ArticlePDF AvailableLiterature Review

Molecular approaches to brain asymmetry and handedness

Authors:

Abstract

In the human brain, distinct functions tend to be localized in the left or right hemispheres, with language ability usually localized predominantly in the left and spatial recognition in the right. Furthermore, humans are perhaps the only mammals who have preferential handedness, with more than 90% of the population more skillful at using the right hand, which is controlled by the left hemisphere. How is a distinct function consistently localized in one side of the human brain? Because of the convergence of molecular and neurological analysis, we are beginning to consider the puzzle of brain asymmetry and handedness at a molecular level.
The human brain is a complex structure that controls
sophisticated cognitive behaviour. Anatomically,
the cerebral cortex is divided into frontal, temporal,
parietal and occipital lobes, and these regions control
thinking, language, movement, sensation, vision and
other functions. The formation of these distinct func-
tional regions during cortical development is called
regionalization (or arealization)
1–4
. Two models for
the formation of cortical functional regions have been
proposed
5–7
. The protomap model suggests that intrinsic
signals from the ‘proliferative units’ in the ventricular
zone regulate functional regionalization, whereas the
protocortex model argues the importance of extrinsic
influences, such as the thalamocortical inputs
5–7
.
Accumulating evidence indicates that both models are
applicable to the regulation of cortical patterning and
the establishment of cortical regionalization
2–4
.
The cerebral cortex is also divided into left and
right hemispheres. The left hemisphere is normally
dominant for language and logical processing, whereas
the right hemisphere is specified for spatial recogni-
tion
8,9
. Additionally, the segregation of human brain
functions between the left and right hemispheres is
associated with asymmetries of anatomical structures,
such as the
Sylvian fissures and the planum tempo-
rale
10,11
. One of the striking features of motor control
in humans is that more than 90% of the population is
more skilful with the right hand, which is controlled
by the left hemisphere
12
. Similar to the left-hemi-
sphere dominance of handedness, language ability is
dominant in the left hemisphere in more than 95% of
the right-handed population but in only 70% of the
left-handed population
12
.
Is it coincidental that both language ability and
hand use are dominant in the left hemisphere in most
of humans? Is there genetic control of both brain asym-
metry and handedness? Using molecular and neuro-
logical approaches, we are beginning to tackle these
questions and discover the neurological circuitries
that regulate brain asymmetry. Here, we describe brain
asymmetries that have been measured using modern
imaging techniques and discuss the genetic correlation
between brain asymmetry and preferential hand use.
Furthermore, we propose evolutionary and molecular
mechanisms that might regulate brain asymmetry and
handedness.
Functional and anatomical brain asymmetries
The first detailed description of functional asymmetry
in the human brain was made in the 1860s by a French
doctor named Paul Broca. He found that there was a
lesion in the left hemisphere of the post-mortem brain
of a patient with a one-word vocabulary. Broca claimed
that language ability in the human brain is lateral-
ized and supplied perhaps the first strong evidence of
functional asymmetry in the brain
13
. This brain region,
which controls speech, is called
Broca’s area in honour
of his discovery. In 1874, a German neurologist named
Carl Wernicke discovered that damage to a region of
the left hemisphere could cause a type of aphasia that
resulted in an impairment of language comprehension
14
.
This area is called
Wernicke’s area.
Brain functional asymmetry is not limited to lan-
guage ability. Whereas the right cerebral cortex regu-
lates movement of the left side of the body (and the left
cerebral cortex regulates movement of the right side),
*Department of Cell and
Developmental Biology,
Cornell University Weill
Medical College, Box 60,
W820A, 1300 York Avenue,
New York 10021, USA.
Department of Neurology,
Howard Hughes Medical
Institute, Beth Israel
Deaconess Medical Center,
New Research Building Room
0266, 77 Avenue Louis
Pasteur, Boston,
Massachusetts 02115, USA.
Correspondence to T.S. or
C.A.W. e-mails:
tas2009@med.cornell.edu;
cwalsh@bidmc.harvard.edu
doi:10.1038/nrn1930
Protomap model
Proposed by Pasko Rakic. He
suggested that regionalization
is mainly controlled by
molecular determinants that
are intrinsic to the proliferative
zone of the neocortex. The
‘proliferative units’ in the
ventricular zone form a
protomap of prospective
cortical regions. Postmitotic
neurons migrating from the
ventricular zone maintain the
regional properties of the
proliferative units.
Molecular approaches to brain
asymmetry and handedness
Tao Sun* and Christopher A. Walsh
Abstract | In the human brain, distinct functions tend to be localized in the left or right
hemispheres, with language ability usually localized predominantly in the left and spatial
recognition in the right. Furthermore, humans are perhaps the only mammals who have
preferential handedness, with more than 90% of the population more skilful at using the right
hand, which is controlled by the left hemisphere. How is a distinct function consistently
localized in one side of the human brain? Because of the convergence of molecular and
neurological analysis, we are beginning to consider the puzzle of brain asymmetry and
handedness at a molecular level.
REVIEWS
NATURE REVIEWS
|
NEUROSCIENCE VOLUME 7
|
AUGUST 2006
|
655
© 2006 Nature Publishing Group
© 2006 Nature Publishing Group
© 2006 Nature Publishing Group
ab c
AA
R
L
R
L
R
L
DD
VV
PP
Protocortex model
Proposed by Dennis O’Leary.
He suggested that
regionalization is controlled in
large part by extrinsic
influences, such as
thalamocortical inputs.
Sylvian fissures
The deepest and most
prominent of the cortical
fissures (clefts). They separate
the frontal lobes and temporal
lobes in both hemispheres.
Broca’s area
The left inferior frontal gyrus of
the frontal lobe of the human
cortex. This area is responsible
for speech and for
understanding language.
Injuries to this area can cause
Broca’s aphasia, which is
characterized by non-fluent
speech, few words, short
sentences and many pauses.
Patients normally lose the
ability to understand or
produce grammatically
complex sentences.
Wernicke’s area
The left posterior section of the
superior temporal gyrus, where
the temporal lobe and parietal
lobe meet. It is involved in the
comprehension of written or
spoken language. People with
damage in this area speak
fluently, but often using words
or jumbled syllables that make
no sense; this is known as
Wernicke’s aphasia.
more than 90% of the human population is naturally
more skilled with the right hand than with the left
12
.
Cognitive studies on patients with unilateral lesions and
on patients with split-brain surgery have revealed many
other differences between the left and right cerebral
cortex
15
. For example, the left hemisphere is dominant
for mathematical and logical reasoning, whereas the
right hemisphere excels at shape recognition, spatial
attention, emotion processing and musical and artistic
functions
15–17
.
Using modern imaging techniques, particularly
MRI, scientists can map the asymmetries of anatomical
structures in the human brain. Among the most studied
regions are the Sylvian fissures, which separate the fron-
tal and temporal lobes. For example, the posterior end
of the Sylvian fissure in the right hemisphere is higher
than in the left, whereas the left fissure has a more gentle
slope
10,11,18
(FIG. 1a,b). The planum temporale, a region in
the posterior portion of the superior temporal sulcus, is
larger in the left hemisphere than in the right in more
than 65% of adult brains and 56–79% of fetal or infant
brains examined
19–22
. More recently, digital brain maps
have generated three-dimensional images of human
brains and further revealed cortical asymmetries
10
.
Moreover, a new population average, landmark- and
surface-based (PALS) atlas approach has shown the most
consistent asymmetries to be in and near the Sylvian
fissures
(FIG. 1a,b) and the superior temporal sulcus
11
(FIG. 1c)
.
The differences in neuronal cell type or cell
organization that might underlie these gross anatomi-
cal differences are unclear. Studies have shown that
language-related areas of the left cortex might contain
more and larger layer 3 pyramidal cells than corre-
sponding areas in the right hemisphere
23
. Rosen
24
and Galaburda
25
used histological studies to suggest
that the asymmetrical regions in the cortex might be
the results of differences in neuron numbers but not
packing density. However, the tremendous size of the
human cortex and its extensive and variable folding
pattern make corresponding areas difficult to compare
with certainty.
In addition to the asymmetries related to language
abilities, such as those of the Sylvian fissures and the pla-
num temporale, anatomical asymmetries associated with
hand use have also been detected in other regions in the
human cerebral cortex. In the primary somatosensory
cortex (S1), studies using
magnetic source imaging have
shown that the cortical representation of the right hand
is larger than the one of the left hand in right-handers,
and vice versa in left-handers
26
. Moreover, the left central
sulcus, a large inward fold marking the division between
the frontal and parietal lobes, is deeper than the right
central sulcus in right-handers
27
. Inter-hemispheric
comparison has further revealed a significant increase
of the hand and finger movement representation in the
primary motor cortex opposite to the preferred hand
28
.
In contrast to these findings, other reports have
shown no obvious correlation of handedness and brain
asymmetries. For example, using voxel-based morpho-
metry, Good et al.
