ArticlePDF Available

The Relationship betweenSkewed X-chromosome Inactivation and Neurological Disorders Development: A Review

Authors:

Abstract and Figures

X-chromosome inactivation (XCI) is a process by which one of the copies of the X chromosome in mammalian female cells is inactivated. The XCI causes a balanced X-linked gene quantity between male and females; moreover, it results mosaic females which have paternal active X in some cells and maternal active X in others. Cellular mosaicism is a noteworthy phenomenon and lowers the risk of X-linked diseases in women because the presentation of a mutation on both X chromosomes is unlikely. Therefore, in heterozygous females, the XCI will be present only on the half of the X genome. In contrast, a similar mutation will present in all of the cells of men.Female carriers of some neurological disorders such as autism, Rett syndrome, adreno-leukodystrophyand X-linked mental retardation are reported to present XCI. These observations underscore the important role of X chromosome in the brain which may be related to the existence of a chromosomal signature of gene expression associated with the X-chromosome for neurological conditions not normally associated with that chromosome.In this review, we focused on latestinvestigations on the role of XCI in neurodevelopmental disorders and how these investigations can be effective in the treatment of neurodevelopmental disorders.
Content may be subject to copyright.
81
Review Article
International Clinical Neuroscience Journal Vol 3, No 2, Spring 2016
The Relationship between Skewed X-chromosome
Inactivation and Neurological Disorders Development
Mahdi Taherian1, Hossein Maghsoudi2, Kazem Bidaki2, Reza Taherian3
1 Food and Drug Administration, Reference Laboratory for Food and Drug Control, Tehran, Iran
2 Department of Biology, Payame Noor University, Tehran, Iran
3 Students’ Research Committee, School of M edicine, Shahid Beheshti University of Medical Sciences, Tehran, Iran
ABSTRACT
X-chromosome inactivation (XCI) is a process by which one of the copies of the X chromosome
in mammalian female cells is inactivated. The XCI causes a balanced X-linked gene quantity
between male and females; moreover, it results mosaic females which have paternal active X
in some cells and maternal active X in others. Cellular mosaicism is a noteworthy phenomenon
and lowers the risk of X-linked diseases in women because the presentation of a mutation on
both X chromosomes is unlikely. Therefore, in heterozygous females, the XCI will be present
only on the half of the X genome. In contrast, a similar mutation will present in all of the cells
of men. Female carriers of some neurological disorders such as autism, Rett syndrome, adreno-
leukodystrophy and X-linked mental retardation are reported to present XCI. These observations
underscore the important role of X chromosome in the brain which may be related to the
existence of a chromosomal signature of gene expression associated with the X-chromosome
for neurological conditions not normally associated with that chromosome. In this review, we
focused on latest investigations on the role of XCI in neurodevelopmental disorders and how
these investigations can be effective in the treatment of neurodevelopmental disorders.
Keywords: Adreno-leukodystrophy; Rett Syndrome; X Chromosome Inactivation, X-linked mental retardation
INTRODUCTION
According to the Lyon hypothesis, after the random
inactivation of paternal or maternal X chromosome in a
cell, all of the daughters of that cell will have a similar
inactivated X. This process results a mosaic females which
have paternal active X in some cells and maternal active
X in others 1. Somatic cells have a stable X chromosome
inactivation (XCI); however, germline cells have a cyclic
XCI. If one of the X was heterochromatinized and
compressed during meiosis, pairing and recombination
would not completely achieved. Hence, while both X are
active during Oogenesis, one of them will be inactivated
during mitotic phase of gametogenesis (Figure 1).
Skewing of the normal pattern of XCI
Intermediate phenotypes in women who carry X-linked
diseases are considered as evidence of skewed XCI 2,3.
More than 80% (95% in severe cases) of cells show
preferred inactivation of one X chromosome in skewed
XCI 4. So that, these heterozygote women mainly will
have a single cell population, i.e. paternal or maternal
active X 5,6. Almost 10% of skewed XCI is by chance,
meaning that, for example, a heterozygous woman for
hemophilia has the least amount of factor VIII and
shows typical form of the disease similar to homozygous
women for the mutant allele 6,7 (Figure 2). Moreover,
chromosomal abnormalities and mutations that give a
ICNSJ 2016; 3 (2) :81-91 www.journals.sbmu.ac.ir/neuroscience
Correspondence to: Reza Taherian, Students’ Research Committee, School of Medicine, Shahid Beheshti University of Medical
Sciences, Tehran, Iran; Tel: +98-9123314094; E-mail: r.taherian@sbmu.ac.ir
Received: April 13, 2016 Accepted: May 21, 2016
Skewed X-chromosome Inactivation and Neurological Disorders—Taherian et al
82 International Clinical Neuroscience Journal Vol 3, No 2, Spring 2016
cell selective advantage or negative attributes may play
a role in this skewness. For instance, if one of the X
chromosomes carry a mutant allele, which opposes
cell survival or inhibits the growth of cells, the cells
will inactivate that X chromosome in order to provide
cell survival and growth 8. Also, mutations in the XIC
(X inactivation center), which is a 13 Mb area in Xq
chromosome, mutation in Xist (X-inactive specific
transcript) promotor 8,9 or Tsix gene (a gene antisense
to Xist) deletion 10 may all cause skewed XCI. Promoter
single nucleotide polymorphisms or mutations affecting
the start of transcription of Tsix, Xist and Xite also
can change start of a correct XCI. For instance, single
nucleotide polymorphism in the second binding protein
of CTCF in Xist causes skewed XCI in order to inactivate
the other polymorphic allele 9-11. It should be noted that
in most cases, the reason for this skewing is unknown.
This skewing may occur during inactivating, which
is called primary skewing, or it may be the result of
selection after the XCI which is called secondary
skewing. For example, the slower proliferation of cells
expressing a mutant allele is considered as a secondary
skewed XCI 3,5. Cross of different strains of mice Shows
Xce (X-controlling element) locus to be the cause of a
primary, but not complete, skewed XCI and suggests
that these alleles cannot disrupt inactivation process but
they can disrupt the randomness of the beginning phases
of inactivation process 12. There are not any evidence
to confirm the presence of Xce locus in humans 13. As
mentioned before, in heterozygous women with a skewed
XCI, a single cell population with similar active X is
observed. Thus, the mutant gene on that chromosome will
be expressed in all cells and the normal allele will be
covered on the inactivated X. Hence, these heterozygous
women can show some diseases which are normally
presented only in men. This phenomenon is known as
manifesting heterozygote 7,14.
Despite some previous reports of familial skewed XCI,
such evidences are insufficient to confirm the heritable
pattern of skewing in humans 16. In carriers of X-linked
diseases, degree of skewing of XCI pattern, can affect the
severity of disease. In general, investigating the primary
phases of XCI is very difficult the spread of XCI in
human studies early events is very difficult. A delayed
replication is a good cytological marker for studying XCI
process. In X:autosome translocations, XCI can spread to
autosomal areas; however, this spread seems less efficient
compared to spreading on X chromosome 17. In women
with Balanced X:autosome translocation, the normal X
will be inactivated completely, but not persistently, to
avoid partial autosomal monosomy or X disomy 3,18. This
selective inactivation of one of the X chromosomes will
cause an X-linked disease in carrier women 4,16,19. In
imbalanced X:autosome translocations, the abnormal X
will preferably be inactivated 3. Investigating the spreading
of XCI into autosomal areas will provide analysis of the
DNA elements involved in the XCI signal propagation.
In some cases of XCI spreading, the translocation stops
in the breaking sites and cannot spread to autosomal
areas 20, while, it will spread incompletely to that areas 21.
Study of spread of Xist RNA into autosomal areas has
shown the variable abilities of Xist to cover autosomal
areas 22. Moreover, gene expression investigations have
shown that most genes in a X:autosome translocation
escape inactivation 3. The incomplete spreading of XCI to
autosomal areas suggests the presence of some functional
areas that contribute to the spread of XCI and are more
frequent on X chromosome compared to autosomal
chromosomes 23. Comparing the abilities of translocation
in spreading XCI raised the “repeat hypothesis” in which
the LINE-1 retrotransposons, which are frequent on X
Figure 2. Schematic representation of XCI in female somatic cells 15.
