You are on page 1of 56

!

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

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

EMOTIONS & FEELINGS

One$of$the$important$concepts$in$marketing$and$consumer$insights$is$emotions,$or$feelings...?$
What$is$really$the$basic$difference$between$emotions$and$feelings?$Are$they$just$two$concepts$
for$the$same$phenomenon,$or$do$they$reveal$something$more$fundamental$about$how$our$
minds$are$organized?

To$make$matters$worse,$emotions$and$feelings$are$used$interchangably$in$academia$and$
industry.$Here,$I$will$tease$apart$what$the$basic$differences$are,$and$serve$some$central$articles$
pertaining$to$this$distinction.

Put$simple,$emotions$can$be$said$to$be$neural$and$bodily$responses$to$inner$or$outer$events.$
your$startle$response$to$a$sudden$noise$or$sweating$in$your$palms$when$nervous$are$good$
examples$of$emotional$responses.$Feelings,$on$the$other$hand,$should$be$reserved$to$the$
process$of$having$a$conscious$experience$of$being$in$a$particular$emotional$state.$This$
distinction$suggests$that$we$can$have$many$emotional$responses$without$an$accompanying$
feeling,$but$we$cannot$have$feelings$without$an$accompanying$emotion.
ARTICLE
Received 12 Nov 2013 | Accepted 1 Jul 2014 | Published 29 Jul 2014 DOI: 10.1038/ncomms5567 OPEN

Audience preferences are predicted by


temporal reliability of neural processing
Jacek P. Dmochowski1,w, Matthew A. Bezdek2, Brian P. Abelson3, John S. Johnson3,
Eric H. Schumacher2 & Lucas C. Parra1

Naturalistic stimuli evoke highly reliable brain activity across viewers. Here we record
neural activity from a group of naive individuals while viewing popular, previously-broadcast
television content for which the broad audience response is characterized by social media
activity and audience ratings. We find that the level of inter-subject correlation in the evoked
encephalographic responses predicts the expressions of interest and preference among
thousands. Surprisingly, ratings of the larger audience are predicted with greater accuracy
than those of the individuals from whom the neural data is obtained. An additional functional
magnetic resonance imaging study employing a separate sample of subjects shows that the
level of neural reliability evoked by these stimuli covaries with the amount of blood-oxyge-
nation-level-dependent (BOLD) activation in higher-order visual and auditory regions. Our
findings suggest that stimuli which we judge favourably may be those to which our brains
respond in a stereotypical manner shared by our peers.

1 Department of Biomedical Engineering, City College of New York, 160 Convent Avenue, New York, New York 10027, USA. 2 School of Psychology, Georgia

Institute of Technology, 654 Cherry Street, Atlanta, Georgia 30332, USA. 3 Harmony Institute, 54 West 21st Street, New York, New York 10010, USA.
w Present address: Department of Psychology, Stanford University, 450 Serra Mall, Stanford, California 94305, USA. Correspondence and requests for
materials should be addressed to J.P.D. (email: jdmochow@stanford.edu).

NATURE COMMUNICATIONS | 5:4567 | DOI: 10.1038/ncomms5567 | www.nature.com/naturecommunications 1


& 2014 Macmillan Publishers Limited. All rights reserved.
ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms5567

P
redicting the behaviour of large groups is inherent to such considered the premiere broadcast of a popular television series
diverse processes as forecasting election results, anticipating (‘The Walking Dead’, AMC, 2010) in conjunction with two
the reception to upcoming films, and foreseeing the effects metrics which capture the audience’s response to the original
of changes to laws or policies. Meanwhile, the prediction of broadcast in a time-resolved manner.
individual behaviour is a pillar of neuroscience, with a recent An online service which collects Twitter traffic information was
focus on the study of naturally occurring behaviours. Previous employed to obtain a comprehensive listing of time-stamped,
investigations have identified the neural correlates of individual stimulus-relevant tweets, which originated during the airing of
preferences1–5, subjective values6 and choices7–9 by measuring the episode. Meanwhile, 16 study participants representative of
the functional magnetic resonance imaging (fMRI)-derived the series’ target demographic were recruited to view the episode
blood-oxygenation-level-dependent (BOLD) signal in regions- while having their neural activity recorded with high-density
of-interest while subjects perform experimental tasks. Here we EEG.
ask whether the neural activity of multiple individuals may The stimulus was partitioned into its 190 constituent scenes
collectively predict the behaviour of large groups. (ranging in duration from 1.4 to 300.5 s, with a median length of
Previous works aimed at predicting population trends from 17 s), where a scene was defined as an aggregate of shots (that is,
brain activity have employed the amplitude of a neural signal, uninterrupted sequences of frames) comprising a distinct
typically the BOLD, as a readout of future behaviour4,9,10. Such narrative event. For each scene, we computed the frequency of
an approach implicitly assumes that the strength of neural elicited tweets. To account for the non-negativity and heavy-
response in a fixed region correlates with behavioural measures. tailed distribution of Twitter activity18, we logarithmically
More recently, however, a growing link is emerging between the transformed the tweet rate to yield the time series shown in
reliability of neural processing (that is, correlation across repeated Fig. 1a, which defines our dependent measure.
presentations of the stimulus) and natural behaviours. Indeed, Meanwhile, we sought to measure the amount of neural
naturalistic audiovisual stimuli have been shown to elicit highly reliability evoked by each scene in our sample of participants.
reliable neural activity across multiple viewers11, with the level of Rather than computing reliability in an electrode-to-electrode
such inter-subject correlation (ISC) linked to successful memory fashion, we first performed a dimensionality reduction technique
encoding12 and effective communication between individuals13. which projects the neural responses from all subjects onto a space
ISC is increased during scenes marked by high arousal and which maximizes the ISCs across our sample (see Methods for
negative emotional valence11,14, and is strongest for familiar and details of computation). When measured in this optimized space,
naturalistic events15. In addition to these fMRI studies, recent the bulk of the reliability is captured in just a few dimensions
work found that engaging narrative stimuli yield high levels of (that is, 3). The resulting scene-by-scene neural reliability was
ISC in the evoked encephalographic responses of a small sample then regressed onto our dependent measure, yielding the
of viewers16,17. predicted log tweet frequency (see equation (4) in Methods)
Given the evidence linking ISC—inherently a group measure— shown in Fig. 1b.
to brain states characterized by heightened affect, attention and The neural reliability experienced by the sample throughout
memory retention, we suspected that the agreement in neural each scene explains 16% of the variance in audience log tweet
responses may serve as a suitable predictor for subsequent frequency (Fig. 1c; r ¼ 0.40, P ¼ 6.1 " 10 # 7, N ¼ 190, P-value
population behaviour. Specifically, we hypothesized that the level computed using the analytic distribution of the sample correla-
of neural reliability elicited by a naturalistic stimulus in a small tion coefficient19, 95% confidence interval on r: (0.26,0.51)
sample would be predictive, to some degree, of behavioural computed using the bootstrap20). It is worthwhile to note that
responses reflecting engagement or interest of a large population. while tweeting is a delayed behavioural response, the observed
Broadcasts of popular television shows or advertisements serve neural reliability is driven by immediate short-term responses
as a convenient framework for testing our hypothesis: in the (reliability was calculated for activity 0.5 Hz or higher; see
social media age, the responses of large audiences are captured in Methods and Fig. 4).
online networks such as Twitter, Facebook and YouTube. We On the basis of previous findings suggesting an association
leverage this to explore the link between neural and behavioural between ISC and narrative quality, novelty and coherence11,16,21,
responses. Namely, we recruited a sample of 12–16 naive subjects we suspected that neural reliability may also predict viewership
and presented them with stimuli which had been previously aired size. To that end, we obtained minute-by-minute Nielsen ratings
and for which we compiled aggregated measures of the stemming from the original broadcast (including advertisements),
population response. We imaged brain activity during this resulting in a time series conveying audience size and defining
exposure, employing electroencephalography (EEG) which cap- our dependent measure (Fig. 2a).
tures broad patterns of activity on the time scale of neuronal The decision to continue viewing may depend, in part, on
processing, allowing us to measure reliability in short temporal recent viewing history. As a result, we opted not to correlate
segments. To further characterize the observed reliability, we reliability instantaneously with viewership. Instead, we formed
subsequently performed an EEG-informed fMRI activation study our neural reliability measure using ISCs computed over the prior
to identify brain areas which are systematically more (or less) 3 min of viewing. We then regressed the resulting time series onto
active during stimuli which elicit greater ISCs in the EEG. Most the minute-by-minute viewership, yielding the predicted time
importantly, we found a statistically significant link between the series shown in Fig. 2b. Neural reliability explains 36% of
neural reliability in the sample and preferences of large audiences viewership (Fig. 2c; r ¼ 0.60, P ¼ 7.1 " 10 # 8, N ¼ 86; 95%
within and across contemporary audiovisual stimuli. Our findings confidence interval on r: (0.45,0.71)).
suggest that behavioural responses of large groups to natural There are two evident sources of variability in the Nielsen
stimuli may be robustly predicted from the reliability in ratings: a sudden drop in ratings during advertisements, and a
corresponding neural responses of a small sample of individuals. gradual decay due to declining audience retention. To determine if
the measured correlation is driven by the obvious variation from
intervening advertisements, we repeated the calculation but
Results omitting the advertising segments. Reliability explains 34% of of
We sought a stimulus eliciting time-varying and readily available the variance during programming alone (r ¼ 0.58, P ¼ 2.6 " 10 # 5,
viewer responses across a large population. To that end, we N ¼ 62, 95% confidence interval on r: (0.33,0.75)).

2 NATURE COMMUNICATIONS | 5:4567 | DOI: 10.1038/ncomms5567 | www.nature.com/naturecommunications


& 2014 Macmillan Publishers Limited. All rights reserved.
NATURE COMMUNICATIONS | DOI: 10.1038/ncomms5567 ARTICLE

Tweet rate (log count per second)


per second)
1

Tweet rate
(log count
0 1
−1
0.5
50 100 150
0 r =0.4 P =6.1e−07

Predicted tweet
−0.5

rate (log count


per second)
−1
0

−0.5 0 0.5 1
50 100 150
Scene number Predicted tweet rate (log count per second)

Figure 1 | Neural reliability predicts scene-by-scene tweet frequency. (a) Log frequency of tweets related to each scene (N ¼ 190) of a popular
television broadcast. (b) Log tweet frequency as predicted from the scene-by-scene neural reliability measured across 16 participants during subsequent
viewing of the episode in the laboratory (see equation (4) in Methods). (c) Neural reliability explains 16% of the variance in the log tweet rate (r ¼ 0.4,
P ¼ 6.1 " 10 # 7, N ¼ 190; 95% confidence interval on r: (0.26,0.51)). Dashed line represents regression from predicted to actual log tweet rate.

×103 28
Viewership

25
26

Viewership (×103)
20
24
20 40 60 80

×103 22
viewership

r =0.6 P =7.0e−08
Predicted

25
20
prog
ads
20 18
20 22 24 26 28
20 40 60 80
Time (min) Predicted viewership (×103)

Figure 2 | Neural reliability predicts viewership size. (a) Viewership size during broadcast of television show as measured by Nielsen ratings.
Programming (blue) is interrupted by advertising (red). (b) Viewership as predicted from the neural reliability exhibited by 16 participants viewing the same
programming. (c) Neural reliability explains 36% of the variance in viewership size (r ¼ 0.60) when including both periods of programming (blue)
and advertising (red), while accounting for 34% of the variance during programming alone (r ¼ 0.58). Dashed line denotes regression from predicted to
actual viewership (including ads); prog ads, programming advertisements.

The gradual drop in viewership size observed here is typical of sample neural reliability to predict across-stimuli preferences. We
the free-viewing environment of the general audience (that is, thus obtained the results of a popular online survey occurring
being able to change the channel at any time). This contrasts with annually, in which a large number of participants view and
the laboratory environment in which participants are asked to subsequently rate a series of advertisements (SuperBowl com-
view the entire episode. To compensate for this mismatch in mercials). We randomly sampled 10 ads from the 2012 version of
viewing conditions, we removed the linear trend in the viewership this survey and recruited a new set of N ¼ 12 volunteers to view
size and found even stronger correlations (complete broadcast: these ads while recording their EEG. Subjects also provided their
r ¼ 0.68, P ¼ 4.9 " 10 # 11, 95% confidence interval on r: own preference rating following the recording. For each
(0.56,0.77); programming only: r ¼ 0.66, P ¼ 3.1 " 10 # 7, 95% advertisement, we computed the neural reliability from the ISCs
confidence interval on r: (0.45,0.82)). In other words, neural in the neural responses of the sample (see Methods for details).
reliability explains 43% of the variance in viewership size during We found a strong and statistically significant correlation
programming after accounting for the drop in retention. between neural reliability and the population ratings (Fig. 3a,
We also considered the effect of the temporal window size (that circled markers; r ¼ 0.90, P ¼ 9 " 10 # 5, N ¼ 10, 95% confidence
is, 3 min) used to define reliability on the prediction accuracy. As interval on r: (0.76,0.97)). Given this surprisingly strong
shown in Supplementary Fig. 1, the strength of the relationship correlation, we sought to validate the results on a new stimulus
between neural reliability and viewership exhibits a broad peak at set, repeating the experiment with the 2013 series of ads while
a window size of 3–4 min when predicting ratings during both employing the same 12 participants. The neural reliability
programming and advertisements, while increasing monotoni- correlated significantly with the population ratings (Fig. 3a,
cally from 1 to 6 min when excluding ads (see also Supplementary triangle markers; r ¼ 0.73, P ¼ 0.014, N ¼ 10; 95% confidence
Note 1). In addition, the correlation of viewership size with neural interval on r: ( # 0.06,0.95); the drop in correlation from 2012 is
reliability is insensitive to which of the two age categories driven by a single advertisement, see Supplementary Note 3). By
provided by Nielsen is being predicted (Supplementary Table 1 combining all 20 advertisements viewed by each study partici-
and Supplementary Note 2). pant, neural reliability explains 66% of the variance in population
Both the tweet frequencies and Nielsen ratings considered ratings (Fig. 3a; r ¼ 0.81, P ¼ 3 " 10 # 6, N ¼ 20, 95% confidence
above quantify audience response during a single programme. interval on r: (0.50,0.92)). Intriguingly, neural reliability explains
Audience preferences are often expressed not within but across just 26% of the sample’s own preferences (Fig. 3b; r ¼ 0.51,
competing programming. We wanted to test the ability of the P ¼ 0.02, N ¼ 20, 95% confidence interval on r: ( # 0.14,0.78)),

NATURE COMMUNICATIONS | 5:4567 | DOI: 10.1038/ncomms5567 | www.nature.com/naturecommunications 3


& 2014 Macmillan Publishers Limited. All rights reserved.
ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms5567

9 9

8 r =0.81 P =1.2e−05 8 r =0.51 P =0.021

7 7

Population rating

Sample rating
6 6

5 5

4 4

3 2012 3
2012
2013 2013
2 2
0 0.05 0.1 0.15 0.2 0 0.05 0.1 0.15 0.2
Neural reliability Neural reliability

Figure 3 | Neural reliability in small sample is predictive of preference ratings in large audience. (a) Vertical axis: subjective ratings for 2 " 10
SuperBowl advertisements collected from a large online audience (Facebook-USA Today Ad Meter). Horizontal axis: neural reliability experienced across 12
subjects during each 30 s advertisement. Dashed line indicates the linear prediction of population ratings from neural reliability: 66% of variance in
population ratings is explained. (b) Same as a, but with the vertical axis showing the mean ratings of the individuals in the sample (error bars indicate
s.e.m., N ¼ 12). Only 26% of variance in the sample ratings is explained (P ¼ 0.047, N ¼ 20, Fisher r-to-z transformation).

which is significantly lower than the predictability of the


population preferences (P ¼ 0.047, N ¼ 20, Fisher r-to-z LATERAL
transformation).
Could the reduced predictability of the sample ratings result
from the variability due to the smaller sample size? To examine
this, we generated N ¼ 106 random samples of 12 ratings for each
STG
of the 20 ads (assuming normal distributed ratings with the IFG STG
population rating as the mean and variance as observed in the
actual sample). The resulting correlation of these simulated LEFT RIGHT
sample ratings with the neural reliability was significantly higher SPC/
SPC/
than what was observed for the actual sample ratings: a mean of PCun
PCun
mPFC/ mPFC/
r ¼ 0.75 with a 95% confidence interval of (0.66,0.83), leading to a ACC ACC
probability P ¼ 4 " 10 # 5 of drawing the actual value of r ¼ 0.51
from this distribution. We also explored the possibility of a
systematic difference in the ratings of the sample and those of the
population. However, ratings were largely consistent, differing
MEDIAL
significantly for only two of the 20 ads (P40.05 false-discovery
rate, N ¼ 12, Student t-test). A positive bias observed in the –6.0 T 6.0

average rating ( þ 0.65, P ¼ 0.006, N ¼ 20, Student t-test) should


not affect correlation coefficients which are insensitive to a mean Figure 4 | Covariation of BOLD activity with EEG-derived neural
offset. Indeed, our sample ratings explain 59% of the variance in reliability in different brain regions. Significant clusters of activation with a
the population ratings (r ¼ 0.77, Supplementary Fig. 2), further corrected family-wise error rate of 0.05 are mapped on inflated cortices
lending credence to the notion that the reliability in neural using the CARET software54. IFG, inferior frontal gyrus; STG, superior
responses is indeed more strongly linked to preferences of the temporal gyrus; SPC/PCun, superior parietal cortex/precuneus; mPFC/
population. Finally, it is worth noting that stimuli were judged ACC, medial prefrontal cortex/anterior cingulate cortex. Note that the
with high preference heterogeneity: the same advertising was BOLD and EEG data were collected from separate groups of subjects.
judged very differently by different subjects (the range of ratings
for each ad was 6.25±0.97, that is, almost the full range of 1–10
was used by the 12 subjects to rate the ads).
To probe the spatial dimension of the observed neural temporal poles (Fig. 4). Moreover, we observed significantly
reliability, we performed a follow-up fMRI experiment using a larger BOLD activation patterns for high-reliability advertise-
separate sample (N ¼ 14) of individuals, recording the BOLD ments in an area of parietal cortex including the superior parietal
signal evoked by all 20 of the SuperBowl ads. The subsequent lobule and precuneus. Meanwhile, a significant negative covaria-
BOLD activation time series were regressed onto the neural tion between neural reliability and BOLD activation was found in
reliability scores (see horizontal axis of Fig. 3) in a block-design a region of medial prefrontal cortex (mPFC) that includes
fashion. We sought to identify brain regions which exhibit anterior cingulate cortex (ACC), as well as the left inferior frontal
systematically higher levels of activation for stimuli marked by gyrus (IFG). To test if reliability of BOLD activity is also
high levels of neural reliability. predictive of preference ratings, we computed the ISC of the
We found significant covariation of BOLD activity with EEG- spatiotemporal patterns of BOLD activity in the identified regions
derived neural reliability in both left and right lateral temporal for each advertisement. The measured BOLD-ISC did not
cortices: these large clusters stretched from sensory association significantly correlate with the population nor the sample ratings
areas in occipital cortex, along the superior temporal gyrus, to the (r ¼ 0.34 and r ¼ 0.23, respectively, P40.14).

4 NATURE COMMUNICATIONS | 5:4567 | DOI: 10.1038/ncomms5567 | www.nature.com/naturecommunications


& 2014 Macmillan Publishers Limited. All rights reserved.
NATURE COMMUNICATIONS | DOI: 10.1038/ncomms5567 ARTICLE

Discussion signals required subdural electrodes and focused on slow


Here, we showed that measures of behavioural responses modulations (in the order of 10 s) of oscillatory activity, in
aggregated over large audiences correlate significantly with the particular, the gamma band33, which is known to correlate with
neural reliability evoked by the corresponding naturalistic stimuli the BOLD response34. In contrast, here we used fast evoked
in a small group of individuals. In particular, neural reliability is responses measured on the scalp, which generally do not coincide
highly predictive of across-stimuli preferences, and predicts with BOLD or gamma activity35.
preferences of the large audience more accurately than those of We observed that reliability of neural experience is related to
the individuals from whom the neural activity was recorded. subsequent preferences. However, note that in the Twitter and
Our finding differs subtly but importantly from those in which Nielsen studies, our behavioural measures index general response
population responses are better predicted from a sample’s neural independent of valence; strictly speaking, the Tweet rates and
activity than from its self-reports4,9. Such findings may, in part, Nielsen ratings are not reflective of ‘liking’ the stimulus, but
be explained by the fact that the behaviours of the population and rather being compelled to respond to or continue viewing it,
sample are being evaluated with somewhat different measures respectively. Although it may be argued that tweeting about or
(for example, expressing a preference for a stimulus versus tuning into a programme are behaviours consistent with ‘liking’
actually consuming it). In the SuperBowl experiment described it, they are certainly not sufficient conditions for doing so. The
here, the behaviours performed by both the sample and present analysis has implicitly grouped both positive and negative
population are identical, and their responses are well correlated. valences into the dependent measure being predicted: for
Note, however, that it is the population ratings that link most example, Twitter commentary to the episode expressed both
strongly to the reliability of neural responses, even though the positive as well as negative sentiment. It is thus possible that
sample is the source of the measured reliability. We have not reliability correlates more generally with, for example, interest,
found a precedent for the present observation that neural signals rather than preference itself. On an anecdotal level, we do point
explain the population response better than the response of the out that the SuperBowl ad receiving the lowest population rating
sample. (unambiguously denoting a dislike) among all 20 ads also elicited
One may conclude from the results that stimuli which evoke the lowest neural reliability (see Supplementary Table 2).
highly reliable neural responses among a small sample also do so It is interesting to contrast the present results with the
in a larger audience. However, this interpretation does not literature on the neural basis of individual differences36–38. There,
account for the finding of significantly lower predictability of the the focus is on capturing neural features which vary across
sample ratings, which cannot be fully explained from the individuals and thus explain differences in individual behaviour.
reduction in sample size. We conjecture that this finding is Here, we focus on the commonality in neural responses,
related to the preference heterogeneity of the advertising stimuli effectively ignoring individual differences, to obtain a predictor
used: the high variability observed in the sample ratings may be of group behaviour. The component analysis technique used here
attributed to differing subjective values2,22,23 or other variables to compute neural reliability explicitly looks for shared neural
such as social conformity24,25. Such idiosyncratic processes may components, and the resulting quantity links closely to
involve complex reasoning or emotional considerations that take population measures which reflect shared behaviours (that is,
relatively long to evaluate and presumably fail to yield immediate trends) that emerge after aggregation of large samples.
and reliable EEG signals. Through population aggregation, The broad fMRI activations observed in sensory and associa-
however, these idiosyncratic preferences tend to average out tion cortex suggest that modulations of high-level visual and
and one is left with what is shared by the large audience. auditory processing underlie the measured neural reliability. For
Therefore, the surprising finding of this study is that reliability of example, the activations in bilateral temporal cortex may reflect
relatively fast neural processing is a genuine predictor of the processing of complex auditory speech information (both
common preferences of a large population. linguistic and prosodic) during advertisements39. The activated
Preference heterogeneity has been studied extensively in the region also included areas of occipitotemporal cortex recruited by
context of economic risk-taking, often focusing on the neural the processing of dynamic visual stimuli40,41. Increased BOLD
underpinnings of individual differences in decision-making26. In activation was also found in the superior parietal lobules and
the marketing literature, preference heterogeneity has been precuneus, which mediates attention to auditory and visual
reported to affect perception of advertising27,28. However, we stimuli42–44. In addition, this region has also been associated with
are not aware of literature analysing the neural basis of preference self-referential processing, imagery and memory45, processes that
heterogeneity with natural stimuli or, in particular, video may be elicited during the viewing of well-crafted advertising.
advertisements. Meanwhile, activations of the ACC/mPFC have been implicated
Traditional neuroimaging work on the evaluation of preference in the evaluation of conflict and emotions46, which may have
or ‘value’ uses fMRI and points to elevated activity in specific occurred more frequently during the less likable advertisements.
subcortical regions29. In particular, activity in the ventral striatum We caution, however, that all of these observed BOLD activations
and medial prefrontal cortex (mPFC) correlates with individual were found to co-vary with the EEG reliability of a separate
subjective value22 and the purchasing behaviours of a larger group of subjects; as such, we refrain from inferring that the
population4. Such neural activity encodes information that is encephalographic signal components driving our preference-
predictive of decisions following stimulus presentation7,30 even linked measure of neural reliability originate from these fMRI-
when measured in the absence of a choice31, thus pointing to a identified regions. Although the topographies of the EEG
certain level of automatic stimulus evaluation. The present components have been found to be fairly reproducible across
findings highlight the importance of reliable short-latency various stimuli (see Fig. 5), the specific co-varying BOLD
responses, suggesting similarly automatic stimulus processing. activations may be stimulus-dependent. Disparate neuro-
However, the present study points to neural processing of more modulatory processes may manifest in similar patterns of
superficial cortical areas, which are the main contributors to the cortical generators which drive the observable EEG32.
EEG32. Note also that it is the reliability of temporal dynamics, It is possible that personal preferences yield changes in the
and not necessarily the strength of response, that is carrying the individual’s level of attention or engagement. Such ‘top-down’
predictive information here. It is also worthwhile to point out that modulation may then affect the strength47 and thus the reliability
previous efforts at analysing reliability of electrophysiological of neural responses associated with stimulus-locked neural

NATURE COMMUNICATIONS | 5:4567 | DOI: 10.1038/ncomms5567 | www.nature.com/naturecommunications 5


& 2014 Macmillan Publishers Limited. All rights reserved.
ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms5567

Methods
1 2 3
Subjects and stimuli. For the encephalography recordings, 16 healthy subjects
(nine females and seven males, ages 19–32, mean of 26 years) viewed the pilot
episode of ‘The Walking Dead’ along with intermittent commercials as aired in the
original broadcast. An additional 12 subjects (gender balanced, ages 20–29, mean of
25 years) viewed and subsequently rated (on a scale of 1–10) 10 advertisements
initially aired during the 2012 SuperBowl (one subject was common to both
experiments). To validate the results, the same subjects then viewed 10 ads from the
2013 SuperBowl. These 20 video clips were randomly selected and spanned the range
of viewer ratings from the Facebook-USA Today Ad Meter (see Supplementary
Table 2). For the fMRI recordings, a separate 14 subjects (six females, ages 18–22,
mean of 20 years) viewed the same set of 2012 and 2013 SuperBowl advertisements.
Subjects provided written informed consent in accordance with the procedures
approved by the Institutional Review Boards of the City College of New York (EEG
study) and the Georgia Institute of Technology (fMRI study).

EEG data collection. Study participants viewed the stimuli in a darkened,


electrically and acoustically shielded room. Sound was played back with PC
loudspeakers adjusted by each subject to a comfortable listening volume. Subjects
were instructed to pay attention to the stimuli and to minimize overt movement.
0.5 Before viewing, subjects were fitted with a 64-electrode cap placed on the scalp
according to the international 10/10 standard for EEG, which was recorded with a
BioSemi ActiveTwo system (BioSemi, Amsterdam, The Netherlands) at a sampling
0 frequency of 512 Hz and 24 bits per sample. To subsequently correct eye-move-
ment artifacts, we also recorded the electrooculogram (EOG) with four auxiliary
electrodes (one adjacent to and one below each eye).
–0.5
uV

Figure 5 | Scalp projections of the three most-reliable dimensions of EEG preprocessing. All data processing was performed automatically (that is,
with no manual intervention) offline in the MATLAB software (MathWorks,
neural activity. (a) As measured during viewing of ‘The Walking Dead’ Natick, MA, USA). After extracting the EEG/EOG segments corresponding to the
pilot, (b) as measured during viewing of 10 advertisements from the 2012 duration of each stimulus, the signals were high-pass filtered (1 Hz cutoff), notch
SuperBowl (c) same as b but from the 2013 SuperBowl. filtered at 60 Hz, and down sampled to 256 Hz. Eye-movement related artifacts
were corrected by linearly regressing out the four EOG channels from all EEG
channels. The regression was performed on non-overlapping 5-s blocks for The
Walking Dead data set, and on the entire data record for each SuperBowl adver-
processing. However, if individual preferences were to guide tisement (that is, a 30-s ‘window’). After the correction of eye-movement artifacts,
modulation of sensory processing, then we would have expected channels whose average power exceeded the mean channel power by four standard
neural reliability to predict the sample preferences equally well, if deviations were excluded from analysis, with this process repeated four times in an
iterative scheme. Similarly, within each kept channel, samples whose squared-
not better, than the population preferences. Alternatively, it may amplitude exceeded the mean-squared-amplitude of that channel by more than
be that individuals prefer stimuli precisely because the narratives four standard deviations were rejected. Again, this procedure was iterated four
drive the brain strongly and reliably. Such ‘bottom-up’ influence times for each channel. In addition, we rejected every sample within 100 ms of the
would evidently be well reflected in the preferences of large identified artifactual samples. As our viewing paradigm did not constrain the
subjects’ eye movements during the relatively long stimulus durations, the data
audiences; however, in the small sample, this sensory processing contained a larger proportion of artifacts than that seen in conventional, short-
may be masked by the idiosyncratic preferences or biases of trial-based experiments. The proportion of data rejected for each scene of the
particular individuals. Walking Dead episode is shown in Supplementary Fig. 3: there is no significant
The findings of the SuperBowl advertisement study suggest correlation between the time series and the log tweets per scene (r ¼ 0.04, P40.05),
nor between the time series and the prediction of log tweets per scene from
that stimuli which we judge favourably may be those to which neural reliability (r ¼ 0.002, P40.05). Meanwhile, the proportion of data rejected
our brains respond in a stereotypical manner that is shared for each SuperBowl ad is listed in Supplementary Table 2: there is no
by our peers. Viewed in another manner, if one is able to evoke significant correlation between the proportion of data removed for each ad and the
reliable neural activity from one’s audience, then that audience is, prediction of rating from the ISC (r ¼ 0.24, P40.05), nor between the proportion
as a whole, more likely to find one’s message favourable. of data removed for each ad and the population rating (r ¼ 0.21, P40.05), nor
between the proportion of data removed and the sample rating (r ¼ 0.09, P40.05).
However, the present data do not permit causal inference about In summary, the median (across subjects) percentage of samples removed was
the specific variables mediating the reliable patterns of 15.98% for the Walking Dead data set, 16.28% for the 2012 SuperBowl data
activity. One possibility is that narrative elements of the stimuli set and 19.95% for the 2013 SuperBowl data set. Rejected samples were
directly bring about neural reliability. Indeed, disrupting marked as missing data (‘NaN’), and the analysis proceeded by computing means
and covariances with the nanmean() and nancov() MATLAB functions. As
the narrative structure for stimuli is known to reduce ISC for detailed in the next section, the method employed to compute reliability
BOLD48 and evoked responses16. But it is also possible that other is rooted in covariance matrices whose sensitivity to outliers is well known;
aspects of the stimulus (for example, overall production quality) thus, we opted for a stringent outlier rejection to ensure robust covariance
correlate with population preference49, with this hidden variable estimation.
explaining the link between advertisement ratings neural
reliability. In this case, if one were to pinpoint the stimulus Neural reliability computation. To compute the neural reliability elicited by a
features that drive neural reliability, it would be possible to given stimulus, we employed the component analysis approach of Dmochowski
make the prediction of population behaviour directly from a et al.16, whose mathematical details are described below. The technique is similar to
content analysis of the stimulus (that is, without measuring canonical correlation analysis50 and its generalizations to multiple subjects51,
differing in that it uses the same projection for all data sets. It is conceptually
neural responses). similar to the ‘common canonical covariates’ method52, which is based on a
Regardless of the source of the reliability-preference link, the maximum-likelihood formulation, as opposed to the generalized eigenvalue
finding that naturally occurring audience behaviours may be problem developed in ref. 16.
forecast from scalp measurements bears potentially tremendous For a given stimulus viewed by N subjects, we have a set of N data matrices
{X1,y, XN} where Xn conveys the spatiotemporal neural response of subject n. We
relevance for fields outside the basic sciences such as education, seek to project the data of all subjects onto a common space such that the resulting
marketing and media, which stand to gain from the predictive projections exhibit maximal ISCs across the subject pool. To that end, let pi ¼ {pi1,
power of neural reliability. pi2} ¼ {(1,2),(1,3),y,(N " 1,N)} denote the set of all P ¼ N # (N " 1)/2 unique

6 NATURE COMMUNICATIONS | 5:4567 | DOI: 10.1038/ncomms5567 | www.nature.com/naturecommunications


& 2014 Macmillan Publishers Limited. All rights reserved.
NATURE COMMUNICATIONS | DOI: 10.1038/ncomms5567 ARTICLE

subject pairs. We then form the aggregated auto- and cross-covariance matrices as: the population measure:
P
X C
X
1
R11 ¼ Xpi1 XTpi1 predicted population response ¼ b0 þ b i li ; ð4Þ
PT i¼1 i¼1

P
where bi is the regression coefficient relating the aggregated ISC in dimension i to
1 X the dependent population measure, as determined by linear least-squares. Note that
R22 ¼ Xp XT
PT i¼1 i2 pi2 due to the small sample size of the SuperBowl data set, the ISCs there were
uniformly summed across components to yield the estimate of neural reliability
P which was then directly correlated with the population measure:
1 X
R12 ¼ X p XT ; ð1Þ C
X
PT i¼1 i1 pi2
neural reliability ¼ li : ð5Þ
where T is the number of time samples (columns) in Xn and T denotes matrix i¼1

transposition. For The Walking Dead data set, we learned the optimal projections on data
We seek to find a projection vector w which maximizes the ISC between encompassing all scenes, and then applied these projections back onto the data of
subject-aggregated data records: each scene to yield the time-resolved reliability in each component. In other words,
the covariances in equation (1) were formed using data from all scenes, yielding the
w T R12 w optimal w, which was then applied to equation (2) but with the covariances there
1=2
ð2Þ
ðw T R 11 w Þ ðw T R22 wÞ1=2 formed using only data for the desired segment of the stimulus. For the SuperBowl
data set, we learned the optimal projections by concatenating the neural responses
It is shown in ref. 16 that assuming wTR11w ¼ wTR22w, the solution to of all ads into a single data matrix per subject. Once again, this combined data was
equation (2) is a generalized eigenvalue problem: used to construct the covariance matrices and learn the optimal projection vectors.
lðR11 þ R22 Þw ¼ R12 w; ð3Þ We then projected these optimized filters onto the data from each advertisement to
compute the reliability exhibited by the participants’ during that ad.
where l is the generalized eigenvalue corresponding the maximal ISC,
encompassing all subject pairs, elicited by the stimulus. Note that the assumption
wTR11w ¼ wTR22w does not limit generality, as one can simply define Spatiotemporal characteristics of EEG components. Here we detail the spatial
pi0 ¼ {(1, 2),y, (N % 1, N), (N, N % 1),y, (2, 1)} and then substitute pi0 in and temporal properties of the components formed from the optimal spatial
equation (1) to ensure that R11 ¼ R22; this was done in our analysis. Moreover, filters w. Let us construct a weight matrix W whose columns represent the first C
when computing the generalized eigenvalues of equation (3), we regularize the generalized eigenvectors in equation (3). The projections of the resulting compo-
pooled auto-covariance by keeping only the first K ¼ 10 dimensions. This value of nents onto the scalp data are given by Parra et al.53:
K roughly corresponds to the ‘knee’ of the pooled auto-covariance eigenvalue A ¼ RWðWT RWÞ % 1 ; ð6Þ
spectrum in the spectral representation of R11 þ R22.
There are multiple non-orthogonal solutions to equation (3), whose associated where R ¼ R11 þ R22 is the pooled auto-covariance. The columns of A are termed
generalized eigenvalues are ranked in decreasing order of aggregated ISC: l1 4 l2 ‘forward models’ and inform us of the approximate location of the underlying
4 y 4 lD, where D is the number of electrodes. We take the first C ¼ 3 such neuronal sources (up to the inherent limits imposed by volume conduction
solutions and linearly sum their corresponding eigenvalues to yield the estimate of in EEG).

