ArticlePDF Available

Sustained extrastriate cortical activation without visual awareness revealed by FMRI studies of hemianopic patients

Authors:

Abstract and Figures

Patients with lesions in the primary visual cortex (V1) may show processing of visual stimuli presented in their field of cortical blindness even when they report being unaware of the stimuli. To elucidate the neuroanatomical basis of their residual visual functions, we used functional magnetic resonance imaging in two hemianopic patients, FS and GY. In the first experiment, a rotating spiral stimulus was used to assess the responsiveness of dorsal stream areas. Although no response was detectable within denervated or destroyed early visual cortex, motion-sensitive areas (hMT+/V5) ipsilateral to the lesion showed a strong sustained hemodynamic response. In GY, this activation was at least as strong as that of his contralesional hMT+/V5 to the stimulus in the normal hemifield. In the second experiment, coloured images of natural objects were used to assess the responsiveness of ventral stream areas. Again, no activity was detectable in ipsilesional early visual areas, but extrastriate areas in the lateral occipital cortex (hMT+/V5 and LO) and within the posterior fusiform gyrus (V4/V8) showed a robust sustained hemodynamic response. In both experiments, we observed that ipsilesional areas responded to stimuli presented in either hemifield, whereas the normal hemisphere responded preferentially to stimuli in the sighted hemifield. As only one subject occasionally noticed the onset of stimulation in the impaired field, the unexpectedly strong sustained activity in ipsilesional dorsal and ventral cortical areas appears to be insufficient to generate conscious vision.
Content may be subject to copyright.
Vision Research 41 (2001) 14591474
Sustained extrastriate cortical activation without visual awareness
revealed by fMRI studies of hemianopic patients
Rainer Goebel
a,
*, Lars Muckli
b
, Friedhelm E. Zanella
c
, Wolf Singer
b
, Petra Stoerig
d
a
Department of Cogniti6e Neuroscience,Faculty of Psychology,Maastricht Uni6ersity,Postbus
616
,NL-
6200
MD Maastricht,The Netherlands
b
Max Planck Institute for Brain Research,Deutschordenstr.
46
,D-
60528
Frankfurt a.M., Germany
c
Klinikum der Johann Wolfgang Goethe-Uni6ersita¨t,Institut fu¨r Neuroradiologie,
60528
Frankfurt a.M., Germany
d
Institute of Experimental Psychology II,Heinrich-Heine -Uni6ersity,Uni6ersita¨tsstr.
1
,D-
40225
Du¨ sseldorf,Germany
Received 23 January 2001; received in revised form 15 February 2001
Abstract
Patients with lesions in the primary visual cortex (V1) may show processing of visual stimuli presented in their field of cortical
blindness even when they report being unaware of the stimuli. To elucidate the neuroanatomical basis of their residual visual
functions, we used functional magnetic resonance imaging in two hemianopic patients, FS and GY. In the first experiment, a
rotating spiral stimulus was used to assess the responsiveness of dorsal stream areas. Although no response was detectable within
denervated or destroyed early visual cortex, motion-sensitive areas (hMT+/V5) ipsilateral to the lesion showed a strong sustained
hemodynamic response. In GY, this activation was at least as strong as that of his contralesional hMT+/V5 to the stimulus in
the normal hemifield. In the second experiment, coloured images of natural objects were used to assess the responsiveness of
ventral stream areas. Again, no activity was detectable in ipsilesional early visual areas, but extrastriate areas in the lateral
occipital cortex (hMT+/V5 and LO) and within the posterior fusiform gyrus (V4/V8) showed a robust sustained hemodynamic
response. In both experiments, we observed that ipsilesional areas responded to stimuli presented in either hemifield, whereas the
normal hemisphere responded preferentially to stimuli in the sighted hemifield. As only one subject occasionally noticed the onset
of stimuluation in the impaired field, the unexpectedly strong sustained activity in ipsilesional dorsal and ventral cortical areas
appears to be insufficient to generate conscious vision. © 2001 Elsevier Science Ltd. All rights reserved.
Keywords
:
Cortical blindness; Blindsight; Hemianopia; Visual cortical areas; Visual awareness; fMRI; Lesion
www.elsevier.com/locate/visres
1. Introduction
The term ‘blindsight’ describes the ability of neuro-
logical patients with postgeniculate lesions to detect,
localise and discriminate visual stimuli in blind parts of
the visual field, although they do not consciously per-
ceive them (Weiskrantz, Warrington, Sanders, & Mar-
shall, 1974). These residual visual functions must be
mediated by those parts of the visual system that retain
visual responsiveness despite the lesion that destroys or
denervates primary visual cortex (V1) and causes corti-
cal blindness. Data from monkeys have shown that
despite retrograde degeneration of projection cells in
the dorsal lateral geniculate nucleus (dLGN) and the
retina (Van Buren, 1963; Cowey, Stoerig, & Perry,
1989), subcortical retinorecipient nuclei continue to re-
ceive visual input from the parts of the retinae that
represent the cortically blind visual field (for a recent
review, see Stoerig & Cowey, 1997). Some of these
nuclei project directly to extrastriate visual cortex,
which consists of cytoarchitectonically distinct areas
that subserve specific visual functions (Zeki, 1974; Zilles
& Clarke, 1997). Evidence indicates that these areas are
organized within two main visual pathways (Mishkin,
Ungerleider, & Macko, 1983): a ventral stream or
‘what’ system devoted to the fine-grained analysis of the
visual scene, including processing of shape and colour
and a dorsal stream or ‘where’ system, which processes
spatial characteristics of the visual scene and analyses
motion. Emphasizing that the latter pathway is also
involved in visuomotor transformations, Milner and
* Corresponding author. Tel.: +49-69-96769471; fax: +49-69-
96769327.
E-mail address
:
r.goebel@psychology.unimaas.nl (R. Goebel).
0042-6989/01/$ - see front matter © 2001 Elsevier Science Ltd. All rights reserved.
PII: S0042-6989(01)00069-4
R.Goebel et al.
/
Vision Research
41 (2001) 14591474
1460
Goodale (1995) proposed to distinguish between vision
for perception(ventral pathway) and vision for action
(dorsal pathway). The two pathways originate in areas
V1 and V2 and extend into the temporal (the what
system) and parietal (the wheresystem) lobe, respec-
tively. With brain imaging techniques, they have been
convincingly demonstrated in the normal human brain
(e.g. Haxby et al., 1991) and several specialized areas
have been identied on the basis of their stimulus
preference proles.
An extensively studied area in the ventral stream is
area V4, which in the macaque, appears to be involved
in the processing of chromatic and shape information
(see Cowey, 1994 for a review). It has been reported to
mediate colour constancy, but nevertheless, its necessity
for colour vision is debated on the grounds that abla-
tion of this area does not produce marked decits in
colour discrimination (Heywood & Cowey, 1987). In
man, the colour complex has been located in the
fusiform gyrus (Lueck et al., 1989; Hadjikhani, Liu,
Dale, Cavanagh, & Tootell, 1998), in a region that is
generally compromised in patients with achromatopsia
(Meadows, 1974) and is often referred to as V4 (McK-
eefry & Zeki, 1997; Zeki, McKeefry, Bartels, & Frack-
owiak, 1998) or V8 (Hadjikhani et al., 1998); we shall
refer to it here as the colour complexor V4/V8. Other
areas located in close proximity to area V4/V8 have
been characterized, including the fusiform face area
(FFA, Kanwisher, Dermott, & Chun, 1997) and several
overlapping regions sensitive to different object cate-
gories (e.g. Ishai, Ungerleider, Martin, Schouten &
Haxby, 1999). Less specialized, with respect to object
categories but responsive to common objects, abstract
sculptures and famous faces is a more dorsal area in a
region at the lateral-anterior aspect of the occipital lobe
that Malach et al. (1995) called LO (lateral occipital
complex).
Of the dorsal stream areas, those that are specialized
for motion processing have received particular atten-
tion. In monkey cortex, the motion-selective area MT
(middle temporal area) is located in the middle tempo-
ral sulcus, close to the junction of the occipital, tempo-
ral and parietal lobes (Van Essen, Maunsell & Bixby,
1981). Its cells are responsive to the direction and speed
of moving stimuli (Snowden, Treue & Andersen, 1992)
and their inactivation causes deciencies in motion
direction discrimination. Area MT is easily identied in
histological sections because of its high myelination and
it is surrounded by several other specialized areas
(satellites) that process higher-order aspects of stimu-
lus motion. The human visual cortex contains an area
that is likely to correspond to monkey area MT and is
located in the occipito-temporo-parietal pit. Bilateral
lesions, that include this area, cause a severe impair-
ment in detecting the movement of objects (cortical
akinetopsia, see Zeki, 1991). Imaging studies have
shown that metabolism and blood ow in that region
increases more in response to moving than to stationary
stimuli, indicating preferential involvement of this area
in the processing of motion information (e.g. Watson et
al., 1993; Tootell et al., 1995; Goebel, Khorram-Sefat,
Muckli, Hacker, & Singer, 1998a). As functional imag-
ing studies have not yet convincingly differentiated
between the human homologue of area MT and its
satellites, this motion-selective region is generally re-
ferred to as the human motion complex(hMT+)or
as V5.
Although area MT receives its major input directly
or indirectly from the primary visual cortex (V1), de-
struction or temporary inactivation of V1 in monkeys
does not eliminate its visual responsiveness (Rodman,
Gross & Albright, 1989). Although the responses were
markedly reduced, a large part of the neuronal popula-
tion remained visually responsive and even retained
directional tuning. Accordingly, one might expect resid-
ual visual responsiveness in the motion complex of
human patients who have suffered lesions to the pri-
mary visual cortex, and indeed, functional imaging
studies in patients with lesions of the optic radiation or
primary visual cortex have conrmed its continued
responsiveness to moving or ickering stimuli (Barbur,
Watson, Frackowiak, & Zeki, 1993; Stoerig, Goebel,
Muckli, Hacker, & Singer, 1997; Zeki & ffytche, 1998).
