ArticlePDF AvailableLiterature Review

Lissencephaly and the molecular basis of neural migration

Authors:

Abstract and Figures

Migration of post-mitotic neurons from the ventricular zone to the cortical plate during embryogenesis comprises one of the most critical stages in brain development. Deficiency of this process often results in major brain malformations, including human lissencephaly (smooth brain). Since discovery of the first genetic cause of lissencephaly, deletions of chromosome 17p13.3 in Miller-Dieker syndrome, rapid progress in our understanding of neuronal migration has been made based on advances in both brain imaging technology and molecular genetics. This progress has resulted in a new system of classification that began with pathological descriptions and has evolved to include patterns on brain imaging, causative genes and most recently the molecular pathways and proposed modes of migration involved. In this review, we summarize current knowledge regarding five genes that cause or contribute to human lissencephaly, including LIS1, 14-3-3 epsilon, DCX, RELN and ARX. Each of these is associated with a characteristic pattern of malformation that involves the cerebral cortex and sometimes other brain structures. Based on detailed genotype-phenotype analysis, we can now infer the most likely causative gene based on brain imaging and other clinical findings, and inversely are becoming able to predict clinical severity based on the specific mutations detected. We also hypothesize, for the first time, a relationship between the specific type of lissencephaly observed and deficiency of specific modes of neuronal migration.
Axial T 1 - ( B , C , E , H ) and T 2 - ( A , D , F , G ) weighted magnetic resonance images at the level of basal ganglia in fi ve types of lissencephaly. In contrast to a normal control (H), all types of LIS have broad or absent gyri and abnormally thick cortex, except for LIS grade 6 or SBH, in which the sulci separating gyri are very shallow. The anterior to posterior or rostro-caudal gradient of LIS is strictly correlated with the causative gene. Speci fi cally, mutations of DCX or RELN result in an anterior more severe than posterior (a > p) gradient (A – D), while mutations of LIS1 with or without 14-3-3 e or ARX lead to a posterior more severe than anterior (p > a) gradient (E – G). The absolute thickness of the cortex and presence of a cell sparse zone also differ based on the causative gene. In patients with mutations of DCX (A, B) or LIS1 (E, F), the cortex is very thick, typically 10 – 20 mm, and prominent cell sparse zones are seen in areas of agyria (arrowheads in A, E, F). In patients with LIS with cerebellar hypoplasia group b [C, which resembles patients with known RELN mutations (67,94), although a mutation has not been demonstrated in this patient] or ARX (G) mutations, the cortex is only moderately thick, typically 5 – 10 mm, and cell sparse zones are never seen, even in areas of agyria (G). In LIS with cerebellar hypoplasia group b with or without proven RELN mutations, other images show an abnormal hippocampus and severe cerebellar hypoplasia (not shown). In males with X-linked lissencephaly with abnormal genitalia due to ARX mutations, other images demonstrate poorly demarcated basal ganglia often with small cysts, immature white matter, and agenesis of the corpus callosum (not shown). In heterozygous females, mutations of DCX result in SBH (D), while mutations of ARX often result in agenesis of the corpus callosum (not shown). ILS, isolated lissencephaly; LCHb, lissencephaly with cerebellar hypoplasia group b; MDS, Miller – Dieker syndrome; XLAG, X-linked lissencephaly with abnormal genitalia.
… 
Content may be subject to copyright.
Lissencephaly and the molecular basis of
neuronal migration
Mitsuhiro Kato
1,4
and William B. Dobyns
1– 3,
*
1
Department of Human Genetics,
2
Department of Neurology and
3
Department of Pediatrics,
The University of Chicago, Chicago, IL 60637, USA and
4
Department of Pediatrics,
Yamagata University School of Medicine, Yamagata, Japan
Received January 17, 2003; Revised and Accepted February 5, 2003
Migration of post-mitotic neurons from the ventricular zone to the cortical plate during embryogenesis
comprises one of the most critical stages in brain development. Deficiency of this process often results in
major brain malformations, including human lissencephaly (smooth brain). Since discovery of the first
genetic cause of lissencephaly, deletions of chromosome 17p13.3 in Miller–Dieker syndrome, rapid progress
in our understanding of neuronal migration has been made based on advances in both brain imaging
technology and molecular genetics. This progress has resulted in a new system of classification that began
with pathological descriptions and has evolved to include patterns on brain imaging, causative genes and
most recently the molecular pathways and proposed modes of migration involved. In this review, we
summarize current knowledge regarding five genes that cause or contribute to human lissencephaly,
including LIS1, 14-3-3e, DCX, RELN and ARX. Each of these is associated with a characteristic pattern of
malformation that involves the cerebral cortex and sometimes other brain structures. Based on detailed
genotype–phenotype analysis, we can now infer the most likely causative gene based on brain imaging and
other clinical findings, and inversely are becoming able to predict clinical severity based on the specific
mutations detected. We also hypothesize, for the first time, a relationship between the specific type of
lissencephaly observed and deficiency of specific modes of neuronal migration.
INTRODUCTION
Migration of post-mitotic neurons from the ventricular zone to
form the cortical plate comprises one of the most critical stages
in brain development. Our understanding of this complex
process has progressed based on studies of human malformations
and mouse mutants with deficient neuronal migration, particu-
larly the malformation known as lissencephaly (LIS) or ‘smooth-
brain’. LIS is characterized by a smooth or nearly smooth
cerebral surface. It encompasses a spectrum of gyral malforma-
tions from complete agyria (absent gyri) to regional pachygyria
(broad gyri), and merges with subcortical band heterotopia
(SBH). LIS is always associated with an abnormally thick cortex,
reduced or abnormal lamination and diffuse neuronal heterotopia
(1,2). SBH or ‘double cortex’ consists of circumferential bands
of heterotopic neurons located just beneath the cortex and
separated from it by a thin band of white matter (3,4).
Several different types of LIS have been recognized. The
most common type, known as classical LIS (previously type I),
has a very thick 10–20 mm cortex and no other major brain
malformations. This is the only type associated with SBH. Less
common types are associated with agenesis of the corpus
callosum (ACC) or severe cerebellar hypoplasia (5,6).
In this paper, we review recent discoveries regarding the
molecular mechanisms that regulate or effect neuronal
migration to the cerebral cortex, emphasizing those derived
from studies in humans with LIS. Other genes and proteins
identified in humans with periventricular nodular heterotopia
or cobblestone complex malformations (previously type
II LIS), or in mouse mutants have been reviewed elsewhere
(7–11).
HUMAN LISSENCEPHALY GENES
To date, five genes have been identified that cause or contribute
to LIS in humans: LIS1, 14-3-3e, DCX, RELN and ARX. Their
respective subtypes and syndromes are shown in Figures 1
and 2, and can usually be distinguished based on detailed
analysis of the phenotype as shown in Figure 3.
*To whom correspondence should be addressed at: The University of Chicago, Department of Human Genetics, Room 319 CLSC, 920 E. 58th Street,
Chicago, IL 60637, USA. Tel: þ 1 7738343597; Fax: þ1 7738348470; Email: wbd@genetics.bsd.uchicago.edu
Human Molecular Genetics, 2003, Vol. 12, Review Issue 1 R89–R96
DOI: 10.1093/hmg/ddg086
Human Molecular Genetics, Vol. 12, Review Issue 1 # Oxford University Press 2003; all rights reserved
LIS1
LIS1 or PAFAH1B1 was the rst human neuronal migration
gene to be cloned, and encodes the non-catalytic alpha subunit
of the intracellular Ib isoform of platelet-activating factor
acetylhydrolase (12,13). The gene is located in human
chromosome 17p13.3 and consists of 11 exons with a coding
region of 1233 bp. LIS1 protein is expressed predominantly in
fetal and adult brain (14), and interacts with tubulin to suppress
microtubule dynamics (15). It is a highly conserved protein
with near-identity between mouse and human, and 42%
homology to NudF, an ortholog found in Aspergillus
nidulans (16). Studies of NudF and related genes such as
NudE have shown that LIS1 participates in cytoplasmic
dynein-mediated nucleokinesis, somal translocation and cell
motility (17) as well as mitosis (neurogenesis) and chromo-
some segregation (18). In animal experiments including human
cells, Lis1 acts via a signaling pathway that includes NudE,
NudeL, cytoplasmic dynein, dynactin, and CLIP-170 (1926).
