ArticlePDF Available

Accurate statistical tests for smooth classification images

Authors:

Abstract and Figures

Despite an obvious demand for a variety of statistical tests adapted to classification images, few have been proposed. We argue that two statistical tests based on random field theory (RFT) satisfy this need for smooth classification images. We illustrate these tests on classification images representative of the literature from F. Gosselin and P. G. Schyns (2001) and from A. B. Sekuler, C. M. Gaspar, J. M. Gold, and P. J. Bennett (2004). The necessary computations are performed using the Stat4Ci Matlab toolbox.
Content may be subject to copyright.
Accurate statistical tests for smooth
classification images
D
epartement de Psychologie, Universit
e de Montr
eal,
Montr
eal, QC, Canada
Alan Chauvin
Department of Mathematics and Statistics, McGill
University, Montr
eal, QC, Canada
Keith J. Worsley
Department of Psychology, University of Glasgow,
Glasgow, United Kingdom
Philippe G. Schyns
D
epartement de Psychologie, Universit
e de Montr
eal,
Montr
eal, QC, Canada
Martin Arguin
D
epartement de Psychologie, Universit
e de Montr
eal,
Montr
eal, QC, Canada
Fr
ed
eric Gosselin
Despite an obvious demand for a variety of statistical tests adapted to classification images, few have been proposed. We
argue that two statistical tests based on random field theory (RFT) satisfy this need for smooth classification images. We
illustrate these tests on classification images representative of the literature from F. Gosselin and P. G. Schyns (2001) and
from A. B. Sekule r, C. M. Gaspar, J. M. Gold, and P. J. Bennett (2004). The necessary computations are performed using
the Stat4Ci Matlab toolbox.
Keywords: classifi cation images, reverse correlation, Bubbles, random field theory
Introduction
In recent years, vision research has witnessed a tre-
mendous growth of interest for regression techniques ca-
pable of revealing the use of information (e.g., Ahumada,
1996; Eckstein & Ahumada, 2002; Gosselin & Schyns, 2004b).
Reverse correlation, one such technique, has been em-
ployed in a number of domains ranging from electroretino-
grams (Sutter & Tran, 1992), visual simple response time
(Simpson, Braun, Bargen, & Newman, 2000), single pulse
detection (Thomas & Knoblauch, 1998), vernier acuity
(Barth, Beard, & Ahumada, 1999; Beard & Ahumada, 1998),
objects discrimination (Olman & Kersten, 2004), stereop-
sis (Gosselin, Bacon, & Mamassian, 2004; Neri, Parker,
& Blakemore, 1999), letter discrimination (Gosselin &
Schyns, 2003;Watson,1998; Watson & Rosenholtz, 1997),
single neuron’s receptive field (e.g., Marmarelis & Naka,
1972; Ohzawa, DeAngelis, & Freeman, 1990; Ringach &
Shapley, 2004), modal and amodal completion (Gold,
Murray, Bennett, & Sekuler, 2000), face representa-
tions (Gold, Sekuler, & Bennett, 2004; Kontsevich &
Tyler, 2004; Mangini & Biederman, 2004; Sekuler et al., 2004)
to temporal processing (Neri & Heeger, 2002). Bubbles, a
related technique (Gosselin & Schyns, 2001, 2002, 2004b;
Murray & Gold, 2004), has revealed the use of information for
the categorization of face identity, expression, and gender
(Adolphs et al., 2005; Gosselin & Schyns, 2001; Schyns,
Bonnar, & Gosselin, 2002; Smith, Cottrell, Gosselin, &
Schyns, 2005;Vinette,Gosselin,&Schyns,2004), for the
categorization of natural scenes (McCotter, Sowden, Gosse-
lin, & Schyns, in press), for the perception of an
ambiguous figure (Bonnar, Gosselin, & Schyns, 2002),
and for the interpretation of EEG signals (Schyns,
Jentzsch, Johnson, Schweinberger, & Gosselin, 2003;
Smith, Gosselin, & Schyns, 2004).
Both the Bubbles and the reverse correlation
techniques produce large volumes of regression
coefficients that have to be tested individually. As
we will shortly discuss, this raises the issue of false
positives: the risk of accepting an event that occurred
by chance. Surprisingly, few classification image re-
searchers have taken this into account (for exceptions,
see Abbey & Eckstein, 2002; Kontsevich & Tyler, 2004;
Mangini & Biederman, 2004). Here, we argue that two
statistical tests based on random field theory (RFT) satisfy
this need for smooth classification images. The core
ideas of RFT are presented. In particular, the main
equations for the tests are given. Finally, the usage of
a Matlab toolbox implementing the tests is illustrated
on two representative sets of classification images from
Gosselin and Schyns (2001) and Sekuler et al. (2004). But
first, in order to identify the critical properties of the
proposed statistical tests, we shall discuss some limitations
Journal of Vision (2005) 5, 659–667 http://journalofvision.org/5/9/1/ 659
doi: 10.1167/5.9.1 Received October 25, 2004; published October 5, 2005 ISSN 1534-7362 * ARVO
of the two statistical tests that have already been applied
to classification images.
Multiple comparisons
In a typical classification image experiment, an ob-
server has to classify objects partially revealed by additive
(reverse correlation) or multiplicative (Bubbles) noise fields.
The calculation of the classification image amounts quite
simply to summing all the noise fields weighted by the
observer’s responses (Ahumada, 2002; Murray, Bennett,
& Sekuler, 2002). By doing this, the researcher is actu-
ally performing a multiple regression on the observer’s
responses and the noise fields (see Appendix). A sta-
tistical test compares these data against the distribution
of a random process with similar characteristics. Classi-
fication images can thus be viewed, under the null hypoth-
esis, as expressions of a random N-dimensional process
(i.e., a random field). The alternate hypothesis is that a
signalVknown or unknownVis hidden in the regression
coefficients.
So far, researchers have used two statistical tests to
achieve this end: the Bonferroni correction and Abbey
and Eckstein’s (2002) Hotelling test. We will argue that
these tests are not adapted to some classification images.
The former is too conservative when the elements of the
classification images are locally correlated, and the latter
is not suitable in the absence of a priori expectations
about the shape of the signal hidden in the classification
images.
Bonferroni correction
Consider a one-regression coefficient Z-scored classi-
fication image (see Appendix). If this Z score ex-
ceeds a threshold determined by a specified p value, this
regression coefficient differs significantly from the null
hypothesis. For example, a p value of .05 means that if
we reject the null hypothesis, that is, if the Z score ex-
ceeds a threshold t
Z
= 1.64, the probability of a false
alarm (or Type I error) is .05. Now consider a classi-
fication image comprising 100 regression coefficients:
the expected number of false alarms is 100 0.05 = 5.
With multiple Z tests, such as in the previous example,
the overall p value can be set conservatively using the
Bonferroni correction: p
BON
= p(Z 9 t
BON
)N,withN the
number of points in the classification image. Again, con-
sider our hypothetical 100-point classification image. The
corrected threshold, t
BON
, associated with p
BON
=.05,is
3.29. Such high Z scores are seldom observed in classi-
fication images derived from empirical data. In a clas-
sification image of 65,536 data points (typical of those
found in the literature, like the 256 256 classification
images from Gosselin & Schyns, 2001, reanalyzed in the
last section of this article), t
BON
becomes a formidable
4.81! For classification images of low (or reduced) di-
mensionality such as those of Mangini and Biederman
(2004) or Kontsevich and Tyler (2004), the Bonferroni
correction prescribes thresholds that can be (and have
been) attained.