In contrast to dorsal stream areas, neurons in the
monkeys ventral stream become unresponsive to stim-
uli in the blind eld after V1 lesions (Girard, Salin &
Bullier, 1991; Bullier, Girard & Salin, 1993). Thus, one
would expect little, if any, activity in areas of the
ventral stream in patients with lesions to the primary
visual cortex. However, in a recent imaging study of
two hemianopic patients, we have found that presenta-
tion of images of natural objects in the cortically blind
visual eld can lead to strong sustained activity in the
ventral pathway without any detectable activity in pri-
mary visual cortex (Goebel et al., 1998b; Goebel,
Muckli, Zanella, Singer, & Stoerig, submitted).
In the present study, we compare the responsiveness
of dorsal and ventral (Goebel et al., submitted) stream
areas after blind eld stimulation in the two hemianopic
patients, GY and FS, as measured by echoplanar func-
tional magnetic resonance imaging. To activate dorsal
stream areas, a rotating spiral stimulus was presented in
the cortically blind visual eld. We used a rotating
spiral instead of translational motion (Sahraie et al.,
1997; Zeki & ffytche, 1998) because it restricts motion
to a limited region of space for prolonged stimulation
epochs (e.g. 30 s). This allowed us to separate the
hemodynamic response to stimulus onset, which GY
can be aware of (see Barbur, Ruddock, & Watereld,
1980) from the sustained response to the rotatory mo-
tion. For activation of ventral stream areas, coloured
images of natural objects (fruit and vegetables) were
R.Goebel et al.
/
Vision Research
41 (2001) 14591474
1461
used (Goebel et al., submitted). Both stimulus types
were also presented in the sighted visual eld of the
patients to allow for a direct within-subject comparison
of the hemodynamic response to stimulation in the
cortically blind and in the intact visual eld. As a
further control, the same experiments were conducted
with two healthy subjects. In addition, retinotopic map-
ping experiments were performed to describe the precise
spatial layout of the functional responsiveness of early
visual areas in the normal and lesioned hemisphere.
2. Methods
2
.
1
.Patients
FS and GY, who participated in this study, have
long-standing post-geniculate lesions of the left hemi-
sphere (see Fig. 1). FS, born in 1937, suffered a severe
craniocerebral trauma when he was 42 years of age.
GY, born in 1956, was involved in a trafc accident
when he was 8 years, and his left primary visual cortex
was almost completely destroyed by a vascular incident.
The visual eld defects that resulted from the lesions
are shown in Fig. 2. The wedge-shaped region in FSs
right hemield (black) is absolutely blind, while GYs
hemianopia is relative and allows conscious detection of
salient visual events. Over many years, GY and FS
have participated in studies of their residual visual
functions (e.g. for GY: Barbur et al., 1980; Blythe,
Bromley, Kennard, & Ruddock, 1986; Brent, Kennard,
& Ruddock, 1994; Weiskrantz, Barbur, & Sahraie,
1995; Morland, Ogilvie, Ruddock, & Wright, 1996;
Finlay, Jones, Morland, Ogilvie, & Ruddock, 1997;
Marcel, 1998; Guo, Benson, & Blakemore, 1998; Ben-
son, Guo, & Blakemore, 1998; Kentridge, Heywood, &
Weiskrantz, 1999 and for FS: Po¨ppel, 1985, 1986; Sto-
erig, 1993; Stoerig & Cowey, 1997; Stoerig, Klein-
schmidt, & Frahm, 1998).
The lesions were visualized by MR-imaging (Fig. 1)
using a high-resolution T1-weighted 3D FLASH se-
quence (voxel size 1×1×1 mm) recorded with a 1.5 T
scanner (Siemens Magnetom Vision). FSs extensive
lesion primarily affects the temporal lobe and includes
the optic radiation in the vicinity of the LGN. In GY,
the medial occipital lobe of the left hemisphere is
destroyed, including most of V1, as well as surrounding
extrastriate visual cortex and the underlying white mat-
ter, but sparing the occipital pole.
2
.
2
.Tasks
2
.
2
.
1
.Retinotopic mapping
The responsiveness and delineation of early visual
areas V1, V2, V3, VP, V3A, V4v and V4d was investi-
gated with retinotopic mapping scans (cf. Sereno, Dale,
Reppas, Kwong, Belliveau, Brady, Rosen, & Tootell,
1995), sampling 12 contiguous slices approximately in
parallel to the calcarine ssure. Eccentricity and polar
angle mapping experiments were performed in the same
recording session (for details see Linden, Kallenbach,
Heinecke, Singer, & Goebel, 1999). For eccentricity
mapping, a ring-shaped conguration of black and
white contrast-reversing (8 Hz) checkers was presented
centered around the xation point. The ring started
with a radius of 1°visual angle and expanded to a
radius of 12°within 96 s. For polar angle mapping, the
same pattern was congured as a wedge subtending
22.5°in polar angle with the tip at the xation cross.
The wedge started at the left horizontal meridian and
slowly rotated clockwise for a full cycle of 360°within
96 s. Each mapping experiment consisted of four repeti-
tions of a full expansion and a full rotation.
2
.
2
.
2
.Mapping of the motion complex
Motion-selective areas were identied by comparing
the hemodynamic response during presentation of a
moving stimulus with the response during presentation
of a stationary control stimulus. We used a oweld
stimulus, consisting of 400 white dots moving radially
outward on a black background (visual eld: 28°wide
by 20°high, dot size: 0.06×0.06°, dot velocity: 3.6
14.4°/s). It was contrasted with a stationary dot display
to produce a clear response of motion-sensitive areas
(Tootell et al., 1995).
2
.
2
.
3
.Motion experiment
(
Experiment
1)
While the oweld stimulated both hemields simul-
taneously, the main stimulus was presented 6.8°off-axis
in one hemield at a time. It consisted of a blue-and-
red (run 1) or black-and-white (run 2) checkered spiral
(Fig. 2A,C), 7°in diameter, rotating counter-clockwise
around its centre (180°/s). Each stimulation block
lasted for 30 s and was repeated four times within each
run. In an additional condition, a stationary spiral was
shown four times in four blocks of 30 s. All stimulation
blocks were separated by xation blocks of equal
length. The stimulus was presented on a grey back-
ground to prevent light scatter from the stimulus into
the sighted hemield. A schematic drawing of the stim-
ulus conguration is shown in Fig. 2(A,C).
2
.
2
.
4
.Object experiment
(
Experiment
2)
Coloured images of natural objects (fruit and vegeta-
bles) subtending 5.2×5.2°were presented in the left
(normal) or right (impaired) upper visual eld, 7°off-
axis (Fig. 2B). Each stimulation block lasted for 30 s
and was repeated four times within each recording
session. Stimulation blocks were separated by xation
blocks of equal length. Within a stimulation block, an
image was shown for 1 s and then replaced by the next
image without an interstimulus interval. Images con-
R.Goebel et al.
/
Vision Research
41 (2001) 14591474
1462
Fig. 1. Lesions of patients GY (A) and FS (B). (A) Slices of an anatomical T1-weighted MRI scan showing the V1 lesion of GY. The small cross
in (13) indicates the 3D position of the selected slices, the yellow curve delineate the extend of the lesion within the respective slices. (1) Sagittal
slice through the lesion. (2) Axial slice through the lesion. (3) Coronal slice through the lesion. (4) Surface rendering of GYs reconstructed head
(yellow) and of the boundary of the V1 lesion (red). The rendered head was partially opened to reveal the size and position of the lesion. (B)
Rendering of the reconstructed cortical surface at the boundary between grey and white matter of the affected left hemisphere of patient FS. An
image containing a single slice running through the lesion is also shown.
R.Goebel et al.
/
Vision Research
41 (2001) 14591474
1463
Fig. 2. Plots of visual eld sensitivity (A,B) which are based on a combination of static and dynamic perimetry (target: 116, 320 cyc deg
1
m
2
,
background: 10 cyc deg
1
m
2
; white: normal eld, black: absolutely blind eld; grey: relatively blind region, see text). (A) Visual eld sensitivity
plot with indicated position of the moving spiral stimulus for GY (left) and FS (right). (B) Visual eld sensitivity plot with indicated stimulation
position in the objects experiment for GY (left) and FS (right). Stimuli were shown in separate blocks in either the left or the right visual eld.
(C) Stimulus size and location in the spiral experiment (left) and in the objects experiment (right).
R.Goebel et al.
/
Vision Research
41 (2001) 14591474
1464
sisted of coloured drawings of natural objects (fruit and
vegetables, Corel Draw clip art gallery). The stimuli
were presented on a white background to minimize
stray light falling into the sighted hemield. A sche-
matic drawing of the stimulus conguration is shown in
Fig. 2(B,C).
In order to assess whether the subjects were aware of
the stimuli presented in the cortically blind hemield,
we asked them to respond by button press when they
detected anything in the right visual hemield.
2
.
3
.MRI acquisition
Functional magnetic resonance imaging was per-
formed at 1.5 T (Siemens Magnetom Vision) using the
standard head coil and a gradient echo EPI sequence.
The Siemens Magnetom gradient overdrive allowed
functional scans with high spatial and temporal resolu-
tion (TE=69 ms, TR =3000 ms, FA =90°, FOV =
200×200 mm
2
, matrix size: 128×128, voxel size:
1.6×1.6 ×35mm
3
). A T1-weighted 3D MP RAGE
scan lasting 8 min was performed in the same session
(voxel size 1×1×1mm
3
). An additional T1-weighted
3D data set tuned to optimize the contrast between
gray and white matter was recorded in a separate
recording session lasting 24 min (FLASH sequence).
