Open access

Epigenetic Regulation of Neural Differentiation from Embryonic Stem Cells

Written By

Atsushi Shimomura and Eri Hashino

Submitted: 26 March 2012 Published: 27 March 2013

DOI: 10.5772/53650

From the Edited Volume

Trends in Cell Signaling Pathways in Neuronal Fate Decision

Edited by Sabine Wislet-Gendebien

Chapter metrics overview

2,393 Chapter Downloads

View Full Metrics

1. Introduction

During development, ESCs in the inner cell mass of the blastocyst undergo progressive fate restriction and sequentially give rise to tissue-specific multipotent progenitor cells [1-3]. In the nervous system, neural progenitor cells (NPCs) are defined as multipotent cells capable of differentiating into neurons and glia, the latter of which include astrocytes and oligodendrocytes. NPCs undergo sequential fate restriction during nervous system development [4-6]. During early embryonic development, NPCs give rise exclusively to neuronal cell types, however at later stages they differentiate into glial cell types [7, 8]. Understanding how ESCs differentiate into neuronal or glial cell types, thus, constitutes a fundamental aspect of nervous system development.

ESCs possess two cardinal cellular characteristics: self-renewal capacity and pluripotency [2, 3]. ESCs, when undifferentiated, maintain a unique gene expression profile which keeps themselves in a pluripotent state [3, 9-12]. When ESCs begin to differentiate, however, this profile is dramatically altered without changes in the DNA sequence, allowing initial cell lineage specification to take place [11, 13-15]. Studies on lineage choice and differentiation of stem cells suggest that the crosstalk between transcription factors and epigenetic mechanisms plays pivotal roles in maintaining the existing transcriptional profile or changing it into a new transcriptional profile [16, 17].

Chromatin in eukaryotes is organized into arrays of nucleosomes, which are comprised of DNA wrapped around a histone octamer containing 2 copies of individual core histones, H2A, H2B, H3, and H4 [18]. Chromatin can be tightly or loosely packed to alter its accessibility to transcription factors and RNA polymerases, thereby ultimately modulating the efficiency of gene transcription [19-23]. Epigenetic mechanisms act to change this accessibility both locally and globally through modifications of nucleosomes, and by remodeling of nucleosomes [21, 24, 25]. In addition to epigenetic modifications and chromatin remodeling, epigenetic regulators have been recently extended to non-coding RNAs (ncRNAs), because ncRNAs can affect chromatin structure and transcriptional activation by regulating expression of key nucleosome modifiers [26]. These epigenetic controls appear to influence gene expression profiles, which are essential for self-renewal and differentiation capacities of ESCs. Thus, a clear understanding of the epigenetic mechanisms underlying gene expression patterns will provide significant and novel insights into cell fate specification of ESCs directed to differentiate into neurons. Furthermore, the epigenetic mechanisms are believed to be capable of responding to extrinsic signals such as morphogens and cytokines [8]. Therefore, knowledge of the epigenetic mechanisms is also important for our understanding of neural differentiation by extrinsic factors. In this review, we will describe the major epigenetic processes that underlie the acquisition of the NPC fate from ESCs as well as the subsequent neuronal subtype specification. The focus of this review is weighted on neuronal cell lineage specification, and not on glial cell specification.

Advertisement

2. Epigentic mechanisms

2.1. Post-translational modifications of histones

Post-translational chemical modifications of histones in chromatin are diverse and include methylation, acetylation, phosphorylation, sumoylation, ubiquitilation, and ADP-ribosylation [25, 27]. Here, we will focus on histone methylation and acetylation because these modifications are considered to be most important and widespread for influencing biological processes during neural differentiation.

Histone methylation occurs mainly at lysine and arginine residues on the tails of the histone H3 and H4 [28]. Histone methylation is reversible. Methylation at these sites has been associated with not only transcriptional activation and repression [27, 29] but also multiple biological processes, including heterochromatin formation and genomic imprinting [27, 28]. The differences in the effects of methylation depend on the lysine residues. For example, methylation of histone H3 lysine 9 (H3K9), H3K27, or H4K20 is generally linked to formation of gene silencing [30, 31], whereas methylation on H3K4, H3K36, and H3K79 is associated with actively transcribed regions and gene activation [29, 32, 33]. In addition, lysine residues can be mono-(me1), di-(me2), or tri-methylated (me3). These differentially methylated lysine residues lead to different levels of transcriptional activation or repression, resulting in diverse functional outcomes. For example, H4K20me1 plays an important role in transcriptional repression and X inactivation, while H4K20me2 and H4K20me3 are linked to DNA repair of double-stranded DNA damage [34].

All core histones can be dynamically acetylated on lysine residues in their tails and occasionally within the globular core. Histone acetylation removes the positive charge on the histones, thereby weakening the interaction with negatively charged DNA [25, 29]. As a consequence, chromatin is transformed into a more relaxed structure, which is associated with transcriptional activation. Like methylation, histone acetylation is reversible. This reversible acetylation in histones is an important mechanism of controlling gene expression because histone acetylation and deacetylation are linked to transcriptional activation and inactivation, respectively [25, 27].

2.2. DNA methylation

DNA methylation is one of the major repressive epigenetic pathways. Methylation occurs at the cytosine residues followed by a guanine (CpG dinucleotides) in the DNA sequence. CpG DNA methylation of gene promoters is a well-known hallmark for transcriptionally inactive genes, and is generally associated with stable gene silencing, such as genomic imprinting and X chromosome inactivation [35, 36]. The DNA methylation state is established during embryogenesis by several DNA methyltransferases (DNMTs) [35]. In mammals, 2 types of DNMTs have been identified. DNMT3A and DNMT3B establish de novo DNA methylation, while DNMT1 maintains DNA methylation patterns during DNA replication [37]. These DNA methylation sites then recruit methyl-CpG-binding proteins, including methyl-CpG-binding domain (MBD) proteins [38, 39], which bind the histone deacetylase (HDAC)-containing repressor complex, and consequently repress transcription [40, 41].

2.3. Chromatin remodeling

Chromatin structure is not static, but subject to change in response to internal and external developmental signals [23]. Dynamic changes in the chromatin structure are regulated by ATP-dependent chromatin remodelers, which allow the transcriptional machinery to access its targets more or less effectively [42-44]. Using energy derived from ATP hydrolysis, ATP-dependent chromatin remodelers relocate nucleosomes either by mobilizing or restructuring nucleosomes [45, 46]. Thus, ATP-dependent chromatin remodelers can function both in transcriptional activation and repression via their nucleosome remodeling activity. Nearly all ATP-dependent chromatin remodelers are multi-protein complexes that contain an ATPase subunit, which belongs to the sucrose non-fermenting 2 (SNF2) family of ATPases. Based on the homology between their ATPase domains, ATP-dependent chromatin remodeling complexes are divided into 4 groups: switch/sucrose non-fermenting (SWI/SNF), imitation switch (ISWI), chromo helicase DNA binding (CHD), and inositol auxotroph 80 (INO80) [42-44].

