You are on page 1of 49

!

SELECTED'READINGS'IN''
CONSUMER)NEUROSCIENCE)&)
NEUROMARKETING)
!
2nd$edition$

!
Compiled$by$
Thomas$Zoëga$Ramsøy$$
2014

CONSUMER PREFERENCE: WANTING & LIKING

A"recurring"theme"in"this"compendium"is"whether"choices"are"conscious"or"unconscious."Both"
from"psychology"and"neuroscience,"we"know"that"we"have"(at"least)"two"different"motivation"
systems."Some"scholars"have"distinguished"between"a"wanting"system"and"a"liking"system."The"
wanting'system"operates"unconsciously,"and"is"known"to"reflect"the"operations"of"a"separate,"
deep"brain"system,"such"as"the"basal"ganglia."

Conversely,"the"liking'system"is"related"to"our"overt,"conscious"hedonic"experience,"and"is"
reflected"in"our"verbal"utterances."This"system"is"thought"to"be"based"on"the"brain’s"prefrontal"
regions,"such"as"the"orbitofrontal"cortex,"and"possibly"the"anterior"insula.

In"most"cases,"what"you"want"is"what"you"like."But"in"some"instances,"those"processes"break"
down."Consumer"choice"is"packed"with"such"examples."Think"of"the"choice"of"eating"a"
chocolate"while"on"a"diet."You"feel"the"urge"for"a"snack"and"the"sugar"boost,"but"consciously"
think"it’s"a"bad"thing"to"do."Which"of"the"two"will"win?"

This"is"a"tell"tale"sign"of"consumer"motivation,"and"why"the"whole"concept"of"neuromarketing"
and"consumer"neuroscience"makes"sense."By"only"asking"people,"we"only"have"access"to"the"
liking"system."By"exploring"the"hidden"mental"value"operations"of"the"brain,"we"can"also"
understand"the"wanting"system.
ORIGINAL RESEARCH ARTICLE
published: 03 June 2011
doi: 10.3389/fnins.2011.00077

Investigating the role of the ventromedial prefrontal cortex in


the assessment of brands
José Paulo Santos1,2*, Daniela Seixas 3,4,5, Sofia Brandão 6 and Luiz Moutinho 7
1
Superior Institute of Maia, Maia, Portugal
2
Socius – Research Centre in Economic and Organizational Sociology, Lisbon, Portugal
3
Faculty of Medicine, Department of Experimental Biology, University of Porto, Porto, Portugal
4
Instituto de Biologia Molecular e Celular, Porto, Portugal
5
Department of Imaging, Centro Hospitalar de Vila Nova de Gaia e Espinho, Vila Nova de Gaia, Portugal
6
Magnetic Resonance Imaging Unit, Department of Radiology, São João Hospital, Porto, Portugal
7
Department of Management, University of Glasgow, Glasgow, Scotland

Edited by: The ventromedial prefrontal cortex (vmPFC) is believed to be important in everyday preference
Julia Trommershaeuser, New York
University, USA
judgments, processing emotions during decision-making. However, there is still controversy in
Reviewed by:
the literature regarding the participation of the vmPFC. To further elucidate the contribution of
Ming Hsu, University of California the vmPFC in brand preference, we designed a functional magnetic resonance imaging (fMRI)
Berkeley, USA study where 18 subjects assessed positive, indifferent, and fictitious brands. Also, both the
Jamie D. Roitman, University of period during and after the decision process were analyzed, hoping to unravel temporally the
California San Francisco, USA
role of the vmPFC, using modeled and model-free fMRI analysis. Considering together the
*Correspondence:
José Paulo Santos, Instituto Superior
period before and after decision-making, there was activation of the vmPFC when comparing
da Maia, Avenida Carlos Oliveira positive with indifferent or fictitious brands. However, when the decision-making period was
Campos, Castelo da Maia, 4475-690 separated from the moment after the response, and especially for positive brands, the vmPFC
Avioso (São Pedro), Maia, Portugal. was more active after the choice than during the decision process itself, challenging some of
e-mail: jpsantos@ismai.pt
the existing literature. The results of the present study support the notion that the vmPFC may
be unimportant in the decision stage of brand preference, questioning theories that postulate
that the vmPFC is in the origin of such a choice. Further studies are needed to investigate in
detail why the vmPFC seems to be involved in brand preference only after the decision process.
Keywords: neuromarketing, brands, emotion, preference, multivariate analysis, FMRI

INTRODUCTION in the vmPFC persisted in their original choice, ignoring brand


In the last few years several articles were published involving a new information, showing that the vmPFC is also necessary in the inte-
approach to the study of brands using neuroscientific techniques. gration of information in the decision-making process.
One of these first studies used photographs of soft drinks where In addition the vmPFC was found to be important in signaling
brands figured explicitly, inducing preference judgments (Paulus risk probabilities (Tom et al., 2007; Rangel et al., 2008). Fellows and
and Frank, 2003). These authors hypothesized that a specific Farah (2007) again working with patients with vmPFC impairment,
area in the prefrontal cortex, the ventromedial prefrontal cortex suggested that this brain region is necessary for all sorts of choice
(vmPFC), was critical for everyday preference judgments. In fact, tasks, either uncertain (including risky or ambiguous situations),
they found important activations in this brain region when par- or certain.
ticipants selected preferred soft drinks in contrast with a visual However, there is still controversy in the literature regarding the
discrimination task of the same stimuli (liquids contained in bottles function of the vmPFC in decision-making in general, and in brand
or glasses). Also investigating brands, Deppe et al. (2005), largely preference in particular. For example, Schaefer and Rotte (2007) did
based on the work of Damásio, Bechara, and co-workers (Damásio, not report activations in this brain region when sport and luxury
1994; Bechara et al., 1997, 1999; Bechara and Damásio, 2005), pro- car brands (rewarding stimuli) were compared with rational choices
posed a dichotomic theory in economic decision-making, “(…) one of car brands. In another study, using fNIRS to compare luxury
chain involving emotional experience (…) and another one based and common handbags assessed individually, Lin et al. (2010) sug-
on reasoning strategies” (p. 180). Deppe et al. (2005) propose the gest that the cognitive subprocesses that underlie the assessment of
vmPFC to be central in the processing of emotions during decision- branded handbags were only important after the choice was made.
making, whereas brain regions associated with working memory To further elucidate the contribution of the vmPFC in brand
could sustain reasoning. preference, we designed a functional magnetic resonance imaging
Koenigs and Tranel (2008) recruited patients with a specific (fMRI) study where subjects assessed positive, indifferent and ficti-
damage of the vmPFC to select soft drinks in two conditions: tious brands, testing the participation of the vmPFC in the process-
blinded or brand-cued. While healthy controls and patients with ing of these different hedonic categories of brands. Moreover, both
damage in other brain areas changed their soda preference from the period during and after the decision process were analyzed,
the blinded drinks to the brand-cued ones, patients with a lesion hoping to unravel temporally the role of the vmPFC.

www.frontiersin.org June 2011 | Volume 5 | Article 77 | 1


Santos et al. The vmPFC in brands’ assessments

The designation vmPFC is ambiguous in the literature. The present


study relies on the probabilistic atlases Harvard-Oxford Cortical
Structural Atlas and Harvard-Oxford Subcortical Structural Atlas pro-
vided by the Harvard Centre for Morphometric Analysis1. We have
considered the vmPFC to include the ventral medial frontal pole,
frontal medial cortex, ventral paracingulate gyrus, ventral anterior
cingulate gyrus, and subcallosal cortex, limited dorsally by the plane
z = +10, and laterally by the planes x = ±20 (MNI152 coordinates).

MATERIALS AND METHODS


GENERAL STRUCTURE
To explore the research question, an event-related fMRI experiment FIGURE 1 | Examples of some of the logos used as fictitious stimuli.
was designed. There were four different stimuli categories, plus the
interstimuli interval. Each category was composed by 35 slides (6 s
each). The interstimuli interval ranged from 4 until 9 s, in 0.5 s NON-EMOTIONAL WORDS
steps. The experiment duration was 1200 s, plus 9 s added in the The fourth stimulus (a second baseline) was non-emotional words:
end to ensure that all of the hemodynamic response was included. determiners, conjunctions, prepositions, or adverbs. Importantly,
The sequence was optimized with Optseq2 software (Athinoula A. nouns or verbs that could evoke emotions, objects, or actions were
Martinos Center for Biomedical Imaging, USA)2. not used. With this stimulus we hoped to avoid meditation during
Three of the four stimuli were brands’ logos grouped in the fol- the fMRI task (Gusnard and Raichle, 2001; Beckmann and Smith,
lowing categories: positive, indifferent, and fictitious brands. The 2005; De Luca et al., 2006), that could cloud possible self-reflexive
fourth stimulus was non-emotional words. During the interstimuli processes elicited by brands (Yoon et al., 2006).
interval participants fixated a cross.
STRUCTURING THE PARADIGM
BRAND SELECTION The structure of the paradigm was the same for all participants. The
In order to select the logos for the positive and indifferent brand paradigm sequences were programmed with SuperLab 4.0 software
categories, participants completed an electronic survey in which (version 4.0.6b; Cedrus Corporation, USA)3.
were shown 200 brand logos, that they had to rate using the pleasure
and arousal dimensions of the PAD – pleasure, arousal, dominance INSTRUCTIONS FOR THE SCANNING SESSION
scale (Russell and Mehrabian, 1977; Mehrabian and De Wetter, 1987; Depending on the type of stimulus visualized, the participants were
Mehrabian, 1995), and the SAM – self assessment manikin, explained instructed to either rate hedonically the brand (as positive, negative,
in detail elsewhere (Morris, 1995; Bradley and Lang, 2007). Self- indifferent, or unknown), to read covertly non-emotional words, or
reporting emotions is a complex task for most individuals, mainly just to fixate a cross. Participants made their choices using a but-
due to the difficulty in verbalizing such inner states (Chamberlain ton box (model Lumina LU400-PAIR; Cedrus Corporation, USA)4.
and Broderick, 2007). SAM is a non-verbal pictorial assessment tech-
nique designed to represent each dimension of the PAD scale associ- HUMAN SUBJECTS
ated with a person’s affective reaction to a certain stimuli. Dominance The participants were 18, 7 healthy male and 11 healthy female
was not included in the brand assessment because with static pictures volunteers, right handed, with neither history of neurological nor
this dimension correlates with pleasure (Bradley and Lang, 2007). psychiatric disturbances (mean age 28.2 ± 6.9 years, 19–41 years).
After this task, the responses were screened and categorized Informed consent was obtained in all cases. A safety form for
according to the following criteria: positive brands if the score magnetic resonance imaging was filled by every participant. After
was ≥7 in the pleasure dimension, and ≥5 in the arousal dimen- each session the participants were debriefed. This research project
sion; indifferent brands if the score was ≥4 and ≤6 in the pleasure was performed according to the Declaration of Helsinki and was
dimension, and ≤5 in the arousal dimension. With this procedure 35 approved by the local Ethics Committee.
positive and 35 indifferent brands were chosen for each participant,
and were randomized to enter the fMRI paradigm. DATA ACQUISITION
Functional images with axial orientation were obtained using a
FICTITIOUS LOGOS T2∗-weighted EPI sequence in a Siemens® Magnetom Trio high
The fictitious brands were brands’ logos that did not exist in the field (3 T) MRI scanner (Siemens AG, Germany; TR = 3000 ms,
market. Each logo was designed by a marketer made to resemble a TE = 30 ms, 64 × 64 matrix, FOV = 192 mm, 3.0 mm axial slices).
real one, making it plausible for the consumer. The fictitious brands The order of acquisition of the slices was interleaved, and they
did not represent a particular type of product. Instead, logos with covered the whole brain. The study consisted in one session where
assorted shapes, colors, and fonts suggesting different products and 407 volumes were acquired. The first four volumes were discarded
services were used (examples in Figure 1). to ensure pulses stabilization.

1
http://www.cma.mgh.harvard.edu http://www.superlab.com
3

2
http://surfer.nmr.mgh.harvard.edu/optseq/ http://www.cedrus.com
4

Frontiers in Neuroscience | Decision Neuroscience June 2011 | Volume 5 | Article 77 | 2


Santos et al. The vmPFC in brands’ assessments

A whole brain anatomical structural scan was acquired also for At the individual level two different strategies of analysis were
each volunteer, using a T1-weighted MPRAGE protocol (256 × 256 used for comparison. The first strategy was a traditional approach
matrix, FOV = 256 mm, 3.0 mm axial slices), for co-registration where the hemodynamic response was investigated during the
purposes. Gradient field mapping was additionally acquired for complete time window of the stimulus (6 s). In the second approach
image quality control. the stimulus duration was divided in two: the period before the
response (decision-making), and the period after the response
IMAGE ANALYSIS (passive period; see Figure 2).
Functional magnetic resonance imaging data processing was car-
ried out using FEAT (FMRI Expert Analysis Tool) version 5.98, General linear model analysis – conventional stimulus analysis
a model-based GLM (general linear model) analysis tool, and Before the scanning session, participants assessed a set of 200 brand
also using probabilistic independent component analysis (PICA; logos, from which the positive and indifferent stimuli were extracted.
Beckmann and Smith, 2004) as implemented in MELODIC Then, during the scanning, participants rated again the brands. In the
(Multivariate Exploratory Linear Decomposition into Independent first model, in which the whole of the stimulus duration was consid-
Components) version 3.09, a model-free analysis tool, both part ered, 13 EVs were included: the three types of stimulus (positive, indif-
of FSL – FMRIB’s Software Library5 (Smith et al., 2004; Woolrich ferent, and fictitious logos) times the four possible ratings (positive,
et al., 2009). indifferent, negative, and unknown), and the non-emotional words.
Most of the assessments were consistent between the two study
General linear model analysis – common procedures sessions, before and during the scanning (see Brand Selection), but
In the FEAT analysis, the following pre-statistics processing was some of the possible combinations received little or even no rat-
applied: motion correction using MCFLIRT (Jenkinson et al., ings. Although all the possibilities were modeled with EVs aiming
2002); slice-timing correction using Fourier-space time-series to explain most of the variance, only those that were consistent
phase-shifting; non-brain removal using BET (Smith, 2002); between sessions, i.e., positive brands that were rated as positive
spatial smoothing using a Gaussian kernel of FWHM 5 mm; during the scanning session (PosPos), indifferent brands that were
grand-mean intensity normalization of the entire 4D dataset rated as indifferent (IndInd), or fictitious logos that were rated as
by a single multiplicative factor; high pass temporal filtering unknown inside the scanner (NoBUnk) were considered in the anal-
(Gaussian-weighted least-squares straight line fitting, with ysis. Hence, at the individual level analysis, stimuli, and baseline were
sigma = 30.0 s). Stimuli were convolved with a gamma func- compared, resulting in the following contrasts: positive > indifferent,
tion with canonical values (phase 0 s, SD 3 s, and mean lag 6 s). positive > unrecognized logos, and indifferent > unrecognized logos.
To account for variations, temporal derivatives were added for
every explanatory variable (EV), in order to achieve a better General linear model analysis – stimulus detailed analysis
fit between the signal and the stimuli convolved hemodynamic In the second model, 25 EVs were considered: the three types of
responses. Time-series statistical analysis was performed using stimulus (positive, indifferent, and fictitious logos), times the four
FILM with local autocorrelation correction (Woolrich et al., possible ratings (positive, indifferent, negative, and unknown),
2001). Registration to high-resolution structural and/or stand- times the two epochs (before and after button pressing), and the
ard space images was done using FLIRT (Jenkinson and Smith, non-emotional words.
2001; Jenkinson et al., 2002). At the individual level and as before, stimuli and baseline
were subtracted, resulting in the following six contrasts (ar: after
5
http://www.fmrib.ox.ac.uk/fsl response; br: before response): positive br > indifferent br, positive

FIGURE 2 | Splitting the duration of the stimulus for one subject. The figure represents the splitting of the first five stimuli of each category (positive, indifferent,
and fictitious logos). Lighter areas represent the period until the response (during decision), and darker areas represent the period after the response (passive
visualization of the stimulus).

www.frontiersin.org June 2011 | Volume 5 | Article 77 | 3


Santos et al. The vmPFC in brands’ assessments

br > unrecognized logos br, indifferent br > unrecognized logos br, RESULTS
positive ar > indifferent ar, positive ar > unrecognized logos ar, and CONSISTENCY IN THE ASSESSMENTS BETWEEN SESSIONS
indifferent ar > unrecognized logos ar. Most of the ratings were coherent from one session to the other.
Results are summarized in Table 1. Five hundred fifty-four fictitious
General linear model analysis – group analysis brands’ logos out of 630 (87.9%) were rated as unknown, 590 posi-
For both models, group analysis was performed with FLAME tive brands out of 630 (93.7%) were rated as positive, and 427 indif-
(FMRIB’s Local Analysis of Mixed Effects) stage 1 and stage 2 with ferent brands out of 630 (67.8%) were again rated as indifferent.
automatic outlier detection (Beckmann et al., 2003; Woolrich et al.,
2004; Woolrich, 2008). At this level, group means were calculated RESPONSE TIME
from the individual level contrasts. The graph in Figure 4 depicts the distribution of the subjects’
Z (Gaussianized T/F) statistic images were thresholded using choices by response time. Response times were shorter with positive
clusters determined by z > 2.3 and a (corrected) cluster significance ratings (1546 ms) than indifferent (2370 ms) or fictitious ratings
threshold of p = 1.00 (Worsley, 2001). Only clusters with more than (2334 ms), suggesting a delayed decision process with the last two
50 voxels survived the threshold. ratings. These differences are significant between positive and indif-
ferent ratings (F426,589,0.01 = 1.702 – p-value < 0.000001), and posi-
Probabilistic independent component analysis tive and fictitious ratings (F553,589,0.01 = 1.708 – p-value < 0.000001),
The following data pre-processing was applied: masking of non- but not significant between indifferent and fictitious ratings
brain voxels, voxel-wise de-meaning of the data, and normaliza- (F553,426,0.01 = 1.004 – p-value = 0.969508).
tion of the voxel-wise variance. Pre-processed data were whitened
and projected into a 164-dimensional subspace using probabilistic GENERAL LINEAR MODEL ANALYSIS
Principal Component Analysis where the number of dimensions In the conventional GLM analysis (the whole of the stimulus dura-
was estimated using the Laplace approximation to the Bayesian tion), the vmPFC was significantly and extensively activate for the
evidence of the model order (Minka, 2000; Beckmann and Smith, contrasts positive versus indifferent or fictitious logos (Figure 5).
2004). The whitened observations were decomposed into sets of Figure 6 represents the stimulus detailed analysis for the same
vectors, which describe signal variation across the temporal domain contrasts. For the period before the response (decision stage) the
(time-courses), the session/subject domain and across the spatial vmPFC tendentiously deactivated. Conversely, after button press-
domain (maps) by optimizing for non-Gaussian spatial source ing, i.e., after the decision was made and while subjects were pas-
distributions using a fixed-point iteration technique (Hyvärinen, sively visualizing the stimulus, the vmPFC was active.
1999). Estimated component maps were divided by the SD of the Four local maxima from the cluster in the vmPFC in the contrast
residual noise and thresholded by fitting a mixture model to the positive versus indifferent in the conventional analysis were selected
histogram of intensity values (Beckmann and Smith, 2004). for further analysis. The parameter estimates of these voxels are rep-
The EVs basic shapes convolved with a gamma function and resented in Figure 7 both for the conventional analysis and for the
including temporal derivatives were concatenated for all the par- stimulus detailed analysis. For the conventional GLM analysis, all
ticipants in the same order that time-courses were entered in the four local maxima significantly activated when positive brands
MELODIC, and the same contrasts used in FEAT were computed. were involved. On the contrary, in the stimulus detailed analysis
The parameter estimates of each spatial independent component there were deactivations, more prominent in the anterior subre-
(164 total) were then calculated and tested using GLM for each case gions (ventral paracingulate gyrus and ventral medial frontal pole).
(see Figure 3), and so the selection of significant spatial independ- After the response, however, the vmPFC was extensively activate.
ent components was based on statistical criteria.
PROBABILISTIC INDEPENDENT COMPONENT ANALYSIS
The 164 ICs yielded by PICA account for 86.95% of the variability.
To select the relevant ICs the criteria were: the z statistics of the
contrast between the parameter estimates of the positive brands
versus the parameter estimates of the indifferent brands, the

Table 1 | Assessments made during the scanning sessions separated


according to the type of stimuli.

Recorded ratings
Stimuli Total
Positive Indifferent Negative Unknown No
FIGURE 3 | Illustration of the application of a GLM analysis to each of the answer
164 independent components yielded by MELODIC. For each IC, 25
independent variables were modeled: the three types of stimulus (positive, Positive 590 29 3 6 2 630
indifferent, and fictitious logos), times the four possible ratings (positive, Indifferent 82 427 74 44 3 630
indifferent, negative, and unknown), times the two epochs (before and after
Fictitious 33 36 2 554 5 630
button pressing), and the non-emotional words.
Total 705 492 79 604 10 1890

Frontiers in Neuroscience | Decision Neuroscience June 2011 | Volume 5 | Article 77 | 4


Santos et al. The vmPFC in brands’ assessments

FIGURE 4 | Relative frequency of response times obtained during the scanning session grouped in 500 ms intervals.

FIGURE 5 | Statistical z maps (unthresholded in the upper row and thresholded in the lower row) for the contrasts positive versus indifferent brands and
positive versus fictitious logos in a conventional GLM analysis. In the unthresholded images the significant clusters are outlined in white (for z > 2.3), and the
vmPFC is outlined in green. Sagittal views for x = −04 (MNI152 coordinates).

fictitious logos, and the non-emotional words had to be superior Neither before the response nor after the response there were
to 2.3 in all the three cases, or inferior to −2.3 in all three cases. This ICs with z value inferior to −2.3. On the contrary, two ICs (#17 and
procedure was implemented in the two situations, before and after #152) had all the considered z values superior to 2.3 in the situation
the response. In this way it was guaranteed that the ICs selected before the response, and four other (#24, #49, #96, and #135) had
would be significantly more active or more deactivated for positive z values superior to 2.3 in the situation after the response. Only IC
brands than in the remaining cases. #24 included brain activations in the vmPFC. The z values for the

www.frontiersin.org June 2011 | Volume 5 | Article 77 | 5


Santos et al. The vmPFC in brands’ assessments

ICs #15 and #22 were significantly positive for the situation after the
response in the contrast with other logos (indifferent or fictitious).
In the situation before the response, IC #132 had significantly nega-
tive z values for the contrasts with logos.

DISCUSSION
Most of the neuroimaging studies involving brands use paradigms
consisting of choices between pairs of brands or products, i.e., both
stimuli are presented simultaneously and subjects have to choose
one or the other. However, the structure adopted in our study is
different, we believe closer to everyday life; each brand is presented
one at a time, meaning that subjects decide about the hedonic value
of a particular brand not by comparison. For example, when a
consumer chooses a product from a supermarket shelf, s/he does
not collect first all the available items and then choose. On the
contrary, there is a previous intention summarized in a concept
named consideration set, or evoked set (Roberts and Lattin, 1991;
Shocker et al., 1991; Petrof and Daghfous, 1996). The consumer
confronts each possibility in the shelf against the consideration
set until one brand/product is preferred. Thus, the process is not
a simple choice among several options, but instead an assessment
of the fit between one option and the inner expectations that were
previously constructed.
Damásio (1994) from his observations in neurologically
impaired patients, proposed that the prefrontal cortex is a crucial
structure in decision-making; the vmPFC in particular is thought
to be important in decisions of preference including preference for
certain brands (Paulus and Frank, 2003; Deppe et al., 2005; Knutson
et al., 2007; Koenigs and Tranel, 2008; Luu and Chau, 2009). The
results of our conventional GLM analysis, which included data
acquired both before and after decision of brand preference, cor-
roborate these findings: activation of the vmPFC was found when
comparing positive with indifferent or fictitious brands. However,
when we dissected the subjects’ responses and isolated the decision-
making period from the moment after the response, we found that,
especially for positive brands, the vmPFC was more active after the
choice than during the decision process itself, challenging some of
the existing literature. And this result was supported both by the
GLM time-split analysis and by the PICA analysis.
FIGURE 6 | Statistical z maps (unthresholded and thresholded) for the During the decision process itself, i.e., before the response, the
contrasts between positive versus indifferent brands and positive versus
fictitious logos in the stimulus detailed analysis. The two top row maps
vmPFC was less active for positive brands than for indifferent or
represent the decision stage (before the response), and the two bottom row fictitious logos. Conversely, the vmPFC was more active after the
maps represent the period after the response (passive visualization). In the brand choice was made. Considering the four local maxima in
unthresholded images the significant clusters are outlined in white (for the vmPFC (the subcallosal cortex, the frontal medial cortex, the
z > 2.3), the vmPFC is outlined in green. Sagittal views for x = −04 (MNI152 ventral paracingulate gyrus, and the ventral medial frontal pole),
coordinates).
although they were also involved in the conventional analysis when
it corresponded to all the period when the stimulus was present,
the same voxels of the vmPFC were deactive during the decision
three cases are reported in Table 2. Three slices of IC #24 are repre- period until the response, but active after the response. This pattern
sented in Figure 8. Besides the vmPFC activation this network also was not found with indifferent brands (that subjects recognized
includes active voxels in the precuneus, posterior cingulate gyrus, as having some meaning to them, but that were not preferred),
right and left anterior divisions of the middle temporal gyrus, and with fictitious logos (visualized for the first time and about which,
deactivation in the occipital fusiform gyrus. likewise, subjects could not have a preformed opinion), and also
Table 2 also reports the z values for all the ICs that encompass with non-emotional words.
at least one of the local maxima voxels considered in Figure 7 (acti- One of the ICs obtained with the multivariate model-free analy-
vated or deactivated). Only IC #24 has this statistic consistently and sis (PICA) was significantly more relevant in the choice of positive
significantly positive (for the situation after response). However, brands than indifferent brands, fictitious logos, or non-emotional

Frontiers in Neuroscience | Decision Neuroscience June 2011 | Volume 5 | Article 77 | 6


Santos et al. The vmPFC in brands’ assessments

FIGURE 7 | Parameter estimates for the stimuli in four local maxima in the participation of the voxel before the response (decision stage). The bar graphs
vmPFC (subcallosal cortex: −6, 32, −10; frontal medial cortex: 2, 36, −14; identified with the suffix ar refer to the participation after the decision instant but
ventral paracingulate gyrus: −2, 48, −2; ventral medial frontal pole: −2, 58, before the stimulus offset. Pos: positive; Ind: indifferent; Fic: fictitious; NEW:
4). The bar graphs identified with the suffix 6 s are the conventional GLM-based non-emotional words (baseline). MNI152 coordinates. Error bars correspond to
analysis of fMRI data. The bar graphs identified with the suffix br refer to the confidence intervals of 95%.

Table 2 | Statistic z for all the ICs that had at least one voxel activated or deactivated among those considered in Figure 7.

IC

15 22 24 41 50 89 104 110 132

Pos > Ind br −0.967 0.227 −1.876 −1.940 −1.588 0.476 −1.581 −0.239 −3.143
Pos > Fic br −2.560 0.329 1.441 2.296 0.471 0.573 −0.463 3.269 −2.497
Pos > NEW br −7.146 −1.021 0.275 2.961 1.417 0.388 0.413 −0.358 −1.753
Pos > Ind ar 4.805 3.136 2.562 −2.241 0.819 3.103 −0.173 1.348 −1.353
Pos > Fic ar 4.423 5.432 5.169 2.282 0.520 1.389 2.588 1.487 −0.448
Pos > NEW ar −4.001 0.693 2.892 −1.790 −1.562 −3.278 −4.711 3.839 −3.542

Pos, positive; Ind, indifferent; Fic, fictitious; NEW, non-emotional words (baseline); br refers to the participation of the voxel before the response (decision stage); ar
refers to the participation after the decision instant but before the stimulus offset.

words. IC 24 showed extensive activations in the vmPFC, among


other brain structures (Figure 8). This network was significantly
more active with preferred brands than with indifferent brands,
fictitious logos, or non-emotional words only after the response,
which reinforces the fact that although important in decisions of
preference, the vmPFC is only so after the decision-making pro-
cess itself. The analysis of the participation of the vmPFC in brain
networks represented in other ICs corroborates this hypothesis,
because none of the ICs had consistent or significant statistics
to support the participation of the vmPFC in the period before
the response.
The results of the present study seem to contradict some
of the existing theories on the role of the vmPFC in the deci-
sion process. On the other hand, our data are supported by
FIGURE 8 | Two views of the network that constitutes the independent Lin et al., (2010) work in which the brand stimuli were also
component #24: sagittal (x = −04), and axial (z = −12). The vmPFC is
presented one at a time, suggesting as well a late participation
outlined in green. Radiological convention. MNI152 coordinates.
of the vmPFC in preference decision-making; or by Li et al.

www.frontiersin.org June 2011 | Volume 5 | Article 77 | 7


Santos et al. The vmPFC in brands’ assessments

(2010) study that used fMRI and the Iowa Gambling Task to In summary, the results of the present study converge in
investigate the neural correlates of decision-making. They supporting the notion that the vmPFC may be unimportant in the
have demonstrated a group of brain regions that included the decision stage concerning brand preference, questioning theories
dorsolateral prefrontal cortex for working memory, and the that postulate that the vmPFC is in the origin of brand choice. To
insula and posterior cingulate cortex for representations of complement our findings, further studies that challenge as well
emotional states. However, the vmPFC was not part neither conventional research design and neuroimaging methodologies are
of the memory nor the emotional networks, but instead was need to investigate in detail why the vmPFC seems to be involved
coupling the two processes. in brand preference only after the decision process.

