Interplay between Alpha-Synuclein and Epigenetic Modification: History
Please note this is an old version of this entry, which may differ significantly from the current revision.

Alpha-synuclein (αS) is a small, presynaptic neuronal protein encoded by the SNCA gene. Point mutations and gene multiplication of SNCA cause rare familial forms of Parkinson’s disease (PD). Misfolded αS is cytotoxic and is a component of Lewy bodies, which are a pathological hallmark of PD. Because SNCA multiplication is sufficient to cause full-blown PD, gene dosage likely has a strong impact on pathogenesis. In sporadic PD, increased SNCA expression resulting from a minor genetic background and various environmental factors may contribute to pathogenesis in a complementary manner. With respect to genetic background, several risk loci neighboring the SNCA gene have been identified, and epigenetic alterations, such as CpG methylation and regulatory histone marks, are considered important factors. These alterations synergistically upregulate αS expression and some post-translational modifications of αS facilitate its translocation to the nucleus. 

  • alpha-synuclein
  • bioinformatics
  • epigenome
  • organoids

1. Introduction

α-Synuclein (αS) is a relatively small protein with a molecular weight of 14.5 kDa. It consists of 140 amino acids encoded by the SNCA gene on chromosome 4q22.1 [1]. The first synuclein cDNA was isolated from the electric organ of Torpedo californica and was primarily detected in the nuclear envelope and presynaptic nerve terminals [2]. Thus, the name αS is derived from a combination of the prefixes for synapse (“syn”), nucleus (“nucl”), and the suffix for protein (“ein”). αS is considered an important molecule that triggers the neurodegenerative process in synucleinopathies, including Parkinson’s disease (PD) [3]. Most cases of PD are sporadic; however, less than 10% have a family history and several genes associated with the inherited forms of PD have been identified [4]. SNCA was the first familial PD gene (PARK1) discovered, and patients harboring missense mutations exhibit classic adult-onset forms of PD [5]. SNCA can cause PD not only through a point mutation, but also by gene multiplication, with the latter designated as PARK4 [6][7]. These copy number changes correlate with elevated transcript levels of SNCA and subsequent increase of αS protein production [6]. In addition, increased SNCA mRNA was observed predominantly in the midbrain or substantia nigra of patients with PD [8], while a decrease is detected in other tissues and regions, such as temporal and frontal cortexes and cerebrospinal fluid [9]. It is assumed that this tissue-specific increase in SNCA is closely related to neurodegeneration. Thereby, the idea that increased αS expression in the nervous system can cause dopaminergic neuron loss is the rationale for using αS overexpressing cell and animal models for PD studies.
Despite the historical findings of its localization in the nucleus and presynaptic nerve terminals [2], the physiological function of αS has been primarily focused in the synaptic area, where it is involved in regulating synaptic vesicle exocytosis and plasticity [10][11]. Although the majority of αS is localized to the cytoplasm of neurons, expression has been found in other cellular compartments, such as the endoplasmic reticulum [12][13], endosome/lysosomes [14], and mitochondria [15][16]. The transcriptional coactivator PGC-1α is a master regulator of mitochondrial biogenesis and oxidative metabolism [17]. PGC-1α-response genes are significantly associated with PD pathology [18], and interestingly, PGC-1α expression directly influences the oligomerization of αS in cell culture models [19].
Although the pathological processes underlying neurodegenerative diseases are different for every disease, several neurodegenerative molecules are known to function in the nucleus. For example, TAR DNA-binding protein 43 (TDP-43) and fused in sarcoma (FUS), which are associated with amyotrophic lateral sclerosis, are DNA/RNA-binding proteins that contain nuclear localization sequences (NLS) [20][21]. αS does not contain a canonical NLS; however, several studies have indicated that αS is located in the nucleus in the mammalian central nervous system [22]. Moreover, an autopsy revealed that αS-positive glial nuclear and neuronal nuclear inclusions are present in approximately 80% of patients with multiple system atrophy, a type of synucleinopathy [23]. These nuclear inclusions are ubiquitinated, phosphorylated, and composed of fibrillar filaments that are 10–20 nm in diameter [24]. Although it is inconclusive whether this finding is a cause or consequence, the appearance of abnormally modified and structurally altered αS in neural and glial nuclei may be associated with the progression of synucleinopathy, which is a group of neurodegenerative disorders characterized by fibrillary aggregates of αS protein.
The mechanisms by which αS localizes to the nucleus remains elusive. Because of its small size, αS can pass through nuclear pores and does not require transport carriers, such as importin [25]. After internalization into the nucleus, αS may be retained through its interaction with DNA or histones [26][27][28]. Alternatively, several active mechanisms have been suggested to regulate the nuclear translocation of αS, including its interaction with TRIM28 [29] or Ras-related nuclear protein [30]. The function of αS in the nucleus has remained enigmatic, but is becoming clearer with recent progress in epigenetics [31].

2. Interaction between αS and Epigenetic Factors

2.1. Transcriptional Regulation of SNCA

Comprehensive genome-wide association studies (GWASs) have examined the association between SNPs and the development of idiopathic PD. Several risk loci in the SNCA gene have been identified [32]. A meta-analysis revealed that rs356182 is the most significant SNPs associated with PD [32]; however, it is debatable whether the rs356168 risk variant acts directly on SNCA regulation or is involved in a neurodegenerative process unrelated to αS function [33][34]. In addition, the complex polymorphic microsatellite repeat site Rep1, located approximately 10 kb upstream of the SNCA transcription start site, has been reported [35]. Longer Rep1 alleles increase SNCA gene expression compared with shorter alleles [35]; however, the effect of the Rep allele on disease severity and the risk of cognitive decline remains controversial [36][37][38].
In addition to gene mutations, epigenetic regulation is important to the regulation of SNCA expression. DNA methylation occurs on cytosine residues located adjacent to guanine, which are known as CpG sites. The CpG site near the transcription start site, known as the CpG island, and its methylation, represses the transcription of the associated gene. The SNCA CpG island is located in intron 1, which is upstream of the initiation codon [39]. It serves as a binding site for several transcription factors that regulate SNCA expression [40] (Figure 1). Hypomethylation of this region results in increased expression of SNCA, which leads to the accumulation of αS and ultimately, neurotoxicity [41]. In a study of post-mortem brains, a marked reduction in SNCA methylation was observed in the substantia nigra, putamen, and cortex of PD patients [42]. Although PD-related alterations in methylation status in intron 1 is controversial [43][44][45][46][47][48][49][50][51], the induction of robust hypermethylation of SNCA CpG island results in a reduction in SNCA expression, which may be exploited as a therapeutic approach to prevent pathogenic αS accumulation [52]. This intron 1 domain is also associated with histone modification, and another epigenetic modulator, H3K4me3 transcriptional active mark, was more prevalent in the substantia nigra of PD brains [53]. Based on these observations, attempts have been made to regulate epigenetically active histone marks by gene editing in an experimental model of PD. Precisely, a deletion of H3K4me3 by locus-specific editing successfully reduced αS in SH-SY5Y neuronal cells and iPSC-derived dopaminergic neurons [53].
Figure 1. Changes in methylation status of CpG islands located in SNCA intron 1. (A) The indexed number of CpG sites is based on the report by Jowaed et al. [42][43][44][45][46][47][48][49][50][51]. The numbering of CpG sites is not identical in each study and is, therefore, corrected and displayed accordingly. Differences in methylation rates of CpG sites are compared between Parkinson’s disease (PD) or dementia with Lewy bodies (DLB) and healthy subjects or disease controls. The analyses are varied, but do not show consistent results. Nevertheless, CpG numbers 8 and 18 are likely more sensitive to the methylation modulating system or nuclear environment. (B) Schematic presentation of the human SNCA gene. Rep1, a dinucleotide repeat site, the length of which affects SNCA expression, located approximately 10 kb upstream from the start codon in exon 2. The CpG island located in intron 1 includes 23 CpG sites. Primer sequences for pyrosequencing by Jowaed et al. are underlined. The abbreviations used are as follows: PWBCs, peripheral white blood cells; PBMC, peripheral blood mononuclear cells.

