Next Article in Journal / Special Issue
Intercalated Cells: More than pH Regulation
Previous Article in Journal / Special Issue
Pathological Mutations of the Mitochondrial Human Genome: the Instrumental Role of the Yeast S. cerevisiae
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

MeCP2-Related Diseases and Animal Models

Regenerative Medicine Program, Department of Biochemistry and Medical Genetics, Faculty of Medicine, University of Manitoba, 745 Bannatyne Avenue, Winnipeg, Manitoba, R3E 0J9, Canada
*
Author to whom correspondence should be addressed.
Submission received: 6 December 2013 / Revised: 19 January 2014 / Accepted: 20 January 2014 / Published: 27 January 2014
(This article belongs to the Special Issue Feature Papers)

Abstract

:
The role of epigenetics in human disease has become an area of increased research interest. Collaborative efforts from scientists and clinicians have led to a better understanding of the molecular mechanisms by which epigenetic regulation is involved in the pathogenesis of many human diseases. Several neurological and non-neurological disorders are associated with mutations in genes that encode for epigenetic factors. One of the most studied proteins that impacts human disease and is associated with deregulation of epigenetic processes is Methyl CpG binding protein 2 (MeCP2). MeCP2 is an epigenetic regulator that modulates gene expression by translating epigenetic DNA methylation marks into appropriate cellular responses. In order to highlight the importance of epigenetics to development and disease, we will discuss how MeCP2 emerges as a key epigenetic player in human neurodevelopmental, neurological, and non-neurological disorders. We will review our current knowledge on MeCP2-related diseases, including Rett Syndrome, Angelman Syndrome, Fetal Alcohol Spectrum Disorder, Hirschsprung disease, and Cancer. Additionally, we will briefly discuss about the existing MeCP2 animal models that have been generated for a better understanding of how MeCP2 impacts certain human diseases.

1. Introduction

Methyl CpG binding protein 2 (MeCP2) was first identified and characterized as a DNA binding protein that specifically binds to methyl-CpG dinucleotides [1]. The MECP2 gene is localized on the X-chromosome and contains complex regulatory elements that control its precise expression levels [2,3,4,5]. In both human and mouse, the MECP2/Mecp2 gene consists of four exons encoding for two different protein isoforms, MeCP2E1 and MeCP2E2 [6,7]. The two Mecp2/MeCP2 isoforms show differential temporal and brain region-specific differences in their distribution [7], and we previously showed that MeCP2E1 isoform is similarly distributed compared to total MeCP2 in murine brain cells [8]. The main functional protein domains of MeCP2 include the methyl binding domain (MBD), the transcriptional repression domain (TRD), the C-terminal domain (CTD), and the inter domain (ID) [9,10]. The MBD facilitates binding to methylated CpG dinucleotides and the preference for adjacent A/T-rich motifs [9,11]. It is also capable of binding to non-methylated DNA sequences such as the four-way DNA junctions [12,13]. The TRD domain mediates the transcriptional repressor role of MeCP2, and interacts with co-repressor complexes, such as c-Ski, mSin3A, HDAC1, and HDAC2 [5,9,14,15]. The CTD is suggested to be important for MeCP2 function and its chromatin binding activities [12]. Inter domain interactions significantly impacts MeCP2 structure and increases the stability of the protein [16,17].
MeCP2 is a nuclear protein that is mainly localized to methylated pericentromeric heterochromatin, referred to as chromocenters [18]. MeCP2 is widely expressed in several tissues, and is suggested to mediate transcriptional regulation through association with 5-methyl cytosine (5mC) and 5-hydroxymethlcytosine (5hmC), the two major types of DNA modifications [19,20,21]. MeCP2 binding to 5mC is associated with repressive functions through its interaction with co-repressor complexes, while binding to 5hmC is suggested to facilitate gene expression through the organization of dynamic chromatin [15,22,23,24]. Research studies on MeCP2 have shown diverse functions of this protein including transcriptional repression, transcriptional activation [25,26,27], RNA splicing [28,29], long range chromatin remodeling [30,31], modulation of chromatin architecture, and maintenance of DNA methylation [32]. This diversity in the function of MeCP2 underscores its role in many disorders.
MeCP2 is a multifunctional epigenetic regulator that is involved in transcriptional regulation as well as modulating chromatin structure [17,33]. Epigenetics control gene expression without altering the corresponding DNA sequences and impact development and disease [21,34,35,36]. Epigenetic mechanisms control the expression of many neurodevelopmentally important genes through chromatin remodeling, histone modifications and DNA methylation [37,38,39,40,41]. The prevailing view of MeCP2 as an epigenetic modulator is related to its ability to associate with different epigenetic marks (either through direct DNA binding or recruiting other transcription factors). Thus, MeCP2 acts as an ‘epigenetic reader’ that contributes in the establishment of functional states of chromatin structure, a process that is fundamental for normal cellular function. However, it is still not fully understood whether MeCP2 acts as a genome-wide epigenetic regulator or as a gene-specific transcriptional modulator, and evidence exists in favor of both mechanisms [15,42].
MeCP2 protein is highly expressed in the brain compared to other tissues [18,43,44]. The protein expression pattern within the brain follows a defined pattern, with early appearing structures, such as brainstem and thalamus, expressing the protein before more rostral structures, such as the cortex [18,43]. Temporally, MeCP2 protein expression profile is low at birth, and increases dramatically at specific time periods coinciding with the process of neuronal maturation and synaptogenesis in different parts of the brain [43,45]. Cell type-specific studies have revealed that MeCP2 levels are highest in mature neurons, while astrocytes and immature neurons express lower levels of MeCP2 [8,46,47]. The increased levels of MeCP2 expression in mature neurons are maintained throughout adulthood, implying a requirement for postmitotic neuronal function [47,48,49]. In neurons, MeCP2 is involved in neuronal maturation, dendrite formation, and synaptic functions [15,49,50]. Studies of mouse models lacking Mecp2 expression in neurons further demonstrate the critical role of MeCP2 for normal brain function, especially with regard to synaptic modulation and maintenance [18,51].
Most MeCP2 mutations are de novo and can be grouped into three general categories; severe loss-of-function mutations, mild loss-of-function mutations and a broad group of duplications and other non-coding mutations [52,53]. Each category of mutation is associated with (a) subset(s) of neurological symptoms. Genotype-phenotype analyses have shown that there is no direct and simple correlation with the type of observed mutations and the resulting disease phenotypes. The difficulty in attributing a particular type of mutation to a specific phenotype might be in part due to the pattern of X-chromosome inactivation (XCI) [54,55].
Overall, MeCP2-related diseases that are associated with protein dysfunction are mainly characterized by cognitive impairment and intellectual disabilities [33]. However, dysregulation of the protein are also frequently observed in cases of inherited disorders and autoimmune diseases such as systemic lupus erythematosus (SLE) [56]. In this review, we will primarily focus on human diseases that are associated with MeCP2 dysfunction, and will aim to highlight the role of MeCP2 in neurological/neuropsychiatric and non-neurological disorders. In addition, animal models that have enabled a better understanding of the mechanism of MeCP2 action will be discussed.

2. The Role of MeCP2 in Neurological Disorders

2.1. Rett Syndrome

Rett Syndrome (RTT) is a progressive X-linked neurological disorder that mainly affects young girls with an incidence rate of 1:10,000-1:15,000 [15,57]. RTT is, perhaps, the most common cause of mental retardation in females [58,59]. Although it was initially thought to occur exclusively in females, males have also been identified with classic RTT [60]. In approximately 90% of RTT cases, the disease is due to MECP2 gene mutations [61]. MECP2 mutations that cause classical RTT in females typically result in neonatal encephalopathy and death in the first year of life in males. However, similar MECP2 mutations may result in RTT phenotypes in males with Kleinfelter syndrome (47, XXY), or somatic mosaicism [62,63].
There is the classic RTT as well as atypical forms of RTT that deviate from the classical clinical presentation. Classic RTT is noticed primarily during infancy and can be divided into four stages that reflect the characteristic abnormalities displayed in RTT patients [64]. The first stage takes place after a period of normal development, during the first 6-18 months after birth, when developmental progression ceases and acquisition of new skills is delayed. General motor performance such as crawling, sitting, and walking are also severely impaired at this stage [58,65,66]. The second stage of RTT starts at approximately one to four years, with developmental stagnation accompanied by general growth retardation, loss of purposeful hand movements and speech, tongue protrusion, abnormal facial expression, weight loss, and gait ataxia/apraxia [58,64,67]. Autonomic dysfunction such as irregular breathing patterns, forced expulsion of air and saliva, and apnea can also be observed at this stage [68]. The duration of this stage is from weeks to approximately a year. The third RTT stage is regarded as a relatively ‘quiet’ period as stabilization of some of the symptoms occurs. However, neuromotor regression and stereotypic hand movements still persists [58,64,67]. A defining feature at this stage is the occurrence of seizures, which ranges from easily controlled to intractable epilepsy [69]. The duration of this stage is usually from years to decades [67]. The last RTT stage takes place from ages 5-15 years and beyond. Motor deterioration continues and results in loss of mobility and dependence on wheel chair. Additional abnormalities include dystonia, severe constipation, oropharyngeal dysfunction, and cardiac abnormalities [68]. As patients become older they often develop Parkinson’s-like features [64,68]. The duration of this stage varies with individuals, and usually lasts for decades.
Atypical forms of RTT deviate from classic RTT with respect to age of disease onset, clinical profile and severity of symptoms. These forms of RTT occur due to skewing of XCI, and may range from milder forms to more severe manifestations than classic RTT [70,71]. Mild variants of RTT are characterized by a later age of onset, typically occurring between one to three years of age and display mild stereotypic movements and neurologic symptoms. The preserved speech or the Zappella variant is a mild variant of RTT characterized by the ability of patients to speak a few words. Patients with this variant have a normal head size and are usually overweight [72]. The more severe variants include the congenital form (also known as the Rolando variant) that lack the early period of normal development, and a form of classical RTT with early onset seizures before the age of six months (also known as the Hanefeld variant) [73,74]. Mutations in cyclin-dependent kinase-like 5 (CDKL5) are associated with the early-onset seizure variant form of RTT, while mutations in forkhead box protein G1 (FOXG1) are associated with congenital RTT Syndrome variant [75,76].
Magnetic resonance imaging (MRI) and autopsy examination have shown that major morphological abnormalities detected in the central nervous system of RTT patients include an overall decrease in the size of the brain and of individual neurons [77]. The reduction in brain size is distributed throughout the brain and affects both white, and to a greater extent the grey matter in different brain regions [78,79]. Cortical and cerebellar degeneration have also been demonstrated to progressively occur with increasing age in RTT patients [80].

2.2. MECP2 Duplication Disorder

MECP2 duplication syndrome was predicted by the observation that mice engineered to overexpress MECP2 develop a progressive neurological disorder including stereotyped and repetitive movements, epilepsy, spasticity, hypoactivity, and early death [81,82]. This gain of function mutation occurs mostly in males [83]. Males with this disorder present clinical features, such as infantile hypotonia, severe to profound mental retardation, poor speech development, recurrent infections, epilepsy, and progressive spasticity [84,85]. The incidence rate of MECP2 duplication disorder is estimated to be 1% of unexplained X-linked diseases and most of the reported cases are inherited [86]. However, de novo cases have also been documented [87]. There is considerable clinical overlap in patients with classic RTT and MECP2 duplication disorder specifically in behavioral phenotypes, such as stereotypic hand/body movements, anxiety, and social avoidance [88]. However, in contrast to RTT, immunodeficiency is observed in patients with MECP2 duplication disorder. It remains unclear if this phenotype occurs due to secondary effects from increased dosage of MeCP2 protein [89]. Approximately 40% of males with MECP2 duplication reported so far have died before their 25th birthday, usually from respiratory infections [90].

