MANUFACTURING & SERVICE

OPERATIONS MANAGEMENT informs ® Vol. 12, No. 3, Summer 2010, pp. 371–392 doi 10.1287/msom.1090.0275 issn 1523-4614 eissn 1526-5498 10 1203 0317 © 2010 INFORMS

OM Practice Choice-Based Revenue Management: An Empirical Study of Estimation and Optimization

Gustavo Vulcano Leonard N. Stern School of Business, New York University, New York, New York 10012, [email protected] Garrett van Ryzin Graduate School of Business, Columbia University, New York, New York 10027, [email protected] Wassim Chaar Research Group, Sabre Holdings, Southlake, Texas 76092, [email protected]

iscrete choice models are appealing for revenue management (RM) because they offer a means to Dprofitably exploit preferences for attributes such as time of day, routing, brand, and price. They are also good at modeling demand for unrestricted fare class structures, which are widespread throughout the industry. However, there is little empirical research on the practicality and effectiveness of choice-based RMmodels. Toward this end, we report the results of a study of choice-based RMconducted with a major U.S. airline. Our study had two main objectives: (1) to assess the extent to which choice models can be estimated well using readily available airline data, and (2) to gauge the potential impact that choice-based RMcould have on a sample of test markets. We developed a maximum likelihood estimation algorithm that uses a variation of the expectation- maximization method to account for unobservable data. The procedure was applied to data for a test market from New York City to a destination in Florida. The outputs are promising in terms of the quality of the com- puted estimates, although a large number of departure instances may be necessary to achieve highly accurate results. These choice model estimates were then used in a simulation study to assess the revenue performance of the EMSR-b (expected marginal seat revenue, version b) capacity control policies and the current controls used by the airline relative to controls optimized to account for choice behavior. Our simulation results show 1%–5% average revenue improvements using choice-based RM. Although such simulated results must be taken with caution, overall our study suggests that choice-based revenue management is both feasible to execute and economically significant in real-world airline environments. Key words: choice behavior; multinomial logit; EMmethod; maximum likelihood; capacity control History: Received: February 2, 2008; accepted: July 8, 2009. Published online in Articles in Advance November 13, 2009.

1. Introduction This so-called independent demand model does not Quantity-based revenue management (RM) involves endogenize customer behavior, such as choice behav- optimally allocating capacity of resources to demand ior and purchase-timing behavior (see Talluri and by controlling the availability of fare products. This van Ryzin 2004b for further discussion of the indepen- allocation is done dynamically as demand materi- dent demand model). Clearly, this is a somewhat unre- alizes, and with considerable uncertainty about the alistic assumption. For example, the probability of sell- quantity or composition of future requests. ing a full-fare ticket may very well depend on whether Traditionally, both researchers and practitioners a discount fare is available at the same time, the prob- have made the assumption that demand for each ability that a customer buys at all may depend on the fare class is an independent stochastic process that lowest available fare, etc. When a customer buys a is not influenced by the firm’s availability controls. higher fare when a discount class is closed it is called

371 Vulcano, van Ryzin, and Chaar: OM Practice 372 Manufacturing & Service Operations Management 12(3), pp. 371–392, © 2010 INFORMS a buy-up; when she chooses another flight from the algorithm of Miranda Bront et al. (2009), and the DP same carrier when a discount class is closed, it is called decomposition scheme of Kunnumkal and Topaloglu diversion. When a customer willing to pay a high fare (2009) also belong to this line of research. instead buys a low-fare ticket because it is available, it However, there is little empirical understanding of is called a buy-down. These behaviors have important how choice behavior impacts airline RM. That is, revenue consequences. how significant is choice behavior in real airline mar- Such choice behavior is well recognized in RMprac- kets? Can it be estimated well using available data? tice, and several ad hoc corrections to the independent And what do real-world estimates of choice behav- demand model have been proposed to capture some ior have to say about the potential revenue improve- of these effects (e.g., see Belobaba 1987a, b, Belobaba ments from using choice-based RM? Exceptions are 1989; Belobaba and Weatherford 1996). The success the work of Andersson (1989, 1998) and Algers and of low-cost offering simplified, undifferenti- Besser (2001), who reported development efforts at ated fare structures has rekindled interest in customer SAS to apply logit choice models to estimate buy-up choice models, because with minimal rules and restric- and recapture factors1 at one of their hubs. Work close tions separating fare classes (i.e., in so-called “fence- to ours is that of Ratliff et al. (2008), who proposed less” fare environments), customers have considerable a heuristic methodology for estimating recapture and latitude to buy up, buy down, or divert. Moreover, demand untruncation for parallel flights in the same point-to-point operators often have many flight depar- market (e.g., different departure times between the tures per day in a given origin-destination (O-D) mar- same O-D pair). ket, so customers have a variety of departures at dif- Here, we report the results of a research study ferent price levels from which to choose. RMdecisions conducted in collaboration with a major U.S. commer- can potentially be improved by properly accounting cial airline. The focus of the study was on O-D markets for this environment of flexibility and choice. between the New York metropolitan area and various Choice-based RMhas been an active research area in Florida. The study had two main objectives: of late. Belobaba and Hopperstad (1999) conducted the first objective was to assess the feasibility of esti- simulation studies to understand the impact that mating choice behavior from readily available airline passenger choice behavior has on traditional RM data; the second objective was to use these estimates of choice behavior to assess the potential improvement methods. Talluri and van Ryzin (2004a) provided of using choice-based optimization methods in the tar- an exact analysis of the optimal control policy for geted test markets. a single-leg model of RMunder a general discrete Toward these ends, in a first phase we developed a choice model of demand. Zhang and Cooper (2006) maximum likelihood estimation (MLE) algorithm for analyzed choice among parallel flights in the same inferring customer choice behavior from an airline’s market (e.g., different departure times between the available operational data—namely, capacity avail- same O-D pair). Their model assumes that the cus- ability, revenue accounting, and flight schedule data. tomer chooses within the same fare class among The first challenge here is constructing relevant different flights but not between fare classes. They choice models. Given a booking instance, one must developed bounds and approximations to the result- infer which alternatives the customer was consider- ing dynamic program. Boyd and Kallesen (2004) ing when making their purchase decision and the illustrated the effect of considering demand models relevant attributes that influenced that customer’s for price-sensitive customers in a single-leg setting, decision. This requires determining the consideration where customers are price sensitive and not perfectly set of flight alternatives, the relevant attributes of segmented, and therefore may buy down. Gallego each alternative, and how these attributes are to be et al. (2004) and Liu and van Ryzin (2008) studied a parameterized. deterministic network RMproblem using a customer- choice-based linear programming model. The recent 1 Recapture is the amount of demand that is retained by the firm’s dynamic programming (DP) approximation proposal substitute products when a fare class is closed down or the cabin of Zhang and Adelman (2009), the column generation is sold out. Vulcano, van Ryzin, and Chaar: OM Practice Manufacturing & Service Operations Management 12(3), pp. 371–392, © 2010 INFORMS 373

The second problem of estimating customer choice the real airline markets that we analyzed. This was behavior is the available data. Airlines only observe due mainly to the more limited number of departure bookings and not shopping behavior; i.e., they only instances in the real data set. Overall, the results indi- record the outcomes of customers who decide to cate that with sufficient data, choice behavior can be buy.2 This means it is impossible to distinguish a reasonably accurately estimated from readily avail- small time period without an arrival from a period able airline data. in which there was an arrival but the arriving cus- The second phase of the project involved conduct- tomer did not purchase from our airline (i.e., the ing a simulation study based on the estimated choice customer purchased from a competitor or did not models. Here, we compared the expected revenue purchase at all). One must therefore infer the real, obtainable from the RMcontrols used by the air- uncensored volume of potential customers that the line and assessed the potential revenue improvements airline was facing in each market using only pur- that could be achieved by optimizing to account chase data. Ignoring this censoring can cause a severe for choice behavior effects. To measure this impact, bias in estimation. Moreover, with this incompleteness we simulated demand under the estimated choice in the data, the standard MLE procedure is diffi- behavior parameters and then processed that demand cult because of the complexity of the log-likelihood under the RMcapacity controls used by the airline, function. To circumvent this problem and account for EMSR-b (expected marginal seat revenue, version b) unobservable no-purchase data, we used a variation controls that we calculated based on the indepen- of the expectation-maximization (EM) method (see dent demand assumption, and an alternative set of Dempster et al. 1977, or the textbook by McLachlan controls computed using our previous algorithmic and Krishnan 1996). The EMprocedure we devel- work on simulation-based optimization, which explic- oped is an implementation of the generic framework itly accounts for choice behavior (see van Ryzin and described in Talluri and van Ryzin (2004a, §5).3 Vulcano 2008). Our simulation study showed that To test our estimation method, we first applied it to the benefits obtained from optimizing RMcontrols to a set of simulated data. This step was taken to assess account for choice behavior are significant: the aver- the quality of the estimation procedure in an environ- age revenue gains ranged from 1.4% to 5.3% in the ment where the volume of data and true underlying markets we tested. Although in some cases the con- fidence interval (CI) for the revenue improvements parameters of the choice behavior are controllable and did include zero, the range was always strongly on known. We then applied the estimation procedure to the positive side. We also tested the robustness of our airline data set. We computed specific estimates these results by perturbing the parameters to mimic of customer preferences for price, arrival time, and varying degrees of estimation error. The results show departure day. The accuracy of our estimation was that the gains in revenue can be affected by estima- validated through the calculation of asymptotic stan- tion accuracy, particularly in cases where the value dard errors and 2 goodness-of-fit tests between the added is relatively low (e.g., around 1%). For bigger observed number of bookings and the expected num- revenue gap cases, the revenue benefits are preserved ber of bookings predicted by our model. The results even when parameter estimates have errors on the of the tests showed the method was quite accurate order of 25%. Although these revenue improvements for the simulated data cases when we generated large and robustness studies are only simulation estimates, volumes of booking records. The quality of the esti- they are based on models fit to real-world data and mates was acceptable, though of lower accuracy, in hence give some sense of the potential performance of choice-based RM. 2 Although the airline’s Web data provide more complete shopping In summary, our analysis shows that choice behav- information, direct sales through the company’s own website only ior and customer preferences for price, flight time, represent a small fraction of total sales. For this reason, we focused exclusively on bookings data. and departure date can be estimated relatively well 3 The EMmethod has also been used in other RMcontexts, in par- from readily available airline data, although the qual- ticular by McGill (1995), to estimate multivariate normal demand ity of these estimates may vary across different mar- data with censoring. kets. The simulated benefit of using choice-based RM Vulcano, van Ryzin, and Chaar: OM Practice 374 Manufacturing & Service Operations Management 12(3), pp. 371–392, © 2010 INFORMS methods relative to incumbent, independent-demand The probability that an individual n selects alterna-

