REST/NRSF in Health and Disease: History
Please note this is an old version of this entry, which may differ significantly from the current revision.
Contributor: , , , ,

Chromatin modifications play a crucial role in the regulation of gene expression. The repressor element-1 (RE1) silencing transcription factor (REST), also known as neuron-restrictive silencer factor (NRSF) and X2 box repressor (XBR), was found to regulate gene transcription by binding to chromatin and recruiting chromatin-modifying enzymes. Earlier studies revealed that REST plays an important role in the development and disease of the nervous system, mainly by repressing the transcription of neuron-specific genes. Subsequently, REST was found to be critical in other tissues, such as the heart, pancreas, skin, eye, and vascular. Dysregulation of REST was also found in nervous and non-nervous system cancers. In parallel, multiple strategies to target REST have been developed. 

  • chromatin modification
  • REST/NRSF
  • transcriptional regulation
  • cancer

1. Introduction

The repressor element-1 (RE1) silencing transcription factor (REST), also known as neuron-restrictive silencer factor (NRSF) and X2 box repressor (XBR), is a transcriptional repressor first reported in 1995. REST binds to the RE1 motif, also named neural restrictive silencing element (NRSE), and recruits corepressor complexes [1][2]. Initially identified as a regulator of neuron-specific genes in nonneuronal tissues, REST has been found to play crucial roles in neuronal development, function, and related diseases. Over nearly three decades of research, REST has been revealed to have diverse functions in other tissues such as the heart, pancreas, skin, eye, and vascular, as well as in various cancers.

1.1. Gene and Protein Structure

The full-length human REST contains a DNA binding domain (DBD), an N-terminal repressor domain (NRD), and a C-terminal repressor domain (CRD) (Figure 1). REST is highly conserved in mammals, and there are various mRNA alternative splice variants predictive of multiple protein isoforms. Studies showed that REST might produce at least 45 mRNA variants encoding REST4, REST1, RESTC, and small peptides [3].
Figure 1. Gene structure and protein structure of REST. (A) Human REST is located in Chromosome 4q12, and it has five exons. The exon 5 was found to be mutually exclusive of exon 4. (B) The full-length human REST consists of 1097 aa, and there is a DNA binding domain (DBD), an N-terminal repressor domain (NRD), and a C-terminal repressor domain (CRD) for its regulatory activities. ZF indicates zinc finger. (C) The protein structure was predicated using Alphafold2.

1.2. Nuclear Location

The nuclear location of transcription factors is essential for its modulatory function. Classically, REST binds to DNA and recruits coregulators to modulate gene expression upon entering the nucleus. The deletion of residues 512–522 in the full-length REST resulted in cytosol location. However, the isoform REST4, which lacks this putative nuclear localization signal (NLS), was localized to the nucleus, while another isoform REST1, lacking zinc finger 5, was localized in the cytosol [4]. Via deletion and mutation, the NLS around zinc finger 5 was verified to be the sole NLS within REST [5].

1.3. DNA Binding

The DBD containing eight noncanonical Cys2-His2 zinc fingers (zinc finger 1–8 from N-terminus to C-terminus in Figure 1B) is crucial for DNA binding [2]. The deletion of zinc finger 3 or 5 abolished the DNA binding activity of REST, and the segment of amino acids 196–207 is essential for DNA binding [6]. Generally, each Cys2-His2 zinc finger domain binds three nucleotides within DNA, and a point mutation that changes a Cys residue to an Arg residue disrupts the Cys2-His2 zinc finger. The single mutation in zinc finger 7 and combinations of mutations in any two of zinc finger 6–8 decreases DNA binding, while a single mutation in zinc finger 6 or 8 has little effect [4]. Further, the residue Cys397 plays an important role in the global folding of multiple zinc fingers domain [7].

1.4. Coregulator Recruitment

The transcriptional regulation mediated by REST requires the assembly of multimeric complexes [8]. Sin3 is constitutively required for REST repression, while CoREST is recruited for more specialized repressor functions [9] (Figure 2).
Figure 2. REST modulates gene activation via histone modification. REST binds to the RE1 motif within target genes and recruits Sin3A/B and CoREST as beacons for the recruitment of coregulators to form regulatory complexes. The chromatin was silenced via histone 3/4 deacetylation by HDACs, H3K9 methylation by G9a and SUV39H1, and H3K4 demethylation by LSD1 and H4K8 acetylation by BRG1. Further, the H3K36 methylation mediated by Tet1/3 might activate gene transcription.
The CRD containing a zinc finger (zinc finger 9) recruits CoREST as a beacon for the recruitment of repressors [10][11], including HDAC1/2, LSD1, BRG1, CtBP, G9a, SUV39H1, HP1 and MeCP2 [12]. HDACs remove acetyl groups on H3 and H4, and H3K9 deacetylation triggers localized domains of histone methyltransferase G9a-mediated H3K9 methylation [13]. The K9 residues methylated by G9a and another histone, H3K9 methyltransferase SUV39H1, are binding sites for heterochromatin protein 1 (HP1), which causes chromatin condensation [11].

2. REST in Nervous System

2.1. Embryogenesis and Neurogenesis

During mouse embryogenesis, REST was found to be required in repressing neuronal gene expression in both non-neural and neural precursor tissues [14]. The postnatal development of the hippocampus, especially the formation of the subgranular zone (SGZ), was influenced by conditional knockout (cKO) of REST in mice [15]. During early development, the loss of REST induces global hypermethylation [16][17], and the deregulation of REST was found to be an early event in Down syndrome (DS) [18][19]. In neural progenitor cells, REST was under the regulation of Wnt signaling, which controls various aspects of early development [20][21].

2.2. Neuronal Differentiation

REST is a critical regulator of differentiation in embryonic stem cells (ESCs), neural stem cells (NSCs), and mature cell types [22][23]. In ESCs, a single REST allele disruption reduced alkaline phosphatase activity and several pluripotency-associated genes expression [24]. However, partial or complete loss of REST could not abrogate differentiation potential as reflected by marker gene expression [25]. REST knockdown in mouse ESCs could induce neuronal lineage differentiation [25], and its knockdown in human ESCs increased cell survival and the expression of mesendoderm differentiation markers [26].
The repression of neuron-specific genes by REST in nonneuronal tissues highlights its crucial function during neuronal differentiation. Using ChIP-seq data mining, REST was found to mediate more than 2000 gene silencing in ESCs but not in ESC-derived neurons [27]. REST is critical for maintaining the NSCs pool, and the activation of REST target genes by REST-VP16 transgene was sufficient to induce neuronal differentiation in NSCs [28][29]. The nuclear REST was reduced during neuronal differentiation of embryonic cells, and its target genes N-methyl-D-aspartate receptor subunit 1 (NR1) and tyrosine hydroxylase (TH) were upregulated [30][31]

2.3. Neuronal Survival

REST expression is typically low in differentiated neurons, whereas its elevation frequently exhibits protective effects. For instance, the mice with neural crest-specific KO of REST presented aberrations of the myenteric plexus, leading to neonatal lethality [32]. In differentiated neural cells, overexpression of REST promoted proliferation and disrupted neurosecretion via the REST-TSC2-β-catenin signaling pathway [33]. BMP/RA-inducible neural-specific protein 1 (BRINP1) was a negative regulator of cell cycle progression, and its promotor activity was inhibited by the RE1 motif within its promotor [34]. REST inhibited BAF53a to quit mitosis by targeting miR-9* and miR-124 in post-mitotic neurons [35].
The expression of REST in SH-SY5Y cells was induced by polychlorinated biphenyls (PCBs), and ERK2/Sp1/Sp3/REST signaling mediated the neuronal toxicity of PCBs [36]. REST was increased during monosodium-glutamate-induced excitotoxicity in rats [37]. REST could protect cath.-a-differentiated (CAD) neuronal cells from Mn-induced toxicity by enhancing the expression of the dopamine-synthesizing enzyme tyrosine hydroxylase (TH) [38].

2.4. Neuronal Transmission and Synaptic Plasticity

REST maintains neuronal function balance by dynamically regulating the synaptic efficiency. To facilitate neurite outgrowth, REST restricted the cell adhesion molecule L1 in the embryonic nervous system [39] and positively regulated the exocytosis of enlargeosomes [40]. However, the constitutive expression of REST led to axon guidance mistakes in the spinal cord of chicken embryos, with the depression of the target genes N-tubulin and Ng-CAM [41]. Syn1 was negatively regulated by REST to modulate synaptogenesis and neurotransmitter release [42], and Sec6 has a REST binding site and function on synapse formation and synaptic plasticity [43].
REST protects neurons from hyperexcitability via modulating the transcriptional activity of voltage-gated ion channel genes. REST elevated by hyperactivity inhibited Nav1.2 expression to maintain homeostasis, and silencing of REST disrupted this homeostatic response [44].

