foundations of stated preference elicitation: consumer ...train/foundations.pdf · foundations of...

144
Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis Moshe Ben-Akiva 1 , Daniel McFadden 2,3 and Kenneth Train 2 1 Department of Civil and Environmental Engineering, MIT Cambridge, MA 02139, USA 2 Department of Economics, University of California, Berkeley, CA 94720-3880, USA 3 Department of Economics, University of Southern California, CA 90089-921, USA ABSTRACT Stated preference elicitation methods collect data on consumers by “just asking” about tastes, perceptions, val- uations, attitudes, motivations, life satisfactions, and/or intended choices. Choice-Based Conjoint (CBC) analysis asks subjects to make choices from hypothetical menus in experiments that are designed to mimic market experiences. Stated preference methods are controversial in economics, particularly for valuation of non-market goods, but CBC analysis is accepted and used widely in marketing and policy analysis. The promise of stated preference experiments is that they can provide deeper and broader data on the structure of consumer preferences than is obtainable from revealed market observations, with experimental control of the choice environment that circumvents the feedback found in real market equilibria. The risk is that they give pictures of consumers that do not predict real market behavior. It Moshe Ben-Akiva, Daniel McFadden and Kenneth Train (2019), “Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analy- sis”, Foundations and Trends R in Econometrics: Vol. 10, No. 1-2, pp 1–144. DOI: 10.1561/0800000036. The version of record is available at: http://dx.doi.org/10.1561/0800000036

Upload: dinhcong

Post on 07-Jun-2019

236 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

Foundations of Stated PreferenceElicitation: Consumer Behavior andChoice-based Conjoint AnalysisMoshe Ben-Akiva1, Daniel McFadden2,3 and Kenneth Train2

1Department of Civil and Environmental Engineering,MIT Cambridge, MA 02139, USA2Department of Economics, University of California,Berkeley, CA 94720-3880, USA3Department of Economics, University of Southern California,CA 90089-921, USA

ABSTRACTStated preference elicitation methods collect data onconsumers by “just asking” about tastes, perceptions, val-uations, attitudes, motivations, life satisfactions, and/orintended choices. Choice-Based Conjoint (CBC) analysisasks subjects to make choices from hypothetical menus inexperiments that are designed to mimic market experiences.Stated preference methods are controversial in economics,particularly for valuation of non-market goods, but CBCanalysis is accepted and used widely in marketing and policyanalysis. The promise of stated preference experimentsis that they can provide deeper and broader data on thestructure of consumer preferences than is obtainable fromrevealed market observations, with experimental control ofthe choice environment that circumvents the feedback foundin real market equilibria. The risk is that they give picturesof consumers that do not predict real market behavior. It

Moshe Ben-Akiva, Daniel McFadden and Kenneth Train (2019), “Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analy-sis”, Foundations and Trends©R in Econometrics: Vol. 10, No. 1-2, pp 1–144. DOI: 10.1561/0800000036.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 2: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

2

is important for both economists and non-economists tolearn about the performance of stated preference elicitationsand the conditions under which they can contribute tounderstanding consumer behavior and forecasting marketdemand. This monograph re-examines the discrete choicemethods and stated preference elicitation procedures thatare commonly used in CBC, and provides a guide totechniques for CBC data collection, model specification,estimation, and policy analysis. The aim is to clarify thedomain of applicability and delineate the circumstancesunder which stated preference elicitations can providereliable information on preferences.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 3: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

Preface

Information on consumer preferences and choice behavior is needed toforecast market demand for new or modified products, estimate theeffects of product changes on market equilibrium and consumer welfare,develop and test models of consumer behavior, and reveal determinantsand correlates of tastes. Direct elicitation of stated preferences, per-ceptions, expectations, attitudes, motivations, choice intentions, andwell-being, supplementing or substituting for information on revealedchoices in markets is potentially a valuable source of data on consumerbehavior, but can mislead if the information environments and decision-making processes invoked by direct elicitations differ from the settingsfor revealed choices in real markets.

The purpose of this monograph is to provide the reader with statedpreference data collection methods, discrete choice models, and statisti-cal analysis tools that can be used to forecast demand and assess welfareimpacts for new or modified products or services in real markets, andsummarize the conditions under which the reliability of these methodshas been demonstrated or can be tested. One focus is the collection ofpreference and related data from consumer responses in hypotheticalchoice experiments, particularly choice-based conjoint analysis (CBC)methods that have proven useful in market research. Another is theeconomic theory and statistical analysis of choice behavior, revealedor stated, and an economic framework for forecasting market demand

3

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 4: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

4

and measuring consumer welfare. Stated choice data can be collectedand combined with a broad spectrum of models of consumer behavior.This monograph will focus on the standard economic model of utility-maximizing consumers. Our treatment is informed by and benefits fromexperiments on perceptions and decision-making behavior in cognitivescience and behavioral economics, and includes methods that can accom-modate features of consumer choice that impact forecast reliability suchas anchoring, adaptation to the status quo, and sensitivity to context.However, we will only touch on the implications of behavioral consumertheory for elicitation and analysis of stated preference data.

There are a number of good introductions to discrete choice analysis(Ben-Akiva and Lerman, 1985; Train, 1986; Train, 2009; McFadden,1999; McFadden, 2001; McFadden, 2014b; Brownstone, 2001; Boyce andWilliams, 2015), and to stated preference and conjoint analysis methodsand market research applications (Louviere, 1988; Fischhoff and Manski,1999; Louviere et al., 2000; Wittink and Bergestuen, 2001; Hauser andRao, 2002; Rossi et al., 2005; Chandukala et al., 2007; Raghavarao et al.,2010; Rao, 1977; Rao, 2014; Green and Srinivasan, 1978; Green andSrinivasan, 1990). This monograph complements these introductions byfilling in technical and behavioral backgrounds for these topics.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 5: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

1Some History of Stated Preference Elicitation

Stated preference methods date from the 1930s. The originator ofempirical demand analysis, Henry Schultz, persuaded his University ofChicago colleague, the iconic psychologist Leon Thurstone (1931), topresent a paper at the second meeting of the Econometric Society withthis proposal for direct elicitation of indifference curves:

“Perhaps the simplest experimental method that comes tomind is to ask a subject to fill in the blank space [to achieveindifference] in a series of choices of the following type:

‘eight hats and eight pairs of shoes’ versus ‘six hatsand___pairs of shoes’

One of the combinations such as eight hats and eight pairsof shoes is chosen as a standard and each of the othercombinations is compared directly with it.”

Thurstone postulated that responses would obey Fechner’s law, a com-mon psychophysical regularity in the sensation produced by a stimulus.This turns out to be equivalent to postulating that respondents maxi-mize a log-linear utility function. He collected experimental data on hats

5

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 6: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

6 Some History of Stated Preference Elicitation

vs. shoes, hats vs. overcoats, and shoes vs. overcoats, fit the parametersof the log-linear utility function to data from each comparison, treatingresponses as bounds on the underlying indifference curves, and testedthe consistency of his fits across the three comparisons.

At the time of Thurstone’s presentation, empirical demand analysiswas in its early days. Pioneering studies of market demand for a singlecommodity (sugar) had been published by Frisch (1926), Schultz (1925),and Schultz (1928), but there were no empirical studies of multi-productdemands. Least-squares estimation was new to economics, and requiredtedious hand calculation. Consolidation of the neoclassical theory ofdemand for multiple commodities by Hicks (1939) and Samuelson (1947)was still in the future. Given this setting, Thurstone’s approach waspath-breaking. Nevertheless, his estimates were rudimentary, and hefailed to make a connection between his fitted indifference curves andmarket demand forecasts.1 Most critically, he did not examine whetherthe cognitive tasks of stating indifferent quantities in his experimentand of choosing best bundles subject to a budget constraint in realmarkets were sufficiently congruent so that responses with respect tothe first would be predictive for the second.

According to Moscati (2007), Thurstone’s presentation was crit-icized from the floor by Harold Hotelling and Ragnar Frisch. First,they objected that Thurstone’s indifference curves as constructed wereinsufficient to forecast market demand response to price changes; thisobjection failed to recognize that extending Thurstone’s comparisons toinclude residual expenditure could have solved the problem. Second, theypointed out that the knife-edge of indifference is not well determined

1In retrospect, these two flaws were correctable: Denote by H, S, C, respectively,the numbers of hats, pairs of shoes, and coats consumed, let M denote the moneyremaining for all other goods and services after paying for the haberdashery, andlet Y denote total income. Suppose Thurstone had asked subjects for the amountsM that made a comparison bundle (H,S,C,M) indifferent to a standard bundle(H0, S0, C0,M0). Then, he could have estimated the parameters of the log-linearutility function u = logM +θH logH+θS logS+θC logC by least squares regressionof log(M/M0) on log(H0/H), log(S0/S), and log(C0/C). From this, he could haveforecast demands, e.g., hat demand at price pH and income Y would have been givenby the formula H = θHY/pH(1+θH +θS +θC) that comes from utility maximizationsubject to the budget constraint.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 7: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

7

in comparisons of bundles of discrete commodities.2 Beyond theseobjections, Frisch and Hotelling were generally skeptical that statedindifference points, or non-market responses more generally, could beused to predict market behavior. The orthodoxy of that era, formedpartly as a reaction to the casual introspections of Bentham and theutilitarians, was that empirical economics should rely solely on revealedmarket data; in the words of Irving Fisher (1892), “To fix the ideaof utility, the economist should go no further than is serviceable inexplaining economic facts. It is not his province to build a theory ofpsychology.” Wallis and Friedman (1942) summarized this attitude inan attack that forcefully dismissed Thurstone’s method or any otherattempt to use experimental data for market demand analysis, pointingout difficulties in designing experiments that mimic the environment ofreal market choices: “[Thurstone’s] fundamental shortcomings probablycannot be overcome in any experiment involving economic stimuli andhuman beings.”

For 40 years following Thurstone’s paper, consumer experimentswere mostly limited to testing axioms for choice under uncertainty,and there was no systematic attempt to incorporate stated preference(SP) methods in demand analysis. There was some reason for thislack of interest. The language of economic analysis, then and now, isprediction of market demand and assessment of market failures in termsof dollars of equivalent lost income. Any measurement method that usesexperimental data on preferences has to produce convincing results inthis language by showing that stated preferences collected outside themarket have the same predictive power for market behavior as impliedpreferences reconstructed from market data. With the advent of behav-ioral economics, we have learned that people are often not relentlessutility maximizers, either in markets or in experiments, undermining thetight links neoclassical consumer theory postulates between consumerutility and demand behavior. This has led to calls for less focus on

2Additional objections could have been raised about the applicability of Fechner’slaw and the restrictiveness and realism of the log linear utility specification, lackof accounting for heterogeneity across consumers, and lack of explicit treatment ofconsumer response errors. Decades later, when this demand system was fitted torevealed preference (RP) data, these issues did arise.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 8: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

8 Some History of Stated Preference Elicitation

assessment of consumer welfare in dollars of lost income deduced frommarket demand behavior, and more research on benchmarking statedsatisfaction to psychological or neurological measures of well-being. Thisapproach may eventually succeed, but at present market prediction andvaluation remain the yardsticks against which any method for elicitingconsumer preferences and inferring consumer welfare has to be judged.

The first sustained use of stated preference methods came out of thetheory of conjoint measurement pioneered by Luce and Tukey (1964)and Luce and Suppes (1965), and developed as conjoint analysis bymarket researchers like Green (1974), Johnson (1974, 1999), Huber(1975, 1987), Srinivasan (1988) and Louviere (1988) and applied to thestudy of consumer preferences among familiar market products (e.g.,carbonated beverages, automobiles). Good introductions to conjointexperiments, data, and analysis methods are Louviere et al. (2000)and Rossi et al. (2005). A central feature of conjoint analysis is theuse of experimental designs that allow at least a limited mapping ofthe preferences of each subject, and multiple measurements that allowestimates of preferences to be tested for consistency and refined whennecessary.

Early conjoint analysis experiments described hypothetical productsin terms of price and levels of attributes in various dimensions, andasked subjects to rank attributes in importance, and rate attributelevels. These measurements were used by market researchers to classifyand segment buyers, and target advertising, but they were not reliabletools for predicting market demand. However, Louviere and Woodworth(1983) and Hensher and Louviere (1983) introduced choice-based con-joint (CBC) elicitations that directly mimicked market choice tasks,and McFadden et al. (1986) and McFadden (1986) showed how theseelicitations could be analyzed using the tools of discrete choice analysisand the theory of random utility maximization (RUM).3 Subjects would

3The term “CBC” is used in marketing to include stated choice elicitationswithout an underlying framework of utility-maximizing discrete choice. More explicitterminology for the approach discussed in this monograph would be “CBC/RUM”,or as suggested by Carson and Louviere (2011), “discrete choice experiments” (DCE).However, we will continue to use the umbrella term “CBC”, leaving it to the readerto distinguish our economic approach from other forms of conjoint analysis.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 9: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

9

be presented with a series of menus of products. Each product offered ineach menu would be described in terms of price and levels of attributes.Subjects would be asked to choose their most preferred product in eachmenu. For example, subjects would be offered a menu of paper towels,with each product described in terms of price, number of towel sheets, ameasure of the absorption capacity, a measure of strength when wet, andbrand name. Choice data from these menus, within and across subjects,could then be handled in the same way as the real market choice data.Choice-based conjoint (CBC) surveys analyzed using discrete choicemethods have become widely used and accepted in market researchto predict the demand for consumer products, with a sufficient trackrecord so that it is possible to identify some of the necessary conditionsfor successful prediction; see McFadden et al. (1988) and Green et al.(2001), Cameron and DeShazo (2013), and McFadden (2014, 2017).

Environmental economists developed independently a simple statedpreference method termed contingent valuation (CV), and appliedit to valuing environmental damage. This method traces its begin-nings to a proposal by Ciriacy-Wantrup (1947) and a PhD thesis byDavis (1963a, 1963b) on the use-value of Maine woods. Its first pub-lished applications for values of environmental public goods seem tohave been Brookshire et al. (1976), Bishop and Heberlein (1979), andBrookshire et al. (1980). CV can be viewed as a truncated form ofconjoint analysis with two important differences. First, it does nothave the experimental design features of conjoint analysis that al-low extensive tests for the structure and consistency of stated prefer-ences. Second, because of its applications, it usually does not havepredictive accuracy in markets as a direct yardstick for reliability.Instead, it relies indirectly on the links between preferences, mar-ket demands, and valuations that hold when neoclassical consumertheory is valid, on analogies with stated preference studies of con-sumer goods in markets, and on limited internal consistency checks.The particular challenges of using contingent valuation for natural re-source valuation are discussed by Carson et al. (2001), Carson (2012),Hausman (2012), and Kling et al. (2012). Its reliability in relation tostated preference elicitation for market goods is discussed by McFadden(2017).

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 10: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

10 Some History of Stated Preference Elicitation

Other elicitation methods for stated preferences, termed focus groups,vignette analysis, and measurement of subjective well-being, are popularamong some applied economists and political scientists, see Rossi (1979),King et al. (2004), Caro et al. (2012), and Kahneman and Krueger (2013).Focus group methods are directed open-end discussions of productsand their features in small samples of consumers. Focus groups do notprovide direct data for market demand forecasts, but they can be quiteuseful in designing CBC experiments because of the insights they pro-vide on product perceptions, levels of consumer knowledge, experience,and understanding, and the attributes consumers consider in formingtheir preferences. Vignette analysis uses detailed story descriptions ofalternatives, often visual. Vignette presentations of alternatives can beused within conjoint analysis experiments, and may improve subjectattention and understanding of alternatives. Subjective well-being meth-ods elicit overall self-assessments of welfare, often on Likert or ratingscales similar to those used in the early days of conjoint analysis. In theinstances where vignette and subjective well-being methods have beentested, they prove to be strongly influenced by context and anchoringeffects, see Deaton (2012). These effects compromise forecast accuracy inmarket demand forecasting applications. Psychometrics has developedan array of additional methods for measuring perceptions, attitudes,and motivations. Their usefulness for economic demand forecasting hasnot been demonstrated, but at least for perceptions and intentions it isclear that further development is potentially quite valuable for economicapplications.

The focus of this monograph is market demand forecasting for newor modified products; we do not attempt here any overall assessmentof the reliability of contingent valuation, vignette analysis, or subjec-tive well-being methods in their primary uses. We urge readers to notcasually pre-judge the use of stated preference methods in economicapplications, but rather to acquire the tools needed to conduct andcritique CBC studies, investigate how well these methods work un-der various conditions, and make a reasoned scientific judgement onwhen their use advances our understanding of consumer behavior andwell-being.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 11: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

2Choice-Based Conjoint (CBC) Analysis

A choice-based conjoint (CBC) study offers a consumer a series of menusof alternative products with profiles giving levels of their attributes,including price, and asks the subject to choose the most preferred prod-uct in each menu.1 The menus of products and their descriptions aredesigned to realistically mimic a market experience where a consumercan choose among various competing products. By changing the at-tribute levels for the included products and presenting each consumerwith several menus, a CBC experiment obtains information on therelative importance that consumers place on each of the attributes.A classic CBC setup in marketing might be a laboratory experimentwhere subjects are offered a sequence of binary menus of actual productswith varying attributes and prices, and asked to indicate the preferred

1This setup assumes that menu alternatives are mutually exclusive. Applicationswhere the consumer buys more than one unit of the same or multiple products canbe handled in this framework by listing each possible combination of purchases as aseparate menu item. Alternately, CBC subjects could be allowed to check the numberof units of each product they would purchase, with zero indicating “no purchase”,as they do when placing product orders on Amazon. This alternative requires thatthe discrete choice models used to analyze stated preference data be modified toencompass utility-maximizing counts, see Burda et al. (2012), Bhat et al. (2014),and Chen (2017).

11

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 12: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

12 Choice-Based Conjoint (CBC) Analysis

alternative in each menu. For example, subjects might be trained bytasting cola drink “brands” with various degrees of sweetness, carbona-tion, and flavor, and then asked in each menu to choose one brand–pricecombination or the other. They may be promised that at the end ofthe experiment, they will receive a six-pack of their chosen brand fromone of the menus they faced, plus a participation payment less themenu price of the six-pack. Often, CBC is used in marketing studiesof familiar products whose features can be described in words andpictures, with subjects asked to choose from a menu of products basedon these descriptions. These studies often employ telephone, internet,or mail survey instruments, and recruit subjects using sampling framesthat range from stratified random samples of the target population tosamples obtained from internet volunteers, intercepts at retail outlets,or registered product buyers. For example, experiments on automobilebrand and model choice have described alternatives in terms of priceand attributes such as horsepower, fuel consumption, number of seats,and cargo space, and collected data using interactive internet sessionswith randomly sampled consumers, see Urban et al. (1990), Urbanet al. (1997), and Brownstone and Train (1999). These studies havedetermined with considerable predictive accuracy in the distributionsof preference weights that consumers give to various vehicle features.

2.1 Issues in CBC study design

Based on the authors’ experience in conducting stated preference ex-periments, and our review of the literature, we conclude that there area number of key issues that need to be addressed in any CBC study,see also Carson et al. (1994) and McFadden (2018). We summarizethese in the checklist given in Table 2.1, and discuss them further inthe paragraphs that follow.

2.1.1 Familiarity

The selection of subjects, the setup and framing of the elicitation, train-ing of subjects where necessary, and testing of subjects’ understandingand consistency are critical in a CBC study, particularly when someproducts are novel with attributes that are not easily experienced or

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 13: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

2.1. Issues in CBC study design 13

Table 2.1: CBC study design checklist.

Issue Discussion1 Familiarity: Subjects should be familiar with the class of

products studied and their attributes, after training ifnecessary, and experienced in making market choicesamong them

2.1.1

2 Sampling, Recruitment, Background: Subjects should besampled randomly from the target population, andcompensated sufficiently to assure participation andattention. Background on socioeconomic status andpurchase history should be collected for sample correctionand modeling of heterogeneity in choice behavior

2.1.2

3 Outside Option: Menus should include a meaningfullyspecified “no purchase” alternative, or the study should bedesigned to use equivalently identifying externalbackground or real market data

2.1.3

4 Menu Design: The number of menus, products per menu,attributes per product, and span and realism of attributelevels must balance market realism and the independentvariation needed for statistical accuracy

2.1.4

5 Attribute Formatting: The clarity and prominence ofattributes should mimic, to the extent possible given thegoals of the analysis, the information environment theconsumer will face in the real market

2.1.5

6 Elicitation Frame: Elicitation formats other thanhypothetical market choice must balance information valueagainst the risk of invoking incompatible cognitive frames

2.1.6

7 Incentive Alignment: Elicitations with a positive incentive fora truthful response reduce the risk of carelessness or casualopinion

2.1.7

8 Subject Training: Training may be necessary to providesufficient familiarity and experience with products, but forreliable forecasting it needs to mimic the “street” trainingprovided by real markets and minimize manipulation thatis not realistic

2.1.8

9 Calibration and Testing: Consistency and reality checksshould be built into the study, and forecasting accuracyshould, if possible, be tested against external data

2.1.9

understood, or subjects have not had market experience with similarproducts. CBC studies can forecast market demand relatively well whenthe subjects’ task is choice among a small number of realistic, familiar,

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 14: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

14 Choice-Based Conjoint (CBC) Analysis

and fully described products, and subjects have had experience buyingproducts of this type and can realistically expect to make further pur-chases in the future. However, there is a sharp reliability gradient, andstated preferences become problematic when the products are unfamil-iar or incompletely described, or involve public good or social aspectsthat induce respondents to consider social welfare judgments or peeropinions. For example, when stated preferences for automobile modelsare collected in an elicitation that emphasizes the energy footprint andenvironmental consequences of the alternatives, subjects may give theseattributes more weight than they would in a real market, or state moresocially responsible preferences than they would reveal in actual marketchoices.2

Familiarity may be a given when the CBC study is of ordinaryhousehold products such as cola drinks, but not when novel productsor novel or highly technical product features are outside the subject’srange of experience. Where possible, subjects should have an opportunityto “test-drive” novel products. For example, in a study of consumerchoice among streaming music services, it should improve prediction togive subjects hands-on experience with different features, in a settingthat mimics their opportunities to investigate and experience thesefeatures in real shopping. In some cases, this can be accomplished withactual or mock-up working models of the offered products, or withcomputer simulation of their operation. There is a tradeoff: Attemptingto train consumers and providing mock-ups can inadvertently createanchoring effects. Consumers who are unfamiliar with a product maytake the wording in the training exercises about the attributes, orthe characteristics of the mockups, as clues to what they should feelabout each attribute. Even the mention of an attribute can give it moreprominence in a subject’s mind than it would have in real shopping. The

2This is not an argument that energy footprint should be excluded from theattribute list in studying automobile choice, but rather a caution that when it isincluded, responses are likely to be quite sensitive to how statements about energyfootprints and environmental consequences are framed. In general, valuation ofnon-use aspects of natural resources is challenging for conjoint methods becausethese applications seek to measure preference that lies outside the normal marketexperience of consumers and may invoke sentiments regarding social responsibilitythat are partially formed and unstable.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 15: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

2.1. Issues in CBC study design 15

researcher needs to weigh the often conflicting goals of making subjectsknowledgeable about products and avoiding distortion of the relativevaluations of attributes these subjects would exhibit in real markets.

2.1.2 Sampling, recruitment, background

Target populations may differ depending on the objectives of the study —e.g., all current users of a class of products, only technologically savvycurrent users, the general population. For many policy applications, itis a major blunder to sample only current buyers of a class of products,as the preferences and behavior of non-buyers are normally also neededto determine the effect of policy changes on market equilibrium, thepreferences of marginal buyers in equilibrium, and welfare consequencesfor both initial non-buyers and buyers. It is important that the samplingframe draws randomly from the target population, without excessiveweighting to correct for stratification and non-response. It is generallya bad idea to use opportunistic pools of potential respondents, such assubjects recruited from volunteers at internet survey sites or interceptedat shopping malls or transportation nodes, as these pools may differ sub-stantially from the target population in unobserved characteristics, evenafter weighting (“racking”) to match the target population distributionon observed dimensions such as age and education.

It is sometimes desirable to sample from a target population thathas experience or expertise with a class of products. Then it may beinformative to study the preferences of experienced users, and separatelystudy the differences in tastes of these users and the general popula-tion. An example is study of consumer demand for relatively esoterictechnical attributes of products, say the levels of encryption built intotelecommunication devices, where only technically savvy device userswill appreciate the meaning of different encryption levels. In this case,a good study design may be to conduct an intensive conjoint analysison technically knowledgeable users, and separately survey the targetpopulation to estimate the extent and depth of technical knowledge,and the impact of technical information on the purchase propensities ofcurrent general users and non-users.

Background information on socioeconomic status and product pur-chase history should be collected on subjects in a CBC study, to enable

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 16: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

16 Choice-Based Conjoint (CBC) Analysis

weighting to correct for sampling imperfections such as uneven attritionrates, and to provide the foundation for estimation of choice modelsthat are accurate for various demographic groups. Importantly, wage,asset, and transfer income are needed to model income effects in choicebehavior. Relatively universal internet access has led to inexpensiveand effective surveying using internet panels, rather than telephone,mail, or personal interviews. However, it is risky to recruit internetsurvey subjects from volunteers, as they can look quite different thana target population of possible product buyers. Better practice is touse a reliable sample frame such as random sampling of addresses, thenrecruit internet panel subjects from the sampled addresses. It is alsoimportant to compensate subjects for participation at sufficient levelsto minimize selection due to attrition; see McFadden (2012). Experiencewith “professional” subjects who are carefully recruited and paid toparticipate in internet panels, with extensive background informationcollected prior to the CBC study, is positive: Subjects who view re-sponding as a continuing “job” with rewards for effort seem to be moreattentive and responsive than casual respondents.

2.1.3 Outside option

Obviously, a consumer’s history and future opportunities influenceproduct choice. If the consumer has in the past purchased substitutesor complements that are still on hand, this affects the desirability ofitems on a current menu. Further, if a current menu choice opens, closes,or changes the attractiveness of options in the future, the utility ofa product purchase will include allowance for its future impacts. Forexample, faced with a menu of cars with various attributes and prices,a consumer will consider how a current choice will influence the useor disposition of vehicles already owned, and how this choice would fitnot only with anticipated needs, but also with future purchases. Thetiming of consumer durables and perishable restocking purchases is animportant aspect of consumer behavior in real markets. A reliable CBCstudy must give subjects a decision-making context that is realistic,taking account of their market history, and providing clear instructionson purchase timing and how current stated menu choices affect future

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 17: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

2.1. Issues in CBC study design 17

options. Critically, if subjects strategically “game” their experimentalresponses based on their perceived market opportunities in other timesand places, and these perceptions do not match strategic opportunitiesin a future real market, then CBC forecasts for this real market willtend to be unreliable.

For a neoclassical consumer, the decision to buy at a particulardecision point is made by comparing the maximum utility from apurchase now and the utility of “no purchase”, where the latter is theoption value of opportunities at future times and places for choices thatcomplement or substitute for a purchase now. A CBC study that offersmenus of products without a “no purchase” option can predict marketshares, but not overall product demand. To go further and answerimportant questions such as the extent to which new products expanddemand rather than simply capture market share, one can either turnto external market data to calibrate product demand, or include a “nopurchase” option in the CBC menus. Both strategies present designissues.

Consider first external calibration. If real market demand data withexisting products, attributes, and prices are available at the degree ofgranularity on demographics and purchase history attainable from CBCsubjects, then it is straightforward to start from a model fitted to CBCdata without a “no purchase” option, calibrate the model’s “no purchase”utility as a function of history and socioeconomics so that it matchescurrent demand under current market conditions, and then use thiscalibrated model to forecast future demand. A more general approachstacks CBC product choice data and baseline real market data, with thelatter broken down by available demographics and history, and estimatesa pooled model for forecasting. The accuracy of external calibrationis limited by the granularity of real market data, as it cannot capturevariations in purchase probability arising from unobserved differencesin consumer histories. For example, it is unusual to have real marketdemand broken down by buyers’ income, age, or purchase history, whilethese are commonly collected as background in CBC studies. Reliableexternal calibration obviously requires consistency in the framing ofstudy questions on real market behavior and the definitions used incollecting real market demand data.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 18: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

18 Choice-Based Conjoint (CBC) Analysis

Consider next the strategy of including a “no purchase” option inCBC elicitations. One can simply make “no purchase” as an alternativeon some or all menus. Alternately, one can ask first for the best productchoice from a menu without a “no purchase” option, and follow up theresponse by asking if the subject would in fact buy this chosen product.An advantage of the alternative approach is that it keeps subjects fromusing a “no purchase” choice as a way to avoid the time and effortrequired to evaluate the offered products; it may also mimic the wayconsumers make choices in real markets for complex products suchas cars or houses. However, a drawback of the alternative approachis that unless strong separability conditions on preferences are met,choice probabilities among products without a “no purchase” optionwill not in general coincide with product choice probabilities condi-tioned on a decision to purchase. Another drawback is that conditionalproduct choice may unrealistically color subsequent stated purchasechoice.

An issue with any method of including a “no purchase” option inCBC elicitations is that its meaning needs to be carefully specifiedfor subjects. If a subject states “no purchase” now, what are theybeing asked to assume about their possibilities for hypothetical and realmarket purchases in the future? If “no purchase” in the CBC setting isinterpreted differently by different subjects, and inconsistently with realmarket data definitions such as the numbers of units of products soldin a calendar quarter, a CBC study will fail the mechanical accountingrequirement for reliable forecasting that the CBC menu and the realmarket forecast have consistent purchase periods. For example, in a carchoice study, a subject might interpret a “no purchase” option offeredwithout explanation in a CBC study as meaning that they wouldcontinue to use indefinitely a vehicle that the household currently owns;that they would do without a new car in the current model year; or thatthey could at any time go to the real market as an alternative to havinga product choice from a CBC menu fulfilled. These interpretations areall inconsistent with a study goal of forecasting the number of carssold in the next model year. Better practice in CBC design is to makethe meaning of “no purchase” specific and explicit, and establish thatsubjects understand and accept it.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 19: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

2.1. Issues in CBC study design 19

2.1.4 Menu design

The design of a conjoint experiment establishes the number of menusoffered to each subject, the nature and number of products on eachmenu, the number of attributes and attribute levels introduced foreach product, and the design of the profiles of the products placed oneach menu. The levels of attributes of the products offered on differentmenus can be set by experimental design, so that it is possible toseparate statistically the weights that consumers give to the differentattributes. In its early days, menu designs were often of a “completeprofile” form that mimicked classical experimental design and allowedsimple computation of “part-worths” from rating responses, but cur-rently the emphasis is simply on ensuring that menus are realistic andincorporate sufficient independent variation in the attributes, so thatthe impact of each attribute on choice can be isolated statistically. Theclassical statistical literature on experimental design focused on analysisof variance and emphasized orthogonality properties that permittedsimple computation of effects, and treatments that provided minimumvariance estimates. Designs that reduce some measure of the samplingvariance under specified model parameters (such as the determinantof the covariance matrix for “D-efficiency”) have been implemented inmarket research by Kuhfield et al. (1994), Bleimer and Rose (2009),and Rose and Bleimer (2009), and others. It is important that con-joint studies be designed to yield good statistical estimates, but thereis relatively little to be gained from adherence to designs with clas-sical completeness and orthogonality properties. With contemporarycomputers, the computational simplifications from orthogonal designsare usually unimportant. Further, for the nonlinear models used withCBC, orthogonality of attributes does not in general minimize samplingvariance. Unlike classical analysis of variance problems, it is not usuallypossible in nonlinear choice models to specify efficient designs in advanceof knowing the parameters that are the target of the analysis; see alsoWalker et al. (2018).

In real markets, as a result of competition, surviving productsare often similar in their attributes and prices. Much of the powerof a CBC study is the opportunity to identify preferences in menus

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 20: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

20 Choice-Based Conjoint (CBC) Analysis

with independent variation in attributes that are highly correlated inreal markets. However, relatively mechanistic statistical approaches tosetting attribute levels in a CBC study may lead to profiles that are un-realistic or impractical. Menus and their framing that are unlike familiarreal market menus can jar the subject and invite cognitive framing thatdiffers from the frames that express preferences and drive choices in mar-ket settings. Considerable care is needed to balance statistical objectiveswith realism of the experiment, see Huber and Zwerina (1996).

2.1.5 Attribute formatting

The formatting, clarity, and prominence of attributes and prices ofproducts are critical aspects of real market environments, and arecorrespondingly important in the hypothetical market menus in CBCstudies. Advertising and point-of-sale product presentations in realmarkets often feature “hooks” that attract the consumer’s attentionand make products appealing, and understate or shroud attributesthat may discourage buyers. Thus, prices may not be prominentlydisplayed, or may be presented in a format that shrouds the final cost;e.g., promotions of “sales” or “percent-off” discounts without statingprices, statements of prices without add-ons such as sales taxes andbaggage fees, subscriptions at initial “teaser” rates. Products like mobilephones, automobiles, and hospital treatments are often sold with totalcost obscured or shrouded through ambiguous contract terms, througha two-part tariff that combines an access price and a usage fee, orthrough framing (e.g., “pennies a day”). A CBC study that is reliablefor forecasting evidently needs to mimic the market in its presentationof product costs, incorporating the same attention-getting, persuasion,ambiguities and shrouding that consumers see in the real market.3However, a CBC study whose goal is to estimate consumers’ “genuine”willingness to pay for attributes, rather than forecast market demand,needs the subjects to understand the tradeoffs clearly without maskinginformation in the way that often occurs in real markets.