Power spectral density Inter−subject coherence spectrum


0.2 C1
C1
Time reversed
0.15
FDR P<0.01
µV2/Hz

100 0.1

0.05

0
100 101 102 100 101 102

0.2 C2
C2
Time reversed
0.15
FDR P<0.01
µV2/Hz

100 0.1

0.05

0
100 101 102 100 101 102

C3 0.2 C3
0.15 Time reversed
µV2/Hz

FDR P<0.01
100 0.1

0.05

0
100 101 102 100 101 102
Frequency (Hz) Frequency (Hz)

Figure 6 | Temporal properties of the signal components used to measure reliability. (a) Power-spectral density of each dimension of reliability
averaged across all subjects from the 2012 SuperBowl data. A characteristic peak is evident for the alpha band (around 10 Hz). (b) Coherence spectrum
computed for signals from pairs of subjects (66 unique pairs) and averaged across 10 video clips and all pairs (solid curve). Coherence is a frequency-
resolved positive measure of correlation (on a scale of 0–1). Chance level of coherence was estimated by using the same signal pairs with one of
the two reversed in time for the entire 30 s of each video clip (dashed curve). Significant difference was calculated using a bootstrap shuffle by pooling
original and time-reversed coherence values and randomly drawing (105 times) among the pairs of signals. Significant coherence controlling for false-
discovery rate at 0.01 indicated in bold.

NATURE COMMUNICATIONS | 5:4567 | DOI: 10.1038/ncomms5567 | www.nature.com/naturecommunications 7


& 2014 Macmillan Publishers Limited. All rights reserved.
ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms5567

Figure 5 depicts these forward models for the stimuli used in the study. The scalp the results. For the combined analysis, the ratings of the 2012 commercials were
projections stemming from The Walking Dead study bear a close resemblance to linearly transformed under the assumption that the quality was comparable to that of
those found in ref. 16: a symmetric first component with a dipolar distribution 2013: the 2012 ratings were scaled and offset such that the entire set of ratings from
consisting of frontocentral and occipital poles, a second component exhibiting 2012 matched the ratings from 2013 in mean and standard deviation.
bilateral poles at the temporal electrodes, and an asymmetric third component
marked by frontal and right-parietal poles. Meanwhile, the forward models of the
reliability-maximizing components from the SuperBowl study reveal a highly References
congruent first component topography, while deviating somewhat in the second and
1. McClure, S. M. et al. Neural correlates of behavioral preference for culturally
third components. For example, the frontal pole of the third component from the
familiar drinks. Neuron 44, 379–387 (2004).
2012 ads is slightly more posterior. Such disparities in scalp topographies may
2. O’Doherty, J. P., Buchanan, T. W., Seymour, B. & Dolan, R. J. Predictive neural
reflect a re-distribution of canonical sources among the three components.
coding of reward preference involves dissociable responses in human ventral
Meanwhile, Fig. 6 summarizes the temporal properties of the components used to
midbrain and ventral striatum. Neuron 49, 157 (2006).
construct the measure of neural reliability. The 1/f temporal power spectrum of these
components is typical for encephalography (Fig. 6a). A temporal coherence analysis of 3. Plassmann, H., O’Doherty, J., Shiv, B. & Rangel, A. Marketing actions can
the signals used to measure neural reliability indicates that reliability is driven by modulate neural representations of experienced pleasantness. Proc. Natl Acad.
immediate evoked responses of 2 s or less and can be as fast as as 100 ms (Fig. 6b). Sci. USA 105, 1050–1054 (2008).
Coherence across subjects—a frequency-resolved measure of correlation—is strongest 4. Berns, G. S. & Moore, S. E. A neural predictor of cultural popularity. J. Consum.
at low frequencies, but statistically significant values can be found at frequencies as Psychol. 22, 154–160 (2012).
high as 10 Hz, consistent with previous findings using intra-cortical recordings33. 5. Falk, E. B., O’Donnell, M. B. & Lieberman, M. D. Getting the word out: neural
correlates of enthusiastic message propagation. Front. Hum. Neurosci. 6, 313
(2012).
fMRI data collection. For the fMRI recordings, we used the same two sets of ads 6. Levy, D. J. & Glimcher, P. W. Comparing apples and oranges: using reward-
from the 2012 and 2013 SuperBowls. All MRI data were acquired on a Siemens specific and reward-general subjective value representation in the brain.
Magnetom Trio 3T scanner. A high-resolution T1 structural scan (3D MPRAGE, J. Neurosci. 31, 14693–14707 (2011).
TI ¼ 850 ms, flip angle ¼ 9!, 1 mm isotropic resolution) was acquired before each 7. Tusche, A., Bode, S. & Haynes, J. D. Neural responses to unattended products
subject viewed the ads. Before functional scanning, subjects were instructed to pay predict later consumer choices. J. Neurosci. 30, 8024–8031 (2010).
attention to the stimuli. Images were acquired using a whole-brain echo-planar 8. Falk, E. B., Berkman, E. T., Mann, T., Harrison, B. & Lieberman, M. D.
imaging sequence (transverse orientation, TR ¼ 2,000 ms, TE ¼ 30 ms, flip Predicting persuasion-induced behavior change from the brain. J. Neurosci.
angle ¼ 90!, field of view ¼ 204 mm) of 37 interleaved slices with 3 mm isotropic 30, 8421–8424 (2010).
resolution and a 17% gap. Data were preprocessed to correct for slice timing to the 9. Falk, E. B., Berkman, E. T. & Lieberman, M. D. From neural responses to
first slice with a Fourier interpolation, using AFNI’s 3dTshift tool. Head move-
population behavior neural focus group predicts population-level media effects.
ments were then corrected using AFNI’s 3dvolreg routine. Next, the functional data
Psychol. Sci. 23, 439–445 (2012).
were smoothed with a 6 mm full-width half-maximum Gaussian kernel to reduce
10. Ariely, D. & Berns, G. S. Neuromarketing: the hope and hype of neuroimaging
noise. Finally, data were transformed to the MNI standard space using FSL’s FLIRT
in business. Nat. Rev. Neurosci. 11, 284–292 (2010).
software using a 12-parameter trilinear affine transformation. The EEG reliability
measure (see equation (5)) for each advertisement was used as an amplitude- 11. Hasson, U., Nir, Y., Levy, I., Fuhrmann, G. & Malach, R. Intersubject
modulated block-design regressor in a general linear model of the fMRI data synchronization of cortical activity during natural vision. Science 303,
including six motion parameters as covariates, using AFNI’s 3dDeconvolve tool. 1634–1640 (2004).
Whole-brain group level analysis was performed using AFNI’s 3dttest routine, with 12. Hasson, U., Furman, O., Clark, D., Dudai, Y. & Davachi, L. Enhanced
mixed effects inference on a one-sample t-test using individual beta values. AFNI’s intersubject correlations during movie viewing correlate with successful
3dClustSim tool was used with an estimated smoothing of 9.16 mm (obtained with episodic encoding. Neuron 57, 452–462 (2008).
AFNI’s 3dFWHMx routine) to perform 10 000 Monte Carlo simulations to find a 13. Stephens, G. J., Silbert, L. J. & Hasson, U. Speaker-listener neural coupling
cluster size threshold (40 voxels) with a corrected family-wise error rate of 0.05; underlies successful communication. Proc. Natl Acad. Sci. USA 107,
P(uncorrected) ¼ 0.002. To compute the BOLD-ISC for each advertisement, we 14425–14430 (2010).
concatenated the BOLD time series of all significant voxels shown in Fig. 4, and 14. Nummenmaa, L. et al. Emotions promote social interaction by synchronizing
then computed the correlation coefficient between all subject pairs. The resulting brain activity across individuals. Proc. Natl Acad. Sci. USA 109, 9599–9604
aggregated ISCs were then correlated with the population ratings. (2012).
15. Hanson, S. J., Gagliardi, A. D. & Hanson, C. Solving the brain synchrony
eigenvalue problem: conservation of temporal dynamics (fmri) over subjects
Twitter data collection. Through the Crimson Hexagon service, we obtained a doing the same task. J. Comput. Neurosci. 27, 103–114 (2009).
listing of all episode-related tweets which originated during the initial broadcast of 16. Dmochowski, J. P., Sajda, P., Dias, J. & Parra, L. C. Correlated components of
The Walking Dead pilot, that is, all tweets from 10/31/2010 9:00–10:00 PM EST ongoing EEG point to emotionally laden attention–a possible marker of
containing a relevant hashtag, referencing a show-specific Twitter account, or
engagement? Front. Hum. Neurosci. 6, 112 (2012).
simply referencing the show’s name. The listing was filtered to retain only those
17. Lankinen, K. et al. Intersubject consistency of cortical MEG signals during
tweets which directly referenced episode content (that is, 1,947 of 19,000 total
movie viewing. NeuroImage 92, 217–224 (2014).
tweets). Each retained tweet was then manually linked to the corresponding
18. Sasahara, K., Hirata, Y., Toyoda, M., Kitsuregawa, M. & Aihara, K. Quantifying
scene(s) by inspecting the message content as well as the tweet timestamp. This
collective attention from tweet stream. PLoS ONE 8, e61823 (2013).
procedure was performed by three research assistants who were blind to the
hypothesis of the study. Approximately 61% of tweets could be unambiguously 19. Fisher, R. A. Frequency distribution of the values of the correlation coefficient
linked to one scene only; when ambiguous, the tweet was linked to multiple can- in samples from an indefinitely large population. Biometrika 10, 507–521
didate scenes (that is, in general, the mapping from tweet to scene was one-to- (1915).
many). The distribution of number of scenes referenced per tweet is shown in 20. Efron, B. & Tibshirani, R. J. An Introduction to the Bootstrap, Vol. 57 (CRC
Supplementary Fig. 4. The tweet count of each scene was computed by summing press, 1994).
the number of tweets referencing that scene. To subsequently analyse the rela- 21. Hasson, U. et al. Neurocinematics: the neuroscience of film. Projections 2, 1–26
tionship between neural reliability and Twitter reaction, we divided the tweet (2008).
counts (incremented by one to handle forthcoming log operation) by scene 22. Kable, J. W. & Glimcher, P. W. The neural correlates of subjective value during
duration to compensate for varying scene length, yielding a tweet-frequency. intertemporal choice. Nat. Neurosci. 10, 1625–1633 (2007).
Finally, we logarithmically transformed the tweet rate to arrive at the dependent 23. Peters, J. & Büchel, C. Overlapping and distinct neural systems code for
measure onto which neural reliability was regressed. subjective value during intertemporal and risky decision making. J. Neurosci.
29, 15727–15734 (2009).
24. Izuma, K. The neural basis of social influence and attitude change. Curr. Opin.
Nielsen data collection. Courtesy of AMC, we obtained the Nielsen ratings for Neurobiol. 23, 456–462 (2013).
each minute of the initial airing of pilot episode of The Walking Dead. We summed 25. Rilling, J. K. & Sanfey, A. G. The neuroscience of social decision-making. Annu.
the ratings across age categories (18–49 and 25–54) and method of viewing (live Rev. Psychol. 62, 23–48 (2011).
versus digital video recorder). The regression results when predicting the viewer- 26. Huettel, S. A., Stowe, C. J., Gordon, E. M., Warner, B. T. & Platt, M. L. Neural
ship within each age category are reported in Supplementary Table 1.
signatures of economic preferences for risk and ambiguity. Neuron 49, 765–776
(2006).
Ad Meter data collection. We obtained publicly available population-averaged 27. Horsky, D., Misra, S. & Nelson, P. Observed and unobserved preference
scores for all 2012 and 2013 SuperBowl advertisements via the Facebook-USA Today heterogeneity in brand-choice models. Market. Sci. 25, 322–335 (2006).
Ad Meter service (Supplementary Table 2). An online panel of over 7,000 participants 28. Feick, L. & Higie, R. A. The effects of preference heterogeneity and source
rated each advertisement on a scale of 1–5 for the 2012 commercials and on a scale of characteristics on ad processing and judgements about endorsers. J. Advertising
1–10 for the 2013 commercials. We analysed each set separately and then combined 21, 9–24 (1992).

8 NATURE COMMUNICATIONS | 5:4567 | DOI: 10.1038/ncomms5567 | www.nature.com/naturecommunications


& 2014 Macmillan Publishers Limited. All rights reserved.
NATURE COMMUNICATIONS | DOI: 10.1038/ncomms5567 ARTICLE

29. Lebreton, M., Jorge, S., Michel, V., Thirion, B. & Pessiglione, M. An automatic 50. Hotelling, H. Relations between two sets of variates. Biometrika 28, 321–377
valuation system in the human brain: evidence from functional neuroimaging. (1936).
Neuron 64, 431–439 (2009). 51. Kettenring, J. R. Canonical analysis of several sets of variables. Biometrika 58,
30. Ma, Y., Wang, C. & Han, S. Neural responses to perceived pain in others 433–451 (1971).
predict real-life monetary donations in different socioeconomic contexts. 52. Neuenschwander, B. E. & Flury, B. D. Common canonical covariates.
Neuroimage 57, 1273–1280 (2011). Biometrika 82, 553–560 (1995).
31. Levy, I., Lazzaro, S. C., Rutledge, R. B. & Glimcher, P. W. Choice from non- 53. Parra, L. C., Spence, C. D., Gerson, A. D. & Sajda, P. Recipes for the linear
choice: predicting consumer preferences from blood oxygenation level-dependent analysis of EEG. Neuroimage 28, 326–341 (2005).
signals obtained during passive viewing. J. Neurosci. 31, 118–125 (2011). 54. Van Essen, D. C. A population-average, landmark-and surface-based (pals)
32. Nunez, P. L. Electric Fields of the Brain: The Neurophysics of EEG (Oxford atlas of human cerebral cortex. Neuroimage 28, 635–662 (2005).
University Press, 2006).
33. Honey, C. J. et al. Slow cortical dynamics and the accumulation of information
over long timescales. Neuron 76, 423–434 (2012).
Acknowledgements
We thank the following individuals: Joanna Raczkiewicz and Clint Beharry for propos-
34. Lachaux, J. P. et al. Relationship between task-related gamma oscillations and
ing, together with B.P.A., the use of the Walking Dead episode, and, together with Kelly
bold signal: new insights from combined fmri and intracranial eeg. Hum. Brain
Creighton, Andrew Bowe and Claris Chang, developing the coding schemes for collecting
Mapp. 28, 1368–1375 (2007).
and labelling Twitter data; Jason Sherwin for establishing the collaboration between the
35. Foucher, J. R., Otzenberger, H. & Gounot, D. The bold response and the
CCNY and the HI; Victoria Romero for establishing the collaboration between CCNY
gamma oscillations respond differently than evoked potentials: an interleaved
and GATech; and Tony Norcia, Paul Sajda, Ofer Tchernichovski, Davide Reato and
eeg-fmri study. BMC Neurosci. 4, 22 (2003).
Simon Kelly for reading and providing feedback on earlier versions of the manuscript;
36. Mohr, P. N. & Nagel, I. E. Variability in brain activity as an individual
Uri Hasson and his students for fruitful discussions; Dovid Fein for help with the neural
difference measure in neuroscience? J. Neurosci. 30, 7755–7757 (2010).
recordings. Initial brainstorming meetings that led to the Walking Dead analysis
37. Hamann, S. & Canil, T. Individual differences in emotion processing. Curr.
included J.P.D., B.P.A., J.S.J., Jason Sherwin, Joanna Raczkiewicz, Clint Beharry and Paul
Opin. Neurobiol. 14, 233–238 (2004).
Sajda. We also acknowledge the support from the Defense Advanced Research Projects
38. Berkman, E. T. & Falk, E. B. Falk Beyond brain mapping: using neural measures
Agency (government grant numbers D13AP00003 and D12PC00397).
to predict real-world outcomes. Curr. Dir. Psychol. Sci. 22, 45–50 (2013).
39. Hickok, G. & Poeppel, D. The cortical organization of speech processing. Nat.
Rev. Neurosci. 8, 393–402 (2007). Author contributions
40. Ebisch, S. J. et al. Differential involvement of somatosensory and interoceptive J.P.D. and L.C.P. designed EEG experiments, analysed the data and wrote the paper.
cortices during the observation of affective touch. J. Cogn. Neurosci. 23, M.A.B., E.H.S. and L.C.P. designed the fMRI experiment and wrote the corresponding
1808–1822 (2011). portion of the paper. J.P.D. collected and processed the EEG data. MAB collected and
41. Petit, L. & Haxby, J. V. Functional anatomy of pursuit eye movements in analysed the fMRI data. B.P.A. compiled the Twitter data, proposed the use of and
humans as revealed by fMRI. J. Neurophysiol. 82, 463–471 (1999). compiled the Nielsen data. J.S.J. facilitated access to the Twitter and Nielsen data sets.
42. Nardo, D., Santangelo, V. & Macaluso, E. Stimulus-driven orienting of visuo-
spatial attention in complex dynamic environments. Neuron 69, 1015–1028
(2011). Additional information
43. Shomstein, S. & Yantis, S. Parietal cortex mediates voluntary control of spatial Supplementary Information accompanies this paper at http://www.nature.com/
and nonspatial auditory attention. J. Neurosci. 26, 435–439 (2006). naturecommunications
44. Wise, S. P., Boussaoud, D., Johnson, P. B. & Caminiti, R. Premotor and parietal Competing financial interests: The authors declare no competing financial interests.
cortex: Corticocortical connectivity and combinatorial computations 1. Annu.
Rev. Neurosci. 20, 25–42 (1997). Reprints and permission information is available online at http://npg.nature.com/
45. Cavanna, A. E. & Trimble, M. R. The precuneus: a review of its functional reprintsandpermissions/
anatomy and behavioural correlates. Brain 129, 564–583 (2006).
46. Etkin, A., Egner, T. & Kalisch, R. Emotional processing in anterior cingulate How to cite this article: Dmochowski, J. P. et al. Audience preferences are
and medial prefrontal cortex. Trends. Cogn. Sci. 15, 85–93 (2011). predicted by temporal reliability of neural processing. Nat. Commun. 5:4567
47. Hopfinger, J. B. & West, V. M. Interactions between endogenous and doi: 10.1038/ncomms5567 (2014).
exogenous attention on cortical visual processing. Neuroimage 31, 774–789
(2006). This work is licensed under a Creative Commons Attribution 4.0
48. Hasson, U., Malach, R. & Heeger, D. J. Reliability of cortical activity during International License. The images or other third party material in this
natural stimulation. Trends Cogn. Sci. 14, 40 (2010). article are included in the article’s Creative Commons license, unless indicated otherwise
49. Prag, J. & Casavant, J. An empirical study of the determinants of revenues in the credit line; if the material is not included under the Creative Commons license,
and marketing expenditures in the motion picture industry. J. Cult. Econ. 18, users will need to obtain permission from the license holder to reproduce the material.
217–235 (1994). To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/

NATURE COMMUNICATIONS | 5:4567 | DOI: 10.1038/ncomms5567 | www.nature.com/naturecommunications 9


& 2014 Macmillan Publishers Limited. All rights reserved.
ARTICLES

Anatomically distinct dopamine release during


anticipation and experience of peak emotion to music
Valorie N Salimpoor1–3, Mitchel Benovoy3,4, Kevin Larcher1, Alain Dagher1 & Robert J Zatorre1–3
Music, an abstract stimulus, can arouse feelings of euphoria and craving, similar to tangible rewards that involve the striatal
dopaminergic system. Using the neurochemical specificity of [11C]raclopride positron emission tomography scanning, combined
with psychophysiological measures of autonomic nervous system activity, we found endogenous dopamine release in the striatum
at peak emotional arousal during music listening. To examine the time course of dopamine release, we used functional magnetic
resonance imaging with the same stimuli and listeners, and found a functional dissociation: the caudate was more involved
© 2011 Nature America, Inc. All rights reserved.

during the anticipation and the nucleus accumbens was more involved during the experience of peak emotional responses to
music. These results indicate that intense pleasure in response to music can lead to dopamine release in the striatal system.
Notably, the anticipation of an abstract reward can result in dopamine release in an anatomical pathway distinct from that
associated with the peak pleasure itself. Our results help to explain why music is of such high value across all human societies.

Humans experience intense pleasure to certain stimuli, such as food, peak emotional responses to music5,12–14. Chills involve a clear and
psychoactive drugs and money; these rewards are largely mediated discrete pattern of autonomic nervous system (ANS) arousal15, which
by dopaminergic activity in the mesolimbic system, which has been allows for objective verification through psychophysiological meas-
implicated in reinforcement and motivation (see ref. 1 for a review). urements. Thus, the chills response can be used to objectively index
These rewarding stimuli are either biological reinforcers that are pleasure, a subjective phenomenon that would otherwise be difficult
necessary for survival, synthetic chemicals that directly promote to operationalize, and allows us to pinpoint the precise time of maxi-
dopaminergic neurotransmission, or tangible items that are secondary mal pleasure.
rewards. However, humans have the ability to obtain pleasure from Previous studies have typically used experimenter-selected musi-
more abstract stimuli, such as music and art, which are not directly cal stimuli6–8. However, musical preferences are highly individual-
essential for survival and cannot be considered to be secondary ized; thus, to ensure maximal emotional responses, participants were
or conditioned reinforcers. These stimuli have persisted through asked to select their own highly pleasurable music. After extensive
cultures and generations and are pre-eminent in most people’s lives. screening (Online Methods), we recruited a group of people who
Notably, the experience of pleasure to these abstract stimuli is highly consistently experienced objectively verifiable chills during their peak
specific to cultural and personal preferences, which can vary tremen- emotional responses so that we could quantify both the occurrence
dously across individuals. and the timing of the most intense pleasurable responses. We also
Most people agree that music is an especially potent pleasurable collected psychophysiological measurements (heart rate, respiration
stimulus2 that is frequently used to affect emotional states. It has rate, electrodermal skin conductance, blood volume pulse amplitude
been empirically demonstrated that music can effectively elicit highly and peripheral temperature) during the PET scans to verify ANS dif-
pleasurable emotional responses3,4 and previous neuroimaging stud- ferences between conditions. To account for psychoacoustical differ-
ies have implicated emotion and reward circuits of the brain during ences across self-selected stimuli, we matched musical excerpts using
pleasurable music listening5–8, particularly the ventral striatum5–7, a previously established procedure5, such that participants listened to
suggesting the possible involvement of dopaminergic mechanisms9. one another’s choices, which served as either pleasurable or neutral
However, the role of dopamine has never been directly tested. We stimuli. We predicted that if the rewarding aspects of music listen-
used ligand-based positron emission tomography (PET) scanning to ing are mediated by dopamine, substantial [11C]raclopride binding
estimate dopamine release specifically in the striatum on the basis of potential differences would be found between neutral and pleasurable
the competition between endogenous dopamine and [11C]raclopride conditions in mesolimbic regions.
for binding to dopamine D2 receptors10. Pleasure is a subjective pheno- The second aim of our study was to explore the temporal dynamics
menon that is difficult to assess objectively. However, physiological of any dopaminergic activity, as distinct anatomical circuits are thought
changes occur during moments of extreme pleasure, which can be to underlie specific phases of reward responses16,17. That is, if there is
used to index pleasurable states in response to music. We used the dopamine release, we wanted to examine whether it is associated with
‘chills’ or ‘musical frisson’11 response, a well-established marker of the experience of the reward or with its anticipation18. Music provides

1Montreal Neurological Institute, McGill University, Montreal, Quebec, Canada. 2International Laboratory for Brain, Music and Sound Research, Montreal, Quebec,

Canada. 3Centre for Interdisciplinary Research in Music Media and Technology, Montreal, Quebec, Canada. 4Centre for Intelligent Machines, McGill University,
Montreal, Quebec, Canada. Correspondence should be addressed to V.N.S. (valorie.salimpoor@mail.mcgill.ca) or R.J.Z. (robert.zatorre@mcgill.ca).

Received 7 October 2010; accepted 25 November 2010; published online 9 January 2011; doi:10.1038/nn.2726

NATURE NEUROSCIENCE VOLUME 14 | NUMBER 2 | FEBRUARY 2011 257


ARTICLES

Figure 1 Positive correlation between emotional arousal and intensity Skin


2
conductance
of chills during PET scanning. The mean intensity of chills reported 1

∆ µS
by each participant during the PET scanning session was significantly 0
correlated with psychophysiological measurements that were also acquired –1
–2
during the scan. These are indicative of increased sympathetic nervous
4 6 8 10
system activity, suggesting that the intensity of chills is a good marker of
Temperature
peak emotional arousal (Supplementary Table 1). The y axis represents 2
standardized z scores for each biosignal. See main text for P-values. 1

∆ °C
0
–1
–2
an innovative means of assessing this distinction because the temporal 4 6 8 10
unveiling of tonal arrangements elicits anticipatory responses that are Blood volume pulse
2
amplitude

reflectance
based on cognitive expectations and prediction cues11,19,20. These can 1
be examined to isolate the functional components that precede peak 0


–1
pleasurable responses. As PET does not afford the temporal resolu- –2
tion required to examine this distinction, we combined the temporal 4 6 8 10
specificity of functional magnetic resonance imaging (fMRI) with the Heart rate
2
neurochemical specificity of PET. We acquired fMRI scans with the 1

∆ beats
per min
0
same participants and stimuli to examine the temporal profile of blood –1
oxygenation level–(BOLD) response specifically in those regions that –2
also showed dopamine release with PET. Striatal dopamine release and 4 6 8 10
© 2011 Nature America, Inc. All rights reserved.

BOLD responses are known to be correlated, although the relationship Respiration


2
is complex9,21. We predicted that regions revealing dopamine activity

∆ breaths
per min
1
0
in the PET data would show the largest increases in hemodynamic
–1
response during peak emotional experiences. We separately analyzed –2
the BOLD data from epochs of peak pleasure and the time immediately 4 6 8 10
preceding these responses (that is, anticipation), based on participants’ Intensity of chills

real-time behavioral responses of when chills were experienced. Spatial


conjunction analyses were used to confine the analysis to those striatal experienced from each excerpt. The mean number of chills for each
voxels showing both dopamine release from PET and increased BOLD pleasurable music excerpt was 3.7 (s.d. = 2.8). A paired-samples t test
during fMRI, which ensured that we were measuring the hemodynamic confirmed that greater pleasure was experienced during the pleasur-
signal only from regions known to release dopamine in response to able music condition over the neutral music condition (t(49) = 25.0,
the same stimuli. This multimodal procedure revealed a temporally P < 0.001). Notably, there was a significant positive correlation
mediated distinction in dopamine release to anticipatory and consum- between the reported intensity of chills and the reported degree of
matory responses in the dorsal and ventral striatum, respectively. pleasure (r = 0.71, P < 0.001), suggesting that the chills response is a
good representation of pleasure experienced amongst this group.
RESULTS Objective measures of psychophysiological signals indicative of emo-
PET data: dopamine release and emotional arousal tional arousal collected during the two PET scanning sessions showed
PET scanning took place over two sessions. Participants listened to significantly higher ANS activity during the pleasurable music condition
either pleasurable music or neutral music during the entire session in all of the variables that we measured: namely, increases in heart rate
while both subjective and objective indicators of emotional arousal (P < 0.05), respiration (P < 0.001) and electrodermal response (P < 0.05),
were collected. Subjective responses from rating scales included self- and decreases in temperature (P < 0.01) and blood volume pulse ampli-
reports of number of chills, intensity of chills and degree of pleasure tude (P < 0.001; for values, see (Supplementary Table 1). Subjective
reports of the intensity of the chills response
a b collected via rating scales during PET scanning
0 6 were significantly correlated with the degree of
Left caudate Right caudate
3.6 3.6
ANS arousal on all measures: increases in heart
∆6.4% ∆7.9%
3.0 3.0 rate (P < 0.05), respiration (P < 0.05) and elec-
∆BP

∆BP

Caudate 2.4 2.4 trodermal response (P < 0.01), and decreases


1.8 1.8
Neutral Chills Neutral Chills Figure 2 Evidence for dopamine release during
x = 10 y = 19 z=7
Left putamen Right putamen pleasurable music listening. (a) Statistical
3.6 ∆6.6% 3.6 ∆7.4% parametric maps (t statistic on sagittal,
3.0 3.0 coronal and axial slices) reveal significant
∆BP

∆BP

Putamen 2.4 2.4 (P < 0.001) [11C]raclopride binding potential (BP)


1.8 1.8 decreases bilaterally in the caudate, putamen
Neutral Chills Neutral Chills and NAcc (white arrows) during pleasurable
x = 23 y=1 z=1
Left NAcc/ventral Right NAcc compared with neutral music listening
putamen (Supplementary Table 2), indicating increased
3.6 ∆6.5% 3.6 ∆9.2% dopamine release during pleasurable music.
NAcc 3.0 3.0
(b) Changes in binding potential (BP) values
∆BP

∆BP

2.4 2.4
plotted separately for each individual; note that
1.8 1.8
the change was consistent for the majority of
x = 10 y = 12 z = –10 Neutral Chills Neutral Chills
people at each site.

258 VOLUME 14 | NUMBER 2 | FEBRUARY 2011 NATURE NEUROSCIENCE


ARTICLES

Figure 3 Combined fMRI and PET results


reveal temporal distinctions in regions showing
a b
Anticipation Experience Temporally mediated BOLD response
dopamine release. (a) [11C]raclopride PET fMRI
in dorsal and ventral striatum
3
scan results were spatially conjoined with the Hemodynamic
fMRI results by creating a mask of significant (BOLD)
fMRI
dopamine release overlayed on BOLD response
t maps during each condition. (b) Hemodynamic PET
responses and dopamine activity were maximal PET
Neurochemical y=4 y = 11
in the caudate during anticipatory phases, (dopamine
binding) Anticipation Experience
but shifted more ventrally to NAcc during 0
peak emotional responses. (c) Percent signal
change in BOLD response relative to the
c
VOI 1
mean was calculated from the peak voxel of

Percent signal
Right caudate
the caudate and NAcc clusters based on the

change
Right NAcc
[11C]raclopride PET data. Voxels showing 0

maximum dopamine release in the caudate and


NAcc (Supplementary Table 2) were identified –1
and percent BOLD signal change was calculated

–5
–4
–3
–2
–1
A1
A2
A3
A4
A5
A6
A7
A8
A9
A10
A11
A12
A13
A14
A15
C1
C2
C3
C4
Mean neutral
during the fMRI epochs associated with peak
emotional responses; values were interpolated Time series of peak dopaminergic voxels
for each second preceding this response for
each individual, up to 15 s, which was defined as the anticipatory period based on previous findings 15 (see Online Methods for additional details). We
found increased activity during anticipation (A1-A15) and decreased activity during peak emotional response (C1-C4) for the caudate, but a continuous
© 2011 Nature America, Inc. All rights reserved.

increase in activity in NAcc with a maximum during peak emotional responses. The mean signal for neutral epochs for the NAcc and caudate clusters
are also plotted for reference, as are the 5 s preceding the anticipation epochs.

in temperature (P < 0.05) and blood volume pulse amplitude (P < 0.05; were then used post hoc to identify anticipation and peak experience
(Fig. 1 and Supplementary Table 1). This finding further verified that time periods (Fig. 3a). Anticipation epochs were defined as 15 s before
the chills response is a good objective representation of peak emotional the peak experiences. BOLD responses for each of these epochs were
arousal in this group. compared with periods in which participants reported feeling neutral
Analysis of PET data (Supplementary Methods) revealed increased during the same musical excerpts. The result of this contrast for each
endogenous dopamine transmission, as indexed by decreases in of the events was then spatially conjoined with a mask of regions that
[11C]raclopride binding potential, bilaterally in both the dorsal and had released dopamine according to the [11C]raclopride PET scan. We
ventral striatum (P < 0.001; Fig. 2a) when contrasting the pleasurable found that hemodynamic activity in the regions showing dopamine
music with the neutral music condition. The percentage of dopamine release was not constant throughout the excerpt, but was restricted
binding potential change was highest in the right caudate and the right to moments before and during chills and, critically, was anatomically
nucleus accumbens (NAcc; Fig. 2b and Supplementary Table 2). These distinct. During peak pleasure experience epochs, as compared with
results indicate that the experience of pleasure while listening to music neutral epochs, there was increased BOLD response in the right NAcc
is associated with dopamine release in striatal reward systems. (x, y, z = 8, 10, −8; t = 2.8; Fig. 3b). In contrast, increased BOLD
response was also found during the anticipation epochs, but was largely
fMRI data: temporal specificity of reward responses confined to the right caudate (x, y, z = 14, −6, 20; t = 3.2; Fig. 3b).
To gain information about the dynamics of dopamine release over time, The temporal dynamics of the reward response and its relationship
we acquired fMRI scans during presentation of pleasurable and neutral to the caudate and NAcc clusters can be more specifically analyzed
music excerpts. Listeners indicated by button press when they experi- by examining the percent BOLD signal change occurring over time
enced chills (mean = 3.1 chills per excerpt, s.d. = 0.9); these responses in relation to peak pleasure. To avoid the ‘circularity’ problem22, we

[11C]raclopride PET
Figure 4 Brain and behavior relationships
Combined hemodynamic and Number Intensity Reported
involving temporal components of pleasure
neurochemical activity of chills of chills pleasure
during music listening. Left, coronal slices
Percentage change

Immediately preceding chills


showing binding potential differences in dorsal 20
(top) and ventral (bottom) striatum that also
in BP

3 Caudate
show hemodynamic activity during anticipation 10
versus experience of chills, respectively. Right, y=4 r = 0.71*
behavioral ratings of the number and intensity 0
of chills and pleasure reported during the
During experience of chills
PET scans plotted against [11C]raclopride
Percentage change

20
binding potential changes in the two clusters. r = 0.84**
The number of chills reported was positively NAcc
in BP

10
correlated with percent binding potential 0 y = 11 r = 0.80*
change in the caudate (*P < 0.05), which was 0
linked to BOLD response immediately preceding 10 20 30 40 6 8 10 6 8 10
chills (that is, anticipatory periods), consistent
with the idea that a greater number of chills
would result in greater anticipation and result in more activity in the areas associated with anticipation. The mean intensity of chills and reported
pleasure were positively correlated with the NAcc (**P < 0.01), which was linked to BOLD response during chills, confirming that this region is involved
in the experience of the highly pleasurable component of music listening.