2.5. Pain

REST regulates μ-opioid receptor (MOR) transcriptional activity [45][46][47], contributing to MOR agonist remifentanil inducing postoperative hyperalgesia [48]. PACAP was regulated by REST to mediate neuropathic pain after spinal nerve injury [49]. The negative regulation of NR2B by REST was involved in bone cancer pain [50]. The repression of MOR by REST reduced morphine analgesia in sarcoma-induced bone cancer pain [51]. REST mediated diabetic neuropathic pain in db/db mice [52]. REST also plays a major role in the acute-to-chronic pain transition upon nerve injury [53]. REST KO reverted injury-induced hyperalgesia, and REST regulates peripheral somatosensory neuron programming, leading to chronic pain [54].

2.6. Neuroendocrine

The cocaine- and amphetamine-regulated transcript (CART) peptide is under strict control of REST in neuroendocrine cells [55]. The neurosecretory process involving dense-core vesicles (DCVs) was governed by REST in PC12 and astrocyte cells [56][57]. The depression of REST by Kaposi’s sarcoma-associated herpesvirus (KSHV) latent open reading frame K12 (ORFK12) gene (kaposin A) regulated neuroendocrine gene expression in infected endothelial cells [58].

2.7. Intelligence and Memory

Compared with mice, human-specific genes with REST occupancy were enriched in learning and memory functions [59][60]. In early life adversity (ELA) rats, REST contributes to the memory deficits and its blockage rescued spatial memory [61]. Additionally, in APP/PS1 mice, REST is essential for the maintenance of memory performance during aging [62]. The elevated REST by Pb explore inhibited SV2C affected synaptic plasticity, leading to impairment of learning and memory [63]. During the memory impairment induced by a flame retardant PBDE-209, REST was decreased and the inhibition of NR1 was released. CDRI-08, the ethanolic extract of a nootropic plant Bacopa monnieri, could attenuate the impairment and restore the binding of REST to NR1 promotor [64].

2.8. Aging and Alzheimer’s Disease

The REST-RE1 machinery is well maintained during aging in rats [65]. By regulating autophagy and senescence in neurons, REST plays a protective role during physiological aging [66]. The homologous gene of REST in C. elegans, Spr3, is crucial for lifespan regulation, as its mutation and silencing led to shortened lifespan [67]. REST was elevated during neuronal senescence induced by high glucose and palmitic acid [68]. Upon peripheral nerve injury, REST was elevated in young mice but not in aged mice [69].

2.9. Parkinson’s Disease

Similar to AD, REST enters the nucleus of aged dopaminergic neurons, and it is mostly absent in neurons of Parkinson’s disease (PD) [70]. REST and REST4 were increased in the neurotoxin 1-methyl-4-phenyl-pyridinium ion (MPP+) induced PD cell model [71]. REST KO mice presented higher sensitivity to the dopaminergic neurotoxin MPTP [72], while the neuroprotective effects of Trichostatin A (TSA) were mediated by REST in MPTP PD models [73]. REST deficiency aggravated dopaminergic neurodegeneration and impaired neurogenesis in mice with MPTP-induced PD [74][75].

2.10. Huntington’s Disease

In Huntington’s disease (HD), the increased binding of REST to RE1 reduced its target gene expression. The binding was regulated by huntingtin, and the control was lost in HD, leading to the loss of neuronal gene expression [76][77][78]. The blockage of REST binding rescued BDNF [79]. Huntingtin and REST formed a complex to facilitate REST translocation to the nucleus [80]. The transcriptional regulation of REST by huntingtin interacting protein 1 protein interactor (HIPPI) contributed to the deregulation of transcription in HD [81]. The exon-3 skipping triggered by antisense oligonucleotides (ASOs) treatment rescued neuronal genes [82].

2.11. Epilepsy

The role of REST in epilepsy is controversial, and in the literature, opposite views coexist. The elevated REST during aging was abolished in patients with drug-resistant mesial temporal lobe epilepsy (MTLE) [83]. REST cKO exhibited dramatically accelerated seizure progression and prolonged afterdischarge duration [84]. However, several investigations suggested that REST is proepileptogenic. REST cKO mice showed higher resistance to convulsions induced by PTZ [85]. The interference with the chromatin binding of REST rescued spatial memory impaired by febrile status epilepticus [86]. In kainic acid-induced status epilepticus, the repression of the HCN1 channel was mediated by REST binding [87], and mild inactivation of REST reduces susceptibility [88]. The antiepileptic effect of 2-deoxy-d-glucose (2DG), which inhibits BDNF and its receptor TrkB to reduce the progression of kindling, was mediated by REST [89], and this effect was abolished in REST cKO mice [90]. The mutation of RILP could block its interaction with REST, contributing to the progressive myoclonus epilepsy progression [91]. REST and REST4 controlled the expression of preprotachykinin-A (TAC1), which encodes the neuropeptide substance P [92][93][94]. REST mediated the contribution of Sit1 to epileptogenesis [95]. Sirt1 reduced miR-124, which targeted REST mRNA [96].

2.12. Ischemia

In ischemic damage to neurons, REST appears more like a mediator. Global ischemia induced REST expression to repress GluR2 and MOR-1, and REST knockdown protected neurons from ischemia-induced death [97][98][99][100]. Moreover, the elevated REST antagonized the CREB signaling on CART activation, leading to increased cell death [101]. Pyrvinium pamoate, an activator of casein kinase 1, can suppress REST expression and rescue neurons that would otherwise undergo cell death [102].

2.13. Psychiatric Disorders

REST contributes to genome-wide epigenetic changes and is implicated in schizophrenia [103]. The aberrant increase of REST at the Grin2b promoter caused deficient synaptic physiology and PFC-dependent cognitive dysfunction, a hallmark of schizophrenia [104]. THP2 is the rate-limiting enzyme for 5-HT, and it is under the regulation of REST [105]. The mood-stabilizing drugs lithium and sodium valproate have been shown to regulate REST expression [106][107]. miR-9 targeted REST and PDYN, and miR-9 was targeted by REST [108]. REST was elevated in care-augmented rats, and its target CRH was inhibited [109]. The compound C737 mimics the helical structure of REST to regulate target genes involved in neuronal function and pain modulation and reduce stress-induced weight loss in female Tupaia [110]. REST signaling contributes to stress resilience [111]. The improvement in diabetes mellitus-related cognitive impairment by 9-PAHSA was mediated by REST [112].

3. REST in Other Systems

3.1. Heart

REST regulates heart development, structure, and function [113]. During embryonic heart development, the elevated REST inhibited an embryonic cardiac gene HCN4 [114][115][116][117]. The training-induced REST contributes to the pathological heart rate adaptation to exercise training by reducing HCN4 [118]. REST chronically represses T channels in cardiomyocytes [119]. REST also maintains low alpha 1H expression [120] and regulates CACNA1H encoding alpha-subunit of Cav3.2 [121], thereby regulating aldosterone and cortisol production, which are independent predictors of mortality risk for heart failure patients. Gαo, an inhibitory G protein encoded by GNAO1, was regulated by REST. In several mouse models of heart failure, Gαo is decreased in the ventricles, leading to increased surface sarcolemmal L-type Ca2+ channel activity and impairing Ca2+ handling in ventricular myocytes, resulting in cardiac dysfunction [122].

3.2. Pancreas

REST is absent in insulinoma, insulin- and glucagon-producing cells [123]. Similar to neuronal differentiation, REST is expressed in pancreas progenitors and decreased in differentiated endocrine cells [124]. RE1 motif was found within various genes involved in pancreas development, including pdx-1, Beta2/NeuroD, and pax4 [125]. Cx36 is also controlled by REST in insulin-producing cells [126]. In addition, the rs2518719 SNP could alter the RE1 motif involving pancreatic neuroendocrine tumor progression [127]. It was revealed that the REST silencing could induce human amniotic fluid-derived stem cells (hAFSCs) and bone marrow-derived mesenchymal stem cells (bmMSCs) to differentiate into insulin-producing cells [128][129].

3.3. Skin

REST is expressed in skin keratinocytes, and NRP1 was found to be under its regulation to promote cell migration by blocking the inhibition of semaphorin 3A [130]. During the early neural crest stage, REST is essential for melanoblast development [131]. In melanoma, the decreased REST/RE1 activity leads to aberrant metabotropic glutamate receptor 1 (mGluR1) expression, whose ectopic expression in mouse melanocytes was sufficient to induce melanoma development [132]. These findings provide valuable insights into the functions of REST in the skin, although much more research is needed to fully understand its role in skin biology and pathology.

3.4. Eye

The eye, being a specialized structure derived from the skin, also relies on the proper regulation of REST. In REST cKO mice generated by the Sox1-Cre allele and the floxed REST gene, REST was excised in the early neural tissues, including the lens and retinal primordia. The slightly reduced proliferation of lens epithelial cells after birth was observed, and vacuoles formed without inducing cell apoptosis. The augmented Notch signaling and the reduced lens fiber regulator genes may contribute [133]. The REST-like gene in Drosophila was named Chn, and it promotes D1 for photoreceptor cell development [134].