14-3-3e
This gene belongs to the 14-3-3 family of proteins that bind
to phosphoserine and phosphothreonine motifs in a wide
variety of proteins (2731). It is also located in chromosome
Figure 1. Axial T
1
-(B, C, E, H) and T
2
-(A, D, F, G) weighted magnetic resonance images at the level of basal ganglia in ve types of lissencephaly. In contrast to
a normal control (H), all types of LIS have broad or absent gyri and abnormally thick cortex, except for LIS grade 6 or SBH, in which the sulci separating gyri are
very shallow. The anterior to posterior or rostro-caudal gradient of LIS is strictly correlated with the causative gene. Specically, mutations of DCX or RELN result
in an anterior more severe than posterior (a > p) gradient (AD), while mutations of LIS1 with or without 14-3-3e or ARX lead to a posterior more severe than
anterior (p > a) gradient (EG). The absolute thickness of the cortex and presence of a cell sparse zone also differ based on the causative gene. In patients with
mutations of DCX (A, B) or LIS1 (E, F), the cortex is very thick, typically 1020 mm, and prominent cell sparse zones are seen in areas of agyria (arrowheads in A,
E, F). In patients with LIS with cerebellar hypoplasia group b [C, which resembles patients with known RELN mutations (67,94), although a mutation has not been
demonstrated in this patient] or ARX (G) mutations, the cortex is only moderately thick, typically 510 mm, and cell sparse zones are never seen, even in areas of
agyria (G). In LIS with cerebellar hypoplasia group b with or without proven RELN mutations, other images show an abnormal hippocampus and severe cerebellar
hypoplasia (not shown). In males with X-linked lissencephaly with abnormal genitalia due to ARX mutations, other images demonstrate poorly demarcated basal
ganglia often with small cysts, immature white matter, and agenesis of the corpus callosum (not shown). In heterozygous females, mutations of DCX result in SBH
(D), while mutations of ARX often result in agenesis of the corpus callosum (not shown). ILS, isolated lissencephaly; LCHb, lissencephaly with cerebellar hypo-
plasia group b; MDS, MillerDieker syndrome; XLAG, X-linked lissencephaly with abnormal genitalia.
R90 Human Molecular Genetics, 2003, Vol. 12, Review Issue 1
17p13.3, about 40 kb telomeric to LIS1, and has six exons
that encode 255 amino acids. The gene codes for a highly
conserved protein that binds to and protects phosphorylated
NUDEL from dephosphorylation by protein phosphatase
2A (PP2A). It is required for NUDEL localization and
cytoplasmic dynein function, and appears to be important
for neuronal migration based on mouse studies (Tokyo-oka
et al., submitted).
DCX
DCX or XLIS was rst identied as the gene causing X-linked
LIS and SBH (32,33). It is located in chromosome Xq22.3q23
and has nine exons (six coding exons) that code for a 360
amino acid protein. The DCX protein contains two tandem
evolutionarily conserved repeats (doublecortin domain) that
form a b-grasp superfold. Each of the repeat binds to tubulin
but not to microtubules, so that both repeats are necessary for
microtubule polymerization and stabilization (3437), and pos-
sibly for direct interactions with LIS1 (38). DCX is expressed
exclusively in fetal brain, including forebrain and cerebellum.
RELN
Reln was cloned as the causative gene for the reeler mutant
mouse, which has abnormal lamination of the cerebral and
cerebellar cortices including inversion of the normal inside-
out pattern found in mammals (39,40). The human gene is
located in chromosome 7q22 and consists of 65 exons covering
more than 400 kb of genomic sequence. It encodes a large
extracellular matrix protein with 3460 amino acids that is
secreted by CajalRetzius cells in the preplate (41) and has
94.2% homology with the mouse ortholog (42). Reln functions
in a signal transduction pathway via the apolipoprotein E2
Figure 2. Correlation of LIS grade and gradient and associated brain malformations with causative genes. LIS grades 1 and 2 consist of diffuse agyria, although
patients with LIS grade 2 have a few shallow sulci over the frontal or occipital pole. LIS grade 3 comprises mixed agyria and pachygyria, grade 4 pachygyria only,
grade 5 mixed pachygyria and SBH, and grade 6 SBH only. The most severe form of LIS (grade 1) is typically caused by either combined LIS1 and 14-3-3e
deletion or a severe mutation of DCX. Mutations of LIS1 alone typically cause LIS grades 24, most often grade 3, with very rare patients having posterior
SBH. Mutations of DCX cause a wide range of LIS, although grade 3 has proven to be rare. Mutations of RELN cause both cerebral (LIS) and cerebellar mal-
formations, which resembles the defects found in the reeler mouse. Mutations of ARX most often result in LIS grade 3, in addition to the basal ganglia, white
matter and callosal abnormalities. At least half of all female carriers of ARX mutations that cause XLAG in males have ACC. Black box, most frequently observed;
dark grey box, sometimes observed; light grey box, rarely observed. CBLH, cerebellar hypoplasia. The asterisk indicates it may result from somatic mosaicism.
Human Molecular Genetics, 2003, Vol. 12, Review Issue 1 R91
(ApoER2) and very low-density lipoprotein receptors (Vldlr)
(43,44), which activate the downstream cytoplasmic protein
Dab1 (45,46). The brains of mutant mice strains with
disruptions of mDab1 or of both ApoER1 and Vldlr closely
resemble the brain of the reeler mouse (44,47).
ARX
ARX is a paired-class homeobox gene that shows signicant
homology with the Drosophila al (aristaless) gene in the
homeodomain and C-peptide or aristaless domain (48). The
gene is located in human chromosome Xp22.13 and consists of
ve exons that encode a protein of 562 amino acids (49,50).
ARX is specically expressed in interneurons of the forebrain
and in the interstitium of the male gonad (48,50). It is involved
in differentiation of the testes and the embryonic forebrain,
especially in proliferation of neural precursors and differentia-
tion and tangential migration of interneurons (50).
GENOTYPE–PHENOTYPE CORRELATION
WITH MUTATIONS OF KNOWN
LISSENCEPHALY GENES
LIS1
The most important characteristics of LIS in patients with LIS1
mutations are the very thick 1020 mm cortex, gyral mal-
formations that are more severe in posterior than anterior (p > a
gradient) brain regions (51,52) and prominent cell sparse zone
in the cortex. The corpus callosum and cerebellum appear
normal or mildly hypoplastic on brain MRI (6,52).
Mutations of LIS1 cause isolated lissencephaly sequence
(ILS) or rarely isolated SBH, both with a clear p > a gradient.
We found a signicant correlation between LISSBH severity
and mutation type in these disorders. Patients with submicro-
scopic deletions of 17p13.3 that include LIS1 and those with
null mutations in the coiled coil (MAP1B homology) domain
in exons 25 usually had LIS grade 23, with a mean grade of
2.43 for the latter group (52,53). Patients with null mutations
distal to the coiled coil domain usually had LIS grades 34,
with a mean grade of 3.18. Patients with missense mutations of
LIS1 had even less severe LIS unless a critical amino acid
residue was involved, with a mean grade of 4.20 (5355).
All but one of the missense mutations are associated with
milder phenotypes than patients with null mutations, which
suggests some residual LIS1 protein function. However, in vitro
assays show that all mutant proteins completely lose the
capacity to interact with NUDE as well as the 29 and 30 kDa
subunits of platelet activating factor acetylhydrolase,
PAFAH1B2 and PAFAH1B3 (22,56). Thus, the basis for
differing severity among patients with missense mutations is
unclear. The mildest LIS1-associated phenotype known con-
sists of infrequent seizures, mild clumsiness and normal
intelligence with pachygyria limited to the posterior parietal
and occipital lobes. The affected boy had a missense mutation
in the second WD repeat resulting in substitution of serine for
glycine (G162S). This is predicted to be a mild mutation, as
serine is found at the same position in other WD family
proteins (55,57).
Figure 3. Flow chart to predict the most likely causative gene in patients with LIS by analysis of brain MRI.