NATURE NEUROSCIENCE VOLUME 14 | NUMBER 2 | FEBRUARY 2011 259


ARTICLES

Figure 5 Brain and behavior relationships involving parametric increases 1.1 L caudate L putamen L Nacc/ventral
in pleasure during music listening. Relationship between real-time ratings (–13, 11, 7) (–29, –12, –8) putamen (–21, 9, –10)
0.5
of pleasure during music listening and percent BOLD signal change
relative to the mean in regions showing dopamine release as identified via 0

Percent signal change


PET. The chills epochs (shaded) were excluded from the analysis (values
–0.5
shown here only for reference) to examine activity related to increases in
pleasure irrespective of chills. A regression analysis revealed that the NAcc,
1.1 R caudate R putamen R NAcc
and to a lesser extent the left and right caudate, significantly predicted
(12, 6, 15) (26, –2, 0) (14, 10, –10)
increases in pleasure ratings during each of the conditions (P < 0.05 and 0.5
P < 0.001, respectively; Supplementary Table 4). This analysis indicates
0
that activity in these regions increased with pleasure even when no chills
were experienced.LP, low pleasure; HP, high pleasure. –0.5

Neut LP HP Chills Neut LP HP Chills Neut LP HP Chills


derived our voxels of interest (VOIs) from the PET data, which are Real-time subjective pleasure rating
independent of the fMRI data. This procedure also allowed us to
better integrate the hemodynamic and neurochemical results. We
found that activity in both the caudate and NAcc was increased during revealed a significant linear trend in which the percent signal change
anticipation as compared with the mean signal during the neutral in the right NAcc accounted for 67% of the variability in subjective
epochs for the same pieces of music, with larger increases occurring in pleasure ratings (t(19) = 6.18, P < 0.001). This finding suggests that
the caudate (Fig. 3c). During the peak emotional response, however, increases in subjective pleasure correspond to increases in neural
activity in the caudate decreased, whereas activity in the NAcc con- activity in the NAcc, in the same regions as those involved in the
© 2011 Nature America, Inc. All rights reserved.

tinued to increase. These findings support our fMRI contrast results chills responses and those that showed dopamine release in the PET
and provide temporal information as to how hemodynamic activity study, even though this analysis excluded all chills epochs.
in the regions showing dopamine release may contribute to reward Next, to ensure that increases in pleasure, irrespective of chills, are
processing in real time. not better predicted by activity in regions of the striatum other than
the right NAcc, we performed a similar analysis in all anatomical
Brain-behavior relationships clusters that had shown dopamine release in the PET study. We first
Once we had identified, via fMRI, the caudate and NAcc as contrib- selected peak voxels from each cluster showing dopamine release from
uting to the anticipation and experience, respectively, of peak plea- the PET data and then extracted the percent BOLD signal change as
sure moments during music listening, we used our PET scan data to listeners reported increases in pleasure from the fMRI data; as before,
further explore the brain and behavior relationships in these clus- all chills epochs were excluded. A stepwise multiple regression analy-
ters. Mean [11C]raclopride binding potential values from the NAcc sis was performed to examine which cluster’s hemodynamic responses
and caudate clusters was plotted against behavioral data obtained were best able to predict pleasure states. We found that hemodynamic
during PET scanning, which required participants to indicate the increase in the NAcc cluster was the most significant predictor
total number of chills, mean intensity of chills and mean subjective (P < 0.01) of increasing subjective pleasure (Supplementary Table 4).
pleasure experienced during each piece of music. We found that the However, at a lower statistical threshold of P < 0.05, bilateral caudate
number of chills was significantly correlated (P < 0.05) with binding clusters and the left NAcc/ventral putamen cluster could also pre-
potential differences in the right caudate, but not the NAcc, whereas dict pleasure states, but to a lower degree (31% and 43% for left and
the intensity of chills and overall degree of pleasure experienced were right caudate, respectively, and 37% for the NAcc/ventral putamen).
most significantly correlated (P < 0.01) with binding potential change Recruitment of the caudate is not surprising considering that anticipa-
in the right NAcc, but not the caudate (Fig. 4 and Supplementary tory periods result in a culmination of pleasurable emotional experi-
Table 3). This finding further supports a functional dissociation in ences and the caudate was recruited during these pleasant anticipatory
the contribution of these anatomical regions to pleasure associated moments. Indeed, the mean subjective pleasure rating provided by
with music listening. listeners during the anticipatory epochs was 2.51 (s.d. = 0.55), which
An additional question is whether increases in pleasure alone, in was significantly higher than that of the entire excerpt (mean = 2.11,
the absence of chills, result in increased hemodynamic responses in s.d. = 0.019; t(246) = 8.5, P < 0.001).
the same areas as during the experience of chills, although perhaps Finally, when the percent BOLD signal change during the chills
not to the same extent. We examined this question by determining epochs was included in the multiple regression analysis (Fig. 5), it was
whether there was a linear relationship between increases in plea- apparent that the experience of chills represents the highest point of
sure and hemodynamic activity in the right NAcc, irrespective of hemodynamic activity in the NAcc. These findings converge to sug-
chills, and how this compared with other striatal regions showing gest that the dorsal and ventral subdivisions of the striatum are most
dopamine release. This analysis was done by excluding all of the involved during anticipation and experience of the peak emotional
epochs during which individuals experienced chills and examining responses during music listening, respectively.
BOLD signal changes that related to increasing pleasure in the right
NAcc. Using the voxel that showed the maximum dopamine release DISCUSSION
in the NAcc during the [11C]raclopride scan, we calculated the per- Our results provide, to the best of our knowledge, the first direct evi-
cent BOLD signal change as subjective pleasure ratings increased dence that the intense pleasure experienced when listening to music
from neutral to low pleasure to high pleasure (excluding chills) for is associated with dopamine activity in the mesolimbic reward system,
each individual. Note that this analysis, unlike the one presented including both dorsal and ventral striatum. This phylogenetically
above, does not take into account the temporal component, as all ancient circuitry has evolved to reinforce basic biological behaviors
epochs rated as having the same pleasure were averaged, regardless with high adaptive value. However, the rewarding qualities of music
of when they occurred with respect to chills. A regression analysis listening are not obviously directly adaptive. That is, musical stimuli,

260 VOLUME 14 | NUMBER 2 | FEBRUARY 2011 NATURE NEUROSCIENCE


ARTICLES

similar to other aesthetic stimuli, are perceived as being rewarding It is important to note that chills are not necessarily pleasurable
by the listener, rather than exerting a direct biological or chemical per se, as they can be unpleasant in other contexts (for example, as
influence. Furthermore, the perception that results in a rewarding a result of intense fear). Instead, chills are physiological markers of
response is relatively specific to the listener, as there is large vari- intense ANS arousal5,15,33,34, which in turn is believed to underlie
ability in musical preferences amongst individuals. Thus, through peak pleasure during music listening5,15; we used chills here only to
complex cognitive mechanisms, humans are able to obtain pleasure allow objective quantification of a highly subjective response that
from music2, a highly abstract reward consisting of just a sequence would be otherwise difficult to measure and because they afford preci-
of tones unfolding over time, which is comparable to the pleasure sion as to the time at which the peak pleasure occurred. As such, chills
experienced from more basic biological stimuli. are byproducts, and not a cause of the emotional responses. Thus,
One explanation for this phenomenon is that it is related to enhance- it is important to clarify that, although chills index peak emotional
ment of emotions3,15,20. The emotions induced by music are evoked, responses in this group of people, the specific experience of chills is
among other things, by temporal phenomena, such as expectations, not necessary to result in neural activity in the striatum, a finding
delay, tension, resolution, prediction, surprise and anticipation11,19. that is consistent with less-specific analyses performed in previous
Indeed, we found a temporal dissociation between distinct regions studies6–8. This conclusion is confirmed by our findings that, even
of the striatum while listening to pleasurable music. The combined when the chills epochs were excluded from the analysis, there was still
psychophysiological, neurochemical and hemodynamic procedure a significant linear relationship between increases in self-reported
that we used revealed that peaks of ANS activity that reflect the expe- pleasure and increases in hemodynamic activity in the regions that
rience of the most intense emotional moments are associated with showed dopamine release (Fig. 5). Furthermore, when chills were
dopamine release in the NAcc. This region has been implicated in reported, maximal signal was seen in the NAcc voxels that showed
the euphoric component of psychostimulants such as cocaine23 and a linear increase as participants progressed from neutral to low plea-
© 2011 Nature America, Inc. All rights reserved.

is highly interconnected with limbic regions that mediate emotional sure to high pleasure, further confirming that chills represented the
responses, such as the amygdala, hippocampus, cingulate and ven- peak of pleasure in this group. This finding is also consistent with the
tromedial prefrontal cortex24. In contrast, immediately before the finding that the degree of binding potential decrease in the NAcc for
climax of emotional responses there was evidence for relatively greater each participant was positively correlated with the degree of pleasure
dopamine activity in the caudate. This subregion of the striatum is reported from listening to the musical excerpts, irrespective of the
interconnected with sensory, motor and associative regions of the number of chills that were experienced (Fig. 4).
brain24,25 and has been typically implicated in learning of stimulus- It should be noted that there was some activity in the ventral stria-
response associations24,26 and in mediating the reinforcing qualities tum during the anticipation phase at lower statistical thresholds,
of rewarding stimuli such as food27. Our findings indicate that a sense consistent with other studies using different stimuli24. However, we
of emotional expectation, prediction and anticipation in response to found that, during the anticipatory phase, there was also increased
abstract pleasure can also result in dopamine release, but primarily in BOLD response in the caudate (more so than the NAcc), which then
the dorsal striatum. Previous studies have found that amphetamine- shifted more ventromedially as participants reported experiencing
induced dopamine release in the NAcc spreads to more dorsal regions peak reward (Fig. 3). This is an important finding because the stimu-
after repeated exposure to the drug28, which suggests that this area lus that we are using is a dynamic reward with a temporal component,
may be involved in improved predictability and anticipation of a allowing examination of the reward in real time as it progresses from
reward. Similarly, previous studies involving rewards such as food anticipation to peak pleasure states, which is generally not possible
and smoking that contain a number of contextual predicting cues because of limitations with movement inside the PET scanner. Some
(for example, odor and taste) also found dorsal striatum dopamine studies administered the pleasurable stimulus (for example, food)
release27,29. Conversely, in studies in which there were no contextual immediately before the scan and measured subsequent dopamine
cues or experience with the drugs involved, dopamine release was release27, in which case anticipation and consumption cannot be dis-
largely observed in the ventral striatum30,31. Finally, evidence from tinguished. Other studies measured the anticipation phase online,
animal research also suggests that, as rewards become better pre- with the promise of the delivery of the tangible reward after the scan,
dicted, the responses that initiated in the ventral regions move more in which case the consumption phase is missed35,36. Music is a unique
dorsally in the striatum32. These results are consistent with a model in reward that allows assessment of all reward phases online, from the
which repeated exposure to rewards associated with a specific context point that a single note is heard to the point at which maximum plea-
gradually shift the response from ventral to dorsal and further suggest sure is reached.
that contextual cues that allow prediction of a reward, in our case the The anatomical dissociation between the anticipatory and consum-
sequences of tones leading up to the peak pleasure moments, may also matory phases during intensely pleasurable music listening suggests
act as reward predictors mediated via the dorsal striatum. that distinct mechanisms are involved. This distinction may map onto
Another noteworthy finding is the correspondence between behavio- the ‘wanting’ and ‘liking’ phases of a reward in an error prediction
ral and imaging results, which strengthens the evidence for the distinct model37. The anticipatory phase, set off by temporal cues signaling
roles of dorsal and ventral striatum. We found a positive correlation that a potentially pleasurable auditory sequence is coming, can trigger
between subject-reported intensity of chills and dopamine release in the expectations of euphoric emotional states and create a sense of want-
NAcc during [11C]raclopride PET scanning (Fig. 4), which confirms the ing and reward prediction. This reward is entirely abstract and may
fMRI results that peak pleasure responses are associated with this region. involve such factors as suspended expectations and a sense of resolu-
Furthermore, the number of chills reported by listeners during the PET tion. Indeed, composers and performers frequently take advantage
scan was correlated with dopamine release in the caudate (Fig. 4), of such phenomena, and manipulate emotional arousal by violating
which is consistent with the fMRI results showing increased activity in expectations in certain ways or by delaying the predicted outcome
this region during anticipation of peak emotional responses; as greater (for example, by inserting unexpected notes or slowing tempo) before
number of chills suggests increased incidence of anticipation, greater the resolution to heighten the motivation for completion. The peak
dopamine release would be expected in this area. emotional response evoked by hearing the desired sequence would

NATURE NEUROSCIENCE VOLUME 14 | NUMBER 2 | FEBRUARY 2011 261


ARTICLES

represent the consummatory or liking phase, representing fulfilled 4. Krumhansl, C.L. An exploratory study of musical emotions and psychophysiology.
Can. J. Exp. Psychol. 51, 336–353 (1997).
expectations and accurate reward prediction. We propose that each 5. Blood, A.J. & Zatorre, R.J. Intensely pleasurable responses to music correlate with
of these phases may involve dopamine release, but in different sub- activity in brain regions implicated in reward and emotion. Proc. Natl. Acad. Sci.
circuits of the striatum, which have different connectivity and func- USA 98, 11818–11823 (2001).
6. Menon, V. & Levitin, D.J. The rewards of music listening: response and physiological
tional roles. connectivity of the mesolimbic system. Neuroimage 28, 175–184 (2005).
The notion that dopamine can be released in anticipation of an 7. Koelsch, S., Fritz, T., Cramon, D., Muller, K. & Friederici, A.D. Investigating emotion
abstract reward (a series of tones) has important implications for with music: an fMRI study. Hum. Brain Mapp. 27, 239–250 (2006).
8. Mitterschiffthaler, M.T., Fu, C.H.Y., Dalton, J., Andrew, C.M. & Williams, S.
understanding how music has become pleasurable. However, the pre- A functional MRI study of happy and sad affective states induced by classical
cise source of the anticipation requires further investigation. A sense music. Hum. Brain Mapp. 28, 1150–1162 (2007).
9. Knutson, B. & Gibbs, S.E. Linking nucelus accumbens dopamine and blood
of anticipation may arise through one’s familiarity with the rules that oxygenation. Psychopharmacology (Berl.) 191, 813–822 (2007).
underlie musical structure, such that listeners are anticipating the next 10. Laruelle, M. Imaging synaptic neurotransmission with in vivo binding competition
note that may violate or confirm their expectations, in turn leading techniques: a critical review. J. Cereb. Blood Flow Metab. 20, 423–451 (2000).
11. Huron, D. & Hellmuth Margulis, E. Musical expectancy and thrills. in Music and
to emotional arousal, or alternatively it may arise through familiarity Emotion (eds. Juslin, P.N. & Sloboda, J.) (Oxford University Press, New York, 2009).
with a specific piece and knowing that a particularly pleasant section 12. Grewe, O., Nagel, F., Kopiez, R. & Altenmuller, E. Emotions over time: synchronicity
is coming up11. These components are not mutually exclusive, as the and development of subjective, physiological, and facial affective reactions to music.
Emotion 7, 774–788 (2007).
second likely evolves from the first, and the overall anticipation is 13. Panksepp, J. The emotional source of “chills” induced by music. Music Percept. 13,
likely to be a combination of both. Nonetheless, the subtle differences 171–207 (1995).
14. Sloboda, J. Music structure and emotional response: some empirical findings.
that exist between them will need to be disentangled through future Psychol. Music 19, 110–120 (1991).
experiments that are specifically designed to parse out this distinc- 15. Salimpoor, V.N., Benovoy, M., Longo, G., Cooperstock, J.R. & Zatorre, R.J. The
tion. Abstract rewards are largely cognitive in nature and our results rewarding aspects of music listening are related to degree of emotional arousal.
© 2011 Nature America, Inc. All rights reserved.

PLoS ONE 4, e7487 (2009).


pave the way for future work to examine nontangible rewards that 16. O’Doherty, J.P., Deichmann, R., Critchley, H.D. & Dolan, R. Neural responses during
humans consider rewarding for complex reasons. anticipation of a primary taste reward. Neuron 33, 815–826 (2002).
Dopamine is pivotal for establishing and maintaining behavior. If 17. Schultz, W., Dayan, P. & Montague, P.R. A neural substrate of prediction and reward.
Science 275, 1593–1599 (1997).
music-induced emotional states can lead to dopamine release, as our 18. Wise, R.A. Dopamine, learning and motivation. Nat. Rev. Neurosci. 5, 483–494 (2004).
findings indicate, it may begin to explain why musical experiences 19. Huron, D. Sweet Anticipation: Music and the Psychology of Expectation (MIT Press,
Cambridge, Massachusetts, 2006).
are so valued. These results further speak to why music can be effec- 20. Meyer, L.B. Emotion and Meaning in Music. (University of Chicago Press, Chicago,
tively used in rituals, marketing or film to manipulate hedonic states. 1956).
Our findings provide neurochemical evidence that intense emotional 21. Schott, B.H. et al. Mesolimbic functional magnetic resonance imaging activations
during reward anticipation correlate with reward-related ventral striatal dopamine
responses to music involve ancient reward circuitry and serve as a release. J. Neurosci. 28, 14311–14319 (2008).
starting point for more detailed investigations of the biological sub- 22. Kriegeskorte, N., Simmons, W.K., Bellgowan, P.S. & Baker, C.I. Circular analysis in
strates that underlie abstract forms of pleasure. systems neuroscience: the dangers of double dipping. Nat. Neurosci. 12, 535–540
(2009).
23. Volkow, N.D. et al. Relationship between subjective effects of cocaine and dopamine
METHODS transporter occupancy. Nature 386, 827–830 (1997).
24. Haber, S. & Knutson, B. The reward circuit: linking primate anatomy and human
Methods and any associated references are available in the online imaging. Neuropharmacology 35, 4–26 (2010).
version of the paper at http://www.nature.com/natureneuroscience/. 25. Haber, S.N., Kim, K.S., Mailly, P. & Calzavara, R. Reward-related cortical inputs
define a large striatal region in primates that interface with associative cortical
Note: Supplementary information is available on the Nature Neuroscience website. connections, providing a substrate for incentive-based learning. J. Neurosci. 26,
8368–8376 (2006).
ACKNOWLEDGMENTS 26. Valentin, V.V. & O’Doherty, J.P. Overlapping prediction errors in dorsal striatum
We thank the staff of the Montreal Neurological Institute PET and MR Units during instrumental learning with juice and money reward in the human brain.
and the staff of the Centre for Interdisciplinary Research in Music Media and J. Neurophysiol. 102, 3384–3391 (2009).
Technology for help with data acquisition, M. Ferreira and M. Bouffard for their 27. Small, D.M., Jones-Gotman, M. & Dagher, A. Feeding-induced dopamine release in
assistance with data analysis, and G. Longo for assistance with stimulus preparation. dorsal striatum correlates with meal pleasantness ratings in healthy human
This research was supported by funding from the Canadian Institutes of Health volunteers. Neuroimage 19, 1709–1715 (2003).
28. Boileau, I. et al. Modeling sensiitization to stimulants in humans: an [11C]raclopride/
Research to R.J.Z., a Natural Science and Engineering Research Council stipend to
positron emission tomography study in healthy men. Arch. Gen. Psychiatry 63,
V.N.S., a Jeanne Timmins Costello award to V.N.S. and Centre for Interdisciplinary 1386–1395 (2006).
Research in Music Media and Technology awards to V.N.S. and M.B. 29. Barrett, S.P., Boileau, I., Okker, J., Pihl, R.O. & Dagher, A. The hedonic response
to cigarette smoking is proportional to dopamine release in the human striatum as
AUTHOR CONTRIBUTIONS measured by positron emission tomography and [11C]raclopride. Synapse 54,
V.N.S., R.J.Z. and A.D. designed the study. V.N.S. and M.B. performed all experiments. 65–71 (2004).
V.N.S., M.B. and K.L. analyzed the data. V.N.S. and R.J.Z. wrote the manuscript. 30. Leyton, M. et al. Amphetamine-induced increases in extracellular dopamine, drug
wanting, and novelty seeking: a PET/[11C]raclopride study in healthy men.
COMPETING FINANCIAL INTERESTS Neuropsychopharmacology 27, 1027–1035 (2002).
The authors declare no competing financial interests. 31. Boileau, I. et al. Alcohol promotes dopamine release in the human nuclus
accumbens. Synapse 49, 226–231 (2003).
Published online at http://www.nature.com/natureneuroscience/. 32. Everitt, B.J. & Robbins, T. Neural systems of reinforcement for drug addiction: from
Reprints and permissions information is available online at http://www.nature.com/ actions to habits to compulsion. Nat. Neurosci. 8, 1481–1489 (2005).
reprintsandpermissions/. 33. Rickard, N.S. Intense emotional responses to music: a test of the physiological
arousal hypothesis. Psychol. Music 32, 371–388 (2004).
34. Grewe, O., Kopiez, R. & Altenmuller, E. Chills as an indicator of individual emotional
1. Egerton, A. et al. The dopaminergic basis of human behaviors: a review of molecular peaks. Ann. NY Acad. Sci. 1169, 351–354 (2009).
imaging studies. Neurosci. Biobehav. Rev. 33, 1109–1132 (2009). 35. Zald, D.H. et al. Dopamine transmission in the human striatum during monetary
2. Dube, L. & Lebel, J. The content and structure of laypeople’s concept of pleasure. reward tasks. J. Neurosci. 24, 4105–4112 (2004).
Cogn. Emot. 17, 263–295 (2003). 36. Koepp, M.J. et al. Evidence for striatal dopamine release during a video game.
3. Sloboda, J. & Juslin, P.N. Psychological perspectives on music and emotion. in Nature 393, 266–268 (1998).
Music and Emotion: Theory and Research (ed. Sloboda, J.) 71–104 (Oxford 37. Zald, D.H. & Zatorre, R.J. On music and reward. in The Neurobiology of Sensation
University Press, Oxford, 2001). and Reward (ed. Gottfried, J.A.) (CRC Press, 2011).

262 VOLUME 14 | NUMBER 2 | FEBRUARY 2011 NATURE NEUROSCIENCE


ONLINE METHODS matching algorithm to the MNI template39. All transformed images were visually
Participant screening and stimulus selection. 217 individuals responded to inspected to ensure that there were no alignment errors.
advertisements requesting people who experience chills to music; after five rounds Parametric images were generated in the native PET space by computing
of screening, the final group included eight participants. First, individuals pro- [11C]raclopride binding potential (binding potential = BAvail / KD, where BAvail
vided ten pieces of instrumental music to which they experience intense pleasure is the density of available receptors and KD is the dissociation constant) at each
and “chills” without restrictions to the genre of music, which included classical, voxel of interest40,41. Voxelwise [11C]raclopride binding potential was calculated
folk, jazz, electronica, rock, punk, techno and tango (see http://www.zlab.mcgill. using a simplified reference region method40,41, with the cerebellum chosen as
ca/supplements/supplements_intro.html for samples). Next, an email question- reference region because it does not contain specific D2 receptor–like binding
naire was completed to determine whether their chills were experienced at times sites and can be used for the determination of nonspecific binding and free radio-
of extreme pleasure, consistently at the same point in the music without diminish- ligand in the brain42. The gray matter of the cerebellum assigned as reference
ing on multiple listening, in different environments, and the selected music was region was initially segmented in Talairach space from a probabilistic atlas43
not specifically or generally associated with an episodic memory. 45 individuals and a neural net classifier44. The [11C]raclopride binding potential maps were
continued to the third screening session, where a history of medical, psychiatric then transformed into MNI space39 using the previously determined transforma-
illness, or substance abuse was ruled out. 40 participants continued to the fourth tion parameters. Statistical parametric t maps of binding potential change were
screening session, where control stimuli were selected for each individual using produced by comparing the parametric binding potential maps of the two scan
a paradigm where one individual’s pleasurable music is used as another person’s sessions (pleasurable music and neutral music), using a previously described
neutral music5,15. This way, group-averaged data analysis involves comparison method45. This calculation uses the residuals of the least-squares fit of the com-
of similar sets of stimuli. Although we were not able to match perfectly between partmental model, which improves the sensitivity to small changes by providing
the control and pleasurable pieces used for all participants, efforts were made to better estimates of the s.d. at the voxel and by increasing the degrees of freedom.
ensure that pieces were as evenly distributed as possible. Each individual rated It is assumed that a reduction in [11C]raclopride binding potential is indicative
other participants’ music on a scale of 1–10 (neutral to extremely pleasurable). of an increase in extracellular dopamine concentration46. Clusters of significant
From the pieces rated neutral, the ones that were most familiar to that subject change were defined as all contiguous striatal voxels on the t map exceeding a
© 2011 Nature America, Inc. All rights reserved.

were selected to minimize differences in familiarity between pleasurable music magnitude threshold of 3.11. This threshold was considered to be significant
and neutral music conditions. Individuals whose music was found to be “neutral” (P < 0.05, corrected for multiple comparisons) for a search volume equal to the
by at least one other participant were asked to continue. Participants were asked striatum and an effective spatial resolution of 8-mm full-width at half maximum
not to listen to those pieces anymore during the course of the study to ensure (FWHM)47. Mean binding potential values were extracted from each significant
maximal responses during testing. 28 individuals participated in the final screen- cluster for each individual and percent change in binding potential was calculated
ing session to verify the chills response at prespecified times through subjective as [(BPneutral − BPpleasurable) × 100 / BPneutral], and compared with subjectively
and physiological responses. Participants listened to their chills-inducing music reported post-listening ratings of the number of chills, intensity of chills and
while providing subjective ratings of pleasure through button presses and indi- degree of pleasure experienced.
cating when they experienced a chill (see ref. 15 for additional details). The ten
participants (five female, five male) who most reliably experienced chills during fMRI. One scan was terminated because of claustrophobia. fMRI data were cor-
their peak pleasure responses to music accompanied by clear increases in ANS rected for motion using in-house software. To increase the signal-to-noise ratio,
activity were selected for the study. The final group of participants was between we spatially smoothed the images (or low-pass filtered) with an 8-mm FWHM
the ages of 19 and 24 (M = 20.8, o 1.9 years) and had a wide range of musical isotropic Gaussian kernel. Image analyses were performed with fMRISTAT, which
experiences from no training to 15 years of experience. consists of a series of MATLAB scripts that utilize the general linear model for
analyses48. The general linear model (Y = XB + E) expresses the response vari-
Procedures. Ethical approval for the study was granted by the Montreal able (BOLD signal) Y in terms of a linear combination of explanatory variables
Neurological Institute (MNI) Research Ethics Board. All individuals gave writ- (events) X, the parameter estimates (effects of interest) B and the error term E.
ten informed consent before participating in the study. Testing took place over Temporal drift was modeled as cubic splines and removed by inclusion into the
three sessions. The first two sessions involved PET scanning (Supplementary general linear model as a variable of non-interest. The linear model was solved
Fig. 1) and psychophysiological recording (Supplementary Fig. 2) and the third for the parameter estimates B with least squares, yielding estimates of effects,
session involved fMRI scanning (Supplementary Fig. 3). standard errors and t statistics for each contrast and for each run.
Before group statistical maps for each contrast of interest were generated, in-
Statistical analysis. Signal filtering was performed to remove noise and artifacts house software was used to linearly transform anatomical and functional images
(see ref. 15 for additional details). Data were downsampled to 1-s epochs and from each subject into standard MNI stereotaxic coordinate space using the MNI
compared across neutral music and pleasurable music conditions. To account 305 template39. A mixed-effects linear model was subsequently used to combine
for unequal variances across conditions, we used Welch’s t test. A second analysis data across subjects; the s.d. images were smoothed with a Gaussian filter so that
was performed to examine the relationship between the intensity of chills expe- the ratio of the random-effects variance divided by the fixed-effects variance
rienced and psychophysiological responses. Outliers beyond four s.d. from the results in approximately 100 degrees of freedom. Because the main purpose of
mean were removed for each excerpt and for each participant individually (2–5% the fMRI analyses was to measure BOLD activity in predefined striatal regions,
of the data points). Subjective ratings for one individual were not recorded and we adopted an uncorrected statistical threshold of P < 0.01.
BVP amplitude data for one participant demonstrated excessive artifacts, thus For the main analysis, three events were defined: the peak emotional response
these data were not included in the analysis. Z score values of each biosignal (PER) condition represented all epochs during which the participant was pressing
were calculated for each excerpt and plotted against subjective ratings of chills the chills, the anticipation condition represented 15-s epochs immediately pre-
intensity that subjects reported after hearing each excerpt (Fig. 1). Correlation ceding the onset of the PER condition defined post hoc, and the neutral condition
coefficients were calculated for the intensity of chills and changes in each of the represented all epochs during which participants were pressing down the neutral
psychophysiological measures (Supplementary Table 1). button. Note that these neutral epochs are different from the neutral music condi-
For [11C]raclopride PET, we discarded two datasets because of participant tion, which were not used in this case, as the neutral music condition contrasted
discomfort during the first session. Data from the remaining eight participants with the pleasurable music condition shows less activity in the striatum. As such,
were analyzed. PET emission frames were reconstructed and corrected for any epoch selected from the pleasurable music condition, even those not related
gamma ray attenuation and scatter. All PET images were corrected for head to peak pleasure, could have shown increased striatal activity and overestimated
motion using a co-registration–based method, which performs interframe the results of the study. The anticipation period was defined as the 15 s before
realignment and compensates for emission-transmission mismatches38. The the PER based on previous findings that this is the time frame during which
motion-corrected PET data were summed over the time dimension and aligned psychophysiological responses begin to increase significantly relative to mean
to the subject’s anatomical magnetic resonance image. Anatomical MRI were responses throughout the excerpt15. The times at which participants pressed the
transformed into standardized stereotaxic space by means of automated feature low pleasure and high pleasure buttons were also included in the model to ensure

doi:10.1038/nn.2726 NATURE NEUROSCIENCE


that they did not contribute to baseline. A 0.1-s epoch was incorporated into the by areas that showed dopamine release. A spatial conjunction analysis was
model each time a button was pressed to account for neural activity involved in performed to examine the temporal aspects of hemodynamic activity in areas
button pressing. The BOLD data from times when participants were responding that had shown changes in [11C]raclopride binding potential on PET. A mask of
to questions were excluded from the analysis. The planned comparisons for the striatal areas that had revealed substantial changes in binding potential using the
main analysis were then entered into the analysis: anticipation of PER = anticipa- stated threshold (t q 3.11) was created to spatially mask both contrasts (outlined
tion condition minus neutral condition and experience of PER = PER condition in the fMRI data analysis section): anticipation of PER and experience of PER.
minus neutral condition. This procedure allowed us to measure BOLD changes only in voxels that had
shown binding potential differences in the PET study.
Time series analysis. To further investigate the temporal dynamics of the reward
response, we calculated the time series of hemodynamic activity in the caudate and
NAcc clusters. To avoid the circularity problem22, we derived our VOIs from the 38. Costes, N. et al. Motion correction of multi-frame PET data in neuroreceptor
mapping: simulation based validation. Neuroimage 47, 1496–1505 (2009).
PET data, which are independent of the fMRI data. We first identified the voxel
39. Collins, D.L., Peters, T.M. & Evans, A.C. Automatic 3D intersubject registration of
showing the maximum dopamine release during the [11C]raclopride PET scan, MR volumetric data in standardized Talirach space. J. Comput. Assist. Tomogr. 18,
in the caudate and NAcc clusters. We then extracted the mean signal for each 192–205 (1994).
VOI during the entire fMRI run obtained from each volume and calculated the 40. Gunn, R.N., Lammertsma, A.A., Hume, S.P. & Cunningham, V.J. Parametric imaging
of ligand-receptor binding in PET using a simplified reference region model.
percent BOLD signal change relative to the mean of the run during the epochs in
Neuroimage 6, 279–287 (1997).
which PERs were reported. Participants often experienced multiple chills one after 41. Lammertsma, A.A. & Hume, S.P. Simplified reference tissue model for PET receptor
another. For the purposes of this analysis, the percent signal change during the studies. Neuroimage 4, 153–158 (1996).
first chill of the series was used, which ranged in duration from 1–4 s. The BOLD 42. Litton, J.E., Hall, H. & Pauli, S. Saturation analysis in PET-analysis of errors due
to non-perfect reference regions. J. Cereb. Blood Flow Metab. 14, 358–361
response for each of those seconds is plotted in Figure 3c. Mean signal change
(1994).
for each second preceding this response for each individual, up to 15 s, was also 43. Collins, D.L. & Evans, A.C. ANIMAL: Validation and application of nonlinear
plotted to demonstrate hemodynamic time series during the anticipation period. registration-based segmentation. Intern. J. Pattern Recognit. Artif. Intell. 11,
© 2011 Nature America, Inc. All rights reserved.