Genetic and biochemical studies indicate that some ATP-dependent chromatin remodeling complexes contain epigenetic factors such as HDAC and MBD proteins. For example, nucleosome-remodeling deacetylase (NuRD) is a multi-subunit complex that includes a SWI2/SNF2 helicase/ATPase domain-containing Mi2 protein, HDAC1, HDAC2, and MBD3 [47]. The NuRD complexes promote the establishment of a specific chromatin structure at rRNA genes that are transcriptionally inactive but are poised for transcriptional activation and control transcription of these genes [48]. Thus, ATP-dependent chromatin complexes play essential roles in epigenetic regulation of transcription along with several histone-modifying enzymes and/or modified histone codes.

2.4. Non-coding RNAs

Apart from the role of histone modifications and DNA methylation, another form of epigenetic regulation involves non-coding RNAs (ncRNAs). A large variety of ncRNAs can be classified into 2 major classes based on their transcript size: small ncRNAs (less than 200 nucleotides) and long ncRNAs (greater than 200 nucleotides) [49]. Each of these classes can be further divided into subclasses. Micro RNA (miRNA) is a subgroup of small ncRNA molecules ~22 nucleotides in length. miRNAs post-transcriptionally regulate gene expression by binding to complimentary or uncomplimentary sequences on target mRNA transcripts, which results in either mRNA degradation or inhibition of translation [50]. In animal cells, miRNA genes tend to be clustered in the genome, and are widely distributed [51]. Approximately half of the human miRNA genes are located in the introns of protein-coding genes. Expression studies have revealed that the clustered miRNA genes are often co-expressed, suggesting that they are jointly transcribed as a polycistron [52, 53].

Advertisement

3. Transition from ESCS to NPCS

3.1. Histone modifications

Expression of pluripotency genes, such as OCT4 or NANOG, is a hallmark of undifferentiated ESCs. The promoters and/or enhancers of these pluripotency genes are marked by H3K4me3, which is strongly associated with transcriptionally active genes. In contrast, the majority of genes, whose upregulation leads to differentiation, are inactivated or expressed at very low levels [54, 55]. These genes loci are maintained in a transcriptionally competent but inactive state characterized by both active (H3-K4me3) and repressive (H3-K27me3) histone marks, a configuration described as a ‘‘bivalent domain’’ (Figure 1) [54, 55]. Bivalent histone methylation in the promoters of proneural genes, such as Neurogenins (Ngns), Pax6, and Mash1, has been reported in undifferentiated ESCs [56]. H3K27 and H3K4 methylation is catalyzed by Polycomb-group (PcG) and Trithorax-group (TrxG) proteins, respectively [57]. PcG proteins form a complex referred to as the Polycomb Repressor Complex (PRC). PRCs can be biochemically subdivided into 2 groups: PRC1 and PRC2. PRC1 and PRC2 are essential for the repression of key developmental genes and maintenance of pluripotency in ESCs [58, 59]. The PRC2 complexes, which contain Enhancer of Zeste Homolog (EZH), a histone methyltransferase (HMTase), catalyze tri-methylation of H3K27 [60]. This histone mark leads to the recruitment of PRC1, thereby contributing to a repressive chromatin state [61, 62]. Consistent with this, ESCs deficient in a PRC2 component display de-repression of tissue-specific genes, including neural-associated genes [59]. TrxG proteins also act as large multimeric complexes [57]. TrxG complexes possess methyltransferase activity directed specifically towards H3K4, thereby leading to increased levels of H3K4me3 [57]. Dpy-30 is a mammalian homolog of the Drosophila TrxG protein, and a core component of myeloid/lymphoid or mixed-lineage leukemia (MLL) histone methyltransferase complexes. Depletion of Dpy-30 in ESCs results in a defect in their neural lineage specification through reduced H3K4 methylation at bivalent domains of key developmental loci [63]. Collectively, these findings suggest that the bivalent domain mediated by PcG and TrxG proteins plays a key epigenetic role in maintenance of the undifferentiated state and suppress neural differentiation.

Figure 1.

Schematic diagram of transcriptional states during neuronal fate acquisition from ESCs. In ESCs, pluripotency genes are transcriptionally activated (green "ON”). Key developmental genes are transcriptionally silent, yet competent for expression (orange “OFF”). This silencing ensures rapid reactions to extracellular inductive signals. Upon neural lineage choice, the transcriptional states of various developmental genes are altered. Pluripotency genes and genes associated with other lineages become repressed during the transition from ESCs to neurogenic NPCs (red “OFF”). Such a state of transcriptional repression is maintained over a long period. Gene activation and repression correlates with the presence of H3K4 tri-methylation (green flags) and H3K27 tri-methylation (red flags). DNA methylation (shocking pink stars) contributes to repression in combination with H3K27 tri-methylation.

Genome-wide mapping by chromatin immunoprecipitation has revealed that differentiation of ESCs is generally accompanied by global changes in histone methylation [55, 56]. In the course of neural lineage commitment, the promoters of many neural lineage genes have been shown to lose H3K27me3 from the bivalent domain, retain H3K4me3, and become activated. At the same time, the promoter loci of non-neural lineage genes maintain H3K27me3 in the bivalent domain, while removing H3K4me3, which results in a stable silent state. The promoters of some pluripotency genes such as SOX2, POU5F1, and NANOG shift from modification by H3K4me3 alone to neither H3K4 nor H3K27 methylation as they are repressed during differentiation. As ESCs differentiate into NPCs, H3K27 specific demethylase JMJD3 is recruited to and resolves the bivalent domain at the promoters of neuron-specific genes such as nestin [64]. The majority of the JMJD3 target genes are key inducers of neurogenesis, including Pax6 and Sox1. Knockdown of the H3K4me2/3 demethylase JARID1B results in upregulation of stem cell-specific genes in ESC-derived NPCs [65]. Furthermore, JARID1B knockdown ESCs fail to progress beyond the NPC stage [65]. These results suggest that JARID1B promotes ESCs to differentiate towards a neural lineage by silencing genes associated with pluripotency. Thus, when ESCs are committed to a neural lineage, bivalent domains appear to be resolved in a lineage-specific fashion by leaving methyl marks either being activated or repressed. This mechanism is believed to allow rapid transcription of developmental genes in response to a variety of extrinsic cues.

Other repressive histone marks have also been reported to play important roles during differentiation. A sustained increase in silent chromatin marked by H3K9 methylation is observed in ESC-derived cells undergoing differentiation [66], suggesting that repressive histone marks H3K9me2/3 is essential for promoting differentiation. Additionally, it is possible that H3K9me2/3 marks play an important role in the establishment of an expression profile of neuron specific genes in response to extracellular signals [67]. Ciliary neurotrophic factor (CNTF), an astrocyte differentiation factor, is incapable of inducing expression of glial fibrillary acidic protein (GFAP), an astrocytes-specific marker gene, in NPCs, because the GFAP promoter is marked by H3K9 methylation. However, fibroblast growth factor 2 (FGF2) confer NPCs with responsiveness to CNTF by adding active H3K4 methylation and removing H3K9 methylation at the GFAP promoter. Thus, H3K9 methylation controls the timing of astrogliogenesis through regulation of CNTF-mediated signaling.