R92 Human Molecular Genetics, 2003, Vol. 12, Review Issue 1
LIS1 and 14-3-3e
The association between LIS and deletions of 17p13.3 was rst
recognized in patients with MillerDieker syndrome (MDS),
which consists of LIS and facial abnormalities including
prominent forehead, bitemporal hollowing, short nose with
upturned nares, prominent upper lip with downturned vermi-
lion border and small jaw, and sometimes other congenital
anomalies (58,59). Our previous studies suggested that patients
with MDS have more severe LIS than patients with ILS, and
that deletions in MDS extend further toward the 17p telomere
than in ILS, suggesting that another gene involved in brain
development is located distal to LIS1 (52,60).
We recently completed a contig across most of this 400 kb
region, and studied 30 patients with MDS or ILS and deletions of
the region with a set of FISH probes and somatic cell hybrids
(61). MDS was always associated with LIS grade 1 (essentially
complete agyria), and the deletion in MDS extends from LIS1 to
include all or part of BAC RPCI11-818O24, which contains the
CRK and 14-3-3e genes. The 14-3-3e gene is the best candidate
to account for the more severe LIS phenotype in MDS, as the
mouse knockout has mild defects of neuronal migration, although
a Crk knockout has not yet been reported. As would be predicted
from our results in humans, mice compound heterozygous for
14-3-3e and Lis1 display more severe brain defects than either
heterozygous mutant mouse alone (Toyo-oka et al.,submitted).
DCX
The most important characteristics of LIS in patients with
DCX mutations are the very thick 1020 mm cortex, gyral
malformations that are more severe in anterior than posterior
(a > p gradient) brain regions, prominent cell sparse zone and
the occurrence of SBH rather than LIS in heterozygous females
(3,51,52). Hypoplasia of the cerebellar vermis is sometimes
associated with LIS caused by either LIS1 or DCX mutation,
classied as LIS with cerebellar hypoplasia group a (6,52).
In contrast to LIS1, missense mutations of DCX are more
common than truncations. However, genotypephenotype
correlation is complex. Our experience has shown that the
same (presumably severe) mutations that cause LIS grade 1 or
agyria in males cause diffuse thick SBH in females. Similarly,
the same (presumably mild) mutations that cause LIS grade 4
or frontal pachygyria in males cause diffuse thin or partial
frontal SBH in females. A few missense mutations have led to
partial frontal SBH in males and either very mild or normal
phenotypes in their carrier mothers, who had germline
mutations and random X inactivation (54,62).
Despite substantial experience, predicting the phenotype
based on the genotype remains difcult due to incomplete
knowledge regarding function of specic amino acids within
the two microtubule-binding or doublecortin domains in exons
46, and relatively frequent post-zygotic mosaicism. Our recent
data have shown that truncation mutations in all but the last two
exons (exons 8 and 9) cause a severe phenotype, while
truncation mutations in exon 9 cause a variable phenotype
suggesting nonsense-mediated mRNA decay in DCX tran-
scripts. No truncations in exon 8 have been reported. Missense
mutations of DCX cluster in the doublecortin domains in exons
46, where they cause a variable phenotype (36,37,51,63).
Missense mutations in exons 79 have never been observed,
and so presumably cause a very mild or no phenotype (63).
Finally, post-zygotic mosaicism appears to be an important
mechanism in both sexes. In males, post-zygotic mosaicism
ameliorates the phenotype, resulting in SBH or mild LIS
(grades 5 and 6) rather than severe LIS (grade 1 or 2) (6466).
In females, post-zygotic mosaicism has been found in mothers
of several affected patients (64). Owing to the difculty in
detecting mosaicism in heterozygous females, a high index of
suspicion is required, and must be taken into account when
providing genetic counseling.
RELN
Mutations of RELN have been reported in six children with a
LIS variant from two unrelated families (67). Both families
showed exon skipping resulting in undetectable or reduced
levels of RELN protein. The same pattern of LIS was described
in four patients from two unrelated Japanese families, one of
whom had reduced levels of RELN protein in serum (6,68).
These patients have been classied as LIS with cerebellar
hypoplasia group b (6). In these children, the malformation is
characterized by a moderately thick 510 mm cortex, LIS that
appears more severe in anterior than posterior (a > p gradient)
brain regions, malformed hippocampus and very small
cerebellum virtually lacking folia.
ARX
Mutations of ARX cause a wide range of phenotypes that
correlate closely with the type of mutation. Hemizygous males
with null and non-conservative missense mutations have
a well-delineated syndrome known as X-linked lissencephaly
with abnormal genitalia (XLAG) (5). The most important
characteristics of LIS in patients with XLAG are a moderately
thick 510 mm cortex, gyral malformations that are more
severe in posterior than anterior (p > a gradient) brain regions,
deciency of small granular neurons throughout the cerebral
cortex, abnormal signal of white matter, ACC, and cystic or
fragmented basal ganglia (50,69) (Kato et al., submitted). Rare
patients with null mutations or missense mutation in the
homeobox have a variant with hydranencephaly or isolated
ACC, both with abnormal genitalia (Kato et al., submitted).
Missense or in frame expansion mutations of ARX cause
familial X-linked infantile spasms or West syndrome, sporadic
cryptogenic or non-symptomatic West syndrome, X-linked
myoclonic epilepsy with spasticity and mental retardation,
Partington syndrome (mental retardation and dystonia), non-
syndromic X-linked mental retardation, and possibly autism
(49,7074) (Kato et al., submitted). We also found several
females with symptomatic or asymptomatic ACC, who were
carriers of the same ARX mutations that caused XLAG in
males. Thus, mutations of ARX are clearly associated with
remarkable pleiotropy that includes seemingly disparate
phenotypes both with and without malformations.
LIS AND MECHANISMS OF MIGRATION
For many years, neuronal migration along radial glial bers has
been the most widely accepted mechanism of cortical
Human Molecular Genetics, 2003, Vol. 12, Review Issue 1 R93
formation, and disturbance of neuronal migration the cause of
LIS (2,75,76). However, recent studies have demonstrated three
different modes of migration, including two forms of radial
migration (somal translocation and glia-guided locomotion), as
well as tangential or non-radial migration (7779).
Somal translocation consists of movement of the soma and
nucleus toward the cortical plate in a long, radially oriented
basal process of the cell that terminates at the pial surface
(79,80). Glia-guided migration consists of slower movement
along the scaffold of radial glial bers. In general, early-
generated cells such as preplate neurons use somal trans-
location only. Later-migrating pyramidal neurons rst use
glia-guided locomotion and subsequently somal translocation
as they move past earlier-generated neurons to form the inside-
out pattern of the mammalian neocortex (80,81). Tangential
migration is used by GABAergic neurons to migrate from the
ventral to the dorsal telencephalon along corticofugal bers. On
reaching the dorsal telencephalon, the cells change direction to
migrate into the cortex along either corticofugal projection
bers or radial glial bers (50,8284). Some of these cells
briey migrate toward the ventricular zone, before reversing
direction to migrate into the neocortex (85).
From the limited data available so far, it appears likely that
the three known modes of migration are affected in different
ways in the various LIS subtypes. In the cerebral cortex of
reeler mice (Reln deciency), the preplate appears to form
normally, but later-migrating neurons fail to split the preplate
into marginal zone and subplate. Instead, they form layers
below the preplate with a reverse outside-inpattern compared
with normal cortex (39,86,87). Reln is also required for the
formation of the radial glial scaffold in the hippocampus (88).
These observations suggest a role for Reln, and presumably
human RELN, in glia-guided locomotion.
As mentioned above, LIS1 interacts with tubulin and
participates in nucleokinesis and extension of the leading
process as well as mitosis (10). In compound heterozygous Lis1
mutant mice, splitting of the preplate is defective, leaving a
broad and poorly dened subplate, and cortical lamination is
completely disrupted as in humans with LIS (89). In vitro
studies of Lis1-decient cerebellar granule cells demonstrate
decient migration along neurites of other cells in culture (90).
