Epigenetic regulators sculpt the plastic brain

Ji-Song Guan , Hong Xie , San-Xiong Liu

Front. Biol. ›› 2017, Vol. 12 ›› Issue (5) : 317 -332.

PDF (559KB)
Front. Biol. ›› 2017, Vol. 12 ›› Issue (5) : 317 -332. DOI: 10.1007/s11515-017-1465-z
REVIEW
REVIEW

Epigenetic regulators sculpt the plastic brain

Author information +
History +
PDF (559KB)

Abstract

BACKGROUND: Epigenetic regulation is a level of transcriptional regulation that occurs in addition to the genetic programming found in biological systems. In the brain, the epigenetic machinery gives the system an opportunity to adapt to a given environment to help not only the individual but also the species survive and expand. However, such a regulatory system has risks, as mutations resulting from epigenetic regulation can cause severe neurological or psychiatric disorders.

OBJECTIVE: Here, we review the most recent findings regarding the epigenetic mechanisms that control the activity-dependent gene transcription leading to synaptic plasticity and brain function and the defects in these mechanisms that lead to neurological disorders.

METHODS: A search was carried out systematically, searching all relevant publications up to June 2017, using the PubMed search engine. The following keywords were used: “activity induced epigenetic,” “gene transcription,” and “neurological disorders.”

RESULTS: A wide range of studies focused on the roles of epigenetics in transgenerational inheritance, neural differentiation, neural circuit assembly and brain diseases. Thirty-one articles focused specifically on activity-induced epigenetic modifications that regulated gene transcription and memory formation and consolidation.

CONCLUSION: Activity-dependent epigenetic mechanisms of gene expression regulation contribute to basic neuronal physiology, and defects were associated with an elevated risk for brain disorders.

Keywords

epigenetic / activity-dependent gene expression / memory / neurological diseases

Cite this article

Download citation ▾
Ji-Song Guan, Hong Xie, San-Xiong Liu. Epigenetic regulators sculpt the plastic brain. Front. Biol., 2017, 12(5): 317-332 DOI:10.1007/s11515-017-1465-z

登录浏览全文

4963

注册一个新账户 忘记密码

Introduction

The central nervous system is a dynamic system. For any stable dynamic system, a stimulus triggers a response to activate an “engaged” state and then the system returns to a normal state. For example, at the level of the neural circuit, stimuli trigger a chain reaction of neuronal firing; however, even the most extensive stimuli-induced neuronal firing in the brain would not last long. After the stimulus, an inhibitory circuit or a homeostasis regulatory network would be engaged and balance the network activity, thus helping the circuit to return to a state of normal activity. At the cellular level, stimuli induce a chain reaction of cell signaling molecules to activate the cell and produce the energy needed for the stimuli-induced neuronal firing. Stimuli-induced responses also remodel the cellular structure, including pruning excessive synapses and building new synapses. By any means, molecular activation in the cell would eventually return to a normal state.

The brain is not like other stable dynamic systems in that external stimuli can gradually optimize the system to function efficiently within the external environment. One essential feature of the brain is that it can remember past experiences and optimize its responses to stimuli according to prior experiences. In other words, instead of producing one stable point in a dynamic system, the brain can create a chain of stable points according to the history of each individual (Fig. 1). This feature requires additional regulatory mechanisms as well as the genetic programming. Such a regulatory mechanism must have the following three properties: be long-lasting, so that experiences can be memorized; have a capacity to be switched on and off by environmental signals; and create a stable point in the dynamic system that requires a feedback loop to maintain this regulatory state.

Accumulating evidence suggests that epigenetic regulation is a candidate mechanism for a second level of response regulation in the central nervous system. Epigenetic regulation has been found to be essential for maintaining cell fate after differentiation during neural development and for regulating learning and memory in the mature brain. Epigenetic regulation can even modulate brain function via transgenerational inheritance.

Importantly, recent discoveries in many neurological and most psychiatric disorders (De Rubeis et al., 2014;Devor et al., 2017; Stessman et al., 2017) did not identify single dominant genetic mutations but, instead, studies have identified hundreds of genes associated with increased risks for a single disease, implicating that malfunction of the central nervous system can be induced by a combination of concurrent environmental stresses and genetic mutations. Interestingly, many of the risk factors involve epigenetic regulators. Here, we summarize the major findings of epigenetic mechanisms in the central nervous system. We hypothesize that epigenetic regulation is essential in creating a chain of stimuli-engaged stable states in the dynamic system of the brain; defects in mechanisms of epigenetic regulation can lead to either instability of the system or the loss of environmentally engaged stable states, thereby leading to an increased risk of neurological and psychiatric disorders.

Long-lasting epigenetic modifications regulate gene transcription

In the adult brain, epigenetic mechanisms play an important role in generating and maintaining changes in synaptic plasticity and memory formation through the modulation of gene expression (Guan et al., 2009; Gräff et al., 2012). Epigenetic modifications can be defined as the structural adaptations of chromosomal regions to induce changes in the phenotype without a change to the genotype (Bird, 2007; Waddington, 2012). Epigenetic regulation often includes three aspects: 1) modifications at the level of the nucleotides, which includes DNA methylation and RNA interference (RNAi) (Ramsahoye et al., 2000; Bird, 2002; Laird, 2003; Matzke and Birchler, 2005); 2) post-translational modifications at the level of histones (PTMs) and the incorporation of histone variants (Jenuwein and Allis, 2001; Pusarla and Bhargava, 2005; Kouzarides, 2007; Talbert and Henikoff, 2010); and 3) nucleosome remodeling, which refers to ATP-dependent processes that modify the structure of chromatin (Becker and Hörz, 2002; Clapier and Cairns, 2009). These modifications regulate gene transcription and are essential for normal chromosome architecture and function. The epigenetic marks are variable over time and can be remodeled in response to external stimuli (Borrelli et al., 2008). Above all, epigenetics offers potential mechanisms for sustained changes in the transcriptional activity in the central nervous system as induced by neuronal activity (Roth and Sweatt, 2009; Sweatt, 2013; Zovkic and Sweatt, 2015; Sweatt, 2016).

Activity-dependent switch of epigenetic regulation

Transcriptional regulation by three types of epigenetic regulators in the brain

Activity-dependent alterations in the transcriptional program is governed by histone acetylation (Table 1). Histone acetylation is catalyzed by histone acetyl transferases (HATs), and acetyl marks are removed by histone deacetylases (HDACs) (Kouzarides, 2007; Guan et al., 2009). Interestingly, neuronal depolarization leads to a widespread recruitment of CBP, an enhancer that is correlated with an increase in expression of target genes (Kim et al., 2010). High frequency stimulation leads to a global increase in H3 and H4 acetylation on the reelin and bdnf promoters, correlating with a higher level of targeted gene expression and LTP induction (Sui et al., 2012). Membrane depolarization of cortical neurons increases H3K27ac at a subset of enhancers and regulates activity-dependent transcription (Malik et al., 2014). In the mouse brain, histone acetylation is associated with learning. Object memory formation enriches H3K14ac at thezif268 promoter and promotes zif268 expression during memory consolidation (Gräff et al., 2012). Strikingly, light pulse stimulation induced rapid histone acetylation in the promoters of mPer1 or mPer2 in the suprachiasmatic nucleus (SCN) and increased mouse Per1 (mPer1) expression (Naruse et al., 2004).

Histone methylation is also regulated by neural activity and causes bi-directional effects on chromatin structure and transcriptional activity (Kouzarides, 2007). Contextual fear learning increases global levels of H3K9me2 in area CA1 and the EC (Gupta-Agarwal et al., 2012) and increases H3K4me3 to promote memory consolidation (Gupta et al., 2010). Histone methylation can specifically regulate gene transcription. Nrxn1 is one of the neurexin family members that encodes presynaptic adhesion molecules and is essential for synapse formation (Südhof, 2008). Neuronal activity triggered binding of Ash1L to the promoter to enrich H3K36me2 at the nrxn1a promoter region, leading to the activity-dependent transcriptional repression (Zhu et al., 2016).

Histone phosphorylation represents another form of transcriptional regulation (Banerjee and Chakravarti, 2011). Kainic acid treatment in the mouse hippocampus leads to an increase in H3 phosphorylation at serine 10 (S10), which correlates temporally withc-fos induction (Crosio et al., 2003). Nighttime light exposure increases H3S10 phosphorylation, paralleling c-fos and Per1 induction (Crosio et al., 2000). Moreover, contextual fear conditioning increases the global level of histone H3 phosphorylation in area CA1, which is required for memory formation (Chwang et al., 2006) (Fig. 2).

In addition to histone modifications, DNA methylation is another important form of epigenetic regulation. DNA methylation is catalyzed by DNMT3a and DNMT3b via the addition of a methyl group to a cytosine residue; this methylation mark is maintained by DNMT1(Goll and Bestor, 2005). Synchronous neuronal activation modifies the DNA methylome and the chromatin accessibility landscape in the dentate granule neurons of adult mice (Guo et al., 2011; Su et al., 2017). Accordingly, contextual fear conditioning induces changes in DNA methylation in plasticity genes that are required for the formation and maintenance of memory (Halder et al., 2016). Learning also induced exon-specific methylation in the bdnf gene, which correlated with increased bdnf gene expression and memory consolidation (Lubin et al., 2008). Serotonin-dependent methylation in the promoter of CREB2 results in a reduction in CREB2 expression and enhances memory-related synaptic plasticity (Rajasethupathy et al., 2012). Additionally, learning induces persistent DNA methylation of the memory suppressors CaN (Miller et al., 2010) and PP1 (Miller and Sweatt, 2007), resulting in a reduction of target gene expression and facilitates memory consolidation. Active DNA demethylation has also been observed in the regulatory region ofbdnf, Npas4, and fgf1 (Martinowich et al., 2003; Nelson et al., 2008; Ma et al., 2009; Rudenko et al., 2013). For example, contextual fear conditioning induces demethylation of the synaptic plasticity gene reelin, which is correlated with upregulation of reelin expression and memory formation (Miller and Sweatt, 2007).

Activity-dependent modification of cytosine hydroxymethylation (5hmC) is widely observed in the brain. This modification is another DNA modification that is carried out via hydroxylation of methylated cytosines (5mC) by members of the ten-11 translocation (TET) protein family (Guo et al., 2011). Neuronal activity regulates TET1 expression, thus leading to global changes in modified cytosine levels and expression of activity-dependent genes and memory formation (Kaas et al., 2013).

Activity-dependent epigenetic regulation also involves the binding of chromatin regulators. The methyl-CpG binding protein 2 (MeCP2) recognizes and represses methylated genes by recruiting chromatin remodeling factors such as HDACs, REST and CoREST (59. Neural activity induces a Ca2+-dependent phosphorylation of MeCP2 that results in its release from the bdnf gene promoter IV and increased bdnf transcription, correlating with a reduction in DNA methylation at the promoter region (Chen et al., 2003; Martinowich et al., 2003; Zhou et al., 2006).