In addition to histone methylation, histone acetylation also appears to be an important epigenetic modification during neural differentiation from ESCs, as ESCs generally undergo striking changes in the global pattern of histone acetylation during neural differentiation [68]. Histone acetylation is catalyzed by histone acetyltransferase (HAT), while histone deacetylation is catalyzed by histone deacetylase (HDAC) [69]. Generally, HATs induce the transcriptional activation of their target genes. However, the 60-kDa HIV-Tat interactive protein (Tip60) histone acetyltransferase has been implicated in both transcriptional repression and activation [70]. Tip60-p400 chromatin remodeling complexes containing Tip60 are necessary to maintain characteristic features of ESCs [71]. Tip60-p400 complexes acetylate histone H4 on the promoters of both activated and repressed genes. Additionally, distribution patterns of p400 in ESCs strongly correlate with H3K4me3 marks on the promoters of both active and inactive genes. These results suggest that Tip60-p400 complex-mediated histone acetylation functions as an active mark in the bivalent domain together with the active H3K4me3 mark in ESCs.

3.2. DNA methylation

Genome-wide mapping of DNA methylation patterns has revealed dynamic DNA methylation states at gene promoters during ESC differentiation (Figure 1) [72, 73]. The most pronounced changes in the DNA methylation state occur during neural lineage commitment [74], suggesting that alterations in the DNA methylation state more strongly correlate with neural lineage commitment and loss of pluripotency than with terminal neural differentiation. The promoters of highly expressed housekeeping and pluripotency genes in ESCs exhibit low methylation levels [75]. By contrast, most key developmental and tissue-specific genes exhibit high methylation levels, and are transcriptionally repressed [75]. As ESCs undergo differentiation, significant changes in DNA methylation patterns are observed. De novo methylation occurs on the promoter regions of pluripotency-associated factors [74].

DNA methylation potentially accompanies histone modifications during neural differentiation from ESCs. Cross-referencing DNA methylation patterns with mapping of histone H3K27me3 in ESCs and ESC-derived neurons has revealed that, upon differentiation, the regions marked by H3K27me3 acquire DNA methylation in a sequence-independent manner [74, 75]. H3K27me3 and DNA methylation are compatible throughout most of the genome [76]. Furthermore, deficiency of DNMT in ESCs causes widespread H3K27me3 genomic changes [76]. Taken together, these data suggest that DNA methylation, in cooperation with histone modifications, may function as a protective gear by repressing pluripotency and other lineage-specific genes during differentiation.

3.3. Chromatin remodeling

In addition to histone modifications and DNA methylation, the ATP-dependent chromatin remodeling complexes also play a pivotal role in the maintenance of ESC pluripotency. As previously described, Tip60-p400 chromatin remodeling complexes appear to be necessary for the maintenance of ESCs, including pluripotency [71]. Subunits of the NuRD complexes have also been shown to be important for ESC pluripotency and differentiation [77, 78]. Furthermore, ESCs contain another specialized ATP-dependent chromatin remodeling complex: the Brahma-associated factor (BAF) complex. The BAF complexes are characterized by 2 SWI2/SNF2-like ATPases, BRG1 and BRM. The BAF complexes in ESCs have a unique subunit composition (termed esBAF) that is not seen in other tissues, such as NPCs and post-mitotic neurons [79]. This specialized subunit composition is essential for establishment and maintenance of ESCs. The esBAF complexes contain BRG1 but not BRM, and BAF155 but not BAF170 (Figure 2). As ESCs differentiate into NPCs, these complexes undergo several subunit exchanges. The esBAF complexes incorporate BRM and excludes BAF60B, thereby forming the neural progenitor-specific BAF (npBAF) complexes in NPCs. The npBAF complex-specific subunit is necessary and sufficient for amplification of NPCs [80].

3.4. Non-coding RNAs

Increasing evidence demonstrates contributions of specific miRNAs in establishing ESC properties and their transitioning to NPCs. The miR-290-295 cluster codes for miRNAs are the most abundant in mouse ESCs and constitute over 70% of their entire miRNA population [81]. Consistent with their high expression levels, miR-290-295 miRNAs are involved in many functions in ESCs. For example, miR-290-295 miRNAs promote the transition from mitosis to S phase by targeting G1/S transition inhibitors such as Cdkn1a [82]. Furthermore, miR-290-295 miRNAs have been shown to target the Rbl2 gene, which controls the expression of the DNA methyltransferases, DNMT3A and DNMT3B, thereby establishing de novo DNA methylation in ESCs [83]. miR-9 is gradually upregulated during neural differentiation from ESCs [84, 85]. Furthermore, miR-9 performs diverse functions in different aspects of neuronal differentiation [86]. For example, during neural differentiation of human ESCs, miR-9 expression is not detectable in embryoid bodies, but is turned on in NPCs [84]. Inhibition of miR-9 activity has been found to suppress proliferation and simultaneously promote migration of NPCs [84]. Collectively, miRNAs could regulate multiple developmental processes at the post-transcriptional level.

Figure 2.

A switch in subunit composition of the BAF complexes during neuronal differentiation from ESCs. The exchange of the components is essential for the transition from ESCs to post-mitotic neurons. The exchangeable subunits are colored as follows: esBAF complex-specific subunit, ocher; npBAF complex-specific subunit, red; and nBAF complex-specific subunit, yellow. Cell type-specific BAF complexes have distinct functions that are indispensable for their properties. The BAF complex in ESCs, NPCs, and neurons are defined as esBAF, npBAF, and nBAF, respectively. The microRNA miR-124, binds to BAF53A mRNA transcripts to suppress its expression, thereby facilitating the replacement of BAF53 in the npBAF complexes.

Advertisement

4. Transition from NPCs to neurons

4.1. Histone modifications

After being committed to a neural lineage, NPCs exit the cell cycle and sequentially undergo neural and glial differentiation. Wnt signaling, which plays important roles in the maintenance of embryonic and adult stem cells, promotes differentiation of neurogenic ESCs to neurons, but does not promote differentiation of astrogenic NPCs into neurons [8]. In neurogenic NPCs, one of the target genes for Wnt signaling is the proneural gene Ngn1, which promotes neurogenesis and inhibits astrocytic differentiation [87]. During the transition from neurogenic NPCs to astrogenic NPCs, the level of repressive H3K27me3 mark at the promoter of the Ngn1 gene gradually increases, leading to gene silencing [88]. Therefore, activation of the Wnt signaling does not lead to transactivation of Ngn1 at later developmental stages. This illustrates how histone methylation can modulate responsiveness of NPCs to extracellular cues, thereby rendering NPCs to switch from neurogenesis to astrogenesis.

CREB-binding protein (CBP), one of the most extensively studied HAT proteins, has been found to play an essential role in motor neuron differentiation by interacting with retinoic acid (RA) signaling [89]. Retinoic acid receptors (RARs) are DNA-binding proteins and form a complex with Neurogenin2 (Ngn2) on the promoter of a motor neuron enhancer gene. Binding of RA to RARs triggers recruitment of CBP on the promoter, which leads to acetylation of core histone proteins and activation of the motor neuron enhancer gene. These results indicate that neuronal subtype specification is regulated, at least in part, by the interplay between intrinsic epigenetic mechanisms and extrinsic cues.