While available data is less clear than for Reln, the abnormality
of the preplate and the severe, and thus most likely early-onset,
cortical disruption seem to implicate a defect of somal
translocation. The defective migration along neurites suggests
that migration along radial glia may also be decient. Based on
these results, we hypothesize that both somal translocation and
glia-guided migration are disrupted by mutations of LIS1.
Studies of tangential migration have not yet been reported. The
remarkably thick cortex also supports a more severe defect of
migration with LIS1 compared to RELN mutations.
DCX is a microtubule-associated protein that may interact
with LIS1, and mutations cause severe LIS, similar to that
observed with LIS1 mutations (34,35,91). However, its function
remains poorly understood. Expression of DCX in both radial
columns and tangentially directed neurons of human fetal
cortex suggests that it may be involved in both radial and
tangential migration (92).
Arx is expressed in the ganglionic eminences and the
neocortical ventricular zone, so both radial and tangential
migration could be affected. However, the data available to date
implicate primarily tangential migration. In Arx mutant mice,
mutant GABAergic interneurons originating from the gang-
lionic eminence remain near the subplate for at least 3 days,
and then migrate to the cortex where they are aberrantly
scattered throughout the cortical plate (50). In humans with
ARX mutations, the cortex contains almost exclusively
pyramidal neurons (69), which also implicates tangential
migration. The different pathological ndings between humans
and mice may reect differences in the origin of GABAergic
interneurons (93). We therefore hypothesize that ARX
deciency results in defects primarily of tangential migration.
REFERENCES
1. Crome, L. (1956) Pachygyria. J. Pathol. Bacteriol., 71, 335352.
2. Friede, R.L. (1989) Developmental Neuropathology, 2nd edn. Springer,
Berlin.
3. Dobyns, W.B., Andermann, E., Andermann, F., Czapansky-Beilman, D.,
Dubeau, F., Dulac, O., Guerrini, R., Hirsch, B., Ledbetter, D.H., Lee, N.S.
et al. (1996) X-linked malformations of neuronal migration. Neurology, 47,
331339.
4. Barkovich, A.J., Guerrini, R., Battaglia, G., Kalifa, G., NGuyen, T.,
Parmeggiani, A., Santucci, M., Giovanardi-Rossi, P., Granata, T. and
DIncerti, L. (1994) Band heterotopia: correlation of outcome with
magnetic resonance imaging parameters. Ann. Neurol., 36, 609617.
5. Dobyns, W.B., Berry-Kravis, E., Havernick, N.J., Holden, K.R. and
Viskochil, D. (1999) X-linked lissencephaly with absent corpus callosum
and ambiguous genitalia. Am. J. Med. Genet., 86, 331337.
6. Ross, M.E., Swanson, K. and Dobyns, W.B. (2001) Lissencephaly with
cerebellar hypoplasia (LCH): a heterogeneous group of cortical
malformations. Neuropediatrics, 32, 256263.
7. Sheen, V.L., Dixon, P.H., Fox, J.W., Hong, S.E., Kinton, L., Sisodiya, S.M.,
Duncan, J.S., Dubeau, F., Scheffer, I.E., Schachter, S.C. et al. (2001)
Mutations in the X-linked lamin 1 gene cause periventricular nodular
heterotopia in males as well as in females. Hum. Mol. Genet., 10,
17751783.
8. Ross, M.E. and Walsh, C.A. (2001) Human brain malformations and
their lessons for neuronal migration. A. Rev. Neurosci., 24, 10411070.
9. Herz, J. and Bock, H.H. (2002) Lipoprotein receptors in the nervous
system. A. Rev. Biochem., 71, 405434.
10. Gupta, A., Tsai, L.H. and Wynshaw-Boris, A. (2002) Life is a journey: a
genetic look at neocortical development. Nat. Rev. Genet., 3, 342355.
11. Ross, M.E. (2002) Full circle to cobbled brain. Nature, 418, 376377.
12. Reiner, O., Carrozzo, R., Shen, Y., Wehnert, M., Faustinella, F.,
Dobyns, W.B., Caskey, C.T. and Ledbetter, D.H. (1993) Isolation of a
MillerDieker lissencephaly gene containing G protein beta-subunit-like
repeats. Nature, 364, 717721.
13. Hattori, M., Adachi, H., Tsujimoto, M., Arai, H. and Inoue, K. (1994)
MillerDieker lissencephaly gene encodes a subunit of brain
platelet-activating factor acetylhydrolase. Nature, 370, 216218.
14. Mizuguchi, M., Takashima, S., Kakita, A., Yamada, M. and Ikeda, K.
(1995) Lissencephaly gene product. Localization in the central nervous
system and loss of immunoreactivity in MillerDieker syndrome.
Am. J. Pathol., 147, 11421151.
15. Sapir, T., Elbaum, M. and Reiner, O. (1997) Reduction of microtubule
catastrophe events by LIS1, platelet-activating factor acetylhydrolase
subunit. EMBO J., 16, 69776984.
16. Xiang, X., Osmani, A.H., Osmani, S.A., Xin, M. and Morris, N.R. (1995)
NudF, a nuclear migration gene in Aspergillus nidulans, is similar to the
human LIS-1 gene required for neuronal migration. Mol. Biol. Cell, 6,
297310.
17. Smith, D.S., Niethammer, M., Ayala, R., Zhou, Y., Gambello, M.J.,
Wynshaw-Boris, A. and Tsai, L.H. (2000) Regulation of cytoplasmic
dynein behaviour and microtubule organization by mammalian Lis1. Nat.
Cell Biol., 2, 767775.
18. Faulkner, N.E., Dujardin, D.L., Tai, C.Y., Vaughan, K.T., OConnell, C.B.,
Wang, Y. and Vallee, R.B. (2000) A role for the lissencephaly gene LIS1 in
mitosis and cytoplasmic dynein function. Nat. Cell Biol., 2, 784791.
R94 Human Molecular Genetics, 2003, Vol. 12, Review Issue 1
19. Sasaki, S., Shionoya, A., Ishida, M., Gambello, M.J., Yingling, J.,
Wynshaw-Boris, A. and Hirotsune, S. (2000) A LIS1/NUDEL/cytoplasmic
dynein heavy chain complex in the developing and adult nervous system.
Neuron, 28, 681696.
20. Sweeney, K.J., Prokscha, A. and Eichele, G. (2001) NudE-L, a novel
Lis1-interacting protein, belongs to a family of vertebrate coiled-coil
proteins. Mech. Dev., 101,2133.
21. Hoffmann, B., Zuo, W., Liu, A. and Morris, N.R. (2001) The LIS1-related
protein NUDF of Aspergillus nidulans and its interaction partner NUDE
bind directly to specic subunits of dynein and dynactin and to alpha- and
gamma-tubulin. J. Biol. Chem., 276, 3887738884.
22. Feng, Y., Olson, E.C., Stukenberg, P.T., Flanagan, L.A., Kirschner, M.W.
and Walsh, C.A. (2000) LIS1 regulates CNS lamination by interacting with
mNudE, a central component of the centrosome. Neuron, 28, 665679.
23. Emov, V.P. and Morris, N.R. (2000) The LIS1-related NUDF protein of
Aspergillus nidulans interacts with the coiled-coil domain of the NUDE/
RO11 protein. J. Cell Biol., 150, 681688.
24. Kitagawa, M., Umezu, M., Aoki, J., Koizumi, H., Arai, H. and Inoue, K.
(2000) Direct association of LIS1, the lissencephaly gene product, with a
mammalian homologue of a fungal nuclear distribution protein, rNUDE.
FEBS Lett., 479,5762.
25. Tai, C.Y., Dujardin, D.L., Faulkner, N.E. and Vallee, R.B. (2002) Role of
dynein, dynactin, and CLIP-170 interactions in LIS1 kinetochore function.
J. Cell Biol., 156, 959968.
26. Coquelle, F.M., Caspi, M., Cordelieres, F.P., Dompierre, J.P.,
Dujardin, D.L., Koifman, C., Martin, P., Hoogenraad, C.C.,
Akhmanova, A., Galjart, N. et al. (2002) LIS1, CLIP-170s key to the
dynein/dynactin pathway. Mol. Cell Biol., 22, 30893102.