3.5. Vascular

In vascular tissue, REST was found to regulate the proliferation of vascular smooth muscle cells by targeting the K(Ca)3.1 (IKCa) potassium channel. REST was downregulated when there was cellular proliferation, indicating an inverse relationship with IKCa [135]. In addition, REST declines in the medial smooth muscle of atherosclerotic carotids, and REST overexpression inhibits smooth muscle migration. Inhibition of REST and overexpression of IKCa determine smooth muscle and matrix content of atherosclerotic lesions [136].

4. REST in Cancer

4.1. Nervous System Cancer

The dysregulation of REST was found in various nervous system cancers, often playing promoting roles. In neuroblastoma cells, REST was found to be reduced during differentiation [137]. REST was downregulated in mouse neuroblastoma, leading to increased neurite length [138]. It was found that the loss of lncRNA neuroblastoma-associated transcript-1 (NBAT-1) contributed to aggressive neuroblastoma by activating REST [139]. REST is highly expressed in glioma, and its interference inhibits tumors [140]. REST mediated the inhibition of proliferation of glioma by pioglitazone [141]. In medulloblastoma, REST promotes cancer progression via epigenetic modification and AKT activation [142]. The abnormal expression of REST in NSCs blocks their differentiation and promotes medulloblastoma formation [143]. The transient expression of a competitive form of REST, REST-VP16, inhibited medulloblastoma [144][145].

4.2. Non-Nervous System Cancer

The regulation pattern of REST varies in different types of lung cancers. In small-cell lung cancer cells (SCLCs), A splice variant of REST with a 50-bp insert predicting a truncated REST was highly expressed [146]. The decreased REST permits its target gene expression, including glycine receptor alpha1 (GLRA1) subunit, migration-promoting gene ITPKA [147][148][149], leading to de-regulation of AKT activation, promoting malignant progression [150]. The altered REST splicing was found to mediate the anti-tumor effects of SRRM4 targeting in SCLC [151]. Further, miR-9 inhibits the proliferation and migration of lung squamous cell carcinoma cells by blocking REST/EGFR signaling [152]. However, in neuroendocrine lung cancer, REST mediates neuroendocrine to non-neuroendocrine fate switch to play a pro-tumorigenic role [153]. REST-regulated SCG3 provides a sensitive prognostic biomarker for neuroendocrine lung cancer [154].

5. Conclusion

REST is a key transcription regulator that primarily governs gene transcription via the modification of chromatin. REST plays vital roles in both physiological and pathological processes within the nervous system as well as other non-nervous systems. The imbalance of REST dynamic regulation often leads to dysplasia or diseases. Therefore, it is necessary to conduct in-depth research on its involvement in different diseases and the underlying mechanisms is imperative to gain novel insights into potential interventions for related diseases.