Activity-dependent regulation of alternative splicing by epigenetic factors

Besides transcriptional regulation, epigenetic regulators also determine the choice between alternative splice isoforms. First, neural activation can regulate splicing by directly recruiting snRNP proteins via epigenetic factors. Chromatin remodelers in SWI/SNF complexes have an effect on alternative splicing that depends on their physical interactions and recruitment of snRNPs U1 and U5 but that is independent of their ATPase remodeling activity (Batsché et al., 2006). Furthermore, the histone acetyltransferase STAGA shows direct interaction with U2 snRNPs (Martinez et al., 2001; Cheng et al., 2007), and the histone arginine methyltransferase CARM1 physically interacts with U1 snRNP proteins (Cheng et al., 2007).

Histone modifications are also associated with alternative splicing. A genome-wide analysis showed that histone marks are non-randomly distributed and several types of histone modifications are enriched in exons compared to the flanking introns (Andersson et al., 2009; Kolasinska-Zwierz et al., 2009; Spies et al., 2009; Schwartz et al., 2009). Specifically, FGFR2 showed tissue-specific splicing isoforms, and H3K36me3 and H3K4me3 was specifically enriched along the alternatively spliced region according to the respective splicing pattern (Luco et al., 2010). Interestingly, modulation of H3K36me3 or H3K4me3 levels by interfering with the expression of their respective histone methyltransferases causes splice site switching in a predictable fashion(Luco et al., 2010), suggesting that histone modifications are critical for the regulation of alternative splicing. Additionally, treatment of cell culture with the histone deacetylase inhibitor TSA induces skipping of the alternatively splicedfibronectin exon 33 and NCAM exon 18 (Nogues et al., 2002; Schor et al., 2009).

Furthermore, histone marks affect splicing through the recruitment of splicing regulators via chromatin binding proteins to form chromatin-splicing adaptor systems. High levels of H3K36me3 along the alternatively spliced region attract MRG15 which, in turn, interacts with PTB and recruits it to the nascent RNA and thereby regulating splicing (Luco et al., 2010). H3K4me3 levels affect the pattern of splicing through the bridging of the spliceosome and the alternative spliced region via CHD1(Sims et al., 2007). Another example of chromatin-splicing adaptor systems is that H3K9 trimethylation plays a role in recruiting splicing factors hnRNPs via the chromatin-adaptor protein HP1(Piacentini et al., 2009).

Recent studies showed that epigenetic regulation of alternative splicing can be triggered by neural activity (Ding et al., 2017). Especially in the memory trace neurons, neural activity induced phosphorylation of p66avia 5′AMP-activated protein kinase (AMPK) to recruit HDAC2 and Suv39h1, thereby establishing repressive epigenetic markers on theNrxn1 SS4 site and affecting co-transcriptional Nrxn1 SS4 splicing. Furthermore, disrupting the build-up of intragenic H3K9me3 by knockout of Suv39h1 abolished the activity-dependent splicing changes (Ding et al., 2017). This finding suggests that local histone modification is a one of the key regulators of activity-dependentNrxn1 SS4 splicing, which supports the idea that alternative splicing is epigenetically regulated through a transcription-kinetic-dependent mechanism (Luco et al., 2011). Thus, neural activity regulates gene transcription and modulates splicing in the central nervous system and thus modifies the epigenetic landscape.

Bi-directional regulation of epigenetic modifications creates stable states of epigenetic landscapes in the cellular genome

DNA methylation and post-translational histone modifications constitute a layer of stable epigenetic information. Many of these markers can be maintained for a long time (Luco et al., 2011; Probst et al., 2009). DNA (cytosine-5)-methyltransferase 3 alpha (DNMT3a)/DNA (cytosine-5)-methyltransferase 3 beta (DNMT3b) in association with a cofactor DNMT3L are essential for the establishment of DNA methylation patterns through de novo methylation (Okano et al., 1999; Jia et al., 2007). In vitro methylated DNA templates in cell culture retained methylated-DNA regardless of the DNA sequence, even after many rounds of cell divisions (Pollack et al., 1980; Wigler et al., 1981). Because the DNMT1 enzyme has a high specificity for hemimethylated CpG dinucleotides, unmethylated sites will not be recognized, thus maintaining the precise methylation pattern on the newly synthesized DNA (Gruenbaum et al., 1982). The maintenance of DNA methylation not only depends on the properties of DNMT1 itself (Cheng, 2014), but that the methyl-CpG binding protein, MeCP2, interacts directly with DNMT1 within the hemimethylated sites to perform maintenance methylation in vivo (Kimura and Shiota, 2003). Although DNA methylation can be stably maintained, it can also be erased or re-modified in response to external cues (Reik, 2007).

For the maintenance of post-translational histone modifications, parental histones are used as templates to guide the modification of new histones (Nakatani et al., 2004; Probst et al., 2009). Such a mechanism is widely used in repetitive regions, such as in heterochromatin, in which high density of H3K9me3 is bound by HP1 proteins. HP1, together with DNA methylation and H3K9me3, contribute to the maintenance of repressed state (Bannister et al., 2001; Lachner et al., 2001). A subset of the histone H3 lysine 9 methyltransferases Suv39h1, G9a, GLP, and SETDB1 also participate in the process of heterochromatin formation as a multimeric complex (Fritsch et al., 2010). With respect to the regions of heterochromatin, the process for maintaining H3K27me3 is carried by PRC2, the enzyme that catalyzes this modification and directly binds to H3K27me3 (Hansen and Helin, 2009; Margueron et al., 2009). Histone deacetylation and DNA methylation often work together to stabilize the epigenetic modifications (Vaute et al., 2002; Scharf et al., 2009). HDACs together with H3K9me3 ensure the maintenance of a deacetylated state of chromatin, while in regions without HDAC targeting, it may be easier to maintain histone acetylation (Vaute et al., 2002). Histone modifications also work in concert with DNA methylation. DNMT1 associates with histone deacetylase (Fuks et al., 2000); additionally, methyl-CpG binding protein 1 (MBD1) interacts with the histone methyltransferase Suv39h1(Fujita et al., 2003), which provides an additional means to perpetuate the exact chromatin states.

Epigenetic regulators in neural differentiation

Epigenetic regulators play essential roles in neural differentiation. The central nervous system consists of many cell types, including neurons, astrocytes, and oligodendrocytes, all of which are differentiated from the same neural stem cells (NSCs) (Dietrich et al., 2006). The migration and differentiation of neural cells must be strictly regulated during development by extracellular cues and intracellular gene expression programs, which are, in part, regulated by epigenetic mechanisms(Mizutani et al., 2007; Namihira et al., 2008). Epigenetic mechanisms are key regulators for both pluripotency maintenance and cell fate specification (Hirabayashi and Gotoh, 2010).

The promoter region of genes such as sodium channel type II, BDNF or calbindin are highly methylated to prevent differentiation into neurons during the early stage (Lunyak et al., 2002; Ballas et al., 2005). Neuronal specification of NSCs requires the de-repression of neuronal genes such as Sox2 via DNA demethylation in the promoter region (Sikorska et al., 2008). Astrocytic gene loci are also silenced by DNA methylation during neuronal commitment, and this silencing is attenuated by demethylation of the genes coding for the astrocytic markers such asGFAP (Sun et al., 2001; Takizawa et al., 2001; Namihira et al., 2004; Fan et al., 2005). The proper DNA methylation pattern during each stage of development is coordinated with a tight regulation of the DNA methyltransferases, DNMT1, DNMT3a, and DNMT3b.DNMT1 is highly expressed in NSCs (Brooks et al., 1996), and studies in DNMT1-deficient NSCs showed enhanced astrogliogenesis (Fan et al., 2005). Furthermore, depletion of DNMT3a and DNMT3b lead to precocious glial differentiation (Wu et al., 2012) and failed neuronal differentiation in vitro (Bai et al., 2005).

In addition to DNA methylation, histone modifications are also regulated during development, both spatially and temporally. Histone deacetylation represses the expression ofMash1, an important regulator of cell fate decision in NSCs; however, upon neural differentiation, Mash1 is actively expressed via histone acetylation (Williams et al., 2006). Other neural genes such as NeuroD and Cdkn1c are also activated via histone acetylation during neuronal fate commitment (Sun et al., 2001; Attia et al., 2007).

Epigenetic regulation of the neural circuit assembly

Epigenetic regulators also play an essential role in the formation of functional neural circuit assemblies in the mature brain. Failure to assemble proper neural circuits is associated with neurodevelopmental disorders including intellectual disability and autism spectrum disorders (Geschwind and Levitt, 2007; Gogolla et al., 2009; Wood and Shepherd, 2010). In addition to transcription factors, epigenetic mechanisms are critical for regulation of gene expression during the development of the neural assembly (Ho and Crabtree, 2010; Ronan et al., 2013). In the cerebellum, as well as in the hippocampus, granule neurons form ample dendrites and then prune them. Presynaptic boutons were formed along axons of granule neurons during maturation and integration of the circuits (Yamada et al., 2014; Yang et al., 2016). Conditional knockout of the NuRD component or Suv39h1 in granule neurons disrupts the formation of presynaptic boutons and dendrite elimination. Conditional knockout of Chd4, a chromatin helicase in the NuRD complex, impairs synaptic neurotransmission between granule neurons and its downstream target in Purkinje cells responding to sensorimotor stimuli in the cerebellum.

Epigenetic regulation of learning and memory

Activity-dependent synaptic dynamics are crucial for learning and memory formation (Fortin et al., 2012). Early research on learning and memory formation emphasized the role of transcription factors (Chen et al., 2012); however, epigenetic mechanisms have recently emerged as a key player in the regulation of synaptic plasticity and memory formation (Gräff et al., 2012). Histone deacetylases such as HDAC2, HDAC3, and HDAC6, has been shown to regulate the critical genes related to regulation of plasticity and synapse formation, thereby modulating memory formation and consolidation (Levenson et al., 2004; Chwang et al., 2006; Gupta et al., 2010; Gräff et al., 2012; Gupta-Agarwal et al., 2012; Ding et al., 2017). The histone variant H2A.Z is altered in response to fear conditioning in the hippocampus and the cortex, inhibiting memory formation through downstream effects on gene expression(Zovkic et al., 2014). Moreover, a significant increase in the incorporation of H3.3 into the chromatin of active genes was observed in hippocampal neurons, while blocking the turnover of H3.3 impairs hippocampus-based learning (Maze et al., 2015). Genome-wide analyses of the hippocampal DNA methylation status after learning show that DNA methylation is dynamically regulated, including both de novo methylation and demethylation (Martinowich et al., 2003; Miller and Sweatt, 2007; Lubin et al., 2008; Miller et al., 2010; Halder et al., 2016). Both pharmacological inhibition and conditional knockout of DNA methyltransferases resulted in an impairment in synaptic plasticity and long-term memory formation and maintenance (Levenson et al., 2006; Miller and Sweatt, 2007; Lubin et al., 2008; Miller et al., 2010; Morris et al., 2014; Mitchnick et al., 2015). Therefore, epigenetic regulation encompasses different mechanisms, including histone modification, histone variants (Kamakaka and Biggins, 2005; Zovkic et al., 2014), and DNA methylation, all of which help regulate gene expression for memory formation and consolidation.