Histone deacetylation by HDACs also plays essential roles in neuronal differentiation, as evidenced by studies using histone deacetylase inhibitors, such as valproic (VP) and trichostatin A (TSA) [90-92]. VPA promotes neuronal differentiation from adult hippocampal NPCs by inducing the proneural genes, Ngn1, Math1, and NeuroD, and histone H4 acetylation, while inhibiting glial differentiation [92]. A combination of TSA with sonic hedgehog (Shh), fibroblast growth factor 8 (FGF8) and Wnt instructs non-mesencephalic NPCs to give rise to dopaminergic neurons [93]. Inhibiting the activity of all HDAC1, 2, and 3 in NPCs leads to suppression of oligodendrocyte differentiation, while HDAC2 activity alone inhibits astrocyte differentiation. On the other hand, the HDAC1 activity is required for neural differentiation [94]. HDACs are generally present within large multi-subunit protein complexes in the nucleus [69]. Among them in association with neuronal differentiation is the HDAC/CoREST/REST repressor complex. The repressor element 1-silencing transcription factor (REST) is a vertebrate zinc finger transcriptional repressor protein, which plays a fundamental role in neurogenesis [95-97]. REST is expressed in both ESCs and NSCs, but is not expressed in ES-derived neurons [98]. REST binds to an evolutionally conserved DNA motif known as the repressor element 1 (RE1) [96, 97]. REST represses the expression of RE1-containing neuronal genes via recruitment of HDAC/CoREST complexes containing HDAC1 and HDAC 2 [97]. Thus, RE1 motif-associated neuronal genes in ESCs and NPCs appear to be suppressed by REST. This repression could block premature expression of genes associated with terminal differentiation at earlier stages than needed. Recently, a genome-wide binding site analysis revealed the target genes of REST during cholinergic, GABAergic, glutamatergic, and medium spiny projection neuronal specification from NPCs [99]. A large number of the identified REST target genes are unique for each neuronal subtype, strongly suggesting that histone deacetylase plays essential roles in epigenetic control of neuronal subtype specification as well as neuronal lineage commitment.

4.2. DNA methylation

DNA methylation has been shown to suppress astrocyte-specific genes in NPCs during early stages of development. Promoters of many astrocyte-specific genes contain the signal transducer and activator of transcription (STAT) binding elements. With this, astrocyte-specific genes are transcriptionally activated through the Janus kinase (JAK)-STAT pathway, one of whose ligands is the cytokine leukemia inhibitory factor (LIF). However, neurogenic NPCs are not competent to differentiate into astrocytes even when they are grown with LIF, because the STAT-binding elements within astrocyte-specific genes promoters are methylated [100, 101]. This DNA methylation inhibits the association of activated STATs with the promoter of astrocyte-specific genes, thereby repressing their transcription. Conditional deletion of DNMT1 in embryonic NPCs results in DNA hypomethylation and alteration of the timing of astrocytogenesis [101]. In addition, knockdown of DNMT3B in ESCs alters the timing of their neural differentiation [102]. These findings suggest that DNA methylation controls the timing and developmental switch from neurogenesis to astrogliogenesis of NPCs by altering responsiveness to their extracellular developmental cues.

4.3. Chromatin remodeling

As NPCs exit the cell cycle and differentiate into mature neurons, BAF60C is incorporated into the npBAF complexes [80, 103]. BAF45A and BAF53A in the complexes give way to BAF45B/C and BAF53B, respectively (Figure 2) [80, 103], establishing the post-mitotic neuron-specific nBAF complexes. Preventing the exchange of npBAF and nBAF components impairs neuronal differentiation, indicating that a switch in subunit composition of the BAF complexes is required for the transition from pluripotent ESCs to post-mitotic neurons [80, 103]. The nBAF complexes, along with Ca2+-responsive dendritic regulator CREST, also play a role in regulating the activation of genes essential for activity-dependent dendritic outgrowth, suggesting that the nBAF complexes are required for morphological/synaptic development of neurons [103, 104].

The BAF complexes incorporate the BAF57 subunit containing DNA-binding HMG-box domains [105]. In addition, the BAF complex subunits contain motifs known to bind to modified histones, including chromo-, bromo-, and PHD domains [103]. The bromodomain can bind acetylated histones [106]. The chromo- and PHD domains function as lysine-methylated histone-binding domains [106]. The esBAF and npBAF complexes contain different chromodomain proteins (BAF155 or BAF170) [103], whereas the npBAF and nBAF complexes contain different PHD domain proteins (BAF45a or BAF45b) [80, 103]. Thus, changes in subunit composition could alter targets of the BAF complexes, thereby causing changes in gene expression patterns during neuronal differentiation.

4.4. Non-coding RNAs

A number of miRNAs involved in cell fate decision during stem cell differentiation is also highly expressed in the nervous systems. Among these is miR-9, which is expressed specifically in neurogenic areas of the embryonic and adult brains [107, 108]. TLX, an orphan nuclear receptor, is essential for maintaining a self-renewable and undifferentiated state [109], as well as cell cycle progression [110] of NPCs in the developing brain. TLX is highly expressed in NPCs, but its expression is down-regulated upon neural differentiation [111]. Conversely, miR-9 expression increases during neural differentiation [111]. Furthermore, miR-9 has been shown to negatively regulate NPC proliferation and accelerate differentiation through targeting TLX transcription [111]. These findings suggest that miR-9 can switch from undifferentiated to differentiated state of NPCs by downregulating TLX expression. Another miRNA, miR-124, has been reported to promote neuronal differentiation by targeting several genes involved in selection of non-neuronal cell fates. For example, miR-124 regulates the BAF complex subunit composition during differentiation from NPCs to neurons (Figure 2) by binding and thus suppressing BAF53A mRNA transcripts, which allows facilitation of a switch between BAF53 subunits [112]. These studies indicate that miRNA expression should be strictly controlled to ensure proper differentiation of ESCs into neurons. Consistent with this assumption, miR-124 is expressed in neurons, but not in astrocytes [113], and the miR-124 level increases during neural differentiation [84]. Remarkably, the control of miR-124 expression itself is mediated by an epigenetic mechanism. The promoter of the miR-124 gene contains a functional RE1 site. In ESCs and NPCs, REST occupies the miR-124 gene locus and represses its expression [114], allowing persistent expression of non-neuronal mRNAs. However, once NPCs start to differentiate into neurons, REST is downregulated, thereby disinhibiting miR-124 expression. miR-124 then triggers degradation of non-neuronal mRNA transcripts, which promotes differentiation towards a neuronal lineage. One of the known target genes for miR-124 is C-terminal domain phosphatase 1 (SCP1), which represses transcription of RE1-containing neuronal genes by REST, thereby preventing cells from adopting a neuronal lineage and producing non-neural tissues [115]. Together, miR-124 is a target for REST, but at the same token also targets the REST co-repressor. This represents the presence of a negative feedback loop between miRNA and a REST silencing complex, and such a mechanism may be broadly used to ensure proper cell fate transitions during development.

Neurons derived from different human ESC lines exhibit distinctive cellular properties due to the fact that human ESC lines were established under diverse conditions and from embryos with different genetic backgrounds. Comparison of neurons derived from HSF1 and HSF6 ESC human lines has revealed that the HSF1 line produces forebrain neurons with GABAergic and dopaminergic neurotransmitter phenotypes, while HSF6-derived neurons produce midbrain/hindbrain neurons bearing dopaminergic, cholinergic, serotonergic, non-forebrain GABAergic, and glutamatergic phenotypes [116]. Significant differences in the miRNA expression profile was noted between these 2 human ESC lines [116], suggesting that miRNA expression patterns might dictate in defining various neuronal subtypes arisen from ESCs.

