If i, j ∈ S and pi ≥ pj, then δi ≤ δj.

This proposition implies that it is better to include a product with a higher price

if they have the same utility. However, if prices and utilities are ordered differently,

it is unclear which one is better to be included. We then define assortment S as an

efficient set, if S satisfies the following three conditions.

1. ΠS ≥ ΠS+{i} for any i ∈ S.

2. ΠS ≥ ΠS−{j} for any j ∈ S.

3. ΠS ≥ ΠS−{j}+{i} for any j ∈ S and i ∈ S.

According to the definition, it is not optimal to add, remove or replace any

product from an efficient set. Obviously, the optimal assortment will be an efficient

set. Moreover, we have the following property for these efficient sets.

Proposition 3.3.2. In any efficient set S, for any i ∈ S, if pj > pi and θj ≥ θi or

pj ≥ pi and θj > θi, then j ∈ S.

55

Page 64: Assortment Planning with Inventory Considerations in

Figure 3.1: Efficient set

This proposition says that, for any product in an efficient set, all other products

who both have a larger price and utility than this product are contained in this set as

well. As show in Figure 3.3, if product A is included in an efficient set, then product

B and C are also in this efficient set.

We then can get the results below for two special cases easily by using Proposition

3.3.2.

Proposition 3.3.3. 1. If all products have the same utility, the optimal assort-

ment is a set which includes the most profitable products.

2. (van Ryzin and Mahajan (1999)) If all products have the same price, the opti-

mal assortment is a set which includes the most popular products.

56

Page 65: Assortment Planning with Inventory Considerations in

Appendix A

Proofs of Main Results of Chapter 1

Analysis of the Dedicated System

In a system with dedicated components, the profit function is separable. For

a given assortment S, the profit function under the independent population model

is ΠD,IPS =

∑i∈S π

D,IPi (qSi )

def=∑

i∈S maxyi{E[pmin{yi, Di}] − kyi}, where πi is the

profit function and yi is the stocking quantity for variant i. The optimal stocking

quantities are given by y∗i = λi+zIPσi, where zIP = Φ−1(

1− kp

). The resulting op-

timal profit is ΠD,IPS =

∑i∈S(p−k)λi−pφ(zIP )σi. Under the trend-following model,

the optimal expected profit, given an assortment S, is ΠD,TFS =

∑i∈S π

D,TFi (qSi )

def=∑

i∈S maxyi{E[pqSi min{yi, D}]−kyi}. The optimal stocking quantity for variant i is

given by y∗i = λ+ zTF (qSi )σ, where the value of zTF (qSi ) depends on the profitability

of variant i as follows:

zTF (qSi ) =

{Φ−1

(1− k

qSi p

), if qSi p ≥ k

−µσ, otherwise.

If qSi p < k, then the optimal stocking level for variant i is zero. As a result, the

57

Page 66: Assortment Planning with Inventory Considerations in

profit for variant i is (qSi p− k)λ− qSi pσφ(zTF (qSi )) if qSi p ≥ k, and zero otherwise.

Given an assortment S, we let S(δ) = S ∪ {m} denote the set of variants in

S plus an additional variant m 6∈ S with utility δ. The choice probability for

i ∈ S is qS(δ)i = θi∑

j∈S θj+δ+θ0and q

S(δ)m = δ∑

j∈S θj+δ+θ0. The resulting optimal profit is

ΠDS (δ) =

∑i∈S(δ) π

Di (q

S(δ)i ) under either demand model. Proposition B.1 in Appendix

B shows that ΠDS (δ) is a quasi-convex and decreasing-increasing function and that

the optimal assortment is a popular set. The proof is similar to that in van Ryzin

and Mahajan (1999).

Theorem A.1. ΠDn is quasi-concave in n.

Proof of Theorem A.1: Quasi-concavity of ΠDn in n is established if the following

are satisfied: 1) If ΠDn+1 ≥ ΠD

n , then ΠDn ≥ ΠD

n−1; and 2) If ΠDn+1 ≤ ΠD

n , then ΠDn+2 ≤

ΠDn+1. For (1), ΠD

n+1 ≥ ΠDn implies that δn ≤ θn+1 ≤ θn. Under Condition 1 we have

δn−1− δn ≤ θn− θn+1. Combining the two, we have that δn−1 ≤ δn + θn− θn+1 ≤ θn,

which implies that ΠDn ≥ ΠD

n−1. For (2), ΠDn+1 ≤ ΠD

n implies that δn ≥ θn+1. There

are two possibilities. If δn+1 > θn+1, then ΠDn+2 ≤ ΠD

n+1 because θn+1 ≥ θn+2. If

δn+1 ≤ θn+1, then Condition 1 implies that δn+1 ≥ δn − θn+1 + θn+2 ≥ θn+2, which

implies that ΠDn+2 ≤ ΠD

n+1. �

Proof of Theorem 1.4.1: Because all variants are symmetric with identical util-

ities, the optimal order quantities are the same for all variants under model D. In

model D, ΠD,IPn = maxy{pnEmin{y,Di} − n(kc + kd)y} where the optimal order

quantity y∗ satisfies Pr(Di ≤ y∗) ≥ 1 − kc+kdp

. We consider settings with y∗ > 0,

which can be guaranteed by assuming that λ >> σ. In turn, this implies that

Pr{DSi = 0} = σ2

λqS+σ2 < 1 − kc+kdp≤ Pr{DS

i = λqS + σ2} = λqS

λqS+σ2 and therefore

y∗ = λqS + σ2. Hence, ΠD,IPn = n(pλqS − (kc + kd)(λq

S + σ2)). Now consider

model C. The optimal order quantities for the dedicated components are equal to

y∗ in model C as well, because unit cost is smaller than that in model D due to

58

Page 67: Assortment Planning with Inventory Considerations in

risk pooling effect of the common component and the dedicated components face

the same two-point demand distribution as in model D. The common component,

however, faces the total demand. Moreover, y∗c = m∗y∗ where m∗ is an integer

between 1 and n, since the demand for each variant is either zero or y∗. Thus,

ΠC,IPn = (λqS + σ2) maxm{pmin{m,

∑ni=1 min{1, DS

i /(λqS + σ2)}} −mkc − nkd}.

Define Nn =∑n

i=1 min{1, DSi /(λq

S + σ2)} which subjects to a binomial distribu-

tion with parameters n and λqS

λqS+σ2 . Then the optimal m∗ satisfies Pr(Nn ≤ m∗) ≥

1− kcp

. We now use a normal distribution with mean E[Nn] = nλqS

λqS+σ2 and standard

deviation σ(Nn) =√n λqSσ2

(λqS+σ2)2to approximate the binomial distribution. The

optimal profit function in model C then can be expressed as

ΠC,IPn = (λqS + σ2)[(p− kc)E[Nn]− pφ(Φ−1(1− kc/p))σ(Nn)− nkd]

To ensure that the normal approximation of the binomial distribution is accurate,

we focus on cases that satisfy ΠC,IPn ≥ ΠD,IP

n . This inequality states that, for a

given assortment, the profit in the system with commonality is no smaller than that

in the system with dedicated components. This inequality is equivalent to

kcσ√λθpφ(Φ−1(1− kc/p))

√1

n(nθ + θ0). (A.1)

This condition is satisfied for sufficiently large n. We now proceed to prove that

ΠC,IPn+1 − ΠC,IP

n > ΠD,IPn+1 − ΠD,IP

n .

ΠC,IPn+1 − ΠC,IP

n − ΠD,IPn+1 + ΠD,IP

n

= σpφ(Φ−1(1−kc/p))√λθ

[√n

nθ + θ0

√n+ 1

(n+ 1)θ + θ0

+kcσ√

λθpφ(Φ−1(1− kc/p))

](A.2)

59

Page 68: Assortment Planning with Inventory Considerations in

Using the inequality (A.1) in expression (A.2), we have that

(A.2) ≥ σpφ(Φ−1(1− kc/p))√λθ

[√n

nθ + θ0

√n+ 1

(n+ 1)θ + θ0

+

√1

n(nθ + θ0)

]

= σpφ(Φ−1(1− kc/p))√λθn(nθ + θ0)

[n+ 1−

√(n+ 1)n(nθ + θ0)

(n+ 1)θ + θ0

]> 0.

This completes the proof. �

Proposition A.1. For a given assortment S, there exists a popular subset Se ⊂ S

such that y∗i = y∗c for all i ∈ Se and the optimal stocking quantities of the remaining

variants decrease in order of their popularity (i.e., y∗ir ≤ y∗is < y∗c for ir, is ∈ Sldef=

S \ Se with θir ≤ θis). Hence, the optimal profit is given by

ΠC,TFS =

∑i∈Sl

[(qSi p− kd)µ− qSi pσφ(zSi )

]

+

[(∑i∈Se

(qSi p− kd)− kc)µ−∑i∈Se

qSi pσφ(Φ−1(t∗c(S)))

]

with zSi = 1− kdpqSi

and t∗c(S) = 1− kc+|Se|kdp∑i∈Se q

Si

.

Proof of Proposition A.1: The following are the KKT optimality conditions for

the profit maximization problem subject to the constraints 0 ≤ yi ≤ yc for i ∈ S

(with dual variables αi):

αi ≥ 0, αi(y∗c − y∗i ) = 0,

∑j∈S

αj = kc

Φ

(y∗i − λσ

)= 1− kd + αi

pqSi. (A.3)

Based on the KKT conditions, we partition the variants in S into two sets, Sl and Se.

The set Sl corresponds to the product variants i for which αi = 0. A variant i ∈ Se

60

Page 69: Assortment Planning with Inventory Considerations in

has αi > 0, implying that y∗c = y∗i . Because∑

j∈S αj = kc, we have that Se 6= ∅.

This and (A.3) imply that there exists a fractile t∗c(S) such that y∗c = λ+Φ−1(t∗c(S))σ.

In other words, 0 < αi = pqSi (1− t∗c(S))− kd ≤ kc for all i ∈ Se. Using the fact that∑j∈Se αj = kc, we have that t∗c(S) = 1− kc+|Se|kd

p∑i∈Se q

Si

.

Suppose that S = {i1, · · · , ir} with i1 < · · · < ir, so that θi1 > · · · > θir and

therefore

1− kdpqSi1

> · · · > 1− kdpqSir

> 0.

Then, there exists a variant ia such that

1− kdpqSi1

> · · · > 1− kdpqSia

> t∗c(S) ≥ 1− kdpqSia+1

> · · · > 1− kdpqSir

and the assortment is partitioned as Se = {i1, · · · , ia} and Sl = {ia+1, · · · , ir}. (That

is, the set Se is a popular subset of S.) To see this, note that for i ∈ Sl, αi = 0 and

y∗i = λ+ Φ−1(

1− kd+αipqSi

)σ = λ+ Φ−1

(1− kd

pqSi

)σ ≤ λ+ Φ−1(t∗c(S))σ = y∗c (because

y∗i ≤ y∗c ). This implies that t∗c(S) ≥ 1 − kdpqSi

for all i ∈ Sl. Finally, note that for

i ∈ Se, the values of αi are such that

1− kd + αi1pqSi1

= · · · = 1− kd + αiapqSia

= t∗c(S). �

Proof of Theorem 1.4.2: Quasi-convexity of ΠC,TFS (δ) in δ can be demonstrated

by showing that the first-order derivative of ΠC,TFS (δ) is continuous and crosses zero

(from negative to positive) only once. (The function ΠC,TFS (δ) is evaluated at the

optimal stocking quantities for all variants in S.) As δ increases, we have from

Lemma B.4 that Se(δ) and Sl(δ) change as well, leading to different expressions of

the first-order derivative of ΠC,TFS (δ) with respect to δ. Consider the threshold values

δi, i = 1, 2, 3, 4, defined in Lemma B.4.

61

Page 70: Assortment Planning with Inventory Considerations in

If 0 ≤ δ < δ1, then it is not profitable to include the new variant in the assortment.

Therefore,

∂ΠC,TFS (δ)

∂δ=∑i∈S

∂ΠC,TFS (δ)

∂qS(δ)i

∂qS(δ)i

∂δ=

1

(∑

i∈S θi + δ + θ0)2

[−∑i∈S

θi∂ΠC,TF

S (δ)

∂qS(δ)i

](A.4)

If δ1 ≤ δ < δ2, then the new variant enters the set Sl, and

∂ΠC,TFS (δ)

∂δ=∑i∈S

∂ΠC,TFS (δ)

∂qS(δ)i

∂qS(δ)i

∂δ+∂ΠC,TF

S (δ)

∂qS(δ)m

∂qS(δ)m

∂δ

=1

(∑

i∈S θi + δ + θ0)2

[−∑i∈S

θi∂ΠC,TF

S (δ)

∂qS(δ)i

+∂ΠC,TF

S (δ)

∂qS(δ)m

(∑i∈S

θi + θ0

)](A.5)

If δ2 ≤ δ < δ3, then the new variant enters the set Se, and

∂ΠC,TFS (δ)

∂δ=∑i∈S

∂ΠC,TFS (δ)

∂qS(δ)i

∂qS(δ)i

∂δ+∂ΠC,TF

S (δ)

∂qS(δ)m

∂qS(δ)m

∂δ

=1

(∑

i∈S θi + δ + θ0)2

[−∑i∈S

θi∂ΠC,TF

S (δ)

∂qS(δ)i

+∂ΠC,TF

S (δ)

∂qS(δ)m

(∑i∈S

θi + θ0

)](A.6)

Note that, because m ∈ Se for δ2 ≤ δ < δ3, the expression for∂ΠC,TFS (δ)

∂qS(δ)i

in (A.6) is

different from that in (A.5). For δ ≥ δ3, the general form of the derivative is as

in (A.6), again with different expressions for∂ΠC,TFS (δ)

∂qS(δ)i

. From Lemma B.5, ΠC,TFS is

increasing and convex in qSi . In addition, for i ∈ S, qSi (δ) is decreasing in δ and qSm(δ)

is increasing in δ. Therefore, the terms within square brackets in equations (A.4),

(A.5) and (A.6) are continuous and increasing in δ. We then have that ΠC,TFS (δ)

is quasi-convex in δ for each of the δ-intervals. Also, note that the expression in

(A.4) is negative. In addition, for large δ, we have that Se = {m} and Sl = ∅, so

that∂ΠC,TFS (δ)

∂δis positive. It only remains to prove that the first-order derivative is

62

Page 71: Assortment Planning with Inventory Considerations in

continuous at each threshold δi. That will establish the result that the derivative

crosses zero from negative to positive only once. In what follows, we prove that

the derivative is continuous at δ1 and δ2, and similar results can be obtained for

δ ≥ δ3. For δ = δ1, note that y∗m(δ1) = 0, which implies from (B.2) of Lemma B.1

that∂ΠC,TFS (qSm(δ))

∂qSm(δ)|δ=δ1 = 0. Thus, the first-order derivative is continuous at δ1. For

δ = δ2, we have that y∗m(δ2) = y∗ia(δ2) = y∗c , which shows that (A.5) and (A.6) are

equal at δ2.

Next, we prove part (ii) . Given a popular set An, the first-order profit difference

in system D is

ΠD,TFAn+1

(k)− ΠD,TFAn

(k) =∑i∈An

[πD,TF (pqAn+1

i , k)− πD,TF (pqAni , k)] + πD,TF (pqAn+1

n+1 , k)

For system C, Lemma B.6 suggests that we need to consider the following three

cases: (1) If An ⊂ Ne, then (An)e = An and (An+1)e = An+1; (2) If An = Ne, then

(An)e = An and (An+1)e = An; (3) If An ⊃ Ne, then (An)e = Ne and (An+1)e = Ne.