2.3. Angelman Syndrome

MECP2 mutations that cause RTT have also been reported in cases of Angelman Syndrome (AS) [91,92,93]. Angelman Syndrome is primarily caused by mutations or imprinting errors of the UBE3A gene located on chromosome 15, and is characterized by intellectual disability; severe speech impairment and gait ataxia. Considerable phenotypic overlap exists between AS and RTT; however they differ with respect to timing of symptom onset. Angelman Syndrome has an earlier onset and patients are characterized with low hypotonicity at birth [91].

2.4. X-linked Mental Retardation

X-linked mental retardation (XLMR) is a genetic disorder arising from mutations or duplication of genes across the X chromosome, including the MECP2 gene. MECP2 point mutations have been identified in up to 2% of individuals with XLMR and duplications of the gene are also implicated in approximately 1% to 2% of XLMR cases [86,94]. MECP2 mutations that cause RTT or severe neonatal encephalopathy are not identified in XLMR patients and vice versa [86,95]. In addition, the molecular lesions underlying MECP2 duplications that result in XLMR are different from those observed in MECP2 duplication syndrome [96]. Males with XLMR show phenotypes, such as severe intellectual disability, speech impairment, and motor abnormalities, whereas females display mild intellectual disability or are unaffected [86].

2.5. Severe Neonatal Encephalopathy

Males with a normal karyotype and a mutation in the MECP2 gene present with a distinct clinical condition, and severe neonatal encephalopathy (SNE) [97]. Severe neonatal encephalopathy is a disorder characterized by a static encephalopathy, severe developmental delays and respiratory abnormalities. The mutations are usually inherited from mothers with favorable XCI skewing that display mild/no RTT symptoms. Males with SNE often die within the first years of their life due to autonomic dysfunction [95,98]. Some MECP2 mutations that do not cause RTT in females have also been implicated in moderate to profound mental retardation, deficits in language and motor skills, obesity, autistic features, and psychiatric disorders in males [54].

2.6. Autism

Autism and RTT are similar developmental disorders that belong to the spectrum of autism disorders classified as pervasive developmental disorders, or Autism Spectrum Disorders (ASD) [99]. Although they are similar developmental disorders, autism differs from RTT with respect to its genetic basis. While RTT is caused by MECP2 mutations, the genetic basis of autism is not fully clear and is proposed to involve multiple genes [100]. Mutations in the MECP2 regulatory elements (resulting in decreased expression of the protein) are commonly associated with autism [101]. Reduced expression of MeCP2 protein has been shown to occur frequently in the frontal cortex of autistic patients and is correlated with increased MECP2 promoter DNA methylation [102]. The silencing of autism-related genes through promoter DNA hypermethylation is commonly associated with autism, and drugs that can demethylate promoters might be useful in activating these genes [103,104]. In a recent study, we show that reduced DNA demethylation at the Mecp2 promoter and intron 1 regulatory elements treated with Decitabine is associated with increased Mecp2 expression. Our results provide insight on use of such drugs for future therapeutic interventions of autism [105]. Moreover, MECP2 mutations that are associated with classic RTT have been identified in a number of autistic females who do not meet the diagnostic criteria for RTT [100]. This makes it difficult to determine if a MECP2 mutation that is associated with autism diagnosis is a different disorder from RTT, or if both disorders are simply different representations on a spectrum associated with MECP2 mutations. MECP2 mutations have also been reported in patients with mild cognitive and motor difficulties and early-onset schizophrenia [106].

2.7. Fetal Alcohol Spectrum Disorders

Prenatal exposure to alcohol is associated with adverse effects on neurodevelopment, and results in a set of severe neurodevelopmental disorders known as fetal alcohol spectrum disorders (FASD) [107,108]. The incidence of FASD is estimated to be as high as 1 in 100 births and this disorder commonly manifests as cognitive and intellectual disabilities [108]. Accumulating evidence-implicating MeCP2 in FASD pathogenesis further reinforces the critical role of MeCP2 for central nervous system function. Several studies have demonstrated aberrant expression levels of MeCP2 in rodent FASD models, and this is suggested to be an important epigenetic determinant in FASD [109,110]. For example, prenatal exposure to ethanol has been demonstrated to significantly decrease MeCP2 expression in both prefrontal cortex and striatum of rodent offspring [109]. Research in animal models suggests that the global epigenetic changes due to ethanol are related to variations in the levels, duration as well as time of exposure [108,111]. Further supporting the potential role of MeCP2 in FASD, RTT-causing mutations have been reported in a FASD patient [112]. MeCP2 has been also shown to modulate the alcohol intake and sensitivity to alcohol, demonstrating the role of MeCP2 in alcoholism [113].

2.8. Huntington’s Disease

Huntington's disease (HD) is a progressive neurodegenerative disorder that affects muscle coordination and results in cognitive decline and psychiatric disorders [114,115]. It is one of several tri-nucleotide repeat disorders caused by the length of a repeated section of a gene exceeding a normal range. Expansion of a CAG triplet repeat stretch within the Huntingtin gene (HTT) results in a mutant form of the protein, which gradually damages brain cells [114,116]. Transcriptional dysregulation has been suggested to play major roles in HD pathology, and it was recently demonstrated that the huntingtin protein (Htt) directly interacts with MeCP2 in mouse and cellular models of HD. Aberrant interactions between Htt and MeCP2 is suggested to contribute to aberrant transcription in Huntington’s disease by regulating brain-derived neurotrophic factor (BDNF) levels [117].
The implication of MeCP2 in such diverse range of neurological disorders necessitates a complete understanding of its relationship to brain development and function, as well as its interaction with other epigenetic factors that mediate dysregulation of normal epigenetic program of the brain.

3. Non-Neurological Disorders Associated with MeCP2

3.1. Cancer

Apart from to its role in neurological disorders, MeCP2 has also been shown to play a role in many cancers such as breast, colorectal, lung, liver, and prostate cancer [118,119,120,121,122,123]. MeCP2 role in cancer is related to the epigenetic regulation of cancer-related genes, particularly mechanisms that involve hypermethylation of gene promoters [118]. The growth-promoting role of MeCP2 in prostate cancer cells has been demonstrated previously, where it was shown to control mechanisms, such as cell proliferation and apoptosis [119]. In gastric cancer, the depth of invasion has been shown to be associated with MeCP2 protein levels. In gastric carcinoma cells, microRNA miR-212 was shown to suppress translation of MECP2 transcripts, which in turn resulted in reduced depth of cellular invasion [124]. In addition, MeCP2 has been linked to other cancers, such as myeloma [125], hematological malignancies [126], ductal carcinomas [127], and cervical cancers [128].

3.2. Systemic Lupus Erythematosus

Systemic lupus erythematosus (SLE) is a systemic autoimmune disease that affects multiple organs. The disease predominantly affects females, with a female to male ratio that ranges from 4.3–13.6 to 1 [56,129]. Although the cause of SLE is still not clear, evidence supports an important role for abnormal T cell DNA methylation in the pathogenesis of SLE, as methylation sensitive genes show increased expression in T cells of SLE patients [130]. In active SLE T-cells, the expression of DNA methyltransferase 1 (DNMT1), the main enzyme that maintains DNA methylation during cell division, is reduced, resulting in promoter hypomethylation of these methylation-sensitive genes [56,131,132]. MeCP2 is suggested to play an important role in this process as it is critical for the epigenetic regulation of methylation-sensitive genes, and genetic polymorphisms in MECP2 have also been identified in patients with SLE [56,133,134]. Moreover, MeCP2 associates with DNMT1, the association of which is required to maintain DNA methylation [32]. The association between MeCP2, DNMT, and methylation-sensitive genes suggests an important role for epigenetic regulation in the pathogenesis of SLE and other autoimmune diseases.

3.3. Rheumatoid Arthritis

Rheumatoid arthritis (RA) is a systemic autoimmune disease that results in chronic inflammation and destruction of many tissues and organs, primarily flexible synovial joints [135,136,137]. It is characterized by the presence of auto antibodies, which may be detected in the blood long before the onset of disease [138,139]. Increasing evidence suggests that DNA methylation and histone modifications regulate the progression of Rheumatoid arthritis. Interestingly, it has been shown that MeCP2 expression levels were up-regulated in rodent models of RA, and it is hypothesized that the increased MeCP2 protein levels play a role in the pathogenesis of RA through the canonical Wnt pathway [140,141].

3.4. Hirschsprung's Disease

Hirschsprung’s disease (HSCR) is a congenital disorder of the colon characterized by the absence of certain nerve cells known as ganglion cells [142,143]. The incidence of HSCR is about 1:2000–5000, with males being affected 4 times more frequently than females [144]. The lack of ganglion cells is associated with impaired craniocaudal migration of neural crest cells, and results in severe constipation or intestinal obstruction [145,146]. Defects in the differentiation of neuroblasts into ganglion cells may also contribute to the disorder [142]. A recent study implicating MeCP2 in the pathogenesis of HSCR suggests that aberrant reduced levels of MeCP2 may play an important role in suppressing the proliferative ability of cells in patients with Hirschsprung’s disease [147]. Table 1 represents a list of MeCP2-associated diseases and the genders that are mostly affected.
Table 1. Human diseases associated with Methyl CpG binding protein 2 (MeCP2) and gender mostly affected.
Table 1. Human diseases associated with Methyl CpG binding protein 2 (MeCP2) and gender mostly affected.
DiseaseGender mostly affectedReferences
Rett SyndromeFemales (also males with kleinfelter syndrome 47 XXY, or somatic mosaicism)[15,62,63,148]
MECP2 duplication disorderMales[81,82,83]
Angelman SyndromeFemales[91,93]
X-linked mental retardationMales[86,94,95]
Severe neonatal encephalopathyMales[95,97,98]
AutismBoth[100,101,102]
Fetal alcohol spectrum disordersBoth (Studies from animal models only)[109,110,113]
Huntington’s diseaseBoth[117]
Early-onset schizophreniaBoth[106]
CancersBoth[119,120,121,122,123,124,125,126,127,128,149]
Systemic lupus erythematosusFemales[56,133,134]
Rheumatoid arthritisFemales[140,141]
Hirschsprung’s diseaseMales[147]

4. Animal Models of MeCP2 Dysfunction

Several animal models have been generated in order to better understand the molecular mechanisms, progression and pathology of disorders that are associated with MeCP2 dysfunction. These include mostly rodent models (particularly mouse models) as well as zebrafish and Drosophila models. Different strategies are often employed to alter the expression and function of MeCP2 in these animal models.

4.1. Mecp2 Null Mouse Models

Mecp2 knockout mice have been generated using the Cre recombinase-loxP system. These mice appear normal until approximately three to four weeks of age, when they begin to exhibit behavioral phenotypes such as unusual gait, hindlimb clasping, seizures, tremors, anxiety, learning and memory deficits, and irregular breathing. Brain neuropathology observed in Mecp2 knockout mice is similar to that observed in RTT individuals. Mecp2 knockout mice have smaller densely packed neurons, reduced dendritic spine density, deficits in axonal fasciculation and reduced number of mature synapses. Mecp2 knockout mice die at approximately 10 weeks of life [150,151,152]. The reactivation of Mecp2 in the brains of Mecp2 null mice rescued many RTT-phenotypes seen in the null mice, including delayed disease progression, increased life span, deceased mortality, and restored neurological functions [153,154].