3A CBC study may embed alternative presentations of product features andcosts, so that the influence of presentation format on choice can be quantified andpredicted.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 21: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

2.1. Issues in CBC study design 21

Prominence and ease of comparison are known to be factors thatinfluence the attention subjects give to different aspects of decision prob-lems; for example, market researchers sometimes report that subjects intheir stated choices systematically place more weight on price relative toother product attributes than they do in real markets,4 perhaps becausethis dimension is clearly visible and comparisons are easy in a conjointanalysis menu, whereas prices in real markets often come with qualifica-tions and may not be displayed side-by-side. Widespread folklore is thatsubjects have trouble processing more than six attributes and more thanfour or five products, and begin to exhibit fatigue when making choicesfrom more than 20 menus, see Johnson and Orme (1996). However,other researchers argue that many more attributes and products can beincluded, as long as the subjects understand how the attributes affectthem, see Louviere et al. (2010) and Hensher et al. (2011).5 To achieveverisimilitude with complex markets such as those for houses, cars, andjobs, and induce similar decision protocols, it can be desirable to allowCBC menus to be comparably complex. Often conjoint analysts willlimit the dimensionality of the attribute profile, presenting incompleteprofiles in which subjects are explicitly or implicitly asked to assumethat in excluded attribute dimensions, the products are comparable tobrands currently in the market and/or the omitted attributes are heldequal across products. This fill-in problem leaves subjects free to makepossibly heterogeneous and unrealistic assumptions about these omittedattributes, or asks them to digest and remember lengthy specificationsfor omitted attributes and their assumed levels. These design restrictionsmay make responses easier to analyze, but they will degrade predictionif correlations of included attributes with the fill-ins actually assumed

4Alternatively, subjects may place less weight on price if payment is hypothetical,or real but made from discounted house money, in the CBC exercises.

5In a personal communication, David Hensher says that what matters in bothan SP and an RP setting is attribute relevance. Complexity is often in the eyes ofthe respondent, defined by what matters to them. To exclude relevant attributesin itself creates an element of complexity, since there is a mismatch between whatpeople take into account and what the analyst imposes on them. Hensher says thathe often uses a lot of attributes and 13–14 is not uncommon, and that “What makesit relevant is the way that we design the survey instrument so that respondents canactivate attribute processing rules to accommodate what matters to them.”

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 22: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

22 Choice-Based Conjoint (CBC) Analysis

by respondents are not reproduced in the real market; see Islam et al.(2007).

In real markets with many complex products, such as the marketfor houses, consumers often use filtering heuristics, screening out alter-natives that fail to pass thresholds on selected aspects before choosingamong the survivors, see Swait and Adamowicz (2001). If the primaryfocus of the conjoint study is prediction, then the best design may beto make the experiment as realistic as possible, with approximately thesame numbers and complexity of products as in a real market, and possi-bly sequential search, so that consumers face similar cognitive challengesand respond similarly even if decision-making is less single-minded thanneoclassical preference maximization. However, if the primary focus ismeasurement of consumer welfare, there are deeper problems in linkingwell-being to demand behavior influenced by filtering. While it may bepossible to design simple choice experiments that eliminate filtering andgive internally consistent estimates of willingness-to-pay for productattributes, there is currently no good theoretical or empirical founda-tion for applying these tradeoffs to assess well-being in real markets forcomplex products where filtering influences consumer behavior.

2.1.6 Elicitation frame

The canonical elicitation frame for a CBC study asks the subject whichalternative they would choose in a menu if it were offered in a real market,sometimes with the incentive that with some probability their selectionwill be fulfilled. This resembles real market choice environments, andexperience is that it elicits responses that for familiar products arepredictive for future market behavior. Studies of elicitation frames otherthan first choice, such as rating products on some scale, adjustingsome attribute (e.g., price) to make alternatives indifferent, or rankingproducts from best to worst seem to induce cognitive “task-solving”responses different from the task of maximizing preferences.6 Subjectsseem to approach less familiar choice tasks as if they were school exams —they cast about for “correct” answers by making inferences on what

6See Wright and Kriewall (1980), Chapman and Staelin (1982), and Elrod et al.(1992).

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 23: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

2.1. Issues in CBC study design 23

the experimenter is looking for. While some may use their responsesto air opinions, most give honest answers, but not necessarily to thequestion posed by the experimenter. They may “solve” problems otherthan recovering and stating their true preferences, indicating insteadthe alternative that seems the most familiar, the most feasible, the leastcost, the best bargain, or the most socially responsible, see Schkadeand Payne (1993).

For some products, it would seem that unfolding menu designs withelicitation of stated choices within groups of similar products, followedby elicitation of stated choices across these groups, or alternately designsin which subjects state conditional choices across groups, followed bystated choices that “customize” a preferred product in a chosen group,mimic market decision-making experience. For example, the last formatmimics the real market choice process for products like automobiles orcomputers where the buyer picks a product provisionally, then choosesoptions, with recourse if customization is unsatisfactory. However, evenelicitation formats that are apparently reasonably close to a marketexperience, such as asking for the next choice from a menu if the first isunavailable, may generate inconsistent responses as a result of anchoringto the initial response, self-justification, or a “bargaining” mind-set thatinvites strategic responses, see McFadden (1981), Green et al. (1998),Hurd and McFadden (1998), List and Gallet (2001), Louviere (1988),and Hanemann et al. (1991).

2.1.7 Incentive alignment

The idea behind incentives is that when subjects have a realistic chanceof really getting what they say they prefer, and they understand this,they have a positive disincentive to misrepresent their preferences andrisk getting an inferior outcome. Economic theorists have developed anumber of mechanisms for incentive-compatible elicitation of preferences.The simplest offers the subject a positive probability that every statedchoice will result in a real transaction. If subjects understand the offer,the probabilities are sufficiently high so that the subject does not dismissthem as negligible, and subjects view the transactions as being paid forfrom their own budgets rather than in terms of “house money” that

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 24: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

24 Choice-Based Conjoint (CBC) Analysis

they feel is not really theirs, then it is a dominant strategy for thesubject to honestly state whether or not a product with a given profileof attributes is worth more to them than its price; this is shown byGreen et al. (1998) for a leading variant of this mechanism due toBecker et al. (1964). However, experimental evidence is that peoplehave difficulty understanding, accepting, and acting on the incentives inthese mechanisms.7 Thus, it can be quite difficult in practice to ensureincentive alignment in a CBC, or to determine in the absence of strongincentives whether subjects are responding truthfully. Fortunately, thereis also considerable evidence that while it is important to get subjectsto pay attention and answer carefully, they are mostly honest in theirresponses irrespective of the incentives offered or how well they areunderstood, see Bohm (1972), Bohm et al. (1997), Camerer and Hogarth(1999), Yadav (2007), and Dong et al. (2010).

Where possible, conjoint studies should be designed to be incentive-compatible, so that subjects have a positive incentive to be truthful intheir responses. For example, suppose subjects are promised a Visarcash card, and then offered menus and asked to state whether or not theywould purchase a product with a profile of attributes and price, with theinstruction that at the end of the experiment, one of their choices fromone of their menus will be delivered, and the price of the product in thatmenu deducted from their cash card balance. If they never choose theproduct, then they get no product and the full Visa balance. If subjects

7Several mechanisms have been shown to be incentive-compatible in circumstanceswhere choices involve social judgments as well as individual preferences; for example,McFadden (2012) shows how the Clark–Groves mechanism can be used in an economicjury drawn at random from the population to decide on public projects. Smalltransaction probabilities are an issue. Suppose a CBC experiment on choice amongautomobiles offers a one in ten-thousand chance of receiving either a car with aselling price of $40,000 or $40,000 in cash, depending on stated choice. If the truevalue of the car to this subject is $V, then a truthful response is to choose the car ifand only if $V > $40,000. If a consumer declines the car when $V > $40,000, thenhis expected loss is $V/10,000–$4, a small number. This incentive is still enoughin principle to induce the rational consumer to state truthfully whether or not heprefers the car. However, misperceptions of low-probability events and attitudestoward risk may in practice lead the consumer to ignore this incentive or view it asinsufficient to overcome other motivations for misleading statements.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 25: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

2.1. Issues in CBC study design 25

understand, perhaps through training or experience, that it is in theirinterest to say they would choose a product if and only if its valueto them is higher than its price, then they have a positive incentiveto be truthful, and the experiment is said to be incentive-aligned orincentive-compatible.

In many situations it will not be practical to provide an incentive-compatible format while maintaining the objectives of the analysis.For example, the researcher might want to consider combinations ofattributes that have not yet been manufactured. Or, the cost of providinga product at any meaningful probability is prohibitive. For example,automobile manufacturers might want to test consumers’ reactions tonew features during the design phase of the manufacturers’ productdevelopment. For existing but expensive products such as cars, it isimpractical to offer subjects enough money to buy a new car and thenprovide them one of their chosen cars at the listed price. A lottery maybe incentive-compatible in principle, but the probabilities required tomake it practical (e.g., a one in a thousand chance of being offered thevehicle you chose in one menu at its stated price) may be so low thatsubjects do not take the offer seriously. While incentive compatibility isdesirable, CBC experiments without incentive compatibility have beenfound to be predictive in many settings, see Dong et al. (2010).

2.1.8 Subject training

Extensive experiments from cognitive psychology show that context,framing, and subject preparation can have large, even outsize, effectson subject response. It is particularly important that subjects havefamiliarity with the products and features they are being asked toevaluate that is comparable to their real market experiences, as attention,context, and framing effects are particularly strong when subjects areasked to respond in unusual or unfamiliar circumstances. Familiaritymay be automatic if the target population is experienced users of aparticular line of products. For inexperienced users, tutorials on theproducts and hands-on experience can reduce careless or distortedresponses, but may also influence stated preferences in ways that reduceforecasting accuracy.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 26: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

26 Choice-Based Conjoint (CBC) Analysis

It is useful to recognize that training of subjects can occur at severallevels, and that training can manipulate as well as educate, leadingto unreliable demand predictions. First, subjects have to get used toanswering questions that may be difficult or intrusive, and learn thatit is easier or more rewarding to be truthful than to fabricate. Some ofthis is mechanical: practice with using a computer for an internet-basedconjoint survey, and moving through screens, buttons, and branches ina survey instrument. Second, subjects need to be educated as to whatthe task of stating preferences is. Subjects can be trained in “Decision-Making 101” how to optimize outcomes with assigned preferences, andhow to avoid mistakes such as confusing the intrinsic desirability ofa product with process issues such as availability or dominance byalternatives. Such training can be highly manipulative, leading to be-havior that is very different from and not predictive for real marketchoices. But real markets are also manipulative, providing the “street”version of “Decision Making 101” that teaches by experience the con-sequences of poor choices. The goal of a conjoint study designed forprediction should be to anticipate and mimic the training that realmarkets provide. Third, the study designer needs to determine whatinformation will be conveyed to the subject, in what format, and as-sess what information the subject retains and understands. Typically aconjoint survey begins by describing the types of products the subjectwill be asked to evaluate, their major attributes, and the structure ofthe elicitation, asking for most preferred alternatives from a series ofmenus. Details may be given on the nature and relevance of particularattributes. Particularly important are sufficient explanations of unfa-miliar attributes, and where possible a cross-walk between describedattribute levels and historical or trial experience, to allow informedchoice. Instructions may be given on the time the subject has to re-spond, and what rules they should follow in answering. For example, thesurvey may either encourage or discourage the subject from consultingwith other family members, finding and operating past products in thesame line, or consulting outside sources of information such as internetsearches. Finally, subjects need to be instructed on the incentives theyface, and the consequences of their stated choices. At various stages

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 27: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

2.1. Issues in CBC study design 27

in subject training, they may be monitored or tested to determine ifthey have acquired information and understood it. For example, a pro-tocol in market research called information acceleration gives subjectsthe opportunity to interactively access product descriptions, consumerreviews, and media reports, and through click-stream recording andinquiries during the choice process collects data on time spent viewinginformation sources and impact on interim propensities. This protocolseems to improve subject attention and understanding of product fea-tures, and also identify the sources and content of information that hashigh impact on stated choices, see Urban et al. (1990) and Urban et al.(1997).

In summary, while training may educate subjects so they are familiarwith the products being compared, it is difficult to design trainingthat is neutral and non-manipulative. Real markets are in fact oftenmanipulative, via advertising and peer advice, and one goal for CBCis to achieve accurate prediction by mimicking the advertising andother elements of persuasion the consumer will encounter in the realmarket. One caution is that particularly in cases where preferences arenot well formed in advance, subjects will be particularly vulnerableto manipulation, and training that embodies manipulation that is notrealistic risks inducing stated responses that are not predictive for realmarket behavior.

2.1.9 Calibration and testing

Where possible, CBC results should be tested against and calibratedto consumer behavior in real markets. In some cases, CBC menuswill coincide with product offerings in existing markets, or can bedesigned to include such products. In this case, it is useful to comparemodels estimated from the CBC study and the market data to assesswhether people are weighing attributes similarly. Calibration is notlimited to specific targeted products or policies; it may in fact behelpful to test or impose consistency in stated preference data acrossdisparate choice situations when consistency is expected from rational

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 28: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

28 Choice-Based Conjoint (CBC) Analysis

consumers.8 Improved forecasts may be obtained by imposing realmarket constraints such as product shares on the estimation of choicemodels from CBC data, by calibrating CBC model parameters to satisfymarket constraints, or by combining CBC and market choice dataand estimating a combined model with scaling and shift parametersfor CBC data as needed. CBC studies should when possible embedtests for response distortions that are commonly observed in cognitiveexperiments, such as anchoring to cues in the elicitation format, referencepoint or status quo bias, extension neglect, hypersensitivity to context,and shadowing from earlier questions and elicitations. While some ofthese cognitive effects also appear to influence market choices, manyare specific to the CBC experience and have the potential to reduceforecasting accuracy. Ideally, a CBC study with adequate calibrationwill not show much sensitivity in its bottom-line forecasts to thesesources of possible response distortion.

An advantage of CBC experimental designs is that through thepresentation of a slate of menus, they give an opportunity to test theconsistency of individual stated choices with neoclassical preference the-ory, to confront respondents and ask them to explain and reconcile statedchoices, and to incorporate menus that allow direct cross-validationbetween stated and revealed market choices. For example, menus canbe offered that allow for the possibility of testing whether stated choicesare consistent with the axioms of revealed preference, and specificallywhether they violate the transitivity property of preferences. For exam-ple, even under the relaxed standard that consumers have stochasticpreferences with new preference draws for each choice, their responsescan be tested for the regularity property that adding alternatives cannotincrease the probability that a previous alternative is chosen. If menuscontain current market alternatives, and past purchase behavior ofthe subjects is known, then one can test whether revealed and stated

8In a personal communication, Danny Kahneman suggests that stated preferencesfor environmental policies might be improved by eliciting rankings of values fordisparate policy interventions, possibly including past social decisions that establishcalibrating precedents. The task of constructing valuation scales from such data issimilar to that of developing psychological scales from test inventories where itemresponses are noisy.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 29: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

2.1. Issues in CBC study design 29

preferences for the same alternatives are consistent in their weightingof attributes. For example, Morikawa et al. (2002) find that there aresystematic differences in weights given to attributes between stated andrevealed choices, and that predictions from stated choices can be sharp-ened by calibrating them to revealed preferences; see also Ben-Akivaand Morikawa (1990) and Brownstone et al. (2000), and Hensher andBradley (1993). This step of testing and validating CBC is importantparticularly in studies where verisimilitude of the conjoint menus andcongruity of the cognitive tools respondents use in experimental andreal situations are in question, for example, when the products beingstudied are complex and unfamiliar, such as choices of college, house topurchase, cancer treatment to pursue, or remedies for environmentaldamages.9 For example, the battery of consistency tests conducted onthe contingent valuation method of eliciting stated preferences givenin McFadden and Train (2017) could in many cases be generalized andembedded in any CBC study.

In marketing applications, it is possible to validate CBC forecastsagainst actual market performance of new or modified products, judgedby market shares in the population or in population segments. We havenot found a comprehensive survey of the performance of forecasts fromCBC studies. Natter and Feurstein (2002) compare consumer-level CBCforecasts with scanner data on actual purchases, and conclude thataccounting for individual heterogeneity leads to market-level forecastsno better than aggregate models. However, they do not use a statis-tical method that accounts for the unreliable estimation of individualpreference weights. Moore (2004) compares CBC with other elicitationand forecasting methods, and concludes that CBC data analyzed usingthe methods described in this paper out-performed the alternatives.Wittink and Bergestuen (2001) cite studies in which CBC forecastsof market shares of data terminals, commuter modes, state lottery

9A large literature compares and tests stated preference elicitation methods, andis relevant to questions of CBC reliability, see Acito and Jain (1980), Akaah andKorgaonkar (1983), Andrews et al. (2002), Carmone et al. (1978), Hauser and Urban(1979), Hauser and Rao (2002), Huber et al. (1993), Louviere (1988), Neslin (1981),Jain et al. (1979), McFadden (1994), Orme (1999), and Segal (1982), Srinivasan(1988, 1998), Bateson et al. (1987), and Reibstein et al. (1988).

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 30: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

30 Choice-Based Conjoint (CBC) Analysis

products, personal computers, and fork-lift trucks are close to actualresults, and conclude that “These results provide strong support forthe validity of self-explicated conjoint models in predicting marketplacechoice behavior”. At a disaggregate level, Wittink and Montgomery(2009) use CBC data to estimate individual weights on eight attributesof jobs open to MBA graduates, and four months later observe actualjob choices and the actual attributes of jobs offered. They report 63%accuracy (percent hits) in predicting the jobs students chose out ofthose offered, compared to a 26% expected hit rate if the studentshad chosen randomly. They attribute the failure to achieve higheraccuracy to noise in estimates of weights for individuals, and to theinfluence of job attributes not included in the CBC study. They con-clude: “On balance, the published results of forecast accuracy are verysupportive of the value of conjoint results.” They do caution that “Oneshould keep in mind that positive results (conjoint analysis providingaccurate forecasts) are favored over negative results for publication.Nevertheless, the evidence suggests that marketplace forecasts havevalidity.”

2.2 A conjoint study example

To make CBC more concrete, consider the problem of estimating con-sumer preferences for table grapes, to guide growing and supermarketpricing decisions. In preliminary focus groups, it is found that in ad-dition to price per pound, the attributes that consumers mention asimportant are sweetness, crispness, grape size, and whether or notthe grapes are organically grown. Subjects interviewed at a mall aregiven eight menus, with each menu containing a no-buy alternativeand three alternative types of grapes pictured and described in wordsgiving their price per pound and attribute levels. The CBC experimen-tal design varies price and product attributes over various levels, andassigns these randomly to bunches in each menu, making sure that thebunches offered all differ from each other in two or more dimensions.The levels in the study are $1.00 to $4.00 for the Price of a one-poundbunch, Tart, Sweet for Sweetness, Soft, Crisp for Crispness, Small,Large for Size, and No, Yes for Organic. In a parametric model of

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 31: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

2.2. A conjoint study example 31

the utility of alternatives, these “raw” attributes may enter linearly,and also as transformations such as an interaction of the sweetnessand crispness attributes, (Sweetness = Tart)*(Crispness = Soft) oran interaction of a product attribute such as sweetness and subjectgender, (Gender = Female)*(Sweetness = Tart). Subjects are told thatat the end of the interview, one of the menus they have respondedto will be drawn, and they will receive their choice from this menuplus a Visa cash card for $10 less the price of their choice. This CBCdesign has strong incentive alignment, provided subjects believe thatthe promised transaction will be carried through. The study assumesthat consumers buy at most one bunch of grapes per day, and doesnot ask for purchase intensity. Subjects are told that checking the “nopurchase” option on a menu means that they would not buy any grapestoday if this were the only menu available to them in a food market orin this study, but that after the experiment they are free to purchasetable grapes or not as they wish. If subjects understand and accept thisframe, and do not respond strategically in the experiment even if theyanticipate better deals for table grapes outside, then the experimentshould predict daily market demand. Table 2.2 gives an example of amenu.

Table 2.3 illustrates data obtained from this CBC, with the attributesand levels coded as indicated in Table 2.2. “Bunch intercept” is a dummyvariable that is one except for the no-purchase alternative, and “Choice”is a choice indicator for each subject and menu. The data continueswith the remaining menus for the first subject, next the data for thesecond subject, and so on.

Table 2.2: Table grape CBC menu.

Menu 1 Bunch #1 Bunch #2 Bunch #3 No PurchasePrice/lb. $2.50 $2.75 $3.00 $0.00Sweetness Tart (0) Sweet (1) Sweet (1)Crispness Crisp (1) Soft (0) Crisp (1)Size (count per lb.) Small (0) Large (1) Small (0)Organic? No (0) No (0) Yes (1)Check One:

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 32: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

32 Choice-Based Conjoint (CBC) Analysis

Tab

le2.

3:Extract

ofda

tafrom

thetablegrap

eCBC.

Sweet*

Sweet*

Bun

chSu

bject

Menu

Alte

rnative

Pric

eSw

eetness

Cris

pness

Size

Organ

ic?

Cris

pGen

der

intercep

tGen

der

Cho

ice

11

12.98

00

01

00

10

01

12

1.51

01

01

00

10

01

13

1.42

10

00

00

10

11

14

00

00

00

00

00

12

11.12

10

01

00

10

11

22

1.63

11

11

10

10

01

23

2.77

10

00

00

10

01

24

00

00

00

00

00

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 33: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

2.2. A conjoint study example 33

This example and its incentive scheme show some of the problemsin designing a conjoint survey to closely mimic the real market, whileproviding the information needed to guide supplier decisions. First, thetechnical attributes of table grapes relevant to a supplier are measure-ments such as sugar content (in percent by volume), pH (unbufferedacidity on a scale of 0 for highly acid to 7 for distilled water), and berrycount per pound. Consumers rely on familiarity and experience to giveterms like “tart” and “soft” meaning, and not all subjects will have thesame interpretation. Hence, tying stated preferences among attributelevels such as “tart” and “soft” to specific technical attributes is a prob-lem, and it is difficult to translate stated preferences phrased in terms ofthese words into demand for products with specific technical attributes.Untrained subjects are unlikely to recognize technical attribute levelsand be able to reliably state preferences among them. However, it maybe possible, and important, to train subjects by having them tastegrapes with specific technical attribute levels and attach words to them;thus, “tart” may be associated with grapes of 9% sugar, and “sweet”with grapes of 12% sugar; “crisp” associated with 4 pH grapes and “soft”with 6 pH grapes, “small” with 40 count/pound and “large” with 20count/pound. This training could be part of a “practice” introductionto the experiment in which menus of currently available grape productswith associated verbal attribute descriptions are offered, and subjectsare asked to taste the different grapes and associate their preferences forgrapes with various verbal attribute profiles. This “vignette” approachallows a direct association of implicit technical attributes and tastes.The feasibility of this approach depends on the availability of existingproducts, the variability of attribute levels in these products, the extentto which potential future products will have technical attributes whosedesirability can be reasonably extrapolated from existing products, andthe ability of subjects to taste grapes through multiple menus withoutbecoming satiated or bored.

Second, the incentive alignment in the grape study, offering a chosenalternative from one menu a subject faces, will be feasible only ifone of the stated choices is an existing product. Otherwise, it maynot be possible to fulfill a promised transaction. Incentives can bealigned through more general promises to deliver an available product

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 34: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

34 Choice-Based Conjoint (CBC) Analysis

(or cash) consistent with the subject’s stated preferences, but this makesit more difficult for the subject to recognize that this makes it in theirinterest to be truthful. Third, whether subjects view the offered pricesas reasonable, or grapes as tempting, will depend on their shoppinghistories and grape inventories, their expectations regarding availabilityand prices of grapes elsewhere, particularly awareness of and responseto real market promotions and sales, and their anticipations of how theywill feel if they receive grapes. Overall, the predictive power of the “nopurchase” option for grapes will depend on its being given a sufficientlyspecific description and context, so that it corresponds realistically tothe product and purchase timing options the consumer will face in theforecast market.

2.3 Stated perceptions, expectations, and well-being

The focus of CBC analysis in the literature is on preferences amongproducts described in terms of their objective attributes, includingobjective probabilities of events when the nature or utility of a productis uncertain. In principle, recovery of preferences can incorporate subjec-tive probabilities, and perceptions through studies that include existingproducts with objective attributes, and products whose delivery iscontingent on verifiable future events. As in Ramsey (1931), Savage(1954), and Arrow (1971), if markets or experiments offer a rich familyof lotteries on different events, consumers have complete, transitive,continuous preferences over contingent goods and services, and satisfysome plausible axioms on the separation of subjective probabilities ofevents and preferences over consequences of events, then these con-sumers have consistent preferences and subjective probabilities thatshould be recoverable using stated preference methods. However, ithas been recognized since the work of Allais (1953), and emphasizedby the experiments of Kahneman and Tversky (1979) and Kahnemanand Tversky (2000), that consumers are hypersensitive to context whenasked to perform cognitive tasks requiring personal probabilities orperceptions. Consequently, the task of recovering consistent, stable per-sonal probabilities, and perceptions is at least as challenging as thetask of recovering preferences. At this point, there is no systematic

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 35: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

2.3. Stated perceptions, expectations, and well-being 35

approach to measuring perceptions that is comparable to CBC analysisof preferences, although one can in principle conduct CBC studies ofchoice among contingent events. Manski has found that stated prob-abilities on a 100-point rating scale can be relatively consistent andfairly predictive in some circumstances, although coding of responsesto focal points is an issue, see Manski (1999), Manski (2004), Blasset al. (2010), and Delavande and Manski (2015). Manski’s elicitationformat allows subject uncertainty in decision-making by asking for thechances of picking particular products rather than identifying a singlechoice. His argument is that if subjects consider subjective probabilitiesover completing partially profiled products, then their most realisticresponses are their personal probabilities that each alternative is chosen.He points out that subjective uncertainty here may include uncertaintyabout the subject’s state of mind at the time of actual future choice,and points out that these uncertainties are resolvable; i.e., they willbe known in an actual choice setting. Hudomiet et al. (2018) intro-duce a paired-menu elicitation format, asking for the probabilities ofa choice at two levels of a single attribute, with other attributes left(implicitly or explicitly) unchanged, that shows considerable promise forcontrolling the fill-in problem and identifying predictive causal effects. Aquestion for future research is whether elicitation of conditional choiceprobabilities, which no longer force the subject to resolve uncertain-ties in some manner and state a single choice, introduces cognitiveelements that are different from those operating for choice in actualmarkets. The answer will depend in part on the degree to which subjectsallow the elicitation format to influence their definition of the cogni-tive task they have been given. The Kahneman–Tversky experimentsindicate that errors in the arithmetic of personal probabilities, andsystematic biases, are likely to be intrinsic to subjective probabilityjudgments, and a problem for the future is to combine Manski-typeelicitations, a framework for incentive alignment, and experimentaldesigns like those of conjoint analysis, to map out structural modelsof beliefs, personal probabilities, and risk response that are predic-tive for market demand for new products with shrouded contingentattributes.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 36: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

3Choice Behavior

The neoclassical economic theory of consumer choice behavior is basedon the “consumer sovereignty” postulate that utility is predeterminedand stable, and consumer choice maximizes utility subject to a bud-get constraint determined by income and prices. Utility functions aredefined over bundles of personally consumed continuous and discretemarket goods. The utility of a particular bundle of goods is independentof the existence or availability of alternatives; thus, independent ofprices and income, and of attributes of unchosen bundles. Intertem-poral dependence is allowed, with current preferences influenced byconsumption history and expectations, but interpersonal influences(e.g., altruism, social network effects) are usually excluded.1 When con-sumers face uncertainty regarding events that affect product attributes,additional postulates for strict neoclassical consumer theory are thatconsumers have realistic, statistically consistent perceptions regarding

1Most economic studies of consumer behavior treat households as unitary decision-making units with a single household utility function. While there is an extensiveliterature on household and social group shared and individual tastes, production andconsumption, communication, bargaining, reciprocity, and decision-making, there isno widely accepted model for analyzing demand behavior and well-being of thesegroups.

36

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 37: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

37

future events, and that they handle uncertainty by maximizing expectedutility. We use the term “direct utility” for an index of desirability ofbundles of goods and services with specified attributes, and “indirectutility” for the level of maximum direct utility achievable with a budgetdetermined by income and the prices of goods and services. We alsouse the term “decision utility” for the index of desirability that neo-classical consumers anticipate they will achieve through their marketpurchases, and “experience utility” for the level of satisfaction theyachieve in retrospect. In general, experience utility cannot be recoveredfrom revealed or stated demand behavior, but there are importantexceptions where the gap between decision and experience utility isdue solely to removal of deception or resolution of uncertainty. Thetarget of most stated or revealed preference studies is “indirect decisionutility” and its implications for choice and well-being. Preference het-erogeneity leads to a distribution of market good demands and discretechoices among consumers facing the same incomes and prices. However,a single consumer’s choices from different menus offered at differenttimes or in different stages of a stated choice experiment will in theneoclassical theory be explained by maximization of a single decision-utility function. It is the structural stability of neoclassical decisionutility that makes it possible to forecast the effects of policy changeson the choices of each individual, and the distribution of choices in thepopulation.

Consumer choices from repeated menus in laboratory and marketexperiments often deviate from strict neoclassical theory, e.g., Mc-Fadden (1999).2 Consumers may have trouble discriminating amongsimilar alternatives, particularly if they have ambiguous or shroudedattributes. They may be poor statisticians, with inconsistent perceptions

2McFadden (2006) summarizes the testable implications on choice probabilitiesthat come from a population of preference maximizers. For example, the probabilityof a specified alternative being chosen cannot rise when a new alternative is addedto the menu, and the sum of the probabilities of choices in an intransitive cycle oflength K cannot exceed K−1. When these testable conditions are satisfied, behavioris consistent with a stable distribution of preferences over the population. Thesetests are silent on the extent to which there is heterogeneity in individual consumerpreferences across different choice occasions as opposed to heterogeneity in tastesacross consumers.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 38: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

38 Choice Behavior

and expectations about the attributes of alternatives, e.g., unrealisticexpectations about the probability of a product defect. They may not berelentless and exact preference maximizers, instead making choices thatare influenced by whimsy, fatigue, inattention, bias toward immediategratification, and a willingness to settle for “good enough”. Also, prefer-ences may encompass unmeasured attributes of alternatives, and maybe conditioned on previous choices, experience, and social motivationsthat are not fully accounted for. Finally, preferences may be situational,anchored or adapted to the status quo, and sensitive to context. There isa simple device for accommodating behavior alternatives to neoclassicaltheory. The population of consumers may represent a mixture of peo-ple who conform to the neoclassical model, others who adopt decisionalgorithms that provide economies in decision-making effort or protec-tion from severe mistakes, such as heuristics that may be broadly butnot universally consistent with utility maximization (e.g., Ariely et al.,2006), reductive processes such as filtering (e.g., Lleras et al., 2017) orelimination-by-aspects (Tversky, 1972), or regret-minimization (Manski,2007). In principle, the array of choice models for these consumer classescan be mixed, with latent shares of consumers in each class, to forman omnibus model of consumer behavior. There are technical problemsof identification of the parameters of such a model in the extreme thatsome consumer classes are empty. However, econometric methods fortesting non-nested hypotheses, which also use the device of mixing thenon-nested models, provide solutions, see Pesaran and Weeks (2001) andHong and Preston (2012). Further, in a Bayesian approach to analyzingCBC data, it is rather natural to consider mixtures of models, and withsuitable priors, identification is not an issue.