As a result of cardiac gating, a different number of frames were acquired for each 1271–1294 (1997).
44. Zijdenbos, A., Forghani, R. & Evans, A.C. Automatic quantification of MS lesions
person during this 15-s period and acquisition time varied from 2.1 to 3 s depend-
in 3D MRI Brain data sets: validation of INSECT. in Medical Image Computing and
ing on the individual’s heart rate. As such, the VOI values obtained at each frame Computer-Assisted Intervention (eds. Wells, W.M., Colchester, A. & Delp, S.)
were interpolated to provide an estimate of signal during each second preceding 439–448 (Springer-Verlag, Cambridge, Massachusetts, 1998).
the peak response. The mean number of frames sampled for calculating time 45. Aston, J.A. et al. A statistical method for the analysis of positron emission
tomography neuroreceptor ligand data. Neuroimage 12, 245–256 (2000).
series was 5.3 (s.d. = 1.3) during anticipation and 1.6 (s.d. = 1.2) during chills. The
46. Endres, C.J. et al. Kinetic modeling of [11C]raclopride: combined PET microdialysis
mean signal change during neutral button presses was also calculated for each VOI studies. J. Cereb. Blood Flow Metab. 17, 932–942 (1997).
separately and plotted in Figure 3d for reference. Finally, the percent signal change 47. Worsley, K.J. et al. A unified statistical approach for determining significant signals
for 5 s preceding the anticipatory response were also plotted for reference. in images of cerebral activation. Hum. Brain Mapp. 4, 58–73 (1996).
48. Worsley, K.J. et al. A general statistical analysis for fMRI data. Neuroimage 15,
1–15 (2002).
Conjunction analysis. Because [11C]raclopride binds with D2 receptors mainly 49. Slifstein, M. et al. Striatal and extrastriatal dopamine release measured with PET
in the striatum49, our fMRI data analysis was also limited to this region, masked and [(18)F]fallypride. Synapse 64, 350–362 (2010).

NATURE NEUROSCIENCE doi:10.1038/nn.2726


Marketing actions can modulate neural
representations of experienced pleasantness
Hilke Plassmann*, John O’Doherty*, Baba Shiv†, and Antonio Rangel*‡
*Division of the Humanities and Social Sciences, California Institute of Technology, MC 228-77, Pasadena, CA 91125; and †Stanford Graduate School
of Business, Stanford University, 518 Memorial Way, Littlefield L383, Stanford, CA94305

Edited by Leslie G. Ungerleider, National Institutes of Health, Bethesda, MD, and approved December 3, 2007 (received for review July 24, 2007)

Despite the importance and pervasiveness of marketing, almost experiences and, through this, the actual quality of experiences
nothing is known about the neural mechanisms through which it (2, 7, 8). Consider, for example, the experience of an individual
affects decisions made by individuals. We propose that marketing sampling a wine for which he or she has information about its retail
actions, such as changes in the price of a product, can affect neural price. Because perceptions of quality are known to be positively
representations of experienced pleasantness. We tested this hy- correlated with price (9), the individual is likely to believe that a
pothesis by scanning human subjects using functional MRI while more expensive wine will probably taste better. Our hypothesis goes
they tasted wines that, contrary to reality, they believed to be beyond this by stipulating that higher taste expectations would lead
different and sold at different prices. Our results show that to higher activity in the medial orbitofrontal cortex (mOFC), an
increasing the price of a wine increases subjective reports of flavor area of the brain that is widely thought to encode for actual
pleasantness as well as blood-oxygen-level-dependent activity in experienced pleasantness (6, 10–16). The results described below
medial orbitofrontal cortex, an area that is widely thought to are consistent with this hypothesis. We found that the reported
encode for experienced pleasantness during experiential tasks. The price of wines markedly affected reported EP and, more impor-
paper provides evidence for the ability of marketing actions to tantly, also modulated the blood-oxygen-level-dependent (BOLD)
modulate neural correlates of experienced pleasantness and for signal in mOFC.
the mechanisms through which the effect operates. To investigate the impact of price on the neural computations
associated with EP, we scanned human subjects (n ! 20) using
orbitofrontal cortex ! modulation by marketing actions ! fMRI while they sampled different wines and an affectively neutral
neuroeconomics ! taste control solution, which consisted of the main ionic components of
human saliva (17). We chose wine as a stimulus, because it is
relatively easy to administer inside the scanner using computerized
A basic assumption in economics is that the experienced
pleasantness (EP) from consuming a good depends only on
its intrinsic properties and on the state of the individual (1).
pumps, it induces a pleasurable flavor sensation in most subjects,
and it varies widely in quality and retail price. Subjects were told
they were sampling five different Cabernet Sauvignons, that the
Thus, the pleasure derived from consuming a soda should
purpose of the experiment was to study the effect of degustation
depend only on the molecular composition of the drink and the
time on perceived flavors, and that the different wines would be
level of thirst of the individual. In opposition to this view, a
identified by their retail prices (see Fig. 1A). Unbeknown to the
sizable number of marketing actions attempt to influence EP by
subjects, the critical manipulation was that there were only three
changing properties of commodities, such as prices, that are different wines, and two of them (wines 1 and 2) were administered
unrelated to their intrinsic qualities or to the consumer’s state. twice, one identified at a high price and one at a low price. For
This type of influence is valuable for companies, because EP example, wine 2 was presented half of the time at $90, its retail price,
serves as a learning signal that is used by the brain to guide future and half of the time at $10. Thus, the task consisted of six trial types:
choices. For example, when facing the choice between previously $5 wine (wine 1), $10 wine (wine 2), $35 wine (wine 3), $45 wine
experienced restaurants, one would tend to avoid locales where (wine 1), $90 wine (wine 2), and neutral solution. The wines were
previously meals were unsavory. Contrary to the basic assump- administered in random order, simultaneously with the appearance
tions of economics, several studies have provided behavioral of the price cue. Subjects were asked to focus on the flavor of the
evidence that marketing actions can successfully affect EP by wine during the degustation period and entered taste pleasantness
manipulating nonintrinsic attributes of goods. For example, or taste intensity ratings in every other trial (Fig. 1B).
knowledge of a beer’s ingredients and brand can affect reported
taste quality (2, 3), and the reported enjoyment of a film is Results
influenced by expectations about its quality (4). Even more Modulation of Reported Pleasantness and Taste Intensity by Price. We
intriguingly, changing the price at which an energy drink is measured the impact of price information on EP by comparing
purchased can influence the ability to solve puzzles (5). the mean reported liking rating for wines 1 and 2 when admin-
Despite the importance and pervasiveness of various market- istered at a high vs. a low price. We found significant differences
ing actions, very little is known about the neural mechanisms for both wines (P " 0.001, Fig. 1C). In addition, reported
through which they affect decisions made by individuals. An pleasantness was correlated with wine prices (r ! 0.59, P "
exception is a previous study demonstrating that knowledge of 0.000). We could not find a similar behavioral effect for intensity
the brand of a culturally familiar drink, such as Coke, increases ratings (Fig. 1D). To explore further the role of prices on
activation in the hippocampus, parahipoccampus, midbrain,
dorsolateral prefrontral cortex, and thalamus (6). The authors of
the previous study interpreted such activity as evidence for Author contributions: H.P., J.O., B.S., and A.R. designed research; H.P. performed research;
retrieval of brand information during the consumption H.P. analyzed data; and H.P., J.O., B.S., and A.R. wrote the paper.

experience. The authors declare no conflict of interest.


Here, we propose a mechanism through which marketing actions This article is a PNAS Direct Submission.
can affect decision making. We hypothesized that changes in the ‡To whom correspondence should be addressed. E-mail: rangel@hss.caltech.edu.
price of a product can influence neural computations associated This article contains supporting information online at www.pnas.org/cgi/content/full/
with EP. This hypothesis is based on previous findings showing that 0706929105/DC1.
affective expectations influence appraisals made about hedonic © 2008 by The National Academy of Sciences of the USA

1050 –1054 ! PNAS ! January 22, 2008 ! vol. 105 ! no. 3 www.pnas.org"cgi"doi"10.1073"pnas.0706929105
Fig. 2. The effect of price on each wine. (A) Wine 1: averaged time courses
in the medial OFC voxels shown in B (error bars denote standard errors). (B)
Fig. 1. Experimental design and behavioral results. (A) Time course for a Wine 1: activity in the mOFC was higher for the high- ($45) than the low-price
typical trial. (B) Reported pleasantness and intensity rating scales. (C) Reported condition ($5). Activation maps are shown at a threshold of P ! 0.001 uncor-
pleasantness for the wines during the cued price trials. (D) Taste intensity rected and with an extend threshold of five voxels. (C) Wine 1: activity in the
ratings for the wines during the cued price trials. (E) Reported pleasantness for vmPFC was also selected by the same contrast. (D) Wine 2: averaged time
the wines obtained during a postexperimental session without price cues. courses in the medial OFC voxels shown in E. (E) Wine 2: activity in the mOFC

NEUROSCIENCE
was higher for the high- ($90) than for the low-price condition ($10). (F) Wine
2: activity in the vmPFC was higher for the same contrast.
experienced pleasantness, we administered a follow-up behav-
ioral session 8 weeks after the main experiment, during which
wines were presented without price information. As expected, in
this case, there were no reported differences among the wines (see Materials and Methods and SI Text for details). We found
(Fig. 1E). Interestingly, while the pleasantness ratings were that the effect of price on mOFC activity was higher for the
increasing in price in the main experimental task (Fig. 1C), they cheap $5 wine than for the expensive $90 wine. This suggests that
were not in the postscanning blind test (Fig. 1D). A potential the effect of a price increase on mOFC activity might be larger
concern with these behavioral results is they might exhibit at low than at high prices.
‘‘experimenter demand’’ effects. In particular, some subjects
might deem it inappropriate to report to the experimenter that
a cheaper wine tastes better.

Modulation of Brain Correlates of Experienced Pleasantness by Price.


We analyzed brain imaging data using two different general linear
models [see Materials and Methods and supporting information (SI)
Text for details]. First, we looked for brain areas whose activity
increased with the price of wine. More concretely, we estimated the
BOLD response to each of the liquids at degustation and swallow-
ing and then analyzed the contrasts ‘‘high–low price’’ at degustation
separately for wines 1 and 2. Fig. 2 B and C describe the results of
this contrast for wine 1. We found increased activation in the left
mOFC and the left ventromedial prefrontal cortex (vmPFC).
Another cluster was found in a superior part of the vmPFC
adjoining the rostral anterior cingulate cortex (rACC). We also
found increased activation in the dorsolateral prefrontal cortex,
visual cortex, middle temporal gyrus, and cingulate gyrus (see SI
Table 2, upper). As shown in Fig. 2 E and F, the contrast generated
similar results for wine 2: increased activation was observed in the
bilateral mOFC, vmPFC, and rACC. In addition, for the wine 2, we
also found activation changes in the amygdala, lateral parts of the
OFC, dorsolateral prefrontal cortex, inferior and middle temporal
gyrus, and posterior cingulate cortex (see SI Table 3, upper).
A comparison of SI Tables 2, upper, and 3, lower, or of the
relevant figures, suggests there might be small differences on the
areas of the medial prefrontal cortex activated by the two wines. SI
Fig. 5 shows that, as the statistical threshold is lowered, these Fig. 3. The effect of price on both wines. (A) Conjunction analysis. Activity
in the mOFC/rACC was higher in the high- than in the low-price condition for
differences disappear. To investigate this further, we performed a
both wines 1 and 2. (B) Correlation of behavioral and BOLD responses (r " 0.49,
conjunction analysis to identify areas in which brain activity was P ! 0.001). Each point denotes an individual wine pair. The horizontal axis
higher on the high price condition for both wines. As shown in Fig. measures the change in reported pleasantness between the high- and low-
3A, bilateral mOFC and adjoining rACC exhibited this pattern. price conditions. The vertical axis computes an analogous measure using the
To investigate whether the increase in price had a differential betas from the general linear model in a 5-mm spherical volume surrounding
effect on the two wines, we performed an interaction analysis the area depicted in A.

Plassmann et al. PNAS ! January 22, 2008 ! vol. 105 ! no. 3 ! 1051
tions. Interestingly, an analogous mechanism has been proposed
for pain placebo effects (7).
Our results have implications for several disciplines. First, the
EP signal plays a central role in neuroeconomics, because it
serves as a teaching signal that guides future behavior. Unfor-
tunately, very little is known about the factors that affect the
neural computation of this signal. A natural starting hypothesis
is the economic view, which states that EP depends only on the
sensory properties of the item being consumed (i.e., its molec-
ular properties) and the state of the consumer. Our results
suggest that the brain might compute EP in a much more
sophisticated manner that involves integrating the actual sensory
properties of the substance being consumed with the expecta-
tions about how good it should be. It is important to emphasize
that it might be adaptive for the brain to do this. To make good
decisions in the future, the brain needs to carry out good
measurements of the quality of current experiences. In a world
of noisy measurements, the use of prior knowledge about the
quality of an experience provides additional valuable informa-
tion. A related study (13) provides additional supporting evi-
dence for this point by showing that giving a cognitive label to
an ambiguous odor (‘‘cheddar cheese’’ or ‘‘body odor’’) can
Fig. 4. Neural correlates of liking ratings. (A) Activity in the mOFC and the affect both subjective pleasantness reports and neural activity
midbrain correlated with the reported pleasantness of the six liquids at degus- related to EP. Unlike the current paper, however, de Araujo et
tation time. For illustration purposes, the contrast is shown both at P " 0.001 and al. (13) do not provide evidence that marketing actions, such as
P " 0.005 uncorrected and with an extend threshold of five voxels. (B) Correlation
pricing, can affect neural correlates of EP.
of pleasantness ratings and BOLD responses (r ! 0.593, P " 0.000). Each point
denotes a subject-price pair. The horizontal axis measures the reported pleasant-
Second, our findings also have implications for marketing.
ness. The vertical axis computes the betas from the general linear model in a Whereas there is ample behavioral evidence that various mar-
5-mm spherical volume surrounding the area depicted in A. keting actions are successful in influencing the EP of individuals,
that they can modulate neural representations of this signal had
not been reported before. Furthermore, the neural findings also
Discussion provide some clues about the mechanisms involved. In particu-
The main hypothesis of this study was that an increase in the lar, it seems that price changes modulate the representations of
perceived price of a wine should, through an increase in taste experienced utility but not the encoding of the sensory proper-
expectations, increase activity in the mOFC. The results de- ties of taste in the primary gustatory cortex.
scribed above provide evidence consistent with the hypothesis. Third, our results have implications for economics. EP is an
The hypothesis was motivated by several previous studies, which important component of experienced utility, which is the econ-
have shown that activity in the mOFC is correlated with behav- omist’s term for subjective well being. We show that, contrary to
ioral pleasantness ratings for odors (10–13), tastes (6, 14, and the standard economic view, EP depends on nonintrinsic prop-
15), and even music (16). This, together with our behavioral erties of products, such as the price at which they are sold. It then
results and the additional imaging results described below, follows that marketing manipulations might affect subjective
support the interpretation that, by modulating the activity in the perceptions of well being. This raises several difficult questions
for the field. Should the effect of prices on experienced utility be
mOFC, changes in the price of a wine might lead to a change in
counted as real economic well being or as a mistake made by
the actual EP derived from its consumption.
individuals? To what extent are measurable differences in pref-
We performed two additional analyses to provide further support
erences based on intrinsic differences between products and
for this interpretation. First, for each individual and wine, we
price effects we have identified? What happens to the efficiency
computed the change in reported EP between the high and low of competitive markets when firms can influence experienced
price conditions. We also computed the analogous difference in utility by changing the price of items?
parameter estimates for the BOLD response from the general An important task for future research is to develop a more
linear model in an area surrounding the mOFC. Fig. 3B shows that complete characterization of the range of marketing actions that
the neural and behavioral estimates were positively and highly can influence the neural computation of EP. We conjecture that
correlated (r ! 0.49, P " 0.001). Second, we verified that the results any action affecting expectations of product quality, such as
of the previous literature also held in our study by estimating a expert quality ratings; peer reviews; information about country
different general linear model and looking for brain regions whose of origin, store, and brand names (especially those associated
activity was correlated with reported EP from sampling the differ- with luxury products); and repeated exposure to advertisements
ent stimuli (see SI Text for details). The results replicated the might lead to effects similar to those identified here.
findings of previous studies: activity in the mOFC was correlated
with absolute reports of pleasantness (Fig. 4). Materials and Methods
Importantly, we did not find evidence for an effect of prices Subjects. Twenty normal-weight subjects participated in the experiment (11
on areas of the primary taste areas such as the insula cortex, the males, ages 21–30; mean age, 24.5 yr). One additional subject participated in
ventroposterior medial nucleus of the thalamus, or the prabra- the experiment but was excluded from the analysis, because he reported
chial nuclei of the pons. A natural interpretation is that the being confused about the task during a debriefing at the end of the experi-
ment. All subjects were right-handed and healthy; had normal or corrected-
top-down cognitive processes that encode the flavor expectan-
to-normal vision and no history of alcohol abuse, psychiatric diagnoses, or
cies are integrated with the bottom-up sensory components of neurological or metabolic illnesses; and were not taking any medications that
the wine in the mOFC, thus modulating the hedonic experience interfere with the performance of fMRI. All subjects were screened for liking,
of flavor, but that the flavor expectancies generated by the and at least occasionally drinking, red wine. At the beginning of each exper-
change in prices do not impact more basic sensory representa- iment, subjects were required to show an official form of identification to

1052 ! www.pnas.org"cgi"doi"10.1073"pnas.0706929105 Plassmann et al.


provide evidence they were !21 yr old. Subjects were informed about the sequence designed to optimize functional sensitivity in orbitofrontal cortex
experiment and gave written consent before participating. California Insti- (18). This consisted of a tiled acquisition in an oblique orientation at 30° to the
tute of Technology’s institutional review board approved the study. AC-PC line. In addition, we used an eight-channel phased array coil that yields
a 40% signal increase in OFC over the standard head coil. The sequence
Stimuli. During the course of the fMRI experiment, subjects sampled three enabled 32 axial slices of 3-mm thickness and 3 mm in-plane resolution that
different Cabernet Sauvignon wines and an affectively neutral tasteless con- could be acquired with a TR of 2 s. A T1-weighted structural image was also
trol solution that consisted of the main ionic components of human saliva (25 acquired for each subject. Functional imaging data were acquired in four
mM KCl and 2.5 mM NaHCO3). The wines were administered in random order separate sessions of "13 min each.
and simultaneously with the appearance of a price identifier. Two of the three To detect transient head movements due to swallowing, we attached a
wines were administered twice, once identified by their actual retail price and 1.5-cm-long copper coil with a radius of 0.5 cm to the neck of each subject. The
once by a 900% markup (wine 1: $5 real retail price, $45 fictitious price) or a setup was similar to those used by previous studies in which liquid food had been
900% reduction (wine 2: $90 real retail price, $10 fictitious price). The third administered (14). Small movements of the coil induced a current in the magnetic
wine was used as a distracter and was identified by its retail price (wine 3: $35). field that could be detected after amplification using an EEG system positioned
We also carried out a follow-up behavioral tasting session 8 weeks after the in the scanner room (Biopac Systems). This produced a time series of events
main experiment. In this postexperimental session, the wines were presented reflecting transient larynx movement that was used in the general linear model
without price information (see SI Table 1 for a complete description). (GLM) in two ways. First, the time series of the signal detected by the coil was
In each trial, 1 ml of each stimulus was delivered by a system of electronic added as an additional motion regressor in the GLM as regressor of no interest;
syringe pumps (one for each stimulus) positioned in the scanning control this was done to take out the variance due to swallowing induced head move-
room. These pumps transferred the stimuli to the subjects via "10-m polyeth- ment. Second, swallowing events were either assigned to experimental condi-
tions or classified as noninstructed swallowing. The former were used to correct
ylene plastic tubes (6.4-mm diameter) and a perfusion manifold. The perfusion
the timing of swallowing onsets for each stimulus type that were entered in the
manifold allowed six incoming tubes to be connected to one output tube with
GLM as a regressor of interest (see SI Text for details about the models). The latter
a minimum of dead space to avoid the mixing of the wines. The subjects were
were entered into the data analysis as a regressor of no interest.
instructed to hold the output tube between their lips like a straw while they
fMRI data analysis was performed by using the Statistical Parametric Map-
lay in a supine position in the scanner. We made an effort to keep the tubes
ping software (SPM05; Wellcome Department of Imaging Neuroscience). We
free of air bubbles to avoid wine oxygenation. Between experiments, the
applied the following preprocessing steps to the imagining data: (i) slice-
wines were preserved with a special corking system, which is typically used by

NEUROSCIENCE
timing correction (centered at TR/2) (ii) realignment to the last volume, (iii)
wineries and wine shops to prevent oxidation.
spatial normalization to a standard T2* template with a resampled voxel size
The stimulus presentation and response recording was controlled by Co-
of 3 mm3, (iv) spatial smoothing using a Gaussian kernel with full width at half
gent 2000 (Wellcome Department of Imaging Neuroscience) installed on a
maximum of 8 mm, and (5) intensity normalization and high-pass temporal
computer positioned in the control room that also received trigger pulses
filtering (filter width 128 s). The structural T1 images were coregistered to the
from the scanner to control timing. The visual stimuli were presented by using
mean functional EPI images for each subject and normalized using parameters
video goggles (Resonance Technologies).
derived from the EPI images.

Task. The task consisted of six trial types: $5, $10, $35, $45, and $90 wines and
fMRI Data Analysis 1: Influence of Price on Wine Sampling. We estimated a
a neutral liquid. Each trial type occurred 16 times during the course of the general linear model in which the delivery of each of the six different liquids
experiment, resulting on a total of 96 trials. Subjects were instructed to sample was entered as a regressor of interest. In addition, the swallows of each liquid
the liquid on each trial while it was on their mouth (for a period of 10 s), to type were also entered as separate regressors. Each of these regressors plus
evaluate its pleasantness during this time, and to swallow only when in- additional regressors of no interests were convolved with a canonical hemo-
structed. Between every wine administration, there was a rinse period in dynamic response function. We then calculated first-level single-subject con-
which the neutral solution was delivered. The rinse period was implemented trasts to compare the administration of an identical wine at a high minus a low
to avoid taste spillovers across trials. Trials were separated by a random price. Finally, for each of these first-level contrasts, we calculated a second-
intertrial interval drawn from a Poisson distribution with a mean of 10 s (see level group contrast using a one-sample t test (see SI Text for details).
Fig. 1 A for a detailed description of the timing of each trial). We also performed an interaction analysis to compare the effect of prices
In every other trial, subjects were instructed to enter a rating of either on the two wines. We calculated a first-level single-subject contrast to com-
flavor pleasantness or taste intensity. Thus, a total of four pleasantness ratings pare the administration of the low-quality wine at a high minus a low price
and four taste intensity ratings were sampled for each liquid. We used a minus the administration of the high quality wine at a high minus a low price.
six-point rating scale (1 # do not like it at all/not intense at all; 6 # like it very We also calculated a first-level single-subject contrast to compare the admin-
much/very intense) (see Fig. 1B). The timing of rating trials was identical to istration of the high-quality wine at a high minus a low price minus the
nonrating trials, except that, after swallowing the liquid and before rinsing, administration of the low-quality wine at a high minus a low price. For each
subjects were given 6 s to enter their ratings. of these first-level contrasts, we calculated a second-level group contrast using
After an initial instruction period, subjects were trained on the use of the a one-sample t test (see SI Text for details).
response boxes to enter the ratings. Given the large number of potential
ratings, subjects entered ratings with both hands (three ratings for each fMRI Data Analysis 2: Neural Representation of Experienced Pleasantness. We
hand). The assignment of ratings to buttons was counterbalanced across estimated a second general linear model in which the delivery and swallow of
subjects to avoid motor artifacts. all liquids (independently of type) were entered as regressors. In addition, we
Unbeknown to the subjects, the critical manipulation was that the $5 and $45 entered the reported pleasantness and intensity ratings as parametric mod-
wines and the $10 and $90 wines were identical. This manipulation was not ulators for both regressors. We then computed first-level single-subject con-
revealed to the subjects. Instead, the subjects were told they would be sampling trasts for the parametric modulators at both liquid sampling and swallowing
five different Cabernet Sauvignons, that the purpose of the experiment was to by reported EP and taste-intensity ratings. Finally, for each of these first-level
study the effect of sampling time on perceived flavors, and that the different contrasts, we calculated a second-level group contrast using a one-sample t
wines will be identified by their retail prices. Evidence for the success of our cover test (see SI Text for details).
story was that all subjects reported at the end of the experiment being able to
taste five different wines. Although the experiment contains an element of
deception, subjects were not debriefed after the experiment to avoid contami- ACKNOWLEDGMENTS. We thank Vivian Valentin, Jan Glaescher, and Axel
nating California Institute of Technology’s small subject pool. Linder for their help with this project and Hackjin Kim, Sam Huang, and Shawn
Wagner for developing the coil we used to detect swallowing movement and for
providing support to process the swallowing signal. Financial support from the
Data Acquisition and Preprocessing. The brain imaging was conducted in a 3-T National Science Foundation (Grant SES-0134618) is gratefully acknowledged.
Siemens Trio MRI scanner (Siemens). We acquired gradient echo T$2 This work was also supported by a grant from the Gordon and Betty Moore
weighted echoplanar images (EPI) with BOLD contrast and used a special Foundation to the California Institute of Technology Brain Imaging Center.

1. Kahneman D, Wakker PP, Sarin R (1997) Back to Bentham? Explorations of Experienced 3. Allison RI, Uhl KP (1964) Influence of Beer Brand Identification on Taste Perception. J
Utility. Q J Econ 112:375– 405. Market Res 1:36 –39.
2. Lee L, Frederick S, Ariely D (2006) Try it, you’ll like it: The influence of expectation, 4. Klaaren KJ, Hodges SD, Wilson TD (1994) The role of affective expectations in subjective
consumption, and revelation on preferences for beer. Psychol Sci 17:1054 –1058. experience and decision-making. Soc Cognit 12:77–101.

Plassmann et al. PNAS ! January 22, 2008 ! vol. 105 ! no. 3 ! 1053
5. Shiv B, Carmon Z, Ariely D (2005) Placebo Effects of Marketing Actions: Consumers May 13. de Araujo IE, Rolls ET, Velazco MI, Margot C, Cayeux I (2005) Cognitive modulation of
Get What They Pay For. J Market Res XLII:383–393. olfactory processing. Neuron 46:671– 679.
6. McClure SM, et al. (2004) Neural correlates of behavioral preference for culturally 14. Kringelbach ML, O’Doherty J, Rolls ET, Andrews C (2003) Activation of the human
familiar drinks. Neuron 44:379 –387. orbitofrontal cortex to a liquid food stimulus is correlated with its subjective pleas-
7. Wager T (2005) The neural bases of placebo effects in pain. Curr Dir Psychol Sci 14:175–179. antness. Cereb Cortex 13:1064 –1071.
8. Wilson TD, Klaaren KJ (1992) in Review of Personality and Social Psychology, ed Clark 15. Small DM, et al. (2003) Dissociation of neural representation of intensity and affective
MS (Sage, Newbury Park, CA), Vol 14, pp 1–31. valuation in human gustation. Neuron 39:701–711.
9. Rao A, Monroe KB (1989) The effect of price, brand name, and store name on buyers’
16. Blood AJ, Zatorre RJ (2001) Intensely pleasurable responses to music correlate with
perceptions of product quality: An integrative review. J Market Res 26:351–357.
activity in brain regions implicated in reward and emotion. Proc Natl Acad Sci USA
10. de Araujo IE, Kringelbach ML, Rolls ET, McGlone F (2003) Human cortical responses to
98:11818 –11823.
water in the mouth, and the effects of thirst. J Neurophysiol 90:1865–1876.
11. Anderson AK, et al. (2003) Dissociated neural representations of intensity and valence 17. Francis S, et al. (1999) The representation of pleasant touch in the brain
in human olfaction. Nat Neurosci 6:196 –202. and its relationship with taste and olfactory areas. NeuroReport 10:453–
12. de Araujo IE, Rolls ET, Kringelbach ML, McGlone F, Phillips N (2003) Taste-olfactory 459.
convergence, and the representation of the pleasantness of flavour, in the human brain. 18. Deichmann R, Gottfried JA, Hutton C, Turner R (2003) Optimized EPI for fMRI studies
Eur J Neurosci 18:2059 –2068. of the orbitofrontal cortex. NeuroImage 19:430 – 441.

1054 ! www.pnas.org"cgi"doi"10.1073"pnas.0706929105 Plassmann et al.


Brain Research Bulletin 67 (2005) 428–437

Arousal and consumer in-store behavior


Andrea Groeppel-Klein ∗
Chair of International Marketing, Consumer Behavior, and Retailing at the European University Viadrina,
Grosse Scharrnstrasse 59, 15230 Frankfurt, Oder, Germany

Available online 18 July 2005

Abstract

From a psychophysiological point of view, arousal is a fundamental feature of behavior. As reported in different empirical studies based
on insights from theories of consumer behavior, store atmosphere should evoke phasic arousal reactions to attract consumers. Most of these
empirical investigations used verbal scales to measure consumers’ perceived phasic arousal at the point-of-sale (POS). However, the validity
of verbal arousal measurement is questioned; self-reporting methods only allow a time-lagged measurement. Furthermore, the selection of
inappropriate items to represent perceived arousal is criticized, and verbal reports require some form of cognitive evaluation of perceived
arousal by the individual, who might (in a non-measurement condition) not even be aware of the arousal. By contrast, phasic electrodermal
reaction (EDR) has proven to be the most appropriate and valid indicator for measuring arousal [W. Boucsein, Physiologische Grundlagen und
Meßmethoden der dermalen Aktivität. In: F. Rösler (Ed.), Enzyklopädie der Psychologie, Bereich Psychophysiologie, Band 1: Grundlagen
and Methoden der Psychophysiologie, Kapitel, Vol. 7, Hogrefe, Göttingen, 2001, pp. 551–623] that could be relevant to behavior. EDR can be
recorded simultaneously to the perception of stimuli. Furthermore, telemetric online device can be used, which enables physiological arousal
measurement while participants can move freely through the store and perform the assigned task in the experiments. The present paper delivers
insights on arousal theory and results from empirical studies using EDR to measure arousal at the POS.
© 2005 Elsevier Inc. All rights reserved.

Keywords: Electrodermal reactions; Buying behavior; In-store stimuli

1. Introduction electrodermal activity (EDA) is considered to be a valid and


also very sensitive indicator that responds clearly to the small-
As reported in different empirical studies based on insights est variation in arousal. Latest technical developments offer
from theories of consumer decision-making and on insights larger storage capacity enabling telemetric measurement of
from environmental psychology, store atmosphere must electrodermal activity, which in turn makes field experiments
evoke phasic arousal reactions to attract consumers. Most of in retail stores possible. The present paper delivers insights
these empirical investigations used verbal scales to measure on arousal theory and presents empirical results from studies
consumers’ perceived arousal at the point-of-sale. However, using EDA to measure arousal at the point-of-sale.
the validity of verbal arousal measurement can be questioned.
On the one hand, traditional interviews are mostly done
after the shopping trip through a store or a particular part
of the store so that customers have to remember their per- 2. A psychophysiological perspective: insights from
ceived arousal and emotions. Thus, self-reporting methods arousal theory
only allow a time-lagged measurement. On the other hand,
the preferential use of verbal scales to measure psychophys- From a psychophysiological point of view, arousal is a
iological responses is discussed controversially. By contrast, fundamental feature of behavior. It can be defined as the
neurophysiological basis underlying all processes in the
human organism. Thus, arousal is the basis of emotions,
∗ Tel.: +49 335 5534 2870; fax: +49 335 5534 2275. motivation, information processing, and behavioral reactions
E-mail address: groeppel@euv-frankfurt-o.de. [2,3,10,27,32]). Arousal can vary from deep sleep through

0361-9230/$ – see front matter © 2005 Elsevier Inc. All rights reserved.
doi:10.1016/j.brainresbull.2005.06.012
A. Groeppel-Klein / Brain Research Bulletin 67 (2005) 428–437 429

moderate stages of increase up to panic. Basically, a distinc- more complex arousal theory [5,6,16,28]. Boucsein [6], for
tion can be made between tonic and phasic arousal. Tonic example, presented a three-dimensional theoretical frame-
arousal refers to a relatively long-term state of conscious- work based on neurophysiological insights on arousal and
ness that changes slowly due to long-lasting or extremely information processing (see Fig. 1). This model has been con-
intensive stimuli. Phasic arousal arises in response to spe- firmed by a PET-study by Fredrikson et al. [19]. The model
cific stimuli, resulting in short-term variations in the arousal also considers the dependency between arousal and emo-
level. It indicates a ready state of the body for reaction. It is tion/motivation as well as the effects on central and peripheral
closely related to attention, i.e. enhanced sensitivity of the psychophysiological parameters. The first dimension of this
organism to relevant stimuli and stimuli processing, while framework is in accordance with the concept of the previous
irrelevant stimuli are filtered and not processed. Both inter- unidimensional approach. That means that the RF forms the
nal (thought and metabolic processes) and external stimuli physiological basis of this dimension. EEG can serve as an
can cause arousal [6,7]. Phasic arousal might be the driving indicator. Results from the first dimension are perceived as
force for decision-making processes and approach behavior general activation with a vigilant feeling and an alert state
(i.e. time and money spent in a store) at the point-of-sale (see of mind. The second dimension, the “affect–arousal” sys-
Section 3). tem, comprises primarily emotional components of arousal.
Early unidimensional concepts of arousal theory describe Headed by the Amygdala, attention, orienting reflex, and
the reticular activating system (RAS), which comprises the overall behavior is enhanced via hypothalamic reactions. The
sensoric inflow, the reticular formation (RF), as well as corti- physiological outcome results in phasic cardiovascular (heart
cal, hypothalamic, and thalamic areas, as the fundamental rate) and/or tonic electrodermal variations. When it comes to
component for the formation of arousal [5,6]. The RF is behavior and perception, processes of this second dimension
a complex network of fibers and cell bodies in the core lead to defense reactions and negative emotions. Attention
of the brain stem. It is involved in the filtering process of that turns into an orienting reflex may also directly impinge
sensory information from the central nervous system (e.g. on the third system. This so-called “preparatory activation”
from visual, haptic, and acoustic stimulation). According to system is basically encompassing motivational aspects of
these early theoretical approaches in arousal research [13,29], arousal. Expectations are transformed into a ready state for
all sensory and motor nerve fibers should enhance general reaction. This part of the system interacts, especially, with
arousal of the reticular formation via collaterals. In turn, motor and pre-motor activation of behavior, and with posi-
the RF should indistinctly activate large parts of the central tive emotion. Phasic electrodermal amplitude, for example,
nervous system [6]. Thus, the unidimensional approach sug- can serve as an indicator.
gested a correlation between diverse physiological outcomes For marketing purposes, and especially, when testing the
of arousal, such as heart rate, blood pressure, electroen- impact of in-store stimuli on customers’ arousal and buying
cephalogram (EEG), or electrodermal activity. behavior, the “preparatory activation” system, and in part,
Since empirical findings were not consistent with this the “affect–arousal” system (here, especially, attention and
unidimensional approach, recent efforts yielded towards a orienting reflex), are of major relevance [23].