4.2. Mecp2 Mutant Mice

Mecp2 mutant mice differ from the above-mentioned knockout mice as they express a truncated/mutant MeCP2 protein. One of such mutant mice is the Jaenisch strain (Mecp2tm1.1jae), generated by deletion of exon 3 of the Mecp2 gene. Jaenisch mice display behavioral phenotypes similar to the Mecp2 knockout mice, although the phenotype is slightly milder and the lifespan is longer (12 weeks) [82,155]. Another Mecp2 mutant mouse model has been generated that expresses a protein truncated at amino acid 308 (Mecp2308). It is assumed that the truncated MeCP2 protein produced in these mice has residual unknown functions. Mecp2308 mice appear normal until approximately six weeks of age when they develop similar but milder neurological phenotype as Mecp2 knockout mice [156].

4.3. Mecp2 Conditional-Mutant Mice

Brain-specific deletion of Mecp2 in neural precursor cells, beginning at embryonic day 12 (Nestin-cre line), results in mice that display phenotypes similar to Mecp2 knockout mice [150,155]. This suggests that MeCP2 dysfunction in the CNS is sufficient to cause phenotypes observed in RTT. Selective loss of Mecp2 in hypothalamic and amygdala neurons (Sim1-cre line) results in mice that display similar abnormal physiological stress response as that observed in brain-specific conditional mice. These mice were also obese and aggressive, suggesting a role for MeCP2 in regulating social and feeding behaviors [157]. Deletion of Mecp2 in other parts of the brain such as in brainstem neurons (TH-cre line) significantly affects locomotor activity, with no effect on social interaction, breathing patterns, learning and memory [158,159].

4.4. Mice Overexpressing MECP2

Mice overexpressing human MECP2 at approximately twice the endogenous levels exhibit delayed neurological symptoms at approximately 10 weeks. These symptoms include enhanced motor learning and synaptic plasticity in the hippocampus. However at 20 weeks, these mice develop seizures and become hypoactive and the majority of them die by approximately one year of age [81]. Increased MeCP2 expression in other transgenic lines also results in motor abnormalities and neurological symptoms [160]. These transgenic mice mimic MECP2 duplication syndrome that is observed in humans, and reinforces the notion of a critical requirement for precise dosage of MeCP2 protein.

4.5. Mouse Models Carrying Rett Syndrome-Associated Mutations

A mouse model of Rett syndrome was generated by introducing a premature STOP codon at the amino acid position 168, and resulted in a truncated MeCP2 protein (Mecp2R168X). These mice display neurological phenotypes such as hind limb clasping and breathing irregularities similar to other mouse models [161]. Another RTT mouse model is the A140V Mecp2 mutant mice (Mecp2A140V). This MeCP2 mutation has also been described in cases of X-linked mental retardation and manic-depressive behaviors. A140V mutant mice produce a mutant MeCP2 protein that lacks the ability to bind to Alpha thalassemia/mental retardation syndrome X-linked (ATRX) protein. These mice have an apparently normal life span, and lack specific phenotypes that are displayed in other mouse models including seizures, tremors, and breathing irregularities. However they show increased cellular packing and reduced dendrite branching similar to what is observed in autopsy brains from RTT individuals [162]. MeCP2 partners with ATRX and is involved in the silencing of imprinted genes in brain [163]. Recently, a Mecp2e1-deficient mouse model was developed with a point mutation in exon 1 changing translational start site of the first exon from “ATG” into “TTG”. These mice show many phenotypes observed in Mecp2 null mice including hindlimb clasping, forelimb stereotypy, and excessive grooming followed by death within 7 to 31 weeks [164]. In order to demonstrate the functions of neuronal activity-dependent phosphorylation at S80, S421 and S424, two knock-in mice models were generated; Mecp2S80A and Mecp2S421A;S424A abolishing the phosphorylation at S80 and S421 sites. The Mecp2S80A mice showed reduced locomotion while opposite locomotors behaviors were seen in Mecp2S421A;S424A, providing insights on the functions of S80 phosphorylation in resting neurons and S421 in active neurons [165].
A rat model of Rett Syndrome was also generated with reduced Mecp2 expression in the brain. This RTT rodent model revealed reduced expression of Bdnf, with no significant phenotypes that mimic RTT [166]. Other RTT animal models include the Drosophila and zebrafish models. The Drosophila RTT model was generated by overexpressing known RTT mutations such as R294X and R106W, which resulted in locomotary dysfunction [167]. The recently reported zebrafish model of RTT was generated by introducing a C to T transition-mutation at position 187 of the Mecp2 coding sequence. This resulted in a nonsense mutation and a truncation of the MeCP2 protein at the position 63 (Mecp2Q63*). RTT zebrafish model display altered motor behaviors, although the phenotype is weaker in comparison to other Mecp2-deficient animal models. In contrast to MeCP2-null mouse models, Mecp2-null zebrafish are viable and fertile [168].
Other than animal models, in vitro cellular models of Mecp2 dysfunction have also been established in order to better understand the how the effects of MeCP2 deficiency impairs normal brain function. Loss of Mecp2 from cultured neuronal cell populations obtained from embryonic or postnatal Mecp2-deficient mice indicate biochemical and morphological abnormalities similar to Mecp2 null animals [49,159,169]. Moreover, other studies have revealed putative roles of MeCP2 in astrocytes and microglia despite the low levels of MeCP2 in these cell types. Mecp2-deficient astrocytes and microglia have been demonstrated to produce aberrant levels of soluble factors, such as glutamate, which inhibit dendrite branching from co-cultured neurons in vitro [47,48,170]. In addition, loss of Mecp2 from astrocytes has been demonstrated to affect astrocytic gap junction function, thereby resulting in their failure to adequately support dendritic development [48]. Table 2 presents an overview of animal models that are developed to study the molecular function of MeCP2.

5. Closing Remarks

Epigenetic regulation of gene expression is fundamental for proper organism development and function. MeCP2 is a multifunctional epigenetic regulator and the malleability in its function underscores its role in many human diseases. Since its discovery, significant progress has been made to understand its dynamic molecular properties, and increasing evidence reveals its central position in many neurological, neurodevelopmental, neuropsychiatric and non-neurological disorders. Despite tremendous progress in understanding the molecular mechanisms by which dysregulation of MeCP2 expression and function results in these disorders, we are still far from translating this knowledge towards effective therapeutic approaches. The availability of excellent animal models promises, not only hope, but also a better strategy to overcome the challenge of translational research. It is very likely that the number of diseases associated with MeCP2 dysfunction will grow rapidly in near future, thus, a better understanding of how MeCP2 functions in complex regulatory networks will pave the way for the discovery of better disease biomarkers as well as novel targets for treatments.
Table 2. Animal models of MeCP2 dysfunction.
Table 2. Animal models of MeCP2 dysfunction.
Animal modelsDescriptionPhenotypeReferences
Mouse
Mecp2 null mouse models:
Mecp2tm1.1bird
Mecp2tm1.Tam
Exon 3 and 4 deletion. MeCP2 expression and function are abolishedUnusual gait, hindlimb clasping, seizures, irregular breathing[150,151]
Mecp2lox−Stop/y or Mecp2tm2.birdGene silencing by Cre recombinase insertion into intron 2. No protein is detected (behaves as a null allele)Phenotypes similar to the null mice with abnormal behavior, RTT-like phenotypes and breathing irregularities.[153]
Mecp2 mutant mice:
Mecp2tm1.1jaeExon 3 deletion. MeCP2 expression and function are abolished.Neurological phenotype similar to Mecp2 null mice, however lifespan is longer.[155]
Mecp2308Introduction of a premature STOP codon in exon 4.
Truncated MeCP2 protein with residual unknown function
Milder neurological phenotype compared to Mecp2 null mice[156]
Mecp2 conditional-mutant mice:
Nestin-cre knockoutBrain-specific deletionSimilar to Mecp2 null mice except for breathing phenotype[150,155]
Sim 1-cre knockoutSelective deletion in neurons of hypothalamus and amygdalaAbnormal stress response, stranger aggression[157]
TH-cre knockoutSelective deletion in dopaminergic and noradrenergic neurons.Hypoactivity, reduced expression of tyrosine hydroxylase[158,159]
CamKII-cre knockoutForebrain-specific deletionImpaired motor co-ordination, anxiety[171]
Pet1-cre knockoutSelective deletion in serotonergic neuronsIncreased aggression, hyperactivity[159]
Viaat-cre knockoutSelective deletion in GABAergic neuronsReduced lifespan, self-mutilation[169]
Mice overexpressing MECP2
MeCP2Tg1MECP2 overexpression in all cells Seizures, premature death, abnormal social behaviors, hypoactivity[160]
Tau-MECP2-rescueMECP2 overexpression in neuronsHypoactivity, impaired cognition[81]
Rett Syndrome mouse models:
Mecp2R168XPremature STOP codon at amino acid 168Hindlimb clasping, breathing irregularities[161]
Mecp2A140VMissense mutation that produces mutant MeCP2 proteinNormal life span, reduced dendrite branching[162]
Mecp2T308AKnock-in mutation that causes loss of interaction with NCoR complexMotor abnormalities, hindlimb clasping[172]
Mecp2R306CKnock-in mutation that causes loss of interaction with NCoR/SMRTImpaired motor function, hindlimb clasping[173]
Mecp2T158AKnock-in mutation that disrupts protein stabilityDevelopmental regression, hypoactivity[174]
Mecp2-e1Point mutation of ATG in exon 1 to TTGForelimb stereotypy, hindlimb clasping, excessive grooming, and hypoactivity[164]
Mecp2S80AKnock-in mouse model with abolished phosphorylation at S80.Reduced locomotion similar to that of Mecp2 null mice and RTT patients[165]
Mecp2S421A;S424ADouble mutant mouse model which lacks phosphorylation at both S421A and S424APhenotypes opposite to Mecp2S80A mice (increased locomotion)[165]
Mecp2 Mouse models of phenotypic rescue
Mecp2lox−Stop/y;cre−ERActivation of Mecp2 gene in Mecp2l°x−St°p/y mouse model by cre-ER and Tamoxifen injections.Rescued majority of RTT phenotypes including increased lifespan, delayed disease progression[153]
Mecp2+/−; CAGGS LSL Mecp2Conditional activation (rescue) of Mecp2 gene in brain using synthetic CAGGS promoter Partial rescue of RTT phenotypes, including delayed disease progression, reduced lethality and improved behaviors [154]
Rat
Mecp2-sh-1Viral mediated RNAi-induced downregulation of Mecp2Transient neurobehavioral abnormalities, reduced Bdnf expression in hippocampus[166]
Zebrafish
Mecp2Q63 *.Nonsense mutation and a truncation of MeCP2 at position 63Altered motor behaviors, however viable and fertile [168]
Drosophila
GMR-Gal4:UAS-MeCP2 R106W/+.Overexpression of mutant MeCP2 proteinLocomotar dysfunction, external eye disruption [167]
GMR-Gal4:UAS-MeCP2 R294X/+.

