O ine Assortment Optimization in the Presence of an...

51
Oine Assortment Optimization in the Presence of an Online Channel Daria Dzyabura Stern School of Business, New York University, New York, NY 10012, [email protected] Srikanth Jagabathula Stern School of Business, New York University, New York, NY 10012, [email protected] Firms are increasingly selling through both oine and online channels, allowing customers to experience the touch and feel of product attributes before purchasing those products. Consequently, the selection of products oered oine aects the demand in both channels. We address how firms should select an optimal oine assortment to maximize profits across both channels; we call this the showcase decision problem. We incorporate the impact of physical evaluation on preferences into the consumer demand model. Under this model, we show that the decision problem is NP-hard. Analytically, we derive optimal results for special cases and near-optimal approximations for general cases. Empirically, we use conjoint analysis to identify changes in consumer preferences resulting from physically evaluating products. For this application, we demonstrate gains in expected revenue of up to 40% due to accounting for the impact of oine assortment on the online sales. Key words : Assortment optimization, multichannel, choice modeling, conjoint analysis 1. Introduction A growing number of firms are selling to their customers through both online and oine (or brick- and-mortar) channels. Selling through multiple channels allows the firm to reach customers who have diering channel preferences for purchasing. In addition, the firm can oer a wide selection of its products (at lower inventory costs) through its online channel and showcase its product line through its oine channel. Despite the growing significance of the online channel, maintaining oine channels is essential for firms because customers visit oine stores to physically inspect the products and gain tactile (or “touch-and-feel”) information, before making the purchase. Examples include furniture purchase (from firms such as Crate & Barrel, West Elm, etc.) and apparel pur- chase 1 (from firms such as Bonobos, MM Lafleur, etc.). Therefore, the selection of products that a customer is exposed to in the oine channel impacts her purchase behavior, and the firm faces the key operational problem of optimizing its oine selection with the objective of maximizing overall sales or profits. 1 In fact, there is a growing trend of firms such as Bonobos, Warby Parker, Birchbox, etc., which started online, but then opened brick-and-mortar stores to showcase their product lines. 1

Transcript of O ine Assortment Optimization in the Presence of an...

Page 1: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

O✏ine Assortment Optimization in the Presence ofan Online Channel

Daria DzyaburaStern School of Business, New York University, New York, NY 10012, [email protected]

Srikanth JagabathulaStern School of Business, New York University, New York, NY 10012, [email protected]

Firms are increasingly selling through both o✏ine and online channels, allowing customers to experience

the touch and feel of product attributes before purchasing those products. Consequently, the selection of

products o↵ered o✏ine a↵ects the demand in both channels. We address how firms should select an optimal

o✏ine assortment to maximize profits across both channels; we call this the showcase decision problem. We

incorporate the impact of physical evaluation on preferences into the consumer demand model. Under this

model, we show that the decision problem is NP-hard. Analytically, we derive optimal results for special

cases and near-optimal approximations for general cases. Empirically, we use conjoint analysis to identify

changes in consumer preferences resulting from physically evaluating products. For this application, we

demonstrate gains in expected revenue of up to 40% due to accounting for the impact of o✏ine assortment

on the online sales.

Key words : Assortment optimization, multichannel, choice modeling, conjoint analysis

1. Introduction

A growing number of firms are selling to their customers through both online and o✏ine (or brick-

and-mortar) channels. Selling through multiple channels allows the firm to reach customers who

have di↵ering channel preferences for purchasing. In addition, the firm can o↵er a wide selection

of its products (at lower inventory costs) through its online channel and showcase its product line

through its o✏ine channel. Despite the growing significance of the online channel, maintaining

o✏ine channels is essential for firms because customers visit o✏ine stores to physically inspect the

products and gain tactile (or “touch-and-feel”) information, before making the purchase. Examples

include furniture purchase (from firms such as Crate & Barrel, West Elm, etc.) and apparel pur-

chase1 (from firms such as Bonobos, MM Lafleur, etc.). Therefore, the selection of products that a

customer is exposed to in the o✏ine channel impacts her purchase behavior, and the firm faces the

key operational problem of optimizing its o✏ine selection with the objective of maximizing overall

sales or profits.

1 In fact, there is a growing trend of firms such as Bonobos, Warby Parker, Birchbox, etc., which started online, butthen opened brick-and-mortar stores to showcase their product lines.

1

Page 2: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel2 Article submitted to Management Science; manuscript no.

Existing work in operations and marketing provides guidance on how firms should optimize their

o↵erings, but most such work focuses on single-channel settings. While several such proposals exist,

at the core, they rely on restricting customer choices to trade o↵ losing profits by not o↵ering

low-profit products for gaining profits from switches to higher-profit products. These proposals,

however, do not extend to multichannel settings in which the o✏ine o↵er set does not necessarily

restrict choices as (some) customers may purchase from the (the typically larger) assortment o↵ered

online. Furthermore, they do not account for the fact that the assortment may change the product

the customer will purchase because of the “touch-and-feel” information provided by the o✏ine

channel. For example, consider a customer looking to purchase a messenger bag with a laptop

compartment and the store o↵ers a blue bag without the laptop compartment. In the absence of a

store visit, the customer would have purchased the black bag online, but after the store visit, the

customer realizes that she prefers blue to black and purchases the blue bag with a laptop compart-

ment online. In other words, the o✏ine channel is not only a sales channel but also an information

channel. Ignoring such interactions between channels will result in suboptimal decisions.

Motivated by the above considerations, this paper studies a firm’s showcase decision – that of

determining the subset of products from the online channel to o↵er in the o✏ine channel, in order

to maximize aggregate sales or profits across both channels. We focus on the following setup. A

firm is selling products through an online and an o✏ine channel. The products are di↵erentiated

but close substitutes – a single consumer purchases at most one product and the rest of the market

does not o↵er perfect substitutes. The products are generally infrequently purchased, or comprise a

large variety, so that customers can benefit from a visit to the o✏ine store to examine the products

physically. The customer is utility maximizing and purchases her maximum utility product if its

utility is greater than her no-purchase utility. Products are multi-attribute, and a product’s utility

is the sum of its attribute partworths, which capture consumers’ preferences for each attribute.

Product categories such as furniture, apparel & accessories, consumer electronics, etc., satisfy these

assumptions. The selection of products o↵ered online is large and fixed. The profit associated

with each product is exogenously specified and decomposes into the sum of the profits from the

constituent attributes. The objective of the firm is to choose a selection of products from the online

assortment to o↵er o✏ine to maximize the expected sales or profits from both channels.

In the context of the above setup, this work makes three key contributions: (a) a novel utility-

based model to capture the interactions between the online and o✏ine channels, (b) analytical

results on the structure of the optimal o↵er sets, and (c) a scalable integer-programming (IP)

based optimization algorithm to solve the firm’s showcase decision. We also validate our modeling

assumptions and our methods using real-world preference data on messenger bags.

Page 3: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online ChannelArticle submitted to Management Science; manuscript no. 3

The modeling contribution of this work is to extend the standard utility model to capture the

impact of physical evaluation through changing partworths. The standard utility-based models

suppose that the utility obtained from a product decomposes into the sum of the partworths (or

valuations) of the attributes comprising the product. They assume that customers arrive at the

partworths by evaluating the products, so they are fixed and known to the customers. However,

when products are sold online, customers may find it di�cult to evaluate some attributes based

only on their online descriptions or pictures; for instance, it may be di�cult ascertain how large

the large size is, how bright the blue color is, etc. For such attributes, the consumer may learn her

preferences by physically evaluating products with those attributes. We capture this learning by

allowing the partworth of an attribute to change upon physical evaluation. Thus, for each attribute,

the customer has an online- (or pre-evaluation) and an o✏ine- (or post-evaluation) partworth, and

the di↵erence between them quantifies the information gained from physical examination of that

attribute.

Under the above model, we derive analytical results for the structure of the optimal solution

to obtain insights on how the firm’s decisions di↵er from the single-channel setting when there is

also an online channel. We consider two separate settings: (a) all customers visit both channels,

and (b) only a portion of customers visits both channels (the online segment), while the others

purchase only what is available in the o✏ine store (the o✏ine segment). In the former case – which

we term the pure showcase setting – the o✏ine channel acts only as an information channel, while

in the latter case – which we term the general showcase setting – it acts as both an information

and sales channel. We make the standard assumption that when given the attribute partworths,

the customers make choices according to a multinomial logit (MNL) model. In the pure showcase

setting, we show that to maximize sales (Theorem 3.1), the firm must o↵er the attributes that

are under-valued (more attractive after physical evaluation) and hide the attributes that are over-

valued (less attractive after physical evaluation) by the customers. On the other hand, to maximize

profits (Theorem 3.2), the firm must o↵er the most profitable under-valued attributes and the least

profitable over-valued attributes. In doing this, the firm provides information to its customers,

resulting in an increase (decrease) in the attractiveness of the most (least) profitable attributes; this

shifts the demand to the most profitable attributes. In the general showcase setting, we show that

to maximize sales (Theorem 3.5), the firm must o↵er all the under-valued attributes and hide only

a subset of the over-valued attributes. Precisely which over-valued attributes are hidden depends

on their corresponding magnitudes of attractiveness. Intuitively, the over-valued attributes that

drive a large amount of o✏ine sales should be o↵ered but the rest hidden. In contrast, when there

is no online channel, the firm must o↵er all the attributes to maximize sales and o↵er the most

profitable attributes to maximize profits.

Page 4: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel4 Article submitted to Management Science; manuscript no.

Our second set of results addresses computation. We show that the pure showcase sales maxi-

mization problem can be e�ciently solved. The pure showcase profit maximization and the general

showcase sales (and, hence, profit) maximization problems are NP-hard to solve (Theorems 3.3

and 3.6). However, we show that the pure showcase profit maximization and a natural relaxation

of the general showcase profit maximization problems admit fully polynomial time approximation

schemes (FPTAS); see Theorems 3.4 and 3.7. Applying the ideas used to construct the FPTAS,

we propose an integer-programming-based (IP-based) heuristic to determine the profit maximizing

subset. Using a simulation study, we show that (a) our IP-based heuristic scales to large, practical-

sized problems and (b) the solutions obtained from the IP-based heuristic provide significantly

higher profits and sales when compared to the solutions from the standard revenue-ordered or

greedy heuristics, described in detail in Section 4.

Finally, to illustrate the value of our methodology, we analyzed customer preference data on

messenger bags. The data were obtained from a conjoint study and demonstrate that customers’

valuations of many attributes change significantly when they evaluate the products in an o✏ine,

as opposed to an online channel (see Table 3). Using the proposed IP-based heuristic, we then

computed optimal sales/revenue maximizing assortments for various sizes of the o✏ine segment.

We found that significant gains in sales/revenues (up to 40% in our study) are attained. We also

gain the following broad insight into the structure of the sales maximizing assortment when there

is a constraint on the size of the o↵ered assortment: it is optimal to o↵er a mix of ‘popular’

and ‘informative’ products; the popular products have high utilities and drive sales in the o✏ine

channel, whereas the informative products expose customers to under-valued attributes and drive

sales in the online channel.

Related work. This research builds on existing research in marketing, operations, and informa-

tion systems. We categorize relevant existing research into work pertaining to single and multiple

channel settings.

In the single channel setting, our work is related to work on product line optimization in mar-

keting and work on assortment optimization in operations. Product line optimization focuses on

a canonical manufacturer selecting a set of products to o↵er with the objective of maximizing

consumer welfare, market share, or profit. In this literature, products are represented in a multi-

attribute space, and the optimal product line is constructed directly from the attribute levels, given

their corresponding partworths. The problem has been shown to be NP-hard to solve in general.

Several researchers have proposed heuristic solutions: Kohli and Krishnamurti (1987, 1989) and

Kohli and Sukumar (1990) proposed DP-based greedy heuristics; Balakrishnan and Jacob (1996)

and Fruchter et al. (2006) proposed genetic-algorithm-based heuristics; Dobson and Kalish (1988,

Page 5: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online ChannelArticle submitted to Management Science; manuscript no. 5

1993) and Green and Krieger (1985) proposed a priori selecting some candidate products from a

large number of feasible products and then selecting a product line from only this candidate set.

Belloni et al. (2008) presented a comparison of the performance of di↵erent heuristics for product

line design and found that greedy heuristics perform well.

The work on assortment optimization in operations management (OM) focuses on a canonical

retailer selecting a subset of products to o↵er to maximize expected profits (see Kok and Fisher

(2007) for a review). Because the viewpoint is that of a retailer, the problem is typically solved

in product space rather than in attribute space. The focus of this body of work has been on

deriving either exact or approximate optimization algorithms under various choice models: multi-

nomial logit (Talluri and Van Ryzin 2004, Rusmevichientong et al. 2010, Davis et al. 2013), nested

logit (Gallego and Topaloglu 2014, Alptekinoglu and Grasas 2014, Davis et al. 2014, Feldman and

Topaloglu 2014), d-level nested logit (Li et al. 2015), mixed logit (Rusmevichientong et al. 2014),

and the locational choice model (Gaur and Honhon 2006, Alptekinoglu and Corbett 2010, Ulu

et al. 2012). Jagabathula and Rusmevichientong (2015) jointly optimize assortment and prices,

Ghoniem and Maddah (2015) jointly optimize assortment, prices, and inventories, and Ghoniem

et al. (2013) focus on the problem of jointly optimizing the assortments and prices of a firm selling

products belonging to complementary categories.

We contribute to both of the above bodies of work by focusing on two channels rather than a

single channel. Unlike most work in OM, we focus on the attribute space rather than the product

space. Similarly, unlike most work in marketing, we focus on optimization and computational issues.

In the multiple channel setting, our work is related to the literature in marketing and information

systems that focuses on the interaction between online and o✏ine channels. Most of this work has

focused on the setting in which a customer has the option to buy the exact same product online and

o✏ine and on determining firms’ pricing decisions and equilibrium price outcomes (Brynjolfsson

et al. 2009, Forman et al. 2009, Mehra et al. 2013). We contribute to this literature by focusing on

multi-attribute products that are sold by a single firm and studying assortment rather than pricing

decision.

2. Model

This section provides the precise details of the decision problem and the corresponding consumer

choice model we consider. Our objective is to solve a firm’s showcase decision of determining

the assortment of products to ‘showcase’ in the o✏ine channel in order to maximize the profit

from both o✏ine and online channels. We focus on a firm selling products through an online and

an o✏ine channel. The products are reasonably expensive, infrequently purchased, and require a

high-involvement buying decision; consequently, customers can benefit from an o✏ine store visit.

Page 6: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel6 Article submitted to Management Science; manuscript no.

Furniture, messenger bags, apparel, etc. are good examples. The products are close substitutes but

are di↵erentiated along K pre-specified attributes. We assume that attribute k takes values from

the set Lk := {0,1, . . . ,Lk � 1} of Lk discrete levels. We let X denote the set of all feasible products,

with each product represented by a length K feature vector x2L1

⇥L2

⇥ · · ·⇥LK , where xk 2Lk

denotes the level of attribute k in the product. For example, suppose products are described by

two attributes, Size and Color, with Size being either Small or Large and Color being either Blue

or Black. Then, K = 2 and L1

= L2

= 2, with the vectors (0,0), (0,1), (1,0), and (1,1) denoting

Small-Blue, Small-Black, Large-Blue, and Large-Black products, respectively.

Omnichannel selling. The objective of the firm is to determine the selection of products from

the online channel to showcase in the o✏ine store in order to maximize the total profit across

both the online and the o✏ine channels. We distinguish two settings: the pure showcase and the

general showcase. In a pure showcase setting, the firm does not carry inventory in o✏ine stores and

sells products only through the online channel; examples include: bulky or customized purchases

such as furniture, kitchen cabinetry, or apparel from firms that sell online but showcase their

products in o✏ine showrooms (such as Bonobos). In the general showcase setting, by contrast, the

firm carries inventory and sells through both the channels; examples include most firms that sell

products through multiple channels. Mathematically, the pure showcase decision is a special case

of the general showcase decision. However, we study them separately because the (simpler) pure

showcase problem is more tractable, while encompassing a wide range of practically important

applications.

Customer choice model. The showcase decision a↵ects the purchase behavior of only the cus-

tomers who visit the o✏ine store. Among these customers, we distinguish two types: the o✏ine-type

and the online-type. The o✏ine-type choose from only the selection of products o↵ered o✏ine,

whereas the online-type choose from the entire o↵ered selection (online or o✏ine) and are willing

to purchase from either channel. Note that both types may purchase from either channel, depend-

ing on their channel purchase preferences and their preferred products. We do not distinguish

purchases from di↵erent channels and focus on maximizing the combined profit/sales from both

channels. We let ↵ and 1� ↵, for some ↵ 2 [0,1], denote the sizes of the o✏ine and online seg-

ments, respectively. For the pure showcase setting described above, all the customers are required

to choose from the selection of products online; therefore, the customers are composed entirely

of the online-type (↵= 0). The general showcase setting corresponds to ↵ 2 (0,1]. The case with

↵ = 1, in which none of the customers purchase from the online channel, is similar to the single

channel assortment optimization that has been studied in the operations management literature.

Page 7: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online ChannelArticle submitted to Management Science; manuscript no. 7

However, as discussed in greater detail below, existing results don’t apply because of the presence

of product features.

To model the purchase behavior of customers, we start with the standard multi-attribute utility

model; see Green and Rao (1971), Green and Srinivasan (1990). The customer’s utility for product

x2X is equal to the sum of the utility partworths of the features present in the product: U(x) =PK

k=1

uk(xk)+�price

·⇡x

+ "x

, where ⇡x

is the product price, �price

is the price coe�cient, uk(xk) =P

`2Lk

wk` · 1l`[xk] is the utility obtained from attribute k, wk` is the utility partworth assigned to

level ` of attribute k, and 1l`[xk] is the indicator variable taking value 1 if xk = ` and 0 otherwise.