The presence of non-neoclassical behavioral elements in consumerbehavior is not necessarily a major impediment to forecasting newproduct demand, which requires only that the distribution of decision-rules in the population be recoverable from observed behavior and bestationary between the observation and forecast environments; decisionrules are not required to be realistic, or to map nicely into realizedwell-being. However, behavioral elements break the tight neoclassicallink between the utility and observed choices, complicating welfareanalysis.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 39: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

39

The early models of individual choice behavior that came out ofpsychophysics, Thurstone (1927), Luce (1959), and Marschak (1960),focused on stochastic elements in individual preferences originating fromimperfect psychophysical discrimination. McFadden’s (1974) adaptationof these models for econometric analysis of choice behavior in marketsrecognized both inter-consumer and intra-consumer sources of preferenceheterogeneity, but lumped these together since they are indistinguish-able when observing a single choice by each consumer. However, whenconsidering stated choice data from subjects confronted with a panel ofmenus, intra-consumer and inter-consumer preference heterogeneitiescan be identified, and the question is whether it is theoretically and sta-tistically useful to do so. For one-shot forecasting of market demand ata single point in time, one modeling strategy is to ignore the panel datastructure generated by multiple menu choices of multiple consumers,and treat each observation as the result of an independent draw from adistribution of preferences. This loses information and statistical effi-ciency, but is nevertheless a consistent model that in many applicationswill have good predictive power for market behavior. Another strategyis to retain the neoclassical assumption of stable preferences within con-sumers, with minor allowance for the effects of imperfect psychophysicaldiscrimination. This nearly neoclassical approach exploits the informa-tion contained in multiple choices from a neoclassical consumer, andis particularly valuable if a target of the analysis is recovery of infor-mation on “permanent” individual preferences, with drift and whimsytreated as nuisance factors. However, it will give a misleading picture ofindividual preferences if intra-consumer heterogeneity is in fact signifi-cant, and is more than is needed for one-shot forecasting. For dynamicdemand forecasting, where the behavior of preferences over time iscritical, a third strategy is to introduce a structural system with bothintra-consumer and inter-consumer stochastic elements in preferences,see Hess and Rose (2009), Hess and Train (2011), Yanez et al. (2011),Daly et al. (2012), and Rose et al. (2015). This structural approachcan provide valuable information about systematic drifts in tastes thatwill influence choice in temporal sequences of real markets, and offersa solution to the Heckman (1981) initial values problem of identifyingobserved and unobserved state dependence in panel data. All the issues

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 40: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

40 Choice Behavior

and opportunities that arise in time series analysis appear in the struc-tural approach: Is the stochastic taste process Markovian, ergodic, ordescribed by an ARMA process? In this monograph, we will set up theconsumer model using the structural approach, and give an application.However, for the most part we will concentrate on the nearly neoclassicalmodeling strategy, assuming that aside from errors due to the limitsof psychophysical discrimination, a consumer’s choices over multiplemenus maximize the same preferences. In our view, this is usually themost parsimonious modeling approach to forecasting, focusing on pre-dictive features of consumer behavior and avoiding the complexity ofmodeling features that are descriptive but not predictive. This approachalso covers the first strategy through the expedient of treating multiplechoices from one consumer as if they were single choices from multipleconsumers.

3.1 Utility

We set notation and review the elements of random utility models (RUM)and utility-maximizing choices that can be applied to either revealedor stated choice data, and can be used to forecast demand for new ormodified or repriced products. We follow the RUM setup of McFadden(1974) that is a common starting point for discrete choice analysis, withthe generalizations and regularities developed in McFadden (2018) andinnovations to facilitate analysis of CBC data from multiple menus.Table 3.1 summarizes notation, which in general also follows McFadden(2018).

Suppose a consumer faces m = 1, . . . ,M menus containing a setJm of mutually exclusive alternative products or services, includingin general a “no purchase” alternative, and makes one choice fromeach menu.3 Let zjm denote the “raw” attributes and pjm denote thereal unit price of product j in menu m. The attributes and price ofa “no purchase” alternatives will be normalized to zero unless notedotherwise.

3Assume that one alternative is the designated default if the consumer refusesto make a unique choice. The number of alternatives |Jm| can vary with m, but tosimplify notation we often suppress m and let J denote the number of alternatives.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 41: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

3.1. Utility 41

Table 3.1: Summary of notation.

W Wage incomeA Net Asset incomeT Net transfer income (subsidies minus taxes)I = W +A+ T Disposable incomes Other individual characteristics, past

experience, background pricesm = 1, . . . ,M Menus of alternatives offered to a consumerj ∈ Jm Alternatives offered in menu mzm Vector of attributes of menu m products,

components zjm for j ∈ Jmpm Vector of prices of menu m products,

components pjm for j ∈ Jmρm = ζ + ηm Tastes (an abstract vector, finite in

applications) at menu m, where ζ representsneoclassical tastes that are invariant acrossmenus, and ηm represents transitory tastesor whims that are i.i.d. across menus

ε Vector of i.i.d. standard Extreme Value Type 1(EV1) psychometric discrimination noise,with components εj for j ∈ Jm

σ(ρm) Positive inattentiveness coefficient that scalesdiscrimination noise

ujm = U(W,A, T−pjm, s, zjm|ρm) + σ(ρm)εj

Conditional decision utility of alternative j inmenu m

xjm = X(W,A, T − pjm, s, zjm) Vector of predetermined transformations ofobserved choice environments

β(ρm) Vector of taste coefficients, commensuratewith X, depends on ρm

vjm = vjm(ρm) ≡ V (W,A, T,s, zjm, pjm, ρm) ≡ X(W,A,T − pjm, s, zjm)β(ρm)− pjm

Quality-adjusted net value, the negative ofquality-adjusted price

F (ζ|s) and H(ηm|ζ, s) CDF of ζ and conditional CDF of i.i.d. ηmgiven ζ

F (ρ1, . . . , ρM |s) =∫ζ

F (dζ|s)∏M

j=1 H(ρm−ζ|ζ,s)CDF of the taste vectors

n ≡ (W,A, T, s) All observed socioeconomic characteristics,experience

ujm = I + vjm(ρm) + σ(ρm)εjm≡ I +X(W,A, T −pjm, s, zjm)β(ρm)−pjm+σ(ρm)εj

Linear additive money-metric utility

ujm = I + vjm(ζ) + σ(ζ)εj Linear additive money-metric utility withneoclassical tastes; i.e., ηm ≡ 0.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 42: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

42 Choice Behavior

The consumer has a latent taste vector ρm (abstract in general,finite-dimensional in practice) when confronted with a menu m that isa function of neoclassical “permanent” tastes ζ that are fixed acrosschoice situations and non-neoclassical “transitory” whims ηm that varyacross menus. The tastes ρm will in general vary across consumers.The consumer is also characterized by a vector of socioeconomic char-acteristics s that include household demographics, location, and pastconsumption experience, and by real disposable income I, written asI ≡ W + A + T , where the components are wage and salary incomeW , net asset income A, and net lump-sum transfer income T , thelast typically negative due to taxes. The reason for the income de-composition is that in a population, choice probabilities can vary withincome for three reasons: First, a consumer with given preferences mayshift choices and consumption patterns as disposable income rises, aneoclassical income effect. Second, changes in wage or asset incomehave links to unobserved effective prices of discrete alternatives or at-tractiveness of features, e.g., through (unobserved) opportunity costof leisure or the cost of financing durable purchases. Consequently,observed response of choice to changes in these income componentsconfounds neoclassical income effects and unobserved price effects onchoice. Third, observed income may be ecologically correlated withtastes across the population due to common causes, e.g., consumerswith high (unobserved) rates of impatience may have lower incomesbecause of unwillingness to delay consumption to invest in human capi-tal, and also have higher probabilities of choosing less durable products.This can induce a correlation of income and demand across the popula-tion even if individual demands exhibit no neoclassical income effect.Then, to the extent that the measured impact of income changes onchoice is different for T than it is for W , A, or I, the response to T ismore likely to reflect a pure neoclassical income effect. The reasons thismatters are that neoclassical welfare analysis is greatly simplified if neo-classical income effects are sufficiently small to be neglected, and thateither hypothetical or fulfilled compensation are typically adjustmentsto transfer income, so that compensating values are appropriately calcu-lated with required income adjustments in terms of the transfer incomecomponent.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 43: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

3.1. Utility 43

We assume that consumer tastes ρm are represented for each dis-crete alternative j in menu m by a conditional indirect utility functionU(W,A, T − pjm, s, zjm|ρm) + σ(ρm)εj , obtained by optimization overthe remaining market goods subject to the disposable income I − pjmleft after buying j. In this function, εj is psychometric noise, scaledby σ(ρm) > 0, that comes from imperfect attention, diligence, andacuity.4,5 Given the previous discussion, we allow the W and A com-ponents of income to enter this indirect utility as separate argumentsto accommodate their possible effects on utility other than throughneoclassical income effects. We assume that consumers make discretechoices j from the alternatives available to maximize conditional indirectutility. In neoclassical consumer theory, utility functions are ordinal,and interpersonal comparisons of utility differences are meaningless.However, in a population of heterogeneous consumers, it is importantto represent preferences so that utilities vary regularly with ρm and aredenominated in a metric that permits interpersonal economic compar-isons. We choose to scale utility in money-metric or WTP-space units,with the implication that the marginal utility of a dollar is the samefor every consumer. This regularizes the modeling of choice behaviorand facilitates applications where compensating transfers are calcu-lated for benefit–cost or welfare analysis (with social weights appliedwhen deemed appropriate). The term “money-metric” utility is dueto Samuelson (1950), who observed that indirect neoclassical utilityevaluated at given prices could with an increasing transformation beexpressed in units of income. Hurwicz and Uzawa (1971) relate thisform of utility to measurement of consumer surplus, see also McFadden

4The price pjm will be treated in the following analysis as the purchase priceof product j, but it is also possible to interpret pjm as an “access fee”, with theconditional indirect utility U also depending on “user fees” or other nonlinear pricingstructures, as well as on the underlying prices of goods available in continuousquantities. It is also possible to interpret pjm as a market or implicit rental cost of aconsumer durable, or as the cost of auxiliary activities required to consume productj, e.g., the cost of travel to a site where the product is located. We omit thesebackground and nonlinear prices to simplify notation, but they can be reintroducedwhen interactions of discrete and continuous choices or pricing structures are a focusof study.

5We assume the psychometric noise εj for the product indexed by j is the samein every menu.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 44: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

44 Choice Behavior

(2014a, 2014b, 2018). We standardize money-metric utility using totaldisposable income, but in our suggested specification with wage andasset income included as separate factors to capture unobserved pricesof leisure and capital, this is equivalent to standardization using transferincome. Train and Weeks (2005) use the term “WTP-space” utilityto refer to money-metric utility, as the incremental utility associatedwith a change in an attribute equals the consumer’s WTP for thischange. They recommend this scaling for stabilizing empirical measuresof consumer welfare in discrete choice problems, see also Miller et al.(2011), Daly et al. (2012), and Carson and Czajkowski (2013). Trainand Weeks use the terminology “preference-space” utility when it iswritten with an additive stochastic with a coefficient of one, a normal-ization introduced by McFadden (1974) and employed in most previousapplications. Train and Weeks point out that the two normalizationsare formally equivalent in the sense that any distribution of coefficientsin preference-space translates into a distribution of WTP coefficientsin WTP-space, and vice versa. However, commonly assumed distribu-tions are often inconsistent. For example, a model in preference-spacethat assumes a non-degenerate log-normal price coefficient and normalnon-price coefficients is not equivalent to a model in WTP-space withan assumption of normally distributed WTP coefficients. Instead, theequivalent distribution in WTP-space is of the ratio of a normal toa log-normal, which relative to a conventional normal is skewed inthe direction of the mean of the normal with a heavier tail, see Yang(2008). While a normal log-normal specification could be assumed forWTP-space analysis, we have not seen this in applications. If some con-sumers filter alternatives based on the presence of a discrete attributebefore completing utility maximization, their behavior is consistentwith a high WTP for this attribute. Then, specification of a normallog-normal distribution of WTP coefficients may capture a real featureof behavior. Alternatives to money-metric scaling are to express utilityin units of leisure time (“minute-metric” utility) or a market basket ofgoods (“basket-metric” utility). The last alternative could be used forexample to scale utility in multiples of a designated poverty level. Again,these scalings are all equivalent with appropriate (non-conventional)assumptions on the distributions of coefficients. Without equivalent

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 45: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

3.1. Utility 45

distributional specifications, the various scaling alternatives give non-nested models of utility that could be discriminated using non-nestedtests, or plausibly combined by mixing in an omnibus model of thedistribution of WTP coefficients.

Our money-metric or Willingness-To-Pay-space specification of con-ditional indirect utility forces U(W,A, T, s, j0|ρm) ≡ I for a designated“no purchase” alternative j0, and adopts an additive linear approxima-tion to this utility,

ujm = U(W,A, T − pjm, s, zjm|ρm) + σ(ρm)εj

≡ I − pjm +X(W,A, T − pjm, s, zjm)β(ρm) + σ(ρm)εj . (3.1)

Then Equation (3.1) is denominated in real monetary units with theproperty that uj0m = I + σ(ρm)εj0m. The function xjm = X(W,A, T −pjm, s, zjm) is a vector of predetermined transformations (e.g., logs,polynomials), weighted by taste coefficients β = β(ρm), that caninclude nonlinear transformations and interactions of raw productattributes, interactions of raw product attributes and consumer charac-teristics, and dummy variables that identify types or classes of products(e.g., indicators for “brand”), with X(W,A, T, s, j0) ≡ 0 by normal-ization in most applications. Tastes ρm are in general heterogeneousacross consumers, and may be heterogeneous across menus, and willbe treated in our most general analysis as random effects with a CDFF (ρ1, . . . , ρM |s).6 By construction, the coefficients σ(ρm) and β(ρm)will be predetermined transformations (e.g., linear, exponential) of theunderlying taste vector ρm. The consumer’s Willingness-To-Pay (WTP)vector for unit increases in corresponding components of X is givenby dpjm

dxjm|ujm=const. = β(ρm)/1 + [∂X(W,A, T − pjm, s, zjm)/∂T ]β(ρm).

In marketing, models are often specified without imposing the money-metric scaling in Equation (3.1), and the coefficients β(ρm) are termed

6The socioeconomic effects s may enter as arguments in X, as conditioningvectors in the distribution of taste parameters, or both. The former is more naturalif socioeconomic characteristics change the perceived quality of attributes and levels,while the latter is more natural if socioeconomic factors shift the location of theconsumer in the field of preferences. If choices influence tastes through learning,then F is structurally stable, but if the conditioning comes from selection on seriallycorrelated tastes, one has the challenging initial values problem of Heckman (1981).

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 46: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

46 Choice Behavior

part-worths. If the functions X do not depend on net transfer incomeT − pjm, then Equation (3.1) is of Gorman Polar Form, Engle curvesfor continuous goods are linear and parallel for all consumers, discretechoices are independent of T , showing no neoclassical income effect, andWTP simply equals the dollar-denominated vector β(ρm), see Gorman(1953), Gorman (1961), Chipman and Moore (1980), Chipman andMoore (1990), McFadden (1981), McFadden (2004), and McFadden(2014b). Welfare analysis of discrete choice simplifies considerably whena neoclassical income effect is absent or sufficiently minor to be ignored.The β(ρm) coefficients will play a central role in business or public policyanalysis of the demand and consumer welfare impacts of modifications inthe offerings of products. However, simple measures of central tendencysuch as population mean or median β’s do not account for choice selec-tion effects or neoclassical income effects, and a more careful analysisis required in many cases to translate the distributions of β coefficientsinto demand and welfare impacts, see McFadden and Train (2018).

Examples of X(W,A, T −pjm, s, zjm) functions from a study of auto-mobile choice are horsepower divided by vehicle weight and the numberof seats in the vehicle less the number of persons in the consumer’shousehold. Another example is the fuel efficiency of an automobilealternative, with a coefficient interpreted as the value of a unit increasein fuel efficiency, reflecting both tastes and subjective expectations onthe future price of fuel. In travel mode choice, W and A may influenceutility even with no neoclassical income effect —W interacts with tastesfor time-saving attributes and A interacts with credit requirements forowning a car, but if there is no neoclassical income effect, X does notdepend on T . X can also include dummy variables for “brand name” orfor commodity classes (e.g., “buses” or “public transit” in a study ofmode choice), with the associated components of β(ρm) capturing thecommon impact of these factors on the utilities of various choices. Oneimplication of X transformations that interact with consumer charac-teristics is that a product modification that changes a raw attribute zcan have an impact on multiple components of X, and these impactscan vary across consumers.

The requirements of first-stage utility maximization are that theindirect utility Equation (3.1) be homogeneous of degree zero and quasi-

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 47: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

3.1. Utility 47

convex in nominal income and prices, including the market prices of thegoods demanded in the first-stage optimization, increasing in income,and non-increasing in all prices, see McFadden (1974), Deaton andMuellbauer (1980), Fosgerau et al. (2013), McFadden (2014b), andMcFadden (2018). The homogeneity requirement is met by writingall income components and prices in real rather than nominal terms,but the quasi-convexity requirement is a substantive restriction onthe function forms of the X transformations. Quasi-convexity can beforced on Equation (3.1) by requiring that components of X for whichthe corresponding component of β(ρm) is unrestricted in sign (resp.,restricted positive, restricted negative) be linear (resp., convex, concave)in background prices (and in T − pjm if it is an argument in X).However, these restrictions are stronger than necessary, and it will bemore practical to not impose them a priori, and instead test estimatedutility functions for the required quasi-convexity on the empiricallyrelevant domain.

An issue that is important for predicting demand for new productsis whether X depends only on “generic” attributes of products, or alsoon “branded” attributes. Here, “generic” means that consumers respondto the attribute in a new product in the same way as in an existingproducts, e.g., “schedule delay time” (the time interval between whenthe consumer would like to leave and when the mode is scheduled toleave) in a study of transportation mode choice has the same definitionfor both existing and new modes. Class indicators in xjm may be eithergeneric (e.g., “public transit”) or branded (e.g., “Red Bus Company”).A “branded” attribute interacts with a “brand” class dummy, so thatthe associated WTP can depend on the “brand” of the product, e.g.,WTPs may differ for “Red Bus” and “Blue Bus” schedule delay time.While branded attributes may provide a parsimonious “explanation” oftastes, they are shorthand for deeper generic attributes, such as thecomfort of the waiting environment in the mode choice example. Whenbrand-specific effects are not easily quantified in measures of perceivedreputation, these shorthand variables may be convenient, or expedientwhen it is clear how to extend branded attributes to new products, e.g.,when new products are introduced as new lines in an existing brand.However, the analyst should recognize that recasting “brand” in terms of

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 48: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

48 Choice Behavior

generic attributes such as reputation for reliability or durability extendsthe ability of analysis to forecast demand for innovative products.

Consider the roles of the latent taste parameter vector ρm and thedisturbances εj in Equation (3.1). The εj are interpreted as pertur-bations in utility that vary from one consumer and alternative to thenext, and come from the psychometric difficulty consumers have in dis-tinguishing between products and attribute levels, from whimsy, fromlack of consumer attention and acuity, and from mental mechanismsconsumers use to “break ties” when the utility levels of alternatives seemindistinguishable. The ρm locates the consumer in the field of possiblepreference orders. They reflect “permanent” tastes, and for a strictlyneoclassical consumer are the same for each menu faced. However, moregenerally they may vary across menus due to non-neoclassical decision el-ements. If the preference field can be represented as a linear vector space,the combination of inter-consumer and intra-consumer heterogeneitycan be modeled as ρm = ζ+ηm, where ζ characterizes permanent tasteswith CDF F (ζ|s), and the ηm, independently identically distributedwith CDF H(ηm|ζ, s), capture perturbations in tastes from one menu tothe next.7 The joint CDF of the taste parameters in Equation (3.1) isthen F (ρ1, . . . , ρM |s) =

∫+∞−∞ H(ρ1−ζ|ζ, s) . . . H(ρM−ζ|ζ, s)F (dζ|s). A

strictly neoclassical consumer with sharp psychophysical discriminationcorresponds to H(ηm|ζ, s) having unit probability at ηm = 0 and atiny positive scaling factor σ(ζ) in (3.1) so that the εjm just provide atie-breaking mechanism. This setup will also describe a consumer withsome non-neoclassical elements in decision-making, such as unrealisticperceptions or dependence of tastes on social motives that lie outside theneoclassical assumptions, so long as this consumer behaves consistentlyacross menus “as if” she had fixed tastes. This is the case favorable forforecasting future market demand on which we will concentrate, withobservations from each consumer for multiple menus consistent with a

7The assumption that the ηm are i.i.d. could be generalized to allow a stochasticprocess governed by the rate of decay of memory, and perhaps influenced by externalevents. However, identification and estimation of the features of a more generalprocess is a challenging task in time series analysis. Later, both F and H will bemade functions of a “deep” parameter vector θ, and ζ and the ηm will be termed“intermediate” parameters. In a Bayesian interpretation, these parameters then havea hierarchical structure.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 49: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

3.1. Utility 49

single decision-utility function that has the stability property that itwill also be the decision-utility function in the future market. However,it is also possible to use the utility setup (3.1) to model joint inter-consumer and intra-consumer heterogeneity in decision utility, withcoefficients written as σ(ζ + ηm) and β(ζ + ηm). One limiting case thatmay be expeditious for forecasting applications is “no taste memory”with F (ζ|s) having unit probability at ζ = 0, so that every consumerfor every menu draws tastes from the same distribution without regardfor previous taste history. It would have been possible to write themodel (3.1) more compactly by absorbing the terms σ(ρm)εj into thevector ρm; if these effects are needed to rationalize choice, they wouldthen appear as components of β(ρm) associated with components of Xthat are dummy variables for the individual alternatives j = 0, . . . , Jm.The justification for introducing the “redundant” σ(ρm)εj terms sepa-rately is that isolating these as independent additive terms is useful forobtaining relatively simple expressions for choice probabilities and forcomputation. While the perturbations εj may be needed to reconcileobserved choices with utility maximization, they are not an explanationof choice behavior. If their presence looms large, this says that many ofthe determinants of choice are unobserved and unaccounted for, andthis will limit the accuracy of predictions. Ideally the σ(ρm)εj termshave a modest effect on utility differences relative to the “systematic”component vjm ≡ X(W,A, T − pjm, s, zjm)β(ρm) − pjm that containsthe relevant information on the consumer’s choice environment. Whenthe distribution of the taste coefficients β(ρm) across consumers andmenus is sufficiently broad, and X contains a sufficiently rich arrayof dummy variables whose coefficients capture idiosyncratic tastes forvarious product classes, the disturbances εj are not really needed to“explain” correlations between utilities of different alternatives and id-iosyncratic choices. Then one can treat the εj as elements that simplysmooth responses and facilitate computation. In this case, it should beharmless to assume that the εj are independently, identically distributedwith a convenient Extreme Value type 1 (EV1) distribution, and toassume that the scaling coefficient σ = σ(ρm) is constant. However,if choice is poorly predicted in practice by the systematic componentof Equation (3.1), specification of the distribution of εj can have a

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 50: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

50 Choice Behavior

significant impact on predicted choice probabilities, and it will becomeimportant to avoid restricted specifications such as i.i.d. EV1 that resultin unrealistic choice probabilities.

The additive linear specification of the systematic component inEquation (3.1) seems quite restrictive, but McFadden and Train (2000,Theorem 1) and McFadden (2018, Theorem 5.1) prove that Equation(3.1) with sufficiently flexible specifications of the X transformationsand the distribution F (ρ1, . . . , ρM |s) can mimic any preference fieldand choice situation that meets mild regularity conditions.8 Of course,for Equation (3.1) to be a good model of decision utility, the analystmust be prepared to introduce complex, high-order terms in X andto allow flexible characterizations of the distribution of preferences. Itwill be an empirical question as to whether any specific truncated basisfunction series approximation to random utility is sufficient to achievedesired degrees of accuracy and avoid forecasting bias due to modelmisspecification. In McFadden (2018), the scaling factor σ on the EV1noise can be taken to be any sufficiently small positive constant. Aconstant σ assumption is particularly convenient for consumer welfarecalculations. However, in applications, it may be parsimonious, or usefulfor interpretation in terms of heterogeneity in consumer acuity, to allowσ(ρm) to vary across consumers. For the sake of generality, we will allowheterogeneous σ(ρm).

The specification (3.1) of utility is predicated on the structuralassumption that consumers are decision-utility maximizers, with ad-ditional restrictions if consumer behavior is assumed to be almost orstrictly neoclassical. Without further restrictions, the utility is a “par-tial reduced form” that convolves the effects of true tastes, subjectiveperceptions of product attributes, and subjective probabilities of con-

8The assumptions needed for the result are (1) the maximum numbers of discretealternatives and menus are finite, J ≤ Jmax < +∞ and M < +∞; (2) the domainsof (W,A, T, p1, . . . , pM , s, z1, . . . , zM ) and (ρ1, . . . , ρM ) are finite-dimensional, closed,and bounded; (3) the probability measure on the domain of (ρ1, . . . , ρM ) is absolutelycontinuous; (4) the preference field has Lipschitz-continuity properties; and (5) thereis zero probability of ρm such that U(W,A, T − pjm, s, zjm|ρm) = U(W,A, T −pkm, s, zkm|ρm) for j 6= k. Sufficient conditions for the absolute continuity and zeroprobability of ties conditions are that all alternatives in menus are distinct in someattribute and that tastes for this attribute are continuously distributed.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 51: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

3.2. Choice probabilities 51

tingent events (e.g., “Will these Bluetooth headphones work reliably,and will I enjoy using them if I buy them?”). When it is important forpolicy analysis to identify true tastes, perceptions of attributes thatmay respond to advertising or information campaigns, and subjectiveprobabilities that may be altered by product warranties and guarantees,a combination of CBC experiments focused on subjective perceptionsand probabilities and structural assumptions on Equation (3.1) mayallow these components of decision-utility to be isolated. Developmentof a consistent system of hypotheses on choice behavior, structuralassumptions on Equation (3.1), and CBC experimental designs to ac-complish these tasks is beyond the scope of the current monograph, butis an important topic for future research.

3.2 Choice probabilities

Consumers with decision utility functions of the form (3.1) choose alter-natives j = 0, 1, . . . , J in menu m = 1, . . . ,M that give maximum utility.Abbreviate notation by defining vjm = vjm(ρm) ≡ V (W,A, T, s, zjm,pjm, ρm) ≡ X(W,A, T − pjm, s, zjm)β(ρm) − pjm, then ujm = I +vjm + σ(ρm)εj . We will call vjm “quality-adjusted net value” or thenegative of quality-adjusted price. Define zm = (z1m, . . . , zJm), pm =(p1m, . . . , pJm), and vm = (v0m, . . . , vJm), with v0m = 0 by construction.The next step in discrete choice analysis is to go from (3.1) to a choiceprobability model consistent with its maximization. A useful tool forthis analysis is the expected maximum money metric utility,

G(I,vm) ≡ Eρm|sEε maxj=0,...,J

I + vjm + σ(ρm)εj

= I + Eρm|sσ(ρm) ·

logJ∑j=0

exp(vjm(ρm)σ(ρm)

)) + γ0

. (3.2)

The first form of this expression holds for any distribution of the psy-chometric noise, while the second form holds, with γ0 denoting Euler’sconstant, when the εjm are i.i.d. EV1 distributed, see McFadden (2018,Appendix B). The argument I in G denotes direct dependence throughthe linear term in Equation (3.2); there will also be indirect dependenceon W or A if they influence the effective prices of alternatives, and on

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 52: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

52 Choice Behavior

T through the function vjm = V (W,A, T, s, zjm, pjm, ρm) when there isa neoclassical income effect. The derivative of G with respect to vjm,obtained by differentiating inside the expectations and the maximiza-tion operator,9 gives the choice probability Pj(W,A, T, s,xm,pm) foralternative j:

∂G(I,vm)/∂vkm ≡ Pj(W,A, T, s,xm,pm)

≡ Eρm|sEε1(vjm + σ(ρm)εj > vkm

+σ(ρm)εk for k 6= j). (3.3)

Due to this property, G is called a choice probability generating function(CPGF). Fosgerau et al. (2013) show that every well-behaved randomutility model has a CPGF, and give necessary and sufficient conditionsfor a function to be a CPGF. Their approach embeds a given randomutility model in a random preference field with additive perturbationsof the utilities of various alternatives, and shows that the gradient ofexpected maximum utility with respect to these perturbations givesthe choice probabilities. When there is no neoclassical income effect,G coincides with the social surplus function utilized by McFadden(1981) to characterize choice probabilities when the effect of income onindirect utility is additive, the derivative ∂G(I,vm)/∂vkm is identicalto −∂G(I,vm)/∂pkm, the CPGF is the utility function of a “represen-tative” consumer, and the CPGF gradient property is equivalent toapplication of Roy’s identity to obtain demand shares that equal thechoice probabilities. The CPGF approach is useful because it providesa tight link from expected maximum utility to choice probabilities thatholds whether or not there is a neoclassical income effect. This linkcan be used to generate systems of choice probabilities from functionsthat meet the conditions to be a CPGF, and to test systems of choiceprobabilities with random utility maximization.

In the case of i.i.d. EV1 psychometric noise, the choice probabilitiesgiven by the gradient of the last form in Equation (3.2) are of mixed

9We use vjm to denote quality-adjusted value, a scalar variable, and also asshorthand notation for the function vjm(ρm) = Vjm(W,A, T, s, zjm, pjm, ρm). Thederivative ∂G(I,vm)/∂vjm is taken with respect to the scalar variable. A reader whofinds the dual use of the vjm notation confusing should add an explicit perturbationwjm to vkm, differentiate G with respect to w, and evaluate the result at w = 0.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 53: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

3.2. Choice probabilities 53

(or, random parameters) multinomial logit form,

Pj(W,A, T, s,xm,pm) ≡ Eρm|sPj(W,A, T, s,xm,pm, ρm), (3.4)

where

Pj(W,A, T, s,xm,pm, ρm)

=exp

(xjmβ(ρm)−pjm

σ(ρm)

)∑Jk=0 exp(xkmβ(ρm)−pkm

σ(ρm) )for j = 0, 1, . . . , J (3.5)

is a multinomial logit (MNL) model. A cross-price elasticity from Equa-tion (3.5), ∂ logPj/∂ log pkm = −(δjk−Pk)pkm/σ(ρm), where δjk is oneif j = k, zero otherwise, will be useful in later analysis.

A strong and potentially seriously limiting property of MNL modelsis Independence of Irrelevant Alternatives (IIA): The odds that onealternative is chosen over a second are unchanged when a third alter-native is added to the menu. When the third alternative shares manyfeatures with the second alternative, but not the first (e.g., the first twoalternatives are “car” and “red bus”, and the new alternative is “bluebus”), it may primarily split bus choices rather than drawing from thecar choice, and the IIA property is unrealistic. To avoid poor predictionsdue to IIA, Equation (3.4) must be specified with flexible xjm, andlatent tastes ρm must be allowed to be heterogeneous across consumers,as in Ben-Akiva and Bolduc (1996). With this flexibility, latent tasteheterogeneity overwhelms the IIA property and the probability (3.4) canapproximate the choice probability from any random utility model. How-ever, often in applications the taste parameters (σ(ρm), β(ρm)) = (σ, β)are assumed to be the same for all consumers and menus, and Equa-tion (3.4) reduces to a flat MNL model where IIA is an importantissue,

Pj(W,A, T, s,xm,pm, σ, β)

=exp

(xjmβ−pjm

σ

)∑Jk=0 exp

(xkmβ−pkm

σ

) for j = 0, 1, . . . , J. (3.6)

This model is often successful in forecasting despite its restrictive IIAfeature, but if IIA is rejected empirically, better modeling options are to

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 54: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

54 Choice Behavior

alter the assumed distribution of the disturbances εj , say by adopting aMultivariate Extreme Value distribution for these disturbances, leadingin applications to various nested logit models, or more generally, toallow random parameters preference heterogeneity as in Equation (3.4).Current estimation software makes the model (3.4) practical.

When a population of consumers is considered, inter-consumer het-erogeneities in tastes will appear in Equations (3.1) and (3.4) as het-erogeneities in σ(ρm) and β(ρm). Possible modeling strategies whenconfronted with choice data generated by Equation (3.4) are (A) to im-pose homogeneous (σ, β), obtaining the flat MNL model (3.6) in whichany inter-consumer heterogeneities that are present are “subsumed” inthe εj ; (B) to assume ρm = ζ is a separate “fixed effects” parametervector for each consumer, with no intra-consumer homogeneity, so thatηm = 0; or (C) to specify the ρm = ζ + ηm as “random effects” withCDF’s F (ζ|s, θ) and H(ηm|ζ, s, θ) depending on deep parameters θ.In strategies (B) and (C), unobserved variations across consumers intastes for product classes (e.g., “SUV’s” in automobile choice, “buses” incommuter mode choice) can be introduced as additive fixed or randomeffects embodied in the coefficients of indicators in xjm. In (C), theintermediate parameter vectors ζ and ηm that characterize preferencescan be interpreted as a hierarchical mixture, given θ. In working withchoice data where there is a single or limited number of observed choicesfor each consumer, strategy (A) is practical and often surprisingly goodat predicting demand despite IIA rigidity and the failure to capturetaste heterogeneity. The reason is that this model will often successfullycapture predictive central tendencies in tastes even when it does a poorjob of representing dispersion. The alternative (B) is unworkable, withno identification of individual consumer taste parameters ηm and pooridentification of taste parameters ζ in choice data from a limited numberof menus. In (C) the mixture distributions F (ζ|s, θ) and H(ηm|ζ, s, θ)will be identified under fairly mild conditions on the distributions ofattributes and the structure of utility. Alternative (C) is practical andsufficient for many purposes, particularly in the nearly neoclassical casewhere intra-consumer heterogeneity is absent; i.e., the H have unitprobability at ηm = 0 and ζ characterizes tastes. Train (2009, Chs. 11and 12) discusses estimation of individual consumer parameters under

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 55: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

3.2. Choice probabilities 55

classical estimation and under Bayesian estimation. Huber and Train(2001) describe and compare both procedures, see also Hensher andGreene (2003).