29
did not detect effects of hand use on
asymmetrical morphology in sensorimotor regions of
more than 465 normal adult brains. Although different
methodologies used in these studies could lead to oppo-
site conclusions, the analyses of anatomical asymmetries
associated with handedness in the primary sensory and
motor cortices are compelling.
Handedness and language ability are two of the most
obvious lateralized behaviours in humans. Taking note
of the convergence of functional and anatomical studies,
the asymmetrical cortical controls that regulate handed-
ness are tightly correlated with those for language ability.
But how are these controls established in humans?
Correlation of hand use and language ability
That most humans (more than 90%) prefer to use their
right hand has been observed in almost all cultures and
ethnicities throughout history
12,30
. Statistical studies sug-
gest that handedness might be under genetic control.
There are at least two well-known genetic models of
handedness
31,32
(BOX 1), and although these models seem
to reflect genetic mechanisms of cortical asymmetry
and handedness, genes that regulate these asymmetries
have not been identified. Furthermore, the question of
whether a single gene can control such complex processes
in the CNS is still unanswered
31
. Nevertheless, the single-
gene models proposed by Annett
33
and McManus
34
fit
statistical data of cerebral dominance for handedness
in humans. Identifying the gene(s) that regulates brain
asymmetry and handedness remains an appealing but
challenging task.
Why is there a left-hemisphere bias for handed-
ness and language ability? Preferred hand use has been
observed even at embryonic and fetal stages in humans,
long before language ability is developed. For example, in
most human embryos, the right hand is more developed
than the left at 7 weeks
35
. Using ultrasound, it has been
observed that at 15 weeks most fetuses prefer to suck their
right thumb, hinting that handedness is present prior
to birth
36
. Interestingly, Hepper et al.
37
followed up this
study of 75 individuals. They found that the 60 fetuses
that preferred to suck their right thumb were indeed
right-handed as teenagers, and of the 15 fetuses that
Figure 1 | Anatomical asymmetries in the human cerebral cortex. Coronal (a) and
horizontal (b) MRI of the Sylvian fissures (red arrows) in the human brain. The Sylvian
fissures separate the frontal lobes and temporal lobes in both hemispheres. The left (L)
fissure is more ventral (V) and extends further towards the posterior (P) than does the
right (R). Illustration of differences of sulcal depth between the left and right
hemispheres (c). Cortical regions that are deeper in the right hemisphere are shown in
red and yellow, whereas regions that are deeper in the left are shown in blue and green.
A, anterior; D, dorsal. Modified, with permission, from
REF. 11 © (2005) Elsevier Science.
REVIEWS
656
|
AUGUST 2006
|
VOLUME 7 www.nature.com/reviews/neuro
© 2006 Nature Publishing Group
© 2006 Nature Publishing Group
© 2006 Nature Publishing Group
Magnetic source imaging
The detection of the changing
magnetic fields that are
associated with brain activity
and their subsequent
overlaying on magnetic
resonance images to identify
the precise source of the signal.
Paw preference
In a food-reaching task, paw
preference measures the
frequency of using either the
left or the right front paw to
reach food. It has been
observed in mice, rats, cats and
dogs.
preferred to suck their left thumb, 5 were right-handed
and 10 were left-handed. Moreover, several early studies
have shown that some cortical sulci and gyri, such as the
temporal gyri, are asymmetrical in human fetal brains
from 10–44 weeks
22,38,39
. Using the measurement of cer-
ebral blood flow, Chiron et al.
40
found that the maturation
of the right hemisphere precedes the left in the brains of
human infants between 1 and 3 years of age. This asym-
metrical pattern shifts towards the left hemisphere during
the process of development of language abilities at about
the age of 3
(REF.40). Additionally, Trevarthen
41
observed
that expressive gestures, such as communicative hand
movement, were asymmetrical in infants.
These results imply that anatomical and functional
brain asymmetry precedes uptake of information from
the environment and cognitive development. This in
turn suggests the existence of intrinsic controls that reg-
ulate brain asymmetries at early stages. Although these
kinds of study are interesting, one has to keep in mind
whether this early right-hand preference is controlled by
high-level regulation in the left hemisphere of the cortex,
or by spontaneous movement regulation in the spinal
cord. Furthermore, whether early brain asymmetries
contribute more to handedness or to language ability still
remains an intriguing and challenging question
12
.
Because anatomical asymmetries of certain areas in
the human brain are associated with language ability,
several researchers have made efforts to map asym-
metries of the planum temporale and Brocas area in the
brains of chimpanzees and great apes
42–44
. They found
that simian brains also have asymmetries, which resem-
ble those of humans. These studies suggest that brain
structures associated with language ability might have
existed before humans evolved. However, it is not clear
whether vocal communication is asymmetrical in non-
human primate brains or how these asymmetrical struc-
tures are involved in vocal processing
45
. Furthermore,
because of the complex structure of the cerebral cortex,
the mapping of areas that correspond in human and
primate brains is difficult
46
.
Which hemispheric asymmetry (for handedness
or for language ability) appeared first in evolution still
remains a puzzle. Further comparative studies of brain
asymmetry and handedness in non-human primates will
help us to understand the relationship between handed-
ness and language ability in humans
47
.
Evolutionary mechanisms of biased hand use
More than 90% of the human population is right-handed,
and biased hand use is also observed in non-human
primates and other mammals. But whether there is a
dominant preference for one hand at a population level
is still debatable. What has made most humans right-
handed during evolution is still unknown.
Handedness in non-human primates. There are many
contradictory reports about hand use in non-human
primates, such as chimpanzees. A broad range of
manual tasks have been observed in chimpanzees,
including simple reaching, bimanual feeding, coordi-
nated bimanual actions, throwing, manual gestures and
so on
48,49
. Although these observations have led to the
argument that, for some measures, chimpanzees are
right-handed, most of these findings are from captive
great apes; evidence of population-level handedness in
wild apes is extremely sparse
48,49
. Therefore, these stud-
ies do not conclude that there is a dominant preference
for hand use in non-human primates at the population
level.
A recent report of hand preferences during termite-
fishing/probing actions of chimpanzees is interesting.
First, Lonsdorf and Hopkins
50
studied wild chimpanzees
living in the Gombe National Park, Tanzania, but not
captive chimpanzees. Second, they observed termite-
fishing actions, which require fine motor skill. They
claimed that directional biases in hand use vary depend-
ing on the type of tool use. Therefore, the question of
whether there is strong handedness in non-human
primates might be confounded by biases in the types
of motor skill required. Tests that better discriminate
behavioural biases in wild primates are needed before
any definitive conclusions can be drawn.
Paw preference in other mammals. A well-studied
lateralized manual behaviour of many mammals is the
food-reaching task, defined by
paw preference. Although
paw preference has been observed and studied in mice,
rats, cats and dogs, it does not seem biased to either the
left or the right front paw at a population level
51–56
. For
example, paw preference was observed among domestic
cats, but no significant bias in preference was found at
the level of the group
53
. In mice, although there is paw
preference in each individual mouse, approximately half
of the mice studied preferred to use the left paw and half
preferred to use the right
57,58
. There are also differences
in the strength and direction of paw preference between
mouse strains, indicating that genetic background is an
important influence on this behaviour
55
.
Paw preference in mice has encouraged scientists
to find the genetic causes of this manual lateralization.
Collins
59
attempted to breed left- or right-handed mice,
and although he was unable, by inbreeding, to create a
mouse strain that prefers to use only the left or the right
front paw, he did succeed in generating mice that show
a strong lateralization. For example, the HI strain was
bred using mice that showed consistent right or left paw
use in a food-reaching task, and the LO strain was bred
using mice with little overall paw preference
59
.
Box 1 | Genetic models of human handedness
Marian Annett
33
has proposed that the inheritance of the right-shift (RS) gene shifts the
manual skills in favour of the right hand instead of the left. She emphasizes that RS
influences left cerebral dominance rather than handedness; the effect of RS is to impair
the control of speech systems in the right hemisphere, allowing language abilities to
function in the left side. Handedness is just the secondary consequence of the left-
cerebral cortical dominance
93
.
The other model was proposed by Chris McManus
34
. He suggests that handedness is
controlled by two alleles: D (dextral) and C (chance). According to his model, the
homozygous DD genotype produces only right-handers, whereas the homozygous CC
genotype produces a random mixture of 50% right-handers and 50% left-handers.
Furthermore, the heterozygote, DC, produces 25% left-handers and 75% right-handers.
This model reflects the Mendelian model of genotype and phenotype distribution.