Figure 1. X chromosome inactivation. Reproduced from https://
www.studyblue.com.
Skewed X-chromosome Inactivation and Neurological Disorders—Taherian et al
83
International Clinical Neuroscience Journal Vol 3, No 2, Spring 2016
chromosome, act as inactivation propagation elements 24.
Bioinformatics studies have shown that LINEs are
frequent around transcription start sites of the genes that
are incurred XCI 25,26. Repeats of polymorphic CAG in
exon 1 of the androgen receptor gene, which is located
on the Xq 11.2, and methylation of active and inactive X
chromosomes can be used to investigate the XCI pattern.
Figure 3 shows the detection method of inactivated X
chromosome by two polymorphic repeats (a and b). The
PCR is done after the effect of methylation-sensitive
enzymes, such as HpaII, which do not break inactive X
chromosome because they are methylated before 4, 27.
Skewed XCI and X-linked diseases
Although gene expression studies are an important
source of data in investigating neurodegenerative disorders,
lists of differentially expressed genes are rarely helpful
in understanding the underlying mechanisms of these
diseases 29. Hence, investigating single chromosomes with
a high role in developing neuro-developmental disorders
Figure 3. Method of identifying inactive X chromosome 28.
Skewed X-chromosome Inactivation and Neurological Disorders—Taherian et al
84 International Clinical Neuroscience Journal Vol 3, No 2, Spring 2016
is more helpful. The X-chromosome contains a large
number of genes that are essential for brain development
and function 30. The expression of chromosome X is
generally higher in brain compared to other tissues
31.
This higher expression is linked to “X dosage
compensation” mechanism; a mechanism that associates
the expression of X-linked genes with the expression of
genes on autosomal chromosomes 32-35. Several brain
disorders are also associated with mutations of genes on
X chromosome 36-38. Heterozygous women show skewed
XCI in several X-linked diseases such as Wiskott-Aldrich,
Lesch–Nyhan, Barth, Muscular dystrophy Duchesne,
hemophilia, and some immune deficiency syndrome 39-41.
In these cases, a higher prevalence of skewed aging is
observed by aging 6,7. The skewed XCI causes variable
manifestations of many neurological disorders including
autism, Rett syndrome, X-linked adreno-leukodystrophy,
X-linked mental retardation and etc 8,27,42-44.
Cell removal is the most common phenomenon
occurred in heterozygote women with two cell populations
and can yield many advantages for them 40. For example,
in heterozygotes older than 10 years who carry the
Lesch-Nyhan syndrome gene, the mutant cells could
not be detected in their blood and they are completely
removed due to presence of cells expressing normal
allele 7. Although the presence of the cells with normal
allele may not be dominant, it is sufficient to preclude
the mutant allele phenotype.
In heterozygous women, dermal cells with normal
allele compete well with Lesch-Nyhan mutation. Because,
normal cells, transmit inosinic acid, which is the product
of hypoxanthine guanine phosphoribosyl transferase
(HGPRT) enzyme, and is absent in this disease. Indeed,
heterozygous women have channels that pass molecules
such as inosinic acid and so, with the removal of cells
with mutant alleles that do not have these channels,
deficiency in HGPRT is compensated 7.
X-linked adrenoleukodystrophy
X-linked adreno-leukodystrophy (X-ALD) was first
described by Haberfeld and Spieler in 1910. However,
Siemerling and Creutzfeld described combination of
adrenocortical atrophy, cerebral demyelination and
lymphocytic infiltration in a case of what is now considered
the first true report of X-ALD. The name X-ALD was first
introduced in 1970 45. In this disorder, the characteristic
accumulation of very long chain fatty acids (VLCFA) is
observed which was first observed as the presence of lipid
inclusions in adrenal cells of X-ALD. Now, detection of
the accumulation of VLCFAs in blood, red blood cells,
fibroblasts and amniocytes with new assays enables
better and earlier diagnosis of X-ALD. The adult form
of X-ALD was named adrenomyeloneuropathy (AMN) by
Griffin et al. in 1977 46. As our knowledge about X-ALD
increases, the diagnosis of this disorder will be more
probable and consequently its incidence will increase.
Like other X-linked disorders, referring the X-ALD as
an X-chromosomal recessive disorder is inappropriate
and it should be simply called as an X-linked disorder 47.
It is estimated that about half of the heterozygous
X-ALD females will develop adrenomyeloneuropathy-
like syndrome
48. Thus, heterozygote X-ALD is more
prevalent than homozygote
49. Thus, X-ALD is both
the most frequent peroxisomal disorder and also the
most frequent monogenetically inherited demyelinating
disorder. This monogenic inherit is now considered to be
related to the mutations in the ABCD1 gene 50 (Figure 4).
The inherited defect in X-ALD is linked to a mutation
in G-6-PD gene which is located in the Xq28 51 with
polymorphic markers 52,53. Using positional cloning,
the G-6-PD was cloned and after that it was termed
adreno-leukodystrophy gene. As this gene encodes a
peroxisomal transmembrane protein with a structure
similar to the ATP-binding cassette transporter, the
gene was renamed to ATP-Binding Cassette transporter
subfamily D member1 gene (ABCD1). However, the
protein is still termed as adreno-leukodystrophy protein
(ALDP). Until now, 431 different mutant ABCD1 alleles
have been reported. Of these mutations, 221 are missense
mutations, 53 nonsense mutations, 27 amino acid
insertions and deletions, 113 frame shift mutations and
17 deletions of one or more exons. These mutations are
equally distributed among the entire coding region of the
ABCD1 gene, however, investigating the 221 missense
Figure 4. X-linked adreno-leukodystrophy genetic. Reproduced from
http://genetics4medics.com
Skewed X-chromosome Inactivation and Neurological Disorders—Taherian et al
85
International Clinical Neuroscience Journal Vol 3, No 2, Spring 2016
mutations showed that there is no disease-associated
mutation within the first 88 N-terminal amino acids and
in the last 45 C-terminal amino acids. Interestingly two
single base pair substitutions in exon 1 is reported in
one of X-ALD cases which causes amino acid exchanges
(N13T and K217E). Previous studies have shown that
K217E amino acid exchange was the only ineffective
exchange in restoration of defected β-oxidation in X-ALD
fibroblasts 54. However, ALDP function is not affected by
the N13T amino acid exchange which is consistent with
the hypothesis that a reduced functional importance in the
first 66 N-terminal amino acids of ALDP is present. No
correlation is observed ABCD1 gene mutation type and
the disease clinical presentation. This report is consistent
with previous studies 55-58. These studies showed that
clinical presentation of X-ALD occur within the same
nuclear family 55, mutations causing complete loss of
ALDP are associated with various clinical presentations 56,
a similar deletion in exon 5 leads to a wide spectrum
of X-ALD 56,57 and finally, monozygotic twins present
completely different clinical presentations 58. However,
these studies do not lower the possibility of the preventive
role of residual ALDP in development of inflammatory
cerebral form of X-ALD and its consequent milder
phenotype. In the p-glycoprotein multidrug resistance
transporter, some mutations can lower the transport rate.
ABCA4 (ABCR) gene may also has residual functional
activity in age-related macular degeneration and Stargardt
disease 59. Thus, although a general genotype–phenotype
association cannot be observed, rare cases with this
association may exist.
Besides the presence of functional ABCD1 gene on
Xq28, several autosomal pseudogenes are detectable of
different chromosomes. ABCD1 pseudogenes are located
in 1, 2, 20, 22, and possibly 16 chromosomes which
can be detected by PCR analysis of human monochro-
mosomal mapping panel 56. The pseudogenes on several
different chromosomes harden the mutation analysis 60
and may show non-homologous interchromosomal
exchange pericentromeric plasticity 61.