2.2. Effect of αS Nuclear Localization

As noted above, αS does not have a canonical NLS; but, post-translational modifications and other factors affect its nuclear localization. In neuronal inclusion bodies (i.e., Lewy bodies) of PD brains, greater than 90% of the αS is highly phosphorylated at serine 129 (S129), which is used as a marker for pathological diagnosis. In vivo experiments using mouse brain and primary cortical neurons have revealed that S129 phosphorylated αS is rapidly translocated to the nucleus of laser-induced focal lesions [54]. Similarly, in H4 human neuroglioma cells, S129 phosphorylated αS exhibited higher affinity toward the nucleus; whereas, downregulated cyclin B1 and E2F transcription factor 8 genes were involved in the cell cycle [55]. Polo-like kinase 2 (PLK2), an enzyme that catalyzes the S129 phosphorylation of αS, may be involved in the nuclear-cytoplasmic shuttling of αS [56]. The presence of S129 phospho-αS may be associated with aging, because S129 phosphorylated αS was observed in the nucleus of aged mice, but not in young A30P human αS transgenic mice [57].
Besides phosphorylation, SUMOylation, which is generated by mature small ubiquitin-related modifiers (SUMO), may occur during the nuclear translocation of αS [58]. Interestingly, distinct αS species translocate from the nucleus to neuronal processes during neuronal differentiation, which suggests that the maturation process of the nervous system may affect subcellular localization [59]. With respect to protein conformational change and aggregation, exposure to human fibrotic αS seed facilitates the formation of intranuclear inclusions in mouse primary cortical neurons [60]. Furthermore, the inoculation of αS preformed fibrils into the stomach wall of wild-type mice resulted in the formation of a small number of nuclear inclusions in the dorsal motor nucleus of the vagal nerve [61]. These results suggest that the progression of PD may contribute to the aberrant accumulation of misfolded αS in neuronal nuclei.

2.3. Interplay between Alpha-Synuclein and DNA

2.3.1. Direct Binding to DNA

αS can bind to supercoiled DNA in a conformation-specific manner and alter the bending properties and stability of DNA, which in turn modulates the conformation status of αS itself [62][63][64]. αS also directly binds to a large subset of DNA promoter sequences, including the PPARGC1A (PGC-1α) and NOTCH1 gene promoters, thereby downregulating the transcription of their target genes [55][65][66]. Interestingly, αS colocalizes with DNA damage response components to form discrete foci in the neuronal nuclei and the removal of αS decreases the repair of DNA double-strand breaks [54], suggesting a neuroprotective function of endogenous αS.

2.3.2. Interaction with DNA-Modifying Proteins

DNA methylation, which is an important epigenetic process, is catalyzed by multiple DNA methyltransferases (DNMTs: DNMT1, DNMT2, DNMT3, DNMT3L) through the transfer of a methyl group from S-adenyl methionine (SAM) to the fifth carbon of cytosine, resulting in 5mC [67]. During DNA replication, DNMT1 retrieves the DNA methylation prototype from the parent DNA strand and transfers it to the newly synthesized daughter strand [68]. Conversely, DNA demethylation is often catalyzed by enzymes of the ten-eleven translocation (TET) family, which counteract the activity of DNMTs [69]. In recent years, DNA methylation has been a significant area of interest in PD research [70]. Given that CpG methylation profiles associated with PD matched approximately 30% between brain tissue and blood samples [71], specific loci could be identified as candidate biomarkers for PD in peripheral blood mononuclear cells [72]. An early study demonstrated that DNMT1, normally located in the nucleus, was sequestered in the cytoplasm following αS overexpression, causing global DNA hypomethylation and transcriptional activation of downstream genes [73]. A subsequent study also found that altered DNA methylation at CpG sites affected gene expression associated with locomotor behavior and the glutamate signaling pathway [74].

2.4. Interplay between Alpha-Synuclein and Histones

A nucleosome is a fundamental unit of chromatin consisting of 147 base pairs of DNA and an octamer of core histone proteins containing two copies of each of the histones: H2A, H2B, H3, and H4 [75]. The amino acid sequences of the histone proteins are conserved across species from Archaeum to Homo sapiens [76]. The chromosome structure of eukaryotic cells may be divided into two regions: heterochromatin, which is tightly condensed and transcriptionally repressed, and euchromatin, which is untangled and transcriptionally active [76]. The condensation of histones is essentially responsible for the organization of euchromatin and heterochromatin. The chromatin status is defined by the types of post-translational modifications in histone tails, including acetylation, dopaminylation, methylation, phosphorylation, serotonylation, SUMOylation, and ubiquitination [77][78][79]. These histone modifications may alter the affinity to DNA and other histones by altering the surface charge; thereby, reversibly regulating the entry of transcription factors and other transcription-related proteins [76].