The intrasession MP RAGE 3D data set was automati-
cally aligned with the extrasession T1 FLASH data set.
Visual stimuli were delivered under computer control
(Digital DECpc Celebris XL 590) to a high-luminance
LCD projector (EIKI LC-6000). The image was back-
projected onto a frosted screen positioned at the foot of
the scanner. Visual stimuli were generated using the
Microsoft DirectX graphics library and a Matrox Mys-
tique graphics board.
2
.
4
.Data analysis
FMRI data analysis was performed using BrainVoy-
ager 3.×/2000 (Brain Innovation, Maastricht, The
Netherlands, www.BrainVoyager.com) and included re-
moval of low-frequency drifts, 3D motion detection
and correction, determination of Talairach coordinates,
multiple regression analysis, cortex reconstruction,
ination and attening. For the retinotopic mapping
experiments, cross-correlation analysis was applied (for
details see Linden et al., 1999). The eccentricity and
polar angle represented by a given cortical site was
determined by nding the lag value maximizing the
cross-correlation. The obtained lag values at each
voxel, corresponding to the eccentricity or polar angle
of optimal stimulation, were encoded in pseudocolours
on slices as well as on surface patches (triangles) of the
reconstructed cortical sheet (see below). Pixels were
included into the statistical map if the obtained cross-
correlation value rwas \0.4 (PB0.0001, uncorrected).
In order to detect weak activity within and surrounding
the lesioned or denervated regions, retinotopic mapping
experiments were also analysed with a lowered correla-
tion threshold of r\0.2. Based on the polar angle,
mapping the boundaries of retinotopic cortical areas
V1, V2, V3, VP, V3A and V4v were estimated on the
attened cortical surface maps (see below) of each
patients non-lesioned hemisphere and of both hemi-
spheres in each control subject. The motion complex
mapping experiment was analysed with a simple corre-
lation analysis contrasting the oweld stimulus with
the stationary control stimulus.
The two main experiments were analysed with a
multiple regression model consisting of two predictors,
one for the presentation of the stimulus (moving spiral/
images of natural object) in the left and one for the
presentation of the stimulus in the right visual eld.
The overall model t was assessed using an Fstatistic.
The obtained P-values were corrected for multiple com-
parisons using a cortex-based Bonferroni adjustment,
i.e. the number of voxels included for correction was
limited to grey matter voxels (Goebel & Singer, 1999).
The relative contribution of each of the two predictors
RC=(b
1
b
2
)/(b
1
+b
2
) was visualised with a red-yel-
low-green pseudocolour scale. Statistical maps were
superimposed on the original functional scans and in-
corporated into the high-resolution 3D MRI data sets
through interpolation to the same resolution (voxel size:
1×1×1 mm). Since the 2D functional and 3D struc-
tural measurements were acquired in the same record-
ing session, coregistration of the respective data sets
could be performed on the basis of the Siemens slice
position parameters of the T2*-weighted measurement
(number of slices, slice thickness, distance factor, Tra-
Cor angle, FOV, shift mean, offcenter read, offcenter
phase, in-plane resolution) and the T1-weighted 3D MP
RAGE measurement (number of sagittal partitions,
shift mean, offcenter read, offcenter phase, resolution).
In order to compare activated brain regions across
sessions, anatomical and functional 3D data sets were
transformed into Talairach space (Talairach & Tour-
naux, 1988). Results were visualised by superimposing
3D statistical maps on reconstructions of each patients
cortical sheet. Areas were identied based on their
anatomical position and Talairach coordinates.
The recorded high-resolution T1-weighted 3D
recordings were used for surface reconstruction of both
cortical hemispheres of the control subjects and patients
(for details see Linden et al., 1999; Kriegeskorte &
Goebel, submitted). The white/grey matter border was
segmented with a region-growing method preceded by
inhomogeneity correction of signal intensity across
space. The border of the resulting segmented subvol-
ume was tessellated to produce a polygon-mesh surface
reconstruction of each cortical hemisphere. The tessella-
tion of the white/grey matter boundary of a single
R.Goebel et al.
/
Vision Research
41 (2001) 14591474
1465
hemisphere typically consists of :250000 triangles. An
iterative 3D morphing algorithm (Goebel et al., 1998a)
was used to move the vertices outward along the sur-
face normals into the grey matter. Through visual
inspection, this process was halted when the surface
reached the middle of the gray matter corresponding
approximately to layer 4 of the cortex. The resulting
surface was used as the reference mesh for projecting
functional data on folded, inated or attened repre-
sentations. A morphed surface was always linked to its
folded reference mesh, so that functional data could be
shown at the correct location of an inated or attened
representation. This link was also used to minimize
geometric distortions during ination and attening
(Goebel, 2000) by inclusion of a morphing force that
keeps the distances between vertices of each triangle of
the morphed surface as close as possible to the respec-
tive values of the folded reference mesh.
3. Results
Results from a control subject of the retinotopic
mapping, motion complex mapping and objects experi-
ment are shown in Fig. 3. Results of the two patients
are shown in Figs. 46.
3
.
1
.Retinotopic mapping
The topography of the early visual areas is shown in
Fig. 3(A) for a normal observer. In the patients, the
contralesional V1 appeared normal, but no visual re-
sponsiveness was seen in its lesioned (GY) or dener-
vated (FS) counterpart (r`0.4, PB0.00001). Only a
small region at the occipital pole was activated in GY
(Fig. 5(A), see also Baseler, Morland, & Wandell,
1999). This is in agreement with GYs visual eld
perimetry that shows a 3°macular sparing in the lower
quadrant of the hemianopic right visual eld (Barbur et
al., 1980).
3
.
2
.Motion complex mapping
The comparison of the oweld stimulus with the
stationary dot stimulus revealed several motion-selec-
tive areas in both hemispheres including the motion
complex, area V3A and an area at the border between
occipital and parietal lobes (Fig. 3B). In GY, the mo-
tion complex in the ipsilesional hemisphere is located
more posteriorly and more ventrally than in the con-
tralesional hemisphere. The position of the motion
complex in both patients (Table 1) agrees with previous
reports in normal observers (Tootell et al., 1995; Goe-
bel et al., 1998a).
3
.
3
.Motion experiment
In the control subjects, the foci of activation in the
lateral occipital cortex produced by passive viewing of
the moving spiral were functionally identied as
hMT+/V5 because they overlapped with the regions
activated by the oweld stimulus (not shown). When
the moving spiral stimulus was presented in the corti-
cally blind (right) visual eld of the patients, the ipsile-
sional (left) motion complex was strongly activated
(Fig. 4) despite the fact that there was no detectable
activation within and around the lesioned or dener-
vated region. A weaker but signicant response was
also observed in the contralesional motion complex.
When the spiral stimulus was presented in the left
(sighted) hemield, both the contralesional motion
complex, as well as the ipsilesional motion complex,
responded strongly (see Fig. 4). Thus, the motion com-
plex at the ipsilesional side responded equally strong
(FS) or stronger (GY) to the moving spiral presented in
the cortically blind visual eld, whereas the response of
the contralesional motion complex was stronger to the
spiral in the sighted (left) visual eld in both patients.
In B20% of the presentations of the rotating spiral
in the cortically blind hemield, GY pressed the button
for a short duration following the onset of the spiral,
indicating that he was aware of a stimulus change. An
analysis of the data restricted to the rst 15 s of the 30-s
stimulation epochs produced almost identical results as
the analysis restricted to the last 15 s. FS never indi-
cated awareness when stimuli were presented in his
blind visual eld.
In GY, the comparison of the hemodynamic response
to the presentation of the moving spiral in the cortically
blind visual eld and the sighted visual eld revealed
that the extent of the functionally dened motion com-
plex in the lesioned hemisphere was greater than in the
intact hemisphere (Table 2). Interestingly, this differ-
ence was more pronounced in the black-white spiral
experiment than in the red-blue spiral experiment. The
red-blue spiral, although not isoluminant, possessed
much less luminance contrast than the black-white spi-
ral. We quantied the observation of an extended ip-
silesional motion complex in GY using the number of
activated voxels as well as the maximal correlation
values in the obtained statistical maps (Table 2). Dur-
ing presentation of the moving spiral in the cortically
blind visual eld, the maximum single-voxel correlation
value was r
max
=0.62 (black-white spiral) and r
max
=
0.61 (red-blue spiral) and located in the ipsilesional
motion complex. With a xed single-voxel correlation
value of r\0.4 (PB0.00001, uncorrected), the cluster
size (CS=number of activated voxels) was CS
r\0.4
=
93 (black-white spiral) and CS
r\0.4
=61 (red-blue spi-
ral). During presentation of the moving spiral in the
R.Goebel et al.
/
Vision Research
41 (2001) 14591474
1466
sighted visual hemield, the maximum single-voxel cor-
relation value was r
max
=0.55 (black-white spiral) and
r
max
=0.60 (red-blue spiral) and located in the contrale-
sional motion complex. With a value of CS
r\0.4
=40,
the cluster size was also larger for the red-blue than the
black-white spiral (CS
r\0.4
=26), indicating that in the
normal hemisphere, chromatic contrast is more effec-
tive than in the lesioned one.
Fig. 3. Results of a control subject of the retinotopic mapping experiment (A), motion complex mapping experiment (B) and objects experiment
(C). (A) Boundaries of retinotopic cortical areas were determined through eld sign map computations (cf. Sereno et al., 1995) based on
eccentricity and polar angle mapping. (B) Lateral view of the reconstructed cortical sheet. The statistical map comparing the oweld condition
with the stationary dots condition is projected on the surface revealing motion-selective areas including V3A and hMT+/V5. (C) Lateral and
ventral view of inatedcortical hemispheres. Activated regions responding solely to objects presented in the left visual eld are coloured green;
regions responding solely to objects presented in the right visual eld are coloured red. Areas responding with equal strength to stimuli in either
hemield are shown in yellow (for graded values, see colour scale).