RMmethods appears quite significant. Together, these tive i from the set Cn ∪ 0 is given by results suggest choice-based RMis both a practi- cally feasible and economically significant improve- Pn i =  Uin ≥ Ujn ∀ j ∈ Cn ∪ 0  (3) ment over current airline RMpractice. 2.2. The MNL Model The remainder of this paper is organized as follows: The most common random utility model, widely used For completeness, we begin in §2 by reviewing some in the economics and marketing, is the MNL model general background on choice models and the multi- (see Ben-Akiva and Lerman 1994, Chapter 5). It is nomial logit (MNL) model used in our study. In §3 we derived by assuming that the  s in the utility func- specialize the choice model for an airline O-D market in tions are independent and identically distributed ran- and explain our construction of choice sets and how dom variables with a Gumbel (or double-exponential) we parameterized the relevant flight attributes. The distribution having cumulative distribution function EMestimation algorithm is described in §4. In §5 we describe the simulation study to assess the RMcon- F x=  in ≤ x = exp − exp − x −  sequences of embedding choice behavior in capacity control decisions. Finally, we present our concluding where  is the location parameter (mode) and  is a remarks in §6. positive scale parameter. The mean and variance of in, respectively, are 2. Choice Modeling and Estimation 2 E  =  + / and Var  = (4) We first provide an overview of discrete-choice mod- in in 62 els and then discuss in detail the parametric choice model used in our study. where  ≈ 31416 and  is the Euler constant (i.e.,  ≈ 05772). The Gumbel distribution is used because 2.1. General Choice Models it is closed under maximization and hence can be Choice behavior can be modeled by assuming that applied to (3) (see Gumbel 1958). customers are utility maximizers and individual cus- For the MNL model, the probability that customer n tomer utilities for alternatives are random variables. chooses alternative i ∈ Cn is given by Specifically, consider a set of alternatives offered by evin a firm, denoted C. Each customer n has a choice P i =  n vjn j∈C e + 1 (or consideration) set Cn ⊂ C. We denote by “0” the n no-purchase alternative, which is normalized to have where the one in the denominator stands for the zero utility of zero. Let U be the utility of customer n for v0n in no-purchase utility (i.e., v0n = 0) leading to e = 1. alternative i ∈ Cn. Without loss of generality, we can Note that in the case of linear-in-parameters utili- decompose this utility into two parts, a representa- ties (Equation (2)),  cannot be distinguished from the tive component vin that is deterministic (also called overall scale of . Thus, for convenience, we generally the nominal or expected utility) and a random com- make an arbitrary assumption that  = 1. Although ponent in. This leads to a utility function this is operationally necessary, from (4) we have to keep in mind that we are implicitly assuming equal Uin = vin + in (1) variance random utilities. For future reference, we set The representative component vin is often modeled as a linear-in-parameters combination of observable eTxin Pn i =  i∈ Cn (5) attributes, eTxjn + 1 j∈Cn v = Tx (2) in in We let zero denote the set of null alternatives, and where  is an unknown vector of parameters (to be let P 0 be the probability that customer n does not estimated), and x is a vector of attributes (explana- n in purchase from our airline; i.e., tory deterministic values) of customer n for alterna- tive flight i such as price paid, arrival time, departure 1 Pn 0 =   (6) day, etc. eTxjn + 1 j∈Cn Vulcano, van Ryzin, and Chaar: OM Practice Manufacturing & Service Operations Management 12(3), pp. 371–392, © 2010 INFORMS 375

Section A1 in the online appendix provides a brief dates during the booking horizon. Typically, there is description of a standard approach for estimating daily snapshot data, or one snapshot every two days parametric choice models: maximum likelihood esti- in cases where the snapshot date is further than one mation, which is indeed the approach we followed here. week from the departure date of the flight. A fourth source of data available for our study was 3. A Choice Model for Airline screen scrape data. These are data obtained from third Markets parties that automatically sample information about We first describe the data available to airlines for esti- alternatives and fares offered by competitors at differ- mating customer choice behavior. Then, we describe ent points in time during the booking horizon. This the design decisions we took in constructing our dis- can be obtained by Web crawler programs. Similar crete choice model. data can be obtained via global distributions systems (GDSs). GDSs communicate with the host reservation 3.1. Airline Data system of each airline to periodically obtain availabil- At our sponsor firm, there were three readily avail- ity information at the bucket level and hence are in able sources of data to use in estimating cus- a position to track market-level prices and availabil- tomer choice behavior: flight schedule data, revenue ity. However, airlines must pay to acquire these GDS accounting data, and availability data. The flight data, their request is typically delayed by a month or schedule database lists the flights offered by the car- more, and even with GDS data the problem of esti- rier: flight number, origin/destination, day and time mating customer choice behavior is still a challenge. of departure and arrival, aircraft type, etc. The rev- Although we obtained screen scrape data provid- enue accounting database has one record per ticket ing samples of competitor fares for our study, these issued, i.e., a customer booking record or passenger name data turned out to be quite incomplete and “dirty.” record (PNR) in the airline industry. The relevant fields In our first attempts at model building and esti- for our study were ticket number, issue date, coupon mation, we incorporated competitor alternatives and number4 (this field should be the sequence number of attributes in our choice model and estimation proce- the coupon in the ticket), prorate (portion of the total dure to account for competitive effects. Our expec- ticket value assigned to a particular coupon, which tation was that this would improve the accuracy of henceforth will be called simply the fare), coupon the choice model. However, our preliminary results origin and destination, and flight number. We com- yielded very poor quality estimates. A single generic plete the description of the purchase with information outside alternative produced more stable estimates about the arrival time of the flight, which is taken with lower standard errors and better predictive per- from the flight schedule database. formance. Hence, we decided to omit these com- The availability data are taken from the RMsystem. petitive data. Nevertheless, we strongly believe that It describes the number of available seats for each of with more complete and clean data, explicitly incor- the demand classes (or buckets) for every flight offered porating competitive data could add precision to by the airline. There is also a field for the average the quality of estimates. A short survey with point- revenue value corresponding to each bucket.5 There ers to different applications of discrete choice mod- is one record per flight bucket for different snapshot els under competitive considerations is provided in Dubé et al. (2002). 4 A coupon represents a single flight of a multiflight itinerary.For exam- ple, a round-trip ticket based on two nonstop flights has two coupons. 3.2. Choice Model Design 5 Airline RMsystems book reservations in buckets (or fare classes). We specialized the general choice model described Each compartment (first, business, and coach) has a number of above to our airline setting. One major design deci- fare classes—typically 15 or so for coach, 1 or 2 for business, and sion was the definition of the choice sets C s. In our 1 or 2 for first. Each bucket is used to book tickets sold under n context, C includes all the alternatives from our air- different fare-basis codes. Each of these fare-basis codes has spe- n cific fares associated with them and involves somewhat different line that customer n evaluates at the moment of mak- requirements, like the number of days of advance purchase. ing her purchase decision. Without shopping data, we Vulcano, van Ryzin, and Chaar: OM Practice 376 Manufacturing & Service Operations Management 12(3), pp. 371–392, © 2010 INFORMS