X-Linked Mental Retardation
Mental retardation (MR) is known as the impairment
of cognitive function which affects about 3% of the
population in the US. This disorder is often manifested
before adulthood and 20–30% of MR may be inherited
or have a genetic background by mutations on the X
chromosome X-linked mental retardation (XLMR). Based
on their clinical presentation, XLMRs are considered as
two types of non-syndromic and syndromic. However,
since there are several mutations responsible for these
clinical presentation, the difference between these
two types is now less notable. About 73 dysfunction
of PAK are considered to be involved in the incidence
of both types of XLMR. Several X-linked mutations
causing non-syndromic MR (MRX) are detected on the
X chromosome. This type of MR is clinically similar
but they are genetically diverse. Now, more than 65
MRX pedigrees are detected on different loci on the
X chromosome and it is estimated that these pedigrees
represent about 10 genetic loci 62,63. Among these loci,
probably about two MRX genes have the key roles in
the mentioned signaling. The first gene is named GDI1
which encodes a GDP-dissociation inhibitor for Rab3a
and regulates vesicular transport. Hence, GDI1 mutations
changes the exocytic activity which is in part related to
the synaptic transmission. Another gene which is mutated
in MRX encodes a protein which includes a GTPase
activation domain (GAP) for Rho GTPases and is called
oligophrenin. Since these proteins induce the GTPase
activity, inactivation of them leads to permanent activity
of the corresponding G protein. Oligophrenin have Gap
activity for the signals that control actin cytoskeletal
organization and cell shape 64. One of these proteins
is Rho GTPase which directly controls the axon and
dendritic shape and activity 65-66. The role of a Rho GAP
in human mental retardation implicates its involvement
in neural plasticity. Moreover, PAK3 is reported to be
isolated from Xq22 in MRX families 67. The PAK3 gene
is mainly expressed in the brain, and it is known as an
important downstream effector of Rho GTPases through
actin cytoskeleton and MAPK cascades. Two diseases
are reported to have mutations in MRX pedigrees, one
of them includes a nonsense 68 and another a missense
mutation 69 which implicate the association between
PAK3 and pathogenesis of MRX.
Considering the role of PAK in fragile X syndrome
(FXS), recent studies have shown a defective activation
of synaptic RAC/PAK signaling in the mouse model
of FXS 70. Expectantly, inhibiting PAK activity reduces
various cellular and behavioral deficits, including FXS-
related abnormalities 71. These results implicate the role
of PAK signaling in the FXS pathogenesis.
Fragile X syndrome is the most common inherited form
of mental retardation 72,73. Although the pathogenesis of
FXS is not well understood, it is probably the result of
the expansion of the CGG repeat in the 5-untranslated
region of the fragile X mental retardation 1 (FMR1) gene
which is located on the X chromosome 74. This expansion
silences the transcription of FMR1 gene and leads to the
Skewed X-chromosome Inactivation and Neurological Disorders—Taherian et al
86 International Clinical Neuroscience Journal Vol 3, No 2, Spring 2016
FXS phenotype. The resultant change of the length of
the CGG is the main factor in determining FXS disease
or its carrier form. Having more than 200 CGG repeats
causes FXS-associated cognitive deficits and abnormal
cortical dendritic spines 75.
FMR1 protein (FMRP) plays an important role in
regulation of mRNA translation, transport and stability 76,77.
In the brain, FMRP can regulate translation of a group of
mRNAs at synapses which are essential for learning and
intellectual development. Consequently, in the absence
of FMRP, reduced mRNA translation causes the defect
in synaptic function and synaptic plasticity 76,78,79. FMR1
Knocked-out mice show behavioral defects similar to FXS
patients. These behavioral disturbances include hyper-
reactivity to auditory stimuli, anxiety, impaired spatial
learning and impairment in motor coordination 80-82.
Hence, the association between PAK and FMRP 71 or
defected synaptic RAC/PAK signaling implicate the
reduced synaptic plasticity in patients with FXS. Thus,
targeting PAK signaling may be a potential therapeutic
strategy in formulating new drugs to treat FXS.
In summary, previous data show that normal memory
and leaning is achieved by functional PAK and pathways
which are disturbed in FXS and similar disorders with
developmental cognitive deficits such as dementia of
aging (AD) and Huntington disease (HD). Inhibition of
PAK seems to be an effective approach in treating FXS,
AD and HD. Inhibitors of PAK likely have important
effects on improving cognition by improving dendritic
spine morphology and/or synaptic plasticity. This
hypothesis considers PAK activation as a contributing
factor in the incidence of various neurological disorders
and probably suggest a common treatment for them based
on the correcting PAK dysregulation.
Rett Syndrome
Rett syndrome (RTS) is considered as a progressive
disorder affecting neuro-development in girls; however,
some of the features of the disease appears slowing are not
prominent at initial stages of the disease. These patients
develop normally up to 6-18 months of age. During this
development, patients will have a normal milestones
including walking and saying some words. One of the
early signs of neuro-developmental involvement is
microcephaly which is the result of the deceleration of
head growth between ages one and two and is associated
with growth retardation and muscular hypotonia.
Progression of the disease causes a lost purposeful
use of hands which interrupts normal hand functions.
The patient will be withdrawn socially and inability to
speech become more apparent by aging which turns the
patient irritable. The patients is also hypersensitive to
surrounding sounds, cannot make an eye-to-eye contact
and is unresponsiveness to social cues 83.
Genetic Basis of RTS
Considering that most of the RTS patients are female,
early studies have hypothesized an X-linked dominant
model of inheritance in RTS. However, almost all of
the RTS cases are sporadic and mapping the generic
inheritance of the disease was difficult. Using data from
rare familial cases, Xq28 was identified as the candidate
region and subsequently, mutations in MECP2 were
identified in RTS patients 84 (Figure 5).
These mutations can be detected in about 96% of RTS
cases and arise de novo in the paternal germline with a C
to T transition at CpG dinucleotides 85. These mutations
include missense, nonsense, and frameshift mutation types
with over 300 unique pathogenic nucleotide changes
described 86 as well as deletions 87,88. Eight missense
and nonsense mutations are responsible for more than
75% of all of the mutations in RTS. C-terminal deletions
causes 10% and complex rearrangements are responsible
for about 6% of all mutations in RTS. Missense mutations
cause a more mild phenotype than mutations affecting
the NLS of MeCP2; Similarly, C-terminal deletions are
associated with milder phenotypes 89. Moreover, a mild
phenotype is present in the R133C mutation 90,91.
Figure 5. Rett syndrome genetic. Reproduced from http://
genetics4medics.com.
Skewed X-chromosome Inactivation and Neurological Disorders—Taherian et al
87
International Clinical Neuroscience Journal Vol 3, No 2, Spring 2016
Autism
The diagnosis of autism is based on the presence of
abnormality in a triad of behavioral domains: social
development, communication, and repetitive behavior/
obsessive interests. While autism can occur at any point on
the intelligence quotient (IQ) continuum, IQ can predict
the outcome of autism. The disease is associated with a
language delay. Autism has a spectrum which Asperger
syndrome is a subgroup of it. Patients with this syndrome
have similar features of common autism; however, they
have no history of language delay and moreover, they
have an average or higher than average IQ. The features
of autism spectrum is highly probable to be caused by
genetic factors. The risk of presence of autism in the
sibling of a patient is 4.5% which emphasizes a genetic
inheritance 92 (Figure 6).
In a study of same sex autistic twins, while no
dizygotic twins were concordant for autism, concordance
was present in about 60% of monozygotic pairs 93. This
concordance in monozygotic twins shows a high degree
of genetic inheritance of autism. Different genetic studies
are done to determine the candidate regions involved in
autism. Although these regions are not fully understood,
two regions are the most probable candidates. These
regions are 15q11-13, near the GABAAb3 receptor
subunit gene (GABRB3) and 17q11.2, near the serotonin
transporter gene (SLC6A4). The second region is under
high investigations because of the previous reports about
the increased serotonin levels of platelets in autism.