R.Goebel et al.
/
Vision Research
41 (2001) 14591474
1467
Fig. 4.
R.Goebel et al.
/
Vision Research
41 (2001) 14591474
1468
3
.
4
.Object experiment
In the control subjects, activated regions in early
visual areas responded solely to objects presented in the
contralateral visual eld (Fig. 3C). With increasing
distance from V1, visual areas responded also to stimuli
in the ipsilateral visual eld as indicated in Fig. 3(C) by
the change from red to yellow color (left hemisphere)
and green-to-yellow color (right hemisphere). Regions
responding equally well to stimuli in both visual elds
(yellow) were more extended in the right hemisphere.
The human motion complex, which is known to re-
spond not only to motion but also to ickering stimuli
(Tootell et al., 1995) also responded strongly to the
presentation of the images. The motion complex in
either hemisphere responded to stimuli in both visual
elds albeit with a stronger response to the respective
contralateral hemield.
The normal hemisphere in the patients responded to
stimuli in the normal hemield with a similar activation
pattern as that seen in the control subjects, and includes
early visual areas, the lateral occipital cortical region
and a region in the fusiform gyrus (Fig. 5(B) and Fig.
6(B), right hemisphere). Activity in the lesioned hemi-
sphere caused by stimuli in the normal eld (Fig.
5(B,C) and Fig. 6(B,C), green curves) was less extended
than in the controls, but in GY, the region in the
fusiform gyrus responded as well as it did to stimuli in
the normal hemield. Much to our surprise, ventral
areas in the lesioned hemishere responded quite
strongly to images presented in the impaired (right)
hemield (Fig. 5C and Fig. 6C, red curves),. In GY,
both a region in the fusiform gyrus and the lateral
occipital region were activated. These foci presumably
correspond to V4/V8 and area LO, respectively, as their
Talairach coordinates (Table 1) closely match those
reported in the literature (Malach et al., 1995; Had-
jikhani et al., 1998). Time courses show that the ipsile-
sional response strength in area V4/V8 is nearly the
same for stimuli in the blind and sighted eld, but the
response of area LO is weaker for stimuli in the blind
eld (Fig. 5C). In FS, the lateral occipital region in the
lesioned hemisphere responded roughly half as strong
to images in the impaired hemield compared to the
response after sighted eld stimulation (Fig. 6C), but
there was only little activation in V4/V8. In both sub-
jects, some activation was also seen in hMT+/V5. In
Table 1
Talairach coordinates (x,y,z) of regions hMT+/V5, LO, and V4/V8
in the left (LH) and right (RH) hemisphere of patients GY and FS.
FS LHGY RH FS RHGY LH
36, 67, 1hMT+/V5 47, 63, 0 48, 76, 0 47, 64, 3
45, 65, 20LO 38, 67, 839, 73, 15 43, 65, 16
22, 62, 21V4/V8 35, 62, 1825, 65, 12 28, 53, 17
Table 2
Two measures characterizing the response strenght in hMT+/V5 in
the left, lesioned hemisphere after blind eld stimulation and in the
right hemisphere after stimulation in the sighted hemield: maximum
observed correlation value (r
max
) and cluster size (number of con-
nected, activated pixels) at a specied correlation value of r=0.4
(CS
r=0.4
)
Black-white spiral Red-blue spiral
r
max
CS
r=0.4
r
max
CS
r=0.4
26 0.60 40RH 0.55
610.6193LH 0.62
GY, bilateral hMT+/V5 activity was observed after
stimulation in either hemield (Fig. 5B,C), but in FS,
only stimulation of the normal hemield elicited bilat-
eral hMT+/V5; no signicant activity was observed
after stimulation in the blind visual eld (Fig. 6B,C).
Throughout, the normal hemisphere showed very little
activation in response to stimuli in the impaired eld,
indicating that the coactivation of corresponding areas
commonly seen in higher extrastriate cortical areas is
compromised. In contrast to control subjects, no activ-
ity was observed anterior to V4/V8 in the left, lesioned
hemisphere in either patient.
In B10% of the presentations of the objects in the
cortically blind hemield, GY pressed the button for a
short duration following the onset of the image se-
quence, indicating that he was aware of a stimulus
change. An event-related analysis of awareversus
unawareepochs produced almost identical results. As
in the motion experiment, FS never indicated awareness
when stimuli were presented in his blind visual eld.
Fig. 4. Multiple regression results of the black-and-white spiral experiment of patient GY (A) and patient FS (B), shown on inated representation
of the cortical sheet. Activated regions responding solely to the spiral presented in the left visual eld are coloured green; regions responding solely
to the spiral presented in the right visual eld are coloured red. Areas responding with equal strength to the spiral in either hemield are shown
in yellow (for graded values, see colour scale). (A) Activated regions in the right hemisphere are dominated by green colors whereas regions in
the lesioned left hemisphere are dominated by yellow colors. The extent of the functionally dened motion complex is greater in the ipsilesional
hemisphere than in the intact hemisphere. Time courses of left and right hemispheric motion complex (hMT+/V5) for left hemield stimulation
(green curves) and right hemield stimulation (red curves). (B) Activated regions in the right hemisphere are dominated by green colors whereas
regions in the lesioned left hemisphere are dominated by yellow colors.
R.Goebel et al.
/
Vision Research
41 (2001) 14591474
1469
Fig. 5. Multiple regression results of object experiment of patient GY. Dark red regions in (A) and (B) correspond to the boundary of the cortex
and the lesion; LH=left hemisphere, RH=right hemisphere. (A) Folded (small insets) and attened representation of the reconstructed cortical
sheet with superimposed results of eccentricity mapping; foveal to peripheral visual eld representations are coded in colour from red to yellow
to blue to green. (B) Results of the object experiment superimposed on a subpart of the attened cortex representation of each hemisphere as
indicated with white rectangles in (A). Activated regions responding solely to objects presented in the left visual eld are coloured green, regions
responding solely to objects presented in the right visual eld are coloured red. Areas responding with equal strength to stimuli in either hemield
are shown in yellow (for graded values, see colour scale); FG=fusiform gyrus. (C) Time courses of right hemispheric area V1/V2 and left and
right hemispheric areas hMT+/V5, presumed LO and presumed V4/V8 for left hemield stimulation (green curves) and right hemield
stimulation (red curves); location of plotted areas are indicated by numbers 14in(B).
R.Goebel et al.
/
Vision Research
41 (2001) 14591474
1470
Fig. 6. Multiple regression results of object experiment of patient FS. Dark red regions in (A) and (B) correspond to the boundary of the cortex
and the lesion; LH=left hemisphere, RH=right hemisphere. (A) Folded (small insets) and attened representation of the reconstructed cortical
sheet with superimposed results of the main experiment. (B) Results of the object experiment superimposed on a subpart of the attened cortex
representation of each hemisphere as indicated with white rectangles in (A). Activated regions responding solely to objects presented in the left
visual eld are coloured green, regions responding solely to objects presented in the right visual eld are coloured red (for graded values, see colour
scale); FG=fusiform gyrus. (C) Time course plots of areas of interest as in Fig. 5; location of plotted areas are indicated by numbers 14in(B).
R.Goebel et al.
/
Vision Research
41 (2001) 14591474
1471
4. Discussion
4
.
1
.Summary
We studied activation patterns in dorsal and ventral
cortical areas in response to a rotating spiral and to
images of natural objects presented in the intact and
cortically blind elds of two hemianopic patients. We
found surprisingly strong responses to blind-eld stimu-
lation in ipsilesional extrastriate areas including the
motion complex, but also in ventral areas, which prob-
ably correspond to area LO and V4/V8. There was no
detectable activity in the early visual cortical areas of
the lesioned hemisphere and with the exception of GYs
occasional button press at stimulus onset, the patients
indicated no awareness of the stimuli. In addition, the
data revealed an unexpected asymmetry: the areas of
the intact hemisphere responded well to stimulation of
the contralateral, sighted hemield but very little to
stimulation of the impaired hemield. In contrast, the
corresponding areas in the lesioned hemisphere re-
sponded to stimulation of either hemield.
4
.