REFERENCES influences economic decision making. Minka, T. P. (2000). Automatic Choice of Woolrich, M. W., Behrens, T. E.,
Bechara, A., and Damásio, A. R. (2005). J. Neuroimaging 15, 171–182. Dimensionality for PCA. Technical Beckmann, C. F., Jenkinson, M., and
The somatic marker hypothesis: a Fellows, L. K., and Farah, M. J. (2007). The Report 514, Cambridge: MIT Media Smith, S. M. (2004). Multilevel linear
neural theory of economic decision. role of ventromedial prefrontal cortex Lab Vision and Modeling Group. modelling for FMRI group analysis
Games Econ. Behav. 52, 336–372. in decision making: judgment under Morris, J. D. (1995). Observations: SAM: the using Bayesian inference. Neuroimage
Bechara, A., Damásio, H., Damásio, A. R., uncertainty or judgment per se? Cereb. self-assessment manikin – an efficient 21, 1732–1747.
and Lee, G. P. (1999). Different con- Cortex 17, 2669–2674. cross-cultural measurement of emo- Woolrich, M. W., Jbabdi, S., Patenaude,
tributions of the human amygdala Gusnard, D. A., and Raichle, M. E. (2001). tional response. J. Advert. Res. 35, 63–68. B., Chappell, M., Makni, S., Behrens,
and ventromedial prefrontal cortex Searching for a baseline: functional Paulus, M. P., and Frank, L. R. (2003). T. E., Beckmann, C. F., Jenkinson, M.,
to decision-making. J. Neurosci. 19, imaging and the resting human brain. Ventromedial prefrontal cortex acti- and Smith, S. M. (2009). Bayesian
5473–5481. Nat. Rev. Neurosci. 2, 685–694. vation is critical for preference judg- analysis of neuroimaging data in FSL.
Bechara, A., Damásio, H., Tranel, D., Hyvärinen, A. (1999). Fast and robust ments. Neuroreport 14, 1311–1315. Neuroimage 45, S173–S186.
and Damásio, A. R. (1997). Deciding fixed-point algorithms for independ- Petrof, J. V., and Daghfous, N. (1996). Woolrich, M. W., Ripley, B. D., Brady, J.
advantageously before knowing the ent component analysis. IEEE Trans. Evoked set: myth or reality? Bus. Horiz. M., and Smith, S. M. (2001). Temporal
advantageous strategy. Science 275, Neural Netw. 10, 626–634. 39, 72–77. autocorrelation in univariate linear
1293–1295. Jenkinson, M., Bannister, P. R., Brady, Rangel, A., Camerer, C. F., and Montague, modeling of FMRI data. Neuroimage
Beckmann, C. F., Jenkinson, M., and J. M., and Smith, S. M. (2002). P. R. (2008). A framework for study- 14, 1370–1386.
Smith, S. M. (2003). General multilevel Improved optimization for the robust ing the neurobiology of value-based Worsley, K. J. (2001). “Statistical analysis
linear modeling for group analysis in and accurate linear registration and decision making. Nat. Rev. Neurosci. of activation images,” in Functional
FMRI. Neuroimage 20, 1052–1063. motion correction of brain images. 9, 545–556. MRI: An Introduction to Methods,
Beckmann, C. F., and Smith, S. M. (2004). Neuroimage 17, 825–841. Roberts, J., and Lattin, J. (1991). eds P. Jezzard, P. M. Matthews, and
Probabilistic independent compo- Jenkinson, M., and Smith, S. M. (2001). A Development and testing of a model S. M. Smith (New York, NY: Oxford
nent analysis for functional magnetic global optimisation method for robust of consideration set composition. J. University Press), 251–270.
resonance imaging. IEEE Trans. Med. affine registration of brain images. Mark. Res. 28, 429–440. Yoon, C., Gutchess, A. H., Feinberg, F.,
Imaging 23, 137–152. Med. Image Anal. 5, 143–156. Russell, J. A., and Mehrabian, A. (1977). and Polk, T. A. (2006). A functional
Beckmann, C. F., and Smith, S. M. (2005). Knutson, B., Rick, S., Wimmer, G. E., Evidence for a three-factor theory of magnetic resonance imaging study of
Tensorial extensions of independent Prelec, D., and Loewenstein, G. (2007). emotions. J. Res. Pers. 11, 273–294. neural dissociations between brand
component analysis for multisub- Neural predictors of purchases. Schaefer, M., and Rotte, M. (2007). Favorite and person judgments. J. Consum.
ject FMRI analysis. Neuroimage 25, Neuron 53, 147–156. brands as cultural objects modulate Res. 33, 31–40.
294–311. Koenigs, M., and Tranel, D. (2008). reward circuit. Neuroreport 18, 141–145.
Bradley, M. M., and Lang, P. J. (2007).“The Prefrontal cortex damage abolishes Shocker, A., Ben-Akiva, M., Boccara, Conflict of Interest Statement: The
international affective picture system brand-cued changes in cola prefer- B., and Nedungadi, P. (1991). authors declare that the research was con-
(IAPS) in the study of emotion and ence. Soc. Cogn. Affect. Neurosci. 3, 1–6. Consideration set influences on con- ducted in the absence of any commercial
attention,” in Handbook of Emotion Li, X., Lu, Z.-L., D’argembeau, A., Ng, sumer decision-making and choice: or financial relationships that could be
Elicitation and Assessment, eds J. A. M., and Bechara, A. (2010). The iowa issues, models, and suggestions. Mark. construed as a potential conflict of interest.
Coan and J. J. B. Allen (New York: gambling task in fMRI images. Hum. Lett. 2, 181–197.
Oxford University Press), 29–46. Brain Mapp. 31, 410–423. Smith, S. M. (2002). Fast robust auto- Received: 15 December 2010; accepted: 21
Chamberlain, L., and Broderick, A. Lin, C.-H., Tuan, H.-P., and Chiu, Y.-C. mated brain extraction. Hum. Brain May 2011; published online: 03 June 2011.
(2007). The application of physiologi- (2010). Medial frontal activity in Mapp. 17, 143–155. Citation: Santos JP, Seixas D, Brandão
cal observation methods to emotion brand-loyal consumers: a behav- Smith, S. M., Jenkinson, M., Woolrich, M. S and Moutinho L (2011) Investigating
research. Qual. Mark. Res. Int. J. 10, ior and near-infrared ray study. J. W., Beckmann, C. F., Behrens, T. E., the role of the ventromedial prefron-
199–216. Neurosci. Psychol. Econ. 3, 59–73. Johansen-Berg, H., Bannister, P. R., De tal cortex in the assessment of brands.
Damásio, A. R. (1994). Descarte’s Error – Luu, S., and Chau, T. (2009). Decoding Luca, M., Drobnjak, I., Flitney, D. E., Front. Neurosci. 5:77. doi: 10.3389/
Emotion, Reason and the Human Brain. subjective preference from single-trial Niazy, R. K., Saunders, J., Vickers, J., fnins.2011.00077
New York: Penguin Putnam. near-infrared spectroscopy signals. J. Zhang, Y., De Stefano, N., Brady, J. M., This article was submitted to Frontiers
De Luca, M., Beckmann, C. F., De Stefano, Neural Eng. 6, 016003. and Matthews, P. M. (2004). Advances in Decision Neuroscience, a specialty of
N., Matthews, P. M., and Smith, S. M. Mehrabian, A. (1995). Framework for in functional and structural MR image Frontiers in Neuroscience.
(2006). FMRI resting state networks a comprehensive description and analysis and implementation as FSL. Copyright © 2011 Santos, Seixas, Brandão
define distinct modes of long-distance measurement of emotional states. Neuroimage 23(Suppl. 1), S208–S219. and Moutinho. This is an open-access
interactions in the human brain. Genet. Soc. Gen. Psychol. Monogr. 121, Tom, S. M., Fox, C. R., Trepel, C., and article subject to a non-exclusive license
Neuroimage 29, 1359–1367. 339–361. Poldrack, R. A. (2007). The neural between the authors and Frontiers Media
Deppe, M., Schwindt, W., Kugel, H., Mehrabian, A., and De Wetter, R. (1987). basis of loss aversion in decision-mak- SA, which permits use, distribution and
Plassmann, H., and Kenning, P. Experimental test of an emotion- ing under risk. Science 315, 515–518. reproduction in other forums, provided
(2005). Nonlinear responses within based approach to fitting brand Woolrich, M. W. (2008). Robust group the original authors and source are cred-
the medial prefrontal cortex reveal names to products. J. Appl. Psychol. analysis using outlier inference. ited and other Frontiers conditions are
when specific implicit information 72, 125–130. Neuroimage 41, 286–301. complied with.

Frontiers in Neuroscience | Decision Neuroscience June 2011 | Volume 5 | Article 77 | 8


JCPS-00233; No. of pages: 7; 4C:

Available online at www.sciencedirect.com


Journal of
CONSUMER
PSYCHOLOGY
Journal of Consumer Psychology xx (2011) xxx – xxx

A neural predictor of cultural popularity


Gregory S. Berns ⁎, Sara E. Moore
Economics Department and Center for Neuropolicy, Emory University, Atlanta, GA 30322, USA

Received 7 February 2011; accepted 5 May 2011

Abstract

We use neuroimaging to predict cultural popularity — something that is popular in the broadest sense and appeals to a large number of
individuals. Neuroeconomic research suggests that activity in reward-related regions of the brain, notably the orbitofrontal cortex and ventral
striatum, is predictive of future purchasing decisions, but it is unknown whether the neural signals of a small group of individuals are predictive of
the purchasing decisions of the population at large. For neuroimaging to be useful as a measure of widespread popularity, these neural responses
would have to generalize to a much larger population that is not the direct subject of the brain imaging itself. Here, we test the possibility of using
functional magnetic resonance imaging (fMRI) to predict the relative popularity of a common good: music. We used fMRI to measure the brain
responses of a relatively small group of adolescents while listening to songs of largely unknown artists. As a measure of popularity, the sales of
these songs were totaled for the three years following scanning, and brain responses were then correlated with these “future” earnings. Although
subjective likability of the songs was not predictive of sales, activity within the ventral striatum was significantly correlated with the number of
units sold. These results suggest that the neural responses to goods are not only predictive of purchase decisions for those individuals actually
scanned, but such responses generalize to the population at large and may be used to predict cultural popularity.
© 2011 Published by Elsevier Inc. on behalf of Society for Consumer Psychology.

Keywords: fMRI; Neuroeconomics; Neuromarketing; Music; Prediction

Introduction popularity is a valuable skill that also can translate into


economic success.
How can we predict popularity? Although superficially a In the domain of economic goods, traditional approaches to
trivial question, the desire for popularity consumes a great forecasting popularity rely on standard marketing techniques.
portion of the lives of many youths and adults. More than the These include focus groups, questionnaires, simulated choice
superficial teenager's quest for popularity, being popular is a tests, and market tests. More recently, however, the widespread
marker for social status. Consequently, popularity would seem use of neuroimaging has raised the possibility of using
to confer a reproductive advantage in the evolution of the functional magnetic resonance imaging (fMRI) in the marketing
human species, thus explaining its importance to humans. Such process (Ariely & Berns, 2010). Neuroeconomic research
importance extends to economic success as well because goods suggests that activity in reward-related regions of the brain,
and services that are popular command higher prices. Although notably the orbitofrontal cortex and ventral striatum is
there are good economic and evolutionary rationales for predictive of future purchasing decisions of the individuals
pursuing popularity, predicting who or what becomes popular who are scanned (Hare, O'Doherty, Camerer, Schultz, &
is a challenging problem. Even so, the ability to predict Rangel, 2008; Knutson, Rick, Wimmer, Prelec, & Loewenstein,
2007; Plassmann, O'Doherty, & Rangel, 2007; Plassmann,
O'Doherty, Shiv, & Rangel, 2008). For neuroimaging to be
⁎ Corresponding author at: 1602 Fishburne Dr., Emory University, Atlanta, useful in either a marketing or branding application, however,
GA 30322, USA. these neural signals would need to generalize to a larger group
E-mail address: gberns@emory.edu (G.S. Berns). of individuals who themselves were not the direct object of
1057-7408/$ - see front matter © 2011 Published by Elsevier Inc. on behalf of Society for Consumer Psychology.
doi:10.1016/j.jcps.2011.05.001

Please cite this article as: Berns, G.S., & Moore, S.E., A neural predictor of cultural popularity, Journal of Consumer Psychology (2011), doi:10.1016/
j.jcps.2011.05.001
2 G.S. Berns, S.E. Moore / Journal of Consumer Psychology xx (2011) xxx–xxx

brain scanning. Currently, it is unknown whether the neural Moore, & Noussair, 2010). Using fMRI, we found that although
signals of a small group of individuals are predictive of the an individual's musical preferences were strongly correlated
purchasing decisions of the population at large. with activity in the caudate nucleus, the effect of social
Neuroimaging is often touted as a hot new tool for branding information varied between participants. The tendency to
(Lindstrom, 2008). Although branding and advertising have change one's evaluation of a song was positively correlated
been considered in a few neuroimaging papers (Kenning & with activation in the anterior insula and anterior cingulate, two
Plassmann, 2008; Lee, Broderick, & Chamberlain, 2007; regions that are associated with physiological arousal and
Yoon, Gutchess, Feinberg, & Polk, 2006), it is still unknown negative affective states. While this earlier study examined the
whether neuroimaging can prospectively reveal whether a effect of popularity information on individual preferences, here
particular ad or brand campaign will be effective. In a well- we report a longitudinal analysis in which we examine the
known Coke–Pepsi study, participants who described them- relationship between brain responses and popularity of music
selves as Coke-drinkers showed significant activation in the from the other direction: do neural responses to music in an
hippocampus and right DLPFC when they were cued about the fMRI study predict subsequent commercial success of the song
upcoming drink of Coke (McClure et al., 2004). Self-described and artist?
Pepsi-drinkers did not have this response. In the absence of
brand information, there was no significant difference in pref- Material and methods
erence during a test-taste. This study suggested that any dif-
ferences in the neural response to the two brands must be A total of 32 adolescent participants were studied. Five were
culturally derived. Although these results demonstrate that excluded from the fMRI analyses due to either excessive
branding does affect brain responses to nominally similar goods, movement or susceptibility artifacts. Although this was a
the question of whether brand effectiveness can be predicted in relatively high exclusion rate compared to adult studies, it was
advance remains an open question. comparable to previous fMRI studies in children and adoles-
To demonstrate the efficacy of an fMRI study for branding, cents, who tend to move more than adults (Galvan et al., 2006).
three conditions must be met. First, the study participants—i.e. Prior to the experiment, they were screened for the presence of
the cohort of individuals who are actually scanned—should be medical and psychiatric diagnoses, and none were taking
representative of the population that is the target of a brand medications. There were 14 female and 13 male participants
campaign. Second, to truly test whether the neural signals are between the ages of 12 and 17.9 (mean 14.6). Fifteen were
predictive of brand effectiveness, the scanning must be done Caucasian, eight were African-American, one was Hispanic,
before the campaign is launched. Third, metrics of brand and three were “Other.” The primary stimuli used were 15-s clips
effectiveness must be readily available for the target population. from songs downloaded from MySpace.com. Songs were
For example, these might include sales data, web page views, downloaded between October 23 and November 8, 2006. In
downloads, internet searches, etc. Finally, although not strictly a order to minimize the possibility that participants would recog-
condition, it is an open question as to what should actually be nize the artists, songs from unsigned musicians or relatively
scanned during fMRI. If the product can be consumed in the unknown artists were used. A total of 20 songs were downloaded
scanner, then the product itself becomes the target. Alterna- in each of the following genres: Rock, Country, Alternative/
tively, an ad or branding campaign might be presented in the Emo/Indie, Hip-Hop/Rap, Jazz/Blues, and Metal (identified
scanner, in which case an abstract association between an ad by the MySpace category). At the time of download, the number
and a product becomes the scanned target. of times each song had been played was recorded, and this was
One product that meets these requirements is music. used to calculate the popularity of each song among MySpace
Everyone has musical preferences, and most people spend users. Each song was converted from MP3 to WAV format
money on this product. Thus, it is straightforward to find people and a 15-s clip was extracted that included either the hook or
to scan who are representative of the music-consuming public, chorus of the song. These 15-s clips were subsequently used in the
which is almost everyone. Second, the rise of sites like experiment.
myspace.com has created a large repository of music which is At the beginning of each session, individuals' rankings of
largely unadvertised and unbranded. Because much of this musical genres were elicited. Participants were provided with
music is provided directly by the artist, it can be used well in a list of the six musical genres, and were instructed to rank
advance of any ad campaign; moreover, the band is the brand. the genres from 1 (“the type you like the best”) to 6 (“the type
Third, metrics of music success are simple and straightforward: you like the least”). Each participant's top three genres were
downloads, sales, and ticket receipts. Finally, music is ideally subsequently used in the experiment. Emory University's
suited to scanning because the act of listening to it is the same as Institutional Review Board approved all procedures. Individ-
consuming it. Thus, imaging the neural response to music is a uals then entered the scanner, and the total scan time was
direct measure of the consumption experience. Subsequent approximately one hour. The scanning was performed on a
success is then a combination of quality, branding, and Siemens 3T Trio. Each subject received a T1-weighted structural
marketing. image (TR = 2600 ms, TE = 3.93 ms, flip angle = 8, 224 × 256
In a previous study of adolescents, we measured the matrix, 176 sagittal slices, 1 mm cubic voxel size), a DTI scan
interaction of social influence in the form of popularity ratings (TR = 6500 ms, TE = 90 ms, flip angle = 90, FOV = 220 mm,
with the consumption experience of music (Berns, Capra, 128 × 128 matrix, 34 axial slices, 1.7 × 1.7 × 2.5 mm voxel size,

Please cite this article as: Berns, G.S., & Moore, S.E., A neural predictor of cultural popularity, Journal of Consumer Psychology (2011), doi:10.1016/
j.jcps.2011.05.001
G.S. Berns, S.E. Moore / Journal of Consumer Psychology xx (2011) xxx–xxx 3

6 sets of 12 directional b = 1000 and 1 b = 0 images), and 3 each of the subject's top-three genres were presented in random
functional runs of BOLD-weighting (TR = 2000 ms, TE = 31 ms, order throughout the experiment. In 2/3 of the trials, during the
flip angle = 90, FOV = 192 mm, 64 × 64 matrix, 28 axial slices, second listen, the song's popularity was displayed in the 1–5 star
3 mm cubic voxel size). Each individual participated in 60 trials. scaling system. The 40 trials in which the popularity display
Each trial was divided into two stages; in each stage the subject appeared were sequenced randomly among the 60 trials. Only
listened to the same 15-s song clip (Fig. 1a). During stage 1, no brain activation data from the first listening period was used in
popularity information was shown. After listening, subjects the subsequent analysis. As an incentive to accurately reveal their
were required to rate the song based on (a) how familiar it was and song preferences, each subject received a CD with their top-rated
(b) how much they liked it. Both ratings used a 1–5 star scaling songs.
system. To prevent the subject from passively accepting a default Nielsen SoundScan was used as the source of post factum
rating, each rating screen began with 0 stars, which could not be popularity information over the three years since the songs
accepted as a final selection. After the rating was entered, stage 2 were originally chosen for the study. The SoundScan database
of the trial took place. The clip was played again, after which was searched for information on the performers of each of the
the subject provided another likability rating. Twenty songs in 120 songs in the study. Sales data were available for 87 songs,

Fig. 1. Design of popular music experiment and brain regions correlated with the average likability of each song. a) Timing of events for a typical trial. Each song was
played for 15 s. Each participant heard 60 songs from their favorite 3 genres. Following the song, participants rated the song for familiarity and likability. The trial was
then repeated with the average popularity shown on 2/3 of the trials and blocked for 1/3 of the trials. Only the initial listening period (red) was used for subsequent
analyses. The first listening period for each of the songs was modeled separately, for a total of sixty 15 second variable duration events for each participant. Second-
level models for each of the 120 songs were constructed as one-sample t-tests from the contrast images of the first listening period from the first-level model. A third-
level model was then built using the positive contrast images from the second-level model. This model also included a covariate of the first likability rating for each
song, averaged over the participants who heard that song. b) Brain regions positively correlated with the average likability of the song from the third-level model
(p b 0.005, cluster extent ≥ 56, yielding whole-brain FDR b 0.05) were limited to three areas: cuneus, orbitofrontal cortex, and ventral striatum. (For interpretation of
the references to color in this figure legend, the reader is referred to the web version of this article.)

Please cite this article as: Berns, G.S., & Moore, S.E., A neural predictor of cultural popularity, Journal of Consumer Psychology (2011), doi:10.1016/
j.jcps.2011.05.001
4 G.S. Berns, S.E. Moore / Journal of Consumer Psychology xx (2011) xxx–xxx

and this data was extracted from the SoundScan database during tail (Fig. 2a), with only three songs' albums meeting the
May 2010. The obtained metric aggregated all recorded sales of industry standard for “gold” (500,000 units). Given the large
the song from release through May 2010, including singles, number of songs released annually, this is not surprising. To
albums, and compilations. normalize the distribution, sales data were log-transformed for
Preprocessing of the fMRI data was executed in SPM5 subsequent analyses. First, we checked if either of the subjective
(Functional Imaging Laboratory, UCL, London). The preproces- song ratings were predictive of future sales, but neither of the
sing pipeline consisted of slice timing correction, motion two average ratings obtained for each song was correlated with
correction, spatial normalization, and smoothing (with an 8 mm sales data (likability: R = 0.110, p = 0.313; familiarity: R = 0.106,
Gaussian kernel). A first-level GLM was constructed in SPM5 p = 0.330), nor was the average genre ranking of each song
for each of the 27 participants. The first listening period for (p= 0.102). This indicates that simple subjective reports collected
each of the 60 songs was modeled separately, for a total of sixty from study participants may not be good predictors of commercial
15-second variable duration events. All of the second listening success.
periods, also 15 s variable duration events, were collapsed into a Although subjective ratings of songs did not correlate with
single condition for each run. All three variable duration rating future sales, the activation within the NACC did (Fig. 2b). Log
phases of the trial (familiarity, first likability, and second (sales + 1) was significantly correlated with the average activation
likability) were also collapsed into one condition to model the in NACC during the 15-sec listening period (R = 0.32, p = 0.004).
act of rating including the button presses. The motion parameters To see which part of the 15-sec period was responsible for this
were also included in the model as an effect of non-interest. correlation (e.g. initial or final reactions), we tested an alternative
Second-level models for each of the 120 songs were constructed model with the listening period divided into three 5-sec segments.
as one-sample t-tests in SPM5 using contrast images of the first None of these three segments exhibited a greater correlation to
listening period from the first-level model above. Since every sales than the whole 15-sec period. This indicates that the
participant did not hear every song, the number of contrast images mechanism driving the correlation between NACC activity and
in each of these second-level models ranged from 3 to 23. A sales was integrated over the entire listening period.
third-level model, also a one-sample t-test, was then built in To further understand the interrelationship between song
SPM5 using the positive contrast images from the second-level likability, NACC activity, and sales, we constructed a structural
model above. This model also included a covariate of the first equation model (SEM) (Fig. 3a). The SEM was based on known
likability rating for each song, averaged over the participants anatomical connections between the OFC and NACC and their
who heard that song. We used the likability covariate to identify relationships to subjective likability and purchase decisions
ROIs. Statistical thresholds were determined based on the (Chib, Rangel, Shimojo, & O'Doherty, 2009; Knutson, et al.,
estimated smoothness of the 3rd-level contrasts. Using the 2007; Montague & Berns, 2002; O'Doherty, Kringelbach, Rolls,
AlphaSim routine in AFNI, we estimated the combination of Hornak, & Andrews, 2001; Plassmann et al., 2007; Rolls, 2000).
height and extent thresholds that yielded a whole-brain Consistent with this literature, the average song likability had
FDR b 0.05 (10,000 iterations). First, white matter and CSF significant path coefficients to both the OFC and NACC, and
were masked out using the SPM probabilistic gray matter map. because of the direct connection between OFC and NACC, this
With a gray matter probability N 0.6, this results in a mask that path was also significant. However, the final pathway linking the
retains most gray matter while effectively eliminating most brain to album sales was mediated only through the NACC. When
white matter and CSF, which would otherwise inflate the these relationships were visualized, it became clear that “hit”
required cluster size. Second, we used 3dFWHMx to estimate songs did not result from a specific combination of NACC and
the image smoothness from the square root of the masked SPM- OFC activity, but that “non-hits” were associated with a com-
generated ResMS image and input into AlphaSim. Finally, bination of both low OFC and low NACC activity (Fig. 3b). This
using a voxel level threshold of p b 0.005, the extent threshold relationship was quantified through logistic regressions on
that yielded a cluster level alpha of 0.05 was determined to different hit/non-hit thresholds of sales (Fig. 3c). With thresholds
be k ≥ 56 (Logan & Rowe, 2004; Zhang, Nichols, & Johnson, in the range of 15,000 to 35,000 units sold, the logistic model
2009). The 60% gray matter mask was applied to all contrasts achieved reasonable accuracy in correctly classifying hits and
before using these thresholds. Only three brain regions showed non-hits. For example, with a hit-threshold of 15,000 units, the
a significant correlation between activation and average logistic model correctly classified 80% of the non-hits; however,
song likability: cuneus, orbitofrontal cortex (OFC) and ventral this came at a cost of missing true hits (but still correctly classified
striatum/nucleus accumbens (NACC) (Fig. 1b). The activations 30% of the hits). To test the possibility that consistency of brain
for each song were then extracted from these regions of interest responses might also be predictive of sales, we formulated a
(ROIs) for subsequent analyses with song sales data. This model that included a term for the reciprocal of the variance of the
approach ensures that the ROIs are defined independently from brain response across subjects that heard each song; however, this
the variable of interest (song sales). term was not significant [F(1,83) = 0.28, p = 0.597].