Recent progress showed that Suv39h1, a histone methyltransferase for H3K9me3, regulates Nrxn1 SS4 inclusion in response to learning engaged neural activity, that is essential for memory preservation probably through constraining their connection specificities as well as their plasticity in the memory circuit (Ding et al., 2017). Knockout of HDAC2 or inhibition of HDAC2 with HDACi converts stable memory into an alterable state (Gräff et al., 2014).

Transgenerational Inheritance

Many psychiatric and neurological disorders have strong genetic heritable components (Kendler, 2001; Millan et al., 2012); however, the genome-wide association studies performed to date have failed to identify the causal genetic basis of these disorders (Gibson, 2012), and the key factors of their heritability are still unknown (Eichler et al., 2010; Gershon et al., 2011; So et al., 2011). It has recently been recognized that, in addition to genetically inherited information, non-genetic components, such as epigenetics, may also contribute to disease heritability (Danchin et al., 2011; Bohacek and Mansuy, 2013). Previously, epigenetic modifications were considered to be completely erased between generations; however, several studies have confirmed that epigenetic information can be transmitted to subsequent generations through the germline, a process termed transgenerational epigenetic inheritance (Horsthemke, 2007; Daxinger and Whitelaw, 2010).

DNA methylation is well known to be involved in one form of epigenetic inheritance: genomic imprinting, a process governed by DNA methylation that allows the selective expression of only one parental allele (maternal or paternal) (Paoloni-Giacobino and Chaillet, 2006; Sha, 2008). Sex-specific DNA hypermethylated imprints are essential for silencing of the inactive allele and are protected from global demethylation activity during the following fertilization (Feng et al., 2010; Bartolomei and Ferguson-Smith, 2011). In addition to genomic imprinting, DNA methylation in germ cells that is altered by environmental stresses at specific gene loci can be transmitted between generations (Franklin et al., 2010).

Histone modifications modify chromatin structure and play an essential role in gene expression in somatic tissues, while their function in sperm cells is less clear. Histones are mainly replaced by protamines (up to 98% in mice and 85% in human) in sperm cells (Hammoud et al., 2009; Johnson et al., 2011); however, the remaining histones are specifically retained at genetic loci that are essential for embryogenesis (Puri et al., 2010). H3K4me2 and H3K27me3 are present at functional genes in spermatogenesis and developmental regulation and maintain genes in either an activated or repressed state (Hammoud et al., 2009; Brykczynska et al., 2010). Interestingly, genes repressed by H3K27me3 in sperm cells are maintained in a repressed state in the early embryo (Brykczynska et al., 2010), suggesting that H3K27me3 in sperm can be inherited across generations.

Epigenetic dysregulation in neurological diseases

Environmentally induced epigenetic modifications in neurological disorders

The environment exerts some of its effects on disease progression through epigenetics. Genetically identical individuals in monozygotic twins can show discordancy for neurological diseases such as autism (Hallmayer et al., 2011), schizophrenia (Cannon et al., 1998), and Alzheimer’s disease (Gatz et al., 1997), suggesting a contribution from the environment to the disease through alterations of the epigenome of the individual. Studies have demonstrated the existence of epigenetic differences in monozygotic twins (Fraga et al., 2005).

Exposure to early life stress (ELS) has been thought to alter gene expression programs and enhance the risk of psychopathologies such as schizophrenia, bipolar disorder, depression, and PTSD (Gershon et al., 2013). Epigenetic modifications have the ability to modulate gene expression in response to external factors, and provide a potential mechanism for such programming. ELS pups show DNA hypermethylation and reduction of H3K9ac in promoter region ofGrm1 and Gad1, correlating with decreased expression of targeted genes (Zhang et al., 2010; Bagot et al., 2012). Additionally, maltreated pups show an increase in DNA methylation at the bdnf promoter and a corresponding decrease in bdnf expression in the adult prefrontal cortex (Roth et al., 2009), which has been found in schizophrenia patients (Hashimoto et al., 2005).

Monogenetic neurological diseases associated with epigenetic defects

Notably, many neurological disorders are linked to mutations of epigenetic regulators, including DNA methyltransferase and histone modifying enzymes (Table 2). Those mutations indicate the essential role of epigenetic regulation in these diseases. Interestingly, the list includes not only embryonic defects, but also neurological disorders with symptom onset in childhood and in adults, suggesting the epigenetic regulation is essential both for the development of the nerve system and for the physiologic function in matured brain.

Disordered chromatin in neurological disease

Many neurological diseases showed correlation to the activity-dependent epigenetic regulations. Interestingly, reports showed that treatment with sodium butyrate, the histone deacetylase inhibitor, ameliorated the neurodegenerative phenotype in Huntington's disease mice (Steffan et al., 2001; Ferrante et al., 2003), suggesting the epigenetic regulators might also serve as the drug targets to treat neurological diseases.

Autism spectrum disorder

A significant number of genetic syndromes of ASD are associated with mutations in epigenetic regulators (Crawford et al., 2001; Beyer et al., 2002; Richards et al., 2015). For example, mutations of CHD8, which encodes the chromatin modifier, show strong association (>87%) with the ASD phenotype (Bernier et al., 2014; Merner et al., 2016). Interestingly, mutations in CHD7 also showed a significant risk (40%) for ASD (Smith et al., 2005; Johansson et al., 2006). Furthermore, abnormal DNA methylation is also detected in ASD patients, including the dysregulation of DNA methylation in the 3′ untranslated region of PRRT1, TSPAN32 and C11orf21(Ladd-Acosta et al., 2014; Nardone et al., 2014) and in promoters of GAD65, OXTR, SHANK3, reelin, UBE3A and MECP2 (Jiang et al., 2004; Nagarajan et al., 2006; Gregory et al., 2009; Zhu et al., 2014; Elagoz Yuksel et al., 2016). For histone modifications, excess expansion of H3K4me3 and H3K27Ac from the transcription start sites into downstream gene bodies and upstream promoters have been identified in ASD patients (Shulha et al., 2012; Sun et al., 2016).

Alzheimer disease

Alteration of the epigenome has been understood in the aging process and in age-related neurodegenerative diseases (Hernandez et al., 2011; Lu et al., 2013). Recent findings implicate the dysregulation of REST and REST-dependent epigenetic remodelling is associated with cognitive impairment and Alzheimer disease (Lu et al., 2014). Epigenome-wide association studies assessed the methylation state of the brain's DNA in relation to Alzheimer's disease (De Jager et al., 2014; Lunnon et al., 2014). Differentially methylated regions are found at four loci: ANK1, CDH23, RHBDF2 and RPL13, inducing a corresponding RNA expression alteration. Moreover, alterations in global levels of DNA methylation and hydroxymethylation were found in monozygotic twins discordant for AD (Mastroeni et al., 2009; Chouliaras et al., 2013). Additionally, DNA hypo- and hyper-methylation of genes that are implicated in AD pathology have been found in AD brains, such as TMEM59 and PSEN1 (Wang et al., 2008; Bakulski et al., 2012). Considering the later onset of these diseases, the defects of activity-dependent epigenetic regulation might contribute to its pathology.

Huntington disease

The interaction of mutant huntingtin with the transcriptional machinery and miRNA-mediated gene silencing complexes results in transcriptional silencing, which is central to the pathophysiology of Huntington disease (Savas et al., 2008). The dysregulation of REST and REST-dependent epigenetic remodelling described above was associated with Huntington disease through regulating gene silencing (Buckley et al., 2010). Huntingtin binds to REST, thus inhibiting its translocation to the nucleus to regulate neuronal survival (Zuccato et al., 2003). The CAG repeats in mutant huntingtin disrupts the binding of huntingtin to REST, resulting in reduced transcription of REST target genes (Zuccato et al., 2003, 2007). Furthermore, polycomb proteins have also been implicated in Huntington disease. Loss of PRC2 induces the upregulation of genes involved in Huntington disease (von Schimmelmann et al., 2016). Lastly, genome-wide association studies have shown a deficiency in 5-hydroxymethylcytosine in a mouse model of Huntington disease. The aberrant methylation and silencing of genes are involved in neurogenesis, neuronal function and survival in HD brain (Wang et al., 2013).

Conclusion and remarks

Although epigenetic changes are observed in many diseases, further research is required to better understand the learning induced epigenetic modifications in the genome of functional neural circuits. The various states of the transcriptional program, which are determined by epigenetic regulators, need to be further characterized. Furthermore, there is still a lack of studies dissecting the signaling pathways that transduce the external signals to the epigenetic modifications.

Nonetheless, accumulating evidence has shown the critical role of epigenetics in regulating transcription and disease conditions. Drugs targeting epigenetic factors show efficiency in treating multiple brain disorders, including AD and depression. To develop more potent and selective epigenetic treatments against brain disorders, a better understanding of the epigenetic machinery in disease pathogenesis is required.