could not observe Cn directly and therefore had to no-purchase outcomes and just a few will have book- make a somewhat arbitrary, albeit plausible, defini- ings occurring. From our numerical experience, this tion of the set of relevant alternatives. For our study, situation leads to bad behavior in the estimation pro- and based on our conversations with managers at the cedure. We found a good compromise was to parti- sponsor airline, we tried different definitions of Cn, tion the day into T = 140 small time periods (i.e., each including all flights offered by our airline on a given time period corresponds to a 10-minute interval). departure day, all flights offered on a given set of For the purpose of representing the arrival time of consecutive departure days, and/or all flights offered a flight, a day is split into four time slots: morning from different subsets of arriving or departing air- (between 5 a.m. and 11 a.m.), noon (between 9 a.m. ports (e.g., all New York metro airports versus JFK and 3 p.m.), afternoon (between 1 p.m. and 7 p.m.), alone). Because of the sparsity of the airline and evening (between 5 p.m. and midnight). These real data, we decided on the simpler definition of Cn intervals overlap, and hence convex weights were as being all flights on a given day between a spe- assigned to different times slots to represent arrival cific pair of airports. This choice yielded the highest times that fall in multiple slots. For example, the quality and most intuitive estimates. We used both morning and noon slots overlap from 9 a.m. to 11 a.m. the schedule and the availability files to build this set, Therefore, a flight arriving at 7 a.m. is a morning assuming that for some k, the k lowest available buck- flight with weight 1, a flight arriving at 11:30 a.m. is a ets (i.e., the k cheapest fares) on each flight are the noon flight with weight 1, a flight arriving at 9:30 a.m. alternatives considered by an arriving customer. is considered a morning flight with weight 075 and Although the availability database gives the aver- a noon flight with weight 025, a flight arriving at age revenue per bucket, in general there was a 10:30 a.m. is considered a morning flight with weight mismatch between these revenue numbers and the 025 and a noon flight with weight 075. fares listed in the revenue accounting database (the We acknowledge that this use of overlapping time sequence of PNRs). Moreover, it is not necessarily slots and “fuzzy” indicator variables is nonstandard true that the fare in the PNR corresponds to the low- from a modeling point. Traditional discrete choice for- est open bucket available at the time of booking.6 mulation approaches would typically define nonover- To guarantee consistency within the MNL model, lapping slots, leading to four dummy variables, one we therefore replaced the true fare paid (taken from of which would be eliminated by setting it as a the revenue accounting file) by the average revenue reference alternative (e.g., see Coldren et al. 2003 reported in the availability file.7 In this way, we and Coldren and Koppelman 2005 for characteriza- ensured that the fares for the alternatives that are pur- tions of time of day preferences in airline passenger chased and those that are not purchased are consis- choice models using dummy variables).8 However, tent, and hence the choice probabilities in (5) and (6) the overlapping time windows were suggested by are well defined and add up to one. Still, this ad hoc the sponsor airline because these are the time frames fixing of the data introduces some noise that could used to define “morning,” “noon,” “afternoon,” and hurt the estimation procedure. “evening” flights in the airline’s website search engine. Another major design decision relates to the time We felt, therefore, that they best represented the time unit of analysis. In the extreme case where the time differences as presented to customers. We also used unit is very small—say, at the level of a second—for the attribute of arrival time for our set of outbound a given O-D market, most periods in a day will have flights from New York to Florida rather than the alter- native of departure time. This was also based on dis- 6 The most consistent approach would be to use the fare-basis code cussions with the airline sponsor, who felt time spent to map a booking request to an open bucket. Unfortunately, accord- in Florida was the primary concern of travelers in ing to our experience, this was not straightforward even for the these markets. airline research staff itself to unravel this mapping because of the diversity of sales channels, agency agreements, etc. 8 Yet an alternative approach is used by Carrier (2008), who defines 7 We also tried the reverse substitution, but the results were of a time-based cyclical utility for airline passenger choice based on worse quality. a sine/cosine formulation. Vulcano, van Ryzin, and Chaar: OM Practice Manufacturing & Service Operations Management 12(3), pp. 371–392, © 2010 INFORMS 377

Departure day may be considered as a choice along Table 1 Choice Behavior: Attribute Definitions with arrival time of day, in which case the choice Attribute Description Value set Cn includes the flights of two or more consecutive x Base fare Base fare divided by 1,000 days. This is the formulation we used for our simu- n1 x Morning flight (before 11 a.m.) Weight as morning arrival lated data example. The day of a flight is represented n2 xn3 Noon flight (9 a.m.–3 p.m.) Weight as noon arrival by dummy variables so that if the choice set Cn con- xn4 Afternoon flight (1 a.m.–7 p.m.) Weight as afternoon arrival x Evening flight (5 p.m.–midnight) Weight as evening arrival tains days d1 d2  dC , then we introduce dC − 1 n5 n n x Indicator for flying on day d 1 if flight departs on day d , dummy variables, setting one reference alternative as n6 1 1 0 if it departs on d the null alternative. Our data covered four days dur- 2 ing the peak spring break vacation period, March 23 to March 26, 2005, with d representing March 23 and 1 least with respect to price sensitivity, consisting over- d4 representing March 26. The other attribute included in the model is the whelmingly of vacation travelers. Still, within these base fare paid divided by 1,000 (because of numerical markets the airline could potentially face different scaling issues). All flights considered in these markets customer segments for whom the relative weights of are nonstop flights, and so we did not add an indica- different attributes may vary. In our model-building tor to account for this attribute. phase, we in fact tried a multisegment (or latent class) To illustrate this parametrization, consider a market MNL model, assuming that customers belong to dis- between New York and Florida,9 with possible depar- crete segments l = 1  L, for which we estimated both arrival rates l and preference parameters l (e.g., ture days d1 or d2. The complete specification of the model attributes is provided in Table 1. see Train 2003, Chapter 6). However, as with the com- For example, consider a flight from New York petitor data and alternative specific constants, the vol- departing March 23 with arrival time 9:15 a.m. and ume of data proved too limited, and we were not able fare $180. This alternative is represented as a tuple: to obtain good estimates even with L = 2 latent seg- ments (note that this doubles the number of model xin = 0180 0875 0125 0 0 1. Note that in our discrete choice model we do parameters). not include alternative specific constants. The reason Even though our final single-segment MNL model we omitted alternative specific constants is that the yielded good results (as will be shown in the fol- attributes that we consider are sufficient to distin- lowing sections), it also raises the usual concerns guish a flight, because all flights in a set are operated of the MNL model, most noticeably the property by the same carrier between the same , and of independence from irrelevant alternatives, which they differ only in terms of arrival time, price, and implies proportional substitution across alternatives. departure date. Furthermore, when we tried using Briefly, this property establishes that the ratio of prob- alternative specific constants, we got very poor qual- abilities for two alternatives is constant regardless of ity estimates, most likely because of the increased the consideration set containing them. Other choice number of parameters and the relatively low volume models are more flexible with respect to the variety of of data we had available. substitution patterns that they exhibit (e.g., see Train We emphasize here that our choice model assumes a 2003, Chapter 4). Among them, the nested logit model homogeneous market; i.e., customers’ preferences are has been widely used in the marketing literature. In described by a single set of parameters  for all cus- our case, we could think of a nested structure where tomers. According to the airline staff’s assessment, this “buy/no buy” is at the top of the hierarchy and the was a reasonable assumption; the New York–Florida different flight options are under the “buy” branch. market was believed to be relatively homogeneous, at Yet the nested logit model can be more challenging to estimate than the basic MNL model because the log-likelihood function is not globally concave. 9 In our terminology, market refers to particular origin-destination airport combinations, like origin in LaGuardia (LGA) and destina- In summary, exploring alternative choice model tion in Palm Beach (PBI). specifications within the airline industry context is Vulcano, van Ryzin, and Chaar: OM Practice 378 Manufacturing & Service Operations Management 12(3), pp. 371–392, © 2010 INFORMS certainly worthy of further study. Nevertheless, we Consider a linear-in-parameters mean utility of the T believe that the single-segment MNL model used form vin = &0 +  xin. Our objective is to estimate the in our project provides a good first-cut sense parameters  = &0   from purchase data, where of the possibilities and benefits of implementing  = &1  &K . As discussed above, given only pur- choice-based RM. chase data it is impossible to distinguish a period without an arrival from a period with an arrival but 4. Estimation where the arriving customer chooses the no-purchase Having defined the consideration sets and the para- alternative 0 (i.e., she bought from the competitor or metrization of the attributes, we can compute the did not buy at all). The incomplete data log-likelihood MLE estimates for our model. We first describe the function for our model is given by log-likelihood function with incomplete data for our log x  model, where we recall that the incompleteness comes    from the fact that only purchase transaction data are H B   = log +& +Tx −log e&0+Txj +1 available and no-purchase outcomes are unobserv- 0 i h=1b=1i∈b h j∈Cb able. To overcome this problem, we implement a vari-       B H ation of the EMmethod that estimates parameters for 1 + b h log  + 1−   the utility function (2) and for the arrival rate of cus- e&0+Txj +1 b=1 h=1 j∈Cb tomers jointly. Then, we present a preliminary simu- (7) lated data example to illustrate the quality of the esti- mates produced by the procedure. Finally, we present Unfortunately, this function is difficult to maximize two examples based on real airline data. directly because of the complexity of the second term above. To overcome this problem, we used the EM 4.1. The Incomplete Data Log-Likelihood method of Dempster et al. (1977). Function Consider a collection of booking histories, h = 4.2. The EM Method 1  H, representing statistical replicas of a given The EMmethod operates on the complete data log- booking horizon. Each booking horizon consists of likelihood function, which has a simpler form than (7) B days, and bookings occur for flights departing on and is constructed assuming that all the arrivals, pur- any day 1  D. For instance, we could consider the chases, and nonpurchases can be observed. Specifi- last D = 5 days of a month, a booking horizon of cally, for the periods in b h, let ab = 1 if there is an B = 30 days prior to those D days, and have data for arrival in an observation period of day b, and ab = 0 each month during a calendar year, giving us H = 12 otherwise. Note that ab = 1 accounts for an arrival that histories. either buys from our competitors or decides not to We assume a (discretized) Poisson arrival process buy at all.10 of customers: Each booking day b is broken up into We can now write the complete data log-likelihood T small time intervals. In each small time period function: of day b, an arrival occurs with probability . Note that because of the Poisson assumption, and assum- log  x     ing intervals of time are small, there is at most one H B   T &0+Txj arrival per period with probability 1. Each arrival = log +&0 + xi −log e +1 selects among the products available on day b (or h=1b=1i∈b h j∈Cb does not purchase at all) according to the MNL model 10 For notational simplicity, we omit the fact that each booking day described by (5) and (6). Let b h denote the set of periods in which customers purchase on day b of b is itself divided into many smaller time periods, so in fact there is an indicator a for each interval t of day b. Because this extra bt booking history h, and let b h denote the set of peri- set of indices does not change the form of the likelihood function, ods in which there are no purchase transactions on to minimize the complexity of the displayed equations, we omit day b of booking history h (i.e., b h+b h=T ). them. Vulcano, van Ryzin, and Chaar: OM Practice Manufacturing & Service Operations Management 12(3), pp. 371–392, © 2010 INFORMS 379