Fig. 1. Simplified own illustration of the three-dimensional arousal model of Boucsein and Aktivierung [6].
430 A. Groeppel-Klein / Brain Research Bulletin 67 (2005) 428–437

3. Arousal and in-store behavior On the one hand, of course, appealing store designs can
evoke positive feelings at the POS for both, buyers and non-
For more than 20 years, the empirical investigation of buyers. Since we regard arousal as the physiological basis of
consumers’ responses to store atmosphere and visual mer- emotions, we can reasonably conclude that a positive store
chandising concepts has been an interesting research area atmosphere will also increase consumer arousal indepen-
and it has now become widely accepted that an appealing dently of any purchase. On the other hand, we know that
store atmosphere has a major impact on consumers’ in-store highly appealing stores designs also provoke more purchases
behavior and store assessment [11,18,20–22,24,30,36,38]. (e.g. [11,20]). This leads to the notion that “hedonic buyers”
In summary, the results clearly confirm that a store atmo- will reveal even higher arousal than “hedonic non-buying
sphere that evokes pleasure, can increase consumers’ spend- browsers”. In addition, according to the previous research
ing, time spent in the store, and the desire to come back and to cited earlier, making a real decision implies more effort
recommend the store to others. However, the role of perceived than leaving it, which is another reason for higher phasic
arousal at the point-of-sale is discussed controversially. Some arousal reactions of buyers. During economic downturns,
of the cited studies show a significant positive influence of many retailers complain about the tendency of consumers
an arousal stimulating-atmosphere on store-assessment and to avoid stores and often to refrain altogether from buying.
in-store behavior (e.g. [11,23]), others do not support these Due to this “consumption lethargy”, non-buyers might be less
findings or even show a converse relationship (e.g. [12]). activated irrespective of the store ambience. This argument
One reason for these contradictory results may be that the lends further support to the thesis that on average, buyers are
studies used different methods to measure arousal, which more aroused than non-buyers. Thus, this paper focuses on
questions the validity of the results (see Section 4). A second two research questions:
reason could be that not all investigated environmental set-
tings really offered arousal-increasing stimuli. Three groups (1) Do different visual merchandising concepts evoke dif-
of arousal inducing stimuli can be distinguished [27]: (1) ferent phasic arousal responses of consumers, and can
affective stimuli evoke pleasant or unpleasant emotions due arousal be related to emotions, and in-store behavior?
to innate stimulus–response mechanisms, or preceding condi- (2) Are buyers more aroused at the point-of-sale than non-
tioning. Especially, key-stimuli, such as little children, erotic buyers, and can different arousal patterns be detected?
stimuli, and symbols form nature increase the arousal appeal-
ing mannequins, scents, warm colors or plants are possible
affective in-store stimuli. (2) Intense stimuli produce an effect 4. Measurement of arousal
because of their physical properties. As salient information,
such stimuli automatically engender orienting responses. At Arousal is definitely considered as being fundamental to
the POS intense stimuli could be flashy colors, spotlights, behavior that is related to emotions and behavior [3,8] and
visually, striking price-tags, or some retailers use huge flat as previously shown, arousal plays a major role for cus-
screens presenting movies, music, etc. (3) Collative stimuli tomers’ interaction with the store environment. Yet, despite
are novel, startling, or different stimuli arrangements (e.g. diverse methods applied in order to measure arousal, the vari-
vivid-looking mannequins or unexpected decorations). able is extremely difficult to indicate validly. Among the
Furthermore, little contribution can be found in the liter- applied methods are verbal scales like the PAD scale [33],
ature neither on the impact of store atmosphere on buying non-verbal scales (e.g. color–pattern scale [4,31], and self-
decisions nor on the impact of arousal on decision processes reporting methods.
evoked by the store atmosphere. Therefore, in our studies, The color and pattern scale contains more or less intense
we do not only analyze the influence of store atmosphere colors or complex patterns. Test persons have to select from
and visual merchandising concepts on arousal perceived at the scale one color or a pattern that reflects their subjec-
the point-of-sale (e.g. [23]), but we are also interested in tively perceived arousal level. Meyer-Hentschel [31] assumes
comparing phasic arousal reactions of buyers with those of that excited consumers choose instinctively different col-
non-buyers [25]. ors or patterns than less aroused test persons. However,
The relevance of taking non-buyers into consideration, the color–pattern scale has not yet been implemented often
results from diverse studies reporting that shopping trips can enough to support valid conclusions. Furthermore, irrespec-
have a hedonic value irrespective of whether or not something tive of their arousal reaction, test persons may have an
is purchased [1,23,26] and that consumer behavior does not enduring preference for special colors or patterns. Responses
always satisfy functional or economic needs [17,34]. Accord- from self-reporting methods often suffer from biases [40].
ing to Babin et al. [1], hedonic shopping results “more from Another way to investigate arousal at the point-of-sale – but
fun and playfulness than from task completion (. . .)” and in fact, rather measuring emotion – is the observation of non-
“reflects shopping’s potential entertainment and emotional verbal communication, like orienting reflexes, or especially,
worth (. . .); increased arousal, heightened involvement, per- facial features, e.g. via facial affect scoring technique (FAST)
ceived freedom, fantasy fulfillment, and escapism may all [15] or facial acting coding system (FACS) [14]. Today,
indicate a hedonically valuable shopping experience”. also computer-aided face-recognition methods are used, e.g.
A. Groeppel-Klein / Brain Research Bulletin 67 (2005) 428–437 431

Zlochower and co-workers [9,37]. Furthermore – as men- dermal reactivity is the most crucial indicator for research
tioned before – a variety of different verbal methods has concerning arousal at the point-of-sale.
been used to measure perceived arousal at the POS. How- Basically, EDA can be measured via endosomatic or exo-
ever, the results derived from these scales are controversial, somatic recording. While endosomatic recording does not
and their discriminant validity has been criticized [38,40]. use any external voltage to measure potential differences on
Basically, there are three reasons for these criticisms: (1) the skin, exosomatic recording uses either direct current (dc)
selection of appropriate items to represent perceived arousal; or alternating current (ac). Exosomatic recording with direct
(2) verbal reports require some form of cognitive evaluation current is most frequently used to measure either skin con-
of perceived arousal by the individual, who might (in a non- ductance (SC), measured in microsiemens (!S). Siemens is
measurement condition) not even be aware of the arousal; (3) the unit to measure conductance between two objects or skin
the fact that verbal measures are taken some time later than resistance (SR), measured in kiloohm (k!) (ohm is the unit to
the actual experience of the situation [23]. measure resistance between two objects). Each of these meth-
In contrast to these methods, psychophysiological mea- ods measures both, skin conductance/resistance level (tonic
sures, such as heart rate, EEG, and EDA, or studies with arousal) as well as skin conductance/resistance reactivity
brain-imaging technologies (e.g. fMRI or PET) offer most (phasic arousal) from the EDA-signal [5]. For practicabil-
valid indicators. However, at present neither EEG nor brain- ity and safety reasons for test subjects, exosomatic recording
imaging technology can be used in field experiments. must strongly be recommended for in-store investigations
Electrodermal activity in general “is regarded as a sensi- since endosomatic recording requires one of the electrodes
tive and valid indicator for the lower arousal range, reflecting being placed in an “inactive” site on the skin that has been
small, mostly cognitively conditioned variations in arousal” pretreated (so-called “skin-drilling”). For exosomatic mea-
[5]. Willingly influencing test results is almost impossible. surement, both electrodes are simply attached to the left palm
Contrary to the heart rate, the electrodermal reactivity (EDR) of right-handed test persons, or vice versa for left-handers.
points to even the very smallest psychological change [5]. It
is thus delivering the most sensitive indicator of arousal that
might potentially be relevant to behavior. Furthermore, as we 6. Measurement of EDR at the point-of-sale
will see later, EDR can also be measured at the point-of-sale.
EDR can, therefore, deliver a valuable contribution to fur- In order to investigate the research questions arousal was
ther research on consumers’ shopping behavior induced by measured by recording EDR, while the test subjects per-
arousal. formed the assigned task in the experiments. In all studies,
a telemetric online device was used, which enables physi-
ological arousal measurement while participants can move
5. Electrodermal activity as psychophysiological freely through the store. Rapid movements, pressure on the
indicator for arousal electrodes, change in intimacy between electrodes and skin,
as well as speech activity, may lead to artifacts resulting
The human skin basically consists of two layers, the der- in specific patterns in the EDR curve. Such artifacts need
mis, and the epidermis. The epidermis is located at the surface to be excluded from the data [5,23]. However, due to new
and consists of epithelial tissue. This layer is more callous, EDR-technology, “normal walking” does not lead to artifacts,
the closer to the surface. The comparatively thicker and therefore, we can use this method in field experiments. For
deeper lying dermis consists of taut, fibrous connective tis- registration, an exosomatic approach was chosen applying
sue. Underneath the dermis, lies the subcutis, which contains dc (0.4 V) and measuring skin conductance. The technical
the secretory part of the sweat glands, fatty tissue, and vessels equipment runs with a 12-bit analog-to-digital (A/D) con-
that supply the body surface [5]. verter. Two Ag/AgCl electrodes were used and filled with a
The epidermis is of great importance to EDA. Although 0.5% NaCl electrode creme.
it becomes dryer towards its outside layer as the regularly The following EDR parameters are of particular relevance
arranged cells become less tightly packed, there is a perma- to all experiments. The amplitude describes the strength of
nent insensible perspiration from the dermis via the epidermis each phasic arousal reaction [5]. In order to calculate ampli-
even while no sweat gland activity occurs. This hydration tudes, for all studies, we chose a minimum amplitude criterion
depends on external and internal factors and leads to good of 0.05 !S so as to exclude recording artifacts from the signal-
electric conductivity of the skin, thus, making it possible to to-voice ratio [5,39]. While registering EDR, overlapping
measure it by means of two electrodes attached to the skin. amplitudes may occur, i.e. a second amplitude follows a first
The conductivity of the skin is then transmitted to the com- one, although the original baseline has not yet been reached.
puter by means of an amplifier. The conductivity of the skin is For our data analysis, overlapping amplitudes were evaluated,
primarily responsible for tonic EDA, while active membrane each by means of its own baseline, regardless of the recovery
processes following upon a nerve impulse bring about phasic time of the preceding amplitude [5]. According to Steiger
EDA which turns into electrodermal reactivity [5]. Following [35], the intensity of perceived arousal over a certain period
the three-dimensional model of Boucsein [6], phasic electro- of time, can be received by summing up all single ampli-
432 A. Groeppel-Klein / Brain Research Bulletin 67 (2005) 428–437

tudes so as to obtain the total amplitude which is the most Table 2


important phasic arousal parameter in field experiments. The Correlation between emotion “joy” and arousal
second important parameter is the so-called “frequency” of Emotion Total amplitude Frequency
responses, which subsumes up all single reactions. Each skin Joy (n = 27)
conductance response demonstrates the particular attention Pearson correlation .464 (.015) .420 (.029)
of the individual towards an object in its environment [35]. (two-tailed significance)

subject’s pocket. The data were transmitted via the telemetric


7. Empirical studies device to the computer that was installed at the end of the fruit
and vegetable department for online registration. From that
7.1. Studies 1 and 2 part of the store, we also observed the participants. In both
stores, the fruit and vegetable section is the first part of the
The first empirical study investigates the influence of dif- store the test subjects walked through. Before they went on to
ferent visual merchandising concepts on arousal, emotions, another part of the store, electrodes were detached, and reg-
and buying behavior (for details, see Ref. [23]). The investi- istration stopped. Each individual was then asked to answer a
gation was conducted in the fruit and vegetable department of standardized questionnaire. The questionnaire was the same
two Austrian grocery stores. Both stores belong to the same for both stores.
retail chain and are located in a distance of a 10 min walk In order to gain valid parameters, we first excluded all arti-
from another. The management of the retail chain pursues facts from the recorded curves. Furthermore, the first part of
different marketing strategies with both stores. For “Store the curves were not rated up to the point when the test subjects
1” (experimental store), insights from environmental psy- were used to the electrodes and their curve showed a consis-
chology have been considered. Here, fruits and vegetables tent level (individual baseline = 0). When summing up the
are presented in large-scale on broad and deep carriers and amplitudes of the test persons generated over their controlled
are arranged according to colors. Exotic fruits, fresh herbs, shopping time, we found significant difference in total ampli-
and flowers are used as eye-catchers (“affective stimuli”). tudes for the two stores by means of a Mann–Whitney U-test
Parts of the products are presented on sort of an “island” in (see Table 1). A Mann–Whitney U was chosen to replace
the middle. The whole store is spacious and high-ceilinged, independent samples T-tests since small samples do not sat-
and is brightly illuminated. In “Store 2” (control store), the isfy the normal distribution assumption. This means that the
same assortment of fruits and vegetables is offered. Qual- experimental store evokes higher arousal than the control
ity and prices are exactly identical. However, insights from store. Furthermore, data show significantly higher frequency
environmental psychology have not been considered for the in the response rate of test persons in the experimental store
store decoration. The light appears less bright. Shelves are compared to those in the control store. This indicates that the
put alongside the wall, and display tables are arranged in a environment of the experimental store provides more infor-
row in the middle of a long and narrow aisle. Contrary to mation that has the potential to attract enough attention to be
“Store 1”, an extraordinary product presentation (for exam- cognitively processed.
ple, by an arrangement of the products according to colors) EDA measures arousal, however, the perceived emotion
was not applied. (subjectively interpreted arousal) that is derived from arousal
It can be assumed that “Store 1” which is decorated accord- cannot be indicated from any EDA parameter. Therefore, also
ing to the principles of environmental psychology would verbal scales were used. Both, the total amplitude and the
evoke higher arousal than the ordinary “Store 2”. In the exper- frequency show significant correlation with the emotion “joy”
imental store, 15 test persons were asked to do their usual (Table 2).
shopping in the vegetable department, while EDR was regis- Furthermore, in the experimental store (85.5%), signifi-
tered. In the control store, we collected EDR data from 12 test cantly (p = 0.000), more fruits and vegetables were bought
persons. Test customers were asked to participate right after than in control store (46.6%).
entering the store. Both, the EDR apparel and the telemetric Study 2 (replication study for non-food products) was con-
device, were either laid in the shopping cart, or put in the test ducted in a shopping mall, where we set up a bookstand in

Table 1
Arousal reactions: experiment 1
Items Store Mean (for illustration) N Mean rank Mann–Whitney Monte Carlo significance
Total amplitude Experimental 12.753 !S 15 17.07 44.000 .024
Control 6.752 !S 12 10.17
Frequency Experimental 84.73 15 17.20 42.000 .016
Control 47.25 12 10.00
A. Groeppel-Klein / Brain Research Bulletin 67 (2005) 428–437 433

front of the shop window of a bookstore. Different versions


of the bookstand were designed. The first version (C) simply
has a long table with many books piled on it, on topics about
foreign countries, vacation trips and books for vacation read-
ing. In the second group, the table was decorated with sand
and shells, a straw hat, beach towels, and a vivid-looking,
attractive mannequin in a deckchair (affective stimuli A). The
third variant (A and I) provided the same stimuli as the sec-
ond group, but here, also posters signaling a bargain in flashy
colors were added, in an attempt to attract consumers with
extremely low prices. The bargain posters belong to the cate-
gory of intense stimuli. A “marker” was set to the EDR curve
when the test subject reached the bookstand and another one
when he or she left. Later on, we analyzed only that part
of the EDR curve, which had been recorded in between the
two markers. Between the groups, there were no significant
differences in the socio-demographic data.
The results show that both parameters, total amplitude and
frequency, are highest in the group in which test persons were
exposed to all stimuli (A and I), whereas the parameters are
lowest in the control group (C), in which only books were
presented. Behavioral reactions were tested by a “coupon-
question”. Here, test persons were asked to imagine that they
received a coupon worth 100D , which they could cash in
any bookstore they liked. The subjects were then asked to
name the share of the coupon value they would be willing to Fig. 2. Test person in the mall experiment.
spend in the specific book store, where we conducted the
study. Again, we found significant differences (p < 0.005)
between the control group on the hand (Mc = 64D ), the group longer felt like shopping or when they had decided to choose
with affective stimuli (MA = 79D ), and the special-offer group a product. Then, electrodes were immediately detached, and
(MAI = 88D ). The design of the bookstand was evaluated best test persons were asked to report their product choice. In
in the 2nd group (A); price image was best in the 3rd group both studies, time markers were set when the participants
(A and I). started and ended the shopping trip, so as to detect the exact
EDR-registration time. After the exclusion of artifacts (e.g.
7.2. Studies 3 and 4 pressure on the electrodes, for details, see Boucsein [5]), we
obtained 82 valid EDA data sets for Study 3 (mall) and 105
Studies 3 and 4 were conducted in 2002, Study 3 on the first for Study 4 (department store).
floor of a local, medium-sized German shopping mall. Study In both studies, respondents were asked whether or not
4 took place in a large department store for sports goods. For they decided to buy something in order to identify “buyers”
both groups, randomly chosen customers entering the mall or and “non-buyers”. The products they decided to buy were
the store were asked to participate in our study. Respondents noted on the questionnaire. In Study 4, we also controlled
of all age, income, and education classes were involved. The whether the selected product had really been bought sub-
electrodes were attached to one hand to register the EDR. sequent to the experiment. This is an important difference
Both the telemetric and registering devices were conveniently between the two experimental settings. In the mall-study,
packed in a shoulder bag so that the respondents could walk we were only able to establish test persons’ intention to buy
through the retail environment unhindered by the technical something, since, on the first-floor of the mall with its more
equipment (see Fig. 2). than 25 different shops, 4 main entrances, and crowded floors,
For a start, participants were asked some “ice-breaking etc., it would have been impossible to observe test persons’
questions”. In the meantime, they got used to the electrodes shopping behavior after the interview. In contrast to the sit-
and the individual baseline of the EDR curve reached “zero” uation in the mall, in the third study, the cashier zone, and
so that subsequently, all phasic arousal reactions could be thus, actual purchases of participants could be observed eas-
measured from the same starting-point, thus, rendering EDR ily and without problems. Consequently, only in our second
curves of different participants comparable. Next, the test study, “real” buying decisions were registered. Furthermore,
persons were instructed to walk up and down the mall or the in the third study, some participants mentioned products that
department store as if they were on a normal shopping trip. could not be bought in shops on the first floor of the mall
They were asked to raise their right arm to signal when they no (like computers or washing machines). We assumed these
434 A. Groeppel-Klein / Brain Research Bulletin 67 (2005) 428–437

Fig. 3. Typical arousal curves.


A. Groeppel-Klein / Brain Research Bulletin 67 (2005) 428–437 435

Table 3
Phasic arousal of “buyers” and “non-buyers”
EDA parameters Group Mean Mean rank n Non-parametric or parametric test Significance
Study 3 Chi-square; n < 30 Kruskal–Wallis
Total amplitude Non-buyers (mall) 698.59 !S 27.52 44 4.795 .029
Buyers (mall) 1036.73 !S 38.69 16
Frequency Non-buyers (mall) 971.64 28.26 44 2.711 n.s.
Buyers (mall) 1237.50 36.66 16
Study 4 F-value (n > 30) ANOVA
Total amplitude Non-buyers (department store) 614.75 !S – 70 9.522 .003
Buyers (department store) 1136.39 !S – 35
Frequency Non-buyers (department store) 1040.21 – 70 18.938 .000
Buyers (department store) 1751.31 – 35

persons had already made their decision at home. Therefore, he/she stays at the POS, irrespective of the decision to buy
these cases were excluded from further analysis. Most “buy- something (or not to). According to the theory, time spent at
ers” in the first study tended to buy convenience goods (like the POS is dependent on arousal evoked by the store atmo-
chocolate, cake, flowers, etc.). As shown in Table 3, in both sphere [11,38].
studies, a comparison between buyers and non-buyers reveals To investigate different arousal patterns in the sports
significant differences with respect to total amplitude, thus, department store, also data mining–cluster analyses were
supporting the hypothesis. In the second study, the frequency conducted with total amplitude, frequency and time as cluster
of EDR reactions also varied significantly (for details, see variables. Within these segments, different arousal patterns
Ref. [25]; Table 3). can be detected (Fig. 3). In a second step, statements mea-
In the 3rd study, there was no significant difference suring buying behavior, where compared between the four
between buyers and non-buyers with regard to time spent in groups (see Fig. 4). The results can explain possible causes of
the mall (on average, EDR-registration lasted about 6 min), the different arousal pattern. “Elaborating buyers” declared
whereas in the department store, buyers (14 min) spent 6 min that before they decided to buy something they examined
longer than non-buyers (8 min). In both settings, the exper- if the products matched their life-styles. “Long-staying non-
imental design gave test persons the opportunity to expose buyers” forgot the time while browsing (therefore, they could
themselves to any stimuli as long as they liked. This means also be characterized as “browsers”) and “fast deciders” liked
the individual may gather more EDR responses, the longer the special-offers and cheap prices. In contrast, “fast refusers”

Fig. 4. Attitudes of buyers and non-buyers.


436 A. Groeppel-Klein / Brain Research Bulletin 67 (2005) 428–437

can be characterized by a more negative attitude towards the [5] W. Boucsein, Electrodermal Activity, Plenum Press, NY/London,
store. 1992.
[6] W. Boucsein, Aktivierung, in: H. Luczak, W. Volpert (Eds.), Hand-
buch Arbeitswissenschaft, Schaeffer-Poeschel, Stuttgart, 1997, pp.
8. Discussion 309–312.
[7] W. Boucsein, Physiologische Grundlagen und Meßmethoden der der-
malen Aktivität, in: F. Rösler (Ed.), Enzyklopädie der Psychologie,
The results show that recording EDA at the point-of-sale Bereich Psychophysiologie, Band 1: Grundlagen and Methoden der
is a practicable way to measure arousal in a valid manner. The Psychophysiologie, Kapitel, Vol. 7, Hogrefe, Göttingen, 2001, pp.
findings of Study 1 confirmed the hypothesis that the experi- 551–623.
mental store represents the higher arousing store environment [8] R. Buck, Human Motivation and Emotion, second ed., John Wiley
and Sons, NY, 1988.
than the control store and that arousal was significantly cor-
[9] J.F. Cohn, A.J. Zlochower, J. Lien, T. Kanade, Automated face analy-
related with “joy” (similar results in Study 2). EDR thus sis by feature point tracking has high concurrent validity with manual
seems to be a future research method to work out visual mer- FACS coding, Psychophysiology 36 (1999) 35–43.
chandising concepts that evoke an optimum level of arousal. [10] H.D. Critchley, R. Elliott, C.J. Mathias, R.J. Dolan, Neural activity
EDR recording can also be applied to explore arousal and relating to generation and representation of galvanic skin conduc-
tance responses: a functional magnetic resonance imaging study, J.
orienting reflexes derived from different environmental stim-
Neurosci. 15, 20 (8) (2000) 3033–3040.
uli settings, and the arousal potential of elements designed [11] R.J. Donovan, J.R. Rossiter, Store atmosphere: an environmental psy-
to be visually striking. Furthermore, EDR records arousal chology approach, J. Retail. 28 (1982) 34–57.
evoked at the point-of-sale simultaneously to the perception [12] R.J. Donovan, J.R. Rossiter, G. Marcoolyn, A. Nesdale, Store atmo-
of stimuli during shopping. sphere and purchasing behavior, J. Retail. 70 (3) (1994) 283–294.
[13] E. Duffy, Activation and Behavior, John Wiley and Sons,
Summarizing the essence of Studies 2 und 3, it is evident
NY/London, 1962.
that arousal is an important construct for the explanation of [14] P. Ekman, W.V. Friesen, Facial Action Coding System (FACS): A
buying behavior. EDR shows that buyers are more aroused Technique for the Measurement of Facial Actions, Consulting Psy-
than non-buyers. Moreover, clusters with different arousal chologists Press, Palo Alto, 1978.
patterns can be detected that also show different decision- [15] P. Ekman, W.V. Friesen, S.S. Tomkins, Facial affect scoring tech-
nique: a first validity study, Semiotica 3 (1) (1971) 37–58.
making patterns. Thus, arousal measurement will be useful
[16] M.W. Eysenck, Attention and Arousal, Springer, Berlin, 1982.
to distinguish different problem-solving types (see Ref. [25]). [17] E. Fischer, S.J. Arnold, More than a labor of love: gender roles and
Furthermore, it would be fascinating to combine market- Christmas shopping, J. Consum. Res. 17 (1990) 333–345.
ing relevant results from arousal research at the point-of-sale [18] M.H. Flicker, W.C. Speer, Emotional responses to store layout and
with fMRI or PET research in the lab [10]. An interesting design: an experimental approach, in: A. Parasuraman (Ed.), Enhanc-
ing Knowledge Development in Marketing, Vol. 1, AMA Educators’
research question could be to find out first which products
Proceedings, AMA, Chicago, 1990, pp. 1–5.
evoke high phasic arousal reactions at the point-of sale and [19] M. Fredrikson, T. Furmark, M.T. Olsson, H. Fischer, J. Andersson, B.
then to analyze via brain-imaging, which areas in the brain Langström, Functional neuroanatomical correlates of electrodermal
are activated by these products. activity: a positron emission tomographic study, Psychophysiology
However, EDA recording has also limitations, especially, 35 (1998) 179–185.
[20] A. Groeppel-Klein, The influence of the dominance perceived at the
concerning the time that is needed in order to collect satis-
PoS on the price-assessment, in: B.G. Englis, A. Olofsson (Eds.),
fying sample sizes. Also, as long as computerized artifact European Advances in Consumer Research, Vol. 3, Association for
detection is not possible, artifact detection needs to be done Consumer Research, Provo, UT, 1997, pp. 304–311.
manually by screening each single curve. Another restriction [21] A. Groeppel-Klein, The avoidance of usury-images evoked by exclu-
is the fact that EDA cannot reveal whether an arousing situa- sive store-designs, in: E. Breivik, A.W. Falkenberg, K. Grønhaug
(Eds.), Rethinking European Marketing, Proceedings From the 30th
tion was perceived with positive or with negative emotions or
EMAC conference, The Norwegian School of Economics and Busi-
attitudes, whether a situation was rather pleasing or unpleas- ness Administration, Bergen, 2001, pp. 193–200.
ing. Verbal control of perceived emotions and attitudes by a [22] A. Groeppel-Klein, Aktivierungsforschung und Konsumentenverhal-
standardized questionnaire, therefore, is still necessary. ten, in: A. Groeppel-Klein (Ed.), Konsumentenverhaltensforschung
im, Vol. 21, Jahrhundert, Wiesbaden, 2004, pp. 29–66.
[23] A. Groeppel-Klein, D. Baun, The role of customers’ arousal for retail
stores—results from an experimental pilot study using electrodermal
References activity as indicator, in: M.C. Gilly, J. Meyers-Levy (Eds.), Adv.
Consum. Res. XXVIII (2001) 412–419.
[1] B.J. Babin, W.R. Darden, M. Griffin, Work and/or fun: measuring [24] A. Groeppel, B. Bloch, An investigation of experience-orientated
hedonic and utilitarian shopping value, J. Consum. Res. 10 (1994) consumers in retailing: the international review of retail, Distrib.
644–656. Consum. Res. 1 (4) (1990) 101–118.
[2] R.P. Bagozzi, The role of psychophysiology in consumer research, [25] A. Groeppel-Klein, C.C. Germelmann, A. Domke, H. Woratschek,
in: T.B. Robertson, H.H. Kassarjian (Eds.), Handbook of Consumer Arousal as a driving force for decision making—empirical results
Behavior, Prentice Hall, Englewood Cliffs, 1991, pp. 124–161. from measuring electrodermal reactions at the point-of-sale, Adv.
[3] R.P. Bagozzi, M. Gopinath, P.U. Nyer, The role of emotions in mar- Consum. Res. XXXII, in press.
keting, J. Acad. Market. Sci. 27 (1999) 184–206. [26] M.B. Holbrook, E.C. Hirschmann, The experiential aspects of con-
[4] S. Bekmeier, Nonverbale Kommunikation in der Fernsehwerbung, sumption: consumer fantasies, feelings, and fun, J. Consum. Res. 9
Physica, Heidelberg, 1989. (1982) 132–140.
A. Groeppel-Klein / Brain Research Bulletin 67 (2005) 428–437 437

[27] W. Kroeber-Riel, Activation Research—psychobiological approaches [35] A. Steiger, Computergestützte Aktivierungsmessung in Der Market-
in consumer research, J. Consum. Res. 5 (1979) 240– ingforschung, Peter Lang, Frankfurt am Main, Bern, NY/Paris, 1988.
250. [36] S.H.C. Tai, A.M.C. Fung, Application of an environmental psychol-
[28] J. Le Doux, The Emotional Brain. The Mysterious Underpinnings ogy model to in-store buying behaviour: The International Review
of Emotional Life, Simon, Schuster, NY, 1996. of Retail, Distrib. Consum. Res. 7 (4) (1997) 311–337.
[29] D.B. Lindsley, Emotions, in: S.S. Stevens (Ed.), Handbook of Exper- [37] Y. Tian, T. Kanade, J.F. Cohn, Recognizing action units for facial
imental Psychology, Wiley, NY, 1951, pp. 473–516. expression analysis, IEEE Trans. Pattern Anal. Mach. Intell. 23 (2)
[30] A. Mehrabian, J.A. Russell, An Approach to Environmental Psy- (2001) 97–115.
chology, MIT, Cambridge, MA, 1974. [38] P. van Kenhove, P. Desrumaux, The relationship between emotional
[31] G. Meyer-Hentschel, Aktivierungswirkung von Anzeigen, Physica, states and approach or avoidance responses in a retail environment:
Würzburg, Wien, 1983. The International Review of Retail, Distrib. Consum. Res. 7 (1997)
[32] C.M. Portas, G. Rees, A.M. Howseman, O. Josephs, R. Turner, C.D. 351–368.
Frith, A specific role for the thalamus in mediating the interaction [39] P.H. Venables, M.J. Christie, Eectrodermal activity, in: I. Martin,
of attention and arousal in humans, J. Neurosci. 18 (21) (1998) P.H. Venables (Eds.), Techniques in Psychophysiology, Wiley, NY,
8979–8989. 1980, pp. 3–67.
[33] J.A. Russell, A. Mehrabian, Evidence for a three-factor theory of [40] P. Vitouch, Psychophysiological methods in media research, in:
emotions, J. Res. Perspect. 11 (1977) 273–294. P. Winterhoff-Spurk, T.H.A. van der Voort (Eds.), New Horizons
[34] J.F. Sherry Jr., A sociocultural analysis of a midwestern flea market, in Media Psychology, Westdeutscher Verlag, Opladen, 1997, pp.
J. Consum. Res. 17 (1990) 13–30. 116–124.
Neuron, Vol. 44, 379–387, October 14, 2004, Copyright ©2004 by Cell Press