References

[1]

Devor AAndreassen  O AWang  YMäki-Marttunen T, Smeland O B Fan C C Schork A J Holland D Thompson W K Witoelar A Chen C H Desikan R S McEvoy L K Djurovic S Greengard P Svenningsson P Einevoll G T Dale A M  (2017). Genetic evidence for role of integration of fast and slow neurotransmission in schizophrenia. Mol Psychiatry22(6): 792–801

[2]

De Rubeis SHe  XGoldberg A P Poultney C S Samocha K Cicek A E Kou YLiu  LFromer M Walker S Singh T Klei LKosmicki  JShih-Chen F Aleksic B Biscaldi M Bolton P F Brownfeld J M Cai JCampbell  N GCarracedo  AChahrour M H Chiocchetti A G Coon HCrawford  E LCurran  S RDawson  GDuketis E Fernandez B A Gallagher L Geller E Guter S J Hill R S Ionita-Laza J Jimenz Gonzalez P Kilpinen H Klauck S M Kolevzon A Lee ILei  ILei J Lehtimäki T Lin C F Ma’ayan A Marshall C R McInnes A L Neale B Owen M J Ozaki N Parellada M Parr J R Purcell S Puura K Rajagopalan D Rehnström K Reichenberg A Sabo ASachse  MSanders S J Schafer C Schulte-Rüther M Skuse D Stevens C Szatmari P Tammimies K Valladares O Voran A Li-San W Weiss L A Willsey A J Yu T W Yuen R K Cook E H Freitag C M Gill MHultman  C MLehner  TPalotie A Schellenberg G D Sklar P State M W Sutcliffe J S Walsh C A Scherer S W Zwick M E Barett J C Cutler D J Roeder K Devlin B Daly M J Buxbaum J D , and the DDD Study, and the Homozygosity Mapping Collaborative for Autism, and the UK10K Consortium (2014). Synaptic, transcriptional and chromatin genes disrupted in autism. Nature515(7526): 209–215

[3]

Stessman H AXiong  BCoe B P Wang THoekzema  KFenckova M Kvarnung M Gerdts J Trinh S Cosemans N Vives L Lin JTurner  T NSanten  GRuivenkamp C Kriek M van Haeringen A Aten EFriend  KLiebelt J Barnett C Haan EShaw  MGecz J Anderlid B M    Nordgren A Lindstrand A Schwartz C Kooy R F Vandeweyer G Helsmoortel C Romano C Alberti A Vinci M Avola E Giusto S Courchesne E Pramparo T Pierce K Nalabolu S Amaral D G Scheffer I E    Delatycki M B Lockhart P J Hormozdiari F Harich B Castells-Nobau A Xia KPeeters  HNordenskjöld MSchenck A Bernier R A Eichler E E  (2017). Targeted sequencing identifies 91 neurodevelopmental-disorder risk genes with autism and developmental-disability biases. Nat Genet49(4): 515–526

[4]

Guan J SHaggarty  S JGiacometti  EDannenberg J H Joseph N Gao JNieland  T JZhou  YWang X Mazitschek R Bradner J E DePinho R A Jaenisch R Tsai L H  (2009). HDAC2 negatively regulates memory formation and synaptic plasticity. Nature459(7243): 55–60

[5]

Gräff JRei  DGuan J S Wang W Y Seo JHennig  K MNieland  T JFass  D MKao  P FKahn  MSu S C Samiei A Joseph N Haggarty S J Delalle I Tsai L H  (2012). An epigenetic blockade of cognitive functions in the neurodegenerating brain. Nature483(7388): 222–226

[6]

Waddington C H  (2012). The epigenotype. 1942. Int J Epidemiol41(1): 10–13

[7]

Bird A (2007). Perceptions of epigenetics. Nature447(7143): 396–398

[8]

Bird A (2002). DNA methylation patterns and epigenetic memory. Genes Dev16(1): 6–21

[9]

Laird P W (2003). The power and the promise of DNA methylation markers. Nat Rev Cancer3(4): 253–266

[10]

Ramsahoye B H Biniszkiewicz D Lyko FClark  VBird A P Jaenisch R  (2000). Non-CpG methylation is prevalent in embryonic stem cells and may be mediated by DNA methyltransferase 3a. Proc Natl Acad Sci USA97(10): 5237–5242

[11]

Matzke M ABirchler  J A (2005). RNAi-mediated pathways in the nucleus. Nat Rev Genet6(1): 24–35

[12]

Jenuwein TAllis  C D (2001). Translating the histone code. Science293(5532): 1074–1080

[13]

Kouzarides T (2007). Chromatin modifications and their function. Cell128(4): 693–705

[14]

Pusarla R HBhargava  P (2005). Histones in functional diversification. Core histone variants. FEBS J272(20): 5149–5168

[15]

Talbert P BHenikoff  S (2010). Histone variants--ancient wrap artists of the epigenome. Nat Rev Mol Cell Biol11(4): 264–275

[16]

Becker P BHörz  W (2002). ATP-dependent nucleosome remodeling. Annu Rev Biochem71(1): 247–273

[17]

Clapier C RCairns  B R (2009). The biology of chromatin remodeling complexes. Annu Rev Biochem78(1): 273–304

[18]

Borrelli ENestler  E JAllis  C DSassone-Corsi  P (2008). Decoding the epigenetic language of neuronal plasticity. Neuron60(6): 961–974

[19]

Roth T LSweatt  J D (2009). Regulation of chromatin structure in memory formation. Curr Opin Neurobiol19(3): 336–342

[20]

Sweatt J D (2016). GENE EXPRESSION. Chromatin controls behavior. Science353(6296): 218–219

[21]

Zovkic I BSweatt  J D (2015). Memory-Associated Dynamic Regulation of the “Stable” Core of the Chromatin Particle. Neuron87(1): 1–4

[22]

Sweatt J D (2013). The emerging field of neuroepigenetics. Neuron80(3): 624–632

[23]

Kim T KHemberg  MGray J M Costa A M Bear D M Wu JHarmin  D ALaptewicz  MBarbara-Haley K Kuersten S Markenscoff-Papadimitriou E Kuhl DBito  HWorley P F Kreiman G Greenberg M E  (2010). Widespread transcription at neuronal activity-regulated enhancers. Nature465(7295): 182–187

[24]

Sui LWang  YJu L H Chen M (2012). Epigenetic regulation of reelin and brain-derived neurotrophic factor genes in long-term potentiation in rat medial prefrontal cortex. Neurobiol Learn Mem97(4): 425–440

[25]

Malik A NVierbuchen  THemberg M Rubin A A Ling ECouch  C HStroud  HSpiegel I Farh K K Harmin D A Greenberg M E  (2014). Genome-wide identification and characterization of functional neuronal activity-dependent enhancers. Nat Neurosci17(10): 1330–1339

[26]

Gräff JWoldemichael  B TBerchtold  DDewarrat G Mansuy I M  (2012). Dynamic histone marks in the hippocampus and cortex facilitate memory consolidation. Nat Commun3: 991

[27]

Naruse YOh-hashi  KIijima N Naruse M Yoshioka H Tanaka M  (2004). Circadian and light-induced transcription of clock gene Per1 depends on histone acetylation and deacetylation. Mol Cell Biol24(14): 6278–6287

[28]

Martinowich KHattori  DWu H Fouse S He FHu  YFan G Sun Y E  (2003). DNA methylation-related chromatin remodeling in activity-dependent BDNF gene regulation. Science302(5646): 890–893

[29]

Guo J UMa  D KMo  HBall M P Jang M H Bonaguidi M A Balazer J A Eaves H L Xie BFord  EZhang K Ming G L Gao YSong  H (2011). Neuronal activity modifies the DNA methylation landscape in the adult brain. Nat Neurosci14(10): 1345–1351

[30]

Crosio CHeitz  EAllis C D Borrelli E Sassone-Corsi P  (2003). Chromatin remodeling and neuronal response: multiple signaling pathways induce specific histone H3 modifications and early gene expression in hippocampal neurons. J Cell Sci116(Pt 24): 4905–4914

[31]

Levenson J MO’Riordan  K JBrown  K DTrinh  M AMolfese  D LSweatt  J D (2004). Regulation of histone acetylation during memory formation in the hippocampus. J Biol Chem279(39): 40545–40559

[32]

Crosio CCermakian  NAllis C D Sassone-Corsi P  (2000). Light induces chromatin modification in cells of the mammalian circadian clock. Nat Neurosci3(12): 1241–1247

[33]

Dyrvig MHansen  H HChristiansen  S HWoldbye  D PMikkelsen  J DLichota  J (2012). Epigenetic regulation of Arc and c-Fos in the hippocampus after acute electroconvulsive stimulation in the rat. Brain Res Bull88(5): 507–513

[34]

Gupta SKim  S YArtis  SMolfese D L Schumacher A Sweatt J D Paylor R E Lubin F D  (2010). Histone methylation regulates memory formation. J Neurosci30(10): 3589–3599

[35]

Gupta-Agarwal SFranklin  A VDeramus  TWheelock M Davis R L McMahon L L Lubin F D  (2012). G9a/GLP histone lysine dimethyltransferase complex activity in the hippocampus and the entorhinal cortex is required for gene activation and silencing during memory consolidation. J Neurosci32(16): 5440–5453

[36]

Chwang W BO’Riordan  K JLevenson  J MSweatt  J D (2006). ERK/MAPK regulates hippocampal histone phosphorylation following contextual fear conditioning. Learn Mem13(3): 322–328

[37]

Miller C ASweatt  J D (2007). Covalent modification of DNA regulates memory formation. Neuron53(6): 857–869

[38]

Lubin F DRoth  T LSweatt  J D (2008). Epigenetic regulation of BDNF gene transcription in the consolidation of fear memory. J Neurosci28(42): 10576–10586

[39]

Miller C AGavin  C FWhite  J AParrish  R RHonasoge  AYancey C R Rivera I M Rubio M D Rumbaugh G Sweatt J D  (2010). Cortical DNA methylation maintains remote memory. Nat Neurosci13(6): 664–666

[40]

Ma D KJang  M HGuo  J UKitabatake  YChang M L Pow-Anpongkul N Flavell R A Lu BMing  G LSong  H (2009). Neuronal activity-induced Gadd45b promotes epigenetic DNA demethylation and adult neurogenesis. Science323(5917): 1074–1077

[41]

Chen W GChang  QLin Y Meissner A West A E Griffith E C Jaenisch R Greenberg M E  (2003). Derepression of BDNF transcription involves calcium-dependent phosphorylation of MeCP2. Science302(5646): 885–889

[42]

Zhou ZHong  E JCohen  SZhao W N Ho H Y Schmidt L Chen W G Lin YSavner  EGriffith E C Hu LSteen  J AWeitz  C JGreenberg  M E (2006). Brain-specific phosphorylation of MeCP2 regulates activity-dependent Bdnf transcription, dendritic growth, and spine maturation. Neuron52(2): 255–269

[43]

Kaas G AZhong  CEason D E Ross D L Vachhani R V Ming G L King J R Song HSweatt  J D (2013). TET1 controls CNS 5-methylcytosine hydroxylation, active DNA demethylation, gene transcription, and memory formation. Neuron79(6): 1086–1093

[44]

Zhu TLiang  CLi D Tian MLiu  SGao G Guan J S  (2016). Histone methyltransferase Ash1L mediates activity-dependent repression of neurexin-1α. Sci Rep6(1): 26597160;

[45]

Ding XLiu  STian M Zhang W Zhu TLi  DWu J Deng HJia  YXie W Xie HGuan  J S (2017). Activity-induced histone modifications govern Neurexin-1 mRNA splicing and memory preservation. Nat Neurosci20(5): 690–699

[46]

Su YShin  JZhong C Wang SRoychowdhury  PLim J Kim DMing  G LSong  H (2017). Neuronal activity modifies the chromatin accessibility landscape in the adult brain. Nat Neurosci20(3): 476–483