List of Abbreviations

AS
Angelman Syndrome
ASD
Autism Spectrum Disorders
ATRX
Alpha Thalassemia/Mental Retardation Syndrome X-linked
BDNF
Brain Derived Neurotrophic Factor
CDKL5
Cyclin-Dependent Kinase-Like 5
CTD
C-terminal Domain
DNMT
DNA Methyl Transferase
FASD
Fetal Alcohol Spectrum Disorders
FOXG1
Forkhead Box Protein G1
HD
Huntington's Disease
HSCR
Hirschsprung’s Disease
HTT
Huntingtin Gene
Htt
Huntingtin Protein
MBD
Methyl Binding Domain
MeCP2
Methyl CpG Binding Protein
MECP2
Human MECP2 Gene
Mecp2
Mouse Mecp2 Gene
miRNA
microRNA
RA
Rheumatoid Arthritis
RTT
Rett Syndrome
SLE
Systemic Lupus Erythematosus
TRD
Transcription Repression Domain
XCI
X Chromosome Inactivation
XLMR
X-linked Mental Retardation
5mC
5-methylcytosine
5hmC
5-hyrdoxymethylcytosine

Acknowledgements

We apologize that many excellent papers are not included in this paper due to space limitation. The authors would like to thank members of the Rastegar laboratory (specially Vichithra RB Liyanage) for helpful discussions. The research of M. Rastegar is supported by funds from the Scottish Rites Charitable Foundation of Canada (SRCFC, Grant 10110), Natural Sciences and Engineering Research Council of Canada (NSERC Discovery Grant 372405-2009), Canadian Institute of Health Research (CIHR) Grants TEC-128094 and CEN-132383, as well as Health Sciences Centre Foundation (HSCF). CD Ezeonwuka was a recipient of MHRC-UMGF studentship award.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lewis, J.D.; Meehan, R.R.; Henzel, W.J.; Maurer-Fogy, I.; Jeppesen, P.; Klein, F.; Bird, A. Purification, sequence, and cellular localization of a novel chromosomal protein that binds to methylated DNA. Cell 1992, 69, 905–914. [Google Scholar] [CrossRef]
  2. Reichwald, K.; Thiesen, J.; Wiehe, T.; Weitzel, J.; Poustka, W.A.; Rosenthal, A.; Platzer, M.; Stratling, W.H.; Kioschis, P. Comparative sequence analysis of the mecp2-locus in human and mouse reveals new transcribed regions. Mamm. Genome 2000, 11, 182–190. [Google Scholar] [CrossRef]
  3. Liu, J.; Francke, U. Identification of cis-regulatory elements for mecp2 expression. Hum. Mol. Genet. 2006, 15, 1769–1782. [Google Scholar] [CrossRef]
  4. D'Esposito, M.; Quaderi, N.A.; Ciccodicola, A.; Bruni, P.; Esposito, T.; D'Urso, M.; Brown, S.D. Isolation, physical mapping, and northern analysis of the x-linked human gene encoding methyl cpg-binding protein, mecp2. Mamm. Genome 1996, 7, 533–535. [Google Scholar] [CrossRef]
  5. Singh, J.; Saxena, A.; Christodoulou, J.; Ravine, D. Mecp2 genomic structure and function: Insights from encode. Nucleic Acids Res. 2008, 36, 6035–6047. [Google Scholar] [CrossRef]
  6. Mnatzakanian, G.N.; Lohi, H.; Munteanu, I.; Alfred, S.E.; Yamada, T.; MacLeod, P.J.; Jones, J.R.; Scherer, S.W.; Schanen, N.C.; Friez, M.J.; et al. A previously unidentified mecp2 open reading frame defines a new protein isoform relevant to rett syndrome. Nat. Genet. 2004, 36, 339–341. [Google Scholar] [CrossRef]
  7. Kriaucionis, S.; Bird, A. The major form of mecp2 has a novel n-terminus generated by alternative splicing. Nucleic Acids Res. 2004, 32, 1818–1823. [Google Scholar] [CrossRef]
  8. Zachariah, R.M.; Olson, C.O.; Ezeonwuka, C.; Rastegar, M. Novel mecp2 isoform-specific antibody reveals the endogenous mecp2e1 expression in murine brain, primary neurons and astrocytes. PLoS One 2012, 7, e49763. [Google Scholar]
  9. Hite, K.C.; Adams, V.H.; Hansen, J.C. Recent advances in mecp2 structure and function. Biochem. Cell Biol. 2009, 87, 219–227. [Google Scholar] [CrossRef]
  10. Adams, V.H.; McBryant, S.J.; Wade, P.A.; Woodcock, C.L.; Hansen, J.C. Intrinsic disorder and autonomous domain function in the multifunctional nuclear protein, mecp2. J. Biol. Chem. 2007, 282, 15057–15064. [Google Scholar]
  11. Klose, R.J.; Sarraf, S.A.; Schmiedeberg, L.; McDermott, S.M.; Stancheva, I.; Bird, A.P. DNA binding selectivity of mecp2 due to a requirement for a/t sequences adjacent to methyl-cpg. Mol. Cell 2005, 19, 667–678. [Google Scholar] [CrossRef]
  12. Hansen, J.C.; Ghosh, R.P.; Woodcock, C.L. Binding of the rett syndrome protein, mecp2, to methylated and unmethylated DNA and chromation. IUBMB life 2010, 62, 732–738. [Google Scholar] [CrossRef]
  13. Galvao, T.C.; Thomas, J.O. Structure-specific binding of mecp2 to four-way junction DNA through its methyl cpg-binding domain. Nucleic Acids Res. 2005, 33, 6603–6609. [Google Scholar] [CrossRef]
  14. Kokura, K.; Kaul, S.C.; Wadhwa, R.; Nomura, T.; Khan, M.M.; Shinagawa, T.; Yasukawa, T.; Colmenares, C.; Ishii, S. The ski protein family is required for mecp2-mediated transcriptional repression. J. Biol. Chem. 2001, 276, 34115–34121. [Google Scholar] [CrossRef]
  15. Zachariah, R.M.; Rastegar, M. Linking epigenetics to human disease and rett syndrome: The emerging novel and challenging concepts in mecp2 research. Neural Plast. 2012, 2012, 415825. [Google Scholar]
  16. Ghosh, R.P.; Nikitina, T.; Horowitz-Scherer, R.A.; Gierasch, L.M.; Uversky, V.N.; Hite, K.; Hansen, J.C.; Woodcock, C.L. Unique physical properties and interactions of the domains of methylated DNA binding protein 2. Biochemistry 2010, 49, 4395–4410. [Google Scholar] [CrossRef]
  17. Stuss, D.P.; Cheema, M.; Ng, M.K.; Martinez de Paz, A.; Williamson, B.; Missiaen, K.; Cosman, J.D.; McPhee, D.; Esteller, M.; Hendzel, M.; et al. Impaired in vivo binding of mecp2 to chromatin in the absence of its DNA methyl-binding domain. Nucleic Acids Res. 2013, 41, 4888–4900. [Google Scholar] [CrossRef]
  18. Shahbazian, M.D.; Antalffy, B.; Armstrong, D.L.; Zoghbi, H.Y. Insight into rett syndrome: Mecp2 levels display tissue- and cell-specific differences and correlate with neuronal maturation. Hum. Mol. Genet. 2002, 11, 115–124. [Google Scholar]
  19. Jakovcevski, M.; Akbarian, S. Epigenetic mechanisms in neurological disease. Nat. Med. 2012, 18, 1194–1204. [Google Scholar] [CrossRef]
  20. Liyanage, V.R.B.; Zachariah, R.M.; Delcuve, G.P.; Davie, J.R.; Rastegar, M. New developments in chromatin research: An epigenetic perspective. In New Developments in Chromatin Research; Simpson, N.M., Stewart, V.J., Eds.; Nova Science Publishers: Hauppauge, NY, USA, 2012; pp. 29–58. [Google Scholar]
  21. Delcuve, G.P.; Rastegar, M.; Davie, J.R. Epigenetic control. J. Cell. Physiol. 2009, 219, 243–250. [Google Scholar] [CrossRef]
  22. Mellen, M.; Ayata, P.; Dewell, S.; Kriaucionis, S.; Heintz, N. Mecp2 binds to 5hmc enriched within active genes and accessible chromatin in the nervous system. Cell 2012, 151, 1417–1430. [Google Scholar] [CrossRef]
  23. Nan, X.; Ng, H.H.; Johnson, C.A.; Laherty, C.D.; Turner, B.M.; Eisenman, R.N.; Bird, A. Transcriptional repression by the methyl-cpg-binding protein mecp2 involves a histone deacetylase complex. Nature 1998, 393, 386–389. [Google Scholar] [CrossRef]
  24. Jones, P.L.; Veenstra, G.J.; Wade, P.A.; Vermaak, D.; Kass, S.U.; Landsberger, N.; Strouboulis, J.; Wolffe, A.P. Methylated DNA and mecp2 recruit histone deacetylase to repress transcription. Nat. Genet. 1998, 19, 187–191. [Google Scholar]
  25. Jordan, C.; Li, H.H.; Kwan, H.C.; Francke, U. Cerebellar gene expression profiles of mouse models for rett syndrome reveal novel mecp2 targets. BMC Med. Genet. 2007, 8, 1–16. [Google Scholar]
  26. Chahrour, M.; Jung, S.Y.; Shaw, C.; Zhou, X.; Wong, S.T.; Qin, J.; Zoghbi, H.Y. Mecp2, a key contributor to neurological disease, activates and represses transcription. Science 2008, 320, 1224–1229. [Google Scholar] [CrossRef]
  27. Yasui, D.H.; Peddada, S.; Bieda, M.C.; Vallero, R.O.; Hogart, A.; Nagarajan, R.P.; Thatcher, K.N.; Farnham, P.J.; Lasalle, J.M. Integrated epigenomic analyses of neuronal mecp2 reveal a role for long-range interaction with active genes. Proc. Natl. Acad. Sci. U.S.A. 2007, 104, 19416–19421. [Google Scholar] [CrossRef]
  28. Jeffery, L.; Nakielny, S. Components of the DNA methylation system of chromatin control are rna-binding proteins. J. Biol. Chem. 2004, 279, 49479–49487. [Google Scholar] [CrossRef]
  29. Young, J.I.; Hong, E.P.; Castle, J.C.; Crespo-Barreto, J.; Bowman, A.B.; Rose, M.F.; Kang, D.; Richman, R.; Johnson, J.M.; Berget, S.; et al. Regulation of rna splicing by the methylation-dependent transcriptional repressor methyl-cpg binding protein 2. Proc. Natl. Acad. Sci. U.S.A. 2005, 102, 17551–17558. [Google Scholar] [CrossRef]
  30. Nikitina, T.; Shi, X.; Ghosh, R.P.; Horowitz-Scherer, R.A.; Hansen, J.C.; Woodcock, C.L. Multiple modes of interaction between the methylated DNA binding protein mecp2 and chromatin. Mol. Cell. Biol. 2007, 27, 864–877. [Google Scholar] [CrossRef]