REVIEWS
NATURE REVIEWS
|
NEUROSCIENCE VOLUME 7
|
AUGUST 2006
|
657
© 2006 Nature Publishing Group
© 2006 Nature Publishing Group
© 2006 Nature Publishing Group
Is paw preference associated with lateralized brain
anatomy and/or function in mice? The direction of paw
preference seems to correlate with the dominance of
dopamine expression levels in the brain: a mouse that
prefers to use the left front paw has a higher dopamine
level in the left hemisphere than in the right, although
the physiological implications of this correlation are
unknown
60,61
. Selectively bred mice (the O/AP strain)
with supernumerary whiskers on the right side of the
face and corresponding supernumerary barrels in the left
barrel field showed a higher preference for using the left
front paw. Likewise, mice with supernumerary whiskers
on the left side of the face preferred to use the right front
paw
62
. This biased front-paw use might be the result of
competition for cortical representation between the size
of the motor cortex and the somatosensory barrel field in
which the whiskers are represented. A larger S1 (the bar-
rel field) in the left hemisphere, for example, might make
the size of the left motor cortex smaller, and lead to biased
left front-paw use controlled by the right hemisphere.
Moreover, the areas of whisker-pad representation in
the S1 between the left and right hemispheres of adult rats
have shown striking variations (that is, asymmetries) in
individuals. However, these asymmetries are not biased
to either the left or the right hemispheres
63,64
. Does the
asymmetry of the S1 provide a hint that paw preference
has corresponding and asymmetrical cortical structures
that control it? It will be interesting to map the sizes of the
S1 regions and the direction of paw preference in mice.
In terms of the distribution of hand use, there is a
consistent 9:1 ratio of right/left hand preference reported
in humans, higher than has been reported for any other
mammal
12,50
. The directional manual task in mice, like
paw preference, might be regulated by the formation of
stable neural circuits, but it has a random distribution at
the population level. Given these examples of low or no
bias in chimpanzees and mice, it is an intriguing puzzle
as to how consistent right-hand dominance evolved in
humans. Corballis
31
has proposed that there was a genetic
mutation in hominid evolution that promoted preferen-
tial use of the right hand and is now seen in modern
humans. This evolutionary bias might be advantageous
as it could increase brain capacity and social cohesion
65
.
Applying genomic approaches, particularly the complete
sequencing of the human and chimpanzee genomes,
will provide a considerable insight into the evolution-
ary mechanisms of lateralized human behaviours and
human brain development and asymmetry
47,66–69
.
Body and brain asymmetries
Little is known about the genetic causes of brain ana-
tomical and functional asymmetries
70
. By contrast, stud-
ies of molecular regulation of asymmetries in the visceral
organs, such as the heart and lungs, have made encour-
aging progress
(BOX 2). Inspired by the identification
of molecules that have essential roles in visceral organ
asymmetry, researchers have succeeded in identifying
molecules that regulate brain asymmetry in zebrafish,
perhaps the only species that has been well studied
with respect to brain asymmetry
(BOX 3). Conserved
molecules that regulate body asymmetry, such as
Nodal
and ion channel related gene products, are also essential
for regulating asymmetry of the epithalamus, a small
structure of the diencephalons
71,72
.
Do the same molecular mechanisms that regulate
body asymmetry also cause human brain asymmetry?
The complete reversal of normal organ position, such as
heart and lungs, is called situs inversus. With the excep-
tion of the reversed frontal and occipital petalia observed
using anatomical and functional MRI techniques, the
left-hemisphere dominance for language was still found
to be similar in individuals with situs inversus and in
normal subjects
73
. Nor did lateralization of auditory
processing show any differences between individuals
with situs inversus and normal subjects
74
. Moreover,
50% of individuals with
Kartagener’s syndrome, a dis-
order caused by cilia with a decreased or total absence
of motility, have been found to have situs inversus
75
. This
disorder might confirm the function of cilia in regulat-
ing visceral organ asymmetry, as defects in cilia mobility
might result in the random distribution of NODAL mol-
ecules
(BOX 2). However, the situs inversus patients with
Kartagener’s syndrome developed normal handedness
76
.
Therefore, it seems reasonable that the molecules and
mechanisms that regulate visceral organ asymmetries
might be distinct from those that regulate brain asym-
metries and handedness
70,77
.
Box 2 | Molecular regulation of visceral organ asymmetry
Several studies have elegantly addressed the molecular regulation of the left–right
asymmetry of internal organs, such as the heart, stomach, lungs and intestines of
vertebrate bodies
94,95
. Three signalling pathways (SHH, FGF8 and NODAL) have crucial
roles in left–right body determination
96,97
. In the chick embryo, sonic hedgehog (SHH) and
its target gene caronte (CAR ) are expressed to the left of the chick node (a structure of
the body organizer in chicks) , whereas FGF8 is expressed to the right of the chick node
98–
100
. Misexpression of SHH on the right side of the node is sufficient to induce heart
formation on the right
98
. In the mouse embryo, neither SHH nor FGF8 is expressed
asymmetrically
97
. Instead, the unidirectional rotation of monocilia on the surface of the
mouse node directs the NODAL molecule to the left and activates its downstream genes,
such as Lefty2 and Pitx
101–103
. Moreover, early differential ion flux, such as that driven by
the H
+
and K
+
ATPase transporter, was shown to cause early body asymmetry
104
. The
neurotransmitter serotonin was recently reported upstream of asymmetrically expressed
genes (such as SHH) in chick and frog embryos, and has a role in early patterning of the
left–right body axis
105,106
.
Box 3 | Molecular regulation of zebrafish brain asymmetry
The asymmetries of the epithalamus, which are exemplified by the habenula and the
pineal complex, are well studied in zebrafish
71
. Although the functional consequences of
epithalamus asymmetry are still unclear, it seems to be involved in regulating sexual
activities, photoreception and communication
71,107
. Both the habenula and the pineal
complex show asymmetries on the left side in zebrafish. Interestingly, genes involved in
the Nodal pathway, such as the Nodal-related gene cyclops and Nodal downstream
genes lefty1 and pitx2, were shown to control the laterality of the asymmetry, suggesting
that a conserved signalling pathway that regulates visceral laterality also underlies an
anatomical asymmetry of the zebrafish brain
72,108–110
. Moreover, a recent report has shown
that the frequent situs inversus (fsi) line of zebrafish displayed concordant reversal of
visceral organ and neuroanatomical asymmetries in the diencephalons
111
. Interestingly,
fsi zebrafish also showed a reversal of some behavioural responses, which has not been
detected in mammals with situs inversus
111
. These results indicate that the molecular
regulation of brain and body asymmetries can be species specific.
REVIEWS
658
|
AUGUST 2006
|
VOLUME 7 www.nature.com/reviews/neuro
© 2006 Nature Publishing Group
© 2006 Nature Publishing Group
© 2006 Nature Publishing Group
RABL2B
SUFU
SH3GL2
NEUROD6
NEDD4
TMSB4X
MAGOH
HEY1
LAMR1
IGF1
IFIT4
ID2
DAPPER1
CUTL1
BICD2
BAIAP2
ATP2B3
Le
IGFBP5
MYT1L
MEF2C
MATR3
P311
STMN4
LMO4
BAI1
AUTS2
ARHA
Right
8 76543210
0 0.5 1 1.5 2 2.5 3
Relative gene expression levels
abc
Epidermal ectoderm
Roof plate
Floor plate
Notochord
Neural
tube
Anterior
neural ridge
Brain cortical
hemispheres
Serial analysis of gene
expression
(SAGE). A method for
comprehensive analysis of
gene expression levels and
patterns using PCR
amplification and generating
SAGE libraries.
Our recent studies, using a genomic screening
approach, further support this idea. Using a
serial analysis
of gene expression
(SAGE) technique, we measured gene
expression levels in the left and right hemispheres of
human fetal brains
78
. We verified 27 genes that are dif-
ferentially expressed in the hemispheres of 12-week-old
human fetal brains by using either real-time reverse
transcription (RT)-PCR or in situ hybridization
(FIG. 2).
Most genes identified using SAGE analyses function in
signal transduction and gene expression regulation
47
.
Among them, the transcription factor Lim domain only 4
(
LMO4) showed consistent asymmetry of expression in
human fetal brains at 12 weeks and 14 weeks, and less so
at 16 and 17 weeks
78
. However, we did not detect genes
that have essential roles in visceral organ asymmetry,
such as genes involved in the sonic hedgehog (
SHH)
or NODAL pathways, which are also differentially
expressed in human fetal brains
78
. Because the earliest
stage analysed was in the human fetus at 12 weeks, it
cannot be ruled out that molecules regulating body
asymmetry might also be differentially expressed in
human embryonic brains (for example, at 8–10 weeks).
It will be interesting to measure gene expression levels
in the left and right hemispheres in human embryonic
brains (8–10 weeks) using SAGE or cDNA microarray
approaches.
Molecular regulation of brain asymmetry
An essential step leading towards asymmetry is to break
symmetry
79
. Although the initiation mechanisms of
breaking symmetry are still unknown, an uneven dis-
tribution of molecules that are essential for left–right
body axis patterning could be important for this bio-
logical event.