Also, the involvement of serotonin in autism is highly
probable because it innervates the limbic system which
has known roles in emotion recognition and empathy. At
least four loci on the X chromosome are also detected to
be involved in autism and are high interest due to their
ability to clarify sex differences in autism. These genes
are the neuroligin (NLGN3, NLGN4), FMR1 (which
causes fragile X syndrome), and MECP2. Despite there
are candidate regions, specific genes for autism have not
been detected yet. Further research is needed to detect
these specific genes and also to clarify their function
and ultimately the relation between different influential
factors in autism 92.
Charcot-Marie-Tooth
Charcot-Marie-Tooth (CMT) which is known as
hereditary motor and sensory neuropathy consists of
clinically and pathologically heterogeneous group of
disorders. This disease is considered as a common form
of peripheral inherited neuropathy in humans. The disease
includes slowly progressive atrophy; weakness of the
distal muscles; sensory loss in the feet, lower legs, and
hands; and reduced tendon reflexes 94. CMT is classified
into types 1 and 2 based on the histopathology and nerve
conduction studies 95. Although some X-linked and
autosomal recessive forms of the disease are reported,
the most common form of CMT is inherited autosomal
dominant 96. CMT 1A is a demyelinating neuropathy
related to a duplication of piece of gene on 17p11
chromosome which encodes myelin protein 97. It is
the most common form of the disease inherited as an
autosomal dominant disease. The clinical expression of
this disease depends on the age of the patient and the
average age of the onset of clinical features is 12.2+7.3
years. At least three loci are included in CMT1: the
CMT1A locus maps to human chromosome 17 and the
CMT1B locus maps to human chromosome 1 (region
q23-q25) 98. Since the clinical phenotypes of CMT1
syndromes cannot be distinguished, unless in cases of
male to male inheritance, three genes must be sequenced
which include connexin gene, po gene and PMP22 gene
for X-linked CMT, CMT1B and CMT1A respectively 99.
The X-linked CMT due to the connexin 32 (Cx32) gene
mutation is the second most common form of CMT 100
(Figure 7).
This gene is expressed in several tissues including
both peripheral nerve axons and CNS glia and neurons.
Families with dominant inheritance, which is revealed by
a male-to-male inheritance, PMP22 and po genes must be
screened for mutations. Families with no male-to-male
inheritance or CMT1A may be screened for connexin 32
gene first and then Po and PMP22 102. In an X-linked
pattern of inheritance, male are more severely affected
and not father-to-son transmission is seen. Females who
are heterozygous and may be asymptomatic. Intermediate
Figure 6. Autism genetic prospect 92.
Skewed X-chromosome Inactivation and Neurological Disorders—Taherian et al
88 International Clinical Neuroscience Journal Vol 3, No 2, Spring 2016
range motor conduction velocities is found in X-linked
CMT. Accordingly, prolonged brainstem auditory evoked
potentials (BAEPs) may be used as a distinguishing
feature X-linked CMT in males and females due to XCI
in females 103.
DISCLOSURE STATEMENT
The authors have nothing to disclose.
REFERENCES
1. Yen ZC, Meyer IM, Karalic S, Brown CJ. A cross-species
comparison of X-chromosome inactivation in Eutheria.
Genomics. 2007;90(4):453-63.
2. Willard H, Brown C, Carrel L, Hendrich B, Miller A,
editors. Epigenetic and chromosomal control of gene
expression: molecular and genetic analysis of X chromosome
inactivation. Cold Spring Harbor symposia on quantitative
biology; 1993: Cold Spring Harbor Laboratory Press.
3. Yang C, Chapman AG, Kelsey AD, Minks J, Cotton
AM, Brown CJ. X-chromosome inactivation: molecular
mechanisms from the human perspective. Human genetics.
2011;130(2):175-85.
4. Minks J, Robinson WP, Brown CJ. A skewed view of
X chromosome inactivation. The Journal of clinical
investigation. 2007;118(1).
5. Amos-Landgraf JM, Cottle A, Plenge RM, Friez M, Schwartz
CE, Longshore J, et al. X Chromosome–Inactivation Patterns
of 1,005 Phenotypically Unaffected Females. The American
Journal of Human Genetics. 2006;79(3):493-9.
6. Muers MR, Sharpe JA, Garrick D, Sloane-Stanley J,
Nolan PM, Hacker T, et al. Defining the cause of skewed
X-chromosome inactivation in X-linked mental retardation
by use of a mouse model. The American Journal of Human
Genetics. 2007;80(6):1138-49.
7. Migeon BR. Why females are mosaics, X-chromosome
inactivation, and sex differences in disease. Gender medicine.
2007;4(2):97-105.
8. Wang Z, Yan A, Lin Y, Xie H, Zhou C, Lan F. Familial
skewed x chromosome inactivation in adrenoleukodystrophy
manifesting heterozygotes from a chinese pedigree. PloS
one. 2013;8(3):e57977.
9. Pugacheva EM, Tiwari VK, Abdullaev Z, Vostrov AA,
Flanagan PT, Quitschke WW, et al. Familial cases of point
mutations in the XIST promoter reveal a correlation between
CTCF binding and pre-emptive choices of X chromosome
inactivation. Human molecular genetics. 2005;14(7):953-65.
10. Jeon Y, Sarma K, Lee JT. New and Xisting regulatory
mechanisms of X chromosome inactivation. Current opinion
in genetics & development. 2012;22(2):62-71.
11. Barakat TS, Jonkers I, Monkhorst K, Gribnau J. X-changing
information on X inactivation. Experimental cell research.
2010;316(5):679-87.
12. Cattanach B, Isaacson J. Controlling elements in the mouse
X chromosome. Genetics. 1967;57(2):331.
13. Tomkins DJ, McDonald HL, Farrell SA, Brown CJ. Lack
of expression of XIST from a small ring X chromosome
containing the XIST locus in a girl with short stature, facial
dysmorphism and developmental delay. European Journal
of Human Genetics. 2002;10(1):44-51.
14. Salsano E, Tabano S, Sirchia SM, Colapietro P, Castellotti B,
Gellera C, et al. Preferential expression of mutant ABCD1
allele is common in adrenoleukodystrophy female carriers
but unrelated to clinical symptoms. Orphanet journal of
rare diseases. 2012;7(1):1.
15. Franco, Brunella, Andrea Ballabio. X-inactivation and human
disease: X-linked dominant male-lethal disorders. Current
opinion in genetics & development. 2006;16(3):254-259.
16. Bolduc V, Chagnon P, Provost S, Dubé M-P, Belisle C,
Gingras M, et al. No evidence that skewing of X chromosome
inactivation patterns is transmitted to offspring in humans.
The Journal of clinical investigation. 2008;118(1):333-41.
17. Sharp AJ, Spotswood HT, Robinson DO, Turner BM, Jacobs
PA. Molecular and cytogenetic analysis of the spreading
of X inactivation in X; autosome translocations. Human
molecular genetics. 2002;11(25):3145-56.
18. Schmidt M, Du Sart D. Functional disomies of the X
chromosome influence the cell selection and hence the X
inactivation pattern in females with balanced X-autosome
translocations: a review of 122 cases. American journal of
medical genetics. 1992;42(2):161-9.
19. Jacobs PA, Hunt PA, Mayer M, Bart RD. Duchenne
muscular dystrophy (DMD) in a female with an X/autosomal
Figure 7. Genetic heterogeneity in X-linked Charcot–Marie–Tooth
disease 101.
Skewed X-chromosome Inactivation and Neurological Disorders—Taherian et al
89
International Clinical Neuroscience Journal Vol 3, No 2, Spring 2016
translocation: Further evidence that the DMD locus is at
Xp21. American journal of human genetics. 1981;33(4):513.
20. Solari A, Rahn I, Ferreyra M, Carballo M. The behavior of
sex chromosomes in two human X-autosome translocations:
failure of extensive X-inactivation spreading. Biocell:
official journal of the Sociedades Latinoamericanas de
Microscopia Electronica et al. 2001;25(2):155-66.
21. Caiulo A, Bardoni B, Camerino G, Guioli S, Minelli A,
Piantanida M, et al. Cytogenetic and molecular analysis of an
unbalanced translocation (X; 7)(q28; p15) in a dysmorphic
girl. Human genetics. 1989;84(1):51-4.