2
.Extrastriate cortical acti6ation without V
1
A strong and sustained activity in extrastriate areas is
not what one would have expected on the basis of
previous monkey experiments which showed that re-
sponses of neurons in MT, while still present, are
markedly reduced (Rodman et al., 1989) and that neu-
rons in ventral stream areas become unresponsive to
visual stimuli presented in the affected hemield
(Girard et al., 1991; Bullier et al., 1993). It appears
unlikely that the observed activation is mediated by
only partially damaged V1 tissue (Fendrich, Wessinger,
& Gazzaniga, 1992; Wessinger, Fendrich, & Gazzaniga,
1997) since several other functional imaging studies
have also shown an absence of detectable activation in
the lesioned cortex in GY (see Barbur et al., 1993;
Sahraie et al., 1997; Zeki & ffytche, 1998; Baseler,
Morland, & Wandell, 1999; Kleiser, Wittsack, Niedeg-
gen, Goebel, & Stoerig, 2001) and the denervated re-
gion in FS (Stoerig et al., 1998; Kleiser et al., 2001)
after blind eld stimulation. Nevertheless, of the two
ndings, the dorsal activation of hMT+is less surpris-
ing. It has been reported before (Barbur et al., 1993;
Sahraie et al., 1997; Stoerig et al., 1997; Zeki & ffytche,
1998) and it agrees well with psychophysical data show-
ing that dorsalfunctions, such as localization (Po¨ppel,
Frost & Held, 1973; Weiskrantz et al., 1974) and mo-
tion processing (Barbur et al., 1980; Po¨ppel, 1985;
Blythe et al., 1986; Perenin, 1991; Benson et al., 1998)
can be demonstrated in cortically blind elds. In view
of the physiological results from monkey studies (Rod-
man, Gross & Albright, 1990), it is likely that the
information reaches hMT+via the colliculo-pulvino-
extrastriate cortical pathway. What is unexpected is the
extent of ipsilesional hMT+activation in GY, which,
different from FS, is actually as pronounced as that of
the motion complex in the intact hemisphere when it is
stimulated from the normal hemield. The strength of
activation runs counter the reduction seen in the physi-
ological data from monkeys with ablated or cooled V1
and may indicate plastic changes of the system compen-
sating for the loss of V1 which is normally the major
(direct and indirect) source of hMT+input. When
compared to FS, in whom the activation was weaker on
the lesioned side, one may suggest that GYs much
earlier lesion allowed for better reorganization. That it
is luminance rather than chromatic contrast that reveals
the extent of GYs hMT+activation could be seen as
a consequence of the degenerative effects of a V1 lesion
on the colour-opponent retino-geniculo-striate cortical
system that originates in the Pb-ganglion cells of the
retina (Cowey et al., 1989): The magnocellular lumi-
nance- and motion-processing system feeding into the
dorsal pathway is much less compromised.
In view of the data from monkeys, whose V1 had
been deactivated (Bullier et al., 1993), the activation of
ventral cortical areas from stimulation of the impaired
eld is most unexpected. In the absence of V1, the
information could reach the ventral areas either via
direct subcortical projections, as those from the degen-
erated lateral geniculate nucleus (Yukie & Iwai, 1981;
Cowey & Stoerig, 1989), or it could arise via lateral
dorso-ventral connections. Our data do not allow refu-
tation of either hypothesis, although the latter is ren-
dered somewhat less likely by the weakness of the
object-induced activity seen in hMT+. Preliminarily,
we have attributed this hMT+activation to the image
presentation frequency of 1 Hz which represents a slow
icker. Ventralfunctions, such as chromatic (Po¨ppel,
1986; Stoerig, 1987; Stoerig & Cowey, 1992; Brent et
al., 1994) and shape discrimination (Weiskrantz et al.,
1974; Weiskrantz, 1987; Marcel, 1998; Stoerig, in
preparation, 1998) have also been demonstrated in cor-
tically blind visual elds, although they are generally
more difcult to elicit than the dorsal functions, and
require more testing to reveal weaker, albeit sometimes
highly signicant, results. Our results indicate that it
may be the human colour complex V4/V8 that mediates
the residual processing of chromatic information, as
chromatic information activates this region (Lueck et
al., 1989; Hadjikhani et al., 1998) and its destruction
entails achromatopsia (Meadows, 1974). This assump-
tion ts well with ndings showing good chromatic
processing in GY (Brent et al., 1994 Barbur et al., 1999
Stoerig, in preparation), but poorer processing in FS
(Stoerig, 1987; in preparation) whose V4/V8 was not
detectably activated during blind eld stimulation. The
ipsilesional lateral occipital area LO that responded in
both subjects could be involved in residual form pro-
R.Goebel et al.
/
Vision Research
41 (2001) 14591474
1472
cessing, although the psychophysical predictions sug-
gested by the response properties of LO better
response to objects than textures, comparable responses
to faces, objects, and abstract sculptures (Malach et al.,
1995), and reduction of activity in response to scram-
bled natural objects (Grill-Spector et al., 1998) have
not yet been tested.
4
.
3
.Asymmetry of extrastriate cortical acti6ation
While we found that each hemisphere was activated
by its contralateral visual hemield, activation by the
ipsilateral hemield was asymmetrical for the two hemi-
spheres. Though this pattern was less obvious in dorsal
areas, the ventral foci of the intact hemisphere re-
sponded much less to stimulation of the impaired
hemield than did those of the lesioned hemisphere to
stimulation of the normal hemield. In addition, the
activation in the lesioned ventral areas is quite focal: it
does not extend forward or backward as it does in the
intact hemisphere. It is possible that these extrastriate
areas, while able to respond to stimuli in the impaired
eld, are incapacitated with respect to conveying activ-
ity forward, backward, or to the other hemisphere.
Such an inability to effectively transmit signals to other
areas might be caused by inactivation of recurrent
loops between higher and early visual areas. Recurrent
loops can organize neuronal activity into stable reso-
nant states, which have been proposed as the neural
correlate of conscious vision (e.g. Tononi & Edelman,
1998; Grossberg, 1999; Engel & Singer, 2001). As we
found sustained and pronounced cortical activity in
both subjects without awareness of the presented stim-
uli, our data are in agreement with such resonance
theories, but not with simple activation threshold theo-
ries(Palmer, 1999), which assume that any cortical
neural activity sufciently strong or lasting will produce
conscious experience of the content it represents.
Acknowledgements
It is a pleasure to thank FS and GY for participation
in this study. We thank Claudia I. Goebel for her help
in performing this study and Niko Kriegeskorte for
helpful comments on the manuscript. This work was
supported by the Max Planck Society and a Human-
Capital and Mobility network grant from the European
Community.
References
Barbur, J. L., Ruddock, K. H., & Watereld, V. A. (1980). Human
visual responses in the absence of the geniculo-calcarine projec-
tion. Brain,
103
, 905928.
Barbur, J. L., Watson, J. D., Frackowiak, R. S., & Zeki, S. (1993).
Conscious visual perception without V1. Brain,
116
, 12931302.
Baseler, H. A., Morland, A. B., & Wandell, B. A. (1999). Topo-
graphic organization of human visual areas in the absence of
input from primary cortex. Journal of Neuroscience,
19
, 2619
2627.
Benson, P. J., Guo, K., & Blakemore, C. (1998). Direction discrimi-
nation of moving gratings and plaids and coherence in dot
displays without primary visual cortex (V1). European Journal of
Neuroscience,
10
, 37673772.
Blythe, I. M., Bromley, J. M., Kennard, C., & Ruddock, K. H.
(1986). Visual discrimination of target displacement remains after
damage to the striate cortex in humans. Nature,
320
,619621.
Brent, P. J., Kennard, C., & Ruddock, K. H. (1994). Residual colour
vision in a human hemianope: spectral responses and colour
discrimination. Proceedings of the Royal Society London.Series B
:
Biology Science,
256
, 219225.
Bullier, J., Girard, P., & Salin, P.-A. (1993). The role of area 17 in the
transfer of information to extrastriate visual cortex. In A. Peters,
& K. S. Rockland, Cerebral cortex, vol. 10 (pp. 301330). New
York: Plenum Press.
Cowey, A. (1994). Cortical visual areas and the neurobiology of
higher visual processes. In M. J. Farah, & G. Ratcliff, The
Neuropsychology of high-le6el 6ision (pp. 3 31). Hillsdale, NJ:
Lawrence Erlbaum.
Cowey, A., & Stoerig, P. (1989). Projection patterns of surviving
neurons in the dorsal lateral genicualte nucleus following discrete
lesions of striate cortex: implications for residual vision. Experi-
mental Brain Research,
75
, 631638.
Cowey, A., Stoerig, P., & Perry, V. H. (1989). Transneuronal retro-
grade degeneration of retinal ganglion cells after damage to striate
cortex in macaque monkeys: selective loss of P beta cells. Journal
of Neuroscience,
29
,6580.
Engel, A. K., & Singer, W. (2001). Temporal binding and the neural
correlates of sensory awareness. Trends in Cogniti6e Sciences,
5
,
1625.
Fendrich, R., Wessinger, C. M., & Gazzaniga, M. S. (1992). Residual
vision in a scotoma. Implications for blindsight. Science,
258
,
14891491.
Finlay, A. L., Jones, S. R., Morland, A. B., Ogilvie, J. A., &
Ruddock, K. H. (1997). Movement in the normal visual hemield
induces a percept in the blindhemield of a human hemianope.
Proceedings of the Royal Society London.Series B,
264
, 267275.
Girard, P., Salin, P. A., & Bullier, J. (1991). Visual activity in
macaque area V4 depends on area 17 input. Neuroreport,
2
,
8184.
Goebel, R. (2000). A fast automated method for attening cortical
surfaces. NeuroImage,
11
, S680 Abstract.
Goebel, R., & Singer, W. (1999). Cortical surface-based statistical
analysis of functional magnetic resonance imaging data. NeuroIm-
age,
9
, S64 Abstract.
Goebel, R., Khorram-Sefat, D., Muckli, L., Hacker, H., & Singer, W.
(1998a). The constructive nature of vision: direct evidence from
functional magnetic resonance imaging studies of apparent mo-
tion and motion imagery. European Journal of Neuroscience,
10
,
15631573.
Goebel, R., Stoerig, P., Muckli, L., Zanella, F. E., & Singer, W.
(1998b). Ipsilesional visual activation in ventral extrastriate cortex
in patients with blindsight. Social Neuroscience Abstracts,
24
,
1508.
Goebel, R., Muckli, L, Zanella, F.E., Singer, W. & Stoerig, P.
(submitted). Sustained activation without visual awareness in
ipsilesional ventral cortical areas of hemianopic patients.
Grill-Spector, K., Kushnir, T., Hendler, T., Edelman, S., Itzchak, Y.,
& Malach, R. (1998). A sequence of object-processing stages
revealed by fMRI in the human occipital lobe. Human Brain
Mapping,
6
, 316328.
R.Goebel et al.
/
Vision Research
41 (2001) 14591474
1473
Grossberg, S. (1999). The link between brain learning, attention, and
consciousness. Conscious Cognition,
8
,144.