Results Discussion

The vast majority of songs in our sample were not com- Our results demonstrate that not only are signals in reward-
mercially successful. The distribution of sales exhibited a long related regions of the human brain predictive of individual

Please cite this article as: Berns, G.S., & Moore, S.E., A neural predictor of cultural popularity, Journal of Consumer Psychology (2011), doi:10.1016/
j.jcps.2011.05.001
G.S. Berns, S.E. Moore / Journal of Consumer Psychology xx (2011) xxx–xxx 5

Fig. 2. Distribution of number of albums sold and correlation with nucleus accumbens activation. a) Song sales were determined by sales data reported by Nielsen
SoundScan from the album release through May 2010. The distribution was positively skewed, indicating that most songs did not sell many units and with a long tail to
the right. Log-transformation of the sales data normalized this distribution for subsequent correlations (inset). b) The log of the number of units sold of a song was
significantly correlated with the average nucleus accumbens [MNI coordinates: 9, 6, − 9; p b 0.005, 59 voxels] activation during the listening of the song (R = 0.32,
p = 0.004). Exclusion of the far left outlier resulted in a decreased correlation but which was still significant (R = 0.27, p = 0.013).

purchase decisions, they are also modestly predictive of accounts for approximately 20% of music sales (www.riaa.com/
population effects. While the nascent field of neuromarketing keystatistics.php). To test whether the musical tastes of our cohort
has made claims to this effect, truly prospective data has been were representative of the population, we compared our cohort's
lacking (Ariely & Berns, 2010). Surprisingly, our data suggest pre-scan genre rankings to the 2009 Nielsen sales by category and
some validity to these claims. Why might this be? If the specific found a significant correlation (Kendall's τ = −0.733, p = 0.05;
cohort of participants in an imaging study is representative of assuming that our hip-hop category is equivalent to Nielsen's
a particular population, then it follows that the results should R&B category), showing that our cohort had similar tastes as the
generalize. When it comes to music, however, it may be difficult national population.
to know on which dimensions of a population to match (e.g. age, Our results also raise the question of why the brain activation
gender, SES, region). The Recording Industry Association of was predictive of future sales but the self-reported likability
America (RIAA) estimates that the age range of our cohort ratings were not. One possibility is that the questions were not

Please cite this article as: Berns, G.S., & Moore, S.E., A neural predictor of cultural popularity, Journal of Consumer Psychology (2011), doi:10.1016/
j.jcps.2011.05.001
6 G.S. Berns, S.E. Moore / Journal of Consumer Psychology xx (2011) xxx–xxx

adequately specified to differentiate future success (Cronbach &


Meehl, 1955). A more detailed debriefing of the reactions to each
song might have yielded better predictive data. For example, a
choice-based conjoint model might have been superior to simple
rankings (Green & Srinivasan, 1990; Griffin & Hauser, 1993).
Another possibility is that above a certain quality threshold,
songs are too similar to prospectively differentiate them, but slight
differences in quality become magnified in superstar markets
(Rosen, 1981). Although our results do not invalidate any of these
approaches, they do suggest that brain imaging can augment
them. The SEM showed that even though likability was not
directly correlated with future sales, the OFC and NACC mod-
erated this relationship as essentially hidden variables within
the brain. Asking an individual how much they like something
requires several cognitive operations, including the initial
processing of the stimulus, referencing similar items with which
the individual has experience, projection of future utility, all of
which may be subject to framing effects of the experiment. In
contrast, brain responses in reward-related regions are likely to
reflect sub-conscious processes and may yield measurements that
are less subject to cognitive strategies. This would be especially
true during the consumption of music, which occurred during the
listening phase of our experiment. Thus, while the act of rating
something requires metacognition, the brain response during the
consumption of the good does not, and the latter may prove
superior to rating approaches.
Although our data raise the possibility of predicting future
popularity in the form of commercial sales, the actual per-
formance of such a model depended on the definition of a “hit.”
The scarcity of true hits (e.g. 500,000 units) in our sample,
underscores the difficulty in evaluating a hit-predictor and
confirms the previously noted shift toward superstars (Krueger,
2005). The logistic model performed well in identifying non-
hits, which may itself be valuable information, but given the
widely varying marketing approaches that are invested in bands
(Vogel, 2007), it is surprising that we found any predictive
power at all. The fact that we used a wide variety of songs in
different genres certainly made the prediction of hits more
difficult. A more focused presentation of songs, perhaps within
a single genre and pre-screened for minimal quality, would
increase the likelihood of hit-prediction. A more targeted group
of study participants that is representative of a particular music
consuming demographic might also increase predictive power.
However, predicting hit-songs may always be a particularly
difficult task. A recent study of internet searches found good
predictability for revenue of movies and video games but less
so for music (Goel, Hofman, Lahaie, Pennock, & Watts, 2010),
Fig. 3. Structural equation model linking likability ratings to albums sold through which makes our results even more surprising and refutes the
OFC and NACC, and logistic regression to categorize “hits.” a) Although the average idea that hits are random (Bielby & Bielby, 1994).
likability of each song was not directly correlated with the number of albums sold, the
relationship was moderated by activity in the OFC and NACC. All path coefficients Our results may also have implications for branding. In
and main effects were significant (pb 0.05) except that linking OFC to albums sold. commercial music, the band is the brand. We calculated com-
b) The sales of each are represented by the area of each circle and plotted as a function mercial success based on the number of units sold, but this
of both OFC and NACC activity. Although the “hits” (big circles) are scattered number included all sources of a particular song. As a result, the
throughout the activation space, there are a higher proportion of “non-hits” (small
sales numbers were dominated by album sales, which of course
circles) toward the lower left. c) To test the ability to correctly categorize “hits” and
“non-hits,” logistic regressions were performed. The threshold of hit vs. non-hit was contain many songs and may been heavily influenced by the
varied from 1000 to 50,000 units. Thresholds in the range 15,000 to 35,000 correctly band/artist reputation (i.e. the band brand). It is hard to know
categorize a sizable fraction of hits while also correctly rejecting most non-hits. what marketing efforts might have been done to promote a

Please cite this article as: Berns, G.S., & Moore, S.E., A neural predictor of cultural popularity, Journal of Consumer Psychology (2011), doi:10.1016/
j.jcps.2011.05.001
G.S. Berns, S.E. Moore / Journal of Consumer Psychology xx (2011) xxx–xxx 7

band. However, it has been estimated that only 10% of new Griffin, A., & Hauser, J. R. (1993). The voice of the customer. Marketing
releases end up making a profit for a record label (Vogel, 2007). Science, 12(1), 1–27.
Hare, T. A., O'Doherty, J., Camerer, C. F., Schultz, W., & Rangel, A. (2008).
Consequently, marketing and branding efforts tend to be Dissociating the role of the orbitofrontal cortex and the striatum in the
minimal until a band shows signs of popularity. With more and computation of goal values and prediction errors. The Journal of Neuroscience,
more artists having access to quality production equipment and 28(22), 5623–5630.
being able to release songs directly to the public, neuroimaging Kenning, P. H., & Plassmann, H. (2008). How neuroscience can inform consumer
research. IEEE Transactions on Neural Systems and Rehabilitation Engineer-
tools may have real utility to help labels decide how to invest
ing, 16(6), 532–538.
limited marketing and branding resources. Knutson, B., Rick, S., Wimmer, G. E., Prelec, D., & Loewenstein, G. (2007).
Neural predictors of purchases. Neuron, 53, 147–156.
Acknowledgments Krueger, A. B. (2005). The economics of real superstars: The market for
rock concerts in the material world. Journal of Labor Economics, 23(1),
1–30.
This work was supported by grants from the National Lee, N., Broderick, A. J., & Chamberlain, L. (2007). What is neuromarketing? A
Institute for Drug Abuse, the National Science Foundation, the discussion and agenda for future research. International Journal of
Air Force Office of Scientific Research, and the Office of Naval Psychophysiology, 63, 199–204.
Research. Lindstrom, M. (2008). Buyology. Truth and lies about why we buy. New York:
Doubleday.
Logan, B. R., & Rowe, D. B. (2004). An evaluation of thresholding techniques
References in fMRI analysis. NeuroImage, 22, 95–108.
McClure, S. M., Li, J., Tomlin, D., Cypert, K. S., Montague, L. M., &
Ariely, D., & Berns, G. S. (2010). Neuromarketing: The hope and hype of Montague, P. R. (2004). Neural correlates of behavioral preference for
neuroimaging in business. Nature Reviews Neuroscience, 11, 284–292. culturally familiar drinks. Neuron, 44, 379–387.
Berns, G. S., Capra, C. M., Moore, S., & Noussair, C. (2010). Neural Montague, P. R., & Berns, G. S. (2002). Neural economics and the biological
mechanisms of the influence of popularity on adolescent ratings of music. substrates of valuation. Neuron, 36, 265–284.
NeuroImage, 49, 2687–2696. O'Doherty, J., Kringelbach, M. L., Rolls, E. T., Hornak, J., & Andrews, C.
Bielby, W. T., & Bielby, D. D. (1994). “All hits are flukes”: Institutionalized (2001). Abstract reward and punishment representations in the human
decision making and the rhetoric of network prime-time program orbitofrontal cortex. Nature Neuroscience, 4(1), 95–102.
development. American Journal of Sociology, 99(5), 1287–1313. Plassmann, H., O'Doherty, J., & Rangel, A. (2007). Orbitofrontal cortex
Chib, V. S., Rangel, A., Shimojo, S., & O'Doherty, J. P. (2009). Evidence for a encodes willingness to pay in everyday economic transactions. The Journal
common representation of decision values for dissimilar goods in human of Neuroscience, 27(37), 9984–9988.
ventromedial prefrontal cortex. The Journal of Neuroscience, 29(39), Plassmann, H., O'Doherty, J., Shiv, B., & Rangel, A. (2008). Marketing actions
12315–12320. can modulate neural representations of experienced pleasantness. Proceedings
Cronbach, L. J., & Meehl, P. E. (1955). Construct validity in psychological tests. of the National Academy of Sciences of the United States of America, 105(3),
Psychological Bulletin, 52, 281–302. 1050–1054.
Galvan, A., Hare, T. A., Parra, C. E., Penn, J., Voss, H., Glover, G., et al. (2006). Rolls, E. T. (2000). The orbitofrontal cortex and reward. Cerebral Cortex, 10,
Earlier development of the accumbens relative to orbitofrontal cortex might 284–294.
underlie risk-taking behavior in adolescents. The Journal of Neuroscience, Rosen, S. (1981). The economics of superstars. American Economic Review, 71
26, 6885–6892. (5), 845–858.
Goel, S., Hofman, J. M., Lahaie, S., Pennock, D. M., & Watts, D. J. (2010). Vogel, H. L. (2007). Entertainment industry economics. A guide for financial
Predicting consumer behavior with Web search. Proceedings of the analysis (7th ed.). Cambridge: Cambridge University Press.
National Academy of Sciences of the United States of America, 107(41), Yoon, C., Gutchess, A. H., Feinberg, F., & Polk, T. A. (2006). A functional
17486–17490. magnetic resonance imaging study of neural dissociations between brand
Green, P. E., & Srinivasan, V. (1990). Conjoint analysis in marketing: New and person judgments. Journal of Consumer Research, 33, 31–40.
developments with implications for research and practice. Journal of Zhang, H., Nichols, T. E., & Johnson, T. D. (2009). Cluster mass inference via
Marketing, 54(4), 3–19. random field theory. NeuroImage, 44, 51–61.

Please cite this article as: Berns, G.S., & Moore, S.E., A neural predictor of cultural popularity, Journal of Consumer Psychology (2011), doi:10.1016/
j.jcps.2011.05.001
8024 • The Journal of Neuroscience, June 9, 2010 • 30(23):8024 – 8031

Behavioral/Systems/Cognitive

Neural Responses to Unattended Products Predict Later


Consumer Choices
Anita Tusche,1,2 Stefan Bode,1,2,3 and John-Dylan Haynes1,2
1Bernstein Center for Computational Neuroscience Berlin, Charité–Universitätsmedizin Berlin, 10115 Berlin, Germany, 2Max-Planck-Institute for Human
Cognitive and Brain Sciences, 04103 Leipzig, Germany, and 3Department of Neurology, Otto-von-Guericke University, 39120 Magdeburg, Germany

Imagine you are standing at a street with heavy traffic watching someone on the other side of the road. Do you think your brain is
implicitly registering your willingness to buy any of the cars passing by outside your focus of attention? To address this question, we
measured brain responses to consumer products (cars) in two experimental groups using functional magnetic resonance imaging.
Participants in the first group (high attention) were instructed to closely attend to the products and to rate their attractiveness. Partici-
pants in the second group (low attention) were distracted from products and their attention was directed elsewhere. After scanning,
participants were asked to state their willingness to buy each product. During the acquisition of neural data, participants were not aware
that consumer choices regarding these cars would subsequently be required. Multivariate decoding was then applied to assess the
choice-related predictive information encoded in the brain during product exposure in both conditions. Distributed activation patterns
in the insula and the medial prefrontal cortex were found to reliably encode subsequent choices in both the high and the low attention
group. Importantly, consumer choices could be predicted equally well in the low attention as in the high attention group. This suggests
that neural evaluation of products and associated choice-related processing does not necessarily depend on attentional processing of
available items. Overall, the present findings emphasize the potential of implicit, automatic processes in guiding even important and
complex decisions.

Introduction selective processing (Peelen et al., 2009). To date, no study has


Brain responses obtained during active evaluation of products directly compared neural responses to attended versus unat-
and explicit deliberation about purchases have been found to tended products, and the impact of spatial attention on the pre-
predict consumer choices (Knutson et al., 2007). Interestingly, diction of economic decisions. Here, we investigated whether
some evidence suggests that automatic brain processes might brain responses predict consumer choices even when products
guide human judgments and choices, even in the absence of ex- are entirely task-irrelevant and presented outside the focus of
plicit deliberation and attention to the choice task. Brain re- attention.
sponses were shown to engage automatically in assessing facial To examine the role of attention in the prediction of product
attractiveness and preferences even when such judgments were choices from brain activity, we performed an experiment with
not part of the designated task (O’Doherty et al., 2003; Kim et al., two different groups of male participants. In each trial, partici-
2007). Likewise, brain activation was reported to reflect prefer- pants were presented with an image of a car while their brain
ences when participants evaluate stimuli with respect to other, responses were measured using functional magnetic resonance
non-preference-related aspects (Lebreton et al., 2009). However, imaging (fMRI). Participants in group 1 (high attention) were
the precise role of stimulus-related attention in mediating such instructed to actively evaluate and rate the attractiveness of each
automatic valuation processes remains unclear. On the one hand, particular car after its presentation (Fig. 1A). Hence, functional
sensory responses to unattended stimuli have been shown be brain responses were acquired while products were task-relevant
strongly reduced (Rees et al., 1997; Martínez et al., 1999; Kastner and in the focus of attention. In contrast, participants in group 2
and Ungerleider, 2000). On the other hand, spatially unattended (low attention), were engaged in a demanding visual fixation task
stimuli have been reported to undergo substantial category- while task-irrelevant images of cars were passively presented out-
side the focus of attention in the background of the screen (Fig.
1B). After the scanning session, participants from both groups
Received Jan. 6, 2010; revised March 16, 2010; accepted April 8, 2010. were instructed to realistically picture themselves in a consumer
This work was funded by the Max Planck Society, the German Research Foundation, and the Bernstein Compu-
tational Neuroscience Program of the German Federal Ministry of Education and Research. We thank Christian
setting where they had to decide on a new car. For each of the
Kalberlah, Chun Siong Soon, and Carsten Bogler for their contributions to this project and Thorsten Kahnt for previously presented products, participants were then asked to
comments on an earlier version of the manuscript. state whether they would like to purchase this car or not (Fig. 1C).
Correspondence should be addressed to John-Dylan Haynes or Anita Tusche, Charité–Universitätsmedizin Berlin, During scanning, participants from both groups were unaware
Bernstein Center for Computational Neuroscience, Haus 6, Philippstrasse 13, 10115 Berlin, Germany. E-mail:
haynes@bccn-berlin.de. or anita.tusche@bccn-berlin.de.
that they would later be asked about their potential purchases. Mul-
DOI:10.1523/JNEUROSCI.0064-10.2010 tivariate pattern classification (Haxby et al., 2001; Kriegeskorte et al.,
Copyright © 2010 the authors 0270-6474/10/308024-08$15.00/0 2006; Norman et al., 2006; Haynes et al., 2007) was then applied
Tusche et al. • Attention-Independent Prediction of Consumer Choices J. Neurosci., June 9, 2010 • 30(23):8024 – 8031 • 8025

Figure 1. Experimental paradigms. A, In each trial, participants in group 1 (high attention) were presented with images of single cars for 2.4 s followed by a randomized response-button mapping
screen (displayed for a variable duration of 7.2–12 s). Participants were instructed to closely attend to products and to actively evaluate the attractiveness of the particular car on a four-point scale
by pressing the corresponding response button. B, Participants in group 2 (low attention) were asked to perform a demanding fixation task. They responded to the opening of a centrally presented
fixation square (every 800 ms) with a corresponding left- or right-hand button press. The timing parameters applied in group 1 were retained. Every 7.6 –12 s, a car was passively presented on the
background of the screen for 2.4 s while the fixation task continued. C, After the scanning session, participants from both groups were instructed to realistically picture themselves in a consumer
setting where they had to decide upon a new car. Participants were then asked to state their willingness to buy each product. During scanning, participants of both groups were not informed that
such a consumer choice would subsequently be required.

to the brain responses obtained during product exposure to pre- Experimental paradigms. During each trial of both event-related fMRI
dict their stated product choices. paradigms, participants were presented with a single image of a car for
The present study examined the impact of attention on neural 2.4 s. Participants in group 1 (high attention) were instructed to judge the
predictors of economic decision making. More precisely, we in- subjective attractiveness of the specific car on a four-point-scale via a
vestigated the predictability of consumer choices from brain re- button-press (1 ! “dislike it a lot,” 2 ! “dislike it a little,” 3 ! “like it a
sponses for unattended products. A successful prediction would little,” 4 ! “like it a lot”). The mapping of buttons to attractiveness values
was randomized on a trial-by-trial basis to avoid motor preparation
indicate that product valuation and associated choice-related
during product exposure. Thus, each presentation of a car was followed
processing can take place automatically— even with attention
by a randomized response-mapping screen, which was presented for
diverted from products. 7.2–12 s. Responses were given using the index and middle fingers of
both hands operating separate button boxes. Participants in group 2 (low
Materials and Methods attention) were instructed to attend to a black square presented centrally
Participants and stimuli. Group 1 consisted of 17 participants (aged be-
on a white screen. Every 800 ms, the square opened to either the left or the
tween 24 and 32 years) and group 2 of 15 participants (aged between 22
right side. Participants had to respond to each opening with a matching
and 27 years). Participants were healthy male volunteers, had normal or
left or right button press using the index fingers of both hands. After a
corrected-to-normal vision, were right-handed and stated that they were
pseudo-randomized duration of 7.2–12 s, a single image of a car was
interested in cars before the experiment. The sample was limited to males
presented for 2.4 s in the background of the screen while the fixation
interested in cars to ensure that participants were familiar with and pos-
sess stable mental representations about a wide range of items from the task continued. Thus, the visual presentation of cars in group 2 mir-
product group. Both paradigms were approved by the local ethics com- rored that of group 1 but differed in the direction of attention that was
mittee. All participants were paid €12 to take part and all gave written diverted from products. Participants in group 2 were explicitly in-
informed consent. structed to maximize performance on the fixation task throughout
In both fMRI paradigms, monochrome images of 10 real life cars were the entire experiment.
used as stimuli. The selection of cars was based on a behavioral pretest For both paradigms, scanning was performed in a single measurement
with an independent sample of participants in a way that maximized session during which seven independent runs were acquired. The runs
variability in ratings between participants. All pictures were obtained were separated by breaks of "1 min during which no scanning data were
from the Internet and were normalized with regard to size and contrast. obtained. Within a run, each of the 10 cars (see stimuli section above) was
During both fMRI paradigms, images were centrally presented against a presented three times, resulting in a total number of 30 trials per run. For
white background using MATLAB 7.0 (The MathWorks) and the Cogent each run, the presentation order of cars was pseudo-randomized such
toolbox (http://www.vislab.ucl.ac.uk/Cogent). Data from four partici- that the same product was never shown in two consecutive trials. Pseudo-
pants (in group 1 and 2) as well as data for one car (group 1) and two cars randomized durations between presentations of cars varied between 7.2
(group 2) was excluded because of missing variance in consumer choices. to 12 s, meaning that each car was combined with different interstimulus
8026 • J. Neurosci., June 9, 2010 • 30(23):8024 – 8031 Tusche et al. • Attention-Independent Prediction of Consumer Choices

intervals equally often. This procedure ensured that the onset time of the N-dimensional pattern vector. For each run, two pattern vectors were
next stimulus was unpredictable and event-related brain responses were separately created for the purchase and no purchase conditions (see sup-
clearly separable between trials. plemental Fig. 2A, available at www.jneurosci.org as supplemental ma-
Subsequent to the scanning sessions in both groups, participants were terial). Initially, the pattern vectors of six of the seven runs were used for
given a questionnaire with the question “Would you buy this car?” and the training (“training dataset”) of the nonlinear support vector machine
the response options “No/Not sure/Yes.” Importantly, during the acqui- classifier with a fixed regularization parameter C ! 1. This provided the
sition of brain responses participants were unaware that it would be basis of the subsequent classification of the pattern vectors of the remain-
necessary to make such a choice later on. In addition to consumer ing seventh run (“test dataset”) as belonging to either the purchase or no
choices, participants in group 2 had to explicitly judge the attractiveness purchase condition (see supplemental Fig. 2B, available at www.
of each particular car on a four-point scale. Moreover, for both groups jneurosci.org as supplemental material). The procedure was indepen-
familiarity ratings were obtained for all of the products presented. A dently repeated seven times with a different run used as the test dataset to
between-subject design was chosen to ensure that participants in the achieve a sevenfold cross-validation. (A fivefold cross-validation was
“low attention” condition were completely unaware of the task-relevance realized for two participants in group 2 because complete data were
of the products presented. acquired only for 5 of the 7 runs). The amount of purchase-related in-
Image acquisition. For both groups, functional imaging was performed formation of the spatial activation pattern of each spherical cluster was
on a 3-Tesla Siemens TRIO scanner with a standard head coil. T2*- represented by the average decoding accuracy across all cross-validation
weighted functional images were obtained using an echoplanar imaging steps and was assigned to the central voxel vi of the cluster. Given that the
(EPI) sequence [repetition time (TR) ! 2.4 s, echo time ! 30 ms]. For number of considered dimensions (equivalent to number of voxels in a
each run, 152 EPI volumes were collected (36 ascending axial slices per searchlight) exceeded the number of acquired data points, we have to
volume, slice thickness 2 mm, in-plane resolution 3 mm " 3 mm, 1 mm consider that noise was fitted to the training of the classifier. The use of
interslice gap, matrix size 64 " 64). The whole session consisted of seven independent data for training and testing, however, controlled for the
runs. Due to technical problems, only five runs were acquired for two impact of potential overfitting, and consequently for the overestimation
participants in group 2. in the prediction accuracy (Vul et al., 2009). Moreover, to control for
Data analysis. Data from both groups were analyzed in a similar man- “peaking” problems, decoding accuracies were calculated by estimating
ner. In a first step, the acquired volumes were slice timed and realigned. the mean across seven cross-validation steps.
Preprocessed data were then analyzed using a general linear model The described classification was successively performed for all clusters
(GLM) (Friston et al., 1994) as implemented in SPM2 (http://www.fil. created around every measured voxel, resulting in a three-dimensional
ion.ucl.ac.uk/spm). For every run, parameter estimates for “purchase” map of average classification accuracies for each participant. These accu-
and “no purchase” choices were estimated, based on stated consumer racy maps were then spatially normalized to a standard brain [Montreal
choices obtained after the scanning session. Additionally, biometric Neurological Institute (MNI) EPI template as implemented in SPM2]
pulse data acquired during the scanning were included as coregressors of and resampled to an isotropic spatial resolution of 3 " 3 " 3 mm 3.
no interest to account for pulse artifacts. Four regressors were created by Finally, a standard second-level statistical analysis as implemented in
using the integral of the pulse curve of four successive time bins (TR/4 ! SPM2 was performed to identify brain regions that allowed classification
0.6 s) during the product presentation phase. of consumer choices across participants in each group. For both groups,
In the second step, multivariate pattern classification using a support analyses were based on the normalized three-dimensional accuracy maps
vector machine (SVM) was applied to the parameter estimates of the of each subject. Each single point of an individual accuracy map repre-
consumer choices (Cox and Savoy, 2003; Mitchell et al., 2004; Kamitani sented the average decoding accuracy of a searchlight surrounding this
and Tong, 2005; Haynes and Rees, 2006). To realize the classification a position across all cross-validation steps. Thus, each value represented
standard radial basis function kernel as implemented in LIBSVM (http:// the amount of choice-related information of the spatial activation pat-
www.csie.ntu.edu.tw/#cjlin/libsvm) was used. In contrast to a classifier tern of a surrounding spherical cluster. To assess the statistical signifi-
restricted to linear effects, this approach allowed interactions between cance of these accuracy values across subjects, individual accuracy maps
features and nonlinear functions to drive the prediction of subsequent of one group were submitted to a voxelwise one sample t test and con-
choices. A comparative figure illustrating prediction accuracies obtained trasted against chance level. Since the classification was based on two
with a linear classification compared with our nonlinear approach is alternatives (purchase vs no purchase), chance level was 50%. Familywise
provided in supplemental Figure 1 (available at www.jneurosci.org as error (FWE) correction for multiple comparisons was implemented to
supplemental material). control for false positives. Only regions passing this stringent statistical
Multivariate pattern classification techniques take advantage of infor- threshold ( p $ 0.05, FWE-corrected, whole brain) and showing signifi-
mation contained in multiple voxels distributed across space. They allow cant decoding accuracies above chance were considered relevant for in-
investigating whether spatial patterns of brain activation contain stable formation encoding (Haynes et al., 2007; Soon et al., 2008).
information about different experimental conditions (e.g., purchase vs Supplemental data analysis. After each product presentation, partici-
no purchase). To achieve best predictive accuracy, the classifier weights pants in group 1 (high attention) had to rate the attractiveness of the
the contributions of the different voxels optimally. In our case, some particular car on a four-point scale via a button-press. To ensure that
voxels within a searchlight cluster were weighted positively and others brain regions predictive for subsequent consumer choices do not mainly
negatively by the classification algorithm. Moreover, the nonlinear clas- reflect attractiveness, we conducted an additional multivariate decoding
sifier also takes their interactions into account. analysis. Except for the GLM parameter estimates representing attrac-
To ensure an unbiased analysis of the neural activation patterns through- tiveness ratings instead of consumer choices, the multivariate searchlight
out the whole brain, a “searchlight” approach was used (Kriegeskorte et al., decoding was identical to the one described for the main analysis. An-
2006; Haynes et al., 2007). Given that this approach does not depend on other separate decoding analysis was performed on the button-presses to
a priori assumptions about informative brain regions or prior voxel se- confirm that regions predicting product choices did not simply encode
lection, the problem of circular analysis (or “double dipping”) can be subsequent motor responses. Here, parameter estimates of the GLM
avoided (Kriegeskorte et al., 2009). For each participant, a sphere with a were created based on motor responses indicating subjective attractive-
radius of 4 voxels was created around every voxel vi of the measured ness judgments. Apart from that, the analysis was similar to the one
volume. For each sphere, we investigated whether the local pattern of described above to predict consumer choices. In both additional decod-
activation during product exposure predicted the willingness to buy that ing analyses chance level was 25%, because the classification was always
was stated after scanning (purchase vs no purchase). It should be noted based on four alternatives of either attractiveness or button-presses. We
that this analysis is predictive because it uses brain activity during expo- considered only regions showing significant decoding accuracies above
sure to predict purchase ratings obtained after scanning. chance as relevant for encoding of attractiveness and motor responses.
For every run, parameter estimates from the GLM were extracted for Furthermore, classic univariate analyses were conducted for data from
each of the N voxels in the sphere around voxel vi and transformed in an both groups. This enabled us to investigate whether there are single vox-
Tusche et al. • Attention-Independent Prediction of Consumer Choices J. Neurosci., June 9, 2010 • 30(23):8024 – 8031 • 8027

els whose mean activation is significantly more strongly activated in one while 10% of products were unknown. For the remaining 3% in
experimental condition of consumer choices compared with another both groups the participants stated they were unsure whether
(purchase vs no purchase). The functional imaging data were prepro- they were acquainted with the product before the experiment.
cessed using slice time correction, motion correction, were spatially nor- For both groups, these findings confirmed that most of the prod-
malized to a standard stereotaxic space (MNI EPI template), resampled
ucts presented were well known before the experiment.
to an isotropic spatial resolution of 3 ! 3 ! 3 mm 3 and smoothed with a
Gaussian kernel of 6 mm full-width half-maximum. Except for these
differences in the preprocessing, parameter estimates were created as
Task performance and recognition rates
described in the first step of the multivariate decoding approach. Param-
To examine whether attention was effectively removed from
eter estimates of purchase and no purchase choices were then contrasted products in the low attention condition as implemented in group
against each other on a single-subject level. Subsequently, a second-level 2, we analyzed the behavioral performance in the visual fixation
statistical analysis was performed to identify regions that were signifi- task during scanning. The high mean percentage of 89% correct
cantly more strongly activated in one of the two conditions across responses indicates that participants attended to the assigned de-
participants. manding fixation task as instructed. Moreover, we compared the
To test whether engagement of attention in one region of the visual recognition performance in a memory test for both groups (high
field (corresponding to the square of the fixation task in group 2) can and low attention) conducted upon completion of the experi-
strongly decrease the processing of task-irrelevant background stimuli, ment (independent sample of 41 healthy male volunteers: high
we compared the blood oxygenation level-dependent (BOLD) signal
attention group: N $ 20, 26.5 # 2.8 years; low attention group:
change during product exposure of group 2 (low attention) and group 1
(high attention). In the first step, GLM parameter estimates for visual
N $ 21, 25.8 # 1.9 years, mean # SD). Subsequent to the com-
responses during product exposure were created for each participant and pletion of either the high attention or the low attention condi-
contrasted against baseline. Based on these contrasts, the individual tion, participants were presented with single images of 20 cars in
BOLD signal change was estimated for each participant. In the second randomized order. Ten of these cars were stimuli used in the
step, these estimates were used to calculate the average BOLD signal experiment while the remaining 10 images displayed previously
change across participants by implementing a second-level analysis for unseen cars. For each product, participants had to state whether
Groups 1 and 2 individually. Subsequently to contrasting the resulting this car had been presented during the experiment or not. Size
parameter estimates against baseline, we identified the peak activation and position of the products on the computer screen were kept
area of both hemispheres across these two contrast images (located in the constant to facilitate recognition performance. Individual hit-
left [MNI "30, "81, "21] and right visual cortex [MNI 27, "84, "18]).
rates [HR(high attention) $ mean 0.97 # SD 0.07, HR(low at-
Assuming that the visual information was encoded in both hemispheres,
we pooled the signal change from both peak activation areas. This was
tention) $ mean 0.63 # SD 0.26] as well as false-alarm-rates
performed for group 1 and group 2 individually. Finally, a t test for [FAR(high attention) $ mean 0.02 # SD 0.04, FAR(low atten-
independent samples was applied to compare the average signal change tion) $ mean 0.22 # SD 0.18] were found to differ significantly
in group 1 and group 2. between high and low attention conditions (HR: t $ "5.41, p %
0.001; FAR: t $ 4.74, p % 0.001). Participants in the low attention
condition showed lower correct recognition of previously seen
Results cars than participants in the high attention condition. Moreover,
Behavioral results they were more likely to incorrectly rate a previously unseen car
Consumer choices as having been presented during the experiment. These findings
In group 1 (high attention), we obtained functional brain responses strongly suggest that attention was effectively removed from
while participants closely attended to and actively evaluated prod- products in the low attention condition of our experiment.
ucts. In group 2 (low attention), participants performed a dis-
traction task at fixation while task-irrelevant products were
presented outside the focus of attention on the back of the screen. fMRI results
Subsequent to scanning, participants in both groups had to state Multivariate decoding of subsequent consumer choices
In group 1 (high attention), spatial activation patterns in the
their willingness to purchase each particular car. The number of
prefrontal cortex (PFC), namely in the left medial frontal gyrus
trials assigned to either the purchase or no purchase condition
(82% decoding accuracy), the right dorsomedial PFC (75% de-
was found to be well balanced within both groups and compara-
coding accuracy) and the bilateral ventromedial PFC (73% decod-
ble across them. In group 1, the mean distribution of the declared
ing accuracy) predicted subsequent consumer choices. Moreover,
consumer choices for all products was 54% no purchase, 5%
the left insula (73% decoding accuracy) and the right parahip-
“maybe,” and 41% purchase. In group 2, 46% of the products
pocampal gyrus (72% decoding accuracy) were found to contain
were chosen to be no purchase trials, 4% maybe, and 50% purchase
stable information about later product choices (see Table 1 for a
trials across participants. For both groups, profiles of product selec-
complete list of results).
tion were found to vary across participants (see supplemental Table
In group 2 (low attention), activation patterns in the left me-
1, available at www.jneurosci.org as supplemental material). Addi-
dial PFC (76% decoding accuracy) and the bilateral insula (right:
tionally, the results of a t test for independent samples confirmed that
82% decoding accuracy, left: 72% decoding accuracy) predicted
product-specific means (group 1: mean 1.87 # SD 0.83; group 2:
subsequent consumer choices. Neural responses in the left infe-
mean 1.86 # SD 0.93) of consumer choices were comparable across
rior parietal lobe (82% decoding accuracy) and the bilateral su-
both groups (t $ "0.43, p $ 0.67).
perior temporal gyrus (left: 74% decoding accuracy, right: 70%
Familiarity decoding accuracy) also encoded choices between cars (see Table
After scanning, participants from both groups were asked 2 for a complete list of results). It should be noted that decoding
whether they had been familiar with the presented products be- accuracies in brain regions predicting subsequent consumer
fore the experiment. In group 1 (high attention), participants choices under high and low attention conditions were found to be
reported the products were familiar in 85% of all cases and as comparable (Fig. 2).
being unknown in 12%. In group 2 (low attention), participants To provide further evidence for the statistical validity of the
were acquainted with 87% of the products before the experiment results obtained by this approach, an additional decoding analy-
8028 • J. Neurosci., June 9, 2010 • 30(23):8024 – 8031 Tusche et al. • Attention-Independent Prediction of Consumer Choices

Table 1. Brain regions encoding the subsequent consumer choices among actively evaluated products in group 1 (high attention)
Accuracy MNI coordinates
Brain region Side BA M SE t value x y z
Frontal lobe
Medial frontal gyrus L 9 82 3.3 9.49 "21 27 30
mPFC (dorsal) R 10 75 2.6 9.53 9 57 21
mPFC (ventral) L 10 73 2.3 9.96 "9 63 12
Limbic lobe
Insula L 73 2.3 9.80 "36 "6 0
Parahippocampal gyrus R 30 72 1.9 10.98 30 "57 3
Occipital lobe
Inferior occipital lobe L 18 73 2.2 10.44 "15 "105 "6
L 18 70 2.0 9.96 "27 "93 "15
Cerebellum R 77 2.7 9.69 9 "66 "36
Results are reported on a statistical level of p ! 0.05, FWE-corrected; only peak activations of clusters are listed. L, Left hemisphere; R, right hemisphere; M, mean; BA, Brodmann area.