For simplicity, we only analyze the first case here (the other two cases can be analyzed

similarly to obtain the same results). The first-order profit difference in model C is

ΠC,TFAn+1

− ΠC,TFAn

=∑i∈An

[πD,TF (pqAn+1

i , kd + αAn+1

i )− πD,TF (pqAni , kd + αAni )]

+ πD,TF (pqAn+1

n+1 , kd + αAn+1

n+1 ) (A.7)

where αAni and αAn+1

i are the KKT multipliers from the optimization in (A.3). (Re-

call also that∑

i∈An αAni =

∑i∈An+1

αAn+1

i = kc.) Note that kc > αAni > αAn+1

i > 0

for i ∈ An. To see this, suppose that αAni ≤ αAn+1

i for some i. Because the stock-

ing quantities for all products in (An)e = An are equal, we have thatkd+αAni

θi=

kd+αAnjθj

for all i, j ∈ An. Similarly,kd+α

An+1i

θi=

kd+αAn+1j

θjfor all i, j ∈ An+1.

63

Page 72: Assortment Planning with Inventory Considerations in

Then, αAni ≤ αAn+1

i implies that αAnj ≤ αAn+1

j for any other j ∈ An. Thus,

kc =∑

i∈An αAni ≤

∑i∈An α

An+1

i <∑

i∈An+1αAn+1

i = kc, a contradiction. It also

follows that αAn+1

j > αAn+1

i if θi ≤ θj.

From Lemma B.1, because πD,TF (x, k) is decreasing in k, we have that (A.7) > (A.8),

where

∑i∈An

[πD,TF (pqAn+1

i , kd+αAn+1

i )−πD,TF (pqAni , kd+αAn+1

i )]+πD,TF (pqAn+1

n+1 , kd+αAn+1

n+1 ).

(A.8)

Because πD,TF (x, k) is submodular in (x, k), we also have that (A.8) > (A.9), where

∑i∈An

[πD,TF (pqAn+1

i , kd+αAn+1

n+1 )−πD,TF (pqAni , kd+αAn+1

n+1 )]+πD,TF (pqAn+1

n+1 , kd+αAn+1

n+1 ).

(A.9)

Note that the expression in (A.9) is equal to ΠD,TFAn+1

(kd +αAn+1

n+1 )−ΠD,TFAn

(kd +αAn+1

n+1 )

and that kd + αAn+1

n+1 < k. From Proposition B.2, the optimal assortment in system

D with unit cost k1 is smaller than that with unit cost k2. In other words, if

ΠD,TFAn+1

(k1)−ΠD,TFAn

(k1) ≥ 0, then ΠD,TFAn+1

(k2)−ΠD,TFAn

(k2) ≥ 0. Therefore, if ΠD,TFAn+1

(k)−

ΠD,TFAn

(k) ≥ 0, then ΠC,TFAn+1

− ΠC,TFAn

> ΠD,TFAn+1

(kd + αAn+1

n+1 ) − ΠD,TFAn

(kd + αAn+1

n+1 ) > 0,

which implies that the optimal assortment in system C is no smaller than that in

system D. �

Proof of Proposition 1.5.1: Because ΠDS (δ) is quasi-convex and θi + θj >

max {θi, θj}, if it is optimal to include variant i or j, then it is optimal to include both

of them in the optimal assortment. Thus, the optimal assortment includes either all

the variants that share a common component or none of them. If these variants are

not in the optimal assortment, then the results for the dedicated system show that

the optimal assortment is a popular set. Otherwise, note that the profit function

for any assortment S not containing variants i and j, but containing a new variant

64

Page 73: Assortment Planning with Inventory Considerations in

m with utility δ, is quasi-convex in δ. We then have that the optimal assortment,

excluding variants i and j, is a popular subset of the remaining variants. �

Proof of Proposition 1.5.2: We prove the result for the IP model. A similar

argument applies to the TF model.

1. If i, j ≤ n, then i, j ∈ Ai+j. To see this, note that from part 1 we have that

either both variants or neither of them are in Ai+j. Suppose that i, j 6∈ Ai+j. Then,

ΠDAn

< ΠPCAn

< ΠPCAi+j = ΠD

Ai+j , where the superscript ‘PC’ denote the profit in the

system with partial commonality. The first inequality follows because commonality

increases the profit for a given assortment, the second inequality follows because i

and j are not in the optimal assortment in the system with partial commonality,

and the equality follows because, in that case, Ai+j only contains products that

use dedicated components. Because An is the optimal assortment in the dedicated

system, we reach a contradiction. Therefore, i, j ∈ Ai+j. Also, the profit function

of any assortment that excludes variants i and j, but contains a new variant m with

utility δ, is quasi-convex in δ. Therefore, Ai+j \ {i, j} is a popular subset of the

remaining variants. Note that1

∂ΠD,IPAn−1

(δ)

∂δ=

1∑l∈An−1

θl + δ + θ0

− ∑l∈An−1

qAn−1

l

∂πD,IP (λqAn−1

l , k)

∂qAn−1

l

+

∑l∈An−1

qAn−1

l

∂πD,IP (λqAn−1m , k)

∂qAn−1m

(A.10)

A similar derivative is obtained when the vector of utilities is given by Θi+j:

∂ΠD,IPAn−1

(δ/Θi+j)

∂δ=

1∑l∈An−1

θl + δ + θ0

− ∑l∈An−1\{i,j}

qAn−1

l

∂πD,IP (λqAn−1

l , k)

∂qAn−1

l

1 Recall that An−1(δ) = An−1 ∪ {m}. We omit the dependence of An−1 on δ to simplify thenotation in the expressions below.

65

Page 74: Assortment Planning with Inventory Considerations in

−(qAn−1

i + qAn−1

j )∂πD,IP (λq

An−1

i + λqAn−1

j , k)

∂(qAn−1

i + qAn−1

j )+

∑l∈An−1

qAn−1

l

∂πD,IP (λqAn−1m , k)

∂qAn−1m

.(A.11)

From equation (B.1) of Lemma B.1, we have that

x∂πD,IP (x, k)

∂x= (p− k)x− pφ(z(p, k))σβxβ.

Moreover, Lemma B.2 implies that (qAn−1

i +qAn−1

j )β < (qAn−1

i )β+(qAn−1

j )β. Therefore,

(qAn−1

i + qAn−1

j )∂πD,IP (λq

An−1

i + λqAn−1

j , k)

∂(qAn−1

i + qAn−1

j )>

qAn−1

i

∂πD,IP (λqAn−1

i , k)

∂qAn−1

i

+ qAn−1

j

∂πD,IP (λqAn−1

j , k)

∂qAn−1

j

,

which implies that (A.10) > (A.11). Define t1(θ1, · · · , θi, · · · , θj, · · · , θn−1)

= δn−1(Θi+j). (Recall that, under the vector of utilities Θi+j, this value of δ is

defined so that ΠDn−1(0) = ΠD

n−1(δn−1(Θi+j)).) Because (A.10) > (A.11), we have

that δn−1(Θi+j) > δn−1(Θ) (≤ θn). Therefore, if θn ≥ δn−1(Θi+j) > δn−1(Θ), we

have that Ai+j = An. Otherwise, Ai+j is a strict subset of An. Similar expressions

as in (A.10) and (A.11) as well as a similar argument apply to the set An, which

proves the result for the case j = n.

2. If i < n and j > n, then, with a similar argument as in case 1, we have that

i, j ∈ Ai+j. Define t2(θ1, · · · , θi, · · · , θn−1, θj) = δn−1(Θi+j). From Lemma B.3, we

have that t2(θ1, · · · , θi, · · · , θn−1, θj) ≥ δn−1(Θ). Thus, from part (1) of Proposition

B.5, we conclude that if θn < t2(θ1, · · · , θi, · · · , θn−1, θj), then Ai+j \ {j} is a strict

subset of An, and otherwise, Ai+j = An ∪{j}. A similar argument proves part 3. �

66

Page 75: Assortment Planning with Inventory Considerations in

Numerical Study

The parameters of the numerical study are as follows. N = 10, p = 10, and

β = 0.5. We consider two demand rates, λ = 200 and λ = 400. The costs of dedi-

cated and common components are given by one of the following nine combinations:

(kd, kc) = {(1, 7) , (4, 4) , (7, 1) , (1, 5) , (3, 3) , (5, 1) , (1, 3) , (2, 2) , (3, 1)}. The utilities

are given by one of the following five vectors:

Θ1 = (20, 18, 16, 14, 12, 10, 8, 6, 4, 2), Θ2 = (20, 18, 16, 14, 12, 5, 4, 3, 2, 1),

Θ3 = (20, 19, 12, 11, 10, 9, 8, 3, 2, 1), Θ4 = (20, 19, 18, 10, 9, 8, 4, 3, 2, 1),

Θ5 = (20, 19, 18, 17, 16, 15, 14, 13, 12, 11). For each of these, the no-purchase utility

takes one of the following four values: θ0 = {10, 20, 40, 60} . Tables A.1 and A.2

summarize the results2.

Table A.1: Optimal profit and variety levels in systems with commonality relativeto the corresponding systems with dedicated components under the IP model.

Average change in Average change in Cases with a strictoptimal profit optimal variety level increase in variety

Θ λ = 200 λ = 400 λ = 200 λ = 400 λ = 200 λ = 400Θ1 4.03% 3.32% 1.31 0.89 69.44% 55.56%Θ2 3.41% 2.91% 1.19 1.22 61.11% 61.11%Θ3 3.91% 2.94% 0.94 1.36 47.22% 69.44%Θ4 3.49% 2.90% 1.25 1.33 69.44% 69.44%Θ5 5.19% 3.42% 1.00 0.67 47.22% 33.33%

Average 4.01% 3.10% 1.14 1.09 58.89% 57.78%

2 The “Average change in optimal variety level” denotes the average increase in the optimalnumber of variants offered in the system with commonality relative to the system with dedicatedcomponents.

67

Page 76: Assortment Planning with Inventory Considerations in

Table A.2: The effect of component commonality on optimal profit and variety levelsunder IP and TF demand models.

Independent Population Model Trend-Following ModelAve. change in Ave. change in Ave. change in Ave. change in

Θ opti. profit opti. variety level opti. profit opti. variety levelΘ1 3.68% 1.10 4.51% 2.67Θ2 3.16% 1.21 3.42% 3.79Θ3 3.42% 1.15 3.85% 3.08Θ4 3.20% 1.29 3.59% 3.5Θ5 4.31% 0.83 5.38% 1.17

Ave. 3.55% 1.12 4.15% 2.84

68

Page 77: Assortment Planning with Inventory Considerations in

Appendix B

Proofs of Supplementary Results of Chapter 1

This appendix contains 1) definition and properties of some functions used through-

out the proofs, 2) intermediate results that are used in the proofs in Appendix A,

and 3) a few supplementary results that were not included in the paper.

Let Φ(·) and φ(·) denote the cumulative and probability density functions of the

standard normal distribution. Define z(x, k) = Φ−1(1 − kx). Differentiating with

respect to x and k, we get the following.

φ(z(x, k))∂z(x, k)

∂x=

k

x2,

φ(z(x, k))∂z(x, k)

∂k= −1

x.

Using the property φ′ (z) = −φ(z)z, we get

∂φ(z(x, k))

∂x= −φ(z(x, k))z(x, k)

∂z(x, k)

∂x= −z(x, k)

k

x2,

∂φ(z(x, k))

∂k= −φ(z(x, k))z(x, k)

∂z(x, k)

∂k= z(x, k)

1

x.

Following the profit functions for the two demand models in Section 1.3, let k = kd+kc

69

Page 78: Assortment Planning with Inventory Considerations in

and redefine

πD,IP (x, k) = (p− k)x− pφ(z(p, k))σxβ, where x = λqi and

πD,TF (x, k) = (x− k)λ− xφ(z(x, k))σ, where x = pqi.

The profit expressions are different because the choice model determines the propor-

tion of demand allocated to variants under the IP demand model and the probability

of having the full demand and revenue for a particular variant under the TF demand

model. Given assortment S, the optimal profit functions for the IP and TF models

in system D are given by

ΠD,IPS =

∑i∈S

πD,IP (λqSi , kd + kc)and ΠD,TFS =

∑i∈S

πD,TF (pqSi , kd + kc).

As mentioned, we use the notation ΠD,IPS (δ) = ΠD,IP

S∪{m} and ΠD,TFS (δ) = ΠD,TF

S∪{m},

where θm = δ.

Lemma B.1. Both ΠD,IP (x, k) and ΠD,TF (x, k) are increasing and convex in x,

decreasing and convex in k, and submodular in (x, k).

Proof of Lemma B.1: The first-order derivatives of ΠD,IP (x, k) and ΠD,TF (x, k)

are

∂ΠD,IP (x, k)

∂x= (p− k)− pφ(z(p, k))σβxβ−1 > 0, (B.1)

∂ΠD,TF (x, k)

∂x= λ− σφ(z(x, k)) + σz(x, k)

k

x

=1

x(πD,TF (x, k) + k(λ+ σz(x, k)))

= Emin {y∗, D} ≥ 0, (B.2)

∂ΠD,IP (x, k)

∂k= −x− z(p, k)σxβ < 0,

∂ΠD,TF (x, k)

∂k= −y∗ = −λ− σz(x, k) < 0. (B.3)

70

Page 79: Assortment Planning with Inventory Considerations in

Thus, both ΠD,IP (x, k) and ΠD,TF (x, k) are increasing in x and decreasing in k. The

second-order derivatives are given by:

∂2ΠD,IP (x, k)

∂x2= (1− β)pφ(z(p, k))σβxβ−2 > 0,

∂2ΠD,TF (x, k)

∂x2= σ

1

φ(z(x, k))

k2

x3> 0.

∂2ΠD,IP (x, k)

∂k2= −∂z(p, k)

∂kσxβ > 0,

∂2ΠD,TF (x, k)

∂k2= −∂z(x, k)

∂kσ > 0.

∂2ΠD,IP (x, k)

∂x∂k= −1− βz(p, k)σxβ−1 < 0,

∂2ΠD,TF (x, k)

∂x∂k= −σ∂z(x, k)

∂x< 0.

Thus, ΠD,IP (x, k) and ΠD,TF (x, k) are convex in x and k and submodular in (x, k).�

Proposition B.1. ΠDS (δ) is quasi-convex in δ. As a result, the optimal assortment

set under either demand model in the system with dedicated components is a popular

set.

Proof of Proposition B.1: See van Ryzin and Mahajan (1999) for the proof of

the IP demand model. For the TF demand model,

ΠD,TFS (δ) =

∑i∈S

πD,TF (pqS(δ)i , k) + πD,TF (pqS(δ)

m , k).

Differentiating with respect to δ, we obtain

∂ΠD,TFS (δ)

∂δ=

∑i∈S

∂πD,TF (pqS(δ)i , k)

∂qS(δ)i

∂qS(δ)i

∂δ+∂πD,TF (pq

S(δ)m , k)

∂qS(δ)m

∂qS(δ)m

∂δ

=

[−∑i∈S

θi∂πD,TF (pq

S(δ)i , k)

∂qS(δ)i

+ (∑i∈S

θi + θ0)∂πD,TF (pq

S(δ)m , k)

∂qS(δ)m

]/(∑

)2.