This entry is adapted from the peer-reviewed paper 10.3390/biom13101477

References

  1. Chong, J.A.; Tapia-Ramirez, J.; Kim, S.; Toledo-Aral, J.J.; Zheng, Y.; Boutros, M.C.; Altshuller, Y.M.; Frohman, M.A.; Kraner, S.D.; Mandel, G. REST: A mammalian silencer protein that restricts sodium channel gene expression to neurons. Cell 1995, 80, 949–957.
  2. Schoenherr, C.J.; Anderson, D.J. The neuron-restrictive silencer factor (NRSF): A coordinate repressor of multiple neuron-specific genes. Science 1995, 267, 1360–1363.
  3. Chen, G.L.; Miller, G.M. Alternative REST Splicing Underappreciated. Eneuro 2018, 5, 1–6.
  4. Shimojo, M.; Lee, J.H.; Hersh, L.B. Role of zinc finger domains of the transcription factor neuron-restrictive silencer factor/repressor element-1 silencing transcription factor in DNA binding and nuclear localization. J. Biol. Chem. 2001, 276, 13121–13126.
  5. Shimojo, M. Characterization of the nuclear targeting signal of REST/NRSF. Neurosci. Lett. 2006, 398, 161–166.
  6. Lee, J.H.; Shimojo, M.; Chai, Y.G.; Hersh, L.B. Studies on the interaction of REST4 with the cholinergic repressor element-1/neuron restrictive silencer element. Brain Res. Mol. Brain Res. 2000, 80, 88–98.
  7. Zhang, Y.; Hu, W.; Shen, J.; Tong, X.; Yang, Z.; Shen, Z.; Lan, W.; Wu, H.; Cao, C. Cysteine 397 plays important roles in the folding of the neuron-restricted silencer factor/RE1-silencing transcription factor. Biochem. Biophys. Res. Commun. 2011, 414, 309–314.
  8. Yu, H.B.; Johnson, R.; Kunarso, G.; Stanton, L.W. Coassembly of REST and its cofactors at sites of gene repression in embryonic stem cells. Genome Res. 2011, 21, 1284–1293.
  9. Grimes, J.A.; Nielsen, S.J.; Battaglioli, E.; Miska, E.A.; Speh, J.C.; Berry, D.L.; Atouf, F.; Holdener, B.C.; Mandel, G.; Kouzarides, T. The co-repressor mSin3A is a functional component of the REST-CoREST repressor complex. J. Biol. Chem. 2000, 275, 9461–9467.
  10. Andres, M.E.; Burger, C.; Peral-Rubio, M.J.; Battaglioli, E.; Anderson, M.E.; Grimes, J.; Dallman, J.; Ballas, N.; Mandel, G. CoREST: A functional corepressor required for regulation of neural-specific gene expression. Proc. Natl. Acad. Sci. USA 1999, 96, 9873–9878.
  11. Lunyak, V.V.; Burgess, R.; Prefontaine, G.G.; Nelson, C.; Sze, S.H.; Chenoweth, J.; Schwartz, P.; Pevzner, P.A.; Glass, C.; Mandel, G.; et al. Corepressor-dependent silencing of chromosomal regions encoding neuronal genes. Science 2002, 298, 1747–1752.
  12. Ballas, N.; Mandel, G. The many faces of REST oversee epigenetic programming of neuronal genes. Curr. Opin. Neurobiol. 2005, 15, 500–506.
  13. Roopra, A.; Qazi, R.; Schoenike, B.; Daley, T.J.; Morrison, J.F. Localized domains of G9a-mediated histone methylation are required for silencing of neuronal genes. Mol. Cell 2004, 14, 727–738.
  14. Chen, Z.F.; Paquette, A.J.; Anderson, D.J. NRSF/REST is required in vivo for repression of multiple neuronal target genes during embryogenesis. Nat. Genet. 1998, 20, 136–142.
  15. Wang, Y.C.; Liu, P.; Yue, L.Y.; Huang, F.; Xu, Y.X.; Zhu, C.Q. NRSF deficiency leads to abnormal postnatal development of dentate gyrus and impairment of progenitors in subgranular zone of hippocampus. Hippocampus 2021, 31, 935–9563.
  16. Jin, S.; Lee, Y.K.; Lim, Y.C.; Zheng, Z.; Lin, X.M.; Ng, D.P.; Holbrook, J.D.; Law, H.Y.; Kwek, K.Y.; Yeo, G.S.; et al. Global DNA hypermethylation in down syndrome placenta. PLoS Genet. 2013, 9, e1003515.
  17. El Hajj, N.; Dittrich, M.; Bock, J.; Kraus, T.F.; Nanda, I.; Muller, T.; Seidmann, L.; Tralau, T.; Galetzka, D.; Schneider, E.; et al. Epigenetic dysregulation in the developing down syndrome cortex. Epigenetics 2016, 11, 563–578.
  18. Canzonetta, C.; Mulligan, C.; Deutsch, S.; Ruf, S.; O’Doherty, A.; Lyle, R.; Borel, C.; Lin-Marq, N.; Delom, F.; Groet, J.; et al. DYRK1A-dosage imbalance perturbs NRSF/REST levels, deregulating pluripotency and embryonic stem cell fate in Down syndrome. Am. J. Hum. Genet. 2008, 83, 388–400.
  19. Lepagnol-Bestel, A.M.; Zvara, A.; Maussion, G.; Quignon, F.; Ngimbous, B.; Ramoz, N.; Imbeaud, S.; Loe-Mie, Y.; Benihoud, K.; Agier, N.; et al. DYRK1A interacts with the REST/NRSF-SWI/SNF chromatin remodelling complex to deregulate gene clusters involved in the neuronal phenotypic traits of Down syndrome. Hum. Mol. Genet. 2009, 18, 1405–1414.
  20. Nishihara, S.; Tsuda, L.; Ogura, T. The canonical Wnt pathway directly regulates NRSF/REST expression in chick spinal cord. Biochem. Biophys. Res. Commun. 2003, 311, 55–63.
  21. Li, H.; Malbon, C.C.; Wang, H.Y. Gene profiling of Frizzled-1 and Frizzled-2 signaling: Expression of G-protein-coupled receptor chimeras in mouse F9 teratocarcinoma embryonal cells. Mol. Pharmacol. 2004, 65, 45–55.
  22. Johnson, R.; Teh, C.H.; Kunarso, G.; Wong, K.Y.; Srinivasan, G.; Cooper, M.L.; Volta, M.; Chan, S.S.; Lipovich, L.; Pollard, S.M.; et al. REST regulates distinct transcriptional networks in embryonic and neural stem cells. PLoS Biol. 2008, 6, e256.
  23. Ballas, N.; Grunseich, C.; Lu, D.D.; Speh, J.C.; Mandel, G. REST and its corepressors mediate plasticity of neuronal gene chromatin throughout neurogenesis. Cell 2005, 121, 645–657.
  24. Singh, S.K.; Kagalwala, M.N.; Parker-Thornburg, J.; Adams, H.; Majumder, S. REST maintains self-renewal and pluripotency of embryonic stem cells. Nature 2008, 453, 223–227.
  25. Gupta, S.K.; Gressens, P.; Mani, S. NRSF downregulation induces neuronal differentiation in mouse embryonic stem cells. Differentiation 2009, 77, 19–28.
  26. Thakore-Shah, K.; Koleilat, T.; Jan, M.; John, A.; Pyle, A.D. REST/NRSF Knockdown Alters Survival, Lineage Differentiation and Signaling in Human Embryonic Stem Cells. PLoS ONE 2015, 10, e0145280.
  27. Satoh, J.; Kawana, N.; Yamamoto, Y. ChIP-Seq Data Mining: Remarkable Differences in NRSF/REST Target Genes between Human ESC and ESC-Derived Neurons. Bioinform. Biol. Insights 2013, 7, 357–368.
  28. Gao, Z.; Ure, K.; Ding, P.; Nashaat, M.; Yuan, L.; Ma, J.; Hammer, R.E.; Hsieh, J. The master negative regulator REST/NRSF controls adult neurogenesis by restraining the neurogenic program in quiescent stem cells. J. Neurosci. 2011, 31, 9772–9786.
  29. Su, X.; Kameoka, S.; Lentz, S.; Majumder, S. Activation of REST/NRSF target genes in neural stem cells is sufficient to cause neuronal differentiation. Mol. Cell. Biol. 2004, 24, 8018–8025.
  30. Bai, G.; Zhuang, Z.; Liu, A.; Chai, Y.; Hoffman, P.W. The role of the RE1 element in activation of the NR1 promoter during neuronal differentiation. J. Neurochem. 2003, 86, 992–1005.
  31. Kim, S.M.; Yang, J.W.; Park, M.J.; Lee, J.K.; Kim, S.U.; Lee, Y.S.; Lee, M.A. Regulation of human tyrosine hydroxylase gene by neuron-restrictive silencer factor. Biochem. Biophys. Res. Commun. 2006, 346, 426–435.
  32. Aoki, H.; Hara, A.; Oomori, Y.; Shimizu, Y.; Yamada, Y.; Kunisada, T. Neonatal lethality of neural crest cell-specific Rest knockout mice is associated with gastrointestinal distension caused by aberrations of myenteric plexus. Genes Cells 2014, 19, 723–742.
  33. Tomasoni, R.; Negrini, S.; Fiordaliso, S.; Klajn, A.; Tkatch, T.; Mondino, A.; Meldolesi, J.; D’Alessandro, R. A signaling loop of REST, TSC2 and beta-catenin governs proliferation and function of PC12 neural cells. J. Cell Sci. 2011, 124, 3174–3186.
  34. Toshiyuki, N.; Ichiro, M. Molecular mechanisms regulating cell type specific expression of BMP/RA Inducible Neural-specific Protein-1 that suppresses cell cycle progression: Roles of NRSF/REST and DNA methylation. Brain Res. Mol. Brain Res. 2004, 125, 47–59.