2.4.1. Histone Modification via Acetylation

αS binds to the N-terminal flexible tails of histones H3, H4, and H1 [80]. Its fibrillation is accelerated by H1 released from the nucleus during apoptosis [81]. Lysine acetylation is a reversible process, which is post-translational modification that alters the charge of lysine residues and modifies protein structure to influence protein function [82]. Histone acetylation modulates fundamental cell processes, such as transcriptional regulation and chromatin remodeling. The balance between the activities of lysine acetyltransferases (KAT) and histone deacetylases (HDAC) is strictly controlled, but can be disrupted in neurodegenerative diseases, such as PD [83].
KAT catalyzes the transfer of an acetyl group from acetyl-CoA to the ε-amino group of an internal lysine residue [84]. Acetylated histones counteract the positive charge on the residue, which prevents DNA-histone interactions and activates transcription, resulting in a loose chromatin state that facilitates transcription. In contrast, histone deacetylation results in a tighter chromatin structure, which suppresses transcriptional activity [84]. Mammalian KATs are classified into two groups according to their cellular localization: Nuclear KATs (type A) and cytoplasmic KATs (type B). Type A KATs are primarily involved in transcriptional regulation and may be further classified into five families: Gcn5-related N-acetyltransferase (GNAT), p300/cAMP response element binding (CREB) binding protein (CBP), MYST basal transcription factors, and the nuclear receptor coactivator (NCoA) family [82]. The reverse reaction, deacetylation, is catalyzed by histone deacetylases. To date, 18 mammalian HDACs have been identified and classified into four classes (class I, II, III, and IV) based on their sequence similarity to yeast HDACs. Class I HDACs include HDAC1, 2, 3, and 8; Class II HDACs are subdivided into class IIa (HDAC4, 5, 7, and 9) and IIb (HDAC6 and 10); Class III HDACs are members of the sirtuin family; and class IV HDAC includes only HDAC11 [85]. Of note, nucleosome proteins are not a specific substrate for these KATs and HDACs [86]. For example, class IIa HDACs (HDAC4, HDAC5, and HDAC7) shuttle between the nucleus and the cytoplasm [87]. HDAC4 recognizes a variety of extra-nuclear proteins as substrates, including forkhead transcription factors of the O class (FOXO), myosin heavy chain isoforms (MyHC), PGC-1α, and heat shock cognate 71 kDa (Hsc70) [88].
In general, histone acetylation is associated with gene activation; whereas, the removal of the acetyl mark induces a closed chromatin structure. Several studies suggest that αS reduces histone acetylation, which inhibits the expression of certain genes (Figure 2). Although the intracellular function of αS in neurodegenerative processes remains unclear, nuclear-localized αS increases cytotoxicity [27] whereas cytosolic αS is neuroprotective [89]. Cytosolic αS reduces p300 levels and its KAT activity, resulting in a reduction of histone acetylation in dopaminergic neural cell lines [89]. Alternatively, the A53T mutant αS modulates histone acetylation by interacting with transcriptional adapter 2-α (TADA2a), a component of the major histone acetyltransferase p300/CBP [90].
Figure 2. Schematic illustration of post-translational modifications of histone proteins affected by alpha-synuclein (αS). Lysine methylation has different functions depending on its residues. Methylation (Met) of H3 lysine 9 (K9), lysine 27 (K27), and H4 lysine 20 (K20) is often associated with transcriptional repression. In contrast, the methylation of H3 lysine 4 (K4), lysine 36 (K36), or lysine 79 (K79) is largely responsible for transcriptional activation. In addition, histone acetylation (Ac) usually promotes gene expression. Combined stimulation with αS and retinoic acid (RA) enhances K9 methylation of H3 through the activation of euchromatic histone lysine N-methyltransferase 2 (EHMT2). In addition, the inactivation of lysine acetyltransferases (KATs) decreases histone acetylation. The other transcriptional active mark, K36 methylation of H3, is also downregulated by αS. Both increased repressive signals and disruption of active marks result in the transcriptional repression of downstream genes.
Class IIa HDACs have NLS and shuttle between the nucleus and cytoplasm [91]. HDAC4, a member of the class IIa HDACs, is abundantly expressed in neurons and accumulates in the nucleus following stimulation with MPTP (1-methyl-4-phenyl-1, 2, 3, 6-tetrahydropyridines), a dopaminergic neurotoxin. Nuclear HDAC4 also mediates cell death in A53T mutant αS-expressing cells by inhibiting CREB and myocyte enhancer factor 2A (MEF2A) [92]. The global HDAC inhibitors, sodium butyrate (NaB) and suberoylanilide hydroxamic acid (SAHA), protect against αS-mediated neurotoxicity in cell and transgenic Drosophila models [27]. Decreased H3 acetylation and altered RNA-seq gene expression profiles through αS in LUHMES (Lund human mesencephalic) dopaminergic cells were attenuated by adding NaB, which may be mediated by DNA repair [93]. In contrast, the neuroprotective effects of NaB in PC12 cells are dependent on the activation of PGC-1α via hyperacetylation of its promoter region [94]. Class I HDAC-specific inhibitors and class IIb HDAC6 inhibitors failed to alleviate the αS-induced neurite outgrowth defects in SH-SY5Y cells; whereas, the class IIa HDAC4/5 inhibitor LMK235 successfully promoted neurite outgrowth [95]. The sirtuin family of NAD(+)-dependent class III HDACs are also candidate therapeutic targets in PD [96]. Blockade of sirtuin 2 resulted in a dose-dependent protective effect against αS-induced toxicity [97][98], suggesting the potential therapeutic applications of targeting specific HDACs. The neuroprotective effects of HDAC inhibitors are currently under preclinical investigation as disease-modifying therapy for PD [99]; however, some of their effects may be mediated by mechanisms unrelated to histone acetylation, such as microtubule stabilization [100].

2.4.2. Histone Modification via Methylation

Lysine residues can accept mono-, di-, and tri-methylation (me1, me2, and me3) modifications; whereas, arginine residues accept asymmetric or symmetric di-methylation or mono-methylation. Although histone acetylation usually promotes gene expression, the function of histone methylation depends on the context. Methylation of histone H3 on lysine 4 (H3K4), lysine 36 (H3K36), or lysine 79 (H3K79) is largely responsible for transcriptional activation. In contrast, methylation of histone H3 at lysine 9 (H3K9) and lysine 27 (H3K27) or histone H4 on lysine 20 (H4K20) is often associated with transcriptional repression [101]. Lysine methylation is catalyzed by lysine methyltransferases (KMTs), known as “Writers,”; whereas, histone demethylation is catalyzed by lysine demethylases (KDMs), known as “Erasers”.
Protein arginine methyltransferases (PRMTs) transfer methyl groups from S-Adenosylmethionine (SAM) to arginine residues on histone proteins. They are classified into three types based on their catalytic activity [102]. Type I PRMTs (PRMT1, 2, 3, 4, 6, and 8) asymmetrically di-methylate arginine residues (ADMA) whereas type II PRMTs (PRMT5 and PRMT9) symmetrically di-methylate arginine residues (SDMA). Type III (PRMT7) catalyzes only mono-methylated arginine formation (MMA) [102]. The major targets of arginine methylation in histone proteins are histone H3 on arginine 2 (H3R2), arginine 8 (H3R8), and histone H4 on arginine 3 (H4R3). The histone code of arginine methylation is rather complicated. Both ADMA and SDMA are di-methylated, but the asymmetric type, ADMA, is often associated with transcriptional activation, whereas the symmetric type, SDMA, results in transcriptional repression [103]. Therefore, the structural behavior of chromatin differs depending on the type of PRMT acting on the target histone tails. Consequently, SDMA occupying regional histones corresponds to the signature of type II PRMTs.
αS has been found to interact with several epigenetic writers. In an αS yeast model, the altered histone marks including H3K36 di-methylation were distinct from histone marks affected by TDP-43 or FUS [104]. Additionally, overexpression of αS in transgenic Drosophila and SH-SY5Y cells resulted in H3K9 di-methylation through upregulation of euchromatic histone lysine N-methyltransferase 2 (EHMT2) in the presence of retinoic acid [105]. The chromatin immunoprecipitation with antibodies against repressor element-1 (RE1)-silencing transcription factor (REST) inactivated transcription, one of the EHMT2 interacting protein, revealed the repressed downstream genes SNAP25 and L1CAM. SNAP25 is a major component of SNARE complex involved in synaptic function; thereby, these changes may contribute to the synaptic dysfunction that occurs in PD brain [106].