71

Page 80: Assortment Planning with Inventory Considerations in

where∑

=∑

i∈S θi + δ + θ0. Define LD,TFS (δ) as the expression inside the square

brackets. From Lemma B.1, the first term of LD,TFS (δ) is itself increasing in qS(δ)i

and qS(δ)i is itself decreasing in δ for i ∈ S. The second term is increasing in q

S(δ)m

and qS(δ)m is increasing in δ. Thus, LD,TFS (δ) is increasing in δ, which implies that

ΠD,TFS (δ) is quasi-convex with respect to δ. Hence, for any starting assortment that

is not a popular set, it is possible to improve the profit by dropping the least popular

item in the assortment and possibly including the most popular product outside the

assortment. With an argument similar to the proof of the IP case in van Ryzin and

Mahajan(1999), this suggests that the optimal assortment is a popular set. �

Proposition B.2. The optimal assortment set becomes smaller and the optimal

profit decreases as k increases. Furthermore, ΠS(k) is convex in k.

Proof of Proposition B.2: Given an assortment S, ΠD,IPS =

∑i∈S π

D,IP (λqSi , k)

and ΠD,TFS =

∑i∈S π

D,TF (pqSi , k) are decreasing and convex in k by Lemma 1. Next,

we prove that the optimal assortment set becomes smaller as k increases for both

demand models. Assume that k1 > k2 and that the optimal assortments are An1 and

An2 for the settings with k1 and k2, respectively. By definition of n1, δn1(k1) > θn1+1

and θn1 ≥ δn1−1(k1). Furthermore, from Condition 2, θn1 ≥ δn1−1(k1) ≥ δn1−1(k2),

which implies that An1 ⊂ An2 . If δn1(k2) > θn1+1, then n1 = n2. Otherwise, n2 > n1.

Therefore the optimal assortment becomes smaller as k increases. �

The next result derives upper and lower bounds for the optimal inventory levels

in system C under IP demand model. These bounds are useful for the optimiza-

tion of the profit function under the independent population model in the numerical

experiments. In particular, using the bounds on the inventory values, we use an iter-

ative optimization algorithm (optimizing along each variable, one at a time) which is

72

Page 81: Assortment Planning with Inventory Considerations in

guaranteed to converge to the optimal inventory levels because of the joint concavity

of the profit function. ΠP,IPS denotes the total profit in the pooled system.

Proposition B.3. Consider system C under IP demand model. The following

comparisons hold for any given assortment S.

(i) y∗c ≥ y∗i > λqSi + zIPσ(λqSi )β, which is the stocking level for components

associated with product variant i in the dedicated system.

(ii) y∗c <∑

i∈S y∗i .

(iii) y∗c ≤ λ∑

i∈S qSi + Φ−1(1 − kc

p)σλβ(

∑i∈S q

Si )β, which is the stocking level of

the common component in the pooled system with unit cost kc.

(iv) ΠD,IPS ≤ ΠC,IP

S ≤ ΠP,IPS .

Proof of Proposition B.3: The optimal stocking levels satisfy the following first-

order conditions:

Pr{y∗i ≤ Di, y∗c >

∑i∈S

min{y∗i , Di}} = kd/p (B.4)

Pr{∑i∈S

min{y∗i , Di} > y∗c} = kc/p (B.5)

For part (i), because Pr{y∗i ≤ Di or y∗c >∑

i∈S min{y∗i , Di}} ≤ 1, then

Pr{y∗i ≤ Di}+ Pr{y∗c >∑i∈S

min{y∗i , Di}} − Pr{y∗i ≤ Di, y∗c >

∑i∈S

min{y∗i , Di}} ≤ 1.

With some algebra, we get

Pr{y∗i ≤ Di} ≤ Pr{y∗i ≤ Di, y∗c >

∑i∈S

min{y∗i , Di}}+ Pr{y∗c <∑i∈S

min{y∗i , Di}}.

Using (B.4) and (B.5), we then have that

Pr{y∗i ≤ Di} ≤k

p,

73

Page 82: Assortment Planning with Inventory Considerations in

which implies that y∗i > λqSi +zIPσ(λqSi )β. The proof of part (ii) is by contradiction.

Suppose that y∗c ≥∑

i∈S y∗i . Then, y∗c ≥

∑i∈S y

∗i ≥

∑i∈S min{y∗i , Di}, which implies

that Pr{∑

i∈S min{y∗i , Di} > y∗c} = 0. This contradicts equation (B.5). For part

(iii), we have that

Pr{y∗c >∑i∈S

Di} ≤ Pr{y∗c >∑i∈S

min{y∗i , Di}} = 1− kc/p,

where the equality follows by (B.5). Hence, y∗c should be larger than the stocking

quantity for the sum of the demand for all the components, which is the stocking

level in the pooled system. That is, y∗c < λ∑

i∈S qSi + Φ−1(1 − kc

p)σλβ(

∑i∈S q

Si )β.

Finally, for part (iv), let ΠCS denote the profit in system C with the optimal stocking

quantities of system D (i.e., y∗i = λqSi + zIPσ(λqSi )β, y∗c =∑

i∈S λqSi + zIPσ(λqSi )β)

and ΠCS denote the optimal profit in system C. Clearly, ΠD,IP

S ≤ ΠC,IPS ≤ ΠC,IP

S .

With a similar argument, we obtain ΠC,IPS ≤ ΠP,IP

S . �

Lemma B.2.

(∑i∈S

θi)β <

∑i∈S

θβi , for any S and β ∈ [0, 1] .

Proof of Lemma B.2: We show the result for the case with |S| = 2. Define

f(x) = aβ + xβ − (a + x)β where a, x ≥ 0. Since ∂f(x)∂x

= β[ 1x1−β− 1

(a+x)1−β] > 0, we

have that f(x) > f(0) = 0 for any x > 0. Therefore aβ + xβ > (a + x)β. This can

be extended to the cases with |S| > 2. �

Lemma B.3. For any n and i ≤ n, if δn−1 ≤ θn, then δn is increasing in θi.

Proof of Lemma B.3: By the definition of δn, ΠDn (δn) = ΠD

n (0). Note that δn is

a function of (θ1, · · · , θn) and independent of (θn+1, · · · , θN). Also, both ΠDn (0) and

74

Page 83: Assortment Planning with Inventory Considerations in

ΠDn (δn) are functions of (θ1, · · · , θn) through the choice probabilities (qAni )i∈An and

(qAn(δn)i )i∈An(δn), respectively. Differentiate both sides of the equality with respect

to θi, for i ∈ An, to obtain

∂ΠDn (δn)

∂δ

∂δn∂θi

+∑

j∈An(δn)

∂ΠDn (δn)

∂qAn(δn)j

∂qAn(δn)j

∂θi=∑j∈An

∂ΠDn

∂qAnj

∂qAnj∂θi

(B.6)

(Recall that An(δn) denotes the set An plus a variant m with utility δn.) Because

ΠDn (δ) is quasi-convex, its first-order derivative at δn is positive from the definition

of δn, i.e., ∂ΠDn (δ)/∂δ

∣∣δ=δn

> 0. Hence, ∂δn/∂θi > 0 if we show that (B.7) > (B.8)

below:

∑j∈An

∂ΠDn

∂qAnj

∂qAnj∂θi

=1∑

j∈An θj + θ0

[∂πD(aqAni , k)

∂qAni−∑j∈An

qAnj∂πD(aqAnj , k)

∂qAnj

](B.7)

∑j∈An(δn)

∂ΠDn (δn)

∂qAn(δn)j

∂qAn(δn)j

∂θi=

1∑j∈An(δn) θj + θ0

[∂πD(aq

An(δn)i , k)

∂(qAn(δn)i

)

−∑

j∈An(δn)

qAnj (δn)∂πD(aq

An(δn)j , k)

∂(qAn(δn)j

) ], (B.8)

where a = λ for the IP model and a = p for the TF model. We prove that the

expression in (B.7) is positive by contradiction. If it was negative, then that would

imply that ΠDAn\{i} > ΠD

An. However, Condition 1 and δn−1 < θn (assumed in the

statement of the Lemma) imply that An is the optimal assortment when the set

of possible products is restricted to the n most popular variants. This leads to a

contradiction. In addition, we have that∑

j∈An(δn) θj + θ0 >∑

j∈An θj + θ0. Hence,

we can obtain the desired result by comparing the terms in the brackets in equations

(B.7) and (B.8). Convexity of πD(x, k) in x together with ∂qAni (δ) /∂δ < 0 imply

that the first term inside the bracket in (B.7) is larger than the first term inside the

75

Page 84: Assortment Planning with Inventory Considerations in

bracket in (B.8) . Next, we compare the second terms in each equation for each

demand model separately.

For the IP model, from (B.1), we have

∑j∈An

qAnjπD,IP (λqAnj , k)

∂qAnj=

∑j∈An

qAnj λ((p− k)− pφ(z(p, k))σβ(λqAnj )β−1

)= (p− k)λ

∑j∈An

qAnj − βpσφ(z)∑j∈An

(λqAnj )β, (B.9)

∑j∈An(δn)

qAn(δn)j

πD,IP (λqAn(δn)j , k)

∂(qAn(δn)j

) = (p−k)λ∑

j∈An(δn)

qAn(δn)j −βpσφ(z)

∑j∈An(δn)

(λqAn(δn)j )β.

(B.10)

On the other hand, ΠDAn

(δn) = ΠDAn

implies that

(p− k)λ∑j∈An

qAnj − pσφ(z)∑j∈An

(λqAnj )β = (p− k)λ∑

j∈An(δn)

qAnj (δn)

−pσφ(z)∑

j∈An(δn)

(λqAnj (δn))β

Hence,

(B.9)− (B.10) = (1− β) (p− k)λ[∑j∈An

qAnj −∑

j∈An(δn)

qAn(δn)j ] < 0.

Thus, (B.7) > (B.8) for the IP model.

For the TF model, recall that ΠD,TFn =

∑j∈An q

Anj E[min{yAnj , D}] − kyAnj , where

yAnj = y∗j given qAnj . Note that δn and ΠDn are functions of k. By definition of δn,

we have ΠD,TFn (δn) = ΠD,TF

n . Differentiating both sides of the equation with respect

to k, we obtain

∂ΠD,TFAn

(δ)

∂δ

∣∣∣∣∣x=δn

∂δn∂k

+∑

j∈An(δn)

∂πD,TF (pqAn(δn)j , k)

∂k=∑j∈An

∂πD,TF (pqAn(δn)j , k)

∂k.

76

Page 85: Assortment Planning with Inventory Considerations in

The first term on the left is positive because of the definition of δn and Condition 2.

Thus, we have

∑j∈An

∂πD,TF (pqAn(δn)j , k)

∂k>

∑j∈An(δn)

∂πD,TF (pqAn(δn)j , k)

∂k.

By (B.3) , this is equivalent to ∑j∈An(δn)

yAn(δn)j >

∑j∈An

yAnj

At the same time, ΠD,TFn (δn) = ΠD,TF

n is equivalent to∑j∈An(δn)

pqAn(δn)j E[min{yAn(δn)

j , D}]−∑

j∈An(δn)

yAn(δn)j =

∑j∈An

pqAnj E[min{yAni , D}]−∑j∈An

yAnj .

As a result, we have∑j∈An(δn)

qAn(δn)j E[min{yAn(δn)

j , D}] >∑j∈An

qAnj E[min{yAnj , D}].

This implies the desired inequality:

∑j∈An(δn)

qAn(δn)j

∂πD,TF (pqAn(δn)j , k)

∂qAn(δn)j

>∑j∈An

qAnj∂πD,TF (pqAnj , k)

∂qAnj

because, by (B.2), E[min{yAn , D}] = ∂πD,TF (x,k)∂x

|x=pqAnj. This completes the proof.

Lemma B.4. Se(δ) is a popular subset of S that decreases (in terms of the “ ⊂ ”

order) to the set {m} as δ increases. At the same time, Sl(δ) = S ∪ {m} − Se(δ)

increases to the set S as δ increases. The common-component fractile t∗c(S ∪ {m})

is continuous and first decreasing and then increasing in δ.

77

Page 86: Assortment Planning with Inventory Considerations in

Proof of Lemma B.2: This lemma explores the structure of the sets Se and Sl

when a new variant m, with utility δ, is added to the assortment S. We let Se(δ)

and Sl(δ) denote the sets (S ∪ {m})e and (S ∪ {m})l, respectively. We study how

Sl(δ), Se(δ), and t∗c(S ∪ {m}) change as δ increases. Let a be the cardinality of Se.

We define the following threshold values of δ:

δ1def=

kd(∑

i∈S θi + θ0)

p− kd, δ2

def=

kd∑

i∈Se θi

(kc + akd), δ3

def=

kc + (a+ 1)kdkd

θia −∑i∈Se

θi,

and δ4def=

kc + akdkd

θia−1 −∑i∈Se

θi − θia .

Case 1: The fractile 1− kd

pqS(δ)m

< 0, or equivalently δ < δ1. In this case, the new variant

m is in the assortment S, but with a stocking quantity equal to zero. Therefore

Sl(δ) = Sl ∪ {m} and Se(δ) = Se. Also, t∗c(S ∪ {m}) = 1 − (kc+akd)(∑i∈S θi+δ+θ0)

p∑al=1 θil

is

continuous and decreasing in δ.

Case 2: When 0 ≤ 1 − kd

pqS(δ)m

< t∗c(S ∪ {m}), or equivalently δ1 ≤ δ < δ2, the

new variant m has a positive optimal stocking level which is smaller than that of

the common component. Therefore, Sl(δ) = Sl ∪ {m} and Se(δ) = Se. Again,

t∗c(S ∪ {m}) = 1− (kc+akd)(∑i∈S θi+δ+θ0)

p∑al=1 θil

, which is continuous and decreasing in δ.

Case 3: When 1− kd

pqS(δ)m

> t∗c(S ∪ {m}) and 1− kd

pqS(δ)ia

≥ t∗c(S ∪ {m}), or equivalently

δ2 ≤ δ < δ3, the new variant m has the same optimal stocking level as that of the

common component. Therefore, Se(δ) = Se ∪{m} (which continues to be a popular

subset of S ∪ {m}) and Sl(δ) = Sl. Also, t∗c(S ∪ {m}) = 1− (kc+(a+1)kd)(∑i∈S θi+δ+θ0)

p∑al=1 θil+pδ

is continuous and increasing in δ.

Case 4: When 1 − kd

pqS(δ)m

> t∗c(S ∪ {m}) > 1 − kd

pqS(δ)ia

and 1 − kd

pqS(δ)ia−1

≥ t∗c(S ∪ {m}),

or equivalently δ3 ≤ δ < δ4, the new variant m is still in Se(δ) and variant ia,

78

Page 87: Assortment Planning with Inventory Considerations in

which has the lowest utility in Se, moves to Sl(δ). (Note that δ ≥ δ3 implies that

δ ≥ θia .) Therefore, Se(δ) = Se∪{m}\{ia}, which continues to be a popular subset

of S∪{m} and Sl(δ) = Sl∪{ia}. Also, t∗c(S∪{m}) = 1− (kc+akd)(∑i∈S θi+δ+θ0)

p∑al=1 θil+pδ−pθia

, which

is continuous and increasing in δ.

Following a similar reasoning, as δ increases beyond δ4, the set Se(δ) (which includes

m) continues to be a popular subset of S ∪ {m}. For large enough δ, the set Se(δ)

reduces to {m}. On the other hand, Sl(δ) = S∪{m}\Se(δ) increases in δ, eventually

becoming equal to S. The threshold t∗c(S ∪ {m}) is continuous and increasing in δ.