27. Yaffe, M.B. and Elia, A.E. (2001) Phosphoserine/threonine-binding
domains. Curr. Opin. Cell Biol., 13, 131138.
28. Tzivion, G. and Avruch, J. (2002) 14-3-3 proteins: active cofactors in
cellular regulation by serine/threonine phosphorylation. J. Biol. Chem., 277,
30613064.
29. Shaw, A. (2000) The 14-3-3 proteins. Curr. Biol., 10, R400.
30. Muslin, A.J. and Xing, H. (2000) 14-3-3 proteins: regulation of subcellular
localization by molecular interference. Cell. Signal., 12, 703709.
31. Fu, H., Subramanian, R.R. and Masters, S.C. (2000) 14-3-3 proteins:
structure, function, and regulation. A. Rev. Pharmac. Toxicol., 40,617647.
32. des Portes, V., Pinard, J.M., Billuart, P., Vinet, M.C., Koulakoff, A.,
Carrie, A., Gelot, A., Dupuis, E., Motte, J., Berwald-Netter, Y. et al. (1998)
A novel CNS gene required for neuronal migration and involved in
X-linked subcortical laminar heterotopia and lissencephaly syndrome. Cell,
92,5161.
33. Gleeson, J.G., Allen, K.M., Fox, J.W., Lamperti, E.D., Berkovic, S.,
Scheffer, I., Cooper, E.C., Dobyns, W.B., Minnerath, S.R., Ross, M.E. et al.
(1998) Doublecortin, a brain-specic gene mutated in human X-linked
lissencephaly and double cortex syndrome, encodes a putative signaling
protein. Cell, 92,6372.
34. Horesh, D., Sapir, T., Francis, F., Wolf, S.G., Caspi, M., Elbaum, M.,
Chelly, J. and Reiner, O. (1999) Doublecortin, a stabilizer of microtubules.
Hum. Mol. Genet., 8, 15991610.
35. Gleeson, J.G., Lin, P.T., Flanagan, L.A. and Walsh, C.A. (1999)
Doublecortin is a microtubule-associated protein and is expressed widely
by migrating neurons. Neuron, 23, 257271.
36. Sapir, T., Horesh, D., Caspi, M., Atlas, R., Burgess, H.A., Wolf, S.G.,
Francis, F., Chelly, J., Elbaum, M., Pietrokovski, S. et al. (2000)
Doublecortin mutations cluster in evolutionarily conserved functional
domains. Hum. Mol. Genet., 9, 703712.
37. Taylor, K.R., Holzer, A.K., Bazan, J.F., Walsh, C.A. and Gleeson, J.G.
(2000) Patient mutations in doublecortin dene a repeated tubulin-binding
domain. J. Biol. Chem., 275, 3444234450.
38. Caspi, M., Atlas, R., Kantor, A., Sapir, T. and Reiner, O. (2000) Interaction
between LIS1 and doublecortin, two lissencephaly gene products. Hum.
Mol. Genet., 9, 22052213.
39. DArcangelo, G., Miao, G.G., Chen, S.C., Soares, H.D., Morgan, J.I. and
Curran, T. (1995) A protein related to extracellular matrix proteins deleted
in the mouse mutant reeler. Nature, 374, 719723.
40. Hirotsune, S., Takahara, T., Sasaki, N., Hirose, K., Yoshiki, A., Ohashi, T.,
Kusakabe, M., Murakami, Y., Muramatsu, M., Watanabe, S. et al. (1995)
The reeler gene encodes a protein with an EGF-like motif expressed by
pioneer neurons. Nat. Genet., 10,7783.
41. Ogawa, M., Miyata, T., Nakajima, K., Yagyu, K., Seike, M., Ikenaka, K.,
Yamamoto, H. and Mikoshiba, K. (1995) The reeler gene-associated
antigen on Cajal-Retzius neurons is a crucial molecule for laminar
organization of cortical neurons. Neuron, 14, 899912.
42. DeSilva, U., DArcangelo, G., Braden, V.V., Chen, J., Miao, G.G.,
Curran, T. and Green, E.D. (1997) The human reelin gene: isolation,
sequencing, and mapping on chromosome 7. Genome Res., 7, 157164.
43. Hiesberger, T., Trommsdorff, M., Howell, B.W., Gofnet, A.,
Mumby, M.C., Cooper, J.A. and Herz, J. (1999) Direct binding of
Reelin to VLDL receptor and ApoE receptor 2 induces tyrosine
phosphorylation of disabled-1 and modulates tau phosphorylation. Neuron,
24, 481489.
44. Trommsdorff, M., Gotthardt, M., Hiesberger, T., Shelton, J., Stockinger, W.,
Nimpf, J., Hammer, R.E., Richardson, J.A. and Herz, J. (1999)
Reeler/Disabled-like disruption of neuronal migration in knockout mice
lacking the VLDL receptor and ApoE receptor 2. Cell, 97, 689701.
45. Rice, D.S., Sheldon, M., DArcangelo, G., Nakajima, K., Goldowitz, D. and
Curran, T. (1998) Disabled-1 acts downstream of Reelin in a signaling
pathway that controls laminar organization in the mammalian brain.
Development, 125, 37193729.
46. Howell, B.W., Herrick, T.M. and Cooper, J.A. (1999) Reelin-induced
tryosine phosphorylation of disabled 1 during neuronal positioning. Genes
Dev., 13, 643648.
47. Sheldon, M., Rice, D.S., DArcangelo, G., Yoneshima, H., Nakajima, K.,
Mikoshiba, K., Howell, B.W., Cooper, J.A., Goldowitz, D. and Curran, T.
(1997) Scrambler and yotari disrupt the disabled gene and produce a
reeler-like phenotype in mice. Nature, 389, 730733.
48. Miura, H., Yanazawa, M., Kato, K. and Kitamura, K. (1997) Expression of
a novel aristaless related homeobox gene Arx in the vertebrate
telencephalon, diencephalon and oor plate. Mech. Dev., 65,99109.
49. Stromme, P., Mangelsdorf, M.E., Shaw, M.A., Lower, K.M., Lewis, S.M.,
Bruyere, H., Lutcherath, V., Gedeon, A.K., Wallace, R.H., Scheffer, I.E.
et al. (2002) Mutations in the human ortholog of Aristaless cause X-linked
mental retardation and epilepsy. Nat. Genet., 30, 441445.
50. Kitamura, K., Yanazawa, M., Sugiyama, N., Miura, H., Iizuka-Kogo, A.,
Kusaka, M., Omichi, K., Suzuki, R., Kato-Fukui, Y., Kamiirisa, K. et al.
(2002) Mutation of ARX causes abnormal development of forebrain and
testes in mice and X-linked lissencephaly with abnormal genitalia in
humans. Nat. Genet., 32, 359369.
51. Pilz, D.T., Matsumoto, N., Minnerath, S., Mills, P., Gleeson, J.G.,
Allen, K.M., Walsh, C.A., Barkovich, A.J., Dobyns, W.B., Ledbetter, D.H.
et al. (1998) LIS1 and XLIS (DCX) mutations cause most classical
lissencephaly, but different patterns of malformation. Hum. Mol. Genet., 7,
20292037.
52. Dobyns, W.B., Truwit, C.L., Ross, M.E., Matsumoto, N., Pilz, D.T.,
Ledbetter, D.H., Gleeson, J.G., Walsh, C.A. and Barkovich, A.J. (1999)
Differences in the gyral pattern distinguish chromosome 17-linked and
X-linked lissencephaly. Neurology, 53, 270277.
53. Cardoso, C., Leventer, R.J., Matsumoto, N., Kuc, J.A., Ramocki, M.B.,
Mewborn, S.K., Dudlicek, L.L., May, L.F., Mills, P.L., Das, S. et al. (2000)
The location and type of mutation predict malformation severity in isolated
lissencephaly caused by abnormalities within the LIS1 gene. Hum. Mol.
Genet., 9, 30193028.
54. Pilz, D.T., Kuc, J., Matsumoto, N., Bodurtha, J., Bernadi, B., Tassinari, C.A.,
Dobyns, W.B. and Ledbetter, D.H. (1999) Subcortical band heterotopia in
rare affected males can be caused by missense mutations in DCX (XLIS) or
LIS1. Hum. Mol. Genet., 8, 17571760.