     B H  where, from (6), &0+Txj + b h ab log − log e + 1 1 b=1 h=1 j∈Cb ˆ Pb 0  =     &ˆ +ˆ Tx B H e 0 j + 1 j∈Cb + b h 1 − ab log 1 −  (8) b=1 h=1 ˆ Observe that the probability Pb 0  for the nonob- The first line of (8) in accounts for the observed book- servable periods does not depend on a particular ings, and there is one term for every record in the booking history h, h = 1  H, because the choice revenue accounting database for each booking history. sets for a given booking day b are the same for all h. Substituting aˆ into (8), we obtain the expected, The second line accounts for customers that either b complete data log-likelihood function buy from competitors or do not purchase, and the third line refers to periods with no arrivals. These last Q  ˆ*= Elog  x  x ˆ two are unobserved data.    H B   The broad strategy of the EMmethod consists of = & +Tx −log e&0+Txj +1 ˆ ˆ ˆ ˆ 0 i starting with arbitrary initial estimates  ≡ &0   h=1b=1i∈b h j∈Cb of the parameters, where ˆ = &ˆ  &ˆ . Then     1 K B H B H we use these estimates to compute the conditional + b h log + b h expected value of log  x , Elog  x  x ˆ (the b=1 h=1 b=1 h=1    expectation step, or E-step). Effectively, this replaces  &0+Txj · aˆb log − log e + 1 the missing data (the indicators ab for the no-purchase j∈Cb time periods) by their expected values conditioned on   B H the current estimates. We then maximize the resulting + b h 1 − aˆb log 1 −  (9) log-likelihood function (which has the same form as b=1 h=1 the complete log-likelihood function) to generate new This function is relatively easy to maximize, as shown ˆ estimates  (the maximization step, or M-step). These next. steps are repeated until the procedure converges. We next describe these steps in complete detail. 4.2.2. The M-Step. We next determine a maxi- mizer, ˆ ∗, of the expected log-likelihood function (9). 4.2.1. The E-Step. In the E-step of each iteration, Note that the function (9) is separable in and &0 . the unknown data are the values ab in the second and Define the function F  as third lines of (8), for all b, corresponding to periods     B H B H with nonpurchase from our airline. However, given F *=  log +  aˆ log ˆ ˆ ˆ ˆ b h b h b current estimates  = &0  , we can determine the b=1 h=1 b=1 h=1   expected values of these indicators. Specifically, let B H

Pb i be the probability that a customer arriving in a + b h 1 − aˆb log 1 −  small period of day b chooses alternative i (irrespec- b=1 h=1 tive of the booking history). By the Bayesian formula, Taking the derivative with respect to b,weget we get for b = 1  B, irrespective of the booking ,F  B H 1 B H 1 horizon h: =  +  aˆ , b h b h b ˆ b=1 h=1 b=1 h=1 aˆb *= Eab t ∈ b  B H 1 ˆ −  1 − aˆ   =  ab = 1 t ∈ b  b h b 1 − b=1 h=1  t ∈  a = 1 ˆ a = 1 ˆ = b b b Setting this derivative equal to zero, we obtain the ˆ  t ∈ b  updated estimate ˆ ˆ ˆ     Pb 0 &0  B H  + B H  aˆ = ∗ b=1 h=1 b h b=1 h=1 b h b ˆ ˆ ˆ ˆ ˆ =  Pb 0 &0  + 1 −  B × H × T Vulcano, van Ryzin, and Chaar: OM Practice 380 Manufacturing & Service Operations Management 12(3), pp. 371–392, © 2010 INFORMS

This says that our current best estimate of is the 4.2.3. Convergence and Implementation. To number of observed purchases for flights booked check for convergence, in each iteration of the on any day b across the H booking histories, algorithm we compare the value of the expected   B H log-likelihood function (9) before and after maximiz- b=1 h=1 b h, plus the estimated number of arrivals from unobservable periods on any day b across all ing with respect to and &0 . We also check for   B H the norm of the difference between two consecutive the booking histories, b=1 h=1 b haˆb, divided by sets of these estimates. Thus, if  = & i  i i the total number of periods under consideration, B × i 0 H × T . The function F  is concave in , and hence is the optimal set of parameters from iteration i, we check if Q  ˆ  − Q  ˆ  0. We also set a of F . i+1 i maximum of 300 iterations for the EMmethod. Next, we maximize the terms in (9) contain- Because the expected log-likelihood function (9) is ing &  to obtain the updated estimates &ˆ∗ ˆ ∗. 0 0 continuous in both &  and &ˆ ˆ ˆ ,11 then The- Define the function G &  as 0 0 0 orem 2 in Wu (1983) guarantees that if the sequence of estimates converges, the resulting value will be a G &0  stationary point of the incomplete data log-likelihood    H B   function (7). Whether the sequence diverges or con- T &0+Txj *= &0 +  xi − log e + 1 verges to something other than the global maximum h=1 b=1 i∈b h j∈Cb is more difficult to determine. In practice, the EM     B H  method has proved to be a robust and efficient way −  aˆ log e&0+Txj + 1  (10) b h b to compute maximum likelihood estimates for incom- b=1 h=1 j∈C b plete data problems (e.g., see McLachlan and Krish- nan 1996, Chapter 3, for further discussions on con- To find a critical point &ˆ∗ ˆ ∗ of G & , we solve 0 0 vergence properties of the EMmethod). this nonlinear optimization problem by using the Although the standard implementation of the EM conjugate gradient method. The partial derivative of method is computationally intensive because it makes G &  with respect to & is 0 0 a pass through all of the available data in every iter-    ation, in our experience it is computationally feasible H B &0+Txj ,G &    j∈C e with the volume of data that a real airline may han- 0 = 1−  b & +Tx ,&0 e 0 j +1 dle during a booking period of few weeks for a single h=1b=1i∈b h j∈Cb    O-D pair and for a small number of departure days. B H e&0+Txj j∈Cb Our implementation of the estimation algorithm − b h aˆb   (11) e&0+Txj +1 was developed under Windows XP and coded in b=1 h=1 j∈Cb C++ using the Microsoft Visual C++ 6.0 compiler For all k = 1  K, we compute the partial deriva- and the STL (Standard Template Library) as a source tives of G &0 : of basic classes and containers. To optimize the function (10), we selected from the library Numer-    ,G &  H B  e&0+Txj x ical Recipes in C++ the Broyden-Fletcher-Goldfarb- 0 j∈Cb j k = xi k −  Shanno variant of the Davidon-Fletcher-Powell min- ,& &0+Txj k h=1b=1i∈ e +1 b h j∈Cb imization algorithm, a quasi-Newton-type method    B H e&0+Txj x that requires computing (11) and (12). j∈Cb j k − b h aˆb   (12) e&0+Txj +1 b=1 h=1 j∈Cb 4.3. Preliminary Estimation Case: Example 0 We first applied our EMmethod over a set of sim- Although the function G &0  is data dependent, ulated data. We did this as an initial test to get a it is jointly concave for linear-in-parameters utilities under weak regularity conditions on the data (see 11 Note that the expected log-likelihood function (9) is also a func- ˆ ˆ ˆ Train 2003, p. 65; McFadden 1974). tion of &0   through aˆb . Vulcano, van Ryzin, and Chaar: OM Practice Manufacturing & Service Operations Management 12(3), pp. 371–392, © 2010 INFORMS 381 sense of how many data are necessary to get good we would have estimates and how closely those estimates match the v = & x + & x +···+& x (known) parameters that generated the data. Given in 1 in1 2 in2 8 in8 a known underlying MNL choice model (i.e., given = &1 xin1 + &2 − &5xin2 + &3 − &5xin3 + &4 − &5xin4 values for and ) for a fixed O-D market, we used + & + & − & x + & − & x + &  (13) Monte Carlo simulation to generate H = 100 streams 5 6 8 in6 7 8 in7 8 of data for a booking horizon of B = 10 days for By defining &0 = &5 +&8 and relabeling the parameters flights departing on any day d1 d2 or d3. We set to have a consecutive numbering, we can equivalently T = 03 and used the MNL probabilities (5) and (6) consider a mean utility of the form vin = &0 +  xin to determine the product chosen by each arrival. We for the different purchase options (recall that the no- have around 42,000 arrivals throughout the gener- purchase option has been normalized to have zero ated streams, out of which around 40,300 booked a mean utility). Here, we are choosing as reference vari-

flight. The data about the flights offered were taken ables both evening flight and departure day d3. In gen- from both the schedule and the availability files for eral, one of the criteria for picking the reference vari- the market under consideration. In this preliminary ables is that there is at least one alternative in every example, we assume that the customer considers the choice set that matches it (e.g., in our instance this lowest available bucket of each flight in her consid- would mean that in every booking day there is an eration set (i.e., following the notation in §3.2, we evening flight offered with departure day d3). This is set k = 1). indeed the case in our data. We refer to &0 as the base The parameters  used for this base case are given utility. in Table 2 and were based on estimates obtained from We applied the EMmethod starting from esti- ˆ ˆ ˆ the real airline data set. Descriptive statistics of the mates &0 = 1 and , set equal to zero. The out- data used for this example can be found in Table A1 put after reindexing the  parameters is included in in the online appendix. Table 3. To determine the mean utility (2) for a given The third column in this table reports the true parameter values, following the adjustment described flight, we need one parameter &j for every attribute in (13). The fourth column reports the estimates com- value xin. However, we modified this parameter defi- nition. Observe that both the set of attributes for the puted by the EMmethod. The fifth column reports time slots and the set for the departure days add up the percentage bias between the estimated and true 12 to one, leading to a degree of freedom in the  param- values, followed by the asymptotic standard error eters and hence to the identifiability problem (i.e., the (ASE) of the corresponding estimate (see §A5 in the inability to fully recover the original parameters from online appendix for further details). Note that for the estimated ones). In symbols, for customer n, given all the coefficients except the indicator for morning flight, we can reject the null hypothesis that the true x and , and noting that in value is zero at the 0.01 significance level.13 How- ever, the true value of the indicator for morning flight xin2 + xin3 + xin4 + xin5 = 1 and xin6 + xin7 + xin8 = 1 is indeed zero. The number of iterations performed in the EMmethod in this case was 36, taking 15 Table 2 Input  Values for Example 0 minutes of computing time. Given the identifiabil- Attribute Description Value ity problem inherent to the data, it is not possible