  35. Yoo, A.S.; Staahl, B.T.; Chen, L.; Crabtree, G.R. MicroRNA-mediated switching of chromatin-remodelling complexes in neural development. Nature 2009, 460, 642–646.
  36. Formisano, L.; Guida, N.; Laudati, G.; Boscia, F.; Esposito, A.; Secondo, A.; Di Renzo, G.; Canzoniero, L.M. Extracellular signal-related kinase 2/specificity protein 1/specificity protein 3/repressor element-1 silencing transcription factor pathway is involved in Aroclor 1254-induced toxicity in SH-SY5Y neuronal cells. J. Neurosci. Res. 2015, 93, 167–177.
  37. Rivera-Cervantes, M.C.; Castaneda-Arellano, R.; Castro-Torres, R.D.; Gudino-Cabrera, G.; Feria y Velasco, A.I.; Camins, A.; Beas-Zarate, C. P38 MAPK inhibition protects against glutamate neurotoxicity and modifies NMDA and AMPA receptor subunit expression. J. Mol. Neurosci. 2015, 55, 596–608.
  38. Pajarillo, E.; Rizor, A.; Son, D.S.; Aschner, M.; Lee, E. The transcription factor REST up-regulates tyrosine hydroxylase and antiapoptotic genes and protects dopaminergic neurons against manganese toxicity. J. Biol. Chem. 2020, 295, 3040–3054.
  39. Kallunki, P.; Edelman, G.M.; Jones, F.S. Tissue-specific expression of the L1 cell adhesion molecule is modulated by the neural restrictive silencer element. J. Cell Biol. 1997, 138, 1343–1354.
  40. Schulte, C.; Racchetti, G.; D’Alessandro, R.; Meldolesi, J. A new form of neurite outgrowth sustained by the exocytosis of enlargeosomes expressed under the control of REST. Traffic 2010, 11, 1304–1314.
  41. Paquette, A.J.; Perez, S.E.; Anderson, D.J. Constitutive expression of the neuron-restrictive silencer factor (NRSF)/REST in differentiating neurons disrupts neuronal gene expression and causes axon pathfinding errors in vivo. Proc. Natl. Acad. Sci. USA 2000, 97, 12318–12323.
  42. Schoch, S.; Cibelli, G.; Thiel, G. Neuron-specific gene expression of synapsin I. Major role of a negative regulatory mechanism. J. Biol. Chem. 1996, 271, 3317–3323.
  43. Chin, L.S.; Weigel, C.; Li, L. Transcriptional regulation of gene expression of sec6, a component of mammalian exocyst complex at the synapse. Brain Res. Mol. Brain Res. 2000, 79, 127–137.
  44. Pozzi, D.; Lignani, G.; Ferrea, E.; Contestabile, A.; Paonessa, F.; D’Alessandro, R.; Lippiello, P.; Boido, D.; Fassio, A.; Meldolesi, J.; et al. REST/NRSF-mediated intrinsic homeostasis protects neuronal networks from hyperexcitability. EMBO J. 2013, 32, 2994–3007.
  45. Kim, C.S.; Hwang, C.K.; Song, K.Y.; Choi, H.S.; Kim, D.K.; Law, P.Y.; Wei, L.N.; Loh, H.H. Novel function of neuron-restrictive silencer factor (NRSF) for posttranscriptional regulation. Biochim. Biophys. Acta 2008, 1783, 1835–1846.
  46. Kim, C.S.; Hwang, C.K.; Choi, H.S.; Song, K.Y.; Law, P.Y.; Wei, L.N.; Loh, H.H. Neuron-restrictive silencer factor (NRSF) functions as a repressor in neuronal cells to regulate the mu opioid receptor gene. J. Biol. Chem. 2004, 279, 46464–46473.
  47. Kim, C.S.; Choi, H.S.; Hwang, C.K.; Song, K.Y.; Lee, B.K.; Law, P.Y.; Wei, L.N.; Loh, H.H. Evidence of the neuron-restrictive silencer factor (NRSF) interaction with Sp3 and its synergic repression to the mu opioid receptor (MOR) gene. Nucleic Acids Res. 2006, 34, 6392–6403.
  48. Kong, M.; Shi, L.; Zhou, Y.; He, J.; Zhang, W.; Gu, X.; Zhang, J.; Ma, Z. Changes of Mu-opioid receptor and neuron-restrictive silencer factor in periaquductal gray in mouse models of remifentanil-induced postoperative hyperalgesia. Zhong Nan Da Xue Xue Bao Yi Xue Ban 2014, 39, 901–906.
  49. Shudo, Y.; Shimojo, M.; Fukunaga, M.; Ito, S. Pituitary adenylate cyclase-activating polypeptide is regulated by alternative splicing of transcriptional repressor REST/NRSF in nerve injury. Life Sci. 2015, 143, 174–181.
  50. Wang, D.; Yu, J. Negative regulation of REST on NR2B in spinal cord contributes to the development of bone cancer pain in mice. Oncotarget 2016, 7, 85564–85572.
  51. Zhu, C.; Tang, J.; Ding, T.; Chen, L.; Wang, W.; Mei, X.P.; He, X.T.; Wang, W.; Zhang, L.D.; Dong, Y.L.; et al. Neuron-restrictive silencer factor-mediated downregulation of mu-opioid receptor contributes to the reduced morphine analgesia in bone cancer pain. Pain 2017, 158, 879–890.
  52. Xiao-Die, X.; Xiao-Hong, W.; Cheng-Feng, H.; Zhong-Yu, Y.; Jian-Tao, W.; Hou-Guang, Z.; Jing-Chun, G. Increased NRSF/REST in anterior cingulate cortex contributes to diabetes-related neuropathic pain. Biochem. Biophys. Res. Commun. 2020, 527, 785–790.
  53. Zhang, J.; Chen, S.R.; Chen, H.; Pan, H.L. RE1-silencing transcription factor controls the acute-to-chronic neuropathic pain transition and Chrm2 receptor gene expression in primary sensory neurons. J. Biol. Chem. 2018, 293, 19078–19091.
  54. Zhang, F.; Gigout, S.; Liu, Y.; Wang, Y.; Hao, H.; Buckley, N.J.; Zhang, H.; Wood, I.C.; Gamper, N. Repressor element 1-silencing transcription factor drives the development of chronic pain states. Pain 2019, 160, 2398–2408.
  55. Li, Y.; Liu, Q.; Yang, Y.; Lv, Y.; Chen, L.; Bai, C.; Nan, X.; Wang, Y.; Pei, X. Regulatory role of neuron-restrictive silencing factor in the specific expression of cocaine- and amphetamine-regulated transcript gene. J. Neurochem. 2008, 106, 1314–1324.
  56. D’Alessandro, R.; Klajn, A.; Stucchi, L.; Podini, P.; Malosio, M.L.; Meldolesi, J. Expression of the neurosecretory process in PC12 cells is governed by REST. J. Neurochem. 2008, 105, 1369–1383.
  57. Prada, I.; Marchaland, J.; Podini, P.; Magrassi, L.; D’Alessandro, R.; Bezzi, P.; Meldolesi, J. REST/NRSF governs the expression of dense-core vesicle gliosecretion in astrocytes. J. Cell Biol. 2011, 193, 537–549.
  58. Valiya Veettil, M.; Krishna, G.; Roy, A.; Ghosh, A.; Dutta, D.; Kumar, B.; Chakraborty, S.; Anju, T.R.; Sharma-Walia, N.; Chandran, B. Kaposi’s Sarcoma-Associated Herpesvirus Infection Induces the Expression of Neuroendocrine Genes in Endothelial Cells. J. Virol. 2020, 94, e01692-19.
  59. Rockowitz, S.; Zheng, D. Significant expansion of the REST/NRSF cistrome in human versus mouse embryonic stem cells: Potential implications for neural development. Nucleic Acids Res. 2015, 43, 5730–5743.
  60. Wang, P.; Zhao, D.; Rockowitz, S.; Zheng, D. Divergence and rewiring of regulatory networks for neural development between human and other species. Neurogenesis 2016, 3, e1231495.
  61. Bolton, J.L.; Schulmann, A.; Garcia-Curran, M.M.; Regev, L.; Chen, Y.; Kamei, N.; Shao, M.; Singh-Taylor, A.; Jiang, S.; Noam, Y.; et al. Unexpected Transcriptional Programs Contribute to Hippocampal Memory Deficits and Neuronal Stunting after Early-Life Adversity. Cell Rep. 2020, 33, 108511.
  62. Yang, Y.; Zhang, X.; Li, D.; Fang, R.; Wang, Z.; Yun, D.; Wang, M.; Wang, J.; Dong, H.; Fei, Z.; et al. NRSF regulates age-dependently cognitive ability and its conditional knockout in APP/PS1 mice moderately alters ad-like pathology. Hum. Mol. Genet. 2022, 32, 2558–2575.
  63. Yang, M.; Li, Y.; Hu, L.; Luo, D.; Zhang, Y.; Xiao, X.; Li, G.; Zhang, L.; Zhu, G. Lead exposure inhibits expression of SV2C through NRSF. Toxicology 2018, 398–399, 23–30.
  64. Verma, P.; Gupta, R.K.; Gandhi, B.S.; Singh, P. CDRI-08 Attenuates REST/NRSF-Mediated Expression of NMDAR1 Gene in PBDE-209-Exposed Mice Brain. Evid. Based Complement. Alternat Med. 2015, 2015, 403840.
  65. Mori, N.; Mizuno, T.; Murai, K.; Nakano, I.; Yamashita, H. Effect of age on the gene expression of neural-restrictive silencing factor NRSF/REST. Neurobiol. Aging 2002, 23, 255–262.
  66. Rocchi, A.; Carminati, E.; De Fusco, A.; Kowalska, J.A.; Floss, T.; Benfenati, F. REST/NRSF deficiency impairs autophagy and leads to cellular senescence in neurons. Aging Cell 2021, 20, e13471.