Neural Correlates of Behavioral Preference


for Culturally Familiar Drinks

Samuel M. McClure,1,2 Jian Li,1 Damon Tomlin, neural responses, and the modulation of both by non-
Kim S. Cypert, Latané M. Montague, odor or nonflavor stimuli—that is, the sensory problem.
and P. Read Montague* Ultimately, such sensory discriminations and the vari-
Department of Neuroscience ables that influence them serve to influence expressed
Menninger Department of Psychiatry behavioral preferences. Hence, there is another large
and Behavioral Sciences piece of the problem to understand. For modern hu-
Baylor College of Medicine mans, behavioral preferences for food and beverages
1 Baylor Plaza are potentially modulated by an enormous number of
Houston, Texas 77030 sensory variables, hedonic states, expectations, seman-
tic priming, and social context. This assertion can be
illustrated with a quote from Anderson and Sobel (2003)
Summary profiling the work of Small et al. (2003) on taste intensity
and pleasantness processing:
Coca-Cola! (Coke!) and Pepsi! are nearly identical in
chemical composition, yet humans routinely display “A salad of perfectly grilled woodsy-flavored cala-
mari paired with subtly bitter pale green leaves of
strong subjective preferences for one or the other.
curly endive and succulent petals of tomato flesh in
This simple observation raises the important question a deep, rich balsamic dressing. Delicate slices of
of how cultural messages combine with content to pan-roasted duck breast saturated with an assertive,
shape our perceptions; even to the point of modifying tart-sweet tamarind-infused marinade.”
behavioral preferences for a primary reward like a sug-
ared drink. We delivered Coke and Pepsi to human The text goes on further, but note that the sheer lush-
subjects in behavioral taste tests and also in passive ness of the description adds somehow to the appeal of
experiments carried out during functional magnetic the food described. Also notice one implicit point of the
resonance imaging (fMRI). Two conditions were exam- description: many levels of social, cognitive, and cultural
ined: (1) anonymous delivery of Coke and Pepsi and influences combine to produce behavioral preferences
(2) brand-cued delivery of Coke and Pepsi. For the for food and drink. The above description likely would
anonymous task, we report a consistent neural re- not appeal to a strict vegan or an owner of a pet duck.
sponse in the ventromedial prefrontal cortex that cor- Anderson and Sobel point out that the preferences in-
related with subjects’ behavioral preferences for these dexed by their prose originated from the economic de-
beverages. In the brand-cued experiment, brand knowl- mands on our early forebears and were unlikely to have
edge for one of the drinks had a dramatic influence been strictly about aesthetic responses to food and
on expressed behavioral preferences and on the mea- drink.
sured brain responses. However, the modern problem is different. Cultural
influences on our behavioral preferences for food and
Introduction drink are now intertwined with the biological expediency
that shaped the early version of the underlying prefer-
Perceptual constructs are generally multidimensional, ence mechanisms. In many cases, cultural influences
integrating multiple physical and cognitive dimensions dominate what we eat and drink. Behavioral evidence
to generate coherent behavioral preferences. In sensory suggests that cultural messages can insinuate them-
processing, the idea of multidimensional integration has selves into the decision-making processes that yield
long been used to frame a range of questions about preferences for one consumable or another. Conse-
cross-modal interactions in physiological and behav- quently, the appeal or repulsion of culturally relevant
ioral responses (Stein et al., 1996; 1999; Wallace and sights, sounds, and their associated memories all con-
Stein, 1997; Armony and Dolan, 2001; Dolan et al., 2001; tribute to the modern construction of food and drink
Laurienti et al., 2002, 2003). This same multidimensional preferences. The neural substrates underlying food and
perspective has also been developed for olfactory and drink preferences and their influence by cultural images
gustatory processing, where the detection, discrimina- have not been explored. As alluded to above, the major-
tion, and perceived intensity of stimuli are not only func- ity of work on olfaction and gustation has focused on
tions of the primary physical properties (odors, flavors) sensory processing. In this paper, we combine simple
but are also modulated “cross-modally” by visual input taste tests and event-related functional magnetic reso-
(Gottfried and Dolan, 2003), auditory input, and current nance imaging (fMRI) to probe the neural responses that
reward value (Gottfried et al., 2003). correlate with the behavioral preference for noncarbon-
The work just described has focused on the percep- ated versions of Coke! and Pepsi!. We further investi-
tual discrimination of odors and flavors, the correlated gate the influence of the brand image on behavioral
choice and brain response to both drinks.
These two stimuli were chosen for three reasons. (1)
*Correspondence: read@bcm.tmc.edu
1
These authors contributed equally to this work. They are culturally familiar to subjects. (2) They are both
2
Present address: Department of Psychology, Princeton University, primarily composed of brown, carbonated sugar water,
Princeton, New Jersey 08544. and sugar water serves as a primary reward in many
Neuron
380

animal and human experiments. (3) Despite their similari- mental Procedures for details). In the taste test, the first
ties, they generate a large subjective preference differ- two groups chose between two unmarked cups, one of
ence across human subjects, which might correlate with which contained Pepsi and the other Coke in a 3-trial and
fMRI-measured brain responses. We pursued three pri- a 15-trial behavioral task, correspondingly (anonymous
mary questions using the experiments presented in this taste test; Figure 1A). The other two groups (semianony-
paper. (1) What is the behavioral and neural response mous taste tests) made three preference decisions, but
to these drinks when presented anonymously? (2) What in this case both cups contained the same drink (either
is the behavioral and neural influence of knowledge Pepsi or Coke); however, one cup was unlabeled and
about which drink is being consumed? (3) In questions the other indicated the brand of the drink contained in
1 and 2, is there a correlation between the expressed the cup. In these semianonymous taste tests, subjects
behavioral preference and the neural response as mea- were told that the unlabeled cups contained either Pepsi
sured using fMRI? or Coke, and hence no deception was involved. Prefer-
The medical importance of understanding these ques- ences exhibited during the taste tests are referred to as
tions is straightforward—there is literally a growing crisis behavioral preferences.
in obesity, type II diabetes, and all their sequelae that To determine how subjects’ preferences interact with
result directly from or are exacerbated by overconsump- brand information at the level of brain activity, the four
tion of calories (for recent work see Chacko et al., 2003; subject populations completed scanning experiments
Ford et al., 2003; Wyatt, 2003; Zimmet, 2003; Popkin analogous to the taste test manipulations. The taste test
and Nielsen, 2003). It is now strongly suspected that outside the scanner and drink delivery paradigm inside
one major culprit is sugared colas (Popkin and Nielsen, the scanner are illustrated in Figures 1A and 1B. The
2003). The possibility of obtaining coherent answers to design of these experiments was complicated by two
these questions derives from the growing fMRI work on facts: (1) blood oxygenation level-dependent (BOLD) re-
reward processing. sponses in reward-related brain areas are significantly
Recent work using fMRI has identified reward-related affected by whether stimuli are predictable in time
brain responses that scale with the degree to which (Berns et al., 2001; O’Doherty et al., 2003a; McClure et
subjects find stimuli pleasing or rewarding (Knutson et al., 2003), and (2) we wished to study differential brain
al., 2001; Aharon et al., 2001). With this information in responses to soda delivery alone. Because of these con-
hand, it is tempting to suggest that humans will choose straints, we could not show brand information simulta-
more pleasing stimuli over less pleasing stimuli by evalu- neous with soda delivery since this would make it impos-
ation and comparison and that, for our two sugared sible to separate the perception of brand information
drinks, the most pleasing drink is the one that subjec- from the response to soda delivery. However, with brand
tively tastes better than its competitor. This perspective information displayed prior to soda delivery, we had
offers the simplest model that connects reward-related to contend with making the soda delivery predictable.
brain responses to expressed behavioral preferences. There were two possible strategies: (1) randomize the
However, most real-world settings present numerous time between the display of brand information and deliv-
primary sensations and top-down influences that act to ery of soda or (2) construct a design in which brain
organize a coherent behavioral preference. Studies have responses can still be elicited when brain information
indeed shown that cultural information can modulate consistently (fixed time) precedes soda delivery. Since
reward-related brain response (Erk et al., 2002). This the effect of unpredictability is still a matter of investiga-
general observation is particularly true for Coke and tion and there is some evidence that brain responses
may be elicited in reward-related brain structures based
Pepsi; that is, there are visual images and marketing
on unpredictability alone (Zink et al., 2004), we opted
messages that have insinuated themselves into the ner-
for the second option. In particular, we adopted the
vous systems of humans that consume the drinks. It
design of McClure et al. (2003) in which strong BOLD
is possible that these cultural messages perturb taste
responses were elicited even after a large number of
perception; however, no direct neural probes of this
event repetitions.
possibility have been carried out. It is this issue and its
We trained subjects to expect Coke and Pepsi at fixed
implications that we sought to address, and our results
times (6 s) following distinct visual cues (Figures 1B, 4A,
suggest that there might be parallel mechanisms in the
and 4C, training period). After training, we then studied
brain cooperating to bias preference.
brain responses evoked by delayed, unexpected (10 s
following cue) cola delivery. For the first and second
Results subject groups (anonymous), the predictive visual cues
were flashes of yellow and red light, counterbalanced
A total of 67 subjects participated in the study. They and paired with subsequent Coke and Pepsi delivery.
were separated into four groups (n1 ! 16, n2 ! 17, n3 ! In the third and fourth subject groups (semianonymous),
16, n4 ! 18). Each group was given a separate taste test the two fluids were the same (both either Pepsi or Coke).
outside the scanner and drink delivery paradigm while One of the cues was anonymous (yellow or red light),
in the scanner. and the other contained brand information (picture of a
For all four groups of subjects, two separate measures Coke can or a Pepsi can; Figures 4A and 4C).
of behavioral preferences were obtained. First, subjects In all fMRI experiments in which subjects must swal-
were asked “Which drink you prefer to consume: Coke, low, head motion is a concern. Using head constraints,
Pepsi, or no preference?” Their answers are referred to subjects’ head movements remained no larger than
as their stated preferences. Next, subjects engaged in 2 mm during the entire experiment. In addition, we per-
three rounds of a forced-choice taste test (see Experi- formed two separate analyses that both ruled out the
Brain Response to Culturally Familiar Drinks
381

Figure 1. Anonymous Coke and Pepsi Task


Settings and Behavioral Results
(A) Anonymous taste test. Each subject was
given a taste test outside the scanner. The
test required subjects to make 3 separate
choices (groups 1 and 3) or 15 choices (group
2) in which they indicated their preference for
the soda in one of two unmarked cups. One
of the cups contained 10 mL of Coke, and
the other contained an equal volume of Pepsi.
(B) In the scanner, subjects were trained to
expect soda delivery at a 6 s delay following
light illumination using a Pavlovian condition-
ing paradigm. Twenty light-drink pairings
were used for training, separated by random
time intervals of between 6 and 16 s. Follow-
ing training, 6 test pairings for each liquid
were randomly interleaved during the suc-
ceeding 25 pairings in which soda delivery
was unexpectedly delayed for 4 s; evoked
brain responses were studied for these 12
delayed deliveries.
(C) Correlation of Coke preference in behav-
ioral taste tasks between the original 3 trials
and later 15 trials (groups 1 and 3; red; r2 !
0.51, n ! 15), between first 3 trials and the
whole 15 trials in the other independent 15-
trial anonymous taste test (group 2; green;
r2 ! 0.78, n ! 16).
(D) Average Coke preference in Coke and Pepsi drinkers in a 15-trial carbonated Coke-Pepsi taste task. Average Coke selection is 7.5 " 0.8
(mean " SE) for Coke drinkers and 6.8 " 0.5 (mean " SE) for Pepsi drinkers. Our data do not support that stated preference is correlated
with behavioral preference in the carbonated state (two-tailed Student’s t test, p ! 0.46, n1 ! 18, n2 ! 18).

possibility that any of the following results were contami- For technical reasons (carbonation builds up in deliv-
nated by this possible confound (see Experimental Pro- ery tubes causing unreliable soda delivery in the scanner
cedures for more details). First, a two-way ANOVA test experiment), all the taste tests were originally conducted
of the event-related head movements and six parame- with decarbonated soft drinks. However, we now show
ters of head movements (x, y, z, pitch, yaw, roll) in all that the behavioral results are unaffected by carbon-
of the four groups was performed. We found that move- ation. A separate anonymous taste task of 15 forced-
ment was uncorrelated with any event (p # 0.6 for all choice trials was presented to each subject with carbon-
movement parameters). Second, in those areas which ated drinks (Coke and Pepsi). Stated and behavioral
are identified as “active,” we used head movement pa- preferences were not correlated in this condition (Figure
rameters as regressors to determine whether their activi- 1D, two-tailed Student’s t test, p ! 0.46).
ties were significantly correlated with head movements. Behavioral preferences measured in the 3-trial taste
This revealed no statistical significance (p # 0.1 for all task are a potentially unreliable measure of subjects’
movement parameters). true preferences due to the small number of measure-
All of the subsequent findings relate to brain struc- ments involved in the test. To account for this, we re-
tures that are not directly involved in processing gusta- called our subjects at a delay of several months and had
tory stimuli. Therefore, it is important to note that signifi- them repeat the taste test with 15 trials. The outcome
cant brain activity was evoked by the delivery of Coke of these two separate tests are strongly correlated, as
or Pepsi in gustatory cortical regions (insular cortex; p $ shown in Figure 1C (subjects recalled from groups 1
0.01 for both drinks) but was unaffected by any of our and 3; red; y ! 3.6x % 1.5, r2 ! 0.51, n ! 15). Consistency
experimental factors. of preferences in 3-trial and 15-trial taste tests was fur-
ther confirmed within session. In a separate group of
Group 1-2: Anonymous Taste Test subjects (Figure 1C, green), the outcome of their first 3
The results of the anonymous taste test (3-trial version, trials well predicted the outcome of the full 15-trial test
group 1) are shown in Figure 3A. A histogram is shown (y ! 3.8x % 2.2, r2 ! 0.78, n ! 16).
which summarizes the behavioral preferences for all 16
subjects. Subjects were balanced, with a nearly equal Group 1-2: Scanning, Anonymous Drink Delivery
number preferring Pepsi, preferring Coke, or showing A linear regression analysis using behavioral prefer-
no distinct preference. Similarly, there was no difference ences from the 3-trial anonymous taste task as a re-
in subjects’ stated preferences (data not shown), with gressor indicated that the difference in brain responses
an equal number of subjects declaring a preference for evoked by Coke and Pepsi in the ventromedial prefrontal
Coke (n ! 7) and Pepsi (n ! 6; p ! 0.79). However, cortex (VMPFC; MNI coordinates [8, 56, 0]; peak z score
the correlation between subjects’ stated and behavioral 3.44) (Figure 2B) scaled monotonically with the results of
preferences does not reach statistical significance (r2 ! the behavioral taste test (Figure 2A) (no other significant
0.14; p ! 0.16). regions at p $ 0.001, uncorrected for multiple compari-
Neuron
382

Figure 2. Neural Correlates of Preference for Anonymous Coke and Pepsi Delivery in 3-Trial and 15-Trial Anonymous Taste Tasks
(A) Behavioral preferences expressed in the 3 trial taste test varied linearly with brain responses in the ventromedial prefrontal cortex (group
1). The vertical axis is the contrast (delayed Coke response ! delayed Pepsi response) for the voxels shown in (B).
(B) SPM of neural correlates of behavior preference shown in (A) (thresholded at p " 0.001; uncorrected for multiple comparisons).
(C) Correlation between behavioral preferences expressed in the 15 trial taste and brain responses in the ventromedial prefrontal cortex (group 2).
(D) SPM of neural correlates of behavior preference shown in (C) (thresholded at p " 0.001; uncorrected for multiple comparisons).

sons). As with behavioral preferences, the brain re- Group 3: Semianonymous Taste Test, Coke
sponse in the VMPFC was independent of the subjects’ As before, three pairs of cups were presented to the
stated preferences (paired Student’s t test, p # 0.87). subjects. However, in each pair one of the cups was
As with the behavioral results, it is possible that this labeled “Coke” and the other was left unlabeled. For
finding may suffer from noise in our estimates of sub- the unlabeled cups, the subjects were told that they
jects’ preferences. However, the correlation of BOLD could contain either Coke or Pepsi. A Mann-Whitney
responses in the VMPFC with preference was replicated U test showed that the effect of the Coke label was
in the subjects from group 2 whose preference measures significant when compared with the anonymous taste
were based on 15-round taste tests (Figures 2C and 2D, test, with subjects showing a strong bias in favor of the
MNI coordinates [8, 60, 0]; peak z score 3.83; p " 0.001, labeled cup (Figure 3C, p " 0.05). This was not likely to
uncorrected). The same region of the VMPFC is also be a result of spurious subject sampling, because, when
significantly correlated with behavioral preference when the subjects were later asked to complete the anony-
data from group 1 and group 2 are combined (p " 0.001, mous taste test, their results were not significantly differ-
uncorrected for multiple comparisons; data not shown). ent from the group 1 (anonymous) results (Figure 3A,
Individual subjects generally had a strong stated pref- p # 0.84). Furthermore, these behavioral effects did not
erence for either Coke or Pepsi, and, at any particular correlate with subjects’ stated preferences (p # 0.92).
time, a guess of which soda they were receiving may
have influenced evoked neural responses. The results Group 3: Scanning, Semianonymous Coke
presented here are not likely to be due exclusively to Figure 4A shows the stimulus paradigm for the Coke
such top-down influences of brand preference, since label task. In this condition, one cue was a depiction of
stated preference did not correlate with the behavioral a Coke can followed 6 s later by Coke delivery. The
preference (taste test results) (r2 # 0.14). However, to other stimulus was a light followed by Coke delivery.
explicitly test for effects of brand knowledge, this influ- The number of cue-drink pairings, the number of catch
ence was directly modulated in the following two tasks. trials per cue, and the pseudorandom times between
We developed the working hypothesis that the label pairings were exactly the same as in the anonymous
of either or both drinks would influence the expressed drink delivery task described above (Group 1-2: Scan-
behavioral preference of the subjects. In particular, we ning; see Experimental Procedures). We contrasted the
tested whether knowledge of which cola was being con- brain response to surprising delivery of Coke when it
sumed influenced subjects’ responses. was known to be Coke with the surprising delivery of
Brain Response to Culturally Familiar Drinks
383

Figure 3. Effect of Brand Knowledge on Be-


havioral Preferences
(A) Histogram of subjects’ preference in dou-
ble anonymous task. The x axis indicates the
number of selections made to Coke (maxi-
mum of three). Subjects showed no bias for
either Coke or Pepsi.
(B) Histogram of subjects’ behavior prefer-
ence in semianonymous Pepsi task. The x
axis indicates the number of selections to the
Pepsi-labeled cup. Subjects showed no bias
for either the labeled or unlabeled drink.
(C) Histogram of subjects’ behavior prefer-
ence in the semianonymous Coke task. The
x axis indicates the number of selections to
the labeled Coke. This preference distribution
is different from the double anonymous task
(Mann-Whitney U task, n1 ! 16, n2 ! 16, U !
191.5, p " 0.05) and semianonymous Pepsi
task (n1 ! 18, n2 ! 18, U ! 225.5, p " 0.005),
with subjects demonstrating a strong bias in
favor of the labeled drink.
(D) Average scores of subjects’ preference
(number of selections to Coke, labeled Pepsi,
and labeled Coke, respectively) in the three
behavioral tasks (A–C). Subjects tended to
prefer the labeled Coke drink over anony-
mous Coke (one-way Student’s t test, p "
0.01). The Coke label had a bigger effect in biasing subjects’ preferences than the Pepsi label (one-way Student’s t test, p " 0.005).
(E) Subjects who participated in the semianonymous Coke task later completed the anonymous taste test. The distribution of people’s
preference is significantly different from the Coke-labeled task (Mann-Whitney U test, n1 ! 16, n2 ! 13, U ! 142.5, p " 0.01) but no different
from the results in (A).

Coke when it could have been Coke or Pepsi. The results effect of brand knowledge for Pepsi (Figure 4D). As in
are shown in Figure 4B. Significant differential activity is the semianonymous Coke task, activity in the VMPFC
observed in several brain areas (p " 0.001, uncorrected): did not show a significant effect of brand knowledge
bilateral hippocampus, parahippocampus, midbrain, (p ! 0.89). Further analysis of the hippocampus and
dorsolateral prefrontal cortex (DLPFC), thalamus, and DLPFC revealed that these areas were not significant
left visual cortex. Details are listed in Table 1. At p " even at lower thresholds (p " 0.01, uncorrected). In
0.005 (uncorrected), the activation in the left hippocam- particular, the p value within the area of the hippocam-
pus, left parahippocampus, and midbrain are contigu- pus identified in the semianonymous Coke task was
ous (Table 1). BOLD signal changes in the area of the 0.43, while in the DLPFC it was 0.41. Further, exclusively
VMPFC identified in the anonymous task were unaf- masking the results in the semianonymous task (at p "
fected by brand knowledge (two-tailed paired Student’s 0.01) with the results from the semianonymous Coke
t test, p ! 0.96). task revealed no common areas of activation. Thus, it
seems that brand knowledge for Coke and Pepsi have
Group 4: Semianonymous Taste Test, Pepsi truly different responses both in terms of affecting be-
havioral preference and in terms of modifying brain re-
The taste test for this group was conducted exactly as
sponses.
for group 3 (semianonymous Coke), except both cups
in each pair contained Pepsi, and one was labeled as
Discussion
Pepsi. Again, subjects were told that the unlabeled cup
could contain either Coke or Pepsi. Unlike the Coke
In these experiments, we used functional brain scanning
label, the existence of the Pepsi label did not change
to find correlates of people’s preferences for two similar
the distribution of choices significantly relative to the
sugared drinks: Coke and Pepsi. We report the finding
anonymous taste test (Figure 3B, Mann-Whitney U test,
that two separate systems are involved in generating
p ! 0.82). Furthermore, selections were biased in favor
preferences. When judgments are based solely on sen-
of the Coke label (in the semianonymous task, above)
sory information, relative activity in the VMPFC predicts
to a significantly greater degree than they were in favor
people’s preferences. However, in the case of Coke and
of the Pepsi label (Figure 3D, p " 0.005). Pepsi, sensory information plays only a part in determin-
ing people’s behavior. Indeed, brand knowledge (at least
Group 4: Scanning, Semianonymous Pepsi in the case of Coke in our study) biases preference
Figure 4C shows the stimulus paradigm for the Pepsi decisions and recruits the hippocampus, DLPFC, and
label task. As with group 3, we contrasted the brain midbrain. Our results suggest that the VMPFC and hip-
response elicited by the unexpected delivery of labeled pocampus/DLPFC/midbrain might function indepen-
versus unlabeled Pepsi. At a threshold of p " 0.001 dently to bias preferences based on sensory and cultural
(uncorrected), no brain areas showed a significant main information, respectively.
Neuron
384

Figure 4. Effect of Brand Knowledge on Brain Responses in Semianonymous Tasks


(A) An image of a Coke can was used to cue the occurrence of Coke. A red or yellow circle (randomized across subjects) predicted the other.
Both sodas delivered were Coke.
(B) Coke delivered following an image of a Coke can evoked significantly greater activity in several regions when contrasted against Coke
delivered following a neutral flash of light. Significant activations (p ! 0.001, uncorrected) were found bilaterally in the hippocampus (MNI
coordinates ["24, "24, "20] and [20, "20, "16]), in the left parahippocampal cortex (MNI coordinates ["20, "32, "8]), midbrain (MNI
coordinates ["12, "20, "16]), and dorsolateral prefrontal cortex (MNI coordinates [20, 30, 48]). See Table 1 for details.
(C) In the scanner, an image of a Pepsi can was used to cue the occurrence of Pepsi. A red or yellow circle predicted the other soda, and
both sodas delivered were Pepsi.
(D) No voxels survive p ! 0.001 threshold (uncorrected) for the equivalent contrast in the semianonymous Pepsi experiment.

Coke and Pepsi are special in that, while they have (Figure 3A). The functional brain imaging results corrob-
very similar chemical composition, people maintain orate the behavioral taste test results. The BOLD signal
strong behavioral preferences for one over the other. in the VMPFC correlated strongly with the behavior re-
We initially measured these behavioral preferences ob- sults of the double-blind taste tests. This area of the
jectively, by administering double-blind taste tests. We brain is strongly implicated in signaling basic appetitive
found that subjects split equally in their preference for aspects of reward. Imaging data in healthy subjects
Coke and Pepsi in the absence of brand information indicated with BOLD signal changes scale in the VMPFC

Table 1. Location of Brain Areas that Respond Preferentially to Brand-Cued versus Light-Cued Coke Delivery

Activations are for the semianonymous (Coke) experiment (p ! 0.001 and p # 0.005 in parentheses). L, left hemisphere; R, right hemisphere.
Brain Response to Culturally Familiar Drinks
385

with reward value (Knutson et al., 2001; O’Doherty et portant in recalling affect-related information (Iidaka et
al., 2003b). Furthermore, patients with lesions in the al., 2003; Markowitsch et al., 2003). Our finding supports
VMPFC are insensitive to future reward or punishment these data and suggests that the hippocampus may
value in making decisions (Bechara et al., 1994). A re- participate in recalling cultural information that biases
lated brain region, the medial orbitofrontal cortex (MOFC), preference judgments.
is strongly related to the VMPFC in terms of function. Importantly, the hippocampus and DLPFC are only
Our imaging parameters resulted in significant signal two of several brain areas that have been implicated in
loss in this area due to magnetic field inhomogeneities. biasing behavior based on affect. Other areas include
It remains untested, therefore, whether the MOFC shows the amygdala, ventral striatum, anterior cingulate cortex,
preference-related responses posterior cingulate cortex, and orbitofrontal cortex (for
The other special characteristic of Coke and Pepsi is example, Greene et al., 2001; Schall et al., 2002; Ochsner
that both possess a wealth of cultural meaning. One et al., 2002; Sanfey et al., 2003). To our knowledge, the
important fact that may account for the lack of correla- experiments linking these brain areas to judgments all
tion between subjects’ stated and behavioral prefer- involve subjects making decisions through expressed
ences is the effect of presumed brand knowledge. In motor behavior. In our experiment, by contrast, the
accord with this, subjects expressed strong preferences scanning session involved only the percept of soda, with
for either Coke or Pepsi when asked which type of soda no instruction to make preference decisions. It is an
they normally drink and commonly demonstrated a de- interesting possibility that the hippocampus and DLPFC
sire to prove this preference in the anonymous taste test. are specifically involved in biasing perception based on
To test for effects of brand knowledge, we conducted prior affective bias, whereas the other brain areas listed
a series of semianonymous taste tests and imaging ex- above are more involved in altering behavioral output.
periments. In the taste tests, we found no significant Determining preferences in our experiment appears to
influence of brand knowledge for Pepsi contrasted with result from the interaction of two separate brain systems
the anonymous task. However, there is a dramatic effect situated principally in the prefrontal cortex. The ventro-
of the Coke label on subjects’ behavioral preference. medial region of the prefrontal cortex plays a prominent
Despite the fact that there was Coke in all cups during role when preferences are determined solely from sen-
the taste test, subjects in this part of the experiment sory information. The relative activity in the VMPFC is
preferred Coke in the labeled cups significantly more a very good indicator of which sensory stimulus is pre-
than Coke in the anonymous task and significantly more ferred by the subject. However, cultural influences have
than Pepsi in the parallel semianonymous task. a strong influence on expressed behavioral preferences.
The effects of brand knowledge for Pepsi and Coke We found this to be particularly the case with Coca-Cola,
were reflected in the imaging experiments as well. When for which brand information significantly influences sub-
an image of a Coke can preceded Coke delivery, signifi- jects’ expressed preferences. We hypothesize that cul-
cantly greater brain activity was observed in the DLPFC, tural information biases preference decisions through
hippocampus, and midbrain relative to Coke delivery the dorsolateral region of the prefrontal cortex, with the
preceded by a circle of light. As with the taste test, hippocampus engaged to recall the associated informa-
equivalent knowledge about Pepsi delivery had no such tion. These two systems appear to function indepen-
effect; indeed, no brain areas showed a significant differ- dently in our experiment, since VMPFC activity was un-
ence to Pepsi delivered with versus without brand knowl- affected by brand knowledge. In judging stimuli based
edge. The hippocampus and DLPFC have both been on multifaceted sensory and cultural influences, inde-
previously implicated in modifying behavior based on
pendent brain systems appear to cooperate to bias pref-
emotion and affect. The DLPFC is commonly implicated
erences.
in aspects of cognitive control, including working mem-
ory (e.g., Watanabe, 1996). Lesions to the DLPFC are
Experimental Procedures
also known to result in depression (Davidson, 2002),
which is hypothesized to result from a decreased ability A total of 67 subjects participated in the study (38 male, 29 female;
to use positive affect to modify behavior (Mineka et al., aged 19–50 years, mean ! SD: 28.0 ! 7.6 years old). All subjects
1998). It has been proposed that the DLPFC is necessary gave informed consent to participate in the study; the Baylor Institu-
for employing affective information in biasing behavior tional Review Board approved the experimental paradigm. Each
(Watanabe, 1996; Davidson and Irwin, 1999). This is con- subject participated in one of three similarly designed experiments
(group 1: anonymous, n1 " 16, one failed to finish fMRI scanning
sistent with our findings: labeling Coke in taste and im-
due to technical problems; group 2: anonymous, n2 " 17; group 2:
aging tasks both biases behavior and recruits DLPFC semianonymous Coke, n3 " 16; group 3: semianonymous Pepsi,
activity. Furthermore, both of these effects are lost when n4 " 18). For all four groups, subjects were first given a taste test
compared with the semianonymous Pepsi tasks. and then completed the fMRI study. Subjects were not instructed
The hippocampus has also been implicated in pro- to abstain from drinking prior to the experiment, but all subjects
cessing affective information, but this association is tied reported that they enjoyed the drinks.
to its role in the acquisition and recall of declarative
memories (Eichenbaum, 2000; Markowitsch et al., 2003). Taste Tests
The hippocampus is especially important in relating in- Taste tests consisted of 3 rounds (groups 1, 3, and 4) or 15 rounds
(group 2) of forced-choice preference decisions between two cups
formation to “discontiguous” sensory cues, as in the
of cola drinks (10 mL each). Coke and Pepsi were decarbonated in
case of our task in which there is a 10 s period (catch both the taste tests and scanning experiments in order to ensure
trials) between the cue and soda delivery (McEchron reliable delivery through the plastic tubes required for the scanning
and Disterhoft, 1999; Christian and Thompson, 2003). experiment. In each round of the taste tests, cups were presented
Imaging data indicate that the hippocampus is also im- in random order. In the anonymous test (groups 1 and 2), both cups
Neuron
386

in each round were unlabeled; one cup of each pair contained Coke, and ensured to satisfy a false discovery rate of q ! 0.05 for all of
while the other contained Pepsi (Figure 1A). In group 3 (semianony- our result images (Genovese et al., 2002).
mous Coke), one cup in each pair was labeled “Coke” and the other The motion parameters derived from image realignment were
was unlabeled; both cups contained Coke. The semianonymous used to ensure that head movement did not corrupt our results.
Pepsi experiment (group 4) was the same as the semianonymous This analysis was performed in two separate ways: (1) mean event-
Coke experiment except the cups both contained Pepsi and one of related movement parameters were calculated and tested against
each pair of cups was labeled “Pepsi.” During the test, the Coke the null hypothesis of no significant movement at any time point,
and Pepsi bottles (2 liter versions) were explicitly visible on the table and (2) the movement parameters were entered into a general linear
in front of the subject. For each pair, subjects selected the drink model and regressed to the MRI data in the regions of interest (ROI).
they preferred; we refer to this in the text as behavioral preference. Neither of these efforts revealed any significant effect.
Between each taste-pair, subjects waited at least 40 s and rinsed
with equal volumes of bottled water. Acknowledgments
Subjects who participated in the 3-trial anonymous (group 1 and
group 3) taste test were recalled after a period of several months to We thank Ann Harvey, Kristen Pfeiffer, and Nathan Apple for their
perform the 15-trial anonymous taste test. The number of selections help conducting this experiment. Further thanks go to Leigh Nystrom
each subject made to Coke in the 3-trial and in the 15-trial version for his consultation on data analysis. This work was supported by
of the test are compared in Figure 1C. the Kane Family Foundation (PRM) and NIDA grant DA-11723
(P.R.M.). All trademarks appearing in this article are the property of
fMRI Task their respective owners.
Subjects lay supine with their head in the scanner bore and viewed
a back-projected computer-generated image via a 45" mirror. Sub- Received: February 2, 2004
jects were instructed to watch the screen and swallow the colas Revised: July 22, 2004
when they were delivered; there was no other task to perform. Accepted: September 2, 2004
Whole-brain echo planar images were acquired with BOLD contrast Published: October 13, 2004
and a repetition time of 2 s (performed on a Siemens Allegra 3T
scanner; echo time 40 ms, flip angle 90"). References
We employed a paradigm described in previous work and dis-
cussed more fully in the Results section (Figure 1B, McClure et al., Aharon, I., Etcoff, N., Ariely, D., Chabris, C.F., O’Connor, E., and
2003). Subjects were given a series of training pairings in which Breiter, H.C. (2001). Beautiful faces have variable reward value: fMRI
cues predicted the squirt of cola at a fixed delay of 6 s. After 20 and behavioral evidence. Neuron 32, 537–551.
training pairings, six catch trials, where drink delivery was delayed Anderson, A.K., and Sobel, N. (2003). Dissociating intensity from
by 4 s, were randomly interspersed in the last 25 pairings. This valence as sensory inputs to emotion. Neuron 39, 581–583.
produced a total of 35 pairings for each cue, or a grand total of 70
Armony, J.L., and Dolan, R.J. (2001). Modulation of auditory neural
squirts of cola. The light-drink pairings were separated by a random
responses by a visual context in human fear conditioning. Neurore-
time delay ranging from 6 s to 16 s. We focused on the amplitude
port 12, 3407–3411.
of evoked BOLD responses to Pepsi and Coke delivered at delayed
(unexpected) times. The paradigm was completed in two scanning Bechara, A., Damasio, A.R., Damasio, H., and Anderson, S.W. (1994).
runs of 291 and 296 scans, respectively. Insensitivity to future consequences following damage to human
For the group 1 and 2 subjects, the light stimuli were a red circle prefrontal cortex. Cognition 50, 7–15.
and a yellow circle, with the association of light color to soda flavor Berns, G.S., McClure, S.M., Pagnoni, G., and Montague, P.R. (2001).
randomized across subjects (i.e., red→Coke, yellow→Pepsi or Predictability modulates human brain response to reward. J. Neu-
red→Pepsi, yellow→Coke). For the semianonymous experiments, rosci. 21, 2793–2798.
an image of a Coke can (group 3) and an image of a Pepsi can Chacko, E., McDuff, I., and Jackson, R. (2003). Replacing sugar-
(group 4) were used to cue the occurrence of one of the sodas. A based soft drinks with sugar-free alternatives could slow the prog-
red or yellow circle (randomized across subjects) predicted the ress of the obesity epidemic: have your Coke and drink it too. N. Z.
other. Both sodas were the same (group 3, both Coke; group 4, Med. J. 116, U649.
both Pepsi).
Christian, K.M., and Thompson, R.F. (2003). Neural substrates of
eyeblink conditioning: acquisition and retention. Learn. Mem. 10,
Drink Delivery in Scanner 427–455.
Individual squirts of Coke and Pepsi (0.8 mL each) were delivered
Davidson, R.J. (2002). Anxiety and affective style: role of prefrontal
to subjects through cooled plastic tubes held in the subjects’ mouths
cortex and amygdala. Biol. Psychiatry 51, 68–80.
with plastic mouthpieces. The volume of soda delivered on each
squirt was sufficient to allow the subjects to fully taste the soda but Davidson, R.J., and Irwin, W. (1999). The functional neuroanatomy
were small enough to allow them to easily swallow while lying in the of emotion and affective style. Trends Cogn. Sci. 3, 11–21.
scanner. A computer-controlled syringe pump (Harvard Apparatus, Dolan, R.J., De Gleder, B., and Morris, J.S. (2001). Cross-modal
Holliston, MA) allowed for precise delivery of the colas. binding of fear in voice and face. Proc. Natl. Acad. Sci. USA 98,
10006–10010.
Image Analysis Eichenbaum, H. (2000). A cortical-hippocampal system for declara-
Analysis of functional imaging data was performed using SPM2 tive memory. Nat. Neurosci. 1, 41–50.
(Friston et al., 1995). Raw data were realigned, corrected for slice- Erk, S., Spitzer, M., Wunderlich, A.P., Galley, L., and Walter, H.
timing artifacts, normalized to a canonical spatial axis (resampled (2002). Cultural objects modulate reward circuitry. Neuroreport
at 4 mm # 4 mm # 4 mm), and smoothed with an 8 mm Gaussian 13, 2499–2503.
kernel. In the general linear model analysis, individual events were
Ford, E.S., Mokdad, A.H., and Giles, W.H. (2003). Trends in waist
assumed to evoke known changes in hemodynamic response
circumference among U.S. adults. Obes. Res. 11, 1223–1231.
through time. Linear regression of these assumed response dynam-
ics to measured BOLD signals produced amplitude estimates for Forman, S.D., Cohen, J.D., Fitzgerald, M., Eddy, W.F., Mintun, M.A.,
each class of effect in the experiment. Student’s t tests were per- and Noll, D.C. (1995). Improved assessment of significant activation
formed over these amplitude estimates, constituting a random ef- in functional magnetic resonance imaging (fMRI): use of a cluster-
fects analysis over the subjects. Regions are reported as active that size threshold. Magn. Reson. Med. 33, 636–647.
contain a minimum of three contiguous voxels each significant at Friston, K.J., Holmes, A.P., Worsley, K., Poline, J.P., Frith, C.D., and
p ! 0.001. To protect against type I errors, we further ensured that Frackowiak, R.S.J. (1995). Statistical parametric maps in functional
these regions were composed of at least seven voxels at p ! 0.005 brain imaging: a general linear approach. Hum. Brain Mapp. 2,
(Forman et al., 1995). Both of these thresholds were further tested 189–210.
Brain Response to Culturally Familiar Drinks
387