Advertisement

5. Conclusions

Epigenetic mechanisms are regulatory processes that control gene expression via changes in chromatin structure without alterations in the DNA sequence. Changes in chromatin structure alter the accessibility of transcription factors and RNA polymerase to genes packed into chromatin, thereby modulating the efficiency of gene transcription. Epigenetic mechanisms act to control this accessibility through histone modifications, DNA methylation, chromatin remodeling, and non-coding RNAs. Each of these epigenetic events interacts with intrinsic (ex. transcription factors) and/or extrinsic factors (ex. developmental cues such as morphogens and cytokines). Studies so far have suggested that, during sequential transitions from pluripotent ESCs to terminally differentiated neurons, epigenetic mechanisms play critical roles in not only maintaining self-renewal capacity and pluripotency of ESCs, but also restricting cell lineage choices. Further investigation will therefore help clarifying the mechanisms that control pluripotency and neuronal/glial fate specification. Furthermore, the knowledge will be used in harnessing ESCs safely and effectively for clinical applications.

Advertisement

Ackknowledgments

The authors wish to thank the financial support of JSPS KAKENHI Grant Number 24592567 (to A. S.) and NIH RC1 DC010706 (to E.H.).

References

  1. 1. Evans MJ, Kaufman MH. Establishment in culture of pluripotential cells from mouse embryos. Nature 1981;292(5819) 154-156.
  2. 2. Spivakov M, Fisher AG. Epigenetic signatures of stem-cell identity. Nat Rev Genet 2007;8(4) 263-271.
  3. 3. Chen L, Daley GQ. Molecular basis of pluripotency. Hum Mol Genet 2008;17(R1) R23-27.
  4. 4. Temple S. The development of neural stem cells. Nature 2001;414(6859) 112-117.
  5. 5. Temple S. Division and differentiation of isolated CNS blast cells in microculture. Nature 1989;340(6233) 471-473.
  6. 6. Kilpatrick TJ, Bartlett PF. Cloning and growth of multipotential neural precursors: requirements for proliferation and differentiation. Neuron 1993;10(2) 255-265.
  7. 7. Hirabayashi Y, Gotoh Y. Stage-dependent fate determination of neural precursor cells in mouse forebrain. Neurosci Res 2005;51(4) 331-336.
  8. 8. Hirabayashi Y, Gotoh Y. Epigenetic control of neural precursor cell fate during development. Nat Rev Neurosci 2010;11(6) 377-388.
  9. 9. Loh YH, Wu Q, Chew JL, Vega VB, Zhang W, Chen X, Bourque G, George J, Leong B, Liu J, Wong KY, Sung KW, Lee CW, Zhao XD, Chiu KP, Lipovich L, Kuznetsov VA, Robson P, Stanton LW, Wei CL, Ruan Y, Lim B, Ng HH. The Oct4 and Nanog transcription network regulates pluripotency in mouse embryonic stem cells. Nat Genet 2006;38(4) 431-440.
  10. 10. Zhou Q, Chipperfield H, Melton DA, Wong WH. A gene regulatory network in mouse embryonic stem cells. Proc Natl Acad Sci U S A 2007;104(42) 16438-16443.
  11. 11. Walker E, Ohishi M, Davey RE, Zhang W, Cassar PA, Tanaka TS, Der SD, Morris Q, Hughes TR, Zandstra PW, Stanford WL. Prediction and testing of novel transcriptional networks regulating embryonic stem cell self-renewal and commitment. Cell Stem Cell 2007;1(1) 71-86.
  12. 12. Young RA. Control of the embryonic stem cell state. Cell 2011;144(6) 940-954.
  13. 13. Heo J, Lee JS, Chu IS, Takahama Y, Thorgeirsson SS. Spontaneous differentiation of mouse embryonic stem cells in vitro: characterization by global gene expression profiles. Biochem Biophys Res Commun 2005;332(4) 1061-1069.
  14. 14. Loring JF, Porter JG, Seilhammer J, Kaser MR, Wesselschmidt R. A gene expression profile of embryonic stem cells and embryonic stem cell-derived neurons. Restor Neurol Neurosci 2001;18(2-3) 81-88.
  15. 15. Bhattacharya B, Puri S, Puri RK. A review of gene expression profiling of human embryonic stem cell lines and their differentiated progeny. Curr Stem Cell Res Ther 2009;4(2) 98-106.
  16. 16. Stein GS, Stein JL, Van Wijnen AJ, Lian JB, Montecino M, Croce CM, Choi JY, Ali SA, Pande S, Hassan MQ, Zaidi SK, Young DW. Transcription factor-mediated epigenetic regulation of cell growth and phenotype for biological control and cancer. Adv Enzyme Regul 2010;50(1) 160-167.
  17. 17. Lunyak VV, Rosenfeld MG. Epigenetic regulation of stem cell fate. Hum Mol Genet 2008;17(R1) R28-36.
  18. 18. Luger K, Mader AW, Richmond RK, Sargent DF, Richmond TJ. Crystal structure of the nucleosome core particle at 2.8 A resolution. Nature 1997;389(6648) 251-260.
  19. 19. Yuan GC, Liu YJ, Dion MF, Slack MD, Wu LF, Altschuler SJ, Rando OJ. Genome-scale identification of nucleosome positions in S. cerevisiae. Science 2005;309(5734) 626-630.
  20. 20. Muthurajan UM, Bao Y, Forsberg LJ, Edayathumangalam RS, Dyer PN, White CL, Luger K. Crystal structures of histone Sin mutant nucleosomes reveal altered protein-DNA interactions. EMBO J 2004;23(2) 260-271.
  21. 21. Lieb JD, Clarke ND. Control of transcription through intragenic patterns of nucleosome composition. Cell 2005;123(7) 1187-1190.
  22. 22. Bell O, Tiwari VK, Thoma NH, Schubeler D. Determinants and dynamics of genome accessibility. Nat Rev Genet 2011;12(8) 554-564.
  23. 23. Bai L, Morozov AV. Gene regulation by nucleosome positioning. Trends Genet 2010;26(11) 476-483.
  24. 24. Anderson JD, Lowary PT, Widom J. Effects of histone acetylation on the equilibrium accessibility of nucleosomal DNA target sites. J Mol Biol 2001;307(4) 977-985.
  25. 25. Strahl BD, Allis CD. The language of covalent histone modifications. Nature 2000;403(6765) 41-45.
  26. 26. Lewis BP, Burge CB, Bartel DP. Conserved seed pairing, often flanked by adenosines, indicates that thousands of human genes are microRNA targets. Cell 2005;120(1) 15-20.
  27. 27. Kouzarides T. Chromatin modifications and their function. Cell 2007;128(4) 693-705.
  28. 28. Zhang Y, Reinberg D. Transcription regulation by histone methylation: interplay between different covalent modifications of the core histone tails. Genes Dev 2001;15(18) 2343-2360.