[47]

Rudenko ADawlaty  M MSeo  JCheng A W Meng JLe  TFaull K F Jaenisch R Tsai L H  (2013). Tet1 is critical for neuronal activity-regulated gene expression and memory extinction. Neuron79(6): 1109–1122

[48]

Gräff JJoseph  N FHorn  M ESamiei  AMeng J Seo JRei  DBero A W Phan T X Wagner F Holson E Xu JSun  JNeve R L Mach R H Haggarty S J Tsai L H  (2014). Epigenetic priming of memory updating during reconsolidation to attenuate remote fear memories. Cell156(1-2): 261–276

[49]

Zovkic I BPaulukaitis  B SDay  J JEtikala  D MSweatt  J D (2014). Histone H2A.Z subunit exchange controls consolidation of recent and remote memory. Nature515(7528): 582–586

[50]

Halder RHennion  MVidal R O Shomroni O Rahman R U Rajput A Centeno T P van Bebber F Capece V Garcia Vizcaino J C Schuetz A L Burkhardt S Benito E Navarro Sala M Javan S B Haass C Schmid B Fischer A Bonn S (2016). DNA methylation changes in plasticity genes accompany the formation and maintenance of memory. Nat Neurosci19(1): 102–110 

[51]

Nelson E DKavalali  E TMonteggia  L M (2008). Activity-dependent suppression of miniature neurotransmission through the regulation of DNA methylation. J Neurosci28(2): 395–406

[52]

Rajasethupathy PAntonov  ISheridan R Frey SSander  CTuschl T Kandel E R  (2012). A role for neuronal piRNAs in the epigenetic control of memory-related synaptic plasticity. Cell149(3): 693–707

[53]

Meadows J PGuzman-Karlsson  M CPhillips  SHolleman C Posey J L Day J J Hablitz J J Sweatt J D  (2015). DNA methylation regulates neuronal glutamatergic synaptic scaling. Sci Signal8(382): ra61

[54]

Südhof T C  (2008). Neuroligins and neurexins link synaptic function to cognitive disease. Nature455(7215): 903–911

[55]

Banerjee TChakravarti  D (2011). A peek into the complex realm of histone phosphorylation. Mol Cell Biol31(24): 4858–4873

[56]

Goll M GBestor  T H (2005). Eukaryotic cytosine methyltransferases. Annu Rev Biochem74(1): 481–514

[57]

Su, Y.Shin,  J.Zhong, C.  &  Wang,  S. Neuronal activity modifies the chromatin accessibility landscape in the adult brain. 20, 476–483, doi:10.1038/nn.4494 (2017).

[58]

Guo J USu  YZhong C Ming G L Song H (2011). Emerging roles of TET proteins and 5-hydroxymethylcytosines in active DNA demethylation and beyond. Cell Cycle10(16): 2662–2668

[59]

59 (!!! INVALID CITATION !!!).

[60]

Batsché EYaniv  MMuchardt C  (2006). The human SWI/SNF subunit Brm is a regulator of alternative splicing. Nat Struct Mol Biol13(1): 22–29

[61]

Martinez EPalhan  V BTjernberg  ALymar E S Gamper A M Kundu T K Chait B T Roeder R G  (2001). Human STAGA complex is a chromatin-acetylating transcription coactivator that interacts with pre-mRNA splicing and DNA damage-binding factors in vivo. Mol Cell Biol21(20): 6782–6795

[62]

Cheng DCôté  JShaaban S Bedford M T  (2007). The arginine methyltransferase CARM1 regulates the coupling of transcription and mRNA processing. Mol Cell25(1): 71–83

[63]

Kolasinska-Zwierz P Down TLatorre  ILiu T Liu X S Ahringer J  (2009). Differential chromatin marking of introns and expressed exons by H3K36me3. Nat Genet41(3): 376–381

[64]

Spies NNielsen  C BPadgett  R ABurge  C B (2009). Biased chromatin signatures around polyadenylation sites and exons. Mol Cell36(2): 245–254

[65]

Andersson REnroth  SRada-Iglesias A Wadelius C Komorowski J  (2009). Nucleosomes are well positioned in exons and carry characteristic histone modifications. Genome Res19(10): 1732–1741

[66]

Schwartz SMeshorer  EAst G  (2009). Chromatin organization marks exon-intron structure. Nat Struct Mol Biol16(9): 990–995

[67]

Luco R FPan  QTominaga K Blencowe B J Pereira-Smith O M Misteli T  (2010). Regulation of alternative splicing by histone modifications. Science327(5968): 996–1000

[68]

Nogues GKadener  SCramer P Bentley D Kornblihtt A R  (2002). Transcriptional activators differ in their abilities to control alternative splicing. J Biol Chem277(45): 43110–43114

[69]

Schor I ERascovan  NPelisch F Alló M Kornblihtt A R  (2009). Neuronal cell depolarization induces intragenic chromatin modifications affecting NCAM alternative splicing. Proc Natl Acad Sci USA106(11): 4325–4330

[70]

Sims R J 3rd, Millhouse SChen  C FLewis  B AErdjument-Bromage  HTempst P Manley J L Reinberg D  (2007). Recognition of trimethylated histone H3 lysine 4 facilitates the recruitment of transcription postinitiation factors and pre-mRNA splicing. Mol Cell28(4): 665–676

[71]

Piacentini LFanti  LNegri R Del Vescovo V Fatica A Altieri F Pimpinelli S  (2009). Heterochromatin protein 1 (HP1a) positively regulates euchromatic gene expression through RNA transcript association and interaction with hnRNPs in Drosophila. PLoS Genet5(10): e1000670

[72]

Luco R FAllo  MSchor I E Kornblihtt A R Misteli T  (2011). Epigenetics in alternative pre-mRNA splicing. Cell144(1): 16–26

[73]

Rountree M RBachman  K EHerman  J GBaylin  S B (2001). DNA methylation, chromatin inheritance, and cancer. Oncogene20(24): 3156–3165

[74]

Probst A VDunleavy  EAlmouzni G  (2009). Epigenetic inheritance during the cell cycle. Nat Rev Mol Cell Biol10(3): 192–206

[75]

Okano MBell  D WHaber  D ALi  E (1999). DNA methyltransferases Dnmt3a and Dnmt3b are essential for de novo methylation and mammalian development. Cell99(3): 247–257

[76]

Jia DJurkowska  R ZZhang  XJeltsch A Cheng X  (2007). Structure of Dnmt3a bound to Dnmt3L suggests a model for de novo DNA methylation. Nature449(7159): 248–251

[77]

Pollack YStein  RRazin A Cedar H  (1980). Methylation of foreign DNA sequences in eukaryotic cells. Proc Natl Acad Sci USA77(11): 6463–6467

[78]

Wigler MLevy  DPerucho M  (1981). The somatic replication of DNA methylation. Cell24(1): 33–40

[79]

Gruenbaum YCedar  HRazin A  (1982). Substrate and sequence specificity of a eukaryotic DNA methylase. Nature295(5850): 620–622

[80]

Cheng X (2014). Structural and functional coordination of DNA and histone methylation. Cold Spring Harb Perspect Biol6(8): a018747

[81]

Kimura HShiota  K (2003). Methyl-CpG-binding protein, MeCP2, is a target molecule for maintenance DNA methyltransferase, Dnmt1. J Biol Chem278(7): 4806–4812

[82]

Reik W (2007). Stability and flexibility of epigenetic gene regulation in mammalian development. Nature447(7143): 425–432

[83]

Nakatani YRay-Gallet  DQuivy J P Tagami H Almouzni G  (2004). Two distinct nucleosome assembly pathways: dependent or independent of DNA synthesis promoted by histone H3.1 and H3.3 complexes. Cold Spring Harb Symp Quant Biol69(0): 273–280

[84]

Bannister A J Zegerman P Partridge J F Miska E A Thomas J O Allshire R C Kouzarides T  (2001). Selective recognition of methylated lysine 9 on histone H3 by the HP1 chromo domain. Nature410(6824): 120–124

[85]

Lachner MO’Carroll  DRea S Mechtler K Jenuwein T  (2001). Methylation of histone H3 lysine 9 creates a binding site for HP1 proteins. Nature410(6824): 116–120

[86]

Fritsch LRobin  PMathieu J R Souidi M Hinaux H Rougeulle C Harel-Bellan A Ameyar-Zazoua M Ait-Si-Ali S  (2010). A subset of the histone H3 lysine 9 methyltransferases Suv39h1, G9a, GLP, and SETDB1 participate in a multimeric complex. Mol Cell37(1): 46–56

[87]

Hansen K HHelin  K (2009). Epigenetic inheritance through self-recruitment of the polycomb repressive complex 2. Epigenetics4(3): 133–138

[88]

Margueron RJustin  NOhno K Sharpe M L Son JDrury  W J 3rdVoigt  PMartin S R Taylor W R De Marco V Pirrotta V Reinberg D Gamblin S J  (2009). Role of the polycomb protein EED in the propagation of repressive histone marks. Nature461(7265): 762–767

[89]

Vaute ONicolas  EVandel L Trouche D  (2002). Functional and physical interaction between the histone methyl transferase Suv39H1 and histone deacetylases. Nucleic Acids Res30(2): 475–481

[90]

Scharf A NMeier  KSeitz V Kremmer E Brehm A Imhof A  (2009). Monomethylation of lysine 20 on histone H4 facilitates chromatin maturation. Mol Cell Biol29(1): 57–67

[91]

Fuks FBurgers  W ABrehm  AHughes-Davies L Kouzarides T  (2000). DNA methyltransferase Dnmt1 associates with histone deacetylase activity. Nat Genet24(1): 88–91

[92]

Fujita NWatanabe  SIchimura T Tsuruzoe S Shinkai Y Tachibana M Chiba T Nakao M  (2003). Methyl-CpG binding domain 1 (MBD1) interacts with the Suv39h1-HP1 heterochromatic complex for DNA methylation-based transcriptional repression. J Biol Chem278(26): 24132–24138

[93]

Dietrich JHan  RYang Y Mayer-Pröschel M Noble M  (2006). CNS progenitor cells and oligodendrocytes are targets of chemotherapeutic agents in vitro and in vivo. J Biol5(7): 22

[94]

Mizutani KYoon  KDang L Tokunaga A Gaiano N  (2007). Differential Notch signalling distinguishes neural stem cells from intermediate progenitors. Nature449(7160): 351–355

[95]

Namihira MKohyama  JAbematsu M Nakashima K  (2008). Epigenetic mechanisms regulating fate specification of neural stem cells. Philos Trans R Soc Lond B Biol Sci363(1500): 2099–2109

[96]

Hirabayashi YGotoh  Y (2010). Epigenetic control of neural precursor cell fate during development. Nat Rev Neurosci11(6): 377–388