Let djm denote an indicator that is one if the consumer choosesalternative j from menu m, zero otherwise, dm = (d0m, . . . , dJm), andd ≡ (d1, . . . ,dM ) denote a portfolio of choices for this consumer. Definex = (x1, . . . ,xM ) and p = (p1, . . . ,pM ). The probability of a portfoliod implied by the model (3.4) under the random-parameters modelingstrategy (C) is

P (d|W,A, T, s,x,p)

≡∫ζ

M∏m=1

J∏j=0

[∫ηmPj(W,A, T, s,xm,pm, ζ + ηm)H(dηm|ζ, s)

]djm·F (dζ|s)

≡ Eζ|s

M∏m=1

J∏j=0

[Eηm|ζ,sPj(W,A, T, s,xm,pm, ζ + ηm)]djm

≡ Eζ|s

M∏m=1

J∏j=0

Eηm|ζ,s exp(xjmβ(ζ+ηm)−pjm

σ(ζ+ηm)

)∑Jk=0 exp(xkmβ(ζ+ηm)−pkm

σ(ζ+ηm) )

djm . (3.7)

In the nearly neoclassical case with the ηm ≡ 0, Equation (3.7) reducesto

P (d|W,A, T, s,x,p)

≡ Eζ|s

M∏m=1

J∏j=0

exp(xjmβ(ζ)−pjm

σ(ζ)

)∑Jk=0 exp

(xkmβ(ζ)−pkm

σ(ζ)

)djm , (3.8)

and when there is a single menu m so that the expectation can bemoved inside, reduces further to

P (dm|W,A, T, s,xm,pm)

≡J∏j=0

Eζ|sexp

(xjmβ(ζ)−pjm

σ(ζ)

)∑Jk=0 exp

(xkmβ(ζ)−pkm

σ(ζ)

)djm . (3.9)

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 56: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

56 Choice Behavior

Then Equations (3.7) and (3.8) are mixed MNL models for portfoliosof choices. The nearly neoclassical model (3.8), with its generalization(3.7) to allow non-neoclassical intra-consumer heterogeneity, or its spe-cialization (3.6) to flat MNL, will be the models we will estimate usingCBC data, and then use in Equation (3.9) to forecast demand for newor altered products. A consequence of Equation (3.8) that is useful forsome applications is that the probability that nearly neoclassical tastesζ are contained in any measurable set C, given a vector of choices d, is

F (C|s,d) =

∫ζ∈C

∏Mm=1

∏Jj=0

exp(xjmβ(ζ)−pjm

σ(ζ)

)1+∑J

k=1 exp(xkmβ(ζ)−pkmσ(ζ) )

djm F (dζ|s)

∫ζ

∏Mm=1

∏Jj=0

exp(xjmβ(ζ)−pjm

σ(ζ)

)1+∑J

k=1 exp(xkmβ(ζ)−pkm

σ(ζ)

)djm F (dζ|s)

.

(3.10)

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 57: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

4Choice Model Estimation and Forecasting with

CBC Data

The CBC elicitation format produces data on choices from hypotheticalmarket experiments which must then be analyzed to model preferencesand forecast demand. A natural approach to these tasks is to treatthe choice data from a CBC study like observed market choice data,with allowance for the portfolios of choice observations that come fromresponses to multiple menus. Such data may circumvent two limitationsof real market observations — lack of independent variation in someattribute dimensions and levels and in prices, due to the fact that inequilibrium markets offer only products that are sufficiently attractiveto consumers to be viable, and endogeneity of product features andprices as a result of market equilibrium.1 The key questions are whetherexperiments with hypothetical markets and incentive structures can bedesigned to give context, incentives, and information similar enough toreal markets to elicit the same perceptions, expectations, and cognitiveprocesses, so that stated choice data from these hypothetical markets

1When the disturbances that determine an individual consumer’s choice havecomponents that do not “net out” at the market level, equilibrium market levelprices will depend on these components, and cannot be treated as predetermined inestimating the choice model parameters, see Berry et al. (1995), Berry et al. (1998),and Berry et al. (2004).

57

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 58: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

58 Choice Model Estimation and Forecasting with CBC Data

are predictive for choice behavior in real markets, and whether tests ofthe predictive validity of forecasts based on stated choice data establisha domain in which they are reliably accurate.

When the data generation process for subject n is described bythe nearly neoclassical model (3.8) written as a non-linear function ofdeep parameters θ, a traditional econometric approach to estimatingθ is maximum likelihood estimation. Evaluation of Equation (3.8) re-quires a relatively high-dimensional integration that cannot in generalbe performed analytically, requiring either numerical integration orsimulation methods such as maximum simulated likelihood (MSL) ormethod of simulated scores (MSS). An alternative (and asymptoticallyequivalent) approach is hierarchical Bayesian estimation, which avoidsthe iterations necessary to maximize likelihood, but requires computa-tionally intensive (and usually iterative) sampling from the Bayesianposterior. We discuss and compare these approaches with estimation inSections 5–7. For forecasting, it is necessary to consider a sample fromthe population of interest, which may be the sample of CBC subjectsweighted to match known population characteristics (e.g., size, agedistribution, income distribution), or another real or simulated popu-lation sample. For each person in this forecasting sample, the choiceprobabilities (3.9) are simulated for values of the explanatory variablesthat will prevail in the forecast marketplace, including the attributelevels and prices of the products offered in this marketplace. Generally,the required simulations are similar to those used in the estimationstage.

A first step in an analysis program is to make the model (3.8) moreconcrete by specifying the predetermined transformations X of rawattributes and consumer characteristics, the predetermined transforma-tions σ(ζ) and β(ζ) of the intermediate parameter ζ, and the mixingdistribution F (ζ|s, θ) and the socioeconomic characteristics s and deepparameter vector θ on which it depends. Often, these specifications willbe application-specific, selected to be parsimonious in parameters andfocused on attributes of primary interest. They can be modified throughspecification testing during estimation. While exact model specificationstrategies may be application-specific, a reasonable general approach isto start with simple models including raw attributes and interactions

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 59: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

59

that are likely to matter, and test for the presence of significant omittedterms.2

A typical (but not the most general) random parameters specifica-tion assumes component-wise one-to-one mappings σ(ζ), β(ζ) that arerestricted in sign for taste parameters that are signed a priori. An impor-tant and practical case for applications is F (ζ|s, θ) multivariate normal,with the deep parameters θ specifying the mean and variance of this dis-tribution; often with the mean a linear function of s. Alternately, a quitegeneral specification is F (ζ|s, θ) = C(F1(ζ1|s, θ), . . . , FT (ζT |s, θ); s, θ),where C is any copula3 and the Fi are the univariate marginal CDFsof the components of ζ, typically exponential, log normal, or normalwith or without censoring. For example, let Φ(ζ; Ω) denote a multi-variate normal CDF with standard normal marginals and a correlationmatrix Ω. Then C(u1, . . . , uT ) ≡ Φ(Φ−1(u1), . . . ,Φ−1(uT ); Ω) is a mul-tivariate normal copula that generates the multivariate distributionF (ζ|s, θ) = Φ(Φ−1(F1(ζ1|s, θ)), . . . ,Φ−1(FT (ζT |s, θ)); Ω(s, θ)). In appli-cations, it is sometimes parsimonious to restrict Ω(s, θ) to have a low-dimensional factor-analytic structure: Let Λ be a T×K matrix of rankKwhose columns are factor loadings, scaled so that

∑Tt=1 Λ2

ti < 1, and let Ψdenote a diagonal matrix with diagonal elements Ψii =

√1−

∑Tt=1 Λ2

ti.Then Ω(s, θ) = Ψ(s, θ)2 + Λ(s, θ)Λ(s, θ)′ is the factor-analytic corre-lation matrix. Sampling from the multivariate normal copula with acorrelation matrix Ω(s, θ) can be carried out by setting ς = Ω(s, θ)1/2ξ,or in the case of a factor-analytic structure, ς = Ψ(s, θ)ξ + Λ(s, θ)ω,where ξ and ω are vectors of standard normal variates, and then settingζi = F−1

i (ςi|s, θ) for each component.It is also sometimes important to allow segmentation of the pop-

ulation into classes, with each class having its own distribution of ζ,see Desarbo et al. (1995). Segmentation may be unobserved, with the

2In the framework of maximum likelihood estimation, Lagrange Multiplier teststhat can be calculated at fitted base model parameters are often practical for modelspecification tests.

3A copula is a multivariate CDF C(u1, . . . , uT ) whose univariate marginals areall uniform [0, 1]. Any multivariate distribution can be written in terms of its copulaand marginals, and there are necessary and sufficient conditions for a function to bea copula, see Joe (1997), and Fosgerau et al. (2013).

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 60: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

60 Choice Model Estimation and Forecasting with CBC Data

probability of consumers falling in a “latent class” becoming a param-eter of the problem, or may be observed, e.g., classification by ageor gender. An example would be a population that is segmented bypreferences for “brands” or “classes” of products. The vector xjmn cancontain indicators that are turned on for specific categories of products,and a class might be defined as the sub-population that has positiveor negative preferences for a specific category (i.e., the component ofβ(ζn) is non-zero for sample members in this class, and the remainingcomponents of β(ζn) for product categories are zero). The probabilityof a person falling in this class would be another parameter in θ, andwithin each class, all other taste parameters would have a (conventional)distribution conditioned on the class.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 61: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

5Maximum Simulated Likelihood (MSL) Analysis

of CBC Data

Consider CBC data in the format described in Section 2. Supposeconsumers are nearly neoclassical, with utility (3.1) and tastes ρm = ζ

that are homogeneous across menus. Then their choices are generatedby the discrete choice model (3.8). Assume this model is parameterizedwith deep parameters θ, and consider maximum likelihood estimationof θ. In the simplest case, the parameters σ(ζ) and β(ζ) in this model areassumed to be homogenous over the population, so that (3.8) reducesto the flat MNL model (3.6). If a flat MNL model is believed to besufficient to explain and forecast real market choices, then, the loglikelihood of the stated choice data is

logL(σ, β) =N∑n=1

M∑m=1

J∑j=0

djmnlog

exp(xjmnβ−pjmn

σ

)∑Ji=0 exp

(ximnβ−pimn

σ

).

(5.1)

Next allow for preference heterogeneity across consumers. As dis-cussed earlier, CBC studies normally cannot make the number of menusM large enough to make estimation of “fixed effects” taste coefficients foreach consumer reliable. The most practical alternative is to make σ(ζ)and β(ζ) “random effects” drawn for each consumer from a distribution

61

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 62: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

62 Maximum Simulated Likelihood (MSL) Analysis of CBC Data

F (ζ|s, θ) that depends on deep parameters θ and may depend on somecomponents of the socioeconomic vector s. For simulation, it is some-times useful to assume that ζ can be written as a transformationζ = W (ω, s, θ) of a random vector ω that has a “standard” CDF Φ(ω).Assume the distribution F has a density f(ζ|s, θ). Then the likelihoodfor this consumer is

Ln(dn|xn, pn, θ) =∫ζPn(ζ) · f(ζ|s, θ)dζ

≡∫ωPn(W (ω, s, θ))·Φ(dω), (5.2)

where Pn(ζ) ≡∏Mm=1

∏Jj=0

exp(vjmn(ζ)σ(ζ)

)∑J

i=0 exp(vimn(ζ)σ(ζ)

)djmn and vjmn(ζ) ≡

X(Wn, An, Tn − pjmn, sn, zjmn)β(ζ)− pjmn. The gradient of Equation(5.2) is

∂Ln(dn|xn, pn, θ)/∂θ =∫ζPn(ζ) · ∂ log f(ζ|sn, θ)

∂θ· f(ζ|sn, θ)dζ

≡∫ω[∂Pn(W (ω, sn, θ))/∂θ] · Φ(dω). (5.3)

Standard maximum likelihood theory states that under mild regularityconditions that will ordinarily be satisfied in this application, iterationof the BHHH condition1

√N(θi+1 − θi) = (ΣN )−1 1√

N

N∑n=1

(∂ logLn

(dn|xn, pn, θi

)∂θ

),

(5.4)

where

ΣN = 1N

N∑n=1

(∂ logLn(dn|xn, pn, θi)

∂θ

)(∂ logLn(dn|xn, pn, θi)

∂θ

)′,

(5.5)

1Berndt et al. (1974). Alternately, Gauss–Newton or Newton–Raphson iterationto the maximum can be used.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 63: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

63

with an added step-size adjustment to avoid overshooting and accelerateconvergence, is guaranteed to converge from any initially consistentestimator θ0 to an estimator θ that maximizes the likelihood of thesample L(θ) ≡

∏Nn=1 Ln(dn|xn, pn, θ). Further,

√N(θ − θ∗) is asymp-

totically normal with mean zero for the true θ*, and a covariancematrix that is consistently estimated by (ΣN )−1. Note that this resultdoes not guarantee convergence to the maximum likelihood estimatorfrom all starting values for θ, nor can one in general be assured ofthe global strict concavity that would imply such global convergence.Consequently, it is useful to start the MLE iteration from a grid ofstarting values, and perhaps introduce annealing (discussed in the nextsection) during the iterations, to assure that search spans the parameterspace and has a high probability of finding a global rather than a localmaximum.

Maximum simulated likelihood (MSL) replaces Ln(dn|xn, pn, θ) =∫ω P

n(W (ω, sn, θ)) · Φ(dω) with the approximation 1R

∑Rr=1 P

n

(W (ωr, sn, θ)), where the ωr are R independent pseudo-randomdraws from the CDF Φ(ω), and uses a commensurate approximation1R

∑Rr=1

∂Pn(W (ωr,sn,θ))∂θ to the gradient ∂Ln(dn|xn,pn,θ)

∂θ to iterate to themaximum simulated likelihood using Equation (5.4), see McFadden andRuud (1994), and Hajivassiliou and Ruud (1994). Alternately, a methodof simulated score (MSS) uses Equation (5.4) to find a root of the score ofthe likelihood, with Ln(dn|xn, pn, θ) and ∂Ln(dn|xn, pn, θ)/∂θ approxi-mated respectively by 1

R

∑Rr=1 P

n(ζr) and 1R

∑Rr=1 P

n(ζr) · ∂ log f(ζr|sn,θ)∂θ ,

where the ζr are independent pseudo-random draws from F (ζ|sn, θ),see Hajivassiliou and McFadden (1998). If R rises at least as rapidly as√MN , then either of these simulation estimators will be asymptotically

equivalent to the (intractable) exact maximum likelihood estimator.Train (2007) points out that the gradient approximations in MSL andMSS are not identical for fixed R, even when generated by the samerandom draws, and that for numerical stability, MSL should use itscommensurate gradient approximation.

Suppose F (ζ|s, θ) is multivariate normal with density n(ζ;µ,Ω),where Ω = ΛΛ’ and θ = (µ,Λ), and σ(ζ), β(ζ), δ(ζ) are predeterminedcomponent-wise transformations of ζ; µ may be a linear function µ0 +

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 64: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

64 Maximum Simulated Likelihood (MSL) Analysis of CBC Data

µ1 ·s. Differentiating, logn(ζ;µ,Ω) ∝ − log |Λ|−(ζ−µ)′(ΛΛ′)−1(ζ−µ)/2implies ∂ logn(ζ;µ,ΛΛ′)

∂µ = Ω−1(ζ−µ) and ∂ logn(ζ;µ,ΛΛ′)∂Λ = Ω−1(ζ−µ)(ζ−

µ)’ Ω−1 − Ω−1Λ, and the gradient of Equation (5.2), given by Train(2007), is

∂Ln(dn|xn, pn, θ)∂µ

=∫ζPn(ζ) · Ω−1(ζ − µ) · F (dζ|sn, θ)

∂Ln(dn|xn, pn, θ)∂Λ =

∫ζPn(ζ) · Ω−1(ζ − µ)(ζ − µ)′Ω−1 − Ω−1Λ

·F (dζ|sn, θ). (5.6)

A straightforward MSS approach to computation of Equations (5.2)–(5.6) in the multivariate normal case is to start from a large repositoryr = 1, . . . , R of draws of i.i.d. standard normal vectors ωr, which areheld fixed during estimation to avoid chatter and maintain stochasticequicontinuity, and iteratively consider trial values θ = (µ,Λ), withΩ = ΛΛ’, and values ζr ≡ µ + Λωr for the intermediate parametervectors ζ that enter Pn(ζ). Then approximate Equation (5.2) at θ byLn(dn|xn, pn, θ) = 1

R

∑Ri=1 P

n(ζr), Equation (5.3) by

∇µLn = 1R

R∑i=1

Pn(ζr)Ω−1(ζr−µ)

∇ΛLn = 1R

R∑i=1

Pn(ζr)Ω−1(ζr−µ)(ζr−µ)′Ω−1 − Ω−1Λ,

(5.7)

and Equation (5.5) by an average over n of the outer product of thegradient (5.7), reshaped into a vector; i.e., define gn to be a column vectorthat concatenates ∇µLn and a vector containing the lower triangularelements of ∇ΛLn, column by column, and ΣN = 1

N

∑Nn=1 gngn

′.2 Othertreatments of the problem of estimating mixed MNL models, using MSLrather than MSS, are Ben-Akiva et al. (1993), Revelt and Train (1998),Train (1999), Train (2007), Train (2009), Brownstone and Train (1999),

2Technically, the independent elements in the lower triangular array Λ arevectorized and concatenated with µ to form the vector θ; the R command for doingthis is θ < −c(µ,Γ [lower.tri(Γ, diag = TRUE)]). The corresponding vector gn isformed by the R command ∇ΛLn [lower.tri(∇ΛLn, diag = TRUE)]).

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 65: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

65

Brownstone et al. (2000), Chesher and Santos-Silva (2002), Hensherand Greene (2002), and Walker et al. (2007).3

When the target of the CBC is market prediction rather than esti-mates of individual preferences, there is only a modest loss of generalityor efficiency in ignoring σ, β homogeneity across menus for the sameconsumer. Then model (5.2) reduces to the mixed MNL model (3.9)that ignores the panel structure of stated choices from multiple menus.The simulation method described above still applies, but it is necessaryto use a “sandwich” formula to obtain an estimated covariance matrixof the estimates that is consistent when there are common tastes thatinduce correlation in choices across menus.

3Prof. Arne Rise Hole has developed STATA modules to estimate mixedlogit models in preference-space and wtp-space. The modules, called “mixlogit”for preference space and “mixlogitwtp” for wtp space, are available at https://www.sheffield.ac.uk/economics/people/hole/stata or by typing “ssc install mixlogit”and “ssc install mixlogitwtp” in the STATA command window. The two modulesshare syntax and include options for estimation by various maximization techniques,including Newton–Raphson and BHHH; for correlated or uncorrelated random terms(coefficients in preference space, WTPs and alpha in wtp-space); and for normal andlognormal distributions.

Matlab, Gauss, and R codes to estimate mixed logits in preference-space by MSLare available on K. Train’s website at: https://eml.berkeley.edu/~train/software.html.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 66: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

6Hierarchical Bayes Estimation

An alternative and/or complement to classical statistical estimation isthe use of Bayes methods. These turn out to be particularly convenientfor analysis of CBC data, and have been refined and applied widelyin marketing, led by the contributions of Allenby and Rossi (1993,2006), Allenby and Lenk (1994) and Rossi et al. (2005). Textbooktreatments of Bayesian statistics can be found in Koop (2003), Gelmanet al. (2004), and McFadden (2015) gives a primer on decision-theoreticand statistical properties. Under general conditions, the mean of theBayesian posterior of a parameter is asymptotically equivalent to theclassical maximum likelihood estimator (MLE), and the variance of theBayesian posterior is asymptotically equivalent to the sampling varianceof the MLE. This equivalence implies that a researcher can implement aBayesian estimation procedure and treat the mean and variance of theposterior the same as classical estimates and their covariances, withoutnecessarily adopting a Bayesian approach to statistics. This section firstreviews the basics of Bayes estimation, then shows its use for CBCanalysis.

66

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 67: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

6.1. Bayesian estimation 67

6.1 Bayesian estimation

Observations y are realizations of a random vector that are governedby a probability law or data generation process (DGP) indexed by aparameter vector θ and described by a density or likelihood L(y, θ). Forexample, in a study of automobile stated purchase choices, one mightspecify a flat MNL data generation process, so that L(y, θ) would havethe form (5.1). Bayesian statistical approach postulates that θ can itselfbe treated as a random quantity, see for example Zellner and Rossi(1984). Prior to observing y, the researcher has beliefs on what valuesof θ are likely. These beliefs can be expressed in terms of a prior densityk(θ). Then, the joint density of y and θ is L(y, θ)k(θ). By Bayes Law,the conditional density of θ given y, which is called the researcher’sposterior belief about θ given the data, is

K(θ|y) = L(y, θ)k(θ)∫θ′ L(y, θ′)k(θ′)dθ′ . (6.1)

The researcher can use the posterior to evaluate payoffs that dependon θ, or to calculate an estimate t(y) of θ. Usually, the mean of theposterior is taken as the estimator of θ. The Bayes estimator balancesthe information on θ contained in the likelihood L(y, θ) for the observeddata and that contained in the prior k(θ). Priors can be either diffuseor tightly concentrated, and if they are tight, they can be “good”,concentrated near the true value of θ, or “bad”, concentrated far awayfrom it. The Bayes estimator is asymptotically equivalent to the classicalML and MSL estimators under general conditions, see Train (2001).To see this, assume for simplicity that θ is a one-dimensional variationfrom a true value θ0, and start from the log likelihood function

logL(y, θ) =N∑n=1

log f(yn, θ), (6.2)

where yn is the data, f(yn, θ) is the density for observation n, and N issample size. Take the identity

1 ≡∫ +∞

−∞f(yn, θ)dyn ≡

∫ +∞

−∞exp(log f(yn, θ))dyn (6.3)

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 68: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

68 Hierarchical Bayes Estimation

and differentiate it in θ to get

0 ≡∫ +∞

−∞

∂ log f(yn, θ)∂θ

f(yn, θ)dyn and

0 ≡∫ +∞

−∞

[(∂ log f(yn, θ)

∂θ

)2+ ∂2 log f(yn, θ)

∂θ2

]f(yn, θ)dyn. (6.4)

For regular problems, I(θ′) =∫+∞−∞

(∂ log f(yn,θ′)

∂θ

)2f(yn, θ)dyn, termed

the Fisher Information, is positive for θ′ near θ. A Taylor’s expansionaround the true value θ0 of θ gives

K(θ|y)/K(θ0|y) = exp[SN ·√N(θ − θ0)− 1

2IN (√N(θ − θ0))2

+ 1√Nk′ ·√N(θ − θ0)]. (6.5)

By a central limit theorem, SN ≡ 1√N

∑Nn=1

∂ log f(yn,θ0)∂θ converges to

a normal random variable with mean zero and variance I(θ0). Bya law of large numbers and additional technical arguments, IN =1N

∑Nn=1

(∂ log f(yn,θ)

∂θ

)2converges to I(θ). Then, for N large, the argu-

ment in the exponent on the right-hand-side of Equation (6.5) behaveslike√N(θ − θ0) times the expression SN + 1√

Nk′ − 1

2IN√N(θ − θ0).

But this expression has the negative of the sign of√N(θ − θ0) when√

N(θ− θ0) is large in magnitude, establishing from Equation (6.5) thatthe posterior mode has

√N(θ − θ0) bounded. Further, the contribution

of the prior, 1√Nk′, is asymptotically negligible, so that the mode of

the posterior and the MLE converge and have the same asymptoticdistribution. Finally, the concentration of the density K(θ|y) in a neigh-borhood where

√N(θ − θ0) is bounded, along with an assumption that

the tails of k(θ) are sufficiently thin, so that the mean of the posterioris not dominated by its tails, implying that the mean of K(θ|y) andits γ-quantiles for γ in the interior of (0, 1) all are in a

√N(θ − θ0)

bounded neighborhood of θ0, and hence 1√N

close to θ0 and the MLEfor N large. More precise statements can be found in Ibragimov andHas’minskii (1973) and Strasser (1975). Huber and Train (2001) findthat in a dataset of typical size, the MLE and Bayes procedures pro-duce very similar estimates for a mixed logit model in preference space.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 69: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

6.2. Selecting priors and computing posteriors 69

Scarpa et al. (2008) obtain similar results from Bayesian and MLE formixed logit models in willingness-to-pay space. Train (2009) providesfurther discussion and examples.

6.2 Selecting priors and computing posteriors

For any given likelihood function, if the prior and posterior distribu-tions have the same form but different parameters, then the prior iscalled “conjugate.” Statistical models with conjugate priors are special,but convenient when applicable. Fink (1997) describes the conditionsunder which models admit conjugate priors, notably distributions inthe exponential family, and discusses various transformations of themodels that allow conjugate priors to be used. Gelman et al. (2004)give a compendium of known priors, see also the Wikipedia entry onconjugate priors. Of value in CBC applications are the Dirichlet priorthat is conjugate for a multinomial likelihood,1 and the Normal orWishart prior that is conjugate for a multivariate normal likelihoodwith unknown mean or unknown covariance matrix.2 Drawbacks ofconcentrating on conjugate priors are, first, that even if the posteriordensity is analytically determined, it is not necessarily simple to com-pute moments or draw simulated values from it, and, second, thatnatural parameterizations of models for CBC data often do not admitconjugate priors.

Usually a CBC analysis can start with a prior density and a choicemodel likelihood that can be expressed in terms of computationally

1If the likelihood of counts (k1, . . . , kJ ) from N trials is(

Nk1 · · · kJ

)pk1

1 · · · pkJJ

and the prior is the Dirichlet density Γ(α1+···+αJ )Γ(α1)···Γ(αJ ) p

α1−11 · · · pαJ−1

J , then the posterioris Dirichlet with parameters αj + kj .

2If the likelihood is formed from n draws from a normal N(µ,Σ−1) density withunknown µ or Σ, the prior for µ is normal, or the prior for Σ−1 is independentlyinverse Wishart, then the posterior distribution for µ is multivariate normal andfor Σ−1 is independently inverse Wishart. There are alternative parameterizationsand conjugate priors for Σ that may have better computational and statisticalproperties. For example, an LU decomposition expresses Σ as Σ = LL′, where L islower triangular with a positive diagonal, with a normal/variance–gamma conjugateprior. For analyses of alternative distributions to the inverse Wishart assumption,see Song et al. (2018) and Akinc and Vanderbroek (2018).

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 70: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

70 Hierarchical Bayes Estimation

tractable functions. Even so, except for special cases of conjugate priors,it is usually impossible to obtain analytically the mean

∫θ θ ·K(θ|y)dθ

or other key characteristics of the posterior density, or even to computeeasily the scaling factor A =

∫θ L(y|θ)k(θ)dθ that makes K(θ|y) =

L(y|θ)k(θ)/A integrate to one. However, with contemporary computersit is practical in most CBC applications to use some combination ofnumerical integration and simulation methods to estimate features ofthe posterior density such as its mean. There is a large literature onmethods for these calculations and their computational efficiency, seefor example Abramowitz and Stegum (1964, chap 25.4), Press et al.(2007), Quarteroni et al. (2007, Chap 9), Neal (2011), Carpenter et al.(2017), and Geweke (1997). This monograph will describe several basicmethods, particularly those that have proven useful for Bayes estimation,but we refer readers interested in mathematical and computationaldetails back to these sources.

Deterministic numerical integration is often practical when thedimension of θ is low. Consider integrals

A ·E(θj |y) ≡∫θ·θj · L(y|θ)k(θ)dθ

≡∫θ·θj ·

L(y|θ)k(θ)

h(θ)

h(θ)dθ, (6.6)

for j = 0, 1, with the last formula obtained by multiplying and dividingby a positive function h(θ) selected to facilitate computation. Forexample, h may be a probability density with a CDF that is easyto compute. Quadrature is a standard approach to computing Equation(6.6): For a finite collection of points θi, i = 1, . . . , I, and associatedweights hi, approximate Equation (6.6) by

∑Ii=1 θ

ji ·L(y|θi)k(θi)

h(θi)

hi.

For example, the θi might enumerate the centroids of a finite grid thatpartitions θ-space, with hi the probability that a random vector withdensity h falls in the partition with centroid θi. This is the computationalanalogue of Reimann integration, with accuracy limited by the degreeof variation of the integrand L(y|θ)k(θ)/h(θ) on the partition; e.g.,the accuracy with which the integrand can be approximated by stepfunctions. Computation efficiency is improved by selecting a practicalh such that the integrand varies slowly with θ. There are a variety of

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 71: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

6.2. Selecting priors and computing posteriors 71

mathematical methods that exploit smoothness features of the integrandto improve computational efficiency, e.g., Quarteroni et al. (2007, 9.1–9.7). A general idea behind quadrature is that various series of functionsother than step functions, such as orthogonal polynomials or splines,have analytic integrals with respect to h, and points and weights can beselected so that the finite approximation to Equation (6.6) is exact for theinitial terms in the specified series. For example, a widely used methodtermed Gaussian quadrature selects points and weights such that theapproximation is exact for the family of Hermite orthogonal polynomialsup to a given order. This can be computationally quite efficient, but twodrawbacks are that the number of functional evaluations required risesto the power of the dimension of θ and quickly becomes intractable,and that the approximation is poor when the integrand varies rapidlyrelative to the approximating Hermite polynomials.

A simple simulation method for approximating Equation (6.6) isimportance sampling. Start from a density h(θ) from which it is compu-tationally easy to draw random deviates and for which L(y|θ)k(θ)/h(θ)is bounded for all θ. An average of the integrand in Equation (6.6) overa sample of simulated independent random deviates from h will by alaw of large numbers converge to the integral.3 The primary drawbackof importance sampling is that unless the bracketed integrand is quitesmooth and slowly varying in θ, the estimates will be relatively inef-ficient (as measured by precision achieved per unit of computationalresources). However, variance reduction methods such as quota samplingfrom partitions of θ-space and use of antithetic variates that exploitsymmetries in h can improve efficiency, e.g., Rubinstein et al. (1985).There are a variety of simulation methods that may improve substan-tially on the computational efficiency of importance sampling, some

3When θj ·L(y|θ)k(θ)

h(θ)

is bounded by a constant C for j = 0, 1, Hoeffd-

ing’s inequality (Pollard, 1984, Appendix B) establishes that Prob(∣∣∣ 1I ∑I

i=1 θji ·

L(y|θi)k(θi)h(θi)

−A ·E(θj |y)

∣∣∣ > η)< 2 · e−Iη

2/2C2. Then with some manipulation,

one can show that for η < A/2, Prob(∣∣∣∣∑I

i=1θi

L(y|θi)k(θi)

h(θi)

∑I

i=1

L(y|θi)k(θi)

h(θi)

−E(θ|y)∣∣∣∣ > η

)<

4 · e−Iη2/32·max(C4,1), so convergence is exponential.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 72: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

72 Hierarchical Bayes Estimation

utilizing mathematical analogies to conservation of energy in physics,tempering and annealing in metallurgy, and minimization of entropy inprobability theory. We explain several simple methods that are widelyused in Bayesian analysis, and conclude with a survey of more advancedmethods.

One simple alternative to the i.i.d. sampling from the density h usedin importance sampling is termed acceptance/rejection (A/R) sampling.This method gives a sample from the posterior K(θ|y) without requiringestimation of the normalizing scale factor A =

∫θ L(y|θ)k(θ)dθ. Let

f(θ|y) ≡ L(y|θ)k(θ) ∝ K(θ|y), and find a comparison density h(θ)from which it is easy to make pseudo-random draws, with the propertyBh(θ) ≥ f(θ|y) for some positive scalar B.4 Have the computer drawa sequence of pairs of pseudo-random numbers (θi, ui), where θi is adraw from h and ui is a draw from a uniform density on [0, 1]. Acceptand retain θi if uiBh(θi) ≤ f(θi|y); otherwise, reject and discard it.Then

Prob(θ|accept) = Prob(θ& accept)∫Prob(θ′& accept)dθ′ = h(θ) · f(θ|y)/Bh(θ)∫

h(θ′) · f(θ′|y)/Bh(θ′)dθ′

= f(θ|y)/B∫f(θ′|y)/Bdθ′ = K(θ|y), (6.7)

and the sequence of accepted points θi will be a random sample fromK(θ|y) as claimed. To illustrate, consider a choice experiment in whichn = 1, . . . , N subjects are asked if they will purchase a product at pricexn, and dn is an indicator that is one if the response is “Yes”, zerootherwise; then yn ≡ (xn, dn). Suppose the probability model is logistic,Prob(yn|xn) = exp(−dnxnθ)/(1+exp(−xnθ)), and the prior is standardexponential, eθ for θ < 0. Then K(θ|y) ∝ f(θ|y) ≡

∏Nn=1

exp(−ynxnθ)1+exp(−xnθ) ·

4In the R core statistical package, vectors of draws from the beta, binomial,Cauchy, Chi-squared, Exponential, F, Gamma, Hypergeometric, Logistic, Log-normal,Multinomial, negative binomial, normal, Poisson, T, uniform, and Weibull distri-butions are obtained from the respective functions rbeta, rbinom, rcauchy, rchisq,rexp, rf, fgamma, rhyper, rlogis, rlnorm, rmultinom, rnbinom, rnorm, rpois, rt, runif,rweibull, with arguments specifying distribution parameters. These functions canalso be used, with prefixes “d”, “p”, or “q” substituted for “r”, to get densities, CDFs,or quantiles, respectively. Additional distributions are available in packages fromCRAN.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 73: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

6.2. Selecting priors and computing posteriors 73

0

0.2

0.4

0.6

0.8

1

-1 -0.8 -0.6 -0.4 -0.2 0

h(θ)

f(θ|y)/B

Figure 6.1: Illustration of acceptance/rejection method.

eθ1(θ < 0), an expression whose integral does not have a closed form.Note that f(θ|y) ≤ eθ1(θ < 0); take h(θ) ≡ eθ1(θ < 0) and B = 1.