How, then, is symmetry broken in the CNS?
Neuroepithelial cells divide vigorously, fold dorsally and
form a neural groove during early embryonic develop-
ment
80
. The neural groove continues to grow, and the
dorsal parts meet at the midline and fuse to form a neural
tube
80
. Neural tube development is accompanied by the
formation of the
notochord and the induction of the floor
plate
81
. Numerous studies have shown that the notochord
is a patterning centre for the ventral neural tube
82
. In the
forebrain, a structure anterior to the notochord is called
the prechordal plate
83
. Molecules secreted from the noto-
chord, such as SHH, function as
morphogens to induce
and maintain the ventral property and neural cell types
in the spinal cord
82
. Similar morphogens also induce
and pattern the forebrain, and are probably secreted
from the notochord or the prechordal plate
84
. Moreover,
the patterning centres are not limited to the notochord
and the ventral neural tube. Morphogens, such as bone
morphogenetic proteins and WNTs, are secreted from
the roof plate in the neural tube
85
.
One possible mechanism for breaking symmetry
in the brain is that the morphogens secreted from the
ventral (floor plate or prechordal plate) and/or dorsal
(roof plate) midlines are distributed differently between
the left and right (
FIG. 3a,b). The different expression
levels of morphogens induce differential expression
of downstream transcription factors, such as LMO4
(REF. 78), and eventually lead to brain asymmetry.
Recent studies of the molecular regulation of corti-
cal regionalization have identified a patterning centre
in the anterior cortex
2–4
. An important molecule that is
secreted from the anterior cortical region is fibroblast
growth factor 8 (
FGF8)
86
. The ectopic expression of
FGF8 can expand the motor cortex and shift the func-
tional regions of the cortex caudally
86
. It is possible that
the expression levels of morphogens secreted from the
anterior cortical region might be different in the left and
right hemispheres
(FIG. 3c). The asymmetrical expression
Figure 2 | Asymmetrically expressed genes in 12-week-old human fetal brains,
detected by serial analysis of gene expression and real-time reverse
transcription (RT)-PCR. The cDNA made from the perisylvian regions of the left and
right hemispheres of two 12-week-old human fetal cortices were used as templates for
real-time RT-PCR. The relative gene expression levels are average ratios of gene
expression detected by RT-PCR between the left and right hemispheres of two brains.
These differential gene expression levels also match those measured by serial analysis of
gene expression (SAGE)
78
. 27 genes showing consistent differential expression are listed.
Among them, 17 genes were highly expressed in the left perisylvian regions, whereas 10
genes were highly expressed in the right.
Figure 3 | Three models of molecular induction of brain asymmetry. a | Whereas the
neural tube (red) is derived from the ectoderm, the notochord (blue) is derived from the
mesoderm and accompanies the neural tube formation. During neural tube
development, the most dorsal part of the neural tube becomes the roof plate (green),
whereas the most ventral part of the neural tube becomes the floor plate (blue). The
forebrain is developed from the most rostral region in the neural tube. In the forebrain,
the morphogens secreted from the notochord or the prechordal plate (not shown) might
be differentially distributed between the left and right neural tube. b | Similarly, uneven
secretion of molecules might also occur in the roof plate. Different morphogen
expression levels in the left and right neural tube might break the symmetry of brain
patterning and induce asymmetrical expression of downstream genes. c | Anterior
signals might also induce cortical asymmetry. The patterning centre in the most rostral
neural tube — for example, the anterior neural ridge (green) — could be a source for
cortical asymmetrical patterning. The distribution of molecules secreted from this area
might be different in the left and right hemispheres.
REVIEWS
NATURE REVIEWS
|
NEUROSCIENCE VOLUME 7
|
AUGUST 2006
|
659
© 2006 Nature Publishing Group
© 2006 Nature Publishing Group
© 2006 Nature Publishing Group
RLR L
Notochord
A structure composed of cells
derived from the mesoderm
and defines the primitive axis
of the embryo. It lies between
the neural tube (spinal cord)
and the gut.
Morphogen
A diffusible substance that
carries information influencing
the movement and
organization of cells during
morphogenesis. It normally
forms a concentration gradient.
of regional markers, such as LMO4, could reflect asym-
metrical topographic mapping of functional regions
along the anterior–posterior axis in the cortex
78
.
Future work will include the identification of more of
the morphogens that pattern the early cortex and their
downstream targets. Consistent with regionalization
in the cortex, which involves complex gene expression
and regulation
4
, brain asymmetry and handedness are a
conjugated result of molecular regulation, neural con-
nections and plasticity.
Conclusions and future perspectives
The challenge of studying brain asymmetry is that
because the obvious anatomical and functional asym-
metries have been identified largely in humans, we can-
not carry out direct experiments. The recent behavioural
studies in non-human primates, such as the investigation
of handedness in chimpanzees, might help us to better
understand how human handedness has evolved
50
. In
particular, the comparison of human and chimpanzee
genomes has enriched our knowledge of the evolution-
ary mechanisms of human brain development
47,66–69
.
Similar studies might help us to understand the evolu-
tionary regulation of human brain asymmetry.
Several human neurological disorders show dis-
rupted normal brain asymmetry. For example, reduced
and reversed anatomical brain asymmetry has been
reported in individuals with schizophrenia, autism or
dyslexia
87–90
, suggesting a potential indirect relationship
between the causes of these disorders and the asym-
metrical development of the human cerebral cortex.
Recently, several studies have reported clinical cases of
polymicrogyria — a malformation of cortical develop-
ment that is characterized by many small gyri in the
cortex — that occurs only on one side of the cortex;
this is known as unilateral polymicrogyria
91,92
(FIG. 4).
Patients have seizures, motor dysfunction and mental
retardation. A genetic cause of unilateral right-sided
polymicrogyria is suggested by the existence of several
pedigrees in which the disorder is present in more than
one individual of an affected family
92
. These studies
indicate that unilateral polymicrogyria can be inherited
as a Mendelian trait, suggesting that there might be a
gene that is required for the development of the right
perisylvian region
92
. Using forward genetic approaches
to map genes that cause disrupted brain asymmetry
might reveal their normal function in asymmetrical
development of the brain.
Faster development and improvement of large-scale
screening approaches at the genomic level could make
the identification of asymmetrically expressed genes in
human and mouse brains easier and quicker. Generating
genetically engineered mice can help us to understand
the functions of these genes in brain development. Using
these mouse models can also help to reveal the neural
circuitries that regulate brain asymmetry and lateral-
ized behaviours. However, unlike visceral organ asym-
metries, which are easy to detect, brain asymmetry relies
largely on fine brain mapping and reliable behavioural
tests. Therefore, the development of molecular imaging
techniques and an improved understanding of lateral-
ized behaviours in rodents will be extremely useful for
studies of brain asymmetry.
Figure 4 | Unilateral polymicrogyria detected using MRI. Polymicrogyria (indicated
by arrows) is detectable in the right hemispheres in both brains shown. An apparent
increase in cortical thickness is observed in the right (R) hemispheres, whereas the
cortices of the left (L) hemispheres appear entirely normal. Modified, with permission,
from
REF. 92 © (2006) Lippincott Williams & Wilkins.
1. Levitt, P., Barbe, M. F. & Eagleson, K. L. Patterning
and specification of the cerebral cortex. Annu. Rev.
Neurosci. 20, 1–24 (1997).
2. O’Leary, D. D. & Nakagawa, Y. Patterning centers,
regulatory genes and extrinsic mechanisms controlling
arealization of the neocortex. Curr. Opin. Neurobiol.
12, 14–25 (2002).
3. Grove, E. A. & Fukuchi-Shimogori, T. Generating the
cerebral cortical area map. Annu. Rev. Neurosci. 26,
355–380 (2003).
4. Sur, M. & Rubenstein, J. L. Patterning and plasticity of
the cerebral cortex. Science 310, 805–810 (2005).
5. Rakic, P. Neuronal migration and contact guidance in
primate telencephalon. Postgrad. Med. J. 54, 25–40
(1978).
6. O’Leary, D. D. Do cortical areas emerge from a
protocortex? Trends Neurosci. 12, 400–406 (1989).
7. Rakic, P. Specification of cerebral cortical areas.
Science 241, 170–176 (1988).
8. Geschwind, D. H. & Miller, B. L. Molecular approaches
to cerebral laterality: development and
neurodegeneration. Am. J. Med. Genet. 101,
370–381 (2001).
Examines expression patterns of genes that regulate
body asymmetries in human fetal brains and
discusses molecular regulation of brain asymmetry.
9. Riss, W. Testing a theory of brain function by computer
methods. III. Detecting cerebral asymmetry in normal
adults. Brain Behav. Evol. 24, 13–20 (1984).
10. Toga, A. W. & Thompson, P. M. Mapping brain
asymmetry. Nature Rev. Neurosci. 4, 37–48 (2003).