[97]

Lunyak V VBurgess  RPrefontaine G G Nelson C Sze S H Chenoweth J Schwartz P Pevzner P A Glass C Mandel G Rosenfeld M G  (2002). Corepressor-dependent silencing of chromosomal regions encoding neuronal genes. Science298(5599): 1747–1752

[98]

Ballas NGrunseich  CLu D D Speh J C Mandel G  (2005). REST and its corepressors mediate plasticity of neuronal gene chromatin throughout neurogenesis. Cell121(4): 645–657

[99]

Sikorska MSandhu  J KDeb-Rinker  PJezierski A Leblanc J Charlebois C Ribecco-Lutkiewicz M Bani-Yaghoub M Walker P R  (2008). Epigenetic modifications of SOX2 enhancers, SRR1 and SRR2, correlate with in vitro neural differentiation. J Neurosci Res86(8): 1680–1693

[100]

Sun YNadal-Vicens  MMisono S Lin M Z Zubiaga A Hua XFan  GGreenberg M E  (2001). Neurogenin promotes neurogenesis and inhibits glial differentiation by independent mechanisms. Cell104(3): 365–376

[101]

Takizawa TNakashima  KNamihira M Ochiai W Uemura A Yanagisawa M Fujita N Nakao M Taga T (2001). DNA methylation is a critical cell-intrinsic determinant of astrocyte differentiation in the fetal brain. Dev Cell1(6): 749–758

[102]

Namihira MNakashima  KTaga T  (2004). Developmental stage dependent regulation of DNA methylation and chromatin modification in a immature astrocyte specific gene promoter. FEBS Lett572(1-3): 184–188

[103]

Fan GMartinowich  KChin M H He FFouse  S DHutnick  LHattori D Ge WShen  YWu H ten Hoeve J Shuai K Sun Y E  (2005). DNA methylation controls the timing of astrogliogenesis through regulation of JAK-STAT signaling. Development132(15): 3345–3356

[104]

Brooks P JMarietta  CGoldman D  (1996). DNA mismatch repair and DNA methylation in adult brain neurons. J Neurosci16(3): 939–945

[105]

Wu ZHuang  KYu J Le TNamihira  MLiu Y Zhang J Xue ZCheng  LFan G  (2012). Dnmt3a regulates both proliferation and differentiation of mouse neural stem cells. J Neurosci Res90(10): 1883–1891

[106]

Bai SGhoshal  KDatta J Majumder S Yoon S O Jacob S T  (2005). DNA methyltransferase 3b regulates nerve growth factor-induced differentiation of PC12 cells by recruiting histone deacetylase 2. Mol Cell Biol25(2): 751–766

[107]

Williams R RAzuara  VPerry P Sauer S Dvorkina M Jørgensen H Roix JMcQueen  PMisteli T Merkenschlager M Fisher A G  (2006). Neural induction promotes large-scale chromatin reorganisation of the Mash1 locus. J Cell Sci119(Pt 1): 132–140

[108]

Attia MRachez  CDe Pauw A Avner P Rogner U C  (2007). Nap1l2 promotes histone acetylation activity during neuronal differentiation. Mol Cell Biol27(17): 6093–6102

[109]

Gogolla NLeblanc  J JQuast  K BSüdhof  T CFagiolini  MHensch T K  (2009). Common circuit defect of excitatory-inhibitory balance in mouse models of autism. J Neurodev Disord1(2): 172–181

[110]

Geschwind D H Levitt P  (2007). Autism spectrum disorders: developmental disconnection syndromes. Curr Opin Neurobiol17(1): 103–111

[111]

Wood LShepherd  G M (2010). Synaptic circuit abnormalities of motor-frontal layer 2/3 pyramidal neurons in a mutant mouse model of Rett syndrome. Neurobiol Dis38(2): 281–287

[112]

Ho LCrabtree  G R (2010). Chromatin remodelling during development. Nature463(7280): 474–484

[113]

Ronan J LWu  WCrabtree G R  (2013). From neural development to cognition: unexpected roles for chromatin. Nat Rev Genet14(5): 347–359

[114]

Yamada TYang  YHemberg M Yoshida T Cho H Y Murphy J P Fioravante D Regehr W G Gygi S P Georgopoulos K Bonni A  (2014). Promoter decommissioning by the NuRD chromatin remodeling complex triggers synaptic connectivity in the mammalian brain. Neuron83(1): 122–134

[115]

Yang YYamada  THill K K Hemberg M Reddy N C Cho H Y Guthrie A N Oldenborg A Heiney S A Ohmae S Medina J F Holy T E Bonni A  (2016). Chromatin remodeling inactivates activity genes and regulates neural coding. Science353(6296): 300–305

[116]

Fortin D ASrivastava  TSoderling T R  (2012). Structural modulation of dendritic spines during synaptic plasticity. Neuroscientist18(4): 326–341

[117]

Chen D YBambah-Mukku  DPollonini G Alberini C M  (2012). Glucocorticoid receptors recruit the CaMKIIα-BDNF-CREB pathways to mediate memory consolidation. Nat Neurosci15(12): 1707–1714

[118]

Ding, X.Activity-induced histone modifications govern Neurexin-1 mRNA splicing and memory preservation. 20, 690–699160;(2017).

[119]

Maze IWenderski  WNoh K M Bagot R C Tzavaras N Purushothaman I Elsässer S J Guo YIonete  CHurd Y L Tamminga C A Halene T Farrelly L Soshnev A A Wen DRafii  SBirtwistle M R Akbarian S Buchholz B A Blitzer R D Nestler E J Yuan Z F Garcia B A Shen LMolina  HAllis C D  (2015). Critical Role of Histone Turnover in Neuronal Transcription and Plasticity. Neuron87(1): 77–94

[120]

Levenson J MRoth  T LLubin  F DMiller  C AHuang  I CDesai  PMalone L M Sweatt J D  (2006). Evidence that DNA (cytosine-5) methyltransferase regulates synaptic plasticity in the hippocampus. J Biol Chem281(23): 15763–15773

[121]

Morris M JAdachi  MNa E S Monteggia L M  (2014). Selective role for DNMT3a in learning and memory. Neurobiol Learn Mem115: 30–37

[122]

Mitchnick K A Creighton S O’Hara M Kalisch B E Winters B D  (2015). Differential contributions of de novo and maintenance DNA methyltransferases to object memory processing in the rat hippocampus and perirhinal cortex--a double dissociation. Eur J Neurosci41(6): 773–786

[123]

Kamakaka R TBiggins  S (2005). Histone variants: deviants? Genes Dev19(3): 295–310

[124]

Kendler K S (2001). Twin studies of psychiatric illness: an update. Arch Gen Psychiatry58(11): 1005–1014

[125]

Millan M JAgid  YBrüne M Bullmore E T Carter C S Clayton N S Connor R Davis S Deakin B DeRubeis R J Dubois B Geyer M A Goodwin G M Gorwood P Jay T M Joëls M Mansuy I M Meyer-Lindenberg A Murphy D Rolls E Saletu B Spedding M Sweeney J Whittington M Young L J  (2012). Cognitive dysfunction in psychiatric disorders: characteristics, causes and the quest for improved therapy. Nat Rev Drug Discov11(2): 141–168

[126]

Gibson G (2012). Rare and common variants: twenty arguments. Nat Rev Genet13(2): 135–145

[127]

Eichler E EFlint  JGibson G Kong ALeal  S MMoore  J HNadeau  J H (2010). Missing heritability and strategies for finding the underlying causes of complex disease. Nat Rev Genet11(6): 446–450

[128]

Gershon E SAlliey-Rodriguez  NLiu C  (2011). After GWAS: searching for genetic risk for schizophrenia and bipolar disorder. Am J Psychiatry168(3): 253–256

[129]

So H CGui  A HCherny  S SSham  P C (2011). Evaluating the heritability explained by known susceptibility variants: a survey of ten complex diseases. Genet Epidemiol35(5): 310–317

[130]

BohacekJMansuy  I M(2013).Epigenetic inheritance of disease and disease risk.  Neuropsychopharmacology38: 220–236

[131]

Danchin ÉCharmantier  AChampagne F A Mesoudi A Pujol B Blanchet S  (2011). Beyond DNA: integrating inclusive inheritance into an extended theory of evolution. Nat Rev Genet12(7): 475–486

[132]

Daxinger LWhitelaw  E (2010). Transgenerational epigenetic inheritance: more questions than answers. Genome Res20(12): 1623–1628

[133]

Horsthemke B (2007). Heritable germline epimutations in humans. Nat Genet39(5): 573–574, author reply 575–576

[134]

Sha K (2008). A mechanistic view of genomic imprinting. Annu Rev Genomics Hum Genet9(1): 197–216

[135]

Paoloni-Giacobino A Chaillet J R  (2006). The role of DMDs in the maintenance of epigenetic states. Cytogenet Genome Res113(1-4): 116–121

[136]

Bartolomei M S Ferguson-Smith A C  (2011). Mammalian genomic imprinting. Cold Spring Harb Perspect Biol3(7): a002592

[137]

Feng SJacobsen  S EReik  W (2010). Epigenetic reprogramming in plant and animal development. Science330(6004): 622–627

[138]

Franklin T BRussig  HWeiss I C Gräff J Linder N Michalon A Vizi SMansuy  I M (2010). Epigenetic transmission of the impact of early stress across generations. Biol Psychiatry68(5): 408–415

[139]

Johnson G DLalancette  CLinnemann A K Leduc F Boissonneault G Krawetz S A  (2011). The sperm nucleus: chromatin, RNA, and the nuclear matrix. Reproduction141(1): 21–36

[140]

Hammoud S SNix  D AZhang  HPurwar J Carrell D T Cairns B R  (2009). Distinctive chromatin in human sperm packages genes for embryo development. Nature460(7254): 473–478 

[141]

Puri DDhawan  JMishra R K  (2010). The paternal hidden agenda: Epigenetic inheritance through sperm chromatin. Epigenetics5(5): 386–391

[142]

Brykczynska UHisano  MErkek S Ramos L Oakeley E J Roloff T C Beisel C Schübeler D Stadler M B Peters A H  (2010). Repressive and active histone methylation mark distinct promoters in human and mouse spermatozoa. Nat Struct Mol Biol17(6): 679–687

[143]

Hallmayer JCleveland  STorres A Phillips J Cohen B Torigoe T Miller J Fedele A Collins J Smith K Lotspeich L Croen L A Ozonoff S Lajonchere C Grether J K Risch N  (2011). Genetic heritability and shared environmental factors among twin pairs with autism. Arch Gen Psychiatry68(11): 1095–1102

[144]