  31. Skene, P.J.; Illingworth, R.S.; Webb, S.; Kerr, A.R.; James, K.D.; Turner, D.J.; Andrews, R.; Bird, A.P. Neuronal mecp2 is expressed at near histone-octamer levels and globally alters the chromatin state. Mol. Cell 2010, 37, 457–468. [Google Scholar] [CrossRef]
  32. Kimura, H.; Shiota, K. Methyl-cpg-binding protein, mecp2, is a target molecule for maintenance DNA methyltransferase, dnmt 1. J. Biol. Chem. 2003, 278, 4806–4812. [Google Scholar] [CrossRef]
  33. Gos, M. Epigenetic mechanisms of gene expression regulation in neurological diseases. Acta Neurobiol. Exp. 2013, 73, 19–37. [Google Scholar]
  34. Olynik, B.M.; Rastegar, M. The genetic and epigenetic journey of embryonic stem cells into mature neural cells. Front. Genet. 2012, 3, (81). 1–16. [Google Scholar]
  35. Barber, B.A.; Rastegar, M. Epigenetic control of hox genes during neurogenesis, development, and disease. Ann. Anat. 2010, 192, 261–274. [Google Scholar] [CrossRef]
  36. Rastegar, M.; Delcuve, G.P.; Davie, J.R. Epigenetic analysis of pluripotent cells. In Human Stem Cell Technology and Biology: A Research Guide and Laboratory Manual. Perspectives in Human Stem Cell Technologies, 1st ed.; Wiley-Blackwell: New Jersey, NY, USA, 2011; pp. 273–288. [Google Scholar]
  37. Rastegar, M.; Kobrossy, L.; Kovacs, E.N.; Rambaldi, I.; Featherstone, M. Sequential histone modifications at hoxd4 regulatory regions distinguish anterior from posterior embryonic compartments. Mol. Cell. Biol. 2004, 24, 8090–8103. [Google Scholar] [CrossRef]
  38. Nolte, C.; Rastegar, M.; Amores, A.; Bouchard, M.; Grote, D.; Maas, R.; Kovacs, E.N.; Postlethwait, J.; Rambaldi, I.; Rowan, S.; et al. Stereospecificity and pax6 function direct hoxd4 neural enhancer activity along the antero-posterior axis. Dev. Biol. 2006, 299, 582–593. [Google Scholar]
  39. Huang, H.; Rastegar, M.; Bodner, C.; Goh, S.L.; Rambaldi, I.; Featherstone, M. Meis c termini harbor transcriptional activation domains that respond to cell signaling. J. Biol. Chem. 2005, 280, 10119–10127. [Google Scholar]
  40. Kobrossy, L.; Rastegar, M.; Featherstone, M. Interplay between chromatin and trans-acting factors regulating the hoxd4 promoter during neural differentiation. J. Biol. Chem. 2006, 281, 25926–25939. [Google Scholar] [CrossRef]
  41. Barber, B.A.; Liyanage, V.R.; Zachariah, R.M.; Olson, C.O.; Bailey, M.A.; Rastegar, M. Dynamic expression of meis1 homeoprotein in e14.5 forebrain and differentiated forebrain-derived neural stem cells. Ann. Anat. 2013, 195, 431–440. [Google Scholar]
  42. Guy, J.; Cheval, H.; Selfridge, J.; Bird, A. The role of mecp2 in the brain. Annu. Rev. Cell. Dev. Biol. 2011, 27, 631–652. [Google Scholar] [CrossRef]
  43. LaSalle, J.M.; Goldstine, J.; Balmer, D.; Greco, C.M. Quantitative localization of heterogeneous methyl-cpg-binding protein 2 (mecp2) expression phenotypes in normal and rett syndrome brain by laser scanning cytometry. Hum. Mol. Genet. 2001, 10, 1729–1740. [Google Scholar] [CrossRef]
  44. Balmer, D.; Goldstine, J.; Rao, Y.M.; LaSalle, J.M. Elevated methyl-cpg-binding protein 2 expression is acquired during postnatal human brain development and is correlated with alternative polyadenylation. J. Mol. Med. 2003, 81, 61–68. [Google Scholar]
  45. Mullaney, B.C.; Johnston, M.V.; Blue, M.E. Developmental expression of methyl-cpg binding protein 2 is dynamically regulated in the rodent brain. Neuroscience 2004, 123, 939–949. [Google Scholar] [CrossRef]
  46. Yasui, D.H.; Xu, H.; Dunaway, K.W.; Lasalle, J.M.; Jin, L.W.; Maezawa, I. Mecp2 modulates gene expression pathways in astrocytes. Mol. Autism 2013, 4, 1–11. [Google Scholar] [CrossRef] [Green Version]
  47. Ballas, N.; Lioy, D.T.; Grunseich, C.; Mandel, G. Non-cell autonomous influence of mecp2-deficient glia on neuronal dendritic morphology. Nat. Neurosci. 2009, 12, 311–317. [Google Scholar] [CrossRef]
  48. Maezawa, I.; Swanberg, S.; Harvey, D.; LaSalle, J.M.; Jin, L.W. Rett syndrome astrocytes are abnormal and spread mecp2 deficiency through gap junctions. J. Neurosci. 2009, 29, 5051–5061. [Google Scholar] [CrossRef]
  49. Rastegar, M.; Hotta, A.; Pasceri, P.; Makarem, M.; Cheung, A.Y.; Elliott, S.; Park, K.J.; Adachi, M.; Jones, F.S.; Clarke, I.D.; et al. Mecp2 isoform-specific vectors with regulated expression for rett syndrome gene therapy. PLoS One 2009, 4, e6810. [Google Scholar] [CrossRef]
  50. Wood, L.; Gray, N.W.; Zhou, Z.; Greenberg, M.E.; Shepherd, G.M. Synaptic circuit abnormalities of motor-frontal layer 2/3 pyramidal neurons in an rna interference model of methyl-cpg-binding protein 2 deficiency. J. Neurosci. 2009, 29, 12440–12448. [Google Scholar] [CrossRef]
  51. Akbarian, S.; Chen, R.Z.; Gribnau, J.; Rasmussen, T.P.; Fong, H.; Jaenisch, R.; Jones, E.G. Expression pattern of the rett syndrome gene mecp2 in primate prefrontal cortex. Neurobiol. Dis. 2001, 8, 784–791. [Google Scholar] [CrossRef]
  52. Trappe, R.; Laccone, F.; Cobilanschi, J.; Meins, M.; Huppke, P.; Hanefeld, F.; Engel, W. Mecp2 mutations in sporadic cases of rett syndrome are almost exclusively of paternal origin. Am. J. Hum. Genet. 2001, 68, 1093–1101. [Google Scholar] [CrossRef]
  53. Wan, M.; Lee, S.S.; Zhang, X.; Houwink-Manville, I.; Song, H.R.; Amir, R.E.; Budden, S.; Naidu, S.; Pereira, J.L.; Lo, I.F.; et al. Rett syndrome and beyond: Recurrent spontaneous and familial mecp2 mutations at cpg hotspots. Am. J. Hum. Genet. 1999, 65, 1520–1529. [Google Scholar] [CrossRef]
  54. Chahrour, M.; Zoghbi, H.Y. The story of rett syndrome: From clinic to neurobiology. Neuron 2007, 56, 422–437. [Google Scholar] [CrossRef]
  55. Dragich, J.; Houwink-Manville, I.; Schanen, C. Rett syndrome: A surprising result of mutation in mecp2. Hum. Mol. Genet. 2000, 9, 2365–2375. [Google Scholar]
  56. Sawalha, A.H.; Webb, R.; Han, S.; Kelly, J.A.; Kaufman, K.M.; Kimberly, R.P.; Alarcon-Riquelme, M.E.; James, J.A.; Vyse, T.J.; Gilkeson, G.S.; et al. Common variants within mecp2 confer risk of systemic lupus erythematosus. PLoS One 2008, 3, e1727. [Google Scholar] [CrossRef]
  57. Rett, A. [on a unusual brain atrophy syndrome in hyperammonemia in childhood]. Wiener medizinische Wochenschrift 1966, 116, 723–726. [Google Scholar]
  58. Shahbazian, M.D.; Zoghbi, H.Y. Rett syndrome and mecp2: Linking epigenetics and neuronal function. Am. J. Hum. Genet. 2002, 71, 1259–1272. [Google Scholar] [CrossRef]
  59. Burd, L.; Gascon, G.G. Rett syndrome: Review and discussion of current diagnostic criteria. J. Child Neurol. 1988, 3, 263–268. [Google Scholar] [CrossRef]
  60. Jan, M.M.; Dooley, J.M.; Gordon, K.E. Male rett syndrome variant: Application of diagnostic criteria. Pediatr. Neurol. 1999, 20, 238–240. [Google Scholar] [CrossRef]
  61. Amir, R.E.; Van den Veyver, I.B.; Wan, M.; Tran, C.Q.; Francke, U.; Zoghbi, H.Y. Rett syndrome is caused by mutations in x-linked mecp2, encoding methyl-cpg-binding protein 2. Nat. Genet. 1999, 23, 185–188. [Google Scholar] [CrossRef]
  62. Budden, S.S.; Dorsey, H.C.; Steiner, R.D. Clinical profile of a male with rett syndrome. Brain Dev. 2005, 27 Suppl 1, S69–S71. [Google Scholar] [CrossRef]
  63. Dayer, A.G.; Bottani, A.; Bouchardy, I.; Fluss, J.; Antonarakis, S.E.; Haenggeli, C.A.; Morris, M.A. Mecp2 mutant allele in a boy with rett syndrome and his unaffected heterozygous mother. Brain Dev. 2007, 29, 47–50. [Google Scholar] [CrossRef]
  64. Hagberg, B.; Hanefeld, F.; Percy, A.; Skjeldal, O. An update on clinically applicable diagnostic criteria in rett syndrome. Comments to rett syndrome clinical criteria consensus panel satellite to european paediatric neurology society meeting, baden baden, germany, 11 september 2001. Euro. J. Paediatr. Neurol. 2002, 6, 293–297. [Google Scholar]
  65. Nomura, Y.; Segawa, M. Natural history of rett syndrome. J. Child Neurol. 2005, 20, 764–768. [Google Scholar]
  66. Segawa, M.; Nomura, Y. Rett syndrome. Curr. Opin. Neurol. 2005, 18, 97–104. [Google Scholar] [CrossRef]
  67. Hagberg, B.; Witt-Engerstrom, I. Rett syndrome: A suggested staging system for describing impairment profile with increasing age towards adolescence. Am. J. Med. Genet. Supplement 1986, 1, 47–59. [Google Scholar] [CrossRef]
  68. Hagberg, B. Rett syndrome: Long-term clinical follow-up experiences over four decades. J. Child Neurol. 2005, 20, 722–727. [Google Scholar]
  69. Jian, L.; Nagarajan, L.; de Klerk, N.; Ravine, D.; Bower, C.; Anderson, A.; Williamson, S.; Christodoulou, J.; Leonard, H. Predictors of seizure onset in rett syndrome. J. Pediatr. 2006, 149, 542–547. [Google Scholar] [CrossRef]
  70. Young, J.I.; Zoghbi, H.Y. X-chromosome inactivation patterns are unbalanced and affect the phenotypic outcome in a mouse model of rett syndrome. Am. J. Hum. Genet. 2004, 74, 511–520. [Google Scholar] [CrossRef]