55. Leventer, R.J., Cardoso, C., Ledbetter, D.H. and Dobyns, W.B. (2001) LIS1
missense mutations cause milder lissencephaly phenotypes including a
child with normal IQ. Neurology, 57, 416422.
56. Sweeney, K.J., Clark, G.D., Prokscha, A., Dobyns, W.B. and Eichele, G.
(2000) Lissencephaly associated mutations suggest a requirement for the
PAFAH1B heterotrimeric complex in brain development. Mech. Dev., 92,
263271.
57. Smith, T.F., Gaitatzes, C., Saxena, K. and Neer, E.J. (1999) The WD repeat:
a common architecture for diverse functions. Trends Biochem Sci., 24,
181185.
58. Dobyns, W.B., Stratton, R.F., Parke, J.T., Greenberg, F., Nussbaum, R.L.
and Ledbetter, D.H. (1983) MillerDieker syndrome: lissencephaly and
monosomy 17p. J. Pediatr., 102, 552558.
59. Dobyns, W.B., Curry, C.J., Hoyme, H.E., Turlington, L. and Ledbetter, D.H.
(1991) Clinical and molecular diagnosis of MillerDieker syndrome.
Am. J. Hum. Genet., 48, 584594.
60. Chong, S.S., Pack, S.D., Roschke, A.V., Tanigami, A., Carrozzo, R.,
Smith, A.C., Dobyns, W.B. and Ledbetter, D.H. (1997) A revision of the
Human Molecular Genetics, 2003, Vol. 12, Review Issue 1 R95
lissencephaly and MillerDieker syndrome critical regions in chromosome
17p13.3. Hum. Mol. Genet., 6, 147155.
61. Cardoso, C., Leventer, R.J., Ward, H.L., Toyo-oka, K., Chung, J., Gross, A.,
Martin, C.L., Allanson, J., Pilz, D.T., Olney, A.H. et al. (2003) Renement
of a 400 kb critical region allows genotypic differentiation between isolated
lissencephaly, MillerDieker syndrome and other phenotypes secondary to
deletions of 17p13.3. Am. J. Hum. Genet. (in press).
62. Demelas, L., Serra, G., Conti, M., Achene, A., Mastropaolo, C.,
Matsumoto, N., Dudlicek, L.L., Mills, P.L., Dobyns, W.B., Ledbetter, D.H.
et al. (2001) Incomplete penetrance with normal MRI in a woman with
germline mutation of the DCX gene. Neurology, 57, 327330.
63. Matsumoto, N., Leventer, R.J., Kuc, J.A., Mewborn, S.K., Dudlicek, L.L.,
Ramocki, M.B., Pilz, D.T., Mills, P.L., Das, S., Ross, M.E. et al. (2001)
Mutation analysis of the DCX gene and genotype/phenotype correlation in
subcortical band heterotopia. Eur. J. Hum. Genet., 9,512.
64. Gleeson, J.G., Minnerath, S., Kuzniecky, R.I., Dobyns, W.B., Young, I.D.,
Ross, M.E. and Walsh, C.A. (2000) Somatic and germline mosaic
mutations in the doublecortin gene are associated with variable phenotypes.
Am. J. Hum. Genet., 67, 574581.
65. Kato, M., Kanai, M., Soma, O., Takusa, Y., Kimura, T., Numakura, C.,
Matsuki, T., Nakamura, S. and Hayasaka, K. (2001) Mutation of the
doublecortin gene in male patients with double cortex syndrome: somatic
mosaicism detected by hair root analysis. Ann. Neurol., 50, 547551.
66. Poolos, N.P., Das, S., Clark, G.D., Lardizabal, D., Noebels, J.L.,
Wyllie, E. and Dobyns, W.B. (2002) Males with epilepsy, complete
subcortical band heterotopia, and somatic mosaicism for DCX.
Neurology, 58, 15591562.
67. Hong, S.E., Shugart, Y.Y., Huang, D.T., Shahwan, S.A., Grant, P.E.,
Hourihane, J.O., Martin, N.D. and Walsh, C.A. (2000) Autosomal recessive
lissencephaly with cerebellar hypoplasia is associated with human RELN
mutations. Nat. Genet., 26,9396.
68. Kato, M., Takizawa, N., Yamada, S., Ito, A., Honma, T., Hashimoto, M.,
Saito, E., Ohta, T., Chikaoka, H. and Hayasaka, K. (1999) Diffuse
pachygyria with cerebellar hypoplasia: a milder form of microlissencephaly
or a new genetic syndrome? Ann. Neurol., 46, 660663.
69. Bonneau, D., Toutain, A., Laquerriere, A., Marret, S., Saugier-Veber, P.,
Barthez, M.A., Radi, S., Biran-Mucignat, V., Rodriguez, D. and Gelot, A.
(2002) X-linked lissencephaly with absent corpus callosum and ambiguous
genitalia (XLAG): clinical, magnetic resonance imaging, and neuropatho-
logical ndings. Ann. Neurol., 51, 340349.
70. Frints, S.G., Froyen, G., Marynen, P., Willekens, D., Legius, E. and
Fryns, J.P. (2002) Re-evaluation of MRX36 family after discovery of an
ARX gene mutation reveals mild neurological features of Partington
syndrome. Am. J. Med. Genet., 112, 427428.
71. Stromme, P., Mangelsdorf, M.E., Scheffer, I.E. and Gecz, J. (2002) Infantile
spasms, dystonia, and other X-linked phenotypes caused by mutations in
Aristaless related homeobox gene, ARX. Brain Dev., 24, 266268.
72. Scheffer, I.E., Wallace, R.H., Phillips, F.L., Hewson, P., Reardon, K.,
Parasivam, G., Stromme, P., Berkovic, S.F., Gecz, J. and Mulley, J.C. (2002)
X-linked myoclonic epilepsy with spasticity and intellectual disability:
mutation in the homeobox gene ARX. Neurology, 59, 348356.
73. Turner, G., Partington, M., Kerr, B., Mangelsdorf, M. and Gecz, J. (2002)
Variable expression of mental retardation, autism, seizures, and dystonic
hand movements in two families with an identical ARX gene mutation.
Am. J. Med. Genet., 112, 405411.
74. Bienvenu, T., Poirier, K., Friocourt, G., Bahi, N., Beaumont, D.,
Fauchereau, F., Ben Jeema, L., Zemni, R., Vinet, M.C., Francis, F. et al.
(2002) ARX, a novel Prd-class-homeobox gene highly expressed in the
telencephalon, is mutated in X-linked mental retardation. Hum. Mol.
Genet., 11, 981991.
75. Rakic, P. (1972) Mode of cell migration to the supercial layers of fetal
monkey neocortex. J. Comp. Neurol., 145,6183.
76. Jellinger, K. and Rett, A. (1976) Agyria-pachygyria (lissencephaly
syndrome). Neuropadiatrie, 7,6691.
77. Walsh, C. and Cepko, C.L. (1988) Clonally related cortical cells show
several migration patterns. Science, 241, 13421345.
78. Anderson, S.A., Eisenstat, D.D., Shi, L. and Rubenstein, J.L. (1997)
Interneuron migration from basal forebrain to neocortex: dependence on
Dlx genes. Science, 278, 474476.
79. Nadarajah, B., Brunstrom, J.E., Grutzendler, J., Wong, R.O. and
Pearlman, A.L. (2001) Two modes of radial migration in early development
of the cerebral cortex. Nat. Neurosci., 4, 143150.
80. Nadarajah, B. and Parnavelas, J.G. (2002) Modes of neuronal migration in
the developing cerebral cortex. Nat. Rev. Neurosci., 3, 423432.
81. Angevine, J.B. and Sidman, R.L. (1961) Autoradiographic study of cell
migration during histogenesis of cerebral cortex in the mouse. Nature, 192,
766768.
82. Denaxa, M., Chan, C.H., Schachner, M., Parnavelas, J.G. and
Karagogeos, D. (2001) The adhesion molecule TAG-1 mediates the
migration of cortical interneurons from the ganglionic eminence along the
corticofugal ber system. Development, 128, 46354644.