  29. 29. Peterson CL, Laniel MA. Histones and histone modifications. Curr Biol 2004;14(14) R546-551.
  30. 30. Nishioka K, Rice JC, Sarma K, Erdjument-Bromage H, Werner J, Wang Y, Chuikov S, Valenzuela P, Tempst P, Steward R, Lis JT, Allis CD, Reinberg D. PR-Set7 is a nucleosome-specific methyltransferase that modifies lysine 20 of histone H4 and is associated with silent chromatin. Mol Cell 2002;9(6) 1201-1213.
  31. 31. Schotta G, Lachner M, Sarma K, Ebert A, Sengupta R, Reuter G, Reinberg D, Jenuwein T. A silencing pathway to induce H3-K9 and H4-K20 trimethylation at constitutive heterochromatin. Genes Dev 2004;18(11) 1251-1262.
  32. 32. Kouzarides T. Histone methylation in transcriptional control. Curr Opin Genet Dev 2002;12(2) 198-209.
  33. 33. Martin C, Zhang Y. The diverse functions of histone lysine methylation. Nat Rev Mol Cell Biol 2005;6(11) 838-849.
  34. 34. Wang Y, Jia S. Degrees make all the difference: the multifunctionality of histone H4 lysine 20 methylation. Epigenetics 2009;4(5) 273-276.
  35. 35. Suzuki MM, Bird A. DNA methylation landscapes: provocative insights from epigenomics. Nat Rev Genet 2008;9(6) 465-476.
  36. 36. Eden S, Cedar H. Role of DNA methylation in the regulation of transcription. Curr Opin Genet Dev 1994;4(2) 255-259.
  37. 37. Law JA, Jacobsen SE. Establishing, maintaining and modifying DNA methylation patterns in plants and animals. Nat Rev Genet 2010;11(3) 204-220.
  38. 38. Ballestar E, Wolffe AP. Methyl-CpG-binding proteins. Targeting specific gene repression. Eur J Biochem 2001;268(1) 1-6.
  39. 39. Wade PA. Methyl CpG-binding proteins and transcriptional repression. Bioessays 2001;23(12) 1131-1137.
  40. 40. Jones PL, Veenstra GJ, Wade PA, Vermaak D, Kass SU, Landsberger N, Strouboulis J, Wolffe AP. Methylated DNA and MeCP2 recruit histone deacetylase to repress transcription. Nat Genet 1998;19(2) 187-191.
  41. 41. Nan X, Ng HH, Johnson CA, Laherty CD, Turner BM, Eisenman RN, Bird A. Transcriptional repression by the methyl-CpG-binding protein MeCP2 involves a histone deacetylase complex. Nature 1998;393(6683) 386-389.
  42. 42. Cairns BR. Chromatin remodeling: insights and intrigue from single-molecule studies. Nat Struct Mol Biol 2007;14(11) 989-996.
  43. 43. Hargreaves DC, Crabtree GR. ATP-dependent chromatin remodeling: genetics, genomics and mechanisms. Cell Res 2011;21(3) 396-420.
  44. 44. Smith CL, Peterson CL. ATP-dependent chromatin remodeling. Curr Top Dev Biol 2005;65 115-148.
  45. 45. Varga-Weisz PD, Blank TA, Becker PB. Energy-dependent chromatin accessibility and nucleosome mobility in a cell-free system. EMBO J 1995;14(10) 2209-2216.
  46. 46. Ito T, Bulger M, Pazin MJ, Kobayashi R, Kadonaga JT. ACF, an ISWI-containing and ATP-utilizing chromatin assembly and remodeling factor. Cell 1997;90(1) 145-155.
  47. 47. Zhang Y, Ng HH, Erdjument-Bromage H, Tempst P, Bird A, Reinberg D. Analysis of the NuRD subunits reveals a histone deacetylase core complex and a connection with DNA methylation. Genes Dev 1999;13(15) 1924-1935.
  48. 48. Xie W, Ling T, Zhou Y, Feng W, Zhu Q, Stunnenberg HG, Grummt I, Tao W. The chromatin remodeling complex NuRD establishes the poised state of rRNA genes characterized by bivalent histone modifications and altered nucleosome positions. Proc Natl Acad Sci U S A 2012;109(21) 8161-8166.
  49. 49. Brosnan CA, Voinnet O. The long and the short of noncoding RNAs. Curr Opin Cell Biol 2009;21(3) 416-425.
  50. 50. Cannell IG, Kong YW, Bushell M. How do microRNAs regulate gene expression? Biochem Soc Trans 2008;36(Pt 6) 1224-1231.
  51. 51. Lagos-Quintana M, Rauhut R, Lendeckel W, Tuschl T. Identification of novel genes coding for small expressed RNAs. Science 2001;294(5543) 853-858.
  52. 52. Lee Y, Jeon K, Lee JT, Kim S, Kim VN. MicroRNA maturation: stepwise processing and subcellular localization. EMBO J 2002;21(17) 4663-4670.
  53. 53. He L, Thomson JM, Hemann MT, Hernando-Monge E, Mu D, Goodson S, Powers S, Cordon-Cardo C, Lowe SW, Hannon GJ, Hammond SM. A microRNA polycistron as a potential human oncogene. Nature 2005;435(7043) 828-833.
  54. 54. Azuara V, Perry P, Sauer S, Spivakov M, Jorgensen HF, John RM, Gouti M, Casanova M, Warnes G, Merkenschlager M, Fisher AG. Chromatin signatures of pluripotent cell lines. Nat Cell Biol 2006;8(5) 532-538.
  55. 55. Bernstein BE, Mikkelsen TS, Xie X, Kamal M, Huebert DJ, Cuff J, Fry B, Meissner A, Wernig M, Plath K, Jaenisch R, Wagschal A, Feil R, Schreiber SL, Lander ES. A bivalent chromatin structure marks key developmental genes in embryonic stem cells. Cell 2006;125(2) 315-326.
  56. 56. Mikkelsen TS, Ku M, Jaffe DB, Issac B, Lieberman E, Giannoukos G, Alvarez P, Brockman W, Kim TK, Koche RP, Lee W, Mendenhall E, O'Donovan A, Presser A, Russ C, Xie X, Meissner A, Wernig M, Jaenisch R, Nusbaum C, Lander ES, Bernstein BE. Genome-wide maps of chromatin state in pluripotent and lineage-committed cells. Nature 2007;448(7153) 553-560.
  57. 57. Ringrose L, Paro R. Epigenetic regulation of cellular memory by the Polycomb and Trithorax group proteins. Annu Rev Genet 2004;38 413-443.
  58. 58. Lee TI, Jenner RG, Boyer LA, Guenther MG, Levine SS, Kumar RM, Chevalier B, Johnstone SE, Cole MF, Isono K, Koseki H, Fuchikami T, Abe K, Murray HL, Zucker JP, Yuan B, Bell GW, Herbolsheimer E, Hannett NM, Sun K, Odom DT, Otte AP, Volkert TL, Bartel DP, Melton DA, Gifford DK, Jaenisch R, Young RA. Control of developmental regulators by Polycomb in human embryonic stem cells. Cell 2006;125(2) 301-313.
  59. 59. Boyer LA, Plath K, Zeitlinger J, Brambrink T, Medeiros LA, Lee TI, Levine SS, Wernig M, Tajonar A, Ray MK, Bell GW, Otte AP, Vidal M, Gifford DK, Young RA, Jaenisch R. Polycomb complexes repress developmental regulators in murine embryonic stem cells. Nature 2006;441(7091) 349-353.