A priori expectations
Two possibilities should be considered: either these
classification images really do not contain anything statis-
tically significant (which seems unlikely given the robust-
ness of the results obtained with no Bonferroni correction;
e.g., Gosselin & Schyns, 2001; Schyns et al., 2002, 2003),
or the Bonferroni correction is too conservative. Do we
have a priori expectations that support the latter and can
we use these expectations to our advantage? Abbey and
Eckstein (2002), for example, have derived a statistical
test far more sensitive than the Bonferroni correction for
classification images derived from a two-alternative forced-
choice paradigm when the signal is perfectly known. Al-
though we often do not have such perfect a priori knowledge
about the content of classification images, we do expect
them to be relatively smooth.
The goal of Bubbles and reverse correlation is to reveal
clusters of points that are associated with the measured
response; for example, the mouth or the eyes of a face
(Gold et al., 2004; Gosselin & Schyns, 2001; Mangini &
Biederman, 2004; Schyns et al., 2002; Sekuler et al., 2004),
illusory contours (Gold et al., 2000), and so on. In other
words, it is expected that the data points of classification
images are correlated, introducing Bsmoothness[ in the
solutions. The Bonferroni correction, adequate when data
points are independent, becomes far too conservative (not
sensitive enough) for classification images with a corre-
lated structure.
In the next section, we present two statistical tests based
on RFT that provide accurate thresholds for smooth, high-
dimensional classification images.
Random field theory
Adler (1981) and Worsley (1994, 1995a, 1995b, 1996)
have shown that the probability of observing a cluster
of pixels exceeding a threshold in a smooth Gaussian
random field is well approximated by the expected Euler
characteristic (EC). The EC basically counts the number
of clusters above a sufficiently high threshold in a
smooth Gaussian random fields. Raising the threshold
until only one cluster remains brings the EC value to 1;
raising it still further until no cluster exceeds the thresh-
old brings it to 0. Between these two thresholds, the
expected EC approximates the probability of observing one
cluster. The formal proof of this assertion is the centerpiece
of RFT.
Journal of Vision (2005) 5, 659–667 Chauvin et al. 660
Next we present the main equations of two statistical
tests derived from RFT: the so-called pixel and cluster
tests, which have already been successfully applied for
more than 15 years to brain imaging data. Crucially, these
tests take into account the spatial correlation inherent to
the data set, making them well suited for classification
images.
Pixel test
Suppose that Z is a Z<scored classification image (see
Appendix). In RFT, the subset of Z searched for un-
likely clusters of regression coefficientsVe.g., the face
areaVis called the search space (S). The probability of
observing at least one regression coefficient exceeding t is
well approximated by
Pðmax Z 9 tÞ
X
D
d¼0
Resels
d
ðSÞ I EC
d
ðtÞð1Þ
where D is the dimensionality of S;EC
d
(t )isthe
d-dimensional EC density that depends partly on the type
of statistic (for EC densities of other random fields, see
Cao & Worsley, 2001; Worsley, Marrett, Neelin, Vandal,
Friston, & Evans, 1996); Resels
d
(S) is the d-dimensional
Resels (resolution elements), which varies with the size
and the shape of S. The EC densities of a D = two-
dimensional Gaussian random field Z are
EC
0
ðtÞ¼
Z
1
t
ð2Þ
1
2
e
u
2
2
du ¼ pðZ 9 tÞ; ð2Þ
EC
1
ðtÞ¼
ffiffiffiffiffiffiffiffiffiffiffiffiffiffi
4lnð2Þ
p
2
I e
t
2
2
; ð3Þ
and
EC
2
ðtÞ¼
4lnð2Þ
ð2Þ
3
2
I t I e
t
2
2
: ð4Þ
The Resels is given by
Resels
d
ðSÞ¼
V
d
ðSÞ
FWHM
d
; ð5Þ
where V
0
(S) = 1 for a connected search region, V
1
(S)=
half perimeter length of S, V
2
(S) = caliper area of S (a
disk of the same area as S gives a good approximation and
allows to derive the volumes of the lower dimension; see
Cao & Worlsey, 2001). The FWHM is the full width at
half maximum of the filter f used to smooth the inde-
pendent error noise in the image. If the filter is Gaussian
with standard deviation
b
then
FWHM ¼
b
ffiffiffiffiffiffiffiffiffiffi
8ln2
p
: ð6Þ
The filter f should be chosen to give the best discrimi-
nation, or in other words to maximize the detection of sig-
nal in Z. There is a classic theorem in signal processing,
the matched filter theorem, which states that to detect sig-
nal added to white noise, the optimum filter should match
the shape of the signal. This implies that to optimally
detect, say 10 pixel features, we should smooth the data
with 10 pixels FWHM filter. But if for instance it was felt
that larger contiguous areas of the image were involved in
discrimination, then this might be better detected by using
a broader filter at the statistical analysis stage (see Worsley
et al., 1996).
This dependency of the pixel test on the choice of an
adequate filter has led to a generalization of the test in
which an extra dimension, the scale of the filter, is added
to the image to create a scale space image (Poline &
Mazoyer, 1994; Siegmund & Worsley, 1995). The scale
space search reduces the uncertainty of choosing a filter
FWHM but at the cost of higher thresholds.
Cluster test
The pixel test computes a statistical threshold based
on the probability of observing a single pixel above the
threshold. This test has been shown to be best suited for
detecting focal signals with high Z scores (Poline, Worsley,
Evans & Friston, 1997). But if the region of interest in the
search space (the mouth in a face for example) is wide, it
has usually a lower Z score and cannot be detected. We
could improve detection by applying more smoothing to
the image. The amount of smoothing will depend on the
extent of the features we wish to detect (by the matched
filter theorem), but we do not know this in advance.
Friston, Worsley, Frackowiak, Mazziotta, and Evans
(1994) proposed an alternative to the pixel test to improve
the detection of wide signals with low Z scores (for a
review, see Poline et al., 1997). The idea is to set a low
threshold (t Q 2.3Vin the next section, we used t = 2.7)
and base the test on the size of clusters of connected
pixels above the threshold. The cluster test is based on the
probability that, above a threshold t, a cluster of size K (or
more) pixels has occurred by chance that is calculated in
the D = 2 case as follows (Cao & Worsley, 2001; Friston
et al., 1994):
PðK 9 kÞ1 e
ðResels
2
ðS ÞEC
2
ðtÞpÞ
; ð7Þ
where
p ¼ e
ðð
ffiffiffi
2
p
EC
2
ðtÞkÞ=ðFWHM
2
pðZ 9 tÞÞÞ
ð8Þ
Cluster versus pixel test
The cluster and the pixel test presented above pro-
vide accurate thresholds but for different types of signal.