  67. Yang, P.; Sun, R.; Yao, M.; Chen, W.; Wang, Z.; Fei, J. A C-terminal truncated mutation of spr-3 gene extends lifespan in Caenorhabditis elegans. Acta Biochim. Biophys. Sin. 2013, 45, 540–548.
  68. Xue, W.J.; He, C.F.; Zhou, R.Y.; Xu, X.D.; Xiang, L.X.; Wang, J.T.; Wang, X.R.; Zhou, H.G.; Guo, J.C. High glucose and palmitic acid induces neuronal senescence by NRSF/REST elevation and the subsequent mTOR-related autophagy suppression. Mol. Brain 2022, 15, 61.
  69. Obata, H.; Naito, K.; Kikui, A.; Nakamura, S.; Suzuki, K.; Kawakita, S.; Suzuki, T.; Goto, K.; Nagura, N.; Sugiyama, Y.; et al. Age-related differences for expression of the nerve-specific proteins after peripheral nerve injury. Exp. Ther. Med. 2022, 24, 682.
  70. Kawamura, M.; Sato, S.; Matsumoto, G.; Fukuda, T.; Shiba-Fukushima, K.; Noda, S.; Takanashi, M.; Mori, N.; Hattori, N. Loss of nuclear REST/NRSF in aged-dopaminergic neurons in Parkinson’s disease patients. Neurosci. Lett. 2019, 699, 59–63.
  71. Yu, M.; Cai, L.; Liang, M.; Huang, Y.; Gao, H.; Lu, S.; Fei, J.; Huang, F. Alteration of NRSF expression exacerbating 1-methyl-4-phenyl-pyridinium ion-induced cell death of SH-SY5Y cells. Neurosci. Res. 2009, 65, 236–244.
  72. Yu, M.; Suo, H.; Liu, M.; Cai, L.; Liu, J.; Huang, Y.; Xu, J.; Wang, Y.; Zhu, C.; Fei, J.; et al. NRSF/REST neuronal deficient mice are more vulnerable to the neurotoxin MPTP. Neurobiol. Aging 2013, 34, 916–927.
  73. Suo, H.; Wang, P.; Tong, J.; Cai, L.; Liu, J.; Huang, D.; Huang, L.; Wang, Z.; Huang, Y.; Xu, J.; et al. NRSF is an essential mediator for the neuroprotection of trichostatin A in the MPTP mouse model of Parkinson’s disease. Neuropharmacology 2015, 99, 67–78.
  74. Huang, D.; Li, Q.; Wang, Y.; Liu, Z.; Wang, Z.; Li, H.; Wang, J.; Su, J.; Ma, Y.; Yu, M.; et al. Brain-specific NRSF deficiency aggravates dopaminergic neurodegeneration and impairs neurogenesis in the MPTP mouse model of Parkinson’s disease. Aging 2019, 11, 3280–3297.
  75. Li, H.; Liu, Z.; Wu, Y.; Chen, Y.; Wang, J.; Wang, Z.; Huang, D.; Wang, M.; Yu, M.; Fei, J.; et al. The deficiency of NRSF/REST enhances the pro-inflammatory function of astrocytes in a model of Parkinson’s disease. Biochim. Biophys. Acta Mol. Basis Dis. 2020, 1866, 165590.
  76. Zuccato, C.; Tartari, M.; Crotti, A.; Goffredo, D.; Valenza, M.; Conti, L.; Cataudella, T.; Leavitt, B.R.; Hayden, M.R.; Timmusk, T.; et al. Huntingtin interacts with REST/NRSF to modulate the transcription of NRSE-controlled neuronal genes. Nat. Genet. 2003, 35, 76–83.
  77. Marullo, M.; Valenza, M.; Mariotti, C.; Di Donato, S.; Cattaneo, E.; Zuccato, C. Analysis of the repressor element-1 silencing transcription factor/neuron-restrictive silencer factor occupancy of non-neuronal genes in peripheral lymphocytes from patients with Huntington’s disease. Brain Pathol. 2010, 20, 96–105.
  78. Schiffer, D.; Caldera, V.; Mellai, M.; Conforti, P.; Cattaneo, E.; Zuccato, C. Repressor element-1 silencing transcription factor (REST) is present in human control and Huntington’s disease neurones. Neuropathol. Appl. Neurobiol. 2014, 40, 899–910.
  79. Zuccato, C.; Belyaev, N.; Conforti, P.; Ooi, L.; Tartari, M.; Papadimou, E.; MacDonald, M.; Fossale, E.; Zeitlin, S.; Buckley, N.; et al. Widespread disruption of repressor element-1 silencing transcription factor/neuron-restrictive silencer factor occupancy at its target genes in Huntington’s disease. J. Neurosci. 2007, 27, 6972–6983.
  80. Shimojo, M. Huntingtin regulates RE1-silencing transcription factor/neuron-restrictive silencer factor (REST/NRSF) nuclear trafficking indirectly through a complex with REST/NRSF-interacting LIM domain protein (RILP) and dynactin p150 Glued. J. Biol. Chem. 2008, 283, 34880–34886.
  81. Datta, M.; Bhattacharyya, N.P. Regulation of RE1 protein silencing transcription factor (REST) expression by HIP1 protein interactor (HIPPI). J. Biol. Chem. 2011, 286, 33759–33769.
  82. Chen, G.L.; Ma, Q.; Goswami, D.; Shang, J.; Miller, G.M. Modulation of nuclear REST by alternative splicing: A potential therapeutic target for Huntington’s disease. J. Cell. Mol. Med. 2017, 21, 2974–2984.
  83. Navarrete-Modesto, V.; Orozco-Suarez, S.; Alonso-Vanegas, M.; Feria-Romero, I.A.; Rocha, L. REST/NRSF transcription factor is overexpressed in hippocampus of patients with drug-resistant mesial temporal lobe epilepsy. Epilepsy Behav. 2019, 94, 118–123.
  84. Hu, X.L.; Cheng, X.; Cai, L.; Tan, G.H.; Xu, L.; Feng, X.Y.; Lu, T.J.; Xiong, H.; Fei, J.; Xiong, Z.Q. Conditional deletion of NRSF in forebrain neurons accelerates epileptogenesis in the kindling model. Cereb. Cortex 2011, 21, 2158–2165.
  85. Liu, M.; Sheng, Z.; Cai, L.; Zhao, K.; Tian, Y.; Fei, J. Neuronal conditional knockout of NRSF decreases vulnerability to seizures induced by pentylenetetrazol in mice. Acta Biochim. Biophys. Sin. 2012, 44, 476–482.
  86. Patterson, K.P.; Barry, J.M.; Curran, M.M.; Singh-Taylor, A.; Brennan, G.; Rismanchi, N.; Page, M.; Noam, Y.; Holmes, G.L.; Baram, T.Z. Enduring Memory Impairments Provoked by Developmental Febrile Seizures Are Mediated by Functional and Structural Effects of Neuronal Restrictive Silencing Factor. J. Neurosci. 2017, 37, 3799–3812.
  87. McClelland, S.; Flynn, C.; Dubé, C.; Richichi, C.; Zha, Q.; Ghestem, A.; Esclapez, M.; Bernard, C.; Baram, T.Z. Neuron-restrictive silencer factor-mediated hyperpolarization-activated cyclic nucleotide gated channelopathy in experimental temporal lobe epilepsy. Ann. Neurol. 2011, 70, 454–464.
  88. Carminati, E.; Buffolo, F.; Rocchi, A.; Michetti, C.; Cesca, F.; Benfenati, F. Mild Inactivation of RE-1 Silencing Transcription Factor (REST) Reduces Susceptibility to Kainic Acid-Induced Seizures. Front. Cell. Neurosci. 2019, 13, 580.
  89. Garriga-Canut, M.; Schoenike, B.; Qazi, R.; Bergendahl, K.; Daley, T.J.; Pfender, R.M.; Morrison, J.F.; Ockuly, J.; Stafstrom, C.; Sutula, T.; et al. 2-Deoxy-D-glucose reduces epilepsy progression by NRSF-CtBP-dependent metabolic regulation of chromatin structure. Nat. Neurosci. 2006, 9, 1382–1387.
  90. Hu, X.L.; Cheng, X.; Fei, J.; Xiong, Z.Q. Neuron-restrictive silencer factor is not required for the antiepileptic effect of the ketogenic diet. Epilepsia 2011, 52, 1609–1616.
  91. Bassuk, A.G.; Wallace, R.H.; Buhr, A.; Buller, A.R.; Afawi, Z.; Shimojo, M.; Miyata, S.; Chen, S.; Gonzalez-Alegre, P.; Griesbach, H.L.; et al. A homozygous mutation in human PRICKLE1 causes an autosomal-recessive progressive myoclonus epilepsy-ataxia syndrome. Am. J. Hum. Genet. 2008, 83, 572–581.
  92. Quinn, J.P.; Bubb, V.J.; Marshall-Jones, Z.V.; Coulson, J.M. Neuron restrictive silencer factor as a modulator of neuropeptide gene expression. Regul. Pept. 2002, 108, 135–141.
  93. Spencer, E.M.; Chandler, K.E.; Haddley, K.; Howard, M.R.; Hughes, D.; Belyaev, N.D.; Coulson, J.M.; Stewart, J.P.; Buckley, N.J.; Kipar, A.; et al. Regulation and role of REST and REST4 variants in modulation of gene expression in in vivo and in vitro in epilepsy models. Neurobiol. Dis. 2006, 24, 41–52.
  94. Gillies, S.; Haddley, K.; Vasiliou, S.; Bubb, V.J.; Quinn, J.P. The human neurokinin B gene, TAC3, and its promoter are regulated by Neuron Restrictive Silencing Factor (NRSF) transcription factor family. Neuropeptides 2009, 43, 333–340.
  95. Hall, A.M.; Brennan, G.P.; Nguyen, T.M.; Singh-Taylor, A.; Mun, H.S.; Sargious, M.J.; Baram, T.Z. The Role of Sirt1 in Epileptogenesis. Eneuro 2017, preprint.