Guo, K., Benson, P. J., & Blakemore, C. (1998). Residual motion
discrimination using colour information without primary visual
cortex. NeuroReport,
9
, 21032106.
Hadjikhani, N., Liu, A. K., Dale, A. M., Cavanagh, P., & Tootell, R.
B. H. (1998). Retinotopy and color sensitivity in human visual
cortical area V8. Nature Neuroscience,
1
, 235241.
Haxby, J. V., Grady, C. L., Horwitz, B., Ungerleider, L. B., Mishkin,
M., Carson, R. E., Herscovitch, P., Schapiro, M. B., & Rapoport,
S. I. (1991). Dissociation of spatial and object visual processing
pathways in human extrastriate cortex. Proceedings of the Natural
Academy of Science USA,
88
, 16211625.
Heywood, C. A., & Cowey, A. (1987). On the role of cortical area V4
in the discrimination of hue and pattern in macaque monkeys.
Journal of Neuroscience,
7
, 26012617.
Ishai, A., Ungerleider, L. G., Martin, A., Schouten, J. L., & Haxby,
J. V. (1999). Distributed representation of objects in the human
ventral visual pathway. Proceedings of the Natural Academy of
Science USA,
96
(16), 93799384.
Kanwisher, N., McDermott, J., & Chun, M. M. (1997). The fusiform
face area: a module in human extrastriate cortex specialized for
face perception. Journal of Neuroscience,
17
, 43024311.
Kentridge, R. W., Heywood, C. A., & Weiskrantz, L. (1999). Effects
of temporal cueing on residual visual discrimination in blindsight.
Neuropsychologia,
37
, 479483.
Kleiser, R., Wittsack, J., Niedeggen, M., Goebel, R. & Stoerig, P.
(2001). Is V1 necessary for conscious vision in areas of relative
cortical blindness? NeuroImage, in press.
Kriegeskorte, N. & Goebel, R. (submitted). An efcient algorithm for
topologically correct segmentation of the cortical sheet in anatom-
ical MR volumes.
Linden, D. E. J., Kallenbach, U., Heinecke, A., Singer, W., &
Goebel, R. (1999). The myth of upright vision. A psychophysical
and functional imaging study of adaptation to inverting specta-
cles. Perception,
28
, 469481.
Lueck, C. J., Zeki, S., Friston, K. J., Deiber, M. P., Cope, P.,
Cunningham, V. J., Lammertsma, A. A., Kennard, C., & Frack-
owiak, R. S. (1989). The colour centre in the cerebral cortex of
man. Nature,
340
, 386389.
Malach, R., Reppas, J. B., Benson, R. R., Kwong, K. K., Jiang, H.,
Kennedy, W. A., Ledden, P. J., Brady, T. J., Rosen, B. R., &
Tootell, R. B. (1995). Object-related activity revealed by func-
tional magnetic resonance imaging in human occipital cortex.
Proceedings of the Natural Academy of Science USA,
92
, 8135
8139.
Marcel, A. J. (1998). Blindsight and shape perception: decit of visual
consciousness or of visual function? Brain,
121
, 15651588.
McKeefry, D. J., & Zeki, S. (1997). The position and topography of
the human colour centre as revealed by functional magnetic
resonance imaging. Brain,
120
, 22292242.
Meadows, J. C. (1974). Disturbed perception of colours associated
with localized cerebral lesions. Brain,
97
, 615632.
Milner, A. D., & Goodale, M. A. (1995). The 6isual brain in action.
Oxford: Oxford University Press.
Mishkin, M., Ungerleider, L. G., & Macko, K. A. (1983). Object
vision and spatial vision: two cortical pathways. Trends in Neuro-
science,
6
, 414417.
Morland, A. B., Ogilvie, J. A., Ruddock, K. H., & Wright, J. R.
(1996). Orientation discrimination is impaired in the absence of
the striate cortical contribution to human vision. Proceedings of
the Royal Society London,Series B,
263
, 633640.
Palmer, S. E. (1999). Vision science. Cambridge, MA: MIT Press.
Perenin, M. T. (1991). Discrimination of motion direction in perimet-
rically blind elds. Neuroreport,
2
, 397400.
Po¨ppel, E. (1985). Bridging a neuronal gap. Perceptual completion
across a cortical scotoma is dependent on stimulus motion. Natur-
wissenschaften,
72
, 599600.
Po¨ppel, E. (1986). Long-range colour-generating interactions across
the retina. Nature,
320
, 523525.
Po¨ppel, E., Frost, D., & Held, R. (1973). Residual visual function
after brain wounds involving the central visual pathways in man.
Nature,
243
, 295296.
Rodman, H. R., Gross, C. G., & Albright, T. D. (1989). Afferent
basis of visual response properties in area MT of the macaque. I.
Effects of striate cortex removal. Journal of Neuroscience,
9
,
20332050.
Rodman, H. R., Gross, C. G., & Albright, T. D. (1990). Afferent
basis of visual response properties in area MT of the macaque: 2.
Effects of superior colliculus removal. Journal of Neuroscience,
10
,
11541164.
Sahraie, A., Weiskrantz, L., Barbur, J. L., Simmons, A., Williams, S.
C. R., & Brammer, M. J. (1997). Pattern of neuronal activity
associated with conscious and unconscious processing of visual
signals. Proceedings of the Natural Academy of Science USA,
94
,
94069411.
Sereno, M. I., Dale, A. M., Reppas, J. B., Kwong, K. K., Belliveau,
J. W., Brady, T. J., Rosen, B. R., & Tootell, R. B. H. (1995).
Borders of multiple visual areas in humans revealed by functional
magnetic resonance imaging. Science,
268
, 889893.
Snowden, R. J., Treue, S., & Andersen, R. A. (1992). The response of
neurons in areas V1 and MT of the alert rhesus monkey to
moving random dot patterns. Experimental Brain Research,
88
,
389400.
Stoerig, P. (1987). Chromaticity and achromaticity. Evidence for a
functional differentiation in visual eld defects. Brain,
110
, 869
886.
Stoerig, P. (1993). Spatial summation in blindsight. Vision Neuro-
science,
10
, 11411149.
Stoerig, P. (in prep.) Apples and oranges: Discrimination of images of
natural objects in patients with retrogeniculate lesions.
Stoerig, P., & Cowey, A. (1992). Wavelength discrimination in blind-
sight. Brain,
115
, 425444.
Stoerig, P., & Cowey, A. (1997). Blindsight in man and monkey.
Brain,
120
, 535559.
Stoerig, P., Goebel, R., Muckli, L., Hacker, H., & Singer, W. (1997).
On the functional neuroanatomy of blindsight. Social Neuro-
science Abstracts,
23
, 845.
Stoerig, P., Kleinschmidt, A., & Frahm, J. (1998). No visual re-
sponses in denervated V1: high-resolution functional magnetic
resonance imaging of a blindsight patient. Neuroreport,
9
,2125.
Talairach, J., & Tournaux, P. (1988). Co-planar stereotaxic atlas of
the human brain. New York, NY: Thieme.
Tononi, G., & Edelman, G. M. (1998). Consciousness and complex-
ity. Science,
282
, 18461851.
Tootell, R. B. H., Reppas, J. B., Kwong, K. K., Malach, R., Born, R.
T., Brady, T. J., Rosen, B. R., & Belliveau, J. W. (1995).
Functional analysis of human MT and related visual cortical
areas using magnetic resonance imaging. Journal of Neuroscience,
15
, 32153230.
Van Buren, J. M. (1963). Trans-synaptic retrograde degeneration in
the visual system of primates. Journal of Neurology,Neurosurgery
and Psychiatry,
34
, 140147.
Van Essen, D. C., Maunsell, J. H. R., & Bixby, J. L. (1981). The
middle temporal visual area in the macaque: myeloarchitecture,
connections, functional properties and topographic organization.
Journal of Cogniti6e Neurology,
201
,8198.
Watson, J. D. G., Myers, R., Frackowiak, R. S. J., Hajnal, J. V.,
Woods, R. P., Mazziotta, J. C., Shipp, S., & Zeki, S. (1993). Area
V5 of the human brain: evidence from a combined study using
positron emmission tomography and magnetic resonance imaging.
Cerebral Cortex,
3
,7994.
Wessinger, C. M., Fendrich, R., & Gazzaniga, M. S. (1997). Islands
of residual vision in hemianopic patients. Journal of Comparati6e
Neuroscience,
9
, 203221.
R.Goebel et al.
/
Vision Research
41 (2001) 14591474
1474
Weiskrantz, L. (1987). Residual vision in a scotoma. A follow-up study
of formdiscrimination. Brain,
110
,7792.
Weiskrantz, L., Warrington, E. K., Sanders, M. D., & Marshall, J.
(1974). Visual capacity in the hemianopic eld following a restricted
occipital ablation. Brain,
97
, 709728.
Weiskrantz, L., Barbur, J. L., & Sahraie, A. (1995). Parameters
affecting conscious versus unconscious visual discrimination in a
patient with damage to the visual cortex (V1). Proceedings of the
Natural Academy of Science USA,
92
, 61226126.
Yukie, M., & Iwai, E. (1981). Direct projection from the dorsal lateral
geniculate nucleus to the prestriate cortex in macaque monkeys.
Journal of Comparati6e Neurology,
291
,8197.
Zeki, S. (1974). In R. G. E. G. Bellairs, Essays on the ner6ous system
A festschrift for Professor J.Z.Young (pp. 327343). Oxford:
Clarendon Press.
Zeki, S. (1991). Cerebral akinetopsia (visual motion blindness): a
review. Brain,
114
, 811824.