  71. Amir, R.E.; Van den Veyver, I.B.; Schultz, R.; Malicki, D.M.; Tran, C.Q.; Dahle, E.J.; Philippi, A.; Timar, L.; Percy, A.K.; Motil, K.J.; et al. Influence of mutation type and x chromosome inactivation on rett syndrome phenotypes. Ann. Neurol. 2000, 47, 670–679. [Google Scholar] [CrossRef]
  72. Zappella, M.; Meloni, I.; Longo, I.; Hayek, G.; Renieri, A. Preserved speech variants of the rett syndrome: Molecular and clinical analysis. American J. Med. Genet. 2001, 104, 14–22. [Google Scholar] [CrossRef]
  73. Ariani, F.; Hayek, G.; Rondinella, D.; Artuso, R.; Mencarelli, M.A.; Spanhol-Rosseto, A.; Pollazzon, M.; Buoni, S.; Spiga, O.; Ricciardi, S.; et al. Foxg1 is responsible for the congenital variant of rett syndrome. Am. J. Hum. Genet. 2008, 83, 89–93. [Google Scholar] [CrossRef]
  74. Mari, F.; Azimonti, S.; Bertani, I.; Bolognese, F.; Colombo, E.; Caselli, R.; Scala, E.; Longo, I.; Grosso, S.; Pescucci, C.; et al. Cdkl5 belongs to the same molecular pathway of mecp2 and it is responsible for the early-onset seizure variant of rett syndrome. Hum. Mol. Genet. 2005, 14, 1935–1946. [Google Scholar] [CrossRef]
  75. Chen, Q.; Zhu, Y.C.; Yu, J.; Miao, S.; Zheng, J.; Xu, L.; Zhou, Y.; Li, D.; Zhang, C.; Tao, J.; et al. Cdkl5, a protein associated with rett syndrome, regulates neuronal morphogenesis via rac1 signaling. J. Neurosci. 2010, 30, 12777–12786. [Google Scholar] [CrossRef]
  76. Ellaway, C.J.; Ho, G.; Bettella, E.; Knapman, A.; Collins, F.; Hackett, A.; McKenzie, F.; Darmanian, A.; Peters, G.B.; Fagan, K.; et al. 14q12 microdeletions excluding foxg1 give rise to a congenital variant rett syndrome-like phenotype. Eur. J. Hum. Genet. 2013, 21, 522–527. [Google Scholar] [CrossRef]
  77. Subramaniam, B.; Naidu, S.; Reiss, A.L. Neuroanatomy in rett syndrome: Cerebral cortex and posterior fossa. Neurology 1997, 48, 399–407. [Google Scholar] [CrossRef]
  78. Armstrong, D.D. Neuropathology of rett syndrome. J. Child Neurol. 2005, 20, 747–753. [Google Scholar] [CrossRef]
  79. Jellinger, K.; Armstrong, D.; Zoghbi, H.Y.; Percy, A.K. Neuropathology of rett syndrome. Acta Neuropathol. 1988, 76, 142–158. [Google Scholar] [CrossRef]
  80. Oldfors, A.; Sourander, P.; Armstrong, D.L.; Percy, A.K.; Witt-Engerstrom, I.; Hagberg, B.A. Rett syndrome: Cerebellar pathology. Pediatr. Neurol. 1990, 6, 310–314. [Google Scholar] [CrossRef]
  81. Collins, A.L.; Levenson, J.M.; Vilaythong, A.P.; Richman, R.; Armstrong, D.L.; Noebels, J.L.; David Sweatt, J.; Zoghbi, H.Y. Mild overexpression of mecp2 causes a progressive neurological disorder in mice. Hum. Mol. Genet. 2004, 13, 2679–2689. [Google Scholar] [CrossRef]
  82. Samaco, R.C.; Fryer, J.D.; Ren, J.; Fyffe, S.; Chao, H.T.; Sun, Y.; Greer, J.J.; Zoghbi, H.Y.; Neul, J.L. A partial loss of function allele of methyl-cpg-binding protein 2 predicts a human neurodevelopmental syndrome. Hum. Mol. Genet. 2008, 17, 1718–1727. [Google Scholar] [CrossRef]
  83. Meins, M.; Lehmann, J.; Gerresheim, F.; Herchenbach, J.; Hagedorn, M.; Hameister, K.; Epplen, J.T. Submicroscopic duplication in xq28 causes increased expression of the mecp2 gene in a boy with severe mental retardation and features of rett syndrome. J. Med. Genet. 2005, 42, e12. [Google Scholar] [CrossRef]
  84. del Gaudio, D.; Fang, P.; Scaglia, F.; Ward, P.A.; Craigen, W.J.; Glaze, D.G.; Neul, J.L.; Patel, A.; Lee, J.A.; Irons, M.; et al. Increased mecp2 gene copy number as the result of genomic duplication in neurodevelopmentally delayed males. Genet. Med. 2006, 8, 784–792. [Google Scholar] [CrossRef]
  85. Friez, M.J.; Jones, J.R.; Clarkson, K.; Lubs, H.; Abuelo, D.; Bier, J.A.; Pai, S.; Simensen, R.; Williams, C.; Giampietro, P.F.; et al. Recurrent infections, hypotonia, and mental retardation caused by duplication of mecp2 and adjacent region in xq28. Pediatrics 2006, 118, e1687–1695. [Google Scholar] [CrossRef]
  86. Lugtenberg, D.; Kleefstra, T.; Oudakker, A.R.; Nillesen, W.M.; Yntema, H.G.; Tzschach, A.; Raynaud, M.; Rating, D.; Journel, H.; Chelly, J.; et al. Structural variation in xq28: Mecp2 duplications in 1% of patients with unexplained xlmr and in 2% of male patients with severe encephalopathy. Eur. J. Hum. Genet. 2009, 17, 444–453. [Google Scholar] [CrossRef]
  87. Smyk, M.; Obersztyn, E.; Nowakowska, B.; Nawara, M.; Cheung, S.W.; Mazurczak, T.; Stankiewicz, P.; Bocian, E. Different-sized duplications of xq28, including mecp2, in three males with mental retardation, absent or delayed speech, and recurrent infections. Am. J. Med. Genet. B 2008, 147B, 799–806. [Google Scholar] [CrossRef]
  88. Moretti, P.; Zoghbi, H.Y. Mecp2 dysfunction in rett syndrome and related disorders. Curr. Opin. in Genet. Dev. 2006, 16, 276–281. [Google Scholar] [CrossRef]
  89. Ramocki, M.B.; Tavyev, Y.J.; Peters, S.U. The mecp2 duplication syndrome. Am. J. Med. Genet. A 2010, 152A, 1079–1088. [Google Scholar] [CrossRef]
  90. Ramocki, M.B.; Peters, S.U.; Tavyev, Y.J.; Zhang, F.; Carvalho, C.M.; Schaaf, C.P.; Richman, R.; Fang, P.; Glaze, D.G.; Lupski, J.R.; et al. Autism and other neuropsychiatric symptoms are prevalent in individuals with mecp2 duplication syndrome. Ann. Neurol. 2009, 66, 771–782. [Google Scholar] [CrossRef]
  91. Watson, P.; Black, G.; Ramsden, S.; Barrow, M.; Super, M.; Kerr, B.; Clayton-Smith, J. Angelman syndrome phenotype associated with mutations in mecp2, a gene encoding a methyl cpg binding protein. J. Med. Genet. 2001, 38, 224–228. [Google Scholar] [CrossRef]
  92. Turner, H.; MacDonald, F.; Warburton, S.; Latif, F.; Webb, T. Developmental delay and the methyl binding genes. J. Med. Genet. 2003, 40, E13. [Google Scholar] [CrossRef]
  93. Milani, D.; Pantaleoni, C.; D'Arrigo, S.; Selicorni, A.; Riva, D. Another patient with mecp2 mutation without classic rett syndrome phenotype. Pediatr. Neurol. 2005, 32, 355–357. [Google Scholar] [CrossRef]
  94. Couvert, P.; Bienvenu, T.; Aquaviva, C.; Poirier, K.; Moraine, C.; Gendrot, C.; Verloes, A.; Andres, C.; Le Fevre, A.C.; Souville, I.; et al. Mecp2 is highly mutated in x-linked mental retardation. Hum. Mol. Genet. 2001, 10, 941–946. [Google Scholar] [CrossRef]
  95. Gonzales, M.L.; LaSalle, J.M. The role of mecp2 in brain development and neurodevelopmental disorders. Curr. Psychiat. Rep. 2010, 12, 127–134. [Google Scholar] [CrossRef]
  96. Van Esch, H.; Bauters, M.; Ignatius, J.; Jansen, M.; Raynaud, M.; Hollanders, K.; Lugtenberg, D.; Bienvenu, T.; Jensen, L.R.; Gecz, J.; et al. Duplication of the mecp2 region is a frequent cause of severe mental retardation and progressive neurological symptoms in males. Am. J. Hum. Genet. 2005, 77, 442–453. [Google Scholar] [CrossRef]
  97. Schanen, N.C.; Kurczynski, T.W.; Brunelle, D.; Woodcock, M.M.; Dure, L.S.t.; Percy, A.K. Neonatal encephalopathy in two boys in families with recurrent rett syndrome. J. Child Neurol. 1998, 13, 229–231. [Google Scholar] [CrossRef]
  98. Villard, L.; Kpebe, A.; Cardoso, C.; Chelly, P.J.; Tardieu, P.M.; Fontes, M. Two affected boys in a rett syndrome family: Clinical and molecular findings. Neurology 2000, 55, 1188–1193. [Google Scholar] [CrossRef]
  99. Johnson, C.P.; Myers, S.M.; American Academy of Pediatrics Council on Children With, D. Identification and evaluation of children with autism spectrum disorders. Pediatrics 2007, 120, 1183–1215. [Google Scholar] [CrossRef]
  100. Carney, R.M.; Wolpert, C.M.; Ravan, S.A.; Shahbazian, M.; Ashley-Koch, A.; Cuccaro, M.L.; Vance, J.M.; Pericak-Vance, M.A. Identification of mecp2 mutations in a series of females with autistic disorder. Pediatr. Neurol. 2003, 28, 205–211. [Google Scholar] [CrossRef]
  101. Shibayama, A.; Cook, E.H., Jr.; Feng, J.; Glanzmann, C.; Yan, J.; Craddock, N.; Jones, I.R.; Goldman, D.; Heston, L.L.; Sommer, S.S. Mecp2 structural and 3'-utr variants in schizophrenia, autism and other psychiatric diseases: A possible association with autism. Am. J. Med. Genet. B 2004, 128B, 50–53. [Google Scholar] [CrossRef]
  102. Nagarajan, R.P.; Hogart, A.R.; Gwye, Y.; Martin, M.R.; LaSalle, J.M. Reduced mecp2 expression is frequent in autism frontal cortex and correlates with aberrant mecp2 promoter methylation. Epigenetics 2006, 1, e1–e11. [Google Scholar] [CrossRef]
  103. Chiurazzi, P.; Pomponi, M.G.; Willemsen, R.; Oostra, B.A.; Neri, G. In vitro reactivation of the fmr1 gene involved in fragile x syndrome. Hum. Mol. Genet. 1998, 7, 109–113. [Google Scholar]
  104. Nguyen, A.; Rauch, T.A.; Pfeifer, G.P.; Hu, V.W. Global methylation profiling of lymphoblastoid cell lines reveals epigenetic contributions to autism spectrum disorders and a novel autism candidate gene, rora, whose protein product is reduced in autistic brain. FASEB J. 2010, 24, 3036–3051. [Google Scholar] [CrossRef]