Cannon T DKaprio  JLönnqvist J Huttunen M Koskenvuo M  (1998). The genetic epidemiology of schizophrenia in a Finnish twin cohort. A population-based modeling study. Arch Gen Psychiatry55(1): 67–74

[145]

Gatz MPedersen  N LBerg  SJohansson B Johansson K Mortimer J A Posner S F Viitanen M Winblad B Ahlbom A  (1997). Heritability for Alzheimer’s disease: the study of dementia in Swedish twins. J Gerontol A Biol Sci Med Sci52(2): M117–M125

[146]

Fraga M FBallestar  EPaz M F Ropero S Setien F Ballestar M L Heine-Suñer D Cigudosa J C Urioste M Benitez J Boix-Chornet M Sanchez-Aguilera A Ling CCarlsson  EPoulsen P Vaag AStephan  ZSpector T D Wu Y Z Plass C Esteller M  (2005). Epigenetic differences arise during the lifetime of monozygotic twins. Proc Natl Acad Sci USA102(30): 10604–10609

[147]

Gershon ASudheimer  KTirouvanziam R Williams L M O’Hara R  (2013). The long-term impact of early adversity on late-life psychiatric disorders. Curr Psychiatry Rep15(4): 352

[148]

Bagot R CZhang  T YWen  XNguyen T T Nguyen H B Diorio J Wong T P Meaney M J  (2012). Variations in postnatal maternal care and the epigenetic regulation of metabotropic glutamate receptor 1 expression and hippocampal function in the rat. Proc Natl Acad Sci USA109(Suppl 2): 17200–17207

[149]

Zhang T YHellstrom  I CBagot  R CWen  XDiorio J Meaney M J  (2010). Maternal care and DNA methylation of a glutamic acid decarboxylase 1 promoter in rat hippocampus. J Neurosci30(39): 13130–13137

[150]

Roth T LLubin  F DFunk  A JSweatt  J D (2009). Lasting epigenetic influence of early-life adversity on the BDNF gene. Biol Psychiatry65(9): 760–769

[151]

Hashimoto TBergen  S ENguyen  Q LXu  BMonteggia L M Pierri J N Sun ZSampson  A RLewis  D A (2005). Relationship of brain-derived neurotrophic factor and its receptor TrkB to altered inhibitory prefrontal circuitry in schizophrenia. J Neurosci25(2): 372–383

[152]

Jin BTao  QPeng J Soo H M Wu WYing  JFields C R Delmas A L Liu XQiu  JRobertson K D  (2008). DNA methyltransferase 3B (DNMT3B) mutations in ICF syndrome lead to altered epigenetic modifications and aberrant expression of genes regulating development, neurogenesis and immune function. Hum Mol Genet17(5): 690–709

[153]

Jowaed ASchmitt  IKaut O Wüllner U  (2010). Methylation regulates alpha-synuclein expression and is decreased in Parkinson’s disease patients’ brains. J Neurosci30(18): 6355–6359

[154]

Winkelmann JLin  LSchormair B Kornum B R Faraco J Plazzi G Melberg A Cornelio F Urban A E Pizza F Poli FGrubert  FWieland T Graf EHallmayer  JStrom T M Mignot E  (2012). Mutations in DNMT1 cause autosomal dominant cerebellar ataxia, deafness and narcolepsy. Hum Mol Genet21(10): 2205–2210

[155]

Chestnut B AChang  QPrice A Lesuisse C Wong MMartin  L J (2011). Epigenetic regulation of motor neuron cell death through DNA methylation. J Neurosci31(46): 16619–16636

[156]

Amir R EVan den Veyver  I BWan  MTran C Q Francke U Zoghbi H Y  (1999). Rett syndrome is caused by mutations in X-linked MECP2, encoding methyl-CpG-binding protein 2. Nat Genet23(2): 185–188

[157]

Mnatzakanian G N Lohi HMunteanu  IAlfred S E Yamada T MacLeod P J Jones J R Scherer S W Schanen N C Friez M J Vincent J B Minassian B A  (2004). A previously unidentified MECP2 open reading frame defines a new protein isoform relevant to Rett syndrome. Nat Genet36(4): 339–341

[158]

Chen R ZAkbarian  STudor M Jaenisch R  (2001). Deficiency of methyl-CpG binding protein-2 in CNS neurons results in a Rett-like phenotype in mice. Nat Genet27(3): 327–331

[159]

Collins A LLevenson  J MVilaythong  A PRichman  RArmstrong D L Noebels J L David Sweatt J Zoghbi H Y  (2004). Mild overexpression of MeCP2 causes a progressive neurological disorder in mice. Hum Mol Genet13(21): 2679–2689

[160]

Guy JHendrich  BHolmes M Martin J E Bird A (2001). A mouse Mecp2-null mutation causes neurological symptoms that mimic Rett syndrome. Nat Genet27(3): 322–326

[161]

Carney R MWolpert  C MRavan  S AShahbazian  MAshley-Koch A Cuccaro M L Vance J M Pericak-Vance M A  (2003). Identification of MeCP2 mutations in a series of females with autistic disorder. Pediatr Neurol28(3): 205–211

[162]

Kleefstra Tvan Zelst-Stams  W ANillesen  W MCormier-Daire  VHouge G Foulds N van Dooren M Willemsen M H Pfundt R Turner A Wilson M McGaughran J Rauch A Zenker M Adam M P Innes M Davies C López A G Casalone R Weber A Brueton L A Navarro A D Bralo M P Venselaar H Stegmann S P Yntema H G van Bokhoven H Brunner H G  (2009). Further clinical and molecular delineation of the 9q subtelomeric deletion syndrome supports a major contribution of EHMT1 haploinsufficiency to the core phenotype. J Med Genet46(9): 598–606

[163]

Kirov GPocklington  A JHolmans  PIvanov D Ikeda M Ruderfer D Moran J Chambert K Toncheva D Georgieva L Grozeva D Fjodorova M Wollerton R Rees ENikolov  Ivan de Lagemaat  L NBayés  AFernandez E Olason P I Böttcher Y Komiyama N H Collins M O Choudhary J Stefansson K Stefansson H Grant S G Purcell S Sklar P O’Donovan M C Owen M J  (2012). De novo CNV analysis implicates specific abnormalities of postsynaptic signalling complexes in the pathogenesis of schizophrenia. Mol Psychiatry17(2): 142–153

[164]

Roelfsema J H Peters D J  (2007). Rubinstein-Taybi syndrome: clinical and molecular overview. Expert Rev Mol Med9(23): 1–16

[165]

Zollino MOrteschi  DMurdolo M Lattante S Battaglia D Stefanini C Mercuri E Chiurazzi P Neri GMarangi  G (2012). Mutations in KANSL1 cause the 17q21.31 microdeletion syndrome phenotype. Nat Genet44(6): 636–638

[166]

Michelson D J Shevell M I Sherr E H Moeschler J B Gropman A L Ashwal S  (2011). Evidence report: Genetic and metabolic testing on children with global developmental delay: report of the Quality Standards Subcommittee of the American Academy of Neurology and the Practice Committee of the Child Neurology Society. Neurology77(17): 1629–1635

[167]

Adegbola AGao  HSommer S Browning M  (2008). A novel mutation in JARID1C/SMCX in a patient with autism spectrum disorder (ASD). Am J Med Genet A146A(4): 505–511

[168]

Berdasco MRopero  SSetien F Fraga M F Lapunzina P Losson R Alaminos M Cheung N K Rahman N Esteller M  (2009). Epigenetic inactivation of the Sotos overgrowth syndrome gene histone methyltransferase NSD1 in human neuroblastoma and glioma. Proc Natl Acad Sci USA106(51): 21830–21835

[169]

Kleine-Kohlbrecher D Christensen J Vandamme J Abarrategui I Bak MTommerup  NShi X Gozani O Rappsilber J Salcini A E Helin K  (2010). A functional link between the histone demethylase PHF8 and the transcription factor ZNF711 in X-linked mental retardation. Mol Cell38(2): 165–178

[170]

Pereira P MSchneider  APannetier S Heron D Hanauer A  (2010). Coffin-Lowry syndrome. Eur J Hum Genet18(6): 627–633

[171]

Gibson W THood  R LZhan  S HBulman  D EFejes  A PMoore  RMungall A J Eydoux P Babul-Hirji R An JMarra  M AChitayat  DBoycott K M Weaver D D Jones S J , and the FORGE Canada Consortium (2012). Mutations in EZH2 cause Weaver syndrome. Am J Hum Genet90(1): 110–118

[172]

Jones W DDafou  DMcEntagart M Woollard W J Elmslie F V Holder-Espinasse M Irving M Saggar A K Smithson S Trembath R C Deshpande C Simpson M A  (2012). De novo mutations in MLL cause Wiedemann-Steiner syndrome. Am J Hum Genet91(2): 358–364

[173]

Ng S BBigham  A WBuckingham  K JHannibal  M CMcMillin  M JGildersleeve  H IBeck  A ETabor  H KCooper  G MMefford  H CLee  CTurner E H Smith J D Rieder M J Yoshiura K Matsumoto N Ohta TNiikawa  NNickerson D A Bamshad M J Shendure J  (2010). Exome sequencing identifies MLL2 mutations as a cause of Kabuki syndrome. Nat Genet42(9): 790–793

[174]

Campeau P MKim  J CLu  J TSchwartzentruber  J AAbdul-Rahman  O ASchlaubitz  SMurdock D M Jiang M M Lammer E J Enns G M Rhead W J Rowland J Robertson S P Cormier-Daire V Bainbridge M N Yang X J Gingras M C Gibbs R A Rosenblatt D S Majewski J Lee B H  (2012). Mutations in KAT6B, encoding a histone acetyltransferase, cause Genitopatellar syndrome. Am J Hum Genet90(2): 282–289

[175]

Lederer DGrisart  BDigilio M C Benoit V Crespin M Ghariani S C Maystadt I Dallapiccola B Verellen-Dumoulin C  (2012). Deletion of KDM6A, a histone demethylase interacting with MLL2, in three patients with Kabuki syndrome. Am J Hum Genet90(1): 119–124

[176]

Williams S RAldred  M ADer Kaloustian  V MHalal  FGowans G McLeod D R Zondag S Toriello H V Magenis R E Elsea S H  (2010). Haploinsufficiency of HDAC4 causes brachydactyly mental retardation syndrome, with brachydactyly type E, developmental delays, and behavioral problems. Am J Hum Genet87(2): 219–228

[177]

Iossifov IRonemus  MLevy D Wang ZHakker  IRosenbaum J Yamrom B Lee Y H Narzisi G Leotta A Kendall J Grabowska E Ma BMarks  SRodgers L Stepansky A Troge J Andrews P Bekritsky M Pradhan K Ghiban E Kramer M Parla J Demeter R Fulton L L Fulton R S Magrini V J Ye KDarnell  J CDarnell  R BMardis  E RWilson  R KSchatz  M CMcCombie  W RWigler  M (2012). De novo gene disruptions in children on the autistic spectrum. Neuron74(2): 285–299