Lemma B.5. For any i ∈ S, ΠC,TFS is increasing and convex in qSi .

Proof of Lemma B.5: Based on equation (B.2) of Lemma B.1, the result is

established if we prove that y∗i is continuous and increasing in qSi regardless of how

the sets Se and Sl change with qSi . (Note that, based on the expression of ΠC,TFS ,

the form of the derivative in (B.2) also applies in system C.) Assume that i ∈ Sl.

Then, y∗i is continuous and increasing qSi beyond the point at which variant i enters

the set Se because the common-component fractile is continuous and increasing in

qSi as implied by the proof of Lemma B.4. Before that point, the stocking level is

y∗i = λ+σz(pqSi , kd). This is continuous and increasing in qSi . Moreover, it is easy to

verify that y∗i is continuous at the point in which variant i enters the set Se. Thus,

y∗i is increasing and convex in qSi . For i ∈ Se, the result can be proved similarly. �

The following lemma presents a technical property regarding the set of variants

Se that results from the stocking optimization problem when the assortment is given

by set S. The result is used to prove Theorem 1.4.2. For a popular set An, (An)e

is itself a popular set (following Proposition B.4). Note also that, when S = N , the

set of all products, we denote Se as Ne. Finally, let |S| denote the cardinality of set

79

Page 88: Assortment Planning with Inventory Considerations in

S.

Lemma B.6. Consider a popular set An. If n ≤ |Ne|, then (An)e = An. Otherwise,

(An)e = Ne.

Proof of Lemma B.6: Let r = |Ne| so that Ne = Ar. By definition of Ne, r

satisfies the following:

1− kd∑N

i=0 θipθj

|j∈Ar ≥ t∗c(N ) = 1− (kc + rkd)∑N

i=0 θip∑r

i=1 θi> 1− kd

∑Ni=0 θi

pθr+1

. (B.11)

We complete the proof by contradiction. If n < r, then (An)e ⊂ An, but suppose

that (An)e 6= An. Let z = |(An)e| < n < r, which satisfies the following:

1− kd∑n

i=0 θipθj

|j∈Az ≥ t∗c(An) = 1− (kc + zkd)∑n

i=0 θip∑z

i=1 θi> 1− kd

∑ni=0 θi

pθz+1

. (B.12)

By combining (B.11) and (B.12), we have

kc + rkd∑ri=1 θi

>kdθr> · · · > kd

θz+1

>kc + zkd∑z

i=1 θi.

Inverting all the terms, reversing the order of the inequalities, and taking the average

of all the terms with θr to θz + 1, we have that

kc + zkd∑zi=1 θi

<(r − z)kd∑r

i=z+1 θi<kc + rkd∑r

i=1 θi(B.13)

Let a = kc + zkd, b =∑z

i=1 θi, c = (r − z)kd and d =∑r

i=z+1 θi. The inequality

(B.13) can then be expressed as

a

b<c

d<a+ c

b+ d

80

Page 89: Assortment Planning with Inventory Considerations in

The first inequality shows that ad < bc. However, the second inequality implies that

bc < ad, a contradiction. Therefore (An)e = An when n < r. Similarly, we can

prove that (An)e = Ne when n ≥ r. �

Proposition B.4. The optimal profit is decreasing in the unit costs kc and kd under

both demand models. Moreover, the optimal profit is more sensitive to changes in kd

than in kc.

Proof of Proposition B.4: Under the IP model, the optimal stocking levels y∗i and

y∗c , with i ∈ S, are derived from equations (B.4) and (B.5) of Proposition B.3. The

optimal profit function is given by

ΠC,IPS ({y∗i }i∈S, y∗c , kc, kd) = pE[min{y∗c ,

∑i∈S

min{y∗i , Di}}]−∑i∈S

kdy∗i − kcy∗c .

Differentiating with respect to kc and kd, respectively, we obtain

∂ΠC,IPS ({y∗i }i∈S, y∗c , kc, kd)

∂kc= −y∗c ≤ 0, (B.14)

∂ΠC,IPS ({y∗i }i∈S, y∗c , kc, kd)

∂kd= −

∑i∈S

y∗i ≤ 0. (B.15)

Hence, the optimal profit function is decreasing in kc and kd. Furthermore, from

Proposition B.3, (B.14) > (B.15). Therefore, the rate of change in the optimal profit

with respect to kc is less significant than with respect to kd. Under the TF model,

the optimal profit function is given by

ΠC,TFS ({y∗i }i∈S, y∗c , kc, kd) =

∑i∈Se

[pqiEmin{y∗c , D}−kdy∗c ]+∑i∈Sl

[pqiEmin{y∗i , D}−kdy∗i ]

−kcy∗c .

81

Page 90: Assortment Planning with Inventory Considerations in

Then, we have

∂ΠC,TFS ({y∗i }i∈S, y∗c , kc, kd)

∂kc= −y∗c , (B.16)

∂ΠC,TFS ({y∗i }i∈S, y∗c , kc, kd)

∂kd= −ay∗c −

∑i∈Sl

y∗i (B.17)

where a = |Se|. Therefore the optimal profit function is decreasing in kc and kd.

Because (B.16) > (B.17), we again have that the rate of change in the optimal profit

with respect to kc is less significant than with respect to kd. �

We now proceed with comparative statics results regarding the structure of the

optimal assortment in the dedicated system. These results will be instrumental

in characterizing the optimal assortment structure under general bill-of-materials

configurations. Suppose that the utility for variant j increases from θj to θ′j, with

θ′j > θj. Let Θ = (θ1, · · · , θN) be the original vector of utilities and Θ′ be the

vector Θ with θj replaced by θ′j and reorganized in decreasing order. Suppose that

An is the optimal popular set under the original utilities and let A′ be the optimal

popular set under the utilities in Θ′. The following result characterizes the set A′

both for the independent population and trend-following population models. The

result states that if the utility of a variant in the optimal assortment increases, then

the resulting optimal assortment may contain fewer variants than the original. On

the other hand, if the utility of a variant outside the optimal assortment increases,

then it will be included in the optimal assortment as long as the increase is large

enough.

Proposition B.5. Consider the dedicated system.

(i) If j < n, then j ∈ A′. In addition, if δ′n−1 > θn then A′ is a strict subset of

An. Otherwise, A′ = An. (The quantity δ′n−1 corresponds to the vector of

utilities Θ′.)

82

Page 91: Assortment Planning with Inventory Considerations in

(ii) If j = n, then A′ = An when θ′n ≤ θn−1 or when θ′n > θn−1 and δ′n−1 ≤ θn−1.

When θ′n > θn−1 and δ′n−1 > θn−1, A′ is a strict subset of An.

(iii) If j ≥ n + 1, then An ∪ {j} ⊂ A′ when θn > θ′j ≥ δn > θn+1, and j ∈ A′ and

A′ \ {j} ⊂ An when θ′j ≥ θn.

(iv) If j ≥ n+ 1 and δn > θ′j, then A′ = An.

Proof of Proposition B.5: For (i), because j < n and θ′j > θj, variant j is still in

the optimal assortment, i.e., j ∈ A′. The n− 1 most popular variants are the same

as An−1 which includes variant j. From Lemma B.3, δ′n−1 > δn−1. In addition,

θn ≥ δn−1 because An is the optimal assortment under the original utilities. Thus,

if δ′n−1 which is a function of (θ1, · · · , θ′j, · · · , θn−1) is larger than θn, then A′ is a

strict subset of An. Otherwise, A′ = An. For (ii), if j = n and θ′n ≤ θn−1,

then θ′n > θn ≥ δn−1, Moreover, Lemma B.3 says that δ′n > δn > θn+1. Thus,

A′ = An. If θ′n > θn−1, then the n − 1 most popular variants are (1, · · · , n − 2, n).

From Lemma B.3, δ′n−1 which is a function of (θ1, · · · , θn−2, θ′n) is larger than δn−1.

Then A′ is a strict subset of An if δ′n−1 > θn−1. For (iii), because j ≥ n + 1 and

θn > θ′j ≥ δn > θn+1, variant j as well as An are included in the optimal assortment

A′. When θ′j ≥ θn, similarly as case (i) and (ii), we have j ∈ A′ and A′ \ {j} ⊂ An.

For (iv), because j ≥ n+ 1 and δn > θ′j, it is apparent that A′ = An. �

Proposition B.6. Let Πi+j denote the optimal profit of a system in which variants i

and j (i < j) are produced using a set of common components and all other variants

use dedicated components. Under both the IP and the TF demand models, Π1+2 ≥

Πi+j for any i, j.

Proof of Proposition B.6: Let Θi+j denote the utility vector of a system in which

variants i and j (i < j) are produced using a set of common components and all

83

Page 92: Assortment Planning with Inventory Considerations in

other variants use dedicated components. For variants i and j, Θ1+2 � Θi+j in a

majorization order. Van Ryzin and Mahajan (1999) shows that Π1+2 ≥ Πi+j since

Θ1+2 � Θi+j for the IP demand model. With a similar argument, the result also

holds for the TF demand model. �

Proposition B.7. Consider the trend-following demand model. Let Ai+j be the op-

timal assortment of a system with partial component commonality in which variants

i and j use a common component and a dedicated component each, and all other

variants use only dedicated components, as depicted in Figure 1(d). Let An be the

optimal (popular) assortment in the corresponding dedicated system. The following

results hold.

1. If i, j ≤ n, then Ai+j = A′ ∪ {i, j} where A′ is a popular subset of An\ {i, j}.

There exists a threshold value t7 such that A′ is a strict subset of An\ {i, j} if

and only if θn < t7,.

2. If i < n and j > n, then Ai+j = A′∪{i, j} where A′ is a popular subset of An.

3. If i > n and j > n, and assuming that θj >kd

kd+kcθi, then there exists a threshold

value t8 such that Ai+j = A′ ∪ {i, j} if θi + θj ≥ t8, and Ai+j = An otherwise

(A′ is a popular subset of An).

Proof of Proposition B.7: We will denote by ‘PC’ the system described in the

statement of the proposition.

1. If i, j < n, then i, j ∈ Ai+j (following a similar argument to that in Proposition

1.5.2). Consider any assortment S that contains variants i and j, and let m be a

variant not in S with utility δ. Because πD,TF (x, k) is submodular in (x, k) from

Lemma B.1, we have

84

Page 93: Assortment Planning with Inventory Considerations in

∂πD,TF (pqS(δ)i , k)

∂qS(δ)i

<∂πD,TF (pq

S(δ)i , kd + αi(θi, θj))

∂qS(δ)i

∂πD,TF (pqS(δ)j , k)

∂qS(δ)j

<∂πD,TF (pq

S(δ)j , kd + αj(θi, θj))

∂qS(δ)j

Therefore,

∂ΠD,TFS (δ)

∂δ

=p

(∑

l∈S θl + δ + θ0)2

[−∑l∈S

θl∂πD,TF (pq

S(δ)l , k)

∂qS(δ)l

+

(∑l∈S

θl + θ0

)∂πD,TF (pq

S(δ)m , k)

∂qS(δ)m

]

>−p

(∑

l∈S θl + δ + θ0)2

∑l∈S\{i,j}

θl∂πD,TF (pq

S(δ)l , k)

∂qS(δ)l

+θi∂π

D,TF (pqS(δ)i , kd + αi(θi, θj))

∂qS(δ)i

+θj∂πD,TF (pq

S(δ)j , kd + αj(θi, θj))

∂qS(δ)j

(∑l∈S

θl + θ0

)∂πD,TF (pq

S(δ)m , k)

∂qS(δ)m

]=∂ΠPC,TF

S (δ)

∂δ.

(Note that ΠPC,TFS (δ) is quasi-convex in δ which implies that Ai+j \{i, j} is a popular

subset of the remaining variants.) The inequality above implies that δl(Θ|PC) >

δl(Θ) for any l ≥ j. We then define t7(θ1, · · · , θi, · · · , θj, · · · , θn−1) = δn−1(Θ|PC).

If θn < t7(θ1, · · · , θi, · · · , θj, · · · , θn−1), then Ai+j is a strict subset of An. Otherwise,

Ai+j = An. The case with j = n follows similarly.

2. If i < n, then, following a similar argument as before, we can show that i ∈ Ai+j.

If j > n is not in Ai+j, we then have that Ai+j = An. Suppose now that j ∈ Ai+j.

For a given assortment S and a variant m 6∈ S, we have proved that ΠPC,TFS (δ) is

quasi-convex in δ. Therefore, Ai+j = A′ ∪ {i, j} where A′ is a popular subset of An.

3. Consider now the case with i > n and j > n (and θj >kd

kd+kcθi). We first note

that δn(Θ) > θn+1 ≥ θi, θj. The threshold value δn(Θ) arises in the setting in which

all variants have unit production costs equal to k. Because θj >kd

kd+kcθi, we have

85

Page 94: Assortment Planning with Inventory Considerations in

from Lemma B.7 that αi, αj > 0. Then, we can write the combined profit function

for variants i and j as πTFi,j , being a function of p(qi + qj) and 2kd + kc (essentially,

the two variants jointly play the role of a single variant with utility θi + θj and total

cost 2kd + kc). Let S be any assortment not containing variants i and j, and let

ΠPC,TFS∪{i,j} =

(∑l∈S π

D,TFl

)+ πTFi,j . If m 6∈ S with utility δ, using similar arguments

as before, we have that ΠPC,TFS (δ) is quasi-convex in δ. We then define δ(Θ|PC)

such that ΠPC,TFS (0) = ΠPC,TF

S (δ(Θ|PC)). (Note that now δ(Θ|PC) may depend

on 2kd + kc.) Let m denote a new variant with utility δ and An(δ) = An ∪ {m}.

In system D, this new variant has a cost of k = kd + kc and utility δ, while in the

system with partial commonality, this new variant represents the combined variants

i and j with cost 2kd + kc and utility δ. Because πD,TF (x, k) is increasing in x and

submodular in (x, k) (from Lemma B.1), we have

∂ΠD,TFAn

(δ)

∂δ=

p

(∑

l∈An θl + δ + θ0)2

[−∑l∈An

θl∂πD,TF (pq

An(δ)l , k)

∂qAn(δ)l

+

(∑l∈An

θl + θ0

)∂πD,TF (pq

An(δ)m , k)

∂qAn(δ)m

]

≥∂ΠPC,TF

An(δ)

∂δ=

p

(∑

l∈An θl + δ + θ0)2

[−∑l∈An

θl∂πD,TF (pq

An(δ)l , k)

∂qAn(δ)l

+

(∑l∈An

θl + θ0

)∂πD,TF (pq

An(δ)m , 2kd + kc)

∂qAn(δ)m

].

This implies that δn(Θ|PC) > δn(Θ). Therefore, we define t8(θ1, · · · , θn, αi) =

δn(Θ|PC). If t8(θ1, · · · , θn, αi) ≤ θi + θj, then i, j ∈ Ai+j and Ai+j = A′ ∪ {i, j}

where A′ is a popular subset of An. Otherwise, Ai+j = An. �

86

Page 95: Assortment Planning with Inventory Considerations in

Lemma B.7. Consider the system described in Proposition B.7. Let S be a subset

of N , and assume that variants i > j are both included in S. There exist positive

values αi(θi, θj) and αj(θi, θj) such that αi(θi, θj) + αj(θi, θj) = kc. If θj ≤ kdkd+kc

θi,

then αi(θi, θj) = kc and αj(θi, θj) = 0. Otherwise,kd+αi(θi,θj)

θi=

kd+αj(θi,θj)

θj.