Genovese, C.R., Lazar, N.A., and Nichols, T. (2002). Thresholding of Wyatt, H.R. (2003). The prevalence of obesity. Prim. Care 30,
statistical maps in functional neuroimaging using the false discovery 267–279.
rate. Neuroimage 15, 870–878. Zimmet, P. (2003). The burden of type 2 diabetes: are we doing
Gottfried, J.A., and Dolan, R.J. (2003). The nose smells what the eye enough? Diabetes. Metab. 29, 6S9–18.
sees: crossmodal visual facilitation of human olfactory perception. Zink, C.F., Pagnoni, G., Martin-Skurski, M.E., Chappelow, J.C., and
Neuron 39, 375–386. Berns, G.S. (2004). Human striatal responses to monetary reward
Gottfried, J.A., O’Doherty, J., and Dloan, R.J. (2003). Encoding pre- depend on saliency. Neuron 42, 509–517.
dictive reward value in human amygdala and orbitofrontal cortex.
Science 301, 1104–1107.
Greene, J.D., Sommerville, R.B., Nystrom, L.E., Darley, J.M., and
Cohen, J.D. (2001). An fMRI investigation of emotional engagement
in moral judgment. Science 293, 2105–2108.
Iidaka, T., Terashima, S., Yamashita, K., Okada, T., Sadato, N., and
Yonekura, Y. (2003). Dissociable neural responses in the hippocam-
pus to the retrieval of facial identity and emotion: an event-related
fMRI study. Hippocampus 13, 429–436.
Knutson, B., Adams, C.M., Fong, G.W., and Hommer, D. (2001).
Anticipation of increasing monetary reward selectively recruits nu-
cleus accumbens. J. Neurosci. 21, RC159.
Laurienti, P.J., Burdette, J.H., Wallace, M.T., Yen, Y.-F., Field, A.S.,
and Stein, B.E. (2002). Activity in visual and auditory cortex can be
modulated by influences from multiple senses. J. Cogn. Neurosci.
14, 420–429.
Laurienti, P.J., Wallace, M.T., Maldjian, J.A., Susi, C.M., Stein, B.E.,
and Burdette, J.H. (2003). Cross-modal sensory processing in the
anterior cingulate and medial prefrontal cortices. Hum. Brain Mapp.
19, 213–223.
McClure, S.M., Berns, G.S., and Montague, P.R. (2003). Temporal
prediction errors in a passive learning task activate human striatum.
Neuron 38, 339–346.
McEchron, M.D., and Disterhoft, J.F. (1999). Hippocampal encoding
of non-spatial trace conditioning. Hippocampus 9, 385–396.
Markowitsch, H.J., Vandekerckhovel, M.M., Lanfermann, H., and
Russ, M.O. (2003). Engagement of lateral and medial prefrontal areas
in the ecphory of sad and happy autobiographical memories. Cortex
39, 643–665.
Mineka, S., Watson, D., and Clark, L.A. (1998). Comorbidity of anxiety
and unipolar mood disorders. Annu. Rev. Psychol. 49, 377–412.
Ochsner, K.N., Bunge, S.A., Gross, J.J., and Gabrieli, J.D. (2002).
Rethinking feelings: an FMRI study of the cognitive regulation of
emotion. J. Cogn. Neurosci. 14, 1215–1229.
O’Doherty, J.P., Dayan, P., Friston, K.J., Critchley, H., and Dolan, R.J.
(2003a). Temporal difference models and reward-related learning in
the human brain. Neuron 38, 329–337.
O’Doherty, J.P., Critchley, H., Deichmann, R., and Dolan, R.J.
(2003b). Dissociating valence of outcome from behavioral control
in human orbital and ventral prefrontal cortices. J. Neurosci. 23,
7931–7939.
Popkin, B.M., and Nielsen, S.J. (2003). The sweetening of the world’s
diet. Obes. Res. 11, 1325–1332.
Sanfey, A.G., Rilling, J.K., Aronson, J.A., Nystrom, L.E., and Cohen,
J.D. (2003). The neural basis of economic decision-making in the
Ultimatum Game. Science 300, 1755–1758.
Schall, J.D., Stuphorn, V., and Brown, J.W. (2002). Monitoring and
control of action by the frontal lobes. Neuron 36, 309–322.
Small, D.M., Gregory, M.D., Mak, Y.E., Gitelman, D., Mesulam, M.M.,
and Parrish, T. (2003). Dissociation of neural representation of inten-
sity and affective valuation in human gustation. Neuron 39, 701–711.
Stein, B.E., London, N., Wilkinson, L.K., and Price, D.D. (1996). En-
hancement of perceived visual intensity by auditory stimuli: A psy-
chophysical analysis. J. Cogn. Neurosci. 8, 497–506.
Stein, B.E., Wallace, M.T., and Stanford, T.R. (1999). Development
of multisensory integration: transforming sensory input into motor
output. Ment. Retard. Dev. Disabil. Res. Rev. 5, 72–85.
Wallace, M.T., and Stein, B.E. (1997). Development of multi-sensory
integration in cat superior colliculus. J. Neurosci. 17, 2429–2444.
Watanabe, M. (1996). Reward expectancy in primate prefrontal neu-
rons. Nature 382, 629–632.
www.elsevier.com/locate/ynimg
NeuroImage 23 (2004) 64 – 74

Dissociable effects of arousal and valence on prefrontal activity


indexing emotional evaluation and subsequent memory: an
event-related fMRI study
Florin Dolcos, a,b,* Kevin S. LaBar, a and Roberto Cabeza a
a
Center for Cognitive Neuroscience, Duke University, Durham, NC 27708-0999, USA
b
Centre for Neuroscience, University of Alberta, Edmonton, AB, Canada, T6G 2S2
Received 3 November 2003; revised 9 April 2004; accepted 18 May 2004

Prefrontal cortex (PFC) activity associated with emotional evaluation ment, various effects of emotion on brain activation associated
and subsequent memory was investigated with event-related functional with different perceptual and cognitive functions have been
MRI (fMRI). Participants were scanned while rating the pleasantness revealed (reviewed in Davidson and Irwin, 1999; Lane and Nadel,
of emotionally positive, negative, and neutral pictures, and memory for 2000; Phan et al., 2002). To understand these diverse effects,
the pictures was tested after scanning. Emotional evaluation was researchers have often divided emotion into its basic underlying
measured by comparing activity during the picture rating task relative constructs. One dimensional approach to emotion emphasizes the
to baseline, and successful encoding was measured by comparing
contribution of two orthogonal components, namely arousal and
activity for subsequently remembered versus forgotten pictures (Dm
effect). The effect of arousal on these measures was indicated by greater valence (Lang et al., 1993; Russell, 1980). Arousal refers to a
activity for both positive and negative pictures than for neutral ones, continuum that varies from calm to excitement, whereas valence
and the effect of valence was indicated by differences in activity refers to a continuum that varies from positive to negative with
between positive and negative pictures. The study yielded three main neutral in the middle (for methods to assess these two dimensions,
results. First, consistent with the valence hypothesis, specific regions in see Bradley and Lang, 1994). The vast majority of studies has
left dorsolateral PFC were more activated for positive than for negative focused on the limbic system and particularly on the amygdala,
picture evaluation, whereas regions in right ventrolateral PFC showed whereas other components of the emotional processing network,
the converse pattern. Second, dorsomedial PFC activity was sensitive to such as the prefrontal cortex (PFC), have received relatively less
emotional arousal, whereas ventromedial PFC activity was sensitive to attention. Recent studies of amygdala function have attempted to
positive valence, consistent with evidence linking these regions,
tease apart the relative contributions of the aforementioned affec-
respectively, to emotional processing and self-awareness or appetitive
behavior. Finally, successful encoding (Dm) activity in left ventrolateral tive dimensions to task performance (e.g., Anderson et al., 2003;
and dorsolateral PFC was greater for arousing than for neutral Hamann et al., 2002; Phelps and Anderson, 1997; Phelps et al.,
pictures. This finding suggests that the enhancing effect of emotion on 1998). However, the contribution of these factors to emotional
memory formation is partly due to an augmentation of PFC-mediated processing in other frontolimbic regions is not well understood,
strategic, semantic, and working memory operations. These results and the available evidence is contradictory. To address this
underscore the critical role of PFC in emotional evaluation and imbalance, the present functional MRI (fMRI) study focused on
memory, and disentangle the effects of arousal and valence across PFC the role of PFC regions in emotional processing.
regions associated with different cognitive functions. In particular, we investigated the effects of arousal and valence
D 2004 Elsevier Inc. All rights reserved. on emotional evaluation and emotional memory. Emotional eval-
uation refers to the perception and categorization of emotional
Keywords: Hemispheric asymmetry; Frontal lobes; Affect; Declarative
stimuli, and emotional memory refers to the modulatory effect of
memory; Semantic encoding; Self-referential processing; Neuroimaging
emotion on different stages of memory processing, including
encoding, consolidation, and retrieval. In the domain of emotional
evaluation, the amygdala is assumed to be involved in the rapid
Introduction detection of the basic emotional properties of incoming stimuli,
whereas PFC is assumed to be involved in higher-order emotional
The domain of cognitive neuroscience of emotion has grown evaluation processes, which operate in close interaction with other
dramatically during the last decade. As a result of this develop- cognitive functions and with behavioral goals (Davidson and Irwin,
1999). In the domain of emotional memory, the existing studies
* Corresponding author. Brain Imaging and Analysis Center, Room
have focused on the amygdala and identified arousal-mediated
163, Bell Building, PO Box 3918, Duke University, Durham, NC 27710. effects at encoding that predict subsequent memory (Cahill et al.,
Fax: +1-919-681-7033. 1996; Canli et al., 2000, 2002; Dolcos et al., 2003, 2004; Hamann
E-mail address: fdolcos@duke.edu (F. Dolcos). et al., 1999), but there is little understanding of the contribution of
Available online on ScienceDirect (www.sciencedirect.com.) other brain regions, such as PFC regions. Although it is assumed

1053-8119/$ - see front matter D 2004 Elsevier Inc. All rights reserved.
doi:10.1016/j.neuroimage.2004.05.015
F. Dolcos et al. / NeuroImage 23 (2004) 64–74 65

that activation during emotional evaluation tasks plays a role in irrespective of valence (Lane et al., 1997a,b,c; Reiman, 1997;
memory, few studies have explicitly examined the relationship Reiman et al., 1997; Schneider et al., 1995; Teasdale et al., 1999).
between emotional evaluation and memory. Thus, the overarching On the other hand, there is also evidence suggesting valence-
goal of the present study was to carefully investigate the contri- related specificity in medial PFC (e.g., George et al., 1995;
bution of the PFC to arousal and valence effects on emotional Paradiso et al., 1999b). In particular, medial PFC has been
evaluation and memory. associated with affiliative behaviors and appetitive or reward
Different PFC subregions are likely to make distinct contribu- circuits (e.g., Rolls, 2000). A recent metaanalysis of functional
tions to emotional evaluation, but information about this issue is neuroimaging studies of emotion (Wager et al., 2003) found that,
scarce. A basic anatomical distinction in this domain is between overall, medial PFC activity was associated with approach or
lateral and medial PFC regions. According to one prevailing view, appetitive tasks. To address this issue, the second goal of the
the role of lateral PFC regions in emotional evaluation is primarily present study was to determine whether the role of medial PFC in
related to valence. The valence hypothesis states that the left PFC emotional processing is primarily related to arousal or to valence,
is dominant in the processing of positive emotions, whereas the or whether there are subregions within medial PFC differently
right PFC is dominant in the processing of negative emotions involved in arousal and valence.
(Davidson, 1995; Davidson and Irwin, 1999). This hypothesis is Turning to emotional memory, the most basic phenomenon to
inspired by evidence from lesion literature and is mainly supported explain in this domain is why arousing events (both positive and
by electrophysiological evidence from EEG recordings. Neuropsy- negative) are better remembered than neutral events (Bradley et al.,
chological evidence shows that patients with left hemisphere 1992; Christianson, 1992). This effect has been attributed to the
lesions tend to experience negative emotions, such as sadness modulatory effect of the amygdala on the medial temporal lobe
(Morris et al., 1996; Paradiso et al., 1999a), whereas patients with (MTL) memory system (McGaugh et al., 2002), and this modula-
right hemisphere damage are biased toward experiencing positive tion hypothesis has been confirmed by functional neuroimaging
emotions, such as euphoria (e.g., Starkstein et al., 1989). The studies (Cahill et al., 1996; Canli et al., 2000, 2002; Dolcos et al.,
results of some electrophysiological studies are consistent with the 2003, 2004; Hamann et al., 1999; Kilpatrick and Cahill, 2003). For
valence hypothesis and support the idea that this valence-related example, we investigated this hypothesis using event-related fMRI
PFC lateralization may depend either on transiently induced and the subsequent memory paradigm (Paller and Wagner, 2002).
affective states or on stable personality traits (Aftanas et al., In this paradigm, memory performance on a subsequent memory
2001; Davidson, 1995; Davidson and Irwin, 1999; Tomarken et task is used to sort encoding items into two categories: remem-
al., 1992; Wheeler et al., 1993). bered versus forgotten. Greater encoding activity for remembered
However, electrophysiological studies do not provide an accu- than forgotten items, sometimes known as ‘‘the Dm (difference in
rate localization of the sources of these valence effects, and, memory) effect’’, is assumed to reflect successful encoding oper-
overall, the evidence supporting the valence hypothesis has been ations. Consistent with the modulation hypothesis, we found that
mixed. First, neither lesion nor electrophysiological evidence has the Dm effects in the amygdala and the MTL memory regions were
always been consistent with the valence hypothesis (e.g., Borod, greater for emotionally arousing pictures than for neutral pictures,
1992; Borod et al., 1998; Dolcos and Cabeza, 2002; Hagemann et and that the two Dm effects were more strongly correlated in the
al., 1998). Second, functional neuroimaging evidence is also case of arousing pictures than in the case of neutral pictures
inconclusive. Whereas some studies support the valence hypothesis (Dolcos et al., 2004).
(e.g., Canli et al., 1998), some studies do not report valence-related Although this evidence strongly links the memory-enhancing
hemispheric asymmetry in PFC (e.g., Baker et al., 1997; George et effect of emotion to an MTL mechanism, it does not exclude the
al., 1995; Lane et al., 1997a,b,c; Pardo et al., 1993; Teasdale et al., possibility that other brain regions, such as PFC, also play a major
1999). One possible reason why the results have been mixed is that role. In fact, in functional neuroimaging studies, PFC regions are as
arousal and valence are often not distinguished carefully (but see strongly associated with successful encoding operations as MTL
Canli et al., 1998); hence, valence effects might have been regions (e.g., Brewer et al., 1998; Paller and Wagner, 2002;
confounded with arousal effects. Thus, the first goal of the present Wagner et al., 1998). Also, the effects of several factors affecting
study was to investigate the valence hypothesis and identify the encoding success, such as organizational strategies and attention,
specific PFC regions involved, in conditions where positive and have been found to be mediated by changes in PFC activity
negative stimuli were matched in arousal and other potentially (Anderson et al., 2000; Fletcher et al., 1998; Kensinger et al.,
confounding factors were controlled. 2003). Moreover, studies using transcranial magnetic stimulation
As for the role of medial PFC regions in emotional evaluation, (TMS) have shown that PFC activity is actually necessary for
different hypotheses have been suggested. For instance, orbito- successful encoding (Epstein et al., 2002; Grafman and Wasser-
frontal areas of medial PFC have been linked to the rewarding mann, 1999; Rossi et al., 2001). Thus, it is quite likely that the
nature of stimuli (e.g., O’Doherty et al., 2001; Rolls, 2000), and enhancing effect of emotion on encoding is mediated not only by
anteromedial areas have been related to more personal and sub- MTL but also by PFC. Yet, very little is known about the role of
jective aspects of experiencing internal states (e.g., Frith and Frith, PFC on emotional memory formation (see, however, Canli et al.,
1999). Although medial PFC regions have been strongly associated 2002; Kilpatrick and Cahill, 2003). For example, it is uncertain if
with emotional processing, it is unclear whether the role of these the Dm effect in PFC is enhanced by emotion, similar to what we
regions is related to arousal or to valence. Given that medial PFC found in MTL (Dolcos et al., 2004). It is also unclear whether this
regions are systematically activated by emotional stimuli, regard- putative effect is due to arousal or valence, and which specific PFC
less of their valence (for a review, see Phan et al., 2002), PFC regions are involved. Thus, the third goal of the study was to
involvement could be attributed to its role in the processing of investigate the role of PFC in the formation of emotional memory.
arousal. This notion is consistent with evidence supporting medial The method we employed has two main features: (1) it
PFC involvement in processing emotionally arousing stimuli distinguishes between activity associated with emotional evalua-
66 F. Dolcos et al. / NeuroImage 23 (2004) 64–74

tion and emotional memory, and (2) it distinguishes between the saki et al., 2002). Also, given the evidence that dorsal – ventral PFC
effects of arousal and valence. Participants were scanned while regions are differently involved in the processing of visual stimuli
rating the pleasantness of arousing pictures (positive and negative) depending on their spatial content (e.g., Goldman-Rakic, 1995), it
and non-arousing pictures (neutral), and after scanning, they was important to determine that spatial/scene content did not differ
recalled the contents of the pictures. Stimuli were selected from between emotional and neutral stimuli. To investigate this idea, we
a standardized set of pictures that allows experimental control over asked 10 participants to rate our stimuli using a scene content scale
arousal and valence characteristics (Lang et al., 1997), which has (1 = no scene, 4 = very high scene content). The ratings for
been largely used in neuroimaging studies of emotion (e.g., Dolan emotional (1.86) and neutral (1.93) pictures were similar (T =
et al., 2000; Hamann et al., 1999, 2002; Lane et al., 1997a,b,c, 0.64, P > 0.5), suggesting an equivalent amount of scene/spatial
1999; Liberzon et al., 2000; Paradiso et al., 1999b; Taylor et al., information across the stimulus categories.
1998, 2000). Evaluation activity was measured by comparing
activity during picture rating to the baseline activity, and successful Experimental design
encoding activity was measured by comparing activity for subse-
quently remembered versus subsequently forgotten pictures (Dm The pool of 180 pictures was divided into six sets of 30 pictures
effect). Given that positive and negative pictures were both more (10 positive, 10 negative, and 10 neutral), which were randomly
arousing than neutral pictures, arousal effects should affect both assigned to six study blocks. Six different block orders were
positive and negative pictures. Given that positive and negative randomly assigned to the participants. To avoid the induction of
pictures were matched in arousal, differences between them should long-lasting mood states, the pictures within each block where
reflect valence effects rather than arousal effects. Thus, the effect of pseudo-randomized so that no more than two pictures of the same
arousal was defined as greater activity for both positive and valence were consecutively presented. Functional MR images were
negative pictures than for neutral pictures, and the effect of recorded while subjects viewed emotional and neutral pictures. The
valence, as differences between activity for positive and negative pictures were presented, using an LCD projector, to a screen
pictures. located behind the subjects’ crown that subjects could see via an
To summarize, we investigated three main issues. First, we angled mirror. Each picture was presented for 3 s and followed by a
investigated the valence hypothesis, and in particular, what specific 12-s fixation cross. Participants were instructed to experience any
left and right PFC subregions would be sensitive to valence effects. feelings or thoughts the pictures might elicit in them, and to rate
Second, we investigated the role of medial PFC in emotional each picture in a 3-point pleasantness scale (1 = negative, 2 =
evaluation, and specifically, whether activity in this region is neutral, 3 = positive). Rating the emotional valence of stimuli was
primarily sensitive to arousal or valence, or whether subregions employed because paying attention to emotional responses elicited
can be distinguished. Finally, we investigated the role of PFC in by various stimuli is associated with deep encoding, which results
emotional memory, and particularly, the relationship of Dm effects in better subsequent memory performance. This task also provides
with stimulus arousal and valence, and their localization within PFC. an estimation of subjects’ emotional responses. Nothing was
mentioned about a subsequent memory test, before or during the
encoding task, and hence learning was incidental. Incidental
Methods learning was preferred because intentional learning may interfere
with the experience of emotions, and because differences in
Subjects voluntary attention may complicate the interpretation of subse-
quent memory effects.
Sixteen young (25 + 4.6 years), right-handed, healthy women Forty-five minutes after the scanning session, subjects per-
participated in the study. Female participants were chosen because formed an unexpected cued-recall test conducted outside the
evidence suggests that they are more likely to display strong MRI suite. Subjects were provided with one- or two-word written
physiological responses to emotional stimuli (Lang et al., 1993) cues for each picture (e.g., snake, building, skydivers), and had to
and report more intense emotional experiences (Shields, 1991) than describe in writing, and in as much detail as they could, the
men. All participants consented to a protocol approved by Duke pictures that they remembered. Similar to the procedure employed
University Institutional Review Board. in our previous ERP study (Dolcos and Cabeza, 2002), participants
were asked to provide enough relevant details (e.g., about the
Materials number of elements, color, action, etc.) so that an outsider could
identify each picture and discriminate it from similar studied
Stimuli consisted of 60 positive, 60 negative, and 60 neutral pictures (e.g., a brown snake facing viewer vs. several small green
pictures selected from the International Affective Picture System snakes). The test lasted until participants could not recall any
(IAPS) picture database (Lang et al., 1997), on the basis of their additional pictures or until a maximum of 50 min had elapsed. Two
normative arousal and valence scores. The mean arousal scores (1 = raters were involved in scoring participants’ responses, and only
calm, 9 = excited) were 6.0 for positive (SD = 2.2), 6.15 for those pictures whose description was detailed enough to allow both
negative (SD = 2.2), and 3.15 for neutral pictures (SD = 2.0). Thus, identification and discrimination were classified as remembered.
positive and negative pictures had similar high arousal scores,
whereas neutral pictures had low arousal scores. The mean valence MRI data acquisition
scores (1 = negative, 5 = neutral, 9 = positive) were 7.1 for positive
(SD = 1.7), 2.3 for negative (SD = 1.5), and 5.2 for neutral (SD = Anatomical scanning
1.4). To equate the emotional and neutral categories for visual Neuroimaging was performed using a 1.5 T GE scanner. A T1-
complexity and content (e.g., human presence), the IAPS pictures weighted sagittal localizer series was first acquired. The anterior
were supplemented with neutral pictures from other sources (Yama- (AC) and posterior commissures (PC) were identified in the
F. Dolcos et al. / NeuroImage 23 (2004) 64–74 67

midsagittal slice, and 34 contiguous oblique slices were prescribed results, with a focus on lateral and medial cortices, excepting motor
parallel to the AC-PC plane. High-resolution T1-weighted struc- and cingulate regions. The MTL results were reported in different
tural images were acquired with a 450-ms TR (repetition time), a 9- manuscripts (Dolcos et al., 2003, 2004). Conjunction analyses were
ms TE (echo time), a 24-cm FOV (field of view), a 2562 matrix, used to identify brain regions more activated in two conditions (e.g.,
and a slice thickness of 3.75 mm. A second series of 46 oblique positive and negative) than in a third condition (e.g., neutral). This
T1-weighted images perpendicular to the AC-PC was then ac- was done using the ImCalc feature in SPM, and according to the
quired using the same imaging parameters. following formula: [(Condition 1 T score > 2.01) ! (Condition 2 T
score > 2.01)]. This procedure yields a mask containing only those
Functional scanning voxels that were significantly activated above T = 2.01 (P = 0.0316)
Thirty-four contiguous gradient-echo echoplanar images (EPIs) in each and both contrasts. The probability of finding a voxel that is
sensitive to blood oxygen level dependent (BOLD) contrast were independently significant in each and both contrasts (i.e., the joint
acquired parallel to the AC-PC plane, using the same slice probability) can be estimated by multiplying the probabilities for
prescription described above for the near-axial structural images. each contrast: 0.0316 ! 0.0316 = P < 0.001 (e.g., Allan et al., 2000;
The EPIs were acquired with a 3-s TR, 40-ms TE, one radio Cabeza et al., 2002).
frequency excitation, 24-cm FOV, 642 image matrix, and a 90j flip The resulting conjunction masks provided information about
angle. Slice thickness was 3.75 mm, resulting in cubic 3.75 mm3 the extent of the overlapping activations associated with the
isotropic voxels. conditions involved in the conjunction, but not about the intensity
of the overlapping activity. For this, conjunction T maps were
fMRI data analysis calculated by multiplying the T values for the conditions of interest
in the overlapping regions. The conjunction T maps were calcu-
Image preprocessing lated according to the following formula: [(Condition 1 T score) !
Image preprocessing and statistical analyses were performed (Condition 2 T score) ! (Conjunction mask of Condition 1 and
using SPM99 (http://www.fil.ion.ucl.ac.uk/spm/). Functional Condition 2)]. These are the T values we report.
images were corrected for acquisition order and realigned to For emotional evaluation, the effect of arousal was measured
correct for motion artifacts. Anatomical images were coregistered as greater activity (compared to baseline) for arousing stimuli
with the first functional images for each subject, and then both (positive and negative pictures) than for non-arousing stimuli
anatomical and functional images were spatially normalized to a (neutral pictures). This was done by identifying regions that
standard stereotactic space, using the Montreal Neurological Insti- showed both (1) greater activity for positive than for neutral
tute (MNI) templates implemented in SPM99. Subsequently, pictures, and (2) greater activity for negative than for neutral
functional images were spatially smoothed using an 8-mm isotro- pictures. That is, the effect of arousal on emotional evaluation was
pic Gaussian kernel. defined as [(positive > neutral) conj (negative > neutral)]. The
effect of valence was measured as significant differences between
Statistical analyses positive and negative pictures. Since the valence scales had
Statistical analyses were separately performed to assess both neutral as an intermediate value between positive and negative,
emotion-related (emotional evaluation) and memory-related differ- we further required that regions associated with positive or
ences (emotional encoding) between emotional and neutral pic- negative valence had to be more activated in these conditions
tures. The images were defined as unpleasant, neutral, or pleasant than in the neutral condition. That is, the effect of positive valence
based on the IAPS ratings. The use of the IAPS score was justified on emotional evaluation was defined as [(positive > negative) conj
by two reasons. First, different from the rating scores, which due to (positive > neutral)] and the effect of negative valence on
technical limitations during scanning (i.e., the response box had emotional evaluation was defined as [(negative > positive) conj
only three response options) did not allow fine evaluations of the (negative > neutral)].
subjects’ emotional response, the IAPS scores are based on more The same kind of analyses were performed for emotional
sophisticated methods of assessing emotional arousal and valence, encoding, except that instead of using activity during picture rating
and allow much finer dissociations. Second, although some sub- compared to baseline, we used differences in activity between
jects classified some of the images differently than the norms, the remembered and forgotten items (Dm = activity for remembered
high correlation between the average picture scores as rated by the pictures " activity for forgotten pictures). Dm activity was sepa-
subjects and the IAPS valence scores (R = 0.9, P < 0.0001) suggest rately calculated for each picture category (e.g., Dm positive =
that participants’ classification was highly consistent with the positive remembered " positive forgotten), and then the effects of
normative data. Therefore, the latter scores were used to dissociate arousal and valence were identified by comparing the three types of
the effect of arousal and valence on brain activity associated with Dm activity. The effect of arousal on emotional memory was
emotional evaluation and emotional memory formation. For each defined as [(Dm positive > Dm neutral) conj (Dm negative > Dm
subject, task-related activity was identified by convolving a vector neutral)]. The effect of positive valence on emotional memory was
of the onset times of the stimuli with a synthetic hemodynamic defined as [(Dm positive > Dm negative) conj (Dm positive > Dm
response (HDR) and its temporal derivative. The general linear neutral)] and the effect of negative valence on emotional memory
model, as implemented in SPM99, was used to model the effects of was defined as [(Dm negative > Dm positive) conj (Dm negative >
interest and other confounding effects (e.g., session effects and Dm neutral)]. To make sure that the differences between Dms
magnetic field drift). Functional images were proportionally scaled occurred due to positive activations in the condition of interest and
to the whole-brain signal. were not driven by deactivations in the other conditions, the
Group analyses were conducted using random-effects models to conjunction maps were inclusively masked with the activation
assess the effect of arousal and valence on emotional evaluation and maps showing the main effect of memory (Dm) for the condition
emotional encoding. In the present manuscript, we report the PFC of interest at P < 0.05. For instance, for the latter comparison (i.e.,
68 F. Dolcos et al. / NeuroImage 23 (2004) 64–74

[(Dm negative > Dm positive) conj (Dm negative > Dm neutral)]), Memory performance
the resulting map was masked with the activation map for Dm Arousing pictures, both positive and negative, were better
negative. Thus, the final conjunction map contained only the voxels recalled than neutral pictures. Out of 60 pictures per category,
that showed a significant Dm for negative pictures. participants recalled an average of 52% positive, 53% negative,
The bar graphs of fMRI activations were examined by extract- and 38% neutral pictures (SDs were 4.5, 4.8, and 4.8, respectively).
ing the mean effect size from the peak voxel of each region, as An ANOVA yielded a significant picture type effect (F(2,15) =
identified by the SPM conjunction analyses for each condition of 41.21, P < 0.0001), and post-hoc contrasts showed that recall of
interest and subject. The data extraction was accomplished using positive and negative pictures was similar (P > 0.05), and greater
SPM99. The xyz coordinates provided by SPM, which are in than recall of neutral pictures (P < 0.0001).
Montreal Neurological Institute (MNI) brain space, were converted
to xyz coordinates in Tailarach and Tournoux’s brain space (Talai- fMRI results
rach and Tournoux, 1988).
The analyses on activity associated with emotional evaluation
(rating-baseline) yielded dissociable PFC regions showing effects
Results of arousal and valence. In lateral PFC, the main goal was to test
the valence hypothesis. Consistent with this hypothesis, a va-
Behavioral results lence-related hemispheric asymmetry was found: a left dorsolat-
eral PFC region (BA 8/9; xyz = !41, 21, 48; T = 8.63) showed
Valence ratings an effect of positive valence, whereas a right ventrolateral PFC
The average valence scores (1 = negative, 2 = neutral, 3 = region (BA 47; xyz = 49, 33, !2; T = 17.03) showed an effect of
positive) as rated by the participants in the scanner were 1.14 (SD = negative valence. As illustrated by Fig. 1, the left PFC region was
0.16) for negative pictures, 2.18 (SD = 0.40) for neutral pictures, more activated for positive than for negative pictures, whereas the
and 2.64 (SD = 0.26) for positive pictures. All pairwise compari- right PFC region was more activated for negative than for
sons were significant (P < 0.0001). Thus, the subjects’ rating scores positive pictures. However, this asymmetry was not present in
were consistent with those provided in the IAPS norms (Lang et al., the entire lateral PFC: other dorsolateral PFC areas (BA 9)
1997). Further validating this consistency, the correlation between showed a bilateral (xyz = 49, 9, 24/!38, 9, 24; T = 20.02/T =
our subjects’ average scores and the normed IAPS scores of the 15.41) effect of negative valence. The effect of arousal in lateral
pictures used in the present study was highly significant (R = 0.90, PFC was evident in an area of the right inferior frontal gyrus (BA
P < 0.0001). 47; xyz = 30, 18, !17; T = 9.57).

Fig. 1. Activity in lateral PFC showed evidence for the valence hypothesis. In the left hemisphere, a dorsolateral PFC region (BA 8/9) was more activated
during evaluation of positive pictures than during evaluation of negative pictures, whereas in the right hemisphere, ventrolateral regions (BA 47) were more
activated for negative than for positive pictures. The upper panels show the activation maps overlapped on high-resolution coronal anatomical images, and the
bar graphs at the bottom show the effect size as extracted from the peak-voxels identified in the conjunction analyses (see Methods). The numbers at the left-
bottom corner of the upper panels (e.g., y = 16) indicate the coordinate in MNI space. The color bar located between the upper panels indicates the conjunction
T values. L = left, R = right; Pos = positive, Neg = negative, Neu = neutral; BA = Brodmann Area.
F. Dolcos et al. / NeuroImage 23 (2004) 64–74 69

Fig. 2. Activity in medial PFC identified dissociable regions associated with arousal and positive valence. Dorsomedial PFC (BA 9) activity was sensitive to
arousal (Pos and Neg > Neu), whereas ventromedial PFC (BA 10) activity was sensitive to positive valence (Pos > Neg and Neu). L = left, R = right; Pos =
positive, Neg = negative, Neu = neutral; BA = Brodmann Area.