Figure 6.1 illustrates how the acceptance/rejection method works inthis case. In this figure, points are drawn with equal probability fromthe area below the blue curve that plots Bh(θ), extending indefinitelyto the left, and are accepted if they lie below the red curve that plotsf(θ|y), with the scale factor B set so that the red curve is never higherthan the blue curve. The efficiency or yield of the procedure, the shareof sampled points accepted, is given by the ratio of the area under thered curve to the area under the blue curve.

A limitation of the simple acceptance/rejection method is that itcan have low efficiency unless the h curve and scale factor B can betuned to f(θ|y); this is particularly true when the parameter vector isof high dimension. There are several modifications to A/R methodsthat may attain higher yields. First, in acceptance/rejection sampling,it may be possible to partition the support of K(θ|y) into a finitenumber of intervals or rectangles Aj , and define a value Bj that isa fairly tight upper bound on f(θ|y)/h(θ) for θ ∈ Aj . Define weightswj = h(Aj)/Bj∑

j′ h(Aj′ )/Bj′. Draw θ from h and u from uniform [0, 1], determine

the rectangle Aj in which θ lies, and accept θ with weight 1/wj ifuBjf(θ|y) ≤ h(θ). Then the weighted accepted sample has an empiricalweighted CDF that converges to the CDF of K(θ|y), and the yieldis high when the bounds within each rectangle are good. Second, theBj can be adjusted adaptively. If a “burn-in” sample of θ’s falls inrectangle j, then the maximum of f(θ|y)/h(θ) in this sample is aconsistent estimator of the best, or least upper bound, Bj . Using this

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 74: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

74 Hierarchical Bayes Estimation

best estimate may under-accept θ from regions where the acceptedsubsample maximum is less than the true least upper bound, butadaptively adjusted Bj will converge very fast (e.g., at a 1/N ratherthan a 1/

√N rate), so the bias will be negligible in simulations of

reasonable size. Third, the partition into rectangles Aj can be refinedrecursively and adaptively to reduce the gap between the maximumand the minimum of f(θ|y)/h(θ) in each rectangle. If these refinementsare made as a rate that is asymptotically negligible relative to theaccepted sample size, they will introduce only asymptotically negligiblenoise into the simulation. A variant of A/R sampling is termed slicesampling. It introduces an auxiliary variable τ with a uniform density on[0, L(y|θ)k(θ)] and a density π(θ|τ) = 1(τ ≤ L(y|θ)k(θ)). Starting fromθ−1, the sampler first draws τ from [0, L(y|θ−1)k(θ−1)], next places arandom rectangle in θ-space around the point (τ, θ−1), steps out thisrectangle until it covers the set θ|τ ≤ L(y|θ)k(θ), and then samples θfrom the resulting rectangle until τ ≤ L(y|θ)k(θ) is satisfied. In Figure6.1, this corresponds to drawing horizontal strips at random, and thensampling from the portion of these strips that lie under the curve.An interpretation of slice sampling is that it is a stochastic version ofLebesgue integration, while simple A/R sampling is a stochastic versionof Reimann integration.

Additional efficiency may be achievable using Monte Carlo MarkovChain (MCMC) methods, see Tierney (1994), Albert and Chib (1996),and Chib and Greenberg (1996), and McFadden (1996). A Metropolis–Hastings sampler is a Markov process that recursively draws points fromthe parameter space Θ using a positive conditional density q(θ|θ−1)from which it is convenient to sample. Given θ−1, draw a trial θ∗from q(θ∗|θ−1) and define γ(θ∗, θ−1|y) ≡ min

(1, f(θ∗|y)·q(θ−1|θ∗)

f(θ−1|y)·q(θ∗|θ−1)

). A

key feature of the Metropolis–Hastings sampler is the symmetry of thefunction γ(θ∗, θ−1|y) that enters the acceptance criterion. This makesthe sampler reversible in time, a defining characteristic of a successfulsampler. This critical value reduces to min

(1, f(θ∗|y)

f(θ−1|y)

)when q(θ|θ−1)

is symmetric, the original, more restrictive Metropolis sampler; e.g.,specify θ − θ−1 multivariate normal with mean zero. If a random scalaru from a uniform density on [0, 1] satisfies u < γ(θ∗, θ−1|y) set θ = θ∗;

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 75: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

6.2. Selecting priors and computing posteriors 75

otherwise set θ = θ−1. This defines a Markov transition probability

P (θ|θ−1) =q (θ|θ−1) γ (θ, θ−1|y) if θ 6= θ−1

1−R(θ−1) if θ = θ−1(6.8)

where R(θ−1) =∫θ′∈Θ q(θ′|θ−1) · γ(θ′, θ−1|y)dθ′ is the probability of

acceptance. The posterior K(θ|y) = A · f(θ|y) for some constant A isan invariant of this Markov process. If θ−1 has density A · f(θ−1|y), thejoint density of θ and θ−1 is P (θ|θ−1) ·A · f(θ−1|y). Then, the marginaldensity of θ is

A ·∫θ−1∈Θ

f(θ−1|y)q(θ|θ−1) · γ(θ, θ−1|y)dθ−1 + 1(θ = θ−1)

·A · f(θ−1|y) · (1−R(θ−1))

= A ·∫θ−1∈Θ

f(θ|y)q(θ−1|θ) ·min(

1, f(θ−1|y) · q(θ|θ−1)f(θ|y) · q(θ−1|θ)

)dθ−1

+∫θ−1∈Θ

1(θ = θ−1) ·A · f(θ−1|y) · (1−R(θ−1))

= A · f(θ|y) ·R(θ) +A · f(θ|y) · (1−R(θ−1)) = A · f(θ|y),(6.9)

and K(θ|y) is an invariant density. In Equation (6.14), 1(θ = θ−1)denotes a unit probability on the set of points in Θ×Θ that meet thecondition. The requirement that q(θ|θ−1) > 0 on Θ×Θ, plus technicalconditions on the tails of q(θ|θ−1) relative to f(θ|y) when Θ is notcompact, are sufficient to assure that there is a unique invariant densityand the Markov process converges to this invariant density.5 A drawbackof MCMC methods is that they can have high serial correlation, so thatvery large numbers of draws and substantial “burn-in” iterations areneeded to get reliable empirical approximations to the posterior.

It is fairly common in applications to require a sample from a compu-tationally intractable multivariate distribution whose one-dimensional

5The conditions that must be met on q(θ|θ−1) and Θ are irreducibility (i.e., theprocess can always go from one point in Θ to any other point in a finite number ofsteps), acyclicity (i.e., the process does not always return to its starting point in afinite number of steps), and positive recurrent (i.e., the expected time until a returnto a subset of positive measure is finite), see Tierney (1994), and McFadden (1996).

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 76: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

76 Hierarchical Bayes Estimation

conditional distributions are relatively manageable. A specializationof the Metropolis–Hastings sampler called Gibbs sampling can beused in this case. Suppose θ is of dimension d, and θ = (θ1, . . . , θd)has a multivariate density g(θ1, . . . , θd) with univariate condition-als gi(θi|θ1, . . . , θi−1, θi+1, . . . , θd) from which it is computationallytractable to sample using, say, the quantile transformations of thesedensities, acceptance/rejection sampling, or a MCMC procedure. Froman initial vector θ′ = (θ′1, . . . , θ′d), the sampler loops through draws of θ′′ifrom gi(θi|θ′′1 , . . . , θ′′i−1, θ

′i+1, . . . , θ

′d) for i = 1, . . . , d. The procedure then

proceeds recursively, starting next from the vector θ′′. The sampler canbe started from any initial θ vector. A leading example is a truncatedmultivariate normal distribution g whose one-dimensional conditionalsgi are (truncated) univariate normal. To illustrate, in the R statisticalpackage draws from gi subject to a truncation θi ∈ [ci,+∞) are givenby qnorm(runif(1,pnorm(ci, µi, σi, 1, 0), 1)).

There has been considerable study of the conditions under whichMetropolis–Hastings and Gibbs sampling methods perform poorly, seeWu and Fitzgerald (1996), Neal (2011), Hoffman and Gelman (2014),and Betancourt and Girolami (2013). WhenK(θ|y) has a strong gradientrelative to the conditional density q(θ|θ−1) in the Metropolis–Hastingssampler, or large differences between conditional and marginal variancesin Gibbs sampling, sampler output can resemble local random walksthat are highly serially correlated and fail to span the θ-space well ina reasonable number of iterations. The problem is particularly criticalwhen K(θ|y) has several isolated sharp peaks. One straightforwardmethod of dealing with this problem is called annealing, a term bor-rowed from metallurgy where a material is repeatedly heated above itscrystallization temperature to allow atoms to migrate, and then slowlycooled. The analogy here replaces the target f(θ|y) ≡ L(y|θ)k(θ) byfτ (θ|y) = f(θ|y)τ , which is flatter than f(θ|y) for τ < 1, and then overthe course of the iterations periodically allows τ to drop below one for afew iterations. Used in a Metropolis–Hastings sampler with θ’s selectedfrom parts of the chain where τ = 1, this will lead a broader span ofaccepted values. However, the frequency, magnitude, and duration ofτ < 1 deviations, and “burn-in” times following these deviations, allhave to be selected to tune the sampler to the application.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 77: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

6.2. Selecting priors and computing posteriors 77

The most sophisticated versions of Monte Carlo Markov Chainmethods currently in wide use are Hamiltonian Monte Carlo methods,particularly the No-U-Turn Sampler (NUTS) of Carpenter et al. (2017).These methods accomplish the same things as well-tuned annealing,and utilize a combination of mathematical theory and tinkering toautomate some or all of the tuning process. A good introduction tothese methods is Neal (2011). Hamilton Monte Carlo methods introduceauxiliary variables, in this case a vector ω that is of the same dimensionas θ, two functions V (θ) = − log(L(y|θ)k(θ)) and W (ω) = ω′M−1ω/2,where M is a positive definite matrix, and dynamic equations of motionin continuous time,

dθ/dt = ∂W (ω)/∂ω ≡M−1ω and

dω/dt = −∂V (θ)/∂θ ≡ ∂ log(L(y|θ))/∂θ + ∂ log(k(θ))/∂θ. (6.10)

Interpret ω as a momentum vector,W (ω) as kinetic energy, and V (θ) aspotential energy. This dynamic system is time reversible and preservesthe Hamiltonian H(θ, ω) ≡ V (θ) +W (ω); i.e., dH(θ, ω)/dt ≡ 0. This isanalogous to conservation of energy in physical systems. The dynamicsystem moves rapidly out of regions where potential energy (i.e., poste-rior probability) is low, and slowly out of regions where potential energyis high. It has a strong mathematical property, called symplecticness,that the Jacobian of the transformation from (θ, ω) at time t to timet + s is one, see Neal (2011, page 116). An implication of this prop-erty is that this dynamic system is probability preserving; i.e., it willmove along contours of constant probability density. In computationalapproximation to this dynamic system, continuous time is replaced bydiscrete time moving in small steps ∆. Start from θ(t) = θ−1 and anew independent draw ω(t) = M−1/2ξ, where ξ is a draw of a standardnormal vector, independent of the previous state. Advance θ and ω fors = 0, . . . , S − 1 steps using what is termed the leapfrog method:

ω(t+ (s+ 1/2)∆) = ω(t+ s∆)− (∆/2) · ∂V (θ(t+ s∆))/∂θ,

θ(t+ (s+ 1)∆) = θ(t+ s∆) + ∆ ·Mω(t+ (s+ 1/2)∆),

ω(t+ (s+ 1)∆) = ω(t+ (s+ 1/2)∆)− (∆/2)

· ∂V (θ(t+ (s+ 1)∆))/∂θ. (6.11)

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 78: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

78 Hierarchical Bayes Estimation

The half-steps have to be computed only for steps 1 and S, since inbetween the final half-step for s combines with the initial half-stepfor s+ 1. With this method, the discrete time approximation retainsthe time-reversibility and symplecticness properties of the continuoustime process. At the end of the S steps, one has a proposed state(θ(t+ S∆), ω(t+ S∆)). This proposal is accepted as the next state if auniform u drawn from [0, 1] satisfies

u < min1, exp(V (θ(t)) +W (ω(t))− V (θ(t+ S∆))

−W (ω(t+ S∆)); (6.12)

otherwise the process starts again from the previous value θ−1.6 The pri-mary limitation of this Hamiltonian Monte Carlo method is that thestep size ∆ and path length S have to be specified through tuning of thesampler to the application; Neal (2011) gives an extensive discussionof methods and properties in his Section 5.4, and of alternatives toand variations on Hamiltonian Monte Carlo in his Section 5.5. TheNUTS procedure of Carpenter et al. (2017) is an extension of Hamil-tonian Monte Carlo that makes the specification of S automatic andadaptive. The key idea is to stop the leapfrog iteration (8.1) when theEuclidean distance between θ(t) and θ(t+ s∆) stops increasing at steps, a “U-turn”. To preserve time reversibility, NUTS uses an algorithmthat runs forward or backward one step, then forward or backwardtwo steps, and so on until a U-turn occurs. At that point, it samplesfrom the points collected through the iteration in a way that preservessymmetry, and applies a criterion like Equation (6.12) to determine ifthe selected point is accepted. Carpenter et al. (2017) give a detaileddiscussion of this method, including proof of its time reversibility prop-erties, recommendations and code for its implementation, a methodfor adaptively setting the step size ∆, and empirical experience withNUTS for Bayesian analysis of some standard test data. On the basisof the reported performance and our experience with this algorithm,and its availability in the open-software, R-compatible STAN statistical

6The symmetry of the Hamiltonian process requires that the sign of ω(t+ S∆)be reversed if the proposed state is accepted. However, this step is not needed unlessthis process is interwoven with other Monte Carlo methods.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 79: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

6.3. Hierarchical Bayes Methods 79

package developed by the authors and associates (Carpenter et al.,2017), we believe it to be the best general-purpose method currentlyavailable for Bayesian analysis, particularly hierarchical Bayes analysisof CBC studies. However, it is out-performed by the Allenby–Trainmethod described below in applications where normality assumptionsallow use of efficient Gibbs sampling steps. The code for an R har-ness for organizing CBC data and accessing the STAN package can bedownloaded from https://eml.berkeley.edu/~train/foundations_R.txt.This site also contains code for other data and procedures used in thismonograph.

6.3 Hierarchical Bayes Methods

In marketing, a widely used model for CBC data is a mixed (or randomcoefficients) multinomial logit model like Equation (3.8) for the choiceprobabilities, with a Bayesian prior k(θ) on the deep parameters, seeRossi et al. (2005), Bąk and Bartłomowicz (2012), Hofstede et al. (2002),and Lenk et al. (1996). Let Ln(dn|xn, pn, ζ) denote the likelihood, fromEquation (5.2), that respondent n will state a vector of choices dn fromthe offered menus. Then the posterior density on these parameters is

K(θ|x,p) ∝N∏n=1

∫ζLn(dn|xn, pn, ζ) · F (dζ|sn, θ) · k(θ). (6.13)

To illustrate the use of Equation (6.13) to produce Bayesian estimatesof its parameters, consider a specification with a fairly conventional rela-tively uninformative prior k and a density f(ζ|sn, θ) that is multivariatenormal except for a monotonic transformation (e.g., lognormal) or cen-soring in some components, as described by Train and Sonnier (2005)and summarized in the discussion around Equation (5.6). Let (σ, β)be a transformation of a commensurate latent vector ζ of normallydistributed terms with mean µ and covariance matrix Ω, so that theparameters to be estimated are θ = (µ,Ω). Alternately, write Ω = ΛΛ′,where Λ is a lower triangular matrix. In this case, θ = (µ,Λ). Forthe illustration, assume that sn does not shift the mean µ. If t is thedimension of (σ, β), then the dimension of θ is t(t+ 3)/2; this is alsothe dimension of ζ. The procedure will repeatedly use the following

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 80: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

80 Hierarchical Bayes Estimation

operation for drawing (σ, β), referred to hereafter as the taste parameterdraw: Given a trial value θ and a vector ω of i.i.d. standard normaldeviates, let ζ = µ + Ω1/2ω ≡ µ + Λω be a t × 1 multivariate normalvector. Each component of (σ, β) that is unrestricted in sign is setequal to the corresponding component of ζ; components like σ thatare necessarily positive are set to the exponent of the correspondingelement of ζ; and components that are censored to be non-negative areset to the maximum of 0 and the corresponding element of ζ. Othertransformations are of course possible. We describe two widely usedprocedures that have been applied for this specification.

Allenby–Train Procedure: (Allenby, 1997, Train, 2009, pages 299–305):7Assume that the prior k(θ) is the product of (a) multivariate normaldensity for µ with zero mean and sufficiently large variance, so that thedensity is flat from a numerical perspective, and (b) an inverted Wishartfor Ω with T degrees of freedom and parameter T IT , where IT is aT-dimensional identity matrix. Draws from the posterior distributionare obtained by iterative Gibbs sampling from three conditional poste-riors, with a Monte Carlo Markov Chain algorithm used for one of theconditional posteriors. Let µi, Ωi, and ζni be values of the parametersin iteration i, with the taste parameter draw implied by ζni. Recallthat N is the number of subjects. Gibbs sampling from the conditionalposteriors for each parameter is implemented in three layers:

µ|Ω, ζn for all n: Given Ωi and ζni for all n, the conditional posterior onµ is N

(ζi,

ΩiN

)where ζi = 1

N

∑n ζni. A new draw of µ is obtained

as µi+1 = ζi + Ψiω where Ψi is the lower-triangular Choleskyfactor of Ωi/N and ω is a draw of standard normal deviates.

Ω|µ, ζn for all n: Given µi+1 and vni for all n, the conditional posterioron Ω is Inverted Wishart with degrees of freedom T + N andparameter TI + NV , where V = 1

N (ζi − µi+1)(ζi − µi+1)′ Take7Matlab, Gauss, and R codes to estimate mixed logits in preference-space by the

Allenby–Train procedure are available on Train’s website at https://eml.berkeley.edu/~train/software.html. The codes can be readily adapted for mixed logits inWTP-space. Sawtooth Software offers codes with this procedure for mixed logitin preference-space with normally distributed coefficients and the option of signrestrictions.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 81: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

6.3. Hierarchical Bayes Methods 81

T +N draws of T -dimensional vectors of i.i.d. standard normaldeviates, labeled ηr r = 1, . . . , T +N . A new draw of Ω is obtainedas Ωi+1 = (

∑r(Γηr)(Γηr)′)−1 where Γ is the Cholesky factor of

(TI +NV )−1.

ζn|µ, Ω: Given µi+1 and Ωi+1, for taste parameters σn, βn, δn im-plied by ζn, the conditional posterior of ζn,i+1 is proportional toPn(ζn,i+1)ϕ(ζn,i+1|µi+1,Ωi+1), where ϕ denotes the multivariatenormal density with mean µi+1 and covariance matrix Ωi+1, andPn is defined following Equation (5.2). A draw of ζn,i+1 is ob-tained by the following Metropolis–Hastings procedure: Draw ωnfrom a jumping distribution, which we specify as normal withmean zero and variance mΩ, where scalar m is chosen as describedbelow. Calculate a trial value of ζn as ζn = ζni + ωn. Using thetaste parameters (σn, βn) implied by ζn, calculate the probabilityPn(ζn). If a [0, 1] uniform random variate u satisfies

u <Pn(ζn)ϕ(ζn|µi+1,Ωi+1)Pn(ζni)ϕ(ζni|µi+1,Ωi+1)

, (6.14)

then the new value of ζn,i+1 is this trial ζn; otherwise, discard thetrial ζn and retain the old value as the new value, ζn,i+1 = ζni Thevalue of m is adjusted over iterations to maintain an acceptancerate near 0.30 over n and iterations.

For a model with intra-consumer as well as inter-consumer heterogeneity,the only change in the HB procedure described is to introduce anadditional layer of multivariate normal intermediate parameters ρmnthat vary by menu as well as subject and have a density centered at ζn,with a covariance matrix adjusted in an additional Gibbs layer.

NUTS procedure: This procedure uses the NUTS sampler developedby Carpenter et al. (2017) and implemented in the R-STAN statisticalpackage. This is open-source software, with both R and STAN availablein versions that run on Linux, Windows, and Mac platforms. Thekey acceptance stage of the sampler utilizes a Metropolis–Hastingsalgorithm, and the sampler incorporates an adaptive adjustment ofHamiltonian discrete time step size, and the interactive determination

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 82: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

82 Hierarchical Bayes Estimation

of number of steps to optimize acceptance rates. The sampler is flexiblein specification of priors and likelihoods, and is not limited to themultivariate normal model described above. A drawback of this flexibilityis that STAN may run more slowly for this model than procedures thatuse computationally efficient Gibbs samplers for part of the multivariatenormal iteration. Suppose an intermediate parameter vector ζ thatis multivariate normal with mean µ and a covariance matrix Ω =DΓΓ′D, where D is diagonal and Γ is the lower triangular Choleskyfactor of the correlation matrix (so that Γ is characteristically a lowertriangular matrix with rows whose sums of squared elements are one).Recommended priors are a relatively diffuse multivariate normal priorfor µ, relatively diffuse exponential densities for the elements of D, anda LKJ(1) density for the correlation matrix ΓΓ′, see Lewandowski et al.(2009).8

Terminal Steps: Consider sampler data µi, Ωi, and vni for subjectsn = 1, . . . , N and iterations i = 1, . . . , I after discarding a few thousand“burning in” iterations; I is typically at least 10,000, and perhaps aslarge as 1 million. The Bayes estimates of the deep parameters are theposterior means µ and Ω formed by averaging µi and Ωi respectivelyover i = 1, . . . , I. A first question in applications of sampling schemeslike the one described above is whether one has achieved convergenceto a stationary distribution that spans θ-space, with sufficiently strongmixing so that serial correlations between distant iterations are smalland sampler means are accurate estimates of true posterior means.The recommendations of the developers of these samplers are to startthem from alternative parameter values, vary burn-in times and tuningparameters, and jitter tuning parameters to avoid accidental cyclicity.In the special case of test data with known deep parameters, onecan examine bias and mean square error as a function of sampler

8All correlation matrices can be written as ΓΓ′, where Γ is a lower triangularmatrix with the property that the squares of the elements in each row sum to one.Then the elements (Γi1, . . . ,Γii) are points in a unit sphere of dimension i, and themethod of Muller (1959) and Marsaglia (1972) can be used to draw points uniformlyfrom this sphere, i.e., draw i i.i.d. standard normal deviates, and scale the vector ofthese deviates by dividing by the square root of their sum of squares.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 83: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

6.4. A Monte Carlo example 83

characteristics.9 Econometric time series diagnostic tests for stationarityand for mixing conditions can be applied more generally to the sampleroutput, see Azizzadeh and Rezakhah (2014), Delgado (2005), Domowitzand El-Gamal (1993), Corradi et al. (2000), and Ljung and Box (1978).

From a classical perspective, the posterior means µ and Ω are pointestimates of true parameter values µ0 and Ω0. Suppose the analystwishes to treat these as classical estimators and carry out conventionalsteps such as associating standard errors with the estimates and pro-viding confidence bounds or hypothesis tests. Due to the asymptoticequivalence of MLE and hierarchical Bayes estimators, this can be doneusing the formulas (5.2)–(5.7). To improve the asymptotic approxi-mation, we recommend executing one or more BHHH steps startingfrom the HB posterior means, using the result from Equation (5.4) toobtain MLE-equivalent estimates, and using Equation (5.5) to estimatethe covariance matrix of the resulting estimators. In most cases, it ispractical to draw new intermediate parameter vectors ζr for Equation(5.7), but it is also possible to reuse the array of vectors ζni from iter-ations in the hierarchical Bayes process in the calculation of Equations(5.2)–(5.6).

6.4 A Monte Carlo example

To demonstrate the procedures given in the previous two sections,consider a Monte Carlo example where a known decision-making processis used to generate an artificial CBC data set, and the methods aboveare then applied to estimate the parameters behind these “observed”choices. For this exercise, suppose we have CBC data on preferencesfor table grapes, the example discussed earlier in Section 2.6. Suppose1,000 subjects are each presented with eight menus, and on each menuthey have the option of choosing one of three bunches of grapes with

9In test data where the ζ for each subject n are drawn from a known normaldistribution N(µ,Ω), and choice data are then generated from a random utilitymodel with these taste parameters, the mean and covariance matrix used for thesecomparisons should be µ0 = 1

N

∑N

n=1 vn and Ω0 = 1N

∑N

n=1(vn − µ0)(vn − µ0)′rather than the underlying µ and Ω; as the target of the analysis is recovery ofcharacteristics of realized tastes, and the analysis cannot identify the process thatgenerated the realized tastes.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 84: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

84 Hierarchical Bayes Estimation

Table 6.1: Table grape CBC attributes and levels.

Attribute Symbol LevelsPrice P $1.00 to $4.00Sweetness S Sweet (1) or Tart (0)Crispness C Crisp (1) or Soft (0)Size L Large (1) or Small (0)Organic O Organic (1) or Non-organic (0)

various attributes and prices, or choosing to buy no grapes on thisoccasion. Assume that the subjects are prompted to think of this asthe opportunities they will face on their next trip to the supermarket,so that existing stocks of grapes and purchasing opportunities outsidethe experiment are similar to those faced by these subjects in the realmarket. The product attributes and levels in the experiment are givenin Table 6.1.

The gender (G) or each subject is observed, coded G = 1 for femaleand G = 0 for male. Let In denote disposable income, which is notmeasured, but which drops out of the utility maximization. Define theinteractions SC ≡ S∗C and SG ≡ S∗G. Index a menu’s alternativesby j = 1, 2, 3 for the three offered bunches, and j = 4 for the nopurchase option. Let Bjmn be a dummy variable that is one for purchasealternatives (j = 1, 2, 3) and zero for the no-purchase alternative j = 4.Assume the true utility has the WTP form

Ujmn ≡ In − Pjmn + SjmnβSn + CjmnβCn + LjmnβLn +OjmnβOn

+SCjmnβSCn + SGjmnβSGn +Bjmnβqn + σnεjn, (6.15)

where the εjn are i.i.d. standard EV1 distributed, and ζn = (βSn,βCn, βLn, βOn, βSCn, βSGn, βBn, log(σn)) are the preference param-eters. Define Djmn = 1 for the alternative within a menu that max-imizes utility, Djmn = 0 otherwise. Assume that the true parame-ters are distributed multivariate normal in the population with themeans, standard deviations, and correlation/covariance matrix givenin Table 6.2; R code that generates the data is available at https://eml.berkeley.edu/~train/foundations_R.txt.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 85: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

6.4. A Monte Carlo example 85

Tab

le6.

2:Tr

uepa

rameter

distrib

utionβS,β

C,β

L,β

O,β

SC,β

SG,β

qn,l

og(αn).

Mean

Std.

Dev.

Covariance(bottom)an

dcorrelation(top

)matric

esβS

βC

βL

βO

βSC

βSG

δ nlog(σn)

βS

1.0

0.3

0.09

0.6

00

0.3

00

0βC

0.3

0.1

0.018

0.01

0.48

00.42

00

0βL

0.2

0.1

00.0048

0.01

00.42

00

0βO

0.1

0.2

00

00.04

0.3

00

0βSC

0.0

0.05

0.0045

0.0021

0.0021

0.003

0.0025

00

0βSG

0.1

0.2

00

00

00.04

00

βBn

2.0

1.0

00

00

00

10

log(σn)−0.5

0.3

00

00

00

00.09

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 86: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

86 Hierarchical Bayes Estimation

Model (6.15) is estimated with independence of log(σn) and theremaining parameters imposed, but all other covariances allowed.Table 6.3 give estimates by hierarchical Bayes using the Allenby–Trainprocedure (HB-AT), hierarchical Bayes using the NUTS procedure(HB-NUTS), and maximum simulated likelihood (MSL). Note that thetrue population parameter values given in Table 6.1 are reproducedclosely but not exactly by their sample counterparts for the 1,000subjects in the CBC study; these are given in the third column ofthe table, and are the relevant values for comparison with estimates.The maximum simulated likelihood estimates use 250 Halton draws toapproximate the terms in Equations (5.5) and (5.6), and use Newton–Raphson iteration to the estimator. Starting values were obtainedby estimating a model with fixed coefficients, using the estimatedWTPs as starting values in a model with random but uncorrelatedWTPs, and then using these estimates as starting values in the modelwith correlated WTPs. The three steps combined took about threehours of run time. The HB-AT estimates are obtained with 100,000iterations used to tune and burn-in the sampler, and every tenthdraw taken from 100,000 subsequent draws to form the posteriormeans given in the table.10 It has been observed that this proceduretends to inflate variances that are numerically very small (see, e.g.,Balcombe and Fraser, 2009); to counteract this tendency, we divided thenonprice variables by ten and then rescaled the estimates accordingly.Estimation took 12 minutes. The HB-NUTS estimates are obtainedfrom 11,000 iterations, with the first 5,500 discarded for “burn-in”, andevery 10th draw from the remainder used to obtain the estimates inTable 6.3.

For the NUTS sampler, we also calculated posterior means (notreported) using all iterations including burn-in iterations, using everyiteration rather than every 10th iteration, and using every 100th itera-tion including burn-in iterations. The results were very close, indicatingthat in this application NUTS seems to show rapid convergence withrelatively little drift. However, time-series graphs do show some long

10We set T , the degrees of freedom for the inverted Wishart prior, equal to thenumber of random coefficients in the model. We have found little numerical effect ofusing a larger degrees of freedom.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 87: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

6.4. A Monte Carlo example 87

Tab

le6.

3:Estim

ates

from

tablegrap

esCBC

data.

P anelA

Parameter

True

values

HB-AT

HB-N

UTS

MSL

Popu

latio

nSa

mple

Mean

Std.

Dev.

Mean

Std.

Dev.

Estim

ate

Std.

Error

Means

βS

1.0

1.00

50.

9682

(0.0

445)

0.97

84(0.0

407)

0.96

94(0.0

457)

βC

0.3

0.29

90.

3495

(0.0

358)

0.36

10(0.0

342)

0.35

54(0.0

391)

βL

0.2

0.19

70.

1781

(0.0

251)

0.18

98(0.0

258)

0.17

60(0.0

238)

βO

0.1

0.10

90.

0910

(0.0

238)

0.09

64(0.0

224)

0.09

988

(0.0

248)

βSC

0.0

0.00

1−

0.04

50(0.0

485)−

0.05

15(0.0

464)

−0.

0579

(0.0

490)

βSG

0.1

0.09

30.

1369

(0.0

540)

0.14

46(0.0

469)

0.15

23(0.0

536)

βBn

2.0

2.01

22.

0049

(0.0

498)

2.01

59(0.0

583)

1.98

93(0.5

12)

log(σn)

−0.

5−

0.48

1−

0.52

77(0.0

253)−

0.50

38(0.0

228)

−0.

5687

(0.2

81)

(Con

tinued)

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 88: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

88 Hierarchical Bayes Estimation

Tab

le6.

3:(C

ontin

ued)

P anelB

Parameter

True

values

HB-AT

HB-N

UTS

MSL

Popu

latio

nSa

mple

Mean

Std.

Dev.

Mean

Std.

Dev.

Estim

ate

Std.

Error

Std.

Devs.

βS

0.3

0.30

30.

3086

(0.0

633)

0.32

99(0.0

610)

0.30

44(0.8

889)

βC

0.1

0.09

90.

2149

(0.0

354)

0.12

99(0.0

539)

0.26

61(0.1

085)

βL

0.1

0.10

10.

1621

(0.0

370)

0.13

25(0.0

497)

0.14

96(0.0

517)

βO

0.2

0.21

00.

2157

(0.0

378)

0.18

63(0.0

489)

0.24

40(0.0

467)

βSC

0.05

0.05

10.

2227

(0.0

516)

0.09

74(0.0

738)

0.25

64(0.1

511)

βSG

0.2

0.20

20.

2518

(0.0

758)

0.29

60(0.1

325)

0.39

55(0.1

084)

BBn

1.0

0.98

80.

7924

(0.0

764)

0.92

88(0.0

548)

0.89

17(0.0

606)

log(σn)

0.3

0.29

60.

4220

(0.0

257)

0.27

57(0.0

415)

0.34

06(0.0

415)

(Con

tinued)

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 89: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

6.4. A Monte Carlo example 89T

able

6.3:

(Con

tinue

d)

P ane

lCP a

rameter

True

values

HB-

ATHB-

NUTS

MSL

P opu

latio

nSa

mple

Mean

Std.

Dev

Mean

Std.

Dev

Estim

ate

Std.

Error

Co varianc

esβS,βC

0.01

80.

0185

0.03

480.

0183

0.00

44(0.0

119)

0.05

47(0.0

375)

βS,βL

00.

0007−

0.00

540.

0166

−0.

0020

(0.0

125)

−0.

0038

(0.0

169)

βS,βO

0−

0.00

010.

0077

0.01

690.

0012

(0.0

148)

0.00

03(0.0

176)

βS,βSC

0.00

450.

0046−

0.02

440.

0237

0.00

00(0.0

151)

−0.

0286

7(0.0

449)

βS,βSG

0−

0.00

040.

0002

0.02

30−

0.01

50(0.0

439)

−0.

0152

(0.0

411)

βS,βB

0−

0.00

320.

1430

0.05

660.

0289

(0.0

432)

0.10

28(0.0

453)

βC,βL

0.00

480.

0050

0.00

670.