The term "x

is the error term that captures any unexplained variance. Customers are utility

maximizing, so they purchase the product with the maximum utility from any choice set.

We extend the standard model to capture how exposure to a (subset) of products in the o✏ine

channel impacts the purchase behavior of customers. To model this impact, we suppose that cus-

tomers associate di↵erent utility partworths with each feature, depending on whether they were

exposed to the feature in the o✏ine channel or not. Particularly, with respect to the feature cor-

responding to attribute k and level `, the customer associates utility partworth wo↵

k` if she was

exposed to feature (k, `) in the o✏ine store and won

k` if she was not exposed. The di↵erence between

the online and o✏ine partworths for an attribute-level may be interpreted as being caused by the

information gained by the customer from “touching and feeling” the particular attribute-level in

the o✏ine store. For example, a customer may think she likes the “large” size but change her mind

upon physical inspection. Then, for (k, `) representing the large size, we have that won

k` > 0 but

wo↵

k` < 0, indicating that after physical inspection, the customer incorporates the information that

she dislikes the large size in her purchases, be they online or o✏ine.

To be precise, we let S = (S1

, S2

, . . . , SK) denote the collection of attribute levels that are o↵ered

in the o✏ine store, with Sk ✓ Lk denoting the subset of levels for attribute k that are o↵ered.

Because the firm selects the o✏ine assortment from the online assortment, we suppose that the

firm o↵ers the universe X of feasible products online. Now, consider a customer who has visited

the o✏ine store. For any product x2X , the customer assigns the utility

US

(x) =

(

PKk=1

uo↵

k (xk)+�price

·⇡x

+ "x

, if x o↵ered o✏ine,PK

k=1

uon

k (xk, Sk)+�price

·⇡x

+ "x

, otherwise,(1)

where

uon

k (xk, Sk) =X

`2Sk

wo↵

k` · 1l`[xk] +X

`/2Sk

won

k` · 1l`[xk] and uo↵

k (xk) =X

`2Lk

wo↵

k` · 1l`[xk]

We use the notation uo↵

k (xk) to emphasize that the utility of a product, when the customer is

evaluating o✏ine, is independent of what else is on o↵er. On the other hand, the utility partworth

Page 8: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel8 Article submitted to Management Science; manuscript no.

uon

k (xk, Sk) used for a product that is o↵ered online, but not o✏ine, depends on whether xk is o↵ered

o✏ine (as part of Sk) or not. Our notation is consistent: indeed, uon

k (xk, Sk) = uo↵

k (xk) whenever xk

is o↵ered o✏ine, i.e., xk 2 Sk. We note that according to our model assumptions, consumers use

the (same) o✏ine price coe�cient both online and o✏ine. This is because we are focusing only on

consumers who visit the o✏ine channel, where they are exposed to the price attribute.

A key aspect of our model is that physical exposure to product A a↵ects the utility of product

B if B shares attributes with A, even in the absence of physical exposure to B. For example,

exposure to a large-black bag a↵ects the utility of a large-blue bag, even if the blue bag was not

physically evaluated. Because the utilities and, hence, the purchase probabilities of products in

the online channel are a↵ected by features the customer was exposed to in the o✏ine channel, the

profit and revenue from both channels are a↵ected by the o✏ine assortment; this interaction makes

the assortment problem challenging.

We assume that the price of a product can be decomposed into the sum of the prices of its

constituent attributes. Letting ⇡k` � 0 denote the price associated with level ` of attribute k, the

price ⇡x

of product x is equal toPK

k=1

PLk

`=1

⇡k` · 1l`[xk]. This price structure arises when the firm

adopts a “hedonic price model,” expressing product price in terms of included attributes, to obtain

a simple pricing scheme for an exponentially large configurable product space. It is commonly used

in literature (Cohen et al. 2016, Randall et al. 1998, Rodrıguez and Aydın 2011) and practice

as “optional product pricing” or “feature-based pricing” (e.g., $10 extra for the black color, $20

extra for the large size, etc.) for pricing configurable products such as computers, furniture, cars,

etc. With this assumption, the utility expressions above simplify as follows. Letting wc

k` denote

wc

k` +�price

⇡k`, for c2 {on,o↵}, and defining

uon

k (xk, Sk) =X

`2Sk

wo↵

k` · 1l`[xk] +X

`/2Sk

won

k` · 1l`[xk] and uo↵

k (xk) =X

`2Lk

wo↵

k` · 1l`[xk],

we obtain that US(x) =PK

k=1

uo↵

k (x)+ "x

if x is o↵ered o✏ine andPK

k=1

uon

k (x)+ "x

, otherwise.

We make the standard logit assumption that the idiosyncratic terms "x

are i.i.d. standard Gumbel

distributed for x 2 X . Further, suppose that M ✓ X is the assortment o↵ered o✏ine and the

universe X is o↵ered online. Then, the probability that an online-type customer chooses product

x from the selection X is given by

Px

(M) =exp

PKk=1

uon

k (xk, SMk )

1+P

y2X exp⇣

PKk=1

uon

k (yk, SMk )

⌘ , (2)

where SM = (SM1

, SM2

, . . . , SMK ) denotes the set of features that the products in M are composed of,

i.e., `2 SMk if and only if xk = ` for some x2M . Note that we have made the standard assumption

Page 9: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online ChannelArticle submitted to Management Science; manuscript no. 9

that the mean utility of the no-purchase option is 0. On the other hand, the probability that an

o✏ine-type customer purchases product x2M is given by

Qx

(M) =exp

PKk=1

uo↵

k (xk)⌘

1+P

y2M exp⇣

PKk=1

uo↵

k (yk)⌘

We make the following remarks. The choice probability Qx

(M) is similar to classical choice

probability expression under the multinomial logit (MNL) choice model in which the customer is

restricted to choose from the subset M . The choice probability expression Px

(M) for the online-

segment di↵ers from Qx

(M) in two key ways: (a) the utility partworths depend on whether a

customer is exposed to a feature o✏ine or not, and (b) the customer chooses from the entire

collection X of products o↵ered online and o✏ine. Due to these distinctions of choice probability

expressions, the optimization problems we consider are di↵erent in structure from the classical

assortment optimization problems studied in the literature. Finally, for the above expressions for

choice probabilities to be valid, we only require that the union of the sets of products o↵ered online

and o✏ine be equal to X , and not necessarily that the online assortment be equal to X . However,

because online assortments tend to be larger than o✏ine assortments, we make the assumption

that the online assortment is equal to X .

Firm’s showcase decision. In the context of the above model, we consider the following decision

problems:

maxM✓X

X

x2X

Px

(M), (Pure Showcase Sales Max)

maxM✓X

X

x2X

px

Px

(M), (Pure Showcase Profit Max)

and

maxM✓X

↵X

x2M

px

Qx

(M)+ (1�↵)X

x2X

px

Px

(M), (General Showcase Profit Max)

where px

is the net profit obtained from the sale of product x. The Pure Showcase Sales

Max

2 is the simplest non-trivial decision problem we consider. When the firm is selling through a

single channel, it is always optimal (for maximizing sales) to o↵er all the products when there is

no capacity constraint. As we show below, this simple structure no longer holds in the presence of

two channels. For profit maximization problems, we assume that the profit from a product can be

decomposed into the sum of the profits from its constituent attributes. Letting rk` � 0 denote the

profit margin associated with level ` of attribute k, the net profit px

obtained from selling product

2 This problem maximizes the firm’s market share. However, for brevity we use the term “sales” interchangeably for“market shares.”

Page 10: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel10 Article submitted to Management Science; manuscript no.

x is equal toPK

k=1

PLk

`=1

rk` ·1l`[xk]. Because product prices are assumed to decompose into sums of

attribute prices, the profits also decompose if we assume that the product cost decomposes into the

sum of individual attribute costs. Such cost structure is commonly assumed in literature (Belloni

et al. 2008) and may be backed out from the product’s “bill of materials.” We show that the profit

maximization decision problems are NP-hard to solve in general but admit fully polynomial time

approximation schemes (FPTAS).

3. The product showcase decision

We now discuss our results for solving the showcase decision problems introduced above. For the

development below, we make the common assumption (e.g., see Kohli and Krishnamurti (1989))

that the product universe X is full-factorial, i.e., X = X1

⇥X2

⇥ · · ·⇥XK so that every non-price

feature combination is feasible. With the full-factorial assumption, the Pure Showcase Sales

Max problem becomes tractable, but as we show below both Pure Showcase Profit Max and

General Showcase Profit Max remain NP-complete. In fact, even the sales maximization

problem (with px

set to 1 for all x in General Showcase Profit Max) under the general

showcase setting is NP-complete. However, there may be cases in which the full-factorial assumption

is not reasonable. The optimization problems become significantly harder if we allow for arbitrary

constraints on feature combinations, even in the single channel setting.

3.1. Pure showcase decision

In the pure showcase setting, all customers who visit the o✏ine store are assumed to also visit the

online store before making a purchase. Because all customers choose from the same selection X of

products, only the o↵ered attributes, and not the o↵ered products, impact the choice probabilities

and, consequently, the profits. Therefore, the choice probability expression in (2) may be simplified

to

Px

(S) =exp

PKk=1

uon

k (xk, Sk)⌘

1+P

y2X exp⇣

PKk=1

uon

k (yk, Sk)⌘ , (3)

where as above, S = (S1

, . . . , SK) with Sk 2 Lk denoting the subset of levels for attribute k that

are o↵ered in the o✏ine store and we use the notation Px

(S), instead of the more general Px

(M).

The decision problem now reduces to determining the optimal subset of attribute levels to o↵er

in the store. Once the optimal vector of attribute-level sets S is determined, the assortment M

of products to o↵er in the store is given by the cartesian product S1

⇥ S2

⇥ · · ·⇥ SK . Of course,

multiple assortments3 achieve S, and the eventual decision may be driven by other considerations4

not modeled in our work.

3 For instance, it can be shown that the minimum cardinality assortment M that achieves S is of size maxK

k=1

|Sk

|.4 For instance, cardinality or capacity constraints, variety requirements, etc.

Page 11: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online ChannelArticle submitted to Management Science; manuscript no. 11

With the assumption that the profit from product px

from product x decomposes as px

=PK

k=1

PLk

`=1

rk` · 1l`[xk], the expected profit from o↵ering the attribute-levels S in the o✏ine store

equals

Rpure(S) =X

x2X

px

Px

(S) =

P

x2X

P

k,` rk`1l`[xk]⌘

exp⇣

PKk=1

uon

k (xk, Sk)⌘

1+P

y2X exp⇣

PKk=1

uon

k (yk, Sk)⌘ .

The denominator of the choice probability can be simplified by noting that

X

y2X

exp

KX

k=1

uon

k (yk, Sk)

!

=X

y2X

KY

k=1

0

@

X

`2Sk

1l`[yk]ewo↵

k` +X

`/2Sk

1l`[yk]ewon

k`

1

A=KY

k=1

0

@

X

`2Sk

ewo↵

k` +X

`/2Sk

ewon

k`

1

A ,

where the last equality follows from interchanging the sum and the product operators.

Using a similar simplification of the numerator, we obtain the following expression for the

expected pure showcase profit:

Lemma 3.1 (Simplified pure showcase profit). Suppose we o↵er the collection of attribute

levels represented by S = (S1

, . . . , SK) in the o✏ine store. Then, the pure showcase expected profit

function can be simplified as

Rpure(S) =

PKk=1

Rk(Sk)

1+ 1/D(S), where

Rk(Sk) =bk +

P

`2Sk

rk`�k`

Dk(Sk),Dk(Sk) = dk +

X

`2Sk

�k`, and D(S) =KY

k=1

Dk(Sk),

(4)

with bk :=P

`2Lk

rk`ewon

k` , dk :=P

`2Lk

ewon

k` , and �k` := ewo↵

k` � ewon

k` .

The lemma is proved in Appendix A.1.

Sales maximization. The expected sales function is obtained by setting px

= 1 for all x 2 Xin the expression for Rpure(S). Because p

x

=P

k,` rk`1l`[xk] andP

`2Lk

1l`[xk] = 1 for all k, setting

rk` = 1/K for all k, `, yields px

= 1 for all x 2 X . Setting rk` = 1/K for all k, ` in the expression

in (4) results in the expression (1+1/D(S))�1 for the expected sales from o↵ering S in the o✏ine

store. The sales maximization problem now reduces to

maxS22

L1⇥···⇥2

LK

D(S) = maxS22

L1⇥···⇥2

LK

KY

k=1

Dk(Sk) =KY

k=1

0

@ maxSk

22

Lk

2

4dk +X

`2Sk

�k`

3

5

1

A ,

where 2S denotes the power set {S0 ✓ S : S0 6=?} for any set S and the last equality follows because

the optimization problem is separable in k.

It is immediately seen that an optimal solution S

⇤ = (S⇤1

, . . . , S⇤k) of the above optimization

problem is such that S⇤k = {`2Lk : �k` � 0} if {`2Lk : �k` � 0} 6= ? and S⇤

k = {`⇤k}, where `⇤k 2Lk achieves the maximum i.e., �k,`⇤

k

= max`2Lk

�k`. To simplify notation, let L+

k denote the set

Page 12: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel12 Article submitted to Management Science; manuscript no.

{`2Lk : �k` � 0} and L�k denote the set {`2Lk : �k` < 0}. Because �k` = ew

o↵

k` � ewon

k` , L+

k comprises

the set of under-valued attributes (for which wo↵

k` �won

k` ) and L�k comprises the set of over-valued

attributes.

Our argument above shows that the sales maximizing subset of attribute levels has the following

intuitive structure: include under-valued attribute levels and exclude over-valued attribute levels;

if for a particular attribute k, all the levels are over-valued, then o↵er the least over-valued level.

We summarize this result as the following theorem:

Theorem 3.1 (Pure showcase sales max solution). The optimal solution to the Pure

Showcase Sales Max problem is S

⇤ = (S⇤1

, . . . , S⇤k) such that

S⇤k =

(

{`2Lk : �k` � 0} , if L+

k 6=?{`⇤k} , otherwise.

where �k` = ewo↵

k` � ewon

k` , L+

k = {`2Lk : �k` � 0}, and `⇤k 2Lk is such that �k`⇤k

=max`2Lk

�k`.

Profit maximization. The structure of the profit-maximizing subset of attribute levels is more

complex. Exploiting the profit expression in (4), we establish the following result:

Theorem 3.2 (Pure showcase profit max solution structure). Any optimal solution

S

⇤ = (S⇤1

, . . . , S⇤k) to the Pure Showcase Profit Max problem must satisfy

`2L+

k : rk` > t⇤k

✓ S⇤+k ✓

`2L+

k : rk` � t⇤k

`2L+

k : rk` < t⇤k

✓ S⇤�k ✓

`2L+

k : rk` t⇤k

where t⇤k :=Rk(S⇤k)�R(S⇤)/D(S⇤)

where S+

k denotes Sk \L+

k and S�k denotes Sk \L�

k for any subset Sk ✓Lk.

To understand this result, suppose that the optimal solution to Pure Showcase Profit Max is

unique. Then, Theorem 3.2 establishes that for each attribute k, the optimal set of levels consists

of a profit-ordered (PO) subset of under-valued attribute levels and a reverse profit-ordered (RPO)

subset of over-valued attribute levels. We call a subset of levels a PO subset if it consists of the

top-m most profitable levels for some m and an RPO subset if it consists of the bottom-m least

profitable levels for some m. Because o↵ering under-valued levels increases their attractiveness and

o↵ering over-valued levels decreases their attractiveness, our result provides the following intuitive

suggestion: increase the attractiveness of the most profitable levels and decrease the attractiveness

of the least profitable levels.

It is instructive to contrast the result of Theorem 3.2 with that of the classical single-channel

setting, but when the universe X = L1

⇥ · · · ⇥ LK , consisting of all possible feature combina-

tions. It is known that the profit maximizing assortment M⇤ satisfies M⇤ = {x2X : px

�Z⇤}, for

some Z⇤ (Talluri and Van Ryzin 2004). To understand how this result di↵ers from the result of

Page 13: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online ChannelArticle submitted to Management Science; manuscript no. 13

Theorem 3.2, consider the following example. Suppose a firm is selling horizontally di↵erentiated

products, concretely, shirts that di↵er only in color (so that K = 1). There are two types of colors:

“base” colors such as black, blue, etc., that customers are familiar with and “fashion” colors such

as orange, pink, etc., that are newly introduced. Suppose that base colors are over-valued and

fashion colors are under-valued by the customers. Then, in the absence of an online channel, it is

optimal for the firm to o↵er the most profitable base and fashion colors. In contrast, in the presence

of an online channel, it is optimal for the firm to o↵er the most profitable fashion colors and the

least profitable base colors. By doing this, the firm is providing information to the customer that

the profitable fashion colors are being under-valued and the least profitable base colors are being

over-valued, shifting the demand to more profitable products. This distinction makes it clear that

(a) in the pure showcase setting, the o✏ine channel acts as an “information” channel, as opposed to

the single channel setting in which the o✏ine channel acts as a sales channel; and (b) an algorithm

that can find the best subset for the single channel case will necessarily be sub-optimal for the

pure showcase problem in general.

A consequence of Theorem 3.2 is that Pure Showcase Profit Max reduces to

maxS2Z

1

⇥···⇥ZK

PKk=1

Rk(Sk)

1+ exp⇣

�PK

k=1

logDk(Sk)⌘ , (5)

where Zk is the collection of all possible subsets S of Lk such that S+ is a PO subset and S� is an

RPO subset:

Zk =�

S ✓Lk : S+ =

`2L+

k : �k` � t+

, S� =�

`2L�k : �k` < t�

for some t+, t�

.