1 Base fare −10 12 Note the results suggest an apparent bias in the estimates, 2 Morning flight (before 11 a.m.)05

3 Noon flight (9 a.m.–3 p.m.)07 which is not unexpected because the MLE is only asymptotically

4 Afternoon flight (1 a.m.–7 p.m.)03 unbiased.  Evening flight (5 p.m.–midnight) 05 5 13 The quasi-t statistic is computed as the ratio between the esti-  Indicator for flying on day d 06 6 1 mated value of the parameter and the ASE. Recall that for a two- 7 Indicator for flying on day d2 03  Indicator for flying on day d 09 tailed test, the critical values of this statistic are ±165, ±196, and 8 3 ±258 for the 0.10, 0.05, and 0.01 significance levels, respectively. Vulcano, van Ryzin, and Chaar: OM Practice 382 Manufacturing & Service Operations Management 12(3), pp. 371–392, © 2010 INFORMS

Table 3Output Parameters for Example 0

Parameter Description True value Est. value Bias (%) ASE t-stat ˆ 0 Base utility for any purchase option 1415881 1343 00131 12134 ˆ 1 Base fare −10 −12265 2265 00644 −1906 ˆ 2 Morning flight (before 11 a.m.)0000169 00273 062 ˆ 3 Noon flight (9 a.m.–3 p.m.)0202376 1880 00128 1860 ˆ 4 Afternoon flight (1 p.m.–7 p.m.) −02 −02178 889 00132 −1639 ˆ 5 Indicator for flying on day d1 −03 −02715 −951 00155 −1751 ˆ 6 Indicator for flying on day d2 −06 −05756 −406 00126 −4562 ˆ Arrival rate 0302988 −041 2.3E−05 1322940

ˆ chooses flight i, we have that × P i is the Poisson to distinguish how much of the value &0 is driven n rate of arrivals per unit of time on booking day b that by each reference (evening slot versus day d3). How- ever, the quality of the estimates is quite good, as choose flight i departing on day d. Therefore, indicated by the last three columns of Table 3. In the online appendix (§A3), we include another simu- Enumber of bookings on day b for flight i lated example that does not suffer from this identifi- departing on day d = T × × Pn i ability problem, although the bias of the estimates is higher. where T = 140 is the number of booking periods per Figure 1 shows a measure of the goodness of fit of day. Based on the true parameters of Table 2, we our estimates for Example 0 for the first five days of can compute the true expected number of bookings the booking horizon under consideration. For a given and utilities to be observed from a population of cus- pair b d (i.e., booking day b, departure day d), and tomers that choose according to them. Based on the recalling that Pn i is the probability that customer n estimated parameters of Table 3, we can compute the

Figure 1 Goodness of Fit for the Bookings per Combination bd for Booking Days b = 15 in Example 0

True expected vs. predicted bookings 8 True expected

7 Predicted

6

5

4 Bookings 3

2

1

0

ase 1, D1 1, D2 2, D1 B B B1, D3 B B2, D2 B2, D3 B3, D1 B3, D2 B3, D3 B4, D1B4, D2 B4, D3 B5, D1B5, D2B5, D3

1, no purch B B2, no purchase B3, no purchase B4, no purchase B5, no purchase Booking day, departure day Vulcano, van Ryzin, and Chaar: OM Practice Manufacturing & Service Operations Management 12(3), pp. 371–392, © 2010 INFORMS 383

Figure 2 Goodness of Fit for the Utilities per Combination bd for Booking Days b = 15 in Example 0

True expected vs. predicted utilities 1.8 True expected 1.6 Predicted

1.4

1.2

1.0

Utility 0.8

0.6

0.4

0.2

0

D2 D1 1, 1, D3 2, 5, D2 B1, D1 B B B B2, D2 B2, D3 B3, D1B3, D2 B3, D3 B4, D1B4, D2 B4, D3 B5, D1B B5, D3 o purchase , no purchase 2, no purchase B1 B B3, no purchase B4, n B5, no purchase Booking day, departure day number of bookings and utilities predicted by the out- total predicted sales are quite accurate. Moreover, we ˆ ˆ ˆ put of the EMmethod. In Figure 1, for every possible note that multiple values &0   may produce the combination b d there are several tick marks, each same probabilities of sale as observed by Talluri and one representing a flight. The graph shows an excel- van Ryzin (2004a, §5). In such cases, the EMmethod lent quality of fit; clearly, the difference between the finds only one such pair. In the online appendix average number of observable bookings and the num- (Figure A2.1) we show the comparison between the ber predicted by the EMmethod is negligible. true and predicted MNL choice probabilities for dif- We then aggregated the observed number of book- ferent alternatives (including the no-purchase), where ings per flight across the entire booking horizon and we also found a very good quality of fit. the expected number of bookings predicted by the We next studied the sensitivity of the procedure to EMmethod for the 14 flights contained in the three different starting points. For example, starting from ˆ ˆ departure days under consideration (see Table A2 in &0 = 1  = 0, and two different arrival rates (the true the online appendix). A 2-test gives a p-value = 1, value of the arrival rate ˆ = 03, and ˆ = 06), we giving a very strong justification not to reject the null obtain high-quality results, similar to the case when hypothesis that the observed bookings indeed follow the initial ˆ = 0. Details are provided in Table A3 in the distribution estimated via the EMprocedure. the online appendix. Figure 2 shows another measure of the goodness We then checked the behavior of the EMmethod of fit; in this case, we compared the true expected util- when the model does not capture all parameters of ities and the mean utilities predicted by our EMpro- the utility function (e.g., there are omitted variables in cedure. Here, even though the estimated utilities are the specification of the model). Toward this end, tak- slightly lower than the original utilities, the shift pre- ing again the data generated based on the description serves their relative order. Again, this reflects a bias in in Table 2 with = 03, we assume that customers the component estimates, even though the resulting choose based only on the fare and the departure days, Vulcano, van Ryzin, and Chaar: OM Practice 384 Manufacturing & Service Operations Management 12(3), pp. 371–392, © 2010 INFORMS

Table 4 Estimated Parameters for Example 0 When Customers Choose Based on Fare and Departure Day Only

Parameter Description True value Est. value Bias (%) ASE t-stat ˆ 0 Base utility for any purchase option 0916207 8007 00133 12227 ˆ 1 Base fare −10 −06232 −3768 00627 −994 ˆ 5 Indicator for flying on d1 −03 −04283 4276 00150 −2850 ˆ 6 Indicator for flying on d2 −06 −06823 1372 00126 −5425 ˆ Arrival rate 0302977 −076 00000 1318251 and we omit the fact that the flight arrival time also Running a 2-test of the observed bookings versus the factors into their total utility. Table 4 shows the results number of bookings predicted by our model for the obtained from this misspecified model. Note that for 14 flights under consideration, we get a small p-value all the coefficients we can reject the null hypothe- of 0.08. This case shows the importance of defining sis that the true value is zero at the 0.01 signifi- the customers’ consideration sets; accounting for dif- cance level. Another measure of the goodness of fit ferent departure days is important. is given by comparing the values of the incomplete We also tested how the volume of data affects the data log-likelihood function evaluated at the original quality of estimates from our procedure. We did this parameter values and at the estimated parameters, by changing the value of the arrival rate . We could which are −176 323 and −171 629, respectively. Run- also do this by changing the number of histories, H, ning a 2-test of the observed bookings versus the or the booking periods per day, T . However, vary- ing gives a better sense of the impact of the rela- expected number of bookings predicted by the esti- tive mix of periods with observations and nonobser- mated model for the 14 flights under consideration vations in the sample. Fixing H = 100, T = 140, and  gives strong support for the estimated model with a as in Example 0, and taking = 001, we obtain rather p-value = 093. poor estimates with large biases and small t-statistics Next, we assumed that customer utilities are based (see Table 6). This is we have many periods with no only on the fare and the flight arrival times (ignoring observations and only a few observable purchases. In flight departure day) and estimated this misspecified this case, the total number of bookings across the 100 model based on booking data for the D = 3 book- histories is around 1,400, and this does not seem suf- ing days under consideration. The results are reported ficient to obtain good estimates. in Table 5. Note that for all the coefficients (includ- When we increase the value of the arrival rate to ing the indicator for morning flight), we can reject = 005, which corresponds to around 7,000 recorded the null hypothesis that the true value is zero at the bookings, we start getting better estimation results, 0.01 significance level. In this case, using (13), we can and for an even larger value = 07 (corresponding recover the four time coordinates of ˆ from the value to around 98,000 arrivals), we obtain high-quality ˆ of &0. However, in this case, the quality of the esti- estimates (even better than the base case in Table 3 mates is rather poor, with very pronounced biases. for = 03). Running a 2-test of the expected number

Table 5 Estimated Parameters for Example 0 When Customers Choose Based on Fare and Flight Times

Parameter Description True value Est. value Bias (%) ASE t-stat ˆ 0 Base utility 0516138 22276 00131 12343 ˆ 1 Base fare −10 −23879 13879 00607 −3931 ˆ 2 Morning 0001841 00274 672 ˆ 3 Noon 0204030 10148 00126 3194 ˆ 4 Afternoon −02 −01694 −1530 00134 −1268 ˆ Arrival rate 0302982 −059 00000 1320542 Vulcano, van Ryzin, and Chaar: OM Practice Manufacturing & Service Operations Management 12(3), pp. 371–392, © 2010 INFORMS 385