83. Parnavelas, J.G. (2000) The origin and migration of cortical neurones: new
vistas. Trends Neurosci., 23, 126131.
84. Marin, O. and Rubenstein, J.L. (2001) A long, remarkable journey:
tangential migration in the telencephalon. Nat. Rev. Neurosci., 2,
780790.
85. Nadarajah, B., Alifragis, P., Wong, R.O. and Parnavelas, J.G. (2002)
Ventricle-directed migration in the developing cerebral cortex. Nat.
Neurosci., 5, 218224.
86. Caviness, V.S. Jr (1982) Neocortical histogenesis in normal and reeler mice:
a developmental study based upon [3H]thymidine autoradiography. Brain
Res., 256, 293302.
87. Sheppard, A.M. and Pearlman, A.L. (1997) Abnormal reorganization of
preplate neurons and their associated extracellular matrix: an early
manifestation of altered neocortical development in the reeler mutant
mouse. J. Comp. Neurol., 378, 173179.
88. Forster, E., Tielsch, A., Saum, B., Weiss, K.H., Johanssen, C.,
Graus-Porta, D., Muller, U. and Frotscher, M. (2002) Reelin, Disabled 1,
and beta 1 integrins are required for the formation of the radial glial
scaffold in the hippocampus. Proc. Natl Acad. Sci. USA, 99, 1317813183.
89. Gambello, M.J., Darling, D.L., Yingling, J., Tanaka, T., Gleeson, J.G. and
Wynshaw-Boris, A. (2003) Multiple dose dependent effects of Lis1 on
cerebral cortical development. J. Neurosci. (in press).
90. Hirotsune, S., Fleck, M.W., Gambello, M.J., Bix, G.J., Chen, A.,
Clark, G.D., Ledbetter, D.H., McBain, C.J. and Wynshaw-Boris, A. (1998)
Graded reduction of Pafah1b1 (Lis1) activity results in neuronal migration
defects and early embryonic lethality. Nat. Genet., 19, 333339.
91. Francis, F., Koulakoff, A., Boucher, D., Chafey, P., Schaar, B., Vinet, M.C.,
Friocourt, G., McDonnell, N., Reiner, O., Kahn, A. et al. (1999)
Doublecortin is a developmentally regulated, microtubule-associated
protein expressed in migrating and differentiating neurons. Neuron, 23,
247256.
92. Meyer, G., Perez-Garcia, C.G. and Gleeson, J.G. (2002) Selective
expression of doublecortin and LIS1 in developing human cortex suggests
unique modes of neuronal movement. Cereb. Cortex, 12, 12251236.
93. Letinic, K., Zoncu, R. and Rakic, P. (2002) Origin of GABAergic neurons
in the human neocortex. Nature, 417, 645649.
94. Hourihane, J.O., Bennett, C.P., Chaudhuri, R., Robb, S.A. and Martin, N.D.
(1993) A sibship with a neuronal migration defect, cerebellar hypoplasia
and congenital lymphedema. Neuropediatrics, 24,4346.
R96 Human Molecular Genetics, 2003, Vol. 12, Review Issue 1
... During embryogenesis, the migration of post-mitotic neurons from the ventricular zone to the cortical plate stands as a pivotal stage in brain development. When this process is deficient, it frequently leads to significant brain malformations, such as lissencephaly [9]. Despite thorough diagnostic evaluations, a considerable number of individuals with malformations of cortical development continue to lack a molecular diagnosis. ...
... Individuals with PTEN (phosphate and tensin homologue) gene mutations also typically present cortical malformations (Elia et al., 2012). Unlike FCD (arising due to cell proliferation and differentiation), lissencephaly occurs during neuronal migration, triggering a severe malformation with undeveloped or even missing gyri and sulci (Fry et al., 2014;Kato and Dobyns, 2003;Wong and Roper, 2016;Wynshaw-Boris et al., 2010). Clinical lissencephaly is characterized by epilepsy, intellectual disability, as well as cognitive, motor, behavioral and social deficits (Table 1). ...
... ErbB signalling is vital for forming neural circuitry and has been implicated in neuropsychiatric disorder pathology (Brinkman et al., Mei et al.). A particularly interesting target gene is Dcx (Doublecortin), which is required for normal migration of neurons in cortical development and mutated in the brain disorder lissencephaly [35]. As Novel_1 appears to target genes and pathways critical to neurodevelopment, Novel_1 expression may influence crucial processes in the embryonic mouse brain. ...
Article
Full-text available
Many advances in small RNA-seq technology and bioinformatics pipelines have been made recently, permitting the discovery of novel miRNAs in the embryonic day 15.5 (E15.5) mouse brain. We aimed to improve miRNA discovery in this tissue to expand our knowledge of the regulatory networks that underpin normal neurodevelopment, find new candidates for neurodevelopmental disorder aetiology, and deepen our understanding of non-coding RNA evolution. A high-quality small RNA-seq dataset of 458 M reads was generated. An unbiased miRNA discovery pipeline identified fifty putative novel miRNAs, six of which were selected for further validation. A combination of conservation analysis and target functional prediction was used to determine the authenticity of novel miRNA candidates. These findings demonstrate that miRNAs remain to be discovered, particularly if they have the features of other small RNA species.
... Arx is expressed in GABAergic interneurons (INs) (Colombo et al., 2004;Poirier et al., 2004;Friocourt et al., 2006) and binds to b -catenin to positively regulate downstream signaling (Cho et al., 2017). The adenomatous polyposis coli (APC) protein, part of the b -catenin destruction complex, associates with FOXG1 and LIS1, regulating Wnt signaling and directing cell fate in the developing brain (Kato and Dobyns, 2003;McManus et al., 2004;Hebbar et al., 2008;Danesin et al., 2009;Ivaniutsin et al., 2009;Miyoshi and Fishell, 2012). TSC1/2 functions as a repressor of b -catenin, and conditional deletion of TSC1/2 in GABAergic IN progenitors decreases seizure threshold (Mak et al., 2005;J. ...
Article
Full-text available
Infantile and epileptic spasms syndrome (IESS) is a childhood epilepsy syndrome characterized by infantile or late-onset spasms, abnormal neonatal EEG, and epilepsy. Few treatments exist for IESS, clinical outcomes are poor, and the molecular and circuit-level etiologies of IESS are not well understood. Multiple human IESS risk genes are linked to Wnt/β-catenin signaling, a pathway that controls developmental transcriptional programs and promotes glutamatergic excitation via β-catenin’s role as a synaptic scaffold. We previously showed that deleting adenomatous polyposis coli (APC), a component of the β-catenin destruction complex, in excitatory neurons (APC cKO mice, APC fl/fl x CaMKIIα Cre ) increased β-catenin levels in developing glutamatergic neurons and led to infantile behavioral spasms, abnormal neonatal EEG, and adult epilepsy. Here, we tested the hypothesis that the development of GABAergic interneurons (INs) is disrupted in APC cKO male and female mice. IN dysfunction is implicated in human IESS, is a feature of other rodent models of IESS, and may contribute to the manifestation of spasms and seizures. We found that parvalbumin-positive INs (PV ⁺ INs), an important source of cortical inhibition, were decreased in number, underwent disproportionate developmental apoptosis, and had altered dendrite morphology at P9, the peak of behavioral spasms. PV ⁺ INs received excessive excitatory input, and their intrinsic ability to fire action potentials was reduced at all time points examined (P9, P14, P60). Subsequently, GABAergic transmission onto pyramidal neurons was uniquely altered in the somatosensory cortex of APC cKO mice at all ages, with both decreased IPSC input at P14 and enhanced IPSC input at P9 and P60. These results indicate that inhibitory circuit dysfunction occurs in APC cKOs and, along with known changes in excitation, may contribute to IESS-related phenotypes. Significance Statement: Infantile and epileptic spasms syndrome (IESS) is a devastating epilepsy with limited treatment options and poor clinical outcomes. The molecular, cellular, and circuit disruptions that cause infantile spasms and seizures are largely unknown, but inhibitory GABAergic interneuron dysfunction has been implicated in rodent models of IESS and may contribute to human IESS. Here, we use a rodent model of IESS, the APC cKO mouse, in which β-catenin signaling is increased in excitatory neurons. This results in altered parvalbumin-positive GABAergic interneuron development and GABAergic synaptic dysfunction throughout life, showing that pathology arising in excitatory neurons can initiate long-term interneuron dysfunction. Our findings further implicate GABAergic dysfunction in IESS, even when pathology is initiated in other neuronal types.