22. Hall LL, Clemson CM, Byron M, Wydner K, Lawrence JB.
Unbalanced X; autosome translocations provide evidence for
sequence specificity in the association of XIST RNA with
chromatin. Human molecular genetics. 2002;11(25):3157-65.
23. Gartler SM, Riggs AD. Mammalian X-chromosome
inactivation. Annual review of genetics. 1983;17(1):155-90.
24. Lyon MF. X-chromosome inactivation: a repeat hypothesis.
Cytogenetic and Genome Research. 1998;80(1-4):133-7.
25. Wang Z, Willard HF, Mukherjee S, Furey TS. Evidence of
influence of genomic DNA sequence on human X chromosome
inactivation. PLoS Comput Biol. 2006;2(9):e113.
26. Gartler SM, Goldman MA. Biology of the X chromosome.
Current opinion in pediatrics. 2001;13(4):340-5.
27. Bittel DC, Theodoro MF, Kibiryeva N, Fischer W, Talebizadeh
Z, Butler MG. Comparison of X-chromosome inactivation
patterns in multiple tissues from human females. Journal
of medical genetics. 2008;45(5):309-13.
28. Swierczek SI, Agarwal N, Nussenzveig RH, Rothstein G,
Wilson A, Artz A, et al. Hematopoiesis is not clonal in
healthy elderly women. Blood. 2008;112(8):3186-3193.
29. Mitchell AC, Mirnics K. Gene expression profiling of the
brain: pondering facts and fiction. Neurobiology of disease.
2012;45(1):3-7.
30. Laumonnier F, Cuthbert PC, Grant SG. The role of neuronal
complexes in human X-linked brain diseases. The American
Journal of Human Genetics. 2007;80(2):205-20.
31. Disteche CM. High expression of the mammalian X
chromosome in brain. Brain research. 2006;1126(1):46-9.
32. Disteche CM. Dosage compensation of the active X
chromosome in mammals. Nature genetics. 2006;38(1):47-
53.
33. Straub T, Becker PB. Dosage compensation: the beginning
and end of generalization. Nature Reviews Genetics.
2007;8(1):47-57.
34. Deng X, Hiatt JB, Ercan S, Sturgill D, Hillier LW, Schlesinger
F, et al. Evidence for compensatory upregulation of
expressed X-linked genes in mammals, Caenorhabditis
elegans and Drosophila melanogaster. Nature genetics.
2011;43(12):1179-85.
35. Kharchenko PV, Xi R, Park PJ. Evidence for dosage
compensation between the X chromosome and autosomes
in mammals. Nature genetics. 2011;43(12):1167-9.
36. Oostra B, Willemsen R. The X chromosome and fragile
X mental retardation. Cytogenetic and genome research.
2003;99(1-4):257-64.
37. Chadwick LH, Wade PA. MeCP2 in Rett syndrome:
transcriptional repressor or chromatin architectural protein?
Current opinion in genetics & development. 2007;17(2):121-
5.
38. Zanni G, Bertini ES. X-linked disorders with cerebellar
dysgenesis. Orphanet journal of rare diseases. 2011;6(1):1.
39. Migeon BR. The role of X inactivation and cellular
mosaicism in women’s health and sex-specific diseases.
Jama. 2006;295(12):1428-33.
40. Kristiansen M, Langerød A, Knudsen G, Weber B, Børresen-
Dale A, Ørstavik K. High frequency of skewed X inactivation
in young breast cancer patients. Journal of medical genetics.
2002;39(1):30-3.
41. Bicocchi MP, Migeon BR, Pasino M, Lanza T, Bottini F,
Boeri E, et al. Familial nonrandom inactivation linked to
the X inactivation centre in heterozygotes manifesting
haemophilia A. European journal of human genetics.
2005;13(5):635-40.
42. Knudsen GPS, Neilson TC, Pedersen J, Kerr A, Schwartz
M, Hulten M, et al. Increased skewing of X chromosome
inactivation in Rett syndrome patients and their mothers.
European journal of human genetics. 2006;14(11):1189-94.
43. Talebizadeh Z, Bittel D, Veatch O, Kibiryeva N, Butler M.
Brief report: non-random X chromosome inactivation in
females with autism. Journal of autism and developmental
disorders. 2005;35(5):675-81.
44. Plenge RM, Stevenson RA, Lubs HA, Schwartz CE, Willard
HF. Skewed X-chromosome inactivation is a common feature
of X-linked mental retardation disorders. The American
Journal of Human Genetics. 2002;71(1):168-73.
45. Blaw M. Melanodermic type leukodystrophy
(adrenoleukodystrophy). Handbook of clinical neurology.
1970;10:128-33.
46. GRIFFIN JW, GOREN E, SCHAUMBURG H, Engel WK,
LORIAUX L. Adrenomyeloneuropathy A probable variant
of adrenoleukodystrophy I. Clinical and endocrinologic
aspects. Neurology. 1977;27(12):1107-.
47. Dobyns WB, Filauro A, Tomson BN, Chan AS, Ho AW,
Ting NT, et al. Inheritance of most X-linked traits is not
dominant or recessive, just X-linked. American journal of
medical genetics Part A. 2004;129(2):136-43.
48. Restuccia D, Di Lazzaro V, Valeriani M, Oliviero A, Le Pera
D, Colosimo C, et al. Neurophysiological abnormalities
in adrenoleukodystrophy carriers. Evidence of different
degrees of central nervous system involvement. Brain.
1997;120(7):1139-48.
49. Bezman L, Moser AB, Raymond GV, Rinaldo P, Watkins
PA, Smith KD, et al. Adrenoleukodystrophy: incidence, new
mutation rate, and results of extended family screening.
Annals of neurology. 2001;49(4):512-7.
50. Kemp S, Pujol A, Waterham HR, van Geel BM, Boehm CD,
Raymond GV, et al. ABCD1 mutations and the X-linked
adrenoleukodystrophy mutation database: Role in diagnosis
and clinical correlations. Human mutation. 2001;18(6):499-
515.
51. Migeon BR, Moser HW, Moser AB, Axelman J, Sillence
D, Norum RA. Adrenoleukodystrophy: evidence for X
Skewed X-chromosome Inactivation and Neurological Disorders—Taherian et al
90 International Clinical Neuroscience Journal Vol 3, No 2, Spring 2016
linkage, inactivation, and selection favoring the mutant
allele in heterozygous cells. Proceedings of the National
Academy of Sciences. 1981;78(8):5066-70.
52. Oberle I, Drayna D, Camerino G, White R, Mandel J-L. The
telomeric region of the human X chromosome long arm:
presence of a highly polymorphic DNA marker and analysis
of recombination frequency. Proceedings of the National
Academy of Sciences. 1985;82(9):2824-8.
53. Aubourg PR, Sack GH, Meyers DA, Lease JJ, Moser HW.
Linkage of adrenoleukodystrophy to a polymorphic DNA
probe. Annals of neurology. 1987;21(4):349-52.
54. Dvorakova L, Storkanova G, Unterrainer G, Hujová
J, Kmoch S, Zeman J, et al. Eight novel ABCD1 gene
mutations and three polymorphisms in patients with X-linked
adrenoleukodystrophy: The first polymorphism causing an
amino acid exchange. Human mutation. 2001;18(1):52.
55. Berger J, Molzer B, Fae I, Bernheimer H. X-linked
adrenoleukodystrophy (ALD): a novel mutation of the
ALD gene in 6 members of a family presenting with 5
different phenotypes. Biochemical and biophysical research
communications. 1994;205(3):1638-43.
56. Smith KD, Kemp S, Braiterman LT, Lu J-F, Wei H-M,
Geraghty M, et al. X-linked adrenoleukodystrophy: genes,
mutations, and phenotypes. Neurochemical research.
1999;24(4):521-35.