Table 6 Estimated Parameters for Example 0 Under Different Arrival Rates

Arrival rate  = 001 Arrival rate  = 070

Estimated Estimated Param. Description True value value Bias (%) t-stat value Bias (%) t-stat ˆ 0 Base utility 140 143635 92596 000 15905 1361 19829 ˆ 1 Base fare −100 −05423 −4577 −113 −10761 761 −2879 ˆ 2 Morning 000 −01608 −089 −00161 −098 ˆ 3 Noon 020 01549 −2256 166 02194 971 2932 ˆ 4 Afternoon −020 −02357 1783 −307 −02045 227 −2409 ˆ 5 Indicator for d1 −030 −03333 1111 −358 −02965 −117 −3055 ˆ 6 Indicator for d2 −060 −05868 −220 −771 −06160 267 −7924 ˆ Arrival rate 001 00098 −214 43480 06915 −122 3061887 of bookings predicted by our model in the = 07 available fares (i.e., buckets) of each flight. That is, for case versus the observed number of bookings for each flight in the consideration set, we take the five the 14 flights under consideration, we see there is lowest available fares from the availability file (i.e., very strong support for the estimated model, with we set k = 5 in terms of the notation introduced at p-value = 1. Overall, what matters for the quality of the beginning of §3.2). Recall that a flight’s time is the estimates is the volume of data available and defined as its arrival time in Florida. the definition of the base period. In the extreme case The initial set of estimates for the EMmethod ˆ ˆ ˆ where we have infinitely many small periods T , the was &0 =  = 0 and = 01. The EMmethod reached probability of having an arrival per period would the maximum of 300 iterations for this case in 208 be very small, and the quality of the estimates would seconds of computational time.14 The norm differ- be very poor, leading also to some potential numeri- ence between the last two sets of estimates was cal problems in the estimation process. Having longer  300 −  299=0002. base periods increases the probability of arrival, but In Table 7, the three columns under Example 1 give one must be careful to keep the probability of having a summary of the final set of estimates. Note that we two or more arrivals per period small. cannot reject the null hypothesis that the true value of the morning flight indicator is zero. The 2-test 4.4. Estimation Results for Airline Data of the total number of observed bookings versus the We next present the results of the EMmethod applied expected number of bookings per flight shows a rea- to an O-D market in our airline data set for two sonably high p-value = 041, providing good support different departure dates. We describe the estima- for the estimated distribution of sales. tion for each departure date in turn. Table A6 in A few aspects of the model summarized in the online appendix reports descriptive statistics for Table 7 are also worth noting. First, in terms of the each market. linear-in-parameters utility described in (2), the price coefficient has the right (negative) sign. Second, it is 4.4.1. Example 1. In this market, we consider a interesting to interpret the relative value of the coef- booking horizon of B = 39 days, with data from ficients as indicators of the sensitivity of the choices. February 6 to March 17, 2005 (i.e., we take just a sin- For instance, according to the attributes described gle history H = 1), with 549 reservations. The market in Table A6 in the online appendix, Flight 3 is a is defined by an origin airport in New York and a noon flight and Flight 4 is a 0.092 morning/0.908 destination airport in Florida. We consider flights for one departure day: March 25, 2005. There is a total 14 The computational times reported here were obtained with of five flights on this departure day in this market. an Intel Pentium IV, 2.4 GHz, and 512 MB of RAM. Although When building the consideration set Cn, we assume undoubtedly more sophisticated nonlinear solvers and platforms that each customer takes into account the five lowest could be used, this setup proved to be sufficient for our study. Vulcano, van Ryzin, and Chaar: OM Practice 386 Manufacturing & Service Operations Management 12(3), pp. 371–392, © 2010 INFORMS

Table 7 Estimated Parameters for Examples 1 and 2

Example 1 Example 2 Parameter Description Value ASE t-stat Value ASE t-stat ˆ 0 Base utility −20566 01185 −173619 −23468 00888 −264284 ˆ 1 Base fare −17348 02148 −80745 −19175 01601 −119746 ˆ 2 Morning flight (5 a.m.–11 a.m.)10861 12384 08770 33114 13590 24366 ˆ 3 Noon flight (9 a.m.–3 p.m.)13056 01465 89133 02851 01284 22203 ˆ 4 Afternoon flight (1 p.m.–7 p.m.)07751 01538 50406 −03186 01387 −22976 ˆ Arrival rate 01927 00011 1772174 03485 00011 3227402 noon flight. The mean utilities for this flights are, for the five flights under consideration, assuming that respectively, the fare is set at the intermediate point of the fare range open during the booking horizon.15 Although v3n =−20566 − 17348p3 + 13056 and this decision is arbitrary, the numbers obtained give a sense of the explanatory power of the MNL param- v4n =−20566 − 17348p4 + 10861 × 0092 eters when compared with the relative frequencies of + 13056 × 0908 the observed bookings. =−20566 − 17348p + 12854 4 4.4.2. Example 2. For the same market as above, we next considered a booking horizon of B = 39 days where p is the price charged for Flight i i = 3 4. i (between February 7 and March 18, 2005). Here, we Then, v ≥ v when p ≤ p − 0012. Recalling that 4n 3n 4 3 considered flights for March 26, 2005. There is a total the scale of the prices is in thousands, a discount of five flights in this market on this day with a total of $12 over the current price p will make Flight 4 a 3 of 506 recorded bookings. Like in Example 1, when more attractive alternative (on average), indicating a building the consideration set C , we assume that rather similar willingness to pay for both flights. Sim- n each customer takes into account the five lowest avail- ilarly, and just as another example, we can compare able fares of each flight. Again, a flight’s time is the mean utilities of Flights 4 (a late morning flight) defined as its arrival time in Florida. and 5 (an early evening flight): The initial set of estimates for the EMmethod was &ˆ = ˆ = 0 and ˆ = 01. The EMmethod also v4n =−20566 − 17348p4 + 10861 × 0092 0 reached the maximum of 300 iterations for this case. + 13056 × 0908 The norm difference between the last two sets of esti- mates was  − =00022. It took 193 seconds =−07712 − 17348p4 300 299 of computational time to run the procedure. v =−20566 − 17348p + 07751 × 0675 5n 5 In Table 7, the three columns under Example 2 give

=−15334 − 17348p5 a summary of the final coefficient estimates. Note that we can reject the null hypothesis that the true value of Then, v5n ≥ v4n when p5 ≤ p4 − 0439. Thus, a discount the parameters is zero at the 0.05 confidence level for of $439 over the current price p4 will make Flight 5 an all estimates, implying that the attributes under con- alternative more attractive than Flight 4, on average. sideration are relevant for the sake of capturing choice This type of analysis could be useful for the airline behavior. The 2-test of the total number of observed to determine price strategies, or even flight schedule bookings versus the expected number of bookings per strategies, to influence the revenue performances and load factors of the different flights. 15 See descriptive statistics in Table A6 in the online appendix. Columns for Example 1 in Table 8 provide the cal- For instance, for Flight 1, the range of fares for the 63 observed culation of the market shares based on the estimated bookings is 57 212, so we assume that the fare is set at parameters and the linear-in-parameters utilities (2) 57 + 212/2 = 13450. Vulcano, van Ryzin, and Chaar: OM Practice Manufacturing & Service Operations Management 12(3), pp. 371–392, © 2010 INFORMS 387

Table 8 Estimated Market Shares for Examples 1 and 2 Figure 3An Integrative Estimation/Optimization Approach for Choice-Based RM Example 1 Example 2

Utility-based Booking-based Utility-based Booking-based Forecasting/ Revenue optimization estimation Parameters module Flight market share market share market share market share module (Indep. demand assumption) Current protection 10110 0115 0203 0213 levels 20239 0237 0214 0223 Sample 30285 0260 0155 0170 Estimates paths Choice Simulation- 40294 0260 0264 0251 Choice model behavior based 50073 0128 0164 0142 simulator estimator optimizer