Overview of brain anatomical asymmetries mapped
using modern imaging techniques.
11. Van Essen, D. C. A population-average, landmark- and
surface-based (PALS) atlas of human cerebral cortex.
Neuroimage 28, 635–662 (2005).
Maps brain asymmetries and beautifully
demonstrates consistent asymmetries in the
Sylvian fissures and the superior temporal sulcus.
12. Corballis, M. C. From mouth to hand: gesture, speech,
and the evolution of right-handedness. Behav. Brain.
Sci. 26, 199–208; discussion 208–260 (2003).
13. Broca, P. Remarques sur le siège de la faculté du
langage articulé, suivies d’une observation d’aphémie
(perte de la parole). Bull. Soc. Anthropol. 6, 330–357
(1861) (in French).
14. Wernicke, C. Der Aphasische Symptomenkomplex:
Eine Psychologische Studie auf Anatomischer Basis
(Cohn und Welgert, Breslau, 1874) (in German).
15. Gazzaniga, M. S. Forty-five years of split-brain
research and still going strong. Nature Rev. Neurosci.
6, 653–659 (2005).
16. Gazzaniga, M. S. Brain and conscious experience. Adv.
Neurol. 77, 181–192; discussion 192–193
(1998).
17. Borod, J. C., Bloom, R. L., Brickman, A. M.,
Nakhutina, L. & Curko, E. A. Emotional processing
deficits in individuals with unilateral brain damage.
Appl. Neuropsychol. 9, 23–36 (2002).
18. Rubens, A. B., Mahowald, M. W. & Hutton, J. T.
Asymmetry of the lateral (Sylvian) fissures in man.
Neurology 26, 620–624 (1976).
19. Shapleske, J., Rossell, S. L., Woodruff, P. W. & David,
A. S. The planum temporale: a systematic,
quantitative review of its structural, functional and
clinical significance. Brain Res. Brain Res. Rev. 29,
26–49 (1999).
20. Hirayasu, Y. et al. Planum temporale and Heschl gyrus
volume reduction in schizophrenia: a magnetic
REVIEWS
660
|
AUGUST 2006
|
VOLUME 7 www.nature.com/reviews/neuro
© 2006 Nature Publishing Group
© 2006 Nature Publishing Group
© 2006 Nature Publishing Group
resonance imaging study of first-episode patients.
Arch. Gen. Psychiatry 57, 692–699 (2000).
21. Geschwind, N. & Levitsky, W. Human brain: left–right
asymmetries in temporal speech region. Science 161,
186–187 (1968).
22. Chi, J. G., Dooling, E. C. & Gilles, F. H. Left–right
asymmetries of the temporal speech areas of the
human fetus. Arch. Neurol. 34, 346–348 (1977).
23. Hutsler, J. J. The specialized structure of human
language cortex: pyramidal cell size asymmetries
within auditory and language-associated regions of
the temporal lobes. Brain Lang. 86, 226–242
(2003).
24. Rosen, G. D. Cellular, morphometric, ontogenetic and
connectional substrates of anatomical asymmetry.
Neurosci. Biobehav. Rev. 20, 607–615 (1996).
25. Galaburda, A. M., Rosen, G. D. & Sherman, G. F.
Individual variability in cortical organization: its
relationship to brain laterality and implications to
function. Neuropsychologia 28, 529–546 (1990).
26. Soros, P. et al. Cortical asymmetries of the human
somatosensory hand representation in right- and left-
handers. Neurosci. Lett. 271, 89–92 (1999).
27. Amunts, K. et al. Asymmetry in the human motor
cortex and handedness. Neuroimage 4, 216–222
(1996).
28. Volkmann, J., Schnitzler, A., Witte, O. W. & Freund, H.
Handedness and asymmetry of hand representation in
human motor cortex. J. Neurophysiol. 79,
2149–2154 (1998).
29. Good, C. D. et al. Cerebral asymmetry and the effects
of sex and handedness on brain structure: a voxel-
based morphometric analysis of 465 normal adult
human brains. Neuroimage 14, 685–700 (2001).
30. Coren, S. & Porac, C. Fifty centuries of right-
handedness: the historical record. Science 198,
631–632 (1977).
31. Corballis, M. C. The genetics and evolution of
handedness. Psychol. Rev. 104, 714–727 (1997).
32. Klar, A. J. Genetic models for handedness, brain
lateralization, schizophrenia, and manic-depression.
Schizophr. Res. 39, 207–218 (1999).
33. Annett, M. The distribution of manual asymmetry. Br.
J. Psychol. 63, 343–358 (1972).
Proposes a genetic model for the regulation of
cortical asymmetry and handedness.
34. McManus, I. C. Handedness, language dominance and
aphasia: a genetic model. Psychol. Med. Monogr.
Suppl. 8, 1–40 (1985).
Proposes and discusses a genetic model that leads
to preferential handedness in humans.
35. O’Rahilly, R. & Müller, F. Developmental Stages in
Human Embryos (Carnegie Institute of Washington,
Publication 637, Washington DC, 1987).
36. Hepper, P. G., Shahidullah, S. & White, R. Handedness
in the human fetus. Neuropsychologia 29, 11071111
(1991).
37. Hepper, P. G., Wells, D. L. & Lynch, C. Prenatal thumb
sucking is related to postnatal handedness.
Neuropsychologia 43, 313–315 (2005).
38. Chi, J. G., Dooling, E. C. & Gilles, F. H. Gyral
development of the human brain. Ann. Neurol. 1,
86–93 (1977).
39. Galaburda, A. M., LeMay, M., Kemper, T. L. &
Geschwind, N. Right–left asymmetrics in the brain.
Science 199, 852–856 (1978).
40. Chiron, C. et al. The right brain hemisphere is
dominant in human infants. Brain 120, 1057–1065
(1997).
41. Trevarthen, C. Lateral asymmetries in infancy:
implications for the development of the hemispheres.
Neurosci. Biobehav. Rev. 20, 571–586 (1996).
42. Gannon, P. J., Holloway, R. L., Broadfield, D. C. &
Braun, A. R. Asymmetry of chimpanzee planum
temporale: humanlike pattern of Wernicke’s brain
language area homolog. Science 279, 220–222
(1998).
43. Cantalupo, C. & Hopkins, W. D. Asymmetric Broca’s
area in great apes. Nature 414, 505 (2001).
44. Hopkins, W. D., Marino, L., Rilling, J. K. & MacGregor,
L. A. Planum temporale asymmetries in great apes as
revealed by magnetic resonance imaging (MRI).
Neuroreport 9, 2913–2918 (1998).
45. Hutsler, J. & Galuske, R. A. Hemispheric asymmetries
in cerebral cortical networks. Trends Neurosci. 26,
429–435 (2003).
46. Sherwood, C. C., Broadfield, D. C., Holloway, R. L.,
Gannon, P. J. & Hof, P. R. Variability of Broca’s area
homologue in African great apes: implications for
language evolution. Anat. Rec. A Discov. Mol. Cell.
Evol. Biol. 271
, 276–285 (2003).
47. Sun, T., Collura, R. V., Ruvolo, M. & Walsh, C. A.
Genomic and evolutionary analyses of asymmetrically
expressed genes in human fetal left and right cerebral
cortex. Cereb. Cortex 16, i18–i25 (2006).
48. Hopkins, W. D. & Cantalupo, C. Individual and setting
differences in the hand preferences of chimpanzees
(Pan troglodytes): a critical analysis and some
alternative explanations. Laterality 10, 65–80 (2005).
49. McGrew, W. C., Marchant, L. F., Wrangham, R. W. &
Klein, H. Manual laterality in anvil use: wild
chimpanzees cracking Strychnos fruits. Laterality 4,
79–87 (1999).
50. Lonsdorf, E. V. & Hopkins, W. D. Wild chimpanzees
show population-level handedness for tool use. Proc.
Natl Acad. Sci. USA 102, 12634–12638 (2005).
Shows that preferential hand use in wild
chimpanzees depends on the specific task.
51. Wells, D. L. Lateralised behaviour in the domestic dog,
Canis familiaris. Behav. Processes 61, 27–35 (2003).
52. Tan, U. Paw preferences in dogs. Int. J. Neurosci. 32,
825–829 (1987).
53. Fabre-Thorpe, M., Fagot, J., Lorincz, E., Levesque, F. &
Vauclair, J. Laterality in cats: paw preference and
performance in a visuomotor activity. Cortex 29,
15–24 (1993).
54. Bulman-Fleming, M. B., Bryden, M. P. & Rogers, T. T.
Mouse paw preference: effects of variations in testing
protocol. Behav. Brain Res. 86, 79–87 (1997).
55. Waters, N. S. & Denenberg, V. H. Analysis of two
measures of paw preference in a large population of
inbred mice. Behav. Brain Res. 63, 195–204 (1994).
56. Guven, M., Elalmis, D. D., Binokay, S. & Tan, U.
Population-level right-paw preference in rats assessed
by a new computerized food-reaching test. Int. J.