57. Kemp S, Ligtenberg MJ, Vangeel BM, Barth PG, Wolterman
RA, Schoute F, et al. Identification of a two base pair deletion
in five unrelated families with adrenoleukodystrophy: a
possible hot spot for mutations. Biochemical and biophysical
research communications. 1994;202(2):647-53.
58. Korenke GC, Wilichowski E, Hunneman DH, Hanefeld F,
Fuchs S, Krasemann E, et al. Cerebral adrenoleukodystrophy
(ALD) in only one of monozygotic twins with an identical
ALD genotype. Annals of neurology. 1996;40(2):254-7.
59. Yatsenko A, Shroyer N, Lewis R, Lupski J. Late-onset
Stargardt disease is associated with missense mutations that
map outside known functional regions of ABCR (ABCA4).
Human genetics. 2001;108(4):346-55.
60. Kok F, Neumann S, Sarde CO, Zheng S, Wu KH, Wei
HM, et al. Mutational analysis of patients with X-linked
adrenoleukodystrophy. Human mutation. 1995;6(2):104-15.
61. Eichler EE, Budarf ML, Rocchi M, Deaven LL, Doggett
NA, Baldini A, et al. Interchromosomal duplications
of the adrenoleukodystrophy locus: a phenomenon of
pericentromeric plasticity. Human molecular genetics.
1997;6(7):991-1002.
62. Antonarakis S, Van Aelst L. Mind the GAP, Rho, Rab and
GDI. Nature genetics. 1998;19(2):106-8.
63. Gedeon AK, Donnelly AJ, Mulley JC, Kerr B, Turner G.
Letter to the editor: How many X-linked genes for non-
specific mental retardation (MRX) are there? American
journal of medical genetics. 1996;64(1):158-62.
64. Nobes CD, Hall A. Rho, rac, and cdc42 GTPases regulate
the assembly of multimolecular focal complexes associated
with actin stress fibers, lamellipodia, and filopodia. Cell.
1995;81(1):53-62.
65. Kaufmann N, Wills ZP, Van Vactor D. Drosophila
Rac1 controls motor axon guidance. Development.
1998;125(3):453-61.
66. Luo L, Jan LY, Jan Y-N. Rho family small GTP-binding
proteins in growth cone signalling. Current opinion in
neurobiology. 1997;7(1):81-6.
67. Threadgill R, Bobb K, Ghosh A. Regulation of dendritic
growth and remodeling by Rho, Rac, and Cdc42. Neuron.
1997;19(3):625-34.
68. Gu XX, Decorte R, Marynen P, Fryns J-P, Cassiman J-J,
Raeymaekers P. Localisation of a new gene for non-specific
mental retardation to Xq22-q26 (MRX35). Journal of medical
genetics. 1996;33(1):52-5.
69. Ryan SG, Chance PF, Zou C-H, Spinner NB, Golden JA,
Smietana S. Epilepsy and mental retardation limited to
females: an X-linked dominant disorder with male sparing.
Nature genetics. 1997;17(1):92-5.
70. Chen LY, Rex CS, Babayan AH, Kramár EA, Lynch G, Gall
CM, et al. Physiological activation of synaptic Rac> PAK
(p-21 activated kinase) signaling is defective in a mouse
model of fragile X syndrome. The Journal of Neuroscience.
2010;30(33):10977-84.
71. Hayashi ML, Rao BS, Seo J-S, Choi H-S, Dolan BM,
Choi S-Y, et al. Inhibition of p21-activated kinase rescues
symptoms of fragile X syndrome in mice. Proceedings of
the national academy of sciences. 2007;104(27):11489-94.
72. Coffee B, Keith K, Albizua I, Malone T, Mowrey J, Sherman
SL, et al. Incidence of fragile X syndrome by newborn
screening for methylated FMR1 DNA. The American Journal
of Human Genetics. 2009;85(4):503-14.
73. Peprah E. Fragile X syndrome: the FMR1 CGG repeat
distribution among world populations. Annals of human
genetics. 2012;76(2):178-91.
74. Verkerk AJ, Pieretti M, Sutcliffe JS, Fu Y-H, Kuhl DP,
Pizzuti A, et al. Identification of a gene (FMR-1) containing
a CGG repeat coincident with a breakpoint cluster region
exhibiting length variation in fragile X syndrome. Cell.
1991;65(5):905-14.
75. Ma Q-L, Yang F, Frautschy SA, Cole GM. PAK in Alzheimer
disease, Huntington disease and X-linked mental retardation.
Cellular logistics. 2012;2(2):117-25.
76. Bassell GJ, Warren ST. Fragile X syndrome: loss of local
mRNA regulation alters synaptic development and function.
Neuron. 2008;60(2):201-14.
77. De Rubeis S, Bagni C. Fragile X mental retardation protein
control of neuronal mRNA metabolism: Insights into
mRNA stability. Molecular and Cellular Neuroscience.
2010;43(1):43-50.
78. Bassell GJ, Gross C. Reducing glutamate signaling pays off
in fragile X. Nature medicine. 2008;14(3):249-50.
79. Antar L, Dictenberg J, Plociniak M, Afroz R, Bassell
G. Localization of FMRP-associated mRNA granules
and requirement of microtubules for activity-dependent
trafficking in hippocampal neurons. Genes, Brain and
Behavior. 2005;4(6):350-9.
Skewed X-chromosome Inactivation and Neurological Disorders—Taherian et al
91
International Clinical Neuroscience Journal Vol 3, No 2, Spring 2016
80. Baker K, Wray S, Ritter R, Mason S, Lanthorn T, Savelieva
K. Male and female Fmr1 knockout mice on C57 albino
background exhibit spatial learning and memory impairments.
Genes, Brain and Behavior. 2010;9(6):562-74.
81. Mineur YS, Huynh LX, Crusio WE. Social behavior deficits
in the Fmr1 mutant mouse. Behavioural brain research.
2006;168(1):172-5.
82. Peier AM, McIlwain KL, Kenneson A, Warren ST, Paylor
R, Nelson DL. (Over) correction of FMR1 deficiency with
YAC transgenics: behavioral and physical features. Human
Molecular Genetics. 2000;9(8):1145-59.
83. Nomura Y. Early behavior characteristics and sleep
disturbance in Rett syndrome. Brain and Development.
2005;27:S35-S42.
84. Amir RE, Van den Veyver IB, Wan M, Tran CQ, Francke
U, Zoghbi HY. Rett syndrome is caused by mutations in
X-linked MECP2, encoding methyl-CpG-binding protein
2. Nature genetics. 1999;23(2):185-8.
85. Trappe R, Laccone F, Cobilanschi J, Meins M, Huppke P,
Hanefeld F, et al. MECP2 mutations in sporadic cases of
Rett syndrome are almost exclusively of paternal origin. The
American Journal of Human Genetics. 2001;68(5):1093-101.
86. Christodoulou J, Grimm A, Maher T, Bennetts B. RettBASE:
the IRSA MECP2 variation database—a new mutation
database in evolution. Human mutation. 2003;21(5):466-72.
87. Archer HL, Whatley SD, Evans JC, Ravine D, Huppke P,
Kerr A, et al. Gross rearrangements of the MECP2 gene are
found in both classical and atypical Rett syndrome patients.
Journal of medical genetics. 2006;43(5):451-6.
88. Pan H, Li MR, Nelson P, Bao XH, Wu XR, Yu S. Large
deletions of the MECP2 gene in Chinese patients with
classical Rett syndrome. Clinical genetics. 2006;70(5):418-9.
89. Smeets E, Terhal P, Casaer P, Peters A, Midro A, Schollen E,
et al. Rett syndrome in females with CTS hot spot deletions:
a disorder profile. American Journal of Medical Genetics
Part A. 2005;132(2):117-20.
90. Neul J, Fang P, Barrish J, Lane J, Caeg E, Smith E, et
al. Specific mutations in methyl-CpG-binding protein 2
confer different severity in Rett syndrome. Neurology.
2008;70(16):1313-21.
91. Kerr A, Archer H, Evans J, Prescott R, Gibbon F. People
with MECP2 mutation-positive Rett disorder who converse.