New protection flight shows a good p-value = 050, confirming the levels quality of the estimated distribution. CRS In this example, the price coefficient again has the right (negative) sign. We could also perform sensitiv- characterized by a preferred list of products, built ran- ity calculations as the ones presented for Example 1. domly using the utilities based on the MNL param- Regarding the consistency of the predicted market eters. Figure 3 sketches the integration of the whole share with the booking frequencies, we see in columns estimation/optimization approach. under Example 2 in Table 8 that, once again, the pre- 5.1. Simulating Customer Choice Behavior dicted shares based on the estimated parameters are Once we have the estimates that characterize cus- remarkably close to the observed bookings. tomer choice behavior from our EMmethod, we can use them to simulate the arrival of customer requests 5. Assessing RM Improvements to the central reservation system (CRS) as follows: On As mentioned above, the second phase of our booking day b, customers arrive in accordance with a study used the choice model estimates fit from Poisson process with rate per unit of time. Hence, the test markets to assess the potential improve- we generate a Poisson random variable with mean ment in revenues from using choice-based RMopti- 140 × to have an instance of the number of arrivals mization methods. Toward this end, we used the in day b. model and optimization algorithm developed in Next, for each generated arrival n, we build a pref- van Ryzin and Vulcano (2008). In this model, we erence list based on the utilities described in §2.1 assume the firm controls the availability of products as follows: First, we construct the set of alternatives 16 using a nested, protection-level-based, capacity con- Cn as we did in the estimation phase. Then, for trol strategy. Using a simulation-based optimization each alternative i ∈ Cn, we compute the utility Uin algorithm, we calculate sample path gradients of the in (1) by taking the vector of attributes xin correspond- network revenue function with respect to the protec- ing to this alternative, and by simulating the ran- tion levels. The gradients are then used sequentially dom term in. We use the standard inverse function to construct a stochastic steepest ascent method. Start- method for the latter; that is, for a Unif(0,1) value u, ing from an initial set of protection levels (possibly we compute computed under the traditional independent demand in =−log − log u −  model assumption), the algorithm is guaranteed to We also simulate the random noise  for the no- converge (in probability) to a stationary point of the 0n expected revenue function under mild conditions. purchase alternative. Recalling that the mean no- purchase utility has been normalized to zero, we The input data to the optimization algorithm are provided by a choice model simulator module that 16 Recall that the consideration set C is built primarily upon the generates streams of demand based on customers’ n schedule and the availability files. The difference from the estima- preferences described by the choice model estimated tion phase is that when simulating we do not take information from from our test markets. Every arriving customer is the revenue accounting file. Vulcano, van Ryzin, and Chaar: OM Practice 388 Manufacturing & Service Operations Management 12(3), pp. 371–392, © 2010 INFORMS rank our airline’s alternatives with utility in excess data and the fact that their value changed during of 0n. This ordered set of alternatives constitute cus- the booking horizon. In this sense, the protection lev- tomer n’s preference list. els given by EMSR-b constitute a more sensible “null This is repeated for all days in the booking horizon model.” under analysis, following the sequence b = 1  B to For the examples below, we simulated 2,000 build one stream of arrivals. The sequence of arrivals streams of customer arrivals. Every arrival is specified is saved in a file. For simplicity, we assume single-unit as a “customer type,” characterized by a particular demand for all customers. preference list (i.e., two customers that have the same preference list are considered the same type). In each 5.2. Optimizing RM Controls example, there are five parallel flights and 80 prod- The last stage in our analysis is to assess the per- ucts. According to the sponsor airline policy, there formance of RMcontrols based on the traditional are 16 buckets per flight (i.e., 15 protection levels independent demand assumption relative to those per leg). incorporating choice behavior effects. To this end, After generating the new set of protection lev- we ran simulation tests using Examples 1 and 2 in els, we checked the revenue obtained with the orig- §4.4. We tested the performance of our procedure inal and new protection levels over 2,000 simulated 17 based on two sets of initial protections levels : one streams of arrivals. We report expected revenues based on the current controls implemented by the under both protection levels policies, 95% confidence airline, and the other using single-leg, EMSR-b con- intervals for the revenue gap and corresponding (net- trols (e.g., see Talluri and van Ryzin 2004b, §2.2, work) load factors, defined as the average of the leg load for a definition and discussion of EMSR-b). For the factors, i.e., the ratio between average number of seats former, we used the “availability” file provided by sold on leg i and its initial capacity ci. the airline and constructed an initial set of nested We ran the numerical experiments under Windows protection levels taking the highest values of the XP, with a CPU Intel Core Duo of 2.0 GHz, and 1 GB controls for each class observed in the file during of RAM. the booking horizon under consideration, preserving 5.2.1. Example 1. This is the example described in the nested property of the protection levels. For the § 4.4.1. Across the 2,000 streams of demand, there are, EMSR-b controls, we first computed statistics for the on average, 841 arrivals per stream that have one of independent demands of the different products. To our flights as the first choice, with a total of 592,605 do so, we took the simulated streams of demand and customer types. The initial capacities available are still retained the first element of each customer’s prefer- significant (between 76 and 155 seats). ence list. In this way, we obtained the mean and vari- First, we ran the stochastic gradient algorithm over ance for the uncensored demand of each product (i.e., the set of protection levels provided by the airline. of each flight-bucket combination). Then, following Comparing the revenue obtained from the protection the usual practice, we assumed normal distribution levels improved by the algorithm and the original of each product’s demand and computed nested pro- ones, we observed an expected revenue gap of 3.04%, tection levels using EMSR-b. We compared these two with a 95% CI of −012% 621%. Even though this sets of protection levels with the improved ones that confidence interval technically includes zero, it is we obtained when applying the choice-based stochas- strongly on the positive side, and the improvement is tic gradient (SG) algorithm described in van Ryzin significant at slightly less than the 95% level. We also and Vulcano (2008). We emphasize here that the pro- observe a small increase in the load factor from 0.68 to tection levels provided by the airline were calibrated 0.70. However, as we explained above, because these in an ad hoc way, because of some dirtiness in the protection levels were built in an approximate way from the airline’s availability file, we place greater sig- 17 A protection level yi is defined as the number of seats reserved for classes i and higher. The labeling of the classes assumes that nificance on the values given by the EMSR-b heuristic. class 1 is the highest (i.e., most expensive fare) one, followed by Table 9 reports the original EMSR-b and updated class 2, class 3, etc. sets of protection levels obtained by the stochastic Vulcano, van Ryzin, and Chaar: OM Practice Manufacturing & Service Operations Management 12(3), pp. 371–392, © 2010 INFORMS 389

Table 9 Protection Levels for Example 1: EMSR-b vs. SG

Protection levels yi

Flight Capacity ci Stage 1 234567891011

1 136 Original 3 8 9 10 12 16 20 39 62 83 105 Updated 0 0 0 17 17 17 136 136 136 136 136 2 155 Original 9 26 56 91 126 155 155 155 155 155 155 Updated 9 10 147 155 155 155 155 155 155 155 155 3 152 Original 49 96 119 134 145 152 152 152 152 152 152 Updated 34 152 152 152 152 152 152 152 152 152 152 4 150 Original 48 86 102 112 113 115 128 138 149 150 150 Updated 80 83 83 83 83 89 125 138 149 150 150 5 76 Original 25 66 76 76 76 76 76 76 76 76 76 Updated 42 76 76 76 76 76 76 76 76 76 76 gradient algorithm. Even though according to the air- To assess the robustness of our combined estimation line system there were 16 virtual classes in the origi- and optimization procedure with respect to estimation ˆ ˆ ˆ nal flights, we report the protection levels of the open noise, we perturbed the original and &0  esti- classes (all the remaining classes were closed for the mates provided by our estimation algorithm, by ran- five flights, and remained closed after the application dom multiplicative factors in the range ±10% ±25%, of the algorithm). Revenue-wise, the stochastic gradi- and ±50%. Then, we simulated 2,000 streams of ent algorithm led to a revenue improvement of 1.42%, demand based on each perturbed set of estimates, and, with a 95% CI for the revenue gap of −247% 530%, starting from the original set of EMSR-b reported in suggesting the improvement in this case is not very Table 9, we computed the corresponding improved set statistically significant. It took just 15 seconds to com- of (noisy) protection levels using the stochastic gradi- pute the new set of protection levels. As we described ent algorithm. Finally, to assess the impact of the ran- in our previous paper (see van Ryzin and Vulcano dom noise in the estimates, we processed the original 2008, §4), the impact of the algorithm over differ- streams of demand (i.e., the demand streams gener- ˆ ˆ ˆ ent buckets is not unidirectional; that is, some classes ated under the “true” and &0 ) and compared become more protected and others become less. Some the revenues obtained with the noisy protection lev- classes are collapsed (e.g., classes 1–4 in Flight 1), and els to the revenue obtained using the original set of others are closed (e.g., classes 8–12 in Flight 1). In EMSR-b protection levels. Note that given the num- this case, the load factor significantly decreased from ber of bookings for the different flights described in 0.90 to 0.75, meaning that the new protection levels Table A6 in the online appendix (between 63 and 143), dramatically reduced the number of tickets sold but these perturbation of the estimated choice parameters improved the mix of passengers by forcing several can indeed have an impact on the distribution of pas- sengers across different flights. The results are summa- buy-ups (e.g., note the increment in y1 for Flights 4 and 5) and rejecting several formerly accepted low- rized in Table A7 in the online appendix, showing the fare tickets. outcome of five different perturbations in each of the The above test assumes the choice-based RMopti- ranges ±10% ±25%, and ±50%. Considering the rela- mization uses the true parameter values of the utility tively narrow, original expected revenue gap of 1.42%, function. However, in reality there will always be esti- we see that even with a 10% perturbation, the addi- mation error as well as specification error. We were tional revenues become negative in two of the five interested in how such errors in estimation effect the cases. revenue results. For example, one might posit that 5.2.2. Example 2. This is the example described in choice-based RMonly “works” if one has highly accu- §4.4.2. Each of the five parallel flights has 16 classes. rate estimates of the parameters of the choice model. Each of the 2,000 demand streams has an average of Vulcano, van Ryzin, and Chaar: OM Practice 390 Manufacturing & Service Operations Management 12(3), pp. 371–392, © 2010 INFORMS

1,097 arrivals that have one of our flights as the first explained by an increased sales volume. This is par- choice, and produce a total of 351,654 customer types. ticularly noticeable in Flights 1 and 4, where pro-

The capacities available vary between 106 and 135. tection levels y1 to y5 were significatively reduced. As in Example 1, first we ran the stochastic gradient This phenomenon occurs jointly with a change of the algorithm over the set of protection levels provided passenger mix that is clear in Flight 3, which now by the airline. Comparing the revenues obtained focuses exclusively on passengers for Buckets 1 and 2, from both sets of protection levels, we observed a and in Flight 5, which closes Bucket 4 but opens more expected revenue gap of 2.49%, with a 95% CI of the highest three classes. −052% 550%. Again, even though the confidence Even though the expected revenue gap of 5.30% interval includes the zero, it is strongly on the posi- might seem very high by RMstandards, it is not sur- tive side and significant at slightly less than 95%. We prising when compared to other results reported in also observe an increase in the load factor from 0.82 the literature. For instance, Liu and van Ryzin (2008, to 0.87. However, as explained above, because these §7.1) report revenue improvements between 0.10% protection levels were built in an approximate way, and 7% in 18 of 20 parallel flight cases when account- we also consider the protection levels given by the ing for choice behavior effects versus traditional meth- EMSR-b heuristic. ods. Also for a parallel flight case, Miranda Bront Table 10 reports the original EMSR-b and the cor- et al. (2009) report revenue improvements of approx- responding updated sets of protection levels obtained imately 10% for similar load factor scenarios. We also by the stochastic gradient algorithm. Even though observed significant gains for simulated parallel flight there were 16 buckets in the original flights, we networks in our previous paper (see Examples 2, 3, report the protection levels of the open classes (all and 4 in van Ryzin and Vulcano 2008, §4.) the remaining classes were closed for the five flights, We again tested the robustness of our estimation/ and remained closed after the application of the algo- optimization procedure with respect to estimation rithm). It took 70 seconds to compute the new set noise. In this case, given the large initial expected rev- enue gap of 5.30%, there is more room for estimation of protection levels. For this market, the stochas- error. As shown in Table A8 in the online appendix, tic gradient algorithm led to a revenue improve- the choice-based protection levels still produce sig- ment of 5.30%, with a 95% CI of 228% 832%. nificant revenue gains with up to 25% errors in the The initial and improved sets of protection levels are parameter estimates. shown in Table 10. In contrast to Example 1, here the load factor significantly increases from 0.82 to 0.89, suggesting that part of the increase in revenues is 6. Conclusions Our analysis shows that it is feasible to estimate choice behavior and customer preferences for price, Table 10 Protection Levels for Example 2: EMSR-b vs. SG arrival time, and departure dates from readily avail-