Proof of Lemma B.7: Suppose that θj ≤ kdkd+kc

θi, which implies that 1− kd+kcθi≥

1 − kdθj

. From the results in system C, we have that if the KKT multipliers αi > 0

and αj > 0, then αi + αj = kc and kd+αiθi

=kd+αjθj

. This equality, together with

θj ≤ kdkd+kc

θi can only occur if αi = kc and αj = 0, a contradiction. We then have

that αi = kc and αj = 0 (and both values are independent of all other parameters,

except for θi and θj). If θj >kd

kd+kcθi, then 1 − kd+kc

θi< 1 − kd

θj. In this case, there

exist positive values αi(θi, θj) and αj(θi, θj) from the solution of the KKT conditions,

such that αi(θi, θj) +αj(θi, θj) = kc andkd+αi(θi,θj)

θi=

kd+αj(θi,θj)

θj. Here again, αi and

αj only depend on kc, kd, θi, and θj. �

We also study a system in which there is a component common to all products and

another component common to only products i and j. (All other variants require,

in addition, a dedicated component.) Figure 1.5(e) illustrates an example of such

system. In the next proposition, we compare this system to the related system C

(illustrated in Figure 1.5(a)).

Proposition B.8. Consider a system with partial component commonality in which

there is a component common to all products and another one common to only prod-

ucts i and j, while all other variants also require a dedicated component, as depicted

in Figure 1(e). Consider the trend-following model. Let Ai+j be the optimal assort-

ment in that system. Let An be the optimal assortment in the corresponding system

C in which all variants are produced using a dedicated component and a common

87

Page 96: Assortment Planning with Inventory Considerations in

component. Assume that Ne ⊂ An.1

1. If i, j ≤ n, then Ai+j = A′ ∪ {i, j} where A′ is a popular subset of An\ {i, j}.

There exists a threshold value t5 such that A′ is a strict subset of An\ {i, j} if

θn < t5.

2. If i < n and j > n, then Ai+j = A′∪{i, j} where A′ is a popular subset of An.

3. If i > n and j > n, then there exists a threshold value t6 such that Ai+j =

A′ ∪ {i, j} if θi ≥ t6, and Ai+j = An otherwise. A′ is a popular subset of An.

Proof of Proposition B.8: Based on the assumption that Ne ⊂ An, we consider

the case in which i ∈ Ne and j ∈ N \ Ne (applicable to parts 1 and 2). The

other cases can be proved similarly. From the results of the pooled system, it is

optimal to include both variants i and j in the optimal assortment. In this setting

with partial commonality, we have that the (unordered) utility vector is Θi+j =

(θ1, · · · , θi + θj, · · · , 0, · · · , θn). For an assortment S, we denote the corresponding

set of equal stocking quantities in this setting as SPCe . From Lemma B.4, we have

that N PCe ⊆ Ne.

1. If i, j < n, then i, j ∈ Ai+j. Given any assortment S including variants i and j,

ΠPC,TFS (δ) is quasi-convex in δ by Lemma B.5. This implies that Ai+j \ {i, j} is a

popular subset of the remaining variants. Consider An−1 = {1, 2, · · · , n − 1}, with

i, j ∈ An−1. Then, we have

∂ΠPC,TFS (δ)

∂δ

1 Recall thatNe is the popular subset of variants with equal stocking levels as the stocking quantityof the common component that results from the profit maximization problem with respect to thestocking quantities assuming an assortment that contains all possible variants in N .

88

Page 97: Assortment Planning with Inventory Considerations in

=p

(∑

l∈S θl + δ + θ0)

[−∑l∈S

qS(δ)l

∂ΠPC,TF (δ)

∂qS(δ)l

+ (∑l∈S

qS(δ)l + q

S(δ)0 )

∂ΠPC,TF (δ)

∂qS(δ)m

](B.18)

∂ΠC,TFS (δ)

∂δ=

p

(∑

l∈S θl + δ + θ0)

[−∑l∈S

qS(δ)l

∂ΠC,TF (δ)

∂qS(δ)l

+ (∑l∈S

qS(δ)l + q

S(δ)0 )

∂ΠC,TF (δ)

∂qS(δ)m

]. (B.19)

We want to prove that (B.19) > (B.18) when δ < δn−1(Θ) (< θn), so that

δn−1(Θi+j) > δn−1(Θ). For δ < δn−1(Θ) (< θn), the new variant m is not in Se(δ),

and therefore not in SPCe (δ) either as SPCe (δ) ⊂ Se(δ). Therefore, the second terms

in the square brackets in (B.19) and (B.18) are equal. We then compare the first

terms in the square brackets. We denote the optimal stocking levels in the system

with partial commonality as yPC∗l . The proof of Lemma B.4 implies that yPC∗l > y∗l

for any l ∈ Se ∪ {j} and yPC∗l = y∗l for any l ∈ S \ (Se ∪ {j}). Moreover, from (B.2)

of Lemma B.1, we have that

−∑l∈S

qS(δ)l

∂ΠPC,TF (δ)

∂qS(δ)l

= −∑l∈S

qS(δ)l Emin{yPC∗l , D}

and

−∑l∈S

qS(δ)l

∂ΠC,TF (δ)

∂qS(δ)l

= −∑l∈S

qS(δ)l Emin{y∗l , D}.

Therefore, (B.19)> (B.18), which implies that δn−1(Θi+j) > δn−1(Θ). Then, defining

t5(θ1, · · · , θi, · · · , θj, · · · , θn−1) = δn−1(Θi+j),

we have that if t5(θ1, · · · , θi, · · · , θj, · · · , θn−1) > θn, then A′ is a strict subset of An.

Otherwise, A′ = An.

2. If i < n, then i ∈ Ai+j. Because it is optimal to include both variants i and j

89

Page 98: Assortment Planning with Inventory Considerations in

together, we also have that j ∈ Ai+j. Given assortment S including both i and j,

we have proved that ΠPC,TFS (δ) is quasi-convex in δ. Therefore, Ai+j = A′ ∪ {i, j},

where A′ is a popular subset of An.

3. Because i > n and j > n, we have that δn(Θ) > θn+1 ≥ θi, θj. Let t6(θ1, · · · , θn) =

δn(Θ). If δn(Θ) > θi + θj, then Ai+j = An. Otherwise, Ai+j = A′ ∪ {i, j} where A′

is a popular subset of An. �

90

Page 99: Assortment Planning with Inventory Considerations in

Appendix C

Proofs of Chapter 2

Proposition 2.3.1

Proof: The optimality equation (2.3.2) can be rewritten as

Vt(y|m) = maxS⊂S(y)

{∑i∈S

λqmi(S)pit(y)

}+ Vt+1(y) (B.1)

If Vt+1(·), the value function in period t + 1, is known, then (pit(y))i∈N can be

derived. Without loss of generality, assume pi1t ≥ · · · ≥ piNt and S(y) = N . Note

that these marginal revenues are the same for all customer segments. Suppose the

arriving customer belongs to segment m. We prove the result by contradiction.

Suppose the optimal assortment is S∗ such that there exists k and l where ik ∈ S∗,

il ∈ N \S∗, and pilt (y) > pikt (y). Then pilt (y) > pikt (y) ≥∑

j∈S∗ qmj(S)pijt (y) because

ik ∈ S∗. For simplicity, let a =∑

j∈S∗ θmjpijt (y) and b =

∑j∈S∗ θmj + θm0. Because

(a+ θx)/(b+ θ) is larger than a/b for any positive θ if and only if x ≥ a/b, then we

have ∑j∈S∗\{ik}

qmj(S∗ ∪ {il})p

ijt (y) + qmil(S

∗ ∪ {il})pilt (y) >∑j∈S∗

qmj(S)pijt (y).

91

Page 100: Assortment Planning with Inventory Considerations in

Hence, assortment S∗ ∪ {il} is strictly better than S∗. This is a contradiction. �

Parts (i) and (ii) of Proposition 2.4.1 and Theorem 2.4.1 follow from these tech-

nical properties:

Property 1 Vt(y − ej)− Vt(y − ei − ej)− Vt(y) + Vt(y − ei) ≥ 0, i, j ∈ N .

Property 2 Vt(y − ei) ≤ Vt(y − ej) where yi = yj ≥ 1 and θi ≥ θj.

Property 3 For any S ⊂ N \ {i}, p−∆it(y) ≥

∑j∈S αj(p−∆j

t (y))∑j∈S αj+θ0

. In other words,

θ0p+ (∑j∈S

αj + θ0)Vt(y − ei)−∑j∈S

αjVt(y − ej)− θ0Vt(y) ≥ 0

Properties 1 and 2 are (i) and (ii) of Proposition 2.4.1, respectively. From the

proof of Proposition 2.3.1, Property 3 claims that the optimal assortment in period

t−1 is to offer all available products, which is Theorem 2.4.1. We prove these results

together by induction. First, because the optimal assortment in period T is to offer

all available products, it is easy to verify that the results hold for period T . Second,

we assume these three results hold in period t + 1. Note that Property 3 in period

t+1 implies that the optimal assortment in period t is to offer all available products.

Finally, we prove Properties 1 and 2 in period t in the proof of Proposition 2.4.1

and Property 3 in period t in the proof of Theorem 2.4.1. Then the entire proof is

completed by induction. Without loss of generality, assume S(y) = N .

Proposition 2.4.1

Proof:

(i) and (ii): We begin with the proof of Property 1. There are in total six cases.

When i = j, there are two cases: (1) yi = 2; (2) yi > 2. When i 6= j, there are

four cases: (3) yi ≥ 2 and yj ≥ 2; (4) yi = 1 and yj ≥ 2; (5) yi ≥ 2 and yj = 1; (6)

92

Page 101: Assortment Planning with Inventory Considerations in

yi = 1 and yj = 1. Here, we prove case 6 only. The other cases are either easier or

similar. Define∑N = (

∑k∈N θk + θ0),

∑N\{i} = (

∑k∈N\{i} θk + θ0), ei,j = ei + ej,

and ei,j,k = ei + ej + ek.

Vt(y − ej)− Vt(y − ei,j)− Vt(y) + Vt(y − ei) =

θ0θiθj(∑N\{i}+

∑N\{j})∑

N∑N\{i}

∑N\{j}

∑N\{i,j}

p

+

∑k∈N\{j} θkVt+1(y − ej,k)∑

N\{j}−∑

k∈N\{i,j} θkVt+1(y − ei,j,k)∑N\{i,j}

−∑

k∈N θkVt+1(y − ek)∑N

+

∑k∈N\{i} θkVt+1(y − ei,k)∑

N\{i}+θ0Vt+1(y − ej)∑

N\{j}

−θ0Vt+1(y − ei,j)∑N\{i,j}

− θ0Vt+1(y)∑N

+θ0Vt+1(y − ei)∑

N\{i}

=

∑k∈N\{i,j} θk (Vt+1(y − ej,k)− Vt+1(y − ei,j,k)− Vt+1(y − ek) + Vt+1(y − ei,k))∑

N

+θ0 (Vt+1(y − ej)− Vt+1(y − ei,j)− Vt+1(y) + Vt+1(y − ei))∑

N

+θj∑

k∈N\{i,j} θkVt+1(y − ej,k)∑N\{j}

∑N

−(θi + θj)

∑k∈N\{i,j} θkVt+1(y − ei,j,k)∑N\{i,j}

∑N

+θi∑

k∈N\{i,j} θkVt+1(y − ei,k)∑N\{i}

∑N

+θjθ0Vt+1(y − ej)∑

N\{j}∑N

−(θi + θj)θ0Vt+1(y − ei,j)∑N\{i,j}

∑N

+θiθ0Vt+1(y − ei)∑

N\{i}∑N

+θiVt+1(y − ei,j)∑

N\{j}+θjVt+1(y − ei,j)∑

N\{i}− θiVt+1(y − ei) + θjVt+1(y − ej)∑

N

+θiθjθ0p∑

N∑N\{j}

∑N\{i,j}

+θiθjθ0p∑

N∑N\{i}

∑N\{i,j}

(B.2)

93

Page 102: Assortment Planning with Inventory Considerations in

The first two terms are positive due to Property 1 for period t+ 1, then

(B.2) ≥

θj∑

k∈N\{i,j} θk(Vt+1(y − ei,j)− Vt+1(y − ei,j,k)− Vt+1(y − ej) + Vt+1(y − ej,k)∑N\{j}

∑N

+θi∑

k∈N\{i,j} θk(Vt+1(y − ei,j)− Vt+1(y − ei,j,k)− Vt+1(y − ei) + Vt+1(y − ei,k))∑N\{i}

∑N

− θiθjθ0Vt+1(y − ei)∑N\{i,j}

∑N\{i}

∑N− θiθjθ0Vt+1(y − ej)∑

N\{i,j}∑N\{j}

∑N

+θiθj

∑k∈N\{i,j} θk(Vt+1(y − ei,j)− Vt+1(y − ei,j,k)− Vt+1(y − ei,j,k))∑

N\{i,j}∑N\{i}

∑N

+θjθi

∑k∈N\{i,j} θk(Vt+1(y − ei,j)− Vt+1(y − ei,j,k)− Vt+1(y − ei,j,k))∑

N\{i,j}∑N\{j}

∑N

+θiθj∑

N∑N\{j}

∑N\{i,j}

[θ0p+ (∑

k∈N\{i,j}

θk + θ0)Vt+1(y − ei,j)]

+θiθj∑

N∑N\{i}

∑N\{i,j}

[θ0p+ (∑

k∈N\{i,j}

θk + θ0)Vt+1(y − ei,j)] (B.3)

The first two terms are positive due to Property 1 for period t + 1. For the

last two terms we use the following two inequalities from Property 3 for period

t + 1 : θ0p + (∑N\{i,j})Vt(y − ei,j) ≥

∑k∈N\{i,j} θkVt(y − ej,k) + θ0Vt(y − ej) and

θ0p+ (∑N\{i,j})Vt(y − ei,j) ≥

∑k∈N\{i,j} θkVt(y − ei,k) + θ0Vt(y − ei). Then,

(B.3) ≥

θiθj∑

k∈N\{i,j} θk(Vt+1(y − ei,j)− 2Vt+1(y − ei,j,k) + Vt+1(y − ei,k))∑N\{i,j}

∑N\{i}

∑N

+θjθi

∑k∈N\{i,j} θk(Vt+1(y − ei,j)− 2Vt+1(y − ei,j,k) + Vt+1(y − ej,k))∑

N\{i,j}∑N\{j}

∑N

94

Page 103: Assortment Planning with Inventory Considerations in

This is positive.

We next prove Property 2. We focus on the case where yi = yj = 1. The case

where yi = yj > 1 is easier to prove.

Vt(y − ej)− Vt(y − ei) =(θi − θj)θ0p∑N\{j}

∑N\{i}

+

∑k∈N∪{0}\{j} θkVt+1(y − ej − ek)∑

N\{j}−∑

k∈N∪{0}\{i} θkVt+1(y − ei − ek)∑N\{i}

=(θi − θj)[θ0p+ (

∑k∈N\{i,j} θk + θ0)Vt+1(y − ei,j)−

∑k∈N∪{0}\{i,j} θkVt+1(y − ei,k)]∑

N\{j}∑N\{i}

+

∑k∈N∪{0}\{i,j} θk[Vt+1(y − ej,k)− Vt+1(y − ei,k)]∑

N\{j}

This is positive, because θ0p+(∑N\{i,j})Vt+1(y−ei,j)−

∑k∈N∪{0}\{i,j} θkVt+1(y−

ei,k) ≥ 0 from Property 3 for period t+ 1 and Vt+1(y− ej,k)−Vt+1(y− ei,k) ≥ 0 from

Property 2 for period t+ 1.