Zeki, S., & ffytche, D. H. (1998). The Riddoch syndrome: insights into
the neurobiology of conscious vision. Brain,
121
,2545.
Zeki, S., McKeefry, D. J., Bartels, A., & Frackowiak, R. S. (1998). Has
a new color area been discovered? Nature Neuroscience,
1
, 335336.
Zilles, K., & Clarke, S. (1997). Architecture, connectivity, and trans-
mitter receptors of human extrastriate visual cortex: comparison
with nonhuman primates. In K. S. Rockland, J. H. Kaas, & A.
Peters, Cerebral cortex, vol. 12 (pp. 673742). New York: Plenum
Press.
.
... After 2-month VPL training, the positive association of the MTD score in the affected hemifield remained strongly significant with the interhemispheric visual RSFC; The positive associations of the MTD score in the affected hemifield became significant with the ipsilesional visual RSFC and became non-significant with the contralesional visual RSFC (Fig. 2). In tandem with previous literature [26][27][28] , a lower ipsilesional RSFC compared to contralesional visual RSFC at baseline (ipsilesional, 0.549 ± 0.306; contralesional, 0.718 ± 0.216; p = 0.014) was found in our study. Stroke-related dysfunction is believed to be related to weaker ipsilesional function 23,24 , accompanied by compensatory functional increase in the contralesional hemisphere. ...
... Albeit without statistical significance; this discrepancy could be explained by two speculative hypotheses: (1) VPL may result in sharpened tuning characteristics of the visual Similarly, perceptual learning of orientation discrimination tasks lowered visual functional activation and improved visual performance in previous literature [35][36][37][38] , suggesting that more specific changes may occur in fewer visual cortical neurons [35][36][37][38] . Furthermore, the VPL-responsive subgroup (high visual RSFC subgroup) in our study have efficiently restored ipsilesional dysfunction, not necessarily via increased ipsilesional visual RSFC (Fig. 5D) 42 , but after disentangling the neurobehavioral disassociations underlying VFD (Fig. 2) [26][27][28] . Fig. 5F) as compensation than ipsilesional functional usage, further increasing inter-hemispheric functional imbalance 43 . ...
Article
Full-text available
A reciprocal relationship between perceptual learning and functional brain changes towards perceptual learning effectiveness has been demonstrated previously; however, the underlying neural correlates remain unclear. Further, visual perceptual learning (VPL) is implicated in visual field defect (VFD) recovery following chronic stroke. We investigated resting-state functional connectivity (RSFC) in the visual cortices associated with mean total deviation (MTD) scores for VPL-induced VFD recovery in chronic stroke. Patients with VFD due to chronic ischemic stroke in the visual cortex received 24 VPL training sessions over 2 months, which is a dual discrimination task of orientation and letters. At baseline and two months later, the RSFC in the ipsilesional, interhemispheric, and contralesional visual cortices and MTD scores in the affected hemi-field were assessed. Interhemispheric visual RSFC at baseline showed the strongest correlation with MTD scores post-2-month VPL training. Notably, only the subgroup with high baseline interhemispheric visual RSFC showed significant VFD improvement following the VPL training. The interactions between the interhemispheric visual RSFC at baseline and VPL led to improvement in MTD scores and largely influenced the degree of VFD recovery. The interhemispheric visual RSFC at baseline could be a promising brain biomarker for the effectiveness of VPL-induced VFD recovery.
... Results from a lesion networking mapping (LNM) study, a powerful tool used to make causal inferences from lesions causally linked to symptoms (Fox, 2018), identified the eVC to be implicated in VH . Pathologically elevated eVC activity has also been demonstrated in psychosis (Goebel et al., 2001). Lastly, a study examining the neural basis of motion perception in schizophrenia found that reduced V5/MT activation was associated with lower delta (2 Hz) evoked amplitude during motion related tasks and poorer cognitive performance (Martínez et al., 2018). ...
Article
Background: Transcranial electrical stimulation (tES) may improve psychosis symptoms, but few investigations have targeted brain regions causally linked to psychosis symptoms. We implemented a novel montage targeting the extrastriate visual cortex (eVC) previously identified by lesion network mapping in the manifestation of visual hallucinations. Objective: To determine if lesion network guided High Definition-tES (HD-tES) to the eVC is safe and efficacious in reducing symptoms related to psychosis. Methods: We conducted a single-blind crossover pilot study (NCT04870710) in patients with psychosis spectrum disorders. Participants first received HD-tDCS (direct current), followed by 4 weeks of wash out, then 2 Hz HD-tACS (alternating current). Participants received 5 days of daily (2×20 min) stimulation bilaterally to the eVC. Primary outcomes included the Positive and Negative Syndrome Scale (PANSS), biological motion task, and Event Related Potentials (ERP) from a steady state visual evoked potential (SSVEP) paradigm. Secondary outcomes included the Montgomery-Asperg Depression Rating Scale, Global Assessment of Functioning (GAF), velocity discrimination and visual working memory task, and emotional ERP. Results: HD-tDCS improved PANSS general psychopathology in the short-term (d=0.47; pfdr=0.03), with long-term improvements in general psychopathology (d=0.62; pfdr=0.05) and GAF (d=-0.56; pfdr=0.04) with HD-tACS. HD-tDCS reduced SSVEP P1 (d=0.25; pfdr=0.005), which correlated with general psychopathology (β = 0.274, t = 3.59, p = 0.04). No significant differences in safety or tolerability measures were identified. Conclusion: Lesion network guided HD-tES to the eVC is a safe, efficacious, and promising approach for reducing general psychopathology via changes in neuroplasticity. These results highlight the need for larger clinical trials implementing novel targeting methodologies for the treatments of psychosis.
... Results from a lesion networking mapping (LNM) study, a powerful tool used to make causal inferences from lesions causally linked to symptoms 22 , identified the eVC to be implicated in VH 23 . Pathologically elevated eVC activity has also been demonstrated in psychotic 24 . Lastly, a study examining the neural basis of motion perception in schizophrenia found that reduced V5/MT activation was associated with lower delta (2Hz) evoked amplitude during motion related tasks and poorer cognitive performance 25 . ...
Preprint
Full-text available
Importance: Transcranial electrical stimulation (tES) may improve psychosis symptoms, but few investigations have targeted brain regions causally linked to psychosis symptoms. We implemented a novel montage targeting the extrastriate visual cortex (eVC) previously identified by lesion network mapping in the manifestation of visual hallucinations. Objective: To determine if lesion network guided HD-tES to the eVC is safe and efficacious in reducing symptoms related to psychosis. Design, Setting, and Participants: Single-center, nonrandomized, single-blind trial using a crossover design conducted in two 4-week phases beginning November 2020, and ending January 2022. Participants were adults 18-55 years of age with a diagnosis of schizophrenia, schizoaffective or psychotic bipolar disorder as confirmed by the Structured Clinical Interview for DSM-V, without an antipsychotic medication change for at least 4 weeks. A total of 8 participants consented and 6 participants enrolled. Significance threshold set to <0.1 due to small sample size. Interventions: 6 Participants first received HD-tDCS (direct current), followed by 4 weeks of wash out, then 4 received 2Hz HD-tACS (alternating current). Participants received 5 consecutive days of daily (2 x 20min) stimulation applied bilaterally to the eVC. Main Outcomes and Measures: Primary outcomes included the Positive and Negative Syndrome Scale (PANSS) total, positive, negative, and general scores, biological motion task, and Event Related Potential (ERP) measures obtained from a steady state visual evoked potential (SSVEP) task across each 4-week phase. Secondary outcomes included the Montgomery-Asperg Depression Rating Scale (MADRS), Global Assessment of Functioning (GAF), velocity discrimination task, visual working memory task, and emotional ERP across each 4-week phase. Results: HD-tDCS improved general psychopathology in the short-term (d=0.47; pfdr=0.03), with long-term improvements in general psychopathology (d=0.62; pfdr=0.05) and GAF (d=-0.56; pfdr=0.04) with HD-tACS. HD-tDCS reduced SSVEP P1 (d=0.25; pfdr=0.005), which correlated with general psychopathology (β=0.274, t=3.59, p=0.04). No significant differences in safety or tolerability measures were identified. Conclusions and Relevance: Lesion network guided HD-tES to the eVC is a safe, efficacious, and promising approach for reducing general psychopathology via changes in neuroplasticity. These results highlight the need for larger clinical trials implementing novel targeting methodologies for the treatments of psychosis. Trial Registration: ClinicalTrials.gov Identifier: NCT04870710
Chapter
The Cambridge Handbook of Consciousness is the first of its kind in the field, and its appearance marks a unique time in the history of intellectual inquiry on the topic. After decades during which consciousness was considered beyond the scope of legitimate scientific investigation, consciousness re-emerged as a popular focus of research towards the end of the last century, and it has remained so for nearly 20 years. There are now so many different lines of investigation on consciousness that the time has come when the field may finally benefit from a book that pulls them together and, by juxtaposing them, provides a comprehensive survey of this exciting field. An authoritative desk reference, which will also be suitable as an advanced textbook.
Chapter
The Cambridge Handbook of Consciousness is the first of its kind in the field, and its appearance marks a unique time in the history of intellectual inquiry on the topic. After decades during which consciousness was considered beyond the scope of legitimate scientific investigation, consciousness re-emerged as a popular focus of research towards the end of the last century, and it has remained so for nearly 20 years. There are now so many different lines of investigation on consciousness that the time has come when the field may finally benefit from a book that pulls them together and, by juxtaposing them, provides a comprehensive survey of this exciting field. An authoritative desk reference, which will also be suitable as an advanced textbook.