Protection levels yi able airline data. The main caveat is that the quality of these estimates may vary across markets, and they Flight Capacity c Stage 1 234567 i improve significantly if one has access to a large num- 1 118 Original 27 34 43 51 56 79 102 ber of historical booking records. Having customer- Updated 6 9 10 36 40 118 118 level shopping data would, in our view, provide for 2 122 Original 6 47 85 122 122 122 122 even more robust estimates of choice effects. Updated 1 4 122 122 122 122 122 Our simulation study suggests that the benefits that 3 106 Original 7 40 70 97 106 106 106 Updated 0 106 106 106 106 106 106 could be obtained from optimizing RMcontrols to 4 135 Original 48 76 87 98 109 130 135 account for choice behavior are significant: the rev- Updated 15 65 65 65 68 130 135 enue gains were between 1.4% and 5.3% in the mar- 5 121 Original 22 61 101 121 121 121 121 kets we tested. For the cases of relatively low rev- Updated 5 37 121 121 121 121 121 enue improvements, the accuracy of the choice model Vulcano, van Ryzin, and Chaar: OM Practice Manufacturing & Service Operations Management 12(3), pp. 371–392, © 2010 INFORMS 391

estimates was important for providing these bene- Blankenbaker and Katia Frank for their input and insight fits. When gains are higher (i.e., around 5% revenue during this study. Finally, they are grateful to the associate lift), the benefit of choice-based RMdoes not require editor and three anonymous referees for their helpful and constructive feedback during the review process. highly accurate estimates of the underlying choice model parameters. Although these are only simu- lation estimates and rely on the fact that demand References follows the model exactly (i.e., no model specifica- Algers, S., M. Besser. 2001. Modeling choice of flight and booking class: A study using stated preference and revealed preference tion error), they give some sense of the potential for data. Internat. J. Services Tech. Management 2 28–45. improvement using choice-based RM. More testing Andersson, S.-E. 1989. Operational planning in airline business— with industry-wide data of both the estimation and Can science improve efficiency? Experiences from SAS. Eur. J. Oper. Res. 43(1) 3–12. optimization procedures are needed to confirm these Andersson, S.-E. 1998. Passenger choice analysis for seat capacity initial results. We have worked out the basic technical control: A pilot project in Scandinavian airlines. Internat. Trans. Oper. Res 5(6) 471–486. building blocks for doing so in this paper. Belobaba, P. P. 1987a. demand and inventory Several extensions of this work warrant attention. management. Ph.D. thesis, Flight Transportation Laboratory, It would be desirable to further test a model with MIT, Cambridge, MA. Belobaba, P. P. 1987b. Airline yield management: An overview of unobservable (or latent) segments within the popu- seat inventory control. Transportation Sci. 21(2) 63–73. lation (e.g., business and leisure segments) so that Belobaba, P. P. 1989. Application of a probabilistic decision model customer choice behavior can be predicted more to airline seat inventory control. Oper. Res. 37(2) 183–197. Belobaba, P. P., C. Hopperstad. 1999. Boeing/MIT simulation study: accurately. However, as we discussed in §3.2, more PODS results update. Proc. 1999 AGIFORS Reservations and complex models appear to require more data to esti- Yield Management Study Group Sympos., London. Belobaba, P. P., L. R. Weatherford. 1996. Comparing decision rules mate accurately. Another extension that could prove that incorporate customer diversion in perishable asset rev- useful when dealing with a large volume of data enue management situations. Decision Sci. 27(2) 343–363. is the sampling-based implementation of the EM Ben-Akiva, M., S. Lerman. 1994. Discrete Choice Analysis: Theory and Applications to Travel Demand, 6th ed. The MIT Press, method proposed by Jank (2005), which consists of Cambridge, MA. intelligently incrementing the number of observations Boyd, E. A., R. Kallesen. 2004. Practice papers: The science of rev- enue management when passengers purchase the lowest avail- needed to compute the log-likelihood function toward able fare. J. Revenue Pricing Management 3(2) 171–177. the end of the algorithm instead of using all the avail- Carrier, E. 2008. Modeling the choice of an airline itinerary and able data in each iteration of the EMmethod. Finally, fare product using booking and seat availability data. Ph.D. thesis, Department of Civil and Environmental Engineering, our choice model assumes that customers’ valuations Massachusetts Institute of Technology, Cambridge. for different attributes do not depend on the offer set. Coldren, G., F. Koppelman. 2005. Modeling the competition among air-travel itinerary shares: GEV model development. Trans- It thus ignores framing or reference effects. Account- portation Res. Part A 39(4) 345–365. ing for such behavioral biases would be worth further Coldren, G. M., F. S. Koppelman, K. Kasturirangan, A. Mukher- study. jee. 2003. Modeling aggregate air travel itinerary shares: Logit model development at a major U.S. airline. J. Air Transport Management 9(6) 361–369. Electronic Companion Dempster, A. P., N. M. Laird, D. B. Rubin. 1977. Maximum likeli- An electronic companion to this paper is available on hood from incomplete data via the EMalgorithm. J. Roy. Statist. the Manufacturing & Service Operations Management website Soc. B 39(1) 1–38. Dubé, J.-P., P. Chintagunta, A. Petrin, B. Bronnenberg, P. B. (http://msom.pubs.informs.org/ecompanion.html). Seetharaman, K. Sudhir, R. Thomadsen, Y. Zhao. 2002. Struc- tural applications of the discrete choice model. Marketing Lett. Acknowledgments 13(3) 207–220. Gallego, G., G. Iyengar, R. Phillips, A. Dubey. 2004. Managing Part of this work was conducted while the first author flexible products on a network. Report TR-2004-01, Computa- was visiting the School of Business, Universidad Torcu- tional Optimization Research Center. Department of Industrial ato di Tella, Buenos Aires, Argentina, in 2005. The authors Engineering and Operations Research, Columbia University, especially thank Juan Manuel Chaneton, research assistant New York. at this school, for helping with the coding of the esti- Gumbel, E. J. 1958. Statistics of extremes. Columbia University Press, mation algorithm and processing of the data. Juan José New York. Jank, W. 2005. Fast and Eficient Model-Based Clustering with the Miranda Bront, Departamento de Computación, Universi- Ascent-EMAlgorithm. B. Golden, S. Raghavan, E. Wasil, eds. dad de Buenos Aires, Argentina, helped with some prelim- The Next Wave in Computing, Optimization, and Decision Tech- inary numerical experiments. The authors also thank John nologies. Springer, New York, 201–212. Vulcano, van Ryzin, and Chaar: OM Practice 392 Manufacturing & Service Operations Management 12(3), pp. 371–392, © 2010 INFORMS

Kunnumkal, S., H. Topaloglu. 2009. A new dynamic programming demand from airline bookings. J. Revenue Pricing Management decomposition method for the network revenue management 7(2) 153–171. problem with customer choice behavior. Production Oper. Man- Talluri, K. T., G. J., van Ryzin. 2004a. Revenue management under agement. Forthcoming. a general discrete choice model of consumer behavior. Manage- Liu, Q., G. van Ryzin. 2008. On the choice-based linear program- ment Sci. 50(1) 15–33. ming model for network revenue management. Manufacturing Talluri, K., G. J. van Ryzin. 2004b. The Theory and Practice of Revenue Service Oper. Management 10(2) 288–310. Management. Kluwer Academic Publishers, New York. McFadden, D. 1974. Conditional logit analysis of qualitative choice Train, K. 2003. Discrete Choice Methods with Simulation. Cambridge behavior. P. Zarembka, eds. Frontiers in Econometrics. Academic University Press, New York. Press, New York, 105–142. van Ryzin, G. J., G. Vulcano. 2008. Computing virtual nesting McGill, J. I. 1995. Censored regression analysis of multiclass pas- controls for network revenue management under customer senger demand data subject to joint capacity constraints. Ann. choice behavior. Manufacturing Service Oper. Management 10(3) Oper. Res. 60(1) 209–240. 448–467. McLachlan, G., T. Krishnan. 1996. The EM Algorithm and Extensions. Wu, C. F. J. 1983. On the convergence properties of the EMalgo- John Wiley & Sons, New York. rithm. Ann. Statist. 11(1) 95–103. Miranda Bront, J., I. Méndez-Díaz, G. Vulcano. 2009. A column gen- Zhang, D., D. Adelman. 2009. An approximate dynamic program- eration algorithm for choice-based network revenue manage- ming approach to network revenue management with cus- ment. Oper. Res. 57(3) 769–784. tomer choice. Transportation Sci. 43(3) 381–394. Ratliff, R., B. Rao, C. Narayan, K. Yellepeddi. 2008. A Zhang, D., W. L. Cooper. 2006. Revenue management for parallel multi-flight recapture heuristic for estimating unconstrained flights with customer-choice behavior. Oper. Res. 53(3) 415–431.