2.4.3. SWI/SNF Chromatin Remodeling Complexes

The SWI/SNF family was first functionally characterized in Saccharomyces cerevisiae but is conserved across species throughout eukaryotes. The SWI/SNF family contains an ATPase subunit that utilizes ATP-dependent chromatin remodeling to enhance DNA accessibility during transcription [107]. BAF (Brg1-associated factors) is a mammalian homolog of SWI/SNF, whose central ATPase is composed of SWI/SNF-related, matrix-associated, actin-dependent regulator of chromatin, subfamily A, member 4 (SMARCA4), and SMARCA2. The BAF complexes are not just chromatin remodeling factors, but they can repress or activate gene expression [108]. Interestingly, SMARCA4 harbors SNPs and a meta-analysis yielded p-values of 1 × 10−4 and 0.05, which may be considered a potential risk for PD, although not in the top 10,000 most significant GWAS results [32][109]. Furthermore, in silico disease-associated gene prediction followed by in vivo Drosophila genetic screening identified SMARCA4. Knockdown of Brahma, the Drosophila homolog of SMARCA4, in dopaminergic neurons prolonged the lifespan of human LRRK2 or SNCA transgenic Drosophila [109]. Another component of the BAF complex, Brg-associated factor 57 (BAF57), was modulated in PC12 cells treated with the dopaminergic neurotoxin, 6-OHDA [110].