Table 2. Brain regions encoding the subsequent consumer choices among passively presented, task-irrelevant products in group 2 (low attention)
Accuracy MNI coordinates
Brain region Side BA M SE t value x y z
Frontal lobe
Middle frontal gyrus L 8 79 2.8 10.45 "24 36 42
L 8 76 1.8 10.70 "21 36 51
L 8 72 1.9 11.35 "39 33 42
mPFC L 10 77 2.2 12.38 "15 57 24
L 10 76 2.0 12.83 "6 51 18
mOFC L 11 74 2.0 12.20 "12 42 "18
Inferior frontal gyrus L 45 73 2.1 10.90 "51 24 15
Limbic lobe
Insula R 82 2.6 5.19 39 "3 15
L 72 1.8 12.46 "42 15 "6
Parietal lobe
Inferior parietal lobe L 40 82 2.7 11.81 "57 "36 27
R 40 73 2.1 10.66 60 "39 30
Temporal lobe
Superior temporal gyrus L 38 74 1.8 13.31 "51 12 "18
R 42 75 1.8 13.40 54 "33 15
R 42 70 1.5 13.62 63 "36 18
Results are reported on a statistical level of p ! 0.05, FWE-corrected; only peak activations of clusters are listed. L, Left hemisphere; R, right hemisphere; M, mean; BA, Brodmann area; OFC, orbitofrontal cortex.

sis was conducted for data of both groups. When test datasets stable information about product choices when smaller searchlights
were randomly assigned to either the purchase or no purchase (radius of 2 voxels) were used, an additional decoding analysis was
condition during the testing phase of the classifier, no statistical performed. At a more liberal statistical threshold of p ! 0.00001
significant prediction of consumer choices could be achieved (uncorrected), this analysis revealed predictive information in the
(FDR and FWE-corrected). This was true for data from both striatum (see supplemental Fig. 4, available at www.jneurosci.org as
groups independently. Moreover, decoding accuracies in brain supplemental material).
regions that were informative in the original analysis were at
chance level (50%) when the test data were randomly allocated to Univariate comparisons of subsequent consumer choices
the conditions (see supplemental Fig. 3, available at www. In both groups, classic univariate comparisons did not reveal any
jneurosci.org as supplemental material). This finding speaks activation differences between products that participants were
against potential methodological concerns such as possible biases willing to purchase and those they were not. This strongly sug-
inherent in the testing procedure and insufficient corrections for gests that multivariate pattern classification is capable of extract-
multiple comparisons. Finally, we investigated whether a combi- ing information which conventional analyses fail to detect.
nation of two informative classifiers [i.e., medial prefrontal cor-
tex (mPFC) and insula] could improve the overall prediction Multivariate decoding of attractiveness judgments
accuracy. Compared with decoding results of single searchlights Participants in group 1 were instructed to judge the subjective
in these areas, the weighted classification using decision values of attractiveness of a particular car after each product presentation.
the first-level decoding enhanced the prediction by 7% and by 5% Spatial activation patterns in the right middle frontal gyrus (47%
respectively. decoding accuracy, [MNI 30, 12, 33]), medial frontal gyrus (43%
Neural activation in the ventral striatum has frequently been decoding accuracy, [MNI 15, 33, 45]) and left orbitofrontal cor-
implicated in financial decision-making, preference-related pro- tex (40% decoding accuracy, [MNI "27, 33, 18]) were found to
cessing of products as well as purchases (Erk et al., 2002; Kuhnen encode attractiveness judgments during product presentation.
and Knutson, 2005; Knutson and Bossaerts, 2007; Knutson et al., Activity in the left (51% decoding accuracy, [MNI "15, 24, 30])
2007; Schaefer and Rotte, 2007a). To investigate whether spatial and right dorsal anterior cingulate cortex (49% decoding accu-
activation patterns in these regions would be found to contain racy, [MNI 15, 24, 30]), left (41% decoding accuracy, [MNI "18,
Tusche et al. • Attention-Independent Prediction of Consumer Choices J. Neurosci., June 9, 2010 • 30(23):8024 – 8031 • 8029

decrease the visual processing of task-


irrelevant background stimuli (cars). To
test this assumption, we compared the
BOLD signal change in the visual cortex
during product exposure in the high and
low attention conditions. Consistent with
our hypothesis, we found a significant dif-
ference in the BOLD signal change evoked
by the presentation of cars in the high at-
tention and low attention group (t test for
independent samples, t # 5.15, p "
0.001). More precisely, the results demon-
strated that the responses in the visual
cortex evoked by the task-irrelevant im-
ages of cars which were presented outside
Figure 2. Brain regions encoding subsequent consumer choices in both groups. Multivariate searchlight decoding (radius of 4 the focus of attention were reduced com-
voxels) was applied to functional brain responses obtained during product exposure to predict subsequent consumer choices pared with when the images are actively
(chance level 50%). Spatial activation patterns in the mPFC and the insula were found to encode these choices when participants
evaluated and attended to. This finding
did not explicitly deliberate on purchases ( p " 0.05, FWE-corrected). Importantly, this applied to situations when participants
closely attended to and actively evaluated products (“high attention” group 1, blue) as well as when products were passively
indicates that the engagement of attention
presented outside the focus of attention (“low attention” group 2, red). The amount of predictive information in the brain in one region of the visual field (corre-
responses was found to be comparably high in both groups. The graph displays mean decoding accuracies and SEs across partici- sponding to the square of the fixation
pants for both regions and both groups. For illustrative purposes, the contrasts are shown at p " 0.0001 (uncorrected) with a task) decreased the processing of task-
cluster threshold of 10 voxels (L indicates left hemisphere). irrelevant products in the background.

Discussion
!45, 12]) and right posterior cingulate cortex (42% decoding The present study investigated the impact
accuracy, [MNI 9, !42, 12]) and left insula (40% decoding accu- of product-related attention on neural predictors of consumer
racy, [MNI !42, 9, !3]) were also found to be predictive of choices. Activation patterns in the insula and the mPFC were
evaluation scores ( p " 0.05, FWE-corrected, chance level 25% found to predict these choices under high and low attention pro-
for attractiveness ratings 1– 4). For a complete list of results, see cessing (Fig. 2). Thus, a close match of predictive brain regions
supplemental Table 3, available at www.jneurosci.org as supple- was revealed independent of spatial attention to products. This
mental material. Importantly, no direct overlap of regions pre- demonstrates that processing of unattended stimuli can proceed
dicting consumer choices and those predicting attractiveness beyond object processing (Peelen et al., 2009) to a stage that even
scores could be observed (see supplemental Fig. 5, available at allows the prediction of consumer choices. Importantly, the
www.jneurosci.org as supplemental material). This finding amount of predictive information in these areas was comparably
strongly supports the notion that the information encoding con- as high when task-irrelevant products were presented outside the
sumer choices is not merely based on evaluative judgments of focus of attention as when they were actively evaluated and at-
attractiveness. tended to. Congruent with the notion of unconscious environ-
mental triggers for automatic processes in consumer settings
Multivariate decoding of button presses (Chartrand, 2005), this indicates that a prediction of economic
To ensure that the neural activity during the product viewing
choices from brain responses does not necessarily depend on
phase was not confounded by the preparation of the motor
attentive processing of products.
response, we applied a randomized response-button mapping.
Notably, these findings were achieved with participants who
This mapping scheme was only presented after the removal of the
were naive about the necessity of a subsequent choice. This is in
product to isolate the motor response from the processing of
line with behavioral findings reporting that consumer-related
product information. As expected, we found no brain region that
goals can automatically be activated, guiding subsequent con-
was informative about the subsequent motor response during the
sumer behavior and choices even outside of conscious awareness
product-viewing phase. This finding strongly suggests that the
(Bargh, 2002; Chartrand, 2005; Dijksterhuis et al., 2005). More-
prediction of subsequent consumer choices was not based on
over, it is consistent with previous data showing that brain re-
motor preparation during product exposure. During the presen-
sponses reflect preference choices when participants evaluate
tation of the response-mapping scheme (subsequent to the prod-
stimuli with respect to other, non-preference-related aspects
uct presentation), the bilateral motor cortex ([MNI 48, !21, 57]
(Kim et al., 2007; Lebreton et al., 2009). The current study goes
and [MNI !42, !21, 54]) predicted the current button-press
beyond these results by demonstrating that brain responses pre-
( p " 0.05, FWE-corrected, chance level 25% for buttons 1– 4).
dict subsequent choices even in the absence of spatial attention to
Importantly, no neural structure predictive of subsequent con-
choice options.
sumer choices contained information about the associated motor
Neural activation in the insula and the mPFC has been shown
responses (see supplemental Fig. 6, available at www.jneurosci.
to predict consumer choices when participants closely attend to
org as supplemental material). This finding indicates that the
products and explicitly deliberate about purchases (Knutson et
prediction of consumer choices was not related to neural pro-
al., 2007). Given that the prediction of consumer choices in the
cesses involved in the execution of motor responses.
present study was achieved in the absence of explicit deliberation
Attention modulation in the visual cortex and without a priori assumptions about informative regions, this
Engagement of attention in one region of the visual field (corre- further supports the role of both areas in economic decision-
sponding to the square of the fixation task) has been suggested to making. Spatial activation patterns in the insula and the mPFC
8030 • J. Neurosci., June 9, 2010 • 30(23):8024 – 8031 Tusche et al. • Attention-Independent Prediction of Consumer Choices

were also reported to predict behavioral choices in the following predictive information was based on physical properties of single
trial of a reward-based decision-making task (Hampton and cars (see supplemental Table 1, available at www.jneurosci.org as
O’Doherty, 2007). In line with this finding, both structures have supplemental material). Results of an additional decoding anal-
been suggested to underlie the neural representation of the ex- ysis also demonstrate that the prediction of consumer choices was
pected reward value (Knutson et al., 2005; Preuschoff et al., 2006; not related to neural processes involved in the preparation of
Hampton et al., 2007; Hare et al., 2008; Rolls et al., 2008), a motor responses (see supplemental Fig. 6, available at www.
concept central to microeconomic and psychological models of jneurosci.org as supplemental material).
decision-making (Knutson and Bossaerts, 2007; Loewenstein et Another point that needs to be addressed concerns the ques-
al., 2008; Rangel et al., 2008). Cultural stimuli such as cars and tion of whether attention was effectively attenuated by engaging
logos of car manufacturers can signal potential social dominance in the distraction task in group 2. Evidence of successful distrac-
or wealth and have been shown to modulate activity in the reward tion comes from differences in memory performance, obtained
circuit (Erk et al., 2002; Schaefer and Rotte, 2007a). It can be for both groups after the experiment. Recognition rates of cars
assumed that products in the present experiment differed in their were strongly decreased in participants who performed the dis-
expected reward value depending on whether they were subse- traction task compared with participants who actively evaluated
quently chosen to be purchased or not (Sharot et al., 2009). We products. Moreover, neural responses in the visual cortex evoked
suggest that distinguishable activation patterns in the insula and by task-irrelevant, unattended images of cars were found to be
the mPFC reflected these different reward values and— even reduced compared with images that were actively evaluated and
without full attention— contributed to subsequent hypothetical fully attended to. This is consistent with previous research dem-
consumer choices. onstrating that the engagement of attention in one region of the
Activation in the mPFC—particularly the ventral part— has visual field strongly decreases the processing of irrelevant back-
also been reported to reflect product-related preference and at- ground stimuli (Rees et al., 1997). It can therefore be considered
tractiveness judgments (Paulus et al., 2003; McClure et al., 2004; that the attention to products in group 2 was strongly diminished
Deppe et al., 2005; Plassmann et al., 2008; Lebreton et al., 2009; compared with group 1. As in most attention studies, there is a
Luu and Chau, 2009). The dorsal mPFC, on the other hand, has possibility of weak but residual attention to unattended target
been proposed to be involved in the processing of brand knowl- stimuli. However, our key finding is that a strong reduction in
edge (Schaefer et al., 2006; Schaefer and Rotte, 2007b), the impact attention does not affect the choice-predictive information.
of which on consumer preferences and choices is well known Purchase-related information encoded in the brain was not re-
(McClure et al., 2004; Lee et al., 2006). Considering this evidence, duced, but was comparably high in low and high attention con-
it is likely that the predictive information encoded in the mPFC ditions. Furthermore, we found a close match of predictive brain
might have been influenced by subjective valuation and the brand regions during high and low attentional processing of products.
and price information. Together, this indicates that choice-related processing does not
It can be assumed that global product valuation was a major necessarily require close attention to products and can even occur
source for subsequent choices. This is particularly likely, given in cases where sensory signal processing is attenuated due to a
that no further product-related information was provided during removal of attention. Although it remains possible that the choice
the experiment to avoid drawing participants’ attention to the was not explicitly calculated during product presentation, the
potential choice options. Therefore, it could be speculated that high predictive accuracy reveals that choice-related processes
the predictive information merely reflects global attractiveness or have already reached a high level of completion.
desirability of products. However, many cognitive factors as well The present study implemented hypothetical choices, which
as their interactions with automatic valuation processes might are commonly applied in marketing research. Nevertheless, it
contribute to a complex choice such as those for a new car. This is needs to be explored whether the same activation patterns are
particularly likely for participants who stated an interest in cars predictive for actual purchases— beyond the stated willingness to
and are assumed to possess relevant knowledge on the items. buy. This would allow checking for potential biases such as the
Consistent with this notion product choices were not entirely tendency to overstate the willingness to pay or vote in case of
determined by attractiveness judgments. Thus, only partial cor- hypothetical choices (List and Gallet, 2001; Murphy et al., 2005).
relations between product-specific attractiveness ratings and Requiring expenses would likely involve losses and more relative
consumer choices were found for both groups in the present comparisons between available products, including changes in
experiment (see supplemental Table 2, available at www. the reference point for product valuation (FitzGerald et al., 2009;
jneurosci.org as supplemental material). Additionally, no direct Sharot et al., 2009). Real purchases might also engage processes of
overlap of brain regions predicting attractiveness judgments and perceived justification, anticipated regret, time pressure or self
those being informative about consumer choices could be iden- esteem (Plassmann et al., 2007). Finally, it might be that making
tified (see supplemental Table 3 and supplemental Fig. 5, avail- actual purchases differs from our experimental setting in terms of
able at www.jneurosci.org as supplemental material). Together, the strategies used for information acquisition and information
these findings indicate that the prediction of product choices was integration across multiple relevant dimensions. However, it
not mainly due to global evaluations of attractiveness but might should be noted that informative brain regions as identified with
reflect automatic choice-related processing itself. This is in line the present approach are strikingly consistent with previous find-
with previous results showing that brain activation reflecting ings that implemented actual purchases (Knutson et al., 2007).
subsequent preference decisions were not merely responding to Further research might also address the generalizability of our find-
attractiveness of stimuli (Lebreton et al., 2009). However, more ings to other product categories, including different goods (e.g., cars
research is needed to specifically examine choice-relevant dimen- vs coffee) and different types of purchases (e.g., routine vs new), as
sions with the current paradigms, possibly testing explicit models of well as to people who are not interested in the product group.
multiattribute decision-making (Dijksterhuis et al., 2006; Lassiter et In summary, we found a close match of brain regions predicting
al., 2009). Given that consumer choices for each product were found consumer choices for both high and low attentional processing of
to vary from participant to participant, it is also unlikely that the products. Importantly, the amount of predictive information was
Tusche et al. • Attention-Independent Prediction of Consumer Choices J. Neurosci., June 9, 2010 • 30(23):8024 – 8031 • 8031

found to remain persistently high when task-irrelevant products (2009) The deliberation-without-attention effect: evidence for an arti-
were presented outside the focus of attention. Altogether, these find- factual interpretation. Psychol Sci 20:671– 675.
ings support the notion that even complex and important economic Lebreton M, Jorge S, Michel V, Thirion B, Pessiglione M (2009) An auto-
matic valuation system in the human brain: evidence from functional
choices can be prepared automatically, in the absence of explicit neuroimaging. Neuron 64:431– 439.
deliberation and without attention to products. Lee L, Frederick S, Ariely D (2006) Try it, you’ll like it: the influence of
expectation, consumption, and revelation on preferences for beer. Psy-
References chol Sci 17:1054 –1058.
Bargh JA (2002) Losing consciousness: automatic influences on consumer
List JA, Gallet CA (2001) What experimental protocol influence disparities
judgment, behavior, and motivation. J Consum Res 29:280 –285.
between actual and hypothetical stated values. Environ Res Econ 20:241–
Chartrand TL (2005) The role of conscious awareness in consumer behav-
254.
ior. J Consum Psychol 15:203–210.
Loewenstein G, Rick S, Cohen JD (2008) Neuroeconomics. Annu Rev Psy-
Cox DD, Savoy RL (2003) Functional magnetic resonance imaging (fMRI)
chol 59:647– 672.
“brain reading”: detecting and classifying distributed patterns of fMRI
Luu S, Chau T (2009) Decoding subjective preference from single-trial
activity in human visual cortex. Neuroimage 19:261–270.
Deppe M, Schwindt W, Kugel H, Plassmann H, Kenning P (2005) Nonlin- near-infrared spectroscopy signals. J Neural Eng 6:016003.
ear responses within the medial prefrontal cortex reveal when specific Martínez A, Anllo-Vento L, Sereno MI, Frank LR, Buxton RB, Dubowitz DJ,
implicit information influences economic decision making. J Neuroim- Wong EC, Hinrichs H, Heinze HJ, Hillyard SA (1999) Involvement of
aging 15:171–182. striate and extrastriate visual cortical areas in spatial attention. Nat Neu-
Dijksterhuis A, Smith PK, van Baaren RB, Wigboldus DHJ (2005) The rosci 2:364 –369.
unconscious consumer: effects of environment on consumer behav- McClure SM, Li J, Tomlin D, Cypert KS, Montague LM, Montague PR
ior. J Consum Psychol 15:193–202. (2004) Neural correlates of behavioral preference for culturally familiar
Dijksterhuis A, Bos MW, Nordgren LF, van Baaren RB (2006) On making drinks. Neuron 44:379 –387.
the right choice: the deliberation-without-attention effect. Science 311: Mitchell TM, Hutchinson R, Niculescu RS, Pereira F, Wang X, Just M, Newman
1005–1007. S (2004) Learning to decode cognitive states from brain images. Ma-
Erk S, Spitzer M, Wunderlich AP, Galley L, Walter H (2002) Cultural ob- chine Learn 57:145–175.
jects modulate reward circuitry. Neuroreport 13:2499 –2503. Murphy J, Allen PG, Stevens TH, Weatherhead D (2005) A meta-analysis of
FitzGerald THB, Seymour B, Dolan RJ (2009) The role of human orbito- hypothetical bias in stated preference valuation. Environ Res Econ 30:
frontal cortex in value comparison for incommensurable objects. J Neu- 313–325.
rosci 29:8388 – 8395. Norman KA, Polyn SM, Detre GJ, Haxby JV (2006) Beyond mind-reading:
Friston KJ, Holmes AP, Worsley KJ, Poline JP, Frith CD, Frackowiak RSJ multi-voxel pattern analysis of fMRI data. Trends Cogn Sci 10:424 – 430.
(1994) Statistical parametric maps in functional imaging: a general linear O’Doherty J, Winston J, Critchley H, Perrett D, Burt DM, Dolan RJ (2003)
approach. Hum Brain Mapp 2:189 –210. Beauty in a smile: the role of medial orbitofrontal cortex in facial attrac-
Hampton AN, O’Doherty JP (2007) Decoding the neural substrates of tiveness. Neuropsychologia 41:147–155.
reward-related decision making with functional MRI. Proc Natl Acad Sci Paulus MP, Rogalsky C, Simmons A, Feinstein JS, Stein MB (2003) In-
U S A 104:1377–1382. creased activation in the right insula during risk-taking decision making is
Hampton AN, Adolphs R, Tyszka MJ, O’Doherty JP (2007) Contributions related to harm avoidance and neuroticism. Neuroimage 19:1439 –1448.
of the amygdala to reward expectancy and choice signals in human pre- Peelen MV, Li FF, Kastner S (2009) Neural mechanisms of rapid natural
frontal cortex. Neuron 55:545–555. scene categorization in human visual cortex. Nature 460:94 –97.
Hare TA, O’Doherty J, Camerer CF, Schultz W, Rangel A (2008) Dissociat- Plassmann H, Ambler T, Brautigam S, Kenning P (2007) What can adver-
ing the role of the orbitofrontal cortex and the striatum in the computa- tisers learn from neuroscience? Int J Advertising 26:151–175.
tion of goal values and prediction errors. J Neurosci 28:5623–5630. Plassmann H, O’Doherty J, Shiv B, Rangel A (2008) Marketing actions can
Haxby JV, Gobbini MI, Furey ML, Ishai A, Schouten JL, Pietrini P (2001) modulate neural representations of experienced pleasantness. Proc Natl
Distributed and overlapping representations of faces and objects in ven- Acad Sci U S A 105:1050 –1054.
tral temporal cortex. Science 293:2425–2430. Preuschoff K, Bossaerts P, Quartz SR (2006) Neural differentiation of
Haynes JD, Rees G (2006) Decoding mental states from brain activity in expected reward and risk in human subcortical structures. Neuron
humans. Nat Rev Neurosci 7:523–534.
51:381–390.
Haynes JD, Sakai K, Rees G, Gilbert S, Frith C, Passingham RE (2007) Read-
Rangel A, Camerer C, Montague PR (2008) A framework for studying
ing hidden intentions in the human brain. Curr Biol 17:323–328.
the neurobiology of value-based decision making. Nat Rev Neurosci
Kamitani Y, Tong F (2005) Decoding the visual and subjective contents of
9:545–556.
the human brain. Nat Neurosci 8:679 – 685.
Rees G, Frith CD, Lavie N (1997) Modulating irrelevant motion perception
Kastner S, Ungerleider LG (2000) Mechanisms of visual attention in the
by varying attentional load in an unrelated task. Science 278:1616 –1619.
human cortex. Annu Rev Neurosci 23:315–341.
Kim H, Adolphs R, O’Doherty JP, Shimojo S (2007) Temporal isolation of Rolls ET, McCabe C, Redoute J (2008) Expected value, reward outcome,
neural processes underlying face preference decisions. Proc Natl Acad Sci and temporal difference error representations in a probabilistic decision
U S A 104:18253–18258. task. Cereb Cortex 18:652– 663.
Knutson B, Bossaerts P (2007) Neural antecedents of financial decisions. Schaefer M, Rotte M (2007a) Favorite brands as cultural objects modulate
J Neurosci 27:8174 – 8177. reward circuit. Neuroreport 18:141–145.
Knutson B, Taylor J, Kaufman M, Peterson R, Glover G (2005) Distributed Schaefer M, Rotte M (2007b) Thinking on luxury or pragmatic brand prod-
neural representation of expected value. J Neurosci 25:4806 – 4812. ucts: brain responses to different categories of culturally based brands.
Knutson B, Rick S, Wimmer GE, Prelec D, Loewenstein G (2007) Neural Brain Res 1165:98 –104.
predictors of purchases. Neuron 53:147–156. Schaefer M, Berens H, Heinze HJ, Rotte M (2006) Neural correlates of cul-
Kriegeskorte N, Goebel R, Bandettini P (2006) Information-based func- turally familiar brands of car manufacturers. Neuroimage 31:861– 865.
tional brain mapping. Proc Natl Acad Sci U S A 103:3863–3868. Sharot T, De Martino B, Dolan RJ (2009) How choice reveals and shapes
Kriegeskorte N, Simmons WK, Bellgowan PSF, Baker CI (2009) Circular expected hedonic outcome. J Neurosci 29:3760 –3765.
analysis in systems neuroscience: the dangers of double dipping. Nat Neu- Soon CS, Brass M, Heinze HJ, Haynes JD (2008) Unconscious determinants
rosci 12:535–540. of free decisions in the human brain. Nat Neurosci 11:543–545.
Kuhnen CM, Knutson B (2005) The neural basis of financial risk taking. Vul E, Harris C, Winkielman P, Pashler H (2009) Puzzlingly high correla-
Neuron 47:763–770. tions in fMRI studies of emotion, personality, and social cognition. Per-
Lassiter GD, Lindberg MJ, González-Vallejo C, Bellezza FS, Phillips ND spect Psychol Sci 4:274 –290.
Neuron

Report

Subliminal Instrumental Conditioning


Demonstrated in the Human Brain
Mathias Pessiglione,1,2,* Predrag Petrovic,1 Jean Daunizeau,1 Stefano Palminteri,2 Raymond J. Dolan,1 and Chris D. Frith1
1Wellcome Trust Centre for NeuroImaging, Institute of Neurology, University College London, 12 Queen Square, London WC1N 3BG, UK
2Laboratoire INSERM U610, Centre de NeuroImagerie de Recherche (CENIR), Institut Fédératif de Recherche en Neurosciences, Hôpital
Pitié-Salpêtrière, Université Pierre et Marie Curie (Paris 6), 47 Boulevard de l’Hôpital 75013 Paris, France
*Correspondence: mathias.pessiglione@gmail.com
DOI 10.1016/j.neuron.2008.07.005

SUMMARY conditioning studies in humans have so far been restricted to


Pavlovian paradigms such as fear conditioning (Clark and
How the brain uses success and failure to optimize Squire, 1998; Knight et al., 2003; Morris et al., 1998; Olsson
future decisions is a long-standing question in neuro- and Phelps, 2004), where subliminal stimuli (like unseen faces)
science. One computational solution involves updat- are paired with unpleasant events (like white noise) to increase
ing the values of context-action associations in automatic responses (like skin conductance). To our knowledge,
proportion to a reward prediction error. Previous subliminal instrumental conditioning, where decision making
would be biased by unperceived cues predicting rewards or
evidence suggests that such computations are ex-
punishments, has never been previously demonstrated.
pressed in the striatum and, as they are cognitively
Our subliminal conditioning task was adapted from a published
impenetrable, represent an unconscious learning supraliminal task, wherein subjects selected between visual
mechanism. Here, we formally test this by studying cues so as to learn choices that maximized monetary outcomes
instrumental conditioning in a situation where we (Pessiglione et al., 2006). In our previous study, we modeled sub-
masked contextual cues, such that they were not jects’ behavior by optimizing the free parameters of a standard
consciously perceived. Behavioral data showed machine learning algorithm (termed Q learning), to get maximal
that subjects nonetheless developed a significant likelihoods for the observed decisions. When we regressed key
propensity to choose cues associated with monetary output variables of the optimized model against simultaneously
rewards relative to punishments. Functional neuro- acquired functional neuroimaging data, we showed that predic-
imaging revealed that during conditioning cue values tion errors were expressed in the striatum. Postexperimental
debriefing indicated that some subjects managed to understand
and prediction errors, generated from a computa-
the statistical structure of the task, while others appeared to rely
tional model, both correlated with activity in ventral
on what they referred to as their intuition. These latter reports
striatum. We conclude that, even without conscious suggest that subjects can improve their decisions without
processing of contextual cues, our brain can learn consciously following the incremental steps of the Q-learning
their reward value and use them to provide a bias procedure.
on decision making. The motivating assumption of the current experiment was that
processes associated with striatal learning are not consciously
INTRODUCTION accessible but, nonetheless, influence choice decision making.
Indeed, if contextual cues reach awareness, other brain systems
Humans frequently invoke an argument that their intuition can re- are likely to play a role, as expressed in conscious reasoning or
sult in a better decision than conscious reasoning. Such asser- strategic control, which allows one to develop explicit knowl-
tions may rely on subconscious associative learning between edge of statistical contingencies. However, if the cues remain
subliminal signals present in a given situation and choice out- unseen, learning would solely depend on a subconscious pro-
comes. For instance, clinicians may improve their therapeutic cessing that involves the striatum, with an algorithmic structure
decisions through learned associations between treatment similar to a Q learning, which does not embody explicit informa-
outcomes and subliminal signs presented by their patients. tion about statistical contingencies. Under these assumptions,
Likewise, poker players can improve their gambles through we predicted that, if in our task visual cues were masked, both
a learned association between monetary outcomes and sublim- striatal activity and behavioral choices would still reflect Q-learn-
inal behavioral manifestations of their opponents (the so-called ing outputs.
‘‘gamblers’ tell’’).
The idea that such instrumental conditioning can occur sub- RESULTS
consciously has been around for almost a century (Thorndike,
1911). This assumption originally rested on observations that re- A prerequisite for the present study was to demonstrate efficient
wards and punishments shape behavioral responses in species masking of the visual cues. These cues were novel abstract sym-
allegedly lacking conscious awareness. However, subliminal bols, which were scrambled and mixed to create mask images.