  96. Brennan, G.P.; Dey, D.; Chen, Y.; Patterson, K.P.; Magnetta, E.J.; Hall, A.M.; Dube, C.M.; Mei, Y.T.; Baram, T.Z. Dual and Opposing Roles of MicroRNA-124 in Epilepsy Are Mediated through Inflammatory and NRSF-Dependent Gene Networks. Cell Rep. 2016, 14, 2402–2412.
  97. Calderone, A.; Jover, T.; Noh, K.M.; Tanaka, H.; Yokota, H.; Lin, Y.; Grooms, S.Y.; Regis, R.; Bennett, M.V.; Zukin, R.S. Ischemic insults derepress the gene silencer REST in neurons destined to die. J. Neurosci. 2003, 23, 2112–2121.
  98. Formisano, L.; Noh, K.M.; Miyawaki, T.; Mashiko, T.; Bennett, M.V.; Zukin, R.S. Ischemic insults promote epigenetic reprogramming of mu opioid receptor expression in hippocampal neurons. Proc. Natl. Acad. Sci. USA 2007, 104, 4170–4175.
  99. Liang, H.M.; Geng, L.J.; Shi, X.Y.; Zhang, C.G.; Wang, S.Y.; Zhang, G.M. By up-regulating mu- and delta-opioid receptors, neuron-restrictive silencer factor knockdown promotes neurological recovery after ischemia. Oncotarget 2017, 8, 101012–101025.
  100. He, C.F.; Xue, W.J.; Xu, X.D.; Wang, J.T.; Wang, X.R.; Feng, Y.; Zhou, H.G.; Guo, J.C. Knockdown of NRSF Alleviates Ischemic Brain Injury and Microvasculature Defects in Diabetic MCAO Mice. Front. Neurol. 2022, 13, 869220.
  101. Zhang, J.; Wang, S.; Yuan, L.; Yang, Y.; Zhang, B.; Liu, Q.; Chen, L.; Yue, W.; Li, Y.; Pei, X. Neuron-restrictive silencer factor (NRSF) represses cocaine- and amphetamine-regulated transcript (CART) transcription and antagonizes cAMP-response element-binding protein signaling through a dual NRSE mechanism. J. Biol. Chem. 2012, 287, 42574–42587.
  102. Kaneko, N.; Hwang, J.Y.; Gertner, M.; Pontarelli, F.; Zukin, R.S. Casein kinase 1 suppresses activation of REST in insulted hippocampal neurons and halts ischemia-induced neuronal death. J. Neurosci. 2014, 34, 6030–6039.
  103. Loe-Mie, Y.; Lepagnol-Bestel, A.M.; Maussion, G.; Doron-Faigenboim, A.; Imbeaud, S.; Delacroix, H.; Aggerbeck, L.; Pupko, T.; Gorwood, P.; Simonneau, M.; et al. SMARCA2 and other genome-wide supported schizophrenia-associated genes: Regulation by REST/NRSF, network organization and primate-specific evolution. Hum. Mol. Genet. 2010, 19, 2841–2857.
  104. Gulchina, Y.; Xu, S.J.; Snyder, M.A.; Elefant, F.; Gao, W.J. Epigenetic mechanisms underlying NMDA receptor hypofunction in the prefrontal cortex of juvenile animals in the MAM model for schizophrenia. J. Neurochem. 2017, 143, 320–333.
  105. Patel, P.D.; Bochar, D.A.; Turner, D.L.; Meng, F.; Mueller, H.M.; Pontrello, C.G. Regulation of tryptophan hydroxylase-2 gene expression by a bipartite RE-1 silencer of transcription/neuron restrictive silencing factor (REST/NRSF) binding motif. J. Biol. Chem. 2007, 282, 26717–26724.
  106. Ishii, T.; Hashimoto, E.; Ukai, W.; Tateno, M.; Yoshinaga, T.; Saito, S.; Sohma, H.; Saito, T. Lithium-induced suppression of transcription repressor NRSF/REST: Effects on the dysfunction of neuronal differentiation by ethanol. Eur. J. Pharmacol. 2008, 593, 36–43.
  107. Warburton, A.; Savage, A.L.; Myers, P.; Peeney, D.; Bubb, V.J.; Quinn, J.P. Molecular signatures of mood stabilisers highlight the role of the transcription factor REST/NRSF. J. Affect. Disord. 2015, 172, 63–73.
  108. Henriksson, R.; Backman, C.M.; Harvey, B.K.; Kadyrova, H.; Bazov, I.; Shippenberg, T.S.; Bakalkin, G. PDYN, a gene implicated in brain/mental disorders, is targeted by REST in the adult human brain. Biochim. Biophys. Acta 2014, 1839, 1226–1232.
  109. Korosi, A.; Shanabrough, M.; McClelland, S.; Liu, Z.W.; Borok, E.; Gao, X.B.; Horvath, T.L.; Baram, T.Z. Early-life experience reduces excitation to stress-responsive hypothalamic neurons and reprograms the expression of corticotropin-releasing hormone. J. Neurosci. 2010, 30, 703–713.
  110. Hai-Ying, C.; Nagano, K.; Ezzikouri, S.; Yamaguchi, C.; Kayesh, M.E.; Rebbani, K.; Kitab, B.; Nakano, H.; Kouji, H.; Kohara, M.; et al. Establishment of an intermittent cold stress model using Tupaia belangeri and evaluation of compound C737 targeting neuron-restrictive silencer factor. Exp. Anim. 2016, 65, 285–292.
  111. Singh-Taylor, A.; Molet, J.; Jiang, S.; Korosi, A.; Bolton, J.L.; Noam, Y.; Simeone, K.; Cope, J.; Chen, Y.; Mortazavi, A.; et al. NRSF-dependent epigenetic mechanisms contribute to programming of stress-sensitive neurons by neonatal experience, promoting resilience. Mol. Psychiatry 2018, 23, 648–657.
  112. Wen, X.H.; Guo, Q.L.; Guo, J.C. Effect of 9—PAHSA on cognitive dysfunction in diabetic mice and its possible mechanism. Biochem. Biophys. Res. Commun. 2020, 524, 525–532.
  113. Kuwahara, K.; Saito, Y.; Takano, M.; Arai, Y.; Yasuno, S.; Nakagawa, Y.; Takahashi, N.; Adachi, Y.; Takemura, G.; Horie, M.; et al. NRSF regulates the fetal cardiac gene program and maintains normal cardiac structure and function. EMBO J. 2003, 22, 6310–6321.
  114. Kuratomi, S.; Kuratomi, A.; Kuwahara, K.; Ishii, T.M.; Nakao, K.; Saito, Y.; Takano, M. NRSF regulates the developmental and hypertrophic changes of HCN4 transcription in rat cardiac myocytes. Biochem. Biophys. Res. Commun. 2007, 353, 67–73.
  115. Schweizer, P.A.; Yampolsky, P.; Malik, R.; Thomas, D.; Zehelein, J.; Katus, H.A.; Koenen, M. Transcription profiling of HCN-channel isotypes throughout mouse cardiac development. Basic. Res. Cardiol. 2009, 104, 621–629.
  116. Zhang, D.; Wu, B.; Wang, P.; Wang, Y.; Lu, P.; Nechiporuk, T.; Floss, T.; Greally, J.M.; Zheng, D.; Zhou, B. Non-CpG methylation by DNMT3B facilitates REST binding and gene silencing in developing mouse hearts. Nucleic Acids Res. 2017, 45, 3102–3115.
  117. Liu, J.; Liu, S.; Gao, H.; Han, L.; Chu, X.; Sheng, Y.; Shou, W.; Wang, Y.; Liu, Y.; Wan, J.; et al. Genome-wide studies reveal the essential and opposite roles of ARID1A in controlling human cardiogenesis and neurogenesis from pluripotent stem cells. Genome Biol. 2020, 21, 169.
  118. D’Souza, A.; Bucchi, A.; Johnsen, A.B.; Logantha, S.J.; Monfredi, O.; Yanni, J.; Prehar, S.; Hart, G.; Cartwright, E.; Wisloff, U.; et al. Exercise training reduces resting heart rate via downregulation of the funny channel HCN4. Nat. Commun. 2014, 5, 3775.
  119. Koyama, R.; Mannic, T.; Ito, J.; Amar, L.; Zennaro, M.C.; Rossier, M.F.; Maturana, A.D. MicroRNA-204 Is Necessary for Aldosterone-Stimulated T-Type Calcium Channel Expression in Cardiomyocytes. Int. J. Mol. Sci. 2018, 19, 2941.
  120. Rossier, M.F. The Cardiac Mineralocorticoid Receptor (MR): A Therapeutic Target Against Ventricular Arrhythmias. Front. Endocrinol. 2021, 12, 694758.
  121. Somekawa, S.; Imagawa, K.; Naya, N.; Takemoto, Y.; Onoue, K.; Okayama, S.; Takeda, Y.; Kawata, H.; Horii, M.; Nakajima, T.; et al. Regulation of aldosterone and cortisol production by the transcriptional repressor neuron restrictive silencer factor. Endocrinology 2009, 150, 3110–3117.
  122. Inazumi, H.; Kuwahara, K.; Nakagawa, Y.; Kuwabara, Y.; Numaga-Tomita, T.; Kashihara, T.; Nakada, T.; Kurebayashi, N.; Oya, M.; Nonaka, M.; et al. NRSF-GNAO1 Pathway Contributes to the Regulation of Cardiac Ca2+ Homeostasis. Circ. Res. 2022, 130, 234–248.
  123. Atouf, F.; Czernichow, P.; Scharfmann, R. Expression of neuronal traits in pancreatic beta cells. Implication of neuron-restrictive silencing factor/repressor element silencing transcription factor, a neuron-restrictive silencer. J. Biol. Chem. 1997, 272, 1929–1934.
  124. Martin, D.; Kim, Y.H.; Sever, D.; Mao, C.A.; Haefliger, J.A.; Grapin-Botton, A. REST represses a subset of the pancreatic endocrine differentiation program. Dev. Biol. 2015, 405, 316–327.
  125. Kemp, D.M.; Lin, J.C.; Habener, J.F. Regulation of Pax4 paired homeodomain gene by neuron-restrictive silencer factor. J. Biol. Chem. 2003, 278, 35057–35062.