This entry is adapted from the peer-reviewed paper 10.3390/ijms24076645

References

  1. Kawahata, I.; Finkelstein, D.I.; Fukunaga, K. Pathogenic Impact of alpha-Synuclein Phosphorylation and Its Kinases in alpha-Synucleinopathies. Int. J. Mol. Sci. 2022, 23, 6216.
  2. Maroteaux, L.; Campanelli, J.T.; Scheller, R.H. Synuclein: A neuron-specific protein localized to the nucleus and presynaptic nerve terminal. J. Neurosci. 1988, 8, 2804–2815.
  3. Sulzer, D.; Edwards, R.H. The physiological role of alpha-synuclein and its relationship to Parkinson’s Disease. J. Neurochem. 2019, 150, 475–486.
  4. Thomas, B.; Beal, M.F. Parkinson’s disease. Hum. Mol. Genet. 2007, 16, R183–R194.
  5. Day, J.O.; Mullin, S. The Genetics of Parkinson’s Disease and Implications for Clinical Practice. Genes 2021, 12, 1006.
  6. Singleton, A.B.; Farrer, M.; Johnson, J.; Singleton, A.; Hague, S.; Kachergus, J.; Hulihan, M.; Peuralinna, T.; Dutra, A.; Nussbaum, R.; et al. alpha-Synuclein locus triplication causes Parkinson’s disease. Science 2003, 302, 841.
  7. Farrer, M.; Kachergus, J.; Forno, L.; Lincoln, S.; Wang, D.S.; Hulihan, M.; Maraganore, D.; Gwinn-Hardy, K.; Wszolek, Z.; Dickson, D.; et al. Comparison of kindreds with parkinsonism and alpha-synuclein genomic multiplications. Ann. Neurol. 2004, 55, 174–179.
  8. Chiba-Falek, O.; Lopez, G.J.; Nussbaum, R.L. Levels of alpha-synuclein mRNA in sporadic Parkinson disease patients. Mov. Disord. 2006, 21, 1703–1708.
  9. Kang, J.H.; Irwin, D.J.; Chen-Plotkin, A.S.; Siderowf, A.; Caspell, C.; Coffey, C.S.; Waligorska, T.; Taylor, P.; Pan, S.; Frasier, M.; et al. Association of cerebrospinal fluid beta-amyloid 1-42, T-tau, P-tau181, and alpha-synuclein levels with clinical features of drug-naive patients with early Parkinson disease. JAMA Neurol. 2013, 70, 1277–1287.
  10. Somayaji, M.; Cataldi, S.; Choi, S.J.; Edwards, R.H.; Mosharov, E.V.; Sulzer, D. A dual role for alpha-synuclein in facilitation and depression of dopamine release from substantia nigra neurons in vivo. Proc. Natl. Acad. Sci. USA 2020, 117, 32701–32710.
  11. Burre, J.; Sharma, M.; Tsetsenis, T.; Buchman, V.; Etherton, M.R.; Sudhof, T.C. Alpha-synuclein promotes SNARE-complex assembly in vivo and in vitro. Science 2010, 329, 1663–1667.
  12. Colla, E.; Jensen, P.H.; Pletnikova, O.; Troncoso, J.C.; Glabe, C.; Lee, M.K. Accumulation of toxic alpha-synuclein oligomer within endoplasmic reticulum occurs in alpha-synucleinopathy in vivo. J. Neurosci. 2012, 32, 3301–3305.
  13. Sugeno, N.; Takeda, A.; Hasegawa, T.; Kobayashi, M.; Kikuchi, A.; Mori, F.; Wakabayashi, K.; Itoyama, Y. Serine 129 phosphorylation of alpha-synuclein induces unfolded protein response-mediated cell death. J. Biol. Chem. 2008, 283, 23179–23188.
  14. Sugeno, N.; Hasegawa, T.; Tanaka, N.; Fukuda, M.; Wakabayashi, K.; Oshima, R.; Konno, M.; Miura, E.; Kikuchi, A.; Baba, T.; et al. Lys-63-linked ubiquitination by E3 ubiquitin ligase Nedd4-1 facilitates endosomal sequestration of internalized alpha-synuclein. J. Biol. Chem. 2014, 289, 18137–18151.
  15. Chinta, S.J.; Mallajosyula, J.K.; Rane, A.; Andersen, J.K. Mitochondrial alpha-synuclein accumulation impairs complex I function in dopaminergic neurons and results in increased mitophagy in vivo. Neurosci. Lett. 2010, 486, 235–239.
  16. Devi, L.; Raghavendran, V.; Prabhu, B.M.; Avadhani, N.G.; Anandatheerthavarada, H.K. Mitochondrial import and accumulation of alpha-synuclein impair complex I in human dopaminergic neuronal cultures and Parkinson disease brain. J. Biol. Chem. 2008, 283, 9089–9100.
  17. Lin, J.; Handschin, C.; Spiegelman, B.M. Metabolic control through the PGC-1 family of transcription coactivators. Cell Metab. 2005, 1, 361–370.
  18. Zheng, B.; Liao, Z.; Locascio, J.J.; Lesniak, K.A.; Roderick, S.S.; Watt, M.L.; Eklund, A.C.; Zhang-James, Y.; Kim, P.D.; Hauser, M.A.; et al. PGC-1alpha, a potential therapeutic target for early intervention in Parkinson’s disease. Sci. Transl. Med. 2010, 2, 52ra73.
  19. Eschbach, J.; von Einem, B.; Muller, K.; Bayer, H.; Scheffold, A.; Morrison, B.E.; Rudolph, K.L.; Thal, D.R.; Witting, A.; Weydt, P.; et al. Mutual exacerbation of peroxisome proliferator-activated receptor gamma coactivator 1alpha deregulation and alpha-synuclein oligomerization. Ann. Neurol. 2015, 77, 15–32.
  20. Doll, S.G.; Meshkin, H.; Bryer, A.J.; Li, F.; Ko, Y.H.; Lokareddy, R.K.; Gillilan, R.E.; Gupta, K.; Perilla, J.R.; Cingolani, G. Recognition of the TDP-43 nuclear localization signal by importin alpha1/beta. Cell Rep. 2022, 39, 111007.
  21. Yoshizawa, T.; Ali, R.; Jiou, J.; Fung, H.Y.J.; Burke, K.A.; Kim, S.J.; Lin, Y.; Peeples, W.B.; Saltzberg, D.; Soniat, M.; et al. Nuclear Import Receptor Inhibits Phase Separation of FUS through Binding to Multiple Sites. Cell 2018, 173, 693–705.e22.
  22. Geertsma, H.M.; Suk, T.R.; Ricke, K.M.; Horsthuis, K.; Parmasad, J.A.; Fisk, Z.A.; Callaghan, S.M.; Rousseaux, M.W.C. Constitutive nuclear accumulation of endogenous alpha-synuclein in mice causes motor impairment and cortical dysfunction, independent of protein aggregation. Hum. Mol. Genet. 2022, 31, 3613–3628.
  23. Nishie, M.; Mori, F.; Yoshimoto, M.; Takahashi, H.; Wakabayashi, K. A quantitative investigation of neuronal cytoplasmic and intranuclear inclusions in the pontine and inferior olivary nuclei in multiple system atrophy. Neuropathol. Appl. Neurobiol. 2004, 30, 546–554.
  24. Weston, L.J.; Bowman, A.M.; Osterberg, V.R.; Meshul, C.K.; Woltjer, R.L.; Unni, V.K. Aggregated Alpha-Synuclein Inclusions within the Nucleus Predict Impending Neuronal Cell Death in a Mouse Model of Parkinsonism. Int. J. Mol. Sci. 2022, 23, 15294.
  25. Timney, B.L.; Raveh, B.; Mironska, R.; Trivedi, J.M.; Kim, S.J.; Russel, D.; Wente, S.R.; Sali, A.; Rout, M.P. Simple rules for passive diffusion through the nuclear pore complex. J. Cell Biol. 2016, 215, 57–76.
  26. Goers, J.; Manning-Bog, A.B.; McCormack, A.L.; Millett, I.S.; Doniach, S.; Di Monte, D.A.; Uversky, V.N.; Fink, A.L. Nuclear localization of alpha-synuclein and its interaction with histones. Biochemistry 2003, 42, 8465–8471.
  27. Kontopoulos, E.; Parvin, J.D.; Feany, M.B. Alpha-synuclein acts in the nucleus to inhibit histone acetylation and promote neurotoxicity. Hum. Mol. Genet. 2006, 15, 3012–3023.
  28. Surguchov, A. Protein-DNA interaction: One step closer to understanding the mechanism of neurodegeneration. J. Neurosci. Res. 2019, 97, 391–392.