Journal of Intellectual Disability Research. 2006;50(5):386-
94.
92. Jorde LB, Hasstedt S, Ritvo E, Mason-Brothers A, Freeman
B, Pingree C, et al. Complex segregation analysis of autism.
American journal of human genetics. 1991;49(5):932.
93. Bailey A, Le Couteur A, Gottesman I, Bolton P, Simonoff
E, Yuzda E, et al. Autism as a strongly genetic disorder:
evidence from a British twin study. Psychological medicine.
1995;25(01):63-77.
94. Harding AE, Thomas PK. The clinical features of hereditary
motor and sensory neuropathies type I and II. Brain 1980;
103:259±280.
95. England JD, Garcia CA. Electrophysiological studies in
the different genotypes of Charcot-Marie-Tooth disease.
Curr Opin Neurol 1996;9:338±342.
96. McKusick, V. A. (1990). Mendelian Inheritance in Man
(Baltimore: The Johns Hopkins UniversityPress).
97. Lupski JR, Montes de Oca-Luna R, Slaugenhaupt S, Pentao
L, Guzzetta V, Trask BJ, Saucedo-Cardenas O, Barker
DF, Killian JM, Garcia CA, Chakravarti A, Patel PI. DNA
duplication as-sociated with Charcot-Marie-Tooth disease
type 1A. Cell 1991; 66:219±232.
98. Vance JM, Barker D, Yamaoka LH, Stajich JM, Loprest L,
Hung WY, et al. Localization of Charcot-Marie-Tooth disease
type 1a (CMT1A) to chromosome 17p11.2. Genomics.
1991;9(4):623-8.
99. Nicholson G, Corbett A. Slowing of central conduction
in X-linked Charcot-Marie-Tooth neuropathy shown by
brainstem auditory evoked responses. J Neurol Neurosurg
Psychiatry 1996;61:43-46.
100. Ionasescu W, Ionasescu R, Searby C. Screening of dominantly
inherited Charcot-Marie-Tooth neuropathies. Muscle Nerve
1993;16:1232-1238.
101. Huttner IG, Kennerson ML, Reddel SW, Radovanovic D,
Nicholson GA. Proof of genetic heterogeneity in X-linked
Charcot-Marie-Tooth disease. Neurology. 2006;67(11):2016-
21.
102. Dermietzel R, Spray DC. Gap-junctions in the brain: where,
what type, how many and why? Trends Neurosci 1993;16:
186 -192.
103. Nicholson G, Nash J. Intermediate nerve conduction
velocities define X-linked Charcot-Marie-Tooth neuropathy
families. Neurology 1996;43:2558-2564.
... Among these, it regulates the differentiation of cell lineages within the brain during embryogenesis [51]. It contributes to X chromosome inactivation [101], which may account for the gender bias of some inherited CNS disorders with male predominance, such as ASD [102]. Alterations of SETDB1 or its targeted histone substrate have been associated with the pathogenesis of several diseases of the CNS, including ASD [51,[103][104][105]. ...
Article
Full-text available
Human endogenous retroviruses (HERVs) are relics of ancestral infections and represent 8% of the human genome. They are no longer infectious, but their activation has been associated with several disorders, including neuropsychiatric conditions. Enhanced expression of HERV-K and HERV-H envelope genes has been found in the blood of autism spectrum disorder (ASD) patients, but no information is available on syncytin 1 (SYN1), SYN2, and multiple sclerosis-associated retrovirus (MSRV), which are thought to be implicated in brain development and immune responses. HERV activation is regulated by TRIM28 and SETDB1, which are part of the epigenetic mechanisms that organize the chromatin architecture in response to external stimuli and are involved in neural cell differentiation and brain inflammation. We assessed, through a PCR realtime Taqman amplification assay, the transcription levels of pol genes of HERV-H, -K, and -W families, of env genes of SYN1, SYN2, and MSRV, as well as of TRIM28 and SETDB1 in the blood of 33 ASD children (28 males, median 3.8 years, 25–75% interquartile range 3.0–6.0 y) and healthy controls (HC). Significantly higher expressions of TRIM28 and SETDB1, as well as of all the HERV genes tested, except for HERV-W-pol, were found in ASD, as compared with HC. Positive correlations were observed between the mRNA levels of TRIM28 or SETDB1 and every HERV gene in ASD patients, but not in HC. Overexpression of TRIM28/SETDB1 and several HERVs in children with ASD and the positive correlations between their transcriptional levels suggest that these may be main players in pathogenetic mechanisms leading to ASD.
... We already reported that several X-located genes are implicated in immunity (such as Toll-like receptors, cytokine receptors, genes related to T-cell and B-cell activity, and regulatory factors), and some of them are responsible for X-linked primary immunodeficiencies [116][117][118][119][120][121][122][123][124][125][126][127][128][129][130][131][132][133][134][135] (Figure 2). ...
Article
Full-text available
Many complex traits or diseases, such as infectious and autoimmune diseases, cancer, xenobiotics exposure, neurodevelopmental and neurodegenerative diseases, as well as the outcome of vaccination, show a differential susceptibility between males and females. In general, the female immune system responds more efficiently to pathogens. However, this can lead to over-reactive immune responses, which may explain the higher presence of autoimmune diseases in women, but also potentially the more adverse effects of vaccination in females compared with in males. Many clinical and epidemiological studies reported, for the SARS-CoV-2 infection, a gender-biased differential response; however, the majority of reports dealt with a comparable morbidity, with males, however, showing higher COVID-19 adverse outcomes. Although gender differences in immune responses have been studied predominantly within the context of sex hormone effects, some other mechanisms have been invoked: cellular mosaicism, skewed X chromosome inactivation, genes escaping X chromosome inactivation, and miRNAs encoded on the X chromosome. The hormonal hypothesis as well as other mechanisms will be examined and discussed in the light of the most recent epigenetic findings in the field, as the concept that epigenetics is the unifying mechanism in explaining gender-specific differences is increasingly emerging.
Article
Full-text available
X-linked adrenoleukodystrophy (X-ALD) is an inherited neurodegenerative disorder caused by mutations in the gene. Approximately 20% of X-ALD female carriers may develop neurological symptoms. Skewed X chromosome inactivation (XCI) has been proposed to influence the manifestation of symptoms in X-ALD carriers, but data remain conflicting so far. We identified a three generation kindred, with five heterozygous females, including two manifesting carriers. XCI pattern and the allele expression were assessed in order to determine if symptoms in X-ALD carriers could be related to skewed XCI and whether skewing within this family is more consistent with genetically influenced or completely random XCI. We found a high frequency of skewing in this family. Four of five females had skewed XCI, including two manifesting carriers favoring the mutant allele, one asymptomatic carrier favoring the normal allele, and one female who was not an X-ALD carrier. Known causes of skewing, such as chromosomal abnormalities, selection against deleterious alleles, promoter mutations, were not consistent with our results. Our data support that skewed XCI in favor of the mutant allele would be associated with the manifestation of heterozygous symptoms. Furthermore, XCI skewing in this family is genetically influenced. However, the underlying mechanism remains to be substantiated by further experiments.
Article
Full-text available
Developmental cognitive deficits including X-linked mental retardation (XLMR) can be caused by mutations in P21-activated kinase 3 (PAK3) that disrupt actin dynamics in dendritic spines. Neurodegenerative diseases such as Alzheimer disease (AD), where both PAK1 and PAK3 are dysregulated, may share final common pathways with XLMR. Independent of familial mutation, cognitive deficits emerging with aging, notably AD, begin after decades of normal function. This prolonged prodromal period involves the buildup of amyloid-β (Aβ) extracellular plaques and intraneuronal neurofibrillary tangles (NFT). Subsequently region dependent deficits in synapses, dendritic spines and cognition coincide with dysregulation in PAK1 and PAK. Specifically proximal to decline, cytoplasmic levels of actin-regulating Rho GTPase and PAK1 kinase are decreased in moderate to severe AD, while aberrant activation and translocation of PAK1 appears around the onset of cognitive deficits. Downstream to PAK1, LIM kinase inactivates cofilin, contributing to cofilin pathology, while the activation of Rho-dependent kinase ROCK increases Aβ production. Aβ activation of fyn disrupts neuronal PAK1 and ROCK-mediated signaling, resulting in synaptic deficits. Reductions in PAK1 by the anti-amyloid compound curcumin suppress synaptotoxicity. Similarly other neurological disorders, including Huntington disease (HD) show dysregulation of PAKs. PAK1 modulates mutant huntingtin toxicity by enhancing huntingtin aggregation, and inhibition of PAK activity protects HD as well as fragile X syndrome (FXS) symptoms. Since PAK plays critical roles in learning and memory and is disrupted in many cognitive disorders, targeting PAK signaling in AD, HD and XLMR may be a novel common therapeutic target for AD, HD and XLMR.