  126. Martin, D.; Tawadros, T.; Meylan, L.; Abderrahmani, A.; Condorelli, D.F.; Waeber, G.; Haefliger, J.A. Critical role of the transcriptional repressor neuron-restrictive silencer factor in the specific control of connexin36 in insulin-producing cell lines. J. Biol. Chem. 2003, 278, 53082–53089.
  127. Campa, D.; Capurso, G.; Pastore, M.; Talar-Wojnarowska, R.; Milanetto, A.C.; Landoni, L.; Maiello, E.; Lawlor, R.T.; Malecka-Panas, E.; Funel, N.; et al. Common germline variants within the CDKN2A/2B region affect risk of pancreatic neuroendocrine tumors. Sci. Rep. 2016, 6, 39565.
  128. Li, B.; Wang, S.; Liu, H.; Liu, D.; Zhang, J.; Zhang, B.; Yao, H.; Lv, Y.; Wang, R.; Chen, L.; et al. Neuronal restrictive silencing factor silencing induces human amniotic fluid-derived stem cells differentiation into insulin-producing cells. Stem Cells Dev. 2011, 20, 1223–1231.
  129. Li, H.T.; Jiang, F.X.; Shi, P.; Zhang, T.; Liu, X.Y.; Lin, X.W.; Pang, X.N. In vitro reprogramming of rat bone marrow-derived mesenchymal stem cells into insulin-producing cells by genetically manipulating negative and positive regulators. Biochem. Biophys. Res. Commun. 2012, 420, 793–798.
  130. Kurschat, P.; Bielenberg, D.; Rossignol-Tallandier, M.; Stahl, A.; Klagsbrun, M. Neuron restrictive silencer factor NRSF/REST is a transcriptional repressor of neuropilin-1 and diminishes the ability of semaphorin 3A to inhibit keratinocyte migration. J. Biol. Chem. 2006, 281, 2721–2729.
  131. Aoki, H.; Hara, A.; Kunisada, T. White spotting phenotype induced by targeted REST disruption during neural crest specification to a melanocyte cell lineage. Genes Cells 2015, 20, 439–449.
  132. Lee, H.J.; Wall, B.A.; Wangari-Talbot, J.; Chen, S. Regulation of mGluR1 expression in human melanocytes and melanoma cells. Biochim. Biophys. Acta 2012, 1819, 1123–1131.
  133. Aoki, H.; Ogino, H.; Tomita, H.; Hara, A.; Kunisada, T. Disruption of Rest Leads to the Early Onset of Cataracts with the Aberrant Terminal Differentiation of Lens Fiber Cells. PLoS ONE 2016, 11, e0163042.
  134. Tsuda, L.; Kaido, M.; Lim, Y.M.; Kato, K.; Aigaki, T.; Hayashi, S. An NRSF/REST-like repressor downstream of Ebi/SMRTER/Su(H) regulates eye development in Drosophila. EMBO J. 2006, 25, 3191–3202.
  135. Cheong, A.; Bingham, A.J.; Li, J.; Kumar, B.; Sukumar, P.; Munsch, C.; Buckley, N.J.; Neylon, C.B.; Porter, K.E.; Beech, D.J.; et al. Downregulated REST transcription factor is a switch enabling critical potassium channel expression and cell proliferation. Mol. Cell 2005, 20, 45–52.
  136. Tharp, D.L.; Bowles, D.K. KCa3.1 Inhibition Decreases Size and Alters Composition of Atherosclerotic Lesions Induced by Low, Oscillatory Flow. Artery Res. 2021, 27, 93–100.
  137. Nishimura, E.; Sasaki, K.; Maruyama, K.; Tsukada, T.; Yamaguchi, K. Decrease in neuron-restrictive silencer factor (NRSF) mRNA levels during differentiation of cultured neuroblastoma cells. Neurosci. Lett. 1996, 211, 101–104.
  138. Lepagnol-Bestel, A.M.; Maussion, G.; Ramoz, N.; Moalic, J.M.; Gorwood, P.; Simonneau, M. Nrsf silencing induces molecular and subcellular changes linked to neuronal plasticity. NeuroReport 2007, 18, 441–446.
  139. Pandey, G.K.; Mitra, S.; Subhash, S.; Hertwig, F.; Kanduri, M.; Mishra, K.; Fransson, S.; Ganeshram, A.; Mondal, T.; Bandaru, S.; et al. The risk-associated long noncoding RNA NBAT-1 controls neuroblastoma progression by regulating cell proliferation and neuronal differentiation. Cancer Cell 2014, 26, 722–737.
  140. Conti, L.; Crisafulli, L.; Caldera, V.; Tortoreto, M.; Brilli, E.; Conforti, P.; Zunino, F.; Magrassi, L.; Schiffer, D.; Cattaneo, E. REST controls self-renewal and tumorigenic competence of human glioblastoma cells. PLoS ONE 2012, 7, e38486.
  141. Ren, H.; Gao, Z.; Wu, N.; Zeng, L.; Tang, X.; Chen, X.; Liu, Z.; Zhang, W.; Wang, L.; Li, Z. Expression of REST4 in human gliomas in vivo and influence of pioglitazone on REST in vitro. Biochem. Biophys. Res. Commun. 2015, 463, 504–509.
  142. Dobson, T.H.W.; Tao, R.H.; Swaminathan, J.; Maegawa, S.; Shaik, S.; Bravo-Alegria, J.; Sharma, A.; Kennis, B.; Yang, Y.; Callegari, K.; et al. Transcriptional repressor REST drives lineage stage-specific chromatin compaction at Ptch1 and increases AKT activation in a mouse model of medulloblastoma. Sci. Signal 2019, 12, eaan8680.
  143. Su, X.; Gopalakrishnan, V.; Stearns, D.; Aldape, K.; Lang, F.F.; Fuller, G.; Snyder, E.; Eberhart, C.G.; Majumder, S. Abnormal expression of REST/NRSF and Myc in neural stem/progenitor cells causes cerebellar tumors by blocking neuronal differentiation. Mol. Cell. Biol. 2006, 26, 1666–1678.
  144. Lawinger, P.; Venugopal, R.; Guo, Z.S.; Immaneni, A.; Sengupta, D.; Lu, W.; Rastelli, L.; Marin Dias Carneiro, A.; Levin, V.; Fuller, G.N.; et al. The neuronal repressor REST/NRSF is an essential regulator in medulloblastoma cells. Nat. Med. 2000, 6, 826–831.
  145. Fuller, G.N.; Su, X.; Price, R.E.; Cohen, Z.R.; Lang, F.F.; Sawaya, R.; Majumder, S. Many human medulloblastoma tumors overexpress repressor element-1 silencing transcription (REST)/neuron-restrictive silencer factor, which can be functionally countered by REST-VP16. Mol. Cancer Ther. 2005, 4, 343–349.
  146. Coulson, J.M.; Edgson, J.L.; Woll, P.J.; Quinn, J.P. A splice variant of the neuron-restrictive silencer factor repressor is expressed in small cell lung cancer: A potential role in derepression of neuroendocrine genes and a useful clinical marker. Cancer Res. 2000, 60, 1840–1844.
  147. Gurrola-Diaz, C.; Lacroix, J.; Dihlmann, S.; Becker, C.M.; von Knebel Doeberitz, M. Reduced expression of the neuron restrictive silencer factor permits transcription of glycine receptor alpha1 subunit in small-cell lung cancer cells. Oncogene 2003, 22, 5636–5645.
  148. Neumann, S.B.; Seitz, R.; Gorzella, A.; Heister, A.; Doeberitz, M.; Becker, C.M. Relaxation of glycine receptor and onconeural gene transcription control in NRSF deficient small cell lung cancer cell lines. Brain Res. Mol. Brain Res. 2004, 120, 173–181.
  149. Chang, L.; Schwarzenbach, H.; Meyer-Staeckling, S.; Brandt, B.; Mayr, G.W.; Weitzel, J.M.; Windhorst, S. Expression Regulation of the Metastasis-Promoting Protein InsP3-Kinase-A in Tumor Cells. Mol. Cancer Res. 2011, 9, 497–506.
  150. Kreisler, A.; Strissel, P.L.; Strick, R.; Neumann, S.B.; Schumacher, U.; Becker, C.M. Regulation of the NRSF/REST gene by methylation and CREB affects the cellular phenotype of small-cell lung cancer. Oncogene 2010, 29, 5828–5838.
  151. Yoshida, M.; Oda, C.; Mishima, K.; Tsuji, I.; Obika, S.; Shimojo, M. An antisense amido-bridged nucleic acid gapmer oligonucleotide targeting SRRM4 alters REST splicing and exhibits anti-tumor effects in small cell lung cancer and prostate cancer cells. Cancer Cell Int. 2023, 23, 8.
  152. Wang, K.; Xu, K.; Leng, X.; Han, Y.; Fang, Q. miRNA-9 Inhibits Proliferation and Migration of Lung Squamous Cell Carcinoma Cells by Regulating NRSF/EGFR. Technol. Cancer Res. Treat. 2020, 19, 1533033820945807.
  153. Lim, J.S.; Ibaseta, A.; Fischer, M.M.; Cancilla, B.; O’Young, G.; Cristea, S.; Luca, V.C.; Yang, D.; Jahchan, N.S.; Hamard, C.; et al. Intratumoural heterogeneity generated by Notch signalling promotes small-cell lung cancer. Nature 2017, 545, 360–364.
  154. Moss, A.C.; Jacobson, G.M.; Walker, L.E.; Blake, N.W.; Marshall, E.; Coulson, J.M. SCG3 transcript in peripheral blood is a prognostic biomarker for REST-deficient small cell lung cancer. Clin. Cancer Res. 2009, 15, 274–283.
More
This entry is offline, you can click here to edit this entry!
ScholarVision Creations