0095

0.00

26(0.0

052)

0.01

65(0.0

190)

βC,βO

00.

0006

0.01

610.

0119

0.00

120.

0079

0.01

08(0.0

193)

βC,βSC

0.00

210.

0022−

0.02

910.

0154

−0.

0030

(0.0

061)

−0.

0579

(0.0

639)

βC,βSG

00.

0000−

0.01

290.

0177

0.00

99(0.0

172)

−0.

0038

(0.0

449)

βC,βB

0−

0.00

270.

1073

0.04

020.

0112

(0.0

378)

0.04

39(0.0

533)

βL,βO

00.

0009

0.01

450.

0114

0.01

21(0.0

093)

0.03

15(0.1

43)

βL,βSC

0.00

210.

0023−

0.00

520.

0116

0.00

02(0.0

052)

−0.

0083

(0.0

210)

βL,βSG

00.

0001−

0.01

050.

0148

−0.

0043

(0.0

151)

−0.

0001

(0.0

217)

βL,βB

0−

0.00

40−

0.00

130.

0332

−0.

0094

(0.0

253)

−0.

0169

(0.0

236)

βO,βSC

0.00

300.

0032−

0.01

420.

0131

0.00

060.

0084

0.00

50(0.0

240)

βO,βSG

0−

0.00

16−

0.01

530.

0202

−0.

0076

(0.0

182)

−0.

0213

(0.0

263)

βO,βB

0−

0.00

020.

0492

0.03

310.

0396

(0.0

268)

0.00

76(0.0

326)

βSC,βSG

0−

0.00

020.

0212

0.01

940.

0027

(0.0

140)

0.01

60(0.0

471)

βSC,βB

0−

0.00

20−

0.10

830.

0513

−0.

0045

(0.0

290)

−0.

0420

(0.0

668)

βSG,βB

00.

0008−

0.04

170.

0615

−0.

0145

(0.0

508)

−0.

0869

(0.0

747)

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 90: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

90 Hierarchical Bayes Estimation

cycles in the parameter estimates. The NUTS procedure required com-putation overnight; this may be due to storage of posterior intermediateparameters for each subject in each iteration, output that may be usefulfor some marketing applications such as segmentation, but is not neededfor the primary task of estimating the deep parameters. Comparing theHB-AT, HB-NUTS, and MSL estimates with the true parameter valuesand with each other, we find that all the methods do a good job ofrecovering the true utility model.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 91: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

7An Empirical CBC Study Using MSL and HB

Methods

In this section, we provide an example of a money-metric randomutility model estimated using both Maximum Simulated Likelihoodand Hierarchical Bayes methods. We utilize the data from a conjointexperiment designed and described by Glasgow and Butler (2017) forconsumers’ choice among video streaming services. Their experimentsincluded the monthly price of the service and the following non-priceattributes:

Each choice experiment included four alternative video steamingservices with specified price and attributes plus a fifth alternative ofnot subscribing to any video streaming service. Each respondent waspresented with 11 choice experiments. Glasgow and Butler obtainedchoice from 300 respondents and implemented an estimation procedurethat accounts for “protestors” (mainly consumers who never chose aservice that shared data). In our use of their data, we do not include the40 respondents that they identified as protestors, so that our sampleconsists of 260 respondents.

Table 7.2 gives a model estimated in WTP-space by hierarchicalBayes, using the Allenby–Train procedure with 10,000 burn-in iterations,10,000 iterations after burn-in from which, to reduce serial correlation,

91

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 92: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

92 An Empirical CBC Study Using MSL and HB Methods

every 10th draw from the sampler was retained to calculate the estimates.WTP for each non-price attribute is assumed to be normally distributedover consumers; the price/scaling coefficient (1/α) is assumed to belog-normally distributed; and correlation is allowed among the WTPs aswell as the price/scaling coefficient. The point estimate of the populationmean in the mean of the draws from the posterior distribution, and thestandard error of this estimate is the standard deviation of the draws.Similarly for the standard deviation in the population, the estimateis the mean of the draws from its posterior and the standard erroris the standard deviation of these draws. The point estimates of thecorrelations are shown in the second part of the table. The results inTable 7.1 indicate that people are willing to pay $1.57 per month onaverage to avoid commercials. Fast availability is valued highly, with an

Table 7.1: Non-price attributes.

Attribute LevelsCommercials shownbetween content

Yes (“commercials”)No (baseline category)

Speed of contentavailability

TV episodes next day, movies in 3 months(“fast content”)

TV episodes in 3 months, movies in 6 months(baseline category)

Catalogue 10,000 movies and 5,000 TV episodes(“more content”)

2,000 movies and 13,000 TV episodes(“more TV/fewer movies”)

5,000 movies and 2,500 TV episodes(the baseline category)

Data-sharingpolicies

Information is collected but not shared (baselinecategory)

Usage information is shared with third parties(“share usage”)a

Usage and personal information are shared withthird parties (“share usage and personal”)

Note: aGlasgow and Butler use the terms “non-personally identifiable information (NPPI)” and“personally identifiable information (PII)” for what we are labelling “share usage” and “shareusage and personal”.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 93: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

93

average WTP of $3.50 per month in order to see TV shows and moviessoon after their original showing. On average, people prefer having a mixwith more TV shows and fewer movies, but the mean is not significant.Average willingness to pay for more content of both kinds is $2.57 permonth. Interestingly, people who want fast availability tend to be thosewho prefer more TV shows and fewer movies: the correlation betweenthese two WTPs is 0.51, while the correlation between WTP for fastavailability and more content of both kinds is only 0.04. Apparently,the desire for fast availability mainly applies to TV shows.

Consider now the data-sharing policies of services. The point esti-mate implies that consumers have a WTP of 43 cents per month toavoid having their usage data shared in aggregate form; however, thehypothesis of zero average WTP cannot be rejected. Consumers aremuch more concerned about their personal information being sharedalong with their usage information: the average WTP to avoid suchsharing is $1.83 per month. The correlation between WTP to avoid thetwo forms of sharing is a substantial 0.82.

Table 7.3 gives the same model estimated by the method of maximumsimulated likelihood (MSL). We used STATA’s module for mixed logitin wtp-space, “mixlogitwtp.” We specified 100 Halton draws per person,which has been shown to be more accurate for mixed logit estimationthan 1,000 pseudo-random draws per person (Bhat, 2001; Train, 2000;Train, 2009). We used the HB estimates (Table 6.3) as starting values. Atthe HB estimates, the log-likelihood was −3994.64 and rose to −3903.47at convergence.1

The MSL estimates for the mean WTPs are fairly similar to the HBestimates. In particular, the MSL estimates of mean WTP to: avoidcommercials is $1.56 compared to $1.57 by HB; obtain fast availability is$3.95 compared to $3.50 by HB; obtain more content is $2.96 comparedto $2.57 by HB; to avoid having no service is $27.26 compared to $28.59

1The mixlogitwtp module allows estimation by Newton–Raphson (NR), BHHH,DFP, and BFGS. Using NR, one iteration raised the log-likelihood from −3994.65 to−3969.31. The additional rise to convergence at −3903.47 took 29 iterations. Runtime was about 90 minutes. The NR option utilizes numerical gradients and Hessian.The other procedures utilize numerical gradients but alternative forms of the Hessianand can be expected to run faster.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 94: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

94 An Empirical CBC Study Using MSL and HB MethodsT

able

7.2

PanelA

:HierarchicalB

ayes

mod

elof

WTPs

forvide

ostream

ingservice

Popu

latio

nMean

Std.

Dev.inPo

pulatio

nEs

timate

Std.

Error

Estim

ate

Std.

Error

Log(1/σ

)−

1.89

00.

1188

1.40

10.

2497

WTP

for:

Com

mercials

−1.

571

0.55

473.

4882

0.42

06Fa

stavailability

3.49

60.

7182

3.65

170.

4189

MoreTV/few

ermovies

0.22

70.

5237

3.67

060.

3602

Morecontent

2.56

90.

5170

2.65

930.

3532

Shareusag

eon

ly−

0.43

20.

3015

1.95

040.

3170

Sharepe

rson

alan

dusag

e−

1.83

10.

6325

4.71

890.

6012

Noservice

−28.5

913.

2753

20.7

522.

9424

Pane

lB:C

orrelatio

ns:P

oint

estim

ates

Log(1/σ

)Com

mercials

Fast

availability

MoreTV,

fewer

mov

ies

More

content

Share

usag

e

Share

person

alan

dusag

eNo

service

Log(1/σ

)1.

0000

−0.

3010

−0.

1484

−0.

1414

0.00

72−

0.00

50−

0.14

310.

3865

Com

mercials

1.00

000.

2238

0.14

93−

0.19

12−

0.07

99−

0.02

04−

0.36

39Fa

stavailability

1.00

000.

5094

0.04

18−

0.51

26−

0.51

350.

2197

MoreTV,few

ermovies

1.00

00−

0.45

23−

0.11

51−

0.27

870.

3266

Morecontent

1.00

00−

0.00

570.

1650

0.18

84Sh

areusage

1.00

000.

8155

−0.

2866

Sharepe

rson

alan

dusage

1.00

00−

0.45

35

Noservice

1.00

00

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 95: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

95T

able

7.3

PanelA

:MSL

Estim

ates

ofW

TPs

forVideo

Stream

ingSe

rvices

Popu

latio

nMean

Std.

Dev.inPo

pulatio

nEs

timate

Std.

Error

Estim

ate

Std.

Error

Log(1/σ

)−

2.00

20.

0.94

51.

0637

0.07

55W

TP

for:

Com

mercials

−1.

562

0.42

143.

940

0.53

02Fa

stAv

ailability

3.94

50.

4767

3.63

10.

4138

MoreTV,few

ermovies

−0.

6988

0.47

834.

857

0.55

41Morecontent

2.96

30.

4708

2.52

40.

4434

Shareusag

eon

ly−

0.62

240.

4040

2.49

40.

4164

Sharepe

rson

alan

dusag

e−

2.70

50.

5844

6.75

10.

7166

Noservice

−27.2

62.

662

19.4

22.

333

PanelB

:Correlatio

ns:P

oint

Estim

ates

(*deno

tessig

nific

ance

at5%

level)

Com

mercials

Fast

Availability

Mostly

TV

Mostly

movies

Share

usag

e

Share

person

alan

dusag

eNo

service

Log(1/σ

)−

0.58

13∗

−0.

1371

0.03

580.

0256

0.00

22−

0.12

870.

2801

Com

mercials

1.00

000.

1172

−0.

3473

∗0.

0109

−0.

2562

−0.

0079

−0.

4108

Fast

Availability

1.00

000.

8042

∗−

0.40

19∗−

0.35

42∗−

0.42

06∗

0.23

91∗

Mostly

TV

1.00

00−

0.58

90∗−

0.16

95−

0.33

28∗

0.46

16∗

Mostly

movies

1.00

000.

5141

∗0.

5181

∗−

0.01

47Sh

areusage

1.00

000.

9370

∗−

0.05

63Sh

arepe

rson

alan

dusag

e1.

0000

−0.

0975

Noservice

1.00

00

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 96: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

96 An Empirical CBC Study Using MSL and HB Methods

by HB. The mean WTPs for more TV shows/fewer movies and toavoid sharing of usage data are insignificant under both MSL and HB.The only substantial difference is the mean WTP to avoid sharing ofpersonal and usage information, which is estimated to be $2.71 by MSLand $1.83 by HB.

The estimated standard deviations of WTP are practically the sameby MSL and HB for the following attributes: to avoid commercials,obtain fast service, obtain more content, and avoid not having service.Standard deviations of WTP for more TV/fewer movies and to avoideither form of data sharing are estimated to be higher by MSL thanHB. The correlations that were discussed above in relation to the HBestimates are also evidenced with MSL. However, other correlationsdiffer in magnitude and even sign between MSL and HB.

A preference-space model without the money-metric scaling was alsoestimated with the price coefficient assumed log-normally distributedand the non-price coefficients assumed normal (not shown.)2 In general,the estimated means and standard deviations of WTP were considerablyhigher and less plausible. For example, for sharing of usage and personaldata, the model in WTP-space above gives a mean WTP to avoidsharing of $2.71 with a standard deviation of $6.75. For the model inpreference space, the derived distribution of WTP has a mean $6.99and standard deviation of $98.95. The high values, especially for thestandard deviation, are due to the fact that WTP in the preference spacemodel is calculated as the ratio of the attribute’s coefficient divided bythe price coefficient, such that values of the price coefficient that arearbitrarily close to zero (which are allowed by the distribution of theprice coefficient) imply arbitrarily large WTP. Models in WTP-spaceavoid this issue by specifying and estimating the distribution of WTPdirectly. However, the log-likelihood of the model in preference space ishigher than that in wtp-space: −3863.9 compared to −3903.5. Theseresults, namely, more plausible estimates of the distribution of WTPaccompanied by a lower log-likelihood, mirror those of Train and Weeks(2005). Other studies making this comparison (e.g., Scarpa et al., 2008)

2As noted in section 3.1, a model in preference-space with the assumption of alog normal price coefficient and normal non-price coefficients is not equivalent to aWTP-space model with normally distributed WTP for non-price attributes.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 97: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

97

have found no such tradeoff on their datasets, with their model inWTP-space obtaining a better fit as well as more plausible distributionsof WTP.

As discussed above, the means and standard deviations of the WTPestimates obtained by MSL and HB-AT methods generally agree onsign and magnitude, but are not close enough to conclude that thesemethods have converged to asymptotically equivalent estimates. Thismay reflect finite sampling deviations from the asymptotic limit; thereis no guarantee that the hierarchical Bayes and MSL estimates will beclose in finite samples. However, both methods use simulations, and thehierarchical Bayes methods in particular are known to converge slowlyto their posterior limiting distribution when their tuning is imperfect.Then, some of the differences in Tables 7.2 and 7.3 might disappearwith more complete tuning and more simulation draws.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 98: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

8An Application with Inter and Intra-consumer

Heterogeneity

This section extends the Hierarchical Bayes method to models withinter- and intra-consumer heterogeneity. These models account for intra-and inter-consumer heterogeneity with three levels of parameters:

1. Population-level parameters µ and Ωb: average tastes/ preferencesin the population and the inter-consumer variance–covariancematrix, respectively.

2. Individual-level parameters ζn and Ωw: average tastes/preferencesof a specific individual and the intra-consumer variance–covariancematrix, respectively.

3. Menu-level parameters ρmn: to reflect menu-specific (choice spe-cific) taste perturbations.

Such models have been previously estimated by Hess and Train (2011),and Hess and Rose (2009) using a maximum simulated likelihood esti-mator. Below we describe an extension of the Allenby–Train estimatorthat is based on the same Normality assumptions. For more details, see

98

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 99: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

99

Becker et al. (2018). The extended HB procedure includes the followingfive Gibbs sampling layers.

Five-Step Gibbs Sampling

µ|Ωb, ζn, Ωw, ρmn: Given Ωbi and ζni for all n, the conditional posterior

on µ is N(ζi,

ΩbiN

)where ζı = 1

N

∑n ζni. A new draw of µ is

obtained as µi+1 = ζi + Ψiω where Ψi is the Cholesky factor ofΩbi/N and ω is a draw of the multivariate standard normal.

Ωb|µ, ζn, Ωw, ρmn: Given µi+1 and ζni for all n, the conditional pos-terior on Ωb is Inverted Wishart with degrees of freedom T andparameter TI +NV b, where T is the number of unknown param-eters and V b = 1

N (ζni − µi+1)(ζni − µi+1)′. Take T + N drawsof T -dimensional vectors of i.i.d. standard normal deviates, la-beled υr, r = 1, . . . , T + N . A new draw of Ωb is obtained asΩbi+1 = (

∑r(Γbυr)(Γbυr)′)−1 where Γb is the Cholesky factor of

(TI +NV b)−1.

Ωw|µ, ζn, Ωb, ρmn: Given ρmn,i and ζni for all n, the conditional poste-rior on Ωw is Inverted Wishart with degrees of freedom T+MT andparameter TI +MT V

w, where MT represents the total numberof menus faced by all individuals, and V w = 1

MT

∑Mm=1(ρmn,i −

ζi)(ρmn,i − ζi)′. Take T +MT draws of T -dimensional vectors ofi.i.d. standard normal deviates, labeled υs, s = 1, . . . , T +MT . Anew draw of Ωw is obtained as Ωw

i+1 = (∑s(Γwυs)(Γυs)′)−1 where

Γw is the Cholesky factor of (TI +MT Vw)−1. Here we assume a

single variance–covariance matrix for all individuals; due to thesmall number of choice situations faced by each individual in atypical stated preference survey, it is not possible to estimate avariance–covariance matrix for each individual.

ζn|Ωb, µ, Ωw, ρmn: Given µi, ρmni and Ωbi for all n, the conditional

posterior on ζn with N(µi,Ωbi) as a prior is N(ζn,Σζn) where ζn =

([Ωbi+1]−1 +M [Ωw

i+1]−1)−1([Ωbi+1]−1µi+M [Ωw

i+1]−1 1M

∑Mm=1 ρmni),

and Σζn = ([Ωbi+1]−1 +M [Ωw

i+1]−1)−1. A new draw of µ is obtained

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 100: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

100 An Application with Inter and Intra-consumer Heterogeneity

as µi+1 = ζn + Ψiω where Ψi is the Cholesky factor of Σζn and ωis a draw of the multivariate standard normal.

ρmn|µ, ζn, Ωb, Ωw: Given ζi+1 and Ωwi+1, the conditional posterior of

ρmn,i+1 is proportional to Pmn(ηmn,i+1)ϕ(ηmn,i+1|ζn,i+1,Ωwi+1),

where ϕ denotes the multivariate normal density with mean ζn,i+1and covariance matrix Ωw

i+1, and Pmn(η) =

∏Jj=0

exp(xjmnρmn−pjmn

σ(ρmn)

)∑J

i=0 exp(ximnρmn−pimn

σ(ρmn)

)djmn . A draw of ρmn,i+1 is ob-

tained by the following Metropolis–Hastings procedure: Draw ωnfrom a jumping distribution, which we specify as normal withmean zero and variance mΩw, where scalar m is adjusted overiterations to maintain an acceptance rate of 0.3. Calculate a trialvalue of ρmn,i+1 as ρmn = ρmn,i + ωn. Calculate the probabilityPmn(ρmn). If a [0, 1] uniform random variate u satisfies

u <Pmn(ρmn)ϕ(ρmn|ζn,i+1,Ωw

i+1)Pmn(ρmn,i)ϕ(ρmn,i|ζn,i+1,Ωw

i+1) , (8.1)

then the new value of ρmn,i+1 is this trial ρmn; otherwise, discardthe trial ρmn and retain the old value as the new value, ρmn,i+1 =ρmn,i.

The description of the estimator above assumes that all coefficientshave inter- and intra-consumer distributions. However, the estimatoralso includes coefficients with only inter-consumer heterogeneity, orcoefficients without any heterogeneity, as follows:

1. If some of the coefficients do not vary across subjects or menus, theprocedure described above is only applied to coefficients with inter-consumer heterogeneity, while the sampling of these coefficients isobtained by an additional Metropolis–Hastings procedure.

2. Similarly, if some of the coefficients do not vary across menus, theprocedure described above is only applied to coefficients with intra-consumer heterogeneity, and an additional Metropolis–Hastingsprocedure is used to sample inter only coefficients.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 101: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

101

Similarly, the estimator is also able to impose zero covariance restrictionson Ωb and Ωw. If these matrices are specified to be block-diagonal, thensampling is performed by drawing from the different blocks separatelyusing the steps specified above. For more detail see Becker et al. (2018).

A Monte Carlo ExampleTo demonstrate this HB procedure consider again the Monte Carloexample from Section 6.4. Intra-consumer heterogeneity is introducedto the table grape CBC data for the coefficients of “size” and “organic”.Data is generated using the procedure sourced in part C of the codesat https://eml.berkeley.edu/~train/foundations_R.txt. In addition, anintra-consumer covariance matrix is obtained by defining a Choleskyfactor of the correlation matrix, and intra-consumer standard deviations.The data generation process is presented in Figure 8.1. The correlationbetween the coefficients of “size” and “organic” (βL and βO) is specifiedto be −0.3. Additionally, three levels of intra-consumer heterogeneityare considered as presented in Table 8.1.

Population means and inter-consumer covariances are the sameas those presented in Table 6.2. The values in the samples for intra-consumer heterogeneity are presented in Tables 8.2. The correspondingtrue values are presented in the Appendix.

Figure 8.1: Data Generation Process — Inputs in White Boxes and Outputs inShaded Boxes

Table 8.1: Levels of intra-consumer heterogeneity.

Heterogeneity level βL Std. Dev. βO Std. Dev.Low 0.2 0.1Medium 0.6 0.3High 1 0.5

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 102: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

102 An Application with Inter and Intra-consumer Heterogeneity

Table 8.2

Std. Devs. Covariances and Correlation in top right cornerβL βO

Panel A: Low heterogeneity — values in sample.βL 0.201 0.040 −0.291βO 0.099 −0.006 0.010

Panel B: Medium heterogeneity — values in sampleβL 0.602 0.362 −0.291βO 0.297 −0.052 0.088

Panel C: High heterogeneity — values in sampleβL 1.003 1.006 −0.291βO 0.496 −0.145 0.246

Model Estimation

The utility equation for each menu is similar to that presented inEquation (9.4), but with menu-specific coefficients βLmn and βOmn:

Ujmn ≡ In − Pjmn + SjmnβSn + CjmnβCn + LjmnβLmn +OjmnβOmn

+SCjmnβSCn + SGjmnβSGn +Bjmnβqn + σnεjmn (8.2)

The model is estimated using the five-step procedure described above.After 100,000 burn-in iterations, every tenth draw is taken from 100,000subsequent draws for the estimation of the parameters. The estimationresults are presented in Table 8.3. Models with only inter-consumerheterogeneity were also estimated using data sets with intra-consumerheterogeneity. The results are presented in the Appendix.

Table 8.4 presents the differences in the mean and standard devia-tion of the scale parameter Log(σn) between models estimated with andwithout intra-consumer heterogeneity (using the data generated for thecases of low heterogeneity, medium heterogeneity, and high heterogene-ity). The results indicate that ignoring intra-consumer heterogeneityresults in smaller means and standard deviations of the scale parametercompared to the cases where intra-consumer heterogeneity is accounted

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 103: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

103T

able

8.3:

Estim

ates

from

tablegrap

esCBC

data

with

intra-consum

erhe

terogene

ity.

P arameter

True

values

No

heterogene

ityLo

whe

terogene

ityMed

ium

heterogene

ityHigh

heterogeneity

Means

Popu

latio

nSa

mple

Mean

Std.

Dev.

Mean

Std.

Dev.

Mean

Std.

Dev.

Mean

Std.

Dev.

Pane

lAβ

S1

1.00

50.

973

0.03

60.

995

0.04

91.

027

0.05

81.

002

0.04

C0.

30.

299

0.35

80.

033

0.28

60.

036

0.29

10.

032

0.28

70.

043

βL

0.2

0.19

70.

181

0.02

40.

213

0.02

50.

223

0.02

70.

238

0.03

O0.

10.

109

0.09

40.

027

0.12

00.

024

0.14

60.

026

0.12

80.

028

βS

C0

0.00

1−

0.05

90.

039

0.07

50.

046

0.05

50.

042

0.04

60.

048

βS

G0.

10.

093

0.14

40.

046

0.05

30.

051

0.04

40.

074

0.07

70.

045

βB

N2

2.01

22.

003

0.05

12.

009

0.05

31.

969

0.05

01.

981

0.05

4Lo

g(σ

n)

−0.

5−

0.48

1−

0.52

60.

024−

0.48

60.

025−

0.48

60.

027−

0.47

80.

033

Pane

lBInter-consum

erStd.

Devs.

βS

0.3

0.30

40.

268

0.05

80.

200

0.04

40.

213

0.04

60.

227

0.05

C0.

10.

099

0.16

70.

037

0.19

00.

057

0.15

70.

043

0.19

40.

055

βL

0.1

0.10

10.

137

0.03

30.

165

0.03

90.

168

0.04

50.

171

0.04

O0.

20.

210

0.19

50.

034

0.15

90.

041

0.20

50.

043

0.19

20.

044

βS

C0.

050.

051

0.17

10.

049

0.22

10.

056

0.19

10.

054

0.23

50.

074

βS

G0.

20.

202

0.20

20.

059

0.19

90.

066

0.20

80.

057

0.21

30.

069

βB

N1

0.98

80.

844

0.05

90.

922

0.07

60.

887

0.06

60.

852

0.07

2Lo

g(σ

n)

0.3

0.29

60.

327

0.03

20.

285

0.03

10.

287

0.03

30.

301

0.03

7(C

ontin

ued)

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 104: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

104 An Application with Inter and Intra-consumer Heterogeneity

Tab

le8.

3:(C

ontin

ued)

Noheterogeneity

Low

heterogene

ityMedium

heterogeneity

Highheterogeneity

Intra-consum

erStd.

Devs.

and

covaria

nce

True

value

Mean

Std.

Dev.

True

value

Mean

Std.

Dev.

True

value

Mean

Std.

Dev.

True

value

Mean

Std.

Dev.

Pane

lCβ

Lstd.

dev.

00.

186

0.04

50.

20.

213

0.09

60.

60.

681

0.07

21

1.06

30.

068

βO

std.

dev.

00.

128

0.05

60.

10.

135

0.08

20.

30.

206

0.09

50.

50.

382

0.12

L,β

OCov

00.

010

0.01

2−

0.00

60.

002

0.01

7−

0.05

4−

0.06

60.

051−

0.15

−0.

106

0.07

0(C

ontin

ued)

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 105: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

105T

able

8.3:

(Con

tinue

d)

True

values

Nohe

terogene

ityLo

whe

terogene

ityMed

ium

heterogene

ityHighhe

terogene

ity

Covarianc

esPo

pulatio

nSa

mple

Mean

Std.

Dev.

Covarianc

eStd.

Dev.

Covarianc

eStd.

Dev.

Covarianc

eStd.

Dev.

Pane

lDβS,β

C1.8

0.01

850.

0141

0.18

240.

0978

0.11

670.

0095

0.14

740.

0123

0.19

55βS,β

L0

0.00

07−

0.00

600.

0130

0.00

010.

0153

0.00

890.

0145

0.00

800.

0155

βS,β

O0

−0.

0002−

0.00

090.

0204

0.00

550.

0093

0.00

640.

0152

0.01

270.

0158

βS,β

SC

0.45

0.00

460.

0016

0.01

91−

0.00

270.

0157

−0.

0027

0.01

46−

0.00

910.

0248

βS,β

SG

0−

0.00

040.

0124

0.01

810.

0059

0.01

810.

0105

0.01

490.

0020

0.01

68βS,β

BN

0−

0.00

320.

0975

0.04

370.

0623

0.05

070.

0631

0.04

960.

0850

0.06

59βC,β

L0.48

0.00

500.

0012

0.06

910.

0701

0.13

520.

0016

0.10

640.

0054

0.11

50βC,β

O0

0.00

060.

0069

0.10

230.

0264

0.09

41−

0.00

710.

1453

0.00

580.

1408

βC,β

SC

0.21

0.00

22−

0.01

430.

1361

−0.

1322

0.16

06−

0.00

500.

1225

−0.

0260

0.25

07βC,β

SG

00.

0000−

0.00

230.

1163

0.07

380.

1631

0.00

540.

1308

−0.

0054

0.17

95βC,β

BN

0−

0.00

270.

0786

0.40

080.

3574

0.59

440.

0300

0.41

700.

0759

0.60

75βL,β

O0

0.00

090.

0115

0.00

84−

0.00

480.

0114

0.00

980.

0117

0.01

160.

0126

βL,β

SC

0.21

0.00

23−

0.00

150.

0075

−0.

0020

0.01

82−

0.00

610.

0141

−0.

0074

0.01

47βL,β

SG

00.

0001−

0.00

600.

0111

−0.

0046

0.01

470.

0005

0.01

43−

0.00

260.

0163

βL,β

BN

0−

0.00

400.

0046

0.02

74−

0.02

530.

0378

0.02

690.

0363

0.03

550.

0397

βO,β

SC

0.3

0.00

32−

0.00

560.

0120

−0.

0064

0.01

34−

0.01

210.

0169

−0.

0148

0.01

76βO,β

SG

0−

0.00

16−

0.01

050.

0151

0.00

130.

0105

−0.

0101

0.01

29−

0.00

490.

0213

βO,β

BN

0−

0.00

020.

0535

0.03

200.

0509

0.03

420.

0497

0.04

140.

0677

0.03

56βSC,β

SG

0−

0.00

020.

0120

0.01

170.

0047

0.02

150.

0138

0.01

580.

0121

0.02

34βSC,β

BN

0−

0.00

20−

0.06

070.

0465

−0.

0943

0.06

97−

0.03

710.

0555

−0.

0781

0.09

13βSG,β

BN

00.

0081−

0.04

020.

0592

0.01

370.

0654

−0.

0287

0.06

48−

0.04

820.

0731

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 106: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

106 An Application with Inter and Intra-consumer Heterogeneity

Tab

le8.

4:Differen

cesinthemeanan

dstan

dard

deviationof

thescalepa

rameter

betw

eenmod

elsw

ithan

dwith

outintra-con

sumer

heterogene

ity.

Data

Values

insample

Nohe

terogene

ityLo

wheterogeneity

Med

ium

heterogene

ityHigh

heterogeneity

Estimator

Means

With intra-consum

erheterogeneity

−0.48

1−0.52

6(109

%)

−0.48

6(101

%)

−0.48

6(101

%)

−0.47

8(99%

)

With

out

intra-consum

erheterogeneity

−0.48

1−0.51

6(107

%)

−0.46

8(97%

)−0.40

2(84%

)−0.30

1(63%

)

Std.

Devs.

With intra-consum

erheterogeneity

0.29

60.32

7(110

%)

0.28

5(96%

)0.28

7(97%

)0.30

1(101

%)

With

out

intra-consum

erheterogeneity

0.29

60.32

3(109

%)

0.27

9(94%

)0.25

3(85%

)0.24

7(83%

)

Not

e:*P

ercentages

representde

viations

from

thevalues

inthesample.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 107: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

107

Tab

le8.

5:Differen

cesin

theinter-consum

erstan

dard

deviations

ofthevariab

leswithintra-consum

erhe

terogene

itybe

tween

mod

elswith

andwith

outintra-consum

erhe

terogene

ity.

Std.

Dev.(β

L)+

Std.

Dev.(β

O)

Values

insample

No

heterogene

ityLo

whe

terogene

ityMed

ium

heterogene

ityHigh

heterogeneity

With

intra-consum

erheterogeneity

0.31

10.33

3(107

%)

0.32

4(104

%)

0.37

3(120

%)

0.36

3(117

%)

With

outintra-consum

erheterogene

ity0.31

10.35

8(115

%)

0.38

2(123

%)

0.40

9(132

%)

0.43

8(141

%)

Not

e:*P

ercentages

representde

viations

from

thevalues

inthesample.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 108: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

108 An Application with Inter and Intra-consumer Heterogeneity

for. Smaller values for the scale parameter indicate the bias is due to agreater degree of unobserved effects. In addition, when ignoring intra-consumer heterogeneity, increasing values of the standard deviationsof coefficients with intra-consumer heterogeneity are observed as indi-cated in Table 8.5. Thus, the unspecified intra-consumer heterogeneityresults in over-estimation of inter-consumer heterogeneity. This MonteCarlo example demonstrated the application of the extended Inter-/Intra-HB estimator and indicated that if intra-heterogeneity is presentthen an inter only estimator may overstate the level of inter-consumerheterogeneity.

A careful analysis of the HB Markov chains generated by the MonteCarlo examples above and in Section 6 indicated two issues that requirefurther attention. The first issue is that the Inverted Wishart priorwas found to be informative. The examples required data scaling toreduce the resulting bias caused by properties of the Inverted Wishart.For more details and alternatives to inverse Wishart, see Song et al.(2018) and Akinc and Vanderbroek (2018). The second issue is the needfor unusually long Markov chains in cases of weak identification. Thisindicated a need for systematic procedures to determine the requiredlength (200K iterations in the examples) and thinning (1 every 10 inthe examples) of the Markov chains.

To allow more flexible distributions than the Normality assumption,a straightforward extension is a discrete mixture of Normals obtainedby adding a Multinomial to the likelihood function and a Dirichlet prior.For more details, see Danaf et al. (2018).