Neurosci. 113 , 1675–1689 (2003).
57. Signore, P. et al. An assessment of handedness in
mice. Physiol. Behav. 49, 701–704 (1991).
58. Biddle, F. G., Coffaro, C. M., Ziehr, J. E. & Eales, B. A.
Genetic variation in paw preference (handedness) in
the mouse. Genome 36, 935–943 (1993).
59. Collins, R. L. Reimpressed selective breeding for
lateralization of handedness in mice.
Brain Res. 564,
194–202 (1991).
60. Cabib, S. et al. Paw preference and brain dopamine
asymmetries. Neuroscience 64, 427–432 (1995).
61. Barneoud, P., le Moal, M. & Neveu, P. J. Asymmetric
distribution of brain monoamines in left- and right-
handed mice. Brain Res. 520, 317–321 (1990).
62. Barneoud, P. & Van der Loos, H. Direction of
handedness linked to hereditary asymmetry of a
sensory system. Proc. Natl Acad. Sci. USA 90,
3246–3250 (1993).
63. Riddle, D. R. & Purves, D. Individual variation and
lateral asymmetry of the rat primary somatosensory
cortex. J. Neurosci. 15, 4184–4195 (1995).
64. Chen-Bee, C. H. & Frostig, R. D. Variability and
interhemispheric asymmetry of single-whisker
functional representations in rat barrel cortex.
J. Neurophysiol. 76, 884–894 (1996).
65. Halpern, M. E., Gunturkun, O., Hopkins, W. D. &
Rogers, L. J. Lateralization of the vertebrate brain:
taking the side of model systems. J. Neurosci. 25,
10351–10357 (2005).
Summary of recent studies of vertebrate brain
asymmetries, particularly brain asymmetries in
zebrafish and chicks.
66. Hill, R. S. & Walsh, C. A. Molecular insights into
human brain evolution. Nature 437, 64–67 (2005).
67. Initial sequence of the chimpanzee genome and
comparison with the human genome. Nature 437,
69–87 (2005).
Accomplished the initial sequencing of the
chimpanzee genome. This work will have a
considerable impact on our understanding of the
genomic regulation of human evolution.
68. Khaitovich, P. et al. Parallel patterns of evolution in
the genomes and transcriptomes of humans and
chimpanzees. Science 309, 1850–1854 (2005).
69. Preuss, T. M., Caceres, M., Oldham, M. C. &
Geschwind, D. H. Human brain evolution: insights from
microarrays. Nature Rev. Genet. 5, 850–860 (2004).
70. Hobert, O., Johnston, R. J. Jr & Chang, S. Left-right
asymmetry in the nervous system: the Caenorhabditis
elegans model. Nature Rev. Neurosci. 3, 629–640
(2002).
Overview of the molecular mechanisms that
regulate asymmetries in the Caenorhabditis
elegans nervous system.
71. Concha, M. L. & Wilson, S. W. Asymmetry in the
epithalamus of vertebrates. J. Anat. 199, 63–84
(2001).
72. Halpern, M. E., Liang, J. O. & Gamse, J. T. Leaning to
the left: laterality in the zebrafish forebrain. Trends
Neurosci. 26, 308–313 (2003).
73. Kennedy, D. N. et al. Structural and functional brain
asymmetries in human situs inversus totalis.
Neurology 53, 1260–1265 (1999).
74. Tanaka, S., Kanzaki, R., Yoshibayashi, M., Kamiya, T. &
Sugishita, M. Dichotic listening in patients with situs
inversus: brain asymmetry and situs asymmetry.
Neuropsychologia 37, 869–874 (1999).
75. Carlen, B. & Stenram, U. Primary ciliary dyskinesia: a
review. Ultrastruct. Pathol. 29, 217–220
(2005).
76. McManus, I. C., Martin, N., Stubbings, G. F., Chung, E.
M. & Mitchison, H. M. Handedness and situs inversus
in primary ciliary dyskinesia. Proc. Biol. Sci. 271,
2579–2582 (2004).
77. McManus, C. Reversed bodies, reversed brains, and
(some) reversed behaviors: of zebrafish and men. Dev.
Cell 8, 796–797 (2005).
78. Sun, T. et al. Early asymmetry of gene transcription in
embryonic human left and right cerebral cortex.
Science 308, 1794–1798 (2005).
Measured differential gene expression levels in the
Sylvian fissures between the left and right human
fetal brains and identified 27 asymmetrically
expressed genes using SAGE, real-time reverse
transcription (RT)-PCR and in situ hybridization.
79. Palmer, A. R. Symmetry breaking and the evolution of
development. Science 306, 828–833 (2004).
80. Tanabe, Y. & Jessell, T. M. Diversity and pattern in the
developing spinal cord. Science 274, 1115–1123
(1996).
81. Dodd, J., Jessell, T. M. & Placzek, M. The when and
where of floor plate induction. Science 282,
1654–1657 (1998).
82. Jessell, T. M. Neuronal specification in the spinal cord:
inductive signals and transcriptional codes. Nature
Rev. Genet. 1, 20–29 (2000).
83. Rubenstein, J. L., Shimamura, K., Martinez, S. &
Puelles, L. Regionalization of the prosencephalic
neural plate. Annu. Rev. Neurosci. 21, 445–477
(1998).
84. Tannahill, D., Harris, L. W. & Keynes, R. Role of
morphogens in brain growth. J. Neurobiol. 64, 367–
375 (2005).
85. Chizhikov, V. V. & Millen, K. J. Roof plate-dependent
patterning of the vertebrate dorsal central nervous
system. Dev. Biol. 277, 287–295 (2005).
86. Fukuchi-Shimogori, T. & Grove, E. A. Neocortex
patterning by the secreted signaling molecule FGF8.
Science 294, 10711074 (2001).
87. Herbert, M. R. et al. Brain asymmetries in autism and
developmental language disorder: a nested whole-
brain analysis. Brain 128, 213–226 (2005).
88. Hugdahl, K. et al. Central auditory processing, MRI
morphometry and brain laterality: applications to
dyslexia. Scand. Audiol. Suppl. 49, 26–34 (1998).
89. Galaburda, A. M., Menard, M. T. & Rosen, G. D.
Evidence for aberrant auditory anatomy in
developmental dyslexia. Proc. Natl Acad. Sci. USA 91,
8010–8013 (1994).
90. Falkai, P. et al. Loss of Sylvian fissure asymmetry in
schizophrenia. A quantitative post mortem study.
Schizophr. Res. 7, 23–32 (1992).
91. Pascual-Castroviejo, I., Pascual-Pascual, S. I., Viano, J.,
Martinez, V. & Palencia, R. Unilateral polymicrogyria:
a common cause of hemiplegia of prenatal origin.
Brain Dev. 23, 216–222 (2001).
92. Chang, B. S. et al. A familial syndrome of unilateral
polymicrogyria affecting the right hemisphere.
Neurology 66, 133–135 (2006).
93. Annett, M. Handedness and cerebral dominance: the
right shift theory. J. Neuropsychiatry Clin. Neurosci.
10, 459–469 (1998).
94. Wright, C. V. Mechanisms of left–right asymmetry:
what’s right and what’s left? Dev. Cell 1
, 179–186
(2001).
95. Levin, M. Left–right asymmetry in embryonic
development: a comprehensive review. Mech. Dev.
122, 3–25 (2005).
Gives a comprehensive summary of asymmetries in
invertebrates and vertebrates and discusses
molecular mechanisms that lead to these
asymmetries.
96. Capdevila, J., Vogan, K. J., Tabin, C. J. & Izpisua
Belmonte, J. C. Mechanisms of left–right
determination in vertebrates. Cell 101, 9–21 (2000).
97. Hamada, H., Meno, C., Watanabe, D. & Saijoh, Y.
Establishment of vertebrate left–right asymmetry.
Nature Rev. Genet. 3, 103–113 (2002).
REVIEWS
NATURE REVIEWS
|
NEUROSCIENCE VOLUME 7
|
AUGUST 2006
|
661
© 2006 Nature Publishing Group
© 2006 Nature Publishing Group
© 2006 Nature Publishing Group
98. Levin, M., Johnson, R. L., Stern, C. D., Kuehn, M. &
Tabin, C. A molecular pathway determining left-right
asymmetry in chick embryogenesis. Cell 82, 803–814
(1995).
99. Rodriguez Esteban, C. et al. The novel Cer-like protein
Caronte mediates the establishment of embryonic left-
right asymmetry. Nature 401, 243–251 (1999).
100. Boettger, T., Wittler, L. & Kessel, M. FGF8 functions in
the specification of the right body side of the chick.
Curr. Biol. 9, 277–280 (1999).
101. Nonaka, S., Shiratori, H., Saijoh, Y. & Hamada, H.
Determination of left–right patterning of the mouse
embryo by artificial nodal flow. Nature 418, 96–99
(2002).