  29. Rousseaux, M.W.; de Haro, M.; Lasagna-Reeves, C.A.; De Maio, A.; Park, J.; Jafar-Nejad, P.; Al-Ramahi, I.; Sharma, A.; See, L.; Lu, N.; et al. TRIM28 regulates the nuclear accumulation and toxicity of both alpha-synuclein and tau. Elife 2016, 5, e19809.
  30. Chen, V.; Moncalvo, M.; Tringali, D.; Tagliafierro, L.; Shriskanda, A.; Ilich, E.; Dong, W.; Kantor, B.; Chiba-Falek, O. The mechanistic role of alpha-synuclein in the nucleus: Impaired nuclear function caused by familial Parkinson’s disease SNCA mutations. Hum. Mol. Genet. 2020, 29, 3107–3121.
  31. Surguchov, A. alpha-Synuclein and Mechanisms of Epigenetic Regulation. Brain. Sci. 2023, 13, 150.
  32. Nalls, M.A.; Blauwendraat, C.; Vallerga, C.L.; Heilbron, K.; Bandres-Ciga, S.; Chang, D.; Tan, M.; Kia, D.A.; Noyce, A.J.; Xue, A.; et al. Identification of novel risk loci, causal insights, and heritable risk for Parkinson’s disease: A meta-analysis of genome-wide association studies. Lancet Neurol. 2019, 18, 1091–1102.
  33. Prahl, J.D.; Pierce, S.E.; van der Schans, E.J.C.; Coetzee, G.A.; Tyson, T. The Parkinson’s disease variant rs356182 regulates neuronal differentiation independently from alpha-synuclein. Hum. Mol. Genet. 2023, 32, 1–14.
  34. Pihlstrom, L.; Blauwendraat, C.; Cappelletti, C.; Berge-Seidl, V.; Langmyhr, M.; Henriksen, S.P.; van de Berg, W.D.J.; Gibbs, J.R.; Cookson, M.R.; The International Parkinson Disease Genomics Consortium; et al. A comprehensive analysis of SNCA-related genetic risk in sporadic parkinson disease. Ann. Neurol. 2018, 84, 117–129.
  35. Chiba-Falek, O.; Nussbaum, R.L. Effect of allelic variation at the NACP-Rep1 repeat upstream of the alpha-synuclein gene (SNCA) on transcription in a cell culture luciferase reporter system. Hum. Mol. Genet. 2001, 10, 3101–3109.
  36. Ng, A.S.L.; Tan, Y.J.; Zhao, Y.; Saffari, S.E.; Lu, Z.; Ng, E.Y.L.; Ng, S.Y.E.; Chia, N.S.Y.; Setiawan, F.; Xu, Z.; et al. SNCA Rep1 promoter variability influences cognition in Parkinson’s disease. Mov. Disord. 2019, 34, 1232–1236.
  37. Pedersen, C.C.; Lange, J.; Forland, M.G.G.; Macleod, A.D.; Alves, G.; Maple-Grodem, J. A systematic review of associations between common SNCA variants and clinical heterogeneity in Parkinson’s disease. NPJ Park. Dis. 2021, 7, 54.
  38. Markopoulou, K.; Biernacka, J.M.; Armasu, S.M.; Anderson, K.J.; Ahlskog, J.E.; Chase, B.A.; Chung, S.J.; Cunningham, J.M.; Farrer, M.; Frigerio, R.; et al. Does alpha-synuclein have a dual and opposing effect in preclinical vs. clinical Parkinson’s disease? Park. Relat. Disord. 2014, 20, 584–589; Discussion in 584.
  39. de Boni, L.; Riedel, L.; Schmitt, I.; Kraus, T.F.J.; Kaut, O.; Piston, D.; Akbarian, S.; Wullner, U. DNA methylation levels of alpha-synuclein intron 1 in the aging brain. Neurobiol. Aging 2015, 36, 3334.e7–3334.e11.
  40. Miranda-Morales, E.; Meier, K.; Sandoval-Carrillo, A.; Salas-Pacheco, J.; Vazquez-Cardenas, P.; Arias-Carrion, O. Implications of DNA Methylation in Parkinson’s Disease. Front. Mol. Neurosci. 2017, 10, 225.
  41. Wullner, U.; Kaut, O.; deBoni, L.; Piston, D.; Schmitt, I. DNA methylation in Parkinson’s disease. J. Neurochem. 2016, 139 (Suppl. S1), 108–120.
  42. Jowaed, A.; Schmitt, I.; Kaut, O.; Wullner, U. Methylation regulates alpha-synuclein expression and is decreased in Parkinson’s disease patients’ brains. J. Neurosci. 2010, 30, 6355–6359.
  43. Matsumoto, L.; Takuma, H.; Tamaoka, A.; Kurisaki, H.; Date, H.; Tsuji, S.; Iwata, A. CpG demethylation enhances alpha-synuclein expression and affects the pathogenesis of Parkinson’s disease. PLoS ONE 2010, 5, e15522.
  44. de Boni, L.; Tierling, S.; Roeber, S.; Walter, J.; Giese, A.; Kretzschmar, H.A. Next-generation sequencing reveals regional differences of the alpha-synuclein methylation state independent of Lewy body disease. Neuromol. Med. 2011, 13, 310–320.
  45. Richter, J.; Appenzeller, S.; Ammerpohl, O.; Deuschl, G.; Paschen, S.; Bruggemann, N.; Klein, C.; Kuhlenbaumer, G. No evidence for differential methylation of alpha-synuclein in leukocyte DNA of Parkinson’s disease patients. Mov. Disord. 2012, 27, 590–591.
  46. Ai, S.X.; Xu, Q.; Hu, Y.C.; Song, C.Y.; Guo, J.F.; Shen, L.; Wang, C.R.; Yu, R.L.; Yan, X.X.; Tang, B.S. Hypomethylation of SNCA in blood of patients with sporadic Parkinson’s disease. J. Neurol. Sci. 2014, 337, 123–128.
  47. Tan, Y.Y.; Wu, L.; Zhao, Z.B.; Wang, Y.; Xiao, Q.; Liu, J.; Wang, G.; Ma, J.F.; Chen, S.D. Methylation of alpha-synuclein and leucine-rich repeat kinase 2 in leukocyte DNA of Parkinson’s disease patients. Park. Relat. Disord. 2014, 20, 308–313.
  48. Funahashi, Y.; Yoshino, Y.; Yamazaki, K.; Mori, Y.; Mori, T.; Ozaki, Y.; Sao, T.; Ochi, S.; Iga, J.I.; Ueno, S.I. DNA methylation changes at SNCA intron 1 in patients with dementia with Lewy bodies. Psychiatry Clin. Neurosci. 2017, 71, 28–35.
  49. Guhathakurta, S.; Evangelista, B.A.; Ghosh, S.; Basu, S.; Kim, Y.S. Hypomethylation of intron1 of alpha-synuclein gene does not correlate with Parkinson’s disease. Mol. Brain 2017, 10, 6.
  50. Gu, J.; Barrera, J.; Yun, Y.; Murphy, S.K.; Beach, T.G.; Woltjer, R.L.; Serrano, G.E.; Kantor, B.; Chiba-Falek, O. Cell-Type Specific Changes in DNA Methylation of SNCA Intron 1 in Synucleinopathy Brains. Front. Neurosci. 2021, 15, 652226.
  51. Bakhit, Y.; Schmitt, I.; Hamed, A.; Ibrahim, E.A.A.; Mohamed, I.N.; El-Sadig, S.M.; Elseed, M.A.; Alebeed, M.A.; Shaheen, M.T.; Ibrahim, M.O.; et al. Methylation of alpha-synuclein in a Sudanese cohort. Park. Relat. Disord. 2022, 101, 6–8.
  52. Kantor, B.; Tagliafierro, L.; Gu, J.; Zamora, M.E.; Ilich, E.; Grenier, C.; Huang, Z.Y.; Murphy, S.; Chiba-Falek, O. Downregulation of SNCA Expression by Targeted Editing of DNA Methylation: A Potential Strategy for Precision Therapy in PD. Mol. Ther. 2018, 26, 2638–2649.
  53. Guhathakurta, S.; Kim, J.; Adams, L.; Basu, S.; Song, M.K.; Adler, E.; Je, G.; Fiadeiro, M.B.; Kim, Y.S. Targeted attenuation of elevated histone marks at SNCA alleviates alpha-synuclein in Parkinson’s disease. EMBO Mol. Med. 2021, 13, e12188.
  54. Schaser, A.J.; Osterberg, V.R.; Dent, S.E.; Stackhouse, T.L.; Wakeham, C.M.; Boutros, S.W.; Weston, L.J.; Owen, N.; Weissman, T.A.; Luna, E.; et al. Alpha-synuclein is a DNA binding protein that modulates DNA repair with implications for Lewy body disorders. Sci. Rep. 2019, 9, 10919.
  55. Pinho, R.; Paiva, I.; Jercic, K.G.; Fonseca-Ornelas, L.; Gerhardt, E.; Fahlbusch, C.; Garcia-Esparcia, P.; Kerimoglu, C.; Pavlou, M.A.S.; Villar-Pique, A.; et al. Nuclear localization and phosphorylation modulate pathological effects of alpha-synuclein. Hum. Mol. Genet. 2019, 28, 31–50.