Neuron 59, 561–567, August 28, 2008 ª2008 Elsevier Inc. 561
Neuron
fMRI Study of Subliminal Conditioning

Figure 1. Subliminal Conditioning Task


Successive screenshots displayed during a given
trial are shown from left to right, with durations in
milliseconds. After seeing a masked contextual
cue flashed on a computer screen, subjects
choose to press or not press a response button
and subsequently observe the outcome. In this
example, ‘‘Go’’ appears on the screen because
the subject has pressed the button, following
the cue associated with a rewarding outcome
(winning £1).

To assess visual awareness, we successively displayed two mental Data available online). As they would not see the cues,
masked cues on a computer screen and asked subjects whether we encouraged them to follow their intuition, taking the risky
they perceived a difference or not. We reasoned that if subjects response if they had a feeling they were in a winning trial and
are unable to correctly perceive any difference between the choosing a safe response if they felt it was a losing trial. Note
masked cues, then they are also unable to build conscious rep- that if subjects always made the same response, or if they
resentations of cue-outcome associations. The procedure has performed at chance, their final payoff would be zero.
the advantage of not showing the cues unmasked, so that, by As a dependent variable to assess for conditioning effects, we
the end of the experiment, subjects had no idea what the cues used monetary payoff, which corresponds to the area below the
look like. reward and above the punishment learning curves (Figure 2A).
The perceptual discrimination task was performed outside the Overall subjects won money in this task, on average £7.5 ± £1.8
scanner at the beginning of the experiment, in order to adapt du- (p < 0.001, one-tailed paired t test), indicating that the risky re-
ration of cue display to each individual, and in the scanner at the sponse was more frequently chosen following reward predictive
end of the experiment, to check for any effect of learning or relative to punishment predictive cues. Both reward and punish-
change in visual conditions. For all subjects, cue duration was ment conditions also differed significantly from the neutral condi-
set at either 33 or 50 ms and was kept fixed through the entire tion (p < 0.05, one-tailed paired t test). There was no significant
experiment. In every individual, correct guessing on the final as- difference (p > 0.5, two-tailed paired t test) between the reward
sessment did not differ from chance (p > 0.05, chi-square test). and punishment condition: on average subjects won £24.3 ±
At group level, average percentage of correct responses for £1.9 and avoid losing £23.2 ± £2.1. Learning curves showed
the 20 subjects was 48% ± 3%, which again was not different that responses improved symmetrically for rewards and punish-
from chance (p > 0.5, two-tailed paired t test). Average d0 was ments, ending with 63% ± 5% of correct responses on average.
0.08 ± 0.20, showing that, even when correcting for response Surprisingly, this plateau was reached at around the halfway
bias, signal detection was not different from zero (p > 0.5, two- point of the learning session. The effects of subliminal condition-
tailed paired t test). Thus, subjects remained unable to discrim- ing were subsequently assessed with a preference judgment
inate between the different masked cues, from the beginning task, in which cues were uncovered and rated by the subjects
to the end of conditioning sessions. from the most to least liked (Figure 2B). Ratings were significantly
We employed the same masking procedure in the subliminal higher for reward compared to punishment cues (p < 0.01, one-
conditioning task, in which cues were paired with monetary out- tailed paired t test), consistent with subjects having learned the
comes (Figure 1). From these outcomes (!£1, £0, +£1), subjects affective values of subliminal cues, such that these values were
had to learn when to take the risky response (either ‘‘Go’’ or able to bias their preferences. When uncovering the cues, sub-
‘‘NoGo,’’ depending on subjects). Subjects were also told that, jects were also asked to signal any cue that they may have
for the risky response, the outcome would depend on the cue seen during conditioning sessions; none was reported as previ-
hidden behind the masking image (see instructions in Supple- ously seen.

Figure 2. Behavioral Data


(A) Learning curves. Colors indicate cues for which
button presses are rewarded (green), neutral
(blue), or punished (red). Diamonds represent,
across trials, percentages of subjects that pressed
the button. Left: continuous lines join the dia-
monds to illustrate actual choices made by
subjects. Right: continuous lines represent the
probabilities of button press estimated by an
optimized Q-learning model.
(B) Preferences. After the conditioning phase,
cues were unmasked and subjects rated them,
from the most (3) to the least liked (1). The graph
shows the average rating for reward (green), neu-
tral (blue), and punishment (red) cues. Bars are ±
intersubjects standard errors of the mean.

562 Neuron 59, 561–567, August 28, 2008 ª2008 Elsevier Inc.
Neuron
fMRI Study of Subliminal Conditioning

every trial, the model estimates the likelihood of the risky re-
sponse from the value of the displayed cue. If the risky response
was actually taken, the model then updates the value of the dis-
played cue in proportion to the prediction error. The parameters
of the model were optimized such that likelihoods of risky
responses provided the best fit of subjects’ actual responses
across conditioning sessions (Figure 2A). The Q values and pre-
diction errors generated by this optimized algorithm were then
used as regressors for analysis of brain imaging data (see
Figure S1).
We recorded brain activity while subjects performed the sub-
liminal conditioning task, using functional magnetic resonance
imaging (fMRI). We first examined brain regions reflecting Q
value at the time of cue onset, increasing their response to re-
ward-predicting cues and decreasing their response to punish-
ment-predicting cues, across learning sessions. After correction
for multiple comparisons (family-wise error, p < 0.05), we noted
significant correlated activity in ventral striatum bilaterally (Fig-
ures 3A and 3B, left). The same region was also significantly ac-
tivated at the time of outcome in keeping with prediction errors
being expressed at this time point (Figure 3C, left). In a second
analysis, we computed regression coefficients for the different
conditions at the time of cue and outcome onsets, separately
for the first and second half of conditioning sessions. Contrasts
with the neutral condition were then averaged over all ventral
striatal voxels showing significant activation at the most conser-
vative threshold in the first analysis. This confirmed that from the
first to the second half of conditioning sessions, ventral striatal
responses increased for reward cues and decreased for punish-
ment cues (Figure 4A, left). At the time of outcome onset, the
same ventral striatal region reflected positive prediction errors
in the reward condition and negative prediction errors in the pun-
ishment condition. In keeping with the Q-learning model, both
positive and negative prediction errors decreased from the first
to the second half of conditioning sessions. Thus, across sublim-
inal conditioning, the ventral striatal response was consistent
with the expression of Q values (for unseen cues) and prediction
errors (based on visible outcomes).
We further examined variability in individual performance to
explain why some subjects won more money than others.
More precisely, we searched for brain regions where coefficients
Figure 3. Neuroimaging Data of Q-value regressors correlated with individual payoffs. These
Left: ventral striatal regions isolated by regression of monetary values against regions were confined to extrastriate visual cortex (Figure 3A,
BOLD responses to cue display. Right: visual cortical regions isolated by cor- right) at the most conservative threshold (familywise error, p <
relation of cue-value regression coefficients with individual payoffs. Slices 0.05), spreading into the ventral occipitotemporal stream
were taken at global maxima of interest indicated by red pointers on the above
(Figure 3B, right) with a more liberal threshold (uncorrected, p <
axial glass brains. Areas shown in gray/black on glass brains and in orange/
yellow on coronal slices showed significant effect. The [x y z] coordinates of
0.001). Contrast estimates confirmed that extrastriate voxels
the maxima refer to the Montreal Neurological Institute space. progressively differentiated the reward and punishment cues
(A) Statistical parametric maps using conservative threshold (p < 0.05 after from the first to the second half of conditioning sessions
familywise error correction for multiple comparisons). (Figure 4A, right). At the time of outcome onset, these extrastriate
(B) Statistical parametric maps using liberal threshold (p < 0.001 uncorrected). regions responded positively for both rewards and punishments,
(C) Regression coefficients of Q values (QV) and prediction errors (PE) against
showing no evidence for encoding of prediction errors. Thus,
BOLD responses to cue and outcome display, respectively. Bars are ± inter-
subjects standard errors of the mean.
during the subliminal conditioning task, the extrastriate visual
cortex learned to discriminate between unseen cues according
to their reward value but did not express outcome-related
To model instrumental conditioning, we implemented a stan- prediction errors (Figure 3C, right).
dard Q-learning algorithm (Pessiglione et al., 2006), with inputs To further assess whether the ventral striatum and visual
from individual histories of cues, choices, and outcomes. On cortex were able to discriminate between the subliminal cues,

Neuron 59, 561–567, August 28, 2008 ª2008 Elsevier Inc. 563
Neuron
fMRI Study of Subliminal Conditioning

none was significant; Pearson’s correlation coefficients were


respectively 0.26 and 0.29.

DISCUSSION

We provide evidence that instrumental learning can occur with-


out conscious processing of contextual cues. This finding might
relate to previous evidence for implicit or procedural learning,
where behavioral responses can be adapted to the statistical
structure of stimuli that fails to be reported explicitly (Bayley
et al., 2005; Berns et al., 1997; Destrebecqz and Cleeremans,
2001; Seitz and Watanabe, 2003). Interestingly, implicit/proce-
dural learning has been suggested to involve the basal ganglia,
in contrast with explicit/declarative memory which would involve
the medial temporal lobe (Cohen et al., 1985; Milner et al., 1998;
Poldrack and Packard, 2003; Squire, 1992). In implicit learning
tasks, such as serial reaction time or probabilistic classification,
authors have claimed that subjects can achieve good acquisition
without explicit knowledge of the task structure. However,
methods for assessing awareness of statistical contingencies
have been criticized, principally on the issue that questions
were too demanding in terms of memory (Lagnado et al., 2006;
Lovibond and Shanks, 2002; Wilkinson and Shanks, 2004).
Thus, to formally test whether instrumental conditioning can
occur without awareness, we took a more stringent approach:
masking the cues, so that they remained unperceived.
We believe our methodology avoids most previous problems
related to assessing awareness, by demonstrating that subjects
were not able to discriminate between masked cues (without
Figure 4. Model-free Analyses of Brain Activations the help of rewards and punishments), rather than retrospectively
Ventral striatum (left) and visual cortex (right) correspond to voxels surviving assessing awareness of contingencies. Moreover, postcondi-
familywise error correction on statistical parametric maps. tioning recognition tests would not be sufficient in our case, since
(A) Regression coefficients. Histograms represent contrasts of the reward subjects would not need to identify cues for associative learning
(green) or the punishment (red) condition with the neutral condition, at the to be conscious. Indeed, they could learn associations between
times of cue and outcome display. For every contrast the two joint histograms
risky response outcomes and tiny fragments of the visual dynamic
correspond to the first (empty) and second (filled) halves of the conditioning
sessions.
pattern formed by the mask/cue/mask flashing. However, post-
(B) Time courses. BOLD responses were averaged across trials over condi- conditioning debriefing questions might be informative in explain-
tioning sessions, for the reward (green) and punishment (red) conditions, ing why subjects could not discriminate between masked cues.
relative to the neutral condition. Bars are ± intersubjects standard errors of Thus, when we explicitly asked subjects to state what the cues
the mean. looked like, they reported in majority of cases that they had no
idea. When the subjects were presented with the cues, now
we extracted time courses of BOLD response. These time unmasked, they reported surprise at seeing the symbols while as-
courses were averaged over trials, sessions, and subjects, serting that they had never seen them before. This suggests that
separately for the reward and punishment conditions (Figure 4B). during conditioning, subjects had no a priori representational
We found that BOLD responses to reward and punishment cues knowledge to guide a visual search for cues hidden behind the
significantly differed after two acquisition volumes (3.9 s) in the masks. We believe that absence of an a priori representation is
ventral striatum (one-tailed paired t test, p < 0.01) and after three a crucial feature of our design, which, in addition to visual mask-
(5.85 s) in the visual cortex (one-tailed paired t test, p < 0.01). ing, prevented subjects from consciously seeing the cues.
Finally, we ascertained whether neuroimaging and behavioral Using this methodology, we show that pairing rewards and
effects of subliminal conditioning were driven by subjects scor- punishments can guide behavioral responses and even condi-
ing at the high end in perceptual discrimination performance. tion preferences for abstract cues that subjects could not
We tested for correlations between correct guessing assessed consciously see. Note that if cues were visible, learning curves
in the final perceptual discrimination test and coefficients of Q- would have been optimized in one trial or two; hence we are
value regressors in both the ventral and extrastriate cortex. not claiming that conscious awareness is unhelpful in supralim-
None was significant; Pearson’s correlation coefficients were inal instrumental conditioning. However, in our subliminal condi-
respectively !0.25 and !0.18. We also tested correlation of cor- tioning task, conscious strategies (such as win-stay/lose-switch)
rect guessing with monetary payoffs from conditioning sessions might have been detrimental, which would explain why learning
and differential ratings in the preference judgment task. Again, curves were limited well below the optimum.

564 Neuron 59, 561–567, August 28, 2008 ª2008 Elsevier Inc.
Neuron
fMRI Study of Subliminal Conditioning

We also identified brain circuitry associated with subliminal theory has proven useful to model both brain activity and behav-
instrumental conditioning. The ventral striatum responded to ioral choices in human and nonhuman primates (Daw and Doya,
subliminal cues and to visible outcomes in a manner that closely 2006; McClure et al., 2004; O’Doherty et al., 2007). Brain activity
approximates Q-learning algorithm, expressing reward ex- reflecting sophisticated computations are unlikely to be ac-
pected values and prediction errors, just as was reported in cessed by the conscious mind, which takes minutes to solve
supraliminal instrumental conditioning studies (O’Doherty et al., even simple equations. This brain activity would therefore repre-
2004; Pagnoni et al., 2002; Pessiglione et al., 2006; Yacubian sent unconscious processes, which we formally demonstrated
et al., 2006). Interestingly, there is no need for representing the here in the case of instrumental conditioning. Combining mask-
statistical structure of the task in Q learning, which is an incre- ing and modeling can, in principle, make more tractable the
mental procedure updating the expected values of chosen identification of basic neuronal mechanisms shared within other
actions according to the subject’s history of reward and punish- species, eliminating the use of reportable knowledge that might
ment outcomes. This accords well with views that the striatum is be unique to humans. It might also help assess the integrity of
a major player in implicit/procedural learning (Graybiel, 2005; these same basic mechanisms in patients with neurological or
Hikosaka et al., 1999; Packard and Knowlton, 2002) and with ev- psychiatric conditions, avoiding confounds generated by
idence that ventral striatum encodes reward-related information conscious strategic compensations.
(Delgado, 2007; Knutson and Cooper, 2005; Pecina et al., 2006).
For the sake of simplicity, we have described ventral striatum EXPERIMENTAL PROCEDURES
activity as directly reflecting key outputs of Q-learning algorithm:
Q value at the time of cue onset and prediction error at the time of Subjects
outcome. There are nonetheless other variables in machine The study was approved by the National Hospital for Neurology and Neurosur-
gery and the Institute of Neurology joint Ethics Committee. Subjects were re-
learning literature that would also correlate with ventral striatum
cruited via Gumtree website and screened for exclusion criteria: left handed-
activity and which could provide an alternative interpretational ness, age below 18 or above 39, regular taking of drug or medication,
framework for our study. In particular, it is important to note history of psychiatric or neurological illnesses and contra-indications to MRI
that average Q values (over the reward, neutral, and punishment scanning (pregnancy, claustrophobia, metallic implants). All subjects gave
conditions) remain around zero during our conditioning para- informed consent prior to taking part. We scanned 20 subjects: 11 males
digm. Hence, Q value is approximately equal to Q value minus (mean age 26.8 ± 6.3 years) and 9 females (mean age 23.8 ± 3.3 years). Two
more subjects were scanned but discarded from the analysis, because they
average Q value, which can be seen as equivalent to a cue pre-
eventually could describe parts of the visual cues, were above chance level
diction error (actual Q value minus predicted Q value). Our data in the perception task, and won unusually large amounts of money in the con-
are therefore equally compatible with the notion that the ventral ditioning task. Subjects were told that they would play for real money, but at
striatum encodes prediction errors at the time of both cue and the end of the experiment their winnings were rounded up to a fixed amount.
outcome onsets. However, because prediction errors represent
a function of Q values, the brain has to learn about Q values in Behavioral Task and Analysis
order to signal prediction errors. Thus, whether we consider Subjects first read the instructions (see Supplemental Data) about the different
the ventral striatum as encoding a Q value or a prediction error tasks, which were later explained again step by step. Before scanning, sub-
does not alter our central conclusion: namely, the human brain jects were trained to perform the conditioning task and the perception task
on practice versions. In the scanner, they had to perform three sessions of
can learn the reward value of subliminal cues, so as to later
the conditioning task, each containing 120 trials and lasting 13 min, and one
influence behavioral choices. session of the perception task, containing 120 trials and lasting about 7 min.
It is of interest that extrastriate visual cortex also reflected the The abstract cues were letters taken from the Agathodaimon font. The 12
reward value of subliminal cues, but not outcome-related predic- cues shown in the scanner were randomly assigned to the four task sessions,
tion errors. Modulation of visual cortex activity by monetary in- each session hence employing 3 new cues. The same two masking patterns
centives has already been reported in neuroimaging studies of (see Figure 1), one displayed before and the other after the cue, were used
in all task sessions. The sequence of display and the cue-outcome associa-
supraliminal processes, such as visuomotor transformation, at-
tions were also randomized for every subject.
tentional control, and working memory (Krawczyk et al., 2007; The perceptual discrimination task was used to select the appropriate dura-
Ramnani and Miall, 2003; Small et al., 2005). In our case, the tion for cue display, which was then kept to either 33 or 50 ms for the entire
modulation suggests that conditioning involves an interaction experiment. In this task, subjects were flashed two masked cues, 3 s apart,
between the extrastriate cortex (which would discriminate cues displayed on the center of a computer screen, each following a fixation cross.
according to their visual properties) and the ventral striatum They had to report whether or not they perceived any difference between the
two visual stimulations. The response was given manually, by pressing one of
(which would tag cues with affective values depending on
two buttons assigned to ‘‘same’’ and ‘‘different’’ choices. The perceptual dis-
reward prediction errors). However, we acknowledge that we crimination task was then employed as a control for awareness at the end of
do not as yet have a complete account of how the brain pro- conditioning sessions. We checked with a chi-square test that in all included
duces behavioral effects of subliminal conditioning. Notably, subjects performance was not significantly different from chance level (50%
we failed to identify the brain regions mapping affective values of correct responses). We also calculated d0 measure, which is the difference
onto motor commands, which would complete the circuit from between normalized rates of hits (correct ‘‘different’’ response) and false
visual cues to behavioral responses. Further experiments will alarms (incorrect ‘‘different’’ responses). We ensured that this measure was
not significantly different from zero, at group level, using one-tailed paired t
be necessary to fill in these explanatory gaps.
test.
More generally, our approach, combining perceptual masking The instrumental conditioning task involved choosing between pressing or
and computational modeling, can be extended over the field of not pressing a button, in response to masked cues. After showing the fixation
functional neuroimaging. Computational reinforcement learning cross and the masked cue, the response interval was indicated on the

Neuron 59, 561–567, August 28, 2008 ª2008 Elsevier Inc. 565
Neuron
fMRI Study of Subliminal Conditioning

computer screen by a question mark. The interval was fixed to 3 s and the re- metric mapping software SPM5 (Wellcome Trust center for NeuroImaging,
sponse was taken at the end: ‘‘Go’’ if the button was being pressed, ‘‘No’’ if the London, UK). The first 5 volumes of each session were discarded to allow
button was released. The response was written on the screen as soon as the for T1 equilibration effects. Preprocessing consisted of spatial realignment,
delay had elapsed. Subjects were told that one response was safe (you do not normalization using the same transformation as structural images, and spatial
win or lose anything) while the other was risky (you can win £1, lose £1, or get smoothing using a Gaussian kernel with a full-width at half-maximum of 6 mm.
nothing). The risky response was assigned to Go for half of the subjects, and to We used two different statistical linear regression models for our analyses. In
NoGo for the other half, such that motor aspects were counterbalanced both every trial was modeled as having two time points, corresponding to cue
between reward and punishment conditions. Subjects were also told that and outcome onsets. In the first model, two separate regressors were created
the outcome of the risky response would depend on the cue that was for cues and outcomes, respectively modulated by the Q values and prediction
displayed between the mask images. In fact, three cues were used, one was errors computed by our optimized algorithm. In the second model, 12 separate
rewarding (+£1), one was punishing (!£1), and the last was neutral (£0). Be- regressors were created corresponding to the two time points (cues and
cause subjects were not informed about the associations, they could only outcomes) times the two conditioning phases (first and second half of each
learn them by observing the outcome, which was displayed at the end of the session) times the three conditions (reward, neutral, and punishment). In all
trial. This was either a circled coin image (meaning +£1), a barred coin image cases, the regressors of interest were convolved with a canonical hemody-
(meaning !£1), or a gray square (meaning £0). namic response function (HRF). To correct for motion artifact, subject-specific
Subjects were then debriefed about their visual perceptions and their re- realignment parameters were modeled as covariates of no interest. Linear
sponse strategies. They reported responding either at chance, following their contrasts of regression coefficients were computed at the individual subject
intuition, or following logical rules. None of them had the slightest idea of what level and then taken to a group level random-effects analysis. At group level,
the cues looked like. For the preference judgment task, the cues were then we performed two statistical analyses: first a one-sample t test to find brain re-
shown unmasked on a computer screen. The three cues used for a given ses- gions where regression coefficients were significant across subjects, and sec-
sion were displayed side by side, the position being randomized. Subjects ond a correlation with individual payoffs to find brain regions where regression
were asked to rate them in order of preferences: 3 for the most liked, 2 for coefficients increased with higher conditioning effect. A threshold of p < 0.05
the intermediate, and 1 for the least one. after familywise error (FWE) correction for multiple comparisons was applied to
To assess for instrumental conditioning, we used one-tailed paired t tests avoid any a priori on brain localization. A more liberal threshold (p < 0.001,
comparing individual earnings with chance level (which is £0). Similarly, to as- uncorrected) was also used to observe the extension of significant activations.
sess for preference conditioning, we used one-tailed paired t tests comparing To further illustrate activations, time courses were estimated by fitting a flexible
differential rating of winning and losing cues with chance level (which is 0). basis set of finite impulse responses (FIRs), separated from the next by one
scan (1.95 s). Both regression coefficients and time courses were then aver-
Computational Model aged across subjects, pooling together the voxels that passed the conserva-
We used a standard Q-learning algorithm (Sutton and Barto, 1998), which has tive threshold in statistical parametric maps (SPMs).
been shown previously to offer a good account of instrumental choice in both
humans and monkeys (Daw and Doya, 2006; McClure et al., 2004; O’Doherty
et al., 2007). For each cue, the model estimates the expected value of the risky SUPPLEMENTAL DATA
response, on the basis of individual sequences of choices and outcomes. This
value, termed a Q value, is essentially the amount of reward expected from The Supplemental Data include one figure and supplemental text and can be
choosing the risky response given the contextual cue. These Q values were found with this article online at http://www.neuron.org/cgi/content/full/59/4/
set at 0.1 before learning, and after every risky response the value of the cue 561/DC1/.
was updated according to the Rescorla-Wagner rule: Q(t + 1) = Q(t) + a*d(t).
Following this rule, values are increased if the outcome is better than expected, ACKNOWLEDGMENTS
and decreased in the opposite case. The prediction error was d(t) = R(t) ! Q(t),
with R(t) defined as the reinforcement obtained from choosing the risky This work was funded by the Wellcome Trust research program grants to
response at trial t. In other words, the prediction error d(t) is the difference be- C.D.F. and R.J.D. M.P. received a Young Researcher Prize from the Betten-
tween the expected outcome, i.e., Q(t), and the actual outcome, i.e., R(t). The court-Schuller Foundation, P.P. was supported by the Swedish Research
reinforcement magnitude R was +1 and !1 for winning and losing £1, and 0 for Council, and J.D. is a beneficiary of the Marie Curie Intra-European research
neutral outcomes. Given the Q value, the associated probability of choosing fellowship program. We wish to thank Helen Bates and Maël Lebreton, who
the risky response was estimated by implementing the softmax rule: P(t) = 1/ independently replicated the behavioral findings outside the scanner.
(1 + exp(!Q(t)/b)). This rule ensures that likelihood will be superior to 0.5 for
positive values and inferior to 0.5 for negative values. The learning rate a con-
cerns the amplitude of value changes from one trial to the next. The tempera- Accepted: July 1, 2008
ture b concerns the randomness of decision making. These two free parame- Published: August 27, 2008
ters, a and b, were adjusted to maximize the probability (or likelihood) of the
actual choices under the model. With the constraint that the parameters
should be identical for reward and punishment cues we found: a = 0.1 and REFERENCES
b = 0.9. The model was then used to create statistical regressors correspond-
ing to the Q values and prediction errors, for analysis of brain images. Bayley, P.J., Frascino, J.C., and Squire, L.R. (2005). Robust habit learning in
the absence of awareness and independent of the medial temporal lobe.
Images Acquisition and Analysis Nature 436, 550–553.
T2*-weighted echo planar images (EPI) were acquired with blood oxygen-level Berns, G.S., Cohen, J.D., and Mintun, M.A. (1997). Brain regions responsive to
dependent (BOLD) contrast on a 3.0 Tesla magnetic resonance scanner. We novelty in the absence of awareness. Science 276, 1272–1275.
employed a tilted plane acquisition sequence designed to optimize functional
Clark, R.E., and Squire, L.R. (1998). Classical conditioning and brain systems:
sensitivity in the orbitofrontal cortex and medial temporal lobes (Deichmann
the role of awareness. Science 280, 77–81.
et al., 2003). To cover the whole brain with a short TR (1.95 s), we used the fol-
lowing parameters: 30 slices, 2 mm slice thickness, 2 mm interslice gap. T1- Cohen, N.J., Eichenbaum, H., Deacedo, B.S., and Corkin, S. (1985). Different
weighted structural images were also acquired, coregistered with the mean memory systems underlying acquisition of procedural and declarative knowl-
EPI, normalized to a standard T1 template, and averaged across subjects to edge. Ann. N Y Acad. Sci. 444, 54–71.
allow group level anatomical localization. EPI images were analyzed in an Daw, N.D., and Doya, K. (2006). The computational neurobiology of learning
event-related manner, within a general linear model, using the statistical para- and reward. Curr. Opin. Neurobiol. 16, 199–204.

566 Neuron 59, 561–567, August 28, 2008 ª2008 Elsevier Inc.
Neuron
fMRI Study of Subliminal Conditioning

Deichmann, R., Gottfried, J.A., Hutton, C., and Turner, R. (2003). Optimized O’Doherty, J.P., Hampton, A., and Kim, H. (2007). Model-based fMRI and its
EPI for fMRI studies of the orbitofrontal cortex. Neuroimage 19, 430–441. application to reward learning and decision making. Ann. N Y Acad. Sci.
1104, 35–53.
Delgado, M.R. (2007). Reward-related responses in the human striatum. Ann.
N Y Acad. Sci. 1104, 70–88. Olsson, A., and Phelps, E.A. (2004). Learned fear of ‘‘unseen’’ faces after
Pavlovian, observational, and instructed fear. Psychol. Sci. 15, 822–828.
Destrebecqz, A., and Cleeremans, A. (2001). Can sequence learning be
implicit? New evidence with the process dissociation procedure. Psychon. Packard, M.G., and Knowlton, B.J. (2002). Learning and memory functions of
Bull. Rev. 8, 343–350. the Basal Ganglia. Annu. Rev. Neurosci. 25, 563–593.
Pagnoni, G., Zink, C.F., Montague, P.R., and Berns, G.S. (2002). Activity in hu-
Graybiel, A.M. (2005). The basal ganglia: learning new tricks and loving it. Curr.
man ventral striatum locked to errors of reward prediction. Nat. Neurosci. 5,
Opin. Neurobiol. 15, 638–644.
97–98.
Hikosaka, O., Nakahara, H., Rand, M.K., Sakai, K., Lu, X., Nakamura, K., Miya- Pecina, S., Smith, K.S., and Berridge, K.C. (2006). Hedonic hot spots in the
chi, S., and Doya, K. (1999). Parallel neural networks for learning sequential brain. Neuroscientist 12, 500–511.
procedures. Trends Neurosci. 22, 464–471.
Pessiglione, M., Seymour, B., Flandin, G., Dolan, R.J., and Frith, C.D. (2006).
Knight, D.C., Nguyen, H.T., and Bandettini, P.A. (2003). Expression of condi- Dopamine-dependent prediction errors underpin reward-seeking behaviour
tional fear with and without awareness. Proc. Natl. Acad. Sci. USA 100, in humans. Nature 442, 1042–1045.
15280–15283. Poldrack, R.A., and Packard, M.G. (2003). Competition among multiple
Knutson, B., and Cooper, J.C. (2005). Functional magnetic resonance imaging memory systems: converging evidence from animal and human brain studies.
of reward prediction. Curr. Opin. Neurol. 18, 411–417. Neuropsychologia 41, 245–251.
Krawczyk, D.C., Gazzaley, A., and D’Esposito, M. (2007). Reward modulation Ramnani, N., and Miall, R.C. (2003). Instructed delay activity in the human pre-
of prefrontal and visual association cortex during an incentive working memory frontal cortex is modulated by monetary reward expectation. Cereb. Cortex
task. Brain Res. 1141, 168–177. 13, 318–327.