  60. 60. Cao R, Zhang Y. The functions of E(Z)/EZH2-mediated methylation of lysine 27 in histone H3. Curr Opin Genet Dev 2004;14(2) 155-164.
  61. 61. Cao R, Wang L, Wang H, Xia L, Erdjument-Bromage H, Tempst P, Jones RS, Zhang Y. Role of histone H3 lysine 27 methylation in Polycomb-group silencing. Science 2002;298(5595) 1039-1043.
  62. 62. Schwartz YB, Pirrotta V. Polycomb silencing mechanisms and the management of genomic programmes. Nat Rev Genet 2007;8(1) 9-22.
  63. 63. Jiang H, Shukla A, Wang X, Chen WY, Bernstein BE, Roeder RG. Role for Dpy-30 in ES cell-fate specification by regulation of H3K4 methylation within bivalent domains. Cell 2011;144(4) 513-525.
  64. 64. Burgold T, Spreafico F, De Santa F, Totaro MG, Prosperini E, Natoli G, Testa G. The histone H3 lysine 27-specific demethylase Jmjd3 is required for neural commitment. PLoS One 2008;3(8) e3034.
  65. 65. Schmitz SU, Albert M, Malatesta M, Morey L, Johansen JV, Bak M, Tommerup N, Abarrategui I, Helin K. Jarid1b targets genes regulating development and is involved in neural differentiation. EMBO J 2011;30(22) 4586-4600.
  66. 66. Meshorer E, Yellajoshula D, George E, Scambler PJ, Brown DT, Misteli T. Hyperdynamic plasticity of chromatin proteins in pluripotent embryonic stem cells. Dev Cell 2006;10(1) 105-116.
  67. 67. Song MR, Ghosh A. FGF2-induced chromatin remodeling regulates CNTF-mediated gene expression and astrocyte differentiation. Nat Neurosci 2004;7(3) 229-235.
  68. 68. Meshorer E, Misteli T. Chromatin in pluripotent embryonic stem cells and differentiation. Nat Rev Mol Cell Biol 2006;7(7) 540-546.
  69. 69. Hayakawa T, Nakayama J. Physiological roles of class I HDAC complex and histone demethylase. J Biomed Biotechnol 2011;2011 129383.
  70. 70. Sapountzi V, Logan IR, Robson CN. Cellular functions of TIP60. Int J Biochem Cell Biol 2006;38(9) 1496-1509.
  71. 71. Fazzio TG, Huff JT, Panning B. An RNAi screen of chromatin proteins identifies Tip60-p400 as a regulator of embryonic stem cell identity. Cell 2008;134(1) 162-174.
  72. 72. Meissner A, Mikkelsen TS, Gu H, Wernig M, Hanna J, Sivachenko A, Zhang X, Bernstein BE, Nusbaum C, Jaffe DB, Gnirke A, Jaenisch R, Lander ES. Genome-scale DNA methylation maps of pluripotent and differentiated cells. Nature 2008;454(7205) 766-770.
  73. 73. Isagawa T, Nagae G, Shiraki N, Fujita T, Sato N, Ishikawa S, Kume S, Aburatani H. DNA methylation profiling of embryonic stem cell differentiation into the three germ layers. PLoS One;6(10) e26052.
  74. 74. Mohn F, Weber M, Rebhan M, Roloff TC, Richter J, Stadler MB, Bibel M, Schubeler D. Lineage-specific polycomb targets and de novo DNA methylation define restriction and potential of neuronal progenitors. Mol Cell 2008;30(6) 755-766.
  75. 75. Fouse SD, Shen Y, Pellegrini M, Cole S, Meissner A, Van Neste L, Jaenisch R, Fan G. Promoter CpG methylation contributes to ES cell gene regulation in parallel with Oct4/Nanog, PcG complex, and histone H3 K4/K27 trimethylation. Cell Stem Cell 2008;2(2) 160-169.
  76. 76. Brinkman AB, Gu H, Bartels SJ, Zhang Y, Matarese F, Simmer F, Marks H, Bock C, Gnirke A, Meissner A, Stunnenberg HG. Sequential ChIP-bisulfite sequencing enables direct genome-scale investigation of chromatin and DNA methylation cross-talk. Genome Res 2012;22(6) 1128-1138.
  77. 77. Kaji K, Caballero IM, MacLeod R, Nichols J, Wilson VA, Hendrich B. The NuRD component Mbd3 is required for pluripotency of embryonic stem cells. Nat Cell Biol 2006;8(3) 285-292.
  78. 78. Kaji K, Nichols J, Hendrich B. Mbd3, a component of the NuRD co-repressor complex, is required for development of pluripotent cells. Development 2007;134(6) 1123-1132.
  79. 79. Ho L, Ronan JL, Wu J, Staahl BT, Chen L, Kuo A, Lessard J, Nesvizhskii AI, Ranish J, Crabtree GR. An embryonic stem cell chromatin remodeling complex, esBAF, is essential for embryonic stem cell self-renewal and pluripotency. Proc Natl Acad Sci U S A 2009;106(13) 5181-5186.
  80. 80. Lessard J, Wu JI, Ranish JA, Wan M, Winslow MM, Staahl BT, Wu H, Aebersold R, Graef IA, Crabtree GR. An essential switch in subunit composition of a chromatin remodeling complex during neural development. Neuron 2007;55(2) 201-215.
  81. 81. Marson A, Levine SS, Cole MF, Frampton GM, Brambrink T, Johnstone S, Guenther MG, Johnston WK, Wernig M, Newman J, Calabrese JM, Dennis LM, Volkert TL, Gupta S, Love J, Hannett N, Sharp PA, Bartel DP, Jaenisch R, Young RA. Connecting microRNA genes to the core transcriptional regulatory circuitry of embryonic stem cells. Cell 2008;134(3) 521-533.
  82. 82. Wang Y, Baskerville S, Shenoy A, Babiarz JE, Baehner L, Blelloch R. Embryonic stem cell-specific microRNAs regulate the G1-S transition and promote rapid proliferation. Nat Genet 2008;40(12) 1478-1483.
  83. 83. Sinkkonen L, Hugenschmidt T, Berninger P, Gaidatzis D, Mohn F, Artus-Revel CG, Zavolan M, Svoboda P, Filipowicz W. MicroRNAs control de novo DNA methylation through regulation of transcriptional repressors in mouse embryonic stem cells. Nat Struct Mol Biol 2008;15(3) 259-267.
  84. 84. Delaloy C, Liu L, Lee JA, Su H, Shen F, Yang GY, Young WL, Ivey KN, Gao FB. MicroRNA-9 coordinates proliferation and migration of human embryonic stem cell-derived neural progenitors. Cell Stem Cell 2010;6(4) 323-335.
  85. 85. Krichevsky AM, Sonntag KC, Isacson O, Kosik KS. Specific microRNAs modulate embryonic stem cell-derived neurogenesis. Stem Cells 2006;24(4) 857-864.
  86. 86. Gao FB. Context-dependent functions of specific microRNAs in neuronal development. Neural Dev 2010;5 25.
  87. 87. Hirabayashi Y, Itoh Y, Tabata H, Nakajima K, Akiyama T, Masuyama N, Gotoh Y. The Wnt/beta-catenin pathway directs neuronal differentiation of cortical neural precursor cells. Development 2004;131(12) 2791-2801.