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 109: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

9Policy Analysis

Stated preference elicitations and modeling of consumer tastes andchoices are often undertaken for the purpose of quantifying the impactsof alternative policies on suppliers, consumers, and markets. Clientsfor such policy analysis can include businesses interested in the finan-cial impacts of product innovations, governments who need to knowthe impacts of market regulations on transactions and product stan-dards, and litigants who want to determine the harm caused by “asis” policy compared with “but for” policy that corrects improper con-duct. Economic cost–benefit and business strategic analyses are usuallyprospective, forecasting the impacts of alternative policies on futureoutcomes for broad classes of consumers, e.g., the effect on consumersof food and drug product safety regulations, or the impact on suppliersof new product offerings. In contrast, analyses relevant to antitrust,patent, environmental, and tort law are primarily retrospective, deter-mining the consequences for individuals or specific consumer classesof alternatives to actual past policy, e.g., the harm to buyers of carswith unsafe ignition switches, or to homeowners in a neighborhood withgroundwater contamination. However, the separation between prospec-tive and retrospective analysis is imperfect, as realized outcomes and ex

109

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 110: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

110 Policy Analysis

post equity are also of concern to economists, while jurisprudence is alsoconcerned with the ex ante incentive effects of remedies for harm andwith costs incurred by consumers to mitigate potential harm, see Rawls(1971), Hammond (1981), Kaplow and Shavell (2002), and Fleurbaeyand Maniquet (2011). Revealed and stated preference studies can pro-vide forecasts of average per capita market demands under alternativepolicies for classes of consumers in either prospective or retrospectiveanalysis. There is no additional information in prospective analysis onthe choices that consumers would actually make under each policy, andno direct information on differences between anticipated and realizedoutcomes. However, in retrospective analysis, the analyst also observessome realized outcomes such as revealed product defects, and observesconsumers’ choices of products under actual market conditions, whichoften define relevant consumer classes. In this section, we first describegeneral-purpose simulation methods that line up well the use of hierar-chical Bayes estimates and can be used for a variety of policy purposes.Following this, we discuss policy impacts on suppliers, a motivation forthe development of CBC methods in market research, and concludewith a discussion of measuring consumer well-being.

9.1 Policy simulations

Suppose consumers face policy alternatives m = 1, 2, and note thatthis is a different use of the index m than in Sections 4–6 where mindexed menus presented to each subject in a CBC experiment. Sup-pose consumers have observed characteristics (W,A, T, s), and facemenus containing alternatives j = 0, . . . , Jm with attributes (zjm, pjm).For welfare analysis later, it will be useful to also introduce a netincremental income transfer t from the consumer to the government.Assume utilities are of a fully neoclassical money-metric form Equa-tion (3.1), ujm(t, ζ, ε) = T − t + vjm(t, ζ) + σ(ζ)εj , with net valuesvjm(t, ζ) ≡ X(W,A, T − t− pjm, s, zjm)β(ζ)− pjm, tastes ζ, i.i.d. EV1psychometric noise εj , a parametric taste distribution F (ζ|s, θ), andWTP parameters σ(ζ) and β(ζ) that are predetermined functions of ζ.These consumers are fully neoclassical since ζ and the εj are fixed acrosspolicy regimes. Giving a branded product distinct indices in J1 and J2,

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 111: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

9.1. Policy simulations 111

which may be appropriate in some applications, allows psychometriceffects across policies. However, this device rules out forms of depen-dence in noise across policies intermediate between total independenceand total dependence. The treatment of this specification can havesignificant effects on the distribution of policy impacts across classes ofconsumers making different “as is” choices.1

When preferences are heterogeneous and choices are discrete, re-vealed and stated preference studies in practice provide insufficientinformation to predict reliably the demands and utilities of individ-ual consumers. It is necessary to use bounds, approximations, or sta-tistical averages over classes of consumers, conditioned on what canbe usefully observed and distinguished among them, to obtain reli-able estimates. When consumer choices revealed in the market orstated in CBC experiments allow estimation using the methods ofSections 4–6 of choice probabilities of the form (3.8), then the esti-mated underlying taste distribution F (ζ|s, θ), Equation (3.9) can beused, along with the menus available to consumers under alternativepolicies, to forecast demands under various product or public policyinnovations. Further, F (ζ|s, θ) can be used to predict the distributionof money-metric decision utilities of consumers for these menus ofproducts.

A straightforward but computationally intensive approach to policyanalysis is to construct a synthetic population with the distributions ofutilities and tastes described above, simulate their choices and utilitiesfor the menus they face under each policy, and then process the resultsto obtain measures of interest. Start with a representative samplen = 1, . . . , N of consumer characteristics, (Wn, An, Tn, sn), drawn froma general population survey, or from a CDC sample if it is representative.For each n, make R draws ζ1n, . . . , ζRn from F (ζ|sn, θ), and Q i.i.d. EV1draws of psychometric noise vectors εqn, . . . , εqn.2 If hierarchical Bayes

1Giving a branded product distinct indices in J1 and J2 reintroduces variation inpsychometric effects across policies, which may be appropriate in some applications.Ruled out by this device are forms of dependence in noise across policies intermediatebetween total independence and total dependence.

2The application will determine whether εqj1n and εqj2n are independent, iden-tical, or exhibit some other form of dependence, and this specification can have

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 112: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

112 Policy Analysis

estimation of F (ζ|s, θ) is used, then the values ζ1n, . . . , ζRn may beobtained directly from the estimation process. Each combination ofdraws (n, r, q) defines a synthetic consumer. For policy m, construct theavailable menu of alternatives (zjm, pjm) for j = 1, . . . , Jm. As notedearlier in the table grape CBC example, this step may be non-trivial,particularly if some attributes are ambiguous, non-generic, or brand-specific, or are stated in “consumer-friendly” terms that may be difficultto translate from technical attributes controlled by the product supplier,or if there are questions of how much “brand loyalty” and currentconsumer information will transfer to new menus. Further, productprices and attributes in scenario m may not be predetermined, butrather determined through market equilibrium. Then, the analyst mustmodel producer behavior, and solve for the fixed point in prices andattributes that balance demand and supply. In many applications, thisrequires repetition of the steps below to obtain demands for a grid ofproduct prices and attributes, calculation of supplies at each grid point,and search for the grid point that approximates equilibrium, with gridrefinements as necessary.

For each synthetic consumer (n, r, q), calculate σ(ζrn) andβ(ζrn), the net value functions vjmn(t, ζrn) ≡ X(Wn, An, Tn − t −pjmn, sn, zjmn)β(ζrn)− pjmn, and the utilities ujmn(t, ζrn, εqn) = Tn −t+vjmn(t, ζrn)+σ(ζrn)εjqn When t 6= 0, this calculation will be repeatedfor each t of interest. Finally, calculate the choice that maximizes utility,indicated by djmn(t, ζrn, εqn) for each consumer and policy, the maxi-mum money-metric utility umn(t, ζrn, εqn) ≡ maxj∈Jm ujmn(t, ζrn, εqn),the marginal utility of transfer income given choice j, µjmn(t, ζrn) =1− ∂vjmn(t, ζrn)/∂t, and the unconditional marginal utility of transferincome µmn(t, ζrn, εmqn) ≡

∑j∈Jm djmn(t, ζrn, εmqn)µjmn(t, ζrn).

With the information in hand for this synthetic population, it is amechanical exercise to form statistical averages to estimate measures ofinterest. For example, one can estimate demands for each product andthe revenues generated under each policy, and if it is a target of theinvestigation, the policy impacts on selected sub-populations. When

significant effects on the distribution of policy impacts across classes of consumersmaking different “as is” choices.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 113: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

9.2. Demand analysis 113

necessary, this can be combined with a model of producer behavior todetermine market equilibrium prices and attributes under each policy.Using measures described in Section 9.3, one can also estimate consumerwelfare effects of policy changes. Confidence intervals can be obtainedby a Krinsky and Robb (1986) and Krinsky and Robb (1990) procedurewith the estimated covariance of θ, or by a block bootstrap in whichrepeated estimates of the deep parameters are obtained by resamplingfrom the survey respondents.

Both the Allenby–Train and NUTS procedures produce intermediateparameter draws ζni at iteration i whose empirical distribution overretained trials estimates the posterior density f(ζn|sn, θ) conditionedon the choices dn stated by respondent n in a CBC sample. Given thelimited data a CBC study provides on individual respondents, theseestimates will not be individually statistically consistent, but averagesof smooth functions of the empirical distribution will be statisticallywell-behaved. Then, if the CBC sample is representative and is used toconstruct the synthetic population, these empirical distributions canbe used to estimate expressions such as changes in levels of maximizedmoney-metric utilities for classes of consumers that are large enoughfor laws of large numbers to yield reliable averages.

9.2 Demand analysis

Consider a prospective policy intervention that changes the menus ofproducts that consumers face. For example, for a product like solarpanels, one could forecast the impact of introduction of a new productfeatures by suppliers, or of new government subsidies. Assume thatconsumers know product attributes and prices when they make choicesin the future market. Suppose consumers are described as in Section 9.1,with tastes that are fixed across policies and fully neoclassical. Let “En”denote an expectation over the consumer population; this is equivalentto indexing by n the observations (W,A, T, s) on the characteristics ofconsumers and (zjm, pjm) on the attributes and prices of each productj under an incumbent policy m = 1 and a replacement policy m = 2.Assume that t = 0. Let (zjλ, pjλ) for λ ∈ [1, 2] denote a rectifiable pathbetween policies for product attributes and prices. In the case that a

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 114: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

114 Policy Analysis

product is unavailable under policy m, we set its price to a large positivenumber so that it effectively drops out of the choice set. The estimated(per capita) demand Dj and revenue Rj for product j along this pathare given by

Dj (λ) = EnEζ|sPj (W,A, T, s,xλ,pλ, ζ)Rj(λ) = EnEζ|spjλP j (W,A, T, s,xλ,pλ, ζ) .

(9.1)

where Pj is the MNL model (3.5) evaluated at ρλ = ζ. Note thatPj declines exponentially for large pjλ, so that Rj(λ) is bounded andeventually decreasing as pjλ → +∞. The change in demand alongλ ∈ [1, 2] is

dDj(λ)dλ

= EnEζ|sPj(W,A, T, s,xλ,pλ, ζ)

× δjk − Pk(W,A, T, s,xλ,pλ, ζ)α(ζ)

dxkλndλ

· β(ζ)− dpkλndλ

,

(9.2)

where δjk is one if j = k, zero otherwise. This formula is easy to interpretin special cases. For example, if the price of good k is the same for alln and the policy changes only this price, and no x attributes or otherprices, then Equation (9.2) gives the cross-elasticity of demand for goodj with respect to the price of good k,

∂ logDj(λ)∂ logpkλ

≡ pkλDj(λ)

dDj(λ)/dλdpkλ/dλ

= −EnEζ|s(δjk − Pk(W,A, T, s,xλ,pλ, ζ))pkλ

α(ζ)

· Pj (W,A, T, s,xλ,pλ, ζ)Dj(λ) , (9.3)

a population-and-taste-weighted average (i.e., over n and ζ) of MNLprice elasticities.

A typical economic application of Equation (9.2) is to calculate thechange in profit for the producer of product j resulting from a change∆pj1 in its product price and ∆zj1 in its attributes. The change ∆zj1induces a change that may span multiple components of xj1 : ∆xj1 =

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 115: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

9.3. Consumer welfare analysis 115

X(W,A, T −pj1−∆pj1, s, zj1 +∆zj1)−X(W,A, T −pj1, s, zj1). Assumethat ∆xj1 and ∆pj1 are the same for all consumers in the population.Along a linear path xjλ = xj1 +(λ−1)∆xj1 and pjλ = pj1 +(λ−1)∆pj1for λ ∈ [1, 2], the estimated changes in demand and revenue are

dDj(λ)dλ

= EnEζ|s

((1−Pj (Wn, An, Tn, sn,xλn,pλn, ζ)) pjλ1α (ζ)

)× (∆xj1β (ζ)−∆pj1)Pj (W,A, T, s,xλ,pλ, ζ),

dRj(λ)dλ

= EnEζ|s

(1− (1−Pj (Wn, An, Tn, sn,xλn,pλn, ζ)) pjλ1

α (ζ)

)× (∆xj1β (ζ)−∆pj1)Pj (W,A, T, s,xλ,pλ, ζ)

(9.4)

population-and-taste-weighted averages of MNL elasticity termstimes (∆xj1β(ζ) −∆pj1).3 Combined with estimates of the marginalcost of product modifications and assumptions (e.g., Nash) on theresponse of other suppliers to its innovations, Equation (9.4) allowsestimation of the incremental profitability of the change. Note that whenβ = β(ζ) is homogeneous, demand and revenue are unchanged when∆xj1β−∆pj1 = 0, and the innovation increases profit if and only if it iscost-saving. For innovations that are not uniform in their incidence onconsumers, when β(ζ) is heterogeneous, or when ∆xj1β−∆pj1 6= 0, theprofit analysis requires full calculation of Equation (9.2). Note that theformulas (9.1)–(9.4) can all be estimated using the synthetic populationof consumers n = 1, . . . , N with a distribution of tastes ζrn constructedin Section 8.1, without sampling from the EV1 distributions, and thatthis alternative will reduce sampling variance relative to completereliance on the full synthetic population (n, r, q).

9.3 Consumer welfare analysis

Suppose the population of consumers with nearly neoclassical de-cision utilities described in Section 9.1, and consider an incum-bent/default/“as is” policym = 1 and a replacement/candidate/“but for”

3If the supplier makes multiple products, the sum of revenues in (38) over itsproducts leads to an extension of (41) that incorporates cross-elasticities between itsproducts and allows analysis of the effects of product positioning and cannibalization.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 116: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

116 Policy Analysis

policy m = 2. McFadden (2018) and McFadden and Train (2018) de-velop the consumer theory foundations and practical formulas forapplied welfare analysis. This analysis can be applied to variouscombinations of stated and revealed preference data. In particular,the welfare consequences of policy changes can be calculated from theutility function of each simulated consumer in Section 9.1, and thenaggregated to the population, or to classes that are sufficiently large sothat the aggregates are statistically reliable.

There are several issues, detailed in McFadden (2018), that need tobe resolved in setting up these calculations. First, welfare applicationsmay be either prospective, examining policy alternatives that mightbe implemented in the future, or retrospective, calculating the benefitor harm to consumers from past policies. For example, determiningthe impact on consumer welfare of a proposed merger of suppliers re-quires a prospective analysis, while assessing the harm to consumersfrom past collusion between suppliers requires a retrospective analy-sis. Second, while retrospective analysis is usually conducted for thepurpose of computing transfers to be fulfilled to “make consumerswhole”, the transfers in prospective analysis are often hypothetical,intended to determine whether the policy change could in principlebe a Pareto improvement, rather than fulfilled to ensure an actualPareto improvement. Third, market and stated preference studies usu-ally provide only partial observability, insufficient to identify reliably thetastes of individual consumers, and at best identifying the distributionof tastes conditioned on observed histories and socioeconomic status.Then, any system of fulfilled transfers based on observed characteristicswill generally lead to gainers and losers. Consequently, compensationcalculations must consider both overall welfare and the impact of impre-cision in individual compensating transfers. Fourth, experienced-utilityof consumers may differ from the decision-utility that determines mar-ket demands, as the result of resolution of contingencies regardingattributes of products and interactions with consumer needs, or as theresult of inconsistencies in tastes and incomplete optimizing behavior.There is then a question of whether the target of welfare analysis isexperienced utility, reflecting outcomes, or decision utility, reflectingopportunities.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 117: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

10Conclusions

This monograph has reviewed methods for stated preference elicitation,concentrating on the most successful and widely applied approach,choice-based conjoint (CBC) analysis. It has detailed recommendationsfor the design of CBC studies, including sampling, subject training,incentive alignment, and reconciliation and validation of CBC dataagainst revealed preference data. It has identified the conditions underwhich CBC data provide a reliable basis for forecasting demand for newor modified products, and cautioned readers about the sensitivity of CBCrespondents to context and framing, the importance of accounting forthe manipulability of consumers in both real and hypothetical markets,and in realistic hypothetical markets mimicking the manipulationspresent in real markets.

This monograph has reviewed the choice-theory underpinnings ofCBC, and recommends modeling utility in money-metric or WTP-spaceform, introducing flexible generic attributes and taste heterogeneity inthe population, and controlling carefully how components of incomeenter utility, discrete choice probability models, and measures of con-sumer welfare. It recommends the use of mixed logit choice modelsadapted to the CDC panel data structure to model the behavior of

117

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 118: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

118 Conclusions

nearly neoclassical consumers who aside from psychometric noise havestable tastes across different CBC menus.

The monograph has discussed extensively the use of simulatedmaximum likelihood and hierarchical Bayes methods to estimate thedeep parameters that characterize distributions of tastes and choices inthe population, and described methods that facilitate computation ofboth types of estimators. Hierarchical Bayes methods, particularly theAllenby-Train and NUTS procedures with diffuse priors, are shown tobe very fast, even for high-dimensional problems, and for CBC studies ofpractical size are shown to be very close to classical maximum likelihoodestimators. Therefore, we recommend use of these estimators.

Our overall conclusion is that collection and analysis of stated pref-erence data is a powerful tool for exploring and predicting consumerbehavior, but both the data collection and analysis require considerablecare and caution in inferring that the results are predictive and reliablefor real market behavior. We urge economist and non-economists toevaluate realistically what stated preference methods can do now, avoidoverselling results from stated preference experiments in circumstanceswhere they have not been proven reliable, and to engage in a vigor-ous, wide-ranging, cross-disciplinary research effort to improve thesemethods, expand them to provide structural information on consumerbehavior such as distinctions between perceptions, decision utility, andexperienced utility, and systematically monitor the performance of thesemethods in applications.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 119: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

Appendix

True values of intra-consumer heterogeneity and estimation resultswith inter-only estimator

Intra-consumer heterogeneity — True values:

Table A1: Low heterogeneity — population

Covariances and CorrelationStd. Devs. in top right corner

βL βOβL 0.2 0.040 −0.3βO 0.1 −0.006 0.010

Table A2: Medium heterogeneity — population

Covariances and CorrelationStd. Devs. in top right corner

βL βOβL 0.6 0.360 −0.3βO 0.3 −0.054 0.09

119

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 120: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

120 Conclusions

Table A3: High heterogeneity — population

Covariances and CorrelationStd. Devs. in top right corner

βL βOβL 1 1 −0.3βO 0.5 −0.15 0.25

Estimation results for models without intra-consumer heterogeneity

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 121: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

121

Tab

leA

4:Datawithinteran

dintra-consum

erhe

terogeneity

(three

levels

ofintra-consum

erhe

terogene

ity),mod

elwithon

lyinter-consum

erhe

terogene

ity

P arameter

True

Values

NoHeterog

eneity

Low

Heterog

eneity

Med

ium

Heterog

eneity

HighHeterog

eneity

Means

Popu

latio

nSa

mple

Mean

Std.

Dev.

Mean

Std.

Dev.

Mean

Std.

Dev.

Mean

Std.

Dev.

Pane

lA:P

aram

eter

estim

ates

βS

11.

005

0.95

00.04

40.

983

0.04

11.

020

0.04

61.

015

0.05

C0.

30.

299

0.33

70.03

40.

295

0.03

10.

292

0.04

90.

289

0.04

L0.

20.

197

0.17

60.02

50.

221

0.02

60.

236

0.02

70.

260

0.02

O0.

10.

109

0.09

50.02

60.

123

0.02

60.

148

0.02

50.

127

0.02

SC

00.

001−

0.02

30.03

90.

074

0.03

70.

056

0.06

40.

041

0.05

SG

0.1

0.09

30.

142

0.06

20.

084

0.04

20.

059

0.05

50.

069

0.05

BN

22.

012

2.01

60.05

02.

002

0.05

11.

969

0.05

51.

973

0.05

5Lo

g(α

n)

−0.

5−

0.48

1−

0.51

60.02

3−

0.46

80.02

3−

0.40

20.02

2−

0.30

10.02

2Pa

nelB

:Stand

ardde

viations

forinter-consum

erhe

terogene

ityβ

S0.

30.

304

0.25

70.06

30.

202

0.05

40.

185

0.05

20.

201

0.04

C0.

10.

099

0.17

70.04

70.

152

0.04

20.

154

0.03

60.

183

0.03

L0.

10.

101

0.16

20.03

30.

217

0.05

10.

187

0.05

20.

222

0.05

O0.

20.

210

0.19

60.05

00.

165

0.04

40.

222

0.04

60.

216

0.04

SC

0.05

0.05

10.

181

0.05

80.

180

0.05

10.

212

0.04

80.

214

0.04

SG

0.2

0.20

20.

202

0.05

10.

222

0.06

70.

247

0.07

50.

204

0.06

BN

10.

988

0.83

40.06

40.

970

0.07

10.

926

0.07

70.

898

0.06

4Lo

g(α

n)

0.3

0.29

60.

323

0.03

00.

279

0.02

90.

253

0.02

50.

247

0.02

6(C

ontin

ued)

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 122: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

122 Conclusions

Tab

leA

4:(C

ontin

ued)

T rue

Values

NoHeterog

eneity

Low

Heterog

eneity

Med

ium

Heterog

eneity

High

Heterog

eneity

Co varianc

esPo

pulatio

nSa

mple

Mean

Std.

Dev.

Covarianc

eStd.

Dev.

Covarianc

eStd.

Dev.

Covarianc

eStd.

Dev.

P ane

lC:C

ovarianc

esforinter-consum

erhe

terogene

ityβS,β

C1.

80.

0185

0.12

10.13

80.

032

0.12

80.

001

0.10

70.

010

0.09

9βS,β

L0

0.00

07−

0.00

80.01

30.

003

0.01

80.

007

0.01

40.

001

0.01

6βS,β

O0

−0.

0002−

0.00

70.01

40.

007

0.01

50.

011

0.01

50.

006

0.01

2βS,β

SC

0.45

0.00

460.

002

0.01

60.

005

0.01

50.

003

0.01

5−

0.00

60.01

5βS,β

SG

0−

0.00

040.

013

0.01

7−

0.00

40.02

30.

002

0.02

30.

013

0.01

6βS,β

BN

0−

0.00

320.

097

0.04

50.

016

0.05

30.

021

0.06

90.

073

0.05

1βC,β

L0.

480.

0050

0.06

10.09

70.

044

0.14

50.

001

0.10

30.

003

0.14

3βC,β

O0

0.00

060.

075

0.10

40.

000

0.09

2−

0.00

80.11

3−

0.00

20.14

5βC,β

SC

0.21

0.00

22−

0.18

40.17

1−

0.08

00.14

4−

0.00

20.13

2−

0.01

70.14

2βC,β

SG

00.

0000−

0.01

10.12

20.

094

0.15

70.

002

0.15

4−

0.00

10.15

6βC,β

BN

0−

0.00

270.

835

0.36

70.

101

0.64

80.

002

0.59

70.

078

0.44

9βL,β

O0

0.00

090.

016

0.01

20.

000

0.01

40.

010

0.01

60.

011

0.02

0βL,β

SC

0.21

0.00

23−

0.00

50.01

10.

000

0.01

6−

0.00

20.01

3−

0.00

90.02

3βL,β

SG

00.

0001−

0.01

00.01

50.

005

0.02

50.

002

0.02

4−

0.00

40.02

3βL,β

BN

0−

0.00

40−

0.00

20.03

0−

0.04

20.03

90.

016

0.04

2−

0.00

50.04

1βO,β

SC

0.3

0.00

32−

0.00

60.01

3−

0.00

10.00

8−

0.01

30.01

8−

0.01

40.01

8βO,β

SG

0−

0.00

16−

0.01

30.01

9−

0.00

30.01

7−

0.01

90.02

9−

0.00

80.01

7βO,β

BN

0−

0.00

020.

046

0.03

20.

040

0.03

50.

054

0.04

20.

062

0.03

5βSC,β

SG

0−

0.00

020.

007

0.01

50.

004

0.01

50.

028

0.02

20.

006

0.02

1βSC,β

BN

0−

0.00

20−

0.05

90.04

1−

0.05

10.07

2−

0.00

80.07

4−

0.08

60.05

3βSG,β

BN

00.

0081−

0.00

20.06

20.

023

0.06

80.

003

0.06

5−

0.01

20.05

2

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 123: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

Acknowledgements

Charlie Gibbons made major contributions to our hierarchical Bayesdiscussion and estimation. We also benefited from discussions withMark Berkman, Michel Bierlaire, Lisa Cameron, Andrew Daly, MogensFosgerau, Garrett Glasgow, Stephane Hess, Armando Levy, DouglasMacNair, Charles Manski, Kevin Murphy, Frank Pinter, Joan Walker,and Ken Wise.

123

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 124: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

References

Abramowitz, M. and L. Stegum (1964). Handbook of MathematicalFunctions. National Bureau of Standards, 24.4.

Acito, F. and A. Jain (1980). “Evaluation of Conjoint Analysis Results:A Comparison of Methods”. Journal of Marketing Research. 17:106–112.

Akaah, I. and P. Korgaonkar (1983). “An Empirical Comparison of thePredictive Validity of Self-explicated, Huber-hybrid, Traditional Con-joint, and Hybrid Conjoint Models”. Journal of Marketing Research.20: 187–197.

Akinc, A. and M. Vanderbroek (2018). “Bayesian Estimation of MixedLogit Models: Selecting an Appropriate Prior for the CovarianceMatrix”. Journal of Choice Modeling. 29: 133–151.

Albert, J. and S. Chib (1996). “Computation in Bayesian Econometrics:An Introduction to Markov Chain Monte Carlo”. In: BayesianComputational Methods and Applications. Ed. by C. Hill. London:JAI Press. 3–24.

Allais, M. (1953). “Le Comportement de l’Homme Rationnel Devant LeRisque: Critique Des Postulats Et Axiomes de l’école Américaine”.Econometrica. 21: 503–546.

Allenby, G. (1997). “An Introduction to Hierarchical Bayesian Model-ing”. Tutorial Notes, Advanced Research Techniques Forum, Ameri-can Marketing Association.

124

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 125: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

References 125

Allenby, G. and P. Lenk (1994). “Modeling Household Purchase Be-havior with Logistic Normal Regressions”. Journal of the AmericanStatistical Association. 89: 1218–1231.

Allenby, G. and P. Rossi (1993). “A Bayesian Approach to EstimatingHousehold Parameters”. Journal of Marketing Research. 30: 171–182.

Allenby, G. and P. Rossi (2006). “Hierarchial Bayes Models: A Practi-tioner’s Guide”. In: The Handbook of Marketing Research. Ed. byR. Grover and M. Vriens. Los Angeles: Sage Publications.

Andrews, R., A. Ansari, and I. Currim (2002). “Hierarchical Bayesversus Finite Mixture Conjoint Analysis Models: A Comparison ofFit, Prediction, and Partworth Recovery”. Journal of MarketingResearch. 39: 87–98.

Ariely, D., G. Lowenstein, and D. Prelec (2006). “Tom Sawyer andthe Construction of Value”. Journal of Economic Behavior andOrganization. 60: 1–10.

Arrow, K. (1971). “Exposition of the Theory of Choice Under Uncer-tainty”. In: Theory of Risk-Bearing. Ed. by K. Arrow. Chicago:Markham. 44–89.

Azizzadeh, F. and S. Rezakhah (2014). “The CUSUM Test for DetectingStructural Changes in Strong Mixing Processes”. Communicationsin Statistics. 43(17): 3733–3750.

Bąk, A. and T. Bartłomowicz (2012). “Conjoint Analysis Method and ItsImplementation in Conjoint R Package”. Working paper, WroclawUniversity.

Balcombe, K. and I. Fraser (2009). “Dichotomous-choice ContingentValuation with ‘Don’t Know’ Responses and Misreporting”. Journalof Applied Econometrics. 24.7.

Bateson, J., D. Reibstein, and W. Boulding (1987). “Conjoint AnalysisReliability and Validity: A Framework for Future Research”. In:Review of Marketing. Ed. by M. Houston. 451–481.

Becker, F., M. Danaf, B. A. X. Song, and M. Ben-Akiva (2018). “Hier-archical Bayes Estimator for Logit Mixtures with Inter- and Intra-consumer Heterogeneity”. Transportation Research Part B: Method-ological. 117(A): 1–17.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 126: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

126 References

Becker, G., M. DeGroot, and J. Marschak (1964). “Measuring Utilityby a Single-response Sequential Method”. Behavioral Science. 9:226–32.

Ben-Akiva, M. and D. Bolduc (1996). “Multinomial Probit with a LogitKernel and a General Parametric Specification of the CovarianceStructure”. Working paper, Department of Civil Engineering, MIT.

Ben-Akiva, M., D. Bolduc, and M. Bradley (1993). “Estimation ofTravel Model Choice Models with Randomly Distributed Values ofTime”. Transportation Research Record. 1413: 88–97.

Ben-Akiva, M. and S. Lerman (1985). Discrete Choice Analysis: Theoryand Application to Travel Demand. Cambridge: MIT Press.

Ben-Akiva, M. and T. Morikawa (1990). “Estimation of Switching Mod-els from Revealed Preferences and Stated Intentions”. TransportationResearch A. 24: 485–495.

Berndt, E., B. Hall, R. Hall, and J. Hausman (1974). “Estimation andInference in Nonlinear Structural Models”. Annals of Economic andSocial Measurement. 3(4): 653–665.

Berry, S., J. Levinsohn, and A. Pakes (1995). “Automobile Prices inMarket Equilibrium”. Econometrica. 63(4): 841–890.

Berry, S., J. Levinsohn, and A. Pakes (1998). “Differentiated ProductsDemand Systems from a Combination of Micro and Macro Data:The New Car Market”. Working paper, Yale University.

Berry, S., J. Levinsohn, and A. Pakes (2004). “Differentiated ProductsDemand Systems from a Combination of Micro and Macro Data:The New Car Market”. Journal of Political Economy. 112: 68–105.

Betancourt, M. and M. Girolami (2013). “Hamilton Monte Carlo forHierarchical Models”. Working paper, arXiv:1312.0906.

Bhat, C. (2001). “Quasi-random Maximum Simulated Likelihood Es-timation of the Mixed Multinomial Logit Model”. TransportationResearch B. 35: 677–693.

Bhat, C., S. Astroza, R. Sidharthan, and P. Bhat (2014). “A Multi-variate Hurdle Count Data Model with an Endogenous MultipleDiscrete-continuous Selection System”. Transportation Research B:Methodological. 63: 77–97.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 127: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

References 127

Bishop, R. and T. Heberlein (1979). “Measuring Values of Extramar-ket Goods: Are Indirect Measures Biased?” American Journal ofAgricultural Economics. 61(5): 926–930.

Blass, A., S. Lach, and C. Manski (2010). “Using Elicited Choice Prob-abilities to Estimate Random Utility Models for Electricity Reliabil-ity”. International Economic Review. 51: 421–440.

Bleimer, M. and J. Rose (2009). “Designing Stated Choice Experiments:The State of the Art”. In: The Expanding Sphere of Travel BehaviourResearch. Ed. by R. Kitamura, T. Yoshi, and T. Yamamoto. Bingley,West Yorkshire: Emerald Group Publishing, Ch. 25. 499–538.

Bohm, P. (1972). “Estimating the Demand for Public Goods: An Ex-periment”. European Economic Review. 3: 111–130.

Bohm, P., J. Linden, and J. Sonnegard (1997). “Eliciting ReservationPrices: Becker DeGroot-Marschak Mechanisms vs. Markets”. TheEconomic Journal. 107: 1079–1089.

Boyce, D. and H. Williams (2015). Forecasting Urban Travel Demand.Cheltenham: Edward Elgar.

Brookshire, D., B. Ives, and W. Schmitz (1976). “The Value of AestheticPreferences”. Journal of Environmental Economics and Measure-ment. 3(4): 325–346.

Brookshire, D., A. Randall, and J. Stoll (1980). “Valuing Increments andDecrements in Natural Resource Service Flows”. American Journalof Agricultural Economics. 62(3): 478–488.

Brownstone, D. (2001). “Discrete Choice Modeling for Transportation”.In: Travel Behavior Research: The Leading Edge. Ed. by D. Hensher.Oxford: Elsevier. 97–124.

Brownstone, D., D. Bunch, and K. Train (2000). “Joint Mixed LogitModels of Stated and Revealed Preferences for Alternative-fuelVehicles”. Transportation Research, Part B. 34: 315–338.

Brownstone, D. and K. Train (1999). “Forecasting New Product Penetra-tion with Flexible Substitution Patterns”. Journal of Econometrics.89: 109–129.

Burda, M., M. Harding, and J. Hausman (2012). “A Poisson MixtureModel of Discrete Choice”. Journal of Econometrics. 166(2): 184–203.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 128: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

128 References

Camerer, C. and R. Hogarth (1999). “The Effects of Financial In-centives in Experiments: A Review and Capital-labor-productionFramework”. Journal of Risk and Uncertainty. 19: 7–42.

Cameron, T. and J. DeShazo (2013). “Demand for Health Risk Re-ductions”. Journal of Environmental Economics and Management.65(1): 87–109.

Carmone, F., P. Green, and A. K. Jain (1978). “Robustness of Con-joint Analysis: Some Monte Carlo Results”. Journal of MarketingResearch. 15: 300–303.

Caro, F., T. Ho, D. McFadden, A. Gottleib, C. Yee, T. Chan, and J.Winter (2012). “Using the Internet to Administer more RealisticVignette Experiments”. Social Science Computer Review. 30: 184–201.

Carpenter, B., A. Gelman, M. Hoffman, D. Lee, B. Goodrich, M. Betan-court, M. Brubaker, J. Guo, P. Li, and A. Riddell (2017). “STAN: AProbabilistic Programming Language”. Journal of Statistical Soft-ware. 76(1).

Carson, R. (2012). “Contingent Valuation: A Practical AlternativeWhen Prices Aren’t Available”. Journal of Economic Perspectives.26(4): 27–42.

Carson, R. T., J. J. Louviere, D. A. Anderson, P. Arabie, D. S. Bunch,D. A. Hensher, R. M. Johnson, W. F. Kuhfeld, D. Steinberg, J.Swait, H. Timmermans, and J. B. Wiley (1994). “ExperimentalAnalysis of Choice”. Marketing Letters. 5(4): 351–368.