In medial PFC, a dorsal – ventral distinction was found during more activated during evaluation of both pleasant and unpleasant
emotional evaluation in which arousal effects were in a dorsal pictures than during evaluation of neutral pictures, whereas the
region (BA 9; xyz = !4, 52, 19; T = 11.45), whereas valence effects ventral regions were more activated during the evaluation of
were found in more ventral locations—orbitofrontal and antero- positive pictures.
medial cortices (BA 10; xyz = 0, 58, !10/0, 58, 4; T = 14.72/T = The analysis of activity associated with emotional memory
30.2). As illustrated by Fig. 2, the dorsomedial PFC region was (Dm = remembered ! forgotten) also yielded PFC regions

Fig. 3. Arousal enhanced successful encoding activity (Dm) in left PFC. Compared to the Dm for neutral pictures, the Dm for arousing pictures (both positive
and negative) was greater in left ventrolateral (BA 47) and dorsolateral (BA 9/6) PFC regions. L = left, R = right; Pos = positive, Neg = negative, Neu = neutral;
BA = Brodmann Area; Dm = remembered ! forgotten.
70 F. Dolcos et al. / NeuroImage 23 (2004) 64–74

showing effects of arousal and valence. The areas showing effects (Canli et al., 1998), which could not determine whether hemi-
of arousal were found only in the lateral PFC, and included both spheric asymmetries were stimulus-specific or reflected sustained
ventral (BA 47; xyz = !49, 29, !1; T = 5.48) and dorsal (BA 9/6; changes in affective states, the present study provides evidence
xyz = !38, 2, 31; T = 8.64) locations. As illustrated by Fig. 3, in supporting the valence hypothesis using an event-related design in
these lateral PFC regions, the Dm for positive pictures and the Dm which stimuli were randomized during scanning.
for negative pictures were both greater than the Dm for neutral Second, the present results not only demonstrated a valence-
pictures. The average effect size values for remembered and related hemispheric asymmetry but also identified the specific left
forgotten pleasant, unpleasant, and neutral items, respectively, and right PFC regions associated with positive and negative
are as follows: BA 47 (!0.22, !0.46, !0.13, !0.30, !0.31, valence. The valence hypothesis has been primarily supported
and !0.22) and BA 9/6 (!0.24, !0.41, !0.09, !0.22, !0.31, and using electrophysiological methods (Aftanas et al., 2001; David-
!0.16). The areas showing effects of valence included medial (BA son, 1995; Tomarken et al., 1992; Wheeler et al., 1993), which do
9; xyz = !11, 38, 29; T = 6.41) and lateral (BA 45; xyz = !49, 22, not allow a good localization of neural sources. Evidence for the
3; T = 8.51) locations, both showing an effect of positive valence. valence hypothesis was also found using fMRI (Canli et al., 1998)
but using analyses that collapsed activity over a whole hemisphere.
In contrast, the present result shows that the left PFC region
Discussion specifically sensitive to positive valence is in dorsolateral cortex
(middle frontal gyrus; BA 8/9) whereas the right PFC region
The present study yielded three main findings relevant for particularly sensitive to negative valence is in ventrolateral cortex
understanding PFC contributions to emotional evaluation and (inferior frontal gyrus; BA 47). A more inferior sector of BA9, in
memory. First, consistent with the valence hypothesis, during contrast, did not show such asymmetry effects, and was sensitive to
emotional evaluation, specific left dorsolateral PFC areas showed negative valence bilaterally.
greater activity for positive than for negative pictures, whereas What are the implications of this region-specific hemispheric
right ventrolateral PFC areas showed the converse pattern. Second, asymmetry in the processing of emotional valence? As described
also during emotional evaluation, dorsomedial PFC activity was above, lesion, electrophysiological, as well as the available neuro-
sensitive to arousal (greater activation for both positive and imaging evidence supporting the valence hypothesis lacks the
negative pictures relative to neutral ones), whereas ventromedial regional specificity necessary to link activity in the regions
PFC activity was sensitive to valence (greater activation for identified here with possible differential involvement in processing
positive pictures relative to negative ones). Finally, demonstrating emotional valence. One possibility is to broadly explain the role of
the role of lateral PFC in emotional memory, arousal enhanced these regions based on the available neuroimaging evidence
successful encoding activity in left ventrolateral and dorsolateral concerning their involvement in various tasks. Another possibility
PFC. These results provide evidence for multiple, regionally is to link our findings with more specific evidence concerning
specific emotional influences on PFC function. The implications dorsal – ventral dissociations in the lateral PFC. As concerning the
of the findings are discussed in separate sections below. former, neuroimaging evidence has associated the dorsolateral
sectors of PFC, including the BAs 8 and 9 with a variety of tasks,
Lateral PFC showed a hemispheric asymmetry consistent with the involving perceptual, attentional, imagistic, and mnemonic opera-
valence hypothesis tions, but typically they have been associated with working
memory tasks (see Cabeza and Nyberg, 2000, for a review). In
As illustrated by Fig. 1, a left PFC region was more activated addition, portions of BA 8 are thought to be part of the so-called
for positive than for negative pictures and a right PFC region frontal eye field, although the overlap between this region in
showed the converse pattern. This finding is consistent with other human and nonhuman primates is still controversial (Koyoma et
evidence supporting the valence hypothesis (Aftanas et al., 2001; al., 2004). Interestingly, related to its role in emotional processing,
Canli et al., 1998; Davidson, 1995; Tomarken et al., 1992; Wheeler BA 8 has been identified in tasks involving rating the pleasantness
et al., 1993), but it extends this evidence in two ways. of facial stimuli (Nakamura et al., 1998). As concerning the role of
First, the present finding demonstrates hemispheric asymmetry ventrolateral PFC regions, including BA 47, their function has
effects predicted by the valence hypothesis under conditions in been associated with semantic memory operations, as well as with
which positive and negative stimuli were matched for arousal and interference control and inhibitory processes (Miller and Cohen,
visual properties, and in which the effects of valence could be 2001; Smith and Jonides, 1999; see also Cabeza and Nyberg,
distinguished from the effects of arousal on a trial-by-trial basis. As 2000). Specifically related to the involvement in emotional pro-
noted before (see Methods), positive and negative stimuli had cessing, right ventrolateral PFC has been implicated in the inhibi-
similar normative scores of arousal and were equivalent in terms of tion of negative emotions (Petrovic et al., 2002).
complexity, presence of human figures, and other lower-level Turning to the evidence concerning the asymmetry in the
visual features. Participant ratings obtained on-line during encod- dorsal – ventral dimension, two main views have been identified
ing confirmed the valence manipulation and were highly correlated regarding the role of lateral PFC. According to one view, dorso-
with normative valence scores. Moreover, the inclusion of neutral lateral PFC regions are more involved in manipulating working
pictures allowed us to disentangle valence effects from arousal memory contents whereas ventrolateral PFC regions are more
effects. Since arousal effects are by definition common to positive involved in simple maintenance operations (D’Esposito et al.,
and negative pictures, these effects should appear as greater 2000; Owen et al., 1999; Petrides, 1995). Another view (Davidson
activity for both positive and negative pictures than for neutral and Irwin, 1999) specifically relates the function of lateral PFC to
pictures. As illustrated by Fig. 1, this was not the case in either the the role of emotion in guiding and organizing behavior in a
left or the right PFC regions comprising valence-related hemi- motivationally consistent manner (Frijda, 1988). According to this
spheric asymmetry. Additionally, in contrast to previous findings view, dorsolateral PFC is involved in a particular form of working
F. Dolcos et al. / NeuroImage 23 (2004) 64–74 71

memory—affective working memory—responsible for the repre- a bias toward detecting and processing positive stimuli. Given the
sentation of goal-related emotional states, whereas the ventrolateral fact that, overall, the medial PFC in the present study was more
sector of the PFC is involved in the simple representation of activated for positive than for negative pictures, this interpretation
elementary emotional states (Davidson and Irwin, 1999). may also apply to the current finding.
Combining these ideas, one may speculate that the left dorso- An alternative account is that the sensitivity of medial PFC to
lateral PFC activity for positive stimuli reflects the maintenance positive valence reflected greater self-engagement in the process-
and manipulation of positive information in working memory ing of positive pictures compared to negative and neutral pictures.
during the valence-rating task, whereas the right ventrolateral A number of recent neuroimaging studies associated medial PFC
PFC activity reflects the inhibition (avoidance) of negative infor- activity with self-referential processing (e.g., Cabeza et al., in
mation. Of course, these ideas are ad hoc and require independent press; Frith and Frith, 1999; Gusnard et al., 2001; Kelley et al.,
confirmation. For instance, since other dorsolateral PFC regions 2002). For instance, in a review of the literature, Frith and Frith
were associated with evaluation of negative pictures, it would be (1999) suggested that activity in ventral medial PFC was specif-
important to clarify that working memory-related activity in the left ically associated with emotional aspects of self-processing. It
PFC region specifically associated with positive pictures is related should be noted, however, that the appetitive and self-engagement
to the maintenance of appetitive goals (e.g., detection and process- accounts of the present medial PFC activation are compatible. For
ing of positive stimuli). example, participants could have been more likely to relate the
pictures to their own self and life in the case of positive pictures
Dissociable regions of medial PFC were associated with arousal than in the case of negative and neutral pictures. The medial PFC
and positive valence region is frequently activated in functional neuroimaging studies of
autobiographical memories. For example, we found this region to
As illustrated by Fig. 2, dorsomedial PFC activity was sensitive be more activated during the recognition of photographs taken by
to arousal, whereas ventromedial PFC activity was sensitive to oneself than during the recognition of photographs taken by others
positive valence. Previous evidence is consistent with the idea of (Cabeza et al., in press).
possible segregation among medial PFC regions with respect to
their involvement in emotional processing, but it is unclear whether Arousal enhanced successful encoding activity in left PFC
the role of these regions is related to arousal or to valence. On the
one hand, a metaanalysis of neuroimaging studies of emotion Emotional arousal enhanced successful encoding (Dm) activity
suggests that medial PFC regions are systematically activated by in lateral PFC. As illustrated by Fig. 3, compared to the Dm for
emotional stimuli, regardless of their valence (for a review, see neutral pictures, the Dm for arousing pictures was greater in left
Phan et al., 2002). This finding suggests a nonspecific involvement ventrolateral and dorsolateral PFC. These findings suggest that the
of medial PFC in emotional processing, probably mediated by enhancing effect of emotion on memory formation (i) is partly
arousal, and is consistent with some neuroimaging studies report- mediated by changes in PFC activity, (ii) is mainly related to arousal,
ing dorsomedial activations associated with the processing of and (iii) may involve an amplification of semantic processing and
emotional stimuli, regardless of their valence (Lane et al., working memory operations mediated by lateral PFC regions.
1997a,b,c; Reiman, 1997; Reiman et al., 1997; Teasdale et al., Research on the neural bases of the enhancing effect of emotion
1999). On the other hand, another metaanalysis (Wager et al., on memory formation has emphasized the role of the amygdala and
2003) found that, overall, medial PFC activity was associated with its interactions with MTL memory regions (Cahill et al., 1996;
approach or appetitive tasks. Although approach-withdrawal and Canli et al., 2000, 2002; Dolcos et al., 2003, 2004; Hamann et al.,
valence dimensions are not identical, they do overlap, and thus this 1999; Kilpatrick and Cahill, 2003; McGaugh et al., 2002). The
finding suggests possible valence-related specificity in medial PFC present results expand this line of research by showing that the
function (see also George et al., 1995; Paradiso et al., 1999b). enhancing effect of emotion on memory formation is also mediated
However, comparisons across studies are complicated by differ- by changes in PFC activity. However, the effects of emotion on
ences in stimuli, methods, and participants. MTL and PFC are likely to enhance different memory mecha-
In the present study, we demonstrated a dorsal – ventral disso- nisms. Given the functions typically attributed to these regions
ciation in the medial PFC, within-subjects and under controlled (Moscovitch, 1992; Simons and Spiers, 2003), it is reasonable to
conditions. By identifying specific medial PFC regions sensitive to assume that in MTL, emotion enhances the storage and consoli-
arousal and valence, this finding complements and reconciles dation of memory representations, whereas in PFC, it enhances
previous functional neuroimaging evidence suggesting nonspecific strategic encoding processes.
(Lane et al., 1997a,b,c; Reiman, 1997; Reiman et al., 1997; A second implication of the present findings is that the enhanc-
Teasdale et al., 1999) versus valence-specific involvement of ing effect of emotion on memory formation is primarily related to
medial PFC during emotional processing (George et al., 1995; arousal rather than to valence. In our MTL study (Dolcos et al.,
Paradiso et al., 1999b). This finding is also consistent with the 2003, 2004), we also found that the Dm increase was related to
results of an ERP study (Dolcos and Cabeza, 2002) where we arousal rather than to valence. Thus, although valence-related Dm
found arousal versus valence dissociations at midline frontal increases also occur, it seems fair to conclude that arousal is the
electrodes. Although spatial resolution of ERP did not allow us main factor modulating the neural mechanisms of memory forma-
to separate these effects topographically, it allowed dissociations in tion. This conclusion is consistent with our behavioral results,
timing: there was a faster effect (500 – 800 ms) of positive valence which showed that compared to non-arousing neutral pictures,
(positive > negative = neutral) and a delayed effect (after 800 ms) memory is better for arousing positive and arousing negative
of arousal (positive = negative > neutral). Since the valence-related pictures, with no significant difference between these two condi-
ERP effect occurred in an earlier time-window (see also Cuthbert tions. Thus, from the point of view of memory, a negative event can
et al., 2000; Dillon et al., submitted), we interpreted it as reflecting be as effective as a positive event (see also Talarico et al., in press).
72 F. Dolcos et al. / NeuroImage 23 (2004) 64–74

The specific PFC regions where the Dm was increased by and to valence, and that they play an important role in the
arousal suggest that arousing events are better remembered because evaluation of emotional stimuli and in processes that lead to
they receive deeper semantic processing and working memory better memory for emotional events.
resources during encoding. The Dm was enhanced by arousal in
left ventrolateral (BA 47) and dorsolateral (BA 9/6) PFC regions
(see Fig. 3). The left ventrolateral region is an area that many
Acknowledgments
functional neuroimaging studies have associated with encoding
processes (for a review, see Cabeza and Nyberg, 2000), including
We thank David Beckmann for assistance with data analysis.
event-related fMRI studies using the subsequent memory paradigm
This study was supported by NIH grants R01 AG19731 (RC) and
(Brewer et al., 1998; Kirchhoff et al., 2000; Paller and Wagner,
R01 DA14094 (KSL), and a NARSAD Young Investigator Award
2002; Wagner et al., 1998). Since the role of this region in encoding
(KSL). FD was supported by a Chia PhD Scholarship and a
is generally attributed to semantic processing (Kapur et al., 1996;
Dissertation Fellowship from the University of Alberta (Canada),
Poldrack et al., 1999; Shallice et al., 1994), the present finding
and a Research Assistantship from Duke University (USA).
suggests that arousal facilitated successful encoding by increasing
semantic processing of the information in the pictures. It is possible
that arousal also enhanced perceptual encoding processes mediated
by right PFC (Brewer et al., 1998). However, since picture memory References
was tested using verbal recall, the effects of arousal on perceptual
encoding were probably not detected. To detect such effects, it Aftanas, L., Varlamov, A., Pavlov, S., Makhnev, V., Reva, N., 2001. Event-
would be necessary to test memory with a nonverbal task, such as related synchronization and desynchronization during affective process-
picture recognition (e.g., Brewer et al., 1998). ing: emergence of valence-related time-dependent hemispheric asym-
Finally, the effect of arousal on left dorsolateral PFC is likely metries in theta and upper alpha band. Int. J. Neurosci. 110 (3 – 4),
to reflect the augmentation of the working memory processes 197 – 219.
typically associated with this region (D’Esposito et al., 2000; Allan, K., Dolan, R.J., Fletcher, P.C., Rugg, M.D., 2000. The role of the
Owen et al., 1999; Petrides, 1995). Thus, it is possible that the right anterior prefrontal cortex in episodic retrieval. NeuroImage 11 (3),
217 – 227.
contents of arousing events not only receive deeper semantic
Anderson, N.D., Iidaka, T., Cabeza, R., Kapur, S., McIntosh, A.R., Craik,
processing but are also maintained longer or manipulated more
F.I., 2000. The effects of divided attention on encoding- and retrieval-
intensely in working memory, leading to better retention. It should related brain activity: a PET study of younger and older adults. J. Cogn.
be noted that this interpretation is compatible with the idea of Neurosci. 12 (5), 775 – 792.
dorsolateral PFC involvement during emotional evaluation of Anderson, A.K., Christoff, K., Stappen, I., Panitz, D., Ghahremani, D.G.,
positive pictures, since the specific regions involved during Glover, G., Gabrieli, J.D., Sobel, N., 2003. Dissociated neural repre-
positive evaluation versus successful encoding of arousing pic- sentations of intensity and valence in human olfaction. Nat. Neurosci. 6
tures are slightly different. It is possible that in one case, the (2), 196 – 202.
involvement of working memory operations is related to the Baker, S.C., Frith, C.D., Dolan, R.J., 1997. The interaction between mood
maintenance of appetitive-goals, whereas in the latter, the main- and cognitive function studied with PET. Psychol. Med. 27 (3), 565 – 578.
Borod, J.C., 1992. Interhemispheric and intrahemispheric control of emo-
tenance or manipulation of emotionally arousing information leads
tion: a focus on unilateral brain damage. J. Consult. Clin. Psychol. 60,
to better subsequent memory.
339 – 348.
Borod, J.C., Obler, L.K., Erhan, H.M., Grunwald, I.S., Cicero, B.A.,
Welkowitz, J., Santschi, C., Agosti, R.M., Whalen, J.R., 1998. Right
Conclusions hemisphere emotional perception: evidence across multiple channels.
Neuropsychology 12 (3), 446 – 458.
Using an fMRI paradigm that distinguished between activity Bradley, M.M., Lang, P.J., 1994. Measuring emotion: the self-assessment
related to emotional evaluation and emotional memory and manikin and the semantic differential. J. Behav. Ther. Exp. Psychiatry
between the effects of arousal and valence, the present study 25 (1), 49 – 59.
yielded three main results. First, during emotional evaluation, Bradley, M.M., Greenwald, M.K., Petry, M.C., Lang, P.J., 1992. Remem-
bering pictures: pleasure and arousal in memory. J. Exp. Psychol.,
PFC activity showed a hemispheric asymmetry consistent with
Learn, Mem. Cogn. 18 (2), 379 – 390.
the valence hypothesis. A left dorsolateral PFC region was
Brewer, J.B., Zhao, Z., Desmond, J.E., Glover, G.H., Gabrieli, J.D., 1998.
sensitive to positive valence, possibly reflecting the maintenance Making memories: brain activity that predicts how well visual experi-
of positive information in working memory, whereas a right ence will be remembered. Science 281, 1185 – 1187.
ventrolateral PFC region was sensitive to negative valence, Cabeza, R., Nyberg, L., 2000. Imaging cognition: II. An empirical review
possibly reflecting the inhibition of negative information. Second, of 275 PET and fMRI studies. J. Cogn. Neurosci. 12 (1), 1 – 47.
dorsomedial PFC activity was sensitive to arousal, whereas Cabeza, R., Dolcos, F., Graham, R., Nyberg, L., 2002. Similarities and
ventromedial PFC activity was sensitive to positive valence, differences in the neural correlates of episodic memory retrieval and
possibly reflecting the involvement of these regions in general working memory. NeuroImage 16 (2), 317 – 330.
processing of emotional information (dorsomedial PFC), and self- Cabeza, R., Prince, S.E., Daselaar, S.M., Greenberg, D.L., Budde, M.,
Dolcos, F., LaBar, K.S., Rubin, D.C., 2004. Brain activity during epi-
awareness or appetitive behavior (ventromedial PFC). Finally,
sodic retrieval of autobiographical and laboratory events: an fMRI study
successful encoding activity was enhanced by arousal in left
using a novel photo paradigm. J. Cogn. Neurosci. (in press).
ventrolateral and dorsolateral PFC regions, possibly reflecting Cahill, L., Haier, R.J., Fallon, J., Alkire, M.T., Tang, C., Keator, D., Wu, J.,
an enhancement of strategic, semantic, and working memory McGaugh, J.L., 1996. Amygdala activity at encoding correlated with
operations. Although further research is required, these findings long-term, free recall of emotional information. Proc. Natl. Acad. Sci.
strongly suggest that different PFC regions are sensitive to arousal U. S. A. 93, 8016 – 8021.
F. Dolcos et al. / NeuroImage 23 (2004) 64–74 73

Canli, T., Desmond, J.E., Zhao, Z., Glover, G., Gabrieli, J.D., 1998. Hemi- Kapur, S., Tulving, E., Cabeza, R., McIntosh, A.R., Houle, S., Craik, F.I.,
spheric asymmetry for emotional stimuli detected with fMRI. Neuro- 1996. The neural correlates of intentional learning of verbal materials: a
Report 9 (14), 3233 – 3239. PET study in humans. Conit. Brain Res. 4, 243 – 249.
Canli, T., Zhao, Z., Brewer, J., Gabrieli, J.D.E., Cahill, L., 2000. Event- Kelley, W.M., Macrae, C.N., Wyland, C.L., Caglar, S., Inati, S., Heather-
related activation in the human amygdala associated with later memory ton, T.F., 2002. Finding the self? An event-related fMRI study. J. Cogn.
for individual emotional experience. J. Neurosci. 20 (RC99), 1 – 5. Neurosci. 14 (5), 785 – 794.
Canli, T., Desmond, J.E., Zhao, Z., Gabrieli, J.D.E., 2002. Sex differences Kensinger, E.A., Clarke, R.J., Corkin, S., 2003. What neural correlates
in the neural basis of emotional memories. Proc. Natl. Acad. Sci. U. S. A. underlie successful encoding and retrieval? A functional magnetic res-
99 (16), 10789 – 10794. onance imaging study using a divided attention paradigm. J. Neurosci.
Christianson, S.-A., 1992. The Handbook of Emotion and Memory: Re- 23 (6), 2407 – 2415.
search and Theory. Lawrence Erlbaum Associates, Hillsdale, NJ. Kilpatrick, L., Cahill, L., 2003. Amygdala modulation of parahippocampal
Cuthbert, B.N., Schupp, H.T., Bradley, M.M., Birbaumer, N., Lang, P.J., and frontal regions during emotionally influenced memory storage.
2000. Brain potentials in affective picture processing: covariation with NeuroImage 20 (4), 2091 – 2099.
autonomic arousal and affective report. Biol. Psychol. 52, 95 – 111. Kirchhoff, B.A., Wagner, A.D., Maril, A., Stern, C.E., 2000. Prefrontal –
Davidson, R.J., 1995. Cerebral asymmetry, emotion and affective style. In: temporal circuitry for episodic encoding and subsequent memory.
Davidson, R.J., Hugdahl, K. (Eds.), Brain Asymmetry MIT Press, Cam- J. Neurosci. 20 (16), 6173 – 6180.
bridge, MA, pp. 361 – 387. Koyoma, M., Hasegawa, I., Osada, T., Adachi, Y., Nakahara, K., Miya-
Davidson, R.J., Irwin, W., 1999. The functional neuroanatomy of emotion shita, Y., 2004. Functional magnetic resonance imaging of macaque
and affective style. Trends Cogn. Sci. 3 (1), 11 – 20. monkeys performing visually guided saccade tasks: comparison of cor-
D’Esposito, M., Postle, B.R., Rypma, B., 2000. Prefrontal cortical contri- tical eye field with humans. Neuron 41 (5), 895 – 907.
butions to working memory: evidence from event-related fMRI studies. Lane, D.R., Nadel, L., 2000. Cognitive Neuroscience of Emotion. Oxford
Exp. Brain Res. 133 (3 – 11). Univ. Press, Inc., New York.
Dillon, D.G., Cooper, J.J., Grent-t-Jong, T., Woldorff, M.G., LaBar, K.S. Lane, R.D., Fink, G.R., Chau, P.M., Dolan, R.J., 1997a. Neural activation
Dissociation of event-related potentials indexing arousal, valence, and during selective attention to subjective emotional responses. NeuroRe-
semantic cohesion during emotional stimulus encoding, submitted. port 8 (18), 3969 – 3972.
Dolan, R.J., Lane, L., Chua, P., Fletcher, P., 2000. Dissociable temporal Lane, R.D., Reiman, E.M., Ahern, G.L., Schwartz, G.E., Davidson, R.J.,
lobe activations during emotional episodic memory retrieval. Neuro- 1997b. Neuroanatomical correlates of happiness, sadness, and disgust.
Image 11, 203 – 209. Am. J. Psychiatry 154 (7), 926 – 933.
Dolcos, F., Cabeza, R., 2002. Event-related potentials of emotional mem- Lane, R.D., Reiman, E.M., Bradley, M.M., Lang, P.J., Ahern, G.L., David-
ory: encoding pleasant, unpleasant, and neutral Pictures. Cogn. Affect son, R.J., Schwartz, G.E., 1997c. Neuroanatomical correlates of pleas-
Behav. Neurosci. 2 (3), 252 – 263. ant and unpleasant emotion. Neuropsychologia 35 (11), 1437 – 1444.
Dolcos, F., Graham, R., LaBar, K., Cabeza, R., 2003. Coactivation of the Lane, R.D., Chua, P.M., Dolan, R.J., 1999. Common effects of emotional
amygdala and hippocampus predicts better recall for emotional than for valence, arousal and attention on neural activation during visual pro-
neutral pictures. Brain and Cognition 51, 221 – 223. cessing of pictures. Neuropsychologia 37 (9), 989 – 997.
Dolcos, F., LaBar, K.S., Cabeza, R., 2004. Interaction between the amyg- Lang, P.J., Greenwald, M.K., Bradley, M.M., Hamm, A.O., 1993. Looking
dala and the medial temporal lobe memory system predicts better mem- at pictures: affective, facial, visceral, and behavioral reactions. Psycho-
ory for emotional events. Neuron 42, 855 – 863. physiology 30 (3), 261 – 273.
Epstein, C.M., Sekino, M., Yamaguchi, K., Kamiya, S., Ueno, S., 2002. Lang, P.J., Bradley, M.M., Cuthberg, B.N., 1997. International Affective
Asymmetries of prefrontal cortex in human episodic memory: effects of Picture System [Pictures]. NIMH Center for the Study of Emotion and
transcranial magnetic stimulation on learning abstract patterns. Neuro- Attention, Gainesville.
sci. Lett. 320 (1 – 2), 5 – 8. Liberzon, I., Taylor, S.F., Fig, L.M., Decker, L.R., Koeppe, R.A., Mino-
Fletcher, P.C., Shallice, T., Dolan, R.J., 1998. The functional roles of pre- shima, S., 2000. Limbic activation and psychophysiologic responses to
frontal cortex in episodic memory: I. Encoding. Brain 121, 1239 – 1248. aversive visual stimuli. Interaction with cognitive task. Neuropsycho-
Frijda, N.H., 1988. The laws of emotion. Am. Psychol. 43 (5), 349 – 358. pharmacology 23 (5), 508 – 516.
Frith, C.D., Frith, U., 1999. Interacting minds—A biological basis. Science McGaugh, J.L., McIntyre, C.K., Power, A.E., 2002. Amygdala modulation
286 (5445), 1692 – 1695. of memory consolidation: interaction with other brain systems. Neuro-
George, M.S., Ketter, T.A., Parekh, P.I., Horwitz, B., Herscovitch, P., Post, biol. Learn. Mem. 78 (3), 539 – 552.
R.M., 1995. Brain activity during transient sadness and happiness in Miller, E.K., Cohen, J.D., 2001. An integrative theory of prefrontal cortex
healthy women. Am. J. Psychiatry 152 (3), 341 – 351. function. Annu. Rev. Neurosci. 24, 67 – 202.
Goldman-Rakic, P.S., 1995. Architecture of the prefrontal cortex and the Morris, P.L., Robinson, R.G., Raphael, B., Hopwood, M.J., 1996. Lesion
central executive. Ann. N. Y. Acad. Sci. 769, 71 – 83. location and poststroke depression. J. Neuropsychiatry Clin. Neurosci.
Grafman, J., Wassermann, E., 1999. Transcranial magnetic stimulation can 8 (4), 399 – 403.
measure and modulate learning and memory. Neuropsychologia 37, Moscovitch, M., 1992. Memory and working-with-memory: a component
159 – 167. process model based on modules and central systems. J. Cogn. Neuro-
Gusnard, D.A., Akbudak, E., Shulman, G.L., Raichle, M.E., 2001. Medial sci. 4, 257 – 267.
prefrontal cortex and self-referential mental activity: relation to a de- Nakamura, K., Kawashima, R., Nagumo, S., Ito, K., Sugiura, M., Kato, T.,
fault mode of brain function. Proc. Natl. Acad. Sci. U. S. A. 98 (7), Nakamura, A., Hatano, K., Kubota, K., Fukuda, H., Kojima, S., 1998.
4259 – 4264. Neuroanatomical correlates of the assessment of facial attractiveness.
Hagemann, D., Naumann, E., Becker, G., Maier, S., Bartussek, D., 1998. NeuroReport 9 (4), 753 – 757.
Frontal brain asymmetry and affective style: a conceptual replication. O’Doherty, J., Kringelbach, M.L., Rolls, E.T., Hornak, J., Andrews, C.,
Psychophysiology 35 (4), 372 – 388. 2001. Abstract reward and punishment representations in the human
Hamann, S.B., Ely, T.D., Grafton, S.T., Kilts, C.D., 1999. Amygdala ac- orbitofrontal cortex. Nat. Neurosci. 4 (1), 95 – 102.
tivity related to enhanced memory for pleasant and aversive stimuli. Owen, A.M., Herrod, N.J., Menon, D.K., Clark, J.C., Downey, S.P., Car-
Nat. Neurosci. 2 (3), 289 – 293. penter, T.A., Minhas, P.S., Turkheimer, F.E., Williams, E.J., Robbins,
Hamann, S.B., Ely, T.D., Hoffman, J.M., Kilts, C.D., 2002. Ecstasy and T.W., Sahakian, B.J., Petrides, M., Pickard, J.D., 1999. Redefining the
agony: activation of the human amygdala in positive and negative functional organization of working memory processes within human
emotion. Psychol. Sci. 13 (2), 135 – 141. lateral prefrontal cortex. Eur. J. Neurosci. 11 (2), 567 – 574.
74 F. Dolcos et al. / NeuroImage 23 (2004) 64–74

Paller, K.A., Wagner, A.D., 2002. Observing the transformation of experi- Schneider, F., Gur, R.E., Mozley, L.H., Smith, R.J., Mozley, P.D., Censits,
ence into memory. Trends Cogn. Sci. 6 (2), 93 – 102. D.M., Alavi, A., Gur, R.C., 1995. Mood effects on limbic blood flow
Paradiso, S., Chemerinski, E., Yazici, K.M., Tartaro, A., Robinson, R.G., correlate with emotional self-rating: a PET study with oxygen-15 labeled
1999a. Frontal lobe syndrome reassessed: comparison of patients with water. Psychiatry Res. 61 (4), 265 – 283.
lateral or medial frontal brain damage. J. Neurol., Neurosurg. Psychiatry Shallice, T., Fletcher, P., Frith, C.D., Grasby, P., Frackowiak, R.S., Dolan,
67 (5), 664 – 667. R.J., 1994. Brain regions associated with acquisition and retrieval of
Paradiso, S., Johnson, D.L., Andreasen, N.C., O’Leary, D.S., Watkins, G.L., verbal episodic memory. Nature 368, 633 – 635.
Ponto, L.L., Hichwa, R.D., 1999b. Cerebral blood flow changes associ- Shields, S., 1991. Gender in the psychology of emotion: a selective re-
ated with attribution of emotional valence to pleasant, unpleasant, and search review. In: Strongman, K. (Ed.), International Review of Studies
neutral visual stimuli in a PET study of normal subjects. Am. J. Psychi- on Emotion. Wiley, New York, pp. 227 – 245.
atry 156 (10), 1618 – 1629. Simons, J.S., Spiers, H.J., 2003. Prefrontal and medial temporal lobe inter-
Pardo, J.V., Pardo, P.J., Raichle, M.E., 1993. Neural correlates of self- actions in long-term memory. Nat. Rev., Neurosci. 4 (8), 637 – 648.
induced dysphoria. Am. J. Psychiatry 150 (5), 713 – 719. Smith, E., Jonides, J., 1999. Storage and executive processes in the frontal
Petrides, M., 1995. Functional organization of the human frontal cortex for lobes. Science 283, 657 – 660.
mnemonic processing. Evidence from neuroimaging studies. Ann. N. Y. Starkstein, S.E., Robinson, R.G., Honig, M.A., Parikh, R.M., Joselyn, J.,
Acad. Sci. 769, 85 – 96. Price, T.R., 1989. Mood changes after right-hemisphere lesions. Br. J.
Petrovic, P., Kalso, E., Petersson, K.M., Ingvar, M., 2002. Placebo and Psychiatry 155, 79 – 85.
opioid analgesia—Imaging a shared neuronal network. Science 295, Talairach, J., Tournoux, P., 1988. A Co-Planar Stereotactic Atlas of the
1737 – 1740. Human Brain. Thieme, Stuttgart, Germany.
Phan, K.L., Wager, T., Taylor, S.F., Liberzon, I., 2002. Functional neuro- Talarico, J.T., LaBar, K.S., Rubin, D.C., 2004. Emotional intensity predicts
anatomy of emotion: a meta-analysis of emotion activation studies in autobiographical memory experience. Mem. Cogn. (in press).
PET and fMRI. NeuroImage 16 (2), 331 – 348. Taylor, S.F., Liberzon, I., Fig, L.M., Decker, L.R., Minoshima, S., Koeppe,
Phelps, E.A., Anderson, A.K., 1997. Emotional memory: what does the R.A., 1998. The effect of emotional content on visual recognition mem-
amygdala do? Curr. Biol. 7 (5), R311 – R314. ory: a PET activation study. NeuroImage 8 (2), 188 – 197.
Phelps, E.A., LaBar, K.S., Anderson, A.K., O’Connor, K.J., Fulbright, R.K., Taylor, S.F., Liberzon, I., Koeppe, R.A., 2000. The effect of graded aver-
Spencer, D.D., 1998. Specifying the contributions of the human amyg- sive stimuli on limbic and visual activation. Neuropsychologia 38 (10),
dala to emotional memory: a case study. Neurocase 4, 527 – 540. 1415 – 1425.
Poldrack, R.A., Wagner, A.D., Prull, M.W., Desmond, J.E., Glover, G.H., Teasdale, J.D., Howard, R.J., Cox, S.G., Ha, Y., Brammer, M.J., Williams,
Gabrieli, J.D., 1999. Functional specialization for semantic and phono- S.C., Checkley, S.A., 1999. Functional MRI study of the cognitive
logical processing in the left inferior prefrontal cortex. NeuroImage 10, generation of affect. Am. J. Psychiatry 156 (2), 209 – 215.
15 – 35. Tomarken, A.J., Davidson, R.J., Wheeler, R.E., Doss, R.C., 1992. Individ-
Reiman, E.M., 1997. The application of positron emission tomography to ual differences in anterior brain asymmetry and fundamental dimen-
the study of normal and pathologic emotions. J. Clin. Psychiatry 58 sions of emotion. J. Pers. Soc. Psychol. 62 (4), 676 – 687.
(16), 4 – 12. Wager, T.D., Phan, K.L., Liberzon, I., Taylor, S.F., 2003. Valence, gender,
Reiman, E.M., Lane, R.D., Ahern, G.L., Schwartz, G.E., Davidson, R.J., and lateralization of functional brain anatomy in emotion: a meta-anal-
Friston, K.J., Yun, L.S., Chen, K., 1997. Neuroanatomical correlates of ysis of findings from neuroimaging. NeuroImage 19 (3), 513 – 531.
externally and internally generated human emotion. Am. J. Psychiatry Wagner, A.D., Schacter, D.L., Rotte, M., Koutstaal, W., Maril, A., Dale,
154 (7), 918 – 925. A.M., Rosen, B.R., Buckner, R.L., 1998. Building memories: remem-
Rolls, E.T., 2000. The orbitofrontal cortex and reward. Cereb. Cortex 10 bering and forgetting of verbal experiences as predicted by brain activ-
(3), 284 – 294. ity. Science 281, 1188 – 1191.
Rossi, S., Cappa, S.F., Babiloni, C., Pasqualetti, P., Miniussi, C., Carducci, Wheeler, R.E., Davidson, R.J., Tomarken, A.J., 1993. Frontal brain asym-
F., Babiloni, F., Rossini, P.M., 2001. Prefrontal [correction of Prefontal] metry and emotional reactivity: a biological substrate of affective style.
cortex in long-term memory: an ‘‘interference’’ approach using mag- Psychophysiology 30, 82 – 89.
netic stimulation. Nat. Neurosci. 4 (9), 948 – 952. Yamasaki, H., LaBar, K.S., McCarthy, G., 2002. Dissociable prefrontal brain
Russell, J., 1980. A circumplex model of affect. J. Pers. Soc. Psychol. 39, systems for attention and emotion. Proc. Natl. Acad. Sci. U. S. A. 99 (17),
1161 – 1178. 11447 – 11451.
!
!
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