(iii): Because we have one more period to sell products in period t compared to

period t + 1, we can easily prove that Vt(y) and ∆it(y) are decreasing in time by

sample path analysis. �

Theorem 2.4.1

Proof: We now focus on Property 3. For simplicity, define S ′ as follows: j ∈ S ′

if and only if yj = 1 and j ∈ S. There are two cases: (1) yi > 1; (2) yi = 1. We

only show the proof of case 2 here and case 1 can be proved similarly. Moreover, we

assume S ′ 6= ∅ and θi > θj for any j ∈ S ′.

θ0p+ (∑S

)Vt(y − ei)−∑j∈S

θjVt(y − ej)− θ0Vt(y)

95

Page 104: Assortment Planning with Inventory Considerations in

=

∑k∈N\S\{i} θk[θ0p+ (

∑S)Vt+1(y − ei,k)−

∑j∈S θjVt+1(y − ej,k) + θ0Vt+1(y − ek)]∑

N

+

∑k∈N\S\{i} θkθi(

∑S)Vt+1(y − ei,k)∑

N∑N\{i}

−∑

k∈N\S\{i} θk∑N

[∑j∈S′

θ2j

Vt+1(y − ej,k)∑N\{j}

]

+

∑k∈S\S′ θk[θ0p+ (

∑S)Vt+1(y − ei,k)−

∑j∈S θjVt+1(y − ej,k) + θ0Vt+1(y − ek)]∑

N

+

∑k∈S\S′ θkθi(

∑S)Vt+1(y − ei,k)∑

N∑N\{i}

−∑

k∈S\S′ θk∑N

[∑j∈S′

θ2j

Vt+1(y − ej − ek)∑N\{j}

]

+

∑k∈S′ θk[θ0p+ (

∑S\{k})Vt+1(y − ei,k)−

∑j∈S\{k} θjVt+1(y − ej,k) + θ0Vt+1(y − ek)]∑N

+

∑k∈S′ θkθi(

∑S\{k})Vt+1(y − ei,k)∑N∑N\{i}

−∑

k∈S′ θk∑N

[∑

j∈S′\{k}

θ2j

Vt+1(y − ej,k)∑N\{j}

]

+

∑j∈S′ θ

2j (Vt+1(y − ei,j)∑N\{i}

+θ0[θ0p+ (

∑S)Vt+1(y − ei)−

∑j∈S θjVt+1(y − ej) + θ0Vt+1(y)]∑N

+θ0θi(

∑S)Vt+1(y − ei)∑N∑N\{i}

− θ0∑k∈N θk + θ0

[∑j∈S′

θ2j

Vt+1(y − ej)∑N\{j}

]

+θiθ0

∑k∈N\S\{i} θkp∑N∑N\{i}

+θ0p∑

k∈N θk + θ0

[∑j∈S′

θ2j∑N\{j}

]

−θi∑

j∈S\S′ θjVt+1(y − ei,j)∑N

−∑j∈S′

θiθjVt+1(y − ei,j)∑N\{j}

− θiθ0Vt+1(y − ei)∑N

(B.4)

The first, fourth, seventh, and eleventh terms are positive due to Property 3 for

period t+ 1, then

96

Page 105: Assortment Planning with Inventory Considerations in

(B.4) ≥θiθ0

∑k∈N\S\{i} θkp∑N∑N\{i}

+

∑k∈N\S\{i} θkθi(

∑S)Vt+1(y − ei,k)∑

N∑N\{i}

−∑

k∈N\S\{i} θk∑N

[∑j∈S′

θ2j

Vt+1(y − ej,k)∑N\{j}

]

+

∑k∈S\S′ θkθi(

∑S)Vt+1(y − ei,k)∑

N∑N\{i}

−∑

k∈S\S′ θk∑N

[∑j∈S′

θ2j

Vt+1(y − ej,k)∑N\{j}

]

+

∑k∈S′ θkθi(

∑S\{k})Vt+1(y − ei,k)

(∑N∑N\{i}

−∑

k∈S′ θk∑N

[∑

j∈S′\{k}

θ2j

Vt+1(y − ej,k)∑N\{j}

]

+

∑j∈S′ θ

2j (Vt+1(y − ei,j)∑N\{i}

+θ0θi(

∑S)Vt+1(y − ei)∑N∑N\{i}

− θ0∑N

[∑j∈S′

θ2j

Vt+1(y − ej)∑N\{j}

]

+θ0p∑

k∈N θk + θ0

[∑j∈S′

θ2j∑N\{j}

]

−θi∑

j∈S\S′ θjVt+1(y − ei,j)∑N

−∑j∈S′

θiθjVt+1(y − ei,j)∑N\{j}

− θiθ0Vt+1(y − ei)∑N

(B.5)

Using θ0p + (∑S)Vt+1(y − ei,k) ≥

∑j∈S θjVt+1(y − ei,j) + θ0Vt+1(y − ei) from

Property 3 in period t+ 1 in the first two terms, we get

(B.5) ≥

∑j∈S′

θ2j (θ0p+ (

∑N\{j})Vt+1(y − ei,j)−

∑k∈N\{j} θkVt+1(y − ej,k) + θ0Vt+1(y − ej))∑N\{j}

∑N

This is positive from Property 3 in period t + 1. This completes the proof by

induction. �

97

Page 106: Assortment Planning with Inventory Considerations in

Parts (i) and (ii) of Proposition 2.4.2, Theorem 2.4.2, and part (i) of Proposition

2.4.4 follow from the properties described below. Note that these results are pre-

sented for segment 2 customers only – we can similarly prove the results for segment

1.

Property 4 ∆2t (y1, y2) is decreasing in yi, i ∈ {1, 2}.

Property 5 ∆2t (y1, y2) is decreasing in t.

Property 6 p − ∆2t (y1, y2) ≥ θ21

θ21+θ0(p − ∆1

t (y1, y2)). In other words, θ0p + (θ0 +

θ21)Vt(y1, y2 − 1) ≥ θ0Vt(y1, y2) + θ21Vt(y1 − 1, y2).

Property 7

∆1t (y1, y2)−∆1

t (y1 + 1, y2) ≥ θ22

θ22 + θ0

(∆2t (y1, y2)−∆2

t (y1 + 1, y2)).

In other words,

(θ22 + 2θ0)Vt(y1, y2) + θ22Vt(y1, y2 − 1) ≥

(θ22 + θ0)Vt(y1 − 1, y2) + θ22Vt(y1 + 1, y2 − 1) + θ0Vt(y1 + 1, y2)

Property 8 If p−∆1t (y1, y2) ≥ p−∆2

t (y1, y2), then p−∆1t (y1, y2−1) ≥ p−∆2

t (y1, y2−

1). In other words, If Vt(y1 − 1, y2) ≥ Vt(y1, y2 − 1), then Vt(y1 − 1, y2 − 1) ≥

Vt(y1, y2 − 2).

Property 9 Vt(y1, y2 + 1) > Vt(y1 + 1, y2) if and only if y2 < y1. In addition, if

y2 = y1, Vt(y1, y2 + 1) = Vt(y1 + 1, y2).

Properties 4 and 5 are parts (i) and (ii) of Proposition 2.4.2, respectively. Prop-

erty 6 implies that product 2 is always in the optimal assortment of segment 2 and

Property 7 implies the threshold policy, which are the results in Theorem 2.4.2.

98

Page 107: Assortment Planning with Inventory Considerations in

Property 8 indicates that, if it is optimal to offer {1, 2} for customer segment 2 and

{1} for customer segment 1 under inventory (y1, y2), the optimal assortment for cus-

tomer segment 2 remains to be {1, 2} under inventory (y1, y2−1), which implies that,

if y1 ≥ y∗1(y2) and y2 ≤ y∗2(y1), then y1 ≥ y∗1(y2 − 1). This is part (i) of Proposition

2.4.4. These results are proved together by induction. First, because the optimal

assortment in period T is to offer all available products, it is easy to verify that the

results hold for period T . Second, we assume these four results hold in period t+ 1.

Note that Properties 6 and 7 in period t+ 1 implies that the optimal assortment in

period t is of threshold-type. Finally, we prove Properties 4 and 5 in period t in the

proof of Proposition 2.4.2, Properties 6 and 7 in period t in the proof of Theorem

2.4.2, and Properties 8 and 9 in period t in the proof of Proposition 2.4.4. Then, the

entire proof is completed by induction. Since Vt(y1, y2) is a linear combination of

Vt(y1, y2|m)(m = 1, 2), if we can show that the results hold for Vt(y1, y2|m)(m = 1, 2),

then they are also true for Vt(y1, y2). Without loss of generality, assume S(y) = N

and p−∆1t+1(y1, y2) ≥ p−∆2

t+1(y1, y2).

Proposition 2.4.2

Proof: We start with Property 4 and focus on proving the case where Vt(y1, y2) −

Vt(y1 − 1, y2) − Vt(y1, y2 + 1) + Vt(y1 − 1, y2 + 1) ≥ 0. The other cases are similar.

There are in total twenty four cases. The optimal assortments for both customer

segments can be summarized as follows.

(y1, y2) (y1 − 1, y2) (y1, y2 + 1) (y1 − 1, y2 + 1)S∗t (1) {1, 2} or {1} {1, 2} or {1} {1, 2} or {1} {1, 2} or {1}S∗t (2) {1, 2} {1, 2} or {2} {1, 2} or {2} {1, 2} or {2}

Here, we just show the proof of the case for segment 1 customers with the optimal

assortments below and the other cases are either easier or similar.

(y1, y2) (y1 − 1, y2) (y1, y2 + 1) (y1 − 1, y2 + 1)S∗t (1) {1} {1, 2} {1, 2} {1, 2}

99

Page 108: Assortment Planning with Inventory Considerations in

According to these optimal assortments, we have p − ∆2t+1(y1, y2) < θ11

θ11+θ0[p −

∆1t+1(y1, y2)], p−∆2

t+1(y1−1, y2) ≥ θ11θ11+θ0

[p−∆1t+1(y1−1, y2)], p−∆2

t+1(y1, y2 +1) ≥

θ11θ11+θ0

[p−∆1t+1(y1, y2+1)], and p−∆2

t+1(y1−1, y2+1) ≥ θ11θ11+θ0

[p−∆1t+1(y1−1, y2+1)].

Then

Vt(y1, y2|1)− Vt(y1 − 1, y2|1)− Vt(y1, y2 + 1|1) + Vt(y1 − 1, y2 + 1|1)

=θ12[θ11Vt+1(y1 − 1, y2) + θ0Vt+1(y1, y2)− θ0p− (θ11 + θ0)Vt+1(y1, y2 − 1)]

(θ11 + θ0)(θ11 + θ12 + θ0)

+θ11[Vt+1(y1 − 1, y2)− Vt+1(y1 − 2, y2)− Vt+1(y1 − 1, y2 + 1) + Vt+1(y1 − 2, y2 + 1))]

θ11 + θ12 + θ0

+θ12[Vt+1(y1, y2 − 1)− Vt+1(y1 − 1, y2 − 1)− Vt+1(y1, y2) + Vt+1(y1 − 1, y2)]

θ11 + θ12 + θ0

+θ0[Vt+1(y1, y2)− Vt+1(y1 − 1, y2)− Vt+1(y1, y2 + 1) + Vt+1(y1 − 1, y2 + 1)]

θ11 + θ12 + θ0

(B.6)

The first term is positive because p−∆2t+1(y1, y2) < θ11

θ11+θ0[p−∆1

t+1(y1, y2)] which is

equivalent to θ0p+(θ11 +θ0)Vt+1(y1, y2−1) < θ11Vt+1(y1−1, y2)+θ0Vt+1(y1, y2). The

last three terms are positive from Property 4 in period t+ 1. Therefore, (B.6) ≥ 0.

We then turn our attention to Property 5. There are in total four cases and the

optimal assortments for both customer segments can be summarized as follows.

(y1, y2) (y1, y2 − 1)S∗t (1) {1, 2} or {1} {1, 2} or {1}S∗t (2) {1, 2} {1, 2}

We focus on the case below, the others are similar.

(y1, y2) (y1, y2 − 1)S∗t (1) {1, 2} {1}

100

Page 109: Assortment Planning with Inventory Considerations in

According to the optimal assortments, we have p−∆2t+1(y1, y2 − 1) < θ11

θ11+θ0[p−

∆1t+1(y1, y2 − 1)]. Then

Vt(y1, y2|1)− Vt(y1, y2 − 1|1)− Vt+1(y1, y2) + Vt+1(y1, y2 − 1)

=θ12

(θ11 + θ12 + θ0)(θ11 + θ0)[θ0p+ (θ11 + θ0)Vt+1(y1, y2 − 1)] +

θ11Vt+1(y1 − 1, y2)

θ11 + θ12 + θ0

− θ11

θ11 + θ0

Vt+1(y1 − 1, y2 − 1)− θ11 + θ12

θ11 + θ12 + θ0

Vt+1(y1, y2) +θ11Vt+1(y1, y2 − 1)

θ11 + θ0

≥ θ11[Vt+1(y1 − 1, y2) + Vt+1(y1, y2 − 1)− Vt+1(y1 − 1, y2 − 1)− Vt+1(y1, y2)]

θ11 + θ0

≥ 0

The first inequality follows from using θ0p+(θ11+θ0)Vt+1(y1, y2−1) ≥ θ11Vt+1(y1−

1, y2) + θ0Vt+1(y1, y2) in the first term, which in turn follows from p−∆2t+1(y1, y2) ≥

θ11θ11+θ0

[p−∆1t+1(y1, y2)]. The last inequality is due to Property 4 for period t+ 1. �

Theorem 2.4.2

Proof: We now prove Property 6. There are in total twelve cases and the opti-

mal assortments for both customer segments under different inventory levels can be

summarized as follows.

(y1, y2) (y1 − 1, y2) (y1, y2 − 1)S∗t (1) {1, 2} or {1} {1, 2} or {1} {1, 2} or {1}S∗t (2) {1, 2} {1, 2} or {2} {1, 2} or {2}

Here, we just show the proof for the case below. The other cases are similar.