  56. Weston, L.J.; Stackhouse, T.L.; Spinelli, K.J.; Boutros, S.W.; Rose, E.P.; Osterberg, V.R.; Luk, K.C.; Raber, J.; Weissman, T.A.; Unni, V.K. Genetic deletion of Polo-like kinase 2 reduces alpha-synuclein serine-129 phosphorylation in presynaptic terminals but not Lewy bodies. J. Biol. Chem. 2021, 296, 100273.
  57. Schell, H.; Hasegawa, T.; Neumann, M.; Kahle, P.J. Nuclear and neuritic distribution of serine-129 phosphorylated alpha-synuclein in transgenic mice. Neuroscience 2009, 160, 796–804.
  58. Ryu, S.; Baek, I.; Liew, H. Sumoylated alpha-synuclein translocates into the nucleus by karyopherin 6. Mol. Cell Toxicol. 2019, 15, 103–109.
  59. Pieger, K.; Schmitt, V.; Gauer, C.; Giessl, N.; Prots, I.; Winner, B.; Winkler, J.; Brandstatter, J.H.; Xiang, W. Translocation of Distinct Alpha Synuclein Species from the Nucleus to Neuronal Processes during Neuronal Differentiation. Biomolecules 2022, 12, 1108.
  60. De Giorgi, F.; Abdul-Shukkoor, M.B.; Kashyrina, M.; Largitte, L.A.; De Nuccio, F.; Kauffmann, B.; Lends, A.; Laferriere, F.; Bonhommeau, S.; Lofrumento, D.D.; et al. Neurons with Cat’s Eyes: A Synthetic Strain of alpha-Synuclein Fibrils Seeding Neuronal Intranuclear Inclusions. Biomolecules 2022, 12, 436.
  61. Uemura, N.; Yagi, H.; Uemura, M.T.; Hatanaka, Y.; Yamakado, H.; Takahashi, R. Inoculation of alpha-synuclein preformed fibrils into the mouse gastrointestinal tract induces Lewy body-like aggregates in the brainstem via the vagus nerve. Mol. Neurodegener. 2018, 13, 21.
  62. Vasudevaraju, P.; Guerrero, E.; Hegde, M.L.; Collen, T.B.; Britton, G.B.; Rao, K.S. New evidence on alpha-synuclein and Tau binding to conformation and sequence specific GC* rich DNA: Relevance to neurological disorders. J. Pharm. Bioallied Sci. 2012, 4, 112–117.
  63. Hegde, M.L.; Vasudevaraju, P.; Rao, K.J. DNA induced folding/fibrillation of alpha-synuclein: New insights in Parkinson’s disease. Front. Biosci. Landmark Ed. 2010, 15, 418–436.
  64. Dent, S.E.; King, D.P.; Osterberg, V.R.; Adams, E.K.; Mackiewicz, M.R.; Weissman, T.A.; Unni, V.K. Phosphorylation of the aggregate-forming protein alpha-synuclein on serine-129 inhibits its DNA-bending properties. J. Biol. Chem. 2022, 298, 101552.
  65. Siddiqui, A.; Chinta, S.J.; Mallajosyula, J.K.; Rajagopolan, S.; Hanson, I.; Rane, A.; Melov, S.; Andersen, J.K. Selective binding of nuclear alpha-synuclein to the PGC1alpha promoter under conditions of oxidative stress may contribute to losses in mitochondrial function: Implications for Parkinson’s disease. Free Radic. Biol. Med. 2012, 53, 993–1003.
  66. Desplats, P.; Spencer, B.; Crews, L.; Pathel, P.; Morvinski-Friedmann, D.; Kosberg, K.; Roberts, S.; Patrick, C.; Winner, B.; Winkler, J.; et al. alpha-Synuclein induces alterations in adult neurogenesis in Parkinson disease models via p53-mediated repression of Notch1. J. Biol. Chem. 2012, 287, 31691–31702.
  67. Lyko, F. The DNA methyltransferase family: A versatile toolkit for epigenetic regulation. Nat. Rev. Genet. 2018, 19, 81–92.
  68. Kaur, G.; Rathod, S.S.S.; Ghoneim, M.M.; Alshehri, S.; Ahmad, J.; Mishra, A.; Alhakamy, N.A. DNA Methylation: A Promising Approach in Management of Alzheimer’s Disease and Other Neurodegenerative Disorders. Biology 2022, 11, 90.
  69. Wu, X.; Zhang, Y. TET-mediated active DNA demethylation: Mechanism, function and beyond. Nat. Rev. Genet. 2017, 18, 517–534.
  70. Li, D.; Liang, J.; Guo, W.; Zhang, Y.; Wu, X.; Zhang, W. Integrative analysis of DNA methylation and gene expression data for the diagnosis and underlying mechanism of Parkinson’s disease. Front. Aging Neurosci. 2022, 14, 971528.
  71. Masliah, E.; Dumaop, W.; Galasko, D.; Desplats, P. Distinctive patterns of DNA methylation associated with Parkinson disease: Identification of concordant epigenetic changes in brain and peripheral blood leukocytes. Epigenetics 2013, 8, 1030–1038.
  72. Somayaji, M.; Lanseur, Z.; Choi, S.J.; Sulzer, D.; Mosharov, E.V. Roles for alpha-Synuclein in Gene Expression. Genes 2021, 12, 1166.
  73. Desplats, P.; Spencer, B.; Coffee, E.; Patel, P.; Michael, S.; Patrick, C.; Adame, A.; Rockenstein, E.; Masliah, E. Alpha-synuclein sequesters Dnmt1 from the nucleus: A novel mechanism for epigenetic alterations in Lewy body diseases. J. Biol. Chem. 2011, 286, 9031–9037.
  74. Schaffner, S.L.; Wassouf, Z.; Lazaro, D.F.; Xylaki, M.; Gladish, N.; Lin, D.T.S.; MacIsaac, J.; Ramadori, K.; Hentrich, T.; Schulze-Hentrich, J.M.; et al. Alpha-synuclein overexpression induces epigenomic dysregulation of glutamate signaling and locomotor pathways. Hum. Mol. Genet. 2022, 31, 3694–3714.
  75. Luger, K.; Dechassa, M.L.; Tremethick, D.J. New insights into nucleosome and chromatin structure: An ordered state or a disordered affair? Nat. Rev. Mol. Cell Biol. 2012, 13, 436–447.
  76. Park, J.; Lee, K.; Kim, K.; Yi, S.J. The role of histone modifications: From neurodevelopment to neurodiseases. Signal Transduct. Target. Ther. 2022, 7, 217.
  77. Nitsch, S.; Zorro Shahidian, L.; Schneider, R. Histone acylations and chromatin dynamics: Concepts, challenges, and links to metabolism. EMBO Rep. 2021, 22, e52774.
  78. Farrelly, L.A.; Thompson, R.E.; Zhao, S.; Lepack, A.E.; Lyu, Y.; Bhanu, N.V.; Zhang, B.; Loh, Y.E.; Ramakrishnan, A.; Vadodaria, K.C.; et al. Histone serotonylation is a permissive modification that enhances TFIID binding to H3K4me3. Nature 2019, 567, 535–539.
  79. Lepack, A.E.; Werner, C.T.; Stewart, A.F.; Fulton, S.L.; Zhong, P.; Farrelly, L.A.; Smith, A.C.W.; Ramakrishnan, A.; Lyu, Y.; Bastle, R.M.; et al. Dopaminylation of histone H3 in ventral tegmental area regulates cocaine seeking. Science 2020, 368, 197–201.
  80. Jos, S.; Gogoi, H.; Prasad, T.K.; Hurakadli, M.A.; Kamariah, N.; Padmanabhan, B.; Padavattan, S. Molecular insights into alpha-synuclein interaction with individual human core histones, linker histone, and dsDNA. Protein Sci. 2021, 30, 2121–2131.
  81. Jiang, P.; Gan, M.; Yen, S.H.; McLean, P.J.; Dickson, D.W. Histones facilitate alpha-synuclein aggregation during neuronal apoptosis. Acta Neuropathol. 2017, 133, 547–558.
  82. Li, P.; Ge, J.; Li, H. Lysine acetyltransferases and lysine deacetylases as targets for cardiovascular disease. Nat. Rev. Cardiol. 2020, 17, 96–115.