The pixel test is based on the maximum of a random field
Journal of Vision (2005) 5, 659–667 Chauvin et al. 661
and therefore is best adapted for focal signal (optimally
the size of the FWHM) with high Z scores (Poline et al.,
1997; Siegmund & Worsley, 1995). The cluster test is based
on the size of a cluster above a relatively low threshold
and therefore is more sensitive for detecting wide regions
of contiguous pixels with relatively low Z scores. The
two tests potentially identify different statistically signif-
icant regions in smooth classification images. Figure 1
illustrates this point with a one-dimensional classification
image comprising 257 pixels convolved with a Gaussian
kernel with an FWHM of 11.8 pixels. For a p value of .05,
the pixel test gives a threshold of 3.1 (green line) whereas
the cluster test gives a minimum cluster size of 6.9 above
a threshold of 2.7 (red line).
Furthermore, the interpretation of the results following
the application of the pixel and the cluster test differs
drastically. On the one hand, the cluster test allows the
inference that the clusters of Z scores larger than the
minimum size are significant, not that the individual Z
scores inside these clusters are significant. On the other
hand, the pixel test allows the conclusion that each
individual Z score above threshold is significant (Friston,
Holmes, Poline, Price, Frith, 1996;Fristonetal.,1994;
Poline et al., 1997).
Accuracy
Since the late 1980s, RFT has been used to analyze pos-
itron emission tomography (PET) images, galaxy density
maps, cosmic microwave background data, and functional
magnetic resonance imaging (fMRI) data. In fact, the RFT
is at the heart of two popular fMRI data analysis pack-
ages: SPM2 (Frackowiak et al., 2003) and FMRISTAT
(Worsley, 2003).
Not surprisingly, the accuracy of RFT has been ex-
amined extensively. An accurate statistical test must be
both sensitive (i.e., high hit rate) and specific (i.e., high
correct rejection rate). In particular, RFT has been eval-
uated in the context of so-called Bphantom[ simulations
(Hayasaka, Luan Phan, Liberzon, Worsley, & Nichols,
2004; Hayasaka & Nichols, 2003; Poline et al., 1997;
Worsley, 2005). A Bphantom[ simulation basically consists
of generating a lot of smooth random regression
coefficients, hiding a Bphantom[ in them (i.e., a known
signalVusually a disc or a Gaussian), attempting to detect
the Bphantom[ with various statistical tests, and deriving,
per statistical test, a measure of accuracy such as a
d-prime or an ROC area. We singled-out Bphantom[
simulations for a reason: If we were to compare the
accuracy of various statistical tests for the detection of a
Bphantom[ template (e.g., used by a linear amplifier model)
in a smooth classification images, this is exactly what we
would have to do. In other words, these Bphantom[
simulations inform us just as much about the accuracy of
RFT for fMRI data than about its accuracy for classification
images.
To summarize these assessments, the p values given by
RFT appears to be more accurate than those given by the
Bonferroni, the Hochberg, the Holm, the Sid"k, and the
false discovery rate, provided that the size of the search
space is greater than about three times that of the FWHM
(Hayasaka & Nichols, 2003), that the FWHM is greater
than about five pixels (Taylor, Worsley, & Gosselin,
2005), and that the degree of freedom is greater than about
200. Also, the cluster test is more sensitive and less
specific than the pixel test.
Reanalyzing representative
classification images
In the final section of this article, we apply the pixel
and cluster tests to classification images representative of
the literature from Gosselin and Schyns (2001) and Sekuler
et al. (2004). We give sample commands for the Stat4Ci
Matlab toolbox throughout.
Matlab implementation
A mere four pieces of information are required for the
computation of the significant regions using the pixel and
the cluster tests: a desired p value, a threshold t (only used
for the cluster test), a search space, and the FWHMVor,
equivalently, the sigmaVof the Gaussian kernel used to
smooth the classification image. The main function from
the Stat4Ci Matlab toolboxVStatThresh.mVinputs this
information together with a suitably prepared classifica-
tion image (i.e., smoothed and Z-scored), performs all the
computations described above, and outputs a threshold for
the pixel test as well as the minimum size of a significant
cluster for the cluster test. The StatThresh.m function makes
extensive use of the stat
_
threshold.m function, which was
Figure 1. Regions revealed by the cluster (red) versus the pixel
(green) test. See text for details.
Journal of Vision (2005) 5, 659–667 Chauvin et al. 662
originally written by Keith Worsley for the FMRISTAT
toolbox.
Other functions included in the Stat4Ci toolbox perform
a variety of related computations; for example, ReadCid.m
reads a classification image data (CID) file; BuildCi.m con-
structs classification images from a CID file; SmoothCi.m
convolves a raw classification image with a Gaussian
filter; ExpectedSCi.m computes the expected mean and
standard deviation of a smooth classification image
(see Appendix); ZscoreSCi.m Z<scores a smoothed
classification image (see Equation 9 and Appendix); and
DisplayRes.m displays the thresholded Z-transformed
smooth classification image and outputs a summary table
(see Figure 2). All of these functions include thorough help
sections.
Sekuler et al. (2004)
Sekuler et al. examined the effect of face inversion
on the information used by human observers to resolve
an identification task. Four classification images extracted
using reverse correlation are reanalyzed: one for each com-
bination of two subjects (MAT and CMG) and two con-
ditions (UPRIGHT and INVERTED). Each classification
image cumulates the data from 10,000 trials. We will not
further describe this experiment. Rather we will limit the
presentation to what is required for to application of the
pixel and cluster tests.
First, the raw classification images must be convolved
with a Gaussian filter (i.e., smoothed). The choice of
the appropriate Gaussian filter depends essentially on the
size of the search space (for a discussion, see Worsley,
2005). We chose a Gaussian filter with a standard de-
viation of
b
= 4 pixels; its effect are similar to those of
the filter used by Sekuler et al. (2004). Second, the
smooth classification images must be Z-scored. This can
sometimes be achieved analytically (see Appendix). How-
ever, if the number of trials is greater than 200Vas is
usually the case with classification imagesVthe transforma-
tion can be approximated as follows:
ZSCi ¼
SCi SCi
SCi
; ð9Þ
where the mean and standard deviation are estimated di-
rectly from the data, preferably from signal-less regions
of the classification images (e.g., regions corresponding
to a homogeneous background). In the Stat4Ci toolbox,
classification image preparation can be done as illus-
trated in Figure 3.
Once the classification image has been smoothed and
Z-scored, it must be inputted into the StatThr esh.m func-
tion together with the four additional required pieces
Figure 2. Sample summary table produced by DisplayRes.m
(from the reanalysis of classification images from Gosselin &
Schyns, 2001; see next section). The numbers between brackets
were set by the user. C = cluster test; P = pixel test; t = threshold;
size = size of the cluster; Z
max
, x, and y = maximum Z score and
its coordinates.
Figure 3. Sample commands for the Stat4Ci Matlab toolbox (from the
reanalysis of classification images from Gosselin & Schyns, 2001).
Figure 4. Sekuler et al.’s (2004) classification images reanalyzed
using the Stat4Ci Matlab toolbox.
Journal of Vision (2005) 5, 659–667 Chauvin et al. 663
of information: a p value ( p e .05), the sigma of the
Gaussian filter used during the smoothing phase (
b
=4
pixels for this reanalysis), a threshold for the cluster test
(equal to 2.7 for this reanalysis), and a search space (i.e., the
face region).