  88. 88. Hirabayashi Y, Suzki N, Tsuboi M, Endo TA, Toyoda T, Shinga J, Koseki H, Vidal M, Gotoh Y. Polycomb limits the neurogenic competence of neural precursor cells to promote astrogenic fate transition. Neuron 2009;63(5) 600-613.
  89. 89. Lee S, Lee B, Lee JW, Lee SK. Retinoid signaling and neurogenin2 function are coupled for the specification of spinal motor neurons through a chromatin modifier CBP. Neuron 2009;62(5) 641-654.
  90. 90. Marin-Husstege M, Muggironi M, Liu A, Casaccia-Bonnefil P. Histone deacetylase activity is necessary for oligodendrocyte lineage progression. J Neurosci 2002;22(23) 10333-10345.
  91. 91. Balasubramaniyan V, Boddeke E, Bakels R, Kust B, Kooistra S, Veneman A, Copray S. Effects of histone deacetylation inhibition on neuronal differentiation of embryonic mouse neural stem cells. Neuroscience 2006;143(4) 939-951.
  92. 92. Hsieh J, Nakashima K, Kuwabara T, Mejia E, Gage FH. Histone deacetylase inhibition-mediated neuronal differentiation of multipotent adult neural progenitor cells. Proc Natl Acad Sci U S A 2004;101(47) 16659-16664.
  93. 93. Rossler R, Boddeke E, Copray S. Differentiation of non-mesencephalic neural stem cells towards dopaminergic neurons. Neuroscience 2010;170(2) 417-428.
  94. 94. Humphrey GW, Wang YH, Hirai T, Padmanabhan R, Panchision DM, Newell LF, McKay RD, Howard BH. Complementary roles for histone deacetylases 1, 2, and 3 in differentiation of pluripotent stem cells. Differentiation 2008;76(4) 348-356.
  95. 95. Chong JA, Tapia-Ramirez J, Kim S, Toledo-Aral JJ, Zheng Y, Boutros MC, Altshuller YM, Frohman MA, Kraner SD, Mandel G. REST: a mammalian silencer protein that restricts sodium channel gene expression to neurons. Cell 1995;80(6) 949-957.
  96. 96. choenherr CJ, Paquette AJ, Anderson DJ. Identification of potential target genes for the neuron-restrictive silencer factor. Proc Natl Acad Sci U S A 1996;93(18) 9881-9886.
  97. 97. Ballas N, Mandel G. The many faces of REST oversee epigenetic programming of neuronal genes. Curr Opin Neurobiol 2005;15(5) 500-506.
  98. 98. Sun YM, Greenway DJ, Johnson R, Street M, Belyaev ND, Deuchars J, Bee T, Wilde S, Buckley NJ. Distinct profiles of REST interactions with its target genes at different stages of neuronal development. Mol Biol Cell 2005;16(12) 5630-5638.
  99. 99. Abrajano JJ, Qureshi IA, Gokhan S, Zheng D, Bergman A, Mehler MF. REST and CoREST modulate neuronal subtype specification, maturation and maintenance. PLoS One 2009;4(12) e7936. http://www.plosone.org/article/info%3Adoi%2F10.1371%2Fjournal.pone.0007936 (accessed 7 December 2009)
  100. 100. Takizawa T, Nakashima K, Namihira M, Ochiai W, Uemura A, Yanagisawa M, Fujita N, Nakao M, Taga T. DNA methylation is a critical cell-intrinsic determinant of astrocyte differentiation in the fetal brain. Dev Cell 2001;1(6) 749-758.
  101. 101. Fan G, Martinowich K, Chin MH, He F, Fouse SD, Hutnick L, Hattori D, Ge W, Shen Y, Wu H, ten Hoeve J, Shuai K, Sun YE. DNA methylation controls the timing of astrogliogenesis through regulation of JAK-STAT signaling. Development 2005;132(15) 3345-3356.
  102. 102. Martins-Taylor K, Schroeder DI, Lasalle JM, Lalande M, Xu RH. Role of DNMT3B in the regulation of early neural and neural crest specifiers. Epigenetics 2012;7(1) 71-82.
  103. 103. Yoo AS, Crabtree GR. ATP-dependent chromatin remodeling in neural development. Curr Opin Neurobiol 2009;19(2) 120-126.
  104. 104. Wu JI, Lessard J, Olave IA, Qiu Z, Ghosh A, Graef IA, Crabtree GR. Regulation of dendritic development by neuron-specific chromatin remodeling complexes. Neuron 2007;56(1) 94-108.
  105. 105. Wang W, Chi T, Xue Y, Zhou S, Kuo A, Crabtree GR. Architectural DNA binding by a high-mobility-group/kinesin-like subunit in mammalian SWI/SNF-related complexes. Proc Natl Acad Sci U S A 1998;95(2) 492-498.
  106. 106. Yun M, Wu J, Workman JL, Li B. Readers of histone modifications. Cell Res 2011;21(4) 564-578.
  107. 107. Deo M, Yu JY, Chung KH, Tippens M, Turner DL. Detection of mammalian microRNA expression by in situ hybridization with RNA oligonucleotides. Dev Dyn 2006;235(9) 2538-2548.
  108. 108. Kapsimali M, Kloosterman WP, de Bruijn E, Rosa F, Plasterk RH, Wilson SW. MicroRNAs show a wide diversity of expression profiles in the developing and mature central nervous system. Genome Biol 2007;8(8) R173.
  109. 109. Shi Y, Chichung Lie D, Taupin P, Nakashima K, Ray J, Yu RT, Gage FH, Evans RM. Expression and function of orphan nuclear receptor TLX in adult neural stem cells. Nature 2004;427(6969) 78-83.
  110. 110. Li W, Sun G, Yang S, Qu Q, Nakashima K, Shi Y. Nuclear receptor TLX regulates cell cycle progression in neural stem cells of the developing brain. Mol Endocrinol 2008;22(1) 56-64.
  111. 111. Zhao C, Sun G, Li S, Shi Y. A feedback regulatory loop involving microRNA-9 and nuclear receptor TLX in neural stem cell fate determination. Nat Struct Mol Biol 2009;16(4) 365-371.
  112. 112. Yoo AS, Staahl BT, Chen L, Crabtree GR. MicroRNA-mediated switching of chromatin-remodelling complexes in neural development. Nature 2009;460(7255) 642-646.
  113. 113. Smirnova L, Grafe A, Seiler A, Schumacher S, Nitsch R, Wulczyn FG. Regulation of miRNA expression during neural cell specification. Eur J Neurosci 2005;21(6) 1469-1477.
  114. 114. Conaco C, Otto S, Han JJ, Mandel G. Reciprocal actions of REST and a microRNA promote neuronal identity. Proc Natl Acad Sci U S A 2006;103(7) 2422-2427.
  115. 115. Yeo M, Lee SK, Lee B, Ruiz EC, Pfaff SL, Gill GN. Small CTD phosphatases function in silencing neuronal gene expression. Science 2005;307(5709) 596-600.
  116. 116. Wu H, Xu J, Pang ZP, Ge W, Kim KJ, Blanchi B, Chen C, Sudhof TC, Sun YE. Integrative genomic and functional analyses reveal neuronal subtype differentiation bias in human embryonic stem cell lines. Proc Natl Acad Sci U S A 2007;104(34) 13821-13826.

Written By

Atsushi Shimomura and Eri Hashino

Submitted: 26 March 2012 Published: 27 March 2013