Unlike for the case of sales maximization, however, Theorem 3.2 does not yield an e�cient

algorithm to determine the optimal solution. In particular, for each k, we must search over all

possible combinations of PO subsets of L+

k and RPO subsets of L�k . Because there is a total of at

most |Lk|2 such combinations for each k, in the worst case, brute force search requires searching

over O⇣

QKk=1

|Lk|2⌘

, which scales exponentially in K. In fact, solving Pure Showcase Profit

Max is NP-hard:

Theorem 3.3 (Hardness of Pure Showcase Profit Max). The following decision prob-

lem is NP-complete: for any Q� 0, is there a subset S = (S1

, . . . , SK) such that Rpure(S)�Q?

The theorem is proved in Appendix A.1. The reduction is from the NP-complete partition prob-

lem (Garey and Johnson 1979). The proof focuses on the special case when each attribute k has

only two levels and one of the levels, say the first one, has zero discrepancy: wo↵

k1 =won

k1. The deci-

sion problem then reduces to whether to o↵er the second level in each attribute or not. A brute

Page 14: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel14 Article submitted to Management Science; manuscript no.

forces search has O(2K) complexity. We then obtain a reduction from the partition problem to this

special case.

Despite the fact that the problem is NP-hard in the worst case, we show below that the opti-

mization problem admits a fully polynomial time approximation scheme (FPTAS). An algorithm

is formally defined to be an "-approximation algorithm of a maximization problem if for each prob-

lem instance and tolerance parameter 0< " 1, the algorithm produces a solution with objective

value R such that R⇤ �R� (1� ")R⇤, where R⇤ is the objective value of an optimal solution. An

"-approximation algorithm is called an FPTAS if for any fixed ", the running time of the algorithm

is bounded above by a polynomial in the size of the input and 1/".

In constructing the FPTAS, we use the ideas developed for the construction of an FPTAS for the

classical knapsack problem (Lawler 1979). These ideas have been used in the existing literature to

construct either an FPTAS or a polynomial time approximation scheme5 (PTAS) for assortment

optimization problems (see Rusmevichientong et al. (2009), Desir and Goyal (2014)). However,

whereas the existing body of work has considered objective functions that can be expressed as the

sum of ratios of functions that are linear in the decision variables, our setting results in objective

functions that are sums of ratios of functions that are non-linear in the decision variables. As a

result, constructing an FPTAS requires a treatment di↵erent from the existing work, as presented

below.

We use the following general procedure to solve (5). We guess the values of the numerator and the

denominator of the objective function at the optimal solution and find a solution S that approxi-

mately achieves the guessed values. We show below that for the given values of the numerator and

denominator of the objective function, the solution that achieves them approximately can be found

by solving a dynamic program (DP) in time that is polynomial in the input size and 1/". Because

we do not know the optimal numerator and denominator values, we search over an "-grid of the

region of possible values. We show that the number of possible grid points we need to search over

is polynomial in the input size and 1/". Putting everything together results in the desired FPTAS.

The most challenging step of the above procedure is to find the solution S = (S1

, . . . , Sk) that

approximately achieves the guessed values of the numerator and denominator. In particular, let q

denote our guess of the optimal value of the numerator and t denote our guess ofP

k logDk(Sk)

at the optimal solution. Our goal is to find an S such that

X

k

Rk(Sk)� q andX

k

logDk(Sk)� t. (6)

5 Unlike an FPTAS, the computational complexity of the algorithm may scale exponentially in 1/" in a PTAS.

Page 15: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online ChannelArticle submitted to Management Science; manuscript no. 15

We solve the above problem approximately. In particular, for a given "> 0, we find S (if it exists)

such thatX

k

Rk(Sk)� q andX

k

logDk(Sk)� (1� 2")t. (7)

To find S we use the following DP formulation. We discretize logDk(Sk)s as follows. Let jS,k :=

blogDk(S)/("t/K)c for any S 2Zk, where bxc is the floor function defined as the integer such that

bxc x bxc+1, for any x2R. Further, let ⇢ := bK/"c�K. Now define the DP value function:

V (k,!) = maxS2Z

1

⇥···⇥Zk

kX

k0=1

Rk0(Sk0) subject toX

1k0k

jS,k0 � !.

Our goal is to compute V (K,⇢), for which we use the following DP recursion:

V (k,!) =

8

>

<

>

:

0, if k= 0,! 0

�1, if k= 0,!> 0

maxS2Zk

[Rk(S)+V (k� 1,!� jS,k)] , otherwise.

(8)

We carry out the above DP for integers k and ! such that 0 kK and 0 ! ⇢. Each iteration

requires a search over O(|Zk|) =O(|Lk|2) elements. Therefore, the running time of the DP scales

as O(K⇢maxk |Zk|) =O(K2maxk |Lk|2 /") because ⇢=O(K/"). We first show that the DP above

indeed obtains the desired approximation S.

Lemma 3.2 (Pure showcase DP approximation). For given t and q, if there exists an S

such thatX

k

Rk(Sk)� q andX

k

logDk(Sk)� t,

then the DP (8) terminates with S such that

X

k

Rk(Sk)� q andX

k

logDk(Sk)� (1� 2")t.

The lemma is proved in Appendix A.1. The precise algorithm is summarized below.

Algorithm 1: FPTAS for Pure Showcase Profit Max

Input Tolerance parameter "> 0 and problem inputs Rk

(S) and Dk

(S) for all S 2Zk

and k= 1, . . . ,K.

Step 1 Define ⌧min

:= e�P

k logDk(L+

k ) and ⌧max

:= e�P

k logDk(L�k ). Create the " grid T of the interval [⌧

min

, ⌧max

]

such that T :=�

⌧min

(1+ ")i : i= 0,1, . . . , I

with I = log(⌧max

/⌧min

)/ log(1+ ").

Step 2 For ⌧ 2 T do

define t=� log ⌧ :

determine S⌧

by solving the DP with inputs t and "/ |log ⌧min

|.Output Subset S that maximizes the profit from the collection

n

S⌧

: ⌧ 2 To

.

We now show that the above scheme produces an " approximation of the optimal solution with

computational complexity that is polynomial in 1/" and K.

Page 16: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel16 Article submitted to Management Science; manuscript no.

Theorem 3.4 (FPTAS for pure showcase profit maximization). Algorithm 1 produces a

1� 8" optimal solution in O⇣

K2maxk |Lk|2 |log ⌧min

| log(⌧max

/⌧min

)/(" log(1+ "))⌘

running time,

where ⌧min

:= e�P

k

logDk

(L+

k

) and ⌧max

:= e�P

k

logDk

(L�k

).

3.2. General showcase decision

We now consider the more general setting in which the population comprises both online- and

o✏ine-type customers. Recall that both segments visit the o✏ine store. The o✏ine segment of

customers chooses from the assortment o↵ered o✏ine, whereas the online segment chooses from

the assortments o↵ered both online and o✏ine. Let ↵ and 1�↵ denote the sizes of the o✏ine and

online segments of customers, respectively, for some ↵ 2 (0,1). Our goal then is to find the profit

maximizing assortment of products to carry in the store, i.e., to solve the General Showcase

Profit Max problem, restated here:

maxM✓X

Rgeneral(M) = ↵X

x2M

px

Qx

(M)+ (1�↵)X

x2X

px

Px

(M).

Sales maximization. The sales maximization problem is obtained by setting px

= 1 for all x2X

in the optimization problem above. The sales maximizing assortment has the following structure:

Theorem 3.5 (General showcase sales max solution structure). The optimal solution

M⇤ to the following optimization problem maxM✓X ↵P

x2M Qx

(M) + (1� ↵)P

x2X Px

(M) is of

the form M⇤ = S⇤1

⇥ · · ·S⇤K with L+

k ✓ Sk for all k.

We note the following implications of Theorem 3.5. First, the sales maximizing assortment can be

obtained by searching through the attribute space, as opposed to the space of subsets of products.

Second, it is optimal to o↵er all the under-valued attribute levels. This is because o↵ering an

under-valued attribute level in the o✏ine store increases its attractiveness, resulting in higher sales

from both online and o✏ine customers.

On the other hand, o↵ering an over-valued attribute decreases the attractiveness of the attribute-

level and, hence, decreases sales from the online-type customers but increases the o↵ered selection

and, hence, increases sales from the o✏ine-type. Therefore, the optimal o↵ering of over-valued

attribute levels should balance sales from both customer types. By contrast, if the market consists

of only the online-type customers (reducing the problem to pure showcase sales maximization),

then the o✏ine channel becomes a pure “information” channel (as above) and it is optimal to

“hide” all the over-valued attribute levels. If, on the other hand, the market consists of only the

o✏ine-type customers (reducing the problem to the classic single-channel sales optimization), then

the o✏ine channel becomes a pure “sales” channel (by restricting the selection of products) and it is

optimal to o↵er all the over-valued attribute-levels. When there is a mix of o✏ine- and online-type

Page 17: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online ChannelArticle submitted to Management Science; manuscript no. 17

customers, the o✏ine channel becomes both an information and a sales channel, and the over-valued

attribute-levels should be chosen to balance providing (or hiding) information and driving sales

from the o✏ine channel.

Unfortunately, determining the optimal o↵ering of over-valued attribute levels is computationally

challenging. Specifically, unlike the pure showcase setting, we show that even sales maximization

is NP-hard to solve:

Theorem 3.6 (Hardness of General Showcase Profit Max). The following decision

problem is NP-complete: for any Q� 0, is there a subset M ✓X such that

↵X

x2M

Qx

(M)+ (1�↵)X

x2X

Px

(M)�Q?

The theorem is proved in Appendix A.2. The reduction is from the partition problem.

Profit maximization. The profit maximization problem is more challenging. Because the o✏ine

segment of customers is impacted by particular product configurations, as opposed to only the

attributes, the problem cannot be solved in the attribute space. In fact, even the ↵= 1 case is in

general hard to solve. When ↵= 1, the problem reduces to finding the profit maximizing assortment

for a single channel under the MNL model. It is known that the single-channel profit maximizing

assortment is one of the profit-ordered assortments (Talluri and Van Ryzin 2004, Rusmevichientong

et al. 2010), but because the universe of products is exponentially (in K) large, finding the best

subset is computationally challenging, except when the optimal subset has polynomial (in K)

size (Gallego et al. 2016).

A key source of hardness with the General Showcase Profit Max is that we need to

search over subsets M ✓ X of products as opposed to subsets of attribute-levels; indeed, for a

given subset S of attribute-levels, we obtain di↵erent profits from di↵erent sets M that achieve S.

Therefore, to make the problem tractable, we restrict the search to the collection of assortments

{S1

⇥ · · ·⇥SK : Sk ✓Lk for 1 kK}. This restriction6 can be justified by noting that customers

tend to extrapolate attribute combinations based on what they have been exposed to. For instance,

if an o✏ine-type customer is exposed to a large, red bag and a small, blue bag, then she may infer

the availability of a large, blue bag and choose from the subset S1

⇥ · · ·⇥Sk when the the attribute

levels (S1

, . . . , SK) are o↵ered in the store, even though not all products in S1

⇥ · · ·⇥Sk are o↵ered.

With the above restriction, we arrive at an optimization problem that can be solved in the

attribute space and admits an FPTAS. In particular, the profit function can now be simplified as

follows:

6 For instance, the assortment {(0,1), (1,0)} cannot be expressed as the cartesian product S1

⇥S2

for any two subsetsS1

✓L1

= {0,1} and S2

✓L2

= {0,1}.

Page 18: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel18 Article submitted to Management Science; manuscript no.

Lemma 3.3 (Restricted showcase profit).

Rgeneral(S) =X

c2{on,o↵}

↵c

PKk=1

Rc

k(Sk)

1+ 1/Dc(S), where

Rc

k(Sk) =bck +

P

`2Sk

rk`�c

k`

Dc

k(Sk),Dc

k(Sk) = dck +X

`2Sk

�ck`, and Dc(S) =KY

k=1

Dc

k(Sk),

(9)

with bonk :=P

`2Lk

rk`ewon

k` , donk :=P

`2Lk

ewon

k` , and �onk` := ewo↵

k` � ewon

k` for all k. Further, bo↵k , do↵k = 0

and �o↵k` := ewo↵

k` for all k.

The lemma is proved in Appendix A.2. The decision problem now reduces to

maxS22

L1⇥···⇥2

LK

X

c2{on,o↵}

↵c

P

kRc

k(Sk)

1+ exp⇣

�PK

k=1

logDc

k(Sk)⌘ . (Restricted Showcase Profit Max)

Because sales maximization is a special case of Restricted Showcase Profit Max, it follows

from Theorem 3.6 that the above problem is NP-hard. Therefore, we extend the ideas from the

pure showcase setting to obtain an FPTAS. As above, we first consider the following maximization

problem:

maxS22

L1⇥···⇥2

LK

X

c2{on,o↵}

↵c

P

kRc

k(Sk)

1+ e�tc

s.t.X

k

logDc

k(Sk)� tc

, c2 {on,o↵} .

We solve the above optimization problem by approximately satisfying the constraints through a

DP formulation. For any " > 0, define jcS,k := blogDc

k(S)/("tc/K)c for c 2 {on,o↵}. Further, let

⇢ := bK/"c�K. Now define the DP value function:

V (k,!on

,!o↵

) = maxS22

L1⇥···⇥2

LK

X

c2{on,o↵}

↵c

P

1k0kRc

k0(Sk)

1+ e�tc

s.t.X

1k0k

jcSk

,k � !c

, c2 {on,o↵} .

Our goal is to compute V (K,⇢,⇢), for which we use the following DP recursion:

V (k,!on

,!o↵

)

=

8

>

<

>

:

0, if k= 0,!on

,!o↵

0,

�1, if k= 0,!on

> 0 or !o↵

> 0,

maxS22

Lk

P

c

⌫c,k(S)+V (k� 1,!

on

� jonS,k,!o↵

� jo↵S,k)⇤

, otherwise ,

where we used ⌫c,k(S) := ↵cRc

k(S)/(1+ e�tc) for compactness of notation. We carry out the above

DP for integers k and !c

such that 0 k K and 0 !c

⇢ for c 2 {on,o↵}. Each iteration

requires a search over O(|2Lk |) = O(2|Lk

|) elements. Therefore, the running time of the DP is

O(K⇢2maxk 2|Lk

|) = O(K3maxk 2|Lk

|/"2) because ⇢ = O(K/"). We first show that the DP above

solves the desired optimization problem approximately.

Page 19: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online ChannelArticle submitted to Management Science; manuscript no. 19

Lemma 3.4 (Restricted showcase DP approximation). For given ton

, to↵

, and q, if there

exists a S such that

X

c2{on,o↵}

↵c

P

kRc

k(Sk)

1+ e�tc

� q andKX

k=1

logDc

k(Sk)� tc

, c2 {on,o↵} ,

then the DP described above terminates with solution S such that

X

c2{on,o↵}

↵c

P

kRc

k(Sk)

1+ e�tc

� q andKX

k=1

logDc

k(Sk)� (1� 2")tc

, c2 {on,o↵} .

The lemma is proved in Appendix A.2. The precise algorithm is summarized below.

Algorithm 2: FPTAS for Restricted Showcase Profit Max

Input Tolerance parameter "> 0 and problem inputs Rc

k

(S) and Dc

k

(S) for all S 2 2Lk , 1 kK, and c2 {on,o↵}.Step 1 Define ⌧

on,min

:= e�P

k logD

on

k (L+

k ) and ⌧on,max

:= e�P

k logD

on

k (L�k ). Similarly, ⌧

o↵,min

:= e�P

k logD

off

k (Lk)

and ⌧on,max

:= e�P

k min`2Lklog �

off

k` . Create the " grid Tc

of the interval [⌧c,min

, ⌧c,max

] such that Tc

:=�

⌧c,min

(1+ ")j : j = 0,1, . . . , Jc

with Jc

= log(⌧c,max

/⌧c,min

)/ log(1+ "), for c2 {on,o↵}.Step 2 For ⌧ = (⌧

on

, ⌧o↵

)2 Ton

⇥ To↵

do

define tc

=� log ⌧c

, c2 {on,o↵}:determine S

by solving the DP with inputs ton

, to↵

, and "/ |log ⌧⇤|, where ⌧⇤ =min{⌧on,min

, ⌧o↵,min

}.Output Solution S that maximizes the expected profit from the collection

n

S⌧

: ⌧ 2 Ton

⇥ To↵

o

.

We now show that the algorithm above is indeed an FPTAS.

Theorem 3.7 (FPTAS for Restricted Showcase Profit Max).

Algorithm 2 produces a 1 � 8" optimal solution with a running time of

O�

K3maxk 2|Lk

| log2 ⌧ ⇤ log(⌧on,max

/⌧on,min

) log(⌧o↵,max

/⌧o↵,min

)/("2 log2(1+ "))�

, where ⌧on,min

:=

e�P

k

logDon

k

(L+

k

), ⌧on,max

:= e�P

k

logDon

k

(L�k

), ⌧o↵,min

:= e�P

k

logDo↵

k

(Lk

), ⌧on,max

:= e�P

k

min

`2Lk

log �o↵k` ,

and ⌧ ⇤ =min{⌧on,min

, ⌧o↵,min

}.

The theorem is proved in Appendix A.2.

3.3. Integer-programming-based heuristic

Building on the ideas in the construction of the FPTAS, we now propose an IP-based heuristic

to approximately find the profit maximizing assortment of products to o↵er in the store. Existing

work has used IP formulations to solve assortment problems (Bront et al. 2009, Subramanian and

Sherali 2010) in the product space when the objective function can be expressed as a ratio of

linear functions in decision variables. Instead, we solve the problem in the attribute space and the

objective function is a ratio of linear to non-linear function in the decision variables; hence, we use

the ideas from our FPTAS to convert the non-linear IP (NLIP) into a collection of mixed integer

linear programs (MILPs). We demonstrate the performance of the heuristic on synthetic data in

Section 4 and on real-world data in Section 5.

Page 20: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel20 Article submitted to Management Science; manuscript no.