The statistical threshold obtained using the pixel test is
very low compared with that obtained using the Bonferroni
correction (i.e., 3.67 rather than 4.5228; see Bonferroni
correction). In fact, the stat
_
threshold.m function outputs
the minimum between the Bonferroni and the pixel test
thresholds. Figure 4 displays the thresholded classifica-
tion images for both the pixel and cluster tests. For
the cluster test, only the clusters larger than the
minimum size (i.e., 66.6 pixels) are shown. Red pixels
indicate the regions that attained significance in the
UPRIGHT condition; green pixels, in the INVERTED
condition; and yellow pixels, in both. A face background
was overlaid to facilitate interpretation.
Gosselin and Schyns (2001)
Gosselin and Schyns (Experiment 1) examined the in-
formation used by human observers to resolve a GENDER
and an expressive versus not expressive (EXNEX) face
discrimination task. They employed the Bubbles techni-
que to extract two classification images per observer, one
per task. Each of the two classification images reana-
lyzed in this section combines the data from 500 trials
executed by subject FG. The classification images can be
built either from the opaque masks punctured by Gaussian
holes and applied multiplicatively to a face on each trial,
or from the center of these Gaussian holes. The former
option naturally results in smooth classification images;
and the latter option calls for smoothing with a filter,
just like the classification images of Sekuler et al. (2004).
For this reanalysis, we used a Gaussian filter identical
to the one used to sample information during the actual
experiment (
b
= 20 pixels). In this case, both options
are strictly equivalent. These smooth classification im-
ages (see SCi in Figure 3)wereZ-scored using Equation 9
with estimations of the expected means and standard de-
viations based on the signal-less pixels outside the search
region.
Next, the Z-transformed smooth classification image
(see ZSCi in Figure 3) is inputted into the StatThresh.m
function with the four additional required pieces of infor-
mation: a p value ( p e .05), the sigma of the Gaussian fil-
ter used during the smoothing phase (
b
= 20 pixels for
this reanalysis), a threshold for the cluster test (equal to
2.7 for this reanalysis; see tC in Figure 3), and a search
space (see S
_
r in Figure 3). See Figure 3 for all the rel-
evant Stat4Ci toolbox commands.
Again, the statistical threshold obtained using the pixel
test is extremely low compared with that obtained using
the Bonferroni correction: 3.30 rather than 4.808 (see
Bonferroni correction). Figure 5 displays the thresholded
classification images for both the pixel and cluster tests.
For the cluster test, only the clusters larger than the
minimum size (i.e., 861.7 pixels) are shown. Red pixels
indicate the regions that attained statistical significance. A
face background (see background in Figure 3)was
overlaid to facilitate interpretation.
Take-home message
We have presented two statistical tests suitable for
smooth, high-dimensional classification images in the ab-
sence of a priori expectations about the shape of the sig-
nal. The pixel and the cluster tests, based on RFT, are
accurate within known boundaries discussed in the article.
These tests require only four pieces of information and
their computation can be performed easily using the Stat4Ci
Matlab toolbox. We expect these tests to be most useful
for researchers applying Bubbles or reverse correlation
to complex stimuli.
Appendix: The construction of
a classification image
In a reverse correlation or Bubbles experiment, an ob-
server is presented with a noise field on trial i (i =1,I, n)
and produces the response Y
i
.
Figure 5. Two of Gosselin and Schyns’ (2001) classification images
reanalyzed using the Stat4Ci Matlab toolbox.
Journal of Vision (2005) 5, 659–667 Chauvin et al. 664
At a particular pixel v, we suppose that some feature of
the noise field, X
i
(v), is correlated with the response. In a
reverse correlation experiment, X
i
(v) might be the added
noise at pixel v; in a Bubbles experiment, X
i
(v) might be
the actual bubble mask at pixel v. We aim to detect those
pixels where the response is highly correlated with the
image feature of interest. The sample correlation at pixel v
is
CðvÞ¼
X
i
ðX
i
ðvÞX ðvÞÞðY
i
Y Þ
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
P
i
ðX
i
ðvÞX ðvÞÞ
2
r ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
P
i
ðY
i
Y Þ
2
r
; ðA1Þ
where bar indicates averaging over all n trials. It is straight-
forward to show that if there is no correlation between
image features and response, then
ZðvÞ¼
ffiffiffiffiffiffiffiffiffiffi
n 2
p
CðvÞ
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 CðvÞ
2
q
:
ffiffi
n
p
CðvÞðA2Þ
has a Student-t distribution with n Y 2 degrees of freedom
provided X
i
(v) is Gaussian, but in any case n is usually
very large so the standard normal distribution will be a
very good approximation, by the Central Limit Theorem.
Provided that
P
X
i
= 0 and
P
Y
i
= 0, the ZTransSCi.m
function from the Stat4Ci toolbox implements Equation A1.
In this case, the numerator is simply the sum of all the
noise fields weighted by the observer’s responses. The
remaining term in C(v) can be approximated if X
i
(v)is
white noise W
i
(v) (i.e. independent and identically dis-
tributed noise at each pixel) convolved with a filter f (v),
that is if
X
i
ðvÞ¼
X
u
f ðv uÞW
i
ðuÞ: ðA3Þ
In the case of reverse correlation, W
i
(v) is usually a
Gaussian random variable and often there is no filtering,
so f (v) is zero except for f (0) = 1. In the case of Bubbles,
W
i
(v) is a binary random variable taking the value 1 if
there is a bubble centered at v, and 0 otherwise. The
X
i
ðX
i
ðvÞX ðvÞÞ
2
: n
2
X
u
f ðvÞ
2
; ðA4Þ
where
2
is the variance of the white noise. For reverse
correlation,
2
is the variance of the Gaussian white noise.
For Bubbles, it is the Binomial variance
2
¼
N
b
N
1
N
b
N

:
N
b
N
; ðA5Þ
where N
b
is the number of bubbles and N is the number of
pixels.
The central limit theorem ensures that, at the limit, Z(v)
is a Gaussian random field with an effective FWHM equals
to the FWHM of the filter f (v). The rate of convergence
toward Gaussianity depends partly on the predictive
variable and partly on the total number of bubbles per
Resels. Worsley (2005) has examined the exactness of the
p values given by the Gaussian procedures presented in
this article in function of these two factors: At 10,000
bubbles per Resels, the p values given by the Gaussian
procedures depart from the true p values by less than
T0.04 logarithmic unit; at 500 bubbles per Resels,
a figure more often encountered in practice (e.g.,
Gosselin & Schyns, 2001), the discrepancy can be as
much as T0.3 logarithmic unit. If the predictive variable
has a positively skewed distribution, the Gaussian pro-
cedure is liberal; and if it has a negatively skewed distri-
bution, as is usually the case in practice (e.g., Gosselin &
Schyns, 2001), the Gaussian procedure is conservative.
Acknowledgments
This research was supported by an NSERC (249974) and
a NATEQ (03NC082052) grant awarded to Fr2d2ric Gosselin;
by a NATEQ (84180) grant awarded to Martin Arguin and
Fr2d2ric Gosselin; and by an ESRC grant R000237901 to
Philippe G. Schyns. We thank Allison Sekuler, Carl Gaspar,
Jason Gold, and Patrick Bennett for having kindly given
us access to their classification images.
Commercial relationships: none.
Corresponding author: Fr2d2ric Gosselin.
Email: frederic.gosselin@umontreal.ca.
Address: D2partement de psychologie, Universit2 de
Montr2al, C.P. 6128, Succursale Centre-ville, Montr2al,
Qu2bec, Canada H3C 3J7.
References
Adler, R. J. (1981). The geometry of random fields. New
York: Wiley.