Lagnado, D.A., Newell, B.R., Kahan, S., and Shanks, D.R. (2006). Insight and Seitz, A.R., and Watanabe, T. (2003). Psychophysics: Is subliminal learning
strategy in multiple-cue learning. J. Exp. Psychol. Gen. 135, 162–183. really passive? Nature 422, 36.
Small, D.M., Gitelman, D., Simmons, K., Bloise, S.M., Parrish, T., and Mesu-
Lovibond, P.F., and Shanks, D.R. (2002). The role of awareness in Pavlovian
lam, M.M. (2005). Monetary incentives enhance processing in brain regions
conditioning: empirical evidence and theoretical implications. J. Exp. Psychol.
mediating top-down control of attention. Cereb. Cortex 15, 1855–1865.
Anim. Behav. Process. 28, 3–26.
Squire, L.R. (1992). Memory and the hippocampus: a synthesis from findings
McClure, S.M., York, M.K., and Montague, P.R. (2004). The neural substrates with rats, monkeys, and humans. Psychol. Rev. 99, 195–231.
of reward processing in humans: the modern role of FMRI. Neuroscientist 10,
Sutton, R.S., and Barto, A.G. (1998). Reinforcement Learning (Cambridge, MA:
260–268.
MIT Press).
Milner, B., Squire, L.R., and Kandel, E.R. (1998). Cognitive neuroscience and Thorndike, E.L. (1911). Animal Intelligence: Experimental Studies (New York:
the study of memory. Neuron 20, 445–468. Macmillan).
Morris, J.S., Öhman, A., and Dolan, R.J. (1998). Conscious and unconscious Wilkinson, L., and Shanks, D.R. (2004). Intentional control and implicit
emotional learning in the human amygdala. Nature 393, 467–470. sequence learning. J. Exp. Psychol. Learn. Mem. Cogn. 30, 354–369.
O’Doherty, J., Dayan, P., Schultz, J., Deichmann, R., Friston, K., and Dolan, Yacubian, J., Glascher, J., Schroeder, K., Sommer, T., Braus, D.F., and Bu-
R.J. (2004). Dissociable roles of ventral and dorsal striatum in instrumental chel, C. (2006). Dissociable systems for gain- and loss-related value predic-
conditioning. Science 304, 452–454. tions and errors of prediction in the human brain. J. Neurosci. 26, 9530–9537.

Neuron 59, 561–567, August 28, 2008 ª2008 Elsevier Inc. 567
9984 • The Journal of Neuroscience, September 12, 2007 • 27(37):9984 –9988

Behavioral/Systems/Cognitive

Orbitofrontal Cortex Encodes Willingness to Pay in Everyday


Economic Transactions
Hilke Plassmann, John O’Doherty, and Antonio Rangel
California Institute of Technology, Division of Humanities and Social Sciences, MC 228-77, Pasadena, California 91125

An essential component of every economic transaction is a willingness-to-pay (WTP) computation in which buyers calculate the maxi-
mum amount of financial resources that they are willing to give up in exchange for the object being sold. Despite its pervasiveness, little
is known about how the brain makes this computation. We investigated the neural basis of the WTP computation by scanning hungry
subjects’ brains using functional magnetic resonance imaging while they placed real bids for the right to eat different foods. We found that
activity in the medial orbitofrontal cortex and in the dorsolateral prefrontal cortex encodes subjects’ WTP for the items. Our results
support the hypothesis that the medial orbitofrontal cortex encodes the value of goals in decision making.
Key words: decision making; reward; neuroeconomics; orbitofrontal cortex; fMRI; buying

Introduction ination task (see also Arana et al., 2003). Erk et al. (2002) found
An essential component of every marketplace transaction is a that mOFC activity during a hypothetical liking rating task in-
willingness-to-pay (WTP) computation in which buyers calcu- creased with the reported attractiveness of the stimuli. Finally, a
late the maximum amount of resources that they are willing to series of stimulus– outcome learning studies (in which no deci-
give up in exchange for the object being sold. The WTP compu- sions were made) have shown that the mOFC maintains a repre-
tation is used to evaluate whether a proposed trade is beneficial sentation of the expected reward associated with particular cues
(e.g., when the WTP exceeds the price at which the item is being (Rolls, 1996; Schoenbaum et al., 1998; Tremblay and Schultz,
offered) or to decide how much to bid for an item (e.g., when 1999; Roesch and Olson, 2004). Although all of these studies
competing with other individuals at an auction). To make good suggest that the medial OFC plays a critical role in the evaluation
trades, individuals must be able to assign a WTP to an item that is of choices, none of them have established that activity in the
commensurate to the benefits that it will generate. Otherwise they medial OFC correlates with the economic computation of WTP.
would end up purchasing items for a price that exceeds their We investigated the neural basis of the WTP computation by
worth to them. Despite its pervasiveness and importance for eco- scanning hungry subjects’ brains using fMRI while they placed
nomic well being, little is known about how the brain performs bids for the right to eat different foods in a Becker–DeGroot–
the WTP computation in everyday transactions, or about how its Marshak auction (Becker et al., 1964). The results described be-
ability to do so is affected by diseases such as addiction or obses- low confirmed our hypothesis: we found that activity in the right
sive compulsive disorders. This makes understanding how and medial OFC encodes subjects’ WTP for items.
where the brain makes these computations one of the most im-
portant open questions in the nascent field of neuroeconomics Materials and Methods
(Glimcher and Rustichini, 2004; Camerer et al., 2005). Subjects. Nineteen normal-weight subjects participated in the experi-
Based on the results of several previous studies, we hypothe- ment (16 males, mean age, 25.45; age range, 18 – 46). One additional
sized that activity in the medial orbitofrontal cortex (mOFC) subject participated in the experiment but was excluded from the analysis
encodes WTP. Monkey electrophysiology studies of binary because she did not understand the instructions. All subjects were right-
choice have found that activity in the OFC encodes the value of handed, healthy, had normal or corrected-to-normal vision, had no his-
tory of psychiatric diagnoses, neurological or metabolic illnesses, and
the available actions (Wallis and Miller, 2003; Padoa-Schioppa
were not taking medications that interfere with the performance of fMRI.
and Assad, 2006). Using functional magnetic resonance imaging All subjects had no history of eating disorders and were screened for
(fMRI), Paulus and Frank (2003) found greater medial OFC ac- liking and at least occasionally eating the types of foods that we used.
tivation during hypothetical choices than during a visual discrim- Subjects were told that the goal of the experiment was to study food
preferences and gave written consent before participating. Caltech’s in-
stitutional The review board of the California Institute of Technology
Received May 9, 2007; revised July 2, 2007; accepted July 24, 2007. (Pasadena, CA) approved the study.
This work was supported by the Moore Foundation, National Science Foundation Grant SES-0134618, and Ger-
Stimuli. Subjects bid on 50 different sweet and salty junk foods (e.g.,
man Academic Exchange Service Grant DAAD D/05/47698. We thank Vivian Valentin, Jan Glaescher, Alan Hampton,
and Axel Linder for their help with this work.
chips and candy bars). We selected the foods based on pilot data to satisfy
Correspondence should be addressed to Antonio Rangel, California Institute of Technology, Division of Human- several characteristics. First, we wanted items to be highly familiar and to
ities and Social Sciences, MC 228-77, Pasadena, CA 91125. E-mail: rangel@hss.caltech.edu. be sold in local convenience stores, to remove uncertainty considerations
DOI:10.1523/JNEUROSCI.2131-07.2007 from the WTP computation as much as possible. Familiarity data col-
Copyright © 2007 Society for Neuroscience 0270-6474/07/279984-05$15.00/0 lected at the end of the experiment shows that we were successful [famil-
Plassmann et al. • fMRI of Willingness to Pay J. Neurosci., September 12, 2007 • 27(37):9984 –9988 • 9985

There is also no incentive to increase the bid


above the WTP because this may lead to a situ-
ation in which the subject gets the item but ends
up paying a price larger than his WTP (e.g.,
consider the case WTP " $1, b " $3, and n "
$2). The fact that bidding the WTP is the opti-
mal strategy was explained and emphasized ex-
tensively during the instruction and training
period. We performed an extensive amount of
pilot work to find a set of instructions that led to
100% reported compliance with the best strat-
egy. The instructions, included in supplemental
material (available at www.jneurosci.org), em-
phasized that the subject’s best strategy is to
look at the item, ask themselves how much is
worth, and simply bid that amount. Third, be-
cause individuals always bid their exact WTP,
we got a measure of the WTP computed by the
brain for every bidder and item at the time of
decision making, which we could then compare
with the blood oxygenation level-dependent
(BOLD) measure of neural activity.
To keep the task simple, subjects were only
allowed to bid discrete amounts for the items
($0, $1, $2, or $3). A consequence of this is that
the bids are only approximations of the true
WTP computed by subjects. For example, when
Figure 1. Experimental design. A, Timeline of the experiment. B, Time course for free bid and forced bid trials. Free and forced the true WTP is $2.3, our measure is $2. Simi-
bid trials were identical except that in forced bid trials visual presentation of the food items was paired with the forced bid amount. larly, subjects with a WTP larger than $3 enter a
In addition, the forced bid amount was repeated during the bidding cue. Food items, trial type, and forced bid amounts were fully bid of $3. However, the bids are a monotonic
randomized within subjects. function of the true WTP and highly correlated
with it.
iarity scores: mean, 3.97; SD, 1.34; scale, 1 (not familiar) to 5 (very famil- We used two different kinds of trials: free-bid
iar)]. Second, we wanted items to be positive for the subjects (in the sense trials and forced-bid trials. Each of the 50 items was shown twice, once in
that their WTP for them is greater or equal than zero). The foods were a bid trial and once in a forced trial. These trials were fully randomized
presented to the subjects using high-resolution color pictures (72 dpi). within and across subjects. Both types of trials had an equal probability of
The stimulus presentation and response recording was controlled by being selected to be the trial that counted. The timing for each type of trial
E-prime (Psychology Software Tools, Pittsburgh, PA). The visual stimuli is shown in Figure 1 B. The only difference between the two types of trials
were presented using video goggles. is that whereas subjects were free to select the amount of their bid in the
Task. Figure 1 describes the time structure of the experiment. Subjects free trials, they were told how much to bid in the forced trials. The forced
were instructed not to eat for 4 h before the experiment, which increased bids were drawn uniformly and independently from $0, $1, $2, or $3 on
the value that they placed on the foods. They were also instructed that each trial. The set of rules described above applied to both trials. Note
they would have to remain in the lab for 30 min at the conclusion of the that subjects needed to make a willingness-to-pay computation in free
experiment, and that the only thing that they will be able to eat is what- trials to decide how much to bid, but they did not need to do so in forced
ever food they purchased from us during the task. In addition to a $35 trials.
participation fee, each subject received three $1 bills in “spending After receiving the instructions, subjects were trained on using the
money” to purchase food from us. Whatever money they did not spend response boxes with their right hand and on the bidding procedure. To
was theirs to keep. avoid activation artifacts caused by the assignment of buttons to bid
Subjects placed bids for the right to eat a snack at the end of the amounts, the assignment was counterbalanced across subjects.
experiment in 100 different bidding trials. In each trial they were allowed The existence of two types of bidding trials is a novel and essential
to bid $0, $1, $2, or $3 for each food item. At the end of the experiment, component of the experimental design. A difficulty in searching for the
one of those trials was randomly selected, by drawing a ball from an urn, neural basis of the WTP computation is that, when the brain is exposed to
and only the outcome of that trial was implemented. As a result, subjects a picture of a food item, it might calculate other variables that are corre-
did not have to worry about spreading their $3 dollar budget over the lated with WTP. For example, the brain may simulate the anticipated
different items and they could treat each trial as if it were the only deci- taste of the food, or it may asses its caloric content. If this issue is not
sion that counted. Objects were sold using a Becker–DeGroot–Marschak properly addressed, one could erroneously attribute WTP computations
auction. The rules of the auction are as follows. Let b denote the bid made to areas that are calculating different albeit correlated variables. The pres-
by the subject for a particular item. After the bid is made a random ence of free and forced trials provides a solution to the problem. The only
number n is drawn from a known distribution (in our case, $0, $1, $2, difference between both types of trials is that the subject needs to perform
and $3 were chosen with equal probability). If b ! n, the subject got the a WTP computation in the free trials, because she needs to decide how
item and paid a price equal to n. In contrast, if b ! n, the subject did not much to bid, but not in the forced trials, because she is told what her bid
get the object but also did not have to pay anything. should be. Every other computation, such as the anticipated taste of the
We used this auction institution as our model of market transactions food, should be performed equally in both types of trials. As a result, we
in the laboratory because it has three very useful properties. First, it is can conclude that a brain area encodes the WTP computation whenever
characterized by a simple set of rules. Second, the optimal strategy for a its activity increases with the WTP in the free trials, but not in the forced
buyer is to bid exactly her WTP for the item being sold (Becker et al., trials.
1964). The intuition for why this is the case is as follows. There is no fMRI data acquisition. The functional imaging was conducted using a
incentive to bid less than the WTP because the price paid is determined Siemens (Erlangen, Germany) 3.0 Tesla Trio MRI scanner to acquire
by the random number n and, thus, the bids do not affect the price paid. gradient echo T2*-weighted echoplanar (EPI) images with BOLD con-
9986 • J. Neurosci., September 12, 2007 • 27(37):9984 –9988 Plassmann et al. • fMRI of Willingness to Pay

trast. To optimize functional sensitivity in the OFC, we used a tilted


acquisition in an oblique orientation of 30° to the anterior commissure–
posterior commissure line (Deichmann et al., 2003). In addition, we used
an eight-channel phased array coil which yields a 40% signal increase in
signal in the medial OFC over a standard head coil. Each volume com-
prised 32 axial slices. A total of 1100 volumes (two sessions, !18 min
each) were collected during the experiment in an interleaved-ascending
manner. The imaging parameters were as follows: echo time, 30 ms; field
of view, 192 mm; in-plane resolution and slice thickness, 3 mm; repeti-
tion time, 2 s. Whole-brain high-resolution T1-weighted structural scans
(1 " 1 " 1 mm) were acquired from the 19 subjects and coregistered with
their mean EPI images and averaged together to permit anatomical lo-
calization of the functional activations at the group level. Image analysis
was performed using SPM5 (Wellcome Department of Imaging Neuro-
science, Institute of Neurology, London, UK). Temporal normalization
was applied to the scans with a time of acquisition of 1.9375 referenced to
the last volume. To correct for subject motion, the images were realigned
to the last volume, spatially normalized to a standard T2* template with
a resampled voxel size of 3 mm, and spatially smoothed using a Gaussian
kernel with a full width at half maximum of 8 mm. Intensity normaliza-
tion and high-pass temporal filtering (using a filter width of 128 s) were
Figure 2. Behavioral results. A, Distribution of bids in free and forced trials. B, Reaction times
also applied to the data.
for free and forced trials as a function of bid. Error bars denote SEs.
fMRI data analysis. The data analysis proceeded in three steps. First, we
estimated a general linear model with AR(1) and the following regressors
that capture the main events in our experiment: free bid and picture cluster size of 10. Anatomical localizations were then performed by over-
presentation (R1), free bid and response (R2), forced bid and picture laying the t maps on a normalized structural image averaged across sub-
presentation (R3), forced bid and response (R4), missed bid trial and jects, and with reference to an anatomical atlas (Duvernoy, 1999)
picture presentation (R5), and missed bid trial and response (R6). The
regressors that capture the presentation of the food pictures were mod- Results
eled using 4 s box-car functions. The regressors for the bid responses were Behavioral
modeled using stick functions. Figure 2 A shows the distribution of bids during free- and forced-
To take advantage of the parametric nature of our design, the general bid trials. The average free bid was $1.4 (SD, 0.27) and over 75%
linear model also included the following parametric modulators: free bid of the free bids were greater than zero. The bid amounts for the
and picture presentation modulated by bid (M1), free bid and picture
forced bid trials were randomly drawn from a uniform distribu-
presentation modulated by surplus# (M2), free bid and response mod-
ulated by bid (M3), free bid and response modulated by surplus# (M4), tion on $0, $1, $2, and $3. Although there is substantial variability
forced bid and picture presentation modulated by bid (M5), forced bid on value that subjects place on particular items, the average WTP
and item presentation modulated by surplus# (M6), forced bid and item was significantly greater than zero ( p & 0.001), which suggests
presentation modulated by surplus$ (M7), forced bid and response that most items were rewarding for most subjects.
modulated by bid (M8), forced bid and response modulated by surplus# Figure 2 B shows the distribution of bidding reaction times for
(M9), and forced bid and response modulated by surplus$ (M10). The free and forced trials. The reaction times were entered in a two-
parametric modulators are defined as follows. “Bid” equals the amount way repeated-measures ANOVA with two factors: bid amount
bid for the item sold in that trial during the corresponding free trial and, ($0, $1, $2, or $3) and trial type (free- or forced-bid trial). The
thus, is a measure of the subject’s WTP for the item being shown. “Sur- analysis revealed no significant main effects or interactions.
plus” is a variable that measures the expected “profit” from the trial given
the bid that was placed (and conditional on the trial being selected to be
the one that counts). For example, suppose that a subject’s true value is $2
Neuroimaging results
and that he bids $2. Then the surplus equals 0.25 " $2 # 0.25 " $1 # Identifying the neural correlates of WTP in free trials
0.50 " $0% $0.75, where the first term measures the probability that the We performed a whole-brain analysis to identify areas that cor-
random number in the auction is 0 times the profit made in that case, the related with WTP in the free trials at the time of evaluation (i.e.,
second number measure the probability that the random number is 1 when the food picture is displayed). This contrast is interesting
times the profit made in that case, and so on. Surplus# % max[0,surplus] because it identifies areas that might encode for WTP. Our hy-
and Surplus$ % min[surplus,0]. (Note that for the free trials, the surplus pothesis was that activity in the OFC would be positively corre-
variable is always non-negative, whereas for the forced trials it can be lated with WTP. The hypothesis was supported by the data: ac-
positive or negative.) We orthogonalized the modulators for each of the tivity in the medial OFC (x % 6, y % 30, z % $17; p & 0.001,
main regressors (M1 and M2, M3 and M4, M5 to M7, and M8 to M10).
uncorrected) was correlated with WTP. Other areas identified by
Each of the regressors was convolved with a canonical hemodynamic
response function. We also included a constant term and six motion
this contrast were the dorsal anterior cingulate cortex (x % $4,
parameters as regressors of no interest. Note that the inclusion of the y % 34, z % $20; p & 0.001, uncorrected), and the dorsolateral
surplus modulator is important to avoid confounding areas that code for prefrontal cortex (DLPFC; x % 44, y % 44, z %12; p & 0.001,
WTP with areas that code for economic surplus. uncorrected).
Second, we calculated the following first-level single-subject contrasts:
(1) free-bid trials when exposed to item modulated by bid (regressor Identifying the neural correlates of value during the
M1), (2) forced-bid trials when exposed to item modulated by bid (re- forced-bid trials
gressor M5), and (3) free- minus forced-bid trials when exposed to an Our experimental design is based on the idea that the brain com-
item modulated by bid (regressors M1 minus M5). putes a WTP during free-bid trials, but not during forced-bid
Third, we calculated second-level group contrasts using a one-sample trials. To test this hypothesis, we performed a whole-brain anal-
t test. The figures shown below are constructed using these second-level ysis to identify areas correlated with WTP in the forced-bid trials
contrasts at a threshold of p & 0.001, uncorrected, and a minimum at the time of evaluation (i.e., when the food picture is displayed).
Plassmann et al. • fMRI of Willingness to Pay J. Neurosci., September 12, 2007 • 27(37):9984 –9988 • 9987

Discussion
In this study, we provide evidence that
mOFC encodes subjects’ WTP during sim-
ple economic transactions. Critically, we
used a parametric experimental design
that allowed us to identify areas that en-
code for WTP, as opposed to areas that are
active during economic choice, but that do
not correlate with WTP (Blair et al., 2006;
Arana et al., 2003; Paulus and Frank,
2003).
Our findings are consistent with data
from human lesion studies showing that
lesions to the ventromedial prefrontal cor-
tex impair the ability to make consistent
pairwise choices (Fellows and Farah,
2007). Our findings are also consistent
with monkey electrophysiology studies of
simple choice behavior. For example, sev-
eral studies have found that OFC neurons
encode for the decision or incentive value
of the stimuli at the time of decision mak-
ing (Wallis and Miller, 2003; Roesch and
Figure 3. Neural correlates of WTP. A, B, Activity in the medial OFC and the DLPFC was positively correlated with WTP at the Olson, 2004; Padoa-Schioppa and Assad,
time of evaluation in the free trials more than in the forced trials. Activation maps shown at a threshold of p ! 0.001 uncorrected 2006). In addition to providing cross-
and 10 voxel clusters. C, Averaged time courses for the medial OFC voxels during free trials as a function of WTP (error bars denote species and cross-modality validation, this
SEs). D, Averaged time courses for the medial OFC voxels during forced bid trials as a function of WTP. E, Averaged time courses for experiment shows that the OFC plays a
the medial OFC voxels during forced bid trials as a function of the forced bid. A comparison of the time courses shows that the central role in the encoding of decision
medial OFC encodes WTP in free trials, but not in forced trials, and that it does not encode the forced bid amounts. values in decision-making tasks that are
significantly more abstract and complex
No areas showed the desired correlation at a level of p ! 0.001 that those that can be studied in monkey
(uncorrected) and an extent threshold of 10 voxels. experiments. As a result, we speculate that the mOFC might en-
Another possibility is that during the forced trials the brain code for the decision value of choices in a wide class of economic
encodes either the size of the forced bid, or the disagreement settings. However, because this study and much of the previous
between the forced bid and the WTP. We tested for both possi- literature has focused on the valuation of primary appetitive re-
bilities using minor variants of the general linear model described wards, such as desirable foods, additional work is needed to in-
above and found no regions of correlation at a level of p ! 0.001 vestigate whether the mOFC area also encodes for the value of
(uncorrected) and an extent threshold of 10 voxels (for details, nonprimary rewards, such as a book or a DVD, and of negative or
see the supplemental material, available at www.jneurosci.org). undesirable items, such as electric shocks of different
magnitudes.
Test of the main hypothesis An open question in behavioral neuroscience is which parts of
As discussed above, a limitation of the previous contrast is that it
the OFC play a role in learning the encoding of stimulus-outcome
identifies areas with activity that is correlated with WTP, but also
associations (Tremblay and Schultz, 1999; Rolls, 2000) and which
areas that encode for variables that are correlated with it, such as
anticipatory taste. To address this potential confound, we looked parts are involved in guiding decisions by encoding the value of
for areas that (1) showed increasing activation with WTP in the alternative goals (Arana et al., 2003; Schoenbaum and Roesch,
free trials and (2) were significantly more activated in the free 2005; Feierstein et al., 2006; Padoa-Schioppa and Assad, 2006)?
trials than in the forced trials. As predicted, we found that the Our results suggest that the mOFC plays a critical role in goal
right mOFC satisfied these conditions (x " 4, y " 30, z " #18; directed behavior by encoding economic value.
p ! 0.001, uncorrected). Unexpectedly, we also found that right We also found that activity in the DLPFC was correlated with
DLPFC satisfied them (x " 44, y " 44, z " 18; p ! 0.001, uncor- the subject’s WTP. This is consistent with previous monkey neu-
rected). Figure 3, A and B, describes the results of this contrast. rophysiology studies that have found neurons in the DLPFC that
We also extracted trial averaged time-course data from peak encode aspects of the decision or the incentive value of stimuli
voxels in the mOFC for each subject, which were then averaged (Watanabe, 1996; Wallis and Miller, 2003). This raises an impor-
across subjects (Fig. 3C–E). The time courses show that activity in tant open question in the neuroeconomics literature: what are the
this area during free trials showed an increase in activation that relative contributions of the DLPFC and the mOFC to the valu-
was correlated with the subjects’ bids. The time courses also show ation of stimuli during economic decision making? The neuro-
that activity during forced trials did not discriminate the subjects’ anatomy of these two regions suggests a potential explanation.
WTP for the items (as measured by their bids for that item during The OFC receives inputs from multiple sensory areas, which are
free trials) or the magnitude of the forced bids. We thus con- likely to be used in valuation, whereas the DLPFC does not (Price,
cluded that activity in the right medial OFC and dorsolateral 2006). In contrast, the DLPFC is heavily connected with motor
prefrontal cortex encode for WTP in everyday economic output areas, whereas the OFC is not connected with these areas
decisions. directly (Petrides and Pandya, 1999). Finally, the DLPFC and the
9988 • J. Neurosci., September 12, 2007 • 27(37):9984 –9988 Plassmann et al. • fMRI of Willingness to Pay

OFC are interconnected (Petrides and Pandya, 1999). This pat- frontal cortex to incentive motivation and goal selection. J Neurosci
tern of connectivity suggests that economic values might be first 23:9632–9638.
computed in the mOFC and then passed to the DLPFC to influ- Becker G, DeGroot M, Marschak J (1964) Measuring utility by a single-
response sequential method. Behav Sci 9:226 –232.
ence motor commands. This pathway is not unique as the OFC
Blair K, Marsh AA, Morton J, Vythilingam M, Jones M, Mondillo K, Pine DC,
might also be able to affect actions through its connections to the Drevets WC, Blair JR (2006) Choosing the lesser of two evils, the better
striatum, which, in turn, is also heavily connected to the motor of two goods: specifying the roles of ventromedial prefrontal cortex and
system (Yeterian and Pandya, 1991). This conjecture is consistent dorsal anterior cingulate in object choice. J Neurosci 26:11379 –11386.
with our data, but our experimental design does not allow us to Camerer C, Loewenstein G, Prelec D (2005) Neuroeconomics: how neuro-
reach this directly. science can inform economics. J of Economic Literature 43:9 – 64.
It is interesting to compare our findings with those of Knutson Deichmann R, Gottfried JA, Hutton C, Turner R (2003) Optimized EPI for
et al. (2007), who study which brain areas are involved in making fMRI studies of the orbitofrontal cortex. NeuroImage 19:430 – 441.
economic purchase decisions for unfamiliar items sold at exog- Duvernoy HM (1999) The human brain: surface, three-dimensional sec-
enously given prices. Their task proceeds in three steps: (1) sub- tional anatomy with MRI, and blood supply. Berlin: Springer.
Erk S, Spitzer M, Wunderlich A, Galley L, Walter H (2002) Cultural objects
jects are shown a picture of the item that is for sale, (2) the sale
modulate reward circuitry. NeuroReport 13:2499 –2503.
price is added, and (3) subjects decide whether to purchase the Feierstein CE, Quirk MC, Uchida N, Sosulski DL, Mainen ZF (2006) Rep-
item or not. Knutson et al. (2007) obtained a measure of the value resentation of spatial goals in rat orbitofrontal cortex. Neuron
that the subjects placed on the items, albeit not a WTP, and 51:495–507.
looked for neural correlates of this value at the time of picture Fellows LK, Farah MJ (2007) The role of ventromedial prefrontal cortex in
presentation (step 1). They found a correlation with nucleus ac- decision making: judgment under uncertainty or judgment per se? Cere-
cumbens (NAcc) activity, but not with OFC activity. In contrast, bral Cortex, in press.
in our study activity in the medial OFC, but not in the NAcc, Glimcher PW, Rustichini A (2004) Neuroeconomics: the concilience of
encoded the subject’s WTP for items. Note that the difference in brain and decision. Science 306:4447– 4452.
Knutson B, Adams CM, Fong GW, Hommer D (2001) Anticipation of in-
results is not attributable to an inability to image the NAcc be-
creasing monetary reward selectively recruits nucleus accumbens. J Neu-
cause we get strong striatal activation in other contrasts of interest rosci 21:RC159.
(supplemental Tables 3, 4, available at www.jneurosci.org as sup- Knutson B, Rick S, Wimmer GE, Prelec D, Loewenstein G (2007) Neural
plemental material). Instead, it is likely that there are subtle but predictors of purchases. Neuron 53:147–156.
important differences between the computations that the brain Padoa-Schioppa C, Assad JA (2006) Neurons in the orbitofrontal cortex
makes in the two tasks, and that the mOFC and NAcc play a encode economic value. Nature 441:223–226.
differential role in such computations. For example, in the Knut- Paulus MP, Frank LR (2003) Ventromedial prefrontal cortex activation is
son et al. (2007) task, the value of purchasing an item is the WTP critical for preference judgments. NeuroReport 14:1311–1315.
minus the price and the information needed to compute this “net Petrides M, Pandya DN (1999) Dorsolateral prefrontal cortex: comparative
value” is revealed over time. In contrast, in our experiment all of cytoarchitectonic analysis in the human and the macaque brain and cor-
ticocortical connection patterns. Eur J Neurosci 11:1011–1036.
the information needed to compute the WTP is revealed at the
Price JL (2006) Connections of orbital cortex. In: the orbitofrontal cortex
beginning of the trials. As a result, anticipatory reward signals, (Zald DH, Raunch SL, eds), pp 39 –55. Oxford: Oxford UP.
which are known to be correlated with NAcc activity (Knutson et Roesch MR, Olson CR (2004) Neuronal activity related to reward value and
al., 2001), might be computed in the Knutson et al. (2007) task, motivation in primate frontal cortex. Science 304:307–310.
but not in the current study. Additional experiments are needed Rolls ET (1996) The orbitofrontal cortex. Philos Trans R Soc Lond B Biol Sci
to systematically explore the differences between the computa- 351:1433–1443, discussion 1443–1434.
tions made in the two experiments. Rolls ET (2000) The orbitofrontal cortex and reward. Cereb Cortex
Part of the research agenda in neuroeconomics is to under- 10:284 –294.
stand how the brain evaluates potential goals and outcomes at the Schoenbaum G, Roesch M (2005) Orbitofrontal cortex, associative learn-
ing, and expectancies. Neuron 47:633– 636.
time of decision making, and how other cognitive, emotional,
Schoenbaum G, Chiba AA, Gallagher M (1998) Orbitofrontal cortex and
and visceral processes affect the computation of economic value. basolateral amygdala encode expected outcomes during learning. Nat
A first step in this research agenda is to understand what are the Neurosci 1:155–159.
brain structures responsible for the computation of value in sim- Tremblay L, Schultz W (1999) Relative reward preference in primate or-
ple everyday choices. Our results suggest that the medial OFC is a bitofrontal cortex. Nature 398:704 –708.
place where a variety of variables computed in other brain regions Wallis JD, Miller EK (2003) Neuronal activity in primate dorsolateral and
are integrated into a single representation of value. If this hypoth- orbital prefrontal cortex during performance of a reward preference task.
esis is correct, other brain processes may be able to influence Eur J Neurosci 18:2069 –2081.
decision making by modulating activity in the medial OFC. Watanabe M (1996) Reward expectancy in primate prefrontal neurons. Na-
ture 382:629 – 632.
References Yeterian EH, Pandya DN (1991) Prefrontostriatal connections in relation to
Arana FS, Parkinson JA, Hinton E, Holland AJ, Owen AM, Roberts AC cortical architectonic organization in rhesus monkeys. J Comp Neurol
(2003) Dissociable contributions of the human amygdala and orbito- 312:43– 67.
The Journal of Neuroscience, September 29, 2010 • 30(39):13095–13104 • 13095

Behavioral/Systems/Cognitive

The Architecture of Reward Value Coding in the Human


Orbitofrontal Cortex
Guillaume Sescousse,1,2 Jérôme Redouté,1,2,3 and Jean-Claude Dreher1,2
1Center for Cognitive Neuroscience, Reward and Decision-Making Group, Centre National de la Recherche Scientifique, Unité Mixte de Recherche 5229,
69675 Bron, France, 2Université Lyon 1, 69003 Lyon, France, and 3CERMEP–Imagerie du Vivant, 69003 Lyon, France

To ensure their survival, animals exhibit a number of reward-directed behaviors, such as foraging for food or searching for mates. This
suggests that a core set of brain regions may be shared by many species to process different types of rewards. Conversely, many new brain
areas have emerged over the course of evolution, suggesting potential specialization of specific brain regions in the processing of more
recent rewards such as money. Here, using functional magnetic resonance imaging in humans, we identified the common and distinct
brain systems processing the value of erotic stimuli and monetary gains. First, we provide evidence that a set of neural structures,
including the ventral striatum, anterior insula, anterior cingulate cortex, and midbrain, encodes the subjective value of rewards regard-
less of their type, consistent with a general hedonic representation. More importantly, our results reveal reward-specific representations
in the orbitofrontal cortex (OFC): whereas the anterior lateral OFC, a phylogenetically recent structure, processes monetary gains, the
posterior lateral OFC, phylogenetically and ontogenetically older, processes more basic erotic stimuli. This dissociation between OFC
representations of primary and secondary rewards parallels current views on lateral prefrontal cortex organization in cognitive control,
suggesting an increasing trend in complexity along a postero-anterior axis according to more abstract representations. Together, our
results support a modular view of reward value coding in the brain and propose that a unifying principle of postero-anterior organization
can be applied to the OFC.