Using the simplification of the profit function from Lemma 3.3, our objective is to solve the

following decision problem:

maxS22

L1⇥···2LK

X

c2{on,o↵}

↵c

PKk=1

Rc

k(Sk)

1+ exp(� logDc

k(Sk)). (10)

We first formulate this optimization problem as an NLIP. To do so, we encode subset S using

binary vectors zk, for 1 k K, defined as zk,S = 1 if and only if S = Sk and zk,S = 0 otherwise.

In other words, the binary vector zk is of length 2|Lk

| � 1, with each component associated with

a non-empty subset of Lk. It is clear from the definition that the encoding is a bijection from

{0,1}2|L

k

|�1 to 2Lk . With this binary encoding of S, we can formulate the optimization problem

in (10) as the following NLIP:

maxz

k

,1kK

X

c2{on,o↵}

↵c

PKk=1

P

S22

Lk

zk,SRc

k(S)

1+ exp⇣

�PK

k=1

P

S22

Lk

zk,S logDc

k(S)⌘

subject toX

S22

Lk

zk,S = 1, k= 1,2, . . . ,K,

zk,S 2 {0,1} , k= 1,2, . . . ,K;S 2 2Lk .

Because the above formulation has a non-linear objective function, we solve it by reducing

it to a collection of MILPs at various grid points. To obtain the reduction, we use the ideas

from the FPTAS described above. Let ⌧c,min

and ⌧c,max

denote the minimum and maximum val-

ues of exp⇣

�PK

k=1

P

S22

Lk

zk,S logDc

k(S)⌘

respectively, as zk varies over all the binary vectors in

{0,1}2|L

k

|�1 such thatP

S22

Lk

zk,S = 1 for all 1 k K and c 2 {on,o↵}. Then, for a given "> 0,

we consider the "-grid Ton

⇥To↵

where Tc

= {⌧c,min

(1+ ")j : j = 0,1, . . . , Jc

} with Jc

chosen such that

⌧c,min

(1+ ")Jc�1 ⌧c,max

⌧c,min

(1+ ")Jc . For each grid point ⌧ = (⌧on

, ⌧o↵

)2 Ton

⇥To↵

, we solve the

following MILP:

maxz

k

,1kK

X

c2{on,o↵}

(↵c/(1+ ⌧c

))KX

k=1

X

S22

Lk

zk,SRc

k(S)

subject toKX

k=1

X

S22

Lk

zk,S logDc

k(S)�� log ⌧c

, c2 {on,o↵}

X

S22

Lk

zk,S = 1, k= 1,2, . . . ,K,

zk,S 2 {0,1} , k= 1,2, . . . ,K;S 2 2Lk .

Let zk,⌧ , 1 k K, denote the optimal solution obtained from solving the above MILP for grid

point ⌧ . Let S⌧ ,k denote the subset of attribute levels such that zk,⌧ ,S = 1 for S = S

⌧ ,k and let

S

denote (S⌧ ,1, . . . , S⌧ ,K). We then output the subset S

⇤ from the collection {S⌧

: ⌧ 2 Ton

⇥ To↵

}

Page 21: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online ChannelArticle submitted to Management Science; manuscript no. 21

that maximizes the expected profit Rgeneral(S⌧

), computed from the expression in (9). Instead of

the value of ", we may also specify the number of grid points Jc

, in which case we can back out

the value of " as equal to (⌧c,max

/⌧c,min

)1/Jc � 1.

Note that for each grid point ⌧ , we can find the solution S

by solving the DP in Section 3.2

instead of solving the MILP above. Solving the MILP may be faster because of any structures that

the commercial IP solvers exploit.

4. Numerical study

We carried out two computational experiments using synthetic data. The first study is designed

to assess the optimality gap of the IP-based heuristic on smaller problem instances. On problem

instances with 4 attributes and 2 levels per attribute, the study shows that the IP-based heuristic

obtains profits within 6.45% to 0.11% of the optimal profit, on average, as the number of grid points

is increased from Jon

= Jo↵

= 2 to Jon

= Jo↵

= 32. In all these instances, the average running time

was < 0.5 seconds. The second study is aimed at demonstrating the practical e↵ectiveness of our IP-

based heuristic. The study achieves two objectives: it demonstrates that (a) the computational time

of the IP-based heuristic scales to large, practical-sized problems and (b) the solutions obtained

from the IP-based heuristic provide significantly higher profits and sales when compared to the

solutions from standard revenue-ordered (RO) and greedy heuristics. Our results demonstrate that,

on average, the IP-based heuristic runs in < 3 minutes for problem instances with 100 attributes

and 10 levels per attribute and provides 32% more profit than the best single-channel solution that

ignores the presence of the online channel and 28% more profit than the best standard heuristic.

The broad simulation setup we used is as follows: (a) generate a random instance of the ground-

truth model class; (b) determine the approximate profit/sales maximizing o✏ine assortment using

the IP-based heuristic, the RO heuristic, the greedy heuristic, and the single-channel heuristic that

ignores the impact of the online channel; (c) compare the “true” profits/sales, as computed using

the ground-truth model, from the di↵erent solutions. We repeated the above sequence of steps

for a large number of instances and various parameter combinations to cover the spectrum. For

smaller problem instances, we also determined the optimal profit through exhaustive search and

determined the optimality gaps.

Ground-truth models generated. We considered problem instances of di↵erence sizes by

varying the number of attributes K and the number of levels L per attribute. For each combination

(K,L) of parameters, we randomly generated 100 model instances, with each instance generated

as follows: (a) for each attribute-level combination (k, `), such that 1 k K and 1 ` L,

sample wo↵

k` uniformly at random from [�4,1]; (b) given wo↵

k` , set won

k` = wo↵

k` with probability 1�

⇢ = 0.4 and sample won

k` uniformly at random from [�4,wo↵

k` ) with probability ⇢/2 and (wo↵

k` ,1]

Page 22: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel22 Article submitted to Management Science; manuscript no.

with the remaining probability ⇢/2; (c) sample the profit rk` for the attribute-level combination

(k, `) uniformly at random from the interval [100/K,150/K], ensuring that product profits fall in

the interval [100,150]; (d) sample the size of the o✏ine segment ↵o↵ uniformly at random from

[↵min

,↵max

] and set the size of the online segment ↵on = 1�↵o↵ . We set ↵min

= ↵max

= 0.2 for the

first study and [↵min

,↵max

] = [0,0.5] for the second one.

The above generative model is designed to reflect the situation when the online and o✏ine

partworths are the same for some attribute-levels; for instance, the partworths may be the same

for attributes the customers are familiar with, such as, the color blue (see Table 3). The parameter

⇢ captures the fraction of attribute-levels for which the the online and o✏ine partworths di↵er. By

construction, fractions of about ⇢/2 attribute-levels are overvalued and undervalued, respectively.

4.1. Optimality gaps for smaller problem instances

Table 1 reports the optimality gaps of the IP-based heuristic proposed in Section 3.3 above when

K = 4 and L = 2. We varied the number of grid points from Jon

= Jo↵

= 2 to Jon

= Jo↵

= 32 in

powers of two, which ensured that the grids were nested. We fixed ↵o↵ = 20%, reflecting that 80%

of the customers who visit the o✏ine store also visit the online store. For each of the 100 model

instances, we computed the optimal solution through exhaustive search and then the optimality

gaps of the IP-based heuristic by varying the number of grid points. Table 1 reports the optimality

gaps, averaged over the 100 random instances.

As expected, we observe that the optimality gap shrinks but the running time increases, on

average, as the number of grid points increases. For the instances in our study, we observe that the

gap decreases to within < 0.2% of the optimal revenue with 16 grid points. The IP-based heuristic

runs in < 0.5 seconds7, even when the number of grid points is 32.

4.2. Scaling of the IP-based heuristic to larger problem instances

For the second study, we considered eight larger problem instances by varying K over

{10,20,50,100} and L over {5,10}. Because the universe of the products is exponentially large (for

K = 100 and L= 10, the universe consists of 10100 products), we focused on the problem of finding

the profit maximizing assortment of size at most C = 50. Exhaustive search is no longer computa-

tionally feasible, so we assessed the performance of the IP-based heuristic by comparing its profit

to that obtained by three benchmark methods: (a) the o✏ine heuristic, (b) the revenue-ordered

(RO) heuristic, and (c) the greedy heuristic, described next.

Benchmark methods. The o✏ine heuristic ignores the impact of the online channel and per-

forms single-channel assortment optimization. Under the MNL model, this problem can be solved in

7 The MILPs were solved using Gurobi Optimizer version 6.0.2 on a computer with processor 3.5GHz Intel Core i5,RAM of 16GB, and operating system Mac OSX Yosemite.

Page 23: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online ChannelArticle submitted to Management Science; manuscript no. 23

# grid pts. Opt. gap Run time (s)

2 6.45 0.014 2.97 0.018 0.92 0.0416 0.19 0.1232 0.11 0.43

Table 1 Average optimality gaps

and running times (in seconds) of the

IP-based heuristic as a function of the

number of grid points.

Relative performance Computation times

(K,L) RO Greedy IP RO Greedy IP

(10, 5) 1.00 1.07 1.22 0.07 0.07 0.30(20, 5) 1.00 1.01 1.25 0.21 0.24 0.51(50, 5) 1.00 1.01 1.28 1.17 1.45 1.05(100, 5) 1.00 1.01 1.27 4.58 5.65 1.93(10, 10) 1.00 1.11 1.41 0.09 0.18 7.79(20, 10) 1.00 1.01 1.45 0.30 0.66 15.93(50, 10) 1.00 1.01 1.42 1.74 3.82 54.47(100, 10) 1.00 1.00 1.27 6.79 14.66 131.50

Table 2 The “Relative performance” columns report the average ratio of the

profit from each method to that from the o✏ine heuristic, averaged over 100

problem instances. The “Computation times” columns report the average

computation time, in seconds. The IP heuristic scales to large problem instances and

extracts significantly larger profit than standard heuristics.

O(NC) time, where C is the maximum subset size and N is the number of products in the universe.

Because N is exponentially large, we implemented the following heuristic: find the most profitable

subset from among the subsets of the form {x1

,x2

, . . . ,xm} for 1m C, where x

1

, . . . ,xC are

the C most profitable products such that px

1

� px

2

� · · ·� px

C

. This heuristic returns the optimal

solution when the profit maximizing subset without the capacity constraint has size at most C

because it is known that the unconstrained profit maximizing subset comprises the m most prof-

itable products, for some m. To find the C most profitable products, we used the recent algorithm

proposed by Gallego et al. (2016); details are provided in Appendix B.

The RO heuristic finds the profit maximizing subset from among the N subsets, each comprising

the m most profitable products for m ranging from 1 to N . Because N is exponentially large,

we only search over m= 1,2, . . . ,C. The key di↵erence from the o✏ine heuristic is that while the

o✏ine heuristic picks the subset M that maximizes Ro↵(M), the RO heuristic picks the subset M

that maximizes the profit R(M) from both online and o✏ine channels.

The greedy heuristic is another general-purpose heuristic commonly applied to assortment opti-

mization problems (Jagabathula 2014). While the existing heuristics typically operate in the prod-

uct space, we implemented a natural variant that operates in the attribute space. In each iteration,

we add the feature that results in the maximum increase in the profit. We stop if the capacity is

reached or the profit no longer increases; details are provided in Appendix B.

IP-based heuristic. For the IP-based heuristic, we chose the number of grid points to be

Jon

= Jo↵

= 5, so that we solve a total of Jon

⇥ Jo↵

= 25 IPs for each instance. We enforced the

cardinality constraint by adding the linear constraintPK

k=1

P

S22

Lk

zk,S log|S| logC to the IP

described above.

Results and Discussion. The results from our simulation study are presented in Table 2. The

table reports two metrics: (a) the profits extracted by the RO, greedy, and IP-based heuristics

Page 24: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel24 Article submitted to Management Science; manuscript no.

relative to the profit extracted by the o✏ine heuristic; and (b) the computational times, in sec-

onds, of each of the heuristics. Each row corresponds to one of the eight model types, represented

by the tuple (K,L). The “Relative performance” columns report the average profit from each

heuristic, relative to that from the o✏ine heuristic, with the average computed over the 100 ran-

dom instances for each model type: 1

100

P

100

t=1

R(Mmethod

t )/R(M o✏ine

t ), where method= RO, greedy,

IP and Mmethod

t and M o✏ine

t denote the solutions found by the particular method and the o✏ine

heuristics, respectively, for problem instance t. Higher values are better, and values above 1 indi-

cate the improvements in profits from accounting for the presence of the online channel. The last

three columns report the average computational times, in seconds, averaged over the 100 random

instances for each model type. We draw the following key conclusions:

1. IP-based heuristic scales well. The computational times of the IP-based heuristic scale well

to large, practical-sized problem instances. Even when we stress test our method by applying

it to large instances with 100 attributes and 10 levels in each attribute (making the product

universe consist of 10100 products), the IP-based heuristic provides good quality solutions

within < 3 minutes, on average.

2. IP-based heuristic extracts the most profit. The IP-based heuristic vastly outperforms all the

other heuristics: 32% and 28% more profit on average extracted than the RO and greedy

heuristics, respectively. This shows that because of the problem structure, relying on general-

purpose heuristics can leave a lot of money on the table.

The above results establish the value of the IP heuristic: scales to large problem sizes and extracts

higher profits than existing benchmarks. Finally, we note that the IP-based heuristic can also be

used to solve the single-channel assortment problem by setting ↵on = 0. Recent work (Gallego

et al. 2016) has reduced the single-channel assortment problem (in the feature space) to the K-

shortest path problem in a DAG, which can be solved e�ciently using Yen’s algorithm (Yen 1971).

But existing techniques do not extend to the setting with cardinality constraints. The IP-based

heuristic, on the other hand, can readily accommodate (linear) constraints.

5. Timbuk2 case study

This section describes a case study we conducted to illustrate how our techniques apply to a real-

world application and quantify the value of our methodologies. Particularly, the study demonstrates

that the utility partworths of the same individual can change significantly after physical evaluation.

The case study focuses on messenger bags from Timbuk2 – a San Francisco-based company that

sells customized messenger bags through its online store and also showcases some of the bags in

self-owned or third-party (such as Recreational Equipment Inc., or REI) brick-and-mortar retail

stores. The key findings are (a) the di↵erences between online and o✏ine partworths are statistically

Page 25: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online ChannelArticle submitted to Management Science; manuscript no. 25

significant for 6 of the 9 included product attributes, with the magnitude of some of these di↵erences

being “large”; (b) the gain in sales and revenue from accounting for channel interactions can be

substantial (up to 40% in our case study); and (c) the single channel optimal assortment, which

ignores channel interactions, is substantially di↵erent from the optimal assortment that accounts

for channel interactions.

For our analysis, we conducted a conjoint study to collect preference data on messenger bags.

Using the collected data, we estimated participants’ online and o✏ine partworths to validate our

modeling assumption that online and o✏ine partworths di↵er. Then, using the estimated part-

worths, we computed the sales and revenue8 maximizing subsets using the o✏ine heuristic (which

ignores channel interactions) and the IP heuristic. By comparing the sales/revenues from the

resulting assortments, we show that the gains from accounting for the channel interactions are

substantial.

5.1. Details of the conjoint study

Conjoint analysis is widely used by practitioners for quantitative preference measurement. In a

typical conjoint study, participants are shown a set of products and asked to provide evaluations

by either rating, ranking, or choosing products. These evaluations are then used to back out

individual level attribute partworths by fitting utility or choice models to the responses. The

measured preferences are used by firms for demand predictions, product design decisions (Kohli

and Krishnamurti 1989), and assortment decisions in a single channel (Dobson and Kalish 1988,

1993).

Conjoint studies are typically conducted either online (in which participants evaluate descriptions

of products on a computer) or o✏ine (in which participants evaluate physical prototypes). However,

because our goal is to measure the di↵erences between online and o✏ine partworths, we asked each

participant to complete an online task, followed by an o✏ine task.

Product and attributes. We chose Timbuk2 messenger bags for our study for the following

reasons: (a) they vary on several attributes, some of which are “touch-and-feel” attributes for

which we expect di↵erences between online and o✏ine partworths; (b) they are in the right price

range – expensive enough for participants to take the decision seriously but cheap enough for

the participants to be interested in purchasing them; (c) they are infrequently purchased so that

many participants may be unfamiliar with at least some of the attributes and lack well-formed

preferences; (d) they are configurable through Timbuk2’s website (www.timbuk2.com/customizer),

allowing us to purchase bags to create a balanced orthogonal design, required to e�ciently estimate

8 We couldn’t compute profit maximizing subsets because we don’t have cost data.

Page 26: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel26 Article submitted to Management Science; manuscript no.

• Exterior design: Black, Blue, Reflective, Col-orful (illustrated in Figure 1)

• Size: Small (10⇥19⇥14 in), Large (12⇥22⇥15 in)

• Price: $120, $140, $160, $180• Strap pad: Yes, No• Water bottle pocket: Yes, No• Interior compartments: Empty bucket (nodividers), Divider for files, Crater laptopsleeve

Colorful Blue

Black Reflective

Figure 1 List of attributes and image of exterior designs, shown to participants in the online task.

the attribute partworths; and (e) they are physically small enough to simplify the logistics of

carrying out the study in a behavioral lab. Figure 1 shows the included attributes of the bag.

Study design. Based on the six attributes described above, there is a total of 42 ⇥ 3⇥ 23 = 384

(four levels of exterior design, four levels of price, three levels of interior compartments, and two

levels each of size, strap pad, and water bottle pocket) feasible feature combinations. We used

the “D-optimal” design criterion (Kuhfeld et al. 1994) to select a subset of 20 bags from the

above universe to be included in our study. Our design has a D-e�ciency metric of 0.97, which

is considered su�ciently high for reliable estimation. The configurations of the 20 bags that were

included in the study are presented in Table EC.1 in Appendix C.