102. McGrath, J., Somlo, S., Makova, S., Tian, X. &
Brueckner, M. Two populations of node monocilia
initiate left–right asymmetry in the mouse. Cell 114 ,
61–73 (2003).
103. Nonaka, S. et al. Randomization of left–right
asymmetry due to loss of nodal cilia generating
leftward flow of extraembryonic fluid in mice lacking
KIF3B motor protein. Cell 95, 829–837 (1998).
104. Levin, M., Thorlin, T., Robinson, K. R., Nogi, T. &
Mercola, M. Asymmetries in H
+
/K
+
-ATPase and cell
membrane potentials comprise a very early step in
left–right patterning. Cell 111, 77–89 (2002).
105. Fukumoto, T., Blakely, R. & Levin, M. Serotonin
transporter function is an early step in left–right
patterning in chick and frog embryos. Dev. Neurosci.
27, 349–363 (2005).
106. Fukumoto, T., Kema, I. P. & Levin, M. Serotonin
signaling is a very early step in patterning of the left–
right axis in chick and frog embryos. Curr. Biol. 15,
794–803 (2005).
References 96–106 are excellent studies of
molecular regulation of body asymmetries.
107. Harris, J. A., Guglielmotti, V. & Bentivoglio, M.
Diencephalic asymmetries. Neurosci. Biobehav. Rev.
20, 637–643 (1996).
108. Concha, M. L., Burdine, R. D., Russell, C., Schier, A. F.
& Wilson, S. W. A nodal signaling pathway regulates
the laterality of neuroanatomical asymmetries in the
zebrafish forebrain. Neuron 28, 399–409
(2000).
109. Bisgrove, B. W., Essner, J. J. & Yost, H. J. Multiple
pathways in the midline regulate concordant brain,
heart and gut left–right asymmetry. Development
127, 3567–3579 (2000).
110. Essner, J. J., Branford, W. W., Zhang, J. & Yost, H. J.
Mesendoderm and left–right brain, heart and gut
development are differentially regulated by pitx2
isoforms. Development 127, 1081–1093 (2000).
111. Barth, K. A. et al. fsi zebrafish show concordant
reversal of laterality of viscera, neuroanatomy, and a
subset of behavioral responses. Curr. Biol. 15,
844–850 (2005).
Acknowledgements
Owing to space limitations, we apologize for being unable to
cite many excellent papers in this field. We thank the referees
for critical reading and useful comments, and B. Chang for the
MRI images in figure 4. The authors were supported by grants
from the National Institute of Neurological Disorders and
Stroke, National Institutes of Health. C.A.W. is an investigator
of the Howard Hughes Medical Institute.
Competing interests statement
The authors declare no competing financial interests.
DATABASES
The following terms in this article are linked online to:
Entrez Gene: http://www.ncbi.nlm.nih.gov/entrez/query.
fcgi?db=gene
CAR | FGF8 | LMO4 | Nodal | SHH
OMIM: http://www.ncbi.nlm.nih.gov/entrez/query.
fcgi?db=OMIM
Kartagener’s syndrome
FURTHER INFORMATION
Sun’s homepage: http://www.med.cornell.edu/research/
taosun/index.html
Access to this links box is available online.
REVIEWS
662
|
AUGUST 2006
|
VOLUME 7 www.nature.com/reviews/neuro
© 2006 Nature Publishing Group
© 2006 Nature Publishing Group
© 2006 Nature Publishing Group
... Furthermore, concerning handedness, hemispheric asymmetries may influence functional specialization. Generally, the left hemisphere specializes in communicative language functions and logical reasoning, while the right hemisphere plays a larger role in spatial reasoning and emotional processing (Sun and Walsh, 2006;Hartwigsen et al., 2021). Regarding skilled action, the left hemisphere is considered to be heavily involved in movement selection and coordination, even for movement of the ipsilateral side of the body (Serrien et al., 2006). ...
Article
Full-text available
Background Normative childhood motor network resting-state fMRI effective connectivity is undefined, yet necessary for translatable dynamic resting-state-network-informed evaluation in pediatric cerebral palsy. Methods Cross-spectral dynamic causal modeling of resting-state-fMRI was investigated in 50 neurotypically developing 5- to 13-year-old children. Fully connected six-node network models per hemisphere included primary motor cortex, striatum, subthalamic nucleus, globus pallidus internus, thalamus, and contralateral cerebellum. Parametric Empirical Bayes with exhaustive Bayesian model reduction and Bayesian modeling averaging informed the model; Purdue Pegboard Test scores of hand motor behavior were the covariate at the group level to determine the effective-connectivity-functional behavior relationship. Results Although both hemispheres exhibited similar effective connectivity of motor cortico-basal ganglia-cerebellar networks, magnitudes were slightly greater on the right, except for left-sided connections of the striatum which were more numerous and of opposite polarity. Inter-nodal motor network effective connectivity remained consistent and robust across subjects. Age had a greater impact on connections to the contralateral cerebellum, bilaterally. Motor behavior, however, affected different connections in each hemisphere, exerting a more prominent effect on the left modulatory connections to the subthalamic nucleus, contralateral cerebellum, primary motor cortex, and thalamus. Discussion This study revealed a consistent pattern of directed resting-state effective connectivity in healthy children aged 5–13 years within the motor network, encompassing cortical, subcortical, and cerebellar regions, correlated with motor skill proficiency. Both hemispheres exhibited similar effective connectivity within motor cortico-basal ganglia-cerebellar networks reflecting inter-nodal signal direction predicted by other modalities, mainly differing from task-dependent studies due to network differences at rest. Notably, age-related changes were more pronounced in connections to the contralateral cerebellum. Conversely, motor behavior distinctly impacted connections in each hemisphere, emphasizing its role in modulating left sided connections to the subthalamic nucleus, contralateral cerebellum, primary motor cortex, and thalamus. Motor network effective connectivity was correlated with motor behavior, validating its physiological significance. This study is the first to evaluate a normative effective connectivity model for the pediatric motor network using resting-state functional MRI correlating with behavior and serves as a foundation for identifying abnormal findings and optimizing targeted interventions like deep brain stimulation, potentially influencing future therapeutic approaches for children with movement disorders.
... The left hemisphere tends to depict finer details and intricate structures in object visualization, while the right hemisphere better captures the overall shape and silhouette. These results align with previous discoveries that the left hemisphere is involved in detail-oriented tasks while the right hemisphere in more holistic tasks 43,44 . However, our results challenge the notion of strict functional division. ...
Preprint
Full-text available
Understanding the hidden mechanisms behind human's visual perception is a fundamental quest in neuroscience, underpins a wide variety of critical applications, e.g. clinical diagnosis. To that end, investigating into the neural responses of human mind activities, such as functional Magnetic Resonance Imaging (fMRI), has been a significant research vehicle. However, analyzing fMRI signals is challenging, costly, daunting, and demanding for professional training. Despite remarkable progress in artificial intelligence (AI) based fMRI analysis, existing solutions are limited and far away from being clinically meaningful. In this context, we leap forward to demonstrate how AI can go beyond the current state of the art by decoding fMRI into visually plausible 3D visuals, enabling automatic clinical analysis of fMRI data, even without healthcare professionals. Innovationally, we reformulate the task of analyzing fMRI data as a conditional 3D scene reconstruction problem. We design a novel cross-modal 3D scene representation learning method, Brain3D, that takes as input the fMRI data of a subject who was presented with a 2D object image, and yields as output the corresponding 3D object visuals. Importantly, we show that in simulated scenarios our AI agent captures the distinct functionalities of each region of human vision system as well as their intricate interplay relationships, aligning remarkably with the established discoveries of neuroscience. Non-expert diagnosis indicate that Brain3D can successfully identify the disordered brain regions, such as V1, V2, V3, V4, and the medial temporal lobe (MTL) within the human visual system. We also present results in cross-modal 3D visual construction setting, showcasing the perception quality of our 3D scene generation.
... However, large population-health databases in neurooncology do not consistently present tumor laterality, limiting our evaluation of this possibility (50). Hemispheric asymmetry in neuroanatomic structures (51,52) and gene expression (53) have been identified, suggesting that tumors arising from contralateral hemispheres may have biologic distinctions that warrant consideration when exploring gene-brain associations. ...