Carson, R. and M. Czajkowski (2013). “A New Baseline Model for Esti-mating Willingness to Pay from Discrete Choice Models”. Workingpaper, University of California, San Diego.

Carson, R., N. Flores, and N. Meade (2001). “Contingent Valuation: Con-troversies and Evidence”. Environmental and Resource Economics.19: 173–210.

Carson, R. and J. Louviere (2011). “A Common Nomenclature for StatedPreference Elicitation Approaches”. Environmental and ResourceEconomics. 49(4): 539–559.

Chandukala, S., J. Kim, T. Otter, P. Rossi, and G. Allenby (2007).“Choice Models in Marketing: Economic Assumptions, Challenges,and Trends”. Foundations and Trends in Marketing. 2(2): 97–184.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 129: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

References 129

Chapman, R. and R. Staelin (1982). “Exploiting Rank Ordered ChoiceSet Data within the Stochastic Utility Model”. Journal of MarketingResearch. 14: 288–301.

Chen, R. (2017). “Models of Count with Endogenous Choices”. Trans-portation Research Procedia. 23C: 460–479.

Chesher, A. and J. Santos-Silva (2002). “Taste Variation in DiscreteChoice Models”. Review of Economic Studies. 69: 62–78.

Chib, S. and E. Greenberg (1996). “Markov Chain Monte Carlo Simula-tion Methods in Econometrics”. Econometric Theory. 12(3): 409–431.

Chipman, J. and J. Moore (1980). “Compensating Variation, Consumer’sSurplus, and Welfare”. American Economic Review. 70: 933–348.

Chipman, J. and J. Moore (1990). “Acceptable Indicators of WelfareChange, Consumer’s Surplus Analysis, and the Gorman Polar Form”.In: Preferences, Uncertainty, and Optimality: Essays in Honor ofLeonid Hurwicz. Ed. by D. McFadden and M. Richter. New York:Westview Press. 68–120.

Ciriacy-Wantrup, S. (1947). “Capital Returns from Soil-conservationPractices”. Journal of Farm Economics. 29: 1181–1202.

Corradi, V., N. Swanson, and H. White (2000). “Testing for Stationarity-ergodicity and for Comovements between Nonlinear Discrete TimeMarkov Processes”. Journal of Econometrics. 96: 39–73.

Daly, A., S. Hess, and K. Train (2012). “Assuring Finite Moments forWillingness to Pay in Random Coefficient Models”. Transportation.39(1): 19–31.

Danaf, M., B. Atasoy, and M. Ben-Akiva (2018). “Logit Mixture withInter and Intra-consumer Heterogeneity and Flexible Mixing Distri-butions”. Working paper, MIT.

Davis, R. (1963a). “Recreation Planning as an Economic Problem”.Natural Resources. 3: 239–249.

Davis, R. (1963b). The Value of Outdoor Recreation: An EconomicStudy of the Maine Woods. PhD Dissertation, Harvard University.

Deaton, A. (2012). “The Financial Crisis and the Well-being of Amer-ica”. In: Investigations in the Economics of Aging. Ed. by D. Wise.Chicago: University of Chicago Press. 343–376.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 130: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

130 References

Deaton, A. and J. Muellbauer (1980). Economics and Consumer Be-havior. Cambridge: Cambridge University Press.

Delavande, A. and C. Manski (2015). “Using Elicited Choice Probabili-ties in Hypothetical Elections to Study Decisions to Vote”. ElectoralStudies. 38: 28–37.

Delgado, M. (2005). “Sign Tests for Long-memory Time Series”. Journalof Econometrics. 128(2): 215–251.

Desarbo, W., V. Ramaswamy, and S. Cohen (1995). “Market Segmen-tation with Choice-based Conjoint Analysis”. Marketing Letters. 6:137–147.

Domowitz, I. and M. El-Gamal (1993). “A Consistent Test ofStationarity-ergodicity”. Econometric Theory. 9(4): 589–601.

Dong, S., M. Ding, and J. Huber (2010). “A Simple Mechanism toIncentive-align Conjoint Experiments”. International Journal ofResearch in Marketing. 27: 25–32.

Elrod, T., J. Louviere, and K. Davey (1992). “An Empirical Comparisonof Ratings-based and Choice-based Conjoint Models”. Journal ofMarketing Research. 29: 368–377.

Fink, D. (1997). “A Compendium of Conjugate Priors”. Working paper,Montana State University.

Fischhoff, B. and C. Manski (1999). Elicitation of Preferences. Kluwer:Boston.

Fisher, I. (1892). Mathematical Investigations in the Theory of Valueand Prices. New Haven: Yale.

Fleurbaey, M. and F. Maniquet (2011). A Theory of Fairness and SocialWelfare. Cambridge: Cambridge University Press.

Fosgerau, M., D. McFadden, and M. Bierlaire (2013). “Choice Prob-ability Generating Functions”. Journal of Choice Modelling. 8: 1–18.

Frisch, R. (1926). “Sur Un Probleme of Economie Pure. Norsk Matem-atisk Forenings Sskrifter, 16, 1–40. Translation”. In: Preferences,Utility, and Demand. Ed. by J. Chipman, L. Hurwicz, M. Richter,and H. Sonnenschein. New York: Harcourt.

Gelman, A., J. Carlin, H. Stern, and D. Rubin (2004). Bayesian DataAnalysis (2nd edition). Boca Raton, FL: Chapman & Hall/CRC.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 131: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

References 131

Geweke, J. (1997). Posterior Simulators in Econometrics. EconometricSociety Monographs 28. New York: Cambridge University Press.

Glasgow, G. and S. Butler (2017). “The Value of Non-personally Identi-fiable Information to Consumers of Online Services: Evidence froma Discrete Choice Experiment”. Applied Economics Letters. 24(6):392–395.

Gorman, W. (1953). “Community Preference Fields”. Econometrica. 21:63–80.

Gorman, W. (1961). “On a Class of Preference Fields”. Metroeconomica.13: 53–56.

Green, D., K. Jacowitz, D. Kahneman, and D. McFadden (1998). “Ref-erendum Contingent Valuation, Anchoring, and Willingness to Payfor Public Goods”. Resource and Energy Economics. 20: 85–116.

Green, P. (1974). “On the Design of Choice Experiments InvolvingMultifactor Alternatives”. Journal of Consumer Research. 1: 61–8.

Green, P., A. Krieger, and Y. Wind (2001). “Thirty Years of ConjointAnalysis: Reflections and Prospects”. Interfaces. 31: S56–S73.

Green, P. and V. Srinivasan (1978). “Conjoint Analysis in ConsumerResearch: Issues and Outlook”. Journal of Consumer Research. 5:103–123.

Green, P. and V. Srinivasan (1990). “Conjoint Analysis in Marketing:New Developments with Implications for Research and Practice”.Journal of Marketing. 54: 3–19.

Hajivassiliou, V. and D. McFadden (1998). “The Method of SimulatedScores for the Estimation of LDV Models”. Econometrica. 66: 863–896.

Hajivassiliou, V. and P. Ruud (1994). “Classical Estimation Methodsfor LDV Models Using Simulation”. In: Handbook of Econometrics.Ed. by R. Engle and D. McFadden. Amsterdam: North-Holland.2383–2441.

Hammond, P. (1981). “Ex-ante and Ex-post Welfare Optimality UnderUncertainty”. Economica. 48: 235–250.

Hanemann, M., J. Loomis, and B. Kanninen (1991). “Statistical Ef-ficiency of Double-bounded Dichotomous Choice Contingent Val-uation”. American Journal of Agricultural Economics. 73: 1255–1263.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 132: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

132 References

Hauser, J. and V. Rao (2002). “Conjoint Analysis, Related Modelling,and Applications”. In: Advances in Marketing Research: Progressand Prospects. Ed. by Y. Wind and P. Green. New York: Springer.141–168.

Hauser, J. and G. Urban (1979). “Assessment of Attribute Importancesand Consumer Utility Functions: von Neumann-Morgenstern TheoryApplied to Consumer Behavior”. Journal of Consumer Research. 5:251–262.

Hausman, J. (2012). “Contingent Valuation: From Dubious to Hopeless”.Journal of Economic Perspectives. 26(4): 43–56.

Heckman, J. (1981). “The Initial Parameters Problem and the Problemof Initial Conditions in Estimating a Discrete Time-discrete DataStochastic Process”. In: Structural Analysis of Discrete Data. Ed. byC. Manski and D. McFadden. Cambridge: MIT Press. 179–195.

Hensher, D. and M. Bradley (1993). “Using Stated Response Data toEnrich Revealed Preference Discrete Choice Models”. MarketingLetters. 4: 39–152.

Hensher, D. and W. Greene (2002). “Specification and Estimation ofNested Logit Model”. Transportation Research Part B. 36: 1–17.

Hensher, D. and W. Greene (2003). “The Mixed Logit Model: TheState of Practice and Warnings for the Unwary”. Transportation.30: 133–176.

Hensher, D. and J. Louviere (1983). “Identifying Individual Preferencesfor International Air Travel: An Application of Functional Mea-surement Theory”. Journal of Transport Economics and Policy. 17:225–245.

Hensher, D., J. Rose, and A. Collins (2011). “Identifying CommuterPreferences for Existing Modes and a Proposed Metro in Sydney,Australia, with Special Reference to Crowding”. Public Transport.3: 109–147.

Hess, S. and J. Rose (2009). “Allowing for Intra-respondent Variationsin Coefficients Estimated on Repeated Choice Data”. TransportationResearch, Part B. 43(6): 708–719.

Hess, S. and K. Train (2011). “Recovery of Inter- and Intra-personalHeterogeneity Using Mixed Logit Models”. Transportation Research,Part B. 45(7): 973–990.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 133: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

References 133

Hicks, J. (1939). Value and Capital. Oxford: Clarendon Press.Hoffman, M. and A. Gelman (2014). “The No-U-Turn sampler: Adap-

tively Setting Path Lengths in Hamiltonian Monte Carlo”. Journalof Machine Learning Research. 15: 1351–1381.

Hofstede, F. T., Y. Kim, and M. Wedel (2002). “Bayesian Predictionin Hybrid Conjoint Analysis”. Journal of Marketing Research. 36:253–261.

Hong, H. and B. Preston (2012). “Bayesian Averaging, Prediction, andNonnested Model Selection”. Journal of Econometrics. 167(2): 358–369.

Huber, J. (1975). “Predicting Preferences on Experimental Bundlesof Attributes: A Comparison of Models”. Journal of MarketingResearch. 12: 290–297.

Huber, J. (1987). “Conjoint Analysis: How We Got Here and WhereWe Are”. In: Proceedings of the Sawtooth Software Conference onPerceptual Mapping, Conjoint Analysis, and Computer Interviewing.237–252.

Huber, J. and K. Train (2001). “On the Similarity of Classical andBayesian Estimates of Individual Mean Partworths”. MarketingLetters. 12: 259–269.

Huber, J., D. Wittink, J. Fiedler, and R. Miller (1993). “The Effective-ness of Alternative Preference Elicitation Procedures in PredictingChoice”. Journal of Marketing Research. 30: 105–114.

Huber, J. and K. Zwerina (1996). “The Importance of Utility Balancein Efficient Choice Designs”. Journal of Marketing Research. 33:307–317.

Hudomiet, P., M. Hurd, and S. Rohwedder (2018). “The Causal Effects ofEconomic Incentives, Health, and Job Characteristics on Retirement:Estimates Based on Subjective Conditional Probabilities”. RANDworking paper.

Hurd, M. and D. McFadden (1998). “Consumption and Savings Balancesof the Elderly: Experimental Evidence on Survey Response Bias”.In: Frontiers in the Economics of Aging. Ed. by D. Wise. Chicago:University of Chicago Press. 353–387.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 134: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

134 References

Hurwicz, L. and H. Uzawa (1971). “On the Integrability of DemandFunctions”. In: Preferences, Utility and Demand. Ed. by J. Chipman,L. Hurwicz, M. Richter, and H. Sonnenschein. New York: Harcourt,Brace and Jovanovich. 114–148.

Ibragimov, I. and R. Has’minskii (1973). “On the Approximation ofStatistical Estimators by Sums of Independent Variables”. SovietMath. Dokl. 14: 883–887.

Islam, T., J. J. Louviere, and P. F. Burke (2007). “Modeling the effects ofvincluding/excluding attributes in choice experiments on systematicand randon components”. Intern. J. of Reserch in Markeing. 24:289–300.

Jain, A., F. Acito, N. Malhotra, and V. Mahajan (1979). “A Comparisonof the Internal Validity of Alternative Parameter Estimation Meth-ods in Decompositional Multiattribute Preference Models”. Journalof Marketing Research. 16: 313–322.

Joe, H. (1997). Multivariate Models and Dependence Concepts. London:Chapman & Hall.

Johnson, R. (1974). “Trade-off Analysis of Consumer Values”. Journalof Marketing Research. 11: 121–127.

Johnson, R. (1999). “The Joys and Sorrows of Implementing HB Meth-ods for Conjoint Analysis”. Working Paper, Sawtooth Software.

Johnson, R. and B. Orme (1996). “How Many Questions Should YouAsk in Choice-based Conjoint Studies?” Sawtooth Software technicalreport.

Kahneman, D. and A. Krueger (2013). Developments in the Measure-ment of Subjective Well-Being. Cheltenham: Edward Elgar.

Kahneman, D. and A. Tversky (1979). “Prospect Theory: An Analysisof Decisions Under Risk”. Econometrica. 47: 263–291.

Kahneman, D. and A. Tversky (2000). Choices, Values, and Frames.Cambridge: Cambridge University Press.

Kaplow, L. and S. Shavell (2002). Fairness and Welfare. Cambridge:Harvard University Press.

King, G., C. Murray, J. Salomon, and A. Tandon (2004). “Enhancingthe Validity and Cross-cultural Comparability of Measurement inSurvey Research”. American Political Science Review. 98: 191–207.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 135: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

References 135

Kling, C., D. Phaneuf, and J. Zhao (2012). “From Exxon to BP: HasSome Number Become Better Than No Number?” Journal of Eco-nomic Perspectives. 26(4): 3–26.

Koop, G. (2003). Bayesian Econometrics. Chichester: John Wiley andSons.

Krinsky, I. and L. Robb (1986). “On Approximating the StatisticalProperties of Elasticities”. The Review of Economics and Statistics.68: 715–719.

Krinsky, I. and L. Robb (1990). “On Approximating the StatisticalProperties of Elasticities: a Correction”. The Review of Economicsand Statistics. 72: 189–190.

Kuhfield, W., R. Tobias, and M. Garratt (1994). “Efficient ExperimentalDesign with Marketing Research Applications”. Journal of MarketingResearch. 31: 545–557.

Lenk, P., W. DeSarbo, P. Green, and M. Young (1996). “HierarchicalBayes Conjoint Analysis: Recovery of Partworth Heterogeneity fromReduced Experimental Designs”. Marketing Science. 15: 173–191.

Lewandowski, D., D. Kurowicka, and H. Joe. (2009). “Generating Ran-dom Correlation Matrices Based on Vines and Extended OnionMethod”. Journal of Multivariate Analysis. 100: 1989–2001.

List, J. and C. Gallet (2001). “What Experimental Protocol InfluencesDisparities Between Actual and Hypothetical Stated Values?” Envi-ronmental and Resource Economics. 20(3): 241–254.

Ljung, G. and G. Box (1978). “On a Measure of a Lack of Fit in TimeSeries Models”. Biometrika. 65(2): 297–303.

Lleras, J., Y. Masatlioglu, D. Nakajima, and E. Ozbay (2017). “WhenMore is Less: Limited Consideration”. Journal of Economic Theory.170: 70–85.

Louviere, J. (1988). Analyzing Decision Making: Metric Conjoint Anal-ysis. Newbury Park: Sage.

Louviere, J., T. Flynn, and R. Carson (2010). “Discrete Choice Exper-iments Are Not Conjoint Analysis”. Journal of Choice Modeling.3(3): 57–72.

Louviere, J., D. Hensher, and J. Swait (2000). Stated Choice Methods:Analytics and Applications. Cambridge: Cambridge University Press.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 136: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

136 References

Louviere, J. and G. Woodworth (1983). “Design and Analysis of Simu-lated Consumer Choice or Allocation Experiments: An ApproachBased on Aggregate Data”. Journal of Marketing Research. 20: 350–367.

Luce, D. (1959). Individual Choice Behavior. Hoboken: John Wiley andSons.

Luce, D. and P. Suppes (1965). “Preferences, Utility and SubjectiveProbability”. In: Handbook of Mathematical Psychology. Ed. by R.Luce, R. Bush, and E. Galanter. Hoboken: John Wiley and Sons.249–410.

Luce, D. and J. Tukey (1964). “Simultaneous Conjoint Measurement: ANew Type of Fundamental Measurement”. Journal of MathematicalPsychology. 1: 1–27.

Manski, C. (1999). “Analysis of Choice Expectations in IncompleteScenarios”. Journal of Risk and Uncertainty. 19: 49–66.

Manski, C. (2004). “Measuring Expectations”. Econometrica. 72: 1329–1376.

Manski, C. (2007). “Minimax-regret Treatment Choice with MissingOutcome Data”. Journal of Econometrics. 139: 105–115.

Marsaglia, G. (1972). “Choosing a Point from the Surface of a Sphere”.Annals of Mathematical Statistics. 43(2): 645–646.

Marschak, J. (1960). “Binary Choice Constraints on Random UtilityIndicators”. In: Stanford Symposium on Mathematical Methods inthe Social Sciences. Ed. by K. Arrow. Palo Alto: Stanford UniversityPress.

McFadden, D. (1974). “Conditional Logit Analysis of Qualitative ChoiceBehavior”. In: Frontiers of Econometrics. Ed. by P. Zarembka. NewYork: Academic Press. 105–142.

McFadden, D. (1981). “Econometric Models of Probabilistic Choice”.In: Structural Analysis of Discrete Data. Ed. by C. Manski and D.McFadden. Cambridge: MIT Press. 198–272.

McFadden, D. (1986). “The Choice Theory Approach to Market Re-search”. Marketing Science. 5: 275–297.

McFadden, D. (1994). “Contingent Valuation and Social Choice”. Amer-ican Journal of Agricultural Economics. 76: 689–708.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 137: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

References 137

McFadden, D. (1996). “Lectures on Simulation-assisted Statistical Infer-ence”. Conference Presentation, EC-squared Conference, Florence,Italy; and Working Paper, Department of Economics, University ofCalifornia, Berkeley.

McFadden, D. (1999). “Rationality for Economists?” In: Elicitationof Preferences. Ed. by B. Fischoff and C. Manski. Boston: KAP.73–106.

McFadden, D. (2001). “Disaggregate Behavioral Travel Demand’s RUMSide: A Thirty Year Retrospective”. In: The Leading Edge of TravelBehavior Research. Ed. by D. Hensher. Oxford: Pergamon Press.

McFadden, D. (2004). “Welfare Economics at the Extensive Margin:Giving Gorman Polar Consumers Some Latitude”. Working Paper,University of California, Berkeley.

McFadden, D. (2006). “Revealed Stochastic Preference: A Synthesis”.In: Rationality and Equilibrium. Ed. by C. Aliprantis, R. Matzkin,and D. McFadden. Berlin: Springer.

McFadden, D. (2012). “Economic Juries and Public Project Provision”.Journal of Econometrics. 166(1): 116–126.

McFadden, D. (2014a). “An Economist’s Perspective on EnvironmentalDamages”. The Eighth Annual Advanced Conference on LitigatingNatural Resource Damages. Law Seminars International.

McFadden, D. (2014b). “The New Science of Pleasure: Consumer Be-havior and the Measurement of Well-being”. In: Handbook of ChoiceModelling. Ed. by S. Hess and A. Daly. Cheltenham: Edward Elgar.7–48.

McFadden, D. (2015). A Primer on Bayes Method for Decision-makingand Statistical Analysis. Lecture notes, University of California,Berkeley.

McFadden, D. (2017). “Stated Preference Methods and their Use inEnvironmental Use and Non-use Valuation”. In: Contingent Valua-tion of Environmental Goods. Ed. by D. McFadden and K. Train.Cheltenham: Edward Elgar. 153–187.

McFadden, D. (2018). “Parametric Approximations of PreferenceFields”. Working paper, University of California, Berkeley.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 138: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

138 References

McFadden, D., M. Ben-Akiva, A. Goett, and D. Bolduc (1986). “TheChoice Theory Approach to Analyze the Preferences and Choices ofElectric Utility Consumers”. EPRI Final Report RP2671–1.

McFadden, D., A. Goett, and C.-K. Woo (1988). “Estimating HouseholdValue of Electric Service Reliability with Market Research Data”.Energy Journal: Special Electricity Reliability Issue. 9: 105–120.

McFadden, D. and P. Ruud (1994). “Estimation by Simulation”. Reviewof Economics and Statistics. 76: 591–608.

McFadden, D. and K. Train (2000). “Mixed MNL Models for DiscreteResponse”. Journal of Applied Econometrics. 15: 447–470.

McFadden, D. and K. Train (2017). Contingent Valuation of Environ-mental Goods: A Comprehensive Critique. Cheltenham: EdwardElgar.

McFadden, D. and K. Train (2018). “Welfare Economics in ProductMarkets”. Working paper, University of California, Berkeley.

Miller, K., R. Hofstetter, H. Krohmer, and J. Zhang (2011). “HowShould Consumers’ Willingness to Pay be Measured? An EmpiricalComparison of State-of-the-art Approaches”. Journal of MarketingResearch. 48(1): 172–184.

Moore, W. (2004). “A Cross-validity Comparison of Rating-based andChoice-based Conjoint Analysis Models”. International Journal ofResearch in Marketing. 21: 299–312.

Morikawa, T., M. Ben-Akiva, and D. McFadden (2002). “Discrete ChoiceModels Incorporating Revealed Preferences and Psychometric Data”.In: Econometric Models in Marketing. Ed. by P. Franses and A.Montgomery. Amsterdam: Elsevier. 29–55.

Moscati, I. (2007). “Early Experiments in Consumer Demand Theory:1930–1970”. History of Political Economy. 39: 359–401.

Muller, M. E. (1959). “A Note on a Method for Generating PointsUniformly on N-dimensional Spheres”. Comm. Assoc. Comput. Mach.2: 19–20.

Natter, M. and M. Feurstein (2002). “Real World Performance of Choice-based Conjoint Models”. European Journal of Operations Research.117: 448–458.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 139: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

References 139

Neal, R. (2011). “MCMC using Hamiltonian Dynamics”. In: Handbookof Markov Chain Monte Carlo. Ed. by S. B. et al. Ch. 5. London:Chapman and Hall.

Neslin, S. (1981). “Linking Product Features to Perceptions: Self-statedversus Statistically Revealed Importance Weights”. Journal of Mar-keting Research. 18: 80–86.

Orme, B. (1999). “ACA, CBC, or Both? Effective Strategies for ConjointResearch”. Working paper, Sawtooth Software.

Pesaran, H. and M. Weeks (2001). “Non-nested Hypothesis Testing: AnOverview”. In: A Companion to Theoretical Econometrics. Ed. byB. Baltagi. Oxford: Blackwell. 277–309.

Pollard, D. (1984). Convergence of Stochastic Processes. New York:Springer.

Press, W., W. Vetterling, S. Teukolsky, and B. Flannery (2007). Numer-ical Recipes: The Art of Scientific Computing. Cambridge UniversityPress.

Quarteroni, A., R. Sacco, and F. Saleri (2007). Numerical Mathematics.New York: Springer.

Raghavarao, D., J. Wiley, and P. Chitturi (2010). Choice-Based ConjointAnalysis: Models and Designs. Boca Raton: CRC Press.

Ramsey, F. (1931). “Truth and Probability”. In: The Foundations ofMathematics and Other Logical Essays. Ed. by R. Braithwaite. Lon-don: Kegan, Paul, Trench, Trubner and Co. 156–98.

Rao, V. (1977). “Conjoint Measurement in Marketing Analysis”. In:Multivariate Methods for Market and Survey Research. Ed. by J.Sheth. Chicago: American Marketing Association. 257–286.

Rao, V. (2014). Applied Conjoint Analysis. Berlin: Springer.Rawls, J. (1971). A Theory of Justice. Cambridge: Harvard University

Press.Reibstein, D., J. Batesonm, and W. Boulding (1988). “Conjoint Analysis

Reliability: Empirical Findings”. Marketing Science. 7: 271–286.Revelt, D. and K. Train (1998). “Mixed Logit with Repeated Choices:

Households’ Choices of Appliance Efficiency Level”. The Review ofEconomics and Statistics. LXXX(4): 647–657.

Rose, J. and M. Bleimer (2009). “Constructing Efficient Stated ChoiceExperimental Designs”. Transport Reviews. 29(5): 587–617.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 140: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

140 References

Rose, J., S. Hess, M. Bliemer, and A. Daly (2015). “The Impact ofVarying the Number of Repeated Choice Observations on the MixedMultinomial Logit Model”. Working paper, University of Leeds.

Rossi, P. (1979). “Vignette Analysis: Uncovering the Normative Struc-ture of Complex Judgments”. In: Qualitative and Quantitative SocialResearch: Papers in Honor of Paul Lazarsfeld. Ed. by R. Merton,J. Coleman, and P. Rossi. New York: Macmillan. 175–188.

Rossi, R., G. Allenby, and R. McCulloch (2005). Bayesian Statisticsand Marketing. Hoboken: John Wiley & Sons Ltd.

Rubinstein, R., G. Samorodnitsky, and M. Shaked (1985). “AntitheticVariates, Multivariate Dependence and Simulation of StochasticSystems”. Management Science. 31(1): 66–77.

Samuelson, P. (1947). Foundations of Economic Analysis. Cambridge:Harvard University Press, 1983.

Samuelson, P. (1950). “The Problem of Integrability in Utility Theory”.Economica. 17: 355–385.

Savage, L. (1954). The Foundations of Statistics. Hoboken: John Wiley& Sons Ltd.

Scarpa, R., M. Thiene, and K. Train (2008). “Utility in Willingness toPay Space: A Tool to Address Confounding Random Scale Effects inDestination Choice to the Alps”. American Journal of AgriculturalEconomics. 90: 994–1010.

Schkade, D. and J. Payne (1993). “Where Do the Numbers ComeFrom? How People Respond to Contingent Valuation Questions”. In:Contingent Valuation: A Critical Assessment. Ed. by J. Hausman.Amsterdam: North-Holland. 271–293.

Schultz, H. (1925). “The Statistical Law of Demand as Illustrated bythe Demand for Sugar”. Journal of Political Economy. 33: 481–504.

Schultz, H. (1928). Statistical Laws of Demand and Supply with SpecialApplication to Sugar. Chicago: University of Chicago Press.

Segal, M. (1982). “Reliability of Conjoint Analysis: Contrasting DataCollection Procedures”. Journal of Marketing Research. 19: 139–143.

Song, X., F. Becker, M. Danaf, B. Atasoy, and M. Ben-Akiva (2018).“Enhancement of Hierarchical Bayes Procedure for Logit Mixture”.Working paper, MIT.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 141: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

References 141

Srinivasan, V. (1988). “A Conjunctive-compensatory Approach to theSelf-explication of Multiattributed Preferences”. Decision Sciences.19: 295–305.

Strasser, H. (1975). “The Asymptotic Equivalence of Bayes and Max-imum Likelihood Estimates”. Journal of Multivariate Analysis. 5:206–226.

Swait, J. and W. Adamowicz (2001). “The Influence of Task Complexityon Consumer Choice: A Latent Class Model of Decision StrategySwitching”. Journal of Consumer Research. 28: 135–148.

Thurstone, L. (1927). “A Law of Comparative Judgment”. PsychologicalReview. 34: 273–286.

Thurstone, L. (1931). “The Indifference Function”. Journal of SocialPsychology. 2: 139–167.

Tierney, L. (1994). “Markov Chains for Exploring Posterior Distribu-tions”. Annals of Statistics. 22: 1701–1728.

Train, K. (1986). Qualitative Choice Analysis. Cambridge: MIT Press.Train, K. (1999). “Mixed Logit Models for Recreation Demand”. In:

Valuing Recreation and the Environment. Ed. by J. Herriges andC. Kling. Northampton: Edward Elgar.

Train, K. (2000). “Halton Sequences for Mixed Logit”. Working paper,University of California, Berkeley. url: https://eml.berkeley.edu/wp/train0899.pdf.

Train, K. (2001). “A Comparison of Hierarchical Bayes and MaximumSimulated Likelihood for Mixed Logit”. Working paper, Universityof California, Berkeley. url: https://eml.berkeley.edu/~train/compare.pdf.

Train, K. (2007). “A Recursive Estimator for Random Coefficient Mod-els”. Working paper, University of California, Berkeley. url: http://eml.berkeley.edu/~train/re.pdf.

Train, K. (2009). Discrete Choice Methods with Simulation (2nd edition).Cambridge: Cambridge University Press.

Train, K. and G. Sonnier (2005). “Mixed Logit with Bounded Distri-butions of Correlated Partworths”. In: Applications of SimulationMethods in Environmental and Resource Economics. Ed. by A. Al-berini and R. Scarpa. Dordrecht: Springer. 117–134.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 142: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

142 References

Train, K. and M. Weeks (2005). “Discrete Choice Models in PreferenceSpace and Willingness-to-pay Space”. In: Applications of Simula-tion Methods in Environmental and Resource Economics. Ed. byA. Alberini and R. Scarpa. Dordrecht: Springer. 1–17.

Tversky, A. (1972). “Elimination by Aspects”. Psychological Review.79(4): 281–299.

Urban, G., J. Hauser, W. Qualls, B. Weinberg, J. Bohlmann, and R.Chicos (1997). “Validation and Lessons from the Field: Applicationsof Information Acceleration”. Journal of Marketing Research. 34(1):143–153.

Urban, G., J. Hauser, and J. Roberts (1990). “Prelaunch Forecastingof New Automobiles: Models and Implementation”. ManagementScience. 36: 401–421.

Walker, J. L., Y. Wang, M. Thorhauge, and M. Ben-Akiva (2018). “D-efficient or Deficient? A Robustness of SP Experimental Design in aVOT Estimation Context”. Theory and Decision. 84(2): 215–238.

Walker, J., M. Ben-Akiva, and D. Bolduc (2007). “Identification ofParameters in Normal Error Component Logit-mixture (NECLM)Models”. Journal of Applied Econometrics. 22: 1095–1125.

Wallis, A. and M. Friedman (1942). “The Empirical Derivation ofIndifference Functions”. In: Studies in Mathematical Economics andEconometrics. Ed. by O. Lange, F. McIntyre, T. Yntema, and H.Schuktz. Chicago: University of Chicago Press. 175–89.

Wittink, D. and T. Bergestuen (2001). “Forecasting with ConjointAnalysis”. In: Principles of Forecasting: A Handbook for Researchersand Practitioners. Ed. by J. Armstrong. New York: Springer. 147–167.

Wittink, D. and D. Montgomery (2009). “Models of Job Preferencefor Stanford MBA’s ’78”. Stanford University Graduate School ofBusiness Research Paper No. 2018.

Wright, P. and M. Kriewall (1980). “State-of-mind Effects on Accuracywith Which Utility Functions Predict Marketplace Utility”. Journalof Marketing Research. 17: 277–293.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 143: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

References 143

Wu, M. and W. Fitzgerald (1996). “Bayesian Multimodal EvidenceComputation by Adapti Tempering MCMC”. In: Maximum Entropyand Bayesian Methods. Ed. by K. Hanson and R. Silver. Dordrecht:Springer.

Yadav, L. (2007). Hypothetical Bias in Contingent Valuation. MountainView: Proquest.

Yanez, M., E. Cherchi, B. Heydecker, and J. Dios-Ortuzar (2011). “Onthe Treatment of Repeated Observations in Panel Data: Efficiency ofMixed Logit Parameter Estimates”. Networks and Spatial Economics.11(3): 393–418.

Yang, M. (2008). “Normal Log-normal Mixture: Leptokurtosis andSkewness”. Applied Economics Letters. 15: 737–742.

Zellner, A. and P. Rossi (1984). “Bayesian Analysis of DichotomousQuantal Response Models”. Journal of Econometrics. 25: 365–394.

The version of record is available at: http://dx.doi.org/10.1561/0800000036

Page 144: Foundations of Stated Preference Elicitation: Consumer ...train/foundations.pdf · Foundations of Stated Preference Elicitation: Consumer Behavior and Choice-based Conjoint Analysis

144 References

Web Sites

https://github.com/stan-dev/rstan/wiki/RStan-Getting-Startedhttp://mc-stan.org/http://www.slideshare.net/surveyanalytics/how-to-run-a-conjoint-analysis-project-in-1-hourhttp://artax.karlin.mff.cuni.cz/r-help/library/conjoint/html/00Index.htmlhttp://cran.r-project.org/web/packages/conjoint/index.htmlhttp://www.lawseminars.com/detail.php?SeminarCode=14NRDNM#agendahttps://eml.berkeley.edu/~train/foundations_R.txthttps://eml.berkeley.edu/~train/software.htmlhttps://www.sheffield.ac.uk/economics/people/hole/stata

The version of record is available at: http://dx.doi.org/10.1561/0800000036