Participant tasks. Each participant was asked to complete two ratings-based tasks in sequence:

an online task followed by an o✏ine task. In the online task, the participants were presented with

20 messenger bags, in sequence, on separate screens and asked to rate each bag on a 5-point scale

(Definitely not buy; Probably not buy; May or may not buy; Probably buy; Definitely buy). After

completing the online task, they were taken to a separate room to complete the o✏ine task. They

were presented with the same set of 20 bags, physically laid out on a conference table, with a card

next to each bag displaying a corresponding identifier and price. The experimenter walked them

through all the features, showing each feature on a sample bag, and asked them to evaluate the

bags and rate them on the same 5-point scale. Appendix C presents additional details of the task.

We recruited 122 participants from a university subject pool for the study. To incentivize honest

responses (Ding 2007), participants were told that they will be entered in a ra✏e and if they

Page 27: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online ChannelArticle submitted to Management Science; manuscript no. 27

win, they will receive, for free, a bag configured according to their preferences (inferred from the

responses they provided in the study) plus cash, for a total value of $1809.

5.2. Data analysis: parameter estimation

From the study, we collected two data sets (online and o✏ine), each set consisting of 20 ratings

from the 122 participants. We fitted the following linear model separately to the two data sets to

obtain the online and o✏ine partworths.

ypiz = �z +X

j

�zjxij + "pi,

where z 2 {online,o✏ine} and ypiz is the rating provided by participant p for bag i online or o✏ine.

Price is a continuous variable with one coe�cient and the remaining attributes are categorical

variables, represented using dummy coding10. Table 3 presents the estimated online and o✏ine

partworths.

Let us first focus on the online partworths. All the estimated partworths were statistically sig-

nificantly di↵erent from 0 at p < 0.001. As expected, participants had a negative price coe�cient

(�0.22). Participants also preferred Black to the other Exterior designs. For example, participants

rated Colorful Exterior design 1.06 points lower on average than Black. In the online study, par-

ticipants also preferred Large bags to Small bags, having a water bottle pocket to not having one,

and having a strap pad to not having one. The o✏ine partworths have similar interpretations.

To test whether the partworths online di↵er from those o✏ine, we fitted following model to the

data pooled from the two studies:

ypiz = �+X

j

�jxij + �z+X

j

�jzxij + "piz, (11)

where we abuse notation and let z denote a boolean variable taking the value 0 for the data from

the online study and 1 for the data from the o✏ine study. The coe�cients �j capture the di↵erence

between the o✏ine and online partworths for feature j. We compared this model with a restricted

(and nested) one obtained by restricting the coe�cients �= �j = 0, for all j:

ypiz = �+X

j

�jxij + "piz, (12)

9 The cash component was intended to eliminate any incentive for the participants to provide higher ratings for moreexpensive items to win a more expensive prize.10 We set the levels Black, Small, No strap pad, No water bottle pocket, and Empty bucket with no dividers to zerofor attributes Exterior design, Size, Strap pad, Water bottle pocket, and Interior compartment, respectively. For thecategorical variables, the coe�cients of the “default” levels (set to zero) are not identified. Their combined e↵ect isincluded in the intercept term. In total, 9 coe�cients and one intercept term were estimated for each data-set.

Page 28: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel28 Article submitted to Management Science; manuscript no.

The F-test (ANOVA test) rejected the null hypothesis that all the di↵erences � and �j, for all j, are

zero at p < 0.01, indicating that the online and o✏ine partworths di↵er statistically significantly.

The last column of Table 3 reports the coe�cients �j, which capture the di↵erence in partworths

for feature j. We note that of the 9 partworths estimates, 6 changed statistically significantly at

p < .01 (all except Blue, Price, and Divider for files), and some of the coe�cients changed by a

large amount. In particular, the population preference for Colorful went up, Reflective went down,

and size reversed (from Large to Small) after physical evaluation.

Finally, we carried out individual-level tests. We fitted the models in (11) and (12) to the data

for each individual and compared them using a F-test. We observed that the models di↵ered

statistically significantly for 29.5% (64 out of 122) of them at p < .01 and 51.6% (63 out of 122) at

p < .05. For completeness, we also fitted a mixed model with a random intercept for each participant

and found the partworths to be essentially the same; the likelihood-ratio tests comparing the models

in (11) and (12) but with random intercepts corresponding to participants also resulted in the same

conclusion, but with a slightly di↵erent p-value.

We also tested whether the o✏ine partworths persist when the consumers go back online, using

a smaller study in which we asked a group of 20 other participants to do the tasks in reverse order:

first the o✏ine task, followed by the online task. For this group, we found that an F-test comparing

models in (11) and (12) could not reject the null hypothesis that the online and o✏ine partworths

di↵er at p < .01; see Table EC.2 in Appendix C for the estimated coe�cients. Furthermore, the

individual-level tests revealed that the models in (11) and (12) were not statistically significantly

di↵erent for all of the 20 individuals at p < .01 and p < .05. The results from this second group of

participants provide evidence that the attribute partworth used for the purchase decision depends

only on whether the customer has been exposed to the attribute in a physical product, rather than

on the channel in which the purchase decision is made. Once the customer has been exposed to the

attribute level, he will apply the new partworths to both his online and o✏ine purchasing decisions.

Fitting a mixed model as described above did not change our conclusions.

We draw the following conclusions from the conjoint study. Consumers use di↵erent partworths

when evaluating products online and o✏ine and once a product is examined o✏ine, there is evidence

to suggest that consumers apply the o✏ine partworths to both online and o✏ine product evaluation.

The relevant parameters can be estimated using well-established market research tools, making

our method readily applicable in practice.

5.3. Assortment optimization: impact on sales and revenues

Using the parameters obtained from the conjoint study, we demonstrate how the firm’s sales and

revenues are a↵ected if the o✏ine assortment is selected without taking into account the online

Page 29: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online ChannelArticle submitted to Management Science; manuscript no. 29

Attribute Level Online (won) O✏ine (wo↵) Di↵erence

Exterior design Reflective �0.31 ⇤⇤ �0.60 ⇤⇤ �0.28 ⇤

Colorful �1.06 ⇤⇤ �0.71 ⇤⇤ +0.36 ⇤⇤

Blue �0.22 ⇤⇤ �0.11 +0.11

Black

Size Large 0.27 ⇤⇤ �0.31 ⇤⇤ �0.58 ⇤⇤

Small

Price $120, $140, $160, $180 �0.011 ⇤⇤ �0.008 ⇤⇤ +0.004

Strap pad Yes 0.51 ⇤⇤ 0.25 ⇤⇤ �0.26 ⇤⇤

No

Water bottle pocket Yes 0.45 ⇤⇤ 0.17 ⇤⇤ �0.28 ⇤⇤

No

Interior compartments Divider for files 0.41 ⇤⇤ 0.52 ⇤⇤ +0.11

Crater laptop sleeve 0.62 ⇤⇤ 0.88 ⇤⇤ +0.26 ⇤

Empty bucket/no dividers

Intercept 3.72 ⇤⇤ 3.39 ⇤⇤ �0.33

Notes: ⇤⇤p < 0.001, ⇤p < 0.01Table 3 Statistically significant di↵erences were observed between the online and o✏ine partworths for several

attribute-level combinations. Results are based on the 122 participants who completed the online task first, followed by the

o✏ine task. The levels with no coe�cients were set to zero in dummy encoding.

channel. We used the IP formulation obtained in Section 3.3 to optimize the assortment with

and without the online channel for various sizes of the o✏ine segment of consumers. Our results

demonstrate that the gain in revenues from accounting for the online channel can be significant if

a large portion of the population visits both channels.

Because we are assuming that customer choices are described by the MNL model, we verified

its fit by carrying out a five-fold cross-validation on both the online and o✏ine conjoint data.

We measured the out-of-sample error in terms of the standard mean absolute percentage error

(MAPE) metric, which measures the average relative error in predicting market shares. We found

that the MNL model has about 3.7% error rate, indicating that it is a good fit. The details of the

verification are in Appendix C.2.

Setup. In order to compute the optimal assortments, we used the following parameter values.

The utility parameters won

k` and wo↵

k` , for all non-price attribute levels k, ` (presented in Table 3),

were obtained from the conjoint study. We obtained the partworth revenues from the prices that

Page 30: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel30 Article submitted to Management Science; manuscript no.

Timbuk2 posted on their website for their customizable messenger bags. The price for the base

configuration is $140, which corresponds to the product configuration Black, Small, No strap pad,

No water bottle pocket, Empty bucket. Because of the dummy coding, the utility of the base

configuration is equal to the intercept + �price

· $140. Here we use the o✏ine intercept (because

the intercepts don’t di↵er significantly) and �price

= �0.008 is the o✏ine price coe�cient. The

partworth revenues ⇡k` for the non-dummy attribute levels were obtained from the additional prices

over the base that Timbuk2 charges: $10 for Reflective and Colorful; $10 for Large; $15 for Strap

pad; $5 for Water bottle pocket; $10 for Laptop compartment; and $0 for the Divider for files.

For every non-dummy attribute level k, `, we set wc

k` = wc

k` + �price

· ⇡k,`, for c 2 {on,o↵}. Becauseonly di↵erences in mean utilities matter, we absorbed the utility of the base configuration into

the utility of the no-purchase option, which was set to achieve reasonable market shares of about

40% in the o✏ine channel (range between 41% and 56% in both channels depending on the size

of the o✏ine customer segment). With these parameter values, we carried out sales and revenue

maximization and compared the optimal solutions with the single channel benchmark.

Results. Using the IP-based heuristic (with the number of grid points Jon

= Jo↵

= 10) described

in Section 3.3, we computed both sales and revenue maximizing assortments by varying the online

segment sizes from 0.5 to 1. We compared the optimal sales and revenues against the benchmark

sales and revenues, respectively, obtained from maximizing the o✏ine channel (ignoring channel

interactions). The benchmark solution was obtained by applying our IP-based heuristic with ↵= 1.

Figure 2 reports the gains from accounting for the presence of the online channel in the firm

sales and revenues. The upward trend indicates that as the number of customers visiting both

channels increases, it becomes more important to account for the online channel in determining

the o✏ine assortment. When all customers visit both channels, the gain in both sales and revenue

is about 40%. As the number of these customers decreases and more customers purchase only from

the o✏ine store, accounting for the online channel becomes less relevant.

Figure 2 also shows the attribute levels contained in the sales and revenue maximizing assort-

ments for di↵erent values of ↵. Focusing on the sale maximizing subsets, we note that when the

o✏ine segment is large, all features are included. This is because this regime is dominated by

the o✏ine segment, for which it is indeed optimal to include all feasible products. As the online

segment grows, the assortment changes to exclude over-valued features: Reflective, Large, Strap

pad, and Water bottle pocket. This result is consistent with the result of Theorem 3.1. Excluding

over-valued attributes benefits the sales from the online segment but hurts sales from the o✏ine

segment. This tension is strongest for over-valued attributes that have positive o✏ine partworths

(in our case, Strap pad and Water bottle pocket) because excluding them decreases the o✏ine

Page 31: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online ChannelArticle submitted to Management Science; manuscript no. 31

sales most significantly. As a result, the decision whether to include these attributes is particularly

sensitive to the value of ↵. The interpretation of the results for the revenue maximizing subsets is

similar. When the o✏ine segment is large, all features are included, which is optimal for the o✏ine

channel because the firm’s market share is “small enough.” When the online segment becomes

larger, the firm will exclude over-valued features because they bring in the most revenues and

hiding them increases the chance of their sale.

0

10

20

30

40

0

10

20

30

40

salesrevenu

es

0.5 0.6 0.7 0.8 0.9 1.0

online segment size (1�↵)

%gain

relative

tosinglechan

nel

solution

Attribute-Level sales max. revenue max.

0.5� 0.6 0.7 � 0.8 0.5� 0.6 � 0.7

Exterior design

Reflective⇤ ⇥ ⇥ ⇥Colorful ⇥ ⇥ ⇥ ⇥ ⇥Blue ⇥ ⇥ ⇥ ⇥ ⇥Black ⇥ ⇥ ⇥ ⇥ ⇥Size

Large⇤ ⇥ ⇥Small ⇥ ⇥ ⇥ ⇥ ⇥Strap pad

Yes⇤ ⇥ ⇥ ⇥No ⇥ ⇥ ⇥ ⇥ ⇥Water bottle pocket

Yes⇤ ⇥ ⇥ ⇥No ⇥ ⇥ ⇥ ⇥ ⇥Interior compartments

Divider for files ⇥ ⇥ ⇥ ⇥ ⇥Crater laptop sleeve ⇥ ⇥ ⇥ ⇥ ⇥Empty bucket/no dividers ⇥ ⇥ ⇥ ⇥ ⇥

Figure 2 The figure shows the gains in the revenues and sales from accounting for the online channel. The table marks

by ⇥ the attribute levels that are present in the sales and revenue max. subsets. The features marked as

⇤are over-valued.

To gain insights into the products included in the sales maximizing o✏ine assortment, we con-

strained its cardinality to be 4, which allowed us to examine the individual products included.

Figure 3 shows the sales maximizing subset of products at di↵erent values of the online segment

size. Each product in the optimal subset is represented by a point whose horizontal axis value

denotes the product’s “popularity” and the radius denotes the product’s “informativeness.” We use

the o✏ine utility of the product as a measure of its popularity and di↵erence between its o✏ine and

online utilities as a measure of its informativeness. The broad insight we obtain from our results

is that the optimal assortment is a mix of popular and informative products. The popular prod-

ucts have high utilities and generate sales in the o✏ine channel, whereas the informative products

expose customers to under-valued attributes and generate sales in the online channel. When the

o✏ine segment dominates, i.e., the o✏ine channel generates most of the firm’s sales, it is optimal

to fill the capacity with popular products; we see this on the graph at the second to last row where

Page 32: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel32 Article submitted to Management Science; manuscript no.

the assortment consists of products with high utilities. On the other hand, when the online segment

dominates, it is optimal to o↵er informative products at the expense of popular ones; we see this

on the graph where the points become larger (more informative products) but move to the left

(less popular) in the upper rows.

all prods

0� 0.1

0.2� 0.5

> 0.5

1.5 2.0 2.5 3.0 3.5

utilities

onlinesegm

ent(1

�↵)

informativeness-0.2

0.0

0.2

0.4

0.6

Figure 3 Utilities of products included in the optimal assortments of size at most 6 at di↵erent sizes of the online

segment. The last row plots the utilities of all the 96 products in the universe. Products represented by larger sized points are

more “informative,” measured as the di↵erence between the o✏ine and online partworths.

6. Conclusions and future work

This work focused on a firm’s showcase decision: selecting a subset of products to o↵er in an

o✏ine channel from a larger product line o↵ered through the online channel in order to maximize

expected profits across both channels. A key component of our consumer demand model is that

utility partworths change when customers learn about products by inspecting them physically in

a brick-and-mortar store. In the context of this demand model, we formalized the decision prob-

lem, established computational hardness, and proposed approximation algorithms with theoretical

guarantees. In addition, we used a demonstrative case study with messenger bags to estimate con-

sumers’ utility parameters in a conjoint study. Through this case study, we demonstrated that

accounting for channel interactions can result in substantial gains in expected revenue (up to 40%

in our case); the composition of the optimal assortment can also be significantly di↵erent. By lay-

ing out a framework for product showcasing, this work provides a platform for other interesting

aspects of omnichannel retailing. Next, we discuss two specific directions in which this work can

be extended.

This paper assumes that consumers exogenously decide whether to visit one or both of the

channels. However, a consumer’s decision to visit the o✏ine (online) channel may depend on the

products she examines online (o✏ine) and o↵ers of “in-store exclusives” by the firm to encourage

Page 33: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online ChannelArticle submitted to Management Science; manuscript no. 33

store visits. Accounting for these e↵ects by endogenizing the store visit decision is a promising

future direction that is particularly relevant when the firm sells through multiple channels.

Further, the utility model proposed in this work provides a framework for modeling product

returns. Given that many online retailers, such as Warby Parker, Zappos, or Bonobos, o↵er generous

return policies, the o✏ine channel can be viewed as a way to mitigate costs of product returns.

When consumers purchase from these retailers, they decide what to order based on their online

evaluation of the available items. However, once they receive their order, they determine what they

want to keep based on physical evaluation.

Finally, our utility model ignores interactions between attributes. While our proposed modeling

framework readily extends (the partworths of the interaction terms change upon exposure to one

of the attributes), the algorithmic methods may face computational challenges because the full-

factorial assumption may be violated. Extending our algorithms to handle cases with constraints

on the feasible products is a promising future direction.

Page 34: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel34 Article submitted to Management Science; manuscript no.

ReferencesAlptekinoglu, Aydın, Charles J Corbett. 2010. Leadtime-variety tradeo↵ in product di↵erentiation. Manu-

facturing & Service Operations Management 12(4) 569–582.

Alptekinoglu, Aydın, Alex Grasas. 2014. When to carry eccentric products? Optimal retail assortment under

consumer returns. Production and Operations Management 23(5) 877–892.

Balakrishnan, PV, Varghese S Jacob. 1996. Genetic algorithms for product design. Management Science

42(8) 1105–1117.

Belloni, Alexandre, Robert Freund, Matthew Selove, Duncan Simester. 2008. Optimizing product line designs:

E�cient methods and comparisons. Management Science 54(9) 1544–1552.

Bront, Juan Jose Miranda, Isabel Mendez-Dıaz, Gustavo Vulcano. 2009. A column generation algorithm for

choice-based network revenue management. Operations Research 57(3) 769–784.

Brynjolfsson, Erik, Yu Hu, Mohammad S Rahman. 2009. Battle of the retail channels: How product selection

and geography drive cross-channel competition. Management Science 55(11) 1755–1765.

Cohen, Maxime C, Ilan Lobel, Renato Paes Leme. 2016. Feature-based dynamic pricing. Available at SSRN

.