Article
Full-text available
Introduction Like other forms of neuropathology, gliomas appear to spread along neural pathways. Accordingly, our group and others have previously shown that brain network connectivity is highly predictive of glioma survival. In this study, we aimed to examine the molecular mechanisms of this relationship via imaging transcriptomics. Methods We retrospectively obtained presurgical, T1-weighted MRI datasets from 669 adult patients, newly diagnosed with diffuse glioma. We measured brain connectivity using gray matter networks and coregistered these data with a transcriptomic brain atlas to determine the spatial co-localization between brain connectivity and expression patterns for 14 proto-oncogenes and 3 neural network construction genes. Results We found that all 17 genes were significantly co-localized with brain connectivity (p < 0.03, corrected). The strength of co-localization was highly predictive of overall survival in a cross-validated Cox Proportional Hazards model (mean area under the curve, AUC = 0.68 +/− 0.01) and significantly (p < 0.001) more so for a random forest survival model (mean AUC = 0.97 +/− 0.06). Bayesian network analysis demonstrated direct and indirect causal relationships among gene-brain co-localizations and survival. Gene ontology analysis showed that metabolic processes were overexpressed when spatial co-localization between brain connectivity and gene transcription was highest (p < 0.001). Drug-gene interaction analysis identified 84 potential candidate therapies based on our findings. Discussion Our findings provide novel insights regarding how gene-brain connectivity interactions may affect glioma survival.
... This study showed a correlation between hemispheric dominance and the nasal cycle (Shannahoff-Khalsa, 1993). Cerebral hemispheres exhibit functional and structural asymmetry in terms of performance based on spatial processing and logical processing, which might be due to handedness differences (Goel, 2019;Sun & Walsh, 2006). Previous studies showed a correlation between nasal airflow and handedness (Searleman et al., 2005). ...
Article
Full-text available
Nasal cycle (NC) is a rhythmic change of lateralised nasal airflow mediated by the autonomous nervous system. Previous studies reported the dependence of NC dominance or more patent side on handedness and hemispheric cerebral activity. We aimed to investigate firstly the possible lateralised effect of NC on olfactory bulb volume and secondly the association of NC with the lateralised cerebral dominance in terms of olfactory processing. Thirty‐five subjects (22 women and 13 men, mean age 26 ± 3 years) participated in the study. NC was ascertained using a portable rhino‐flowmeter. Structural and functional brain measurements were assessed using a 3T MR scanner. Vanillin odorant was presented during functional scans using a computer‐controlled olfactometer. NC was found to be independent of the olfactory bulb volumes. Also, cerebral activations were found independent of the NC during odorant perception. NC potency is not associated with lateralised structural or functional differences in the cerebral olfactory system.
... The strongest adjusted relationships for diffusion features were found in the cingulate gyrus tract (β males BRIA−microRD = 0. 25 Supplementary Figs. 8,9). Strongest age associations with T 1 -weighted asymmetries were found for the area of the accumbens (β males = 0.14, β females =0.12) and WM surface (β males =0.13, β females = 0.12), with strongest inverse relationships observed for inferior lateral ventricles (β males = − 0.17, β females = − 0.14) and pallidum (β males = −0.11, ...
Article
Full-text available
The human brain demonstrates structural and functional asymmetries which have implications for ageing and mental and neurological disease development. We used a set of magnetic resonance imaging (MRI) metrics derived from structural and diffusion MRI data in N=48,040 UK Biobank participants to evaluate age-related differences in brain asymmetry. Most regional grey and white matter metrics presented asymmetry, which were higher later in life. Informed by these results, we conducted hemispheric brain age (HBA) predictions from left/right multimodal MRI metrics. HBA was concordant to conventional brain age predictions, using metrics from both hemispheres, but offers a supplemental general marker of brain asymmetry when setting left/right HBA into relationship with each other. In contrast to WM brain asymmetries, left/right discrepancies in HBA are lower at higher ages. Our findings outline various sex-specific differences, particularly important for brain age estimates, and the value of further investigating the role of brain asymmetries in brain ageing and disease development.
... Additionally, a minimum score of 85 in their English listening and speaking classes was required to demonstrate a strong proficiency in English conversation tasks. All participants were right-handed (Sun & Walsh, 2006) with normal or corrected-to-normal vision and hearing. All participants were required to sign informed consent documents prior to the experimental tests and received financial compensation after the experiment. ...
Article
Full-text available
Handedness develops early in life, but the structural and functional brain connectivity patterns associated with it remains unknown. Here we investigate associations between handedness and the asymmetry of brain connectivity in 9- to 10-years old children from the Adolescent Brain Cognitive Development (ABCD) study. Compared to right-handers, left-handers had increased global functional connectivity density in the left-hand motor area and decreased it in the right-hand motor area. A connectivity-based index of handedness provided a sharper differentiation between right- and left-handers. The laterality of hand-motor connectivity varied as a function of handedness in unimodal sensorimotor cortices, heteromodal areas, and cerebellum (P < 0.001) and reproduced across all regions of interest in Discovery and Replication subsamples. Here we show a strong association between handedness and the laterality of the functional connectivity patterns in the absence of differences in structural connectivity, brain morphometrics, and cortical myelin between left, right, and mixed handed children.
Article
Abnormal hemispheric specialization and inter-hemispheric interactions may contribute to the pathogenesis of general anxiety disorder (GAD). The current study investigated these abnormalities in GAD patients based on the two analytic approaches and examined whether such abnormalities are correlated with anxiety symptom severity. Seventy-three patients with GAD and 60 matched healthy controls were recruited. All participants completed anxiety symptoms assessment and resting-state functional magnetic resonance imaging (rs-fMRI). The autonomy index (AI) and Connectivity between Functionally Homotopic voxels (CFH) were applied to measure and compared between groups. Compared to controls, patients showed stronger AI in the right middle temporal gyrus (MTG). Seed-based analysis revealed stronger functional connectivity (FC) of the right MTG with both right precuneus and right dorsolateral prefrontal cortex (dlPFC) in patients. Patients also exhibited greater CFH in right anterior cingulate cortex (ACC) but decreased CFH in bilateral postcentral gyrus (PCG) and superior occipital gyrus (SOG). Further there were significant correlations between these regional CFH and anxiety symptoms severity. GAD patients demonstrate right hemispheric specialization and aberrant inter-hemispheric functional cooperation, and abnormal inter-hemispheric coordination is associated with anxiety symptom severity. These findings provide a clue to understanding the neuropathological mechanisms of GAD.
Article
A brief explanation of embryonic staging is provided, and the system now in use for the Carnegie Collection is outlined. The criteria for the first stages, which are not yet published in extenso, are listed, and a summarizing diagram with 2 graphs included.
Article
The development of functional brain asymmetry during childhood is confirmed by changes in cerebral blood flow measured at rest using dynamic single photon emission computed tomography. Between 1 and 3 years of age, the blood flow shows a right hemispheric predominance, mainly due to the activity in the posterior associative area. Asymmetry shifts to the left after 3 years. The subsequent time course of changes appear to follow the emergence of functions localized initially on the right, but later on the left hemisphere (i.e. visuospatial and later language abilities). These findings support the hypothesis that, in man, the right hemisphere develops its functions earlier than the left.
Article
Background: Magnetic resonance imaging studies in schizophrenia have revealed abnormalities in temporal lobe structures, including the superior temporal gyrus. More specifically, abnormalities have been reported in the posterior superior temporal gyrus, which includes the Heschl gyrus and planum temporale, the latter being an important substrate for language. However, the specificity of the Heschl gyrus and planum temporale structural abnormalities to schizophrenia vs affective psychosis, and the possible confounding roles of chronic morbidity and neuroleptic treatment, remain unclear. Methods: Magnetic resonance images were acquired using a 1.5-T magnet from 20 first-episode (at first hospitalization) patients with schizophrenia (mean age, 27.3 years), 24 first-episode patients with manic psychosis (mean age, 23.6 years), and 22 controls (mean age, 24.5 years). There was no significant difference in age for the 3 groups. All brain images were uniformly aligned and then reformatted and resampled to yield isotropic voxels. Results: Gray matter volume of the left planum temporale differed among the 3 groups. The patients with schizophrenia had significantly smaller left planum temporale volume than controls (20.0%) and patients with mania (20.0%). Heschl gyrus gray matter volume (left and right) was also reduced in patients with schizophrenia compared with controls (13.1%) and patients with bipolar mania (16.8%). Conclusions: Compared with controls and patients with bipolar manic psychosis, patients with first-episode schizophrenia showed left planum temporale gray matter volume reduction and bilateral Heschl gyrus gray matter volume reduction. These findings are similar to those reported in patients with chronic schizophrenia and suggest that such abnormalities are present at first episode and are specific to schizophrenia.
Article
The anatomic pattern and left hemisphere size predominance of the planum temporale, a language area of the human brain, are also present in chimpanzees (Pan troglodytes). The left planum temporale was significantly larger in 94 percent (17 of 18) of chimpanzee brains examined. It is widely accepted that the planum temporale is a key component of Wernicke's receptive language area, which is also implicated in human communication-related disorders such as schizophrenia and in normal variations such as musical talent. However, anatomic hemispheric asymmetry of this cerebrocortical site is clearly not unique to humans, as is currently thought. The evolutionary origin of human language may have been founded on this basal anatomic substrate, which was already lateralized to the left hemisphere in the common ancestor of chimpanzees and humans 8 million years ago.