  105. Liyanage, V.R.; Zachariah, R.M.; Rastegar, M. Decitabine alters the expression of mecp2 isoforms via dynamic DNA methylation at the mecp2 regulatory elements in neural stem cells. Mol. Autism 2013, 4, 1–21. [Google Scholar]
  106. Adegbola, A.A.; Gonzales, M.L.; Chess, A.; LaSalle, J.M.; Cox, G.F. A novel hypomorphic mecp2 point mutation is associated with a neuropsychiatric phenotype. Hum. Genet. 2009, 124, 615–623. [Google Scholar] [CrossRef]
  107. Astley, S.J. Fetal alcohol syndrome prevention in washington state: Evidence of success. Paediatr. Perinat. Ep. 2004, 18, 344–351. [Google Scholar] [CrossRef]
  108. Haycock, P.C. Fetal alcohol spectrum disorders: The epigenetic perspective. Biol. Reprod. 2009, 81, 607–617. [Google Scholar] [CrossRef]
  109. Kim, P.; Park, J.H.; Choi, C.S.; Choi, I.; Joo, S.H.; Kim, M.K.; Kim, S.Y.; Kim, K.C.; Park, S.H.; Kwon, K.J.; et al. Effects of ethanol exposure during early pregnancy in hyperactive, inattentive and impulsive behaviors and mecp2 expression in rodent offspring. Neurochem. Res. 2013, 38, 620–631. [Google Scholar] [CrossRef]
  110. Tunc-Ozcan, E.; Ullmann, T.M.; Shukla, P.K.; Redei, E.E. Low-dose thyroxine attenuates autism-associated adverse effects of fetal alcohol in male offspring's social behavior and hippocampal gene expression. Alcohol. Clin. Exp. Res. 2013, 37, 1986–1995. [Google Scholar] [CrossRef]
  111. Romano-Lopez, A.; Mendez-Diaz, M.; Ruiz-Contreras, A.E.; Carrisoza, R.; Prospero-Garcia, O. Maternal separation and proclivity for ethanol intake: A potential role of the endocannabinoid system in rats. Neuroscience 2012, 223, 296–304. [Google Scholar] [CrossRef]
  112. Zoll, B.; Huppke, P.; Wessel, A.; Bartels, I.; Laccone, F. Fetal alcohol syndrome in association with rett syndrome. Genet. Couns. 2004, 15, 207–212. [Google Scholar]
  113. Repunte-Canonigo, V.; Chen, J.; Lefebvre, C.; Kawamura, T.; Kreifeldt, M.; Basson, O.; Roberts, A.J.; Sanna, P.P. Mecp2 regulates ethanol sensitivity and intake. Addict. Biol. 2013. [Google Scholar] [CrossRef]
  114. Walker, F.O. Huntington's disease. Lancet 2007, 369, 218–228. [Google Scholar] [CrossRef]
  115. van der Burg, J.M.; Bjorkqvist, M.; Brundin, P. Beyond the brain: Widespread pathology in huntington's disease. Lancet Neurol. 2009, 8, 765–774. [Google Scholar] [CrossRef]
  116. Myers, R.H. Huntington's disease genetics. NeuroRx 2004, 1, 255–262. [Google Scholar] [CrossRef]
  117. McFarland, K.N.; Huizenga, M.N.; Darnell, S.B.; Sangrey, G.R.; Berezovska, O.; Cha, J.H.; Outeiro, T.F.; Sadri-Vakili, G. Mecp2: A novel huntingtin interactor. Hum. Mol. Genet. 2013.
  118. Parry, L.; Clarke, A.R. The roles of the methyl-cpg binding proteins in cancer. Genes Cancer 2011, 2, 618–630. [Google Scholar] [CrossRef]
  119. Pulukuri, S.M.; Patibandla, S.; Patel, J.; Estes, N.; Rao, J.S. Epigenetic inactivation of the tissue inhibitor of metalloproteinase-2 (timp-2) gene in human prostate tumors. Oncogene 2007, 26, 5229–5237. [Google Scholar] [CrossRef]
  120. Pampalakis, G.; Prosnikli, E.; Agalioti, T.; Vlahou, A.; Zoumpourlis, V.; Sotiropoulou, G. A tumor-protective role for human kallikrein-related peptidase 6 in breast cancer mediated by inhibition of epithelial-to-mesenchymal transition. Cancer Res. 2009, 69, 3779–3787. [Google Scholar] [CrossRef]
  121. Pancione, M.; Sabatino, L.; Fucci, A.; Carafa, V.; Nebbioso, A.; Forte, N.; Febbraro, A.; Parente, D.; Ambrosino, C.; Normanno, N.; et al. Epigenetic silencing of peroxisome proliferator-activated receptor gamma is a biomarker for colorectal cancer progression and adverse patients' outcome. PLoS One 2010, 5, e14229. [Google Scholar] [CrossRef]
  122. Lin, R.K.; Hsu, H.S.; Chang, J.W.; Chen, C.Y.; Chen, J.T.; Wang, Y.C. Alteration of DNA methyltransferases contributes to 5'cpg methylation and poor prognosis in lung cancer. Lung Cancer 2007, 55, 205–213. [Google Scholar] [CrossRef]
  123. Shin, J.E.; Park, S.H.; Jang, Y.K. Epigenetic up-regulation of leukemia inhibitory factor (lif) gene during the progression to breast cancer. Mol. Cells 2011, 31, 181–189. [Google Scholar] [CrossRef]
  124. Wada, R.; Akiyama, Y.; Hashimoto, Y.; Fukamachi, H.; Yuasa, Y. Mir-212 is downregulated and suppresses methyl-cpg-binding protein mecp2 in human gastric cancer. Int. J. Cancer 2010, 127, 1106–1114. [Google Scholar]
  125. Wang, Z.; Zhang, J.; Zhang, Y.; Srivenugopal, K.S.; Lim, S.H. Span-xb core promoter sequence is regulated in myeloma cells by specific cpg dinucleotides associated with the mecp2 protein. Int. J. Cancer 2006, 119, 2878–2884. [Google Scholar] [CrossRef]
  126. Meklat, F.; Li, Z.; Wang, Z.; Zhang, Y.; Zhang, J.; Jewell, A.; Lim, S.H. Cancer-testis antigens in haematological malignancies. Br. J. Haematol. 2007, 136, 769–776. [Google Scholar] [CrossRef]
  127. Xu, X.; Jin, H.; Liu, Y.; Liu, L.; Wu, Q.; Guo, Y.; Yu, L.; Liu, Z.; Zhang, T.; Zhang, X.; et al. The expression patterns and correlations of claudin-6, methy-cpg binding protein 2, DNA methyltransferase 1, histone deacetylase 1, acetyl-histone h3 and acetyl-histone h4 and their clinicopathological significance in breast invasive ductal carcinomas. Diagn. Pathol. 2012, 7, 33. [Google Scholar] [CrossRef]
  128. Wang, J.T.; Wu, T.T.; Bai, L.; Ding, L.; Hao, M.; Wang, Y. [effect of folate in modulating the expression of DNA methyltransferase 1 and methyl-cpg-bingding protein 2 in cervical cancer cell lines]. Zhonghua Liu Xing Bing Xue Za Zhi 2013, 34, 173–177. [Google Scholar]
  129. Fortuna, G.; Brennan, M.T. Systemic lupus erythematosus: Epidemiology, pathophysiology, manifestations, and management. Dent Clin. North Am. 2013, 57, 631–655. [Google Scholar] [CrossRef]
  130. Okada, M.; Ogasawara, H.; Kaneko, H.; Hishikawa, T.; Sekigawa, I.; Hashimoto, H.; Maruyama, N.; Kaneko, Y.; Yamamoto, N. Role of DNA methylation in transcription of human endogenous retrovirus in the pathogenesis of systemic lupus erythematosus. J. Rheumatol. 2002, 29, 1678–1682. [Google Scholar]
  131. Lu, Q.; Wu, A.; Tesmer, L.; Ray, D.; Yousif, N.; Richardson, B. Demethylation of cd40lg on the inactive x in t cells from women with lupus. J. Immunol. 2007, 179, 6352–6358. [Google Scholar]
  132. Deng, C.; Kaplan, M.J.; Yang, J.; Ray, D.; Zhang, Z.; McCune, W.J.; Hanash, S.M.; Richardson, B.C. Decreased ras-mitogen-activated protein kinase signaling may cause DNA hypomethylation in t lymphocytes from lupus patients. Arthritis Rheum. 2001, 44, 397–407. [Google Scholar] [CrossRef]
  133. Liu, K.; Zhang, L.; Chen, J.; Hu, Z.; Cai, G.; Hong, Q. Association of mecp2 (rs2075596, rs2239464) genetic polymorphisms with systemic lupus erythematosus: A meta-analysis. Lupus 2013, 22, 908–918. [Google Scholar] [CrossRef]
  134. Webb, R.; Wren, J.D.; Jeffries, M.; Kelly, J.A.; Kaufman, K.M.; Tang, Y.; Frank, M.B.; Merrill, J.; Kimberly, R.P.; Edberg, J.C.; et al. Variants within mecp2, a key transcription regulator, are associated with increased susceptibility to lupus and differential gene expression in patients with systemic lupus erythematosus. Arthritis Rheum. 2009, 60, 1076–1084. [Google Scholar]
  135. Choy, E. Understanding the dynamics: Pathways involved in the pathogenesis of rheumatoid arthritis. Rheumatology 2012, 51 Suppl 5, v3–11. [Google Scholar] [CrossRef]
  136. Cooles, F.A.; Isaacs, J.D. Pathophysiology of rheumatoid arthritis. Curr. Opin. Rheum. 2011, 23, 233–240. [Google Scholar] [CrossRef]
  137. Ballestar, E. Epigenetic alterations in autoimmune rheumatic diseases. Nat. Rev. Rheumatol. 2011, 7, 263–271. [Google Scholar] [CrossRef]
  138. Darrah, E.; Rosen, A.; Giles, J.T.; Andrade, F. Peptidylarginine deiminase 2, 3 and 4 have distinct specificities against cellular substrates: Novel insights into autoantigen selection in rheumatoid arthritis. Ann. Rheum. Dis. 2012, 71, 92–98. [Google Scholar] [CrossRef]
  139. Chandrashekara, S.; Sachin, S. Measures in rheumatoid arthritis: Are we measuring too many parameters. Int. J. Rheum. Dis. 2012, 15, 239–248. [Google Scholar] [CrossRef]
  140. Miao, C.G.; Huang, C.; Huang, Y.; Yang, Y.Y.; He, X.; Zhang, L.; Lv, X.W.; Jin, Y.; Li, J. Mecp2 modulates the canonical wnt pathway activation by targeting sfrp4 in rheumatoid arthritis fibroblast-like synoviocytes in rats. Cell. Signalling 2013, 25, 598–608. [Google Scholar] [CrossRef]
  141. Miao, C.G.; Yang, Y.Y.; He, X.; Li, J. New advances of DNA methylation and histone modifications in rheumatoid arthritis, with special emphasis on mecp2. Cell. Signalling 2013, 25, 875–882. [Google Scholar] [CrossRef]