Davis, James, Guillermo Gallego, Huseyin Topaloglu. 2013. Assortment planning under the multinomial

logit model with totally unimodular constraint structures. Department of IEOR, Columbia University.

Available at http://legacy.orie.cornell.edu/huseyin/publications/logit const.pdf .

Davis, James M, Guillermo Gallego, Huseyin Topaloglu. 2014. Assortment optimization under variants of

the nested logit model. Operations Research 62(2) 250–273.

Desir, Antoine, Vineet Goyal. 2014. Near-optimal algorithms for capacity constrained assortment optimiza-

tion. Available at SSRN 2543309 .

Ding, Min. 2007. An incentive-aligned mechanism for conjoint analysis. Journal of Marketing Research 44(2)

214–223.

Dobson, Gregory, Shlomo Kalish. 1988. Positioning and pricing a product line. Marketing Science 7(2)

107–125.

Dobson, Gregory, Shlomo Kalish. 1993. Heuristics for pricing and positioning a product-line using conjoint

and cost data. Management Science 39(2) 160–175.

Feldman, Jacob B, Huseyin Topaloglu. 2014. Capacity constraints across nests in assortment optimiza-

tion under the nested logit model. Tech. rep., Cornell University, School of Operations Research and

Information Engineering.

Forman, Chris, Anindya Ghose, Avi Goldfarb. 2009. Competition between local and electronic markets: How

the benefit of buying online depends on where you live. Management Science 55(1) 47–57.

Fruchter, GE, A Fligler, RS Winer. 2006. Optimal product line design: Genetic algorithm approach to

mitigate cannibalization. Journal of Optimization Theory and Applications 131(2) 227–244.

Page 35: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online ChannelArticle submitted to Management Science; manuscript no. 35

Gallego, Guillermo, Anran Li, Jose Lius Beltran. 2016. Produce line design and pricing under logit model.

Columbia University Working Paper .

Gallego, Guillermo, Huseyin Topaloglu. 2014. Constrained assortment optimization for the nested logit

model. Management Science 60(10) 2583–2601.

Garey, Michael R, David S Johnson. 1979. Computers and intractability: A guide to np-completeness.

Gaur, Vishal, Dorothee Honhon. 2006. Assortment planning and inventory decisions under a locational

choice model. Management Science 52(10) 1528–1543.

Ghoniem, Ahmed, Bacel Maddah. 2015. Integrated retail decisions with multiple selling periods and customer

segments: optimization and insights. Omega 55 38–52.

Ghoniem, Ahmed, Bacel Maddah, Ameera Ibrahim. 2013. Optimizing assortment and pricing of multiple

retail categories with cross-selling. Journal of Global Optimization 1–19.

Green, Paul E, Abba M Krieger. 1985. Models and heuristics for product line selection. Marketing Science

4(1) 1–19.

Green, Paul E, Vithala R Rao. 1971. Conjoint measurement for quantifying judgmental data. Journal of

Marketing research 355–363.

Green, Paul E, Venkat Srinivasan. 1990. Conjoint analysis in marketing: new developments with implications

for research and practice. The Journal of Marketing 3–19.

Jagabathula, Srikanth. 2014. Assortment optimization under general choice. Available at SSRN 2512831 .

Jagabathula, Srikanth, Paat Rusmevichientong. 2015. A nonparametric joint assortment and price choice

model. Available at SSRN 2286923 .

Jagabathula, Srikanth, Gustavo Vulcano. 2015. A model to estimate individual preferences using panel data.

Available at SSRN 2560994 .

Kohli, Rajeev, Ramesh Krishnamurti. 1987. A heuristic approach to product design. Management Science

33(12) 1523–1533.

Kohli, Rajeev, Ramesh Krishnamurti. 1989. Optimal product design using conjoint analysis: Computational

complexity and algorithms. European Journal of Operational Research 40(2) 186–195.

Kohli, Rajeev, Ramamirtham Sukumar. 1990. Heuristics for product-line design using conjoint analysis.

Management Science 36(12) 1464–1478.

Kok, A Gurhan, Marshall L Fisher. 2007. Demand estimation and assortment optimization under substitu-

tion: Methodology and application. Operations Research 55 1001–1021.

Kuhfeld, Warren F, Randall D Tobias, Mark Garratt. 1994. E�cient experimental design with marketing

research applications. Journal of Marketing Research 31(November) 545–557.

Lawler, Eugene L. 1979. Fast approximation algorithms for knapsack problems. Mathematics of Operations

Research 4(4) 339–356.

Page 36: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel36 Article submitted to Management Science; manuscript no.

Li, Guang, Paat Rusmevichientong, Huseyin Topaloglu. 2015. The d-level nested logit model: Assortment

and price optimization problems. Operations Research 62(2) 325–342.

Mehra, Amit, Subodha Kumar, Jagmohan S Raju. 2013. ‘Showrooming’ and the competition between store

and online retailers. SSRN Working paper 29–43.

Randall, Taylor, Karl Ulrich, David Reibstein. 1998. Brand equity and vertical product line extent. Marketing

science 17(4) 356–379.

Rodrıguez, Betzabe, Goker Aydın. 2011. Assortment selection and pricing for configurable products under

demand uncertainty. European Journal of Operational Research 210(3) 635–646.

Rusmevichientong, Paat, Zuo-Jun Max Shen, David B Shmoys. 2009. A PTAS for capacitated sum-of-ratios

optimization. Operations Research Letters 37(4) 230–238.

Rusmevichientong, Paat, Zuo-Jun Max Shen, David B Shmoys. 2010. Dynamic assortment optimization

with a multinomial logit choice model and capacity constraint. Operations Research 58(6) 1666–1680.

Rusmevichientong, Paat, David Shmoys, Chaoxu Tong, Huseyin Topaloglu. 2014. Assortment optimiza-

tion under the multinomial logit model with random choice parameters. Production and Operations

Management 23(11) 2023–2039.

Subramanian, Shivaram, Hanif D Sherali. 2010. A fractional programming approach for retail category price

optimization. Journal of Global Optimization 48(2) 263–277.

Talluri, Kalyan, Garrett Van Ryzin. 2004. Revenue management under a general discrete choice model of

consumer behavior. Management Science 50(1) 15–33.

Ulu, Canan, Dorothee Honhon, Aydın Alptekinoglu. 2012. Learning consumer tastes through dynamic

assortments. Operations Research 60(4) 833–849.

Yen, Jin Y. 1971. Finding the k shortest loopless paths in a network. Management Science 17(11) 712–716.

Page 37: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

e-companion to Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel ec1

This page is intentionally blank. Proper e-companion title

page, with INFORMS branding and exact metadata of the

main paper, will be produced by the INFORMS o�ce when

the issue is being assembled.

Page 38: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

ec2 e-companion to Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel

Online Appendix: O✏ine Assortment Optimization in thePresence of an Online ChannelAppendix A: Proofs for Section 3

A.1. Proofs for Section 3.1

Proof of Lemma 3.1 Recall from Section 2 that the expected profit is given by

Rpure(S) =

P

x2X

P

k,`

rk`

1l`

[xk

]⌘

exp�

P

K

k=1

uon

k

(xk

, Sk

)�

1+P

y2X exp�

P

K

k=1

uon

k

(yk

, Sk

)� . (EC.1)

It follows from the argument above the statement of Lemma 3.1 that the denominator simplifies as

1+X

y2X

exp

K

X

k=1

uon

k

(yk

, Sk

)

!

= 1+K

Y

k=1

X

`2Sk

ewoff

k` +X

`/2Sk

ewon

k`

!

.

Noting that �k`

:= ewoff

k` � ewon

k` and dk

:=P

`2Lkew

on

k` , we can writeX

`2Sk

ewoff

k` +X

`/2Sk

ewon

k` =X

`2Lk

ewon

k` +X

`2Sk

ewoff

k` � ewon

k`

= dk

+X

`2Sk

�k`

:=Dk

(Sk

). (EC.2)

We have thus shown that the denominator is equal to 1 +Q

K

k=1

Dk

(Sk

) = 1 + D(S), where D(S) :=Q

K

k=1

Dk

(Sk

).

We now consider the numerator. We first determine the term multiplying rk`

for some k, `. The term can

be obtained by fixing xk

= ` and varying xk

0 over all possible values in Lk

0 for all k0 6= k. Let X�k denote

⇥k

0 6=k

Lk

0 (where⇥a2A

Sa

denotes the cartesian product of all the sets in the collection {Sa

: a2A} for

any set A) and x

�k denote the vector with the kth component dropped. We then obtain that the term

multiplying rk`

is equal to

euon

k (`,Sk)X

x

�k2X�k

exp

X

k

0 6=k

uon

k

0 (x�k

k

0 , Sk

0)

!

= euon

k (`,Sk)Y

k

0 6=k

0

@

X

`2Sk0

ewoff

k0` +X

`/2Sk0

ewon

k0`

1

A ,

where the equality follows from interchanging sum and product, similar to (EC.1) above. It now follows

from (EC.2) that the term multiplying rk`

is equal to euon

k (`,Sk)Q

k

0 6=k

Dk

0(Sk

0) = euon

k (`,Sk)D(S)/Dk

(Sk

). Now

for each k, collecting all the terms corresponding to rk`

for `2Lk

together, we obtain

X

`2Lk

rk`

euon

k (`,Sk)D(S)/Dk

(Sk

) =D(S)

Dk

(Sk

)

X

`2Sk

rk`

euon

k (`,Sk) +X

`/2Sk

rk`

euon

k (`,Sk)

!

=D(S)

Dk

(Sk

)

X

`2Sk

rk`

ewoff

k` +X

`/2Sk

rk`

ewon

k`

!

,

where the last equality follows from the definition uon

k

(xk

, Sk

) =P

`2Skew

off

k` 1l`

[xk

]+P

`/2Skew

off

k` 1l`

[xk

]. Noting

that bk

:=P

`2Lkrk`

ewon

k` , we can writeX

`2Sk

rk`

ewoff

k` +X

`/2Sk

rk`

ewon

k` =X

`2Lk

rk`

ewon

k` +X

`2Sk

rk`

ewoff

k` � ewon

k`

= bk

+X

`2Sk

rk`

�k`

.

We thus have shown that the numerator is equal toK

X

k=1

D(S)bk

+P

`2Skrk`

�k`

Dk

(Sk

)=D(S)

K

X

k=1

Rk

(Sk

), where Rk

(Sk

) :=bk

+P

`2Skrk`

�k`

Dk

(Sk

).

The expected revenue function now becomes

Rpure(S) =D(S)

P

K

k=1

Rk

(Sk

)

1+D(S)=

P

K

k=1

Rk

(Sk

)

1+ 1/D(S).

The result of the lemma now follows.

Page 39: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

e-companion to Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel ec3

Proof of Theorem 3.2 To prove this result, we first determine the profit impact of adding a feature to

the o✏ine o↵er set. For that, for some subset of features S = (S1

, . . . , SK

), consider attribute k and levels

z1

, z2

2 Lk

such that z1

2 Sk

and z2

/2 Sk

. Let S + z2

denote the subset of features (S1

, . . . , Sk�1

, Sk

[{z

2

} , Sk+1

, . . . , SK

) and S� z1

denote the subset (S1

, . . . , Sk�1

, Sk

\ {z1

} , Sk+1

, . . . , SK

). We claim that

R(S+ z2

)>R(S) () �kz

2

·�

rkz

2

��

Rk

(Sk

)�R(S)/D(S)��

> 0 and

R(S� z1

)>R(S) () �kz

1

·�

rkz

1

��

Rk

(Sk

)�R(S)/D(S)��

< 0.(EC.3)

We establish (EC.3) below. Assuming (EC.3) is true, we establish the result by contradiction.

For that, suppose z 2 L+

k

, so that �kz

> 0. If rkz

> t⇤k

and z /2 S⇤+k

, then, by taking S = S

⇤ and z2

= z, it

follows from (EC.3) that R(S⇤ + z)>R(S⇤) contradicting the fact that S

⇤ is optimal. Therefore, we must

have�

`2L+

k

: rkz

> t⇤k

✓ S⇤+k

. Similarly, suppose z 2 L+

k

is such that rkz

< t⇤k

and z 2 S⇤k

. Then, taking

S = S

⇤ and z1

= z, we get from (EC.3) that R(S⇤ � z)>R(S⇤), contradicting the fact that S⇤ is optimal.

We have thus shown that S⇤+k

✓�

`2L+

k

: rkz

� t⇤k

. It thus follows that

`2L+

k

: rkz

> t⇤k

✓ S⇤+k

✓�

`2L+

k

: rkz

� t⇤k

.

Similarly, for any z 2L�k

, because �kz

< 0, following the same set of arguments yields that

{`2L�k

: rkz

< t⇤k

}✓ S⇤�k

✓ {`2L�k

: rkz

t⇤k

} .

We are now are only left with establishing (EC.3). For that, let z denote z2

and S denote S + z. Then,

noting that Dk

(Sk

)�Dk

(Sk

) = �kz

and Dk

(Sk

)Rk

(Sk

)�Dk

(Sk

)Rk

(Sk

) = rkz

�kz

, it follows that

Rk

(Sk

)�Rk

(Sk

) =�kz

Dk

(Sk

)· (r

kz

�Rk

(Sk

)). (EC.4)

It can also be verified that D(S) =D(S)Dk

(Sk

)/Dk

(Sk

) =D(S)/(1� �kz

(Dk

(Sk

))�1). Therefore, we have

1+1/D(S) = 1+1/D(S)� �kz

/(Dk

(Sk

) ·D(S)). We must thus have

(1+1/D(S))[R(S)�R(S)] =K

X

k

0=1

Rk

0(Sk

0)� (1+ 1/D(S))R(S)+�kz

D(S)Dk

(Sk

)R(S)

=K

X

k

0=1

Rk

0(Sk

0)�K

X

k

0=1

Rk

0(Sk

0)+�kz

D(S)Dk

(Sk

)R(S)

=Rk

(Sk

)�Rk

(Sk

)+�kz

D(S)Dk

(Sk

)R(S)

=1

Dk

(Sk

)· �

kz

· (rkz

� [Rk

(Sk

)�R(S)/D(S)]) ,

where the last equality follows from (EC.4). Because D(S)> 0, it follows that the sign of R(S)�R(S) is

determined by the sign of �kz

· (rkz

� [Rk

(Sk

)�R(S)/D(S)]). This establishes the first part of (EC.3).

In a similar fashion, if S = S � z1

, by replacing �kz

by ��kz

1

, z by z1

, and following the same set of

arguments, we can show that

(1+1/D(S))[R(S)�R(S)]] =� 1

Dk

(Sk

)· �

kz

1

· (rkz

1

� [Rk

(Sk

)�R(S)/D(S)])

from which the second part of (EC.3) follows. We have thus established (EC.3). The result of the theorem

now follows. ⇤

Page 40: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

ec4 e-companion to Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel

Proof of Theorem 3.3 In order to prove the hardness, we consider a particular instance of the decision

problem. Suppose that |Lk

|= 2 and �k1

= 0. Because �k1

= 0, o↵ering level 1 in each attribute does not change

the expected profit. Therefore, without loss of generality, we suppose that 1 2 S⇤k

for all k. The assortment

decision now reduces to determining whether or not to o↵er level 2 for each k. To simplify the notation, let

A✓ {1,2, . . . ,K} denote the subset of attributes for which we o↵er level 2 in the store. We abuse notation and

let Rpure(A) denote Rpure(SA

), where SA

= (SA,1

, . . . , SA,K

) is such that SA,k

= {1,2} if k 2A and SA,k

= {1}

if k /2A. It now follows from the definitions of Rk

(Sk

) and DK

(Sk

) that

Rk

(SA,k

) =

(

(bk

+ rk2

)/ (dk

+ �k2

) , if k 2A

bk

/dk

, if k /2A.and D

k

(SA,k

) =

(

dk

+ �k2

, if k 2A

dk

, if k /2A..

Now, letting

�k

:=bk

+ rk2

dk

+ �k2

� bk

dk

=rk2

dk

� bk

�k2

dk

(dk

+ �k2

)and �

k

:= log

dk

+ �k2

dk

,

we can write

Rpure(A) =b+

P

k2A

�k

1+ de�P

k2A �k, where b=

K

X

k=1

bk

dk

and d= 1/

K

Y

k=1

dk

!

Now suppose �k

= ��k

for all k. The decision problem can now be formulated as: Is there a subset

A✓ {1,2, . . . ,K} such that

b�P

k2A

�k

1+ de�P

k2A �k�Q () b�

X

k2A

�k

� dQe�P

k2A �k �Q. (EC.5)

We obtain a reduction from the popular partition problem, defined as follows.

Partition

Inputs: The set of items indexed by 1,2, . . . ,K and sizes tk

2Z+

associated with each item.

Question: Is there a subset A⇢ {1,2, . . . ,K} such thatP

k2A

tk

=P

k/2A

tk

?

We now obtain a reduction as follows. Given an instance of the Partition problem, let T�

= 1

2

P

K

k=1

tk

.

Without loss of generality, suppose that T 2Z+

. Then, there exists a subset A such thatP

k2A

tk

=P

k/2A

tk

if and only if there exists a subset A such thatP

k2A

tk

= T . Our goal is to determine if there is a subset A

such thatP

k2A

tk

= T .

We create an instance of the decision problem (EC.5) as follows. Given tk

and T , define: �k

= tk

for all k,

the target profit value Q= eT/d, and b=Q+ 1+ T . With this assignment, the Partition problem has a

solution if and only if there exists a subset A such that

b�X

k2A

�k

� dQe�P

k2A �k max�2[0,2T ]

b��� dQe�� = b�T � 1 =Q

where the first equality follows from the fact that the mapping � 7! b���dQe�� is concave in � over [0,2T ]

with a unique maximum at �⇤ = log(dQ) = T and the last equality follows from the choice of b.

Therefore, the answer to the Partition problem is yes if and only if there exists a subset A✓ {1,2, . . . ,K}

such that (EC.5) is satisfied.