(y1, y2) (y1 − 1, y2) (y1, y2 − 1)S∗t (1) {1} {1, 2} {1}

According to the optimal assortments, we have p − ∆2t+1(y1, y2) < θ11

θ11+θ0[p −

∆1t+1(y1, y2)], p−∆2

t+1(y1− 1, y2) ≥ θ11θ11+θ0

[p−∆1t+1(y1− 1, y2)], and p−∆2

t+1(y1, y2−

101

Page 110: Assortment Planning with Inventory Considerations in

1) < θ11θ11+θ0

[p−∆1t+1(y1, y2 − 1)]. Then

θ0p+ (θ0 + θ21)Vt(y1, y2 − 1|1)− θ0Vt(y1, y2|1)− θ21Vt(y1 − 1, y2|1)

= θ0p+θ11θ21p

θ11 + θ0

− θ21(θ11 + θ12)p

θ11 + θ12 + θ0

− θ0[θ11Vt+1(y1 − 1, y2) + θ0Vt+1(y1, y2)]

θ0 + θ11

+(θ0 + θ21)[θ11Vt+1(y1 − 1, y2 − 1) + θ0Vt+1(y1, y2 − 1)]

θ0 + θ11

− θ21[θ11Vt+1(y1 − 2, y2) + θ12Vt+1(y1 − 1, y2 − 1) + θ0Vt+1(y1 − 1, y2)]

θ11 + θ12 + θ0

(B.7)

Because (θ0 +θ21)Vt+1(y1, y2−1) ≥ θ0Vt+1(y1, y2)+θ21Vt+1(y1−1, y2)−θ0p, which

follows from p−∆2t+1(y1, y2) < θ11

θ11+θ0[p−∆1

t+1(y1, y2)], then

(B.7) ≥

θ11[θ0p+ (θ0 + θ21)Vt+1(y1 − 1, y2 − 1)− θ0Vt+1(y1 − 1, y2)− θ21Vt+1(y1 − 2, y2)]

θ11 + θ12 + θ0

+θ0θ12(θ11 − θ21)[p− Vt+1(y1 − 1, y2) + Vt+1(y1 − 1, y2 − 1)]

(θ0 + θ11)(θ11 + θ12 + θ0)

The first term is positive because p−∆2t+1(y1−1, y2) ≥ θ11

θ11+θ0[p−∆1

t+1(y1−1, y2)]

and the second term is also positive. Therefore, the proof is finished.

Finally, we move to Property 7. There are in total twelve cases and the opti-

mal assortments for both customer segments under different inventory levels can be

summarized as follows.

(y1, y2) (y1, y2 − 1) (y1 − 1, y2) (y1 + 1, y2 − 1) (y1 + 1, y2)S∗t (1) {1, 2}/{1} {1, 2} / {1} {1, 2} / {1} {1, 2} /{1} {1, 2} /{1}S∗t (2) {1, 2} {1, 2} / {2} {1, 2} /{2} {1, 2} /{2} {1, 2} / {2}

Here, we just show the proof for the case below. The other cases are similar.

(y1, y2) (y1, y2 − 1) (y1 − 1, y2) (y1 + 1, y2 − 1) (y1 + 1, y2)S∗t (1) {1, 2} {1} {1, 2} {1} {1}

102

Page 111: Assortment Planning with Inventory Considerations in

According to the optimal assortments, we have p−∆2t+1(y1, y2 − 1) < θ11

θ11+θ0[p−

∆1t+1(y1, y2 − 1)], p − ∆2

t+1(y1 + 1, y2 − 1) < θ11θ11+θ0

[p − ∆1t+1(y1 + 1, y2 − 1)], and

p−∆2t+1(y1 + 1, y2) < θ11

θ11+θ0[p−∆1

t+1(y1 + 1, y2)]. Then,

(θ22 + 2θ0)Vt(y1, y2|1) + θ22Vt(y1, y2 − 1|1)− (θ22 + θ0)Vt(y1 − 1, y2|1)

−θ22Vt(y1 + 1, y2 − 1|1)− θ0Vt(y1 + 1, y2|1)

=θ12θ

20p

(θ11 + θ12 + θ0)(θ11 + θ0)+

θ11

θ11 + θ12 + θ0

[(θ22 + 2θ0)Vt+1(y1 − 1, y2)

+θ22Vt+1(y1−1, y2−1)−(θ22 +θ0)Vt+1(y1−2, y2)−θ22Vt+1(y1, y2−1)−θ0Vt+1(y1, y2)]

+θ12

θ11 + θ12 + θ0

[(θ22 + 2θ0)Vt+1(y1, y2 − 1)− (θ22 + θ0)Vt+1(y1 − 1, y2 − 1)]

+θ0

θ11 + θ12 + θ0

[(θ22 + 2θ0)Vt+1(y1, y2) + θ22Vt+1(y1, y2 − 1)

−(θ22 + θ0)Vt+1(y1 − 1, y2)− θ22Vt+1(y1 + 1, y2 − 1)− θ0Vt+1(y1 + 1, y2)]

+θ11θ12

(θ11 + θ0)(θ11 + θ12 + θ0)[θ22Vt+1(y1−1, y2−1)−θ22Vt+1(y1, y2−1)−θ0Vt+1(y1, y2)]

+θ0θ12

(θ11 + θ0)(θ11 + θ12 + θ0)[θ22Vt+1(y1, y2−1)−θ22Vt+1(y1+1, y2−1)−θ0Vt+1(y1+1, y2)]

(B.8)

In the second and fourth term, by using Property 6 in period t+ 1, we get

(B.8) ≥ θ12θ20p

(θ11 + θ12 + θ0)(θ11 + θ0)+

θ12

θ11 + θ12 + θ0

(θ22 + θ0)Vt+1(y1, y2 − 1)

+θ12

θ11 + θ12 + θ0

[θ0Vt+1(y1, y2 − 1)− (θ22 + θ0)Vt+1(y1 − 1, y2 − 1)]

+θ11θ12

(θ11 + θ0)(θ11 + θ12 + θ0)[θ22Vt+1(y1−1, y2−1)−θ22Vt+1(y1, y2−1)−θ0Vt+1(y1, y2)]

103

Page 112: Assortment Planning with Inventory Considerations in

+θ0θ12

(θ11 + θ0)(θ11 + θ12 + θ0)[θ22Vt+1(y1, y2−1)−θ22Vt+1(y1+1, y2−1)−θ0Vt+1(y1+1, y2)]

(B.9)

Using p − ∆2t+1(y1, y2) ≥ θ11

θ11+θ0[p − ∆1

t+1(y1, y2)], which is equivalent to (θ11 +

θ0)Vt+1(y1, y2 − 1) ≥ θ0Vt+1(y1, y2) + θ11Vt+1(y1 − 1, y2)− θ0p in the second term, we

get

(B.9) ≥

θ12θ0θ22[2Vt+1(y1, y2 − 1)− Vt+1(y1 − 1, y2 − 1)− Vt+1(y1 + 1, y2 − 1)]

(θ11 + θ12 + θ0)(θ11 + θ0)

+θ12θ0θ11[Vt+1(y1, y2 − 1)− Vt+1(y1 − 1, y2 − 1)− Vt+1(y1, y2) + Vt+1(y1 − 1, y2)]

(θ11 + θ12 + θ0)(θ11 + θ0)

+θ12θ

20[Vt+1(y1, y2 − 1)− Vt+1(y1 − 1, y2 − 1) + Vt+1(y1, y2)− Vt+1(y1 + 1, y2)]

(θ11 + θ12 + θ0)(θ11 + θ0)

All three terms are positive from Property 4 in period t+ 1. �

Proposition 2.4.3

Proof: For any inventory y, if p1t (y) ≥ p2

t (y) , from Proposition 2.3.1, the optimal

assortments for both customer segments then are either {1} or {1, 2}. However,

product 2 is always offered to customer segment 2. Therefore, the optimal as-

sortment for customer segment 2 is {1, 2}. With a similar argument, the optimal

assortment for customer segment 1 is {1, 2} if p1t (y) < p2

t (y). In conclusion, the

optimal assortment for at least one customer segment is {1, 2}. �.

Proposition 2.4.4

Proof: (i) We now prove Properties 8 and 9. We first show the proof of Property

9 and then go to Property 8. For Property 9, if y2 = y1, because of the symmetry of

both customer segments, adding one more unit to any product leads to the same profit

104

Page 113: Assortment Planning with Inventory Considerations in

functions, that is , Vt(y1, y2+1) = Vt(y1+1, y2). If y2 < y1, consider a new assortment

policy for inventory (y1, y2 + 1) as follows. In period t, the assortment policy follows

the optimal assortment policy of period t under inventory (y1 + 1, y2). From period

t+1 to the end of season, we use the optimal assortment policy. We denote the profit

function under this new assortment policy as V t(y1, y2+1). Therefore, Vt(y1, y2+1) ≥

V t(y1, y2 + 1). In the following, we prove V t(y1, y2 + 1) > Vt(y1 + 1, y2) by induction

and then we have Vt(y1, y2 + 1) > Vt(y1 + 1, y2). When comparing V t(y1, y2 + 1)

and Vt(y1 + 1, y2), if the customer arriving in period t chooses not to purchase, then

V t(y1, y2 + 1) = V t+1(y1, y2 + 1) > Vt+1(y1 + 1, y2) = Vt(y1 + 1, y2) by induction; if

a purchase is made, then V t(y1, y2 + 1) ≥ Vt(y1 + 1, y2) regardless of which product

is sold or which assortment is offered. Therefore, V t(y1, y2 + 1) > Vt(y1 + 1, y2) in

expectation. On the other hand, we prove that Vt(y1, y2 + 1) > Vt(y1 + 1, y2), only

if y2 < y1. Because, y2 ≥ y1 implies that Vt(y1 + 1, y2) ≥ Vt(y1, y2 + 1) based on our

previous argument.

Based on Property 9 in period t, p − ∆1t (y1, y2) ≥ p − ∆2

t (y1, y2) if and only if

y2 ≤ y1 and p−∆1t (y1, y2−1) ≥ p−∆2

t (y1, y2−1) if and only if y2−1 ≤ y1. Property

8 then can be restated as if y2 ≤ y1, then y2 − 1 ≤ y1. This is obviously correct.

(ii) We only consider segment 2 customers without loss of generality. Assume

the optimal assortment is {1, 2} in period t with the inventory level (y1, y2), i.e.,

θ0(p−∆1t+1(y1, y2)) + θ22(∆2

t+1(y1, y2)−∆1t+1(y1, y2)) ≥ 0. If we can prove that the

optimal assortment still is {1, 2} in period t + 1 with the same inventory level, i.e.,

θ0(p − ∆1t+2(y1, y2)) + θ22(∆2

t+2(y1, y2) − ∆1t+2(y1, y2)) ≥ 0., then the proof is com-

pleted. If y1 ≥ y2, then ∆2t+2(y1, y2)−∆1

t+2(y1, y2) = Vt+2(y1 − 1, y2)− Vt+2(y1, y2 −

1) ≥ 0 in this symmetric case. If y1 < y2, Vt+1(y1, y2 − 1) − Vt+1(y1 − 1, y2) ≥

Vt+2(y1, y2 − 1) − Vt+2(y1 − 1, y2), which implies that ∆2t+2(y1, y2) − ∆1

t+2(y1, y2) ≥

∆2t+1(y1, y2)−∆1

t+1(y1, y2). Moreover, from part 2 of Proposition 2.4.2, ∆it+1(y1, y2) >

∆it+2(y1, y2). Therefore, θ0(p − ∆1

t+2(y1, y2)) + θ22(∆2t+2(y1, y2) − ∆1

t+2(y1, y2)) ≥

105

Page 114: Assortment Planning with Inventory Considerations in

θ0(p−∆1t+1(y1, y2)) + θ22(∆2

t+1(y1, y2)−∆1t+1(y1, y2)) ≥ 0. �

Proposition 2.6.1

Proof: Assume µ is a feasible policy in the MDP problem. Let Sµmt(Ft) denote

the assortment controlled by policy µ in period t for segment m customer and an N-

dimensional vector D(Sµmt(Ft)) denote the random demand in period t given segment

m consumer and assortment Sµmt(Ft). To be more specific, Dj(Sµmt(Ft)) = 1 indicates

that one unit of product j is sold and Dj(Sµmt(Ft)) = 0 indicates no sale for product

j. In addition, if j ∈ N \ Sµmt(Ft), Dj(Sµmt(Ft)) = 0. Because of the feasibility of

policy µ, we have

T∑t=1

∑m∈M

D(Sµmt(Ft)) ≤ y0, a.s.

We take the expectation and have∑T

t=1

∑m∈MED(Sµmt(Ft)) ≤ y0. Further-

more,

T∑t=1

∑m∈M

ED(Sµmt(Ft)) =∑m∈M

∑S⊂N

λQm(S)E[tµm(S)]

where tµm(S) indicates the random amount of time for which assortment S is

offered to segment m consumers under policy µ. The objective function then is

V µ1 (y0) =

∑m∈M

∑S⊂N

λRm(S)E[tµm(S)]

Then (E[tµm(S)])m∈M are feasible solution in CDLP and the profits are the same.

Based on this result, we then have V1(y0) = maxµ∈A Vµ

1 (y0) ≤ V CDLP (y0). �

106

Page 115: Assortment Planning with Inventory Considerations in

Appendix D

Proofs of Chapter 3

Proof of Theorem 3.2.1: We start by proving the result for the IP model. Because

cost/price ratios are identical across all products, their percentiles denoted by zIP =

Φ−1(1−r) = Φ−1(1−ki/pi) are identical as well. Define f1(q) = (1−r)q−φ(zIP )σqβ.

The optimal profit for product i then is πD,IP (qi) = pif1(qi). From Lemma B.1,

f1(qi) is increasing and convex in qi. In addition, qi = ui∑j∈S uj+u0

implies that pi =

p0 + γqi(

∑i∈S\{i} uj)

1−qi which is increasing and convex in qi. Therefore,

∂πD,IP (qi)

∂qi= p′if1(qi) + pif

′1(qi) > 0,

∂2πD,IP (qi)

∂2qi= p′′i f1(qi) + 2p′if

′1(qi) + pif

′′1 (qi) > 0.

In other words, πD,IP (qi) is increasing and convex in qi. Hence, as in the proof of

Proposition B.1, the optimal assortment is a popular set for the IP model.

For the TF model, all products have identical percentiles given by zTF = Φ−1(1 −rq). Define f2(q) = (q − r) − qφ(zTF )σ. The optimal profit for product i then is

πD,TF (qi) = pif2(qi). From Lemma B.1, f1(qi) is increasing and convex in qi. Similar

107

Page 116: Assortment Planning with Inventory Considerations in

as the previous proof for the IP model, the optimal assortment is a popular set for

the TF model. �

Proof of Theorem 3.2.2: We first prove that ΠC,TFS is increasing and convex in qi,

which implies that the optimal assortment is a popular set. With a similar argument

to that in the identical price case, the variants in Se have the same inventory levels

as the the common component, whose critical fractile is t∗c(S) = 1 − kc+∑i∈Se kdi∑

i∈Se piqSi

=

1− kc+τr∑i∈Se pi∑

i∈Se piqSi

. The optimal profit of set Se is

ΠC,TFSe

= πD,TF (∑i∈Se

piqSi , kc +

∑i∈Se

kdi)

=∑i∈Se

piqi[t∗c(S)− σφ(Φ−1(t∗c(S)))]

If we can prove that ΠC,TFSe

is increasing and convex in qi where i ∈ Se, then the

optimal assortment is a popular set. This is because ΠC,TFSl

=∑

i∈Sl πD,TF (piq

Si , kdi)

is increasing and convex in qi where i ∈ Sl, the proof of which is similar to the proof

of Theorem 3.2.1. In other words, we need to prove that f(qi) =∑

j∈Se pjqj[t∗c(S)−

σφ(Φ−1(t∗c(S)))] is increasing and convex in qi. First, we show that the common

fractile t∗c(S) = 1 − kc+τr∑j∈Se pj∑

j∈Se pjqjsatisfies two properties below which leads to the

result that f(qi) is increasing and convex in qi. Based on KKT conditions, we have

a KKT multiplier αSi for product i which is in Se. Moreover,

1− t∗c(S) =kdj + αSjpjqj

|j∈Se =kdi + αSipiqi

>kdipiqi

Therefore,

kc =∑j∈Se

αSj ≥ τr

∑j∈Se pjqj

qi− τr

∑j∈Se

pj

108

Page 117: Assortment Planning with Inventory Considerations in

In other words,

qi(kc + τr∑j∈Se

pj) ≥ τr(∑j∈Se

pjqj)

The first order-derivative of t∗c(S) is given by

∂t∗c(S)

∂qi=

(kc + τr∑

j∈Se pj)(pi + p′iqi)− τrp′i(∑

j∈Se pjqj)

(∑

j∈Se pjqj)2

> 0

Moreover,

2(pi + qip′i)∂t∗c(S)

∂qi+ [∑j∈Se

pjqj]∂2t∗c(S)

∂2qi

=(kc + τr

∑j∈Se pj)(2p

′i + p′′i qi)− τrp′′i (

∑j∈Se pjqj)∑

j∈Se pjqj> 0 (B.1)

The first and second order derivatives of f(qi) are as follows.