Chapter
The Cambridge Handbook of Consciousness is the first of its kind in the field, and its appearance marks a unique time in the history of intellectual inquiry on the topic. After decades during which consciousness was considered beyond the scope of legitimate scientific investigation, consciousness re-emerged as a popular focus of research towards the end of the last century, and it has remained so for nearly 20 years. There are now so many different lines of investigation on consciousness that the time has come when the field may finally benefit from a book that pulls them together and, by juxtaposing them, provides a comprehensive survey of this exciting field. An authoritative desk reference, which will also be suitable as an advanced textbook.
Chapter
The Cambridge Handbook of Consciousness is the first of its kind in the field, and its appearance marks a unique time in the history of intellectual inquiry on the topic. After decades during which consciousness was considered beyond the scope of legitimate scientific investigation, consciousness re-emerged as a popular focus of research towards the end of the last century, and it has remained so for nearly 20 years. There are now so many different lines of investigation on consciousness that the time has come when the field may finally benefit from a book that pulls them together and, by juxtaposing them, provides a comprehensive survey of this exciting field. An authoritative desk reference, which will also be suitable as an advanced textbook.
Chapter
The Cambridge Handbook of Consciousness is the first of its kind in the field, and its appearance marks a unique time in the history of intellectual inquiry on the topic. After decades during which consciousness was considered beyond the scope of legitimate scientific investigation, consciousness re-emerged as a popular focus of research towards the end of the last century, and it has remained so for nearly 20 years. There are now so many different lines of investigation on consciousness that the time has come when the field may finally benefit from a book that pulls them together and, by juxtaposing them, provides a comprehensive survey of this exciting field. An authoritative desk reference, which will also be suitable as an advanced textbook.
Chapter
The Cambridge Handbook of Consciousness is the first of its kind in the field, and its appearance marks a unique time in the history of intellectual inquiry on the topic. After decades during which consciousness was considered beyond the scope of legitimate scientific investigation, consciousness re-emerged as a popular focus of research towards the end of the last century, and it has remained so for nearly 20 years. There are now so many different lines of investigation on consciousness that the time has come when the field may finally benefit from a book that pulls them together and, by juxtaposing them, provides a comprehensive survey of this exciting field. An authoritative desk reference, which will also be suitable as an advanced textbook.
Chapter
The Cambridge Handbook of Consciousness is the first of its kind in the field, and its appearance marks a unique time in the history of intellectual inquiry on the topic. After decades during which consciousness was considered beyond the scope of legitimate scientific investigation, consciousness re-emerged as a popular focus of research towards the end of the last century, and it has remained so for nearly 20 years. There are now so many different lines of investigation on consciousness that the time has come when the field may finally benefit from a book that pulls them together and, by juxtaposing them, provides a comprehensive survey of this exciting field. An authoritative desk reference, which will also be suitable as an advanced textbook.
Article
Evidence is reviewed indicating that striate cortex in the monkey is the source of two multisynaptic corticocortical pathways. One courses ventrally, interconnecting the striate, prestriate, and inferior temporal areas, and enables the visual identification of objects. The other runs dorsally, interconnecting the striate, prestriate, and inferior parietal areas, and allows instead the visual location of objects. How the information carried in these two separate pathways is reintegrated has become an important question for future research.
Article
Clinical evidence is presented which links the anterior inferior part of the occipital lobe with colour perception in man. A review of the literature suggests that bilateral lesions at this site may cause persisting impairment of colour perception (achromatopsia) with preservation of primary visual function. This is discussed in relation to relevant experiments in animals. Achromatopsia is commonly associated with prosopagnosia (inability to recognize familiar faces) and impaired topographical memory, but published reports suggest that these three disorders may be dissociated from one another and they therefore appear to be functionally distinct.
Article
Functional magnetic resonance imaging was used in combined functional selectivity and retinotopic mapping tests to reveal object-related visual areas in the human occpital lobe. Subjects were tested with right, left, up, or down hemivisual field stimuli which were composed of images of natural objects (faces, animals, man-made objects) or highly scrambled (1,024 elements) versions of the same images. In a similar fashion, the horizontal and vertical meridians were mapped to define the borders of these areas. Concurrently, the same cortical sites were tested for their sensitivity to image-scrambling by varying the number of scrambled picture fragments (from 16–1,024) while controlling for the Fourier power spectrum of the pictures and their order of presentation. Our results reveal a stagewise decrease in retinotopy and an increase in sensitivity to image-scrambling. Three main distinct foci were found in the human visual object recognition pathway (Ungerleider and Haxby [1994]: Curr Opin Neurobiol 4:157–165): 1) Retinotopic primary areas V1–3 did not exhibit significant reduction in activation to scrambled images. 2) Areas V4v (Sereno et al., [1995]: Science 268:889–893) and V3A (DeYoe et al., [1996]: Proc Natl Acad Sci USA 93:2382–2386; Tootell et al., [1997]: J Neurosci 71:7060–7078) manifested both retinotopy and decreased activation to highly scrambled images. 3) The essentially nonretinotopic lateral occipital complex (LO) (Malach et al., [1995]: Proc Natl Acad Sci USA 92:8135–8139; Tootell et al., [1996]: Trends Neurosci 19:481–489) exhibited the highest sensitivity to image scrambling, and appears to be homologous to macaque the infero-temporal (IT) cortex (Tanaka [1996]: Curr Opin Neurobiol 523–529). Breaking the images into 64, 256, or 1,024 randomly scrambled blocks reduced activation in LO voxels. However, many LO voxels remained significantly activated by mildly scrambled images (16 blocks). These results suggest the existence of object-fragment representation in LO. Hum. Brain Mapping 6:316–328, 1998.
Chapter
The ideas concerning the role of area 17 in the transfer of visual information to the rest of the cerebral cortex have for a long time been influenced by the results of behavioral studies of primates following cortical lesions. Since the last century it has been known that lesions of area 17 lead to blindness in humans (for a review see Weiskrantz, 1986, and Rizzo, this volume). This critical role of area 17 in vision was used in the beginning of the 20th century by Inouye in Japan and Holmes in Great Britain to map the representation of the visual field in area 17 of humans by delimiting the extents of scotomata resulting from focal lesions in area 17 of wounded soldiers. In 1942, the results of an extensive study by Klüver of monkeys with cortical lesions seemed to leave little doubt that, for this species as well, area 17 is necessary for any kind of vision beyond a simple discrimination between light and dark. Rudimentary sensitivity to light had also been noted to persist in humans with lesions of area 17 by Holmes (1918) and Riddoch (1917). Interestingly, this last author noted a weak residual sensitivity to moving targets, but no perception of stationary objects.
Article
The cytoarchitectonic and myelogenetic maps of the mammalian visual cortex (Brodmann, 1903, 1905, 1906, 1908a, 1908b, 1909; Flechsig, 1920; Economo and Koskinas, 1925) represented for decades widely accepted organizational concepts. Lately, they have been losing some of their importance, mainly for two reasons. First, the schematic figures published by these authors are often used for purposes of cortical localization, often by surface landmarks, without recourse to histological identification of areas and without taking into account ambiguities of definitions and interindividual variations. Second, purely architectonic maps tend to be replaced by more functionally relevant parcellations (for review see Felleman and van Essen, 1991; Kaas and Krubitzer, 1991; Kaas, 1993; Zeki, 1993). We aim at reestablishing the importance of the architectonic approach to the human cortex by reviewing critically classic and modern architectonic studies and by relating them to hodological and activation studies.
Article
This study is a follow-up of a patient, D.B., who was reported (Weiskrantz et al ., 1974) to be able to discriminate between simple visual forms within the scotoma caused by a lesion in calcarine cortex. Among other capacities, he was able to discriminate between lines of different orientation in the frontal plane. Given the reported deficits for form discrimination but a high sensitivity for orientation discrimination in primates without striate cortex, the question arises whether D.B.'s apparent ‘form’ discrimination arises from an ability to discriminate the orientations of components of the figures. It is shown in the first experiment, by using the optic disc as a control, that his ability to detect stimuli in his scotoma cannot be due to stray light falling upon the intact field. Next, his ability to discriminate orientations is confirmed, and an orientation threshold determined. A range of form discriminations is presented varying in degree of orientation cues of their components. It is confirmed that he can discriminate those forms originally studied, in which such differences are large, but not when orientation cues are small or minimal. Finally, his ability to compare two forms both projected within his scotoma is examined. Even when the components of the two forms have large orientation differences and are highly discriminable when presented successively, D.B. appears to be unable to make a ‘same-different’ comparison when both are presented simultaneously. The evidence is interpreted against D.B.'s having a residual capacity for form discrimination.
Article
In the visual field defects of 10 patients who had suffered lesions in the postgeniculate part of the primary visual projection, red-green discrimination and achromatic target detection was tested. In addition, 8 of these patients were tested for detection of red and green targets. Targets were presented on a low photopic achromatic background, so that the red and green targets differed from the background both in intensity and in wavelength, whereas the achromatic target differed in intensity only. Six patients showed evidence of discriminating between red and green targets, 5 patients could also detect the colour targets, but none could detect the achromatic one that was presented at the same retinal position. These results imply that wavelength and intensity information are treated differentially, and suggest that these patients possess residual colour-opponent channels that subserve the defective part of the visual field.
Article
Following damage to primary visual cortex, some patients our initial findings. The data reveal a patchy distribution of redemonstrate a limited ability to respond to stimuli they do not sidual visual abilities in the absence of acknowledged awareness. acknowledge seeing. This residual vision, which has been referred to as "blindsight," has been attributed to secondary visual pathways. We previously reported an isolated island of blindsight in one patient and argued it was a likely consequence of cortical sparing in V1. We now report an extension of our initial findings. The data reveal a patchy distribution of residual visual abilities in the absence of acknowledged awareness. Variable patterns of cortical sparing appear to be the most parsimonious way to account for this outcome, suggesting that blindsight is generally mediated by the primary visual pathway.