Adolphs, R., Gosselin, F., Buchanan, T. W., Tranel, D.,
Schyns, P. G., & Damasio, A. R. (2005). A mech-
anism for impaired fear recognition after amygdala
damage. Nature, 433, 68Y72. [PubMed]
Ahumada, A. J., Jr. (1996). Perceptual classification im-
ages from vernier acuity masked by noise [Abstract].
Perception, 26, 18.
Ahumada, A. J. (2002). Classification image weights and
internal noise level estimation. Journal of Vision,
2(1), 121Y131, http://journalofvision.org/2/1/8/,
doi:10.1167/2.1.8. [PubMed][Article]
Abbey, C. K., & Eckstein, M. P. (2002). Classification
image analysis: Estimation and statistical inference
for two-alternative forced-choice experiments. Journal
Journal of Vision (2005) 5, 659–667 Chauvin et al. 665
of Vision, 2(1), 66Y78, http://journalofvision.org/2/1/5/,
doi:10.1167/2.1.5. [PubMed][Article]
Barth, E., Beard, B. L., & Ahumada, A. J. (1999). Non-
linear features in vernier acuity. In B. E. Rogowitz &
T. N. Pappas (Eds.), Human vision and electronic
imaging IV, SPIE Proceedings, 3644, paper 8.
Beard, B. L., & Ahumada, A. J. (1998). A technique to
extract the relevant features for visual tasks. In B. E.
Rogowitz&T.N.Pappas(Eds.),Human vision and
electronic imaging III, SPIE Proceedings, 3299,
79Y 85.
Bonnar, L., Gosselin, F., & Schyns, P. G. (2002).
Understanding Dali’s slave market with the disap-
pearing bust of voltaire: A case study in the scale
information driving perception. Perception, 31,
683Y 691. [PubMed]
Cao, J., & Worsley, K. J. (2001). Applications of random
fields in human brain mapping. In M. Moore (Ed.),
Spatial statistics: Methodological aspects and appli-
cati ons, Springer lecture notes in statistics, 159,
169Y 182.
Eckstein, M. P., & Ahumada, A. J. (Ed.). (2002). Clas-
sification images: A tool to analyze visual strategies
[Special issue]. Journal of Vision, 2(1), iYi, http://
journalofvision.org/2/1/i/, doi:10.1167/2.1.i.
[PubMed][Article]
Frackowiak R., Friston, K. J., Frith, C., Dolan, R., Price,
C., Ashburner, J., et al. (2003). Human brain function
(2nd ed.), Academic Press.
Friston, K. J., Worsley, K. J., Frackowiak, R. S. J.,
MazziottaJ.C.,&EvansA.C.(1994).
Assessing the significance of focal activations
using their spatial extent. Human Brain Mapping,
1, 214Y 220.
Friston, K. J., Holmes, A, Poline, J. B., Price, C. J., Frith,
C. D. (1996). Detecting activations in PET and fMRI:
Levels of inference and power. Neuroimage, 4(3),
223Y 35. [PubMed]
Gold, J. M., Murray, R. F., Bennett, P. J., & Sekuler, A.
B. (2000). Deriving behavioural receptive fields for
visually completed contours. Current Biology, 10,
663Y 666. [PubMed]
Gold, J. M., Sekuler, A. B., & Bennett, P. J. (2004).
Characterizing perceptual learning with external
noise. Cognitive Science, 28, 167Y207. [Abstract]
Gosselin, F., Bacon, B. A., & Mamassian, P. (2004). In-
ternal surface representations approximated by re-
verse correlation. Vision Research, 44, 2515Y2520.
[Abstract][PubMed]
Gosselin, F., & Schyns, P. G. (2001). Bubbles: A technique
to reveal the use of information in recognition. Vision
Research, 41, 2261Y 2271. [PubMed]
Gosselin, F., & Schyns, P. G. (2002). RAP: A new
framework for visual categorization. Trends in
Cognitive Science, 6, 70Y 77. [Abstract][PubMed]
Gosselin, F., & Schyns, P. G. (2003). Superstitious
perceptions reveal properties of memory representa-
tions. Psychological Science, 14, 505Y 509. [PubMed]
Gosselin, F., & Schyns, P. G. (2004a). No troubles with
bubbles: A reply to murray and gold. Vision Research,
44, 471Y 477. [PubMed]
Gosselin, F., & Schyns, P. G. (Ed.). (2004b). A pic-
ture is worth thousands of trials: Rendering the use
of visual information from spiking neurons to rec-
ognition [Special issue]. Cognitive Science, 28,
141Y 146.
Hayasaka, S., & Nichols, T. (2003). Validating cluster size
inference: Random field and permutation methods.
NeuroImage, 20, 2343Y 2356. [PubMed]
Hayasaka, S., Luan Phan, K., Liberzon, I., Worsley, K. J.,
& Nichols, T. (2004). Nonstationary cluster-size in-
ference with random field and permutation methods.
NeuroImage, 22, 676Y 687. [PubMed]
Kontsevich, L. L., & Tyler, C. W. (2004). What makes
Mona Lisa smile? Vision Research, 44, 1493Y 1498.
[PubMed]
Mangini, M. C., & Biederman, I. (2004). Making the in-
effable explicit: Estimating the information employed
for face classifications. Cognitive Science, 28, 209Y 226.
[Abstract]
Marmarelis, P. Z., & Naka, K. I. (1972). White-noise
analysis of a neuron chain: An application of the
wiener theory. Science, 175, 1276Y 1278. [PubMed]
McCotter M., Gosselin, F., Sowden, P., & Schyns, P. G.
(in press). The use of visual information in natural
scenes. Visual Cognition.
Murray, R. F., & Gold, J. M. (2004). Troubles with bub-
bles. Vision Research, 44, 461Y 470. [PubMed]
Murray, R. F., Bennett, P. J., & Sekuler, A. B. (2002).
Optimal methods for calculating classification images:
Weighted sums. Journal of Vision, 2(1), 79Y 104,
http://journalofvision.org/2/1/6/, doi:10.1167/2.1.6.
[PubMed][Article]
Neri, P., & Heeger, D. (2002). Spatiotemporal mecha-
nisms for detecting and identifying image features in
human vision. Nature Neuroscience, 5, 812Y816.
[PubMed][Article]
Neri, P., Parker, A. J., & Blakemore, C. (1999). Probing
the human stereoscopic system with reverse correla-
tion. Nature, 401, 695Y 698. [PubMed]
Ohzawa, I., DeAngelis, G. C., & Freeman, R. D. (1990).
Stereoscopic depth discrimination in the visual
cortex: Neurons ideally suited as disparity detectors.
Science, 249, 1037Y 1041. [PubMed]
Journal of Vision (2005) 5, 659–667 Chauvin et al. 666
Olman, C., & Kersten, D. (2004). Classification objects,
ideal observers & generative models. Cognitive
Science, 28, 141Y 146. [Abstract]
Poline, J-B., & Mazoyer, B. M. (1994). Analysis of
individual brain activation maps using hierarchical
description and multiscale detection. IEEE Transac-
tions on Medical Imaging, 13(4), 702Y 710. [Abstract]
Poline, J-B., Worsley, K. J., Evans, A. C., & Friston, K. J
(1997). Combining spatial extent and peak intensity
to test for activations in functional imaging. Neuro-
image, 5, 83Y 96. [PubMed]
Ringach, D., & Shapley, R. (2004). Reverse correlation in
neurophysiology. Cognitive Science, 28, 247Y 166.