[178]

Steffan J SBodai  LPallos J Poelman M McCampbell A Apostol B L Kazantsev A Schmidt E Zhu Y Z Greenwald M Kurokawa R Housman D E Jackson G R Marsh J L Thompson L M  (2001). Histone deacetylase inhibitors arrest polyglutamine-dependent neurodegeneration in Drosophila. Nature413(6857): 739–743

[179]

Ferrante R JKubilus  J KLee  JRyu H Beesen A Zucker B Smith K Kowall N W Ratan R R Luthi-Carter R Hersch S M  (2003). Histone deacetylase inhibition by sodium butyrate chemotherapy ameliorates the neurodegenerative phenotype in Huntington’s disease mice. J Neurosci23(28): 9418–9427

[180]

Richards CJones  CGroves L Moss JOliver  C (2015). Prevalence of autism spectrum disorder phenomenology in genetic disorders: a systematic review and meta-analysis. Lancet Psychiatry2(10): 909–916

[181]

Beyer K SBlasi  FBacchelli E Klauck S M Maestrini E Poustka A ,Molecular Genetic Study of Autism C I, and the International Molecular Genetic Study of Autism Consortium (IMGSAC) (2002). Mutation analysis of the coding sequence of the MECP2 gene in infantile autism. Hum Genet111(4-5): 305–309

[182]

Crawford D CAcuna J MSherman S L 2001). FMR1 and the fragile X syndrome: human genome epidemiology review. Genet Med3: 359–371

[183]

Bernier RGolzio  CXiong B Stessman H A Coe B P Penn OWitherspoon  KGerdts J Baker C Vulto-van Silfhout A T Schuurs-Hoeijmakers J H Fichera M Bosco P Buono S Alberti A Failla P Peeters H Steyaert J Vissers L E Francescatto L Mefford H C Rosenfeld J A Bakken T O’Roak B J Pawlus M Moon RShendure  JAmaral D G Lein ERankin  JRomano C de Vries B B Katsanis N Eichler E E  (2014). Disruptive CHD8 mutations define a subtype of autism early in development. Cell158(2): 263–276

[184]

Merner NForgeot d’Arc  BBell S C Maussion G Peng HGauthier  JCrapper L Hamdan F F Michaud J L Mottron L Rouleau G A Ernst C  (2016). A de novo frameshift mutation in chromodomain helicase DNA-binding domain 8 (CHD8): A case report and literature review. Am J Med Genet A170A(5): 1225–1235

[185]

Johansson MRåstam  MBillstedt E Danielsson S Strömland K Miller M Gillberg C  (2006). Autism spectrum disorders and underlying brain pathology in CHARGE association. Dev Med Child Neurol48(1): 40–50

[186]

Smith I MNichols  S LIssekutz  KBlake K , and the Canadian Paediatric Surveillance Program (2005). Behavioral profiles and symptoms of autism in CHARGE syndrome: preliminary Canadian epidemiological data. Am J Med Genet A133A(3): 248–256

[187]

Ladd-Acosta CHansen  K DBriem  EFallin M D Kaufmann W E Feinberg A P  (2014). Common DNA methylation alterations in multiple brain regions in autism. Mol Psychiatry19(8): 862–871

[188]

Nardone SSams  D SReuveni  EGetselter D Oron OKarpuj  MElliott E  (2014). DNA methylation analysis of the autistic brain reveals multiple dysregulated biological pathways. Transl Psychiatry4(9): e433

[189]

Elagoz Yuksel M Yuceturk B Karatas O F Ozen MDogangun  B (2016). The altered promoter methylation of oxytocin receptor gene in autism. J Neurogenet30(3-4): 280–284

[190]

Gregory S GConnelly  J JTowers  A JJohnson  JBiscocho D Markunas C A Lintas C Abramson R K Wright H H Ellis P Langford C F Worley G Delong G R Murphy S K Cuccaro M L Persico A Pericak-Vance M A  (2009). Genomic and epigenetic evidence for oxytocin receptor deficiency in autism. BMC Med7(1): 62

[191]

Jiang Y HSahoo  TMichaelis R C Bercovich D Bressler J Kashork C D Liu QShaffer  L GSchroer  R JStockton  D WSpielman  R SStevenson  R EBeaudet  A L (2004). A mixed epigenetic/genetic model for oligogenic inheritance of autism with a limited role for UBE3A. Am J Med Genet A131(1): 1–10 

[192]

Nagarajan R P Hogart A R Gwye YMartin  M RLaSalle  J M (2006). Reduced MeCP2 expression is frequent in autism frontal cortex and correlates with aberrant MECP2 promoter methylation. Epigenetics1(4): e1–e11

[193]

Zhu LWang  XLi X L Towers A Cao XWang  PBowman R Yang HGoldstein  JLi Y J Jiang Y H  (2014). Epigenetic dysregulation of SHANK3 in brain tissues from individuals with autism spectrum disorders. Hum Mol Genet23(6): 1563–1578

[194]

Shulha H PCheung  IWhittle C Wang JVirgil  DLin C L Guo YLessard  AAkbarian S Weng Z (2012). Epigenetic signatures of autism: trimethylated H3K4 landscapes in prefrontal neurons. Arch Gen Psychiatry69(3): 314–324

[195]

Sun WPoschmann  JCruz-Herrera Del Rosario  RParikshak N N Hajan H S Kumar V Ramasamy R Belgard T G Elanggovan B Wong C C Mill JGeschwind  D HPrabhakar  S (2016). Histone Acetylome-wide Association Study of Autism Spectrum Disorder. Cell167(5): 1385–1397.e11

[196]

Hernandez D G Nalls M A Gibbs J R Arepalli S van der Brug M Chong S Moore M Longo D L Cookson M R Traynor B J Singleton A B  (2011). Distinct DNA methylation changes highly correlated with chronological age in the human brain. Hum Mol Genet20(6): 1164–1172

[197]

Lu HLiu  XDeng Y Qing H (2013). DNA methylation, a hand behind neurodegenerative diseases. Front Aging Neurosci5: 85

[198]

Lu TAron  LZullo J Pan YKim  HChen Y Yang T H Kim H M Drake D Liu X S Bennett D A Colaiácovo M P Yankner B A  (2014). REST and stress resistance in ageing and Alzheimer’s disease. Nature507(7493): 448–454

[199]

De Jager P L Srivastava G Lunnon K Burgess J Schalkwyk L C Yu LEaton  M LKeenan  B TErnst  JMcCabe C Tang ARaj  TReplogle J Brodeur W Gabriel S Chai H S Younkin C Younkin S G Zou FSzyf  MEpstein C B Schneider J A Bernstein B E Meissner A Ertekin-Taner N Chibnik L B Kellis M Mill JBennett  D A (2014). Alzheimer’s disease: early alterations in brain DNA methylation at ANK1, BIN1, RHBDF2 and other loci. Nat Neurosci17(9): 1156–1163

[200]

Lunnon KSmith  RHannon E De Jager P L Srivastava G Volta M Troakes C Al-Sarraj S Burrage J Macdonald R Condliffe D Harries L W Katsel P Haroutunian V Kaminsky Z Joachim C Powell J Lovestone S Bennett D A Schalkwyk L C Mill J (2014). Methylomic profiling implicates cortical deregulation of ANK1 in Alzheimer’s disease. Nat Neurosci17(9): 1164–1170

[201]

Chouliaras LMastroeni  DDelvaux E Grover A Kenis G Hof P R Steinbusch H W Coleman P D Rutten B P van den Hove D L  (2013). Consistent decrease in global DNA methylation and hydroxymethylation in the hippocampus of Alzheimer’s disease patients. Neurobiol Aging34(9): 2091–2099

[202]

Mastroeni DMcKee  AGrover A Rogers J Coleman P D  (2009). Epigenetic differences in cortical neurons from a pair of monozygotic twins discordant for Alzheimer’s disease. PLoS One4(8): e6617

[203]

Wang S COelze  BSchumacher A  (2008). Age-specific epigenetic drift in late-onset Alzheimer’s disease. PLoS One3(7): e2698

[204]

Bakulski K MDolinoy  D CSartor  M APaulson  H LKonen  J RLieberman  A PAlbin  R LHu  HRozek L S  (2012). Genome-wide DNA methylation differences between late-onset Alzheimer’s disease and cognitively normal controls in human frontal cortex. J Alzheimers Dis29(3): 571–588 

[205]

Savas J NMakusky  AOttosen S Baillat D Then FKrainc  DShiekhattar R Markey S P Tanese N  (2008). Huntington’s disease protein contributes to RNA-mediated gene silencing through association with Argonaute and P bodies. Proc Natl Acad Sci USA105(31): 10820–10825

[206]

Buckley N JJohnson  RZuccato C Bithell A Cattaneo E  (2010). The role of REST in transcriptional and epigenetic dysregulation in Huntington’s disease. Neurobiol Dis39(1): 28–39

[207]

Zuccato CTartari  MCrotti A Goffredo D Valenza M Conti L Cataudella T Leavitt B R Hayden M R Timmusk T Rigamonti D Cattaneo E  (2003). Huntingtin interacts with REST/NRSF to modulate the transcription of NRSE-controlled neuronal genes. Nat Genet35(1): 76–83

[208]

Zuccato CBelyaev  NConforti P Ooi LTartari  MPapadimou E MacDonald M Fossale E Zeitlin S Buckley N Cattaneo E  (2007). Widespread disruption of repressor element-1 silencing transcription factor/neuron-restrictive silencer factor occupancy at its target genes in Huntington’s disease. J Neurosci27(26): 6972–6983

[209]

von Schimmelmann M Feinberg P A Sullivan J M Ku S M Badimon A Duff M K Wang ZLachmann  ADewell S Ma’ayan A Han M H Tarakhovsky A Schaefer A  (2016). Polycomb repressive complex 2 (PRC2) silences genes responsible for neurodegeneration. Nat Neurosci19(10): 1321–1330

[210]

Wang FYang  YLin X Wang J Q Wu Y S Xie WWang  DZhu S Liao Y Q Sun QYang  Y GLuo  H RGuo  CHan C Tang T S  (2013). Genome-wide loss of 5-hmC is a novel epigenetic feature of Huntington’s disease. Hum Mol Genet22(18): 3641–3653

RIGHTS & PERMISSIONS

Higher Education Press and Springer-Verlag Berlin Heidelberg

AI Summary AI Mindmap
PDF (559KB)

1421

Accesses

0

Citation

Detail

Sections
Recommended

AI思维导图

/