The result of the theorem now follows.

Page 41: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

e-companion to Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel ec5

Proof of Lemma 3.2 Suppose there exists an S 2 Z1

⇥ · · · ⇥ ZK

such thatP

K

k=1

Rk

(Sk

) � q andP

K

k=1

logDk

(Sk

)� t. We claim that S satisfies the constraintP

K

k=1

jSk,k

� ⇢. To see this, consider

K

X

k=1

jSk,k

�K

X

k=1

logDk

(Sk

)

"t/K� 1

=1

"t/K

K

X

k=1

logDk

(Sk

)�K � t

"t/K�K � bK/"c�K = ⇢,

where the first inequality follows from the fact that bxc � x � 1 for any real number x. We have thus

shown that S is a feasible solution to the optimization problem corresponding to the value function V (K,⇢).

Because the DP maximizesP

K

k=1

Rk

(Sk

) over all S 2Z1

⇥ · · ·⇥ZK

satisfyingP

K

k=1

S,k

� ⇢, it follows that

the output of the DP S must satisfyP

K

k=1

Rk

(Sk

)�P

K

k=1

Rk

(S⇤k

)� q. We are now left only to prove thatP

k

logDk

(Sk

)� (1� 2")t. To prove this, we observe

⇢K

X

k=1

Sk,k

K

X

k=1

logDk

(Sk

)

"t/K.

As a result,

K

X

k=1

logDk

(Sk

)� "t

K(bK/"c�K)� "t

K(K/"� 1�K) = t (1� "/K � ")� t(1� 2"),

where the second inequality follows from the fact that bxc � x� 1 and the last inequality follows from the

fact that K � 1.

The result of the lemma now follows.

Proof of Theorem 3.4 We proceed as follows. Let ⌧min

denote e�P

k logDk(L+

k) and ⌧

max

denote

e�P

k logDk(L�k). Let T denote the " grid of the interval [⌧

min

, ⌧max

] such that T = {⌧min

(1+ ")i : i= 0,1, . . . , I}where I =O (log(⌧

max

/⌧min

)/ log(1+ ")). We carry out the DP algorithm with t=� log ⌧ for every grid point

⌧ 2 T . Let S⌧

denote the output obtained from solving the DP for grid point ⌧ and let S denote the subset

from the collectionn

S

: ⌧ 2 To

with maximum expected profit⇣

P

k

Rk

(Sk

)⌘

/⇣

1+ e�P

k logDk(ˆ

Sk)

. We

now show that S is a good approximation of the optimal subset.

For that, let S⇤ denote the optimal profit maximizing set of attribute levels and r⇤ the optimal profit from

o↵ering S

⇤. Define t⇤ =P

k

logDk

(S⇤k

). Now consider the grid point ⌧ 2 T such that ⌧/(1 + ") e�t

⇤ ⌧ .

Now, let q= r⇤(1+ e�t

⇤) and t=� log ⌧ . We must have that

X

k

Rk

(S⇤k

) = r⇤(1+ e�t

⇤) = q and

X

k

logDk

(S⇤k

) = t⇤ �� log ⌧ = t,

where the first set of equalities follow from the definition of r⇤ and q and the second set of equalities follow

from the definitions of t⇤ and ⌧ and the fact that e�t

⇤ ⌧ . It now follows from Lemma 3.2 that running the

DP with t=� log ⌧ and " replaced by "/ |log ⌧min

| outputs S⌧

such that

X

k

Rk

(S⌧,k

)� q= r⇤(1+ e�t

⇤) and

X

k

logDk

(S⌧,k

)��(1� 2"/ |log ⌧min

|) log ⌧.

We can now writeP

k

Rk

(S⌧,k

)

1+ e�P

k logDk(S⌧,k)� r⇤

1+ e�t

1+ e(1�2"/|log ⌧

min

|) log ⌧

� r⇤1+ e�t

1+ e(1�2"/|log ⌧

min

|)(�t

⇤+log(1+"))

,

where the second inequality follows from the fact that ⌧ (1+ ")e�t

⇤. Now consider

e�(1�2"/|log ⌧

min

|)t⇤ = e�t

⇤e2"t

⇤/|log ⌧

min

| e�t

⇤e2" (1+ 7")e�t

⇤,

Page 42: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

ec6 e-companion to Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel

where the first inequality follows from the fact that t⇤ |log ⌧min

| and the last inequality follows from verifying

that e2" 1+7" for 0 " 1. Furthermore, we must have

e(1�2"/|log ⌧

min

|) log(1+") elog(1+") = (1+ ").

It thus follows thatP

k

Rk

(S⌧,k

)

1+ e�P

k logDk(S⌧,k)� r⇤

1+ e�t

1+ (1+ ")(1+ 7")e�t

⇤ � r⇤1

(1+ ")(1+ 7")� r⇤(1� 8").

Because the algorithm outputs the subset S that maximizes the expected profit for all possible grid points,

we can conclude thatP

k

Rk

(Sk

)

1+ e�P

k logDk(ˆ

Sk)�

P

k

Rk

(S⌧,k

)

1+ e�P

k logDk(S⌧,k)� (1� 8")r⇤.

Running time. In order to determine the running time, note that it follows from the discussion before the

statement of Lemma 3.2 that for each grid point, the running time of the DP is O⇣

K2maxk

|Lk

|2 |log ⌧min

|/"⌘

because the DP is run with " replaced by "/ log ⌧min

. Now because the DP is run for each grid point and

there is a total of O(log(⌧max

/⌧min

)/ log(1 + ")) grid points, the total running time of the algorithm is

O⇣

K2maxk

|Lk

|2 |log ⌧min

| log(⌧max

/⌧min

)/(" log(1+ "))⌘

.

The result of the theorem now follows.

A.2. Proofs for Section 3.2

Proof of Theorem 3.5 To simplify notation, let vx

denote exp�

P

k

uo↵

k

(xk

)�

for any x2X . For any assort-

ment M ✓X , let F (M) :=P

x2X Px

(M) and G(M) :=P

x2M

Qx

(M). It thus follows that the expected sales

from o↵ering M is equal to (1�↵)F (M)+↵G(M). Further, define the logistic function f(x) = x/(1+x) for

any x> 0. It can be shown that f(·) is increasing in x.

Now note that G(M) =�

P

x2M

vx

/�

1+P

x2M

vx

= f(P

x2M

vx

). Because f(x) is increasing in x, it

follows that G(·) is an increasing set function i.e., G(M)<G(M 0) for any M ⇢M 0.

Furthermore, by setting rk`

= 1/K, it follows from Lemma 3.1 that F (M) = f(D(SM)), where

D(S) =Q

k

DK

(Sk

) and Dk

(Sk

) = dk

+P

k2Sk�k`

for any S and S

M denotes (SM

1

, . . . , SM

K

) where SM

k

=

{`2Lk

: xk

= ` for some x2M}.

We now establish the result as follows. Let M⇤ denote the optimal assortment and let S⇤ = (S⇤1

, . . . , S⇤K

)

denote SM

⇤, the set of attribute levels covered by the products in M⇤. We first claim that M⇤ = S⇤

1

⇥ · · ·⇥S⇤K

.

For that, let M denote S⇤1

⇥ · · · ⇥ S⇤K

. It immediately follows from the definitions that S

M

⇤= S

˜

M and

M⇤ ✓ M . Now, if M⇤ ⇢ M , then

(1�↵)F (M)+↵G(M) = (1�↵)f(D(S˜

M))+↵G(M)> (1�↵)f(D(SM

⇤))+↵G(M⇤) = (1�↵)F (M⇤)+↵G(M⇤),

where the inequality follows from the fact that S

M

⇤= S

˜

M and G(·) is an increasing set function. This

contradicts that fact that M⇤ maximizes (1�↵)F (M)+↵G(M) over all M ✓X . Therefore, it must be that

M⇤ = S⇤1

⇥ · · ·⇥S⇤K

.

We now show that L+

k

✓ S⇤k

for all k. For that, suppose z 2 S⇤k

\ L+

k

for some k. Then, consider S =

(S1

, . . . , SK

) such that Sk

0 = S⇤k

0 for all k0 6= k and Sk

= S⇤k

[ {z}. Let M = S1

⇥ · · ·⇥SK

. It follows from the

Page 43: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

e-companion to Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel ec7

definitions that M �M⇤. Because G(·) is an increasing set function, we must therefore have that G(M)>

G(M⇤). Furthermore,

F (M) = f(D(S)) = f

0

@

2

4dk

+ �kz

+X

`2S

⇤k

�k`

3

5

Y

k

0 6=k

Dk

0(Sk

0)

1

A> f

0

@

2

4dk

+X

`2S

⇤k

�k`

3

5

Y

k

0 6=k

Dk

0(S⇤k

0)

1

A= F (M⇤),

where the inequality follows because �kz

> 0 and f(·) is increasing.

We have thus shown that (1�↵)F (M)+↵G(M)> (1�↵)F (M⇤)+↵G(M⇤), contradicting the fact that

M⇤ is the optimal solution.

The result of the theorem now follows.

Proof of Theorem 3.6 In order to prove hardness, we consider the following instance of the decision

problem. Suppose that |Lk

| = 2 and won

k1

= wo↵

k1

= 0. To simplify notation, let won

k

:= won

k2

and wo↵

k

:= wo↵

k2

.

Let A✓ {1,2, . . . ,K} denote the subset of attributes for which level 2 is o↵ered in the o✏ine store. For any

assortment M , let A(M) denote the subset {1 kK : xk

= 2 for some x2M}. We must now have

X

x2X

Px

(M) =

P

x2X exp⇣

P

k2A(M)

wo↵

k

1l2

[xk

] +P

k/2A(M)

won

k

1l2

[xk

]⌘

1+P

x2X exp⇣

P

k2A(M)

wo↵

k

1l2

[xk

] +P

k/2A(M)

won

k

1l2

[xk

]⌘ =

Q

k2A(M)

(1+ ewoff

k )Q

k/2A(M)

(1+ ewon

k )

1+Q

k2A(M)

(1+ ewoff

k )Q

k/2A(M)

(1+ ewon

k ),

where the second equality follows from the arguments similar to Lemma 3.1. Because the expected sales

only depend on the subset of attributes A(M) and not on M , it is su�cient to search over subsets A of

{1,2, . . . ,K}.

We now focus on the expected sales from the o✏ine channel. It follows from the arguments in the proof

of Theorem 3.5 that for the problem instance under consideration, the assortment decision reduces to

finding the subset of attributes A ✓ {1,2, . . . ,K} that maximizes expected sales. More precisely, for any

subset of attributes A, the expected sales for the o✏ine segment are maximized by the o↵er set M(A) :=

{x2X : xk

= 1 for all k /2A}. It now follows that for any subset A,

X

x2M(A)

Qx

=

P

x2X : xk=1 8 k/2A

exp�

P

k

1l2

[xk

]wo↵

k

1+P

x2X : xk=1 8 k/2A

exp (P

k

1l2

[xk

]wo↵

k

)=

Q

k2A

(1+ ewoff

k )

1+Q

k2A

(1+ ewoff

k ).

Note that our definitions are consistent: A(M(A)) =A. With the above simplification, the decision problem

reduces to: is there a subset A✓ {1,2, . . . ,K} such that

Q

k2A

(1+ ewoff

k )

1+Q

k2A

(1+ ewoff

k )+ (1�↵)

Q

k2A

(1+ ewoff

k )Q

k/2A

(1+ ewon

k )

1+Q

k2A

(1+ ewoff

k )Q

k/2A

(1+ ewon

k )�Q?

Now suppose that log(1+ewon

k ) = 2 log(1+ewoff

k ). Then, we can further simplify the above problem as follows.

Letting vk

denote log(1+ ewoff

k ), we have log(1+ ewon

k ) = 2vk

. Further, let V denoteP

k

vk

. Then, we have

Y

k2A

(1+ ewoff

k ) = exp

X

k2A

log(1+ ewoff

k )

!

= exp

X

k2A

vk

!

= exp

V �X

k/2A

vk

!

.

In a similar fashion, we can write

Y

k2A

(1+ewoff

k )Y

k/2A

(1+ewon

k ) = exp

X

k2A

log(1+ ewoff

k )+X

k/2A

log(1+ ewon

k )

!

= exp

X

k2A

vk

+2X

k/2A

vk

!

= exp

V +X

k/2A

vk

!

.

Page 44: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

ec8 e-companion to Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel

Putting everything together and subtracting both sides of the inequality from 1, our goal is to solve the

following decision problem: is there a subset S ✓ {1,2, . . . ,K} such that

1+ exp�

V �P

k/2A

vk

� +1�↵

1+ exp�

V +P

k/2A

vk

� 1�Q?

The reduction is from the Partition problem, defined below.

Partition

Inputs: The set of items indexed by 1,2, . . . ,K and sizes tk

2Z+

associated with each item.

Question: Is there a subset A✓ {1,2, . . . ,K} such thatP

k2A

tk

=P

k/2A

tk

?

We now obtain a reduction as follows. Given an instance of the Partition problem, let T�

= 1

2

P

K

k=1

tk

.

Without loss of generality, suppose that T 2Z+

. Then, there exists a subset A such thatP

k2A

tk

=P

k/2A

tk

if and only if there exists a subset A such thatP

k2A

tk

= T . Our goal is to determine if there is a subset A

such thatP

k2A

tk

= T .

We create an instance of our decision problem now. Given tk

and T , define: vk

= tk

for all k, the target sales

function 1�Q= ae

T+b

a

2

+b

2

, and ↵= a2/(a2 + b2), where a= (1+ eT )eT and b= 1+ e3T . With this assignment,

note that V =P

k

vk

=P

k

tk

= 2T . Now, the Partition problem has a solution if and only if there exists a

subset A such that

1+ exp�

V �P

k/2A

vk

� +1�↵

1+ exp�

V +P

k/2A

vk

� minz2[0,2T ]

1+ exp(2T � z)+

1�↵

1+ exp(2T + z)

=↵

1+ eT+

1�↵

1+ e3T

=aeT + b

a2 + b2

=1�Q,

where the first equality follows from the fact that G(z)�

= ↵

1+exp(V �z)

+ 1�↵

1+exp(V+z)

is convex in z over [0,2T ]

and has a unique minimum at z = T and the second equality follows from algebra noting the definition of

↵. It thus follows that the answer to the Partition problem is yes if and only if the answer to our decision

problem is yes. As a result, solving our decision problem is at least as hard as solving the Partition problem.

The result of the theorem now follows.

Proof of Lemma 3.3 The simplification of the part of the profit function corresponding to the online

segment is the same as that of Lemma 3.1. So we focus on the part of the profit function corresponding to the

o✏ine segment. The expected profit from the o✏ine segment from o↵ering the assortment M = S1

⇥ · · ·⇥SK

is given by

P

x2M

px

exp�

P

K

k=1

uo↵

k

(xk

)�

1+P

x2M

exp�

P

K

k=1

uo↵

k

(xk

)� =

P

x2M

P

k,`

rk,`

1l`

[xk

]⌘

exp�

P

K

k=1

uo↵

k

(xk

)�

1+P

x2M

exp�

P

K

k=1

uo↵

k

(xk

)�

Focusing on the denominator, we can write

X

x2M

exp

K

X

k=1

uo↵

k

(xk

)

!

=X

x2S

1

⇥···⇥Sk

K

Y

k=1

euoff

k (xk) =K

Y

k=1

X

`2Sk

ewoff

k`

!

=K

Y

k=1

X

`2Sk

�o↵k`

!

,

Page 45: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

e-companion to Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel ec9

where the second equality follows from an interchange of sum and product and the fact that uo↵

k

(xk

) =wo↵

k,xk

and the last equality follows from the definition of �o↵k`

.

In a similar fashion, we can simplify the numerator by collecting the terms multiplying rk`

to obtain

euoff

k (`)

X

x

�k2S

1

⇥···Sk�1

⇥Sk+1

⇥···SK

Y

k

0 6=k

euoff

k0 (xk0 ) = ewoff

k`

Y

k

0 6=k

0

@

X

`

02Sk0

ewoff

k0`0

1

A=ew

off

k`

P

`

02Skew

off

k`0·

K

Y

k

0=1

0

@

X

`

02Sk0

�o↵k

0`

0

1

A ,

where the first equality follows from an interchange of sum and product and the second equality follows from

the definition of �o↵k

0`

0 . We use x

�k to denote the vector x with the kth component removed.

Putting everything together, we obtain that the expected profit is given by

P

K

k=1

P

`2Skrk`

ewoff

k`

/⇣

P

`2Skew

off

k`

1+1/h

Q

K

k=1

P

`2Sk�o↵k`

⌘i =

P

K

k=1

Ro↵

k

(Sk

)

1+ 1/Do↵(S),

where we used the notation Do↵(S) =Q

K

k=1

Do↵

k

(Sk

) with Do↵

k

(Sk

) :=P

`2Sk�k`

and Ro↵

k

(Sk

) =⇣

P

`2Skrk`

ewoff

k`

/⇣

P

`2Skew

off

k`

.

The result of the lemma now follows.

Proof of Lemma 3.4 The proof is similar to that of Lemma 3.2. Particularly, suppose that S 2 2L1

⇥···⇥LK

is such thatX

c2{on,o↵}

↵c

P

k

Rc

k

(Sk

)

1+ e�t

c

� q andK

X

k=1

logDc

k

(Sk

)� tc

, c2 {on,o↵} ,

We claim that S satisfies the constraintP

K

k=1

jcSk,k

� ⇢ for c2 {on,o↵}. To see this, consider

K

X

k=1

jcSk,k

�K

X

k=1

logDc

k

(Sk

)

("tc

/K)� 1

=1

"tc

/K

K

X

k=1

logDc

k

(Sk

)�K � tc

"tc

/K�K � bK/"c�K = ⇢,

where the first inequality follows from the fact that bxc � x� 1 for any real number x. We have thus shown

that S is a feasible solution to the optimization problem corresponding to the value function V (K,⇢,⇢).