Introduction along a postero-anterior axis, is that the anterior part of the OFC
A basic concern about the functional organization of the prefron- would process secondary rewards, whereas the posterior part
tal cortex is to delineate the functional divisions of the orbito- would process primary rewards (Kringelbach and Rolls, 2004).
frontal cortex (OFC). A number of lesion, electrophysiological, Here, we directly tested this hypothesis with functional mag-
and neuroimaging studies indicate a general role for the OFC in netic resonance imaging (fMRI) by comparing the brain re-
encoding the value assigned to different goods in both human sponses to two experienced rewards: money and erotic pictures.
(O’Doherty et al., 2003a; Plassmann et al., 2007; Chib et al., 2009; These two rewards present significant evolutionary differences
FitzGerald et al., 2009) and nonhuman primates (Tremblay and likely to be reflected at the cerebral level: whereas money is a
Schultz, 1999; Padoa-Schioppa and Assad, 2008). An important secondary reward that appeared recently in human history and
remaining issue, which is key to our understanding of the func- whose abstract value needs to be learned by association with pri-
tional organization of the OFC, is to determine whether distinct mary reinforcers, erotic stimuli can be considered as primary
parts of the OFC encode rewards of different nature. The anterior rewards because they have an innate value and satisfy biological
and posterior parts of the OFC are considered to belong to two needs. We therefore hypothesized that monetary gains would
distinct cytoarchitectonic trends. The anterior part of the OFC, recruit anterior OFC regions and erotic pictures would engage
especially well developed in humans and characterized by a gran- more posterior OFC regions. Despite their critical sociobiological
ular cell layer, is thought to be phylogenetically and ontogeneti- importance, erotic stimuli have never been studied as reinforcers
cally more recent than the posterior and medial parts, which but rather as arousing stimuli in passive viewing paradigms fo-
consist of agranular and dysgranular cortices (Ongür and Price, cusing on sexual function (Redouté et al., 2000; Ponseti et al.,
2000; Wise, 2008). Although never tested empirically, one funda- 2006). However, erotic stimuli are clearly rewarding (Hamann et
mental hypothesis, based on this increasing trend in complexity al., 2004), probably because sexual attractiveness, which may
have evolved to enhance reproductive fitness, is an important cue
for mate choice (Thornhill and Gangestad, 1999; Rhodes, 2006).
Received July 7, 2010; revised July 19, 2010; accepted Aug. 9, 2010.
This work was funded by a Marie Curie International Reintegration Grant (J.-C.D.) and a Fyssen Foundation Grant
In addition to specialized OFC regions processing different
(J.-C.D.). G.S. was funded by a scholarship from the French Ministry of Research and the Medical Research Founda- types of rewards, we hypothesized the existence of common brain
tion. We thank the staff of CERMEP–Imagerie du Vivant for helpful assistance with data collection. structures supporting general hedonic representations indepen-
Correspondence should be addressed to either Guillaume Sescousse or Jean-Claude Dreher, Centre de Neuro- dent of reward type. To evaluate and compare the relative value of
science Cognitive, CNRS UMR 5229, 67 Boulevard Pinel, 69675 Bron Cedex, France, E-mail: gsescousse@isc.cnrs.fr or
dreher@isc.cnrs.fr.
different rewards on a unique scale, it has been proposed that the
DOI:10.1523/JNEUROSCI.3501-10.2010 brain may use a “common neural currency,” likely to be imple-
Copyright © 2010 the authors 0270-6474/10/3013095-09$15.00/0 mented in integrative reward regions such as the ventral striatum
13096 • J. Neurosci., September 29, 2010 • 30(39):13095–13104 Sescousse et al. • Reward Value Coding in the Human OFC

and ventromedial prefrontal cortex (vmPFC) (Montague and tions related to our questions. Experimental trials were divided into two
Berns, 2002; Sugrue et al., 2005; Rangel et al., 2008; Dreher, phases: reward anticipation and outcome. During reward anticipation, a
2009). Consistent with this claim, a number of studies have cue was presented, followed by a delay period and a discrimination task
shown that goal/decision values, reflecting the anticipated re- (Fig. 1). The cue carried three types of information regarding the upcom-
ing reward: the red portion of a pie chart in the background indicated its
warding properties of a stimulus at the time of choice, were en-
probability (25, 50, or 75%), and the pictogram in the foreground indi-
coded in these regions regardless of reward type (McClure et al., cated its type (monetary or erotic) and intensity (high or low, depending
2007; Hare et al., 2008; Chib et al., 2009; Peters and Buchel, 2010). on the size of the pictogram). This led to a total of 12 different cues plus
However, very few studies have investigated whether shared ce- a control condition associated with no chance of winning (supplemental
rebral substrates are engaged when actually experiencing differ- Fig. 1, available at www.jneurosci.org as supplemental material). After a
ent types of rewards, i.e., during the computation of outcome/ variable delay period (question mark representing a pseudorandom draw
hedonic values (Izuma et al., 2008; Smith et al., 2010). Such a depending on probability), subjects were asked to perform a discrimina-
common representation of outcome value may be particularly tion task, in which they had to respond correctly to a target within a
useful to support the trading ability demonstrated by primates. maximum time of 1 s. The shape of the target was drawn at random on
For example, humans are willing to sacrifice money to view at- each trial and could be either a triangle (left button press) or a square
(right button press). Success on this discrimination task (indicated by a
tractive faces (Hayden et al., 2007), and similarly male monkeys
magnified target) allowed the subjects to view the outcome of the pseu-
exchange meat for sex or sacrifice fluid for the opportunity to dorandom draw, whereas erroneous or slow response (indicated by no
view female perinea (Deaner et al., 2005; Gomes and Boesch, change in target size) led to no reward. In rewarded trials, the reward was
2009). either an erotic image (with high or low erotic content) or the picture of
Building on these considerations, we designed a reward para- a safe mentioning the amount of money won (high or low amount). After
digm aiming to (1) determine whether the OFC is functionally each reward outcome, subjects were asked to provide a hedonic rating by
divided depending on reward type and (2) identify common moving a cursor along a 1-to-9 continuous scale (1 for very little pleased;
brain structures processing the experienced value of both mone- 9 for very highly pleased). In non-rewarded and control trials, the sub-
tary and erotic rewards. To further characterize the role of these jects were presented with “scrambled” pictures. A fixation cross was fi-
brain regions in reward processing, we also manipulated reward nally used as an intertrial interval of variable length.
intensity and probability and collected hedonic ratings after re- Stimuli
ward outcomes inside the scanner. Because erotic pictures are not Two categories (high and low intensity) of erotic pictures and monetary
quantifiable like money, it is unclear whether changes in the in- gains were used. Nudity being the main criteria driving the reward value
tensity of erotic pictures would affect the same brain regions as of erotic stimuli, we separated them into a “low intensity” group display-
those responding to changes in monetary amounts. Likewise, it is ing women in underwear or bathing suits and a “high intensity” group
unknown whether the concept of prediction error, reflecting the displaying naked women in an inviting posture. Each erotic picture was
discrepancy between expected and actual rewards and primarily presented only once during the course of the task to avoid habituation. A
used with quantifiable rewards (such as money and juice), can be similar element of surprise was introduced for the monetary rewards by
randomly varying the amounts at stake: the low amounts were €1, €2, or
extended to erotic stimuli.
€3 and the high amounts were €10, €11, or €12. The pictures displayed in
non-rewarded and control trials were scrambled versions of the pictures
Materials and Methods used in rewarded trials and hence contained the same information in
Participants terms of chromaticity and luminance.
Eighteen right-handed volunteers (mean ! SD age, 24 ! 3.3 years) with
no history of neurological or psychiatric disorders participated in this fMRI data acquisition
study. All of them were heterosexual males, because men are generally Imaging was conducted on a 1.5 T Siemens Sonata scanner, using an
more responsive to visual sexual stimuli than women (Hamann et al., eight-channel head coil. The scanning session was divided into four runs.
2004) and to avoid the potential influence of the menstrual cycle known Each of them included four repetitions of each cue, with the exception of
to have an effect on reward processing in women (Caldú and Dreher, the control condition, repeated nine times. This yielded a total of 228
2007; Dreher et al., 2007). All subjects gave written informed consent to trials. Within each run, the order of the different conditions was pseudo-
be part of the experiment, which was approved by the local ethics com- randomized and optimized to improve signal deconvolution. The order
mittee (Centre Léon Bérard, Lyon, France). of the runs was counterbalanced between subjects. Before scanning, all
Motivation, which was a crucial element of our study, was closely subjects were given oral instructions and familiarized with the cognitive
controlled. First, sexual arousability was assessed at the time of screening task in a short training session.
through specific questionnaires, namely the Brief Sexual Function ques- Each of the four functional runs consisted of 296 volumes. Twenty-six
tionnaire (Reynolds et al., 1988) and the Sexual Arousability Inventory interleaved slices parallel to the anterior commissure–posterior commis-
(Hoon and Chambless, 1998). Of an initial pool of 22 subjects, two of sure line were acquired per volume (field of view, 220 mm; matrix, 64 "
them were excluded because they scored too low on the Sexual Arous- 64; voxel size, 3.4 " 3.4 " 4 mm; gap, 0.4 mm), using a gradient-echo
ability Inventory (mean score for all subjects, 91.1 ! 14.6; scores of echoplanar imaging (EPI) T2*-weighted sequence (repetition time, 2500
excluded subjects, 54 and 64). To further ensure that all participants ms; echo time, 60 ms; flip angle, 90°). To improve the local field homo-
would be in a similar state of motivation to see erotic stimuli, we asked geneity and hence minimize susceptibility artifacts in the orbitofrontal
them to avoid any sexual contact during a period of 24 h before the area, a manual shimming was performed within a rectangular region
scanning session. Second, we sought to enhance the motivation for including the OFC and the basal ganglia. A high-resolution T1-weighted
money by telling the subjects that the financial compensation for their structural scan was subsequently acquired in each subject.
participation would be calculated based on their winnings during the
task. We also excluded two subjects presenting symptoms of depression fMRI analysis
as assessed by the 13-item version of the Beck Depression Inventory Preprocessing. Preprocessing of fMRI data was conducted using SPM2.
(Beck and Beck, 1972) (mean score for all subjects, 1.9 ! 2.7; scores of The first four functional volumes of each run were removed, and the
excluded subjects, 6 and 10). remaining images were corrected for slice-timing artifacts and spa-
tially realigned to the first image of each time series. We then searched
Task for residual artifacts in the time series with the tsdiffana utility (http://
Our protocol was inspired from the typical design of incentive delay tasks imaging.mrc-cbu.cam.ac.uk/imaging/DataDiagnostics) and modeled
(Knutson et al., 2005; Abler et al., 2006) but included several modifica- them with dummy regressors in our general linear model (two subjects
Sescousse et al. • Reward Value Coding in the Human OFC J. Neurosci., September 29, 2010 • 30(39):13095–13104 • 13097

Figure 1. Paradigm and behavior. A, Sequence of events during a typical trial. Subjects first saw a cue informing them about the type, probability, and intensity of an upcoming reward
(supplemental Fig. 1, available at www.jneurosci.org as supplemental material). Three cases are represented here: a 75% chance of receiving a high amount of money (top), a 25% chance of seeing
a low erotic content picture (middle), and a sure chance of getting nothing (control trials; bottom). After a short delay and a target discrimination task, subjects saw the outcome, which was
contingent on both the announced probability and their performance on the discrimination task. Reward outcomes consisted either in a monetary amount displayed on a safe (top) or an erotic
picture (middle) and were followed by the rating of their subjective value on a continuous scale. Non-rewarded and control trials displayed a scrambled picture at outcome (bottom). B, Behavioral
results on the discrimination task: mean reaction times according to reward intensity (left) and probability (middle) and mean hit rates according to reward intensity (right). C, Mean subjective
ratings according to reward intensity, on a 1-to-9 scale. Error bars indicate SEM. *p # 0.05; **p # 0.01; ***p # 0.001 by Tukey’s HSD tests.

had three artifacts and one subject had one artifact in their time series). Although the anticipatory period was explicitly modeled in our anal-
The functional images were then normalized to the Montreal Neurolog- ysis, we only report results concerning the outcome phase because our
ical Institute (MNI) stereotaxic space using the EPI template of SPM2 focus was on the coding of experienced reward value. Following the
and spatially smoothed with a 10 mm full-width at half-maximum iso- forward inference approach (Henson, 2006) (see Results), brain regions
tropic Gaussian kernel. Anatomical scans were normalized to the MNI responding specifically to monetary (or erotic) rewards resulted from the
space using the icbm152 template brain and averaged across subjects. contrast MR " ER (or ER " MR), masked inclusively with MR " C (or
Identification of common and specific brain regions. The event-related ER " C) and exclusively with ER " C (or MR " C). The main contrasts
statistical analysis was performed according to the general linear model MR " ER and ER " MR and the masks were thresholded independently
as implemented in SPM2. Anticipation-related responses were modeled using a whole-brain correction for multiple comparisons [p # 0.05 fami-
as boxcar functions time locked to the onset time of the cue with a lywise error (FWE) and p # 0.01 false discovery rate (FDR), respectively],
duration of 2.5 s. The 2 rewards (monetary/erotic) ! 2 intensities (high/ thereby ensuring the absence of any “selection bias” in the analysis
low) were modeled as four separate conditions. For each of them, a (Kriegeskorte et al., 2009). Brain regions activated by both monetary and
first-order parametric regressor modeled reward probability. The con- erotic rewards were identified in two steps. We first performed a con-
trol condition was modeled in a separate regressor. Outcome-related junction analysis of the contrasts MR " C and ER " C based on the
responses were modeled as events time locked to the appearance of the minimum statistic (Nichols et al., 2005) ( p # 0.05 FWE whole-brain
reward (or scrambled picture). Four main conditions were defined: corrected). Because this conjunction may be sensitive to other types of
“monetary reward” (MR), “erotic reward” (ER), “no-monetary reward” computations (such as attention or image processing), we then masked it
(NoMR), and “no-erotic reward” (NoER). Two orthogonalized covari- inclusively with the regions responding parametrically with both mone-
ates linearly modeling the expected probability and the ratings were tary and erotic hedonic ratings (each mask thresholded at p # 0.05 FDR
added (in this order) to the MR and ER regressors. A last regressor whole-brain corrected).
modeled the appearance of a scrambled picture in the control condi- Anatomical localization of functional clusters was performed based on
tion (C). All regressors were subsequently convolved with the canon- a probabilistic atlas (Hammers et al., 2003).
ical hemodynamic response function and entered in a first-level Prediction error model. Positive prediction errors were defined at each
analysis. A high-pass filter with a cutoff of 128 s was applied to the trial t by PE(t) $ V(t) % B(t), where V(t) is the outcome value and B(t) is
time series to remove low-frequency noise and baseline drifts. The the expected value (Yacubian et al., 2006). Whereas monetary amounts
resulting images of parameter estimates were then passed in a second- could have been used to assess V(t) for monetary rewards, erotic rewards
level group analysis in which between-subject variability was treated could not be similarly quantified. Hence, to use an equivalent measure
as a random effect. for both rewards, we used the hedonic rating to assess V(t) on each trial.
13098 • J. Neurosci., September 29, 2010 • 30(39):13095–13104 Sescousse et al. • Reward Value Coding in the Human OFC

Figure 2. Functional postero-anterior dissociation in the orbitofrontal cortex depending on reward type. Brain regions responding specifically to monetary reward outcomes are displayed in
blue– green, and those responding specifically to erotic reward outcomes are displayed in red–yellow. Plots of mean percent signal change, which are not independent of the whole-brain analysis,
are shown only to illustrate the double dissociation between monetary/erotic rewards and anterior (Ant.)/posterior (Post.) OFC. Activations are overlaid on an average anatomical scan of all subjects
( p " 0.05 FWE whole-brain corrected). Error bars indicate SEM.

B(t) was defined as the product of reward probability P(t) by expected


intensity E(t). P(t) was simply the probability given explicitly in the cue
(25, 50, or 75%). E(t), presented as either “high” or “low” to the partic-
ipants, was transformed into a numerical value by using the past ratings:
for instance, E(t) for a high monetary reward was estimated as the average
of all the ratings given to high monetary rewards since the beginning of
the task up to trial t.
Prediction error values were entered into two parametric regressors
separately modeling monetary and erotic reward prediction errors
(PEMR and PEER). Note, however, that prediction error was relatively
correlated with outcome value, i.e., with the hedonic ratings (mean r !
0.72 for monetary rewards and mean r ! 0.75 for erotic rewards). This
correlation is inherent to the nature of these signals and is a classical
shortcoming of fMRI studies on reward processing (Hare et al., 2008). As
a consequence, if prediction errors and hedonic ratings were entered in
the same general linear model, they would both end up with a rather low Figure 3. Specific response of amygdala to erotic rewards. Activations are overlaid on an
explanatory power, because only the orthogonal component of each re- average anatomical scan of all subjects ( p " 0.05 FWE whole-brain corrected). Left and right
gressor would be allowed to compete for variance (Hunt, 2008) (such a plots of mean percent signal change, which are not independent of the whole-brain analysis,
model was estimated and produced poor results in expected brain re- are shown only to illustrate the specificity of amygdalar response. Error bars indicate SEM.
gions such as the ventral striatum and the OFC). For this reason, we built
a separate general linear model, in which reward outcomes were modu-
lated only by prediction errors (instead of probability and ratings). We condition (! value) divided by the mean activity of that ROI and multi-
should emphasize, however, that this procedure makes it difficult to plied by 100.
distinguish the contribution of prediction error and outcome value com-
putations at the brain level. Results
The resulting T-maps showing positive correlations between the Behavior
blood oxygenation level-dependent (BOLD) signal and monetary or Hit rates and reaction times (RTs), obtained at the time of the
erotic prediction errors were subsequently entered in a conjunction discrimination task, as well as hedonic ratings obtained at the
analysis (Nichols et al., 2005) thresholded at p " 0.001 uncorrected time of outcome, were analyzed in separate three-way ANOVAs
for multiple comparisons at the voxel level. including reward type, probability, and intensity as within-
Region of interest analyses. Region of interest (ROI) analyses were con- subject factors. The analysis on hit rates and RT was performed
ducted with the extension of SPM MarsBaR (http://marsbar.sourceforge.
on 17 subjects only, because data were accidentally lost for one
net/) within ROIs defined functionally from the whole-brain analyses.
Each ROI was created by taking the intersection of the functional cluster subject. The mean hit rate across subjects on the discrimination
of interest and a 10-mm-radius sphere centered on the highest peak voxel task was 96%.
of the cluster (to isolate distinct brain areas pertaining to the same clus- There was no significant main effect of reward type on hit
ter). In keeping with the approach of MarsBaR, percent signal change for rates ( p ! 0.38) and RT ( p ! 0.20), suggesting that monetary
a given condition in a given ROI was calculated as the effect size of that gains and erotic pictures had comparable incentive values.
Sescousse et al. • Reward Value Coding in the Human OFC J. Neurosci., September 29, 2010 • 30(39):13095–13104 • 13099

Figure 4. Response pattern of the reward-specific brain regions as a function of reward intensity. Percent signal change is plotted in the circled ROIs for monetary and erotic rewards according
to the following conditions: high intensity, low intensity, and no reward. In each region, brain activity increases with reward intensity only for the reward for which it is specific. Error bars indicate
SEM. The signal is averaged across the right and left hemispheres in each brain region (similar patterns of activity were observed in each hemisphere). Ant., Anterior; Post., posterior.

Subjects were faster (F(1,16) ! 34.2, p " 0.001) and more ac- meet two criteria: (1) they are more activated by one reward com-
curate (F(1,16) ! 7.7, p " 0.05) for high intensity incentives and pared with the other (“dissociation” criterion); (2) they respond to
were also faster for more likely rewards (F(2,32) ! 5.3, p " 0.05) either monetary or erotic rewards, but not to both, compared with a
(Fig. 1 B). These results reflect increased motivation for higher common control condition (“association” criterion).
reward intensity and more certain rewards. They also confirm As hypothesized, monetary rewards specifically recruited the
that subjects were engaged in the task and effectively encoded the anterior lateral OFC (MNI [x y z] [$30, 51, 0], T ! 5.92; [30, 54,
cue information. Importantly, these effects were similar for mon- $3], T ! 6.80), spanning the anterior orbital gyrus, the lateral
etary and erotic rewards, as shown by the absence of significant orbital gyrus, and the ventral part of the middle frontal gyrus (Fig.
interaction between intensity and reward type (hit rate, p ! 0.67; 2). In contrast, erotic rewards elicited activity specifically in the
RT, p ! 0.20) (supplemental Fig. 2, available at www.jneurosci. posterior part of the lateral OFC ([$30, 33, $15], T ! 7.56; [30,
org as supplemental material) and between probability and re- 33, $15], T ! 7.54), straddling the posterior and lateral orbital
ward type (RT, p ! 0.11). gyri (Fig. 2). These results demonstrate a double dissociation
No significant effect of reward type was observed on the he- between monetary/erotic rewards and the anterior/posterior
donic ratings ( p ! 0.40), suggesting that monetary and erotic OFC, which is further illustrated in the bar graphs of Figure 2,
rewards had similar subjective values. Conversely, we found a representing the MR # C and ER # C differences in percent
robust main effect of intensity on the ratings (F(1,17) ! 150.8, p " signal change extracted from these regions. Among erotic-
0.001), which remained significant for each type of reward taken specific areas, a large cluster was also present in the medial OFC
separately [Tukey’s honestly significant difference (HSD) tests: ([$6, 45, $15], T ! 8.90), encompassing the medial orbital gy-
monetary rewards, T(17) ! 20.4, p " 0.001; erotic rewards,
rus, the straight gyrus, and the most ventral part of the superior
T(17) ! 4.4, p " 0.001] (Fig. 1C). This shows that, for both re-
frontal gyrus (Fig. 2). Subcortically, the only structure specifically
wards, the two intensity categories chosen a priori (high vs low)
activated by erotic pictures was the bilateral amygdala ([$21, $6,
were effectively perceived by the subjects. The ratings also showed
$27], T ! 6.94; [24, 0, $27], T ! 5.45) (Fig. 3). Other money-
an interaction between reward type and intensity (F(1,17) ! 111.5,
specific and erotic-specific foci are reported in supplemental Ta-
p " 0.001), simply because the subjects used a smaller portion of the
scale to rate erotic pictures. Finally, the ratings were not influenced bles 1 and 2 (available at www.jneurosci.org as supplemental
by reward probability (F(2,34) ! 1.4, p ! 0.26), confirming that they material), respectively.
reflected a purely hedonic evaluation. This segregated representation of reward types in the OFC was
not merely attributable to visual or hedonic differences between
Neuroimaging data monetary and erotic outcomes, because an identical dissociation
Reward-specific brain regions emerged when we repeated the same analysis using no-reward
Brain regions specific for each type of reward were identified outcomes instead of reward outcomes. That is, when comparing
based on the forward inference approach proposed by Henson the NoMR and NoER conditions, which only differ with respect
(2006) to demonstrate qualitative differences in brain imaging to the type of reward being expected (while comparing visually
data. Specifically, money-specific regions were defined as those identical scrambled pictures), the anterior OFC specifically re-
stemming from the comparison MR # ER, masked inclusively sponded to no-monetary outcomes ([$30, 51, 0], T ! 6.16; [33,
with MR # C and exclusively with ER # C, and conversely erotic- 51, 0], T ! 6.20), whereas the posterior OFC specifically re-
specific regions were defined as those responding in the compar- sponded to no-erotic outcomes ([$21, 33, $12], T ! 6.39) (sup-
ison ER # MR, masked inclusively with ER # C and exclusively plemental Fig. 3, supplemental Tables 1, 2, available at www.
with MR # C. This procedure ensures that the resulting brain areas jneurosci.org as supplemental material). This result excludes a mere
13100 • J. Neurosci., September 29, 2010 • 30(39):13095–13104 Sescousse et al. • Reward Value Coding in the Human OFC

perceptual account of the functional disso-


ciation observed in the OFC and supports
the idea that monetary and erotic rewards
are encoded in distinct OFC regions.
Moreover, ROI analyses on reward
intensity coding brought additional evi-
dence in support of a segregated represen-
tation of monetary and erotic outcomes.
For both rewards, we extracted the per-
cent signal change for the high reward,
low reward, and no-reward conditions
(Fig. 4). The results show that, whereas
activity in the anterior OFC increased with
monetary reward intensity, such a mono- Figure 5. Brain regions reflecting hedonic ratings regardless of reward type. Activations show the brain areas in which activity
tonic variation was not present for increas- positively correlates with both monetary and erotic ratings (intersection of T-maps thresholded at p # 0.01 FDR whole-brain
ing levels of erotic reward intensity. corrected shown in yellow or thresholded at p # 0.05 FDR whole-brain corrected shown in red). The regions displayed in red were
Conversely, in the posterior OFC, medial used as an inclusive mask in the analysis of Figure 6. Activations are overlaid on an average anatomical scan of all subjects.
OFC, and amygdala, activity was found to
increase monotonically with erotic reward
intensity but not with monetary reward in-
tensity. Together, these findings indicate
that reward-specific brain regions only re-
flected intensity for the reward type they
specifically encoded.
Finally, to further confirm the specificity
of these regions with respect to the subjec-
tive experience of monetary and erotic out-
comes, we performed a whole-brain
analysis comparing the coding of their he-
donic value. For each reinforcer, reward and
no-reward outcomes were pooled together,
while the hedonic value was modeled in a
parametric regressor with the correspond-
ing continuous rating (from 1 to 9) or a 0,
respectively. As expected, the contrast of
these parametric regressors between the two
rewards revealed the same brain regions as
the previous categorical analysis. In particu-
lar, BOLD activity in the right anterior lat-
eral OFC was found to scale best with
monetary hedonic value, whereas activity in
the posterior lateral OFC, medial OFC, and
amygdala was found to scale best with erotic
hedonic value (supplemental Fig. 4, avail- Figure 6. Common reward brain regions. A, T-map showing the brain regions encoding experienced reward value for both
able at www.jneurosci.org as supplemental monetary and erotic reward outcomes. Activations are overlaid on an average anatomical scan of all subjects ( p # 0.05 FWE
material). whole-brain corrected). B, Percent signal change is plotted in the circled ROIs for monetary and erotic rewards according to the
following conditions: high intensity, low intensity, and no reward. Note that these plots are not independent of the whole-brain
Common reward brain regions analysis and are only shown as an illustration for easier visual comparison with Figure 4. Error bars indicate SEM. The signal is
To identify the brain regions commonly averaged across the right and left hemispheres in each brain region (similar patterns of activity were observed in each hemisphere).
activated by monetary and erotic out- Ant., Anterior.
comes, we first compared each reward
with the control condition and performed 39], T " 8.42), and the anterior insula ([!27, 21, !6], T " 7.48;
a conjunction of these two comparisons (supplemental Fig. 5, [33, 24, 3], T " 8.14) (Fig. 6 A). Other foci are reported in sup-
available at www.jneurosci.org as supplemental material). Be- plemental Table 3 (available at www.jneurosci.org as supplemen-
cause this analysis remains sensitive to non-reward-related com- tal material).
putations such as attentional or image processing effects, it was The relative coding of reward intensity in these regions was
masked by the brain regions responding parametrically with the further illustrated using the same ROI approach as with reward-
subjective value of both monetary and erotic rewards (i.e., hedo- specific regions (Fig. 6 B). Note, however, that this representation
nic ratings) (Fig. 5). This procedure revealed significant bilateral is not independent of the previous whole-brain analysis, because
activations in a set of brain regions classically involved in reward reward intensity is highly correlated with the hedonic ratings that
processing: the ventral striatum ([!12, 9, !9], T " 6.38; [9, 6, have served in the identification of the “common network.” Con-
!9], T " 6.20), the midbrain ([!3, !24, !24], T " 7.30), the sequently, the resulting bar graphs are purely illustrative of the
anterior cingulate cortex (ACC) [!6, 27, 39], T " 8.15; [9, 18, conclusion drawn from the T-map, which is that activity in the
Sescousse et al. • Reward Value Coding in the Human OFC J. Neurosci., September 29, 2010 • 30(39):13095–13104 • 13101