[Abstract]
Schyns, P. G., Bonnar, L., & Gosselin, F. (2002). Show
me the features! Understanding recognition from the
use of visual information. Psychological Science, 13,
402Y 409. [PubMed]
Schyns, P. G., Jentzsch, I., Johnson, M., Schweinberger,
S. R., & Gosselin, F. (2003). A principled method for
determining the functionality of ERP components.
Neuroreport, 14, 1665Y 1669. [PubMed]
Sekuler, A. B., Gaspar, C. M., Gold, J. M., & Bennett,
P. J. (2004). Inversion leads to quantitative changes
in faces processing. Current Biology, 14, 391Y 396.
[PubMed]
Siegmund, D. O., & Worsley, K. J. (1995). Testing for a
signal with unknown location and scale in a sta-
tionary Gaussian random field. Annals of Statistics,
23, 608Y 639.
Simpson, W. A., Braun, J., Bargen, C., & Newman, A.
(2000). Identification of the eyeYbrainYhand system
with point processes: A new approach to simple
reaction time. Journal of Experimental Psychology:
Human Perception and Performance, 26, 1675Y 1690.
[PubMed]
Smith, M. L., Gosselin, F., & Schyns, P. G. (2004). Re-
ceptive fields for flexible face categorizations. Psy-
chological Science, 15, 753Y 761. [PubMed]
Smith, M. L., Cottrell, G., Gosselin, F., & Schyns, P. G.
(2005). Transmitting and decoding facial expressions
of emotions. Psychological Science, 16, 184Y189.
Sutter, E. E., & Tran, D. (1992). The field topography
of ERG components in manYI: The photopic lumi-
nance response. Vision Resear ch, 32, 433Y 446. [PubMed]
Taylor, J. E., Worsley, K. J., & Gosselin, F. (2005).
Maxima of discretely sampled random fields, with an
application to Fbubbles_. Submitted for publication.
Thomas, J. P., & Knoblauch, K. (1998). What do viewers
look for when detecting a luminance pulse?
[Abstract]. Investigative Opthalmology and Visual
Science, 39, S404.
Vinette, C., Gosselin, F., & Schyns, P. G. (2004). Spatio-
temporal dynamics of face recognition in a flash: It’s
in th eyes! Cognitive Science, 28, 289Y 301. [Abstract]
Watson, A. B. (1998). MultiYcategory classification: Tem-
plate models and classification images [Abstract].
Investigative Opthalmology and Visual Science, 39,
S912.
Watson, A. B., & Rosenholtz, R. (1997). A Rorschach test
for visual classification strategies [Abstract]. Inves-
tigative Opthalmology and Visual Science, 38, S1.
Worsley, K. J. (1994). Local maxima and the expected
Euler characteristic of excursion sets of
2
, F and t
fields. Advances in Applied Probability, 26, 13Y 42.
Worsley, K. J. (1995a). Boundary corrections for the ex-
pected Euler characteristic of excursion sets of random
fields, with an application to astrophysics. Advances in
Applied Probability, 27, 943Y 959.
Worsley, K. J. (1995b). Estimating the number of peaks in
a random field using the Hadwiger characteristic of
excursion sets, with applications to medical images.
Annals of Statistics, 23, 640Y 669.
Worsley, K. J. (1996). The geometry of random images.
Chance, 9, 27Y 40.
Worsley, K. J. (2003). FMRISTAT: A general statistical
analysis for fMRI data. Retrieved from http://
www.math.mcgill.ca/keith/fmristat/.
Worsley, K. J. (2005) An improved theoretical P-value for
SPMs based on discrete local maxima. Manuscript in
preparation
.
Worsley, K. J., Marrett, S., Neelin, P., Vandal, A. C.,
Friston, K. J., & Evans, A. C. (1996). A unified
statistical approach for determining significant sig-
nals in images of cerebral activation. Human Brain
Mapping, 4, 58Y 73.
Journal of Vision (2005) 5, 659–667 Chauvin et al. 667

Supplementary resource (1)

... Aggregate masks were weighted by participant performance to avoid unduly influencing group effect due to outlier performance. Significance testing of classification images was performed using the Stat4Ci toolbox for significance testing and cluster thresholding of classification images (Chauvin et al., 2005). ...
... Image Classification Analysis: Statistical testing of these response maps was performed using the random field theory cluster analysis toolbox Stat4Ci (Chauvin et al., 2005). This toolbox is designed for statistical testing of classification images lacking a priori assumptions, particularly when using bubbled stimuli and reverse correlation paradigms. ...
... Our stimuli was bookended by 100ms within noise to create a truly non-informative, baseline search area. Threshold values are recommended to be greater than t = 2.3 in Chauvin et al. (2005), and we set our thresholding t > 2.5 and probability of p < 0.05. These binary statistical classification images (0 non-significant, 1 significant) then give us significant clusters that we assume are representative of spectrotemporal regions that are informative for accurate affect identification per stimuli and are taken through to the fMRI reverse correlation analysis as described below. ...
Preprint
Full-text available
Decoding affect information encoded within a vocally produced signal is a key part of daily communication. The acoustic channels that carry the affect information, however, are not uniformly distributed across a spectrotemporal space meaning that natural listening environments with dynamic, competing noise may unpredictably obscure some spectrotemporal regions of the vocalisation, reducing the potential information available to the listener. In this study, we utilise behavioural and functional MRI investigations to first assess which spectrotemporal regions of a human vocalisation contribute to affect perception in the listener, and then use a reverse-correlation fMRI analysis to see which structures underpin this perceptually challenging task when categorisation relevant acoustic information is unmasked by noise. Our results show that, despite the challenging task and non-uniformity of contributing spectral regions of affective vocalizations, a distributed network of (non-primary auditory) brain regions in the frontal cortex, basal ganglia, and lateral limbic regions supports affect processing in noise. Given the conditions for recruitment and previously established functional contributions of these regions, we propose that this task is underpinned by a reciprocal network between frontal cortical regions and ventral limbic regions that assist in flexible adaptation and tuning to stimuli, while a hippocampal and parahippocampal regions support the auditory system’s processing of the degraded auditory information via associative and contextual processing.
... CIs of processing efficiency as a function of time for each SF condition. The two horizontal dashed red lines indicate the significance criterion which was determined by the Pixel test29 . All portions of a curve above the upper threshold indicate a processing efficiency significantly superior to zero whereas those below the lower threshold indicate a processing efficiency significantly below zero.Figure 2. CIs of processing efficiency as a function of time and stimulus oscillation frequencies for each SF condition. ...
... All portions of a curve above the upper threshold indicate a processing efficiency significantly superior to zero whereas those below the lower threshold indicate a processing efficiency significantly below zero.Figure 2. CIs of processing efficiency as a function of time and stimulus oscillation frequencies for each SF condition. Points in the CIs that are colored white do not differ significantly from zero, as determined by the Pixel test29 . ...