Article
Full-text available
Approximately 20% of adrenoleukodystrophy (X-ALD) female carriers may develop clinical manifestations, typically consisting of progressive spastic gait, sensory deficits and bladder dysfunctions. A skewing in X Chromosome Inactivation (XCI), leading to the preferential expression of the X chromosome carrying the mutant ABCD1 allele, has been proposed as a mechanism influencing X-linked adrenoleukodystrophy (X-ALD) carrier phenotype, but reported data so far are conflicting. To shed light into this topic we assessed the XCI pattern in peripheral blood mononuclear cells (PBMCs) of 30 X-ALD carriers. Since a frequent problem with XCI studies is the underestimation of skewing due to an incomplete sample digestion by restriction enzymes, leading to variable results, we developed a pyrosequencing assay to identify samples completely digested, on which to perform the XCI assay. Pyrosequencing was also used to quantify ABCD1 allele-specific expression. Moreover, very long-chain fatty acid (VLCFA) levels were determined in the same patients. We found severely (≥90:10) or moderately (≥75:25) skewed XCI in 23 out of 30 (77%) X-ALD carriers and proved that preferential XCI is mainly associated with the preferential expression of the mutant ABCD1 allele, irrespective of the manifestation of symptoms. The expression of mutant ABCD1 allele also correlates with plasma VLCFA concentrations. Our results indicate that preferential XCI leads to the favored expression of the mutant ABCD1 allele. This emerges as a general phenomenon in X-ALD carriers not related to the presence of symptoms. Our data support the postulated growth advantage of cells with the preferential expression of the mutant ABCD1 allele, but argue against the use of XCI pattern, ABCD1 allele-specific expression pattern and VLCFA plasma concentration as biomarkers to predict the development of symptoms in X-ALD carriers.
Article
Utilizing the plasma very long chain fatty acid assay, supplemented by mutation analysis and immunofluorescence assay, we determined the number of X-linked adrenoleukodystrophy (X-ALD) hemizygotes from the United States identified each year in the two laboratories that perform most of the assays in this country: the Kennedy Krieger Institute between 1981 and 1998 and the Mayo Clinic Rochester from 1996 to 1998. The minimum frequency of hemizygotes identified in the United States is estimated to be 1:42,000 and that of hemizygotes plus heterozygotes 1:16,800. Our studies involved 616 pedigrees with a total of 12,787 identified at-risk members. Diagnostic assays were performed in 4,169 at-risk persons (33%) and included members of the extended family. Only 5% of male probands and 1.7% of X-ALD hemizygotes were found to have new mutations. The extended family testing led to the identification of 594 hemizygotes and 1,270 heterozygotes. Two hundred fifty of the newly identified hemizygotes were asymptomatic and represent the group in which therapy has the greatest chance of success. Identification of heterozygotes provides the opportunity for disease prevention through genetic counseling. Diagnostic tests should be offered to all at-risk relatives of X-ALD patients and should include members of the extended family. Ann Neurol 2001;49:512–517
Article
X-linked adrenoleukodystrophy (X-ALD) is caused by mutations in the ABCD1 gene, which encodes a peroxisomal ABC half-transporter (ALDP) involved in the import of very long-chain fatty acids (VLCFA) into the peroxisome. The disease is characterized by a striking and unpredictable variation in phenotypic expression. Phenotypes include the rapidly progressive childhood cerebral form (CCALD), the milder adult form, adrenomyeloneuropathy (AMN), and variants without neurologic involvement. There is no apparent correlation between genotype and phenotype. In males, unambiguous diagnosis can be achieved by demonstration of elevated levels of VLCFA in plasma. In 15 to 20% of obligate heterozygotes, however, test results are false–negative. Therefore, mutation analysis is the only reliable method for the identification of heterozygotes. Since most X-ALD kindreds have a unique mutation, a great number of mutations have been identified in the ABCD1 gene in the last seven years. In order to catalog and facilitate the analysis of these mutations, we have established a mutation database for X-ALD ( http://www.x-ald.nl). In this review we report a detailed analysis of all 406 X-ALD mutations currently included in the database. Also, we present 47 novel mutations. In addition, we review the various X-ALD phenotypes, the different diagnostic tools, and the need for extended family screening for the identification of new patients. Hum Mutat 18:499–515, 2001. © 2001 Wiley-Liss, Inc.
Article
A 9.7 kb segment encompassing exons 7-10 of the adrenoleukodystrophy (ALD) locus of the X chromosome has duplicated to specific locations near the pericentromeric regions of human chromosomes 2p11,10p11, 16p11 and 22q11. Comparative sequence analysis reveals 92-96% nucleotide identity, indicating that the autosomal ALD paralogs arose relatively recently during the course of higher primate evolution (5-10 million years ago). Analysis of sequences flanking the duplication region identifies the presence of an unusual GCTTTTTGC repeat which may be a sequence-specific integration site for the process of pericentromeric-directed transposition. The breakpoint sequence and phylogenetic analysis predict a two-step transposition model, in which a duplication from Xq28 to pericentromeric 2p11 occurred once, followed by a rapid distribution of a larger duplicon cassette among the pericentromeric regions. In addition to facilitating more effective mutation detection among ALD patients, these findings provide further insight into the molecular basis underlying a pericentromeric-directed mechanism for non-homologous interchromosomal exchange.
Article
Fragile X syndrome Is the most frequent form of inherited mental retardation and Is associated with a fragile site at Xg27.3. We identified human YAC clones that span fragile X site-induced translocation breakpoints coincident with the fragile X site. A gene (FMR-1) was identified within 8 four cosmid contig of YAC DNA that expresses a 4.8 kb message in human brain. Within a 7.4 kb EcoFII genomic fragment, containing FMR-1 exonic sequences distal to a CpG island previously shown to be hypermethylated in fragile X patients, is a fragile X site-induced breakpoint cluster region that exhibits length variation in fragile X chromosomes. This fragment contains a lengthy CGG repeat that is 250 by distal of the CpG island and maps within a FMR-1 axon. Localization of the brain-expressed FMR-1 gene to this EcoRl fragment suggests the involvement of this gene in the phenotypic expression of the fragile X syndrome.
Article
Charcot-Marie-Tooth (CMT) disease type 1a has been previously localized to chromosome 17 using the markers D17S58 and D17S71. In that report we were unable to provide unequivocal localization of the CMT1A gene on either the proximal p or the q arm. Therefore, data from one additional CMT1A family and typing of other probes spanning the pericentromeric region of chromosome 17 (D17S73, D17S58, D17S122, D17S125, D17S124) were analyzed. Multipoint analysis demonstrates convincing evidence (log likelihood difference > 5) that the CMT1A gene lies within 17p11.2 and most likely between the flanking markers D17S122 and D17S124.
Article
Equalization of X linked gene expression is necessary in mammalian cells due to the presence of two X chromosomes in females and one in males. To achieve this, all female cells inactivate one of the two X chromosomes during development. This process, termed X chromosome inactivation (XCI), is a quintessential epigenetic phenomenon and involves a complex interplay between noncoding RNAs and protein factors. Progress in this area of study has consequently resulted in new approaches to study epigenetics and regulatory RNA function. Here we will discuss recent developments in the field that have advanced our understanding of XCI and its regulatory mechanisms.