Postero-anterior dissociation in the


orbitofrontal cortex
The segregated responses to monetary
and erotic outcomes along a postero-
anterior axis in the OFC suggest a func-
tional division of experienced reward
value representation according to an ab-
stractness gradient. Paralleling this result
in the domain of cognitive control, recent
theories on the functional divisions of the
human lateral prefrontal cortex proposed
Figure 7. Brain regions reflecting prediction errors regardless of reward type. Activations result from a conjunction analysis that it is organized hierarchically, whereby
showing the brain regions in which activity positively correlates with both monetary and erotic prediction errors (p # 0.001 cognitive control involving temporally
uncorrected). Activations are overlaid on an average anatomical scan of all subjects.
proximate and concrete action represen-
tations is supported by posterior lateral
prefrontal regions, and cognitive control
ventral striatum, midbrain, ACC, and anterior insula reflects ex- involving temporally extended and ab-
perienced reward value regardless of reward type. stract representations is supported by more anterior lateral pre-
frontal regions such as the frontopolar cortex (Koechlin and
Coding of reward prediction errors Summerfield, 2007; Badre, 2008; Dreher et al., 2008a). Our study
Reward-related brain regions are thought to mediate prediction
shows that a similar unifying principle of caudo-rostral hierar-
error signals, coding the difference between expected outcomes
chical organization can be applied to the OFC. Notably, patients
and those effectively delivered (McClure et al., 2003; O’Doherty
with lesions in the anterior OFC have been reported to be specif-
et al., 2003b; Schultz, 2006). These prediction error signals occur
not only in conditioning procedures and can also be computed in ically impaired in making decisions entailing abstract, i.e., dis-
non-learning situations, such as in the present study (Dreher et tant, consequences, and not in making decisions leading to
al., 2006; Yacubian et al., 2006). To determine whether such sig- concrete, i.e., immediate, consequences, further supporting a
nals are supported by similar brain networks for monetary and postero-anterior trend in the representation of abstractness in the
erotic rewards, the fMRI data were fitted with parametric regres- OFC (Bechara and Damasio, 2005).
sors modeling positive prediction errors related to reward out- Although an anatomical gradient in the postero-anterior
comes. Our results revealed that monetary and erotic reward axis of the OFC had been suggested based on cytoarchitectonic
prediction errors were processed in a similar set of brain regions, data (Ongür and Price, 2000; Wise, 2008), its functional rele-
essentially overlapping with the previously identified common vance for reward processing had never been tested empirically.
network. Specifically, a conjunction analysis showed that activity Our data bring strong empirical support to this hypothesis.
in the ventral striatum ([!9, 9, !6], T " 3.98; [6, 9, !9] T " Although a generalization to all primary and secondary re-
4.57), anterior insula ([!33, 18, !18], T " 4.80; [39, 21, !15], wards cannot be ascertained in a single fMRI study, it is con-
T " 3.37), and rostral ACC ([!3, 42, 15], T " 5.57) correlated sistent with the pattern of activation observed in the recent
positively with prediction errors regardless of reward type (Fig. 7) literature for other rewards (Kringelbach and Rolls, 2004). In
(supplemental Table 4, available at www.jneurosci.org as supple- particular, the medial and posterior lateral OFC, responding
mental material). Moreover, the comparison of monetary and to erotic pictures in the current and previous studies (Ponseti
erotic reward prediction errors revealed almost no activity in et al., 2006), were shown to respond to other primary rewards,
reward-related areas (at p # 0.001 uncorrected, only a tiny cluster such as attractive faces (O’Doherty et al., 2003a), pleasant
appeared in the left pallidum for the contrast of erotic minus odors (Gottfried et al., 2006), and pleasant taste (Small et al.,
monetary prediction errors) and especially not in the brain re- 2001). Conversely, other studies manipulating monetary (Re-
gions labeled as “reward specific” (even at a liberal threshold of uter et al., 2005; Vollm et al., 2007) or social (Izuma et al.,
p # 0.05). 2008) rewards have reported a similar anterior OFC region as
the one we found.
Discussion One intrinsic property of erotic pictures is that they are the
As predicted, the OFC was found to be functionally organized reward, whereas monetary rewards delivered in the scanner
along a postero-anterior axis with respect to reward type, with the are a representation of what the participant will receive at the
anterior part responding exclusively to money and the posterior end of the experiment. Thus, it could be argued that the OFC
part responding exclusively to erotic stimuli. Additional erotic- dissociation relates to the immediate rewarding effect of erotic
specific activations were also found in the bilateral amygdala and pictures compared with the delayed rewarding effect of mon-
medial OFC. Importantly, brain activity in these reward-specific etary gains. This is unlikely to be the case, because (1) our
regions only scaled with the hedonic value of the reward they pattern of OFC activation was not observed for immediate
specifically encoded. In parallel, our results support the idea of a core versus delayed rewards in intertemporal choice studies (Kable
reward system processing experienced rewards regardless of their and Glimcher, 2007; Prevost et al., 2010) and (2) the same
nature. In this common network, including the ventral striatum, OFC functional dissociation emerged when monetary and
ACC, anterior insula, and midbrain, functional activity correlated erotic rewards were expected but not effectively delivered.
with hedonic value and prediction error for both monetary gains This finding also suggests that the segregated representation of
and erotic pictures. Together, our results reveal the existence of both reward types in the OFC was not merely attributable to visual
reward-specific and nonspecific brain networks, challenging the or intrinsic differences between monetary and erotic out-
view of a unique reward system for all reinforcers. comes, such as saliency or arousal.
13102 • J. Neurosci., September 29, 2010 • 30(39):13095–13104 Sescousse et al. • Reward Value Coding in the Human OFC

Electrophysiological recordings in monkeys indicate that 2007; Knutson et al., 2007; Chib et al., 2009). The present results
OFC neurons encode the economic value assigned to different suggest that a similar computation might be performed in the
rewarding juices when choosing between them (Padoa- same brain network at the time of reward consumption. Indeed,
Schioppa and Assad, 2008). Some OFC neurons were also along with the ventral striatum, activity in the vmPFC also re-
found to encode taste responses reflecting the identity of a flected the hedonic experience of the participants regardless of
chosen juice. However, no neuronal recording study has yet reward type (Fig. 5). Such a mechanism encoding heterogeneous
investigated whether primary and secondary rewards (e.g., outcome values on a common scale might be helpful for efficient
juice vs social dominance) are coded in distinct OFC subre- comparison of these values during subsequent value-based deci-
gions. Our finding of a clear OFC dissociation in humans sion making.
supports a hierarchical organization along a continuum from Finally, we found prediction-error-related activity in the
the posterior to the anterior part of the OFC. Whether and ventral striatum, anterior insula, and ACC for both monetary
how these distinct representations of value in the OFC are and erotic stimuli. This finding demonstrates that prediction
preserved across species remains an important open question. errors are computed in a network independent of reward type
Our results also suggest that learning through secondary rein- and generalizes this concept to the domain of erotic stimuli,
forcement may depend more on anterior OFC regions, paralleling previous work performed with monetary gains
whereas primary reinforcement may depend more on the pos- (Yacubian et al., 2006), pleasant taste (O’Doherty et al.,
terior OFC. This hypothesis, which remains to be tested, may 2003b), and attractive faces (Bray and O’Doherty, 2007).
shed light on studies across a range of animal species indicat- Thus, prediction errors may be a primitive neural signal com-
ing effects of OFC lesions on behavior maintained or acquired puted in midbrain dopaminergic neurons regardless of reward
through secondary reinforcement (Murray et al., 2007). type and primarily delivered to the common reward network
In addition, our findings clarify how value signals in the that may be responsible for making predictions. However, it is
OFC are integrated with those from other brain structures. important to emphasize that, because of the inherent correla-
Together with the posterior and medial OFC, we found that tion existing between prediction error and hedonic value, it is
the amygdala responded exclusively to erotic pictures. Previ- difficult to disentangle these computations at the brain level
ous neuroimaging studies also reported that erotic pictures (Behrens et al., 2008; Hare et al., 2008). Consequently, al-
evoke amygdala response (Redouté et al., 2000; Karama et al., though the response pattern observed in our common net-
2002; Hamann et al., 2004), whereas money, as a secondary work is consistent with both interpretations, we cannot
reinforcer, often failed to do so in studies using experimental ascribe with certainty one role or the other to these regions.
designs similar to the present paradigm (Knutson et al., 2001). Concerning the ventral striatum, however, a recent study aim-
This is consistent with the underlying anatomy, showing that ing to dissociate reward value from prediction error found
the amygdala is more connected with the posterior and medial that activity in this region best correlated with the latter (Hare
OFC than with the anterior OFC (Carmichael and Price, et al., 2008). Although this study focused on the computation
1995). Moreover, the parametric modulation of the amygdala of goal values in the context of decision making, and therefore
with erotic hedonic value is in accordance with its general role cannot be directly compared with the present one, it brings
in emotional arousal for both appetitive and aversive stimuli. evidence favoring the prediction error hypothesis in the ven-
However, amygdala response to erotic stimuli is not solely tral striatum. Note that midbrain activity was observed in our
determined by arousal, as suggested by a previous study re- common network, but not in the prediction error analysis,
porting higher amygdala response in men than in women whereas single-neuron recordings classically report prediction
viewing sexual stimuli, despite similar arousal ratings in both error signals in midbrain dopaminergic structures (Schultz,
groups (Hamann et al., 2004). 2006). Although further work is needed to gain a better un-
derstanding of the relationship between dopaminergic neuron
Brain regions common to monetary and erotic rewards firing and the BOLD signal observed in reward paradigms,
An important strength of our experimental design, compared recent findings combining fMRI with FDOPA positron emis-
with previous reward studies, is that it made it possible to directly sion tomography measures of midbrain dopamine synthesis
test whether monetary and erotic reward outcomes are truly en- (Dreher et al., 2008b) and analyses of variations in genes in-
coded within the same brain regions. The enhanced response to volved in dopamine transmission established a link between
increasing hedonic value observed in the common network re- higher prefronto-striatal BOLD signal and dopamine synaptic
gardless of reward type suggests that this network processes ex- availability during reward processing (Dreher et al., 2009).
perienced reward value in a general manner. This is consistent
with its reported implication (in separate studies) in the hedonic Conclusion
processing of rewards ranging from primary reinforcers, such as Our results provide plausible functional mechanisms explaining
sexual stimuli (Redouté et al., 2000), attractive faces (Bray and the existence of two separate reward networks in the brain. The
O’Doherty, 2007; Smith et al., 2010), and pleasant taste (Small et nature of their interactions remains to be determined, but one
al., 2001), to secondary rewards, such as money or social approval possibility is that outcome value signals computed in the reward-
(Izuma et al., 2008). Our common network activity is also com- specific OFC regions would be sent to the common network for
patible with an interpretation in terms of general arousal or sa- additional integration and comparison processes. From an evo-
liency (Zink et al., 2004). lutionary perspective, the distinct cytoarchitectonic properties of
To efficiently compare the goal values of different rewards the anterior and posterior parts of the OFC suggest that the ability
during decision making, it has been proposed that the brain may to process primary rewards may occur phylogenetically and on-
convert them into a common neural currency (Montague and togenetically earlier than the ability to process secondary rewards,
Berns, 2002). A wealth of electrophysiological and fMRI studies which represent more evolved adaptive behavior. Our findings
has since confirmed this hypothesis, emphasizing in particular also have important clinical implications for a range of neuropsy-
the role of the ventral striatum and vmPFC (Kable and Glimcher, chopathological disorders characterized by major deficits in mo-
Sescousse et al. • Reward Value Coding in the Human OFC J. Neurosci., September 29, 2010 • 30(39):13095–13104 • 13103

tivation and behavioral control, such as pathological gambling or Hunt LT (2008) Distinctive roles for the ventral striatum and ventral pre-
hypersexuality. The dissociable representation of various rewards frontal cortex during decision-making. J Neurosci 28:8658 – 8659.
Izuma K, Saito DN, Sadato N (2008) Processing of social and monetary
along a postero-anterior axis in the OFC may shed light on this
rewards in the human striatum. Neuron 58:284 –294.
important question. Kable JW, Glimcher PW (2007) The neural correlates of subjective value
during intertemporal choice. Nat Neurosci 10:1625–1633.
Karama S, Lecours AR, Leroux JM, Bourgouin P, Beaudoin G, Joubert S,
References Beauregard M (2002) Areas of brain activation in males and females
Abler B, Walter H, Erk S, Kammerer H, Spitzer M (2006) Prediction error as
during viewing of erotic film excerpts. Hum Brain Mapp 16:1–13.
a linear function of reward probability is coded in human nucleus accum-
Knutson B, Fong GW, Adams CM, Varner JL, Hommer D (2001) Dissocia-
bens. Neuroimage 31:790 –795.
tion of reward anticipation and outcome with event-related fMRI. Neu-
Badre D (2008) Cognitive control, hierarchy, and the rostro-caudal organi-
roreport 12:3683–3687.
zation of the frontal lobes. Trends Cogn Sci 12:193–200.
Knutson B, Taylor J, Kaufman M, Peterson R, Glover G (2005) Distributed
Bechara A, Damasio A (2005) The somatic marker hypothesis: A neural
neural representation of expected value. J Neurosci 25:4806 – 4812.
theory of economic decision. Games Econ Behav 52:336 –372.
Knutson B, Rick S, Wimmer GE, Prelec D, Loewenstein G (2007) Neural
Beck AT, Beck RW (1972) Screening depressed patients in family practice. A
predictors of purchases. Neuron 53:147–156.
rapid technic. Postgrad Med 52:81– 85.
Koechlin E, Summerfield C (2007) An information theoretical approach to
Behrens TE, Hunt LT, Woolrich MW, Rushworth MF (2008) Associative
prefrontal executive function. Trends Cogn Sci 11:229 –235.
learning of social value. Nature 456:245–249.
Kriegeskorte N, Simmons WK, Bellgowan PS, Baker CI (2009) Circular
Bray S, O’Doherty J (2007) Neural coding of reward-prediction error sig-
nals during classical conditioning with attractive faces. J Neurophysiol analysis in systems neuroscience: the dangers of double dipping. Nat Neu-
97:3036 –3045. rosci 12:535–540.
Caldú X, Dreher JC (2007) Hormonal and genetic influences on processing Kringelbach ML, Rolls ET (2004) The functional neuroanatomy of the hu-
reward and social information. Ann NY Acad Sci 1118:43–73. man orbitofrontal cortex: evidence from neuroimaging and neuropsy-
Carmichael ST, Price JL (1995) Limbic connections of the orbital and chology. Prog Neurobiol 72:341–372.
medial prefrontal cortex in macaque monkeys. J Comp Neurol McClure SM, Berns GS, Montague PR (2003) Temporal prediction errors in
363:615– 641. a passive learning task activate human striatum. Neuron 38:339 –346.
Chib VS, Rangel A, Shimojo S, O’Doherty JP (2009) Evidence for a common McClure SM, Ericson KM, Laibson DI, Loewenstein G, Cohen JD (2007)
representation of decision values for dissimilar goods in human ventro- Time discounting for primary rewards. J Neurosci 27:5796 –5804.
medial prefrontal cortex. J Neurosci 29:12315–12320. Montague PR, Berns GS (2002) Neural economics and the biological sub-
Deaner RO, Khera AV, Platt ML (2005) Monkeys pay per view: adaptive strates of valuation. Neuron 36:265–284.
valuation of social images by rhesus macaques. Curr Biol 15:543–548. Murray EA, O’Doherty JP, Schoenbaum G (2007) What we know and do
Dreher JC (2009) Decomposing brain signals involved in value-based deci- not know about the functions of the orbitofrontal cortex after 20 years of
sion making. In: Handbook of reward and decision making (Dreher JC, cross-species studies. J Neurosci 27:8166 – 8169.
Tremblay L, eds), pp 137–164. New York: Academic/Elsevier. Nichols T, Brett M, Andersson J, Wager T, Poline JB (2005) Valid conjunc-
Dreher JC, Kohn P, Berman KF (2006) Neural coding of distinct statistical tion inference with the minimum statistic. Neuroimage 25:653– 660.
properties of reward information in humans. Cereb Cortex 16:561–573. O’Doherty J, Winston J, Critchley H, Perrett D, Burt DM, Dolan RJ (2003a)
Dreher JC, Schmidt PJ, Kohn P, Furman D, Rubinow D, Berman KF (2007) Beauty in a smile: the role of medial orbitofrontal cortex in facial attrac-
Menstrual cycle phase modulates reward-related neural function in tiveness. Neuropsychologia 41:147–155.
women. Proc Natl Acad Sci U S A 104:2465–2470. O’Doherty JP, Dayan P, Friston K, Critchley H, Dolan RJ (2003b) Temporal
Dreher JC, Koechlin E, Tierney M, Grafman J (2008a) Damage to the difference models and reward-related learning in the human brain. Neu-
fronto-polar cortex is associated with impaired multitasking. PLoS ron 38:329 –337.
One 3:e3227. Ongür D, Price JL (2000) The organization of networks within the orbital
Dreher JC, Meyer-Lindenberg A, Kohn P, Berman KF (2008b) Age-related and medial prefrontal cortex of rats, monkeys and humans. Cereb Cortex
changes in midbrain dopaminergic regulation of the human reward sys- 10:206 –219.
tem. Proc Natl Acad Sci U S A 105:15106 –15111. Padoa-Schioppa C, Assad JA (2008) The representation of economic value
Dreher JC, Kohn P, Kolachana B, Weinberger DR, Berner KF (2009) Varia- in the orbitofrontal cortex is invariant for changes of menu. Nat Neurosci
tion in dopamine genes influences responsivity of the human reward 11:95–102.
system. Proc Natl Acad Sci U S A 106:617– 622. Peters J, Büchel C (2010) Neural representations of subjective reward value.
FitzGerald TH, Seymour B, Dolan RJ (2009) The role of human orbitofron- Behav Brain Res 213:135–141.
tal cortex in value comparison for incommensurable objects. J Neurosci Plassmann H, O’Doherty J, Rangel A (2007) Orbitofrontal cortex encodes
29:8388 – 8395. willingness to pay in everyday economic transactions. J Neurosci
Gomes CM, Boesch C (2009) Wild chimpanzees exchange meat for sex on a 27:9984 –9988.
long-term basis. PLoS One 4:e5116. Ponseti J, Bosinski HA, Wolff S, Peller M, Jansen O, Mehdorn HM, Büchel C,
Gottfried JA, Small DM, Zald DH (2006) The chemical senses. The orbito- Siebner HR (2006) A functional endophenotype for sexual orientation
frontal cortex, pp 125–171. Oxford: Oxford UP. in humans. Neuroimage 33:825– 833.
Hamann S, Herman RA, Nolan CL, Wallen K (2004) Men and women differ Prevost C, Pessiglione M, Météreau E, Clery-Melin ML, Dreher JC (2010)
in amygdala response to visual sexual stimuli. Nat Neurosci 7:411– 416. Distinct valuation subsystems in the human brain for effort and delay. J
Hammers A, Allom R, Koepp MJ, Free SL, Myers R, Lemieux L, Mitchell TN, Neurosci, in press.
Brooks DJ, Duncan JS (2003) Three-dimensional maximum probability Rangel A, Camerer C, Montague PR (2008) A framework for studying the neu-
atlas of the human brain, with particular reference to the temporal lobe. robiology of value-based decision making. Nat Rev Neurosci 9:545–556.
Hum Brain Mapp 19:224 –247. Redouté J, Stoléru S, Grégoire MC, Costes N, Cinotti L, Lavenne F, Le Bars D,
Hare TA, O’Doherty J, Camerer CF, Schultz W, Rangel A (2008) Dissociat- Forest MG, Pujol JF (2000) Brain processing of visual sexual stimuli in
ing the role of the orbitofrontal cortex and the striatum in the computa- human males. Hum Brain Mapp 11:162–177.
tion of goal values and prediction errors. J Neurosci 28:5623–5630. Reuter J, Raedler T, Rose M, Hand I, Gläscher J, Büchel C (2005) Patholog-
Hayden BY, Parikh PC, Deaner RO, Platt ML (2007) Economic principles ical gambling is linked to reduced activation of the mesolimbic reward
motivating social attention in humans. Proc Biol Sci 274:1751–1756. system. Nat Neurosci 8:147–148.
Henson R (2006) Forward inference using functional neuroimaging: disso- Reynolds CF 3rd, Frank E, Thase ME, Houck PR, Jennings JR, Howell JR,
ciations versus associations. Trends Cogn Sci 10:64 – 69. Lilienfeld SO, Kupfer DJ (1988) Assessment of sexual function in de-
Hoon E, Chambless D (1998) Sexual arousability inventory and sexual pressed, impotent, and healthy men: factor analysis of a brief sexual func-
arousability inventory— expanded. In: Handbook of Sexuality- tion questionnaire for men. Psychiatry Res 24:231–250.
Related Measures, pp 71–74. Washington, DC: American Psychiatric Rhodes G (2006) The evolutionary psychology of facial beauty. Annu Rev
Association. Psychol 57:199 –226.
13104 • J. Neurosci., September 29, 2010 • 30(39):13095–13104 Sescousse et al. • Reward Value Coding in the Human OFC

Schultz W (2006) Behavioral theories and the neurophysiology of reward. Tremblay L, Schultz W (1999) Relative reward preference in primate or-
Annu Rev Psychol 57:87–115. bitofrontal cortex. Nature 398:704 –708.
Small DM, Zatorre RJ, Dagher A, Evans AC, Jones-Gotman M (2001) Völlm B, Richardson P, McKie S, Elliott R, Dolan M, Deakin B (2007) Neuronal
Changes in brain activity related to eating chocolate: from pleasure to correlates of reward and loss in Cluster B personality disorders: a functional
aversion. Brain 124:1720 –1733. magnetic resonance imaging study. Psychiatry Res 156:151–167.
Smith DV, Hayden BY, Truong TK, Song AW, Platt ML, Huettel SA (2010) Wise SP (2008) Forward frontal fields: phylogeny and fundamental func-
Distinct value signals in anterior and posterior ventromedial prefrontal tion. Trends Neurosci 31:599 – 608.
cortex. J Neurosci 30:2490 –2495. Yacubian J, Gläscher J, Schroeder K, Sommer T, Braus DF, Büchel C (2006)
Sugrue LP, Corrado GS, Newsome WT (2005) Choosing the greater of two Dissociable systems for gain- and loss-related value predictions and errors
goods: neural currencies for valuation and decision making. Nat Rev of prediction in the human brain. J Neurosci 26:9530 –9537.
Neurosci 6:363–375. Zink CF, Pagnoni G, Martin-Skurski ME, Chappelow JC, Berns GS (2004)
Thornhill R, Gangestad SW (1999) Facial attractiveness. Trends Cogn Sci Human striatal responses to monetary reward depend on saliency. Neu-
3:452– 460. ron 42:509 –517.
!
!
ABOUT)THIS)COMPENDIUM)
!
!
The$original$purpose$of$this$compendium$has$been$for$the$use$in$my$own$lectures$in$consumer$
neuroscience$and$neuromarketing$at$the$Copenhagen$Business$School.$However,$I$also$
recognise$that$this$volume$can$also$be$a$potentially$valuable$resource$for$both$newcomers$as$
well$as$experienced$people$within$this$discipline.$Neuromarketing$is$today$very$much$a$
conglomerate$of$divergent$solutions;$hyped$up$talks;$and$a$mixture$of$true$science$and$pop$
science$gone$terribly$wrong.$This$collection$of$papers$represent$my$own$take$on$what$the$
basics$should$entail$
!
This$book$is$also$intended$as$a$supplement$to$my$book$“Introduction$to$Neuromarketing$&$
Consumer$Neuroscience”,$which$you$can$read$more$about$here:$http://neuronsinc.com/
publications/introductionPtoPneuromarketingPconsumerPneuroscience/$(also$see$next$page).$
!
The$selection$of$texts$are$not$intended$to$be$an$exhaustive$listing$of$all$relevant$articles.$I$have$
worked$from$two$basic$premises:$1)$that$the$article$is$available$freely$on$the$web;$and$2)$that$
the$article$represents$some$of$the$leading$thoughts$(and$scholars)$in$this$field.$$
!
If$you$have$suggestions$or$comments,$please$send$me$an$email$at$tzramsoy@gmail.com$$
!
!
DISCLAIMER)
!
All$materials$in$this$compendium$–$texts$and$images$–$have$either$been$collected$from$freely$
available$resources,$or$written$by$myself.$I$claim$no$ownership$or$rights$over$these$materials.$
All$materials$in$this$compendium$–$except$my$own$freely$distributed$materials$–$can$be$
collected$and$compiled$by$any$individual.$Please$note$that$there$may$be$restrictions$on$
materials$in$this$compendium$for$sharing,$distributing$or$selling.$$
!
If$you$find$that$this$compendium$contains$materials$that$are$not$permitted$for$sharing,$please$
send$me$an$email$at$tzramsoy@gmail.com$and$I$will$adjust$accordingly.$
!
!
!
Happy$reading!$
!
!
All$the$best,$
$
!
!
!
!
!
!
!
v"2.0"–"August"2014

WHO)MADE)THIS?)
!
$
Thomas"Zoëga"Ramsøy,"b."1973"in"Oslo,"Norway"
!
Thomas$is$considered$one$of$the$leading$experts$on$
neuromarketing$and$consumer$neuroscience,$and$he$is$an$
innovator$by$heart.$With$a$background$in$economics$and$
neuropsychology,$he$holds$a$PhD$in$neurobiology$from$the$
University$of$Copenhagen.$$
!
Thomas$has$published$extensively$on$the$application$of$
neuroimaging$and$neurophysiology$to$consumer$behaviour$and$
decision$making.$He$is$the$Director$of$the$Center$for$Decision$
Neuroscience,$where$his$research$team$uses$an$eclectic$mix$of$
technologies$and$the$sciences$of$economics,$$psychology$and$
neuroscience.$Beyond$this,$Thomas$is$the$CEO$of$Neurons$Inc,$
where$he$consults$companies$around$the$globe$on$the$use$of$
science$and$technology$in$business.$
!
!
!
More$information$about$Thomas$can$be$found$on$the$following$resources:$
!
Professional)pages)
DNRG$CBS$–$http://cbs.dk/DNRG$$
DNRG$HH$–$http://drcmr.dk/research/DecisionNeuroscience$$
Neurons$Inc$–$http://neuronsinc.com$$
!
Social)media)
Twitter$–$https://twitter.com/NeuronsInc$$
Neurons$Inc$–$http://NeuronsInc.com$$
BrainEthics$–$http://brainethics.org$
!
Societies)
Neuromarketing$Science$&$Business$Association$–$http://www.neuromarketingPassociation.com$$
Society$for$Mind$Brain$Sciences$–$http://mbscience.org/$
!
Publications)
ResearchGate$–$https://www.researchgate.net/profile/Thomas_Z_Ramsoy/$$
!
!
!
!

You might also like