  142. Worman, S.; Ganiats, T.G. Hirschsprung’s disease: A cause of chronic constipation in children. Am. Fam. Physician 1995, 51, 487–494. [Google Scholar]
  143. Martucciello, G.; Ceccherini, I.; Lerone, M.; Jasonni, V. Pathogenesis of hirschsprung's disease. J. Pediatr. Surg. 2000, 35, 1017–1025. [Google Scholar] [CrossRef]
  144. Martucciello, G. Hirschsprung's disease, one of the most difficult diagnoses in pediatric surgery: A review of the problems from clinical practice to the bench. Eur. J. Pediatr. Surg. 2008, 18, 140–149. [Google Scholar] [CrossRef]
  145. Badner, J.A.; Sieber, W.K.; Garver, K.L.; Chakravarti, A. A genetic study of hirschsprung disease. Am. J. Hum. Genet. 1990, 46, 568–580. [Google Scholar]
  146. Parisi, M.A.; Kapur, R.P. Genetics of hirschsprung disease. Curr. Opin. Pediatr. 2000, 12, 610–617. [Google Scholar] [CrossRef]
  147. Zhou, Z.; Qin, J.; Tang, J.; Li, B.; Geng, Q.; Jiang, W.; Wu, W.; Rehan, V.; Tang, W.; Xu, X.; et al. Down-regulation of mecp2 in hirschsprung's disease. J. Pediatr. Surg. 2013, 48, 2099–2105. [Google Scholar] [CrossRef]
  148. Jedele, K.B. The overlapping spectrum of rett and angelman syndromes: A clinical review. Seminars in Pediatr. Neurol. 2007, 14, 108–117. [Google Scholar] [CrossRef]
  149. Sohn, B.H.; Park, I.Y.; Lee, J.J.; Yang, S.J.; Jang, Y.J.; Park, K.C.; Kim, D.J.; Lee, D.C.; Sohn, H.A.; Kim, T.W.; et al. Functional switching of tgf-beta1 signaling in liver cancer via epigenetic modulation of a single cpg site in ttp promoter. Gastroenterology 2010, 138, 1898–1908. [Google Scholar] [CrossRef]
  150. Guy, J.; Hendrich, B.; Holmes, M.; Martin, J.E.; Bird, A. A mouse mecp2-null mutation causes neurological symptoms that mimic rett syndrome. Nat. Genet. 2001, 27, 322–326. [Google Scholar] [CrossRef]
  151. Pelka, G.J.; Watson, C.M.; Radziewic, T.; Hayward, M.; Lahooti, H.; Christodoulou, J.; Tam, P.P. Mecp2 deficiency is associated with learning and cognitive deficits and altered gene activity in the hippocampal region of mice. Brain 2006, 129, 887–898. [Google Scholar] [CrossRef]
  152. Calfa, G.; Percy, A.K.; Pozzo-Miller, L. Experimental models of rett syndrome based on mecp2 dysfunction. Exp. Biol. Med. 2011, 236, 3–19. [Google Scholar]
  153. Guy, J.; Gan, J.; Selfridge, J.; Cobb, S.; Bird, A. Reversal of neurological defects in a mouse model of rett syndrome. Science 2007, 315, 1143–1147. [Google Scholar] [CrossRef]
  154. Giacometti, E.; Luikenhuis, S.; Beard, C.; Jaenisch, R. Partial rescue of mecp2 deficiency by postnatal activation of mecp2. Proc. Natl. Acad. Sci. U.S.A. 2007, 104, 1931–1936. [Google Scholar] [CrossRef]
  155. Chen, R.Z.; Akbarian, S.; Tudor, M.; Jaenisch, R. Deficiency of methyl-cpg binding protein-2 in cns neurons results in a rett-like phenotype in mice. Nat. Genet. 2001, 27, 327–331. [Google Scholar] [CrossRef]
  156. Shahbazian, M.; Young, J.; Yuva-Paylor, L.; Spencer, C.; Antalffy, B.; Noebels, J.; Armstrong, D.; Paylor, R.; Zoghbi, H. Mice with truncated mecp2 recapitulate many rett syndrome features and display hyperacetylation of histone h3. Neuron 2002, 35, 243–254. [Google Scholar] [CrossRef]
  157. Fyffe, S.L.; Neul, J.L.; Samaco, R.C.; Chao, H.T.; Ben-Shachar, S.; Moretti, P.; McGill, B.E.; Goulding, E.H.; Sullivan, E.; Tecott, L.H.; et al. Deletion of mecp2 in sim1-expressing neurons reveals a critical role for mecp2 in feeding behavior, aggression, and the response to stress. Neuron 2008, 59, 947–958. [Google Scholar] [CrossRef]
  158. Lindeberg, J.; Usoskin, D.; Bengtsson, H.; Gustafsson, A.; Kylberg, A.; Soderstrom, S.; Ebendal, T. Transgenic expression of cre recombinase from the tyrosine hydroxylase locus. Genesis 2004, 40, 67–73. [Google Scholar] [CrossRef]
  159. Samaco, R.C.; Mandel-Brehm, C.; Chao, H.T.; Ward, C.S.; Fyffe-Maricich, S.L.; Ren, J.; Hyland, K.; Thaller, C.; Maricich, S.M.; Humphreys, P.; et al. Loss of mecp2 in aminergic neurons causes cell-autonomous defects in neurotransmitter synthesis and specific behavioral abnormalities. Proc. Natl. Acad. Sci. U.S.A. 2009, 106, 21966–21971. [Google Scholar] [CrossRef]
  160. Luikenhuis, S.; Giacometti, E.; Beard, C.F.; Jaenisch, R. Expression of mecp2 in postmitotic neurons rescues rett syndrome in mice. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 6033–6038. [Google Scholar] [CrossRef]
  161. Lawson-Yuen, A.; Liu, D.; Han, L.; Jiang, Z.I.; Tsai, G.E.; Basu, A.C.; Picker, J.; Feng, J.; Coyle, J.T. Ube3a mrna and protein expression are not decreased in mecp2r168x mutant mice. Brain Res. 2007, 1180, 1–6. [Google Scholar]
  162. Jentarra, G.M.; Olfers, S.L.; Rice, S.G.; Srivastava, N.; Homanics, G.E.; Blue, M.; Naidu, S.; Narayanan, V. Abnormalities of cell packing density and dendritic complexity in the mecp2 a140v mouse model of rett syndrome/x-linked mental retardation. BMC Neurosci. 2010, 11, 19. [Google Scholar] [CrossRef]
  163. Kernohan, K.D.; Jiang, Y.; Tremblay, D.C.; Bonvissuto, A.C.; Eubanks, J.H.; Mann, M.R.; Berube, N.G. Atrx partners with cohesin and mecp2 and contributes to developmental silencing of imprinted genes in the brain. Dev. Cell 2010, 18, 191–202. [Google Scholar] [CrossRef]
  164. Yasui, D.H.; Gonzales, M.L.; Aflatooni, J.O.; Crary, F.K.; Hu, D.J.; Gavino, B.J.; Golub, M.S.; Vincent, J.B.; Carolyn Schanen, N.; Olson, C.O.; et al. Mice with an isoform-ablating mecp2 exon 1 mutation recapitulate the neurologic deficits of rett syndrome. Hum. Mol. Genet. 2014. [Google Scholar] [CrossRef]
  165. Tao, J.; Hu, K.; Chang, Q.; Wu, H.; Sherman, N.E.; Martinowich, K.; Klose, R.J.; Schanen, C.; Jaenisch, R.; Wang, W.; et al. Phosphorylation of mecp2 at serine 80 regulates its chromatin association and neurological function. Proc. Natl. Acad. Sci. U.S.A. 2009, 106, 4882–4887. [Google Scholar] [CrossRef]
  166. Jin, J.; Bao, X.; Wang, H.; Pan, H.; Zhang, Y.; Wu, X. Rnai-induced down-regulation of mecp2 expression in the rat brain. Int. J. Dev. Neurosci. 2008, 26, 457–465. [Google Scholar] [CrossRef]
  167. Cukier, H.N.; Perez, A.M.; Collins, A.L.; Zhou, Z.; Zoghbi, H.Y.; Botas, J. Genetic modifiers of mecp2 function in drosophila. PLoS Genet. 2008, 4, e1000179. [Google Scholar] [CrossRef]
  168. Pietri, T.; Roman, A.C.; Guyon, N.; Romano, S.A.; Washbourne, P.; Moens, C.B.; de Polavieja, G.G.; Sumbre, G. The first mecp2-null zebrafish model shows altered motor behaviors. Front. Neural Circuits 2013, 7, 118. [Google Scholar]
  169. Chao, H.T.; Chen, H.; Samaco, R.C.; Xue, M.; Chahrour, M.; Yoo, J.; Neul, J.L.; Gong, S.; Lu, H.C.; Heintz, N.; et al. Dysfunction in gaba signalling mediates autism-like stereotypies and rett syndrome phenotypes. Nature 2010, 468, 263–269. [Google Scholar] [CrossRef]
  170. Maezawa, I.; Jin, L.W. Rett syndrome microglia damage dendrites and synapses by the elevated release of glutamate. J. Neurosci. 2010, 30, 5346–5356. [Google Scholar] [CrossRef]
  171. Gemelli, T.; Berton, O.; Nelson, E.D.; Perrotti, L.I.; Jaenisch, R.; Monteggia, L.M. Postnatal loss of methyl-cpg binding protein 2 in the forebrain is sufficient to mediate behavioral aspects of rett syndrome in mice. Biol. Psychiat. 2006, 59, 468–476. [Google Scholar] [CrossRef]
  172. Ebert, D.H.; Gabel, H.W.; Robinson, N.D.; Kastan, N.R.; Hu, L.S.; Cohen, S.; Navarro, A.J.; Lyst, M.J.; Ekiert, R.; Bird, A.P.; et al. Activity-dependent phosphorylation of mecp2 threonine 308 regulates interaction with ncor. Nature 2013, 499, 341–345. [Google Scholar]
  173. Lyst, M.J.; Ekiert, R.; Ebert, D.H.; Merusi, C.; Nowak, J.; Selfridge, J.; Guy, J.; Kastan, N.R.; Robinson, N.D.; de Lima Alves, F.; et al. Rett syndrome mutations abolish the interaction of mecp2 with the ncor/smrt co-repressor. Nat. Neurosci. 2013, 16, 898–902. [Google Scholar] [CrossRef]
  174. Goffin, D.; Allen, M.; Zhang, L.; Amorim, M.; Wang, I.T.; Reyes, A.R.; Mercado-Berton, A.; Ong, C.; Cohen, S.; Hu, L.; et al. Rett syndrome mutation mecp2 t158a disrupts DNA binding, protein stability and erp responses. Nat. Neurosci. 2012, 15, 274–283. [Google Scholar]

Share and Cite

MDPI and ACS Style

Ezeonwuka, C.D.; Rastegar, M. MeCP2-Related Diseases and Animal Models. Diseases 2014, 2, 45-70. https://0-doi-org.brum.beds.ac.uk/10.3390/diseases2010045

AMA Style

Ezeonwuka CD, Rastegar M. MeCP2-Related Diseases and Animal Models. Diseases. 2014; 2(1):45-70. https://0-doi-org.brum.beds.ac.uk/10.3390/diseases2010045

Chicago/Turabian Style

Ezeonwuka, Chinelo D., and Mojgan Rastegar. 2014. "MeCP2-Related Diseases and Animal Models" Diseases 2, no. 1: 45-70. https://0-doi-org.brum.beds.ac.uk/10.3390/diseases2010045

Article Metrics

Back to TopTop