Article
Full-text available
This study examined the temporal profile of spatial frequency processing in a word reading task in 16 normal adult readers. They had to report the word presented in a 200 ms display using a four-alternative forced-choice task (4AFC). The stimuli were made of an additive combination of the signal (i.e. the target word) and of a visual white noise patch wherein the signal-to-noise ratio varied randomly across stimulus duration. Four spatial frequency conditions were defined for the signal component of the stimulus (bandpass Butterworth filters with center frequencies of 1.2, 2.4, 4.8 and 9.6 cycles per degree). In contrast to the coarse-to-fine theory of visual recognition, the results show that the highest spatial frequency range dominates early processing, with a shift toward lower spatial frequencies at later points during stimulus exposure. This pattern interacted in a complex way with the temporal frequency content of signal-to-noise oscillations. The outcome of individual data patterns classification by a machine learning algorithm according to the corresponding spatial frequency band further shows that the most salient spatial frequency signature is obtained when the time dimension within data patterns is recoded into its Fourier transform.
... In the widely used "subsequent memory paradigm", for example, the brain activity elicited by to-be-encoded stimuli is contrasted for correctly and incorrectly retrieved stimuli on a subsequent memory test (for a review, see Wagner et al., 1999). Likewise, in studies that employ the Bubbles technique, participants are asked to complete a task on partially revealed stimuli on every trial, and the information that leads to incorrect responses is subtracted from the information that leads to correct responses (Chauvin et al., 2005). Here, we present a novel procedure that can increase signal-to-noise ratio (SNR) in such psychological experiments that use accuracy as a selection variable for another dependent variable. ...
... To reveal the use of information of a particular participant, a classification image (CI) was computed. It consists, essentially, of averaging the bubble masks associated with correct responses, on the one hand, and those associated with incorrect responses, on the other hand, and in subtracting element-by-element the latter from the former (Chauvin et al., 2005). This is the ideal experiment to test the reclassification procedure. ...
Article
Full-text available
This paper introduces a novel procedure that can increase the signal-to-noise ratio in psychological experiments that use accuracy as a selection variable for another dependent variable. This procedure relies on the fact that some correct responses result from guesses and reclassifies them as incorrect responses using a trial-by-trial reclassification evidence such as response time. It selects the optimal reclassification evidence criterion beyond which correct responses should be reclassified as incorrect responses. We show that the more difficult the task and the fewer the response alternatives, the more to be gained from this reclassification procedure. We illustrate the procedure on behavioral and ERP data from two different datasets (Caplette et al. NeuroImage 218, 116994, 2020; Faghel-Soubeyrand et al. Journal of Experimental Psychology: General 148, 1834–1841, 2019) using response time as reclassification evidence. In both cases, the reclassification procedure increased signal-to-noise ratio by more than 13%. Matlab and Python implementations of the reclassification procedure are openly available (https://github.com/GroupeLaboGosselin/Reclassification).
... Once transformed to a common scale, the individual CIs were averaged, smoothed, and then submitted to a two-way Pixel test (Chauvin et al., 2005) with α = 0.05 to determine the points in CIs that differed significantly from zero. The Pixel test is derived from random field theory and has been applied for about the last 30 years for the analysis of brain imaging data. ...
Article
Full-text available
The temporal features of visual processing were compared between young and elderly healthy participants in visual object and word recognition tasks using the technique of random temporal sampling. The target stimuli were additively combined with a white noise field and were exposed very briefly (200 ms). Target visibility oscillated randomly throughout exposure duration by manipulating the signal-to-noise ratio (SNR). Classification images (CIs) based on response accuracy were calculated to reflect processing efficiency according to the time elapsed since target onset and the power of SNR oscillations in the 5–55 Hz range. CIs differed substantially across groups whereas individuals of the same group largely shared crucial features such that a machine learning algorithm reached 100% accuracy in classifying the data patterns of individual participants into their proper group. These findings demonstrate altered perceptual oscillations in healthy aging and are consistent with previous investigations showing brain oscillation anomalies in the elderly.
... In one pioneering study, Gosselin & Schyns (2003) sought to find representations of letters by asking participants to identify a letter "S" in visual noise stimuli; critically, the stimuli were purely noise (and thus did not contain any actual letters), yet reverse correlation based on subject responses revealed a definitive "S" image recovered from each participant. Reverse correlation has subsequently been demonstrated to avoid the shortcomings of other techniques that bias expectations (Chauvin et al., 2005). Other groups have used reverse correlation to characterize the top-down processes of perception to abstract psychological categories, including the trustworthiness in voices (Ponsot et al., 2018), self-image (Moon et al., 2020), and "male" vs. "female" faces (Brinkman et al., 2017), and shown that reverse correlation can quantitatively compare human decision-making to that of computers (Jäkel et al., 2009;Okazawa et al., 2018). ...
Article
Full-text available
Uncovering cognitive representations is an elusive goal that is increasingly pursued using the reverse correlation method, wherein human subjects make judgments about ambiguous stimuli. Employing reverse correlation often entails collecting thousands of stimulus-response pairs, which severely limits the breadth of studies that are feasible using the method. Current techniques to improve efficiency bias the outcome. Here we show that this methodological barrier can be diminished using compressive sensing, an advanced signal processing technique designed to improve sampling efficiency. Simulations are performed to demonstrate that compressive sensing can improve the accuracy of reconstructed cognitive representations and dramatically reduce the required number of stimulus-response pairs. Additionally, compressive sensing is used on human subject data from a previous reverse correlation study, demonstrating a dramatic improvement in reconstruction quality. This work concludes by outlining the potential of compressive sensing to improve representation reconstruction throughout the fields of psychology, neuroscience, and beyond. Supplementary Information The online version contains supplementary material available at 10.3758/s13428-023-02281-4.
... In other words, because each pixel of the CI is independently tested, we do not have to arbitrarily select facial areas on which to perform the statistical test. Because of the very high number of t-tests performed, the statistical threshold was adjusted to avoid type-I errors, using the Cluster test from the Stat4Ci toolbox (Chauvin et al., 2005). This test compensates for the multiple comparisons across pixels while taking into account the fact that in structured images-like faces-each pixel is not independent of the others. ...
Article
Unlabelled: Effectively communicating pain is crucial for human beings. Facial expressions are one of the most specific forms of behavior associated with pain, but the way culture shapes expectations about the intensity with which pain is typically facially conveyed, and the visual strategies deployed to decode pain intensity in facial expressions, is poorly understood. The present study used a data-driven approach to compare two cultures, namely East Asians and Westerners, with respect to their mental representations of pain facial expressions (experiment 1, N=60; experiment 2, N=74) and their visual information utilization during the discrimination of facial expressions of pain of different intensities (experiment 3; N=60). Results reveal that compared to Westerners, East Asians expect more intense pain expressions (experiments 1 and 2), need more signal, and do not rely as much as Westerners on core facial features of pain expressions to discriminate between pain intensities (experiment 3). Together, those findings suggest that cultural norms regarding socially accepted pain behaviors shape the expectations about pain facial expressions and decoding visual strategies. Furthermore, they highlight the complexity of emotional facial expressions and the importance of studying pain communication in multicultural settings. Supplementary information: The online version contains supplementary material available at 10.1007/s42761-023-00186-1.