  83. Mazzocchi, M.; Wyatt, S.L.; Mercatelli, D.; Morari, M.; Morales-Prieto, N.; Collins, L.M.; Sullivan, A.M.; O’Keeffe, G.W. Gene Co-expression Analysis Identifies Histone Deacetylase 5 and 9 Expression in Midbrain Dopamine Neurons and as Regulators of Neurite Growth via Bone Morphogenetic Protein Signaling. Front. Cell. Dev. Biol. 2019, 7, 191.
  84. Shukla, S.; Tekwani, B.L. Histone Deacetylases Inhibitors in Neurodegenerative Diseases, Neuroprotection and Neuronal Differentiation. Front. Pharmacol. 2020, 11, 537.
  85. Mazzocchi, M.; Collins, L.M.; Sullivan, A.M.; O’Keeffe, G.W. The class II histone deacetylases as therapeutic targets for Parkinson’s disease. Neuronal Signal. 2020, 4, NS20200001.
  86. Shvedunova, M.; Akhtar, A. Modulation of cellular processes by histone and non-histone protein acetylation. Nat. Rev. Mol. Cell Biol. 2022, 23, 329–349.
  87. Chawla, S.; Vanhoutte, P.; Arnold, F.J.; Huang, C.L.; Bading, H. Neuronal activity-dependent nucleocytoplasmic shuttling of HDAC4 and HDAC5. J. Neurochem. 2003, 85, 151–159.
  88. Kutil, Z.; Meleshin, M.; Baranova, P.; Havlinova, B.; Schutkowski, M.; Barinka, C. Characterization of the class IIa histone deacetylases substrate specificity. FASEB J. 2022, 36, e22287.
  89. Jin, H.; Kanthasamy, A.; Ghosh, A.; Yang, Y.; Anantharam, V.; Kanthasamy, A.G. alpha-Synuclein negatively regulates protein kinase Cdelta expression to suppress apoptosis in dopaminergic neurons by reducing p300 histone acetyltransferase activity. J. Neurosci. 2011, 31, 2035–2051.
  90. Lee, J.Y.; Kim, H.; Jo, A.; Khang, R.; Park, C.H.; Park, S.J.; Kwag, E.; Shin, J.H. alpha-Synuclein A53T Binds to Transcriptional Adapter 2-Alpha and Blocks Histone H3 Acetylation. Int. J. Mol. Sci. 2021, 22, 5392.
  91. Renani, P.G.; Taheri, F.; Rostami, D.; Farahani, N.; Abdolkarimi, H.; Abdollahi, E.; Taghizadeh, E.; Gheibi Hayat, S.M. Involvement of aberrant regulation of epigenetic mechanisms in the pathogenesis of Parkinson’s disease and epigenetic-based therapies. J. Cell Physiol. 2019, 234, 19307–19319.
  92. Wu, Q.; Yang, X.; Zhang, L.; Zhang, Y.; Feng, L. Nuclear Accumulation of Histone Deacetylase 4 (HDAC4) Exerts Neurotoxicity in Models of Parkinson’s Disease. Mol. Neurobiol. 2017, 54, 6970–6983.
  93. Paiva, I.; Pinho, R.; Pavlou, M.A.; Hennion, M.; Wales, P.; Schutz, A.L.; Rajput, A.; Szego, E.M.; Kerimoglu, C.; Gerhardt, E.; et al. Sodium butyrate rescues dopaminergic cells from alpha-synuclein-induced transcriptional deregulation and DNA damage. Hum. Mol. Genet. 2017, 26, 2231–2246.
  94. Zhang, Y.; Xu, S.; Qian, Y.; He, X.; Mo, C.; Yang, X.; Xiao, Q. Sodium butyrate attenuates rotenone-induced toxicity by activation of autophagy through epigenetically regulating PGC-1alpha expression in PC12 cells. Brain Res. 2022, 1776, 147749.
  95. Mazzocchi, M.; Goulding, S.R.; Wyatt, S.L.; Collins, L.M.; Sullivan, A.M.; O’Keeffe, G.W. LMK235, a small molecule inhibitor of HDAC4/5, protects dopaminergic neurons against neurotoxin- and alpha-synuclein-induced degeneration in cellular models of Parkinson’s disease. Mol. Cell. Neurosci. 2021, 115, 103642.
  96. Harrison, I.F.; Powell, N.M.; Dexter, D.T. The histone deacetylase inhibitor nicotinamide exacerbates neurodegeneration in the lactacystin rat model of Parkinson’s disease. J. Neurochem. 2019, 148, 136–156.
  97. Hasegawa, T.; Baba, T.; Kobayashi, M.; Konno, M.; Sugeno, N.; Kikuchi, A.; Itoyama, Y.; Takeda, A. Role of TPPP/p25 on alpha-synuclein-mediated oligodendroglial degeneration and the protective effect of SIRT2 inhibition in a cellular model of multiple system atrophy. Neurochem. Int. 2010, 57, 857–866.
  98. Outeiro, T.F.; Kontopoulos, E.; Altmann, S.M.; Kufareva, I.; Strathearn, K.E.; Amore, A.M.; Volk, C.B.; Maxwell, M.M.; Rochet, J.C.; McLean, P.J.; et al. Sirtuin 2 inhibitors rescue alpha-synuclein-mediated toxicity in models of Parkinson’s disease. Science 2007, 317, 516–519.
  99. Liu, Y.; Zhang, Y.; Zhu, K.; Chi, S.; Wang, C.; Xie, A. Emerging Role of Sirtuin 2 in Parkinson’s Disease. Front. Aging Neurosci. 2019, 11, 372.
  100. Zhang, Y.; Kwon, S.; Yamaguchi, T.; Cubizolles, F.; Rousseaux, S.; Kneissel, M.; Cao, C.; Li, N.; Cheng, H.L.; Chua, K.; et al. Mice lacking histone deacetylase 6 have hyperacetylated tubulin but are viable and develop normally. Mol. Cell. Biol. 2008, 28, 1688–1701.
  101. Di Nisio, E.; Lupo, G.; Licursi, V.; Negri, R. The Role of Histone Lysine Methylation in the Response of Mammalian Cells to Ionizing Radiation. Front. Genet. 2021, 12, 639602.
  102. Blanc, R.S.; Richard, S. Arginine Methylation: The Coming of Age. Mol. Cell 2017, 65, 8–24.
  103. Stopa, N.; Krebs, J.E.; Shechter, D. The PRMT5 arginine methyltransferase: Many roles in development, cancer and beyond. Cell. Mol. Life Sci. 2015, 72, 2041–2059.
  104. Chen, K.; Bennett, S.A.; Rana, N.; Yousuf, H.; Said, M.; Taaseen, S.; Mendo, N.; Meltser, S.M.; Torrente, M.P. Neurodegenerative Disease Proteinopathies Are Connected to Distinct Histone Post-translational Modification Landscapes. ACS Chem. Neurosci. 2018, 9, 838–848.
  105. Sugeno, N.; Jackel, S.; Voigt, A.; Wassouf, Z.; Schulze-Hentrich, J.; Kahle, P.J. alpha-Synuclein enhances histone H3 lysine-9 dimethylation and H3K9me2-dependent transcriptional responses. Sci. Rep. 2016, 6, 36328.
  106. Garcia-Reitbock, P.; Anichtchik, O.; Bellucci, A.; Iovino, M.; Ballini, C.; Fineberg, E.; Ghetti, B.; Della Corte, L.; Spano, P.; Tofaris, G.K.; et al. SNARE protein redistribution and synaptic failure in a transgenic mouse model of Parkinson’s disease. Brain 2010, 133 Pt 7, 2032–2044.
  107. Cenik, B.K.; Shilatifard, A. COMPASS and SWI/SNF complexes in development and disease. Nat. Rev. Genet. 2021, 22, 38–58.
  108. Ho, L.; Jothi, R.; Ronan, J.L.; Cui, K.; Zhao, K.; Crabtree, G.R. An embryonic stem cell chromatin remodeling complex, esBAF, is an essential component of the core pluripotency transcriptional network. Proc. Natl. Acad. Sci. USA 2009, 106, 5187–5191.
  109. Sun, L.; Zhang, J.; Chen, W.; Chen, Y.; Zhang, X.; Yang, M.; Xiao, M.; Ma, F.; Yao, Y.; Ye, M.; et al. Attenuation of epigenetic regulator SMARCA4 and ERK-ETS signaling suppresses aging-related dopaminergic degeneration. Aging Cell 2020, 19, e13210.
  110. Sanphui, P.; Kumar Das, A.; Biswas, S.C. Forkhead Box O3a requires BAF57, a subunit of chromatin remodeler SWI/SNF complex for induction of p53 up-regulated modulator of apoptosis (Puma) in a model of Parkinson’s disease. J. Neurochem. 2020, 154, 547–561.
More
This entry is offline, you can click here to edit this entry!
Video Production Service