Because the DP maximizesP

c2{on,o↵}↵c

Pk R

c

k(˜

Sk)

1+e

�tc

over all S that satisfyP

K

k=1

jc˜

Sk,k� ⇢ for c2 {on,o↵}, it

must be that that output S of the DP satisfies

X

c2{on,o↵}

↵c

P

k

Rc

k

(Sk

)

1+ e�t

c

�X

c2{on,o↵}

↵c

P

k

Rc

k

(S⇤k

)

1+ e�t

c

� q.

We are now left only to prove thatP

K

k=1

logDc

k

(Sk

)� (1� 2")tc

for c 2 {on,o↵}. To prove this, we observe

that

⇢K

X

k=1

jcˆ

Sk,k

K

X

k=1

logDc

k

(Sk

)

"tc

/K.

As a result,

K

X

k=1

logDc

k

(Sk

)� "tc

K(bK/"c�K)� "t

c

K(K/"� 1�K) = t

c

(1� "/K � ")� tc

(1� 2"),

where the second inequality follows from the fact that bxc � x� 1 and the last inequality follows from the

fact that K � 1.

The result of the lemma now follows.

Page 46: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

ec10 e-companion to Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel

Proof of Theorem 3.7 We proceed as follows. For c 2 {on,o↵}, let Tc

denote the " grid of the interval

[⌧c,min

, ⌧c,max

] such that Tc

= {⌧c,min

(1+ ")j : j = 0,1, . . . , Jc

} where Jc

=O (log(⌧c,max

/⌧c,min

)/ log(1+ ")). We

carry out the DP algorithm with tc

=� log ⌧c

for every grid point ⌧ = (⌧on

, ⌧o↵

) 2 Ton

⇥ To↵

. Let S⌧

denote

the solution obtained from solving the DP for grid point ⌧ and let S denote the subset from the collectionn

S

: ⌧ 2 Ton

⇥ To↵

o

with maximum expected profit. We now show that S is a good approximation of the

optimal solution.

For that, let S⇤ denote the profit maximizing solution and r⇤ = ↵onr⇤on

+↵o↵r⇤o↵

the optimal profit, where

r⇤on

and r⇤o↵

are, respectively, the expected profits from the online and o✏ine segments of customers. Define

t⇤c

=P

K

k=1

logDc

k

(S⇤k

) for c 2 {on,o↵}. Now consider the grid point ⌧ 2 Ton

⇥ To↵

such that ⌧c

/(1 + ") e�t

⇤c ⌧

c

. Now note that

X

c2{on,o↵}

↵c

P

K

k=1

Rc

k

(S⇤k

)

1+ exp�

�P

K

k=1

logDc

k

(S⇤k

)� =

X

c2{on,o↵}

↵c

P

K

k=1

Rc

k

(S⇤k

)

1+ e�t

⇤c

�X

c2{on,o↵}

↵c

P

K

k=1

Rc

k

(S⇤k

)

1+ elog ⌧

c

=X

c2{on,o↵}

↵cr⇤c

· 1+ e�t

⇤c

1+ elog ⌧

c

,

where the inequality follows from the definition of ⌧c

and the last equality follows from our definition of

r⇤c

. Furthermore, note thatP

K

k=1

logDc

k

(S⇤k

) = t⇤c

� � log ⌧c

. Therefore, by invoking Lemma 3.4 with q =P

c2{on,o↵}↵cr⇤

c

· 1+e

�t⇤c

1+e

log ⌧c

, tc

=� log ⌧c

, and " replaced by "/ |log ⌧⇤|, where ⌧⇤ =min{⌧on,min

, ⌧o↵,min

}, it followsthat the DP outputs the solution S

such that

X

c2{on,o↵}

↵c

P

K

k=1

Rc

k

(S⌧ ,k

)

1+ e�t

c

� q andK

X

k=1

logDc

k

(S⌧ ,k

)��(1� 2"/ |log ⌧⇤|) log ⌧c

, c2 {on,o↵} . (EC.6)

Using (EC.6), we now show that S provides a good approximation for S⇤. For that, note that

1+ e�t

c

1+ e�PK

k=1

logD

c

k(

ˆ

S⌧,k)� 1+ e�t

c

1+ e(1�2"/|log ⌧

⇤|) log ⌧

c

=1+ e�t

c

1+ e�(1�2"/|log ⌧

⇤|)tc

.

Now consider

e�(1�2"/|log ⌧

min

|)tc = e�t

ce2"tc/|log ⌧

⇤| e�t

ce2" (1+ 7")e�t

⇤c ,

where the first inequality follows from the fact that tc

|log ⌧⇤| and the last inequality follows from verifying

that e2" 1+7" for 0 " 1. Therefore, we must have

1+ e�t

c

1+ e�PK

k=1

logD

c

k(

ˆ

S⌧,k)� 1+ e�t

c

1+ (1+7")e�t

c

� 1

1+7".

We can now write

X

c2{on,o↵}

↵c

P

K

k=1

Rc

k

(S⌧ ,k

)

1+ e�PK

k=1

logD

c

k(

ˆ

S⌧,k)=

X

c2{on,o↵}

↵c

P

K

k=1

Rc

k

(S⌧ ,k

)

1+ e�t

c

· 1+ e�t

c

1+ e�PK

k=1

logD

c

k(

ˆ

S⌧,k)

� 1

1+7"·

X

c2{on,o↵}

↵c

P

K

k=1

Rc

k

(S⌧ ,k

)

1+ e�t

c

� q

1+7",

where the last inequality follows from (EC.6). Now consider

1+ e�t

⇤c

1+ elog ⌧

c

=1+ e�t

⇤c

1+ ⌧c

� 1+ e�t

⇤c

1+ (1+ ")e�t

⇤c

� 1

1+ ",

Page 47: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

e-companion to Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel ec11

where the first inequality follows from our choice of ⌧c

so that ⌧c

(1+ ")e�t

⇤c . It now follows that

q=X

c2{on,o↵}

↵cr⇤c

1+ e�t

⇤c

1+ elog ⌧

c

� 1

1+ "

X

c2{on,o↵}

↵cr⇤c

=r⇤

1+ ".

Putting everything together, we now have

X

c2{on,o↵}

↵c

P

K

k=1

Rc

k

(S⌧ ,k

)

1+ e�PK

k=1

logD

c

k(

ˆ

S⌧,k)� q

1+7"� r⇤

(1+ 7")(1+ ")� (1� 8")r⇤.

Because the algorithm outputs the solution S that maximizes the expected profit over all possible grid points,

we can conclude that

X

c2{on,o↵}

↵c

P

K

k=1

Rc

k

(Sk

)

1+ e�PK

k=1

logD

c

k(

ˆ

Sk)�

X

c2{on,o↵}

↵c

P

K

k=1

Rc

k

(S⌧ ,k

)

1+ e�PK

k=1

logD

c

k(

ˆ

S⌧,k)� (1� 8")r⇤.

Running time. In order to determine the running time, note that it follows from the discussion before the

statement of Lemma 3.4 that for each grid point, the running time of the DP is O�

K3maxk

2|Lk| log2 ⌧⇤/"2�

because the DP is run with " replaced by "/ |log ⌧⇤|. Now because the DP is run for each grid point and

there is a total of O(log(⌧on,max

/⌧on,min

) log(⌧o↵,max

/⌧o↵,min

)/ log2(1+ ")) grid points, the total running time

of the algorithm is O�

K3maxk

2|Lk| log2 ⌧⇤ log(⌧on,max

/⌧on,min

) log(⌧o↵,max

/⌧o↵,min

)/("2 log2(1+ "))�

.

The result of the theorem now follows.

Appendix B: Details of the benchmark methods implemented

The o✏ine heuristic. As mentioned, the o✏ine heuristic ignores the impact of the online channel and

optimizes only the o✏ine channel. As a result, the problem reduces to determining the profit maximizing

subset of size at most C under the MNL model when the retailer sells through only one channel. This problem

can be solved in O(NC) iterations (Rusmevichientong et al. 2010), where N is the number of products in

the universe. Because N can be exponentially large in our setting, we use the following heuristic to solve the

problem in the attribute space. Recent work by Gallego et al. (2016) has shown that when the profit from a

product decomposes in an additive fashion into the profits of the constituent attribute levels, the problem of

finding the m products with the highest profits reduces to the problem of finding the m shortest (minimum

cost) paths from a single source in an appropriately defined directed acyclic graph (DAG) with O(KL) nodes

and O(KL2) directed edges. The single-source m-shortest path problem is well-studied and can be solved in

a polynomial (in K and L) time using Yen’s algorithm (Yen 1971). To find the profit maximizing subset of

size at most C, we implemented Yen’s algorithm to find the C most profitable products {x1

,x2

, . . . ,xC

} such

that px

1

� px

2

� · · ·� pxC

. We then searched through all subsets of the form {x1

,x2

, . . . ,xm

} for 1mC

and chose the subset that maximizes the single channel profit. Note that our heuristic returns the optimal

solution if the profit maximizing subset without the capacity constraint has size at most C because it is

known that the unconstrained profit maximizing subset comprises the m most profitable products, for some

m.

Greedy heuristic. The greedy heuristic is another general-purpose heuristic commonly applied to assort-

ment optimization problems (Jagabathula 2014). The existing heuristics typically operate in the product

space. Because the product space is combinatorially large, we used the following natural extension of the

Page 48: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

ec12 e-companion to Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel

standard greedy heuristic to the attribute space. Each iteration starts with an assortment M of the form

Y1

⇥ · · ·⇥YK

, where Yk

✓Lk

denotes the subset of levels in attribute k that are present in the assortment.

The algorithm then searches over all assortments of the form Mk`

:= Y1

⇥ · · ·⇥ Yk

[ {`}⇥ · · ·⇥ YK

, where

` 2 Lk

\ Yk

and 1 k K, and chooses the subset Mk

⇤`

⇤ with the maximum profit. If R(Mk

⇤`

⇤) R(M)

or |Mk

⇤`

⇤ |>C, then the algorithm terminates with M as the solution. Otherwise, it continues to the next

iteration with Mk

⇤`

⇤ as the starting assortment. We ran the algorithm with the most profitable product

(determined using Yen’s algorithm, as described above) as the initial solution.

Appendix C: Omitted details of the Timbuk2 case study

C.1. Details of the conjoint study

Study procedure: description of the tasks. We invited participants to complete a two-part task: a

web-based conjoint survey, and a paper-and-pencil survey providing evaluations of physical products. Both

parts were ratings-based conjoint tasks, in which participants rate each bag individually with respect to

how likely they would be to purchase the bag. A five point scale was used to rate the bags: Definitely not

buy, Probably not buy, May or may not buy, Probably buy, Definitely buy. We chose to conduct a ratings-

based conjoint rather the more common choice-based conjoint. In a choice-based conjoint, participants are

presented with a series of choice sets and asked to choose one product from each set. Conducting such a

choice-based conjoint is logistically much harder when the choice tasks involve evaluating physical products.

Hence, for the purposes of logistical simplicity, we carried our the ratings-based conjoint study.

Online task. The web-based conjoint was conducted using Sawtooth Softwares CVA tool. The top image in

Figure EC.1 provides a screen shot of the task. Of the six features, five were represented with text, and one,

Exterior design, was represented with an image.

The task proceeded as follows:

1. The experimenter informed the respondent that there would be two parts, one on the computer, one

in an adjacent room with paper-and-pencil, and described the incentive-aligned prize lottery.

2. Initial screens ensured privacy and described the basic study.

3. The next six screens introduced the features one at a time and included a brief description of each

feature.

4. Participants rated one bag as a warm up exercise. They were informed that this response would be

discarded.

5. Participants then provided ratings for the 20 bags on the five-point scale described above.

Paper and Pencil Study. After completing the online study, participants were escorted to a di↵erent room,

where the same set of 20 bags were laid out on a conference room table, as shown in Figure EC.2. The prices

were displayed on stickers on a tag attached to the bag. Each bag had an index card next to it displaying

a number indexing the bag, and the bags were laid out in order 1 through 20. All participants saw the

bags in the same order. The experimenter walked the participant through all the bag features, showing each

feature on a sample bag. Participants were then asked to complete the paper- and-pencil survey, in which

Page 49: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

e-companion to Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel ec13

Figure EC.1 Screenshot of a one of the 20 questions from the online survey (left) and paper-and-pencil

questionnaire that was given to participants during physical evaluation (right).

they provided ratings for each of the bags (see Figure EC.2). The experimenter asked them to take their

time and examine all the bags and rate them with respect to how likely they would be to purchase such a

bag. Participants were also reminded of the incentive aligned lottery.

Figure EC.2 Photo of the actual task faced by respondents. The 20 bags were laid out on a conference room

table and labeled 1 through 20. Participants could look at each bag and provided their evaluations on a paper and

pencil survey.

Configurations of the bags included in the study. Table EC.1 presents the details of the Timbuk2

messenger bags that were included in the conjoint study.

Partworths for the study when participants evaluate bags o✏ine followed by online. Table EC.2

is the counterpart of Table 3 for the smaller study with 20 participants who did the conjoint study in the

reverse order: first the o✏ine task, followed by the online task.

Page 50: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

ec14 e-companion to Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel

C.2. Fit of the logit model to the conjoint data

To test the fit of the logit model, we conducted a five-fold cross-validation on both the online and o✏ine

conjoint data. We describe our testing procedure for the online data-set; the procedure is the same for the

o✏ine data-set.

Let yi

denote the ratings vector of participant i so that yij

is the rating assigned by participant i to product

j, with 1 i I and 1 j n. In our case, I = 122 and n= 20. In order to carry out k-fold cross-validation

(with k= 5), we randomly partitioned the set of participants into k segments of (about) equal size. We chose

four segments to be the training and the remaining segment to be the test (hold-out) data. Let Itraining

and

Itest

denote the set of participants that are part of training and test data-sets.

We trained the MNL model on the training data as follows. For each participant i 2 Itraining

we con-

verted the ratings vector y

i

into the pairwise comparisons vector z

i

, defined as zi,jj0 = 1 if y

ij

> yij

0 or

yij

= yij

0 and yj

� yj

0 and 0 otherwise, where yj

is the average rating of product j, defined as yj

=⇣

P

i2Itraining

yij

/ |Itraining

|. In other words, we conclude that participant i prefers higher rated products to

lower rated products, with ties broken according to the population rating. Each comparisons vector zi

corre-

Task Exterior Design Size Price Strap Pad Water bottle pocket Interior pocket

1 Black Small $160 No Yes Empty

2 Blue Small $140 Yes Yes Divider

3 Colorful Small $120 Yes No Divider

4 Reflective Small $160 Yes No Empty

5 Colorful Large $160 No Yes Laptop Compt

6 Reflective Small $140 No No Laptop Compt

7 Colorful Large $160 No No Divider

8 Blue Small $120 No No Laptop Compt

9 Black Large $120 Yes No Divider

10 Colorful Large $140 No No Empty

11 Colorful Small $180 Yes Yes Laptop Compt

12 Blue Small $160 No Yes Divider

13 Colorful Small $120 Yes Yes Empty

14 Black Small $180 No No Laptop Compt

15 Black Large $120 Yes Yes Laptop Compt

16 Reflective Large $180 No Yes Divider

17 Reflective Large $120 Yes Yes Laptop Compt

18 Blue Large $180 Yes No Empty

19 Blue Large $120 No Yes Empty

20 Blue Large $160 Yes No Laptop Compt

Table EC.1 The list 20 messenger bags that were included in the conjoint study.

Page 51: O ine Assortment Optimization in the Presence of an …people.stern.nyu.edu/ddzyabur/index_files/OfflineAssortment.pdf · Dzyabura and Jagabathula: O ine Assortment Optimization in

e-companion to Dzyabura and Jagabathula: O✏ine Assortment Optimization in the Presence of an Online Channel ec15

Attribute Level Online (won) O✏ine (wo↵) Di↵erence

Exterior design Reflective �0.38 �0.29 0.09

Colorful �0.28 �0.09 +0.19

Blue �0.03 0.14 +0.17

Black

Size Large �0.13 �0.17 �0.05

Small

Price $120, $140, $160, $180 �0.009 ⇤⇤ �0.005 +0.004

Strap pad Yes 0.58 ⇤⇤ 0.37 �0.21

No

Water bottle pocket Yes 0.32 ⇤ �0.05 �0.37

No

Interior compartments Divider for files 0.65 ⇤⇤ 0.52 ⇤⇤ �0.13

Crater laptop sleeve 1.05 ⇤⇤ 1.03 ⇤⇤ �0.03

Empty bucket/no dividers

Intercept 2.78 ⇤⇤ 2.33 ⇤⇤ �0.45

Notes: ⇤⇤p < 0.001, ⇤p < 0.01Table EC.2 The online and o✏ine partworths for the smaller study in which 20 participants completed the

o✏ine task first, followed by the online task. The levels with no coe�cients were set to zero in dummy encoding.

sponds to a total-order over all the products. We estimated the parameters of the logit model by maximizing

the log likelihood (Jagabathula and Vulcano 2015) of the observed data:

max�

X

i2Itraining

n�1

X

j=1

"

>x

j

� log

n

X

j

0=j

e�>xj0

!#

.

The above optimization problem can be shown to be a convex program and, hence, can be e�ciently solved.

Using the estimated �, we predicted the market share of the products mj

= e�>xj/

h

P

n

j

0=1

e�>xj0

i

.

We compared the predictions to the observed market shares on the hold-out data mj

=�

P

i2Itest

1l[zi,jj

0 = 1, 1 j0 n]�

/ |Itest

| using the standard mean absolute percentage error (MAPE) metric,

defined as:1

|N |X

j2N

mj

� mj

mj

,

where N = {1 j n : mj

6= 0}.

We repeated the above procedure with each of the k segments as the test data and the remaining k� 1

segments as the training data. The average of the MAPE score over the k rotations is reported.