Article
Although pain is a commonly experienced and observed affective state, it is frequently misinterpreted, which leads to inadequate caregiving. Studies show the ability at estimating pain in others (estimation bias) and detecting its subtle variations (sensitivity) could emerge from independent mechanisms. While estimation bias is modulated by variables such as empathy level, pain catastrophizing tendency, and overexposure to pain, sensitivity remains unimpacted. The present study verifies if these two types of inaccuracies are partly explained by perceptual factors. Using reverse correlation, we measured their association with participants' mental representation of pain, or more simply put, with their expectations of what the face of a person in pain should look like. Experiment 1 shows that both parameters are associated with variations in expectations of this expression. More specifically, the estimation bias is linked with expectations characterized by salient changes in the middle face region, whereas sensitivity is associated with salient changes in the eyebrow region. Experiment 2 reveals that bias and sensitivity yield differences in emotional representations. Expectations of individuals with a lower underestimation tendency are qualitatively rated as expressing more pain and sadness, and those of individuals with a higher level of sensitivity as expressing more pain, anger, and disgust. Together, these results provide evidence for a perceptual contribution in pain inferencing that is independent of other psychosocial variables and its link to observers' expectations. PERSPECTIVE: This article reinforces the contribution of perceptual mechanisms in pain assessment. Moreover, strategies aimed to improve the reliability of individuals' expectations regarding the appearance of facial expressions of pain could potentially be developed, and contribute to decrease inaccuracies found in pain assessment and the confusion between pain and other affective states.
Article
The classification image (CI) technique has been used to derive templates for judgements of facial emotion and reveal which facial features inform specific emotional judgements. For example, this method has been used to show that detecting an up- or down-turned mouth is a primary strategy for discriminating happy versus sad expressions. We explored the detection of surprise using CIs, expecting widened eyes, raised eyebrows, and open mouths to be dominant features. We briefly presented a photograph of a female face with a neutral expression embedded in random visual noise, which modulated the appearance of the face on a trial-by-trial basis. In separate sessions, we showed this face with or without eyebrows to test the importance of the raised eyebrow element of surprise. Noise samples were aggregated into CIs based on participant responses. Results show that the eye-region was most informative for detecting surprise. Unless attention was specifically directed to the mouth, we found no effects in the mouth region. The eye effect was stronger when the eyebrows were absent, but the eyebrow region was not itself informative and people did not infer eyebrows when they were missing. A follow-up study was conducted in which participants rated the emotional valence of the neutral images combined with their associated CIs. This verified that CIs for 'surprise' convey surprised expressions, while also showing that CIs for 'not surprise' convey disgust. We conclude that the eye-region is important for the detection of surprise.
Article
Full-text available
Reviews the book, Human brain function by R. S. J. Frackowiak, K. J. Friston, C. D. Frith, R. J. Dolan, and J. C. Mazziotta (1997). A significant recent development in the rapidly evolving field of imaging technology has been the relative shift of emphasis from positron emission tomography (PET) activation studies to functional magnetic resonance imaging (MRI). Not only is MRI technology more widely available than PET, it is less invasive and can benefit from improved temporal and spatial resolution. It seems timely, therefore, to take stock and ask what has been learned from human brain imaging so far. This volume, by Frackowiak and colleagues from the Functional Imaging Laboratory in London, seeks to do that through a comprehensive overview of the group's own work in the field to date; that is, mainly through the use of PET. Physiologists, in particular, may find the title somewhat misleading; "Human Brain Function" focuses almost exclusively on the results of PET activation studies, while relevant electrophysiological studies are only given detailed consideration in one chapter on the visual system. In spite of this, the book does demonstrate, unequivocally, that functional neuroimaging has had a dramatic impact on the study of human brain function. What also continues to be clear, however, is that the gulf between technological know how and depth of understanding has never been wider. (PsycINFO Database Record (c) 2012 APA, all rights reserved)
Article
Full-text available
Nonlinear contributions to pattern classification by humans are analyzed by using previously obtained data on discrimination between aligned lines and offset lines. We how that the optimal linear model can be rejected even when the parameters of the model are estimated individually for each observer. We use a new measure of agreement to reject the linear model and to test simple nonlinear operators. The first nonlinearity is position uncertainty. The linear kernels are shrunk to different extents and convolved with the input images. A Gaussian window weights the results of the convolutions and the maximum in that window is selected as the internal variable. The size of the window is chosen such as to maintain a constant total amount of spatial filtering, i.e., the smaller kernels have a larger position uncertainty. The result of two observers indicate that the best agreement is obtained at a moderate degree of position uncertainty, plus-minus one min of arc. Finally, we analyze the effect of orientation uncertainty and show that agreement can be further improved in some cases.
Article
We present a unified statistical theory for assessing the significance of apparent signal observed in noisy difference images. The results are usable in a wide range of applications, including fMRI, but are discussed with particular reference to PET images which represent changes in cerebral blood flow elicited by a specific cognitive or sensorimotor task. Our main result is an estimate of the P-value for local maxima of Gaussian, t, χ2 and F fields over search regions of any shape or size in any number of dimensions. This unifies the P-values for large search areas in 2-D (Friston et al. [1991]: J Cereb Blood Flow Metab 11:690–699) large search regions in 3-D (Worsley et al. [1992]: J Cereb Blood Flow Metab 12:900–918) and the usual uncorrected P-value at a single pixel or voxel.
Article
This book deals primarily with the sample function behaviour of Gaussian, and related, random fields; i.e. stochastic processes whose arguments vary in a continuous fashion over some subset of ℛ N , N-dimensional Euclidean space. The problems that arise in describing this behaviour in the multiparameter setting are qualitatively different to those covered by the one-dimensional theory. Indeed, most of the really interesting problems are of a geometrical nature, and actually disappear in the simple case. The purpose of this book is to collect within one cover most of the contents of the substantial literature devoted to the sample function analysis of random fields, including a reasonably full and self-contained account of the geometry needed for its understanding. While working on this book I tried to write with two readers in mind. One was a mathematician, for whom it was necessary to present what I feel is a rather beautiful theory, incorporating the standards of rigour and exactness that mathematicians are accustomed to. The other reader was an applied scientist, working in one of the many subjects that employs random fields as an element of mathematical modelling. His needs are quite different, for he requires a full statement of useful results, presented in such a way that he is not required to follow long mathematical proofs in order to understand and use them. It is possible that by attempting to please both readers I will have satisfied neither. Nevertheless, I beg the patience of both readers when they come upon parts of the text obviously written for the other. To be able to read the book from cover to cover a reader will need a good working knowledge (at graduate student level) of the basic features of modern probability theory and, for Chapter 8, measure theory. Although a brief review of those aspects of probability theory required is given in Chapter 1, it is unreasonable to assume that this can serve as anything other than a convenient reference for a reader already familiar with this material. However, with the applied scientist in mind, I have attempted to present the main results of the book in such a fashion that they can be understood and used by a reader whose knowledge of probability incorporates little more than the notions of expectation and stochastic convergence.
Article
Here we demonstrate a method for constructing stimulus classification images. These images provide information regarding the stimulus aspects the observer uses to segregate images into discrete response categories. Data are first collected on a discrimination task containing low contrast noise. The noises are then averaged separately for the stimulus-response categories. These averages are then summed with appropriate signs to obtain an overall classification image. We determine stimulus classification images for a vernier acuity task to visualize the stimulus features used to make these precise position discriminations. The resulting images reject the idea that the discrimination is performed by the single best discriminating cortical unit. The classification images show one Gabor-like filter for each line, rejecting the nearly ideal assumption of image discrimination models predicting no contribution from the fixed vernier line.