... LIS spektrumu sınıflandırılması ile ilgili ilk çalışmalarda, giral malformasyonlar ağırdan hafife 1'den 6'ya kadar derecelendirilmişti(5)(6)(7). Bu derecelendirme sistemine ek olarak giral malformasyonun gradientinin beynin farklı bölgelerinde değişkenlik gösterdiği anteriyor (A) veya posteriyor (P) etkilenmenin ağırlığına göre genotip-fenotip ilişkisi kurulmaya çalışılmıştır. ...
Chapter
GMH is a rather common condition of neuronal migration disorders. GMH is associated with seizures and epilepsy, and may cause intellectual disabilities and developmental delay, causing a huge impact on normal childhood development. Even though previous studies have identified genetic and epigenetic risk factors associated with the development of GMH, the actual mechanism causing GMH remains unclear. The severity of the clinical presentation and prognosis of GMH depends on the type, size, and location. Therefore, defining the type and extent of GMH is useful in outlining the management and predicting patients’ prognosis and outcome.
Article
Full-text available
During the development of the cerebral cortex, N-cadherin plays a crucial role in facilitating radial migration by enabling cell-to-cell adhesion between migrating neurons and radial glial fibers or Cajar-Reztius cells. ADP ribosylation factor 4 (Arf4) and Arf5, which belong to the class II Arf small GTPase subfamily, control membrane trafficking in the endocytic and secretory pathways. However, their specific contribution to cerebral cortex development remains unclear. In this study, we sought to investigate the functional involvement of class II Arfs in radial migration during the layer formation of the cerebral cortex using mouse embryos and pups. Our findings indicate that knockdown of Arf4, but not Arf5, resulted in the stalling of transfected neurons with disorientation of the Golgi in the upper intermediate zone (IZ) and reduction in the migration speed in both the IZ and cortical plate. Migrating neurons with Arf4 knockdown exhibited cytoplasmic accumulation of N-cadherin, along with disturbed organelle morphology and distribution. Furthermore, supplementation of exogenous N-cadherin partially rescued the migration defect caused by Arf4 knockdown. In conclusion, our results suggest that Arf4 plays a crucial role in regulating radial migration via N-cadherin trafficking during cerebral cortical development. Significance Statement In the cortical layer formation, the distribution of N-cadherin on cell surface in migrating neurons is tightly regulated by endosomal trafficking system. However, its molecular detail remained fully understood. Here, we demonstrated that Arf4 small GTPase, a critical regulator of membrane trafficking in the trans-Golgi network (TGN), plays distinct roles from Arf5 in radial migration. We further demonstrated that Arf4 regulates N-cadherin trafficking in and out of the TGN and the contact of migrating neurons with radial fibers. Our results suggest that Arf4 regulates radial migration through N-cadherin trafficking in the TGN.
Article
Full-text available
Whereas total loss of Lis1 is lethal, disruption of one allele of the Lis1 gene results in brain abnormalities, indicating that developing neurons are particularly sensitive to a reduction in Lis1 dosage. Here we show that Lis1 is enriched in neurons relative to levels in other cell types, and that Lis1 interacts with the microtubule motor cytoplasmic dynein. Production of more Lis1 in non-neuronal cells increases retrograde movement of cytoplasmic dynein and leads to peripheral accumulation of microtubules. These changes may reflect neuron-like dynein behaviours induced by abundant Lis1. Lis1 deficiency produces the opposite phenotype. Our results indicate that abundance of Lis1 in neurons may stimulate specific dynein functions that function in neuronal migration and axon growth.
Article
Full-text available
Mutations in the LIS1 gene cause gross histological disorganization of the developing human brain, resulting in a brain surface that is almost smooth. Here we show that LIS1 protein co-immunoprecipitates with cytoplasmic dynein and dynactin, and localizes to the cell cortex and to mitotic kinetochores, which are known sites for binding of cytoplasmic dynein. Overexpression of LIS1 in cultured mammalian cells interferes with mitotic progression and leads to spindle misorientation. Injection of anti-LIS1 antibody interferes with attachment of chromosomes to the metaphase plate, and leads to chromosome loss. We conclude that LIS1 participates in a subset of dynein functions, and may regulate the division of neuronal progenitor cells in the developing brain.
Article
Article
PLATELET-ACTIVATING factor (PAF) is involved in a variety of biological and pathological processes1 and PAF acetylhydrolase, which inactivates PAF by removing the acetyl group at the sn-2 position, is widely distributed in plasma and tissue cytosols2,3. One isoform of PAF acetylhydrolase present in bovine brain cortex is a heterotrimer comprising subunits with relative molecular masses of 45K, 30K and 29K (ref. 4). We have now isolated the comple-mentary DNA for the 45K subunit. Sequence analysis revealed a striking identity (99%) of the subunit with a protein encoded by the causative gene (LIS-1) for Miller-Dieker lissencephaly5, a human brain malformation manifested by a smooth cerebral surface and abnormal neuronal migration. This indicates that the LIS-1 gene product is a human homologue of the 45K subunit of intracellular PAF acetylhydrolase. Our results raise the possibility that PAF and PAF acetylhydrolase are important in the formation of the brain cortex during differentiation and development.
Article
The formation of the distinct layers of the cerebral cortex begins when cortical plate neurons take up positions within the extracellular matrix (ECM)-rich preplate, dividing it into the marginal zone above and the subplate below. We have analyzed this process in the reeler mutant mouse, in which cortical lamination is severely disrupted. The recent observation that the product of the reeler gene is an ECM-like protein that is expressed by cells of the marginal zone indicates a critical role for ECM in cortical lamination. We have found that preplate cells in normal cortex that are tagged during their terminal division with bromodeoxyuridine (BrdU) are closely associated with chondroitin sulfate proteoglycans (CSPGs), which were identified by immunolabeling; this association is maintained in the marginal zone and subplate after the preplate is divided by cortical plate formation. Cortical plate cells do not aggregate within the preplate in reeler; instead, preplate cells remain as an undivided superficial layer containing abundant CSPGs, and cortical plate neurons accumulate below them. These findings indicate that preplate cells are responsible for the formation of a localized ECM, because the association of CSPGs with preplate cells is maintained even when these cells are in abnormal positions. The failure of cortical plate neurons to aggregate within the framework of the preplate and its associated ECM and to divide it is one of the earliest structural abnormalities detectable in reeler cortex, suggesting that this step is important for the subsequent formation of cortical layers. J. Comp. Neurol. 378:173–179, 1997. © 1997 Wiley-Liss, Inc.
Article
X-linked lissencephaly with absent corpus callosum and ambiguous genitalia is a newly recognized syndrome responsible for a severe neurological disorder of neonatal onset in boys. Based on the observations of 3 new cases, we confirm the phenotype in affected boys, describe additional MRI findings, report the neuropathological data, and show that carrier females may exhibit neurological and magnetic resonance imaging abnormalities. In affected boys, consistent clinical features of X-linked lissencephaly with absent corpus callosum and ambiguous genitalia are intractable epilepsy of neonatal onset, severe hypotonia, poor responsiveness, genital abnormalities, and early death. On magnetic resonance imaging, a gyration defect consisting of anterior pachygyria and posterior agyria with a moderately thickened brain cortex, dysplastic basal ganglia and complete agenesis of the corpus callosum are consistently found. Neuropathological examination of the brain shows a trilayered cortex containing exclusively pyramidal neurons, a neuronal migration defect, a disorganization of the basal ganglia, and gliotic and spongy white matter. Finally, females related to affected boys may have mental retardation and epilepsy, and they often display agenesis of the corpus callosum. These findings expand the phenotype of X-linked lissencephaly with absent corpus callosum and ambiguous genitalia, may help in the detection of carrier females in affected families, and give arguments for a semidominant X-linked mode of inheritance.