∂f(qi)

∂qi= (pi+qip

′i)(t∗c(S)µ−σφ(Φ−1(t∗c(S))))+

∑j∈Se

pjqj∂t∗c(S)

∂qi(µ+σΦ−1(t∗c(S))) > 0

∂2f(qi)

∂2qi=

(2p′i + p′′i qi)(t∗c(S)µ− σφ(Φ−1(t∗c(S)))) + 2(pi + qip

′i)∂t∗c(S)

∂qi(µ+ σΦ−1(t∗c(S)))

+[∑j∈Se

pjqj]∂2t∗c(S)

∂2qi(µ+ σΦ−1(t∗c(S))) +

∑j∈Se

pjqj

(∂t∗c(S)

∂qi

)2

σ1

φ(Φ−1(t∗c(S)))> 0

The second inequality is due to (B.1). Thus, f(qi) is increasing and convex in qi.

Similarly, results analogous to Theorem 1.4.2 hold in the case of increasing prices.

Therefore, the optimal assortment is a popular set and the optimal assortment in

the system with commonality is no smaller than that in the dedicated system. �

109

Page 118: Assortment Planning with Inventory Considerations in

Proof of Theorem 3.2.3: The profit function with order quantity Q in model P is

ΠP (S,Q) =∑i∈S

piqiEmin{D,Q} − kQ.

After optimizing the order quantity, the optimal profit function is

ΠP (S) = πD,TF (∑

i∈S piqi, k). From Lemma B.1, this optimal profit function is

increasing and convex in qi. Thus, the optimal assortment is a popular set. The

proof of the result that the optimal assortment in model P is no smaller than that in

model D is similar to and easier than the proof of the comparison between models P

and C. Next, we prove that the optimal assortment in model P is no smaller than

that in model C. In other words, we need prove that ΠP,TFAn+1−ΠP,TF

An≥ ΠC,TF

An+1−ΠC,TF

An

for any popular set An.

When An+1 ⊂ Ne,

ΠP,TFAn+1− ΠP,TF

An= πD,TF (

∑i∈An+1

piqi, k)− πD,TF (∑i∈An

piqi, k)

≥ πD,TF (∑

i∈An+1

piqi, kc +∑

i∈An+1

kdi)− πD,TF (∑i∈An

piqi, kc +∑i∈An

kdi) = ΠC,TFAn+1

− ΠC,TFAn

This is because πD,TF (x, k) is decreasing in k and is submodular in (x, k). When

Ne ⊂ An,

ΠP,TFAn+1− ΠP,TF

An=

πD,TF (∑i∈Ne

piqAn+1

i , k

∑i∈Ne piq

An+1

i∑i∈An+1

piqAn+1

i

)− πD,TF (∑i∈Ne

piqAni , k

∑i∈Ne piq

Ani∑

i∈An piqAni

)

+∑

i∈An\Ne

[πD,TF (piqAn+1

i , kpiq

An+1

i∑i∈An+1

piqAn+1

i

)− πD,TF (piqAni , k

piqAni∑

i∈An piqAni

)]

+ πD,TF (pn+1qAn+1

n+1 , kpn+1q

An+1

n+1∑i∈An+1

piqAn+1

i

) (B.2)

110

Page 119: Assortment Planning with Inventory Considerations in

Because k ≤ kc+kd1 , k∑i∈Ne piq

An+1i∑

i∈An+1piq

An+1i

≤ k∑i∈Ne piq

Ani∑

i∈An piqAni

, kpiq

An+1i∑

i∈An+1piq

An+1i

≤ kpiq

Ani∑

i∈An piqAni

when i ∈ An \ Ne, and kpn+1q

An+1n+1∑

i∈An+1piq

An+1i

< kdn+1 . Moreover, since πD,TF (x, k) is

decreasing in k, then

(B.2) > πD,TF (∑i∈Ne

piqAn+1

i , k

∑i∈Ne piq

Ani∑

i∈An piqAni

)− πD,TF (∑i∈Ne

piqAni , k

∑i∈Ne piq

Ani∑

i∈An piqAni

)

+∑

i∈An\Ne

[πD,TF (piqAn+1

i , kpiq

Ani∑

i∈An piqAni

)− πD,TF (piqAni , k

piqAni∑

i∈An piqAni

)]

+πD,TF (pn+1qAn+1

n+1 , kdn+1) (B.3)

Because πD,TF (x, k) is submodular in (x, k), k∑i∈Ne piq

Ani∑

i∈An piqAni

≤ kc +∑

i∈Ne kdi , and

kpiq

Ani∑

i∈An piqAni

≤ kdi when i ∈ An \ Ne, we have

(B.3) ≥ πD,TF (∑i∈Ne

piqAn+1

i , kc +∑i∈Ne

kdi)− πD,TF (∑i∈Ne

piqAni , kc +

∑i∈Ne

kdi)

+∑

i∈An\Ne

[πD,TF (piqAn+1

i , kdi)− πD,TF (piqAni , kdi)] + πD,TF (pn+1q

An+1

n+1 , kdn+1) (B.4)

Note that the expression in (B.4) is equal to ΠC,TFAn+1−ΠC,TF

An. Thus, we have ΠP,TF

An+1−

ΠP,TFAn

≥ ΠC,TFAn+1−ΠC,TF

An, which implies that the optimal assortment in model P is no

smaller than that in model C. �

Proof of Proposition 3.3.1: According to the definition, ΠS = ΠS+{δj},pj =

ΠS+{δi},pi . Since the profit function is linear and increasing in price, ΠS+{δj},pi ≥

ΠS+{δj},pj = ΠS+{δi},pi . From Lemma B.1, then δj ≥ δi.�

Proof of Proposition 3.3.2: We prove the result by contradiction. Suppose

product j is not in assortment S. We include product j instead of product i. Then

ΠS−{i}+{j} > ΠS. This is a contradiction.�

111

Page 120: Assortment Planning with Inventory Considerations in

Proof of Proposition 3.3.3: The result follows from Proposition 3.3.2.�

112

Page 121: Assortment Planning with Inventory Considerations in

Bibliography

[1] S. Anderson, A. de Palma, and J. Thisse. Discrete Choice Theory of ProductDifferentiation. MIT Press, Cambridge, MA, 1992.

[2] Y. Bassok, R. Anupindi, and R. Akella. Single-period multiproduct inventorymodels with substitution. Operations Research, 47(4):632–642, 1999.

[3] F. Bernstein, G. DeCroix, and Y. Wang. Incentives and commonality in adecentralized multiproduct assembly system. Operations Research, 55(4):630–646, 2007.

[4] J. Miranda Bront, I. Mendez-Diaz, and G. Vulcano. A column generationalgorithm for choice-based network revenue management. Operations Research,57(3):769–784, 2009.

[5] G. P. Cachon, C. Terwiesch, and Y. Xu. Retail assortment planning in thepresence of consumer search. Manufacturing & Service Operations Management,7(4):330–346, 2005.

[6] F. Caro and J. Gallien. Dynamic assortment with demand learning for seasonalconsumer goods. Management Science, 53(2):276–292, 2007.

[7] L. Chen and T. Homem de Mello. Re-solving stochastic programming modelsfor airline revenue management. Annals of Oper. Res., 2009. Forthcoming.

[8] J. Chod and N. Rudi. Resource flexibility with responsive pricing. OperationsResearch, 53(3):532–548, 2005.

[9] F. de Vericourt, F. Karaesmen, and Y. Dallery. Optimal stock allocation fora capacitated supply system. Management Science, 48(11):1486–1501, 2002.

[10] P. Desai, S. Kekre, S. Radhakrishnan, and K. Srinivasan. Product differentiationand commonality in design: Balancing revenue and cost drivers. ManagementScience, 47(1):37–51, 2001.

113

Page 122: Assortment Planning with Inventory Considerations in

[11] G. Dobson and S. Kalish. Positioning and pricing a product line. MarketingScience, 7(2):107–125, 1988.

[12] F.Chen and Y. Bassok. Substitution and variety. Working paper, 2009. Univer-sity of Southern California.

[13] C. Fine and R. Freund. Optimal investment in product-flexible manufacturingcapacity. Management Science, 36(4):449–466, 1990.

[14] M. Fisher and C. Ittner. The impact of product variety on automobile assemblyoperations: Empirical evidence and simulation analysis. Management Science,45(6):771–786, 1999.

[15] G. Gallego, G. Iyengar, R. Phillips, and A. Dubey. Managing flexible productson a network. CORC Technical Report TR-2004-01, 2004.

[16] V. Gaur and D. Honhon. Product variety and inventory decisions under a loca-tional consumer choice model. Management Science, 52(10):1528–1543, 2006.

[17] G.P.Cachon and A.G. Kok. Category management and coordination in retailassortment planning in the presence of basket shopping consumers. ManagementScience, 53(6):934–951, 2007.

[18] S. Gupta and P. K. Chintagunta. On using demographic variables to determinesegment membership in logit mixture models. Journal of Marketing Research,31(1):128–136, 1994.

[19] A. Ha. Inventory rationing in a make-to-stock production system with severaldemand classes and lost sales. Management Science, 43(8):1093–1103, 1997a.

[20] A. Ha. Stock-rationing policy for a make-to-stock production system with twopriority classes and backordering. Naval Res. Logist., 44(5):458–472, 1997b.

[21] S. Heese and J. Swaminathan. Product line design with component commonalityand cost reduction effort. Manufacturing & Service Operations Management,8(2):206–219, 2006.

[22] S. Holmes. Nike. Businessweek Online, 2003.

[23] D. Honhon, V. Gaur, and S. Seshadri. Assortment planning and inventorymanagement under dynamic stockout-based substitution. Working paper, 2008.University of Texas at Austin.

114

Page 123: Assortment Planning with Inventory Considerations in

[24] W. Hopp and X. Xu. Product line selection and pricing with modularity indesign. Manufacturing & Service Operations Management, 7(3):172–187, 2005.

[25] B. Kahn. Dynamic relationships with customers: High-variety strategies. Jour-nal of the Academy of Marketing Science, 26(1):45–53, 1998.

[26] K. Kim and D. Chhajed. Commonality in product design: Cost saving, valuationchange and cannibalization. European J. of Operations Research, 125(3):602–621, 2000.

[27] A. G. Kok and M. Fisher. Demand estimation and assortment optimiza-tion under substitution: Methodology and application. Operations Research,55(6):1001–1021, 2007.

[28] A. G. Kok, M. Fisher, and R. Vaidyanathan. Assortment planning: Review ofliterature and industry practice. Retail Supply Chain Management, 2008. Eds.N. Agrawal,S.A. Smith. Springer.

[29] H. Lee and C. Tang. Modeling the costs and benefits of delayed productiondifferentiation. Management Science, 43(1):40–53, 1997.

[30] Q. Liu and G. J. van Ryzin. On the choice-based linear programming modelfor network revenue management. Manufacturing & Service Operations Man-agement, 10(2):288–310, 2008.

[31] J. Van Mieghem. Investment strategies for flexible resources. ManagementScience, 44(8):1071–1078, 1998.

[32] J. Van Mieghem. Commonality strategies: value drivers and equivalence withflexible capacity and inventory substitution. Management Science, 50(3):419–424, 2004.

[33] K. Ramdas. Managing product variety: An integrative review and researchdirections. Production and Operations Management, 12(1):79–101, 2003.

[34] K. Ramdas, M. Fisher, and K. Ulrich. Components sharing in the manage-ment of product variety: A study of automotive braking systems. ManagementScience, 45(5):297–315, 1999.

[35] B. Rodriguez and G. Aydin. Assortment selection and pricing for configurableproducts under demand uncertainty. Working paper, 2009. University of Michi-gan.

115

Page 124: Assortment Planning with Inventory Considerations in

[36] P. Singh, H. Groenevelt, and N. Rudi. Product variety and supply chain struc-tures. Working paper, 2005. University of Rochester.

[37] S. Smith and N. Agrawal. Management of multi-item retail inventory systemswith demand substitution. Operations Research, 48(1):50–64, 2000.

[38] J-S. Song and Y. Zhao. The value of commonality in a dynamic inventory systemwith lead times. Manufacturing & Service Operations Management, 11(3):493–508, 2009.

[39] J-S. Song and P. Zipkin. Handbooks in Operations Research and ManagementScience, Vol. 11. Supply Chain Management: Design, Coordination and Opera-tions. Eds. A. G. de Kok and S. Graves, chapter 11, pages 561–596. Elsevier,2003.

[40] K.T. Talluri and G. J. van Ryzin. The Theory and Practice of Revenue Man-agement. Kluwer Academic, New York, 2004a.

[41] K.T. Talluri and G. J. van Ryzin. Revenue management under a general discretechoice model of consumer behavior. Management Science, 50(1):15–33, 2004b.

[42] G. van Ryzin and S. Mahajan. On the relationship between inventory costs andvariety benefits in retail assorment. Management Science, 45(11):1496–1509,1999.

[43] G. van Ryzin and S. Mahajan. Stocking retail assortments under dynamicconsumer substitution. Operations Research, 49:334–351, 2001.

[44] G. J. van Ryzin and G. Vulcano. Simulation-based optimization of virtual nest-ing controls for network revenue management. Operations Research, 56(4):865–880, 2008a.

[45] G. J. van Ryzin and G. Vulcano. Computing virtual nesting controls for networkrevenue management under customer choice behavior. Manufacturing ServiceOperations Management, 10(3):448–467, 2008b.

[46] M. Wedel and W.A. Kamakura. Market segmentation: conceptual and method-ological foundations. Kluwer Academic, 1998.

[47] D. Zhang and D. Adelman. An approximate dynamic programming approachto network revenue management with customer choice. Transportation Science,43(3):381–394, 2009.

116

Page 125: Assortment Planning with Inventory Considerations in

[48] D. Zhang and W. L. Cooper. Revenue management for parallel flights withcustomer-choice behavior. Operations Research, 53(3):415–431, 2005.

117

Page 126: Assortment Planning with Inventory Considerations in

Biography

Lei Xie was born at Hunan province of P.R. China on 06 July, 1981. He entered

University of Science and Technology of China in 1999. After four-year study, he

gained his bachelor degree of science in Mathematics and Applied Mathematics. In

2003, he was admitted to the department of System Engineering and Engineering

Management in Chinese University of Hong Kong with postgraduate scholarship. It

took his two years to get a master degree of philosophy in Operations Management.

In 2005, Lei entered Duke University as a doctoral student and majored in Operations

Management in the department of Business Administration. His research interests

include product variety, revenue management, and the interface between supply chain

and finance. After he graduates in September 2010, Lei plans to move to Montreal,

Canada, where he will continue his research as a post-doc research fellow in McGill

University.

118